paper_id
stringlengths
9
16
version
stringclasses
26 values
yymm
stringclasses
311 values
created
timestamp[s]
title
stringlengths
6
335
secondary_subfield
sequencelengths
1
8
abstract
stringlengths
25
3.93k
primary_subfield
stringclasses
124 values
field
stringclasses
20 values
fulltext
stringlengths
0
2.84M
1804.04335
1
1804
2018-04-12T06:14:07
Sparse Reconstruction with Multiple Walsh matrices
[ "math.FA", "eess.SP" ]
The problem of how to find a sparse representation of a signal is an important one in applied and computational harmonic analysis. It is closely related to the problem of how to reconstruct a sparse vector from its projection in a much lower-dimensional vector space. This is the setting of compressed sensing, where the projection is given by a matrix with many more columns than rows. We introduce a class of random matrices that can be used to reconstruct sparse vectors in this paradigm. These matrices satisfy the restricted isometry property with overwhelming probability. We also discuss an application in dimensionality reduction where we initially discovered this class of matrices.
math.FA
math
SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES ENRICO AU-YEUNG Abstract. The problem of how to find a sparse representation of a signal is an important one in applied and computational harmonic analysis. It is closely related to the problem of how to reconstruct a sparse vector from its projection in a much lower-dimensional vector space. This is the setting of compressed sensing, where the projection is given by a matrix with many more columns than rows. We introduce a class of random matrices that can be used to reconstruct sparse vectors in this paradigm. These matrices satisfy the restricted isometry property with overwhelming probability. We also discuss an application in dimensionality reduction where we initially discovered this class of matrices. 8 1 0 2 r p A 2 1 ] . A F h t a m [ 1 v 5 3 3 4 0 . 4 0 8 1 : v i X r a 1. Introduction and Motivation In an influential survey paper by Bruckstein, Donoho, and Elad, the problem of finding a sparse solution to an underdetermined linear system is discussed in great detail [6]. This is an important problem in applied and computational harmonic analysis. Their survey provides plenty of inspiration for future directions of research, with both theoretical and practical consideration. To make this presentation complete, we provide a brief overview. To motivate our discussion, we start by reviewing how sparsity and redundancy are brought to use. Suppose we have a signal which we regard as a nonzero vector y ∈ Rn and there are two available orthonormal bases Ψ and Φ. Then the vector can be expressed as a linear combination of the columns of Ψ or as a linear combination of the columns of Φ, y = Ψα = Φβ. An important example is to take Ψ to be the identity matrix, and Φ to be the matrix for discrete cosine transform. In this case, α is the representation of the signal in the time domain (or space domain) and β is the representation in the frequency domain. For some pairs of orthonormal bases, such as the ones we have just mentioned, either the coefficients α can be sparse, or the β can be sparse, but they cannot both be sparse. This interesting phenomenon is sometimes called the Uncertainty Principle : √ (cid:107)α(cid:107)0 + (cid:107)β(cid:107)0 ≥ 2 n. Here, we have written (cid:107)α(cid:107)0 to denote the sparsity of α, which is the number of nonzero n nonzero entries entries in the vector. This means that a signal cannot have fewer than in both the time domain and the frequency domain. Since the signal is sparse in either the time domain or the frequency domain, but not in both, this leads to the idea of combining the two bases by concatenating the two matrices into one matrix A = [Ψ Φ]. √ By a representation of the signal y, we mean a column vector x so that y = Ax. The representation of the signal is not unique because the column vectors of A are not linearly independent. From this observation, we are naturally led to consider a matrix A formed 1 2 ENRICO AU-YEUNG by combining more than two bases. The hope is that among the many possible ways of representing the signal y, there is at least one representation that is very sparse, i.e. most entries of x are zero. We want the vector x to be s-sparse, which means that at most s of the entries are nonzero. A natural question that arises is how to find the sparse represen- tation of a given signal y. There is a closely related problem that occurs commonly in signal and image processing. Suppose we begin with a vector x ∈ RN that is s-sparse, which we consider to be our com- pressible signal. Using the matrix A ∈ Rn×N , we observe the vector y from the projection y = Ax. This leads to the following problem: given a matrix A ∈ Rn×N , where typically N is much larger than n, and given y ∈ Rn, how to recover the s-sparse vector x ∈ RN from the observation y = Ax. The term most commonly used in this setting is compressed sensing. This problem is NP-hard, i.e. the natural approach to consider all possible s-sparse vectors in RN is not feasible. The reconstruction of the vector x is accomplished by a non-linear operator ∆ : RN → Rn that solves the minimization problem, (P 1) min(cid:107)x(cid:107)1 subject to y = Ax. The following definition plays a central role in this paper. Definition 1.1. A matrix A ∈ Rm×N is said to have the restricted isometry property (RIP) of order s and level δs ∈ (0, 1) if (1 − δs)(cid:107)x(cid:107)2 2 ≤ (cid:107)Ax(cid:107)2 2 ≤ (1 + δs)(cid:107)x(cid:107)2 for all s-sparse x ∈ RN . 2 The restricted isometry property says that the columns of any sub-matrix with at most s columns are close to being orthogonal to each other. If the matrix A satisfies this property, then the solution to (P1) is unique, i.e. it is possible to reconstruct the s-sparse vector by minimizing the l1 norm of x, subject to y = Ax. For this reason, matrices that satisfy the RIP play a key role in compressed sensing. Some examples of random matrices that satisfy the RIP are the Gaussian, Bernoulli, or partial random Fourier matrices. From the foundational papers of Donoho [11] and Candes, Romberg, and Tao [8, 9], the field of compressed sensing has been studied and extended by many others to include a broad range of theoretical issues and applications; see, for example, [7, 10, 2, 5, 19, 1, 12, 13, 25, 26, 20, 23, 22, 16, 27], and the comprehensive treatment found in [14]. The search for structured matrices that can be used in compressed sensing continues to be an active research area (see, e.g., [21].) Towards that goal, our contribution is to introduce a class of random matrices that satisfy the RIP with overwhelming probability. We also describe an application where we initially discovered this class of matrices. 1.1. Application. Dimensionality reduction is another area where matrices that satisfy the RIP play an important role. A powerful tool is the Johnson-Lindenstrauss (JL) lemma. This lemma tells us that the distance between each pair of points in a high-dimensional space is nearly preserved if we project the points into a much lower-dimensional space using a random linear mapping. Krahmer and Ward [16] showed that if a matrix satisfies the RIP, then we can use it to create such a mapping if we randomize the column signs of the matrix. For a precise statement, see [16]. Together with our matrix that satisfies the RIP, SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 3 their result allows one to create a matrix to be used in a JL-type embedding. To demon- strate this in a concrete setting, let us turn to an application in robust facial recognition. The goal of object recognition is to use training samples from k distinct object classes to determine the class to which a new sample belongs. We arrange the given nj training samples from the j-th class as columns of a matrix Yj ≡ [vj,1, vj,2, . . . , vj,nj ] ∈ Rm×nj . In the context of a face recognition system, we identify a w×h facial image with the vector v ∈ Rm (m = wh) given by stacking its columns. Therefore, the columns of Yj are the training facial images of the j-th person. One effective approach for exploiting the structure of the Yj in object recognition is to model the samples from a single class as lying on a linear subspace. Subspace models are flexible enough to capture the structure in real data, where it has been demonstrated that the images of faces under varying lighting and expressions lie on a low-dimensional subspace [3]. For our present discussion, we will assume that the training samples from a single class do lie on a single subspace. Suppose we are given sufficient training samples of the j-th object class, Yj ≡ [vj,1, vj,2, . . . , vj,nj ] ∈ Rm×nj . Then, any new sample ynew ∈ Rm from the same class will approximately lie in the linear span of the training samples associated with object j, i.e. ynew = cj,1vj,1 + cj,2vj,2 + . . . cj,nj vj,nj , for some coefficients cj,k ∈ R, 1 ≤ k ≤ nj. We define a new matrix Φ for the entire training set as the concatenation of the n training samples of all k object classes, Φ = [Y1, Y2, Y3, . . . , Yk] = [v1,1, v1,2, . . . , v2,1, v2,2, . . . , vk,nk]. The new sample ynew ∈ Rm can be expressed as a linear combination of all training samples, ynew = Φx, where the transpose of the vector x is of the form, x = [0, 0, . . . , 0, cj,1, cj,2, . . . , cj,nj , 0, 0, . . . , 0] ∈ Rn, i.e. x is the coefficient vector whose entries are zero, except those entries associated with the j-th class. The sparse vector x encodes the identity of the new sample ynew. The task of classifying a new sample amounts to solving the linear system ynew = Φx to recover the sparse vector x. For more details, see [28], where the authors presented strong experimental evidence to support this approach to robust facial recognition. One practical issue that arises is that for face images without any pre-processing, the corresponding linear system y = Φx is very large. For example, if each face image is given at a typical resolution of 640 × 480 pixels, then the matrix Φ has m rows, where m is in the order of 105. Using scalable algorithms, such as linear programming, applying this directly to high-resolution images still requires enormous computing power. Dimensionality reduction becomes indispensable in this setting. The projection from the image space to the much lower-dimensional feature space can be represented by a matrix P , where P has many more columns than rows. The linear system y = Φx then becomes (cid:101)y ≡ P y = P Φx. 4 The new sample y is replaced by its projection(cid:101)y. The sparse vector x is reconstructed by ENRICO AU-YEUNG solving the minimization problem, min(cid:107)x(cid:107)1 subject to y = P Φx. In the past, enormous amount of effort was spent to develop feature-extraction methods for finding projections of images into lower-dimensional spaces. Examples of feature-extraction methods include EigenFace, FisherFace, and a host of creative techniques; see, e.g. [4]. For the approach to facial recognition that we have described, choosing a matrix P is no longer a difficult task. We can select a matrix P so that it nearly preserves the distance between every pair of vectors, i.e. (cid:107)P x − P y(cid:107)2 ≈ (cid:107)x − y(cid:107)2. As mentioned earlier, beginning with a matrix A that satisfies the RIP, the result of Krahmer and Ward allows one to create a matrix P to be used in a JL-type embedding. 1.2. Notation. Before continuing further, we need to define some terminology. The Rademacher system {rn(x)} on the interval [0, 1] is a set of orthogonal functions defined by rn(x) = sign(sin(2n+1πx)); n = 0, 1, 2, 3, . . . The Rademacher system does not form a basis for L2([0, 1]), but this can be remedied by considering the Walsh system of functions. Each Walsh function is a product of Rademacher functions. The sequence of Walsh functions is defined as follows. Every positive integer n can be written in the binary system as: where the integers nj are uniquely determined by nj+1 < nj. The Walsh functions {Wn(x)}∞ are then given by n=0 n = 2n1 + 2n2 + . . . + 2nk, W0(x) = 1, Wn(x) = rn1(x)rn2(x) . . . rnk(x). The Walsh system forms an orthogonal basis for L2([0, 1]). There is a convenient way to represent these functions as vectors. Define the matrices H0 = 1, and for n ≥ 1, (cid:20) Hn−1 −Hn−1 Hn−1 Hn−1 (cid:21) . Hn = 1√ 2 Then, the column vectors of Hn form an orthogonal basis on R2n. Note that the matrix Hn has 2n rows and 2n columns. Because of its close connection to the Walsh system, a matrix of the form Hn is called a Hadamard-Walsh matrix. The inner product of two vectors x and y is denoted by (cid:104)x, y(cid:105). The Euclidean norm of a vector x is denoted by (cid:107)x(cid:107)2. If a vector has at most s nonzero entries, we say that the vector is s-sparse. For clarity, we often label constants by C1, C2, . . ., but we do not keep track of their precise values. For a matrix A ∈ Rm×N , its operator norm is (cid:107)A(cid:107) = sup{(cid:107)Ax(cid:107)2 : (cid:107)x(cid:107)2 = 1}. If x ∈ RN , then we say that Γ is the support set of the vector if the entries of x are nonzero only on the set Γ, and we write supp(x) = Γ. We define BΓ = {x ∈ RN : (cid:107)x(cid:107)2 = 1, supp(x) = Γ}. We write A∗ for the adjoint (or transpose) of the matrix. Working with s-sparse vectors, there is another norm defined by (cid:107)A(cid:107)Γ = sup{(cid:104)Ax, y(cid:105) : x ∈ BΓ, y ∈ BΓ, Γ ≤ s}. SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 5 This norm is important because if the matrix A obeys the relation (cid:107)I − A∗A(cid:107)Γ ≤ δs, then A satisfies the RIP of order s and level δs. Let us introduce a model called Sparse City. From now on, we fix a positive integer m that is a power of 2, i.e. m = 2k, and focus on a single Hadamard-Walsh matrix W with m n be the m× n matrix formed by selecting the first n columns rows and m columns. Let W m of the matrix W . Let Θ be a bounded random variable with expected value zero, i.e. E(Θ) = 0,Θ ≤ B. To be precise, the random variable Θ is equally likely to take one of the four possible values, {1,−1, 3,−3}. Define the random vectors x1, x2, . . . xb ∈ Rm, so that the entries of each vector are independent random variables drawn from the same probability distribution as Θ. For each vector xj = (θj1, θj2, . . . , θjm), we have E(θjw) = 0 and θjw ≤ B, for 1 ≤ w ≤ m. We associate a matrix Dj to each vector xj, so that each Dj ∈ Rm×m is a diagonal matrix with the entries of xj along the diagonal. To construct the matrix A, we concatenate b blocks of DjW m n , so that written out in block form, n D2W m n D3W m n D4W m n . . . . . . DbW m n A =(cid:2) D1W m Note that the matrix A has m rows and nb columns. In our application, Walsh matrices are more appropriate than other orthogonal matrices, such as discrete cosine transforms (DCT). For illustration, if A ∈ R1024×20480, with b = 320 blocks, then each entry of 64A is one of the four values {1,−1, 3,−3}. Consider any vector y that contains only integer values, ranging from 0 to 255, which are typical for facial images. The product Ay can be computed from 64× Ay; the calculation of 64× Ay uses only integer-arithmetic operations. 1.3. Main results. Our first result is that the matrix satisfies the RIP in expectation. Theorem 1.2. Let W be the Hadamard-Walsh matrix with m rows and m columns. Let n be the m × n matrix formed by selecting the first n columns of the matrix W . The W m matrix A ∈ Rm×nb is constructed by concatenating b blocks of DjW m n D2W m n D3W m n D4W m n , so that n . . . . . . DbW m n A =(cid:2) D1W m Each Dj ∈ Rm×m is a diagonal matrix, as defined in section (1.2). Then, there exists a constant C > 0 such that for any 0 < δs ≤ 1, we have E(cid:107)I − A∗A(cid:107)Γ ≤ δs (cid:3) . (cid:3) . provided that and m ≤ nb. More precisely, there are constants C1 and C2 so that m ≥ C · δ−2 · s · log4(nb) s (cid:115) (1) provided that E(cid:107)I − A∗A(cid:107)Γ ≤ C1 · s · log2(s) · log(mb) · log(nb) m m ≥ C2 · s · log2(s) log(mb) log(nb). The next theorem tells us that the matrix satisfies the restricted isometry property with overwhelming probability. 6 ENRICO AU-YEUNG Theorem 1.3. Fix a constant δs > 0. Let A be the matrix specified in Theorem 1.2. Then, there exists a constant C > 0 such that with probability at least 1 −  provided that (cid:107)I − A∗A(cid:107)Γ < δs m ≥ C · δ−2 · s · log4(nb) · log(1/) s and m ≤ nb. 1.4. Related Work. Gaussian and Bernoulli matrices satisfy the restricted isometry prop- erty (RIP) with overwhelmingly high probability, provided that the number of measure- ments m satisfies m = O(s log( N s )). Although these matrices require the least number of measurements, they have limited use in practical applications. Storing an unstructured matrix, in which all the entries of the matrix are independent of each other, requires a prohibited amount of storage. From a computation and application view point, this has motivated the need to find structured random matrices that satisfy the RIP. Let us review three of the most popular classes of random matrices that are appealing alternatives to the Gaussian matrices. For a broad discussion and other types of matrices, see [18] and [21]. The random subsampled Fourier matrix is constructed by randomly choosing m rows from the N × N discrete Fourier transform (DFT) matrix. In this case, it is important to note that the fast Fourier transform (FFT) algorithm can be used to significantly speed up the matrix-by-vector multiplication. A random subsampled Fourier matrix with m rows and N columns satisfies the RIP with high probability, provided that m ≥ C · δ2s log(N )4, where C is a universal constant; see [23] for a precise statement. The next type of structured random matrices are partial random Toeplitz and circulant matrices. These matrices naturally arise in applications where convolutions are involved. Recall that for a Toeplitz matrix, each entry aij in row i and column j is determined by the value of i− j, so that for example, a11 = a22 = a33 and a21 = a32 = a43. To construct a ran- dom m×N Toeplitz matrix A, only N +m−2 random numbers are needed. Haupt, Bajwa, Raz, Wright and Nowak [15] showed that the matrix A satisfies the RIP of order 3s with high probability for every δ ∈ (0, 1 s ), where C is a constant. 3), provided that m > C·s3 log( N There are many situations in signal processing where we encounter signals that are band-limited and are sparse in the frequency domain. The random demodulator matrix is suitable in this setting [27]. For motivation, imagine that we try to acquire a single high-frequency tone that lies within a wide spectral band. Then, a low-rate sampler with an antialiasing filter will be oblivious to any tone whose frequency exceeds the passband of the filter. To deal with this problem, the random demodulator smears the tone across the entire spectrum so that it leaves a signature that a low-rate sampler can detect. Consider a signal whose highest frequency does not exceed W 2 hertz. We can give a mathematical description of the system. Let D be a W × W diagonal matrix, with random numbers along the diagonal. Next, we consider the action of the sampler and suppose the sampling rate is R, where R divides W . Each sample is then the sum of W R consecutive entries of the demodulated signal. The action of the sampling is specified by a matrix G with R rows and W columns, such that the r-th row has W R ) + 1 R consecutive ones, beginning in column ( rW SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 7 for each r = 0, 1, 2, . . . , R − 1. For example, when W = 12 and R = 3, we have  1 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1  . G = Define the R × W matrix A = GDF , where F is the W × W discrete Fourier transform (DFT) matrix with the columns permuted; see [27] for further detail. For a fixed δ > 0, if the sampling rate R is greater than or equal to Cδ−2 · s log(W )6, then an R × W random demodulation matrix A has the RIP of order s with constant δs ≤ δ, with probability at least 1 − O( 1 W ). A =(cid:2) D1W m In contrast to the classes of matrices described above, the class of structured random matrices introduced in Sparse City has a block form. The matrix A ∈ Rm×nb is constructed by concatenating b blocks. More precisely, n D3W m n D4W m n D2W m n , the same m × n matrix W m n . . . . . . DbW m In each block of DjW m n is used, but each block has its own random diagonal matrix Dj. For compressed sensing to be useful in applications, we need to have suitable hardware and a data acquisition system. In seismic imaging, the signals are often measured by multiple sensors. A signal can be viewed as partitioned into many parts. Different sensors are responsible for measuring different parts of the signal. Each sensor is equipped with its own scrambler which it uses to randomly scramble the measurements. The block structure of the sensing matrix A facilitates the design of a suitable data acquisition scheme tailored to this setting. (cid:3) . n 8 ENRICO AU-YEUNG 2. Mathematical tools We collect together the tools we need to prove the main results. We begin with a fundamental result by Rudelson and Vershynin [23], followed by an extension of this result, and then a concentration inequality. In what follows, for vectors x, y ∈ Rn, the tensor x⊗ y is the rank-one operator defined by (x⊗y)(z) = (cid:104)x, z(cid:105)y. For a given subset Γ ⊆ {1, 2, . . . , n}, the notation xΓ is the restriction of the vector x on the coordinates in the set Γ. Let x1, x2, x3, . . . , xm, with m ≤ n, be vectors in Rn with uniformly bounded entries, Lemma 2.1. (Rudelson and Vershynin) (cid:107)xi(cid:107)∞ ≤ K for all i. Then (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) m(cid:88) i=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ M · sup (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) m(cid:88) √ log n Γ≤s √ log m. i=1 i xΓ i i ⊗ xΓ √ s log(s) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 i ⊗ xΓ xΓ i (2) E sup Γ≤s where the constant M equals C1(K) Since our next lemma is an extension of this lemma, we provide a review of the main ideas in the proof of Lemma (2.1). Let E1 denote the left-hand side of (2). We will bound E1 by the supremum of a Gaussian process. Let g1, g2, g3, . . . , gm be independent standard normal random variables. The expected value of gi is a constant that does not depend on the index i. E1 ≤ C3 · E sup Γ≤s ≤ C3 · E sup Γ≤s = C3 · E sup i i ⊗ xΓ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) : Γ ≤ s, x ∈ BΓ (cid:41) E gi i xΓ gi xΓ i ⊗ xΓ i gi(cid:104)xi, x(cid:105)2 i=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) m(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) m(cid:88) (cid:40)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) m(cid:88) m(cid:88) i=1 i=1 Az = gi(cid:104)xi, z(cid:105)xi To see that the last equality is true, consider an operator A on Rn defined by i=1 and since A is a self-adjoint operator, it follows that (cid:104)Az, z(cid:105) = sup (cid:107)z(cid:107)=1 (cid:107)A(cid:107)op = sup (cid:107)z(cid:107)=1 m(cid:88) i=1 gi(cid:104)xi, z(cid:105)2. For each vector u in Rn, we consider the Gaussian process m(cid:88) G(u) = gi(cid:104)xi, u(cid:105)2. This Gaussian process is a random process indexed by vectors in Rn. i=1 SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 9 Thus to obtain an upper bound on E1, we need an estimate on the expected value of the supremum of a Gaussian process over an arbitrary index set. We use Dudley's Theorem (see [24], Proposition 2.1) to obtain an upper bound. Theorem 2.2. Let (X(t) : t ∈ T ) be a Gaussian process with the associated pseudo-metric d(s, t) =(cid:0)E X(s) − X(t)2(cid:1)1/2 . Then there exists a constant K > 0 such that X(t) ≤ K (cid:112)log N (T, d, u) du. (cid:90) ∞ E sup t∈T Here, T is an arbitrary index set, and the covering number N (T, d, u) is the smallest number of balls of radius u to cover the set T with respect to the pseudo-metric. T = BΓ 0 Γ≤s (cid:91) (cid:91) Γ≤s 1/2 BΓ,(cid:107) · (cid:107)G, u du, There is a semi-norm associated with the Gaussian process, so that if x and y are any By applying Dudley's inequality with the above calculations show that, (3) E1 ≤ C4 where N is the covering number. 0 (cid:90) ∞ log N two fixed vectors in Rn, then i=1 = ≤ ((cid:104)xi, x(cid:105) + (cid:104)xi, y(cid:105))2 (cid:34) m(cid:88) (cid:34) m(cid:88) (cid:107)x − y(cid:107)G = (cid:0)E G(x) − G(y)2(cid:1)1/2 (cid:35)1/2 (cid:0)(cid:104)xi, x(cid:105)2 − (cid:104)xi, y(cid:105)2(cid:1)2 (cid:35)1/2 (cid:35)1/2 (cid:34) m(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) m(cid:88)  1√ ≤ 2 max Γ≤s,z∈BΓ R ≡ sup Γ≤s E3 ≤ C5R (cid:90) ∞ (cid:104)xi, z(cid:105)2 log1/2 N 2 i=1 √ s i=1 i=1 0 where (4) Here, the semi-norm (cid:107)x(cid:107)X is defined by (cid:107)x(cid:107)X = max i≤m (cid:104)xi, x(cid:105) . Thus, by a change of variable in the integral in (3), we see that · max i≤m (cid:104)xi, x − y(cid:105) = 2 R max i≤m (cid:104)xi, x − y(cid:105) , · max i≤m (cid:104)xi, x − y(cid:105) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 . i i ⊗ xΓ xΓ (cid:91) s T≤s  du. BΓ,(cid:107) · (cid:107)X, u 10 It is sufficient to show the integral in (4) is bounded by C11(K) · log(s) · √ log n · √ ENRICO AU-YEUNG log m. This concludes our review of the main ideas in the proof of Lemma (2.1). We extend the fundamental lemma of Rudelson and Vershynin. The proof follows the strategy of the proof of the original lemma, with an additional ingredient. The Gaussian process involved is replaced by a tensorized version, with the appropriate tensor norm. Let u1, u2, u3, . . . , uk, and v1, v2, v3, . . . , vk, with k ≤ n, be vectors in Rn with uniformly Lemma 2.3. (Extension of the fundamental lemma of Rudelson and Vershynin) bounded entries, (cid:107)ui(cid:107)∞ ≤ K and (cid:107)vi(cid:107)∞ ≤ K for all i. Then (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) i=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ M ·  sup Γ≤s (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) i=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) i=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (5) E sup Γ≤s i uΓ i ⊗ vΓ i i ⊗ uΓ uΓ i + sup Γ≤s i ⊗ vΓ vΓ i where the constant M depends on K and the sparsity s. Proof. Let E1 denote the left-hand side of (5). Our plan is to bound E1 by the supremum of a Gaussian process. Let g1, g2, g3, . . . , gk be independent standard normal random variables. Then (cid:40)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k(cid:88) (cid:40)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k(cid:88) i=1 i=1 E1 = E sup i(cid:104)xp, ui(cid:105)(cid:104)vi, xq(cid:105) ≤ C3 · E sup gi(cid:104)xp, ui(cid:105)(cid:104)vi, xq(cid:105) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) : Γ ≤ s, xp ∈ BΓ, xq ∈ BΓ (cid:41) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) : Γ ≤ s, xp ∈ BΓ, xq ∈ BΓ (cid:41) When G(x) is a Gaussian process indexed by the elements x in an arbitrary index set T , Dudley's inequality states that G(x) ≤ C · E sup x∈T log1/2 N (T, d, u) du, (cid:90) ∞ 0 (cid:88) (cid:91) i Γ≤s with the pseudo-metric d given by d(x, y) =(cid:0)E G(x) − G(y)2(cid:1)1/2 . Our Gaussian process is indexed by two vectors xp and xq so that and the index set is T = G(xp, xq) = gi(cid:104)xp, ui(cid:105)(cid:104)vi, xq(cid:105) BΓ ⊗ BΓ. SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 11 ((cid:104)xp + yp, ui(cid:105)(cid:104)vi, xq − yq(cid:105) + (cid:104)xp − yp, ui(cid:105)(cid:104)vi, xq + yq(cid:105)) (cid:35)1/2 (cid:35)1/2 ((cid:104)xp + yp, ui(cid:105) + (cid:104)xq + yq, vi(cid:105))2 The pseudo-metric on T is given by d ((xp, xq), (yp, yq)) ((cid:104)xp, ui(cid:105)(cid:104)vi, xq(cid:105) − (cid:104)yp, ui(cid:105)(cid:104)vi, yq(cid:105))2 (cid:34) k(cid:88) (cid:34) k(cid:88) i=1 1 2 i=1 = = · max ≤ 1 2 ≤ Q · max i i ((cid:104)ui, xp − yp(cid:105) ,(cid:104)vi, xq − yq(cid:105)) · ((cid:104)ui, xp − yp(cid:105) ,(cid:104)vi, xq − yq(cid:105)) , where the quantity Q is defined by (cid:35)1/2 (cid:34) k(cid:88) i=1  (cid:34) k(cid:88) i=1 (cid:35)1/2  : (xp, xq) ∈ Γ (cid:41) (cid:104)xp + yp, ui(cid:105) · (cid:104)xq + yq, vi(cid:105) Q = 1 2 sup ((cid:104)xp + yp, ui(cid:105) + (cid:104)xq + yq, vi(cid:105))2 We bound the quantity Q in the following calculations. Q2 = 1 4 sup (xp,xq)∈Γ i=1 ≤1 4 sup (xp,xq)∈Γ i=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) i=1 i=1 ≤ ≤ = i=1 ui ⊗ ui + ui ⊗ ui + ui ⊗ ui i=1 i=1 + 1 2 vi ⊗ vi k(cid:88) i=1 k(cid:88) (cid:104)xp + yp, ui(cid:105)2 + (cid:104)xp + yp, vi(cid:105)2 + 2 ((cid:104)xp + yp, ui(cid:105) + (cid:104)xq + yq, vi(cid:105))2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ (cid:32) k(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ k(cid:88) (cid:40) k(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ  sup (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) d((xp, xq), (yp, yq)) ≤ S · max(cid:0)(cid:107)xp − yp(cid:107)(U )∞ ,(cid:107)xq − yq(cid:107)(V )∞ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) 2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:33)1/2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:32) k(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k(cid:88) i=1 (cid:104)xp + yp, ui(cid:105)2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 Γ ui ⊗ ui ui ⊗ ui vi ⊗ vi vi ⊗ vi ≡ S2. (xp,xq)∈Γ Γ i=1 Γ i=1 Γ + 2 i=1 i=1 + + · i=1 (cid:104)xq + yq, vi(cid:105)2 (cid:1) . (cid:33)1/2 We now define two norms. Let (cid:107)x(cid:107)(U )∞ = maxi (cid:104)x, ui(cid:105) and (cid:107)x(cid:107)(V )∞ = maxi (cid:104)x, vi(cid:105). The above calculations show that the pseudo-metric satisfies the next inequality, 12 ENRICO AU-YEUNG Let (cid:101)T =(cid:83)Γ≤s BΓ. Then T ⊆ (cid:101)T ⊗ (cid:101)T . Moreover, the covering number of the set T and the covering number of the set (cid:101)T must satisfy the relation Here, d and (cid:101)d are the pseudo-metrics for the corresponding index sets. Consequently, we N (T, d, u) ≤ N ((cid:101)T ⊗ (cid:101)T ,(cid:101)d, u). have (cid:90) ∞ log1/2 N (T, d, u) du log1/2 N ((cid:101)T ,(cid:107) · (cid:107)(U )∞ , u) du + S · 0 ≤ S · (cid:90) ∞ 0 (cid:90) ∞ 0 log1/2 N ((cid:101)T ,(cid:107) · (cid:107)(V )∞ , u) du. We have completed all the necessary modification to the proof of the original lemma. The rest of the proof proceeds in exactly the same manner as the proof of the original lemma, almost verbatim, and we omit the repetition. (cid:3) In order to show that, with high probability, a random quantity does not deviate too much from its mean, we invoke a concentration inequality for sums of independent symmetric random variables in a Banach space. (See [27], Proposition 19, which follows from [17], Theorem 6.17). Proposition 2.4. (Concentration Inequality) Let Y1, Y2, . . . , YR be independent, symmetric random variables in a Banach space X. Assume that each random variable satisfies the bound (cid:107)Yj(cid:107)X ≤ B almost surely, for 1 ≤ j Yj(cid:107)X. Then there exists a constant C so that for all u, t ≥ 1, j ≤ R. Let Y = (cid:107)(cid:80) P (Y > C[uE(Y ) + tB]) ≤ e−u2 + e−t. We define a sequence of vectors that depend on the entries in the matrix W m n . Let ykw ∈ Rnb, where the entries indexed by (k − 1)n + 1, (k − 1)n + 2, (k − 1)n + 3, . . . , kn are from row w of the matrix W m n , while all other entries are zero. The next example illustrates the situation. Example 2.5. Consider the matrix W m n with m rows and n columns. W m n = a(2, 1) a(2, 2) a(3, 1) a(3, 2) a(4, 1) a(4, 2)  a(1, 1) a(1, 2)   a(3, 1)  y13 =  a(2, 1) a(2, 2) a(3, 2) 0 0  y14 =  a(4, 1) a(4, 2) 0 0  0 0  a(1, 1) a(1, 2) 0 0  y12 = y11 = Here, m = 4 and n = 2. We define the vectors y11, y12, y13, y14 by SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES and we define the vectors y21, y22, y23, y24 by  y23 =   y24 =  0 0 a(4, 1) a(4, 2) 13  y21 = a(1, 1) a(1, 2) Since the columns of W m n come from an orthogonal matrix, we have the following relations (a(k, 1))2 = 1, (a(k, 2))2 = 1, a(k, 1)a(k, 2) = 0. The rank-one operator y11⊗ y11 is defined by (y11 ⊗ y11) (z) = (cid:104)z, ykw(cid:105)ykw, for every z ∈ R4. Explicitly in matrix form, this rank-one operator is 0 0  4(cid:88) k=1 y11 ⊗ y11 = 0 0  y22 = a(2, 1) a(2, 2)  4(cid:88)  a(1, 1) · a(1, 1) m(cid:88) a(1, 2) · a(1, 1) 0 0 k=1 b(cid:88) k=1 w=1 0 0 a(3, 1) a(3, 2) 4(cid:88) k=1 a(1, 1) · a(1, 2) a(1, 2) · a(1, 2) 0 0  0 0 0 0 0 0 0 0 We can directly compute and verify that : ykw ⊗ ykw = I, the identity matrix. Remark 2.6. The vectors ykw ∈ Rnb may seem cumbersome at first but they enable us to write the matrix A∗A in a manageable form. The matrix A is constructed from b blocks of DjW m n and so the matrix has the form n D2W m n D3W m n D4W m n . . . . . . DbW m n A =(cid:2) D1W m which means that when b = 3, the matrix A∗A has the form, 0 ∗ (W m n ) 0 0 0 ∗ (W m n ) 1D1 D∗ D∗ 2D1 D∗ D∗ 3D1 D∗ 1D2 D∗ 2D2 D∗ 3D2 D∗ 1D3 2D3 3D3 0 0 0 n 0 W m n 0 W m 0 n For clarity, we have written out the form of A∗A when b = 3. The pattern extends to the general case with b blocks. The key observation is that we can now write (cid:3)  .  W m  (W m ∗ n ) 0 0  D∗ b(cid:88) b(cid:88) m(cid:88) A∗A = θkwθjw ykw ⊗ yjw. This expression for A∗A plays a crucial role in the proof of Theorem 1.2. k=1 j=1 w=1 To show that a quantity P is bounded by some constant, it is enough, as the next lemma √ P + 1). tells us, to show that P is bounded by some constant multiplied by (2 + Lemma 2.7. Fix a constant c1 ≤ 1. If P > 0 and √ (cid:16) (cid:17) P ≤ c1 2 + P + 1 , then P < 5c1. 14 ENRICO AU-YEUNG Proof. Let x = (P + 1)1/2 and note that x is an increasing function of P . The hypothesis of the lemma becomes x2 − 1 ≤ c1(2 + x) which implies that The polynomial on the left is strictly increasing when x ≥ c1/2. Since α ≤ 1 and x ≥ 1 for P ≥ 0, it is strictly increasing over the entire domain of interest, thus x2 − c1x − (2c1 + 1) ≤ 0. x ≤ c1 +(cid:112)(c1)2 + 4(2c1 + 1) (cid:112)(c1)2 + 4(2c1 + 1) c1 2 . + 2 (cid:112)(c1)2 + 4(2c1 + 1) 4 By substituting (P + 1)1/2 back in for x, this means P + 1 ≤ (c1)2 4 + Since c1 < 1, this implies that P < 5c1. 3. Proof of RIP in expectation (Theorem 1.2) The rank-one operators ykw ⊗ ykw are constructed so that (6) ykw ⊗ ykw = I. . (cid:3) b(cid:88) m(cid:88) k=1 w=1 b(cid:88) b(cid:88) m(cid:88) k=1 j=1 w=1 As explained in Remark (2.6) from the last section, we have (7) A∗A = θkwθjw ykw ⊗ yjw. The proof of the theorem proceeds by breaking up I − A∗A into four different parts, then bounding the expected norm of each part separately. By combining equations (6) and (7), we see that (8) I − A∗A = (1 − θkw2) ykw ⊗ ykw + θkwθjw ykw ⊗ yjw. (cid:88) m(cid:88) j(cid:54)=k w=1 b(cid:88) m(cid:88) k=1 w=1 For the two sums on the right hand side of (8), we will bound the expected norm of each sum separately. Define two random quantities Q1 and Q(cid:48) (9) and Q1 = Q(cid:48) 1 = b(cid:88) b(cid:88) k=1 m(cid:88) m(cid:88) w=1 k=1 w=1 1 by (1 − θkw2) ykw ⊗ ykw (1 − θ(cid:48) kw2) ykw ⊗ ykw SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 15 kw} is an independent copy of {θkw}. This implies that Q1 has the same probability where {θ(cid:48) distribution as Q(cid:48) 1. To bound the expected norm of Q1, 1)(cid:107)Γ 1 Q1] (cid:107)Γ 1(cid:107)Γ Q1] 1(cid:107)Γ). E(cid:107)Q1(cid:107)Γ = E(cid:107)Q1 − E(Q(cid:48) = E(cid:107)E[Q1 − Q(cid:48) ≤ E[E(cid:107)Q1 − Q(cid:48) = E((cid:107)Q1 − Q(cid:48) In the above equations, the first equality holds because Q(cid:48) equality holds by the independence of Q1 and Q(cid:48) by Jensen's inequality. Let 1 has mean zero. The second 1. The inequality in the third line is true (10) Y = Q1 − Q(cid:48) 1 = (θ(cid:48) kw2 − θkw2)ykw ⊗ ykw. We randomize this sum. The random variable Y has the same probability distribution as (11) Y (cid:48) = : kw(θ(cid:48) kw2 − θkw2)ykw ⊗ ykw where {kw} are independent, identically distributed Bernoulli random variables. m(cid:88) w=1 b(cid:88) m(cid:88) k=1 b(cid:88) k=1 w=1 E(cid:107)Y (cid:107)Γ = E(cid:107)Y (cid:48)(cid:107)Γ = E [E ((cid:107)Y (cid:48)(cid:107)Γ {θkw},{θ(cid:48) kw})] . B ≥ max (cid:0)θ(cid:48) kw2 − θkw2(cid:1)1/2 , kw2 − θkw2(cid:1)1/2 · (cid:107)ykw(cid:107)∞ ≤ B√ (cid:0)θ(cid:48) k,w Let xkw = (θ(cid:48) To see that each xkw is bounded, we note that kw2 − θkw2)1/2 ykw in order to apply the lemma of Rudelson and Vershynin. and so With the {θkw},{θ(cid:48) E [ (cid:107)Y (cid:48)(cid:107)Γ {θkw},{θ(cid:48) where L ≡ log2(s)·log(nb)·log(mb). To remove the conditioning, we apply Cauchy-Schwarz inequality and the law of double expectation, m . m (cid:107)xkw(cid:107)∞ ≤ max kw} fixed, and with K = B/ k,w √ k=1 w=1 · B · (cid:114) (cid:114) kw}] ≤ m(cid:88) C · s · L m, we apply Lemma (2.1) to obtain (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) m(cid:88) (cid:0)θ(cid:48) kw2 − θkw2(cid:1) · ykw ⊗ ykw (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ (cid:0)θ(cid:48) kw2 − θkw2(cid:1) · ykw ⊗ ykw (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:0)θ(cid:48) kw2 − θkw2(cid:1) · ykw ⊗ ykw (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ (cid:33)1/2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ θkw2 · ykw ⊗ ykw C · s · L m(cid:88) ≤ 2E (cid:32) · B k=1 w=1 k=1 w=1 m Γ . E By using the triangle inequality, (12) E(cid:107)Y (cid:107)Γ ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) k=1 m(cid:88) w=1 E 16 ENRICO AU-YEUNG and so the bound in (12) becomes C · s · L (cid:114) Since(cid:80)b k=1 E(cid:107)Y (cid:107)Γ ≤ (cid:80)m w=1 ykw ⊗ ykw = I and since E(cid:107)I(cid:107)Γ = 1, we have w=1 m E θkw2 · ykw ⊗ ykw · m(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) (cid:32) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) m(cid:88) k=1 w=1 k=1 (cid:32) · E · (E(cid:107)Q1(cid:107)Γ + 1)1/2 · (E(cid:107)Y (cid:107)Γ + 1)1/2 . (cid:114) (cid:114) (cid:114) C · s · L m C · s · L m C · s · L m E(cid:107)Y (cid:107)Γ ≤ = ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ (cid:33)1/2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ . (cid:33)1/2 (1 − θkw)2 · ykw ⊗ ykw + 1 (17) (18) where ukw and vkw are defined by ukw = θkw · ykw, vkw = θ(cid:48) kw · ykw Solutions to the equation E ≤ α(E + 1)1/2 satisfy E ≤ 2α, where α ≤ 1. Hence, the above inequalities show that there exist constants C10, C11 such that if m ≥ C10 · s · L, then (13) E(cid:107)Q1(cid:107)Γ ≤ E(cid:107)Y (cid:107)Γ ≤ (cid:114) C · s · L . m We have now obtained a bound on the expected norm of the first sum in equation (8). To control the norm of the second sum, we next define (14) Q2 = θkw θjw · ykw ⊗ yjw and we will apply decoupling inequality. Let (15) Q(cid:48) 2 = θkw θ(cid:48) jw · ykw ⊗ yjw (cid:88) (cid:88) j(cid:54)=k j(cid:54)=k m(cid:88) m(cid:88) w=1 w=1 where {θ(cid:48) kw} is an independent sequence with the same distribution as {θkw}. Then We will break up Q(cid:48) 2 into two terms and control the norm of each one separately. (16) Q(cid:48) 2 = θkw θ(cid:48) θkw θ(cid:48) jw · ykw ⊗ yjw. Denote the first term on the right by Q3 and the second term on the right by Q4. To bound (cid:107)Q4(cid:107)Γ, note that the random quantity Q4 has the same distribution as b(cid:88) b(cid:88) m(cid:88) j=1 k=1 w=1 2(cid:107)Γ. m(cid:88) w=1 E(cid:107)Q2(cid:107)Γ ≤ C12 · E(cid:107)Q(cid:48) jw · ykw ⊗ yjw − b(cid:88) b(cid:88) m(cid:88) k=1 Q(cid:48) 4 = kw · ukw ⊗ vkw, k=1 w=1 SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 17 and {kw} is an independent Bernoulli sequence. Since max{θkw, θ(cid:48) kw} ≤ B, we have and (cid:107)ukw(cid:107)∞ ≤ B√ m (cid:107)vkw(cid:107)∞ ≤ B√ m . Apply Lemma (2.3) with {θkw, θ(cid:48) E [(cid:107)Q(cid:48) kw} fixed, (cid:114) ≤ 4(cid:107)Γ {θkw},{θ(cid:48) C · s · L · B · m kw}] (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) k=1 m(cid:88) w=1 θkw2 · ykw ⊗ ykw (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 Γ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) k=1 m(cid:88) w=1 + θ(cid:48) kw2 · ykw ⊗ ykw  . (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 Γ Then, we use the law of double expectation and the Cauchy-Schwarz inequality, as in (12), to remove the conditioning : E(cid:2)(cid:107)Q(cid:48) 4(cid:107)Γ (cid:114) (cid:3) ≤ E · C · s · L  m (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) k=1 · B m(cid:88) w=1 θkw2 · ykw ⊗ ykw kw2 · ykw ⊗ ykw θ(cid:48) The two sequences of random variables {θkw} and {θ(cid:48) using Jensen inequality, we get kw} are identically distributed, so E [(cid:107)Q(cid:48) 4(cid:107)Γ] ≤ θkw2 · ykw ⊗ ykw To bound the expected value on the right-hand side, we note that 2  1/2 . (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:33)1/2 Γ . (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ Γ + k=1 w=1 (cid:32) m(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) m(cid:88)  (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1/2 (cid:0)1 − θkw2(cid:1) · ykw ⊗ ykw w=1 k=1 E Γ · B · θkw2 · ykw ⊗ ykw m C · s · L (cid:114) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) m(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) b(cid:88) (cid:32) m(cid:88) w=1 k=1 w=1 E k=1 E ≤ = (E(cid:107)Q1(cid:107)Γ + 1)1/2 . (cid:33)1/2 + 1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Γ Recall that from equation (13), if m ≥ C · s · L, then E(cid:107)Q1(cid:107)Γ is bounded. The random variables {θkw} are bounded by the constant B, so we can conclude there exist constants C13, C14 such that if m ≥ C13 · s · L, then (19) E(cid:107)Q4(cid:107)Γ ≤ C14 · (cid:114) s · L m . 18 ENRICO AU-YEUNG Recall that the right side of (16) has two terms, Q3 and Q4. expected norm of the other term, It remains to bound the Q3 = θkw θ(cid:48) jw · ykw ⊗ ykw. To bound E(cid:107)Q3(cid:107)Γ, note that Q3 has the same probability distribution as j=1 j=1 k=1 w=1 m(cid:88) b(cid:88) b(cid:88) b(cid:88) b(cid:88) m(cid:88) (cid:32) b(cid:88) m(cid:88) m(cid:88) b(cid:88) w w=1 w=1 k=1 k=1 w uw ⊗ vw, k=1 Q(cid:48) 3 = = = w θkw θ(cid:48) θkwykw jw · ykw ⊗ ykw (cid:33) (cid:32) b(cid:88) ⊗ θ(cid:48) jwyjw (cid:33) j=1 b(cid:88) j=1 θ(cid:48) jw yjw. w=1 where uw and vw are defined by (20) uw = θkw ykw and vw = The ykw have disjoint support for different values of k, so that Also note that m(cid:88) w=1 (cid:107)uw(cid:107)∞ ≤ B√ m , (cid:107)vw(cid:107)∞ ≤ B√ m . uw ⊗ vw = θkw θjw ykw ⊗ yjw, and so by comparing equation (7) with the above expression, we see that b(cid:88) j=1 uw ⊗ uw vw ⊗ vw k=1 w=1 m(cid:88) b(cid:88) m(cid:88) m(cid:88) (cid:114) w=1 w=1 C · s · L(cid:48) (cid:114) m C · s · L(cid:48) m ≤ C15 (cid:114) and are independent copies of A∗A. By Lemma (2.3) and Cauchy-Schwarz inequality, E(cid:107)Q3(cid:107)Γ = E(cid:107)Q(cid:48) 3(cid:107)Γ ≤ · B · (E(cid:107)A∗A(cid:107)Γ)1/2 · (E(cid:107)I + A∗A(cid:107)Γ + 1)1/2 where we have written L(cid:48) = log2(s) · log(m) · log(nb). Since L(cid:48) < L(s, n, m, b) ≡ log2 s · log(mb) · log(nb), we can conclude that E(cid:107)Q3(cid:107)Γ ≤ C15 s · L m · (E(cid:107)I + A∗A(cid:107)Γ + 1)1/2 . SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 19 To recapitulate, we have shown that E(cid:107)I − A∗A(cid:107)Γ ≤ E(cid:107)Q1(cid:107)Γ + E(cid:107)Q2(cid:107)Γ 2(cid:107)Γ ≤ E(cid:107)Q1(cid:107)Γ + C12 · E(cid:107)Q(cid:48) ≤ E(cid:107)Q1(cid:107)Γ + C12 (C15E(cid:107)Q(cid:48) 3(cid:107)Γ + C14E(cid:107)Q4(cid:107)Γ) . When m ≥ max(C10, C13) · s · L, we have established that · (E (cid:107)I − A∗A(cid:107)Γ + 1)1/2 E(cid:107)Q3(cid:107)Γ ≤ C15 · E(cid:107)Q4(cid:107)Γ ≤ C14 · E(cid:107)Q1(cid:107)Γ ≤ C11 · (cid:114) (cid:114) (cid:114) (cid:114) s · L m s · L m s · L m Therefore, in conclusion, we have proven that · E (cid:107)I − A∗A(cid:107)Γ ≤ C · s · L m By Lemma (2.7), with P = E(cid:107)I − A∗A(cid:107)Γ and c1 = that if m ≥ C6 · s · L, then we have (cid:113) E (cid:107)I − A∗A(cid:107)Γ + 1 (cid:19) . (cid:113) C·s·L (cid:18) 2 + (cid:114) E (cid:107)I − A∗A(cid:107)Γ ≤ C · s · L . m m , there exits constant C6 such Recall that by definition, L = log2 s · log(mb) · log(nb). This proves that the assertion (1) is true. It remains to prove that equation (1) implies that for any 0 < δs ≤ 1, we have provided that m ≥ C7 · δ−2 s E(cid:107)I − A∗A(cid:107)Γ ≤ δs · s · log4(nb) and m ≤ nb. (cid:115) If m ≤ nb, then log(mb) = log(m) + log(b) ≤ 2 log(nb). Hence, for s ≤ nb, C5 · s · log2(s) · log(mb) · log(nb) (21) The right-hand side is bounded by δs when m ≥ C7 · δ−2 m s This concludes the proof Theorem 1.2. 4C5 · s log4(nb) m · s · log4(nb). . (cid:115) ≤ 4. Proof of the tail bound (Theorem 1.3) Recall that I − A∗A = Q1 + Q2, where the expressions for Q1 and Q2 are given in equations (9) and (15). To obtain a bound for P ((cid:107)I − A∗A(cid:107)Γ > δ), we will first find a bound for P ((cid:107)Q1(cid:107)Γ > δ 1 from (10), and Y (cid:48) has the same probability distribution as Y from (11). For any β > 0 and for any λ > 0, (22) This equation holds because if (cid:107)Q1(cid:107)Γ > β + λ and if (cid:107)Q(cid:48) P ((cid:107)Q(cid:48) 2). Recall that Y = Q1 − Q(cid:48) 1(cid:107)Γ < β)P ((cid:107)Q1(cid:107)Γ > β + λ) ≤ P ((cid:107)Y (cid:107)Γ > λ). 1(cid:107)Γ < β, then 1(cid:107)Γ + (cid:107)Y (cid:107)Γ < β + (cid:107)Y (cid:107)Γ, β + λ < (cid:107)Q1(cid:107)Γ ≤ (cid:107)Q(cid:48) Since E((cid:107)Q1(cid:107)Γ) = E(cid:107)Q(cid:48) Let Vk,w = θk(w)(cid:48)2−θk(w)2· yk,w⊗ yk,w. Then (cid:107)Vk,w(cid:107) ≤ K · s By the Proposition 2.4, the concentration inequality gives us 1(cid:107)Γ, we obtain P ((cid:107)Q1(cid:107)Γ > 2E(cid:107)Q1(cid:107)Γ + λ) ≤ 2P ((cid:107)Y (cid:107)Γ > λ). (cid:20) (cid:18) (cid:21)(cid:19) (cid:107)Y (cid:48)(cid:107)Γ > C uE(cid:107)Y (cid:48)(cid:107)Γ + P ≤ e−u2 + e−t. Ks m t From (13), we have the bound (cid:114) C · s · KL (cid:34) (cid:114) m E(cid:107)Y (cid:48)(cid:107)Γ ≤ (cid:32) where m ≥ C · s · L · log(mb). (cid:35)(cid:33) Combining the last two inequalities, we see that m, where we define K = B2. 20 ENRICO AU-YEUNG and so (cid:107)Y (cid:107)Γ must be greater than λ. Note that the median of a positive random variable is never bigger than two times the mean, therefore 1(cid:107)Γ ≤ 2E(cid:107)Q(cid:48) 1(cid:107)Γ so that (22) becomes 1(cid:107)Γ)P ((cid:107)Q1(cid:107)Γ > 2E(cid:107)Q(cid:48) 1(cid:107)Γ + λ) ≤ P ((cid:107)Y (cid:107)Γ > λ). We can choose β = 2E(cid:107)Q(cid:48) 1(cid:107)Γ < 2E(cid:107)Q(cid:48) 1(cid:107)Γ) ≥ 1 2 P ((cid:107)Q(cid:48) P ((cid:107)Q(cid:48) . Fix a constant α, where 0 < α < 1/10. If we pick t = log(1/α) and u =(cid:112)log(1/α), then m + P u t ≤ e−u2 + e−t. (cid:107)Y (cid:48)(cid:107)Γ > C Ks m sKL the above equation becomes (cid:32) (cid:34)(cid:114) (cid:33)(cid:35) (cid:107)Y (cid:48)(cid:107)Γ > C P sKL log(1/α) m + sK m log(1/α) ≤ 2α. That means for some constant λ, we have P ((cid:107)Y (cid:48)(cid:107)Γ > λ) ≤ 2α. With this constant λ, we use the bound for E(cid:107)Q1(cid:107)Γ in (13) to conclude that sK m 2E(cid:107)Q1(cid:107)Γ + λ ≤ C sKL log(1/α) + log(1/α) sL log(mb) m m + (cid:32)(cid:114) (cid:114) (cid:33) . Inside the bracket on the right side, when we choose m so that all three terms are less than 1, the middle term will dominate the other two terms. That means there exists a constant C16 such that if m ≥ C16 · δ−2 s · s · K · L · log(1/α) then Combining the above inequalities, we obtain (cid:107)Q1(cid:107)Γ > P (cid:18) 2E(cid:107)Q1(cid:107)Γ + λ ≤ δs 2 . (cid:19) δs 2 ≤ 4α. SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 21 Next, we want to obtain a bound for (cid:107)Q2(cid:107)Γ. We saw that from (16), Q(cid:48) so to obtain a bound for (cid:107)Q2(cid:107)Γ, we will proceed to find bounds for (cid:107)Q3(cid:107)Γ and (cid:107)Q4(cid:107)Γ. 2 = Q3 + Q4, and Recall that Q4 has the same probability distribution as Q(cid:48) C · s · L · log(mb) E(cid:107)Q(cid:48) 4(cid:107) ≤ With the same uk,w and vk,w defined in (18), we have (cid:107)uk,w ⊗ vk,w(cid:107) ≤ K s m . Apply the concentration inequality (Proposition 2.4) with t = log(C/α) and u =(cid:112)log(C/α), m . 4 and we showed that (cid:114) (cid:32) we obtain P (cid:34)(cid:114) (cid:107)Q(cid:48) 4(cid:107)Γ > C s · KL · log(C/α) m s · K · log(C/α) m + ≤ 2α C . (cid:35)(cid:33) Inside the probability bracket, when we choose m so that both terms on the right side are less than 1, the first term will dominate the second term. That means there exists a constant C18 such that if m ≥ C18 · δ−2 s · s · K · L · log(1/α) then Since Q4 and Q(cid:48) 4 have the same probability distribution, we finally arrive at P ((cid:107)Q(cid:48) 4(cid:107)Γ > δ4 · C) ≤ 2α C17 . P ((cid:107)Q4(cid:107)Γ > δ4 · C) ≤ 2α C17 . To obtain a bound for (cid:107)Q3(cid:107), we will follow similar strategy. Recall that Q3 has the same probability distribution as Q(cid:48) 3 and we showed that E(cid:107)Q3(cid:107)Γ ≤ E ((cid:107)I − A∗A(cid:107)Γ + 1)1/2 CsL log(mb) . With the same uw and vw defined in (20), we have (cid:107)uw ⊗ vw(cid:107) ≤ K s m . By the concentration inequality (Proposition 2.4) with t = log(C/α) and u =(cid:112)log(C/α), m (cid:114) (cid:32) we obtain P (cid:34)(cid:114) (cid:107)Q3(cid:107)Γ > C s · L · log(mb) · log(C/α) m s · K · log(C/α) m + ≤ 2α C17 . (cid:35)(cid:33) That means there exists a constant C19 such that if m ≥ C19 · δ−2 s · s · KL · log(1/α) then To summarize, we have shown for any 0 < α < 1 m ≥ C20δ−2 · s · K · L · log(1/α), all the following inequalities hold: s P ((cid:107)Q3(cid:107)Γ > C · δs) ≤ 2α C17 10, there exists C20 such that when . (cid:18) (cid:19) (cid:107)Q1(cid:107)Γ > δs 2 ≤ 4α. (23) P 22 ENRICO AU-YEUNG (24) P ((cid:107)Q3(cid:107)Γ > C · δs) ≤ 2α C17 P ((cid:107)Q4(cid:107)Γ > δ4 · C) ≤ 2α (25) . C17 Finally, to find a tail bound on (cid:107)I − A∗A(cid:107)Γ, we note that . P ((cid:107)I − A∗A(cid:107)Γ > δs) ≤ P ((cid:107)Q1(cid:107)Γ > ≤ P ((cid:107)Q1(cid:107)Γ > ≤ P ((cid:107)Q1(cid:107)Γ > δs 2 δs 2 δs 2 ) + P ((cid:107)Q2(cid:107)Γ > δs ) 2 2(cid:107)Γ > ) + C12 · P ((cid:107)Q(cid:48) ) + C12 · P ((cid:107)Q3(cid:107)Γ > δs 2 δs 4 C12) C12) + C12 · P ((cid:107)Q4(cid:107)Γ > δs 4 C12) Combining this observation with inequalities (23), (24), (25), we can conclude that for any 0 < α < 1 10, we have P ((cid:107)I − A∗A(cid:107) > δs) ≤ 8α when m ≥ C20δ−2 · s· K · L· log(1/α). Recall that by definition, L = log2(s) log(mb) log(nb). Note that m ≤ nb, and hence equation (21) remains valid. To complete the proof, we s replace 8α with . This completes the proof of the theorem. References [1] Richard Baraniuk, Volkan Cevher, Marco F. Duarte, and Chinmay Hegde. Model-based compressive sensing. IEEE Trans. Inform. Theory, 56(4):1982–2001, 2010. [2] Richard Baraniuk, Mark Davenport, Ronald DeVore, and Michael Wakin. A simple proof of the restricted isometry property for random matrices. Constr. Approx., 28(3):253–263, 2008. [3] Ronen Basri and David W. Jacobs. Lambertian reflection and linear subspaces. IEEE Trans. Pattern Anal. Mach. Intell., 25(3):218–233, 2003. [4] Peter Belhumeur, J. Hespanha, and David Kriegman. Eigenfaces versus Fisherfaces: Recognition using class specific linear projection. IEEE Trans. Pattern Anal. Mach. Intell., 19(7):711–720, 1997. [5] Thomas Blumensath and Mike Davies. Iterative hard thresholding for compressed sensing. Appl. Com- put. Harmon. Anal., 27(3):265–74, 2009. [6] Alfred Bruckstein, David L. Donoho, and Michael Elad. From sparse solutions of systems of equations to sparse modeling of signals and images. SIAM Rev., 51(1):34–81, 2009. [7] Emmanuel Candes. The restricted isometry property and its implications for compressed sensing. C. R. Math. Acad. Sci. Paris, 346(9-10):589–592, 2008. [8] Emmanuel Candes, Justin Romberg, and Terence Tao. Robust uncertainty principles: exact signal reconstruction from highly incomplete frequency information. IEEE Trans. Inform. Theory, 52(2):489– 509, 2006. [9] Emmanuel Candes, Justin Romberg, and Terence Tao. Stable signal recovery from incomplete and inaccurate measurements. Comm. Pure Appl. Math., 59(8):1207–1223, 2006. [10] Albert Cohen, Wolfgang Dahmen, and Ronald DeVore. Compressed sensing and best k-term approx- imation. J. Amer. Math. Soc., 22(1):211–231, 2009. [11] David L. Donoho. Compressed sensing. IEEE Trans. Inform. Theory, 52(4):1289–1306, 2006. [12] David L. Donoho and Jared Tanner. Counting faces of randomly projected polytopes when the pro- jection radically lowers dimension. J. Amer. Math. Soc., 22(1):1–53, 2009. SPARSE RECONSTRUCTION WITH MULTIPLE WALSH MATRICES 23 [13] Simon Foucart. Sparse recovery algorithms: sufficient conditions in terms of restricted isometry con- stants. In Approximation Theory XIII: San Antonio 2010, volume 13 of Springer Proceedings in Math- ematics, pages 65–77. Springer, New York, 2012. [14] Simon Foucart and Holger Rauhut. A Mathematical Introduction to Compressive Sensing. Applied and Numerical Harmonic Analysis. Birkhauser, New York, 2013. [15] J. Haupt, W. U. Bajwa, G. Raz, S.J. Wright, and R. Nowak. Toeplitz-structured compressed sensing matrices. In Proc. 14th IEEE/SP Workshop Stat. Signal Process. (SSP'07), Madison, WI, Aug 2007, pages 294–298. [16] Felix Krahmer and Rachael Ward. New and improved Johnson-Lindenstrauss embeddings via the restricted isometry property. SIAM J. Math. Anal., 43(3):1269–1281, 2011. [17] Michel Ledoux and Michel Talagrand. Probability in Banach spaces. A Series of Modern Surveys in Mathematics. Springer-Verlag, Berlin, 1991. [18] Kezhi Li and Shuang Cong. State of the art and prospects of structured sensing matrices in compressed sensing. Front. Comput. Sci., 9(5):665–677, 2015. [19] Shahar Mendelson, Alain Pajor, and Nicole Tomczak-Jaegermann. Uniform uncertainty principle for Bernoulli and Subgaussian ensembles. Constr. Approx., 28(3):277–289, 2008. [20] Holger Rauhut. Random sampling of sparse trigonometric polynomials. Appl. Comput. Harmon. Anal., 22(1):16–42, 2007. [21] Holger Rauhut. Compressive sensing and structured random matrices. In Theoretical foundations and numerical methods for sparse recovery, Radon Ser. Comput. Appl. Math., 9. Walter de Gruyter, Berlin, 2010. [22] Holger Rauhut and Rachael Ward. Sparse Legendre expansions via l1-minimization. J. Approx. Theory, 164(5):517–533, 2012. [23] Mark Rudelson and Roman Vershynin. On sparse reconstruction from Fourier and Gaussian measure- ments. Comm. Pure Appl. Math., 61(8):1025–1045, 2008. [24] Michel Talagrand. Majorizing measures: the generic chaining. Ann. Probab., 24(3):1049–1103, 1996. [25] Joel A. Tropp. Greed is good: algorithmic results for sparse approximation. IEEE Trans. Inform. Theory, 50(10):2231–2242, 2004. [26] Joel A. Tropp. Just relax: convex programming methods for identifying sparse signals in noise. IEEE Trans. Inform. Theory, 52(3):1030–1051, 2006. [27] Joel A. Tropp, Jason N. Laska, Marco F. Duarte, Justin Romberg, and Richard Baraniuk. Beyond Nyquist: Efficient sampling of sparse bandlimited signals. IEEE Trans. Inform. Theory, 56(1):540, 2010. [28] John Wright, Allen Y. Yang, Arvind Ganesh, S. Shankar Sastry, and Yi Ma. Robust face recognition via sparse representation. IEEE Trans. Pattern Anal. Mach. Intell., 31(2):210–227, 2009. Enrico Au-Yeung, Department of Mathematical Sciences, DePaul University, Chicago E-mail address: [email protected]
1501.02493
2
1501
2016-05-03T18:03:02
Traces for Besov spaces on fractal h-sets and dichotomy results
[ "math.FA" ]
We study the existence of traces of Besov spaces on fractal $h$-sets $\Gamma$ with the special focus laid on necessary assumptions implying this existence, or, in other words, present criteria for the non-existence of traces. In that sense our paper can be regarded as an extension of [Br4] and a continuation of the recent paper [Ca2]. Closely connected with the problem of existence of traces is the notion of dichotomy in function spaces: We can prove that -- depending on the function space and the set $\Gamma$ -- there occurs an alternative: either the trace on $\Gamma$ exists, or smooth functions compactly supported outside $\Gamma$ are dense in the space. This notion was introduced by Triebel in [Tr7] for the special case of $d$-sets.
math.FA
math
Traces for Besov spaces on fractal h-sets and dichotomy results António M. Caetano Center for R&D in Mathematics and Applications, Department of Mathematics, University of Aveiro 3810-193 Aveiro, Portugal Email: [email protected] Dorothee D. Haroske Institute of Mathematics, Friedrich-Schiller-University Jena 07737 Jena, Germany Email: [email protected] 6 1 0 2 y a M 3 ] . A F h t a m [ 2 v 3 9 4 2 0 . 1 0 5 1 : v i X r a Abstract We study the existence of traces of Besov spaces on fractal h-sets Γ with the special focus laid on necessary assumptions implying this existence, or, in other words, present criteria for the non-existence of traces. In that sense our paper can be regarded as an extension of [Br4] and a continuation of the recent paper [Ca2]. Closely connected with the problem of existence of traces is the notion of dichotomy in function spaces: We can prove that -- depending on the function space and the set Γ -- there occurs an alternative: either the trace on Γ exists, or smooth functions compactly supported outside Γ are dense in the space. This notion was introduced by Triebel in [Tr7] for the special case of d-sets. 2010 Mathematics Subject Classification: Primary 46E35; Secondary 28A80. Key words and phrases: fractal h-sets, traces, Besov spaces of generalised smoothness, c(cid:13)2015. density of test functions, dichotomy. under CC-BY-NC-ND Licensed 4.0 license http://creativecommons.org/licenses/by-nc-nd/4.0/. Formal publication: http://dx.doi.org/10.4064/sm8171-1-2016. the 1 Traces for Besov spaces on fractal h-sets and dichotomy results 2 Introduction 1 The paper is devoted to a detailed study of traces of regular distributions taken on fractal sets. Such questions are of particular interest in view of boundary value problems of elliptic operators, where the solutions belong to some appropriate Besov (or Sobolev) space. One standard method is to start with assertions about traces on hyperplanes and then to transfer these results to bounded domains with sufficiently smooth boundary af- terwards. Further studies may concern compactness or regularity results, leading to the investigation of spectral properties. However, when it comes to irregular (or fractal) boundaries, one has to circumvent a lot of difficulties following that way, such that another method turned out to be more appro- priate. This approach was proposed by Edmunds and Triebel in connection with smooth boundaries first in [ET1] and then extended to fractal d-sets in [ET3, ET2, Tr4]. Later the setting of d-sets was extended to (d, Ψ)-sets by Moura [Mo2] and finally to the more general h-sets by Bricchi [Br2]. The idea is rather simple to describe, but the details are much more compli- cated: at first one determines the trace spaces of certain Besov (or Sobolev) spaces as precisely as possible, studies (compact) embeddings of such spaces into appropriate target spaces together with their entropy and approxima- tion numbers afterwards, and finally applies Carl's or Weyl's inequalities to link eigenvalues and entropy or approximation numbers. If one is in the lucky situation that, on one hand, one has atomic or wavelet decomposition results for the corresponding spaces, and, on the other hand, the irregular- ity of the fractal can be characterised by its local behaviour (within 'small' cubes or balls), then there is some chance to shift all the arguments to appro- priate sequence spaces which are usually easier to handle. This is one reason for us to stick to fractal h-sets and Besov spaces at the moment. But still the problem is not so simple and little is known so far. Dealing with spac! es on h-sets we refer to [CaL2,CaL1,KZ,Lo], and, probably closest to our ap- proach here, [Tr6, Chapter 8]. There it turns out that one first needs a sound knowledge about the existence and quality of the corresponding trace spaces. Returning to the first results in that respect in [Br2], see also [Br3,Br4,Br1], we found that the approach can (and should) be extended for later applica- tions. More precisely, for a positive continuous and non-decreasing function h : (0, 1] → R (a gauge function) with limr→0 h(r) = 0, a non-empty com- pact set Γ ⊂ Rn is called h-set if there exists a finite Radon measure µ in Rn with supp µ = Γ and µ(B(γ, r)) ∼ h(r), r ∈ (0, 1], γ ∈ Γ, see also [Ro, Chapter 2] and [Ma, p. 60]. In the special case h(r) = rd, 0 < d < n, Γ is called d-set (in the sense of [Tr4, Def. 3.1], see also [JW,Ma] -- be aware that this is different from [Fa]). Recall that some self-similar fractals are outstanding examples of d-sets; for instance, the usual (middle- third) Cantor set in R1 is a d-set for d = ln 2/ ln 3, and the Koch curve in Traces for Besov spaces on fractal h-sets and dichotomy results 3 ing that for 0 < p < ∞ we have in addition (cid:107)ϕ(cid:12)(cid:12)Γ R2 is a d-set for d = ln 4/ ln 3. The trace is defined by completion of pointwise traces of ϕ ∈ S(Rn), assum- p,q(Rn)(cid:107) for suitable parameters t ∈ R and 0 < q < ∞. In case of a compact d-set Γ, 0 < d < n, this results in Lp(Γ)(cid:107) (cid:46) (cid:107)ϕBt n−d p,q (Rn) = Lp(Γ) p if 0 < q ≤ min{p, 1} (1.1) trΓB and for s > n−d p , p,q(Rn) = Bs− n−d p,q p trΓBs (Γ), 1 p p,q(Rn) = Lp(Rn−1). p,q(Rn) which naturally extend Bs p,q(Rn) are the see [Tr4] with some later additions in [Tr5, Tr6]. Here Bs usual Besov spaces defined on Rn. In the classical case d = n − 1, 0 < p < ∞, 0 < q ≤ min{p, 1} this reproduces the well-known trace result trRn−1B In case of h-sets Γ one needs to consider Besov spaces of generalised smooth- p,q(Rn): instead of the smoothness ness Bσ parameter s ∈ R one now admits sequences σ = (σj)j∈N0 of positive numbers which satisfy σj ∼ σj+1, j ∈ N0. Such spaces are special cases of Bσ,N p,q (Rn) studied in [FaLe] recently, but they are known for a long time: apart from the interpolation approach (with a function parameter), see [Me,CoF], there is the rather abstract approach (approximation by series of entire ana- lytic functions and coverings) developed independently by Gol'dman and Kalyabin in the late 70's and early 80's of the last century; we refer to the survey [KL] and the appendix [Li] which cover the extensive (Russian) lit- erature at that time. We shall rely on the Fourier-analytical approach as presented in [FaLe]. It turns out that the classical smoothness s ∈ R has to be replaced by certain regularity indices s (σ), s (σ) of σ. In case of p,q(Rn) coincide and s (σ) = s (σ) = s σ = (2js)j the spaces Bσ in that case. Dealing with traces on h-sets Γ in a similar way as for d-sets, one obtains p,q(Rn) and Bs trΓBτ p,q(Rn) = Bσ p,q(Γ), where the sequence τ (representing smoothness) depends on σ, h (rep- resenting the geometry of Γ) and the underlying Rn; in particular, with h := (h(2−j))j, hp = , the counterpart of (1.1) reads as h(2−j) p 2j n p 1 (cid:16) (cid:17) j trΓBhp p,q (Rn) = Lp(Γ), 0 < p < ∞, 0 < q ≤ min{p, 1}. These results were already obtained in [Br4] under some additional restric- tions. In [Ca2] we studied sufficient conditions for the existence of such traces again (in the course of dealing with growth envelopes, characteris- ing some singularity behaviour) and return to the subject now to obtain 'necessary' conditions, or, more precisely, conditions for the non-existence of traces. This problem is closely connected with the so-called dichotomy: Traces for Besov spaces on fractal h-sets and dichotomy results 4 Triebel coined this notion in [Tr7] for, roughly speaking, the following al- ternative: the existence of a trace on Γ (by completion of pointwise traces) on the one hand, and the density of the set of smooth functions compactly supported outside Γ, denoted by D(Rn \ Γ), on the other hand. Though it is rather obvious that the density of D(Rn \ Γ) in some space prevents the existence of a properly defined trace, it is not clear (and, in fact,! not true in general) that there is some close connection vice versa. However, in some cases there appears an alternative that either we have an affirmative answer to the density question or traces exist. The criterion which case occurs nat- urally depends on the function spaces and the set Γ. Our main outcome in this respect, Theorem 3.18, establishes the following: if the h-set Γ satisfies, in addition, some porosity condition, σ is an admissible sequence and either 1 ≤ p < ∞, 0 < q < ∞, or 0 < q ≤ p < 1, then either Bσ p,q(Γ) = trΓBσhp or D(Rn \ Γ) p,q (Rn) exists is dense in Bσhp p,q (Rn) and, therefore, trΓBσhp p,q (Rn) cannot exist. This result is later reformulated in terms of the dichotomy introduced in [Tr7]. Note that there are further related approaches to trace and dichotomy questions in [Sch1, Sch2] for Besov spaces defined by differences, and in [Pi, Ha] referring to weighted settings. The paper is organised as follows. In Section 2 we collect some funda- mentals about h-sets and Besov spaces of generalised smoothness, including their atomic decomposition. In Section 3 we turn to trace questions with our main result being Theorem 3.18, before we finally deal with the di- chotomy and obtain Corollary 3.30. Throughout the paper we add remarks, discussions and examples to illustrate the (sometimes technically involved) arguments and results. 2 Preliminaries 2.1 General notation As usual, Rn denotes the n-dimensional real Euclidean space, N the collec- tion of all natural numbers and N0 = N ∪ {0}. We use the equivalence '∼' in ak ∼ bk or ϕ(x) ∼ ψ(x) always to mean that there are two positive numbers c1 and c2 such that c1 ak ≤ bk ≤ c2 ak or c1 ϕ(x) ≤ ψ(x) ≤ c2 ϕ(x) for all admitted values of the discrete variable k or the continuous variable x, where (ak)k, (bk)k are non-negative sequences and ϕ, ψ are non-negative functions. If only one of the inequalities above is meant, we use the symbol Traces for Besov spaces on fractal h-sets and dichotomy results 5 (cid:46) instead. Given two quasi-Banach spaces X and Y , we write X (cid:44)→ Y if X ⊂ Y and the natural embedding of X into Y is continuous. All unimportant positive constants will be denoted by c, occasionally with additional subscripts within the same formula. If not otherwise indicated, log is always taken with respect to base 2. For some κ ∈ R let (cid:98)κ(cid:99) = max{k ∈ Z : k ≤ κ} . κ+ = max{κ, 0} and (2.1) Moreover, for 0 < r ≤ ∞ the number r(cid:48) is given by 1 Lebesgue measure in the sequel. The notation · of an n-tuple in Nn for the corresponding infinity norm. For convenience, let both dx and · stand for the (n-dimensional) is also used for the size 0 and the Euclidean norm in Rn, while · ∞ is reserved r(cid:48) :=(cid:0)1 − 1 Given x ∈ Rn and r > 0, B(x, r) denotes the closed ball (cid:1) + . r (2.2) B(x, r) = {y ∈ Rn : y − x ≤ r} . 2.2 h-sets Γ A central concept for us is the theory of so-called h-sets and corresponding measures, we refer to a comprehensive treatment of this concept in [Ro]. Certainly one of the most prominent sub-class of these sets are the famous d- sets, see also Example 2.7 below, but it is also well-known that in many cases more general approaches are necessary, cf. [Ma, p. 60]. Here we essentially follow the presentation in [Br2,Br3,Br4,Br1], see also [Ma] for basic notions and concepts. Definition 2.1. (i) Let H denote the class of all positive continuous and h(r) = non-decreasing functions h : (0, 1] → R (gauge functions) with lim r→0 0. (ii) Let h ∈ H. A non-empty compact set Γ ⊂ Rn is called h-set if there exists a finite Radon measure µ in Rn with supp µ = Γ, µ(B(γ, r)) ∼ h(r), (2.3) (2.4) If for a given h ∈ H there exists an h-set Γ ⊂ Rn, we call h a measure function (in Rn) and any related measure µ with (2.3) and (2.4) will be called h-measure (related to Γ). r ∈ (0, 1], γ ∈ Γ. We quote some results on h-sets and give examples afterwards; we refer to the above-mentioned books and papers for proofs and a more detailed account on geometric properties of h-sets. In view of (ii) the question arises which h ∈ H are measure functions. We give a necessary condition first, see [Br2, Thm. 1.7.6]. Traces for Besov spaces on fractal h-sets and dichotomy results 6 Proposition 2.2. Let h ∈ H be a measure function. Then there exists some c > 0 such that for all j, k ∈ N0, (2.5) h(2−k−j) h(2−j) ≥ c 2−kn. Remark 2.3. Note that every h-set Γ satisfies the doubling condition, i.e. there is some c > 0 such that (2.6) µ(B(γ, 2r)) ≤ c µ(B(γ, r)), r ∈ (0, 1], γ ∈ Γ. Obviously one can regard (2.5) as a refined version of (2.6) for the function h, in which the dimension n of the underlying space Rn is taken into account (as expected). The complete characterisation for functions h ∈ H to be measure functions is given in [Br3]: There is a compact set Γ and a Radon measure µ with (2.3) and (2.4) if, and only if, there are constants 0 < c1 ≤ c2 < ∞ and a function h∗ ∈ H such that c1h∗(t) ≤ h(t) ≤ c2h∗(t), t ∈ (0, 1], and h∗(2−j) ≤ 2knh∗(2−k−j), (2.7) Proposition 2.4. Let Γ be an h-set in Rn. All h-measures µ related to Γ are equivalent to HhΓ, where the latter stands for the restriction to Γ of the generalised Hausdorff measure with respect to the gauge function h. for all j, k ∈ N0. Remark 2.5. A proof of this result is given in [Br2, Thm. 1.7.6]. Concerning the theory of generalised Hausdorff measures Hh we refer to [Ro, Chapter 2] and [Ma, p. 60]; in particular, if h(r) = rd, then Hh coincides with the usual d-dimensional Hausdorff measure. We recall a description of measure functions and explicate a few examples afterwards. Proposition 2.6. Let n ∈ N. (i) Let ξ : (0, 1] → [0, n] be a measurable function. Then the function (2.8) h(r) = exp is a measure function. ξ(s) ds s r ∈ (0, 1] (cid:110)− (cid:90) 1 r (cid:110)− (cid:90) 1 r (cid:111) , (cid:111) , (ii) Conversely, let h be a given measure function. Then for any ε > 0 there exists a measurable function ξ : (0, 1] → [−ε, n + ε] such that (2.9) h(r) ∼ exp ξ(s) ds s r ∈ (0, 1]. Traces for Besov spaces on fractal h-sets and dichotomy results 7 This version of the theorem is given in [Br4, Thm. 3.7]; it can also be identified as a special case of a result in [BGT, pp. 74]. Example 2.7. We restrict ourselves to a few examples only, but in view of Proposition 2.6 one can easily find further examples, we refer to [Br4, Ex. 3.8], too. All functions are defined for r ∈ (0, ε), suitably extended on (0, 1] afterwards. Let Ψ be a continuous admissible function or a continuous slowly varying function, respectively. An admissible function Ψ in the sense of [ET2,Mo2] is a positive monotone function on (0, 1] such that Ψ (2−2j) ∼ Ψ (2−j), j ∈ N. A positive and measurable function Ψ defined on the interval (0, 1] is said to be slowly varying (in Karamata's sense) if (2.10) lim t→0 Ψ(st) Ψ(t) = 1, s ∈ (0, 1]. c sδ ≤ Ψ(st) h(r) = rd Ψ(r), r ∈ (0, 1], x ∈ (0, 1], b ∈ R, Ψb(x) = (1 + log x)b , For such functions it is known, for instance, that for any δ > 0 there exists Ψ(t) ≤ c s−δ, for t, s ∈ (0, 1], and for each c = c(δ) > 1 such that 1 ε > 0 there is a non-increasing function φ and a non-decreasing function ϕ with t−ε Ψ(t) ∼ φ(t), and tε Ψ(t) ∼ ϕ(t); we refer to the monograph [BGT] for details and further properties; see also [Zy, Ch. V], [EKP], and [Ne1,Ne2]. In particular, (2.11) may be considered as a prototype both for an admissible function and a slowly varying function. Let 0 < d < n. Then (2.12) is a typical example for h ∈ H. The limiting cases d = 0 and d = n can be included, assuming additional properties of Ψ in view of (2.5) and h(r) → 0 for r → 0, e.g. (2.13) referring to (2.11). Later on we shall need to consider measure functions of this kind when b ∈ [−1, 0), so we would like to mention here that with this restriction the proof that the above is a measure function is a simple application of the characterisation given just before Proposition 2.4 (take h∗ = h and observe that, for b ∈ [−1, 0), (1 + x)b2xn ≥ 1 for x = 0 and for x ≥ 1, from which follows (2.7) taking x = k). ied in [ET2, Mo2], whereas the special setting Ψ ≡ 1 leads to (2.14) connected with the famous d-sets. Apart from (2.12) also functions of type h(r) = exp (b log rκ Such functions h given by (2.12) are related to so-called (d, Ψ)-sets stud- ), b < 0, 0 < κ < 1, are admitted. h(r) = (1 + log r)b, b < 0, r ∈ (0, 1], h(r) = rd, r ∈ (0, 1], 0 < d < n, Traces for Besov spaces on fractal h-sets and dichotomy results 8 We shall need another feature of h-sets, namely the so-called 'porosity' condition, see also [Ma, p.156] and [Tr5, Sects. 9.16-9.19]. Definition 2.8. A Borel set Γ (cid:54)= ∅ satisfies the porosity condition if there exists a number 0 < η < 1, such that for any ball B(γ, r) centered at γ ∈ Γ and with radius 0 < r ≤ 1, there is a ball B(x, ηr) centered at some x ∈ Rn satisfying (2.15) B(γ, r) ⊃ B(x, ηr), B(x, ηr) ∩ Γ = ∅. Replacing η by η 2, we can complement (2.15) by dist (cid:0)B(x, ηr), Γ(cid:1) ≥ ηr, 0 < r ≤ 1. (2.16) This definition coincides with [Tr4, Def. 18.10]. There is a complete char- acterisation for measure functions h such that the corresponding h-sets Γ satisfy the porosity condition; this can be found in [Tr5, Prop. 9.18]. We recall it for convenience. Proposition 2.9. Let Γ ⊂ Rn be an h-set. Then Γ satisfies the porosity condition if, and only if, there exist constants c > 0 and ε > 0 such that (2.17) ≥ c 2−(n−ε)k, j, k ∈ N0. h(cid:0)2−j−k(cid:1) h (2−j) Note that an h-set Γ satisfying the porosity condition has Lebesgue mea- sure Γ = 0, but the converse is not true. This can be seen from (2.17) and the result [Tr6, Prop. 1.153], (2.18) Γ = 0 if, and only if, lim r→0 rn h(r) = 0 for all h-sets Γ. Remark 2.10. In view of our above examples and (2.17) it is obvious that h from (2.12) and (2.14) with d = n does not satisfy the porosity condition, unlike in case of d < n. Let Lp(Ω), Ω ⊆ Rn, 0 < p ≤ ∞, stand for the usual quasi-Banach space of p-integrable (measurable, essentially bounded if p = ∞) functions with respect to the Lebesgue measure, quasi-normed by (cid:16)(cid:90) (cid:17)1/p , (cid:107)f Lp(Ω)(cid:107) := f (x)p dx with the obvious modification if p = ∞. Moreover, when Γ ⊂ Rn is an h-set in the sense of Definition 2.1, then we consider Lp(Γ) = Lp(Γ, µ) as the Ω Traces for Besov spaces on fractal h-sets and dichotomy results 9 usual quasi-Banach space of p-integrable (measurable, essentially bounded if p = ∞) functions on Γ with respect to the measure µ, quasi-normed by (cid:16)(cid:90) (cid:17)1/p (cid:107)f Lp(Γ)(cid:107) = f (γ)pµ( dγ) < ∞ for 0 < p < ∞, and Γ (cid:107)f L∞(Γ)(cid:107) = inf {s > 0 : µ ({γ ∈ Γ : f (γ) > s}) = 0} < ∞. In view of Proposition 2.4 all (possibly different) measures µ corresponding to h result in the same Lp(Γ) space. 2.3 Function spaces of generalised smoothness We deal with the concept of admissible sequences first. Definition 2.11. A sequence σ = (σj)j∈N0 of positive numbers is called admissible if there are two positive constants d0, d1 such that (2.19) d0 σj ≤ σj+1 ≤ d1 σj, j ∈ N0. σr := (cid:0)σr (cid:1) notation j j (a) :=(cid:0)2ja(cid:1) Remark 2.12. If σ and τ are admissible sequences, then στ := (σjτj)j and , r ∈ R, are admissible, too. For later use we introduce the where a ∈ R, j∈N0 (2.20) that is (a) = σ with σj = 2ja, j ∈ N0. Obviously, for a, b ∈ R, r > 0, and σ admissible, we have (a)(b) = (a + b), ( a r ) = (a)1/r, and (a)σ = (2jaσj)j∈N0 . Example 2.13. Let s ∈ R, Ψ be an admissible function in the sense of Example 2.7 above. Then σ = (2jsΨ (2−j))j is admissible, including, in particular, σ = (s), s ∈ R. We refer to [FaLe] for a more general approach and further examples. We introduce some 'regularity' indices for σ. Definition 2.14. Let σ be an admissible sequence, and (2.21) and (2.22) s (σ) := lim inf j→∞ log s (σ) := lim sup j→∞ log (cid:18) σj+1 (cid:18) σj+1 σj (cid:19) (cid:19) . σj Traces for Besov spaces on fractal h-sets and dichotomy results 10 Remark 2.15. These indices were introduced and used in [Br2]. For admissi- ble sequences σ according to (2.19) we have log d0 ≤ s (σ) ≤ s (σ) ≤ log d1. One easily verifies that in case of σ =(cid:0)2jsΨ(cid:0)2−j(cid:1)(cid:1) s (σ) = s (σ) = s (2.23) for all admissible functions Ψ and s ∈ R. On the other hand one can find examples in [FaLe], due to Kalyabin, showing that an admissible sequence has not necessarily a fixed main order. Moreover, it is known that for any 0 < a ≤ b < ∞, there is an admissible sequence σ with s (σ) = a and s (σ) = b, that is, with prescribed upper and lower indices. j For later use we fix some observations that are more or less immedi- ate consequences of the definitions (2.21), (2.22). Let σ, τ be admissible sequences. Then s (σ) = −s(cid:0)σ−1(cid:1) , s (σr) = r s (σ) , r ≥ 0, (2.24) and s (στ ) ≤ s (σ) + s (τ ) , (2.25) In particular, for σ = (a), a ∈ R, then (2.25) can be sharpened by (2.26) s (τ (a)) = a + s (τ ) , s (τ (a)) = a + s (τ ) . s (στ ) ≥ s (σ) + s (τ ) . Observe that, given ε > 0, there are two positive constants c1 = c1(ε) and c2 = c2(ε) such that c1 2(s(σ)−ε)j ≤ σj ≤ c2 2(s(σ)+ε)j, (2.27) Plainly this implies that whenever s (σ) > 0, then σ−1 belongs to any space (cid:96)u, 0 < u ≤ ∞, whereas s (σ) < 0 leads to σ−1 (cid:54)∈ (cid:96)∞. Remark 2.16. Note that in some later papers, cf. [Br1], instead of (2.21) and (2.22) the so-called upper and lower Boyd indices of σ are considered, given by j ∈ N0. (cid:18) (cid:18) (2.28) and (2.29) ασ = lim j→∞ βσ = lim j→∞ 1 j 1 j log sup k∈N0 σj+k σk log inf k∈N0 σj+k σk (cid:19) (cid:19) (cid:18) (cid:18) = inf j∈N = sup j∈N 1 j 1 j log sup k∈N0 log inf k∈N0 (cid:19) (cid:19) σj+k σk σj+k σk respectively. In general we have s (σ) ≤ βσ ≤ ασ ≤ s (σ) , but one can construct admissible sequences with s (σ) < βσ and ασ < s (σ). Traces for Besov spaces on fractal h-sets and dichotomy results 11 We want to introduce function spaces of generalised smoothness and need to recall some notation. By S(Rn) we denote the Schwartz space of all complex-valued, infinitely differentiable and rapidly decreasing functions on Rn and by S(cid:48)(Rn) the dual space of all tempered distributions on Rn. If ϕ ∈ S(Rn), then (2.30) (cid:98)ϕ(ξ) ≡ (Fϕ)(ξ) := (2π)−n/2 e−ixξϕ(x) dx, ξ ∈ Rn, (cid:90) Rn denotes the Fourier transform of ϕ. As usual, F−1ϕ or ϕ∨ stands for the inverse Fourier transform, given by the right-hand side of (2.30) with i in place of −i. Here xξ denotes the scalar product in Rn. Both F and F−1 are extended to S(cid:48)(Rn) in the standard way. Let ϕ0 ∈ S(Rn) be such that x ≤ 1 and supp ϕ0 ⊂ {x ∈ Rn : x ≤ 2}, (2.31) ϕ0(x) = 1 and for each j ∈ N let if ϕj(x) := ϕ0(2−jx) − ϕ0(2−j+1x), x ∈ Rn. (2.32) Then the sequence (ϕj)∞ Definition 2.17. Let σ be an admissible sequence, 0 < p, q ≤ ∞, and (ϕj)∞ j=0 a smooth dyadic resolution of unity (in the sense described above). Then (2.33) j=0 forms a smooth dyadic resolution of unity. (cid:18) ∞(cid:88) (cid:13)(cid:13)F−1ϕjFfLp(Rn)(cid:13)(cid:13)q σq j (cid:19)1/q (cid:41) < ∞ p,q(Rn) = Bσ f ∈ S(cid:48)(Rn) : (cid:40) j=0 (with the usual modification if q = ∞). Remark 2.18. These spaces are quasi-Banach spaces, independent of the chosen resolution of unity, and S(Rn) is dense in Bσ p,q(Rn) when p < ∞ and q < ∞. Taking σ = (2js)j , s ∈ R, we obtain the classical Besov spaces p,q(Rn), whereas σ = (2jsΨ (2−j))j , s ∈ R, Ψ an admissible function, Bs (Rn), studied in [Mo1, Mo2] in detail. Moreover, the leads to spaces B(s,Ψ) p,q(Rn) are special cases of the more general approach inves- above spaces Bσ p,q(Rn) we refer to the series of tigated in [FaLe]. For the theory of spaces Bs monographs [Tr2, Tr3, Tr4, Tr5, Tr6]. p,q We recall the atomic characterisation of spaces Bσ p,q(Rn), for later use. Let Zn stand for the lattice of all points in Rn with integer-valued components, Qjm denote a cube in Rn with sides parallel to the axes of coordinates, centered at 2−jm = (2−jm1, . . . , 2−jmn), and with side length 2−j, where m ∈ Zn and j ∈ N0. If Q is a cube in Rn with sides parallel to the axes of coordinates and r > 0, then rQ is the cube in Rn concentric with Q, with sides parallel to the sides of Q and r times the length. Traces for Besov spaces on fractal h-sets and dichotomy results 12 Definition 2.19. Let K ∈ N0 and b > 1. (i) A K times differentiable complex-valued function a(x) in Rn (continuous if K = 0) is called an 1K-atom if supp a ⊂ b Q0m, (2.34) and Dαa(x) ≤ 1, for some m ∈ Zn, for α ≤ K. (ii) Let L+1 ∈ N0, and σ be admissible. A K times differentiable complex- valued function a(x) in Rn (continuous if K = 0) is called an (σ, p)K,L- atom if for some j ∈ N0, (2.35) (2.36) for some m ∈ Zn, for supp a ⊂ b Qjm, p +α) Dαa(x) ≤ σ−1 α ≤ K, x ∈ Rn, 2j( n j and (2.37) (cid:90) Rn xβa(x) dx = 0, if β ≤ L. We adopt the usual convention to denote atoms located at Qjm (which means (2.34) or (2.35)) by ajm, j ∈ N0, m ∈ Zn. For sequences λ = (λjm)j∈N0, m∈Zn of complex numbers the Besov sequence spaces bp,q, 0 < p, q ≤ ∞, are given by (2.38) λ ∈ bp,q if, and only if, (cid:107)λbp,q(cid:107) = λjmp < ∞  ∞(cid:88) j=0 (cid:32)(cid:88) m∈Zn (cid:33)q/p1/q (with the usual modification if p = ∞ or q = ∞). The atomic decomposition p,q(Rn) reads as follows, see [FaLe, Thm. 4.4.3, Rem. 4.4.8]. theorem for Bσ Proposition 2.20. Let σ be admissible, b > 1, 0 < p, q ≤ ∞, K ∈ N0 and L + 1 ∈ N0 with (2.39) be fixed. Then f ∈ S(cid:48)(Rn) belongs to Bσ represented as p,q(Rn) if, and only if, it can be L > −1 + n (cid:18) 1 − s (σ) K > s (σ) , (cid:19) − 1 and p + (2.40) f = λjm ajm(x), convergence being in S(cid:48)(Rn), ∞(cid:88) (cid:88) j=0 m∈Zn where λ ∈ bp,q and ajm(x) are 1K-atoms (j = 0) or (σ, p)K,L-atoms (j ∈ N) according to Definition 2.19. Furthermore, inf (cid:107)λ bp,q(cid:107), where the infimum is taken over all admissible representations (2.40), is an equivalent quasi-norm in Bσ p,q(Rn). Traces for Besov spaces on fractal h-sets and dichotomy results 13 Remark 2.21. For later use it is useful to remark that, for fixed j ∈ N0, the family {Qjm : m ∈ Zn} constitutes a tessellation of Rn. We shall call it a tessellation of step 2−j and denote each cube of it by a (corresponding) grid cube. Remark 2.22. The following shall also be of use later on: given the families of tessellations as above (for any possible j ∈ N0), clearly there exists, for each j ∈ N0, a partition of unity {ϕjm : m ∈ Zn} of Rn by functions ϕjm supported on 3 there exists cγ > 0 independent of j such that 2Qjm and such that, for each γ ∈ Nn 0, Dγϕjm(x) ≤ cγ2jγ, (2.41) We shall call it a partition of unity of step 2−j. x ∈ Rn, m ∈ Zn. 3 Besov spaces on Γ Let Γ be some h-set, h ∈ H. We follow [Br2] and use the abbreviation (3.1) for the sequence connected with h ∈ H. h := (hj)j∈N0 with hj := h(2−j), j ∈ N0, 3.1 Trace spaces Bσ Recall that Lp(Γ) = Lp(Γ, µ), where µ ∼ HhΓ is related to the h-set Γ. Suppose there exists some c > 0 such that for all ϕ ∈ S(Rn), p,q(Γ) (cid:13)(cid:13)ϕΓ Lp(Γ)(cid:13)(cid:13) ≤ c (cid:13)(cid:13)ϕBτ p,q(Rn)(cid:13)(cid:13) , (3.2) where the restriction on Γ is taken pointwisely. By the density of S(Rn) in p,q(Rn) for p, q < ∞ and the completeness of Lp(Γ) one can thus define Bτ for f ∈ Bτ p,q(Rn) its trace trΓf on Γ by completion of pointwise restrictions. This was the approach followed in [Br2, Def. 3.3.5], conjugating general embedding results for Besov spaces on Rn (as seen for example in [CaF, Thm. 3.7] or [Br2, Prop. 2.2.16]) with the fact that (3.2) above holds for τ = h1/p(n)1/p, 0 < p < ∞, 0 < q ≤ min{p, 1} (cf. [Br2, Thm. 3.3.1(i)]. Then we have, following [Br2, Def. 3.3.5], that for 0 < p, q < ∞ and σ admissible with s (σ) > 0, Besov spaces Bσ p,q(Γ) on Γ are defined as (3.3) Bσ p,q(Γ) := trΓBσh1/p(n)1/p p,q (Rn), Traces for Besov spaces on fractal h-sets and dichotomy results 14 more precisely, Bσ p,q(Γ) := (3.4) f ∈ Lp(Γ) : ∃ g ∈ Bσh1/p(n)1/p (cid:110) (cid:110)(cid:13)(cid:13)(cid:13)gBσh1/p(n)1/p p,q equipped with the quasi-norm (3.5)(cid:13)(cid:13)fBσ p,q(Γ)(cid:13)(cid:13) = inf (Rn) p,q (Rn), trΓg = f (cid:111) (cid:13)(cid:13)(cid:13) : trΓg = f, g ∈ Bσh1/p(n)1/p p,q , (cid:111) (Rn) . This was extended in the following way in [Ca2, Def. 2.7]: Definition 3.1. Let 0 < p, q < ∞, σ be an admissible sequence, and Γ be an h-set. Assume that (i) in case of p ≥ 1 or q ≤ p < 1, (3.6) σ−1 ∈ (cid:96)q(cid:48), (ii) in case of 0 < p < 1 and p < q, (3.7) σ−1h 1 r − 1 p ∈ (cid:96)vr for some r ∈ [p, min{q, 1}] and 1 vr = 1 r − 1 q , p,q p,q(Γ) as in (3.3), (3.4), again is satisfied. Then (it makes sense to) define Bσ with the quasi-norm given by (3.5). Remark 3.2. The reasonability of declaring that smoothness (1) (that is, 0, in classical notation) on Γ corresponds to smoothness h1/p(n)1/p on Rn -- as is implicit in (3.3) -- comes from the fact that, at least when Γ also satisfies (Rn) = Lp(Γ) when the porosity condition, we have indeed that trΓBh1/p(n)1/p 0 < p < ∞, 0 < q ≤ min{p, 1} -- cf. [Br2, Thm. 3.3.1(ii)]. Remark 3.3. Both in [Br2] and in [Ca2] the definition of Besov spaces on Γ also covers the cases when p or q can be ∞, following some convenient modifications of the approach given above. However, in view of the main results to be presented in this paper, the restriction to 0 < p, q < ∞ is natural for us here, as will be apparent in the following subsection. Remark 3.4. We briefly compare the different assumptions in [Br2, Def. 3.3.5] and in Definition 3.1 above. Due to the observation made immediately af- ter (2.27), s (σ) > 0 implies σ−1 ∈ (cid:96)v for arbitrary v, i.e. (3.6) and (3.7) κ, with r = p. The converse, however, is not true, take e.g. σj = (1 + j) κ ∈ R, then s (σ) = 0, but σ−1 ∈ (cid:96)q(cid:48) for κ > 1 q(cid:48) , corresponding to (3.6). As for (3.7), say with p = r, for the same σ given above we have also that σ−1 ∈ (cid:96)vp for κ > 1 q , but still s (σ) = 0. So the above definition is in fact a proper extension of the one considered in [Br2] and, as we shall see, will (at least in some cases) be indeed the largest possible extension. Remark 3.5. The definition above applies in particular when Γ is a d-set and σ = (s) with s > 0. In simpler notation, we can write in this case that p − 1 (3.8) This coincides with [Tr4, Def. 20.2]. Bs p,q(Γ) = trΓB s+ n−d p,q p (Rn). Traces for Besov spaces on fractal h-sets and dichotomy results 15 3.2 Criteria for non-existence of trace spaces Here we shall try to get necessary conditions for the existence of the trace. To this end, we explore the point of view that the trace cannot exist in the sense used in Definition 3.1 when D(Rn \ Γ), the set of test functions with (Rn). compact support outside Γ, is dense in Bσh1/p(n)1/p p,q(Rn). Let ϕ ∈ C∞ In fact, assume that D(Rn \ Γ) is dense in Bτ with ϕ ≡ 1 on a neighbourhood of Γ. Clearly, ϕ ∈ Bτ exists a sequence (ψk)k ⊂ D(Rn \ Γ) with (3.9) 0 (Rn) p,q(Rn). Then there p,q(Rn)(cid:13)(cid:13) −−−→ (cid:13)(cid:13)ϕ − ψkBτ p,q k→∞ 0. If the trace were to exist in the sense explained before, this would imply (3.10) 0 = ψkΓ = trΓψk −−−→ k→∞ trΓϕ = ϕΓ = 1 in Lp(Γ), which is a contradiction. Discussion 3.6. So in order to disprove the existence of the trace in certain cases it is sufficient to show the density of D(Rn\Γ) in Bτ p,q(Rn). Further, we may restrict ourselves to functions ϕ ∈ C∞ 0 (Rn) because of their density in p,q(Rn), 0 < p, q < ∞, and approximate them by functions ψk ∈ D(Rn\Γ). Bτ We shall construct such ψk based on finite sums of the type (cid:88) r∈Ik λrϕr, where Ik is some finite index set, λr ∈ C and ϕr are compactly supported smooth functions with (3.11) λrϕrϕ = ϕ (cid:88) r∈Ik on a neighbourhood (depending on k) of Γ, ϕ as above, and (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) −−−→ k→∞ 0. λrϕrϕBτ p,q(Rn) (3.12) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:88) Plainly, then ψk := ϕ −(cid:80) p,q(Rn)(cid:13)(cid:13) = (cid:13)(cid:13)ϕ − ψkBτ r∈Ik r∈Ik λrϕrϕ ∈ D(Rn \ Γ), k ∈ N, λrϕrϕBτ p,q(Rn) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:88) r∈Ik (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) −−−→ k→∞ 0, and therefore the required density is proved. The limit in (3.12) above is going to be computed with the help of atomic representations for the functions considered. In order to explain how this will be done, we need first to consider a preliminary result which connects the Definition 2.1 of h-sets with the structure of the atomic representation theorem given in Proposition 2.20. Traces for Besov spaces on fractal h-sets and dichotomy results 16 Definition 3.7. Given a set Γ ⊂ Rn and r > 0, we shall denote by Γr the neighbourhood of radius r of Γ. That is, Γr := {x ∈ Rn : dist(x, Γ) < r}. Lemma 3.8. Let Γ be an h-set in Rn with µ a corresponding Radon mea- sure. Let j ∈ N. There is a cover {Q(j, i)}Nj (i) Nj ∼ h−1 (ii) Q(j, i) are cubes in Rn of side length ∼ 2−j, (iii) each Q(j, i) is, in fact, the union of ∼ 1 grid cubes of a tessellation of i=1 of Γ2−j such that , j Rn of step 2−j in accordance with Remark 2.21, (iv) each Q(j, i) contains a ball of radius ∼ 2−j centred at a point of Γ such that any pair of them, for different values of i, are disjoint. Denote by Q(j) the family of all grid cubes obtained in this way (that is, Q(j) = {Qjm : Qjm ⊂ Q(j, i) for some i}) and consider the correspond- ing functions of a related partition of unity {ϕjm} of step 2−j of Rn in accordance with Remark 2.22. That is, consider only the functions of that partition which are supported on 3 2Qjm such that Qjm ∈ Q(j). The cover of Γ2−j considered above can be chosen in such a way that the family {ϕjm : Qjm ∈ Q(j)} is a partition of unity of Γ2−j. All equivalence constants in the estimates above are independent of j and i. Proof. For each j ∈ N, start by considering an optimal cover of Γ in the sense of [Br2, Lemma 1.8.3], that is, a cover by balls B(γi, 2−j−1) centred ) = ∅ for i (cid:54)= k. As at points γi ∈ Γ and such that B(γi, 2−j−1 was pointed out in [Br2, Lemma 1.8.3], the number N2−j−1 of balls in such a cover satisfy ) ∩ B(γk, 2−j−1 3 3 N2−j−1 ∼ h−1 j+1. (3.13) It is not difficult to see that each B(γi, 2−j−1) is contained in the union of 2n grid cubes of the type Qjm which together form a cube of side length 2−j+1 and which we shall provisionally denote by Q(j, i). Then assertions (ii)-(iv) of the lemma are clear. As to (i), it is also clear, using (3.13) and Proposition 2.2, that Nj (cid:46) h−1 . The reverse inequality follows from the fact that there can only be ∼ 1 different B(γk, 2−j−1) giving rise to the same Q(j, i), because each one of those produces also a ball B(γk, 2−j−1 ) which is disjoint from all the other balls obtained in a similar way. Eliminating the repetitions among the previously considered Q(j, i) and redefining the i accordingly, we get a possible cover of Γ satisfying (i)-(iv) of the lemma. 3 j Traces for Besov spaces on fractal h-sets and dichotomy results 17 In order to get from here a corresponding cover of Γ2−j, we just need to enlarge each one of the previous Q(j, i) by adding to it all the 4n − 2n surrounding grid cubes which touch it. It is clear that the new cubes Q(j, i) obtained in this way also satisfy (i)-(iv) above. Finally, if we further enlarge the preceding Q(j, i) by adding to it all the 6n − 4n surrounding grid cubes which touch it, clearly we do not destroy properties (i)-(iv) for the new cubes Q(j, i) obtained in this way, this cover of Γ2−j satisfying also the property that {ϕjm : Qjm ∈ Q(j)} is a partition of unity of Γ2−j. Remark 3.9. It follows from the construction above that Q(j) is also a cover of Γ2−j with (cid:93)Q(j) ∼ h−1 (though a property like (iv) above cannot be guaranteed for each grid cube in Q(j)). Remark 3.10. We shall use the expression "optimal cover of Γ2−j" when referring to a cover of Γ2−j satisfying all the requirements of the lemma above. j We return now to the question of calculating the limit (3.12) in Discus- Ik = {(j, m) ∈ N × Zn : j ∈ Jk, m ∈ Mj} , sion 3.6. Discussion 3.11. The index set Ik, k ∈ N, will have the structure (3.14) where Jk and Mj are appropriately chosen finite subsets of N and Zn re- spectively. In any case, for each (j, m) ∈ Ik the relation dist(Qjm, Γ) (cid:46) 2−j should be satisfied. With r = (j, m) ∈ Ik, we shall also require that (3.15) for some constant b > 1, and, for arbitrary fixed K ∈ N, supp ϕr ⊂ b Qjm, x ∈ Rn, Dγϕr(x) ≤ cK2jγ, 0 with γ ≤ K. (3.16) We claim that then (apart from constants) the functions a(j,m) = τ−1 j 2j n are (τ , p)K,−1-atoms (no moment conditions) located at Qjm: the support condition is obvious; as for the derivatives we calculate for α ∈ Nn 0, α ≤ K, that γ ∈ Nn p ϕ(j,m)ϕ (cid:19) (cid:18)α p(cid:88) K2jα(cid:13)(cid:13)ϕC K(Rn)(cid:13)(cid:13) γ Dγϕ(j,m)(x)Dα−γϕ(x) j 2j n Dαa(j,m)(x) ≤ τ−1 γ≤α ≤ τ−1 p c(cid:48) j 2j n ≤ CK,ϕτ−1 j 2j( n fitting the needs of Definition 2.19. p +α), Traces for Besov spaces on fractal h-sets and dichotomy results 18 Since (cid:88) r∈Ik λrϕrϕ = (cid:88) (j,m)∈Ik (λ(j,m)CK,ϕτj2 −j n p )(C−1 K,ϕa(j,m)), to the effect of obtaining (3.12) we can then estimate from above the quasi- norm of this function by applying Proposition 2.20, in case the parameters considered do not require moment conditions. If moment conditions are required, in the presence of porosity of the h-set Γ a standard procedure can be applied in order to identify the above sum at least in a smaller neighbourhood of Γ (smaller than the one considered apropos of (3.11)) with an atomic representation (cid:88) (j,m)∈Ik (c−1 2 λ(j,m)CK,ϕτj2 −j n p )(c2C−1 K,ϕa(j,m)) in this case we replace, in Discussion 3.6,(cid:80) with the appropriate moment conditions and, as is apparent, producing essentially the same upper estimate by application of Proposition 2.20. So, λrϕrϕ by the above sum, keeping a property like (3.11). In any case, following Proposition 2.20, the quasi-norms of both sums in Bτ p,q(Rn) are estimated by r∈Ik (3.17) (cid:46) λ(j,m)τj2 pp −j n (cid:88) j∈Jk (cid:88) m∈Mj q/p1/q . The standard procedure referred to above, leading to the creation of atoms with appropriate moment conditions, though still coinciding with the former atoms in a somewhat smaller neighbourhood of a set satisfying the porosity condition, comes from [TW] and was, for example, used in [Ca1] and, more recently, in [Tr7]. It is quite technical, but we give here a brief description, for the convenience of the reader: For each Qjm as above, namely with dist(Qjm, Γ) (cid:46) 2−j, fix an element of Γ, xj,m say, at a distance (cid:46) 2−j of Qjm. Clearly, there is a constant c1 > 0 such that Qjm ⊂ B(xj,m, c12−j). Without loss of generality, we can assume that 0 < c12−j < 1, so that, from the fact Γ satisfies the porosity condition given in Definition 2.8, there exists yj,m ∈ Rn such that B(yj,m, ηc12−j) ⊂ B(xj,m, c12−j) and B(yj,m, ηc12−j) ∩ Γ = ∅, where 0 < η < 1 is as in Definition 2.8. Obviously, as mentioned in (2.16), we can also say that dist(B(yj,m, ηc12−j−1), Γ) ≥ ηc12−j−1. γ ∈ Nn ◦ B(0, 1) and satisfying the property Fix a natural L as in (2.39) -- with τ instead of σ -- and let ψγ, with 0 and γ ≤ L, be C∞-functions ψγ with support in the open ball ∀β, γ ∈ Nn 0 with β,γ ≤ L, xβψγ(x) dx = δβγ, (cid:90) Rn Traces for Besov spaces on fractal h-sets and dichotomy results 19 where δβγ stands for the Kronecker symbol (for a proof of the existence of such functions, see [TW, p. 665]). Define, for each j, m as above, γ ≡ dj,m xγa(j,m)(ηc12−j−1x + yj,m) dx, γ ∈ Nn 0 with γ ≤ L, γ ψγ((ηc1)−12j+1(z − yj,m)), z ∈ Rn. dj,m and a(j,m)(z) = a(j,m)(z) − (cid:88) (cid:90) γ≤L It is easy to see that (cid:90) Rn (cid:90) Rn ((ηc1)−12j+1(z − yj,m))βa(j,m)(z) dz = 0, β ∈ Nn 0 with β ≤ L, Rn and, consequently (by Newton's binomial formula), zβa(j,m)(z) dz = 0, β ∈ Nn 0 with β ≤ L. This means that each a(j,m) has the required moment conditions for the p,q(Rn). Actually, it atoms in the atomic representations of functions of Bτ is not difficult to see that there exists a positive constant c2 such that c2C−1 K,ϕa(j,m) is a (τ, p)K,L-atom located at Qjm. Since, from the hypotheses and choices that have been made, a(j,m) = a(j,m) on Γηc12−j−1, we are through. Proposition 3.12. Let Γ be an h-set satisfying the porosity condition, 0 < p, q < ∞ and τ an admissible sequence. Assume that (3.18) Then D(Rn \ Γ) is dense in Bτ τ −1h1/p(n)1/p (cid:54)∈ (cid:96)∞. p,q(Rn), therefore p,q(Rn) cannot exist. trΓBτ (3.19) Proof. By standard reasoning we conclude from (3.18) that there exists a divergent subsequence of τ −1h1/p(n)1/p. More precisely, there is a strictly increasing sequence (jk)k∈N ⊂ N such that −jk 2 (3.20) For each k ∈ N, consider an optimal cover of Γ2−jk , in the sense of Remark 3.10, and follow our Discussions 3.6 and 3.11 with Ik from (3.14) given by lim k→∞ τjk h p = 0. n − 1 jk p Ik = {(j, m) ∈ N × Zn : j = jk, Qjm ∈ Q(j)} Traces for Besov spaces on fractal h-sets and dichotomy results 20 (that is, Jk = {jk} and Mj = Mjk = {m ∈ Zn : Qjkm ∈ Q(jk)}), ϕ(j,m) = ϕjm and λ(j,m) = 1. Due to Remark 2.22, this fits nicely into Discussions 3.6 and 3.11, in particular (3.11), (3.15) and (3.16) hold. It then follows, specially from (3.17), that the quasi-norm of the sum in (3.12) (or of an alternative similar sum, as discussed in Discussion 3.11 apropos of the consideration of moment conditions) is  (cid:88) (cid:46) Qjkm∈Q(jk)  1 p 2−jkn τ p jk −jk = τjk 2 n p ((cid:93)Q(jk)) 1 p ∼ τjk 2 −jk n p h − 1 jk p −−−→ k→∞ 0, where we have also used Remark 3.9 and (3.20). Thus (3.12) holds (possi- bly with an alternative similar sum if moment conditions are required, as mentioned above) and Discussion 3.6 then concludes the proof. Remark 3.13. The porosity condition was only used in the proof above to guarantee that atoms with appropriate moment conditions can be consid- ered. Thus the proposition holds without assuming porosity of Γ as long as p,q(Rn) L can be taken equal to -1 in the atomic representation theorem for Bτ (cf. Proposition 2.20 with τ instead of σ). The next result shows that when p < q we can conclude as in the pre- ceding proposition with an hypothesis weaker than (3.18). Proposition 3.14. Let Γ be an h-set satisfying the porosity condition, 0 < p < q < ∞, τ admissible. Assume that (3.21) Then D(Rn \ Γ) is dense in Bτ lim sup τ −1h1/p(n)1/p > 0. p,q(Rn), therefore p,q(Rn) trΓBτ cannot exist. (3.22) Proof. Denote σ := τ h−1/p(n)−1/p. From (3.21) it follows the existence of a constant c > 0 and a strictly increasing sequence (j(cid:96))(cid:96)∈N ⊂ N such that (3.23) (cid:96) ∈ N. ≥ c, σ−1 j(cid:96) Given k ∈ N, let k2 ∈ N be such that (3.24) 1 k + 1 k + 1 . . . + ≥ 2, 1 k2 which clearly exists. Then choose j(k) large enough in order that it is pos- sible to consider an optimal cover of Γ2−j(k) in the sense of Remark 3.10 2 Nj(k) ≥ 2. Again, this is clearly with a number Nj(k) of cubes such that k−1 possible, because of Nj(k) ∼ h−1 j(k) (cf. Lemma 3.8) and properties of gauge Traces for Besov spaces on fractal h-sets and dichotomy results 21 functions (cf. Definition 2.1). Actually, we shall also require that j(k) be chosen in such a way it coincides with one of the j(cid:96)'s above, which is also possible. 2i−1Nj(k) and Hence, for i = k, k + 1, . . . , k2,(cid:4)i−1Nj(k) therefore, using (3.24), (cid:5) > i−1Nj(k) − 1 ≥ 1 (cid:5) ≥ Nj(k). (cid:5) ≥ Nj(k). (cid:5) + . . . +(cid:4)k−1 (cid:5) + . . . +(cid:4)k−1 2 Nj(k) 1 Nj(k) Let k1 ∈ N be the smallest number such that (cid:4)k−1Nj(k) (cid:4)k−1Nj(k) (cid:5) +(cid:4)(k + 1)−1Nj(k) (cid:5) +(cid:4)(k + 1)−1Nj(k) (cid:88) (3.25) (3.26) On the other hand, following Lemma 3.8 and Discussion 3.6, ϕj(k)mϕ = ϕ on Γ2−j(k). Qj(k)m∈Q(j(k)) Now we associate the terms of this sum in the following way: first consider the m's such that Qj(k)m are grid subcubes of the first(cid:4)k−1Nj(k) subcubes of the following(cid:4)(k + 1)−1Nj(k) the at most (cid:4)k−1 (cid:5) cubes of (cid:5) cubes of the same cover. And so (cid:5) cubes of the optimal cover. This process certainly the optimal cover above; next consider the m's such that Qj(k)m are grid on until the consideration of the m's such that Qj(k)m are grid subcubes of leads to repetition of grid cubes, so, in order not to affect the sum above we make the convention that ϕj(k)m is replaced by zero any time it corresponds to a grid cube that has already been considered before. That is, ϕj(k)m is replaced by ϕj(k)m which can either be ϕj(k)m or zero, according to what was just explained. 1 Nj(k) Denote by ψi, i = k, k + 1, . . . , k1, the sum of the ϕj(k)m's corresponding to each association made above, so that the sum in (3.26) can be written as k1(cid:88) ψiϕ, (3.27) (3.29) i=k which, of course, still equals ϕ on Γ2−j(k). The next thing to do is to choose, for each i = k + 1, . . . , k1, j(i) in such a way that (3.28) j(k) < j(k + 1) < . . . < j(k1) and, moreover, each j(i) coincides with one of the j(cid:96)'s used to get (3.23). Then consider optimal covers of Γ2−j(i) and the sum k1(cid:88) (cid:88) i=k Qj(i)m∈Q(j(i)) ϕj(i)mψiϕ, which, by Lemma 3.8 and (3.28), equals ϕ on Γ2−j(k1). We know, following j(i) terms. However, because Remark 3.9, that each inner sum above has ∼ h−1 Traces for Besov spaces on fractal h-sets and dichotomy results 22 of the product structure of the latter and the support of each ψi, actually only a smaller number of terms are non-zero. We claim that the latter number is (cid:46) i−1h−1 j(i). This can be seen in the following way: for fixed i = k, k + 1, . . . , k1, only the ϕj(i)m such that supp ϕj(i)m ∩ supp ψi (cid:54)= ∅ are of interest; notice that this implies that the cubes of the optimal cover of Γ2−j(i) which contain the grid cubes at which such functions ϕj(i)m are located are contained in a neighbourhood of radius ∼ 2−j(i) of supp ψi which, in turn, is contained in the union of neighbourhoods of radius ∼ 2−j(k) of the (cid:46) i−1h−1 j(k) cubes of the optimal cover of Γ2−j(k) which were used to form ψi; since these union measures (cid:46) i−1h−1 j(k)hj(k), that is, (cid:46) i−1, and the cubes of the optimal cover of Γ2−j(i) contain disjoint balls of radius ∼ 2−j(i), so measuring ∼ hj(i) each, the number of such cubes that are being considered here must be (cid:46) i−1h−1 j(i); since, clearly, the same estimate holds for the family of grid subcubes of such cubes, our claim is proved. We write then (3.29) in the form (3.30) k1(cid:88) (cid:48)(cid:88) i=k Qj(i)m∈Q(j(i)) ϕj(i)mψiϕ, where the prime over the sum means that we are in fact taking only (cid:46) i−1h−1 j(i) of the terms, according to the discussion above, without changing the value of (3.29). Now we remark that, for i = k, k + 1, . . . , k1, supp ϕj(i)mψi ⊂ supp ϕj(i)m ⊂ 3 2 Qj(i)m and, due to Leibniz formula, (3.28) and (2.41), for a fixed K ∈ N there is a positive constant cK such that Dγ(ϕj(i)mψi)(x) ≤ cK 2j(i)γ, x ∈ Rn, γ ∈ Nn 0 with γ ≤ K. That is, (3.11), (3.15) and (3.16) hold for λr = λ(j(i),m) = 1 and ϕr = ϕ(j(i),m) = ϕj(i)mψi with r = (j(i), m) belonging to a set Ik as in (3.14), though more involved to describe: here we have Jk = {j(k), j(k+1), . . . , j(k1)} and each Mj(i), with i = k, k + 1, . . . , k1, is the set of m's considered in the corresponding inner sum in (3.30). It then follows, specially from (3.17), that the quasi-norm of the sum in (3.12) (or of an alternative similar sum, as discussed in Discussion 3.11 Traces for Besov spaces on fractal h-sets and dichotomy results 23 apropos of the consideration of moment conditions) is q/p1/q j(i)2−j(i)n τ p (cid:33)1/q − q p h − q p j(i) i=k  k1(cid:88) (cid:32) k1(cid:88) (cid:32) k1(cid:88) (cid:32) ∞(cid:88) i=k i=k (cid:46) (cid:46) = (cid:46)  (cid:48)(cid:88) Qj(i)m∈Q(j(i)) τ q j(i)2 − q p σq i −j(i)n q p i (cid:33)1/q (cid:33)1/q j(i) − q p i , i=k the last estimate being a consequence of (3.23) and the choice of the j(i) in (3.28). Using now the hypothesis 0 < p < q < ∞, we have that the last expres- sion above tends to zero when k tends to infinity, so that the result follows from Discussion 3.6. Remark 3.15. An observation corresponding to the one made in Remark 3.13 holds also here. The next result shows that when 1 < q we can also conclude as in Proposition 3.12 with an hypothesis weaker than (3.18). Proposition 3.16. Let Γ be an h-set satisfying the porosity condition, 1 < q < ∞, τ admissible. Assume that (3.31) Then D(Rn \ Γ) is dense in Bτ τ −1h1/p(n)1/p /∈ (cid:96)q(cid:48). p,q(Rn), therefore p,q(Rn) cannot exist. trΓBτ jk+1−1(cid:88) Proof. Denote σ := τ h−1/p(n)−1/p. From (3.31) it follows the existence of some strictly increasing sequence (jk)k∈N ⊂ N such that (3.32) l ≥ 1 , −q(cid:48) σ k ∈ N. l=jk For each k ∈ N, consider optimal covers of Γ2−i, in the sense of Remark 3.10, for all i = jν, jν + 1, . . . , jν+1 − 1 and all ν = 1, . . . , k, and follow our Discussions 3.6 and 3.11 with Ik from (3.14) defined such that Jk = {j ∈ Traces for Besov spaces on fractal h-sets and dichotomy results 24 N : ∃ ν = 1, . . . , k : j ∈ {jν, jν + 1, . . . , jν+1 − 1}} and Mj = {m ∈ Zn : Qjm ∈ Q(j)}, ϕ(j,m) = ϕjm and (3.33) λ(j,m) = σ −q(cid:48) j k (cid:33)−1 (cid:32)jν+1−1(cid:88) l=jν −q(cid:48) l σ j = jν, . . . , jν+1 − 1, ν = 1, . . . , k. , Due to Remark 2.22, this fits nicely into Discussions 3.6 and 3.11. In par- ticular, (3.15) and (3.16) immediately hold. As to (3.11), we have, for x ∈ Γ2 −jk+1, that j=j1 Qjm∈Q(j) (cid:88) jk+1−1(cid:88) jν+1−1(cid:88) k(cid:88) jν+1−1(cid:88) k(cid:88) ϕ(x) 1 k j=jν ν=1 σ −q(cid:48) j k = = λ(j,m)ϕ(j,m)(x)ϕ(x) (cid:32)jν+1−1(cid:88) (cid:32)jν+1−1(cid:88) l=jν −q(cid:48) σ l (cid:33)−1 (cid:88) (cid:33)−1 Qjm∈Q(j) −q(cid:48) σ j −q(cid:48) σ l = ϕ(x). ϕjm(x)ϕ(x) ν=1 j=jν l=jν It then follows, specially from (3.17), that the quasi-norm of the sum in (3.12) (or of an alternative similar sum, as discussed in Discussion 3.11 apropos of the consideration of moment conditions) is λ(j,m)τj2 pp −j n  q p q (cid:33) 1 (cid:33)−q q p q  1  1 q −q(cid:48) σ l σq j −j n p q((cid:93)Q(j)) (cid:32)jν+1−1(cid:88) (cid:33)1−q 1 l=jν q −q(cid:48) σ l Qjm∈Q(j) (j,m)τ q λq j 2 j=j1  (cid:88) jk+1−1(cid:88) (cid:32)jk+1−1(cid:88)  k(cid:88) jν+1−1(cid:88)  k(cid:88) (cid:32)jν+1−1(cid:88) (cid:33) 1 (cid:32) k(cid:88) 1 k j=j1 j=jν l=jν ν=1 ν=1 σ q −q(cid:48)q j k (cid:46) = ∼ = ≤ 1 k 1 ν=1 = k q(cid:48) −−−→ − 1 k→∞ 0, since q > 1, where we have also used Remark 3.9, (3.33) and (3.32). Thus (3.12) holds (possibly with an alternative similar sum if moment conditions are required, as mentioned above) and Discussion 3.6 then concludes the proof. Traces for Besov spaces on fractal h-sets and dichotomy results 25 Remark 3.17. An observation corresponding to the one made in Remark 3.13 holds also here. Theorem 3.18. Let Γ be an h-set satisfying the porosity condition, σ an admissible sequence and let either 1 ≤ p < ∞, 0 < q < ∞, or 0 < q ≤ p < 1. Then either (i) Bσ p,q(Γ) = trΓBσh1/p(n)1/p p,q (Rn) exists or (ii) D(Rn \ Γ) is dense in Bσh1/p(n)1/p p,q (Rn) and, therefore, trΓBσh1/p(n)1/p p,q (Rn) cannot exist. Proof. Either (3.6) holds or not. If it holds, then from Definition 3.1 it follows that (i) above holds. If (3.6) fails, then, with τ := σh1/p(n)1/p, τ −1h1/p(n)1/p /∈ (cid:96)q(cid:48), therefore, from Propositions 3.12 (in the case 0 < q ≤ 1) and 3.16 (in the case 1 < q < ∞), (ii) above holds. Remark 3.19. The only use of the porosity condition in the proof above is in guaranteeing the existence of suitable atoms when moment condi- tions for these are required. Therefore the result of the theorem above holds without assuming porosity of Γ whenever the atomic representation of Bσh1/p(n)1/p Remark 3.20. From the proof above it also follows that, under the conditions of the theorem, (i) is equivalent to (3.6), that is, σ−1 ∈ (cid:96)q(cid:48). Conjecture 3.21. Let Γ be an h-set satisfying the porosity condition, σ an admissible sequence and 0 < p < 1, 0 < p < q < ∞. Define vp by the identity 1 vp (Rn) does not require atoms with moment conditions. q . With the extra assumption p − 1 = 1 p,q (3.34) j→∞ hjσvp lim j = 0 the alternative in the conclusion of the theorem above also holds. More pre- cisely, assertion (i) holds iff assertion (ii) does not hold iff (3.35) σ−1 ∈ (cid:96)vp. We discuss a little bit this conjecture. Clearly, if (3.35) holds then Defini- tion 3.1 guarantees that (i) holds, and therefore (ii) does not, even without the extra assumption (3.34). On the other hand, Proposition 3.14, with τ := σh1/p(n)1/p, guarantees that at least in the special subcase of σ−1 /∈ (cid:96)vp given by lim sup σ−1 > 0 (ii) holds, and therefore (i) does not. Traces for Besov spaces on fractal h-sets and dichotomy results 26 In other words, we can get the alternative in the conclusion of the the- orem above in the case 0 < p < 1, 0 < p < q < ∞ if instead of the extra assumption (3.34) we assume that lim sup σ−1 > 0. The drawback of this restriction is that it immediately implies that (3.35) never holds, so (3.35) is not a real alternative under that extra assumption. From this point of view, (3.34) is more interesting. Notice also that, in order that this conjec- ture does not contradict Definition 3.1, under (3.34) it must be true that (3.35) holds whenever any of the conditions (3.7) does. In fact, by standard comparison criteria for series it is easy to see directly that this is indeed the case (take also into account that the case vr = ∞ in the conditions (3.7) never holds in presence of (3.34)). Remark 3.22. It is worth noticing that in the case when Γ is a d-set with 0 < d < n, (3.34) holds whenever (3.35) fails, so in that setting, and by the above discussion, there is no need to impose the extra condition (3.34), the conjecture then turning out to be indeed a known result (cf. [Tr7, (1.5)]). However, for general h-sets we cannot dispense with an extra assumption (such as (3.34)), as the Conjecture above then fails, at least in the part of the equivalence with (3.35). We give a class of examples where this is the case: Consider h(r) = (1 + log r)b, r ∈ (0, 1], for any given b ∈ (−1, 0), recall (2.13) and the discussion afterwards. Let 0 < p < 1 and 0 < p < q < ∞. Given any κ ∈ (−b( 1 q ], consider σ = ((1 + j)κ)j. It is easy to see that, though (3.35) fails, by Definition 3.1(ii) with r = min{q, 1} (i) of the theorem above holds, i.e., the trace exists (and therefore the corresponding assertion (ii) fails, as was discussed in the beginning of Section 3.2). q ) + (1 + b) 1 p − 1 p − 1 q(cid:48) , 1 Nevertheless, our conjecture as stated above resists this class of coun- terexamples: as is easily seen, for such a class the assumption (3.34) is violated. 3.3 Dichotomy results We combine our results from the previous subsections and deal with the so-called dichotomy of trace spaces. First we briefly describe the idea. p,q(Rn) with 0 < p, q < ∞. So Recall that D(Rn) is dense in all spaces Bτ removing from Rn only 'small enough' Γ one can ask whether (still) (3.36) D(Rn \ Γ) is dense in Bτ p,q(Rn). Conversely, we have the affirmative trace results mentioned in Section 3.1, but one can also ask for what ('thick enough') Γ (3.37) there exists a trace of Bτ p,q(Rn) in Lp(Γ) (for sufficiently high smoothness and q-regularity). Though these questions may arise independently, it is at least clear that whenever D(Rn \ Γ) is Traces for Besov spaces on fractal h-sets and dichotomy results 27 p,q(Rn), then there cannot exists a trace according to (3.2); see dense in Bτ our discussion at the beginning of Section 3.2 and [Tr7] for the corresponding argument in the classical case. Remark 3.23. It is not always true that one really has an alternative in the sense that either there is a trace or D(Rn \ Γ) is dense in Bτ p,q(Rn). Triebel studied such questions in [Tr7] for spaces of type Bs p,q and described an example of a set Γ where a gap remains: traces can only exist for spaces p,q with smoothness s ≥ s0, whereas density requires s ≤ s1 and s1 < s0. Bs However, if one obtains an alternative between (3.36) and (3.37), then following Triebel in [Tr7] we call this phenomenon dichotomy. First we recall p,q(Rn) and point out necessary modifications this notion for spaces of type Bs for our setting afterwards. Let (3.38) trΓ : Bs p,q(Rn) → Lp(Γ) be the trace operator defined by completion from the pointwise trace ac- cording to (3.2), and Bp(Rn) = {Bs 0 < p < ∞. p,q(Rn) : 0 < q < ∞, s ∈ R}, (3.39) Definition 3.24. Let n ∈ N, Γ ⊂ Rn, 0 < p < ∞. The dichotomy of the scale Bp(Rn) with respect to Lp(Γ), denoted by D(Bp(Rn), Lp(Γ)), is defined by D(Bp(Rn), Lp(Γ)) = (sΓ, qΓ), sΓ ∈ R, 0 < qΓ < ∞, (3.40) if (cid:40) s > sΓ, 0 < q < ∞, s = sΓ, 0 < q ≤ qΓ, (3.41) (i) trΓBs p,q(Rn) exists for and (3.42) (ii) D(Rn \ Γ) is dense in Bs p,q(Rn) for (cid:40) qΓ < q < ∞, s = sΓ, s < sΓ, 0 < q < ∞. Remark 3.25. The notion applies to spaces of type F s p,q in a similar way, cf. [Tr7]. Then one has to define the borderline cases qΓ = 0 and qΓ = ∞, too. But this will not be needed at the moment in our setting. We briefly explain why it is reasonable to look for the 'breaking point' (sΓ, qΓ). In the diagram below we sketch this situation, where spaces of type p,q(Rn) are indicated by their parameters ( 1 q , s) (while p is always assumed Bs Traces for Besov spaces on fractal h-sets and dichotomy results 28 s sΓ trΓ exists Bs2 p,q2 (Rn) Bs1 p,q1 (Rn) p,u(Rn) p,q2(Rn). , s2) referring to Bs2 to be fixed). Assume that D(Rn \ Γ) is dense in some space Bs2 Then, since D(Rn) is dense in all p,u(Rn), we immediately ob- spaces Bt tain that D(Rn \ Γ) is also dense in all spaces Bt in which p,q2(Rn) is continuously embedded Bs2 (the shaded area left and below of p,q2(Rn) in the ( 1 q2 diagram). This explains why we look for the largest possible s2 and small- est possible q2 in (3.42). Conversely, if the trace exists for some space p,q1(Rn), then it exists likewise for Bs1 p,q1(Rn) (the shaded area right all spaces which embed continuously into Bs1 p,q1(Rn) in the diagram); hence we now and above of ( 1 q1 search for the smallest possible s1 and largest possible q1 in (3.41). Di- chotomy in the above-defined sense happens, if the two 'extremal' points , sΓ) in the diagram exists. merge, that is, the common breaking point ( 1 qΓ Then we denote the couple of parameters (sΓ, qΓ) by D(Bp(Rn), Lp(Γ)). In [Tr8, Sect. 6.4.3] Triebel mentioned already, that it might be more rea- sonable in general to exclude the limiting case q = qΓ in (3.41) or shift it to (3.42). But as will turn out below, (also) in our context the breaking point q = qΓ is always on the trace side. , s1) referring to Bs1 D(Rn \ Γ) dense 1 qΓ 1 q Now we collect what is known in the situation of spaces Bs p,q for hyper- planes Γ = Rm or d-sets Γ, 0 < d < n. Proposition 3.26. Let 0 < p < ∞. (i) Let m ∈ N, m ≤ n − 1. Then D(Bp(Rn), Lp(Rm)) = (cid:19) . , min{p, 1} (cid:19) (cid:18) n − m (cid:18) n − d p p (ii) Let Γ be a compact d-set, 0 < d < n. Then D(Bp(Rn), Lp(Γ)) = , min{p, 1} . Remark 3.27. The result (i) is proved in this explicit form in [Tr7], see also [Tr8, Cor. 6.69] and [Sch1]. The second part (ii) can be found in [Tr7] and [Tr8, Thm. 6.68] (with forerunners in [Tr4, Thm. 17.6] and [Tr5, Prop. 19.5]). As for the classical case of a bounded C∞ domain Ω in Rn with boundary Γ = ∂Ω (referring to d = n−1), 1 < p < ∞, 1 ≤ q < ∞, s ∈ R, the situation (whether D(Ω) is dense in Bs p,q(Ω)) is known for long, cf. [Tr1, Thm. 4.7.1], even in a more general setting, we refer to [Sch2]. Moreover, we dealt with dichotomy questions for weighted spaces corresponding to (ii) in [Pi, Ha]. Traces for Besov spaces on fractal h-sets and dichotomy results 29 More precisely, when Γ is again a compact d-set, 0 < d < n, 0 < p < ∞, and the weight wκ,Γ is given by (cid:40) κ dist (x, Γ) 1, , if dist (x, Γ) ≤ 1, if dist (x, Γ) ≥ 1, wκ,Γ(x) = with κ > −(n − d), then (3.43) D(Bp(Rn, wκ,Γ), Lp(Γ)) = , min{p, 1} . (cid:18) n − d + κ p (cid:19) We return to our setting to study traces of Bτ In all the cases mentioned above there are parallel results for F -spaces, too. p,q(Rn) on fractal h-sets. Obviously, definitions (3.39) -- (3.42) have to be adapted now. The following extension seems appropriate. (Recall that we put v = ∞ if 1 Definition 3.28. Let Γ be a porous h-set, and (3.44) Bp(Rn) = {Bτ Then the dichotomy of the scale Bp(Rn) with respect to Lp(Γ) is defined by (3.45) if τΓ is admissible, 0 < qΓ < ∞, and p,q(Rn) : 0 < q < ∞, τ admissible}, D(Bp(Rn), Lp(Γ)) = (τΓ, qΓ), 0 < p < ∞. v = 0.) (cid:18) 1 qΓ (cid:19) − 1 q , + (3.46) (i) trΓBτ p,q(Rn) exists for τ −1τΓ ∈ (cid:96)v, where 1 v = and (3.47) (ii) D(Rn \ Γ) is dense in Bτ p,q(Rn) for τ −1τΓ (cid:54)∈ (cid:96)v. Remark 3.29. One immediately verifies that (3.44) -- (3.47) with h(r) = rd, τ = (s) for s ∈ R, and 0 < p < ∞, 0 < q < ∞, coincides with (3.39) -- (3.42), that is, the notion is extended (and we may thus and in this sense retain the same symbols in a slight abuse of notation). Using the continuous embedding (3.48) for σ−1τ ∈ (cid:96)q∗ with 1 in Remark 3.25 to motivate the definition. p,q1(Rn) (cid:44)→ Bτ Bσ − 1 q∗ = ( 1 q2 q1 p,q2(Rn) )+, cf. [CaF, Thm. 3.7], we can argue as Part of the results contained in Theorem 3.18 and Remark 3.20 can then be rephrased in terms of this notion of dichotomy in the following way. Corollary 3.30. Let Γ be an h-set satisfying the porosity condition and 1 ≤ p < ∞. Then D(Bp(Rn), Lp(Γ)) =(cid:0)h1/p(n)1/p, 1(cid:1) . Traces for Besov spaces on fractal h-sets and dichotomy results 30 Remark 3.31. Again, with the special setting h(r) = rd, τ = (s), Corol- lary 3.30 coincides with Proposition 3.26(ii) for p ≥ 1. So one might expect some parallel result with qΓ = p for 0 < p < 1, see also (3.43). However, we are not yet able to (dis)prove this claim. More precisely, if 0 < p < 1 and p < q < ∞, (3.46) is satisfied with τΓ = h1/p(n)1/p and qΓ = p, see (3.7). The gap that remains at the moment is to confirm (3.47) in that case (possibly with some additional assumptions), recall Conjecture 3.21 and the discussion afterwards. But we can obtain some weaker version as follows. Introducing a notion of dichotomy where also q can be fixed beforehand, hence adapting accordingly (3.44), (3.45) and denoting the new versions respectively by Bp,q(Rn) and D(Bp,q(Rn), Lp(Γ)) = (τΓ, qΓ), while (3.46) and (3.47) are kept unchanged, we can also cast the case 0 < q ≤ p < 1 of Theorem 3.18 and Remark 3.20 in terms of dichotomy. Corollary 3.32. Let Γ be an h-set satisfying the porosity condition and 0 < q ≤ p < 1. Then D(Bp,q(Rn), Lp(Γ)) =(cid:0)h1/p(n)1/p, p(cid:1) . Acknowledgements It is our pleasure to thank our colleagues in the Department of Mathematics at the University of Coimbra for their kind hospitality during our stays there. The first author was partially supported by FEDER funds through COM- PETE -- Operational Programme Factors of Competitiveness ("Programa Op- eracional Factores de Competitividade") and by Portuguese funds through the Center for Research and Development in Mathematics and Applica- tions (University of Aveiro) and the Portuguese Foundation for Science and Technology ("FCT -- Fundação para a Ciência e a Tecnologia"), within project PEst-C/MAT/UI4106/2011 with COMPETE number FCOMP-01- 0124-FEDER-022690 and projects PEst-OE/MAT/UI4106/2014 and UID/MAT/04106/2013. The second author was also supported by the DFG Heisenberg fellowship HA 2794/1-2. References [BGT] N.H. Bingham, C.M. Goldie, and J.L. Teugels. Regular variation, volume 27 of Encyclopedia of Mathematics and its Applications. Cambridge Univ. Press, Cambridge, 1987. [Br1] M. Bricchi. Existence and properties of h-sets. Georgian Math. J., 9(1):13 -- 32, 2002. Traces for Besov spaces on fractal h-sets and dichotomy results 31 [Br2] [Br3] [Br4] [Ca1] [Ca2] [CaF] M. Bricchi. Tailored function spaces and related h-sets. PhD thesis, Friedrich-Schiller-Universität Jena, Germany, 2002. M. Bricchi. Complements and results on h-sets. In D.D. Haroske, Th. Runst, and H.J. Schmeisser, editors, Function Spaces, Dif- ferential Operators and Nonlinear Analysis - The Hans Triebel Anniversary Volume, pages 219 -- 230. Birkhäuser, Basel, 2003. M. Bricchi. Tailored Besov spaces and h-sets. Math. Nachr., 263- 264(1):36 -- 52, 2004. A. Caetano. Approximation by functions of compact support in Besov-Triebel-Lizorkin spaces on irregular domains. Studia Math., 142(1):47 -- 63, 2000. A.M. Caetano. Growth envelopes of Besov spaces on fractal h-sets. Math. Nachr., 286(5-6):550 -- 568, 2013. A.M. Caetano and W. Farkas. Local growth envelopes of Besov spaces of generalized smoothness. Z. Anal. Anwendungen, 25:265 -- 298, 2006. [CaL1] A.M. Caetano and S. Lopes. The fractal Dirichlet Laplacian. Rev. Mat. Complut., 24(1):189 -- 209, 2011. [CaL2] A.M. Caetano and S. Lopes. Spectral theory for the fractal Lapla- cian in the context of h-sets. Math. Nachr., 284(1):5 -- 38, 2011. [CoF] [EKP] [ET1] [ET2] [ET3] [Fa] F. Cobos and D.L. Fernandez. Hardy-Sobolev spaces and Besov spaces with a function parameter. In M. Cwikel and J. Peetre, editors, Function spaces and applications, volume 1302 of Lecture Notes in Mathematics, pages 158 -- 170. Proc. US-Swed. Seminar held in Lund, June, 1986, Springer, 1988. D.E. Edmunds, R. Kerman, and L. Pick. Optimal Sobolev imbed- dings involving rearrangement-invariant quasinorms. J. Funct. Anal., 170:307 -- 355, 2000. D.E. Edmunds and H. Triebel. Function spaces, entropy numbers, differential operators. Cambridge Univ. Press, Cambridge, 1996. D.E. Edmunds and H. Triebel. Spectral theory for isotropic fractal drums. C. R. Acad. Sci. Paris, 326(11):1269 -- 1274, 1998. D.E. Edmunds and H. Triebel. Eigenfrequencies of isotropic frac- tal drums. Oper. Theory Adv. Appl., 110:81 -- 102, 1999. K.J. Falconer. The geometry of fractal sets. Cambridge Univ. Press, Cambridge, 1985. Traces for Besov spaces on fractal h-sets and dichotomy results 32 [FaLe] W. Farkas and H.-G. Leopold. Characterisation of function spaces of generalised smoothness. Ann. Mat. Pura Appl., 185(1):1 -- 62, 2006. [Ha] [JW] [KL] [KZ] [Li] [Lo] [Ma] [Me] [Mo1] [Mo2] [Ne1] [Ne2] D.D. Haroske. Dichotomy in Muckenhoupt weighted function space: A fractal example. In B.M. Brown, J. Lang, and I. Wood, editors, Spectral Theory, Function Spaces and Inequalities. New Techniques and Recent Trends, volume 219 of Operator Theory: Advances and Applications, pages 69 -- 89. Springer, Basel, 2012. A. Jonsson and H. Wallin. Function spaces on subsets of Rn. Math. Rep. Ser. 2, No.1, xiv + 221 p., 1984. G.A. Kalyabin and P.I. Lizorkin. Spaces of functions of generalized smoothness. Math. Nachr., 133:7 -- 32, 1987. V. Knopova and M. Zähle. Spaces of generalized smoothness on h-sets and related Dirichlet forms. Studia Math., 174(3):277 -- 308, 2006. P.I. Lizorkin. Spaces of generalized smoothness. Mir, Moscow, pages 381 -- 415, 1986. Appendix to Russian ed. of [Tr2]; Russian. S. Lopes. Besov spaces and the Laplacian on fractal h-sets. PhD thesis, Universidade de Aveiro, Portugal, 2009. P. Mattila. Geometry of sets and measures in euclidean spaces. Cambridge Univ. Press, Cambridge, 1995. C. Merucci. Applications of interpolation with a function param- eter to Lorentz, Sobolev and Besov spaces. In M. Cwikel and J. Peetre, editors, Interpolation spaces and allied topics in analy- sis, volume 1070 of Lecture Notes in Mathematics, pages 183 -- 201. Proc. Conf., Lund/Swed. 1983, Springer, 1984. S.D. Moura. Function spaces of generalised smoothness. Disser- tationes Math., 398:88 pp., 2001. S.D. Moura. Function spaces of generalised smoothness, entropy numbers, applications. PhD thesis, Universidade de Coimbra, Por- tugal, 2002. J.S. Neves. Lorentz-Karamata spaces, Bessel and Riesz potentials and embeddings. Dissertationes Math., 405:46 pp., 2002. J.S. Neves. Spaces of Bessel-potential type and embeddings: the super-limiting case. Math. Nachr., 265:68 -- 86, 2004. Traces for Besov spaces on fractal h-sets and dichotomy results 33 [Pi] [Ro] [Sch1] [Sch2] [Tr1] [Tr2] [Tr3] [Tr4] [Tr5] [Tr6] [Tr7] [Tr8] [TW] [Zy] I. Piotrowska. Traces on fractals of function spaces with Muck- enhoupt weights. Funct. Approx. Comment. Math., 36:95 -- 117, 2006. C.A. Rogers. Hausdorff measures. Cambridge Univ. Press, Lon- don, 1970. C. Schneider. Trace operators in Besov and Triebel-Lizorkin spaces. Z. Anal. Anwendungen, 29(3):275 -- 302, 2010. C. Schneider. Traces of Besov and Triebel-Lizorkin spaces on domains. Math. Nachr., 284(5-6):572 -- 586, 2011. H. Triebel. Interpolation theory, function spaces, differential op- erators. North-Holland, Amsterdam, 1978. H. Triebel. Theory of function spaces. Birkhäuser, Basel, 1983. Reprint (Modern Birkhäuser Classics) 2010. H. Triebel. Theory of function spaces II. Birkhäuser, Basel, 1992. Reprint (Modern Birkhäuser Classics) 2010. H. Triebel. Fractals and spectra. Birkhäuser, Basel, 1997. Reprint (Modern Birkhäuser Classics) 2011. H. Triebel. The structure of functions. Birkhäuser, Basel, 2001. H. Triebel. Theory of function spaces III. Birkhäuser, Basel, 2006. H. Triebel. The dichotomy between traces on d-sets Γ in Rn and the density of D(Rn \ Γ) in function spaces. Acta Math. Sinica, 24(4):539 -- 554, 2008. H. Triebel. Function Spaces and Wavelets on domains. EMS Tracts in Mathematics (ETM). European Mathematical Society (EMS), Zürich, 2008. H. Triebel and H. Winkelvoss. Intrinsic atomic characterizations of function spaces on domains. Math. Z., 221(4):647 -- 673, 1996. A. Zygmund. Trigonometric series. Cambridge Univ. Press, Cam- bridge, 2nd edition, 1977.
1604.07587
4
1604
2016-11-04T13:15:20
Weak$^*$ Fixed Point Property in $\ell_1$ and Polyhedrality in Lindenstrauss Spaces
[ "math.FA" ]
The aim of this paper is to study the $w^*$-fixed point property for nonexpansive mappings in the duals of separable Lindenstrauss spaces by means of suitable geometrical properties of the dual ball. First we show that a property concerning the behaviour of a class of $w^*$-closed subsets of the dual sphere is equivalent to the $w^*$-fixed point property. Then, the main result of our paper shows an equivalence between another, stronger geometrical property of the dual ball and the stable $w^*$-fixed point property. The last geometrical notion was introduced by Fonf and Vesel\'{y} as a strengthening of the notion of polyhedrality. In the last section we show that also the first geometrical assumption that we have introduced can be related to a polyhedral concept for the predual space. Indeed, we give a hierarchical structure among various polyhedrality notions in the framework of Lindenstrauss spaces. Finally, as a by-product, we obtain an improvement of an old result about the norm-preserving compact extension of compact operators.
math.FA
math
WEAK∗ FIXED POINT PROPERTY IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES EMANUELE CASINI, ENRICO MIGLIERINA, ŁUKASZ PIASECKI, AND ROXANA POPESCU Abstract. The aim of this paper is to study the w∗-fixed point property for nonexpansive mappings in the duals of separable Lindenstrauss spaces by means of suitable geometrical properties of the dual ball. First we show that a property concerning the behaviour of a class of w∗-closed subsets of the dual sphere is equivalent to the w∗-fixed point property. Then, the main re- sult of our paper shows an equivalence between another, stronger geometrical property of the dual ball and the stable w∗-fixed point property. The last geometrical notion was introduced by Fonf and Veselý as a strengthening of the notion of polyhedrality. In the last section we show that also the first ge- ometrical assumption that we have introduced can be related to a polyhedral concept for the predual space. Indeed, we give a hierarchical structure among various polyhedrality notions in the framework of Lindenstrauss spaces. Fi- nally, as a by-product, we obtain an improvement of an old result about the norm-preserving compact extension of compact operators. 1. Introduction and Preliminaries Let X be an infinite dimensional real Banach space and let us denote by BX its closed unit ball and by SX its unit sphere. We say that a Banach space X is a Lindenstrauss space if its dual is a space L1(µ) for some measure µ. A nonempty bounded closed and convex subset C of X has the fixed point property (shortly, fpp) if each nonexpansive mapping (i.e., the mapping T : C → C such that kT (x) − T (y)k ≤ kx − yk for all x, y ∈ C) has a fixed point. A dual space X ∗ is said to have the σ(X ∗, X)-fixed point property (σ(X ∗, X)-fpp) if every nonempty, convex, w∗-compact subset C of X ∗ has the fpp. The study of the σ(X ∗, X)-fpp reveals to be of special interest whenever a dual space has different preduals. For instance, this situation occurs when we consider the space ℓ1 and its preduals c0 and c where it is well-known (see [11]) that ℓ1 has the σ(ℓ1, c0)-fpp whereas it lacks the σ(ℓ1, c)-fpp. The first result of our paper is devoted to a geometrical characterization of the preduals X of ℓ1 that induce on ℓ1 itself the σ(ℓ1, X)-fpp. This theorem can be seen as an extension of the characterization given in Theorem 8 in [15] and it is based on the studies carried out in [3]. However, the main purpose of the present paper is to investigate the stability of σ(ℓ1, X)- fpp. Generally speaking, stability of fixed point property deals with the following question: let us suppose that a Banach space X has the fixed point property and Y is a Banach space isomorphic to X with "small" Banach-Mazur distance, does Y have fixed point property? This problem has been widely studied for fpp and only occasionally for w∗-topology (see [5,20]). It is worth pointing out that the stability property of w∗-fpp previously studied in the literature considers the renormings of 2010 Mathematics Subject Classification. 47H10, 46B45, 46B25. Key words and phrases. w∗-fixed point property, stability of the w∗-fixed point property, Lindenstrauss spaces, Polyhedral spaces, ℓ1 space, Extension of compact operators. Acknowledgements: The authors thank Stanisław Prus for helpful conversations on and around the article [15]. They also thank Libor Veselý for useful discussions. 1 WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 2 X ∗ whereas it maintains the original w∗-topology on the renormed space. Since a renorming of a given Banach space X ∗ not necessarily is a dual space, we prefer to introduce a more suitable notion of stability for w∗-fpp that takes into account each dual space with the proper w∗-topology induced by its predual (see Definition 3.1). In Section 3 we prove that the stability of σ(ℓ1, X)-fpp is equivalent to a suitable property concerning the behaviour of w∗-limit points of the set of extreme points of Bℓ1. It is worth to mention that the property playing a key role in Section 3 was already introduced in [8] by Fonf and Veselý, in the completely different setting of polyhedral spaces theory. The concept of polyhedrality, originally introduced by Klee in [13], is widely studied (for detailed survey about various definitions of polyhedrality for infinite dimensional spaces see [6, 8]) and it gives a deep insight of geometrical properties of Banach spaces. Beyond its intrinsic interest, polyhedrality has some important applications. For instance, in the framework of Lindenstrauss spaces, it is related to the existence of norm-preserving compact extension of compact operators (see [4] and [7]). In the last section of the paper we compare the geometrical assumptions used to study σ(ℓ1, X)-fpp with the main generalizations of polyhedrality already considered in the literature. Indeed, the assumption characterizing the preduals of ℓ1 satisfying stable w∗-fpp is listed as a generalization of polyhedrality in [8] and we show how the property ensuring the validity of σ(ℓ1, X)-fpp fits very well in a list of several properties related to the original definition of polyhedral space. Our results, in the framework of Lindenstrauss spaces, prove that the notions of polyhedrality play an important role also in fixed point theory. We also give a hierarchical structure among various notions of polyhedrality by restricting our attention to Lindenstrauss spaces. These results allow us to prove a new version of an old uncorrect result (Theorem 3 in [14]) concerning the norm-preserving compact extension of compact operators. Finally, we collect some notations. If A ⊂ X, then we denote by A, conv(A) and ext A the norm closure of A, the convex hull of A and the set of the extreme points ∗ the w∗-closure of of A respectively. Moreover, whenever A ⊂ X ∗, we denote by A A. For x ∈ SX , we call D(x) the image of x by the duality mapping, i.e., D(x) = {x∗ ∈ SX ∗ : x∗(x) = 1} . 2. A characterization of w∗-fixed point property in ℓ1 The aim of this brief section is to characterize the separable Lindenstrauss spaces X such that X ∗ has the σ(X ∗, X)-fpp. The present result adds a new character- ization of σ(X ∗, X)-fpp to those listed in [3]. Moreover, the theorem sheds some new light on the relationships between σ(X ∗, X)-fpp and a geometrical feature of the sphere in X ∗. We will show in Section 4 that this feature can be interpreted as a polyhedrality requirement on X. Theorem 2.1. Let X be a separable Lindenstrauss space. Then the following are equivalent. (i) X ∗ has the σ(X ∗, X)-fpp; (ii) there is no infinite set C ⊂ ext BX ∗ such that conv(C) ∗ ⊂ SX ∗. Proof. Let us suppose that X is a separable Lindenstrauss space with nonseparable dual. Then by Corollary 3.4 in [3] X fails the w∗-fpp and it also fails property (ii) by Theorem 2.3 in [16] and Theorem 2.1 in [4]. Therefore we limit ourselves to consider the case where X ∗ is isometric to ℓ1. Let us suppose that X ∗ fails the w∗-fixed point property. Then, by Theorem 4.1 in [3], there exist a subsequence σ(ℓ1,X) −−−−−→ e∗ (cid:8)e∗ nk(cid:9) of the standard basis of ℓ1 and a point e∗ ∈ Sℓ1 such that e∗ nk WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 3 Corollary 2 in [15], we have property (ii). On the other hand, without loss of generality we can assume that and e∗(nk) ≥ 0. By taking C = (cid:8)e∗ there exists a set C = (cid:8)e∗ nk(cid:9) where (cid:8)e∗ standard basis of ℓ1. Let us denote by e∗ the w∗-limit of (cid:8)e∗ n2 , · · ·(cid:9)(cid:1) ⊂ Sℓ1 . nk(cid:9), it is easy to see that X does not satisfy nk(cid:9) is a w∗-convergent subsequence of the nk(cid:9) then, by recalling By adapting to our setting the method developed in the last part of the proof of Theorem 8 in [15], we easily find a nonexpansive mapping fixed point free from conv(C) (cid:3) = conv (cid:0)(cid:8)e∗, e∗ to conv(C) conv(C) n1 , e∗ ∗ . ∗ ∗ 3. A characterization of stable w∗-fixed point property in ℓ1 This section is devoted to the main result of the paper. We characterize the stable w∗-fpp for the duals of separable Lindenstrauss spaces by means of a suitable property describing the interplay between w∗-topology and the geometry of the dual ball BX ∗ . To our knowledge, the first result dealing with the stability of w∗-fpp for this class of spaces is the following: Theorem 3.1. ( [20]) Let Y be a Banach space such that d(ℓ1, Y ) < 2. Then Y has the w∗-fpp. We remark that the statement of the previous theorem implicitly assume that ℓ1 is endowed with the σ(ℓ1, c0)-topology. Moreover, since Y is not necessarily a dual space, the author of [20] considers the original topology σ(ℓ1, c0) on Y . For this reason, this approach does not allow to consider a true w∗-fpp in the space Y . In order to avoid this undesirable feature we introduce a different definition of stability for the w∗-fpp. Definition 3.1. A dual space X ∗ enjoys the stable σ(X ∗, X)-fpp if there exists a real number γ > 1 such that Y ∗ has the σ(Y ∗, Y )-fpp whenever d(X, Y ) < γ, where d(X, Y ) is the Banach-Mazur distance between X and Y . We recall that every nonseparable dual of a separable Lindenstrauss space fails the w∗-fpp (see Corollary 3.4 in [3]). Therefore, in the sequel of this section, we restrict our attention to the preduals of ℓ1. If A ⊂ X ∗, then we denote by A′ the set of all w∗-limit points of A: A′ =nx∗ ∈ X ∗ : x∗ ∈ (A \ {x∗}) ∗o . The following Lemma shows how the geometrical assumption that will play a crucial role in our characterization of stable w∗-fpp influences the behaviour of some sequences in ℓ1. Lemma 3.2. Let X be a predual of ℓ1. (a) For every sequence {x∗ n} ⊂ ℓ1 coordinatewise converging to x∗ 0 and such that limn→∞ kx∗ n − x∗ 0k exists, it holds lim m→∞ lim n→∞ kx∗ n − x∗ mk = 2 lim n→∞ kx∗ n − x∗ 0k . If, in addition, (ext Bℓ1)′ ⊂ rBℓ1 for some 0 ≤ r < 1, then (b) for every sequence {x∗ n} ⊂ ℓ1 such that {x∗ n} tends to 0 coordinatewise, it holds and {x∗ n} is σ(ℓ1, X)-convergent to x∗ kx∗k ≤ r lim inf n→∞ kx∗ nk ; WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 4 (c) for every sequence {x∗ n} ⊂ ℓ1 such that {x∗ n} is σ(ℓ1, X)-convergent to x∗, up to a subsequence, it holds lim n→∞ kx∗ n − x∗k ≤ 1 + r 2 lim m→∞ lim n→∞ kx∗ n − x∗ mk . Proof. Assertion (a) is a straightforward consequence of the following fact: for every sequence {x∗ n} ⊂ ℓ1 coordinatewise converging to 0 and such that limn→∞ kx∗ nk exists it holds lim n→∞ kx∗ n + x∗k = lim n→∞ kx∗ nk + kx∗k for every x∗ ∈ ℓ1. Now we prove assertion (b). For all m, n ∈ N and x ∈ BX , we have m ∞ x∗ n(x) = x∗ n(i)e∗ i (x) + x∗ n(i)e∗ i (x) ≤ Xi=1 Xi=m+1 m Xi=1 x∗ n(i) + kx∗ nk sup i≥m+1 e∗ i (x) . supi≥m+1 e∗ i (x) ≤ r for every x ∈ BX . There- Since (extBℓ1 )′ ⊆ rBℓ1 , then lim m→∞ fore, we easily see that x∗(x) = lim n→∞ x∗ n(x) ≤ lim inf n→∞ kx∗ nk lim m→∞(cid:18) sup i≥m+1 e∗ i (x)(cid:19) ≤ r lim inf n→∞ kx∗ nk , which proves the thesis of assertion (b). In order to prove the last assertion, without loss of generality, we may assume 0 and 0k exist. By assertions (b) and (a), we n} is coordinatewise convergent to x∗ n − x∗ 0 ∈ ℓ1 such that {x∗ n − x∗k and limn→∞ kx∗ that there exists x∗ that limn→∞ kx∗ get lim n→∞ kx∗ n − x∗k = lim n→∞ ≤ lim n→∞ 1 + r kx∗ kx∗ n − x∗ n − x∗ 0k + kx∗ 0k + r lim n→∞ 0 − x∗k kx∗ n − x∗ 0k = lim m→∞ lim n→∞ kx∗ n − x∗ mk . 2 (cid:3) For the sake of convenience of the reader, we recall the following known result. Lemma 3.3. ( [20]) Let Y ∗ be a dual Banach space, K ⊂ Y ∗ be a convex w∗- compact subset and T : K → K be a nonexpansive mapping. Then, for every x∗ ∈ K there is a closed convex subset H(x∗) ⊂ K which is invariant under T and satisfies (a) diam (H(x∗)) ≤ supn kx∗ − T nx∗k ; (b) supy∗∈H(x∗) kx∗ − y∗k ≤ 2 supn kx∗ − T nx∗k. Theorem 3.4. Let X be a predual of ℓ1. Then the following are equivalent. (i) ℓ1 has the stable σ(ℓ1, X)-fpp; (ii) (ext Bℓ1)′ ⊂ rBℓ1 for some 0 ≤ r < 1. Proof. We first prove that (i) implies (ii). By contradiction, let us suppose that X does not satisfies property (ii). For clarity we divide the proof of this implication into three parts. Step 1. (Renorming.) Let ε ∈ (0, 1) and choose a subsequence (e∗ nk ) is w∗-convergent to e∗ 6= 0, and n) in ℓ1 such that (e∗ standard basis (e∗ nk ) of the (3.1) e∗(nk) < ε 4 ke∗k . ∞ Xk=1 WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 5 Consider the subset C of ℓ1 defined by C =(α0e∗ + Xk=1 ∞ ∞ αke∗ nk : αi ≥ 0 for all i ∈ N ∪ {0} and αi = 1) . Xi=0 It is easy to check that C is convex and w∗-compact (see e.g. Corollary 2 in [15]). From (3.1) it follows that for every x = α0e∗ +P∞ k=1 αke∗ nk ∈ C we have α0e∗(j) + αk kxk = α0 Xj∈N\{nk} ≥ α0 Xj∈N\{nk} ≥ α0 Xj∈N\{nk} ≥ Xj∈N\{nk} e∗(j) + Xj∈{nk} Xk=1 ∞ e∗(j) + αk − α0 Xj∈{nk} e∗(j) + (1 − α0) Xj∈N\{nk} e∗(j) − Xj∈{nk} e∗(j) e∗(j) e∗(j) − α0 Xj∈{nk} e∗(j) ≥ (1 − = (1 − Let ε 4 ke∗k ) ke∗k − ) ke∗k . ε 4 ε 2 K := 1 (1 − ε) ke∗k C. The set K is convex, w∗-compact and K ∩ Bℓ1 = ∅. Next we define the set D ⊂ ℓ1 by D = conv(Bℓ1 ∪ K ∪ −K). It is easy to check that D is convex, symmetric, w∗-compact, and 0 is its interior point. Therefore D is a dual unit ball of an equivalent norm k·k on X. Let Y = (X, k·k). Obviously, D = BY ∗ and so (3.2) Bℓ1 ⊂ D ⊂ 1 (1 − ε) ke∗k Bℓ1 , d(X, Y ) ≤ 1 (1 − ε) ke∗k . Consider a subspace Z ⊂ Y ∗ defined as where n1 , e∗ ∞ e∗ − Z = span(cid:0)(cid:8)e∗, e∗ Pj=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Pj=1 e∗ 0 := e∗ − ∞ ∞ 0, e∗ e∗(nj)e∗ nj n2 , . . .(cid:9)(cid:1) = span(cid:0)(cid:8)e∗ Pj=1 Pj=1 ke∗k − e∗(nj)e∗ e∗ − = ∞ nj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n1 , e∗ n2 , . . .(cid:9)(cid:1), e∗(nj)e∗ nj . e∗(nj) Step 2. We claim that BZ = D ∩ Z has the following property: (♥) BZ = conv(K ∪ −K). Indeed, from definition of K and Z, it follows that BZ = conv(Bℓ1 ∪ K ∪ −K) ∩ Z = conv(Bℓ1 ∪ conv(K ∪ −K)) ∩ Z = conv((Bℓ1 ∩ Z) ∪ conv(K ∪ −K)) WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 6 so, in order to prove property (♥), we have to show that (3.3) Bℓ1 ∩ Z ⊂ conv(K ∪ −K). It is easy to see that ℓ1 and Z are isometrically isomorphic via φ : ℓ1 → Z defined nj−1 , j ≥ 2, where (ej)j≥1 denotes the standard basis in by φ(e1) = e∗ ℓ1. Since Bℓ1 = conv ({±e1, ±e2, ±e3, . . . }), it follows that 0, φ(ej ) = e∗ Consequently, in order to prove (3.3), it is enough to show that 0, ±e∗ n1 , ±e∗ Bℓ1 ∩ Z = conv(cid:0)(cid:8)±e∗ (cid:8)±e∗ n2 , . . .(cid:9)(cid:1). n2 , . . .(cid:9) ⊂ conv(K ∪ −K). n1 , ±e∗ 0, ±e∗ e∗ n1 (1 − ε) ke∗k , ± e∗ n2 (1 − ε) ke∗k (cid:26)± , . . .(cid:27) ⊂ conv(K ∪ −K) Obviously (3.4) so n1 , ±e∗ Therefore it remains to prove that (cid:8)±e∗ n2 , . . .(cid:9) ⊂ conv(K ∪ −K). ±e∗ 0 ∈ conv(K ∪ −K). The case P∞ we have j=1 e∗(nj ) = 0 is trivial. Suppose that P∞ j=1 e∗(nj ) 6= 0. From (3.4) conv(K ∪ −K) ⊃ conv(cid:18)(cid:26)± e∗ n1 e∗ n2 , ± (1 − ε) ke∗k (1 − ε) ke∗k , . . .(cid:27)(cid:19) 1 (1 − ε) ke∗k Bℓ1 ∩ span(cid:0)(cid:8)e∗ n1 , e∗ n2 , . . .(cid:9)(cid:1) and, consequently, ± = 1 (1 − ε) ke∗k ∈ conv(K ∪ −K). e∗(nj) e∗(nj)e∗ nj · ∞ ∞ Pj=1 Pj=1 Therefore, we easily see that for t :=(cid:16)P∞ Pj=1 conv(K ∪ −K) ∋ t · − ∞ e∗(nj)e∗ nj j=1 e∗(nj)(cid:17) ·(cid:16)1 +P∞ j=1 e∗(nj)(cid:17)−1 , + (1 − t) · e∗ (1 − ε) ke∗k (1 − ε) ke∗k e∗(nj) e∗ − e∗(nj)e∗ nj (1 − ε) ke∗k (1 + e∗(nj)) ∞ Pj=1 ∞ Pj=1 ∞ Pj=1 ∞ Pj=1 ke∗k − e∗(nj) (1 − ε) ke∗k (1 + e∗ 0 e∗(nj)) ∞ Pj=1 = = WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 7 and, since conv(K ∪ −K) is symmetric, we get ke∗k − ± ∞ Pj=1 (1 − ε) ke∗k (1 + e∗(nj) e∗ 0 ∈ conv(K ∪ −K). e∗(nj)) ∞ Pj=1 From (3.1) it follows that ke∗k − (1 − ε) ke∗k (1 + e∗(nj )) > 1 − ε 4 (1 − ε)(1 + ε 4 ) > 1. e∗(nj) ∞ Pj=1 ∞ Pj=1 Thus ±e∗ 0 ∈ conv(K ∪ −K) which finishes the proof of property (♥). Step 3. (Fixed point free nonexpansive map.) We now define the operator T : Z → Z by T  t0e∗ + ∞ Xj=1 tje∗ nj  = tj−1e∗ nj , ∞ Xj=1 ∞ Xj=0 tj < ∞. It is easy to see (via (3.1)) that the mapping T is well-defined and linear. We claim that T (BZ) ⊂ BZ . Indeed, property (♥) ensures that every x ∈ BZ has the form x = (1 − λ)y + λz for some λ ∈ [0, 1], y ∈ K and z ∈ −K. Therefore, since T (K) ⊂ K and T (−K) ⊂ −K, we see that T x = T ((1 − λ)y + λz) = (1 − λ)T y + λT z ∈ BZ . Consequently, since K ⊂ SZ, kT k = 1. Clearly, the restriction T K of operator T to the w∗-compact convex set K ⊂ Y ∗ is a fixed point free nonexpansive map. Finally, by taking e∗ arbitrarily close to the unit sphere Sℓ1 and ε ց 0 in (3.2), we get a contradiction. 0,n − x∗ 0 − T kx∗ 0,n − T x∗ Now, we proceed to prove the implication (ii) ⇒ (i). This part of the proof is an appropriate adaptation of the proof of the main result in [20]. Let Y be a Banach space isomorphic to X and A be any isomorphism from X onto Y . Let T :K → K be a nonexpansive mapping, where K ⊂ Y ∗ is a convex w∗- compact set. is w∗-convergent to x∗ Since T is a nonexpansive mapping, for every k ∈ N we have that limn→∞(cid:13)(cid:13)x∗ (cid:13)(cid:13)x∗ Therefore, by (a) in Lemma 3.3, there exists a closed convex invariant set H(x∗ K such that diam(H(x∗ such that It is well known that there exists a sequence (cid:8)x∗ 0,n(cid:9) ⊂ K such 0,n(cid:13)(cid:13)Y ∗ = 0 (see, e.g, [10]). We may assume that (cid:8)x∗ 0,n(cid:9) 0 ∈ K and that the limit α0 = limn→∞(cid:13)(cid:13)x∗ 0(cid:13)(cid:13)Y ∗ exists. n→∞ (cid:13)(cid:13)x∗ 0(cid:13)(cid:13)Y ∗ ≤ lim sup 0(cid:13)(cid:13)Y ∗ ≤ α0. 0)) ≤ α0. Then there exists a sequence (cid:8)x∗ 1,n(cid:13)(cid:13)Y ∗ = 0, 1,n(cid:9) is w∗-convergent to x∗ 1(cid:13)(cid:13)Y ∗ exists. 1,n(cid:9) ⊂ ℓ1 is σ(ℓ1, X)-convergent to A∗x∗ 1(cid:13)(cid:13)ℓ1 (1) limn→∞(cid:13)(cid:13)x∗ (2) (cid:8)x∗ (3) α1 = limn→∞(cid:13)(cid:13)x∗ Now, we have that (cid:8)A∗x∗ is no loss of generality in assuming that limn→∞(cid:13)(cid:13)A∗x∗ 1. Moreover, there exists. Hence, 1,n(cid:9) ⊂ H(x∗ 0,n − T kx∗ 1,n − A∗x∗ 1,n − T x∗ 0) ⊂ 0) 1,n − x∗ 1 ∈ K, WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 8 by (c) in Lemma 3.2 we obtain α0 ≥ lim sup m→∞ 1 kAk 1 kAk ≥ ≥ ≥ lim sup m→∞ 2 lim sup lim sup 1,m − x∗ 1,m − A∗x∗ n→∞ (cid:13)(cid:13)x∗ n→∞ (cid:13)(cid:13)A∗x∗ n→∞(cid:13)(cid:13)A∗x∗ n→∞(cid:13)(cid:13)x∗ 1,n(cid:13)(cid:13)Y ∗ 1,n(cid:13)(cid:13)ℓ1 1(cid:13)(cid:13)ℓ1 1(cid:13)(cid:13)Y ∗ = 1,n − A∗x∗ 1,n − x∗ 1 + r lim lim 2 1 + r 1 kA−1k kAk 1 kA−1k kAk 2 1 + r α1. 1 + r From the inequalities above, we conclude that α1 ≤ kA−1k kAk since {x∗ 2α0. 1,n} is w∗-convergent to x∗ 0 −x∗ 0 −x∗ 1, (b) in Lemma 3.3 yields kx∗ 2 α0. Moreover, 1kY ∗ ≤ 0 −x∗ Repeated applications of this construction give us a sequence {αn} of non neg- ative numbers such that (3.5) αn+1 ≤(cid:20)kA−1k kAk 1 + r 2 (cid:21)n+1 α0 n} ⊂ Y ∗ such that kx∗ and a sequence {x∗ nkY ∗ ≤ αn. From inequality (3.5), we see at once that the sequence {αn} converges to 0 if kA−1k kAk < 2 n} strongly converges to a fixed point of T . This fact concludes the proof by showing that Y ∗ has the w∗-fpp whenever d(X, Y ) < 2 (cid:3) 1+r . Moreover, if this condition holds, the sequence {x∗ n+1kY ∗ ≤ 2αn and kx∗ n − T x∗ n − x∗ 1+r . Remark 3.5. It is easy to observe that, from the proof of Theorem 3.4, we obtain also a quantitative estimation of the stability constant γ = 2 1+r . Moreover, it is worth pointing out that the estimation above is sharp when r = 0 as shown by the example contained in [18]. On the other hand the proof of the sharpness of the present estimation when 0 < r < 1 remains as an open problem. 4. Polyhedrality in Lindenstrauss spaces The aim of this section is to show that the geometrical properties used in Theo- rems 2.1 and 3.4 are strictly related to polyhedrality. The starting point to consider polyhedrality in an infinite-dimensional setting is the definition given by Klee in [13], where he extended the notion of convex finite-dimensional polytope for the case of the closed unit ball BX of an infinite-dimensional Banach space X. Thenceforth polyhedrality was extensively studied and several different definitions have been stated. For a detailed account about these definitions and their relationships see [6] and [8]. Here we restrict our attention to some of the definitions collected in [8]. Moreover, we show that the property introduced in Theorem 2.1 can be considered as a new definition of polyhedrality (namely, property (pol-iii)) since we will prove that this property is placed in an intermediate position between other polyhedrality notions that are already considered in the literature. Definition 4.1. Let X be a Banach space. We consider the following properties of X: (pol - i) (ext BX ∗ )′ ⊂ {0} ( [19]); (pol - ii) (ext BX ∗ )′ ⊂ rBX ∗ for some 0 < r < 1 ( [8]); (pol - iii) there is no infinite set C ⊂ ext BX ∗ such that conv(C) (pol - iv) there is no infinite-dimensional w∗-closed proper face of BX ∗ ( [14]); (pol - v) x∗(x) < 1 whenever x ∈ SX and x∗ ∈ (ext BX ∗ )′ ( [9]); (pol - vi) the set ext D(x) is finite for each x ∈ SX (property (∆) in [8]); (pol - vii) sup {x∗(x) : x∗ ∈ ext BX ∗ \ D(x)} < 1 for each x ∈ SX ( [1]); ⊂ SX ∗; ∗ WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 9 (pol-K) the unit ball of every finite-dimensional subspace of X is a polytope ( [13]). The following theorem clarifies the relationships between the various notions of polyhedrality stated in Definition 4.1 in the framework of Lindenstrauss spaces. Theorem 4.1. Let X be a Lindenstrauss space. The following relationships hold: (pol − v) m (pol − i) ⇒ (pol − ii) ⇒ (pol − iii) ⇒ (pol − iv) ⇒ (pol − vii)⇔ (pol − K). m (pol − vi) Proof. The implications (pol − i) ⇒ (pol − ii) ⇒ (pol − iii) ⇒ (pol − iv), (pol − v) ⇒ (pol − vi) are trivial. The proof of Theorem 1.2 (p. 402-403) in [9] shows that The implication (pol − iv) ⇒ (pol − v). (pol − vi) ⇒ (pol − iv) follows easily from Lemma 4.2 below. The implication (pol − v) ⇒ (pol − vii) is proved in Theorem 1 in [6]. Finally the equivalence (pol − vii) ⇔ (pol − K) is proved in Theorem 4.3 in [4]. (cid:3) The previous proof needs the following lemma, that is interesting in itself as it Indeed, this property is strictly gives a property of the w∗-closed faces of BX ∗ . related to the compact norm-preserving extension of compact operators (see [4]). Lemma 4.2. Let X be a Lindenstrauss space and let F be a w∗-closed proper face of BX ∗ . Then there exists x ∈ SX such that F ⊆ D(x). Proof. Let us fix an element ¯x∗ ∈ F . Then we consider the subspace V = span(¯x∗ − F ). It is easy to prove that for every h∗ ∈ H = conv (F ∪ (−F )) there exists a unique real number α ∈ [−1, 1] such that h∗ ∈ α¯x∗ + V. Now, let A0(H) denotes the Banach space of all w∗-continuous affine symmetric (i.e., f (−x) = −f (x)) functions on H. We introduce the function ¯y : H → R defined by ¯y(h∗) = α. It is easy to recognize that ¯y ∈ A0(H) and that ¯yF = 1. By considering the separable subspace Y = span ({¯y}) of A0(H), we can apply Proposition 1 in [14] to show that there exists an isometry T : Y → X such that h∗(T (y)) = y(h∗) for every h∗ ∈ H and for every y ∈ Y . Therefore T (¯y) ∈ SX and F ⊆ D(T (¯y)). (cid:3) The equivalence among (pol-iv), (pol-v) and (pol-vi) allows us to clarify the situation about the existence of compact norm-preserving extension of a compact operator with values in a Lindenstrauss space. From [?] it is known that a Lin- denstrauss space X is polyhedral if and only if for every Banach spaces Y ⊂ Z and every operator T : Y → X with dim T (Y ) ≤ 2 there exists a compact ex- = kT k (combine Theorem 7.9 in [?] and Theorem 4.7 in [?]). Moreover, Lazar provided (see Theorem 3 in [14]) the following more general extension property that he asserted to be equivalent to polyhedrality. tension T : Z → X with (cid:13)(cid:13)(cid:13) T(cid:13)(cid:13)(cid:13) WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 10 Fact 4.3. If X is a Lindenstrauss space, then the following properties are equiva- lent: (1) X is a polyhedral space; (2) X does not contain an isometric copy of c; (3) there are no infinite-dimensional w∗- closed proper faces of BX ∗ ; (4) for every Banach spaces Y ⊂ Z and every compact operator T : Y → X there exists a compact extension T : Z → X with (cid:13)(cid:13)(cid:13) T(cid:13)(cid:13)(cid:13) However, in [4] it is shown that the equivalences between (1) and (4) and between (1) and (3) in Fact 4.3 are false. Therefore, some of the considered implications remain unproven. On the other hand, a characterization of norm-preserving ex- tendability of a compact operators is provided. Indeed, the following result holds. = kT k. Theorem 4.4 (Theorem 5.3 in [4]). For an infinite-dimensional Banach space X, the following assertions are equivalent. (1) X is a Lindenstrauss space such that each D(x) (x ∈ SX ) is finite-dimensional. (2) For every Banach spaces Y ⊂ Z, every compact operator T : Y → X admits a compact norm-preserving extension T : Z → X. Since the set D(x) is finite-dimensional if and only if ext D(x) is finite (see Remark 5.2 in [4]), Theorem 4.1 gives a correct version of the result of Lazar quoted in Fact 4.3. Theorem 4.5. Let X be an infinite-dimensional Banach space. The following assertions are equivalent: (1) X is a Lindenstrauss space enjoying property (pol-iv); (2) X is a Lindenstrauss space enjoying property (pol-v); (3) X is a Lindenstrauss space enjoying property (pol-vi); (4) For every Banach spaces Y ⊂ Z, every compact operator T : Y → X admits a compact norm-preserving extension T : Z → X. It is worth to pointing out that the original statement of Lazar (see Fact 4.3) has a correct version where two separate groups of equivalent properties are recognized. Namely, Theorem 4.3 in [4] shows that properties (1) and (2) are equivalent, while the equivalence between (3) and (4) follows from Theorem 5.3 in [4] and Theorem 4.5 above. Now, we prove that none the one-side implications of Theorem 4.1 can be re- versed. In [6, 8] there are many examples proving that the considered implications cannot be reversed when a general Banach space is considered, but most of them are not Lindenstrauss spaces. On the other hand, the following examples show that the implications cannot be reversed even if we restrict our attention to the class of Lindenstrauss spaces. All of them are based on suitable hyperplanes of the space c of the convergent sequences. Let α = (α(1), α(2), . . .) ∈ Bℓ1, we define the space Wα =(x = (x(1), x(2), . . .) ∈ c : lim i→∞ x(i) = α(i)x(i)) . +∞ Xi=1 A detailed study of this class of spaces was developed in [2] and in Section 2 of [3]. Here we recall only that Wα is a predual of ℓ1 and that the standard basis {e∗ n} of ℓ1 is σ(ℓ1, Wα)-convergent to α for every α ∈ Bℓ1. Example 4.2. [(pol − ii) ; (pol − i)] Let α = ( r Then the standard basis {e∗ 2 , − 1 Example 4.3. [(pol − iii) ; (pol − ii)] Let α = (− 1 to see that Wα satisfies (pol-iii), but the standard basis {e∗ to α. 4 , − 1 8 , . . .) ∈ ℓ1. It is easy n} is σ(ℓ1, Wα)-convergent 2 , 0, 0, . . .) ∈ ℓ1 for 0 < r < 1. n} is σ(ℓ1, Wα)-convergent to α. 2 , r WEAK∗-FPP IN ℓ1 AND POLYHEDRALITY IN LINDENSTRAUSS SPACES 11 Example 4.4. [(pol − iv) ; (pol − iii)] Let α = ( 1 considering the set C = {e∗ (pol − iii). However Wα satisfies (pol-iv). 16 , . . .) ∈ ℓ1. By 5, . . .}, it is easy to recognize that Wα fails property 8 , − 1 3, e∗ 1, e∗ 2 , − 1 4 , 1 Example 4.5. [(pol − vii) ; (pol − iv)] Let α = ( 1 8 , . . .) ∈ ℓ1. A detailed study of the properties of Wα is carried out in [4]. Here, we recall only that there we proved that Wα satisfies (pol − vii) (and hence it enjoys also (pol − K)) but it lacks (pol − iv). 2 , 1 4 , 1 References [1] B. Brosowski, F. Deutsch. On some geometric properties of suns. J. Approx. Theory 10 (1974), 245-267. [2] E. Casini, E. Miglierina, Ł. Piasecki. Hyperplanes in the space of convergent sequences and preduals of ℓ1. Canad. Math. Bull. 58 (2015), 459-470. [3] E. Casini, E. Miglierina, Ł. Piasecki. Separable Lindenstrauss spaces whose duals lack the weak∗ fixed point property for nonexpansive mappings. To appear in Studia Math. [4] E. Casini, E. Miglierina, Ł. Piasecki and L. Veselý. Rethinking polyhedrality for Lindenstrauss spaces. Israel J. Math. 216 (2016), 355-369. [5] T. Domínguez Benavides, J. García Falset, M. A. Japón Pineda. The τ -fixed point property for nonexpansive mappings. Abstr. Appl. Anal. 3 (1998), 343-362. [6] R. Durier, P. L. Papini. Polyhedral norms in an infinite dimensional space. Rocky Mt. J. Math. 23 (1993), 863-875. [7] V.P. Fonf, J. Lindenstrauss, R.R. Phelps. Infinite dimensional convexity. In Handbook of the Geometry of Banach Spaces 1 (eds.: W.B. Johnson and J. Lindenstrauss), Elsevier Science, 2001, 599-670. [8] V. P. Fonf, L. Veselý. Infinite-dimensional polyhedrality. Canad. J. Math. 56 (2004), 472-494. [9] A. Gleit, R. McGuigan. A note on polyhedral Banach spaces. Proc. Amer. Math. Soc. 33 (1972), 398-404. [10] K. Goebel, W.A. Kirk. Topics in metric fixed point theory. Cambridge Univ. Press. 1990. [11] L. A. Karlovitz. On nonexpansive mappings. Proc. Amer. Math. Soc. 55 (1976), 321-325. [12] V. Klee. Some characterizations of convex polyhedra, Acta Math. 102 (1959), 79-107. [13] V. Klee. Polyhedral sections of convex bodies. Acta Math. 103 (1960), 243-267. [14] A. J. Lazar. Polyhedral Banach spaces and extensions of compact operators. Israel J. Math. 7 (1969), 357-364. [15] M. A. Japón-Pineda, S. Prus. Fixed point property for general topologies in some Banach spaces. Bull. Austral. Math. Soc. 70 (2004), 229-244. [16] A. J. Lazar, J. Lindenstrauss. Banach spaces whose duals are L1-spaces and their representing matrices. Acta Math. 126 (1971), 165-194. [17] J. Lindenstrauss. Extension of compact operators, Mem. Amer. Math. Soc. 48 (1964). [18] T. C. Lim. Asymptotic centers and nonexpansive mappings in conjugate Banach spaces. Pacific J. Math. 90 (1980), 135-143. [19] P. H. Maserick. Convex polytopes in linear spaces. Illinois J. Math. 9 (1965), 623-635. [20] P. M. Soardi. Schauder basis and fixed points of nonexpansive mappings. Pacific. J. Math. 101 (1982), 193-198. Emanuele Casini, Dipartimento di Scienza e Alta Tecnologia, Università dell'Insubria, via Valleggio 11, 22100 Como, Italy. E-mail address: [email protected] Enrico Miglierina, Dipartimento di Discipline Matematiche, Finanza Matematica ed Econome- tria, Università Cattolica del Sacro Cuore, Via Necchi 9, 20123 Milano, Italy. E-mail address: [email protected] Łukasz Piasecki, Instytut Matematyki, Uniwersytet Marii Curie-Skłodowskiej, Pl. Marii Curie-Skłodowskiej 1, 20-031 Lublin, Poland. E-mail address: [email protected] Roxana Popescu, Department of Mathematics, University of Pittsburgh, Pittsburgh, PA 15260, USA. E-mail address: [email protected]
1711.06163
4
1711
2018-03-26T08:15:09
A breakdown of injectivity for weighted ray transforms in multidimensions
[ "math.FA", "math.CA" ]
We consider weighted ray-transforms $P\_W$ (weighted Radon transforms along straight lines) in $\mathbb{R}^d, \, d\geq 2,$ with strictly positive weights $W$. We construct an example of such a transform with non-trivial kernel in the space of infinitely smooth compactly supported functions on $\mathbb{R}^d$. In addition, the constructed weight $W$ is rotation-invariant continuous and is infinitely smooth almost everywhere on $\mathbb{R}^d \times \mathbb{S}^{d-1}$. In particular, by this construction we give counterexamples to some well-known injectivity results for weighted ray transforms for the case when the regularity of $W$ is slightly relaxed. We also give examples of continous strictly positive $W$ such that $\dim \ker P\_W \geq n$ in the space of infinitely smooth compactly supported functions on $\mathbb{R}^d$ for arbitrary $n\in \mathbb{N}\cup \{\infty\}$, where $W$ are infinitely smooth for $d=2$ and infinitely smooth almost everywhere for $d\geq 3$.
math.FA
math
A breakdown of injectivity for weighted ray transforms in multidimensions F.O. Goncharov∗ R. G. Novikov∗† September 24, 2018 Abstract We consider weighted ray-transforms PW (weighted Radon transforms along straight lines) in Rd, d ≥ 2, with strictly positive weights W . We construct an example of such a transform with non-trivial kernel in the space of infinitely smooth compactly supported functions on Rd. In addition, the constructed weight W is rotation-invariant continuous and is infinitely smooth almost everywhere on Rd × Sd−1. In particular, by this construction we give counterexamples to some well-known injectivity results for weighted ray transforms for the case when the regularity of W is slightly relaxed. We also give examples of continous strictly positive W such that dim ker PW ≥ n in the space of infinitely smooth compactly supported functions on Rd for arbitrary n ∈ N ∪ {∞}, where W are infinitely smooth for d = 2 and infinitely smooth almost everywhere for d ≥ 3. Keywords: Radon transforms, ray transforms, integral geometry, injectivity, non-injectivity AMS Mathematics Subject Classification: 44A12, 53C65, 65R32 1 Introduction We consider the weighted ray transforms PW defined by PW f (x, θ) =ZR T Sd−1 = {(x, θ) ∈ Rd × Sd−1 : xθ = 0}, W (x + tθ, θ)f (x + tθ) dt, (x, θ) ∈ T Sd−1, d ≥ 2, (1.1) (1.2) where f = f (x), W = W (x, θ), x ∈ Rd, θ ∈ Sd−1. Here, W is the weight, f is a test function on Rd. In addition, we interpret T Sd−1 as the set of all rays in Rd. As a ray γ we understand a straight line with fixed orientation. If γ = γ(x, θ), (x, θ) ∈ T Sd−1, then γ(x, θ) = {y ∈ Rd : y = x + tθ, t ∈ R} (up to orientation), where θ gives the orientation of γ. We assume that W = W ≥ c > 0, W ∈ L∞(Rd × Sd−1), where W denotes the complex conjugate of W , c is a constant. Note also that PW f (x, θ) =Zγ W (x, γ)f (x) dx, γ = γ(x, θ), where W (x, γ) = W (x, θ) for x ∈ γ, γ = γ(x, θ), (x, θ) ∈ T Sd−1. ∗CMAP, Ecole Polytechnique, CNRS, Universit´e Paris-Saclay, 91128, Palaiseau, France; email: [email protected] †IEPT RAS, 117997 Moscow, Russia; email: [email protected] (1.3) (1.4) (1.5) (1.6) 1 The aforementioned transforms PW arise in various domains of pure and applied [Sh92], mathematics; see [LB73], [Kun92], [BQ93], [B93], [Sh93], [KLM95], [Pa96], [ABK98], [Na01], [N02a], [N02b], [BS04], [Bal09], [Gi10], [BJ11], [PG13], [N14], [I16], [Ng17] and references therein. [MQ85], [BQ87], [Fi86], [TM80], [Q83], [Be84], In particular, the related results are the most developed for the case when W ≡ 1. In this case PW is reduced to the classical ray-transform P (Radon transform along straight lines). The transform P arises, in particular, in the X-ray transmission tomography. We refer to [R17], [J38], [C64], [GGG82], [H01], [Na01] and references therein in connection with basic results for this classical case. At present, many important results on transforms PW with other weights W satis- fying (1.4) are also known; see the publications mentioned above with non-constant W and references therein. In particular, assuming (1.4) we have the following injectivity results. Injectivity 1 (see [Fi86]). Suppose that d ≥ 3 and W ∈ C 2(Rd × Sd−1). Then PW is injective on Lp 0 denotes compactly supported functions from Lp. 0(Rd) for p > 2, where Lp Injectivity 2 (see [MQ85]). Suppose that d = 2, W ∈ C 2(R2 × S1) and 0 < c0 ≤ W, kWkC2(R2×S1) ≤ N, (1.7) for some constants c0 and N. Then, for any p > 2, there is δ = δ(c0, N, p) > 0 such that PW is injective on Lp(B(x0, δ)) for any x0 ∈ R2, where Lp(B(x0, δ)) = {f ∈ Lp(R2) : supp f ⊂ B(x0, δ)}, B(x0, δ) = {x ∈ R2 : x − x0 ≤ δ}. Injectivity 3 (see [Q83]). Suppose that d = 2, W ∈ C 1(R2 × S1) and W is rotation invariant (see formula (2.18) below). Then PW is injective on Lp 0(R2) for p ≥ 2. In a similar way with [Q83], we say that W is rotation invariant if and only if W (x, γ) is independent of the orientation of γ, W (x, γ) = W (Ax, Aγ) for x ∈ γ, γ ∈ T Sd−1, A ∈ O(d), (1.8) where T Sd−1 is defined in (1.2), O(d) denotes the group of orthogonal transformations of Rd. Note also that property (1.8) can be rewritten in the form (2.18), (2.19) or (2.20), (2.21); see Section 2. Injectivity 4 (see [BQ87]). Suppose that d = 2, W is real-analytic on R2× S1. Then PW is injective on Lp 0(R2) for p ≥ 2. Injectivity 1 is a global injectivity for d ≥ 3. Injectivity 2 is a local injectivity for d = 2. Injectivity 3 is a global injectivity for d = 2 for the rotation invariant case. Injectivity 4 is a global injectivity for d = 2 for the real-analytic case. The results of Injectivity 1 and Injectivity 2 remain valid with C α, α > 1, in place of C 2 in the assumptions on W ; see [I16]. Injectivity 1 follows from Injectivity 2 in the framework of the layer-by-layer re- construction approach. See [Fi86], [N02a], [I16] and references therein in connection with the layer-by-layer reconstruction approach for weighted and non-abelian ray transforms in dimension d ≥ 3. 2 The work [B93] gives a counterexample to Injectivity 4 for PW in C ∞ 0 (R2) for the case when the assumption that W is real-analytic is relaxed to the assumption that W is infinitely smooth, where C ∞ 0 denotes infinitely smooth compactly supported functions. In somewhat similar way with [B93], in the present work we obtain counterexamples to Injectivity 1, Injectivity 2 and Injectivity 3 for the case when the regularity of W is slightly relaxed. In particular, by these counterexamples we continue related studies of [MQ85], [B93] and [GN17]. More precisely, in the present work we construct W and f such that where W satisfies (1.4), W is rotation-invariant (i.e., satisfies (1.8)), PW f ≡ 0 on T Sd−1, d ≥ 2, W is infinitely smooth almost everywhere on Rd × Sd−1 and W ∈ C α(Rd × Sd−1), at least, for any α ∈ (0, α0), where α0 = 1/16; f is a non-zero spherically symmetric infinitely smooth and compactly supported function on Rd; (1.9) (1.10) (1.11) see Theorem 1 of Section 3. These W and f directly give the aforementioned counterexamples to Injectivity 1 and Injectivity 3. Our counterexample to Injectivity 1 is of particular interest (and is rather surpris- ing) in view of the fact that the problem of finding f on Rd from PW f on T Sd−1 for known W is strongly overdetermined for d ≥ 3. Indeed, dim Rd = d, dim T Sd−1 = 2d − 2, d < 2d − 2 for d ≥ 3. This counterexample to Injectivity 1 is also rather surprising in view of the aforemen- tioned layer-by-layer reconstruction approach in dimension d ≥ 3. Our counterexample to Injectivity 3 is considerably stronger than the preceeding counterexample of [MQ85], where W is not yet continuous and is not yet strictly positive (i.e., is not yet separated from zero by a positive constant). Using our W and f of (1.10), (1.11) for d = 3 we also obtain the aforementioned counterexample to Injectivity 2; see Corollary 1 of Section 3. Finally, in the present work we also give examples of W satisfying (1.4) such that 0 (Rd) for arbitrary n ∈ N ∪ {∞}, where W ∈ C ∞(R2 × S1) for dim ker PW ≥ n in C ∞ d = 2 and W satisfy (1.10) for d ≥ 3; see Theorem 2 of Section 3. To our knowledge, examples of W satisfying (1.4), where dim ker PW ≥ n (for example in L2 0(Rd)) were not yet given in the literature even for n = 1 in dimension d ≥ 3 and even for n = 2 in dimension d = 2. In the present work we adopt and develop considerations of the famous work [B93] and of our very recent work [GN17]. In Section 2 we give some preliminaries and notations. Main results are presented in detail in Sections 3. Related proofs are given in Sections 4-9. 3 2 Some preliminaries Notations. Let Ω = Rd × Sd−1, r(x, θ) = x − (xθ)θ, (x, θ) ∈ Ω, Ω0(δ) = {(x, θ) ∈ Ω : r(x, θ) > δ}, Ω1(δ) = Ω\Ω0(δ) = {(x, θ) ∈ Ω : r(x, θ) ≤ δ}, δ > 0, Ω(Λ) = {(x, θ) ∈ Rd × Sd−1 : r(x, θ) ∈ Λ}, Λ ⊂ [0, +∞), T0(δ) = {(x, θ) ∈ T Sd−1 : x > δ}, T1(δ) = {(x, θ) ∈ T Sd−1 : x ≤ δ}, δ > 0, T (Λ) = {(x, θ) ∈ T Sd−1 : x ∈ Λ}, Λ ⊂ [0, +∞), Jr,ε = (r − ε, r + ε) ∩ [0, +∞), r ∈ [0, +∞), ε > 0. (2.1) (2.2) (2.3) (2.4) (2.5) (2.6) (2.7) (2.8) (2.9) The set T0(δ) in (2.6) is considered as the set of all rays in Rd which are located at distance greater than δ from the origin. The set T1(δ) in (2.7) is considered as the set of all rays in Rd which are located at distance less or equal than δ. We also consider the projection (2.10) (2.11) (2.12) In addition, r(x, θ) of (2.2) is the distance from the origin {0} ∈ Rd to the ray π : Ω → T Sd−1, π(x, θ) = (πθx, θ), (x, θ) ∈ Ω, πθx = x − (xθ)θ. γ = γ(π(x, θ)) (i.e., r(x, θ) = πθx). The rays will be also denoted by γ = γ(x, θ) def = γ(π(x, θ)), (x, θ) ∈ Ω. PW f (x, θ) = PW f (π(x, θ)) for (x, θ) ∈ Ω. We also consider We also define B(x0, δ) = {x ∈ Rd : x − x0 < δ}, B(x0, δ) = {x ∈ Rd : x − x0 ≤ δ}, x0 ∈ Rd, δ > 0, B = B(0, 1), B = B(0, 1). (2.13) (2.14) (2.15) (2.16) For a function f on Rd we denote its restriction to a subset Σ ⊂ Rd by fΣ. By C0, C ∞ 0 we denote continuous compactly supported and infinitely smooth com- pactly supported functions, respectively. By C α(Y ), α ∈ (0, 1), we denote the space of α-Holder functions on Y with the norm: kukC α(Y ) = kukC(Y ) + kuk′ kuk′ C α(Y ) = sup y1,y2∈Y y1−y2≤1 C α(Y ), u(y1) − u(y2) y1 − y2α , (2.17) where kukC(Y ) denotes the maximum of u on Y . 4 Rotation invariancy. Using formula (1.6), for positive and continous W , property (1.8) can be rewritten in the following equivalent form: W (x, θ) = U(x − (xθ)θ, xθ), x ∈ Rd, θ ∈ Sd−1, for some positive and continuous U such that U(r, s) = U(−r, s) = U(r,−s), r ∈ R, s ∈ R. In addition, symmetries (2.18), (2.19) of W can be also written as W (x, θ) = U(x, xθ), (x, θ) ∈ Ω, U(r, s) = U (−r, s) = U (r,−s), r ∈ R, s ∈ R. (2.21) where U is positive and continuous on R×R. Using the formula x2 = xθ2 +r2(x, θ), one can see that symmetries (2.18), (2.19) and symmetries (2.20), (2.21) of W are equivalent. (2.18) (2.19) (2.20) i=1 be an open locally-finite cover of M. Then there exists a C ∞-smooth locally-finite partition of unity {ψi}∞ Partition of unity. We recall the following classical result (see, e.g., Theorem 5.6 in [M92]): Let M be a C ∞-manifold, which is Hausdorff and has a countable base. Let also {Ui}∞ i=1 on M, such that (2.22) In particular, any open interval (a, b) ⊂ R and Ω satisfy the conditions for M of supp ψi ⊂ Ui. this statement. It will be used in Subsection 3.1. 3 Main results Theorem 1. There exist a weight W satisfying (1.4) and a non-zero function f ∈ C ∞ 0 (Rd), d ≥ 2, such that PW f ≡ 0 on T Sd−1, In addition, W is rotation invariant, i.e., satisfies (3.1) where PW is defined in (1.1). (2.18), and f is spherically symmetric with supp f ⊆ B. Moreover, W ∈ C ∞(Ω\Ω(1)), W ∈ C α(Rd × Sd−1) for any α ∈ (0, α0), α0 = 1/16, W ≥ 1/2 on Ω and W ≡ 1 on Ω([1, +∞)), W (x, θ) ≡ 1 for x ≥ R > 1, θ ∈ Sd−1, (3.2) (3.3) (3.4) (3.5) where Ω, Ω(1), Ω([1, +∞)) are defined by (2.1), (2.5), R is a constant. The construction of W and f proving Theorem 1 is presented below in Subsec- tions 3.1, 3.2. In addition, this construction consists of its version in dimension d = 2 (see Subsection 3.1) and its subsequent extension to the case of d ≥ 3 (see Subsec- tion 3.2). Theorem 1 directly gives counterexamples to Injectivity 1 and Injectivity 3 of Introduction. Theorem 1 also implies the following counterexample to Injectivity 2 of Introduction: 5 Corollary 1. For any α ∈ (0, 1/16) there is N > 0 such that for any δ > 0 there are Wδ, fδ satisfying Wδ ≥ 1/2, Wδ ∈ C α(R2 × S1), kWδkC α(R2×S1) ≤ N fδ ∈ C ∞(R2), fδ 6≡ 0, supp fδ ⊆ B(0, δ), PWδfδ ≡ 0 on T S1. (3.6) (3.7) (3.8) The construction of Wδ, fδ proving Corollary 1 is presented in Subsection 5.1. Theorem 2. For any n ∈ N ∪ {∞} there exists a weight Wn satisfying (1.4) such that (3.9) dim ker PWn ≥ n in C ∞ 0 (Rd), d ≥ 2, where PW is defined in (1.1). Moreover, Wn ∈ C ∞(R2 × S1) for d = 2, Wn is infinitely smooth almost everywhere on Rd × Sd−1 and Wn ∈ C α(Rd × Sd−1), α ∈ (0, 1/16) for d ≥ 3, Wn(x, θ) ≡ 1 for x ≥ R > 1, θ ∈ Sd−1 for n ∈ N, d ≥ 2, (3.10) (3.11) (3.12) where R is a constant. The construction of Wn proving Theorem 2 is presented in Section 4. In this con- struction we proceed from Theorem 1 of the present work for d ≥ 3 and from the result of [B93] for d = 2. In addition, for this construction it is essential that n < +∞ in (3.12). 3.1 Construction of f and W for d = 2 In dimension d = 2, the construction of f and W adopts and develops considerations of [B93] and [GN17]. In particular, we construct f , first, and then W (in this con- struction we use notations of Section 2 for d = 2). In addition, this construction is commented in Remarks 1-5 below. Construction of f . The function f is constructed as follows: f = fk k! , ∞Xk=1 fk(x) = efk(x) = Φ(2k(1 − x)) cos(8kx2), x ∈ R2, k ∈ N, for arbitrary Φ ∈ C ∞(R) such that supp Φ = [4/5, 6/5], 0 < Φ(t) ≤ 1 for t ∈ (4/5, 6/5), Φ(t) = 1, for t ∈ [9/10, 11/10], Φ monotonously increases on [4/5, 9/10] and monotonously decreases on [11/10, 6/5]. (3.13) (3.14) (3.15) (3.16) (3.17) (3.18) Properties (3.15), (3.16) imply that functions efk (and functions fk) in (3.14) have disjoint supports: suppefi ∩ suppefj = ∅ if i 6= j, suppefk = [1 − 2−k(cid:18) 6 5(cid:19) , 1 − 2−k(cid:18)4 5(cid:19)], i, j, k ∈ N. This implies the convergence of series in (3.13) for every fixed x ∈ R2. (3.19) 6 Lemma 1. Let f be defined by (3.13)-(3.17). Then f is spherically symmetric, f ∈ C ∞ fγ has non-constant sign. 0 (R2) and supp f ⊆ B. In addition, if γ ∈ T S1, γ ∩ B 6= ∅, then fγ 6≡ 0 and Lemma 1 is similair to Lemma 1 of [GN17] and it is, actually, proved in Section 4.1 of [GN17]. Remark 1. Formulas (3.13)-(3.17) for f are similar to the formulas for f in [B93], where PW was considered in R2, and also to the formulas for f in [GN17], where the weighted Radon transform RW along hyperplanes was considered in R3. The only difference between (3.13)-(3.17) and the related formulas in [GN17] is the dimension d = 2 in (3.13)-(3.17) instead of d = 3 in [GN17]. At the same time, the important difference between (3.13)-(3.17) and the related formulas in [B93] is that in formula (3.14) the factor cos(8kx2) depends only on x, whereas in [B93] the corresponding factor is cos(3kφ) which depends only on the angle φ in the polar coordinates in R2. In a similar way with [B93], [GN17], we use the property that the restriction of the function cos(8kx2) to an arbitrary ray γ intersecting the open ball oscillates sufficiently fast (with change of the sign) for large k. Construction of W . In our example W is of the following form: W (x, θ) = φ1(x) NXi=0 ξi(r(x, θ))Wi(x, θ)! + φ2(x) = φ1(x) ξ0(r(x, θ))W0(x, θ) + ξi(r(x, θ))Wi(x, θ)! + φ2(x), (x, θ) ∈ Ω, NXi=1 (3.20) where φ1 = φ1(x), φ2 = φ2(x) is a C ∞-smooth partition of unity on R2 such that, φ1 ≡ 0 for x ≥ R > 1, φ1 ≡ 1 for x ≤ 1, φ2 ≡ 0 for x ≤ 1, {ξi(s), s ∈ R}N ξi(s) = ξi(−s), s ∈ R, i = 0, N, Wi(x, θ) are bounded, continuous, strictly positive and rotation invariant (according to (2.18)), (2.21) on i=0 is a C ∞- smooth partition of unity on R, (3.21) (3.22) (3.23) (3.24) the open vicinities of supp ξi(r(x, θ)), i = 0, N, respectively. From the result of Lemma 1 and from (3.21) it follows that PW f (x, θ) = ξ0(x)PW0f (x, θ) + where W is given by (3.20). NXi=1 ξi(x)PWif (x, θ), (x, θ) ∈ T S1, (3.25) From (3.20)-(3.24) it follows that W of (3.20) satisfies the conditions (1.4), (2.20), (2.21). 7 The weight W0 is constructed in next paragraph and has the following properties: W0 is bounded, continuous and rotation invariant on Ω(1/2, +∞), W0 ∈ C ∞(Ω ((1/2, 1) ∪ (1, +∞))) and W0 ∈ C α(Ω(1/2, +∞)) for α ∈ (0, 1/16), there exists δ0 ∈ (1/2, 1) such that: W0(x, θ) ≥ 1/2 if r(x, θ) > δ0, if r(x, θ) ≥ 1, W0(x, θ) = 1 PW0f (x, θ) = 0 on Ω((1/2, +∞)), (3.26) (3.27) (3.28) (3.29) where PW0 is defined according to (1.1) for W = W0, f is given by (3.13), (3.14). In addition, supp ξ0 ⊂ (−∞,−δ0) ∪ (δ0, +∞), ξ0(s) = 1 for s ≥ 1, where δ0 is the number of (3.28). In particular, from (3.28), (3.30) it follows that W0(x, θ)ξ0(r(x, θ)) > 0 if ξ0(r(x, θ)) > 0. In addition, ξi(r(x, θ))Wi(x, θ) are bounded, rotation invariant and C ∞ on Ω, Wi(x, θ) ≥ 1/2 if ξi(r(x, θ)) 6= 0, PWif (x, θ) = 0 on (x, θ) ∈ T S1, such that ξi(r(x, θ)) 6= 0, i = 1, N, (x, θ) ∈ Ω. (3.30) (3.31) (3.32) (3.33) (3.34) (3.35) Theorem 1 for d = 2 follows from Lemma 1 and formulas (3.20)-(3.29), (3.32)- i=0 are constructed in Subsection 3.1. Weights W1, . . . , WN of (3.20) and {ξi}N (3.35). We point out that the construction of W0 of (3.20) is substantially different from the construction of W1, . . . , WN . The weight W0 is defined for the rays γ ∈ T S1 which can be close to the boundary ∂B of B which results in restrictions on global smoothness of W0. Remark 2. The construction of W summarized above in formulas (3.20)-(3.35) arises in the framework of finding W such that PW f ≡ 0 on T S1 for f defined in (3.13)-(3.18), (3.36) under the condition that W is strictly positive, sufficiently regular and rotation in- variant (see formulas (1.4), (2.18), (2.19)). In addition, the weights Wi, i = 0, . . . , N, in (3.20) are constructed in a such a way that PWif = 0 on Vi, i = 0, . . . , N, (3.37) under the condition that Wi = Wi(x, γ) are strictly positive, sufficiently regular and rotation invariant for x ∈ γ, γ ∈ Vi ⊂ T S1, i = 0, . . . , N, where {Vi}N i=0 is an open cover of T S1 and V0 = T0(δ0), Vi = T (Λi) for some open Λi ⊂ R, i = 0, . . . , N, (3.38) (3.39) where T0 is defined in (2.6), δ0 is the number of (3.28), T (Λ) is defined in (2.8). In addition, the functions ξi, i = 0, . . . , N, in (3.20) can be interpreted as a partition of unity on T S1 subordinated to the open cover {Vi}N i=0. The aforementioned con- struction of W is a two-dimensional analog of the construction developed in [GN17], 8 where the weighted Radon transform RW along hyperplanes was considered in R3. At the same time, the construction of W of the present work is similar to the construc- tion in [B93] with the important difference that in the present work f is spherically symmetric and W, Wi, i = 0, . . . , N, are rotation invariant. Construction of W0. Let {ψk}∞ k=1 be a C ∞ partition of unity on (1/2, 1) such that supp ψk ⊂ (1 − 2−k+1, 1 − 2−k−1), k ∈ N, first derivatives ψ′ k satisfy the bounds: supψ′ k ≤ C2k, (3.40) (3.41) where C is a positive constant. Actually, functions {ψk}∞ were used in considerations of [B93]. k=1 satisfying (3.40), (3.41) Note that Therefore, 1 − 2−(k−2)−1 < 1 − 2−k(6/5), k ≥ 3. for all s0, t0 ∈ R, s0 ∈ supp ψk−2, t0 ∈ supp Φ(2k(1 − t)) ⇒ s0 < t0, k ≥ 3. Weight W0 is defined by the following formulas (3.42) (3.43) W0(x, θ) = G(x, θ) =Zγ(x,θ) 1 − G(x, θ) 1, r(x, θ) ≥ 1 ∞Pk=3 f (y) dy, Hk(x, θ) =Zγ(x,θ) k!fk(x) ψk−2(r(x, θ)) Hk(x, θ) , 1/2 < r(x, θ) < 1, , (3.44) k (y) dy, x ∈ R2, θ ∈ S1, f 2 (3.45) where f, fk are defined in (3.13), (3.14), respectively, rays γ(x, θ) are given by (2.13). Formula (3.44) implies that W0 is defined on Ω0(1/2) ⊂ Ω. Due to (3.14)-(3.17), (3.40), (3.43), in (3.45) we have that Hk(x, θ) 6= 0 if ψk−2(r(x, θ)) 6= 0, (x, θ) ∈ Ω, ψk−2(r(x, θ)) Hk(x, θ) ∈ C ∞(Ω(1/2, 1)), (3.46) (3.47) where r(x, θ) is defined in (2.2), Ω, Ω(·) are defined in (2.1), (2.5), d = 2. Also, for any fixed (x, θ) ∈ Ω, 1/2 < r(x, θ), the series in the right hand-side of (3.44) has only a finite number of non-zero terms (in fact, no more than two) and, hence, the weight W0 is well-defined. By the spherical symmetry of f , functions G, Hk in (3.44) are of the type (2.18) (and (2.20)). Therefore, W0 is rotation invariant (in the sense of (2.18) and (2.20)). Actually, formula (3.29) follows from (3.13), (3.14), (3.44), (3.45) (see Subsec- tion 6.2 for details). Using the construction of W0 and the assumption that r(x, θ) > 1/2 one can see that W0 is C ∞ on its domain of definition, possibly, except points with r(x, θ) = 1. Note also that due to (3.13), (3.14), the functions fk, G, Hk, used in (3.44), (3.45) can be considered as functions of one-dimensional arguments. Formulas (3.26)-(3.28) are proved in Subsection 6.1. Remark 3. Formulas (3.44), (3.45) given above for the weight W0 are considered for the rays from T0(δ0) (mentioned in Remark 2) and, in particular, for rays close to the tangent rays to ∂B. These formulas are direct two-dimensional analogs of the related formulas in [GN17]. At the same time, formulas (3.44), (3.45) are similar to the related formulas in [B93] with the important difference that f, fk are spherically symmetric in the present work and, as a corollary, W0 is rotation invariant. Also, in 9 a similar way with [B93], [GN17], in the present work we show that G(x, θ) tends to zero sufficiently fast as r(x, θ) → 1. This is a very essential point for continuity of W0 and it is given in Lemma 3 of Subsection 6.1. Construction of W1, . . . , WN and ξ0, . . . , ξN Lemma 2. Let f ∈ C ∞ 0 (R2) be spherically symmetric, (x0, θ0) ∈ T S1, fγ(x0,θ0) 6≡ 0 and fγ(x0,θ0) changes the sign. Then there exist ε0 > 0 and weight W(x0,θ0),ε0 such that f = 0 on Ω(Jr(x0,θ0),ε0), PW(x0,θ0),ε0 W(x0,θ0),ε0 is bounded, infinitely smooth, strictly positive and rotation invariant on Ω(Jr(x0,θ0),ε0), (3.48) (3.49) Lemma 2 is proved in Section 7. This lemma is a two-dimensional analog of the where Ω(Jr,ε0),Jr,ε0 are defined in (2.5) and (2.9), respectively. related lemma in [GN17]. Remark 4. In Lemma 2 the construction of W(x0,θ0),ε0 arises from 1. finding strictly positive and regular weight W(x0,θ0),ε on the rays γ = γ(x, θ) with fixed θ = θ0, where r(x, θ0) ∈ Jr(x0,θ0),ε for some ε > 0, such that (3.48) holds for θ = θ0 and under the condition that W(x0,θ0),ε(y, γ) = W(x0,θ0),ε(yθ0, γ), y ∈ γ = γ(x, θ0), r(x, θ0) ∈ Jr(x0,θ0),ε; (3.50) 2. extending Wr(x0,θ0),ε to all rays γ = γ(x, θ), r(x, θ) ∈ Jr(x0,θ0),ε, θ ∈ S1, via for- We recall that r(x, θ) is defined in (2.2). mula (1.8). Let f be the function of (3.13), (3.14). Then, using Lemmas 1, 2 one can see that for all δ ∈ (0, 1) there exist {Ji = Jri,εi, Wi = W(xi,θi),εi}N such that Ji, i = 1, N, is an open cover of [0, δ] in R, and Wi satisfy (3.48) and (3.49) on Ω(Ji), respectively. i=1 (3.51) Actually, we consider (3.51) for the case of δ = δ0 of (3.28). Note that in this case {Ω(Ji)}N To the set Ω0(δ0) we associate the open set i=1 for Ji of (3.51) is the open cover of Ω1(δ0). (3.52) Therefore, the collection of intervals {±Ji, i = 0, N} is an open cover of R, where −Ji is the symmetrical reflection of Ji with respect to {0} ∈ R. J0 = (δ0, +∞) ⊂ R. We construct the partition of unity {ξi}N i=0 as follows: ξi(s) = ξi(s) = supp ξi ⊂ Ji ∪ (−Ji), i = 0, N, ( ξi(s) + ξi(−s)), s ∈ R, 1 2 (3.53) (3.54) i=0 is a partition of unity for the open cover {Ji ∪ (−Ji)}N where { ξi}N Partition of unity, for Ui = Ji ∪ (−Ji)). symmetry of Ji ∪ (−Ji), i = 1, N, choice of J0 in (3.52) and from (3.53). (see the proof of Lemma 2 and properties (3.51) in Section 7 for details). Properties (3.30), (3.54) follow from (2.22) for { ξi}N i=0 with Ui = Ji ∪ (−Ji), the In turn, (3.31) follows from (3.52) and the construction of Ji, i = 1, N, from (3.51) i=0 (see Section 2, Properties (3.33)-(3.35) follow from (3.51) for δ = δ0 and from (3.52)-(3.54). This completes the description of W1, . . . , WN and {ξi}N i=0. Remark 5. We have that Ji = Λi, i = 1, . . . , N, where Λi are the intervals in formula (3.39) of Remark 2 and Ji are the intervals considered in (3.51), (3.52). 10 3.2 Construction of W and f for d ≥ 3 Consider f and W of Theorem 1, for d = 2, constructed in Subsection 3.1. For these f and W consider f and U such that Proposition 1. Let W and f , for d ≥ 3, be defined as f (x) = ef (x), W (x, θ) = U (x,xθ), x ∈ R2, θ ∈ S1. W (x, θ) = U (x,xθ), (x, θ) ∈ Rd × Sd−1, f (x) = f (x), x ∈ Rd, where U , f are the functions of (3.55). Then PW f ≡ 0 on T Sd−1. (3.55) (3.56) (3.57) (3.58) In addition, weight W satisfies properties (3.2)-(3.5), f is spherically symmetric in- finitely smooth and compactly supported on Rd, f 6≡ 0. Proposition 1 is proved in Subsection 5.2. This completes the proof of Theorem 1. 4 Proof of Theorem 2 4.1 Proof for d ≥ 3 Let W be the weight of Theorem 1 for d ≥ 3, R be the number in (3.5) for d ≥ 3, {yi}∞ for i 6= j, i, j ∈ N, {Bi}∞ The weight Wn is defined as follows i=1 be a sequence of vectors in Rd such that y1 = 0, yi − yj > 2R i=1 be the closed balls in Rd of radius R centered at yi (see (4.2), (4.3)). (4.4) (4.1) (4.2) (4.3) (4.5) Wn(x, θ) = nSi=1 Bi, 1 if x 6∈ W (x − y1, θ) = W (x, θ) if x ∈ B1, W (x − y2, θ) if x ∈ B2, ..., W (x − yk, θ) if x ∈ Bk, ..., W (x − yn, θ) if x ∈ Bn,  θ ∈ Sd−1, n ∈ N ∪ {∞}, d ≥ 3, where W is defined in (4.1), yi and Bi are defined in (4.3), (4.4), respectively. (3.2)-(3.5), (4.1), (4.2). Properties (1.4), (3.11) and (3.12) for Wn, defined in (4.5), for d ≥ 3, follow from Let f1(x) def = f (x), f2(x) def = f (x − y2), . . . , fn(x) def = f (x − yn), x ∈ Rd, d ≥ 3, (4.6) where yi are defined in (4.3) and f is the function of Theorem 1 for d ≥ 3. (4.7) 11 One can see that fi ∈ C ∞ 0 (Rd), d ≥ 3, fi 6≡ 0, supp fi ⊂ Bi, Bi ∩ Bj = ∅ for i 6= j, where Bi are defined in (4.4), i = 1, . . . , n. The point is that PWnfi ≡ 0 on T Sd−1, d ≥ 3, i = 1, . . . , n, fi are linearly independent in C ∞ 0 (Rd), d ≥ 3, i = 1, . . . , n, where Wn is defined in (4.5), fi are defined in (4.6). To prove (4.9) we use, in particular, the following general formula: W (y′ − y, θ)f (y′ − y)dy′ PWyfy(x, θ) = Zγ(x,θ) = Zγ(x−y,θ) W (y′, θ)f (y′)dy′ = PW f (x − y, θ), x ∈ Rd, θ ∈ Sd−1, Wy(x, θ) = W (x − y, θ), fy = f (x − y), x, y ∈ Rd, θ ∈ Sd−1. where W is an arbitrary weight satisfying (1.4), f is a test-function, γ(x, θ) is defined according to (2.13). Formula (4.9) follows from formula (3.1), definitions (4.5), (4.6), (4.7), properties (4.8) and from formulas (4.11), (4.12). Formula (4.10) follows from definitions (4.6), (4.7) and properties (4.8). This completes the proof of Theorem 2 for d ≥ 3. 4.2 Proof for d = 2 In [B93], there were constructed a weight W and a function f for d = 2, such that: PW f ≡ 0 on T S1, W = W ≥ c > 0, W ∈ C ∞(R2 × S1), f ∈ C ∞ 0 (R2), f 6≡ 0, suppf ⊂ B, where c is a constant, B is defined in (2.16). We define where W is the weight of (4.13), (4.14), c is a constant of (4.14). fW (x, θ) = c−1φ1(x)W (x, θ) + φ2(x), x ∈ R2, θ ∈ S1, φ1 = φ1(x), φ2 = φ2(x) is a C ∞-smooth partition of unity on R2 such that, φ1 ≡ 0 for x ≥ R > 1, φ1 ≡ 1 for x ≤ 1, φ1 ≥ 0 on R2, φ2 ≡ 0 for x ≤ 1, φ2 ≥ 0 on R2, where R is a constant. From (4.13)-(4.17) it follows that (4.8) (4.9) (4.10) (4.11) (4.12) (4.13) (4.14) (4.15) (4.16) (4.17) (4.18) (4.19) The proof of Theorem 2 for d = 2 proceeding from (4.15), (4.16), (4.18), (4.19) is completely similar to the proof of Theorem 2 for d ≥ 3, proceeding from Theorem 1. Theorem 2 is proved. PfW f ≡ 0 on T S1, fW ≥ 1,fW ∈ C ∞(R2 × S1), fW (x, θ) ≡ 1 for x ≥ R > 1, θ ∈ S1. 12 5 Proofs of Corollary 1 and Proposition 1 5.1 Proof of Corollary 1 Let Xr = {x1e1 + x2e2 + re3 : (x1, x2) ∈ R2}, 0 ≤ r < 1, S = X0 ∩ S2 = {(cos φ, sin φ, 0) ∈ R3 : φ ∈ [0, 2π)} ≃ S1. (5.1) (5.2) where (e1, e2, e3) is the standard orthonormal basis in R3. Without loss of generality we assume that 0 < δ < 1. Choosing r so that √1 − δ2 ≤ r < 1, we have that the intersection of the three dimensional ball B(0, 1) with Xr is the two-dimensional disk B(0, δ′), δ′ ≤ δ (with respect to the coordinates (x1, x2) induced by basis (e1, e2) on Xr). We define N, Wδ on R2 × S1 and fδ on R2 as follows: N = kWkC α(R3×S2), Wδ := WXr×S, fδ := fXr, for r = √1 − δ2, where W and f are the functions of Theorem 1 for d = 3. Due to (3.2)-(3.4), (5.3), (5.4) we have that Wδ ≥ 1/2, kWδkC α(R2×S1) ≤ N. (5.3) (5.4) (5.5) (5.6) Properties (5.6) imply (3.6). In view of Lemma 1 for the function f of Theorem 1, we have that fδ is spherically symmetric, fδ ∈ C ∞ 0 (B(0, δ′)), fδ 6≡ 0. Using (3.1), (5.4), (5.5) one can see that (3.8) holds. This completes the proof of Corollary 1. 5.2 Proof of Proposition 1 Let I(r) =Zγr U(y, r) f(y) dy, r ≥ 0, γr = γ(re2, e1), (5.7) where γ(x, θ) is defined by (1.3), (e1, . . . , ed) is the standard basis in Rd. Due to formula (3.1) of Theorem 1 for d = 2 and formulas (3.55), (5.7) we have that Next, using (1.1), (3.55), (5.8) we have also that I(r) = PW f (re2, e1) = 0 for r ≥ 0. (5.8) PW f (x, θ) = Zγ(x,θ) U (y,y − (yθ)θ) f (y) dy = I(x) = 0 for (x, θ) ∈ T Sd−1, (5.9) where γ(x, θ) is defined in (1.3). Formula (5.9) implies (3.58). Properties of W and f mentioned in Proposition 1 follow from properties (3.2)-(3.5) of W and of f of Theorem 1 for d = 2. This completes the proof of Proposition 1. 13 6 Proofs of formulas (3.26)-(3.29) 6.1 Proof of formulas (3.26)-(3.28) Lemma 3. Let W0 be defined by (3.44), (3.45). Then W0 admits the following rep- resentation: W0(x, θ) = U0(xθ,x − (xθ)θ), (x, θ) ∈ Ω((1/2, +∞)), ψk−2(r) U0(s, r) = ∞Pk=3 1 − eG(r) = Zγr ef (y) dy, eHk(r) eG(r) s ∈ R, x ∈ R2, γr is an arbitrary ray in T (r), r > 1/2, k!efk((s2 + r2)1/2) eHk(r) = Zγr ef 2 ∞Xk=1 efk k (y) dy, ef = 1, r ≥ 1 def def , 1/2 < r < 1, (6.1) , (6.2) , (6.3) k! where efk are defined by (3.14), T (r) is defined by (2.8), d = 2. In addition: U0 is infinitely smooth on R × {(1/2, 1) ∪ (1, +∞)}, U0(s, r) → 1 as r → 1 (uniformly in s ∈ R), U0(s, r) = 1 if s2 + r2 ≥ 1, 1 − U0(s, r) ≤ C0(1 − r)1/2 log4 2(cid:18) 1 1 − r(cid:19) , for s ∈ R, 1/2 < r < 1, U0(s, r) − U0(s′, r′) ≤ C1s − s′1/α + C1r − r′1/α, for α ∈ (0, 1/16), s, s′ ∈ R, r, r′ > 1/2, (6.4) (6.5) (6.6) (6.7) (6.8) where C0, C1 are positive constants depending on Φ of (3.15)-(3.17). Lemma 3 is proved Section 8. Lemma 3 implies (3.26)-(3.28) as follows. The continuity and rotation invariancy of W0 in (3.26) follow from (2.18), (2.19), (6.1), (6.8). Due to (3.40), (6.1), (6.2), (6.3) we have also that U0 admits a continuous extension to R × [1/2, +∞). (6.9) Properties (6.6), (6.9) imply the boundedness of W0 on Ω0(1/2), where Ω0(·) is defined in (2.3), d = 2. This completes the proof of (3.26). Formula (3.27) follows from (6.1), (6.4), (6.8) and from the fact that xθ, x − (xθ)θ Formula (3.28) follows from (3.26), (6.1), (6.2), (6.5), (6.6). This completes the proof of (3.26)-(3.28). are infinitely smooth functions on Ω0(1/2) and are Lipshitz in (x, θ) for x ∈ B(0, R), R > 1. 14 6.2 Proof of formula (3.29) From (1.1), (3.13)-(3.16), (3.40), (3.44), (3.45) it follows that: PW0f (x, θ) = Zγ(x,θ) = Zγ(x,θ) = Zγ(x,θ) f (y) dy − G(x, θ) f (y) dy − Zγ(x,θ) f (y) dy − Zγ(x,θ) ∞Xk=3 f (y) dy k!ψk−2(r(x, θ)) Rγ(x,θ) ψk−2(r(x, θ)) Rγ(x,θ) Rγ(x,θ) ∞Xk=3 ∞Xk=3 f (y)fk(y)dy Hk(x, θ) f 2 k (y)dy f 2 k (y) dy (6.10) f (y) dy ψk−2(r(x, θ)) = 0 for (x, θ) ∈ Ω0(1/2), where γ(x, θ) is defined in (1.3), Ω0(·) is defined in (2.3), d = 2. Formula (3.29) is proved. 7 Proof of Lemma 2 By u ∈ R we denote the coordinates on a fixed ray γ(x, θ), (x, θ) ∈ Ω, d = 2, taking into account the orientation, where u = 0 at the point x−(xθ)θ ∈ γ(x, θ); see notation (2.13). Using Lemma 1, one can see that (7.1) Using (7.1) and the assumption that fγ(x0,θ0)(u) changes the sign, one can see that 0 (R), fγ(x,θ)(u) = fγ(x,θ)(u), u ∈ R. fγ(x,θ) ∈ C ∞ there exists ψ(x0,θ0) such that ψ(x0,θ0) ∈ C ∞ 0 (R), ψ(x0,θ0) ≥ 0, ψ(x0,θ0)(u) = ψ(x0,θ0)(u), u ∈ R, Zγ(x0,θ0) f ψ(x0,θ0) dσ 6= 0, and if then also f dσ 6= 0 Zγ(x0,θ0) f dσ) sgn(Zγ(x0,θ0) sgn(Zγ(x0,θ0) f ψ(x0,θ0) dσ) = −1, where dσ = du (i.e., σ is the standard Euclidean measure on γ(x, θ)). (7.2) (7.3) (7.4) (7.5) Let W(x0,θ0)(x, θ) = 1 − ψ(x0,θ0)(xθ) f dσ Rγ(x,θ) Rγ(x,θ)f ψ(x0,θ0) dσ where dσ = du, where u is the coordinate on γ(x, θ). Lemma 1 and property (7.2) imply that , x ∈ R2, θ ∈ S1, (7.6) Zγ(x,θ) f dσ and Zγ(x,θ) f ψ(x0,θ0) dσ depend only on r(x, θ), where (x, θ) ∈ Ω, (7.7) 15 where r(x, θ) is defined in (2.2), Ω is defined in (2.1), d = 2. From (7.2), (7.6), (7.7) it follows that W(x0,θ0) is rotation-invariant in the sense (2.18). Formulas (7.3), (7.6), (7.7), properties of f of Lemma 1 and properties of ψ(x0,θ0) of (7.2) imply that ∃ε1 > 0 : Zγ(x,θ) f ψ(x0,θ0) dσ 6= 0 for (x, θ) ∈ Ω(Jr(x0,θ0),ε1), (7.8) where sets Ω(Js,ε), Js,ε are defined in (2.5), (2.9), respectively. (7.6), (7.8), one can see that In addition, using properties of f of Lemma 1 and also using (3.13), (3.19), (7.2), In addition, from (7.1)-(7.7) it follows that W(x0,θ0) ∈ C ∞(Ω(Jr(x0,θ0),ε1)). (7.9) if r(x, θ) = r(x0, θ0) then W(x0,θ0)(x, θ) = 1 − ψ(x0,θ0)(xθ) = 1 − ψ(x0,θ0)(xθ) f ψ(x0,θ0) dσ f dσ f dσ Rγ(x0,θ0) Rγ(x0,θ0) Rγ(x0,θ0) Rγ(x0,θ0) f ψ(x0,θ0) dσ ≥ 1, (7.10) where r(x, θ) is defined in (2.2), d = 2. From properties of f of Lemma 1, properties of ψ(x0,θ0) of (7.2) and from formulas (7.6), (7.8), (7.9), (7.10) it follows that ∃ε0 > 0 (ε0 < ε1) : W(x0,θ0)(x, θ) ≥ 1/2 for (x, θ) ∈ Ω(Jr(x0,θ0),ε0). Let where W(x0,θ0) is defined in (7.6). W(x0,θ0),ε0 := W(x0,θ0) for (x, θ) ∈ Ω(Jr(x0,θ0),ε0), Properties (7.7), (7.9), (7.11) imply (3.49) for W(x0,θ0),ε0 of (7.12). Using (1.1), (7.6), (7.8), (7.12) one can see that (7.11) (7.12) PW(x0,θ0),ε0 W(x0,θ0)(·, θ)f dσ f (x, θ) = Zγ(x,θ) = Zγ(x,θ) f dσ Rγ(x,θ) f ψ(x0,θ0) dσ Zγ(x,θ) Rγ(x,θ) f dσ − f ψ(x0,θ0) dσ = 0 for (x, θ) ∈ Ω(Jr(x0,θ0),ε0), (7.13) where Ω(·) is defined in (2.5), d = 2, Jr,ε is defined in (2.9). Formula (3.48) follows from (7.13). Lemma 2 is proved. 16 8 Proof of Lemma 3 Proof of (6.1)-(6.3). Using (2.2), (3.13), (3.14), (3.45), (6.3) we obtain f (x) dx, G(x, θ) = eG(r(x, θ)) =Zγ(x,θ) Hk(x, θ) = eHk(r(x, θ)) =Zγ(x,θ) efk(x) = efk((xθ2 + x − (xθ)θ2)1/2), (x, θ) ∈ Ω0(1/2), f 2 k (x) dx, where Ω0(·) is defined in (2.3), d = 2, γ(x, θ) is defined as in (2.13). Formulas (3.44), (3.45), (8.1)-(8.3) imply (6.1)-(6.3). Proof of (6.4). Let From (3.40) it follows that, for k ≥ 4: Λk = (1 − 2−k+3, 1 − 2−k+1), k ∈ N, k ≥ 4. supp ψk−1 ⊂ (1 − 2−k+2, 1 − 2−k), supp ψk−2 ⊂ (1 − 2−k+3, 1 − 2−k+1) = Λk, supp ψk−3 ⊂ (1 − 2−k+4, 1 − 2−k+2). Due to (6.2), (6.3), (8.5)-(8.7), we have the following formula for U0: U0(s, r) = 1 − eG(r) (k − 1)!efk−1((s2 + r2)1/2) +k!efk((s2 + r2)1/2) eHk(r) +(k + 1)!efk+1((s2 + r2)1/2) From (6.3), (8.8) it follows that ∂nU0 ∂sn (s, r) = − eG(r) (k − 1)! ∂nefk((s2 + r2)1/2) +k! ∂sn +(k + 1)! ψk−3(r) ψk−3(r) ψk−2(r) ψk−1(r) eHk−1(r) eHk+1(r)! for r ∈ Λk, s ∈ R, k ≥ 4. ∂nefk−1((s2 + r2)1/2) eHk−1(r) eHk+1(r)! , +∞Z−∞ ∂rnef 2 m((s2 + r2)1/2) ds, ∂rn (r) = ψk−2(r) ψk−1(r) ∂sn ∂n ∂sn eHk(r) ∂nefk+1((s2 + r2)1/2) ∂neHm (8.10) (8.9) ∂n ∂rn (r) = +∞Z−∞ ∂neG ∂rnef ((s2 + r2)1/2) ds, where eG, eHm are defined in (6.3). r ∈ Λk, s ∈ R, m ≥ 1, n ≥ 0, k ≥ 4, Using Lemma 1 and formulas (3.13), (3.14), (3.40)-(3.47), (6.3) one can see that: (8.1) (8.2) (8.3) (8.4) (8.5) (8.6) (8.7) (8.8) 0 (R), ψm−2 belongs to C ∞ ef , efm−2, eG, eHm belong to C ∞ eHm 17 0 ((1/2, 1)) for any m ≥ 3. (8.11) From (8.9)-(8.11) it follows that U0(s, r) has continuous partial derivatives of all orders with respect to r ∈ Λk, s ∈ R. It implies that U0 ∈ C ∞(R × Λk). From the fact that Λk, k ≥ 4, is an open cover of (1/2, 1) and from definition (6.2) of U0, it follows that U0 ∈ C ∞(R × {(1/2, 1) ∪ (1, +∞)}). This completes the proof of (6.4). Proof of (6.6). From (3.14)-(3.17) it follows that (8.12) Formula x2 = xθ2 + x − (xθ)θ2, x ∈ R2, θ ∈ S1, and formulas (6.2), (8.12) imply (6.6). efk(x) = 0 if x ≥ 1 for k ∈ N. Proofs of (6.7)-(6.8). Lemma 4. There are positive constants c, k1 depending on Φ of (3.15)-(3.17), such that (i) for all k ∈ N the following estimates hold: Lemma 5. Let U0 be defined by (6.2)-(6.3). Then the following estimates are valid: where C is a constant depending only on Φ of (3.15)-(3.17). Lemmas 4, 5 are proved in Subsections 9.1, 9.2, respectively. C (1 − r)5 for s ∈ R, r ∈ (1/2, 1), (ii) for k ≥ k1 and 1/2 < r ≤ 1 the following estimates hold: ψk−2(r) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) d where ef ′ efk ≤ 1, ef ′ k ≤ c8k, k denotes the derivative of efk defined in (6.3). eHk(r) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c2k, eHk(r) !(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c25k, dr ψk−2(r) where ψk are defined in (3.40), eHk is defined in (6.3). eG(r) ≤ c (r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) deG (s, r)(cid:12)(cid:12)(cid:12)(cid:12) ≤ where eG is defined in (6.3). (1 − r)3 ,(cid:12)(cid:12)(cid:12)(cid:12) (s, r)(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:12)(cid:12)(cid:12)(cid:12) (2√2)−k ∂U0 ∂r k! 8k k! , ∂U0 ∂s C dr (iii) for k ≥ 3 and r ≥ 1 − 2−k the following estimates hold: , Proof of (6.7). From (8.15), (8.17) it follows that ψk−2(r) eG(r) ≤ c(2√2)−k+3/(k − 3)!, eHk(r) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c2k, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) for r ∈ Λk, k ≥ max(4, k1), 18 (8.13) (8.14) (8.15) (8.16) (8.17) (8.18) (8.19) (8.20) (8.21) where Λk is defined in (8.4). Properties (8.5)-(8.7) and estimate (8.15) imply that r ∈ (1 − 2−k+3, 1 − 2−k+2), (8.22) ψk−1(r) = 0, ψk−3(r)  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) eHk−1(r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c2k−1  eHk+1(r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c2k+1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (ψk−1(r) = 0, ψk−2(r) = 0, ψk−1(r) ψk−3(r) = 0 if if r ∈ (1 − 2−k+2, 1 − 2−k+1), if r = 1 − 2−k+2, (8.23) (8.24) ψk−2(r) eHk(r) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) eHk+1(r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ψk−1(r) (8.25) (8.26) (8.27) (8.28) for k ≥ max(4, k1). (8.22), (8.23), (8.24). 1 − U0(s, r) = eG(r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 − U0(s, r) = eG(r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 − U0(s, r) = eG(r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Note that the assumption that r ∈ Λk is splitted into the assumptions on r of Using formulas (8.8), (8.20)-(8.24), we obtain the following estimates: (k − 1)!efk−1((s2 + r2)1/2) ≤ c(2√2)−k+3(c(k − 2)(k − 1)2k−1 + c(k − 2)(k − 1)k2k) ≤ 25√2c22−k/2k3 r ∈ (1 − 2−k+3, 1 − 2−k+2), if + k!efk((s2 + r2)1/2) ψk−3(r) eHk−1(r) ψk−2(r) k!efk((s2 + r2)1/2) + (k + 1)!efk+1((p2 + r2)1/2) ≤ c(2√2)−k+3(c2k(k − 2)(k − 1)k + c2k+1(k − 2)(k − 1)k(k + 1)) ≤ 210√2c22−k/2k4 if eHk(r) r ∈ (1 − 2−k+2, 1 − 2−k+1), eHk(r) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) r = 1 − 2−k+2, ψk−2(r) k!efk((s2 + r2)1/2) if ≤ 24√2c22−k/2k3 for s ∈ R, k ≥ max(4, k1). Estimates (8.25)-(8.27) imply that 1 − U0(s, r) ≤ C 2−k/2k4, r ∈ Λk, s ∈ R, k ≥ max(4, k1), where C is a positive constant depending on c of Lemma 4. (8.28) imply (6.7). In addition, for r ∈ Λk we have that 2−k+1 < (1 − r) < 2−k+3, which together with This completes the proof of (6.7). Proof of (6.8). We consider the following cases of s, s′, r, r′ in (6.8): 1. Let Due to (6.2) we have that s, s′ ∈ R and r, r′ ≥ 1. U0(s, r) = 1, U0(s′, r′) = 1. Identities in (8.30) and assumption (8.29) imply (6.8) for this case. (8.29) (8.30) 19 2. Let Then, due to (6.2), (6.7) we have that s, s′ ∈ R, 1/2 < r < 1 and r′ ≥ 1. 1 − U0(s, r) ≤ C(1 − r)1/3, U0(s′, r′) = 1, (8.31) (8.32) (8.33) where s, s′, r, r′ satisfy assumption (8.31), C is a constant depending only on Φ. In particular, inequality (8.32) follows from (6.7) due to the following simple property of the logarithm: loga 2(cid:18) 1 1 − r(cid:19) ≤ C(a, ε)(1 − r)−ε for any ε > 0, r ∈ [0, 1), a > 0, where C(a, ε) is some positive constant depending only on a and ε. Due to (8.31), (8.32), (8.33) we have that U0(s′, r′) − U0(s, r) = 1 − U0(s, r) ≤ C(1 − r)1/3 ≤ Cr − r′1/3 ≤ C(r − r′1/3 + s − s′1/3), where C is a constant depending only on Φ. Estimate (8.35) and assumptions (8.31) imply (6.8) for this case. 3. Let In addition, without loss of generality we assume that r > r′. Next, using (6.4) one can see that s, s′ ∈ R and r, r′ ∈ (1/2, 1). (8.34) (8.35) (8.36) (8.37) U0(s, r) − U0(s′, r′) = U0(s, r) − U0(s′, r) + U0(s′, r) − U0(s′, r′) ≤ U0(s, r) − U0(s′, r) + U0(s′, r) − U0(s′, r′) ∂U0 ∂r for s, s′ ∈ R, r, r′ >1/2, and for appropriate s, r. (s, r)(cid:12)(cid:12)(cid:12)(cid:12)s − s′ +(cid:12)(cid:12)(cid:12)(cid:12) (s′, r)(cid:12)(cid:12)(cid:12)(cid:12)r − r′, ∂U0 ∂s ≤(cid:12)(cid:12)(cid:12)(cid:12) Note that s, r belong to open intervals (s, s′), (r′, r), respectively. Using (6.7), (8.19), (8.32), (8.37) and the property that 1/2 < r′ < r < r < 1 we obtain U0(s, r) − U0(s′, r′) ≤ C((1 − r)1/3 + (1 − r′)1/3), U0(s, r) − U0(s′, r′) ≤ (1 − r)5 (s − s′ + r − r′), C (8.38) (8.39) where C is a constant depending only on Φ. We have that (1 − r)1/3 + (1 − r′)1/3 = (1 − r)1/3 + ((1 − r) + (r − r′))1/3 ≤ 2(1 − r)1/3 + r − r′1/3 ≤( 3r − r′1/3 if 1 − r ≤ r − r′, 3(1 − r)1/3 if 1 − r > r − r′, (8.40) where r, r′ satisfy (8.36). Note that in (8.40) we used the following inequality: (a + b)1/m ≤ a1/m + b1/m for a ≥ 0, b ≥ 0, m ∈ N. (8.41) 20 In particular, using (8.38), (8.40) we have that U0(s, r) − U0(s′, r′)15 ≤ 315C 15(1 − r)5 if 1 − r > r − r′, (8.42) where s, s′, r, r′ satisfy assumption (8.36), C is a constant of (8.38), (8.39). Multiplying the left and the right hand-sides of (8.39), (8.42) we obtain U0(s, r) − U0(s′, r′)16 ≤ 315C 16(s − s′ + r − r′), if 1 − r > r − r′. (8.43) Using (8.38), (8.40) we obtain U0(s, r) − U0(s′, r′) ≤ 3Cr − r′1/3, if 1 − r ≤ r − r′, (8.44) where C is a constant of (8.38), (8.39) depending only on Φ. Using (8.43) and (8.41) for m = 16, a = s − s′, b = r − r′, we have that U0(s, r) − U0(s′, r′) ≤ 3C(s − s′1/16 + r − r′1/16), if 1 − r > r − r′, (8.45) where s, s′, r, r′ satisfy assumption (8.36), C is a constant of (8.38), (8.39) which depends only on Φ. Formulas (8.44), (8.45) imply (6.8) for this case. Note that assumptions (8.29), (8.31), (8.36) for cases 1, 2, 3, respectively, cover all possible choices of s, s′, r, r′ in (6.8). This completes the proof of (6.8). This completes the proof of Lemma 3. 9 Proofs of Lemmas 4, 5 9.1 Proof of Lemma 4 Proof of (8.13), (8.14). Estimates (8.13), (8.14) follow directly from (3.14)-(3.17). Proof of (8.17). We will use the following parametrization of the points y on γ(x, θ) ∈ T S1, (x, θ) ∈ Ω, r(x, θ) 6= 0 (see notations (2.1), (2.2), (2.13) for d = 2): y(β) = x − (xθ)θ + tan(β)r(x, θ) θ, β ∈ (−π/2, π/2), (9.1) where β is the parameter. We have that: dσ(β) = r d(tan(β)) = r dβ cos2 β , r = r(x, θ), (9.2) where σ is the standard Lebesgue measure on γ(x, θ). From definitions (3.13), (6.3) it follows that ∞Xk=1 eGk(r) eG(r) = eGk(r) =Zγr efk(y) dy, γr ∈ T (r), r > 1/2, (9.3) (9.4) where T (r) is defined by (2.8). , k! 21 r r2 cos2 β = {u = tan(β)} = 2 r Φ(cid:18)2k(cid:18)1 − cos2 β(cid:19) dβ Using (3.14), (9.1), (9.2), (9.4) we obtain the following formula for eGk: π/2Z−π/2 eGk(r) = r cos β(cid:19)(cid:19) cos(cid:18)8k +∞Z0 Φ(cid:16)2k(cid:16)1 − r√u2 + 1(cid:17)(cid:17) cos(cid:0)8kr2(u2 + 1)(cid:1) du +∞Z0 Φ(cid:16)2k(cid:16)1 − r√t + 1(cid:17)(cid:17) cos(cid:0)8kr2(t + 1)(cid:1) dt = {t = u2} = r +∞Z0 Φ(2k(1 − r√t + 1)) +∞Z0 Φ(2k(1 − r√t + 1)) sin(8kr2t) √t dt − r sin(8kr2) cos(8kr2t) dt √t = r cos(8kr2) √t = 8−k/2r−1 cos(8kr2) − 8−k/2r−1 sin(8kr2) +∞Z0 +∞Z0 Φk(t, r) cos(t) √t dt Φk(t, r) sin(t) √t dt, r > 1/2, where Φk(t, r) = Φ(2k(1 − rp8−kr−2t + 1)), t ≥ 0, r > 1/2, k ∈ N. For integrals arising in (9.5) the following estimates hold: (9.5) (9.6) (9.7) (9.8) Φk(t, r) Φk(t, r) +∞Z0 +∞Z0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) sin(t) √t cos(t) √t dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) for 1/2 < r < 1, k ≥ 1. ≤ C1 < +∞, ≤ C2 < +∞, where Φk is defined in (9.6), C1, C2 are some positive constants depending only on Φ and not depending on k and r. Estimates (9.7), (9.8) are proved in Subsection 9.3. From (9.5)-(9.8) it follows that Note that for y ∈ γr, the following inequality holds: eGk(r) ≤ 2 · 8−k/2(C1 + C2) for r > 1/2, k ∈ N. 2k(1 − y) ≤ 2k(1 − r) ≤ 2k−m ≤ 1/2 < 4/5 for 1 − 2−m ≤ r < 1, k < m, m ≥ 3, where γr is a ray in T (r) (see notations of (2.8), d = 2). Formulas (3.14), (3.15), (6.3), (9.10) imply that γr ∩ supp fk = ∅ if r ≥ 1 − 2−m, k < m, 22 (9.9) (9.10) (9.11) In turn, (9.4), (9.11) imply that Due to (9.3), (9.4), (9.9), (9.12) we have that: eGk(r) = 0 for r ≥ 1 − 2−m, k < m, m ≥ 3. ∞Xk=m (2√2)−m eG(r) ≤ ≤ 2(C1 + C2) eGk(r)/k! 4√2 2√2 − 1 , c1 =(C1 + C2) for r ≥ 1 − 2−m, m ≥ 3. m! (2√2)−k = c1 (2√2)−m m! ∞Xk=0 (9.12) , (9.13) . (9.14) Formulas (3.14), (8.10) for n = 1, (8.14), (9.4) imply that deGk(r) dr 1 k!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) This completes the proof of estimate (8.17). Proof of (8.18). Using (9.3), (9.4) we have that: dr (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) deG (r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ ∞Xk=1 =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ds(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +∞Z−∞ ref ′ +∞Z−∞ k((s2 + r2)1/2) ds =Zγr ef ′ k((s2 + r2)1/2) √r2 + s2 ≤ deGk dr (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (r)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) where B(0, 1) is defined in (2.16), d = 2. At the same time, formula (9.12) implies that k(y) dy ≤ c8k Zγr∩B(0,1) ef ′ dy ≤ 2c8k, Formulas (9.14), (9.15), (9.16) imply the following sequence of inequalities: = 0 for r ≥ 1 − 2−m, k < m, m ≥ 3. 1 dr deGk(r) k!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) deGk(r) ∞Xk=0 8k k! dr (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c ∞Xk=m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ m!8k (k + m)! ≤ dr (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) deG(r) ∞Xk=0 8m m! ∞Xk=0 m!8k (k + m)! , r ≥ 1 − 2−m, m ≥ 3. (9.17) The series in the right hand-side in (9.17) admits the following estimate: = e8 and the estimate does not depend on m. Formulas (9.17), (9.18) imply (8.18). Proof of (8.15). For each ψk from (3.40) we have that Therefore, it is sufficient to show that ψk ≤ 1. eHk ≥ C2−k for k ≥ k1, C = c−1. 23 (9.15) (9.16) (9.18) (9.19) (9.20) Proceeding from (6.3) and in a similar way with (9.5) we obtain the formulas Φ2(2k(1 − r√t + 1)) √t cos2(8kr2(t + 1)) dt = eHk,1(r) + eHk,2(r), r > 1/2, (9.21) +∞Z0 +∞Z0 +∞Z0 r 2 r 2 eHk(r) = r eHk,1(r) = eHk,2(r) = Φ2(2k(1 − r√t + 1)) √t dt, Φ2(2k(1 − r√t + 1)) √t cos(2 · 8kr2(t + 1)) dt. (9.22) (9.23) (9.24) (9.25) (9.26) (9.27) (9.28) (9.29) In addition, we have that: supptΦ2(2k(1 − r√t + 1) ⊂ [0, 3] for 1/2 < r ≤ 1 − 2−k+1, k ≥ 3, where suppt denotes the support of the function in variable t. Property (9.24) is proved below in this paragraph (see formulas (9.26)-(9.29)). Note that From (3.15), (3.16) and from (9.25) we have that: 2k(1 − r) ≥ 2k · 2−k+1 ≥ 2 > 6/5 for 1/2 < r ≤ 1 − 2−k+1, k ≥ 3. supptΦ2(2k(1 − r√t + 1) ⊂ [0, +∞) for 1/2 < r ≤ 1 − 2−k+1, k ≥ 3. We have that 2 = t(k)  1 + 1) = 11/10, 1 (r) ≥ 0, t(k) ∃t(k) 1 = t(k) 2k(1 − rqt(k) 2k(1 − rqt(k) 2 + 1 −qt(k) t(k) 2 − t(k) for 1/2 < r ≤ 1 − 2−k+1, k ≥ 3. 1 ≥(cid:18)qt(k) 2 + 1) = 9/10, 2 (r) ≥ 0, t(k) 2 > t(k) 1 , such that 1 + 1(cid:19) = 2−k 5 r−1 ≥ 2−k 5 , In addition, from (9.27) it follows that t(k) 1 = (1 − 2−k 11 10 )2 (1 − 2−k 9 10 )2 r2 − 1 ≤ 4(1 − 2−k 11 − 1 ≤ 4(1 − 2−k 11 10 10 )2 − 1 ≤ 3, )2 − 1 ≤ 3, t(k) 2 = for 1/2 < r ≤ 1 − 2−k+1, k ≥ 3. r2 Using (3.15)-(3.17), (9.22), (9.24), (9.27)-(9.29) we have that eHk,1(r) ≥ dt √t ≥ 3+t(k) r 2 2 −t(k) 1 dt √t Z3 r 2 t(k) 1 2Zt(k) Z3 2 −t(k) 1 3+t(k) r ≥ 6 dt = r 6t(k) 2 − t(k) 1 ≥ for 1/2 < r ≤ 1 − 2−k+1, k ≥ 3. (9.30) 2−k 30 24 On the other hand, proceeding from using (9.23) and, in a similar way with (9.5)-(9.9), we have eHk,2(r) = ≤ + cos(2 · 8kr2(t + 1)) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) r r √t Φ2(2k(1 − r√t + 1)) 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +∞Z0 2 cos(2 · 8kr2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +∞Z0 Φ2(2k(1 − r√t + 1)) 2 sin(2 · 8kr2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +∞Z0 Φ2(2k(1 − r√t + 1)) 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +∞Z0 ≤ 8−k/2 r−1 ≤ 8−k/2C, for 1/2 < r < 1 − 2−k+1, k ≥ 3, dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) cos(2t) √t k(t, r) Φ2 r √t +∞Z0 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) cos(2 · 8kr2t) √t sin(2 · 8kr2t) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (9.31) dt(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) + 8−k/2 r−1 Φ2 k(t, r) sin(2t) √t where Φk(t, r) is defined in (9.6), C is some constant depending only on Φ and not depending on k, r. In (9.31) we have also used that Φ2(t) satisfies assumptions (3.15)- (3.17). Note also that Φ2(t) satisfies assumptions (3.15)-(3.17) for Φ(t). Using (9.21)-(9.23), (9.30), (9.31) we obtain ≥ eHk(r) ≥ eHk,1(r) − eHk,2(r) 2−k 30 − C ′ · 8−k/2 ≥ 2−k(cid:18) 1 (√2)k(cid:19) C ′ ≥ C2−k for 1/2 < r < 1 − 2−k+1, k ≥ k1 ≥ 3, 30 − C ′(√2)−k1, 30 − C = 1 (9.32) where C ′ depends only on Φ, k1 is arbitrary constant such that k1 ≥ 3 and C is positive. Formulas (8.15) follows from (3.40), (9.32). This completes the proof (8.15). Proof of (8.16). The following formula holds: d dr ψk−2(r) eHk(r) ! = −eH ′ k−2 denote the derivatives of eHk, ψk, defined in (6.3), (3.40), respectively. k(r)ψk−2(r) − eHk(r)ψ′ , 1/2 < r < 1, eH 2 k−2(r) k, ψ′ (9.33) k (r) where eH ′ 25 Using (3.14), (6.3), (8.10), n = 1, (8.13), (8.14) we have that r k(r) = 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +∞Z−∞ √r2 + s2efk(√r2 + s2)ef ′ eH ′ +∞Z−∞ (cid:12)(cid:12)(cid:12)efk(√r2 + s2)ef ′ ≤ 2c8k Zγr∩B(0,1) k(√r2 + s2) ds(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k(√r2 + s2)(cid:12)(cid:12)(cid:12) ds = 2Zγr efk(y)ef ′ dy ≤ 4c8k, γr ∈ T (r), k ≥ 3, r > 1/2, ≤ 2 k(y) dy where we use notations (2.8), (2.16), d = 2. Using (3.40), (3.41), (8.15), (9.32)-(9.34) we have that d eHk(r) !(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ C22k(eH ′ dr ψk−2(r) for 1/2 < r < 1 − 2−k+1, k ≥ k1 ≥ 3, k(r) + eHk(r) · ψ′ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k(r)) ≤ C ′25k, where C ′ is a constant not depending on k and r and depending only on Φ. This completes the proof of Lemma 4. 9.2 Proof of Lemma 5 It is sufficient to show that (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ∂U0(s, r) ∂s ∂U0(s, r) ∂r C (1 − r)3 , (1 − r)5 , C (cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:12)(cid:12)(cid:12)(cid:12) ≤ for s ∈ R, r ∈ Λk, k ≥ max(4, k1), ∂U0(s, r) ∂s ∂U0(s, r) ∂r (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ≤ C · 8k, (cid:12)(cid:12)(cid:12)(cid:12) ≤ C · (32)k, for s ∈ R, r ∈ Λk, (9.34) (9.35) (9.36) (9.37) (9.38) (9.39) where C is a positive constant depending only on Φ of (3.14), Λk is defined in (8.4), k1 is a constant arising in Lemma 4 and depending only on Φ. is an open cover of (1/2, 1). Indeed, estimates (8.19) follow from (6.4), (9.36), (9.37) and the fact that Λk, k ≥ 4, In turn, estimates (9.36), (9.37) follow from the estimates and from the fact that 2−k+1 < 1 − r < 2−k+3, k ≥ max(4, k1), for r ∈ Λk, where C is a positive constant depending only on Φ. Estimate (9.38) follows from formula (8.9) for n = 1 and estimates (8.14), (8.15), (8.20)-(8.24). Estimate (9.39) follows from (8.8), (8.13)-(8.16), (8.20)-(8.24) and from the esti- 26 where c is a constant arising in Lemma 4. Estimate (9.40) follows from (8.16) (used with k−1, k, k+1 in place of k). Estimate This completes the proof of Lemma 5. (9.41) follows from (8.18) (used with k − 3 in place of k). 9.3 Proof of estimates (9.7), (9.8) We use the following Bonnet's integration formulas (see, e.g., [F59], Chapter 2): bZa bZa f1(t)h(t) dt = f1(a) f2(t)h(t) dt = f2(b) ξ1Za bZξ2 h(t) dt, h(t) dt, for some appropriate ξ1, ξ2 ∈ [a, b], where f1 is monotonously decreasing on [a, b], f1 ≥ 0, f2 is monotonously increasing on [a, b], f2 ≥ 0, h(t) is integrable on [a, b]. Let g1(t) = sin(t) √t , g2(t) = cos(t) √t , t > 0, G1(s) = sZ0 sin(t) √t dt, G2(s) = dt, s ≥ 0. We recall that From (9.45), (9.46), (9.47) it follows that lim s→+∞ G1(s) = lim s→+∞ cos(t) √t sZ0 G2(s) =rπ 2 . (9.40) (9.41) (9.42) (9.43) (9.44) (9.45) (9.46) (9.47) (9.48) (9.49) (9.50) mates: d eHk−i+2(r)!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c25(k+1), dr ψk−i(r) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c deG(r) for r ∈ Λk, i ∈ {1, 2, 3}, 8−k+3 (k − 3)! (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dr , G1, G2 are continuous and bounded on [0, +∞). Due to (3.15)-(3.18), (9.6) and monotonicity of the function 2k(1 − r√8−kr−2t + 1) in t on [0, +∞) it follows that Φk(t, r) is monotonously decreasing on [0, +∞), if 2k(1 − r) ≤ 11/10, Φk(t, r) is monotonously increasing on [0, t0] for some t0 > 0 and is monotonously decreasing on [t0, +∞), if 2k(1 − r) > 11/10. for r > 1/2, k ∈ N, 27 Moreover, due to (3.15)-(3.17), (9.6), for Tk = 8k, k ∈ N, we have that √1 + r−2)) = 0, Φk(Tk, r) = Φ(2k(1 − r√r−2 + 1)) = Φ(2k(1 − Φk(t, r) = 0 for t ≥ Tk, Φk(t, r) ≤ 1 for t ≥ 0, r > 1/2, k ∈ N. Using (9.6), (9.45)-(9.50), (9.52) and (9.42)-(9.44) we obtain (9.51) (9.52) (9.53) +∞Z0 ∞Z0 Φk(t, r)gi(t) dt = (9.54) ξZ0 gi(t) dt Φk(t, r)gi(t) dt = Φk(0, r) TkZ0 = Φk(0, r)Gi(ξ) for appropriate ξ ∈ [0, Tk], if 2k(1 − r) ≤ 11/10, TkZ0 Φk(t, r)gi(t) dt = Φk(t, r)gi(t) dt + Φk(t, r)gi(t) dt Φk(t, r)gi(t) dt = t0Z0 TkZt0 = Φk(t0, r) t0Zξ′ gi(t) dt + Φk(t0, r) ξ′′Zt0 gi(t) dt = Φk(t0, r)(Gi(ξ′′) − Gi(ξ′)) for appropriate ξ′ ∈ [0, t0], ξ′′ ∈ [t0, Tk], if 2k(1 − r) > 11/10, (9.55) where i = 1, 2. Estimates (9.7), (9.8) follow from (9.45), (9.46), (9.48), (9.53)-(9.55). References [ABK98] Arbuzov, E.V., Bukhgeim, A.L., Kazantsev, S. G., Two-dimensional tomography problems and the theory of A-analytic functions. Siberian Adv. Math., 8:1-20, 1998. [Bal09] Bal, G., Inverse transport theory and applications. Inverse Problems, 25(5), 053001 (48pp), 2009. [BJ11] Bal, G., Jollivet, A., Combined source and attenuation reconstructions in SPECT. To- mography and Inverse Transport Theory, Contemp. Math., 559, 13-27, 2011. [Be84] Beylkin, G., The inversion problem and applications of the generalized Radon transform. Communications on pure and applied mathematics, 37(5):579-599, 1984. [BQ87] Boman, J., Quinto, E.T., Support theorems for real-analytic Radon transforms. Duke Mathematical J., 55(4):943-948, 1987. [BQ93] Boman, J., Quinto, E. T., Support theorems for Radon transforms on real analytic line complexes in three-space. Trans. Amer. Math. Soc., 335(2):877-890, 1993. [B93] Boman, J., An example of non-uniqueness for a generalized Radon transform. Journal d'Analyse Mathematique, 61(1):395 -- 401, 1993. [BS04] Boman, J., Stromberg, J.-O., Novikov's inversion formula for the attenuated Radon trans- form -- a new approach, The Journal of Geometric Analysis, 14(2): 185-198, 2004. [C64] Cormack, A.M., Representation of a function by its line integrals, with some radiological applications I, II. J. Appl. Phys. 34:2722 -- 2727; 35:2908 -- 2912, 1963-1964. [M92] Do Carmo, M. P., Riemannian Geometry. Birkhauser Basel, 1992. [F59] [Fi86] Fichtenholz, G.M., A course of differential and integral calculus, volume II, Moscow, 1959. Finch, D., Uniqueness for the attenuated X-ray transform in the physical range. Inverse Problems 2(2), 1986. [GGG82] Gel'fand, I. M., Gindikin, S. G., Graev, M. I., Integral geometry in affine and projective spaces. Journal of Soviet Mathematics, 18(2):39 -- 167, 1982. 28 [Gi10] Gindikin, S., A remark on the weighted Radon transform on the plane. Inverse Problems and Imaging, 4(4):649-653, 2010. [GN17] Goncharov F.O., Novikov, R.G. An example of non-uniqueness for Radon transforms with continuous positive rotation invariant weights. The Journal of Geometric Analysis (doi:10.1007/s12220-018-0001-y), 2017. [H01] [I16] [J38] Helgason, S., Differential geometry and symmetric spaces. American Mathematical Soc., volume 341, 2001. Ilmavirta, J., Coherent quantum tomography. SIAM Journal on Mathematical Analysis, 48(5):3039 -- 3064, 2016. John, F., The ultrahyperbolic differential equation with 4 independent variables. Duke Math. J., 4:300 -- 322, 1938. [KLM95] Kuchment, P., Lancaster, K., Mogilevskaya L., On local tomography. Inverse Problems, 11(3):571-589, 1995. [Kun92] Kunyansky, L., Generalized and attenuated Radon transforms: restorative approach to the numerical inversion. Inverse Problems, 8(5):809-819, 1992. [LB73] Lavrent'ev, M. M., Bukhgeim, A. L., A class of operator equations of the first kind. Functional Analysis and Its Applications, 7(4):290-298, 1973. [MQ85] Markoe, A., Quinto, E.T., An elementary proof of local invertibility for generalized and attenuated Radon transforms. SIAM Journal on Mathematical Analysis, 16(5):1114 -- 1119, 1985. [Na01] Natterer, F., The Mathematics of Computerized Tomography. SIAM, 2001. [Ng17] Nguyen, L. V., On the strength of streak artifacts in filtered back-projection reconstructions for limited angle weighted X-ray transform. J. Fourier Anal. Appl., 23(3):712 -- 728, 2017. [N02a] Novikov, R.G., On determination of a gauge field on Rd from its non-abelian Radon transform along oriented straight lines. Journal of the Institute of Mathematics of Jussieu, 1(4), 559-629., 2002. [N02b] Novikov, R.G., An inversion formula for the attenuated X-ray transformation. Arkiv for matematik, 40(1):145-167, 2002. [N14] [Pa96] Novikov, R.G., Weighted Radon transforms and first order differential systems on the plane. Moscow Mathematical Journal, 14(4):807 -- 823, 2014. Palamodov, V.P., An inversion method for an attenuated x-ray transform. Inverse Prob- lems, 12(5):717-729, 1996. [PG13] Puro, A., Garin, A., Cormack-type inversion of attenuated Radon transform, Inverse Problems, 29(6), 065004 (14pp), 2013. [R17] [Sh92] Radon, J., Uber die Bestimmung von Funktionen durch ihre Integralwerte langs gewisser Mannigfaltigkeiten. Ber. Saechs Akad. Wiss. Leipzig, Math-Phys, 69:262 -- 267, 1917. Sharafutdinov, V. A., On the problem of emission tomography for non-homogeneous media. Dokl. Akad. Nauk. 326:446-448, 1992 (Russian). English transl.: Soviet Phys. Dokl. 37: 469-470, 1992. [Sh93] Sharafutdinov, V. A., Uniqueness theorems for the exponential X-ray transform. Journal of Inverse and Ill-Posed Problems, 1(4):355-372, 1993. [TM80] Tretiak O.J., Metz, C., The exponential Radon transform. SIAM J. Appl. Math., 39:341- 354, 1980. [Q83] Quinto, E.T., The invertibility of rotation invariant Radon transforms. Journal of Mathe- matical Analysis and Applications, 91(2):510 -- 522, 1983. 29
1501.07136
4
1501
2017-05-08T12:32:08
Density of bounded maps in Sobolev spaces into complete manifolds
[ "math.FA" ]
Given a complete noncompact Riemannian manifold $N^n$, we investigate whether the set of bounded Sobolev maps $(W^{1, p} \cap L^\infty) (Q^m; N^n)$ on the cube $Q^m$ is strongly dense in the Sobolev space $W^{1, p} (Q^m; N^n)$ for $1 \le p \le m$. The density always holds when $p$ is not an integer. When $p$ is an integer, the density can fail, and we prove that a quantitative trimming property is equivalent with the density. This new condition is ensured for example by a uniform Lipschitz geometry of $N^n$. As a byproduct, we give necessary and sufficient conditions for the strong density of the set of smooth maps $C^\infty (\overline{Q^m}; N^n)$ in $W^{1, p} (Q^m; N^n)$.
math.FA
math
DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES INTO COMPLETE MANIFOLDS PIERRE BOUSQUET, AUGUSTO C. PONCE, AND JEAN VAN SCHAFTINGEN ABSTRACT. Given a complete noncompact Riemannian manifold N n, we inves- tigate whether the set of bounded Sobolev maps (W 1,p ∩ L∞)(Qm; N n) on the cube Qm is strongly dense in the Sobolev space W 1,p(Qm; N n) for 1 ≤ p ≤ m. The density always holds when p is not an integer. When p is an integer, the den- sity can fail, and we prove that a quantitative trimming property is equivalent with the density. This new condition is ensured for example by a uniform Lipschitz geometry of N n. As a byproduct, we give necessary and sufficient conditions for the strong density of the set of smooth maps C∞(Qm; N n) in W 1,p(Qm; N n). 1. INTRODUCTION Bounded maps from the unit cube Qm = (−1, 1)m ⊂ Rm with m ∈ N∗ = {1, 2, . . .} are dense in the class of Sobolev maps W 1,p(Qm; Rν), and this follows from a straightforward truncation argument. In the setting of Sobolev maps with values into manifolds, this elementary approach is unable to handle additional constraints on the target. class of Sobolev maps with values into N n as More precisely, given a closed smooth submanifold N n ⊂ Rν, we define the W 1,p(Qm; N n) =nu ∈ W 1,p(Qm; Rν) : u ∈ N n almost everywhereo; the space L∞(Qm; N n) of essentially bounded maps is defined similarly. The ques- tion addressed in the present work is whether the set (W 1,p ∩ L∞)(Qm; N n) is dense in W 1,p(Qm; N n) with respect to the strong W 1,p topology. When N n is an abstract complete smooth Riemannian manifold, there exists an isometric embedding ι : N n → Rν such that ι(N n) is closed [22, 24]. This allows one to define the functional spaces W 1,p(Qm; N n) and L∞(Qm; N n), and different embeddings of N n yield homeomorphic spaces; see Section 2 below. We thus con- sider indifferently the case where N n is an embedded closed smooth submanifold of Rν or an abstract complete smooth Riemannian manifold. One of our motivations to the question above comes from the density problem of the set C∞(Qm; N n) of smooth maps in Sobolev spaces W 1,p(Qm; N n) with values into complete manifolds. Even when N n is compact, this is a delicate prob- lem that has been studied by many authors. Schoen and Uhlenbeck [26] estab- lished the density when p ≥ m, and Bethuel [2] (see also Hang and Lin [16]) proved in the case 1 ≤ p < m that density holds if and only if the homotopy Date: May 9, 2017. 2010 Mathematics Subject Classification. 58D15 (46E35, 46T20). Key words and phrases. Strong density; Sobolev maps; bounded maps; complete manifolds. 1 2 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN group π⌊p⌋(N n) is trivial. The latter condition means that every continuous map f : S⌊p⌋ → N n on the ⌊p⌋-dimensional sphere is homotopic to a constant map, where ⌊p⌋ is the largest integer less than or equal to p. For complete manifolds, the same conclusions hold provided that W 1,p(Qm; N n) is replaced by the smaller space (W 1,p ∩ L∞)(Qm; N n); see Section 3 below. The strong density of smooth maps in W 1,p(Qm; N n) is thus equivalent to the density of (W 1,p ∩ L∞)(Qm; N n) in the latter space. When p > m, Sobolev maps on the cube Qm are bounded, and even Hölder continuous, whence W 1,p ∩ L∞ = W 1,p. When p ≤ m, we establish that the density of bounded Sobolev maps depends on whether p is an integer or not. Theorem 1. For every 1 ≤ p ≤ m such that p 6∈ N, the set (W 1,p ∩ L∞)(Qm; N n) is dense in W 1,p(Qm; N n). The case where p is an integer is more subtle and the answer involves analytical properties of the manifold N n. This surprising phenomenon arises even in the case p = m. In the related problem of density of smooth maps in W 1,m(Qm; N n) when N n is a compact manifold, this critical case always has an affirmative answer, regardless of πm(N n), and is a straightforward consequence of the fact that W 1,m maps embed into the class of vanishing mean oscillation (VMO) maps [8, 26]. For noncompact manifolds, this VMO property is not sufficient to imply the density of bounded maps in W 1,m, even if N n is diffeomorphic to the Euclidean space Rn; see Section 4 below. In fact, for integer exponents p this density problem is equivalent to the following analytical assumption on the target N n: Definition 1.1. Given p ∈ N∗, the manifold N n satisfies the trimming property of dimension p whenever there exists a constant C > 0 such that each map f ∈ C∞(∂Qp; N n) with a Sobolev extension u ∈ W 1,p(Qp; N n) also has a smooth ex- tension v ∈ C∞(Qp; N n) such that kDvkLp(Qp) ≤ CkDukLp(Qp). The use of C∞ maps is not essential, and other classes like Lipschitz maps or continuous Sobolev maps (W 1,p ∩ C0) yield equivalent definitions of the trimming property; see e.g. Proposition 6.1. The trimming property is satisfied by any mani- fold N n with uniform Lipschitz geometry in the sense of Definition 6.1 below, for example when N n is the covering space of a compact manifold [3, 25] or when N n is a Lie group or a homogeneous space [10, 27]. We also observe that every com- plete manifold satisfies the trimming property of dimension 1: it suffices to take as v any shortest geodesic connecting the points f (−1) and f (1). follows: The answer to the density problem for integer exponents can now be stated as Theorem 2. For every p ∈ {1, . . . , m}, the set (W 1,p ∩ L∞)(Qm; N n) is dense in W 1,p(Qm; N n) if and only if N n satisfies the trimming property of dimension p. As a consequence of Theorems 1 and 2, we characterize the class of complete manifolds N n for which smooth maps are dense in the space W 1,p(Qm; N n). For non-integer values of the exponent p we have: DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 3 Corollary 1.1. For every 1 ≤ p ≤ m such that p 6∈ N, the set C∞(Qm; N n) is dense in W 1,p(Qm; N n) if and only if π⌊p⌋(N n) is trivial. For integer values of p, the characterization becomes: Corollary 1.2. Case p = 1: The set C∞(Qm; N n) is dense in W 1,1(Qm; N n) if and only if π1(N n) is trivial. Case p ∈ {2, . . . , m − 1}: The set C∞(Qm; N n) is dense in W 1,p(Qm; N n) if and only if πp(N n) is trivial and N n satisfies the trimming property of dimension p. Case p = m: The set C∞(Qm; N n) is dense in W 1,m(Qm; N n) if and only if N n satis- fies the trimming property of dimension m. In the more general setting of Sobolev maps into noncompact metric spaces [14, 19], Hajłasz and Schikorra [15] have recently given a necessary condition for the density of Lipschitz maps in terms of an (m− 1)-Lipschitz connectedness property. The strategy of the proofs of Theorems 1 and 2 above is based on the good and bad cube method introduced by Bethuel [2] for a compact manifold N n. More pre- cisely, we divide the domain Qm as a union of small cubes and we approximate a map u ∈ W 1,p(Qm; N n) in two different ways, depending on the properties of u on each cube. On the good cubes, we approximate u by convolution with a smooth kernel. In general, such an approximation does not take its values in N n, so that we must project it back on the target manifold using a retraction Π on N n. Such a strategy works when • the retraction Π is well-defined on a tubular neighborhood of positive uniform • the convolution of u with a smooth kernel takes its values in this tubular neigh- borhood. The first condition is automatically satisfied when N n is compact, while the second one holds true when u has small mean oscillation. In Bethuel's and Hang– Lin's works, the latter condition on u is used to define good cubes (see also [5, η of inradius η is good if the rescaled energy p. 797]) in the sense that a cube σm satisfies width around N n, 1 ηm−p σm η Dup . δ, for some small parameter δ > 0 depending on the width of the tubular neighbor- hood; the inradius is half the edge-length of the cube. The connection between such a condition and the oscillation of u on σm η can be made using the Poincaré– Wirtinger inequality: σm η σm η u(x) − u(y) dx dy ≤ C ηm−p σm η Dup. Noncompact submanifolds, however, need not have a global tubular neighbor- hood with positive uniform width. For instance, it is impossible to find such a tubular neighborhood for any isometric embedding of the hyperbolic space in a Euclidean space because the volumes of hyperbolic balls grow exponentially with respect to the radius. Since geodesic balls in N n do have a tubular neighborhood 4 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN with uniform width, we thus modify the classical definition of a good cube by fur- ther requiring that most of the points of u lie in a fixed geodesic ball. It is then possible to perform the approximation by convolution and projection. The param- eter scale of this convolution is not constant but depends on the distance to the bad cubes. Such an adaptive smoothing (Section 5.2) is used to smoothen the transition from good cubes to bad cubes. On the bad cubes, we modify u using the opening technique (Section 5.1). This op- eration was introduced by Brezis and Li [7], and was then pursued by the authors in [5] in the framework of higher order Sobolev spaces. More precisely, given an ℓ-dimensional grid in the bad cubes with ℓ ∈ {0, . . . , m − 1}, we use the opening technique to slightly modify u near the grid to obtain a new map uop that locally depends on ℓ variables and whose restriction to the grid belongs to W 1,p. Taking ℓ = ⌊p⌋ for p < m, we next perform a zero-degree homogenization (Sec- tion 5.3) which consists in propagating the values of uop on the ⌊p⌋-dimensional grid to all m-dimensional bad cubes [2, 4, 16]. When p 6∈ N∗, the Morrey–Sobolev embedding implies that uop is bounded on the ⌊p⌋-dimensional grid, and we end up with a bounded Sobolev map on the bad cubes. When p ∈ N∗, the resulting map need not be bounded, so before applying the zero-degree homogenization, we need to do some preliminary work that relies on the trimming property of dimen- sion p. Indeed, since uop is continuous on the (p − 1)-dimensional grid, the trim- ming property allows one to replace uop on the p-dimensional grid by a bounded map which coincides with uop on the (p − 1)-dimensional grid; see Section 6. The resulting map obtained by zero-degree homogenization is now bounded as in the non-integer case. Although we obtain a function which can be quite different from u on the bad cubes, we can conclude using the fact that most of the cubes are good. We now describe the plan of the paper. In Section 2, we explain why the def- inition of the Sobolev space W 1,p(Qm; N n) is independent of the specific isomet- ric embedding of N n. In Section 3, we investigate the density of smooth maps C∞(Qm; N n) in (W 1,p ∩ L∞)(Qm; N n) using results from the case of compact- In Section 4, we present a counterexample to the density of target manifolds. bounded maps in the class W 1,m(Qm; N n), where N n is a suitable embedding of Rn in Rn+1. In Section 5, we have collected the main tools that will be used in the proofs of Theorems 1 and 2: the opening technique, the adaptive smoothing and the zero-degree homogenization. In Section 6, we give equivalent formulations of the trimming property and we establish the necessity of this condition for the den- sity of bounded maps when p ∈ N∗. Finally, we present the proofs of Theorems 1 and 2 in Section 7. 2. MAPS INTO A COMPLETE MANIFOLD 2.1. Sobolev maps. In the definition of the Sobolev spaces introduced above, we have used an isometric embedding ι : N n → Rν of the target smooth Riemannian manifold N n. We claim that such a definition does not depend on the embedding, in the following sense: if ι1 and ι2 are two isometric embeddings of N n into Rν1 and Rν2 respectively, then the two resulting Sobolev spaces are homeomorphic. For the convenience of the reader, we provide below a self-contained proof of this fact. Alternatively, it is possible to define intrinsically Sobolev spaces of maps into DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 5 complete Riemannian manifolds without reference to any isometric embedding of the target manifold N n. Such an approach turns out to be equivalent to the definition that relies on an embedding of N n, see [9, Propositions 2.7 and 4.4] and also [13]. Proposition 2.1. Let ιk : N n → Rνk , with k ∈ {1, 2}, be two isometric embeddings of a complete Riemannian manifold N n and let 1 ≤ p < ∞. (i) Given a measurable map u : Qm → N n, ι1 ◦ u ∈ W 1,p(Qm; Rν1 ) if and only if ι2 ◦ u ∈ W 1,p(Qm; Rν2). (ii) Given a sequence of measurable maps uj : Qm → N n, with j ∈ N, we have that (ι1◦uj)j∈N converges to ι1◦u in W 1,p(Qm; Rν1 ) if and only if (ι2◦uj)j∈N converges to ι2 ◦ u in W 1,p(Qm; Rν2 ). In the proof of Proposition 2.1, we need a specific version of the chain rule for functions of the form Φ ◦ w, where w is a Sobolev map and Φ is a C1 map such that DΦ is bounded on the range of w, in the spirit of the composition formula in [21, Theorem 2.1]: Lemma 2.2. Let M n ⊂ Rν1 be an embedded complete Riemannian submanifold and Φ ∈ C1(M n; Rν2). Then, for every w ∈ W 1,p(Qm; M n) such that DΦ ◦ w ∈ L∞(Qm), we have Φ ◦ w ∈ W 1,p(Qm; Rν2 ) and, for almost every x ∈ Qm, D(Φ ◦ w)(x) = DΦ(w(x)) ◦ Dw(x). The proof is based on an adaptation of the argument in [21] and relies on Mor- rey's characterization of Sobolev maps. To deduce Proposition 2.1, we apply this lemma to the map Φ = ι2 ◦ ι−1 1 , which need not be globally Lipschitz-continuous as in usual versions of the composition formula. Proof of Lemma 2.2. The fact that Φ ◦ w ∈ Lp(Qm; Rν2) is a consequence of the fol- lowing Poincaré–Wirtinger inequality: (2.1) Qm QmΦ ◦ w(x) − Φ ◦ w(y)p dx dy ≤ C QmDw(x)p dx, where C > 0 depends on kDΦ ◦ wkL∞(Qm) = kDΦkL∞(F ), and F is the essential range of w. We recall that the essential range is the smallest closed subset F ⊂ M n such that w(x) ∈ F for almost every x ∈ Qm, see [8, Section I.4]. To prove (2.1), we observe that, for almost every x′ ∈ Qm−1, we have w(·, x′) ∈ W 1,p(Q1; M n). By the Morrey–Sobolev embedding, w(·, x′) can be identified to a continuous (and thus bounded) map on Q1 with values into M n. Hence, we can apply the classical chain rule to Φ ◦ w(·, x′), because the restriction of Φ to w(·, x′)(Q1) is globally Lipschitz-continuous. Hence, Φ ◦ w(·, x′) ∈ W 1,p(Q1; Rν2 ) and (2.2) d dx1 (Φ ◦ w)(x1, x′) = DΦ(w(x1, x′))[∂1w(x1, x′)]. Since DΦ is bounded on the essential range F , this implies (2.3) d dx1 (cid:12)(cid:12)(cid:12) (Φ ◦ w)(x1, x′)(cid:12)(cid:12)(cid:12) ≤ C1∂1w(x1, x′), 6 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN for some constant C1 > 0 independent of x′. The standard one-dimensional Poincaré–Wirtinger inequality applied to Φ ◦ w(·, x′) thus yields Q1 Q1Φ ◦ w(t, x′) − Φ ◦ w(s, x′)p dt ds ≤ C2Q1∂1w(t, x′)p dt. By integrating over x′ ∈ Qm−1, one gets Qm−1 Q1 Q1Φ ◦ w(t, x′) − Φ ◦ w(s, x′)p dt ds dx′ ≤ C2Qm∂1w(t, x′)p dt dx′. The same calculation can be performed for every coordinate. Using the triangle inequality, one has the estimate Φ ◦ w(x) − Φ ◦ w(y)p ≤ C3 mXi=1 Φ ◦ w(y1, . . . , yi−1, xi, xi+1, . . . , xm) − Φ ◦ w(y1, . . . , yi−1, yi, xi+1, . . . , xm)p. By integration, one obtains the Poincaré–Wirtinger inequality (2.1). We now prove that Φ ◦ w ∈ W 1,p(Qm; Rν2 ). For almost every x′ ∈ Qm−1, by estimate (2.3) we have that d dx1 Qm(cid:12)(cid:12)(cid:12) (Φ ◦ w)(x1, x′)(cid:12)(cid:12)(cid:12) p dx ≤ (C1)pQm∂1w(x)p dx. Since this is true for every coordinate, Morrey's characterization of Sobolev maps [11, Theorem 4.21] implies that Φ ◦ w ∈ W 1,p(Qm; Rν2 ) and, for almost every x ∈ Qm, it follows from the counterpart of identity (2.2) for each coordinate that D(Φ ◦ w)(x) = DΦ(w(x)) ◦ Dw(x). (cid:3) Proof of Proposition 2.1. Let u : Qm → N n and let us assume that ι1◦u ∈ W 1,p(Qm; Rν1 ). The smooth map Φ := ι2 ◦ ι−1 is defined on the embedded complete submanifold ι1(N n) ⊂ Rν1 with values into ι2(N n) ⊂ Rν2. Since ι2 ◦ ι−1 is an isometry, DΦ is bounded on the tangent bundle T (ι1(N n)). We can thus apply Lemma 2.2 to w = ι1 ◦ u. This implies that ι2 ◦ u ∈ W 1,p(Qm; Rν2) and, for almost every x ∈ Qm, we have 1 1 D(ι2 ◦ u)(x) = DΦ((ι1 ◦ u)(x)) ◦ D(ι1 ◦ u)(x). Let uj : Qm → N n be a sequence of measurable maps such that (ι1 ◦ uj)j∈N converges to ι1 ◦ u in W 1,p(Qm; Rν1 ). This implies the convergence in measure of the functions ι1 ◦ uj and their derivatives. Since Φ is C1, one deduces the same convergence in measure for ι2 ◦ uj = Φ(ι1 ◦ uj). For every j ≥ 1 and almost every x ∈ Qm, the quantity D(ι2 ◦ uj)(x) − D(ι2 ◦ u)(x)p, which is equal to D(Φ ◦ ι1 ◦ uj)(x) − D(Φ ◦ ι1 ◦ u)(x)p, is dominated by C1(cid:0)D(Φ ◦ ι1 ◦ uj)(x)p + D(Φ ◦ ι1 ◦ u)(x)p(cid:1). Lemma 2.2 and the boundedness of DΦ on T (ι1(N n)) yield D(ι2 ◦ uj)(x) − D(ι2 ◦ u)(x)p ≤ C2(cid:0)D(ι1 ◦ uj)(x)p + D(ι1 ◦ u)(x)p(cid:1). Since (D(ι2 ◦ uj))j∈N converges to D(ι2 ◦ u) in measure and the right-hand side converges in L1(Qm), the dominated convergence theorem implies that (D(ι2 ◦ uj))j∈N converges to D(ι2 ◦ u) in Lp(Qm; Rm×ν2). DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 7 The convergence of (ι2 ◦ uj)j∈N to ι2 ◦ u in Lp(Qm; Rν2 ) follows from the conver- gence in measure and the convergence of the derivatives in Lp. Indeed, for every ε > 0 we have Qmι2 ◦ uj − ι2 ◦ up ≤ (2ε)pQm + ι2 ◦ uj − ι2 ◦ up. {ι2◦uj −ι2◦u≥2ε} By the convergence in measure, for every j ∈ N large enough we have {ι2 ◦ uj − ι2 ◦ u ≥ ε} ≤ Qm/2. Defining θε : [0, +∞) → R by if t ≤ ε, if ε < t < 2ε, if t ≥ 2ε, 0 2(t − ε) t θε(t) := it follows from the Poincaré inequality for functions vanishing on a set of positive measure that {ι2◦uj −ι2◦u≥2ε} ι2 ◦ uj − ι2 ◦ up ≤ Qm(cid:0)θε(ι2 ◦ uj − ι2 ◦ u)(cid:1)p ≤ C3QmD(ι2 ◦ uj − ι2 ◦ u)p. By the convergence of (D(ι2 ◦ uj))j∈N to D(ι2 ◦ u) in Lp(Qm; Rm×ν2 ), we then have lim sup j→∞ Qmι2 ◦ uj − ι2 ◦ up ≤ (2ε)pQm. Since ε is arbitrary, this proves the convergence of (ι2◦uj)j∈N to ι2◦u in Lp(Qm; Rν2 ). The proof is complete. (cid:3) 2.2. Bounded maps. Let N n be an (abstract) complete Riemannian manifold. We say that a measurable map u : Qm → N n is essentially bounded if it is essentially bounded for the geodesic distance distN n induced by the Riemannian metric on N n: there exists C > 0 such that, for almost every x, y ∈ Qm, distN n (u(x), u(y)) ≤ C. Since N n is complete, this is equivalent to the existence of a compact set K ⊂ N n such that, for almost every x ∈ Qm, u(x) ∈ K. We now consider an isometric embedding ι : N n → Rν. If the map u is es- sentially bounded, then ι ◦ u ∈ L∞(Qm; Rν). In general, the converse is not true because there exists ν ∈ N∗, depending on the dimension n, such that the mani- fold N n can be isometrically embedded inside a ball of any radius r > 0 in Rν, see [23, Theorem 2; 24, Theorem 3]. We can discard this phenomenon under the additional assumption that the em- bedding ι : N n → Rν is a proper map from N n to Rν; that is, for every compact set L ⊂ Rν, the set ι−1(L) is a compact subset of N n. Since ι is a homeomorphism from N n to ι(N n), this amounts to the property that ι(N n) is a closed subset of Rν. Such proper isometric embeddings of complete Riemaniann manifolds have been constructed by a shrewd application of the classical Nash embedding theo- rem [22]. The next proposition summarizes the preceding discussion: 8 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN Proposition 2.3. If N n is a complete Riemannian manifold, then there exists an isometric embedding ι : N n → Rν such that ι(N n) is closed. For such an embedding, the map u : Qm → N n is essentially bounded if and only if the map ι◦ u : Qm → Rν is essentially bounded. In the sequel, we work exclusively with proper isometric embeddings ι : N n → Rν, and we systematically identify N n with ι(N n). 2.3. The nearest point projection. An important tool for the approximation of Sobolev maps when N n is a closed submanifold of Rν is the fact that N n is a smooth retraction of a neighborhood of itself. More precisely, the nearest point projection Π is well-defined and smooth on an open neighborhood O ⊂ Rν of N n. The map Π ∈ C∞(O; N n) satisfies the following properties: (a) for every (y, z) ∈ O × N n, distRν (y, Π(y)) ≤ distRν (y, z); (b) in particular, Π(y) = y for every y ∈ N n. It follows that DΠ is bounded on N n. For our purposes in this paper, the map Π could be replaced by any retraction eΠ ∈ C1(O; N n) with bounded derivative. The existence of such a map only requires N n to be C1, as a consequence of [28, Theorem 10A]. In contrast, the nearest point projection onto a C1 submanifold is merely continuous in general. Reducing the size of O if necessary, we note that (1) by continuity of DΠ one can assume that DΠ is bounded on O, although this does not imply that the map Π is globally Lipschitz-continuous on O, (2) by closedness of N n, the map Π : O → N n can be extended as a smooth map from Rν to Rν, whence the chain rule of Marcus and Mizel [21, Theorem 2.1] or Lemma 2.2 above with M m = Rν can be applied to Π ◦ w when w is a Sobolev map with values into a closed subset F ⊂ O. We assume henceforth that O ⊃ N n is chosen so that the smooth map Π : O → N n satisfies Properties (1) and (2) above. 3. APPROXIMATION OF BOUNDED MAPS BY SMOOTH MAPS 3.1. High-integrability case. When p > m, maps in W 1,p(Qm; N n) are essentially bounded and continuous, and can be approximated uniformly via convolution by smooth maps with values in Rν. The nearest point projection Π defined in Section 2.3 allows one to project the sequence back to N n. Proposition 3.1. If p > m, then, for every u ∈ W 1,p(Qm; N n), there exists a sequence in C∞(Qm; N n) converging strongly to u in W 1,p(Qm; Rν). 2 with inra- Proof. We first extend the map u by reflection on the larger cube Qm dius equal to 2. We still denote by u the resulting map, which now belongs to W 1,p(Qm 2 ; N n). Given a family of mollifiers (ϕε)ε>0 of the form ϕε(z) = ϕ(z/ε)/εm with ϕ ∈ c (Bm 1 ), we have C∞ and thus, by the Morrey–Sobolev embedding, ε→0 kϕε ∗ u − ukW 1,p(Qm) = 0, lim ε→0kϕε ∗ u − ukL∞(Qm) = 0. lim DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 9 Since the map u is essentially bounded on Qm, there exists a compact set K ⊂ N n ιK ⊂ O, such that u(x) ∈ K for almost every x ∈ Qm. Take ιK > 0 such that K + Bν where O is the tubular neighborhood on which the nearest point projection Π is defined and smooth, see Section 2.3. For every ε > 0 sufficiently small, (ϕε ∗ ιK . Then, the C1 regularity of Π implies that Π ◦ (ϕε ∗ u) ∈ u)(Qm) ⊂ K + Bν W 1,p(Qm; N n) and such a family of maps converges to Π ◦ u = u in W 1,p(Qm; Rν). This completes the proof. (cid:3) 3.2. Critical-integrability case. When p = m, maps in W 1,m(Qm; N n) need not be continuous nor even bounded. However, smooth maps taking their values into the manifold N n are always dense in (W 1,m ∩ L∞)(Qm; N n), and this essentially follows from the seminal work of Schoen and Uhlenbeck for compact manifolds [26]. Proposition 3.2. For every u ∈ (W 1,m ∩ L∞)(Qm; N n), there exists a sequence in C∞(Qm; N n) converging strongly to u in W 1,m(Qm; Rν). Once again, the proof of Proposition 3.2 is based on an approximation by a family of mollifiers. In contrast to the case p > m considered in Proposition 3.1, the family (ϕε ∗ u)ε>0 need not converge uniformly to u. However, as in the setting of compact manifolds [26], one can exploit the vanishing mean oscillation property satisfied by maps in the critical integrability space, as has been observed in [8]. Indeed, this property guarantees that the approximating sequence takes its values in a small neighborhood of N n. One can then project it back on N n by using the nearest point projection Π. Proof of Proposition 3.2. Let u ∈ (W 1,m ∩ L∞)(Qm; N n). As in Proposition 3.1, we may assume that u ∈ (W 1,m ∩ L∞)(Qm 2 ; N n) and consider the maps ϕε ∗ u. Let K ⊂ N n be a compact subset such that u(x) ∈ K for almost every x ∈ Qm. By the Poincaré–Wirtinger inequality, for every x ∈ Qm and every 0 < ε < 1, (cid:0) distRν ((ϕε ∗ u)(x), K)(cid:1)m dy ≤ C1Bm ε (x)Dum, for some constant C1 > 0 independent of ε. The quantity in the right-hand side converges uniformly to 0 with respect to x as ǫ tends to 0. ε (x)(cid:12)(cid:12)ϕε ∗ u(x) − u(y)(cid:12)(cid:12)m ≤ Bm Taking ιK > 0 such that K + Bν ιK ⊂ O, we deduce from the estimate above that there exists ¯ε > 0 such that, for every 0 < ε ≤ ¯ε and every x ∈ Qm, we have distRν ((ϕε ∗ u)(x), K) ≤ ιK. We can then consider the family(cid:0)Π ◦ (ϕε ∗ u)(cid:1)0<ε≤¯ε and conclude as in the proof of Proposition 3.1. (cid:3) 3.3. Low-integrability case. The low integrability case p < m is the most delicate, but can be settled by the results and methods used to handle the density of smooth maps when the target manifold N n is compact. In general, smooth maps are not dense without an additional topological assumption on N n, but a larger class of maps admitting (m − 1 − ⌊p⌋)-dimensional singularities is dense. 10 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN Proposition 3.3. Let 1 ≤ p < m. For every u ∈ (W 1,p ∩ L∞)(Qm; N n), there exists a sequence in (Rm−⌊p⌋−1 ∩ L∞)(Qm; N n) converging strongly to u in W 1,m(Qm; Rν). If moreover π⌊p⌋(N n) is trivial, then such a sequence can be taken in C∞(Qm; N n). In the statement above we denote by Ri(Qm; N n), for i ∈ {0, . . . , m − 1}, the set of maps u : Qm → N n which are smooth on Qm \ T and such that, for every x ∈ Qm \ T , Du(x) ≤ dist(x, T ) , C where T is a finite union of i-dimensional hyperplanes. Here, the set T and the constant C > 0 depend on u. The proof of Proposition 3.3 relies on the next lemma which allows one to iden- tify a bounded map into a complete manifold as a map into a compact manifold. Lemma 3.4. Let N n for every compact set K in the relative interior of N n submanifold Ln without boundary of Rν × R such that 0 be a smooth compact submanifold of Rν with boundary. Then, 0 , there exists a smooth compact and P (Ln) ⊂ N n 0 , where P : Rν × R → Rν denotes the projection P (x, s) = x. K × {0} ⊂ Ln The idea of the proof of Lemma 3.4 is to glue together two identical copies of 0 along the boundary. To avoid the creation of singularities, the gluing is done N n along a tube diffeomorphic to ∂N0 × [0, 1]. To avoid the intersection between the two copies, they are placed in distinct ν-dimensional affine hyperplanes of the space Rν+1. 0 . By the collar Proof of Lemma 3.4. Let K be a compact subset in the interior of N n neighborhood theorem (see for example [20, Theorem 1.7.3]), there exist a relative open neighborhood U of ∂N n 0 in N n 0 and a smooth diffeomorphism 0 × [0, 1] → N n f : ∂N n 0 ∩ U 0 × {1} and f −1(∂U ∩ N n 0 × {0}. By reducing 0 ) = ∂N n such that f −1(∂N n the size of U if necessary, we can assume that U ∩ K = ∅. Let α, β : [0, 1] → [0, 1] be two smooth functions such that 0 ) = ∂N n α(t) =(t 1 − t if t < 1/4, if t > 3/4, and β(t) =(0 if t < 1/8, 1 if t > 7/8. We also require that β be nondecreasing and β′ > 0 on the interval [1/4, 3/4]. We now define the set We observe that 0 \ U ) × {0, 1}(cid:1) ∪(cid:8)(cid:0)f (z, α(t)), β(t)(cid:1) : z ∈ ∂N n Ln =(cid:0)(N n 0 × {0}) ∩(cid:8)(cid:0)f (z, α(t)(cid:1), β(t)(cid:1) : z ∈ ∂N n 0 , t ∈ (0, 1)(cid:9) =(cid:8)(f (z, t), 0) : z ∈ ∂N n (N n 0 , t ∈ (0, 1)(cid:9). 0 , t ∈ (0, t0](cid:9), DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 11 where t0 = max{t : β(t) = 0}. Similarly, 0 × {1}) ∩(cid:8)(cid:0)f (z, α(t)(cid:1), β(t)) : z ∈ ∂N n (N n 0 , t ∈ (0, 1)(cid:9) =(cid:8)(cid:0)f (z, 1 − t), 1(cid:1) : z ∈ ∂N n 0 , t ∈ [t1, 1)(cid:9), where t1 = min{t : β(t) = 1}. This implies that Ln is a smooth submanifold of Rν+1. By construction, Ln is compact and has no boundary. Moreover, K × {0} 0 follows from the fact that Ln ⊂ is contained in Ln. The inclusion P (Ln) ⊂ N n 0 × [0, 1]. N n (cid:3) In the proof of Proposition 3.3 that we present below, we first observe that the range of a map u ∈ (W 1,p∩L∞)(Qm; N n) is contained in a compact set. Hence, the map u can be identified to an element of W 1,p(Qm; Ln), where Ln is the compact manifold given by Lemma 3.4. For a compact target manifold, the density of the class Rm−⌊p⌋−1(Qm; Ln) in W 1,p(Qm; Ln) has been proved in [2, 5, 16]. The retrac- tion P from Lemma 3.4 then allows one to bring an approximating sequence back to the original manifold N n. This approach cannot work for the density of C∞(Qm; N n) in the space (W 1,p∩ L∞)(Qm; N n). Indeed, there is no guarantee that the manifold Ln inherits the topo- logical assumption satisfied by N n. For example, by gluing two balls Bn one gets a manifold which is diffeomorphic to the sphere Sn, while the homotopy group πn(Sn) is nontrivial. Instead, once the density of the class Rm−⌊p⌋−1(Qm; N n) is proved, one can proceed along the lines of the proof in the compact setting [2,5,16]. Proof of Proposition 3.3. Given u ∈ (W 1,p∩ L∞)(Qm; N n), the essential range of u is contained in a compact subset K of a smooth compact submanifold N n 0 ⊂ N n with boundary. Let Ln be a compact smooth submanifold of Rν+1 satisfying the prop- erties of Lemma 3.4. Then u belongs to W 1,p(Qm; Ln). By [5, Theorem 2], there exists a sequence of maps (uj)j∈N in Rm−⌊p⌋−1(Qm; Ln) which converges to u in W 1,p(Qm; Rν+1). This implies that the sequence (P ◦ uj)j∈N in Rm−⌊p⌋−1(Qm; N n) still converges to P ◦ u = u in W 1,p(Qm; Rν). Since P (Ln) ⊂ N n 0 , the sequence (P ◦ uj)j∈N is also contained in the space L∞(Qm; N n). This completes the proof of the first part of the proposition. If we further assume that π⌊p⌋(N n) is trivial, then we can approximate each map P (uj) ∈ Rm−⌊p⌋−1(Qm; N n) by a sequence of smooth maps in C∞(Qm; N n). As in the case of compact-target manifolds, one relies here on the existence of a smooth projection from a tubular neighborhood of a compact subset of N n into N n, see Section 7 and the Claim in Section 9 of [5]. By a diagonal argument, this implies that u itself belongs to the closure of C∞(Qm; N n). (cid:3) 4. LACK OF STRONG DENSITY IN W 1,n(Qm; N n) In this section we give an example of a complete manifold for which (W 1,n ∩ L∞)(Qm; N n) is not strongly dense in W 1,n(Qm; N n) with n ≤ m. We state the main result of this section as follows: Proposition 4.1. Let ν ∈ N∗, m, n ∈ N∗ be such that m ≥ n ≥ 2, and let a ∈ Sn. For every smooth embedding F : Sn \ {a} → Rν such that lim x→aF (x) = +∞ and DF ∈ Ln(Sn \ {a}), the closed submanifold N n = F (Sn \ {a}) equipped with the 12 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN Riemannian metric inherited from Rν is complete, but (W 1,n ∩ L∞)(Qm; N n) is not strongly dense in W 1,n(Qm; N n). For instance, one may take F (x) = λ(x)x, where λ : Sn \ {a} → R is a positive λ(x) = +∞. This is always smooth function such that λ ∈ W 1,n(Sn; R) and lim possible in dimension n ≥ 2, and an example is given by setting distSn (x, a)(cid:17)α λ(x) =(cid:16) log for x in a neighborhood of a and any exponent 0 < α < n−1 n . x→a 1 Proposition 4.1 follows from the fact that F cannot be approximated in W 1,n(Sn; Rn+1) by a sequence in C∞(Sn; N n). In turn, the latter is proved by contradiction: the ranges of such approximating maps when restricted to a small neighborhood V of a in Sn would contain a fixed compact subset K of N n. By the area formula, the possibility of taking V arbitrarily small contradicts the equi-integrability of the sequence in W 1,n. Proof of Proposition 4.1. Given a diffeomorphism f : Qn → Sn between Qn and a closed neighborhood of a in Sn, the function u = F ◦ f (4.1) belongs to W 1,n(Qn; N n). We first handle the case m = n by proving that u cannot be approximated by bounded maps in W 1,n(Qn; N n). We proceed by contradic- tion: if u is in the closure of the set (W 1,n∩ L∞)(Qn; N n) then, by density of the set C∞(Qn; N n) in the former space (Proposition 3.2), there exists a sequence of maps (uk)k∈N in C∞(Qn; N n) converging strongly to u in W 1,n(Qn; Rν). Without loss of generality, we may assume that f (0) = a. Given a compact subset K ⊂ N n with Hn(K) > 0, since the embedding F diverges at the point a, there exists 0 < δ < 1 such that K ∩ u(Qn δ ) = ∅. By a Fubini-type argument, there exists a subsequence (ukj )j∈N such that, for al- r ; Rν), whence most every r ∈ (0, 1), (ukj∂Qn also uniformly by the Morrey–Sobolev inequality. Since for each such r ≤ δ we r )j∈N there exists Jr ∈ N have K ∩ u(∂Qn such that, for every j ≥ Jr, r ) = ∅, by uniform convergence of (ukj∂Qn r )j∈N converges to u∂Qn r in W 1,n(∂Qn kukj − ukL∞(∂Qn r ) < dist(K, u(∂Qn r )). In particular, K ∩ ukj (∂Qn r ) = ∅. We claim that (4.2) K ⊂ ukj (Qn r ). To prove this, we take a homeomorphism g : Qn f∂Qn r . Since K ∩(cid:0)F ◦ f (Qn δ )(cid:1) = ∅, this implies that r ) = F (Sn \ f (Qn F ◦ g(Qn r )) ⊃ K. r → Sn \ f (Qn r ) such that g∂Qn r = By continuity of the Brouwer degree with respect to the uniform convergence, for every y ∈ K and for every j ≥ Jr we have deg (ukj , Qn r , y) = deg (F ◦ g, Qn r , y) 6= 0. DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 13 FIGURE 1. The algebraic manifold N n This implies Claim (4.2). By monotonicity of the Hausdorff measure and by the area formula, we then have Hn(K) ≤ Hn(cid:0)ukj (Qn r )(cid:1) ≤ Qn r jac ukj , where jac ukj =(cid:0)det ((Dukj )∗ ◦ Dukj )(cid:1)1/2. Using the pointwise inequality jac ukj ≤ C1Dukjn, as j tends to infinity we get Hn(K) ≤ C1Qn r Dun. Since the right-hand side tends to zero as r tends to zero, we have a contradiction. Hence the density of bounded maps fails in the space W 1,n(Qn; N n), and there exists ε > 0 such that, for every w ∈ (W 1,n ∩ L∞)(Qn; N n), (4.3) QnDw − Dun ≥ ε. When m > n, we consider the map u ◦ P , where u is defined by (4.1) and P : Qm = Qn × Qm−n → Qn is the projection on the first component. Given v ∈ (W 1,n ∩ L∞)(Qm; N n), it follows for almost every y′′ ∈ Qm−n that v(·, y′′) ∈ (W 1,n ∩ L∞)(Qn; N n). By Fubini's theorem and the lower bound (4.3) applied to w = v(·, y′′), we get QmDv − D(u ◦ P )n = Qm−n(cid:18) QnDv(y′, y′′) − Du(y′)n dy′(cid:19) dy′′ ≥ 2m−nε. Hence, u ◦ P does not belong to the closure of (W 1,n ∩ L∞)(Qm; N n). (cid:3) An alternative example, this time of an algebraic complete manifold N n for which (W 1,n ∩ L∞)(Qn; N n) is not strongly dense in W 1,n(Qn; N n), is (1 + y1)2β−1o, N n =ny = (y1, y′) ∈ R+ × Rn : y′2 = y1 n−1 , see Figure 1. The lack of density can be obtained using the map where β > n u : Qn → N n defined for x ∈ Qn \ {0} by (4.4) u(x) =(cid:18)ϕ(x), pϕ(x) (1 + ϕ(x))β− 1 2 x x(cid:19), where ϕ : (0,∞) → R+ is a smooth function such that ϕ(r) = log rγ for r ∈ (0, 1/3), ϕ(r) = 0 for r ∈ (2/3,∞) and ϕ′/√ϕ is bounded on (1/3, 2/3). Then, u belongs to W 1,n(Qn; N n) provided that n . An adaptation of the proof of Proposition 4.1 shows that there exists no sequence of maps in C∞(Qn; N n) converging strongly to u in W 1,n(Qn; N n). n(β−1) < γ < n−1 1 14 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN Hajłasz and Schikorra [15, Section 3] have provided examples of noncompact manifolds N n for which Lipschitz maps are not strongly dense in W 1,n(Qn; N n). Instead of taking an embedding F that blows up at some point a as we do, they construct an embedding that is not proper but strongly oscillates in a neighbor- hood of the point a. 5. MAIN TOOLS In this section we explain the main tools used in the proofs of Theorems 1 and 2. 5.1. The opening technique. We recall the technique of opening of maps that has been introduced by Brezis and Li [7] and pursued in [5, Section 2]. To il- lustrate the main idea, we explain this tool in a model situation. Given a map u ∈ W 1,p(Rm; Rν ), we wish to construct a smooth map Φ : Rm → Rm such that (a) u ◦ Φ is constant in Qm 1 ; (b) u ◦ Φ = u in Rm \ Qm 4 ; (c) u ◦ Φ ∈ W 1,p(Rm; Rν ) and ku ◦ ΦkW 1,p(Rm) ≤ CkukW 1,p(Rm). The opening construction is based on the following elementary inequality [5, Lemma 2.5]: 1 (cid:18)Qm Qm f ◦ Φz(cid:19) dz ≤ 6mRm whose proof is based on Fubini's theorem and is valid for every nonnegative func- tion f ∈ L1(Rm), where Φz : Rm → Rm is defined by Φz(x) = ζ(x + z) − z f, 5 and ζ : Rm → Rm is any smooth function. Assuming that ζ = 0 in Qm 1 that Φz is constant in Qm in Rm \ Qm in Rm \ Qm 3 , we then have for every z ∈ Qm 4 . Formally taking f = up + Dup in the inequality above, one gets 2 and ζ = Id 1 and Φz = Id ku ◦ Φzkp and then it suffices to take z ∈ Qm 1 Qm 1 such that W 1,p(Qm 5 ) dz ≤ 6mkukp W 1,p(Rm), ku ◦ Φzkp W 1,p(Qm 5 ) ≤ 2 · 6mkukp W 1,p(Rm). This formal argument can be rigorously justified using an approximation of u by smooth functions in W 1,p(Rm; Rν), see [5, Lemma 2.4] and also [17, Section 7]. Such an averaging procedure is reminiscent of the work of Federer and Flem- ing [12] and was adapted to Sobolev functions by Hardt, Kinderlehrer and Lin [18]. More generally, one can open a map around a small neighborhood of 0, or along the normals to a planar set. For example, the singleton {0} may be replaced by a relative open subset of an ℓ-dimensional plane, with ℓ ∈ {1, . . . , m − 1}. In this case, we obtain an opened Sobolev map depending locally on ℓ variables, and constant along m − ℓ normal directions. The following statement coincides with [5, Proposition 2.2] and we omit the proof. Proposition 5.1. Let ℓ ∈ {0, . . . , m − 1}, η > 0, 0 < ρ < ρ and A ⊂ Rℓ be an open set. For every u ∈ W 1,p(A × Qm−ℓ ; Rν), there exists a smooth map ζ : Rm−ℓ → Rm−ℓ such that ρη DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 15 , ρη (i) ζ is constant in Qm−ℓ (ii) {x ∈ Rm : ζ(x) 6= x} ⊂ Qm−ℓ (iii) if Φ : Rm → Rm is defined for every x = (x′, x′′) ∈ Rℓ × Rm−ℓ by ) ⊂ Qm−ℓ and ζ(Qm−ℓ ρη ρη ρη , Φ(x) =(cid:0)x′, ζ(x′′)(cid:1), ; Rν) and then u ◦ Φ ∈ W 1,p(A × Qm−ℓ ρη kD(u ◦ Φ)kLp(A×Qm−ℓ ρη ) ≤ CkDukLp(A×Qm−ℓ ρη ), for some constant C > 0 depending on m, p, ρ and ρ. We observe that (iii) implies that Φ is constant on the (m−ℓ)-dimensional cubes of inradius ρη which are orthogonal to A. The map u ◦ Φ thus only depends on ℓ variables in a neighborhood of A. In order to present the opening technique in the framework of cubications, we first need to introduce some vocabulary. First, given a set A ⊂ Rm and η > 0, a cubication of A of inradius η > 0 is a family of closed cubes Sm of inradius η such that (1) Sσm∈Sm dimension i ∈ {0, . . . , m}. For ℓ ∈ {0, . . . , m}, the set Sℓ of all ℓ-dimensional faces of all cubes in Sm is called the skeleton of dimension ℓ. We then denote by Sℓ the union of all elements of Sℓ: 2 ∈ Sm which are not disjoint, σm 2 is a common face of (2) for every σm 1 ∩ σm σm = A, 1 , σm A subskeleton E ℓ of Sℓ is simply a subfamily of Sℓ and the associated subset of Rm is Sℓ = [σℓ∈S ℓ σℓ. Eℓ = [σℓ∈E ℓ σℓ. Accordingly, given a set A in Rm equipped with a cubication, a subskeleton of A is a subfamily of the ℓ-dimensional skeleton of the given cubication. We proceed to state the main result of this section, which is essentially [5, Propo- sition 2.1]. Proposition 5.2. Let p ≥ 1, ℓ ∈ {0, . . . , m − 1}, η > 0, 0 < ρ < 1 subskeleton of Rm of inradius η. Then, for every u ∈ W 1,p(Eℓ + Qm a smooth map Φ : Rm → Rm such that (i) for every i ∈ {0, . . . , ℓ} and for every i-dimensional face σi ∈ E i, Φ is constant on (ii) {x ∈ Rm : Φ(x) 6= x} ⊂ Eℓ + Qm 2ρη; Rν), and (iii) u ◦ Φ ∈ W 1,p(Eℓ + Qm the (m − i)-dimensional cubes of inradius ρη which are orthogonal to σi, σℓ + Qm 2ρη and, for every σℓ ∈ E ℓ, Φ(σℓ + Qm 2 , and E ℓ be a 2ρη; Rν), there exists 2ρη) ⊂ 2ρη , kD(u ◦ Φ)kLp(Eℓ+Qm 2ρη ) ≤ CkDukLp(Eℓ+Qm 2ρη ), for some constant C > 0 depending on m, p and ρ, 16 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN (iv) for every σℓ ∈ E ℓ, kD(u ◦ Φ)kLp(σℓ+Qm 2ρη ) ≤ C′kDukLp(σℓ+Qm 2ρη ), for some constant C′ > 0 depending on m, p and ρ. Here Qm r (x) = x + rQm is the cube centered at x with inradius r. For the convenience of the reader, and also because Assertion (ii) in Proposition 5.2 is slightly more precise than the corresponding statement in [5], we sketch its proof. dimensional cubes of inradius ρiη which are orthogonal to σr, Proof of Proposition 5.2. Let ρ = ρℓ < ··· < ρi < ··· < ρ0 < 2ρ. We define by induction on i ∈ {0, . . . , ℓ} a map Φi : Rm → Rm such that (a) for every r ∈ {0, . . . , i} and every σr ∈ E r, Φi is constant on the (m − r)- (b) {x ∈ Rm : Φi(x) 6= x} ⊂ Ei + Qm (c) u ◦ Φi ∈ W 1,p(Eℓ + Qm (d) for every σi ∈ E i, 2ρη and, for every σi ∈ E i, Φi(σi + Qm 2ρη) ⊂ 2ρη; Rν), σi + Qm 2ρη , kD(u ◦ Φi)kLp(σi+Qm 2ρη ) ≤ CkDukLp(σi+Qm 2ρη ), for some constant C > 0 depending on m, p and ρ. The map Φℓ will satisfy the conclusion of the proposition. For i = 0, E 0 is the set of vertices of cubes in E m. The map Φ0 is obtained by applying Proposition 5.1 to u around each σ0 ∈ E 0 with parameters ρ0 < 2ρ and ℓ = 0. Assume that the maps Φ0, . . . , Φi−1 have been constructed. We then apply Proposition 5.1 to u ◦ Φi−1 around each σi ∈ E i with parameters ρi < ρi−1. This gives a smooth map Φσi : Rm → Rm which is constant on the (m− i)-dimensional cubes of inradius ρiη orthogonal to σi. Let σi 1, σi 2 ∈ E i such that (σi 1 + Qm ρi−1η) ∩ (σi 2 + Qm ρi−1η) 6= ∅. We claim that for every x in this set, (5.1) Φi−1(Φσi 1 (x)) = Φi−1(Φσi 2 (x)) = Φi−1(x). j 1 and σi By the formula of Φσi ρi−1η for some face τ r ∈ E r with τ r ⊂ σi (x), Φσi Indeed, since σi 2 are not disjoint, we can take the smallest integer r ∈ {0, . . . , i − 1} such that x ∈ τ r + Qm 2. 1 ∩ σi given in Proposition 5.1, the points Φσi (x) and x belong to the same (m−r)-dimensional cube of inradius ρi−1η which is orthogonal to τ r. By induction, Φi−1 is constant on the (m − r)-dimensional cubes of inradius ρi−1η which are orthogonal to τ r. This proves claim (5.1). We can thus define the map Φi : Rm → Rm as follows: if x ∈ σi + Qm otherwise. Φi(x) =(Φi−1(Φσi (x)) Φi−1(x) ρi−1η, 1 2 Assertion (a) follows from the above discussion which implies in particular that Φi = Φi−1 on Ei−1 + Qm ρi−1η. DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 17 We proceed with the proof of assertion (b). By definition of the map Φi, we ρi−1η). By induction, Φi−1 agrees with the identity have Φi = Φi−1 on Rm \ (Ei + Qm outside Ei−1 + Qm 2ρη. Hence, {x ∈ Rm : Φi(x) 6= x} ⊂ Ei + Qm 2ρη. Moreover, by induction, for every τ i−1 ∈ E i−1, 2ρη. Since Φi−1(x) = x for 2ρη) ⊂ τ i−1 + Qm 2ρη. 2ρη) ⊂ ∂σi + Qm 2ρη) ∩ (σi + Qm 2ρη) ⊂ ∂σi + Qm 2ρη , it follows that 2ρη and (Ei−1 + Qm Φi−1(τ i−1 + Qm Thus for every σi ∈ E i, Φi−1(∂σi + Qm x 6∈ Ei−1 + Qm (5.2) Now, let x ∈ σi + Qm it follows that Φi(x) ∈ σi + Qm τ i ∈ E i be such that x ∈ τ i + Qm x ∈ τ r + Qm particular, Φτ i(x) ∈ σi + Qm This completes the proof of (b). Φi−1(σi + Qm 2ρη. If x 6∈ Ei + Qm ρi−1η, then Φi(x) = Φi−1(x). From (5.2), 2ρη. We now assume that x ∈ Ei + Qm ρi−1η. Let ρi−1η. Then the cube τ r := τ i ∩ σi is not empty, 2ρη. In 2ρη and thus by (5.2), Φi(x) = Φi−1(Φτ i(x)) ∈ σi + Qm 2ρη. The proofs of Assertions (c) and (d) are the same as in [5] and we omit them. (cid:3) 2ρη and from the form of Φτ i, we deduce that Φτ i(x) ∈ τ r + Qm 2ρη) ⊂ σi + Qm 2ρη. When ℓ ≤ p + 1, the function u ◦ Φ given by Proposition 5.2 satisfies (5.3) 1 rm−p Qm r (x)D(u ◦ Φ)p ≤ C ηm−p τ ℓ−1+Qm ρη D(u ◦ Φ)p, ρη and every face τ ℓ−1 ∈ E ℓ−1. This estimate for every cube Qm follows from the fact that Φ is constant on the (m − ℓ + 1)-dimensional cubes of inradius ρη which are orthogonal to τ ℓ−1. r (x) ⊂ τ ℓ−1 + Qm Indeed, without loss of generality, we can assume that τ ℓ−1 has the form [−η, η]ℓ−1× {0′′}, where 0′′ ∈ Rm−ℓ+1. Accordingly, we write every y ∈ τ ℓ−1 + Qm (y′, y′′) ∈ Rℓ−1 × Rm−ℓ+1. By construction, for every y ∈ τ ℓ−1 + Qm (u ◦ Φ)(y′, 0′′). This implies ρη as y = ρη , (u ◦ Φ)(y) = Qm r (x)D(u ◦ Φ)(y)p dy ≤ C1rm−ℓ+1Qℓ−1 ηm−ℓ+1 Qℓ−1 ≤ C2 ηm−ℓ+1 τ ℓ−1+Qm ≤ C2 rm−ℓ+1 rm−ℓ+1 r r ρη (x′)D(u ◦ Φ)(y′, 0′′)p dy′ (x′)×Qm−ℓ+1 ρη (0′′)D(u ◦ Φ)(y′, y′′)p dy′ dy′′ D(u ◦ Φ)(y′, y′′)p dy′ dy′′. (x′) × Qm−ℓ+1 ρη r (x) ⊂ τ ℓ−1 + Qm η(cid:17)p−ℓ+1τ ℓ−1+Qm ρη C3 ηm−p(cid:16) r D(u ◦ Φ)p, (0′′) ⊂ τ ℓ−1 + Qm ρη ρη and the explicit In the last line, we have used the fact that Qℓ−1 which in turn follows from the assumption Qm form of τ ℓ−1. Hence, r 1 rm−p Qm r (x)D(u ◦ Φ)p ≤ which proves estimate (5.3) since r ≤ η and ℓ ≤ p + 1. 18 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN 5.2. Adaptive smoothing. A second tool is the adaptive smoothing, in which the function is smoothened by mollification at a variable scale. More precisely, given ϕ = 1, u ∈ L1 define the map ϕψ ∗ u : Rm → Rν (5.4) loc(Rm; Rν) and a nonnegative function ϕ ∈ C∞ 1 ) such that ´Bm c (Bm 1 u(x − ψ(x)y)ϕ(y) dy, (ϕψ ∗ u)(x) = Bm 1 where ψ ∈ C∞(Ω) is a nonnegative function that plays the role of the variable parameter. Observe that if ψ(x) > 0, then by an affine change of variable we have (ϕψ ∗ u)(x) = otherwise, ψ(x) = 0 and since ´Bm 1 1 ψ(x)m Ω u(z)ϕ(cid:16) x − z ψ(x)(cid:17) dz; ϕ = 1 we have ϕψ ∗ u(x) = u(x). The adaptive smoothing has an immediate counterpart for functions u ∈ L1 in an open subset Ω ⊂ Rm. In this case, we define ϕψ ∗ u by (5.4) at points x in the open set loc(Ω; Rν) (5.5) For Sobolev maps u the following property holds [5, Propositions 3.1 and 3.2]: ω =(cid:8)x ∈ Ω : dist (x, ∂Ω) > ψ(x)(cid:9). Proposition 5.3. If u ∈ W 1,p(Ω; Rν ) and kDψkL∞(Ω) < 1, then ϕψ ∗ u ∈ W 1,p(ω; Rν) where ω is given by (5.5). Moreover, the following estimates hold: 1 kτψv(u) − ukLp(ω), kϕψ ∗ u − ukLp(ω) ≤ sup v∈Bm kD(ϕψ ∗ u)kLp(ω) ≤ C (1 − kDψkL∞(ω)) 1 p kDukLp(Ω), and kD(ϕψ ∗ u) − DukLp(ω) ≤ sup v∈Bm 1 kτψv(Du) − DukLp(ω) + for some constants C, C′ > 0 depending on m and p, where A = C′ (1 − kDψkL∞(ω)) 1 p kDukLp(A), Bm ψ(x)(x). Sx∈ω∩supp Dψ In this statement, τv(u) : Ω + v → Rν denotes the translation with respect to the vector v ∈ Rm defined for each x ∈ Ω + v by τv(u)(x) = u(x − v). 5.3. Zero-degree homogenization. This tool has been used in problems involv- ing compact-target manifolds [2, 4, 16], and allows one to extend a Sobolev map u defined on the boundary of a star-shaped domain to the whole domain, by pre- serving the range of u. We first recall the notion of a Sobolev map on skeletons [16]: Definition 5.1. Given p ≥ 1, ℓ ∈ {0, . . . , m}, and an ℓ-dimensional skeleton Sℓ, we say that a map u belongs to W 1,p(Sℓ; Rν) whenever (i) each σℓ ∈ Sℓ, the map uσℓ belongs to W 1,p(σℓ; Rν), (ii) each σℓ on σℓ 2 ∈ Sℓ such that σℓ 1 ∩ σℓ 2 ∈ Sℓ−1, we have uσℓ = uσℓ 1, σℓ 1 ∩ σℓ 2 in 1 2 the sense of traces. DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 19 We then denote W 1,p(Sℓ;Rν ) = Xσℓ∈S ℓkukp kukp W 1,p(σℓ;Rν ). Given ℓ ∈ {1, . . . , m}, η > 0 and a ∈ Rm, we may consider the boundary of the η(a); Rν) has a cube Qℓ well-defined meaning in the sense of Definition 5.1. η(a) as an (ℓ − 1)-dimensional skeleton, and then W 1,p(∂Qℓ Given i ∈ N∗ and η > 0, the homogenization of degree 0 of a map u : ∂Qi η(a) → Rν is the map v : Qi (5.6) η(a) → Rν defined for x ∈ Qi x − a v(x) = u(cid:16)a + η x − a∞(cid:17), η(a) by where y∞ = max(cid:8)y1, . . . ,yi(cid:9) denotes the maximum norm in Ri. The basic property satisfied by this construction is the following: Proposition 5.4. If 1 ≤ p < i, then, for every u ∈ W 1,p(∂Qi η(a) → Rν defined in (5.6) belongs to W 1,p(Qi Qi η(a); Rν ), the map v : η(a); Rν ) and kDvkLp(Qi η (a)) ≤ Cη pkDukLp(∂Qi η(a)). 1 Iterating the zero-degree homogenization described above, one extends Sobolev functions defined on lower-dimensional subskeletons of Rm to an m-dimensional subskeleton. We apply this strategy to prove the following proposition that will be used in the proof of Theorems 1 and 2: Proposition 5.5. Let ℓ ∈ {0, . . . , m − 1}, η > 0, E m be a subskeleton of Rm of inradius η and Sm−1 be a subfamily of E m−1. If p < ℓ + 1, then for every continuous function u : Eℓ ∪ Sm−1 → Rν such that (i) uEℓ ∈ W 1,p(Eℓ; Rν), (ii) uSi ∈ W 1,p(Si; Rν), for every i ∈ {ℓ + 1, . . . , m − 1}, there exists v ∈ W 1,p(Em; Rν) such that v(Em) ⊂ u(Eℓ ∪ Sm−1), v = u on Sm−1 in the sense of traces, and kDvkLp(Em) ≤ C(cid:18)η m−ℓ p kDukLp(Eℓ) + η m−i p kDukLp(Si)(cid:19). m−1Xi=ℓ+1 Proof. Let vℓ : Eℓ → Rν be defined by vℓ = u in Eℓ. We define by induction on i ∈ {ℓ + 1, . . . , m − 1} a map vi : Ei → Rν as follows: (a) for every σi ∈ E i \ Si, we apply the zero-degree homogenization on the face σi (Proposition 5.4) to define vi on σi, so that vi = vi−1 on ∂σi in the sense of traces and σiDvip ≤ C1η∂σiDvi−1p. (b) for every σi ∈ Si, we take vi = u in Si, whence σiDvip = σiDup. With this definition, we have vi ∈ W 1,p(Ei; Rν) since for any given σi such that σi 1 ∩ σi From (a) and (b), 2 ∈ E i−1 we have viσi 2 ∈ E i 2 in the sense of traces. = viσi 1 ∩ σi on σi 1, σi 1 2 EiDvip ≤ C2ηEi−1Dvi−1p +SiDup. 20 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN Iterating these estimates we get Em−1Dvm−1p ≤ C3(cid:18)ηm−1−ℓEℓDvℓp + ηm−1−iSiDup(cid:19). m−1Xi=ℓ+1 From the construction of vi we also have Iterating these inclusions we deduce that vi(Ei) ⊂ vi−1(Ei−1) ∪ u(Si). vm−1(Em−1) ⊂ vℓ(Eℓ) ∪ m−1[i=ℓ+1 u(Si) ⊂ u(Eℓ ∪ Sm−1). The map vm−1 : Em−1 → Rν extends by zero-degree homogenization on each cube σm ∈ E m to vm : Em → Rν, with vm(Em) = vm−1(Em−1) and EmDvmp ≤ C4 ηEm−1Dvm−1p. The function vm thus satisfies the required properties. (cid:3) 6. TRIMMING PROPERTY The next proposition reformulates the trimming property (Definition 1.1), re- placing smooth maps by continuous Sobolev maps. Proposition 6.1. Let p ∈ N∗. The manifold N n satisfies the trimming property of di- mension p if and only if there exists a constant C′ > 0 such that, for each map u ∈ W 1,p(Qp; N n) with u ∈ W 1,p(∂Qp; N n), kDvkLp(Qp) ≤ C′kDukLp(Qp), for some v ∈ W 1,p(Qp; N n)∩ C0(Qp; N n) such that u = v on ∂Qp in the sense of traces. Proof. We begin with the direct implication. For this purpose, let u ∈ W 1,p(Qp; N n) be such that u ∈ W 1,p(∂Qp; N n). We regularize u in two different ways: near the boundary of Qp, this is done by zero-degree homogenization of u∂Qp; far from the boundary, we use the trimming property. We paste the two different approxima- tions with a mollification and cut-off argument in such a way that the approximat- ing function takes its values in a neighborhood of N n. This allows us to project it back on N n. More precisely, given 1 for x ∈ Qp by 2 ≤ λ < 1, we set w : Qp → N n to be the function defined w(x) =(u(x/λ) if x ∈ Qp λ, if x ∈ Qp \ Qp λ. u(x/x∞) Then w ∈ W 1,p(Qp; N n), w is continuous in Qp \ Qp equality and λ by the Morrey–Sobolev in- QpDwp ≤ C1(cid:18)QpDup + (1 − λ)∂QpDup(cid:19). DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 21 Take λ < r1 < r2 < 1. If (ϕε)ε>0 is a family of smooth mollifiers with supp ϕε ⊂ ε , then the function ϕε∗w is smooth on Qp Bp Given θ ∈ C∞ r1), then for ε > 0 small the function v : Qp → N n such that r1, for every 0 < ε ≤ min{r1 − λ, 1 − r2}. r2\Qp c (Qp r2 \ Qp v(x) =(w(x) Π(cid:0)(1 − θ(x))w(x) + θ(x)ϕε ∗ w(x)(cid:1) r2 \ Qp if x ∈ Qp \ (Qp r2 \ Qp if x ∈ Qp r1, r1), is well-defined and belongs to W 1,p(Qp; N n). Remember that Π is the nearest point projection onto the submanifold N n, which is smooth on a neighborhood O ⊂ Rν of N n and satisfies DΠ ∈ L∞(O), see Section 2.3. Here, we also use the continuity r2\Qp of w on Qp\Qp r1) is contained in a compact subset of O. Moreover, by Lemma 2.2, we have the estimate λ, which implies that for ε small, the set(cid:0)(1−θ)w+θϕε∗w(cid:1)(Qp w − ϕε ∗ wp(cid:19). r, there exists a map v ∈ C∞(Qp QpDvp ≤ C2(cid:18)QpDwp + kDθkL∞(Qp)Qp r for some r1 < r < r2, then v ∈ C∞(∂Qp r; N n). Applying r; N n) Setting θ = 1 on ∂Qp r2 \Qp r1 the trimming property to the map v on Qp r and is such that that coincides with v on ∂Qp Qp rDvp ≤ C3Qp rDvp. r, we deduce from the estimates above that Extending v as v on Qp \ Qp QpDvp ≤ C4(cid:18)QpDup + (1 − λ)∂QpDup + kDθkL∞(Qp)Qp w − ϕε ∗ wp(cid:19). To conclude the proof we may assume that ´QpDup > 0. Choosing λ close to 1 and then ε > 0 small, the second and third terms in the right-hand side can be controlled by ´QpDup and the direct implication follows. To prove the converse implication, we take f ∈ C∞(∂Qp; N n) having an exten- sion u ∈ W 1,p(Qp; N n). By assumption, there exists a map v ∈ W 1,p(Qp; N n) ∩ C0(Qp; N n) such that v∂Qp = f and r2 \Qp r1 kDvkLp(Qp) ≤ C′kDukLp(Qp). Once again, the idea of the proof is to smoothen v in two different ways. Far from the boundary, this is done by mollification and projection. Near the boundary, we work with a smooth extension of f. More specifically, given 0 < λ < 1, we fix a smooth extension f ∈ C∞(Qp \ Qp λ; N n) of f. Given 0 < λ < r < r < 1, we take θ ∈ C∞ r) such that θ = 1 in Qp \ Qp r. We note that, for r close to 1 and for ε > 0 small, the function v : Qp → N n such that f (x) c (Qp \ Qp Π(cid:0)(1 − θ(x))ϕε ∗ v(x) + θ(x) f (x)(cid:1) Π(ϕε ∗ v(x)) if x ∈ Qp \ Qp r, if x ∈ Qp r \ Qp r, if x ∈ Qp r, v(x) = 22 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN is well-defined and satisfies the estimate Qp(cid:12)(cid:12)Dv(cid:12)(cid:12)p ≤ C5(cid:18)QpDvp +Qp\Qp L∞(Qp)Qp L∞(Qp)Qp\Qp Since v − f = 0 on ∂Qp, it follows from the Poincaré inequality that rD fp + kDθkp + kDθkp r \Qp rϕε ∗ v − vp rv − fp(cid:19). Qp\Qp r(cid:12)(cid:12)v − f(cid:12)(cid:12)p . ≤ C6(1 − r)p Qp\Qp r(cid:12)(cid:12)D(v − f )(cid:12)(cid:12)p rD fp + (1 − r)−p Qp Taking r < r < 1 and θ such that (1 − r)kDθkL∞(Qp) ≤ C, we get ≤ C7(cid:18)QpDvp +Qp\Qp rϕε ∗ v − vp(cid:19). Qp(cid:12)(cid:12)Dv(cid:12)(cid:12)p We now assume that ´QpDup > 0. The first integral in the right-hand side is by assumption estimated by ´QpDup. Taking r close to 1 and then ε > 0 small the second and third integrals are also bounded by ´QpDup, and the conclusion follows. r \Qp (cid:3) We prove the necessity part in Theorem 2. Proposition 6.2. If p ∈ {2, . . . , m} and if the set (W 1,p ∩ L∞)(Qm; N n) is dense in W 1,p(Qm; N n), then N n satisfies the trimming property of dimension p. Proof. We first consider the case p = m. Let u ∈ W 1,m(Qm; N n) be such that u ∈ W 1,m(∂Qm; N n). By the characterization given by Proposition 6.1, it suffices to prove that there exists a map v ∈ W 1,m(Qm; N n) ∩ C0(Qm; N n) such that u = v on ∂Qm and such that kDvkLm(Qm) ≤ CkDukLm(Qm). The idea of the proof is to rely on the density of bounded maps to replace u by a smooth approximation of u in the interior of the cube. Such an approximation can be made uniform on the boundary of cubes ∂Qm r with r close to 1. We can thus create a transition layer using a zero-degree homogenization of u and an averaging procedure, obtaining a function that is sufficiently close to N n. More precisely, given 0 < λ < 1, we introduce the same map w as in the proof of Proposition 6.1. By assumption, the map w is the limit of a sequence of bounded maps in W 1,m(Qm; N n). It follows from Proposition 3.2 that there exists a se- quence (uk)k∈N in C∞(Qm; N n) converging strongly to w in W 1,m(Qm; Rν). Then, r )j∈N converging to for almost every λ < r < 1, there exists a subsequence (ukj∂Qm r ; Rν). By the Morrey–Sobolev embedding, this convergence w∂Qm is also uniform on the set ∂Qm r ) is a compact subset of N n, there exist a compact set K in the domain O of the nearest point projection Π and an integer J ∈ N such that for every j ≥ J, every z ∈ ∂Qm tukj (z) + (1 − t)w(z) ∈ K. r and every t ∈ [0, 1], r . Since w(∂Qm in W 1,m(∂Qm r We also introduce a cut-off function θ ∈ C∞ θ = 1 in Qm c (Qm) such that 0 ≤ θ ≤ 1 in Qm and r , with (1−r)kDθkL∞(Qm) ≤ C. For every j ≥ J, the map vj : Qm → N n DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 23 defined by vj(x) =(Π(cid:0)θ(x)ukj (rx/x∞) + (1 − θ(x))w(rx/x∞)(cid:1) ukj (x) if x ∈ Qm \ Qm r , if x ∈ Qm r , is such that vj ∈ W 1,m(Qm; N n) ∩ C0(Qm; N n), vj = u on ∂Qm, and QmDvjm ≤ C1(cid:18)Qm Dukjm + kDθkL∞(Qm)(1 − r)∂Qm r r ukj − wm (Dukjm + Dwm)(cid:19). Without loss of generality, we can assume that ´QmDum > 0. We take j ≥ J large enough so that the second term in the right-hand side is bounded from above by ´QmDum. By convergence of the sequence (ukj )j∈N we may also assume that + (1 − r)∂Qm r Qm r Dukjm ≤ 2Qm r Dwm and ∂Qm Dukjm ≤ 2∂Qm Dwm. r r In view of the definition of w in terms of u we deduce from the estimates above that QmDvjm ≤ C2(cid:18)QmDum + (1 − λ)∂QmDum(cid:19). To conclude the case p = m, it suffices to choose λ sufficiently close to 1 so that the second term in the right-hand side is bounded from above by ´QmDum. We now consider the case where p < m. Under the assumption of the propo- sition, we claim that (W 1,p ∩ L∞)(Qp; N n) is also dense in W 1,p(Qp; N n). We are thus led to the first situation where p equals the dimension of the domain, and we conclude that the manifold N n satisfies the trimming property of dimension p. It thus suffices to prove the claim. For this purpose, take u ∈ W 1,p(Qp; N n) and define the function v : Qm → N n for x = (x′, x′′) ∈ Qp × Qm−p by v(x) = u(x′). By assumption, there exists a sequence of maps (vk)k∈N in (W 1,p ∩ L∞)(Qm; N n) converging to v in W 1,p(Qm; N n). Hence, there exist a subsequence (vkj )j∈N and a ∈ Qm−p such that (vkj (·, a))j∈N converges to u in W 1,p(Qp; N n). This proves the claim, and concludes the proof of the proposition. (cid:3) Definition 6.1. A Riemannian manifold N n has uniform Lipschitz geometry (or N n is uniformly Lipschitz) whenever there exist κ, κ′ > 0 and C > 0 such that, for every ξ ∈ N n, kDΨkL∞(BN n (ξ;κ)) + kDΨ−1kL∞(BRn (Ψ(ξ);κ′)) ≤ C, for some local chart Ψ : BN n(ξ; κ) → Rn with BRn (Ψ(ξ); κ′) ⊂ Ψ(BN n(ξ; κ)). Here, for ξ ∈ N n and κ ≥ 0, we have denoted by BN n(ξ; κ) the geodesic ball in N n of center ξ and radius κ. A natural candidate for Ψ is the inverse of the exponential map when the manifold N n has a positive global injectivity radius and the exponential and its inverse are uniformly Lipschitz maps on balls of a fixed radius. If the injectivity radius of N n is uniformly bounded from below and the Riemann curvature of N n is uniformly bounded, then N n has uniform Lipschitz geometry. By relying on harmonic coordinates instead of the normal coordinates 24 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN given by the exponential maps, it can be proved that it is sufficient to bound the Ricci curvature instead of the Riemann curvature [1]. Proposition 6.3. If N n is uniformly Lipschitz, then N n satisfies the trimming property of any dimension p ∈ N∗. The proof of Proposition 6.3 is based on the following lemma that reduces the problem to a trimming property for maps with small energy on the boundary kDukLp(∂Qp). By the Morrey–Sobolev embedding, the range of u∂Qp is then con- tained on a small geodesic ball, and one can perform the extension in a suitable local chart for manifolds having uniform Lipschitz geometry. Lemma 6.4. Let p ∈ N∗ and α > 0. Assume that for every map u ∈ W 1,p(Qp; N n) satisfying u∂Qp ∈ W 1,p(∂Qp; N n) and kDukLp(Qp) + kDukLp(∂Qp) ≤ α, there exists v ∈ (W 1,p ∩ L∞)(Qp; N n) such that v = u on ∂Qp and kDvkLp(Qp) ≤ C′′(cid:0)kDukLp(Qp) + kDukLp(∂Qp)(cid:1) for some constant C′′ > 0 independent u. Then, N n satisfies the trimming property of dimension p. Proof of Lemma 6.4. Let u ∈ W 1,p(Qp; N n) be a map such that u∂Qp ∈ W 1,p(∂Qp; N n). The idea of the proof is to subdivide the domain Qp into smaller cubes, and to apply the opening technique on each cube. The resulting map also has small (rescaled) Sobolev energy on the boundaries of the small cubes, and so we can locally apply the small-energy trimming property to obtain a bounded extension of u∂Qp. We then use an approximation argument by convolution to get a contin- uous extension of u∂Qp. More precisely, for 1 2 < λ < 1, we introduce the map w as in the proof of Proposition 6.1. Then, w∂Qp = u∂Qp, w is bounded on Qp \ Qp λ and kDwkLp(Qp) ≤ C1(cid:0)kDukLp(Qp) + (1 − λ) 1 pkDukLp(∂Qp)(cid:1). Without loss of generality, we can assume that kDukLp(Qp) > 0. We take λ > 0 such that This implies (6.1) (1 − λ) 1 pkDukLp(∂Qp) ≤ kDukLp(Qp). kDwkLp(Qp) ≤ 2C1kDukLp(Qp). We fix 0 < ρ < 1 2 . For every 0 < µ < 1 sufficiently small, we consider a cubication Kp µ of inradius µ such that Qp λ+2ρµ ⊂ K p We open the map w around Kp−1 the smooth map given by Proposition 5.2 above, we consider µ µ + Qp 2ρµ ⊂ Qp. µ ⊂ K p . More precisely, denoting by Φop : Rm → Rm wop = w ◦ Φop. DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 25 In particular, wop ∈ W 1,p(Qp; N n), wop = w outside K p−1 σp ∈ Kp µ, we have µ + Qp 2ρµ and, for every kDwopkLp(∂σp+Qp 2ρµ) ≤ C2kDwkLp(σp+Qp This implies that wop∂Qp = u∂Qp and, for every σp ∈ Kp µ, (6.2) 2ρµ) ≤ C3kDwkLp(σp+Qp Raising both sides to the power p and summing over all σp ∈ Kp, we also get kDwopkLp(σp+Qp 2ρµ). 2ρµ). kDwopkLp(Qp) ≤ C4kDwkLp(Qp). We also need the fact that the opening construction preserves the ranges of the maps. More precisely, for every σp−1 ∈ Kp−1 µ wop(σp−1 + Qp , we have 2ρµ) ⊂ w(σp−1 + Qp 2ρµ). µ + Qp 2ρµ, this proves We apply this remark to every σp−1 ⊂ ∂K p wop(∂K p µ + Qp µ to get 2ρµ) ⊂ w(∂K p Together with the fact that w is bounded on Qp \ Qp that wop is bounded on Qp \ K p µ. Since wop is (p − 1)-dimensional on ∂σp + Qp C5 p kDwopkLp(∂σp+Qp µ kDwopkLp(∂σp) ≤ 1 (6.3) µ + Qp 2ρµ). λ ⊃ ∂K p µ, we have ρµ for every σp ∈ Kp 2ρµ) ≤ µ, we have C6 p kDwkLp(σp+Qp µ 1 2ρµ). We take µ > 0 such that, for every σp ∈ Kp (C3 + C6)kDwkLp(σp+Qp 2ρµ) ≤ α; this is possible by equi-integrability of the summable function Dwp. Then, by estimates (6.2) and (6.3) we have kDwopkLp(σp) + µ 1 pkDwopkLp(∂σp) ≤ α. By the small-energy trimming assumption applied to wopσp for every σp ∈ Kp and by a scalling argument, there exists a map wσp ∈ (W 1,p ∩ L∞)(σp; N n) which agrees with wop on ∂σp and is such that µ (6.4) kDwσpkLp(σp) ≤ C′′(cid:0)kDwopkLp(σp) + µ 1 pkDwopkLp(∂σp)(cid:1). ew(x) = wσp (x) when x ∈ σp and σp ∈ Kp µ u∂Qp. By additivity of the integral and by estimates (6.3) and (6.4), we also have µ. Then, ew ∈ (W 1,p∩ L∞)(Qp; N n) and ew∂Qp = Lp(σp) + kDwopkp Lp(Qp\K p µ) We then define the map ew by and we extend ew by wop outside K p kDwσpkp 2p−1(C′′)p(cid:0)kDwopkp kDewkp Lp(Qp) = Xσp∈Kp ≤ Xσp∈Kp ≤ C7kDwkp Lp(Qp). µ µ Lp(σp) + µkDwopkp Lp(∂σp)(cid:1) + kDwopkp Lp(Qp\K p µ) 26 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN Applying estimate (6.1), we deduce that 1 (6.5) pkDukLp(Qp). 2ρµ) since it agrees with the map w c (Qp) such that 0 ≤ θ ≤ 1 in Qm 2ρµ. Given a family of mollifiers (ϕε)ε>0, kDewkLp(Qp) ≤ 2C1(C7) µ + Qp The map ew is continuous on Qp \ (K p there. We introduce a cut-off function θ ∈ C∞ and θ = 1 on a neighborhood of K p as a consequence of the VMO-property of ew in the critical-integrability case (see exists ε > 0 such that, for every 0 < ε ≤ ε, the set (ϕε ∗ ew)(supp θ) is contained in the neighborhood O where the nearest point projection Π is defined. Since ew is the proof of Proposition 3.2) the Poincaré-Wirtinger inequality implies that there continuous on Qp \ (K p µ + Qp 2ρµ), we can define µ + Qp v = Π(cid:0)θ(ϕε ∗ ew) + (1 − θ)ew(cid:1) for ε sufficiently small. This map v is an extension of u∂Qp in the space W 1,p(Qp; N n)∩ C0(Qp; N n). By the same calculation as in the proof of Proposition 6.1, one has the estimate (6.6) for ε small enough. By estimates (6.5) and (6.6), we have constructed a continuous extension of u∂Qp such that kDvkLp(Qp) ≤ C8kDewkLp(Qp) kDvkLp(Qp) ≤ 2C1(C7) Proposition 6.1 now yields the conclusion. 1 p C8kDukLp(Qp). (cid:3) We now apply Lemma 6.4 to prove that manifolds with uniform Lipschitz ge- ometry satisfy the trimming property. Another application of such a lemma in con- nection with the problem of weak sequential density of bounded Sobolev maps is investigated in [6]. Proof of Proposition 6.3. Let u ∈ W 1,p(Qp; N n) be such that u∂Qp ∈ W 1,p(∂Qp; N n). Take 0 < κ′′ < min{κ, κ′/C}, where throughout the proof we refer to the notation of Definition 6.1. By the Morrey–Sobolev embedding, there exists α > 0 such that if kDukLp(∂Qp) ≤ α, then there exists x ∈ ∂Qp such that, for almost every y ∈ ∂Qp, we have distN n (u(y), u(x)) ≤ κ′′. Given a local chart Ψ on BN n (u(x); κ) as in Definition 6.1, the function Ψ◦u belongs to W 1,p(∂Qp; Rn). By the classical extension property of Sobolev functions, there exists w ∈ W 1,p(Qp; Rn) such that w = Ψ ◦ u on ∂Qp in the sense of traces, and the following estimate holds QpDwp ≤ C1 ∂QpD(Ψ ◦ u)p ≤ C1Cp ∂QpDup. Observe that Ψ ◦ u(∂Qp) ⊂ BRn (Ψ(u(x)); κ′). Indeed, by the mean value in- equality and the choice of κ′′, for almost every y ∈ ∂Qp we have distRn (Ψ(u(y)), Ψ(u(x))) ≤ C distN n (u(y), u(x)) ≤ Cκ′′ < κ′. Thus, truncating w with a retraction on the ball BRn(Ψ(u(x)); κ′) if necessary, we may further assume that the image of the extension w satisfies w(Qp) ⊂ BRn (Ψ(u(x)); κ′), since this does not modify the values of Ψ ◦ u∂Qp. Defining the map v = Ψ−1 ◦ DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 27 w, by composition of Sobolev maps with smooth functions it follows that v ∈ W 1,p(Qp; N n) and QpDvp ≤ CpQpDwp ≤ C2∂QpDup. In view of Lemma 6.4, the proof is complete. (cid:3) 7. PROOFS OF THEOREMS 1 AND 2 Let 1 ≤ p ≤ m and u ∈ W 1,p(Qm; N n). We begin by extending u in a neighbor- hood of Qm and then by taking a cubication that contains Qm. More precisely, by using reflexions across the boundary of Qm, we can extend u as a map in η be a cubica- W 1,p(Qm tion of Qm 1+2γ; N n) for some γ > 0. We also fix 0 < ρ < 1 1+γ of inradius 0 < η ≤ γ such that 2ρη ≤ γ. 2 . Let Km 1+2γ, the function t 7→ u(tx + (1 − t)y) is an abso- For almost every x, y ∈ Qm lutely continuous path in N n between u(x) and u(y). Hence, the geodesic distance distN n (u(x), u(y)) between u(x) and u(y) can be estimated as follows: distN n (u(x), u(y)) ≤ 1 0 Du(tx + (1 − t)y)[x − y] dt. As in the proof of the Poincaré-Wirtinger inequality, this implies that Qm 1+2γ Qm 1+2γ distN n (u(x), u(y)) dx dy ≤ C1 Qm 1+2γ Du. It follows that, for almost every y ∈ Qm 1+2γ, Qm 1+2γ distN n (u(x), u(y)) dx < ∞. 1+2γ. We now distinguish the cubes in the cubication Km This implies that, for every a ∈ N n, the function x 7→ distN n (u(x), a) is summable on Qm η in terms of good cubes and bad cubes. In a good cube, most of the values of the function u lie in a geodesic ball centered at some fixed point a ∈ N n, and u does not oscillate too much; the latter is quantified in terms of the rescaled Lp norm of Du. We fix a point a ∈ N n. For every R > 0 and λ > 0, we define the subskeleton η of Km Gm η in the sense that η as the set of good cubes σm ∈ Km σm+Qm distN n (u(x), a) dx ≤ R and 2ρη η m−p 1 p kDukLp(σm+Qm η defined as the complement of 2ρη ) ≤ λ. We also introduce the subskeleton of bad cubes E m η in Km Gm η . Thus, by definition of E m R < σm+Qm η we have η , for every σm ∈ E m 1 p kDukLp(σm+Qm distN n (u(x), a) dx or λ < m−p η 2ρη 2ρη ). In the proof, we do not explicitly indicate the dependence of Gm parameters R and λ. η and E m η on the On the bad cubes E m η , we wish to replace the function u by some nicer, bounded, func- tion. To reach this goal, we would like to use the values of u on the lower-dimensional 28 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN η, where ℓ = ⌊p⌋, assuming that u ∈ W 1,p(Eℓ η. We can thus propagate the values of u inside Em η; N n). Indeed, for p noninte- skeleton E ℓ ger we have p > ℓ and then the Morrey–Sobolev embedding implies that u is continuous on Eℓ η by zero-degree homogenization. When p is integer, we have p = ℓ and we cannot rely on the Morrey–Sobolev embedding. In this case, we apply the trimming property of dimension p to modify the function u on Eℓ η, keeping its values on the lower-dimensional skeleton Eℓ−1 We first quantify the total volume of bad cubes. More precisely, the Lebesgue 2ρη can be made as small as we want by a suitable choice measure of the set Em of parameters R and η. This is a consequence of the following estimate: η + Qm η . Proof of the claim. By finite subadditivity of the Lebesgue measure, we have distN n(u(x), a) dx + Claim 1. The Lebesgue measure of the set Em η + Qm 1+2γ η + Qm (cid:12)(cid:12)Em 2ρη(cid:12)(cid:12) ≤ C(cid:18) 1 R Qm 2ρη(cid:12)(cid:12) ≤ Xσm∈E m (cid:12)(cid:12)Em From the definition of E m η(cid:18) η ≤ Xσm∈E m #E m σm + Qm η + Qm 1 η(cid:12)(cid:12)σm + Qm 2ρηR σm+Qm 2ρη 2ρη satisfies ηp λp Qm 2ρη(cid:12)(cid:12) ≤ C2ηm (#E m 1+2γ η ). Dup(cid:19). η , we estimate the number #E m η of bad cubes as follows: distN n (u(x), a) dx Dup(cid:19) + 1 ηm−pλp σm+Qm Dup(cid:19). λp Qm ηp 1+2γ 2ρη C3 ηm(cid:18) 1 R Qm 1+2γ ≤ distN n(u(x), a) dx + Combining both estimates, we get the conclusion. Since the cubication Km ⋄ η is prescribed independently of u, the Lp norm of Du on the η could be very large. We thus begin by opening u in a neighborhood of Eℓ η, η depending on, at most, ℓ components around each face skeleton Eℓ which provides a new function uop of Eℓ η. Throughout the proof, we denote by ℓ = ⌊p⌋ η if the integer part of p. We begin by opening the map u in a neighborhood of Eℓ p < m and in a neighborhood of Em−1 if p = m. More precisely, if Φop : Rm → Rm is the smooth map given by Proposition 5.2 with the parameter ρ, we consider the opened map η uop η = u ◦ Φop. When p < m, we have that uop of Eℓ η + Qm η ∈ W 1,p(Qm 2ρη. Moreover, there exists C > 0 such that, for every σℓ ∈ E ℓ η, we have 1+2γ; N n) and uop η = u in the complement kDuop η kLp(σℓ+Qm 2ρη ) ≤ CkDukLp(σℓ+Qm 2ρη ), (7.1) and also (7.2) When p = m, the integer ℓ must be replaced by m − 1 in the above estimates. 1+2γ ) ≤ CkDukLp(Eℓ η − DukLp(Qm kDuop 2ρη ). η+Qm DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 29 We now consider a convolution of the opened map uop η . The convolution parameter is not constant: it is small on the good cubes and quickly becomes zero as we enter the bad cubes. Such a transition is made in a region having width of order η. 1+2γ) More precisely, given 0 < ρ < ρ, we consider a smooth function ψη ∈ C∞(Qm such that (a) 0 ≤ ψη < (ρ − ρ)η, η , for some parameter 0 < t < ρ − ρ , (b) ψη = tη on Gm ρη , η + Qm (c) supp ψη ⊂ Gm 1+2γ ) < 1, (d) kDψηkL∞(Qm The parameter t is fixed throughout the proof and is independent of η, R and λ. Condition (b) gives an upper bound on t, while Condition (d) imposes t to be typically smaller than ρ and this can be achieved independently of the geometry of the cubication Gη. Given a mollifier ϕ ∈ C∞ c (Bm η (x) = (ϕψη ∗ uop usm 1+γ let 1 ), for every x ∈ Qm η )(x) = Bm uop 1 η (cid:0)x − ψη(x)y(cid:1)ϕ(y) dy. Since 0 < ψη ≤ ρη < γ, the smoothened map usm Claim 2. The map usm η satisfies the estimates η : Qm 1+γ → Rν is well-defined. (7.3) (7.4) kusm η − ukLp(Qm 1+γ ) ≤ sup η − DukLp(Qm 1 kτψηv(u) − ukLp(Qm 1+γ ) ≤ sup 1+γ ) + Ckuop 1 kτψηv(Du) − DukLp(Qm v∈Bm v∈Bm 1+γ ) kDusm η − ukLp(Qm 1+2γ ), Proof of the claim. By Proposition 5.3 with ω = Qm 1 kτψη v(uop 1+γ ) ≤ sup We also observe that, for every v ∈ Bm η − uop η kLp(Qm 1 , we have kusm v∈Bm η +Qm 2ρη ). + CkDukLp(Em 1+γ, we have η ) − uop η kLp(Qm 1+γ ). kτψη v(uop η ) − uop η kLp(Qm ≤ kτψηv(uop 1+γ ) η ) − τψη v(u)kLp(Qm 1+γ ) + kτψηv(u) − ukLp(Qm 1+γ ) + kuop ≤ kτψηv(u) − ukLp(Qm 1+γ ) + Ckuop η − ukLp(Qm 1+γ ) 1+2γ ), η − ukLp(Qm and this proves (7.3). We now consider the second estimate. Since kDψηkL∞(Qm 1+2γ ) < 1, it also fol- lows from Proposition 5.3 that (7.5) kDusm η − Duop η kLp(Qm 1+γ ) ≤ sup v∈Bm where A = x∈Qm 1+γ ∩supp Dψη S η ) − Duop 1 kτψηv(Duop ψη(x)(x). From Property (b), we have Bm η kLp(Qm 1+γ ) + C1kDuop η kLp(A), supp Dψη ∩ Qm 1+γ ⊂ Qm 1+γ \ Gm η ⊂ Em η 30 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN and, since ψη ≤ ρη, we deduce that A ⊂ Em get η + Qm ρη. By Proposition 5.2, we then kDuop (7.6) η kLp(A) ≤ C2kDukLp(Em As in the proof of the first estimate, for every v ∈ Bm (7.7) kτψηv(Duop η ) − Duop η kLp(Qm 2ρη ). η +Qm 1 we also have 1+γ ) ≤ kτψηv(Du) − DukLp(Qm + CkDuop 1+γ ) η − DukLp(Qm 1+2γ ). η (Gm Although the smoothened map usm Combining estimates (7.5)–(7.7) and (7.2), we complete the proof of (7.4). ⋄ η need not lie on the manifold N n, we now quantify how far the set usm η ) is with respect to some large geodesic ball BN n (a; R) with R > R. Since there are many points of u(Gm η ) on the geodesic ball BN n (a; R), we can apply the Poincaré-Wirtinger inequality to establish such an estimate. By choosing the parameter λ sufficiently small, we will later on be able to project back usm to N n, at least on the good η cubes Gm η . Claim 3. There exists R > R such that, for every η > 0 and λ > 0, the directed Hausdorff distance to the geodesic ball BN n (a; R) satisfies C′ DistBN n (a;R)(usm η (Gm η )) ≤ m−p p η max σm∈Gm η kDukLp(σm+Qm 2ρη ), for some constant C′ > 0 depending on m and p. Here, the directed Hausdorff distance from a set S ⊂ Rν to the geodesic ball BN n (a; R) is defined as DistBN n (a;R) (S) = supn distRν(cid:0)x, BN n (a; R)(cid:1) : x ∈ So, where distRν denotes the Euclidean distance in Rν. Proof of the claim. Given σm ∈ Gm η and R > 0, we consider the sets WR =nz ∈ σm + Qm ZR =nz ∈ σm + Qm =nz ∈ σm + Qm =nz ∈ σm + Qm 2ρη : distN n (u(z), a) < Ro, 2ρη : distN n (u(z), a) ≥ Ro, η (z), a) < Ro, η (z), a) ≥ Ro. η (z) ∈ BN n(a; R) for every z ∈ W op 2ρη : distN n (uop 2ρη : distN n (uop W op R Z op R R and their counterparts for the map uop η obtained by the opening construction, Observe that by definition uop W op BN n (a; R) in terms of an average integral as follows R > 0, then for every x ∈ σm we may estimate the distance from usm . Assuming that η (x) to distRν(cid:0)usm η (x), BN n (a; R)(cid:1) ≤ kϕkL∞(Bm η (x), BN n (a; R)(cid:1) ≤ W op 1 ) W op For every x ∈ σm, we then have distRν(cid:0)usm R R usm η (x) − uop η (z) dz. Bm ψη (x)(x)uop η (y) − uop η (z) dy dz, DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 31 where ϕ is the mollifier used in the definition of usm Bm deduce that ψη(x)(x) are contained in σm + Qm R and 2ρη, by the Poincaré–Wirtinger inequality we η . Since both sets W op (7.8) distRν (usm Since ψη = tη on Gm η (x), BN n (a; R)) ≤ C1η2m R Bm η , for every x ∈ σm we have W op ψη(x)(x) 1 p kDuop m−p η kLp(σm+Qm 2ρη ). η (7.9) ψη(x)(x) ≥ C2ηm. We now estimate from below the quantity W op Bm η the average integral satisfies nition of Gm R . Since σm ∈ Gm η , then by defi- σm+Qm 2ρη distN n(u(x), a) dx ≤ R, hence by the Chebyshev inequality we have (7.10) ZR σm + Qm 2ρη R ≤ R. We now proceed with the choice of R. Taking any R > R such that (7.11) we have σm + Qm 2ρη R ≤ (cid:12)(cid:12)(σm + Qm ZR ≤ (cid:12)(cid:12)(σm + Qm 2 2ρη) \ (∂σm + Qm 2ρη) \ (∂σm + Qm R, 2 2ρη)(cid:12)(cid:12) 2ρη)(cid:12)(cid:12) . Since σm is a cube of inradius η, by a scaling argument with respect to η this choice of R is independent of η. Since the maps uop 2ρη)\(∂σm + Qm η and u coincide in (σm +Qm 2ρη), we have By subadditivity of the Lebesgue measure and by the choice of R we get Z op R ⊂ ZR ∪ (∂σm + Qm 2ρη). 2ρη) \ (∂σm + Qm 2 +(cid:12)(cid:12)∂σm + Qm 2ρη(cid:12)(cid:12), hence the measure of the complement set W op 2ρη) \ (∂σm + Qm W op (7.12) By estimates (7.8), (7.9) and (7.12) for every x ∈ σm we deduce that with the above choice of R we have = 2m−1(η − 2ρη)m = C3ηm. 2 R satisfies 2ρη)(cid:12)(cid:12) 2ρη)(cid:12)(cid:12) Z op R ≤ (cid:12)(cid:12)(σm + Qm R ≥ (cid:12)(cid:12)(σm + Qm distRν(cid:0)usm η (x), BN n (a, R)(cid:1) ≤ C4 p kDuop m−p η kLp(σm+Qm 2ρη ). η By subadditivity of the Lebesgue measure and by the properties of the opening construction, when p < m we have η kp kDuop Lp(σm+Qm η kp 2ρη ) ≤ kDuop Lp((σm+Qm 2ρη )\(Eℓ η +Qm 2ρη )) + Xσℓ∈E ℓ η σℓ⊂σm η kp kDuop Lp(σℓ+Qm 2ρη ) ≤ C5kDukp Lp(σm+Qm 2ρη ). (7.13) (7.14) 32 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN When p = m, then ℓ must be replaced by m − 1 in the above inequality. Together with (7.13), this implies the estimate we claimed. ⋄ is from the large geodesic ball BN n (a; R) on the part of the bad set Eℓ η that lies in the transition between good and bad cubes. The estimate uses the fact that the opened map uop depends on ℓ components nearby Eℓ We now quantify how far the smoothened map usm η η and that the convolution parameter is chosen very small in this region. Claim 4. There exists R > R such that, for every η > 0 and λ > 0, the directed Hausdorff distance to the geodesic ball BN n (a; R) satisfies DistBN n (a;R)(cid:0)usm η (Eℓ η ∩ supp ψη)(cid:1) ≤ η for some constant C′′ > 0 depending on m and p. C′′ m−p p max σm∈Gm η kDukLp(σm+Qm 2ρη ), Proof of the claim. Using Property (c) satisfied by the function ψη (see page 29), one gets ρη(cid:1). By Claim 3 above it thus suffices to prove that, for every τ ℓ−1 ∈ E ℓ−1 have η ∩ supp ψη ⊂ (Eℓ Eℓ η ∩ Gℓ−1 η ∩ Gℓ ) + Qm η η ∩ Gℓ−1 η , we (7.15) DistBN n (a;R)(cid:0)usm η (τ ℓ−1 + Qm η kDukLp(σm+Qm For this purpose, we observe that there exists R > R such that the map uop η can be constructed with the following additional property: for every τ ℓ−1 ∈ E ℓ−1 , η ∩ Gℓ−1 (7.16) η kLp(τ ℓ−1+Qm ρη ). C2 p kDuop η (τ ℓ−1 + Qm max σm∈Gm 2ρη ). m−p m−p η η) ∪(cid:0)(Eℓ−1 ρη)(cid:1) ≤ C1 η p DistBN n (a;R)(cid:0)uop Indeed, for every σm ∈ Gm (7.17) η ρη)(cid:1) ≤ 2ρη R ≤ Qm ρη 2 R, σm + Qm η and for every R > R such that we have, by (7.10), ZR ≤ Qm ρη 2 . Again by a scaling argument with respect to η, this choice of R is independent of η. For each vertex v of the cube σm and for at least half of the points x of the ρη(v), we thus have u(x) ∈ BN n(a; R). Since the opening construction is cube Qm based on a Fubini-type argument (see the explanation preceding Proposition 5.1), η, the common value of uop we may thus assume that for each vertex v of ∂σm ∩ E0 η in Qm . Since p > ℓ− 1, by the η ∩Gℓ−1 Morrey–Sobolev inequality we have, for every y, z ∈ τ ℓ−1, p kDuop ρη(v) belongs to BN n (a; R). Consider an (ℓ− 1)-dimensional face τ ℓ−1 ∈ E ℓ−1 η (z)(cid:1) ≤ C3η1− ℓ−1 distRν(cid:0)uop η (y), uop On the other hand, since the map uop map in τ ℓ−1 + Qm ρη, we have uop η (τ ℓ−1 + Qm ρη) = uop η η kLp(τ ℓ−1). η is, by construction, an (ℓ − 1)-dimensional η (τ ℓ−1) and also kDuop η kLp(τ ℓ−1) ≤ C4 m−(ℓ−1) p η kDuop η kLp(τ ℓ−1+Qm ρη ). DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 33 This implies that, for every y, z ∈ τ ℓ−1 + Qm ρη, C3C4 distRν(cid:0)uop η (y), uop η (z)(cid:1) ≤ η m−p p kDuop η kLp(τ ℓ−1+Qm ρη ). η, we thus obtain estimate (7.16). Taking as z any vertex of τ ℓ−1 in G0 We now complete the proof of (7.15). Recall that the map usm η η by convolution with parameter ψη. Hence, for every τ ℓ−1 ∈ Gℓ−1 uop for every x ∈ τ ℓ−1 + Qm distRν(cid:0)usm is obtained from and η ∩ E ℓ−1 ρη such that ψη(x) 6= 0, by the triangle inequality we have η η (x), BN n (a, R)(cid:1) ≤ C5 Qm η (z) − uop η (y) dy dz ψη (x)(x) Qm η (y), BN n (a, R)(cid:1) dy. ρη. ψη(x)(x) ⊂ τ ℓ−1 + Qm ψη (x)(x) distRν(cid:0)uop Since x ∈ τ ℓ−1 + Qm ρη and ψη(x) < (ρ − ρ)η, we have Qm ψη(x)(x), Together with (7.16), this implies that, for every y ∈ Qm distRν(cid:0)uop η (y), BN n (a, R)(cid:1) ≤ C6 p kDuop By the Poincaré–Wirtinger inequality, we deduce that η (x), BN n (a, R)) ≤ C7(cid:18) p kDuop distRν (usm (7.18) ψη(x) m−p m−p 1 η η kLp(τ ℓ−1+Qm ρη ). ψη (x)(x) uop + Qm η kLp(Qm ψη (x)(x)) + η 1 p kDuop m−p ρη)(cid:19). η kLp(τ ℓ−1+Qm r (x) ⊂ τ ℓ−1 + Qm ρη we Next, from the opening construction, for every cube Qm have (7.19) 1 rm−p Qm r (x)Duop η p ≤ C8 ηm−p τ ℓ−1+Qm ρη Duop η p. Indeed, this follows directly from (5.3) when p < m. When p = m, one can proceed along the lines of the proof of estimate (5.3) with ℓ replaced by m. Combining inequalities (7.18) and (7.19) with r = ψη(x), we get distRν (usm η (x), BN n (a, R)) ≤ η C9 p kDuop m−p η kLp(τ ℓ−1+Qm ρη ). In view of the estimates satisfied by the opening construction and the fact that τ ℓ−1 ∈ Gℓ−1 η , for every x ∈ τ ℓ−1 + Qm η (x), BN n (a, R)(cid:1) ≤ distRν(cid:0)usm C10 ρη such that ψη(x) 6= 0 we have η kDukLp(σm+Qm η (x) = uop max σm∈Gm m−p η p If ψη(x) = 0, then usm from which (7.15) follows. inequality remains true by (7.16). This completes the proof of the claim. η (x), and the above ⋄ Up to now, the parameters λ and η were arbitrary. In the following, they will be subject η ∪ Gm η , η back to the manifold N n. We then extend the projected map to to some restrictions. Our aim is to make usm η so that we can project usm Em sufficiently close to N n on the set Eℓ η using the zero-degree homogenization. 2ρη ), 34 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN For a given R > 0, we take R > R satisfying the conclusions of Claims 3 and 4. For any such R, let ιR > 0 be such that BN n(a; R) + Bν ιR ⊂ O. Remember that the geodesic ball BN n(a; R) is a relatively compact subset of N n and that O is an open neighborhood of N n in Rν on which can be defined a smooth retraction Π : O → N n such that DΠ ∈ L∞(O); see Section 2.3. We also take λ > 0 depending on R > 0, whence also on R > 0, such that (7.20) λ ≤ ιR , max{C′, C′′} where C′, C′′ > 0 are the constants given by Claims 3 and 4, respectively. On the one hand, for every good cube σm ∈ Gm 1 p kDukLp(σm+Qm η we have 2ρη ) ≤ By the estimate from Claim 3, this implies that max{C′, C′′} ιR m−p η . On the other hand, Claim 4 implies that usm η (Gm η ) ⊂ BN n(a; R) + Bν ιR ⊂ O. usm η (Eℓ η ∩ supp ψη) ⊂ BN n (a; R) + Bν ιR ⊂ O. On K m η \ supp ψη, we have usm η = uop usm η (Eℓ η . In particular, η \ supp ψη) ⊂ N n. This proves that (7.21) usm η (Eℓ η) ⊂ N n ∪ BN n (a; R) + Bν ιR ⊂ O. We now define the projected map upr η ∪ Eℓ η : Gm upr η = Π ◦ usm η . η → N n by setting On Gm η , the map upr η is smooth and we have: Claim 5. The map upr η satisfies kDupr η ) η − DukLp(Gm ≤ kDΠkL∞(O)kDusm Proof of the claim. Since upr compact subset of O, Lemma 2.2 and the triangle inequality imply η − DukLp(Gm η = Π ◦ usm η ) +(cid:13)(cid:13)DΠ(usm η , u = Π ◦ u and usm η ) − DΠ(u)Du(cid:13)(cid:13)Lp(Gm η (Gm η ) is contained in a η ). kDupr η − DukLp(Gm η ) ≤ kDΠ(usm η )kL∞(Gm η )kDusm η − DukLp(Gm η ) +(cid:13)(cid:13)DΠ(usm η ) − DΠ(u)Du(cid:13)(cid:13)Lp(Gm η ). ⋄ We now make sure that the zero-degree homogenization can be successfully performed η using the values of upr η. For this purpose, we need some local control of the η in terms of the rescaled Lp norm of Du. This is enough to get the strong on Em Lp norm of Dupr convergence in W 1,p since the bad set Em η on Eℓ η is small. DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 35 η belongs to W 1,p(Eℓ η; N n) and, for every τ ℓ ∈ E ℓ η , we have η Eℓ Claim 6. The map upr upr η ∂τ ℓ ∈ W 1,p(∂τ ℓ; N n) and also (7.22) kDupr η kLp(τ ℓ) ≤ C p kDukLp(τ ℓ+Qm m−ℓ 2ρη ), η for some constant C > 0 depending on m and p. Proof of the claim. By Lemma 2.2, the inclusion (7.21) and the fact that upr it is enough to prove the claim for usm η τ ℓ ∈ E ℓ, the restriction usm above. For this purpose, we can assume that η , η = Π◦usm η . We first prove that, for every η τ ℓ belongs to W 1,p(τ ℓ; Rν) and satisfies the estimate instead of upr τ ℓ = (−η, η)ℓ × {0′′}, where 0′′ ∈ Rm−ℓ. Accordingly, we write every vector y ∈ Rm as y = (y′, y′′) ∈ ρη, Rℓ × Rm−ℓ. Since, for y ∈ τ ℓ + Qm uop η (y) = uop η (y′, 0′′), and since Dψη is uniformly bounded with respect to η, for every x′ ∈ τ ℓ we have Dusm η (x′, 0′′)p ≤ C1Bm ≤ C2Bℓ 1 1 Duop η (x′ − ψη(x′, 0′′)y′, 0′′)p dy Duop η (x′ − ψη(x′, 0′′)y′, 0′′)p dy′. Hence, by Fubini's theorem, τ ℓDusm η (x′, 0′′)p dx′ ≤ C2 Bℓ 1 τ ℓDuop η (x′ − ψη(x′, 0′′)y′, 0′′)p dx′ dy′. Using the change of variables z′ = x′ − ψη(x′, 0′′)y′ with respect to the variable x′, we get τ ℓDusm η (x′, 0′′)p dx′ ≤ 1 − kDψηkL∞(Km C3 η ) Bℓ 1 dy′ (−(1+ρ)η,(1+ρ)η)ℓ Duop η (z′, 0′′)p dz′ ≤ C4 (−(1+ρ)η,(1+ρ)η)ℓ Duop η (z′, 0′′)p dz′. We observe that Duop η (z′, 0′′)p dz′ ≤ (−(1+ρ)η,(1+ρ)η)ℓ Combining both inequalities, we get C5 ηm−ℓ τ ℓ+Qm ρη Duop η p. τ ℓDusm C6 ηm−ℓ τ ℓ+Qm η p ≤ η τ ℓ belongs to W 1,p(τ ℓ; Rν) and (7.22) holds by When p < m, we deduce that usm η . the estimate in Assertion (iv) of Proposition 5.2 satisfied by the opened map uop When p = m, we rely instead on the estimate (7.14), which holds with ℓ replaced by m − 1. Duop η p. ρη 36 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN η We can prove in a similar way that usm . Moreover, the map uop η η τ ℓ−1 belongs to W 1,p(τ ℓ−1; Rν ) for ev- ery τ ℓ−1 ∈ E ℓ−1 is (ℓ − 1)-dimensional on Eℓ−1 η + Qm ρη and thus continuous by the Morrey–Sobolev embedding. This implies that usm η . Hence, the first part of the claim also is continuous on a neighborhood of Eℓ−1 follows. ⋄ η so as to preserve η on the (higher-dimensional) common faces with good cubes. Such a the values of upr construction naturally yields a Sobolev map on the entire domain Em The zero-degree homogenization of upr η should be performed inside Em η By construction, the map usm η i ∈ {ℓ, . . . , m − 1}. In particular, upr now estimate the Lp norm of D(upr η ∪ Gm η . is smooth on a neighborhood of Ei η for every η ∩ Gi η; Rν), and we η ∩ Ei η Ei η Ei η belongs to W 1,p(Gi ) when p < m (and thus ℓ ≤ m − 1): η∩Gi η η∩Gi Claim 7. Assume that p < m. For every i ∈ {ℓ, . . . , m − 1}, we have kDupr η kLp(Ei η∩Gi η) ≤ η C p kDukLp(Ei m−i η+Qm 2ρη ), for some constant C > 0 depending on m and p. Proof of the claim. Once again, we only need to prove the estimate with usm η of upr instead η, we have ψη(x) = tη and thus η . Fix i ∈ {ℓ, . . . , m − 1}. For every x ∈ Ei η ∩ Gi usm η (x) = Bm 1 uop η (x − tηy)ϕ(y) dy. Hence, by Jensen's inequality and a change of variable, C1 Dusm η (x)p ≤ C1 Bm 1 Duop η (x − tηy)p dy = (tη)m Bm η p. tη (x)Duop Since the parameter t < ρ is fixed, we can incorporate it in the constant. Integrating both members with respect to the i-dimensional Hausdorff measure over Ei η, η ∩ Gi by Fubini's theorem we get By construction of uop Ei η+Qm 2ρη Duop η∩Gi η Dusm Ei η p ≤ η , we also have η p ≤ (Ei ≤ C3 Ei η +Qm η+Qm 2ρη 2ρη )\(Eℓ η+Qm Dup, C2 ηm−i Ei η +Qm 2ρη Duop η p. 2ρη )Dup + Xτ ℓ∈E ℓ η τ ℓ+Qm 2ρη Duop η p η to Em and the conclusion follows. η is bounded on Eℓ η. We now proceed to construct a bounded extension ube ⋄ η . It follows from Claim 6 and the Morrey–Sobolev embedding that, when ℓ < p, the projected map upr In this case, we apply the zero-degree homogenization to extend upr η is bounded on the lower-dimensional skeleton Eℓ−1 . We then apply the trimming property of dimension p, restated in terms of Sobolev maps by Proposition 6.1, to obtain a new, bounded and continuous, function on Eℓ η. The quantitative character of the trimming property ensures that this new function satisfies the same energy bounds. η. When ℓ = p, we can only infer that upr η using its values on Eℓ η of the map upr η inside Em η Em η ∩Gm η DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 37 Claim 8. If ℓ < p or if ℓ = p and N n satisfies the trimming property of dimension p, then there exists a map ube η and η ; N n) such that ube η on Em η = upr η ∈ (W 1,p ∩ L∞)(Em η ) ≤ C(cid:18)η p kDupr m−ℓ η kLp(Eℓ η) + kDube η kLp(Em for some constant C > 0 depending on m, p and N n. m−1Xi=ℓ+1 η m−i p kDupr η kLp(Ei η ∩ Gm η)(cid:19), η∩Gi η, the map upr η ∩ Gℓ η is continuous on Eℓ η. On every Proof of the claim. We first define the extension ube η in τ ℓ. When ℓ < p, face τ ℓ ∈ E ℓ the map upr η. When ℓ = p, we assume that the trimming property holds, whence by Proposition 6.1 we to get may replace upr a continuous map ube η \ Gℓ η without changing its trace on Eℓ−1 η on each face τ ℓ ∈ E ℓ η is smooth, and we set ube η on the subskeleton Eℓ η, and we also set ube η on τ ℓ ∈ E ℓ η = upr η = upr η \Gℓ η η ∈ W 1,p(τ ℓ; N n) such that η kLp(τ ℓ) ≤ C1kDupr η is continuous and belongs to W 1,p(Eℓ kDube η thus defined in Eℓ η kLp(τ ℓ). (7.23) The map ube η; N n). η on Em The definition of the extension ube proceed assuming that p < m; thus ℓ ≤ m − 1. Let Sm−1 as a continuous Sobolev map by upr we extend ube η on (Eℓ ) ⊂ (Eℓ upr η = ube η ∪ Sm−1 η : Eℓ ube η ; N n), agrees with upr W 1,p(Em η is complete when p = m. We now , and η . This is possible since η). We now apply Proposition 5.5 to the map η , belongs to η → N n. The resulting map, that we still denote by ube η ∩ Gℓ η on Em η ∩ Sm−1 = E m−1 and satisfies η to Sm−1 ∩ Gm−1 η η η η η η η ∩ Gm η (Eℓ η = Sm−1 η ∪ Sm−1 η ) ⊂ ube η ; N n). Finally, η ). Dube η p + ηm−iSi η η p(cid:19). Dupr m−1Xi=ℓ+1 In particular, we have ube η (Em ube η ∈ L∞(Em η p ≤ C2(cid:18)ηm−ℓEℓ Dube η Em η ⊂ Ei η Since Si (7.23). η ∩ Gi η, the required estimate follows from the above inequality and ⋄ We deduce from Claims 6–8 that (7.24) kDube η kLp(Em η ) ≤ C1kDukLp(Em η +Qm 2ρη ). We now complete the proof of the theorem. For this purpose, let (Ri)i∈N be a sequence of positive numbers diverging to infinity. Accordingly, Claims 3 and 4 yield a sequence (Ri)i∈N from which we define a sequence of positive numbers (λRi )i∈N satisfying (7.20). Finally, we take a sequence of positive numbers (ηi)i∈N converging to zero such that lim i→∞ ηi λRi = 0. By Claim 1, we have (7.25) We proceed to prove that (7.26) i→∞ Em lim ηi + Qm 2ρηi = 0. i→∞ kDupr lim ηi − DukLp(Gm ηi ) = 0. 38 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN Indeed, from estimate (7.3) in Claim 2, we have kusm ηi − ukLp(Qm 1+γ ) ≤ sup v∈Bm 1 kτψηi v(u) − ukLp(Qm 2ρηi, by the Poincaré inequality for functions vanish- ηi − ukLp(Qm 1+γ ) + Ckuop 1+2γ ). Since u = uop ing on a set of positive measure and by property (7.25) we have ηi outside Em ηi + Qm kuop ηi − ukLp(Qm 1+2γ ) ≤ C2kDuop ηi − DukLp(Qm 1+2γ ) ≤ C3kDukLp(Em ηi +Qm 2ρηi ). Hence, kusm ηi − ukLp(Qm 1+γ ) ≤ sup v∈Bm 1 kτψηi v(u) − ukLp(Qm 1+γ ) + C4kDukLp(Em ηi +Qm 2ρηi ), which proves that (7.27) i→∞kusm lim ηi − ukLp(Qm 1+γ ) = 0. By the dominated convergence theorem, we thus get i→∞kDΠ(usm lim ηi ) − DΠ(u)DukLp(Gm ηi ) = 0. By estimate (7.4) in Claim 2, we also have i→∞kDusm lim Both limits and Claim 5 imply (7.26). ηi − DukLp(Gm ηi ) = 0. ηi = ube ηi on Gm ηi , the function obtained by juxtaposing upr ηi and ube ηi Since upr defined by ηi ∩ Em ηi(x) =(upr ujx ηi (x) ube ηi (x) ηi, if x ∈ Gm ηi , if x ∈ Em 1+γ; N n). Moreover, by (7.24)–(7.26) we have belongs to (W 1,p ∩ L∞)(Qm i→∞kDujx (7.28) lim ηi − DukLp(Qm 1+γ ) = 0. Finally, we establish that (7.29) i→∞kujx lim ηi − ukLp(Qm 1+γ ) = 0. To this aim, we introduce the auxiliary sequence (vi)i∈N in the space Lp(Qm defined by 1+γ; Rν) vi(x) =(usm ηi (x) ube ηi (x) ηi, if x ∈ Gm ηi . if x ∈ Em Observe that ujx ηi = Π ◦ vi. ηi )i∈N converges to u in measure on As a consequence of (7.27), the sequence (usm 1+γ. In view of (7.25), this is also true for (vi)i∈N. Since Π is continuous, (ujx Qm ηi)i∈N 1+γ. Together with (7.28), this implies converges in measure to Π ◦ u = u on Qm (7.29) as in the end of the proof of Proposition 2.1. This completes the proofs of Theorem 1 and the sufficiency part of Theorem 2; the necessity part follows from Proposition 6.2 above. (cid:3) DENSITY OF BOUNDED MAPS IN SOBOLEV SPACES 39 Proofs of Corollaries 1.1 and 1.2. (=⇒) Let 1 ≤ p ≤ m, and assume that the set C∞(Qm; N n) is dense in W 1,p(Qm; N n). When p < m, this implies that π⌊p⌋(N n) is trivial as in the case when N n is compact [4, 26], with the same proof. When p ∈ {2, . . . , m}, the set (W 1,p ∩ L∞)(Qm; N n) is then dense in W 1,p(Qm; N n), and it follows from Proposition 6.2 that N n satisfies the trimming property of dimen- sion p. (⇐=) Conversely, if 1 ≤ p ≤ m is not an integer and π⌊p⌋(N n) is trivial, then Proposition 3.3 implies that C∞(Qm; N n) is dense in (W 1,p ∩ L∞)(Qm; N n). It also follows from Theorem 1 that the set (W 1,p∩L∞)(Qm; N n) is dense in W 1,p(Qm; N n). Hence, C∞(Qm; N n) is dense in W 1,p(Qm; N n). This completes the proof of Corol- lary 1.1. Finally, the sufficiency part of Corollary 1.2 follows from Theorem 2 and Propo- sition 3.3 when p ∈ {1, . . . , m − 1}, and from Theorem 2 and Proposition 3.2 when p = m. This completes the proof of Corollary 1.2. (cid:3) ACKNOWLEDGMENTS The authors would like to thank the referee for his or her detailed reading and comments that have improved the presentation of the paper and H. Brezis for calling their attention to Marcus and Mizel's paper [21]. Part of this work was carried out while the first author (PB) was visiting IRMP with support from UCL. The second (ACP) and third (JVS) authors were supported by the Fonds de la Recherche scientifique-FNRS; ACP: Crédit de Recherche (CDR) J.0026.15 and JVS: Mandat d'Impulsion scientifique (MIS) F.452317. REFERENCES [1] M. T. Anderson, Convergence and rigidity of manifolds under Ricci curvature bounds, Invent. Math. 102 (1990), 429–445. [2] F. Bethuel, The approximation problem for Sobolev maps between two manifolds, Acta Math. 167 (1991), 153–206. [3] F. Bethuel and D. Chiron, Some questions related to the lifting problem in Sobolev spaces, Perspectives in nonlinear partial differential equations, Contemp. Math., vol. 446, Amer. Math. Soc., Providence, RI, 2007, pp. 125–152. [4] F. Bethuel and X. M. Zheng, Density of smooth functions between two manifolds in Sobolev spaces, J. Funct. Anal. 80 (1988), 60–75. [5] P. Bousquet, A. C. Ponce, and J. Van Schaftingen, Strong density for higher order Sobolev spaces into compact manifolds, J. Eur. Math. Soc. (JEMS) 17 (2015), 763–817. [6] , Weak approximation by bounded Sobolev maps with values into complete manifolds, available at arxiv:1701.07627. Submitted for publication. [7] H. Brezis and Y. Li, Topology and Sobolev spaces, J. Funct. Anal. 183 (2001), 321–369. [8] H. Brezis and L. Nirenberg, Degree theory and BMO. I. Compact manifolds without boundaries, Selecta Math. (N.S.) 1 (1995), 197–263. [9] A. Convent and J. Van Schaftingen, Intrinsic colocal weak derivatives and Sobolev spaces between mani- folds, Ann. Sc. Norm. Super. Pisa Cl. Sci. (5) 16 (2016), 97–128. [10] Y.-J. Dai, M. Shoji, and H. Urakawa, Harmonic maps into Lie groups and homogeneous spaces, Differ- ential Geom. Appl. 7 (1997), 143–160. [11] L. C. Evans and R. F. Gariepy, Measure theory and fine properties of functions, 2nd ed., Textbooks in Mathematics, CRC Press, Boca Raton, FL, 2015. [12] H. Federer and W. H. Fleming, Normal and integral currents, Ann. of Math. (2) 72 (1960), 458–520. [13] M. Focardi and E. Spadaro, An intrinsic approach to manifold constrained variational problems, Ann. Mat. Pura Appl. (4) 192 (2013), 145–163. 40 P. BOUSQUET, A. C. PONCE, AND J. VAN SCHAFTINGEN [14] P. Hajłasz, Sobolev mappings between manifolds and metric spaces, Sobolev spaces in mathematics. I, Int. Math. Ser. (N.Y.), vol. 8, Springer, New York, 2009, pp. 185–222. [15] P. Hajłasz and A. Schikorra, Lipschitz homotopy and density of Lipschitz mappings in Sobolev spaces, Ann. Acad. Sci. Fenn. Math. 39 (2014), 593–604. [16] F. Hang and F.-H. Lin, Topology of Sobolev mappings, II, Acta Math. 191 (2003), 55–107. [17] [18] R. Hardt, D. Kinderlehrer, and F.-H. Lin, Stable defects of minimizers of constrained variational princi- , Topology of Sobolev mappings, III, Comm. Pure Appl. Math. 56 (2003), 1383–1415. ples, Ann. Inst. H. Poincaré Anal. Non Linéaire 5 (1988), 297–322. [19] J. Heinonen, P. Koskela, N. Shanmugalingam, and J. T. Tyson, Sobolev classes of Banach space-valued functions and quasiconformal mappings, J. Anal. Math. 85 (2001), 87–139. [20] A. A. Kosinski, Differential manifolds, Pure and Applied Mathematics, vol. 138, Academic Press, Inc., Boston, MA, 1993. [21] M. Marcus and V. J. Mizel, Absolute continuity on tracks and mappings of Sobolev spaces, Arch. Rational Mech. Anal. 45 (1972), 294–320. [22] O. Müller, A note on closed isometric embeddings, J. Math. Anal. Appl. 349 (2009), 297–298. [23] J. Nash, C 1 isometric embeddings, Ann. of Math. (2) 60 (1954), 383–396. [24] [25] M. R. Pakzad and T. Rivière, Weak density of smooth maps for the Dirichlet energy between manifolds, , The embedding problem for Riemannian manifolds, Ann. of Math. (2) 63 (1956), 20–63. Geom. Funct. Anal. 13 (2003), 223–257. [26] R. Schoen and K. Uhlenbeck, Boundary regularity and the Dirichlet problem for harmonic maps, J. Dif- ferential Geom. 18 (1983), 253–268. [27] K. Uhlenbeck, Harmonic maps into Lie groups: classical solutions of the chiral model, J. Differential Geom. 30 (1989), 1–50. [28] H. Whitney, Geometric integration theory, Princeton University Press, Princeton, NJ, 1957. PIERRE BOUSQUET UNIVERSITÉ DE TOULOUSE INSTITUT DE MATHÉMATIQUES DE TOULOUSE, UMR CNRS 5219 UNIVERSITÉ PAUL SABATIER TOULOUSE 3 118 ROUTE DE NARBONNE 31062 TOULOUSE CEDEX 9 FRANCE AUGUSTO C. PONCE UNIVERSITÉ CATHOLIQUE DE LOUVAIN INSTITUT DE RECHERCHE EN MATHÉMATIQUE ET PHYSIQUE CHEMIN DU CYCLOTRON 2, BTE L7.01.02 1348 LOUVAIN-LA-NEUVE BELGIUM JEAN VAN SCHAFTINGEN UNIVERSITÉ CATHOLIQUE DE LOUVAIN INSTITUT DE RECHERCHE EN MATHÉMATIQUE ET PHYSIQUE CHEMIN DU CYCLOTRON 2, BTE L7.01.02 1348 LOUVAIN-LA-NEUVE BELGIUM
1308.3215
3
1308
2013-09-14T14:16:40
Parseval frames with n+1 vectors in R^n
[ "math.FA" ]
We construct a Parseval frame with $n+1$ vectors in $\R^n$ that contains a given vector. We also provide a characterization of unit-norm frames that can be scaled to a Parseval frame.
math.FA
math
Parseval frames with n + 1 vectors in Rn Laura De Carli and Zhongyuan Hu Abstract We prove a uniqueness theorem for triangular Parseval frame with n + 1 vectors in Rn. We also provide a characterization of unit-norm frames that can be scaled to a Parseval frame. Mathematical subject classification: 42C15, 46C99 1 Introduction Let B = {v1, ..., vN} be a set of vectors in Rn. We say that B is a frame if it contains a basis of Rn, or equivalently, if there exist constants A, B > 0 for which Av2 ≤ PN j=1 < v, vj >2≤ Bv2 for every v ∈ Rn. Here and throughout the paper, < , > and are the usual scalar product and norm in Rn. In general A < B, but we say that a frame is tight if A = B, and is Parseval if A = B = 1. ties v = PN j=1 < v, vj > vj and v2 = PN Parseval frames are nontrivial generalizations of orthonormal bases. Vec- tors in a Parseval frame are not necessarily orthogonal or linearly indepen- dent, and do not necessarily have the same length, but the Parseval identi- j=1 < v, vj >2 still hold. In the applications, frames are more useful than bases because they are resilient against the corruptions of additive noise and quantization, while providing numerically stable reconstructions ([1], [7], [9]). Appropriate frame decom- position may reveal hidden signal characteristics, and have been employed as detection devices. Specific types of finite tight frames have been studied to solve problems in information theory. The references are too many to cite, but see [3], the recent book [1] and the references cited there. In recent years, several inquiries about tight frames have been raised. In particular: how to characterize Parseval frames with N elements in Rn (or Parseval N frames), and whether it is possible to scale a given frame so that the resulting frame is Parseval. Following [2] and [11], we say that a frame B = {v1, ..., vN} is scalable if there exists positive constants ℓ1, ..., ℓN such that {ℓ1v1,..., ℓN vN} is a Parseval frame. Two Parseval N-frames are equivalent if one can be trans- formed into the other with a rotation of coordinates and the reflection of one or more vectors. A frame is nontrivial if no two vectors are parallel. In 1 the rest of the paper, when we say "unique" we will always mean "unique up to an equivalence", and we will often assume without saying that frames are nontrivial. It is well known that Parseval n-frames are orthonormal (see also Corol- lary 3.3). Consequently, for given unit vector w, there is a unique Parseval n-frame that contains w. If w 6= 1, no Parseval n-frame contains w. When N > n and w ≤ 1, there are infinitely many non-equivalent Parseval N -frames that contain w 1. By the main theorem in [5], it is possible to construct a Parseval frame {v1, ..., vN} with vectors of prescribed lengths 0 < ℓ1, ..., ℓN ≤ 1 that satisfy ℓ1 N = n. We can let ℓN = w and, after a rotation of coordinates, assume that vN = w, thus proving that the Parseval frames that contain w are as many as the sets of constants ℓ1, ... ℓN−1. But when N = n + 1, there is a class of Parseval frames that can be uniquely constructed from a given vector: precisely, all triangular frames, that is, frames {v1, ..., vN} such that the matrix (v1, ... vn) whose columns are v1, ..., vn is right triangular. We recall that a matrix {ai,j}1≤i, j≤n is right-triangular if ai,j = 0 if i > j. 1 + ... + ℓ2 The following theorem will be proved in Section 3. Theorem 1.1. Let B = {v1, ..., vn, w} be a triangular Parseval frame, with w < 1. Then B is unique, in the sense that if B′ = {v′1, ..., v′n, w} is another triangular Parseval frame, then v′j = ±vj. Every frame is equivalent, through a rotation of coordinates ρ, to a triangular frame, and so Theorem 1.1 implies that every Parseval (n + 1)- frame that contains a given vector w is equivalent to one which is uniquely determined by ρ(w). However, that does not imply that the frame itself is uniquely determined by w because the rotation ρ depends also on the other vectors of the frame. We also study the problem of determining whether a given frame B = {v1, ..., vn, vn+1} ⊂ Rn is scalable or not. Assume vj = 1, and let θi,j ∈ [0, π) be the angle between vi and vj. If B contains an orthonormal basis, then the problem has no solution, so we assume that this not the case. We prove the following Theorem 1.2. B is scalable if and only there exist constants ℓ1, ..., ℓn+1 such that for every i 6= j j cos θ2 i,j. (1.1) (1 − ℓ2 i )(1 − ℓ2 j ) = ℓ2 i ℓ2 1We are indebted to P. Casazza for this remark. 2 The identity (1.1) has several interesting consequences (see corollary 3.2). First of all, it shows that ℓ2 i,j = 0 for every j, and so vi is orthogonal to all other vectors. This interesting fact is also true for other Parseval frames, and is a consequence of the following j ≤ 1; if ℓi = 1 for some i, we also have cos θ2 Theorem 1.3. Let B = {v1, ..., vN} be a Parseval frame. Let vj = ℓj. Then NXj=i NXj=i j cos2 θij = 1 ℓ2 j sin2 θij = n − 1 ℓ2 (1.2) The identities (1.2) are probably known, but we did not find a reference in the literature. It is worthwhile to remark that from (1.2) follows that NXj=i ℓ2 j cos2 θij − NXj=i ℓ2 j sin2 θij = NXj=i ℓ2 j cos(2θij) = 2 − n. When n = 2, this identity is proved in Proposition 2.1. Another consequence of Theorem 1.1 is the following Corollary 1.4. If B = {v1, ..., vn, vn+1} is a scalable frame, then, for every 1 ≤ i ≤ n + 1, and every j 6= k 6= i and k′ 6= j′ 6= i, cos θk,j cos θk,j + cos θk,i cos θj,i = cos θk′,j ′ cos θk′,j ′ + cos θk′,i cos θj ′,i . (1.3) We prove Theorems 1.1, 1.2 and 1.3 and their corollaries in Section 3. In Section 2 we prove some preliminary results and lemmas. Acknowledgement. We wish to thank Prof. P. Casazza and Dr. J. Cahill for stimulating conversations. 2 Preliminaries We refer to [4] or to [10] for the definitions and basic properties of finite frames. We recall that B = {v1, ..., vN} is a Parseval frames in Rn if and only if rows of the matrix (v1, ..., vN ) are orthonormal. Consequently,PN i=1 vi2 = n. If the vectors in B have all the same length, then vi =pn/N . See e.g. [4]. 3 We will often let ~e1 = (1, 0, ..., 0), ... ~en = (0, ..., 0, 1), and we will denote by (v1, ..., vk, ...vN ) the matrix with the column vk removed. To the best of our knowledge, the following proposition is due to P. Casazza (unpublished-2000) but can also be found in [8] and in the recent preprint [6]. Proposition 2.1. B = {v1, ..., vN} ⊂ R2 is a tight frame if and only if for some index i ≤ N , NXj=1 vj2e2iθi,j = 0. (2.1) It is easy to verify that if (2.1) is valid for some index i, then it is valid for all other i's. Proof. Let ℓj = vj. After a rotation, we can let vi = v1 = (ℓ1, 0) and θ1,j = θj, so that vj = (ℓj cos θj, ℓj sin θj). B is a tight frame with frame constant A if and only if the rows of the matrix (v1, ..., vN ) are orthogonal and have length A. That implies NXj=1 ℓ2 j cos2 θj = NXj=1 ℓ2 j sin2 θj = A ℓ2 j cos θj sin θj = 0. (2.2) (2.3) and NXj=1 From (2.2) follows that PN and from (2.3) that PN follows that B is a tight frame. j=1 ℓ2 j (cos2 θj − sin2 θj) = PN j=1 ℓ2 j sin(2θj) = 0, and so we have proved (2.1). j=1 ℓ2 j cos(2θj) = 0, If (2.1) holds, then (2.2) and (2.3) hold as well, and from these identities Corollary 2.2. Let B = {v1, v2, v3} ⊂ R2 be a tight frame. Assume that the vi's have all the same length. Then, θ1,2 = π/3, and θ1,3 = 2π/3. So, every such frame B is equivalent to a dilation of the "Mercedes-Benz frame" n(1, 0), (− 1 √3 2 ), (− 1 2 )o . 2 ,− √3 2 , 4 Proof. Let v1 = (1, 0), and θ1,i = θi for simplicity. By Proposition 2.1, It is easy to 1 + cos(2θ2) + cos(2θ3) = 0, and sin(2θ2) + sin(2θ3) = 0. verify that these equations are satisfied only when θ2 = π 3 or viceversa. 3 and θ3 = 2π The following simple proposition is a special case of Theorem 1.1, and will be a necessary step in the proof. Lemma 2.3. Let w = (α1, α2) be given. Assume w < 1. There exists a unique nontrivial Parseval frame {v1, v2, w} ⊂ R2, with v1 = (a11, 0), v2 = (a1,2, a2,2), and a1,1, a2,2 > 0. Proof. We find a1,1, a1,2 and a2,2 so that the rows of the matrix (cid:18) a11, a12, α1 a22, α2(cid:19) are orthonormal. That is, 0, α2 1 + a2 1,1 + a2 1,2 = 1, α2 2 + a2 2,2 = 1, α1α2 + a1,2a2,2 = 0. (2.4) From the second equation, a2,2 = ±p1 − α2 the third equation we obtain a1,2 = − α1α2√1−α2 2 2; if we can chose a2,2 > 0, from and from the first equation 1,1 = 1 − α2 a2 1 − a2 12 = 1 − α2 1 − Note that a2 1,1 > 0 because w2 = α2 1 + α2 2 = 1 − α2 1 − α2 1α2 α2 1 − α2 2 < 1. We can chose then 1 − α2 . 2 2 2 a1,1 = p1 − α2 2 − α2 p1 − α2 2 1 . Note also that v and v2 cannot be parallel; otherwise, a1,2 a2,2 α1 α2 ⇐⇒ −α2 Remark. The proof shows that v1 and v2 are uniquely determined by w. It shows also that if w = 1, then a1,1 = 0, and consequently v1 = 0. 2, which is not possible. 2 = 1 − α2 = − α1α2 1−α2 = 2 3 Proofs In this section we prove Theorem 1.1 and some of its corollaries. 5 Proof of Theorem 1.1 . Let w = (α1, ..., αn). We construct a nontrivial Parseval frame M = {v1, ..., vn, w} ⊂ Rn with the following properties: the matrix (v1, ..., vn) = {ai,j}1≤i,j≤n is right triangular, and aj,j =  p1 − α2 q1−Pn q1−Pn n k=j α2 k k=j+1 α2 k if j = n if 1 ≤ j < n. (3.1) The proof will show that M is unique, and also that the assumption that w < 1 is necessary in the proof. To construct the vectors vj we argue by induction on n. When n = 2 we have already proved the result in Lemma 2.3. We now assume that the lemma is valid in dimension n − 1, and we show that it is valid also in dimension n. Let w = (α2, ..., αn). By assumptions, there exist vectors v2, ..., vn such that the set fM = {v2, ..., vn, w} ⊂ Rn−1 is a Parseval frame, and the matrix (v2, ..., vn, w) is right triangular and invertible. If we assume that the ele- ments of the diagonal are positive, the vj's are uniquely determined by w. We let vj = (a2,j, ..., an,j), with ak,j = 0 if k < j and aj,j > 0. We show that fM is the projection on Rn−1 of a Parseval frame in Rn = R× Rn−1 that satisfies the assumption of the theorem. To this aim, we prove that there exist scalars x1,..., xn so that the vectors {v1, ..., vn+1} which are defined by v1 = (x1, 0, ..., 0), vj = (xj, vj) if 2 ≤ j ≤ n, vn+1 = w (3.2) form a Parseval frame of Rn. The proof is in various steps: first, we construct a unit vector (y2, ..., yn+1) which is orthogonal to the rows of the matrix (v2, ..., vn, w). Then, we show that there exists −1 < λ < 1 so that λyn+1 = α1. Finally, we chose x1 = √1 − λ2, xj = λyj, and we prove that the vectors v1, ..., vn+1 defined in (3.2) form a Parseval frame that satisfies the assumption of the lemma. First of all, we observe that {v1, ..., vn+1} is a Parseval frame if and only if ~x = (x1, x2, ..., xn, α1) is a unit vector that satisfies the orthogonality conditions: (v2, ...vn, w)~x = = . (3.3)   a22 a23 a33 0 ... ... 0 0 ... a2,n α2 ... a3,n α3 ... ... ... an,n αn ... 6     x2 ... xn α1     0 0 ... 0   By a well known formula of linear algebra, the vector ~y = y2~e2 + ... + ~en+1yn+1 = det   ~e2 ~e3 a2,2 a2,3 a3,3 ... 0 0 0 ... 0 0 ~en a2,n a3,n ... ~en+1 α2 α3 ... ... ... ... ... ... an−1,n αn−1 αn ... an,n   (3.4) is orthogonal to the rows of the matrix in (3.3), and so it is a constant multiple of ~x. That is, ~x = λ~y for some λ ∈ R. Let us prove that ~y = 1. The rows of the matrix (v2, ..., vn, w) are orthonormal, and so after a rotation (v2, ..., vn, w) =   0 1 ... 0 0 0 0 ... 0 0 ... ... 0 0 ... 1 0 0 0 ... 0 1 ... ... ...   . (3.5) The formula in (3.4) applied with the matrix in (3.5) produces the vector ~e1 = (1, 0, ..., 0). Thus, ~y in (3.4) is a rotation of ~e1, and so it is a unit vector as well. We now prove that λ < 1. From ~x = (x2, ..., xn, α1) = λ(y2, ..., yn, yn+1), we obtain λ = α1/yn+1. By (3.4), yn+1 = (−1)n+1det   a2,2 a2,3 a3,3 ... 0 0 0 ... 0 0 a2,n a3,n ... a2,n−1 a3,n−1 ... ... ... ... an−1,n−1 an−1,n an,n ... 0 ...   = (−1)n+1 nYj=2 aj,j. Recalling that by (3.1), aj,j = q1−Pn q1−Pn k=j α2 k k=j+1 α2 k , we can see at once that nYj=2 aj,j = (−1)n+1q1 − α2 yn+1 = (−1)n+1 =(−1)n+1q1 − w2 + α2 1. 2 − ... − α2 n−1 − α2 n (3.6) 7 In view of λyn+1 = α1, we obtain Clearly, λ < 1 because w < 1. We now let α1 1 . λ = (−1)n+1 p1 − w2 + α2 x1 =p1 − λ2 = p1 − w2 p1 − w2 + α2 1 , (3.7) and we define the vj's as in (3.2). The first rows of the matrix (v1, ..., vn+1) is (√1 − λ2, ~x) = (√1 − λ2, λ~y), and so it is unitary and perpendicular to the other rows. Therefore, the {vj} form a tight frame that satisfies the assumption of the theorem. The proof of Theorem 1.1 shows the following interesting fact: By (3.6) and (3.7) det(v1, ..., vn) = nYj=1 ajj = x1 nYj=2 ajj =p1 − w2. This formula does not depend on the fact that (v1, ..., vn) is right triangular, because every n × n matrix can be reduced in this form with a rotation that does not alter its determinant and does not alter the norm of w. This observation proves the following Corollary 3.1. Let {w1, ..., wn+1} be a Parseval frame. Then, det(w1, ..., wj , ..., wn+1) = ±q1 − wj2. Proof of Theorem 1.2. Let ℓj = vj. If {v1, ..., vn+1} is a Parseval frame, then the rows of the matrix B = (v1, ..., vn+1) are orthonormal. While proving Theorem 1.1, we have constructed a vector ~x = (x1, ..., xn+1), with xj = (−1)j+1λ det(v1, ..., vj , ..., vn+1), which is perpendicular to the rows of B. By Corollary 3.1, xj = ±q1 − ℓ2 j . Since B is a Parseval frame, 1 + ... + ℓ2 ℓ2 n+1 = n, and so ~x2 = x2 1 + ... + x2 n+1 = (1 − ℓ2 1) + ... + (1 − ℓ2 n+1) = 1. 8 So, the (n+1)×(n+1) matrix B which is obtained from B with the addition of the row ~y, is unitary, and therefore also the columns of B are orthonormal. For every i, j ≤ n + 1, hvi, vji ±q1 − ℓ2 j = ℓiℓj cos θij ±q1 − ℓ2 iq1 − ℓ2 iq1 − ℓ2 j = 0 (3.8) which implies (1.1). Conversely, suppose that (1.1) holds. By (3.8), the vectors vj = (±q1 − ℓ2 are orthonormal for some choice of the sign ±; therefore, the columns of the matrix B are orthonormal, and so also the rows are orthonormal, and B is a Parseval frame. j , vj) 1 Corollary 3.2. Let B = {v1, ...vn+1} be a nontrivial Parseval frame. Then, n+1 < ℓ2 j for every j. Moreover, for all j with the possible exception of one, 2 < ℓ2 j . 1 Proof. The identity (1.1) implies that, for i 6= j, That implies ℓ2 exception of one, are ≥ 1 j + ℓ2 i ℓ2 i + ℓ2 j − ℓ2 j sin2 θij = 0. 1 − ℓ2 2 . Recalling that Pn+1 i ≥ 1 for every i 6= j, and so all ℓ2 i = n, i=1 ℓ2 (3.9) j 's, with the possible 1 − (n + 1)ℓ2 j + and so ℓ2 j > 1 n+1 . n+1Xi=1 ℓ2 j ℓ2 i sin2 θij = 0, Proof of Theorem 1.3. After a rotation, we can assume vi = v1 = (ℓ1, 0, ..., 0). We let θ1,j = θj for simplicity. With this rotation vj = (ℓj cos θj, ℓj sin θjwj) where wj is a unitary vector in Rn−1. The rows of the matrix (v1, ...vN ) are orthonormal, and so the norm of the first row is ℓ2 j cos2 θj + ℓ2 1 = 1 (3.10) Xj≥1 9 The projections of v2, .... vN over a hyperplane that is orthogonal to v1 form a tight frame on this hyperplane. That is to say that {ℓ2 sin θ2w2, ..., ℓN sin θN wN} is a tight frame in Rn−1, and so it satisfies ℓ2 2 sin θ2 = ℓ2 2w22 2 sin θ2 2 + ... + ℓ2 N sin θ2 2 + ... + ℓ2 N sin θ2 2 NwN2 N = n − 1. (3.11) Corollary 3.3. {v1, v2, ..., vn} is a Parseval frame in Rn if and only if the vi's are orthonormal Proof. By (3.10), all vectors in a Parseval frame have length ≤ 1. By (3.11) nXj=1 ℓ2 j sin2 θi,j = n − 1 which implies that ℓj = 1 and sin θij = 1 for every j 6= i, and so all vectors are orthonormal. Proof of Corollary 1.4. Assume that B is scalable; fix i < n + 1, and chose j 6= k 6= i. By 1.1 (1 − ℓ2 (1 − ℓ2 (1 − ℓ2 i )(1 − ℓ2 i )(1 − ℓ2 k)(1 − ℓ2 j ) = ℓ2 k) = ℓ2 j ) = ℓ2 i ℓ2 i ℓ2 kℓ2 j cos θ2 k cos θ2 j cos θ2 i,j, i,k, k,j. These equations are easily solvable for for ℓ2 1, ℓ2 j and ℓ2 k; we obtain ℓ2 i = cos θk,j cos θk,j + cos θk,i cos θj,i . This expression for ℓi must be independent of the choice of j and k, and so (1.3) is proved References [1] Z. Cvetkovic. Resilience properties of redundant expansions under ad- ditive noise and quantization. IEEE Trans. Inform. Th., 2002 [2] J. Cahill and X. Chen. A note on scalable frames, arXiv:1301.7292v1 (2013) 10 [3] P.G. Casazza, M. Fickus, J. Kovacevic, M.T. Leon, and J.C. Tremain. A physical interpretation of finite frames, Appl. Numer. Harmon. Anal. 2-3 (2006), 51 -- 76 [4] P. Casazza and G. Kutyniok. Finite Frames: Theory and Applications, Birkhauser 2013 [5] P. Casazza and M. T. Leon, Existence and construction of finite tight frames. J. Concr. Appl. Math. 4 (2006), no. 3, 277289. [6] M. S. Copenhaver, Y. H. Kim, C. Logan, K. Mayfield, S. K. Narayan, M. J. Petro, J. Sheperd. Diagram vectors and Tight Frame Scaling in Finite Dimensions arXiv:1303.1159 (2013) [7] I. Daubechies. Ten Lectures on Wavelets. SIAM, Philadelphia, PA, 1992 [8] V. K. Goyal and J. Kovacevic. quantized frames expansion with era- sures Applied and Computational Harmonic Analysis 10, 203233 (2001) [9] V. K Goyal, M. Vetterli, and N. T. Thao. Quantized overcomplete expansions in RN : Analysis, synthesis, and algorithms.IEEE Trans. Inform. Th., 44(1):1631, January 1998 [10] D. Han, K. Kornelson, D. Larson and E. Weber. Frames for Under- graduates, Student Mathematical Library Series, AMS, 2007 [11] G.Kutyniok, K. A. Okoudjou, F. Philipp, and E. K. Tuley. Scalable frames, arXiv:1204.1880v3 (2012) 11
1902.04537
1
1902
2019-02-12T18:32:00
Gabor windows supported on $[-1,1]$ and construction of compactly supported dual windows with optimal frequency localization
[ "math.FA" ]
We consider Gabor frames $\{e^{2\pi i bm \cdot} g(\cdot-ak)\}_{m,k \in \mathbb{Z}}$ with translation parameter $a=L/2$, modulation parameter $b \in (0,2/L)$ and a window function $g \in C^n(\mathbb{R})$ supported on $[x_0,x_0+L]$ and non-zero on $(x_0,x_0+L)$ for $L>0$ and $x_0\in \mathbb{R}$. The set of all dual windows $h \in L^2(\mathbb{R})$ with sufficiently small support is parametrized by $1$-periodic measurable functions $z$. Each dual window $h$ is given explicitly in terms of the function $z$ in such a way that desirable properties (e.g., symmetry, boundedness and smoothness) of $h$ are directly linked to $z$. We derive easily verifiable conditions on the function $z$ that guarantee, in fact, characterize, compactly supported dual windows $h$ with the same smoothness, i.e., $h \in C^n(\mathbb{R})$. The construction of dual windows is valid for all values of the smoothness index $n \in \mathbb{Z}_{\ge 0} \cup \{\infty\}$ and for all values of the modulation parameter $b<2/L$; since $a=L/2$, this allows for arbitrarily small redundancy $(ab)^{-1}>1$. We show that the smoothness of $h$ is optimal, i.e., if $g \notin C^{n+1}(\mathbb{R})$ then, in general, a dual window $h$ in $C^{n+1}(\mathbb{R})$ does not exist.
math.FA
math
Gabor windows supported on [−1, 1] and construction of compactly supported dual windows with optimal frequency localization Jakob Lemvig∗ , Kamilla Haahr Nielsen† February 13, 2019 Abstract: We consider Gabor frames(cid:8)e2πibm·g(· − ak)(cid:9)m,k∈Z with translation parameter a = L/2, modulation parameter b ∈ (0, 2/L) and a window function g ∈ C n(R) supported on [x0, x0 + L] and non-zero on (x0, x0 + L) for L > 0 and x0 ∈ R. The set of all dual windows h ∈ L2(R) with sufficiently small support is parametrized by 1-periodic measurable functions z. Each dual window h is given explicitly in terms of the function z in such a way that desirable properties (e.g., symmetry, boundedness and smoothness) of h are directly linked to z. We derive easily verifiable conditions on the function z that guarantee, in fact, characterize, compactly supported dual windows h with the same smoothness, i.e., h ∈ C n(R). The construction of dual windows is valid for all values of the smoothness index n ∈ Z≥0 ∪ {∞} and for all values of the modulation parameter b < 2/L; since a = L/2, this allows for arbitrarily small redundancy (ab)−1 > 1. We show that the smoothness of h is optimal, i.e., if g /∈ C n+1(R) then, in general, a dual window h in C n+1(R) does not exist. 1 Introduction One of the central tasks in signal processing and time-frequency analysis is to find convenient series expansions of functions in L2(R). A popular choice of such series expansions is by use of Gabor frames, which are function systems of the form {MbmTakg}m,k∈Z =ne2πibm·g(· − ak)om,k∈Z , where a, b > 0 and g ∈ L2(R), and where Tλf = f (· − λ) and Mγf = e2πiγ·f , λ, γ ∈ R, denote the translation and modulation operator on L2(R), respectively. Now, a Gabor frame for L2(R) is a Gabor system {MbmTakg}m,k∈Z for which there exists constants A, B > 0 such that A kf k2 ≤ Xm,k∈Z hf, MbmTakgi2 ≤ B kf k2 for all f ∈ L2(R). (1.1) 2010 Mathematics Subject Classification. Primary 42C15. Secondary: 42A60 Key words and phrases. dual frame, dual window, Gabor frame, optimal smoothness, redundancy, time- frequency localization ∗Technical University of Denmark, Department of Applied Mathematics and Computer Science, Matem- atiktorvet 303B, 2800 Kgs. Lyngby, Denmark, E-mail: [email protected] †Technical University of Denmark, Department of Applied Mathematics and Computer Science, Matem- atiktorvet 303B, 2800 Kgs. Lyngby, Denmark, E-mail: [email protected] 1 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . If the upper bound holds, we say that {MbmTakg}m,k∈Z is a Bessel system with bound B. In case ab < 1 and {MbmTakg}m,k∈Z satisfies (1.1), there exists infinitely many functions h ∈ L2(R) such that {MbmTakh}m,k∈Z is a Bessel system and f = Xm,k∈Z hf, MbmTakgiMbmTakh for all f ∈ L2(R) (1.2) holds with unconditionally L2-convergence. The function g generating the Gabor frame {MbmTakg}m,k∈Z is called the window, while h is called a dual window. For a Bessel sys- tem {MbmTakg}m,k∈Z, the linear operator Sg : L2(R) → L2(R) defined by Sgf = Xm,k∈Z hf, MbmTakgiMbmTakg is called the frame operator, and the canonical dual window is given by h = S−1 consider windows g supported on an interval of length L. g g. We will 1.1 Painless non-orthonormal expansions The most classical method of constructing dual windows is by painless non-orthonormal ex- pansions by Daubechies et al. [11, Theorem 2]. Assume s(x) is nonnegative, bounded and sup- ported on an interval of length L > 0, and that s has constant periodizationPn∈Z s(x+an) = 1 almost everywhere. Then defining g = sp and h = s1−p, where 0 < p < 1 classically is taken to be p = 1/2, generate dual frames {MbmTakg}m,k∈Z and {MbmTakh}m,k∈Z for any 0 < b ≤ 1/L and 0 < a ≤ L. A variant of the painless construction without assuming a partition of unity property (i.e., constant periodization) is a follows. We will again consider g ∈ L∞(R) having compact support in an interval of length L. If a ≤ L, and b ≤ 1/L, the frame operator Sg becomes a multiplication operator: Sgf (x) = g(x + an)2 · f (x). 1 bXn∈Z It follows that the Gabor system {MbmTakg}m,k∈Z is a frame with bound A and B if and only if A ≤ 1 g g In this bPn∈Z g(· + an)2 ≤ B, in which case the canonical dual window h := S−1 is compactly supported on supp g and given by h(x) = bg(x)/Pn∈Z g(x + an)2. case {MbmTakg}m,k∈Z andnMbmTakhom,k∈Z are canonical dual Gabor frames with compact support in an interval of length L. 1.2 Our contribution On the borderline a = L and 0 < b < 1/L of the painless expansions region (a, b) ∈ (0, L] × (0, 1/L], the discontinuous window g = L−1/2χ[0,L] generates a tight Gabor frame. In this case, the Gabor system becomes a union of Fourier series Sk∈Z {MbmTakg}m∈Z with no support overlap between the different Fourier systems {MbmTakg}m∈Z indexed by k ∈ Z. This may lead to unwanted mismatch artifacts at the seam points aZ when truncating the Gabor expansion (1.2). To diminish such artifacts, we will use a 2-overlap (or 2-covering) condition, namely, a = L/2. This means that for almost any time x ∈ R, two Fourier-like systems {MbmTakg}m∈Z, k ∈ Z, represent the signal f at time x. Phrased differently, for each k ∈ Z, the Fourier- like system {MbmTakg}m∈Z has an overlap of length L/2 with (cid:8)MbmTa(k−1)g(cid:9)m∈Z and with (cid:8)MbmTa(k+1)g(cid:9)m∈Z. date/time: 13-Feb-2019/1:32 2 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . Under the 2-overlap condition a = L/2, the painless construction is only applicable for redundancies (ab)−1 ≥ 2 since b ≤ 1/L. We are interested in small redundancies (ab)−1 < 2, that is, large modulations b > 1/L; note that, as is standard in Gabor analysis, in fact, necessary once continuity of the window g is imposed, we always take (ab)−1 > 1. Hence, we are interested in the redundancy range (ab)−1 ∈ (1, 2). However, once outside the region of painless non-orthonormal expansions, computing the canonical dual h = S−1 g g becomes much more cumbersome since it requires inverting the frame operator Sg, and one often resort to numerical approaches [2, 13, 15, 18, 19]. For n ∈ Z≥0 ∪ {∞}, we consider the window class of n times continuously differentiable functions g : R → C, i.e., g ∈ C n(R), that are supported on a closed interval [x0, x0 + L] (x0 ∈ R) of length L and nonzero on the open interval (x0, x0 + L). For any function g in this window class, the Gabor system(cid:8)MbmTLk/2g(cid:9)m∈Z is a frame for L2(R) if and only if b ∈ (0, 2/L) [7, Corollary 2.8]. The objective of our work is to characterize and construct compactly supported (alternate) dual windows h of g with good, even optimal, frequency localization in a setting beyond the painless expansions region with arbitrarily small redundancy, but without having to invert the frame operator. The main features of our approach can be summarized as follows: (I) It uses 2-overlap (a = L/2) and works outside the region of painless non-orthonormal expansions and with arbitrarily small redundancy (ab)−1 > 1 of the Gabor frames. (II) It provides a natural parametrization of all dual windows (via an explicit formula) with sufficiently small support, given in terms of a measurable function z defined on a compact interval. (III) It provides optimal smoothness of the dual window h, e.g., dual windows h with the same smoothness as the original window, i.e., g, h ∈ C n(R). (IV) It yields support size of h only depended on the modulation parameter b ∈ (0, 2/L), not on properties of g (or h, e.g., smoothness). Note that since we work beyond the painless expansions region (that is, b > 1/L), the canonical dual may not even have compact support. By a result of Bölcskei [3], assuming rational redundancy (ab)−1, the canonical dual window has compact support if and only if the Zibulski-Zeevi representation of the frame operator Sg is unimodular in the frequency variable. 1.3 Results in the literature By dilation and translation of the Gabor system, we may without loss of generality take the translation parameter a = 1, the modulation parameter b ∈ (0, 1), and supp g = [−1, 1]. Christensen, Kim and Kim [7] characterize the frame property of {MbmTkg}m,k∈Z for g ∈ C 0(R) with supp g = [−1, 1] and finitely many zeros in (−1, 1). In particular, they inductively construct a continuous and compactly supported dual window once such a g generates a Gabor frame {MbmTkg}m,k∈Z. While the focus in [7] is on existence questions, we aim for explicit constructions for dual windows with symmetry and higher order smoothness, albeit for a smaller class of window functions as we do not allow g to have zeros inside the support. Gabor systems with continuous windows supported on an interval of length 2 have been considered in several recent papers [1,9] typically the dual windows constructed in these works are not continuous. There exist a considerable amount of work in Gabor analysis on explicit constructions of alternate dual windows, see, e.g., [5 -- 8, 10, 14, 16]. These constructions have the desirable date/time: 13-Feb-2019/1:32 3 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . feature that the dual window shares or inherits many of the properties of the window, e.g., smoothness and compact support. However, the methods from all the above cited works are restricted to the painless expansions region, e.g., redundancies (ab)−1 ≥ 2 or even (ab)−1 ≥ 3. The construction in [16] of C n-smooth dual windows supported on [−1, 1] is not guaranteed to work, however, when it does both windows are spline polynomials. Our method always works, but if g is a piecewise polynomial, the dual window will in general be a piecewise rational function of polynomials. The construction of Laugesen [16] is generally limited to the painless region, but by using a trick of non-linear dilation by a soft-thresholding type function, Laugesen is able to handle smaller redundancies, i.e., 1 < (ab)−1 < 2. However, this has the effect of making the windows constant of most of their support, which may not be desirable. In short, our work can be seen as a a continuation of [7, 8], but with the objective of [16] to construct smooth dual pairs of Gabor windows. 1.4 Outline In Section 2 we introduce the family of dual windows. Section 3 is the main contribution with a detailed analysis of properties (smoothness, symmetry, etc.) of the dual windows. In Section 4 we present examples of the construction. 2 The construction of the dual windows For each n ∈ Z≥0 ∪ {∞} we define the window classes: + (R) = {f ∈ C n(R) : supp f = [−1, 1] and f (x) > 0 for all x ∈ (−1, 1)} V n (2.1) Observe that the window classes are nested V n g in the largest of these window classes V 0 system {MbmTkg}m,k∈Z is a frame for L2(R) for any b ∈ (0, 1). + (R) ⊂ V n−1 + (R) for n ∈ Z>0, and that even for +(R), it is known [7, Corollary 2.8] that the Gabor We now introduce compactly supported functions hz that will serve as dual windows of +(R). Assume 0 < b < 1. Let kmax ∈ Z≥0 be the largest integer strictly smaller than g ∈ V 0 b/(1 − b), that is, kmax = max(cid:26)k ∈ Z≥0 : k < b 1 − b(cid:27) . Note that kmax ≥ 1 when 1/2 < b < 1. For any k ∈ {0, 1, . . . , kmax} we have k(1/b − 1) < 1. We define, for k ∈ {1, 2, . . . , kmax}, [k] = {1, 2, . . . , k}, and set [0] = ∅. For g ∈ V 0 +(R), define ψ : R → C by ψ(x) = for x ∈ R. (2.2) 1 Pn∈Z g(x + n) By the assumptions on g, the 1-periodic function ψ is well-defined, continuous, and satisfies c ≤ ψ(x) ≤ C for all x ∈ R (2.3) for some positive, finite constants c, C > 0. We will often consider ψ as a function on [0, 1] with ψ(0) = ψ(1) = 1/g(0) given by ψ(x) = 1/(g(x) + g(x − 1)) for x ∈ [0, 1]. date/time: 13-Feb-2019/1:32 4 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . Let z : [0, 1] → C be a measurable function. For each k ∈ {0, 1, . . . , kmax} we define the following auxiliary functions: ηk(x − k) = (−1)k Yj∈[k] g(x + 1 + j(1/b − 1)) g(x + j(1/b − 1)) for x ∈ [−1, −k(1/b − 1)] and [−g(x + 1)z(x + 1) + bψ(x + 1)] (2.4) γk(x + k) = (−1)k Yj∈[k] g(x − 1 − j(1/b − 1)) g(x − j(1/b − 1)) [g(x − 1)z(x) + bψ(x)] (2.5) for x ∈ [k(1/b − 1), 1]. We finally define hz : R → C by hz(x) = kmaxXk=0 ηk(x) 1[−k−1,−k/b](x) + kmaxXk=0 γk(x) 1[k/b,k+1](x) for x ∈ R \ {0}, (2.6) and hz(0) = bψ(0). More explicitly, hz is given as:  hz(x) = x ∈ [−k − 1, −k/b] , k = 1, . . . , kmax, ηk(x) −g(x + 1)z(x + 1) + bψ(x + 1) x ∈ [−1, 0) , g(x − 1)z(x) + bψ(x) γk(x) 0 x ∈ [0, 1] , x ∈ [k/b, k + 1] , k = 1, . . . , kmax, otherwise. 1 g(·−j(1/b−1)) We remark that the function γk(· + k) is indeed well-defined on [k(1/b − 1), 1] since the is well-defined on (−1 + k(1/b − 1), 1/b) and [k(1/b − 1), 1] ⊂ product Qj∈[k] (−1 + k(1/b − 1), 1/b). Note also that the productQj∈[k] g(· − 1 − j(1/b − 1)) has support on [k(1/b − 1), 1 + 1/b]. A similar consideration shows that ηk is well-defined on its domain. The function hz defined in (2.6) has compact support in [−kmax − 1, kmax + 1]. More precisely, supp hz ⊂ kmax[k=1 kmax[k=1 [−k − 1, −k/b] ∪ [−1, 1] [k/b, k + 1] . (2.7) The function hz is piecewise defined with ∪kmax k=0 {±k/b, ±(k + 1)} being the seam points of hz. The seam point x = 0 is special as it is the only seam point, where two nonzero functions in the definition of hz meet. In all other seam points, i.e., x 6= 0, the function hz is zero on one side of the seam points. Remark 2.1. For later use, we remark that Christensen, Kim, and Kim [7] consider dual windows h ∈ C 0(R) of g ∈ V 0 +(R) with compact support supp h ⊂ [−N , N ], where N := max {n ∈ Z>0 : n ≤ b/(1/b)} + 1 and b ≥ 1/2. When b/(1 − b) /∈ Z>0, then N = kmax + 1, and the support supp h ⊂ [−N, N ] corresponds to the support of hz in (2.7). On the other hand, if b/(1 − b) ∈ Z>0, there is a mismatch in the support relation as N = kmax + 2. However, in case b/(1 − b) ∈ Z>0 we have (N − 1)/b = N , and the interval [(N − 1)/b, N ] collapses to a point {N }. As a consequence, we can, in any case, use N := kmax + 1 when applying results from [7]. date/time: 13-Feb-2019/1:32 5 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . 3 Properties of the dual windows We first prove that hz indeed is a dual window of g as soon as {MbmTkhz}m,k∈Z is a Bessel system in Section 3.1. In Section 3.2 we show that the chosen parametrization z 7→ hz has several desirable properties. In Section 3.3 we show how to construct smooth dual windows h ∈ C n(R) for any g ∈ V n In Section 3.4 we discuss optimal smoothness of dual windows and show that the results in Section 3.3, in general, are optimal. In Section 3.5 we minimize the support length of the dual window in the sense of [8] while preserving the optimal smoothness. + (R) and n ∈ Z≥0 ∪ {∞}. However, we first introduce some notation used below. We use f (x+ 0 ) to denote the one- sided limit from the right limxցx0 f (x) and similarly f (x− 0 ) to denote the one-sided limit from the left limxրx0 f (x). For n ∈ Z≥0, we let Dn[f ] and f (n) denote the nth derivative of a function f : R → C, with the convention f (x) = D0[f ](x) = f (0)(x). We will repeatedly use the (general) Leibniz rule for differentiation of products: (f g)(n)(x) = nXℓ=0(cid:18)n ℓ(cid:19)f (n−ℓ)(x)g(ℓ)(x), (3.1) where(cid:0)n ℓ(cid:1) = n! ℓ!(n−ℓ)! is the binomial coefficient. A function f : [c, d] → C is said to piecewise C n if there exists a finite subdivision {x0, . . . , xk} of [c, d], x0 = c, xn = d such that f is C n on [xi−1, xi], the derivatives at xi−1 understood as right-handed and the derivatives at xi understood as left-handed, for ev- ery i ∈ {1, . . . , k}. Hence, if f is piecewise C n, the one-sided limits of f (m) exists everywhere for all m = 0, . . . , n, but the left and right limits may differ in a finite number of points. The definition of piecewise C n functions can be extended to functions on R, we will, however, only need it for compactly supported functions, where the modification is obvious. 3.1 Duality The following duality condition for two Gabor systems by Ron and Shen [20, 21] is central to our work; we use the formulation due to Janssen [12]. Theorem 3.1 ( [20]). Let b > 0 and g, h ∈ L2(R). Suppose {MbmTkg}m,k∈Z and {MbmTkh}m,k∈Z are a Bessel sequences. Then {MbmTkg}m,k∈Z and {MbmTkh}m,k∈Z are dual frames for L2(R), if and only if, for all k ∈ Z, g(x + k/b + n)h(x + n) = δ0,kb for a.e. x ∈ R. (3.2) Xn∈Z Since the infinite sums in (3.2) are 1-periodic, it suffices to verify (3.2) on any interval of length one. Furthermore, for supp g ⊂ [−1, 1], the duality conditions (3.2) become, for k 6= 0, g(x − k/b)h(x) + g(x − k/b − 1)h(x − 1) = 0 for a.e. x ∈ [k/b, k/b + 1] (3.3) and, for k = 0, g(x)h(x) + g(x − 1)h(x − 1) = b for a.e. x ∈ [0, 1] . (3.4) The following theorem shows that hz is a convenient representation of dual windows of g and justifies our interest in hz. date/time: 13-Feb-2019/1:32 6 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . Theorem 3.2. Let b ∈ (0, 1), let g ∈ V 0 +(R), and let z : [0, 1] → C be a measurable function. Then g and hz satisfy the characterizing equations (3.2). Hence, if {MbmTkhz}m,k∈Z is a Bessel sequence, e.g., if hz ∈ L∞(R), then {MbmTkg}m,k∈Z and {MbmTkhz}m,k∈Z are dual frames for L2(R). Proof. First, we check that (3.2) holds for k = 0. Since g(x)hz (x) = Xn∈Z it follows that, for a.e. x ∈ [0, 1), g(x) [−g(x + 1)z(x + 1) + bψ(x + 1)] a.e. x ∈ [−1, 0) g(x) [g(x − 1)z(x) + bψ(x)] 0 a.e. x ∈ [0, 1) otherwise, g(x + n)hz(x + n) = −g(x − 1)g(x)z(x) + g(x − 1)bψ(x) + g(x)g(x − 1)z(x) + g(x)bψ(x) = bψ(x) (g(x − 1) + g(x)) = b, where the final equality follows from the definition of ψ(x). This shows that (3.2) holds for k = 0. For k ≥ kmax + 1, the functions g(· − k/b) and hz have disjoint support. This follows from (2.7) and the fact that kmax + 1 ≤ (kmax + 1)/b. Consequently, equation (3.2) holds for k ≥ kmax + 1. To show that (3.2) holds k ∈ {1, . . . , kmax}, we will verify (3.2) on [k, k + 1). We first compute For a.e. x ∈ [k, k/b), it is trivial that g(x − k/b)hz(x) = Xn∈Z g(x − k/b)γk−1(x) a.e. x ∈ [k/b − 1, k) , g(x − k/b)γk(x) a.e. x ∈ [k/b, k + 1) , otherwise. 0 g(x − k/b + n)hz(x + n) = 0. On the other hand, for a.e. x ∈ [k/b, k + 1), g(x + k b + n)hz(x + n) = g(x − k b )γk(x) + g(x − k b − 1)γk−1(x − 1) (3.5) Xn∈Z We focus on the first term of the right hand side, which, by definition, is given as: g(x− k b )γk(x) = g(x− k b )(−1)k Yj∈[k] g(x − k − 1 − j( 1 b − 1)) g(x − k − j( 1 b − 1)) [g(x − k − 1)z(x − k) + bψ(x − k)] . Since g(x − k − 1 − j( 1 b − 1)) g(x − k − j( 1 b − 1)) = Yj∈[k] we can rewrite g(x − k b )γk(x) as follows: g(x − k g(x − k b − 1) b ) Yj∈[k−1] g(x − k − 1 − j( 1 b − 1)) g(x − k − j( 1 b − 1)) , g(x − k b )γk(x) = g(x − k b − 1)(−1)k Yj∈[k−1] g(x − k − 1 − j( 1 b − 1)) g(x − k − j( 1 b − 1)) date/time: 13-Feb-2019/1:32 7 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . · [g(x − k − 1)z(x − k) + bψ(x − k)] = −g(x − k b − 1)γk−1(x − 1). By inserting this back into (3.5), we obtain, for a.e. x ∈ [k/b, k + 1), g(x + k Xn∈Z b + n)hz(x + n) = −g(x − k b − 1)γk−1(x − 1) + g(x − k b − 1)γk−1(x − 1) = 0, which verifies (3.2) for k ∈ {1, . . . , kmax}. The calculations for k ∈ {−kmax, . . . , −1} are similar to the above calculations, hence we leave this case for the reader. The next result, Lemma 3.3, shows that not only is hz a convenient expression of dual windows of g ∈ V 0 +(R), it is in fact a parametrization by 1-periodic measurable functions z of all dual windows with sufficiently small support. The structure of the proof of Lemma 3.3 is somewhat similar to the proof of Lemma 3.3 in [7]. As the two results are quite different, we give the proof of Lemma 3.3. Lemma 3.3 (Parametrization of all compactly supported dual windows). Let b ∈ (0, 1) and g ∈ V 0 If {MbmTkg}m,k∈Z and {MbmTkh}m,k∈Z are dual frames, then h = hz for some measurable func- tion z : [0, 1] → C. +(R). Suppose h ∈ L2(R) has compact support in [−kmax − 1, kmax + 1]. Proof. Let h ∈ L2(R) be a dual window of g with supp h ⊂ [−kmax − 1, kmax + 1]. Lemma 3.2 in [7] says, see Remark 2.1, that supp h ⊂ kmax[k=1 [−k − 1, −k/b] ∪ [−1, 1] [k/b, k + 1] . kmax[k=1 Define a measurable function z on [0, 1] by: z(x) = h(x) − bψ(x) g(x − 1) a.e. x ∈ (0, 1) . This definition gives immediately that h(x) = hz(x) for a.e. x ∈ [0, 1]. By (3.2) with k = 0, we have for a.e. x ∈ (−1, 0) h(x) = −g(x + 1)h(x + 1) + b g(x) We continue, using that h(x) = hz(x) for a.e. x ∈ [0, 1], h(x) = −g(x + 1)(cid:2)−g(x)z(x + 1) + bψ(x + 1)(cid:3) + b g(x) = −g(x + 1)z(x + 1) + bψ(x + 1) = hz(x) for x ∈ (−1, 1) . Thus, also h(x) = hz(x) for a.e. x ∈ [−1, 1]. We will complete the proof by induction, showing that h(x) = hz(x) for a.e. x ∈ [−k − 1, −k/b] ∪ [−1, 1] k0[k=1 [k/b, k + 1] . k0[k=1 date/time: 13-Feb-2019/1:32 8 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . holds for all integers k0 in 1 ≤ k0 ≤ kmax. The base case k0 = 0 was verified above so we only have to show the induction step k0 − 1 → k0. We first consider x > 0. By induction hypothesis, we have h(x) = hz(x) = γk0−1(x) for a.e. x ∈ [(k0 − 1)/b, k0]. We aim to prove h(x) = γk0(x) for a.e. x ∈ [k0/b, k0 + 1], or rather h(x + 1) = γk0(x + 1) for a.e. x ∈ [k0/b − 1, k0]. Since [k0/b − 1, k0] ⊂ [k0/b − 1, k0/b], it follows by (3.2) for k = −k0: h(x + 1) = − g(x − k0/b) g(x − k0/b + 1) · h(x) for a.e. x ∈ [k0/b − 1, k0]. Since also [k0/b − 1, k0] ⊂ [(k0 − 1)/b, k0], using the induction hypothesis yields further: h(x + 1) = − g(x − k0 − k0(1/b − 1)) g(x − k0 + 1 − k0(1/b − 1)) · γk0−1(x) = γk0(x + 1) = hz(x + 1) for a.e. x ∈ [k0/b − 1, k0]. The argument for x < 0 is similar, hence we omit it. Obviously, there are choices of an unbounded function z that leads to hz not generating a Gabor Bessel sequence in L2(R), e.g., if hz /∈ L2(R), in which case {MbmTkg}m,k∈Z and {MbmTkhz}m,k∈Z are not dual frames. This is not contradicting Lemma 3.3. Lemma 3.3 should be compared to the well-known parametrization of all dual windows h by functions ϕ generating Bessel Gabor systems, see, e.g., [4, Proposition 12.3.6]; the parametriza- tion formula (3.6) is due to Li [17] and reads: h = S−1 g g + ϕ − Xm,n∈Z(cid:10)S−1g, MbmTakg(cid:11)MbmTakϕ. (3.6) The parametrization by Bessel generators ϕ is not injective nor explicit as one needs to com- pute both S−1 g g and an infinite series. On the other hand, the mapping z 7→ hz (z measur- able on [0, 1]) is injective and its range contains all dual windows with compact support on [−kmax − 1, kmax + 1]. Moreover, by Lemma 3.4 below, the mapping L∞([0, 1]) ∋ z 7→ hz is a bijective and explicit parametrization of all bounded dual windows with sufficiently small support, i.e., supp h ⊂ [−kmax − 1, kmax + 1]. Of course, the parametrization in Lemma 3.3 only works for our setting, in particular, only under the 2-overlap condition, i.e., a = L/2, while formula (3.6) works for all Gabor frames. 3.2 Basic properties: symmetry, boundedness, and continuity The next two lemmas show that the chosen parametrization z 7→ hz is rather natural. Indeed, both symmetry and boundedness properties of g and z are transferred to hz. Lemma 3.4 (Boundedness). Let b ∈ (0, 1). Suppose g ∈ V 0 only if z ∈ L∞(R). +(R). Then hz ∈ L∞([0, 1]) if and Proof. To show the "only if"-implication, note that for x ∈ [0, 1], we have hz(x) = g(x − 1)z(x) + bψ(x). Since g(x) is bounded (on R) and positive on (−1, 1), it follows that if hz is bounded on [0, 1], then so is z on [c, 1] for any c > 0. Using the boundedness of hz on [−1, 0] leads to the conclusion that z is bounded on [0, 1 − c] for any c > 0. For the converse assertion, let k ∈ {1, . . . , kmax}. Once we argue for the boundedness of γk and ηk on [k/b, k + 1] and [−k − 1, −k/b], respectively, the assertion is clear. We only consider date/time: 13-Feb-2019/1:32 9 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . γk as the argument of ηk is similar. Now, to argue for boundedness of γk, it suffices to show that the product g(x − j(1/b − 1)) for x ∈ [k(1/b − 1), 1] 1 Yj∈[k] − 1(cid:19) − k(cid:18) 1 0 = k(cid:18) 1 b is bounded. For all j ∈ [k] and all x ∈ [k(1/b − 1), 1] we have − 1(cid:19) ≤ x − j(cid:18) 1 b − 1(cid:19) ≤ 2 − 1 b < 1. b Let c denote the positive minimum of the continuous function g on the compact interval [−2 + 1/b, 2 − 1/b]. Then sup x∈[k(1/b−1),1] Yj∈[k](cid:12)(cid:12)(cid:12)(cid:12) which completes the proof. 1 g(x − j(1/b − 1))(cid:12)(cid:12)(cid:12)(cid:12) ≤ k/c, Lemma 3.5 (Symmetry). Let b ∈ (0, 1). Suppose g ∈ V 0 only if z is antisymmetric around x = 1/2, i.e., z(x) = −z(1 − x) for a.e. x ∈ [0, 1/2]. +(R) is even. Then hz is even if and Proof. Assume first z(x) = −z(1− x) for a.e. x ∈ R. If g is even, then so is ψ, and it follows by straightforward verification in the definition (2.6) that hz(x) = hz(−x) holds for a.e. x ∈ R. On the other hand, if hz is even, then using the definition of hz on [−1, 1], it follows easily that z is antisymmetric around x = 1/2. The last lemma of this subsection characterizes continuity of the dual windows hz in terms of easy verifiable conditions on z. Lemma 3.6 (Continuity). Let b ∈ (0, 1). Suppose g ∈ V 0 if z : [0, 1] → C is a continuous function satisfying +(R). Then hz ∈ C 0(R) if and only and z(0) = b ψ(0) g(0) = b g(0)2 z(1) = − b ψ(1) g(0) = − b g(0)2 . (3.7) (3.8) Proof. Suppose first that hz ∈ C 0(R). Then since for x ∈ (0, 1], we have g(x − 1) 6= 0 and hz(x) = g(x − 1)z(x) + bψ(x) with hz, g, ψ ∈ C 0(R), it follows that z is continuous on (0, 1]. Continuity of z at x = 0, i.e., existence of limxց0 z(x), follows by similar considerations of hz[−1,0). By continuity of hz and (2.7), we have hz(±1) = 0. Hence, 0 = hz(−1) = −g(0)z(0) + bψ(0) and 0 = hz(1) = g(0)z(1) + bψ(1) which shows the "only if"-implication. To show the other implication, we have to work a little harder. On the open set R \ kmax[k=0 {±k/b, ±(k + 1)}! , the function hz is continuous since it is a sum and product of continuous functions on this set. date/time: 13-Feb-2019/1:32 10 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . To show continuity at the seam points ∪kmax k=0 {±k/b, ±(k + 1)}, it suffices to show hz(0−) = hz(0+), hz(±k/b) = 0, k 6= 0, and hz(±(k + 1)) = 0. for k ∈ {0, 1, . . . , kmax}. Fix k ∈ {0, 1, . . . , kmax}. We first focus on the seam points in (−∞, −1]. For x = −k − 1 we immediately have hz(−k − 1) = (−1)k Yj∈[k] g(j(1/b − 1)) g(−1 + j(1/b − 1)) [−g(0)z(0) + bψ(0)] = 0 as the expression in the square brackets is zero by our assumption on z(0) in (3.7). For x = −k/b, k 6= 0, we note that hz(−k/b), by definition, contains the product g(−k/b + k + 1 + j(1/b − 1)) g(−k/b + k + j(1/b − 1)) Yj∈[k] as a factor. Further, since k ∈ [k], this product has g(−k/b+k+1+k(1/b−1)) factors. Since g(1) = 0 by the support and continuity assumption g ∈ V 0 hz(−k/b) = 0 for k = 1, . . . , kmax. g(−k/b+k+k(1/b−1)) = g(1) g(0) as one of its +(R), it follows that Now, we consider seam points in [1, ∞). For x = k + 1 we have by (3.8) that hz(k + 1) = (−1)k Yj∈[k] g(−j(1/b − 1)) g(1 − j(1/b − 1)) [g(0)z(1) + bψ(1)] = 0. To see hz(k/b) = 0 we note that the product defining hz(k/b) contains the factor g(−1)/g(0) which is zero due to the assumption g ∈ V 0 +(R). Finally, we show continuity of hz(x) at x = 0. However, this follows readily by considering the two one-sided limits x ր 0 and x ց 0 of hz(x): hz(0−) = −g(1)z(1) + b ψ(1) = b ψ(1) = b/g(0) and respectively. hz(0+) = g(−1)z(0) + b ψ(0) = b ψ(0) = b/g(0), 3.3 Higher order smoothness The main result of this section, Theorem 3.9, characterizes C n-smoothness of hz in terms of conditions on z. As these conditions involves derivatives of ψ, more precisely, ψ(m)(0), the next lemma shows how to operate with this condition. Lemma 3.7. Let n ∈ Z>0, and let g ∈ V n + (R). Then ψ := 1 Pn∈Z g(·+n) ∈ C n(R) and ψ(n)(0) = Xm∈M n! (−1)m1+···+mn(m1 + · · · + mn)! m1!m2! · · · mn! g(0)m1+···+mn+1 j! !mj nYj=1 g(j)(0) , (3.9) where M := {(m1, . . . , mn) ∈ (Z≥0)n1 · m1 + 2 · m2 + · · · + n · mn = n}. date/time: 13-Feb-2019/1:32 11 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . Proof. Since g ∈ V n constant. Hence, since the mapping x 7→ 1/x is C ∞ on (0, ∞), we have ψ ∈ C n(R). + (R), the sumPn∈Z g(·+n) is in C n(R) and is bounded below by a positive As usual, we consider ψ on [0, 1] where the function is given by x 7→ 1/(g(x) + g(x − 1)). We will use the following version of Faà di Bruno's formula: dn dxn f (h(x)) = Xm∈M n! m1!m2! · · · mn! f (m1+···+mn) (h(x)) Taking f (x) = 1 x and h(x) = g(x) + g(x − 1) in Faà di Bruno's formula yields . nYj=1 h(j)(x) j! !mj nYj=1 g(j)(0) + g(j)(−1) nYj=1 g(j)(0) j! !mj j! , (3.10) !mj ψ(n)(0) = Xm∈M = Xm∈M n! n! (−1)m1+···+mn(m1 + · · · + mn)! m1!m2! · · · mn! (g(0) + g(−1))m1+···+mn+1 (−1)m1+···+mn(m1 + · · · + mn)! m1!m2! · · · mn! g(0)m1+···+mn+1 using dm(x−1)/dxm = (−1)mm!x−(m+1) for m ∈ Z>0 in the first equality and g(ℓ)(−1) = 0 for ℓ = 0, 1, . . . , n in the second. Example 3.8. We illustrate the computation of Lemma 3.7 for n = 1, 2. For n = 1, since M = {1}, formula (3.9) simply yields ψ(1)(0) = − g(1)(0) g(0)2 . For n = 2, we have M = {(2, 0), (0, 1)}, whereby ψ(2)(0) = 2 g(1)(0)2 g(0)3 − g(2)(0) g(0)2 . Theorem 3.9. Let n ∈ Z>0∪{∞}, and let g ∈ V n + (R). The following assertions are equivalent: (i) z ∈ C n([0, 1]) satisfies (3.7), (3.8), and, for each m = 1, . . . , n, z(m)(0) = − z(m)(1) = − and (ii) hz ∈ C n(R). mXℓ=1(cid:18)m ℓ(cid:19) g(ℓ)(0) g(0) mXℓ=1(cid:18)m ℓ(cid:19) g(ℓ)(0) g(0) z(m−ℓ)(0) + b ψ(m)(0) g(0) z(m−ℓ)(1) − b ψ(m)(0) g(0) , . (3.11) (3.12) Proof. We first prove the assertion (i)⇒(ii). Consider the open set J := R \ kmax[k=0 {±k/b, ±(k + 1)}! . The function hzJ is in C n(J) since it is a sum and product of C n(J) functions. date/time: 13-Feb-2019/1:32 12 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . To prove C n-smoothness at the seam points ∪kmax k=0 {±k/b, ±(k + 1)}, we need to show that h(m) z (0−) = h(m) z (0+), h(m) z (±k/b) = 0, k 6= 0, and h(m) z (±(k + 1)) = 0 for m = 1, . . . , n and k ∈ {0, 1, . . . , kmax}. The proof of this is split into the three cases: x = 0, x ∈ kmax[k=0 {±k/b} and x ∈ {±(k + 1)} . kmax[k=0 However, we first note that g ∈ V n + (R) implies g(m)(±1) = 0 for m = 0, . . . , n, C n-smoothness at x = 0. For x ∈ [0, 1] h(m) z (x) = mXℓ=0(cid:18)m ℓ(cid:19)g(ℓ)(x + 1)z(m−ℓ)(x + 1) + bψ(m)(x). Since g(m)(1) = 0 for m = 0, . . . , n, it is readily seen that mXℓ=0(cid:18)m ℓ(cid:19)g(ℓ)(1)z(m−ℓ)(1) + bψ(m)(0) h(m) z (0−) = = bψ(m)(0) for all m ≤ n. Similarly, by considering x ∈ [−1, 0), we see that h(m) z (0+) = − mXℓ=0(cid:18)m ℓ(cid:19)g(ℓ)(−1)z(m−ℓ)(0) + bψ(m)(1) = bψ(m)(0), where the last equality follows from g(m)(−1) = 0 for m = 0, . . . , n and periodicity of ψ. In the two remaining cases, we will use the following easy consequence of the Leibniz rule (3.1). If f, g ∈ C n(R) and g(m)(x0) = 0 for m = 0, 1, . . . , n, then (f g)(m)(x0) = 0 for m = 0, 1, . . . , n, where f g = x 7→ f (x)g(x). C n-smoothness at x = ±k/b, k = 1, . . . , kmax. We first consider x = k/b, k = 1, . . . , kmax. By rearranging the terms in the definition of the auxiliary function γk(x + k) in (2.5), we obtain γk(x + k) = g(x − 1 − k(1/b − 1)) · (−1)k g(x − k(1/b − 1))) Yj∈[k−1] g(x − 1 − j(1/b − 1)) g(x − j(1/b − 1)) [g(x − 1)z(x) + bψ(x)] , for x ∈ [k(1/b − 1), 1]. Thus, for x ∈ [k/b, k + 1], we will consider γk(x) a product of the functions g(x − 1 − k/b) and (−1)k g(x − k/b) Yj∈[k−1] g(x − k − 1 − j(1/b − 1)) g(x − k − j(1/b − 1)) [g(x − k − 1)z(x − k) + bψ(x − k)] . date/time: 13-Feb-2019/1:32 13 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . Focusing on the derivatives of g(x − 1 − k/b)) at x = k/b, we get Dm [g(x − 1 − k/b)] (k/b) = g(m)(−1) = 0 for m = 0, 1, . . . , n. Hence, as a consequence of the Leibniz rule (3.1), γ(m) k (k/b) = 0, for m = 1, 2, . . . , n. The calculations for x = −k/b, k = 1, . . . , kmax are similar to the ones above hence we leave these to the reader. C n-smoothness at x = ±(k + 1), k = 1, . . . , kmax. Again, we first consider x = k + 1, k = 1, . . . , kmax. For x ∈ [k/b, k + 1], we consider γk(x) as a product of the two functions and g(x − k − 1 − j(1/b − 1)) g(x − k − j(1/b − 1)) Yj∈[k] g(x − k − 1)z(x − k) + bψ(x − k). We focus on the second of the two and observe that, for each m = 1, 2, . . . , n, Dm [g(· − k − 1)z(· − k) + bψ(· − k)] (k + 1) = = g(0)"z(m)(1) + ψ(m)(0) g(0) # = 0, mXℓ=0(cid:18)m ℓ(cid:19)g(ℓ)(0)z(m−ℓ)(1) + bψ(m)(0) mXℓ=1(cid:18)m ℓ(cid:19)g(ℓ)(0) z(m−ℓ)(1) + b g(0) where the final equality follows from the assumption (3.12). Thus, as a consequence of the Leibniz rule (3.1), we arrive at γ(m) k (k + 1) = 0, for m = 1, 2, . . . , n. The calculations for x = −k − 1, k = 1, . . . , kmax are similar to the ones above, hence we leave these to the reader. The proof of assertion (ii)⇒(i) is similar to the argument in the proof of Lemma 3.6, hence we will omit the proof. Example 3.10. We compute (3.11) and (3.12) for n = 2 using Example 3.8. For m = 1, using that z also has to satisfy (3.7) and (3.8), we get z(1)(0) = −z(1)(1) = −2b g(1)(0) g(0)3 . Similarly, for m = 2, using that z satisfies (3.7), (3.8), and (3.13), we get z(2)(0) = −z(2)(1) = 6b g(1)(0)2 g(0)4 − 2b g(2)(0) g(0)3 . (3.13) (3.14) It is easy to find a function z : [0, 1] → C satisfying the conditions in (i) in Theorem 3.9. E.g., if n < ∞, a polynomial z of degree 2n + 2 will always do. Further, as the conditions on z ∈ C n(R) only concern the derivatives of z(x) at the boundary points x = 0 and x = 1, there is an abundance of C n dual windows of each g ∈ V n + (R) for any value of b ∈ (0, 1). Hence, date/time: 13-Feb-2019/1:32 14 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . given a window g ∈ V n example of such constructions will be given in Section 4. + (R), we can easily construct dual windows in C n(R) using Theorem 3.9; We now exhibit a large class of window functions g ∈ V n + (R) containing, e.g., all symmetric windows and all windows forming a partition of unity, for which the issue of computing ψ(m)(0) used in Theorem 3.9 disappears. Corollary 3.11. Let n ∈ Z>0 ∪ {∞}. Suppose g ∈ V n 1, . . . , n. Then the following assertions are equivalent: + (R) satisfies g(m)(0) = 0 for m = (i) z ∈ C n([0, 1]) satisfies (3.7), (3.8), and, for each m = 1, . . . , n, and (ii) hz ∈ C n(R). In particular, if then hz ∈ C n(R). z(m)(0) = 0, z(m)(1) = 0, z(x) = b g(0)3(cid:2)2g(x) − g(0)(cid:3), (3.15) (3.16) (3.17) Proof. By Lemma 3.7, it follows that ψ(m)(0) = 0 for m = 1, . . . , n. With g(m)(0) = 0 and ψ(m)(0) = 0 for all m = 1, . . . , n conditions (3.11) and (3.12) reduce to (3.15) and (3.16), respectively. Finally, it is straightforward to verify that z defined by (3.17) satisfies the 2n + 2 conditions in (i). Remark 3.12. (a) Suppose g ∈ V n 0 for m = 1, . . . , n. Now, + (R) satisfies the assumptions of Corollary 3.11, i.e., g(m)(0) = if h ∈ C n(R) and b ∈ C satisfy the window condition condition we see that h(m)(0) = 0 for m = 1, . . . , n. Thus, any dual window in C n(R), not necessarily with compact support, will also have this property. Pn∈Z g(x + n)h(x + n) = b for x ∈ R, then by term-wise differentiating the window + (R) satisfies eitherPn∈Z g(x+n) = 1 or g(x) = g(−x) for x ∈ R (or both). Then g satisfies the assumptions of Corollary 3.11, i.e., g(m)(0) = 0 for m = 1, . . . , n. For g symmetric, this is obvious. If g ∈ V n + (R) forms a partition of unity, then, by differentiating g(x) + g(x − 1) = 1 for x ∈ [0, 1], one will see that g(m)(0) = 0 for m = 1, . . . , n. (b) Suppose g ∈ V n (c) The dual window hz defined by (3.17) in Corollary 3.11 is often a convenient choice as it guarantees that the dual window hz is defined only in terms of the window g. Hence, if g is, e.g., a piecewise polynomial, then hz becomes a piecewise rational function of polynomials. However, if g is symmetric, hz defined by (3.17) will only be symmetric, if g(x) = g(0) − g(1 − x) for x ∈ [0, 1], that is, if the graph of g, restricted to [0, 1] × C, is anti-symmetric around (1/2, g(0)/2). date/time: 13-Feb-2019/1:32 15 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . 3.4 Optimality of smoothness The first result of this section shows that, even though there are an abundance of dual windows in C n(R) for g ∈ V n + (R) \ C n+1(R), additional smoothness, e.g., dual windows in C n+1(R), is in general not possible. Proposition 3.13. Let b ∈ (0, 1). Let n ∈ Z≥0, and let g ∈ V n for x ∈ R. Assume g is a real-valued, piecewise C n+1-function for which g(n+1) has a simple If h ∈ L2(R) is compactly supported in discontinuity at x = −1, x = 0, and/or x = 1. [−kmax − 1, kmax + 1], and {MbmTkg}m,k∈Z and {MbmTkh}m,k∈Z are dual frames, then h /∈ C n+1(R). + (R) with Pn∈Z g(x + n) = 1 Proof. Assume towards a contradiction that h = hz ∈ C n+1(R). Recall that g ∈ V n + (R) implies g(m)(±1) = 0 for m = 0, . . . , n. We only consider the case g(x) > 0 for x ∈ (−1, 1) as the argument for g(x) < 0 is similar. Since g(x) > 0 for x ∈ (−1, 1), it follows that g(n+1)(−1+) ≥ 0 and g(n+1)(1−) ≤ 0. Assume that g(n+1) is discontinuous at x = −1 and/or x = 1, i.e., g(n+1)(−1+) > 0 and/or g(n+1)(1−) < 0. The case x = 0 follows from Theorem 3.16 below. As in the proof of Theorem 3.9, we see that h(n+1) z (0−) = gn+1(1−)z(1) = −gn+1(1−)b/g(0)2 and h(n+1) z (0+) = −gn+1(−1+)z(0) = −gn+1(−1+)b/g(0)2 Since, by assumption, h(n+1) is a contradiction. z (0−) = h(n+1) z (0+), it follows that gn+1(1−) = gn+1(−1+), which Example 3.14. Let g = max(0, 1 − x) be the second cardinal B-spline with uniform knots. For any b ∈ (0, 1), the Gabor system {MbmTkg}m,k∈Z is a frame for L2(R). Since g′(x) is discontinuous at x = −1 (and at x = 0 and x = 1), it follows by Proposition 3.13 that no dual windows h ∈ C 1(R) ∩ L2(R) with support in [−kmax − 1, kmax + 1] exists for any value of b. Let us comment on the assumptions of Proposition 3.13. The location of the discontinuity of g(n+1) is important. In fact, if discontinuities of g(n+1) avoid certain points, the conclusion h /∈ C n+1(R) of Proposition 3.13 may not hold. On the other hand, we assume the partition of unity property of g only for convenience as to simplify the proof. Furthermore, as we see by the next two results, positivity of g(x) on (−1, 1) and compact support of the dual window h are also not essential for obstructions results on the smoothness of dual windows. For n = −1, we ignore the requirement g ∈ C n(R) in the formulation below. Lemma 3.15. Let n ∈ Z≥−1, and let g ∈ C n(R) be a piecewise C n+1-function. Let {xr}r∈[R] = {x1, . . . , xR} denote the finite set of points, where g(n+1) has simple discontinuities. Assume h ∈ C n+1(R) and the constants a > 0, b ∈ C satisfy the window condition g(x + an)h(x + an) = b, for all x ∈ [−a/2, a/2] . (3.18) Xn∈Z Let {tr}r∈[R] = {xr}r∈[R] (mod a) so that tr ∈ [−a/2, a/2) and tr ≤ tr+1. Set t0 = −a/2 and tR+1 = a/2. Suppose Pn∈Z Dm[g(· + an)h(· + an)] converges uniformly on [tr, tr+1] for m = 1, . . . , n + 1 and r = 0, . . . , R. Then, for each r = 1, . . . , R, (3.19) X{s∈[R]:xs−xr∈aZ}hg(n+1)(x+ s ) − g(n+1)(x− s )i h(xr) = 0. date/time: 13-Feb-2019/1:32 16 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . In particular, if {s ∈ [R] : xs − xr ∈ aZ} = {r}, then h(xr) = 0. Proof. The following argument is inspired by the proof of Lemma 1 in [16]. Fix r ∈ [R]. Uniform convergence ofPn∈Z Dm[g(· + an)h(· + an)] for each m = 0, 1, . . . , n + 1 allows us to differentiate the window condition (3.18) term by term [22, Theorem 7.17]. Differentiating m times then gives: Xn∈Z>0 mXℓ=0(cid:18)m ℓ(cid:19)g(ℓ)(x + an)h(m−ℓ)(x + an) = 0, for all x ∈ (tr, tr+1) . (3.20) for all r = 0, 1, . . . , R and m = 1, . . . , n. Note that the sum (3.20) is a periodic. Hence, by subtracting the two one-sided limits x ր tr and x ց tr of (3.20), we obtain (3.19). Recall that if {MbmTakg}m,k∈Z and {MbmTakh}m,k∈Z with g, h ∈ L2(R) are dual frames for L2(R), then (3.18) holds. Hence, under the assumptions of Lemma 3.15, duality of g and h restricts the possible values of h on {xj}j∈J . For windows g with support in [−1, 1], Lemma 3.15 leads to the following general obstruction result. Theorem 3.16. Let b ∈ (0, 1) and h ∈ L2(R). Let n ∈ Z≥−1, and let g ∈ C n(R) be a piecewise C n+1-function with supp g ⊂ [−1, 1], and let {xj}j∈J ⊂ [−1, 1] denote the finite set of points, where g(n+1) is discontinuous. Assume either (i) 0 ∈ {xj}j∈J and supp h ⊂ [−kmax − 1, kmax + 1], or (ii) 0 ∈ {xj}j∈J and ±1 /∈ {xj}j∈J . If {MbmTkg}m,k∈Z and {MbmTkh}m,k∈Z are dual frames, then h /∈ C n+1(R). Proof. Assume towards a contradiction that h ∈ C n+1(R). From supp g ⊂ [−1, 1], it follows that g(k) = 0 for all k ∈ Z \ {0} and that {xj}j∈J ⊂ [−1, 1]. Hence, equation (3.2) for k = 0 implies that h(0) = b/g(0) > 0. Depending on whether we use assumption (i) or (ii), the points x = ±1 may or may not belong to {xj}j∈J . In either case, g(n+1)(−1−) = g(n+1)(1+) = 0 and we have from (3.19) that 0 =hg(n+1)(−1+) − 0)i h(−1) +hg(n+1)(0+) − g(n+1)(0−)i h(0) +h0 − g(n+1)(1−)i h(1) If we use assumption (i), then, by the compact support of h, we have from [7, Lemma 3.2] that h(±1) = 0. On the other hand, from assumption (ii), we have g(n+1)(−1+) = g(n+1)(1−) = 0. In either case, we get which is a contradiction to 0 ∈ {xj}j∈J and h(0) > 0. hg(n+1)(0+) − g(n+1)(0−)i h(0) = 0, The conditions on the window g in the above results should be understood as follows. In order to make the statement as strong as possible, we want generators g just shy of being in the C n+1-class. Hence, the function g is assumed to be C n everywhere and piecewise C n+1 except at a finite number of points, where both one-sided limits of g(n+1) exist, but do not agree. The following example shows a general, but typical, obstruction of the smoothness of dual windows. Example 3.17. Let b ∈ (0, 1), n ∈ Z>0, and let g ∈ C n(R) with supp g ⊂ [−1, 1] be a C ∞-function except at x = 0, where g(n+1) fails to be continuous. Suppose the Gabor system {MbmTkg}m,k∈Z is a frame for L2(R). Then, by Theorem 3.16, it follows that no dual windows h ∈ C n+1(R) ∩ L2(R) exists. Note that this conclusion holds whether or not h is assumed to have compact support. date/time: 13-Feb-2019/1:32 17 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . 3.5 Small support While the previous section considered optimality of the smoothness of the dual windows, we are here concerned with optimizing, that is, minimizing, the support length. Such questions were considered in [8], where the authors characterized the existence of continuous dual windows with short support for continuous windows g with finitely many zeros inside their support [−1, 1]. In the following result we consider the possibility of higher order smoothness of dual windows with short support. Theorem 3.18. Let n ∈ Z≥0, let b ∈ [ N Define N +1 , 2N 2N +1 ) for some N ∈ Z>0, and let g ∈ V n + (R). z(x) = b − 1), N ( 1 b ψ(x) g(x) zmid (x) −b ψ(x) b − 1)(cid:3) , x ∈(cid:2)0, 1 − N ( 1 x ∈(cid:0)1 − N ( 1 b − 1), 1(cid:3) , g(x−1) x ∈(cid:2)N ( 1 b − 1)(cid:1) → C is a measurable function. The following assertions b − 1)(cid:1) , b − 1), N ( 1 (3.21) where zmid :(cid:0)1 − N ( 1 hold: (a) The dual window hz has compact support in [−N , N ]. (b) hz ∈ C n(R) if and only if z ∈ C n([0, 1]). (c) Suppose g is even. Then hz is even if and only if zmid is antisymmetric around x = 1/2, i.e., zmid (x) = −zmid (1 − x) for a.e. x ∈(cid:0)1 − N(cid:0) 1 Proof. By definition of N , we have N ( 1 b − 1(cid:1), 1/2(cid:3). b − 1) ≤ 1, hence, z is well-defined. (a): If kmax < N , then supp hz ⊂ [−kmax − 1, kmax + 1] ⊂ [−N, N ] by (2.7). Assume now that kmax ≥ N and consider the dual windows hz on [N, kmax + 1]. For x > N , it suffices to show that hz(x) = 0 for x ∈ (k/b, k + 1) and k = N, . . . , kmax. Recall that, for any k = 1, . . . , kmax, for x ∈ (k/b, k + 1), which can be rewritten as hz(x) = γk(x) hz(x + k) = γk(x + k) = (−1)k Yj∈[k] g(x − 1 − j(1/b − 1)) g(x − j(1/b − 1)) [g(x − 1)z(x) + bψ(x)] for x ∈ (k/b − k, 1). Since k ≥ N , we have the inclusion(cid:0)k( 1 g(x − 1)z(x) + bψ(x) = g(x − 1)(cid:18)−b b − 1), 1(cid:1) ⊂(cid:0)N ( 1 g(x − 1)(cid:19) + bψ(x) = 0. ψ(x) b − 1), 1(cid:1). Hence Thus hz(x) = 0 for x > N . The argument for x < −N is similar so we omit it. (b): By inserting x = 0 and x = 1 into (3.21), it can easily be seen that z satisfies (3.7) and (3.8), respectively. Therefore, the result for n = 0 simply follows from Lemma 3.6. For n > 0, the "only if"-assertion follows directly from Theorem 3.9. To prove the other direction, we assume that z ∈ C n([0, 1]). From definition (3.21) we have: g(x)z(x) − bψ(x) = 0 for all x ∈(cid:2)0, 1 − N ( 1 b − 1)(cid:3) . By differentiating both sides m times, isolating z(m)(x) and inserting x = 0, we see that z satisfies (3.11). In a similar way, one proves that z satisfies (3.12). Hence, by Theorem 3.9, hz ∈ C n(R). date/time: 13-Feb-2019/1:32 18 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . (c): From (3.21) we see that, for x ∈(cid:2)0, 1 − N(cid:0) 1 b − 1(cid:1)(cid:3), the function z satisfies −z(1 − x) = b ψ(1 − x) g(−x) = b ψ(x) g(x) = z(x), where the second equality uses that g and ψ are even and that ψ is 1-periodic. Hence, if g is even, then zmid is antisymmetric around x = 1/2 if and only if z defined by (3.21) is antisymmetric around x = 1/2. The conclusion now follows from Lemma 3.5. The short support of Theorem 3.18 is optimal in the following sense of [8]: If a dual window h with support supp h ⊂ [−N , N ] exists, then necessarily b ≤ 2N/(2N + 1), see [8, Theorem 2.3]. If, in addition, h is assumed to be continuous, then b < 2N/(2N + 1), see [8, Theorem 2.5]. 4 Examples of the construction In this section, we present two examples of the construction procedure of dual windows using the results from the previous sections. In Example 4.1 we construct dual windows of the clas- sical and widely used Hann and Blackman window, respectively. In Example 4.2 we consider a smoother, but non-symmetric window; the setup is more complicated than Example 4.1 and perhaps less useful for applications, but it serves as a proof of concept of the flexibility of our method. Example 4.1. The Hann window ghann ∈ C 1(R) is defined by 1 1 2 cos(π(x + 1)) x ∈ [−1, 0) 2 cos(π(x)) 2 − 1 2 + 1 0 ghann (x) = cos2(πx/2)χ[−1,1](x) = gblac(x) =(cid:2)0.42 + 0.5 cos(πx) + 0.08 cos(2πx)(cid:3)χ[−1,1](x) x ∈ [0, 1] otherwise, and the Blackman window gblac ∈ C 1(R) is defined by for x ∈ R, see Figure 1. Both these widows are continuously differentiable, symmetric, non- negative, and normalized g(0) = 1, and the Hann window even has the partition of unity property. Both of the windows belong to V 1 +(R); hence, the optimal smoothness of compactly supported dual windows are h ∈ C 1(R). +(R), but not V 2 As an example, let us consider the modulation parameter b = 3/5. By definition of kmax we get kmax = 1. Thus, the dual windows hz defined in (2.6) will have support in [−2, 2]. Since g is a trigonometric polynomial on [−1, 1], it is natural to take z to be a trigonometric polynomial as well. For the Hann window the standard choice (3.17) is: zhann (x) = b cos(πx) for x ∈ [0, 1] , (4.1) while (3.17) for the Blackman window becomes: zblac(x) = b(cid:2)−0.16 + 0.5 cos(πx) + 0.08 cos(2πx)(cid:3) for x ∈ [0, 1] . (4.2) Figure 2 shows dual windows of the Hann window hhann and of the Blackman window hblac defined using z from (4.1) and (4.2), respectively. While zhann is anti-symmetric around x = 1/2, this is not the case for the chosen zblac; see Lemma 3.5 and Remark 3.12(c). date/time: 13-Feb-2019/1:32 19 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . 1 1 -1 0 x 1 0 1 2 γ 3 Figure 1: Left: the Hann window ghann ∈ V 1 +(R) (red) and the Blackman window gblac ∈ V 1 +(R) (blue). Right: The Fourier transform of the Hann window bghann (red) and of the Blackman windowbgblac (blue). Both windows and their Fourier transforms are real and symmetric. 1 -3 -2 -5/3 -1 0 1 5/3 2 x 3 Figure 2: Dual windows hhann (red) and hblac (blue) of the Hann and Blackman window based on zhann and zblac defined in (4.1) and (4.2), respectively. Both windows are in C 1(R) and with support supp h = [−2, −5/3] ∪ [−1, 1] ∪ [5/3, 2]. We can actually decrease the support size of the dual windows without sacrificing the C 1-smoothness. By applying Theorem 3.18 with N = 1 and taking zmid to be the unique third degree trigonometric polynomial zmid (x) = c1 cos(πx) + c3 cos(3πx) so that z ∈ C 1(R), we obtain dual windows in C 1(R) with support in [−1, 1]. It turns out that the support of the dual windows even shrink to [−2/3, 2/3] for this specific setup. Since the two constructed functions zmid are anti-symmetric around x = 1/2, it follows by Theorem 3.18(c) that both these dual windows will be symmetric. The short-support dual windows of the Hann and Blackman window and their Fourier transform are shown in Figure 3. The next example illustrates the construction of dual windows when g does not have zero derivatives at the origin and the redundancy (ab)−1 = 3π/7 is irrational. Example 4.2. We take β to be a spline defined as: β(x) = x ∈ [−1, −4/5] , x ∈ [−4/5, 4/5] , p(x) 1 p(−x) x ∈ [4/5, 1] , 0 otherwise,  where p(x) = 10625 − 60000x + 135000x2 − 151250x3 + 84375x4 − 18750x5 is the five-degree polynomial satisfying p(1) = p′(1) = p′′(1) = p′(4/5) = p′′(4/5) = 0 and p(4/5) = 1. Then date/time: 13-Feb-2019/1:32 20 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . 1 1 0 x 1 1 2 3 4 γ 5 Figure 3: Left: Dual windows in C 1(R) with short support on [−2/3, 2/3]. The dual of the Hann window is shown in red, while the dual of the Blackman window is shown in blue. Right: The Fourier transform the two dual windows shown left: the Fourier transform of the dual Hann window (red) and the Fourier transform of the dual Blackman window (blue). Both dual windows and their Fourier transforms are real and symmetric. 1 0.5 -0.5 -3 -2 -1 1 2 x 3 Figure 4: The window function g ∈ V 2 g and h are C 2-functions, and h has support in [−3, 3]. +(R) (red) and a dual window hz (blue) for b = 7 3π . Both β ∈ C 2(R) is a bump function supported on [−1, 1]. We consider the window g ∈ V 2 defined by +(R) g(x) = 1 16(cid:0)2 − (x − 5)(x + 3)(cid:1)β(x). As an example of an irrational modulation parameter, let us consider b = 7 so the support of hz is: 3π . Then kmax = 2 supp hz = [−3, −6π/7] ∪ [−2, −3π/7] ∪ [−1, 1] ∪ [3π/7, 2] ∪ [6π/7, 3] ⊂ [−3, 3] . We chose z to be the unique polynomial of degree five that satisfies the six conditions of Theorem 3.9 for n = 2; these conditions are explicitly computed in Example 3.10. It follows that hz ∈ C 2(R). The graphs of g and the dual window hz are shown in Figure 4. References [1] A. G. D. Atindehou, Y. B. Kouagou, and K. A. Okoudjou. On the frame set for the 2-spline, preprint 2018. arXiv:1806.05614. [2] P. Balazs, H. G. Feichtinger, M. Hampejs, and G. Kracher. Double preconditioning for gabor frames. IEEE Transactions on Signal Processing, 54(12):4597 -- 4610, dec 2006. doi:10.1109/tsp.2006.882100. date/time: 13-Feb-2019/1:32 21 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . [3] H. Bölcskei. A necessary and sufficient condition for dual weyl-heisenberg frames to be compactly supported. The Journal of Fourier Analysis and Applications, 5(5):409 -- 419, sep 1999. doi:10.1007/bf01261635. [4] O. Christensen. An introduction to frames and Riesz bases. Applied and Nu- second edition, 2016. [Cham], merical Harmonic Analysis. Birkhäuser/Springer, doi:10.1007/978-3-319-25613-9. [5] O. Christensen. Pairs of dual Gabor frame generators with compact support and desired frequency localization. Appl. Comput. Harmon. Anal., 20(3):403 -- 410, 2006. doi:10.1016/j.acha.2005.10.003. [6] O. Christensen and S. S. Goh. From dual pairs of Gabor frames to dual pairs of wavelet frames and vice versa. Appl. Comput. Harmon. Anal., 36(2):198 -- 214, 2014. doi:10.1016/j.acha.2013.03.005. [7] O. Christensen, H. O. Kim, and R. Y. Kim. Gabor windows supported on [−1, 1] and compactly supported dual windows. Appl. Comput. Harmon. Anal., 28(1):89 -- 103, 2010. doi:10.1016/j.acha.2009.07.004. [8] O. Christensen, H. O. Kim, and R. Y. Kim. Gabor windows supported on [−1, 1] and dual windows with small support. Adv. Comput. Math., 36(4):525 -- 545, 2012. doi:10.1007/s10444-011-9189-0. [9] O. Christensen, H. O. Kim, and R. Y. Kim. On Gabor frames generated by sign- changing windows and B-splines. Appl. Comput. Harmon. Anal., 39(3):534 -- 544, 2015. doi:10.1016/j.acha.2015.02.006. [10] O. Christensen and R. Y. Kim. Pairs of explicitly given dual Gabor frames in L2(Rd). J. Fourier Anal. Appl., 12(3):243 -- 255, 2006. doi:10.1007/s00041-005-5052-3. [11] I. Daubechies, A. Grossmann, and Y. Meyer. Painless nonorthogonal expansions. J. Math. Phys., 27(5):1271 -- 1283, 1986. doi:10.1063/1.527388. [12] A. J. E. M. Janssen. The duality condition for Weyl-Heisenberg frames. In Gabor analysis and algorithms, Appl. Numer. Harmon. Anal., pages 33 -- 84. Birkhäuser Boston, Boston, MA, 1998. [13] A. J. E. M. Janssen and P. L. Søndergaard. Iterative algorithms to approximate canonical Gabor windows: computational aspects. J. Fourier Anal. Appl., 13(2):211 -- 241, 2007. doi:10.1007/s00041-006-6069-y. [14] I. Kim. Gabor frames with trigonometric spline dual windows. Asian-Eur. J. Math., 8(4):1550072, 32, 2015. doi:10.1142/S1793557115500722. [15] T. Kloos, J. Stöckler, and K. Gröchenig. Implementation of discretized Ga- IEEE Trans. Inform. Theory, 62(5):2759 -- 2771, 2016. bor frames and their duals. doi:10.1109/TIT.2016.2541918. [16] R. S. Laugesen. Gabor dual spline windows. Appl. Comput. Harmon. Anal., 27(2):180 -- 194, 2009. doi:10.1016/j.acha.2009.02.002. [17] S. Li. On general frame decompositions. Numer. Funct. Anal. Optim., 16(9-10):1181 -- 1191, 1995. doi:10.1080/01630569508816668. date/time: 13-Feb-2019/1:32 22 of 23 Lemvig, Nielsen Gabor windows supported on [−1, 1] and construction of . . . [18] S. Moreno-Picot, F. J. Ferri, M. Arevalillo-Herraez, and W. Diaz-Villanueva. Efficient analysis and synthesis using a new factorization of the gabor frame matrix. IEEE Transac- tions on Signal Processing, 66(17):4564 -- 4573, sep 2018. doi:10.1109/tsp.2018.2855643. [19] N. Perraudin, N. Holighaus, P. L. Sø ndergaard, and P. Balazs. Designing Ga- bor windows using convex optimization. Appl. Math. Comput., 330:266 -- 287, 2018. doi:10.1016/j.amc.2018.01.035. [20] A. Ron and Z. Shen. Frames and stable bases for shift-invariant subspaces of L2(Rd). Canad. J. Math., 47(5):1051 -- 1094, 1995. doi:10.4153/CJM-1995-056-1. [21] A. Ron and Z. Shen. Weyl-Heisenberg frames and Riesz bases in L2(Rd). Duke Math. J., 89(2):237 -- 282, 1997. doi:10.1215/S0012-7094-97-08913-4. [22] W. Rudin. Principles of mathematical analysis. McGraw-Hill Book Co., New York, third edition, 1976. International Series in Pure and Applied Mathematics. date/time: 13-Feb-2019/1:32 23 of 23
1112.1960
1
1112
2011-12-08T21:02:16
Convergence of the alternating split Bregman algorithm in infinite-dimensional Hilbert spaces
[ "math.FA", "math.NA" ]
We prove results on weak convergence for the alternating split Bregman algorithm in infinite dimensional Hilbert spaces. We also show convergence of an approximate split Bregman algorithm, where errors are allowed at each step of the computation. To be able to treat the infinite dimensional case, our proofs focus mostly on the dual problem. We rely on Svaiter's theorem on weak convergence of the Douglas-Rachford splitting algorithm and on the relation between the alternating split Bregman and Douglas-Rachford splitting algorithms discovered by Setzer. Our motivation for this study is to provide a convergent algorithm for weighted least gradient problems arising in the hybrid method of imaging electric conductivity from interior knowledge (obtainable by MRI) of the magnitude of one current.
math.FA
math
Convergence of the alternating split Bregman algorithm in infinite-dimensional Hilbert spaces Amir Moradifam∗ Adrian Nachman† June 6, 2018 Abstract We prove results on weak convergence for the alternating split Bregman algorithm in infinite dimensional Hilbert spaces. We also show convergence of an approximate split Bregman algorithm, where errors are allowed at each step of the computation. To be able to treat the infinite dimensional case, our proofs focus mostly on the dual problem. We rely on Svaiter's theorem on weak convergence of the Douglas-Rachford splitting algorithm and on the relation between the alternating split Bregman and Douglas- Rachford splitting algorithms discovered by Setzer. Our motivation for this study is to provide a convergent algorithm for weighted least gradient problems arising in the hybrid method of imaging electric conductivity from interior knowledge (obtainable by MRI) of the magnitude of one current. 1 Introduction Split Bregman and alternating split Bregman algorithms were proposed by Goldstein and Osher [8] for solving problems of the form min u f (d) + g(u) subject to d = Du, where D ∈ Rm×n is a linear transform acting from Rn to Rm and f, g are convex func- tions. These algorithms have been shown to be very successful in various PDE based image restoration approaches and compressed sensing [8, 25]. In particular, the alternating split Bregman algorithm is very efficient for large scale l1-norm minimization and TV minimiza- tion problems problems [8, 25]. The convergence of the split Bregman algorithm was proved in [8]. Later in [5] and [20, 21] the authors independently proved convergence of the alter- nating split Bregman algorithm in finite dimensional Hilbert spaces. In this paper we shall ∗Department of Mathematics, University of Toronto, Toronto, Ontario, Canada M5S 2E4. E-mail: [email protected]. The author is supported by a MITACS Postdoctoral Fellowship. †Department of Mathematics and the Edward S. Rogers Sr. Department of Electrical and Computer En- gineering, University of Toronto, Toronto, Ontario, Canada M5S 2E4. E-mail: [email protected]. The author is supported in part by an NSERC Discovery Grant. 1 prove weak convergence results for the alternating split Bregman algorithm in infinite di- mensional Hilbert spaces. Our motivation for this study is to provide a convergent algorithm for weighted least gradient problems arising in the hybrid method of imaging electric con- ductivity from interior knowledge (obtainable by MRI) of the magnitude of one current. Our proof relies on a recent result of Svaiter [22] about weak convergence of the Douglas-Rachford splitting algorithm. Let H1 and H2 be real Hilbert spaces and consider the minimization problem (P ) min u∈H1 {g(u) + f (Lu)}, where L : H1 → H2 is a bounded linear operator and both functions g : H1 → R ∪ {∞} and f : H2 → R ∪ {∞} are proper, convex and lower semi-continuous. The problem (P ) can be written as a constrained minimization problem min u∈H1,d∈H2 g(u) + f (d) subject to Lu = d, which leads to an unconstrained problem: min u∈H1,d∈H2 g(u) + f (d) + λ 2 kLu − dk2. (1) (2) To solve the above problem, Goldstein and Osher [8] introduced the split Bregman method: (uk+1, dk+1) = argminu∈H1,d∈H2{g(u) + f (d) + bk+1 = bk + Luk+1 − dk+1. λ 2 k bk + Lu − d k2 2}, (3) Yin et al [24] (see also [23]) showed that the split Bregman algorithm can be viewed as an augmented Lagrangian algorithm [9, 16, 18]. Since the joint minimization problem (3) in both u and d could sometimes be hard to solve exactly, Goldstein and Osher [8] proposed the following algorithm for solving the problem (P ). Alternating split Bregman algorithm: Initialize b0 and d0. For k ≥ 1: 1. Find a minimizer uk of I k 1 (u) := g(u) + λ 2 k bk−1 + Lu − dk−1 k2 2, on H1. 2. Find the minimizer dk of on H2. k bk−1 + Luk − d k2 2, I k 2 (d) := f (d) + λ 2 2 (4) (5) 3. Let bk = bk−1 + Luk − dk. Cai, Osher, and Shen [5] proved that if the primal problem (P ) has a unique solution then the sequence uk in the above algorithm will converge to the minimizer of (P ). Independently, Setzer [20] showed that the alternating split Bregman algorithm coincides with the Douglas- Rachford splitting algorithm applied to the dual problem (D) − min b∈H2 {g ∗(−L∗b) + f ∗(b)}, and proved the convergence of the alternating split Bregman algorithm in finite-dimensional Hilbert spaces (See [21], Proposition 1 and Theorem 4). We note that for general L, the functional I k 1 (u) in the above algorithm may not have a minimizer in H1 and therefore the alternating split Bregman algorithm may not be well defined. We thus make the following definition, which will be used throughout this paper. Definition 1 We say that the alternating split Bregman algorithm is well defined if for all k ≥ 1 the functionals I k 2 (d) have minimizers in H1 and H2, respectively. 1 (u) and I k Note that I k tinuous, then I k a sufficient condition for existence of a minimizer of I k 2 (u) is strictly convex and coercive in H2. Hence if f is weakly lower semi con- 2 (d) will have a unique minimizer in H1. The following proposition provides 1 (u). Proposition 1.1 Let H1 and H2 be two Hilbert spaces, g : H1 → R ∪ {∞} be a proper, convex, and lower semi-continuous function. Assume L : H1 → H2 is a bounded linear operator. If L∗L : H1 → H1 is surjective, then for every c ∈ H2 the functional I(u) = g(u) + λ 2 k Lu + c k2 2, (6) has a unique minimizer on H1. We include a proof of Proposition 1.1 in Section 2 of this paper. We are now ready to state our main theorems. Below is a special case of a more general result (Theorem 2.6) that we will prove in Section 2. We let L∗ denote the Hermitian adjoint of L. Throughout the paper, we make the usual identification of a dual of a Hilbert space H with H itself. Theorem 1.1 Let H1 and H2 be two Hilbert spaces (possibly infinite dimensional) and as- sume that both primal (P ) and dual (D) problems have optimal solutions and that L∗L : H1 → H1 is surjective. Then the alternating split Bregman algorithm is well defined and the sequences {bk}k∈N and {dk}k∈N converge weakly to some b and d, respectively. Moreover λb is a solution of the dual problem, {dk − Luk+1}k∈N converges strongly to zero, and dk − Luk+12 2 < ∞. (7) ∞ Xk=0 Furthermore, there exists a unique u ∈ H1 such that Lu = d and u is a solution of the primal problem (P ). In particular {uk}k∈N has at most one weak cluster point ¯u = u. 3 Under additional assumptions on the functionals f and g we will also prove the following result. Theorem 1.2 Let H1 and H2 be two Hilbert spaces (possibly infinite dimensional). Suppose that alternating split Bregman is well defined, f is continuous, and both f and g are weakly lower semi continuous. Then the sequences {bk}k∈N and {dk}k∈N converge weakly to some b and d, respectively. Moreover λb is a solution of the dual problem (D), {dk − Luk+1}k∈N converges strongly to zero, and Furthermore dk − Luk+12 2 < ∞. ∞ Xk=0 lim k→∞ f (Luk) + g(uk) = min u∈H1 f (Lu) + g(u), (8) (9) every weak cluster point of {uk}k∈N is a solution of the problem (P ), and L(u) = d for every cluster point u of {uk}k∈N . In particular if (P) has a unique solution then the sequence {uk}k∈N has at most one weak cluster point which is a solution of the problem (P). Comparing Theorem 1.2 with Theorem 1.1 (and with our more general result Theorem 2.6) one can see the effect of the operator L on the convergence behaviour of the alternating split Bregman algorithm. Indeed if L is injective then Theorem 2.6 guarantees that the sequence {uk}k∈N has at most one weak cluster point without assuming uniqueness of minimizers of the problem (P), continuity of f , or weak lower semi continuity of the functionals f, g. When L is injective and the problem (P ) has more than one solution then, depending on the initial values of b0 and d0, the alternating split Bregman algorithm may converge to different solutions of the primal problem. On the other hand when L is not injective and the primal problem (P) has a unique solution then Theorem 1.2 guarantees that {uk}k∈N has at most one weak cluster point while Theorem 2.6 only says L−1( d) contains a solution of the primal problem, which is a weaker conclusion. In many applications L∗L is surjective and Theorem 1.1 guarantees the convergence of the alternating split Bregman algorithm. For instance in weighted least gradient problems H1 = H 1 0 (Ω) is surjective. Recently, our group studied the problem of recovering an isotropic conductivity from the interior measurement of the magnitude of one current density field [12, 14, 15]. We showed that the conductivity is uniquely determined by the magnitude of the current gener- ated by imposing a given boundary voltage. Moreover the corresponding voltage potential is the unique minimizer of the infinite-dimensional minimization problem 0 (Ω), H2 = (L2(Ω))n, and L = ∇u. It is easy to check that L∗ : (L2(Ω))n → H −1 u = argmin{ZΩ J∇v : v ∈ H 1(Ω), v∂Ω = f }, (10) where J is the magnitude of the current density vector field generated by imposing the voltage f on the boundary of the connected bounded region Ω ⊂ Rn, n ≥ 2. The results presented in this paper lead to a convergent split Bregman algorithm for com- puting the unique minimizer of the least gradient problem (10). The details will be presented in a forthcoming paper [13], along with a number of successful numerical experiments for recovering the electric conductivity. 4 2 Convergence of the alternating split Bregman algo- rithm Recall that, by Fenchel duality [19], the dual problem corresponding to the problem (P ) can be written as (D) − min b∈H2 {g ∗(−L∗b) + f ∗(b)}. Let v(P ) and v(D) be the optimal values of the primal (P) and dual problem (D), respectively. Weak duality always holds, that is v(P ) ≥ v(D) [1, 19]. To guarantee strong duality, i.e., the equality v(P ) = v(D) together with existence of a solution to the dual problem, several regularity conditions are available in the literature (see [1], Chapter 7). In particular, if ∃x ∈ dom(g) ∩ dom(f oL) such that f is continuous at Lx, or 0 ∈ int(Ldom(g) − dom(f )), then strong duality holds. In this paper we will always assume that both primal and dual problems have optimal solutions, and at least one of the above conditions is satisfied. Then by Rockafellar-Fenchel duality [19], if b is any solution of the dual problem, then the entire set of solutions of the primal problem is obtained as ∂g ∗(−L∗b) ∩ L−1∂f ∗(b). (11) The above representation of solutions of the primal problem is the key for our understanding of the alternating split Bregman algorithm. To explain this, we first prove the following simple lemma. Lemma 2.1 Let g : H1 → R ∪ {∞} and assume L : H1 → H2 is a bounded linear operator. Assume one of the following conditions hold: 1. There exists a point L¯u where g ∗ is continuous and finite. 2. L∗ : H2 → H1 is surjective. Then for all b ∈ H2. In particular if L−1(d) 6= ∅ for some d ∈ H2, then ∂(g ∗o(−L∗))(b) = −L∂g ∗(−L∗b), − d ∈ ∂(g ∗o(−L∗))(b) ⇔ L−1d ∩ ∂g ∗(−L∗b) 6= ∅. (12) (13) Proof. If 1) holds, then (12) follows from Proposition 5.7 in [7]. Now assume L∗ is surjective and let −L(u) = −d ∈ ∂(g ∗o(−L∗))(b). Then g ∗(−L∗(c)) − g ∗(−L∗(b)) ≥ h−Lu, c − bi = hu, −L∗(c − b)i, for all c ∈ H2. Since L∗ is surjective, u ∈ ∂g ∗(−L∗b) and consequently L−1d ⊆ ∂g ∗(−L∗b). 5 Now assume u ∈ ∂(−L∗(b)). Then g ∗(−L∗(c)) − g ∗(−L∗(b)) ≥ hu, −L∗(c − b)i = h−Lu, c − bi = h−d, c − bi, for all c ∈ H2. Hence −d ∈ ∂(g ∗o(−L∗))(b). (cid:3) By above lemma, if one can find d ∈ L(H1) such that d ∈ ∂f ∗(b) and − d ∈ ∂(g ∗o(−L∗))(b), for some solution b of the dual problem, then L−1( d) ⊆ ∂g ∗(−L∗b) ∩ L−1∂f ∗(b). (14) Consequently, by Rockafellar-Fenchel duality, every u ∈ L−1 d will be a solution of the primal problem (P ). One can find d ∈ H2 satisfying the above conditions by solving the inclusion problem (15) Indeed if b ∈ H2 is a solution of the dual problem and d ∈ ∂f ∗(b), then − d ∈ ∂(g ∗o(−L∗))(b) and (14) follows from Lemma 2.1. This is summarized in the following lemma. 0 ∈ ∂(g ∗o(−L∗))(b) + ∂f ∗(b). Lemma 2.2 Let g : H1 → R ∪ {∞} and f : H2 → R ∪ {∞} be two proper lower semi- continuous convex functions and assume that hypothesis of Lemma 2.1 hold. Also suppose both primal (P) and dual (D) problems have optimal solutions and let b be an arbitrary solution of the dual problem, then every u ∈ L−1(∂f ∗(b)) is a solution of the primal problem (P). We thus focus on computing a solution of the problem (15). This can be written in the form of an inclusion problem 0 ∈ A(b) + B(b), (16) where A := ∂(g ∗o(−L∗)) and B := ∂f ∗ are maximal monotone operators on H2. If both primal and dual problems have optimal solutions, then the above inclusion problem has at least one solution. Therefore, if we can find a solution b of the problem (16) as well as d ∈ B(b), then by Lemma 2.2 every u ∈ L−1( d) will be a solution of the primal problem (D). The Douglas-Rachford splitting method in Convex Analysis provides precisely such a pair (b, d). Following this route leads to the alternating split Bregman algorithm. Indeed it is shown by Setzer in [20] that the alternating split Bregman algorithm for the primal problem (P) coincides with Douglas-Rachford spliting algorithm applied to (16) (see Theorem 2.4). We now explain this in more detail.Let H a real Hilbert space and A, B : H → 2H be two maximal monotone operators. For a set valued function P : H → 2H, let JP to be its resolvent i.e., JP = (Id + P )−1. It is well known that sub-gradient of convex, proper, lower semi-continuous functions are maximal monotone [19] and if P is maximal monotone then JP is single valued. 6 Lions and Mercier [10] showed that for any general maximal monotone operators A, B and any initial element x0 the sequence Defined by the Douglas-Rachford recursion: xk+1 = (JA(2JB − Id) + Id − JB)xk, (17) converges weakly to some point x ∈ H such that p = JB(x) solves the inclusion problem (16). Much more recently, Svaiter [22] proved that the sequence pk = JB(xk) also converges weakly to p. Theorem 2.3 (Svaiter [22]) Let H be a Hilbert space and A, B : H → 2H be maximal monotone operators and assume that a solution of (16) exists. Then, for any initial elements x0 and p0 and any λ > 0, the sequences pk and xk generated by the following algorithm xk+1 = JλA(2pk − xk) + xk − pk pk+1 = JλB(xk+1), (18) converges weakly to some x and p respectively. Furthermore, p = JλB(x) and p satisfies 0 ∈ A(p) + B(p). To apply Douglas-Rachford splitting algorithm one needs to evaluate the resolvents JλA(2pk− xk) and JλB(xk+1) at each iteration. To evaluate the resolvents we are led to find minimizers of I k 2 (d) in the alternating split Bregman algorithm. Indeed if we let 1 (u) and I k A = ∂(g ∗o(−L∗)), B = ∂f ∗, x0 = λ(b0 + d0), and p0 = λb0. Then the resolvents JλA(2pk − xk) and JλB(xk+1) can be computed as follows and JλA(2pk − xk) = λ(bk + Luk+1 − dk), JλB(xk+1) = λ(bk + Luk+1 − dk+1), where uk+1 and dk+1 are minimizers of I k 2 (d), respectively (see [20, 21] for a proof). The following theorem gives the precise relation between the sequences generated by the alternating split Bregman algorithm and those generated by the Douglas-Rachford splitting algorithm. 1 (u) and I k Theorem 2.4 (Setzer [20]) The Alternating Split Bregman Algorithm coincides with Douglas- Rachford splitting algorithm applied to (D) with A := ∂(g ∗o(−L∗)) and B := ∂f ∗, where xk = λ(bk + dk), pk = λbk, k ≥ 0. The operator T := JA(2JB − Id) + Id − JB is known to be firmly non-expansive, T = 1 2R, with R satisfying: 2Id + 1 k Rx − Ry k≤k x − y k for all x, y ∈ H. We will need the following lemma in our convergence proof. 7 (19) i.e., Lemma 2.5 If T : H → H is a firmly non-expansive operator and xk+1 = T (xk) with x0 ∈ H, then k xk+1 − x k2 + k xk+1 − xk k2≤k xk − x k2 . Proof. Since T is firmly non-expansive, R = 2T − Id is a non-expansive operator. Hence k Rxk − Rx k2= − k xk − x k2 +2 k T xk − T x k2 −2 k (Id − T )xk − (Id − T )x k2 . Therefore we have 1 2 (k xk − x k2 − k Rxk − Rx k2) =k xk − x k2 − k xk+1 − x k2 − k xk+1 − xk k2 . Since R is non-expansive, the left hand side of the above inequality is non-negative, and this completes the proof. (cid:3) Now we are ready to prove our main theorem. Theorem 2.6 Let H1 and H2 be two Hilbert spaces (possibly infinite dimensional) and as- sume that both the primal (P ) and the dual (D) problems have optimal solutions. Suppose that (12) holds and the alternating split Bregman algorithm is well defined. Let {uk}k∈N , {dk}k∈N , and {bk}k∈N be the three sequences generated by the alternating Split Bregman algo- rithm. Then {dk}k∈N , and {bk}k∈N converge weakly to some d, and b, respectively. Moreover λb is a solution of the dual problem, the sequence {dk −Luk+1}k∈N converges strongly to zero, and ∞ dk − Luk+12 2 < ∞. (20) Xk=0 Furthermore, L−1( d) contains a solution u of the primal problem (P ). In particular if L is injective, then {uk}k∈N has at most one weak cluster point ¯u = u. Proof. The weak convergence of the sequences dk, and bk follows from Theorems 2.3 and 2.4. To prove the estimate (3.3), let T = JA(2JB −Id) + Id −JB. Since T is firmly non-expansive, by Lemma 2.5 we have k xk+1 − x k2 + k xk+1 − xk k2≤k xk − x k2, where x is the weak limit of xk with T (x) = x. By the above inequality, we have k xk+1 − xk k2< ∞. ∞ Xk=0 Now observe that xk − xk−1 = λ(bk+1 + dk+1 − bk − dk) = λ(Luk+1 − dk), and hence (3.3) follows. 8 (21) (22) By Theorem 2.4 and Theorem 2.3, p = λb is a minimizer of the dual problem and Jλ∂f ∗(λ( d+ b)) = λb. Therefore λb + λ∂f ∗(λb) = λ( d + b) ⇔ d ∈ ∂f ∗(λb) ⇔ − d ∈ ∂(g ∗o(−L∗))(λb). By Lemma 2.1 there exists u ∈ H1 such that u ∈ ∂g ∗(−L∗(p)) and L(u) = d. Therefore u ∈ ∂g ∗(−L∗(p)) ∩ L−1(∂f ∗(p)). Since p is a minimizer of the dual problem, it follows from the Fenchel-Rockafellar duality theorem that u is a minimizer of the primal problem. If L is injective then the sequence uk has at most one weak cluster point ¯u and necessarily ¯u = u. The proof is now complete. (cid:3) Proof of Theorem 1.2: Since f is continuous by Theorem 4.1 in [7] the dual problem (D) has an optimal solution. Thus it follows from an argument similar to that of Theorem 2.6 that the sequences bk and dk weakly converge to some b and d where λb is a solution of the dual problem and (8) holds. In particular if u is a weak cluster point of uk then L(u) = d. To prove (9), we can now use the argument of Cai, Osher, and Shen in the proof of Theorem 3.2 in [5] Let u be a solution of the problem (P ) and set d = Lu, p ∈ ∂f ( d), and b = p λ. Define uk e = uk − u, dk e = dk − d, bk e = bk − b. Then, as in [5], = + Xk=0 2 K Xk=0 λ (k b0 e k2 − k bK+1 e λ 2 K k2 + k d0 e k2 − k bK+1 e k2) h∂g(uk+1) − ∂g(u), uk+1 − ui + h∂f (dk+1) − ∂f ( d), dk+1 − di K Xk=0 k Luk+1 e − dk+1 e k2 + K Xk=0 k Luk+1 e − dk e k2! . (23) Now (9) follows from (23) as in the proof of Theorem 3.2 in [5]. Finally since both f and g are weakly lower semi continuous, in view of (9), every weak cluster point of {uk}k∈N is a solution of the primal (P). (cid:3) Remark 2.7 Notice that Theorem 2.3 is crucial for the proof of the convergence of the se- quences bk and dk. Proof of Proposition 1.1: First note that u is a minimizer of (6) if and only if 0 ∈ (1/λ)∂g(u) + L∗Lu + L∗c ⇔ u ∈ ((1/λ)∂g + L∗L)−1 (−L∗c). (24) Therefore to guarantee existence of a solution of (6) it is enough to prove that (1/λ)∂g + L∗L is surjective. Let A := L∗L. Since A is the subgradient of the convex lower semi-continuous 9 functional k Lu k2, it is a maximal monotone operator. We claim that A is 3∗−monotone, i.e. ∀(x, x∗) ∈ H1 × H1 sup hx − y, y ∗ − x∗i < ∞. (25) (y,y∗)∈Graph(A) Since A is surjective and (y, y ∗) ∈ Graph(A), y ∗ = A(y), and x∗ = A(x0) for some x0 ∈ H1. Hence hx − y, y ∗ − x∗i = hL(x − y), L(y − x0)i = −kL(x0 − y)k2 + hL(x − x0), L(y − x0)i ≤ −kL(x0 − y)k2 + kL(x − x0)kkL(y − x0)k ≤ kL(x − x0)k 4 < ∞. Thus (25) holds and A is 3∗−monotone. It follows from Theorem 8 in [11] and Lemma 2.2 in [3] that the operator (1/λ)∂g + L∗L is surjective and therefore (6) has a solution. Now notice that since L∗L is surjective, L∗ is surjective and hence L is injective. Conse- quently the functional I(u) is strictly convex and has a unique minimizer. The proof is now complete. (cid:3) Remark 2.8 If A = L∗L is surjective then one can show that A is also invertible. To see this note that if A is surjective then L∗ is surjective and hence L is injective. Now assume L∗L(u) = 0. Then 0 = hL∗L(u), ui = kLuk2. Therefore u = 0. Theorem 1.1 now follows from Theorem 2.6. 3 Approximate alternating split Bregman algorithm In this section we show that the alternating split Bregman algorithm is stable with respect to possible errors at each step in the calculation of minimizers of I k 2 (d). The proof relies on the following theorem about the Douglas-Rachford splitting algorithm. 1 (u) and I k Theorem 3.1 (Svaiter [22]) Let λ > 0, and let {αk}k∈N and {βk}k∈N be sequences in a and set Hilbert space H. Suppose 0 ∈ ran(A + B), and Pk∈N (k αk k + k βk k) < ∞. Take x0 ∈ H xk+1 = xk + JγA(2(JλBxk + βn) − xk) + αk − (JλBxk + βk), k ≥ 1. Then xk and pk = JλBxk converge weakly to x ∈ H and p ∈ H, respectively and p = JλB x ∈ (A + B)−1(0). 10 (26) The proof of the above theorem in infinite-dimensional Hilbert spaces is due to Svaiter [22] (see also [4]). Approximate alternating split Bregman algorithm: Initialize b0 and d0. For k ≥ 1: 1. Find uk such that where uk ex is a minimizer of k Luk − Luk ex kH2≤ αk, I k 1 (u) = {g(u) + λ 2 k bk−1 + Lu − dk−1 k2 2}, on H1. 2. Find dk such that where dk ex is the minimizer of k dk − dk ex kH2≤ βk, I k 2 (d) = argmind∈H2{f (d) + λ 2 k bk−1 + Luk − d k2 2}, on H2. 3. Let bk = bk−1 + Luk − dk. By Theorem 3.1 and an argument similar to that of Theorem 2.6 we can prove the following theorem about convergence of the sequences uk, dk, and bk produced by the above algorithm. Theorem 3.2 Let H1 and H2 be two Hilbert spaces (possibly infinite dimensional) and as- sume that both primal (P) and dual (D) problems have optimal solutions. Suppose that (12) holds, the perturbed alternating split Bregman algorithm is well defined, and (αk + βk) < ∞. ∞ Xk=1 Let {uk}k∈N , {dk}k∈N , and {bk}k∈N be the three sequences generated by the approximate alternating Split Bregman algorithm. Then {dk}k∈N , and {bk}k∈N converge weakly to some d, and b, respectively. Moreover λb is a solution of the dual problem, the sequence {dk − Luk+1}k∈N converges strongly to zero, and dk − Luk+12 2 < ∞. ∞ Xk=0 Furthermore, L−1( d) contains a solution u of the primal problem (P ). In particular if L is injective, then {uk}k∈N has at most one weak cluster point ¯u = u. 11 If L∗L is surjective, we have the following stronger result. Corollary 3.3 Let H1 and H2 be two Hilbert spaces (possibly infinite dimensional) and assume that both the primal (P) and the dual (D) problems have optimal solutions and that L∗L : H1 → H1 is surjective. Suppose that (αk + βk) < ∞. ∞ Xk=1 Let {uk}k∈N , {dk}k∈N , and {bk}k∈N be the three sequences generated by the approximate al- ternating Split Bregman algorithm, then the sequences {dk}k∈N , and {bk}k∈N converge weakly to some d, and b, respectively. Moreover λb is a solution of the dual problem, the sequence {dk − Luk+1}k∈N converges strongly to zero, and dk − Luk+12 2 < ∞. ∞ Xk=0 Furthermore there exists a unique u ∈ H1 such that Lu = d and u is a solution of the primal problem (P ). In particular, {uk}k∈N has at most one weak cluster point ¯u = u. References [1] H.H. Bauschke et al., Fixed-Point Algorithms for Inverse Problems in Science and En- gineering, Springer-Verlag, 2011. [2] H. H. Bauschke, New Demiclosedness Principles for (firmly) nonexpansive operators, arXiv:1103.0991v1 (2011). [3] H.H. Bauschke, X. Wang, and L. Yao: General resolvents for monotone operators: char- acterization and extension, Biomedical Mathematics: Promising Directions in Imaging, Therapy Planning and Inverse Problems (Huangguoshu 2008), in press. [4] P. L. Combettes, Solving monotone inclusions via compositions of nonexpansive aver- aged operators, Optimization (2004) 53(56):475-504. [5] J. F. Cai, S. Osher, Z. Shen, Split Bregman methods and frame based image restoration. Technical report (2009), UCLA Computational and Applied Mathematics. [6] J. Douglas, H. H. Rachford. On the numerical solution of heat conduction problems in two and three space variables. Trans. Americ. Math. Soc., 82(2):421-439, 1956. [7] I. Ekeland, R. T´emam, Convex analysis and variational problems, North-Holland- Elsevier, 1976. [8] T. Goldstein, S. Osher, The Split Bregman method for L1-regularized problems. SIAM Journal on Imaging Sciences (2009) 2(2):323-343. 12 [9] M.R. Hestenes, Multiplier and gradient methods. Journal of Op- timization Theory and Applications (1969) 4:303320. [10] P.-L. Lions and B. Mercier, Splitting algorithms for the sum of two nonlinear operators. SIAM J. Numer. Anal., 16(6):964-979, 1979. [11] J.-E. Mart´nez-Legaz, Some generalizations of Rockafellars surjectivity theorem, Pacific Journal of Optimization 4 (2008), pp. 527-535. [12] A. Moradifam, A. Nakhman, A. Tamasan, Conductivity imaging from one interior mea- surement in the presence of perfectly conducting and insulating inclusions, submitted (2011). [13] A. Moradifam, A. Nakhman, A. Timonov, A convergent alternating split Bregman algorithm for conductivity imaging from one interior measurement, in preparation. [14] A. Nachman, A. Tamasan, and A. Timonov, Conductivity imaging with a single measurement of boundary and interior data, Inverse Problems, 23 (2007), pp. 2551-2563. [15] A. Nachman, A. Tamasan, and A. Timonov, Recovering the conductivity from a single measurement of interior data, Inverse Problems, 25 (2009) 035014 (16pp). [16] M. J. D. Powell, A method for nonlinear constraints in minimization problems. Academic Press, London (1969) . [17] S. Osher and M. Burger, D.Goldfarb, J. Xu, and W. Yin, An Iterative Regularization Method for Total Variation-Based Image Restoration, Multiscale Model. Simul. 4 (2005) 460-489. [18] R. T. Rockafellar. Augmented Lagrangians and applications of the proximal point algo- rithm in convex programming. Mathematics of Operations Research, 1(2):97-116, 1976. [19] R.T. Rockafellar, Convex Analysis, Princeton University Press, 1996. [20] S. Setzer, Split Bregman Algorithm, Douglas-Rachford Splitting and Frame Shrinkage, Proc. of the Second International Conference on Scale Space Methods and Variational Methods in Computer Visio, Springer, 2009. [21] S. Setzer, Operator Splittings, Bregman Methods and Frame Shrinkage in Image Pro- cessing, International Journal of Computer Vision, 92(3), pp. 265-280, (2011). [22] B.F. Svaiter, On weak convergence of the Douglas-Rachford method, SIAM Journal on Control and Optimization, vol. 49, pp. 280-287, 2011. [23] X. C. Tai, C. Wu, Augmented Lagrangian method, dual methods and split Bregman iteration for ROF model. In: Lie A, Lysaker M, Morken K, Tai XC (eds) Second In- ternational Conference on Scale Space Methods and Variational Methods in Computer Vision, SSVM 2009, Voss, Norway, June 1-5, 2009. Proceedings, Springer, Lecture Notes in Computer Science, vol 5567, pp 502-513. 13 [24] W. Yin, S. Osher, D. Goldfarb, J. Darbon, Bregman iterative algorithms for l1- minimization with applications to compressed sensing. SIAM Journal on Imaging Sci- ences (2008)1(1):143-168 [25] X. Zhang, M. Burger, X. Bresson, and S. Osher, Bregmanized Nonlocal Regularization for Deconvolution and Sparse Reconstruction, 2009. UCLA CAM Report (09-03). 14
1107.1670
3
1107
2011-08-04T15:55:45
On p-Compact mappings and p-approximation
[ "math.FA" ]
The notion of $p$-compact sets arises naturally from Grothendieck's characterization of compact sets as those contained in the convex hull of a norm null sequence. The definition, due to Sinha and Karn (2002), leads to the concepts of $p$-approximation property and $p$-compact operators, which form a ideal with its ideal norm $\kappa_p$. This paper examines the interaction between the $p$-approximation property and the space of holomorphic functions. Here, the $p$-compact analytic functions play a crucial role. In order to understand this type of functions we define a $p$-compact radius of convergence which allow us to give a characterization of the functions in the class. We show that $p$-compact holomorphic functions behave more like nuclear than compact maps. We use the $\epsilon$-product, defined by Schwartz, to characterize the $p$-approximation property of a Banach space in terms of $p$-compact homogeneous polynomials and also in terms of $p$-compact holomorphic functions with range on the space. Finally, we show that $p$-compact holomorphic functions fit in the framework of holomorphy types which allows us to inspect the $\kappa_p$-approximation property. Along these notes we solve several questions posed by Aron, Maestre and Rueda in [2].
math.FA
math
ON p-COMPACT MAPPINGS AND p-APPROXIMATION SILVIA LASSALLE AND PABLO TURCO Abstract. The notion of p-compact sets arises naturally from Grothendieck's character- ization of compact sets as those contained in the convex hull of a norm null sequence. The definition, due to Sinha and Karn (2002), leads to the concepts of p-approximation property and p-compact operators, which form a ideal with its ideal norm κp. This paper examines the interaction between the p-approximation property and certain space of holo- morphic functions, the p-compact analytic functions. In order to understand these functions we define a p-compact radius of convergence which allow us to give a characterization of the functions in the class. We show that p-compact holomorphic functions behave more like nuclear than compact maps. We use the ǫ-product of Schwartz, to characterize the p- approximation property of a Banach space in terms of p-compact homogeneous polynomials and in terms of p-compact holomorphic functions with range on the space. Finally, we show that p-compact holomorphic functions fit into the framework of holomorphy types which allows us to inspect the κp-approximation property. Our approach also allows us to solve several questions posed by Aron, Maestre and Rueda in [2]. 1 1 0 2 g u A 4 ] . A F h t a m [ 3 v 0 7 6 1 . 7 0 1 1 : v i X r a Introduction In the Theory of Banach spaces (or more precisely, of infinite dimensional locally con- vex spaces), three concepts appear systematically related since the foundational articles by Grothendieck [21] and Schwartz [28]. We are referring to compact sets, compact operators and the approximation property. A Banach space E has the approximation property when- ever the identity map can be uniformly approximated by finite rank operators on compact sets. Equivalently, if E′ ⊗ E, the subspace of finite rank operators, is dense in Lc(E; E), the space of continuous linear operators considered with the uniform convergence on com- pact sets. The other classical reformulation states that E has the approximation property if F ′ ⊗ E is dense in K(F ; E), the space of compact operators, for all Banach spaces F . It was not until 1972 that Enflo provided us with the first example of a Banach space without the approximation property [19]. In the quest to a better understanding of this property and the delicate relationships inherit on different spaces of functions, important variants of the approximation property have emerged and were intensively studied. For the main develop- ments on the subject we quote the comprehensive survey [9] and the references therein. Key words and phrases. p-compact sets, holomorphic mappings, approximation properties. This project was supported in part by UBACyT X218 and UBACyT X038. 1 2 SILVIA LASSALLE AND PABLO TURCO Inspired in Grothendieck's result which characterize relatively compact sets as those con- tained in the convex hull of a norm null sequence of vectors of the space, Sinha and Karn [29] introduced the concept of relatively p-compact sets. Loosely speaking, these sets are determined by norm p-summable sequences. Associated to relatively p-compact sets we have naturally defined the notions of p-compact operators and the p-approximation property (see definitions below). Since relatively p-compact sets are, in particular, relatively compact, the p-approximation property can be seen as a way to weaken the approximation property. These three concepts were first studied by Sinha and Karn [29, 30] and, more recently, several authors continued the research on this subject [10, 12, 13, 14, 20]. This paper examines the interaction between the p-approximation property and the class of p-compact holomorphic functions. The connection between the approximation property and the space of holomorphic functions is not without precedent. The pioneer article on this topic is due to Aron and Schottenloher [5], who prove that a Banach space E has the approximation property if and only if (H(E), τ0), the space of the entire functions with the compact open topology, has the approximation property. Since then, many authors studied the approximation property for spaces of holomorphic functions in different contexts, see for instance [6, 7, 17, 18, 24]. Recently, Aron, Maestre and Rueda [2] prove that E has the p-approximation property if and only if (H(E), τ0p) has the approximation property, here τ0p denotes the topology of the uniform convergence on p-compact sets. The relation between the approximation property and holomorphic mappings was studied in detail in [5], where the class of compact holomorphic functions play a crucial role. The article is organized as follows. In the first section we fix the notation and state some basic results on p-compact mappings. In Section 2 we study the behavior of p-compact ho- mogeneous polynomials which can be considered as a polynomial Banach ideal with a natural norm denoted by κp. Following [10] we show that any p-compact homogeneous polynomial factors through a quotient of ℓ1 and express the κp-norm in terms of an infimum of certain norms of all such possible factorizations. This result slightly improves [10, Theorem 3.1.]. We also show that the Aron-Berner extension preserves the class of p-compact polynomials with the same norm. Finally we show that there is an isometric relationship between the adjoint of p-compact polynomials and quasi p-nuclear operators, improving the analogous result for operators [14]. Section 3 is devoted to the study of p-compact holomorphic mappings. Since p-compact functions are compact, we pay special attention to the results obtained by Aron and Schot- tenloher [5], where the authors prove that any holomorphic function is compact if and only if each polynomial of its Taylor series expansion at 0 is compact, [5, Proposition 3.4]. Then, Aron, Maestre and Rueda [2, Proposition 3.5] show that each component of the Taylor series expansion of a p-compact holomorphic mapping has to be p-compact. We define a natural p-compact radius of convergence and, in Proposition 3.4, we give a characterization of this type of functions. Surprisingly, we found that p-compact holomorphic functions behave more 3 like nuclear than compact mappings. We show this feature with two examples. Example 3.7 shows that Proposition 3.4 cannot be improved and also that it is possible to find an entire function whose polynomials at 0 are p-compact but the function fails to be p-compact at 0, which answers by the negative the question posed in [2, Problem 5.2]. In Example 3.8 we construct an entire function on ℓ1 which is p-compact on the open unit ball, but it fails to be p-compact at the first element of the canonical basis of ℓ1, giving an answer to [2, Problem 5.1]. We apply the results of Section 2 and 3 to study the p-approximation property in Section 4. We characterize the p-approximation property of a Banach space in terms of p-compact homogeneous polynomials with range on the space. Our proof requires the notion of the ǫ-product of Schwartz [28]. We also show that a Banach space E has the p-approximation property if and only if p-compact homogeneous polynomials with range on E can be uniformly approximated by finite rank polynomials. Then, we give the analogous result for p-compact holomorphic mappings endowed with the Nachbin topology, Proposition 4.7. The final section is dedicated to the p-compact holomorphic mappings withing the frame- work of holomorphy types, concept introduced by Nachbin [25, 26] . This allows us to inspect the κp-approximation property introduced, in [13], in the spirit of [5, Theorem 4.1]. For general background on the approximation property and its different variants we refer the reader to [9, 22]. 1. Preliminaries Throughout this paper E and F will be Banach spaces. We denote by BE the closed unit ball of E, by E′ its topological dual, and by ℓp(E) the Banach space of the p-summable sequences of elements of E, endowed with the natural norm. Also, c0(E) denotes the space of null sequences of E endowed with the supremum norm. Following [29], we say that a subset K ⊂ E is relatively p-compact, 1 ≤ p ≤ ∞, if there exists a sequence (xn)n ⊂ ℓp(E) so that K is contained in the closure of {P αnxn : (αn)n ∈ Bℓq}, where Bℓq denotes the closed unit ball of ℓq, with 1 q = 1. We denote this set by p-co{xn} and its closure by p-co{xn}. Compact sets are considered ∞-compact sets, which corresponds to q = 1. When p = 1, the 1-convex hull is obtained by considering coefficients in Bℓ∞ or, if necessary, with some extra work by coefficients in Bc0, see [14, Remark 3.3]. p + 1 Since the sequence (xn)n in the definition of a relatively p-compact set K converges to zero, any p-compact set is compact. Such a sequence is not unique, then we may consider mp(K; E) = inf{k(xn)nkp : K ⊂ p-co{xn}} which measures the size of K as a p-compact set of E. For simplicity, along this work we write mp(K) instead of mp(K; E). When K is relatively p-compact and (xn)n ⊂ ℓp(E) is a sequence so that K ⊂ p-co{xn}, any x ∈ K has the form x = P αnxn with (αn)n 4 SILVIA LASSALLE AND PABLO TURCO some sequence in Bℓq . By Holder's inequality, we have that kxk ≤ k(xn)nkℓp(E) and in consequence, kxk ≤ mp(K), for all x ∈ K. We will use without any further mention the following equalities: mp(K) = mp(K) = mp(Γ(K)), where Γ(K) denotes the absolutely convex hull of K, a relatively p-compact set. The space of linear bounded operators from E to F will be denoted by L(E; F ) and E′ ⊗ F will denote its subspace of finite rank operators. As in [29], we say that an operator T ∈ L(E; F ) is p-compact, 1 ≤ p ≤ ∞, if T (BE) is a relatively p-compact set in F . The space of p-compact operators from E to F will be denoted by Kp(E; F ). If T belongs to Kp(E; F ), we define κp(T ) = inf(cid:8)k(yn)nkp : (yn)n ∈ ℓp(F ) and T (BE) ⊂ p-co{yn}(cid:9), where κ∞ coincides with the supremum norm. It is easy to see that κp is a norm on Kp(E; F ) and following [27] (see also [14]) it is possible to show that the pair (Kp, κp) is a Banach operator ideal. The Banach ideal of p-compact operators is associated by duality with the ideal of quasi- p-nuclear operators, introduced and studied by Persson and Pietsch [27]. A linear operator T ∈ L(E; F ) is said to be quasi p-nuclear if jF T : E → ℓ∞(BF ′) is a p-nuclear operator, It is known that an where jF is the natural isometric embedding from F into ℓ∞(BF ′). operator T is quasi p-nuclear if and only if there exists a sequence (x′ n)n ⊂ ℓp(E′), such that kT xk ≤(cid:16)Xn x′ n(x)p(cid:17) 1 p , for all x ∈ E and the quasi p-nuclear norm of T is given by νQ p (T ) = inf{k(x′ n)nkp : kT xkp ≤Xn x′ n(x)p, ∀x ∈ E}. The space of quasi p-nuclear operators from E to F will be denoted by QN p(E; F ). The duality relationship is as follows. Given T ∈ L(E; F ), T is p-compact if and only if its adjoint is quasi p-nuclear. Also, T is is quasi p-nuclear if and only if its adjoint is p-compact, see [14, Corollary 3.4 ] and [14, Proposition 3.8]. A mapping P : E → F is an m-homogeneous polynomial if there exists a (unique) sym- metric m-linear form ∨ P : E × · · · × E → F such that m {z } P (x) = ∨ P (x, . . . , x), for all x ∈ E. The space of m-homogeneous continuous polynomials from E to F will be denoted by P(mE; F ), which is a Banach space considered with the supremum norm kP k = sup{kP (x)k : x ∈ BE}. 5 by LP ∈ L(Nm πs E; F ), where Nm Every P ∈ P(mE, F ) has associated two natural mappings: the linearization denoted πs E stands for the completion of the symmetric m-tensor product endowed with the symmetric projective norm. Also we have the polynomial P ∈ P(mE′′, F ′′), which is the canonical extension of P from E to E′′ obtained by weak-star density, known as the Aron-Berner extension of P [1]. While kLP k ≤ mm m! kP k, P is an isometric extension of P [11]. A mapping f : E → F is holomorphic at x0 ∈ E if there exists a sequence of polynomials Pmf (x0) ∈ P(mE, F ) such that f (x) = Pmf (x0)(x − x0), ∞Xm=0 uniformly for all x in some neighborhood of x0. We say that P∞ m=0 Pmf (x0), is the Taylor series expansion of f at x0 and that Pmf (x0) is the m-component of the series at x0. A function is said to be holomorphic or entire if it is holomorphic at x for all x ∈ E. The space of entire functions from E to F will be denote by H(E; F ). We refer the reader to [16, 23] for general background on polynomials and holomorphic functions. 2. The p-compact polynomials We want to understand the behavior of of p-compact holomorphic mappings. The defini- tion, due to Aron, Maestre and Rueda [2] was introduced as a natural extension of p-compact operators to the non linear case. In [2] the authors show that for any p-compact holomorphic function each m-homogeneous polynomial of its Taylor series expansion must be p-compact. Motivated by this fact we devote this section to the study of polynomials. 1 (yn)n ∈ ℓp(F ), (yn)n ∈ c0 if p = ∞, such that P (BE) ⊂ {P∞ Recall that P ∈ P(mE; F ) is said to be p-compact, 1 ≤ p ≤ ∞, if there exists a sequence n=1 αnyn : (αn)n ∈ Bℓq}, where p + 1 q = 1. In particular, any p-compact polynomial is compact. We denote by PKp(mE; F ) the space of p-compact m-homogeneous polynomials and by PK(mE; F ) the space of compact polynomials. Following [13], for P ∈ PKp(mE; F ) we may define κp(P ) = inf(cid:8)k(yn)nkp : (yn)n ∈ ℓp(F ) and P (BE) ⊂ p-co{yn}(cid:9). In other words, κp(P ) = mp(P (BE)). It is easy to see that κp is, in fact, a norm satisfying that kP k ≤ κp(P ), for any p-compact homogeneous polynomial P , and that (PKp(mE; F ), κp) is a polynomial Banach ideal. Furthermore, we will see that any p-compact m-homogeneous polynomial factors through a p-compact operator and a continuous m-homogeneous polyno- mial. Also, the κp-norm satisfies the natural infimum condition. Lemma 2.1. Let E and F be Banach spaces and let P ∈ P(mE; F ). The following state- ments are equivalent. 6 SILVIA LASSALLE AND PABLO TURCO (i) P is p-compact. (ii) LP : Nm Moreover, we have κp(P ) = κp(LP ). πs E → F , the linearization of P , is a p-compact operator. Proof. To show the equivalence, we appeal to the familiar diagram, where Λ is a norm one homogeneous polynomial (Λ(x) = xm) and P = LP Λ: E Λ F P LP <yyyyyyyyy Nm Note that the open unit ball of Nm πs πs E E is the absolutely convex hull Γ{xm : kxk < 1}. Then, we have that P (BE) ⊂ Γ({LP (xm) : kxk < 1}) = Γ(P (BE)). Now, the equality LP (Γ{xm : kxk < 1}) = Γ(P (BE)) shows that any sequence (yn)n ∈ ℓp(F ) involved in the definition of κp(P ) is also involved in the definition of κp(LP ) and vice versa, which completes the proof. (cid:3) Sinha and Karn [29, Theorem 3.2] show that a continuous operator is p-compact if and only if it factorizes via a quotient of ℓq and some sequence y = (yn)n ∈ ℓp(F ). Their construction is as follows. Given y = (yn)n ∈ ℓp(F ) there is a canonical continuous linear operator θy : ℓq → F associated to y such that on the unit basis of ℓq satisfies θy(en) = yn. Then, θy is extended by continuity and density. Associated to θy there is a natural injective Now, if we start with T belonging to Kp(E; F ), there exists y = (yn)n ∈ ℓp(F ) so that T (BE) ⊂ p-co{yn} and, via θy, it is possible to define an operator Ty : E → ℓq/ker θy by n=1 αnyn, which exists by the p-compactness of T . Note that the operator Ty is well defined, kTyk ≤ 1 and T operator eθy : ℓq/ker θy → F such that eθy[(αn)n] = θy((αn)n) with keθyk = kθyk ≤ kykℓp(F ). Ty(x) = [(αn)n] where (αn)n ∈ ℓq is a sequence satisfying that T (x) = P∞ satisfies the factorization T = eθyTy, where eθy is p-compact and κp(eθy) = kykℓp(F ). It is clear that, if an operator S admits such factorization, then S is p-compact. Given a sequence y = (yn)n ∈ ℓp(F ), Choi and Kim [10, Theorem 3.1] factorize the operator θy through ℓ1 as the composition of a compact and a p-compact operator, concluding that any p-compact operator T factors through a quotient space of ℓ1 as a composition of a compact mapping with a p-compact operator. The behavior of p-compact polynomials is similar to that described for p-compact opera- tors. In the proposition below a slight improvement of [10, Theorem 3.1] is obtained. We prove that the corresponding factorizations via a quotient of ℓ1 suffice to characterize κp(P ) for P a p-compact polynomial and therefore, for a p-compact operator. In order to do so, we will use the following technical lemma. Although we believe it should be a known basic / /   < result, we have not found it mentioned in the literature as we need it. Thus, we also include a proof. 7 First, we fix some notation: for σ ⊂ N a finite set, we write Sσ(r) = Pn∈σ rn. The following fact has a direct proof. Let σ1, σ2, . . . be a disjoint sequence of finite sets such that its union, ∪nσn, is infinite. Take 0 < r < t < 1 and consider the sequence β = (βn)n defined by βn = Sσn (r) Sσn (t) , for all n. Then, β belongs to Bc0. Lemma 2.2. Let 1 ≤ p < ∞. Given (xn)n ∈ ℓp and ε > 0, there exists β = (βn)n ∈ Bc0 such that ( xn βn )nkp ≤ k(xn)nkp(1 + ε). )n ∈ ℓp and k( xn βn Proof. It is enough to prove the lemma for (xn)n ∈ ℓ1 with k(xn)nk1 = 1. Indeed, suppose the result holds for this case. Fix (xn)n ∈ ℓp, a nonzero sequence and consider the sequence (zn)n defined by zn = xp , which is a norm one element of ℓ1. Given ε > 0, take (αn)n ∈ Bc0 such that k( zn αn )nk1 ≤ 1 + ε and the conclusion follows with β = (α1/p k(xn)nkp p n )n. n Now, suppose (xn)n ∈ ℓ1 and k(xn)nk1 = 1, we also may assume that xn 6= 0 for all n. Choose δ > 0 such that 1+δ 1−δ < 1 + ε. We construct β inductively. Since x1 < 1, there exists m1 ∈ N such that if σ1 = {1, 2, . . . , m1}, then x1 < Sσ1( 1 Let n1 ≥ 2 be the integer so that Pn<n1 exists 0 < tn1 ≤ 1 such that 2) and Pn1 xn < Sσ1( 1 n=1 xn ≥ Sσ1( 1 2 ). 2). Then, there Xn<n1 xn + tn1xn1 = Sσ1( 1 2) and (1 − tn1)xn1 + Xn>n1 xn = Xn>m1 2)n. ( 1 {m1 + 1, . . . , m2} we have that (1 − tn1)xn1 + xn1+1 < Sσ2( 1 Now, since (1 − tn1)xn1 + xn1+1 < Pn>m1 integer such that (1−tn1)xn1+Pn2−1 (1 − tn1)xn1 + Xn1<n<n2 Then, there exists 0 < tn2 ≤ 1 such that xn + tn2xn2 = Sσ2( 1 2) n1+1 xn < Sσ2( 1 ( 1 2)n, there exists m2 > m1 such that if σ2 = 2). Let n2 > n1 + 1 be the 2 ). n1+1 xn ≥ Sσ2( 1 2 ) and (1−tn1)xn1+Pn2 and (1 − tn2)xn2 +Xn>n2 xn = Xn>m2 2 )n. ( 1 Continuing this procedure, we can find nj, mj ∈ N, nj + 1 < nj+1, mj < mj+1 ∀j, and 0 < tnj ≤ 1 such that, if σj = {mj−1 + 1, . . . , mj}, then we obtain (1 − tnj−1)xnj−1 + Xnj−1<n<nj Now, with n0 = 0, choose β such that xn + tnj xnj = Sσj ( 1 2 ), for all j ≥ 2. β−1 n =   tnj S σj ( 1+δ 2 ) Sσj ( 1 2 ) + (1 − tnj ) Sσj+1 ( 1+δ 2 ) Sσj+1 ( 1 2 ) Sσj ( 1+δ 2 ) Sσj ( 1 2 ) if if n = nj, nj−1 < n < nj. 8 SILVIA LASSALLE AND PABLO TURCO Put cj = we have Sσj ( 1+δ 2 ) 2 ) , as we mentioned above cj ≥ 1 and cj → ∞, then (βn)n ∈ Bc0. Finally, Sσj ( 1 k( xn βn )nk1 = Xn<n1 = c1 Xn<n1 β−1 n xn + β−1 n1 xn1 +Xj≥2 Xnj−1<n<nj β−1 n xn +Xj≥2 β−1 nj xnj xn + tn1c1xn1 + (1 − tn1)c2xn1 + Xj≥2 Xnj−1<n<nj cjxn +Xj≥2 xn + tn1xn1(cid:17) +Xj≥2 2 ) = 1+δ 1−δ < 1 + ε. = c1(cid:16) Xn<n1 =P∞ j=1 Sσj ( 1+δ Thus, the lemma is proved. tnj cjxnj + (1 − tnj )cj+1xnj cjh(1 − tnj−1)xnj−1 + Xnj−1<n<nj xn + tnj xnj i (cid:3) Proposition 2.3. Let E and F be Banach spaces, 1 ≤ p < ∞, and P ∈ P(mE; F ). The following statements are equivalent. (i) P ∈ PKp(mE; F ). (ii) There exist a sequence (yn)n ∈ ℓp(F ), a polynomial Q ∈ P(mE; ℓq/ker θy) and an operator T ∈ Kp(ℓq/ker θy; F ) such that P = T ◦ Q. In this case κp(P ) = inf{kQkκp(T )}, where the infimum is taken over all the possible factorizations as above. (iii) There exist a closed subspace M ⊂ ℓ1, a polynomial Q ∈ P(mE; ℓq/ker θy) and op- erators R ∈ Kp(ℓq/ker θy; ℓ1/M) and S ∈ K(ℓ1/M; F ) such that P = SRQ. In this case κp(P ) = inf{kQkκp(R)kSk}, where the infimum is taken over all the possible factorizations as above. Proof. It is clear that either (ii) or (iii) implies (i). Now, assume that P is a p-compact polynomial. By Lemma 2.1, we have that P = LP Λ, where LP is a p-compact operator and Λ is a norm one polynomial. Applying [29, Theorem πs E) ⊆ 3.2] to Lp we have that LP = eθy(LP )y where y = (yn)n ∈ ℓp(F ) is such that LP (BNm p-co{yn}, with κp(eθy) = kykℓp(F ) and k(LP )yk ≤ 1. Then, P = T Q for T = eθy and Q = (LP )yΛ, see the diagram below. Also, we have that whenever P = T Q as in (ii), κp(P ) ≤ κp(T )kQk. Again by Lemma 2.1, κp(LP ) = κp(P ) and, since kΛk = 1, the infimum is attained. To show that (i) implies (iii), we again consider the p-compact operator LP and the fac- torization LP = eθy(LP )y. Now we modify the proof in [10, Theorem 3.1] to obtain the infimum equality. Fix ε > 0 and suppose y = (yn)n is such that k(yn)nkℓp(F ) ≤ κp(LP ) + ε. Using Lemma 2.2, choose (βn)n ∈ Bc0 such that ( yn )nkℓp ≤ βn k(yn)nkℓp(F ) + ε. )n belongs to ℓp(F ) with k( yn βn 9 s((γn)n) = Pn γnyn For simplicity, we suppose that yn 6= 0 , for all n. Now, define the operator s : ℓ1 → F as βn kynk which satisfies ksk ≤ kβk∞ ≤ 1, and consider the closed subspace M = ker s ⊂ ℓ1. Then, we may set R : ℓq/ker θy → ℓ1/M the linear operator such that R([(αn)n]) = [(αn )n] and, with S the quotient map associated to s we get the following diagram: kynk βn F P LP 9sssssssssss eKKKKKKKKKKK S eθy / ℓq/ker θy (LP )y ℓ1/M R E Λ πs E Nm Then, P = SRQ, with Q = (LP )yΛ and kSk, kQk ≤ 1. Since R(Bℓq/ker θy) ⊂ p-co{[ kynk βn by the choice of β, R is p-compact and κp(R) ≤ k( kynk βn )nkp ≤ k(yn)nkℓp(F ) + ε. en]}, Finally, κp(P ) ≤ kSkκp(R)kQk ≤ k(yn)nkℓp(F ) + ε ≤ κp(LP ) + 2ε. By Lemma 2.1, the proof is complete. (cid:3) It was shown in [14], that an operator T : E → F is p-compact if and only if its bitrans- pose T ′′ : E′′ → F ′′ is p-compact with κp(T ′′) ≤ κp(T ). In [20], it is shown that, in fact, κp(T ′′) = κp(T ) regardless T ′′ is considered as an operator on F ′′ or, thanks to the Gant- macher theorem, as an operator on F . This result, allows us to show that the Aron-Berner extension is a κp-isometric extension which preserves the ideal of p-compact homogeneous polynomials. Recall that P denotes the Aron-Berner extension of P . Proposition 2.4. Let E and F be Banach spaces, 1 ≤ p < ∞, and P ∈ P(mE; F ). Then P is p-compact if and only if P is p-compact. Moreover, κp(P ) = κp(P ). Proof. Clearly, P is p-compact whenever P is and also κp(P ) ≤ κp(P ). Now, suppose that P is p-compact. By Lemma 2.1, we can factorize P via its linearization as P = LP Λ, with kΛk = 1 and LP a p-compact operator. Since P = L′′ P Λ, by [14], P is p-compact with κp(P ) ≤ κp(L′′ P ) = κp(LP ) = κp(P ), which gives the reverse inequality. (cid:3) P ). Now, applying [20] and Lemma 2.1, κp(L′′ We finish this section by relating the transpose of p-compact polynomials with quasi p- nuclear operators. Given an homogeneous polynomial P its adjoint is defined as the linear operator P ′ : F ′ → P(mE) given by P ′(y′) = y′ ◦ P . In [20], it is shown that the transpose of a p-compact linear operator satisfies the equality κp(T ) = νQ P , where LP is the linearization of P , using Lemma 2.1 we immediately have: p (T ′). Since P ′ = L′ / /   9 / / / O O e 10 SILVIA LASSALLE AND PABLO TURCO Corollary 2.5. An homogeneous polynomial P ∈ P(mE; F ) is p-compact if and only if its transpose P ′ ∈ L(F ′; P(mE)) is quasi p-nuclear, and κp(P ) = νQ p (P ′). When this manuscript was completed we learned that R. Aron and P. Rueda were also been working on p-compact polynomials [3]. They obtained Lemma 2.1 and a non isometric version of the corollary above. 3. The p-compact holomorphic mappings In this section we undertake a detailed study of p-compact holomorphic mappings, whose definition recovers the notion of compact holomorphic mappings for p = ∞, [2]. Recall that for E and F Banach spaces, 1 ≤ p ≤ ∞, a holomorphic function f : E → F is said to be p-compact at x0 if there is a neighborhood Vx0 of x0, such that f (Vx0) ⊂ F is a relatively p-compact set. Also, f ∈ H(E; F ) is said to be p-compact if it is p-compact at x for all x ∈ E. We denote by HKp(E; F ) the space of p-compact entire functions and by HK(E; F ) the space of compact holomorphic mappings. For homogeneous polynomials, it is equivalent to be compact (p-compact) at some point of E or to be compact (p-compact) at every point of the space [2, 5]. The same property remains valid for compact holomorphic mappings [5, Proposition 3.4]. We will see that the situation is very different for p-compact holomorphic functions, 1 ≤ p < ∞. Furthermore, we will show that p-compact holomorphic mappings, 1 ≤ p < ∞, behave more like nuclear than compact holomorphic functions. Having in mind that (PKp(mE; F ), κp) is a polynomial Banach ideal with κp(P ) = mp(P (BE)), and that all polynomials in the Taylor series expansion of a p-compact holomorphic function at x0 are p-compact [2, Proposition 3.5], we propose to connect the concepts of p-compact holomorphic mappings and the size of p-compact sets measured by mp. We start with a simple but useful lemma. Lemma 3.1. Let E be a Banach space and consider K1, K2, . . . a sequence of relatively p- j=1 xj is absolutely j=1 xj : xj ∈ Kj} is relatively j=1 mp(Kj) < ∞, then the series P∞ compact sets in E, 1 ≤ p < ∞. If P∞ convergent for any choice of xj ∈ Kj and the set K = {P∞ p-compact with mp(K) ≤P∞ P∞ j=1 mp(Kj) < ∞. j=1 mp(Kj) < ∞. Proof. Note that K is well defined since for xj ∈ Kj, kxjk ≤ mp(Kj), for all j and First, suppose that p > 1 and fix ε > 0. For each j ∈ N, we may assume that Kj n : n ∈ N} with p + 1 q = 1 and n, following the standard is nonempty and we may choose (xj k(xj define the sequence (zk)k ⊂ E such that each term is of the form λjxj n)n ∈ ℓp(E) such that Kj ⊂ p-co{xj mp(Kj)−1)1/p. Now, take λj = mp(Kj)−1/q, where 1 n)nkp ≤ mp(Kj)(1 + ε 2j 11 λ1x1 2 / λ1x1 3 . . . λ1x1 1 <yyyyyyyy λ2x2 1 λ2x2 2 yttttttttt 9ttttttttt λ2x2 3 . . . λ3x3 1 λ3x3 2 λ3x3 3 . . . order: Then yyyyyyyy :uuuuuuuuu ∞Xj=1 ∞Xn=1 ∞Xj=1 ∞Xj=1 ∞Xj=1 ∞Xk=1 kzkkp = = ≤ = λp j kxj nkp mp(Kj)−p/qk(xj n)nkp ℓp(E) mp(Kj)−p/q mp(Kj)p(1 + ε 2j mp(Kj)−1) mp(Kj) + ε. j=1 mp(Kj) + ε)1/p. Hence, (zk)k belongs to ℓp(E) and k(zk)kkℓp(E) ≤ (P∞ Now, if K = {P∞ if x ∈ K, then x = P∞ xj = P∞ x =P mp(Kj)1/qαj j=1 xj : xj ∈ Kj}, we claim that K ⊂ (P∞ n. Then, x = P∞ j=1P∞ partial sums of αj nkxj nλjxj j=1 xj with xj ∈ Kj. Fix j ∈ N, there exists (αj j=1 mp(Kj))1/qp-co{zk}. Indeed, n)n ∈ Bℓq such that n and the series converges absolutely as the nk are clearly convergent with the order given above. We may write n with n=1 αj nxj n=1 αj nxj X mp(Kj)1/qαj nq = ∞Xj=1 ∞Xn=1 αj nq mp(Kj) ≤ ∞Xj=1 mp(Kj). j=1 mp(Kj))1/qp-co{zk}. It follows that K is p-compact and Then K ⊂ (P∞ mp(K) ≤ ( mp(Kj))1/qk(zk)kkℓp(E) ≤ ( ∞Xj=1 ∞Xj=1 mp(Kj))1/q( ∞Xj=1 mp(Kj) + ε)1/p. Letting ε → 0, we conclude that mp(K) ≤P∞ j=1 mp(Kj). With the usual modifications, the case p = 1 follows from the above construction consid- (cid:3) ering λj = 1, for all j. Aron, Maestre and Rueda [2, Proposition 3.5] prove that if f is a p-compact holomorphic mapping at some x0 ∈ E, every homogeneous polynomial of the Taylor series expansion of f at x0 is p-compact. At the light of the existent characterization for compact holomorphic   / y <   9 : 12 SILVIA LASSALLE AND PABLO TURCO mappings [5], they also wonder if the converse is true [2, Problem 5.2]. To tackle this question we need to define the p-compact radius of convergence of a function f at x0 ∈ E. Definition 3.2. Let E and F be Banach spaces, f ∈ H(E; F ) and x0 ∈ E. IfP∞ is the Taylor series expansion of f at x0, we say that m=0 Pmf (x0) rp(f, x0) = 1/ lim sup κp(Pmf (x0))1/m is the radius of p-compact convergence of f at x0, for 1 ≤ p < ∞. As usual, we understand that whenever lim sup κp(Pmf (x0))1/m = 0, the radius of p- compact convergence is infinite. Also, if Pmf (x0) fails to be p-compact for some m, f fails to be p-compact and rp(f, x0) = 0. The following lemma is obtained by a slight modification of the generalized Cauchy formula given in the proof of [2, Proposition 3.5], which asserts that if f ∈ H(E; F ) and x0 ∈ E, fixed ε > 0 we have that Pmf (x0)(Bε(0)) ⊂ co{f (Bε(x0))}, where Bε(x0) stands for the open ball of center x0 and radius ε. We state the result as it will be used in Section 4, also we are interested in measuring the mp-size of Pmf (x0)(V ) in terms of the mp-size of f (x0 + V ) for certain absolutely convex open sets V ⊂ E. Lemma 3.3. Let E and F be Banach spaces, let x0 ∈ E and let V ⊂ E be an abso- lutely convex open set. Let f ∈ H(E; F ) whose Taylor series expansion at x0 is given by m=0 Pmf (x0). Then P∞ (a) Pmf (x0)(V ) ⊂ co{f (x0 + V )}, for all m. (b) If f (x0 + V ) is relatively p-compact then mp(Pmf (x0)(V )) ≤ mp(f (x0 + V )), for all m. Now we are ready to give a characterization of p-compact functions in terms of the poly- nomials in its Taylor series expansion and the p-compact radius of convergence. Proposition 3.4. Let E and F be Banach spaces and let f ∈ H(E; F ) whose Taylor series m=0 Pmf (x0). For 1 ≤ p < ∞, the following statements are expansion at x0 is given by P∞ equivalent. (i) f is p-compact at x0. (ii) Pmf (x0) ∈ PKp(mE; F ), for all m and lim sup κp(Pmf (x0))1/m < ∞. and f (x) = P∞ Proof. To prove that (i) implies (ii), take ε > 0 such that f (Bε(x0)) is relatively p-compact m=1 Pmf (x0)(x − x0), with uniform convergence in Bε(x0). By [2, Proposi- tion 3.5], Pmf (x0)(εBE) ⊂ co{f (Bε(x0))} and Pmf (x0) is p-compact, for all m. Moreover, by the lemma above, κp(Pmf (x0)) = mp(Pmf (x0)(BE)) = 1 εm mp(Pmf (x0)(εBE)) ≤ 1 εm mp(co{f (Bε(x0))}). It follows that lim sup κp(Pmf (x0))1/m ≤ 1 ε , as we wanted to prove. 13 Conversely, suppose that lim sup κp(Pmf (x0))1/m = C > 0 and choose 0 < ε < rp(f, x0) m=1 Pmf (x0)(x − x0), with uniform convergence. Now we have such that, for all x ∈ Bε(x0), f (x) = P∞ ∞Xm=1 f (Bε(x0)) ⊂ { xm : xm ∈ Pmf (x0)(εBE)}. By Lemma 3.1, we obtain the result if we prove thatP∞ follows from the equality m=1 mp(Pmf (x0)(εBE)) < ∞, which ∞Xm=1 mp(Pmf (x0)(εBE)) = ∞Xm=1 εmκp(Pmf (x0)), and the choice of ε. (cid:3) Remark 3.5. Let f be a p-compact holomorphic mapping at x0 and let P∞ its Taylor series expansion at x0. Then, if ε < rp(f, x0), m=0 Pmf (x0) be mp(f (Bε(x0)) ≤ ∞Xm=1 mp(Pmf (x0)(εBE)), where the right hand series is convergent. The p-compact radius has the following natural property. Proposition 3.6. Let E and F be Banach spaces, 1 ≤ p < ∞, and f ∈ H(E; F ). Suppose that f is p-compact at x0 with positive p-compact radius r = rp(f, x0). Then f is p-compact for all x ∈ Br(x0). Also, if f is p-compact at x0 with infinite p-compact radius, then f is p-compact at x, for all x ∈ E. Proof. Without loss of generality, we can assume that x0 = 0. For r = rp(f, 0), take x ∈ E, kxk < r. By [25, Proposition 1, p.26], there exists ε > 0 such that f (y) = m=1 Pmf (0)(y) converges uniformly for all y ∈ Bε(x). We also may assume that kxk+ε < r. P∞ As in Proposition 3.4, we have that f (Bε(x)) ⊂ {P∞ if we prove that P∞ Indeed, m=1 xm : xm ∈ Pmf (0)(Bε(x))}. Now, m=1 mp(Pmf (0)(Bε(x))) < ∞, the result will follow from Lemma 3.1. ∞Xm=1 mp(Pmf (0)(Bε(x))) = ≤ = kxk+εBε(x))(cid:17) 1 (kxk + ε)m mp(cid:16)Pmf (0)( ∞Xm=1 ∞Xm=1 mp(cid:0)Pmf (0)(BE)(cid:1) ∞Xm=1(cid:16)(kxk + ε)κp(Pmf (0))1/m(cid:17)m (kxk + ε)m . Since (kxk + ε)r−1 < 1, the last series is convergent and the claim is proved. (cid:3) 14 SILVIA LASSALLE AND PABLO TURCO We recently learnt that R. Aron and P. Rueda defined, in the context of ideals of holomor- phic functions [4], a radius of I-boundedness which for p-compact holomorphic functions coincides with Definition 3.2. With the radius of I-boundedness they obtained a partial version of Proposition 3.4. Thanks to the Josefson-Nissenzweig theorem we have, for any Banach spaces E and F , a p-compact holomorphic mapping, f ∈ HKp(E; F ), whose p-compact radius of convergence at the origin is finite. It is enough to consider a sequence (x′ mk = 1 ∀m ∈ N, and (x′ m(x)m belongs to H(E), is 1-compact (hence, p-compact for p > 1) and rp(f, 0) = 1 since κp((x′ mk = 1. The example can be modified to obtain a vector valued holomorphic function with similar properties. m)m point-wise convergent to 0. Then, f (x) = P∞ m)m ⊂ E′ with kx′ m=1 x′ m)m) = kx′ There are two main questions related to p-compact holomorphic functions which where stated as Problem 5.1 and Problem 5.2 by Aron, Maestre and Rueda [2]. Both arise from properties that compact holomorphic functions satisfy. Recall that we may consider compact sets as ∞-compact sets and compact mappings as ∞-compact functions, where κ∞(P ) = kP k, for any compact m-homogeneous polynomial P . Let us consider f ∈ H(E; F ), by [5, Proposition 3.4] it is known that if f is compact at one point, say at the origin, then f is m=0 Pmf (0) is the Taylor series expansion of f at 0, and for each m the homogeneous polynomial Pmf (0) : E → F is compact, then f is compact. With Example 3.7 we show that this later result is no longer true for 1 ≤ p < ∞. Note that lim sup kPmk1/m < ∞ is fulfilled by the Cauchy's inequalities whenever f is compact. Example 3.7 also shows that, in Proposition 3.4, the hypothesis lim sup κp(Pmf (x0))1/m < ∞ cannot be ruled out. For our purposes, we adapt [15, Example 10]. compact at x for all x ∈ E. Also, if P∞ Example 3.7. For every 1 ≤ p < ∞, there exists a holomorphic function f ∈ H(ℓ1; ℓp) such that for all m ∈ N, Pmf (0) are p-compact, but f is not p-compact at 0. Furthermore, every polynomial Pmf (y) in the Taylor series expansion of f at any y ∈ ℓ1 is 1-compact, and therefore p-compact for all 1 ≤ p < ∞, but f is not p-compact at any y. Proof. Consider the partition of the natural numbers given by {σm}m, where each σm is a finite set of m! consecutive elements determined as follows: σ1 = {1}; }; σ2 = { 2, 3 {z}2! σ3 = { 4, 5, 6, 7, 8, 9 }; }; · · · σ4 = { . . .{z}4! } 3! {z Let (ej)j be the canonical basis of ℓp and denote by (e′ j)j the sequence of coordinate functionals on ℓ1. Fixed m ≥ 1, consider the polynomial Pm ∈ P(mℓ1; ℓp), defined by Pm(x) = ( mm/2 m! )1/p Xj∈σm e′ j(x)mej. Then kPmk = ( mm/2 m! )1/p sup x∈Bℓ1 k Xj∈σm e′ j(x)mejkℓp ≤ ( mm/2 m! )1/p sup x∈Bℓ1 kxk1/p 1 = ( mm/2 m! )1/p. 15 First, note that Pm is p-compact since it is of finite rank. Now, as lim sup kPmk1/m ≤ lim( m1/2 m!1/m )1/p = 0, we may define f as the series P∞ m=1 Pm, obtaining that f ∈ H(ℓ1; ℓp). In order to show that f fails to be p-compact at 0, by Proposition 3.4, it is enough to prove that lim sup κp(Pm)1/m = ∞. Fix m ∈ N and take (xn)n ∈ ℓp(ℓp), such that k ek. For each j ∈ σm, there Pm(Bℓ1) ⊂ p-co{xn}. Each xn may be written by xn = P∞ is a sequence (αj n)n ∈ Bℓq such that k=1 xn Pm(ej) = ( mm/2 m! )1/pej = ∞Xn=1 αj nxn = ∞Xn=1 ∞Xk=1 αj nxn k ek = ∞Xk=1 ( ∞Xn=1 αj nxn k )ek. Therefore, we have that ( mm/2 n=1 αj nxn j , for each j ∈ σm. Then m! )1/p =P∞ )1/p(cid:12)(cid:12)p mm/2 m! mm/2 = Xj∈σm(cid:12)(cid:12)( αj nxn j(cid:12)(cid:12)p αj nxn j )p ( ∞Xn=1 = Xj∈σm(cid:12)(cid:12) ∞Xn=1 ≤ Xj∈σm ∞Xn=1 ≤ Xj∈σm ∞Xn=1 ≤ Xj∈σm ( αj nq)p/q ∞Xn=1 xn j p xn j p ≤ k(xn)nkp ℓp(ℓp). We have shown that for any sequence (xn)n ∈ ℓp(ℓp) such that Pm(Bℓ1) ⊂ p-co{xn}, the inequality k(xn)nkℓp(ℓp) ≥ mm/2p holds. Then, κp(Pm) ≥ mm/2p for all m ∈ N. Hence, we conclude that lim sup κp(Pm)1/m = ∞ and, by Proposition 3.4, f cannot be p-compact at 0, which proves the first statement of the example. To show the second assertion, take any nonzero element y ∈ ℓ1 and fix m0 ∈ N. For all x ∈ Bℓ1, Pm0f (y)(x) = = ∞Xm=m0 ∞Xm=m0 m0(cid:19) ∨ (cid:18) m m0(cid:19)(cid:18)mm/2 (cid:18) m m! (cid:19)1/p Xj∈σm P m(ym−m0, xm0) e′ j(y)m−m0e′ j(x)m0ej. 16 SILVIA LASSALLE AND PABLO TURCO We claim that the sequence (cid:18)(cid:0) m m! (cid:19)1/p Xj∈σm Xm>m0 m0(cid:19)(cid:18)mm/2 (cid:18) m m! (cid:17)1/p m0(cid:1)(cid:16) mm/2 j(y)m−m0 ≤ Xm>m0 j(y)m−m0ej(cid:19) j∈σm m0(cid:19)(cid:18) mm/2 (cid:18) m m>m0 e′ e′ m! (cid:19)1/p belongs to ℓ1(ℓp). In fact, kykm−m0 1 < ∞. Then, since (e′ j(x)m) j∈σm m≥m0 hull of (cid:26)(cid:0) m m0(cid:1)(cid:16) mm/2 m! (cid:17)1/p belongs to Bc0, the set Pm0f (y)(Bℓ1) is included in the 1-convex j(y)m−m0ej : m ≥ m0, j ∈ σm(cid:27), which proves that Pm0f (y) is 1- e′ compact and, therefore, p-compact for every 1 ≤ p, for any m0. To show that f is not p-compact at y, note that fixed m, it is enough to choose j ∈ σm, to obtain that Pmf (y)(ej) = (cid:16) mm/2 ej. Now, we can proceed as in the first part of the example to show that lim sup κp(Pmf (y))1/m = ∞. And, again by Proposition 3.4, we have that f cannot be p-compact at y. (cid:3) m! (cid:17)1/p The following example gives a negative answer to [2, Problem 5.1]. We show an entire func- tion which is p-compact at 0, but this property does not extend beyond the ball Brp(f,0)(0). Example 3.8 proves, in addition, that Proposition 3.6 cannot be improved. We base our construction in [15, Example 11]. Example 3.8. For every 1 ≤ p < ∞, there exists a holomorphic function f ∈ H(ℓ1; ℓp) such that f is p-compact at 0, with lim sup κp(Pmf (0))1/m = 1, but f is not p-compact at e1. Proof. Consider {σm}m, the partition of the natural numbers, as in Example 3.7. Let (ej)j be the canonical basis of ℓp and denote (e′ j)j the sequence of coordinate functionals on ℓ1. Fixed m ≥ 2, define Pm ∈ P(mℓ1; ℓp), the m-homogeneous polynomial Then kPmk = ( ≤ ( 1 m! 1 m! )1/p sup x∈Bℓ1 )1/p sup x∈Bℓ1 Pm(x) = ( 1 m! )1/pe′ e′ j(x)2ej. 1(x)m−2 Xj∈σm (Xj∈σm (Xj∈σm e′ e′ 1(x)m−2e′ j(x)2p)1/p j(x)2p)1/p ≤ ( 1 m! )1/p. Since lim kPmk1/m ≤ lim( 1 belongs to H(ℓ1; ℓp) and Pm≥2 Pm is its Taylor series expansion at 0. computing kPmk, we showed that α(x) = (cid:0)e′ m! )1/pm = 0, we may define f as f (x) = Pm≥2 Pm(x), which j(x)2(cid:1)j ∈ Bℓq for all x ∈ Bℓ1. Then m! )1/pp-co{ej : j ∈ σm} and since k(ej)j∈σmkℓp(ℓp) = (Pj∈σm 1)1/p = (m!)1/p, Note that each Pm is p-compact, as it is of finite rank, for all m ≥ 2. Moreover, when Pm(Bℓ1) ⊂ ( 1 1(x)m−2e′ we have that κp(Pm) ≤ ( 1 tion 3.4, we have that f is p-compact at 0. m! )1/p(m!)1/p = 1. Then, lim sup κp(Pm)1/m ≤ 1 and, by Proposi- 17 To show that rp(f, 0) = 1, fix m ≥ 2 and ε > 0. Take xj ∈ Bℓ1 such that e′ e′ j(xj) = ε and e′ k(xj) = 0 for j ∈ σm and k 6= j. 1(xj) = 1 − ε, P∞ Now, fix any sequence (yn)n ∈ ℓp(ℓp) such that Pm(Bℓ1) ⊂ p-co{yn} and write yn = k=1 yn Then, for each j ∈ σm there exists (αj n)n ∈ Bℓq so that k ek. Pm(xj) = ( 1 m! nyn. αj )1/p(1 − ε)m−2ε2ej = ∞Xn=1 m! )1/p(1 − ε)m−2ε2 =P∞ Thus, for each j ∈ σm, the equality ( 1 n=1 αj nyn j holds. In consequence ((1 − ε)m−2ε2)p = Xj∈σm = Xj∈σm = Xj∈σm ≤ Xj∈σm ≤ Xj∈σm 1 m! 1 m! ( ∞Xn=1 ∞Xn=1 ∞Xn=1 ( ((1 − ε)m−2ε2)p )1/p(1 − ε)m−2ε2p αj nyn j p αj nyn j )p yn j p ≤ k(yn)nkp ℓp(ℓp). Finally, we get that κp(Pm) ≥ (1 − ε)m−2ε2 which implies that lim sup κp(Pm)1/m ≥ 1 − ε. Since ε > 0 was arbitrary, we obtain that rp(f, 0) = 1. Now, we want to prove that f is not p-compact at e1. By Proposition 3.4, it is enough to show, for instance, that the 2-homogeneous polynomial P2f (e1) : ℓ1 → ℓp is not p-compact. We have (1) P2f (e1)(x) = ∞Xm=2 2(cid:19) ∨ (cid:18)m Pm(em−2 1 , x2) where ∨ Pm is the symmetric m-linear mapping associated to Pm. By the definition of Pm we easily obtain a multilinear mapping Am ∈ L(mℓ1; ℓp) satisfying Pm(x) = Am(x, . . . , x), defined by Am(x1, . . . , xm) = ( 1 m! )1/pe′ 1(x1) · · · e′ 1(xm−2) Xj∈σm e′ j(xm−1)e′ j(xm)ej. 18 SILVIA LASSALLE AND PABLO TURCO Let Sm be the symmetric group on {1, . . . , m} and denote, for each ξ ∈ Sm, the multilinear mapping Aξ m(x1, . . . , xm) = Am(xξ(1), . . . , xξ(m)). Then we have m given by Aξ ∨ Pm(em−2 1 , x2) = 1 m! Xξ∈Sm Aξ m(em−2 1 , x2). Since Am(x1, . . . , xm−2, e1, xm−1) = Am(x1, . . . , xm−1, e1) = 0, for all x1, . . . , xm−1 ∈ ℓ1, and Am(em−2 j(x)2ej, we obtain 1 , x2) =(cid:0) 1 m!(cid:1)1/pPj∈σm e′ ∨ Pm(em−2 1 , x2) = (2) 1 m! 2(m − 2)!( 1 m! )1/p Xj∈σm e′ j(x)2ej. Combining (1) and (2) we get that P2f (e1)(x) = Xm≥2 ( 1 m! )1/p Xj∈σm e′ j(x)2ej. Suppose that P2f (e1) is p-compact. Hence, there exists a sequence (yn)n ∈ ℓp(ℓp), yn = n)n ∈ Bℓq nyn. As in the Example 3.7, we conclude that k ek such that P2f (e1)(Bℓ1) ⊂ p-co{yn}. For each j ∈ σm, there exists (αj n=1 αj such that P2f (e1)(ej) = ( 1 ( 1 n=1 αj nyn j , if j ∈ σm. m! )1/pej = P∞ k=1 yn P∞ m! )1/p =P∞ Hence Xm≥2 Xj∈σm (( 1 m! )1/p)p = Xm≥2 Xj∈σm ≤ Xm≥2 Xj∈σm ≤ Xm≥2 Xj∈σm ( ∞Xn=1 ∞Xn=1 ∞Xn=1 αj nyn j p αj nq)p/q yj np ≤ k(yn)nkp ℓp(ℓp) < ∞, ∞Xn=1 yj np which is a contradiction sincePm≥2Pj∈σm(( 1 be p-compact at e1, and the result is proved. m! )1/p)p is not convergent. Therefore, f cannot (cid:3) 4. The p-approximation property and p-compact mappings The concept of p-compact sets leads naturally to that of p-approximation property. A Banach space E has the p-approximation property if the identity can be uniformly approxi- mated by finite rank operators on p-compact sets. Since p-compact sets are compact, every space with the approximation property has the p-approximation property. Then, this prop- erty can be seen as a way to weaken the classical approximation property. The p-approximation property has been studied in [10, 12] related with p-compact linear operators and in [2] related with non linear mappings. The relation between the approxi- mation property and compact holomorphic mappings was first addressed in [5]. Here, we 19 are concern with the study of the p-approximation property and its relation with p-compact polynomials and holomorphic functions in the spirit of [2] and [5]. We start characterizing the notion of a homogeneous polynomial P being p-compact in terms of different conditions of continuity satisfied by P ′ the transpose of P . The first proposition gives an answer to [2, Problem 5.8] and should be compared with [5, Proposition 3.2]. Before going on, some words are needed on the topologies which we will use. We denote by Pc(mE) the space P(mE) considered with the uniform convergence on compact sets of E, if m = 1 we simply write E′ c. When compact sets are replaced by p-compact sets we use the notation Pcp(mE) and E′ cp. By the Ascoli theorem, any set L ⊂ Pc(mE) is relatively compact if and only if supP ∈L kP k is finite. Also, if L ⊂ Pcp(mE) is relatively compact we have that L is point-wise bounded and then, by the Principle of uniform boundedness, L is relatively compact in Pc(mE). Now we have: Proposition 4.1. Let E and F be Banach spaces, 1 ≤ p < ∞, and P ∈ P(mE; F ). The following statements are equivalent. (i) P ∈ PKp(E; F ). (ii) P ′ : F ′ (iii) P ′ : F ′ (iv) P ′ : F ′ (v) P ′ : F ′ cp → P(mE) is continuous. cp → Pc(mE) is compact. cp → Pcq(mE) is compact for any q, 1 ≤ q < ∞. cp → Pcq(mE) is compact for some q, 1 ≤ q < ∞. Proof. Suppose (i) holds, then P (BE) = K is p-compact and its polar set K ◦ is a neighbor- hood in F ′ cp → P(mE) is continuous. cp. For y′ ∈ K ◦ we have that kP ′(y′)k = supx∈BE y′(P x) ≤ 1, and P ′ : F ′ Now suppose (ii) holds, then there exists a p-compact set K ⊂ F such that P ′(K ◦) is equicontinuous in P(mE). By the Ascoli theorem, P ′(K ◦) is relatively compact in Pc(mE) and P ′ : F ′ cp → Pc(mE) is compact. The continuity of the identity map Pc(mE) ֒→ Pcq(mE) gives that (iii) implies (iv), for all 1 ≤ q < ∞. Obviously, (iv) implies (v). To complete the proof, suppose (v) holds. Then, there exist an absolutely convex p-compact set K ⊂ F and a compact set L ⊂ Pcq(mE) such that P ′(K ◦) ⊂ L and therefore, there exists c > 0 such that supy′∈K ◦ kP ′(y′)k ≤ c. Note that for any x ∈ c− 1 n BE and y′ ∈ K ◦ we have that P ′(y′)(x) = y′(P x) ≤ 1. Then P (x) ∈ K, for all x ∈ c− 1 (cid:3) n BE and P is p-compact. Now, we characterize the p-approximation property on a Banach space in terms of the p-compact homogeneous polynomials with values on it. In order to do so we appeal to the notion of the ǫ-product introduced by Schwartz [28]. Recall that for E and F two locally convex spaces, F ǫE is defined as the space of all linear continuous operators from E′ c to F , 20 SILVIA LASSALLE AND PABLO TURCO endowed with the topology of uniform convergence on all equicontinuous sets of E′. The space F ǫE is also denoted by Lǫ(E′ c; F ). In [5, Proposition 3.3] its shown, for all Banach spaces E and F , that (P(mF ), k.k)ǫE = Lǫ(E′ c; (P(mF ), k.k)) = (PK(mF ; E), k.k), where the isomorphism is given by the transposition P ↔ P ′. As a consequence, it is proved that P(mF ) has the approximation property if and only if P(mF ) ⊗ E is k.k-dense in PK(mF ; E) for all Banach spaces E and all m ∈ N. We have the following result. Proposition 4.2. Let E and F be Banach spaces. Then (PKp(mF ; E), k.k) is isometrically isomorphic to Lǫ(E′ cp; (P(mF ), k.k)). As a consequence, E has the p-approximation property if and only if P(mF )⊗E is k.k-dense in PKp(mF ; E) for all Banach spaces F and all m ∈ N. Proof. Note that [(i) implies (ii)] of Proposition 4.1, says that the transposition operator maps a p-compact polynomial into a linear map in Lǫ(E′ cp; P(mF ), k.k)). Now, take T an cp; P(mF ), k.k)). Since the identity map ι : E′ operator in Lǫ(E′ cp is continuous, T c; P(mF ), k.k)). By [5, Proposition 3.3], we have that T = P ′ for some belongs to Lǫ(E′ P ∈ PK(mF ; E). In particular, P ′ : E′ cp → P(mF ) is continuous and by [(ii) implies (i)] of Proposition 4.1, P is p-compact. c → E′ For the second statement, cp; G), for every locally convex space G, [20]. if E has the p-approximation property, G ⊗ E is dense in Lǫ(E′ In particular we may consider G = (P(mF ), k.k). Conversely, with m = 1 we have that F ′ ⊗ E is k.k-dense in Kp(F ; E) for every Banach space F . Now, an application of [12, Theorem 2.1] completes the proof. (cid:3) At the light of [5, Proposition 3.3], we expected to obtain a result of the type P(mE) has the p-approximation property if and only if P(mE) ⊗ F is k.k-dense in PKp(mE; F ) for all Banach spaces F and all m ∈ N. Unfortunately, our characterization is not as direct as we wanted and requires the following notion. Definition 4.3. Let E be a Banach space, A an operator ideal and α a norm on A . We say that E has the (A, α)-approximation property if F ′ ⊗ E is α-dense in A(F, E), for all Banach spaces F . The relation between an ideal A with the ideal of those operators whose transpose belongs to A leads us to work with the ideal of quasi p-nuclear operators QN p. Proposition 4.4. Let E be a Banach space and fix m ∈ N. Then, (a) P(mE)⊗F is k.k-dense in PKp(mE; F ), for all Banach spaces F if and only if P(mE) has the (QN p, k.k)-approximation property. (b) P(mE) has the p-approximation property if and only if P(mE) ⊗ F is k.k-dense in {P ∈ P(E; F ) : Lp ∈ QN p(⊗m πsE; F )}, for all Banach spaces F . 21 πsE)′ ⊗ F is k.k-dense in Kp(⊗m πsE)′ , has the (QN p, k.k)-approximation prop- Proof. The space P(mE), or equivalently (⊗m erty if and only if (⊗m πsE; F ) for all Banach spaces F , see In virtue of Lemma 2.1, it is equivalent to have that P(mE) ⊗ F is k.k-dense in [20]. PKp(mE; F ). Then, statement (a) is proved. Note that (a) can be reformulated saying that P(mE) has the (QN p, k.k)-approximation property if and only if P(mE) ⊗ F is k.k-dense in {P ∈ P(E; F ) : Lp ∈ Kp(⊗m πsE; F )}, for all Banach spaces F . For the proof of (b), we use that the p-approximation property corresponds to the (A, k.k)- approximation property for the ideal A = Kp, of p-compact operators. The result follows proceeding as before if the ideal Kp and its dual ideal QN p are interchanged. (cid:3) Now, we change our study to that of p-compact holomorphic mappings. Aron and Schot- tenloher described the space of compact holomorphic functions considered with τw, the Nach- bin topology [25], via the ǫ-product. Namely, they show that (HK(E; F ), τω) = Lε(F ′ where the isomorphism is given by the transposition map f 7→ f ′ [5, Theorem 4.1]. The authors use this equivalence to obtain, in presence of the approximation property, results on density similar to that of Proposition 4.2. Recall that f ′ : F ′ → H(E) denotes the linear operator given by f ′(y′) = y′ ◦ f . With the next proposition we try to clarify the relationship between p-compact holomorphic mappings and the ǫ-product. The result obtained gives, somehow, a partial answer to [2, Problem 5.6]. c; H(E), τω), Proposition 4.5. Let E and F be Banach spaces. Then, (a) (HKp(E; F ), τω) is topologically isomorphic to a subspace of Lǫ(F ′ (b) Lǫ(F ′ cp; (H(E), τω)) is topologically isomorphic to a subspace of(cid:8)f ∈ H(E; F ) : Pmf (x) ∈ PKp(mE; F ), ∀x ∈ E, ∀m ∈ N(cid:9), considered with the Nachbin topology, τω. cp; (H(E), τω)). Proof. To prove (a), fix f in HKp(E; F ) and consider q any τω-continuous seminorm on H(E). By [16, Proposition 3.47], we may consider only the seminorms such that, for g ∈ H(E), q(g) = ∞Xm=0 kPmg(0)kK+amBE , with K ⊂ E an absolutely convex compact set and (am)m a sequence in c+ 0 . There exists V ⊂ E, an open set such that 2K ⊂ V and f (V ) ⊂ F is p-compact. Fix m0 ∈ N such that 2K + 2amBE ⊂ V , for all m ≥ m0. Now, choose c > 0 such that c(2K + 2amBE) ⊂ 2K + 2am0BE ⊂ V , for all m < m0. The polar set of f (V ), f (V )◦, is a neighborhood in F ′ cp. By the Cauchy inequalities for entire functions, we have for all y′ ∈ f (V )◦, 22 SILVIA LASSALLE AND PABLO TURCO q(f ′(y′)) = kPm (y′ ◦ f ) (0)kK+amBE = ∞Xm=0 ∞Xm=0 = Xm<m0 ≤ Xm<m0 ≤ Xm<m0 ≤ Xm<m0 ≤ Xm<m0 1 1 1 1 2m kPm (y′ ◦ f ) (0)k2K+2amBE 2m kPm (y′ ◦ f ) (0)k2K+2amBE + Xm≥m0 (2c)m kPm (y′ ◦ f ) (0)kc(2K+2amBE ) + Xm≥m0 (2c)m ky′ ◦ f kc(2K+2amBE) + Xm≥m0 (2c)m ky′ ◦ f kV + Xm≥m0 1 2m ky′ ◦ f kV (2c)m + Xm≥m0 1 2m < ∞. 1 1 1 2m kPm (y′ ◦ f ) (0)k2K+2amBE 1 2m kPm (y′ ◦ f ) (0)k2K+2amBE 1 2m ky′ ◦ f k2K+2amBE Then f ′ ∈ L(F ′ cp; (H(E), τω)). Again, we use the continuity of the identity map ι : F ′ c → F ′ cp now, [5, Theorem 4.1] implies the result. To prove that (b) holds, take T ∈ L(F ′ cp; (H(E), τω)) which, in particular, is an operator c; (H(E), τω)). By [5, Theorem 4.1], T = f ′ for some f ∈ HK(E; F ). By virtue of cp → (P(mE), k.k) is continuous, for x : (H(E), τω) → (P(mE), k.k) the continuous projection given by x ◦ f ′ coincide as linear (cid:3) in L(F ′ Proposition 4.1, it is enough to show that (Pmf (x))′ : F ′ each m ∈ N. Consider Dm Dm operators. Hence, we obtain the result. x (g) = Pmg(x), for all g ∈ H(E). Note that (Pmf (x))′ and Dm Example 3.7 shows that there exists an entire function f : ℓ1 → ℓp, so that every homo- geneous polynomial in its Taylor series expansion at y is q-compact for any y ∈ ℓ1, for all 1 ≤ q < ∞, but f fails to be q-compact at y, for every y and every q ≤ p. However, we have the following result. Lemma 4.6. Let E and F be Banach spaces. Then, HKp(E; F ) is τω-dense in {f ∈ H(E; F ) : Pmf (x) ∈ PKp(mE; F ), ∀x ∈ E, ∀m ∈ N}. Proof. Fix f ∈ H(E; F ) so that Pmf (x) ∈ PKp(mE; F ) for all x ∈ E and for all m. Let ε > 0 and let q be any τω-continuous seminorm on H(E; F ) of the form q(g) = ∞Xm=0 kPmg(0)kK+amBE , with K ⊂ E absolutely convex and compact and (am)m ∈ c+ Pm≥m0 kPmf (0)kK+amBE < ε. Now, let f0 = Pm<m0 that q(f − f0) ≤ ε and the lemma follows. 0 . Consider m0 ∈ N such that Pmf (0), which is p-compact. Note (cid:3) Proposition 4.7. Let E be a Banach space. Then, the following statements are equivalent. 23 (i) E has the p-approximation property. (ii) H(F ) ⊗ E is τω-dense in HKp(F ; E) for all Banach spaces F . Proof. If E has the p-approximation property, E ⊗G is dense in Lǫ(E′ cp; G) for all locally con- vex space G [20], in particular if we consider G = (H(F ), τω). Applying Proposition 4.5 (a), we have the first assertion. For the converse, put H0 = {f ∈ H(F ; E) : Pmf (x) ∈ PKp(mF ; E), ∀x ∈ E, ∀m ∈ N}. By Lemma 4.6, H(F ) ⊗ E is τω-dense in H0. Now, take T ∈ Kp(F ; E) and ε > 0. Since T ∈ H0 and q(f ) = kP1f (0)k is a τω-continuous seminorm, there exists g ∈ H(F ) ⊗ E such that q(T − g) ≤ ε. But q(T − g) = kT − P1g(0)k and since P1g(0) ∈ F ′ ⊗ E, we have shown that F ′ ⊗ E is k.k-dense in Kp(F ; E). By [12, Theorem 2.1], E has the p-approximation property. (cid:3) 5. Holomorphy types and topologies In this section we show that p-compact holomorphic functions fit into the framework of holomorphy types. Our notation and terminology follow that given in [15]. Since, PKp(mE; F ) is a subspace of P(mE; F ) and PKp(0E; F ) = F , the first two conditions in the definition of a holomorphy type are fulfilled. Therefore, we only need to corroborate that the sequence (PKp(mE; F ), κp)m satisfies the third condition. Indeed, this last condi- tion will be also fulfilled if we show (3) κp(Pl(P )(a)) ≤ (2e)mκp(P )kakm−l, for every P ∈ PKp(mE; F ), for all l = 1, . . . , m and for all m, where Pl(P )(a) denotes the l-component in the expansion of P at a. A function f ∈ H(E; F ) is said to be of holomorphic type κp, at a, if there exist c1, c2 > 0 such that each component of its Taylor series expansion, at a, is a p-compact polynomial satisfying that κp(Pmf (a)) ≤ c1cm 2 . To give a simple proof of the fact that (PKp(mE; F ), κp)m satisfy the inequalities given in (3) we use the following notation. Let P ∈ P(mE; F ) and fix a ∈ E, we denote by Pal the (m − l)-homogeneous polynomial defined as Pal(x) : = ∨ P (al, xm−l), for all x ∈ E and l < m. Note that, for any j < l < m, we have that Pal = (Pal−j )aj and that Pl(P )(a) =(cid:0) m m−l(cid:1)Pam−l. We appeal to the description of Pa given in [8, Corollary 1.8, b)]: (4) Pa(x) = ∨ P (a, xm−1) = 1 m2 P ((m − 1)rjx + a). 1 (m − 1)m−1 m−1Xj=1 24 SILVIA LASSALLE AND PABLO TURCO where r ∈ C is such that rm = 1 and rj 6= 1 for j < m. Theorem 5.1. For any Banach spaces E and F , the sequence (PKp(mE; F ), κp)m is a holo- morphy type from E to F . Proof. If P ∈ PKp(mE; F ) by [2, Proposition 3.5] or Proposition 3.4 we have that Pj(P )(a) ∈ PKp(jE; F ) for all a ∈ E, for all j ≤ m. To prove the holomorphy type structure, we will show that κp(Pj(P )(a)) ≤ 2memkakm−jκp(P ), for all j ≤ m. Fix a ∈ E. If we show that κp(Pa) ≤ ekakκp(P ) then the proof is complete using a generalized inductive reasoning. Indeed, suppose that for any p-compact homogeneous poly- nomial Q, of degree less than m, the inequality κp(Qa) ≤ ekakκp(Q) holds. Then, since Pal = (Pal−1)a and Pj(P )(a) =(cid:0) m m−j(cid:1)Pam−j , we obtain m−j(cid:1)κp(Pam−j ) =(cid:0) m κp(Pj(P )(a)) =(cid:0) m ≤(cid:0) m m−j(cid:1)ekakκp((Pam−j−1)) ≤(cid:0) m m−j(cid:1)em−jkakm−jκp(P ) ≤ 2memkakm−jκp(P ). m−j(cid:1)κp((Pam−j−1)a) Now, take P ∈ PKp(mE; F ). Then (5) κp(Pa) = mp( ∨ P (a, Bm−1 E )) = kakmp( Using (4) and Lemma 3.1 we have ∨ P ( a kak , Bm−1 E )). (6) kakmp( ∨ P ( a kak , Bm−1 E )) ≤ kak 1 m2 1 (m−1)m−1 m−1Xj=1 mp(P ((m − 1)rjBE + a kak )). Since sup{kxk : x ∈ (m − 1)rjBE + a kak } = m, ∨ P ( a kak , Bm−1 E kakmp( (7) )) ≤ kak m2(m−1)m−1 = kak m2(m−1)m−1 m−1Xj=1 m−1Xj=1 ≤ kak( m m−1)m−1κp(P ) ≤ ekakκp(P ). mp(P ((m − 1)rjBE + a kak )) mmmp(cid:0)P(cid:0) 1 m((m − 1)rjBE + a kak )(cid:1)(cid:1) Combining (5), (6) and (7) we get that κp(Pa) ≤ ekakκp(P ), as we wanted to show. (cid:3) As a consequence we have the following result. Corollary 5.2. Let f be a function in H(E; F ), then f ∈ HKp(E; F ) if and only if f is of κp-holomorphy type. Proof. It follows from Theorem 5.1 and [2, Proposition 3.5] or Proposition 3.4. (cid:3) 25 Remark 5.3. Theorem 5.1 can be improved. Indeed, the same proof of Theorem 5.1 shows that the sequence (PKp(mE; F ))m is a coherent sequence associated to the operator ideal Kp(E; F ) (see [8] for definitions). Since HKp(E, F ) is a holomorphy type, following [26] we have a natural topology defined on HKp(E, F ) denoted by τω,mp. This topology may be generated by different families of continuous seminorms. The original set of seminorms used to define τω,mp corresponds to the family of seminorms given below in Theorem 5.5, item (c). Our aim is to to characterize the κp-approximation property of a Banach space E in an analogous way to [5, Theorem 4.1]. In order to do so, we will give different descriptions of τω,mp. First, we need the following result. Proposition 5.4. Let E and F be Banach spaces. Then, f ∈ HKp(E; F ) if and only if, for all m, Pmf (0) ∈ PKp(mE; F ) and for any absolutely convex compact set K, there exists ε > 0 such that P∞ n=0 mp(Pmf (0)(K + εBE)) < ∞. Proof. Take f ∈ HKp(E; F ) and K an absolutely convex compact set. Then, 2K is also absolutely convex and compact. For each x ∈ 2K, there exist εx > 0 such that f (x + εxBE) j=1(xj + εxj BE) and with j=1(xj + εxj BE) we have that f (V ) is p-compact. Let d = dist(2K, CV ) > 0, where CV denotes the complement of V . Let us consider W = 2K + dBE, then W is an absolutely convex open set and 2K ⊂ W ⊂ V . Then, applying Proposition 3.3 we have is p-compact. Now, we choose x1, . . . , xn ∈ 2K such that K ⊂ Sn V = Sn ∞Xn=0 mp(Pmf (0)(K + d/2BE) = ∞Xn=0 which proves the first claim. (1/2)m mp(Pmf (0)(W )) ≤ 2mp(f (W )) < ∞, Conversely, let f ∈ H(E; F ) satisfy the conditions in the proposition. We have to show that f is p-compact at x for any fixed x ∈ E. Consider the absolutely convex compact set n=0 mp(Pmf (0)(K + ε1BE)) < ∞. Since f is an entire function, by [25, Proposition 1, p.26], there exists ε2 > 0 m=1 Pmf (0)(y) uniformly for y ∈ Bε2(x). Let ε = min{ε1; ε2}, then K, given by K = {λx : λ ≤ 1}. Then, there exists ε1 > 0 such that P∞ such that f (y) = P∞ f (Bε(x)) ⊂ {P∞ ∞Xm=0 m=0 xm : xm ∈ Pmf (0)(Bε(x))}. mp(Pmf (0)(Bε(x)) ≤ mp(Pmf (0)(K + ε1BE)) < ∞. ∞Xm=0 Also Now, applying Lemma 3.1 we obtain that f is p-compact at x, and the proof is complete. (cid:3) The next characterization of the topology τω,mp associated to the holomorphy type HKp(E; F ) follows that of [15] and [25]. Theorem 5.5. Let E and F be Banach spaces and consider the space HKp(E; F ). Any of the following families of seminorms generate the topology τω,mp. 26 SILVIA LASSALLE AND PABLO TURCO (a) The seminorms p satisfying that there exists a compact set K such that for every open set V ⊃ K there exists CV > 0 so that p(f ) ≤ CV mp(f (V )) ∀f ∈ HKp(E; F ). In this case, we say that p is mp-ported by compact sets. (b) The seminorms p satisfying that there exists an absolutely convex compact set K such that for every absolutely convex open set V ⊃ K there exists CV > 0 so that p(f ) ≤ CV mp(f (V )) ∀f ∈ HKp(E; F ). In this case, we say that p is AC-mp-ported by absolutely convex compact sets. (c) The seminorms p satisfying that there exists an absolutely convex compact set K such that, for all ε > 0 exists C(ε) > 0 so that p(f ) ≤ C(ε) ∞Xm=0 εm sup x∈K κp(Pmf (x)) ∀f ∈ HKp(E; F ). (d) The seminorms p satisfying that there exists an absolutely convex compact set K such that, for all ε > 0 exists C(ε) > 0 so that p(f ) ≤ C(ε) ∞Xm=0 mp(Pmf (0)(K + εBE) ∀f ∈ HKp(E; F ). (e) The seminorms of the form p(f ) = ∞Xm=0 mp(Pmf (0)(K + amBE)), where K ranges over all the absolutely convex compact sets and (am)m ∈ c+ 0 . Proof. First note that if f is p-compact and K is a compact set, there exists a open set V ⊃ K such that f (V ) is p-compact. Then, seminorms in (a) and (b) are well defined on HKp(E; F ). Also, in virtue of Proposition 5.4, seminorms in (d) and (e) are well defined. Standard arguments show that seminorms in (a) and (b) define the same topology. Now we show that seminorms in (b) and (c) coincide. Let p be a seminorm and let K be an absolutely convex compact set satisfying the conditions in (c). Let V ⊃ K be any absolutely convex open set and take d = dist(K, CV ) > 0. By Proposition 3.3, since K + dBE ⊂ V , we get mp(Pmf (x)(dBE)) ≤ mp(f (x + dBE)) ≤ mp(f (V )), for all f ∈ HKp(E; F ). Thus, for each m. Hence dm sup x∈K κp(Pmf (x)) ≤ mp(f (V )), 27 p(f ) ≤ C( d 2) ∞Xm=0 2)m supx∈K κp(Pmf (x)) ≤ 2C( d ( d 2)mp(f (V )), which shows that p is AC-mp-ported by K. conditions in (b). Fix ε > 0 and take x1, . . . , xn in K such that K ⊂ V with V =Sn Conversely, let p be a seminorm, let K be an absolutely convex compact set satisfying the j=1 Bε(xj). As we did before, we may find an absolutely convex open set W so that K ⊂ W ⊂ V . Let f ∈ HKp, without loss of generality we may assume that ε < rp(f, x) for all x ∈ K. By Remark 3.5, we obtain mp(f (Bε(xj))) ≤ ∞Xm=0 εmκp(Pmf (xj)) ≤ ∞Xm=0 εm sup x∈K κp(Pmf (x)). As p is AC-mp-ported by K, we have that p(f ) ≤ CW mp(f (W )) ≤ CW mp(f (V )) and therefore p(f ) ≤ CW ≤ CW = nCW nXj=1 nXj=1 ∞Xm=0 ∞Xm=0 mp(f (Bε(xj))) εm sup x∈K κp(Pmf (x)) εm sup x∈K κp(Pmf (x)). Thus p belongs to the family in (c). If ε ≥ rp(f, x), then Pm≥0 εm supx∈K κp(Pmf (x)) = ∞ and the inequality follows. By the proof of [15, Proposition 4], we have that seminorms in (d) and (e) generate the same topology. Finally, we show that seminorms in (d) and (b) are equivalent. The proof of Proposition 5.4 shows that seminorms in (d) are AC-mp-ported by absolutely convex compact sets. To conclude the proof, consider a seminorm p and an absolutely convex compact set K satisfying conditions in (b). We borrow some ideas of [16, Chapter 3]. For each m, let Wm be the absolutely convex open set defined by Wm = K + ( 1 2)mBE. Since p is AC-mp-ported by K, for each m ∈ N, there exists a constant Cm = CWm such that p(f ) ≤ Cmmp(f (Wm)), every p-compact function f . For m = 1, there exists n1 ∈ N, such that for all n > n1, C 1/n 1 < 2. Take V1 = 2W1. Now, if n > n1 and Q ∈ PKp(nE; F ), p(Q) ≤ C1mp(Q(W1)) = mp(Q(C 1/n 1 W1)) ≤ mp(Q(V1)). For m = 2, there exists n2 > n1 such that C 1/n and, as before, we have for any Q ∈ PKp(nE; F ), with n > n2, 2 ≤ 2, for all n > n2. Now, take V2 = 2W2 p(Q) ≤ C2mp(Q(W2)) = mp(Q(C 1/n 2 W2)) ≤ mp(Q(V2)). 28 SILVIA LASSALLE AND PABLO TURCO Repeating this procedure we obtain a sequence of absolutely convex open sets Vj satisfying p(f ) ≤ Xm≥0 p(Pmf (0)) = Xm<n1 p(Pmf (0)) +Xj≥1 Xnj ≤m<nj+1 p(Pmf (0)) ≤ CV1 Xm<n1 ≤ C Xm<n1 mp(Pmf (0)(V1)) +Xj≥1 Xnj≤m<nj+1 mp(Pmf (0)(V1)) +Xj≥1 Xnj ≤m<nj+1 mp(Pmf (0)(Vj)) mp(Pmf (0)(Vj))  where C = min{1, CV1} and the result follows since Vj = 2K + ( 1 2)j−1BE and the seminorm p is bounded above by a seminorm of the family of the form (e). Now, the proof is complete. (cid:3) We finish this section by inspecting the κp-approximation property introduced in [13]. We will show that p-compact homogeneous polynomials from F to E can be κp-approximated by polynomials in P(mF ) ⊗ E whenever E has the κp-approximation property. We then obtain a similar result for p-compact holomorphic functions. What follows keeps the spirit of [5, Theorem 4.1]. Recall that a Banach space E has the κp-approximation property if for every Banach space F , F ′ ⊗ E is κp-dense in Kp(F ; E). Theorem 5.6. Let E be a Banach space. The following statements are equivalent. (i) E has the κp-approximation property. (ii) For all m ∈ N, P(mF ) ⊗ E is κp-dense in PKp(mF, E), for every Banach space F . (iii) H(F ) ⊗ E is τω,mp-dense in HKp(F ; E) for all Banach spaces F . Proof. First, suppose that E has the κp-approximation property and fix m ∈ N. Then, πs F )′⊗E is κp-dense in Kp(Nm (Nm via the isomorphism given by P 7→ LP . Thus, (ii) is satisfied. πs F ; E) which, by Proposition 2.1, coincides with (PKp(mF, E), κp), K ⊂ F is an absolutely convex compact set and (am)m ∈ c+ Now, assume (ii) holds. Take f ∈ HKp(F, E), ε > 0. By Theorem 5.5, we may con- m=0 mp(Pmf (0)(K + amBF )), where 0 . Let m0 ∈ N be such that C (K + amBF ) ⊂ BF , for all m ≤ m0. Given δ > 0, to be chosen later, by hypothesis, we may find Qm ∈ P(mF ) ⊗ E m=0 Qm, which belongs to sider a τω,mp-continuous seminorm of the form q(f ) =P∞ Pm>m0 such that κp(Pmf (0) − Qm) ≤ δ, for all m ≤ m0. Define g = Pm0 2 and let C > 0 be such that 1 mp(Pmf (0)(K + amBF )) ≤ ε H(F ) ⊗ E, then q(f − g) = ≤ m0Xm=0 m0Xm=0 mp((Pmf (0) − Qm)(K + amBF )) + Xm>m0 C mκp((Pmf (0) − Qm)) + ε 2 . mp(Pmf (0)(K + amBF )) Thus, q(f − g) < ε for a suitable choice of δ, which proves (iii). 29 Finally, suppose we have (iii). Take T ∈ Kp(F, E), ε > 0 and the seminorm on HKp(F ; E) defined by q(f ) = κp(P1f (0)). Since q is τω,mp-continuous, by assumption, there exist j=1 fj ⊗ xj) < ε. In other words, j=1 P1fj(0) ⊗ xj) < ε which proves that F ′ ⊗ E is κp-dense in Kp(F, E). Whence, (cid:3) f1, . . . , fn ∈ H(F ) and x1, . . . , xn ∈ E, such that q(T −Pn κp(T −Pn the proof is complete. Acknowledgements. We wish to express our gratitude to Richard Aron for introducing us to the problem during his visit to Buenos Aires in 2007 and for many useful conversations thereafter. Also, we are grateful to Manuel Maestre for sharing with us the open problems in a preliminary version of [2]. Finally, we thank Nacho Zalduendo for the suggestions he made to consider holomorphy types and to Christopher Boyd for his careful reading and comments on the manuscript. References [1] Aron, R., Berner, P. A Hahn-Banach extension theorem for analytic mappings, Bull. Math. Soc. France, 106 (1978), 3 -- 24. [2] Aron R., Maestre M., Rueda P. p-compact holomorphic mappings, RACSAM 104 (2) (2010), 353 -- 364. [3] Aron R., Rueda P. p-Compact homogeneous polynomials from an ideal point of view, Contemporary Mathematics, Amer. Math. Soc. To appear. [4] Aron R., Rueda P. I-bounded holomorphic functions. Preprint. [5] Aron R., Schottenloher M., Compact holomorphic mappings and the approximation property, J. Funct. Anal. 21, (1976), 7-30 [6] Boyd C., Dineen S., Rueda P. Weakly uniformly continuous holomorphic functions and the approxima- tion property, Indag. Math. (N.S.) 12 (2) (2001), 147 -- 156. [7] C¸ ali¸skan E. The bounded approximation property for spaces of holomorphic mappings on infinite dimen- sional spaces, Math. Nachr. 279 (7) (2006), 705 -- 715. [8] Carando D., Dimant V.; Muro S. Coherent sequences of polynomial ideals on Banach spaces, Math. Nachr.282 (8) (2009), 1111 -- 1133. [9] Casazza P. Approximation properties. Handbook of the geometry of Banach spaces, Vol. I, 271 -- 316, North-Holland, Amsterdam, 2001. [10] Choi Y.S., Kim J.M, The dual space of (L(X, Y ); τp) and the p-approximation property, J. Funct. Anal 259, (2010) 2437-2454. [11] Davie A. M., Gamelin T. W. A theorem on polynomial-star approximation. Proc. Amer. Math. Soc. 106 (2) (1989), 351 -- 356. [12] Delgado, J. M., Oja, E., Pineiro, C., Serrano, E. The p-approximation property in terms of density of finite rank operators, J. Math Anal, Appl. 354 (2009), 159-164. [13] Delgado, J. M., Pineiro, C., Serrano, E. Density of finite rank operators in the Banach space of p-compact operators, J. Math. Anal. Appl. 370 (2010), 498-505. [14] Delgado, J. M., Pineiro, C., Serrano, E. Operators whose adjoints are quasi p-nuclear, Studia Math. 197 (3) (2010), 291-304. [15] Dineen, S. Holomorphy types on a Banach space, Studia Math. 39 (1971), 241-288. [16] Dineen, S. Complex analysis of infinite dimensional spaces, S.M.M., Springer, 1999. [17] Dineen S., Mujica J., The approximation property for spaces of holomorphic functions on infinite- dimensional spaces. I. J. Approx. Theory 126 (2) (2004), 141 -- 156. 30 SILVIA LASSALLE AND PABLO TURCO [18] Dineen S., Mujica J., The approximation property for spaces of holomorphic functions on infinite- dimensional spaces. II, J. Funct. Anal. 259 (2) (2010), 545 -- 560. [19] Enflo, P. A counterexample to the approximation problem in Banach spaces. Acta Math. 130 (1), (1973), 309 -- 317. [20] Galicer D., Lassalle S., Turco P. On p-compact operators, duality and tensor norms. Preprint [21] Grothendieck A. Produits tensoriels topologiques et espaces nucl´eaires, Mem. Amer. Math. Soc. (1955), no. 16, 140 pp. [22] Lindenstrauss J., Tzafriri L., Classical Banach Spaces I., vol. 92, Springer-Verlag, Berlin, New York, 1977. [23] Mujica, J. Complex Analysis in Banach Spaces, Math. Studies, vol. 120, North-Holland, Amsterdam, 1986. [24] Mujica, J. Linearization of bounded holomorphic mappings on Banach spaces, Trans. Amer. Math. Soc. 324 (2) (1991), 867 -- 887. [25] Nachbin, L. Topology on Spaces of Holomorphic Mappings, Erg. d. Math. 47, Springer-Verlag, Berlin, 1969. [26] Nachbin, L. Concernig holomorphy types for Banach Spaces, Studia Math. 38, (1970) 407-412. [27] Persson, A., Pietsch, A. p-nukleare une p-integrale Abbildungen in Banachraumen, (German) Studia Math. 33 1969 19 -- 62. [28] Schwartz, L. Th´eorie des distributions `a valeurs vectorielles. I. (French) Ann. Inst. Fourier, Grenoble 7, (1957) 1 -- 141. [29] Sinha, D.P., Karn, A. K. Compact operators whose adjoints factor through subspaces of ℓp, Studia Math. 150 (2002), 17-33. [30] Sinha, D.P., Karn, A. K. Compact operators which factor through subspaces of ℓp, Math. Nach. 281 (2008), 412-423. Departamento de Matem´atica - Pab I, Facultad de Cs. Exactas y Naturales, Universidad de Buenos Aires, (1428) Buenos Aires, Argentina E-mail address: [email protected], [email protected]
1706.08047
1
1706
2017-06-25T07:40:14
Convexity of parameter extensions of some relative operator entropies with a perspective approach
[ "math.FA" ]
In this paper, we introduce two notions of a relative operator $(\alpha, \beta)$-entropy and a Tsallis relative operator $(\alpha, \beta)$-entropy as two parameter extensions of the relative operator entropy and the Tsallis relative operator entropy. We apply a perspective approach to prove the joint convexity or concavity of these new notions, under certain conditions concerning $\alpha$ and $\beta$. Indeed, we give the parametric extensions, but in such a manner that they remain jointly convex or jointly concave.
math.FA
math
Convexity of parameter extensions of some relative operator entropies with a perspective approach1 Ismail Nikoufar Department of Mathematics, Payame Noor University, P.O. Box 19395-3697 Tehran, Iran e-mail: [email protected] Abstract. In this paper, we introduce two notions of a relative operator (α, β)-entropy and a Tsallis relative operator (α, β)-entropy as two parameter extensions of the relative operator entropy and the Tsallis relative operator entropy. We apply a perspective approach to prove the joint convexity or concavity of these new notions, under certain conditions concerning α and β. Indeed, we give the parametric extensions, but in such a manner that they remain jointly convex or jointly concave. Significance Statement. What is novel here is that we convincingly demonstrate how our techniques can be used to give simple proofs for the old and new theorems for the functions that are relevant to quantum statistics. Our proof strategy shows that the joint convexity of the perspective of some functions plays a crucial role to give simple proofs for the joint convexity (resp. concavity) of some relative operator entropies. Mathematics Subject Classification. 81P45, 15A39, 47A63, 15A42, 81R15. Key words and phrases: perspective function, generalized perspective function, relative operator entropy, Tsallis relative operator entropy. ⋆ The notions introduced here were used in our published paper [15], when this paper was a draft. 1. Introduction Let H be an infinite-dimensional (separable) Hilbert space. Let B(H) denote the set of all bounded linear operators on H, B(H)sa the set of all self -- adjoint operators, B(H)+ the set of all positive operators, and B(H)++ the set of all strictly positive operators. A continuous real function f on [0, ∞) is said to be operator monotone (more precisely, operator monotone increasing) if A ≤ B implies f (A) ≤ f (B) for A, B ∈ B(H)sa. For a self -- adjoint operator A, the value f (A) is defined via functional calculus as usual. The function f is called operator convex if f (cA1 + (1 − c)A2) ≤ cf (A1) + (1 − c)f (A2) (1.1) 1to appear in Glasgow Mathematical Journal 2 I. Nikoufar for all A1, A2 ∈ B(H)sa and c ∈ [0, 1]. Moreover, the function f is operator concave if −f is operator convex. The function g of two variables is called jointly convex if g(cA1 + (1 − c)A2, cB1 + (1 − c)B2) ≤ cg(A1, B1) + (1 − c)g(A2, B2) (1.2) for all A1, A2, B1, B2 ∈ B(H)sa and c ∈ [0, 1], and jointly concave if the sign of inequality (1.2) is reversed. Let f and h be two functions defined on [0, ∞) and (0, ∞), respectively and let h be a strictly positive function, in the sense that, h(A) ∈ B(H)++ for A ∈ B(H)++. We introduced in [2] a fully noncommutative perspective of two variables (associated to f ), by choosing an appropriate ordering, as follows: Πf (A, B) := A1/2f (A−1/2BA−1/2)A1/2 for A ∈ B(H)++ and B ∈ B(H)sa. We also introduced the operator version of a fully noncommutative generalized perspective of two variables (associated to f and h) as follows: Πf ∆h(A, B) := h(A)1/2f (h(A)−1/2Bh(A)−1/2)h(A)1/2 for A ∈ B(H)++ and B ∈ B(H)sa. This beautiful contribution can surely affect quantum information theory and quantum statistical mechanics. Noncommutative functional analysis gives an appropriate framework for many of the calculations in quantum information theory and nonclassical techniques that clarify some of the conceptual problems in operator convexity theory. Note that the introduced perspective Πf with the operator monotone function f is the operator mean introduced by Kubo and Ando in [12]. By recalling that if for every continuous function f , f (A) commutes with every operator commuting with A (including A itself) and when we restricted to positive commuting matrices, i.e., [A, B] = 0, it becomes Effros's approach which is considered in [3] as follows: Πf (A, B) := f ( Πf ∆h(A, B) := f ( B A )A, B h(A) )h(A). Afterwards, we introduced the notion of the non-commutative perspective in [2], this notion was studied by Effros and Hansen in [4]. They proved that the non-commutative perspective of an operator convex function is the unique extension of the corresponding commutative perspective that preserves homogeneity and convexity. In [2], we proved several striking matrix analogues of a classical result for operator convex functions. Indeed, we proved the following two theorems that entail the necessary and sufficient conditions for the joint convexity of a fully noncommutative perspective and generalized perspective function. We applied the affine version of Hansen -- Pedersen -- Jensen inequality [9, Theorem 2.1] to prove the following result: Theorem 1.1. The function f is operator convex (concave) if and only if the perspective function Πf is jointly convex (concave). We also used Hansen -- Pedersen -- Jensen inequality [8, Theorem 2.1] to prove the following result: Convexity of parameter extensions of some relative operator entropies 3 Theorem 1.2. Suppose that f and h are continuous functions with f (0) < 0 and h > 0. Then f is operator convex and h is operator concave if and only if the generalized perspective function Πf ∆h is jointly convex. In the 'if' part of the above theorem, I would remark that we could allow f (0) ≤ 0. However, for the next applications in the 'only if' part the condition f (0) 6= 0 is essential. So, this theorem and its reverse one can be modified as follows. Note that part (ii) of Theorem 1.3 is a correct version of [2, Corollary 2.6 (i)] and parts (iii) and (v) are a complete and correct version of [2, Corollary 2.6 (ii)]. We include the proofs for the convenience of the readers. Theorem 1.3. Suppose that f and h are continuous functions and h > 0. (i) If f is operator convex and h is operator concave with f (0) ≤ 0, then the generalized perspective function Πf ∆h is jointly convex. (ii) If f and h are operator concave with f (0) ≥ 0, then the generalized perspective function Πf ∆h is jointly concave. (iii) If the generalized perspective function Πf ∆h is jointly convex (concave), then f is operator convex (con- cave). (iv) If f (0) > 0 and the generalized perspective function Πf ∆h is jointly convex (concave), then h is operator convex (concave). (v) If f (0) < 0 and the generalized perspective function Πf ∆h is jointly convex (concave), then h is operator concave (convex). Proof. (i) For the strictly positive operators A1, A2, the self -- adjoint operators B1, B2, and c ∈ [0, 1] set A := cA1 +(1− c)A2 and B := cB1 +(1− c)B2. Define T1 := (ch(A1))1/2h(A)−1/2 and T2 := ((1− c)h(A2))1/2h(A)−1/2. The concavity of h gives T ∗ 1 T1 + T ∗ 2 T2 ≤ 1 and the operator convexity of f together with Hansen -- Pedersen -- Jensen inequality [8] imply Πf ∆h(A, B) = h(A)1/2f (h(A)−1/2Bh(A)−1/2)h(A)1/2 1 h(A1)−1/2B1h(A1)−1/2T1 + T ∗ 2 h(A2)−1/2B2h(A2)−1/2T2(cid:17) h(A)1/2 = h(A)1/2f (cid:16)T ∗ ≤ h(A)1/2(cid:16)T ∗ 1 f (h(A1)−1/2B1h(A1)−1/2)T1 2 f (h(A2)−1/2B2h(A2)−1/2)T2(cid:17) h(A)1/2 = ch(A1)1/2f (h(A1)−1/2B1h(A1)−1/2)h(A1)1/2 + T ∗ + (1 − c)h(A2)1/2f (h(A2)−1/2B2h(A2)−1/2)h(A2)1/2 = cΠf ∆h(A1, B1) + (1 − c)Πf ∆h(A2, B2). (ii) It follows from (i) by replacing −f with f . 4 I. Nikoufar (iii) A simple computation shows that f (A) = 1 h(1) Πf ∆h(1, h(1)A). Then, by using the joint convexity of Πf ∆h, for the self -- adjoint operators A1, A2 and 0 ≤ c ≤ 1 we have f (cA1 + (1 − c)A2) = = ≤ 1 h(1) 1 h(1) 1 h(1) Πf ∆h(1, h(1)(cA1 + (1 − c)A2)) Πf ∆h(1, ch(1)A1 + (1 − c)h(1)A2) (cΠf ∆h(1, h(1)A1) + (1 − c)Πf ∆h(1, h(1)A2)) = cf (A1) + (1 − c)f (A2). (iv) It is obvious that h(A) = 1 f (0) Πf ∆h(A, 0). By using f (0) > 0, for the strictly positive operators A1, A2 and 0 ≤ c ≤ 1 we get h(cA1 + (1 − c)A2) = ≤ 1 f (0) 1 f (0) Πf ∆h(cA1 + (1 − c)A2, 0) (cΠf ∆h(A1, 0) + (1 − c)Πf ∆h(A2, 0)) = ch(A1) + (1 − c)h(A2). (v) The proof is similar to that of (iv). (cid:3) 2. Parametric relative operator entropies Generalized entropies are used as alternate measures of an informational content. Studies of generalized entropies allow to treat properties of the standard entropy in more general setting. The connection between strong subadditivity of the von Neumann entropy and the Wigner -- Yanase -- Dyson conjecture is a remarkable example (see [10, 11]). In this section, we show usefulness of the notions of the perspective and the generalized perspective to obtain the joint convexity of the (quantum) relative operator entropy, the joint concavity of the Fujii -- Kamei relative operator entropy and the Tsallis relative operator entropy, and moreover the joint convexity (concavity) of some other well-known operators. Yanagi et al. [17] defined the notion of the Tsallis relative operator entropy and gave its properties and the generalized Shannon inequalities. Furuichi et al. [6] defined this notion as a parametric extension of the relative operator entropy and proved some operator inequalities related to the Tsallis relative operator entropy. For the strictly positive matrices A, B and 0 < λ ≤ 1, Tλ(AB) := A 1 2 (A− 1 2 BA− 1 2 )λA 1 2 − A λ is called the Tsallis relative operator entropy between A and B [17]. We often rewrite the Tsallis relative operator entropy Tλ(AB) as Tλ(AB) = A 1 2 lnλ(A− 1 2 BA− 1 2 )A 1 2 , where lnλ X ≡ X λ−1 λ for the positive operator X [6]. Convexity of parameter extensions of some relative operator entropies 5 We give a generalized notion of the Tsallis relative operator entropy and call it a Tsallis relative operator (α, β)-entropy. We define Tα,β(AB) := A β 2 lnα(A− β 2 BA− β 2 )A β 2 for the strictly positive operators A, B and the real numbers α 6= 0, β. It is clear that every Tsallis relative operator (λ, 1)-entropy is the Tsallis relative operator entropy, i.e., Tλ(AB) = Tλ,1(AB). We want to establish the joint convexity or concavity of Tα,β with a perspective approach, namely, we find the functions f, h such that Tα,β(AB) = Πf ∆h(A, B). In particular, we reach a simple result on the joint convexity or concavity of the Tsallis relative operator entropy. Lemma 2.1. The function lnλ(t) is operator convex for λ ∈ [1, 2] and operator concave for λ ∈ [−1, 0) ∪ (0, 1]. Proof. The result follows from the operator convexity or concavity of the elementary function tλ. (cid:3) Theorem 2.2. The Tsallis relative operator (α, β)-entropy is jointly convex for α ∈ [1, 2] and β ∈ [0, 1]. Proof. Note that the Tsallis relative operator (α, β)-entropy Tα,β is the generalized perspective of the functions lnα(t) and tβ, in the sense that, Tα,β(AB) = Πlnα t∆tβ (A, B). Therefore, we obtain the result from Lemma 2.1 and Theorem 1.3 (i). (cid:3) We remark that the concavity assertion in Theorem 2.2 is doubtful. In fact, lnα(t) is operator concave for α ∈ [−1, 0) ∪ (0, 1], where lnα(0) < 0, so Theorem 1.3 (ii) can not be applied. Applying Theorem 1.1 the concavity assertion for the Tsallis relative operator entropy is not doubtful. Theorem 2.3. The Tsallis relative operator entropy is jointly convex for α ∈ [1, 2] and jointly concave for α ∈ [−1, 0) ∪ (0, 1]. Proof. We have Πlnα t(A, B) = Tα(AB) and the result follows from Lemma 2.1 and Theorem 1.1. (cid:3) The notion of the relative operator entropy was introduced on strictly positive matrices in noncommutative information theory by Fujii and Kamei [5] as an extension of the operator entropy considered by Nakamura and Umegaki [14] and the relative operator entropy considered by Umegaki [16] as follows: S(AB) := A 1 2 (log A− 1 2 BA− 1 2 )A 1 2 . Fujii et al. [5] estimated the value of the relative operator entropy S(AB) by applying the Furuta's inequality and obtained the upper and lower bounds of S(AB). It is obvious that S(AB) = limα→0 Tα(AB) for A, B > 0. Hence, S(A, B) is jointly concave by Theorem 2.3. We show that the joint concavity of the relative operator entropy is a simple consequence of the joint concavity of the perspective of the elementary function f (t) = log t. Theorem 2.4. The Fujii -- Kamei relative operator entropy S(AB) is jointly concave on the strictly positive operators A, B. 6 I. Nikoufar Proof. The relative operator entropy S(AB) is the perspective of log t in the sense of our definition and so Theorem 1.1 and the operator concavity of log t imply the result. (cid:3) Effros gave a new interesting proof for Lieb and Ruskai's result [13] (see Corollary 2.1 of [3]) and now we provide simple proofs for the same results. Theorem 2.5. (i) The (quantum) relative entropy (ρ, σ) 7→ H(ρkσ) = T race ρ log ρ − ρ log σ is jointly convex on the commutative strictly positive operators ρ, σ. (ii) Part (i) holds for the noncommutative strictly positive operators ρ, σ. Proof. The following equalities show that parts (i) and (ii) are a simple application of Theorem 2.4. (i) We have H(ρkσ) = −T race S(ρσ). (ii) For the commuting operators Lρ and Rσ we have h−S(LρRσ)(I), Ii = T race ρ log ρ − ρ log σ, where h·, ·i is the Hilbert-Schmidt inner product, Lρ is the Left multiplication by ρ and Rσ is the right multipli- cation by σ. (cid:3) Furuta [7] defined the generalized relative operator entropy for the strictly positive operators A, B and q ∈ R by Sq(AB) = A1/2(A−1/2BA−1/2)q(log A−1/2BA−1/2)A1/2. Using the notion of the generalized relative operator entropy, Furuta obtained the parametric extension of the operator Shannon inequality and its reverse one. Note that for q = 0, we get the relative operator entropy between A and B, i.e., S0(AB) = S(AB). A natural question now arises: What can we say about the joint convexity or concavity of the generalized relative operator entropy? We will find a function f such that Sq(AB) = Πf (A, B). We discuss this in the next section. 3. Generalized transpose operator functions and its applications A motivation to write this section is to prove the joint convexity of the generalized relative operator entropy introduced by Furuta [7]. We also give a parametric extension of this notion, namely, we introduce the notion of a relative operator (α, β)-entropy and prove that Sα,β is jointly convex for α, β ∈ [0, 1]. Definition 3.1. Let f and h be continuous functions and h > 0. We define a generalized transpose function with respect to the functions f and h by f ∗ h (t) := h(t)f ( 1 h(t) ). Convexity of parameter extensions of some relative operator entropies 7 In particular, the transpose function with respect to the function f is defined by f ∗(t) := tf (t−1). The following result is a straight forward consequence of Theorem 1.1. Indeed, we have f ∗(A) = Af (A−1) = A1/2f (A−1/2A−1/2)A1/2 = Πf (A, 1). Theorem 3.2. Suppose that f is a continuous function. Then, f is operator convex (concave) if and only if so is f ∗. The trace operation plays a central role in quantum statistical mechanics. The mapping A 7→ T raceKf (A) is certainly convex when K > 0 and f is operator convex. Corollary 3.3. The von Neumann entropy S(ρ) = −T raceρ log ρ is operator concave on the strictly positive operator ρ. Proof. For the operator concave function f (t) = log t we have f ∗(t) = −t log t. Using Theorem 3.2, we deduce the function −t log t is operator concave and hence we obtain the desired result. (cid:3) Theorem 3.4. Suppose that f and h are continuous functions and h > 0. (i) If f is operator convex with f (0) ≤ 0 and h is operator concave, then f ∗ h is operator convex. (ii) If f and h are operator concave with f (0) ≥ 0, then f ∗ h is operator concave. Proof. (i) Let f be operator convex and h operator concave. Then, it follows from Theorem 1.3 (i) that Πf ∆h is jointly convex. The fact that a jointly convex function is convex in each of its arguments separately and Πf ∆h(A, 1) = f ∗ h (A) imply f ∗ h is operator convex. (ii) The result comes from Theorem 1.3 (ii). (cid:3) The proof of the following lemma is straightforward. Lemma 3.5. The function f is operator convex (concave) if and only if so is fε for every ε > 0, where fε(t) := f (t + ε). Let f be a twice differentiable function on [0, ∞). Define k(t) := tqf (t) for 0 ≤ q ≤ 1 and consider Iq := {t ≥ 0 : k′′(t) ≥ 0}. Clearly, k is not convex on R+ − Iq and hence is not operator convex on outside of Iq. We show that under some assumptions k is operator convex on Iq. Lemma 3.6. If f is operator monotone on [0, ∞) such that f (0) ≤ 0 and limt→∞ f (t) t = 0, then the function k is operator convex on Iq. Proof. The operator monotone function f on [0, ∞) can be represented as f (t) = f (0) + βt + Z ∞ 0 λt λ + t dµ(λ), 8 I. Nikoufar where β ≥ 0 and µ is a positive measure on [0, ∞); see [1, Chapter V]. Since limt→∞ f (t) t = 0, β = 0. So by multiplying both sides to tq we have tqf (t) = f (0)tq + Z ∞ 0 λt1+q λ + t dµ(λ). The function f (0)tq is operator convex. Indeed, it is sufficient to prove that the function λt1+q Define g(t) = tq and consider two cases: (i) For λ > 1 the function h(t) = t+1/λ λ+t is operator convex. g(t) is operator monotone by t [1, Corollary V.3.12]. So [8, Theorem 2.4] and [1, Problem V.5.7] show that the function th(t−1)−1 is operator convex. (ii) For 0 < λ < 1 the function h1(t) = t+λ is also operator monotone. Therefore, the function λ2qts( t t g(t) is operator monotone by [1, Corollary V.3.12]. So [1, Problem V.5.7] entails that the function s(t) = h1(λt−1)−1 is operator monotone. This implies the function s( t λ ) λ ) is operator convex by [8, Theorem 2.4]. In each of λ ) in the the cases a simple calculation shows that the function th(t−1)−1 in the case (i) and the function λ2qts( t case (ii) are equal to the function λt1+q λ+t . (cid:3) Lemma 3.7. The function k(t) = tq log t is operator convex on Jq := [0, e 2q−1 q(1−q) ] for 0 ≤ q ≤ 1. Proof. Let ε ∈ (0, 1). Then, the function fε(t) = log(t + ε) satisfies in the assumptions of Lemma 3.6 and so kε(t) = tq log(t + ε) is operator convex on an interval Jq,ε ⊆ Jq. Hence, when ε → 0 we see that the function k is operator convex on Jq. (cid:3) We now prove the main result of this section and its generalization. Corollary 3.8. The generalized relative operator entropy Sq(AB) is jointly convex on the strictly positive oper- ators A, B with spectra in Jq and 0 ≤ q ≤ 1. Proof. The generalized relative operator entropy Sq(AB) is the perspective of the operator convex function tq log t, t > 0 and so Lemma 3.7 and Theorem 1.1 ensure that Sq(AB) is jointly convex. (cid:3) We introduce a relative operator (α, β)-entropy (two parameters relative operator entropy) as follows: Sα,β(AB) = A β 2 (A− β 2 BA− β 2 )α(log A− β 2 BA− β 2 )A β 2 for the strictly positive operators A, B and the real numbers α, β. We consider its convexity or concavity properties. In particular, we have Sq,1(AB) = Sq(AB) and S0,1(AB) = S(AB). Theorem 3.9. The relative operator (α, β)-entropy Sα,β(AB) is jointly convex on the strictly positive operators A, B with spectra in Jα and 0 ≤ α, β ≤ 1. Proof. Consider f (t) := tα log t and h(t) := tβ. Then, f (0) = 0 and Sα,β(AB) = Πf ∆h(A, B). Using Theorem 1.3 (i) and Lemma 3.7 we deduce the generalized perspective of the operator convex function f and the operator concave function h is jointly convex so that the relative operator (α, β)-entropy Sα,β(AB) is jointly convex. (cid:3) Convexity of parameter extensions of some relative operator entropies 9 References [1] R. Bhatia, Matrix Analysis, Springer-Verlag, 1996. [2] A. Ebadian, I. Nikoufar, and M. Eshagi Gordji, Perspectives of matrix convex functions, Proc. Natl. Acad. Sci., vol. 108, no. 18, (2011), 7313 -- 7314. [3] E. G. Effros, A matrix convexity approach to some celebrated quantum inequalities, Proc. Natl. Acad. Sci. U S A., 106(4), (2009), 1006 -- 1008. [4] E. G. Effros and F. Hansen, Non-commutative perspectives, Ann. Funct. Anal., 5(2) (2014), 74 -- 79. [5] J. I. Fujii and E. Kamei, Relative operator entropy in noncommutative information theory, Math. Japonica, vol. 34, (1989), 341 -- 348. [6] S. Furuichi, K. Yanagi, and K. Kuriyama, A note on operator inequalities of Tsallis relative operator entropy, Linear Alg. Appl., vol. 407, (2005), 19 -- 31. [7] T. Furuta, Parametric extensions of Shannon inequality and its reverse one in Hilbert space operators, Linear Algebra Appl. vol. 381, (2004) 219-235. [8] F. Hansen and G. Pedersen, Jensen's Inequality for Operators and Lowner's Theorem, Math. Ann., 258, (1982), 229 -- 241. [9] F. Hansen and G. Pedersen, Jensen's operator inequality, Bull. London Math. Soc., 35 (2003), 553 -- 564. [10] F. Hiai, M. Mosonyi, D. Petz, and C. Beny, Quantum f -divergences and error correction, Rev. Math. Phys., 23 (2011), 691-747. [11] A. Jencova and M. B. Ruskai, A unified treatment of convexity of relative entropy and related trace functions, with conditions for equality, Rev. Math. Phys. 22 (2010), 1099-1121. [12] F. Kubo and T. Ando, Means of positive linear operators, Math. Ann., vol. 246, (1979-1980), 205 -- 224. [13] E. Lieb and M. Ruskai, Proof of the strong subadditivity of quantum-mechanical entropy, With an appendix by B. Simon, J. Math. Phys., vol. 14, (1973), 1938 -- 1941. [14] M. Nakamura and H. Umegaki, A note on the entropy for operator algebras, Proc. Japan Acad., vol. 37, (1967), 149 -- 154. [15] I. Nikoufar, On operator inequalities of some relative operator entropies, Adv. Math., vol. 259, (2014), 376-383. [16] H. Umegaki, Conditional expectation in operator algebra IV (entropy and information), Kodai Math. Sem. Rep., vol. 14, (1962), 59 -- 85. [17] K. Yanagi, K. Kuriyama, and S. Furuichi, Generalized Shannon inequalities based on Tsallis relative operator entropy, Linear Alg. Appl., vol. 394, (2005), 109 -- 118.
1503.06893
1
1503
2015-03-24T02:27:21
Detecting Fourier subspaces
[ "math.FA", "math.CO", "math.GR", "math.OA" ]
Let G be a finite abelian group. We examine the discrepancy between subspaces of l^2(G) which are diagonalized in the standard basis and subspaces which are diagonalized in the dual Fourier basis. The general principle is that a Fourier subspace whose dimension is small compared to |G| = dim(l^2(G)) tends to be far away from standard subspaces. In particular, the recent positive solution of the Kadison-Singer problem shows that from within any Fourier subspace whose dimension is small compared to |G| there is standard subspace which is essentially indistinguishable from its orthogonal complement.
math.FA
math
DETECTING FOURIER SUBSPACES CHARLES A. AKEMANN AND NIK WEAVER Abstract. Let G be a finite abelian group. We examine the discrepancy be- tween subspaces of l2(G) which are diagonalized in the standard basis and subspaces which are diagonalized in the dual Fourier basis. The general principle is that a Fourier subspace whose dimension is small compared to G = dim(l2(G)) tends to be far away from standard subspaces. In particular, the recent positive solution of the Kadison-Singer problem shows that from within any Fourier subspace whose dimension is small compared to G there is standard subspace which is essentially indistinguishable from its orthogonal complement. The purpose of this note is to describe a simple application of the recent solution of the Kadison-Singer problem [6, 7] to a question in harmonic analysis and signal analysis. Let G be a finite abelian group equipped with counting measure, and for each g ∈ G let eg ∈ l2(G) be the function which takes the value 1 at g and is zero elsewhere. Then {eg : g ∈ G} is an orthonormal basis of l2(G); call it the standard basis. Another nice basis of l2(G) comes from the dual group G, the set of characters of G, i.e., homomorphisms from G into the circle group T. Every character has l2 norm equal to G1/2, where G is the cardinality of G, and the normalized set {eφ = G−1/2φ : φ ∈ G} is also an orthonormal basis of l2(G). We call this the Fourier basis. (Note that when every element of G has order 2, then the Fourier basis forms the rows of a Hadamard matrix [11]. The method of this paper applies to bases of this type as well even if they don't arise from a Fourier transform.) Say that a subspace of l2(G) is standard if it is the span of some subset of the standard basis, and Fourier if it is the span of some subset of the Fourier basis. Now each eφ is as far away from the standard basis as possible in the sense that heφ, egi = G−1/2 for all g ∈ G and φ ∈ G. However, Fourier subspaces can certainly intersect standard subspaces -- trivially, l2(G) is itself both a standard subspace and a Fourier subspace. A more interesting question is whether Fourier subspaces whose dimensions are "small" compared to G can intersect standard subspaces which are small in the same sense. This could be of interest in relation to signal analysis, say, if we are trying to detect a signal by measuring a relatively small number of frequencies. (By interchanging the roles of a group and its dual group, we see that the problem of detecting standard subspaces using Fourier subspaces is equivalent to the problem of detecting Fourier subspaces using standard subspaces. But in keeping with the signal analysis perspective, we will stick with the first formulation.) Date: March 23, 2015. Second author partially supported by NSF grant DMS-1067726. 1 2 CHARLES A. AKEMANN AND NIK WEAVER The basic obstacle to having small standard and Fourier subspaces which inter- sect is the uncertainty principle for finite abelian groups [5, 9, 10]. According to this principle, if a nonzero function f ∈ l2(G) is supported on a set S ⊆ G and its Fourier transform is supported on T ⊆ G -- meaning that hf, eφi 6= 0 only for φ ∈ T -- then S · T ≥ G. In the special case where G has prime order p, we have the much stronger inequality S+T ≥ p+ 1, and this inequality is absolutely sharp [9]. Intuitively, if f is very localized with respect to the standard basis then it must be "spread out" with respect to the Fourier basis. In terms of subspaces, the result can be stated as follows: Proposition 0.1. Let G be a finite abelian group and let E and F respectively be standard and Fourier subspaces of l2(G). If dim(E)·dim(F ) < G then E∩F = {0}. If G is prime and dim(E) + dim(F ) ≤ G then E ∩ F = {0}. This follows immediately from the uncertainty principles described above be- cause the dimensions of E and F equal the number of basis elements which span them, so that the support of any element of E (respectively, F ) has cardinality at most dim(E) (respectively, dim(F )). (Uncertainty principles are related to signal reconstruction in a different way in [4] and related papers.) So standard and Fourier subspaces must intersect only in {0} if both are suffi- ciently small. However, according to the multiplicative bound in the last propo- sition, both dimensions could be as small as G1/2, which is small compared to G when G is large. This means that the multiplicative bound does not prevent standard and Fourier subspaces whose dimensions are small compared to G from intersecting, although the additive bound when G is prime certainly does. Indeed, there are easy examples of intersecting standard and Fourier subspaces whose di- mensions are both equal to G1/2. Example 0.2. Let G = Z/n2Z, where G = n2 and the characters have the form φb : a 7→ e2πiab/n2 b=0 φnb ∈ l2(G) satisfies for a, b ∈ Z/n2Z. Here the function f = Pn−1 so that f belongs both to an n-dimensional Fourier subspace (directly from its defi- nition) and to an n-dimensional standard subspace (by the preceding calculation). From the point of view of signal analysis, however, we are probably not so interested in intersecting a single standard subspace. If we do not know where the signal we want to detect is supported, we would presumably want to intersect, if not every standard subspace, at least every standard subspace whose dimension is greater than some threshold value. But unless the relevant dimensions are large, this is impossible for elementary linear algebra reasons. We have the following simple result: Proposition 0.3. Let G be a finite abelian group and let F be a Fourier subspace of l2(G). Then there exists a standard subspace E with dim(E) = G − dim(F ) such that E ∩ F = {0}. The proof is easy. Starting with the standard basis B = {eg : g ∈ G} of l2(G) and the basis B′ = {eφ : φ ∈ T} of F , we can successively replace distinct elements of B with elements of B′ in such a way that the set remains linearly f (a) = n−1 Xb=0 e2πiab/n =(n if a ≡ 0 (mod n) if a 6≡ 0 (mod n), 0 DETECTING FOURIER SUBSPACES 3 independent at each step. The end result will be a (non orthogonal) basis of l2(G) of the form B′ ∪ {eg : g ∈ S} for some set S ⊆ G with S = G − dim(F ). Then E = span{eg : g ∈ S} is the desired standard subspace which does not intersect F . Thus, if the dimension of a Fourier subspace is small compared to G, there will be many standard subspaces whose dimensions are small compared to G which it does not intersect. But requiring subspaces to intersect is a very strong condition. Merely asking that the Fourier subspace contain unit vectors which are close to the standard subspace makes the problem much more interesting and difficult. Nonetheless, we can still get something from simply counting dimensions. Namely, if eφ is any Fourier basis vector and E is a standard subspace, then, letting P be the orthogonal projection onto E = span{eg : g ∈ S}, we have kP eφk2 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xg∈S heφ, egieg(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 = dim(E) G . Thus, even a single Fourier basis vector can "detect" arbitrary standard subspaces to the extent that those subspaces have dimension comparable to G. Can Fourier subspaces do better? The relevant gauge here is the quantity kP Qk = kQPk = sup{kQvk : v ∈ E,kvk = 1} where P and Q are the orthogonal projections onto a standard subspace E and a Fourier subspace F , respectively. It effectively measures the minimal angle be- tween E and F . The surprisingly strong result is that, by this measure, so long as the dimension of a Fourier subspace is small compared to G, there are standard subspaces which it detects only marginally better than a single Fourier basis vector does. One might suspect that a randomly chosen standard subspace would demonstrate that claim. Maybe this technique would work for most Fourier subspaces, but it does not in general, even for Fourier subspaces whose dimension is small compared to G. Example 0.4. Recall Example 0.2 where an n-dimensional Fourier subspace F intersected an n-dimensional standard subspace E. Here dim(F )/G = 1/n, which can be as small as we like. Now consider the group G′ = Z/n2Z × Z/N Z where N is large compared to n. We have a natural identification l2(G′) ∼= l2(Z/n2Z) ⊗ l2(Z/N Z), under which identification F ⊗ l2(Z/N Z) is a Fourier subspace. The ratio of its dimension to the cardinality of G′ is still 1/n. But for sufficiently large N , a ran- domly chosen subset of G′ will contain at least one element of S × Z/N Z with high probability, where E = span{eg : g ∈ S}. One can even say this of a randomly cho- sen subset of G′ of cardinality G′/n. This means that a randomly chosen standard subspace of dimension G′/n will intersect the Fourier subspace F ⊗ l2(Z/N Z) with high probability -- depending on the value of N , with probability as close to 1 as we like. Nonetheless, if a Fourier subspace F has relatively small dimension, there will always exist standard subspaces of arbitrary dimension which are barely closer to F than they are to a single Fourier basis vector. We have the following theorem: 4 CHARLES A. AKEMANN AND NIK WEAVER Theorem 0.5. Let G be a finite abelian group, let F be a Fourier subspace of l2(G), and let Q be the orthogonal projection onto F . Then for any k ≤ G there is a set S ⊆ G with S = k and such that kP Qk2 ≤ k G + O(√ǫ), (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xg∈S ugu∗g − k G = O(√ǫ). Q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) where P is the orthogonal projection onto span{eg : g ∈ S} and ǫ = dim(F )/G. This result is strengthened further by Theorem 0.6 below, but Theorem 0.5 has an easier proof which is of independent interest. We include this proof in the appendix. The strengthened version of Theorem 0.5 simultaneously asks the same question about the complementary standard subspace. If our goal is detection, then the worst that could happen here is that a Fourier subspace F does essentially no better than a single Fourier basis vector at detecting either some standard subspace E or its orthocomplement E⊥. In fact, this worst-case scenario is realized: we can show that every Fourier subspace of small dimension relative to G fails to do significantly better than a single Fourier basis vector at detecting both a sequence of standard subspaces of varying dimension and their orthocomplements. In this case, constructing the undetectable standard subspaces is no longer merely a matter of controlling the largest eigenvalue of QP Q (which suffices be- cause kQP Qk = kP Qk2). Now we also have to control the largest eigenvalue of Q(1− P )Q, and this is a Kadison-Singer type setup. Although the problem we con- sider here is not as general as the full Kadison-Singer problem, the core difficulty is clearly already present. Thus, one should not expect any easier proof than the ones that appear in [6], [7] or the remarkable generalization in [3]. Theorem 0.6. Let G be a finite abelian group, let F be a Fourier subspace of l2(G), and let Q be the orthogonal projection onto F . Then for any k ≤ G there is a set S ⊆ G with S = k and such that both kP Qk2 ≤ + O(√ǫ) k G and k(I − P )Qk2 ≤ G − k G + O(√ǫ) where P is the orthogonal projection onto span{eg : g ∈ S} and ǫ = dim(F )/G. Proof. For each g ∈ G let ug = Qeg ∈ F . Then the rank one operators ugu∗g : f 7→ hf, ugiug satisfy Xg∈G hf, Qegi · Qeg = Q Xg∈G for all f ∈ l2(G); that is, Pg∈G ugu∗g = Q. We also have kugk2 = kQegk2 = ǫ for all g. So by ([1], comment following Corollary 1.2), for any k ≤ G there is a set S ⊆ G such that ugu∗gf = Xg∈G hQf, egieg  = Qf DETECTING FOURIER SUBSPACES 5 Letting P be the orthogonal projection onto span{eg : g ∈ S}, we have P = Pg∈S ege∗g, and it follows that Q(cid:13)(cid:13)(cid:13)(cid:13) k G which also implies = O(√ǫ), ugu∗g − QP Q − (cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xg∈S k G Q(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) = O(√ǫ). (cid:13)(cid:13)(cid:13)(cid:13) Q(I − P )Q − G − k G kP Qk2 = kQP Qk ≤ Q(cid:13)(cid:13)(cid:13)(cid:13) k G + O(√ǫ) We therefore have and k(I − P )Qk2 = kQ(I − P )Qk ≤ G − k G + O(√ǫ), as desired. The set S might not have cardinality exactly k, but it cannot have cardinality greater than k + O(√ǫ) or less than k − O(√ǫ), so it can be adjusted to have cardinality k without affecting the order of the estimate, if needed. (cid:3) In particular, taking k ≈ G/2 in Theorem 0.6, we get kP Qk ≈ k(1 − P )Qk ≈ 1 √2 . This implies that every nonzero vector in ran(P ) is roughly 45◦ away from F , and the same is true of every nonzero vector in ran(P )⊥. That is, if ǫ is small then from within F the two subspaces are essentially indistinguishable. Another way to say this is that every vector in F has roughly half of its l2 norm supported on S and roughly half supported on G \ S. Appendix A. We prove Theorem 0.5. The argument is a straightforward application of the spectral sparsification technique introduced in Srivastava's thesis [8]. Let {u1, . . . , un} be a finite set of vectors in Cm, each of norm √ǫ, satisfying n uiu∗i = I Xi=1 where uiu∗i is the rank one operator on Cm defined by uiu∗i : v 7→ hv, uiiui and I is the identity operator on Cm. Note that Tr(uiu∗i ) = Tr(u∗i ui) = kuik2 = ǫ; since Tr(I) = m, it follows that nǫ = m. Let k < n. As in [8], we will build a sequence ui1 , . . . , uik one element at a time. The construction is controlled by the behavior of the operators Aj = Pj d=1 uid u∗id using the following tool. For any positive operator A and any a > kAk, define the upper potential Φa(A) to be Φa(A) = Tr((aI − A)−1); then, having chosen the vectors ui1 , . . . , uij−1 , the plan will be to select a new vector uij so as to minimize Φaj (Aj), where the aj are an increasing sequence of upper bounds. This potential function disproportionately penalizes eigenvalues which are 6 CHARLES A. AKEMANN AND NIK WEAVER close to aj and thereby controls the maximum eigenvalue, i.e., the norm, of Aj. The key fact about the upper potential is given in the following result. Lemma A.1. ([8], Lemma 3.4) Let A be a positive operator on Cm, let a, δ > 0, and let v ∈ Cm. Suppose kAk < a. If h((a + δ)I − A)−2v, vi + h((a + δ)I − A)−1v, vi ≤ 1 Φa(A) − Φa+δ(A) then kA + vv∗k < a + δ and Φa+δ(A + vv∗) ≤ Φa(A). The proof relies on the Sherman-Morrison formula, which states that if A is positive and invertible then (A + vv∗)−1 = A−1 − A−1(vv∗)A−1 I+hA−1v,vi We also require a simple inequality. . Proof. Let M = 1 Lemma A.2. Let a1 ≤ ··· ≤ am and b1 ≥ ··· ≥ bm be sequences of positive real numbers, respectively increasing and decreasing. Then P aibi ≤ 1 mP aiP bi. mP bi. We want to show that P aibi ≤ P aiM , i.e., that P ai(bi − M ) ≤ 0. Since the sequence (bi) is decreasing, we can find j such that bi ≥ M for i ≤ j and bi < M for i > j. Then Pj i=1 ai(bi − M ) ≤ ajPj i=1(bi − M ) (since the ai are increasing and the values bi − M are positive) i=j+1 ai(bi − M ) ≤ ajPm and Pm i=j+1(bi − M ) (since the ai are increasing and the values bi − M are negative). So m m as desired. ai(bi − M ) ≤ aj Xi=1 (bi − M ) = 0, Xi=1 (cid:3) Theorem A.3. Let m ∈ N and ǫ > 0, and suppose {u1, . . . , un} is a finite sequence of vectors in Cm satisfying kuik2 = ǫ for 1 ≤ i ≤ n and Pn i=1 uiu∗i = I. Then for any k ≤ n there is a set S ⊆ {1, . . . , n} with S = k such that yielding the desired conclusion. We start with A0 = 0, so that Φa0 (A0) = Φ√ǫ(0) = Tr((√ǫI)−1) = m/√ǫ. To carry out the induction step, suppose ui1, . . . , uij have been chosen. Let λ1 ≤ ··· ≤ λm be the eigenvalues of Aj . Then the eigenvalues of I − Aj are 1 − λ1 ≥ ··· ≥ 1 − λm and the eigenvalues of (aj+1I − Aj )−1 are aj+1−λ1 ≤ ··· ≤ aj+1−λm . Thus by Lemma A.2 1 1 Tr((aj+1I − Aj)−1(I − Aj)) = m Xi=1 1 aj+1 − λi (1 − λi) ≤ k n + O(√ǫ). uiu∗i(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi∈S 1−√ǫ · j n for 0 ≤ j ≤ k. We will find a sequence of distinct d=1 uid u∗id , 0 ≤ j ≤ k, satisfy Proof. Define aj = √ǫ + 1 indices i1, . . . , ik such that the operators Aj = Pj kAjk < aj and Φa0(A0) ≥ ··· ≥ Φak (Ak). Thus k = n kAkk < √ǫ + k n 1 1 − √ǫ · + O(√ǫ), DETECTING FOURIER SUBSPACES 7 ≤ = = ≤ ≤ = m m 1 1 m 1 m 1 m 1 m 1 m 1 √ǫ Xi=1 (1 − λi) aj+1 − λi Xi=1 Tr((aj+1I − Aj )−1)Tr(I − Aj) Φaj+1 (Aj )Tr(I − Aj) Φaj (Aj )Tr(I − Aj) Φa0(A0)Tr(I − Aj) Tr(I − Aj ). Next, aj+1 − aj = 1 1−√ǫ · 1 n , so we can estimate Φaj (Aj) − Φaj+1 (Aj) = Tr((ajI − Aj)−1 − (aj+1I − Aj)−1) = > 1 1 − √ǫ · 1 − √ǫ · 1 1 n 1 n Tr((aj I − Aj )−1(aj+1I − Aj)−1) Tr((aj+1I − Aj )−2) of the operator (ajI − Aj)−1(aj+1I − (aj+1−λi)2 of the operator 1 1 1 aj−λi aj+1−λi since each of the eigenvalues Aj)−1 is greater than the corresponding eigenvalue (aj+1I − Aj )−2. Combining this with Lemma A.2 yields Tr((aj+1I − Aj)−2(I − Aj )) ≤ < 1 m 1 ǫ Tr((aj+1I − Aj)−2)Tr(I − Aj) (1 − √ǫ)(Φaj (Aj ) − Φaj+1 (Aj))Tr(I − Aj). Thus 1 ǫ ≤ Φaj (Aj ) − Φaj+1 (Aj) Tr((aj+1I − Aj )−2(I − Aj)) (1 − √ǫ)Tr(I − Aj). Now let S′ ⊆ {1, . . . , n} be the set of indices which have not yet been used. Observe that hAu, ui = Tr(A(uu∗)) and that Pi∈S ′ uiu∗i = I − Pj = I − Aj. Thus + h(aj+1I − Aj)−1ui, uii(cid:19) + Tr((aj+1I − Aj)−1(I − Aj)) Tr(I − Aj) Xi∈S ′(cid:18)h(aj+1I − Aj)−2ui, uii Φaj (Aj ) − Φaj+1 (Aj) Tr((aj+1I − Aj)−2(I − Aj )) Φaj (Aj) − Φaj+1 (Aj ) (1 − √ǫ)Tr(I − Aj) + 1 ǫ 1 Tr(I − Aj). ǫ d=1 uid u∗id 1 √ǫ ≤ = = But 1 ǫ Tr(I − Aj ) = 1 ǫ (m − Tr(Aj )) = n − j is exactly the number of elements of S′. So there must exist some i ∈ S′ for which h(aj+1I − Aj )−2ui, uii Φaj (Aj) − Φaj+1 (Aj) + h(aj+1I − Aj)−1ui, uii ≤ 1. 8 CHARLES A. AKEMANN AND NIK WEAVER Therefore, by Lemma A.1, choosing uij+1 = ui allows the inductive construction to proceed. (cid:3) Theorem 0.5 follows by taking m = dim(F ), ǫ = m/G, and ui = Qei for 1 ≤ i ≤ n = G, and identifying F with Cm. Letting P be the orthogonal projection onto span{ei : i ∈ S}, we then have kP Qk2 = kQP Qk = kPi∈S uiu∗ik. References [1] C. Akemann and N. Weaver, A Lyapunov-type Theorem from Kadison-Singer, Bull. Lond. Math. Soc. 46 (2014), 517-524. [2] J. Bourgain, N. Katz, and T. Tao, A Sum-Product Estimate in Finite Fields and Applications, GAFA Geometric and Functional Analysis 14 (2004) 27-57. [3] P. Branden, Hyperbolic Polynomials and the Marcus-Spielman-Shrivastava Theorem, arXiv:1412.0245v1. [4] E. J. Cand`es, J. Romberg, and T. Tao, Robust uncertainty principles: exact signal reconstruction from highly incomplete frequency information, IEEE Trans. Inform. Theory 52 (2006), 489-509. [5] D. L. Donoho, P.B. Stark, Uncertainty principles and signal recovery, SIAM J. Appl. Math. 49 (1989), 906-931. [6] A. W. Marcus, D. A. Spielman, and N. Srivastava, Interlacing families II: mixed char- acteristic polynomials and the Kadison-Singer problem, arXiv:1306.3969. (To appear in Annals of Math, 2015) [7] -- -- -- , Ramanujan graphs and the solution of the Kadison-Singer problem, arXiv:1408.4421. [8] N. Srivastava, Spectral Sparsification and Restricted Invertibility, Ph.D. Thesis, Yale University (2010). [9] T. Tao, An Uncertainty Principle for Cyclic Groups of Prime Order, Mathematics Research Letters 12, (2005) 121-127. [10] A. Terras, Fourier Analysis on Finite Groups and Applications. London Mathematical Society Student Texts, 43. Cambridge University Press, Cambridge (1999). [11] Y. Yang, X. Nu, and C. Xu, Theory and Applications of Higher Dimensional Hadamard Matrices (2 ed.), CRC Press, Boca Raton, 2010. Department of Mathematics, UCSB, Santa Barbara, CA 93106, Department of Math- ematics, Washington University, Saint Louis, MO 63130 E-mail address: [email protected], [email protected]
1208.1883
1
1208
2012-08-09T12:04:03
Gevrey functions and ultradistributions on compact Lie groups and homogeneous spaces
[ "math.FA", "math.RT" ]
In this paper we give global characterisations of Gevrey-Roumieu and Gevrey-Beurling spaces of ultradifferentiable functions on compact Lie groups in terms of the representation theory of the group and the spectrum of the Laplace-Beltrami operator. Furthermore, we characterise their duals, the spaces of corresponding ultradistributions. For the latter, the proof is based on first obtaining the characterisation of their $\alpha$-duals in the sense of Koethe and the theory of sequence spaces. We also give the corresponding characterisations on compact homogeneous spaces.
math.FA
math
GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS AND HOMOGENEOUS SPACES APARAJITA DASGUPTA AND MICHAEL RUZHANSKY Abstract. In this paper we give global characterisations of Gevrey-Roumieu and Gevrey-Beurling spaces of ultradifferentiable functions on compact Lie groups in terms of the representation theory of the group and the spectrum of the Laplace- Beltrami operator. Furthermore, we characterise their duals, the spaces of corre- sponding ultradistributions. For the latter, the proof is based on first obtaining the characterisation of their α-duals in the sense of Kothe and the theory of sequence spaces. We also give the corresponding characterisations on compact homogeneous spaces. 1. Introduction The spaces of Gevrey ultradifferentiable functions are well-known on Rn and their characterisations exists on both the space-side and the Fourier transform side, lead- ing to numerous applications in different areas. The aim of this paper is to obtain global characterisations of the spaces of Gevrey ultradifferentiable functions and of the spaces of ultradistributions using the eigenvalues of the Laplace-Beltrami oper- ator LG (Casimir element) on the compact Lie group G. We treat both the cases of Gevrey-Roumieu and Gevrey-Beurling functions, and the corresponding spaces of ultradistributions, which are their topological duals with respect to their inductive and projective limit topologies, respectively. If M is a compact homogeneous space, let G be its motion group and H a stationary subgroup at some point, so that M ≃ G/H. Our results on the motion group G will yield the corresponding characterisations for Gevrey functions and ultradistributions on the homogeneous space M. Typical examples are the real spheres Sn = SO(n + 1)/SO(n), complex spheres (complex projective spaces) CPn = SU(n + 1)/SU(n), or quaternionic projective spaces HPn. Working in local coordinates and treating G as a manifold the Gevrey(-Roumieu) class γs(G), s ≥ 1, is the space of functions φ ∈ C ∞(G) such that in every local coordinate chart its local representative, say ψ ∈ C ∞(Rn), is such that there exist constants A > 0 and C > 0 such that for all multi-indices α, we have that ∂αψ(x) ≤ CAα (α!)s Date: September 24, 2018. 1991 Mathematics Subject Classification. Primary 46F05; Secondary 22E30. Key words and phrases. Gevrey spaces, ultradistributions, Fourier transform, compact Lie groups, homogenous spaces. The first author was supported by the Grace-Chisholm Young Fellowship from London Mathemat- ical Society. The second author was supported by the EPSRC Leadership Fellowship EP/G007233/1. 1 2 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY holds for all x ∈ Rn. By the chain rule one readily sees that this class is invariantly defined on (the analytic manifold) G for s ≥ 1. For s = 1 we obtain the class of analytic functions. This behaviour can be characterised on the Fourier side by being equivalent to the condition that there exist B > 0 and K > 0 such that bψ(η) ≤ Ke−Bhηi1/s holds for all η ∈ Rn. We refer to [5] for the extensive analysis of these spaces and their duals in Rn. However, such a local point of view does not tell us about the global properties of φ such as its relation to the geometric or spectral properties of the group G, and this is the aim of this paper. The characterisations that we give are global, i.e. they do not refer to the localisation of the spaces, but are expressed in terms of the behaviour of the global Fourier transform and the properties of the global Fourier coefficients. Such global characterisations will be useful for applications. For example, the Cauchy problem for the wave equation (1.1) ∂2 t u − a(t)LGu = 0 is well-posed, in general, only in Gevrey spaces, if a(t) becomes zero at some points. However, in local coordinates (1.1) becomes a second order equation with space- dependent coefficients and lower order terms, the case when the well-posedness results are, in general1, not available even on Rn. At the same time, in terms of the group Fourier transform the equation (1.1) is basically constant coefficients, and the global characterisation of Gevrey spaces together with an energy inequality for (1.1) yield the well-posedness result. We will address this and other applications elsewhere, but we note that in these problems both types of Gevrey spaces appear naturally, see e.g. [3] for the Gevrey-Roumieu ultradifferentiable and Gevrey-Beurling ultradistributional well-posedness of weakly hyperbolic partial differential equations in the Euclidean space. In Section 2 we will fix the notation and formulate our results. We will also recall known (easy) characterisations for other spaces, such as spaces of smooth functions, distributions, or Sobolev spaces over L2. The proof for the characterisation of Gevrey spaces will rely on the harmonic analysis on the group, the family of spaces ℓp(bG) on the unitary dual introduced in [8], and to some extent on the analysis of globally defined matrix-valued symbols of pseudo-differential operators developed in [8, 9]. The analysis of ultradistributions will rely on the theory of sequence spaces (echelon and co-echelon spaces), see e.g. Kothe [6], Ruckle [7]. Thus, we will first give charac- terisations of the so-called α-duals of the Gevrey spaces and then show that α-duals and topological duals coincide. We also prove that both Gevrey spaces are perfect spaces, i.e. the α-dual of its α-dual is the original space. This is done in Section 4, and the ultradistributions are treated in Section 5. We note that the case of the periodic Gevrey spaces, which can be viewed as spaces on the torus Tn, has been characterised by the Fourier coefficients in [11]. However, 1The result of Bronshtein [1] holds but is, in general, not optimal for some types of equations or does not hold for low regularity a(t). GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 3 that paper stopped short of characterising the topological duals (i.e. the correspond- ing ultradistributions), so already in this case our characterisation in Theorem 2.5 appears to be new. In the estimates throughout the paper the constants will be denoted by letter C If we want to emphasise the which may change value even in the same formula. change of the constant, we may use letters like C ′, A1, etc. We first fix the notation and recall known characterisations of several spaces. We refer to [8] for details on the following constructions. 2. Results classes of) continuous irreducible unitary representations of G. Since G is compact, Let G be a compact Lie group of dimension n. Let bG denote the set of (equivalence bG is discrete. For [ξ] ∈ bG, by choosing a basis in the representation space of ξ, we can view ξ as a matrix-valued function ξ : G → Cdξ×dξ , where dξ is the dimension of the representation space of ξ. For f ∈ L1(G) we define its global Fourier transform at ξ by bf (ξ) =ZG f (x) = X[ξ]∈ bG f (x)ξ(x)∗dx, dξ Tr(cid:16)ξ(x)bf (ξ)(cid:17). where dx is the normalised Haar measure on G. The Peter-Weyl theorem implies the Fourier inversion formula (2.1) For each [ξ] ∈ bG, the matrix elements of ξ are the eigenfunctions for the Laplace- Beltrami operator LG with the same eigenvalue which we denote by −λ2 −LGξij(x) = λ2 [ξ]ξij(x), for all 1 ≤ i, j ≤ dξ. [ξ], so that Different spaces on the Lie group G can be characterised in terms of comparing the Fourier coefficients of functions with powers of the eigenvalues of the Laplace- Beltrami operator. We denote hξi = (1 + λ2 [ξ])1/2, the eigenvalues of the elliptic first-order pseudo-differential operator (I − LG)1/2. Then, it is easy to see that f ∈ C ∞(G) if and only if for every M > 0 there exists C > 0 such that kbf (ξ)kHS ≤ Chξi−M , and u ∈ D′(G) if and only if there exist M > 0 and C > 0 such that kbu(ξ)kHS ≤ ChξiM , where we define bu(ξ)ij = u(ξji), 1 ≤ i, j ≤ dξ. For this and other occasions, we can write this as bu(ξ) = u(ξ∗) in the matrix notation. The appearance of the Hilbert-Schmidt norm is natural in view of the Plancherel identity so that (f, g)L2(G) = X[ξ]∈ bG kf kL2(G) =X[ξ]∈ bG dξ Tr(cid:16)bf (ξ)bg(ξ)∗(cid:17), HS dξkbf (ξ)k2 =: kbf kℓ2( bG) 1/2 4 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY It is convenient to use the sequence space Σ = {σ = (σ(ξ))[ξ]∈ bG : σ(ξ) ∈ Cdξ×dξ}. can be taken as the definition of the space ℓ2(bG). Here, of course, kAkHS =pTr(AA∗). In [8], the authors introduced a family of spaces ℓp(bG), 1 ≤ p < ∞, by saying that σ ∈ Σ belongs to ℓp(bG) if the norm kσkℓp( bG) :=X[ξ]∈ bG HS if finite. There is also the space ℓ∞(bG) for which the norm − 1 ξ kσ(ξ)kHS kσ(ξ)kp p( 2 ξ d (2.2) 2) p − 1 1/p kσkℓ∞( bG) := sup [ξ]∈ bG d 2 is finite. These are interpolation spaces for which the Hausdorff-Young inequality holds, in particular, we have (2.3) kbf kℓ∞( bG) ≤ kf kL1(G) and kF −1σkL∞(G) ≤ kσkℓ1( bG), with (F −1σ)(x) = P[ξ]∈ bG dξ Tr (ξ(x)σ(ξ)). We refer to [8, Chapter 10] for further details on these spaces. Usual Sobolev spaces on G as a manifold, defined by locali- sations, can be also characterised by the global condition (2.4) f ∈ H t(G) if and only if hξitbf (ξ) ∈ ℓ2(bG). For a multi-index α = (α1, . . . , αn), we define α = α1 + · · · + αn and α! = α1! · · · αn!. We will adopt the convention that 0! = 1 and 00 = 1. 1 ≤ j ≤ α, and Pj:Yj=Xk Let X1, . . . , Xn be a basis of the Lie algebra of G, normalised in some way, e.g. with respect to the Killing form. For a multi-index α = (α1, . . . , αn), we define the left- invariant differential operator of order α, ∂α := Y1 · · · Yα, with Yj ∈ {X1, · · · , Xn}, 1 = αk for every 1 ≤ k ≤ n. It means that ∂α is a composition of left-invariant derivatives with respect to vectors X1, · · · , Xn, such that each Xk enters ∂α exactly αk times. There is a small abuse of notation here since we do not specify in the notation ∂α the order of vectors X1, · · · , Xn entering in ∂α, but this will not be important for the arguments in the paper. The reason we define ∂α in this way is to take care of the non-commutativity of left-invariant differential operators corresponding to the vector fields Xk. We will distinguish between two families of Sobolev spaces over L2. The first one (2.5) The second one is defined for k ∈ N0 ≡ N ∪ {0} by is defined by H t(G) =(cid:8)f ∈ L2(G) : (I − LG)t/2f ∈ L2(G)(cid:9) with the norm kf kH t(G) := k(I − LG)t/2f kL2(G) = khξitbf (ξ)kℓ2( bG). W k,2 = ∂αf L2(G) < ∞ f ∈ L2(G) : kf kW k,2 := Xα≤k . GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 5 Obviously, H k ≃ W k,2 for any k ∈ N0 but for us the relation between norms will be of importance, especially as k will tend to infinity. Let 0 < s < ∞. We first fix the notation for the Gevrey spaces and then formulate the results. In the definitions below we allow any s > 0, and the characterisation of α-duals in the sequel will still hold. However, when dealing with ultradistributions we will be restricting to s ≥ 1. Definition 2.1. Gevrey-Roumieu(R) class γs(G) is the space of functions φ ∈ C ∞(G) for which there exist constants A > 0 and C > 0 such that for all multi-indices α, we have (2.6) ∂αφL∞ ≡ sup x∈G ∂αφ(x) ≤ CAα (α!)s. Functions φ ∈ γs(G) are called ultradifferentiable functions of Gevrey-Roumieu class of order s. For s = 1 we obtain the space of analytic functions, and for s > 1 the space of Gevrey-Roumieu functions on G considered as a manifold, by saying that the function is in the Gevrey-Roumieu class locally in every coordinate chart. The same is true for the other Gevrey space: Definition 2.2. Gevrey-Beurling(B) class γ(s)(G) is the space of functions φ ∈ C ∞(G) such that for every A > 0 there exists CA > 0 so that for all multi-indices α, we have ∂αφL∞ ≡ sup x∈G ∂αf (x) ≤ CAAα (α!)s. Functions φ ∈ γ(s)(G) are called ultradifferentiable functions of Gevrey-Beurling class of order s. Theorem 2.3. Let 0 < s < ∞. (R) We have φ ∈ γs(G) if and only if there exist B > 0 and K > 0 such that holds for all [ξ] ∈ bG. (2.7) (2.8) bφ(ξ)HS ≤ Ke−Bhξi1/s bφ(ξ)HS ≤ KBe−Bhξi1/s (B) We have φ ∈ γ(s)(G) if and only if for every B > 0 there exists KB > 0 such that Expressions appearing in the definitions can be taken as seminorms, and the spaces are equipped with the inductive and projective topologies, respectively2. We now turn to ultradistributions. holds for all [ξ] ∈ bG. Definition 2.4. The space of continuous linear functionals on γs(G)(cid:0)or γ(s)(G)(cid:1) is (s)(G)(cid:17) , respec- called the space of ultradistributions and is denoted by γ′ tively. s(G)(cid:16)or γ′ 2See also Definition 5.1 for an equivalent formulation. 6 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY For any v ∈ γ′ s(G)(cid:16)or γ′ (s)(G)(cid:17), for [ξ] ∈ bG, we define the Fourier coefficients bv(ξ) := hv, ξ∗i ≡ v(ξ∗). These are well-defined since G is compact and hence ξ(x) are actually analytic. Theorem 2.5. Let 1 ≤ s < ∞. (R) We have v ∈ γ′ s(G) if and only if for every B > 0 there exists KB > 0 such that 1 s kbv(ξ)kHS ≤ KBeBhξi (s)(G) if and only if there exist B > 0 and KB > 0 such that (2.9) (2.9) (B) We have v ∈ γ′ holds for all [ξ] ∈ bG. holds for all [ξ] ∈ bG. The proof of Theorem 2.5 follows from the characterisation of α-duals of3 the Gevrey spaces in Theorem 4.2 and the equivalence of the topological duals and α- duals in Theorem 5.2. The result on groups implies the corresponding characterisation on compact homo- geneous spaces M. First we fix the notation. Let G be a compact motion group of M and let H be the stationary subgroup of some point. Alternatively, we can start with a compact Lie group G with a closed subgroup H. The homogeneous space M = G/H is an analytic manifold in a canonical way (see, for example, [2] or [10] as textbooks on this subject). We normalise measures so that the measure on H is a probability one. Typical examples are the spheres Sn = SO(n + 1)/SO(n) or complex spheres (complex projective spaces) PCn = SU(n + 1)/SU(n). We denote by bG0 the subset of bG of representations that are class I with respect to the subgroup H. This means that [ξ] ∈ bG0 if ξ has at least one non-zero invariant vector a with respect to H, i.e. that ξ(h)a = a for all h ∈ H. Let Hξ denote the representation space of ξ(x) : Hξ → Hξ and let Bξ be the space of these invariant vectors. Let kξ = dim Bξ. We fix an orthonormal basis of Hξ so that its first kξ vectors are the basis of Bξ. The matrix elements ξij(x), 1 ≤ j ≤ kξ, are invariant under the right shifts by H. We refer to [12] for the details of these constructions. We can identify Gevrey functions on M = G/H with Gevrey functions on G which are constant on left cosets with respect to H. Here we will restrict to s ≥ 1 to see the equivalence of spaces using their localisation. This identification gives rise to the corresponding identification of ultradistributions. Thus, for a function f ∈ γs(M) we can recover it by the Fourier series of its canonical lifting ef (g) := f (gH) to G, ef ∈ γs(G), and the Fourier coefficients satisfy bef (ξ) = 0 for all representations with [ξ] 6∈ bG0. Also, for class I representations [ξ] ∈ bG0 we havebef (ξ)ij = 0 for i > kξ. With this, we can write the Fourier series of f (or of ef , but as we said, from now ξij of the representations ξ, [ξ] ∈ bG0, with respect to the subgroup H. Namely, the on we will identify these and denote both by f ) in terms of the spherical functions 3The characterisation of α-duals is valid for all 0 < s < ∞. GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 7 Fourier series (2.1) becomes (2.10) f (x) = X[ξ]∈ bG0 dξ dξXi=1 kξXj=1 bf (ξ)jiξij(x). In view of this, we will say that the collection of Fourier coefficients {bφ(ξ)ij : [ξ] ∈ bG, 1 ≤ i, j ≤ dξ} is of class I with respect to H if bφ(ξ)ij = 0 whenever [ξ] 6∈ bG0 or i > kξ. By the above discussion, if the collection of Fourier coefficients is of class I with respect to H, then the expressions (2.1) and (2.10) coincide and yield a function f such that f (xh) = f (h) for all h ∈ H, so that this function becomes a function on the homogeneous space G/H. The same applies to (ultra)distributions with the standard distributional interpretation. With these identifications, Theorem 2.3 immediately implies Theorem 2.6. Let 1 ≤ s < ∞. (R) We have φ ∈ γs(G/H) if and only if its Fourier coefficients are of class I with respect to H and, moreover, there exist B > 0 and K > 0 such that (B) We have φ ∈ γ(s)(G) if and only if its Fourier coefficients are of class I with respect to H and, moreover, for every B > 0 there exists KB > 0 such that (2.11) holds for all [ξ] ∈ bG0. (2.12) holds for all [ξ] ∈ bG0. bφ(ξ)HS ≤ Ke−Bhξi1/s bφ(ξ)HS ≤ KBe−Bhξi1/s It would be possible to extend Theorem 2.6 to the range 0 < s < ∞ by adopting Definition 2.1 starting with a frame of vector fields on M, but instead of obtaining the result immediately from Theorem 2.3 we would have to go again through arguments similar to those used to prove Theorem 2.3. Since we are interested in characterising the standard invariantly defined Gevrey spaces we decided not to lengthen the proof in this way. On the other hand, it is also possible to prove the characterisations on homogeneous spaces G/H first and then obtain those on the group G by taking H to be trivial. However, some steps would become more technical since we would have to deal with frames of vector fields instead of the basis of left-invariant vector fields on G, and elements of the symbolic calculus used in the proof would become more complicated. We also have the ultradistributional result following from Theorem 2.5. Theorem 2.7. Let 1 ≤ s < ∞. (R) We have v ∈ γ′ respect to H and, moreover, for every B > 0 there exists KB > 0 such that s(G/H) if and only if its Fourier coefficients are of class I with (2.13) holds for all [ξ] ∈ bG0. (B) We have v ∈ γ′ 1 s kbv(ξ)kHS ≤ KBeBhξi (s)(G/H) if and only if its Fourier coefficients are of class I with 8 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY respect to H and, moreover, there exist B > 0 and KB > 0 such that (2.13) holds for Finally, we remark that in the harmonic analysis on compact Lie groups sometimes all [ξ] ∈ bG0. another version of ℓp(bG) spaces appears using Schatten p-norms. However, in the context of Gevrey spaces and ultradistributions eventual results hold for all such norms. Indeed, given our results with the Hilbert-Schmidt norm, by an argument similar to that of Lemma 3.2 below, we can put any Schatten norm k·kSp, 1 ≤ p ≤ ∞, instead of the Hilbert-Schmidt norm k · kHS in any of our characterisations and they still continue to hold. 3. Gevrey classes on compact Lie groups We will need two relations between dimensions of representations and the eigen- values of the Laplace-Beltrami operator. On one hand, it follows from the Weyl character formula that (3.1) with the latter4 also following directly from the Weyl asymptotic formula for the eigenvalue counting function for LG, see e.g. [8, Prop. 10.3.19]. This implies, in particular, that for any 0 ≤ p < ∞ and any s > 0 and B > 0 we have 2 ≤ Chξi dξ ≤ Chξi 2 , n−rankG n (3.2) ξe−Bhξi1/s dp < ∞. sup [ξ]∈ bG d2 On the other hand, the following convergence for the series will be useful for us: Lemma 3.1. We haveP[ξ]∈ bG d2 Proof. We notice that for the δ-distribution at the unit element of the group,bδ(ξ) = Idξ is the identity matrix of size dξ × dξ. Hence, in view of (2.4) and (2.5), we can write ξ hξi−2t < ∞ if and only if t > n 2 . ξhξi−2t = X[ξ]∈ bG dξhξi−2tkbδ(ξ)k2 By using the localisation of H −t(G) this is finite if and only if t > n/2. X[ξ]∈ bG We denote by bG∗ the set of representations from bG excluding the trivial represen- tation. For [ξ] ∈ bG, we denote ξ := λξ ≥ 0, the eigenvalue of the operator (−LG)1/2 corresponding to the representation ξ. For [ξ] ∈ bG∗ we have ξ > 0 (see e.g. [4]), and for [ξ] ∈ bG\bG∗ we have ξ = 0. From the definition, we have ξ ≤ hξi. On the other 1 > 0 be the smallest positive eigenvalue of −LG. Then, for [ξ] ∈ bG∗ we hand, let λ2 have λξ ≥ λ1, implying HS = k(I − LG)−t/2δk2 L2(G) = kδk2 H −t(G). (cid:3) 1 + λ2 so that altogether we record the inequality (3.3) ξ ≤ hξi ≤(cid:18)1 + 4Namely, the inequality dξ ≤ Chξi n 2 . + 1(cid:19)λ2 ξ, λ2 1 ξ ≤(cid:18) 1 1(cid:19)1/2 1 λ2 ξ, for all [ξ] ∈ bG∗. GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 9 We will need the following simple lemma which we prove for completeness. Let a ∈ Cd×d be a matrix, and for 1 ≤ p < ∞ we denote by ℓp(C) the space of such matrices with the norm kakℓp(C) = dXi,j=1 aijp!1/p , and for p = ∞, kakℓ∞(C) = sup1≤i,j≤d aij. We note that kakℓ2(C) = kakHS. We adopt the usual convention c ∞ = 0 for any c ∈ R. Lemma 3.2. Let 1 ≤ p < q ≤ ∞ and let a ∈ Cd×d. Then we have (3.4) kakℓp(C) ≤ d2( 1 p − 1 q )kakℓq(C) and kakℓq(C) ≤ d 2 q kakℓp(C). Proof. For q < ∞, we apply Holder's inequality with r = q p and r′ = q q−p to get kakp ℓp(C) = aijp ≤ dXi,j=1 aijpr!1/r dXi,j=1 dXi,j=1 1!1/r′ = kakp ℓq(C)d2 q−p q , implying (3.4) for this range. Conversely, we have kakq ℓq(C) = aijq ≤ dXi,j=1 dXi,j=1 kakq ℓp(C) = d2kakq ℓp(C), proving the other part of (3.4) for this range. For q = ∞, we have kakℓp(C) ≤ ≤ kakℓ∞(C)d2/p. Conversely, we have trivially kakℓ∞(C) ≤ kakℓp(C), (cid:3) (cid:16)Pd completing the proof. i,j=1 kakp ℓ∞(C)(cid:17)1/p We observe that the Gevrey spaces can be described in terms of L2-norms, and this will be useful to us in the sequel. Lemma 3.3. We have φ ∈ γs(G) if and only if there exist constants A > 0 and C > 0 such that for all multi-indices α we have (3.5) k∂αφkL2 ≤ CAα (α!)s . We also have φ ∈ γ(s)(G) if and only if for every A > 0 there exists CA > 0 such that for all multi-indices α we have k∂αφkL2 ≤ CAAα (α!)s . Proof. We prove the Gevrey-Roumieu case (R) as the Gevrey-Beurling case (B) is similar. For φ ∈ γs(G), (3.5) follows in view of the continuous embedding L∞(G) ⊂ L2(G) with kf kL2 ≤ kf kL∞ since the measure is normalised. 10 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY Now suppose that for φ ∈ C ∞(G) we have (3.5). In view of (2.3), and using Lemma 3.1 with an integer k > n/2, we obtain5 kφkL∞ ≤ X[ξ]∈ bG ≤ X[ξ]∈ bG ≤ Ck Xβ≤k d3/2 ξ kbφ(ξ)kHS dξkbφ(ξ)k2 k∂βφkL2, ≤ Ck(I − LG)k/2φkL2 HShξi2k 1/2X[ξ]∈ bG 1/2 d2 ξhξi−2k with constant Ck depending only on G. Consequently we also have (3.6) Using the inequalities k∂αφkL∞ ≤ Ck Xβ≤k k∂α+βφkL2. (3.7) α! ≤ α!, α! ≤ nαα! and (α + k)! ≤ 2α+kk!α!, in view of (3.6) and (3.5) we get ∂αφL∞ ≤ CkAα+k Xβ≤k ≤ CkAα+k Xβ≤k ((α + β)!)s ((α + k)!)s ≤ C ′ ≤ C ′′ ≤ C ′′ kAα+k(2α+kk!)s(α!)s k Aα k Aα 1 (nαα!)s 2 (α!)s, with constants C ′′ the proof. k and A2 independent of α, implying that φ ∈ γs(G) and completing (cid:3) The following proposition prepares the possibility to passing to the conditions formulated on the Fourier transform side. Proposition 3.4. We have φ ∈ γs(G) if and only if there exist constants A > 0 and C > 0 such that (3.8) (−LG)k φL∞ ≤ CA2k ((2k)!)s holds for all k ∈ N0. Also, φ ∈ γ(s)(G) if and only if for every A > 0 there exists CA > 0 such that for all k ∈ N0 we have (−LG)k φL∞ ≤ CAA2k ((2k)!)s . 5Note that this can be adopted to give a simple proof of the Sobolev embedding theorem. GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 11 Proof. We prove the Gevrey-Roumieu case (3.8) and indicate small additions to the argument for γ(s)(G). Thus, let φ ∈ γs(G). Recall that by the definition there exist some A > 0, C > 0 such that for all multi-indices α we have ∂αφ(x) ≤ CAα (α!)s. ∂αφL∞ = sup x∈G We will use the fact that for the compact Lie group G the Laplace-Beltrami operator LG is given by LG = X 2 n, where Xi, i = 1, 2, . . . , n, is a set of left-invariant vector fields corresponding to a normalised basis of the Lie algebra of G. Then by the multinomial theorem6 and using (3.7), with Yj ∈ {X1, . . . , Xn}, 1 ≤ j ≤ α, we can estimate 2 + ... + X 2 1 + X 2 1 . . . Y 2 k! α!(cid:12)(cid:12)Y 2 (−LG)kφ(x) ≤ C Xα=k ≤ C Xα=k ≤ CA2k[(2k)!]s Xα=k k! α! [(2α)!]sA2α αφ(x)(cid:12)(cid:12) k!nα α! ≤ C1A2k[(2k)!]snkkn−1 ≤ C2A2k 1 [(2k)!]s, (3.9) with A1 = 2nA, implying (3.8). For the Gevrey-Beurling case γ(s)(G), we observe that we can obtain any A1 > 0 in (3.9) by using A = A1 2n in the Gevrey estimates for φ ∈ γ(s)(G). Conversely, suppose φ ∈ C ∞(G) is such that the inequalities (3.8) hold. First we note that for α = 0 the estimate (2.7) follows from (3.8) with k = 0, so that we can assume α > 0. Following [9], we define the symbol of ∂α to be σ∂α(ξ) = ξ(x)∗∂αξ(x), and we have σ∂α(ξ) ∈ Cdξ×dξ is independent of x since ∂α is left-invariant. For the in-depth analysis of symbols and symbolic calculus for general operators on G we refer to [8, 9] but we will use only basic things here. In particular, we have ∂αφ(x) = X[ξ]∈ bG dξ Tr(cid:16)ξ(x)σ∂α(ξ)bφ(ξ)(cid:17). First we calculate the operator norm σ∂α(ξ)op of the matrix multiplication by σ∂α(ξ). Since ∂α = Y1 · · · Yα and Yj ∈ {X1, . . . , Xn} are all left-invariant, we have σ∂α = σY1 · · · σYα, so that we get kσ∂α(ξ)kop ≤ kσX1(ξ)kα1 op · · · kσXn(ξ)kαn op . Now, since Xj are operators of the first order, one can show (see e.g. [9, Lemma 8.6], or [8, Section 10.9.1] for general arguments) that σXj (ξ)op ≤ Cjhξi for some 6The form in which we use it is adapted to non-commutativity of vector fields. Namely, although the coefficients are all equal to one in the non-commutative form, the multinomial coefficient appears once we make a choice for α = (α1, · · · , αn). 12 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY constants Cj, j = 1, . . . , n. Let C0 = supj Cj + 1, then we have (3.10) kσ∂α(ξ)kop ≤ C α 0 hξiα. Let us define σPα ∈ Σ by setting σPα(ξ) := ξ−2kσ∂α(ξ) for [ξ] ∈ bG∗, and by σPα(ξ) := 0 for [ξ] ∈ bG\bG∗. This gives the corresponding operator (3.11) (3.12) 1 λ2 . From (3.10) we obtain Using (3.11) and the Plancherel identity, we estimate Now, for [ξ] ∈ bG∗, from (3.3) we have Together with (3.12), and the trivial estimate for [ξ] ∈ bG\bG∗, we obtain kσPα(ξ)kop ≤ C α ξ−2k ≤ C 2k 0 C 2k (3.13) kσPα(ξ)kop ≤ C α (Pαφ)(x) = X[ξ]∈ bG dξ Tr(cid:16)ξ(x)σPα(ξ)bφ(ξ)(cid:17). 0 hξiαξ−2k for all [ξ] ∈ bG∗. 1 hξi−2k, C1 =(cid:18)1 + 1(cid:19)1/2 1 hξiα−2k for all [ξ] ∈ bG. Pαφ(x) ≤ X[ξ]∈ bG dξkξ(x)σPα(ξ)kHSkbφ(ξ)kHS ≤ X[ξ]∈ bG HS 1/2X[ξ]∈ bG dξkbφ(ξ)k2 = kφkL2X[ξ]∈ bG op 1 X[ξ]∈ bG ξhξi−2(2k−α) Pαφ(x) ≤ kφkL2C α ξkσPα(ξ)k2 d2 dξkσPα(ξ)k2 From this and (3.13) we conclude that 0 C 2k d2 1/2 . 1/2 . opkξ(x)k2 1/2 HS Now, in view of Lemma 3.1 the series on the right hand side converges provided that 2k − α > n/2. Therefore, for 2k − α > n/2 we obtain (3.14) kPαφkL2 ≤ CC 2k 2 kφkL2, with some C and C2 = C0C1 independent of k and α. We note that here we used that α ≤ 2k and that we can always have C0 ≥ 1. We now observe that from the definition of σPα we have (3.15) σ∂α(ξ) = σPα(ξ)ξ2k for all [ξ] ∈ bG∗. On the other hand, since we assumed α 6= 0, for [ξ] ∈ bG\bG∗ we have σ∂α(ξ) = ξ(x)∗∂αξ(x) = 0, so that (3.15) holds true for all [ξ] ∈ bG. This implies GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 13 that in the operator sense, we have ∂α = Pα ◦ (−LG)k. Therefore, from this relation and (3.14), for α < 2k − n/2, we get k∂αφk2 L2 = kPα ◦ (−LG)kφk2 L2 ≤ CC 4k (−LG)kφ(x)2dx 2 ZG ≤ C ′C 4k ≤ C ′A4k 2 A4k((2k)!)2s 1 ((2k)!)2s, where we have used the assumption (3.8), and with C ′ and A1 = C2A independent of k and α. Hence we have k∂αφkL2 ≤ CA2k 1 ((2k)!)s for all α < 2k −n/2. Then, for every β, by the above argument, taking an integer k such that β + 4n ≥ 2k > β + n/2, if A1 ≥ 1, we obtain k∂βφkL2 ≤ CAβ+4n 1 ((β + 4n)!)s ≤ C ′Aβ ≤ C ′′Aβ 2 (β!)s, 1 (cid:0)2β+4n(4n)!β!(cid:1)s in view of inequalities (3.7). By Lemma 3.3 it follows that φ ∈ γs(G). If A1 < 1 (in the case of γ(s)(G)), we estimate k∂βφkL2 ≤ CAβ+n/2 1 ((β + 4n)!)s ≤ C ′′Aβ 3 (β!)s by a similar argument. The relation between constants, namely A1 = C2A and A3 = 2nA1, implies that the case of γ(s)(G) also holds true. (cid:3) We can now pass to the Fourier transform side. Lemma 3.5. For φ ∈ γs(G), there exist constants C > 0 and A > 0 such that (3.16) ξ−2mA2m ((2m)!)s ξ bφ(ξ)HS ≤ Cd1/2 bφ(ξ)HS ≤ CAd1/2 ξ CA > 0 such that holds for all m ∈ N0 and [ξ] ∈ bG∗. Also, for φ ∈ γ(s)(G), for every A > 0 there exists holds for all m ∈ N0 and [ξ] ∈ bG∗. Fourier transform is a bounded linear operator from L1(G) to l∞(bG), see (2.3), and Proof. We will treat the case γs since γ(s) is analogous. Using the fact that the using Proposition 3.4, we obtain ξ−2mA2m ((2m)!)s ξ2mbφ(ξ)l∞( bG) ≤ ZG ≤ CA2m ((2m)!)s (−LG)m φ(x)dx for all [ξ] ∈ bG and m ∈ N0. Recalling the definition of ℓ∞(bG) in (2.2) we obtain (3.16). (cid:3) We can now prove Theorem 2.3. Proof of Theorem 2.3. (R) "Only if" part. Let φ ∈ γs(G). Using k! ≤ kk and Lemma 3.5 we get (3.17) bφ(ξ)HS ≤ Cd1/2 ξ inf 2m≥0 ξ−2mA2m (2m)2ms 14 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY for all [ξ] ∈ bG∗. We will show that this implies the (sub-)exponential decay in (2.7). It is known that for r > 0, we have the identity (3.18) xsxr−x = e−(s/e)r1/s . inf x>0 So for a given r > 0 there exists some x0 = x0(r) > 0 such that (3.19) inf x>0 xsx(cid:16) r 8s(cid:17)−x = xsx0 0 (cid:16) r 8s(cid:17)−x0 . We will be interested in large r, in fact we will later set r = ξ A , so we can assume that r is large. Consequently, in (3.18) and later, we can assume that x0 is sufficiently large. Thus, we can take an even (sufficiently large) integer m0 such that m0 ≤ x0 < m0 + 2. Using the trivial inequalities and (m0)sm0 r−(m0+2) ≤ xsx0 0 r−x0, r ≥ 1, (k + 2)k+2 ≤ 8kkk for any k ≥ 2, we obtain (m0 + 2)s(m0+2) r−(m0+2) ≤ 8sm0msm0 0 It follows from this, (3.18) and (3.19), that r−(m0+2) ≤ xsx0 0 (cid:16) r 8s(cid:17)−x0 . (3.20) inf 2m≥0 (2m)2smr−2m ≤ xsx0 0 (cid:16) r 8s(cid:17)−x0 = e−(s/e)( r 8s )1/s . Let now r = ξ A . From (3.17) and (3.20) we obtain kbφ(ξ)kHS ≤ Cd1/2 = Cd1/2 ξ ξ A2m ξ2m (2m)2ms r−2m (2m)2ms inf 2m≥0 inf 2m≥0 ≤ Cd1/2 = Cd1/2 ≤ Cd1/2 8s )1/s ξ e−(s/e)( r ξ e−(s/e) ξ ξ e−2Bξ1/s , 1/s 8A1/s (3.21) with 2B = s 8e 1 A1/s . From (3.2) it follows that d1/2 ξ e−Bξ1/s ≤ C. Using (3.3), we obtain (2.7) for all [ξ] ∈ bG∗. On the other hand, for trivial [ξ] ∈ bG\bG∗ the estimate (2.7) is just the condition of the boundedness. This completes the proof of the "only if" part. Now we prove the "if" part. Suppose φ ∈ C ∞(G) is such that (2.7) holds, i.e. we have bφ(ξ)HS ≤ Ke−Bhξi1/s . GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 15 The ℓ1(bG) − L∞(G) boundedness of the inverse Fourier transform in (2.3) implies k(−LG)kφkL∞(G) ≤ kξ2kbφkℓ1( bG) d3/2 ξ = X[ξ]∈ bG ≤ K X[ξ]∈ bG ≤ K X[ξ]∈ bG ξ2kbφ(ξ)HS hξi2ke−Bhξi1/s d3/2 ξ d3/2 ξ e −Bhξi 1/s 2 (cid:18)hξi2ke −Bhξi 1/s 2 (cid:19). (3.22) Now we will use the following simple inequality, m = 2k and a = B 2 , we estimate tN N ! ≤ et for t > 0. Setting later (m!)−shξim = (ahξi1/s)m m! !s a−sm ≤ a−smeahξi1/s , which implies e− B equality and (3.22) we obtain 2 hξi1/s hξi2k ≤ A2k((2k)!)s, with A = a−s = (2/B)s. Using this in- (3.23) k(−LG)kφkL∞ ≤ K X[ξ]∈ bG d3/2 ξ e −Bhξi1/s 2 A2k((2k)!)s ≤ CA2k((2k)!)s with A = 2s Therefore, φ ∈ γs(G) by Proposition 3.4. Bs , where the convergence of the series in [ξ] follows from Lemma 3.1. (B) "Only if" part. Suppose φ ∈ γ(s)(G). For any given B > 0 define A by solving A1/s . By Lemma 3.5 there exists KB > 0 such that 2B =(cid:0) s 8e(cid:1) 1 Consequently, arguing as in case (R) we get (3.21), i.e. bφ(ξ)HS ≤ KBd1/2 ξ ξ−2mA2m (2m)2ms . inf 2m≥0 kbφ(ξ)kHS ≤ KBd1/2 ξ e−2Bξ1/s for all [ξ] ∈ bG. The same argument as in the case (R) now completes the proof. "If" part. For a given A > 0 define B > 0 by solving A = 2s enough as in the case of (R), so that we get Bs and take CA big k(−LG)kφkL∞ ≤ CAA2k((2k)!)s. Therefore, φ ∈ γ(s)(G) by Proposition 3.4. (cid:3) 4. α-duals γs(G)∧ and γ(s)(G)∧, for any s, 0 < s < ∞. First we analyse α-duals of Gevrey spaces regarded as sequence spaces through their Fourier coefficients. 16 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY cients and Theorem 2.3. We denote the α-dual of such the sequence space γs(G) (or γ(s)(G)) as We can embed γs(G)(cid:0)or γ(s)(G)(cid:1) in the sequence space Σ using the Fourier coeffi- (vξ)ijbφ(ξ)ij < ∞ for all φ ∈ γs(G) [γs(G)]∧ = with a similar definition for γ(s)(G). Lemma 4.1. (R) We have v ∈ [γs (G)]∧ if and only if for every B > 0 the inequality v = (vξ)[ξ]∈ bG ∈ Σ : X[ξ]∈ bG dξXi,j=1 , e−Bhξi 1 s kvξkHS < ∞ (4.1) X[ξ]∈ bG holds for all [ξ] ∈ bG. (B) Also, we have v ∈ (cid:2)γ(s) (G)(cid:3)∧ inequality (4.1) holds for all [ξ] ∈ bG. if and only if there exists B > 0 such that the The proof of this lemma in (R) and (B) cases will be different. For (R) we can show this directly, and for (B) we employ the theory of echelon spaces by Kothe [6]. Proof. (R) "Only if" part. Let v ∈ [γs (G)]∧. For any B > 0, define φ by setting s ≤ 1 1 ξe−Bhξi s by (3.2), which implies that φ ∈ γs (G) by Theorem 2.3. Using Lemma 3.2, its Fourier coefficients to be bφ(ξ)ij := dξe−Bhξi Ce− B we obtain 2 hξi 1 s , so that kbφ(ξ)kHS = d2 X[ξ]∈ bG e−Bhξi 1 s kvξkHS ≤ X[ξ]∈ bG dξe−Bhξi 1 s kvξkℓ1(C) = X[ξ]∈ bG dξXi,j=1 by the assumption v ∈ [γs (G)]∧, proving the "only if" part. (vξ)ijbφ(ξ)ij < ∞ "If" part. Let φ ∈ γs(G). Then by Theorem 2.3 there exist some B > 0 and C > 0 such that which implies that X[ξ]∈ bG dξXi,j=1 1 s , kbφ(ξ)kHS ≤ Ce−Bhξi kvξkHSkbφ(ξ)kHS ≤ C X[ξ]∈ bG e−Bhξi 1 s kvξkHS < ∞ (vξ)ijbφ(ξ)ij ≤ X[ξ]∈ bG EB = is finite by the assumption (4.1). But this means that v ∈ [γs(G)]∧. (B) For any B > 0 we consider the so-called echelon space, v = (vξ) ∈ Σ : X[ξ]∈ bG e−Bhξi dξXi,j=1 1 s (vξ)ij < ∞ . GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 17 ∼= l∞, and it is easy to Now, by diagonal transform we have EB ∼= l1 and hence cEB check that cEB is given by for all 1 ≤ i, j ≤ dξo . cEB =nw = (wξ) ∈ Σ ∃K > 0 : By Theorem 2.3 we know that φ ∈ γ(s)(G) if and only if (cid:16)bφ(ξ)(cid:17)[ξ]∈ bG ∈ TB>0 cEB. consequently, that v ∈ γ(s)(G)∧ if and only if (vξ)[ξ]∈ bG ∈SB>0 EB. But this means Using Kothe's theory relating echelon and co-echelon spaces [6, Ch. 30.8], we have, that for some B > 0 we have (wξ)ij ≤ Ke−Bhξi1/s X[ξ]∈ bG dξXi,j=1 e−Bhξi 1 s (vξ)ij < ∞. Finally, we observe that this is equivalent to (4.1) if we use Lemma 3.2 and (3.2). (cid:3) We now give the characterisation for α-duals. Theorem 4.2. Let 0 < s < ∞. (R) We have v ∈ [γs (G)]∧ if and only if for every B > 0 there exists KB > 0 such that (4.2) vξHS ≤ KBeBhξi 1 s if and only if there exist B > 0 and KB > 0 such that Proof. We prove the case (R) only since the proof of (B) is similar. First we deal with "If" part. Let v ∈ Σ be such that (4.2) holds for every B > 0. Let ϕ ∈ γs(G). Then holds for all [ξ] ∈ bG. (B) We have v ∈(cid:2)γ(s) (G)(cid:3)∧ (4.2) holds for all [ξ] ∈ bG. by Theorem 2.3 there exist some constants A > 0 and C > 0 such that kbφ(ξ)kHS ≤ X[ξ]∈ bG kvξkHSkbφ(ξ)kHS ≤ CKB X[ξ]∈ bG so that v ∈ [γs (G)]∧. "Only if" part. Let v ∈ [γs(G)]∧ and let B > 0. Then by Lemma 4.1 we have that . Taking B = A/2 in (4.2) we get that dξXi,j=1 Ce−Ahξi1/s 2 hξi1/s e− A < ∞, (vξ)ijbφ(ξ)ij ≤ X[ξ]∈ bG X[ξ]∈ bG e−Bhξi1/s vξHS < ∞. This implies that the exists a constant KB > 0 such that e−Bhξi1/s yielding (4.2). v(ξ)HS ≤ KB, (cid:3) We now want to show that the Gevrey spaces are perfect in the sense of Kothe. We define the α−dual of [γs(G)]∧ as [\γs(G)]∧ = w = (wξ)[ξ]∈ bG ∈ Σ : X[ξ]∈ bG dξXi,j=1 (wξ)ij(vξ)ij < ∞ for all v ∈ [γs(G)]∧ , 18 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY and similarly for [γ(s)(G)]∧. First, we prove the following lemma. Lemma 4.3. (R) We have w ∈h\γs (G)i∧ 1 (4.3) eBhξi s kwξkHS < ∞. X[ξ]∈ bG if and only if there exists B > 0 such that . Let B > 0, and define (vξ)ij := if and only if for every B > 0 the series (4.3) converges. (B) We have w ∈h \γ(s) (G)i∧ (B) "Only if" part. We assume that w ∈h \γ(s) (G)i∧ 1 1 1 Proof. We first show the Beurling case as it is more straightforward. s . Then kvξkHS = d2 dξeBhξi by Theorem 4.2. Consequently, using Lemma 3.2 we can estimate s ≤ Ce2Bhξi ξeBhξi s by (3.2), which implies v ∈ [γ(s)(G)]∧ X[ξ]∈ bG eBhξi 1 s kwξkHS ≤ X[ξ]∈ bG 1 s dξeBhξi dξXi,j=1 (wξ)ij = X[ξ]∈ bG dξXi,j=1 (vξ)ij(wξ)ij < ∞, implying (4.3). "If" part. Here we are given w ∈ Σ such that for every B > 0 the series (4.3) converges. Let us take any v ∈ [γ(s)(G)]∧. By Theorem 4.2 there exist B > 0 and K > 0 such that kvξkHS ≤ KeBhξi s . Consequently, we can estimate 1 ∼= l∞, and since l∞ is a perfect sequence space, . dξXi,j=1 1 eBhξi X[ξ]∈ bG s kwξkHS < ∞ (vξ)ij ≤ KeBhξi1/s (R) For B > 0 we consider the echelon space (vξ)ij(wξ)ij ≤ X[ξ]∈ bG By diagonal transform we have DB ∼= l1, and it is given by kvξkHSkwξkHS ≤ K X[ξ]∈ bG by the assumption (4.3), which shows that w ∈h \γ(s) (G)i∧ DB =nv = (vξ) ∈ Σ ∃K > 0 : for all 1 ≤ i, j ≤ dξo . we have cDB s (wξ)ij < ∞ By Theorem 4.2 we know that γs(G)∧ =TB>0 DB, and henceh\γs (G)i∧ This means that w ∈h\γs (G)i∧ P[ξ]∈ bGPdξ X[ξ]∈ bG s kwξkHS ≤ X[ξ]∈ bG w = (wξ) ∈ Σ : X[ξ]∈ bG s kwξkℓ1(C) ≤ C X[ξ]∈ bG s (wξ)ij < ∞. Consequently, by Lemma 3.2 we get cDB = s (wξ)ij < ∞, dξXi,j=1 dξXi,j=1 i,j=1 e2Bhξi 1 e2Bhξi 1 dξeBhξi 1 eBhξi . 1 eBhξi 1 if and only if there exists B > 0 such that we have =SB>0 cDB. GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 19 completing the proof of the "only if" part. Conversely, given (4.3) for some 2B > 0, we have X[ξ]∈ bG dξXi,j=1 eBhξi 1 s (wξ)ij ≤ X[ξ]∈ bG dξeBhξi 1 s kwξkHS ≤ C X[ξ]∈ bG e2Bhξi 1 s kwξkHS < ∞, . (cid:3) implying w ∈h\γs (G)i∧ Now we can show that the Gevrey spaces are perfect spaces (sometimes called Kothe spaces). Theorem 4.4. γs(G) and γ(s)(G) are perfect spaces, that is, γs(G) = [\γs(G)]∧ and γ(s)(G) = [ \γ(s)(G)]∧. Proof. We will show this for γs(G) since the proof for γ(s)(G) is analogous. From the definition of [\γs(G)]∧ we have γs(G) ⊆ [\γs(G)]∧. We will prove the other direction, i.e., [\γs(G)]∧ ⊆ γs(G). Let w = (wξ)[ξ]∈ bG ∈ [\γs(G)]∧ and define φ(x) := X[ξ]∈ bG dξ Tr (wξξ(x)) . The series makes sense due to Lemma 4.3, and we have kbφ(ξ)kHS = kwξkHS. Now since w ∈ [\γs(G)]∧ by Lemma 4.3 there exists B > 0 such thatP[ξ]∈ bG eBhξi which implies that for some C > 0 we have s wξHS < ∞, 1 By Theorem 2.3 this implies φ ∈ γs(G). Hence γs(G) = [\γs(G)]∧, i.e. γs(G) is a perfect space. (cid:3) wξHS < C ⇒ bφ(ξ)HS ≤ Ce−Bhξi1/s . eBhξi1/s 5. Ultradistributions γ′ s(G) and γ′ (s)(G) Here we investigate the Fourier coefficients criteria for spaces of ultradistributions. The space γ′ (s)(G)) of the ultradistributions of order s is defined as the dual of γs(G) (resp. γ(s)(G)) endowed with the standard inductive limit topology of γs(G) (resp. the projective limit topology of γ(s)(G)). s(G) (resp. γ′ Definition 5.1. The space γ′ (s)(G)(cid:17) is the set of the linear forms u on γs(G)(cid:0)resp. γ(s)(G)(cid:1) such that for every ǫ > 0 there exists Cǫ (resp. for some ǫ > 0 s(G)(cid:16)resp. γ′ and C > 0) such that u(φ) ≤ Cǫ sup α ǫα(α!)−s sup x∈G (−LG)α/2φ(x) holds for all φ ∈ γs(G) (resp. φ ∈ γ(s)(G)). We can take the Laplace-Beltrami operator in Definition 5.1 because of the equiv- alence of norms given by Proposition 3.4. We recall that for any v ∈ γ′ bv(ξ) := hv, ξ∗i ≡ v(ξ∗). s(G), for [ξ] ∈ bG, we define the Fourier coefficients 20 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY We have the following theorem showing that topological and α-duals of Gevrey spaces coincide. Theorem 5.2. Let 1 ≤ s < ∞. Then v ∈ γ′ v ∈ γs(G)∧(cid:0)resp. γ(s)(G)∧(cid:1) . s(G)(cid:16)resp. γ′ (s)(G)(cid:17) if and only if Proof. (R) "If" part. Let v ∈ γs(G)∧. For any φ ∈ γs(G) define (5.1) v(φ) := X[ξ]∈ bG dξ Tr(cid:16)bφ(ξ)vξ(cid:17) . Since by Theorem 2.3 there exist some B > 0 such that bφ(ξ)HS ≤ Ce−Bhξi1/s can estimate , we X[ξ]∈ bG dξ Tr(cid:16)bφ(ξ)vξ(cid:17) ≤ X[ξ]∈ bG dξkbφ(ξ)kHSkvξkHS ≤ C X[ξ]∈ bG dξe−Bhξi1/s kvξkHS < ∞ by Lemma 4.1 and (3.2). Therefore, v(φ) in (5.1) is a well-defined linear functional on γs(G). It remains to check that v is continuous. Suppose φj → φ in γs(G) as j → ∞, that is, in view of Proposition 3.4, there is a constant A > 0 such that sup α A−α(α!)−s sup x∈G (−LG)α/2(φj(x) − φ(x)) → 0 as j → ∞. It follows that k (−LG)α/2 (φj − φ)k∞ ≤ CjAα ((α)!)s , for a sequence Cj → 0 as j → ∞. From the proof of Theorem 2.3 it follows that we then have where B > 0 and Kj → 0 as j → ∞. Hence we can estimate kbφj(ξ) −bφ(ξ)kHS ≤ Kje−Bhξi1/s , v(φj − φ) ≤ X[ξ]∈ bG ≤ Kj X[ξ]∈ bG dξkbφj(ξ) −bφ(ξ)kHSkv(ξ)kHS dξe−Bhξi1/s kvξkHS → 0 as j → ∞ since Kj → 0 as j → ∞ andP[ξ]∈ bG dξe−Bhξi1/s and (3.2). Therefore, we have v ∈ γ′ s(G). kvξkHS < ∞ by Lemma 4.1 "Only if" part. Let us now take v ∈ γ′ s(G). This means that for every ǫ > 0 there exists Cǫ such that v(φ) ≤ Cǫ sup α ǫα(α!)−s sup x∈G (−LG)α/2φ(x) GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 21 holds for all φ ∈ γs(G). So then, in particular, we have v(ξ∗ ij) ≤ Cǫ sup α ǫα(α!)−s sup x∈G (−LG)α/2ξ∗ ij(x) = Cǫ sup α ≤ Cǫ sup α = Cǫ sup α This implies ξ∗ ij(x) kξ∗(x)kHS ǫα(α!)−sξα sup x∈G ǫα(α!)−shξiα sup x∈G ǫα(α!)−shξiαd1/2 . ξ Setting r = ǫhξi and using inequalities ij)2 ≤ Cǫd3/2 ξ v(ξ∗ v(ξ∗)HS =vuut dξXi,j=1 α! (cid:19)s α! ≥ α!n−α and(cid:18)(r1/sn)α ǫα(α!)−shξiα. sup α ≤(cid:16)er1/sn(cid:17)s = ensr1/s , we obtain (5.2) kv(ξ∗)kHS ≤ Cǫd3/2 ξ ≤ Cǫd3/2 ξ (rns)α (α!)−s ensr1/s sup α sup α = Cǫd3/2 ξ ensǫ1/shξi1/s v ∈ γs(G)∧ by Theorem 4.2 and (3.2). for all ǫ > 0. We now recall that v(ξ∗) =bv(ξ) and, therefore, with vξ :=bv(ξ), we get (B) This case is similar but we give the proof for completeness. "If" part. Let v ∈ γ(s)(G)∧ and for any φ ∈ γ(s)(G) define v(φ) by (5.1). By a similar argument to the case (R), it is a well-defined linear functional on γ(s)(G). To check the continuity, suppose φj → φ in γ(s)(G), that is, for every A > 0 we have sup α A−α(α!)−s sup x∈G (−LG)α/2(φj(x) − φ(x)) → 0 as j → ∞. It follows that k (−LG)α/2 (φj − φ)k∞ ≤ CjAα ((α)!)s , for a sequence Cj → 0 as j → ∞, for every A > 0. From the proof of Theorem 2.3 it follows that for every B > 0 we have kbφj(ξ) −bφ(ξ)kHS ≤ Kje−Bhξi1/s , 22 APARAJITA DASGUPTA AND MICHAEL RUZHANSKY where Kj → 0 as j → ∞. Hence we can estimate v(φj − φ) ≤ X[ξ]∈ bG ≤ Kj X[ξ]∈ bG dξkbφj(ξ) −bφ(ξ)kHSkvξkHS dξe−Bhξi1/s kvξkHS → 0 as j → ∞ since Kj → 0 as j → ∞, and where we now take B > 0 to be such kvξkHS < ∞ by Lemma 4.1 and (3.2). Therefore, we have that P[ξ]∈ bG dξe−Bhξi1/s (s)(G). v ∈ γ′ "Only if" part. Let v ∈ γ′ (s)(G). This means that there exists ǫ > 0 and C > 0 such that v(φ) ≤ C sup α ǫα(α!)−s sup x∈G (−LG)α/2φ(x) holds for all φ ∈ γ(s)(G). Then, proceeding as in the case (R), we obtain (5.3) i.e. kbv(ξ)kHS ≤ Ceδhξi1/s kv(ξ∗)kHS ≤ Cd3/2 ξ ensǫ1/shξi1/s , , for some δ > 0. Hence v ∈ γ(s)(G)∧ by by Theorem 4.2. (cid:3) References [1] M. D. Bronshtein, The Cauchy problem for hyperbolic operators with characteristics of variable multiplicity. (Russian) Trudy Moskov. Mat. Obshch. 41 (1980), 83 -- 99; Trans. Moscow Math. Soc. 1 (1982), 87 -- 103. [2] F. Bruhat, Lectures on Lie groups and representations of locally compact groups. Tata Institute of Fundamental Research, Bombay, 1968. [3] C. Garetto and M. Ruzhansky, On the well-posedness of weakly hyperbolic equations with time dependent coefficients, J. Differential Equations, 253 (2012), 1317 -- 1340. [4] J. Faraut, Analysis on Lie groups. An introduction. Cambridge University Press, Cambridge, 2008. [5] H. Komatsu, Ultradistributions, I, II, III, J. Fac. Sci. Univ. of Tokyo, Sec. IA, 20 (1973), 25 -- 105, 24 (1977), 607 -- 628, 29 (1982), 653 -- 718. [6] G. Kothe, Topological vector spaces. I. Springer, 1969. [7] W. H. Ruckle, Sequence Spaces, Pitman, 1981. [8] M. Ruzhansky and V. Turunen, Pseudo-differential operators and symmetries, Birkhauser, Basel, 2010. [9] M. Ruzhansky and V. Turunen, Global quantization of pseudo-differential operators on com- pact Lie groups, SU(2) and 3-sphere, Int Math Res Notices IMRN (2012), 58 pages, doi: 10.1093/imrn/rns122. [10] E. M. Stein, Topics in harmonic analysis related to the Littlewood-Paley theory. Princeton University Press, Princeton, 1970. [11] Y. Taguchi, Fourier coefficients of periodic functions of Gevrey classes and ultradistributions. Yokohama Math. J. 35 (1987), 51 -- 60. [12] N. Ja. Vilenkin and A. U. Klimyk, Representation of Lie groups and special functions. Vol. 1. Simplest Lie groups, special functions and integral transforms. Kluwer Academic Publishers Group, Dordrecht, 1991. Aparajita Dasgupta: Department of Mathematics Imperial College London GEVREY FUNCTIONS AND ULTRADISTRIBUTIONS ON COMPACT LIE GROUPS 23 180 Queen's Gate, London SW7 2AZ United Kingdom E-mail address [email protected] Michael Ruzhansky: Department of Mathematics Imperial College London 180 Queen's Gate, London SW7 2AZ United Kingdom E-mail address [email protected]
1701.01478
1
1701
2017-01-05T21:15:03
New Multirectional Mean Value Inequality
[ "math.FA" ]
We establish new and stronger inequality of Clarke-Ledyaev type by direct construction.
math.FA
math
New Multirectional Mean Value Inequality∗ M. Hamamdjiev, M. Ivanov January, 2017 Abstract We establish new and stronger inequality of Clarke-Ledyaev type by direct construction. 2010 AMS Subject Classification: 49K27, 49J53, 49J52. Keywords and phrases: Multidirectional Mean Value Theorem, Subdifferential. 1 Introduction. In theories on the notion subdifferential it is often cumbersome to list for which of the many subdifferntials a given statement holds. A way around this issue, proposed by Ioffe and established by Thibault and others, is to consider the notion of subdifferential abstractly: as defined by set of axioms rather than by construction. Some of these axioms will be principal, like (P1), (P2) and (P3) below, and some technical, like (P4). As it is well known, a subdifferential operator ∂ applied to a lower semi- continuous function f : X → R ∪ {∞} on a Banach space X produces a multivalued map ∂f : X → 2X ∗ . Definition 1. We call the subdifferential ∂ feasible if the following properties hold: (P1) ∂f (x) = ∅ if x 6∈ dom f . ∗Partially supported by Bulgarian National Scientific Fund under Grant DFNI-I02/10. 1 (P2) ∂f (x) = ∂g(x) whenever f and g coincide in a neighbourhood of x. (P3) If f is convex and continuous in a neighbourhood of x then ∂f (x) co- incides with the standard subdifferential in Convex Analysis. (P4) If g is convex and continuous in a neighbourhood of z and f +g has local minimum at z then for each ε > 0 there are p ∈ ∂f (x) and q ∈ ∂g(y) such that kx − zk < ε, f (x) − f (z) < ε, ky − zk < ε, and kp + qk < ε. We discuss these axioms in Section 2. There we point out that most of the known subdifferentials satisfy them under natural conditions on the space. Here we state our main result. Let Bδ := B + δBX , where B is any subset of X and BX is the closed unit ball. For A, B ⊆ X let [A, B] be the convex hull of A and B. Theorem 2. Let X be a Banach space and let ∂ be a feasible subdifferential. Let A and B be non-empty closed, bounded and convex subsets of X. Let f : X → R ∪ {∞} be a proper lower semicontinuous function such that A ∩ dom f 6= ∅. Let f be bounded below on C := [A, B]δ for some δ > 0. Let Let r, s ∈ R be such that µ < inf C f. r = inf A f, s < inf Bδ f. Then for each ε > 0 there are ξ ∈ [A, B]δ and p ∈ ∂f (ξ) such that and f (ξ) < inf [A,B] f + r − s + ε kpk < max{r, s} − µ δ + ε, inf B p − inf A p > s − r. 2 (1) (2) (3) (4) (5) We will compare our result to what is known. Historically, the original multidirectional inequality can be found in [1]. It compares the values of a locally Lipschitz function on two bounded, closed and convex subsets of a Banach space, one of which is compact. The next work in the field is [2]. There we can find a multidirectional inequality, which compares the values of a lower semicontinuous and bounded below function on a point and a closed, convex and bounded set in the setting of a Hilbert space by using the prox- imal subgradient. In [7] we find a multidirectional inequality on β-smooth Banach spaces in a configuration of a point and a closed, bounded and convex set, a lower semicontinuous bounded below function and the corresponding β subdifferential. There are a number of subsequent developments, for example [3], [4], where we can find different kind of relaxations: non-convexness of a set, function not bounded below, etc. In this work, one can see that compared to the results from [1],[2], [7] we obtain an inequality for two sets for a lower semicontinuous bounded below function on a Banach space and a feasible subdifferential. Moreover, in the conditions of [1] we have a stronger inequality, as inf B f − sup A f ≤ inf B f − inf A f Furthermore, from the construction is clear that our inequality (4) is close to the optimal. The main tool for our proof is the function ϕK constructed in Section 3. By sketching the graph of ϕK the reader would readily grasp the idea. The reminder of the article is organised as follows: In Section 2 we discuss the axioms of subdifferential. Section 3 is devoted to the main construction. Finally, the proof of Theorem 2 is presented in Section 4. 2 On axioms of subdifferential. In this section we discuss Definition 1. First, observe that (P1), (P2) and (P3) are very common. The form here is essentially that found in [6]. The fuzzy sum rule (P4) here is formally stronger than the corresponding one in [6], but we have no example showing that it is actually stronger. This might be an interesting open question. It immediately follows from Corollary 4.64 [5, p.305] that the smooth sub- differential in smooth Banach space (the types of smoothness synchronized, 3 of course) is feasible. For the experts it would be obvious that (P4) has the typical "fuzzy" touch of smooth subdifferential. Indeed, (P4) is modeled after the corresponding property for smooth subdifferentials. Corollary 5.52 [5, p.385] and Theorem 7.23 [5, p.475] imply that Clarke subdifferential and G-subdifferential of Ioffe satisfy more than (P4), that is, if z is an local minimum of f + g then 0 ∈ ∂f (z) + ∂g(z). Similarly, it is easy to show that the limiting subdifferential on Asplund space, considered by Morduchovich and others, is feasible. This means that most of the subdifferentials are feasible under natural assumptions on their underlying spaces. And we put 'most' here just to be on the safe side, because we know of no meaningful counterexample. This implies that the result of this paper is very general. For simplicity we will derive an equivalent form of (P4) which encapsulates the standard application of Ekeland Variational principle. Proposition 3. Let X be a Banach space. Let f : X → R ∪ {∞} be a proper and lower semicontinuous function. If the subdifferential ∂ is feasible then it satisfies (P4′) If g is convex and continuous and if f + g is bounded below, then there are pn ∈ ∂−f (xn) and qn ∈ ∂−g(yn) such that kxn − ynk → 0, (f (xn) + g(yn)) → inf(f + g), kpn + qnk → 0. Proof. Let (zn)∞ 1 be a minimizing sequence for f + g, that is, f (zn) + g(zn) < inf(f + g) + εn, for some εn → 0. By Ekeland Variational Principle, see e.g. Theorem 1.88 [5, p.62], there are un ∈ X such that kun − znk ≤ εn and f (x) + g(x) + εnkx − unk ≥ f (un) + g(un). Set gn(x) := g(x) + εnkx − unk. Since f + gn has a minimum at un and gn is convex and continuous, by (P4) there are pn ∈ ∂f (xn) and qn ∈ ∂gn(yn) such that kxn −unk < εn, f (xn)−f (un) < εn, kyn −unk < εn and kpn + qnk < εn. All conclusions of (P4′) except the last one follow from triangle inequality. For the last one we note that any subdifferential of εnk · −unk is of norm less or equal to εn. By Sum Theorem of Convex Analysis, see e.g. [5, p. 206], there is qn ∈ ∂g(yn) such that kqn − qnk ≤ εn. We have kpn + qnk ≤ 2εn → 0. 4 3 Main construction. We will recall few notions. Let f and g be functions defined on the Banach space X. The supremal convolution (sup-convolution) of f and g is the function f ∗ g defined by f ∗ g(x) := sup{f (x − y) + g(y) : y ∈ X} The ε-subdifferential and ε-superdifferential of a function f are: ∂− ε f (x) := {p ∈ X ∗ : p(· − x) ≤ f (·) − f (x) + ε}, ∂+ ε f (x) := {p ∈ X ∗ : p(· − x) ≥ f (·) − f (x) − ε}, respectively. Note that we will use ∂ instead of ∂0 as usual. The hypograph of a function f is hyp f := {(x, t) : t ≤ f (x)} ⊆ X × R. Note also that the convex hull of two convex sets A and B can be written as: [A, B] = {z = λu + (1 − λ)v : u ∈ A, v ∈ B, λ ∈ [0, 1]}. Lemma 4. If p ∈ ∂+(f ∗ g)(x) and y ∈ X is such that f ∗ g(x) − ε ≤ f (x − y) + g(y), (6) then p ∈ ∂+ ε g(y). Proof. By the definition of sup-convolusion for all h ∈ X f ∗ g(x + h) ≥ f ((x + h) − (y + h)) + g(y + h) = f (x − y) + g(y + h). From this, p ∈ ∂+(f ∗ g)(x) and (6) it follows that p(h) ≥ f ∗ g(x + h) − f ∗ g(x) ≥ (f (x − y) + g(y + h)) − (f (x − y) + g(y) + ε) ≥ g(y + h) − g(y) − ε. That is, p ∈ ∂+ ε g(y). 5 Proposition 5. Let A and B be convex subsets of the Banach space X. Let r, s ∈ R be such that r 6= s. We define the concave function ψ by hyp ψ := co {A × [−∞; r], B × [−∞; s]}. Let x0 ∈ dom ψ = [A, B] be such that ψ(x0) 6= s. Let p ∈ ∂+ ε ψ(x0). Then inf A p − inf B p ≤ r − s + r − s ψ(x0) − s ε. (7) (8) Proof. Let p ∈ ∂+ ε ψ(x0). Set l(x) := ψ(x0) + p(x − x0). Note that by definition we have l ≥ ψ − ε. First we note that as (x0, ψ(x0)) ∈ hyp ψ (the reader is advised to draw a simple picture of ψ), we can find points (u, ¯r) ∈ A × [−∞, r], (v, ¯s) ∈ B × [−∞, s] such that: (x0, ψ(x0)) = λ(u, ¯r) + (1 − λ)(v, ¯s), for some λ ∈ [0; 1]. Suppose that λ = 0. Then (x0, ψ(x0)) = (v, ¯s), or ψ(x0) = ψ(v) = ¯s ≤ s. On the other hand, ψ(v) ≥ s. So, ψ(x0) = s, a contradiction to ψ(x0) 6= s. Thus, λ ∈ (0; 1]. Furthermore, as x0 = λu + (1 − λ)v; ψ(x0) = λ¯r + (1 − λ)¯s and ¯r ≤ r and ¯s ≤ s, we have that ψ(x0) ≤ λr + (1 − λ)s and at the same time (x0, λr + (1 − λ)s) = λ(u, r) + (1 − λ)(v, s) ∈ hyp ψ. It follows that λ¯r + (1 − λ)¯s = λr + (1 − λ)s, or equivalently λ(r − ¯r) + (1 − λ)(s − ¯s) = 0. There are two cases: Case 1: λ ∈ (0; 1). In this case we have that r = ¯r and s = ¯s, or (x0, ψ(x0)) = (λu + (1 − λ)v, λr + (1 − λ)s) = λ(u, r) + (1 − λ)(v, s) (9) It follows that p(v − x0) ≥ ψ(v) − ψ(x0) − ε ≥ s − ψ(x0) − ε = λ(s − r) − ε. 6 As we have we get u − x0 = − 1 − λ λ (v − x0), p(u − x0) ≤ 1 − λ λ It follows that [−λ(s − r) + ε] = (λ − 1)(s − r) + 1 − λ λ ε. l ≤ l(u) = ψ(x0) + p(u − x0) inf A ≤ ψ(x0) + (λ − 1)(s − r) + 1 − λ λ ε = r + 1 − λ λ ε. On the other hand, as l ≥ ψ − ε, we have inf B l ≥ inf B (ψ − ε) = inf B ψ − ε ≥ s − ε and after combining the last two inequalities, we obtain inf B p − inf A p = inf B l − inf A l ≥ s − ε − r − 1 − λ λ ε = s − r − ε λ . Finally, s − ψ(x0) = λ(s − r), so λ−1 = (s − r)/(s − ψ(x0)) and we get (8). Case 2: λ = 1. Here r = ¯r. Then (x0, ψ(x0)) = (u, r). We have l ≤ l(u) = ψ(x0) + p(u − x0) = r + p(0) = r, inf A inf B l ≥ inf B ψ − ε ≥ s − ε Here we get inf B p − inf A p = inf B l − inf A l ≥ s − r − ε. Since s − ψ(x0) = s − ψ(u) = s − r, this is equivalent to (8). The following function plays in our proof the role of the linear function in the standard proof of Lagrange Mean Value Theorem. Proposition 6. Let A and B be convex subsets of the Banach space X and let r, s ∈ R be such that r 6= s. We consider the function ψ as defined in Proposition 5. Let K > 0 and ϕK(x) := (−Kk · k) ∗ ψ(x) = sup{ψ(y) − Kkx − yk : y ∈ X}. 7 Then ϕK is K-Lipschitz and concave. Let ¯x be such that there exists c > 0 for which the sets U := {z ∈ [A, B] : ψ(z) − Kkz − ¯xk > ϕK(¯x) − c} and V := {z ∈ [A, B] : s − ψ(z)} < c do not intersect: U ∩ V = ∅. If p ∈ ∂+ϕK(¯x) then inf B p − inf A p ≥ s − r. (10) Proof. Since ψ is bounded from above, ϕK is well defined. As a sup-convolution of two concave functions ϕK is itself concave, see for example [5, p. 41]. As −Kk · k is a K-Lipschitz function, it easily follows that ϕK is also K-Lipschitz. For the main part note that from U ∩ V = ∅ it easily follows that for any sequence (xn)∞ 1 ⊆ [A, B] such that ϕK(¯x) = lim n→∞ (ψ(xn) − Kk¯x − xnk) it holds s − ψ(xn) ≥ c, ∀n large enough. We can assume that the latter is fulfilled for all n. The definition of ϕK is equivalent to: ϕK(x) = sup{ψ(y) − Kkx − yk : y ∈ [A, B]}, (11) (12) since ψ = −∞ outside [A, B]. From (11) it follows that we can find εn ↓ 0 with ψ(xn) − Kk¯x − xnk ≤ ϕK(xn) < ψ(xn) − Kk¯x − xnk + εn. In particular, we have: (−Kk · k) ∗ ψ(¯x) − εn < −Kk¯x − xnk + ψ(xn). For p ∈ ∂+ϕK(¯x) we apply Lemma 4. It follows that p ∈ ∂+ Proposition 5 and (12) we get εnψ(xn). So, from inf A p − inf B p ≤ r − s + r − s ψ(xn) − s εn ≤ r − s + s − r c εn. Since εn ↓ 0, we are done. 8 We will also need few more preparatory claims. Lemma 7. Let A, B be convex and bounded subsets of the Banach space X. Let r 6= s ∈ R and let ψ be constructed as in Proposition 5. If the sequence (xn)∞ n=1 is such that ψ(xn) → s as n → ∞, then d(xn, B) → 0. Proof. Let (xn, ψ(xn)) = λn(un, r) + (1 − λn)(vn, s), for some un ∈ A, vn ∈ B and λn ∈ [0, 1]. From r 6= s and λnr + (1 − λn)s → s it immediately follows that λn → 0. But d(xn, B) ≤ kxn − vnk = λnkun − vnk → 0, since A and B are bounded. Lemma 8. Let A, B be convex and bounded subsets of the Banach space X. Let r 6= s ∈ R and let ϕK be constructed as in Proposition 6 for some K > 0. Then min{r, s} ≤ ϕK(x) ≤ max{r, s}, ∀x ∈ [A, B]. (13) Proof. Since (A ∪ B) × (−∞, min{r, s}] ⊆ A × [−∞; r] ∪ B × [−∞; s] ⊆ (A ∪ B) × (−∞, max{r, s}], taking convex envelopes we get [A, B] × (−∞, min{r, s}] ⊆ hyp ψ ⊆ [A, B] × (−∞, max{r, s}], or, in other words, min{r, s} ≤ ψ ≤ max{r, s} on [A, B]. Since ψ ≤ max{r, s}, from the definition of sup-convolution it readily follows that ϕK ≤ max{r, s}. On the other hand, if x ∈ [A, B] then ϕK(x) ≥ ψ(x) ≥ min{r, s}. Lemma 9. Let A, B be convex subsets of the Banach space X and δ > 0. For the set C = [A, B]δ we have that its topological boundary ∂C satisfies ∂C ⊆ {x ∈ C : d(x, [A, B]) = δ}. Proof. First note that since C is closed, we have ∂C = C\intC, where the latter denotes the topological interior of C. Observe that ∀x : d(x, [A, B]) > δ =⇒ x /∈ C. 9 Indeed, if d(x, [A, B]) > δ for some x ∈ X then ∃ε > 0 with d(x, [A, B]) > δ + ε. Then for each y ∈ B(x, ε) := {y ∈ X : ky − xk < ε} ky − zk ≥ kz − xk − ky − xk > kz − xk − ε, ∀z ∈ X. This implies inf [A,B] ky − ·k ≥ inf [A,B] kx − ·k − ε > δ + ε − ε = δ, which means that [A, B]δ ∩ B(x, ε) = ∅. Thus x /∈ C. Next, it is obvious that ∀x : d(x, [A, B]) < δ =⇒ x ∈ intC. 4 Proof of the main result In this section we prove Theorem 2. Fix ε > 0. Fix s1 such that s < s1 < s + min{ε, εδ}, s1 < inf Bδ f and s1 6= r. Note that Also, r − s1 < r − s + ε. (14) max{r, s1} − µ δ < max{r, s} − µ δ + ε. Take δ1 ∈ (0, δ) such that K := max{r, s1} − µ δ1 < max{r, s} − µ δ + ε. (15) By Lemma 9 we have ∂C ⊆ {x ∈ C : d(x, [A, B]) = δ}. Let ϕK be the function constructed in Proposition 6 with these r, s1 and K. If x ∈ ∂C then ϕK(x) ≤ sup{ψ(y) : y ∈ [A, B]} + sup{−Kky − xk : y ∈ [A, B]} ≤ max{r, s1} − K inf{ky − xk : y ∈ [A, B]} = max{r, s1} − Kδ < µ. That is, ∀x ∈ ∂C ⇒ ϕK(x) < µ. (16) Set f1(x) :=(f (x), x ∈ C, x 6∈ C. ∞, 10 Since C is closed, f1 is lower semicontinuous. Also, inf f1 > µ from (1). From (P2) we have that ∂f1(x) = ∂f (x) for x ∈ C \ ∂C. Consider g(x) := f1(x) − ϕK(x). and note that dom f1 = dom g ⊆ C. From the above and (16) we have that the lower semicontinuous function g (ϕK is K-Lipschitz, Proposition 6) is bounded below and, moreover, We claim that inf{g(x) : x ∈ ∂C} > 0. inf g ≤ 0. (17) (18) Indeed, from (2) for any t > 0 there is x ∈ A such that f1(x) = f (x) < r + t. On the other hand, ϕK(x) ≥ ψ(x) ≥ r by the very construction of ψ, see (7). Therefore, g(x) < t. Since −ϕK is convex and continuous, we can apply (P 4′) from Proposi- tion 3 to f1 and−ϕK to get pn ∈ ∂f1(xn) and qn ∈ ∂(−ϕK )(yn) = −∂+ϕK(yn), such that kxn − ynk → 0, (f1(xn)−ϕK(yn)) → inf(f1−ϕK), kpn + qnk → 0. (19) We will next show that for all n ∈ N large enough (ξ, p) = (xn, pn) satisfies the conclusions of Theorem 2. Lemma 10. xn ∈ intC for all n ∈ N large enough and, therefore, pn ∈ ∂f (xn). Proof. Note that inf(f1−ϕK) = inf g ≤ 0 by (18), so lim(f1(xn)−ϕK(yn)) ≤ 0 by (19). Assume that there exists subsequense (xni)∞ i=1 ⊆ ∂C. Then (f1(xni)−ϕK(yni)) = g(xni) − (ϕK(xni) − ϕK(yni)) ≥ inf ∂C g − Kkxni − ynik, since ϕK is K-Lipschitz (see Proposition 6). From (17) and (19) it follows that the latter tends to strictly positive limit. Contradiction. The estimate (4) is easy to check: from (19) and the K-Lipschitz conti- nuity of ϕK, which implies kqnk ≤ K, it follows that lim sup kpnk ≤ K and we need only recall (15). 11 Lemma 11. For all n ∈ N large enough f (xn) < inf [A,B] f + r − s + ε. Proof. Let ν := r − s + ε − r − s1. From (14) we have ν > 0. As in the proof of Lemma 10 we use the Lipschitz continuity of ϕK to see that for all n large enough f (xn) − ϕK(yn) < inf{f (x) − ϕK(x) : x ∈ [A, B]} + ν. But from (13) we have min{r, s1} ≤ ϕK ≤ max{r, s1} on [A, B]. Therefore, f (xn) − max{r, s1} < inf [A,B] f − min{r, s1} + ν. Obviously, max{r, s1}−min{r, s1} = r−s1 and, therefore, f (xn) < inf [A,B] f + r − s1 + ν = inf [A,B] f + r − s + ε. Lemma 12. For all n ∈ N large enough there exist cn > 0 such that the sets Un,cn := {z ∈ [A, B] : ψ(z) − Kkz − ynk > ϕK(yn) − cn} and Vcn := {z ∈ [A, B] : s1 − ψ(z) < cn} do not intersect, that is Un,cn ∩ Vcn = ∅. Proof. Fix ¯ε > 0 such that ¯ε < inf C f − µ, ¯ε < inf Bδ f − s1, ¯ε < δ − δ1 1 + K −1 . Let n be so large that for ¯x = xn and ¯y = yn it is fulfilled k¯x − ¯yk < ¯ε, f1(¯x) − ϕK(¯y) < ¯ε, (20) (21) see (19). For this fixed n assume the contrary, that is, for any positive cn > 0 the sets Un,cn and Vcn defined with this cn, intersect. For any m ∈ N chose zm ∈ Un,m−1 ∩ Vm−1. Then the sequence {zm} ⊆ [A, B] satisfies ϕK(¯y) = lim m→∞ (ψ(zm) − Kkzm − ¯yk), lim m→∞ ψ(zm) = s1. (22) 12 It follows that ϕK(¯y) = s1−K limm→∞ kzm − ¯yk. But by (21) and (1) we have ϕK(¯y) > f1(¯x)−¯ε > µ−¯ε. So, K limm→∞ kzm−¯yk < s1−µ+ ¯ε ≤ Kδ1+ ¯ε from (15). But Lemma 7 implies d(zm, B) → 0, thus d(¯y, B) ≤ δ1 + ¯ε/K < δ − ¯ε from (20). From this, (21) and triangle inequality it follows that d(¯x, B) < δ. There- fore, ¯x ∈ Bδ and f1(¯x) ≥ inf Bδ f . But from (22) it is obvious that ϕK(¯y) ≤ s1 and, therefore, f1(¯x) − ϕK(¯y) ≥ inf Bδ f − s1 > ¯ε from (20). This, however, contradicts (21). From Proposition 6 and Lemma 12 it follows that inf B (−qn) − inf A (−qn) ≥ s1 − r for all n large enough. Since A and B are bounded, from (19) we get inf A pn − inf A (−qn) ≤ kpn + qnk sup X ∈A kxk → 0, inf B pn − inf B (−qn) ≤ kpn + qnk sup x∈B kxk → 0. From the three above and s1 > s it follows that inf B pn − inf A pn > s − r for all n large enough. The proof of Theorem 2 is thus completed. Acknowledgment. The second named author would like to thank Prof. R. Deville for poining out the potential for exploration of then new Clarke-Ledyaev inequality. Thanks are also due to Prof. N. Zlateva for valuable suggestions. References [1] F. H. Clarke, Yu. S. Ledyaev, Mean value inequalities, PAMS, 122(4), 1994, 1075 -- 1083. 13 [2] F. H. Clarke, Yu. S. Ledyaev, Mean value inequalities in Hilberrt spaces, TAMS, 344(1), 1994, 307 -- 324. [3] M. Ivanov, N. Zlateva, On nonconvex version of the inequality of Clarke and Ledyaev, Nonlinear Analysis: Theory, Methods and Applications, 49(8), 2002, 1023 -- 1036. [4] Yu. S. Ledyaev, Q. Zhu, Multidirectional mean value inequalities and weak monotonicity, Journal of the London Mathematical Society 71(1), 2005, 187 -- 202. [5] J.-P. Penot, Calculus without derivatives, GTM 266, Springer, 2012. [6] L. Thibault, N. Zlateva, Integrability of subdifferentials of directionally Lipschitz functions, PAMS, 133(10), 2939 -- 2948, 2005. [7] Q. Zhu, Clarke-Ledyaev mean value inequalities in smooth Banach spaces, Nonlinear Analysis: Theory, Methods and Applications, 32(3), 1998, 315 -- 324. 14
1306.3498
1
1306
2013-06-14T09:51:28
Coincidence and Common Fixed Point Results for Generalized $\alpha$-$\psi$ Contractive Type Mappings with Applications
[ "math.FA" ]
A new, simple and unified approach in the theory of contractive mappings was recently given by Samet \emph{et al.} (Nonlinear Anal. 75, 2012, 2154-2165) by using the concepts of $\alpha$-$\psi$-contractive type mappings and $\alpha$-admissible mappings in metric spaces. The purpose of this paper is to present a new class of contractive pair of mappings called generalized $\alpha$-$\psi$ contractive pair of mappings and study various fixed point theorems for such mappings in complete metric spaces. For this, we introduce a new notion of $\alpha$-admissible w.r.t $g$ mapping which in turn generalizes the concept of $g$-monotone mapping recently introduced by $\acute{C}$iri$\acute{c}$ et al. (Fixed Point Theory Appl. 2008(2008), Article ID 131294, 11 pages). As an application of our main results, we further establish common fixed point theorems for metric spaces endowed with a partial order as well as in respect of cyclic contractive mappings. The presented theorems extend and subsumes various known comparable results from the current literature. Some illustrative examples are provided to demonstrate the main results and to show the genuineness of our results.
math.FA
math
COINCIDENCE AND COMMON FIXED POINT RESULTS FOR GENERALIZED α-ψ CONTRACTIVE TYPE MAPPINGS WITH APPLICATIONS PRIYA SHAHI∗, JATINDERDEEP KAUR, S. S. BHATIA School of Mathematics and Computer Applications, Thapar University, Patiala-147004, Punjab, India. Email addresses: [email protected]∗, [email protected], [email protected] Abstract. A new, simple and unified approach in the theory of contractive mappings was re- cently given by Samet et al. (Nonlinear Anal. 75, 2012, 2154-2165) by using the concepts of α-ψ-contractive type mappings and α-admissible mappings in metric spaces. The purpose of this paper is to present a new class of contractive pair of mappings called generalized α-ψ con- tractive pair of mappings and study various fixed point theorems for such mappings in complete metric spaces. For this, we introduce a new notion of α-admissible w.r.t g mapping which in turn generalizes the concept of g-monotone mapping recently introduced by ´Ciri´c et al. (Fixed Point Theory Appl. 2008(2008), Article ID 131294, 11 pages). As an application of our main results, we further establish common fixed point theorems for metric spaces endowed with a par- tial order as well as in respect of cyclic contractive mappings. The presented theorems extend and subsumes various known comparable results from the current literature. Some illustrative examples are provided to demonstrate the main results and to show the genuineness of our results. Keywords: Common fixed point; Contractive type mapping; Partial order; Cyclic mappings. Mathematics Subject Classification (2000): 54H25, 47H10, 54E50. 1. Introduction Fixed point theory has fascinated many researchers since 1922 with the celebrated Banach fixed point theorem. There exists a vast literature on the topic and this is a very active field of research at present. Fixed point theorems are very important tools for proving the existence and uniqueness of the solutions to various mathematical models (integral and partial differential equations, variational inequalities etc). It is well known that the contractive-type conditions are very indispensable in the study of fixed point theory. The first important result on fixed points for contractive-type mappings was the well-known Banach-Caccioppoli theorem which was published in 1922 in [1] and it also appears in [2]. Later in 1968, Kannan [3] studied a new type of contractive mapping. Since then, there have been many results related to mappings satisfying various types of contractive inequality, we refer to ([4], [5], [6], [7], [8] etc) and references therein. Recently, Samet et al. [9] introduced a new category of contractive type mappings known as α-ψ contractive type mapping. The results obtained by Samet et al. [9] extended and generalized the existing fixed point results in the literature, in particular the Banach contraction principle. Further, Karapinar and Samet [10] generalized the α-ψ-contractive type mappings and obtained various fixed point theorems for this generalized class of contractive mappings. The study related to common fixed points of mappings satisfying certain contractive conditions has been at the center of vigorous research activity. In this paper, some coincidence and common fixed point theorems are obtained for the generalized α-ψ contractive pair of mappings. Our results ∗Corresponding author: Priya Shahi, School of Mathematics and Computer Applications, Thapar University, Patiala 147004, Punjab, India. Email address: [email protected]. 1 2 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA unify and generalize the results derived by Karapinar and Samet [10], Samet et al. [9], ´Ciri´c et al. [11] and various other related results in the literature. Moreover, from our main results, we will derive various common fixed point results for metric spaces endowed with a partial order and that for cyclic contractive mappings. The presented results extend and generalize numerous related results in the literature. 2. Preliminaries First we introduce some notations and definitions that will be used subsequently. Definition 2.1. (See [9]). Let Ψ be the family of functions ψ : [0,∞) → [0,∞) satisfying the following conditions: (i) ψ is nondecreasing. (ii) +∞ Xn=1 ψn(t) < ∞ for all t > 0, where ψn is the nth iterate of ψ. These functions are known as (c)-comparison functions in the literature. It can be easily verified that if ψ is a (c)-comparison function, then ψ(t) < t for any t > 0. Recently, Samet et al. [9] introduced the following new notions of α-ψ-contractive type mappings and α-admissible mappings: Definition 2.2. Let (X, d) be a metric space and T : X → X be a given self mapping. T is said to be an α-ψ-contractive mapping if there exists two functions α : X × X → [0, +∞) and ψ ∈ Ψ such that α(x, y)d(T x, T y) ≤ ψ(d(x, y)) for all x, y ∈ X. Definition 2.3. Let T : X → X and α : X × X → [0, +∞). T is said to be α-admissible if x, y ∈ X, α(x, y) ≥ 1 ⇒ α(T x, T y) ≥ 1. The following fixed point theorems are the main results in [9]: Theorem 2.1. Let (X, d) be a complete metric space and T : X → X be an α-ψ-contractive mapping satisfying the following conditions: (i) T is α-admissible; (ii) there exists x0 ∈ X such that α(x0, T x0) ≥ 1; (iii) T is continuous. Then, T has a fixed point, that is, there exists x∗ ∈ X such that T x∗ = x∗. Theorem 2.2. Let (X, d) be a complete metric space and T : X → X be an α-ψ-contractive mapping satisfying the following conditions: (i) T is α-admissible; (ii) there exists x0 ∈ X such that α(x0, T x0) ≥ 1; (iii) if {xn} is a sequence in X such that α(xn, xn+1) ≥ 1 for all n and xn → x ∈ X as n → +∞, then α(xn, x) ≥ 1 for all n. Then, T has a fixed point. Samet et al. [9] added the following condition to the hypotheses of Theorem 2.1 and Theorem 2.2 to assure the uniqueness of the fixed point: (C): For all x, y ∈ X, there exists z ∈ X such that α(x, z) ≥ 1 and α(y, z) ≥ 1. Recently, Karapinar and Samet [10] introduced the following concept of generalized α-ψ-contractive type mappings: Definition 2.4. Let (X, d) be a metric space and T : X → X be a given mapping. We say that T is a generalized α-ψ-contractive type mapping if there exists two functions α : X × X → [0,∞) and ψ ∈ Ψ such that for all x, y ∈ X, we have 3 where M (x, y) = max(cid:26)d(x, y), α(x, y)d(T x, T y) ≤ ψ(M (x, y)), d(x, T x) + d(y, T y) d(x, T y) + d(y, T x) 2 , 2 (cid:27). Further, Karapinar and Samet [10] established fixed point theorems for this new class of con- tractive mappings. Also, they obtained fixed point theorems on metric spaces endowed with a partial order and fixed point theorems for cyclic contractive mappings. Definition 2.5. [12] Let X be a non-empty set, N is a natural number such that N ≥ 2 and T1, T2, ..., TN : X → X are given self-mappings on X. If w = T1x = T2x = ... = TN x for some x ∈ X, then x is called a coincidence point of T1, T2, ..., TN −1 and TN , and w is called a point of coincidence of T1, T2, ..., TN −1 and TN . If w = x, then x is called a common fixed point of T1, T2, ..., TN −1 and TN . Let f, g : X → X be two mappings. We denote by C(g, f ) the set of coincidence points of g and f ; that is, C(g, f ) = {z ∈ X : gz = f z} 3. Main results We start the main section by introducing the new consepts of α-admissible w.r.t g mapping and generalized α-ψ contractive pair of mappings. Definition 3.1. Let f, g : X → X and α : X × X → [0,∞). We say that f is α-admissible w.r.t g if for all x, y ∈ X, we have α(gx, gy) ≥ 1 ⇒ α(f x, f y) ≥ 1. Remark 3.1. Clearly, every α-admissible mapping is α-admissible w.r.t g mapping when g = I. The following example shows that a mapping which is α-admissible w.r.t g may not be α- admissible. Example 3.2. Let X = [1,∞). Define the mapping α : X × X → [0,∞) by if x > y α(x, y) =  2 1 3 Also, define the mappings f, g : X → X by f (x) = Suppose that α(x, y) ≥ 1. This implies from the definition of α that x > y which further implies that Now, we prove that f is α-admissible w.r.t g. Let us suppose that α(gx, gy) ≥ 1. So, . Thus, α(f x, f y) (cid:3) 1, that is, f is not α-admissible. and g(x) = e−x for all x ∈ X. otherwise 1 x < 1 y 1 x α(gx, gy) ≥ 1 ⇒ gx > gy ⇒ e−x > e−y ⇒ Therefore, f is α-admissible w.r.t g. 1 x > 1 y ⇒ α(f x, f y) ≥ 1 In what follows, we present examples of α-admissible w.r.t g mappings. Example 3.3. Let X be the set of all non-negative real numbers. Let us define the mapping α : X × X → [0, +∞) by 4 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA α(x, y) =( 1 if x ≥ y, 0 if x < y. and define the mappings f, g : X → X by f (x) = ex and g(x) = x2 for all x ∈ X. Thus, the mapping f is α-admissible w.r.t g. Example 3.4. Let X = [1,∞). Let us define the mapping α : X × X → [0, +∞) by α(x, y) =( 3 1 2 if x, y ∈ [0, 1], otherwise. and define the mappings f, g : X → X by f (x) = ln(cid:16)1 + the mapping f is α-admissible w.r.t g. x 3(cid:17) and g(x) = √x for all x ∈ X. Thus, Next, we present the new notion of generalized α-ψ contractive pair of mappings as follows: Definition 3.5. Let (X, d) be a metric space and f, g : X → X be given mappings. We say that the pair (f, g) is a generalized α-ψ contractive pair of mappings if there exists two functions α : X × X → [0, +∞) and ψ ∈ Ψ such that for all x, y ∈ X, we have (1) α(gx, gy)d(f x, f y) ≤ ψ(M (gx, gy)), where M (gx, gy) = max(cid:26)d(gx, gy), d(gx, f x) + d(gy, f y) 2 , d(gx, f y) + d(gy, f x) 2 (cid:27). Our first result is the following coincidence point theorem. Theorem 3.1. Let (X, d) be a complete metric space and f, g : X → X be such that f (X) ⊆ g(X). Assume that the pair (f, g) is a generalized α-ψ contractive pair of mappings and the following conditions hold: (i) f is α-admissible w.r.t. g; (ii) there exists x0 ∈ X such that α(gx0, f x0) ≥ 1; (iii) If {gxn} is a sequence in X such that α(gxn, gxn+1) ≥ 1 for all n and gxn → gz ∈ g(X) as n → ∞, then there exists a subsequence {gxn(k)} of {gxn} such that α(gxn(k), gz) ≥ 1 for all k. Also suppose g(X) is closed. Then, f and g have a coincidence point. Proof. In view of condition (ii), let x0 ∈ X be such that α(gx0, f x0) ≥ 1. Since f (X) ⊆ g(X), we can choose a point x1 ∈ X such that f x0 = gx1. Continuing this process having chosen x1, x2, ..., xn, we choose xn+1 in X such that (2) f xn = gxn+1, n = 0, 1, 2, ... Since f is α-admissible w.r.t g, we have α(gx0, f x0) = α(gx0, gx1) ≥ 1 ⇒ α(f x0, f x1) = α(gx1, gx2) ≥ 1 Using mathematical induction, we get (3) α(gxn, gxn+1) ≥ 1,∀ n = 0, 1, 2, ... If f xn+1 = f xn for some n, then by (2), f xn+1 = gxn+1, n = 0, 1, 2, ... that is, f and g have a coincidence point at x = xn+1, and so we have finished the proof. For this, we suppose that d(f xn, f xn+1) > 0 for all n. Applying the inequality (1) and using (3), we obtain (4) d(f xn, f xn+1) ≤ α(gxn, gxn+1)d(f xn, f xn+1) ≤ ψ(M (gxn, gxn+1)) On the other hand, we have M (gxn, gxn+1) = max(cid:26)d(gxn, gxn+1), d(gxn, f xn) + d(gxn+1, f xn+1) 2 ≤ max{d(f xn−1, f xn), d(f xn, f xn+1)} 5 , d(gxn, f xn+1) + d(gxn+1, f xn) 2 (cid:27) Owing to monotonicity of the function ψ and using the inequalities (2) and (4), we have for all n ≥ 1 (5) d(f xn, f xn+1) ≤ ψ(max {d(f xn−1, f xn), d(f xn, f xn+1)} If for some n ≥ 1, we have d(f xn−1, f xn) ≤ d(f xn, f xn+1), from (5), we obtain that d(f xn, f xn+1) ≤ ψ(d(f xn, f xn+1)) < d(f xn, f xn+1), a contradiction. Thus, for all n ≥ 1, we have (6) max{d(f xn−1, f xn), d(f xn, f xn+1)} = d(f xn−1, f xn) Notice that in view of (5) and (6), we get for all n ≥ 1 that (7) d(f xn, f xn+1) ≤ ψ(d(f xn−1, f xn)). Continuing this process inductively, we obtain (8) d(f xn, f xn+1) ≤ ψn(d(f x0, f x1)), ∀n ≥ 1. From (8) and using the triangular inequality, for all k ≥ 1, we have d(f xn, f xn+k) ≤ d(f xn, f xn+1) + ... + d(f xn+k−1, f xn+k) n+k−1 (9) ≤ ≤ ψp(d(f x1, f x0)) ψp(d(f x1, f x0)) +∞ Xp=n Xp=n Letting p → ∞ in (9), we obtain that {f xn} is a Cauchy sequence in (X, d). Since by (2) we have {f xn} = {gxn+1} ⊆ g(X) and g(X) is closed, there exists z ∈ X such that (10) gxn = gz. lim n→∞ Now, we show that z is a coincidence point of f and g. On contrary, assume that d(f z, gz) > 0. Since by condition (iii) and (10), we have α(gxn(k), gz) ≥ 1 for all k, then by the use of triangle inequality and (1) we obtain d(gz, f z) ≤ d(gz, f xn(k)) + d(f xn(k), f z) (11) On the other hand, we have ≤ d(gz, f xn(k)) + α(gxn(k), gz)d(f xn(k), f z) ≤ d(gz, f xn(k)) + ψ(M (gxn(k), gz) M (gxn(k), gz) = max(cid:26)d(gxn(k), gz), d(gxn(k), f xn(k)) + d(gz, f z) 2 , d(gxn(k), f z) + d(gz, f xn(k)) 2 (cid:27) Owing to above equality, we get from (11), d(gz, f z) ≤ d(gz, f xn(k)) + ψ(M (gxn(k), gz) ≤ d(gz, f xn(k)) + ψ(cid:18)max(cid:26)d(gxn(k), gz), d(gxn(k), f xn(k)) + d(gz, f z) 2 , d(gxn(k), f z) + d(gz, f xn(k)) 2 (cid:27)(cid:19) 6 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA Letting k → ∞ in the above inequality yields d(gz, f z) ≤ ψ(cid:18) d(f z, gz) 2 (cid:19) < d(f z, gz) 2 , which is a contradiction. Hence, our supposition is wrong and d(f z, gz) = 0, that is, f z = gz. This shows that f and g have a coincidence point. (cid:3) The next theorem shows that under additional hypotheses we can deduce the existence and uniqueness of a common fixed point. Theorem 3.2. In addition to the hypotheses of Theorem 3.1, suppose that for all u, v ∈ C(g, f ), there exists w ∈ X such that α(gu, gw) ≥ 1 and α(gv, gw) ≥ 1 and f, g commute at their coinci- dence points. Then f and g have a unique common fixed point. Proof. We need to consider three steps: Step 1. We claim that if u, v ∈ C(g, f ), then gu = gv. By hypotheses, there exists w ∈ X such that (12) α(gu, gw) ≥ 1, α(gv, gw) ≥ 1 Due to the fact that f (X) ⊆ g(X), let us define the sequence {wn} in X by gwn+1 = f wn for all n ≥ 0 and w0 = w. Since f is α-admissible w.r.t g, we have from (12) that (13) α(gu, gwn) ≥ 1, α(gv, gwn) ≥ 1 for all n ≥ 0. Applying inequality (1) and using (13), we obtain d(gu, gwn+1) = d(f u, f wn) ≤ α(gu, gwn)d(f u, f wn) ≤ ψ(M (gu, gwn)) (14) On the other hand, we have M (gu, gwn) = max(cid:26)d(gu, gwn), 2 ≤ max {d(gu, gwn), d(gu, gwn+1)} d(gu, f u) + d(gwn, f wn) (15) , d(gu, f wn) + d(gwn, f u) 2 (cid:27) Using the above inequality, (14) and owing to the monotone property of ψ, we get that (16) d(gu, gwn+1) ≤ ψ(max {d(gu, gwn), d(gu, gwn+1)}) for all n. Without restriction to the generality, we can suppose that d(gu, gwn) > 0 for all n. If max{d(gu, gwn), d(gu, gwn+1)} = d(gu, gwn+1), we have from (16) that (17) d(gu, gwn+1) ≤ ψ(d(gu, gwn+1)) < d(gu, gwn+1) which is a contradiction. Thus, we have max{d(gu, gwn), d(gu, gwn+1)} = d(gu, gwn), and d(gu, gwn+1) ≤ ψ(d(gu, gwn)) for all n. This implies that (18) d(gu, gwn) ≤ ψn(d(gu, gw0)), ∀n ≥ 1 Letting n → ∞ in the above inequality, we infer that (19) d(gu, gwn) = 0 lim n→∞ Similarly, we can prove that (20) lim n→∞ d(gv, gwn) = 0 It follows from (19) and (20) that gu = gv. Step 2. Existence of a common fixed point: Let u ∈ C(g, f ), that is, gu = f u. Owing to the commutativity of f and g at their coincidence points, we get (21) g2u = gf u = f gu Let us denote gu = z, then from (21), gz = f z. Thus, z is a coincidence point of f and g. Now, from Step 1, we have gu = gz = z = f z. Then, z is a common fixed point of f and g. Step 3. Uniqueness: Assume that z ∗ is another common fixed point of f and g. Then z ∗ ∈ C(g, f ). By step 1, we have z ∗ = gz ∗ = gz = z. This completes the proof. (cid:3) In what follows, we furnish an illustrative example wherein one demonstrates Theorem 3.2 on the existence and uniqueness of a common fixed point. 7 Example 3.6. Consider X = [0, +∞) equipped with the usual metric d(x, y) = x − y for all x, y ∈ X. Define the mappings f : X → X and g : X → X by if x > 2, if 0 ≤ x ≤ 2. 2x − x 3 3 2 and f (x) =  g(x) = x 2 ∀x ∈ X. Now, we define the mapping α : X × X → [0, +∞) by α(x, y) =( 1 0 if (x, y) ∈ [0, 1], otherwise. Clearly, the pair (f, g) is a generalized α-ψ contractive pair of mappings with ψ(t) = t ≥ 0. In fact, for all x, y ∈ X, we have 4 5 t for all α(gx, gy).d(f x, f y) = 1.(cid:12)(cid:12)(cid:12) x 3 − y 3(cid:12)(cid:12)(cid:12) ≤ = ≤ 4 5 4 5 y 4 x 2 − d(gx, gy) 2(cid:12)(cid:12)(cid:12) 5(cid:12)(cid:12)(cid:12) M (gx, gy) = ψ(M (gx, gy)) , 1 2 Moreover, there exists x0 ∈ X such that α(gx0, f x0) ≥ 1. Infact, for x0 = 1, we have α(cid:18) 1 3(cid:19) = 1. In so doing, let x, y ∈ X such that Now, it remains to show that f is α-admissible w.r.t g. α(gx, gy) ≥ 1. This implies that gx, gy ∈ [0, 1] and by the definition of g, we have x, y ∈ [0, 2]. Therefore, by the definition of f and α, we have x y 3 ∈ [0, 1], f (y) = 3 ∈ [0, 1] and α(f x, f y) = 1. Thus, f is α-admissible w.r.t g. Clearly, f (X) ⊆ g(X) and g(X) is closed. Finally, let {gxn} be a sequence in X such that α(gxn, gxn+1) ≥ 1 for all n and gxn → gz ∈ g(X) as n → +∞. Since α(gxn, gxn+1) ≥ 1 for all n, by the definition of α, we have gxn ∈ [0, 1] for all n and gz ∈ [0, 1]. Then, α(gxn, gz) ≥ 1. Now, all the hypotheses of Theorem 3.1 are satisfied. Consequently, f and g have a coincidence point. Here, 0 is a coincidence point of f and g. Also, clearly all the hypotheses of Theorem 3.2 are satisfied. In this example, 0 is the unique common fixed point of f and g. f (x) = 4. Consequences In this section, we will show that many existing results in the literature can be obtained easily from our Theorem 3.2. 4.1. Standard Fixed Point Theorems. By taking α(x, y) = 1 for all x, y ∈ X in Theorem 3.2, we obtain immediately the following fixed point theorem. 8 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA Corollary 4.1. Let (X, d) be a complete metric space and f, g : X → X be such that f (X) ⊆ g(X). Suppose that there exists a function ψ ∈ Ψ such that (22) d(f x, f y) ≤ ψ(M (gx, gy)), for all x, y ∈ X. Also, suppose g(X) is closed. Then, f and g have a coincidence point. Further, if f , g commute at their coincidence points, then f and g have a common fixed point. By taking g = I in Corollary 4.1, we obtain immediately the following fixed point theorem. Corollary 4.2. (see Karapinar and Samet [10]). Let (X, d) be a complete metric space and f : X → X. Suppose that there exists a function ψ ∈ Ψ such that (23) d(f x, f y) ≤ ψ(M (x, y)), for all x, y ∈ X. Then f has a unique fixed point. The following fixed point theorems can be easily obtained from Corollaries 4.1 and 4.2. Corollary 4.3. Let (X, d) be a complete metric space and f, g : X → X be such that f (X) ⊆ g(X). Suppose that there exists a function ψ ∈ Ψ such that (24) d(f x, f y) ≤ ψ(d(gx, gy)), for all x, y ∈ X. Also, suppose g(X) is closed. Then, f and g have a coincidence point. Further, if f , g commute at their coincidence points, then f and g have a common fixed point. Corollary 4.4. (Berinde [13]). Let (X, d) be a complete metric space and f : X → X. Suppose that there exists a function ψ ∈ Ψ such that (25) d(f x, f y) ≤ ψ(d(x, y)), for all x, y ∈ X. Then f has a unique fixed point. Corollary 4.5. ( ´Ciri´c [14]). Let (X, d) be a complete metric space and T : X → X be a given mapping. Suppose that there exists a constant λ ∈ (0, 1) such that d(f x, f y) ≤ λ max(cid:26)d(x, y), 2 for all x, y ∈ X. Then T has a unique fixed point. Corollary 4.6. (Hardy and Rogers [15]) Let (X, d) be a complete metric space and T : X → X be a given mapping. Suppose that there exists constants A, B, C ≥ 0 with (A + 2B + 2C) ∈ (0, 1) such that 2 d(x, f x) + d(y, f y) d(x, f y) + d(y, f x) , (cid:27) d(f x, f y) ≤ Ad(x, y) + B[d(x, f x) + d(y, f y)] + C[d(x, f y) + d(y, f x)], for all x, y ∈ X. Then T has a unique fixed point. Corollary 4.7. (Banach Contraction Principle [1]) Let (X, d) be a complete metric space and T : X → X be a given mapping. Suppose that there exists a constant λ ∈ (0, 1) such that d(f x, f y) ≤ λd(x, y) for all x, y ∈ X. Then T has a unique fixed point. Corollary 4.8. (Kannan [3]) Let (X, d) be a complete metric space and T : X → X be a given mapping. Suppose that there exists a constant λ ∈ (0, 1/2) such that for all x, y ∈ X. Then T has a unique fixed point. d(f x, f y) ≤ λ[d(x, f x) + d(y, f y)], 9 Corollary 4.9. (Chatterjee [16]) Let (X, d) be a complete metric space and T : X → X be a given mapping. Suppose that there exists a constant λ ∈ (0, 1/2) such that d(f x, f y) ≤ λ[d(x, f y) + d(y, f x)], for all x, y ∈ X. Then T has a unique fixed point. 4.2. Fixed Point Theorems on Metric Spaces Endowed with a Partial Order. Recently, there have been enormous developments in the study of fixed point problems of contractive map- pings in metric spaces endowed with a partial order. The first result in this direction was given by Turinici [17], where he extended the Banach contraction principle in partially ordered sets. Some applications of Turinici's theorem to matrix equations were presented by Ran and Reurings [18]. Later, many useful results have been obtained regarding the existence of a fixed point for contraction type mappings in partially ordered metric spaces by Bhaskar and Lakshmikantham [4], [20], Lakshmikantham and ´Ciri´c [6] and Samet [21] etc. Nieto and Lopez [7, 19], Agarwal et al. In this section, we will derive various fixed point results on a metric space endowed with a partial order. For this, we require the following concepts: Definition 4.1. [10] Let (X,(cid:22)) be a partially ordered set and T : X → X be a given mapping. We say that T is nondecreasing with respect to (cid:22) if x, y ∈ X, x (cid:22) y ⇒ T x (cid:22) T y. Definition 4.2. [10] Let (X,(cid:22)) be a partially ordered set. A sequence {xn} ⊂ X is said to be nondecreasing with respect to (cid:22) if xn (cid:22) xn+1 for all n. Definition 4.3. [10] Let (X,(cid:22)) be a partially ordered set and d be a metric on X. We say that (X,(cid:22), d) is regular if for every nondecreasing sequence {xn} ⊂ X such that xn → x ∈ X as n → ∞, there exists a subsequence {xn(k)} of {xn} such that xn(k) (cid:22) x for all k. Definition 4.4. [11] Suppose (X,(cid:22)) is a partially ordered set and F, g : X → X are mappings of X into itself. One says F is g-non-decreasing if for x, y ∈ X, (26) implies g(x) (cid:22) g(y) F (x) (cid:22) F (y). Definition 4.5. Let (X,(cid:22)) be a partially ordered set and d be a metric on X. We say that (X,(cid:22), d) is g-regular where g : X → X if for every nondecreasing sequence {gxn} ⊂ X such that gxn → gz ∈ X as n → ∞, there exists a subsequence {gxn(k)} of {gxn} such that gxn(k) (cid:22) gz for all k. We have the following result. Corollary 4.10. Let (X,(cid:22)) be a partially ordered set and d be a metric on X such that (X, d) is complete. Assume that f, g : X → X be such that f (X) ⊆ g(X) and f be a g-non-decreasing mapping w.r.t (cid:22). Suppose that there exists a function ψ ∈ Ψ such that (27) d(f x, f y) ≤ ψ(M (gx, gy)), for all x, y ∈ X with gx (cid:22) gy. Suppose also that the following conditions hold: (i) there exists x0 ∈ X s.t gx0 (cid:22) f x0; (ii) (X,(cid:22), d) is g-regular. Also suppose g(X) is closed. Then, f and g have a coincidence point. Moreover, if for every pair (x, y) ∈ C(g, f ) × C(g, f ) there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, and if f and g commute at their coincidence points, then we obtain uniqueness of the common fixed point. 10 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA Proof. Define the mapping α : X × X → [0,∞) by (28) α(x, y) =( 1 if x (cid:22) y or x (cid:23) y 0 otherwise Clearly, the pair (f, g) is a generalized α-ψ contractive pair of mappings, that is, α(gx, gy)d(f x, f y) ≤ ψ(M (gx, gy)), for all x, y ∈ X. Notice that in view of condition (i), we have α(gx0, f x0) ≥ 1. Moreover, for all x, y ∈ X, from the g-monotone property of f , we have (29) α(gx, gy) ≥ 1 ⇒ gx (cid:22) gy or gx (cid:23) gy ⇒ f x (cid:22) f y or f x (cid:23) f y ⇒ α(f x, f y) ≥ 1. which amounts to say that f is α-admissible w.r.t g. Now, let {gxn} be a sequence in X such that α(gxn, gxn+1) ≥ 1 for all n and gxn → gz ∈ X as n → ∞. From the g-regularity hypothesis, there exists a subsequence {gxn(k)} of {gxn} such that gxn(k) (cid:22) gz for all k. So, by the definition of α, we obtain that α(gxn(k), gz) ≥ 1. Now, all the hypotheses of Theorem 3.1 are satisfied. Hence, we deduce that f and g have a coincidence point z, that is, f z = gz. Now, we need to show the existence and uniqueness of common fixed point. For this, let x, y ∈ X. By hypotheses, there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, which implies from the definition of α that α(gx, gz) ≥ 1 and α(gy, gz) ≥ 1. Thus, we deduce the existence and uniqueness of the common fixed point by Theorem 3.2. (cid:3) The following results are immediate consequences of Corollary 4.10. Corollary 4.11. Let (X,(cid:22)) be a partially ordered set and d be a metric on X such that (X, d) is complete. Assume that f, g : X → X and f be a g-non-decreasing mapping w.r.t (cid:22). Suppose that there exists a function ψ ∈ Ψ such that (30) d(f x, f y) ≤ ψ(d(gx, gy)), for all x, y ∈ X with gx (cid:22) gy. Suppose also that the following conditions hold: (i) there exists x0 ∈ X s.t gx0 (cid:22) f x0; (ii) (X,(cid:22), d) is g-regular. Also suppose g(X) is closed. Then, f and g have a coincidence point. Moreover, if for every pair (x, y) ∈ C(g, f ) × C(g, f ) there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, and if f and g commute at their coincidence points, then we obtain uniqueness of the common fixed point. Corollary 4.12. Let (X,(cid:22)) be a partially ordered set and d be a metric on X such that (X, d) is complete. Assume that f, g : X → X and f be a g-non-decreasing mapping w.r.t (cid:22). Suppose that there exists a constant λ ∈ (0, 1) such that (31) d(f x, f y) ≤ λ max(cid:26)d(gx, gy), for all x, y ∈ X with gx (cid:22) gy. Suppose also that the following conditions hold: (i) there exists x0 ∈ X s.t gx0 (cid:22) f x0; (ii) (X,(cid:22), d) is g-regular. Also suppose g(X) is closed. Then, f and g have a coincidence point. Moreover, if for every pair (x, y) ∈ C(g, f ) × C(g, f ) there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, and if f and g commute at their coincidence points, then we obtain uniqueness of the common fixed point. d(gx, f x) + d(gy, f y) (cid:27) , 2 d(gx, f y) + d(gy, f x) , 2 Corollary 4.13. Let (X,(cid:22)) be a partially ordered set and d be a metric on X such that (X, d) is complete. Assume that f, g : X → X and f be a g-non-decreasing mapping w.r.t (cid:22). Suppose that there exists constants A, B, C ≥ 0 with (A + 2B + 2C) ∈ (0, 1) such that (32) d(f x, f y) ≤ Ad(gx, gy) + B[d(gx, f x) + d(gy, f y)] + C[d(gx, f y) + d(gy, f x)], 11 for all x, y ∈ X with gx (cid:22) gy. Suppose also that the following conditions hold: (i) there exists x0 ∈ X s.t gx0 (cid:22) f x0; (ii) (X,(cid:22), d) is g-regular. Also suppose g(X) is closed. Then, f and g have a coincidence point. Moreover, if for every pair (x, y) ∈ C(g, f ) × C(g, f ) there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, and if f and g commute at their coincidence points, then we obtain uniqueness of the common fixed point. Corollary 4.14. Let (X,(cid:22)) be a partially ordered set and d be a metric on X such that (X, d) is complete. Assume that f, g : X → X and f be a g-non-decreasing mapping w.r.t (cid:22). Suppose that there exists a constant λ ∈ (0, 1) such that (33) d(f x, f y) ≤ λ(d(gx, gy)), for all x, y ∈ X with gx (cid:22) gy. Suppose also that the following conditions hold: (i) there exists x0 ∈ X s.t gx0 (cid:22) f x0; (ii) (X,(cid:22), d) is g-regular. Also suppose g(X) is closed. Then, f and g have a coincidence point. Moreover, if for every pair (x, y) ∈ C(g, f ) × C(g, f ) there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, and if f and g commute at their coincidence points, then we obtain uniqueness of the common fixed point. Corollary 4.15. Let (X,(cid:22)) be a partially ordered set and d be a metric on X such that (X, d) is complete. Assume that f, g : X → X and f be a g-non-decreasing mapping w.r.t (cid:22). Suppose that there exists constants A, B, C ≥ 0 with (A + 2B + 2C) ∈ (0, 1) such that (34) d(f x, f y) ≤ λ[d(gx, f x) + d(gy, f y)], for all x, y ∈ X with gx (cid:22) gy. Suppose also that the following conditions hold: (i) there exists x0 ∈ X s.t gx0 (cid:22) f x0; (ii) (X,(cid:22), d) is g-regular. Also suppose g(X) is closed. Then, f and g have a coincidence point. Moreover, if for every pair (x, y) ∈ C(g, f ) × C(g, f ) there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, and if f and g commute at their coincidence points, then we obtain uniqueness of the common fixed point. Corollary 4.16. Let (X,(cid:22)) be a partially ordered set and d be a metric on X such that (X, d) is complete. Assume that f, g : X → X and f be a g-non-decreasing mapping w.r.t (cid:22). Suppose that there exists constants A, B, C ≥ 0 with (A + 2B + 2C) ∈ (0, 1) such that (35) d(f x, f y) ≤ λ[d(gx, f y) + d(gy, f x)], for all x, y ∈ X with gx (cid:22) gy. Suppose also that the following conditions hold: (i) there exists x0 ∈ X s.t gx0 (cid:22) f x0; (ii) (X,(cid:22), d) is g-regular. Also suppose g(X) is closed. Then, f and g have a coincidence point. Moreover, if for every pair (x, y) ∈ C(g, f ) × C(g, f ) there exists z ∈ X such that gx (cid:22) gz and gy (cid:22) gz, and if f and g commute at their coincidence points, then we obtain uniqueness of the common fixed point. Remarks • Letting g = IX in Corollary 4.11 we obtain Corollary 3.12 in [10]. • Letting g = IX in Corollary 4.12 we obtain Corollary 3.13 in [10]. • Letting g = IX in Corollary 4.13 we obtain Corollary 3.14 in [10]. • Letting g = IX in Corollary 4.14 we obtain Corollary 3.15 in [10]. • Letting g = IX in Corollary 4.15 we obtain Corollary 3.16 in [10]. • Letting g = IX in Corollary 4.16 we obtain Corollary 3.17 in [10]. 12 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA 4.3. Fixed Point Theorems for Cyclic Contractive Mappings. As a generalization of the Banach contraction mapping principle, Kirk et al. [22] in 2003 introduced cyclic representations and cyclic contractions. A mapping T : A∪ B → A∪ B is called cyclic if T (A) ⊆ B and T (B) ⊆ A, where A, B are nonempty subsets of a metric space (X, d). Moreover, T is called a cyclic contraction if there exists k ∈ (0, 1) such that d(T x, T y) ≤ kd(x, y) for all x ∈ A and y ∈ B. Notice that although a contraction is continuous, cyclic contractions need not be. This is one of the important gains of this theorem. In the last decade, several authors have used the cyclic representations and cyclic contractions to obtain various fixed point results. see for example ([23, 24, 25, 26, 27, 28]). Corollary 4.17. Let (X, d) be a complete metric space, A1 and A2 are two nonempty closed subsets of X and f, g : Y → Y be two mappings, where Y = A1 ∪ A2. Suppose that the following conditions hold: (i) g(A1) and g(A2) are closed; (ii) f (A1) ⊆ g(A2) and f (A2) ⊆ g(A1); (iii) g is one-to-one; (iv) there exists a function ψ ∈ Ψ such that (36) d(f x, f y) ≤ ψ(M (gx, gy)), ∀(x, y) ∈ A1 × A2. Then, f and g have a coincidence point z ∈ A1 ∩ A2. Further, if f , g commute at their coincidence points, then f and g have a unique common fixed point that belongs to A1 ∩ A2. Proof. Due to the fact that g is one-to-one, condition (iv) is equivalent to (37) d(f x, f y) ≤ ψ(M (gx, gy)), ∀(gx, gy) ∈ g(A1) × g(A2). Now, since A1 and A2 are closed subsets of the complete metric space (X, d), then (Y, d) is complete. Define the mapping α : Y × Y → [0,∞) by (38) α(x, y) =( 1 if (x, y) ∈ (g(A1) × g(A2)) ∪ (g(A2) × g(A1)) 0 otherwise Notice that in view of definition of α and condition (iv), we can write (39) α(gx, gy)d(f x, f y) ≤ ψ(M (gx, gy)) for all gx ∈ g(A1) and gy ∈ g(A2). Thus, the pair (f, g) is a generalized α-ψ contractive pair of mappings. By using condition (ii), we can show that f (Y ) ⊆ g(Y ). Moreover, g(Y ) is closed. Next, we proceed to show that f is α-admissible w.r.t g. Let (gx, gy) ∈ Y ×Y such that α(gx, gy) ≥ 1; that is, (40) (gx, gy) ∈ (g(A1) × g(A2)) ∪ (g(A2) × g(A1)) Since g is one-to-one, this implies that (41) (x, y) ∈ (A1 × A2) ∪ (A2 × A1) So, from condition (ii), we infer that (42) (f x, f y) ∈ (g(A2) × g(A1)) ∪ (g(A1) × g(A2)) that is, α(f x, f y) ≥ 1. This implies that f is α-admissible w.r.t g. Now, let {gxn} be a sequence in X such that α(gxn, gxn+1) ≥ 1 for all n and gxn → gz ∈ g(X) as n → ∞. From the definition of α, we infer that (43) (gxn, gxn+1) ∈ (gA1 × gA2) ∪ (gA2 × gA1) 13 Since (gA1 × gA2) ∪ (gA2 × gA1) is a closed set with respect to the Euclidean metric, we get that (44) (gz, gz) ∈ (gA1 × gA2) ∪ (gA2 × gA1), thereby implying that gz ∈ g(A1) ∩ g(A2). Therefore, we obtain immediately from the definition of α that α(gxn, gz) ≥ 1 for all n. Now, let a be an arbitrary point in A1. We need to show that α(ga, f a) ≥ 1. Indeed, from condition (ii), we have f a ∈ g(A2). Since ga ∈ g(A1), we get (ga, f a) ∈ g(A1) × g(A2), which implies that α(ga, f a) ≥ 1. Now, all the hypotheses of Theorem 3.1 are satisfied. Hence, we deduce that f and g have a coincidence point z ∈ A1 ∪ A2, that is, f z = gz. If z ∈ A1, from (ii), f z ∈ g(A2). On the other hand, f z = gz ∈ g(A1). Then, we have gz ∈ g(A1) ∩ g(A2), which implies from the one-to-one property of g that z ∈ A1 ∩ A2. Similarly, if z ∈ A2, we obtain that z ∈ A1 ∩ A2. Notice that if x is a coincidence point of f and g, then x ∈ A1 ∩ A2. Finally, let x, y ∈ C(g, f ), that is, x, y ∈ A1 ∩ A2, gx = f x and gy = f y. Now, from above observation, we have w = x ∈ A1 ∩ A2, which implies that gw ∈ g(A1 ∩ A2) = g(A1) ∩ g(A2) due to the fact that g is one-to-one. Then, we get that α(gx, gw) ≥ 1 and α(gy, gw) ≥ 1. Then our claim holds. Now, all the hypotheses of Theorem 3.2 are satisfied. So, we deduce that z = A1∩ A2 is the unique common fixed point of f and g. This completes the proof. (cid:3) The following results are immediate consequences of Corollary 4.17. Corollary 4.18. Let (X, d) be a complete metric space, A1 and A2 are two nonempty closed subsets of X and f, g : Y → Y be two mappings, where Y = A1 ∪ A2. Suppose that the following conditions hold: (i) g(A1) and g(A2) are closed; (ii) f (A1) ⊆ g(A2) and f (A2) ⊆ g(A1); (iii) g is one-to-one; (iv) there exists a function ψ ∈ Ψ such that d(f x, f y) ≤ ψ(d(gx, gy)), ∀(x, y) ∈ A1 × A2. Then, f and g have a coincidence point z ∈ A1 ∩ A2. Further, if f , g commute at their coincidence points, then f and g have a unique common fixed point that belongs to A1 ∩ A2. Corollary 4.19. Let (X, d) be a complete metric space, A1 and A2 are two nonempty closed subsets of X and f, g : Y → Y be two mappings, where Y = A1 ∪ A2. Suppose that the following conditions hold: (i) g(A1) and g(A2) are closed; (ii) f (A1) ⊆ g(A2) and f (A2) ⊆ g(A1); (iii) g is one-to-one; (iv) there exists a constant λ ∈ (0, 1) such that (cid:27) ∀(x, y) ∈ A1 × A2. d(f x, f y) ≤ λ max(cid:26)d(gx, gy), Then, f and g have a coincidence point z ∈ A1 ∩ A2. Further, if f , g commute at their coincidence points, then f and g have a unique common fixed point that belongs to A1 ∩ A2. Corollary 4.20. Let (X, d) be a complete metric space, A1 and A2 are two nonempty closed subsets of X and f, g : Y → Y be two mappings, where Y = A1 ∪ A2. Suppose that the following conditions hold: (i) g(A1) and g(A2) are closed; (ii) f (A1) ⊆ g(A2) and f (A2) ⊆ g(A1); d(gx, f x) + d(gy, f y) 2 d(gx, f y) + d(gy, f x) , 2 14 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA (iii) g is one-to-one; (iv) there exists a constant λ ∈ (0, 1) such that d(f x, f y) ≤ Ad(gx, gy) + B[d(gx, f x) + d(gy, f y)] + C[d(gx, f y) + d(gy, f x)], ∀(x, y) ∈ A1 × A2. Then, f and g have a coincidence point z ∈ A1 ∩ A2. Further, if f , g commute at their coincidence points, then f and g have a unique common fixed point that belongs to A1 ∩ A2. Corollary 4.21. Let (X, d) be a complete metric space, A1 and A2 are two nonempty closed subsets of X and f, g : Y → Y two mappings, where Y = A1 ∪ A2. Suppose that the following conditions hold: (i) g(A1) and g(A2) are closed; (ii) f (A1) ⊆ g(A2) and f (A2) ⊆ g(A1); (iii) g is one-to-one; (iv) there exists a constant λ ∈ (0, 1) such that d(f x, f y) ≤ λ(d(gx, gy)), ∀(x, y) ∈ A1 × A2. Then, f and g have a coincidence point z ∈ A1 ∩ A2. Further, if f , g commute at their coincidence points, then f and g have a unique common fixed point that belongs to A1 ∩ A2. Corollary 4.22. Let (X, d) be a complete metric space, A1 and A2 are two nonempty closed subsets of X and f, g : Y → Y be two mappings, where Y = A1 ∪ A2. Suppose that the following conditions hold: (i) g(A1) and g(A2) are closed; (ii) f (A1) ⊆ g(A2) and f (A2) ⊆ g(A1); (iii) g is one-to-one; (iv) there exists a constant λ ∈ (0, 1) such that d(f x, f y) ≤ λ[d(gx, f x) + d(gy, f y)], ∀(x, y) ∈ A1 × A2. Then, f and g have a coincidence point z ∈ A1 ∩ A2. Further, if f , g commute at their coincidence points, then f and g have a unique common fixed point that belongs to A1 ∩ A2. Corollary 4.23. Let (X, d) be a complete metric space, A1 and A2 are two nonempty closed subsets of X and f, g : Y → Y be two mappings, where Y = A1 ∪ A2. Suppose that the following conditions hold: (i) g(A1) and g(A2) are closed; (ii) f (A1) ⊆ g(A2) and f (A2) ⊆ g(A1); (iii) g is one-to-one; (iv) there exists a constant λ ∈ (0, 1) such that d(f x, f y) ≤ λ[d(gx, f y) + d(gy, f x)], ∀(x, y) ∈ A1 × A2. Then, f and g have a coincidence point z ∈ A1 ∩ A2. Further, if f , g commute at their coincidence points, then f and g have a unique common fixed point that belongs to A1 ∩ A2. Remarks • Letting g = IX in Corollary 4.18 we obtain Corollary 3.19 in [10]. • Letting g = IX in Corollary 4.19 we obtain Corollary 3.20 in [10]. • Letting g = IX in Corollary 4.20 we obtain Corollary 3.21 in [10]. • Letting g = IX in Corollary 4.21 we obtain Corollary 3.22 in [10]. • Letting g = IX in Corollary 4.22 we obtain Corollary 3.23 in [10]. • Letting g = IX in Corollary 4.23 we obtain Corollary 3.24 in [10]. 5. Acknowledgement 15 The first author gratefully acknowledges the University Grants Commission, Government of India for financial support during the preparation of this manuscript. References [1] Banach, S.: Surles operations dans les ensembles abstraits et leur application aux equations itegrales, Funda- menta Mathematicae 3, 133-181 (1922). [2] Caccioppoli, R.: Un teorema generale sullesistenza di elementi uniti in una trasformazione funzionale, Rendi- contilincei: Matematica E Applicazioni. 11, 794-799 (1930). (in Italian). [3] Kannan, R.: Some results on fixed points, Bull. Calcutta. Math. Soc. 10, 71-76 (1968). [4] Bhaskar, T. G., Lakshmikantham, V.:Fixed Point Theory in partially ordered metric spaces and applications, Nonlinear Analysis 65, 1379-1393 (2006). [5] Branciari, A.:A fixed point theorem for mappings satisfying a general contractive condition of integral type, Int. J. Math. Math. Sci. 29, 531-536 (2002). [6] Lakshmikantham, V., ´Ciri´c, L.:Coupled fixed point theorems for nonlinear contractions in partially ordered metric spaces, Nonlinear Analysis 70, 4341 -- 4349 (2009). [7] Nieto, J. J., Lopez, R. R.:Contractive mapping theorems in partially ordered sets and applications to ordinary differential equations, Order 22, 223-239 (2005). [8] Saadati, R., Vaezpour, S. M.:Monotone generalized weak contractions in partially ordered metric spaces, Fixed Point Theory 11, 375-382 (2010). [9] Samet, B., Vetro, C., Vetro, P.: Fixed point theorem for α-ψ contractive type mappings, Nonlinear Anal. 75, 2154-2165 (2012). [10] Karapinar, E., Samet, B.:Generalized α-ψ-contractive type mappings and related fixed point theorems with applications, Abstract and Applied Analysis 2012 Article ID 793486, 17 pages doi:10.1155/2012/793486. [11] ´Ciri´c, L., Cakic, N., Rajovic, M., Ume, J.S.: Monotone generalized nonlinear contractions in partially ordered metric spaces, Fixed Point Theory Appl. 2008(2008), Article ID 131294, 11 pages. [12] Aydi, H., Nashine, H.K., Samet, B., Yazidi, H.: Coincidence and common fixed point results in partially ordered cone metric spaces and applications to integral equations, Nonlinear Analysis 74, 6814-6825 (2011). [13] Berinde, V.:Iterative Approximation of fixed points, Editura Efemeride, Baia Mare, 2002. [14] ´Ciri´c, L.: Fixed points for generalized multi-valued mappings, Mat. Vesnik. 24, 265-272 (1972). [15] Hardy, G. E., Rogers, T. D.: A generalization of a fixed point theorem of Reich, Canad. Math. Bull. 16, 201-206 (1973). [16] Chatterjea, S.K.: Fixed point theorems, C. R. Acad. Bulgare Sci. 25, 727-730 (1972). [17] Turinici, M.: Abstract comparison principles and multivariable Gronwall-Bellman inequalities, J. Math. Anal. Appl. 117, 100 -- 127 (1986). [18] Ran, A. C. M., Reurings, M. C. B.: A fixed point theorem in partially ordered sets and some applications to matrix equations, Proc. Amer. Math. Soc. 132, 1435-1443 (2004). [19] Nieto, J. J., Lopez, R. R.:Existence and uniqueness of fixed point in partially ordered sets and applications to ordinary differential equations, Acta Math. Sinica, Eng. Ser. 23 2205-2212 (2007). [20] Agarwal, R. P., El-Gebeily, M. A., Regan, D. O':Generalized contractions in partially ordered metric spaces, Applicable Analysis 87, 1-8 (2008). 16 PRIYA SHAHI, JATINDERDEEP KAUR, S. S. BHATIA [21] Samet, B.: Coupled fixed point theorems for a generalized Meir-Keeler contraction in partially ordered metric spaces, Nonlinear Anal. TMA (2010) doi:10.1016/j.na.2010.02.026 [22] Kirk, W. A., Srinivasan, P. S., Veeramani, P.: Fixed points for mappings satisfying cyclical contractive condi- tions, Fixed Point Theory 4, 79-89 (2003). [23] Agarwal, R. P., Alghamdi, M. A., Shahzad, N.: Fixed point theory for cyclic generalized contractions in partial metric spaces, Fixed Point Theory Appl. (2012), 2012:40. [24] Karapinar E., Fixed point theory for cyclic weak φ-contraction, Appl. Math. Lett. 24, 822-825 (2011). [25] Karapinar, E. and Sadaranagni, K.: Fixed point theory for cyclic (φ − ψ)-contractions, Fixed point theory Appl. 2011, 2011:69. [26] Pacurar, M., Rus I. A.:Fixed point theory for cyclic ϕ-contractions, Nonlinear Anal. 72, 1181-1187 (2010). [27] Petric, M. A.:Some results concerning cyclic contractive mappings, General Mathematics 18, 213-226 (2010). [28] Rus, I. A.:Cyclic representations and fixed points, Ann. T. Popovicin. Seminar Funct. Eq. Approx. Convexity 3, 171-178 (2005).
1602.01902
1
1602
2016-02-05T00:59:32
Sharp pointwise estimates for functions in the Sobolev spaces Hs(Rn)
[ "math.FA" ]
We obtain the optimal value of the constant K(n,s) in the Sobolev-Nirenberg-Gagliardo inequality $ \|\,u\,\|_{L^{\infty}(\mathbb{R}^{n})} \leq K(n,s) \,\|\, u \,\|_{L^{2}(\mathbb{R}^{n})}^{1 - n/(2s)} \|\, u \,\|_{\dot{H}^{s}(\mathbb{R}^{n})}^{n/(2s)} $ where $ s > n/2 $.
math.FA
math
Sharp pointwise estimates for functions in the Sobolev spaces Hs(Rn) Lineia Schutz 1, Juliana S. Ziebell2, Jana´ına P. Zingano1 and Paulo R. Zingano1 1 Instituto de Matem´atica e Estat´ıstica Universidade Federal do Rio Grande do Sul Porto Alegre, RS 91509-900, Brazil 2 Instituto de Matem´atica, Estat´ıstica e F´ısica Universidade Federal do Rio Grande Rio Grande, RS 96201-900, Brazil Abstract We provide the optimal value of the constant K(n, m) in the Gagliardo-Nirenberg L2(Rn), m > n/2, and supnorm inequality k u kL∞(Rn) ≤ K(n, m)k u k its generalizations to the Sobolev spaces Hs(Rn) of arbitrary order s > n/2 as well. L2(Rn)k Dmu k 1 − n 2m n 2m 1. Introduction In recent decades there has been a growing interest in determining the sharpest form of many important inequalities in analysis, see e.g. [1, 2, 3, 4, 5, 8, 10] and references therein. A noticeable miss is the fundamental Gagliardo-Nirenberg supnorm inequality k u kL∞(Rn) ≤ K(n, m) k u k 1 − n 2m L2(Rn) k Dmuk n 2m L2(Rn) (1.1) for functions u ∈ H m(Rn) when m > n/2 (see [7, 9]), as well as some of its generalizations. Here, as usual, H m(Rn) is the Sobolev space of functions u ∈ L2(Rn) with all derivatives of order up to m in L2(Rn), which is a Banach space under its natural norm defined by 6 1 0 2 b e F 5 ] . A F h t a m [ 1 v 2 0 9 1 0 . 2 0 6 1 : v i X r a k u kHm(Rn) + k Dm uk2 L2(Rn)o1/2 , L2(Rn) where k Dm ukL2(Rn) = n k u k2 nXi2 = 1 ... = (cid:26) nXi1 = 1 nXim = 1 k Di1 Di2 ... Dim L2(Rn)(cid:27)1/2 u k2 (1.2a) (1.2b) (with Di denoting the weak derivative with respect to the variable xi). In this brief note, we will review some basic results in order to derive the optimal (i.e., minimal) value for the constant K(m, n) in (1.1) above. It will be seen in Section 2 that it turns out to be 1 4(cid:26) n K(n, m) = (cid:8)4π(cid:9)− n where σ (r) := rπ and Γ(·) is the Gamma function (for its definition, see e.g. [6], p. 7). For example, we get, with m = 2 and n = 1, 2, 3, the sharp pointwise estimates 4m(cid:26) 2m 2m − n(cid:27)1 2m − n(cid:27)− n 2m ) (cid:27)− 1 2(cid:26) 2(cid:17)(cid:27)− 1 Γ(cid:16) n 2(cid:26) sin σ( n σ( n 2m ) (1.3) n 2 2 k u kL∞(R) ≤ 4√2 8√27 k u k 3 4 L2(R)k D2u k 1 4 L2(R) , k u kL∞(R2) ≤ k u kL∞(R3) ≤ 1 2 1 2 k u k 8√12 √ 6π k u k 1 4 L2(R2)k D2u k 1 2 L2(R2) , L2(R3)k D2u k 3 4 L2(R3) (1.4a) (1.4b) (1.4c) , and so forth. (In [10], it is obtained that k u kL∞(R3) ≤ which is also shown to be optimal.) Also, setting 1 √ 2π k D u k L2(R3)k D2uk 1 2 1 2 L2(R3) , k u k Hs(Rn) := (cid:26)ZRn ξ2s u(ξ)2 dξ(cid:27)1 2 for real s > 0,1 where u(·) denotes the Fourier transform of u(·), that is, 2ZRn u(ξ) = (cid:0)2π(cid:1)− n − i x · ξ e u(x) dx, ξ ∈ Rn, (1.5) (1.6) and letting Hs(Rn) = (cid:8) u ∈ L2(Rn) : k u k Hs(Rn) < ∞(cid:9) be the Sobolev space of order s, we get the following generalization of (1.1), (1.3) above. If s > n/2, then k u kL∞(Rn) ≤ K(n, s) k u k 1 − n 2s L2(Rn)k u k n 2s Hs(Rn) for all u ∈ Hs(Rn), with the optimal value of the constant K(n, s) being given by K(n, s) = (cid:8)4π(cid:9)− n 2s − n(cid:27)− n 2(cid:26) n 2s − n(cid:27)1 4s(cid:26) 2s 2(cid:17)(cid:27)− 1 Γ(cid:16) n 2(cid:26) sin σ( n 2s ) (cid:27)− 1 4(cid:26) n σ( n 2s ) 2 2 (1.7a) (1.7b) for any s > n/2, where, as before, σ (r) := r π. The proof of (1.1), (1.3), (1.7) and of the sharpness of the values for K given in (1.3), (1.7b) above is provided in the sequel; when u ∈ Hs(Rn), s > n/2, is also in addition, a second classical estimate for k u kL∞(Rn) reexamined here (see (2.4) below), so as to be similarly presented in its sharpest form. 1Note that (1.5) corresponds to (1.2b) when s = m (m integral), that is: k u k Hm(Rn) = k D m 2 u k L 2(Rn) . 2. Proof of (1.1), (1.3), (1.7) and other optimal supnorm results in Hs(Rn) To obtain the results stated in Section 1, we first review the following basic lemma. We recall that u denotes the Fourier transform of u, cf. (1.6) above. Lemma 2.1. Let u ∈ L2(Rn). If u ∈ L1(Rn), then u ∈ L∞(Rn) and k u kL∞(Rn) ≤ (cid:0) 2π(cid:1)− n/2 k u kL1(Rn) . (2.1) Moreover, equality holds in (2.1) if u is real-valued and of constant sign (say, nonnegative). Proof: Clearly, (2.1) is valid if u ∈ S(Rn), where S(Rn) denotes the Schwartz class of smooth, rapidly decreasing functions at infinity ([6], p. 4), since we have, in this case, the representation (see e.g. [6], p. 16) 2ZRn um(x) = (cid:0)2π(cid:1)− n e i x · ξ um(ξ) dξ, ∀ x ∈ Rn. (2.2) For general u ∈ L2(Rn) with u ∈ L1(Rn), let { um } be a sequence of Schwartz approximants to u ∈ L1(Rn) ∩ L2(Rn) such that k um − ukL1(Rn) → 0 and k um − ukL2(Rn) → 0 as m → ∞, and let um ∈ S(Rn) be the Fourier inverse of um, for each m. Applying (2.1) to { um }, we see that { um} is Cauchy in L∞(Rn), so that k um − v kL∞(Rn) → 0 for some v ∈ L∞(Rn) ∩ C0(Rn). Since we have k um − u kL2(Rn) → 0, it follows that u = v. This shows that u ∈ L∞(Rn) ∩ C0(Rn) and, letting m → ∞ in (2.2), we also have 2ZRn u(x) = (cid:0)2π(cid:1)− n e i x ·ξ u(ξ) dξ, ∀ x ∈ Rn, (2.2′) from which (2.1) immediately follows. In particular, in the event that u(ξ) ≥ 0 for all ξ, we get from (2.2′) that u(0) = (2π)− n/2k u kL1(Rn), so that (2.1) becomes an identity in this case. (cid:3) We observe that the hypotheses of Lemma 2.1 are satisfied for u ∈ Hs(Rn) if s > n/2. An important consequence of this fact is the fundamental embedding property revisited next, where the norm in Hs(Rn) is set to be (in accordance with (1.2a) above): k u kHs(Rn) = n k u k2 L2(Rn) + k u k2 Hs(Rn)o1/2 = (cid:26)ZRn(cid:0) 1 + ξ2s(cid:1) u(ξ)2 dξ(cid:27)1/2 . (2.3) Theorem 2.1. Let s > n/2. If u ∈ Hs(Rn), then u ∈ L∞(Rn) ∩ C 0(Rn) and k u kL∞(Rn) ≤ (cid:8)4π(cid:9)− n 4(cid:26) n 2 2(cid:17)(cid:27)− 1 Γ(cid:16) n 2(cid:26) sin σ( n 2s ) (cid:27)− 1 σ( n 2s ) 2 , · k u kHs(Rn) (2.4) where σ(r) = rπ. Moreover, equality holds in (2.4) when u(ξ) = c / ( 1 + ξ2s ) ∀ ξ ∈ Rn for some constant c ≥ 0, so that the constant given in (2.4) above is optimal. 3 Proof: Let u ∈ Hs(Rn), with s > n/2. By (2.1) and Cauchy-Schwarz's inequality, we have k u kL∞(Rn) ≤ (cid:0)2π(cid:1)− n/2ZRn(cid:0) 1 + ξ2s(cid:1)− 1/2(cid:0) 1 + ξ2s(cid:1)1/2 u(ξ) dξ ≤ (cid:0)2π(cid:1)− n/2(cid:26)ZRn(cid:0) 1 + ξ2s(cid:1)− 1 = (cid:8)4π(cid:9)− n/4(cid:26) n 2 2(cid:17)(cid:27)− 1/2(cid:26) sin σ( n Γ(cid:16) n 2s ) (cid:27)− 1/2 σ( n 2s ) · k u kHs(Rn) dξ(cid:27)1/2(cid:26)ZRn(cid:0) 1 + ξ2s(cid:1) u(ξ)2 dξ(cid:27)1/2 (2.5a) (2.5b) by (2.3), since, using polar coordinates and the change of variable t =(cid:0) 1 + r2s(cid:1)− 1 ZRn(cid:0) 1 + ξ2s(cid:1)− 1 0 (cid:0) 1 + r2s(cid:1)− 1 dξ = ωnZ ∞ 2sZ 1 2s (1 − t) rn−1 dr = n 2s − 1 dt = − n ωn t 0 , we obtain ωn n σ(cid:0) n 2s(cid:1) 2s(cid:1) sin σ(cid:0) n where ωn = 2 πn/2/ Γ(n/2) is the surface area of the unit ball in Rn (see [6], p. 8), and σ(r) = rπ. This shows (2.4). Finally, if u(ξ) = c/(1 + ξ2s) for all ξ, for some c ≥ 0 constant, then equality holds in both steps (2.5a) and (2.5b) above, so that (2.4) is an identity in this case, as claimed. (cid:3) We are now in very good standing to obtain (1.1), (1.7a) with their sharpest constants. Theorem 2.2. Let s > n/2. If u ∈ Hs(Rn), then k u kL∞(Rn) ≤ K(n, s) k u k 1 − n 2s L2(Rn)k u k n 2s Hs(Rn) (2.6) with K(n, s) defined in (1.7b). Moreover, equality holds in (2.6) if u(ξ) = c / ( 1 + ξ2s ) ∀ ξ ∈ Rn, c ≥ 0 constant, so that the numerical value of K given in (1.7b) is optimal. Proof: Let s > n/2, u ∈ Hs(Rn) fixed. Given λ > 0, setting uλ ∈ Hs(Rn) by uλ(x) := u(λx), we have, by (2.4), Theorem 2.1, k u kL∞(Rn) = k uλkL∞(Rn) ≤ (cid:8)4π(cid:9)− n 2(cid:17)(cid:27)− 1 Γ(cid:16) n = (cid:8)4π(cid:9)− n 4 (cid:26) n 2 2 2s ) 2s ) (cid:27)− 1 2 2(cid:17)(cid:27)− 1 4 (cid:26) n 2(cid:26) sin σ( n Γ(cid:16) n 2s ) (cid:27)− 1 2(cid:26) sin σ( n · (cid:26) λ σ( n σ( n 2s ) − n k u k 2 · k uλkHs(Rn) 2 L2(Rn) 2s − n + λ k u k 2 2 Hs(Rn)(cid:27)1 for λ > 0 arbitrary. Choosing λ that minimizes the last term on the right of the above expression gives us (2.6), with K(n, s) defined by (1.7b), as claimed. Now, to show that the value provided in (1.7b) is the best possible, we proceed as follows. First, we observe that, by Young's inequality, 2s − n(cid:27)− n (cid:26) n 2s − n(cid:27)1 4s(cid:26) 2s 2 1 − n 2s L2(Rn) k u k k u k n 2s Hs(Rn) ≤ n k u k 2 L2(Rn) + k u k 4 2 2 Hs(Rn)o1 = k u kHs(Rn) (2.7) for all u ∈ Hs(Rn). Therefore, taking w ∈ Hs(Rn) defined by bw(ξ) = c/(1 + ξ2s), c ≥ 0, we get, with K(n, s) given in (1.7b): k w kL∞(Rn) ≤ K(n, s) k w k 4 (cid:26) n ≤ (cid:8)4π(cid:9)− n 2 = k w kL∞(Rn) , n 2s Hs(Rn) 1 − n 2s L2(Rn)k w k 2(cid:17)(cid:27)− 1 Γ(cid:16) n 2(cid:26) sin σ( n 2s ) (cid:27)− 1 σ( n 2 2s ) · k w kHs(Rn) [ by (2.6) ] [ by (1.7b), (2.7) ] [ by Theorem 2.1 ] thus showing that we have k w kL∞(Rn) = K(n, s) k w k 1 − n 2s L2(Rn)k w k n 2s Hs(Rn) , as claimed. (cid:3) References [1] M. Agueh, Gagliardo-Nirenberg inequalities involving the gradient L2-norm, C. R. Math. Acad. Sci. Paris, 346 (2008), 757-762. [2] W. Beckner, Inequalities in Fourier analysis, Ann. Math, 102 (1975), 159-182. [3] E. A. Carlen and M. Loss, Sharp constant in Nash's inequality, Intern. Math. Research Notices, 7 (1993), 213-215. [4] D. Cordero-Erausquin, B. Nazaret and C. Villani, A mass-transportation approach to sharp Sobolev and Gagliardo-Nirenberg inequalities, Advances in Math., 182 (2004), 307-332. [5] M. Del Pino and J. Dolbeault, Best constants for Gagliardo-Nirenberg inequal- ities and applications to nonlinear diffusions, J. Math. Pures Appliqu´ees, 81 (2002), 847-875. [6] G. B. Folland, An Introduction to Partial Differential Equations (2nd ed.), Princeton University Press, Princeton, 1995. [7] E. Gagliardo, Propriet`a di alcune classi di funzioni in pi`u variabili, Ricerche Mat., 7 (1958), 102-137. [8] E. H. Lieb, Inequalities (Selecta), Springer, New York, 2003. [9] L. Nirenberg, On elliptic partial differential equations, Ann. Scuola Norm. Sup. Pisa, 13 (1959), 115-162. [10] W. Xie, A sharp pointwise bound for functions with L2-Laplacians and zero bound- ary values of arbitrary three-dimensional domains, Indiana Univ. Math. Journal, 40 (1991), 1185-1192. 5
1702.07182
2
1702
2017-04-08T11:47:46
Tingley's problem for spaces of trace class operators
[ "math.FA", "math.OA" ]
We prove that every surjective isometry between the unit spheres of two trace class spaces admits a unique extension to a surjective complex linear or conjugate linear isometry between the spaces. This provides a positive solution to Tingley's problem in a new class of operator algebras.
math.FA
math
TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS FRANCISCO J. FERN ´ANDEZ-POLO, JORGE J. GARC´ES, ANTONIO M. PERALTA, AND IGNACIO VILLANUEVA Abstract. We prove that every surjective isometry between the unit spheres of two trace class spaces admits a unique extension to a surjective complex linear or conjugate linear isometry between the spaces. This provides a positive solution to Tingley's problem in a new class of operator algebras. 1. Introduction In 1987, D. Tingley published a study on surjective isometries between the unit spheres of two finite dimensional Banach spaces, showing that any such mappings preserves antipodes points (see [38]). A deep and difficult geometric problem has been named after Tingley's contribution. Namely, let f : S(X) → S(Y ) be a surjective isometry between the unit spheres of two Banach spaces X and Y . Is f the restriction of a (unique) isometric real linear surjection T : X → Y ? This problem remains unsolved even if we asume that X and Y are 2-dimensional spaces. Despite Tingley's problem still being open for general Banach spaces, positive solutions have been found for specific cases (see, for example, [3, 6, 7, 8, 9, 10, 5, 11, 12, 13, 17, 19, 20, 21, 24, 27, 31, 32, 33, 34, 35, 36, 37, 39] and [40]), and each particular case has required strategies and proofs which are more or less independent. 0 = ℓ1, and ℓ∗ For the purposes of this note, we recall that positive solutions to Tingley's prob- lem include the following cases: f : S(c0) → S(c0) [7], f : S(ℓ1) → S(ℓ1) [9], f : S(ℓ∞) → S(ℓ∞) [8], f : S(K(H)) → S(K(H ′)), where H and H ′ are complex Hilbert spaces [27], and f : S(B(H)) → S(B(H ′)) [20]. It is well known that the natural dualities c∗ 1 = ℓ∞ admit a non-commutative counterparts in the dualities K(H)∗ = C1(H) and C1(H)∗ = B(H), where C1(H) is the space of all trace class operators on H. So, there is a natural open question concerning Tingley's problem in the case of surjective isometries between the unit spheres of two trace class spaces. In this paper we explore this problem and we prove that every surjective isometry between the unit spheres of two trace class spaces ad- mits a unique extension to a surjective complex linear or conjugate linear isometry between the spaces (see Theorem 4.1). The results are distributed in three main sections. In section 2 we establish new geometric properties of a surjective isometry f : S(X) → S(Y ) in the case in which norm closed faces of the closed unit balls of X and Y are all norm-semi-exposed, and weak∗ closed faces of the closed unit balls of X ∗ and Y ∗ are all weak∗-semi- exposed (see Corollary 2.5). Applying techniques of geometry and linear algebra, in 2010 Mathematics Subject Classification. Primary 47B49, Secondary 46A22, 46B20, 46B04, 46A16, 46E40, . Key words and phrases. Tingley's problem; extension of isometries; trace class operators. 2 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA section 3 we present a positive answer to Tingley's problem for surjective isometries f : S(C1(H)) → S(C1(H ′)) when H and H ′ are finite dimensional (see Theorem 3.7). The result in the finite dimensional case play a fundamental role in the proof of our main result. 2. Facial stability for surjective isometries between the unit spheres of trace class spaces 1 Let H be a complex Hilbert space. We are interested in different subclasses of the space K(H) of all compact operators on H. We briefly recall the basic termi- nology. For each compact operator a, the operator a∗a lies in K(H) and admits a unique square root a = (a∗a) 2 . The characteristic numbers of the operator a are precisely the eigenvalues of a arranged in decreasing order and repeated according to multiplicity. Since a belongs to K(H), only an at most countably number of its eigenvalues are greater than zero. According to the standard terminology, we usually write µn(a) for the n-th characteristic number of a. It is well known that (µn(a))n → 0. The symbol C1 = C1(H) will stand for the space of trace class operators on H, that is, the set of all a ∈ K(H) such that kak1 := ∞Xn=1 µn(a)! < ∞. λn. ∞Xn=1 We set kak∞ = kak, where the latter stands for the operator norm of a. The set C1 is a two-sided ideal in the space B(H) of all bounded linear operators on H, and (C1,k.k1) is a Banach algebra. If tr(.) denotes the usual trace on B(H) and a ∈ K(H), we know that a ∈ C1 if, and only if, tr(a) < ∞ and kak1 = tr(a). It is further known that the predual of B(H) and the dual of K(H) both can be identified with C1(H) under the isometric linear mapping a 7→ ϕa, where ϕa(x) := tr(ax) (a ∈ C1(H), x ∈ B(H)). The dualities K(H)∗ = C1(H) and C1(H)∗ = B(H) can be regarded as a non-commutative version of the natural dualities between c0, ℓ1 and ℓ∞. It is known that every element a in C1(H) can be written uniquely as a (possibly finite) sum (1) a = ∞Xn=1 λnηn ⊗ ξn, where (λn) ⊂ R+ 0 , (ξn), (ηn) are orthonormal systems in H, and kak1 = Along the paper, we shall try to distinguish between C1(H) and C1(H)∗ ≡ B(H), however the reader must be warned that we shall regard K(H) and C1(H) inside B(H). For example, when an element a in C1(H) is regarded in the form given in (1), the element s(a) = ηn ⊗ ξn, ∞Xn=1 is a partial isometry in B(H) (called the support partial isometry of a in B(H)), and it is precisely the smallest partial isometry e in B(H) satisfying e(a) = kak1. TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 3 We recall that two elements a, b ∈ B(H) are orthogonal (a ⊥ b in short) if and only if ab∗ = b∗a = 0. The relation "being orthogonal" produces a partial order ≤ in the set U(B(H)) of all partial isometries given by w ≤ s if and only if s − w is a partial isometry with s− w ⊥ w (this is the standard order employed, for example, in [1, 16]). We refer to [23, Chapter III], [14, §9], [30, Chapter II] and [29, §1.15] for the basic results and references on the spaces K(H), C1(H) and B(H). It is worth recalling that, by a result due to B. Russo [28] every surjective complex linear isometry T : C1(H) → C1(H) is of the form T (x) = vxu or of the form T (x) = vxtu (x ∈ C1(H)), where u and v are unitary elements in B(H) and xt denotes the transpose of x (compare [22, Theorem 11.2.2]). The non-commutative Clarkson-McCarthy inequalities (see [25, Theorem 2.7]) can be written as follows: (2) (kak1 + kbk1) ≤ ka + bk1 + ka − bk1 ≤ 2 (kak1 + kbk1) holds for every a and b in C1(H). It is further known that equality ka + bk1 + ka − bk1 = 2 (kak1 + kbk1) holds in (2) if and only if (a∗a)(b∗b) = 0, which is equivalent to say that a and b are orthogonal as elements in B(H) (a ⊥ b in short), or in other words s(a) and s(b) are orthogonal partial isometries in B(H) (i.e. (s(a)∗s(a), s(b)∗s(b)) and (s(a)s(a)∗, s(b)s(b)∗) are two pairs of orthogonal projections in B(H)). Conse- quently, if we fix a, b ∈ S(C1(H)) we can conclude that (3) ka ± bk1 = 2 ⇔ a ⊥ b (in C1(H)) ⇔ s(a) ⊥ s(b) (in B(H)). Let us recall a technical result due to X.N. Fang, J.H. Wang and G.G. Ding, who established it in [17] and [10], respectively. Lemma 2.1. ([17, Corollary 2.2], [10, Corollary 1]) Let X and Y be normed spaces and let f : S(X) → S(Y ) be a surjective isometry. Then for any x, y in S(X), we have kx + yk = 2 if and only if kf (x) + f (y)k = 2. (cid:3) Throughout the paper, the extreme points of a convex set C will be denoted by ∂e(C), and the symbol BX will stand for the closed unit ball of a Banach space X. We shall write T for the unit sphere of C. Following standard notation, the elements in ∂e(BC1(H)) are called pure atoms. It is known that every pure atom in C1(H) is an operator of the form η ⊗ ξ, where ξ and η are elements in S(H). Given ξ, η in a Hilbert space H, the symbol η⊗ξ will denote the rank one operator on H defined by η ⊗ ξ(ζ) = hζξiη (ζ ∈ H). Clearly η ⊗ ξ ∈ C1(H). When η ⊗ ξ is regarded as an element in C1(H), we shall identify it with the normal functional on B(H) given by η ⊗ ξ(x) = hx(η)ξi (x ∈ B(H)). As it is commonly assumed, given φ ∈ B(H)∗ and z ∈ B(H) we define φz, zφ ∈ B(H)∗ by (φz)(x) = φ(zx) and (zφ)(x) = φ(xz) (x ∈ B(H)). Accordingly with this notation, for η ⊗ ξ in C1(H) = B(H)∗, we have (η ⊗ ξ)z = η ⊗ z∗(ξ) and z(η ⊗ ξ) = z(η) ⊗ ξ, for every z ∈ B(H). We also recall an inequality established by J. Arazy in [2, Proposition in page 48]: For each projection p in B(H) and every x ∈ C1(H), we have (4) 1 + kpx(1 − p)k2 1 + k(1 − p)xpk2 kxk2 1 ≥ kpxpk2 1 + k(1 − p)x(1 − p)k2 1 4 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA Suppose {ξi}i∈I is an orthonormal basis of H. The elements in the set {ξi ⊗ ξi : i ∈ I} are mutually orthogonal in C1(H). Actually, the dimension of H is precisely the cardinal of the biggest set of mutually orthogonal pure atoms in C1(H). We can state now a non-commutative version of [9, Lemma 3]. Lemma 2.2. Let H and H ′ be complex Hilbert spaces, and let f : S(C1(H)) → S(C1(H ′)) be a surjective isometry. Then f preserves orthogonal elements in both directions, that is, a ⊥ b in S(C1(H)) if and only if f (a) ⊥ f (b) in S(C1(H ′)). Proof. Take a, b in S(C1(H)). We have already commented that a ⊥ b if and only if ka ± bk1 = 2 (compare (3)). Since f is an isometry we deduce that kf (a) − f (b)k1 = 2. Lemma 2.1 implies that kf (a) + f (b)k1 = 2, and hence kf (a)± f (b)k1 = 2, which (cid:3) assures that f (a) ⊥ f (b). Among the ingredients and prerequisites needed in our arguments we highlight the following useful geometric result which is essentially due to L. Cheng and Y. Dong [3, Lemma 5.1] and R. Tanaka [35] (see also [34, Lemma 3.5], [36, Lemmas 2.1 and 2.2]). Proposition 2.3. ([3, Lemma 5.1], [35, Lemma 3.3], [34, Lemma 3.5]) Let f : S(X) → S(Y ) be a surjective isometry between the unit spheres of two Banach spaces, and let M be a convex subset of S(X). Then M is a maximal proper face of BX if and only if f (M) is a maximal proper (closed) face of BY . (cid:3) The previous result emphasizes the importance of a "good description" of the facial structure of a Banach space. A basic tool to understand the facial structure of the closed unit ball of a complex Banach space X and that of the unit ball of its dual space is given by the "facear" and "pre-facear" operations. Following [16], for each F ⊆ BX and G ⊆ BX ∗, we define F ′ = {a ∈ BX ∗ : a(x) = 1 ∀x ∈ F}, G′ = {x ∈ BX : a(x) = 1 ∀a ∈ G}. Then, F ′ is a weak∗ closed face of BX ∗ and G′ is a norm closed face of BX . The subset F is said to be a norm-semi-exposed face of BX if F = (F ′)′, while the subset G is called a weak∗-semi-exposed face of BX ∗ if G = (G′)′. The mappings F 7→ F ′ and G 7→ G′ are anti-order isomorphisms between the complete lattices Sn(BX) of norm-semi-exposed faces of BX and Sw∗ (BX ∗) of weak∗-semi-exposed faces of BX ∗ and are inverses of each other. Our next result is a generalization of the above Proposition 2.3. Proposition 2.4. Let f : S(X) → S(Y ) be a surjective isometry between the unit spheres of two Banach spaces, and let C be a convex subset of S(X). Suppose that for every extreme point φ0 ∈ ∂e(BX ∗), the set {φ0} is a weak∗-semi-exposed face of BX ∗ . Then C is a norm-semi-exposed face of BX if and only if f (C) is a norm-semi-exposed face of BY . Proof. We begin with an observation. By Eidelheit's separation theorem [26, Theo- rem 2.2.26], every maximal proper face of BX is a norm-semi-exposed face (compare [35, Lemma 3.3]). Suppose C ∈ Sn(BX ). We set Λ+ C :=nM : M is a maximal proper face of BX containing Co. TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 5 Let us observe thatTM∈Λ+ C M is a proper face of BX containing C. Since a non- empty intersection of proper norm-semi-exposed faces of BX (respectively, weak∗- semi-exposed face of BX ∗ ) is a proper norm-semi-exposed face of BX (respectively, a proper weak∗-semi-exposed face of BX ∗ ), the set TM∈Λ+ C M is a proper norm- semi-exposed face of BX . We shall next show that (5) C M. C = \M∈Λ+ C M. Therefore, C′ % (cid:16)TM∈Λ+ C M(cid:17)′ The inclusion ⊆ is clear. To see the other inclusion we argue by contradiction, and thus we assume that C $ TM∈Λ+ . The C M(cid:17)′ sets C′ and (cid:16)TM∈Λ+ are weak∗-closed convex faces of BX ∗, and hence, by the Krein-Milman theorem C′ = cow∗ (∂e(C′)). Therefore, having in mind that C′ is a face, we can find an extreme point φ0 ∈ ∂e(C′) ⊂ ∂e(BX ∗) such that φ0 /∈ (cid:16)TM∈Λ+ C M(cid:17)′ . Since, by hypothesis, {φ0} is a weak∗-semi-exposed face of BX ∗, we can easily check that M0 = {φ0}′ is a maximal proper face of BX. Furthermore, φ0 ∈ C′ implies that C ⊆ {φ0}′ and hence M0 ∈ Λ+ C . Clearly 0 ⊆(cid:16)TM∈Λ+ C M(cid:17)′ φ0 ∈ M′ , which is impossible. We have thus proved (5). C, and thus f (C) ⊆ TM∈Λ+ C ⊆ f −1 \M∈Λ+ f (M) ⊆ \M∈Λ+ f −1 (f (M)) = \M∈Λ+ Therefore, the identity f (C) =TM∈Λ+ Since, by Proposition 2.3, for each M ∈ Λ+ C , f (M) is a maximal proper face of BX and hence norm-semi-exposed, the set f (C) coincides with a non-empty intersection of norm-semi-exposed faces, and hence f (C) is a norm-semi-exposed face too. (cid:3) Clearly, f (C) ⊆ f (M) for every M ∈ Λ+ f (M) follows from the bijectivity of f . Applying f −1 and (5) we get f (M) . C C C C M = C. C In certain classes of Banach spaces where norm closed faces are all norm-semi- exposed and weak∗ closed faces in the dual space are all weak∗-semi-exposed, the previous proposition becomes meaningful and guarantees the stability of the facial structure under surjective isometries of the unit spheres. For example, when X is a C∗-algebra or a JB∗-triple, every proper norm closed face of BX is norm-semi- exposed, and every weak∗ closed proper face of BX ∗ is weak∗-semi-exposed (see [1], [15], and [18]). The same property holds when X is the predual of a von Neumann algebra or the predual of a JBW∗-triple (see [16]). Suppose X and Y are Banach spaces satisfying the just commented property, and f : S(X) → S(Y ) is a surjective isometry. Clearly f maps proper norm closed faces of BX to proper norm closed faces of BY and preserves the order given by the natural inclusion. In this particular setting, for each extreme point e ∈ BX , the set {e} is a minimal norm-semi-exposed face of BX and hence {f (e)} = f ({e}) must be a minimal norm closed face of BY , and thus f (e) ∈ ∂e(BY ). All these facts are stated in the next corollary. 6 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA Corollary 2.5. Let X and Y be Banach spaces satisfying the following two prop- erties (1) Every norm closed face of BX (respectively, of BY ) is norm-semi-exposed; (2) Every weak∗ closed proper face of BX ∗ (respectively, of BY ∗ ) is weak∗-semi- exposed. Let f : S(X) → S(Y ) be a surjective isometry. The following statements hold: (a) Let F be a convex set in S(X). Then F is a norm closed face of BX if and (b) Given e ∈ S(X), we have that e ∈ ∂e(BX ) if and only if f (e) ∈ ∂e(BY ). (cid:3) only if f (F ) is a norm closed face of BY ; We have already commented that Corollary 2.5 holds when X and Y are von Neumann algebras, or predual spaces of von Neumann algebras, or more generally, JBW∗-triples or predual spaces of JBW∗-triples. It is well known that C1(H) is the predual of B(H). Proposition 2.6. Let f : S(C1(H)) → S(C1(H ′)) be a surjective isometry, where H and H ′ are complex Hilbert spaces. Then the following statements hold: (a) A subset F ⊂ S(C1(H)) is a proper norm-closed face of BC1(H) if and only if (b) f maps ∂e(BC1(H)) into ∂e(BC1(H ′)); (c) dim(H) =dim(H ′). (d) For each e0 ∈ ∂e(BC1(H)) we have f (ie0) = if (e0) or f (ie0) = −if (e0); (e) For each e0 ∈ ∂e(BC1(H)) if f (ie0) = if (e0) (respectively, f (ie0) = −if (e0)) then f (λe0) = λf (e0) (respectively, f (λe0) = λf (e0)) for every λ ∈ C with λ = 1. f (F ) is. Proof. (a) and (b) are consequences of Corollary 2.5. Having in mind that the dimension of H is precisely the cardinal of the biggest set of mutually orthogonal pure atoms in C1(H), statement (c) follows from (b) and Lemma 2.2. (d) Let e0 ∈ ∂e(BC1(H)). Let us set {e0}⊥ := {x ∈ C1(H) : x ⊥ e0}. By Lemma 2.2 we have f(cid:0){e0}⊥ ∩ S(C1(H))(cid:1) = {f (e0)}⊥ ∩ S(C1(H ′)) and f (Te0) = f(cid:0){e0}⊥⊥ ∩ S(C1(H))(cid:1) = {f (e0)}⊥⊥ ∩ S(C1(H ′)) = Tf (e0). Therefore f (ie0) = µf (e0) for a suitable µ ∈ T. Since 1 − µ = kf (e0) − f (ie0)k1 = ke0 − ie0k1 = 1 − i = √2, we deduce that µ ∈ {±i}, as desired. (e) Suppose e0 ∈ ∂e(BC1(H)) and f (ie0) = if (e0). Let λ ∈ T. Arguing as above, we prove that f (λe0) = µf (e0) for a suitable µ ∈ T. The identities 1 − µ = kf (e0) − f (λe0)k1 = ke0 − λe0k1 = 1 − λ, and i − µ = kif (e0) − f (λe0)k1 = kf (ie0) − f (λe0)k1 = kie0 − λe0k1 = i − λ, prove that µ = λ. (cid:3) Whenever we have a surjective isometry f : S(C1(H)) → S(C1(H ′)), we deduce from the above proposition that H and H ′ are isometrically isomorphic, we can therefore restrict our study to the case in which H = H ′. TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 7 We complete this section by recalling a result established by C.M. Edwards and G.T. Ruttimann in [16] (later rediscovered in [1]). More concretely, as a consequence of the result proved by C.M. Edwards and G.T. Ruttimann in [16, Theorem 5.3], we know that every proper norm-closed face F of BC1(H) is of the form (6) for a unique partial isometry w ∈ B(H). Furthermore the mapping w 7→ {w}′ is an order preserving bijection between the lattices of all partial isometries in B(H) and all norm closed faces of BC1(H). F = {w}′ = {x ∈ C1(H) = B(H)∗ : kxk1 = 1 = x(w)}, If H is a finite dimensional complex Hilbert space, a maximal (or complete) partial isometry w ∈ B(H) is precisely a unitary element. Therefore by the just commented result ([16, Theorem 5.3], see also [1, Theorem 4.6]) every maximal proper (norm-closed) face M of BC1(H) is of the form (7) for a unique unitary element u ∈ B(H). M = {u}′ = {x ∈ C1(H) = B(H)∗ : kxk1 = 1 = x(u)}, 3. Surjective isometries between the unit spheres of two finite dimensional trace class spaces In this section we present a positive solution to Tingley's problem for surjective isometries f : S(C1(H)) → S(C1(H)), in the case in which H is a finite dimensional complex Hilbert space. Our next result is a first step towards a solution to Tingley's conjecture in M2(C) when the latter is equipped with the trace norm. Proposition 3.1. Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is a two dimensional complex Hilbert space. Suppose e1, e2 is a (maximal) set of mutually orthogonal pure atoms in S(C1(H)) and λ1, λ2 are real numbers satisfying λ1 + λ2 = 1. Then f (λ1e1 + λ2e2) = λ1f (e1) + λ2f (e2). Proof. Under these assumptions B(H) is M2(C) with the spectral or operator norm, and C1(H) is M2(C) with the trace norm. We can assume the existence of orthonor- λ1λ2 = 0. Proposition 2.6((d) and (e)) shows that the desired statement is true when mal basis of H {η1, η2}, {ξ1, ξ2}, {eη1,eη2} and {eξ1,eξ2} such that ej = ηj ⊗ ξj and f (ηj ⊗ ξj) =eηj ⊗eξj for every j = 1, 2 (compare Proposition 2.6(b) and Lemma 2.2). basis {η1, η2}, {eη1,eη2} and {eξ1,eξ2} to the basis {ξ1, ξ2}, respectively. Let T1, T2 : C1(H) → C1(H) be the surjective complex linear isometries defined by T1(x) = u1x and T2(x) = u2xv2 (x ∈ C1(H)). We set g = T2S(C1(H)) ◦ f ◦ T −1 S(C1(H)) : S(C1(H)) → S(C1(H)). Then g is a surjective isometry satisfying To simplify the notation, let u1, u2 and v∗ 2 be the unitaries in B(H) mapping the 1 g(φj ) = φj, where φj = ξj ⊗ ξj, ∀1 ≤ j ≤ 2. We can chose a matricial representation (i.e. {ξ1, ξ2}) such that φ1 =(cid:18) 1 0 0 0 (cid:19) and φ2 =(cid:18) 0 0 0 1 (cid:19). the representation on the basis 8 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA We recall that every maximal proper (norm-closed) face M of BC1(H) is of the form M = {u}′ = {x ∈ C1(H) = B(H)∗ : kxk1 = 1 = x(u) = 1}, for a unique unitary element u ∈ B(H) (see (7) or [16, Theorem 5.3]). We note that φ1, φ2 ∈ {1}′ and {1}′ is precisely the set of all normal states on B(H). c By Proposition 2.3 we know that g({1}′) = {u}′, for a unique unitary element u ∈ B(H). Since φj = g(φj) ∈ {u}′ we can easily check that u = 1. We shall first assume that 0 ≤ λj for every j. Since λ1 + λ2 = 1 we have g(λ1φ1 + λ2φ2) ∈ g({1}′) = {1}′. Therefore, the element g(λ1φ1 + λ2φ2) must be a positive matrix (cid:18) t 0 ≤ t ≤ 1 and(cid:13)(cid:13)(cid:13)(cid:18) t (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) t 1 − t (cid:19) with trace and trace norm equal to one (i.e., 1 − t (cid:19)(cid:13)(cid:13)(cid:13)1 0 0 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 1 − t (cid:19) −(cid:18) 1 0 c = kλ1φ1 + λ2φ2 − φ1k1 = 1 − λ1 + λ2 = 2(1 − λ1). c = kg(λ1φ1 + λ2φ2) − g(φ1)k1 Since, by hypothesis, we have = 1. c c c c c c c (8) Therefore, The equality It is easy to check that the eigenvalues of the matrix(cid:18) t − 1 1 − t (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 {(1−t)2 +c2, (t−1)2 +c2}}, and hence(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) t − 1 1 − λ1 =p(1 − t)2 + c2. 2pt2 + c2 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) t 1 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 1 − t (cid:19) −(cid:18) 0 λ1 = 1 − λ2 =pt2 + c2. implies (9) 0 0 c c = kλ1φ1 + λ2φ2 − φ2k1 = λ1 + 1 − λ2 = 2(1 − λ2) = 2λ1, are exactly 1 − t (cid:19)2 = 2p(1 − t)2 + c2. = kg(λ1φ1 + λ2φ2) − g(φ2)k1 Combining (8) and (9) we get λ1 = t and c = 0, which shows that g(λ1φ1 + λ2φ2) =(cid:18) λ1 0 0 λ2 (cid:19) = λ1φ1 + λ2φ2 = λ1g(φ1) + λ2g(φ2), equivalently, since g = T2S(C1(H)) ◦ f ◦ T −1 (10) for every λ1, λ2 ≥ 0, with λ1 + λ2 = 1 and every set {e1, e2} of mutually orthogonal rank one elements in S(C1(H)). f (λ1e1 + λ2e2) = λ1f (e1) + λ2f (e2), S(C1(H)) we obtain 1 Finally, suppose λ1, λ2 ∈ R, with λ1 + λ2 = 1. Set σi ∈ {±1} such that λi = σiλi. Given and arbitrary set {e1, e2} of mutually orthogonal rank one elements in S(C1(H)), the set {σ1e1, σ2e2} satisfies the same properties. It follows from (10) that f (λ1e1 + λ2e2) = f (λ1σ1e1 + λ2σ2e2) = λ1f (σ1e1) + λ2f (σ2e2) TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 9 = (by Proposition 2.6(d) and (e)) = λ1σ1f (e1)+λ2σ2f (e2) = λ1f (e1)+λ2f (e2). (cid:3) Corollary 3.2. Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is a two dimensional complex Hilbert space. Suppose e1, e2 is a (maximal) set of mutually orthogonal rank one elements in S(C1(H)). The following statements hold: (a) If f (ie1) = if (e1) then f (ie2) = if (e2); (b) f (ie1) = −if (e1) then f (ie2) = −if (e2); Proof. (a) Suppose f (ie1) = if (e1). Let us find two orthonormal basis {η1, η2}, {ξ1, ξ2} such that e1 = η1 ⊗ ξ1 and e2 = η2 ⊗ ξ2. We set u1 = η1 ⊗ ξ2 and u2 = η2 ⊗ ξ1. The elements x = 1 2 (e1 + e2 − u1 − u2) are rank one elements in S(C1(H)). 2 (e1 + e2 + u1 + u2) and y = 1 By Proposition 2.6(d), f (ix) = ±if (x) and f (iy) = ±if (y). If f (ix) = −if (x) we have 1 1 2 1 2 1 2 1 2 1 2 1 2 and 1 = 1 = = √2 kf (ix) − f (ie1)k1 = k−if (x) − if (e1)k1 = kf (x) + f (e1)k1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) −1 1 1 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 kf (ix) − f (ie1)k1 = kix − ie1k1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) − 1 2 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) 3 1 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 = kx + e1k1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) 3 2(cid:18)q6 + 4√2 +q6 − 4√2(cid:19) , y(cid:19) = f(cid:18) i y(cid:19) = if(cid:18) 1 2 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 e2(cid:19) = f(cid:18) i = if(cid:18) 1 which gives a contradiction. Therefore, f (ix) = if (x). We similarly prove f (iy) = if (y). Since ix ⊥ iy in S(C1(H)), we can apply Proposition 3.1 to deduce that f (iy) = i f (x) + i f (ix) + 1 = x + 2 x + 2 e1 + 2 f (y) e1 + 2 i 2 1 1 i 2 1 2 Since a new application of Proposition 3.1 gives f(cid:0) 1 and f(cid:0) i 2 e1 + i f (ie2) = if (e2). 2 e2(cid:1) = 1 2 f (ie1) + 1 2 f (e2) 2 f (ie2), the equality f (ie1) = if (e1) proves that 2 f (e1) + 1 2 e1 + 1 1 2 e2(cid:19) . 2 e2(cid:1) = 1 1 2 1 2 1 2 1 2 We can actually prove that f (iu1) = if (u1) and f (iu2) = if (u2). Statement (b) follows by similar arguments. (cid:3) Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is a two dimensional complex Hilbert space. Suppose e1, e2 is a (maximal) set of mutually orthogonal rank one elements in S(C1(H)) and λ1, λ2 are real numbers satisfying λ1 + λ2 = 1. Then we have proved in Proposition 3.1 that f (λ1e1 + λ2e2) = λ1f (e1) + λ2f (e2). Arguing as in the proof of Proposition 3.1 we can find two surjective complex linear isometries T1, T2 : S(C1(H)) → S(C1(H)) such that g = T2S(C1(H)) ◦ f ◦ 10 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA T −1 1 0 0 0 0 0 1 1 γ γ or (12) (11) By hypothesis, φj, where φ1 =(cid:18) 1 0 S(C1(H)) : S(C1(H)) → S(C1(H)) is a surjective isometry satisfying g(φj) = 0 0 (cid:19) and φ2 =(cid:18) 0 0 0 (cid:19)(cid:19) =(cid:18) 0 µ 0 (cid:19)(cid:19) =(cid:18) 0 0 1 (cid:19) . Now, we claim that µ 0 (cid:19) 1 0 (cid:19)(cid:19) =(cid:18) 0 0 (cid:19) 1 0 (cid:19)(cid:19) =(cid:18) 0 µ 0 (cid:19) and g(cid:18)(cid:18) 0 0 µ 0 (cid:19) and g(cid:18)(cid:18) 0 0 g(cid:18)(cid:18) 0 g(cid:18)(cid:18) 0 for suitable µ ∈ T. Indeed, we know from previous arguments that g(cid:18)(cid:18) 0 1 0 0 (cid:19)(cid:19) and g(cid:18)(cid:18) 0 0 1 0 (cid:19)(cid:19) are rank one orthogonal elements in S(C1(H)). Let us write g(cid:18)(cid:18) 0 1 0 0 (cid:19)(cid:19) = η⊗ ξ =(cid:18) α β δ (cid:19), with αδ = γβ and α2 +β2 +γ2 +δ2 = 1. (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) α β 0 0 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 =(cid:13)(cid:13)(cid:13)(cid:13)g(cid:18)(cid:18) 0 1 δ (cid:19) ±(cid:18) 1 0 0 0 (cid:19)(cid:19) ± g(cid:18)(cid:18) 1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) 0 1 0 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 0 0 (cid:19) ±(cid:18) 1 Since(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) α ± 1 β δ (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) 2 ≥ α ± 12 + β2 + γ2 + δ2 = 1 + α2 ± 2ℜe(α) + β2 + γ2 + δ2, which assures that 0 ≥ ±2ℜe(α) and hence ℜe(α) = 0. Replacing g(cid:18)(cid:18) 1 0 with g(cid:18)(cid:18) i 0 (cid:19)(cid:19) ∈(cid:26)±ig(cid:18)(cid:18) 1 Similar arguments applied to g(cid:18)(cid:18) 0 δ = 0. Since γβ = 0 and β2 + γ2 = 1 we deduce that 0 (cid:19) and g(cid:18)(cid:18) 0 0 (13) µ 0 (cid:19) and g(cid:18)(cid:18) 0 1 (cid:19)(cid:19) instead of g(cid:18)(cid:18) 1 1 0 (cid:19)(cid:19) =(cid:18) 0 0 (cid:19)(cid:19) =(cid:18) 0 0 0 (cid:19)(cid:19) 0 (cid:19)(cid:19)(cid:27), 0 (cid:19)(cid:19) prove 0 (cid:19) 0 0 (cid:19) , 0 0 (cid:19)(cid:19) and having in mind that g(cid:18)i(cid:18) 1 g(cid:18)(cid:18) 0 1 g(cid:18)(cid:18) 0 0 0 (cid:19)(cid:19) =(cid:18) 0 µ 0 (cid:19)(cid:19) =(cid:18) 0 we obtain ℑm(α) = 0, and thus α = 0. = 2, we deduce from the inequality in (4) that 0 (cid:19)(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 = √2. (14) or γ 2 1 0 δ 0 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 δ with δ, µ ∈ T. 2 2 − 1 − 1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) 1 =(cid:13)(cid:13)(cid:13)(cid:13)f(cid:18) 1 We shall show next that δ = µ. Indeed, let us assume that g satisfies (13). Then, applying Proposition 3.1 we have 1 2 0 1 2(cid:18)(cid:18) 1 2 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 =(cid:13)(cid:13)(cid:13)(cid:13) 2(cid:18)(cid:18) 1 0 0 0 (cid:19) +(cid:18) 0 0 0 (cid:19) +(cid:18) 0 0 1 (cid:19)(cid:19)(cid:19) − f(cid:18) 1 1 (cid:19)(cid:19) − 2(cid:18)(cid:18) 0 0 0 0 0 0 1 1 1 0 2(cid:18)(cid:18) 0 0 (cid:19) +(cid:18) 0 0 (cid:19)(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 0 (cid:19) +(cid:18) 0 0 (cid:19)(cid:19)(cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 1 0 1 TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 11 Since we have 2 − µ − δ 1 2 2 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) 1 1 = 2 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 2(cid:18)q2 − δ + µ +q2 + δ + µ(cid:19) . 2 (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) 1 q2 − δ + µ +q2 + δ + µ = 2, 2 − 1 − 1 = 1, 2 2 1 which implies 2 = δ + µ, and hence δ = µ. Similarly, when we are in case (14) we get δ = µ and hence (12) holds. Let us assume we are in the case derived from (11), that is, 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 µ 0 (cid:19) . g(cid:18)(cid:18) 0 1 1 (cid:19) , T(cid:18)(cid:18) 0 µ 0 (cid:19)(cid:19) =(cid:18) 0 1 0 0 (cid:19) and T(cid:18)(cid:18) 0 0 0 (cid:19)(cid:19) =(cid:18) 0 µ 0 (cid:19) and g(cid:18)(cid:18) 0 0 (cid:19)(cid:19) =(cid:18) 0 Consider the unitary v =(cid:18) √µ 0 √µ (cid:19) and the surjective linear isometry T : S1(H) → S1(H), T (x) = v∗xv. It is easy to see that T(cid:18)(cid:18) 1 0 0 0 (cid:19)(cid:19) =(cid:18) 1 0 0 0 (cid:19) , T(cid:18)(cid:18) 0 1 (cid:19)(cid:19) =(cid:18) 0 µ 0 (cid:19)(cid:19) = (cid:18) 0 0 (cid:19) . Therefore, replacing g with h = TS(S1(H))g, we obtain a surjective isometry h : S(S1(H)) → S(S1(H)) satisfying h(cid:18)(cid:18) 1 0 0 0 (cid:19)(cid:19) = (cid:18) 1 0 0 0 (cid:19) , h(cid:18)(cid:18) 0 1 (cid:19)(cid:19) =(cid:18) 0 0 (cid:19) and h(cid:18)(cid:18) 0 0 (cid:19)(cid:19) = (cid:18) 0 0 (cid:19) . such that h = TS(S1(H))g is a surjective isometry satisfying h(cid:18)(cid:18) 1 0 (cid:18) 1 0 (cid:19)(cid:19) = (cid:18) 0 h(cid:18)(cid:18) 0 1 When (12) holds we can find surjective linear isometry T : S1(H) → S1(H) 0 (cid:19) , h(cid:18)(cid:18) 0 0 (cid:19)(cid:19) =(cid:18) 0 1 (cid:19)(cid:19) = (cid:18) 0 0 (cid:19) . 1 (cid:19) , h(cid:18)(cid:18) 0 1 0 0 (cid:19)(cid:19) =(cid:18) 0 0 0 (cid:19)(cid:19) = 0 (cid:19) and 1 (cid:19) , h(cid:18)(cid:18) 0 We shall establish now a technical proposition to measure the distance between two pure atoms in C1(H). For each pure atom e = η⊗ ξ in S(C1(H)), as before, let s(e) = η⊗ ξ be the unique minimal partial isometry in B(H) satisfying e(s(e)) = 1. For any other x ∈ S(C1(H)) the evaluation x(s(e)) ∈ C. For each partial isometry s in B(H), the Bergmann operator P0(s) : B(H) → B(H), x 7→ (1− ss∗)x(1− s∗s), is weak∗ continuous and hence P0(s)∗(x) ∈ C1(H) for every x ∈ C1(H). Lemma 3.3. Let H be a complex Hilbert space. Suppose e1, e2 are two rank one pure atoms in S(C1(H)). Then the following formula holds: 0 0 1 1 0 0 1 0 0 1 0 0 0 0 1 0 1 0 1 ke2 − e1k1 = Xk=1,2r(1 − ℜe(α)) + (−1)kq(1 − ℜe(α))2 − δ2, where α = α(e1, e2) = e2(s(e1)) and δ = δ(e1, e2) = kP0(s(e1))∗(e2)k1. 12 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA e2 = αv11 + βv12 + δv22 + γv21, Proof. By choosing an appropriate matrix representation we can find two orthonor- mal systems {η1, η2} and {ξ1, ξ2} to represent e1 and e2 in the form e1 = η1 ⊗ ξ1, e2 =eη1 ⊗eξ1 and where e1 = v11, v12 = η2 ⊗ ξ1, v21 = η1 ⊗ ξ2, v22 = η2 ⊗ ξ2, α = hξ1/eξ1iheη1/η1i, β = hξ1/eξ1iheη1/η2i, γ = hξ2/eξ1iheη1/η1i, δ = hξ2/eξ1iheη1/η2i ∈ C, with α2 + β2 + γ2 + δ2 = hξ1/eξ1i2keη1k2 + hξ2/eξ1i2keη1k2 = keξ1k2 = 1, and αδ = βγ. In an appropriate matrix representation we can identify e1 and e2 with (cid:18) 1 0 (cid:19) , and (cid:18) α β eigenvalues of the element (e2 − e1)(e2 − e1)∗ =(cid:18) α − 1 β precisely (1 − ℜe(α)) ±p(1 − ℜe(α))2 − δ2 and hence δ (cid:19)(cid:18) α − 1 β δ (cid:19)∗ ke2 − e1k1 = Xk=1,2q(1 − ℜe(α)) + (−1)kp(1 − ℜe(α))2 − δ2, γ Following the arguments in the proof of [19, Proposition 3.3] we deduce that the δ (cid:19) , respectively. are γ γ 0 0 which proves the desired formula. (cid:3) We are now in position to solve Tingley's problem for the case of trace class operators on a two dimensional Hilbert space. Theorem 3.4. Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is a two dimensional complex Hilbert space. Then there exists a surjective complex linear or conjugate linear isometry T : C1(H) → C1(H) satisfying f (x) = T (x) for every x ∈ S(C1(H)). More concretely, there exist unitaries u, v ∈ M2(C) such that one of the following statements holds: (a) f (x) = uxv, for every x ∈ S(C1(H)); (b) f (x) = uxtv, for every x ∈ S(C1(H)); (c) f (x) = uxv, for every x ∈ S(C1(H)); (d) f (x) = ux∗v, for every x ∈ S(C1(H)), where (xij ) = (xij ). Proof. By the comments preceding this theorem, we can find two surjective linear isometries U, V : C1(H) → C1(H) such that the mapping h = US(C1(H)) ◦ f ◦ V S(C1(H)) : S(C1(H)) → S(C1(H)) is a surjective isometry satisfying precisely one of the next statements (15) or (16) 0 0 1 h(cid:18)(cid:18) 1 0 0 0 (cid:19)(cid:19) =(cid:18) 1 0 h(cid:18)(cid:18) 0 0 1 0 (cid:19)(cid:19) =(cid:18) 0 0 h(cid:18)(cid:18) 1 0 0 0 (cid:19)(cid:19) =(cid:18) 1 0 h(cid:18)(cid:18) 0 0 1 0 (cid:19)(cid:19) =(cid:18) 0 1 0 0 (cid:19) , h(cid:18)(cid:18) 0 1 0 (cid:19) , and h(cid:18)(cid:18) 0 0 0 0 (cid:19) , h(cid:18)(cid:18) 0 0 0 (cid:19) , and h(cid:18)(cid:18) 0 0 0 (cid:19)(cid:19) =(cid:18) 0 0 1 (cid:19)(cid:19) =(cid:18) 0 0 (cid:19)(cid:19) =(cid:18) 0 0 1 (cid:19)(cid:19) =(cid:18) 0 0 0 1 1 0 0 1 0 (cid:19) , 1 (cid:19) ; 0 (cid:19) , 1 (cid:19) . 0 0 TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 13 Proposition 2.6(d) and (e) implies that f (−z) = −f (z) and h(−z) = −h(z) for every pure atom z in C1(H). We assume that (15) holds. Let us denote v11 =(cid:18) 1 0 v21 =(cid:18) 0 0 1 0 (cid:19), and v22 =(cid:18) 0 1 (cid:19). 0 0 0 0 (cid:19), v12 =(cid:18) 0 1 0 0 (cid:19), By Proposition 2.6(d) and Corollary 3.2, we know that one of the next statements holds: (a) h(iv11) = ih(v11) and h(iv22) = ih(v22); (b) h(iv11) = −ih(v11) and h(iv22) = −ih(v22). The proof will be splitted into two cases corresponding to the above statements. γ′ 1 2 1 2 1 2 1 2 By the hypothesis on h and Lemma 3.3 we get Case (a). Let us assume that (a) holds. We consider the pure atom e1 = 2 (cid:19). Proposition 2.6 assures that h(e1) is a pure atom, and hence it must δ′ (cid:19) with α′2 + β′2 + γ′2 + δ′2 = 1, α′δ′ = β′γ′. ) + (−1)ks(1 − (cid:18) 1 be of the form h(v) =(cid:18) α′ β′ vuut(1 − √2 = Xk=1,2 = kh(e1) − v11k1 = Xk=1,2q(1 − ℜe(α′)) + (−1)kp(1 − ℜe(α′))2 − δ′2, + √2 +r 3 r 3 = kh(e1) − h(−v11)k1 = Xk=1,2q(1 + ℜe(α′)) + (−1)kp(1 + ℜe(α′))2 − δ′2. ) + (−1)ks(1 + √2 = Xk=1,2 2(cid:19)2 )2 −(cid:18) 1 2(cid:19)2 )2 −(cid:18) 1 vuut(1 + = ke1 − v11k1 = ke1 + v11k1 2 − and 1 2 1 2 1 2 2 Taking squares in both sides we get and 2 = 2(1 − ℜe(α′)) + 2δ′, 4 = 2(1 + ℜe(α′)) + 2δ′, 2 and δ′ = 1 2 . 2 and ℜe(δ′) = 1 which gives ℜe(α′) = 1 2 , ℜe(γ′) = 1 ℜe(β′) = 1 deduce that α′ = β′ = γ′ = δ′ = 1 We can similarly show that taking e2 =(cid:18) 1 2 − 1 − 1 When in the above arguments we replace v11 with v12, v21 and v22 we obtain 2 . Since α′2 + β′2 + γ′2 + δ′2 = 1, we 2 (cid:19), we have h(e2) = e2. 2 , and hence h(e1) = e1 =(cid:18) 1 Now since e1 and e2 are orthogonal pure atoms, Proposition 2.6(d) and (e) and 2 (cid:19) . 2 1 2 1 2 1 2 2 1 Corollary 3.2 imply that exactly one of the next statements holds: (a.1) h(ie1) = ih(e1) and h(ie2) = ih(e2); (a.2) h(ie1) = −ih(e1) and h(ie2) = −ih(e2). 14 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA Let us show that the conclusion in (a.2) is impossible. Indeed, if (a.2) holds, Proposition 3.1 implies 1 2 iv11 + 1 2 iv22 = 1 2 h (iv11) + 1 2 h (iv22) = h(cid:18) 1 2 iv11 + 1 2 iv22(cid:19) = h(cid:18) 1 2 ie1 + 1 2 ie2(cid:19) = which is impossible. 1 2 h(ie1) + 1 2 h(ie2) = − 1 2 ie1 − 1 2 ie2 = − 1 2 iv11 − 1 2 iv22, Since (a.1) holds, we deduce, via Proposition 3.1, that 1 2 h (iv12) + 1 2 h (iv21) = h(cid:18) 1 2 1 2 iv21(cid:19) = h(cid:18) 1 2 1 2 ie2(cid:19) iv12 + 1 2 ie1 − 1 2 ie1 − 1 2 1 2 h (ie2) = ie2 = iv12 + iv21. = 1 2 h (ie1) − 1 2 We know from Corollary 3.2 that h (ivjk) ∈ {±ih(vjk)} = {±ivjk}, for every k, j = 1, 2. Thus, (17) h (iv12) = ih (v12) = iv12, and h (iv21) = ih (v21) = iv21. We shall prove that (18) h(v) = v, for every pure atom v ∈ S(C1(H)). γ γ′ Let v be a pure atom (i.e. a rank one partial isometry) in S(C1(H)). By Proposition 2.6(b), h(v) is a pure atom in S(C1(H)). Arguing as in the proof of Applying the hypothesis on h and Lemma 3.3 we get the following equations Lemma 3.3, we may assume that v =(cid:18) α β δ′ (cid:19) , with α2 +β2 +γ2 +δ2 = 1, α′2 +β′2 +γ′2 +δ′2 = 1, α′δ′ = β′γ′, and αδ = βγ. δ (cid:19) , and h(v) =(cid:18) α′ β′ Xk=1,2q(1 − ℜe(α)) + (−1)kp(1 − ℜe(α))2 − δ2 = kv − v11k1 = kh(v) − v11k1 = Xk=1,2q(1 − ℜe(α′)) + (−1)kp(1 − ℜe(α′))2 − δ′2, Xk=1,2q(1 + ℜe(α)) + (−1)kp(1 + ℜe(α))2 − δ2 = kv + v11k1 = kh(v) − h(−v11)k1 = Xk=1,2q(1 + ℜe(α′)) + (−1)kp(1 + ℜe(α′))2 − δ′2. and Taking squares in both sides we get 2(1 − ℜe(α)) + 2δ = 2(1 − ℜe(α′)) + 2δ′, and 2(1 + ℜe(α)) + 2δ = 2(1 + ℜe(α′)) + 2δ′, which gives ℜe(α′) = ℜe(α) and δ = δ′. ℑm(α), and hence α = α′. Then applying the above arguments to v, iv11 and −iv11 we obtain ℑm(α′) = TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 15 Having in mind that in case (a) we have h(ivjk) = ivjk for every j, k (see (17)), a similar reasoning applied to v, v22 and iv22 (respectively, v, v12 and iv12 or v, v21 and iv21) gives δ = δ′ (respectively, β = β′ or γ = γ′). We have therefore shown that h(v) = v, for every pure atom v. Proposition 3.1 assures that h(x) = x for every x ∈ S(C1(H)), and hence f = U −1V −1S(C1(H)), where U −1V −1 = C1(H) → C1(H) is a surjective complex linear isometry. In Case (b), we can mimic the above arguments to show that h(v) = v, for every pure atom v ∈ S(C1(H)), (19) x21 x21 (cid:19) =(cid:18) x11 x12 where(cid:18) x11 x12 x21 x21 (cid:19), and consequently, f (x) = U −1(V −1(x)) = T (x), for every x ∈ S(C1(H)), where T : C1(H) → C1(H), T (x) = U −1(V −1(x)) (x ∈ C1(H)) is a surjective conjugate linear isometry. Finally, if we assume (16), then there exist two surjective linear isometries U, V : C1(H) → C1(H) such that or for every x ∈ S(C1(H)). f (x) = U −1(V −1(xt)) f (x) = U −1(V −1(x∗)) (cid:3) Before dealing with surjective isometries between the unit spheres of trace class spaces over a finite dimensional complex Hilbert space, we shall present a technical result. Proposition 3.5. Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is a complex Hilbert space with dim(H) = n. Suppose f satisfies the follow- ing property: given a set {e1, e2, . . . , ek} of mutually orthogonal pure atoms in λj = 1 we S(C1(H)) with k < n, and real numbers λ1, λ2, . . . , λk satisfying kXj=1 λjej. Then f (x) = x, for every x ∈ S(C1(H)). Proof. Let {ξ1, . . . , ξn} be an orthonormal basis of H. We set vj = ξj ⊗ ξj (j ∈ {1, . . . , n}). We claim that the identity have f kXj=1 λj ej = kXj=1 (20) µjvj, nXj=1 f nXj=1 µjvj = nXj=1 holds for every µ1, . . . , µn in R+ 0 with µj = 1. We observe that we can assume that µj > 0 for every j, otherwise the statement is clear from the hypothesis on f . By the hypothesis on f we know that f (vj ) = vj for every j ∈ {1, . . . , n}. Let 1 denote the identity in B(H). Clearly {v1, v2, . . . , vn} ⊆ {1}′. By Proposition 2.6 (see also (7)) there exists a unique unitary w ∈ B(H) such that f ({1}′) = {w}′. Since, for each j = 1, . . . , n, vj = f (vj ) ∈ {w}′, we can easily deduce that w = 1 16 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA and hence f ({1}′) = {1}′ is the face of all states (i.e., positive norm-one functional on B(H)). The element f matrix a = (aij) ∈ Mn(C) with kak1 = 1 = tr(a) such that f µjvj ∈ f ({1}′) = {1}′, and hence there exists a positive µjvj = a. It should be remarked that we can also identify each vj with the matrix in Mn(C) with entry 1 in the (j, j) position and zero otherwise. nXj=1 nXj=1 Let us fix a projection p ∈ B(H). The mapping Mp : C1(H) → C1(H), Mp(x) = pxp+ (1− p)x(1− p) is linear, contractive and positive. Let pj denote the projection ξj ⊗ ξj = s(vj ) ∈ B(H). By hypothesis µjvj − f (vj)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1 ka − vjk1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) f nXj=1 =Xk6=j µk + 1 − µj = 2Xk6=j =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 µjvj − vj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1 µk = 2(1 − µj). Having in mind that (1 − pj)a(1 − pj), pjapj, and a are positive functionals with a(1) = kak1 = 1, we get 2(1 − µj) = ka − vjk1 ≥ kMpj (a − vj)k1 = kpj(a − vj)pj + (1 − pj)a(1 − pj)k1 = (by orthogonality) = kpj(a − vj)pjk1 + k(1 − pj)a(1 − pj)k1 = 1 − ajj + ((1 − pj)a(1 − pj))(1 − pj) = 1 − ajj + a(1 − pj) = 2(1 − ajj ). This shows that ajj ≥ µj for every j = 1, . . . , n. Since 1 = a(1) = a11 + . . . + ann ≥ µ1 + . . . + µn = 1, we deduce that ajj = µj for every j = 1, . . . , n. We shall now show that aij = 0 for every i 6= j. For this purpose, fix i 6= j and set q = pi + pj ∈ B(H) and x = µivi + (1 − µi)vj ∈ S(C1(H)). By hypothesis f (x) = x, and Let us observe that whose eigenvalues are precisely 0 0 aij µk + 1 − µi − µj = 2(1 − µi − µj). = Xk6=i,j 1 − µi (cid:19)(cid:19)2 ka − xk1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) µjvj − x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)1 nXj=1 aij µj (cid:19) −(cid:18) µi aij −1 + µi + µj (cid:19)2 (cid:18)(cid:18) µi s (1 − µi − µj)2 + 2aij2 ±p((1 − µi − µj)2 + 2aij2)2 − 4aij4 (cid:13)(cid:13)(cid:13)(cid:13)(cid:18) µi aij µj (cid:19) −(cid:18) µi 1 − µi (cid:19)(cid:13)(cid:13)(cid:13)(cid:13) = (1 − µi − µj)2 + 4aij2. =(cid:18) 0 = (1 − µi − µj)2 + 2aij2 + 2aij2 aij aij 2 2 1 0 0 , and thus TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 17 Therefore, we have 2(1 − µi − µj) ≥ kMq(a − x)k1 = kq(a − x)q + (1 − q)a(1 − q)k1 = (by orthogonality) = kq(a − x)qk1 + k(1 − q)a(1 − q)k1 0 0 aij + a(1 − q) 1 − µi (cid:19)(cid:13)(cid:13)(cid:13)(cid:13)1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:18) µi aij µj (cid:19) −(cid:18) µi =q(1 − µi − µj)2 + 4aij2 + a(1) − a(pi) − a(pj) =q(1 − µi − µj)2 + 4aij2 + 1 − µi − µj, f µjvj = a = nXj=1 nXj=1 µjvj, which implies that 1 − µi − µj ≥p(1 − µi − µj)2 + 4aij2, and hence aij = 0 as desired. We have thus proved that which concludes the proof of (20). Finally, let us take x ∈ S(C1(H)). We can find a set {e1, e2, . . . , en} of mutually orthogonal pure atoms in S(C1(H)) and real numbers λ1, λ2, . . . , λn such that x = nXj=1 λj ej. We observe that, replacing each ej with ±ej we can always assume that λj ≥ 0 for every j. Let us pick two unitary matrices u1, w1 ∈ B(H) satisfying u1vj w1 = ej for every j = 1, . . . , n. The operator Tu1,w1 : C1(H) → C1(H), T (y) = u1yw1 is a surjective isometry mapping elements of rank k to elements of the same rank. Consequently, the mapping f2 : S(C1(H)) → S(C1(H)), f2(y) = u∗ 1f (u1yw1)w∗ 1 is a surjective isometry. For each y ∈ S(C1(H)) with rank(y) < n, we have u1yw1 ∈ S(C1(H)) with rank(u1yw1) < n and thus, by the hypothesis on f , f (u1yw1) = u1yw1, which implies f2(y) = y. Therefore f2 satisfies the same hypothesis of f . Applying (20) we get 1 = u∗ 1 = u∗ u∗ 1f (x)w∗ λj ej w∗ 1f nXj=1 λj vj = = f2 nXj=1 λjej = which proves that f (x) = f nXj=1 nXj=1 1fu1 nXj=1 nXj=1 λjvj , λjvj w1 w∗ 1 λjej = x, as desired. (cid:3) Remark 3.6. Let T : C1(H) → C1(H) be a surjective real linear isometry, where H is a complex Hilbert space. Since T ∗ : B(H) → B(H) is a surjective real linear isometry, T ∗ and T must be complex linear or conjugate linear (see [4, Proposition 2.6]). We therefore deduce from [28] (see also [22, §11.2]) that there exist unitary matrices u, v ∈ B(H) such that one of the next statements holds: (a.1) T (x) = uxv, for every x ∈ S(C1(H)); (a.2) T (x) = uxtv, for every x ∈ S(C1(H)); (a.3) T (x) = uxv, for every x ∈ S(C1(H)); 18 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA (a.4) T (x) = ux∗v, for every x ∈ S(C1(H)), where (xij ) = (xij ). We can find a more concrete description under additional hypothesis. Suppose dim(H) = n. The symbol eij ∈ C1(H) will denote the elementary matrix with entry 1 at position (i, j) and zero otherwise. Case A Suppose that T (ζekk) = ζekk for every k = 1, . . . , n − 1, and all ζ ∈ T, T (en1) = αen1, and T (e1n) = µe1n with α, µ ∈ T. Then T has the form described in case (a.1) above with u = 0 1 . . . 0 ... ... ... 0 . . . 1 0 0 . . . 0 α  , and v = 1 ... 0 0 0 . . . 0 ... ... . . . 1 0 . . . 0 µ .  If we also assume T (enn) = enn, then α = µ. For the proof we simply observe that since T (ie11) = ie11, we discard cases (a.3) and (a.4). The assumption T (e1n) = µe1n shows that case (a.2) is impossible too. Since T has the form described in (a.1) for suitable u, v. Now, T (ekk) = ekk implies that ukkvkk = 1, for all k = 1, . . . , n− 1. Finally, it is straightforward to check that T (e1n) = µe1n and T (en1) = αen1 give the desired statement. We can present now the main result of this section. Theorem 3.7. Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is a finite dimensional complex Hilbert space. Then there exists a surjective complex linear or conjugate linear isometry T : C1(H) → C1(H) satisfying f (x) = T (x) for every x ∈ S(C1(H)). More concretely, there exist unitary elements u, v ∈ Mn(C) = B(H) such that one of the following statements holds: (a) f (x) = uxv, for every x ∈ S(C1(H)); (b) f (x) = uxtv, for every x ∈ S(C1(H)); (c) f (x) = uxv, for every x ∈ S(C1(H)); (d) f (x) = ux∗v, for every x ∈ S(C1(H)), where (xij ) = (xij ). Proof. We shall argue by induction on n =dim(H). The case n = 2 has been proved in Theorem 3.4. Let us assume that the desired conclusion is true for every surjective isometry f : S(C1(K)) → S(C1(K)), where K is a finite dimensional complex Hilbert space of dimension ≤ n. Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is an (n + 1)-dimensional complex Hilbert space. Let {ξ1, . . . , ξn+1} be an orthonormal basis of H, and let eij = ξj ⊗ ξi. Clearly eij is a pure atom for every i, j, and {e11, . . . , e(n+1)(n+1)} is a maximal set of mutually orthogonal pure states in C1(H). By Proposition 2.6 and Lemma 2.2 {f (e11), . . . , f (e(n+1)(n+1))} is a maximal set of mutually orthogonal pure atoms in C1(H) too. We can find unitary matrices u1, w1 ∈ Mn+1(C) such that u1f (eii)w1 = eii for every i = 1, . . . , n + 1. We set f1 = u1f w1. We observe that f admits an extension to a surjective real linear isometry if and only if f1 does. Let us note that {e(n+1)(n+1)}⊥ := {x ∈ C1(H) : x ⊥ e(n+1)(n+1)} ∼= C1(K) for a suitable n-dimensional complex Hilbert subspace of H. We regard C1(K) as a TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 19 complemented subspace of C1(H) under the appropriate identification. Lemma 2.2 implies that f1(S(C1(K))) = f1(cid:0){e(n+1)(n+1)}⊥ ∩ S(C1(H))(cid:1) = {f1(e(n+1)(n+1))}⊥ ∩ S(C1(H)) = {e(n+1)(n+1)}⊥ ∩ S(C1(H)) = S(C1(K)), and hence f1S(C1(K)) : S(C1(K)) → S(C1(K)) is a surjective isometry. By the induction hypothesis, there exist unitaries un, vn ∈ Mn(C) = B(K) such that one of the following statements holds: (1) f1(x) = unxvn, for every x ∈ S(C1(K)); (2) f1(x) = unxtvn, for every x ∈ S(C1(K)); (3) f1(x) = unxvn, for every x ∈ S(C1(K)); (4) f1(x) = unx∗vn, for every x ∈ S(C1(K)). Let un+1 =(cid:18) un 1 (cid:19) and vn+1 =(cid:18) vn 0 1 (cid:19). In each case from (1) to (4), we can define via the unitaries un+1, vn+1 in B(H) = Mn+1(C), the involution ∗, the transposition and the conjugation · , a surjective complex linear or conjugate linear isometry T1 : C1(H) → C1(H) such that T1f1(x) = x, for every x ∈ S(C1(K)) and T1f1(e(n+1)(n+1)) = e(n+1)(n+1). 0 0 0 We deal now with the mapping f2 = T1f1, which is a surjective isometry from S(C1(H)) onto itself and satisfies (21) f2(x) = x, for every x ∈ S(C1(K)) = {e(n+1)(n+1)}⊥ ∩ S(C1(H)), and f2(e(n+1)(n+1)) = e(n+1)(n+1). We claim that (22) f2(e1(n+1)) = µe1(n+1), and f2(e(n+1)1) = µe(n+1)1, for a suitable µ in T. e22, . . . , enn, e12, Lemma 2.2 implies that Indeed, since e1(n+1) ⊥ e22, . . . , enn, e21 and e(n+1)1 ⊥ f2(e1(n+1)) ⊥ f2(e22) = e22, . . . , f2(e1(n+1)) ⊥ enn, f2(e1(n+1)) ⊥ f2(e21) = e21, and f2(e(n+1)1) ⊥ f2(e22) = e22, . . . , f2(e1(n+1)) ⊥ enn, f2(e1(n+1)) ⊥ f2(e12) = e12, which implies that and f2(e1(n+1)) = µe1(n+1) + λe(n+1)(n+1) f2(e(n+1)1) = αe1(n+1) + βe(n+1)(n+1) with α2+β2 = 1 and λ2+µ2 = 1 (compare Proposition 2.6(b)). Since e(n+1)1 ⊥ e1(n+1), a new application of Lemma 2.2 proves that f2(e1(n+1)) ⊥ f2(e(n+1)1), and consequently λ = β = 0. We have thus proved that f2(e1(n+1)) = µe1(n+1), and f2(e(n+1)1) = αe(n+1)1, with µ, α in T. We shall show next that α = µ. To this end, we observe that the subspace {e22, . . . , enn}⊥ := {x ∈ C1(H) : x ⊥ ejj, ∀j ∈ {2, . . . , n}} is isometrically isomorphic to C1(K2) for a suitable 2-dimensional complex Hilbert subspace K2 of H, contains e11, e1(n+1), e(n+1)1, and e(n+1)(n+1), and by Lemma 20 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA 2.2, f2S(C1(K2)) : S(C1(K2)) → S(C1(K2)) is a surjective isometry. So, by the induction hypothesis, there exist unitaries u3, v3 ∈ B(K2) such that f2S(C1(K2)) satisfies one of the statements from (a) to (d) in our theorem. Having in mind that f2(ζe11) = ζe11 for every ζ ∈ T, f2(e(n+1)(n+1)) = e(n+1)(n+1), f2(e1(n+1)) = µe1(n+1) and f2(e(n+1)1) = αe(n+1)1, it can be easily seen that we can identify u3 and v3 with (cid:18) 1 0 u33 (cid:19) and (cid:18) 1 0 u33 (cid:19), where u33 ∈ T, respectively (see Remark 3.6 Case A). This shows that µ = u33 and α = u33, which finishes the proof of (22). 0 0 Now, the subspace {e(n+1)1}⊥ ⊂ C1(H) is isometrically isomorphic to C1(K3) for a suitable n-dimensional complex Hilbert space K3, and since f2(e(n+1)1) = µe(n+1)1 (see (22)), Lemma 2.2 implies that f2S(C1(K3)) : S(C1(K3)) ∼= Mn(C) → S(C1(K3)) ∼= Mn(C) is a surjective isometry. By the induction hypothesis there exists a surjective real linear isometry T3 : C1(K3) → C1(K3) such that f2S(C1(K3)) ≡ T3S(C1(K3)). Since T3(ζeij ) = f2(ζeij ) = ζeij for every i, j ∈ {2, . . . , n} and all ζ ∈ T (see (21)) and T3(e1(n+1)) = f2(e1(n+1)) = µe1(n+1), Remark 3.6 Case A shows that T3(x) = xv3, where v3 identifies with the matrix (23) f2(z) = T3(z) = µz, 0 1 . . . 0 ... ... ... 0 0 . . . 1 0 . . . 0 µ  . This implies that for every z ∈ S(C1(K3)) with z = zpn+1, where pn+1 is the rank one projection ξn+1 ⊗ ξn+1 ∈ B(H), that is, for every z ∈ S(C1(H)) with z ⊥ e(n+1)1 and z = zpn+1. Similar arguments prove that (24) f2(z) = µz, 1 ... 0 0  0 1 . . . 0 ... ... ... 0 . . . 1 0 0 . . . 0 µ for every z ∈ S(C1(H)) with z ⊥ e1(n+1) and z = pn+1z. 0 . . . 0 ... ... . . . 1 0 . . . 0 µ Let us consider the unitaries u4 =   ∈ B(H), and v4 =  ∈ B(H), and the surjective complex linear isometry defined by T4(x) = u4xv4. Let f3 : S(C1(H)) → S(C1(H)) be the surjective isome- try defined by f3(x) = T4(f2(x)) (x ∈ S(C1(H))). Since T4(y) = y for every y ∈ {e(n+1)(n+1)}⊥, T4(e(n+1)(n+1)) = e(n+1)(n+1), T4(z) = µz, for all z ∈ S(C1(H)) with z ⊥ e(n+1)1 and z = zpn+1, and T4(z) = µz, for all z ∈ S(C1(H)) with z ⊥ e1(n+1) and z = pn+1z, we deduce from (21), (23), and (24) that (25) f3(e(n+1)(n+1)) = e(n+1)(n+1), and f3(x) = x, TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 21 for all x in the intersection of S(C1(H)) with the set {e(n+1)(n+1)}⊥ ∪ {z ⊥ e(n+1)1&z = zpn+1} ∪ {z ⊥ e1(n+1)&z = pn+1z}. We shall show next that (26) f3(v) = v, for every pure atom v ∈ S(C1(H)). Let v = η ⊗ ζ be a pure atom in C1(H). We can always write ζ = λ1ζ1 + λ2ξn+1 and η = µ1η1 + µ2ξn+1, with λj , µj ∈ C, µ12 + µ22 = 1, λ12 + λ22 = 1, and η1, ζ1 are norm one elements in {ξn+1}⊥. As before, we can write v = αv11 + βv12 + δv22 + γv21, where v11 = η1 ⊗ ζ1, v12 = η1 ⊗ ξn+1, v21 = ξn+1 ⊗ ζ1, v22 = ξn+1 ⊗ ξn+1, α2 + β2 + γ2 + δ2 = 1, and αδ = βγ. Let us note that Tv11 ⊆ {e(n+1)(n+1)}⊥, Tv12 ⊆ {z ⊥ e(n+1)1&z = zpn+1} ∩ S(C1(H)), and Tv12 ⊆ {z ⊥ e1(n+1)&z = pn+1z} ∩ S(C1(H)), it follows from (25) that (27) f3(µvij ) = µf3(vij ) for all (i, j) ∈ {(1, 1), (1, 2), (2, 1)}, and µ = 1. Let us consider the space v ∈ {v11, v12, v21, v22}⊥⊥ = Span{v11, v12, v21, v22} ∼= M2(C) whose unit sphere is fixed by f3 (compare Lemma 2.2). It follows from the in- duction hypothesis (i.e. f3 satisfies one of the statements from (a) to (d) in the state- ment of the theorem for every element in the unit sphere of {v11, v12, v21, v22}⊥⊥) and (27) that f3(v) = v = αv11 + βv12 + δv22 + γv21, which finishes the proof of (26). We shall next prove that f3 satisfies the hypothesis of the above Proposition 3.5. Let {v1, v2, . . . , vk} be a set of mutually orthogonal pure atoms in S(C1(H)) with λj = 1. Since dim(H) = k < n + 1, and real numbers λ1, λ2, . . . , λk satisfying n + 1, we can find a non-empty finite set of pure atoms {vk+1, . . . , vn+1} such that {vk+1, . . . , vn+1}⊥ = {v1, . . . , vk}. By (26) f3(vj) = vj for every j, and then Lemma 2.2 assures that f3(cid:0){vk+1, . . . , vn+1}⊥ ∩ S(C1(H))(cid:1) = {vk+1, . . . , vn+1}⊥ ∩ S(C1(H)). Having in mind that {vk+1, . . . , vn+1}⊥ ∼= C1(H ′), where H ′ is a complex Hilbert space with dim(H ′) = k < n+1, and f3S(C1(H ′)) : S(C1(H ′)) → S(C1(H ′)), it follows from the induction hypothesis the existence of a surjective real linear isometry R : C1(H ′) → C1(H ′) such that f3(x) = R(x) for all x ∈ S(C1(H ′)). Applying (26) we get kXj=1 f3 nXj=1 µjvj = R nXj=1 µjvj = nXj=1 µjR(vj) = µjf3(vj) = nXj=1 µjvj. nXj=1 Finally, since f3 satisfies the hypothesis of the above Proposition 3.5, we deduce (cid:3) from this result that f3(x) = x, for every x ∈ S(C1(H)). 22 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA 4. Surjective isometries between the unit spheres of two arbitrary trace class spaces In this section we consider the trace class operators on a complex Hilbert space H of arbitrary dimension. The answers obtained in the finite dimensional case can be now applied to simplify the study. Theorem 4.1. Let f : S(C1(H)) → S(C1(H)) be a surjective isometry, where H is an arbitrary complex Hilbert space. Then there exists a surjective complex linear or conjugate linear isometry T : C1(H) → C1(H) satisfying f (x) = T (x) for every x ∈ S(C1(H)). Proof. Let {ξk : k ∈ I} be an orthonormal basis of H. As before, we set ek := ξk ⊗ ξk. Then the set {ek : k ∈ I} is a maximal set of mutually orthogonal pure atoms in S(C1(H)). By Lemma 2.2 and Proposition 2.6(b) the elements in the set {f (ek) : k ∈ I} are mutually orthogonal pure atoms in C1(H). We can therefore find orthonormal systems {ηk : k ∈ I} and {ζk : k ∈ I} in H such that f (ek) = ηk ⊗ ζk for every k ∈ I. We claim that at least one of {ηk : k ∈ I} and {ζk : k ∈ I} must be an orthonormal basis of H. Otherwise, we can find norm one elements η0 and ζ0 in H such that η0 ⊥ ηk and ζ0 ⊥ ζk (in H) for every k. Then the element v0 = η0 ⊗ ζ0 is a pure atom in S(C1(H)) which is orthogonal to every f (ek). Applying Lemma 2.2 to f −1 we deduce that f −1(v0) ⊥ ek for every k ∈ I, which is impossible because {ξj : j ∈ I} is a basis of H. We can therefore assume that {ηk : k ∈ I} is an orthonormal basis of H. In a second step we shall show that {ζk : k ∈ I} also is an orthonormal basis of H. If that is not the case, there exists ζ0 in H such that ζ0 ⊥ ζk (in H) for every k. Fix an index k0 in I and set v0 := ηk0 ⊗ ζ0 ∈ ∂e(BC1(H)). Clearly, v0 ⊥ f (ek) for every k 6= k0. Lemma 2.2 and Proposition 2.6(b) imply that f −1(v0) is a pure atom in C1(H) which is orthogonal to ek for every k 6= k0. Since {ξj : j ∈ I} is a basis of H, we can easily see that f −1(v0) = λek0 for a unique λ ∈ T. We deduce from Proposition 2.6(d) and (e) that ηk0 ⊗ ζ0 = v0 = f f −1(v0) = µf (ek0) = ηk0 ⊗ ζk0 with µ ∈ T, which contradicts the fact ζ0 ⊥ ζk (in H) for every k. We have therefore shown that {ηk : k ∈ I} and {ζk : k ∈ I} both are orthonormal basis of H. Let us pick two unitary elements u1, w1 ∈ B(H) such that u1f (ek)w1 = ek for every k ∈ I. The mapping f2 : S(C1(H)) → S(C1(H)), f2(x) = u1f (x)w1 is a surjective isometry and f2(ek) = ek for every k ∈ I. Let T1 denote the surjective complex linear isometry on C1(H) given by T1(x) = u1xw1 (x ∈ C1(H)). Now, let F be a finite subset of I, and let qF denote the orthogonal projection of H onto HF = span{ξk : k ∈ F}. The set {ek : k /∈ F} is invariant under f2. Lemma 2.2 assures that f2(cid:0){ek : k /∈ F}⊥ ∩ S(C1(H))(cid:1) = {ek : k /∈ F}⊥∩S(C1(H)), where {ek : k /∈ F}⊥ ∩ S(C1(H)) = S(C1(HF )), and f2S(C1(HF )) : S(C1(HF )) → S(C1(HF )) is a surjective isometry. By Theorem 3.7 there exists a surjective real linear isometry TF : C1(HF ) → C1(HF ) such that f2(x) = TF (x) for all x ∈ S(C1(HF )). Let T2 : C1(H) → C1(H) denote the homogeneous extension of f2 defined by kxk1(cid:17) if x 6= 0 and T2(0) = 0. Clearly T2 is surjective. We shall T2(x) := kxk1f2(cid:16) x TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 23 show that T2 is an isometry. To this end, let us fix x, y ∈ C1(H)\{0} and ε > 0. Since every element in C1(H)(⊆ K(H)) can be approximated in norm by elements x ∈ S(C1(H)) with x = qF xqF , where F is a finite subset of I, we can find a finite set F ⊂ I, xε and yε in S(C1(H)) such that xε = qF xεqF , yε = qF yεqF , k x 2(kxk1+kyk1) . By the triangular kxk1 − xεk1 < inequality we have 2(kxk1+kyk1) and k y kyk1 − yεk1 < ε ε and since f2 is an isometry we get < ε 2 , (cid:12)(cid:12)(cid:12) kx − yk1 − kkxk1xε − kyk1yεk1(cid:12)(cid:12)(cid:12) ≤ kx − kxk1xεk1 + ky − kyk1yεk1 x y + kyk1(cid:13)(cid:13)(cid:13)(cid:13) kyk1 − yε(cid:13)(cid:13)(cid:13)(cid:13)1 kxk1 − xε(cid:13)(cid:13)(cid:13)(cid:13)1 ≤ kxk1(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12)(cid:12) kT2(x) − T2(y)k1 − kT2(kxk1xε) − T2(kyk1yε)k1(cid:12)(cid:12)(cid:12) kyk1(cid:19) − f2(yε)(cid:13)(cid:13)(cid:13)(cid:13)1 ≤ kxk1(cid:13)(cid:13)(cid:13)(cid:13)f2(cid:18) x + kyk1(cid:13)(cid:13)(cid:13)(cid:13)f2(cid:18) y kxk1(cid:19) − f2 (xε)(cid:13)(cid:13)(cid:13)(cid:13)1 kyk1 − yε(cid:13)(cid:13)(cid:13)(cid:13)1 kxk1 − xε(cid:13)(cid:13)(cid:13)(cid:13)1 + kyk1(cid:13)(cid:13)(cid:13)(cid:13) = kxk1(cid:13)(cid:13)(cid:13)(cid:13) ≤ kT2(x) − T2(kxk1xε)k1 + kT2(y) − T2(kyk1yε)k1 ε 2 < . x y On the other hand, since yε, xε ∈ S(C1(HF )), we can consider the surjective real linear isometry TF : C1(HF ) → C1(HF ) satisfying f2(x) = TF (x) for all x ∈ S(C1(HF )) to deduce that kT2(kxk1xε) − T2(kyk1yε)k1 = kkxk1f2(xε) − kyk1f2(yε)k1 = kkxk1TF (xε) − kyk1TF (yε)k1 = kTF (kxk1xε − kyk1yε)k1 = kkxk1xε − kyk1yεk1 . Combining this identity with the previous two inequalities we obtain kT2(x) − T2(y)k1 − kx − yk1 < ε. The arbitrariness of ε shows that kT2(x) − T2(y)k1 = kx − yk1, and hence T2 is an isometry. Finally, since T2 is a surjective isometry with T2(0) = 0, the Mazur-Ulam theorem 1 T2(x) (cid:3) guarantees that T2 is a surjective real linear isometry, and hence f (x) = T −1 for all x ∈ S(C1(H)), witnessing the desired conclusion. Acknowledgements First, second and third author were partially supported by the Spanish Ministry of Economy and Competitiveness (MINECO) and Eu- ropean Regional Development Fund project no. MTM2014-58984-P and Junta de Andaluc´ıa grant FQM375. Fourth author partially supported by grants MTM2014- 54240-P, funded by MINECO and QUITEMAD+-CM, Reference: S2013/ICE-2801, funded by Comunidad de Madrid. The authors are indebted to the anonymous reviewer for a thorough report, insightful comments, and suggestions. 24 F.J. FERN ´ANDEZ-POLO, J.J. GARC´ES, A.M. PERALTA, AND I. VILLANUEVA References [1] C.A. Akemann, G.K. Pedersen, Facial structure in operator algebra theory, Proc. Lond. Math. Soc. 64, 418-448 (1992). [2] J. Arazy, More on convergence in unitary matrix spaces, Proc. Amer. Math. Soc., 83, 44-48 (1981). [3] L. Cheng, Y. Dong, On a generalized Mazur-Ulam question: extension of isometries between unit spheres of Banach spaces, J. Math. Anal. Appl. 377, 464-470 (2011). [4] T. Dang, Real isometries between JB∗-triples, Proc. Amer. Math. Soc. 114, 971-980 (1992). [5] G.G. Ding, The 1-Lipschitz mapping between the unit spheres of two Hilbert spaces can be extended to a real linear isometry of the whole space, Sci. China Ser. A 45, no. 4, 479-483 (2002). [6] G.G. Ding, The isometric extension problem in the spheres of lp(Γ) (p > 1) type spaces, Sci. China Ser. A 46, 333-338 (2003). [7] G.G. Ding, On the extension of isometries between unit spheres of E and C(Ω), Acta. Math. Sin. (Engl. Ser.) 19, 793-800 (2003). [8] G.G. Ding, The representation theorem of onto isometric mappings between two unit spheres of l∞-type spaces and the application on isometric extension problem, Sci. China Ser. A 47, 722-729 (2004). [9] G.G. Ding, The representation theorem of onto isometric mappings between two unit spheres of l1(Γ) type spaces and the application to the isometric extension problem, Acta. Math. Sin. (Engl. Ser.) 20, 1089-1094 (2004). [10] G.G. Ding, The isometric extension of the into mapping from a L∞(Γ)-type space to some Banach space, Illinois J. Math. 51 (2) 445-453 (2007). [11] G.G. Ding, On isometric extension problem between two unit spheres, Sci. China Ser. A 52 2069-2083 (2009). [12] G.G. Ding, The isometric extension problem between unit spheres of two separable Banach spaces, Acta Math. Sin. (Engl. Ser.) 31, 1872-1878 (2015). [13] G.G. Ding, J.Z. Li, Sharp corner points and isometric extension problem in Banach spaces, J. Math. Anal. Appl. 405 297-309 (2013). [14] N. Dunford, J.T. Schwartz, Linear operators. Part II: Spectral theory. Self adjoint operators in Hilbert space, Interscience Publishers John Wiley & Sons, New York-London, 1963. [15] C.M. Edwards, F.J. Fern´andez-Polo, C.S. Hoskin, A.M. Peralta, On the facial structure of the unit ball in a JB∗-triple, J. Reine Angew. Math. 641, 123-144 (2010). [16] C.M. Edwards, G.T. Ruttimann, On the facial structure of the unit balls in a JBW∗-triple and its predual, J. Lond. Math. Soc. 38, 317-332 (1988). [17] X.N. Fang, J.H. Wang, Extension of isometries between the unit spheres of normed space E and C(Ω), Acta Math. Sinica (Engl. Ser.), 22, 1819-1824 (2006). [18] F.J. Fern´andez-Polo, A.M. Peralta, On the facial structure of the unit ball in the dual space of a JB∗-triple, Math. Ann. 348, 1019-1032 (2010). [19] F.J. Fern´andez-Polo, A.M. Peralta, Low rank compact operators and Tingley's problem, preprint 2016. arXiv:1611.10218v1 [20] F.J. Fern´andez-Polo, A.M. Peralta, On the extension of isometries between the unit spheres of a C∗-algebra and B(H), preprint 2017. arXiv:1701.02916v1 [21] F.J. Fern´andez-Polo, A.M. Peralta, Tingley's problem through the facial structure of an atomic JBW∗-triple, preprint 2017. arXiv:1701.05112v1 [22] R. Fleming, J. Jamison, Isometries on Banach Spaces: Vector-Valued Function Spaces, vol. 2 Chapman & Hall/CRC Monogr. Surv. Pure Appl. Math., vol. 138, Chapman and Hall/CRC, Boca Raton, London, New York, Washington, DC (2008). [23] I.C. Gohberg, M.G. Krein, Introduction to the theory of linear nonselfadjoint operators. Translations of Mathematical Monographs, Vol. 18 American Mathematical Society, Provi- dence, R.I., 1969. [24] V. Kadets, M. Mart´ın, Extension of isometries between unit spheres of infite-dimensional polyhedral Banach spaces, J. Math. Anal. Appl., 396, 441-447 (2012). [25] C.A. McCarthy, Cp, Israel J. Math. 5, 249-271 (1967). [26] R.E. Megginson, An Introduction to Banach Space Theory, Springer-Verlag, New York, 1998. [27] A.M. Peralta, R. Tanaka, A solution to Tingley's problem for isometries between the unit spheres of compact C∗-algebras and JB∗-triples, preprint 2016. arXiv:1608.06327v1. TINGLEY'S PROBLEM FOR SPACES OF TRACE CLASS OPERATORS 25 [28] B. Russo, Isometries of the trace class, Proc. Amer. Math. Soc. 23, 213 (1969). [29] S. Sakai, C∗-algebras and W ∗-algebras. Springer Verlag. Berlin 1971. [30] M. Takesaki, Theory of operator algebras I, Springer, New York, 2003. [31] D. Tan, Extension of isometries on unit sphere of L∞, Taiwanese J. Math. 15, 819-827 (2011). [32] D. Tan, On extension of isometries on the unit spheres of Lp-spaces for 0 < p ≤ 1, Nonlinear Anal. 74, 6981-6987 (2011). [33] D. Tan, Extension of isometries on the unit sphere of Lp-spaces, Acta. Math. Sin. (Engl. Ser.) 28, 1197-1208 (2012). [34] R. Tanaka, A further property of spherical isometries, Bull. Aust. Math. Soc., 90, 304-310 (2014). [35] R. Tanaka, The solution of Tingley's problem for the operator norm unit sphere of complex n × n matrices, Linear Algebra Appl. 494, 274-285 (2016). [36] R. Tanaka, Spherical isometries of finite dimensional C ∗-algebras, J. Math. Anal. Appl. 445, no. 1, 337-341 (2017). [37] R. Tanaka, Tingley's problem on finite von Neumann algebras, preprint 2016. [38] D. Tingley, Isometries of the unit sphere, Geom. Dedicata 22, 371-378 (1987). [39] R.S. Wang, Isometries between the unit spheres of C0(Ω) type spaces, Acta Math. Sci. (Eng- lish Ed.) 14, no. 1, 82-89 (1994). [40] X. Yang, X. Zhao, On the extension problems of isometric and nonexpansive mappings. In: Mathematics without boundaries. Edited by Themistocles M. Rassias and Panos M. Pardalos. 725-748, Springer, New York, 2014. Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain. E-mail address: [email protected] Departamento de Matem´atica, Centro de Ciencias F´ısicas e Matem´aticas, Universi- dade Federal de Santa Catarina, Brazil E-mail address: [email protected] Departamento de An´alisis Matem´atico, Facultad de Ciencias, Universidad de Granada, 18071 Granada, Spain. E-mail address: [email protected] Departamento de An´alisis Matem´atico, Facultad de Matem´aticas, Universidad Com- plutense de Madrid, Madrid 28040, Spain E-mail address: [email protected]
1906.05359
2
1906
2019-10-21T17:39:44
On upper triangular operator matrices over C*-algebras
[ "math.FA" ]
We study adjointable, bounded operators on the direct sum of two copies of the standard Hilbert C*-module over a unital C*-algebra A that are given by upper triangular 2 by 2 operator matrices. Using the definition of A-Fredholm and semi-A-Fredholm operators given in [3], [4], we obtain conditions relating semi-A-Fredholmness of these operators and that of their diagonal entries, thus generalizing the results in [1], [2]. Moreover, we generalize the notion of the spectra of operators by replacing scalars by the center of the C*-algebra A denoted by Z(A).Considering these new spectra in Z(A) of bounded, adjointable operators on Hilbert C*-modules over A related to the classes of A-Fredholm and semi-A-Fredholm operators, we prove an analogue or a generalized version of the results in [1] concerning the relationship between the spetra of 2 by 2 upper triangular operator matrices and the spectra of their diagonal entries.
math.FA
math
ON UPPER TRIANGULAR OPERATOR MATRICES OVER C∗-ALGEBRAS STEFAN IVKOVI ´C Abstract. In this paper we study the operator matrices MA C = (cid:20) F C 0 D (cid:21) acting on HA ⊕ HA, where F, D ∈ Ba(HA). We investigate the relationship between the semi-Fredholm properties of MA C and of F, D, when F, D are fixed and C varies over Ba(HA), as an analogue of the results by Djordjevi´c. KEYWORDS: Hilbert C*- module, semi-A-Fredholm operator, A-Fredholm spec- tra, perturbations of spectra, essential spectra and operator matrices. MSC (2010): Primary MSC 47A53 and Secondary MSC 46L08 1. Introduction Perturbations of spectra of operator matrices were earlier studied in several pa- pers such as [1]. In [1] Djordjevic lets X and Y be Banach spaces and the operator MC : X ⊕ Y −→ X ⊕ Y be given as 2 × 2 operator matrix (cid:20) A C 0 B (cid:21) where A ∈ B(X), B ∈ B(Y ) and C ∈ B(Y, X). Djordjevic investigates the relationship between certain semi-Fredholm properties of A, B and certain semi-Fredholm prop- erties of MC. Then he deduces as corollaries the description of the intersection of ′s, when C varies over all operators in B(Y, X) and A, B are fixed, spectra of MC in terms of spectra of A and B. The spectra which he considers are not in general ordinary spectra, but rather different kind of Fredholm spectra such as essential spectra, left and right Fredholm spectra etc... Some of the main results in [1] are Theorem 3.2, Theorem 4.4 and Theorem 4.6. In Theorem 3.2 Djordjevic gives necessary and sufficient conditions on operators A and B for the operator MC to be Fredholm. Recall that two Banach spaces U and V are isomorpphic up to a finite dimen- sional subspace, if one following statements hold: (a): there exists a bounded below operator J1 : U ⇒ V , such that dim V /J1(U ) < ∞, or ; (b): there exists a bounded below operator J2 : V ⇒ U , such that dim V /J2(V ) < ∞. Recall also that for a Banach space X, the sets Φ(X), Φl(X), Φr(X) denote the sets of all Fredholm, left-Fredholm and right-Fredholm operators on X, respectively. Theorem 1.1. [1, Theorem 3.2] Let A ∈ L(X) and B ∈ L(Y ) be given and consider the statments: (i): MC ∈ Φ(X ⊕ Y ) for some C ∈ L(Y, X); (ii): (a): A ∈ Φl(X); (b): B ∈ Φr(Y ); 1 2 STEFAN IVKOVI ´C (c): N (B) and X/R(A) are isomorphic up to a finite dimensional sus- pace. Then (i) ⇔ (ii). The implication (i) ⇒ (ii) was proved in [2], whereas Djordjevic proves the implication (ii) ⇒ (i). Similarly in Theorem 4.4 and Theorem 4.6 of [1] Djordjevic investigates the case when MC is right and left semi-Fredholm operator, respectively. Here we are going to recall these results as well, but first we repeat the following definition from [1]: Definition 1.2. [1, Definition 4.2] Let X and Y be Banach spaces. We say that X can be embeded in Y and write X (cid:22) Y if and only if there exists a left invertible operator J : X → Y. We say that X can essentially be embedded in Y and write X ≺ Y, if and only if X (cid:22) Y and Y /T (X) is an infinite dimensional linear space for all T ∈ L(X, Y ). Remark 1.3. [1, Remark 4.3] Obviously, X (cid:22) Y if and only if there exists a right invertible operator J1 : Y → X. If H and K are Hibert spaces, then H (cid:22) K if and only if dim H ≤ dim K. Also H ≺ K if and only if dim H < dim K and K is ifinite dimensional. Here dim H denotes the orthogonal dimension of H. Theorem 1.4. [1, Theorem 4.4] Let A ∈ L(X) and B ∈ L(Y ) be given operators and consider the following statements: (i): (a): B ∈ Φr(Y ); (b): (A ∈ Φr(X)), or (R(A) is closed and complemented in X and X/R(A) (cid:22) N (B)). (ii): MC ∈ Φr(X ⊕ Y ) for some C ∈ L(X, Y ). (iii): (a): B ∈ Φr(Y ); (b): A ∈ Φr(X) or (R(A) is not closed, or N (B) ≺ X/R(A) does not hold. Then (i) ⇔ (ii) ⇔ (iii). Theorem 1.5. [1, Theorem 4.6] Let A ∈ L(X) and B ∈ L(Y ) be given operators and consider the following statements: (i): (a): A ∈ Φl(Y ); (b): (B ∈ Φl(X)), or (R(B) and N (B) are closed and complemented subspaces of Y and N (B)) (cid:22) X/R(A). (ii): MC ∈ Φl(X ⊕ Y ) for some C ∈ L(Y, X). (iii): (a): A ∈ Φl(X); (b): B ∈ Φl(X), or (R(B) is not closed, or R(A)◦ ≺ N (B) does not hold. Then (i) ⇔ (ii) ⇔ (iii). Now, Hilbert C∗-modules are natural generalization of Hilbert spaces when the field of scalars is replaced by a C∗-algebra. Fredholm theory on Hilbert C∗-modules as a generalization of Fredholm theory on Hilbert spaces was started by Mishchenko and Fomenko in [4]. They have elaborated the notion of a Fredholm operator on the standard module HA and proved the generalization of the Atkinson theorem. Their definition of A-Fredholm operator on HA is the following: [4, Definition ] A (bounded A linear) operator F : HA → HA is called A- Fredholm if 1) it is adjointable; ON UPPER TRIANGULAR OPERATOR MATRICES OVER C ∗-ALGEBRAS 3 2) there exists a decomposition of the domain HA = M1 ⊕N1, and the range, HA = M2 ⊕N2, where M1, M2, N1, N2 are closed A-modules and N1, N2 have a finite number of generators, such that F has the matrix from (cid:20) F1 0 0 F4 (cid:21) with respect to these decompositions and F1 : M1 → M2 is an isomorphism. The notation ⊕ denotes the direct sum of modules without orthogonality, as given in [5]. In [3] we vent further in this direction and defined semi-A-Fredholm operators on Hilbert C∗-modules. We investigated then and proved several properties of these generalized semi Fredholm operators on Hilbert C∗-modules as an analogue or generalization of the well-known properties of classical semi-Fredholm operators on Hilbert and Banach spaces. In particular we have shown that the class of upper semi-A-Fredholm operators and lower semi-A-Fredholm operators on HA, denoted by MΦ+(HA) and MΦ−(HA), respecively, are exactly those that are one-sided invertible modulo com- pact operators on HA. Hence they are natural generalizations of the classical left and right semi-Fredholm operators on Hilbert spaces. The idea in this paper was to use these new classes semi-A-Freholm of operators on HA and prove that an analogue or a generalized version of [1, Theorem 3.2], [1, Theorem 4.4], [1, Theorem 4.6] hold when one considers these new classes of operators. We let Ba(HA) denote the set of all bounded, adjointable operators on HA and we consider M A C : HA ⊕ HA → HA ⊕ HA given as 2 × 2 operator C = (cid:20) F C 0 D (cid:21) , where F, C, D ∈ Ba(HA). Using this set up and these matrix M A generalized classes of A-Fredholm and semi-A-Fredholm operators on HA defined in [3], [4], we obtain generalizations of Theorem 3.2, Theorem 4.4 and Theorem 4.6 in [1]. Actually, our Theorem 3.2 is a generalization of a result in [2], as the implication in one way in Theorem 3.2 in [1] was already proved in [2]. In addition, we show that in the case when X = Y = H where H is a Hilbert space Theorem 4.4 and Theorem 4.6 in [1] can be simplified. Let us remind now the definition of the essential spectrum of bounded operators on Banach spaces. Namely, for a bounded operator T on a Banach space, the essential spectrum of T denoted σe(T ) is defined to be the set of all λ ∈ C for which T − λI is not Fredholm. In [1] Djordjevic considers the essential spectra of A, B, MC and he describes the situation when σe(MC) = σe(A) ∪ σe(B) in a chain of propositions. He shows first in Proposition 3.1 that σe(MC ) ⊂ σe(A)∪σe(B) in general and then, in Proposition 3.5 he gives sufficient conditions on A and B for the equality to hold. Next, passing from Hilbert space to Hilbert C∗-modules we don't only replace the field of scalars by a C∗-algebra A, but also work with Z(A) valued spectrum instead of the standard one. Namely, given an A-linear, bounded, adjointable operator F on HA, we consider the operators of the form F − α1 as α varies over Z(A) and this gives rise to a different kind of spectra of F in Z(A) as a generalization of ordinary spectra of F in C. Using the generalized definitions of Fredholm and semi-Fredholm operators on HA given in [4] and [3] together with these new, generalized spectra in Z(A). Finally we give a description of the intersection, when C varies over Ba(HA), of generalized essential spectra in Z(A), of the operator matrix M A C . We deduce this description as corollary from our generalizations of Theorem 3.2 in [1]. Similar 4 STEFAN IVKOVI ´C corollaries follows from our generalizations of Theorem 4.4 and Theorem 4.6 in [1], however in these corollaries we consider generalized left and right Fredholm spectra of M A C instead of generalized essential spectrum of M A C . 2. Preliminaries In this section we are going to introduce the notation, the definitions in [3] that are needed in this paper as well as some auxiliary results which are going to be used later in the proofs. Throughout this paper we let A be a unital C∗- algebra, HA be the standard module over A and we let Ba(HA) denote the set of all bounded , adjointable operators on HA. Next, for the C∗-algebra A, we let Z(A) = {α ∈ A αβ = βα for all β ∈ A} and for α ∈ Z(A) we let αI denote the operator from HA into HA given by αI(x) for all x ∈ HA. The operator αI is obviously A-linear since α ∈ Z(A) and it is adjointable with its adjoint α∗I. According to [5, Definition 1.4.1], we say that a Hilbert C∗-module M over A is finitely generated if there exists a finite set {xi} ⊆ M such that M equals the linear span (over C and A) of this set. Definition 2.1. [3, Definition 2.1] Let F ∈ Ba(HA). We say that F is an upper semi-A-Fredholm operator if there exists a decomposition F−→ M2 ⊕N2 = HA HA = M1 ⊕N1 with respect to which F has the matrix (cid:20) F1 0 0 F4 (cid:21) , where F1 is an isomorphism M1, M2, N1, N2 are closed submodules of HA and N1 is finitely generated. Similarly, we say that F is a lower semi-A-Fredholm operator if all the above conditions hold except that in this case we assume that N2 ( and not N1 ) is finitely generated. Set MΦ+(HA) = {F ∈ Ba(HA) F is upper semi-A-Fredholm }, MΦ−(HA) = {F ∈ Ba(HA) F is lower semi-A-Fredholm }, MΦ(HA) = {F ∈ Ba(HA) F is A-Fredholm operator on HA}. Lemma 2.2. Let M, N, W be Hilbert C∗-modules over a unital C∗-algebra A. If F ∈ Ba(M, N ), D ∈ Ba(N, W ) and DF ∈ MΦ(M, W ), then there exists a chain of decompositions M = M ⊥ 2 ⊕ M2 F−→ F(M ⊥ 2 ) ⊕ R D−→ W1 ⊕W2 = W w.r.t. which F, D have the matrices (cid:20) F1 0 0 F4 (cid:21) , (cid:20) D1 D2 D4 (cid:21) , respectively, where 0 F1, D1 are isomorphisms, M2, W2 are finitely generated, F(M ⊥ addition M = M ⊥ DF−→ W1 ⊕W2 = W is an MΦ-decomposition for DF. 2 ) ⊕ R = N and in 2 ⊕ M2 Proof. By the proof of [5, Theorem 2.7.6 ] applied to the operator there exists an MΦ-decomposition DF ∈ MΦ(M, W ), M = M ⊥ 2 ⊕ M2 DF−→ W1 ⊕W2 = W for DF. This is because the proof of [5, Theorem 2.7.6 ] also holds when we consider arbitrary Hilbert C∗-modules M and W over unital C∗-algebra A and not only the ON UPPER TRIANGULAR OPERATOR MATRICES OVER C ∗-ALGEBRAS 5 standard module HA. Then we can proceed as in the proof of Theorem 2.2 [3], part 2) ⇒ 1). (cid:3) Lemma 2.3. If D ∈ MΦ−(HA), then there exists an MΦ−-decompo- sition HA = N ′ 2 = HA for D. Similarly, if F ∈ MΦ+(HA), then there exists 1 an MΦ+-decomposition HA = M ⊥ D−→ M2 ⊕ N ′ 2 ⊕ N2 = HA for F. F−→ N ⊥ ⊥ ⊕ N ′ 1 2 ⊕ N1 Proof. Follows from the proofs of Theorem 2.2 [3] and Theorem 2.3 [3], part 1) ⇒ 2). (cid:3) Definition 2.4. [3, Definition 5.1] Let F ∈ MΦ(HA). We say that F ∈ MΦ− if there exists a decomposition +(HA) HA = M1 ⊕N1 with respect to which F has the matrix F−→ M2 ⊕N2 = HA (cid:20) F1 0 0 F4 (cid:21) , where F1 is an isomorphism, N1, N2 are closed, finitely generated and N1 (cid:22) N2, that is N1 is isomorphic to a closed submodule of N2. We define similarly the class MΦ+ −(HA), the only difference in this case is that N2 (cid:22) N1. Then we set MΦ− +(HA) = ( MΦ− +(HA)) ∪ (MΦ+(HA) \ MΦ(HA)) and MΦ+ −(HA) = ( MΦ+ −(HA)) ∪ (MΦ−(HA) \ MΦ(HA)) 3. Perturbations of spectra in Z(A) of operator matrices acting on HA ⊕ HA It this section we will consider the operator MA C (F, D) : HA ⊕ HA → HA ⊕ HA given as 2 × 2 operator matrix (cid:20) F C 0 D (cid:21) , where C ∈ Ba(HA). To simplify notation, throughout this paper, we will only write MA MA Let σA following proposition. C (F, D) when F, D ∈ Ba(HA) are given. C ) = {α ∈ Z(A)MA e (MA C − αI is not A-Fredholm }. Then we have the C instead of Proposition 3.1. For given F, C, D ∈ Ba(HA), one has σA e (MA C ) ⊂ (σA e (F) ∪ σA e (D)). Proof. Observe first that C − αI =(cid:20) 1 0 D − α1 (cid:21) (cid:20) 1 C 1 (cid:21) (cid:20) F − α1 0 0 0 MA 0 1 (cid:21) . 0 1 (cid:21) is clearly invertible in Ba(HA ⊕ HA) with inverse (cid:20) 1 −C Now (cid:20) 1 C it follows that (cid:20) 1 C (cid:20) 1 0 D − α1 (cid:21) are A-Fredholm, then MA of A-Fredholm operators. But, if F−αI is A-Fredholm, then clearly(cid:20) F − α1 0 1 (cid:21) is A-Fredholm. If, in addition both (cid:20) F − α1 0 0 0 1 0 C − αI is A-Fredholm being a composition (cid:21) , so 1 (cid:21) and 0 0 1 (cid:21) 6 STEFAN IVKOVI ´C is A-Fredholm, and similarly if D − αI is A-Fredholm, then (cid:20) 1 0 D − α1 (cid:21) is A-Fredholm. Thus, if both F − αI and D − αI are A-Fredholm, 0 then MA C − αI is A-Fredholm. The proposition follows. (cid:3) This proposition just gives an inclusion. We are going to investigate in which cases the equality holds. To this end we introduce first the following theorem. Theorem 3.2. Let F, D ∈ Ba(HA). If MA C ∈ Ba(HA), then F ∈ MΦ+(HA), D ∈ MΦ−(HA) and for all decompositions C ∈ MΦ(HA ⊕ HA) for some HA = M1 ⊕N1 1 ⊕N ′ 1 HA = M ′ F−→ M2 ⊕N2 = HA, D−→ M ′ 2 ⊕N ′ w.r.t. which F, D have matrices (cid:20) F1 2 = HA 0 0 F4 (cid:21) , (cid:20) D1 0 D4 (cid:21) , respectively, where 0 F1, D1 are isomorphisms, and N1, N ′ 2 are finitely generated, there exist closed sub- modules N ′ 1, N2, N2 such that N2 ∼= N2, N ′ N ′ 1, 1 are finitely generated and N ′ 1 ∼= N ′ N2 ⊕ N2 ∼= N ′ 1, N2 and N ′ 1 ⊕ 1. C = D′C′F′ where Proof. Again write MA C as MA F′ =(cid:20) F 0 1 (cid:21) , C′ =(cid:20) 1 C 1 (cid:21) , D′ =(cid:20) 1 0 0 D (cid:21) . 0 0 Since MA C is A-Fredholm, if is a decomposition w.r.t. which MA HA⊕HA = M ⊕N MA C−→ M ′ ⊕N ′ = HA⊕HA C has the matrix (cid:20) (MA 0 C )1 C )4 (cid:21) where 0 (MA C )1 is an isomorphism and N, N ′ are finitely generated, then by Lemma 2.2 and (MA also using that C′ is invertible, one may easily deduce that there exists a chain of decompositions HA⊕HA = M ⊕N F′ −→ R1 ⊕R2 C′ −→ C′(R1) ⊕C′(R2) D′ −→ M ′ ⊕N ′ = HA⊕HA w.r.t. which F′, C′, D′ have matrices 1 (cid:20) F′ 4 (cid:21) ,(cid:20) C′ 4 (cid:21) w.r.t. the decomposition 0 F′ 1, C′ 4, D′ 0 1, C′ 0 1 0 D′ (cid:20) D′ 0 1 4 (cid:21) ,(cid:20) D′ 0 0 C′ 1 D′ 2 D′ 4 (cid:21) , respectively, where F′ 1 are isomorphisms. So D′ has the matrix HA⊕HA = WC′(R1) ⊕WC′(R2) D′ −→ M ′ ⊕N ′ = HA⊕HA, where W has the matrix (cid:20) 1 −D′ 0 1 1 −1D′ 2 (cid:21) w.r.t the decomposition C′(R1) ⊕C′(R2) W−→ C′(R1) ⊕C′(R2) and is therefore an isomorphism. It follows from this that F′ ∈ MΦ+(HA⊕HA), D′ ∈ MΦ−(HA⊕HA), as N and N ′ are finitely generated submodules of HA⊕HA . Moreover R2 ∼= WC′(R2), as WC′ is an isomorphism. ON UPPER TRIANGULAR OPERATOR MATRICES OVER C ∗-ALGEBRAS 7 Since there exists an adjointable isomorphism between HA and HA ⊕ HA, using [3, Theorem 2.2 ] and [3, Theorem 2.3] it is easy to deduce that F′ is left invertible and D′ is right invertible in the Calkin algebra on Ba(HA ⊕ HA)/ K(HA ⊕ HA). It follows from this that F is left invertible and D is right invertible in the Calkin algebra Ba(HA)/K(HA), hence F ∈ MΦ+(HA) and D ∈ MΦ−(HA) again by [3, Theorem 2.2 ] and [3, Theorem 2.3 ], respectively. Choose arbitrary MΦ+ and MΦ− decompositions for F and D respectively i.e. HA = M1 ⊕N1 F−→ M2 ⊕N2 = HA, HA = M ′ 1 ⊕N ′ 1 D−→ M ′ 2 ⊕N ′ 2 = HA. Then and HA ⊕ HA = (M1 ⊕ HA) ⊕(N1 ⊕ {0}) ↓ F′ HA ⊕ HA = (M2 ⊕ HA) ⊕(N2 ⊕ {0}) HA ⊕ HA = (HA ⊕ M ′ 1) ⊕({0} ⊕ N ′ 1) ↓ D′ HA ⊕ HA = (HA ⊕ M ′ 2) ⊕({0} ⊕ N ′ 2) are MΦ+ and MΦ− decompositions for F′ and D′ respectively. Hence the decom- position HA ⊕ HA = M ⊕N F′ −→ R1 ⊕R2 = HA ⊕ HA and the MΦ+ decomposition given above for F′ are two MΦ+ decompositions for F′. Again, since there exists an adjointable isomorphism between HA ⊕ HA and HA, we may apply [3, Corollary 2.18 ] to operator F′ to deduce that ((N2⊕{0}) ⊕P ) ∼= (R2 ⊕ P ) for some finitely generated submodules P, P of HA⊕HA. Similarly, since HA ⊕ HA = WC′(R1) ⊕WC′(R2) D′ −→ M ′ ⊕N ′ = HA ⊕ HA and HA ⊕ HA = (HA ⊕ M ′ 1) ⊕({0} ⊕ N ′ 1) ↓ D′ HA ⊕ HA = (HA ⊕ M ′ 2) ⊕({0} ⊕ N ′ 2) are two MΦ− decompositions for D′, we may by the same arguments apply [3, Corollary 2.19 ] to the operator D′ to deduce that (({0} ⊕ N ′ 1) ⊕P ′) ∼= (WC′(R2) ⊕ P ′) for some finitely generated submodules P ′, P ′ of HA ⊕ HA. Since WC′ is an iso- morphism, we get ((WC′(R2) ⊕ P ′) ⊕ P ) ∼= (WC′(R2) ⊕ P ′ ⊕ P ) ∼= (R2 ⊕ P ⊕ P ′) ∼= ((R2 ⊕ P ) ⊕ P ′). Hence (((N2 ⊕ {0}) ⊕P ) ⊕ P ′) ∼= ((({0} ⊕ N ′ 1) ⊕P ′) ⊕ P ). This gives (N2 ⊕ P ⊕ P ′) ∼= (N ′ of modules in the sense of [5, Example 1.3.4 ]). Now 1 ⊕ P ′ ⊕ P ) (Here ⊕ always denotes the direct sum N2 ⊕ P ⊕ P ′ = (N2 ⊕ {0} ⊕ {0}) ⊕({0} ⊕ P ⊕ P ′), 1 ⊕ {0} ⊕ {0}) ⊕({0} ⊕ P ′ ⊕ P ) N1 ⊕ P ′ ⊕P ′ = (N ′ 8 STEFAN IVKOVI ´C and they are submodules of L5(HA) which is isomorphic to HA ( the notation L5(HA) is as in [5, Example 1.3.4 ]). Call the isomorphism betwen HA for and L5(HA) for U and set N2 = U(N2 ⊕ {0} ⊕ {0}), N2 = U({0} ⊕ P ⊕ P ′), N1 = U(N ′ 1 = U({0} ⊕ P ′ ⊕ P ). 1 ⊕ {0} ⊕ {0}), N ′ Since P, P ′, P , P ′ are finitely generated, the result follows. (cid:3) Remark 3.3. [1, Theorem 3.2 ], part (1) ⇒ (2) follows actually as a corollary from our Theorem 3.2 in the case when X = Y = H, where H is a Hilbert space. Indeed, by Theorem 3.2 if MC ∈ Φ(H ⊕ H), then F ∈ Φ+(H) and D ∈ Φ−(H). Hence ImF and ImD are closed, dim ker F, dim ImD⊥ < ∞. W.r.t. the decompositions H = ker F⊥ ⊕ ker F F−→ ImF ⊕ ImF⊥ = H and H = ker D⊥ ⊕ ker D D−→ ImD ⊕ ImD⊥ = H, 0 0 0 0 F4 (cid:21) , (cid:20) D1 D4 (cid:21) , respectively, where F1, D1 are iso- F, D have matrices (cid:20) F1 morphisms. From Theorem 3.2 it follows that there exist closed subspaces N2, N2, N ′ 1, ∼= ker D, dim N2, dim that N2 ∼= ImF⊥, N ′ 1 N ′ ( N2 ⊕ N2) ∼= ( 1). But this just means that ImF⊥ and ker D are isomorphic up to a finite dimensional subspace in the sense of [1, Definition 2.2 ] because we consider Hilbert subspaces now. 1 < ∞ and 1 such N ′ N ′ N ′ 1 ⊕ Proposition 3.4. Suppose that there exists some C ∈ Ba(HA) such that the in- clusion σA e (D) is proper. Then for any e (F) ∪ σA e (MA C ) ⊂ σA α ∈ [σA e (F) ∪ σA e (D)] \ σA e (MA C ) we have Proof. Assume that α ∈ σA e (F) ∩ σA e (D). α ∈ [σA e (F) \ σA e (D)] \ σA e (MA C ). e (MA Then (F − α1) /∈ MΦ(HA) and (D − α1) ∈ MΦ(HA). Moreover, since α /∈ σA C − α1) is A-Fredholm. From Theorem 3.2, it follows that (F − α1) ∈ MΦ+(HA). Since (F − α1) ∈ MΦ+(HA), (D − α1) ∈ MΦ(HA), we can find decompositions C ), then (MA HA = M1 ⊕N1 HA = M ′ 1 ⊕N ′ 1 F−α1−→ M2 ⊕N2 = HA, D−α1−→ M ′ 2 = HA 2 ⊕N ′ w.r.t. which F − α1, D − α1 have matrices (cid:20) (F − α1)1 0 0 (F − α1)4 (cid:21) , (cid:20) (D − α1)1 0 0 (D − α1)4 (cid:21) , N ′ respectively, where (F−α1)1, (D−α1)1 are isomorphisms, N1, N ′ generated. By Theorem 3.2 there exist then closed submodules N2, N2, N ′ 1 such that N2 ∼= N2, N ′ 1 ⊕ 1 are 1, 1 finitely generated. But then, since N ′ 1 is finitely generated (as (D−α1) ∈ MΦ(HA)), we get that N ′ 1) is N ′ 1 are finitely generated). Thus ( N2 ⊕ N2) finitely generated also (as both N ′ is finitely generated as well, so N2 is finitely generated. Therefore N2 is finitely 1 is finitely generated being isomorphic to N ′ 1, ( N2 ⊕ N2) ∼= ( N ′ N ′ 1) and N2, 1. Hence ( N ′ 2 are finitely 1 and N ′ ∼= N ′ 1 and N ′ N ′ 1 ⊕ ON UPPER TRIANGULAR OPERATOR MATRICES OVER C ∗-ALGEBRAS 9 generated, being isomorphic to N2. Hence F − α1 is in MΦ(HA). This contradicts the choice of α ∈ [σA e (F) \ σA e (D)] \ σA e (MA C ). Thus Analogously we can prove [σA e (F) \ σA e (D)] \ σA e (MA C ) = ∅. [σA e (D) \ σA e (F)] \ σA e (MA C ) = ∅. The proposition follows. (cid:3) Next, we define the following classes of operators on HA : MS+(HA) = {F ∈ Ba(HA (F − α1) ∈ MΦ+ −(HA) whenever α ∈ Z(A) and (F − α1) ∈ MΦ±(HA)}, MS−(HA) = {F ∈ Ba(HA (F − α1) ∈ MΦ− +(HA) whenever α ∈ Z(A) and (F − α1) ∈ MΦ±(HA)}. Proposition 3.5. If F ∈ MS+(HA) or D ∈ MS−(HA), then for all C ∈ Ba(HA), we have σA e (MA C ) = σA e (F) ∪ σA e (D) Proof. By Proposition 3.4, it suffices to show the inclusion. Assume that α ∈ [σA e (F) ∪ σA e (D)] \ σA e (MA C ). Then, (MA C − α1) ∈ MΦ(HA ⊕ HA). By Theorem 3.2, we have (F − α1) ∈ MΦ+(HA), (D − α1) ∈ MΦ−(HA). Let again HA = M1 ⊕N1 HA = M ′ 1 ⊕N ′ 1 F−α1−→ M2 ⊕N2 = HA, D−α1−→ M ′ 2 = HA 2 ⊕N ′ be decompositions w.r.t. which F − α1, D − α1 have matrices (cid:20) (F − α1)1 0 0 (F − α1)4 (cid:21), (cid:20) (D − α1)1 0 0 (D − α1)4 (cid:21) , 1 N ′ ∼= N ′ N ′ 1) and N2, ⊕ 1 such that N2 ∼= N2, N ′ respectively, where (F − α1)1, (D − α1)1, are isomorphisms and N1, N ′ 2 are finitely generated submodules of HA. Again, by Theorem 3.2, there exist closed submodules 1, ( N2 ⊕ N2) ∼= ( N ′ N2, N2, N ′ 1 are 1, 1 finitely generated submodules. If F ∈ MS+(HA), then since (F − α1) ∈ MΦ±(HA), we get that (F − α1) ∈ MΦ+ (F − α1) ∈ MΦ−(HA) in particular. So (F − α1) ∈ MΦ+(HA) ∩ MΦ−(HA) and by [3, Corollary 2.4], we know that MΦ+(HA) ∩ MΦ−(HA) = MΦ(HA). Then, by [3, Lemma 2.16], we have that N2 must be finitely generated, hence N2 must be finitely generated. Thus N2 ⊕ N2 is finitely generated. Since ( N2 ⊕ N2) ∼= ( N ′ 1 is finitely generated also. So (D − α1) ∈ MΦ(HA). Similarly, we can show that if D ∈ S−(HA), then (F − α1) ∈ MΦ(HA). In both cases (F − α1) ∈ MΦ(HA) and (D − α1) ∈ MΦ(HA), which contradicts that α ∈ σA (cid:3) 1 is finitely generated, hence N ′ 1), it follows that N ′ −(HA). Thus e (F) ∪ σA e (D). N ′ N ′ 1 ⊕ Theorem 3.6. Let F ∈ MΦ+(HA), D ∈ MΦ−(HA) and suppose that there exist decompositions HA = M1 ⊕N1 HA = N ′ 1 w.r.t. which F, D have matrices ⊥ ⊕ N ′ 1 F−→ N ⊥ D−→ M ′ 2 ⊕ N2 = HA 2 = HA 2 ⊕N ′ 10 STEFAN IVKOVI ´C (cid:20) F1 0 0 F4 (cid:21), (cid:20) D1 0 0 D4 (cid:21) , respectively, where F1, D1 are isomorphims, N1, N ′ sume also that one of the following statements hold: a) There exists some J ∈ Ba(N2, N ′ generated. b) There exists some J′ ∈ Ba(N ′ erated. Then MA C ∈ MΦ(HA ⊕ HA) for some C ∈ Ba(HA). 1, N2) such that N ′ 1 ∼= ImJ′, (ImJ′)⊥ is finitely gen- 2 are finitely generated and as- 1) such that N2 ∼= ImJ and ImJ⊥ is finitely Remark 3.7. ImJ⊥ in part a) denotes the orthogonal complement of ImJ in N ′ 1 and ImJ′⊥ denotes the orthogonal complement of ImJ′ in N2. By [5, Theorem 2.3.3 ], if ImJ is closed, then ImJ is indeed orthogonally comple- mentable, so since in assumption a) above ImJ ∼= N2, it follows that ImJ is closed, so N ′ 1 = ImJ ⊕ ImJ⊥. Similarly, in b) N2 = ImJ′ ⊕ ImJ′⊥. Proof. Suppose that b) holds, and consider the operator J′ = J′PN ′ where PN ′ 1. Then J′ can be considered as a bounded denotes the orthogonal projection onto N ′ adjointable operator on HA (as N2 is orthogonally complementable in (HA). To simplify notation, we let M2 = N ⊥ MA ⊥ and we let J′ = M J′. We claim then that w.r.t. the decomposition 1 = N ′ 1 2 , M ′ 1 1 HA ⊕ HA = (M1 ⊕ HA) ⊕(N1 ⊕ {0}) ↓ M J′ HA ⊕ HA = ((M2 ⊕ImJ′) ⊕ M ′ 2) ⊕(ImJ′⊥ ⊕ N ′ 2), M J′ has the matrix (cid:20) (M J′ )1 (M J′ )3 (M J′ )2 (M J′ )4 (cid:21) , where (M J′)1 is an isomorphism. To see this observe first that (M J′)1 = ⊓(M2 ⊕ImJ′)⊕M ′ 2 M J′ M1⊕HA = (cid:20) FM1 0 J′ D⊓M ′ 1 (cid:21) 2 D = D⊓M ′ ( as ⊓M ′ (M2 ⊕ImJ′) ⊕ M ′ ), where ⊓(M2 ⊕ImJ′)⊕M ′ 2 and ⊓M ′ 1. Clearly, (M J′)1 is onto (M2 ⊕ImJ′ ⊕ M ′ 2 along ImJ′⊥) ⊕ N ′ N ′ 1 2 1 denotes the projection onto denotes the projection onto M ′ 2. Now, if (M J′ )1 (cid:20) x y (cid:21) = (cid:20) 0 1 along 0 (cid:21) for 1 y = 0, so y ∈ N ′ 1 as DM ′ some x ∈ M1, y ∈ HA, then D ⊓M ′ is bounded below. 1, then J′y = J′y, so we get Fx + J′y = 0. Also Fx + J′y = 0. But, since y ∈ N ′ Since Fx ∈ M2, J′y = N2 and M2 ∩ N2 = {0}, we get Fx = J′y = 0. Since FM1 and J′ are bounded below, we get x = y = 0. So (M J′)1 is injective as well, thus an isomorphism. Recall next that N1 ⊕ {0} and ImJ′⊥ ⊕ N ′ 2 are finitely generated. By using the procedure of diagonalisation of M J′ as done in the proof of [5, Lemma 2.7.10], we obtain that M J′ ∈ MΦ(HA ⊕ HA). Assume now that a) holds. Then there exists ι ∈ Ba(ImJ, N2) s.t ιJ = idN2. 1 Letbι = ιPImJ where PImJ denote the orthogonal projection onto ImJ. (notice that ImJ is orthogonally complementable in HA since it is orthogonally complementable ON UPPER TRIANGULAR OPERATOR MATRICES OVER C ∗-ALGEBRAS 11 in N ′ 1 and HA = N ′ 1 ⊕ N ′ 1 claim that w.r.t. the decomposition ⊥). Thus bι ∈ Ba(HA). Consider Mbι = (cid:20) F bι 0 D (cid:21). We HA ⊕ HA = (M1 ⊕ (M ′ 1 ⊕ImJ)) ⊕(N1 ⊕ ImJ⊥)) ↓ Mbι HA ⊕ HA = (HA ⊕ M ′ 2) ⊕({0} ⊕ N ′ 2), Mι has the matrix (cid:20) (Mbι)1 (Mbι)3 this, observe again that (Mbι)1 = ⊓(HA⊕M ′ 2)MbιM1⊕(M ′ 1 (Mbι)2 (Mbι)4 (cid:21), where (Mbι)1 is an isomorphism. To see 1 (cid:21) , so (Mbι)1 is obviously onto =(cid:20) FM1 bι 0 (cid:21) for some x ∈ M1 and y ∈ M ′ 1 ⊕ImJ, we get that D⊓M ′ 0 ⊕ImJ) HA ⊕ M ′ 2. 1 D ⊓M ′ y = 0, so y ∈ ImJ. y (cid:21) =(cid:20) 0 Moreover, if (Mbι)1 (cid:20) x Hencebιy = ιy, so, Fx +bιy = Fx + ιy = 0. Since Fx ∈ M2, ιy ∈ N2 and M2 ∩ N2 = {0}, we get Fx = ιy = 0. As FM1 and ι are bounded below, we deduce that x = y = 0. So (Mbι)1, is also injective, hence an isomorphism. In addition, we recall that N1 ⊕ ImJ⊥ and {0} ⊕ N ′ 2 are finitely generated, so by the same arguments as before, we deduce that Mbι ∈ MΦ(HA ⊕ HA). (cid:3) Remark 3.8. We know from the proofs of [3, Theorem 2.2] and [3, Theorem 2.3], part 1) ⇒ 2) that since F ∈ MΦ+(HA), D ∈ MΦ−(HA), we can find the decompositions HA = M1 ⊕N1 HA = N ′ 1 w.r.t. which F, D have matrices F−→ N ⊥ D−→ M ′ ⊥ ⊕ N ′ 1 2 ⊕ N2 = HA, 2 ⊕N ′ 2 = HA, (cid:20) F1 0 0 F4 (cid:21), (cid:20) D1 0 0 D4 (cid:21) , respectively, where F1, D1 are isomorphisms, N1, N ′ ever, in this theorem we have also the additional assumptions a) and b). 2 are finitely generated. How- Remark 3.9. [1, Theorem 3.2 ], part (ii) ⇒ (i) follows as a direct consequence of Theorem 3.6 in the case when X = Y = H, where H is a Hilbert space. Indeed, if F ∈ Φ+(H), D ∈ Φ−(H), ker D and ImF⊥ are isomorphic up to a finite dimensional subspace, then we may let M1 = ker F⊥, N1 = ker F⊥, N2 ⊥ = ImF, N2 = ImF⊥, N ′ 1 = ker D, M ′ 2 = ImD, N ′ 2 = ImD⊥, N ′ 1 = ker D. Since ker D and ImF⊥ are isomorphic up to a finite dimensional subspace, by [1, Definition 2.2 ] this means that either the condition a) or the condition b) in The- orem 3.6 holds. By Theorem 3.6 it follows then that MC ∈ Φ(H ⊕ H). Let W (F, D) be the set of all α ∈ Z(A) such that there exist decompositions H,A = M1 ⊕N1 1 ⊕N ′ 1 HA, = M ′ F−α1−→ M2 ⊕N2 = HA, D−α1−→ M ′ 2 = HA, 2 ⊕N ′ w.r.t. which F − α1, D − α1 have matrices 12 STEFAN IVKOVI ´C (cid:20) (F − α1)1 0 0 (F − α1)4 (cid:21) , (cid:20) (D − α1)1 0 0 (D − α1)4 (cid:21) , where (F−α1)1, (D−α1)1 are isomorphisms, N1, N ′ ules and such that there are no closed submodules N2, N2, N ′ 1, 1, N2, N1 are finitely generated and that N2 ∼= N2, N ′ 1 ∼= N ′ 2 are finitely generated submod- 1 with the property N ′ Set W (F, D) to be the set of all α ∈ Z(A) such that there are no decompositions ( N2 ⊕ N2) ∼= ( N2 ′ ⊕ N ′ 2). HA = M1 ⊕N1 ⊥ ⊕N ′ 1 HA = N ′ 1 F−α1−→ N ⊥ D−α1−→ M ′ 2 ⊕N2 = HA, 2 ⊕N ′ 2 = HA, w.r.t. which F − α1, D − α1 have matrices (cid:20) (F − α1)1 0 0 (F − α1)4 (cid:21) , (cid:20) (D − α1)1 0 0 (D − α1)4 (cid:21) , where (F−α1)1, (D−α1)1, are isomorphisms N1, N ′ 2 are finitely generated and with the property that a) or b) in the Theorem 3.6 hold. Then we have the following corollary: Corollary 3.10. For given F ∈ Ba(HA) and D ∈ Ba(HA), W (F, D) ⊆ \C∈Ba(HA) σA e (MA C ) ⊆ W (F, D). Theorem 3.11. Suppose MA D ∈ MΦ−(HA) and in addition the following statement holds: Either F ∈ MΦ−(HA) or there exists decompositions C ∈ MΦ−(HA⊕HA) for some C ∈ Ba(HA). Then HA ⊕ HA = M1 ⊕N1 HA ⊕ HA = M ′ 1 ⊕N ′ 1 w.r.t. which F′, D′ have the matrices (cid:20) F′ 0 isomorphisms, N ′ erated, and M2 ∼= M ′ 1, N2 ∼= N ′ 1. F′ −→ M2 ⊕N2 = HA ⊕ HA, D′ −→ M ′ 2 ⊕N ′ 0 F′ 4 (cid:21) , (cid:20) D′ 2 = HA ⊕ HA, 0 D′ 0 1 1 4 (cid:21) , where F′ 1, D′ 1 are 2 is finitely generated, N1, N2, N ′ 1 are closed, but not finitely gen- Proof. If MA C ∈ MΦ−(HA⊕HA), then there exists a decomposition HA ⊕ HA = M1 ⊕ N1 MA C−→ M2 ⊕N2 = HA ⊕HA 0 C )1 0 (MA C )4 (cid:21) , where (MA w.r.t. which MC has the matrix(cid:20) (MA C )1 is an isomor- phism and N2 is finitely generated. By the part of [3, Theorem 2.3], part 1) ⇒ 2) we may assume that M1 = N ⊥ can be viewed as an operator in Ba(M1, (D′C′)−1(M2)), as M1 is orthogonally complementable, by [5, Theorem 2.3.3.], F′(M1) is orthogonally complementable in (D′C′)−1(M2). By the same arguments as in the proof of [3, Theorem 2.2] part 2) ⇒ 1) we deduce that there exists a chain of decompositions is adjointable. Since F′ 1 . Hence F′ M1 M1 M1 ⊕N1 F′ −→ R1 ⊕R2 w.r.t. which F′, C′, D′ have matrices (cid:20) F′ 0 −→ C′(R1) ⊕C′(R2) D′ C′ 0 C′ 4 (cid:21) , (cid:20) C′ 0 F′ 0 1 1 −→ M2 ⊕N2 4 (cid:21) , (cid:20) D′ 0 1 D′ 2 D′ 4 (cid:21) , where F′ 1, C′ 1, C′ 4, D′ 1 are isomorphisms. Hence D′ has the matrix (cid:20) D′ 0 1 4 (cid:21) , 0 D′ ON UPPER TRIANGULAR OPERATOR MATRICES OVER C ∗-ALGEBRAS 13 w.r.t. the decomposition HA ⊕ HA = WC′(R1) ⊕WC′(R2) D′ −→ M2 ⊕N2 = HA ⊕ HA, where W is an isomorphism. It follows that D′ ∈ MΦ−(HA ⊕HA), as N2 is finitely generated. Hence D ∈ MΦ−(HA) (by the same arguments as in the proof of The- orem 3.2). Next, assume that F /∈ MΦ−(HA), then F′ /∈ MΦ−(HA ⊕ HA). Therefore R2 can not be finitely generated (otherwise F′ (cid:3) would be in MΦ−(HA ⊕ HA) ). Now, R1 ∼= WC′(R1), R2 = WC′(R2). Remark 3.12. In case of ordinary Hilbert spaces, [1, Theorem 4.4 ] part 2) ⇒ 3) follows as a corollary from Theorem 3.11. Indeed, suppose that D ∈ B(H) and that F ∈ B(H) (where H is a Hilbert space). If ker D ≺ ImF⊥, this means by [1, Remark 4.4 ] that dim ker D < ∞. So, if (2) in [1, Theorem 4.4 ] holds, that is MC ∈ Φ−(H ⊕ H) for some C ∈ B(H), then by Theorem 3.11 D ∈ Φ−(H) and either F ∈ Φ−(H) or there exist decompositions H ⊕ H = M1 ⊕N1 F′ −→ M2 ⊕N2 = H ⊕ H, H ⊕ H = M ′ 1 ⊕N ′ 1 D′ −→ M ′ 2 ⊕N ′ 2 = H ⊕ H, M ′ 1 1 . Assume that dim ker D = dim ker D′ < ∞ and let N1 which satisfy the conditions described in Theorem 3.11. In particular N2, N ′ 1 are infinite dimensional whereas N ′ 2 is finite dimensional. Suppose that F /∈ Φ−(H) and that the decompositions above exist. Observe that ker D′ = {0} ⊕ ker D. Hence, if dim ker D < ∞, then dim ker D′ < ∞ . Since D′ is an isomorphism, by the same arguments as in the proof of [5, Proposition 3.6.8 ] one can deduce that ker D′ ⊆ N ′ be the orthogonal 1. Now, since ImD′ is closed complement of ker D′ in N1 as D′ ∈ MΦ−(H ⊕ H), then D′ 1 = ∞ and N ′ 1 1 = ∞ . Hence D′( N ′ dim ker D′ < ∞ , we have dim N ′ 1) is infinite dimensional subspace of N ′ 2 is finite. Thus, if F /∈ Φ−(H), we must have that ker D is infinite dimensional. Hence, we deduce, as a corollary, [1, Theorem 4.4 ] in case when X = Y = H, where H is a Hilbert space. In this case, part (3b) in [1, Theorem 4.4 ] could be reduced to the following statement: Either F ∈ Φ−(H) or dim ker D = ∞. 2. This is a contradiction since dim N ′ is an isomorphism. Since dim N ′ 1 = ker D′ ⊕ N ′ ′, that is N ′ ′ Theorem 3.13. Let F, D ∈ Ba(HA) and suppose that D ∈ MΦ−(HA) and either F ∈ MΦ−(HA) or that there exist decompositions HA = M1 ⊕N1 F−→ N ⊥ 2 ⊕N2 = HA, HA = N ′ 1 ⊥ ⊕N ′ 1 D−→ M ′ 2 ⊕N ′ 2 = HA, 0 0 F4 (cid:21) , (cid:20) D1 0 D4 (cid:21) , respectively, where w.r.t. which F, D have the matrices (cid:20) F1 0 F1, D1 are isomorphisms N ′ ι ∈ Ba(N2, N ′ MΦ−(HA ⊕ HA) for some C ∈ Ba(HA). 2, is finitely generated and that there exists some 1) such that ι is an isomorphism onto its image in N ′ 1 . Then MA C ∈ 1 by [5, Theorem 2.3.3 ], that is N ′ Proof. Since Imι is closed and ι ∈ Ba(N2, N ′ in N ′ N ′ 1. ⊥, that is Imι is orthogonally complementable in HA. Hence HA = Imι ⊕ N ′ Also, there exists J ∈ Ba(Imι, N2) such that Jι = idN2, ιJ = idImι. Let PImι be the 1), Imι is orthogonally complementable 1 for some closed submodule 1 = Im ι ⊕ N ′ 1 ⊕ N ′ 1 14 STEFAN IVKOVI ´C orthogonal projection onto Imι and set C = JPImι. Then C ∈ Ba(HA). Moreover, w.r.t. the decomposition HA ⊕ HA = (M1 ⊕ (M ′ 1 ⊕Imι)) ⊕(N1 ⊕ N ′ 1) MA C−→ C has the matrix (cid:20) (MA (MA C )1 C 3) MA (HA ⊕ M ′ 2) = HA ⊕ HA, 2) ⊕({0} ⊕ N ′ (MA (MA C )4 (cid:21) , where (MA C )2 C )1 is an isomorphism. This follows by the same arguments as in the proof of Theorem 3.6. Using that N ′ 2 is finitely generated and proceeding further as in the proof of the above mentiond theorem, we reach the desired conclusion. (cid:3) Remark 3.14. In the case of ordinary Hilbert spaces, [1, Theorem 4.4] part (1) ⇒ (2) can be deduced as a corollary from Theorem 3.13. Indeed, if F is closed and D ∈ Φ−(H), which gives that ImD is closed also, then the pair of decompositions H = (ker F)⊥ ⊕ ker F F−→ ImF ⊕ ImF⊥ = H, H = (ker F)⊥ ⊕ ker D D−→ ImD ⊕ ImD⊥ = H for F and D, respectively, is one particular pair of decompositions that satisfies the hypotheses of Theorem 3.13 as long (ImF)⊥ (cid:22) ker D. Let R(F, D) be the set of all α ∈ Z(A) such that there exists no decompositions HA = M1 ⊕N1 F−αI−→ N2 ⊥ ⊕N2 = HA, HA = N ′ 2 ⊥ ⊕N ′ 1 D−αI−→ M ′ 2 ⊕N ′ 2 = HA that satisfy the hypotheses of the Theorem 3.13. Set R′(F, D) to be the set of all α ∈ Z(A) such that there exist no decompositions HA ⊕ HA = M1 ⊕N1 F′−αI−→ M2 ⊕N2 = HA ⊕ HA, HA ⊕ HA = M ′ D′−αI−→ M ′ that satisfy the hypotheses of the Theorem 3.11. Then we have the following corollary: 1 ⊕N ′ 1 2 ⊕N ′ 2 = HA ⊕ HA Corollary 3.15. Let F, D ∈ Ba(HA). Then σA re(D) ∪ (σA re(F) ∩ R′(F, D)) ⊆ \C∈Ba(HA) σA re(MA C ) ⊆ σA re(D) ∪ (σA re(F) ∩ R(F, D)) Theorem 3.16. Let MA either D ∈ MΦ+(HA) or there exist decompositions C ∈ MΦ+(HA ⊕ HA). Then F′ ∈ MΦ+(HA ⊕ HA) and HA ⊕ HA = M1 ⊕N1 F′ −→ M2 ⊕N2 = HA ⊕ HA, HA ⊕ HA = M ′ 1 ⊕N ′ 1 D′ −→ M ′ 2 ⊕N ′ 2 = HA ⊕ HA, w.r.t. which F′, D′ have matrices (cid:20) F′ 1 1, D′ F′ N2, N ′ 1 are isomorphisms, M2 ∼= M ′ 1 are closed, but not finitely generated. 4 (cid:21) , (cid:20) D′ 0 F′ 1 4 (cid:21) , respectively, where 0 D′ 0 0 1 and N2 ∼= N ′ 1, N1 is finitely generated and ON UPPER TRIANGULAR OPERATOR MATRICES OVER C ∗-ALGEBRAS 15 Proof. Since MA C ∈ MΦ+(HA ⊕HA), there exists an MΦ+ decomposition for MA C , HA ⊕ HA = M1 ⊕N1 MA C−→ M ′ 2 ⊕N ′ 2 = HA ⊕ HA, so N1 is finitely generated. By the proof of [5, Theorem 2.7.6 ], we may assume that M1 = N ⊥ , is adjointable. As in the proof of Lemma 2.2 and Theorem 3.2 we may consider a chain of decompositions 1 . Hence F′ M1 HA ⊕ HA = M1 ⊕N1 F′ −→ R1 ⊕R2 C′ −→ C′(R1) ⊕C′(R2) D′ −→ M ′ w.r.t. which F′, C′, D′ have matrices (cid:20) F′ 0 1 4 (cid:21) , (cid:20) C′ 0 0 F′ 1 2 ⊕M ′ 2 = HA ⊕ HA 4 (cid:21) and (cid:20) D′ 0 0 C′ 1 D′ 2 D′ 4 (cid:21) , respectively, where F′ same way as in the proof of Theorem 3.11. 4, D′ 1, C′ 1, C′ 1 are isomorphisms. Then we can proceed in the (cid:3) Remark 3.17. In the case of Hilbert spaces, the implication (2) ⇒ (3) in [1, Theorem 4.6] follows as a corollary of Theorem 3.16. Indeed, for the implication (2) ⇒ (3b), we may proceed as follows: Since Im(F)0 ∼= Im(F)⊥ and (ker D)′ ∼= ker D when one considers Hilbert spaces, then by [1, Remark 4.3], (ImF)0 ≺ (ker D)′ means simply that dim ImF⊥ < ∞ whereas dim ker D = ∞. If in addition D /∈ Φ+(H), then D′ /∈ Φ+(H ⊕ H). Now, if dim Im(F)⊥ < ∞, then dim ker D = ∞, and F ∈ Φ(H) as F ∈ Φ+(H) and dim Im(F)⊥ < ∞. Then F′ ∈ Φ(H ⊕ H), so by [3, Lemma 2.16 ] N2 must be finitely generated. Thus N ′ 1 must be finitely generated being isomorphic to N2. By the same arguments as earlier, we have that ker D′ ∼= ker D and ker D′ ⊆ N ′ 1 is finitely generated means actually that N ′ 1 is finite dimensional. Hence ker D′ must be finite dimensional, so dim ker D = dim ker D′ < ∞. This is in a contradiction to ImF⊥ ≺ ker D. So, in the case of Hilbert spaces, if MC ∈ Φ+(H ⊕ H), from Theorem 3.16 it follows that F ∈ Φ+(H) and either D ∈ Φ+(H) or ImF⊥ is infinite dimensional. 1. Since we consider Hilbert spaces now, the fact that N ′ Theorem 3.18. Let F ∈ MΦ+(HA) and suppose that either D ∈ MΦ+(HA) or that there exist decompositions HA = M1 ⊕N1 F−→ N ⊥ 2 ⊕N2 = HA, HA = N ′ 1 ⊥ ⊕N ′ 1 D−→ M ′ 2 ⊕N ′ 2 = HA 0 w.r.t. which F, D have matrices (cid:20) F1 0 0 F4 (cid:21) , (cid:20) D1 0 D4 (cid:21) , respectively, where F1, D1 are isomorphisms, N ′ ι ∈ Ba(N ′ 1, N2) such that ι is an isomorphism onto its image. Then 1 is finitely generated and in addition there exists some for some C ∈ Ba(HA). MA C ∈ MΦ+(HA ⊕ HA), Proof. Let C = PN ′ apply similar arguments as in the proof of Theorem 3.6 and Theorem 3.13 denotes the orthogonal projection onto N ′ ι where PN ′ 1 1 1, then (cid:3) Remark 3.19. The implication (1) ⇒ (2) in [1, Theorem 4.6] in case of Hilbert spaces could also be deduced as a corollary from 3.18. Indeed, if ImD is closed, then D is an isomorphism from ker D⊥ onto ImD. Moreover, if F ∈ Φ+(H), then F is also an isomorphism from ker F⊥ onto ImF and dim ker F < ∞. If in addition ker D (cid:22) ImF⊥, then the pair of decompositions H = ker F⊥ ⊕ ker F F−→ ImF ⊕ ImF⊥ = H, H = ker D⊥ ⊕ ker D D−→ ImD ⊕ ImD⊥ = H 16 STEFAN IVKOVI ´C is one particular pair of decompositions that satisfies the hypotheses of Theorem 3.18. Let L′(F, D) be the set of all α ∈ Z(A) such that there exist no decompositions HA ⊕ HA = M1 ⊕N1 HA ⊕ HA = M ′ 1 ⊕N ′ 1 F′−αI−→ M2 ⊕N2 = HA ⊕ HA, D′−αI−→ M ′ 2 = HA ⊕ HA, 2 ⊕N ′ for F′ − αI, D′ − αI respectively, which satisfy the hypotheses of Theorem 3.16. Set L(F, D) to be the set of all α ∈ Z(A) such that there exist no decompositions HA = M1 ⊕N1 ⊥ ⊕N ′ 1 HA = N ′ 1 F−α1−→ N ⊥ D−α1−→ M ′ 2 ⊕N2 = HA, 2 ⊕N ′ 2 = HA, for F − α1, D − α1 respectively which satisfy the hypotheses of Theorem 3.18. Then we have the following corollary: Corollary 3.20. Corollary: Let F, D ∈ Ba(HA). Then C ) ⊆ σA σA le(F) ∪ (σA le (D) ∩ L′(F, D)) ⊆ \C∈Ba(HA) σA le (MA le (F) ∪ (σA le (D) ∩ L(F, D)) Acknowledgements: First of all, I am grateful to Professor Dragan S. Djord- jevic for suggesting the research topic of the paper and for introducing to me the relevant reference books and papers. In addition, I am especially grateful to my supervisors, Professor Vladimir M. Manuilov and Professor Camillo Trapani, for careful reading of my paper and for detailed comments and suggestions which led to the improved presentation of the paper. References [1] Dragan S. Djordjevi´c, Perturbations of spectra of operator matrices, J. Operator Theory 48(2002), 467-486. [2] J.H. Han, H.Y. Lee, W.Y. Lee, Invertible completions of 2 × 2 upper triangular operator matrices, Proc. Amer.Math. Soc. 128 (2000), 119-123. [3] S. Ivkovi´c , Semi-Fredholm theory on Hilbert C*-modules, Banach Journal of Mathematical Analysis, to appear (2019), arXiv: https://arxiv.org/abs/1906.03319 [4] A. S. Mishchenko, A.T. Fomenko, The index of eliptic operators over C*-algebras, Izv. Akad. Nauk SSSR Ser. Mat. 43 (1979), 831 -- 859; English transl., Math. USSR-Izv.15 (1980) 87 -- 112. [5] V. M. Manuilov, E. V. Troitsky, Hilbert C*-modules, In: Translations of Mathematical Mono- graphs. 226, American Mathematical Society, Providence, RI, 2005. [6] N. E. Wegge Olsen, K-theory and C*-algebras, Oxford Univ. Press, Oxford, 1993. [7] S. Zivkovi´c Zlatanovi´c, V. Rakocevi´c, D.S. Djordjevi´c, Fredholm theory, University of Nis Faculty of Sciences and Mathematics, Nis, (2019). Stefan Ivkovi´c ,The Mathematical Institute of the Serbian Academy of Sciences and Arts, Kneza Mihaila 36 p.p. 367, 11000 Beograd, Serbia, Tel.: +381-69-774237 E-mail address: [email protected]
1612.03680
4
1612
2016-12-27T14:14:36
A note on conditional risk measures of Orlicz spaces and Orlicz-type modules
[ "math.FA" ]
We consider conditional and dynamic risk measures of Orlicz spaces and study their robust representation. For this purpose, given a probability space $(\Omega,\mathcal{E},\mathbb{P})$, a sub-$\sigma$-algebra $\mathcal{F}$ of $\mathcal{E}$, and a Young function $\varphi$, we study the relation between the classical Orlicz space $L^\varphi(\mathcal{E})$ and the modular Orlicz-type module $L^\varphi_\mathcal{F}(\mathcal{E})$; based on conditional set theory, we describe the conditional order continuous dual of a Orlicz-type module; and by using scalarization and modular extensions of conditional risk measures together with elements of conditional set theory, we finally characterize the robust representation of conditional risk measures of Orlicz spaces.
math.FA
math
DRAFT: A note on conditional risk measures of Orlicz spaces and Orlicz-type modules Jos´e Orihuela, Jos´e M. Zapata Abstract We consider conditional and dynamic risk measures of Orlicz spaces and study their robust representation. For this purpose, given a prob- ability space (Ω, E , P), a sub-σ-algebra F of E , and a Young function ϕ, we study the relation between the classical Orlicz space Lϕ(E ) and the modular Orlicz-type module Lϕ F (E ); based on conditional set theory, we describe the conditional order continuous dual of a Orlicz-type mod- ule; and by using scalarization and modular extensions of conditional risk measures together with elements of conditional set theory, we finally char- acterize the robust representation of conditional risk measures of Orlicz spaces. 1 Introduction The study of dual representation of risk measures is an active field within the financial mathematics, cf.[4, 11, 18, 19, 22, 31] among many other references. In particular, the study of risk measures in conditional or dynamic discrete time settings has gained an increasing importance in the recent literature cf.[5, 8, 13, 20, 17, 25, 32]. Whereas many works have studied risk measures defined on Orlicz spaces cf.[9, 12, 31, 33, 34], to the best of our knowledge, conditional and dynamic risk measures have been studied only in the case that they are defined on Lp spaces cf.[8, 13] or modular extensions of Lp spaces cf.[17, 25, 32]. As explained by [36], one of the advantages of consider Orlicz spaces is that, in many cases, convex risk measures defined on Lp spaces have effective domain whose interior is empty, and this makes impossible to apply subdifferentiabilty results; how- ever, sometimes one can find a suitable finer Orlicz space on which a convex risk measure is defined, obtaining a domain with non-empty interior, for which subdifferentiabilty results apply. Also, Fritelli et al. [7] found a connection be- tween Orlicz spaces and the theory of utility functions. In addition, this relation has applications to the study of no-arbitrage conditions cf.[21, 30]. Thus, Or- licz spaces have a key role in the study of dynamic problems in mathematical finance. Thereby, the purpose of this paper is to extend the notions of conditional and dynamic convex risk measure, defining the notions of conditional and dy- 1 namic convex risk measure of Orlicz spaces and study their dual representation. For instance, given a probability space (Ω, E, P), a sub-σ-algebra F of E, and two Young functions ϕ, ψ : [0, +∞) → [0, +∞] with ψ ≤ ϕ, a conditional convex risk measure is a function ρ : Lϕ(E) → Lψ(F ) which is monotone (de- creasing), L0(F )-convex and Lϕ(F )-cash invariant. Then, based on the dual pair hLϕ(E), Lϕ∗ (E)i where ϕ∗(y) := supx≥0(xy − ϕ(x)) is the conjugate Young function of ϕ, we will study under which conditions we have a dual represen- tation ρ(x) = W(cid:8)E[xyF ] − α(y) ; y ∈ Lϕ∗ representation is attained. (E)(cid:9) for x ∈ Lϕ(E); and when this The works that study the dual representation of dynamic and conditional risk measures employ mainly two types of strategies. In a first stage, a technique called scalarization is adopted; more specifically, these works try to reduce the conditional case to the static case, which is known and treatable cf.[13, 8]. In a second stage, with the aim of providing a tailored analytic framework for the study of conditional risk measures, some recent works have developed new models of functional analysis: a modular based approach is proposed by Filipovic et al.[16] where some theorems of convex analysis are extended to L0(F )-modules (also see [17, 23, 24, 36, 38]); and in a more abstract level in [14] it was developed the theory of conditional sets, providing a formal formulation, which is in some precise sense stable and consistent with respect to the measure algebra associated to a probability space, and is closely related to the theory of Boolean-valued models of set theory cf.[6, 10, 26]. Therefore, in this paper we will employ a mix of both techniques: where possible we will try to reduce to the static case, where not, we will make use of tools from the modular and conditional approach. The Orlicz-type modules were introduced by Vogelpoth [36]. Based on a modular approach, he considered the L0(F )-module generated by Lϕ(E); more precisely, the Orlicz-type module is given by Lϕ F (E) = L0(F )Lϕ(E). Then, we will show that every conditional convex risk measure ρ : Lϕ(E) → Lψ(F ) can be uniquely extended to a conditional convex risk measure ¯ρ : Lϕ F (E) → L0(F ), and prove that the dual representation of the former is equivalent to the dual representation of the latter. We will also prove that the attainability of the representation is equivalent to the so-called Lebesgue property of ρ, and also to the conditional compactness of the conditional sublevel sets of ¯ρ. The manuscript is structured as follows: Section 2 is devoted to some pre- liminaries. In Section 3 we study the relation between the classical Orlicz space Lϕ(E) and the Orlicz-type module Lϕ F (E). Section 4 is devoted to the study of the conditional order continuous dual of a Orlicz-type module. Finally, in Sec- tion 5 we introduce the notions of conditional and dynamic convex risk measures and study their robust representation. 2 Preliminaries and notation Let us start by describing the setting and notation of this paper. For this purpose, we will review the different elements of the L0(F )-theory (see [16, 2 23, 38]) and, at the same time, we will identify them with the corresponding notions within of the conditional set theory ([14, 38]), describing how the former framework is embedded in the latter when we assume on the different objects the necessary stability properties. Let us fix an underlying probability space (Ω, E, P) and let us also consider some sub-σ-algebra F ⊂ E, which will remain fixed throughout this paper. We will denote by L0(E) the space (of equivalence classes) of F -measurable random variables, which equipped with the order of almost sure dominance is a Dedekind complete Riesz space. We will follow the common practice of identifying random variables and measurable sets which agree almost surely. By doing so, we can also identify the σ-algebra F with the measure algebra associated to F , which is a complete Boolean algebra. We define the set of partitions of Ω to F as follows:1 Π(Ω, F ) :=(cid:26){Ak}k∈N ⊂ F ; Ω = ∨ k Ak, Ai ∧ Aj = ∅, ∀i 6= j, i, j ∈ N(cid:27) . Given a non-empty subset M of L0(E), we define the countable concatenation hull of M by cc M :=(Xk∈N 1Ak xk ; {Ak} ∈ Π(Ω, F ), {xk} ⊂ M) . We will say that a subset M of L0(E) has the countable concatenation prop- . Given a subset M of L0(E) with the countable concatenation erty, if M = M property, we can define an equivalent relation on M ×F where the class of (x, A) is denoted by cc xA := {(y, B) ; A = B, 1Ax = 1By} . By doing so, we find that the quotient set M is a conditional set as it satisfies the axioms of Definition 2.1 of [14]. Namely, inspection shows that 1. xA = yB implies that A = B; 2. (Consistency) xA = yA and B ≤ A implies xB = yB; 3. (Stability) for {xk} ⊂ M and {Ak} ∈ Π(Ω, F ), there exists a unique x ∈ M such that xAk = xkAk for all k ∈ N. Notice that, for {xk} ⊂ M and {Ak} ∈ Π(Ω, F ), x = Pk∈N 1Ak xk satisfies that xAk = xkAk for all k ∈ N. The element x is called the concatenation of {xk} along {Ak}. In particular, L0(E) defines a conditional set, which will be denoted by L0. Also, L0(F ) obviously has the countable concatenation property, therefore it also defines a conditional set; moreover, in [14, Theorem 4.4] it is proved that this conditional set is a copy of the so-called conditional real numbers (see [14, Def- inition 4.3]). Thus, we will use the notation R for denoting the conditional set 1It is known that every partition of Ω to F is at most countable. 3 +(F ) =(cid:8)x ∈ L0(F ) ; x ≥ 0(cid:9); L0 generated by L0(F ). Likewise, the following sets have the countable concatena- tion property: L0 ¯L0(F ), defined as the space of all (equivalence classes of) F -measurable ran- dom variables whose values are in R ∪ {±∞}; and L0(F , N), the set of random variables which take values in N. Then, the corresponding conditional sets are denoted by R+, R++, R and N, respectively. In addition, since the elements of L0(F , N) are step-functions which take values in N, i.e. L0(F , N) = N , one can identify the conditional set N with the conditional natural numbers (see 5 of [14, Examples 2.3]). ++(F ) =(cid:8)x ∈ L0(F ) ; x > 0(cid:9); cc The elements x = xΩ are called conditional elements 3. Throughout this paper, some conditional subsets will be required to be defined by describing their conditional elements. Namely, let φ be certain statement which can be true of false for the conditional elements of some conditional set C. Also, let us suppose that there exists some x ∈ C such that φ(x) is true. Then, we denote by .4 [x ∈ C ; φ(x) is true] the conditional set generated by {x ∈ C ; φ(x) is true} Also, if C has the countable concatenation property and A ∈ F , then CA := {xB ; x ∈ C, B ≤ A} is also a conditional set (considering as σ-algebra FA, the trace of F on A), and is called a conditional subset (on A) of L0 (see [14, Definition 2.8]). A conditional subset CA is said to be conditionally included in DB if CA ⊂ DB, in accordance with [14], we will use the notation CA ⊏ DB. In [14] it is proved that the collection of all objects CA is a complete Boolean algebra, when it is endowed with the operations of conditional union ⊔, conditional intersection ⊓ and conditional complement ⊏ (see [14, Corollary 2.10]). We do not give the construction; instead, we refer to the proof of [14, Theorem 2.9]. cc It is also important to recall the conditional Cartesian product of conditional sets (see [14, Definition 2.14]). Suppose that L, M are two conditional subsets of L0 (on 1). Then, we define their conditional product as follows: L ⋊⋉ M := {(xA, yA) ; A ∈ F , x ∈ L, y ∈ M } , which is a conditional set as it satisfies the axioms of [14, Definition 2.1]. If we have an application f : L → M between conditional subsets with the countable concatenation property, according to [14, Definition 2.17] f is a stable function if it preserves countable concatenations, i.e. f (P 1Ak xk) = P 1Ak f (xk). In this case, if Gf = {(x, f (x)) ; x ∈ L} is the graph of f , then Gf := {(xA, yA) ; (x, y) ∈ Gf , A ∈ F} is the conditional graph of a conditional function f : L → M, as explained in [14, Definition 2.17]. For x, y ∈ L0 and A ∈ F , we will say x ≤ y (resp. x < y) on A whenever P(x ≤ yA) = 1 (resp. P(x < yA) = 1), this defines a conditional partial order as in [14, Definition 2.15]. Further, the conditional supremums and conditional infimums are defined in term of the partial ordered set (L0(E), ≤), i.e. if C is a 3In [14] it is used a different definition of conditional element, but as mentioned in [38] both objects can be identified due to the axiom of consistency 4As explained in [14], we use this set-builder notation, because the set of conditional elements {x ; φ(x) is true} is not necessarily a conditional set. 4 conditional subset of L0 (on 1) we have sup C := W CΩ and inf C := V CΩ. This conditional order, when restricted to R, becomes a conditional total order as in [14, Definition 2.15]; that is, for every r, s ∈ R there exists A, B, C ∈ F such that r = s on A, r > s on B and r < s on C. Let us consider a sequence {xn}n∈N in L0(E). Each n ∈ L0(F , N) is a concatenation of a sequence {nk} ⊂ N along a partition {Ak} ∈ Π(Ω, F ), i.e. n = Pk∈N 1Aknk. Then, we can define xn := Pk∈N 1Ak xnk . It turns out, that the family {xn}n∈L0(F ,N) is a stable family as in [14, Definition 2.20]; therefore, we can construct a conditional sequence {xn}n∈N as described in [14, Definition 2.20]. Conversely, given a conditional sequence {xn}n∈N, we can consider a sequence {xn}n∈N. We see that, there is a bijection between sequences of L0(E), and conditional sequences of L0. Suppose that E is a L0(F )-submodule of L0(E) with the countable con- catenation property, then E is a conditional vector space in the sense of [14, Definition 5.1]. Suppose that f : E → F is a homomorphism of L0(F )-modules, then inspection shows that f is stable and the corresponding conditional func- tion f : E → F is a conditionally linear function as it satisfies [14, Definition 5.1] and, conversely, for any conditionally linear conditional function f : E → F, we have that f : E → L0(F ) is a stable homomorphism of L0(F )-modules. Likewise, let f : E → L0(F ) be an application which is L0(F )-convex, i.e. f (λx + (1 − λ)y) ≤ λf (x) + (1 − λ)f (y) for every λ ∈ L0(F ) with 0 ≤ λ ≤ 1 and x, y ∈ E. Then, by [16, Theorem 3.2], we have that 1Af (1Az) = 1Af (z) for all z ∈ E, A ∈ F . From this fact, we can obtain that f is stable. Consequently, it defines a conditional function f : E → R, which is conditionally convex in the sense of [14, Definition 5.1]. Again, the converse is also true. Suppose that k · k : E → L0 +(F ) is a L0(F )-norm, i.e.: kλxk = λkxk for all λ ∈ L0(F ) and x ∈ E; kx + yk ≤ kxk + kyk for all x, y ∈ E; and, x = 0 whenever kxk = 0. In particular, k · k is stable; therefore, it defines a conditional function k · k : E → R+ which is called a conditional norm (see [14, Definition 5.11]). For instance, the absolute value · : R → R+ is a conditional norm. Other important examples, which are introduced by Filipovic et al.[16], are the Lp-type modules. More specifically, given 1 ≤ p ≤ +∞, we define k · Fkp : L0(E) → ¯L0(F ) by kxFkp =( E [limn xp ∧ nF ] V(cid:8)λ ∈ ¯L0(F ) ; x ≤ λ(cid:9) , which is a L0(F )-norm for the Lp-type module Lp 1 p , if p < +∞; if p = +∞, F (E) :=(cid:8)x ∈ L0(E) ; kxFkp ∈ L0(F )(cid:9). Inspection shows that Lp F (E) has the countable concatenation property. The cor- responding conditional vector space will be denoted by Lp F (E), or simply Lp, and can be endowed with the corresponding conditional norm k · Fkp : Lp → R+. , which allows a better under- Guo et al.[25] proved that Lp F (E) = Lp(E) standing of the relation between Lp(E) and Lp cc F (E). By using this equality for p = 1, the classical conditional expectation E[·F ] : F (E), which is of F (E) as follows: For x ∈ L1 L1(E) → L1(F ) is extended to L1 the form x = Pk∈N 1Ak xk with {xk} ⊂ L1(E) and {Ak} ∈ Π(Ω, F ), we define 5 E[xF ] :=Pk∈N 1Ak E[xkF ]. An easy check shows that this definition does not depend on the choice of {xk} and {Ak}. Clearly, this application is stable, then we consider the conditional function E[·F ] : L1 → R. Also, let us recall the Orlicz-type modules. Let ϕ : [0, +∞) → [0, +∞] be a Young function, i.e. ϕ is an increasing left-continuous convex function finite on a neighborhood of 0 with ϕ(0) = 0 and limx→+∞ ϕ(x) = +∞. The Orlicz-type module corresponding to ϕ is defined by Lϕ F (E) :=(cid:8)x ∈ L0(E) ; E(cid:2)ϕ(cid:0)xλ−1(cid:1) F(cid:3) ∈ L0(F ), It is easy to check that Lϕ the corresponding conditional vector space is denoted by Lϕ as long as F and E are fixed. F (E) has the countable concatenation property. Then, F (E) or simply Lϕ ++(F )(cid:9) . for some λ ∈ L0 In [36], it is introduced (1) kxFkϕ :=^(cid:8)λ ∈ L0 ++(F ) ; E(cid:2)ϕ(cid:0)xλ−1(cid:1) F(cid:3) ≤ 1(cid:9) , which is a L0(F )-norm for Lϕ F (E). Consequently, we have a conditional norm k · Fkϕ : Lϕ → R++ which is a conditional version of the classical Lux- emburg norm; therefore, hereafter k · Fkϕ will be referred to as the conditional Luxemburg norm. An important remark, it is that the Lp-type modules are Orlicz-type mod- Indeed, for 1 ≤ p < +∞ let us define φ(t) := tp and for p = ∞ let ules. us define φ(t) := ∞ whenever t ≥ 1 and 0 otherwise, then inspection shows that Lϕ It is also easy to prove that F (E) ⊂ Lϕ L∞ F (E) and k · Fkϕ = k · Fkp. F (E) = Lp F (E) ⊂ L1 Further, Lϕ is a conditional lattice-normed space, in the sense that x, y ∈ Lϕ F (E). with x ≤ y implies kxFkϕ ≤ kyFkϕ. If (E, k · k) is a conditional normed space, the collection B := {x + Bε ; x ∈ E, ε ∈ L0 ++(F )}, where Bε := {x ∈ E ; kxk ≤ ε} is a topological basis, which generates the topology induced by k · k (see [16]). Also, notice that B is a stable family. Then, according to [14, Proposition 3.5], the collection of conditional subsets OA where O is open for this topology and A ∈ F , defines a conditional topology in the sense of [16, Definition 3.1], which is precisely the topology induced by the conditional norm k · k : E → R. In particular, (R, · ), (Lp, k · Fkp) and (Lϕ, k · Fkϕ) can be endowed with conditional topologies. Consider two L0(F )-submodules E, F of L0(E) with the countable concate- nation property. We say that E, F is a random duality pair with respect to the L0(F )-bilinear form h, i : E × F → L0(F ) if hx, yi = 0 for all y ∈ F if, and only if, x = 0; and hx, yi = 0 for all x ∈ E if, and only if, x = 0 (see [24, Definition 3.19]). Then, if we consider the corresponding conditional vector spaces E, F and the conditional function h, i : E ⋊⋉ F → R, we obtain a conditional dual pair as in [24, Definition 5.6]. For a given random duality pair E, F (both with the countable concatena- F (E)), we can define the weak topologies. ++(F )} is a base for the weak tion property; for instance, Lϕ Namely, U := {x + UQ,ε ; Q ⊂ F finite , ε ∈ L0 F (E), Lϕ∗ 6 topology σ(E, F ). However, we have that U is not necessarily a stable family, and we cannot directly apply [14, Proposition 3.5] just as we have done for L0(F )-normed modules. Instead, we need to stabilize the topology by consid- ering the collection Ucc :=nX 1Ak Uk ; {Ak} ∈ Π(Ω, F ), Uk ∈ U for all k ∈ No , which is the basis for a finer topology; let us denote it by σcc(E, F ). In addition, Example 1.4 of [32], shows a situation in which σ(E, F ) is strictly coarser than σcc(E, F ), which proves that, in fact, U is not stable. By doing so, Ucc is already a stable family and [14, Proposition 3.5] applies. Then, the conditional weak topology 5 σ(E, F) is the collection of conditional sub- sets OA where O is σcc(E, F )-open and A ∈ F . In particular, since σcc(E, F ) is finer than σ(E, F ), we have that for any σ(E, F )-open subset O with the countable concatenation property, it holds that O is conditionally σ(E, F)-open. However, the converse is not true (as [32, Example 1.4] shows). It is also important to recall the notion of conditionally compact conditional set (see [14, Definition 3.24]). Suppose that K is a conditional subset of L0, and consider some conditional topology T on K. Then K is conditionally compact, if for every conditional family of conditionally open subsets {Oi} such that K ⊏ ⊔Oi, there exists a conditional subset [ik ; k ∈ N, 1 ≤ k ≤ n] with n ∈ N, such that K ⊏ ⊔ Oik. 1≤k≤n 3 The relation between Lϕ(E) and Lϕ F (E) Let us fix a Young function ϕ : [0, +∞) → [0, +∞], which will be the same in the remainder of this paper. Let us start by studying the relation between the classical Orlicz space Lϕ(E) F (E). Vogelpoth [36] proved the following product F (E) = L0(F )Lϕ(E). Then, analogously as was made in [25] for Lp- and the Orlicz-type module Lϕ structure Lϕ type modules, we can prove the following result: Proposition 1 The L0(F )-module Lϕ erty; moreover, F (E) has the countable concatenation prop- cc Lϕ(E) = Lϕ F (E). Proof. It is clear that Lϕ Now, since Lϕ(E) ⊂ Lϕ F (E) has the countable concatenation property. F (E), it follows that F (E). For the reverse inclusion, suppose that x ∈ Lϕ Lϕ(E) ⊂ Lϕ cc F (E). Since Lϕ F (E) = L0(F )Lϕ(E), we can put x = ηx0 with η ∈ L0(F ) and x0 ∈ Lϕ(E). Let us define Ak := 5This is not exactly how the conditional weak topologies was introduced in [14]; however, it is not difficult to prove that both approaches are equivalent. 7 1Ak ηx0 ≤ kx0 with x0 ∈ Lϕ(E), and Lϕ(E) is a solid subspace of L0(E), it fol- cc . (k − 1 ≤ η < k) for each k ∈ N. Then, x = Pk∈N 1Akηx0. Besides, since lows that xk := 1Ak ηx0 ∈ Lϕ(E). We conclude that x =Pk∈N 1Ak xk ∈ Lϕ(E) Apart from the Luxemburg norm, there are other relevant equivalent norms for Lϕ(E). For instance, we can consider the so-called Orlicz norm, which has a different formulation given by the so-called Amemiya norm (see [2]). Namely, for x ∈ Lϕ(E), the Ameya norm is given by kxkA ϕ := inf(cid:26) 1 r (1 + E [ϕ(rx)]) ; r > 0(cid:27) In this paper, we will be interested in the modular form of the Amemiya norm. Thereby, for x ∈ Lϕ F (E), we define kxFkA ϕ :=^(cid:26) 1 λ (1 + E [ϕ(λx)F ]) ; λ ∈ L0 ++(F )(cid:27) . (2) In the appendix of this paper, it is proved that the formula above defines a L0(F )-norm, which is equivalent to k · Fkϕ (see Proposition 25). The corre- sponding conditional norm will be referred to as conditional Amemiya norm. For the next result, we also need to recall the following spaces. For 1 ≤ p ≤ ∞ and for a given L0-normed module (E, k · k) Guo et al.[25] defined the following spaces Lp(E) := {x ∈ E ; kkxkkp < +∞} , for which xp := kkxkkp defines a norm. Then, we have the following result: Proposition 2 For every x ∈ Lϕ(E), it holds kkxFkA ϕ k1 ≤ kxkA ϕ . In particular, it follows that Lϕ(E) ⊂ L1(Lϕ F (E)). Proof. By definition of the Amemiya norm, we have that for each real number r > 0, kxFkA ϕ ≤ 1 r (1 + E[ϕ(rx)F ]). By taking expectations kkxFkA ϕ k1 = E[kxFkA ϕ ] ≤ 1 r (1 + E[ϕ(rx)]) ≤ +∞. Finally, by taking infimums on r ∈ R+ kkxFkA ϕ k1 ≤ kxkA ϕ < +∞. and the proof is complete. 8 Let us recall that a conditional sequence {xn} in a conditional normed space (E, k · k) conditionally converges to x ∈ E if for every r ∈ R++ there exists m ∈ N such that kx − xnk ≤ r for all n ≥ m. Also, {xn} is said to be conditionally Cauchy, if for every r ∈ R++, there exists m ∈ N such that kxp − xqk ≤ r for all p, q ≥ m (both notions are introduced in [14]). Then, we have the following result: Proposition 3 The conditional normed space (Lϕ, k · Fkϕ) is conditionally Banach6, i.e. every conditionally Cauchy sequence {xn} conditionally converges to some x ∈ Lϕ. Proof. In [14, Theorem 3.3.3], it is shown that the L0-normed module (Lϕ Fkϕ) is complete in the sense that every Cauchy net converges in Lϕ Now, let {xn} be a conditional Cauchy sequence in Lϕ F (E). Then, we can consider the stable family {xn}n∈L0(F ,N). We have that L0(F , N) is upward di- rected, and therefore {xn}n∈L0(F ,N) is a net indexed by L0(F , N). Furthermore, since {xn} is conditionally Cauchy, it follows that {xn}n∈L0(F ,N) is Cauchy. F (E) is complete, the net {xn}n∈L0(F ,N) converges to some x0 ∈ Lϕ Since Lϕ F (E). If follows that the conditional sequence {xn} conditionally converges to x0. F (E), k · F (E). 4 Duality in Orlicz-type modules Given a subset H ⊂ L0(E), the Kothe dual of H is given by H x :=(cid:8)y ∈ L0(E) ; xy ∈ L1(E), for all x ∈ H(cid:9) . A well-known result of the theory of Orlicz spaces (see [28, 37]) is the fol- lowing identity Lϕ∗ (E) = [Lϕ(E)]x , (3) where ϕ∗(y) := supx≥0(xy − ϕ(x)) is the conjugate Young function of ϕ. We will show that the equality above naturally extends to the modular case. For this purpose, suppose that H is a conditional subset of L0, we can consider a conditional version of the Kothe dual. Namely, Hx :=(cid:2)y ∈ L0 ; xy ∈ L1, for all x ∈ H(cid:3) . Then, we will prove that Lϕ∗ = [Lϕ]x . For a given y ∈ [Lϕ]x, then we can define an application µy : Lϕ F (E) → L0(F ), µy(x) := E[xyF ]. Clearly, µy is stable. Consequently, it generates a conditional function 6The notion on conditionally Banach normed space was also introduced in [14]. µy : Lϕ → R. 9 Proposition 4 Suppose that f : Lϕ F (E) → L0(F ) is a stable function, which is also monotone and convex, then the conditional function f : Lϕ → R is con- ditionally norm continuous and conditionally convex. Moreover, if f is linear, then f is conditionally linear. Proof. By adapting the proof of [36, Theorem 4.1.3]7 to the present setting, we can obtain that f : Lϕ → R is conditionally norm continuous. Let us show that f is conditionally convex. Indeed, suppose that a ∈ R with 0 ≤ a ≤ 1 and x, y ∈ Lϕ. Let us show that f(ax+(1−a)y) ≤ af(x)+(1−a)f(y). If a ∈ R the result is clear, due to the convexity of f . also is clear due to the stability and convexity of f . For arbitrary a, we proceed as follows. Let us pick some dense and countable set H ⊂ R, H = {h1, h2, ...} (for instance H = Q), and for n ∈ N, let A1 := (a ≤ h1 < a + 1 n ) and Ak := (a ≤ hk < a + 1 k−1Ai for k > 1, and put If a ∈ R cc n ) − ∨ i=1 an := 1 ∧Xk∈N 1Ak hk. For n ∈ L0(F , N) of the form Pi∈N 1Bini with {ni} ⊂ N and {Ai} ∈ Π(Ω, F ), we define an :=Pi∈N 1Biani. Then, {an}n∈N is a conditional sequence, which conditionally converges to cc cc a. Also, note that an ∈ H ⊂ R and 0 ≤ an ≤ 1. Then f(anx + (1 − an)y) ≤ anf(x) + (1 − an)f(y), for all n ∈ N. Since f is conditionally continuous, by taking conditional limits, we obtain f(ax + (1 − a)y) ≤ af(x) + (1 − a)f(y). If f is linear, by following a similar argument, we get that f is conditionally linear. Corollary 5 Let y ∈ [Lϕ]x, then the conditional function µy : Lϕ → R, µy(x) := E[xyF ], is conditionally norm continuous. Proof. We can suppose y ≥ 0; otherwise, we can put y = y+ − y− and argue on y+ and y−. Then, the application µy is monotone, convex and stable, and the result follows by Proposition 4. 7In the literature some issues regarding the countable concatenation property have been reported cf.[24, 38]. The proof of [36, Theorem 4.1.3] involves infinite countable concatena- tions; therefore, it seems that in the statement of [36, Theorem 4.1.3] should be added the extra-hypothesis on E of having the countable concatenation property. The module Lϕ F (E) has the countable concatenation; moreover, the proof of [36, Theorem 4.1.3] perfectly works for this particular case, and the results applies to this case as Proposition 4 asserts. 10 Let (E, k · k) be a conditional normed space. In [14] it is introduced the conditional topological dual E∗, which is defined at the conditional set whose conditional elements are conditionally linear continuous functions µ : E → R, and is endowed with the conditional norm kµk := sup [µ(x) ; kxk ≤ 1] . Consequently, in view of Corollary 5, for y ∈ [Lϕ]x, we have that µy ∈ [Lϕ]∗ and E[xyF ] ≤ kµykkxFkϕ, for all x ∈ Lϕ. We have the following result: Proposition 6 It holds Lϕ∗ = [Lϕ]x . Proof. Lϕ∗ Due to Proposition 1, we have Lϕ∗(E) (E) = [Lϕ(E)]x. Thus, it suffices to show that cc = Lϕ∗ F (E), also we know that [Lϕ(E)]xcc = [Lϕ F (E)]x F , where [Lϕ F (E)]x F :=(cid:8)y ∈ L0(E) ; xy ∈ L1 which is the generating set of [Lϕ]x. First, inspection shows that [Lϕ F (E)]x F (E) for all x ∈ Lϕ F (E)(cid:9) , Indeed, for x ∈ Lϕ Let us take y ∈ [Lϕ(E)]x, and let us show that xy ∈ L1 erty. Then, for the inclusion "⊂" it suffices to show that [Lϕ(E)]x ⊂ [Lϕ F has the countable concatenation prop- F (E)]x F . F (E). F (E), due to Proposition 1, we have that x = P 1Ak xk with {xk} ⊂ Lϕ(E) and {Ak} ∈ Π(Ω, F ). Therefore, xy = P 1Ak xky ∈ P 1Ak L1(E) ⊂ L1 P 1Ak rk > kµyk where rk are positive real number for each k ∈ N and {Ak} ∈ We claim that 1Ak y ∈ [Lϕ(E)]x. Indeed, for a given x ∈ Lϕ(E), we have that F . Let us pick some let us take y ∈ [Lϕ F (E)]x F (E) for all x ∈ Lϕ For the reverse inclusion, Π(Ω, F ). F (E). E[x1Ak yF ] = 1Ak E[x1Ak sgn(y)yF ] ≤ 1Ak kµykk1AkxFkA ϕ ≤ rkk1AkxFkA ϕ . By taking expectations E[x1Ak y] ≤ E[rkk1Ak xFkA ϕ ], and the right side of the inequality above is finite, in view of Proposition 2. Consequently, x1Ak y ∈ L1(E). Since x ∈ Lϕ(E) is arbitrary, we obtain that 1Ak yk ∈ [Lϕ(E)]x; hence, y ∈ [Lϕ(E)]xcc . In the following definition we recall some notions from the theory of Riesz spaces. 11 Definition 7 An linear application µ : Lϕ 1. is order bounded, if for every x ∈ Lϕ µ([−x, x]) ⊂ [−y, y]; F (E) → L0(F ): F (E), there exists y ∈ L0 +(F ) such that 2. is order continuous if, µ is order bounded, and for every downward directed subset D of Lϕ F (E) with V D = 0, it holds that Vx∈D µ(x) = 0. 3. is σ-order continuous if, µ is order bounded, and for every decreasing sequence {xn} in Lϕ F (E) with V xn = 0, it holds that V µ(xn) = 0. Let us introduce a conditional version of the classical notion of Riesz space theory of order continuity. Definition 8 A conditionally linear function µ : Lϕ → R is: 1. conditionally order bounded if µ is order bounded; 2. conditionally order continuous if it is conditionally order bounded and for any conditionally downward directed8 subset D of Lϕ with inf D = 0, it holds that inf x∈D µ(x) = 0. We denote by [Lϕ]∼ n the conditional set of conditionally order continuous and conditionally linear functions µ : Lϕ → R, which is referred to as the conditional order dual. The following result is an adaptation of [25, Lemma 2.16(2)] to the present setting. Lemma 9 [25] Let E be a conditional set, C a subset of E, and f : E → R a conditional function, then suphf(x) ; x ∈ C cci =_ {f (x) ; x ∈ C} Ω. F (E) → L0(F ) be a stable linear function, the fol- Proposition 10 Let µ : Lϕ lowing are equivalent: 1. µ is σ-order continuous; 2. µ is order continuous; 3. µ is conditionally order continuous. In this case, µ is conditionally linear and conditionally norm continuous, i.e. [Lϕ]∼ n ⊏ [Lϕ]∗. 8In [14] it is introduced the notion of direction. 12 Proof. 1 ⇔ 2 : It is known that L0(F ) is a Riesz space which is Dedekind complete and order separable (see [37] for all this terminology relative to Riesz spaces). Also it is clear that Lϕ F (E) is a Riesz space. Then, according to [37, Theorem 84.4(i)], we find that µ is order continuous if, and only if, µ is σ-order continuous. Indeed, if µ : Lϕ F (E) → L0(F ) is an order continuous linear stable application. From the theory of Riesz spaces, it has a Jordan Decomposition µ = µ+ − µ− (see [37, Theorem 84.4(i)]) where µ+, µ− are linear functions with 2 ⇒ 3: First, let us show that µ is conditionally linear. µ+(x) :=W {µ(y) ; 0 ≤ y ≤ x} for x ≥ 0, (4) and for arbitrary x, µ+(x) = µ+(x+) − µ−(x−) and µ−(x) = (−µ)+(x). By using this definition we have that µ+ and µ− are also stable. Then, − is conditionally linear and + − µ due to Proposition 4, we obtain that µ = µ conditionally norm continuous. Finally, the result is clear, as every conditionally downward directed set is defined from a downward directed set (with the countable concatenation prop- erty). 3 ⇒ 1: Suppose that {xn} is a decreasing sequence such that V xn = 0. As commented in the section of preliminaries, we can extend this sequence to a stable family {xn}n∈L0(F ,N). Then, the conditional set D := [xn ; n ∈ N] in conditionally downward directed and satisfies inf D = 0, due to Lemma 9 (and using that N = L0(F , N)). Again, by Lemma 9 we have cc ^{µ(xn) ; n ∈ N}Ω =^{µ(xn) ; n ∈ L0(F , N)}Ω = inf Therefore, we have that V{µ(xn) ; n ∈ N} = 0. Finally, suppose that µ ∈ [Lϕ]∼ n . Notices that, in 2 ⇒ 3, we have shown that µ is not only conditionally linear, but also conditionally norm continuous; that is, µ ∈ [Lϕ]∗. µ(x) = 0. x∈D Proposition 11 We have the following [Lϕ]∼ n =hµy ; y ∈ Lϕ∗i . Proof. Let µy with y ∈ Lϕ∗ . If xn ց 0 a.s., then by monotone convergence, we have that µy(xn) = E[xnyF ] → 0 a.s. This means that µy is σ-order continuous and, due to Proposition 10, we have that µy ∈ [Lϕ]∼ n . For the reverse inclusion, suppose that µ ∈ [Lϕ]∼ n . Then, in view of Propo- sition 10, µy ∈ [Lϕ]∗. Now, for each n ∈ N, let An := (n − 1 ≤ kµk < n). For every n ∈ N and x ∈ Lϕ(E), we have that E[1Anµ(x)] ≤ nE[kxFkA ϕ ], 13 and the right hand of the inequality above is finite due to Proposition 2. Thus, for each n ∈ N, we can define an application µn : Lϕ(E) → R, µn(x) := E[1Anµ(x)], which is linear. Moreover, by dominated convergence is easy to show that µn is σ-order continuous (as µ is σ-order continuous). From the theory of Orlicz spaces, we have that, for each n ∈ N, there exists yn ∈ Lϕ∗ (E) such that µn(x) := E[xyn], for all x ∈ Lϕ(E). We claim that, for each n ∈ N, 1An µ(x) = E[ynxF ], for all x ∈ Lϕ(E). Indeed, for fixed x ∈ Lϕ(E) and any A ∈ F , E[1A1Anµ(x)] = E[µn(1Ax)] = E[1Axyn], and this implies the equality above. F (E). Then, for x ∈ Lϕ(E), we have that µ(x) = cc , we can easily extend the equality to x ∈ E[yxF ]. Let y :=Pn∈N 1Anyn ∈ Lϕ∗ Finally, since Lϕ F (E) = Lϕ(E) Lϕ F (E), obtaining that µ = µy. The following result characterizes those linear applications µ : Lϕ(E) → L0(F ) that can be represented by conditional expectations. Corollary 12 Let µ : Lϕ(E) → L0(F ) be linear. Then, the following are equiv- alent: 1. µ is σ-order continuous and µ(1Ax) = 1Aµ(1Ax) for all A ∈ F ; 2. there exists y ∈ Lϕ∗ F (E) such that µ(x) = E[xyF ] for all x ∈ Lϕ(E). Proof. 2 ⇒ 1 is clear. 1 ⇒ 2 : We define ¯µ : Lϕ F (E) → L0(F ), where for x ∈ Lϕ F (E), which is of the form x = Pk∈N 1Ak xk with {xk} ⊂ Lϕ(E) and {Ak} ∈ Π(Ω, F ), we put ¯µ(x) = Pk∈N 1Ak µ(xk). By using that µ(1Ax) = 1Aµ(1Ax) for all A ∈ F , we see that ¯µ is well-defined. Due to Proposition 11, we can pick y ∈ Lϕ∗ (E) such that µ(x) = ¯µ(x) = µy(x) for all x ∈ Lϕ(E). 5 Application to conditional risk measures of Orlicz spaces The notion of conditional convex risk measure was independently introduced in [8] and [13]. This definition naturally extends to Orlicz spaces as follows: 14 Definition 13 Let ψ : [0, +∞) → [0, +∞] be a Young function with ψ ≤ ϕ. A conditional convex risk measure (of Orlicz spaces) is a function ρ : Lϕ(E) → Lψ(F ) with the following properties for x, y ∈ Lϕ(E): 1. monotonicity: i.e. x ≤ y implies ρ(x) ≥ ρ(y); 2. L0(F )-convexity: i.e. for all λ ∈ L0(F ), with 0 ≤ λ ≤ 1, ρ(λx+(1−λ)y) ≤ λρ(x) + (1 − λ)ρ(y); 3. Lϕ(F )-cash invariance: i.e. for all λ ∈ Lϕ(F ), ρ(x + λ) = ρ(x) − λ. Further, given a conditional convex risk measure ρ : Lϕ(E) → Lψ(F ) we define the following version of the Fenchel conjugate ρ∗(y) :=W{E[xyF ] − ρ(x) ; x ∈ Lϕ(E)}, Likewise, the notion of dynamic convex risk measure, naturally extends to for y ∈ Lϕ∗ (E). (5) Orlicz spaces: Definition 14 Let {Ft}t∈T be a filtration of E with T = N0 ∩ [0, T ] and T ∈ N ∪ {+∞}, and {ϕt}t∈T be a sequence of Young functions with ϕ ≥ ϕt ≥ ϕt+1. A dynamic convex risk measure (of Orlicz spaces) is a sequence {ρt}t∈T, where ρt : Lϕ(E) → Lϕt(Ft) is a conditional convex risk measure for each t ∈ T. We say that a function f : Lϕ(E) → Lψ(F ) has the local property if 1Af (1Ax) = ρ(1Ax), for all A ∈ F . Proposition 15 Let f : Lϕ(E) → Lψ(F ) be a L0(F )-convex function, then f has the local property. Proof. A similar argument is followed in the proof of [16, Theorem 3.2]. Let A ∈ F . Then, for x ∈ Lϕ(E) f (1Ax) = f (1Ax + 1Ac0) ≤ 1Af (x) + 1Acf (0) = 1Af (1Ax + 1c Ax) + 1Acf (0) ≤ ≤ 1A(1Af (1Ax) + 1c Af (x)) + 1Acf (0) = 1Af (1Ax) + 1Acf (0). Finally, by multiplying by 1A, the inequalities above become equalities and 1Af (1Ax) = 1Af (x). 15 Proposition 16 Let f : Lϕ(E) → Lψ(F ) be a function with the local property, then there is a unique stable extension ¯f : Lϕ F (E) → L0(F ). Moreover, if f is monotone, then ¯f is monotone; if f is L0(F )-convex, then ¯f is L0(F )-convex; and if f is Lϕ(F )-cash invariant, then ¯f is L0(F )-cash invariant (i.e. ¯f (x + λ) = ¯f (x) − λ for x ∈ Lϕ F (E) and λ ∈ L0(F )). Further, for all y ∈ Lϕ∗ (E): _ {E[xyF ] − f (x) ; x ∈ Lϕ(E)} =_(cid:8)E[xyF ] − ¯f (x) ; x ∈ Lϕ the version of the Fenchel conjugate f ∗ introduced in (5) agrees on i.e. (E) with the modular version of the Fenchel conjugate ¯f ∗ (see [16]). F (E)(cid:9) , (6) Lϕ∗ Proof. Suppose that f : Lϕ(E) → Lψ(F ) has the local property. Due to Proposition F (E) of the form x = . Then, given x ∈ Lϕ F (E) = Lϕ(E) 1, we have that Lϕ cc P 1Ak xk with {xk} ⊂ Lϕ(E) and {Ak} ∈ Π(Ω, F ), let 1Ak f (xk). ¯f (x) :=Xk∈N By using the local property of f , it is not difficult to show that ¯f is well- defined, is stable, and is not possible another stable extension. If f is monotone, it is clear that ¯f is monotone too. Also it is clear that the L0(F )-convexity of f implies the L0(F )-convexity of ¯f . By using that Lϕ(F ) = L0(F ) is easy to show that, if f is Lϕ(F )-cash cc invariant, then ¯f is L0(F )-cash invariant. Finally, for a given y ∈ Lϕ∗ and that the application x 7→ E[xyF ] − ¯f (x) is stable, we obtain (6) by means of Lemma 9. (E), by using that Lϕ F (E) = Lϕ(E) cc Remark 17 Notice that, given a conditional convex risk measure ρ, since ¯ρ and ¯ρ∗ are stable, we can define conditional functions ¯ρ : Lϕ → R and ∗ : Lϕ∗ ¯ρ → R. In particular, ¯ρ is conditionally monotone, conditionally convex and condi- tionally cash-invariant (i.e. ¯ρ(x + r) = ¯ρ(x) − r for x ∈ Lϕ and r ∈ R). Also, ¯ρ ∗ is precisely the conditional Fenchel conjugate of ¯ρ, i.e. ¯ρ ∗(y) := sup [E[xyF ] − ¯ρ(x) ; x ∈ Lϕ] . The following result shows that, if we replace L0(F )-convexity and Lϕ(F )- cash invariance by convexity, cash-invariance (considering only real numbers) and the local property in Definition 13, then we obtain an equivalent definition of conditional convex risk measure. 16 Proposition 18 Let f : Lϕ(E) → Lψ(F ) be a monotone function. The follow- ing properties are equivalent: 1. f is convex and has the local property; 2. f is L0(F )-convex. In this case, if in addition f is cash-invariant, then f is also Lϕ(F )-cash invariant. Proof. 2 ⇒ 1: It is obvious that f is convex. The local property is given by Proposition 15. 1 ⇒ 2: Since f has the local property, in virtue of Proposition 16, it can be extended to a function ¯f : Lϕ F (E) → L0(F ). By Proposition 4, we know that ¯f is conditionally norm continuous and conditionally convex. Consequently, ¯f is L0(F )-convex; hence, f is L0(F )-convex. Suppose now that f is cash- invariant, and let us show that it is also Lϕ(F )-cash invariant. Indeed, given x ∈ Lϕ(E) and a ∈ Lϕ(F ), let us consider a countable and dense subset H of R. Suppose H = {h1, h2, ...}. For n ∈ N, define B1 := (a ≤ h1 < a + 1/n) and Bk := (a ≤ hk < a + 1/n) − ∨ i=1 doing so, we obtain a conditional sequence {an}n∈N. Then, by using again that ¯f is conditionally continuous, together with the fact that f is cash-invariant and stable, we have that k−1Bi for k > 1. Define an :=Pk∈N hk1Bk . By ¯f(x + a) = lim ¯f(x + an) = ¯f(x) − lim an = ¯f(x) − a. n n We conclude that f (x + a) = ¯f (x + a) = ¯f (x) − a = f (x) − a. In what follows we will study the representation of a conditional convex risk measure ρ : Lϕ(E) → Lψ(F ). We have the following result: Theorem 19 Let ρ : Lϕ(E) → Lψ(F ) be a conditional convex risk measure, then the following properties are equivalent: 1. ¯ρ is σ(Lϕ F (E), Lϕ∗ 2. For every x ∈ Lϕ F (E) F (E))-lower semicontinuous; ¯ρ(x) =_nE[xyF ] − ¯ρ∗(y) ; y ∈ Lϕ∗ F (E), E[yF ] = −1, y ≤ 0o ; 3. For every x ∈ Lϕ(E) ρ(x) =_nE[xyF ] − ρ∗(y) ; y ∈ Lϕ∗ (E), E[yF ] = −1, y ≤ 0o . Proof. 2 ⇒ 1 is clear. 2 ⇔ 3: Since Lϕ∗ F (E) = Lϕ∗(E) cc , we have that ny ∈ Lϕ∗ F (E) ; y ≤ 0, E[yF ] = −1o = {y ∈ Lϕ∗(E) ; y ≤ 0, E[yF ] = −1} cc . 17 Besides, since the map y 7→ E[xyF ] − ¯ρ∗(y) is stable and (according to (E), due to Lemma 9 we obtain that Proposition 16) ρ∗(y) = ¯ρ∗(y) for y ∈ Lϕ∗ for any x ∈ Lϕ(E) WnE[xyF ] − ¯ρ∗(y) ; y ∈ Lϕ∗ =W(cid:8)E[xyF ] − ρ∗(y) ; y ∈ Lϕ∗ F (E), E[yF ] = −1, y ≤ 0o = (E), E[yF ] = −1, y ≤ 0(cid:9) . Then, 2 ⇒ 3 is clear from (7). As for 3 ⇒ 2, we see that the map (7) x 7→_nE[xyF ] − ¯ρ∗(y) ; y ∈ Lϕ∗ F (E), E[yF ] = −1, y ≤ 0o , is stable and, due to (7), agrees with ρ on Lϕ(E). Then, due to Proposition 16, we obtain the result. 1 ⇒ 2: It is easy to show that hLϕ F (E)i is a random dual pair in the F (E), Lϕ∗ F (E)9. By using the modular version of the classical Fenchel-Moreau dual sense of [24]. Therefore, due to [24, Theorem 3.22], we have(cid:16)Lϕ Lϕ∗ representation theorem [16, Theorem 3.8] we obtain F (E), σ(Lϕ F (E), Lϕ∗ F (E))(cid:17)∗ = ¯ρ(x) :=WnE[xyF ] − ¯ρ∗(y) ; y ∈ Lϕ∗ Let us show that dom(¯ρ∗) ⊂ ny ∈ Lϕ∗ F (E)o , F (E) ; y ≤ 0, E[yF ] = −1o. Indeed, ++(F ), due to the monotonicity of ¯ρ, we F (E), then for a given λ ∈ L0 for x ∈ Lϕ F (E). let z ∈ Lϕ∗ have ¯ρ∗(z) ≥ E[λ1(z≥0)zF ] − ¯ρ(λ1(z≥0)) ≥ λE[z+F ] − ¯ρ(0). Since λ is arbitrary, we have that the left side of the inequality above is finite only if z ≤ 0. Further, due to the L0(F )-cash invariance of ¯ρ, for any λ ∈ L0(F ) we have the following ¯ρ∗(z) ≥ E[λzF ] − ¯ρ(λ) = λ (E[zF ] + 1) − ¯ρ(0), being λ arbitrary, the left hand of the inequality above is finite only if E[zF ] = −1. Regarding the attainability of the representation provided above, we have the following result: Theorem 20 Let ρ : Lϕ(E) → Lψ(F ) be a conditional convex risk measure, such that ¯ρ is σ(Lϕ F (E))-lower semicontinuous, then the following are equivalent: F (E), Lϕ∗ 9Notice that '∗' denotes here the modular topological conjugate, instead of the conditional topological dual. As commented in the section of preliminaries, this topology can possibly be non-stable. 18 1. ρ has the Lebesgue property, i.e y ∈ Lϕ(E), xn ≤ y for all n ∈ N, xn → x a.s. implies ρ(xn) → ρ(x) a.s.; 2. ¯ρ has the Lebesgue property, i.e y ∈ Lϕ F (E), xn ≤ y for all n ∈ N, xn → x a.s. implies ¯ρ(xn) → ¯ρ(x) a.s.; 3. For every x ∈ Lϕ(E) there exists y ∈ Lϕ∗ (E) with y ≤ 0 and E[yF ] = −1 such that ρ(x) = E[xyF ] − ρ∗(y); 4. For every x ∈ Lϕ F (E) there exists y ∈ Lϕ∗ F (E) with y ≤ 0 and E[yF ] = −1 such that ¯ρ(x) = E[xyF ] − ¯ρ∗(y); 5. For each c ∈ R++, Vc :=hy ∈ Lϕ∗ ; ¯ρ ∗(y) ≤ ci 6. For each c ∈ R, c > 0, Vc :=hy ∈ Lϕ∗ ; ¯ρ ∗(y) ≤ ci is conditionally σ(Lϕ∗ , Lϕ)-compact; is conditionally σ(Lϕ∗ , Lϕ)-compact. Before proving the statement above, we need some preliminary results. In particular, we have the following lemma of scalarization of conditional risk measures: Lemma 21 Let ρ : Lϕ(E) → Lψ(F ) be a conditional convex risk measure. Then the function ρ0 : Lϕ(E) → R, ρ0(x) := E[ρ(x)], is a (static) convex risk measure. Moreover, suppose that ρ∗ 0(y) = sup {E[xy] − ρ0(x) ; x ∈ Lϕ(E)} , for y ∈ Lϕ∗ (E) is the Fenchel conjugate of ρ0, then ρ∗ 0(y) = E[ρ∗(y)] for all y ∈ Lϕ∗ (E). In addition, if ¯ρ is σ(Lϕ F (E), Lϕ∗ F (E))-lower semicontinuous, then ρ0 is σ(Lϕ(E), Lϕ∗ (E))- lower semicontinuous. Proof. Inspection shows that ρ0 is a convex risk measure. Let us show that ρ∗ 0(y) = E[ρ∗(y)]. Indeed, for a given y ∈ Lϕ∗ is upward directed, because the function x 7→ E[xyF ] − ρ(x) has the local prop- erty. (E), the set My := {E[xyF ] − ρ(x) ; x ∈ Lϕ(E)} 19 Therefore, there exists a sequence {xn} ⊂ Lϕ(E), so that E[xnyF ]−ρ(xn) ր 0(y). W My = ρ∗ By monotone convergence, we have E[ρ∗(y)] = E[lim n (E[xnyF ] − ρ(xn))] = lim n (E[xny] − ρ0(xn)) ≤ ρ∗ 0(y). Also, for every x ∈ Lϕ(E), we have E[ρ∗(y)] ≥ E[E[xyF ] − ρ(x)] = E[xy] − ρ0(x), hence E[ρ∗(y)] ≥ sup{E[xy] − ρ0(x) ; x ∈ Lϕ(E)} = ρ∗ 0(y). We conclude that ρ∗ 0(y) = E[ρ∗(y)]. Theorem 19, we have that Finally, let us suppose that ¯ρ is σ(Lϕ F (E), Lϕ∗ F (E))-lower semicontinuous. By ρ(x) =W(cid:8)E[xyF ] − ρ∗(y) ; y ∈ Lϕ∗ Now, for a fixed x ∈ Lϕ(E), we have that the set (E), E[yF ] = −1, y ≤ 0(cid:9) for all x ∈ Lϕ(E). Mx :=nE[xyF ] − ρ∗(y) ; y ∈ Lϕ∗ (E), E[yF ] = −1, y ≤ 0o is upward directed (the function y 7→ E[xyF ]−ρ∗(y) has the local property). (E) with E[ynF ] = −1 and yn ≤ 0 Thus, we can find a sequence {yn} ⊂ Lϕ∗ such that E[xynF ] − ρ∗(yn) րW Mx = ρ(x). By monotone convergence, and by using that E[ρ∗(yn)] = ρ∗ 0(yn), we have ρ0(x) = E[ρ(x)] = E[lim n (E[xynF ] − ρ∗(yn))] = lim n (E[xyn] − ρ∗ 0(yn)). Also, for every y ∈ Lϕ∗ (E), we have ρ0(x) = E[ρ(x)] ≥ E[E[xyF ] − ρ∗(y)] = E[xy] − ρ∗ 0(y), hence ρ0(x) ≥ sup{E[xy] − ρ∗ We conclude that 0(y) ; y ∈ Lϕ∗ (E)}. ρ0(x) = sup{E[xy] − ρ∗ 0(y) ; y ∈ Lϕ∗ (E))-lower semicontinuous. (E)}. This implies that ρ0 is σ(Lϕ(E), Lϕ∗ A similar result can be found in [27, Lemma 2]. Lemma 22 Let {yn} be a sequence in L0(F ) such that lim sup yn = y. Then, there exists a sequence n1 < n2 < ... in L0(F , N), such that {ynk } converges a.s. to y. Notice that if C is a subset of L0(E) with the countable concatenation prop- erty, then the solid hull sol(C) :=(cid:8)x ∈ L0(E) ; there exists y ∈ C, x ≤ y(cid:9) , also has the countable concatenation property. Then it generates a conditional set which is denoted by sol(C), and will be referred to as conditional solid hull of C. The following lemma provides a test for conditional σ(Lϕ∗ , Lϕ)-compactness. 20 Lemma 23 Let C ⊂ Lϕ∗ the following are equivalent: F (E) with the countable concatenation property. Then E[unzF ] → 0 a.s. whenever un ց 0 a.s. in Lϕ F (E); 1. Wz∈C 2. sol(C) is conditionally σ(Lϕ∗ , Lϕ)-relatively compact. Then, in this case, C is conditionally σ(Lϕ∗ , Lϕ)-relatively compact. Proof. 1 ⇒ 2: Put K := sol(C). First, we will prove that K σ([Lϕ]∗,Lϕ) K i.e. every µ ∈ K . For that, it suffices to show that K σ([Lϕ]∗,Lϕ) satisfies that µ is σ-order continuous. σ([Lϕ]∗,Lϕ) σ(Lϕ∗ ,Lϕ) = actually lies on Lϕ∗ , Indeed, for a given µ ∈ K and r ∈ R++, let us fix u ∈ Lϕ. Then, there is y ∈ K such that (µy − µ)(u) ≤ r. Since K := sol(C), we can find z ∈ C so that y ≤ z. Thus, σ([Lϕ]∗,Lϕ) µ(u) ≤ (µ − µy)(u) + µy(u) ≤ r + _z∈C E[uzF ]. Since r is arbitrary, we have that E[uzF ], for all u ∈ Lϕ F (E). µ(u) ≤ Wz∈C Now, let {un} be a decreasing sequence in Lϕ F (E) such thatV un = 0. Then, the inequality above implies that V µ(un) = 0. Also, it is not difficult to show that every L0(F )-linear continuous function is also order bounded. This means that µ : Lϕ F (F ) → L0(F ) is σ-order continuous. The second step is to prove that K is conditionally norm bounded (or equiv- n x ց In particular, since Indeed, given x ∈ Lϕ, we have that 1 alently L0(F )-norm bounded). 0 a.s. Then, by hypothesis, ∨ z∈C n xzF ] → 0 a.s. E[ 1 E[ 1 n xzF ] = 1 ∨ z∈C inspection shows that ∨ z∈C n ∨ z∈C E[xzF ], necessarily ∨ z∈C E[xzF ] is finite. Moreover, E[xzF ] = ∨ z∈K E[xzF ]. This means that there exists rx ∈ R such that E[zxF ] ≤ rx for all z ∈ K. By the conditional version of the Uniform Boundedness Principle [39, The- orem 3.4], we obtain that K is conditionally norm bounded. Finally, the conditional Banach-Alaoglu Theorem ([14, Theorem 5.10]) yields the result. 2 ⇒ 1: Now, let un ց 0 a.s. in Lϕ F (E). For fixed r ∈ R++, for each n ∈ N, we can find zn ∈ C such that ∨ z∈C E[unzF ] ≤ E[unznF ] + r. 21 Then, we can consider the conditional sequence {zn} in K := sol(C). Since K is conditionally relatively compact, due to Remark 3.20 and Proposition 3.25 of [14], we can find some conditional σ(Lϕ∗ of {zn}. Besides, we have that z ≥ 010. , Lϕ)-cluster point z ∈ Lϕ∗ Also, we can find n1 ∈ N such that 0 ≤ E[zunF ] ≤ r for all n ≥ n1. Let us pick n2 ≥ n1 such that E[(zn2 − z)un1F ] ≤ r. Then, for n ≥ n2, we have sup z∈C E[unzF ] ≤ sup z∈C E[un2 zF ] ≤ E[un2 zn2F ] + r ≤ ≤ E[un1zn2 F ] + r = E[(zn2 − z)un1 F ] + E[zun1F ] + r ≤ 3r. This yields that limn sup z∈C E[unzF ] = 0. It is not difficult to show that this implies that ∨ z∈C E[unzF ] → 0 a.s. (considering now n ∈ N). The following result is an adaptation to the present setting of [33, Lemma 2.3]. Lemma 24 Suppose that ¯ρ is σ(Lϕ 0, then for any β ∈ R, x ∈ Lϕ F (E) and y ∈ Lϕ F (E), Lϕ∗ F (E))-lower semicontinuous and ¯ρ(0) = E[xyF ] − ¯ρ∗(y) ≥ −β implies F (E) ¯ρ∗(y) ≤ 2β + 2 ¯ρ(−2x). Proof. Due to Theorem 19, since ¯ρ is lower semicontinuous, we have 0 = ¯ρ(0) =_n− ¯ρ∗(y) ; y ∈ Lϕ∗ = −^n¯ρ∗(y) ; y ∈ Lϕ∗ F (E), y ≤ 0, E[yF ] = −1o = F (E), y ≤ 0, E[yF ] = −1o . Then, given ε ∈ L0 ++(F ), there exists yε ∈ Lϕ∗ F (E) with yε ≤ 0 and E[yεF ] = −1 such that 0 ≤ ¯ρ∗(yε) ≤ ε. Thus, E[xyF ] ≤ E(cid:20)−2x y + yε 2 F(cid:21) ≤ ¯ρ(−2x) + ¯ρ∗(cid:18) y + yε 2 (cid:19) ≤ ≤ ¯ρ(−2x) + 1 2 ¯ρ∗(y) + 1 2 ¯ρ∗(yε) = ¯ρ(−2x) + 1 2 ¯ρ∗(y) + ε. Since ε is arbitrary, we obtain 1 2 Then, if E[xyF ] − ¯ρ∗(y) ≥ −β, we have that E[xyF ] ≤ ¯ρ(−2x) + ¯ρ∗(y). ¯ρ∗(y) ≤ E[xyF ] + β ≤ ¯ρ(−2x) + 1 2 ¯ρ∗(y) + β. 10Indeed, for each k ∈ N we can find nk ∈ L0(N, F ) such that E[(znk − z)1(z≤0)F ] ≤ 1 k , this means that E[znk 1(z≤0)F ] → E[z−F ] a.s. This implies that E[z−F ] ≥ 0, and therefore z ≥ 0. 22 From where we finally obtain ¯ρ∗(y) ≤ 2β + 2 ¯ρ(−2x). Proof of Theorem 20. 1 ⇒ 3 : For a given x ∈ Lϕ(E), let us define ρ0(x) := E[ρ(x)] as in Lemma 21, which is a convex risk measure. Since ρ(x) ≥ E[xyF ]− ρ∗(y) for all y ∈ Lϕ∗ (E) with E[yF ] = −1 and y ≤ 0, it suffices to show that E[ρ(x)] = E[E[xyF ] − ρ∗(y)], or equivalently that ρ0(x) = E[xy] − ρ∗ 0(y). Since ρ0 is σ(Lϕ(E), Lϕ∗ (E))-lower semicontinuous, by [33, Theorem 1.1] it suffices to prove that ρ0 has the Lebesgue property. Indeed, suppose that {xn} satisfies xn ≤ y for some y ∈ Lϕ(E) and xn → x a.s. Since ρ has the Lebesgue property, it holds that ρ(xn) converges to ρ(x) a.s. Also, by monotonicity, we have that ρ(xn) ≤ ρ(y) ∨ ρ(−y). Then, by dominated convergence we have that lim n ρ0(xn) = E[lim n ρ(xn)] = E[ρ(x)] = ρ0(x). 3 ⇒ 1 : Suppose that {xn} satisfies xn ≤ y for some y ∈ Lϕ(E) and xn → x a.s. Then, by dominated convergence we have limn E[xnzF ] = E[xzF ] for all z ∈ Lϕ∗ (E) with E[zF ] = −1 and z ≤ 0. Then E[xnzF ] − ρ∗(z) ; z ∈ Lϕ∗ ρ(x) =_nlim n ≤ lim inf n _nE[xnzF ] − ρ∗(z) ; z ∈ Lϕ∗ = lim inf ρ(xn). n (E), E[zF ] = −1, z ≤ 0o ≤ (E), E[zF ] = −1, z ≤ 0o = It suffices to show that ρ(x) ≥ lim sup ρ(xn). Indeed, suppose that ρ(x) < n lim sup ρ(xn) on some A ∈ F with P(A) > 0. We can suppose A = Ω w.l.g. n Again, we consider the convex risk measure ρ0(x) := E[ρ(x)], for which we have ρ0(x) < E[lim sup ρ(xn)]. By using Lemma 22, we can construct a sequence {zn}, with zn ≤ y for all n ∈ N, such that zn → x a.s. and ρ(zn) → lim supn ρ(xn). Then, by dominated convergence lim n ρ0(zn) = E[lim n ρ(zn)] = E[lim sup n ρ(xn)] > ρ0(x). 23 However, by assumption, for each z ∈ Lϕ(E), we have ρ0(z) = E[E[zyF ] − ρ∗(z)] = E[zy] − ρ∗ 0(y), for some y ∈ Lϕ∗ (E))-lower semicontinuous. Thus, by [33, Theorem 1.1], we have that ρ0 has necessarily the Lebesgue prop- erty; hence, limn ρ0(zn) = ρ0(x). This is a contradiction. (E) and also ρ0 is σ(Lϕ(E), Lϕ∗ 3 ⇒ 4: Given x ∈ Lϕ F (E), we know that there are {xk} ⊂ Lϕ(E) and {Ak} ∈ Π(Ω, F ) such that x = P 1Ak xk. Then, for each k ∈ N there is (E) such that yk ∈ Lϕ∗ ρ(xk) = E[xkykF ] − ρ∗(yk), for all k ∈ N. we have Let y =P 1Ak yk. Then, since ¯ρ is stable and ρ∗(y) = ¯ρ∗(y) for y ∈ Lϕ∗ ¯ρ(x) =X 1Ak ρ(xk) =X 1Ak (E[xkykF ] − ρ∗(yk)) = E[xyF ] − ρ∗(y). 4 ⇒ 3: It is clear, as ρ∗(y) = ¯ρ∗(y) for y ∈ Lϕ∗ (E) (see last part of Proposi- (E), tion 16). 2 ⇒ 1 is clear. 1 ⇒ 2: Let y ∈ Lϕ F (E), xn ≤ y such that xn → x a.s. Let us take ¯ρ(x). Then, for each k ∈ N, 1Ak xn ≤ 1Ak yk and xn → x a.s. Thus, ρ(1Ak xn) → {yk} ⊂ Lϕ(E) and {Ak} ∈ Π(Ω, F ) such the y =P 1Ak yk. ρ(1Ak x) a.s. Then, ¯ρ(xn) =P 1Ak ρ(1Ak xn) converges a.s. to P 1Akρ(1Ak x) = 1 ⇒ 5: We have already shown (see 1 ⇒ 3) that, under this assumption, ρ0 (E))-lower semicontinuous and, for any x ∈ Lϕ(E), there exists (E) such that ρ0(x) = E[xy] − ρ0(y). Then, by [33, Theorem 1.1], we is σ(Lϕ(E), Lϕ∗ y ∈ Lϕ∗ obtain that, for any positive c ∈ R, the set V 0 is σ(Lϕ∗ Let us define (E), Lϕ(E))-compact. c :=(cid:8)y ∈ Lϕ∗ (E) ; ρ∗(y) ≤ c(cid:9) fc(x) :=WnE[xzF ] ; z ∈ Lϕ∗ F (E), ¯ρ∗(z) ≤ co , f0,c(x) := sup(cid:8)E[xz] ; z ∈ Lϕ∗ 0(z) ≤ c(cid:9) , Given x ∈ Lϕ(E), the set Hx := (cid:8)E[xzF ] ; z ∈ Lϕ∗ (E), ¯ρ∗(z) ≤ c(cid:9) is up- such that E[xznF ] րW Hx. Besides, due to Lemma 9, we haveW Hx = fc(x). ward directed; hence, one can find a sequence {zn} ⊂ Lϕ∗ 0(zn) for each n ∈ N, by dominated convergence we Since c ≥ E[¯ρ∗(zn)] = ρ∗ have that E[fc(x)] = E[lim E[xznF ]] = lim E[xzn] ≤ f0,c(x). (E) with ¯ρ∗(zn) ≤ c for x ∈ Lϕ(E). for x ∈ Lϕ F (E); (E), ρ∗ Now, let xn ց 0 a.s. F (E), there is a partition {Ak} ∈ Π(Ω, F ), such that 1Ak x1 ∈ Lϕ(E) for all k ∈ N. Since Lϕ(E) is solid and 0 ≤ xn ≤ x1, we have that 1Ak xn ∈ Lϕ(E) for all F (E). First note that, since Lϕ(E) in Lϕ cc = Lϕ 24 k, n ∈ N. By arguing on each Ak, we can suppose that Ak = Ω, and thus {xn} ⊂ Lϕ(E). Since C := V0,c is σ(Lϕ∗ (E), Lϕ(E))-compact, due to [33, Lemma 2.1], we obtain that f0,c(xn) → 0 a.s. Therefore, we obtain 0 ≤ lim n E[fc(xn)] ≤ lim n f0,c(xn) = 0. Since fc(xn) ≥ fc(xn+1) > 0, necessarily fc(xn) → 0 a.s. (This happens on each Ak; therefore, on Ω). Due to Lemma 23, we conclude that Vc is conditionally relatively σ(Lϕ∗ , Lϕ)- compact. Besides, since ¯ρ is lower semicontinuous, we have that Vc is σ(Lϕ∗ F (E), Lϕ closed. Since Vc has the countable concatenation property, by the comments in section of preliminaries about the weak topologies, we have that Vc is condi- tionally closed; in turn, we also obtain the conditional compactness. cc F (E))- 5 ⇒ 6: For arbitrary c ∈ L0 ++(F ), we can find r ∈ R such that c ≤ r. Suppose that r =P 1Ak rk with {Ak} ∈ Π(Ω, F ) and {rk} ⊂ R. Then, we have that Vrk Ak is conditionally compact. We derive that Vc ⊏ Vr =P Vrk Ak is conditionally compact too. 6 ⇒ 4 : We can suppose w.l.g. that ρ(0) = 0. Due to the conditionally ∗(z) is ∗, the conditional function z 7→ E[xzF ] − ¯ρ , Lϕ)-upper semicontinuous. Then, the conditional set lower semicontinuity of ¯ρ conditionally σ(Lϕ∗ Mx := [z ∈ Lϕ∗ ; E[xzF ] − ¯ρ ∗(z) ≥ ¯ρ(x) − 1] is conditionally σ(Lϕ∗ , Lϕ)-closed and, applying Lemma 24 for β = 1 − ¯ρ(x), we obtain that Mx is conditionally contained in V2−2 ¯ρ(x)+2 ¯ρ(−2x). Therefore, Mx is conditionally σ(Lϕ∗ , Lϕ)-compact. Finally, it is not difficult to show that every conditionally upper semicontin- uous function attains its conditional maximum on a conditionally compact set (for the sake of completeness we have included the proof in the appendix, see Proposition 26), obtaining the result. (cid:4) References [1] B. Acciaio, and I. Penner. Dynamic risk measures. Advanced Mathematical Methods for Finance. Springer Berlin Heidelberg, 2011. 1-34. [2] C. D. Aliprantis and K.C. Border: Infinite Dimensional Analysis. A Hitch- hiker's Guide, third ed., Springer, 2006. [3] C. D. Aliprantis, and O. Burkinshaw: Locally solid Riesz spaces with appli- cations to economics. No. 105. American Mathematical Soc., 2003. [4] P. Artzner, F. Delbaen, J.-M. Eber, and D. Heath: Coherent measures of risk. Math. Finance, 9(3):203 -- 228, 1999. 25 [5] P. Artzner, F. Delbaen, J.-M. Eber, David Heath, and Hyejin Ku. Coherent multiperiod risk adjusted values and Bellman's principle. Ann. Oper. Res., 152:5 -- 22, 2007. [6] J. L. Bell. Set theory: Boolean-valued models and independence proofs. Oxford Logic Guides. Clarendon Press, 2005. [7] S. Biagini, and M. Frittelli. A unified framework for utility maximization problems: an Orlicz space approach. The Annals of Applied Probability 18.3 (2008): 929-966. [8] J. Bion-Nadal: Conditional risk measures and robust representation of con- vex conditional risk measures. Preprint, 2004. [9] P. Cheridito and T. Li: Dual characterization of properties of risk measures on Orlicz hearts. Math. Financ. Econ., 2(1):29 -- 55, 2008. [10] P. J. Cohen: The independence of the continuum hypothesis. Proceed- ings of the National Academy of Sciences of the United States of America, 50:1143 -- 1148, 1963 [11] F. Delbaen. Coherent risk measures. Cattedra Galileiana. [Galileo Chair]. Scuola Normale Superiore, Classe di Scienze, Pisa, 2000. [12] F. Delbaen, and K. Owari. On convex functions on the duals of ∆2-Orlicz spaces. arXiv preprint:1611.06218 (2016). [13] K. Detlefsen, and G. Scandolo: Conditional dynamic convex risk measures. Finance Stoch. 9(4), 539 -- 561 (2005). [14] S. Drapeau, A. Jamneshan, M. Karliczek, and M. Kupper: The algebra of conditional sets, and the concepts of conditional topology and compactness. J. Math. Anal. Appl., 437 (1) (2016), 561 -- 589. [15] S. Drapeau, A. Jamneshan, and M. Kupper: Vector duality via conditional extension of dual pairs. arXiv preprint arXiv:1608.08709 (2016). [16] D. Filipovi´c, M. Kupper, and N. Vogelpoth: Separation and duality in locally L0-convex modules. Journal of Functional Analysis 256.12 (2009): 3996-4029. [17] D. Filipovi´c, M. Kupper, and N. Vogelpoth: Approaches to conditional risk. SIAM Journal on Financial Mathematics 3.1 (2012): 402-432. [18] H. Follmer and A. Schied. Convex measures of risk and trading constraints. Finance Stoch., 6(4):429 -- 447, 2002. [19] H. Follmer and Alexander Schied: Stochastic finance, volume 27 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, extended edition, 2004. An introduction in discrete time 26 [20] Hans Follmer and Irina Penner: Convex risk measures and the dynamics of their penalty functions. Statist. Decisions, 24(1):61 -- 96, 2006. [21] M. Frittelli: and Preferences. bardo -- Accademia di Scienze e Lettere (2007): 179-199. No arbitrage Instituto Lom- [22] M. Frittelli and E. R. Gianin. Putting order in risk measures. Journal of Banking and Finance, 26(7):1473 -- 1486, 2002. [23] T. Guo: Relations between some basic results derived from two kinds of topologies for a random locally convex module, J. Funct. Anal. 258 (2010) 3024-3047. [24] T. Guo, S. Zhao, and X. Zeng: On random convex analysis -- the ana- lytic foundation of the module approach to conditional risk measures, arXiv preprint:1210.1848 (2012). [25] T. Guo, S. Zhao, and X. Zeng: The relations among the three kinds of conditional risk measures. Science China Mathematics 57.8 (2014): 1753- 1764. [26] T. Thomas: Set theory. Springer Science & Business Media, 2013. [27] Y. Kabanov, Ch. Stricker, A teacher's note on no-arbitrage criteria, S´eminaire de probabilit´es de Strasbourg 35 (2001) 149 -- 152. [28] W. A. Luxemburg: Banach Function Spaces, Delft, 1955. [29] M. Nowak: Order continuous linear functionals on non-locally convex Orlicz spaces. Comment. Math. Univ. Carolinae 33 (1992): 465-475. [30] T.M. Offwood: No free lunch and risk measures on Orlicz spaces, Diss. University of the Witwatersrand, South Africa, 2012. [31] J. Orihuela, M. Ru´ız Gal´an, Lebesgue property for convex risk measures on Orlicz spaces, Math. Financ. Econ. 6(1) (2012) 15-35. [32] J. Orihuela, and J. M. Zapata: On conditional Lebesgue property for con- ditional risk measures. arXiv preprint:1509.07081 (2015). [33] K. Owari: On the Lebesgue property of monotone convex functions. Math- ematics and Financial Economics 8.2 (2014): 159-167. [34] K. Owari: Maximum Lebesgue extension of monotone convex functions. Journal of Functional Analysis 266.6 (2014): 3572-3611. [35] M. M. Rao, and Z. D. Ren: Theory of Orlicz spaces. 1991. [36] N. Vogelpoth: L0-convex Analysis and Conditional Risk Measures. PhD thesis, Universitat Wien, 2009. 27 [37] A.C. Zaanen: Riesz spaces II, North-Holland Publ. Co., Amsterdam, New York, Oxford, 1983. [38] J. M. Zapata: On the characterization of locally L0-convex topologies in- duced by a family of L0-seminorms, J. Convex Anal. 24(2) (2017) - to appear. [39] J. M. Zapata: Versions of Eberlein-Smulian and Amir-Lindenstrauss theo- rems in the framework of conditional sets. Applicable Analysis and Discrete Mathematics 10(2) (2016): 231 -- 261. A Appendix Proposition 25 The Amemiya formula defines a L0(F )-norm such that kxFkϕ ≤ kxFkA ϕ ≤ 2kxFkϕ. (8) Proof. Let us first show the inequality (8). For given λ ∈ L0 ++(F ) with E(cid:2)ϕ(cid:0)xλ−1(cid:1) F(cid:3) ≤ 1, we have that ϕ ≤ λ(cid:0)1 + E(cid:2)ϕ(cid:0)xλ−1(cid:1) F(cid:3)(cid:1) ≤ 2λ. kxFkA Then, by taking infimums in λ we obtain kxFkA ϕ ≤ 2kxFkϕ. On the other hand, let us fix λ ∈ L0 ++(F ) and let us take If so, using the convexity of ϕ, we obtain that λ := λ(cid:0)1 + E(cid:2)ϕ(cid:0)xλ−1(cid:1) F(cid:3)(cid:1) . λ (cid:19) F(cid:21) = E ϕ   F  ≤  1 + Ehϕ(cid:16) x λ (cid:17) Fi x λ E(cid:20)ϕ(cid:18) x We derive that kxFkϕ ≤ λ Ehϕ(cid:16) x 1 + Ehϕ(cid:16) x λ (cid:17) Fi λ (cid:17) Fi ≤ 1 and by taking infimums in λ, we conclude kxFkϕ ≤ kxFkA ϕ . Finally, let us show that k · FkA (8), it suffices to show that k · FkA inequality and is L0(F )-homogeneous. ϕ is a L0(F )-norms. In view of the inequality F (E) → L0(F ) satisfies the triangle ϕ : Lϕ Given x1, x2 ∈ Lϕ F (E) and ε ∈ L0 ++(F ), for some λ1, λ2 ∈ L0 ++(F ) we have λi(cid:0)1 + E(cid:2)ϕ(cid:0)xiλi −1(cid:1) F(cid:3)(cid:1) ≤ kxiFkA ϕ + ε 2 , for i = 1, 2. 28 Let λ := λ1λ2 λ1+λ2 . By the convexity of ϕ, we have kx1 + x2FkA λ1 + λ2 ϕ ≤ λ(cid:0)1 + E(cid:2)ϕ(cid:0)x1 + x2λ−1(cid:1) F(cid:3)(cid:1) = −1x2(cid:12)(cid:12)(cid:12)(cid:12) −1x1 + λ1 + λ2 λ2 (cid:19) F(cid:21)(cid:19) ≤ −1(cid:1) F(cid:3)(cid:1) ≤ kx1FkA −1(cid:1) F(cid:3)(cid:1)+λ2(cid:0)1 + E(cid:2)ϕ(cid:0)x2λ2 λ1 λ2 λ1 ++(F ) is arbitrary, we obtain the triangle inequality. = λ1λ2 λ1 + λ2 (cid:18)1 + E(cid:20)ϕ(cid:18)(cid:12)(cid:12)(cid:12)(cid:12) ≤ λ1(cid:0)1 + E(cid:2)ϕ(cid:0)x1λ1 Since ε ∈ L0 Finally, let µ ∈ L0(F ). Then, ϕ +kx2FkA ϕ +ε. kµxFkA ϕ =^(cid:26)λ(cid:18)1 + E(cid:20)ϕ(cid:18) µx = µ^( λ µ 1 + E"ϕ x(cid:18) λ λ (cid:19) F(cid:21)(cid:19) ; λ ∈ L0 µ(cid:19)−1! F#! ; λ ∈ L0 ++(F )(cid:27) = ++(F )) = µkxFkA ϕ Proposition 26 Suppose that (E, T ) is a conditional topological space, f : E → R is a conditionally upper semicontinuous function, and K is a conditionally f(y) ∈ R, then there exists x ∈ K such that f(x) := compact subset of E. If sup y∈K f(y). sup y∈K Proof. Let r := sup y∈K f(y). For each n ∈ K, we define Cn :=(cid:20)x ∈ K ; f(x) ≥ r − 1 n(cid:21) . Clearly Cn is on 1. Besides, since f is conditionally upper semicontinuous, we obtain that Cn is conditionally closed in K. Also, for every m ∈ N, we have that ⊓n≤mCn = Cm. Due to [14, Proposition 3.25], we have that ⊓n∈NCn is on 1. Thus, there exists x ∈ K such that x ∈ Cn for all n. Necessarily, f(x) = r. 29
1701.03745
1
1701
2017-01-13T17:32:32
Conjugates of integral functionals on continuous functions
[ "math.FA", "math.OC" ]
This article gives necessary and sufficient conditions for the dual representation of Rockafellar in (Integrals which are convex functionals. II, Pacific J. Math., 39:439--469, 1971) for integral functionals on the space of continuous functions.
math.FA
math
Conjugates of integral functionals on continuous functions Ari-Pekka Perkkio∗ August 14, 2018 Abstract This article gives necessary and sufficient conditions for the dual rep- resentation of Rockafellar in [Integrals which are convex functionals. II, Pacific J. Math., 39:439-469, 1971] for integral functionals on the space of continuous functions. Keywords. semicontinuity integral functional; convex conjugate; set-valued mapping; inner 1 Introduction Let T be a σ-compact locally compact Hausdorff (σ-lcH) with the topology τ , the Borel-σ-algebra B(T ) and a strictly positive regular measure µ. The space of Rd-valued continuous functions C := C(T ; Rd), that vanish at infinity, is a Banach space with respect to the supremum norm. The Banach dual of C can be identified with the space M of Rd-valued signed regular measures with finite total variation on T via the bilinear form hy, θi :=Z ydθ; see [Fol99, Section 7]. This article studies conjugates of convex integral functionals on C. Such functionals appear in numerous applications in optimal control, variational prob- lems over functions of bounded variation, optimal transport, financial mathe- matics, and in extremal value theory. Applications to singular stochastic control and finance are given in [PP15]. Our main theorems sharpen those of [Roc71] and [Per14] by giving necessary and sufficient conditions for the conjugacy of integral functionals on C and functionals on the space of measures with integral representation involving the ∗Department of Mathematics, Ludwig-Maximilians-Universitat Munchen, Theresienstr. 39, 80333 Munchen, Germany. The author is grateful to the Vaisala Foundation for the financial support. 1 recession function of the conjugate integrand. For closely related results, we refer to [AB88, BV88] and the references therein. Our characterizations are new in the literature and they complete the idea initiated in [Roc71] on expressing the singular parts of the conjugates on the space of measures in terms of recession functions. The paper starts by sharpening [Roc71, Theorem 6] on the integral represen- tation of the support function of continuous selections of a convex-valued map- ping. We show that inner semicontinuity of the mapping is not only sufficient but also a necessary condition for the validity of the representation. Section 3 gives the main results on integral functionals. The conditions appearing in the main results are analyzed further in Section 4. 2 Inner semicontinuity and continuous selections Let S be a set-valued mapping from T to Rd and denote the set of its continuous selections by C(S) := {y ∈ C yt ∈ St ∀t ∈ T }. This section sharpens [Roc71, Theorem 6] on the integral representation of the support function σC(S)(θ) := sup hy, θi y∈C(S) of C(S). Theorem 6 of [Roc71] states that if S is "inner semicontinuous" and T is compact, then σC(S)(θ) =Z σS(dθ/dθ)dθ :=Z σSt ((dθ/dθ)t)dθt, where θ denotes the total variation of θ and σSt is the normal integrand defined pointwise by σSt (x) := sup y∈St x · y. Theorem 1 below extends the result to general σ-lcH T and shows that in- ner semicontinuity of S is a necessary condition for the validity of the integral representation. The proof of sufficiency is analogous of that of Rockafellar, but we repeat it here for completeness since we do not assume compactness of T . To the best of our knowledge, the necessity of the inner semicontinuity has not been recorded in the literature before. We denote the neighborhoods of t ∈ T by Ht and the open neighborhoods x. The mapping S is inner semicontinuous (isc) at t if St ⊆ of x ∈ Rd by Ho lim inf St, where lim inf St := {x ∈ Rd S−1(A) ∈ Ht ∀A ∈ Ho x}. When S is isc at every t, S is inner semicontinuous. By [Mic56, Proposition 2.1], S is isc if and only if the preimage S−1(O) := {t ∈ T St ∩ O 6= ∅} of every open 2 set O is open. Inner semicontinuity is studied in detail in [RW98, Section 5] in the case where T is Euclidean. We often use Michael's continuous selection theorem [Mic56, Theorem 3.1"'] that is valid for a Lindelof perfectly normal T1- space T . We refer to [Dug66] for the definitions, and note here that Hausdorff spaces are T1 and σ-compactness of T implies that every subset of T is Lindelof, so T is a Lindelof perfectly normal T1-space by [Dug66, Theorem XI.6.4 and Problem VIII.6.7]. Given a set A, NA(y) denotes the set of its normals at y, that is, θ ∈ NA(y) if y ∈ A and hθ, y′ − yi ≤ 0 ∀y′ ∈ A, while NA(y) := ∅ for y /∈ A. For a set-valued mapping S : T ⇒ Rd and a function y : T → Rd, we denote the set-valued mapping t 7→ NSt(yt) by NS(y). The mapping t 7→ cl St is called the image closure of S. Theorem 1. Assume that S is a convex-valued mapping from T to Rd such that C(S) 6= ∅. Then σC(S)(y) =Z σS(dθ/dθ)dθ if and only if S is inner semicontinuous. In this case θ ∈ NC(S)(y) if and only if dθ/dθ ∈ NS(y) θ-a.e. Proof. We first prove sufficiency. The mapping defined by Γt := cl{yt y ∈ C(S)} is contained in S, and, since a mapping is isc if and only if its image closure is isc, Γ is isc by part 7 of Theorem 2 below. Assuming that S is not inner semicontinuous, there is t such that Γt is strictly smaller than St. By a separat- ing hyperplane theorem, there is x with supv∈Γt x · v < supv∈St x · v. Defining θ ∈ M so that θ is a Dirac measure at t and dθ/dθ = x, we can write this as sup y∈C(S) hy, θi <Z σS(dθ/dθ)dθ, which is a contradiction. Now we turn to necessity. By the Fenchel inequality, hy, θi ≤Z σS(dθ/dθ)dθ (1) for every y ∈ C(S), so it suffices to show sup y∈C(S) hy, θi ≥Z σS(dθ/dθ)dθ. 3 Denoting L∞(S) = {w ∈ L∞(T, θ) wt ∈ St ∀t}, we have, by [RW98, Theorem 14.60], sup w∈L∞(θ;S)Z wdθ =Z σS(dθ/dθ)dθ. Let ¯y ∈ C(D), α <Z σS(dθ/dθ)dθ and w ∈ L∞(S) be such thatR wdθ > α. By regularity of θ and Lusin's theorem [Fol99, Theorem 7.10], there is an open ¯O ⊂ T such thatR ¯O(¯y+wt)dθ < ǫ/2, ¯OC is compact and w is continuous relative to ¯OC . The mapping Γt =(wt St if t ∈ ¯OC if t ∈ O is isc convex closed nonempty-valued so that, by [Mic56, Theorem 3.1"'], there is a yw ∈ C with yw t ∈ St for all t. Since θ is regular, t dθ < ǫ/2. Since ¯OC is compact and T is locally compact, we may choose O precompact, by [Dug66, Theorem XI.6.2]. there is an open O ⊃ ¯OC such that R O\ ¯OC yw t = wt on ¯OC and yw Since O and ¯O form an open cover of T and since T is normal, there is, by [Mun00, Theorem 36.1], a continuous partition of unity (α, ¯α) subordinate to ( O, ¯O). Defining y := αyw + ¯α¯y, we have y ∈ C(S) and Z ydθ ≥Z ¯OC wdθ −Z O\ ¯OC αywdθ −Z ¯O ¯α¯ydθ ≥Z α − ǫ, which finishes the proof of necessity, since α <R σS(dθ/dθ)dθ was arbitrary. As to the normal vectors, θ ∈ NC(D)(y) if and only if σC(D)(y) = hθ, yi which, by the Fenchel inequality (1), is equivalent to σS(dθ/dθ)dθ =Z y · (dθ/dθ)dθ which is equivalent to condition in the statement, since we also have the point- wise Fenchel inequalities σSt(x) ≥ yt · x ∀x ∈ Rd that are satisfied as equalities if and only if x ∈ NSt(yt). The next theorem lists some useful criteria for the preservation of inner semicontinuity in operations. Given a set A, rint A denotes its relative interior. Theorem 2. If S, S1 and S2 are isc mappings, then 1. AS is isc for any continuous A : Rn → Rm, 4 2. co S is isc, 3. S1 × S2 is isc, 4. S1 + S2 is isc, 5. t 7→ S1 t ∩ S2 t is isc provided G := {(t, x) x ∈ S2 t } is open. 6. t 7→ S1 t ∩S2 t is isc provided T is Euclidean and S1 and S2 are convex-valued with 0 ∈ int(S1 t − S2 t ) for all t, 7. Arbitrary pointwise unions of isc mappings are isc. Proof. 1 follows from (AS)−1(O) = S−1(A−1(O)), since then preimages of open sets are open. 3 can be derived directly from the definition. Choosing A(v1, v2) = v1 + v2 and S = S1 × S2, 4 follows from 1 and 3. To prove 5, let A ⊂ Rd be open, t ∈ (S1 ∩ S2)−1(A) and x ∈ S1 t ∩ A. The set G ∩ (T × A) is open and contains (t, x), so there is A′ ∈ Ho x and O ∈ Ho t such that A′ ⊂ A and O × A′ ⊂ G. Then O ∩ (S1)−1(A′) ⊆ (S1 ∩ S2)−1(A), so (S1 ∩ S2)−1(A) ∈ Ht. Thus (S1 ∩ S2)−1(A) is open. To prove 6, let tν → t. By [RW98, Theorem 4.32], t ∩ S2 lim inf(S1 tν ∩ S2 tν ) ⊇ lim inf S1 tν ∩ lim inf S2 tν . Taking the intersection over all sequences tν → t proves the claim, by [RW98, Exercise 5.6]. To prove 2, notice that co S is an union of all the mappings of the form It Pm i=1 αiS for Pm remains to prove 7. For any mappings (Sα), (Sα Sα)−1(A) = S((Sα)−1(A)). Thus arbitrary unions of isc mappings is isc, since preimages of open sets are open. i=1 αi = 1, αi ≥ 0, so the result follows from 1,4 and 7. 3 Integral functionals of continuous functions This section studies conjugates of functionals of the form Ih + δC(D) : C → R, where h is a convex normal B(T )-integrand on Rd, Ih(v) :=Z ht(vt)dµt and Dt := cl dom ht. Here dom g := {x g(x) < ∞} is the domain of g and cl denotes the closure operation. Recall that h is a convex normal B(T )- integrand on Rd if its epigraphical mapping t 7→ epi ht := {(v, α) ∈ Rd × R ht(v) ≤ α} is closed convex-valued and measurable. A set-valued mapping S is measurable if preimages of open sets are measurable. We refer the reader to 5 [RW98, Chapter 14] for a general study of measurable set-valued mappings and normal integrands. Note that, when ht(v) = δSt(v), we have Ih + δC(D) = δC(D) so we are in the setting of Section 2. On the other hand, when dom Ih ⊆ C(D), we have Ih + δC(D) = Ih. Sufficient conditions for the above inclusion can be found in [Roc71], [Per14] and Section 4 below. Recall that the conjugate of an extended real-valued function F on C is the extended real-valued lower semicontinuous (lsc) convex function on M defined by F ∗(θ) := sup y∈C {hy, θi − F (y)}. Rockafellar [Roc71] and more recently Perkkio [Per14] gave conditions for the validity of the integral representation where the functional Jh∗ : M → R is defined by (Ih)∗ = Jh∗ , Jh∗ (θ) =Z h∗ t ((dθa/dµ)t)dµt +Z (h∗ t )∞((dθs/dθs)t)dθst, where θa and θs denote the absolutely continuous and the singular part, respec- t )∞ denotes the recession function of h∗ tively, of θ with respect to µ, and (h∗ t . Recall that the recession function of a proper lsc convex function g is given by g∞(x) = sup α>0 g(αx + ¯x) − g(¯x) α , where ¯x ∈ dom g is arbitrary; see [Roc70, Chapter 8]. We often use the identity g∞ = σdom g∗ (2) valid for proper lsc convex functions on locally convex topological vector spaces; see [Roc66, Corollary 3D] or [Roc70, Theorem 13.3] for the finite dimensional case. The identity (2) makes it clear that the properties of Jh∗ are related to those of t 7→ Dt. The following lemma is the first part of the proof of [Per14, Theorem 3]. It shows that the functional Ih + δC(D) arises naturally as the conjugate of Jh∗ . Lemma 3. If dom Jh∗ 6= ∅, then J ∗ h∗ = Ih + δC(D). 6 Proof. We have J ∗ h∗ (y) = sup = sup θ∈M(cid:26)Z ydθ − Jh∗ (θ)(cid:27) θ ′∈L1(µ;Rd)(cid:26)Z y · θ′dµ − Ih∗ (θ′)(cid:27) θ∈M,w∈L1(θs;Rd)(cid:26)Z y · wdθs −Z (h∗)∞(w)dθs(cid:27) sup + =(Ih(y) +∞ if yt ∈ cl dom ht ∀t, otherwise. Here the second equality follows from the positive homogeneity of (h∗ t )∞, and the third follows by first applying [RW98, Theorem 14.60] on the second and the third line, where one uses (2) and then takes the supremum over all purely atomic finite measures which are singular with respect to µ. The next lemma combines [Roc71, Theorem 4] with the condition C(D) = cl(dom Ih ∩ C(D)). This condition is part of the sufficient conditions in the theorem below, which will be reduced to the situation in the lemma. We denote the closed ball centered at v with radius r by Br(v) and the interior of a set A by int A. When v = 0, we write simply Br. Lemma 4. Assume that D is isc, C(D) = cl(dom Ih ∩ C(D)) and that y ∈ dom Ih and r > 0 are such that t 7→ ht(yt + v) is finite and belongs to L1 whenever v ∈ Br. Then (Ih + δC(D))∗ = Jh∗ . Proof. By [Roc71, Theorem 4], y ∈ int dom Ih ∩ C(D) and I ∗ h(θ) = min θ ′ {Ih∗ (dθ′/dµ) + σdom Ih (θ − θ′) θ′ ≪ µ}. Since int dom Ih ∩ C(D) 6= ∅, we may apply Fenchel duality [Roc74, Theorem 20] to the sum Ih + δC(D). Thus (Ih + δC(D))∗(θ) = min θ ′′ = min {min θ ′ min θ ′′ θ ′ {Ih∗(dθ′/dµ) + σdom Ih (θ − θ′ − θ′′) θ′ ≪ µ} + σC(D)(θ′′)} {Ih∗ (dθ′/dµ) + σdom Ih (θ − θ′ − θ′′) + σC(D)(θ′′) θ′ ≪ µ} = min θ ′ = min θ ′ = min θ ′ {Ih∗(dθ′/dµ) + σdom Ih∩C(D)(θ − θ′) θ′ ≪ µ} {Ih∗(dθ′/dµ) + JσD (θ − θ′) θ′ ≪ µ} {Z h∗(dθ′/dµ)dµ +Z (h∗)∞(d(θ − θ′)/dµ)dµ} +Z (h∗)∞(d(θs)/dθs)dθs. 7 Here the third equality follows from another application of Fenchel duality, the fourth from the assumption C(D) = cl(dom Ih ∩ C(D)) and from Theorem 1. By [Roc70, Corollary 8.5.1], the last minimum is attained at dθ′/dµ = dθ/dµ, so the last expression equals Jh∗ (θ). The following theorem and its corollary are the main results of the paper. The theorem gives necessary and sufficient conditions for the conjugacy of Ih + δC(D) and Jh∗ . The corollary sharpens [Roc71, Theorem 5] by giving necessary and sufficient conditions for the conjugacy between Ih and Jh∗. Given a function g, ∂g(y) denotes its subgradients at y, that is, θ ∈ ∂g(y) if g(y) + hy′ − y, θi ≤ g(y′) ∀y′. Given a normal integrand h and y ∈ C, we denote the set-valued mapping t 7→ ∂(ht)(yt) by ∂h(y) and the mapping t 7→ NDt (yt) by ∂sh(y). Theorem 5. Assuming Ih + δC(D) and Jh∗ are proper, they are conjugates of each other if and only if dom h is isc and C(D) = cl(dom Ih ∩ C(D)), and then θ ∈ ∂(Ih + δC(D))(y) if and only if dθa/dµ ∈ ∂h(y) µ-a.e., dθs/dθs ∈ ∂sh(y) θs-a.e. (3) Proof. To prove the necessity, assume that Ih + δC(D) and Jh∗ are conjugates of each other. By (2), σcl dom(Ih+δC(D)) = J ∞ h∗ , where, by the monotone convergence theorem, J ∞ h∗ = J(h∗)∞ . Here, by (2) again, (h∗)∞ = σD. By Lemma 3, J ∗ σD = δC(D), so cl(dom Ih ∩ C(D)) = C(D). In particular, σC(D) = JσD , so D is inner semicontinuous, by Theorem 1, which shows the necessity. To prove the sufficiency, it suffices, by Lemma 3 and the biconjugate theorem [Roc74, Theorem 5], to show that Jh∗ is lsc. For ǫ > 0, we define ǫht(x) := min x′∈Bǫ ht(x + x′) and ǫDt := cl dom ǫht. By [RW98, Proposition 14.47], ǫh is a convex normal integrand, and a direct calculation gives ǫh∗ t (v) = h∗ t (v) + ǫv. When ǫ ց 0, both Iǫh + δC(Dǫ) ր Ih + δC(D) and Jǫh∗ ց Jh∗ pointwise, so, Jh∗ is lsc by Lemma 9 in the appendix provided that (Iǫh + δC(Dǫ))∗ = J(ǫh)∗ . To have this, we will apply Lemma 4 to ǫh. We have ǫD = cl dom ht + Bǫ, so ǫD is isc, by Theorem 2. To show that C(ǫD) = cl(dom Iǫh ∩ C(ǫD)), let y ∈ C(ǫD). For any ν = 1, 2, . . . , the mapping t 7→ Dt ∩ int B ǫ(1+2−ν−1)(yt) is convex nonempty-valued and isc, by Theorem 2. By [Mic56, Theorem 3.1"'] and the assumption C(D) = cl(dom Ih ∩ C(D)), there is yν ∈ dom Ih ∩ C(D) such that ky − yνk ≤ ǫ(1 + 2−ν). Defining yν = (1−ǫ2−ν)y +ǫ2−νyν, we have yν ∈ C(ǫD) and yν → y in C. Since kyν − yνk < ǫ, 8 we get Iǫh(yν) ≤ Ih(yν) and yν ∈ dom Iǫh. Thus y ∈ cl(dom Iǫh ∩ C(ǫD). Finally, since Ih and Jh∗ are proper, we have, for any ¯y ∈ dom Ih ∩ C(D), that t 7→ ǫht(¯yt + v) is finite and integrable whenever v ∈ Bǫ. Thus ǫh satisfies the assumptions of Lemma 4 which finishes the proof the sufficiency. As to the subdifferential formulas, we have, for any y ∈ dom(Ih + δcl dom h) and θ ∈ M , the Fenchel inequalities h(y) + h∗(dθa/dµ) ≥ y · (dθa/dµ) µ-a.e., (h∗)∞(dθs/dθs) ≥ y · (dθs/dθs) θs-a.e. (the latter holds by (2)), so θ ∈ ∂Ih(v) if and only if Ih(y)+Jh∗(θ) = hy, θi which is equivalent to having the Fenchel inequalities satisfied as equalities which in turn is equivalent to the given pointwise subdifferential conditions. Remark 1. The properness of Ih + δC(D) and Jh∗ is equivalent to the existence of ¯x ∈ L1(Rd), ¯y ∈ C(D) and α ∈ L1 such that ht(v) ≥ v · ¯xt − αt, h∗ t (x) ≥ ¯yt · x − αt. Indeed, for y ∈ dom Ih ∩ C(D) and θ ∈ Jh∗, we may choose ¯y = y, ¯x = θ/dµ, and α := max{h∗(¯x), h(¯v)}. Choosing y = ¯y and θ by dθ/dµ = ¯x and θs = 0, gives the other direction. Just like in Theorem 1, the necessity of inner semicontinuity in Theorem 5 has not been explicitly stated in the literature before. The domain condition will be analyzed more in detail in the next section. The following corollary extends [Roc71, Theorem 5] and [Per14, Theorem 2] in the case of lcH σ-compact T . It builds on Theorem 5 by giving necessary and sufficient conditions for the equality Ih + δC(D) = Ih. Corollary 6. Assuming Ih and Jh∗ are proper, they are conjugates of each other if and only if dom h is isc and C(D) = cl dom Ih, and then θ ∈ ∂Ih(y) if and only if dθa/dµ ∈ ∂h(y) µ-a.e., dθs/dθs ∈ ∂sh(y) θs-a.e. Proof. Since C(D) = cl dom Ih implies that Ih = Ih + δC(D) and that C(D) = cl(dom Ih ∩ C(D)), we have sufficiency by Theorem 5. Assuming that Ih and Jh∗ are conjugates of each other, Lemma 3 implies that Ih + δC(D) = Ih, so dom Ih ∩ C(D) = dom Ih, and the sufficiency follows from Theorem 5. The related main theorems by Rockafellar and Perkkio in [Roc71] and [Per14] give sufficient conditions for Corollary 6. Rockafellar's assumption involved the notion of "full lower semicontinuity" of set-valued mappings whereas Perkkio formulated the assumption in terms of "outer µ-regularity". We return to these concepts in the next section. 9 4 The domain conditions in the main results In this section we will analyze more in detail the conditions C(D) = cl(dom Ih ∩ C(D)), C(D) = cl dom Ih appearing in Theorem 5 and Corollary 6. The former is characterized in Theo- rem 7 below whereas for the latter we give sufficient conditions in terms of the domain mapping D. The condition in the next result, which extends the condition in [Roc71, The- orem 5], can be found in [BV88, Proposition 6] for metrizable σ-lcH T . Here we generalize to σ-lcH T and give the converse statement for inner semicontinuous domains. Theorem 7. Let Ih + δC(D) and Jh∗ be proper. If, for every t ∈ T and every v ∈ rint Dt, there is O ∈ Ht and y ∈ C(D) such that yt = v andRO h(y)dµ < ∞, then C(D) = cl(dom Ih ∩ C(D)). The converse holds if D is isc. Proof. Let y0 ∈ dom Ih ∩C(D), ǫ > 0 and y ∈ C(D). There is a compact K ⊂ T such that ky0k, kyk ≤ ǫ/2 outside K. By compactness and the assumptions, there exist ti ∈ T , i = 1, . . . , n for which there are yi ∈ C(D) and Oti ∈ Hti such i=1 Oti . Let α0, αi, i = 1, . . . , n be continuous partition of unity subordinate to {K C, Ot1 , . . . Otn}. By h(yi)dµ < ∞ and K ⊂ Sn yk ≤ ǫ. Since ǫ was arbitrary, we see that y ∈ cl(dom Ih ∩ C(D)). i=0 αiyi ∈ (dom Ih ∩ C(D)), and, by construction, kPi=1 αiyti − that yi − y ≤ ǫ on Oti , ROti convexity, Pn Assume now that D is isc and that C(D) = cl(dom Ih ∩ C(D)). Let t ∈ T and v ∈ rint Dt. There exists vi ∈ rint Dt, i = 0, . . . , d, such that v ∈ co{vi i = 0, . . . d}. Let ǫ > 0 be small enough so that v ∈ co{¯vi i = 0, . . . d} whenever, for all i, ¯vi ∈ B vi,ǫ. By Michael selection theorem [Mic56, Theorem 3.1"'], there is, for every i, a continuous selection of D taking a value vi at t, so, there exists vi,ǫ. Let αi ∈ (0, 1) be convex weights such yi ∈ dom Ih ∩ C(D) such that yi i=0 αiyi ∈ C(D) ∩ dom Ih which proves i=0 αiyi t ∈ B that Pd the claim. t = v. By convexity, Pd We now turn to the condition C(D) = cl dom Ih in Corollary 6. We say that a function y : T → Rd is a µ-selection of D if yt ∈ Dt outside a µ-null set. Clearly, every y ∈ dom Ih is a continuous µ-selection of D. Thus, if every continuous µ-selection of D is a selection of D, we have dom Ih ⊆ C(D) and thus, Ih + δC(D) = Ih. If, in adddition, C(D) = cl(dom Ih ∩ C(D)), we thus get C(D) = cl dom Ih. Following [Per14], we say that a set-valued mapping S is outer µ-regular if µ-liminf St ⊆ cl St ∀t ∈ T, 10 where µ-liminf St := {x ∈ Rd ∃O ∈ Ht : µ(S−1(A) ∩ O) = µ(O) ∀A ∈ Ho x}. By [Per14, Theorem 1], every continuous µ-selection of an outer µ-regular map- ping is a selection of its image closure. The necessity of outer µ-regularity is given in [Per14, Theorem 2] for solid convex-valued mappings when the under- lying topological space is Lindelof perfectly normal T1. Here we give sufficient conditions for continuous µ-selections to be selections in terms of familiar con- tinuity notions of set-valued mappings. Given a topology ¯τ on T , we denote the ¯τ neighborhood-system of t by ¯Ht. The mapping S is ¯τ -outer semicontinuous (osc) at t if ¯τ - lim sup St ⊆ St, where ¯τ - lim sup St := {x ∈ Rd S−1(A) ∈ ¯H# t ∀A ∈ Ho x}, and ¯H# := {B ⊂ T B ∩ O 6= ∅ ∀O ∈ ¯Ht}. We denote ¯τ ⊇ τ if ¯τ is finer than t τ , and we say that ¯τ supports µ locally at t if there is O ∈ ¯Ht such that every O-relatively ¯τ -open set has a subset with positive measure. The following definition of µ-fullness is inspired by the notion of full lower semicontinuity introduced in [Roc71] for solid convex-valued mappings. Exam- ple 1 below shows that such mappings are µ-full. Definition 1. A set-valued mapping S is ¯τ -full locally at t if there is O ∈ ¯Ht such that (¯τ - lim inf(¯τ - lim sup S))t′ ⊆ cl St′ ∀ t′ ∈ O. The mapping S is µ-full if, for every t, S is ¯τ -full locally at t with respect to some ¯τ ⊇ τ that supports µ locally at t. Recall that S is ¯τ -continuous at t if it is both isc and osc at t, and that S is ¯τ -continuous locally at t it is continuous at every point on some ¯τ -neighborhood of t. Example 2 below shows that, in the next definition, it is important to allow the topology to depend on t. Definition 2. The mapping S is µ-continuous if, for every t, its image closure is continuous locally at t with respect to some ¯τ ⊇ τ that supports µ locally at t. Theorem 8. Continuous µ-selections of S are selections of the image closure of S under either of the following conditions, 1. S is µ-continuous, 2. S is µ-full. Proof. 1 follows from 2, since when S is ¯τ -continuous locally at t, then ¯τ - lim sup St = cl St, ¯τ - lim inf St = cl St, and ¯τ - lim inf(¯τ - lim sup S)t = cl St. To prove 2, let y be a τ -continuous µ-selection of S. Fix t, and ¯τ and O ∈ ¯Ht in Definition 1. We may assume that every O-relative ¯τ -open set has a set with positive measure, Then every ¯τ -neighborhood of t′ ∈ O contains a 11 subset with positive measure and, thus, a point t′′ with yt′ ∈ St′′ , and hence yt′ ∈ ¯τ - lim sup St′. By continuity again, yt ∈ ¯τ - lim inf(¯τ - lim sup S)t, so yt ∈ cl St, which proves 2. Example 1. A fully lower semicontinuous mappings S is τ -full. Indeed, [Roc71, Lemma 2] characterizes full lower semicontinuity by the property that int G = int cl G for G = {(t, x) x ∈ int S}. It suffices to show that for a set-valued mapping Γ, int gph Γ ⊆ gph ((lim inf Γ)o), where (lim inf Γ)o t := int lim inf Γt. Assume that (t, x) ∈ int gph Γ. Then there exists A ∈ Ho x and O ∈ Ht such that O × A ∈ gph Γ. Thus for every x′ ∈ A and t′ ∈ O, Γ−1(O′) ∈ Hx′ for every O′ ∈ Ht′ , which means that x′ ∈ lim inf Γt′ , and hence (x, t) ∈ gph ((lim inf Γ)o). Example 2. Let T = R be endowed with the Euclidean topology τ . Assume that S : R ⇒ Rd has no removable discontinuities in the sense that, for every y ∈ Rd and t, the distance function t 7→ d(y, St) is either left- or right-continuous locally at t. Then S is µ-continuous. Indeed, by [RW98, Proposition 5.11], the left-continuity of d(y, S) is equivalent to S being left-continuous with respect to the topology generated by {(s, t] s < t}, and the right-continuity is handled symmetrically. 5 Appendix The following lemma gives a general condition for lower semicontinuity of a convex function on the dual of a Frechet space V . Lemma 9. A convex function g on the dual of V is lsc if and only if g + σO is lsc for every element O of a local basis of the origin. Proof. By Krein-Smulian theorem [KN76, Theorem 22.6], it suffices to show that g is lsc relative to every weak∗-compact set K. Assume to the contrary that there is a net (xν ) in K converging to x, α ∈ R, and ǫ > 0 such that g(xν) ≤ α but g(x) ≥ α + ǫ. Let O ⊆ ǫ 2 K ◦ be an element of the local basis, 2 K ◦ ≤ ǫ where K ◦ := {v supx∈K hv, xi ≤ 1} is the polar of K. Then σO ≤ σ ǫ on K, so 2 α + ǫ ≤ g(x) ≤ (g + σO)(x) ≤ lim inf(g + σO)(xν ) < α + ǫ which is a contradiction. References [AB88] L. Ambrosio and G. Buttazzo. Weak lower semicontinuous envelope of functionals defined on a space of measures. Ann. Mat. Pura Appl. (4), 150:311 -- 339, 1988. 12 [BV88] G. Bouchitt´e and M. Valadier. Integral representation of convex func- tionals on a space of measures. J. Funct. Anal., 80(2):398 -- 420, 1988. [BW96] A. Bagh and R. Wets. Convergence of set-valued mappings: equi-outer semicontinuity. Set-Valued Anal., 4(4):333 -- 360, 1996. [Dug66] James Dugundji. Topology. Allyn and Bacon, Inc., Boston, Mass., 1966. [Fol99] G. B. Folland. Real analysis. Pure and Applied Mathematics. John Wiley & Sons Inc., New York, second edition, 1999. Modern techniques and their applications, A Wiley-Interscience Publication. [KN76] J. L. Kelley and I. Namioka. Linear topological spaces. Springer- Verlag, New York, 1976. With the collaboration of W. F. Donoghue, Jr., Kenneth R. Lucas, B. J. Pettis, Ebbe Thue Poulsen, G. Baley Price, Wendy Robertson, W. R. Scott, and Kennan T. Smith, Second corrected printing, Graduate Texts in Mathematics, No. 36. [Mic56] E. Michael. Continuous selections. I. Ann. of Math. (2), 63:361 -- 382, 1956. [Mun00] J. Munkres. Topology. Prentice Hall, 2 edition, jan 2000. [Per14] A.-P. Perkkio. Continuous Essential Selections and Integral Function- als. Set-Valued Var. Anal., 22(1):45 -- 58, 2014. [PP15] T. Pennanen and A.-P. Perkkio. Convex integral functionals of regular processes. arXiv, 2015. [Roc66] R. T. Rockafellar. Level sets and continuity of conjugate convex func- tions. Trans. Amer. Math. Soc., 123:46 -- 63, 1966. [Roc70] R. T. Rockafellar. Convex analysis. Princeton Mathematical Series, No. 28. Princeton University Press, Princeton, N.J., 1970. [Roc71] R. T. Rockafellar. Integrals which are convex functionals. II. Pacific J. Math., 39:439 -- 469, 1971. [Roc74] R. T. Rockafellar. Conjugate duality and optimization. Society for Industrial and Applied Mathematics, Philadelphia, Pa., 1974. [RW98] R. T. Rockafellar and R. J.-B. Wets. Variational analysis, volume 317 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, 1998. 13
1304.3042
2
1304
2013-08-27T11:13:31
Discrete integrals based on comonotonic modularity
[ "math.FA" ]
It is known that several discrete integrals, including the Choquet and Sugeno integrals as well as some of their generalizations, are comonotonically modular functions. Based on a recent description of the class of comonotonically modular functions, we axiomatically identify more general families of discrete integrals that are comonotonically modular, including signed Choquet integrals and symmetric signed Choquet integrals as well as natural extensions of Sugeno integrals.
math.FA
math
DISCRETE INTEGRALS BASED ON COMONOTONIC MODULARITY MIGUEL COUCEIRO AND JEAN-LUC MARICHAL Abstract. It is known that several discrete integrals, including the Choquet and Sugeno integrals as well as some of their generalizations, are comonotoni- cally modular functions. Based on a recent description of the class of comono- tonically modular functions, we axiomatically identify more general families of discrete integrals that are comonotonically modular, including signed Choquet integrals and symmetric signed Choquet integrals as well as natural extensions of Sugeno integrals. . A F h t a m [ 2 v 2 4 0 3 . 4 0 3 1 : v i X r a 1. Introduction Aggregation functions arise wherever merging information is needed: applied and pure mathematics (probability, statistics, decision theory, functional equations), operations research, computer science, and many applied fields (economics and finance, pattern recognition and image processing, data fusion, etc.). For recent references, see Beliakov et al. [1] and Grabisch et al. [12]. Discrete Choquet integrals and discrete Sugeno integrals are among the best known functions in aggregation theory, mainly because of their many applications, for instance, in decision making (see the edited book [13]). More generally, signed Choquet integrals, which need not be nondecreasing in their arguments, and the Lov´asz extensions of pseudo-Boolean functions, which need not vanish at the origin, are natural extensions of the Choquet integrals and have been thoroughly investi- gated in aggregation theory. For recent references, see, e.g., [3, 8]. The class of n-variable Choquet integrals has been axiomatized independently by means of two noteworthy aggregation properties, namely comonotonic additivity (see, e.g., [11]) and horizontal min-additivity (originally called "horizontal additiv- ity", see [2]). Function classes characterized by these properties have been recently described by the authors [8]. Quasi-Lov´asz extensions, which generalize signed Cho- quet integrals and Lov´asz extensions by transforming the arguments by a 1-variable function, have also been recently investigated by the authors [9] through natural aggregation properties. Lattice polynomial functions and quasi-Sugeno integrals generalize the notion of Sugeno integral [4 -- 7, 10]: the former by removing the idempotency requirement and the latter also by transforming arguments by a 1-variable function. Likewise, these functions have been axiomatized by means of well-known properties such as comonotonic maxitivity and comonotonic minitivity. Date: June 6, 2013. 2010 Mathematics Subject Classification. Primary 39B22, 39B72; Secondary 26B35. Key words and phrases. Aggregation function, discrete Choquet integral, discrete Sugeno in- tegral, functional equation, comonotonic additivity, comonotonic modularity, axiomatization. 1 2 MIGUEL COUCEIRO AND JEAN-LUC MARICHAL All of these classes share the feature that its members are comonotonically mod- ular. These facts motivated a recent study that led to a description of comonoton- ically modular functions [9]. In this paper we survey these and other results and present a somewhat typological study of the vast class of comonotonically modular functions, where we identify several families of discrete integrals within this class using variants of homogeneity as distinguishing feature. The paper is organized as follows. In Section 2 we recall basic notions and terminology related to the concept of signed Choquet integral and present some preliminary characterization results. In Section 3 we survey several results that culminate in a description of comonotonic modularity and establish connections to other well studied properties of aggregation functions. These results are then used in Section 4 to provide characterizations of the various classes of functions considered in the previous sections as well as of classes of functions extending Sugeno integrals. We employ the following notation throughout the paper. The set of permutations on X = {1, . . . , n} is denoted by SX . For every σ ∈ SX , we define Rn σ = {x = (x1, . . . , xn) ∈ Rn ∶ xσ(1) ⩽ ⋯ ⩽ xσ(n)}. Let R+ = [0, +∞[ and R− = ]−∞, 0]. We let I denote a nontrivial (i.e., of positive Lebesgue measure) real interval, possibly unbounded. We also introduce the nota- tion I+ = I ∩ R+, I− = I ∩ R−, and I n σ. For every S ⊆ X, the symbol 1S denotes the n-tuple whose ith component is 1, if i ∈ S, and 0, otherwise. Let also 1 = 1X and 0 = 1∅. The symbols ∧ and ∨ denote the minimum and maximum func- tions, respectively. For every x ∈ Rn, let x+ be the n-tuple whose ith component is xi ∨ 0 and let x− = (−x)+. For every permutation σ ∈ SX and every i ∈ X, we set σ(i) = {σ(i), . . . , σ(n)}, S↓ S↑ σ(i) = {σ(1), . . . , σ(i)}, and S↑ σ(n + 1) = S↓ σ(0) = ∅. σ = I n ∩ Rn 2. Signed Choquet integrals In this section we recall the concepts of Choquet integral, signed Choquet inte- gral, and symmetric signed Choquet integral. We also recall some axiomatizations of these function classes. For general background, see [3, 8, 9] A capacity on X = {1, . . . , n} is a set function µ∶ 2X → R such that µ(∅) = 0 and µ(S) ⩽ µ(T) whenever S ⊆ T . Definition 1. The Choquet integral with respect to a capacity µ on X is the function Cµ∶ Rn + → R defined as Cµ(x) = n Q i=1 xσ(i) µ(S↑ σ(i)) − µ(S↑ σ(i + 1)) , x ∈ (Rn +)σ, σ ∈ SX . The concept of Choquet integral can be formally extended to more general set functions and n-tuples of Rn as follows. A signed capacity on X is a set function v∶ 2X → R such that v(∅) = 0. Definition 2. The signed Choquet integral with respect to a signed capacity v on X is the function Cv∶ Rn → R defined as (1) Cv(x) = n Q i=1 xσ(i) v(S↑ σ(i)) − v(S↑ σ(i + 1)) , x ∈ Rn σ, σ ∈ SX . From (1) it follows that Cv(1S) = v(S) for every S ⊆ X. Thus Eq. (1) can be rewritten as 3 (2) Cv(x) = n Q i=1 xσ(i) Cv(1S↑ σ(i)) − Cv(1S↑ σ(i+1)) , x ∈ Rn σ, σ ∈ SX . Thus defined, the signed Choquet integral with respect to a signed capacity v on X is the continuous function Cv whose restriction to each region Rn σ (σ ∈ SX ) is the unique linear function that coincides with v (or equivalently, the corresponding pseudo-Boolean function v∶{0, 1}n → R) at the n+1 vertices of the standard simplex [0, 1]n ∩ Rn σ of the unit cube [0, 1]n. As such, Cv is called the Lov´asz extension of v. From this observation we immediately derive the following axiomatization of the class of n-variable signed Choquet integrals over a real interval I. A function f ∶ I n → R is said to be a signed Choquet integral if it is the restriction to I n of a signed Choquet integral. Theorem 3 ( [3]). Assume that 0 ∈ I. A function f ∶ I n → R satisfying f(0) = 0 is a signed Choquet integral if and only if λ ∈ [0, 1], x, x′ ∈ I n f(λ x + (1 − λ) x′) = λ f(x) + (1 − λ) f(x′), The next theorem provides an axiomatization of the class of n-variable signed Choquet integrals based on comonotonic additivity, horizontal min-additivity, and horizontal max-additivity. Recall that two n-tuples x, x′ ∈ I n are said to be comono- σ . A function f ∶ I n → R is said to tonic if there exists σ ∈ SX such that x, x′ ∈ I n be comonotonically additive if, for every comonotonic n-tuples x, x′ ∈ I n such that x + x′ ∈ I n, we have σ , σ ∈ SX . f(x + x′) = f(x) + f(x′). A function f ∶ I n → R is said to be horizontally min-additive (resp. horizontally max-additive) if, for every x ∈ I n and every c ∈ I such that x − x ∧ c ∈ I n (resp. x − x ∨ c ∈ I n), we have f(x) = f(x ∧ c) + f(x − x ∧ c) resp. f(x) = f(x ∨ c) + f(x − x ∨ c). Theorem 4 ( [8]). Assume [0, 1] ⊆ I ⊆ R+ or I = R. Then a function f ∶ I n → R is a signed Choquet integral if and only if the following conditions hold: max-additive if I = R). (i) f is comonotonically additive or horizontally min-additive (or horizontally (ii) f(c x1S) = c f(x1S) for all x ∈ I and c > 0 such that c x ∈ I and all S ⊆ X. Remark 1. It is easy to see that condition (ii) of Theorem 4 is equivalent to the following simpler condition: f(x1S) = sign(x) x f(sign(x) 1S) for all x ∈ I and S ⊆ X. We now recall the concept of symmetric signed Choquet integral. Here "sym- metric" does not refer to invariance under a permutation of variables but rather to the role of the origin of Rn as a symmetry center with respect to the function values. Definition 5. Let v be a signed capacity on X. The symmetric signed Choquet integral with respect to v is the function Cv∶ Rn → R defined as x ∈ Rn. (3) Cv(x) = Cv(x+) − Cv(x−), 4 MIGUEL COUCEIRO AND JEAN-LUC MARICHAL Thus defined, a symmetric signed Choquet integral is an odd function in the sense that Cv(−x) = − Cv(x). It is then not difficult to show that the restriction of Cv to Rn σ is the function Cv(x) = p Q i=1 xσ(i) Cv(1S↓ σ(i)) − Cv(1S↓ σ(i−1)) (4) + n Q i=p+1 xσ(i) Cv(1S↑ σ(i)) − Cv(1S↑ σ(i+1)), x ∈ Rn σ, where the integer p ∈ {0, . . . , n} is given by the condition xσ(p) < 0 ⩽ xσ(p+1), with the convention that xσ(0) = −∞ and xσ(n+1) = +∞. The following theorem provides an axiomatization of the class of n-variable sym- metric signed Choquet integrals based on horizontal median-additive additivity. Assuming that I is centered at 0, recall that a function f ∶ I n → R is said to be horizontally median-additive if, for every x ∈ I n and every c ∈ I+, we have f(x) = fmed(−c, x, c) + fx − x ∧ c + fx − x ∨ (−c), (5) where med(−c, x, c) is the n-tuple whose ith component is the middle value of {−c, xi, c}. Equivalently, a function f ∶ I n → R is horizontally median-additive if and only if its restrictions to I n − are comonotonically additive and + and I n f(x) = f(x+) + f(−x−), x ∈ I n. A function f ∶ I n → R is said to be a symmetric signed Choquet integral if it is the restriction to I n of a symmetric signed Choquet integral. Theorem 6 ( [8]). Assume that I is centered at 0 with [−1, 1] ⊆ I. Then a func- tion f ∶ I n → R is a symmetric signed Choquet integral if and only if the following conditions hold: (i) f is horizontally median-additive. (ii) f(c x1S) = c f(x1S) for all c, x ∈ I such that c x ∈ I and all S ⊆ X. Remark 2. It is easy to see that condition (ii) of Theorem 6 is equivalent to the following simpler condition: f(x1S) = x f(1S) for all x ∈ I and S ⊆ X. We end this section by recalling the following important formula. For every signed capacity v on X, we have Cv(x) = Cv(x+) − Cvd(x−), (6) where vd is the capacity on X, called the dual capacity of v, defined as vd(S) = v(X) − v(X ∖ S). x ∈ Rn, 3. Comonotonic modularity Recall that a function f ∶ I n → R is said to be modular (or a valuation) if f(x) + f(x′) = f(x ∧ x′) + f(x ∨ x′) (7) for every x, x′ ∈ I n, where ∧ and ∨ are considered componentwise. It was proved [16] that a function f ∶ I n → R is modular if and only if it is separable, that is, there exist n functions fi∶ I → R (i = 1, . . . , n) such that f = ∑n i=1 fi. In particular, any 1-variable function f ∶ I → R is modular. More generally, a function f ∶ I n → R is said to be comonotonically modular (or a comonotonic valuation) if (7) holds for every comonotonic n-tuples x, x′ ∈ I n; 5 see [9, 15]. It was shown [9] that a function f ∶ I n → R is comonotonically modular if and only if it is comonotonically separable, that is, for every σ ∈ SX , there exist functions f σ i ∶ I → R (i = 1, . . . , n) such that f(x) = n Q i=1 i (xσ(i)) = f σ n Q i=1 σ−1(i)(xi), f σ x ∈ I n σ . We also have the following important definitions. For every x ∈ Rn and every c ∈ R+ (resp. c ∈ R−) we denote by [x]c (resp. [x]c) the n-tuple whose ith component is 0, if xi ⩽ c (resp. xi ⩾ c), and xi, otherwise. Recall that a function f ∶ I n → R, where 0 ∈ I ⊆ R+, is invariant under horizontal min-differences if, for every x ∈ I n and every c ∈ I, we have (8) Dually, a function f ∶ I n → R, where 0 ∈ I ⊆ R−, is invariant under horizontal max- differences if, for every x ∈ I n and every c ∈ I, we have f(x) − f(x ∧ c) = f([x]c) − f([x]c ∧ c). (9) f(x) − f(x ∨ c) = f([x]c) − f([x]c ∨ c). The following theorem provides a description of the class of functions which are comonotonically modular. Theorem 7 ( [9]). Assume that I ∋ 0. For any function f ∶ I n → R, the following assertions are equivalent. (i) f is comonotonically modular. (ii) fI n fI n and we have f(x) + f(0) = f(x+) + f(−x−) for every x ∈ I n. + is comonotonically modular (or invariant under horizontal min-differences), − is comonotonically modular (or invariant under horizontal max-differences), + → R and h∶ I n − → R such that, for every σ ∈ SX and every (iii) There exist g∶ I n x ∈ I n σ , f(x) = f(0)+ p Q i=1 h(xσ(i)1S↓ σ(i))−h(xσ(i)1S↓ σ(i−1))+ n Q i=p+1 g(xσ(i)1S↑ σ(i))−g(xσ(i)1S↑ σ(i+1)), where p ∈ {0, . . . , n} is such that xσ(p) < 0 ⩽ xσ(p+1), with the convention that xσ(0) = −∞ and xσ(n+1) = +∞. In this case, we can choose g = fI n and h = fI n − . + We finish this section with remarks on some properties subsumed by comono- tonic modularity, namely the following relaxations of maxitivity and minitivity properties. Recall that a function f ∶ I n → R is said to be maxitive if (10) f(x ∨ x′) = f(x) ∨ f(x′), x, x′ ∈ I n, and it is said to be minitive if (11) f(x ∧ x′) = f(x) ∧ f(x′), x, x′ ∈ I n. As in the case of modularity, maxitivity and minitivity give rise to noteworthy decompositions of functions into maxima and minima, respectively, of 1-variable functions. In the context of Sugeno integrals (see Section 4), de Campos et al. [11] proposed the following comonotonic variants of these properties. A function f ∶ I n → R is said to be comonotonic maxitive (resp. comonotonic minitive) if (10) (resp. (11)) holds for any two comonotonic n-tuples x, x′ ∈ I n. It was shown in [7] that any of these 6 MIGUEL COUCEIRO AND JEAN-LUC MARICHAL properties implies nondecreasing monotonicity, and it is not difficult to observe that comonotonic maxitivity together with comonotonic minitivity imply comonotonic modularity; the converse is not true (e.g., the arithmetic mean). Explicit descriptions of each one of these properties was given in [4] for functions over bounded chains. For the sake of self-containment, we present these descriptions here. To this end, we now assume that I = [a, b] ⊆ R, and for each S ⊆ X, we denote by eS the n-tuple in {a, b}n whose i-th component is b if i ∈ S, and a otherwise. Theorem 8 ( [4]). Assume I = [a, b] ⊆ R. A function f ∶ I n → R is comonotonic maxitive (resp. comonotonic minitive) if and only if there exists a nondecreasing function g∶ I n → R such that f(x) =  S⊆X geS ∧  i∈S xi (resp. f(x) =  S⊆X geX∖S ∨  i∈S xi). In this case, we can choose g = f . These descriptions are further refined in the following corollary. Corollary 9. Assume I = [a, b] ⊆ R. For any function f ∶ I n → R, the following assertions are equivalent. (i) f is comonotonic maxitive (resp. comonotonic minitive). (ii) there are unary nondecreasing functions ϕS∶ I → R (S ⊆ X) such that f(x) =  S⊆X ϕS  i∈S xi (resp. f(x) =  S⊆X ϕS  i∈S xi). In this case, we can choose ϕS(x) = f(eS ∧ x) (resp. ϕS(x) = f(eX∖S ∨ x)) for every S ⊆ X. (iii) for every σ ∈ SX , there are nondecreasing functions f σ i ∶ I → R (i = 1, . . . , n) (resp. f(x) =  i∈X i (xσ(i))). f σ i (x) = f(eS↑ σ(i) ∧x) (resp. f σ i (x) = f(eS↓ σ(i−1) ∨ such that, for every x ∈ I n σ , i (xσ(i)) f σ f(x) =  i∈X In this case, we can choose f σ x)). Remark 3. (i) Note that the expressions provided in Theorem 8 and Corollary 9 greatly differ from the additive form given in Theorem 7. (ii) An alternative description of comonotonic maxitive (resp. comonotonic minitive) functions was obtained in Grabisch et al. [12, Ch. 2]. 4. Classes of comonotonically modular integrals In this section we present axiomatizations of classes of functions that naturally generalize Choquet integrals (e.g., signed Choquet integrals and symmetric signed Choquet integrals) by means of comonotonic modularity and variants of homogene- ity. From the analysis of the more stringent properties of comonotonic minitivity and comonotonic maxitivity, we also present axiomatizations of classes of functions generalizing Sugeno integrals. 7 4.1. Comonotonically modular integrals generalizing Choquet integrals. The following theorem provides an axiomatization of the class of n-variable signed Choquet integrals. Theorem 10. Assume [0, 1] ⊆ I ⊆ R+ or [−1, 1] ⊆ I. Then a function f ∶ I n → R is a signed Choquet integral if and only if the following conditions hold: (i) f is comonotonically modular. (ii) f(0) = 0 and f(x1S) = sign(x) x f(sign(x) 1S) for all x ∈ I and S ⊆ X. (iii) If [−1, 1] ⊆ I, then f(1X∖S) = f(1) + f(−1S) for all S ⊆ X. Proof. (Necessity) Assume that f is a signed Choquet integral, f = Cv. Then condition (ii) is satisfied in view of Theorem 4 and Remark 1. If [−1, 1] ⊆ I, then by (6) we have Cv(−1S) = −Cvd(1S) = Cv(1X∖S) − Cv(1), which shows that condition (iii) is satisfied. Let us now show that condition (i) is σ, setting p ∈ {0, . . . , n} such that also satisfied. For every σ ∈ SX and every x ∈ Rn xσ(p) < 0 ⩽ xσ(p+1), by (2) and conditions (iii) and (ii), we have Cv(x) = = = n Q i=1 p Q i=1 p Q i=1 xσ(i) Cv(1S↑ σ(i)) − Cv(1S↑ σ(i+1)) xσ(i) Cv(−1S↓ σ(i−1)) − Cv(−1S↓ σ(i)) + n Q i=p+1 xσ(i) Cv(1S↑ σ(i)) − Cv(1S↑ σ(i+1)) Cv(xσ(i) 1S↓ σ(i)) − Cv(xσ(i) 1S↓ σ(i−1)) + n Q i=p+1 Cv(xσ(i) 1S↑ σ(i)) − Cv(xσ(i) 1S↑ σ(i+1)) which shows that condition (iii) of Theorem 7 is satisfied. Hence Cv is comonoton- ically modular. (Sufficiency) Assume that f satisfies conditions (i) -- (iii). By condition (iii) of σ we have Theorem 7 and conditions (ii) and (iii), for every σ ∈ SX and every x ∈ Rn f(x) = = = p Q i=1 p Q i=1 n Q i=1 f(xσ(i)1S↓ σ(i)) − f(xσ(i)1S↓ σ(i−1)) + n Q i=p+1 f(xσ(i)1S↑ σ(i)) − f(xσ(i)1S↑ σ(i+1)) xσ(i) f(−1S↓ σ(i−1)) − f(−1S↓ σ(i)) + n Q i=p+1 xσ(i) f(1S↑ σ(i)) − f(1S↑ σ(i+1)) xσ(i) f(1S↑ σ(i)) − f(1S↑ σ(i+1)) which, combined with (2), shows that f is a signed Choquet integral. Remark 4. Condition (iii) of Theorem 10 is necessary. Indeed, the function f(x) = Cv(x+) satisfies conditions (i) and (ii) but fails to satisfy condition (iii). Theorem 11. Assume I is centered at 0 with [−1, 1] ⊆ I. Then a function f ∶ I n → R is a symmetric signed Choquet integral if and only if the following conditions hold: (cid:3) (i) f is comonotonically modular. (ii) f(x1S) = x f(1S) for all x ∈ I and S ⊆ X. Proof. (Necessity) Assume that f is a symmetric signed Choquet integral, f = Cv. Then condition (ii) is satisfied in view of Theorem 6 and Remark 2. Let us now 8 MIGUEL COUCEIRO AND JEAN-LUC MARICHAL show that condition (i) is also satisfied. For every σ ∈ SX and every x ∈ Rn p ∈ {0, . . . , n} such that xσ(p) < 0 ⩽ xσ(p+1), by (4) and condition (ii), we have σ, setting Cv(x) = = p Q i=1 p Q i=1 xσ(i) Cv(1S↓ σ(i)) − Cv(1S↓ σ(i−1)) + n Q i=p+1 xσ(i) Cv(1S↑ σ(i)) − Cv(1S↑ σ(i+1)) Cv(xσ(i) 1S↓ σ(i)) − Cv(xσ(i) 1S↓ σ(i−1)) + n Q i=p+1 Cv(xσ(i) 1S↑ σ(i)) − Cv(xσ(i) 1S↑ σ(i+1)) which shows that condition (iii) of Theorem 7 is satisfied. Hence Cv is comonoton- ically modular. (Sufficiency) Assume that f satisfies conditions (i) and (ii). By condition (iii) of Theorem 7 and condition (ii), for every σ ∈ SX and every x ∈ Rn σ we have f(x) = = p Q i=1 p Q i=1 f(xσ(i)1S↓ σ(i)) − f(xσ(i)1S↓ σ(i−1)) + n Q i=p+1 f(xσ(i)1S↑ σ(i)) − f(xσ(i)1S↑ σ(i+1)) xσ(i) f(1S↓ σ(i)) − f(1S↓ σ(i−1)) + n Q i=p+1 xσ(i) f(1S↑ σ(i)) − f(1S↑ σ(i+1)) which, combined with (4), shows that f is a symmetric signed Choquet integral. (cid:3) The authors [9] showed that comonotonically modular functions also include the class of signed quasi-Choquet integrals on intervals of the forms I+ and I− and the class of symmetric signed quasi-Choquet integrals on intervals I centered at the origin. Definition 12. Assume I ∋ 0 and let v be a signed capacity on X. A signed quasi-Choquet integral with respect to v is a function f ∶ I n → R defined as f(x) = Cv(ϕ(x1), . . . , ϕ(xn)), where ϕ∶ I → R is a nondecreasing function satisfying ϕ(0) = 0. We now recall axiomatizations of the class of n-variable signed quasi-Choquet integrals on I+ and I− by means of comonotonic modularity and variants of homo- geneity. Theorem 13 ( [9]). Assume [0, 1] ⊆ I ⊆ R+ (resp. [−1, 0] ⊆ I ⊆ R−) and let f ∶ I n → R be a nonconstant function such that f(0) = 0. Then the following assertions are equivalent. + is comonotonically modular (or invariant under horizontal min-differences), − is comonotonically modular (or invariant under horizontal max-differences), (i) f is a signed quasi-Choquet integral and there exists S ⊆ X such that f(1S) ≠ 0 (resp. f(−1S) ≠ 0). (ii) fI n fI n and there exists a nondecreasing function ϕ∶ I → R satisfying ϕ(0) = 0 such that f(x1S) = sign(x) ϕ(x) f(sign(x) 1S) for every x ∈ I and every S ⊆ X. Remark 5. If I = [0, 1] (resp. I = [−1, 0]), then the "nonconstant" assumption and the second condition in assertion (i) of Theorem 13 can be dropped off. The extension of Theorem 13 to functions on intervals I centered at 0 and con- taining [−1, 1] remains an interesting open problem. We now recall the axiomatization obtained by the authors of the class of n- variable symmetric signed quasi-Choquet integrals. 9 Definition 14. Assume I is centered at 0 and let v be a signed capacity on X. A symmetric signed quasi-Choquet integral with respect to v is a function f ∶ I n → R defined as f(x) = Cv(ϕ(x1), . . . , ϕ(xn)), where ϕ∶ I → R is a nondecreasing odd function. Theorem 15 ( [9]). Assume that I is centered at 0 with [−1, 1] ⊆ I and let f ∶ I n → R − is nonconstant and f(0) = 0. Then the following be a function such that fI n assertions are equivalent. + or fI n (i) f is a symmetric signed quasi-Choquet integral and there exists S ⊆ X such that f(1S) ≠ 0. (ii) f is comonotonically modular and there exists a nondecreasing odd function ϕ∶ I → R such that f(x1S) = ϕ(x) f(1S) for every x ∈ I and every S ⊆ X. Remark 6. If I = [−1, 1], then the "nonconstant" assumption and the second con- dition in assertion (i) of Theorem 15 can be dropped off. 4.2. Comonotonically modular integrals generalizing Sugeno integrals. In this subsection we consider natural extensions of the n-variable Sugeno integrals on a bounded real interval I = [a, b]. By an I-valued capacity on X we mean an order preserving mapping µ∶ 2X → I such that µ(∅) = a and µ(X) = b. Definition 16. Assume that I = [a, b]. The Sugeno integral with respect to an I-valued capacity µ on X is the function Sµ∶ I n → I defined as Sµ(x) =  i∈X xσ(i) ∧ µ(S↑ σ(i)), x ∈ I n σ , σ ∈ SX . As the following proposition suggests, Sugeno integrals can be viewed as idem- potent "lattice polynomial functions" (see [14]). Proposition 17. Assume that I = [a, b]. A function f ∶ I n → I is a Sugeno integral if and only if f(e∅) = a, f(eX) = b, and for every x ∈ I n xi . f(x) =  f(eS) ∧  S⊆X i∈S As mentioned, the properties of comonotonic maxitivity and comonotonic mini- tivity were introduced by de Campos et al. in [11] to axiomatize the class of Sugeno integrals. However, without further assumptions, they define a wider class of func- tions that we now define. Definition 18. Assume that I = [a, b] and J = [c, d] are real intervals and let µ be an I-valued capacity on X. A quasi-Sugeno integral with respect to µ is a function f ∶ J n → I defined by f(x) = Sµ(ϕ(x1), . . . , ϕ(xn)), where ϕ∶ J → I is a nondecreasing function. Using Proposition 11 and Corollary 17 in [5], we obtain the following axiomati- zation of the class of quasi-Sugeno integrals. Theorem 19. Let I = [a, b] and J = [c, d] be real intervals and consider a function f ∶ J n → I. The following assertions are equivalent. (i) f is a quasi-Sugeno integral. (ii) f is comonotonically maxitive and comonotonically minitive. (iii) f is nondecreasing, and there exists a nondecreasing function ϕ∶ J → I such that for every x ∈ J n and r ∈ J, we have (12) f(r ∨ x) = ϕ(r) ∨ f(x) and f(r ∧ x) = ϕ(r) ∧ f(x), 10 MIGUEL COUCEIRO AND JEAN-LUC MARICHAL where r ∨ x (resp. r ∧ x) is the n-tuple whose ith component is r ∨ xi (resp. r ∧ xi). In this case, ϕ can be chosen as ϕ(x) = f(x, . . . , x). Remark 7. The two conditions given in (12) are referred to in [5] as quasi-max homogeneity and quasi-min homogeneity, respectively. As observed at the end of the previous section, condition (ii) (and hence (i) or (iii)) of Theorem 19 implies comonotonic modularity. As the following result shows, the converse is true whenever f is nondecreasing and verifies any of the following weaker variants of quasi-max homogeneity and quasi-min homogeneity: f(x ∨ eS) = f(x, . . . , x) ∨ f(eS), f(x ∧ eS) = f(x, . . . , x) ∧ f(eS), (13) (14) Theorem 20. Let I = [a, b] and J = [c, d] be real intervals and consider a function f ∶ J n → I. The following conditions are equivalent. x ∈ J, S ⊆ X, x ∈ J, S ⊆ X. f(x, . . . , x). (i) f is a quasi-Sugeno integral, f(x) = Sµ(ϕ(x1), . . . , ϕ(xn)), where ϕ(x) = (ii) f is a quasi-Sugeno integral. (iii) f is comonotonically modular, nondecreasing, and satisfies (13) or (14). (iv) f is nondecreasing and satisfies (13) and (14). Proof. (i) ⇒ (ii) Trivial. (ii) ⇒ (iii) Follows from Theorem 19. (iii) ⇒ (iv) Suppose that f is comonotonically modular and satisfies (13). Then, f(x ∧ eS) = f(x, . . . , x) + f(eS) − f(x ∨ eS) = f(x, . . . , x) + f(eS) − f(x, . . . , x) ∨ f(eS) = f(x, . . . , x) ∧ f(eS). Hence f satisfies (14). The other case can be dealt with dually. (iv) ⇒ (i) Define ϕ(x) = f(x, . . . , x). By nondecreasing monotonicity and (14), for every S ⊆ X we have f(x) ⩾ feS ∧  xi = f(eS) ∧ ϕ  i∈S xi = f(eS) ∧  i∈S ϕ(xi) i∈S and thus f(x) ⩾ ⋁S⊆X f(eS) ∧ ⋀i∈S ϕ(xi). To complete the proof, it is enough to establish the converse inequality. Let S∗ ⊆ X be such that f(eS∗) ∧ ⋀i∈S∗ ϕ(xi) is maximum. Define T = j ∈ X ∶ ϕ(xj) ⩽ f(eS∗) ∧  i∈S∗ ϕ(xi). We claim that T ≠ ∅. Suppose this is not true, that is, ϕ(xj) > f(eS∗)∧⋀i∈S∗ ϕ(xi) for every j ∈ X. Then, by nondecreasing monotonicity, we have f(eX) ⩾ f(eS∗), and since f(eX) ⩾ ⋀i∈X ϕ(xi), f(eX) ∧  i∈X ϕ(xi) > f(eS∗) ∧  i∈S∗ ϕ(xi) which contradicts the definition of S∗. Thus T ≠ ∅. Now, by nondecreasing monotonicity and (13) we have f(x) ⩽ feX∖T ∨  j∈T xj = f(eX∖T)∨ϕ  j∈T xj = f(eX∖T)∨  j∈T ϕ(xj) = f(eX∖T). Indeed, we have ϕ(xj) ⩽ f(x) for every j ∈ T and x ⩽ eX∖T ∨ ⋁j∈T xj. 11 Note that f(eX∖T) ⩽ f(eS∗) ∧ ⋀i∈S∗ ϕ(xi) since otherwise, by definition of T , we would have f(eX∖T) ∧  i∈X∖T ϕ(xi) > f(eS∗) ∧  i∈S∗ ϕ(xi), again contradicting the definition of S∗. Finally, ϕ(xi) =  f(x) ⩽ f(eS∗) ∧  i∈S∗ f(eS) ∧  i∈S ϕ(xi), S⊆X and the proof is thus complete. (cid:3) Remark 8. An axiomatization of the class of Sugeno integrals based on comonotonic modularity can be obtained from Theorems 19 and 20 by adding the idempotency property. 5. Conclusion In this paper we analyzed comonotonic modularity as a feature common to many well-known discrete integrals. In doing so, we established its relation to many other noteworthy aggregation properties such as comonotonic relaxations of addi- tivity, maxitivity and minitivity. In fact, the latter become equivalent in presence of comonotonic modularity. As a by-product we immediately see that, e.g., the so-called discrete Shilkret integral lies outside the class of comonotonic modular functions since this integral is comonotonically maxitive but not comonotonically minitive. Albeit such an example, the class of comonotonically modular functions is rather vast and includes several important extensions of the Choquet and Sugeno integrals. The results presented in Section 4 seem to indicate that suitable variants of homo- geneity suffice to distinguish and fully describe these extensions. This naturally asks for an exhaustive study of homogeneity-like properties, which may lead to a complete classification of all subclasses of comonotonically modular functions. Another question that still eludes us is the relation between the additive forms given by comonotonic modularity and the max-min forms. As shown in Theorem 20, the latter are particular instances of the former; in fact, proof of Theorem 20 pro- vides a procedure to construct max-min representations of comonotonically modu- lar functions, whenever they exist. However, we were not able to present a direct translation between the two. This remains as a relevant open question since its answer will inevitably provide a better understanding of the synergy between these intrinsically different normal forms. Acknowledgments This research is partly supported by the internal research project F1R-MTH- PUL-12RDO2 of the University of Luxembourg. References [1] G. Beliakov, A. Pradera, and T. Calvo. Aggregation Functions: A Guide for Practitioners. Studies in Fuziness and Soft Computing. Springer, Berlin, 2007. [2] P. Benvenuti, R. Mesiar, and D. Vivona. Monotone set functions-based integrals. In Handbook of measure theory, Vol. II, pages 1329 -- 1379. North-Holland, Amsterdam, 2002. [3] M. Cardin, M. Couceiro, S. Giove, and J.-L. Marichal. Axiomatizations of signed discrete Choquet integrals. Int. J. Uncertainty, Fuzziness and Knowledge-Based Systems, 19(2):193 -- 199, 2011. 12 MIGUEL COUCEIRO AND JEAN-LUC MARICHAL [4] M. Couceiro and J.-L. Marichal. Axiomatizations of quasi-polynomial functions on bounded chains. Aeq. Math., 78(1-2):195 -- 213, 2009. [5] M. Couceiro and J.-L. Marichal. Quasi-polynomial functions over bounded distributive lat- tices. Aeq. Math., 80(3):319 -- 334, 2010. [6] M. Couceiro and J.-L. Marichal. Characterizations of discrete Sugeno integrals as polynomial functions over distributive lattices. Fuzzy Sets and Systems, 161(5):694 -- 707, 2010. [7] M. Couceiro and J.-L. Marichal. Representations and characterizations of polynomial func- tions on chains. J. Mult.-Valued Logic Soft Comput., 16(1-2):65 -- 86, 2010. [8] M. Couceiro and J.-L. Marichal. Axiomatizations of Lov´asz extensions of pseudo-Boolean functions. Fuzzy Sets and Systems, 181:28 -- 38, 2011. [9] M. Couceiro and J.-L. Marichal. Axiomatizations of quasi-Lov´asz extensions of pseudo- Boolean functions. Aeq. Math., 82:213 -- 231, 2011. [10] M. Couceiro and J.-L. Marichal. Polynomial functions over bounded distributive lattices. J. Mult.-Valued Logic Soft Comput., 18(3-4):247 -- 256, 2012. [11] L. M. de Campos and M. J. Bolanos. Characterization and comparison of Sugeno and Choquet integrals. Fuzzy Sets and Systems, 52(1):61 -- 67, 1992. [12] M. Grabisch, J.-L. Marichal, R. Mesiar, and E. Pap. Aggregation functions. Encyclopedia of Mathematics and its Applications 127. Cambridge University Press, Cambridge, UK, 2009. [13] M. Grabisch, T. Murofushi, and M. Sugeno, editors. Fuzzy measures and integrals - Theory and applications, volume 40 of Studies in Fuzziness and Soft Computing. Physica-Verlag, Heidelberg, 2000. [14] J.-L. Marichal. Weighted lattice polynomials. Discrete Mathematics, 309(4):814 -- 820, 2009. [15] R. Mesiar and A. Mesiarov´a-Zem´ankov´a. The ordered modular averages. IEEE Trans. Fuzzy Syst., 19(1):42 -- 50, 2011. [16] D. M. Topkis. Minimizing a submodular function on a lattice. Operations Research, 26(2):305 -- 321, 1978. Lamsade, University Paris-Dauphine, Place du Mar´echal de Lattre de Tassigny, F- 75775 Paris cedex 16, France E-mail address: miguel.couceiro[at]dauphine.fr Mathematics Research Unit, FSTC, University of Luxembourg, 6, rue Coudenhove- Kalergi, L-1359 Luxembourg, Luxembourg E-mail address: jean-luc.marichal[at]uni.lu
1805.01461
2
1805
2018-10-11T11:28:06
Fredholm operators and essential S-spectrum in the quaternionic setting
[ "math.FA" ]
For bounded right linear operators, in a right quaternionic Hilbert space with a left multiplication defined on it, we study the approximate $S$-point spectrum. In the same Hilbert space, then we study the Fredholm operators and the Fredholm index. In particular, we prove the invariance of the Fredholm index under small norm operator and compact operator perturbations. Finally, in association with Fredholm operators, we develop the theory of essential S-spectrum. We also characterize the $S$-spectrum in terms of the essential S-spectrum and Fredholm operators. In the sequel we study left and right S-spectrums as needed for the development of the theory presented in this note.
math.FA
math
FREDHOLM OPERATORS AND ESSENTIAL S-SPECTRUM IN THE QUATERNIONIC SETTING B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ Abstract. For bounded right linear operators, in a right quaternionic Hilbert space with a left multiplication defined on it, we study the approximate S-point spectrum. In the same Hilbert space, then we study the Fredholm operators and the Fredholm index. In particular, we prove the invariance of the Fredholm index under small norm operator and compact operator perturbations. Finally, in association with the Fred- holm operators, we develop the theory of essential S-spectrum. We also characterize the S-spectrum in terms of the essential S-spectrum and Fredholm operators. In the sequel we study left and right S-spectrums as needed for the development of the theory presented in this note. 1. Introduction In the complex theory Fredholm operators play an important role in the investigations of various classes of singular integral equations, in the theory of perturbations of Her- mitian operators by Hermitian and non-Hermitian operators, and in obtaining a priori estimate in determining properties of certain differential operators [24, 15, 8, 14, 18]. Fredholm alternative theorem is used to derive an adjoint equation to the linear stability equations in Fluid dynamics, and useful in scattering of a 1-D particle on a small poten- tial barrier, see [15] and the many references therein. In the complex case, studies of Fredholm theory and perturbation results are of a great importance in the description of the essential spectrum. In several applications, such as essential spectrum of the Schrodinger equations, essential spectrum of the per- turbed Hamiltonians, in general, essential spectrum of the differential operators, essential spectrum of the transport operator (transport of the neutron, photons, molecules in gas, etc), in particular, the time dependent transport equations arise in a number of differ- ent applications in Biology, Chemistry and Physics, information about it is important [3, 14, 15, 16, 12, 24]. Further, near the essential spectrum, numerical calculations of eigenvalues become difficult. Hence they have to be treated analytically [15]. There are several distinct definitions of the essential spectrum. However they all coincide for the self-adjoint operators on Hilbert spaces [14, 3]. In this note we only consider the essential spectrum associated with Fredholm operators in the quaternionic setting. In the complex setting, in a Hilbert space H, a bounded linear operator, A, is not invertible if it is not bounded below (same is true in the quaternion setting, see theorem 3.7 below). The set of approximate eigenvalues which are λ ∈ C such that A−λIH, where Date: October 12, 2018. 1991 Mathematics Subject Classification. Primary 47A10, 47A53, 47B07. Key words and phrases. Quaternions, Quaternionic Hilbert spaces, S-spectrum, Fredholm operator, Essential S-spectrum. 1 2 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ IH is the identity operator on H, is not bounded below, equivalently, the set of λ ∈ C for which there is a sequence of unit vectors φ1, φ2, · · · such that lim kAφn −λφnk = 0. The n→∞ set of approximate eigenvalues is known as the approximate spectrum. In the quater- nionic setting, let V R H be a separable right Hilbert space, A be a bounded right linear operator, and Rq(A) = A2 − 2Re(q)A + q2I , with q ∈ H, the set of all quaternions, be the pseudo-resolvent operator, the set of right eigenvalues of Rq(A) coincide with the point S-spectrum (see proposition 4.5 in [13]). In this regard, it will be appropriate to define and study the quaternionic approximate S-point spectrum as the quaternions for which Rq(A) in not bounded below. V R H Due to the non-commutativity, in the quaternionic case there are three types of Hilbert spaces: left, right, and two-sided, depending on how vectors are multiplied by scalars. This fact can entail several problems. For example, when a Hilbert space H is one-sided (either left or right) the set of linear operators acting on it does not have a linear struc- ture. Moreover, in a one sided quaternionic Hilbert space, given a linear operator A and a quaternion q ∈ H, in general we have that (qA)† 6= qA† (see [20] for details). These re- strictions can severely prevent the generalization to the quaternionic case of results valid in the complex setting. Even though most of the linear spaces are one-sided, it is possible to introduce a notion of multiplication on both sides by fixing an arbitrary Hilbert basis of H. This fact allows to have a linear structure on the set of linear operators, which is a minimal requirement to develop a full theory. Thus, the framework of this paper, is in part, is a right quaternionic Hilbert space equipped with a left multiplication, introduced by fixing a Hilbert basis. As far as we know, the Fredholm operator theory and the essential S-spectrum and the approximate S-point spectrum have not been studied in the quaternionic setting yet. In this regard, in this note we investigate the quaternionic S-point spectrum, Fredholm operators and associated S-essential spectrum for a bounded right linear operator on a right quaternionic separable Hilbert space. Since the pseudo-resolvent operator, Rq(A) has real coefficients the left multiplication defined on a right quaternionic Hilbert space play a little role. Even the non-commutativity of quaternions does not play an essential role. Even though the S-approximate point spectrum, the S-essential spectrum and the Fredholm operators are structurally different from its complex counterparts, the results we obtain and their proofs are somewhat similar to those in the corresponding complex theory. The article is organized as follows. In section 2 we introduce the set of quaternions and quaternionic Hilbert spaces and their bases, as needed for the development of this article, which may not be familiar to a broad range of audience. In section 3 we define and investigate, as needed, right linear operators and their properties. We have given proofs for some results which are not available in the literature. In section 3.1 we define a basis dependent left multiplication on a right quaternionic Hilbert space. In section 3.2 we deal with the right S-spectrum, left S-spectrum, S-spectrum and its major partitions. In section 4 we provide a systematic study of compact operators which has not yet been done in the literature. In section 5 we study the approximate S-point spectrum, σS ap(A), of a bounded right linear operator, A, on a right quaternionic Hilbert space. In particular we prove that σS ap(A) is a non-empty closed subset of H and the S-spectrum FREDHOLM OPERATOR 3 is the union of the σS ap(A) and the continuous S-spectrum. In section 6 we study the In particular, Fredholm operators and its index for a bounded right linear operator. we prove the invariance of the Fredholm index under small norm operator and compact operator perturbations. In this section, since a quaternionic multiple of an operator is not involved, the proofs of these results are almost verbatim copies of the complex ones. In section 7, we study the essential S-spectrum as the S-spectrum of the quotient map image of a bounded right linear operator on the quaternionic version of the Calkin algebra and then characterize the S-essential spectrum in terms of Fredholm operators. We also establish the so-called Atkinsons theorem, prove the invariance of the essential S-spectrum under compact perturbations, and give a characterization to S-spectrum in terms of Fredholm operators and its index (see proposition 7.18). Section 8 ends the manuscript with a conclusion. 2. Mathematical preliminaries In order to make the paper self-contained, we recall some facts about quaternions which may not be well-known. For details we refer the reader to [1, 13, 23]. 2.1. Quaternions. Let H denote the field of all quaternions and H∗ the group (under quaternionic multiplication) of all invertible quaternions. A general quaternion can be written as q = q0 + q1i + q2j + q3k, q0, q1, q2, q3 ∈ R, where i, j, k are the three quaternionic imaginary units, satisfying i2 = j2 = k2 = −1 and ij = k = −ji, jk = i = −kj, ki = j = −ik. The quaternionic conjugate of q is q = q0 − iq1 − jq2 − kq3, while q = (qq)1/2 denotes the usual norm of the quaternion q. If q is non-zero element, it has inverse q−1 = q q2 . Finally, the set S = {I = x1i + x2j + x3k x1, x2, x3 ∈ R, x2 1 + x2 2 + x2 3 = 1}, contains all the elements whose square is −1. It is a 2-dimensional sphere in H identified with R4. 2.2. Quaternionic Hilbert spaces. In this subsection we discuss right quaternionic Hilbert spaces. For more details we refer the reader to [1, 13, 23]. 2.2.1. Right quaternionic Hilbert Space. Let V R cation by quaternions. For φ, ψ, ω ∈ V R H and q ∈ H, the inner product H be a vector space under right multipli- satisfies the following properties h· ·iV R H : V R H × V R H −→ H = hψ φiV R H = hφ φiV R H H H V R (i) hφ ψiV R (ii) kφk2 (iii) hφ ψ + ωiV R (iv) hφ ψqiV R (v) hφq ψiV R H H H = hφ ψiV R q = qhφ ψiV R H H > 0 unless φ = 0, a real norm = hφ ψiV R H + hφ ωiV R H 4 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ where q stands for the quaternionic conjugate. It is always assumed that the space V R H is complete under the norm given above and separable. Then, together with h· ·i this defines a right quaternionic Hilbert space. Quaternionic Hilbert spaces share many of the standard properties of complex Hilbert spaces. In this paper, more than one Hilbert spaces will be involved in the results, but one notices that every Hilbert space, involved in results, is right quaternionic Hilbert space. The next two Propositions can be established following the proof of their complex counterparts, see e.g. [13, 23]. Proposition 2.1. Let O = {ϕk k ∈ N } be an orthonormal subset of V R a countable index set. Then following conditions are pairwise equivalent: H , where N is (a) The closure of the linear combinations of elements in O with coefficients on the right is V R H . (b) For every φ, ψ ∈ V R and it holds: H , the series Pk∈N hφ ϕkiV R H hϕk ψiV R H converges absolutely hφ ψiV R H = X k∈N hφ ϕkiV R H hϕk ψiV R H . (c) For every φ ∈ V R H , it holds: kφk2 V R H = X k∈N hϕk φiV R H 2 . (d) O⊥ = {0}. Definition 2.2. The set O as in Proposition 2.1 is called a Hilbert basis of V R H . Proposition 2.3. Every quaternionic Hilbert space V R Hilbert bases of V R H have the same cardinality. Furthermore, if O is a Hilbert basis of V R H , then every φ ∈ V R H can be uniquely H has a Hilbert basis. All the decomposed as follows: where the series Pk∈N ϕkhϕk φiV R H ϕkhϕk φiV R φ = X k∈N converges absolutely in V R H . , H It should be noted that once a Hilbert basis is fixed, every left (resp. right) quaternionic Hilbert space also becomes a right (resp. left) quaternionic Hilbert space [13, 23]. The field of quaternions H itself can be turned into a left quaternionic Hilbert space by defining the inner product hq q′i = qq′ or into a right quaternionic Hilbert space with hq q′i = qq′. 3. Right quaternionic linear operators and some basic properties In this section we shall define right H-linear operators and recall some basis properties. Most of them are very well known. In this manuscript, we follow the notations in [2] and [13]. We shall also recall some results pertinent to the development of the paper. Definition 3.1. A mapping A : D(A) ⊆ V R of A, is said to be right H-linear operator or, for simplicity, right linear operator, if H , where D(A) stands for the domain H −→ U R A(φa + ψb) = (Aφ)a + (Aψ)b, if φ, ψ ∈ D(A) and a, b ∈ H. FREDHOLM OPERATOR 5 The set of all right linear operators from V R H to U R H will be denoted by I H will be denoted by L(V R H , U R . For a given A ∈ L(V R H ) and H , U R H ), V R H the identity linear operator on V R the range and the kernel will be ran(A) = {ψ ∈ U R ker(A) = {φ ∈ D(A) Aφ = 0}. H Aφ = ψ for φ ∈ D(A)} We call an operator A ∈ L(V R (3.1) H ) bounded if H , U R kAk = sup kφkV R H =1 kAφkU R H < ∞, or equivalently, there exist K ≥ 0 such that kAφkU R set of all bounded right linear operators from V R Set of all invertible bounded right linear operators from V R G(V R Assume that V R on it. Then, there exists a unique linear operator A† such that H ). We also denote for a set ∆ ⊆ H, ∆∗ = {q q ∈ ∆}. H to U R H , U R H H ≤ KkφkV R H will be denoted by B(V R for all φ ∈ D(A). The H ). H will be denoted by H to U R H , U R H is a right quaternionic Hilbert space, A is a right linear operator acting (3.2) hψ AφiU R H = hA†ψ φiV R H ; for all φ ∈ D(A), ψ ∈ D(A†), where the domain D(A†) of A† is defined by D(A†) = {ψ ∈ U R H ∃ϕ such that hψ AφiU R H = hϕ φiV R H }. The following theorem gives two important and fundamental results about right H-linear bounded operators which are already appeared in [13] for the case of V R H . Point (b) of the following theorem is known as the open mapping theorem. H = U R Theorem 3.2. Let A : D(A) ⊆ V R H −→ U R H be a right H-linear operator. Then H , U R (a) A ∈ B(V R (b) if A ∈ B(V R H , U R then A−1 ∈ B(V R H ) if and only if A is continuous. H ) is surjective, then A is open. In particular, if A is bijective H , U R H ). Proof. The proof is the same as the proof in a complex Hilbert space, (see e.g. [8]). (cid:3) The following proposition provides some useful aspects about the orthogonal comple- ment subsets. Proposition 3.3. Let M ⊆ V R H . Then ⊥ is closed. (a) M (b) if M is a closed subspace of V R (c) if dim(M ) < ∞, then M is a closed subspace. H then V R H = M ⊕ M ⊥. Proof. The proof is similar to the proof of the complex case, (see e.g. [22]). (cid:3) The points (a) and (b) of the following proposition are already appeared in [13] for the case V R H = U R H . Proposition 3.4. Let A ∈ B(V R H , U R H ). Then (a) ran(A)⊥ = ker(A†). (b) ker(A) = ran(A†)⊥. (c) ker(A) is closed subspace of V R H . Proof. The proofs are elementary. (cid:3) 6 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ Proposition 3.5. Let A : D(A) : V R A is closed and satisfies the condition that there exists c > 0 such that H −→ U R H be a right quaternionic linear operator. If then ran(A) is closed. kAφkU R H ≥ ckφkV R H , for all φ ∈ D(A), Proof. The proof can be manipulated from the proof of proposition 2.13 in [19]. Proposition 3.6. If A ∈ B(V R H ) and if kAk < 1, then the right linear operator I (cid:3) − A V R H is invertible and the inverse is given by (I V R H − A)−1 = n X k=0 Ak. Proof. The proof is exactly the same as its complex version, see theorem 2.17, in [10] for a complex proof. (cid:3) Theorem 3.7. (Bounded inverse theorem) Let A ∈ B(V R are equivalent. H , U R H ), then the following results (a) A has a bounded inverse on its range. (b) A is bounded below. (c) A is injective and has a closed range. Proof. The proof is exactly similar to its complex version. See theorem 1.2 in [18] for a complex proof. (cid:3) Proposition 3.8. Let A ∈ B(V R is closed in V R H . H , U R H ), then ran(A) is closed in U R H if and only if ran(A†) Proof. The proof is the same as for the complex Hilbert spaces, see lemma 1.5 in [18] for a complex proof. (cid:3) Proposition 3.9. Let A ∈ B(V R H ). Then, (a) A is invertible if and only if it is injective with a closed range (i.e., ker(A) = {0} and ran(A) = ran(A)). (b) A is left (right) invertible if and only if A† is right (left) invertible. (c) A is right invertible if and only if it is surjective (i.e., ran(A) = V R H ). Proof. The proof is also exactly similar to its complex version. See lemma 5.8 in [18] for a complex proof. (cid:3) Definition 3.10. Let A ∈ B(V R under A if A(M ) ⊆ M , where A(M ) = {Aφ φ ∈ M }. H ). A closed subspace M ⊆ V R H is said to be invariant Following results hold in any quaternionic normed linear space. Definition 3.11. Let X R H be a quaternionic normed linear space. A subset C is said to be totally bounded, if for every ε > 0, there exist N ∈ N and ϕi ∈ C : i = 1, 2, 3, · · · , N such that C ⊆ B(ϕi; ε), N [ i=1 where B(ϕi; ε) = {φ ∈ X R H kφ − ϕikX R H < ε} for all i = 1, 2, 3, · · · , N . Proposition 3.12. Let C be a subset in a right quaternionic normed linear space X R H . Then FREDHOLM OPERATOR 7 (a) C is compact if and only if C is complete and totally bounded. (b) if dim(X R H ) < ∞, C is compact if and only if C is closed and bounded. Proof. The proof is exactly similar to the complex proof. For a complex proof of (a) and (b), see theorem 6.19 in [21] and theorem 2.5-3 in [17] respectively. (cid:3) 3.1. Left Scalar Multiplications on U R H . We shall extract the definition and some properties of left scalar multiples of vectors on U R H from [13] as needed for the development of the manuscript. The left scalar multiple of vectors on a right quaternionic Hilbert space is an extremely non-canonical operation associated with a choice of preferred Hilbert basis. From the proposition 2.3, U R H has a Hilbert basis (3.3) O = {ϕk k ∈ N }, where N is a countable index set. The left scalar multiplication on U R defined as the map H × U R H ∋ (q, φ) 7−→ qφ ∈ U R H given by H induced by O is (3.4) for all (q, φ) ∈ H × U R H . qφ := X k∈N ϕkqhϕk φi, Proposition 3.13. [13] The left product defined in the equation 3.4 satisfies the follow- ing properties. For every φ, ψ ∈ U R H and p, q ∈ H, (a) q(φ + ψ) = qφ + qψ and q(φ p) = (qφ) p. (b) kqφk = qkφk. (c) q( pφ) = (q p)φ. (d) hqφ ψi = hφ qψi. (e) rφ = φr, for all r ∈ R. (f) qϕk = ϕkq, for all k ∈ N . Remark 3.14. (1) The meaning of writing pφ is p · φ, because the notation from the equation 3.4 may be confusing, when U R H = H. However, regarding the field H itself as a right H-Hilbert space, an orthonormal basis O should consist only of a singleton, say {ϕ0}, with ϕ0 = 1, because we clearly have θ = ϕ0hϕ0 θi, for all θ ∈ H. The equality from (f) of proposition 3.13 can be written as pϕ0 = ϕ0 p, for all p ∈ H. In fact, the left hand may be confusing and it should be understood as p · ϕ0, because the true equality pϕ0 = ϕ0 p would imply that ϕ0 = ±1. For the simplicity, we are writing pφ instead of writing p · φ. (2) Also one can trivially see that ( p + q)φ = pφ + qφ, for all p, q ∈ H and φ ∈ U R H . Furthermore, the quaternionic left scalar multiplication of linear operators is also H −→ U R H , defined in [5], [13]. For any fixed q ∈ H and a given right linear operator A : V R the left scalar multiplication of A is defined as a map qA : V R H by the setting H −→ U R (3.5) (qA)φ := q(Aφ) = X k∈N ϕkqhϕk Aφi, for all φ ∈ V R for all φ ∈ V R A : V R H −→ U R H ). It is straightforward that qA is a right linear operator. If qφ ∈ V R H , H , one can define right scalar multiplication of the right linear operator H as a map Aq : V R H by the setting H −→ U R (3.6) (Aq)φ := A(qφ), 8 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ for all φ ∈ D(A). It is also right linear operator. One can easily obtain that, if qφ ∈ V R H , for all φ ∈ V R H is dense in U R H and V R H , then (3.7) (qA)† = A†q and (Aq)† = qA†. 3.2. S-Spectrum. For a given right linear operator A : D(A) ⊆ V R we define the operator Rq(A) : D(A2) −→ H by H −→ V R H and q ∈ H, Rq(A) = A2 − 2Re(q)A + q2I V R H , where q = q0 + iq1 + jq2 + kq3 is a quaternion, Re(q) = q0 and q2 = q2 2 + q2 3. In the literature, the operator is called pseudo-resolvent since it is not the resolvent operator of A but it is the one related to the notion of spectrum as we shall see in the next definition. For more information, on the notion of S-spectrum the reader may consult e.g. [4, 5, 7], and [13]. Definition 3.15. Let A : D(A) ⊆ V R H be a right linear operator. The S- resolvent set (also called spherical resolvent set) of A is the set ρS(A) (⊂ H) such that the three following conditions hold true: H −→ V R 1 + q2 0 + q2 (a) ker(Rq(A)) = {0}. (b) ran(Rq(A)) is dense in V R H . (c) Rq(A)−1 : ran(Rq(A)) −→ D(A2) is bounded. The S-spectrum (also called spherical spectrum) σS(A) of A is defined by setting σS(A) := H r ρS(A). For a bounded linear operator A we can write the resolvent set as ρS(A) = {q ∈ H Rq(A) ∈ G(V R H )} = {q ∈ H Rq(A) has an inverse in B(V R = {q ∈ H ker(Rq(A)) = {0} H )} and ran(Rq(A)) = V R H } and the spectrum can be written as σS(A) = H \ ρS(A) = {q ∈ H Rq(A) has no inverse in B(V R = {q ∈ H ker(Rq(A)) 6= {0} or H )} ran(Rq(A)) 6= V R H } The right S-spectrum σS r (A) and the left S-spectrum σS l (A) are defined respectively as σS r (A) = {q ∈ H Rq(A) in not right invertible in B(V R σS l (A) = {q ∈ H Rq(A) in not left invertible in B(V R H ) }. H ) } The spectrum σS(A) decomposes into three major disjoint subsets as follows: (i) the spherical point spectrum of A: σpS(A) := {q ∈ H ker(Rq(A)) 6= {0}}. (ii) the spherical residual spectrum of A: σrS(A) := {q ∈ H ker(Rq(A)) = {0}, ran(Rq(A)) 6= V R H }. (iii) the spherical continuous spectrum of A: σcS(A) := {q ∈ H ker(Rq(A)) = {0}, ran(Rq(A)) = V R H , Rq(A)−1 /∈ B(V R H ) }. If Aφ = φq for some q ∈ H and φ ∈ V R r {0}, then φ is called an eigenvector of A with right eigenvalue q. The set of right eigenvalues coincides with the point S-spectrum, see [13], proposition 4.5. H FREDHOLM OPERATOR 9 Proposition 3.16. [6, 13] For A ∈ B(V R set and the spectrum σS(A) is a non-empty compact set. H ), the resolvent set ρS(A) is a non-empty open Remark 3.17. For A ∈ B(V R That is, ∂σS(A) = ∂ρS(A) 6= ∅. H ), since σS(A) is a non-empty compact set so is its boundary. Proposition 3.18. Let A ∈ B(V R H ). (3.8) (3.9) σS l (A) = {q ∈ H ran(Rq(A)) is closed or ker(Rq(A)) 6= {0}}. r (A) = {q ∈ H ran(Rq(A)) is closed or ker(Rq(A†)) 6= {0}}. σS H ) and q ∈ H. Set S = Rq(A) ∈ B(V R Proof. Let A ∈ B(V R H ). By proposition 3.9, S is not left invertible if and only if ran(S) is not closed or ker(S) 6= {0}. Thus we have equation 3.8. Again by proposition 3.9, A is not right invertible if and only if ran(S) 6= V R H . ran(S) 6= V R H if and only if ran(S) 6= ran(S) or ker(S†)⊥ 6= V R (cid:3) H if and only if ran(S) 6= ran(S) or ran(S) 6= V R H . That is, ran(S) 6= V R H . Hence we have equation 3.9. 4. Quaternionic Compact Operators A systematic study of quaternionic compact operators has not appeared in the lit- erature. In this regard, in this section, as needed for our purpose, we provide certain significant results about compact right linear operators on V R H . Definition 4.1. Let V R operator K : V R H −→ U R That is, K(U ) is compact in U R for all bounded sequences {φn}∞ subsequence in U R H [11]. H and U R H be right quaternionic Hilbert spaces. A bounded H is compact if K maps bounded sets into precompact sets. ≤ 1}. Equivalently, n=0 has a convergence H , where U = {φ ∈ V R n=1 in V R H kφkV R H the sequence {Kφn}∞ H We denote the set of all compact operators from V R H to U R H will be denoted by B0(V R H by B0(V R H ). H , U R H ) and the compact operators from V R H from V R Proposition 4.2. The following statements are true: (a) If (q, A) ∈ H × B0(V R H , U R H ), then qA, Aq ∈ B0(V R H , U R H ) which are defined by the equations 3.5 and 3.6 respectively. H ) is a vector space under left-scalar multiplication. H , U R H ) and A ∈ B0(V R H , U R H ) such that kAn − Ak −→ 0, then H , U R (b) B0(V R (c) If {An} ⊆ B(V R H ). A ∈ B0(V R (d) If A ∈ B(U R H , U R H ), B ∈ B(U R H ) and K ∈ B0(V R H , U R H ), then AK and KB are compact operators. Proof. Let {φn} ⊆ V R H be a bounded sequence. Then {Aφn} has a convergent sub- sequence. Thus {qAφn} also has a convergent subsequence, and so qA is a compact operator. Now {qφn} ⊆ V R H is also bounded sequence. Thus {(Aq)φn} has a convergent subsequence, as A is compact. Therefore point (a) follows. And point (b) is straightfor- ward. Now suppose that {An} ⊆ B0(V R H ) such that kAn − Ak −→ 0 and let ε > 0. Then by proposition 3.12, it is enough to show that A(U ) is totally bounded, as A(U ) is a complete set, where U = {φ ∈ V R ≤ 1}. Let ε > 0, then H ) and A ∈ B(V R H , U R H , U R H kφkV R H 10 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ there exists N ∈ N such that kA − Ank < ε/6, for all n > N . Since An is compact, there are φj ∈ U : j = 1, 2, 3, · · · , m such that An(U ) ⊆ An(U ) ⊆ m [ i=1 B(Anφj; ε/6); < ε/6}, for all j = 1, 2, 3, · · · , m and where B(Anφj; ε/6) = {Anφ kAnφ − AnφjkU R n ∈ N. Let ψ ∈ A(U ). Then there exists a sequence {φ(ℓ)} ⊆ U such that Aφ(ℓ) −→ ψ as ℓ −→ ∞. So for each (n, ℓ) ∈ N2, there is a φj ∈ U such that kAnφj − Anφ(ℓ)kU R < ε/6 as Anφ(ℓ) ∈ An(U ), for all (n, ℓ) ∈ N2. Therefore, for each (n, ℓ) ∈ N2 with n > N , kAφj − Aφ(ℓ)kU R + kAnφ(ℓ) − Aφ(ℓ)kU R + kAnφj − Anφ(ℓ)kU R H H H H H ≤ kAφj − AnφjkU R < 2kA − Ank + ε/6 < ε/2. H Now taking limit ℓ −→ ∞ both side of the inequality kAφj − Aφ(ℓ)kU R kAφj − ψkU R ≤ ε/2 < ε as the norm k · k is a continuous maps. Thus H H < ε/2, gives A(U ) ⊆ m [ i=1 B(Aφj; ε). This concludes the result of (c). Now suppose that A ∈ B(U R H ). Then for any bounded sequence {φn} ⊆ V R H has a convergent subsequence {Kφnk }, as K is com- pact. Thus {AKφnk} is a convergent subsequence of {AKφn}, because A is continuous. That is AK is compact. In a similar fashion, we can obtain KB is compact. Hence the results hold. (cid:3) H ) and K ∈ B0(V R H , {Kφn} ⊆ U R H ), B ∈ B(V R H , U R Definition 4.3. An operator A : V R is finite dimensional. H −→ U R H is said to be of finite rank if ran(A) ⊆ U R H Lemma 4.4. Following statements are equivalent: (a) The operator A : V R (b) For each φ ∈ V R H −→ U R H is a finite rank operator. H , there exist (vi, ui) ∈ V R H × U R H : i = 1, 2, 3, · · · , n such that (4.1) hui ujiU R H = δij and Aφ = where δij is Kronecker delta. n X i=1 uihvi φiV R H , Proof. Suppose that (a) holds. Then dim(ran(A)) < ∞ and so there exists an orthonor- mal basis ei : i = 1, 2, 3, · · · , n for ran(A). Thus for each φ ∈ V R H , Aφ = n X i=1 eihei AφiU R H = n X i=1 eihA†ei φiV R H . Point (b) follows as taking ui = ei and vi = A†ei. On the other hand, suppose (b) holds. Then it immediately follows that ran(A) = right span over H{ui i = 1, 2, 3, · · · , n} as ui ∈ ran(A), for all i = 1, 2, 3, · · · , n. Assume that ∃ qi ∈ H : i = 1, 2, 3, · · · , n such that n X i=1 uiqi = 0. FREDHOLM OPERATOR 11 Since Pn 1, 2, 3, · · · , n, we have for each j = 1, 2, 3, · · · , n, i=1(A†ui)qi = 0 and there exists φi ∈ V R H such that ui = Aφi, for all i = qj = n X i=1 qihui AφjiU R H = n X i=1 h(A†ui)qi φjiV R H = 0. That is, the set {ui i = 1, 2, 3, · · · , n} is an orthonormal basis for ran(A), and the point (a) holds. Hence the lemma follows. (cid:3) i = 1, 2, 3, · · · , n} is linearly independent. Thus {ui Lemma 4.5. In the lemma 4.4, the set {ui i = 1, 2, 3, · · · , n} of vectors can be chosen to be an orthonormal basis for ran(A), and the adjoint of A can be written as (4.2) A†ψ = n X i=1 vihui ψiU R H , for all ψ ∈ U R 1, 2, 3, · · · , n} of vectors can be chosen to be a basis for ran(A†). H . Furthermore, A†ui = vi for all i = 1, 2, 3, · · · , n and the set {vi i = Proof. In the lemma 4.4, one can trivially observe that the set {ui i = 1, 2, 3, · · · , n} of vectors can be chosen to be an orthonormal basis for ran(A). Now let (φ, ψ) ∈ V R H × U R H , and from equation 4.1, we have hψ AφiU R H = n X i=1 hψ uiiU R H hvi φiV R H . = h Pn This implies that, hA†ψ φiV R and so the equation 4.2 holds. From equation 4.2, it immediately follows that A†ui = vi for all i = 1, 2, 3, · · · , n, = δij. This guarantees that for each i = 1, 2, 3, · · · , n, vi ∈ ran(A†), as hui ujiU R and it suffices, together with (4.2), to say that ran(A†) = right span over H{vi i = 1, 2, 3, · · · , n}. Now assume that i=1 vihψ uiiU R φiV R H H H H ∃ pi ∈ H : i = 1, 2, 3, · · · , n such that n X i=1 vipi = 0. i=1(A†ui)pi = 0, and Pn Then Pn i=1 uipi ∈ ker(A†) = ran(A)⊥ by (a) of proposition 3.4. Thus Pn i = 1, 2, 3, · · · , n} forms a (orthonormal) basis, we have for each i = 1, 2, 3, · · · , n, pi = 0. That is, {vi i = 1, 2, 3, · · · , n} is a linear independent set in ran(A†). This completes the proof. (cid:3) i=1 uipi = 0 as ui ∈ ran(A), for all i = 1, 2, 3, · · · , n. Since {ui Theorem 4.6. Let A ∈ B(V R H , U R H ) be a finite rank operator. Then (a) A is compact. (b) A† ∈ B(U R H , V R H ) is a finite rank operator and dim(ran(A)) = dim(ran(A†)). Proof. Since the operator A : V R dim(ran(A)) < ∞, where U = {φ ∈ V R A(U ) is closed. Now for each φ ∈ U , H −→ U R H is a finite rank, we have dim(A(U )) < ≤ 1}. Hence by proposition 3.3, H kφkV R H kAφkU R H ≤ kAkkφkV R H ≤ kAk. 12 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ That is, A(U ) is closed and bounded set. Hence by proposition 3.12, A(U ) = A(U ) is compact, and so A is a compact operator. This concludes the point (a). From the lemmas 4.4 and 4.5, the point (b) immediately follows. (cid:3) Definition 4.7. Let M ⊂ V R H be a closed subspace, then codim(M ) = dim(V R H /M ). Definition 4.8. Let A : V R and dim(coker(A)) = dim(U R H ) − dim(ran(A)). H −→ U R H be a bounded operator, then coker(A) := U R H /ran(A) Theorem 4.9. If A ∈ B(U R H , V R H ), then the following statements are equivalent: (a) A is compact. (b) There exist finite rank operators An : V R H −→ U R H such that kA − Ank −→ 0 as n −→ ∞. (c) A† is compact. Proof. Suppose that point (a) holds. Then A(U ) is a compact set, where U = {φ ∈ V R ≤ 1}. So A(U ) is separable, and there exists an orthonormal basis {ei i = H kφkV R 1, 2, 3, · · · } for A(U ). Now, for any n ∈ N, define An : V R H −→ U R H by H Anφ = n X i=1 eihA†ei φiV R H , for all φ ∈ V R H . Then by lemma (4.4), An is a finite rank operator, for all n ∈ N. Further, for every φ ∈ U , we have kAφ−Anφk2 eihA†ei φiV R eihei AφiU R hei AφiU R 2 . H k2 U R H H U R = k X i>n Thus kAφ − AnφkU R H H = k X i>n k2 U R H H = X i>n −→ 0 as n −→ ∞, for all φ ∈ U , since for each φ ∈ U , hei AφiU R H 2= X i>n as n −→ ∞. Therefore, ∞ X i=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) hei AφiU R H 2 − n X i=1 hei AφiU R H −→ 0 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) kA − Ank = sup kφkV R =1 H kAφ − AnφkU R H −→ 0 as n −→ ∞. Hence point (b) holds. Suppose that point (b) holds. Then kA† − A† point (c) follows from theorem (4.6) and (c) of proposition (4.2). The implication (c) ⇒ (a) can be obtained, by applying the implications (a) ⇒ (b) and (b) ⇒ (c) for A† and using the fact that A†† = A. Hence the theorem follows. (cid:3) nk = kA − Ank −→ 0 as n −→ ∞. Thus Proposition 4.10. If A ∈ B(V R q ∈ H r {0}. H ) is compact, then dim(ker(qI − A)) < ∞, for all V R H Proof. Let q ∈ H r {0}. Assume that dim(ker(qI sequence {φn} ⊆ ker(qI − A) such that V R H V R H − A)) = ∞. Then there exists a kφnkV R H = 1 and kφn − φmkV R H > 1 2 , for any m, n ∈ N with m 6= n, as ker(qI V R H − A) is separable. Now for each n ∈ N, Aφn = qφn implies FREDHOLM OPERATOR kAφnkV R H = q and kAφn − AφmkV R H > q 2 , for any m, n ∈ N with m 6= n. This suffices to say that A is not compact. This concludes the result. 13 (cid:3) 5. Approximate S-point spectrum Following the complex case, we define the approximate spherical point spectrum of A ∈ B(V R H ) as follows. H ). The approximate S-point spectrum of A, denoted by Definition 5.1. Let A ∈ B(V R σS ap(A), is defined as σS ap(A) = {q ∈ H there is a sequence {φn}∞ Proposition 5.2. Let A ∈ B(V R H ), then σpS(A) ⊆ σS ap(A). n=1 such that kφnk = 1 and kRq(A)φnk −→ 0}. Proof. Let q ∈ σpS(A), then ker(Rq(A)) 6= {0}. That is {φ ∈ D(A2) Rq(A)φ = 0} 6= {0}. Therefore, there exists φ ∈ D(A2) and φ 6= 0 such that Rq(A)φ = 0. Set φn = φ/kφk for all n = 1, 2, · · · . Then kφnk = 1 for all n and kRq(A)φnk −→ 0 as n −→ ∞. Thus q ∈ σS (cid:3) ap(A). Proposition 5.3. If A ∈ B(V R H ) and q ∈ H, then the following statements are equivalent. ap(A). (a) q 6∈ σS (b) ker(Rq(A)) = {0} and ran(Rq(A)) is closed. (c) There exists a constant c ∈ R, c > 0 such that kRq(A)φk ≥ ckφk for all φ ∈ D(A2). Proof. (a)⇒ (c): Suppose (c) fails to hold. Then for every n, there exists non-zero vectors φn such that For each n, let ψn = kRq(A)φnk ≤ kφnk n . , then kψnk = 1 and kRq(A)ψnk < 1 n −→ 0 as n −→ ∞. φn kφnk ap(A). Therefore q ∈ σS (c)⇒ (b): Suppose that there is a constant c > 0 such that kRq(A)φk ≥ ckφk for all φ ∈ D(A2). Therefore, Rq(A)φ 6= 0 for all φ 6= 0 and Rq(A)0 = 0. Therefore ker(Rq(A)) = {0}. Let φ ∈ ran(Rq(A)), where the bar stands for the closure. Then there exists a sequence {φn} ⊆ ran(Rq(A)) such that φn −→ φ as n −→ ∞. From (c) we have a constant c > 0 such that kφn − φmk ≤ {φn} is a Cauchy sequence as {Rq(A)φn} is a Cauchy sequence. Let ψ = lim n→∞ then Rq(A)ψ = lim n→∞ kRq(A)φn − Rq(A)φmk. Therefore φn, Rq(A)φn as Rq is continuous. Therefore Rq(A)ψ = φ and hence φ ∈ ran(Rq(A)), from which we get ran(Rq(A)) ⊆ ran(Rq(A)). Hence ran(Rq(A)) is closed. (b)⇒(a): Suppose that ker(Rq(A)) = {0} and ran(Rq(A)) is closed. Since A ∈ B(V R H ), Rq(A) is bounded and Rq(A) : D(A2) −→ ran(Rq(A)) is a bijection. Therefore, by theorem 3.2, there is a bounded linear operator B : ran(Rq(A)) −→ D(A2) such that BRq(A)φ = φ for all φ ∈ D(A2). Thus, kφk = kBRq(A)φk ≤ kBkkRq(A)φk for all φ ∈ D(A2). Therefore, if there is a sequence {φn} ⊆ D(A2) with kφnk = 1, then 1 ≤ 1 c 14 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ kBkkRq(A)φnk. That is, 1 kBk ≤ kRq(A)φnk. Hence, there is no sequence {φn} ⊆ D(A2) with kφnk = 1 such that kRq(A)φnk −→ 0 as n −→ ∞. Therefore q 6∈ σS ap(A). (cid:3) Theorem 5.4. Let A ∈ B(V R ∂σS(A) ⊆ σS ap(A), where ∂σS(A) is the boundary of σS(A). H ), then σS ap(A) is a non-empty closed subset of H and Proof. Let q ∈ ∂σS(A). Since ρS(A) 6= ∅ and σS(A) is closed, ∂σS(A) = ∂ρS(A) = ρS(A) ∩ σS(A). Therefore, there exists a sequence {qn} ⊆ ρS(A) such that qn −→ q as n −→ ∞. Since, Rqn(A) − Rq(A) = (kqnk − kqk)I V R H − 2(Re(qn) − Re(q))A, for all n, we have Rqn(A) −→ Rq(A) in B(V R Claim: sup kRqn (A)−1k = ∞. n H ) as kqnk −→ kqk and Re(qn) −→ Re(q). Since σS(A) is closed, qn ∈ ρS(A), q ∈ ∂σS(A) and q 6∈ ρS(A) we have Rqn(A) −→ Rq(A) in B(V R Case-I: q ∈ σS kRqn (A)−1k < ∞. Since q ∈ σS H ) and Rq(A) ∈ B(V R ap(A), ker(Rq(A)) 6= {0}. H ) \ G(V R H ). ap(A). Assume sup n Therefore. there exits φ ∈ D(A2) with φ 6= 0 such that Rq(A)φ = 0. Thus, 0 6= kφk = kRqn (A)−1Rqn(A)φk ≤ sup n ≤ sup n kRqn (A)−1k lim sup n kRqn (A)φk kRqn (A)−1kkRq(A)φk = 0, which is a contradiction. Therefore sup kRqn (A)−1k = ∞. n Case-II: q 6∈ σS exists. For each n, we have ap(A). Then ker(Rq(A)) = {0}. Thus Rq(A)−1 : ran(Rq(A)) −→ V R H Rqn(A)−1 − Rq(A)−1 = Rqn (A)−1[Rq(A) − Rqn(A)]Rq(A)−1. kRqn (A)−1k < ∞. Then, since Rqn(A) −→ Rq(A) in B(V R H ), Rqn(A)−1 −→ H ), that is, Rq(A) ∈ G(V R H ), which is a contradiction to kRqn(A)−1k = ∞. The claim is proved. kRqn(A)−1ψk, we can take vectors φn ∈ V R H with kφnk = 1, Assume sup n Rq(A)−1 and Rq(A)−1 ∈ B(V R q ∈ σS(A). Therefore, sup n Since kRqn(A)−1k = sup kψk=1 such that, for each n, kRqn (A)−1k − 1 n ≤ kRqn (A)−1φnk ≤ kRqn(A)−1k. From which we have sup n kRqn (A)−1φnk = ∞, and hence inf n kRqn (A)−1φnk−1 = 0. There- fore, there are subsequences of {φn} and {qn}, call them {φk} and {qk} such that (5.1) kRqk (A)−1φkk−1 −→ 0, as k −→ ∞. Set ψk = Rqk (A)−1φk kRqk (A)−1φkk , then {ψk} is a sequence of unit vectors in V R H with kRqk (A)ψkk = kRqk (A)Rqk (A)−1φkk kRqk (A)−1φkk = kRqk (A)−1φkk−1. FREDHOLM OPERATOR 15 Now, from (5.1) and as qk −→ q, we have kRq(A)ψkk = k(A2 − 2Re(q)A + q2I )ψkk H V R = k(A2 − 2Re(qk)A + qk2I = k(Rqk (A) − (qk2 − q2)I ≤ k(Rqk (A)ψkk + qk2 − q2 + 2Re(qk) − Re(q) kAk = kRqk (A)−1φkk−1 + qk2 − q2 + 2Re(qk) − Re(q) kAk −→ 0 as k −→ ∞. + 2(Re(qk) − Re(q))A)ψkk + 2Re(qk)A − qk2I − 2Re(q)A + q2I )ψkk V R V R V R V R H H H H That is, kRq(A)ψkk −→ 0 as k −→ ∞. Hence q ∈ σS Further, from remark (3.17) ∂σS(A) 6= ∅. Thus σS Finally, take an arbitrary q ∈ H \ σS all φ ∈ D(A2) and p ∈ H, there exists α > 0 (α ∈ R) such that ap(A) and therefore ∂σS(A) ⊆ σS ap(A) 6= ∅. ap(A). ap(A) such that Rq(A) is bounded below. Then, for αkφk ≤ kRq(A)φk ≤ kRp(A)φk + k(p2 − q2)φk + 2k(Re(p) − Re(q))Aφk, which implies (α − p2 − q2 − 2(Re(p) − Re(q))kAk )kφk ≤ kRp(A)φk. Hence Rp(A) is bounded below for each p ∈ H that satisfies α > p2 − q2 + 2(Re(p) − Re(q))kAk. From this we can say that p ∈ H\σS the open ball centered at q with radius α, Bα(q) ⊆ H \ σS open and σS ap(A) for every p which is sufficiently close to q. Hence, ap(A) is (cid:3) ap(A). Therefore H \ σS ap(A) is closed. Theorem 5.5. Let A ∈ B(V R H ) and q ∈ H, then the following statements are equivalent. (a) q 6∈ σS (b) q 6∈ σS (c) q 6∈ σS (d) ran(Rq(A†)) = V R H . ap(A). l (A). r (A†). H V R H ), we have: BRq(A) = I if and only if Rq(A†)B† = I ap(A). Then by proposition (5.3), ker(Rq(A)) = {0}. Hence Proof. For B ∈ B(V R fore (b) and (c) are equivalent. (a)⇒ (b): Suppose q 6∈ σS Rq(A) is invertible, and therefore Rq(A) is left invertible. Thus q 6∈ σS (c)⇒ (d): Since q 6∈ σS C ∈ B(V R Therefore V R (d)⇒ (a): Suppose ran(Rq(A†)) = V R H . Define T : ker(Rq(A†))⊥ −→ V R Rq(A†)φ. Then ker(T ) = {0}, that is, T is bijective and hence invertible. Let r (A†), Rq(A†) is right invertible. Therefore, there is an operator H ) ⊆ ran(Rq(A†)). H ) such that Rq(A†)C = I H = Rq(A†)C(V R H = ran(Rq(A†)). . Hence, V R H by T φ = . There- l (A). V R V R H H C : V R H −→ V R H by Cφ = T −1φ. Then CV R φ ∈ V R H , H = ker(Rq(A†))⊥ and Rq(A†)C = I . Hence, C †Rq(A) = I V R H and, for V R H kφk = kC †Rq(A)φk ≤ kC †k kRq(A)φk, 16 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ from this we get inf{kRq(A)φk kφk = 1} ≥ kC †k−1. Therefore, there is no sequence {φn} ⊆ V R (cid:3) H with kφnk = 1 and kRq(A)φnk −→ 0. Hence q 6∈ σS Corollary 5.6. If A ∈ B(V R H ), then ∂σS(A) ⊆ σS l (A) ∩ σS ap(A). r (A) = σS ap(A) ∩ σS ap(A†)∗. ap(A) = σS ap(A) ⇔ q 6∈ σS r (A) ⇔ q 6∈ σS ap(A†) ⇔ q 6∈ σS Proof. From theorem 5.5, we have q 6∈ σS Also q 6∈ σS q ∈ ∂σS (A), then by theorem 5.4 q ∈ σS 5.5 ran(Rq(A†)) = V R invertible. Therefore Rq(A) is invertible as A ∈ B(V R is a contradiction. Therefore, q ∈ σS q ∈ σS l (A). r (A). Let ap(A†), then by theorem H and by proposition 5.3 ker(Rq(A†)) = {0}. Thus Rq(A†) is H ), and hence q 6∈ σS(A), which r (A). Thus (cid:3) ap(A†), which implies q ∈ σS r (A)∗, and therefore σS l (A), and hence σS l (A). If q 6∈ σS ap(A†)∗ = σS ap(A†)∗ = σS ap(A) = σS l (A) ∩ σS r (A). Following the complex formalism in the following we define the S-compression spec- trum for an operator A ∈ B(V R H ). Definition 5.7. The spherical compression spectrum of an operator A ∈ B(V R by σS c (A), is defined as H ), denoted σS c (A) = {q ∈ H ran(Rq(A)) is not dense in V R H }. Proposition 5.8. Let A ∈ B(V R H ) and q ∈ H. Then, c (A) if and only if q ∈ σpS(A). (a) q ∈ σS c (A) ⊆ σS (b) σS (c) σS(A) = σS r (A). ap(A) ∪ σS c (A). H . If q 6∈ σS H ⇔ ker(Rq(A†)) 6= r (A) and hence σS c (A), then ran(Rq(A)) is not dense in V R c (A) ⇔ ran(Rq(A)) = ker(Rq(A†))⊥ is not dense in V R ap(A†), thus, again by theorem 5.5, ran(Rq(A)) = V R Proof. (a) q ∈ σS {0} ⇔ q ∈ σpS(A†). (b) Let q ∈ σS 5.5 q 6∈ σS Therefore q ∈ σS (c) Let q ∈ σS(A), then Rq(A) is not invertible in B(V R or ran(Rq(A)) 6= V R σS ap(A) ∪ σS Let q ∈ σS ran(Rq(A)) is not closed. Hence, ker(Rq(A)) 6= {0} or ran(Rq(A)) 6= V R σS(A). If q ∈ σS and hence q ∈ σS(A). In summary σS ap(A), then by proposition 5.3 ker(Rq(A)) 6= {0} or H . Thus q ∈ H . Therefore ran(Rq(A)) 6= V R H (cid:3) H ). That is, ker(Rq(A)) 6= {0} ap(A). Hence σS(A) ⊆ r (A), then by theorem H , which is contradiction. c (A), then ran(Rq(A)) is not dense in V R H . Therefore, by proposition 5.3, q ∈ σS c (A) ⊆ σS(A). Hence (c) holds. c (A). ap(A) ∪ σS c (A). If q ∈ σS ap(A) ∪ σS c (A) ⊆ σS r (A). 6. Fredholm operators in the quaternionic setting In the complex setting Fredholm operators are studied in Banach spaces and Hilbert spaces for bounded and even to unbounded linear operators. In this section we shall study the theory for quaternionic bounded linear operators on separable Hilbert spaces. In this regard let V R H be two separable right quaternionic Hilbert spaces. H and U R Definition 6.1. A Fredholm operator is an operator A ∈ B(V R and coker(A) = U R called the codimension, and it is denoted by codim(A). H ) such that ker(A) H /ran(A) are finite dimensional. The dimension of the cokernel is H , U R Proposition 6.2. If A ∈ B(V R H , U R H ) is a Fredholm operator, then ran(A) is closed. FREDHOLM OPERATOR 17 Proof. Restrict A to ker(A)⊥, that is, Aker(A)⊥ : ker(A)⊥ −→ U R H . Then we can assume that ker(A) = {0}. Since A is Fredholm, assume that codim(A) = dim(U R H /ran(A)) < ∞. Therefore, there exist a finite dimensional subspace V ⊆ U R H such that U R H = ran(A)⊕ V . H −→ V ⊥ Since V is finite dimensional, V is closed. Hence U R be the orthogonal projection onto V ⊥. Hence P is a bounded bijection. Let G := P A : ker(A)⊥ −→ V ⊥, which is a composition of two bounded operators and so G is a bounded bijection as A and P are bounded bijections and thereby continuous. Therefore by the open mapping theorem G−1 : V ⊥ −→ ker(A)⊥ is bounded. Let ψ ∈ ran(A). Then, there exist a sequence {φn} ⊂ V R Thus lim n φ := lim n A(φn) = ψ. G(φn) = P (ψ) exists in V ⊥. Hence, H . Therefore, ψ = A(φ) ∈ ran(A). Hence (cid:3) P A(φn) = P (ψ) exists in V ⊥. That is lim n φn = G−1P (ψ) = A−1(ψ) exists in V R H = V ⊕ V ⊥. Let P : U R H such that lim n ran(A) is closed. Definition 6.3. Let A ∈ B(V R the integer, ind(A) = dim(ker(A)) − dim(coker(A)). H , U R H ) be a Fredholm operator. Then the index of A is Remark 6.4. Since ran(A) is closed, we have U R Therefore, coker(A) = U R H /ran(A) ∼= ker(A†). Thus, ind(A) = dim(ker(A)) − dim(ker(A†)). H = ran(A)⊕ran(A)⊥ = ran(A)⊕ker(A†). H , U R Theorem 6.5. Let A ∈ B(V R Then A + K is a Fredholm operator. H ) be bijective, and let K ∈ B0(V R H , U R H ) be compact. Proof. Since A is bijective, A−1 exists. Since the kernel of a bounded right linear operator is closed, ker(A + K) is a Hilbert space. Thus, for φ ∈ ker(A + K), we have Aφ = −Kφ. Let {φn}∞ n=1 ⊆ ker(A + K) be a bounded sequence. Since K is a compact operator, {Kφn}∞ k=1. Since, for all k, φnk ∈ ker(A+K), we have n=1 has a convergence subsequence, {Kφnk}∞ {Kφnk}∞ k=1 = {−Aφnk }∞ k=1. lim lim k→∞ Aφnk = ψ exists. Hence, as A−1 is continuous, we get Therefore, φnk = A−1ψ. That is, any bounded sequence in ker(A + K) has a convergence subsequence. Since an infinite dimensional Hilbert space has an infinite orthonormal sequence with no convergence subsequence, dim(ker(A + K)) < ∞. Since A is invertible and K is compact, A† is invertible and K † is compact. Therefore, by H = ran(A + K) ⊕ ker(A† + the above argument dim(ker(A† + K †)) < ∞. Thus, since U R K †), to show that codim(A + K) < ∞, it is enough to show that ran(A + K) is closed. Split V R H such that H ⊥ ker(A+K), and consider the restriction of A + K to H. Claim: For φ ∈ H, the inequality H = H⊕ker(A+K), where H is a subspace of V R H as V R k−→∞ kφkV R H ≤ ck(A + K)φkU R H holds for some c > 0. If for all c > 0 there exists φ ∈ H such that kφkV R sequences {cn}∞ n=1 ⊆ (0, ∞) and {φn}∞ cn −→ ∞ as n −→ ∞. Further, for all n, H n=1 ⊆ H with kφnkV R H > ck(A + K)φkU R , then there exists = 1 for all n such that H 1 = kφnkV R H > cn(A + K)φnkU R H . 18 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ 1 cn H H < −→ 0 as n −→ ∞. Since K is compact and, for each n, = 1, there exists a convergent subsequence {Kφnk}∞ Thus k(A + K)φnkU R kφnkV R Kφnk −→ ψ ∈ U R A−1Aφnk −→ η = −A−1ψ as k −→ ∞, where η ∈ H and kηkV R all k. But (A + K)η = lim H as k −→ ∞. Then, Aφnk −→ −ψ ∈ U R k=1 of {Kφn}∞ n=1. Assume H as k −→ ∞. Hence φnk = = 1 for 0 = 0. Therefore η ∈ ker(A + K), = 1 as kφnk kV R H H (Aφnk + Kφnk ) = lim k−→∞ k−→∞ which is a contradiction to H ⊥ ker(A+K). Hence the claim holds. Thus, by proposition 3.5, ran(A + K) is closed. Hence codim(A + K) < ∞ and A + K is Fredholm. (cid:3) Proposition 6.6. If A ∈ B(V R H , U R H ) is Fredholm then A† ∈ B(U R H , V R H ) is Fredholm. Proof. Since A is Fredholm, dim(ker(A)) < ∞ and dim(U R proposition 6.2, ran(A) is closed. Therefore, U R Thus, U R 3.4, ker(A) is closed, we have coker(A†) = V R dim(coker(A†)) = dim(ker(A)) < ∞. Hence A† is Fredholm. H /ran(A)) < ∞. Then by H = ran(A)⊕ran(A)⊥ = ran(A)⊕ker(A†). H /ran(A) ∼= ker(A†). Hence dim(ker(A†)) < ∞. Also, since by proposition H /ker(A)⊥ ∼= ker(A) Thus (cid:3) H /ran(A†) = V R Remark 6.7. From propositions 6.2 and 6.6, ran(A†) is closed. Therefore, ind(A†) = dim(ker(A†)) − dim(ker((A†)†)) = dim(ker(A†)) − dim(ker(A)) = −ind(A). H , V R Theorem 6.8. A ∈ B(V R H ) and compact operators K1 and K2, on V R H ) is Fredholm if and only if there exist S1, S2 ∈ B(U R H respectively, such that H , U R S1A = I V R H + K1 H and U R and AS2 = I + K2. U R H Proof. (⇒) Suppose A ∈ B(V R operator H , U R H ) is a Fredholm operator. Then A defines a bijective A : H1 = (ker(A))⊥ −→ H2 = ran(A) = (ker(A†))⊥. H U R H , V R H ) and AS2 = Ai A−1Pr = I H −→ H2 and the inclusion operator i : H1 −→ V R Then A−1 : H2 −→ H1 exists and bounded. Consider the orthogonal projection operator H . Let S2 = i A−1Pr : U R H −→ V R Pr : U R H , then S2 ∈ B(U R − Pa, where Pa is the orthogonal projection operator Pa : U R H −→ ker(A†). Let K2 = −Pa. Since ker(A†) is finite dimensional, K2 is a finite rank operator so it is compact. That is, AS2 = I From proposition 6.6, A† is Fredholm. Therefore, from the above argument, there exists H , U R S3 ∈ B(V R + K3. Hence, + K † S† H ), K1 is compact on 3A = I V R H and S1A = I (⇐) Suppose that there are operators S1, S2 ∈ B(U R V R H and K2 on U R inclusions, H ) and compact operators K1 on H + K2. We have the obvious 3. Set S1 = S† V R H ) and a compact operator on V R H such that A†S3 = I 3, then S1 ∈ B(U R + K1 and AS2 = UR 3 and K1 = K † H such that S1A = I H , V R H , V R + K2. + K1. V R V R V R U R H H H H H (6.1) (6.2) ker(A) ⊆ ker(S1A) = ker(I V R H ran(A) ⊇ ran(AS2) = ran(I U R H + K1) + K2). H V R +K1 and I By theorem 6.5, I 6.1, dim(ker(A)) ≤ dim(I K2) < ∞. Hence A is a Fredholm operator. Remark 6.9. Let A ∈ B(V R H ), then H , U R V R H H U R + K1) < ∞ and by equation 6.2, codim(A) ≤ codim(I +K2 are Fredholm operators. Therefore from equation + (cid:3) U R H FREDHOLM OPERATOR 19 (a) A is said to be left semi-Fredholm if there exists B ∈ B(U R H ) and a compact + K1. The set of all left semi-Fredholm H , V R operator K1 on V R operators are denoted by Fl(V R H such that BA = I H , U R V R H ) [8]. H (b) A is said to be right semi-Fredholm if there exists B ∈ B(U R H ) and a compact + K2. The set of all right semi-Fredholm H , V R operator K2 on U R operators are denoted by Fr(V R H such that AB = I H , U R U R H H ) [8]. (c) By theorem 6.8, the set of all Fredholm operators, F(V R H , U R H ) = Fl(V R H , U R H ) ∩ Fr(V R H , U R H ). (d) From theorem 6.8 it is also clear that every invertible right linear operator is Fredholm. (e) Let SF(V R H ) = Fl(V R H ). From theorem 6.8 and theorem 4.9, we have H ) ⇔ A† ∈ Fr(V R H ) H ) ⇔ A† ∈ SF(V R H ) H ) ⇔ A† ∈ F(V R H ). H ) ∪ Fr(V R A ∈ Fl(V R A ∈ SF(V R A ∈ F(V R H ) is Fredholm if and only if there exists S ∈ B(U R H , U R V R are both finite rank operators. H , V R H ) H Proposition 6.10. A ∈ B(V R such that AS − I and SA − I U R H Proof. (⇒) Suppose A is a Fredholm operator, then ran(A) is closed and A : ker(A)⊥ −→ ran(F ) is a bijection between Hilbert spaces. Let A be the inverse of this bijection. By the open mapping theorem, A ∈ B(U R H , V R H −→ ran(A) be the orthogonal projection onto ran(A). Set S = AP , then = AP A − I H ). Let P : U R = AA − I SA − I = −Pk, V R H V R H V R H where Pk is the orthogonal projection onto ker(A), hence a finite rank operator. Also AS − I U R H = A AP − I U R H = −(I U R H − P ), which is also a finite rank operator (because ran(A)⊥ = ker(A†) is finite dimensional as A† is Fredholm). Therefore, SA − I are finite rank operators and hence compact. (⇐) Now suppose that SA − I rank operators on V R H and U R operators. Therefore by theorem 6.8, A is a Fredholm operator. = F2 with F1 and F2 being finite H respectively. By proposition 4.6, F1 and F2 are compact (cid:3) = F1 and AS − I and AS − I V R V R U R U R H H H H Definition 6.11. Let V0, · · · , Vn be right quaternionic vector spaces and let Aj : Vj −→ Vj+1, 0 ≤ j ≤ n − 1, be right linear operators. Then the sequence An−1−−−→ Vn An−2−−−→ Vn−1 A0−−→ V1 A1−−→ V2 A2−−→ · · · V0 is called exact if ran(Aj) = ker(Aj+1) for j = 0, 1, · · · n − 2. Lemma 6.12. Let Vj be a finite dimensional right quaternionic vector space for each j = 0, 1, · · · , n. If {0} = V0 A0−−→ V1 A1−−→ V2 A2−−→ · · · An−2−−−→ Vn−1 An−1−−−→ Vn = {0} be an exact sequence with dim(Vj) < ∞ for all j = 0, 1, · · · , n, then n−1 X j=0 (−1)j dim(Vj) = 0. 20 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ Proof. For each j, decompose Vj as Vj = ker(Aj) ⊕ Yj, where Yj is an orthogonal com- plement of ker(Aj). The exactness of the sequence implies that Aj : Yj −→ ker(Aj+1) = ran(Aj) is an isomorphism. Therefore, dim(Yj) = dim(ker(Aj+1)) for each j = 0, 1, · · · , n− 1. Hence dim(Vj) = dim(ker(Aj)) + dim(ker(Aj+1)). Further, dim(ker(A0)) = 0 and dim(Vn−1) = dim(ker(An−1)). Hence the result follows. (cid:3) Theorem 6.13. Let V R B(V R is also a Fredholm operator, and it satisfies ind(A2A1) = ind(A1) + ind(A2). H ) are two Fredholm operators, then A2A1 ∈ B(V R H be right quaternionic Hilbert spaces. H ) and A2 ∈ B(U R If A1 ∈ H , W R H ) H and W R H , W R H , U R H , U R Proof. Suppose that A1 and A2 be Fredholm operators. Clearly, ker(A1) ⊆ ker(A2A1) and A1 : ker(A2A1) −→ ker(A2). Therefore, there is an isomorphism S : ker(A2A1)/ker(A1) −→ B ⊆ ker(A2), where B is a subspace of ker(A2). Hence, dim(ker(A2A1)/ker(A1)) ≤ dim(ker(A2)) < ∞, and thus dim(ker(A2A1)) ≤ dim(ker(A1)) + dim(ker(A2)) < ∞. Since ran(A2A1) ⊆ ran(A2) ⊆ W R H , we have (cid:0)W R H /ran(A2A1)(cid:1) / (ran(A2)/ran(A2A1)) = W R H /ran(A2). Hence, (6.3) Further, A2 : U R (6.4) dim(coker(A2A1)) = dim(ran(A2)/ran(A2A1)) + dim(coker(A2)). H /ran(A1) −→ ran(A2)/ran(A2A1) is surjective, and therefore dim(ran(A2)/ran(A2A1)) ≤ dim(coker(A1)) < ∞. From equations 6.3,6.4, we have dim(coker(A2A1)) ≤ dim(coker(A1)) + dim(coker(A2)) < ∞. Thus A2A1 is a Fredholm operator. Let i : ker(A1) −→ ker(A2A1) be the inclusion map, Q : ker(A2) −→ U R the quotient map, and E : W R modulo ran(A2A1) to equivalence classes modulo ran(A2). Consider the sequence H /ran(A1) be H /ran(A2) maps equivalence classes H /ran(A2A1) −→ W R {0} O1−−→ ker(A1) i−→ ker(A2A1) A1−−→ ker(A2) H /ran(A2A1) Q −→ U R E−→ W R A2−−→ W R H /ran(A1) → H /ran(A2) O2−−→ {0}. Since ran(O1) = ker(i) = {0}, ran(i) = ker(A1), ran(A1) = ker(Q), ran(Q) = ker(A2) and ran(A2) = ker(E), the above sequence is exact. Therefore, by lemma 6.12, we have 0 = dim({0}) − dim(ker(A1)) + dim(ker(A2A1)) − dim(ker(A2)) H /ran(A2A1)(cid:1) + dim (cid:0)W R H /ran(A1)(cid:1) − dim (cid:0)W R + dim (cid:0)U R H /ran(A2)(cid:1) H /ran(A2A1)(cid:1)] − [dim(ker(A1)) − dim (cid:0)U R = [dim(ker(A2A1)) − dim (cid:0)W R H /ran(A1)(cid:1)] −[dim(ker(A2)) − dim (cid:0)W R = ind(A2A1) − ind(A1) − ind(A2). H /ran(A2)(cid:1)] Hence the result follows. (cid:3) Corollary 6.14. Let A ∈ B(V R H ) such that AB = F and BA = G are Fredholm operators. Then A and B are Fredholm operators and ind(AB) = ind(A) + ind(B). H ) and B ∈ B(U R H , U R H , V R FREDHOLM OPERATOR 21 Proof. Since ran(F ) ⊆ ran(A) and ker(A) ⊆ ker(G), we have dim(ker(A)) < ∞ and dim(U R H /ran(F )) < ∞. Hence A is Fredholm. Similarly B is also Fredholm. The index equality follows from theorem 6.13. (cid:3) H /ran(A)) ≤ dim(U R Lemma 6.15. Let F ∈ B(V R H ) be a finite rank operator, then ind(I + F ) = 0. V R H Proof. Suppose that F ∈ B(V R H ) be a finite rank operator. Then, dim(ran(F )) < ∞. Also by theorem 4.6, ker(F )⊥ = ran(F †) is also a finite dimensional subspace. Define H = L ⊕ L⊥. L := ran(F ) + ker(F )⊥ with dim(L) < ∞, and hence L is closed. Thus V R Further, (I V R H + F )L ⊆ L + F L ⊆ L, Hence L and L⊥ are invariant under I V R H + F , and therefore, (I V R H + F )L⊥ = I L⊥. ind(I V R H + F ) = ind((I V R H + F )L) + ind((I H V R + F )L⊥) = {0} and L⊥/ran((I Since ker((I Further, since dim(L) < ∞, dim(L) = dim(ker((I and therefore codim((I which concludes the proof. +F )L)) = dim(ker((I V R V R H H H V R H V R + F )L⊥). + F )L⊥) ∼= {0}, ind((I V R +F )L)). Hence, ind(I + F )L)) + dim(ran((I V R H H V R H + F )L⊥) = 0. + F )L)), V R +F )L) = 0, (cid:3) H Theorem 6.16. Let A ∈ B(V R operator K ∈ B(V R H , U R H , U R H ) be a Fredholm operator, then for any compact H ), A + K is a Fredholm operator and ind(A + K) = ind(A). Proof. Let A ∈ B(V R S1, S2 ∈ B(U R S1A = I H , V R V R H H , U R H ) be a Fredholm operator, then by theorem 6.8, there exists H respectively such that H ) and compact operators K1, K2 on V R H , U R + K1 and AS2 = I + K2. Hence we have U R H S1(A + K) = S1A + S1K = I V R H + K1 + S1K = I V R H (A + K)S2 = AS2 + KS2 = I U R H + K2 + KS2 = I U R H + K ′ 1 + K ′ 2, where, by proposition 4.2, K ′ Therefore by theorem 6.8, A + K is a Fredholm operator. Now by proposition 6.10, there exists S ∈ B(U R F2 respectively on V R 1 = K1 + S1K and K ′ H such that H , V R 2 = K2 + KS2 are compact operators. H ) and finite rank operators F1 and H and U R SA = I V R H + F1 and AS = I U R H + F2. Hence by proposition 4.6 and theorem 6.5, SA and AS are Fredholm operators. Thus by corollary 6.14, S is a Fredholm operator. Hence by lemma 6.15 and theorem 6.13, (6.5) 0 = ind(SA) = ind(S) + ind(A). Since S and A + K are Fredholm operators, by theorem 6.13, S(A + K) is a Fredholm operator. Therefore by theorem 6.10, there exists a finite rank operator F3 on V R H such that S(A + K) = I + F3. Thus by lemma 6.15 and theorem 6.13, we have V R H (6.6) 0 = ind(S(A + K)) = ind(S) + ind(A + K). From equations 6.5 and 6.6, we have ind(A + K) = ind(A). (cid:3) Corollary 6.17. Let K ∈ B(V R 0. H ) be compact then I +K is Fredholm and ind(I +K) = V R H 22 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ Proof. Let K ∈ B0(V R theorem 6.16, I V R H is Fredholm and ind(I + K) = ind(I V R H V R H V R H ) = 0. Therefore, by ) = 0. (cid:3) H ). Clearly I V R H + K is Fredholm and ind(I Corollary 6.18. Every invertible operator A ∈ B(V R H ) is Fredholm and ind(A) = 0. Proof. Let A ∈ B(V R is Fredholm. Therefore, by corollary 6.14, A and A−1 are Fredholm operators. Since A is invertible so is A†. Therefore, ker(A) = ker(A†) = {0}. Hence ind(A) = 0. (cid:3) H ) is invertible. Then AA−1 = A−1A = I V R H Corollary 6.19. Let n be a non-negative integer. If A ∈ F(V R ind(An) = n ind(A). H ) then An ∈ F(V R H ) and Proof. The result is trivial for n = 0, 1. n = 2 follows from theorem 6.13, and by induction we have the result. (cid:3) Theorem 6.20. Let A ∈ B(V R stant c > 0 such that for every operator S ∈ B(V R Fredholm operator and ind(A + S) = ind(A). H , U R H ) be a Fredholm operator. Then there exists a con- H ) with kSk < c, A + S is a H , U R H H , U R H , V R H and U R − Pk and AR2 = I H respectively. Let R1, R2 ∈ B(U R − Pk†, where Pk : V R H ) be a Fredholm operator, then by proposition 6.6, A† ∈ Proof. Let A ∈ B(V R B(U R H ) is Fredholm. Therefore, ker(A) and ker(A†) are finite dimensional subspaces of V R H ) be operators such that R1A = H −→ ker(A†) I V R be orthogonal projections. Further Pk and Pk† are finite rank operators hence compact. + R1S − Pk. For kSk < min{kR1k−1, kR2k−1}, we Now R1(A + S) = R1A + R1S = I have kR1Sk < 1. Therefore by proposition 3.6, I + R1S is invertible and by the open H ). Now mapping theorem (I H , V R H −→ ker(A) and Pk† : U R + R1S)−1 ∈ B(V R V R V R U R H H H V R H (I V R H + R1S)−1R1(A + S) = I V R H − (I V R H + R1S)−1Pk = I V R H + K1, where, by proposition 4.2, K1 = −(I Further, (A + S)R2 = AR2 + SR2 = I the same argument as above I V R H U R H + R1S)−1Pk is a compact operator on V R H . + SR2 − Pk†. Again kSR2k < 1, therefore by H ). Hence. + SR2)−1 ∈ B(U R U R H + SR2 is invertible and (I U R H (A + S)R2(I U R H + SR2)−1 = I U R H − (I U R H + SR2)−1Pk† = I U R H + K2, where K2 = −(I S1 = (I and K2, on V R V R H +SR2)−1Pk† is a compact operator on U R H , V R + SR2)−1 ∈ B(U R + R1S)−1R1, S2 = R2(I U R H . That is, we have operators H ) and compact operators K1 H U R H H and U R H respectively, such that S1(A + S) = I V R H + K1 and (A + S)S2 = I U R H + K2. Therefore, by theorem 6.8, A + S is a Fredholm operator. The identity, ind(A + S) = ind(A) follows similar to the proof in theorem 6.16. (cid:3) Theorem 6.21. An operator A ∈ B(V R closed and ker(A) is finite dimensional. Hence H ) is left semi-Fredholm if and only if ran(A) is (6.7) (6.8) Fl(V R Fr(V R H ) = {A ∈ B(V R H ) = {A ∈ B(V R H ) ran(A) is closed and dim(ker(A)) < ∞} H ) ran(A) is closed and dim(ker(A†)) < ∞} FREDHOLM OPERATOR 23 + K) and ran(SA) = ran(I H ) is left semi-Fredholm. Then there exists S ∈ B(V R H ) +K. We have ker(A) ⊆ ker(SA) = + K is Fredholm and + K)) < ∞ and which implies dim(ker(A)) < ∞. By proposition 6.2, + K) is closed. That is, dim(ker(A)) < ∞, dim(A(ker(SA))) < ∞ and ran(SA) Proof. (⇒) Suppose that A ∈ B(V R and a compact operator K in V R ker(I hence dim(ker(I ran(I is closed.Consider the operator, + K). By theorem 6.5, I H such that SA = I V R V R V R V R V R V R H H H H H H (SA)ker(SA)⊥ : ker(SA)⊥ −→ V R H . Since this operator is injective with a closed range, by theorem 3.7, it is bounded below. Hence, there exists c > 0 such that ≤ kSAφkV R for all φ ∈ ker(SA)⊥. ≤ kSkkAφkV R ckφkV R H H H Hence, Aker(SA)⊥ : ker(SA)⊥ −→ V R H is bounded below, therefore by theorem 3.7, Aker(SA)⊥ has a closed range. Since V R ker(SA) ⊕ ker(SA)⊥, we have H = ran(A) = A(V R H ) = A(ker(SA) ⊕ ker(SA)⊥) = A(ker(SA)) ⊕ A(ker(SA)⊥)), which is closed because the direct sum of a closed subspace and a finite dimensional space is closed. (⇐) Suppose that dim(ker(A)) < ∞ and ran(A) is closed. Hence, Aker(A)⊥ : ker(A)⊥ −→ V R H is injective and has closed range, ran(Aker(A)⊥) = ran(A). Therefore by theorem 3.7, it has a bounded inverse. Let Pr : V R H −→ ran(A) be the orthogonal projection onto H ) by T = (Aker(A)⊥)−1Pr. If φ ∈ ker(A), then T Aφ = 0. If ran(A). Define T ∈ B(V R ψ ∈ ker(A)⊥, then T Aψ = (Aker(A)⊥)−1Pr(Aker(A)⊥)ψ = (Aker(A)⊥)−1(Aker(A)⊥)ψ = ψ. Therefore, for every η = φ + ψ ∈ V R H = ker(A) ⊕ ker(A)⊥, where Pn : V R H −→ ker(A)⊥ is the orthogonal projection onto ker(A)⊥. Thus, T Aη = T Aφ + T Aψ = T Aψ = ψ = Pnψ, T A = Pn = I V R H − (I V R H − Pn) = I V R H + K, where K = −(I the finite dimensional subspace ker(A). That is, T A = I hence A is left semi-Fredholm. Therefore, − Pn) is a finite rank operator as it is an orthogonal projection onto + K and K is compact, and V R V R H H Fl(V R H ) = {A ∈ B(V R (6.9) Now by remark 6.9, A† ∈ Fl(V R closed if and only if ran(A†) is closed. Hence, from equation 6.9, we have H ) ran(A) is closed and dim(ker(A)) < ∞}. H ) ⇔ A ∈ Fr(V R H ), and by proposition 3.8, ran(A) is Fr(V R H ) = {A ∈ B(V R H ) ran(A) is closed and dim(ker(A†)) < ∞}. Remark 6.22. Let A ∈ B(V R H ). (a) The so-called Weyl operators are Fredholm operators on V R H with null index. That is, the set of all Weyl operators, (cid:3) W(V R H ) = {A ∈ F(V R H ) ind(A) = 0}. 24 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ (b) Since, by remark 6.9 (e), A ∈ F(V R H ) ⇔ A† ∈ F(V R H ) and ind(A) = −ind(A†), A ∈ W(V R H ) ⇔ A† ∈ W(V R H ). (c) By proposition 6.5 and lemma 6.15, if F is a finite rank operator, then I V R H + F ∈ W(V R H ). H ) ⇒ AB ∈ W(V R H ). (d) By theorem 6.13, A, B ∈ W(V R (e) By theorem 6.16, A ∈ W(V R (f) By corollary 6.18, A ∈ B(V R (g) Suppose dim(V R H ) < ∞, then dim(V R H = ran(A) + ran(A)⊥. Hence dim(V R V R H ), K ∈ B0(V R H ) is invertible, then A ∈ W(V R H ) H ) ⇒ A + K ∈ W(V R H ). H ) = dim(ker(A)) + dim(ran(A)). Also H ) = dim(ran(A)) + dim(ran(A)⊥). Thus, ind(A) = dim(ker(A)) − dim(ker(A†)) = dim(ker(A)) − dim(ran(A)⊥) = dim(ker(A)) + dim(ran(A)) − dim(V R H ) = 0. Therefore, in the finite dimensional space, every operator in B(V R operator with index zero. In this case, W(V R H ) = B(V R H ). H ) is a Fredholm Remark 6.23. In the complex case, the Fredholm theory can also be extended to un- bounded operators using the graph norm. For such a complex treatment we refer the reader to, for example, [14]. Following the complex theory we may be able to extend the quaternionic Fredholm theory to quaternionic unbounded operators. 7. essential S-spectrum In this section we define and study the essential S-spectrum in V R H in terms of the Fredholm operators. In particular, we give an interesting characterization to the S- spectrum in terms of the Fredholm operators and the Fredholm index (see proposition 7.18). Theorem 7.1. [13] Let V R H be a right quaternionic Hilbert space equipped with a left scalar multiplication. Then the set B(V R H ) equipped with the point-wise sum, with the left and right scalar multiplications defined in equations 3.5 and 3.6, with the composition as product, with the adjunction A −→ A†, as in 3.2, as ∗− involution and with the norm defined in 3.1, is a quaternionic two-sided Banach C ∗-algebra with unity I . V R H Remark 7.2. In the above theorem, if the left scalar multiplication is left out on V R H , then B(V R H ) becomes a real Banach C ∗-algebra with unity I . V R H Theorem 7.3. The set of all compact operators, B0(V R and is closed under adjunction. H ) is a closed biideal of B(V R H ) Proof. The theorem 7.1 gives together with proposition 4.2 that B0(V R of B(V R Hence the theorem holds. H ). Theorem 4.9 is enough to conclude that B0(V R H ) is a closed biideal H ) is closed under adjunction. (cid:3) On the quotient space B(V R H )/B0(V R H ) the coset of A ∈ B(V R H ) is [A] = {S ∈ B(V R H ) S = A + K for some K ∈ B0(V R H )} = A + B0(V R H ). On the quotient space define the product [A][B] = [AB]. FREDHOLM OPERATOR 25 H ), with the above product, B(V R H ) is ]. We call this algebra the quaternionic Calkin H )/B0(V R H ) is a closed subspace of B(V R Since B0(V R a unital Banach algebra with unit [I algebra. Define the natural quotient map H )/B0(V R H ) −→ B(V R π : B(V R V R H H ) by π(A) = [A] = A + B0(V R H ). Note that [0] = B0(V R H ) and hence ker(π) = {A ∈ B(V R H ) π(A) = [0]} = B0(V R H ). Since B0(V R H ) is an ideal of B(V R (a) π(A + B) = (A + B) + B0(V R (b) π(AB) = AB + B0(V R (c) π(I ) = [I ]. V R H V R H H ), for A, B ∈ B(V R H ) = (A + B0(V R H ), we have H )) + (B + B0(V R H ) = (A + B0(V R H ))(B + B0(V R H )) = π(A)π(B). H )) = π(A) + π(B). Hence π is a unital homomorphism. The norm on B(V R H )/B0(V R H ) is given by k[A]k = inf K∈B0(V R H ) kA + Kk ≤ kAk. Therefore π is a contraction. Theorem 7.4. Let A ∈ B(V R H ). The following are pairwise equivalent. H ). (a) A ∈ Fl(V R (b) There exist S ∈ B(V R (c) There exist S ∈ B(V R H ) and K ∈ B0(V R H ) and K1, K2, K3 ∈ B0(V R H ) such that SA = I + K. V R H H ) such that (S + K1)(A + K2) = I V R H + K3. (d) π(S)π(A) = π(I (e) π(A) is left invertible in B(V R ), the identity in B(V R H ). H )/B0(V R V R H H )/B0(V R H ), for some π(S) ∈ B(V R H )/B0(V R H ). H H V R V R ] = π(I + K + SK2 + K1A + K1K2 = I Proof. By the definition of left semi-Fredholm operators (a) and (b) are equivalent. (b)⇒ (c): Suppose (b) holds. Since (S + K1)(A + K2) = SA + SK2 + K1A + K1K2 = + K3, where, by proposition 4.2, K3 = K + SK2 + I V R K1A + K1K2 is compact. (c)⇒ (d): Suppose (c) holds. Then there exists S1 ∈ [S] = π(S), A1 ∈ [A] = π(A) and J ∈ [I ) such that S1A1 = J. Hence π(S)π(A) = [S][A] = [SA] = [J] = [I (d)⇒ (b): Suppose (d) holds. B ∈ π(S)π(A) = π(SA) if and only if B = SA + K1 for some K1 ∈ Bo(V R + K2 for some K2 ∈ Bo(V R H ). Thus A ∈ Fl(V R (a) and (e) are equivalent by the definition of left invertibility. +K, where K = K2−K1 ∈ B0(V R ). Therefore (d) is established. V R +K2−K1 = I H ). Therefore, SA = I ] if and only if B = I H ). Also B ∈ π(I ] = π(I ) = [I H ). V R V R V R V R V R V R V R (cid:3) H H H H H H H H H H ). A ∈ Fl(V R Corollary 7.5. Let A ∈ B(V R left (or right) invertible in the Calkin algebra B(V R H )/B0(V R Proof. It is straightforward from the theorem 7.4 for Fl(V R theorem 7.4 holds for Fr(V R H ) (or A ∈ Fr(V R H ). H ). H )) if and only if π(A) is H ). Also similar version of (cid:3) Definition 7.6. The essential S-spectrum (or the Calkin S-spectrum) σS B(V R H )/B0(V R H ) is the S-spectrum of π(A) in the unital Banach algebra B(V R e (A) of A ∈ H ). That is, σS e (A) = σS(π(A)). 26 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ Similarly, the left essential S-spectrum σS are the left and right S-spectrum of π(A) respectively. That is, el(A) and the right essential S-spectrum σS er(A) σS el(A) = σS l (π(A)) and σS er(A) = σS r (π(A)) H ). H )/B0(V R in B(V R Clearly, by definition, σS e (A) = σS Proposition 7.7. Let A ∈ B(V R el(A) ∪ σS H ), then er(A) and σS e (A) is a compact subset of H. (7.1) (7.2) el(A) = {q ∈ H Rq(A) ∈ B(V R σS σS er(A) = {q ∈ H Rq(A) ∈ B(V R H ) \ Fl(V R H ) \ Fr(V R H )} H )} H ). Then by corollary 7.5, Rq(A) 6∈ Fl(V R Proof. Let A ∈ B(V R not left invertible in B(V R q ∈ σS equation 7.1. A similar argument proves equation 7.2. H ) if and only if π(Rq(A)) is H ), which means, by the definition of left S-spectrum, el(A) if and only if Rq(A) 6∈ Fl(V R H ). Hence we have (cid:3) l (π(A)). That is, q ∈ σS H )/B0(V R Corollary 7.8. (Atkinson theorem) Let A ∈ B(V R H ), then (7.3) σS e (A) = {q ∈ H Rq(A) ∈ B(V R H ) \ F(V R H )}. Proof. Let A ∈ B(V R e (A) = σS it is straightforward from proposition 7.7. H ). Since σS el(A) ∪ σS er(A) and F(V R H ) = Fl(V R H ) ∩ Fr(V R H ), (cid:3) Corollary 7.9. (Atkinson theorem) Let A ∈ B(V R is invertible in the Calkin algebra B(V R H ). H )/B0(V R H ). A is Fredholm if and only if π(A) Proof. A ∈ B(V R π(A) is invertible in B(V R in B(V R H ) is Fredholm if and only if A is left and right semi-Fredholm. Further, H ) if and only if π(A) is both left and right invertible (cid:3) H ). Therefore, from corollary 7.5, we have the desired result. H )/B0(V R H )/B0(V R Proposition 7.10. Let A ∈ B(V R H ), then σS el(A) and σS er(A) are closed subsets of H. Proof. In the unital algebra B(V R σS r (A) = σS B(V R ap(A) and ap(A†) are closed subsets of H. By the same argument, in the Calkin algebra r (π(A)) are closed subsets of H. (cid:3) H ), by theorems 5.4 and 5.5, σS l (A) = σS er(A) = σS H )/B0(V R el(A) = σS H ), σS Proposition 7.11. Let A ∈ B(V R l (π(A)) and σS H ), then (7.4) (7.5) σS el(A) = {q ∈ H ran(Rq(A)) is not closed or dim(ker(Rq(A))) = ∞}. σS er(A) = {q ∈ H ran(Rq(A)) is not closed or dim(ker(Rq(A†))) = ∞}. H ). By proposition 7.7, q ∈ σS Proof. Let A ∈ B(V R el(A) if and only if Rq(A) is not left semi-Fredholm. By theorem 6.21, Rq(A) is not left semi-Fredholm if and only if ran(Rq(A)) is not closed or dim(ker(Rq(A))) = ∞. Hence we have equation 7.4. In the same way equation 7.5 can be obtained. (cid:3) Corollary 7.12. Let A ∈ B(V R H ), then σS er(A) = σS el(A) = σS σS er(A†), el(A†) and hence σS e (A) = σS e (A†). Proof. Since Rq(A) = Rq(A), by proposition 3.8, ran(Rq(A)) is closed if and only if ran(Rq(A†)) is closed, and σS el(A), proposition 7.11 conclude the results. (cid:3) e (A) = σS er(A)∪σS FREDHOLM OPERATOR 27 Remark 7.13. If dim(V R and hence compact. Therefore, the Calkin algebra B(V R case, σS H ) < ∞, then all operators in B(V R e (A) = ∅. H ) are finite rank operators H ) is null. Hence, in this H )/B0(V R Proposition 7.14. For A ∈ B(V R H ), σS e (A) 6= ∅ if and only if dim(V R H ) = ∞. H ) = ∞. B(V R Proof. (⇐) Suppose dim(V R H ) is a unital Banach algebra. By proposition 3.16, the S-spectrum on a unital Banach algebra is not empty. Therefore σS e (A) = σS(π(A)) 6= ∅. (⇒) By remark 7.13, dim(V R dim(V R e (A) 6= ∅ implies (cid:3) e (A) = ∅. That is, σS H ) < ∞ implies σS H )/B0(V R H ) = ∞. Proposition 7.15. For every A ∈ B(V R In the same way, σS el(A + K) = σS el(A) and σS er(A + K) = σS er(A). H ) and K ∈ B0(V R H ), we have σS e (A+K) = σS e (A). Proof. For every A ∈ B(V R Therefore by the definition, we get H ) and K ∈ B0(V R H ) we have π(A+K) = π(A) in B(V R H )/B0(V R H ). σS e (A + K) = σS(π(A + K)) = σS(π(A)) = σS e (A). The other two equalities follow in the same way. Proposition 7.16. Let A ∈ B(V R H ). Then, σS el(A) ⊆ σS l (A) and hence σS e (A) ⊆ σS(A). and σS er(A) ⊆ σS r (A), Proof. It is straightforward from proposition 3.18 and 7.11. Definition 7.17. Let A ∈ B(V R H ) and k ∈ Z \ {0}. Define, σS k (A) = {q ∈ H Rq(A) ∈ F(V R H ) and ind(Rq(A)) = k}. Also σS 0 = {q ∈ σS(A) Rq(A) ∈ W(V R H )}. (cid:3) (cid:3) Proposition 7.18. Let A ∈ B(V R H ), then σS(A) = σS e (A) ∪ [ k∈Z σS k (A). k (A)}k∈Z is pair-wise disjoint. Let 0 6= k ∈ Z and q ∈ σS Proof. Clearly the family {σS k (A). If k > 0, then 0 < ind(Rq(A)) < ∞, hence 0 < dim(ker(Rq(A)))−dim(ker(Rq(A†))) < ∞. Thus, ker(Rq(A)) 6= {0} and therefore q ∈ σS(A). If k < 0, then dim(ker(Rq(A))) < dim(ker(Rq(A†))). Hence ker(Rq(A†)) 6= {0} and therefore Rq(A†) is not invertible. Thus, by proposition 3.9, Rq(A) is not invertible. That is, q ∈ σS(A). Altogether we get k (A) = {q ∈ σS(A) Rq(A) ∈ F(V R σS H )}. [ k∈Z Also, by proposition 7.16, we have, σS {q ∈ σS(A) Rq(A) 6∈ F(V R H )}. Therefore, σS(A) = σS e (A) ⊆ σS(A) and by corollary 7.8, σS e (A) = σS k (A), clearly with e (A) ∪ [ k∈Z e (A) ∩ [ σS k∈Z σS k (A) = ∅. (cid:3) 28 B. MURALEETHARAN† AND K. THIRULOGASANTHAR‡ 8. conclusion We have studied the approximate S-point spectrum, Fredholm operators and essential S-spectrum of a bounded right linear operator on a right quaternionic Hilbert space V R H . The left multiplication defined on V R H does not play a role in the approximate S-point spectrum and in the Fredholm theory. However, in getting a unital Calkin algebra it has appeared and hence played a role in the S-essential spectrum. Also an interesting characterization to the S-spectrum is given in terms of the Fredholm operators and its index. In the complex case, the exact energy spectrum is obtainable and known only for a handful of Hamiltonians, and for the rest only approximate spectrum is computed numerically. In these approximate theories the Fredholm operators and the essential spectra are used frequently. Further, these theories are successfully used in transport operator, integral operator and differential operator problems [3, 12, 15, 16, 18, 8]. In the same manner, the theories developed in this note may be useful in obtaining the approximate S-point spectrum or essential S-spectrum for quaternionic Hamiltonians and their small norm perturbations. For quaternionic Hamiltonians and potentials see [1, 9] and the references therein. 9. Acknowledments K. Thirulogasanthar would like to thank the FRQNT, Fonds de la Recherche Nature et Technologies (Quebec, Canada) for partial financial support under the grant number 2017-CO-201915. Part of this work was done while he was visiting the University of Jaffna to which he expresses his thanks for the hospitality. References [1] Adler, S.L., Quaternionic Quantum Mechanics and Quantum Fields, Oxford University Press, New York, 1995. [2] Alpay, D., Colombo, F., Kimsey, D.P., The spectral theorem for quaternionic unbounded normal operators based on the S-spectrum, J. Math. Phys. 57 (2016), 023503. [3] Benharrat, M., Comparison between the different definitions of the essential spectrum and ap- plications, Ph.D. thesis (2013), University of Oran. [4] Colombo, F., Sabadini, I., On Some Properties of the Quaternionic Functional Calculus, J. Geom. Anal., 19 (2009), 601-627. [5] Colombo, F., Sabadini, I., On the Formulations of the Quaternionic Functional Calculus, J. Geom. Phys., 60 (2010), 1490-1508. [6] Colombo, F., Gentili, G., Sabadini, I., Struppa, D.C., Non commutative functional calculus: Bounded operators, Complex Analysis and Operator Theory, 4 (2010), 821-843. [7] Colombo, F., Sabadini, I., Struppa, D.C., Noncommutative Functional Calculus, Birkhauser Basel, 2011. [8] Conway, J.B., A course in functional analysis, 2nd Ed., Springer-Verlag, New York (1990). [9] De Leo, S., Giardino, S.,Dirac solutions for quaternionic potentials, J. Math. Phys.55 (2014), 022301. [10] Diagana, T., Almost Automorphic Type and Almost Periodic Type Functions in Abstract Spaces, Springer International Publishing : Imprint: Springer, 2013. [11] Fashandi, M., Compact operators on quaternionic Hilbert spaces, Ser. Math. Inform. 28 (2013), 249-256. [12] Garding, L., Essential spectrum of Schrodinger operators, J. Funct. Anal. 52 (1983), 1-10. [13] Ghiloni, R., Moretti, W. and Perotti, A., Continuous slice functional calculus in quaternionic Hilbert spaces, Rev. Math. Phys. 25 (2013), 1350006. FREDHOLM OPERATOR 29 [14] Gohberg, I., Goldberg, S., Kaashoek, M.A., Basic classes of linear operators, Springer, Basel AG (2003). [15] Jeribi, A., Spectral theory and applications of linear operators and block operator matrices, Springer International Publishing, Switzerland (2015). [16] Jeribi, A., Mnif, M., Essential spectra and application to transport equations, Acta Applicandae Mathematica, 89 (2005), 155-176. [17] Kreyszig, E. O., Introductory functional analysis with applications, John Wiley & Sons. Inc, 1978. [18] Kubrusly, C.S., Spectral theory of operators on Hilbert spaces, Springer, New York (2012). [19] Muraleetharan, B., Thirulogasanthar, K., Deficiency Indices of Some Classes of Unbounded H-Operators, Complex Anal. Oper. Theory (2017), 1-29. https://doi.org/10.1007/s11785-017- 0702-4. [20] Muraleetharan, B, Thirulogasanthar, K., Coherent state quantization of quaternions, J. Math. Phys., 56 (2015), 083510. [21] Muscat, J., Functional Analysis: An Introduction to Metric Spaces, Hilbert Spaces, and Banach Algebras, Springer, 2014. [22] Rudin, W., Functional Analysis, International Series in Pure and Applied Mathematics, McGraw-Hill Inc., New York, 1991. [23] Viswanath, K., Normal operators on quaternionic Hilbert spaces, Trans. Amer. Math. Soc. 162 (1971), 337-350. [24] Willians, V., Closed Fredholm and semi-Fredholm operators, essential spectra and perturbations, J. Funct. Anal. 20 (1975), 1-25. † Department of mathematics and Statistics, University of Jaffna, Thirunelveli, Sri Lanka. ‡ Department of Computer Science and Software Engineering, Concordia University, 1455 De Maisonneuve Blvd. West, Montreal, Quebec, H3G 1M8, Canada. E-mail address: [email protected] and [email protected]
1204.0606
1
1204
2012-04-03T06:48:27
Anisotropic Function Spaces on Singular Manifolds
[ "math.FA", "math.AP", "math.DG" ]
A rather complete investigation of anisotropic Bessel potential, Besov, and H\"older spaces on cylinders over (possibly) noncompact Riemannian manifolds with boundary is carried out. The geometry of the underlying manifold near its 'ends' is determined by a singularity function which leads naturally to the study of weighted function spaces. Besides of the derivation of Sobolev-type embedding results, sharp trace theorems, point-wise multiplier properties, and interpolation characterizations particular emphasize is put on spaces distinguished by boundary conditions. This work is the fundament for the analysis of time-dependent partial differential equations on singular manifolds.
math.FA
math
Anisotropic Function Spaces on Singular Manifolds H. Amann∗ Math. Institut, Universitat Zurich, Winterthurerstr. 190, CH -- 8057 Zurich, Switzerland Key words Anisotropic weighted Sobolev spaces, Bessel potential spaces, Besov spaces, Holder spaces, non- complete Riemannian manifolds with boundary, embedding theorems, traces, interpolation, boundary operators Subject classification 46E35, 54C35, 58A99, 58D99, 58J99 A rather complete investigation of anisotropic Bessel potential, Besov, and Holder spaces on cylinders over (possibly) noncompact Riemannian manifolds with boundary is carried out. The geometry of the underlying manifold near its 'ends' is determined by a singularity function which leads naturally to the study of weighted function spaces. Besides of the derivation of Sobolev-type embedding results, sharp trace theorems, point-wise multiplier properties, and interpolation characterizations particular emphasize is put on spaces distinguished by boundary conditions. This work is the fundament for the analysis of time-dependent partial differential equations on singular manifolds. 1 Introduction In [5] we have performed an in-depth study of Sobolev, Bessel potential, and Besov spaces of functions and tensor fields on Riemannian manifolds which may have a boundary and may be noncompact and noncomplete. That as well as the present research is motivated by -- and provides the basis for -- the study of elliptic and parabolic boundary value problems on piece-wise smooth manifolds, on domains in Rm with a piece-wise smooth boundary in particular. A singular manifold M is to a large extent determined by a 'singularity function' ρ ∈ C∞(cid:0)M, (0,∞)(cid:1). The behavior of ρ at the 'singular ends' of M , that is, near that parts of M at which ρ gets either arbitrarily small or arbitrarily large, reflects the singular structure of M . The basic building blocks for a useful theory of function spaces on singular manifolds are weighted Sobolev spaces based on the singularity function ρ. More precisely, we denote by K either R or C. Then, given k ∈ N, λ ∈ R, and p ∈ (1,∞), the weighted Sobolev space W k,λ (M, K) is the completion of D(M ), the space of smooth functions with compact support in M , in L1,loc(M ) with respect to the norm (M ) = W k,λ p p 2 1 0 2 r p A 3 ] . A F h t a m [ 1 v 6 0 6 0 . 4 0 2 1 : v i X r a p(cid:17)1/p u 7→(cid:16) kXi=0(cid:13)(cid:13)ρλ+i ∇iug(cid:13)(cid:13)p . (1.1) Here ∇ denotes the Levi-Civita covariant derivative and ∇iug is the 'length' of the covariant tensor field ∇iu naturally derived from the Riemannian metric g of M . Of course, integration is carried out with respect to the volume measure of M . It turns out that W k,λ (M ) is well-defined, independently -- in the sense of equivalent norms -- of the representation of the singularity structure of M by means of the specific singularity function. p A very special and simple example of a singular manifold is provided by a bounded smooth domain whose boundary possesses a conical point. More precisely, suppose Ω is a bounded domain in Rm whose topolog- ical boundary, bdry(Ω), contains the origin, and Γ := bdry(Ω)\{0} is a smooth (m − 1)-dimensional sub- manifold of Rm lying locally on one side of Ω. Also suppose that Ω ∪ Γ is near 0 diffeomorphic to a cone ∗ e-mail: [email protected] 2 H. Amann: Anisotropic Function Spaces on Singular Manifolds { ry ; 0 < r < 1, y ∈ B }, where B is a smooth compact submanifold of the unit sphere in Rm. Then, en- dowing M := Ω ∪ Γ with the Euclidean metric, we get a singular manifold with a single conical singularity, as considered in [35] and [27], for example. In this case the weighted norm (1.1) is equivalent to u 7→(cid:16)Xα≤k krλ+α∂αukp Lp(Ω)(cid:17)1/p , where r(x) is the Euclidean distance from x ∈ M to the origin. Moreover, W k,λ (M ) coincides with the space p,λ+k(Ω) employed by S.A. Nazarov and B.A. Plamenevsky [35, p. 319] and, in the case p = 2, by V.A. Kozlov, V k V.G. Maz′ya, and J. Rossmann (see Section 6.2 of [27], for instance). p In [5] we have exhibited a number of examples of singular manifolds. For more general classes, comprising notably manifolds with corners and non-smooth cusps, we refer to H. Amann [6]. It is worthwhile to point out that our concept of singular manifolds encompasses, as a very particular case, manifolds with bounded geometry (that is, Riemannian manifolds without boundary possessing a positive injectivity radius and having all covariant derivatives of the curvature tensor bounded). In this case we can set ρ = 1, the function constantly equal to 1, so that W k,λ (M ) is independent of λ and equal to the standard Sobolev space W k p (M ). p The weighted Sobolev spaces W k,λ (M ) and their fractional order relatives, that is, Bessel potential and Besov spaces, come up naturally in, and are especially useful for, the study of elliptic boundary value problems for differential and pseudodifferential operators in non-smooth settings. This is known since the seminal work of V.A. Kondrat′ev [22] and has since been exploited and amplified by numerous authors in various levels of generality, predominantly however in the Hilbertian case p = 2 (see [5] for further bibliographical remarks). p For an efficient study of evolution equations on singular manifolds we have to have a good understanding of function spaces on space-time cylinders M × J with J ∈ {R, R+}, where R+ = [0,∞). Then, in general, the functions (or distributions) under consideration posses different regularity properties with respect to the space and time variables. Thus we are led to study anisotropic Sobolev spaces and their fractional order relatives. Anisotropic weighted Sobolev spaces depend on two additional parameters, namely r ∈ N× := N\{0} and µ ∈ R. More precisely, we denote throughout by ∂ = ∂t the vector-valued distributional 'time' derivative. Then, given k ∈ N×, W (kr,k),(λ,µ) p (M × J) is the linear subspace of L1,loc(M × J) consisting of all u satisfying ρλ+i+jµ ∇i∂jug ∈ Lp(M × J) endowed with its natural norm. for i + jr ≤ kr, (1.2) It is a Banach space, a Hilbert space if p = 2. Spaces of this type, as well as fractional order versions thereof, provide the natural domain for an Lp-theory of linear differential operators of the form Xi+jr≤kr aij · ∇i∂j, where aij is a time-dependent contravariant tensor field of order i and · indicates complete contraction. In this connection the values µ = 0, µ = 1, and µ = r are of particular importance. If µ = 1, then space and time derivatives carry the same weight. If also r = 1, then we get isotropic weighted Sobolev spaces on M × J. If µ = 0, then the intersection space characterization W (kr,k),(λ,0) p (M × J) . = Lp(cid:0)J, W kr,λ p (M )(cid:1) ∩ W k p(cid:0)J, Lλ p(M )(cid:1) is valid, where . = means: equal except for equivalent norms. Spaces of this type (with k = 1) have been used by S. Coriasco, E. Schrohe, and J. Seiler [9], [10] for studying parabolic equations on manifolds with conical points. In this case ρ is (equivalent to) the distance from the singular points. Anisotropic spaces with µ = 0 are also important for certain classes of degenerate parabolic boundary value problems (see [6]). 3 The spaces W (kr,k),(λ,r) p p (M × J) constitute, perhaps, the most natural extension of the 'stationary' spaces (M ) to the space-time cylinder M × J. They have been employed by V.A. Kozlov [23] -- [26] -- in the W k,λ Hilbertian setting p = 2 -- for the study of general parabolic boundary value problems on a cone M . (Kozlov, as well as the authors mentioned below, write W (kr,k) (M × J) oc- curs in the works on second order parabolic equations on smooth infinite wedges by V.A. Solonnikov [45] and A.I. Nazarov [34] (also see V.A. Solonnikov and E.V. Frolova [46], [47]), as well as in the studies of W.M. Zaj aczkowski [28], [29], and K. Pileckas [36] -- [38] (see the ref- ' erences in these papers for earlier work) on Stokes and Navier-Stokes equations. In all these papers, except the ones of Pileckas, ρ is the distance to the singularity set, where in Zaj aczkowski's publications M is obtained from ' a smooth subdomain of Rm by eliminating a line segment. Pileckas considers subdomains of Rm with outlets to infinity and ρ having possibly polynomial or exponential growth. aczkowski [52] -- [55], A. Kubica and W.M. Zaj ' λ+kr for W (kr,k),(λ,r) .) The space W (2,1),(λ,2) 2 p In this work we carry out a detailed study of anisotropic Sobolev, Bessel potential, Besov, and Holder spaces on singular manifolds and their interrelations. Besides of this introduction, the paper is structured by the following sections on whose principal content we comment below. Isotropic Bessel Potential and Besov Spaces 2 Vector Bundles 3 Uniform Regularity 4 Singular Manifolds 5 Local Representations 6 7 The Isotropic Retraction Theorem 8 Anisotropic Bessel Potential and Besov Spaces 9 The Anisotropic Retraction Theorem 10 Renorming of Besov Spaces 11 Holder Spaces in Euclidean Settings 12 Weighted Holder Spaces Point-Wise Multipliers 13 14 Contractions 15 Embeddings 16 Differential Operators 17 Extensions and Restrictions 18 Trace Theorems 19 20 Boundary Operators 21 22 Bounded Cylinders Interpolation Spaces With Vanishing Traces We have already pointed out in [5] that it is not sufficient to study function spaces on singular manifolds since spaces of tensor fields occur naturally in applications. In order to pave the way for a study of systems of differen- tial and pseudodifferential operators it is even necessary to deal with tensor fields taking their values in general vector bundles. This framework is adopted here. Sections 2 and 3 are of preparatory character. In the former, besides of fixing notation and introducing conven- tions used throughout, we present the background material on vector bundles on which this paper is based. We emphasize, in particular, duality properties and local representations which are fundamental for our approach. Since we are primarily interested in noncompact manifolds we have to impose suitable regularity conditions 'at infinity'. This is done in Section 3 where we introduce the class of 'fully uniformly regular' vector bundles. They constitute the 'image bundles' for the tensor fields on the singular manifolds which we consider here. After these preparations, singular manifolds are introduced in Section 4. There we also install the geometrical frame which we use from thereon without further mention. Although we study spaces of tensor fields taking their values in uniformly regular vector bundles, the vector bundles generated by these tensor fields are not uniformly regular themselves, in general. In fact, their metric and their covariant derivative depend on the metric g of the underlying singular Riemannian manifold. Since the singularity behavior of g is controlled by the singularity function ρ, due to our very definition of a singular manifold, we have to study carefully the dependence of all relevant parameters on ρ as well. This is done in Section 5. On its basis we can show in later sections that the various function spaces are independent of particular representations; they depend on the underlying geometric structure only. Having settled these preparatory problems we can then turn to the main subject of this paper, the study of function spaces (more precisely, spaces of vector-bundle-valued tensor fields) on singular manifolds. We begin in Sections 6 and 7 by recalling and amplifying some results from our previous paper [5] on isotropic spaces. On the one hand this allows us to introduce some basic concepts and on the other hand we can point out the changes which have to be made to cover the more general setting of of vector-bundle-valued tensor fields. The actual study of anisotropic weighted function spaces begins in Section 8. First we introduce Sobolev spaces which can be easily described invariantly. They form the building blocks for the theory of anisotropic 4 H. Amann: Anisotropic Function Spaces on Singular Manifolds weighted Bessel potential and Besov spaces. The latter are invariantly defined by interpolation between Sobolev spaces and by duality. This being done, it has to be shown that these spaces coincide in the most simple situation in which M is either the Euclidean space Rm or a closed half-space Hm thereof with the 'usual' anisotropic Bessel potential and Besov spaces, respectively. In the Euclidean model setting a thorough investigation has been carried out in H. Amann [4] by means of Fourier analytic techniques. That work is the fundament upon which the present research is built. The basic result which settles this identification and is fundamental for the whole theory as well as for the study of evolution equations is Theorem 9.3. In particular, it establishes isomorphisms between the function spaces on M × J and certain countable products of corresponding spaces on model manifolds. By these isomorphisms we can transfer the known properties of the 'elementary' spaces on Rm × J and Hm × J to M × J. With this method we establish the most fundamental properties of anisotropic Bessel potential and Besov spaces which are already stated in Section 8. In Section 10 we take advantage of the fact that the anisotropic spaces we consider live in cylinders over M so that the 'time variable' plays a distinguished role. This allows us to introduce some useful semi-explicit equivalent norms for Besov spaces. It is well-known that spaces of Holder continuous functions are intimately related to the theory of partial differential equations on Euclidean spaces. They occur naturally, even in the Lp-theory, as point-wise multiplier spaces, in particular as coefficient spaces for differential operators. Although it is fairly easy to study Holder continuous functions on subsets of Rm, it is surprisingly difficult to do this on manifolds. Our approach to this problem is similar to the way in which we defined Bessel potential and (Lp-based) Besov spaces on manifolds. Namely, first we introduce spaces of bounded and continuously differentiable functions. Then we define Holder spaces, more generally Besov-Holder spaces, by interpolation. This is not straightforward since we can only interpolate between spaces of bounded Ck-functions whose derivatives are uniformly continuous. Due to the fact that we are mainly interested in noncompact manifolds, the concept of uniform continuity is not a priori clear and has to be clarified first. Then the next problem is to show that Holder spaces introduced in this invariant way can be described locally by their standard anisotropic counterparts on Rm × J and Hm × J. Such representations in local coordinates are, of course, fundamental for the study of concrete equations, for example. In order to achieve these goals we set up the preliminary Section 11 in which we establish the needed properties of (vector-valued) Holder and Bessel-Holder spaces in Euclidean settings. In Section 12 we can then settle the problems alluded to above. It should be mentioned that in these two sections we consider time-independent isotropic as well as time-dependent anisotropic spaces, thus complementing the somewhat ad hoc results on Holder spaces in [5]. Having introduced all these spaces and established their basic properties we proceed now to more refined features. In Section 13 we show that, similarly as in the Euclidean setting, Holder spaces are universal point-wise multiplier spaces for Bessel potential and Besov spaces modeled on Lp. For this we establish the rather general (almost) optimal Theorem 13.5. In practice point-wise multiplications occur, as a rule, through contractions of tensor fields. For this reason we carry out in Section 14 a detailed study of mapping properties of contractions of tensor fields, one factor belonging to a Holder space and the other one to a Bessel potential or a Besov space, in particular. It should be noted that we impose minimal regularity assumptions for the multiplier space. The larger part of Section 14 is, however, devoted to the problem of the existence of a continuous right inverse for a multiplier operator induced by a complete contraction. The main result of this section thus is Theorem 14.9. It is basic for the theory of boundary value problems. Section 15 contains general Sobolev-type embedding theorems for parameter-dependent weighted Bessel po- tential and Besov spaces. They are natural extensions of the corresponding classical results in the Euclidean setting. Making use of our point-wise multiplier and Sobolev-type embedding theorems we study in Section 16 map- ping properties of differential operators in anisotropic spaces. In view of applications to quasilinear equations we strive for minimal regularity requirements for the coefficient tensors. All results established up to this point hold both for J = R and J = R+. In contrast, Section 17 is specifically concerned with anisotropic spaces on the half-line R+. It is shown that in many cases properties of function 5 spaces on R+ can be derived from the corresponding results on the whole line R. This can simplify the situation since M × R is a usual manifold (with boundary), whereas M × R+ has corners if ∂M 6= ∅. In Section 18 we consider the important case where M has a nonempty boundary and establish the fundamental trace theorem for anisotropic weighted Bessel potential and Besov spaces, both on the 'lateral boundary' ∂M × J and on the 'initial boundary' M × {0} if J = R+. In the next section we characterize spaces of functions having vanishing initial traces. Section 20 is devoted to extending the boundary values. Here we rely, besides the trace theorem, in particular on the 'right inverse theorem' established in Section 14. The results of this section are of great importance in the theory of boundary value problems. Section 21 describes the behavior of anisotropic weighted Bessel potential, Besov, and Holder spaces under interpolation. In addition to this, we also derive interpolation theorems for 'spaces with vanishing boundary conditions'. These results are needed for a 'weak Lp-theory' of parabolic evolution equations. Our investigation of weighted anisotropic function spaces is greatly simplified by the fact that we consider full and half-cylinders over M . In this case we can take advantage of the dilation invariance of J. In practice, cylinders of finite height come up naturally and are of considerable importance. For this reason it is shown, in the last section, that all embedding, interpolation, trace theorems, etc. are equally valid if J is replaced by [0, T ] for some T ∈ (0,∞). In order to cover the many possibilities due to the (unavoidably) large set of parameters our spaces depend upon, and to eliminate repetitive arguments, we use rather condensed notation in which we exhibit the locally rel- evant information only. This requires a great deal of concentration on the part of the reader. However, everything simplifies drastically in the important special case of Riemannian manifolds with bounded geometry. In that case there are no singularities and all spaces are parameter-independent. Readers interested in this situation only can simply ignore all mention of the parameters λ, µ, and ~ω and set ρ = 1. Needless to say that even in this 'simple' situation the results of this paper are new. 2 Vector Bundles First we introduce some notation and conventions from functional analysis. Then we recall some relevant facts from the theory of vector bundles. It is the main purpose of this preparatory section to create a firm basis for the following. We emphasize in particular duality properties and local representations, for which we cannot refer to the literature. Background material on manifolds and vector bundles is found in J. Dieudonn´e [12] or J. Jost [21], for example. Given locally convex (Hausdorff topological vector) spaces X and Y, we denote by L(X ,Y) the space of con- tinuous linear maps from X into Y, and L(X ) := L(X ,X ). By Lis(X ,Y) we mean the set of all isomorphisms in L(X ,Y), and Laut(X ) := Lis(X ,X ) is the automorphism group in L(X). If X and Y are Banach spaces, then L(X ,Y) is endowed with the uniform operator norm. In this situation Lis(X ,Y) is open in L(X ,Y). We write h·, ·iX for the duality pairing between X ′ := L(X , K) and X , that is, hx′, xiX is the value of x′ ∈ X ′ at x ∈ X . Let H =(cid:0)H, (··)(cid:1) be a Hilbert space. Then the Riesz isomorphism is the conjugate linear isometric isomor- phism ϑ = ϑH : H → H ′ defined by where h·, ·i = h·, ·iH. Then hϑx, yi = (yx), x, y ∈ H, (x′y′)∗ := (ϑ−1y′ϑ−1x′), x′, y′ ∈ H ′, defines the adjoint inner product on H ′, and H ∗ :=(cid:0)H ′, (··)∗(cid:1) is the adjoint Hilbert space. Denoting by k·k and k·k∗ the inner product norms associated with (··) and (··)∗, respectively, we obtain from (2.1) and (2.2) hx′, xi ≤ kx′k∗ kxk, x′ ∈ H ′, x ∈ H. (2.3) (2.1) (2.2) 6 H. Amann: Anisotropic Function Spaces on Singular Manifolds It follows from (2.1) -- (2.3) and the fact that ϑ is an isometry that kx′k∗ = sup(cid:8) hx′, xi, kxk ≤ 1(cid:9) for x′ ∈ H ′. Thus k·k∗ is the norm in H ′ = L(H, K), the dual norm. In other words, H ′ = H ∗ as Banach spaces. For this and historical reasons we use the 'star notation' for the dual space in the finite-dimensional setting and in connection with vector bundles, whereas the 'prime notation' is more appropriate in functional analytical considerations. 2, H ′ 1), defined by hA′x′ 2, x1iH1 = hx′ 2 ∈ H ′ 2. If H1 and H2 are Hilbert spaces and A ∈ L(H1, H2), then it has to be carefully distinguished between 2, Ax1iH2 , and the adjoint A∗ ∈ L(H2, H1), given by the dual A′ ∈ L(H ′ (A∗x2x1)H1 = (x2Ax1)H2 for xi ∈ Hi and x′ Suppose H1 and H2 are finite-dimensional. Then L(H1, H2) is a Hilbert space with the Hilbert-Schmidt inner product (··)HS defined by (AB)HS := tr(B∗A) for A, B ∈ L(H1, H2), where tr denotes the trace. The corresponding norm ·HS is the Hilbert-Schmidt norm A 7→ptr(A∗A). Throughout this paper, we use the summation convention for indices labeling coordinates or bases. This means that such a repeated index, which appears once as a superscript and once as a subscript, implies summation over its whole range. By a manifold we always mean a smooth, that is, C∞ manifold with (possibly empty) boundary such that its underlying topological space is separable and metrizable. Thus, in the context of manifolds, we work in the smooth category. A manifold need not be connected, but all connected components are of the same dimension. Let M be an m-dimensional manifold and V = (V, π, M ) a K vector bundle of rank n over M . For a nonempty subset S of M we denote by VS, or VS, the restriction π−1(S) of V to S. If S is a submanifold, or S = ∂M , then VS is a vector bundle of rank n over S. As usual, Vp := V{p} is the fibre π−1(p) of V over p. Occasionally, we use the symbolic notation V =Sp∈M Vp. By Γ(S, V ) we mean the KS module of all sections of V over S (no smoothness). If S is a submanifold, or S = ∂M, then Ck(S, V ) is for k ∈ N ∪ {∞} the Fr´echet space of Ck sections over S. It is a Ck(S) := Ck(S, K) module. In the case of a trivial bundle M × E = (M × E, pr1, M ) for some n-dimensional Banach space E, a section over S is a map from S into E, that is, Γ(S, M × E) = ES. Accordingly Ck(S, M × E) = Ck(S, E) is the Fr´echet space of all Ck maps from S into E. As usual, pri denotes the natural projection onto the i-th factor of a Cartesian product (of sets). Let eV = (eV ,eπ,fM ) be a vector bundle over a manifold fM . A Ck map (f0, f ) : (M, V ) → (fM ,eV ), that is, f0 ∈ Ck(M,fM ) and f ∈ Ck(V,eV ), is a Ck bundle morphism if the diagram f ✲ eV V π ❄ M f0 ✲ eπ ❄ fM is commuting, and f Vp ∈ L(Vp, Vf0(p)) for p ∈ M. It is a conjugate linear bundle morphism if f Vp is a conjugate linear map. By defining compositions of bundle morphisms in the obvious way one gets, in particular, the category of smooth, that is C∞, bundles in which we work. Thus a bundle isomorphism is an isomorphism in the category of smooth vector bundles. If M = fM , then f is called bundle morphism if (idM , f ) is one. A bundle metric on V is a smooth section h of the tensor product V ∗ ⊗ V ∗ such that h(p) is an inner product on Vp for p ∈ M . Then the continuous map is the bundle norm derived from h. ·h : V → C(M ), v 7→ph(v, v) Suppose V = (V, h) is a metric vector bundle, that is, V is endowed with a bundle metric h. Then Vp is an p , h∗(p)), where h∗(p) is the adjoint inner p is endowed with the adjoint n-dimensional Hilbert space with inner product h(p). Hence V ∗ product on V ′ bundle metric h∗ satisfying h∗(V ∗ ⊕ V ∗)p = h∗(p) for p ∈ M, where ⊕ is the Whitney sum. p as a Banach space. The dual bundle V ∗ =Sp∈M V ∗ p , equals V ′ p = (V ′ 7 p ∈ M. It follows The (bundle) duality pairing h·, ·iV is the smooth section of V ⊗ V ∗ defined by h·, ·iV (p) = h·, ·iVp hv∗, viV ≤ v∗h∗ vh, p the Riesz isomorphism for(cid:0)Vp, h(p)(cid:1) and by h♯(p) its inverse. This defines We denote by h♭(p) : Vp → V ∗ the C∞(M )-conjugate linear (bundle) Riesz isomorphism h♭ : V → V ∗ and its inverse h♯ : V ∗ → V , given by h♭Vp = h♭(p) and h♯V ∗ p = h♯(p), respectively, for p ∈ M. Thus (v∗, v) ∈ Γ(M, V ∗ ⊕ V ). for The canonical identification of V ∗∗ hh♭v, wiV = h(w, v), p with Vp implies (v, w) ∈ Γ(M, V ⊕ V ). V ∗∗ = V, hv, v∗iV ∗ = hv∗, viV , (v, v∗) ∈ Γ(M, V ⊕ V ∗). We fix an n-dimensional Hilbert space E =(cid:0)E, (··)E(cid:1), a model fiber for V . We also fix a basis (e1, . . . , en) of E and denote by (ε1, . . . , εn) the dual basis. Of course, without loss of generality we could set E = Kn. However, for notational simplicity it is more convenient to use coordinate-free settings. Let U be open in M . A local chart for V over U is a map κ⋉ϕ : VU → κ(U ) × E, vp ∈ Vp, p ∈ U, vp 7→(cid:0)κ(p), ϕ(p)vp(cid:1), such that (κ, κ⋉ϕ) : (U, VU ) →(cid:0)κ(U ), κ(U ) × E(cid:1) is a bundle isomorphism, where κ(U ) is open in the closed half-space Hm := R+ × Rm−1 of Rm (and R0 := {0}). In particular, κ is a local chart for M . Suppose κ⋉ϕ andeκ⋉eϕ are local charts of V over U and eU, respectively. Then the coordinate change is given by (x, ξ) 7→(cid:0)eκ ◦ κ−1(x), ϕκeκ(x)ξ(cid:1), where (eκ⋉eϕ) ◦ (κ ◦ ϕ)−1 : κ(U ∩eU ) × E →eκ(U ∩eU ) × E ϕκeκ ∈ C∞(cid:0)κ(U ∩eU ),Laut(E)(cid:1) is the corresponding bundle transition map. It follows ϕeκbκϕκeκ = ϕκbκ, ϕκκ = 1E, (2.4) 1E being the identity in L(E). We set ϕ−⊤(p) :=(cid:0)ϕ−1(p)(cid:1)′ Then κ⋉ϕ−⊤ : V ∗ p ∈ U. U → κ(U ) × E∗ is the local chart for V ∗ over U dual to κ⋉ϕ. and κ∗f = f ◦ κ−1. The push-forward by κ⋉ϕ is the vector space isomorphism ∈ Lis(V ∗ p , E∗), In the following, we use standard notation for the pull-back and push-forward of functions, that is, κ∗f = f ◦ κ (κ⋉ϕ)∗ : Γ(U, V ) → Eκ(U), Its inverse is the pull-back, defined by (κ⋉ϕ)∗ : Eκ(U) → Γ(U, V ), It follows that (κ⋉ϕ)∗ is a vector space isomorphism from C∞(U, V ) onto C∞(cid:0)κ(U ), E(cid:1), and Furthermore, (eκ⋉eϕ)∗(κ⋉ϕ)∗ξ = ϕκeκ(cid:0)ξ ◦ (eκ ◦ κ−1)(cid:1), κ∗(cid:0)hv∗, viV(cid:1) =(cid:10)(κ⋉ϕ−⊤)∗v∗, (κ⋉ϕ)∗v(cid:11)E, (2.5) (2.6) v 7→(cid:0)x 7→ ϕ(cid:0)κ−1(x)(cid:1)v(cid:0)κ−1(x)(cid:1)(cid:1). ξ(cid:0)κ(p)(cid:1)(cid:1). ξ 7→(cid:0)p 7→(cid:0)ϕ(p)(cid:1)−1 ξ ∈ E eκ(Uκ∩Ueκ). (v∗, v) ∈ Γ(U, V ∗ ⊕ V ). 8 In addition, H. Amann: Anisotropic Function Spaces on Singular Manifolds v ∈ Γ(U, V ). We define the coordinate frame (b1, . . . , bn) for V over U associated with κ⋉ϕ by (κ⋉ϕ)∗(f v) = (κ∗f )(κ⋉ϕ)∗v, f ∈ KU , bν := (κ⋉ϕ)∗eν, 1 ≤ ν ≤ n. Then βν := (κ⋉ϕ−⊤)∗εν, 1 ≤ ν ≤ n, defines the dual coordinate frame for V ∗ over U . In fact, it follows from (2.6) that hβµ, bνiV = κ∗(cid:0)hεµ, eνiE(cid:1) = δµ ν , 1 ≤ µ, ν ≤ n. (2.7) (2.8) (2.9) Let (eb1, . . . ,ebn) be the coordinate frame for V over eU associated witheκ⋉eϕ. Then (2.5) and (2.6) imply ν ∈ C∞(cid:0)κ(U ∩eU )(cid:1). Hence we infer fromebν = hβµ,ebνiV bµ on U ∩eU and (2.7) that κ∗hβµ,ebνiV =(cid:10)εµ, (κ⋉ϕ)∗(eκ⋉eϕ)∗eν(cid:11)E = hεµ, ϕeκκeνiE =: (ϕeκκ)µ ν eµ, 1 ≤ ν ≤ n. The push-forward of the bundle metric h is the bundle metric (κ⋉ϕ)∗h on κ(U ) × E defined by (κ⋉ϕ)∗ebν = (ϕeκκ)µ (κ⋉ϕ)∗h(ξ, η) := κ∗(cid:0)h(cid:0)(κ⋉ϕ)∗ξ, (κ⋉ϕ)∗η(cid:1)(cid:1), ξ, η ∈ Eκ(U). Since h is a smooth section of V ∗ ⊗ V ∗ it has a local representation with respect to the dual coordinate frame: (2.10) h = hµνβµ ⊗ βν , hµν = h(bµ, bν) ∈ C∞(U ). In the following, we endow Kr×s with the Hilbert-Schmidt norm by identifying it with L(Ks, Kr) by means of the standard bases. Then we call [h] := [hµν] ∈ C∞(U, Kn×n) representation matrix of h with respect to the local coordinate frame (b1, . . . , bn). Letf[h] be the representation matrix of h with respect to the local coordinate frame associated witheκ⋉eϕ. It follows from (2.8) that where [ϕeκκ] is the representation matrix of ϕeκκ ∈ C∞(cid:0)U,L(E)(cid:1) with respect to (e1, . . . , en) and a⊤ is the κ∗f[h] = [ϕeκκ]⊤κ∗[h][ϕeκκ] on κ(U ∩eU ), transposed of the matrix a. (2.11) It should also be noted that (2.9) implies κ∗(vh) = (κ⋉ϕ)∗v(κ⋉ϕ)∗h, v ∈ Γ(U, V ). (2.12) Let [h∗] be the representation matrix of h∗ with respect to the dual coordinate frame on U . Denote by [hµν] the inverse of [h]. It is a consequence of hbν, h♭bµiV ∗ = hh♭bµ, bνiV = h(bν, bµ) = hνµ that Hence h♯βν = hνρbρ. This implies h∗µν = h∗(βµ, βν) = h(h♯βν , h♯βµ) = hνρhµσhρσ = hµν, that is, h♭bµ = hbν, h♭bµiV ∗ βν = hνµβν = hµν βν. [h]−1 = [h∗]. (2.13) Let Vi = (Vi, hi) be a metric vector bundle of rank ni over M , where i = 1, 2. Assume U is open in M and ni) the coordinate frame for Vi over U associated with 1, . . . , bi κ⋉ϕi is a local chart for Vi over U . Denote by (bi κ⋉ϕi and by (β1 i , . . . , βni i ) its dual frame. Suppose a ∈ Γ(cid:0)U, Hom(V1, V2)(cid:1). Then ν1iV2 ∈ KU . ν1 = hβν2 aν2 ν2 ⊗ βν1 1 , 2 , ab1 a = aν2 ν1 b2 (2.14) Hence, given ui = uνi i bi νi ∈ Γ(U, Vi), it follows from (2.10) that h2(au1, u2) = aν2 ν1 h2,ν2 eν2uν1 1 ueν2 2 . 9 For the adjoint section a∗ = a∗ν1 ν2 b1 ν1 ⊗ βν2 2 ∈ Γ(cid:0)U, Hom(V2, V1)(cid:1) we find analogously h1(u1, a∗u2) = a∗eν1 eν2 h1,ν1 eν1uν1 1 ueν2 2 . From h2(au1, u2) = h1(u1, a∗u2) for all ui in Γ(U, Vi) we thus get aν2 from (2.13) ν1 h2,ν2 eν2 = a∗eν1 eν2 h1,ν1 eν1. Hence it follows ν2 = h∗ν1 eν1 a∗ν1 1 aeν2 eν1 h2,eν2ν2 , 1 ≤ νi ≤ ni. (2.15) The following well-known basic examples of vector bundles are included for later reference and to fix notation. Examples 2.1 (a) (Trivial bundles) Consider the trivial vector bundle V = (V, h) :=(cid:0)M × E, (··)E(cid:1) with the usual identification of the inner product of E with the bundle metric M × E. For any local chart κ of M , the trivial bundle chart over κ is given by κ⋉1E. Thus (κ⋉1E)∗v = κ∗v for v ∈ Γ(cid:0)dom(κ), M × E) = Edom(κ). (b) (Tangent bundles) Let M = (M, g) be an m-dimensional Riemannian manifold. Throughout this pa- per we denote by T M the tangent bundle if K = R and the complexified tangent bundle if K = C. Then g, respectively its complexification, is a bundle metric on T M (also denoted by g if K = C). Thus is the (complexified, if K = C) cotangent bundle of M . T ∗M := (T M )∗ = (T ∗M, g∗) We use Km as the model fiber for T M and choose for (e1, . . . , em) the standard basis ei j, 1 ≤ i, j ≤ m. Furthermore, (··) = (··)Km is the Euclidean (Hermitean) inner product on Km and · = ·Km the correspond- ing norm. We identify (Km)∗ with Km by means of the duality pairing j = δi hη, ξi = hη, ξiKm := ηiξj, η = ηiεi, ξ = ξjej, (2.16) push-forward of w, denoted by κ∗w also. so that εi = ei for 1 ≤ i ≤ m. Suppose κ is a local chart for M and set U := dom(κ). Denote by T κ : TU M = (T M )U → κ(U ) × Km the (complexified, if K = C) tangent map of κ. Then κ⋉T κ is a local chart for T M over U , the canonical chart for T M over κ. It is completely determined by κ. For this reason (κ⋉T κ)∗v is denoted, as usual, by κ∗v for v ∈ Γ(U, T M ). Then the push-forward(cid:0)κ⋉(T κ)−⊤(cid:1)∗w of a covector field w ∈ Γ(U, T ∗M ) is the usual Note that the bundle transition map for the coordinate change (eκ⋉Teκ) ◦ (κ⋉T κ)−1 equals ∂x(eκ ◦ κ−1), The coordinate frame for T M on U associated with κ, that is, with κ⋉T κ, equals (∂/∂x1, . . . , ∂/∂xm). Its dual frame is (dx1, . . . , dxm). The representation matrix of g with respect to this frame is the fundamental matrix [gij ] ∈ C∞(U, Km×m) of M on U . is the C∞(M ) module of all (complexified, if K = C) smooth vector, respectively covector, fields on M . For abbreviation, we set T M := C∞(M, T M ) and T ∗M := C∞(M, T ∗M ). Then T M , respectively T ∗M , where ∂x denotes the (Fr´echet) derivative (on Rm). (cid:3) Let Vi = (Vi, hi) be a metric vector bundle over M for i = 1, 2. Then the dual (V1 ⊗ V2)∗ of the tensor product V1 ⊗ V2 is identified with V ∗ 1 ⊗ V ∗ 2 by means of the duality pairing h·, ·iV1⊗V2 defined by 1 ⊗ v∗ hv∗ 2 , v2iV2 , By h1 ⊗ h2 we denote the bundle metric for V1 ⊗ V2, given by 2, v1 ⊗ v2iV1⊗V2 := hv∗ 1 , v1iV1hv∗ i , vi) ∈ Γ(M, V ∗ (v∗ i ⊕ Vi). (2.17) h1 ⊗ h2(v1 ⊗ v2, w1 ⊗ w2) := h1(v1, w1)h2(v2, w2), (vi, wi) ∈ Γ(M, Vi ⊕ Vi). (2.18) We always equip V1 ⊗ V2 with this metric. 10 H. Amann: Anisotropic Function Spaces on Singular Manifolds Suppose that κ is a local chart for M and κ⋉ϕi is a local chart for Vi over dom(κ). Then κ⋉(ϕ1 ⊗ ϕ2) denotes the local chart for V1 ⊗ V2 over dom(κ) induced by κ⋉ϕi, i = 1, 2, that is, (cid:0)κ⋉(ϕ1 ⊗ ϕ2)(cid:1)∗(v1 ⊗ v2) = (κ⋉ϕ1)∗v1 ⊗ (κ⋉ϕ2)∗v2, (v1, v2) ∈ Γ(M, V1 ⊕ V2). (2.19) It is obvious how these concepts generalize to tensor products of more than two vector bundles over M . A connection on V is a map ∇ : T M × C∞(M, V ) → C∞(M, V ), (X, v) 7→ ∇X v which is C∞(M ) linear in the first argument, additive in its second, and satisfies the 'product rule' f ∈ C∞(M ), where Xf := df (X) = hdf, Xi := hdf, XiT M . Equivalently, ∇ is considered as a K linear map, ∇X (f v) = (Xf )v + f∇Xv, v ∈ C∞(M, V ), X ∈ T M, (2.20) called covariant derivative, defined by ∇ : C∞(M, V ) → T ∗M ⊗ C∞(M, V ), h∇v, X ⊗ v∗iT M⊗V ∗ = hv∗,∇X viV , v∗ ∈ C∞(M, V ∗), v ∈ C∞(M, V ), X ∈ T M, (2.21) and satisfying the product rule. Here and in similar situations, T M is identified with the 'real' subbundle of the complexification T M + i T M if K = C. (In other words: We consider 'real derivatives' of complex-valued sections.) A connection is metric if it satisfies v, w ∈ C∞(M, V ). Let ∇ be a metric connection on V . Then we define a connection on V ∗, again denoted by ∇, by Xh(v, w) = h(∇X v, w) + h(v,∇X w), X ∈ T M, h∇X v∗, viV := Xhv∗, viV − hv∗,∇X viV (2.22) (2.23) for v∗ ∈ C∞(M, V ∗), v ∈ C∞(M, V ), and X ∈ T M. It follows for v, w ∈ C∞(M, V ) and X ∈ T M that, due to (2.22), Xh(v, w) = Xhh♭w, viV =(cid:10)∇X (h♭w), v(cid:11)V + hh♭w,∇X viV =(cid:10)∇X (h♭w), v(cid:11)V + h(∇X v, w) =(cid:10)∇X (h♭w), v(cid:11)V + Xh(v, w) − h(v,∇X w) =(cid:10)∇X (h♭w), v(cid:11)V + Xh(v, w) −(cid:10)h♭(∇X w), v(cid:11)V . h♯ ◦ ∇ = ∇ ◦ h♯. ∇ ◦ h♭ = h♭ ◦ ∇, This and h♯ = (h♭)−1 imply Consequently, Xh∗(v∗, w∗) = Xh(h♯w∗, h♯v∗) = h(h♯∇X w∗, h♯v∗) + h(h♯w∗, h♯∇X v∗) = h∗(∇X v∗, w∗) + h∗(v∗,∇X w∗) for v∗, w∗ ∈ C∞(M, V ∗). This shows that ∇ is a metric connection on (V ∗, h∗). Let (Vi, hi) be a metric vector bundle over M for i = 1, 2. Suppose ∇i is a metric connection on Vi. Then ∇X (v1 ⊗ v2) := ∇1X v1 ⊗ v2 + v1 ⊗ ∇2X v2, vi ∈ C∞(M, Vi), X ∈ T M, (2.24) defines a metric connection ∇ = ∇(∇1,∇2) on V1 ⊗ V2, the connection induced by ∇1 and ∇2. In the particular case where either V2 = V1 or V2 = V ∗ 1 and ∇2 = ∇1, we write again ∇1 for ∇(∇1,∇1). Let ∇ be a connection on V . Suppose κ⋉ϕ is a local chart for V over U . The Christoffel symbols Γν iµ, 1 ≤ i ≤ m, 1 ≤ µ, ν ≤ n, of ∇ with respect to κ⋉ϕ are defined by iµbν. ∇∂/∂xibµ = Γν 11 (2.25) Here and in similar situations, it is understood that Latin indices run from 1 to m and Greek ones from 1 to n. It follows (2.26) Let V1 and V2 be metric vector bundles over M with metric connections ∇1 and ∇2, respectively. For a smooth section a of Hom(V1, V2) we define v = vνbν ∈ C∞(U, V ). ∂xi + Γν ∇v =(cid:16) ∂vν iµvµ(cid:17) dxi ⊗ bν, u ∈ C∞(M, V1). (∇12a)u := ∇2(au) − a∇1u, (2.27) Then ∇12 is a metric connection on Hom(V1, V2), the one induced by ∇1 and ∇2, where Hom(V1, V2) is endowed with the (fiber-wise defined) Hilbert-Schmidt inner product. It is verified that this definition is consistent with (2.14) and (2.24). Hence we also write ∇(∇1,∇2) for ∇12. 3 Uniform Regularity Let M be an m-dimensional manifold. We set Q := (−1, 1) ⊂ R. If κ is a local chart for M , then we write Uκ for the corresponding coordinate patch dom(κ). A local chart κ is normalized if κ(Uκ) = Qm whenever Uκ ⊂ M , the interior of M , whereas κ(Uκ) = Qm ∩ Hm if Uκ ∩ ∂M 6= ∅. We put Qm An atlas K for M has finite multiplicity if there exists k ∈ N such that any intersection of more than k coordi- nate patches is empty. In this case κ := κ(Uκ) if κ is normalized. N(κ) := {eκ ∈ K ; Ueκ ∩ Uκ 6= ∅ } κ ) ; κ ∈ K(cid:9) is a cover of M . has cardinality ≤ k for each κ ∈ K. An atlas is uniformly shrinkable if it consists of normalized charts and there exists r ∈ (0, 1) such that(cid:8) κ−1(rQm Given an open subset X of Rm or Hm and a Banach space X over K, we write k·kk,∞ for the usual norm of BCk(X,X ), the Banach space of all u ∈ Ck(X,X ) such that ∂αuX is uniformly bounded for α ∈ Nm with α ≤ k (see Section 11). By c we denote constants ≥ 1 whose numerical value may vary from occurrence to occurrence; but c is always independent of the free variables in a given formula, unless an explicit dependence is indicated. Let S be a nonempty set. On RS we introduce an equivalence relation ∼ by setting f ∼ g iff there exists c ≥ 1 such that f /c ≤ g ≤ cf. Inequalities between bundle metrics have to be understood in the sense of quadratic forms. An atlas K for M is uniformly regular if (i) K is uniformly shrinkable and has finite multiplicity. (ii) keκ ◦ κ−1kk,∞ ≤ c(k), κ,eκ ∈ K, k ∈ N. In (ii) and in similar situations it is understood that only κ,eκ ∈ K with Uκ ∩ Ueκ 6= ∅ are being considered. Two uniformly regular atlases K andeK are equivalent, K ≈eK, if card{eκ ∈eK ; Ueκ ∩ Uκ 6= ∅ } ≤ c, κ ∈ K. keκ ◦ κ−1kk,∞ ≤ c(k), κ ∈ K, eκ ∈eK, k ∈ N. Let V be a vector bundle of rank n over M with model fiber E. Suppose K is an atlas for M and κ⋉ϕ is for each κ ∈ K a local chart for V over Uκ. Then K⋉Φ := { κ⋉ϕ ; κ ∈ K} is an atlas for V over K. It is uniformly regular if (3.2) (i) (ii) (3.1) (i) (ii) K is uniformly regular; kϕκeκkk,∞ ≤ c(k), κ⋉ϕ,eκ⋉eϕ ∈ K⋉Φ, k ∈ N, (3.3) 12 H. Amann: Anisotropic Function Spaces on Singular Manifolds where ϕκeκ is the bundle transition map corresponding to the coordinate change (eκ⋉eϕ) ◦ (κ⋉ϕ)−1. Two atlases K⋉Φ andeK⋉eΦ for V over K andeK, respectively, are equivalent, K⋉Φ ≈eK⋉eΦ, if K ≈eK; kϕκeκkk,∞ ≤ c(k), κ⋉ϕ ∈ K⋉Φ, eκ⋉eϕ ∈eK⋉eΦ, k ∈ N. Suppose h is a bundle metric for V . Let K⋉Φ be a uniformly regular atlas for V over K. Then h is uniformly (3.4) (ii) (i) regular over K⋉Φ if (i) (ii) (κ⋉ϕ)∗h ∼ (··)E, κ⋉ϕ ∈ K⋉Φ; k(κ⋉ϕ)∗hkk,∞ ≤ c(k), κ⋉ϕ ∈ K⋉Φ, k ∈ N. (3.5) Let [h]κ⋉ϕ = [hµν ]κ⋉ϕ be the representation matrix of h with respect to the local coordinate frame associated with κ⋉ϕ. Then it follows from (2.10) that Hence (3.5)(i) is equivalent to κ∗(cid:0)[h]κ⋉ϕ(cid:1) = [κ∗hµν ] =(cid:2)(κ⋉ϕ)∗h(cid:3). ζ ∈ Kn, x ∈ Qm κ , (3.6) κ⋉ϕ ∈ K⋉Φ. ζ2/c ≤ κ∗hµν(x)ζµζν ≤ cζ2, , If K⋉Φ ≈eK⋉eΦ and h is uniformly regular over K, then we see from (2.4) and (2.11) that h is uniformly regular overeK. Assume ∇ is a connection on V . Let K⋉Φ be an atlas for V over K. For κ⋉ϕ ∈ K⋉Φ we denote by Γν iµ[κ⋉ϕ] the Christoffel symbols of ∇ with respect to the coordinate frame for V over Uκ induced by κ⋉ϕ. Then ∇ is uniformly regular over K⋉Φ if K⋉Φ is uniformly regular; (i) (ii) (cid:13)(cid:13)κ∗(cid:0)Γν iµ[κ⋉ϕ](cid:1)(cid:13)(cid:13)k,∞ ≤ c(k), 1 ≤ i ≤ m, 1 ≤ µ, ν ≤ n, κ⋉ϕ ∈ K⋉Φ, k ∈ N. Suppose ∇ is uniformly regular over K⋉Φ and eK⋉eΦ ≈ K⋉Φ. Then it follows from (2.8), (2.26), (3.2), and (3.4) that ∇ is uniformly regular overeK⋉eΦ. A uniformly regular structure for M is a maximal family of equivalent uniformly regular atlases for it. We say M is a uniformly regular manifold if it is endowed with a uniformly regular structure. In this case it is understood that each uniformly regular atlas under consideration belongs to this uniformly regular structure. Let M be uniformly regular and V a vector bundle over M . A uniformly regular bundle structure for V is a maximal family of equivalent uniformly regular atlases for V . Then V is a uniformly regular vector bundle over M , if it is equipped with a uniformly regular bundle structure. Again it is understood that in this case each atlas for V belongs to the given uniformly regular bundle structure. A uniformly regular metric vector bundle is a uniformly regular vector bundle endowed with a uniformly regular bundle metric. By a fully uniformly regular vector bundle V = (V, hV ,∇V ) over M we mean a uniformly regular vector bundle V over M equipped with a uniformly regular bundle metric hV and a uniformly regular metric connection ∇V . As earlier, it is the main purpose of the following examples to fix notation and to prepare the setting for further investigations. Examples 3.1 (a) (Trivial bundles) Let E =(cid:0)E, (··)E(cid:1) be an n -- dimensional Hilbert space. Suppose M is a uniformly regular manifold. It is obvious from Example 2.1(a) that the trivial bundle M × E is uniformly regular over M and (··)E is a uniformly regular bundle metric. We consider E as a manifold of dimension n if K = R, and of dimension 2n if K = C (using the standard identification of C = R + i R with R2) whose smooth structure is induced by the trivial chart 1E. We identify T E canonically with E × E. Then T v : T M → T E = E × E, the tangential of v ∈ C∞(M, E), is well-defined. We set dX v := pr2 ◦ T v(X), X ∈ T M, v ∈ C∞(M, E). Then d : T M × C∞(M, E) → C∞(M, E), 13 (X, v) 7→ dX v is a connection on M × E, the E-valued differential on M . Then it follows that Let (e1, . . . , en) be a basis for E and use the same symbol for the constant frame p 7→ (e1, . . . , en) of M × E. Thus, since all Christoffel symbols are identically zero, d is trivially uniformly regular. df = df ν eν, f = f νeν ∈ C∞(M, E). (b) (Subbundles) Let V be a vector bundle of rank n over a manifold M , endowed with a bundle metric h and a metric connection ∇. Suppose W is a subbundle of rank ℓ. Denote by ι : W ֒→ V the canonical injection. Let hW := ι∗h be the pull-back metric on W . We write P for the the orthogonal projection onto W in V . Then P ∈ C∞(cid:0)M, Hom(V, V )(cid:1) and it is verified that is a metric connection on (W, hW ), the one induced by ∇. Let E be a model fiber of V and (e1, . . . , en) a basis for it. Suppose V is uniformly regular and there exists an atlas K⋉Φ for V such that (κ⋉ϕ)∗(e1, . . . , eℓ) is for each κ ∈ K a frame for W over Uκ. Then it is checked that W = (W, hW ,∇W ) is a fully uniformly regular vector bundle over M . (X, w) 7→ P∇X(cid:0)ι(w)(cid:1) ∇W : T M × C∞(M, W ) → C∞(M, W ), Suppose Vi = (Vi, hi,∇i), i = 1, 2, are fully uniformly regular vector bundles over M . Set and (h1 ⊕ h2)(v1 ⊕ v2, ev1 ⊕ev2) := h1(v1,ev1) + h2(v2,ev2), (∇1 ⊕ ∇2)(v1 ⊕ v2) := ∇1v1 ⊕ ∇2v2, (vi,evi) ∈ Γ(M, Vi ⊕ Vi), (v1, v2) ∈ C∞(M, V1 ⊕ V2). Then (V1 ⊕ V2, h1 ⊕ h2, ∇1 ⊕ ∇2) is a fully uniformly regular vector bundle over M . Furthermore, Vi is for i = 1, 2 a fully uniformly regular subbundle of V . (c) (Riemannian manifolds) Let M = (M, g) be an m-dimensional Riemannian manifold. We denote by gm = (dx1)2 + ··· + (dxm)2 the Euclidean metric on Rm and use the same symbol for its complexification as well as for the restriction thereof to open subsets of Rm and Hm. Then M is a uniformly regular Riemannian manifold, if T M is uniformly regular and g is a uniformly regular bundle metric on T M . It follows from Example 2.1(b) that M is a uniformly regular Riemannian manifold iff (i) (ii) (iii) M is uniformly regular; κ∗g ∼ gm, κ ∈ K; kκ∗gkk,∞ ≤ c(k), κ ∈ K, k ∈ N, (3.7) for some uniformly regular atlas K for M . Of course, κ∗g := (κ⋉T κ)∗g in conformity with standard usage. We denote by ∇g the (complexified, if K = C) Levi-Civita connection for M , that is, for T M . Its Christoffel symbols with respect to the coordinate frame (∂/∂x1, . . . , ∂/∂xm) over Uκ admit the representation 2Γk ij = gkℓ(∂igℓj + ∂jgℓi − 2∂ℓgij ), (3.8) where ∂i := ∂/∂xi. From this and (3.7)(ii) and (iii) it follows that ∇g is uniformly regular if (M, g) is a uniformly regular Riemannian manifold. In addition, ∇g is metric and Γk (d) Every compact Riemannian manifold is a uniformly regular Riemannian manifold. (e) It has been shown in Example 2.1(c) of [5] that Rm = (Rm, gm) and Hm = (Hm, gm) are uniformly ji. ij = Γk regular Riemannian manifolds. (f) (Homomorphism bundles) For i = 1, 2 let (Vi, hi) be a uniformly regular metric vector bundle of rank ni over M . We denote by (V12, h12) the homomorphism bundle V12 := Hom(V1, V2) endowed with the Hilbert- Schmidt bundle metric h12 = (··)HS. 14 H. Amann: Anisotropic Function Spaces on Singular Manifolds Assume K⋉Φi is a uniformly regular atlas for Vi, and Ei is a model fiber for Vi with basis (ei 1, . . . , ei ni) and dual basis (ε1 i , . . . , εni i ). For κ⋉ϕi ∈ K⋉Φi we define a bundle isomorphism (κ, κ⋉ϕi) : (cid:0)Uκ, (V12)Uκ(cid:1) →(cid:0)κ(Uκ), κ(Uκ) × L(E1, E2)(cid:1) by setting (κ⋉ϕ12)ap :=(cid:0)κ(p), ϕ12(p)ap(cid:1) for p ∈ Uκ and ap ∈ (V12)p, where ϕ12(p)ap(x) := ϕ2(p)apϕ−1 x = κ(p). 1 (x), It follows (eκ⋉eϕ12)∗(κ⋉ϕ12)∗b = (eκ⋉eϕ2)∗(κ⋉ϕ2)∗b(κ⋉ϕ1)∗(eκ⋉eϕ1)∗, ifeκ⋉eϕi belongs to a uniformly regular atlas for Vi. From this we deduce that K12 := { κ⋉ϕ12 ; κ⋉ϕi ∈ K⋉Φi, i = 1, 2 } b ∈ L(E1, E2), is a uniformly regular atlas for V12 and that any two such atlases are equivalent. Hence V12 is a uniformly regular vector bundle over M . The coordinate frame of V12 over Uκ associated with κ⋉ϕ12 is given by { b2 ν2 ⊗ βν1 1 ; 1 ≤ νi ≤ ni, i = 1, 2 }, (3.9) where (bi frame. By (2.15) and (3.9) we find 1, . . . , bi ni) is the coordinate frame of Vi over Uκ associated with κ⋉ϕi and (β1 i , . . . , βni i ) is its dual (3.10) From this, (2.10), (3.5), and (3.6) we deduce (κ⋉ϕ12)∗h12(a, a) = κ∗h∗ν1 eν1 1 κ∗h2,ν2 eν2aν2 h2,eν2ν2(cid:3). 1 [h12] =(cid:2)h∗ν1 eν1 eν1 ∼Xν2 ν1 aeν2 κ∗h∗ν1 eν1 1 ν1aν2 aν2 eν1 ∼ Xν1,ν2 ν1aν2 aν2 ν1 = (a, a)HS for a ∈ L(E1, E2), as well as k(κ⋉ϕ12)∗h12kk,∞ ≤ c(k) for κ⋉ϕ12 ∈ K⋉Φ12 and k ∈ N. Hence (V12, h12) is a uniformly regular metric vector bundle over M . Suppose ∇i is a uniformly regular metric connection on Vi. Then it is a consequence of the consistency of (2.27) with (2.24) that ∇12 is a uniformly regular metric connection on V12. (g) (Tensor products) Let (Vi, hi), i = 1, 2, be uniformly regular metric vector bundles over M . Then it follows from (2.17) -- (2.19) that (V1 ⊗ V2, h1 ⊗ h2) is a uniformly regular metric vector bundle over M . If ∇i is a uniformly regular metric connection on Vi, then we see from (2.24) that ∇(∇1,∇2) is a uniformly regular metric connection on V1 ⊗ V2. 4 Singular Manifolds (cid:3) Let M = (M, g) be an m-dimensional Riemannian manifold. Suppose ρ ∈ C∞(cid:0)M, (0,∞)(cid:1). Then (ρ, K) is a singularity datum for M if (i) (ii) (iii) (iv) (M, g/ρ2) is a uniformly regular Riemannian manifold. K is a uniformly regular atlas for M which is orientation preserving if M is oriented. kκ∗ρkk,∞ ≤ c(k)ρκ, κ ∈ K, k ∈ N, where ρκ := κ∗ρ(0) = ρ(cid:0)κ−1(0)(cid:1). ρκ/c ≤ ρ(p) ≤ cρκ, p ∈ Uκ, κ ∈ K. Two singularity data (ρ, K) and (eρ,eK) are equivalent, (ρ, K) ≈ (eρ,eK), if ρ ∼eρ and K ≈eK. (4.1) (4.2) Note that (4.1)(iv) and (4.2) imply 1/c ≤ ρκ/ρeκ ≤ c, κ ∈ K, eκ ∈eK, Uκ ∩ Ueκ 6= ∅. 15 (4.3) A singularity structure, S(M ), for M is a maximal family of equivalent singularity data. A singularity function for M is a function ρ ∈ C∞(cid:0)M, (0,∞)(cid:1) such that there exists an atlas K with (ρ, K) ∈ S(M ). The set of all singularity functions is the singularity type, T(M ), of M . By a singular manifold we mean a Riemannian manifold M endowed with a singularity structure S(M ). Then M is said to be singular of type T(M ). If ρ ∈ T(M ), then it is convenient to set [[ρ]] := T(M ). Let M be singular of type [[ρ]]. Then M is a uniformly regular Riemannian manifold iff ρ ∼ 1. If ρ /∼ 1, then either inf ρ = 0 or sup ρ = ∞, or both. Hence M is not compact but has singular ends. It follows from (4.1) that the diameter of the coordinate patches converges either to zero or to infinity near the singular ends in a manner controlled by the singularity type T(M ). We refer to [5] and [6] for examples of singular manifolds which are not uniformly regular Riemannian mani- folds. Throughout the rest of this paper we assume M = (M, g) is an m-dimensional singular manifold. W = (W, hW , D) is a fully uniformly regular vector bundle of rank n over M. σ, τ ∈ N. (4.4) It follows from the preceding section that the uniform regularity of W , hW , and D is independent of the particular choice of the singularity datum (ρ, K). Henceforth, T M and T ∗M have to be interpreted as the complexified tangent and cotangent bundles, respec- tively, if K = C. Accordingly, h·, ·iT M , g, and ∇g are then the complexified duality pairing, Riemannian metric, and Levi-Civita connection, respectively. As usual, T σ τ M = T M ⊗σ ⊗ T ∗M ⊗τ is the (σ, τ )-tensor bundle, that is, the vector bundle of all K-valued 1 M = T ∗M , tensors on M being contravariant of order σ and covariant of order τ. In particular, T 1 and T 0 0 M = T M, T 0 0 M = M × K. Then τ (W ) = T σ is the vector bundle of W -valued (σ, τ )-tensors on M . V = V σ τ (M, W ) := T σ τ M ⊗ W call its elements E-valued (σ, τ )-tensors. Furthermore, T σ viation, we set If W = M × E with an n-dimensional Hilbert space E, then we write T σ τ (M, W )(cid:1). τ (M, W ) := C∞(cid:0)M, T σ T σ It is the C∞(M ) module of smooth W -valued (σ, τ )-tensor fields on M . τ (M, K) is naturally identified with T σ τ (M, E) for T σ τ (M, M × E) and τ M . For abbre- The canonical identification of (T σ (bundle) duality pairing We endow V with the bundle metric h·, ·iV := h·, ·iT σ τ M ⊗ h·, ·iW . τ M )∗ with T τ σ M leads to T σ τ (M, W )∗ = T τ σ (M, W ∗) with respect to the h := (··)τ σ := g⊗σ ⊗ g∗⊗τ is the bundle metric on T σ σ ⊗ hW , (4.5) τ M induced by g (denoted by (··)g in Section 3 of [5]). where (··)τ Finally, we equip V with the metric connection ∇ := ∇(∇g, D) 16 H. Amann: Anisotropic Function Spaces on Singular Manifolds induced by the Levi-Civita connection of M and connection D of W . In summary, in addition to (4.4), V = (V, h,∇) :=(cid:0)T σ τ (M, W ), (··)τ σ ⊗ hW , ∇(∇g, D)(cid:1) τ (M, W ) into T σ 0 (M, W ) = C∞(M, W ). is a standing assumption. In particular, ∇ is a K-linear map from T σ and ∇k+1 := ∇ ◦ ∇k for k ∈ N. Note ∇u = Du for u ∈ T 0 5 Local Representations τ +1(M, W ). We set ∇0 := id Although W is a fully uniformly regular vector bundle over M this is not true for V , due to the fact that h involves the singular Riemannian metric g. For this reason we have to study carefully the dependence of various local representations on the singularity datum. This is done in the present section. For a subset S of M and a normalized atlas K we let KS := { κ ∈ K ; Uκ ∩ S 6= ∅ }; hence K∅ = ∅. Then, given κ ∈ K, Xκ :=( Rm if κ ∈ K\K∂M , Hm otherwise, (5.1) considered as an m-dimensional uniformly regular Riemannian manifold with the Euclidean metric. Furthermore, Qm κ is an open Riemannian submanifold of Xκ. Let F be a finite-dimensional Hilbert space. Then, using standard identifications, Of course, we identify (Km)∗ with Km by means of (2.16), but continue to denote it by (Km)∗ for clarity. We endow T σ κ , F ) with the inner product τ (Qm T σ τ (Qm κ , F ) = (Km)⊗σ ⊗(cid:0)(Km)∗(cid:1)⊗τ ⊗ F. (··)T σ τ (Qm κ ,F ) := (··)⊗σ Km ⊗ (··)⊗τ (Km)∗ ⊗ (··)F . (5.2) For ν ∈ N× we set Jν := {1, . . . , m}ν and denote its general point by (i) = (i1, . . . , iν). The standard basis (e1, . . . , em) of Km, that is, ei j, and its dual basis (ε1, . . . , εm) induce the standard basis j = δi of T σ τ Qm (cid:8) e(i) ⊗ ε(j) ; (i) ∈ Jσ, (j) ∈ Jτ(cid:9) κ , where e(i) = ei1 ⊗ ··· ⊗ eiσ and ε(j) = εj1 ⊗ ··· ⊗ εjτ . Then ⊗ (Km)⊗τ , F(cid:1) κ , F ) = L(cid:0)(cid:0)(Km)∗(cid:1)⊗σ (e)(cid:3)(cid:1)HS,F := X(i)∈Jσ, (j)∈Jτ(cid:0)a(i) (j)(cid:12)(cid:12) b(i) (j)(cid:1)F a ∈ T σ (j)(cid:3) ∈ F mσ×mτ . We endow F mσ ×mτ with the inner product has the representation matrix(cid:2)a(i) (j)(cid:3)(cid:12)(cid:12)(cid:2)b(eı) (cid:0)(cid:2)a(i) which coincides with the Hilbert-Schmidt inner product if F = K. For abbreviation, we set τ (Qm It follows from (5.2) that a 7→(cid:2)a(i) We assume E = Eσ τ = Eσ τ (F ) := F mσ×mτ , (··)E := (··)HS,F . τ (Qm κ , F ), (··)T σ (j)(cid:3) defines an isometric isomorphism by which (ρ, K) is a singularity datum for M ; κ ,F )(cid:1) with(cid:0)E, (··)E(cid:1). we identify(cid:0)T σ • • K⋉Φ is a uniformly regular atlas for W over K; • F =(cid:0)F, (··)F(cid:1) is a model fiber for W with basis (e1, . . . , en). τ (Qm (5.3) (5.4) Suppose κ⋉ϕ ∈ K⋉Φ and κ = (x1, . . . , xm). Then κ⋉ϕσ τ : VUκ → Qm κ × E, the local chart for V over Uκ induced by κ⋉ϕ, is defined by ϕσ τ (p)vp := (Tpκ)X 1 p ⊗ ··· ⊗ (Tpκ)X σ τ (p)vp(cid:1), vp 7→(cid:0)κ(p), ϕσ p ⊗ (Tpκ)−⊤α1,p ⊗ ··· ⊗ (Tpκ)−⊤ατ,p ⊗ ϕ(p)wp vp ∈ Vp, p ∈ Uκ, for vp = X 1 longing to Wp. p ⊗ ··· ⊗ X σ p ⊗ α1,p ⊗ ··· ⊗ ατ,p ⊗ wp ∈ T σ τ (M, W )p with X i p ∈ TpM, αj,p ∈ T ∗ p M , and wp be- Set ∂ ∂x(i) := ∂ ∂xi1 ⊗ ··· ⊗ ∂ ∂xiσ , dx(j) := dxj1 ⊗ ··· ⊗ dxjτ , (i) ∈ Jσ, (j) ∈ Jτ . Furthermore, let (b1, . . . , bn) be the coordinate frame for W over Uκ associated with κ⋉ϕ and (β1, . . . , βn) its dual frame. Then n ∂ ∂x(i) ⊗ dx(j) ⊗ bν ; (i) ∈ Jσ, (j) ∈ Jτ , 1 ≤ ν ≤ no is the coordinate frame for V over Uκ associated with κ⋉ϕσ τ . Hence v ∈ Γ(Uκ, V ) has the local representation 17 (5.5) (5.6) (5.7) ϕσ and v = v(i),ν (j) ∂ ∂x(i) ⊗ dx(j) ⊗ bν (j) (cid:0)κ−1(x)(cid:1)eν(cid:3) ∈ F mσ ×mτ τ v(x) =(cid:2)v(i),ν τ )−1 =(cid:0)eκ ◦ κ−1, (ϕσ Assumeeκ⋉eϕ ∈ K⋉Φ. Then (eκ⋉eϕσ (cid:0)(ϕσ τ )κeκξ(cid:1)(i),ν j1 ··· Beτ ∂(eκ ◦ κ−1)i ∂(κ ◦eκ−1)e τ ) ◦ (κ⋉ϕσ (j) = A(i) eı1 ··· Aiσ (j) = Be1 (j)(ϕκeκ)ν (eı) = Ai1 with A(i) and B(e) (eı)B(e) eνξ(eı),eν , Be , and j = eı = ∂xeı , (e) ∂yj Ai eıσ jτ x ∈ Qm κ . = E, τ )κeκ(cid:1), where ξ ∈ E, ◦ (eκ ◦ κ−1) for 1 ≤ i,eı, j,e ≤ n and y =eκ ◦ κ−1(x). Hence (3.1), (3.3), and assumption (4.4) imply that τ ; κ⋉ϕ ∈ K⋉Φ} τ := { κ⋉ϕσ K⋉Φσ is a uniformly regular atlas for V over K. From (3.2) and (3.4) we also infer that K⋉Φ ≈eK⋉eΦ =⇒ K⋉Φσ τ ≈eK⋉eΦσ τ . The local chart κ⋉ϕσ the push-forward and pull-back by κ⋉ϕσ the use of κ∗ for the push-forward of vector fields by κ⋉ϕ (see Example 2.1(b)). τ is completely determined by κ⋉ϕ. For this reason, and to simplify notation, we denote τ simply by (κ⋉ϕ)∗ and (κ⋉ϕ)∗, respectively. This is consistent with We set g(j)(ℓ) (i)(k) := gi1k1 ··· giσkσ gj1ℓ1 ··· gjτ ℓτ with (i), (k) running through Jσ and (j), (ℓ) through Jτ . Then (4.5) and (2.13) imply h(u, v) = g(j)(ℓ) (i)(k)u(i),ν (j) v(k),µ (ℓ) hW (bν, bµ), u, v ∈ Γ(Uκ, V ). Hence, setting uκ := (κ⋉ϕ)∗u etc., we get from (2.9) (κ⋉ϕ)∗h(uκ, vκ) = κ∗g(j)(ℓ) (i)(k)κ∗u(i),ν (j) κ∗v(k),µ (ℓ) κ∗hWνµ . (5.8) 18 H. Amann: Anisotropic Function Spaces on Singular Manifolds Lemma 3.1 of [5] guarantees and κ∗g ∼ ρ2 κgm, κ∗g∗ ∼ ρ−2 κ gm, κ ∈ K, From (2.12), the uniform regularity of hW over K⋉Φ, (5.8), and (5.9) we deduce ρ−2 κ kκ∗gkk,∞ + ρ2 κ kκ∗g∗kk,∞ ≤ c(k), κ ∈ K, k ∈ N. (5.9) (5.10) κ∗(uh) = (κ⋉ϕ)∗u(κ⋉ϕ)∗h ∼ ρσ−τ κ (κ⋉ϕ)∗uEσ τ , κ⋉ϕ ∈ K⋉Φ, u ∈ Γ(M, V ). (5.11) Suppose u ∈ T σ τ (M, V ) has the local representation u = u(i),ν (j) ∂ ∂x(i) ⊗ dx(j) ⊗ bν. Then it follows from (2.20), (2.21), (2.23), (2.24), and (2.25), denoting by Dν that kµ the Christoffel symbols of D, ∇u = ∂ ∂x(i) ⊗ dx(j) ⊗ dxk ⊗ bν u(i1,...,is,...,iσ ),ν (j) Γℓ kis ∂ ∂x(i1,...,ℓ,...,iσ ) ⊗ dx(j) ⊗ dxk ⊗ bν ∂ ∂x(i) ⊗ dx(j1,...,ℓ,...,jτ ) ⊗ dxk ⊗ bν u(i),ν (j1,...,jt,...,jτ ) Γjt kℓ ∂u(i),ν (j) ∂xk + − σXs=1 τXt=1 (5.12) + u(i),µ (j) Dν kµ ∂ ∂x(i) ⊗ dx(j) ⊗ dxk ⊗ bν, with ℓ being at position s in (i1, . . . , ℓ, . . . , iσ) and position t in (j1, . . . , ℓ, . . . , jτ ). We endow the trivial bundle Qm τ with the Euclidean connection, denoted by ∂x and being naturally κ × Eσ (j)(cid:3) : Qm κ → F mσ×mτ , then, . ℓ ∈ N, (5.13) (5.14) ∂ℓ τ ), κ , Eσ xv ∈ C∞(cid:0)Qm identified with the Fr´echet derivative. Thus, given v ∈ C∞(Qm τ )(cid:1), κ ,Lℓ(Rm; Eσ where Lℓ(Rm; Eσ setting ∂(k) := ∂kℓ ◦ ··· ◦ ∂k1 for (k) ∈ Jℓ with ∂i = ∂/∂xi, (j)(cid:3) : Qm ℓ ∈ N×, If v =(cid:2)v(i) τ . τ ) is the space of ℓ-linear maps from Rm into Eσ xv =(cid:2)∂(k)v(i) κ → F mσ×mτ +ℓ Hence, using the latter interpretation, ∂ℓ where ∂0 x := id. ∂ℓ x ∈ Lℓ(cid:0)C∞(Qm κ , Eσ τ ), C∞(Qm κ , Eσ τ +ℓ)(cid:1), We define the push-forward (κ⋉ϕ)∗∇ℓ : C∞(Qm κ , Eσ τ ) → C∞(Qm κ , Eσ τ +ℓ) of ∇ℓ by κ⋉ϕ by for ℓ ∈ N. Then (κ⋉ϕ)∗∇ is a metric connection on(cid:0)T σ (κ⋉ϕ)∗∇ℓ := (κ⋉ϕ)∗ ◦ ∇ℓ ◦ (κ⋉ϕ)∗ (κ⋉ϕ)∗∇ℓ+1 =(cid:0)(κ⋉ϕ)∗∇(cid:1) ◦ (κ⋉ϕ)∗∇ℓ, κ , F ), (κ⋉ϕ)∗h(cid:1) and τ (Qm ℓ ∈ N. Suppose r ∈ N× and u ∈ Cr(M, V ). Set v := (κ⋉ϕ)∗u ∈ Cr(Qm induction, and from (5.13) and (5.14) that there exist κ , Eσ τ ). Then we infer from (5.12) by 19 such that aℓ ∈ C∞(cid:0)Qm κ ,L(Eσ τ +ℓ, Eσ (κ⋉ϕ)∗∇rv = ∂r τ +r)(cid:1), r−1Xℓ=0 xv + 0 ≤ ℓ ≤ r − 1, aℓ∂ℓ xv. (5.15) More precisely, the entries of the matrix representation of aℓ are polynomials in the derivatives of order at most r − ℓ − 1 of the Christoffel symbols of ∇g and D. Hence assumption (4.4) implies κ⋉ϕ ∈ K⋉Φ, kaℓkk,∞ ≤ c(k), 0 ≤ ℓ ≤ r − 1, (5.16) due to (3.8), (5.9), and (5.10). By solving system (5.15) for 0 ≤ ℓ ≤ r 'from the bottom' we find τ +ℓ, Eσ κ ,L(Eσ whereeaℓ ∈ C∞(cid:0)Qm xv = (κ⋉ϕ)∗∇rv + ∂r τ +r)(cid:1) satisfy keaℓkk,∞ ≤ c(k), From (5.15) -- (5.18) we infer that, given r ∈ N×, rXi=0(cid:12)(cid:12)(κ⋉ϕ)∗∇i(cid:0)(κ⋉ϕ)∗u(cid:1)(cid:12)(cid:12)Eσ for κ⋉ϕ ∈ K⋉Φ and u ∈ Cr(M, V ). 6 Isotropic Bessel Potential and Besov Spaces r−1Xℓ=0eaℓ(κ⋉ϕ)∗∇ℓv, 0 ≤ ℓ ≤ r − 1, κ⋉ϕ ∈ K⋉Φ. τ +i ∼ Xα≤r(cid:12)(cid:12)∂α x(cid:0)(κ⋉ϕ)∗u(cid:1)(cid:12)(cid:12)Eσ τ (5.17) (5.18) (5.19) Weighted (isotropic) function spaces on singular manifolds have been studied in detail in [5], where, however, only scalar-valued tensor fields are considered. In this and the next section we recall the basic definitions and notation on which we shall build in the anisotropic case, and describe the needed extensions to the case of vector- bundle-valued tensor fields. We denote by D := D(V ) := D( M , V ), respectively D := D(V ) := D(M, V ), the LF-space of smooth sec- tions of V which are compactly supported in M , respectively M . Then D′ = D′(V ) := D(V ′)′ w∗ is the dual of D(V ′) endowed with the w∗-topology, the space of distribution sections on M , whereby V ′ = T τ σ (M, W ′). As usual, we identify v ∈ L1,loc( M , V ) with the distribution section(cid:0)u 7→ hu, viM(cid:1) ∈ D′, where hu, viM :=ZMhu, viV dVg, u ∈ D( M , V ′), v ∈ L1,loc( M , V ), and dVg is the volume measure of M . Hence where ֒→ means 'continuous' and d In addition to (4.4) we suppose throughout d D ֒→ D u7→u M −−−−−→ L1,loc( M , V ) ֒→ D′, ֒→ L1,loc(M, V ) ֒→ 'continuous and dense' embedding. ρ ∈ T(M ), 1 < p < ∞, λ ∈ R. (6.1) 20 H. Amann: Anisotropic Function Spaces on Singular Manifolds Assume k ∈ N. The weighted Sobolev space W k,λ p = W k,λ p (V ) = W k,λ p (V ; ρ) of W -valued (σ, τ )-tensor fields on M is the completion of D in L1,loc(V ) with respect to the norm p(cid:17)1/p u 7→ kukk,p;λ :=(cid:16) kXi=0(cid:13)(cid:13)ρλ+τ −σ+i ∇iuh(cid:13)(cid:13)p . It is independent of the particular choice of ρ in the sense that W k,λ . = means 'equal except for equivalent norms'. (V ; ρ) for ρ′ ∈ [[ρ]], where For simplicity, we do not indicate the dependence of these norms, and of related ones to be introduced below, (V ; ρ′) p p . = W k,λ on (σ, τ ). This has to be kept in mind. Note that W 0,λ p = Lλ p = Lλ p (V ) :=(cid:0)(cid:8) u ∈ Lp,loc ; kukp;λ < ∞(cid:9), k·kp;λ(cid:1), d where k·kp;λ := k·k0,p;λ. Also observe W k,λ p ֒→ W ℓ,λ p for k > ℓ. Given 0 < θ < 1, we write [·, ·]θ for the complex, and (·,·)θ,q, 1 ≤ q ≤ ∞, for the real interpolation functor of exponent θ (see [2, Section I.2] for definitions and a summary of the basic facts of interpolation theory of which we make free use). Then, given k ∈ N, and H s,λ p = H s,λ Bs,λ p = Bs,λ p W k,λ p (V ) :=( [W k,λ p (V ) :=( (W k,λ (W k,λ p p p , W k+1,λ p ]s−k, k < s < k + 1, , s = k, p , W k+1,λ , W k+2,λ p )s−k,p, k < s < k + 1, )1/2,p, s = k. In favor of a unified treatment, throughout the rest of this paper F ∈ {H, B}, Fs,λ p := Fs,λ p (V ). We denote by Fs,λ p the closure of D in Fs,λ p for s > 0 and set (V ′)(cid:1)′ with respect to the duality pairing induced by h·, ·iM . We also set , W 1,λ (V ) :=(cid:0)Fs,−λ := (W −1,λ F−s,λ B0,λ p′ p , p p )1/2,p. p s > 0, This defines the weighted Bessel potential space scale [ H s,λ [ Bs,λ p ; s ∈ R ]. p It follows (see the next section) that Fs,λ p is for s ∈ R a reflexive Banach space, and ; s ∈ R ] and the weighted Besov space scale D Denoting, for any s ∈ R, by Fs,λ p d d d p p ֒→ Ft,λ ֒→ D′, ֒→ Fs,λ the closure of D in Fs,λ p , p = Fs,λ p , Fs,λ −∞ < t < s < ∞. s < 1/p. Thus, by reflexivity, 21 , s ∈ R, Fs,λ p (V ) =(cid:0)F−s,−λ p′ (V ′)(cid:1)′ If ρ ∼ 1, then all these spaces are independent of λ. Furthermore, Fs,λ with respect to h·, ·iM . Bessel potential space H s p(V ) and Besov space Bs with g = gm, V = X × E, and D = dF . Then H s p(X, E) the standard (E-valued) Besov space Bs Bs (cf. H. Triebel [50], for example). Thus noting Fs their properties which we shall do without further reference. reduces to the non-weighted (standard) p(V ), respectively. Assume, in addition, M = X ∈ {Rm, Hm} p(X, E) is the classical (E-valued) Bessel potential space and p,p(X, E). In the scalar case these spaces are well investigated p)d with d = dim(E), we can make free use of p(X, E) ≃ (Fs p 7 The Isotropic Retraction Theorem Let Eα be a locally convex space for each α in a countable index set. Then E :=Qα Eα is endowed with the product topology. Now suppose that each Eα is a Banach space. Then we denote for 1 ≤ q ≤ ∞ by ℓq(E) the linear subspace of E consisting of all x = (xα) such that Eα(cid:1)1/q is finite. Then ℓq(E) is a Banach space with norm k·kℓq(E), and kxkℓq(E) :=((cid:0)Pαkxαkq supαkxαkEα, , 1 ≤ q < ∞, q = ∞, ℓp(E) ֒→ ℓq(E), 1 ≤ p < q ≤ ∞. all finitely supported sequences in E equipped with the finest locally convex topology for which all injections We also set cc(E) :=Lα Eα, where L denotes the locally convex direct sum. Thus Lα Eα consists of Eβ →Lα Eα are continuous. It follows d (7.1) (7.2) 1 ≤ q ≤ ∞, Furthermore, c0(E) is the closure of cc(E) in ℓ∞(E). cc(E) ֒→ ℓq(E), cc(E) ֒→ ℓq(E), q < ∞. such that If each Eα is reflexive, then ℓp(E) is reflexive as well, and ℓp(E)′ = ℓp′(E′) with respect to the duality α, and h·, ·iα is the Eα-duality pairing. pairing hh·, ·ii :=Pα h·, ·iα. Of course, E′ :=Qα E′ Let assumption (5.4) be satisfied. A localization system subordinate to K is a family(cid:8) (πκ, χκ) ; κ ∈ K(cid:9) πκ ∈ D(cid:0)Uκ, [0, 1](cid:1) and { π2 subordinate to the covering { Uκ ; κ ∈ K}; χκ = κ∗χ with χ ∈ D(cid:0)Qm, [0, 1](cid:1) and χ supp(κ∗πκ) = 1 for κ ∈ K; κ ; κ ∈ K} is a partition of unity on M kκ∗πκkk,∞ + kκ∗χκkk,∞ ≤ c(k), κ ∈ K, k ∈ N. Lemma 3.2 of [5] guarantees the existence of such a localization system. (7.3) (iii) (ii) (i) In addition to (5.4) we assume (cid:8) (πκ, χκ) ; κ ∈ K(cid:9) is a localization system subordinate to K. For abbreviation, we put for s ∈ R W s p,κ := W s p (Xκ, E), Fs p,κ := Fs p(Xκ, E), where E = Eσ τ (F ). Hence W s p =Qκ W s p,κ is well-defined, as is Fs p. We set Dκ := D(Xκ, E), Dκ := D(Xκ, E), κ ∈ K, 22 H. Amann: Anisotropic Function Spaces on Singular Manifolds as well as It should be noted that, due to (5.1), in W s D = D(X, E) :=Mκ p , Fs Given κ⋉ϕ ∈ K⋉Φ, we put for 1 ≤ q ≤ ∞ ϕλ q,κu := ρλ+m/q κ Dκ, D = D(X, E) :=Mκ Dκ. p, D, and D there occur at most two distinct function spaces. (κ⋉ϕ)∗(πκu), u ∈ C(V ), and q,κv := ρ−λ−m/q ψλ κ πκ(κ⋉ϕ)∗v, v ∈ C(Xκ, E). Here and in similar situations it is understood that a partially defined and compactly supported section of a vector bundle is extended over the whole base manifold by identifying it with the zero section outside its original domain. In addition, ϕλ q u := (ϕλ q,κu) ∈Yκ C(Xκ, E), u ∈ C(V ), and ψλ ψλ q,κvκ, q v :=Xκ A retraction from a locally convex space X onto a locally convex space Y is a map R ∈ L(X ,Y) possessing a right inverse Rc ∈ L(Y,X ), a coretraction. If no confusion seems likely, we use the same symbol for a continuous linear map and its restriction to a linear subspace of its domain, respectively for a unique continuous linear extension of it. Furthermore, in a diagram arrows always represent continuous linear maps. v = (vκ) ∈Yκ C(Xκ, E). The following theorem shows that ψλ p has a unique continuous linear extension to a retraction from ℓp(Fs ψλ coretraction. This holds for any choice of s ∈ R and p ∈ (1,∞). Thus ψλ onto Fs,λ replaced by D and Fs,λ p is a retraction from D onto D, and that ϕλ p) onto Fs,λ p , respectively. in the sense that it is completely determined by its restriction to D. The same holds if D and Fs,λ p , and ϕλ p is a coretraction. Moreover, p extends uniquely to a p is a universal retraction from ℓp(Fs p) are p p Theorem 7.1 Suppose s ∈ R. Then the diagrams D ✄ ✂ id ❅ ϕλ p ❅❘ D ✄ ✂ ψλ p ✠ ❄ D ✄ ✂ d d d ✲ ✲ Fs,λ p D ✄ ✂ ℓp(Fs p) ϕλ p ✠ ❅ ψλ p ❅❘ ✲ ❅ ϕλ p ❅❘ D ✄ ✂ ψλ p ✠ ❄ D ✄ ✂ id id ❄ Fs,λ p d d d ✲ ✲ Fs,λ p ℓp(Fs p) ϕλ p ✠ ❅ ψλ p ❅❘ ✲ id ❄ Fs,λ p are commuting, where s > 0 in the second case. P r o o f. (1) Suppose W = M × K so that V = T σ rem 6.1 of [5] guarantees that τ (M, W ) = T σ τ M. Also suppose k ∈ N. Then Theo- p is a retraction from D onto D and from ℓp(W k ψλ p ) onto W k,λ p , and ϕλ p is a coretraction. (7.4) Furthermore, set and p,κ := ρ−m ϕλ κ √κ∗g ϕλ p,κ, ψλ p,κ := ρm κ (√κ∗g)−1ψλ p,κ (7.5) ϕλ p u := (ϕλ p,κu), ψλ p v :=Xκ ψλ p,κvκ, u ∈ D, v ∈ D. 23 Then it follows from Theorem 11.1 of [5] that ψλ onto W k,λ p is a coretraction. , and ϕλ p p is a retraction from from D onto D and from ℓp( W k p ) From step (2) of the proof of the latter theorem we know ρ−m κ √κ∗g ∼ 1, κ √κ∗gkk,∞ + kρm kρ−m κ (√κ∗g)−1kk,∞ ≤ c(k), κ ∈ K, k ∈ N. (7.6) This implies that we can replace ϕp,κ and ψp,κ in [5, Theorem 11.1] by ϕp,κ and ψp,κ, respectively. Consequently, p is a retraction from D onto D and from ℓp( W k ψλ p ) onto W k,λ p , and ϕλ p is a coretraction. (7.7) (2) Let now W = (W, hW , D) be an arbitrary fully uniformly regular vector bundle over M . Then (5.19) is the analogue of Lemma 3.1(iv) of [5]. Furthermore, (5.11) implies the analogue of [5, part (v) of Lemma 3.1]. If W = M × K, then the proofs of (7.4) and (7.7) are solely based on Lemma 3.1 of [5]. Hence, due to the preceding observations, they apply without change to the general case as well. Thus (7.4) and (7.7) hold if W is an arbitrary fully uniformly regular vector bundle over M . (3) The assertions of the theorem are now deduced from (7.4) and (7.7) by interpolation and duality as in [5]. Let X and Y be Banach spaces, R : X → Y a retraction, and Rc : Y → X a coretraction. Then kykY = kRRcykY ≤ kRk kRcykX ≤ kRk kRck kykY , y ∈ Y. Hence From this and Theorem 7.1 it follows that k·kY ∼ kRc·kX. u 7→ kϕλ p ukℓp(Fs p) (7.8) (7.9) is a norm for Fs,λ p . Furthermore, another choice of K⋉Φ and the localization system leads to an equivalent norm. v 7→ (κ⋉ϕ)∗(eκ⋉eϕ)∗(χv). (7.10) For κ ∈ K andeκ ∈ N(κ) we define a linear map Seκκ : E X eκ → E Xκ, The following lemma will be repeatedly useful. Lemma 7.2 Suppose s ∈ R+ with s > 0 if F = B. Then kSeκκk ≤ c, P r o o f. Note that, by (2.5) and our convention on (κ⋉ϕ)∗, Seκκ ∈ L(Fs p,eκ, Fs p,κ), κ ∈ K. eκ ∈ N(κ), τ )eκκ(cid:0)(χv) ◦ (κ ◦eκ−1)(cid:1). Seκκv = (ϕσ Hence it follows from (3.1), (3.3), (5.6), (5.7), (7.3), and the product rule and Leibniz' formula that the asser- p,κ for s ∈ N. Now we obtain the statement for general s by tion is true if s ∈ N and F = H, since H s interpolation. . = W s p,κ It follows from Theorem 7.1 and the preceding consideration that all results proved in [5] for the Banach space scales [ Fs,λ of scalar-valued (σ, τ )-tensor fields are likewise true for W -valued (σ, τ )-tensor fields, ; s ∈ R ] p (7.11) using obvious adaptions. Thus, in particular, the properties of Fs,λ use (7.11) without further ado and simply refer to [5]. p listed in Section 6 are valid. Henceforth, we 24 H. Amann: Anisotropic Function Spaces on Singular Manifolds 8 Anisotropic Bessel Potential and Besov Spaces Given subsets X and Y of a Hausdorff topological space, we write X ⋐ Y if X is compact and contained in the interior of Y . Let I be an interval with nonempty interior and X a locally convex space. Suppose Q is a family of seminorms for X generating its topology. Then C∞(I,X ) is a locally convex space with respect to the topology induced by the family of seminorms u 7→ sup t∈K q(cid:0)∂ku(t)(cid:1), k ∈ N, K ⋐ I, q ∈ Q. This topology is independent of the particular choice of Q. For K ⋐ I we denote by DK(I,X ) the linear subspace of C∞(I,X ) consisting of those functions which are supported in K. We provide DK(I,X ) with the topology induced by C∞(I,X ). Then D(I,X ), the vector space of smooth compactly supported X -valued functions, is endowed with the inductive topology with respect to the spaces DK(I,X ) with K ⋐ I. If K ⋐ K ′ ⋐ I, then DK ′ (I,X ) induces on DK(I,X ) its original topology. Note, however, that in general D(I,X ) is not an LF-space since DK(I,X ) may not be a Fr´echet space. Given a locally convex space Y, a linear map T : D(I,X ) → Y is continuous iff its restriction to every subspace DK (I,X ) is continuous (e.g., Section 6 of H.H. Schaefer [40]). From now on it is assumed, in addition to (4.4) and (6.1), that r ∈ N×, µ ∈ R, J ∈ {R, R+}. • 1/~r := (1, 1/r) ∈ R2, ~ω := (λ, µ), We set so that s/~r = (s, s/r) for s ∈ R. on M , Suppose k ∈ N. The anisotropic weighted Sobolev space of time-dependent W -valued (σ, τ )-tensor fields p = W kr/~r,~ω W kr/~r,~ω consisting of all u satisfying ∂ku ∈ Lp(J, Lλ+kµ p p (J, V ), is the linear subspace of Lp(J, W kr,λ ) p ), endowed with the norm (8.1) kukkr/~r,p;~ω :=(cid:0)kukp Lp(J,W kr,λ p + k∂kukLp(J,Lλ+kµ p ) . )(cid:1)1/p Thus W 0/~r,~ω p . = Lp(J, Lλ p ). Theorem 8.1 (i) W kr/~r,~ω p is a reflexive Banach space. (ii) kuk∼ (iii) D(J,D) kr/~r,p;~ω :=(cid:0)kukp ֒→ W kr/~r,~ω d p . Lp(J,W kr,λ p ) j=0 k∂jukp Lp(J,W (k−j)r,λ+jµ p +Pk )(cid:1)1/p is an equivalent norm. P r o o f. It follows from Theorem 9.3 below that W kr/~r,~ω is isomorphic to a closed linear subspace of a reflexive Banach space, hence it is complete and reflexive. Proofs for parts (ii) and (iii) are given in the next section. p Observe and kuk∼ kr/~r,p;~ω =(cid:16)ZJ Xi+jr≤kr(cid:13)(cid:13)ρλ+i+jµ+τ −σ ∇i∂juh(cid:13)(cid:13)p p dt(cid:17)1/p W kr/~r,(λ,0) p . = Lp(J, W kr,λ p ) ∩ W k p (J, Lλ p ). (8.2) Note that Theorem 8.1(ii) and (8.2) show that definition (8.1) coincides, except for equivalent norms, with (1.2). Also note that the reflexivity of Lλ p implies 25 with respect to the duality pairing defined by Given 0 < θ < 1, we set W 0/~r,~ω p = Lp(J, Lλ p ) =(cid:0)Lp′ (J, L−λ p′ (V ′))(cid:1)′ hu, viM×J :=ZJ(cid:10)u(t), v(t)(cid:11)M dt. (·,·)θ :=( [·, ·]θ if F = H, if F = B. (·,·)θ,p For s > 0 we define 'fractional order' spaces by Fs/~r,~ω p = Fs/~r,~ω p (J, V ) :=( (W kr/~r,~ω (W kr/~r,~ω p p p , W (k+1)r/~r,~ω , W (k+2)r/~r,~ω p )(s−kr)/r, kr < s < (k + 1)r, )1/2, s = (k + 1)r. (8.3) (8.4) (8.5) (8.6) We denote by Fs/~r,~ω p = Fs/~r,~ω p Then negative order spaces are introduced by duality, that is, F−s/~r,~ω p = F−s/~r,~ω p . p (J, V ) the closure of D( J , D) in Fs/~r,~ω (J, V ) :=(cid:0)Fs/~r,−~ω (J, V ′)(cid:1)′ p′ , s > 0, with respect to the duality pairing induced by h·, ·iM×J . We also set s(p) := 1/2p and )1/2,p. := (H −s(p)/~r,~ω := Lp(J, Lλ , H s(p)/~r,~ω p ), B0/~r,~ω H 0/~r,~ω p p p p This defines the weighted anisotropic Bessel potential space scale [ H s/~r,~ω tropic Besov space scale [ Bs/~r,~ω p p ; s ∈ R ] and the weighted aniso- The proof of the following theorem, which describes the interrelations between these two scales and gives first interpolation results, is given in the next section. Henceforth, ξθ := (1 − θ)ξ0 + θξ1 for ξ0, ξ1 ∈ R and 0 ≤ θ ≤ 1. ; s ∈ R ]. Theorem 8.2 (i) H kr/~r,~ω p . = W kr/~r,~ω p (ii) Bs/~r,~ω 2 . = H s/~r,~ω 2 (iii) (Bs0/~r,~ω p , Bs1/~r,~ω p , k ∈ N. , s ∈ R. )θ,p . = Bsθ /~r,~ω p (iv) [F s0/~r,~ω p , F s1/~r,~ω p ]θ . = F sθ/~r,~ω p , 0 ≤ s0 < s1, 0 < θ < 1. , 0 ≤ s0 < s1, 0 < θ < 1. Next we prove, among other things, an elementary embedding theorem for anisotropic weighted Bessel po- tential and Besov spaces. Theorem 8.3 (i) Suppose −∞ < s0 < s < s1 < ∞. Then D(J,D) d ֒→ H s1/~r,~ω p d ֒→ Bs/~r,~ω p d ֒→ H s0/~r,~ω p . (8.7) (ii) Assume s < 1/p if ∂M 6= ∅, and s < r(1 + 1/p) if ∂M = ∅ and J = R+. Then Fs/~r,~ω p = Fs/~r,~ω p . 26 H. Amann: Anisotropic Function Spaces on Singular Manifolds P r o o f of (i) for s 6= 0. Using reiteration theorems, well-known density properties, and relations between the real and complex interpolation functor (e.g., [2, formula (I.2.5.2)] and Theorem 8.1(iii)), we see that (8.7) is true if s0 ≥ 0. Since D( J , D) is dense in H 0/~r,~ω p = Lp(J, Lλ p ) it follows H s1/~r,~ω p d ֒→ Bs/~r,~ω p d ֒→ H s0/~r,~ω p d ֒→ Lp(J, Lλ p ), s0 ≥ 0. Hence the definition of the negative order spaces implies that (8.7) holds if s1 ≤ 0, where the density of these embeddings follows by reflexivity. This implies assertion (i) if s 6= 0. The proofs for the case s = 0 and for assertion (ii) are given in the next section. Corollary 8.4 Suppose s ∈ R. (i) Fs/~r,~ω is a reflexive Banach space. p (ii) If s > 0, then Fs/~r,~ω p (iii) Fs1/~r,~ω p d ֒→ Fs0/~r,~ω p if s1 > s0. =(cid:0)F−s/~r,−~ω p′ (J, V ′)(cid:1)′ with respect to h·, ·iM×J . P r o o f. Assume s > 0. Then assertion (i) follows from the reflexivity of W kr/~r,~ω properties of the real and complex interpolation functors. Hence Fs/~r,~ω a reflexive Banach space, is reflexive. Thus F−s/~r,~ω We have already seen that H 0/~r,~ω proves (i) for every s ∈ R. is reflexive. The reflexivity of B0/~r,~ω p p p Assertion (ii) is a consequence of (i) and (8.5). Claim (iii) is immediate by (8.7). p for k ∈ N and the duality (J, V ′), being a closed linear subspace of is reflexive since it is the dual of a reflexive Banach space. follows by interpolation as well. This p p , of course. are denoted by Fs/~r 0 = T 0 V 0 If M is uniformly regular, that is, T(M ) = [[1]], then Fs/~r,~ω is independent of ~ω. These non-weighted spaces 0 ). Since 0 M is in this case the trivial vector bundle M × K, whose sections are the K-valued functions on M , this p (M × J) via the identification of u(t) If W = M × K, then we write Fs/~r,~ω p(M )(cid:1) with Lλ notation is consistent with usual identification of Lp(cid:0)J, Lλ with u(·, t). (M × J) for Fs/~r,~ω (J, V 0 p p p 9 The Anisotropic Retraction Theorem rem implies Let { Eα ; α ∈ A} be a countable family of Banach spaces. We set Lp(J, E) :=Qα Lp(J, Eα). Fubini's theo- using obvious identifications. We also set (E, F )θ :=Qα(Eα, Fα)θ for 0 < θ < 1 if each (Eα, Fα) is an inter- ℓp(cid:0)Lp(J, E)(cid:1) = Lp(cid:0)J, ℓp(E)(cid:1), polation couple. (9.1) We presuppose as standing hypothesis (ρ, K) is a singularity datum for M. K⋉Φ is a uniformly regular atlas for W over K. F =(cid:0)F, (··)F(cid:1) is a model fiber for W with basis (e1, . . . , en). (cid:8) (πκ, χκ) ; κ ∈ K(cid:9) is a localization system subordinate to K. On the basis of (7.9) we can provide localized versions of the norms k·kkr/~r,p;~ω and k·k∼ kr/~r,p;~ω. Theorem 9.1 Suppose k ∈ N. Set ·kr/~r,p;~ω :=(cid:0)kϕλ p ukp kr/~r,p;~ω :=(cid:16)kϕλ ·∼ p ukp and Then ·kr/~r,p;~ω ∼ k·kkr/~r,p;~ω and ·∼ ℓp(Lp(J,W kr p )) + kϕλ+kµ p (∂ku)kp ℓp(Lp(J,W kr p )) + kXj=0 kr/~r,p;~ω ∼ k·k∼ kr/~r,p;~ω. kϕλ+jµ p (∂ju)kp 27 . ℓp(Lp(J,Lp))(cid:1)1/p ))(cid:17)1/p (k−j)r p ℓp(Lp(J,W P r o o f. This follows from (7.9) and (9.1). It is worthwhile to note u∼ kr/~r,p;~ω =(cid:16)Xκ ZJ Xα+jr≤kr(cid:0)ρλ+α+jµ+m/q κ x ∂j(κ⋉ϕ)∗(πκu)kp;E(cid:1)p k∂α dt(cid:17)1/p . Together with Theorems 8.1(ii) and 9.1 this gives a rather explicit and practically useful local characterization of anisotropic Sobolev spaces. For abbreviation, we set • Yκ := Xκ × J, κ ∈ K. Hence Yκ = Xκ × J is the interior of Yκ in Rm+1 = Rm × R. We also put D(Yκ, E), D(Y, E) :=Mκ D(Y, E) :=Mκ D(Y, E) and W kr/~r p,κ := W kr/~r p (Yκ, E), Fs/~r p,κ := Fs/~r p (Yκ, E), s ∈ R, More precisely, the 'local' spaces W kr/~r namely with M = (Xκ, gm), ρ = 1, W = Xκ × F , and D = dF . and Fs/~r p,κ p,κ are special instances of W kr/~r,~ω p k ∈ N. and Fs/~r,~ω p , respectively, It is of fundamental importance that these spaces coincide with the anisotropic Sobolev, Bessel potential, and Besov spaces studied by means of Fourier analytical techniques in detail in H. Amann [4], therein de- noted by W kν/ν (Yκ, E), respectively, where ν := r and ν := (1, r). For abbreviation, we set W kν/ν for Lp(J, W kr (Yκ, E), and Bs/ν p (Yκ, E), H s/ν p := W kν/ν (Yκ, E). Furthermore, we write fW kr/~r p (J, Lp,κ) endowed with the norm k·k∼ s/ν p,κ := F kr/~r,p. p,κ) ∩ W k (Yκ, E) and F s/ν p p,κ p,κ p p Lemma 9.2 . = W kν/ν . p,κ p,κ p,κ (i) If k ∈ N, then W kr/~r (ii) If s ∈ R, then Fs/~r P r o o f. (1) If J = R+ and κ ∈ K∂M , then Yκ is isomorphic to the closed 2-corner R+ × R+ × Rm−1 (in the sense of Section 4.3 of [4]) by a permutation isomorphism. Otherwise, Yκ equals either the half-space Hm+1 (except for a possible permutation) or Rm+1. =fW kr/~r p,κ for κ ∈ K. for κ ∈ K. . = Fs/ν p,κ (2) If Yκ = Rm+1, then (i) follows from Theorem 2.3.8 of [4] and the definition of W kν/ν p,κ in the first para- graph of [4, Section 3.5]. If Yκ 6= Rm+1, then we obtain claim (i) by invoking [4, Theorem 4.4.3(i)]. (3) Suppose Yκ = Rm+1. Then statement (ii) follows from [4, Theorem 3.7.1]. Let Yκ 6= Rm+1 and s 6= 0 if F = B. Then we get this claim by employing, in addition, [4, Theorems 4.4.1 and 4.4.4]. If Yκ 6= Rm+1, F = B, and s = 0, then we have to use [4, Theorem 4.7.1(ii) and Corollary 4.11.2] in addition. 28 H. Amann: Anisotropic Function Spaces on Singular Manifolds Due to this lemma we can apply the results of [4] to the local spaces F s/~r p,κ. This will be done in the following usually without referring to Lemma 9.2. Let X be a locally convex space and 1 ≤ q ≤ ∞. For κ ∈ K we consider the linear map Θµ defined by q,κ : X J → X J (9.2) Θµ q,κu(t) := ρµ/q κ u(ρµ κt), u ∈ X J , t ∈ J. Note and Moreover, q,κ = id q,κ = Θ0 q,κ ◦ Θ−µ Θµ q,κ(cid:0)C(J,X )(cid:1) ⊂ C(J,X ). Θµ ∂k ◦ Θµ q,κ = ρkµ κ Θµ q,κ ◦ ∂, k ∈ N, and, if X is a Banach space, kΘµ q,κukLq(J,X ) = kukLq(J,X ). We put and ϕ~ω q,κu := Θµ q,κ ◦ ϕλ q,κu, ϕ~ω q u := (ϕ~ω q,κu), q,κvκ := Θ−µ ψ~ω q,κ ◦ ψλ q,κvκ, ψ~ω q v :=Xκ ψ~ω q,κvκ, u ∈ C(cid:0)J, C(V )(cid:1), v = (vκ) ∈Mκ C(Yκ, E). (9.3) (9.4) (9.5) (9.6) (9.7) (9.8) After these preparations we can prove the following analogue to Theorem 7.1. Not only will it play a funda- mental role in this paper but also be decisive for the study of parabolic equations on singular manifolds. Theorem 9.3 Suppose s ∈ R. Then the diagrams ✄ ✂ D(J, D) ❅ ❄ ✠ D(J, D) ✄ ✂ ϕ~ω p ❅❘ ψ~ω p d d d ✄ ✂ ✲ ✲ Fs/~r,~ω p ϕ~ω p ✠ ❅ ψ~ω p ✲ ❅❘ ❄ Fs/~r,~ω p ✄ ✂ D( J , D) ❅ ❄ ✠ D( J , D) ✄ ✂ ϕ~ω p ❅❘ ψ~ω p d d d ✄ ✂ ✲ ✲ Fs/~r,~ω p ℓp(Fs/~r p ) ϕ~ω p ✠ id ❅ ψ~ω p ✲ ❅❘ ❄ Fs/~r,~ω p id D(Y, E) ℓp(Fs/~r p ) id id D(Y, E) are commuting, where s > 0 in the second case. P r o o f. (1) It is not difficult to see that D(J,Dκ) = D(Yκ, E) by means of the identification u(t) = u(·, t) for t ∈ J (see Corollary 1 in Section 40 of F. Treves [49], for example). Consequently, Similarly, D( J, Dκ) = D(Yκ, E), and thus D(Y, E) =Mκ D(Y, E) =Mκ D(J,Dκ). D( J , Dκ). Using this, (9.4), and (9.5), obvious modifications of the proof of Theorem 5.1 in [5] show that the assertions encoded in the respective left triangles of the diagrams are true. (2) Suppose k ∈ N. From (9.6) we get kϕ~ω p,κukLp(J,W kr p,κ) = kϕλ p,κukLp(J,W kr p,κ). Hence, using (9.1) kϕ~ω From this and Theorem 7.1 we deduce p ukℓp(Lp(J,W kr p )) = kϕλ p ukLp(J,ℓp(W kr p )). that is, p ukℓp(Lp(J,W kr kϕ~ω p ∈ L(cid:0)Lp(J, W kr,λ p,κ ) = kϕλ+jµ Consequently, invoking (9.1) and Theorem 7.1 once more, By means of (9.5) and (9.6) we obtain p,κukLp(J,W k−j k∂jϕ~ω ϕ~ω p p )) ≤ ckukLp(J,W kr,λ p ), ), ℓp(Lp(J, W kr p ))(cid:1). p,κ (∂ju)kLp, 0 ≤ j ≤ k. This, together with (9.9), implies (3) Note that where k∂kϕ~ω p ukℓp(Lp(J,Lp)) ≤ ck∂kukLp(J,Lλ+kµ p ∈ L(cid:0)W kr/~r,~ω , ℓp(W kr/~r )(cid:1). ϕ~ω p p p ). ϕλ p,κψλ p,eκ = aeκκSeκκ, Lemma 7.2, estimate (4.3), and (7.3)(iii) imply aeκκ ∈ BCk(Xκ), Hence we infer from (9.12) and Lemma 7.2 aeκκ := (ρκ/ρeκ)λ+m/p(κ∗πκ)Seκκ(eκ∗πeκ). eκ ∈ N(κ), kϕλ kaeκκkk,∞ ≤ c(k), p,eκ ∈ L(W k p,eκ, W k p,κψλ p,κ), p,eκk ≤ c(k) ϕλ p,κψλ κ ∈ K, k ∈ N. p,eκvkLp(J,W k p,κ) ≤ ckvkLp(J,W k p, eκ) (∂kv)kLp(J,Lp,κ) ≤ ck∂kvkLp(J,Lp, eκ) (9.14) 29 (9.9) (9.10) (9.11) (9.12) (9.13) (9.15) (9.16) kϕλ p,κψλ p,κψ~ω p,κ) = kϕλ p,eκvkLp(J,W k foreκ ∈ N(κ), κ ∈ K, and k ∈ N. By this and (9.6) we find foreκ ∈ N(κ), κ ∈ K, and k ∈ N. Similarly, using (9.5), foreκ ∈ N(κ), κ ∈ K, and k ∈ N. p,eκv)(cid:1)(cid:13)(cid:13)Lp(J,Lp,κ) = kϕλ+kµ (cid:13)(cid:13)ϕλ+kµ (cid:0)∂k(ψ~ω Observe p,κ ψ(λ+kµ,µ) p,eκ p,κ p,κψ~ω ϕλ p v = Xeκ∈N(κ) p,κψ~ω ϕλ p,eκveκ. From (9.13) -- (9.15) and the finite multiplicity of K we infer and Hence Theorem 9.1 implies p (ψ~ω kϕλ (cid:0)∂k(ψ~ω kψ~ω p (cid:13)(cid:13)ϕλ+kµ p v)kℓp(Lp(J,W kr p )) ≤ ckvkℓp(Lp(J,W kr p )) p v)(cid:1)(cid:13)(cid:13)ℓp(Lp(J,Lp)) ≤ ck∂kvkℓp(Lp(J,Lp)). p vkkr/~r,p:~ω ≤ ckvkℓp(W kr/~r p ), 30 that is, It follows from ψλ p ϕλ p = id that ψ~ω p ϕ~ω p = id. Thus we see from (9.11) and (9.17) that the diagram ψ~ω p ∈ L(cid:0)ℓp(W kr/~r p ), W kr/~r,~ω p (cid:1). H. Amann: Anisotropic Function Spaces on Singular Manifolds (9.17) (9.18) W kr/~r,~ω p ❍❍❍❍❥ ϕ~ω p id ✲ W kr/~r,~ω p ✟✟✟✟✯ ψ~ω p is commuting. ℓp(W kr/~r p ) (4) It is a consequence of Lemma 9.2(i) and [4, Theorems 2.3.2(i) and 4.4.1] that D(J,Dκ) = D(Yκ, E) is dense in W kr/~r p,κ . This implies D(Y, E) d ֒→Mκ W kr/~r p,κ = cc(W kr/~r p ). Hence, by (7.2), Thus we deduce from step (1) and (9.18) that D(Y, E) d ֒→ ℓp(W kr/~r p ). (9.19) ✄ ✂ D(J, D) ❅ id ❄ ✠ D(J, D) ✄ ✂ ϕ~ω p ❅❘ ψ~ω p D(Y, E) d ✲ ✄ ✂ ℓp(W kr/~r p ) id ✲ p W kr/~r,~ω ϕ~ω p ✠ ❅ ψ~ω p ✲ ❄ ❅❘ W kr/~r,~ω p is a commuting diagram. From this and [4, Lemma 4.1.6] we obtain (5) Suppose k ∈ N and kr < s ≤ (k + 1)r. If s < (k + 1)r, set θ := (s − kr)/r and ℓ := k + 1. Otherwise, θ := 1/2 and ℓ := k + 2. Then we infer from (9.18) and (8.3) by interpolation that ψ~ω p is a retraction from D(J,D) d ֒→ W kr/~r,~ω p . (9.20) (cid:0)ℓp(W kr/~r p ), ℓp(W ℓr/~r p )(cid:1)θ (9.21) onto Fs/~r,~ω p norms. . By Theorem 1.18.1 in H. Triebel [50], (9.21) equals ℓp(cid:0)(W kr/~r p , W ℓr/~r p )θ(cid:1), except for equivalent p,κ )θ . = Fs/~r It follows from Lemma 9.2 and [4, Theorem 3.7.1(iv), formula (3.3.12), and Theorems 3.5.2 and 4.4.1] that (W kr/~r p,κ , W ℓr/~r p,κ. This shows that the right triangle of the first diagram is commuting if s > 0. Fur- thermore, the density properties of the interpolation functor (·,·)θ, (9.19), and (9.20) imply that the 'horizontal embeddings' of the first diagram of the assertion are dense if s > 0. This proves the first assertion for s > 0. (6) It is a consequence of what has just been shown and step (1) that the second part of the statement is true. (7) Let X be a reflexive Banach space. Then hv, Θµ We define ϕ~ω From this we infer (cf. the proof of Theorem 5.1 in [5]) p,κuiLp(J,X) = hΘ−µ p and ψ~ω p by replacing ϕλ p′,κv, uiLp(J,X), p,κ and ψλ u ∈ Lp(J, X), v ∈ Lp′(J, X ′) =(cid:0)Lp(J, X)(cid:1)′ . p,κ in (9.7) and (9.8) by ϕλ p,κ and ψλ p,κ, defined in (7.5), respectively. hψ−~ω p′ v, uiM×J = hhv, ϕ~ω p uii, v ∈ D(Y, E), u ∈ D(J,D), (9.22) and hhϕ−~ω p′ v, uii = hv, ψ~ω p uiM×J , Moreover, (8.5) implies for s > 0 ℓp(F−s/~r p v ∈ D( J, D), u ∈ D(Y, E). p′ (Y, E′)(cid:1)(cid:1)′ s/~r . ) =(cid:0)ℓp′(cid:0)F It follows from (7.6) that ϕ~ω deduce from (9.22) and (9.23) that, given s > 0, p and ψ~ω p possess the same mapping properties as ϕ~ω p and ψ~ω p , respectively. Hence we kϕ~ω p ukℓp(F −s/~r p ) ≤ ckukF , −s/~r,~ω p u ∈ D(J,D), and kψ~ω p ukF −s/~r,~ω p ≤ ckukℓp(F −s/~r p ), u ∈ D(Y, E). We infer from (9.20), Theorem 8.3(i), and reflexivity that D(J,D) is dense in F−s/~r,~ω p . Hence p ∈ L(cid:0)F−s/~r,~ω Since, as above, D(J,Dκ) = D(Yκ, E) is dense in F−s/~r ϕ~ω p , ℓp(F−s/~r p p,κ we see, by the arguments used to prove (9.19), )(cid:1). (cid:1). D(Y, E) d ֒→ ℓp(F−s/~r p ). ψ~ω p ∈ L(cid:0)ℓp(F−s/~r p ), F−s/~r,~ω p 31 (9.23) (9.24) (9.25) (9.26) (9.27) Thus (9.24) implies Thus, as in step (5), From (9.25) -- (9.27) and step (1) it now follows that the first statement is true if s < 0. (8) Suppose s = 0. If F = H, then assertion (i) is contained in (9.18) (for k = 0). If F = B, then we deduce from Lemma 9.2(ii) and [4, Theorems 3.7.1, 4.4.1, 4.7.1(ii), and Corollary 4.11.2] that (H −s(p)/~r p,κ , H s(p)/~r p,κ )1/2,p . = B0/~r p,κ , (cid:0)ℓp(H −s(p)/~r p ), ℓp(H s(p)/~r p )(cid:1)1/2,p κ ∈ K. . = ℓp(B0/~r p ). Since we have already shown that ψ~ω tion (8.6) that it is a retraction from ℓp(B0/~r p p is a retraction from ℓp(H ±s(p)/~r p ) onto H ±s(p)/~r p , it follows from defini- ) onto B0/~r,~ω p . This proves the theorem. Now we can supply the proofs left out in Section 8. First note that assertion (iii) of Theorem 8.1 has been shown in (9.20). P r o o f of part (ii) of Theorem 8.1. It is a consequence of Lemma 9.2(i) that Hence, due to (7.8) and (9.18), Using (9.9) and (9.10) one verifies kϕ~ω Now the assertion follows from Theorem 9.1. ). . p = ℓp(fW kr/~r ) ∼ k·kkr/~r,p;~ω. ) ∼ ·∼ kr/~r,p;~ω. ℓp(W kr/~r p ) kϕ~ω p p · kℓp(fW kr/~r p · kℓp(fW kr/~r p 32 H. Amann: Anisotropic Function Spaces on Singular Manifolds P r o o f of Theorem 8.2. (1) Lemma 9.2 and [4, Theorems 3.7.1 and 4.4.3(i)] imply H kr/~r p,κ . = W kr/~r p,κ for κ ∈ K and k ∈ N. Hence ℓp(H kr/~r p ) . = ℓp(W kr/~r p ), k ∈ N, and assertion (i) is a consequence of Theorem 9.3. (2) In order to prove (ii) it suffices, due to Theorem 9.3 and Lemma 9.2, to show H s/ν (Yκ, E) . = Bs/~r 2 (Yκ, E). 2 By the results of Section 4.4 of [4] we can assume Yκ = Rm+1. Suppose s > 0 and write H s 2 := H s 2 (Rm, E), etc. Then [4, Theorem 3.7.2] asserts H s/ν 2 = Lp(R, H s 2 ) ∩ H s/ν 2 (R, L2). From Theorem 3.6.7 of [4] we get Bs/ν 2 = L2(R, Bs 2) ∩ Bs/ν 2 (R, L2). By Theorem 2.12 in [50] we know that H s 2 W. Sickel [42] guarantee H s/ν follows by duality. (R, L2) . = Bs/ν 2 2 . = Bs (R, L2). This proves H s/ν 2. Remark 7 and Proposition 2(1) in H.-J. Schmeisser and for s > 0. The case s < 0 . = Bs/ν 2 2 From Lemma 9.2(ii) and [4, (3.4.1) and Theorem 3.7.1] we get [F−s(p)/~r , Fs(p)/~r 2 ]1/2 . = F0/~r 2 . Thus, by what 2 we already know, B0/~r 2 . = [B−s(p)/~r 2 , Bs(p)/~r 2 ]1/2 . = [H −s(p)/~r 2 , H s(p)/~r 2 ]2 . = H 0/~r 2 . This settles the case s = 0 also. (3) By [4, (3.3.12), (3.4.1), and Theorems 3.7.1(iv) and 4.4.1] we know that assertions (iii) and (iv) hold for p,κ. Thus we get (iii) and (iv) in the general case by the arguments of step (5) of the proof of the local spaces Fs/~r Theorem 9.3. P r o o f of Theorem 8.3(i) for s = 0. Since (8.7) has already been established for s ∈ R\{0} it remains to show that H s1/~r,~ω p d ֒→ B0/~r,~ω p d ֒→ H s0/~r,~ω p if −1 + 1/p < s0 < 0 < s1 < 1/p. By [4, Theorems 3.7.1(iii), 4.4.1, 4.7.1(ii), and Corollary 4.11.2] H s1/ν p,κ d ֒→ B0/ν p,κ d ֒→ H s0/ν p,κ . From this and Lemma 9.2 we deduce Now the claim follows from Theorem 9.3. ℓp(H s1/~r p ) d ֒→ ℓp(B0/~r p ) d ֒→ ℓp(H s0/~r p ). P r o o f of Theorem 8.3(ii). If J = R and ∂M = ∅, then the claim is obvious by (8.4), D( J , D) = D(J,D), and (i). Otherwise, we get from [4, Theorem 4.7.1 and Corollary 4.11.2], due to the stated restrictions for s, that s/~r F p,κ = F s/~r p,κ. Here we also used the fact that D(J,Dκ) d ֒→ F0/~r κ d ֒→ F−t/~r κ , t > 0, κ ∈ K. Hence ℓp(Fs/~r p ) = ℓp(Fs/~r p ) and the claim follows from (the right triangles of the diagrams of) Theorem 9.3. 10 Renorming of Besov Spaces Let X be a Banach space and X ∈ {Rm, Hm}. For u : X → X and h ∈ Hm\{0} we put △hu := u(· + h) − u, △k+1 h u := △h△k hu, k ∈ N, △0 hu := u. Given k ≤ s < k + 1 with s > 0, where k·kp;X := k·kLp(X,X ). We set for s > 0 [u]s,p;X :=(cid:16)ZX(cid:16)k△k+1 s,p;X :=(cid:0)k·kp k·k∗ Suppose k ≤ s < k + 1 with k ∈ N and s > 0. Then kukk,p;X :=(cid:16)Xα≤k is the norm of the X -valued Sobolev space W k h ukp;X hs (cid:17)p dh hm(cid:17)1/p , p;X + [·]p . s,p;X(cid:1)1/p p;X(cid:17)1/p x ukp k∂α kuk∗∗ s,p;X := [∂α p (X,X ) and x u]p k,p;X + Xα=k k−1,p;X + Xα=k−1 (cid:16)kukp (cid:16)kukp , s−k,p;X(cid:17)1/p 1,p;X(cid:17)1/p x u]p [∂α  k < s < k + 1, , s = k ∈ N×. Then, given s > 0, is a Banach space, an X -valued Besov space, Bs p(X,X ) :=(cid:0)(cid:8) u ∈ Lp(X,X ) ; [u]s,p;X < ∞(cid:9), k·k∗ s,p;X(cid:1) k·k∗ s,p;X ∼ k·k∗∗ s,p;X , 33 (10.1) d and D(X,X ) valued results (e.g., H.-J. Schmeisser [41] or H. Amann [3]). p(X,X ). These facts can be derived by modifying the corresponding well-known scalar- ֒→ Bs Now we choose X = J. Note that h ◦ Θµ △k q,κ = Θµ q,κ ◦ △k κh, ρµ 1 ≤ q ≤ ∞. Hence (9.6) implies [Θµ p,κu]s,p;X = ρµs κ [u]s,p;X . Suppose s > 0. Then s/~r,p;~ω :=(cid:0)kukp kuk∗ and, if kr < s ≤ (k + 1)r with k ∈ N, s/~r,p;~ω :=(cid:16)kukp +Xj≤k k∂jukp kuk∗∗ p;Bs,λ p p;Bs,λ p + [u]p s/r,p;Lλ+sµ/r p (cid:1)1/p p;W (k−j)r,λ+µj p + [∂ku]p (s−kr)/r,p;Lλ+sµ/r p Besides of these norms we introduce localized versions of them by u∗ s/~r,p;~ω :=(cid:0)kϕλ p ukp p;ℓp(Bs p) + [ϕλ+sµ/r p u]p s/r,p;ℓp(Lp)(cid:1)1/p (10.2) (10.3) (10.4) (10.5) (cid:17)1/p . 34 H. Amann: Anisotropic Function Spaces on Singular Manifolds and, if kr < s ≤ (k + 1)r, u∗∗ s/~r,p;~ω :=(cid:16)kϕλ p ukp p;ℓp(Bs p) +Xj≤k k∂jϕλ+jµ p + [∂kϕλ+sµ/r p u]p ) p;ℓp(W (k−j)r p ukp (s−kr)/r,p;ℓp(Lp)(cid:17)1/p Theorem 10.1 Suppose s > 0. Then (10.3) -- (10.6) are equivalent norms for Bs/~r,~ω . p P r o o f. (1) It follows from (9.6) that kϕλ p ukp;ℓp(Bs p) = kϕ~ω p ukp;ℓp(Bs p). Using (10.2) we get Thus, by Fubini's theorem, From this and (10.7) we obtain [ϕ~ω p,κu]s/r,p;Lp,κ = [ϕλ+sµ/r p,κ u]s/r,p;Lp,κ. [ϕ~ω p u]s/r,p;ℓp(Lp) = [ϕλ+sµ/r p u]s/r,p;ℓp(Lp). . (10.6) (10.7) (10.8) u∗ Similarly, invoking (9.5) as well, s/~r,p;~ω =(cid:0)kϕ~ω p ukp s/~r,p;~ω =(cid:16)kϕ~ω p) +Xj≤k p ukp p;ℓp(Bs u∗∗ p;ℓp(Bs p) + [ϕ~ω p u]p s/r,p;ℓp(Lp)(cid:1)1/p . (10.9) k∂jϕ~ω p ukp p;ℓp(W (k−j)r p ) + [∂kϕ~ω p u]p (s−kr)/r,p;ℓp(Lp)(cid:17)1/p if kr < s ≤ (k + 1)r. (2) Lemma 9.2 and [4, Theorems 3.6.3 and 4.4.3] imply p,κ) ∩ Bs/r . = Lp(J, Bs Bs/~r p,κ p Hence (J, Lp,κ), κ ∈ K. due to Bs k·kBs/~r p,κ ∼ k·kp;Bs + [·]s/r,p;Lp,κ p,κ ֒→ Lp,κ. From this, (10.9), and Fubini's theorem we deduce p · kℓp(Bs/~r p ). Thus (7.8) and Theorem 9.3 guarantee that (10.5) is a norm for Bs/~r,~ω is a norm for Bs/~r,~ω s/~r,p;~ω ∼ kϕ~ω ·∗ p,κ p . p , . Similarly, using (10.1), we see that (10.6) (3) We set α := λ + sµ/r and β := α + τ − σ. Then we deduce from (4.1)(iv), (5.11), (7.6), and [5, Lem- ma 3.1(iii)] dVgm dt dξ ξ1+ps/r [ϕα s/r,p;Lp,κ p,κu]p =Z ∞ 0 ZJZXκ(cid:0)ρα+m/p ∼Z ∞ 0 ZJZXκ =Z ∞ 0 ZJZUκ for u ∈ D(J,D). We insert 1 =Peκ π2 summation. Then ξ κ (cid:0)(κ⋉ϕ)∗(πκu)(cid:1)(cid:12)(cid:12)E(cid:1)p (cid:12)(cid:12)△k+1 ξ uh)p dVg(cid:1) dt κ∗(cid:0)(ρβπκ △k+1 (ρβπκ △k+1 ξ uh)p dVg dt ξ1+ps/r dξ dξ ξ1+ps/r [ϕα p u]p s/r,p;ℓp(Lp) ∼Xeκ Xκ∈N(eκ)Z ∞ 0 ZJZUeκ eκ(ρβπκ △k+1 π2 ξ uh)p dVg dt dξ ξ1+ps/r . eκ in the inner integral, sum over κ ∈ K, and interchange the order of Using (7.3)(iii) and the finite multiplicity of K we see that the last term can be bounded above by cXeκ Z ∞ 0 ZJZUeκ = cZ ∞ 0 ZJZM eκ(ρβ △k+1 π2 ξ uh)p dVg dt dξ ξ1+ps/r (ρβ △k+1 ξ uh)p dVg dt ξ1+ps/r = c [u]p s/r,p;Lα p . dξ Hence, recalling (10.8), [ϕ~ω p u]s/r,p;ℓp(Lp) ≤ c [u]s/r,p;Lλ+sµ/r p , u ∈ D(J,D). (4) It is a consequence of Theorem 7.1 that ϕλ kϕ~ω p ukp;ℓp(Bs p) = kϕλ p ukp;ℓp(Bs p) ≤ ckukp;Bs,λ p , p ∈ L(cid:0)Bs,λ p , ℓp(Bs p)(cid:1). This implies, due to (10.7), u ∈ D(J,D). Thus we obtain from (10.9), (10.10), and (10.11) u∗ s/~r,p;~ω ≤ ckuk∗ s/~r,p;~ω, u ∈ D(J,D). 35 (10.10) (10.11) (10.12) We denote by B∗s/~r,~ω p and step (2) imply the completion of D(J,D) in Lp(J, Lλ p ) with respect to the norm k·k∗ s/~r,p;~ω. Then (10.12) B∗s/~r,~ω p ֒→ Bs/~r,~ω p . (5) Observing ψ~ω p,κ = χκψ~ω p,κ and 0 ≤ χκ ≤ 1, the finite multiplicity of K implies ξ ψ~ω △k+1 p vh =(cid:12)(cid:12)(cid:12)Xκ ≤ c(cid:16)Xκ △k+1 ξ ψ~ω ξ ψ~ω △k+1 p,κvκ(cid:12)(cid:12)(cid:12)h ≤(cid:16)Xκ h(cid:17)1/p p,κvκp ξ ψ~ω △k+1 p,κvκp h(cid:17)1/p(cid:16)Xκ χκ(cid:17)1/p′ p [ψ~ω p v]p s/r,p;Lλ+sµ/r △k+1 for v ∈ D(Y, E). Hence, reasoning as in step (3), ≤ cZ ∞ 0 ZJZM ρβpXκ 0 ZJZXκXκ ≤ cZ ∞ △k+1 ≤ cXκ s/r,p;Lp,κ ≤ cXκ ≤ cXκ (6) Theorem 7.1 and (7.9) guarantee that kϕλ p ≤ ckϕλ p · kℓp(Bs p ψ~ω for v ∈ D(Y, E). p vkp;ℓp(Bs p), p vkp;Bs,λ kvκkp [πκvκ]p kψ~ω (J,Lp,κ) Bs/r ξ p dξ ξ1+ps/r dξ ξ1+ps/r ξ ψ~ω p,κvκp h dVg dt (πκvκ)p [vκ]p gm dVgm dt s/r,p;Lp,κ p) is an equivalent norm for Bs,λ p . This implies v ∈ D(Y, E). (10.13) From (9.16) we infer by interpolation, using the arguments of step (5) of the proof of Theorem 9.3, that kϕλ Hence (10.13) and (9.1) imply p ψ~ω p vkℓp(Lp(J,Bs p)) ≤ ckvkℓp(Lp(J,Bs p)), v ∈ D(Y, E). kψ~ω p vkp;Bs,λ p ≤ ckvkp;ℓp(Bs p), v ∈ D(Y, E). 36 H. Amann: Anisotropic Function Spaces on Singular Manifolds By combining this with the result of step (5) we find, employing (9.1) once more, Thus, by Theorem 9.3, kψ~ω p vk∗ s/~r,p;~ω ≤ ckvkℓp(B s/~r p ), v ∈ D(Y, E). kuk∗ s/~r,p;~ω = kψ~ω p (ϕ~ω p u)k∗ s/~r,p;~ω ≤ ckϕ~ω p ukℓp(Bs/~r p ) = cu∗ s/~r,p;~ω, u ∈ D(J,D), the last estimate being a consequence of (10.9). Since, by step (2), (10.5) is a norm for Bs/~r,~ω p , we get This implies Bs/~r,~ω p ֒→ B∗s/~r,~ω p s/~r,p;~ω ≤ ckukBs/~r,~ω kuk∗ . From this and step (4) it follows that (10.3) is a norm for Bs/~r,~ω u ∈ D(J,D). , p p . (7) The proof of the fact that (10.4) is a norm for Bs/~r,~ω p Corollary 10.2 If s > 0, then Bs/~r,(λ,0) p . = Lp(J, Bs,λ is similar. p ) ∩ Bs/r p (J, Lλ p ). 11 Holder Spaces in Euclidean Settings In [5] it has been shown that isotropic weighted Holder spaces are important point-wise multiplier spaces for weighted isotropic Bessel potential and Besov spaces. In Section 13 we shall show that similar results hold in the anisotropic case. For this reason we introduce and study anisotropic weighted Holder spaces and establish the fundamental retraction theorem which allows for local characterizations. In order to achieve this we have to have a good understanding of Holder spaces of Banach-space-valued functions on Rm and Hm. In this section we derive those properties of such spaces which are needed to study weighted Holder spaces on M . Let X be a Banach space. Suppose X ∈ {Rm, Hm} and X ∈ {X, X × J}. Then B = B(X,X ) is the Banach Throughout this section, k, k0, k1 ∈ N. Then space of all bounded X -valued functions on X endowed with the supremum norm k·k∞ = k·k0,∞. x u ∈ B(X,X ), α ≤ k(cid:9), k·kk,∞(cid:1), BCk = BCk(X,X ) :=(cid:0)(cid:8) u ∈ Ck(X,X ) ; ∂α α≤k k∂α kukk,∞ := max x uk∞, where is a Banach space. As usual, BC = BC0. We write k·kk,∞;X for k·kk,∞ if it seems to be necessary to indicate the image space. Similar conventions apply to the other norms and seminorms introduced below. Note that is a closed linear subspace of BCk. The mean value theorem implies the first embedding of BUCk =(cid:8) u ∈ BCk ; ∂α x u is uniformly continuous for α ≤ k(cid:9) Hence It is a Fr´echet space with the natural projective topology. Thus BCk+1 ֒→ BUCk ֒→ BCk. BC∞ :=TkBCk =TkBUCk. BC∞ ֒→ BUCk, k ∈ N. (11.1) (11.2) In fact, this embedding is dense. For this we recall that a mollifier on Rd is a family { wη ; η > 0 } of nonnegative compactly supported smooth functions on Rd such that wη(x) = η−dw1(x/η) for x ∈ Rd and R w1 dx = 1. Then, denoting by wη ∗ u convolution, wη ∗ u ∈ BC∞(Rd,X ), u ∈ BC(Rd,X ), (11.3) and lim η→0 wη ∗ u = u in BUCk(Rd,X ), u ∈ BUCk(Rd,X ), 37 (11.4) (cf. [7, Theorem X.7.11], for example, whose proof carries literally over to X -valued spaces). From this we get (11.5) BC∞ d ֒→ BUCk if X = Rm and J = R. In the other cases it follows by an additional extension and restriction argument based on the extension map (4.1.7) of [4] (also cf. Section 4.3 therein). From now on X = X. For k ≤ s < k + 1, 0 < δ ≤ ∞, and u : X → X we put [·]s,∞ := [·]∞ s,∞. s,∞ := sup [u]δ , h∈(0,δ)m k△k+1 h uk∞;X hs Furthermore, Note that h ∈ (0,∞)m\(0, δ)m implies δ ≤ h∞ ≤ h ≤ √mh∞. Hence k·k∗ s,∞ := k·k∞ + [·]s,∞, s > 0. If 0 < θ0 < θ ≤ 1, then Consequently, This implies [·]θ,∞ ≤ [·]δ θ,∞ + 4δ−θ k·k∞, θ0,∞ ≤ √m δθ−θ0[·]δ [·]δ 0 < θ ≤ 1, 0 < δ < ∞. θ,∞, 0 < δ < ∞. [·]θ0,∞ ≤ √m [·]1 θ,∞ + 4 k·k∞ ≤ √m [·]θ,∞ + 4 k·k∞. k·k∗ θ0,∞ ≤ c(m)k·k∗ θ,∞, 0 < θ0 < θ ≤ 1. Suppose u ∈ BCk and denote by D the Fr´echet derivative. Then, by the mean value theorem, 0 ···Z 1 where [h]k := (h, . . . , h) ∈ Xk. From this we get hu(x) =Z 1 △k 0 Dku(cid:0)x + (t1 + ··· + tk)h(cid:1)[h]k dt1 ··· dtk, [u]δ θ,∞ ≤ mk/2δk−θ kukk,∞, 0 < θ ≤ 1, θ < k, δ > 0, u ∈ BCk. Thus, by (11.6), We also set for k < s ≤ k + 1 k·k∗ θ,∞ ≤ c(m)k·k1,∞, 0 < θ < 1. (11.6) (11.7) (11.8) (11.9) (11.10) kuk∗∗ If k < s < k + 1, then k·ks,∞ := k·k∗∗ s,∞ and s,∞ := kukk,∞ + max α=k [∂α x u]s−k,∞. BCs = BCs(X,X ) :=(cid:0)(cid:8) u ∈ BCk ; max α=k [∂α x u]s−k,∞ < ∞(cid:9), k·ks,k(cid:1), k < s < k + 1, is a Holder space of order s. Given h = (h1, . . . , hm) ∈ X, we set hj := (0, . . . , 0, hj, . . . , hm) for 1 ≤ j ≤ m, and hm+1 := 0. Then △hu(x) = mXj=1(cid:0)u(x + hj) − u(x + hj+1)(cid:1). 38 H. Amann: Anisotropic Function Spaces on Singular Manifolds From this we infer for 0 < θ < 1 and hj 6= 0 for 1 ≤ j ≤ m ku(· + hjej) − uk∞ ≤ k△huk∞ ≤ hθ hjθ mXj=1 sup hj 6=0 ku(· + hjej) − uk∞ hjθ mXj=1 mXj=1 = Consequently, ku(· + hjej) − uk∞ (hj)θ sup hj >0 ≤ m [u]θ,∞. [u]θ,∞ ≤ sup h6=0 k△huk∞ hθ ≤ m [u]θ,∞, 0 < θ < 1. (11.11) This shows that BCs coincides, except for equivalent norms, with the usual Holder space of order s if s ∈ R+\N. From (11.11) we read off that the last embedding of is valid. The other two follow from (11.8) and (11.10). BCk+1 ֒→ BCs ֒→ BCs0 ֒→ BUCk, k < s0 < s < k + 1, (11.12) We introduce the Besov-Holder space scale [ Bs ∞ ; s > 0 ] by ∞ :=( (BUCk, BUCk+1)s−k,∞, k < s < k + 1, (BUCk, BUCk+2)1/2,∞, s = k + 1. Bs Theorem 11.1 (i) k·k∗ (ii) Bs ∞ s,∞ are norms for Bs ∞. s,∞ and k·k∗∗ . = (BUCk0 , BUCk1 )(s−k0)/(k1−k0),∞ for k0 < s < k1. . = [Bs0 (iii) If 0 < s0 < s1 and 0 < θ < 1, then (Bs0 . = Bsθ ∞ ∞, Bs1 ∞)θ,∞ ∞, Bs1 ∞]θ. P r o o f. (1) For s > 0 we denote by Bs ∞,∞ = Bs ∞,∞(X,X ) the 'standard' Besov space modeled on L∞ for whose precise definition we refer to [4] (choosing the trivial weight vector therein). It is a consequence of [4, (3.3.12), (3.5.2), and Theorem 4.4.1] that . = (BUCk0 , BUCk1 )(s−k0)/(k1−k0),∞, Bs ∞,∞ k0 < s < k1. This implies Bs ∞ . = Bs ∞,∞ (11.13) and, consequently, statement (ii). (2) The first part of (iii) follows by reiteration from (ii). For ξ ∈ Rm we set Λ(ξ) := (1 + ξ2)1/2. Given s ∈ R, we put Λs := F −1ΛsF, where F = Fm is the Fourier transform on Rm. Suppose X = Rm. It follows from [4, Theorem 3.4.1] and (11.13) that Λs ∈ Lis(Bt+s ∞ , Bt ∞), (Λs)−1 = Λ−s, t, s + t > 0. (11.14) We set A := −Λs1−s0, considered as a linear operator in Bs0 ∞. Then [4, Proposition 1.5.2 and Theorem 3.4.2] guarantee the existence of ϕ ∈ (π/2, π) such that the sector Sϕ := { z ∈ C ; arg z ≤ ϕ} ∪ {0} belongs to the resolvent set of A and k(λ − A)−1k ≤ c/λ for λ ∈ Sϕ. Furthermore, by [4, Proposition 1.5.4 ∞) and there exists γ > 0 such that kAzk ≤ ceγ Im z for Re z ≤ 0. and Theorem 3.4.2] we find that Az ∈ L(Bs0 Now Seeley's theorem, more precisely: the proof in R. Seeley [44], and (11.14) imply [Bs0 ∞. This ∞]θ proves the second part of (iii) if X = Rm. The case X = Hm is then covered by [4, Theorem 4.4.1]. ∞ with domain Bs1 . = Bsθ ∞, Bs1 (3) By [4, Theorems 3.3.2, 3.5.2, and 4.4.1] we get Bs proof of [4, Theorem 4.4.3(i)] we infer from [4, Theorem 3.6.1] that k·kBs Theorem 1.13.1] in the proof of [4, Theorem 3.6.1] we obtain similarly k·kBs (11.12) in the usual extension-restriction argument. Due to (11.13) this proves (i). ∞,∞ ֒→ BUC. Using this and the arguments of the s,∞. By appealing to [50, s,∞, making also use of ∞,∞ ∼ k·k∗ ∞,∞ ∼ k·k∗∗ Corollary 11.2 39 . = BCs for s ∈ R+\N. (i) Bs ∞ (ii) BUCk ֒→ Bk P r o o f. (i) is implied by part (i) of the theorem. (ii) The first claim is a consequence of [4, Theorem 3.5.2]. It follows from Example IV.4.3.1 in E. Stein [48] ∞ contains functions which are not uniformly Lipschitz continuous. This proves the ∞ and BUCk 6= Bk ∞. that the 'Zygmund space' B1 second statement. By (11.12) we see that BCs1 ֒→ BCs0 , 0 ≤ s0 < s1. However, these embeddings are not dense. Since dense embeddings are of great importance in the theory of elliptic and parabolic differential equations we introduce the smaller subscale of 'little' Holder spaces which enjoy the desired property. Suppose s ∈ R+. The little Holder space Similarly, the little Besov-Holder space scale [ b∞ ; s > 0 ] is defined by bcs = bcs(X,X ) is the closure of BC∞ in BCs. ∞ is the closure of BC∞ in Bs bs ∞. (11.15) These spaces possess intrinsic characterizations. Theorem 11.3 (i) bck = BUCk. . = bcs for s ∈ R+\N. (ii) bs ∞ (iii) Suppose k < s ≤ k + 1. Then u ∈ Bs lim δ→0 ∞ belongs to bs ∞ iff [∂α x u]δ s−k,∞ = 0, α = k. (11.16) (iv) BCs d ֒→ bs0 ∞ for 0 < s0 < s. P r o o f. (1) Assertion (i) is a consequence of (11.5). Statement (ii) follows from Corollary 11.2(i). ∞ the linear subspace of Bs ∞ of all u satisfying (11.16). Then we infer from (11.9) that (2) Suppose k < s ≤ k + 1. We denote byebs Let u ∈ bs we can find δε > 0 such that [∂α x v]δε ∞ and ε > 0. Then (11.12) implies the existence of v ∈ BUCk+2 with ku − vk∗∗ BC∞ ֒→ BUCk+1 ֒→ebs ∞. [∂α x u]δ s−k,∞ ≤(cid:2)∂α for α = k and δ ≤ δε. This proves bs (3) Suppose X = Rm and u ∈ebs x (wη ∗ u) = wη ∗ ∂α s−k,∞ ≤ ku − vk∗∗ s−k,∞ ≤ ε/2 for α = k and 0 < δ ≤ δε. Hence x (u − v)(cid:3)s−k,∞ + [∂α ∞ ⊂ebs ∞. We claim that wη ∗ u converges in Bs x u we can assume 0 < s ≤ 1 and then have to show [wη ∗ u − u]s,∞ → 0 as η → 0. and ∂α x v]δε ∞. Note wη ∗ u(x) − u(x) =Z (cid:0)u(x − y) − u(x)(cid:1)wη(y) dy. (11.17) s,∞ < ε/2. By (11.17) s,∞ + ε/2 < ε ∞ towards u as η → 0. Using (11.4) (11.18) 40 From this we infer H. Amann: Anisotropic Function Spaces on Singular Manifolds [wη ∗ u − u]δ s,∞ ≤ 2 [u]δ s,∞, δ > 0. (11.19) Fix δε > 0 such that [u]δ s,∞ < ε/4. Then we get from (11.6) and (11.19) that there exists ηε > 0 such that [wη ∗ u − u]s,∞ ≤ ε/2 + 4δ−s ε kwη ∗ u − uk∞ ≤ ε ∞. ∞ ⊂ bs ∞ ֒→ BUC and (11.4). This proves (11.18). Thusebs ∞ from (3) and a standard extension and restriction argument based on the extension operator (4.1.7) of [4]. Together with the result of step (2) this proves claim (iii). The last assertion follows from (11.12) and (11.7). ∞ ⊂ bs for η ≤ ηε, due to Bs (4) If X = Hm, then we getebs It should be remarked that assertion (iii) is basically known (see, for example, Proposition 0.2.1 in A. Lu- nardi [32], where the case m = 1 is considered). The proof is included here for further reference. d Little Besov-Holder spaces can be characterized by interpolation as well. For this we recall that, given Banach θ,∞ of exponent θ ∈ (0, 1) is the closure of X1 spaces X1 in (X0,X1)θ,∞. This defines an interpolation functor of exponent θ in the category of densely injected Banach couples, the continuous interpolation functor. It possesses the reiteration property (cf. [2, Section I.2] for more details and, in particular, G. Dore and A.Favini [13]). ֒→ X0, the continuous interpolation space (X0,X1)0 Theorem 11.4 (i) Suppose k0 < s < k1 with s /∈ N. Then (bck0 , bck1)0 . (ii) If 0 < s0 < s1 and 0 < θ < 1, then (bs0 = bsθ ∞ ∞, bs1 ∞)0 θ,∞ . = [bs0 ∞, bs1 ∞]θ. (s−k0)/(k1−k0),∞ . = bs ∞. P r o o f. (1) The validity of (i) and the first part of (ii) follow from Theorem 11.1(ii) and (iii) and Theo- rem 11.3(i). (2) We deduce from (11.2), (11.12), and Corollary 11.2 that BC∞ =Ts>0 Bs ∞. From this and (11.14) we infer Λs ∈ Laut(BC∞). Hence, using the definition of the little Besov-Holder spaces and once more (11.14) and Corollary 11.2, we find Λs ∈ Lis(bt+s ∞ , bt ∞), (Λs)−1 = Λ−s, t, t + s > 0. Thus the relevant arguments of part (2) of the proof of Theorem 11.1 apply literally to give the second part of (ii). This is due to the fact that the Fourier multiplier Theorem [4, Theorem 3.4.2] holds for bs ∞ also (see [3, Theorem 6.2]). Now we turn to anisotropic spaces. We set where BCkr/~r :=(cid:0)(cid:8) u ∈ C(X × J,X ) ; ∂α x ∂ju ∈ BC(X × J,X ), α + jr ≤ kr(cid:9), k·kkr/~r(cid:1), α+jr≤kr k∂α x ∂juk∞. kukkr/~r := max This space is complete and contains BUCkr/~r :=(cid:8) u ∈ BCkr/~r ; ∂α x ∂ju ∈ BUC(X × J,X ), α + jr ≤ kr(cid:9) as a closed linear subspace. Proposition 11.5 BUCkr/~r =Tk j=0 BUCj (J, BUC(k−j)r). 41 value theorem, claim is immediate for k = 0. x ∂ju(x, t) − h∂α ∂α x ∂ju(x, t + h) − ∂α for x ∈ X and t, h ∈ J. Thus, given ε > 0, the uniform continuity of ∂α such that P r o o f. (1) Due to u(x, t) − u(y, s) = u(x, t) − u(y, t) + u(y, t) − u(y, s) for (x, t), (y, s) ∈ X × J, the (2) Suppose k ∈ N× and u ∈ BUCkr/~r. Suppose also 0 ≤ j ≤ k − 1 and α ≤ (k − j)r. Then, by the mean x ∂j+1u(x, t)(cid:1) dτ x ∂j+1u(x, t) = hZ 1 0 (cid:0)∂α x ∂ju(·, t)(cid:1) − ∂α x ∂j+1u(·, t + τ h) − ∂α x ∂ju(·, t)(cid:1) : J → B(X,X ) is differentiable and its deriva- x ∂j+1u(·, t). From this and step (1) we infer u ∈ BUCj(J, BUC(k−j)r) for 0 ≤ j ≤ k. j=0 BUCj(J, BUC(k−j)r). The converse embedding is an obvious consequence x ∂ju(·, t + h) − ∂α ≤ max for h ∈ J\{0} with h ≤ δ. Hence the map(cid:0)t 7→ ∂α tive equals t 7→ ∂α This implies BUCkr/~r ֒→Tk of step (1). x ∂j+1u implies the existence of δ > 0 x ∂j+1u(·, t)(cid:13)(cid:13)∞;X x ∂j+1u(·, t)k∞ ≤ ε x ∂j+1u(x, t + τ h) − ∂α (cid:13)(cid:13)h−1(cid:0)∂α 0≤τ ≤1k∂α It is an immediate consequence of this lemma that BUCkr/~r ֒→ BUC(J, BUCkr) ∩ BUCk(J, BUC). It follows from Remark 1.13.4.2 in [50], for instance, that BUCkr/~r is a proper subspace of the intersection space on the right hand side. We infer from (11.1) that BC(k+1)r/~r ֒→ BUCkr/~r ֒→ BCkr/~r. Consequently, BC∞/~r :=TkBCkr/~r =TkBUCkr/~r = BC∞(X × J,X ). kuk∗ For s > 0 we set s/~r,∞ := sup s,∞ + sup t ku(·, t)k∗ = kuk∞ + sup x (cid:2)u(x,·)(cid:3)s/r,∞ t (cid:2)u(·, t)(cid:3)s,∞ + sup Suppose 0 < s ≤ r. Then x (cid:2)u(x,·)(cid:3)s/r,∞. x (cid:2)u(x,·)(cid:3)s/r,∞. kuk∗∗ s/~r,∞ := sup t ku(·, t)k∗∗ s,∞ + sup If kr < s ≤ (k + 1)r with k ∈ N×, then kuk∗∗ s/~r,∞ := max α+jr≤kr k∂α x ∂juk∗∗ (s−kr)/~r,∞. The anisotropic Besov-Holder space scale [ Bs/~r ∞ ; s > 0 ] is defined by ∞ :=( (BUCkr/~r, BUC(k+1)r/~r)(s−kr)/r,∞, kr < s < (k + 1)r, (BUCkr/~r, BUC(k+2)r/~r)1/2,∞, s = (k + 1)r. Bs/~r The next theorem is the anisotropic analogue of Theorem 11.1. Theorem 11.6 s/~r,∞ and k·k∗∗ (i) k·k∗ (ii) Suppose k0r < s < k1r. Then (BUCk0r/~r, BUCk1r/~r)(s−k0r)/(k1−k0)r,∞ s/~r,∞ are norms for Bs/~r ∞ . . = Bs/~r ∞ . (11.20) (11.21) (11.22) 42 H. Amann: Anisotropic Function Spaces on Singular Manifolds (iii) If 0 < s0 < s1 and 0 < θ < 1, then (Bs0/~r ∞ , Bs1/~r ∞ )θ,∞ . = Bsθ /~r ∞ . = [Bs0/~r ∞ , Bs1/~r ∞ ]θ. (iv) ∂α x ∂j ∈ L(B(s+α+jr)/~r ∞ , Bs/~r ∞ ) for α ∈ Nm and j ∈ N. P r o o f. (1) We infer from [4, (3.3.12), (3.5.2), and Theorem 4.4.1] that Bs/~r ∞ = Bs/ν ∞,∞ (11.23) and that (ii) is true. (2) The first part of (iii) follows from (ii) by reiteration. [4, Theorem 3.4.1] and (11.23) we get (3) For (ξ, τ ) ∈ Rm × R we seteΛ(ξ, τ ) := (1 + ξ2r + τ 2)1/2r. TheneΛs := F −1 m+1eΛsFm+1 for s ∈ R. From eΛs ∈ Lis(B(t+s)/~r ∞ , Bt/~r ∞ ), (eΛs)−1 = eΛ−s, t, t + s > 0, provided X = Rm and J = R. Now we obtain the second part of (iii) by obvious modifications of the relevant sections of part (2) of the proof of Theorem 11.1. (4) Taking [4, Section 4.4] into account, we get from Theorems 3.3.2 and 3.5.2 therein that Bs/~r ∞ ֒→ BUC. Suppose kr < s ≤ (k + 1)r. By [4, Theorem 3.6.1] kukBs/~r ∞ ∼ kuk∞ + sup h∈(0,∞)m k△[s]+1 (h,0) uk∞ hs k△[s/r]+1 (0,h) uk∞ hs/r , + sup h>0 where [t] is the largest integer less than or equal to t ∈ R. Since u ∈ BC it follows (h,0) uk∞ = sup t k△[s]+1 h u(·, t)k∞, k△[s/r]+1 (0,h) uk∞ = sup x k△[s/r]+1 h u(x,·)k∞. k△[s]+1 ∞ ∼ k·k∗ s/~r,∞. Thus k·kBs/~r (5) Suppose X = Rm and J = R. Then (iv) follows by straightforward modifications of the proof of [4, Lemma 2.3.7] by invoking the Fourier multiplier Theorem 3.4.2 therein. Similarly as in the proof of [4, Theo- rem 2.3.8], we see that, given 0 < s ≤ r and k ∈ N, k·k(s+kr)/~r B∞ ∼ max α+jr≤kr k∂α x ∂j · kBs/~r ∞ (11.24) (cf. [4, Corollary 2.3.4]). In the general case we now obtain the validity of (iv) and (11.24) by extension and restriction, taking Bs/~r ∞ ֒→ BUC into account. (6) Suppose 0 < s ≤ r. Then k·k∗ s/~r,∞ ∼ k·k∗∗ (11.24) we see that the latter equivalence holds for every k ∈ N. This proves the theorem. s/~r,∞ follows from Theorem 11.1(i). By combining this with Corollary 11.7 . = B(J, Bs ∞) ∩ Bs/r ∞ (J, B). (i) Bs/~r ∞ (ii) Set kuk∼ s/~r,∞ := sup t ku(·, t)k∗∗ s,∞ + sup x ku(x,·)k∗∗ s/r, s > 0. s/~r,∞ is a norm for Bs ∞. Then k·k∼ P r o o f. (i) is implied by Theorem 11.6(i), Bs lows from (i) and Theorem 11.1(i). ∞ ֒→ BUCk if k < s ≤ k + 1, and Proposition 11.5. (ii) fol- 43 We define anisotropic Holder spaces by BCs/~r := Bs/~r s/~r,∞, for example, we find, similarly as in the isotropic case, that ∞ for s ∈ R+\rN. By means of the mean value theorem and using the norm k·k∼ BCs/~r ֒→ BCs0/~r, 0 ≤ s0 < s. In order to obtain scales of spaces enjoying dense embeddings we define anisotropic little Holder spaces by Similarly, the anisotropic little Besov-Holder space bcs/~r is the closure of BC∞/~r in BCs/~r, s ∈ R+. (11.25) ∞ is the closure of BC∞/~r in Bs/~r bs/~r ∞ , s > 0. These spaces possess intrinsic characterizations as well. To allow for a simple formulation we denote by [s]− the largest integer strictly less than s. Theorem 11.8 (i) bckr/~r = BUCkr/~r. (ii) bcs/~r = bs/~r (iii) u ∈ bs/~r ∞ if s ∈ R+\N. ∞ and ∞ iff u ∈ Bs/~r as δ → 0. (iv) BCs/~r d ֒→ bs0/~r ∞ for 0 < s0 < s. sup t max α=[s]−(cid:2)∂α x u(·, t)(cid:3)δ s−[s]−,∞ + sup x (cid:2)∂[s/r]−u(x,·)(cid:3)δ s/r−[s/r]−,∞ → 0 (11.26) P r o o f. As in previous proofs it suffices to consider the case X = Rm and J = R. (1) We know from (11.20) that BC∞/~r ֒→ BUCkr/~r. Let { wη ; η > 0 } be a mollifier on Rm+1. If u belongs x ∂ju) that wη ∗ u → u in BCkr/~r as to BUCkr/~r, then it follows from (11.4) and ∂α η → 0. This proves assertion (i). Claim (ii) is trivial. x ∂j(wη ∗ u) = wη ∗ (∂α (2) Let kr ≤ i < s ≤ i + 1 ≤ (k + 1)r with i ∈ N. Suppose u ∈ bs/~r ∞ and ε > 0. Then we can find v belong- ing to ∈ BUC(k+2)r/~r ֒→ Bs/~r ∞ such that ku − vk∗∗ s/~r,∞ < ε/2. By Proposition 11.5 we know Hence it follows from (11.9) that α = i. sup s−i,∞ ≤ cδ kvk(k+2)r/~r,∞, BUC(k+2)r/~r ֒→ BUC(J, BUC(k+2)r) ∩ BUCk+2(J, BUC). x v(·, t)(cid:3)δ t (cid:2)∂α x (cid:2)∂kv(x,·)(cid:3)δ x v(·, t)(cid:3)δ s/r−k,∞ ≤ cδ kvk(k+2)r/~r,∞, 0 < δ ≤ 1, 0 < δ ≤ 1. s−i,∞ + sup sup s/r−k,∞ < ε/2, x (cid:2)∂kv(x,·)(cid:3)δ x (u − v)(·, t)(cid:3)s−i,∞ + sup t max α=i(cid:2)∂α max α=i(cid:2)∂α x u(·, t)(cid:3)δ s−i,∞ ≤ sup t 0 < δ ≤ δε. max α=i(cid:2)∂α x v(·, t)(cid:3)δ s−i,∞ ≤ ε Similarly, Thus we find δε > 0 such that sup t Consequently, sup t max α=i(cid:2)∂α for 0 < δ ≤ δε. This shows that the first term in (11.26) converges to zero. Analogously, we see that this is true for the second summand. 44 H. Amann: Anisotropic Function Spaces on Singular Manifolds (3) Suppose 0 < s ≤ 1 and u ∈ Bs/~r ∞ satisfies (11.26). By (11.3) it suffices to show that It follows from △[s]+1 (h,0) (wη ∗ u) = wη ∗ (△[s]+1 kwη ∗ u − uk∼ (h,0) u) that s/~r,∞ → 0 as η → 0. (11.27) k△[s]+1 (h,0) (wη ∗ u)(x, t)k ≤ sup u(·, t)k∞, h t k△[s]+1 t (cid:2)u(·, t)(cid:3)δ s,∞, (x, t) ∈ X × J. 0 < δ < ∞. Consequently, sup t (cid:2)wη ∗ u(·, t)(cid:3)δ Let ε > 0 and fix δε > 0 with supt(cid:2)u(·, t)(cid:3)δε t (cid:2)(wη ∗ u − u)(·, t)(cid:3)δε s,∞ ≤ sup s,∞ < ε/4. Then sup Thus we infer from (11.6) that s,∞ ≤ 2 sup t (cid:2)u(·, t)(cid:3)δε s,∞ < ε/2. sup t (cid:2)(wη ∗ u − u)(·, t)(cid:3)s,∞ ≤ ε/2 + 4δ−s ε Since u ∈ BUC(X × J,X ) it follows from (11.4) that t k(wη ∗ u − u)(·, t)k∞. sup t k(wη ∗ u − u)(t)k∞ = kwη ∗ u − ukB(X×J,X ) → 0 as η → 0. sup Hence sup Similarly, t (cid:2)(wη ∗ u − u)(·, t)(cid:3)s,∞ → 0 x (cid:2)(wη ∗ u − u)(x,·)(cid:3)s/r,∞ → 0 This proves (11.27), thus, due to step (2), assertion (iii) for 0 < s ≤ 1. sup as η → 0. as η → 0. (4) To prove (iv) assume kr ≤ i < s ≤ i + 1 ≤ (k + 1)r and u ∈ Bs/~r x ∂j(wη ∗ u) = wη ∗ (∂α ∞ , which shows that claim (iii) is always true. x ∂ju) for α + jr < s and step (3) that wη ∗ u → u in Bs/~r ∞ satisfies (11.26). Then it follows ∞ as η → 0. Hence from ∂α u ∈ bs/~r (5) The proof of (iv) is obtained by employing (11.7), (11.9), and Corollary 11.7(ii). Anisotropic little Holder spaces can be characterized by interpolation, similarly as their isotropic relatives. Theorem 11.9 (i) bs/~r ∞ . = (bck0r/~r, bck1r/~r)0 (ii) If 0 < s0 < s1 and 0 < θ < 1, then (bs0/~r (s/r−k0)/(k1−k0),∞ for k0r < s < k1r. . = [bs0/~r . = bsθ /~r ∞ , bs1/~r ∞ )0 ∞ θ,∞ ∞ , bs1/~r ∞ ]θ. (iii) ∂α x ∂j ∈ L(b(s+α+jr)/~r ∞ , bs/~r ∞ ) for α ∈ Nm and j ∈ N. P r o o f. (1) The first assertion as well as the first part of (ii) follow from part (i) of Theorem 11.8. Part two of (ii) and the first claim are implied by part (iii) of Theorem 11.6. (2) The last part of statement (ii) is obtained by replacing BC∞ and Λs in step (2) of the proof of Theorem 11.4 s by BC∞/~r andeΛ theorem, we obtain (iii). (3) Theorem 11.6(iv) implies ∂α , respectively. x ∂j ∈ L(BC∞/~r). Thus, using the definition of bs/~r ∞ and once more the latter 45 this we remind the reader of the notations and conventions introduced at the beginning of Section 7. In the next section we need to employ Holder spaces with a particular choice of X which we discuss now. For Let { Fβ ; β ∈ B} be a countable family of Banach spaces. Then it is obvious that f : F X →Yβ F X β , u 7→ f u := (prβ ◦ u) (11.28) is a linear bijection. Since F carries the product topology u ∈ F X is continuously differentiable iff uβ := prβ ◦ u ∈ C1(X, Fβ), β ∈ B. Then ∂ju = (∂juβ), that is, Setting Ck(X, F) :=Qβ Ck(X, Fβ) etc., it follows x = ∂α x ◦ f , f ◦ ∂α α ∈ Nm. f ∈ Lis(cid:0)Ck(X, F ), C k(X, F)(cid:1). Furthermore, f ∈ L(cid:0)BCk(cid:0)X, ℓ∞(F )(cid:1), ℓ∞(cid:0)BCk(X, F)(cid:1)(cid:1). sup Suppose u ∈ BC1(cid:0)X, ℓ∞(F )(cid:1). Then, given x ∈ X, β∈B(cid:13)(cid:13)t−1(cid:0)uβ(x + tej) − uβ(x)(cid:1) − ∂juβ(x)(cid:13)(cid:13)Fβ as t → 0, with t > 0 if X = Hm and j = 1. From this we see that f maps u ∈ BCk(cid:0)X, ℓ∞(F )(cid:1) into the linear subspace of ℓ∞(cid:0)BC k(X, F)(cid:1) consisting of all v = (vβ) for which vβ is k-times continuously differentiable, uniformly with respect to β ∈ B. Thus (11.31) is not surjective if k ≥ 1. =(cid:13)(cid:13)t−1(cid:0)u(x + tej) − u(x)(cid:1) − ∂ju(x)(cid:13)(cid:13)ℓ∞(F ) → 0 We denote by the linear subspace of ℓ∞(cid:0)BC k(X, F)(cid:1) of all v = (vβ) such that ∂αvβ is uniformly continuous on X for α ≤ k, uniformly with respect to β ∈ B. Lemma 11.10 f is an isomorphism ℓ∞,unif(cid:0)bck(X, F)(cid:1) (11.29) (11.30) (11.31) (11.32) (11.33) and from bck(cid:0)X, ℓ∞(F )(cid:1) onto ℓ∞,unif(cid:0)bck(X, F)(cid:1) ∞(cid:0)X, ℓ∞(F )(cid:1) onto ℓ∞(cid:0)Bs ∞(X, F)(cid:1), from Bs s > 0. P r o o f. (1) Suppose u ∈ bck(cid:0)X, ℓ∞(F )(cid:1). Then, by the above, it is obvious that f u ∈ ℓ∞,unif(cid:0)bck(X, F)(cid:1). Conversely, assume u = (uβ) ∈ ℓ∞,unif(cid:0)bck(X, F)(cid:1). Set u := f −1u, which is defined due to (11.30). Then ku(x)kℓ∞(F ) = sup β kuβ(x)kFβ , x ∈ X, and ku(x) − u(y)kℓ∞(F ) = sup β kuβ(x) − uβ(y)kFβ , x, y ∈ X, show u ∈ bc(cid:0)X, ℓ∞(F )(cid:1). Hence we infer from (11.29) that ∂α (2) Let k ≥ 1 and 1 ≤ j ≤ m. Then, by the mean value theorem, t−1(cid:0)uβ(x + tej) − uβ(x)(cid:1) − ∂juβ(x) =Z 1 x u ∈ bc(cid:0)X, ℓ∞(F )(cid:1) for α ≤ k. 0 (cid:0)∂juβ(x + stej) − ∂juβ(x)(cid:1) ds, 46 H. Amann: Anisotropic Function Spaces on Singular Manifolds where t > 0 if j = 1 and X = Hm. Hence (cid:13)(cid:13)t−1(cid:0)uβ(x + tej) − uβ(x)(cid:1) − ∂juβ(x)(cid:13)(cid:13)Fβ ≤ sup t≤δ x∈Xk∂juβ(x + tej) − ∂juβ(x)kFβ sup t≤δ k∂ju(· + tej) − ∂jukℓ∞(BC(X,F)) ≤ sup for t ≤ δ, x ∈ X, and β ∈ B. Thus (cid:13)(cid:13)t−1(cid:0)u(· + tej) − u(cid:1) − ∂ju(cid:13)(cid:13)B(X,ℓ∞(F )) ≤ sup t≤δ k∂ju(· + tej) − ∂jukℓ∞(BC(X,F)) for t ≤ δ. This implies that u is differentiable in the topology of BC(cid:0)X, ℓ∞(F )(cid:1). From this, step (1), and by induction we infer f −1 ∈ L(cid:0)ℓ∞,unif(cid:0)bck(X, F)(cid:1), bck(cid:0)X, ℓ∞(F )(cid:1)(cid:1). This proves (11.32). to Bs (3) Suppose 0 < s ≤ 1 and set i := [s]−. It is convenient to write h ≫ 0 iff h ∈ (0,∞)m. Given u belonging ∞(cid:0)X, ℓ∞(F )(cid:1), we deduce from △huβ = prβ(△hu) that β (cid:2)prβ(f u)(cid:3)s,∞;Fβ [uβ]s,∞;Fβ = sup β h uβ(x)kFβ k△i+1 sup h≫0 hs = sup sup sup β x k△i+1 h u(x)kℓ∞(F ) k△i+1 h uk∞;ℓ∞(F ) (11.34) = sup h≫0 sup x hs = sup h≫0 hs From (11.31) and (11.34) we infer = [u]s,∞;ℓ∞(F ). ℓ∞(cid:0)Bs f ∈ L(cid:0)Bs ∞(X, F)(cid:1)(cid:1). Now it follows from (11.29) that (11.35) holds for any s > 0. ∞(cid:0)X, ℓ∞(F )(cid:1), ℓ∞(cid:0)Bs It is obvious from (11.11) and (11.29) that, given k < s ≤ k + 1, ∞(X, F)(cid:1) ֒→ ℓ∞,unif(cid:0)bck(X, F)(cid:1). ∞(X, F)(cid:1). Due to (11.30) this proves (11.33). From this, (11.34), and (11.32) we get that f is onto ℓ∞(cid:0)Bs We denote for k < s ≤ k + 1 by ℓ∞,unif(cid:0)bs ∞(X, F)(cid:1) the linear subspace of ℓ∞,unif(cid:0)bck(X, F)(cid:1) of all v = (vβ) such that limδ→0 maxα=k[∂α formly with respect to β ∈ B. ∞(cid:0)X, ℓ∞(F )(cid:1), ℓ∞,unif(cid:0)bs ∞(X, F)(cid:1)(cid:1). Lemma 11.11 f ∈ Lis(cid:0)bs P r o o f. The proof of (11.34) shows that, given k < s ≤ k + 1, β (cid:2)prβ(f u)(cid:3)δ s,∞;ℓ∞(F ), = [u]δ x vβ]δ s−k,∞;Fβ δ > 0. s,∞;Fβ sup Thus the claim follows by the arguments of step (2) of the proof of Lemma 11.10 and from Theorem 11.3. (11.35) = 0, uni- Now we extend f point-wise over J: ef : F X×J →Yβ F X×J β , u 7→ef u :=(cid:0)t 7→ f u(·, t)(cid:1). ∞ (X × J, Fβ) for s > 0. Analogous definitions apply to bs/~r ∞ (X × J, F). As above, Bs/~r ∞ (X × J, F) :=Qβ Bs/~r Clearly, is the closed subspace of ℓ∞(cid:0)BC kr/~r(X × J, F)(cid:1) of all u = (uβ) for which ∂α α + jr ≤ kr, uniformly with respect to β ∈ B. Suppose kr < s ≤ (k + 1)r. We denote by ℓ∞,unif(cid:0)bckr/~r(X × J, F)(cid:1) x ∂juβ ∈ BUC(X × J, Fβ) for 47 ∞ (X × J, F)(cid:1) ℓ∞,unif(cid:0)bs/~r ∞ (X × J, F)(cid:1) satisfying x uβ(·, t)(cid:3)δ s−[s]−,∞;Fβ β + sup sup x (cid:2)∂[s/r]−uβ(x,·)(cid:3)δ s/r−[s/r]−,∞;Fβ → 0 Now we can prove the following anisotropic analogue of Lemmas 11.10 and 11.11. the set of all u = (uβ) ∈ ℓ∞(cid:0)Bs/~r α=[s]−(cid:2)∂α max sup sup β t as δ → 0. Lemma 11.12 ef is an isomorphism and as well as from Bs/~r from bckr/~r(cid:0)X × J, ℓ∞(F )(cid:1) onto ℓ∞,unif(cid:0)bckr/~r(X × J, F)(cid:1) ∞ (X × J, F)(cid:1) ∞ (X × J, F)(cid:1). ∞ (cid:0)X × J, ℓ∞(F )(cid:1) onto ℓ∞(cid:0)Bs/~r ∞ (cid:0)X × J, ℓ∞(F )(cid:1) onto ℓ∞,unif(cid:0)bs/~r from bs/~r P r o o f. Note ∂j ◦ef =ef ◦ ∂j. Hence the first assertion follows from (11.32). The remaining statements are verified by obvious modifications of the relevant parts of the proofs of Lemmas 11.10 and 11.11, taking Corollary 11.7(ii) and Theorem 11.8 into account. 12 Weighted Holder Spaces Having investigated Holder spaces on Rm and Hm in the preceding section we now return to the setting of singular manifolds. First we introduce isotropic weighted Holder spaces and study some of their properties. Afterwards we study to anisotropic Holder spaces of time-dependent W -valued (σ, τ )-tensor fields on M . Making use of the results of Section 11 we can give coordinate-free invariant definitions of these spaces. By B0,λ = B0,λ(V ) we mean the weighted Banach space of all sections u of V satisfying endowed with the norm k·k∞;λ, and B := B0,0. kuk∞;λ = kuk0,∞;λ :=(cid:13)(cid:13)ρλ+τ −σ uh(cid:13)(cid:13)∞ < ∞, For k ∈ N where BCk,λ = BCk,λ(V ) :=(cid:0)(cid:8) u ∈ Ck(M, V ) ; kukk,∞;λ < ∞(cid:9), k·kk,∞;λ(cid:1), kukk,∞;λ := max 0≤i≤k(cid:13)(cid:13)ρλ+τ −σ+i ∇iuh(cid:13)(cid:13)∞. The topologies of B0,λ and BCk,λ are independent of the particular choice of ρ ∈ T(M ). Consequently, this is also true for all other spaces of this section as follows from their definition which involves the topology of BCk,λ for k ∈ N only. It is a consequence of Theorem 12.1 below that BCk,λ is a Banach space. We set BC∞,λ = BC∞,λ(V ) :=TkBCk,λ, 48 H. Amann: Anisotropic Function Spaces on Singular Manifolds endowed with the obvious projective topology. Then bck,λ = bck,λ(V ) is the closure of BC∞,λ in BCk,λ, The weighted Besov-Holder space scale [ Bs,λ ∞ ; s > 0 ] is defined by k ∈ N. Bs,λ ∞ = Bs,λ ∞ (V ) :=( (bck,λ, bck+1,λ)s−k,∞, k < s < k + 1, (bck,λ, bck+2,λ)1/2,∞, s = k + 1. (12.1) It is a scale of Banach spaces. The following fundamental retraction theorem allows to characterize Besov-Holder spaces locally. Theorem 12.1 Suppose k ∈ N and s > 0. Then ψλ ∞ is a retraction from ℓ∞(BC k) onto BCk,λ and from ℓ∞(Bs ∞) onto Bs,λ ∞ , and ϕλ ∞ is a coretraction. P r o o f. (1) The first claim is settled by Theorem 6.3 of [5]. (2) Suppose k ∈ N. It is obvious by the definition of bck, step (1), (4.1), and (7.3) that ∞ is a retraction from ℓ∞,unif(bck) onto bck,λ, and ϕλ ψλ ∞ is a coretraction. (12.2) (3) If ∂M = ∅, then we put M := Rm, B := K, and Fκ := Eκ := E for κ ∈ K. Then, defining f by (11.28) with this choice of Fβ and X := Rm, Lemma 11.10 implies f ∈ Lis(cid:0)bck(cid:0)M, ℓ∞(E)(cid:1), ℓ∞,unif(bck)(cid:1). (12.3) infer from Lemma 11.10 (4) Suppose ∂M 6= ∅. Then we set K0 := K\K∂M and K1 := K∂M . With Eκ := E for κ ∈ K we put Ei :=Qκ∈Ki Eκ and define fi by setting B = Ki and Fκ = Eκ. Then, letting X0 := Rm and X1 := Hm, we with bck(Xi, Ei) :=Qκ∈Ki For bck =Qκ∈K bck fi ∈ Lis(cid:0)bck(cid:0)Xi, ℓ∞(Ei)(cid:1), ℓ∞,unif(cid:0)bck(Xi, Ei)(cid:1)(cid:1), logical direct sum decomposition bck(Xi, Eκ). (12.4) κ we use the natural identification bck = bck(X0, E0) ⊕ bck(X1, E1). It induces a topo- ℓ∞,unif(bck) = ℓ∞,unif(cid:0)bck(X0, E0)(cid:1) ⊕ ℓ∞,unif(cid:0)bck(X1, E1)(cid:1), bck(cid:0)M, ℓ∞(E)(cid:1) := bck(cid:0)X0, ℓ∞(E0)(cid:1) ⊕ bck(cid:0)X1, ℓ∞(E1)(cid:1). f := f0 ◦ pr0 + f1 ◦ pr1 ∈ Lis(cid:0)bck(cid:0)M, ℓ∞(E)(cid:1), ℓ∞,unif(bck)(cid:1). (12.6) (12.5) Denoting by ⊔ the disjoint union, we set M := Rm ⊔ Hm and where on the right side we use the maximum of the norms of the two summands. (5) Returning to the general case, where ∂M may or may not be empty, we set It follows from (12.4) and (12.5) that We deduce from (12.2), (12.3), and (12.6) that ∞ := f −1 ◦ ϕλ Φλ ∞, Ψλ ∞ := ψλ ∞ ◦ f . Ψλ ∞ is a retraction from bck(cid:0)M, ℓ∞(E)(cid:1) onto bck,λ, and Φλ As a consequence of this, Theorem 11.1(ii), definition (12.1), and general properties of interpolation functors (cf. [2], Proposition I.2.3.3) we find ∞ is a coretraction. (12.7) Ψλ ∞ is a retraction from Bs ∞ , and Φλ ∞ is a coretraction. ∞(cid:0)M, ℓ∞(E)(cid:1) onto Bs,λ ∞ ◦ f −1, ϕλ ∞ = f ◦ Φλ ∞ ψλ ∞ = Ψλ Since we get the second assertion from (12.8) and Lemma 11.10. (12.8) (12.9) Corollary 12.2 49 (i) u 7→ uk,∞;λ := supκ⋉ϕ ρλ (ii) Suppose s > 0. Then κ k(κ⋉ϕ)∗(πκu)kk,∞;E is a norm for BCk,λ. u 7→ u∗ s,∞;λ := sup κ⋉ϕ κ k(κ⋉ϕ)∗(πκu)k∗ ρλ s,∞;E and are norms for Bs,λ ∞ . u 7→ u∗∗ s,∞;λ := sup κ⋉ϕ κ k(κ⋉ϕ)∗(πκu)k∗∗ ρλ s,∞;E (iii) Assume k0 < s < k1 with k0, k1 ∈ N. Then (bck0,λ, bck1,λ)(s−k0)/(k1−k0),∞ (iv) If 0 < s0 < s1 and 0 < θ < 1, then (Bs0,λ . = [Bs0,λ . = Bsθ ,λ ∞ , Bs1,λ ∞ )θ,∞ ∞ ∞ , Bs1,λ ∞ ]θ. . = Bs,λ ∞ . P r o o f. (i) and (ii) are implied by (7.9) and Theorem 11.1(i). Assertions (iii) and (iv) follow from (12.7) and (12.8) and parts (ii) and (iii) of Theorem 11.1, respectively, and (12.9) and Lemma 11.10. rem 11.1(ii). Weighted Holder spaces are defined by BCs,λ := Bs,λ ∞ for s ∈ R+\N. This is in agreement with Theo- Parts (i) and (ii) of Corollary 12.2 show that the present definition of weighted Holder spaces is equivalent to the one used in [5]. It should be noted that Corollary 12.2(iii) gives a positive answer to the conjecture of Remark 8.2 of [5], provided BCk,λ and BCk+1,λ are replaced by bck,λ and bck+1,λ, respectively. We define weighted little Holder spaces by Similarly, the weighted little Besov-Holder space scale [ bs,λ ∞ ; s > 0 ] is obtained by bcs,λ is the closure of BC∞,λ in BCs,λ, s ≥ 0. ∞ is the closure of BC∞,λ in Bs,λ bs,λ ∞ . (12.10) Theorem 12.3 ψλ ∞ is a retraction from ℓ∞,unif(bs ∞) onto bs,λ ∞ , and ϕλ ∞ is a coretraction. Ψλ (11.15) and (12.10) we deduce from (12.7) and (12.8) that P r o o f. We infer from (11.20) that BC∞(cid:0)M, ℓ∞(E)(cid:1) =Tk bck(cid:0)M, ℓ∞(E)(cid:1). Hence we get from (12.7) that ∞ is a retraction from BC∞(cid:0)M, ℓ∞(E)(cid:1) onto BC∞,λ, and Φλ ∞(cid:0)M, ℓ∞(E)(cid:1) onto bs,λ ∞ is a coretraction. Due to this and definitions Now the assertion follows from Lemma 11.11 and (12.9). ∞ is a retraction from bs ∞ is a coretraction. ∞ , and Φλ (12.11) Ψλ Corollary 12.4 (i) Suppose k0 < s < k1 with k0, k1 ∈ N. Then (bck0,λ, bck1,λ)0 . (ii) If 0 < s0 < s1 and 0 < θ < 1, then (bs0,λ = bsθ ,λ ∞ , bs1,λ ∞ )0 ∞ θ,∞ (s−k0)/(k1−k0),∞ . = [bs0,λ ∞ , bs1,λ ∞ ]θ. . = bs,λ ∞ . P r o o f. These predications are derived from (12.11) and Theorem 11.4. Now we turn to weighted anisotropic spaces. We set and, for k ∈ N×, BC0/~r,~ω = BC0/~r,(λ,0) :=(cid:0)(cid:8) u ∈ C(cid:0)J, C(V )(cid:1) ; kuk∞;B0,λ < ∞(cid:9), k·k∞;B0,λ(cid:1) BCkr/~r,~ω :=(cid:8) u ∈ C(cid:0)J, C(V )(cid:1) ; ∇i∂ju ∈ BC0/~r,(λ+i+jµ,0), i + jr ≤ kr(cid:9), (12.12) (12.13) 50 H. Amann: Anisotropic Function Spaces on Singular Manifolds endowed with the norm u 7→ kukkr/~r,∞;~ω := max i+jr≤kr k∇i∂juk∞;λ+i+jµ. (12.14) It is a consequence of Theorem 12.6 below that BCkr/~r,~ω is a Banach space. Similarly as in the isotropic case, anisotropic Holder spaces can be characterized by means of local coordi- nates. For this we prepare the following analogue of Lemma 7.2. Lemma 12.5 Suppose k ∈ N and s > 0. Then κ) ∩ L(bck Seκκ ∈ L(BCk eκ , BCk eκ, bck κ) ∩ L(Bs ∞,eκ, Bs ∞,κ) ∩ L(bs ∞,eκ, bs ∞,κ) and kSeκκk ≤ c foreκ ∈ N(κ) and κ ∈ K. P r o o f. As in the proof of Lemma 7.2 we see that the statement applies for the spaces BCk and bck. Now we get the remaining assertions by interpolation, due to Theorems 11.1(ii) and 11.4(i). Theorem 12.6 ψ~ω ∞ is a retraction from ℓ∞(BCkr/~r) onto BCkr/~r,~ω, and ϕ~ω ∞ is a coretraction. P r o o f. (1) From (9.6) and Theorem 12.1 we get ∞,κuk∞;BC kr Similarly, by invoking (9.5) as well, kϕ~ω κ = kϕλ ∞,κuk∞;BC kr κ ≤ ckuk∞;BC kr,λ, κ⋉ϕ ∈ K⋉Φ. k∂jϕ~ω ∞,κuk∞;BC (k−j)r κ = kϕλ+jr ∞,κ ∂juk∞;BC (k−j)r κ ≤ ck∂juk∞;BC (k−j)r,λ+jr for 0 ≤ j ≤ k and κ⋉ϕ ∈ K⋉Φ. From this and definition (12.14) we infer ∞ukℓ∞(BC kr/~r) ≤ ckukkr/~r,∞;~ω. kϕ~ω (2) Given κ ∈ K andeκ ∈ N(κ), with ϕλ ∞,κ ◦ ψλ ∞,eκ = aeκκSeκκ (12.15) It is obvious that the scalar-valued BCk-spaces form continuous multiplication algebras. Hence (4.3), (7.3), and Lemma 12.5 imply aeκκ := (ρκ/ρeκ)λ(κ∗πκ)Seκκ(eκ∗πeκ). eκ ∈ N(κ), Thus we deduce from (12.15), (12.16), and Lemma 12.5 that ∞,κ ◦ ψλ kaeκκkBC kr ≤ c, ∞,eκveκk∞;BC kr ∞,κ ◦ ψ~ω = kϕλ kϕλ κ ∞,eκveκk∞;BC kr κ ≤ ckveκk∞;BC kr eκ κ ∈ K. (12.16) foreκ ∈ N(κ) and κ⋉ϕ,eκ⋉eϕ ∈ K⋉Φ. By this and the finite multiplicity of K we obtain ∞vk∞;BC kr ∞,κ ◦ ψ~ω kϕλ κ =(cid:13)(cid:13)(cid:13) Xeκ∈N(κ) ≤ c max eκ∈N(κ)kϕλ ∞,κ ◦ ψ~ω ϕλ ∞,κ ◦ ψ~ω ∞,eκveκ(cid:13)(cid:13)(cid:13)∞;BC kr ∞,eκveκk∞;BC kr κ κ ≤ ckvkℓ∞(B(J,BCkr)) for κ⋉ϕ ∈ K⋉Φ. Note kϕλ+jµ ∞,κ ◦ ∂j ◦ ψ~ω ∞,eκveκk∞;BC (k−j)r κ = kϕλ+jµ ∞,κ ◦ ψλ+jµ ∞,eκ (∂jveκ)k∞;BC (k−j)r κ for 0 ≤ j ≤ k and κ⋉ϕ,eκ⋉eϕ ∈ K⋉Φ. Thus, as above, ∞ ◦ ∂j ◦ ψ~ω kϕλ+jµ ∞vkℓ∞(BC(k−j)r ) ≤ ck∂jvkℓ∞(B(J,BC (k−j)r )), Now we deduce from Corollary 12.2(i) 0 ≤ j ≤ k. Since ϕ~ω ∞ is a right inverse for ψ~ω ∞ the theorem is proved. kψ~ω ∞vkkr/~r,∞;~ω ≤ ckvkℓ∞(BCkr/~r). 51 Next we introduce a linear subspace of BCkr/~r,~ω by bckr/~r,~ω is the set of all u in BCkr/~r,~ω with ϕ~ω ∞u ∈ ℓ∞,unif(bckr/~r). Due to the fact that ℓ∞,unif(bckr/~r) is a closed linear subspace of ℓ∞(BCkr/~r) it follows from the continuity of ϕ~ω ∞ that bckr/~r,~ω is a closed linear subspace of BCkr/~r,~ω. The next theorem shows, in particular, that bckr/~r,~ω is independent of the particular choice of K⋉Φ and the localization system used in the preceding definition. For this we set equipped with the natural projective topology. BC∞/~r,~ω :=TkBCkr/~r,~ω, Theorem 12.7 (i) ψ~ω ∞ is a retraction from ℓ∞,unif(bckr/~r) onto bckr/~r,~ω, and ϕ~ω ∞ is a coretraction. (ii) bckr/~r,~ω is the closure of BC∞/~r,~ω in BCkr/~r,~ω. P r o o f. (1) Suppose ϕ~ω κ⋉ϕ ∈ K⋉Φ. Hence πκu = 0 for κ ∈ K, and consequently π2 Thus ϕ~ω ∞ is injective. ∞u = 0 for some u ∈ BCkr/~r,~ω. Then it follows from (9.3) that (κ⋉ϕ)∗(πκu) = 0 for κu = 0. κu = 0 for κ ∈ K. This implies u =Pκ π2 ∞. Theorem 12.6 and [4, Lemma 4.1.5] imply (2) We denote by Y the image space of BCkr/~r,~ω under ϕ~ω ψ~ω ∞ ∈ Lis(Y, BCkr/~r,~ω). ℓ∞(BC kr/~r) = Y ⊕ ker(ψ~ω ∞), (12.17) Thus, by step (1) (see Remarks 2.2.1 of [4]), ϕ~ω ∞ ∈ Lis(BCkr/~r,~ω,Y), ∞)−1 = ψ~ω Since X := Y ∩ ℓ∞,unif(bckr/~r) is a closed linear subspace of Y we thus get ∞bckr/~r,~ω)−1 = ψ~ω ∞Y. (ϕ~ω Due to (12.17) we can write w ∈ ℓ∞,unif(bckr/~r) in the form w = u + v with u ∈ X and v ∈ ker(ψ~ω this and (12.18) it follows ψ~ω ∞ ◦ ϕ~ω ψ~ω (3) Using obvious adaptions of the notations of the proof of Theorem 12.1 we deduce from Lemma 11.12 ∞ ∈ L(cid:0)ℓ∞,unif(bckr/~r), bckr/~r,~ω(cid:1) and ∞u = u for u ∈ bckr/~r,~ω. This proves (i). ∞). From ∞X . (ϕ~ω ∞ ∈ Lis(bckr/~r,~ω,X ), ϕ~ω ∞(cid:0)ℓ∞,unif(bckr/~r)(cid:1) ⊂ bckr/~r,~ω. Hence ψ~ω ef ∈ Lis(cid:0)bckr/~r(cid:0)M × J, ℓ∞(E)(cid:1), ℓ∞,unif(bckr/~r)(cid:1). We set (12.18) (12.19) (12.20) Then we infer from (i) and (12.19) that Ψ~ω Φ~ω ∞, Ψ~ω ∞ := ψ~ω ∞ :=ef −1 ◦ ϕ~ω ∞ ◦ef . ∞ is a retraction from bckr/~r(cid:0)M × J, ℓ∞(E)(cid:1) onto bckr/~r,~ω, and Φ~ω ֒→ bckr/~r(cid:0)M × J, ℓ∞(E)(cid:1). BC∞(cid:0)M × J, ℓ∞(E)(cid:1) d Definition (11.25) guarantees ∞ is a coretraction. It is an easy consequence of the mean value theorem that ℓ∞(BC (k+1)r/~r) ֒→ ℓ∞,unif(bckr/~r). From these embeddings, Theorem 12.6, and (i) we infer that the first of the injections BC(k+1)r/~r,~ω ֒→ bckr/~r,~ω ֒→ BCkr/~r,~ω H. Amann: Anisotropic Function Spaces on Singular Manifolds 52 is valid. Thus BC∞/~r,~ω =TkBCkr/~r,~ω =Tkbckr/~r,~ω. Now it follows from (11.20) and (12.20) that Assertion (ii) is implied by (12.20) and [4, Lemma 4.1.6]. Ψ~ω ∞ is a retraction from BC∞(cid:0)M × J, ℓ∞(E)(cid:1) onto BC∞/~r,~ω. We define the weighted anisotropic Besov-Holder space scale [ Bs/~r,~ω ∞ ; s > 0 ] by Bs/~r,~ω ∞ = Bs/~r,~ω ∞ (J, V ) :=( (bckr/~r,~ω, bc(k+1)r/~r,~ω)(s−kr)/r,∞, kr < s < (k + 1)r, (bckr/~r,~ω, bc(k+2)r/~r,~ω)1/2,∞, s = (k + 1)r. (12.21) These spaces allow for a retraction-coretraction theorem as well which provides representations via local coordi- nates. Theorem 12.8 ψ~ω ∞ is a retraction from ℓ∞(Bs/~r ∞ ) onto Bs/~r,~ω ∞ , and ϕ~ω ∞ is a coretraction. P r o o f. We infer from (12.20), Theorem 11.6(ii), and definition (12.21) that Ψ~ω ∞ is a retraction from Bs/~r ∞ , and Φ~ω ∞ is a coretraction. Thus the assertion follows from and Lemma 11.12. Corollary 12.9 ∞ (cid:0)M × J, ℓ∞(E)(cid:1) onto Bs/~r,~ω ∞ =ef ◦ Φ~ω ∞, ψ~ω ∞ = Ψ~ω ϕ~ω ∞ ◦ef −1, (12.22) (12.23) (i) Suppose k0r < s < k1r with k0, k1 ∈ N. Then (bck0r/~r,~ω, bck1r/~r,~ω)(s−k0)/(k1−k0),∞ (ii) If 0 < s0 < s1 and 0 < θ < 1, then (Bs0/~r,~ω . = [Bs0/~r,~ω . = Bsθ /~r,~ω , Bs1/~r,~ω )θ,∞ ∞ ∞ ∞ ∞ , Bs1/~r,~ω ∞ ]θ. . = Bs/~r,~ω ∞ . P r o o f. This is implied by (12.20), (12.22), and Theorem 11.6. Weighted anisotropic Holder spaces are defined by setting BCs/~r,~ω := Bs/~r,~ω ∞ introduce weighted anisotropic little Holder spaces by for s ∈ R\N. Then we bcs/~r,~ω = bcs/~r,~ω(J, V ) is the closure of BC∞/~r,~ω in BCs/~r,~ω for s ≥ 0. Note that this is consistent with Theorem 12.7(ii). Lastly, we get the weighted anisotropic little Besov-Holder space scale [ bs/~r,~ω ∞ ; s > 0 ] by ∞ is the closure of BC∞/~r,~ω in Bs/~r,~ω bs/~r,~ω ∞ . (12.24) ∞ is a retraction from ℓ∞,unif(bs/~r Theorem 12.10 (i) ψ~ω (ii) Suppose k0r < s < k1r with k0, k1 ∈ N. Then (bck0r/~r,~ω, bck1r/~r,~ω)0 (s/r−k0)/(k1−k0),∞ . = bcs/~r,~ω. ∞ ) onto bs/~r,~ω ∞ , and ϕ~ω ∞ is a coretraction. (iii) If 0 < s0 < s1 and 0 < θ < 1, then (bcs0/~r,~ω, bcs1/~r,~ω)0 θ,∞ . = bcsθ /~r,~ω . = [bcs0/~r,~ω, bcs1/~r,~ω]θ. P r o o f. Assertion (ii) and the first part of (iii) follow from Corollary 12.9(i) and definition (12.24). From (ii), Theorem 11.9(i), and (12.20) it follows that Ψ~ω ∞ is a retraction from bcs/~r(cid:0)M × J, ℓ∞(E)(cid:1) onto bcs/~r,~ω, and Φ~ω Due to this the second part of (iii) is now implied by Theorem 11.9(ii). Statement (i) is a consequence of (12.25), (12.23), and Lemma 11.12. ∞ is a coretraction. (12.25) 13 Point-Wise Multipliers In connection with differential and pseudodifferential operators there occur naturally 'products' of tensor fields possessing different regularity of the factors, so called 'point-wise products' or 'multiplications'. Although there is no problem in establishing mapping properties of differential operators say, if the coefficients are smooth, this is a much more difficult task if one is interested in operators with little regularity of the coefficients. Since such low regularity coefficients are of great importance in practice we derive in this and the next section point-wise multiplier theorems which are (almost) optimal. 53 Let Xj, j = 0, 1, 2, be Banach spaces. A multiplication X0 × X1 → X2 from X0 × X1 into X2 is an element of L(X0,X1;X2), the Banach space of continuous bilinear maps from X0 × X1 into X2. Before considering multiplications in tensor bundles we first investigate point-wise products in Euclidean settings. Let Ei = (Ei,·i), i = 0, 1, 2, be finite-dimensional Banach spaces, X ∈ {Rm, Hm}, and Y := X × J. Theorem 13.1 Suppose b ∈ L(E0, E1; E2) and 1 → E Y 2 , 0 × E Y m : E Y is its point-wise extension. Then (u0, u1) 7→ b(u0, u1) and B ∈ {B∞, b∞}. (i) m ∈ L(cid:0)Bs/~r(Y, E0),Bs/~r(Y, E1);Bs/~r(Y, E2)(cid:1) if either s ∈ rN and B ∈ {BC, bc}, or s > 0 (ii) m ∈ L(cid:0)BCs/~r(Y, E0), W s/~r (iii) m ∈ L(cid:0)Bs0/~r P r o o f. (1) Assertion (i) for s ∈ rN and B ∈ {BC, bc} as well as assertion (ii) follow from the product rule. (2) Suppose ui ∈ E Y (Y, E2)(cid:1), s ∈ rN. p (Y, E2)(cid:1), 0 < s < s0. In either case the map b 7→ m is linear and continuous. i , i = 0, 1, and 0 < θ < 1. Then ∞ (Y, E0), Fs/~r p (Y, E1); Fs/~r (Y, E1); W s/~r p p △ξ(cid:0)m(u0, u1)(cid:1) = m(cid:0)△ξu0, u1(· + ξ)(cid:1) + m(u0,△ξu1), sup From this we infer, letting ξ = (h, 0) with h ∈ (0, δ)m, t (cid:2)u0(·, t)(cid:3)δ x (cid:2)u0(x,·)(cid:3)δ t (cid:2)m(u0, u1)(·, t)(cid:3)δ x (cid:2)m(u0, u1)(x,·)(cid:3)δ θ,∞ ≤ c(cid:0)sup θ,∞ ≤ c(cid:0)sup for 0 < δ ≤ ∞. Similarly, sup θ,∞ ku1k∞ + sup θ,∞ ku1k∞ + sup t (cid:2)u1(·, t)(cid:3)δ x (cid:2)u1(x,·)(cid:3)δ θ,∞ ku0k∞(cid:1) θ,∞ ku0k∞(cid:1). ξ ∈ Y. (13.1) By step (1), (11.21), and (11.22) we infer that (i) is true if s ∈ R+\N. Now we fill in the gaps s ∈ N by means of Theorems 11.6(iii) and 11.9(ii) and bilinear complex interpolation (cf. J. Bergh and J. Lofstrom [8, Theorem 4.4.1]). This proves (i) for s > 0 and B ∈ {B∞, b∞}. (3) Assume s ∈ rN and s0 > s. By Theorem 11.8 Bs0/~r ∞ (Y, E0) ֒→ BCs/~r ∞ (Y, E0). Hence we deduce from (ii) ∞ (Y, E0), W s/~r s ∈ rN, Using this, Theorem 8.2(iv), and once more bilinear complex interpolation we obtain ∞ (Y, E0) ֒→ bs/~r (Y, E2)(cid:1), (Y, E2)(cid:1), (4) We assume kr < s < (k + 1)r with k ∈ N. It is well-known that m ∈ L(cid:0)Bs0/~r m ∈ L(cid:0)Bs0/~r ∞(X, E0) × Bs Bs0 p(X, E1) → Bs ∞ (Y, E0), H s/~r (Y, E1); W s/~r (Y, E1); H s/~r p(X, E2), p p p p (v0, v1) 7→ b(v0, v1) 0 < s < s0. s < s0. 54 H. Amann: Anisotropic Function Spaces on Singular Manifolds ∞ is denoted by BUCs0, Th. Runst and W. Sickel is a multiplication (see Remark 4.2(b) in H. Amann [1], where Bs0 [39, Theorem 4.7.1], or V.G. Maz'ya and T.O. Shaposhnikova [33], and H. Triebel [51]), depending linearly and continuously on b. From this we infer km(u0, u1)kp;Bs p(X,E2) ≤ cku0k∞;Bs0 ∞ (X,E0) ku1kp;Bs p(X,E1). By the product rule and (ii) (cid:13)(cid:13)∂ℓ(cid:0)m(u0, u1)(cid:1)(cid:13)(cid:13)p;Lp(X,E2) ≤ c k∂ju0k∞;B(X,E0) k∂ℓ−ju1kp;Lp(X,E1) ℓXj=0 ≤ cku0kk,∞;B(X,E0) ku1kk,p;Lp(X,E1) ≤ cku0k∗∗ s0/r,∞;B(X,E0) ku1ks/r,p;Lp(X,E1) (13.2) (13.3) (13.4) for 0 ≤ ℓ ≤ k. We deduce from (13.1) that, given θ ∈ (0, 1), Hence s0/r,∞;B(X,E0) ku1k∗∗ (s−kr)/r,∞;B(X,E0) k∂k−ju1k∗ s/r,p;Lp(X,E2), (cid:2)m(u0, u1)(cid:3)θ,p;Lp(X,E2) ≤ c(cid:0)[u0]θ,∞;B(X,E0) ku1kp;Lp(X,E1) + ku0k∞;B(X,E0) [u1]θ,p;Lp(X,E1)(cid:1). (cid:2)m(∂ju0, ∂k−ju1)(cid:3)(s−kr)/r,p;Lp(X,E2) ≤ ck∂ju0k∗ ≤ cku0k∗∗ ∞ (cid:0)J, B(X, E0)(cid:1) in the last estimate. Thus s0/r,∞;B(X,E0) ku1k∗∗ p (cid:0)J, Lp(X, E2)(cid:1). (Y, E2)(cid:1), ∞ (cid:0)J, B(X, E0)(cid:1) ֒→ Bs/r (cid:2)∂k(cid:0)m(u0, u1)(cid:3)(s−kr)/r,p;Lp(X,E2) ≤ cku0k∗∗ p(X, E2)(cid:1) ∩ Bs/r m ∈ L(cid:0)Bs0/~r = Lp(cid:0)J, Bs s/r,p;Lp(X,E1). ∞ (Y, E0), Bs/~r p s /∈ rN. (Y, E1); Bs/~r p (s−kr)/r,p;Lp(X,E2) Bs/~r p (Y, E2) . where we used Bs0/r By Corollary 10.2 Thus we infer from (10.1) and (13.2) -- (13.4) that Now we fill in the gaps at s ∈ rN once more by bilinear complex interpolation, which is possible due to Theorems 8.2(iv) and 11.6(iii). Since the last part of the statement is obvious from the above considerations, the theorem is proved. It should be remarked that J. Johnson [20] has undertaken a detailed study of point-wise multiplication in anisotropic Besov and Triebel-Lizorkin spaces on Rn. However, it does not seem to be possible to derive Theo- rem 13.1 from his results. Next we extend the preceding theorem to (x, t)-dependent bilinear operators. Theorem 13.2 Suppose b ∈ L(E0, E1; E2)Y and set m : E Y 0 × E Y 1 → E Y 2 , (u0, u1) 7→(cid:0)(x, t) 7→ b(x, t)(cid:0)u0(x, t), u1(x, t)(cid:1)(cid:1). Then assertions (i) -- (iii) of the preceding theorem are valid in this case also, provided b possesses the same regularity as u0. P r o o f. Consider the multiplication b0 : L(E0, E1; E2) × E0 → L(E1, E2), (b, e0) 7→ b(e0,·) and let m0 be its point-wise extension. By applying Theorem 13.1(i) we obtain m0 ∈ L(cid:0)Bs/~r(cid:0)Y,L(E0, E1; E2)(cid:1),Bs/~r(Y, E0);Bs/~r(cid:0)Y,L(E1, E2)(cid:1)(cid:1), where either s ∈ rN and B ∈ {BC, bc}, or s > 0 and B ∈ {B∞, b∞}. Next we introduce the multiplication and its point-wise extension m1. Then we infer from Theorem 13.1 L(E1, E2) × E1 → E2, (A, e1) 7→ Ae1 if either s ∈ rN and B ∈ {BC, bc}, or s > 0 and B ∈ {B∞, b∞}, m1 ∈ L(cid:0)Bs/~r(cid:0)Y,L(E1, E2)(cid:1),Bs/~r(Y, E1);Bs/~r(Y, E2)(cid:1), (Y, E2)(cid:1), p (Y, E2)(cid:1), m1 ∈ L(cid:0)BCs/~r(cid:0)Y,L(E1, E2)(cid:1), W s/~r m1 ∈ L(cid:0)Bs0/~r ∞ (cid:0)Y,L(E1, E2)(cid:1), Fs/~r m(u0, u1) = m1(cid:0)m0(b, u0), u1(cid:1), (u0, u1) ∈ E Y 0 × E Y 1 . p (Y, E1); Fs/~r (Y, E1); W s/~r p p s ∈ rN, 0 < s < s0. Thus the statement is a consequence of (13.5) -- (13.8). and Note In order to study point-wise multiplications on manifolds we prepare a technical lemma which is a relative of Lemma 12.5. For this we set Teκκu(t) := u(cid:0)(ρκ/ρeκ)µt(cid:1), t ∈ J, Reκκ := Teκκ ◦ Seκκ, κ,eκ ∈ K. Θµ q,κ = (ρκ/ρeκ)µ/qTeκκΘµ q,eκ, 1 ≤ q ≤ ∞. κ,eκ ∈ K, q,κ := ρλ+m/q κ Θµ q,κ(κ⋉ϕ)∗(χκ·), κ⋉ϕ ∈ K⋉Φ. Note We also put bϕ ~ω Then, using u =Peκ π2 eκu, where 55 (13.5) (13.6) (13.7) (13.8) (13.9) (13.10) (13.11) (13.12) (13.13) (13.14) q,κ = Xeκ∈N(κ) bϕ ~ω aeκκReκκϕ~ω q,eκ, kaeκκkk,∞ ≤ c(k), aeκκ := (ρκ/ρeκ)λ+(m+µ)/qχSeκκ(eκ∗πeκ). eκ ∈ N(κ), eκ ∈ N(κ), kReκκk ≤ c, p,κ , BCkr/~r κ κ ∈ K, k ∈ N. p,κ , Bs/~r , bckr/~r κ , Fs/~r ∞,κ, bs/~r ∞,κ}. Then κ⋉ϕ,eκ⋉eϕ ∈ K⋉Φ. ) ∩ L(bckr/~r , bckr/~r ) κ eκ Hence, given q ∈ [1,∞], we deduce from (4.3), Lemma 12.5, and (7.3)(iii) that Lemma 13.3 Suppose k ∈ N and s > 0. Let Gκ ∈ {W kr/~r aeκκ ∈ BCk(Xκ), Reκκ ∈ L(Geκ, Gκ), P r o o f. It is immediate from (9.5), (9.6), (4.3), and Lemma 12.5 that Reκκ ∈ L(W kr/~r p,eκ , W kr/~r p,κ ) ∩ L(BCkr/~r eκ , BCkr/~r κ and that the uniform estimates of (13.14) are satisfied. Now the remaining statements follow by interpolation. 56 H. Amann: Anisotropic Function Spaces on Singular Manifolds Assume Vj = (Vj , hj), j = 0, 1, 2, are metric vector bundles. By a bundle multiplication from V0 ⊕ V1 into V2, denoted by we mean a smooth section m of Hom(V0 ⊗ V1, V2) such that m(v0, v1) := m(v0 ⊗ v1) and m : V0 ⊕ V1 → V2, (v0, v1) 7→ m(v0, v1), m(v0, v1)h2 ≤ cv0h0 v1h1 , vi ∈ Γ(M, Vi), i = 0, 1. Examples 13.4 (a) The duality pairing is a bundle multiplication. h·, ·iV1 : V ∗ 1 ⊕ V1 → M × K, (v∗, v) 7→ hv∗, viV1 (b) Assume σi, τi ∈ N for i = 0, 1. Then the tensor product τ0+τ1 M, ⊗ : T σ0 τ0 M ⊕ T σ1 τ1 M → T σ0+σ1 (a, b) 7→ a ⊗ b is a bundle multiplication where (X ⊗σ0 ⊗ X ∗⊗τ0) ⊗ (X ⊗σ1 ⊗ X ∗⊗τ1) := X ⊗(σ0+σ1) ⊗ X ∗(τ0+τ1), where we set X ⊗σ = X 1 ⊗ ··· ⊗ X σ etc. τ −1 M the contraction with respect to positions i and j, defined by (c) Suppose 1 ≤ i ≤ σ and 1 ≤ j ≤ τ. We denote by Ci τ M → T σ−1 j : T σ Ci j(cid:16) σOk=1 X k ⊗ τOℓ=1 X ∗ ℓ(cid:17) := hX ∗ j , X ii σOk=1 k6=i X k ⊗ X ∗ ℓ , τOℓ=1 ℓ6=j X k ∈ Γ(M, T M ), X ∗ ℓ ∈ Γ(M, T ∗M ). It follows from (a) and (b) that Ci j : T σ1 τ1 M ⊕ T σ2 τ2 M → T σ1+σ2−1 τ1+τ2−1 M, (a, b) 7→ Ci j(a ⊗ b) is a bundle multiplication, where 1 ≤ i ≤ σ1 + σ2 and 1 ≤ j ≤ τ1 + τ2. (d) Let Wj = (Wj, hWj ), j = 0, 1, 2, be metric vector bundles and σj , τj ∈ N. Suppose are bundle multiplications. Set T σj τj (M, Wj) := (T σj w : W0 ⊕ W1 → W2, t : T σ0 τ0 M ⊕ T σ1 τ1 M → T σ2 τ2 M τj M ⊗ Wj, hj) with hj := (·,·)τj τ2 (M, W2), τ1 (M, W1) → T σ2 σj ⊗ hWj . Then t ⊗ w : T σ0 τ0 (M, W0) ⊕ T σ1 Let m be a bundle multiplication from V0 ⊕ V1 into V2. Then defined by t ⊗ w(a0 ⊗ u0, a1 ⊗ u1) := t(a0, a1) ⊗ w(u0, u1), is a bundle multiplication. (cid:0)v0(t), v1(t)(cid:1) 7→ m(cid:0)v0(t), v1(t)(cid:1), Γ(M, V0 ⊕ V1)J → Γ(M, V2)J , is the point-wise extension of m, denoted by m also. (cid:3) t ∈ J, After these preparations we can prove the following point-wise multiplier theorem which is the basis of the more specific results of the next section. Theorem 13.5 Let Wj = (Wj, hWj , Dj), j = 0, 1, 2, be fully uniformly regular vector bundles over M . Assume σj, τj ∈ N satisfy Set σ2 − τ2 = σ0 + σ1 − τ0 − τ1. (13.15) Vj = (Vj, hj,∇j) :=(cid:0)T σj τj (M, Wj), (·,·)τj σj ⊗ hWj ,∇(∇g, Dj)(cid:1) and suppose m : V0 ⊕ V1 → V2 is a bundle multiplication, λ0, λ1 ∈ R, λ2 := λ0 + λ1, and ~ωj := (λj , µ). Then (i) m ∈ L(cid:0)Bs/~r,~ω0(J, V0),Bs/~r,~ω1(J, V1);Bs/~r,~ω2(J, V2)(cid:1), where either s ∈ rN and B ∈ {BC, bc}, or s > 0 and B ∈ {B∞, b∞}. (ii) m ∈ L(cid:0)BCs/~r,~ω0(J, V0), W s/~r,~ω1 (iii) m ∈ L(cid:0)Bs0/~r,~ω0 P r o o f. (1) Suppose (J, V0), F s/~r,~ω1 p ∞ p (J, V1); W s/~r,~ω2 p (J, V1); F s/~r,~ω2 p where dWj is the Fj-valued differential. Set g = gm, M = X ∈ {Rm, Hm}, Ej :=(cid:0)Eσj . τj (Fj ), (··)HS(cid:1), where (··)j := (··)Ej 57 , dWj(cid:1), (J, V2)(cid:1), s ∈ rN. (J, V2)(cid:1), 0 < s < s0. ρ ∼ 1, Wj =(cid:0)M × Fj, (·,·)Fj Vj =(cid:0)X × Ej, (·,·)j, dEj(cid:1), Introducing bases, we define isomorphisms Ej ≃ KNj . By means of them m is transported onto an element (2) Now we consider the general case. We choose uniformly regular atlases K⋉Φj for Wj over K with model KN0 × KN1 ∋ (ξ, η) 7→ (mν2 of L(KN0 , KN1; KN2)X which has the 'matrix representation' ν0ν1(x)ξν0 ην1 )1≤ν2≤N2 ∈ KN2. Assume m ∈ BC∞(cid:0)X,L(E0, E1; E2)(cid:1). Then the assertion follows from Theorem 13.2. fiber Fj. Given κ⋉ϕj ∈ K⋉Φj we define, recalling (5.3), mκ ∈ D(cid:0)Xκ,L(E0, E1; E2)(cid:1) by mκ(η0, η1) := (κ⋉ϕ2)∗(cid:0)χκm(cid:0)(κ⋉ϕ0)∗η0, (κ⋉ϕ1)∗η1(cid:1)(cid:1) for ηj ∈ E Xκ mκ(η0, η1)2 ≤ c ρτ2−σ2 ηj ∈ E Xκ mκ ∈ BCk(cid:0)Xκ,L(E0, E1; E2)(cid:1), . It follows from (5.11) and the fact that m is a bundle multiplication that Hence we infer from (13.15) η00 η11, (3) In the following, it is understood that ϕ~ωj q ρσ0−τ0 κ ρσ1−τ1 κ κ . j j vj ∈ Γ(M, Vj)J , ϕ~ω2 q,κ(cid:0)m(v0, v1)(cid:1) = ρλ0 Consequently, we get from (13.11) κ ρλ1+m/q κ Θµ ϕ~ω2 q,κ(cid:0)m(v0, v1)(cid:1) = Xeκ∈N(κ) k ∈ N. κ⋉ϕj ∈ K⋉Φj, kmκkk,∞ ≤ c(k), is defined by means of K⋉Φj for 1 ≤ q ≤ ∞. Then, given q,κ(κ⋉ϕ2)∗(cid:0)πκm(v0, v1)(cid:1) = mκ(ϕ~ω0 q,κv1). ∞,κv0,bϕ ~ω1 aeκκmκ(ϕ~ω0 ∞,κv0, Reκκϕ~ω1 q,κv1). (13.16) (4) Suppose either s ∈ N and B ∈ {BC, bc}, or s > 0 and B ∈ {B∞, b∞}. Then we infer from (13.13), Lemma 13.3, Theorem 13.2, and steps (1) and (3) that kaeκκmκ(η0, η1)kBs/~r(Yκ,E2) ≤ ckη0kBs/~r(Yκ,E0) kη1kBs/~r(Yκ,E1), uniformly with respect to κ⋉ϕ,eκ⋉eϕ ∈ K⋉Φ2. Hence we get from (13.16) and the finite multiplicity of K ∞(cid:0)m(v0, v1)(cid:1)(cid:13)(cid:13)ℓ∞(Bs/~r(Y,E2)) ≤ ckv0kℓ∞(Bs/~r(Y,E0)) kv1kℓ∞(Bs/~r(Y,E1)). (cid:13)(cid:13)ϕ~ω2 Thus Theorems 12.6 and 12.8 imply, due to (7.8), (13.17) km(v0, v1)kBs/~r,~ω2 (J,V2) ≤ ckv0kBs/~r,~ω0 (J,V0) kv1kBs/~r,~ω1 (J,V1), provided either s ∈ rN and B = BC, or s > 0 and B = B∞. Thanks to Theorems 12.7(i) and 12.10(i) this proves assertion (i). The proofs for (ii) and (iii) are similar. If s ∈ rN and B = bc, or s > 0 and B = b∞, then (13.17) holds with ℓ∞ replaced by ℓ∞,unif everywhere. It is clear that obvious analogues of the results of this section hold in the case of time-independent isotropic spaces. This generalizes and improves [5, Theorem 9.2]. 58 14 Contractions H. Amann: Anisotropic Function Spaces on Singular Manifolds In practice, most pointwise multiplications in tensor bundles occur through contractions of tensor fields. For this reason we specialize in this section the general multiplier Theorem 13.5 to this setting and study the problem of right invertibility of multiplier operators induced by contraction. Let Vi = (Vi, hi), i = 1, 2, be uniformly regular metric vector bundles of rank ni over M with model fiber Ei. Set V0 = (V0, h0) :=(cid:0)Hom(V1, V2), h12(cid:1). By Example 3.1(f), V0 is a uniformly regular vector bundle of rank n1n2 over M with model fiber L(E1, E2). The evaluation map ev : Γ(M, V0 ⊗ V1) → Γ(M, V2), (a, v) 7→ av is defined by av(p) := a(p)v(p) for p ∈ M . Lemma 14.1 The evaluation map is a bundle multiplication. P r o o f. We fix uniformly regular atlases K⋉Φi, i = 1, 2, for Vi over K. Then, using the notation of Section 2, it follows from (2.15) Hence we infer from (3.5) (a∗a)ν1 eν1 = h∗ν1 bν1 1 abν2 bν1 h2,bν2 eν2 aeν2 eν1 . κ∗(a2 uniformly with respect to κ ∈ K. Furthermore, (2.12) and (3.5) imply 1 ν1 κ∗h2,eν2 bν2 κ∗abν2 κ∗aeν2 h0 ) = κ∗(cid:0)tr(a∗a)(cid:1) = κ∗h∗ν1 bν1 κ∗(auh2 ) =(cid:12)(cid:12)(cid:0)(κ⋉ϕ12)∗a(cid:1)(κ⋉ϕ1)∗u(cid:12)(cid:12)(κ⋉ϕ2)∗h2 bν1 ∼ Xν1,ν2 ν12 = tr(cid:0)[κ∗a]∗[κ∗a](cid:1), κ∗aν2 ∼(cid:12)(cid:12)(cid:0)(κ⋉ϕ12)∗a(cid:1)(κ⋉ϕ1)∗u(cid:12)(cid:12)E2 ≤ (κ⋉ϕ12)∗aL(E1,E2) (κ⋉ϕ1)∗uE1 for u ∈ Γ(M, V1) and κ⋉ϕi ∈ K⋉Φi. Since L(E1, E2) is finite-dimensional the operator norm ·L(E1,E2) is equivalent to the trace norm. Hence, using L(E1, E2) ≃ Kn2×n1 and (2.12) and (3.5) once more, we deduce from (14.1) that κ∗(auh2) ≤ cκ∗(ah0 )κ∗(uh1 ) for κ ∈ K. Consequently, (14.1) This proves the lemma. auh2 ≤ cah0 uh1, (a, u) ∈ Γ(M, V0 ⊕ V1). Suppose σ, σi, τ, τi ∈ N for i = 1, 2 with σ + τ > 0. We define the center contraction of order σ + τ, τ +τ1 M ) → Γ(M, T σ1+σ2 as follows: Given (ik) ∈ Jσk, (jk) ∈ Jτk for k = 1, 2, and σ ∈ Jσ, τ ∈ Jτ we set τ2+σ M ⊕ T σ+σ1 [τ ] : Γ(M, T σ2+τ C = C[σ] τ1+τ2 M ), (14.2) etc. Assume a ∈ Γ(M, T σ2+τ τ2+σ M ) is locally represented on Uκ by (i2; j) := (i2,1, . . . , i2,σ2, j1, . . . , jτ ) ∈ Jσ2+τ a = a(i2;j) (j2;i) ∂ ∂x(i2) ⊗ ∂ ∂x(j) ⊗ dx(j2) ⊗ dx(i) and b has a corresponding representation. Then the local representation of C(a, b) on Uκ is given by (j2;i) b(i;i1) a(i2;j) (j;j1) ∂ ∂x(i2) ⊗ ∂ ∂x(i1) ⊗ dx(j2) ⊗ dx(j1). A center contraction (14.2) is a complete contraction (on the right) if σ1 = τ1 = 0. If C is a complete contraction, then we usually simply write a · u for C(a, u). Lemma 14.2 The center contraction associated with the evaluation map ev, 59 C ⊗ ev : Γ(cid:0)M, T σ2+τ τ2+σ (M, V0) ⊕ T σ+σ1 τ +τ1 (M, V1)(cid:1) → Γ(cid:0)M, T σ1+σ2 τ1+τ2 (M, V2)(cid:1), is a bundle multiplication. P r o o f. Note that C is a composition of σ + τ simple contractions of type Ci j. Hence the assertion follows from Lemma 14.1 and Examples 13.4(c) and (d). Henceforth, we write again C for C ⊗ ev, if no confusion seems likely. Furthermore, we use the same symbol for point-wise extensions to time-dependent tensor fields. In addition, we do not indicate notationally the tensor bundles on which C is operating. This will always be clear from the context. Throughout the rest of this section we presuppose • Wi = (Wi, hi, Di), i = 1, 2, 3, are fully uniformly regular vector bundles of rank ni over M with model fiber Fi. For i, j ∈ {1, 2, 3} we set Wij = (Wij , hWij , Dij) :=(cid:0)Hom(Wi, Wj), (··)HS,∇(Di, Dj)(cid:1). Example 3.1(f) guarantees that Wij is a fully uniformly regular vector bundle over M . We also assume for i, j ∈ {1, 2, 3} • σi, τi, σij , τij ∈ N; • Vi = (Vi, hi,∇i) :=(cid:0)T σi • Vij = (Vij, hij ,∇ij) :=(cid:0)T σij τi (M, Wi), (··)τi τij (M, Wij ), (··)τij σi ⊗ hWi,∇(∇g, Di)(cid:1); σij ⊗ hWij ,∇(∇g, Dij)(cid:1); • λi, λij ∈ R, ~ωi = (λi, µ), ~ωij = (λij , µ). Due to Lemma 14.2 we can apply Theorem 13.5 and its corollary with m = C. For simplicity and for their importance in the theory of differential and pseudodifferential operators, we restrict ourselves in the following to complete contractions. It should be observed that condition (14.4) below is void if ∂M = ∅ and J = R. Theorem 14.3 (i) Suppose and λ2 = λ12 + λ1, σ2 = σ12 − τ1, τ2 = τ12 − σ1, s >( −1 + 1/p r(−1 + 1/p) if ∂M 6= ∅, if ∂M = ∅ and J = R+. (14.3) (14.4) Let one of the following additional conditions be satisfied: (α) (β) (γ) (δ) s = t ∈ rN, q := ∞, B = G ∈ {BC, bc}; s = t ∈ rN, q := p, B = BC, G = W ; s = t > 0, s < t, q := ∞, B = G ∈ {B∞, bc∞}; q := p, B = B∞, G = F. Assume a ∈ Bt/~r,~ω12(J, V12). Then A := (u 7→ a · u) ∈ L(cid:0)Gs/~r,~ω1 q (J, V1), Gs/~r,~ω2 q (J, V2)(cid:1), where BC∞ := BC and bc∞ := bc if (α) applies. The map a 7→ A is linear and continuous. 60 H. Amann: Anisotropic Function Spaces on Singular Manifolds (ii) Assume, in addition, λ3 = λ23 + λ2, σ3 = σ23 − τ2, τ3 = τ23 − σ2 and b ∈ Bt/~r,~ω23(J, V23). Set B := (v 7→ b · v). Then BA =(cid:16)u 7→ C[σ2] [τ2] (b, a) · u(cid:17) ∈ L(cid:0)Gs/~r,~ω1 q (J, V1), Gs/~r,~ω3 q (J, V3)(cid:1). Theorem 13.5. P r o o f. (1) Suppose s ≥ 0 with s > 0 if F = B. Then, due to Lemma 14.2, assertion (i) is immediate from (2) Choose uniformly regular atlases K⋉Φi, i = 1, 2, for Wi over K. Let a = a(i12),ν2 (j12),ν1 (t) ∂ ∂x(i12) ⊗ dx(j12) ⊗ b2 ν2 ⊗ βν1 1 , t ∈ J, (14.5) 1, . . . , bi ni) is the local coordinate frame for Wi over Uκ associated with κ⋉Φi, and (β1 be the local representation of a in the local coordinate frame for V12 over Uκ associated with κ⋉ϕ12 ∈ K⋉Φ12, i , . . . , βni where (bi i ) is its dual frame (cf. Example 2.1(b) and (5.5)). Write (ikℓ) = (iℓ; jk) ∈ Jσkℓ and (jkℓ) = (jℓ; ik) ∈ Jτkℓ for k, ℓ ∈ {1, 2} with k 6= ℓ, where (ik) ∈ Jσk and (jk) ∈ Jτk. 1)(cid:1)(cid:1)J by We define a′ ∈ Γ(cid:0)M, T σ12 τ12(cid:0)M, Hom(W ′ 2, W ′ a′(j1;i2),ν2 (i1;j2),ν1 (t) ∂ ∂x(j1) ⊗ ∂ ∂x(i2) ⊗ dx(i1) ⊗ dx(j2) ⊗ βν1 1 ⊗ b2 ν2, t ∈ J, where a′(j1;i2),ν2 (i1;j2),ν1 := a(i2;j1),ν2 (j2;i1),ν1 . It is obvious that 1)(cid:1)(cid:1), the map a 7→ a′ is linear and continuous, and (a′)′ = a. Furthermore, since V ′ (v, u) ∈ Γ(M, V ′ 2 ⊕ V1)J . a′ ∈ Bt/~r,~ω12(cid:0)J, T σ12 hv, a · uiV2 = ha′ · v, uiV1 , τ12(cid:0)M, Hom(W ′ i = T τi 2, W ′ σi (M, W ′ i ), (3) Suppose condition (δ) is satisfied and s < 0. It follows from (14.3), step (1), and (14.6) C(a′) := (v 7→ a′ · v) ∈ L(cid:0)F−s/~r,−~ω2 p′ From Theorem 8.3(ii) and assumption (14.4) we infer (J, V ′ 2 ), F−s/~r,−~ω1 p′ (J, V ′ 1 )(cid:1). F−s/~r,−~ωi p′ (J, V ′ i ) = F−s/~r,−~ωi p′ (J, V ′ i ), i = 1, 2. (14.6) (14.7) (14.8) (4) It is clear that C(b)C(a) : T σ1 Thus we deduce from (8.5), (14.7), and (14.8) that C(a′) = C(a)′. Hence, using (a′)′ = a, we get the remaining part of assertion (i), provided s 6= 0 if F = B. Now this gap is closed by interpolation. [τ2] (b, a)v)(cid:17). [τ2] (b, a), v(cid:17) =(cid:16)v 7→ C[σ2] v 7→ C(cid:0)b, C(a, v)(cid:1) = C(cid:16)C[σ2] τ1 (M, V1) → T σ3 τ3 (M, V3) is given by (14.9) [τ2] in Theorem 13.5. Also set V0 := W23, V1 := W12, and V2 := W13 in Lemma 14.2. Then it Set m = C[σ2] follows from that lemma and Theorem 13.5 that Thus claim (ii) is a consequence of (14.9) and assertion (i). C[σ2] [τ2] (b, a) ∈ Bt/~r,(λ3−λ1,µ)(cid:0)J, T σ3+τ1 τ3+σ1 (M, W13)(cid:1). Next we study the invertibility of the linear map A. We introduce the following definition: Suppose t > 0 and 61 a ∈ Bt/~r,~ω11 ∞ (J, V11), σ11 = τ11 = σ1 + τ1. Then a ia said to be λ11-uniformly contraction invertible if there exists a−1 ∈ Γ(M, V11)J satisfying a−1 · (a · u) = u, a · (a−1 · u) = u, u ∈ Γ(M, V1)J , and ρ−λ11 a−1(t)h11 ≤ c, t ∈ J. (14.10) (14.11) (14.12) Note that the second part of (14.10) guarantees that the complete contractions in (14.11) are well-defined. Also note that there exists at most one a−1 satisfying (14.11), the contraction inverse of a. For abbreviation, we put Bt/~r,~ω11 ∞,inv (J, V11) :=(cid:8) a ∈ Bt/~r,~ω11 ∞ (J, V11) ; a is λ11-uniformly contraction invertible(cid:9). Let X and Y be Banach spaces and let U be open in X . Then f : U → Y is analytic if each x0 ∈ U has a neighborhood in which f can be represented by a convergent series of continuous monomials. If f is analytic, then f is smooth and it can be locally represented by its Taylor series. If K = C, and f is (Fr´echet) differentiable, then it is analytic. For this and further details we refer to E. Hille and R.S. Phillips [19]. To simplify the presentation we restrict ourselves now to the most important cases in which B = B∞. We leave it to the reader to carry out the obvious modifications in the following considerations needed to cover the remaining instances as well. Proposition 14.4 Suppose σ11 = τ11 = σ1 + τ1. Then Bt/~r,~ω11 ∞,inv (J, V11) is open in Bt/~r,~ω11 ∞ (J, V11). If a ∈ Bt/~r,~ω11 ∞,inv (J, V11), then a−1 ∈ Bt/~r,(−λ11,µ) ∞,inv (J, V11). The map is analytic. Bt/~r,~ω11 ∞,inv (J, V11) → Bt/~r,(−λ11,µ) ∞ (J, V11), a 7→ a−1 P r o o f. (1) Without loss of generality we let F1 = Kn and set σ := σ11. Note that E := L(Kn)mσ ×mσ is a Banach algebra with unit of dimension N 2 := (nmσ)2. It is obvious that we can fix an algebra isomorphism from E onto KN ×N by which we identify E with KN ×N . For b ∈ KN ×N we denote by b♮ the (N × N )-matrix of cofactors of b. Thus b♮ = [b♮ ij] with b♮ ij := det[b1, . . . , bi−1, ej, bi+1, . . . , bN ], (14.13) where b1, . . . , bN are the columns of b and ej is the j-th standard basis vector of KN . Then, if b is invertible, (2) Suppose either X := (Qm, gm) or X := (Qm ∩ Hm, gm), and Y = X × J. Set ∞ (cid:0)J, B(X, E)(cid:1), X t/~r(Y, E) := B(cid:0)J, Bt ∞(Rm, E) by restriction, of course. Note ∞(X, E) is obtained from Bt where Bt b−1 =(cid:0)det(b)(cid:1)−1 b♮. ∞(X, E)(cid:1) ∩ Bt/r X t/~r(Y, E) ֒→ B∞(Y, E). (14.14) (14.15) (14.16) It follows from Theorem 13.5 that X t/~r(Y, E) is a Banach algebra with respect to the point-wise extension of the (matrix) product of E. Assume b ∈ X t/~r(Y, E) and b(y) is invertible for y ∈ Y such that y ∈ Y. b−1(y)E ≤ c0, (14.17) 62 H. Amann: Anisotropic Function Spaces on Singular Manifolds plicities. 1/c(c0) ≤(cid:12)(cid:12) det(cid:0)b(y)(cid:1)(cid:12)(cid:12) ≤ c(c0), Then the spectrum σ(cid:0)b(y)(cid:1) of b(y) is bounded and has a positive distance from 0 ∈ C, uniformly with respect to y ∈ Y . Hence due to the fact that det(cid:0)b(y)(cid:1) can be represented as the product of the eigenvalues of b(y), counted with multi- Since det(cid:0)b(y)(cid:1) is a polynomial in the entries of b(y) and X t/~r(Y ) := X t/~r(Y, K) is a multiplication algebra (14.19) Using the chain rule if t ≥ 1 (cf. Lemma 1.4.2 of [4]), we get(cid:0) det(b)(cid:1)−1 ∈ X t/~r(Y ) from (14.18) and (14.19). Now we deduce from (14.13), (14.14), and the fact that X t/~r(Y ) is a multiplication algebra, that det(b) ∈ X t/~r(Y ). y ∈ Y, we infer (14.18) b−1 ∈ X t/~r(Y, E), kb−1kX t/~r(Y,E) ≤ c(c0), whenever b ∈ X t/~r(Y, E) satisfies (14.17). c0 = c0(b) ≥ 1 is open in X t/~r(Y, E). By (14.16) it is obvious that the set of all invertible elements of X t/~r(Y, E) satisfying (14.17) for some (3) Assume K⋉Φ1 is a uniformly regular atlas for W1 over K. Given κ⋉ϕ1 ∈ K⋉Φ1, put κ v := ρλ1 χ~ω1 κ Θµ ∞,κ(κ⋉ϕ1)∗v, χ~ω11 κ a := ρλ11 κ Θµ ∞,κ(κ⋉ϕ11)∗a ∞,inv (J, V11). Then we deduce from (14.11) (see Example 3.1(f)) and Suppose a ∈ Bt/~r,~ω11 for v ∈ Γ(M, V1)J and a ∈ Γ(M, V11)J , respectively, and Yκ := Qm κ (cid:0)a−1 · (a · v)(cid:1) = (χ(−λ11,µ) κ v = χ~ω1 χ~ω1 for κ⋉ϕ1 ∈ K⋉Φ1 and v ∈ Γ(Uκ, V1)J . Note that χ~ω1 follows from (14.20) that χ(−λ11,µ) a right inverse. Hence bκ := χ~ω11 a−1 is a left inverse for χ~ω11 κ a is invertible in B∞(Yκ, E) and κ κ κ × J. κ is a bijection from Γ(Uκ, V1)J onto (Eσ1 τ1 )Yκ. Thus it κ a in B∞(Yκ, E). Similarly, we see that it is also a−1)(χ~ω11 κ a)χ~ω1 κ v (14.20) (14.21) (14.22) (14.23) b−1 κ = χ(−λ11,µ) κ a−1. We infer from (4.1)(iv), (5.11), (3.5), (3.10), (14.10), and (14.12) that b−1 κ E ≤ c Θµ ∞,κκ∗(ρ−λ11 a−1h11 ) ≤ c, κ⋉ϕ1 ∈ K⋉Φ1. Recalling (4.3), (7.10), and (13.10) we find bκ = χ~ω11 κ (cid:16) Xeκ∈N(κ) π2 eκa(cid:17) = Xeκ∈N(κ) Seκκ(eκ∗πeκ)Reκκϕ~ω11 ∞,eκa. Since a ∈ Bt/~r,~ω11 (J, V11) implies ϕ~ω11 13.3, Theorem 13.5, and definition (14.15) ∞ a ∈ ℓ∞(Bt/~r ∞ ∞ ) we deduce from (14.23), (7.3)(iii), Lemmas 12.5 and kbκkX t/~r(Yκ,E) ≤ ckakt/~r,∞;~ω11, κ⋉ϕ1 ∈ K⋉Φ1. (14.24) Set aκ := ρ−λ11 κ bκ. Then it follows from (14.22) and (14.24) that a−1 κ kX t/~r(Yκ,E) ≤ c, Employing (7.3)(iii) and Theorem 13.5 once more we derive from (14.25) a−1 κ ∈ X t/~r(Yκ, E), kρ−λ11 ρ−λ11 κ κ κ⋉ϕ1 ∈ K⋉Φ1. (14.25) ϕ(−λ11,µ) ∞,κ a−1 = χ(−λ11,µ) κ (πκa−1) = (κ∗πκ)b−1 κ ∈ Bt/~r ∞ (Yκ, E) = Bt/~r ∞,κ (14.26) and ϕ(−λ11,µ) ∞ a−1 ∈ ℓ∞(Bt/~r ∞ ). Hence Theorem 12.8 implies a−1 = ψ(−λ11,µ) ∞ (ϕ(−λ11,µ) ∞ a−1) ∈ Bt/~r,(−λ11,µ) ∞ (J, V11). 63 (14.27) (4) Let X be a Banach algebra with unit e. Denote by G the group of invertible elements of X . For b0 ∈ X and δ > 0 let X (b0, δ) be the open ball in X of radius δ, centered at b0. Suppose b0 ∈ G. Then 0 (cid:1)b0 and b = b0 − (b0 − b) =(cid:0)e − (b0 − b)b−1 k(b0 − b)b−1 0 k ≤ kb0 − bk kb−1 imply that b ∈ X (b0,kb−1 0 k/2) is invertible and b ∈ X (b0,kb−1 0 k < 1/2, 0 k/2), b−1 = b−1 0 (cid:1)−1 0 (cid:0)e − (b0 − b)b−1 = b−1 0 ∞Xi=0(cid:0)(b0 − b)b−1 0 (cid:1)i . (14.28) In fact, this Neumann series has the convergent majorantPi 2−i. Note that pi(x) := (−1)ib−1 continuous homogenous polynomial in x ∈ X . Hence it follows from (14.28) 0 (xb−1 0 )i is a b−1 = ∞Xi=0 pi(b − b0), b ∈ X (b0,kb−1 0 k/2), and this series converges uniformly on X (b0,kb−1 b 7→ b−1 is analytic. (5) We set X := B(cid:0)J, B(M, V11)(cid:1) and define a multiplication by (a, b) 7→ C[σ1] algebra with unit e :=(cid:0)(p, t) 7→ idL((V1)p)). Consider the continuous linear map 0 k/2). Thus G is open and the inversion map inv : G → X , [τ1] (a, b). Then X is a Banach Then G := f −1(G) is open in Bt/~r,~ω11 ∞ f : Bt/~r,~ω11 ∞ (J, V11) → X , (J, V11). Consequently, a 7→ ρλ11 a. f0 := inv ◦ (f G) : G → X , a 7→ (ρλ11 a)−1 is continuous (in fact, analytic) by step (4). Note that a−1 = ρλ11 f0(a) is the contraction inverse of a. Further- more, f0(a) ∈ X implies ρ−λ11 a−1(t)h11 = f0(a)(t)h11 ≤ c, t ∈ J. Hence each a ∈ G is λ11-uniformly contraction invertible. Conversely, if a ∈ Bt/~r,~ω11 contraction invertible, then a belongs to G. Thus G = Bt/~r,~ω11 open. ∞,inv (J, V11) which shows that Bt/~r,~ω11 (J, V11) is λ11-uniformly ∞,inv (J, V11) is ∞ (6) We denote by Gκ the group of invertible elements of Bt/~r and (14.25)) guarantees that b0,κ := χ~ω11 ∞,κ. Suppose a0 ∈ G. Then step (3) (see (14.24) κ a0 ∈ Gκ and 0,κkBt/~r + kb−1 kb0,κkBt/~r ∞,κ ∞,κ ≤ c, κ⋉ϕ1 ∈ K⋉Φ1. Hence we infer from step (4) that there exists δ > 0 such that the open ball Bt/~r and the inversion map invκ : Gκ → Bt/~r sense that the series ∞,κ(b0,κ, δ) belongs to Gκ for κ ∈ K ∞,κ(b0,κ, δ), uniformly with respect to κ ∈ K in the ∞,κ is analytic on Bt/~r converges in Bt/~r ∞,κ, uniformly with respect to bκ ∈ Bt/~r ∞,κ(b0,κ, δ) and κ ∈ K. Xi b−1 0,κ(cid:0)(b0,κ − bκ)b−1 0,κ(cid:1)i 64 Note that H. Amann: Anisotropic Function Spaces on Singular Manifolds Bt/~r ∞ (b0, δ) :=Yκ Bt/~r ∞,κ(b0,κ, δ) is open in ℓ∞(Bt/~r ∞ ). The above considerations show that is analytic. It follows from (14.24) that the linear map inv : Bt/~r ∞ (b0, δ) → ℓ∞(Bt/~r ∞ ), b 7→(cid:0)invκ(bκ)(cid:1) v 7→ (χ~ω11 (14.29) χ~ω11 : Bt/~r,~ω11 ∞ (J, V11) → ℓ∞(Bt/~r ∞ ), ∞ (b0, δ)(cid:1) ∩ G is an open neighborhood of a0 in G. It is a consequence (14.30) κ v) Consider the point-wise multiplication operator is continuous. Hence G0 := (χ~ω11 )−1(cid:0)Bt/~r of (14.29) and (14.30) that inv ◦ χ~ω11 is an analytic map from G0 into ℓ∞(Bt/~r ∞ ). b 7→(cid:0)(κ∗πκ)bκ(cid:1). ∞ ) → ℓ∞(Bt/~r ∞ ), It follows from (7.3) and Theorem 13.5 that it is a well-defined continuous linear map. π : ℓ∞(Bt/~r If a ∈ G0, then we know from (14.21) and (14.24) that inv ◦ χ~ω11 (a) = (χ(−λ11,µ) Hence we see by (14.26) and (14.27) that a−1 = ψ(−λ11,µ) κ ∞ ). a−1(cid:1) ∈ ℓ∞(Bt/~r ◦ π ◦ inv ◦ χ~ω11 a. Thus ◦ π ◦ inv ◦ χ~ω11 : G0 → Bt/~r,(−λ11,µ) ∞ ∞ (J, V11) (a 7→ a−1) = ψ(−λ11,µ) ∞ is analytic, being a composition of analytic maps. This proves the proposition. Henceforth, we set F∞ := B∞ so that Fq is defined for 1 < q ≤ ∞. Theorem 14.5 Suppose 1 < q ≤ ∞ and t > 0 and s satisfies (14.4) with s < t if q = p, and s = t if q = ∞. Assume σ11 = τ11 = σ1 + τ1 and λ2 = λ11 + λ1. If a ∈ Bt/~r,~ω11 ∞,inv (J, V11), then A = C(a) ∈ Lis(cid:0)Fs/~r,~ω1 q (J, V1), Fs/~r,~ω2 q (J, V1)(cid:1) (14.31) and A−1 = C(a−1). The map a 7→ A−1 is analytic. P r o o f. It follows from Theorem 14.3(i) and Proposition 14.4 that (14.31) applies and a 7→ C(a−1) is ana- lytic. Part (ii) of that theorem implies A−1 = C(a−1). Next we study the problem of the right invertibility of the operator A of Theorem 14.3. This is of particular importance in connection with boundary value problems. First we need some preparation. We assume σ12 = σ2 + τ1, (14.32) Then, given a ∈ Γ(M, V12)J , there exists a unique a∗ ∈ Γ(M, V21)J , the complete contraction adjoint of a, such that (14.33) τ12 = τ2 + σ1, τ21 = σ12. σ21 = τ12, h2(a · u, v) = h1(u, a∗ · v), (u, v) ∈ Γ(M, V1 ⊕ V2)J . Indeed, recalling (14.5) set (a∗)(i21),ν1 (j21),ν2 := g(i1)(eı1) (j1)(e1) h∗ν1 eν1 W1 a(eı2;e1),eν2 (e2;eı1),eν1 g(e2)(j2) (eı2)(i2) hW2,eν2ν2 . (14.34) Then it follows from (2.15) and hkℓ = (··)τkℓ (a∗)(i21),ν1 (j21),ν2 σkℓ ⊗ hWkℓ that ∂ ∂x(i21) ⊗ dx(j21) ⊗ b1 ν1 ⊗ βν2 2 65 (14.35) is the local representation of a∗ over Uκ with respect to the coordinate frame for V21 over Uκ associated with κ⋉ϕ21, We set and suppose a ∈ Bt/~r,~ω12 ∞ λ∗ 21 := λ12 + σ21 − τ21, ~ω∗ 21 := (λ∗ 21, µ) (14.36) (J, V12). Then it is a consequence of (3.5), (5.9), (5.10), (14.34), and (14.35) that ∞,κakBt/~r From this, Theorem 12.8, (7.8), and (14.34) we infer ∞,κa∗kBt/~r ∞ (Yκ,Eσ21 kϕ~ω∗ 21 τ21 ) ∼ kϕ~ω12 (a 7→ a∗) ∈ L(cid:0)Bt/~r,~ω12 ∞ (J, V12), Bt/~r,~ω∗ ∞ 21 (J, V21)(cid:1). τ12 ), ∞ (Yκ,Eσ12 κ⋉ϕi ∈ K⋉Φi, i = 1, 2. Assume a∗(p, t) ∈ L(cid:0)(V2)p, (V1)p(cid:1) is injective for (p, t) ∈ M × J. Then a(p, t) ∈ L(cid:0)(V1)p, (V2)p(cid:1) is sur- jective. This motivates the following definition: a ∈ Bt/~r,~ω12 ρλ12+(τ12−σ12)/2 a∗(t) · uh1 ≥ uh2/c, (J, V12) is λ12-uniformly contraction surjective if u ∈ Γ(M, V2), t ∈ J. The reason for the specific choice of the exponent of ρ will become apparent below. We set ∞ (14.38) (14.37) For abbreviation, we put Bt/~r,~ω12 ∞,surj (J, V12) :=(cid:8) a ∈ Bt/~r,~ω12 a ⊙ a∗ := C[σ1] [τ1] (a, a∗), ∞ (J, V12) ; a is λ12-uniformly contraction surjective(cid:9). σ22 := τ22 := σ2 + τ2, λ22 := 2λ12 + τ12 − σ12. It follows from (14.37) and Theorem 13.5 that is a well-defined continuous quadratic map. Hence it is analytic. Bt/~r,~ω12 ∞ (J, V12) → Bt/~r,~ω22 ∞ (J, V22), a 7→ a ⊙ a∗ (14.39) ∞,surj (J, V12) iff a ⊙ a∗ ∈ Bt/~r,~ω22 Lemma 14.6 a ∈ Bt/~r,~ω12 P r o o f. It follows from (14.33) that h2(cid:0)(a ⊙ a∗) · u, v(cid:1) = h2(cid:0)a · (a∗ · u), v(cid:1) = h1(a∗ · u, a∗ · v), ∞,inv (J, V22). (u, v) ∈ Γ(M, V2 ⊕ V2)J . (14.40) Hence C(a ⊙ a∗) is symmetric and positive semi-definite. We see from (14.40) that (14.38) is equivalent to ρλ22 h2(cid:0)(a ⊙ a∗)(t) · u, u(cid:1) ≥ u2 h2/c, u ∈ Γ(M, V2), t ∈ J. By symmetry this inequality is equivalent to the λ22-uniform contraction invertibility of a ⊙ a∗. In the next proposition we give a local criterion for checking λ12-uniform surjectivity. Proposition 14.7 Suppose a ∈ Bt/~r,~ω12 (J, V12). Let K⋉Φi, ∞ i = 1, 2, be uniformly regular atlases for Vi over K. Set aκ(t)(ζ, ζ) := X(i1)∈Jσ1 , (j1)∈Jτ1 1≤ν1≤n1 (j2;i1),ν1 (cid:12)(cid:12)(cid:12)κ∗a(i2;j1),ν2 (t) ζ(j2) 2 (i2),ν2(cid:12)(cid:12)(cid:12) for ζ ∈ Eτ2 σ2 (F ∗ 2 )Qm κ and t ∈ J. Then a is λ12-uniformly contraction surjective iff κ ∈ K, aκ(t)(ζ, ζ) ∼ ζ2, ζ ∈ Eτ2 2 )Qm κ , σ2 (F ∗ ρ2λ12 κ t ∈ J. 66 H. Amann: Anisotropic Function Spaces on Singular Manifolds P r o o f. Assume v ∈ Γ(M, V2)J and put w := (h2)♭v ∈ Γ(M, V ′ 2 )J , where V ′ 2 = T τ2 σ2 (M, W ′ 2). Then, lo- cally on Uκ, v = v(i2),ν2 (j2) where ∂ ∂x(i2) ⊗ dx(j2) ⊗ b2 ν2, w = w(j2) (i2),ν2 ∂ ∂x(j2) ⊗ dx(i2) ⊗ βν2 2 , due to h2 = (··)τ2 , σ2 ⊗ hW2. Thus it follows from (14.34) that, locally on Uκ, (i2)(eı2) hW2,ν2 eν2 v(eı2),eν2 = g(j2)(e2) w(j2) (i2),ν2 (e2) Hence (cid:0)(a ⊙ a∗) · v(cid:1)(i2),ν2 h2(cid:0)(a ⊙ a∗) · v, v(cid:1) = g(j2)(b2) (j2) = a(i2;j1),ν2 (j2;i1),ν1 g(i1)(eı1) (j1)(e1) h∗ν1 eν1 W1 a(eı2;e1),eν2 (e2;eı1),eν1 w(e2) (eı2),eν2 . (i2)(bı2) hW2,ν2 bν2(cid:0)(a ⊙ a∗) · v(cid:1)(i2),ν2 =(cid:0)(a ⊙ a∗) · v(cid:1)(i2),ν2 (j2) w(j2) g(i1)(eı1) (j1)(e1) h∗ν1 eν1 = a(i2;j1),ν2 (j2;i1),ν1 w(j2) (i2),ν2 (i2),ν2 (j2) W1 v(bı2),bν2 (b2) a(eı2;e1),eν2 (e2;eı1),eν1 w(e2) (eı2),eν2 . Thus we deduce from (3.5), (4.3), (5.8), (5.9) (applied to Wi) aκ(ζ, ζ) (14.41) for κ⋉ϕi ∈ K⋉Φi, i = 1, 2, and v ∈ Γ(M, V2)J , where κ κ∗(cid:0)ρλ22 h2(cid:0)(a ⊙ a∗) · v, v(cid:1)(cid:1) ∼ ρλ22+2(τ1−σ1) 2 )(cid:1)Qm ζ := (κ⋉ϕ2)∗(cid:0)(h2)♭v(cid:1) ∈(cid:0)Eτ2 2 is the bundle metric of V ∗ ) ∼ ρ2(τ2−σ2) h2 ) = κ∗(w2 ζ2 Eτ2 2 ), σ2 (F ∗ σ2 (F ∗ h∗ 2 κ 2 we get from (5.11) κ ×J . κ ∈ K, Since (h2)♭ is an isometry and h∗ κ∗(v2 (14.42) (14.43) with v and ζ being related by (14.42). Now the assertion follows from (14.32), (14.36), (14.41), and (14.43). (J, V12) and ac ∈ Bt/~r,(−λ12,µ) Γ(M, V2)J . Then ac is a right contraction inverse of a. Suppose a ∈ Bt/~r,~ω12 Proposition 14.8 Let conditions (14.32) be satisfied. Then Bt/~r,~ω12 ∞ ∞ there exists an analytic map (J, V21) are such that a · (ac · v) = v for v belonging to ∞,surj (J, V12) is open in Bt/~r,~ω12 ∞ (J, V12) and I c : Bt/~r,~ω12 ∞,surj (J, V12) → Bt/~r,(−λ12,µ) ∞ (J, V21) such that I c(a) is a right contraction inverse for a. P r o o f. It follows from (14.39), Proposition 14.4, and Lemma 14.6 that S := Bt/~r,~ω12 ∞,surj (J, V12) is open in Bt/~r,~ω12 ∞ (J, V12). Set [τ2](cid:0)a∗, (a ⊙ a∗)−1(cid:1), where (a ⊙ a∗)−1 is the contraction inverse of a ⊙ a∗ ∈ Bt/~r,~ω22 rem 13.5 imply that I c is an analytic map from S into Bt/~r,(−λ12,µ) I c(a) := C[σ2] ∞ ∞ a ∈ S, (J, V22). Then (14.37), (14.39), and Theo- (J, V21). Since a ·(cid:0)I c(a) · v(cid:1) = a ·(cid:0)a∗ ·(cid:0)(a ⊙ a∗)−1 · v(cid:1)(cid:1) = (a ⊙ a∗) ·(cid:0)(a ⊙ a∗)−1 · v(cid:1) = v, the assertion follows. v ∈ Γ(M, V2), After these preparations it is easy to prove the second main theorem of this section. For this it should be noted that definition (14.38) applies equally well if a ∈ Bt/~r,~ω12(J, V12) where either B = b∞, or t ∈ rN and B ∈ {BC, bc}. Hence Bt/~r,~ω12 (J, V12) is defined in these cases also. surj Theorem 14.9 Let assumptions (14.3) and (14.4) be satisfied and 1 < q ≤ ∞. (i) Assume s < t if q = p, and s = t > 0 if q = ∞. Then there exists an analytic map (J, V1)(cid:1) ∞,surj (J, V12) → L(cid:0)Fs/~r,~ω2 (J, V2), Fs/~r,~ω1 Ac : Bt/~r,~ω12 q q such that Ac(a) is a right inverse for A(a) = (v 7→ a · v). 67 (ii) There exists an analytic map Ac : Bs/~r,~ω12 surj (J, V12) → L(cid:0)Bs/~r,~ω2(J, V2),Bs/~r,~ω1(J, V1)(cid:1) such that Ac(a) is a right inverse for A(a) if either s ∈ rN and B ∈ {BC, bc}, or s > 0 and B = b∞. P r o o f. The first assertion is an obvious consequence of Theorem 14.3 and Proposition 14.8. The second claim is obtained by modifying the above arguments in the apparent way. As in the preceding section, the above results possess obvious analogues applying in the isotropic case. 15 Embeddings Now we complement the embedding theorems of Section 8 by establishing further inclusions between anisotropic weighted spaces. Theorem 15.1 Suppose λ0 < λ1 and put ~ωi := (λi, µ) for i = 0, 1. Then Fs/~r,~ω0 p d ֒→ Fs/~r,~ω1 p if ρ ≤ 1, whereas ρ ≥ 1 implies Fs/~r,~ω1 p d ֒→ Fs/~r,~ω0 p for s ∈ R. Similarly, Bs/~r,~ω0 ֒→ Bs/~r,~ω1 if ρ ≤ 1, and Bs/~r,~ω1 ֒→ Bs/~r,~ω0 for ρ ≥ 1, if either s > 0 and B ∈ {B∞, b∞}, or s ∈ rN and B ∈ {BC, bc}. P r o o f. If ρ ≤ 1, then it is obvious that W kr/~r,~ω0 p d ֒→ W kr/~r,~ω1 p , W kr/~r,~ω0 p d ֒→ W kr/~r,~ω1 p , BCkr/~r,~ω0 ֒→ BCkr/~r,~ω1 for k ∈ N. Thus, by duality, W kr/~r,~ω0 p d ֒→ W kr/~r,~ω1 p , k ∈ −N×. By interpolation Fs/~r,~ω0 p d ֒→ Fs/~r,~ω1 p follows. The proof of the other embeddings is similar. The next theorem contains Sobolev-type embedding results. In the anisotropic case they involve the weight exponents as well as the regularity parameters. Theorem 15.2 (i) Suppose s0 < s1 and p0, p1 ∈ (1,∞) satisfy s1 − (m + r)/p1 = s0 − (m + r)/p0. ֒→ Fs0/~r,~ω0 p1 Set ~ω0 :=(cid:0)λ + (m + µ)(1/p1 − 1/p0), µ(cid:1). Then Fs1/~r,~ω (ii) Assume t > 0 and s ≥ t + (m + r)/p. Set ~ω∞ :=(cid:0)λ + (m + µ)/p, µ(cid:1). Then Fs/~r,~ω p0 d p . (15.1) ֒→ bt/~r,~ω∞ ∞ . 68 H. Amann: Anisotropic Function Spaces on Singular Manifolds P r o o f. (1) Note that s1 > s0 and (15.1) imply p0 > p1. Hence it follows from (15.1) and Theorems 3.3.2, 3.7.5, and 4.4.1 of [4] that ℓp1(Fs1/~r p1 ) d ֒→ ℓp1 (Fs0/~r p0 ) d ֒→ ℓp0(Fs0/~r p0 ). Also note that we get ψ~ω p1 = ψ~ω0 p0 from (15.1). Thus we infer from Theorem 9.3 that ϕ~ω p1 ✲ Fs1/~r,~ω p1 ✄ ❄ Fs0/~r,~ω0 p0 ψ~ω0 p0 ✛ ℓp1 (Fs1/~r p1 ✄ ) d ❄ ℓp0 (Fs0/~r p0 ) is commuting. From this we obtain assertion (i). (2) We infer from Lemma 9.2 and [4, Theorem 3.3.2] that Bs/~r p,κ = Bs/ν p,κ ֒→ Bs/ν p,∞,κ ֒→ Bt/ν ∞,∞,κ = Bt/~r ∞,κ and from [4, Theorem 3.7.1] that H s/~r D(Yκ, E) in Fs/~r p,κ it follows, due to D(Yκ, E) ֒→ bt/~r ℓp(Fs/~r p,κ ֒→ Bs/ν ∞,κ, that Fs/~r p,∞,κ. Consequently, Fs/~r p,κ ֒→ bt/~r p ) ֒→ ℓ∞(bt/~r ∞ ). ∞,κ. From this and the density of p,κ ֒→ Bt/~r ∞,κ. Thus, by (7.1), (15.2) It is obvious that D(Y, E) ֒→ ℓ∞,unif(bt/~r this and (15.2) we deduce ℓp(Fs/~r 12.10 that the diagram p ) ֒→ ℓ∞,unif(bt/~r ∞ ). By Theorem 9.3 we know that D(Y, E) is dense in ℓp(Fs/~r ). From ∞ , we infer from Theorems 9.3 and p ∞ ). Observing ψ~ω p = ψ~ω∞ ϕ~ω p ✲ Fs/~r,~ω p ✄ ❄ bt/~r,~ω∞ ∞ ψ~ω∞ ∞ ✛ ℓp(F s/~r p ) ✄ ❄ ℓ∞,unif(bt/~r ∞ ) is commuting. This proves (ii). Remark 15.3 Define the anisotropic small Holder space Cs/~r,~ω ֒→ Ct/~r,~ω∞ ∞ for s > 0. Then the above proof shows Fs/~r,~ω in Bs/~r,~ω p 0 0 = Cs/~r,~ω if the hypotheses of (ii) are satisfied. (J, V ) to be the closure of D(J,D) (cid:3) 0 16 Differential Operators First we establish the mapping properties of ∇ and ∂ in anisotropic weighted Bessel potential and Besov spaces. They are, of course, of fundamental importance for the theory of differential equations. Theorem 16.1 Suppose either s ≥ 0 and G = Fp, or s > 0 and G ∈ {B∞, b∞}. Then τ +1)(cid:1) τ +1)(cid:1) ∩ L(cid:0)G(s+1)/~r,~ω, Gs/~r,~ω(J, V σ ∇ ∈ L(cid:0)Gs+1,λ, Gs,λ(V σ and ∂ ∈ L(cid:0)G(s+r)/~r,~ω, Gs/~r,(λ+µ,µ)(cid:1). P r o o f. We consider the time-dependent case. The proof in the stationary setting is similar. (1) From (5.15) and (5.16) we know that (κ⋉ϕ)∗∇v = ∂xv + aκv, v ∈ C(cid:0)J, C1(Xκ, E)(cid:1), kAκk ≤ c, κ⋉ϕ ∈ K⋉Φ. κ , Gs/~r(Yκ, Eσ By [4, Theorem 4.4.2] and Theorems 11.6 and 11.9 we get Aκ := (v 7→ aκχv) ∈ L(cid:0)G(s+1)/~r τ +1)(cid:1), ∂ ∈ L(G(s+r)/~r (2) Set q := p if G = Fp, and q := ∞ otherwise. Then, given u ∈ G(s+1)/~r,~ω ∂x ∈ L(cid:0)G(s+1)/~r τ +1)(cid:1), , Gs/~r(Yκ, Eσ κ κ κ Hence we get from (13.11) ϕ~ω q,κ(∇u) = ρλ+m/q κ Θµ q,κ(κ⋉ϕ)∗(πκ∇u) = (κ∗πκ)(cid:0)(κ⋉ϕ)∗∇(cid:1)(bϕ ~ω q,κu). beκκ(κ⋉ϕ)∗∇(Reκκϕ~ω q,eκu), ϕ~ω q,κ(∇u) =Xeκ∈K , Gs/~r κ ). , (16.1) where aκ ∈ C∞(cid:0)Qm 13.1 and G(s+1)/~r κ κ ,L(Eσ ֒→ Gs/~r κ τ , Eσ that τ +1)(cid:1) satisfies kaκk ≤ c(k) for κ⋉ϕ ∈ K⋉Φ. Hence it follows from Theorem 69 where beκκ = (κ∗πκ)aeκκ and aeκκ is defined by (13.12). From this, (7.3), (13.13), Lemma 13.3, step (1), Theo- rem 13.5, and the finite multiplicity of K we infer kϕ~ω q (∇u)kℓq(Gs/~r(Y,Eσ τ +1)) ≤ ckϕ~ω q ukℓq(G(s+1)/~r) for u ∈ G(s+1)/~r,~ω. Using Theorems 9.3, 12.8, and 12.10 we thus obtain kϕ~ω q (∇u)kℓq(Gs/~r(Y,Eσ q(cid:0)ϕ~ω q (∇u)(cid:1) by invoking these theorems once more. Thus the first assertion follows from ∇u = ψ~ω ϕ(λ+µ,µ) τ +1)) ≤ ckukG(s+1)/~r,~ω , (3) Since (see (9.5)) u ∈ G(s+1)/~r,~ω. (∂u) = ρµ κϕ~ω q,κ∂u = ∂(ϕ~ω q,κu), q,κ the second assertion is implied by the second part of (16.1) and the arguments of step (2). By combining this result with Theorem 14.3 and embedding theorems of the preceding section we can derive mapping properties of differential operators. To be more precise, for k ∈ N× we consider operators of the form A = Xi+jr≤kr aij · ∇i∂j where aij are suitably regular time-dependent vector-bundle-valued tensor field homomorphisms and aij · ∇i∂j equals(cid:0)u 7→ aij · (∇i∂ju)(cid:1), of course. Recall that F∞ = B∞. Theorem 16.2 Let ¯W = ( ¯W , h ¯W , D ¯W ) be a fully uniformly regular vector bundle over M . Suppose k, ¯σ, ¯τ belong to N and ¯λ ∈ R. For 0 ≤ i ≤ k set σi := ¯σ + τ + i, τi := ¯τ + σ, ~¯ω := (¯λ, µ). (i) Given i, j ∈ N with i + jr ≤ k, put λij := ¯λ − λ − jµ, ~ωij := (λij , µ). Let condition (14.4) be satisfied. Supposebs > s if q = p, andbs = s > 0 if q = ∞, and ∞ i + jr ≤ k. (16.2) (cid:0)J, T σi aij ∈ B bs/~r,~ωij A ∈ L(cid:0)F(s+kr)/~r,~ω τi(cid:0)M, Hom(W, ¯W )(cid:1)(cid:1), ¯τ ( ¯W )(cid:1)(cid:1), (cid:0)J, V ¯σ , Fs/~r,~¯ω q q 1 < q ≤ ∞. Then 70 H. Amann: Anisotropic Function Spaces on Singular Manifolds If B bs/~r,~ωij ∞ in (16.2) is replaced by b bs/~r,~ωij ∞ , then A ∈ L(cid:0)b(s+kr)/~r,~ω ∞ , bs/~r,~¯ω ∞ (cid:0)J, V ¯σ ¯τ ( ¯W )(cid:1)(cid:1). = (m + r)/(kr − i − jr), > p, = p, i + jr > kr − (m + r)/p, i + jr = kr − (m + r)/p, i + jr < kr − (m + r)/p, (ii) Fix and set pij ~ωij := (λij , µ) Then for i + jr ≤ kr. Suppose λij := ¯λ − λ − jµ − (m + µ)/pij, τi (M, Hom(W, ¯W ))(cid:1)(cid:1). aij ∈ Lpij(cid:0)J, Lλij pij(cid:0)T σi ¯τ ( ¯W ))(cid:1)(cid:1). , Lp(cid:0)J, L A ∈ L(cid:0)W kr/~r,~ω (iii) In either case the map (aij 7→ A) is linear and continuous. ∇i∂j ∈ L(cid:0)F(s+kr)/~r,~ω P r o o f. (1) Theorem 16.1 implies , F(s−i+(k−j)r)/~r,(λ+jµ,µ) and this is also true if F∞ is replaced by b∞. Since F(s−i+(k−j)r)/~r,(λ+jµ,µ) ¯λ p (V ¯σ p q q q (J, V σ τ +i)(cid:1) (J, V σ τ +i) ֒→ Fs/~r,(λ+jµ,µ) q (J, V σ τ +i) assertion (i) follows from Theorem 14.3. (2) If i + jr > kr − (m + r)/p, then we get from Theorem 15.2(i) τ +i) ֒→ Lqij(cid:0)J, L H (kr−i−jr)/~r,(λ+jµ,µ) (J, V σ p ¯λ−λij qij (V σ τ +i)(cid:1), (16.3) where 1/qij := 1/p − 1/pij. Theorem 15.2(i) once more, Suppose i + jr = kr − (m + r)/p. Then pij > p implies s := i + jr + (m + r)/pij < kr. Thus, invoking H (kr−i−jr)/~r,(λ+jµ,µ) p (J, V σ τ +i) ֒→ H (s−i−jr)/~r,(λ+jµ,µ) p (J, V σ ¯λ−λij qij (V σ τ +i)(cid:1). τ +i) ֒→ Lqij(cid:0)J, L τ +i)(cid:1). u ∈ H kr/~r,~ω ¯λ−λij ∞ (V σ p . p (J, V σ H (kr−i−jr)/~r,(λ+jµ,µ) Since qij = ∞ if pij = p we get in either case from (16.3) If i + jr < kr − (m + r)/p, then we deduce from Theorem 15.2(i) τ +i) ֒→ L∞(cid:0)J, L τ +i)(cid:1) =: Lqij (J, Xij ), ∇i∂ju ∈ Lqij(cid:0)J, L ¯λ+¯τ −¯σ aij · ∇i∂ju¯h ≤ cρλij +τi−σi aijhij ρ σi ⊗ hW ¯W , and hi := (·,·)τ +i ¯λ−λij qij (V σ ρ where ¯h := (·,·)¯τ Note ¯λ + ¯τ − ¯σ = λij + τi − σi + ¯λ − λij + τ + i − σ implies, due to Lemma 14.2, ¯σ ⊗ h ¯W , hij := (·,·)τi kaij · ∇i∂jukLp(J,L¯λ pij(cid:0)T σi τi(cid:0)M, Hom(W, ¯W )(cid:1)(cid:1). By combining this with (16.3) and using W kr/~r,~ω ¯τ ( ¯W ))) ≤ kaijkLpij (J,Yij ) k∇i∂jukLqij (J,Xij ), p (V ¯σ p where Yij := Lλij get assertion (ii). (3) The last claim is obvious. ¯λ−λij +τ +i−σ ∇i∂juhi, σ ⊗ hW . Hence, by Holder's inequality, . = F kr/~r,~ω p we It is clear which changes have to be made to get analogous results for 'stationary' differential operators in the time-independent isotropic case. Details are left to the reader. 17 Extensions and Restrictions 71 In many situations it is easier to consider anisotropic function spaces on the whole line rather than on the half-line. Therefore we investigate in this section the possibility of extending half-line spaces to spaces on all of R. We fix h ∈ C∞(cid:0)(0,∞), R(cid:1) satisfying ts h(t) dt < ∞, Z ∞ 0 s ∈ R, (−1)kZ ∞ 0 tkh(t) dt = 1, k ∈ Z, (17.1) and h(1/t) = −th(t) for t > 0. Lemma 4.1.1 of [4], which is taken from [18], guarantees the existence of such a function. Let X be a locally convex space. Then the point-wise restriction, r+ : C(R,X ) → C(R+,X ), is a continuous linear map. For v ∈ C(R+,X ) we set εv(t) :=Z ∞ h(s)v(−st) ds, 0 u 7→ u R+, t < 0, and e+v :=( v εv on R+, on (−∞, 0). (17.2) (17.3) (17.4) It follows from (17.1) that e+ is a continuous linear map from C(R+,X ) into C(R,X ), and r+e+ = id. Thus point-wise restriction (17.2) is a retraction, and e+ is a coretraction. By replacing R+ in (17.2) by −R+ and using obvious modifications we get the point-wise restriction r− 'to the negative half-line' and a corresponding extension operator e−. The trivial extension operator is defined by e+ 0 v := v on R+ and e+ e+ 0 : C(0)(R+,X ) :=(cid:8) u ∈ C(R+,X ) ; u(0) = 0(cid:9) → C(R,X ) 0 v := 0 on (−∞, 0). Then is a retraction, and e+ 0 is a coretraction. 0 := r+(1 − e−r−) : C(R,X ) → C(0)(R+,X ) r+ (17.5) We define: Thus Fs/~r,~ω p (cid:0)(0,∞), V(cid:1) is the closure of D(cid:0)(0,∞),D(cid:1) in Fs/~r,~ω p (R+, V ). Fs/~r,~ω p (R+, V ) ֒→ Fs/~r,~ω p (cid:0)(0,∞), V(cid:1) ֒→ Fs/~r,~ω p (R+, V ). Now we can prove an extension theorem 'from the half-cylinder M × R+ to the full cylinder M × R.' Theorem 17.1 (i) Suppose s ∈ R where s > −1 + 1/p if ∂M 6= ∅. Then the diagram ✲ d ✄ ✂ id D(R, D) ✄ ✂ ✲ Fs/~r,~ω p (R, V ) D(R+, D) ❅ e+ ❅❘ r+ ❄ ✠ D(R+, D) ✄ ✂ d d Fs/~r,~ω p e+ ✠ ❅ r+ ❅❘ Fs/~r,~ω p ✲ (R+, V ) id ❄ (R+, V ) is commuting. 72 (ii) If s > 0, then H. Amann: Anisotropic Function Spaces on Singular Manifolds D(cid:0)(0, ∞), D(cid:1) ❅ ✄ ✂ e+ 0 ❅❘ id r+ ✠ 0 D(cid:0)(0, ∞), D(cid:1) ❄ ✄ ✂ D(R, D) ✄ ✂ d d d ✲ Fs/~r,~ω p (R, V ) ✲ (cid:0)(0, ∞), V (cid:1) p Fs/~r,~ω e+ 0 ✠ ❅ r+ 0 ❅❘ Fs/~r,~ω p id ❄ ✲ (cid:0)(0, ∞), V (cid:1) is a commuting diagram as well. P r o o f. (1) Suppose M = (X, gm) with X ∈ {Rm, Hm}, ρ = 1, W = X × F, D = dF . (17.6) If k ∈ N, then it is not difficult to see that r+ is a retraction from W kr/~r (X × R+, E), and e+ is a coretraction. (cf. steps (1) and (2) of the proof of Theorem 4.4.3 of [4]). Thus, if s > 0, the first assertion follows by interpolation. (X × R, E) onto W kr/~r p p (2) Let (17.6) be satisfied. Suppose s > 0 and J = R+. It is an easy consequence of (17.7) (17.8) Note that Lp(V ) = Lp(X, E) is a UMD space (e.g.; [2, Theorem III.4.5.2]). Hence [4, Lemma 4.1.4], defini- tion (17.5), and the arguments of step (1) show that From this it is obvious that e+ Fs/~r Fs/~r p ( J, V ), Fs/~r p(V )(cid:1) ∩ Fs/r p(V )(cid:1) ∩ Fs/r p (J, V ) = Lp(cid:0)J, Fs p ( J, V ) = Lp(cid:0)J, Fs 0 ∈ L(cid:0)Fs/~r p (cid:0)R, Lp(V )(cid:1), Fs/r p (cid:0)J, Lp(V )(cid:1) p (cid:0) J, Lp(V )(cid:1). p (R, V )(cid:1). p (cid:0) J, Lp(V )(cid:1)(cid:1). (3) Assume (17.6) and s < 0 with s > −1 + 1/p if X = Hm. Then F−s (cid:0) J, Lp′(V ′)(cid:1) = Lp′(cid:0)J,F−s =(cid:0)F−s/~r ( J, V ′) = Lp′(cid:0)J, F−s p′ (V ′)(cid:1) ∩ F−s/r 0 ∈ L(cid:0)Fs/r ( J, V ′)(cid:1)′ = Fs/~r p′ (J, V ′). F−s/~r Fs/~r p (J, V ) r+ p′ p′ p′ . . From this, (17.7), and (17.8) we deduce assertion (ii) in this setting. Thus, by (8.5), of [4]. Hence p′ (V ′) = F−s p′ (V ′) by Theorem 4.7.1(ii) p′ (V ′)(cid:1) ∩ F−s/r p′ (cid:0) J, Lp′(V ′)(cid:1) The results of Section 4.2 of [4] imply r+, respectively e+, is the dual of e+ 0 . From this and step (2) it follows (see [4, (4.2.3)] that assertion (i) holds in the present setting if s < 0, provided s > −1 + 1/p if X = Hm. 0 , respectively r+ (4) It follows from (9.2) and (17.2) -- (17.4) that r+ ◦ Θµ q,κ = Θµ q,κ ◦ r+, e+ ◦ Θµ q,κ = Θµ q,κ ◦ e+ for 1 ≤ q ≤ ∞. Hence Thus r+ 0 ◦ Θµ q,κ = Θµ q,κ ◦ r+ 0 , e+ 0 ◦ Θµ q,κ = Θµ q,κ ◦ e+ 0 . (17.9) ϕ~ω q,κ(r+u) = ρλ+m/q κ Θµ q,κ(κ⋉ϕ)∗(πκr+u) = r+(ϕ~ω q,κu) and, similarly, 73 This implies that ϕ~ω and Theorem 9.3. q and ψ~ω q commute with r+, r+ ψ~ω q,κ(r+vκ) = r+(ψ~ω 0 , e+, and e+ q,κvκ). 0 . Hence the statements follow from steps (1) -- (3) The next theorem concerns the extension of Besov-Holder spaces from half- to full cylinders. Theorem 17.2 Suppose either s ∈ rN and B ∈ {BC, bc}, or s > 0 and B ∈ {B∞, b∞}. Then r+ is a re- traction from Bs/~r,~ω(R, V ) onto Bs/~r,~ω(R+, V ), and e+ is a coretraction. P r o o f. (1) Let k ∈ N and B ∈ {BC, bc}. It is obvious that r+ ∈ L(cid:0)Bkr/~r,~ω(R, V ),Bkr/~r,~ω(R+, V )(cid:1). e+ ∈ L(cid:0)Bkr/~r,~ω(R+, V ),Bkr/~r,~ω(R, V )(cid:1). It follows from (17.1) that ε ∈ L(cid:0)Bkr/~r,~ω(R+, V ),Bkr/~r,~ω(−R+, V )(cid:1). Thus, by the second part of (17.1) and (17.4), From this we get the assertion in this case. (2) If s > 0 and B ∈ {B∞, b∞}, then, due to Corollary 12.9, we obtain the statement by interpolation from the results of step (1). Lastly, we consider little Besov-Holder spaces 'with vanishing initial values'. They are defined as follows: If (17.10) (17.11) (s−kr)/r,∞, kr < s < (k + 1)r, 1/2,∞, s = (k + 1)r, k ∈ N, then u ∈ bckr/~r,~ω(cid:0)(0,∞), V(cid:1) iff u ∈ bckr/~r,~ω(R+, V ) and ∂ju(0) = 0 for 0 ≤ j ≤ k. Furthermore, bs/~r,~ω ∞ (cid:0)(0,∞), V(cid:1) is defined by (cid:0)bckr/~r,~ω((0,∞), V ), bc(k+1)r/~r,~ω((0,∞), V )(cid:1)0 (cid:0)bckr/~r,~ω((0,∞), V ), bc(k+2)r/~r,~ω((0,∞), V )(cid:1)0  where k ∈ N. Theorem 17.3 Let k ∈ N and s > 0. Then r+ and from bs/~r,~ω ∞ (R, V ) onto bs/~r,~ω ∞ (cid:0)(0,∞), V(cid:1), and e+ 0 is a retraction from bckr/~r,~ω(R, V ) onto bckr/~r,~ω(cid:0)(0,∞), V(cid:1) 0 is a coretraction. P r o o f. It is easily seen by (17.5) and the preceding theorem that the assertion is true for bckr/~r,~ω spaces. The stated results in the remaining cases now follow by interpolation. 18 Trace Theorems Suppose Γ is a union of connected components of ∂M . We denote by •ι : Γ ֒→ M the natural injection and endow Γ with the induced Riemannian metric •g := •ι∗g. Let (ρ, K) be a singularity datum for M . For κ ∈ KΓ we put U •κ := ∂Uκ = Uκ ∩ Γ and •κ := ι0 ◦ ( •ι∗κ) : U •κ → Rm−1, where ι0 : {0} × Rm−1 → Rm−1, (0, x′) 7→ x′. K := { •κ ; κ ∈ KΓ } is a normalized atlas for Γ, the one induced by K. We set •ρ := •ι∗ρ = ρΓ. It follows Then that ( •ρ, K) is a singularity datum for Γ, so that Γ is singular of type [[ •ρ]]. Henceforth, it is understood that Γ is given this singularity structure induced by T(M ). • • We denote by • W = WΓ the restriction of W to Γ and by h • W := •ι∗hW the bundle metric on Γ induced by hW . Furthermore, the connection D • W for • W , induced by D, is defined by restricting D : T M × C∞(M, W ) → C∞(M, W ) to T Γ × C∞(Γ, • W ), 74 H. Amann: Anisotropic Function Spaces on Singular Manifolds considered as a map into C∞(Γ, over Γ. • W ). Then • • W = ( W , h • W , D • W ) is a fully uniformly regular vector bundle We set • V := T σ bundle metric on T σ • • W ) and endow it with the bundle metric τ (Γ, τ Γ induced by •g. Then we equip τ Γ is the ). Hence V ) is a well-defined anisotropic weighted space with respect to the , where (··)T σ ∇ := ∇(∇•g , D • V with the metric connection h := (··)T σ τ Γ ⊗ h • (J, W W • • • • V , • • • V = ( boundary weight function •ρ. ∇). It follows that Fs/~r,~ω h, p We write n = n(Γ) for the inward pointing unit normal on Γ. In local coordinates, κ = (x1, . . . , xm), n =(cid:0)pg11∂Uκ(cid:1)−1 ∂ ∂x1 . (18.1) • V ), is defined by (18.2) n(Γ)u ∈ D(Γ, V ∗). • Let u ∈ D = D(M, V ) and k ∈ N. The trace of order k of u on Γ, ∂k nu = ∂k h∂k nu, ai • V ∗ :=(cid:10)∇kuΓ, a ⊗ n⊗k(cid:11) • V ∗ , a ∈ D(Γ, n = ∂k n(Γ) over J, that is, (∂k n(Γ) and call it trace operator on Γ. We write again ∂k We also set γΓ := ∂0 n(cid:0)u(t)(cid:1) for t ∈ J and u ∈ D(J,D), and call it lateral trace operator of ∂k of order k on Γ × J. Correspondingly, the lateral trace operator on Γ × J is the point-wise extension of γΓ, denoted by γΓ as well. Moreover, n(Γ) for the point-wise extension nu)(t) := ∂k ∂k n,0 : D(cid:0)(0,∞),D(cid:1) → D(cid:0)(0,∞),D(Γ, • V )(cid:1), u 7→ ∂k nu is the restriction of ∂k n to D(cid:0)(0,∞),D(cid:1). Assume J = R+. Then M0 := M × {0} is the initial boundary of the space-time (half-)cylinder M × R+. The initial trace operator is the linear map where M0 is identified with M . Furthermore, γM0 : D(R+,D) → D, u 7→ u(0), t=0 := γM0 ◦ ∂k : D(R+,D) → D, ∂k u 7→ (∂ku)(0) is the initial trace operator of order k. Suppose s0 > 1/p. The following theorem shows, in particular, that there exists a unique such that extending γΓ and being a retraction. Furthermore, there exists a coretraction (γc there is , B(s0−1/p)/~r,(λ+1/p,µ) (J, p p (γΓ)s0 ∈ L(cid:0)Fs0/~r,~ω Γ)s ∈ L(cid:0)B(s−1/p)/~r,(λ+1/p,µ) Γ)s(cid:12)(cid:12)D(cid:0)J,D(Γ, V )(cid:1) = (γc (γc (γc (γc p • (i) (ii) (J, • V ), Fs/~r,~ω p Γ)s0(cid:12)(cid:12)D(cid:0)J,D(Γ, • V )(cid:1) Γ)s0 such that, for each s ∈ R, (cid:1) V )(cid:1), (18.3) • Γ)s is for each s > 1/p a coretraction for (γΓ)s. Thus (γΓ)s0 is for each s0 > 1/p uniquely determined by γΓ and (γc continuous extension or restriction of (γc and (γc Γ)s, respectively, without fearing confusion. So we can say γc Γ)s can be obtained for any s ∈ R by unique Γ for (γΓ)s γ is a universal coretraction for the retraction Γ)s0 for any s0 > 1/p. Hence we simply write γΓ and γc γΓ ∈ L(cid:0)Fs/~r,~ω p , B(s−1/p)/~r,(λ+1/p,µ) p (J, s > 1/p, • V )(cid:1), herewith expressing properties (18.3). Similar conventions hold for higher order trace operators and traces oc- curring below. Theorem 18.1 Suppose k ∈ N. (i) Assume Γ 6= ∅ and s > k + 1/p. Then ∂k n is a retraction from Fs/~r,(λ,µ) p (J, V ) onto B(s−k−1/p)/~r,(λ+k+1/p,µ) p • (J, V ). It possesses a universal coretraction (γk n)c satisfying ∂i n ◦ (γk n)c = 0 for 0 ≤ i ≤ k − 1. (ii) Suppose s > r(k + 1/p). Then ∂k t=0 is a retraction from Fs/~r,(λ,µ) p (R+, V ) onto Bs−r(k+1/p),λ+µ(k+1/p) p (V ). There exists a universal coretraction (γk t=0)c such that ∂i t=0 ◦ (γk t=0)c = 0 for 0 ≤ i ≤ k − 1. (iii) Let Γ 6= ∅ and s > k + 1/p. Then ∂k n,0 is a retraction from Fs/~r,(λ,µ) p (cid:0)(0,∞), V(cid:1) onto B(s−k−1/p)/~r,(λ+k+1/p,µ) p (cid:0)(0,∞), • V(cid:1). The restriction of (γk n)c to the space on the right side of (18.4) is a universal coretraction. 75 (18.4) P r o o f. (1) Suppose X ∈ {Rm, Hm}, M = (X, gm), ρ = 1, W = X × F , and D = dF so that V = X × E. Put Y := X × J. Assume either Γ 6= ∅ or J = R+. If J = R, then M × J = Y = Hm+1. If J = R+ and Γ = ∅, then M × J = Rm × R+ ≃ Hm+1. Finally, if J = R+ and Γ 6= ∅, then M × J = Hm × R+ ≃ R+ × R+ × Rm−1, that is, M × J is a closed 2-corner in the sense of Section 4.3 of [4]. In each case ≃ is simply a permutation diffeomorphism. If J = R+ and Γ 6= ∅, then assertion (i) follows from Theorem 4.6.3 and the definition of the trace operator for a face of R+ × R+ × Rm−1, that is, formula (4.10.12) of [4]. Claim (iii) is a consequence of Theorem 4.10.3 of [4] (choose any κ therein with κ > s + 1). If either J = R or Γ = ∅, then assertions (i) and (ii) follow from Theorem 4.6.3 of [4]. (2) Now we consider the general case. Suppose Γ 6= ∅. For t > 1/p we set • B(t−1/p)/~r p,κ :=( B(t−1/p)/~r p {0} (∂Yκ, E) if κ ∈ KΓ, otherwise. Let γκ be the trace operator on ∂Yκ = {0} × Rm−1 × J if κ ∈ KΓ, and γκ := 0 otherwise. Set γk,κ := ρk κ(cid:0)pγκ(κ∗g11)(cid:1)−k γκ ◦ ∂k 1 , κ ∈ K. 1 is a retraction from Fs/~r p,κ onto • B(s−1/p)/~r p,κ and that there (18.5) It follows from step (1), (18.1), and (18.2) that γκ ◦ ∂k k,κ satisfying exists a universal coretractioneγc (settingeγc k,κ := 0 if κ ∈ K\KΓ). We put (γκ ◦ ∂i k,κ := ρ−k γc k,κ = 0, 1) ◦eγc κ (cid:0)pγκ(κ∗g11)(cid:1)k 0 ≤ i ≤ k − 1, κ ∈ K. k,κ, eγc Then (3.7) and (4.1) imply γk,κ ∈ L(Fs/~r p,κ , • B(s−k−1/p)/~r p,κ ), γc k,κ ∈ L( • B(s−k−1/p)/~r p,κ , Fs/~r p,κ ) 76 and H. Amann: Anisotropic Function Spaces on Singular Manifolds From (18.5) and Leibniz' rule we thus infer kγk,κk + kγc k,κk ≤ c, κ ∈ K. γi,κ ◦ γc k,κ = δikid, 0 ≤ i ≤ k. (18.6) (3) We set ( •π •κ , •χ •κ ) := (πκ, χκ)U •κ for •κ ∈ • system subordinate to K. We denote by • K. Then it is verified that(cid:8) ( •π •κ , •χ •κ ) ; •κ ∈ • K(cid:9) is a localization • ψ~ω p : ℓp( • B(s−k−1/p)/~r p p ) → B(s−k−1/p)/~r,~ω p . Correspondingly, •ϕ~ω • (J, V ) the 'boundary retraction' defined analogously to ψ~ω p is the 'boundary coretraction'. We write •κ⋉ •ϕ for the restriction of κ⋉ϕ ∈ K⋉Φ to Γ and put n ◦ (κ⋉ϕ)∗, •κ ( •κ⋉ •ϕ)∗ ◦ ∂k Ck,κ := •ρk κ⋉ϕ ∈ KΓ⋉Φ, and Ck,κ := 0 otherwise. Note •ρ •κ = ρκ for κ ∈ KΓ. It follows from (5.15), (18.1), and (18.2) that Ck,κv = γk,κv + v ∈ D(cid:0)J,D(∂Xκ, E)(cid:1), and (5.16) implies kaℓ,kkk−1,∞ ≤ c for 0 ≤ ℓ ≤ k − 1 and κ ∈ K. Hence, using Fs/~r rem 13.5, we find κ ∈ K. Ck,κ ∈ L(Fs/~r p,κ , kCk,κk ≤ c, B(s−k−1/p)/~r aℓ,κγℓ,κv, p,κ ), • k−1Xℓ=0 p,κ ֒→ F(s−k+ℓ)/~r p,κ (18.7) and Theo- (18.8) (4) For u ∈ D(J,D) •π •κ ∂k nu = ∂k n(πκu) − k−1Xj=0(cid:16) k j(cid:17)(∂k−j n πκ)∂j n(χκu), κ ∈ K, (18.9) setting ∂k since Θµ p, •κ p,κ for κ ∈ KΓ. Similarly, using (13.11) and (13.12) also, p, •κ Θµ = Θµ = •ρk κ( •κ⋉ •ϕ)∗ ◦ ∂k n(πκu)(cid:1) nv := 0 if supp(v) ∩ Γ = ∅. Note ( •κ⋉ •ϕ)∗(cid:0)∂k •ρλ+k+1/p+(m−1)/p •κ n ◦ (κ⋉ϕ)∗(cid:0)ρλ+m/p ( •κ⋉ •ϕ)∗(cid:0)(∂k−j = Ck−j,κ(κ∗πκ)Cj,κ(bϕ ~ω •ρλ+k+1/p+(m−1)/p •κ Θµ p, •κ κ Θµ p,κ(κ⋉ϕ)∗(πκu)(cid:1) = Ck,κ(ϕ~ω n(χκu)(cid:1) p,κu) = Xeκ∈N(κ) n πκ)∂j From this, (18.9), and (18.10) we get p,κu), Ck−j,κ(κ∗πκ)Cj,κ(aeκκReκκϕ ~ω p,eκu). •ϕ(λ+k+1/p,µ) p, •κ (∂k nu) = Ck,κ(ϕ~ω Ak−1,eκκ(ϕ~ω p,eκu), p,κu) + Xeκ∈N(κ) k−1Xj=i(cid:16) k j(cid:17)(cid:16) j bi,eκκ := − i(cid:17)Ck−j,κ(κ∗πκ)Cj−iaeκκ. where Ak−1,eκκ := k−1Xi=0 bi,eκκCi,κ ◦ Reκκ, (18.10) (18.11) 77 kCℓ,κk ≤ c(n), κ ∈ K, 0 ≤ ℓ ≤ k, n ∈ N. It is obvious that From this, (7.3), (13.13), and Theorem 13.5 we obtain , BCn(∂Xκ, E)(cid:1), κ Cℓ,κ ∈ L(cid:0)BCn+ℓ bi,eκκ ∈ BC∞(∂Xκ, E), kbi,eκκkn,∞ ∈ c, Hence, using Theorem 13.5 once more, we get from (18.8) and Lemma 13.3 κ ∈ K, 0 ≤ i ≤ k, n ∈ N. eκ ∈ N(κ), Ak−1,eκκ ∈ L(Fs/~r p,eκ , (5) We define Ck by • B(s−k−1/p)/~r p,κ ), kAk−1,eκκk ≤ c, C kv :=(cid:16)Ck,κvκ + Xeκ∈N(κ) Ak−1,eκκveκ(cid:17)κ∈K , Then we deduce from (18.8), (18.12), and the finite multiplicity of K κ ∈ K. (18.12) eκ ∈ N(κ), v = (vκ). k,κ = id. Furthermore, recalling (13.9) and using •ρκ = ρκ for (18.13) • p p ), ℓp( )(cid:1). B(s−k−1/p)/~r C k ∈ L(cid:0)ℓp(Fs/~r Employing (18.6) and (18.7) we infer Ck,κ ◦ γc κ ∈ KΓ, n ◦ (κ⋉ϕ)∗ ◦ Teκκ ◦ (κ⋉ϕ)∗(eκ⋉eϕ)∗(χ·) κ( •κ⋉ •ϕ)∗ ◦ ∂i S • k,eκ = 0 for 0 ≤ i ≤ k − 1. Thus, setting γγc By this, (18.6), and (18.7) it follows Ci,κ ◦ Reκκ ◦ γc p )(cid:1), C i ◦ γγ c ), ℓp(Fs/~r 0 ≤ i ≤ k. From (18.11), (18.13), and the first claim of Theorem 9.3 we get k ∈ L(cid:0)ℓp( Ci,eκ = (ρκ/ρeκ)kR • eκ •κ Ci,κ ◦ Reκκ = •ρk = (ρκ/ρeκ)kT • eκ •κ B(s−k−1/p)/~r k = δikid, Ci,eκ. γγc eκ •κ p • kv := (γc k,κvκ), (18.14) , B(s−k−1/p)/~r,(λ+k+1/p,µ) (J, • V )(cid:1). (6) Given v ∈ ∂i n(ψ~ω ∂k n = • ψ(λ+k+1/p,µ) p • B(s−k−1/p)/~r p , p v) =Xκ =X•κ ρ−(λ+m/p) κ ∂i •ρ−(λ+i+m/p) •κ p p ◦ C k ◦ ϕ~ω p ∈ L(cid:0)Fs/~r,~ω p,κπκ(κ⋉ϕ)∗vκ(cid:1) n(cid:0)Θ−µ p, •κ(cid:16) •π •κ ( •κ⋉ •ϕ)∗Ci,κvκ + •ρi Θ−µ • = ψ(λ+i+1/p,µ) p •ρ−(λ+i+m/p) •κ Θ−µ p, •κ ( •κ⋉ •ϕ)∗ Ci,κvκ +Xκ •κ i−1Xj=0(cid:16) i n πκ)∂j j(cid:17)(∂i−j i−1Xj=0(cid:16) i n(cid:0)(κ⋉ϕ)∗vκ(cid:1)(cid:17) j(cid:17)Ci−j,κ(πκ)Cj,κvκ. Thus we infer from (18.6), (18.7), and (18.14) n(ψ~ω ∂i p γγ c kw) = δik • ψ(λ+i+1/p,µ) p w, w ∈ • B(s−k−1/p)/~r p , 0 ≤ i ≤ k. Now (18.14) and the first part of Theorem 9.3 imply (γk n)c := ψ~ω p ◦ γγ c k ◦ •ϕ(λ+k+1/p,µ) p ∈ L(cid:0)B(s−k−1/p)/~r,(λ+k+1/p,µ) p (J, • V ), Fs/~r,~ω p (cid:1) and ∂i n(γk n)c = δikid for 0 ≤ i ≤ k. This proves assertion (i). 78 H. Amann: Anisotropic Function Spaces on Singular Manifolds (7) By invoking in the preceding argumentation the second statement of Theorem 9.3 we see that assertion (iii) is true. (8) We denote by ∂k t=0,κ the initial trace operator of order k for Yκ = Xκ × R+. It follows from step (1) that p is a retraction and there exists a universal coretraction ∂k t=0 : ℓp(Fs/~r p ) → ℓp(Bs−r(k+1/p) (∂k t=0)c : ℓp(Bs−r(k+1/p) p ) → ℓp(Fs/~r t=0,κvκ) ), v 7→ (∂k p ), w 7→(cid:0)(∂k t=0,κ)cwκ(cid:1) ∂j t=0 ◦ (∂ k t=0)c = δjkid, (9) We deduce from (9.5) and step (1) ∂k t=0,κ ◦ ϕ~ω p,κ = ϕλ+µ(k+1/p) p,κ 0 ≤ j ≤ k. ◦ ∂k t=0, κ ∈ K. (18.15) (18.16) such that Hence ∂ k t=0 ◦ ϕ~ω p = ϕλ+µ(k+1/p) p ◦ ∂k t=0. From this and Theorems 7.1 and 9.3 we infer t=0 ◦ ϕ~ω ∂k t=0 = ψλ+µ(k+1/p) ◦ ∂ k p (10) Set Then, similarly as above, (R+, V ), Bs−r(k+1/p),λ+µ(k+1/p) p (V )(cid:1). p p ∈ L(cid:0)Fs/~r,~ω p ◦ (∂k (γk t=0)c := ψ~ω t=0)c ◦ ϕλ+µ(k+1/p) p . (γk t=0)c ∈ L(cid:0)Bs−r(k+1/p),λ+µ(k+1/p) p (V ), Fs/~r,~ω p For 0 ≤ j ≤ k we get from (9.5) and (18.15) ∂j t=0(γk t=0)cw = ∂j (R+, V )(cid:1). w(cid:17) t=0(cid:16)Xκ =Xκ p,κ ψλ+µ(j+1/p) = δjkψλ+µ(j+1/p) p p,κ t=0,κ)c ◦ ϕλ+µ(k+1/p) ψ~ω p,κ ◦ (∂k ◦ ∂j t=0,κ ◦ (∂k ◦ ϕλ+µ(k+1/p) , assertion (ii) follows. p,κ p,κ t=0,κ)c ◦ ϕλ+µ(k+1/p) w = δjkw w for w ∈ D. Since D is dense in Bs−r(k+1/p),λ+µ(k+1/p) p Suppose M is a compact m-dimensional submanifold of Rm. In this setting and if s = r ∈ 2N× assertions (i) and (ii) reduce to the trace theorems for anisotropic Sobolev spaces due to P. Grisvard [14]; also see O.A. (In Ladyzhenskaya, V.A. Solonnikov, and N.N. Ural'ceva [30] and R. Denk, M. Hieber, and J. Pruss [11]. the latter paper the authors consider vector-valued spaces.) The much simpler Hilbertian case p = 2 has been presented by J.-L. Lions and E. Magenes in [31, Chapter 4, Section 2] following the approach by P. Grisvard [15]. 19 Spaces With Vanishing Traces In this section we characterize Fs/~r,~ω characterize those subspaces of Fs/~r,~ω denote by p p p and Fs/~r,~ω ( J , V ) by the vanishing of certain traces. In fact, we need to (J, V ) whose traces vanish on Γ even if Γ 6= ∂M . More precisely, we (19.1) (J, V ). Fs/~r,~ω p,Γ = Fs/~r,~ω p,∂M = Fs/~r,~ω p,Γ (J, V ) the closure of D(cid:0) J,D(M \Γ, V )(cid:1) in Fs/~r,~ω . By Theorem 8.3(ii) we know already p p Note that Fs/~r,~ω and, trivially, Fs/~r,~ω (Γ, J) 6= (∅, R). p = Fs/~r,~ω p Fs/~r,~ω p = Fs/~r,~ω p , s < 1/p, (19.2) if ∂M = ∅ and J = R. The following theorem concerns the case s > 1/p and Theorem 19.1 (i) If Γ 6= ∅ and k + 1/p < s < k + 1 + 1/p with k ∈ N, then Fs/~r,~ω p,Γ = { u ∈ Fs/~r,~ω p ; ∂i nu = 0, i ≤ k }. (ii) Assume r(ℓ + 1/p) < s < r(ℓ + 1 + 1/p) with ℓ ∈ N. Then Fs/~r,~ω p (cid:0)(0,∞), V(cid:1) =(cid:8) u ∈ Fs/~r,~ω p (R+, V ) ; ∂j 79 (19.3) (19.4) (R+, V ). t=0u = 0, j ≤ ℓ(cid:9). (cid:0)(0,∞), V(cid:1) = Fs/~r,~ω p Suppose s < r/p with s > r(−1 + 1/p) if Γ 6= ∅. Then Fs/~r,~ω P r o o f. (1) Let the assumptions of (i) be satisfied. Since ∂i p n is continuous and vanishes on the dense subset p s/~r,~ω p,Γ it follows that the latter space is contained in the one on the right side of (19.3). D(cid:0) J,D(M \Γ)(cid:1) of F Conversely, let u ∈ Fs/~r,~ω satisfy ∂i borhood of Γ. Then v := αu ∈ Fs/~r,~ω γκ ◦ ∂i that ϕ~ω near ∂M \Γ and Fs/~r part (ii) of that theorem guarantees v = ψ~ω 1(ϕ~ω p,κv ∈ Fs/~r nu = 0 for i ≤ k. Suppose α ∈ D(cid:0) M ∪ Γ, [0, 1](cid:1) and α = 1 in a neigh- nv = 0 for i ≤ k. We infer from (18.7), (18.9), and (18.10) that p,κv) = 0 for i ≤ k and κ⋉ϕ ∈ K⋉Φ. Since γκ = 0 for κ /∈ KΓ it follows from [4, Theorem 4.7.1] p,κ as well. Moreover, v vanishes p ). Now p,κ for κ⋉ϕ ∈ KΓ⋉Φ. If κ ∈ K\KΓ, then ϕ~ω p,κ for κ ∈ K\K∂M . Hence we deduce from Theorem 9.3 that ϕ~ω p,κv belongs to Fs/~r p,κ = Fs/~r p,Γ . This proves claim (i). and ∂i p (ϕ~ω p p v) ∈ Fs/~r,~ω p . Consequently, u ∈ Fs/~r,~ω (2) Assume J = R+ and r(ℓ + 1/p) < s < r(ℓ + 1 + 1/p). As above, we see that Fs/~r,~ω p p v ∈ ℓp(Fs/~r (cid:0)(0,∞), V(cid:1) is con- tained in the space on the right side of (19.4). p t=0,κ(ϕ~ω p,κu) = 0 for (R+, V ) satisfy ∂j j ≤ ℓ and κ⋉ϕ ∈ K⋉Φ. t=0u = 0 for 0 ≤ j ≤ ℓ. We get from (18.16) that ∂j Let u ∈ Fs/~r,~ω Suppose κ ∈ K\K∂M . Then [4, Theorem 4.7.1] implies ϕ~ω p (cid:0)Xκ × (0,∞), E(cid:1). If κ ∈ K∂M , then p,κu first from Hm × R+ to Rm × R+ (as in Section 4.1 of [4]), we obtain the latter result by extending vκ := ϕ~ω then applying [4, Theorem 4.7.1], and restricting afterwards to Hm × R+. From this and Theorems 9.3 and 17.1 we obtain p u) ∈ ℓp(cid:0)Fs/~r(X × R, E)(cid:1). (19.5) Thus, using these theorems once more and the fact that, by (17.9), e+ p,κu ∈ Fs/~r 0 ◦ ψ~ω This implies the first part of claim (ii). 0 commutes with ψ~ω p u ∈ Fs/~r,~ω Assume s < r/p. If ∂M = ∅, then D(cid:0)(0,∞), M(cid:1) = D( J , M ). Hence Fs/~r,~ω Thus, by (19.2), Fs/~r,~ω s > r(−1 + 1/p). Consequently, as above, we deduce Fs/~r p (Xκ × J, E) = Fs/~r Now the second part of assertion (ii) is implied by (19.5) and (19.6). (cid:0)(0,∞), V(cid:1) = Fs/~r,~ω (cid:0)(0,∞), V(cid:1). (cid:0)(0,∞), V(cid:1) = Fs/~r,~ω (R+, V ). (R+, V ) for s < 1/p and ∂M = ∅. This shows that in either case p,κ from [4, Theorem 4.7.1(ii)]. p , we find 0 ◦ ψ~ω 0 ◦ ϕ~ω 0 ◦ e+ p ◦ ϕ~ω p ◦ e+ p u = r+ e+ 0 (ϕ~ω u = r+ (19.6) p p p p p 20 Boundary Operators Throughout this section we suppose Γ 6= ∅. For k ∈ N we consider differential operators on Γ of the form kXi=0 bi( • ∇) ◦ ∂i n, bi( • ∇) := k−iXj=0 bij · • ∇j, where bij · k − i. • ∇j := (u 7→ bij · • ∇ju), of course. Thus bi( • ∇) is a tangential differential operator of order at most 80 H. Amann: Anisotropic Function Spaces on Singular Manifolds In fact, we consider systems of such operators. Thus we assume • k, ri ∈ N with r0 < ··· < rk, • σi, τi ∈ N and λi ∈ R, • Gi = (Gi, hGi, DGi) is a fully uniformly regular vector bundle over Γ (20.1) for 0 ≤ i ≤ k. For abbreviation, 0 ≤ i ≤ k, Then we define boundary operators on Γ of order at most ri by νi := (ri, σi, τi, λi), νk := (ν0, . . . , νk). Bi(bi) := riXj=0 Bij(bij) ◦ ∂j n, Bij(bij ) := ri−jXℓ=0 bij,ℓ · • ∇ℓ, where bi := (bi0, . . . , biri ) and bij := (bij,0, . . . , bij,ri−j) with bij,ℓ being time-dependent Hom( tensor fields on Γ. To be more precise, we introduce data spaces for s > ri by • W , Gi)-valued Bs ij(Γ, Gi) = Bs ij (Γ, Gi, νi, µ) := with general point (bij), and ri−jYℓ=0 B(s−ri)/~r,(λi+ri−j,µ) ∞ (cid:0)R, T σi+τ +ℓ τi+σ (cid:0)Γ, Hom( • W , Gi)(cid:1)(cid:1) Bs i (Γ, Gi) = Bs i (Γ, Gi, νi, µ) := Bs ij (Γ, Gi) riYj=0 whose general point is bi. Remarks 20.1 (a) For the ease of writing we assume that these data spaces are defined on the whole line R. In the following treatment, when studying function spaces on R+ or (0,∞) it suffices, of course, to consider data defined on R+ only. It follows from Theorem 17.2 that this is no restriction of generality to assume that the data are given on all of R. (b) B(s−ri)/~r,(λi+ri−j,µ) ∞ ∞ i (Γ, Gi) is independent of p. Bs It should be observed that everything which follows below remains valid if we replace the data space with ¯s > s − ri − 1/p. The selected choice has the advantage that (cid:3) i (Γ, Gi), it follows from Theorem 16.2, by taking also Theo- by B ¯si/~r,(λi+ri−j,µ) Henceforth, I ∈(cid:8) J, (0,∞)(cid:9). Given bi ∈ Bs rem 19.1(ii) into consideration if I = (0,∞), that Bij(bij ) ∈ L(cid:0)B(s−j−1/p)/~r,(λ+j+1/p,µ) (I, Hence, by Theorem 18.1, p • V ), B(s−ri−1/p)/~r,(λ+λi+ri+1/p,µ) p τi (Γ, Gi)(cid:1)(cid:1). (cid:0)I, T σi τi (Γ, Gi)(cid:1)(cid:1). (20.2) (20.3) Bi(bi) ∈ L(cid:0)Fs/~r,~ω Finally, we set G := G0 ⊕ ··· ⊕ Gk, p (I, V ), B(s−ri−1/p)/~r,(λ+λi+ri+1/p,µ) p (cid:0)I, T σi Bs(Γ, G) = Bs(Γ, G, ν, µ) := Bs i (Γ, Gi) kYi=0 and B(b) :=(cid:0)B0(b0), . . . ,Bk(bk)(cid:1), The boundary operator Bi(bi) is normal if biri := biri,0 is λi-uniformly contraction surjective, and B(b) is nor- mal if each Bi(bi), 0 ≤ i ≤ k, has this property. Then b := (b0, . . . , bk) ∈ Bs(Γ, G). Bs norm(Γ, G) :=(cid:8) b ∈ Bs(Γ, G) ; B(b) is normal(cid:9). the order of the boundary operators may be different on different parts of ∂M . It should be observed that Γ 6= ∂M, in general. This will allow us to consider boundary value problems where Lastly, we introduce the 'boundary space' 81 p ∂Γ×I Fs/~r,~ω (G) = ∂Γ×I Fs/~r,~ω kYi=0 Lemma 20.2 If s > rk + 1/p and b ∈ Bs(Γ, G), then (G, ν, µ) := p B(s−ri−1/p)/~r,(λ+λi+ri+1/p,µ) p (cid:0)I, T σi τi (Γ, Gi)(cid:1). The following lemma shows that it is an image space for the boundary operators under consideration. B(b) ∈ L(cid:0)Fs/~r,~ω p (I, V ), ∂Γ×I Fs/~r,~ω p (G)(cid:1). norm(Γ, G) is open in Bs(Γ, G). The map B(·) =(cid:0)b 7→ B(b)(cid:1) is linear and continuous, and Bs P r o o f. The first assertion is immediate from (20.3). The second one is obvious, and the last one is a conse- quence of Proposition 14.8. retraction from Fs/~r,~ω (J, V ) onto ∂Γ×J Fs/~r,~ω Theorem 20.3 Suppose assumption (20.1) applies. Let s > rk + 1/p and b ∈ Bs (J, V )(cid:1) norm(Γ, G) → L(cid:0)∂Γ×J Fs/~r,~ω (G). There exists an analytic map Bc(·) : Bs (G), Fs/~r,~ω p p p p such that norm(Γ, G). Then B(b) is a P r o o f. (1) We deduce from Theorem 14.9 for 0 ≤ i ≤ k the existence of an analytic map Ac i (·) from If J = R+, then Bc(b)g ∈ Fs/~r,~ω p (i) Bc(b) is a coretraction for B(b), (ii) ∂j p n ◦ Bc(b) = 0 for 0 ≤ j < s − 1/p with j /∈ {r0, . . . , rk}. (cid:0)(0,∞), V(cid:1) whenever g ∈ ∂Γ×(0,∞)Fs/~r,~ω (G). W , Gi)(cid:1)(cid:1) (cid:0)J, T σi+τ τi+σ(cid:0)Γ, Hom( τi (Γ, Gi)(cid:1), B(s−ri−1/p)/~r,(λ+ri+1/p,µ) (cid:0)J, T σi norm(Γ, G). For 0 ≤ i ≤ k we set L(cid:0)B(s−ri−1/p)/~r,(λ+λi+ri+1/p,µ) i (a) is a right inverse for Ai(a) := (u 7→ a · u). B(s−ri)/~r,(λi,µ) ∞,surj p p • into such that Ac (2) Suppose b ∈ Bs (J, • V )(cid:1) Cri(bi) := − ri−1Xj=0 i (biri )Bij(bij) ◦ ∂j Ac n. It follows from (20.2), step (1), and Theorem 18.1 that Cri(bi) ∈ L(cid:0)Fs/~r,~ω p and the map bi → Cri(bi) is analytic. Let N := [s − 1/p]− and define (J, V ), B(s−j−1/p)/~r,(λ+ri+1/p,µ) p (J, • V )(cid:1) (20.4) C = (C0, . . . ,CN ) ∈ L(cid:16)Fs/~r,~ω p (J, V ), NYℓ=0 B(s−ℓ−1/p)/~r,(λ+ℓ+1/p,µ) p (J, • V )(cid:17) by setting Cℓ := 0 for 0 ≤ ℓ ≤ N with ℓ /∈ {r0, . . . , rk}. 82 H. Amann: Anisotropic Function Spaces on Singular Manifolds (3) Assume g = (g0, . . . , gk) ∈ ∂Γ×J F s/~r p (G). Define h = (h0, . . . , hN ) ∈ NYℓ=0 B(s−ℓ−1/p)/~r,(λ+ℓ+1/p,µ) p • (J, V ) by hri := Ac i (biri )gi for 0 ≤ i ≤ k, and hℓ := 0 otherwise. By Theorem 18.1 there exists for j ∈ {0, . . . , N} a universal coretraction (γj n)c for ∂j n satisfying We put u0 := (γ0 Set n)ch0 ∈ F ∂ℓ n ◦ (γj n)c = δℓjid, 0 ≤ ℓ ≤ j. (20.5) s/~r,~ω p (J, V ). Suppose 1 ≤ j ≤ N and u0, u1, . . . , uj−1 have already been defined. uj := uj−1 + (γj (20.6) nuj−1). n)c (hj + Cjuj−1 − ∂j This defines u0, u1, . . . , uN ∈ F s/~r,~ω p (J, V ). It follows from (20.5) and (20.6) ∂j nuj = hj + Cjuj−1, 0 ≤ j ≤ N, (20.7) and The latter relation implies ∂ℓ nuj = ∂ℓ nuj−1, 0 ≤ ℓ ≤ j − 1, 1 ≤ j ≤ N. Hence, since Cj involves ∂0 n, . . . , ∂j−1 n only, we deduce from (20.7) ∂ℓ nuj = ∂ℓ nun, 0 ≤ ℓ < j < n ≤ N. ∂j nun = hj + Cjun, If j = ri, then we apply Ai(biri ) to this equation to find 0 ≤ j ≤ n ≤ N. For 0 ≤ i ≤ k we set Gi := G0 ⊕ ··· ⊕ Gi and ν i := (ν1, . . . , νi) as well as bi := (b0, . . . , bi). Then it follows from (20.3) that Biun = gi, ri ≤ n ≤ N. (20.8) Bi(bi) :=(cid:0)B0(b0), . . . ,Bi(bi)(cid:1) ∈ L(cid:0)Ft/~r,~ω p for ri + 1/p < t ≤ s. We define Bic(bi) by (J, V ), ∂Γ×J Ft/~r,~ω p (Gi, νi, µ)(cid:1) (20.9) It follows from (20.4), (20.6), and Theorem 18.1 that Bic(bi) (g0, . . . , gi) := uri. p Furthermore, (20.8) and the definition of h imply (Gi, ν i, µ), Ft/~r,~ω Bic(bi) ∈ L(cid:0)∂Γ×J Ft/~r,~ω (J, V )(cid:1), Bα(bα)Bic(bi)(g0, . . . , gi) = Bα(bα)Bjc(bj)(g0, . . . , gj) = gα, nBic(bi) = 0 for 0 ≤ j < t − 1/p with j /∈ {r0, . . . , rk}. and ∂j p ri + 1/p < t ≤ s. (20.10) 0 ≤ α ≤ i ≤ j ≤ k, (20.11) Now we set Bc(b) := Bkc(b). Then (20.9) and (20.11) show that it is a right inverse for B(b). It is a conse- quence of step (2) and (20.6) that Bc(·) is analytic. Due to Theorem 18.1 it is easy to see that the last assertion applies as well. There is a similar, though much simpler result concerning the 'extension of initial conditions'. Theorem 20.4 Suppose 0 ≤ j0 < ··· < jℓ and s > r(jℓ + 1/p). Set C := (∂j0 t=0, . . . , ∂jℓ t=0) and Bs−r(jℓ+1/p),λ+µ(jℓ+1/p) p (V ) := ℓYi=0 Bs−r(ji+1/p),λ+µ(ji+1/p) p (V ). (20.12) 83 Then C is a retraction from Fs/~r,~ω satisfying ∂j p t=0 ◦ Cc = 0 for 0 ≤ j < s/r − 1/p with j /∈ {j0, . . . , jℓ}. (R+, V ) onto Bs−~r(jℓ+1/p),λ+µ(jℓ+1/p) p (V ), and there exists a coretraction Cc P r o o f. Theorem 18.1(ii) guarantees that C is a continuous linear map from Fs/~r,~ω (R+, V ) into (20.12). Due to that theorem the assertion follows from step (3) of the proof of Theorem 20.3 using the following modifications: hji := gi for 0 ≤ i ≤ ℓ and p uj := uj−1 + (γj t=0)c(hj − ∂j t=0uj−1) with u−1 := 0. Now we suppose Γ 6= ∅ and J = R+. We write Σ := Γ × R+ for the lateral boundary over Γ and recall that M0 := M × {0} is the initial boundary. Then Σ ∩ M0 = Γ × {0} =: Γ0 is the corner manifold over Γ. We suppose • • t=0, . . . , ∂ℓ assumption (20.1) is satisfied, ℓ ∈ N and s > max(cid:8)rk + 1/p, r(ℓ + 1/p)(cid:9). t=0). Then, by Theorem 20.4, C is a retraction from Fs/~r,~ω p (20.13) (R+, V ) onto We set C := −−→∂ℓ t=0 := (∂0 Bs−r(ℓ+1/p),λ+µ(ℓ+1/p) p (V ) := Bs−r(j+1/p),λ+µ(j+1/p) p (V ). ℓYj=0 norm(Γ, G) a retraction from Fs/~r,~ω p (R+, V ) onto ∂ΣFs/~r,~ω p (G). We put (G) := ∂ΣFs/~r,~ω p (G) × Bs−r(ℓ+1/p),λ+µ(ℓ+1/p) p (V ) By Theorem 20.3 B(b) is for b ∈ Bs ∂Σ∪M0 Fs/~r,~ω p and ~B(·) :=(cid:0)B(·),C(cid:1). Then ~B(·) : Bs norm(Γ, G) → L(cid:0)Fs/~r,~ω However, ~B(b) is not surjective, in general. Indeed, suppose is the restriction of a continuous linear map to the open subset Bs p (R+, V ), ∂Σ∪M0 Fs/~r,~ω p (G)(cid:1) 0 ≤ j ≤ ℓ, Then we deduce from (20.3) and Theorem 18.1(ii) 0 ≤ i ≤ k, s > ri + 1/p + r(j + 1/p) =: rij . norm(Γ, G) of Bs(Γ, G), hence analytic. ∂j t=0 ◦ Bi(b) ∈ L(cid:0)Fs/~r,~ω p (R+, V ), Bs−rij ,λ+λi+rij p τi (Γ0, Gi)(cid:1)(cid:1). (cid:0)T σi Furthermore, ∂j(cid:0)Bi(b)u(cid:1) = B(j) i (b)u, where B(j) i (b)u = jXα=0(cid:16) j α(cid:17)Bi(∂j−αbi) ◦ ∂α. Theorem 16.1 implies ∂j−αbi ∈ Bs i(cid:0)Γ, Gi,(cid:0)ri + r(j − α), σi, τi, λi + µ(j − α)(cid:1), µ(cid:1). 84 H. Amann: Anisotropic Function Spaces on Singular Manifolds From this and (20.3) we infer that B(j) i (b) possesses the same mapping properties as ∂j ◦ Bi(b). Set B(j) i (0)~vj = B(j) i α(cid:17)Bi(∂j−α (b, 0)~vj := jXα=0(cid:16) j (V ). Then B(j) i with vα ∈ Bs−r(α+1/p),λ+µ(α+1/p) p (0) is a continuous linear map t=0 bi)vα, ~vj := (v0, . . . , vj) jYα=0 Bs−r(α+1/p),λ+µ(α+1/p) p (V ) → Bs−rij ,λ+λj +rij p τi (Γ0, Gi)(cid:1) (cid:0)T σi and b 7→ B(j) i (0) is the restriction of a linear and continuous map to Bs norm(Γ, G). Furthermore, ∂j t=0(cid:0)Bi(b)u(cid:1) = B(j) i (0)−−→∂j t=0u, u ∈ Fs/~r,~ω p (R+, V ). (20.14) We denote for b ∈ Bs norm(Γ, G) by ∂cc ~B(b) Fs/~r,~ω p (G) the set of all (g, h) ∈ ∂Σ∪M0 Fs/~r,~ω p (G) satisfying the compatibility conditions for 0 ≤ i ≤ k and 0 ≤ j ≤ ℓ with ri + 1/p + r(j + 1/p) < s. t=0gi = B(j) ∂j i (0)~hj The linearity and continuity of ∂j ∂Σ∪M0 Fs/~r,~ω that, in fact, ∂cc Fs/~r,~ω (G) is a closed linear subspace of (G). By the preceding considerations it contains the range of ~B(b). The following theorem shows (0) guarantee that ∂cc t=0 and B(j) Fs/~r,~ω ~B(b) p p i norm(Γ, G). (G) = im( ~B), provided b ∈ Bs ~B(b) p Theorem 20.5 Let assumption (20.13) be satisfied and suppose s /∈(cid:8) ri + 1/p + r(j + 1/p) ; 0 ≤ i ≤ k, 0 ≤ j ≤ ℓ(cid:9). norm(Γ, G) a retraction from Fs/~r,~ω p (R+, V ) onto ∂cc ~B(b) Fs/~r,~ω p (G). There exists an ana- Then ~B(b) is for b ∈ Bs lytic map ~Bc(·) : Bs Fs/~r,~ω p norm(Γ, G) → L(cid:0)∂Σ∪M0 Fs/~r,~ω p (G), Fs/~r,~ω p (R+, V )(cid:1) (20.15) such that ~Bc(b)∂cc ~B(b) (G) is a coretraction for ~B(b). P r o o f. By the preceding remarks it suffices to construct ~Bc(·) satisfying (20.15) such that its restriction to Fs/~r,~ω ∂cc ~B(b) p (G) is a right inverse for ~B(b). By Theorem 20.3 there exists an analytic map Bc(·) : Bs norm(Γ, G) → L(cid:0)∂ΣFs/~r,~ω p (G), Fs/~r,~ω p v ∈ ∂Γ×(0,∞)Fs/~r,~ω p (G) =⇒ Bc(b)v ∈ Fs/~r,~ω p (R+, V )(cid:1) (cid:0)(0,∞), V(cid:1) such that for b ∈ Bs norm(Γ, G). (20.16) Let Cc be a coretraction for C. Its existence is guaranteed by Theorem 20.4. Given (g, h) ∈ ∂Σ∪M0 F and b ∈ Bs norm(Γ, G), set Then Bc(·) satisfies (20.15) and is analytic. Furthermore, ~Bc(b)(g, h) := Cch + Bc(b)(cid:0)g − B(b)Cch(cid:1). B(b)(cid:0) ~Bc(b)(g, h)(cid:1) = g. 85 s/~r,~ω p (G) (20.17) We fix b ∈ Bs norm(Γ, G) and write B = B(b) and Bc = Bc(b). For (g, h) ∈ ∂cc ~B(b) Fs/~r,~ω p (G) we set Suppose 0 ≤ i ≤ k. Then v := g − BCch ∈ ∂ΣFs/~r,~ω p (G). vi = gi − BiCch ∈ B(s−ri−1/p)/~r,(λ+λi+ri+1/p,µ) p (R+, Vi), where Vi := T σi (20.14), τi (Gi). Let ji ∈ {0, . . . , ℓ} be the largest integer satisfying ri + 1/p + r(ji + 1/p) < s. Then, by t=0vi = ∂j ∂j t=0gi − ∂j t=0(BiCch) = ∂j t=0gi − B(j) i (0)−−→∂j t=0Cch = ∂j t=0gi − B(j) i (0)~hj = 0 for 0 ≤ j ≤ ji. Hence r(ji + 1/p) < s − ri − 1/p < r(ji + 1 + 1/p) and Theorem 19.1(ii) imply vi ∈ B(s−ri−1/p)/~r,(λ+λi+ri+1/p,µ) p (cid:0)(0,∞), Vi(cid:1). (20.18) If there is no such ji, then s − ri − 1/p < r/p. In this case that theorem guarantees (20.18) also. This shows that v ∈ ∂Γ×(0,∞)Fs/~r,~ω (G). Hence CBcv = 0 by (20.16) and Theorem 19.1(ii) and since s > r(j + 1/p) for 0 ≤ j ≤ ℓ. Consequently, C(cid:0) ~Bc(b)(g, h)(cid:1) = h for (g, h) ∈ ∂cc (G). Together with (20.17) this proves the theorem. Fs/~r,~ω ~B(b) p p Remark 20.6 Let assumption (20.1) be satisfied. Suppose r0 + 1/p < s < r/p, s /∈ { ri + 1/p ; 1 ≤ i ≤ k }. Then there is a lateral boundary operator B only, since there is no initial trace. Thus this case is covered by Theorem 20.3. Assume r/p < s < r0 + 1/p, s /∈(cid:8) r(j + 1/p) ; 1 ≤ j ≤ ℓ(cid:9). Then there is no lateral trace operator and we are in a situation to which Theorem 20.4 applies. Lastly, if −1 + 1/p < s < min{r0 + 1/p, r/p}, then there is neither a lateral nor an initial trace operator and (cid:3) s/~r,~ω p s/~r,~ω (R+, V ) = F p F (R+, V ). The theorems on the 'extension of boundary values' proved in this section are of great importance in the theory of nonhomogeneous time-dependent boundary value problems. The only results of this type available in the literature concern the case where M is an m-dimensional compact submanifold of Rm. In this situation an anisotropic extension theorem involving compatibility conditions has been proved by P. Grisvard in [14] for the case where s ∈ rN×, and in [15] if p = 2 (also see J.-L. Lions and E. Magenes [31, Chapter 4, Section 2] for the Hilbertian case) by means of functional analytical techniques. If s = r = 2, then corresponding results are derived in O.A. Ladyzhenskaya, V.A. Solonnikov, and N.N. Ural'ceva [30] by studying heat potentials. In contrast to our work, in all these publications the exceptional values ri + 1/p + r(j + 1/p) for s are considered also. 86 H. Amann: Anisotropic Function Spaces on Singular Manifolds 21 Interpolation In Section 8 the anisotropic spaces Fs/~r,~ω have been defined for s > 0 by interpolating between anisotropic Sobolev spaces. From this we could derive some interpolation properties by means of reiteration theorems. How- ever, such results would be restricted to spaces with one and the same value of λ. In this section we prove general interpolation theorems for anisotropic Bessel potential, Besov, and Besov-Holder spaces involving different val- ues of s and λ. p Reminding that ξθ = (1 − θ)ξ0 + θξ1 for ξ0, ξ1 ∈ R and 0 ≤ θ ≤ 1, we set ~ωθ := (λθ, µ) for λ0, λ1 ∈ R. We also recall that (·,·)θ = [·, ·]θ if F = H, and (·,·)θ = (·,·)θ,p if F = B. ]θ (21.1) (21.2) Theorem 21.1 Suppose −∞ < s0 < s1 < ∞, λ0, λ1 ∈ R, and 0 < θ < 1. (i) Assume s0 > −1 + 1/p if ∂M 6= ∅. Then , Fs1/~r,~ω1 )θ . = [Fs0/~r,~ω0 . = Fsθ /~r,~ωθ (Fs0/~r,~ω0 p p p p , Fs1/~r,~ω1 p and (ii) If s0 > 0, then and (H s0/~r,~ω0 p , H s1/~r,~ω1 p )θ,p . = Bsθ /~r,~ωθ p . [Bs0/~r,~ω0 ∞ , Bs1/~r,~ω1 ∞ ]θ . = Bsθ /~r,~ωθ ∞ [bs0/~r,~ω0 ∞ , bs1/~r,~ω1 ∞ ]θ . = bsθ /~r,~ωθ ∞ . P r o o f. (1) Let X be a Banach space and δ > 0. Then δX := (X,k·kδX ), where kxkδX := kδ−1xkX for x ∈ X. Thus δX is the image space of X under the map x 7→ δx so that this function is an isometric isomorphism from X onto δX. Assume Xβ is a Banach space and δβ > 0 for each β in a countable index set B. Then we set δX :=Qβ δβXβ and δx := (δβxβ). Hence δ := (x 7→ δx) ∈ Lis(X, δX). for j = 0, 1, that is, (X0, X1) is an interpolation couple. Suppose {·,·}θ ∈(cid:8)[·, ·]θ, (·,·)θ,p(cid:9). Then interpolation Let (X0, X1) be a pair of Banach spaces such that Xj is continuously injected in some locally convex space theory guarantees {δ0X0, δ1X1}θ = δ1−θ 0 δθ 1{X0, X1}θ, δ0, δ1 > 0, (21.3) (e.g., [50, formula (7) in Section 3.4.1]). (2) Let J = R and s > −1 + 1/p if ∂M 6= ∅. Put ξ := λ − λ0. Then ϕ~ω imply, due to Theorem 9.3, that the diagram p,κ = ρξ κϕ~ω0 p,κ and ψ~ω p,κ = ρ−ξ κ ψ~ω0 p,κ Fs/~r,~ω p id ✲✲ Fs/~r,~ω p ❍❍❍❍❥ ❆ ❆ ϕ~ω p ❆ ❆ ϕ~ω0 p ❆ ❆ ❆❯ ℓp(Fs/~r p ) ✻ ∼= ρξ ℓp(ρ−ξFs/~r p ) ✟✟✟✟✯ ✁✕ ✁ ψ~ω p ✁ ✁ ψ~ω0 p ✁ ✁ ✁ is commuting. Hence ψ~ω0 ) onto Fs/~r,~ω (3) Let ξ := λ1 − λ0. By Theorem 9.3 and the preceding step each of the maps p is for each s a retraction from ℓp(ρ−ξFs/~r p p , and ϕ~ω0 p is a coretraction. ψ~ω0 p : ℓp(Fs0/~r p ) → Fs0/~r,~ω0 p , ψ~ω0 p : ℓp(ρ−ξFs1/~r p ) → Fs1/~r,~ω1 p is a retraction, and ϕ~ω0 p is a coretraction. Thus, by interpolation, is a retraction, and ϕ~ω0 p is a coretraction. From [50, Theorem 1.18.1] and (21.3) we infer ψ~ω0 p p : (cid:8)ℓp(Fs0/~r (cid:8)ℓp(Fs0/~r p ), ℓp(ρ−ξFs1/~r p , Fs1/~r,~ω1 p p )(cid:9)θ →(cid:8)Fs0/~r,~ω0 = ℓp(cid:0)ρ−θξ{Fs0/~r p . )(cid:9)θ (cid:9)θ }θ(cid:1). ), ℓp(ρ−ξFs1/~r p , Fs1/~r p 87 (21.4) (21.5) (Recall the definition of (E, F )θ after (9.1).) Suppose ∂M = ∅. Then [4, formulas (3.3.12) and (3.4.1) and Theorem 3.7.1(iv)] imply (Fs0/~r p,κ , Fs1/~r p,κ )θ . = Fsθ /~r p,κ , (H s0/~r p,κ , H s1/~r p,κ )θ,p . = Bsθ /~r p,κ . = [Bs0/~r p,κ , Bs1/~r p,κ ]θ. (21.6) This is due to the fact that, on account of [4, Corollary 3.3.4 and Theorem 3.7.1(i)], the definition of Fs/~r s < 0 used in that publication coincides with the definition by duality employed in this paper. p,κ for holds in this case as well. If ∂M 6= ∅, then it follows from s > −1 + 1/p and Theorem 4.7.1(ii) of [4] by the same arguments that (21.6) Thus, in either case, due to (21.4) -- (21.6) ψ~ω0 p is a retraction from ℓp(ρ−θξFsθ /~r ) onto (Fs0/~r,~ω0 , Fs1/~r,~ω1 p s0/~r,~ω0 p , F and onto [F ξ = θ(λ1 − λ0), that ψ~ω0 implies the validity of (21.1) if J = R. The proof for (21.2) is similar. ]θ, and ϕ~ω0 p is a retraction from ℓp(ρ−θξFsθ /~r p p s1/~r,~ω1 p )θ is a coretraction. On the other hand, we infer from step (2), setting is a coretraction. This ) onto Fsθ /~r,~ωθ p p , and ϕ~ω0 p p (4) Assume J = R+. In this case we get assertion (i) by Theorem 17.1(i) in conjunction with what has just been proved. onto Bs0/~r,~ω0 ∞ and from ℓ∞(ρ−ξBs1/~r (5) Set ξ = λ1 − λ0. Then as above, we infer from Theorem 12.8 that ψ~ω0 ∞ )(cid:3)θ → [Bs0/~r,~ω0 ∞ : (cid:2)ℓ∞(Bs0/~r ∞ ), ℓ∞(ρ−ξBs1/~r ∞ ) onto Bs1/~r,~ω0 , and ϕ~ω0 ψ~ω0 ∞ ∞ is a retraction, and ϕ~ω0 ∞ is a coretraction. ∞ is a coretraction. Hence ∞ is a retraction from ℓ∞(Bs0/~r ∞ ) , Bs1/~r,~ω1 ∞ ]θ (21.7) We use the notation of Sections 11 and 12. Then, setting Bs/~r it is not difficult to verify (cf. Lemma 11.12) that ∞ (M × J, ρ−ξE) :=Qκ Bs/~r ∞ (M × J, ρ−ξ κ E), ∞ (21.8) , Bs1/~r,~ω1 ∞ = ψ~ω0 ∞ is a coretraction. is a retraction, and Φ~ω0 for s > 0. Hence we deduce from (21.7) that the map Ψ~ω0 ef ∈ Lis(cid:0)Bs/~r ∞ (cid:0)M × J, ℓ∞(E)(cid:1), Bs/~r (cid:2)Bs/~r (6) Set K0 := { κ ∈ K ; ρκ ≤ 1 } and K1 := K\K0. Let Xκ be a Banach space for κ ∈ K and set X :=Qκ Xκ and Xj :=Qκ∈Kj ∞ (cid:0)M × J, ℓ∞(ρ−ξE)(cid:1), ℓ∞(ρ−ξBs/~r ∞ )(cid:1) ∞ ◦ef ∞ (cid:0)M × J, ℓ∞(ρ−ξBs/~r ∞ )(cid:1)(cid:3)θ → [Bs0/~r,~ω0 ∞(X) := ℓ∞(Xj) for j = 0, 1. Then ℓ∞(X) ∞ (cid:0)M × J, ℓ∞(ρ−ηE)(cid:1) . ∞ (cid:0)M × J, ℓ1 ∞(E)(cid:1). It follows from ρκ ≤ 1 for κ ∈ K0 that Y1 ֒→ Y0. Define a linear operator A0 in Y0 with domain Y1 by A0u = ρ−ξu. Then A0 is closed, −A0 con- tains the sector Sπ/4 in its resolvent set and satisfies k(λ + A0)−1kL(Y0) ≤ c/λ for λ ∈ Sπ/4. Furthermore, ∞(ρ−ηE)(cid:1) ⊕ Bs/~r ∞ (cid:0)M × J, ℓ0 ∞ (cid:0)M × J, ℓ0 ∞(ρ−ξE)(cid:1) and Y1 := Bs/~r k(−A0)zkL(Y0) ≤ sup ∞ (cid:0)M × J, ℓ0 ∞(ρ−ηE)(cid:1) ≤ 1, Re z ≤ 0. for η ∈ {0, ξ}. (7) Put Y0 := Bs/~r ∞(X) ⊕ ℓ1 ∞(X). Conse- Xκ as well as ℓj ρ−ξ Re z κ = Bs/~r quently, . = ℓ0 (21.10) (21.9) Bs/~r ]θ ∞ κ∈K0 88 H. Amann: Anisotropic Function Spaces on Singular Manifolds Hence Seeley's theorem, alluded to in the proof of Theorem 11.1, implies ∞ (cid:0)M × J, ℓ0 (cid:2)Bs/~r ∞(E)(cid:1), Bs/~r ∞ (cid:0)M × J, ℓ0 ∞(ρ−ξE)(cid:1)(cid:3)θ = [Y0, Y1]1−θ . = Bs/~r due to the fact that the space on the right side equals, except for equivalent norms, dom(A1−θ graph norm. ∞ (cid:0)M × J, ℓ0 ∞(ρ−ξE)(cid:1). Then ρκ > 1 for κ ∈ K1 implies ∞(E)(cid:1) and Z1 := Bs/~r Z1 ֒→ Z0. Define a linear map A1 in Z0 with domain Z1 by A1u := ρξu. Then A1 is closed and satisfies k(λ + A1)−1kL(Z0) ≤ c/λ for λ ∈ Sπ/4 as well as k(−A1)zkL(Z0) ≤ sup ∞ (cid:0)M × J, ℓ1 ∞ (cid:0)M × J, ℓ1 ∞(ρ−θξE)(cid:1), (8) Set Z0 := Bs/~r ) equipped with the Re z ≤ 0. ρξ Re z κ ≤ 1, (21.11) 0 κ∈K0 Thus, using Seeley's theorem once more, Now we deduce from (21.10) -- (21.12) that ∞ (cid:0)M × J, ℓ1 ∞(E)(cid:1), Bs/~r (cid:2)Bs/~r (cid:2)Bs/~r ∞ (cid:0)M × J, ℓ∞(E)(cid:1), Bs/~r ∞ (cid:0)M × J, ℓ1 ∞(ρ−ξE)(cid:1)(cid:3)θ ∞ (cid:0)M × J, ℓ∞(ρ−ξE)(cid:1)(cid:3)θ ∞ (cid:0)M × J, ℓ1 ∞(ρ−θξE)(cid:1). ∞ (cid:0)M × J, ℓ∞(ρ−θξE)(cid:1). . = Bs/~r . = Bs/~r Thus (21.9) shows that (21.12) is a retraction, and Φ~ω ∞ is a coretraction. Hence (12.23) and (21.8) imply that Ψ~ω ∞ : Bs/~r ∞ (cid:0)M × J, ℓ∞(ρ−θξE)(cid:1) → [Bs0/~r,~ω0 ∞ , Bs1/~r,~ω1 ∞ ]θ ∞ : ℓ∞(ρ−θξBs/~r ψ~ω ∞ ) → [Bs0/~r ∞ , Bs1/~r ∞ ]θ is a retraction, and ϕ~ω the first part of the second statement is true. ∞ is a coretraction. From this and the observation at the beginning of step (5) we derive that (9) By replacing ℓ∞ in the preceding considerations by ℓ∞,unif and invoking Theorem 12.10 instead of Theo- rem 12.8 we see that the second part of claim (ii) is also true. For completeness and complementing the results of [5] we include the following interpolation theorem for isotropic Besov-Holder spaces. Remark 21.2 Suppose 0 < s0 < s1, λ0, λ1 ∈ R, and 0 < θ < 1. Then ∞ ]θ . = Bsθ ,λθ ∞ , ∞ , Bs1,λ1 ∞ , bs1,λ1 [Bs0,λ0 ∞ ]θ [bs0,λ0 . = bsθ ,λθ ∞ . P r o o f. This follows from the above proof by relying on the corresponding isotropic results of Sections 11 and 12. Throughout the rest of this section we suppose • Γ 6= ∅. assumption (20.1) is satisfied. • ¯s > rk + 1/p and b ∈ B¯s • • B = (B0, . . . ,Bk) := B(b). norm. (21.13) Let I ∈(cid:8)J, (0,∞)(cid:9). For −1 + 1/p < s ≤ ¯s with s /∈ { ki + 1/p ; 0 ≤ i ≤ k } we set (I) = Fs/~r,~ω p (I, V ) ; Biu = 0 for ri < s − 1/p(cid:9). Fs/~r,~ω p,B (I) :=(cid:8) u ∈ Fs/~r,~ω (I) if s < k0 + 1/p. p Thus Fs/~r,~ω p,B (I) = Fs/~r,~ω p Suppose b is independent of t. Then we can define stationary isotropic spaces with vanishing boundary conditions analogously, that is, 89 Fs,λ p,B :=(cid:8) u ∈ Fs,λ p (V ) ; Biu = 0 for ri < s − 1/p(cid:9). Theorem 21.3 Let (21.13) be satisfied. Suppose −1 + 1/p < s0 < s1 ≤ ¯s and 0 < θ < 1 satisfy s0, s1, sθ /∈ { ri + 1/p ; 0 ≤ i ≤ k } and λ0, λ1 ∈ R. Then and . = Fsθ /~r,~ωθ p,B (I) (I), Fs1/~r,~ω1 p,B . =(cid:2)Fs0/~r,~ω0 p,B (I)(cid:3)θ (21.14) p,B (cid:0)Fs0/~r,~ω0 (I), Fs1/~r,~ω1 p,B (I)(cid:1)θ (cid:0)H s0/~r,~ω0 p,B (I), H s1/~r,~ω1 p,B . = Bsθ /~r,~ωθ p,B (I). (I)(cid:1)θ,p (Fs0,λ0 p,B , Fs1,λ1 p,B )θ . = Fsθ ,λθ p,B . = [Fs0,λ0 p,B , Fs1,λ1 p,B ]θ If b is independent of t ∈ R, then and (H s0,λ0 p,B , H s1,λ1 p,B )θ,p . = Bsθ ,λθ p,B . P r o o f. Theorem 20.3 guarantees the existence of a coretraction Bc ∈ L(cid:0)∂Γ×I Fs/~r,~ω (I)(cid:1) is a projection. Note that BcB depends on b and the universal extension opera- Hence BcB ∈ L(cid:0)Fs/~r,~ω tors (20.5) only. Thus we do not need to indicate the parameters s, λ, and p with rk + 1/p < s ≤ ¯s which characterize the domain Fs/~r,~ω rk + 1/p < s ≤ ¯s. (I)(cid:1), (G), Fs/~r,~ω p p p (I). p Taking this into account and using the notation of the proof of Theorem 20.3 we set Xℓ := Fsℓ/~r,~ωℓ p (I) for ℓ ∈ {0, 1, θ} and, putting rk+1 := ∞, Pℓ :=( idℓ, idℓ − BicBi, sℓ < r0 + 1/p, ri + 1/p < sℓ < ri+1 + 1/p. (21.15) Theorem 21.1 guarantees (X0, X1)θ Since Xℓ ֒→ D( J, D)′ the sum space X0 + X1 is well-defined, that is, (X0, X1) is an interpolation couple. It follows from (20.9) -- (20.11) that P0 ∈ L(X0 + X1) and Pℓ ∈ L(Xℓ) with P0Xℓ = Pℓ for ℓ ∈ {0, 1, θ}. . = Xθ. Theorem 20.3, definition (21.15), and [2, Lemma I.2.3.1] (also see [4, Lemma 4.1.5]) imply that Pℓ is a projection onto Xℓ,B := F (I) for ℓ ∈ {0, 1, θ}. Thus it is a retraction from Xℓ onto Xℓ,B possessing the natural injection Xℓ,B ֒→ Xℓ as a coretraction. Consequently, Pθ is . = Xθ,B. This proves a retraction from Xθ the first equivalence of (21.14). The remaining statements for the anisotropic case follow analogously. . = (X0, X1)θ onto (X0,B, X1,B)θ. From this we get (X0,B, X1,B)θ sℓ/~r,~ωℓ p,B Due to the observation at the end of Section 14 it is clear that the above proof applies to the isotropic case as well. There is a similar result concerning interpolations of spaces with vanishing initial conditions. For this we ℓ, j0, . . . , jℓ ∈ N with j0 < j1 < ··· < jℓ. t=0, . . . , ∂jℓ t=0). (21.16) assume Then, given s > −1 + 1/p, we put • • C := (∂j0 (R+) :=(cid:8) u ∈ Fs/~r,~ω p Fs/~r,~ω p,C (R+) ; ∂ji t=0u = 0 if r(ji + 1/p) < s(cid:9). 90 H. Amann: Anisotropic Function Spaces on Singular Manifolds Theorem 21.4 Let (21.16) be satisfied. Suppose −1 + 1/p < s0 < s1 and θ ∈ (0, 1) satisfy and λ0, λ1 ∈ R. Then p,C (cid:0)Fs0/~r,~ω0 (R+), Fs1/~r,~ω1 p,C and s0, s1, sθ /∈(cid:8) r(ji + 1/p) ; 0 ≤ i ≤ ℓ(cid:9) =(cid:2)Fs0/~r,~ω0 (R+)(cid:1)θ . = Fsθ /~r,~ωθ (R+) p,C p,C . (R+), Fs1/~r,~ω1 p,C (R+)(cid:3)θ P r o o f. This is shown by the preceding proof using Theorem 20.4 instead of Theorem 20.3. (cid:0)H s0/~r,~ω0 p,C (R+), H s1/~r,~ω1 p,C . = Bsθ /~r,~ωθ p,C (R+). (R+)(cid:1)θ,p Now we suppose, in addition to (21.13), that ℓ ∈ N and ¯s > r(ℓ + 1/p). Then we set Fs/~r,~ω p, ~B :=(cid:8) u ∈ Fs/~r,~ω p (R+) ; Biu = 0, ∂j t=0u = 0 if ri + 1/p + r(j + 1/p) < s(cid:9) if max{r0 + 1/p, r/p} < s ≤ ¯s and s /∈(cid:8) ri + 1/p + r(j + 1/p) ; 0 ≤ i ≤ k, 0 ≤ j ≤ ℓ(cid:9), Fs/~r,~ω p, ~B := Fs/~r,~ω p,B (R+) if r0 + 1/p < s < r/p and s /∈ { ri + 1/p ; 1 ≤ i ≤ k }, Fs/~r,~ω p, ~B := Fs/~r,~ω −−→ ∂ℓ t=0 p, (R+) if r/p < s < r0 + 1/p with s /∈(cid:8) r(j + 1/p) ; 1 ≤ j ≤ ℓ(cid:9), and Fs/~r,~ω p, ~B := Fs/~r,~ω (R+) p if −1 + 1/p < s < max{r0 + 1/p, r/p}. function spaces with vanishing boundary and initial conditions. The following theorem is analogue to Theorem 21.3. It describes the interpolation behavior of anisotropic Theorem 21.5 Let assumption (21.13) be satisfied. Also assume ℓ ∈ N and r(ℓ + 1/p) < ¯s. Suppose −1 + 1/p < s0 < s1 ≤ ¯s and 0 < θ < 1 satisfy s0, s1, sθ /∈ { ri + 1/p + r(j + 1/p), ri + 1/p, r(j + 1/p) ; 0 ≤ i ≤ k, 0 ≤ j ≤ ℓ } and λ0, λ1 ∈ R. Then and (Fs0/~r,~ω0 p, ~B , Fs1/~r,~ω1 p, ~B )θ . = Fsθ /~r,~ωθ p, ~B . = [Fs0/~r,~ω0 p, ~B , Fs1/~r,~ω1 p, ~B ]θ (H s0/~r,~ω0 p, ~B , H s1/~r,~ω1 p, ~B )θ,p . = Bsθ /~r,~ωθ p, ~B . P r o o f. This follows by the arguments of the proof of Theorem 21.3 by invoking Theorem 20.5 and Re- mark 20.6. The preceding interpolation theorems combined with the characterization statements of Section 19 lead to interpolation results for spaces with vanishing traces. For abbreviation, Fs/~r,~ω p,Γ (I) = Fs/~r,~ω p,Γ (I, V ), etc. Theorem 21.6 Suppose −1 + 1/p < s0 < s1 < ∞, 0 < θ < 1, and λ0, λ1 ∈ R. (i) If s0, s1, sθ /∈ N + 1/p, then (J),Fs1/~r,~ω1 p,Γ . = Fsθ /~r,~ωθ p,Γ (J) (J),Fs1/~r,~ω1 p,Γ (cid:0)Fs0/~r,~ω0 p,Γ (J)(cid:1)θ . =(cid:2)Fs0/~r,~ω0 p,Γ (J)(cid:3)θ 91 . = Bsθ /~r,~ωθ p,Γ (J). (J)(cid:1)θ,p (0,∞) . =(cid:2)Fs0/~r,~ω0 p (0,∞), Fs1/~r,~ω1 p (0,∞)(cid:3)θ (0,∞)(cid:1)θ,p . = Bsθ /~r,~ωθ p (0,∞). and (J), H s1/~r,~ω1 p,Γ p,Γ (cid:0) H s0/~r,~ω0 (ii) Assume s0, s1, sθ /∈ r(N + 1/p). Then (0,∞)(cid:1)θ (0,∞), H s1/~r,~ω1 (0,∞), Fs1/~r,~ω1 (cid:0)H s0/~r,~ω0 (cid:0)Fs0/~r,~ω0 and p p p p p . = Fsθ/~r,~ωθ (iii) Suppose s0, s1, sθ /∈ N + 1/p with s0, s1, sθ /∈ r(N + 1/p) if J = R+. Then ,Fs1/~r,~ω1 . = [Fs0/~r,~ω0 . = Fsθ /~r,~ωθ ,Fs1/~r,~ω1 (Fs0/~r,~ω0 )θ ]θ p p p p p and ( H s0/~r,~ω0 p , H s1/~r,~ω1 p )θ,p . = Bsθ /~r,~ωθ p . P r o o f. To prove (i) we can assume s1 > 1/p, due to Theorem 8.3(ii). Hence k := [s1 − 1/p]− ≥ 0. Set (J) for j ∈ {0, 1, θ}. B := (∂0 Hence assertion (i) is a consequence of Theorem 21.1. The proofs for claims (ii) and (iii) follow analogous lines. n) on Γ × J. Then Theorem 19.1(i) guarantees F n, . . . , ∂k sj /~r,~ωj p,B sj /~r,~ωj p,Γ (J) = F Since, in (8.5), the negative order spaces have been defined by duality we can now prove interpolation theorems for these spaces as well. Theorem 21.7 Suppose −∞ < s0 < s1 < 1/p, 0 < θ < 1, and λ0, λ1 ∈ R. Assume s0, s1, sθ /∈ −N + 1/p and, if J = R+, also s0, s1, sθ /∈ r(−N + 1/p). Then (Fs0/~r,~ω0 p , Fs1/~r,~ω1 p )θ . = Fsθ /~r,~ωθ p . = [Fs0/~r,~ω0 p , Fs1/~r,~ω1 p ]θ and (H s0/~r,~ω0 p , H s1/~r,~ω1 p )θ,p . = Bsθ /~r,~ωθ p . P r o o f. This follows easily from Theorem 21.6(iii), the duality properties of (·,·)θ, and Theorem 8.3(ii) and Corollary 8.4(ii). Suppose M is an m-dimensional compact submanifold of Rm with boundary and W = M × Cn. In this situation it has been shown by R. Seeley [44] that [Lp, H s p,B]θ = H θs p,B, s > 0, (21.17) p,B)θ,p p,B)θ,p . = Bθs . = Bθk p,B and (Lp, Bs with B a normal system of boundary operators (with smooth coefficients). This generalizes the earlier result by P. Grisvard [15] who obtained (21.17) in the case p = 2 and n = 1. The latter author proved in [16] that p,B for k ∈ N× and s > 0. An extension of these results to arbitrary (Lp, W k Banach spaces is due to D. Guidetti [17]. In each of those papers the 'singular values' N + 1/p are considered also. (If s ∈ N + 1/p, then H s p,B are no longer closed subspaces of H s Following the ideas of R. Seeley and D. Guidetti we have given in [4, Theorem 4.9.1] a proof of the anisotropic part of Theorem 21.3 in the special case where M = Hm and J = R, respectively M = Rm and J = R+ (to remain in the setting of this paper), W = M × Cn, and B has constant coefficients. The proof given here, which is solely based on Theorem 20.3 and general properties of interpolation functors, is new even in this simple Euclidean setting. p, respectively.) p,B and Bs p and Bs 92 H. Amann: Anisotropic Function Spaces on Singular Manifolds 22 Bounded Cylinders So far we have developed the theory of weighted anisotropic function spaces on full and half-cylinders, making use of the dilation invariance of J. In this final section we now show that all preceding results not explicitly depending on this dilation invariance remain valid in the case of cylinders of finite height. Throughout this section Furthermore, Fs/~r,~ω (J) = Fs/~r,~ω For k ∈ N we introduce W kr/~r p p p s > 0 analogously to (8.3). Similarly as in (19.1) 0 < T < ∞, • J = R+, (J, V ) etc. (JT ) by replacing J in definition (8.1) by JT . Then Fs/~r,~ω JT := [0, T ]. p (JT ) is defined for Moreover, Fs/~r,~ω p,Γ (JT ) is the closure of D(cid:0)(0, T ],D(M \Γ, V )(cid:1) in Fs/~r,~ω p (JT ) for s > 0. Fs/~r,~ω p (JT ) := Fs/~r,~ω p,∂M (JT ), Fs/~r,~ω p (0, T ] := Fs/~r,~ω p,∅ (JT ). Note that we do not require that u ∈ Fs/~r,~ω define: p,Γ (JT ) approaches zero near T . To take care of this situation also we Fs/~r,~ω p (0, T ) is the closure of D(cid:0)(0, T ), D(cid:1) in Fs/~r,~ω p (JT ). Then p p′ and H 0/~r,~ω F−s/~r,~ω (JT ) :=(cid:0)Fs/~r,~ω (cid:0)(0, T ), V ′(cid:1)(cid:1)′ (JT ) :=(cid:0)B−s(p)/~r,~ω This defines the weighted anisotropic Bessel potential space scale(cid:2) H s/~r,~ω (cid:2) Bs/~r,~ω (JT ) ; s ∈ R(cid:3) on JT . (JT ) := Lp(JT , Lλ p ), B0/~r,~ω p p p p , bckr/~r,~ω(JT ) is the closure of BC∞/~r,~ω(JT ) := \i∈N BCir/~r,~ω(JT ) s > 0, (JT ), Bs(p)/~r,~ω p (JT )(cid:1)1/2,p. (JT ) ; s ∈ R(cid:3) and Besov space scale p As for Holder space scales, BCkr/~r,~ω(JT ) is obtained by replacing J in (12.12) and (12.13) by JT . Then in BCkr/~r,~ω(JT ). Besov-Holder spaces are defined for s > 0 by Bs/~r,~ω ∞ (JT ) := (cid:0)bckr/~r,~ω(JT ), bc(k+1)r/~r,~ω(JT )(cid:1)(s−k)/r,∞, kr < s < (k + 1)r, (cid:0)bckr/~r,~ω(JT ), bc(k+2)r/~r,~ω(JT )(cid:1)1/2,∞, s = (k + 1)r. ∞ (JT ). Lastly, bs/~r,~ω ∞ (0, T ] is obtained by substi- (22.1) ∞ (JT ) is the closure of BC∞/~r,~ω(JT ) in Bs/~r,~ω Moreover, bs/~r,~ω tuting (0, T ] for (0,∞) in (17.10) and (17.11). Given a locally convex space X , the continuous linear map rT : C(J,X ) → C(JT ,X ), u 7→ uJT is the point-wise restriction to JT . As usual, we use the same symbol for rT and any of its restrictions or (unique) continuous extensions. Theorem 22.1 Let one of the following conditions be satisfied: (α) (β) (γ) (δ) p s ∈ R and G = Fs/~r,~ω ; s > 0 and G ∈ {Bs/~r,~ω k ∈ N and G ∈ {BCkr/~r,~ω, bckr/~r,~ω}; s > 0 and G = Fs/~r,~ω p,Γ . ∞ , bs/~r,~ω ∞ }; P r o o f. (1) Suppose k ∈ N. It is obvious that Then rT is a retraction from G(J) onto G(JT ) possessing a universal coretraction eT . It is also a retraction from G(0,∞) onto G(0, T ] with coretraction eT if either s > 0 and G = bs/~r,~ω (JT )(cid:1). Thus we get r+ ∈ L(cid:0)G(J), G(JT )(cid:1) if (α) is satisfied with s > 0 by interpolation, due to the definition of (I) for I ∈ {J, JT}. (2) It is also clear that r+ ∈ L(cid:0)BCkr/~r,~ω(J), BCkr/~r,~ω(JT )(cid:1) for k ∈ N. Hence r+ ∈ L(cid:0)W kr/~r,~ω ∞ or k ∈ N and G = bckr/~r,~ω. (J), W kr/~r,~ω Fs/~r,~ω p p p From this we obtain r+ ∈ L(cid:0)BC∞/~r,~ω(J), BC∞/~r,~ω(JT )(cid:1). r+ ∈ L(cid:0)G(J), G(JT )(cid:1) 93 (22.2) (22.3) if either (β) or (γ) is satisfied. In fact, this is obvious from (22.2) if (γ) applies. If s > 0 and G = Bs/~r,~ω ∞ , then (22.3) is obtained by interpolation on account of (12.21), Corollary 12.9(ii), and (22.1). From this and (22.2) it follows that (22.3) is valid if G = bs/~r,~ω ∞ , due to (12.24) and the definition of bcs/~r,~ω(JT ). Clearly, r+ maps D(cid:0) J,D(M \Γ, V )(cid:1) into D(cid:0)(0, T ],D(M \Γ, V )(cid:1). From this and step (1) we infer that (22.3) is true if (δ) applies. It is equally clear that r+ ∈ L(cid:0)G(0,∞), G(0, T )(cid:1) if either s > 0 and G = bs/~r,~ω ∞ or k ∈ N (3) We set δT (t) := t + T for t ∈ R. We fix α ∈ D(cid:0)(−T, 0], R(cid:1) satisfying α(t) = 1 for −T /2 ≤ t ≤ 0 and T u(t) for t ≤ 0 and u : JT → C(V ). It follows that β ∈ L(cid:0)G(JT ), G(−R+)(cid:1), provided s > 0 put βu(t) := αδ∗ if (α) holds. Indeed, this is easily verified if G is one of the spaces W kr/~r,~ω claim by interpolation, similarly as in steps (1) and (2). and BCkr/~r,~ω. From this we get the and G = bckr/~r,~ω. p (4) We recall from Section 17 the definition of the extension operator e− associated with the point-wise restriction r− to −R+. Then we define a linear map u 7→ δ∗ −T (e−βu). εT : C(cid:0)JT , C(V )(cid:1) → C(cid:0)[T,∞), C(V )(cid:1), eT ∈ L(cid:0)G(JT ), G(J)(cid:1) Finally, we put eT u(t) := u(t) for t ∈ JT and eT u(t) := εT u(t) for T < t < ∞. It follows from step (3) and Theorems 17.1 and 17.2 that (22.4) if one of conditions (α) -- (γ) is satisfied, provided s > 0 if (α) applies. Since α is compactly supported it follows from (17.1) that εT u is smooth and rapidly decreasing if u is smooth. By the density of D(cid:0)[T,∞),D(M \Γ, V )(cid:1) in the Schwartz space of smooth rapidly decreasing D(M \Γ, V )- valued functions on [T,∞) we get eT u ∈ Fs/~r,~ω (δ) is satisfied. It is obvious that rT eT = id. Thus the assertion is proved, provided s > 0 if (α) is satisfied. p,Γ (JT ). From this we see that (22.4) holds if p,Γ (J) if u ∈ Fs/~r,~ω (5) As in (17.5) we introduce the trivial extension map e− 0 : C(0)(−R+,X ) → C(R,X ) by e− 0 u(t) := u(t) if t ≤ 0, and e− 0 u(t) := 0 if t > 0. Then is a retraction possessing e− 0 as coretraction. We also set r0,T := δ∗ −T r− 0 δ∗ T e+ 0 . Then r− 0 := r−(1 − e+r+) : C(R,X ) → C0(−R+,X ) The mapping properties of e+ and r+ described in Theorems 17.1 and 17.2, and the analogous ones for r−, imply, similarly as above, that (22.5) r0,T(cid:0)D(cid:0)(0, T ), D(cid:1)(cid:1) ⊂ D(cid:0)(0,∞), D(cid:1). r0,T ∈ L(cid:0)Fs/~r,~ω (JT )(cid:1), (J), Fs/~r,~ω p p s > 0. 94 H. Amann: Anisotropic Function Spaces on Singular Manifolds Consequently, we get from (22.5) r0,T ∈ L(cid:0)Fs/~r,~ω p (J),Fs/~r,~ω p s > 0. (0, T )(cid:1), tends to a continuous linear map from Fs/~r,~ω r0,T e0,T = id. Thus r0,T is a retraction possessing e0,T as coretraction. We define e0,T : D(cid:0)(0, T ), D(cid:1) → D(cid:0)(0,∞), D(cid:1) by e0,T uJT := u and e0,T u[0,∞) := 0. Then e0,T ex- (6) Let s > 0. For u ∈ D(J,D) and ϕ ∈ D(cid:0)(0, T ),D( M , V ′)(cid:1) we get (J) for s > 0, the trivial extension. Moreover, (0, T ) into Fs/~r,~ω p p Z T 0 ZMhϕ, rT uiV dVg dt =Z ∞ 0 ZMhe0,T ϕ, uiV dVg dt. Hence, by step (5) and the definition of the negative order spaces, hϕ, rT uiM×J ≤ ckϕkF s/~r,~ω p′ ( JT ,V ′) kukF −s/~r,~ω p (J) for u ∈ F−s/~r,~ω p (J) and ϕ ∈ Fs/~r,~ω p′ (7) For v ∈ C(−J,D) we set p (cid:0)(0, T ), V ′(cid:1). Thus rT ∈ L(cid:0)F−s/~r,~ω ε−v(t) :=Z ∞ 0 (J), F−s/~r,~ω p (JT )(cid:1). h(s)v(−st) ds, t ≥ 0. Then, given ϕ ∈ D(cid:0)(0,∞),D( M , V ′)(cid:1), we obtain from h(1/s) = −sh(s) for s > 0 and (17.3) 0 (cid:10)ϕ(t), ε−v(t)(cid:11)M dt =Z ∞ 0 Z ∞ Z ∞ 0 (cid:10)ϕ(t), h(s)v(−st)(cid:11)M ds dt =Z 0 −∞Z ∞ 0 (cid:10)s−1ϕ(−τ /s)h(s), v(τ )(cid:11)M ds dτ =Z 0 −∞Z ∞ σ−1(cid:10)ϕ(−τ σ)h(1/σ), v(τ )(cid:11)M dσ dτ −∞DZ ∞ = −Z 0 h(σ)ϕ(−στ ) dσ, v(τ )EM Thus, by the definition of eT , given u ∈ D(JT ,D), hϕ, eT uiM dt =Z T 0 hϕ, uiM dt +Z ∞ T hϕ, δ∗ The last integral equals, due to e−w(t) = ε−w(t) for t ≥ 0 and (22.6), Z ∞ 0 0 0 dτ = −Z 0 −∞hεϕ, viM dτ. −T e−αδ∗ T uiM dt. Z ∞ 0 T ϕ, ε−αδ∗ hδ∗ T uiM dt = −Z 0 = −Z T T ϕ, αδ∗ T uiM dt −∞hεδ∗ 0 Z ∞ 0 (cid:10)h(σ)ϕ(cid:0)−σ(s − T ) + T(cid:1) dσ, α(s − T )u(s)(cid:11)M ds since α is supported in (−T, 0]. From this we infer as in steps (3) and (4) that, given s > 0, hϕ, eT uiM×J ≤ ckϕkF s/~r,~ω p′ (J,V ′) kukF −s/~r,~ω p (JT ) (22.6) 95 (JT ), F−s/~r,~ω p′ (J)(cid:1) for ϕ ∈ D(cid:0)(0,∞),D( M , V ′)(cid:1) and u ∈ D(JT ,D). Thus eT ∈ L(cid:0)F−s/~r,~ω p for s > 0. This and step (6) imply that the assertion holds if (α) is satisfied with s < 0. The case s = 0 and F = H is covered by step (1). If s = 0 and F = B, we now obtain the claim by interpola- tion, due to the definition of B0/~r,~ω p (I) for I ∈ {J, JT}. Corollary 22.2 Suppose s > 0. There exists a universal retraction r0,T from Fs/~r,~ω p that the trivial extension is a coretraction for it. (J) onto Fs/~r,~ω p ( JT ) such P r o o f. This has been shown in step (5). As a consequence of this retraction theorem we find that, modulo obvious adaptions, everything proved in the preceding sections remains valid for cylinders of finite height. Theorem 22.3 All embedding, interpolation, trace, and point-wise contraction multiplier theorems, as well as the theorems involving boundary conditions, remain valid if J is replaced by JT . Furthermore, all retraction theorems for the anisotropic spaces stay in force, provided ϕ~ω q ◦ eT and rT ◦ ψ~ω q , respectively. q are replaced by ϕ~ω q and ψ~ω P r o o f. This is an immediate consequence of Theorem 22.1 and the fact that all contraction multiplication and boundary operators are local ones. References [1] H. Amann. Multiplication in Sobolev and Besov Spaces. In Nonlinear Analysis, A Tribute in Honour of G. Prodi, pages 27 -- 50. Quaderni, Scuola Norm. Sup. Pisa, 1991. [2] H. Amann. Linear and Quasilinear Parabolic Problems, Volume I: Abstract Linear Theory. Birkhauser, Basel, 1995. [3] H. Amann. Operator-valued Fourier multipliers, vector-valued Besov spaces, and applications. Math. Nachr., 186 (1997), 5 -- 56. [4] H. Amann. Anisotropic Function Spaces and Maximal Regularity for Parabolic Problems. Part 1: Function Spaces. Jindrich Necas Center for Mathematical Modeling, Lecture Notes, 6, Prague, 2009. [5] H. Amann. Function Spaces on Singular Manifolds. Math. Nachr., (2012). arXiv:1106.2033. [6] H. Amann. Parabolic equations on singular manifolds, 2012. In preparation. [7] H. Amann, J. Escher. Analysis III. Birkhauser, Basel, 2008. English Translation. [8] J. Bergh, J. Lofstrom. Interpolation Spaces. An Introduction. Springer Verlag, Berlin, 1976. [9] S. Coriasco, E. Schrohe, J. Seiler. Bounded imaginary powers of differential operators on manifolds with conical singularities. Math. Z., 244 (2003), 235 -- 269. [10] S. Coriasco, E. Schrohe, J. Seiler. Bounded H∞-calculus for differential operators on conic manifolds with boundary. Comm. Partial Differential Equations, 32 (2007), 229 -- 255. [11] R. Denk, M. Hieber, J. Pruss. Optimal Lp-Lq-estimates for parabolic boundary value problems with inhomogeneous data. Math. Z., 257(1) (2007), 193 -- 224. [12] J. Dieudonn´e. El´ements d'Analyse, I -- VII. Gauthier-Villars, Paris, 1969 -- 1978. [13] G. Dore, A. Favini. On the equivalence of certain interpolation methods. Boll. UMI (7), 1-B (1987), 1227 -- 1238. [14] P. Grisvard. Commutativit´e de deux foncteurs d'interpolation et applications. J. Math. Pures Appl., 45 (1966), 143 -- 290. [15] P. Grisvard. Caract´erisation de quelques espaces d'interpolation. Arch. Rational Mech. Anal., 25 (1967), 40 -- 63. [16] P. Grisvard. ´Equations diff´erentielles abstraites. Ann. Sc. ´Ec. Norm. Sup., S´er. 4 (1969), 311 -- 395. [17] D. Guidetti. On interpolation with boundary conditions. Math. Z., 207 (1991), 439 -- 460. [18] R.S. Hamilton. Harmonic Maps of Manifolds with Boundary. Lecture Notes in Mathematics, 471. Springer-Verlag, Berlin, 1975. [19] E. Hille, R.S. Phillips. Functional Analysis and Semi-Groups. Amer. Math. Soc., Providence, R.I., 1957. [20] J. Johnson. Pointwise multiplication of Besov and Triebel-Lizorkin spaces. Math. Nachr., 175 (1995), 85 -- 133. [21] J. Jost. Riemannian Geometry and Geometric Analysis. Springer-Verlag, Berlin, 2008. [22] V.A. Kondrat′ev. Boundary value problems for elliptic equations in domains with conical or angular points. Trudy Moskov. Mat. Obsc., 16 (1967), 209 -- 292. [23] V.A. Kozlov. Asymptotic behavior as t → 0 of the solutions of the heat equation in a domain with a conic point. Mat. Sb. (N.S.), 136(178) (1988), 384 -- 395. 96 H. Amann: Anisotropic Function Spaces on Singular Manifolds [24] V.A. Kozlov. Coefficients in the asymptotics of the solutions of initial-boundary value parabolic problems in domains with a conic point. Sibirsk. Mat. Zh., 29 (1988), 75 -- 89. [25] V.A. Kozlov. Asymptotics of the Green function and Poisson kernels of a mixed parabolic problem in a cone. I. Z. Anal. Anwendungen, 8 (1989), 131 -- 151. [26] V.A. Kozlov. Asymptotics of the Green function and Poisson kernels of a mixed parabolic problem in a cone. II. Z. Anal. Anwendungen, 10 (1991), 27 -- 42. [27] V.A. Kozlov, V.G. Maz′ya, J. Rossmann. Elliptic Boundary Value Problems in Domains with Point Singularities. Amer. Math. Soc., Providence, RI, 1997. aczkowski. A parabolic system in a weighted Sobolev space. Appl. Math. (Warsaw), 34 (2007), ' [28] A. Kubica, W.M. Zaj 169 -- 191. [29] A. Kubica, W.M. Zaj Appl. Math. (Warsaw), 34 (2007), 431 -- 444. aczkowski. A priori estimates in weighted spaces for solutions of the Poisson and heat equations. ' [30] O.A. Ladyzhenskaya, V.A. Solonnikov, N.N. Ural'ceva. Linear and Quasilinear Equations of Parabolic Type. Amer. Math. Soc., Transl. Math. Monographs, Providence, R.I., 1968. [31] J.-L. Lions, E. Magenes. Non-Homogeneous Boundary Value Problems and Applications I -- III. Springer Verlag, Berlin, 1972/1973. [32] A. Lunardi. Analytic Semigroups and Optimal Regularity in Parabolic Problems. Birkhauser, Basel, 1995. [33] V.G. Maz'ya, T.O. Shaposhnikova. Theory of Multipliers in Spaces of Differentiable Functions. Pitman, Boston, 1985. [34] A.I. Nazarov. Lp-estimates for a solution to the Dirichlet problem and to the Neumann problem for the heat equation in a wedge with edge of arbitrary codimension. J. Math. Sci. (New York), 106 (2001), 2989 -- 3014. [35] S.A. Nazarov, B.A. Plamenevskiı. Elliptic Problems in Domains with Piecewise Smooth Boundaries. Walter de Gruyter & Co., Berlin, 1994. [36] K. Pileckas. On the nonstationary linearized Navier-Stokes problem in domains with cylindrical outlets to infinity. Math. Ann., 332 (2005), 395 -- 419. [37] K. Pileckas. Solvability in weighted spaces of the three-dimensional Navier-Stokes problem in domains with cylindrical outlets to infinity. Topol. Methods Nonlinear Anal., 29 (2007), 333 -- 360. [38] K. Pileckas. Global solvability in W 2,1 2 -weighted spaces of the two-dimensional Navier-Stokes problem in domains with strip-like outlets to infinity. J. Math. Fluid Mech., 10 (2008), 272 -- 309. [39] Th. Runst, W. Sickel. Sobolev Spaces of Fractional Order, Nemytskii Operators, and Nonlinear Partial Differential Equations. W. de Gruyter, Berlin, 1996. [40] H.H. Schaefer. Topological Vector Spaces. Springer-Verlag, New York, 1971. [41] H.-J. Schmeisser. Vector-valued Sobolev and Besov spaces. Seminar Analysis 1985/86, Teubner-Texte Math., 96, (1987), 4 -- 44. [42] H.-J. Schmeisser, W. Sickel. Traces, Gagliardo-Nirenberg Inequalities and Sobolev Type Embeddings for Vector-valued Function Spaces. Jenaer Schriften zur Mathematik und Informatik, (2001). [43] H.-J. Schmeisser, H. Triebel. Anisotropic spaces. I. Interpolation of abstract spaces and function spaces. Math. Nachr., 73 (1976), 107 -- 123. [44] R.T. Seeley. Norms and domains of the complex powers Az [45] V.A. Solonnikov. Lp-estimates for solutions of the heat equation in a dihedral angle. Rend. Mat. Appl. (7), 21 (2001), B. Amer. J. Math., 93 (1971), 299 -- 309. 1 -- 15. [46] V.A. Solonnikov, E.V. Frolova. A problem with the third boundary condition for the Laplace equation in a plane angle and its applications to parabolic problems. Algebra i Analiz, 2 (1990), 213 -- 241. [47] V.A. Solonnikov, E.V. Frolova. A nonstationary problem in a dihedral angle. II. Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 188 (1991), 178 -- 185. [48] E.M. Stein. Singular Integrals and Differentiability Properties of Functions. Princeton University Press, Princeton, N.J., 1970. [49] F. Treves. Topologic Vector Spaces, Distributions and Kernels. Academic Press, New York, 1967. [50] H. Triebel. Interpolation Theory, Function Spaces, Differential Operators. North Holland, Amsterdam, 1978. [51] H. Triebel. Theory of Function Spaces. Birkhauser, Basel, 1983. [52] W.M. Zaj distinguished axis. I. Existence near the axis in 2d. Appl. Math. (Warsaw), 34 (2007), 121 -- 141. aczkowski. Existence of solutions to the nonstationary Stokes system in H 2,1 ' aczkowski. The Helmholtz-Weyl decomposition in weighted Sobolev spaces. Math. Methods Appl. Sci., 34 ' aczkowski. Nonstationary Stokes system in weighted Sobolev spaces. Math. Methods Appl. Sci., 34 (2011), ' aczkowski. Solvability of the heat equation in weighted Sobolev spaces. Appl. Math. (Warsaw), 38 (2011), ' −µ, µ ∈ (0, 1), in a domain with a (2011), 191 -- 197. [53] W.M. Zaj [54] W.M. Zaj 544 -- 562. [55] W.M. Zaj 147 -- 171.
1701.07777
1
1701
2017-01-26T17:13:16
Henkin measures for the Drury-Arveson space
[ "math.FA", "math.CV" ]
We exhibit Borel probability measures on the unit sphere in $\mathbb C^d$ for $d \ge 2$ which are Henkin for the multiplier algebra of the Drury-Arveson space, but not Henkin in the classical sense. This provides a negative answer to a conjecture of Clou\^atre and Davidson.
math.FA
math
HENKIN MEASURES FOR THE DRURY-ARVESON SPACE MICHAEL HARTZ Abstract. We exhibit Borel probability measures on the unit sphere in Cd for d ≥ 2 which are Henkin for the multiplier algebra of the Drury- Arveson space, but not Henkin in the classical sense. This provides a negative answer to a conjecture of Clouâtre and Davidson. 1. Introduction Let Bd denote the open unit ball in Cd and let A(Bd) be the ball algebra, which is the algebra of all analytic functions on Bd which extend to be continuous on Bd. A regular complex Borel measure µ on the unit sphere Sd = ∂Bd is said to be Henkin if the functional A(Bd) → C, f 7→ZSd f dµ, extends to a weak-∗ continuous functional on H ∞(Bd), the algebra of all bounded analytic functions on Bd. Equivalently, whenever (fn) is a sequence in A(Bd) which is uniformly bounded on Bd and satisfies limn→∞ fn(z) = 0 for all z ∈ Bd, then n→∞ZSd lim fn dµ = 0. Henkin measures play a prominent role in the description of the dual space of A(Bd) and of peak interpolation sets for the ball algebra, see Chapter 9 and 10 of [17] for background material. Such measures are completely characterized by a theorem of Henkin [14] and Cole-Range [8]. To state the theorem, recall that a Borel probability measure τ on Sd is said to be a representing measure for the origin if ZSd f dτ = f (0) for all f ∈ A(Bd). Theorem 1.1 (Henkin, Cole-Range). A regular complex Borel measure µ on Sd is Henkin if and only if it is absolutely continuous with respect to some representing measure for the origin. 2010 Mathematics Subject Classification. Primary 46E22; Secondary 47A13. Key words and phrases. Drury-Arveson space, Henkin measure, totally singular mea- sure, totally null set. The author was partially supported by a Feodor Lynen Fellowship. 1 2 MICHAEL HARTZ If d = 1, then the only representing measure for the origin is the normal- ized Lebesgue measure on the unit circle, hence the Henkin measures on the unit circle are precisely those measures which are absolutely continuous with respect to Lebesgue measure. In addition to their importance in complex analysis, Henkin measures also play a role in multivariable operator theory [13]. However, it has become clear over the years that for the purposes of multivariable operator theory, the "correct" generalization of H ∞, the algebra of bounded analytic functions on the unit disc, to higher dimensions is not H ∞(Bd), but the multiplier algebra of the Drury-Arveson space H 2 d . This is the reproducing kernel Hilbert space on Bd with reproducing kernel K(z, w) = 1 . 1 − hz, wi A theorem of Drury [11] shows that H 2 d hosts a version of von Neumann's inequality for commuting row contractions, that is, tuples T = (T1, . . . , Td) of commuting operators on a Hilbert space H such that the row operator [T1, . . . , Td] : Hd → H is a contraction. The corresponding dilation theorem is due to Müller-Vasilescu [15] and Arveson [5]. The Drury-Arveson space is also known as symmetric Fock space [5, 10], it plays a distinguished role in the theory of Nevanlinna-Pick spaces [1, 2] and is an object of interest in harmonic analysis [9, 4]. An overview of the various features of this space can be found in [18]. In [7], Clouâtre and Davidson generalize much of the classical theory of Henkin measures to the Drury-Arveson space. Let Md denote the multiplier algebra of H 2 d and let Ad be the norm closure of the polynomials in Md. In particular, functions in Ad belong to A(Bd). Clouâtre and Davidson define a regular Borel measure µ on Sd to be Ad-Henkin if the associated integration functional Ad → C, f 7→ZSd f dµ extends to a weak-∗ continuous functional on Md (see Subsection 2.1 for the definition of weak-∗ topology). Equivalently, whenever (fn) is a sequence in Ad such that fnMd ≤ 1 for all n ∈ N and limn→∞ fn(z) = 0 for all z ∈ Bd, then ZSd fn(z) dµ = 0, see [7, Theorem 3.3]. This notion, along with the complementary notion of Ad-totally singular measures, is crucial in the study of the dual space of Ad and of peak interpolation sets for Ad in [7]. Compelling evidence of the importance of Ad-Henkin measures in multi- variable operator theory can be found in [6], where Clouâtre and Davidson extend the Sz.-Nagy-Foias H ∞ -- functional calculus to commuting row con- tractions. Recall that every contraction T on a Hilbert space can be written HENKIN MEASURES FOR THE DRURY-ARVESON SPACE 3 as T = Tcnu ⊕ U , where U is a unitary operator and Tcnu is completely non- unitary (i.e. has no unitary summand). Sz.-Nagy and Foias showed that in the separable case, T admits a weak-∗ continuous H ∞-functional calculus if and only if the spectral measure of U is absolutely continuous with respect to Lebesgue measure on the unit circle, see [19] for a classical treatment. Clouâtre and Davidson obtain a complete generalization of this result. The appropriate generalization of a unitary is a spherical unitary, which is a tu- ple of commuting normal operators whose joint spectrum is contained in the unit sphere. Every commuting row contraction admits a decomposition T = Tcnu ⊕ U , where U is a spherical unitary and Tcnu is completely non- unitary (i.e. has no spherical unitary summand), see [6, Theorem 4.1]. The following result is then a combination of Lemma 3.1 and Theorem 4.3 of [6]. Theorem 1.2 (Clouâtre-Davidson). Let T be a commuting row contraction acting on a separable Hilbert space with decomposition T = Tcnu⊕U as above. Then T admits a weak-∗ continuous Md-functional calculus if and only if the spectral measure of U is Ad-Henkin. This result shows that for the theory of commuting row contractions, Ad- Henkin measures are a more suitable generalization of absolutely continuous measures on the unit circle than classical Henkin measures. Thus, a charac- terization of Ad-Henkin measures would be desirable. Since the unit ball of Ad is contained in the unit ball of A(Bd), it is trivial that every classical Henkin measure is also Ad-Henkin. Clouâtre and David- son conjectured [7, Conjecture 5.1] that conversely, every Ad-Henkin measure is also a classical Henkin measure, so that these two notions agree. If true, the classical theory would apply to Ad-Henkin measures and in particular, the Henkin and Cole-Range theorem would provide a characterization of Ad- Henkin measures. They also formulate a conjecture for the complementary notion of totally singular measure, which turns out to be equivalent to their conjecture on Henkin measures [7, Theorem 5.2]. Note that the conjecture is vacuously true if d = 1, as M1 = H ∞. The purpose of this note is to provide a counterexample to the conjec- ture of Clouâtre and Davidson for d ≥ 2. To state the main result more precisely, we require one more definition. A compact set K ⊂ Sd is said to be totally null if it is null for every representing measure of the origin. By the Henkin and Cole-Range theorem, a totally null set cannot support a non-zero classical Henkin measure. Theorem 1.3. Let d ≥ 2 be an integer. There exists a Borel probability measure µ on Sd which is Ad-Henkin and whose support is totally null. In fact, every measure which is supported on a totally null set is totally singular (i.e. it is singular with respect to every representing measure of the origin). The measure in Theorem 1.3 therefore also serves at the same time a counterexample to the conjecture of Clouâtre and Davidson on totally singular measures, even without invoking [7, Theorem 5.2]. 4 MICHAEL HARTZ It is not hard to see that if µ is a measure on Sd which satisfies the conclusion of Theorem 1.3, then so does the trivial extension of µ to Sd′ for any d′ ≥ d (see Lemma 2.3), hence it suffices to prove Theorem 1.3 for d = 2. In fact, the construction of such a measure µ is easier in the case d = 4, so will consider that case first. The remainder of this note is organized as follows. In Section 2, we recall some of the necessary background material. Section 3 contains the construc- tion of a measure µ which satisfies the conclusion of Theorem 1.3 in the case d = 4. In Section 4, we prove Theorem 1.3 in general. 2. Preliminaries 2.1. The Drury-Arveson space. As mentioned in the introduction, the Drury-Arveson space H 2 d is the reproducing kernel Hilbert space on Bd with reproducing kernel K(z, w) = 1 . 1 − hz, wi For background material on reproducing kernel Hilbert spaces, see [16] and [3]. We will require a more concrete description of H 2 d . Recall that if α = (α1, . . . , αd) ∈ Nd is a multi-index and if z = (z1, . . . , zd) ∈ Cd, one usually writes zα = zα1 1 . . . zαd d , α! = α1! . . . αd!, α = α1 + . . . + αd. The monomials zα form an orthogonal basis of H 2 d , and α! α! for every multi-index α, see [5, Lemma 3.8]. zα2 H 2 d = Let (xn) and (yn) be two sequences of positive numbers. We write xn ≃ yn to mean that there exist C1, C2 > 0 such that C1yn ≤ xn ≤ C2yn for all n ∈ N. The following well-known result can be deduced from Stirling's formula, see [5, p.19]. Lemma 2.1. Let d ∈ N. Then (z1z2 . . . zd)n2 d ≃ d−nd(n + 1)(d−1)/2 H 2 for all n ∈ N. The multiplier algebra of H 2 d is (cid:3) Md = {ϕ : Bd → C : ϕf ∈ H 2 d for all f ∈ H 2 d}. Every ϕ ∈ Md gives rise to a bounded multiplication operator Mϕ on H 2 d , and we set ϕMd = Mϕ. Moreover, we may identify Md with a unital subalgebra of B(H 2 d . It is not hard to see that Md is WOT-closed, and hence weak-∗ closed, inside of B(H 2 d ). d ), the algebra of bounded operators on H 2 HENKIN MEASURES FOR THE DRURY-ARVESON SPACE 5 Thus, Md becomes a dual space in this way, and we endow it with the resulting weak-∗ topology. In particular, for every f, g ∈ H 2 d , the functional Md → C, ϕ 7→ hMϕf, gi, is weak-∗ continuous. Moreover, it is well known and not hard to see that on bounded subsets of Md, the weak-∗ topology coincides with the topology of pointwise convergence on Bd. 2.2. Henkin measures and totally null sets. Let K ⊂ Sd be a compact set. A function f ∈ A(Bd) is said to peak on K if f = 1 on K and f (z) < 1 for all z ∈ Bd \ K. Recall that K is said to be totally null if it is null for every representing measure of the origin. In particular, if d = 1, then K is totally null if and only if it is a Lebesgue null set. We will make repeated use of the following characterization of totally null sets, see [17, Theorem 10.1.2]. Theorem 2.2. A compact set K ⊂ Sd is totally null if and only if there exists a function f ∈ A(Bd) which peaks on K. If d′ ≥ d, then we may regard Sd ⊂ Sd′ in an obvious way. Thus, every regular Borel measure µ on Sd admits a trivial extension bµ to Sd′ defined by for Borel sets A ⊂ Sd′. The following easy lemma shows that it suffices to prove Theorem 1.3 in the case d = 2. Lemma 2.3. Let µ be a Borel probability measure on Sd, let d′ ≥ d and let bµ be the trivial extension of µ to Sd′. bµ is a totally null subset of Sd′ . (a) If µ is Ad-Henkin, then bµ is Ad′-Henkin. Proof. (a) Let P : Cd′ → Cd denote the orthogonal projection onto the first d coordinates. It follows from the concrete description of the Drury-Arveson space at the beginning of Subsection 2.1 that (b) If the support of µ is a totally null subset of Sd, then the support of bµ(A) = µ(A ∩ Sd) V : H 2 d → H 2 d′, f 7→ f ◦ P, is an isometry. Moreover, V ∗MϕV = MϕBd for every ϕ ∈ Md′, so that Md′ 7→ Md, ϕ 7→ ϕ(cid:12)(cid:12)Bd , is weak-∗-weak-∗ continuous and maps Ad′ into Ad. Suppose now that µ is Ad-Henkin. Then there exists a weak-∗ continuous functional Φ on Md which extends the integration functional given by µ, thus ϕ dbµ =ZSd for ϕ ∈ Ad′. Since the right-hand side defines a weak-∗ continuous functional on Md′, we see that bµ is Ad′-Henkin. ϕ dµ = Φ(ϕ(cid:12)(cid:12)Bd ZS ) d′ 6 MICHAEL HARTZ (b) We have to show that if K ⊂ Sd is totally null, then K is also totally null as a subset of Sd′. But this is immediate from Theorem 2.2 and the observation that if f ∈ A(Bd) peaks on K, then f ◦ P ∈ A(Bd′) peaks on K as well, where P denotes the orthogonal projection from (a). (cid:3) 3. The case d = 4 The goal of this section is to prove Theorem 1.3 in the case d = 4 (and hence for all d ≥ 4 by Lemma 2.3). To prepare and motivate the construc- tion of the measure µ, we begin by considering analogues of Henkin measures for more general reproducing kernel Hilbert spaces on the unit disc. Sup- pose that H is a reproducing kernel Hilbert space on the unit disc D with reproducing kernel of the form (1) K(z, w) = an(zw)n, ∞Xn=0 where a0 = 1 and an > 0 for all n ∈ N. If P∞ n=0 an < ∞, then the series above converges uniformly on D × D, and H becomes a reproducing kernel Hilbert space of continuous functions on D in this way. In particular, eval- uation at 1 is a continuous functional on H and hence a weak-∗ continuous functional on Mult(H). Indeed, ϕ(1) = hMϕ1, K(·, 1)iH for ϕ ∈ Mult(H). Therefore, the Dirac measure δ1 induces a weak-∗ con- tinuous functional on Mult(H), but it is not absolutely continuous with re- spect to Lebesgue measure, and hence not Henkin. (In fact, every regular Borel measure on the unit circle induces a weak-∗ continuous functional on Mult(H).) The main idea of the construction is to embed a reproducing kernel Hilbert space as in the preceding paragraph into H 2 4 . To find the desired space H on the disc, recall that by the inequality of arithmetic and geometric means, sup{z1z2 . . . zd : z ∈ Bd} = d−d/2, and the supremum is attained if and only if z1 = . . . = zd = d−1/2. Hence, r : B4 → D, z 7→ 16z1z2z3z4, indeed takes values in D, and it maps B4 onto D. For n ∈ N, let an = r(z)n−2 H 2 , 4 and let H be the reproducing kernel Hilbert space on D with reproducing kernel K(z, w) = an(zw)n. ∞Xn=0 HENKIN MEASURES FOR THE DRURY-ARVESON SPACE 7 Lemma 3.1. The map H → H 2 4 , n=0 an < ∞. f 7→ f ◦ r, is an isometry, and P∞ Proof. It is well known that for any space on D with kernel as in Equation (2), the monomials zn form an orthogonal basis and zn2 = 1/an for n ∈ N. Thus, with our choice of (an) above, we have zn2 = 1 an = r(z)n2 H 2 4 . Since the sequence r(z)n is an orthogonal sequence in H 2 is an isometry. Moreover, an application of Lemma 2.1 shows that 4 , it follows that V r(z)n2 = 44n(z1, . . . , z4)n2 ≃ (n + 1)3/2, so that an ≃ (n + 1)−3/2, and hence P∞ n=0 an < ∞. Let (cid:3) h : T3 → S4, (ζ1, ζ2, ζ3) 7→ 1/2(ζ1, ζ2, ζ3, ζ1ζ2ζ3) and observe that the range of h is contained in r−1({1}). Let µ be the pushforward of the normalized Lebesgue measure m on T3 by h, that is, µ(A) = m(h−1(A)) for a Borel subset A of S4. We will show that µ satisfies the conclusion of Theorem 1.3. Lemma 3.2. The support of µ is totally null. Proof. Let X = r−1({1}), which is compact, and define f = 1+r 2 . Then f belongs to the unit ball of A(B4) and peaks on X, hence X is totally null by Theorem 2.2. Since h(T3) ⊂ X, the support of µ is contained in X, so the support of µ is totally null as well. (cid:3) The following lemma finishes the proof of Theorem 1.3 in the case d = 4. Lemma 3.3. The measure µ is A4-Henkin. Proof. Let α ∈ N4 be a multi-index. Then zα ◦ h dm = 2−αZT3 zα dµ =ZT3 ZS4 ζ α1−α4 1 ζ α2−α4 2 ζ α3−α4 3 dm. This integral is zero unless α4 = α1 = α2 = α3 = : k, in which case it equals 2−4k. Let g = K(·, 1) ◦ r, where K denotes the reproducing kernel of H. Then 4 by Lemma 3.1, and it is a power series in z1z2z3z4. Thus, zα is g ∈ H 2 orthogonal to g unless α1 = . . . = α4 = : k, in which case = 2−4khr(z)k, giH 2 where we have used Lemma 3.1 again. hzα, giH 2 4 = 2−4khzk, K(·, 1)iH = 2−4k, 4 8 MICHAEL HARTZ Hence, ZS4 ϕ dµ = hMϕ1, giH 2 4 for all polynomials ϕ, and hence for all ϕ ∈ A4. Since the right-hand side obviously extends to a weak-∗ continuous functional in ϕ on M4, we see that µ is A4-Henkin. (cid:3) 4. The case d = 2 In this section, we will prove Theorem 1.3 in the case d = 2 and hence in full generality by Lemma 2.3. To this end, we will also embed a reproducing kernel Hilbert space on D into H 2 2 . Let z 7→ 2z1z2, and observe that r maps B2 onto D. For n ∈ N, let r : B2 → D, an = r(z)n−2 H 2 , 2 and consider the reproducing kernel Hilbert space on D with reproducing kernel K(z, w) = an(zw)n. ∞Xn=0 This space turns out to be the well-known weighted Dirichlet space D1/2, which is the reproducing kernel Hilbert space on D with reproducing kernel (1 − zw)−1/2. This explicit description is not strictly necessary for what follows, but it provides some context for the arguments involving capacity below. Lemma 4.1. The kernel K satisfies K(z, w) = (1 − zw)−1/2. Proof. The formula for the norm of monomials in Section 2 shows that (n!)2 = (−1)n(cid:18)−1/2 n (cid:19), n (cid:19)(zw)n = (1 − zw)−1/2 an = r(z)n−2 = 4−n (2n)! (−1)n(cid:18)−1/2 ∞Xn=0 K(z, w) = so that by the binomial series. (cid:3) The analogue of Lemma 3.1 in the case d = 2 is the following result. Lemma 4.2. The map D1/2 → H 2 2 , f 7→ f ◦ r, is an isometry. Moreover, an ≃ (n + 1)−1/2 and zn2 D1/2 ≃ (n + 1)1/2. HENKIN MEASURES FOR THE DRURY-ARVESON SPACE 9 Proof. As in the proof of Lemma 3.1, we see that V is an isometry. Moreover, Lemma 2.1 shows that zn2 D1/2 = r(z)n2 H 2 2 = 22n(z1z2)n2 2 ≃ (n + 1)1/2 H 2 for n ∈ N. (cid:3) The crucial difference to the case d = 4 is that the functions in D1/2 do not all extend to continuous functions on D. This makes the construction of the measure µ of Theorem 1.3 more complicated. The following lemma provides a measure σ on the unit circle which will serve as a replacement for the Dirac measure δ1, which was used in the case d = 4. It is very likely that this result is well known. Since the measure σ is crucial for the construction of the measure µ, we explicitly indicate how such a measure on the unit circle can arise. Lemma 4.3. There exists a Borel probability measure σ on T such that (a) the support of σ has Lebesgue measure 0, and (b) the functional C[z] → C, p 7→ZT p dσ, extends to a bounded functional on the space D1/2. To prove Lemma 4.3, we require the notion of capacity. Background ma- terial on capacity can be found in [12, Section 2]. Let k(t) = t−1/2. The 1/2-energy of a Borel probability measure ν on T is defined to be Ik(ν) =ZTZT k(x − y)dν(x)dν(y). We say that a compact subset E ⊂ T has positive Riesz capacity of degree 1/2 if there exists a Borel probability measure ν supported on E with Ik(ν) < ∞. Proof of Lemma 4.3. Let E ⊂ T be a compact set with positive Riesz ca- pacity of degree 1/2, but Lebesgue measure 0. For instance, since 1/2 < log 2/ log 3, the circular middle-third Cantor set has this property by [12, Exercise 2.4.3 (ii)]. Thus, there exists a measure σ on T whose support is contained in E with Ik(σ) < ∞. Then (a) holds. To prove (b), for n ∈ Z, let z−n dσ(z) bσ(n) =ZT ∞Xn=0 bσ(n)2 (n + 1)1/2 < ∞. denote the n-th Fourier coefficient of σ. Since Ik(σ) < ∞, an application of [12, Exercise 2.4.4] shows that (2) 10 MICHAEL HARTZ Let now p be a polynomial, say p(z) = NXn=0 αnzn. Then using the Cauchy-Schwarz inequality, we see that (cid:12)(cid:12)(cid:12)ZT p dσ(cid:12)(cid:12)(cid:12) ≤ NXn=0 ≤(cid:16) NXn=0 αnbσ(−n) (n + 1)1/2αn2(cid:17)1/2(cid:16) NXn=0 (n + 1)1/2(cid:17)1/2 bσ(n)2 . Lemma 4.2 shows that the first factor is dominated by CpD1/2 for some constant C, and the second factor is bounded uniformly in N by (2). Thus, (b) holds. (cid:3) Remark 4.4. The last paragraph of the proof of [12, Theorem 2.3.5] in fact shows that the Cantor measure on the circular middle-thirds Cantor set has finite 1/2-energy, thus we can take σ to be this measure. Let now σ be a measure provided by Lemma 4.3 and let E be the support of σ. Let h : T × E → S2, (ζ1, ζ2) 7→ 1 √2 (ζ1, ζ1ζ2), and observe that the range of h is contained in r−1(E). Define µ to be the pushforward of m × σ by h. We will show that µ satisfies the conclusion of Theorem 1.3. Lemma 4.5. The support of µ is totally null. Proof. Let X = r−1(E). Since E has Lebesgue measure 0 by Lemma 4.3, there exists by the Rudin-Carleson theorem (i.e. the d = 1 case of Theorem 2.2) a function f0 ∈ A(D) which peaks on E. Let f = f0 ◦ r. Then f belongs to A(Bd) and peaks on X, so that X is totally null by Theorem 2.2. Finally, the support of µ is contained in X, hence it is totally null as well. (cid:3) The following lemma finishes the proof of Theorem 1.3. Lemma 4.6. The measure µ is A2-Henkin. Proof. For all m, n ∈ N, we have 2 dµ = 2−(m+n)/2ZTZE ZS2 zm 1 zn ζ m−n 1 ζ n 2 dm(ζ1) dσ(ζ2). This quantity is zero unless m = n, in which case it equals 2−nZE ζ ndσ(ζ). HENKIN MEASURES FOR THE DRURY-ARVESON SPACE 11 On the other hand, Lemma 4.3 shows that there exists f ∈ D1/2 such that hp, fiD1/2 =ZE p dσ d by Lemma 4.2 for all polynomials p. Let g = f ◦ r. Then g belongs to H 2 and it is orthogonal to zn = 2−nhzn, fiD1/2 = 2−nZE 2 unless n = m, in which case = 2−nhr(z)n, giH 2 h(z1z2)n, giH 2 2 ζ n dσ(ζ). 1 zm 2 Consequently, ZS2 ϕ dµ = hMϕ1, giH 2 2 for all polynomials ϕ, and hence for all ϕ ∈ A2, so that µ is A2-Henkin. (cid:3) Theorem 1.3 suggests the following problem, which is deliberately stated somewhat vaguely. Problem 4.7. Find a measure theoretic characterization of Ad-Henkin mea- sures. References 1. Jim Agler and John E. McCarthy, Complete Nevanlinna-Pick kernels, J. Funct. Anal. 175 (2000), no. 1, 111 -- 124. 2. 3. , Nevanlinna-Pick kernels and localization, Operator theoretical methods (Tim- işoara, 1998), Theta Found., Bucharest, 2000, pp. 1 -- 20. , Pick interpolation and Hilbert function spaces, Graduate Studies in Mathe- matics, vol. 44, American Mathematical Society, Providence, RI, 2002. 4. N. Arcozzi, R. Rochberg, and E. Sawyer, Carleson measures for the Drury-Arveson Hardy space and other Besov-Sobolev spaces on complex balls, Adv. Math. 218 (2008), no. 4, 1107 -- 1180. 5. William Arveson, Subalgebras of C ∗-algebras. III. Multivariable operator theory, Acta Math. 181 (1998), no. 2, 159 -- 228. 6. Raphaël Clouâtre and Kenneth R. Davidson, Absolute continuity for commuting row contractions, J. Funct. Anal. 271 (2016), no. 3, 620 -- 641. 7. , Duality, convexity and peak interpolation in the Drury-Arveson space, Adv. Math. 295 (2016), 90 -- 149. 8. Brian Cole and R. Michael Range, A-measures on complex manifolds and some appli- cations, J. Functional Analysis 11 (1972), 393 -- 400. 9. Şerban Costea, Eric T. Sawyer, and Brett D. Wick, The corona theorem for the Drury- Arveson Hardy space and other holomorphic Besov-Sobolev spaces on the unit ball in Cn, Anal. PDE 4 (2011), no. 4, 499 -- 550. 10. Kenneth R. Davidson and David R. Pitts, Nevanlinna-Pick interpolation for non- commutative analytic Toeplitz algebras, Integral Equations Operator Theory 31 (1998), no. 3, 321 -- 337. 11. S. W. Drury, A generalization of von Neumann's inequality to the complex ball, Proc. Amer. Math. Soc. 68 (1978), no. 3, 300 -- 304. 12. Omar El-Fallah, Karim Kellay, Javad Mashreghi, and Thomas Ransford, A primer on the Dirichlet space, Cambridge Tracts in Mathematics, vol. 203, Cambridge University Press, Cambridge, 2014. 13. Jörg Eschmeier, Invariant subspaces for spherical contractions, Proc. London Math. Soc. (3) 75 (1997), no. 1, 157 -- 176. 12 MICHAEL HARTZ 14. G. M. Henkin, The Banach spaces of analytic functions in a ball and in a bicylinder are nonisomorphic, Funkcional. Anal. i Priložen. 2 (1968), no. 4, 82 -- 91. 15. V. Müller and F.-H. Vasilescu, Standard models for some commuting multioperators, Proc. Amer. Math. Soc. 117 (1993), no. 4, 979 -- 989. 16. Vern I. Paulsen and Mrinal Raghupathi, An introduction to the theory of reproduc- ing kernel Hilbert spaces, Cambridge Studies in Advanced Mathematics, Cambridge University Press, 2016. 17. Walter Rudin, Function theory in the unit ball of Cn, Classics in Mathematics, Springer-Verlag, Berlin, 2008, Reprint of the 1980 edition. 18. Orr Shalit, Operator theory and function theory in Drury-Arveson space and its quo- tients, arXiv:1308.1081 (2013). 19. Béla Sz.-Nagy, Ciprian Foias, Hari Bercovici, and László Kérchy, Harmonic analysis of operators on Hilbert space, second ed., Universitext, Springer, New York, 2010. Department of Mathematics, Washington University in St. Louis, One Brookings Drive, St. Louis, MO 63130, USA E-mail address: [email protected]
1903.10734
8
1903
2019-09-11T08:59:46
Statistical unbounded convergence in Banach lattices
[ "math.FA" ]
Several recent papers investigated unbounded and statistical versions of order convergence and topology convergence in locally solid Riesz space. In this papers, we study the statistical unbounded order and topology convergence in Riesz spaces and Banach lattices. In particular, we study the relationship of those convergence and characterize order continuous, KB, reflexive Banach lattices in terms of these convergence.
math.FA
math
STATISTICAL UNBOUNDED CONVERGENCE IN BANACH LATTICES ZHANGJUN WANG, ZILI CHEN, AND JINXI CHEN Abstract. Statistical and unbounded convergence have been in- vestigated in the literature. In this paper, we introduce the statisti- cal unbounded order and topology convergence in Banach lattices. We first present some characterizations and related properties for these convergences. Also we using the convergences to characterize order continuous, K B and reflexive Banach lattices. . A F h t a m [ 8 v 4 3 7 0 1 . 3 0 9 1 : v i X r a 1. introduction Several recent papers investigated unbounded convergence. A net (xα) in Riesz space E is unbounded order convergent (uo-convergent for short) to x if xα − x ∧ u o−→ 0 for all u ∈ E+. Similarly, there are unbounded norm convergence, unbounded absolute weak convergence and unbounded absolute weak* convergence in Banach lattices [8, 10, 13]. For further properties about these convergences, see [5-10,13] for details. Statistical convergence has been studied in functional analysis lit- erature, and also been researched in Riesz spaces and Banach lattices [11, 12]. In the past, statistical convergence was usually defined by sequences. In [14], this convergence could be studied for special nets. Combine these convergences, we introduce so called statistical un- bounded convergence. We first establish the basic properties of statisti- cal unbounded order convergence, statistical unbounded norm conver- gence, statistical unbounded absolute weak convergence and statistical unbounded absolute weak* convergence in Banach lattices, then we using these convergences to describe order continuous Banach lattices, KB-spaces and reflexive Banach lattices. For undefined terminology, notations and basic theory of Riesz spaces and Banach lattices, we refer to [1-4]. Date: September 12, 2019. 2010 Mathematics Subject Classification. 46B42,46A40. Key words and phrases. Riesz space, Banach lattice, statistical unbounded or- der convergence, statistical unbounded norm convergence, statistical unbounded absolute weak convergence, statistical unbounded absolute weak* convergence. 1 2 Z. WANG, Z. CHEN, AND J. CHEN 2. statistical unbounded order convergence We start with the definition of statistical order convergence. Let K ⊂ N, the natural density of K is defined by δ(K) = lim n→∞ 1 n {k ∈ K : k ≤ n} where · denotes the cardinality of the set. A sequence (xn) in Riesz space E is statistical monotone increasing (resp. decreasing) if (xk) is increasing (resp. decreasing) for k ∈ K with δ(K) = 1, we write xn ↑st (resp. xn ↓st). A increasing or decreasing sequence will be called a statistical monotonic sequence, moreover, if supk∈K xk = x (resp. infk∈Kxk = x) for some x ∈ E, then (xn) is statistical monotone convergent to x, we write xn ↑st x (resp. xn ↓st x). A sequence (xn) in E is statistical order convergent to x ∈ E provided that there exists a sequence pn ↓st 0 such that xn − x ≤ pn, we call x the st−ord−−−−→ x for short), it is easy to get statistical order limit of (xn) (xn that δ{n ∈ N : xn − x (cid:2) pn} = 0. For a upwards directed set (D, ≤), if for each β ∈ D, the set {α ∈ D : α ≤ β} is finite, we call it lower finite set. To a nonempty subset A∩Dα A of D, the asymptotic density denoted by δ(A; D) = limα∈D , where Dα denotes the cardinality of {β ∈ D : β ≤ α}. If the net is defined on a lower finite set, we call it lower finite net. We are easy to find that sequence is a special lower finite net. If (xn) (resp. (xα)) is a sequence (resp. lower finite net) such that (xn) (resp. (xα)) satisfies property P for all n (resp. α) except a set of natural density zero, then we say that (xn) (resp. (xα)) satisfies property P for almost all n (resp. α) and we abbreviate this by a.a.n (resp. a.a.α). According to these definition, we can define statistical convergence for lower finite net. In this paper, we use D denote lower finite set. Dα A net (xα) in Riesz space E is unbounded order convergent to x ∈ E if (xα − x ∧ u) is order convergent for any u ∈ E+. Combine these convergences, we introduce a new convergence in Riesz space. Definition 2.1. A sequence (xn) in a Riesz space E is statistical un- bounded order convergent to x ∈ E if (xn − x ∧ u) is statistical order convergent for all u ∈ E+ is statistical order convergent (denoted by st−uo−−−→ x). Respectively, for a lower finite net (xα), if (xα − x ∧ u) xn is statistical order convergent for all u ∈ E+, we call x the statistical unbounded order limit of (xα) (xα st−uo−−−→ x for short). Proposition 2.2. For lower finite net (xα) and (yα) in a Riesz space E, we have the following results: ST-UNBOUNDED-CONVERGENCE 3 (1) the st-uo-limit of (xα) is unique; (2) suppose xα for any a, b ∈ R; st−uo−−−→ x and yα st−uo−−−→ y, then axα + byα st−uo−−−→ ax + by (3) xα (4) if xα st−uo−−−→ x iff (xα − x) st−uo−−−→ 0; st−uo−−−→ x,then xα st−uo−−−→ x. st−uo−−−→ x and Proof. (1) Let (xα) be a sequence in E such that xα st−uo−−−→ y. Then we can find (pα), (qα) such that pα ↓st 0 and qα ↓st 0, xα and a set B ⊂ D with δ(B; D) = 1 such that for β of the index subset B of D, we have: xβ − x ∧ u ≤ pβ ↓ 0, xβ − y ∧ u ≤ qβ ↓ 0. Thus, we get 0 ≤ x − y ∧ u ≤ xβ − x ∧ u + xβ − y ∧ u ≤ pβ + qβ for every β ∈ B and u ∈ E+, which shows that x = y; (2) and (3) is obviously; (4) According to(cid:12)(cid:12)xα−x(cid:12)(cid:12)∧u ≤ xα −x∧u, we have the conclusion. It is similar to Theorem 7 of [11] that Squeeze law holds for statistical (cid:3) unbounded order convergence. Proposition 2.3. Let E be a Riesz space; (xα), (yα) and (zα) be lower st−uo−−−→ x finite nets in E such that xα ≤ yα ≤ zα for all α ∈ D, if xα and zα st−uo−−−→ x, then yα st−uo−−−→ x. The st-uo-convergence preverse disjoint property. Proposition 2.4. Let xα then x ⊥ y. st−uo−−−→ x and δ(α ∈ D : xα ∧ y 6= 0) = 0, st−uo−−−→ x, we have xα st−uo−−−→ x. According to xα ∧ Proof. Since xα y = 0 for a.a.α, and xα∧y st−uo−−−→ x∧y, it follows that x∧y = 0, i.e.x ⊥ y. (cid:3) In general, any subsequence of a convergent sequence is convergent, but it does not hold for st-uo-convergence. Example 2.5. In lp(1 ≤ p < +∞), let (en) be the standard basis, and k xn = X1  en, ei2, n = k2(k ∈ N); ortherwise. 4 Z. WANG, Z. CHEN, AND J. CHEN Since (en) is disjoint sequence, by Corollary 3.6 of [7], so it is unbounded order convergent to 0, moreover, (xn) is statistical unbounded order convergent to 0, but the subsequence (xk2) does not. We also can find that (xn) is not st-order convergent and uo-convergent. In Example 2.5, we know that st-uo-convergence is not necessarily uo-convergence. When does the st-uo-convergence imply unbounded order convergence? Next result show that the relation between two convergences. Proposition 2.6. Let (xα) be a lower finite net in Riesz space E, then (xα) is statistical unbounded order convergent to x ∈ E if and only if there is another lower finite net (yα) such that xα = yα for a.a.α and which is unbounded order convergent to the same limit x. Proof. ⇒ Following from definition. ⇐ Assume that xα = yα for a.a.α and yα uo−→ x. Then there exists a lower finite (pα) ↓ 0 (hence pα ↓st 0) and yα − x ∧ u ≤ pα for every u ∈ E+. Thus, we can write {α ≤ β : xβ − x ∧ u (cid:2) pβ} ⊂ {α ≤ β : xβ 6= yβ} ∪ {α ≤ β : yβ − x ∧ u (cid:2) pβ} Since yα have δ{α ∈ D : xα − x ∧ u (cid:2) pα} = 0, i.e: xα uo−→ x, the second set on the right side is empty. Hence we (cid:3) st−uo−−−→ x. For monotonic lower finite net, we have the further result. Proposition 2.7. For monotonic lower finite net (xα) in Riesz space E, xα uo−→ x iff xα st−uo−−−→ x. st−uo−−−→ x, then there exists a B ⊂ D, Proof. Assume that xα ↑ and xα all β ∈ B and δ(B; D) = 1 such that (xβ) is a subnet of (xα) and supβ∈B xβ = x. For all γ ∈ Bc, since xγ ↑, so we have xγ ↑≤ x, factly, if xγ0 > x, then there exists a η0 such that xη0 ≥ xγ0 > x, it is contradiction. And since for any γ we can find a β such that xβ ≤ xγ ≤ x, by supβ∈B xβ = x, it follow that supγ∈B c xγ = x, hence xγ (cid:3) uo−→ x for γ ∈ Bc, so xα uo−→ x. Let (xn) be a lower finite net in Riesz space E, (xα) is called statistical order bounded if there exists an order interval [x, y] such that δ{α ∈ D : xα /∈ [x, y]} = 0. It is clear that every order bounded lower finite net is statistical order bounded, but the converse is not true in general. ST-UNBOUNDED-CONVERGENCE 5 Example 2.8. In l∞, let (en) be the unit vectors, and xn =(nen, n = k2(k ∈ N); ortherwise. en, (xn) is statistical order bounded sequence, but it is not order bounded. According to Example 2.5, st-uo-convergence is not necessarily sta- tistical order convergence. It is natural to ask when the st-uo-convergence implies st-ord-convergence? Next result shows that the st-uo-convergence is the same as st-ord-convergence for statistical order bounded lower finite net. Proposition 2.9. Let (xα) be a statistical order bounded lower finite net in Riesz space E and xα st−uo−−−→ x, then xα st−ord−−−−→ x. st−uo−−−→ x, therefore, xα uo−→ x for a.a.α. And since (xα) Proof. Since xα is statistical order bounded lower finite net, then there exists u ∈ E+ such that xα ≤ u for a.a.α, then fix the u, we have xα − x ∧ 2u = xα − x o−→ x for a.a.α, therefore, xα (cid:3) st−ord−−−−→ x. A norm on a Banach lattice E is called order continuous if kxαk → 0 for xα ↓ 0. A Banach lattice E is called a KB-space, if every monotone bounded sequence is convergent. A lower finite net (xα) is said to be statistical unbounded order Cauchy, if (xα − xβ) st-uo-convergent to 0. Let E be a Banach space and (xα) be a lower finite net in E, (xα) is statistical norm bounded if for any C ∈ R+ there exists a λ > 0 such that δ{α ∈ D : kλxαk > C} = 0. In Example 2.8, it is clear that statistical norm bounded is not necessarily bounded. According to Theorem 4.7 of [5], we decribe KB spaces by st-uo- convergence. Theorem 2.10. Let E be an order continuous Banach lattice. The following are equivalent: (1) E is KB space; (2) every norm bounded uo-Cauchy sequence in E is uo-convergent; (3) every statistical norm bounded st-uo-Cauchy sequence in E is st-uo-convergent. Proof. (1) ⇒ (2) by Theorem 4.7 of [5] and (2) ⇒ (3) is clearly; 6 Z. WANG, Z. CHEN, AND J. CHEN (3) ⇒ (1) Assume that E is not KB space, then E contains a sub- lattice lattice isomorphic to c0. Let (xn) be a sequence in c0, and n = k2(k ∈ N); nen, n ei, ortherwise. X1 xn =  (xn) is statistical norm bounded and st-uo-Cauchy in E but it is not st-uo-convergent. (cid:3) A sequence (xn) in Banach space E is statistical norm convergent to x ∈ E provided that δ{n ∈ N : kxn − xk ≥ ǫ} = 0 for all ǫ > 0. st−n−−−→ x. We call x the statistical norm limit of (xn). A We write xn sequence (xn) in E is statistical weak convergent to x ∈ E provided that δ{n ∈ N : f (xn − x) ≥ ǫ} = 0 for all ǫ > 0, f ∈ E ′. we write st−w−−−→ x. We call x the statistical weak limit of (xn). A sequence xn (x′ is statistical weak* convergent to x′ ∈ E ′ provided that n) in E ′ δ{n ∈ N : (x′ n − x′)(x) ≥ ǫ} = 0 for all ǫ > 0, x ∈ E. we write x′ n st−w∗ −−−→ x. We call x′ the statistical weak limit of (x′ Weak convergent sequence is st-w-convergent and statistical norm bounded, the converse is not hold in general by Example 2.5 for (1 < p < +∞). Now, we consider a Banach space with two properties. Let E be a Banach space, E is said to have statistical weak convergent property (ST W C property, for short) if (xn) is statistical norm bounded and w−→ 0 for all (xn) ⊂ E. E is said to have xn statistical weak* convergent property (ST W ∗C property, for short) if w∗ (x′ −→ 0. Using these properties, we investigate order continuous Banach lattices by st-uo-convergence. n) is statistical norm bounded and x′ st−w∗ −−−→ 0 in E ′, then x′ n st−w−−−→ 0 in E, then xn n). n Theorem 2.11. Let E be an Banach lattice: (1) E has order continuous norm; (2) for any statistical norm bounded lower finite net (x′ st−uo−−−→ 0,then x′ x′ α (3) for any statistical norm bounded lower finite net (x′ st−uo−−−→ 0,then x′ x′ α (4) for any statistical norm bounded sequence (x′ st−σ(E −−−−−−−→ 0; st−σ(E −−−−−−→ 0; ,E) ,E) α α ′ ′ α) in E ′,if α) in E ′,if n) in E ′,if x′ n st−uo−−−→ 0,then x′ n ′ st−σ(E −−−−−−→ 0; ,E) ST-UNBOUNDED-CONVERGENCE (5) for any statistical norm bounded sequence (x′ n) in E ′,if x′ n 7 st−uo−−−→ 0,then x′ n ′ st−σ(E −−−−−−−→ 0. ,E) Then (1) ⇒ (2) ⇔ (3) ⇒ (4) ⇔ (5) hold, and in addition, if E has ST W ∗C property, all are equivalent. Proof. Since x′ α (4) ⇔ (5). st−uo−−−→ 0 ⇔ x′ α st−uo−−−→ 0, it is clear that (2) ⇔ (3) ⇒ (1) ⇒ (2) Assmue that E has order continuous norm, by Theorem α) and β) which asymptotic 2.1 of [6], for any statistical norm bounded lower finite net (x′ x′ α st−uo−−−→ 0 in E ′, there exists a bounded subnet (x′ density of index set is 1 satisfying x′ β uo−→ 0,then x′ β ′ ,E) σ(E −−−−→ 0, hence ′ ,E) st−σ(E x′ −−−−−−→ 0. α (4) ⇒ (1) For any norm bounded disjoint sequence (x′ n) in E ′, obvi- ously, it is statistical norm bounded and uo-null, hence it is st-uo-null, ′ ,E) st−σ(E −−−−−−→ 0. Since E has ST W ∗C property, (x′ so x′ n) is weak* con- n vergent to zero. By Corollary 2.4.3 of [4], E has order continuous norm. (cid:3) 3. statistical unbounded convergence in Banach lattices A net (xα) in Banach lattice E is unbounded norm (resp. absolute weak, absolute weak*) convergent to x ∈ E if (xα − x ∧ u) is norm (resp. weak, weak*) convergent for any u ∈ E+ [8, 10, 13]. We study the statistical unbounded convergence in Banach lattice. Definition 3.1. A lower finite net (xα) in Banach lattice E is statistical unbounded norm convergent to x ∈ E provided that δ{α ∈ D :(cid:13)(cid:13)xα − x ∧ u(cid:13)(cid:13) ≥ ǫ} = 0 for all ǫ > 0 and u ∈ E+, we write xα st−un−−−→ x and call x the statistical unbounded norm limit of (xα). A lower finite net (xα) in Banach lattice E is statistical unbounded absolute weak convergent to x ∈ E provided that δ{α ∈ D : f (xα − x ∧ u) ≥ ǫ} = 0 for all ǫ > 0, st−uaw−−−−→ x and call x the statistical f ∈ E ′ unbounded absolute weak limit of (xα). A lower finite net (x′ α) in dual Banach lattice E ′ is statistical unbounded absolute weak* convergent to x′ ∈ E provided that δ{α ∈ D : (x′ α − x′ ∧ u′)(x) ≥ ǫ} = 0 for all st−uaw∗ ǫ > 0, x ∈ E+ and u′ ∈ E+, we write xα −−−−−→ x and call x′ the statistical unbounded absolute weak* limit of (x′ + and u ∈ E+, we write xα α). 8 Z. WANG, Z. CHEN, AND J. CHEN Proposition 3.2. For a lower finite net (xα) we have the following results (the st-uaw-convergence and st-uaw∗-convergence have analogic conclusions): (1) st-un-limits are unique; (2) suppose xα for any a, b ∈ R; st−un−−−→ x and yα st−un−−−→ y, then axα +byα st−un−−−→ ax+by (3) xα (4) if xα st−un−−−→ x iff (xα − x) st−un−−−→ 0; st−un−−−→ x,then xα st−un−−−→ x; Some subsequences of statistical unbounded convergent sequence is not necessarily convergent. Example 3.3. In Lp[0, 1](1 < p < +∞), let 1 2n−1 ]; 1 2n−1 ]. 1 2n , 1 2n , 0, t /∈ [ 1, t ∈ [ and xn(t) =  yn = X1  2nxn, n ixi, n = k2(k ∈ N); ortherwise. Since (xn) is disjoint sequence in order continuous Banach lattice, by Corollary 3.6 of [7] and Proposition 2.5 of [8], so it is unbounded norm (resp. absolute weak, absolute weak*) convergent to 0, moreover, (yn) is statistical unbounded norm (resp. absolute weak, absolute weak*) convergent to 0, but the subsequence (yk2) does not. We also can find that (yn) is not statistical norm (resp. absolute weak, absolute weak*) convergent and unbounded norm (resp. absolute weak, absolute weak*) convergent. In Example 3.3, we know that st-un(resp. uaw, uaw∗)-convergence is not necessarily un(resp. uaw, uaw∗)-convergence. When does the st-un(resp. uaw, uaw∗)-convergence imply un(resp. uaw, uaw∗)- convergence? It is similar to Proposition 2.6 and 2.7, we have the following results. Lemma 3.4. Let (xα) be a lower finite net in Banach lattice E. Then (xα) is st-un(uaw, uaw∗)-convergent to x ∈ E if and only if there is another lower finite net (yα) such that xα = yα for a.a.α and which is un(uaw, uaw∗)-convergent to the same limit x. Proposition 3.5. For a monotonic lower finite net (xα) in Banach lattice E, we have the following result: ST-UNBOUNDED-CONVERGENCE 9 (1) un-convergence and st-un-convergence agree. (2) uaw-convergence and st-uaw-convergence agree. (3) uaw∗-convergence and st-uaw∗-convergence agree. un−→ x. Since xβ − xα ∈ E+ for all Proof. (1) Suppose that xα ↑ and xα β ≥ α, because lattice operation of un-convergence is continuous, so un−→ (x − xα)+, hence x − xα ≥ 0 for all α. (xβ − xα)(β) = (xβ − xα)+ (β) If x is not supremum, then assume that y ∈ E such that xα ≤ y ≤ x for all α ∈ D, it follows that xα = xα ∧ y un−→ x ∧ y = y Hence, y = x, so x = supα∈D xα. According to proof of Proposition 2.7, if a lower finite net (xα) in a Banach lattice E is increasing and statistical unbounded norm conver- get to x ∈ E, then x = supα∈D xα. Now, assume that xα st−un−−−→ x and xα ↑, we have x = supα∈D xα. Hence there exists a B ⊂ D, all β ∈ B and δ(B; D) = 1 such that (xβ) un−→ x. For any ǫ > 0, is a subnet of (xα) with supβ∈B xβ = x and xβ u ∈ E+, choose a β0 such that (cid:13)(cid:13)xβ0 − x ∧ u(cid:13)(cid:13) < ǫ. Then, for any α > β0, we have 0 ≤ x − xα ∧ u ≤ x − xβ0 ∧ u This shows Hence, xα un−→ x. (cid:13)(cid:13)x − xα ∧ u(cid:13)(cid:13) ≤(cid:13)(cid:13)x − xβ0 ∧ u(cid:13)(cid:13) < ǫ The proof of (2) and (3) is similar. (cid:3) In an order continuous Banach lattice, it is easy to check that st-uo- convergence implies st-un-convergence. So we can describe KB space by statistical unbounded norm convergence. The following result is similar to Theorem 2.10. Theorem 3.6. Let E be an order continuous Banach lattice. The following are equivalent: (1) E is KB space; (2) every norm bounded uo-Cauchy sequence in E is un-convergent; (3) every statistical norm bounded st-uo-Cauchy sequence in E is st-un-convergent. According to Example 3.3, st-un (resp. st-uaw, st-uaw∗)-convergence is not necessarily st-n (resp. st-aw, st-aw∗)- convergence. It is natural to ask when the st-un (resp. st-uaw, st-uaw∗)-convergence imply st-n (resp. st-aw, st-aw∗)- convergence? It is clear that the st-un (resp. 10 Z. WANG, Z. CHEN, AND J. CHEN st-uaw, st-uaw∗)-convergence is the same as st-n(resp. st-aw, st-aw∗)- convergence for statistical order bounded lower finite net. By Theorem 2.3 of [9], the following holds. Proposition 3.7. Let E be a Banach lattice with strong order unit, then the lower finite net (xα) in E is st-norm convergent if and only if it is st-un-convergent. By Lemma 2 of [9], every disjoint lower finite net is uaw-null, hence it is st-uaw-null, moreover it is st-uaw∗-null. Using the result, we can describe order continuous and reflxive Banach lattice by statistical unbounded convergence. Theorem 3.8. Let E be an Banach lattice: α) in E ′,if α) in E ′,if st−uaw∗ −−−−−→ st−uaw∗ −−−−−→ (1) E has order continuous norm; (2) for any statistical norm bounded lower finite net (x′ st−uaw∗ x′ −−−−−→ 0,then x′ α α (3) for any statistical norm bounded lower finite net (x′ st−uaw∗ x′ −−−−−→ 0,then x′ α α (4) for any statistical norm bounded sequence (x′ st−σ(E −−−−−−−→ 0; st−σ(E −−−−−−→ 0; ,E) ,E) ′ ′ n) in E ′,if x′ n 0,then x′ n ′ st−σ(E −−−−−−→ 0; ,E) (5) for any statistical norm bounded sequence (x′ n) in E ′,if x′ n 0,then x′ n ′ st−σ(E −−−−−−−→ 0. ,E) Then (1) ⇒ (2) ⇔ (3) ⇒ (4) ⇔ (5) hold, and in addition, if E has ST W ∗C property, all are equivalent. Proof. Since x′ α (3) ⇒ (4) ⇔ (5). st−uaw∗ −−−−−→ 0 ⇔ x′ α st−uaw∗ −−−−−→ 0, it is clear that (2) ⇔ (1) ⇒ (2) Suppose that E has order continuous norm, by Theorem α) β) which asymptotic 2.15 in [13], for any statistical norm bounded lower finite net (x′ and x′ α st−uaw∗ −−−−−→ 0 in E ′, there exists a subnet (x′ density of index set is 1 satisfying x′ β uaw∗ −−−→ 0,then x′ β ′ ,E) σ(E −−−−→ 0, hence ′ ,E) st−σ(E x′ −−−−−−→ 0. α (4) ⇒ (1) For any norm bounded disjoint sequence (x′ n) in E ′, from the above, we know it is statistical norm bounded and st-uaw∗-null, ′ ,E) st−σ(E −−−−−−→ 0. Since E has ST W ∗C property, (x′ so x′ n) is weak* con- n vergent to zero. By Corollary 2.4.3 of [4], E has order continuous norm. (cid:3) The following result is the dual version of Theorem 3.8. ST-UNBOUNDED-CONVERGENCE 11 Theorem 3.9. Let E be an Banach lattice: (1) E ′ has order continuous norm; (2) for any statistical norm bounded lower finite net (x′ st−uaw−−−−→ 0,then xα (3) for any statistical norm bounded lower finite net (x′ st−uaw−−−−→ 0,then xα (4) for any statistical norm bounded sequence (xn) in E,if xn st−σ(E,E −−−−−−−→ 0; st−σ(E,E −−−−−−→ 0; ) ) ′ ′ xα xα 0,then xn ′ st−σ(E,E −−−−−−→ 0; ) (5) for any statistical norm bounded sequence (xn) in E,if xn 0,then xn ′ st−σ(E,E −−−−−−−→ 0. ) α) in E ′,if α) in E ′,if st−uaw−−−−→ st−uaw−−−−→ Then (1) ⇒ (2) ⇔ (3) ⇒ (4) ⇔ (5) hold, and in addition, if E has ST W C property, all are equivalent. Corollary 3.10. For Banach lattice E, then st-uaw-convergence and st-w-convergence are agree for sequence, if one of the following condi- tions is valid: (1) E is AM-space; (2) E is atomic, E and E ′ is order continuous. Proof. Since AM-space and atomic Banach lattice with order continu- ous norm have lattice operation weakly sequence continuous property, so st-w-convergence implies st-uaw-convergence. For the reverse, the dual space of AM-space is AL-space which is order continuous, hence st-uaw-convergence implies st-w-convergence. (cid:3) Theorem 3.11. For Banach lattice E, the following are equivalent: (1) E is reflexive; (2) every statistical norm bounded st-uaw-Cauchy sequence in E is statistical weakly convergent; (3) every statistical norm bounded st-un-Cauchy sequence in E is statistical weakly convergent; (4) every statistical norm bounded st-uo-Cauchy sequence in E is statistical weakly convergent. Proof. (1) ⇒ (2) Since E is reflexive, then E and E ′ are KB space, so we have every statistical norm bounded st-uaw-Cauchy sequence in E is statistical weakly Cauchy. Since E is KB-space, so it is statistical weakly convergent. (2) ⇒ (1) We show that E contains no lattice copies of c0 or l1. Suppose that E contains a sublattice isomorphic to l1. The (xn) of l1 which is in Example 2.5, it is st-uaw convergent to zero and statistical 12 Z. WANG, Z. CHEN, AND J. CHEN norm bounded but it is not statistical weakly convergence. Suppose that E contains a sublattice isomorphic to c0, the (xn) of c0 which is in the proof of Theorem 2.10, it is st-uaw-Cauchy sequence, but it is not weakly convergence, so we have the conclusion. (2) ⇒ (3) Since un-convergence implies uaw-convergence, so we have the conclusion. (3) ⇒ (1) E contains no sublattice isomorphic to c0 or l1, the proof is similarily to (2) ⇒ (1). (1) ⇔ (4) the proof is similar to (1) ⇔ (2). (cid:3) References [1] W. A. J. Luxemburg and A. C. Zaanen, Riesz Spaces. I, North-Holland, Am- sterdam, 1971. [2] W. A. J. Luxemburg and A. C. Zaanen, Riesz Spaces. II, North-Holland, Amsterdam, 1983. [3] C. D. Aliprantis and O. Burkinshaw, Positive Operators, Springer, 2006. [4] P. Meyer-Nieberg, Banach lattices, Universitext, Springer-Verlag, Berlin, 1991. [5] N. Gao and F. Xanthos, Unbounded order convergence and application to martingales without probability, J. Math. Anal. Appl. 415(2), 2014, 931-947. [6] N. Gao, Unbounded order convergence in dual spaces, J. Math. Anal. Appl. 419(1), 2014, 931-947. [7] N. Gao, V. G. Troitsky and F. Xanthos, Uo-convergence and its applications to Ces`aro means in Banach lattices, Israel J. Math. 220(2), 2017, 649-689. [8] Y. Deng, M. O'Brien and V. G. Troitsky, Unbounded norm convergence in Banach lattices, Positivity.21(3),2017,963-974. [9] M. Kandi´c, M. A. A. Marabeh and V. G. Troitsky, Unbounded norm topology in Banach lattices, J. Math. Anal. Appl. 451(1), 2017, 259-279. [10] Zabeti, Omid, Unbounded absolute weak convergence in Banach lattices, Pos- itivity 22.2 (2018): 501-505. [11] Sencimen, Celaleddin, and Serpil Pehlivan, Statistical order convergence in Riesz spaces, Slovaca J. Math. 62(2), 2012, 257-270. [12] Xue, Xuemei, and Jian Tao, Statistical Order Convergence and Statistically Relatively Uniform Convergence in Riesz Spaces, Journal of Function Spaces 2018. [13] Z.Wang ,Z.Chen and J.Chen , Unbounded absolute weak* convergence in dual Banach lattices, arXiv:1903.02168v2. [14] AR. Murugan, J. Dianavinnarasi and C. Ganesa Moorthy, Statistical conver- gence of nets through directed sets , arXiv:1907.00777. ST-UNBOUNDED-CONVERGENCE 13 School of Mathematics, Southwest Jiaotong University, Chengdu, Sichuan, China, 610000. E-mail address: [email protected] School of Mathematics, Southwest Jiaotong University, Chengdu, Sichuan, China, 610000. E-mail address: [email protected] School of Mathematics, Southwest Jiaotong University, Chengdu, Sichuan, China, 610000. E-mail address: [email protected]
1604.07271
1
1604
2016-04-25T14:21:03
A metric interpretation of reflexivity for Banach spaces
[ "math.FA", "math.MG" ]
We define two metrics $d_{1,\alpha}$ and $d_{\infty,\alpha}$ on each Schreier family $\mathcal{S}_\alpha$, $\alpha<\omega_1$, with which we prove the following metric characterization of reflexivity of a Banach space $X$: $X$ is reflexive if and only if there is an $\alpha<\omega_1$, so that there is no mapping $\Phi:\mathcal{S}_\alpha\to X$ for which $$ cd_{\infty,\alpha}(A,B)\le \|\Phi(A)-\Phi(B)\|\le C d_{1,\alpha}(A,B) \text{ for all $A,B\in\mathcal{S}_\alpha$.}$$ Secondly, we prove for separable and reflexive Banach spaces $X$, and certain countable ordinals $\alpha$ that $\max(\text{ Sz}(X),\text{ Sz}(X^*))\le \alpha$ if and only if $({\mathcal S}_\alpha, d_{1,\alpha})$ does not bi-Lipschitzly embed into $X$. Here $\text{Sz}(Y)$ denotes the Szlenk index of a Banach space $Y$.
math.FA
math
A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES P. MOTAKIS AND TH. SCHLUMPRECHT Abstract. We define two metrics d1,α and d∞,α on each Schreier family Sα, α < ω1, with which we prove the following metric characterization of reflexivity of a Banach space X: X is reflexive if and only if there is an α < ω1, so that there is no mapping Φ : Sα → X for which cd∞,α(A, B) ≤ kΦ(A) − Φ(B)k ≤ Cd1,α(A, B) for all A, B ∈ Sα. Secondly, we prove for separable and reflexive Banach spaces X, and certain countable ordinals α that max(Sz(X), Sz(X ∗)) ≤ α if and only if (Sα, d1,α) does not bi-Lipschitzly embed into X. Here Sz(Y ) denotes the Szlenk index of a Banach space Y . 6 1 0 2 r p A 5 2 ] . A F h t a m [ 1 v 1 7 2 7 0 . 4 0 6 1 : v i X r a Contents Introduction and statement of the main results 1. 2. Regular Families, Schreier Families and Fine Schreier Families 2.1. Regular Families in [N]<ω 2.2. The Schreier families 2.3. The fine Schreier families 2.4. Families indexed by subsets of [N]<ω 3. Repeated averages on Schreier sets 4. Trees and their indices 5. The Szlenk index 6. Estimating certain convex combinations of blocks using the Szlenk index 7. Two metrics on Sα, α < ω1 8. The Szlenk index and embeddings of (Sα, d1,α) into X 9. Refinement argument 10. Some further observation on the Schreier families 10.1. Analysis of a maximal set B in Sβγ 10.2. Components of a set A in Sβγ. 10.3. Special families of convex combinations 11. Conclusion of the proof of Theorems A and C 12. Final comments and open questions References 2 5 5 7 11 12 14 20 22 29 31 35 37 45 45 46 49 50 62 63 2000 Mathematics Subject Classification. 46B03, 46B10, 46B80. The second named author was supported by the National Science Foundation under Grant Numbers DMS -- 1160633 and DMS -- 1464713. 1 2 P. MOTAKIS AND TH. SCHLUMPRECHT 1. Introduction and statement of the main results s : s = 1, 2, . . . , t}, for σ = (σs)i s=1 6= σ′ = (σ′ s)j Bn =Sn infinite binary tree B∞ =S∞ In this paper we seek a metric characterization of reflexivity of Banach spaces. By a metric characterization of a property of a Banach space we mean a characterization which refers only to the metric structure of that space but not its linear structure. In 1976 Ribe [32] showed that two Banach spaces, which are uniformly homeomorphic, have uniformly linearly isomorphic finite-dimensional subspaces. In particular this means that the finite dimensional or local properties of a Banach space are determined by its metric structure. Based on this result Bourgain [11] suggested the "Ribe Program", which asks to find metric descriptions of finite-dimensional invariants of Banach spaces. In [11] he proved the following characterization of super reflexivity: a Banach space X is super reflexive if and only if the binary trees Bn of length at most n, n ∈ N, endowed with their graph metric, are not uniformly bi-Lipschitzly embedded into X. A binary tree of length at most n, is the set k=0{−1, 1}k, with the graph or shortest path metric d(σ, σ′) = i + j − 2 max{t ≥ 0 : σs = σ′ k=0{−1, 1}k. A new and shorter proof of this result was recently obtained by Kloeckner in [19]. In [7] Baudier extended this result and proved that a Banach space X is super reflexive, if and only if the n=0{−1, 1}n (with the graph distance) does not bi-Lipschitzly embed into X. Nowadays this result can be deduced from Bourgain's result and a result of Ostrovskii's [28, Theorem 1.2] which states that a locally finite metric space A embeds bi- Lipschitzly into a Banach space X if all of its finite subsets uniformly bi-Lipschitzly embed into X. In [18] Johnson and Schechtman characterized superflexivity, using the Diamond Graphs, Dn, n ∈ N, and proved that Banach space X is super reflexive if and only if the Dn, n ∈ N, do not uniformly bi-Lipschitzly embed into X. There are several other local properties, i.e., properties of the finite dimensional subspaces of Banach spaces, for which metric characterizations were found. The following are some examples: Bourgain, Milman and Wolfson [12] characterized having non trivial type using Hammond cubes (the sets Bn, together with the ℓ1-norm), and Mendel and Naor [21, 22] present metric characterizations of Banach spaces with type p, 1 < p ≤ 2, and cotype q, 2 ≤ q < ∞. For a more extensive account on the Ribe program we would like refer the reader to the survey articles [6, 23] and the book [29]. s=1 in Sn Instead of only asking for metric characterizations of local properties, one can also ask for metric characterizations of other properties of Banach space, properties which might not be determined by the finite dimensional subspaces. A result in this direction was obtained by Baudier, Kalton and Lancien in [9]. They showed that a reflexive Banach space X has a renorming, which is asymptoticaly uniformly convex (AUC) and asymptoticaly uniformly smooth (AUS), if and only if the countably branching trees of length n ∈ N, Tn do not k=0 Nk, together with the graph metric, i.e., d(a, b) = i+j −max{t ≥ 0 : as = bs, s = 1, 2 . . . t}, for a = (a1, a2, . . . ai) 6= b = (b1, b2, . . . bj) in T n. Among the many equivalent conditions for a reflexive Banach space X to be (AUC)- and (AUS)-renormable (see [25]) one of them states that Sz(X) = Sz(X∗) = ω, where Sz(Z) denotes the Szlenk index of a Banach space Z (see Section 5 for the definition and properties of the Szlenk index). In [13] Dilworth, Kutzarova, Lancien and Randrianarivony, showed that a separable Banach space X is reflexive and (AUC)- and (AUS)-renormable if and only X admits an equivalent norm for which X has Rolewicz's β-property. According to [20] a uniformly bi-Lipschitzly into X. Here Tn =Sn A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 3 Banach space X has Rolewicz's β-property if and only if : n ≥ 1o : (xn)∞ 2 n=1 ⊂ BX , sep(cid:2)(xn)(cid:3) ≥ t, x ∈ BXo > 0, ¯βX (t) = 1 − supn infn kx − xnk for all t > 0, where sep(cid:2)(zn)(cid:3) = inf m6=n kzm − znk, for a sequence (zn) ⊂ X. The function ¯βX is called the β-modulus of X. Using the equivalence between the positivity of the β-modulus and the property that a separable Banach space is reflexive and (AUC)- and (AUS)-renormable, Baudier [8] was able to establish a new and shorter proof of the above cited result from [9]. Metric descriptions of other non-local Banach space properties, for example the Radon-Nikod´ym Property, can be found in [30]. k · ks denotes the summing norm, i.e., kzks = supk∈N(cid:12)(cid:12)Pk In our paper we would like to concentrate on metric descriptions of the property that a Banach space is reflexive, and subclasses of reflexive Banach spaces. In [30], Ostrovskii established a submetric characterization of reflexivity. Let T be the set of all pairs (x, y) in ℓ1 × ℓ1, for which kx − yk1 ≤ 2kx − yks, where k · k1 denotes the usual norm on ℓ1 and [30, Theorem 3.1] states that a Banach space X is not reflexive if and only if there is a map f : ℓ1 → X and a number 0 < c ≤ 1, so that ckx − yk1 ≤ kf (x) − f (y)k ≤ kx − yk1 for all (x, y) ∈ T . In Section 12 we will formulate a similar result, using a discrete subset of ℓ1 ×ℓ1, witnessing the same phenomena. Recently Proch´azka [31, Theorem 3] formulated an interesting metric description of reflexivity. He constructed a unifomly discrete metric space MR with the following properties: If MR bi-Lipschitzly embeds into a Banach space X with distortion less than 2, then X is non reflexive. The distortion of a bi-Lipschitz embedding f of one metric into another is the product of the Lipschitz constant of f and the Lipschitz constant of f −1. Conversely, if X is non reflexive, then there exists a renorming · of X, so that MR embeds into (X, · ) isometrically. j=1 zj(cid:12)(cid:12), for z = (zj) ∈ ℓ1. Our paper has the goal to find a metric characterization of reflexivity. An optimal result would be a statement, similar to Bourgain's result, of the form "all members of a certain family (Mi)i∈I of metric spaces embed uniformly bi-Lipschitzly into a space X if and only if X is not reflexive". In the language, introduced by Ostrovskii [25], this would mean that (Mi)i∈I is a family of test spaces for reflexivity. Instead, our result will be of the form (see Theorem A below),"there is a family of sets (Mi)i∈I , and for i ∈ I, there are metrics d∞,i and d1,i on Mi, with the property, that a given space X is non reflexive if and only if there are injections Φi : Mi → X and 0 < c ≤ 1 so that cd∞,i(x, y) ≤ kx − yk ≤ d1,i(x, y), for all x, y ∈ Mi". In Section 12 we will discuss the difficulties, in arriving to a characterization of reflexivity of the first form. Nevertheless, if we restrict ourselves to the class of reflexive spaces we arrive to a metric characterization for the complexity of a given space, which we measure by the Szlenk index, using test spaces. Roughly speaking, the higher the Szlenk index is of a given Banach space, the more averages of a given weakly null sequence one has to take to arrive to a norm null sequence. For a precise formulation of this statement we refer to Theorem 5.3. For the class of separable and reflexive spaces we will introduce an uncountable family of metric spaces (Mα)α<ω1 for which we will show that the higher the complexity of a given reflexive and separable space X or its dual X∗ is, the more members of (Mα)α<ω1 can be uniformly bi-Lipschitzly embedded into X. The statements of our main results are as follows. The definition of the Schreier families Sα ⊂ [N]<ω, for α < ω1, will be recalled in Section 2, the Szlenk index Sz(X) for a Banach space X in Section 5, and the two metrics d1,α and d∞,α on Sα will be defined in Section 7. The statements of our main results are as follows. 4 P. MOTAKIS AND TH. SCHLUMPRECHT Theorem A. A separable Banach space X is reflexive if and only if there is an α < ω1 for which there does not exist a map Φ : Sα → X, with the property that for some numbers C ≥ c > 0 (1) cd∞,α(A, B) ≤ kΦ(A) − Φ(B)k ≤ Cd1,α(A, B) for all A, B ∈ Sα. Definition 1.1. Assume that X is a Banach space, α < ω1, and C ≥ c > 0. We call a map Φ : Sα → X, with the property that for all A, B ∈ Sα, (2) cd∞,α(A, B) ≤ kΦ(A) − Φ(B)k ≤ Cd1,α(A, B) a c-lower-dα,∞ and C-upper-dα,1 embedding of Sα into X. If A is a subset of Sα, and Φ : A → X, is a map which satisfies (2) for all A, B ∈ A, we call it a c-lower-dα,∞ and C-upper-dα,1 embedding of A into X. Our next result extends one direction (the "easy direction") of the main result of [9] to spaces with higher order Szlenk indices. As in [9] reflexivity is not needed here. Theorem B. Assume that X is a separable Banach space and that max(cid:0)Sz(X), Sz(X∗)(cid:1) > ωα, for some countable ordinal α. Then (Sα, d1,α) embeds bi-Lipschitzly into X and X∗. We will deduce one direction of Theorem A from James's characterization of reflexive Banach spaces [16], and show that for any non reflexive Banach space X and any α < ω1 there is a map Φα : Sα → X which satisfies (1). The converse will follow from the following result. Theorem C. Assume that X is a reflexive and separable Banach space. Let ξ < ω1 and put β = ωωξ . If for some numbers C > c > 0 there exists a c-lower-dα,∞ and C-upper-dα,1 embedding of Sβ2 into X, then Sz(X) > ωβ or Sz(X∗) > β. Theorem C, and thus the missing part of Theorem A, will be shown in Section 11, Theorem 11.6. Combining Theorems B and C, we obtain the following characterization of certain bounds of the Szlenk index of X and its dual X∗. This result extends the main result of [9] to separable and reflexive Banach spaces with higher order Szlenk indices. Corollary 1.2. Assume that ω < α < ω1 is an ordinal for which ωα = α. Then the following statements are equivalent for a separable and reflexive space X a) max(Sz(X), Sz(X∗)) ≤ α, b) (Sα, d1) is not bi-Lipschitzly embeddable into X. Corollary 1.2 and a result in [26] yield the following corollary. We thank Christian Rosendal who pointed it out to us. Corollary 1.3. If α < ω1, with α = ωα, the class of all Banach spaces X for which max(Sz(X), Sz(X∗)) ≤ α is Borel in the Effros-Borel structure of closed subspaces of C[0, 1]. A proof of Corollaries 1.2 and 1.3 will be presented at the end of Section 11. For the proof of our main results we will need to introduce some notation and to make several preliminary observations. The reader who is at first only interested to understand our main results will only need the definition of the Schreier families Sα, α < ω1, given in Subsection 2.2, the definition of repeated averages stated in the beginning of of Section 3, and the definition of the two metrics d1,α and d∞,α on Sα introduced in Section 7. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 5 2. Regular Families, Schreier Families and Fine Schreier Families In this section we will first recall the definition of general Regular Subfamilies of [N]<ω. Then we recall the definition of the Schreier Families Sα and the Fine Schreier Families Fβ,α, α ≤ β < ω1 [1]. The recursive definition of both families depends on choosing for every limit ordinal a sequence (αn), which increases to α. In order to ensure that later our proof will work out, we will need (αn) to satisfy certain conditions. 2.1. Regular Families in [N]<ω. For a set S we denote the subsets, the finite subsets, and the countably infinite subsets by [S], [S]<ω, and [S]ω. We always write subsets of N in increasing order. Thus, if we write A = {a1, a2, . . . an} ∈ [N]<ω, or A = {a1, a2, . . .} ∈ [N]ω we always assume that a1 < a2 < . . .. Identifying the elements of [N] in the usual way with elements of {0, 1}ω , we consider on [N] the product topology of the discrete topology on {0, 1}. Note that it follows for a sequence (An) ⊂ [N]<ω and A ∈ [N]<ω, that (An) converges to A if and only if for all k ≥ max A there is an m so that An ∩ [1, k] = A, for all n ≥ m. For A ∈ [N]<ω and B ∈ [N] we write A < B, if max(A) < min(B). As a matter of convention we put max(∅) = 0 and min(∅) = ∞, and thus A < ∅ and ∅ > A is true for all A ∈ [N]<ω. For m ∈ N we write m ≤ A or m < A, if m ≤ min(A), or m < min(A), respectively. We denote by (cid:22) the partial order of extension on [N]<ω, i.e., A = {a1, a2, . . . al} (cid:22) B = {b1, b2, . . . bm} if l ≤ m and ai = bi, for i = 1, 2 . . . l, and we write A ≺ B if A (cid:22) B and A 6= B. We say that F ⊂ [N]<ω is closed under taking restrictions or is a subtree of [N]<ω if A ∈ F whenever A ≺ B and B ∈ F, hereditary if A ∈ F whenever A ⊂ B and B ∈ F, and F is called compact if it is compact in the product topology. Note that a family which is closed under restrictions is compact if and only if it is well founded, i.e., if it does not contain strictly ascending chains with respect to extensions. Given n, a1 < . . . < an, b1 < . . . < bn in N we say that {b1, . . . , bn} is a spread of {a1, . . . , an} if ai ≤ bi for i = 1, . . . , n. A family F ⊂ [N]<ω is called spreading if every spread of every element of F is also in F. We sometimes have to pass from a family F ⊂ [N]<ω to the subfamily F ∩ [N ]ω = {A ∈ F : A ⊂ N }, where N ⊂ N is infinite. Of course we might lose the property that F is spreading. Nevertheless F ∩ [N ]ω will be called spreading relative to N , if with A ∈ F ∩ [N ]<ω every spread of A which is a subset of N is in F. Note that if F ⊂ [N ]<ω is spreading relatively to N = {n1, n2, . . .} ∈ [N]ω, then F =(cid:8)A ∈ [N]<ω : {nj : j ∈ A} ∈ F(cid:9), F N =(cid:8){nj : j ∈ A} : A ∈ F(cid:9), is spreading. N = {n1, n2, . . .} ∈ [N]ω, then we call A second way to pass to a sub families is the following. Assume that F ⊂ [N]<ω and the spread of F onto N . A family F ⊂ [N]<ω is called regular if it is hereditary, compact, and spreading. Note that if F ⊂ [N]<ω is compact, spreading and closed under restriction it is also hereditary and thus regular. Indeed if B = {b1, b2, . . . , bl} ∈ F and 1 ≤ i1 < i2 < . . . < ik ≤ l, then A = {bi1, bi2, . . . , bik} is a spread of B′ = {b1, b2, . . . , bk} and since B′ ∈ F, it also follows that A ∈ F. If F ⊂ [N ]<ω, for some infinite set N ⊂ N, is called regular relatively to N , if it is compact, spreading relatively to N and hereditary. 6 P. MOTAKIS AND TH. SCHLUMPRECHT If F ⊂ [N]<ω we denote the maximal elements of F, i.e., the elements A ∈ F for which there is no B ∈ F with A ≺ B, by MAX(F). Note that if F is compact every element in F can be extended to a maximal element in F. For F ⊂ [N]<ω and A ∈ [N]<ω we define F(A) = {B ∈ [N]<ω : A < B, A ∪ B ∈ F}. Note that if F is compact, spreading, closed under restrictions, or hereditary, so is F(A). If F ⊂ [N]<ω is compact, we denote by CB(F) its Cantor Bendixson index, which is defined as follows. We first define the derivative of F by F ′ =nA ∈ F : ∃(An) ⊂ F \ {A} An →n→∞ Ao =nA ∈ F : ∃(Bn) ⊂ F(A) \ {∅} Bn →n→∞ ∅o = F \ {A ∈ F : A is isolated in F}. Note that if F is hereditary then F ′ is hereditary, but also note that if F is only closed under restrictions, this is not necessarily true for F ′. Indeed, consider for example Every maximal element A of F is not in F ′ and if F is spreading, then F =(cid:8)∅, {1}, {1, 2}, {1, 2, n}, n > 2(cid:9). For A ∈ [N]<ω it follows that (3) Indeed F ′ = F \ MAX(F). F ′(A) =(cid:0)F(A)(cid:1)′ . B ∈ F ′(A) ⇐⇒ B > A and A ∪ B ∈ F ′ ⇐⇒ B > A and ∃(Cn) ∈ F \{A ∪ B} Cn →n→∞ A ∪ B ⇐⇒ B > A and ∃(Bn) ∈ F(A)\{B} Bn →n→∞ B ⇐⇒ B ∈(cid:0)F(A)(cid:1)′ [Choose Bn = Cn \ A, for n large enough]. By transfinite induction we define for each ordinal α the α-th derivative of F, by F (0) = F, F (α) =(cid:0)F (γ)(cid:1)′ , if α = γ + 1, and F (α) = \γ<α F (γ), if α is a limit ordinal. It follows that F (β) ⊂ F (α) if α ≤ β. By transfinite induction (3) generalizes to (4) F (α)(A) =(cid:0)F(A)(cid:1)(α), for all A ∈ [N]<ω and ordinal α. Assume that F ⊂ [N]<ω is compact. Since F is countable and since every countable and compact metric space has isolated points, it follows that for some α < ω1 the α-th derivative of F is empty and we define CB(F) = min{α : F (α) = ∅}. CB(F) is always a successor ordinal. Indeed, if α < ω1 is a limit ordinal and F (γ) 6= ∅ for all γ < α, it follows that F (α) =T F (γ) 6= ∅. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 7 Definition 2.1. For F,G ⊂ [N] we define (5) (6) F ⊔< G :=(cid:8)A ∪ B : A ∈ F, B ∈ G and A < B(cid:9) F[G] :=( n[i=1 Bi : n ∈ N, B1 < B2 < . . . < Bn, Bi ∈ G, i = 1, 2, . . . n, and {min(Bi) : i = 1, 2 . . . n} ∈ F ) . It is not hard to see that if F and G are regular families so are F ⊔< G and F[G]. 2.2. The Schreier families. We define the Schreier families Sα ⊂ [N]<ω by transfinite induction for all α < ω1, as follows : (7) if α = γ + 1, we let S0 =(cid:8){n} : n ∈ N(cid:9) ∪ {∅} Sα = S1[Sγ] =n n[j=1 (8) Ej : n ≤ min(E1), E1 < E2 < . . . < En, Ej ∈ Sγ, j = 1, 2, . . . , no, and if α is a limit ordinal we choose a fixed sequence (cid:0)λ(α, n) : n ∈ N(cid:1) ⊂ [1, α) which increases to α and put (9) Sα = {E : ∃ k ≤ min(E), with E ∈ Sλ(α,k)}. An easy induction shows that Sα is a hereditary, compact and spreading family for all α < ω1. It is not very hard to see by transfinite induction that Sα is in the following very limited sense backwards spreading: (10) If A = {a1, a2, . . . , an} ∈ Sα , then {a1, a2, . . . , an−1, an − 1} ∈ Sα. So, in particular, if A ∈ Sα \ {∅} is not maximal, then (A ∪ {k})k>max(A) ⊂ Sα. Secondly, by transfinite induction we can easily prove that Sα is "almost" increasing in α, in the following sense: Proposition 2.2. For all ordinals α < β < ω, there is an n ∈ N so that Sα ∩ [n, ∞)<ω ⊂ Sβ. The following formula for CB(Sα) is well know and can easily be shown by transfinite induction for all α < ω1. Proposition 2.3. For α < ω1 we have CB(Sα) = ωα + 1. We now make further assumptions on the approximating sequence (λ(α, n)) ⊂ [1, α), we had chosen to define the Schreier family Sα, for limit ordinals α < ω1. And we will choose (cid:0)λ(α, n)(cid:1) recursively. Assume that α is a countable limit ordinal and that we have defined (cid:0)λ(γ, n)(cid:1), for all limit ordinals γ < α, and thus, Sγ for all γ < α. Recall that α, can be represented uniquely in its Cantor Normal Form (11) where ξk > ξk−1 > . . . ξ1, mk, mk−1, . . . , m1 ∈ N, and, since α is a limit ordinal, ξ1 ≥ 1. α = ωξk mk + ωξk−1mk−1 + . . . + ωξ1m1, We distinguish between three cases: Case 1. k ≥ 2 or m1 ≥ 2. In that case we put for n ∈ N (12) λ(α, n) = ωξk mk + ωξk−1mk−1 + . . . + ωξ1(m1 − 1) + λ(ωξ1, n) Before considering the next cases we need to make the following observation: 8 P. MOTAKIS AND TH. SCHLUMPRECHT Proposition 2.4. Assume that for all limit ordinals γ ≤ α satisfying Case1 the approximat- ing sequences (cid:0)λ(γ, n) : n ∈ N(cid:1) satisfies the above condition (12). It follows for all γ ≤ α, with γ = ωξlml + ωξl−1ml−1 + . . . + ωξ1m1, being the Cantor Normal Form, that (13) Sγ = Sγ2[Sγ1 ], where for some j = 1, . . . l γ1 = ωξlml + ωξl−1ml−1 + . . . ωξj m(1) j with m(1) j , m(2) j ∈ N ∪ {0}, mj = m(1) and γ2 = ωξj m(2) j + m(2) . j j + ωξj−1mj−1 + . . . ωξ1m1, Proof. We will show (13) by transfinite induction for all γ ≤ α. Assume that (13) holds for all γ < γ. If γ = ωξ, then (13) is trivially satisfied. Indeed, in that case γ = γ + 0 = 0 + γ, are the only two choices to write γ as the sum of two ordinals, and we observe that S0[Sγ] = Sγ[S0] = Sγ. It is left to verify (13) in the case that l ≥ 2 or ml ≥ 2. Let γ = γ1 +γ2 be a decomposition of γ as in the statement of (13). We can without loss of generality assume that γ2 > 0. If γ2 = β + 1 for some β (which implies that γ itself is a successor ordinal) it follows from S1(cid:2)Sβ[Sγ1](cid:3) = Sβ+1[Sγ1]. the induction hypothesis and (8) that Sγ1+β+1 = S1(cid:2)Sβ[Sγ1 ](cid:3), so we need to show that If A ∈ S1(cid:2)Sβ[Sγ1](cid:3), we can write A as A = Sm Ai ∈ Sβ[Sγ1], for i = 1, . . . , n which in turn means that Ai =Smi Sβ+1[Sγ1] ⊂ S1(cid:2)Sβ[Sγ1 ](cid:3). i=1 Ai with m ≤ A1 < A2 < . . . < An and j=1 A(i,j), where A(i,1) < A(i,2) < . . . < A(i,li), A(i,j) ∈ Sγ1, for j = 1, 2, . . . , li, and {min(A(i,j) : j = 1, 2, . . . , li} ∈ Sβ, for i = 1, 2, . . . , m. This means that {min A(i,j) : j = 1, 2, . . . , li, i = 1, 2, . . . , m} is in Sβ+1 and thus we conclude that A ∈ Sβ+1[Sγ1 ]. Conversely, we can show in a similar way that If γ2 is a limit ordinal we first observe that λ(γ, n) = λ(γ1 + γ2, n) = γ1 + λ(γ2, n). If A ∈ Sγ1+γ2 it follows that there is an n ≤ min A so that, using the induction hypothesis, we have A ∈ Sγ1+λ(γ2,n) = Sλ(γ2,n)[Sγ1 ]. This means that A =Sm j=1 Aj with, A1 < A2 < . . . < Am, {min(Aj) : j = 1, 2, . . . , m} ∈ Sλ(γ2,n) and Aj ∈ Sγ1, for j = 1, 2, . . . , m. Since n ≤ min(A) = min(A1), it follows that {min(Aj) : j = 1, 2, . . . , m} ∈ Sγ2, and, thus, that A ∈ Sγ2 [Sγ1]. Conversely, we can similarly show that if A ∈ Sγ2 [Sγ1], then it follows that A ∈ Sγ1+γ2. (cid:3) If Case 1 does not hold α must be of the form α = ωγ. , for some κ < ω1. In that cases we make the following requirement on the Case 2. α = ωωκ sequence(cid:0)λ(α, n) : n ∈ N(cid:1): (14) Sλ(α,n) ⊂ Sλ(α,n+1), for all n ∈ N. We can assure (14) as follows: first choose any sequence λ′(α, n), which increases to α. Then we notice that Proposition 2.2 yields that for a fast enough increasing sequence (ln) ⊂ N, it follows that Sλ′(α,n)+ln ⊂ Sλ′(α,n+1)+ln+1 . Indeed, we first note that the only set A ∈ Sγ, γ < α which contains 1, must be the singleton A = {1}. This follows easily by induction. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 9 Secondly we note that by (8) it follows that [{2, 3, . . . , n}] ⊂ Sγ+n, for each γ < α, and n ∈ N, which yields our claim. The remaining case is the following Case 3. α = ωωκ+ξ, where 1 ≤ ξ ≤ ωκ. We first observe that in this case κ and ξ are uniquely defined. Lemma 2.5. Let α be an ordinal number so that there are ordinal numbers κ, ξ with ξ ≤ ωκ and α = ωωκ+ξ. Then for every κ′, ξ′ with ξ′ ≤ ωκ′ , we have κ = κ′ and ξ = ξ′. Proof. Let α = ωωκ+ξ = ωωκ′ + ξ′. ω = ωκ′+1 ≤ ωκ ≤ ωκ + ξ, which is a contradiction. + ξ′ ≤ ωκ′ If κ′ < κ, then ωκ′ We conclude that κ ≤ κ′, and therefore by interchanging the roles of κ and κ′ we obtain that κ = κ′. In conclusion, ωκ + ξ = ωκ + ξ′ and therefore ξ = ξ′ as well. (cid:3) be as above. By [34, Theorem 41, §7.2] ωκ + ξ = ωκ′ so that α = ωωκ′ +ξ′ 2 < ωκ′ +ξ′ We choose now a sequence (θ(ξ, n))n of ordinal numbers increasing to ωξ, so that (15) and define Sωωκ θ(ξ,n) ⊂ Sωωκ θ(ξ,n+1) λ(α, n) = ωωκ (16) We describe how (15) can be obtained. Start with an arbitrary sequence (θ′(ξ, n))n in- creasing to ωξ. We shall recursively choose natural numbers (kn)n∈N, so that setting θ(ξ, n) = θ′(ξ, n) + kn, (15) is satisfied. Assuming that k1, . . . , kn have been chosen choose kn+1,, as in the argument yielding (14), so that θ(ξ, n). We will show that this kn+1 is the desired natural number, i.e. that Sωωκ θ(ξ,n) ⊂ Sωωκ θ′(ξ,n+1)+kn+1. Sωωκ θ(ξ,n) ⊂ Sωωκ (θ′(ξ,n+1)+kn+1). First note that using finite induction and Proposition (2.4), it is easy to verify that for γ < α, with γ = ωξ, for some ξ < ω1, and n ∈ N (17) Sγ·n = Sγ[Sγ · · · Sγ[Sγ]]] . and thus n-times {z } Sωωκ (θ′(ξ,n+1)+kn+1) == Sωωκ Θ′(ξ,n+1)+ωωκ kn+1 = Sωωκ [· · · [Sωωκ [Sωωκ θ′(ξ,n+1)]]] ⊃ S1[· · · [S1 [Sωωκ θ′(ξ,n+1)]]] = Sωωκ θ′(ξ,n+1)+kn+1 ⊃ Sωωκ θ(ξ,n). kn+1-times {z } kn+1-times {z } We point out that the sequence (θ(ξ, n))n also depends on α. Proposition 2.6. Assuming the approximating sequences(cid:0)λ(α, n) : n ∈ N(cid:1) satisfy for all limit ordinals α the above conditions. It follows for all γ < ω1, with γ = ωξlml + ωξl−1ml−1 + . . . + ωξ1m1, being the Cantor Normal Form, that (18) Sλ(γ,n) ⊂ Sλ(γ,n+1) for all n ∈ N, if γ is a limit ordinal. 10 (19) P. MOTAKIS AND TH. SCHLUMPRECHT Sγ = Sγ2[Sγ1 ], where for some j = 1, 2, . . . , l γ1 = ωξlml +ωξl−1ml−1+. . .+ωξj m(1) j with m(1) j , m(2) j ∈ N ∪ {0}, mj = m(1) and γ2 = ωξj m(2) j + m(2) . j j +ωξj−1mj−1+. . .+ωξ1m1, and if β = ωωκ and γ is a limit ordinal with γ ≤ β, then (20) there is a sequence (η(γ, n))n increasing to γ so that λ(βγ, n) = βη(γ, n) (this sequence (η(γ, n))n can depend on β). Proof. We first will prove (18), (19) simultaneously for all γ < ω1. Assume that our claim is true all γ < γ. (19) follows from Proposition 2.4. If l = m1 = 1 we deduce (18) from the choice of λ(γ, n), n ∈ N, in that case (see (14), (15) and (16)). If l ≥ 2 or m2 ≥ 2 we deduce from (13) and the induction hypothesis that Sλ(γ,n) = Sωξk mk+...+ωξ2 m2+ωξ1 m1+λ(ωξ1 ,n) = Sλ(ωξ1 ,n)[Sωξk mk+...+ωξ2 m2+ωξ1 m1 ] ⊂ Sλ(ωξ1 ,n+1)[Sωξk mk+...+ωξ2 m2+ωξ1 m1 ] = Sλ(γ,n+1), which verifies (18) also in that case To verify (20) let κ < ω1 with β = ωωκ ≥ γ. Recall that by (16) λ(ωωκ+ξ1, n) = ωωκ θ(ξ1, n). For each n, define η(γ, n) = ωξlml + ωξl−1ml−1 + · · · + ωξ1(m1 − 1) + θ(ξ1, n). We will show that (η(γ, n))n∈N has the desired property. Note that the Cantor Normal Form of βγ is βγ = ωωκ+ξlml + ωωκ+ξl−1ml−1 + · · · + ωωκ+ξ1m1. Hence, by (12): λ(βγ, n) = ωωκ+ξlml + ωωκ+ξl−1ml−1 + . . . + ωωκ+ξ1(m1 − 1) + λ(ωωκ+ξ1, n) θ(ξ1, n) = ωωκ+ξlml + ωωκ+ξl−1ml−1 + . . . + ωωκ+ξ1(m1 − 1) + ωωκ = ωωκ (ωξlml + ωξl−1ml−1 + . . . + ωξ1(m1 − 1) + θ(ξ1, n)) = βη(γ, n). Remark. The proof of Proposition 2.6, in particular the definition of (η(γ, n))n, implies the following. Let ξ be a countable ordinal number and γ ≤ β = ωωξ be a limit ordinal number. If γ = ωξ1, then (cid:3) (21) η(γ, n) = θ(ξ1, n), for all n ∈ N. Otherwise, if the Cantor normal form of γ is γ = ωξlml + ωξl−1ml−1 + . . . + ωξ1m1 and γ1 = ωξlml + ωξl−1ml−1 + . . . ωξj m(1) j m(1) j ∈ N ∪ {0}, mj = m(1) j + m(2) , m(2) j j , then we have , γ2 = ωξj m(2) j + ωξj−1mj−1 + . . . ωξ1m1, with (22) η(γ, n) = γ1 + η(γ2, n), for all n ∈ N. Corollary 2.7. If α < ω1 is a limit ordinal it follows that (23) Sα =(cid:8)A ∈ [N]<ω \ {∅} : A ∈ Sλ(α,min(A))(cid:9) ∪ {∅}. Remark 2.8. If we had defined Sα by (23), for limit ordinals α < ω1 where(cid:0)λ(α, n) : n ∈ N) is any sequence increasing to α, then we would not have ensured that the family Sα is a regular family. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 11 2.3. The fine Schreier families. We will now define the Fine Schreier Sets. For that we will also need to choose appropriate approximating sequences for limit ordinals. We will define them as a doubly indexed family Fβ,α ⊂ [N]<ω, α ≤ β < ω1. Later in the proof of Theorems A and C, we will fix β, depending on the Banach space X we are considering. Definition 2.9. For a countable ordinal number ξ and β = ωωξ hierarchy of families of finite subsets of the natural numbers (Fβ,γ)γ≤β as follows: , we recursively define an (i) Fβ,0 = {∅}, (ii) if γ < β then Fβ,γ+1 = {{n} ∪ F : F ∈ Fβ,γ, n ∈ N} (i.e., Fβ,γ+1 = Fβ,1 ⊔< Fβ,γ), and (iii) if γ ≤ β is a limit ordinal number, then Fβ,γ = ∪n∈N(Fβ,η(γ,n) ∩ [n, ∞)<ω), where (η(γ, n))n is the sequence provided by Proposition 2.6 (and depends on β). Remark. It can be easily shown by transfinite induction that each family Fβ,γ is regular. In the literature fine Schreier families are usually defined recursively as a singly indexed family (Fα)α<ω1 of subsets of [N]<ω. In that case F0 and Fα are defined for successor ordinals as in (i) and (ii). And if α is a limit ordinal is defined as in (iii) where we do not assume that the approximating sequence (η(α, n))n∈N does not depend of any β ≥ α. Let ξ be a countable ordinal number and ξ1 ≤ ωξ. If and β = ωωξ and γ = ωξ1, it follows by (21) that η(γ, n) = θ(γ, n) for n ∈ N. The choice of (θ(ξ1, n))n∈N may be done so that alongside (15), we also have (24) Fβ,η(γ,n) = Fβ,θ(ξ1,n) ⊂ Fβ,θ(ξ1,n+1) = Fβ,η(γ,n+1). This can be achieved by possibly adding to θ′(ξ1, n) a large enough natural number. Proposition 2.10. Let ξ be a countable ordinal number and β = ωωξ . Assume that for all limit ordinals limit ordinals γ ≤ β the approximating sequence (η(γ, n))n satisfies conditions (21) and (22), and for the case γ = ωξ1 the approximating sequence (θ(ξ1, n))n satisfies condition (24). Then, for all γ ≤ β, whose Cantor normal form is γ = ωξlml + ωξl−1ml−1 + · · · + ωξ1m1 we have that that (25) Fβ,η(γ,n) ⊂ Fβ,η(γ,n+1) for all n ∈ N, if γ is a limit ordinal, and if for some for some 1 ≤ j ≤ l, γ1 = ωξlml + ωξl−1ml−1 + · · · + ωξj m(1) j ωξj m(2) j + m(2) j + ωξj−1mj−1 + · · · + ωξ1m1 with m(1) j , m(2) j ∈ N ∪ {0}, mj = m(1) j and γ2 = , then (26) Fβ,γ = Fβ,γ2 ⊔< Fβ,γ1. Proof. We will show (25) and (26) simultaneously by transfinite induction for all γ ≤ β. Assume that (25) and (26) hold for all γ < γ. If γ = ωκ, then (25) follows from (21) and (24), while (26) is trivially satisfied. Indeed, in that case γ = γ+0 = 0+γ, are the only two choices to write γ as the sum of two ordinals, and we observe that Fβ,0⊔<Fβ,γ = Fβ,γ⊔<Fβ,0 = Fβ,γ. If γ is a limit ordinal (thus ξ1 > 0) and either l ≥ 2 or ml ≥ 2, then (25) follows from the inductive assumption. Indeed, for n ∈ N it follows that η(γ, n) = γ′ + η(ωξ1, n) with γ′ = ωξlml + ωξl−1ml−1 + . . . + ωξ1(m1 − 1) and thus (note that γ′ + η(ωξ1 , n) < γ) Fβ,η(γ,n) = Fβ,η(ωξ1 ,n) ⊔< Fβ,γ′ ⊂ Fβ,η(ωξ1 ,n+1) ⊔< Fβ,γ′ = Fβ,η(γ,n+1). 12 P. MOTAKIS AND TH. SCHLUMPRECHT It is left to verify (26) in the case that l ≥ 2 or ml ≥ 2. Let γ = γ1 +γ2 be a decomposition of γ as in the statement of (26). We can without loss of generality assume that γ2 > 0. If γ2 = γ′ 2 + 1 for some γ′ 2 (which implies that γ itself is a successor ordinal) it follows from the inductive assumption and Definition 2.9 (ii) that Fβ,γ = Fβ,γ1+γ′ 2+1 = Fβ,1 ⊔< Fβ,γ1+γ′ 2 =(cid:16)Fβ,1 ⊔< Fβ,γ′ 2(cid:17) ⊔< Fβ,γ1 = Fβγ′ = Fβ,1 ⊔<(cid:16)Fβ,γ′ 2 ⊔< Fβ,γ1(cid:17) 2+1 ⊔< Fβ,γ1 = Fβ,γ2 ⊔< Fβ,γ1. If γ2 is a limit ordinal then recall that by (22) we have η(γ, n) = γ1 + η(γ2, n). If A ∈ Fβ,γ it follows that there is an n ≤ min A so that, using the inductive assumption, we have A ∈ Fβ,η(γ,n) = Fβ,γ1+η(γ2,n) = Fβ,η(γ2,n) ⊔< Fβ,γ1. This means that A = A1 ∪ A2 with, A1 < A2, A1 ∈ Fβ,η(γ2,n), and A2 ∈ Fβ,γ1. If A1 6= ∅, then min(A1) = min(A) ≥ n, i.e. A1 ∈ Fβ,γ2 and hence A ∈ Fβ,γ2 ⊔< Fβ,γ1. If on the other hand A1 = ∅, then A ∈ Fβ,γ1 ⊂ Fβ,γ2 ⊔< Fβ,γ1. Conversely, we can similarly show that if A ∈ Fβγ1 ⊔< Fβ,γ1, then A ∈ Fβ,γ. (cid:3) Corollary 2.11. Let ξ be a countable ordinal number and γ ≤ β = ωωξ number. Then be a limit ordinal (27) Fβ,γ = {F ∈ [N]<ω : F ∈ Fβ,η(γ,min(F ))} ∪ {∅}. The following formula of the Cantor Bendixson index of Sα and Fβ,α can be easily shown by transfinite induction. Proposition 2.12. For any α, κ < ω1, with α ≤ β = ωωκ , CB(Sα) = ωα + 1 and CB(Fβ,α) = α + 1. Moreover, assuming ωα ≤ β, for every M ∈ [N]ω, there is an M ∈ [N ]ω so that α ⊂ Fβ,ωα and F N S N β,ωα ⊂ Sα. The main result in [15] states that if F and G are two hereditary subsets of [N], then for any M ∈ [N]ω there is an N ∈ [M ]ω so that F ∩ [N ]<ω ⊂ G or G ∩ [N ]<ω ⊂ F. Together with Proposition 2.12 this yields Proposition 2.13. For any α, γ, κ < ω1, with ωα ≤ β = ωωκ is an N ∈ [M ]<ω so that , and any M ∈ [N ]<ω, there S N γ ⊂ Sα ∩ [N ]<ω ⊂ Fβ,ωα, if γ < ωα, and F N β,ωα ⊂ Fβ,ωα ∩ [N ]<ω ⊂ Sγ, if γ > ωα. 2.4. Families indexed by subsets of [N]<ω. We consider families of the form (xA : A ∈ F) in some set X indexed over F ⊂ [N]<ω. If F is a tree, i.e., closed under restrictions, such a family is called an indexed tree. Let us also assume that F is spreading. The passing to a pruning of such an indexed tree is what corresponds to passing to subsequences for sequences. Formally speaking we define a pruning of (xA : A ∈ F) as follows. Let π : F → F be an order isomorphism with the property that if F ∈ F is not maximal, then for any n ∈ N, so that n > max(A) and A ∪ {n} ∈ F, π(A ∪ {n}) is of the form π(A) ∪ {sn}, where (sn) is a sequence which increases with n. We call then the family (xA : A ∈ π(F)) a pruning of (xA : A ∈ F). Let xA = xπ(A) for A ∈ F. (xA : A ∈ F) is then simply a relabeling of the family (xA : A ∈ π(F)), and we call it also a pruning of (xA : A ∈ F). It is important to note that the branches of a pruning of an indexed tree (xA : A ∈ F), are a A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 13 subset of the branches of the original tree (xA : A ∈ F). Here a branch of (xA : A ∈ F), is a set of the form xF = (x{a1}, x{a1,a2}, . . . x{a1,a2,...,al}) for F = {a1, a2, . . . , al} ∈ F . Also the nodes of a pruned tree, namely the sequences of the form (xA∪{n} : A ∪ {n} ∈ F), with A ∈ F not maximal, are subsequences of the nodes of the original tree. Let us finally mention, how we usually choose prunings inductively. Let {An : n ∈ N} be a consistent enumeration of F. By this we mean that if max(Am) < max(An) then m < n. Thus, we also have if Am ≺ An, then m < n, and if Am = A ∪ {s} ∈ F and An = A ∪ {t} ∈ F, for some (non maximal) A ∈ F and s < t in N, then m < n. Of course, then A1 = ∅ and π(∅) = ∅ assuming now that π(A1), π(A2), . . . , π(Am) have been chosen, then Am+1 must be of the form Am = Al ⊂ {k}, with l < m = 1. Moreover if, k > max(Al) + 1 and if Al ∪ {k − 1} ∈ F then Al ∪ {k − 1} = Aj with l < j < m + 1, and π(Aj) = π(Al) ∪ {s} for some s has already been chosen. Thus, we need to choose π(Am+1) to be of the form π(Al) ∪ {t}, where, in case that Al ∪ {k − 1} ∈ F, we need to choose t > s. Proposition 2.14. Assume that F ⊂ [N]<ω is compact. Let r ∈ N and f : MAX(F) → {1, 2, . . . , r}. Then for every M ∈ [N]ω there exists an N ∈ [M ]ω and an i ∈ {1, 2, . . . , r} so that MAX(F) ∩ [N ]ω ⊂ {A ∈ MAX(F) : f (A) = i}. Proposition 2.14 could be deduced from Corollary 2.5 and Proposition 2.6 in [3]. To make the paper as self-contained as possible we want to give a short proof. Proof of 2.14. Without loss of generality we can assume that r = 2. We prove our assump- tion by induction on the Cantor Bendixson index of F. If CB(F) = 1, then F can only consist of finitely many sets. Since F is spreading it follows that F = {∅} and our claim is trivial. Assume that our claim is true for all regular families E, with CB(E) < β. Let F = F1 ∪ F2, where F is a regular family with CB(F) = α + 1, where α is the predecessor of β, if β is a successor ordinal (and thus β = α + 1), and α = β, if β is a limit ordinal, and f : F → {1, 2} with F1 = {A ∈ F : f (A) = 1} and F2 = {A ∈ F : f (A) = 2}, and let M ∈ [N]ω. First we observe that there is a cofinite subset M ′ of M , so that CB(F({m})) ≤ α, for all m ∈ M ′. Indeed, if that were not true, and thus CB(F({m})) = α + 1 for all m in an infinite subsets M ′ of M , we could choose for every m ∈ M ′ an element Am ∈ (cid:0)F({m})(cid:1)(α) = (cid:0)F (α)(cid:1)({m}). Thus {m} ∪ Am ∈ F (α). But this would imply that ∅ = limm∈M ′,m→∞{m} ∪ Am ∈ F (α+1), which is a contradiction. Since CB(F({m})) has to be a a successor ordinal we deduce that CB(F({m})) < β, for m ∈ M , and can therefore apply our induction hypothesis, and choose inductively natural numbers m1 < m2 < m3 < . . . and infinite sets N0 = M ′ ⊃ N1 ⊃ N2 . . ., and elements c1, c2, . . . ∈ {1, 2}, so that mj = min Nj−1 < min(Nj) and F({mj}) ∩ [Nj]<ω ⊂ {A ∈ F : f (A) = cj}, for all j ∈ N. Indeed, if Nj−1 has been defined we let mj = min(Nj−1), and apply our induction hypothesis to the family F({mj}), and the coloring fj : MAX(cid:0)F({mj})(cid:1) → {1, 2}, B 7→ f ({mj} ∪ B), and the set Nj−1 \ {mj}, to obtain an Then take a c for which {j : cj = c} is infinite and N = {mj : cj = c}. If A = {a1, a2, . . . , al} ∈ MAX(F) ∩ [N ]<ω, Then a1 = mj for some j ∈ N, with cj = c, and {a2, a3, . . . , an} ∈ MAX(F({a1}) ∩ [Nj]<ω, and thus f (A) = c, which verifies our claim. (cid:3) infinite set Nj ⊂ Nj−1 \ {mj}. 14 P. MOTAKIS AND TH. SCHLUMPRECHT 3. Repeated averages on Schreier sets We recall repeated averages defined on maximal sets of Sα, α < ω1 (c.f. [4]). As in our previous sections we will assume that Sα is recursively defined using the conditions made in Subsection 2.2. We first need the following characterization of maximal elements of Sα, α < ω1, which can be easily proven by transfinite induction using for the limit ordinal case Corollary 2.7. Proposition 3.1. Let α < ω1 then (i) A ∈ MAX(Sα+1) if and only if A =Sn . . . < An are in MAX(Sα). In this case the Aj are unique. (ii) If α is a limit ordinal then A ∈ MAX(Sα) if and only if A ∈ MAX(Sλ(α,min(A))). j=1 Aj, with n = min(A1) and A1 < A2 < For each α < ω1 and each A ∈ MAX(Sα) we will now introduce an element z(α,A) ∈ Sℓ+ with supp(z(α,A)) = A, which we will call repeated average of complexity α on A ∈ MAX(Sα). If α = 0 then MAX(S0) consists of singletons and for A = {n} ∈ MAX(Sα) we put z(0,{n}) = en, the n-th element of the unit basis of ℓ1. Assume for all γ < α and all 1 with z(γ,A) > 0 for all a ∈ A. A ∈ MAX(Sγ) we already introduced z(γ,A) which we write as z(γ,A) = Pa∈A z(γ,A)(a)ea, write by Proposition 3.1 (i) A in a unique way as A = Sn If α = γ + 1 for some γ < ω1 and if A ∈ MAX(Sα) we j=1 Aj, with n = min A and A1 < A2 < . . . < An are maximal in Sγ. We then define z(α,A) = 1 n z(γ,Aj ) = 1 n nXj=1 nXj=1 Xa∈Aj z(γ,Aj )(a)ea, z(α,A)(a) = 1 n z(γ,Aj )(a) for j = 1, 2, . . . , n and a ∈ Aj. (28) and thus (29) (30) If α is a limit ordinal and A ∈ MAX(Sα) then, by Corollary 2.7, A ∈ Sλ(α,min(A)), and we put z(α,A) = z(λ(α,min(A)),A) =Xa∈A z(λ(α,min(A)),A)(a)ea. The following result was, with slightly different notation, proved in [4]. Lemma 3.2. [4, Proposition 2.15] For all ε > 0, all γ < α, and all M ∈ [N]ω, there is an N = N (γ, α, M, ε) ∈ [M ]ω, so that Pa∈A z(α,B)(a) < ε for all B ∈ MAX(Sα ∩ [N ]ω) and A ∈ Sγ. The following Proposition will be proved by transfinite induction. Proposition 3.3. Assume α < ω1 and A ∈ Sα (not necessarily maximal). If B1, B2 are two extensions of A which both are maximal in Sα then it follows z(α,B1)(a) = z(α,B2)(a) for all a ∈ A. Remark. Proposition 3.3 says the following: If α < ω1 and A = {a1, a2, . . . , an} is in MAX(Sα), then z(α,A)(a1) only depends on a1, z(α,A)(a2), only depends on a1 and a2 etc. Proof of Proposition 3.3. Our claim is trivial for α = 0, assume that α = γ +1 and our claim is true for γ, and let A ∈ Sγ+1. Without loss of generality A 6= ∅, otherwise we would be done. Using Proposition 3.1, we can find an integer 1 ≤ l ≤ min A, sets A1, A2, . . . , Al−1 ∈ A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 15 A =Sl be of the form B =Sl MAX(Sγ), and Al ∈ Sγ (not necessarily maximal in Sγ) so that A1 < A2 < . . . < Al and j=1 Aj. By Proposition 3.1, any extension of A to a maximal element in Sγ will then j=l Bj, where Al < Bl < Bl+1 < . . . < Bmin(A) (Bl may be empty, in which case Al < Bl+1 < . . . < Bmin(A)), so that Al ∪ Bl, Bl+1, . . . , Bmin(A) are in MAX(Sγ). No matter how we extend A to a maximal element B in Sγ+1, the inductive formula (28) yields j=1 Aj ∪Smin(A) z(γ+1,B)(a) = 1 min(A) z(γ,Aj )(a) whenever for some j = 1, 2, . . . , l − 1 we have a ∈ Aj. In the case that a ∈ Al, then, by our induction hypothesis, zγ,Al∪Bl(a) does not depend on the choice of Bl, and z(γ+1,B)(a) = 1 n z(γ,Al∪Bl)(a) whenever a ∈ Al Thus, in both cases, the value of z(γ+1,B)(a) does not depend on how we extend A to a maximal element B in Sγ+1. If α is a limit ordinal and A ∈ Sα is not maximal, we also can assume that A 6= ∅, and thus it follows from the formula (23) in Corollary (2.7) that A ∈ Sλ(α,min(A)). For any two extension B of A into a maximal set of MAX(Sα), it follows from Proposition 3.1 that B is maximal in Sλ(α,min(A)), and that z(α,B) = z(λ(α,min(A),B). Thus, also in this case our claim follows from the induction hypothesis. (cid:3) Using Proposition 3.3 we can define consistently z(α,A) ∈ Bℓ+ for any α < ω1 and any A ∈ Sα by z(α,A) =Xa∈A z(α,B)(a)ea, where B is any maximal extension of A in MAX(Sα). 1 In particular this implies the following recursive definition of z(α,A): If A ∈ Sα+1 \ {∅} we j=1 An, where A1 < A2 < . . . < An, Aj ∈ MAX(Sα), can write in a unique way A as A =Sn for j = 1, 2, . . . , n − 1 and An ∈ Sα \ {∅}, and note that (31) and if α is a limit ordinal z(α+1,A) = 1 min(A) dXj=1 z(α,Aj ) (32) z(α,A) = z(α,λ(α,min(A)). For D ∈ Sα define ζ(α, A) = z(α,D)(max(D). For A ∈ Sα it follows therefore z(α,D) = XD(cid:22)A ζ(α, D)emax(D). We also put ζ(α, ∅) = 0 and emax(∅) = 0. By transfinite induction we can easily show the following estimate fro 1 ≤ α < ω1 (33) ζ(α, A) ≤ 1 min A . From Proposition 2.6 we deduce the following formula for z(α,A) 16 P. MOTAKIS AND TH. SCHLUMPRECHT Proposition 3.4. Assume α < ω1 and that its Cantor Normal Form is Let j = 1, 2 . . . l and m(1) j α = ωξlml + ωξl−1ml−1 + . . . + ωξ1m1. , m(2) j ∈ N ∪ {0}, with m(1) j = mj. Put γ1 = ωξlml + ωξl−1ml−1 + . . . + ωξj−1mj−1 + ωξj m(1) j γ2 = ωξj m(2) j + ωξj−1mj−1 + . . . + ωξ1m1. j + m(2) For A ∈ MAX(Sα) we use Proposition 2.6 and write A = Sn j = 1, 2, . . . , n, A1 < A2 < . . . < An, and B = {min(Aj) : j = 1, 2, . . . , n} ∈ Sγ2. Then it follows that Aj ∈ MAX(Sγ1 ), for j = 1, 2, . . . , n, B ∈ MAX(Sγ2 ) and j=1 Aj, where Aj ∈ Sγ1 for (34) z(α,A) = z(γ2,B)(min(Aj))z(γ1,Aj). In other words, if ∅ ≺ D (cid:22) A, and thus D = Si−1 ∅ ≺ Ai (cid:22) Ai, then j=1 Aj ∪ Ai, for some 0 ≤ i < n, and (35) ζ(α, D) = ζ(γ2, {min(Aj) : j = 1, 2, . . . i}) · ζ(γ1, Ai). Proof. We prove by transfinite induction for all β < ω1, with Cantor Normal Form β = ωξj mj + ωξj−1mj−1 + . . . + ωξ1m1, the following Claim: If γ < ω1 has Cantor Normal Form γ = ωξl ml + ωξl−1 ml−1 + . . . + ωξj mj, where mj could possibly be vanishing, and if A =Sn where Ai ∈ Sγ for i = 1, 2, . . . , n, A1 < A2 < . . . < An, and B = {min(Ai) : i = 1, 2, . . . , n} ∈ Sβ, then it follows that Ai ∈ MAX(Sγ) for i = 1, 2, . . . , n, B ∈ MAX(Sγ1) and i=1 Ai ∈ MAX(Sγ+β) = MAX(cid:0)Sβ[Sγ](cid:1), nXj=1 nXi=1 (36) z(α,A) = z(β,B)(min(Ai))z(γ,Ai) For β = 0 the claim is trivial and for β = 1, our claim follows from Proposition 3.1 and the definition of z(γ+1,A) for A ∈ MAX(Sγ+1). Assume now that the claim is true for all β < β, and that γ < ω1 has above form and let i=1 Ai ∈ MAX(Sγ+β), where Ai ∈ Sγ for i = 1, 2, . . . , n, A1 < A2 < . . . < An, and A =Sn A =Sn i=1 B = {min(Ai) : i = 1, 2, . . . , n} ∈ Sβ. First we note that (10) implies that the Ai are maximal in Sγ. Indeed, if for some i0 = 1, 2, . . . , n, Ai0 is not maximal in Sγ, then if i0 = n, it would directly follow that A cannot be maximal in Sγ+β and if i0 < n we could define Ai = Ai, for i = 1, 2, . . . , , l − 1, Ai0 = Ai0 ∪ {min(Ai0+1)}, Ai = (Ai ∪ {min(Ai+1)}) \ {min(Ai)}, for i = i0, i0 + 1, . . . , l − 1, and Al = Al \ {min Al}. Then, by (10) and the fact that the Schreier families are spreading, Ai is also a decomposition of elements of Sγ with B = {min( Ai) : i = 1, 2, . . . , n} ∈ Sβ. But now An is not maximal in Sγ and we get again a contradiction. It is also easy to see that B is maximal in Sβ. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 17 To verify (36) we first assume that β is a successor ordinal, say β = α + 1. Then we i=1 Bi, where m = min(B) = min(A), B1 < B2 < . . . < Bm, and Bi ∈ MAX(Sα), for i = 1, 2, . . . , m. We can write Bi as can write B as B = Sm with k0 = 0 < k1 < . . . < km = n. We put Ai = Ski s=ki−1+1 As ∈ Sγ+α = Sα[Sγ], for i = 1, 2, , . . . , m. From the definition of z(β+1,B) and from the induction hypothesis we deduce now that Bi = {min(As) : s = ki−1 + 1, ki−1 + 2, . . . , ki}, z(γ+α+1,A) = = 1 m 1 m mXi=1 mXi=1 z(γ+α,Ai) kiXs=ki−1+1 z(α,Bi)(min(As))z(γ,As) = z(β,B)(min(As))z(γ,As), nXs=1 which proves the claim if β is a successor ordinal. If β is a limit ordinal it follows from Corollary 2.7, our definition of z(β,B) and z(γ+β,A), and our choice of the approximating sequence (λ(γ + β), n) that z(γ+β,A) = z(λ(γ+β,min(A)),A) = z(γ+λ(β,min(B)),A) = z(λ(β,min(B)),B)(min(Aj))z(γ,Aj ) = nXj=1 z(β,B)(min(Aj))z(γ,Aj ) nXj=1 (cid:3) which proves our claim also in the limit ordinal case. If α < ω1 and A ∈ MAX(Sα) then z(α,A) is an element of Sℓ1 ∩ ℓ+ 1 and can therefore be seen as probability on N, as well as A. We denote the expected value of a function f defined on A or on all of N as E(α,A)(f ). As done in [33], we deduce the following statement from Lemma 3.2. Corollary 3.5. [33, Corollary 4.10] For each α < ω1 and A ∈ MAX(Sα) let fA : A → [−1, 1] have the property that Eα,A(fA) ≥ ρ, for some fixed number ρ ∈ [−1, 1]. For δ > 0 and M ∈ [N]ω put Aδ,M =(cid:26)A ∈ Sα ∩ [M ]<∞ : ∃B ∈ MAX(Sα ∩ [M ]<∞), A ⊂ B, and fB(a) ≥ ρ − δ for all a ∈ A(cid:27) . Then CB(Aδ,M ) = ωα + 1. We finish this section with an observation, which will be needed later. Definition 3.6. If A ⊂ N \ {∅} is finite, we can write it in a unique way as a union j=1 Aj, where A1 < A2 < . . . < Ad, and Aj ∈ MAX(S1) if j = 1, 2, . . . , d − 1, and Ad ∈ S1 \ {∅}. We call (Aj)d j=1 the optimal S1-decomposition of A, and we define A =Sd For A = ∅ we put l1(A) = 0. l1(A) = min(Ad) − #Ad. The significance of this number and its connection to the repeated averages is explained in the following Lemma. Lemma 3.7. Let α ∈ [1, ω1), A ∈ Sα and let (Aj)d j=1 be its optimal S1-decomposition. (i) l1(A) = 0 if and only if A = ∅ or Ad ∈ MAX(S1). 18 P. MOTAKIS AND TH. SCHLUMPRECHT (ii) If A ∈ MAX(Sα) then Ad ∈ MAX(S1) and, thus, l1(A) = 0. (iii) If l1(A) > 0, then for all max(A) < k1 < k2 < . . . < kl1(A) it follows that A ∪ {k1, k2, . . . , kl1(A)} ∈ Sα and ζ(α, A ∪ {k1, k2, . . . , ki}) = ζ(α, A) for all i = 1, 2 . . . l1(A). (iv) If m > l1(A) and max(A) < k1 < k2 < . . . < km, have the property that A ∪ {k1, k2, . . . km} ∈ Sα then ζ(α, A ∪ {k1, k2, . . . ki}) ≤ 1 kl1(A)+1 . (v) If A 6= ∅, then XD(cid:22)A,l1(D′)=0 ζ(α, D) ≤ 1 min(A) and XD(cid:22)A,l1(D)=0 ζ(α, D) ≤ 1 min(A) (recall that D′ = D \ {max D} for D ∈ [N]<ω \ {∅} and ∅′ = ∅). Proof. We proof (i) through (v) by transfinite induction for all α ∈ [1, ω1). For α = 1, (i), (ii), (iii) and (v) follow from the definition of S1 and the definition of ζ(α, A), for A ∈ S1, while (iv) is vacuous in that case. Assume our claim is true for some α < ω1, and let A ∈ Sα+1. Without loss of generality we can assume that A 6= ∅. Indeed, if A = ∅, then (i) is clear (ii), (iii), and (v) are vacuous, while (iv) follows easily by induction from the fact that always ζ(α, A) ≤ 1 min(A) if A ∈ Sα \ {∅}. By the definition of Sα+1, A can be written in j=1 Bj where Bj ∈ MAX(Sα), for j = 1, 2, . . . , n − 1, and Bn ∈ Sα. For j = 1, 2, . . . , n let (Aj,i)cj i=1 be the optimal S1-decomposition of Bj. From the induction hypothesis (ii) it follows that (Aj,i) are maximal in S1, for j < n or for j = n and i < cn. Therefore it follows that (Aj,i : j = 1, 2, . . . , n, i = 1, 2 . . . cj) (appropriately ordered) is the optimal S1-decomposition of A and it follows l1(A) = l1(Bn), and Ad = An,cn. a unique way as A =Sn We can deduce (i) from the induction hypothesis. If A ∈ MAX(Sα+1), then, in particular, Bn ∈ MAX(Sα), and thus l1(A) = l1(Bn) = 0. Conversely, if l1(A) = l1(Bn) = 0, then Ad = An,cn ∈ MAX(S1). This proves (ii) for α + 1. If l1(A) > 0 and max(A) = max(Bn) < k1 < k2 < . . . < kl1(A), then it follows from the fact that l1(A) = l1(Bn) and our induction hypothesis that Bn ∪ {k1, k2, . . . , kl1(A)} ∈ Sα and ζ(α, Bn ∪ {k1, k2, . . . , ki}) = ζ(α, Bn). Therefore A ∪ {k1, k2, . . . , kl1(A)} ∈ Sα+1 and, using our recursive formula, we obtain ζ(α, A ∪ {k1, k2, . . . , ki}) = 1 min(A) ζ(α, Bn ∪ {k1, k2, . . . , ki}) = 1 min(A) ζ(α, Bn) = ζ(α, A), which verifies (iii). In order to show (iv) let m > l1(A) and max(A) < k1 < k2 < . . . < km, is such that A ∪ {k1, k2, . . . , km} ∈ Sα+1 we distinguish between two cases. Either Bn ∪ {k1, k2, . . . , km} ∈ Sα. In that case we deduce from the induction hypothesis ζ(cid:0)α, A ∪ {k1, k2, . . . , km}(cid:1) = 1 min(A) ζ(α, Bn ∪ {k1, k2, . . . , km}) ≤ Or we can write A ∪ {k1, k2, . . . , km} as A ∪ {k1, k2, . . . , km} =Sn n ∈ MAX(Sα), B′ n+1 < . . . < B′ p, Bn ∪ B′ p > n, Bn < B′ n < B′ 1 . kl1(A)+1 j=n B′ j, where p−1 ∈ MAX(Sα), j=1 Bj ∪Sp n+1, . . . , B′ A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 19 p ∈ Sα \ {∅}. Let s ≤ m such that ks = min(Bp). Then s > l1(Bn), and l1(Bn ∪ B′ and B′ 0. It follows therefore from (31) and the induction hypothesis that n) = ζ(α + 1, A ∪ {k1, k2, . . . , km}) = 1 min(A) ζ(α, {ks, ks+1, . . . , km}) ≤ 1 ks ≤ 1 kl1(A)+1 , This proves (iv) in both cases. Finally, to verify (v) we observe, that by the induction hypothesis and (31) XD(cid:22)A,l1(D′)=0 ζ(α + 1, D) = 1 min(A) nXj=1 XD(cid:22)Bj,l1(D′)=0 nXj=1 min(Bj) ≤ 1 ζ(α, D) n 1 ≤ 1 min A min(A) min(A) ≤ 1 min(A) , which proves the first part of (v), while the second follows in the same way. If α < ω1 is a limit ordinal and assuming that our claim is true for all γ < α we proceed as follows. For A ∈ Sα, we a can assume again that A 6= ∅ and it follows from Corollary (2.7) that A ∈ Sλ(α,min(A)) and, by Proposition 3.1 A is maximal in Sα if and only if it is maximal in Sλ(α,min(A)). Therefore (i) through (v) follow from our claim being true for λ(α, min(A)). (cid:3) Remark. Recall, that if β = ωωξ is a countable ordinal number and γ < β, then by (17) we have Sβ(γ+1) = Sβ[Sβγ]. An argument very similar to what was used in the proof of Lemma 3.7, implies the following: if B1 < · · · < Bd are in MAX(Sβγ) so that ¯B = {min(Bj) : 1 ≤ j ≤ d} is a non-maximal Sβ set, D = ∪d (37) Corollary 3.8. Let A = {a1, a2, . . . , al} and A = {a1, . . . , al} be two sets in [N]<ω whose j=1 and ( Aj)d optimal S1-decompositions (Aj)d j=1, respectively, have the same length and satisfy min(Aj) = min( Aj), for j = 1, 2, . . . , d. j=1Bj and C ∈ Sβγ with D < C, then l1(C) = l1(D ∪ C). Then it follows for α < ω1 that A ∈ Sα if and only if A ∈ Sα and in that case for D (cid:22) A and D (cid:22) A, with #D = # D, it follows that ζ(α, D) = ζ(α, D). Proof. We prove this lemma by transfinite induction on α. A1 = A, and a1 = a1, and thus ζ(1, D) = ζ(1, D) for all D (cid:22) A and D ≺ A. If α = 1 then, A1 = A and Assume that the conclusion holds for some α and let A ∈ Sα+1, A ∈ [N]<ω satisfy the i=1Ci, where C1 < · · · < Cp−1 are in MAX(Sα), whereas Cp ∈ Sα Ci, where C1 < · · · < Cp, and the Cj are chosen assumption. Let A = ∪p and p ≤ min(A). Write also A as A = ∪p such that #Cj = # Cj for j = 1, 2, . . . , p − 1. i=1 From Lemma 3.7 (ii) It follows that for some sequence 0 = d0 < d1 < d2 < . . . < dp = d the sequence (Aj)di j=di−1+1 is the optimal S1-decomposition of Ci for i = 1, 2, . . . , p. Now we can first deduce ( Aj)d1 j=1 is the optimal S1-decomposition of C1, then deduce that ( Aj)d2 j=d1+1is the optimal S1-decomposition of C2, and so on. We are therefore in the position to apply the induction hypothesis and deduce that for all i = 1, 2, . . . , p, and D (cid:22) Ci and D (cid:22) Ci it follows that ζ(α, D) = ζ(α, D). Our claim follows therefore from our recursive formula (31). As usual in the case that α is a limit ordinal the verification follows easily from the (cid:3) definition of Sα in that case. 20 P. MOTAKIS AND TH. SCHLUMPRECHT Lemma 3.9. Let X be a Banach space, α be a countable ordinal number, B ∈ MAX(Sα) and (xA)A(cid:22)B be vectors in BX . Then 2 min(B) . Proof. Using, Lemma 3.7 part (iii), then part (v) we obtain (38) (cid:13)(cid:13)(cid:13)XA(cid:22)B ζ(α, A′)xA(cid:13)(cid:13)(cid:13) ≤ ζ(α, A)xA − XA(cid:22)B ζ(α, A′)xA(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)XA(cid:22)B ζ(α, A)xA − XA(cid:22)B l1(A′)6=0(cid:0)ζ(α, A) − ζ(α, A′)(cid:1) xA(cid:13)(cid:13)(cid:13)+(cid:13)(cid:13)(cid:13) XA(cid:22)B ≤ (cid:13)(cid:13)(cid:13) XA(cid:22)B ≤ XA(cid:22)B ζ(α, A) + XA(cid:22)B ζ(α, A′) ≤ 2 l1(A′)=0 l1(A′)=0 l1(A′)=0 min(B) ζ(α, A)xA(cid:13)(cid:13)(cid:13)+(cid:13)(cid:13)(cid:13) XA(cid:22)B l1(A′)=0 ζ(α, A′)xA(cid:13)(cid:13)(cid:13) . (cid:3) 4. Trees and their indices Let X be an arbitrary set. We set X <ω =S∞ n=0 X n, the set of all finite sequences in X, which includes the sequence of length zero denoted by ∅. For x ∈ X we shall write x instead of (x), i.e., we identify X with sequences of length 1 in X. A tree on X is a non-empty subset F of X <ω closed under taking initial segments: if (x1, . . . , xn) ∈ F and 0 ≤ m ≤ n, then (x1, . . . , xm) ∈ F. A tree F on X is hereditary if every subsequence of every member of F is also in F. Given x = (x1, . . . , xm) and y = (y1, . . . , yn) in X <ω, we write (x, y) for the concatenation of x and y: Given F ⊂ X <ω and x ∈ X <ω, we let (x, y) = (x1, . . . , xm, y1, . . . , yn) . F(x) = {y ∈ X <ω : (x, y) ∈ F} . Note that if F is a tree on X, then so is F(x) (unless it is empty). Moreover, if F is hereditary, then so is F(x) and F(x) ⊂ F. Let X ω denote the set of all (infinite) sequences in X. Fix S ⊂ X ω. For a subset F of S of F consists of all x = (x1, x2, . . . , xl) ∈ X <ω for which there is a X <ω the S-derivative F ′ sequence (yi)∞ i=1 ∈ S with (x, yi) ∈ F for all i ∈ N. Note that if F is a hereditary tree then it follows that F ′ S ⊂ F and that F ′ S is also a hereditary tree (unless it is empty). We then define higher order derivatives F (α) S for ordinals α < ω1 by recursion as follows. S = F, F (α+1) F (0) S S for α < ω1 and F (λ) F (α) S for limit ordinals λ < ω1. =(cid:0)F (α) S (cid:1)′ S = \α<λ It is clear that F (α) S ⊃ F (β) S if α ≤ β and that F (α) S is a hereditary tree (or the empty set) for all α, whenever F is a hereditary tree. An easy induction also shows that (cid:0)F(x)(cid:1)(α) S =(cid:0)F (α) S (cid:1)(x) for all x ∈ X <ω, α < ω1 . A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 21 We now define the S-index IS(F) of F by IS(F) = min(cid:8)α < ω1 : F (α) S = ∅, and IS(F) = ω otherwise. S = ∅(cid:9) if there exists α < ω1 with F (α) Remark. If λ is a limit ordinal and F (α) all α < λ, and hence F (λ) Examples 4.1. 1. A family F ⊂ [N]<ω can be thought of as a tree on N: a set F = {m1, . . . , mk} ∈ [N]<ω is identified with (m1, . . . , mk) ∈ N<ω (recall that m1 < . . . < mk by our convention of always listing the elements of a subset of N in increasing order). S 6= ∅. This shows that IS(F) is always a successor ordinal. 6= ∅ for all α < λ, then in particular ∅ ∈ F (α) for S S Let S be the set of all strictly increasing sequences in N. In this case the S-index of a compact, family F ⊂ [N]<ω is nothing else but the Cantor-Bendixson index of F as a compact topological space, which we will continue to denote by CB(F). We will also use the term Cantor-Bendixson derivative instead of S-derivative and use the notation F ′ and F (α). 2. If X is an arbitrary set and S = X ω, then the S-index of a tree F on X is what is usually called the order of F (or the height of F) denoted by o(F). Note that in this case the S-derivative of F consists of all finite sequences x ∈ X <ω for which there exists y ∈ X such that (x, y) ∈ F. The function o(·) is the largest index: for any S ⊂ X ω we have o(F) ≥ IS (F). We say that S ⊂ X ω contains diagonals if for any sequence (xn) in S with xn = (xn,i)∞ i=1 there exist i1 < i2 < . . . in N so that (xn,in)∞ n=1 belongs to S. In particular every subsequence of every member of S belongs to S. If S contains diagonals, then the S-index of a tree on . X may be measured via the Cantor-Bendixson index of the fine Schreier families(cid:0)Fα(cid:1)α<ω1 Proposition 4.2. [27, Proposition 5] Let X be an arbitrary set and let S ⊂ X ω. If S contains diagonals, then for a hereditary tree A on X and for a countable ordinal α the following are equivalent. a) α < IS(A). b) There is a family (cid:0)xF(cid:1)F ∈Fα\{∅} ⊂ A such that for F = (m1, m2, . . . , mk) ∈ Fα the branch xF =(cid:0)x{m1}, x{m1,m2}, . . . , x{m1,m2,...,mk}(cid:1) is in A, and (cid:0)xF ∪{n}(cid:1)n>max F is in S, if F is not maximal in Fα. Definition 4.3. Let F ⊂ [N]<ω be regular, S a set of sequences in the set X, and and (xA : A ∈ F) a tree in X indexed by F. We call (xA : A ∈ F) an S- tree if for every non maximal A ∈ F the sequence (xA∪{n} : n ∈ N, with A ∪ {n} ∈ F) is a sequence in S. If X is a Banach space and S are the w-null sequences, we call (xA : A ∈ F) a w-null tree. Similarly we define w∗-null trees in in X∗. Remark 4.4. In the case of X = N and S = [N]ω we deduce from Proposition 4.2 that if A ⊂ [N]<ω is hereditary and compact, then CB(A) > α if and only if there is an order isomorphism π : Fα → A, so that for all A ∈ Fα \MAX(Fα) and n > max(A) it follows that π(A ∪ {n}) = π(A) ∪ {sn}, where (sn) is an increasing sequence in {s ∈ N : s > max π(A)}. Examples 4.5. 1. The weak index. Let X be a separable Banach space. Let S be the set of all weakly null sequences in BX, the unit ball of X. We call the S-index of a tree F on X the weak index of F and we shall denote it by Iw(F). We shall use the term weak derivative instead of S-derivative and use the notation F ′ w and F (α) w . 22 P. MOTAKIS AND TH. SCHLUMPRECHT When the dual space X∗ is separable, the weak topology on the unit ball BX , or on any bounded subset of X is metrizable. Hence in this case the set S contains diagonals and Proposition 4.2 applies. 2.The weak∗ index. We can define the weak∗ index similarly to the weak index. If X is a separable Banach space the w∗ topology on B∗ X is metrizable. This implies that the set S of all w∗ null sequences in BX ∗ is diagonalizable. We call the S index of a tree F on X∗ the weak∗ index of F and we shall denote it by Iw∗(F). We shall use the term weak∗ derivative instead of S-derivative and use the notation F ′ 3. The block index. Let Z be a Banach space with an FDD E = (Ei). A block tree of (Ei) in Z is a tree F such that every element of F is a (finite) block sequence of (Ei). Let S be the set of all infinite block sequences of (Ei) in BZ . We call the S-index of a block tree F of (Ei) the block index of F and we shall denote it by Ibl(F). We shall use the term block derivative instead of S-derivative and use the notation F ′ . Note that the set S contains diagonals, and hence Proposition 4.2 applies. w∗ and F (α) w∗ . bl and F (α) bl Note that (Ei) is a shrinking FDD of X if and only if every element of S is weakly null. In this case we have (40) Ibl(F) ≤ Iw(F) for any block tree F of (Ei) in Z. The converse is false in general, but it is true up to perturbations and without the assumption that (Ei) is shrinking. 5. The Szlenk index Here we recall the definition and basic properties of the Szlenk index, and prove further properties that are relevant for our purposes. Let X be a separable Banach space, and let K be a non-empty subset of X∗. For ε ≥ 0 and define K (α) for each countable ordinal α by recursion as follows: n) ⊂ K w∗ − lim n→∞ x∗ n = x∗ and kx∗ n − x∗k > εo. set K ′ ε ε =nx∗ ∈ X∗ : ∃(x∗ =(cid:0)K (α) ε (cid:1)′ ε = \α<λ K (0) ε = K, K (α+1) ε ε for α < ω1, and K (λ) K (α) ε for limit ordinals λ < ω1. Next, we associate to K the following ordinal indices: η(K, ε) = sup{α < ω1 : K (α) ε 6= ∅} , and η(K) = sup ε>0 η(K, ε) . Finally, we define the Szlenk index Sz(X) of X to be η(BX ∗ ), where BX ∗ is the unit ball of X∗. Remark. The original definition of the derived sets K ′ ε in [35] is slightly different. It might lead to different values of Sz(K, ε), but to same values of Sz(K), and in particular to the same values of Sz(X). Szlenk used his index to show that there is no separable, reflexive space universal for the class of all separable, reflexive spaces. This result follows immediately from the following properties of the function Sz(·). Theorem 5.1. [35] Let X and Y be separable Banach spaces. (i) X∗ is separable if and only if Sz(X) < ω1. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 23 (ii) If X isomorphically embeds into Y , then Sz(X) ≤ Sz(Y ). (iii) For all α < ω1 there exists a separable, reflexive space with Szlenk index at least α. We also recall the following observation of [2] about the form of the Szlenk index of a Banach space with separable dual. Proposition 5.2. [2, Theorems 3.22 and 4.2] If X has a separable dual then there is an α < ω1 with Sz(X) = ωα. The following Theorem combines several equivalent description of the Szlenk index of a separable space not containing ℓ1 Theorem 5.3. Assume that X is a separable space not containing ℓ1 and α < ω1. The following conditions are equivalent. (i) Sz(X) > ωα. (ii) There is an ε > 0 and a tree (z∗ A : A ∈ Sα) ⊂ BX ∗, so that for any non maximal (41) A ∈ Sα we have w∗ − lim n→∞ z∗ A∪{n} = z∗ (iii) There is an ε > 0, a tree (z∗ A and(cid:13)(cid:13)z∗ A − z∗ A∪{n}(cid:13)(cid:13) > ε for n > max(A). A : A ∈ Sα) ⊂ B∗ X and a w-null tree (zA : A ∈ Sα) ⊂ BX , so that (42) (43) (44) z∗ B(zA) > ε, for all A, B ∈ Sα \ {∅}, with A (cid:22) B, (cid:12)(cid:12)z∗ B(zA)(cid:12)(cid:12) < ε/2 for all A, B ∈ Sα \ {∅}, with A 6(cid:22) B, and for all non maximal A ∈ Sα we have w∗ − lim n→∞ z∗ A∪{n} = z∗ A. (iv) There is an N ∈ [N]ω, an ε > 0 and a w-null tree tree (xA : A ∈ Sα ∩ [N ]<ω), so that for every maximal B in Sα ∩ [N ]<ω we have Fε = Gε =((x∗ (cid:13)(cid:13)(cid:13)XA(cid:22)B ζ(α, A)xA(cid:13)(cid:13)(cid:13) ≥ ε. j=1 ⊂ [0, 1] (cid:13)(cid:13)(cid:13) j k ≥ ε and(cid:13)(cid:13)(cid:13) jXi=1 x∗ . lXj=1 aj ajxj(cid:13)(cid:13)(cid:13) ≥ ε lXj=1 i(cid:13)(cid:13)(cid:13) ≤ 1, for j = 1, 2, . . . , l) . (v) There is an ε > 0, so that Iw(Fε) > ωα, where (x1, x2, . . . , xl) ⊂ SX : ∀(aj)l (vi) There is an ε > 0 so that Iw∗(Gε) > ωα, where 1, x∗ 2, . . . , x∗ l ) ⊂ BX ∗ : kx∗ Proof. In order to show "(i)⇒(ii)" we first prove the following Lemma 5.4. Let X be a separable Banach space and K ⊂ X∗ be w∗-compact, 0 < ε < 1, and β < ω1. Then for every x∗ ∈ K (β) (45) (x∗,A) : A ∈ Fβ) ⊂ K, so that there is a family (z∗ ε (x∗,∅) = x∗, z∗ (x∗,A) ∈ K and z∗ if A is not maximal in Fβ, then kz∗ if A is not maximal in Fβ, then z∗ (46) (47) (x∗,A∪{n})k > ε for all n > max(A), (x∗,A)−z∗ (x∗,A) = w∗ − lim n→∞ z∗ (x∗,A∪{n}). 24 P. MOTAKIS AND TH. SCHLUMPRECHT Proof. We will prove our claim by transfinite induction for all β < ω1. Let us first assume that β = 1. For x∗ ∈ K ′ nk > ε, for n ∈ N. Thus, we can choose z(x∗,∅) = x∗, and z(x∗,{n}) = x∗ n. This choice satisfies n) which w∗-converges to x∗, with kx∗ − x∗ ε choose a sequence (x∗ Now assume that our claim is true for all γ < β. First assume that β is a successor . Thus there is a sequence n − x∗k > ε, for n ∈ N. By our induction ordinal and let γ < ω1 so that β = γ + 1. Let x∗ ∈ K (γ+1) (x∗ (45), (46), and (47), for β = 1 (recall that F1 =(cid:8){n} : n ∈ N(cid:9) ∪ {∅}). hypothesis we can choose for each n ∈ N a family(cid:0)z∗ n,A) : A ∈ Fγ(cid:1) satisfying (45), (46), n instead of x∗. For every A ∈ Fγ+1 it follows that A \ {min A} ∈ Fγ ε which w∗-converges to x∗, with kx∗ n) ⊂ K (γ) (x∗ ε and (47), for γ and x∗ and we define z∗ (x∗,∅) := x∗ and for A ∈ Fγ+1 \ {∅} z∗ (x∗,A) := z∗ (x∗ min A,A\{min(A)}). (x∗,A) : A ∈ Fγ+1(cid:1) satisfies (45), (46), and (47). It is then easy to see that(cid:0)z∗ Assume that β < ω1 is a limit ordinal and let(cid:0)µ(β, n) : n ∈ N(cid:1) ⊂ (0, β), be the sequence ε = Tγ<β K (γ) of ordinals increasing to β, used to define Fβ. We abbreviate βn = µ(β, n), for n ∈ N. Let x∗ ∈ K (β) . Since βn + 1 < β, we can use for each n ∈ N our induction hypothesis and choose a family ε (n,x∗,A) : A ∈ Fβn+1(cid:1) ⊂ X∗, (cid:0)z∗ satisfying (45), (46), and (47), for βn + 1. In particular it follows that w∗ − lim j→∞ z∗ (n,x∗,{j}) = x∗, for all n ∈ N. Since the w∗-topology is metrizable on K we can find an increasing sequence (jn : n ∈ N) in N, jn > n, for n ∈ N, so that w∗ − lim n→∞ z∗ (n,x∗,{jn}) = x∗. Consider for n ∈ N the set Fβn+1(jn) =(cid:8)A ∈ [N]<ω : jn < min A and {jn} ∪ A ∈ Fβn+1(cid:9) = {A ∈ Fβn : jn < min A}. Since Fβn is spreading, for n ∈ N, we can choose Ln = {l(n) 2 , . . .} ∈ [N]ω so that 1 , l(n) F Ln βn a1 , l(n) a2 , . . . , l(n) =(cid:8){l(n) am } : {a1, a2, . . . , am} ∈ Fβn(cid:9) ⊂ Fβn+1(jn). We define the map φn : Fβn → Fβn+1(jn), {a1, a2, . . . , am} 7→ {l(n) a1 , l(n) a2 , . . . , l(n) am }. Then we put for A ∈ Fβ z∗ x∗ z∗ (x∗,{jn}) z∗ (x∗,{jn}∪φn(B)) (x∗,A) = Sz(X) > β = ωα, it follows that (cid:2)BX ∗(cid:3)(β) ε which has the wanted property also in that case. We now continue with our proof of "(i)⇒(ii)" of Theorem 5.3. Assuming now that and apply Lemma 5.4 to obtain a tree in BX ∗ indexed by Fωα satisfying the conditions (45), (46), and (47). Now Proposition 2.12 and a relabeling of the tree, yields (ii). 6= ∅ for some ε > 0. We choose x∗ ∈(cid:2)BX ∗(cid:3)(β) ε if A = ∅, if A = {n} for some n ∈ N, and if A = {n} ∪ B for some n ∈ N and B ∈ Fαn \ {∅}, (cid:3) A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 25 "(ii)⇒(iii)" For A ∈ Sα\{∅} we define A′ = A \ {max(A)}. Now let (Am)m∈N be a consistent ordering of Sα (see Subsection 2.4) . We write A <lin B or A ≤lin B if A = Am and B = An, for m < n or m ≤ n, respectively. Let ε > 0 and (z∗ A : A ∈ Sα) ⊂ BX ∗, so that (41) is satisfied. Then choose for each A ∈ Sα \ {∅} an element xA ∈ SX so that (z∗ Let 0 < η < ε/8, and let(cid:0)η(A) : A ∈ Sα(cid:1) ⊂ (0, 1) satisfy the following conditions A − z∗ A′)(xA) > ε. (48) (49) (50) (51) η(A) < η, (cid:0)η(A)(cid:1) is decreasing with respect to the linear ordering <lin, XA∈Sα XB∈Sα,B>linA η(B) < η(A), for all A ∈ Sα, and η(Am) < 1 2 η m + 2 , for all m ∈ N. Since X does not contain a copy of ℓ1 we can apply Rosenthal's ℓ1 Theorem, and as- sume, possibly after passing to a pruning, that for each non maximal A ∈ Sα the sequence A)n>max(A) is w∗-null we can assume, (xA∪{n})n>max(A)) is weakly Cauchy. Since (z∗ possibly after passing to a further pruning, that (z∗ A)(xA∪{n−1}) < η(A ∪ {n}), for all non maximal A ∈ Sα and n > 1 + max(A). A∪{n} − z∗ A∪{n} − z∗ Let z∅ = 0. For a non maximal element A ∈ Sα and n > 1 + max(A) let and zA∪{n} = 1 2 (xA∪{n} − xA∪{n−1}) zA∪{max(A)+1} = xA∪{max(A)+1}. Then the families (zA : A ∈ Sα) and (z∗ (zA : A ∈ Sα) is weakly null and (z∗ A : A ∈ Sα) satisfies (44). A : A ∈ Sα) are in BX and BX ∗ respectively, Moreover it follows that (52) (z∗ A − z∗ A′)(zA) ≥ ε 2 − η(A), for all A ∈ Sα \ {∅}. Since w − limn→∞ zB∪{n} = 0 and w∗ − limn→∞ z∗ B∪{n} = z∗ B, for every non maximal B ∈ Sα we can, after passing again to a pruning, assume that (53) B − z∗ B′)(zA)(cid:12)(cid:12) <η(B) and(cid:12)(cid:12)z∗ A(zB)(cid:12)(cid:12) < η(B) for all A, B ∈ Sα, with A <lin B. We are left with verifying (42) and (43) for ε/4 instead of ε. To show (42) let A, B ∈ Sα \ {∅}, with A (cid:22) B. We choose l ∈ N and A = B0 ≺ B1 ≺ B2 ≺ (cid:12)(cid:12)(z∗ . . . ≺ Bl = B, so that B′ j = Bj−1, for j = 1, 2, . . . , l, and deduce from (52) and (53) z∗ B(zA) = (z∗ Bj − z∗ B′ j )(zA) + (z∗ A − z∗ A′)(zA) + z∗ A′(zA) ≥ ε 2 − η(Bj) − 2η(A) > ε 2 − 2η > ε 4 . lXj=1 lXj=1 In order to show (43) let A, B ∈ Sα \ {∅}, with A 6(cid:22) B. We choose l ∈ N and ∅ = j = Bj−1, for j = 1, 2 . . . , l and, since for every B0 ≺ B1 ≺ B2 ≺ . . . ≺ Bl = B so that B′ 26 P. MOTAKIS AND TH. SCHLUMPRECHT j = 1, 2, . . . , l, either A <lin Bj or Bj <lin A we deduce from (53) and the conditions (49) and (51) on η(·), that z∗ B(zA) ≤(cid:12)(cid:12)(cid:12) (z∗ ∅(zA) Bj − z∗ B′ j lXj=1 ≤ XA<linBj(cid:12)(cid:12)(z∗ ≤ XA<linBj )(zA)(cid:12)(cid:12)(cid:12) + z∗ )(zA)(cid:12)(cid:12) + XA>linBj(cid:0)(z∗ η(Bj) + 2 XBj <linA Bj − z∗ B′ j Bj (zA) + z∗ Bj−1 (zA)(cid:1) + η(A) η(A) + η(A) ≤ (2#{j ≤ l : Bj < linA}+2)η(A) < ε 4 which verifies (43) and finishes the proof of our claims. "(iii)⇒(iv)" Let ε > 0, (z∗ it follows for a maximal B ∈ Sα that A : A ∈ Sα) , and (zA : A ∈ Sα) satisfy the condition in (iii). Then (cid:13)(cid:13)(cid:13)XA(cid:22)B ζ(α, A)zA(cid:13)(cid:13)(cid:13) ≥ XA(cid:22)B ζ(α, A)z∗ B(zA) ≥ εXA(cid:22)B ζ(α, A) = ε, which proves our claim. "(iv)⇒(v)" Assume that N ∈ [N]ω, ε > 0, and (xA : A ∈ Sα ∩ [N ]<ω) ⊂ BX satisfy (iv). B ∈ SX ∗ so that y∗ B(yB) = kyBk > ε. For B ∈ MAX(Sα) ∩ [N ]<ω, we define fB : B → [−1, 1], b 7→ y∗ Corollary 3.5 it follows now for δ = ε/2 that CB(Aδ,N ) = ωα + 1 where For B ∈ MAX(Sα ∩ [N ]<ω) put yB = PA≺B ζ(α, A)xA, and choose y∗ A ⊂ B, and fB(a) ≥ δ for all a ∈ A(cid:27) . Aδ,N =(cid:26)A ∈ Sα ∩ [N ]<∞ : ∃B ∈ MAX(Sα ∩ [N ]<∞), B(x{a∈B,a≤b}). From But from Proposition 4.2 and the Remark thereafter we deduce that there is a order iso- morphism π : Fωα → Aδ,N so that for every non maximal A ∈ Fωα and any n > max(A) it follows that π(A ∪ {n}) = π(A) ∪ {sn}, for some increasing sequence (sn) ⊂ N. Putting zA = xπ(A) it follows that (zA)A∈Fωα is a weakly null tree and for every A = {a1, a2, . . . , al} it follows that (z{a1,a2,...,ai})l "(v) ⇐⇒ (i)" This follows from [2, Theorem 4.2] where it was shown that Sz(X) = supε>0 Iw(Fε), if ℓ1 does not embed into X. "(ii) ⇐⇒ (vi)" Follows from Proposition 2.13 and an application of Proposition 4.2 to the tree Gε on BX ∗, and and S = {(x∗ (cid:3) i=1 ∈ Fδ. Applying again Proposition 4.2 yields (v). X : w∗ − limn→∞ = 0}. n) ⊂ B∗ Remark. We note that in the implication (i)⇒(ii) the assumption that ℓ1 does not embed into X was not needed. We will also need the following dual version of Theorem 5.3. Proposition 5.5. Assume that X is a Banach space whose dual X∗ is separable, with Sz(X∗) > ωα. Then there is an ε > 0, a tree (zA : A ∈ Sα) ⊂ BX, and a w∗-null tree (z∗ A : A ∈ Sα) ⊂ B∗ X so that (54) (55) z∗ A(zB) > ε, for all A, B ∈ Sα \ {∅}, with A (cid:22) B, ε 2 for all A, B ∈ Sα \ {∅}, with A 6(cid:22) B. (cid:12)(cid:12)z∗ A(zB)(cid:12)(cid:12) < A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 27 Proof. Recall that, as stated above, the implication (i)⇒(ii) of Theorem 5.3 holds even if the space, to which the theorem is applied, contains ℓ1. Applying this implication to X∗, we find ε > 0 and {z∗∗ A : A ∈ Sα} ⊂ BX ∗∗, so that (41) is satisfied. Then choose for each A ∈ Sα \ {∅} an element x∗ A) > ε. Again let (An) be a consistent enumeration of Sα, and write Am <lin An if m < n. We also assume that (η(A) : A ∈ Sα) ⊂ (0, 1), has the property that A ∈ SX so that (z∗∗ A − z∗∗ A′)(x∗ (56) η(A) < ε 32 . XA∈Sα A∪{n}) w∗-converges, and that for any B ∈ Sα the sequance z∗∗ After passing to a first pruning we can assume that for all non maximal A ∈ Sα the sequence (x∗ A∪{n}) converges so some number rA,B (for fixed A, B ∈ Sα we only need to pass to a subsequence of (A ∪ {n} : n ∈ N, A ∪ {n} >lin B)). Since (z∗∗ A )n>max(A) is w∗-null we can assume, after passing to a second pruning, that we have A∪{n} − z∗∗ B (x∗ (57) B − z∗∗ B′)(x∗ We put z∗ ∅ = 0 and for any non maximal element A ∈ Sα A)(cid:12)(cid:12) < η(B) for all A, B ∈ Sα, with A <lin B, (cid:12)(cid:12)(z∗∗ 1 2 z∗ A∪{max(A)+1} = x∗ A∪{max(A)+1} and z∗ A∪{n} = (x∗ A∪{n} − x∗ A∪{n−1}) if n > max(A) + 1. It follows that (z∗ A : A ∈ Sα) is a w∗-null tree in BX ∗ and that for any A ∈ Sα (z∗∗ A − z∗∗ A′)(z∗ A) > ε 2 − η(A) 2 . Since z∗∗ B (x∗ A∪{n}) converges to rA,B we can assume, after passing to a third pruning, that (58) and hence (59) z∗∗ A (z∗ B) < η(B) 2 whenever A <lin B. z∗∗ A (z∗ A) > ε 2 − η(A) for all A ∈ Sα. Since w∗-limn z∗∗ A we can assume, by passing to a further pruning, that Since BX is w∗-dense in BX ∗∗ we can choose, for every A ∈ Sα, a vector zA ∈ BX so that A∪{n} = z∗∗ B′)(z∗ B − z∗∗ (cid:12)(cid:12)(z∗∗ z∗ A(zB) − z∗∗ B (z∗ A)(cid:12)(cid:12) < η(B) for all A, B ∈ Sα, with A <lin B. A) < η(B), for all A, B ∈ Sα, with A ≤lin B. (60) (61) Combining (60) and (61) we obtain that for all A and B in Sα, with A <lin B we have (62) z∗ A(zB) − z∗∗ B′ (z∗ A) 6 z∗ A(zB) − z∗∗ B (z∗ A) + (z∗∗ B − z∗∗ B′)(z∗ A) < 2η(B). Using that (z∗ A : A ∈ Sα) is a w∗-null tree, we can pass to a further pruning, so that z∗ (63) We deduce from (60) and (61), for A, B ∈ Sα, with A ≤lin B′(<lin B) that (64) B(zA) < η(B), for all A, B ∈ Sα with A <lin B. A(zB −zB′) ≤ z∗ A(zB′ )−z∗∗ A) + (z∗∗ A) + z∗ B′(z∗ z∗ B −z∗∗ B′)(z∗ A) A(zB)−z∗∗ B (z∗ ≤ 2η(B) + η(B′). By (62), (58), and (63) for A, B ∈ Sα, with B′ <lin A < B we obtain (65) A(zB −zB′) 6 z∗ A(zB)−z∗∗ A) + z∗∗ A) + z∗ B′(z∗ B′(z∗ z∗ A(zB′ ) ≤ 2η(B)+2η(A). 28 P. MOTAKIS AND TH. SCHLUMPRECHT We now claim that the families {zA : A ∈ Sα} and {z∗ A : A ∈ Sα} satisfy (54) and (55). In order to verify (54), let A, B ∈ Sα \ {∅}, with A (cid:22) B. Then let k ∈ N and Bj ∈ Sα, j = Bj, for for j = 0, 1, 2, . . . , k, be such that A = B0 ≺ B1 ≺ B2 ≺ · · · ≺ Bk = B, and B′ j = 1, 2, . . . , k. z∗ A(zB) = kXj=1 z∗ A(zBj − zB′ j ) + z∗ A(zA) ≥ z∗∗ A (z∗ A) − z∗∗ A (z∗ A) − z∗ A(zA) − z∗ A(zB1 − zA) − z∗ A(zBj − zB′ ) j kXj=2 (By (59), (61), (64) and (62)) > ε 2 − 3η(A) − 2 η(Bj) ≥ ε 4 . kXj=1 which yields (54) if we replace ε by ε/4. In order to verify (55), let A, B ∈ Sα \ {∅}, with A 6(cid:22) B. If A >lin B we deduce our claim from (63). If A ≤lin B, and thus A <lin B, we choose k ∈ N and B0 ≺ B1 ≺ B2 ≺ · · · ≺ Bk = B, with B′ j = Bj−1, for j = 1, 2, . . . , k, and B0 <lin A <lin B1. Applying (65),(63), (64), and finally (56) we obtain (cid:12)(cid:12)z∗ A(zB)(cid:12)(cid:12) ≤ z∗ A(zB1 − zB0) + z∗ z∗ A(zBj − zB′ j kXj=2 )(cid:12)(cid:12)(cid:12) A(zB0) +(cid:12)(cid:12)(cid:12) kXj=2 ≤ 2η(B1)+η(B0)+3η(A)+ (2η(Bj ) + η(Bj−1)) ≤ ε 8 (cid:3) which proves our claim. Example 5.6. Let us construct an example of families (cid:0)zA : A ∈ [N]<ω(cid:1) ⊂ Bc0 and (cid:0)z∗ A : A ∈ [N]<ω(cid:1) ⊂ Bℓ1 satisfying Proposition 5.5. Let <lin again be a linear consistent ordering of [N]<ω. We first choose a family ( A : A ∈ [N]<ω(cid:1) ⊂ [N]<ω with the following properties (66) (67) (68) A is a spread of A, for each A ∈ [N]<ω, A ≺ B if and only if A ≺ B, if A, B ∈ [N]<ω, ∅ 6= A <lin B, and C ∈ [N]<ω is the maximal element in [N]<ω, so that C (cid:22) A and C (cid:22) B, then ( A \ C) ∩ ( B \ C) = ∅. We define for A ∈ [N]<ω zA =Xa∈ A ea and z∗ A = e∗ max( A), where (ej ) and (e∗ j ) denote the unit vector bases in c0 and ℓ1 respectively. It is now easy to verify that the tree (z∗ A) is w∗-null and that (54) is satisfied for any ε ∈ (0, 1). In order to verify (55) let A, B ∈ [N]<ω with A 6(cid:22) B. If A >lin B, then max( A) 6∈ B, and our claim follows. If A <lin B let C ∈ [N]<ω be the maximal element for which C (cid:22) A and C (cid:22) B. It follows that C ≺ A, but also that C ≺ B, which implies by (68) that max( A) 6∈ B, and, thus our claim. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 29 6. Estimating certain convex combinations of blocks using the Szlenk index as sum x = P∞ In this section we will assume that X has an FDD (Fj ). This means that Fj ⊂ X is a finite dimensional subspace of X, for j ∈ N, and that every x has a unique representation j=1 xj ∈ X we call supp(x) = {j ∈ N : xj 6= 0} the support of x (with respect to (Fj)), and the smallest interval in N containing supp(x) is called the range of x (with respect to (Fj)), and is denoted by ran(x). A (finite or infinite) sequence (xn) ⊂ X is called a block (with respect to (Fj)), if xn 6= 0, for all n ∈ N, and supp(xn) < supp(xn+1), for all n ∈ N for which xn+1 is defined. j=1 xj, with xj ∈ Fj, for j ∈ N. For x = P∞ We call an FDD shrinking if every bounded block (xn)∞ n=1 is weakly null. As in the case of bases, X∗ is separable, and thus Sz(X) < ω1, if X has a shrinking FDD. Theorem 6.1. Let X be a Banach space with a shrinking FDD and α be a countable ordinal number with Sz(X) ≤ ωα. Then for every ε > 0 and M ∈ [N]ω, there exists N ∈ [M ]ω satisfying the following: for every B = {b1, . . . , bd} in Sα ∩ [N ]<ω and sequence (xi)d i=1 in BX, with ran(xj) ⊂ (bj−1, bj+1) for j = 1, . . . , d (where b0 = 0 and bd+1 = ∞), we have (69) (cid:13)(cid:13)(cid:13) dXj=1 ζ(α, Bj)xj(cid:13)(cid:13)(cid:13) < ε, where Bj = {b1, . . . , bj} for j = 1, . . . , d. Proof. It is enough to find N in [M ]ω so that (69) holds whenever B ∈ MAX(Sα) ∩ [N ]<ω. Indeed, if (69) holds for all B in MAX(Sα)∩ [N ]<ω, then for any A ∈ Sα ∩ [N ]<ω and (xi)#A i=1 satisfying the assumption of Theorem 6.1, one may extend A to any maximal set B and extend the sequence (xk)#A k=1 by concatenating the zero vector #B − #A times. Towards a contradiction, we assume that such a set N does not exist. Applying Proposition 2.14 to the Partition (F, Sα \ F) of Sα, where ∃(xj) ⊂ BX , ran(xj) ⊂ (bj−1, bj+1), for j = 1, 2, . . . , n F =(cid:26)B = {b1, b2, . . . , bn} ∈ MAX(Sα) : yields that there is L in [M ]ω, so that for all B = {bB exists a sequence (xB i=1 in BX with ran(xj) ⊂ (bB i )dB (70) ζ(α, BB j )xB (cid:27) , } in MAX(Sα) ∩ [L]<ω, there j+1) for j = 1, . . . , dB, so that 1 , . . . , bB dB j−1, bB (cid:13)(cid:13)Pn j=1 ζ(α, {b1, b2, . . . , bj})xj(cid:13)(cid:13) > ε j(cid:13)(cid:13)(cid:13) ≥ ε, A(cid:13)(cid:13)(cid:13) ≥ ε (cid:13)(cid:13)(cid:13) dBXj=1 (cid:13)(cid:13)(cid:13)XA(cid:22)B ζ(α, A)xB where BB xB j = xB 1 , . . . , bB j = {bB A. Note that, under this notation, (70) takes the more convenient form j } for j = 1, . . . , dB. For A (cid:22) B, if A = BB j , we use the notation (71) and that (72) where A′′ = (A′)′ and max(∅) = 0. A′) ⊂(cid:0)max(A′′), max(A)(cid:1) , for all A (cid:22) B with A′ 6= ∅, ran(xB We will now apply several stabilization and perturbation arguments to show that we may A′, only depends on A., and will the assume that for B ∈ MAX(Sα) and A (cid:22) B the vector xB be renamed xA. 30 P. MOTAKIS AND TH. SCHLUMPRECHT A′ −xA∪D2 Using a compactness argument, Proposition 2.14 yields the following: if A ∈ Sα ∩ [L]<ω 6= ∅, then for δ > 0, there is L′ ∈ [L]ω so that for all D1, D2 in is non-maximal with A′ MAX(Sα(A))∩[L′]<ω, we have kxA∪D1 k < δ. Combining the above with a standard diagonalization argument we may pass to a further infinite subset of L, and a perturbation A′ with B ∈ MAX(Sα) ∩ [L]<ω and {min(B)} ≺ A (cid:22) B (and perhaps of the block vectors xB pass to a smaller ε in (71)), so that for every B1, B2 in MAX(Sα) ∩ [L]<ω and A, with A′ 6= ∅, so that A (cid:22) B1 and A (cid:22) B2, we have xB1 A′ . For every A ∈ Sα ∩ [L]<ω, we call this common vector xA. Note that xA indeed depends on A and not only on A′. For A so that A′ = ∅, i.e. for those sets A that are of the form A = {n} for some n ∈ L, chose any normalized vector xA with supp(xA) = {n}. Note that, using (72), we have A′ = xB2 A′ ran(xA) ⊂ (max(A′′), max(A)) for all A ∈ Sα ∩ [L]<ω with A′ 6= ∅, (73) where max(∅) = 0 and if A′ = 0, i.e. A = {n} for some n ∈ N, then ran(xA) = {n}. Furthermore, fixing 0 < δ < ε/12 and passing to an infinite subset of of L, again denoted by L, satisfying min(L) ≥ 1/δ, (71) and (38) yield that for all B ∈ MAX(Sα) ∩ [L]<ω (74) (cid:13)(cid:13)(cid:13)XA(cid:22)B ζ(α, A′)xA(cid:13)(cid:13)(cid:13) − 2δ ζ(α, A)xA(cid:13)(cid:13)(cid:13) ≥ (cid:13)(cid:13)(cid:13)XA(cid:22)B = (cid:13)(cid:13)(cid:13)0x{min(B)} + X{min(B)}≺A(cid:22)B = (cid:13)(cid:13)(cid:13)XA≺B A(cid:13)(cid:13)(cid:13) − 2δ ≥ ε − 3δ. ζ(α, A)xB For B ∈ MAX(Sα ∩ [L]<ω) and i = 0, 1, 2 define ζ(α, A′)xB A′(cid:13)(cid:13)(cid:13) − 2δ B(i) = {A (cid:22) B : #A mod 3 = i}. By the triangle inequality, for some 0 ≤ i(B) ≤ 2, we have kPA∈B(i(B)) ζ(α, A)xAk ≥ ε/3−δ. By Proposition 2.14, we may pass to some infinite subset of L, again denoted L, so that for all B ∈ MAX(Sα ∩ [L]<ω), we have i(B) = i0, for some common i0 ∈ {0, 1, 2}. We shall assume that i0 = 0, as the other cases are treated similarly. Therefore, for all B ∈ MAX(Sα ∩[L]<ω) we have Lemma 3.7 parts (iii) and (v) also imply the following. If B ∈ MAX(Sα ∩ [L]<ω), then ε 3 − δ. (cid:13)(cid:13)(cid:13) XA∈B(0) ζ(α, A) + XA∈B(0): ζ(α, A)xA(cid:13)(cid:13)(cid:13) ≥ ζ(α, A) + XA∈B(0) l1(A′′)=0 l1(A′′′)=0 (76) XA∈B(0): l1(A′)=0 ζ(α, A) ≤ 3 min(B) ≤ 3δ. (75) then (77) Hence, if for B ∈ MAX(Sα ∩ [L]<ω) we set B(0) = B(0) \ {A (cid:22) B : l1(A′) = 0, or l1(A′′) = 0, or l1(A′′′) = 0}, (cid:13)(cid:13)(cid:13) XA∈ B(0) ζ(α, A)xA(cid:13)(cid:13)(cid:13) ≥ ε 3 − 4δ. If L = {ℓ1, ℓ2, . . . , ℓk, . . .}, define N = {ℓ3, ℓ6, . . . , ℓ3k, ℓ3(k+1), . . .}. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 31 For each A ∈ Sα ∩ [N ]<ω, with (#A)mod 3 = 0, we define A ∈ [L]<ω as described bellow. If A = {a1, . . . , ad}, where aj = ℓ3bj and Aj = {a1, . . . , aj} for 1 ≤ j ≤ d, we define the elements of a set A = {a1, a2, . . . , ad} in groups of three as follows. If jmod 3 = 0 we put (aj−2, aj−1, aj) =((aj−2, aj−1, aj) (ℓ3bj −2ℓ3bj −1, aj) if l1(A′ if l1(A′ j) = 0 or l1(A′ j), l1(A′ j−1), l1(A′ j−2) 6= 0. j−1) = 0 or l1(A′ j−2) = 0 It is not hard to see that A and A satisfy the assumptions of Lemma 3.8, hence A ∈ Sα∩[L]<ω and ζ(α, A) = ζ(α, A). Observe the following: (a) If B ∈ MAX(Sα ∩ [N ]<ω) and A(1) ≺ A(2) are in B(0), then A(1) ≺ A(2). (b) If B ∈ MAX(Sα ∩ [N ]<ω) and A ∈ B(0), if max(A) = ℓ3n then we have max( A′′) = ℓ3n−2. (78) (a) is clear, while (b) follows from the fact that A ∈ B(0), implies that d is divisible by 3 and l1(A′′) 6= 0. We shall define a weakly null tree (zA)A∈Sα∩[N ]<ω so that for all B ∈ MAX(Sα ∩ [N ]∞) we have (79) (cid:13)(cid:13)(cid:13)XA(cid:22)B ζ(α, A)zA(cid:13)(cid:13)(cid:13) ≥ ε 3 − 4δ. ζ(α, A)xA(cid:13)(cid:13)(cid:13) ≥ ε 3 −4δ. The choice of δ and statement (iv) of Theorem 5.3 will yield a contradiction. For A ∈ Sα ∩ [N ]<ω define 0 else. A (cid:22) B, for all A (cid:22) C, with #A = 0 mod 3. Then, one can verify that zA =(cid:26) x A if #A mod 3 = 0 and l1(A′), l1(A′′), l1(A′′′) 6= 0, ζ(α, A)x A(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) XA∈ B(0) ζ(α, A)x A(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) XA∈ C(0) ζ(α, A)zA(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)XA(cid:22)C (cid:13)(cid:13)(cid:13)XA(cid:22)C Let C ∈ MAX(Sα ∩ [N ]<ω) and, by (a) we can find B be in MAX(Sα ∩ [L]<ω) so that Let now A ∈ Sα ∩ [N ]<ω be non-maximal. We will show that w-limn∈N zA∪{n} = 0. By the definition of the vectors zA, we need only treat the case in which (#A + 1)mod 3 = 0, for all n ∈ N with n > max(A). In this case, by (73), we deduce i.e. when zA∪{n} = x ^A∪{n} that if ℓ3n ∈ N , then min supp(zA∪{ℓ3n}) = min supp(x ^A∪{ℓ3n} ) > max(( ^A ∪ {ℓ3n})′′) = ℓ3n−2 where the last equality follows by (b). Hence, limn∈N min supp(zA∪{n}) = ∞. The fact that the FDD of X is shrinking completes the proof. (cid:3) 7. Two metrics on Sα, α < ω1 Since [N]<ω with ≺ is a tree, with a unique root ∅, we could consider on [N]<ω the usual tree distance which we denote by d: For A = {a1, a2, . . . , al} ∈ [N]<ω, or B = {b1, b2, . . . , bm}, we let n = max{j ≥ 0 : ai = bi for i = 1, 2, . . . , j}, and then let d(A, B) = l + m − 2j. But this distance will not lead to to the results we are seeking. Indeed, it was shown in [9, Theorem 1.2] that for any reflexive space X [N]<ω, with the graph metric embeds bi-Lipschitzly into X iff and only if max(Sz(X), Sz(X∗)) > ω. We will need weighted graph metrics on Sα. 32 P. MOTAKIS AND TH. SCHLUMPRECHT (i) The weighted tree distance on Sα. For A, B in Sα let C be the largest element in Sα (with respect to ≺) so that C (cid:22) A and C (cid:22) B (i.e., C is the common initial segment of A and B) and then let d1,α(A, B) = Xa∈A\C z(α,A)(a) + Xb∈B\C z(α,B)(b) = XC≺D(cid:22)A ζ(α, D) + XC≺D(cid:22)B ζ(α, D). (ii) The weighted interlacing distance on Sα can be defined as follows. For A, B ∈ Sα, say A = {a1, a2, . . . , al} and B = {b1, b2, . . . , bm}, with a1 < a2 < . . . < al and b1 < b2 < . . . < bm, we put a0 = b0 = 0 and al+1 = bm+1 = ∞, and define d∞,α(A, B) = max z(α,A)(a) + max z(α,B)(b). i=1,...,m+1 Xa∈A,bi−1<a<bi i=1,...,l+1 Xb∈B,ai−1<b<ai Remark. In order to explain d∞,α let us take some sets A = {a1, a2, . . . , am} and B = {b1, b2, . . . , bm} in Sα and and fix some i ∈ {0, 1, 2, . . . , m}. Now we measure how large the part of B is which lies between ai and ai+1 (as before a0 = 0 and am+1 = ∞) by putting mi(B) := Xj,bj∈(ai,ai+1) ζ(α, {b1, b2, . . . , bj}). Then we define mj(A), for j = 1, 2, . . . , n similarly, and put d∞,α(A, B) = max 1≤i≤m mi(B) + max 1≤j≤n mj(A). We note that if C is maximal so that C (cid:22) A and C (cid:22) B, and if A \ C < B \ C, then d1,α(A, B) = d∞,α(A, B). Proposition 7.1. The metric space (Sα, d1,α,) is stable, i.e., for any sequences (An) and (Bn) in Sα and any ultrafilter U on N it follows that lim m∈U lim n∈U d1,α(Am, Bn) = lim n∈U lim m∈U d1,α(Am, Bn), while (Sα, d∞,α) is not stable. Proof. Since Sα is compact in the product topology we can pass to subsequences and assume that we can write An as An = A ∪ An, with max(A) < min( An) ≤ max( An) < min( An+1), and Bn as Bn = A ∪ Bn, with max(B) < min( Bn) ≤ max Bn < min Bn+1, for n ∈ N. We can also assume the sequences (rA n ) =(cid:16) XA≺D(cid:22) An ζ(α, D)(cid:17) and (rB n ) =(cid:16) XB≺D(cid:22) Bn ζ(α, D)(cid:17) converge to some numbers rA and rB respectively. Let C ∈ Sα be the maximal element for which C (cid:22) A and C (cid:22) B. For m ∈ N large enough and n > m, large enough (depending on m) it follows that Also for n ∈ N large enough and m > n, large enough (depending on n) it follows also that d1,α(Am, Bn) = XC≺D(cid:22)A d1,α(Am, Bn) = XC≺D(cid:22)A ζ(α, D) + XC≺D(cid:22)B ζ(α, D) + XC≺D(cid:22)B ζ(α, D) + rA m + rB n . ζ(α, D) + rA m + rB n . A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 33 Thus we have lim m∈U lim n∈U d1,α(Am, Bn) = XC≺D(cid:22)A = lim n∈U lim m∈U ζ(α, D) + XC≺D(cid:22)B d1,α(Am, Bn). ζ(α, D) + rA + rB To prove our second claim we let (An) be a sequence of nonempty elements in Sα with d(∅, An) = PD(cid:22)An ζ(α, D) < 1/3, and An−1 < min An, for n ∈ N. Then we let B ∈ Sα with d(∅, B) = PD(cid:22)B ζ(α, D) ∈ (1/3, 1/2] , and then choose Bn ∈ Sα so that max B < min Bn ≤ max Bn < min Bn−1, and so that Bn = B ∪ B′ n is maximal in Sα, for n ∈ N. Then it follows and lim m∈U lim n∈U d∞,α(Am, Bn) = 1 − d(∅, B) ≤ 2/3 lim n∈U lim m∈U d∞,α(Am, Bn) ≥ 1. We can now conclude one direction of Theorem A from James' characterization of reflexive spaces. Proposition 7.2. If X is a non reflexive Banach space then for any 0 < c < 1/4 and every α > 0 there is a map Φα : Sα → X, so that (80) cd∞,α(A, B) ≤ kΦ(A) − Φ(B)k ≤ d1,α(A, B) for all A, B ∈ Sα. (cid:3) Remark. Our argument will show that if X is non reflexive there is a sequence (xn) ⊂ BX so that for all α < ω1 the map Φα : Sα → X, A 7→ XD(cid:22)A ζ(α, D)xmax(D), satisfies (80). Proof. Let θ be any number in (0, 1). Then by [16] there is a normalized basic sequence in X whose basic constant is at most 2 Θ satisfying (81) aj for all (aj) ∈ c00, aj ≥ 0, for all j ∈ N. Thus its bimonotonicity constant is at most 4 Θ , which means that for m ≤ n the projection (cid:13)(cid:13)(cid:13) ∞Xj=1 ajxj(cid:13)(cid:13)(cid:13) ≥ θ ∞Xj=1 has norm at most 4 Θ . We define P[m,n] : span(xj) → span(xj), Φ : Sα → X, A 7→ XD(cid:22)A ajxj 7→ ∞Xj=1 ajxj, nXj=m ζ(α, D)xmax(D). For A, B ∈ Sα we let C be the maximal element in Sα for which C (cid:22) A and C (cid:22) B. Then kΦ(A) − Φ(B)k =(cid:13)(cid:13)(cid:13) XC≺D(cid:22)A ζ(α, D)xmax D − XC≺D≺B ζ(α, D)xmax(D)(cid:13)(cid:13)(cid:13) 34 P. MOTAKIS AND TH. SCHLUMPRECHT ≤ XC≺D(cid:22)A ζ(α, D) + XC≺D(cid:22)B ζ(α, D) = d1,α(A, B). On the other hand, if we write A = {a1, a2, . . . , al} and put a0 = 0, and al+1 = ∞, it follows for all i = 1, 2, . . . , l + 1, that kΦ(A) − Φ(B)k ≥ Θ 4(cid:13)(cid:13)P(ai−1,ai)(cid:0)Φ(A) − Φ(B)(cid:1)(cid:13)(cid:13) ≥ Θ2 4 Xai−1<b<ai z(α,B)(b). Similarly, if we write B = {b1, b2, . . . , bm} and put b0 = 0, and bm+1 = ∞, it follows for all j = 1, 2, . . . , m + 1 that kΦ(A) − Φ(B)k ≥ Θ2 4 Xbj−1<a<bj z(α,B)(a). Thus for any i = 1, 2, . . . , l and any j = 1, 2, . . . , m kΦ(A) − Φ(B)k ≥ Θ2 8 " Xai−1<b<ai z(α,B)(b) + Xbj−1<a<bj z(α,B)(a)#, which implies our claim. (cid:3) We finish this section with an observation which we will need later. Lemma 7.3. Let ξ and γ be countable ordinal numbers with γ < β = ωξ. Let B1 < · · · < Bj be in MAX(Sβγ), so that ¯B = {min(Bj) : 1 ≤ j ≤ d} is a non-maximal Sβ set with l1( ¯B) > 0 (l1(A) for A ∈ [N]<ω has been defined before Lemma 3.7) and set D = ∪d j=1Bj ∈ Sβ(γ+1). Then for every A, B in Sβγ with D < A and D < B we have (82) (83) d1,βγ(A, B) = d∞,βγ(A, B) = 1 ζ(β, ¯B) 1 ζ(β, ¯B) d1,β(γ+1)(D ∪ A, D ∪ B) and d∞,β(γ+1)(D ∪ A, D ∪ B). Proof. We will only prove (82), as the proof of (83) uses the same argument. Let C be the maximal element in Sβγ so that C (cid:22) A and C (cid:22) B. Note that C = D ∪ C is the largest element of Sβ(γ+1) so that C (cid:22) D ∪ A and C (cid:22) D ∪ B. Define B1 = ¯B ∪ {min(A)} and B2 = ¯B ∪ {min(B)} and observe that, since l1( ¯B) > 0, ζ(β, B1) = ζ(β, B2) = ζ(β, ¯B). Using (34) in Proposition 3.4 we conclude the following: (84) Xa∈(D∪A)\ C and similarly we obtain z(β(γ+1),D∪A)(a) = ζ(β, B1) Xa∈A\C Xa∈(D∪B)\ C z(β(γ+1),D∪B)(a) = ζ(β, ¯B) Xa∈B\C (85) z(βγ,B)(a). z(βγ,A)(a) = ζ(β, ¯B) Xa∈A\C z(βγ,A)(a) Applying (84) and (85) to the definition of the d1,α metrics, the result easily follows. (cid:3) A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 35 8. The Szlenk index and embeddings of (Sα, d1,α) into X In this section we show the following two results, Theorem 8.1 and 8.3, which establish a proof of Theorem B. Theorem 8.1. Let X be a separable Banach space and α a countable ordinal. Assume that Sz(X) > ωα then (Sα, d1,α) can be bi-Lipschitzly embedded into X and X∗. Before proving Theorem 8.1 we want to first cover the case that ℓ1 embeds into X. Example 8.2. For each α < ω1 we want to define a bi-Lipschitz embedding of (Sα, d1,α) into a Banach space X and and its dual X∗ under the assumption that ℓ1 embeds into X. We first choose for every A ∈ [N]<ω a spread A of A as in Example 5.6. Then we define for α < ω1 Φ : Sα → ℓ1, Since for A, B ∈ Sα it follows that kΦ(A) − Φ(B)k =(cid:13)(cid:13)(cid:13) XC≺D(cid:22)A = XC≺D(cid:22)A ζ(α, D)emax( D). A 7→ XD(cid:22)A ζ(α, D)emax( D) − XC≺D(cid:22)B ζ(α, D) + XC≺D(cid:22)B ζ(α, D)emax( D)(cid:13)(cid:13)(cid:13) ζ(α, D) = d1,α(A, B), where C ∈ Sα is the maximal element for which C (cid:22) A and C (cid:22) B, it follows that Φ is an isometric embedding of (Sα, d1,α) into ℓ1. Thus, if ℓ1 embeds into X then (Sα, d1,α) bi-Lipschitzly embeds into X. Secondly ℓ∞ is a quotient of X∗ in that case, and since ℓ1 embeds into ℓ∞, it follows easily that ℓ1 embeds into X∗, and, thus, that (Sα, d1,α) also bi-Lipschitzly embeds into X∗. Proof of Theorem 8.1. Because of Example 8.2 we can assume that ℓ1 does not embed into X. Thus, we can apply Theorem 5.3, (i) ⇐⇒ (iii) and obtain ε > 0, a tree (z∗ A : A ∈ Sα) ⊂ BX ∗ and a w-null tree (zA : A ∈ Sα) ⊂ BX, so that (86) (87) Then we define B(zA)(cid:12)(cid:12) ≤ (cid:12)(cid:12)z∗ z∗ B(zA) > ε, for all A, B ∈ Sα \ {∅}, with A (cid:22) B, ε 2 for all A, B ∈ Sα \ {∅}, with A 6(cid:22) B, Φ : SA → X, A 7→ XD(cid:22)A ζ(α, D)zD. If A, B ∈ Sα and C ∈ Sα is the maximal element of Sα for which C (cid:22) A and C (cid:22) B, we note that ζ(α, D)zD(cid:13)(cid:13)(cid:13) ζ(α, D) = d1,α(A, B). kΦ(A) − Φ(B)k =(cid:13)(cid:13)(cid:13) XC≺D(cid:22)A ≤ XC≺D(cid:22)A kΦ(A) − Φ(B)k =(cid:13)(cid:13)(cid:13) XC≺D(cid:22)A ζ(α, D)zD − XC≺D(cid:22)B ζ(α, D) + XC≺D(cid:22)B ζ(α, D)zD − XC≺D(cid:22)B ζ(α, D)zD(cid:13)(cid:13)(cid:13) Moreover we obtain 36 P. MOTAKIS AND TH. SCHLUMPRECHT ≥ z∗ A(cid:16) XC≺D(cid:22)A ≥ ε XC≺D(cid:22)A kΦ(A) − Φ(B)k ≥ ε XC≺D(cid:22)B ζ(α, D)zD(cid:17) ζ(α, D), ζ(α, D) − ζ(α, D). ε ζ(α, D)zD − XC≺D(cid:22)B 2 XC≺D(cid:22)B 2 XC≺D(cid:22)A ζ(α, D)! = ζ(α, D) − ε ε 4 d1,α(A, B). ε 4 XC≺D(cid:22)A ζ(α, D) + XC≺D(cid:22)B Similarly we can show that and thus kΦ(A) − Φ(B)k ≥ In order to define a Lipschitz embedding from (Sα, d1,α) into X∗, we also can assume that ℓ1 does not embed into X. Because otherwise ℓ∞ is a quotient of X∗, and thus since ell1 embeds into ℓ∞, it is easy to see that ℓ1 embeds into X∗, and again we let Ψ : Sα → X∗, A 7→ XD(cid:22)A ζ(α, D)z∗ D. If C = A we arrive trivially to the same inequality, Similarly we obtain that kΨ(A) − Ψ(B)k ≥ ε XC≺D(cid:22)B ζ(α, D) − ζ(α, D). This yields kΨ(A) − Ψ(B)k ≥ ε 4 XC≺D(cid:22)A ζ(α, D) + XC≺D(cid:22)B which finishes the proof of our claim. ε 2 XC≺D(cid:22)A ζ(α, D)! = ε 4 d1,α(A, B) (cid:3) Theorem 8.3. Assume that X is a Banach space having a separable dual X∗, with Sz(X∗) > ωα then (Sα, d1,α) can be bi-Lipschitzly embedded into X. Proof. Applying Proposition 5.5 we obtain ε > 0 and trees (z∗ (zA : A ∈ Sα) ⊂ BX , so that (z∗ (88) z∗ A(zB) > ε, for all A, B ∈ Sα \ {∅}, with A (cid:22) B, A : A ∈ Sα) is w∗-null and A : A ∈ Sα) ⊂ B∗ X and As in the case of Φ it is easy to see that Ψ is a Lipschitz function with constant not exceeding the value 1. Again if A, B ∈ Sα let C ∈ Sα be the maximal element of Sα for which C (cid:22) A and C (cid:22) B. In the case that C ≺ A, we let C + ∈ Sα be the minimal element for which C ≺ C + (cid:22) A. We note that C + 6(cid:22) D for any D ∈ Sα with C ≺ D (cid:22) B, and it follows therefore that kΨ(A) − Ψ(B)k =(cid:13)(cid:13)(cid:13) XC≺D(cid:22)A ≥ XC≺D(cid:22)A ≥ ε XC≺D(cid:22)A ζ(α, D)z∗ ζ(α, D)z∗ D − XC≺D(cid:22)B D − XC≺D(cid:22)B 2 XC≺D(cid:22)B ε ζ(α, D)z∗ ζ(α, D)z∗ ζ(α, D) − ζ(α, D). D(cid:13)(cid:13)(cid:13) D!(zC+) A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 37 (89) We define (cid:12)(cid:12)z∗ A(zB)(cid:12)(cid:12) ≤ ε 2 , for all A, B ∈ Sα \ {∅}, with A 6(cid:22) B, Φ : Sα → X, A 7→ XD(cid:22)A ζ(α, D)zA. Again it is easy to see that Φ is a Lipschitz function with constant not exceeding the value 1. Moreover if if A, B ∈ Sα let C ∈ Sα be the maximal element of Sα for which C (cid:22) A and C (cid:22) B, In the case that C ≺ A, we let C + ∈ Sα be the minimal element for which C ≺ C + (cid:22) A. We note that C + 6(cid:22) D for any D ∈ Sα with C ≺ D (cid:22) B, and it follows therefore that If C = A we trivially deduce that Similarly we prove that ζ(α, D)zD(cid:13)(cid:13)(cid:13) ζ(α, D)zD! ζ(α, D)η(D). ζ(α, D)η(D). ζ(α, D)η(D). ≥ z∗ kΦ(A) − Φ(B)k =(cid:13)(cid:13)(cid:13) XC≺D(cid:22)A C+ XC≺D(cid:22)A ≥ ε XC≺D(cid:22)A kΦ(A) − Φ(B)k ≥ ε XC≺D(cid:22)A kΦ(A) − Φ(B)k ≥ ε XC≺D(cid:22)B ζ(α, D)zD − XC≺D(cid:22)B ζ(α, D)zD − XC≺D(cid:22)B ε ζ(α, D) − 2 XC≺D(cid:22)B ζ(α, D)− XC≺D(cid:22)B ζ(α, D) − XC≺D(cid:22)A (ε − η(A))ζ(α, D) + XC≺D(cid:22)B ζ(α, D)(cid:17) = ζ(α, D) + XC≺D(cid:22)B We obtain therefore that kΦ(A) − Φ(B)k ≥ ≥ 1 2(cid:16) XC≺D(cid:22)A 4(cid:16) XC≺D(cid:22)A ε (ε − η(A))ζ(α, D)(cid:17) ε 4 d1,α(A, B) which proves our claim. (cid:3) 9. Refinement argument Before providing a proof of Theorem C, and, thus, the still missing implication of Theorem A, we will introduce in this and the next section some more notation and make some preliminary observations. The will consider maps Φ : Sα → X, satisfying weaker conditions compared to the ones required by Theorem A and C. On the one hand it will make an argument using transfinite induction possible, on the other hand it is sufficient to arrive to the wanted conclusions. Definition 9.1. Let α < ω1. For r ∈ (0, 1] we define S (r) α =nA ∈ Sα : XD(cid:22)A ζ(α, D) ≤ ro. 38 P. MOTAKIS AND TH. SCHLUMPRECHT It is not hard to see that S (r) restrictions. We also put α is a closed subset of Sα, and hence compact, and closed under M(r) α = MAX(S (r) α ) = {A ∈ S (r) α : A is maximal in S (r) α with resp. to "≺"} and for A ∈ Sα, M(r) α (A) =(cid:8)B ∈ Sα(A) : A ∪ B ∈ M(r) α (cid:9). (90) Definition 9.2. Let X be a Banach space, α be a countable ordinal number, L be an infinite subset of N, and A0 be a set in Sα that is either empty or a singleton. A map Φ : Sα(A0) ∩ [L]<ω → X is called a semi-embedding of Sα ∩ [L]<ω into X starting at A0, if there is a number c > 0 so that for all A ∈ Sα(A0) ∩ [L]<ω, with l1(A0 ∪ A) > 0, for all r ∈ (0, 1], and for all B1, B2 in M(r)(A0 ∪ A) ∩ [L]<ω with B1 < B2: (cid:13)(cid:13)(cid:13)Φ(A) − Φ(B)(cid:13)(cid:13)(cid:13) ≤ d1,α(A0 ∪ A, A0 ∪ B) for all A, B ∈ Sα(A0) ∩ [L]<ω, and (cid:13)(cid:13)(cid:13)Φ(A ∪ B1)−Φ(A ∪ B2)(cid:13)(cid:13)(cid:13) ≥ cd1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2). (l1(A) for A ∈ [N]<ω was introduced in Definition 3.6). We call the supremum of all numbers c > 0 so that (91) holds for all A ∈ Sα(A0) ∩ [L]<ω α (A0 ∪ A) ∩ [L]<ω, with B1 < B2, the semi-embedding constant of Φ and (91) and B1, B2 ∈ M(r) denote it by c(Φ). Remark. If Φ : Sα → X is for some 0 < c < C a c-lower d∞,α and C-upper d1,α embedding, we can, after rescaling Φ if necessary, assume that C = 1, and from the definition of d1,α and d∞,α we can easily see that for every A0, that is either empty or a singleton, and L in [N]ω, the restrictions ΦSα(A0)∩[L]<ω X is a semi-embedding. Assume Φ : Sα(A0) ∩ [L]<ω → X, is a semi-embedding of Sα ∩ [L]<ω into X starting at A0. For A ∈ Sα(A0), with A 6= ∅, we put A′ = A \ {max(A)} and define xA0∪A = 1 ζ(α, A0 ∪ A)(cid:0)Φ(A) − Φ(A′)(cid:1) . If A0 = ∅ put x∅ = Φ(∅), whereas if A0 is a singleton, define x∅ = 0 and xA0 = (1/ζ(α, A0))Φ(∅). Note that {x∅} ∪ {xA0∪A : A ∈ Sα(A0) ∩ [L]<ω} ⊂ BX. Recall that ζ(α, ∅) = 0 and hence, for A ∈ Sα(A0) ∩ [L]<ω, we have Φ(A) = x∅ + X∅(cid:22)D(cid:22)A0∪A ζ(α, D)xD. We say in that case that the family {x∅} ∪ {xA0∪A : A ∈ Sα(A0) ∩ [L]<ω} generates Φ. In that case the map Φ0 : Sα(A0) ∩ [L]<ω → X with Φ0 = Φ − x∅, i.e. for A ∈ Sα(A0) ∩ [L]<ω (92) Φ0(A) = X∅(cid:22)D(cid:22)A0∪A ζ(α, D)xD, is also a semi-embedding of Sα ∩ [L]<ω into X starting at A0, with c(Φ0) = c(Φ). A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 39 Lemma 9.3. Let γ, ξ < ω1, with γ < β = ωξ, and let B1 < · · · < Bd be in MAX(Sβγ), so that ¯B = {min(Bj) : 1 ≤ j ≤ d} is a non-maximal Sβ set with l1( ¯B) > 0. Set D = ∪d j=1Bj, let r ∈ (0, 1] and let also A ∈ M(r) ζ(β, ¯B)r, we have that A ∈ M(r0) βγ with D < A. Then, if r0 = PC(cid:22)D ζ(β(γ + 1), C) + β(γ+1)(D). Proof. From Proposition 3.4 and Lemma 3.7 (iii) we obtain that for C (cid:22) A we have ζ(β(γ + 1), D ∪ C) = ζ(β, ¯B ∪ {min(A)})ζ(βγ, C) = ζ(β, ¯B)ζ(βγ, C), β(γ+1). If we assume that D ∪ A is not in M(r0) which implies that D ∪ A ∈ S (r0) B ∈ S (r0) Define B0 = B \ D. Evidently, A ≺ B0 and B′ β(γ+1), there is β(γ+1) with D ∪ A ≺ B. Possibly after trimming B, we may assume that B′ = D ∪ A. 0 = A. We claim that B0 ∈ Sβγ. If we assume that this is not the case, A is a maximal Sβγ set. This yields that PC(cid:22)A ζ(βγ, C) = 1 and hence r = 1 and r0 = PC(cid:22)D ζ(β(γ + 1)) + ζ(β, ¯B) = PC(cid:22)D∪A ζ(β(γ + 1), C), i.e. D ∪ A ∈ M(r0) thus, using Proposition 3.4 and the definition of r0, B0 ∈ S (r) A ∈ M(r) β(γ+1) which we assumed to be false. Thus, we conclude that B0 ∈ Sβγ and βγ , which is a contradiction, as (cid:3) βγ and A ≺ B0. Lemma 9.4. Let α < ω1, N ∈ [N]<ω, A0 be a subset of N that is either empty or a singleton, and c ∈ (0, 1]. Let Ψ : Sα(A0) ∩ [N ]<ω → X be a semi-embedding of Sα ∩ [N ]<ω into X starting at A0, generated by a family of vectors {z∅} ∪ {zA∪A0 : A ∈ Sα(A0) ∩ [N ]<ω}, so that c(Ψ) > c. Let ε < c(Ψ) − c and {z∅} ∪ {zA∪A0 : A ∈ Sα(A0) ∩ [N ]<ω} ⊂ BX , with kzA0∪A − zA0∪Ak < ε, for all A ∈ Sα(A0) ∩ [N ]<ω with A0 ∪ A 6= ∅. Then the map Ψ : Sα(A0) ∩ [N ]<ω → X defined by Ψ(A) = XD(cid:22)A0∪A ζ(α, D)zD, for A ∈ Sα ∩ [N ]<ω, is a semi-embedding of Sα ∩ [N ]<ω into X starting at A0 with c( Ψ) > c. Proof. For any r ∈ (0, 1], any A ∈ Sα(A0) ∩ [L]<ω and B1, B2 ∈ M(r) with B1 < B2 we obtain (cid:13)(cid:13)(cid:13) Ψ(A ∪ B1) − Ψ(A ∪ B2)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) XA0∪A≺D(cid:22)A0∪A∪B1 ≥(cid:13)(cid:13)(cid:13) XA0∪A≺D(cid:22)A0∪A∪B1 − ε XA0∪A≺D(cid:22)A0∪A∪B1 ζ(α, D)zD − ζ(α, D)zD − ζ(α, D) + ≥(c(Ψ) − ε)d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2), α (A0 ∪ A) ∩ [N ]<ω, XA0∪A≺D(cid:22)A0∪A∪B2 XA0∪A≺D(cid:22)A0∪A∪B2 XA0∪A≺D(cid:22)A0∪A∪B2 ζ(α, D)zD(cid:13)(cid:13)(cid:13) ζ(α, D)zD(cid:13)(cid:13)(cid:13) ζ(α, D) which implies (91). (90) follows from the fact that zA0∪A ∈ BX for all A ∈ Sα(A0) ∩ [N ]<ω with A0 ∪ A 6= ∅. (cid:3) For the rest of the section we will assume that X has a bimonotone FDD (En). For finite or cofinite sets A ⊂ N, we denote the canonical projections from X onto span(Ej : j ∈ A) 40 P. MOTAKIS AND TH. SCHLUMPRECHT by PA, i.e., PA : X → X, ∞Xj=1 xj 7→Xj∈A ∞Xj=1 xj, for x = xj ∈ X, with xj ∈ Ej, for j ∈ N, and we write Pj instead of P{j}, for j ∈ N. We denote the linear span of the Ej by c00(Ej : j ∈ N), i.e., c00(Ej : j ∈ N) =n ∞Xj=1 xj : xj ∈ Fj, for j ∈ N and #{j : xj 6= 0} < ∞o. Definition 9.5. Let α be a countable ordinal number, M ∈ [N]ω, and A0 be a subset of N that is either empty or a singleton. A semi-embedding Φ : Sα(A0) ∩ [M ]<ω → X of Sα ∩ [M ]<ω into X starting at A0, is said to be c-refined, for some c ≤ c(Φ), if the following conditions are satisfied: (i) the family {x∅} ∪ {xA0∪A : A ∈ Sα(A0) ∩ [M ]<ω} generating Φ is contained in BX ∩ c00(Ej : j ∈ N), (ii) for all A ∈ Sα(A0) ∩ [M ]<ω with A0 ∪ A 6= ∅ we have max(A0 ∪ A) ≤ max supp(xA0∪A) < min{m ∈ M : m > max(A0 ∪ A)}. (iii) for all r ∈ (0, 1], A ∈ Sα(A0) ∩ [M ]<ω, with l1(A0 ∪ A) > 0, and B1, B2 in M(r) α (A0 ∪ A) ∩ [M ]<ω, with B1 < B2, we have (cid:13)(cid:13)P(max supp(xA0 ∪A),∞) (Φ(A ∪ B1) − Φ(A ∪ B2))(cid:13)(cid:13) ≥ cd1,α (A0 ∪ A ∪ B1, A0 ∪ A ∪ B2) , (iv) for all r ∈ (0, 1], A ∈ Sα(A0) ∩ [M ]<ω and B in M(r) α (A0 ∪ A) ∩ [M ]<ω we have (cid:13)(cid:13)P(max supp(xA0 ∪A),∞)(cid:0)Φ(A ∪ B)(cid:1)(cid:13)(cid:13) ≥ c 2 XA0∪A≺C(cid:22)A0∪A∪B ζ(α, C). Remark 9.6. Let ξ < ω1, γ ≤ β = ωωξ be a limit ordinal, 0 < c ≤ 1, and M ∈ [N]ω. If a0 ∈ N, we note that Sβγ({a0}) ∩ [M ]<ω = Sβη(γ,a0)({a0}) ∩ [M ]<ω, and ζ(βγ, {a0} ∪ D) = ζ(βη(γ, a0), {a0}∪D) for D ∈ Sβγ({a0})∩[M ]<ω, where(cid:0)η(γ, n)(cid:1)n∈N is the sequence provided by Proposition 2.6. It follows that a semi-embedding of Sβγ({a0}) ∩ [M ]<ω into X starting at {a0} that is c-refined, is a c-refined semi-embedding of Sβη(γ,n)({a0}) ∩ [M ]<ω into X. Secondly, if Φ : Sβγ ∩ [M ]<ω → X is a semi-embedding of Sβγ ∩ [M ]<ω into X start- ing at ∅ that is c-refined, then for every a0 ∈ M and N = M ∩ [a0, ∞) the map Ψ = ΦSβη(γ,a0)({a0})∩[N ]<ω , is a semi-embedding of Sβη(γ,a0)∩[N ]<ω into X starting at {a0}, that is c-refined. Furthermore, Ψ is generated by the family {xA0∪A : A ∈ Sβη(γ,a0)({a0}) ∩ [N ]<ω}, where {xA : A ∈ Sβγ ∩ [M ]<ω} is the family generating Φ. Lemma 9.7. Let ξ, γ < ω1, with γ << β = ωωξ , M ∈ [N]ω, and A0 be a subset of N that is either empty or a singleton. Let also Φ : Sβ(γ+1)(A0) ∩ [M ]<ω → X be a semi-embedding of Sβ(γ+1) ∩ [M ]∞ into X starting at A0, that is c-refined. The family generating Φ is denoted by {x∅} ∪ {xA0∪A : A ∈ Sα(A0) ∩ [M ]<ω}. Extend the set A0 to a set A0 ∪ A1, A0 < A1, which can be written as A0 ∪ A1 = ∪k j=1Bj ∈ Sβ(γ+1) ∩ [A0 ∪ M ]<ω, where B1 < · · · < Bk are in MAX(Sβγ) and ¯B = {min(Bj) : 1 ≤ j ≤ k} is a non-maximal Sβ set with l1( ¯B) > 0. Then, for N = M ∩ (max(A0 ∪ A1), ∞) and n0 = max supp(xA0∪A1), the map Ψ : Sβγ ∩ [N ]<ω → X, A 7→ 1 ζ(β, ¯B) P(n0,∞) (Φ(A1 ∪ A)) A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 41 is a semi-embedding of Sβγ ∩ [N ]ω into X starting at ∅, that is c-refined. Moreover, Ψ is generated by the family {zA : A ∈ Sβγ ∩ [N ]<ω}, where (93) z∅ = 0 and zA = P(n0,∞)(xA0∪A1∪A) for A ∈ Sβγ ∩ [N ]<ω \ {∅}. Proof. By Lemma 7.3 we easily obtain that for A, B ∈ Sβγ ∩ [N ]<ω: 1 ζ(β, ¯B) d1,β(γ+1)(A0 ∪ A1 ∪ A, A0 ∪ A1 ∪ B) = d1,(βγ)(A, B), (94) (cid:13)(cid:13)(cid:13)Ψ(A) − Ψ(B)(cid:13)(cid:13)(cid:13) ≤ we set r0 =PC(cid:22)A0∪A1 βγ (A) ∩ [N ]<ω (i.e. A ∪ B1, A ∪ B2 ∈ M(r) (90) from Definition 9.2 is satisfied for Ψ. We will show that (90) from Definition i.e. 9.2 is satisfied for Ψ as well. Let r ∈ (0, 1], A be in Sβγ, with l1(A) > 0, and B1 < B2 in M(r) βγ ). Note that we have l1(A0 ∪ A1 ∪ A) > 0. If ζ(β(γ +1), C)+ζ(β, ¯B)r, by Lemma 9.3 we deduce that A ∪ B1 and A ∪ B2 are in M(r0) β(γ+1)(A0 ∪ (A1 ∪ A)) ∩ [N ]<ω. Definition 9.5 (ii) implies n0 ≤ max supp(xA0∪A1∪A) and, thus, by Definition 9.5 (iii) for Φ we deduce β(γ+1)(A0 ∪ A1) ∩ [N ]<ω, i.e., B1, B2 ∈ M(r0) (95) (cid:13)(cid:13)(cid:13)Ψ(A ∪ B1) − Ψ(A ∪ B2)(cid:13)(cid:13)(cid:13) = 1 ζ(β, ¯B)(cid:13)(cid:13)(cid:13)P(n0,∞) (Φ(A1 ∪ A ∪ B1) − Φ(A1 ∪ A ∪ B2))(cid:13)(cid:13)(cid:13) ζ(β, ¯B)(cid:13)(cid:13)(cid:13)P(max supp(xA0 ∪A1∪A),∞) (Φ(A1 ∪ A ∪ B1) − Φ(A1 ∪ A ∪ B2))(cid:13)(cid:13)(cid:13) d1,β(γ+1)(A0 ∪ A1 ∪ A ∪ B1, A0 ∪ A1 ∪ A ∪ B2) 1 c ≥ ≥ ζ(β, ¯B) = cd1,βγ(A ∪ B1, A ∪ B2), where the last equality follows from Lemma 7.3. In particular, (94) and (95) yield that Ψ : Sβγ ∩ [N ]<ω → X is a semi-embedding of Sβγ ∩ [N ]∞ into X starting at ∅ with c(Ψ) ≥ c. It remains to show that Ψ satisfies Definition 9.5 (i) to (iv). Observe that Definition 9.5 (ii) implies that for C (cid:22) A0 ∪ A1, we have P(n0,∞)(xC ) = 0. We combine this with (34) of Proposition 3.4 to obtain that for A ∈ Sβγ ∩ [N ]<ω: (96) Ψ(A) = = = P(n0,∞)(Φ(A1 ∪ A)) ζ(β, ¯B) 1 ζ(β, ¯B) XC(cid:22)A0∪A1∪A 1 1 XA0∪A1≺C(cid:22)A0∪A1∪A ζ(βγ, C)P(n0,∞)(xA0∪A1∪C). ζ(β, ¯B) = XC(cid:22)A ζ(β(γ + 1), C)P(n0,∞)(xC ) ζ(β(γ + 1), C)P(n0,∞)(xC) For A ∈ Sβγ ∩ [N ]<ω define zA = P(n0,∞)(xA0∪A1∪A) and z∅ = 0. It then easily fol- lows by (96) that Ψ is generated by the family {zA : A ∈ Sα ∩ [N ]<ω}. Moreover, as max supp(zA) = max supp(xA0∪A1∪A), it is straightforward to check that Definition 9.5 (i) and (ii) are satisfied for Ψ. Whereas, observing that for all A ∈ Sβγ ∩ [N ]<ω, with A 6= ∅ (which is the case when l1(A) > 0), we have max supp(zA) = max supp(xA0∪A1∪A) ≥ n0, an 42 P. MOTAKIS AND TH. SCHLUMPRECHT argument similar to the one used to obtain (95) also yields that Ψ satisfies Definition 9.5 (iii) and (iv). (cid:3) The main result of this section is the following Refinement Argument. Lemma 9.8. Let α < ω1, M ∈ [N]ω, and A0 be a subset of N that is either empty or a singleton. Let also Φ : Sα(A0) ∩ [M ]<ω → X be a semi-embedding of Sα ∩ [M ]<ω into X starting at A0. Then, for every c < c(Φ), there exists N ∈ [M ]ω and a semi-embedding Φ : Sα(A0) ∩ [N ]<ω → X of Sα ∩ [N ]<ω into X starting at A0, that is c-refined. Proof. Put c = (c(Φ) + c)/2. Let {x∅} ∪ {xA0∪A : A ∈ Sα(A0) ∩ [M ]<ω} be the vectors generating Φ and choose η > 0 with η < c(Φ) − c. After shifting we can assume without loss of generality that x∅ = 0. Set x∅ = 0 and choose for each A ∈ Sα(A0) ∩ [M ]<ω, with A0 ∪ A 6= ∅, a vector xA0∪A ∈ BX ∩ c00(Ej : j ∈ N) so that: (a) kxA0∪A − xA0∪Ak < η/2, and max(A0 ∪ A) ≤ max supp(xA0∪A). Moreover, recursively choose m1 < · · · < mk < · · · in M so that for all k we have mk+1 > max(cid:8) max supp(xA0∪A) : A ∈ Sα(A0) ∩ [{ m1, . . . , mk}](cid:9). Define M = { mk : k ∈ N} and Φ : Sα(A0) ∩ [ M ]<ω → X so that for all A ∈ Sα(A0) ∩ [ M ]<ω we have Φ(A) = XC(cid:22)A0∪A ζ(α, C)xC. By Lemma 9.4, Φ is a semi-embedding from Sα ∩ [ M ]<ω into X starting at A0 for which c( Φ) > c > c, and for which the conditions (i) and (ii) of Definition 9.5 are satisfied. The goal is to find N ∈ [ M ]ω so that, restricting Φ to Sα(A0) ∩ [N ]<ω, Definition 9.5 (iii) and (iv) are satisfied as well. Put M0 = M . Recursively, we will choose for every k ∈ N an infinite set Mk ⊂ Mk−1 so that for each k ∈ N the following conditions are met: (b) min(Mk−1) < min(Mk), and putting mj = min(Mj) for j = 1, . . . , k − 1, then (c) for every A ∈ Sα(A0) ∩ [{m1, . . . , mk−1}] with l1(A0 ∪ A) > 0, r ∈ (0, 1], and B1, B2 ∈ M(r) α (A0 ∪ A) ∩ [{m1, . . . , mk−1} ∪ Mk]<ω with B1 < B2, we have (cid:13)(cid:13)P(max supp(xA0 ∪A),∞)(cid:0) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) ≥ cd1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2), (d) for every A ∈ Sα(A0) ∩ [{m1, . . . , mk−1}] with l1(A0 ∪ A) > 0, r ∈ (0, 1], B ∈ M(r) α (A0 ∪ A) ∩ [{m1, . . . , mk−1} ∪ Mk]<ω we have (cid:13)(cid:13)P(max supp(xA0 ∪A),∞)(cid:0) Φ(A ∪ B)(cid:1)(cid:13)(cid:13) ≥ (if k = 1, [{m1, . . . , mk−1}] = {∅}). c 2 XA0∪A≺C(cid:22)A0∪A∪B ζ(α, C) If we assume that such a sequence (Mk)k has been chosen, it is straightforward to check that N = {mk : k ∈ N} is the desired set. In the case k = 1, for A ∈ Sα(A0)∩ {∅} we have A = ∅. Hence, if A0 = ∅ then for all A ∈ Sα(A0) ∩ {∅} we have A0 ∪ A = ∅, i.e. l1(A0 ∪ A) = 0 and hence, (c) and (d) are always satisfied. Choosing M1 satisfying (b) completes the first inductive step. If, on the other hand, A0 is a singleton then for all A ∈ Sα(A0) ∩ {∅} we have A0 ∪ A = A0, i.e. l1(A0 ∪ A) > 0. This means that condition (c) and (d) are reduced to the case in which A = ∅. The choice of M1 is done in the same manner as in the general inductive step and we omit it. Assume that we have chosen, for some k ≥ 1, infinite sets Mk ⊂ Mk−1 ⊂ · · · ⊂ M1 ⊂ M0, so that (b), (c), and (d) are satisfied for all 1 ≤ k′ ≤ k. Observe that the inductive A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 43 assumption implies that it is enough to choose Mk+1 ∈ [Mk]ω satisfying (b), and the conditions (c) and (d) for sets A ∈ Sα(A0) ∩ [{m1, . . . , mk−1}] with l1(A0 ∪ A) > 0 and max(A) = mk = min(Mk) (or A = ∅, in the case k = 1, and A0 is a singleton). We set δ = min {ζ(α, A0 ∪ A) : A ∈ Sα(A0) ∩ [{m1, m2, . . . , mk}] and A0 ∪ A 6= ∅} , ε = δ 30 (c − c), d = max{max supp(xA0∪A) : A ∈ Sα(A0) ∩ [{m1, . . . , mk}]}, and choose a finite ε-net R of the interval (0, 1], with 1 ∈ R, which also has the property that for all A ∈ Sα(A0) ∩ [{m1, m2, . . . , mk}] and all j = 0, 1, 2, . . . , l1(A0 ∪ A), j · ζ(α, A0 ∪ A) + XC(cid:22)A0∪A ζ(α, C) ∈ R. (97) (98) Fix a finite ε 2 -net K of the unit ball of the finite dimensional space span(Ej : 1 ≤ j ≤ d). For every r ∈ R and A ∈ Sα(A0) ∩ [{m1, m2, . . . , mk}] we apply Proposition 2.14 to M(r) α (A0 ∪ A) ∩ [Mk]<ω, and find an infinite subset Mk+1 of Mk so that for all A ∈ Sα(A0) ∩ [{m1, m2, . . . , mk}] and r ∈ R, there exists y(r) A in K so that A − P[1,d](cid:0) Φ(A ∪ B)(cid:1)(cid:13)(cid:13) < (cid:13)(cid:13)y(r) ε 2 , for all B in M(r) α (A0 ∪ A) ∩ [ Mk+1]<ω. In particular, note that for all A ∈ Sα(A0) ∩ [{m1, m2, . . . , mk}] and r ∈ R, for any B1, B2 in M(r) α (A0 ∪ A) ∩ [ Mk+1]<ω we have (cid:13)(cid:13)P[1,d](cid:0) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) < ε. Using the property (iv) of l1(·) in Lemma 3.7, we can pass to an infinite subset cMk+1 of Mk+1, so that (b) is satisfied and moreover: (99) ζ(α, A0 ∪ A ∪ B) < ε, if A ∈ Sα(A0) ∩ [{m1, . . . , mk}], and B ∈ Sα(A0 ∪ A) ∩ [Mk+1]<ω with #B > l1(A0 ∪ A) > 0. We will show that (c) is satisfied. To that end, fix 0 < r ≤ 1, A ∈ Sα(A0)∩[{m1, . . . , mk}] with max(A0 ∪ A) = mk and l1(A0 ∪ A) > 0, and B1, B2 ∈ M(r) with B1 < B2. If both sets B1 and B2 are empty, then (c) trivially holds, as the right- hand side of the inequality has to be zero. Otherwise, Bs 6= ∅, where s = 1 or s = 2. Note that max(A) = mk, i.e. A0 ∪ A 6= ∅ and hence, since l1(A0 ∪ A) > 0, putting C = A0 ∪ A ∪ {min(Bs)}, by the definition of δ we obtain ζ(α, C) = ζ(α, A0 ∪ A) ≥ δ. This easily yields: α (A0 ∪ A) ∩ [cMk+1]<ω d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2) = ζ(α, C) + ζ(α, C) XA0∪A≺C(cid:22)A0∪A∪B1 ≥ δ = 30ε c − c XA0∪A≺C(cid:22)A0∪A∪B2 Hence: (100) ε ≤ c − c 30 d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2) Arguing similarly, we obtain: (101) r ≥ XC(cid:22)A0∪A∪Bs ζ(α, C) ≥ XC(cid:22)A0∪A ζ(α, C) + ζ(α, C) ≥ δ > min(R). 44 P. MOTAKIS AND TH. SCHLUMPRECHT Choose r0 to be the maximal element of R with r0 ≤ r. Since r0 ≤ r, we can find B1 and B2 in M(r0) (102) α (A0 ∩ A) ∩ [cMk+1]<ω so that B1 (cid:22) B1 and B2 (cid:22) B2. We will show that d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B1) < 3ε and d1,α(A0 ∪ A ∪ B2, A0 ∪ A ∪ B2) < 3ε. ζ(α, C). The maximality of B1 in S (r0) If B1 = We shall only show that this is the case for B1, for B2 the proof is identical. B1, then there is nothing to prove and we may therefore assume that B1 ≺ B1. Define ζ(α, C), and C1 = B1 ∪ {min(B1 \ B1)}, r1 = PC(cid:22)A0∪A∪B1 r′ =PC(cid:22)A0∪A∪C1 r′ =PC(cid:22)A0∪A ζ(α, C) + (#C1)ζ(α, A0 ∪ A) is in R. This contradicts the maximality of r0. We conclude that #C1 > l1(A0 ∪ A), which by (99) yields r′ − r1 = ζ(α, A0 ∪ A ∪ C1) < ε. Hence, (104) d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B1) = r1 − r1 = (r1 − r′) + (r′ − r1) < (r1 − r0) + ε < 3ε ζ(α, C), r1 = PC(cid:22)A0∪A∪ B1 In this case however, by (97), we obtain that We first assume that #C1 ≤ l1(A0 ∪ A). (103) r1 ≤ r0 < r′ ≤ r1. α (A0 ∪ A) implies We verify (c) now as follows (cid:13)(cid:13)P(d,∞)(cid:0) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) ≥(cid:13)(cid:13)P(d,∞)(cid:0) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) −(cid:16)(cid:13)(cid:13)P(d,∞)(cid:0) Φ(A ∪ B1) − Φ(A ∪ B1)(cid:1)(cid:13)(cid:13) +(cid:13)(cid:13)P(d,∞)(cid:0) Φ(A ∪ B2) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13)(cid:17) ≥(cid:13)(cid:13)P(d,∞)(cid:0) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) −(cid:16)d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B1) + d1,α(A0 ∪ A ∪ B2, A0 ∪ A ∪ B2)(cid:17) >(cid:13)(cid:13)P(d,∞)(cid:0) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) − 6ε ≥(cid:13)(cid:13) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:13)(cid:13) −(cid:13)(cid:13)P[1,d](cid:0) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) − 6ε ≥(cid:13)(cid:13) Φ(A ∪ B1) − Φ(A ∪ B2)(cid:13)(cid:13) − 7ε ≥c( Φ)d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2) − 7ε > cd1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2). Here the second inequality follows from (90), the third from (102), the fifth from (98), the sixth one from (91), and the last one from (100). We will need to pass to a further subset of cMk+1 to obtain (d). An application of the triangle inequality and (c) (for k + 1) yield that for every A ∈ Sα(A0) ∩ [{m1, m2, . . . mk}] with l1(A0 ∪ A), r ∈ (0, 1] and B1, B2 ∈ M(r) α (A0 ∪ A), B1 < B2, it follows that or (cid:13)(cid:13)P(max supp(xA0 ∪A),∞)(cid:0) Φ(A ∪ B1)(cid:1)(cid:13)(cid:13) ≥ (cid:13)(cid:13)P(max supp(xA0 ∪A),∞)(cid:0) Φ(A ∪ B2)(cid:1)(cid:13)(cid:13) ≥ c 2 c 2 d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2), d1,α(A0 ∪ A ∪ B1, A0 ∪ A ∪ B2). We may pass therefore to a further infinite subset Mk+1 of cMk+1, so that for any A ∈ Sα(A0) ∩ [{m1, m2, . . . , mk}], with l1(A0 ∪ A) > 0 and max(A0 ∪ A) = mk, any r ∈ R, and for any B in M(r) α (A0 ∪ A) ∩ [Mk+1]<ω we have (105) (cid:13)(cid:13)P(d,∞) Φ(A ∪ B)(cid:13)(cid:13) ≥ c 2 XA0∪A≺C(cid:22)A0∪A∪B ζ(α, C). A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 45 We will show that (d) is satisfied (meaning that (105) does not only hold if r ∈ R). Fix an arbitrary r ∈ (0, 1], A ∈ Sα(A0)∩[{m1, . . . , mk}], with max(A0∪A) = mk and l1(A0∪A) > 0, and B ∈ M(r) α (A0 ∪ A) ∩ [Mk+1]<ω. If B is empty, then the right-hand side of the inequality in (d) is zero and hence the conclusion holds. Arguing identically as in (100) we obtain ε ≤ c − c 30 XA0∪A≺C(cid:22)A0∪A∪B ζ(α, C) and arguing as in (101) we obtain r > min(R), so we may choose the maximal element r0 of R with r0 ≤ r, and B the maximal element of M(r0) α (A0 ∪ A) ∩ [Mk+1]<ω with B (cid:22) B. The argument yielding (102), also yields d1,α(A0 ∪ A ∪ B, A0 ∪ A ∪ B) = XA0∪A∪ B≺C(cid:22)A0∪A∪B ζ(α, C) < 3ε. Arguing in a very similar manner as above: (cid:13)(cid:13)P(d,∞) Φ(A ∪ B)(cid:13)(cid:13) ≥(cid:13)(cid:13)P(d,∞) ≥ Φ(A ∪ B)(cid:13)(cid:13) −(cid:13)(cid:13)P(d,∞) ζ(α, C) − 3ε Φ(A ∪ B) − P(d,∞) Φ(A ∪ B)(cid:13)(cid:13) c 2 ≥ ≥ = c 2 c 2 c 2 XA0∪A≺C(cid:22)A0∪A∪ B XA0∪A≺C(cid:22)A0∪A∪B XA0∪A≺C(cid:22)A0∪A∪B XA0∪A≺C(cid:22)A0∪A∪B 3c ζ(α, C) −(cid:18)3 + 2(cid:19) ε ζ(α, C) −(cid:18) c − c 2 (cid:19) XA0∪A≺C(cid:22)A0∪A∪B ζ(α, C). ζ(α, C) (cid:3) 10. Some further observation on the Schreier families In this section β will be a fixed ordinal of the form β = ωωξ , with 1 ≤ ξ < ω1. 10.1. Analysis of a maximal set B in Sβγ . Recall that by Proposition 2.6, for every γ ≤ β there exists a sequence η(γ, n) of ordinal numbers increasing to γ, so that λ(βγ, n) = βη(γ, n) (recall that η(γ, n) may also depend on β). For every γ ≤ β and B ∈ MAX(Sβγ) we define a family of subsets of B, which we shall call the β-analysis of B, and denote by Aβ,γ(B). The definition is done recursively on γ. For γ = 1, set (106a) Aβ,γ(B) = {B}. Let γ < β and assume that Aβ,γ(B) has been defined for all B ∈ MAX(Sβγ). For B ∈ MAX(Sβ(γ+1)) = MAX(Sβ[Sβγ]) there are (uniquely defined) B1 < · · · < Bℓ in MAX(Sβγ) with {min Bj : j = 1, . . . , ℓ} in MAX(Sβ) so that B = ∪ℓ j=1Bj. Set (106b) Aβ,γ+1(B) = {B} ∪(cid:16)∪ℓ j=1Aβ,γ(Bj)(cid:17) . 46 P. MOTAKIS AND TH. SCHLUMPRECHT Let γ ≤ β be a limit ordinal and assume that Aβ,γ′(B) has been defined for all γ′ < γ and B ∈ MAX(Sβγ′). If now B ∈ MAX(Sβ(γ)), then B ∈ MAX(Sβη(γ,min(B))). Set (106c) Aβ,γ(B) = Aβ,η(γ,min(B))(B). Remark 10.1. Let γ ≤ β and B ∈ MAX(Sβγ). The following properties are straightfor- ward consequences of the definition of Aβ,γ(B) and a transfinite induction. (i) The set Aβ,γ(B) is a tree when endowed with "⊃". (ii) For C, D in Aβ,γ(B) that are incomparable with respect to inclusion, we have either C < D, or D < C. (iii) The minimal elements (with respect to inclusion) of Aβ,γ(B) are in Sβ. (iv) If D ∈ Aβ,γ(B) is a non-minimal element, then its direct successors (Dj)ℓ j=1 in Aβ,γ(B) can be enumerated so that D1 < · · · < Dℓ and D = ∪ℓ j=1Dj. 10.2. Components of a set A in Sβγ . We recursively define for all non-empty sets A ∈ Sβγ and γ ≤ β, a natural number s(β, γ, A) and subsets Cpβ,γ(A, 1), . . . , Cpβ,γ(A, s(β, γ, A)) of A. We will call (Cpβ,γ(A, i))s(β,γ,A) the components of A in Sβγ, with respect to Sβ. i=1 If γ = 1, i.e. A is a non-empty set in Sβ, define (107a) s(β, γ, A) = 1 and Cpβ,γ(A, 1) = A. Let γ ≤ β and assume that (Cpβ,γ(A, i))s(β,γ,A) has been defined for all non-empty sets A in Sβγ. If now A is a non-empty set in Sβ(γ+1) = Sβ[Sβγ], then there are non-empty sets A1 < A2 < · · · < Ad in Sβγ so that: i=1 i=1Ai, (i) A = ∪d (ii) {min Ai : i = 1, . . . , d} is in Sβ, and (iii) the sets A1, . . . , Ad−1 are in MAX(Sβγ). Note that the set Ad may or may not be in MAX(Sβγ). It may also be the case that d = 1, which in particular happens when A ∈ Sβγ. Set ¯A = Ad, which is always non-empty, and we define (107b) s(β, γ + 1, A) = s(β, γ, ¯A) + 1, Cpβ,γ+1(A, 1) = ∪i<dAi, and Cpβ,γ+1(A, i) = Cpβ,γ( ¯A, i−1), if 2 ≤ i ≤ s(β, γ +1, A). Note that in the case d = 1, Cpβ,γ+1(A, 1) is the empty set. Let γ ≤ β be a limit ordinal and assume that (Cpβ,γ′(A, i))s(β,γ′,A) has been defined for all γ′ < γ and non-empty sets A in Sβγ′. IfA is a non-empty set in Sβγ, then A ∈ Sβη(γ,min(A)) and we define i=1 (107c) s(β, γ, A) = s(β, η(γ, min(A)), A) and Cpβ,γ(A, i)) = Cpβ,η(γ,min(A))(A, i)) for i = 1, 2, . . . , s(β, γ, A). Remark. Let γ ≤ β and A ∈ Sβγ \ {∅}. The following properties follow easily from the definition of (Cpβ,γ(A, i))s(β,γ,A) Cpβ,γ(A, i). (i) A = ∪s(β,γ,A) (ii) For 1 ≤ i < j ≤ s(β, γ, A) so that both Cpβ,γ(A, i) and Cpβ,γ(A, j) are non-empty, and a transfinite induction on γ. i=1 i=1 we have Cpβ,γ(A, i) < Cpβ,γ(A, j). (iii) Cp(A, s(β, γ(A)) 6= ∅. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 47 Lemma 10.2. Let ξ and γ be countable ordinal numbers with γ ≤ β = ωωξ . Let also B be a set in MAX(Sβγ) and ∅ ≺ A (cid:22) B. If Aβ,γ(B) is the β-analysis of B and (Cpβ,γ(A, i))s(β,γ,A) are the components of A in Sβγ with respect to Sβ, then there exists a maximal chain B = D1(A) ) D2(A) ) · · · ) Ds(β,γ,A)(A) in Aβ,γ(B) satisfying the following: i=1 (i) Cpβ,γ(A, i) (cid:22) Di(A) for 1 ≤ i ≤ s(β, γ, A), and (ii) if 1 ≤ i < s(β, γ, A) then Cpβ,γ(A, i) ( Di(A). Proof. We prove the statement by transfinite induction on 1 ≤ γ ≤ β. Aβ,γ(B) = {B} and (Cpβ,γ(A, i))s(β,γ,A) = {A}, and our claim follows trivially. i=1 If γ = 1, then Let γ < β and assume that the statement holds for all non-empty A ∈ Sβγ and B ∈ MAX(Sβγ) with A (cid:22) B. Let now A be a non-empty set in Sβ(γ+1) and B ∈ MAX(Sβ(γ+1)). If B = ∪ℓ i=1Bj, where B1 < · · · < Bℓ are in MAX(Sβγ) with {min(Bj) : 1 ≤ j ≤ ℓ} ∈ MAX(Sβ), then by (106b) we obtain Aβ,γ+1(B) = {B} ∪ (∪ℓ j=1Aβ,γ(Bj)). Define d = max{1 ≤ j ≤ ℓ : Bj ∩ A 6= ∅}, Ai = Bi for 1 ≤ i < d, and Ad = A ∩ Bd. . Define D1(A) = B, and Di(A) = Di−1( ¯A), for 2 ≤ i ≤ s(β, γ, A). Then, by (107b), letting ¯A = Ad, we obtain s(β, γ + 1, A) = s(β, γ, ¯A) + 1, Cpβ,γ+1(A, 1) = ∪i<dAi, and Cpβ,γ+1(A, i) = Cpβ,γ( ¯A, i − 1), for 2 ≤ i ≤ s(β, γ + 1, A). Apply the inductive assumption to Ad = ¯A (cid:22) Bd, to find a maximal chain Bd = D1( ¯A) ) · · · ) Ds(β,γ, ¯A)( ¯A) in Aβγ(Bd) satisfying (i) and (ii) with respect to (Cpβ,γ( ¯A, i))s(β,γ, ¯A) (108) Clearly, (Di(A))s(β,γ, ¯A) It remains to verify that (i) and (ii) are satisfied, with respect to (Cpβ,γ+1(A, i))s(β,γ+1,A) . Assertions (i) and (ii), in the case i = 1, follow trivially from Cpβ,γ+1(A, 1) = ∪j<dAj = ∪j<dBj ≺ ∪j<dBj (cid:22) B. Assertion (i) and (ii) in the case i 6= 1, follow easily from the inductive assumption and Cpβ,γ+1(A, i) = Cpβ,γ(A′, i − 1), for 2 ≤ i ≤ s(β, γ + 1, A). is a maximal chain in Aβ,γ+1(B). i=1 i=1 i=1 To conclude the proof, the case in which γ ≤ β is a limit ordinal number so that the con- clusion is satisfied for all γ′ < γ, we just observe that the result is an immediate consequence of (106c) and (107c). (cid:3) For the next result recall the definition of the doubly indexed fine Schreier families Fβ,γ introduced in Subsection 2.3. Lemma 10.3. Let γ ≤ β. Then for all A ∈ Sβγ, with Cpβγ(A, i) 6= ∅ for 1 ≤ i ≤ s(β, γ, A), we have (109) (cid:8)min(Cpβ,γ(A, i)) : 1 ≤ i ≤ s(β, γ, A)(cid:9) ∈ MAX(Fβ,γ). If γ = 1 and A ∈ Sβ satisfying the Proof. We prove this statement by induction on γ. assumptions of this Lemma, then A = Cpβ,1(A, 1) 6= ∅ and hence the result easily follows from MAX(Fβ,1) = {{n} : n ∈ N}. Assume that the result holds for some γ < β and let A ∈ Sβ(γ+1), with Cpβγ(A, i) 6= ∅ for 1 ≤ i ≤ s(β, γ + 1, A). By the inductive assumption and (107b) we obtain that B = {min(Cpβ,γ(A, i)) : 2 ≤ i ≤ s(β, γ + 1, A)} ∈ MAX(Fβ,γ). We claim that A = spreading property of Fβ,γ+1 there is C ∈ Fβ,γ+1 with A ≺ C. Then, B ≺ C = C \ (cid:8) min(cid:0)Cpβ,γ+1(A, 1)(cid:1)(cid:9) ∪ B ∈ MAX(Fβ,γ+1). Indeed, if this is not the case, then by the (cid:8) min(cid:0)Cpβ,γ+1(A, 1)(cid:1)(cid:9). It follows by Fβ,γ+1 = Fβ,1⊔<Fβ,γ, that C ∈ Fβ,γ. The maximality of B yields a contradiction. (cid:8)A : A (cid:22) B and A 6∈ Eβ(γ+1)(B)(cid:9) = {A : A (cid:22) B1} ∪ ℓ[m=2n(cid:16) m−1[j=1 Bj(cid:17) ∪ A : A (cid:22) Bm and A 6∈ Eβ,γ(Bm)o . 48 P. MOTAKIS AND TH. SCHLUMPRECHT Assume now that γ ≤ β is a limit ordinal number so that the conclusion holds for all γ < γ. Let A ∈ Sβγ so that Cpβ,γ(A, i) 6= ∅ for 1 ≤ i ≤ s(β, γ, A). Note that A ∈ Sβη(γ,min(A)). and by (107c) and the inductive assumption we have A = {min(Cpβ,γ(A, i)) : 1 ≤ i ≤ s(β, γ, A)} ∈ MAX(Fβ,η(γ,min(A))). By (27) we obtain A ∈ MAX(Fβ,γ). (cid:3) For γ ≤ β and a set B in MAX(Sβγ) we define (110) Eβ,γ(B) = {∅ ≺ A (cid:22) B : Cpβ,γ(A, i) 6= ∅ for 1 ≤ i ≤ s(β, γ, A)}. Lemma 10.4. Let γ < β. If B is in MAX(Sβ(γ+1)) = Sβ[Sβγ] and B = ∪ℓ B1 < · · · < Bℓ are in MAX(Sβγ) and {min(Bj) : 1 ≤ j ≤ ℓ} ∈ MAX(Sβ), then j=1Bj, where Proof. Let ∅ 6= D (cid:22) B. Define m = max{1 ≤ j ≤ ℓ : D ∩ Bj 6= ∅} and A = Bm ∩ D. Note that D = (∪j<mBj) ∪ A, where ∪j<mBj = ∅ if m = 1. By (107b) we obtain, s(β, γ + 1, D) = s(β, γ, A)+ 1, Cpβ,γ+1(D, 1) = ∪j<mBj and Cpβ,γ+1(D, i) = Cpβ,γ(A, i− 1) for 2 ≤ i ≤ s(β, γ + 1, D). Observe that Cpβ,γ+1(D, 1) = ∅ if and only if m = 1, i.e. A = D (cid:22) B1. On the other hand, if Cpβ,γ+1(D, 1) 6= ∅, then for some 2 ≤ i ≤ s(β, γ + 1, D), we have Cpβ,γ+1(D, i) = ∅, if and only if Cpβ,γ(A, i − 1) = ∅. These observations yield our claim. (cid:3) Remark. Under the assumptions of Lemma 10.4, if for 1 ≤ j ≤ ℓ − 1 we define E (j) β,γ+1(B) = {A ∈ Eβ,γ+1(B) : Cpβ,γ(A, 1) = ∪i≤jBi}, then, using a similar argument as the one used in the proof of Lemma 10.4, we obtain E (j) β,γ+1(B) =n(∪j i=1Bi) ∪ C : C ∈ Eβ,γ(Bj+1)o and Eβ(γ+1)(B) = ∪ℓ−1 j=1E (j) β,γ+1. Remark. Using the fact S1 ⊂ Sα that for all countable ordinal numbers α, and that MAX(S1) = {F ⊂ N : min(F ) = #F }, it is easy to verify that for all F ∈ MAX(Sα) we have max(F ) ≥ 2 min(F ) − 1. In particular, if B1 < B2 are both in MAX(Sα), then (111) 2 min(B1) ≤ min(B2). Lemma 10.5. Let γ ≤ β and let B be a set in MAX(Sβγ). Then (112) XA(cid:22)B A6∈Eβ,γ(B) ζ(βγ, A) < 2 min(B) . Proof. We prove the statement by transfinite induction for all 1 ≤ γ ≤ β. If γ = 1, then the complement of Eβ,1(B) only contains the empty set and the result trivially holds. Let γ < β and assume that the statement holds for all B ∈ MAX(Sβγ). Let B ∈ MAX(Sβ(γ+1)). Let B1 < · · · < Bℓ in MAX(Sβγ) so that {min(Bj) : 1 ≤ j ≤ ℓ} ∈ MAX(Sβ) and B = ∪ℓ j=1Bj. For m = 1, . . . , ℓ, define Dm = {min(Bj) : 1 ≤ j ≤ m}. Proposition 3.4 implies the following: if C (cid:22) B, m = max{1 ≤ j ≤ ℓ : C ∩ Bj 6= ∅} and A = C ∩ Bm, then A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 49 ζ(β(γ + 1), C) = ζ(β, Dm)ζ(βγ, A). We combine this fact with Lemma 10.4, (111), and (33) to obtain the following: ζ(βγ, A) + ℓXj=2 ζ(β, Dj) X(cid:26) A(cid:22)Bj A6∈Eβγ (Bj )(cid:27) ζ(βγ, A) ζ(β, Dj) 2 min(Bj) ζ(β, Dj) 2 min(Bj) ζ(β, Dj) 1 min(B1) ≤ 2 min(B1) = 2 min(B) . XA(cid:22)B A6∈Eβ,γ(B) < ζ(β, D1) + ζ(βγ, A) = ζ(β, D1) XA(cid:22)B1 ℓXj=2 ℓXj=2 ℓXj=2 min(B1) min(B1) ≤ ≤ + + 1 1 i=1 i=1 k=1 k=1 and (Dk(A(2)))s(β,γ,A(2)) are both maximal chains of Aβ,γ(B) (cid:3) If γ ≤ β is a limit ordinal number so that the conclusion is satisfied for all γ′ < γ, we just observe that the result is an immediate consequence of (107c) and (30). (cid:3) Lemma 10.6. Let γ ≤ β and B ∈ MAX(Sβγ). If A(1), A(2) ∈ Eβ,γ(B), (Dk(A(1)))s(β,γ,A(1)) and (Dk(A(2)))s(β,γ,A(2)) are the maximal chains of Aβ,γ(B) provided by Lemma 10.2, and 1 ≤ i ≤ min{s(β, γ, A(1)), s(β, γ, A(2))} is such that Di(A(1)) = Di(A(2)), then we have Dj(A(1)) = Dj(A(2)) for all 1 ≤ j < i. Proof. As (Dk(A(1)))s(β,γ,A) with Di(A(1)) = Di(A(2)), the result follows from Remark 10.1 (i). If A(1), A(2) ∈ Eβ,γ(B), and 1 ≤ i ≤ Lemma 10.7. Let γ ≤ β and B ∈ MAX(Sβγ). min{s(β, γ, A(1)), s(β, γ, A(2))}, then Di(A(1)) = Di(A(2)) if and only if min(Cpβ,γ(A(1), i)) = min(Cpβ,γ(A(2), i)), where (Dk(A(1)))s(β,γ,A) are the maximal chains of Aβ,γ(B) provided by Lemma 10.2. Proof. Assume that Di(A(1)) = Di(A(2)). Lemma 10.2 (i) and the assumptions Cp(A(1), i) 6= ∅, and Cp(A(2), i) 6= ∅ yield min(Cpβ,γ(A(1), i)) = min(Cpβ,γ(A(2), i)). For the converse let A(1), A(2) ∈ Eβ,γ(B) with min(Cpβ,γ(A(1), i)) = min(Cpβ,γ(A(2), i)). Since all elements of Aβ,γ(B) either compare with respect to "⊂", or are disjoint, Lemma 10.2 (i) and min(Cpβ,γ(A(1), i)) = min(Cpβ,γ(A(2), i)) imply that either Di(A(1)) ⊂ Di(A(2)), or Di(A(2)) ⊂ Di(A(1)). We assume the first, and towards a contradiction assume that Di(A(1)) ( Di(A(2)). The maximality of (Dj(A(2)))s(β,γ,A(2)) in Aβ,γ(B) implies that there is 1 ≤ j < i, so that Dj(A(2)) = Di(A(1)). As A(2) β,γ(j) 6= ∅, we obtain by Lemma 10.2 (i) min(Cpβ,γ(A(1), i)) = min(Di(A(1))) = min(Dj(A(2))) = min(Cpβ,γ(A(2), j)) < min(Cpβ,γ(A(2), i)) which is a con- tradiction. (cid:3) and (Dk(A(2)))s(β,γ,A(2)) k=1 j=1 i=1 10.3. Special families of convex combinations. Definition 10.8. Let γ ≤ β and B ∈ MAX(Sβγ). A family of non-negative numbers {r(A, k) : A ∈ Eβ,γ(B), 1 ≤ k ≤ s(β, γ, A)}, is called a (β, γ)-special family of convex combinations for B if the following are satisfied: 50 P. MOTAKIS AND TH. SCHLUMPRECHT (i) Ps(β,γ,A) k=1 r(A, k) = 1 for all A ∈ Eβ,γ(B) and (ii) If A(1), A(2) are both in Eβ,γ(B), (Dk(A(1)))s(β,γ,A) are the maximal chains in Aβ,γ(B) provided by Lemma 10.2, and for some k we have Dk(A(1)) = Dk(A(2)), then r(A(1), k) = r(A(2), k). and (Dk(A(2)))s(β,γ,A(2)) k=1 i=1 Remark. Let γ ≤ β and B ∈ MAX(Sβγ) and let {r(A, k) : A ∈ Eβ,γ(B), k = 1, 2, . . . , s(β, γ, A)} be a family of (β, γ)-special convex combinations for B. For A ∈ Eβ,γ(B), let (Dk(A)s(β,γ,A) i=1 be the maximal chain in Aβ,γ(B) provided by Lemma 10.2, and let (A(i))s(β,γ,A) be the components of A in Sβγ. i=1 By construction, D1(A) = B for all A ∈ Eβ,γ(B), and thus r(A, 1) does not depend on A. Secondly D2(A), only depends on A(1), thus r(A(1), 2) = r(A(2), 2) if Cpβ,γ(A(1), 1) = Cpβ,γ(A(2), 1), for any A(1), A(2) ∈ Eβ,γ(B). We can continue and inductively we observe that for all k ≤ min(s(β, γ, A(1)), s(β, γ, A(2))) if Cpβ,γ(A(1), i) = Cpβ,γ(A(2), i), for all i = 1, 2 . . . , k − 1, then r(A(1), k) = r(A(2), k). Let γ ≤ β and B ∈ MAX(Sβγ). If γ is a limit ordinal number, then B ∈ MAX(Sβη(γ,min(B))) and any (β, γ)-special family of convex combinations {r(A, k) : A ∈ Eβ,γ(B), 1 ≤ k ≤ s(β, γ, A)} is also a (β, η(γ, min(B)))-special family of convex combinations, as Eβ,γ(B) = Eβ,η(γ,min(B))(B) and for A ∈ Eβ,η(γ,min(B))(B) we have s(β, η(γ, min(B)), A) = s(β, γ, A). Lemma 10.9. We are given γ < β, B ∈ MAX(Sβ(γ+1)), and a (β, γ + 1)-special family of convex combinations {r(A, k) : A ∈ Eβ,γ+1(B), 1 ≤ k ≤ s(β, γ + 1, A)}. Assume that for some D ∈ Eβ,γ+1(B) (and hence for all of them) we have r(D, 1) < 1. Let B = ∪d j=1Bj, where B1 < · · · < Bd are the immediate predecessors of B in Aβγ(B). For every 1 ≤ j < d consider the family {r(j)(C, k) : C ∈ Eβ,γ(Bj+1)}, with r(j)(C, k) = 1 1 − r(D, 1) r(cid:16)(cid:16)∪j i=1Bi(cid:17) ∪ C, k + 1(cid:17) for k = 1, . . . , s(β, γ, C) = s(β, γ, (∪j a (β, γ)-special family of convex combinations. i=1Bi) ∪ C) − 1. Then {r(j)(C, k) : C ∈ Eβ,γ(Bj+1)} is Proof. By (107b), if C ∈ Eβ,γ(Bj+1) then A = (∪j i=1Bi) ∪ C ∈ Eβ,γ+1(B) and s(β, γ + 1, A) = s(β, γ, C) + 1, which implies that Definition 10.8 (i) is satisfied. To see that (ii) holds, let C (1), C (2) be in Eβ,γ(Bj+1) so that for some k we have Dk(C (1)) = Dk(C (2)). Then by Lemma 10.7 we have min(Cpβ,γ(C (1), k) = min(Cpβ,γ(C (2), k). Setting A(1) = (∪j i=1Bt) ∪ C (2), by (107b) we obtain Cpβ,γ+1(A(1), k + 1) = Cpβ,γ(C (1), (k) and Cpβ,γ(A(2), k + 1) = Cpβ,γ(C (2), k), i.e. min(Cpβ,γ(A(1), k + 1)) = min(Cpβ,γ(A(2), k + 1)). By Lemma 10.7 we obtain Dk+1(A(1)) = Dk+1(A(2)) and therefore r(A(1), k + 1) = r(A(2), k + 1), which yields that r(j)(C (1), k) = r(j)(C (2), k). (cid:3) i=1Bt) ∪ C (1) and A(2) = (∪j 11. Conclusion of the proof of Theorems A and C Again, we fix ξ < ω1 and put β = ωωξ . We secondly assume that X is a Banach space X with a bimonotone FDD (Fj ). By the main result in [33] every reflexive Banach space X embeds into a reflexive Banach space Z with basis, so that Sz(Z) = Sz(X) and Sz(Z ∗) = Sz(X∗). The coordinate projections on finitely or cofinitely many coordinates are denoted by PA (see Section 9 after Definition 9.5). A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 51 Definition 11.1. Let γ ≤ β, M ∈ [N ]ω and A0 be a subset of N that is either empty or a singleton. Let also Φ : Sβγ(A0) ∩ [M ]<ω → X be a semi-embedding of Sβγ ∩ [M ]<ω into X, starting after A0, that is c-refined, for some 0 < c ≤ 1. Let {x∅} ∪ {xA0∪A : A ∈ Sβγ(A0) ∩ [M ]<ω} be the family generating Φ (recall that notation from the Remark after Definition 9.2). Let E ∈ MAX(Sβγ(A0) ∩ [M ]<ω). For A (cid:22) E recall the definition of s(β, γ, A) and of . Recall also from (110) Eβ,γ(A0 ∪ E) =(cid:8)∅ ≺ A (cid:22) A0 ∪ E : Cpβ,γ(A, i) 6= i=1 (Cpβ,γ(A, i))s(β,γ,A) ∅, for i = 1, 2 . . . , s(β, γ, A)(cid:9). For each A ∈ Eβ,γ(A0 ∪ E) we will write xA as a sum of a block sequence (113) xA = s(β,γ,A)Xk=1 x(k) Φ,A, Φ,A = PI(A,k)(xA), for k = 1, 2, . . . , s(β, γ, A), where I(A, 1) = [1, max supp(xCpβ,γ(A,1))] i=1Cpβ,γ(A,i))] for 1 < k < s(β, γ, A). i=1 Cpβ,γ(A,i)), max supp(x∪k with x(k) and I(A, k) = (max supp(x∪k−1 We call ((x(k) Φ,A)s(β,γ,A) k=1 )A∈Eβ,γ (A0∪E) the block step decomposition of E with respect to Φ. Remark. Let M ∈ [N]ω, and γ ≤ β be a limit ordinal number and let η(γ, n) be the sequence provided by Proposition 2.6. Assume that A0 a singleton or the empty set and that Φ : Sβγ(A0) ∩ [M ]<ω → X is a semi-embedding of Sβγ ∩ [M ]<ω into X starting at A0 that is c-refined. If A0 is a singleton, say A0 = {a0}, let Ψ : Sβη(γ,a0)(A0)∩[M ]<ω → X, with Ψ(A) = Φ(A), be the semi-embedding of Sβη(γ,a0) ∩ [M ]<ω into X starting at A0, that is c-refined, given by Remark 9.6. Then, for every E ∈ MAX(Sβη(γ,a0)(A0) ∩ [N ]<ω), we have (114) ((x(k) Ψ,A)s(β,η(γ,a0),A) k=1 )A∈Eβ,η(γ,a0)(A0∪E) = ((x(k) Φ,A)s(β,γ,A) k=1 )A∈Eβ,γ(A0∪E). If A0 = ∅, let a0 ∈ M , set A0 = {a0}, N = M ∩ [a0, ∞) and Ψ = ΦSβη(γ,a0)(A0)∩[N ]<ω , which is, by Remark 9.6, a semi-embedding of Sβη(γ,a0) ∩ [N ]<ω into X starting at A0, that is c-refined. Then, again for every E ∈ MAX(Sβη(γ,a0)(A0) ∩ [N ]<ω), (115) ((x(k) )A∈Eβ,η(γ,a0)(A0∪E) = ((x(k) is the step block decomposition of E with respect to Ψ. Ψ,A)s(β,η(γ,a0),A) k=1 Φ,A)s(β,γ,A) k=1 )A∈Eβ,γ (A0∪E) Before formulating and proving the missing parts our Main Theorems A and B (see upcoming Theorem 11.6) we present the argument which is the main inductive step. Let γ < β. For B ∈ MAX(Sβ(γ+1)) = Sβ[Sβγ] (Proposition 2.6), we let B1 < B2 < j=1 Bj. We also define ¯B = {min(Bj) : j = 1, 2, . . . , d} ∈ MAX(Sβ) and for i = 1, . . . , d, ¯Bj = {min(Bj) : j = 1, 2, . . . , i}. i=1 Bi ∪ C, for some j = 1, 2, . . . , d, and some . . . < Bd the (unique) elements of MAX(Sβγ) for which B =Sd If ∅ ≺ A (cid:22) B we can write A as A =Sj−1 ∅ ≺ C (cid:22) Bj, and thus, by Proposition 3.4 ζ(β(γ + 1), A) = ζ(β, ¯Bj)ζ(βγ, C). We define Eβ,γ+1(B) =n(∪j i=1Bj) ∪ C ∈ Eβ,γ+1(B) : 1 ≤ j < d, l1( ¯Bj) > 0 and C (cid:22) Bj+1o . Let M ∈ [N ]ω, A0 be a subset of N that is either empty or a singleton, and Φ : Sβγ(A0) ∩ [M ]<ω → X be a semi-embedding of Sβγ ∩ [M ]<ω into X, starting after A0, that is c-refined, (116) 52 P. MOTAKIS AND TH. SCHLUMPRECHT for some 0 < c ≤ 1. Let also E ∈ MAX(Sβ(γ+1))(A0) ∩ [M ]<ω, and put B = A0 ∪ E. For 1 ≤ j < d, with l1( ¯Bj) > 0, put M (j) = M ∩(max(Bj), ∞), and Φ(j) E : Sβγ ∩[M (j)]<ω → X with (117) Φ(j) E (C) = 1 ζ(β, ¯Bj+1) Φ (((∪1≤i≤jBi) \ A0) ∪ C) . Recall that by Lemma 9.7 Φ(j) that is c refined, recall that for 1 ≤ j < d E is a semi embedding of Sβγ ∩ [M (j)]<ω into X starting at ∅, E (j) β,γ+1(A0 ∪ E) = {A ∈ Eβ,γ+1(A0 ∪ E) : Cpβ,γ(A, 1) = ∪i≤jBi}, and, moreover, if l1( ¯Bj) > 0 define (118) y(j) Φ,E = XA∈E (j) β,γ+1(A0∪E) ζ(β(γ + 1), A) ζ(β, ¯Bj) s(β,γ+1,A)Xk=2 x(k) Φ,A. Remark. 1) By Lemma 9.7, each Φ(j) E is a semi-embedding of Sβγ ∩[M (j)]<ω into X, starting Φ,E)d j=1 is a sequence in X which satisfies the condi- j=1 in Theorem 6.1 with α = β and thus, assuming that 3) By the definition of the components of a set A, we conclude (note that for j = d at ∅, that is c-refined. 2) We note for later use that (y(j) tions of the sequence (xj)d Sz(X) ≤ ωβ, also its conclusion. we have l1( ¯Bd) = 0) (119) (120) Eβ,γ+1(A0 ∪ E) = [1≤j<d l1( ¯Bj )>0 E (j) β,γ+1(A0 ∪ E), which yields XA∈ Eβ,γ+1(A0∪E) ζ(β(γ + 1), A) x(k) Φ,A = s(β,γ+1,A)Xk=1 Φ,E + XA∈ Eβ,γ+1(A0∪E) ζ(β, ¯Bj)y(j) X1≤j<d l1( ¯Bj )>0 ζ(β(γ + 1), A)x(1) Φ,A. Lemma 11.2. Let γ < β, M ∈ [N ]ω, A0 be either empty or a singleton in N, and let Φ : Sβ(γ+1)(A0)∩ [M ]<ω → X be a semi-embedding of Sβ(γ+1) ∩ [M ]<ω into X, starting after A0, that is c-refined, for some 0 < c ≤ 1. Then, for every E ∈ MAX(Sβ(γ+1)(A0) ∩ [M ]<ω) x(k) (121) ζ(β(γ + 1), A) s(β,γ+1,A)Xk=1 Φ,A(cid:13)(cid:13)(cid:13) < (cid:13)(cid:13)(cid:13)Φ(E) − XA∈ Eβ,γ+1(A0∪E) Proof. Recall that for A ∈ Eβ,γ+1(A0 ∪ E), we have xA =Ps(β,γ+1,A) PA(cid:22)A0∪E ζ(β(γ + 1), A)xA. Hence, (cid:13)(cid:13)(cid:13)Φ(E) − XA∈ Eβ,γ+1(A0∪E) Φ,A(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) XA(cid:22)A0∪E: s(β,γ,A)Xk=1 ζ(β(γ + 1), A) x(k) k=1 A6∈ Eβ,γ+1(A0∪E) 3 min(A0 ∪ E) . x(k) Φ,A and that Φ(E) = ζ(β(γ + 1), A)xA(cid:13)(cid:13)(cid:13). A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 53 We calculate: (122) ζ(β(γ + 1), A) A6∈ Eβ,γ+1(A0∪E) XA(cid:22)A0∪E = XA(cid:22)A0∪E = XA(cid:22)A0∪E A6∈Eβ,γ+1(A0∪E) A6∈Eβ,γ+1(A0∪E) l1( ¯Bj )=0 XA∈E (j) ζ(β(γ + 1), A) + X1≤j<d ζ(β(γ + 1), A)+ X1≤j<d ζ(β, ¯Bj+1) XA∈E (j) β,γ+1(A0∪E) β,γ+1(A0∪E) ζ(β(γ + 1), A) ζ(βγ, A \ (∪i≤jBj)) < 2 min(A0 ∪ E) + 1 min( ¯B) = l1( ¯Bj )=0 3 min(A0 ∪ E) , where the last inequality follows from Lemmas 3.7 (v) and 10.5. (cid:3) Lemma 11.3. Let γ, M ∈ [N ]ω, A0, Φ, and c be as in the statement of Lemma 11.2. If E ∈ MAX(Sβ(γ+1)(A0)∩[M ]<ω) and B1 < · · · < Bd are in MAX(Sβγ), with A0∪E = ∪d j=1Bj, and ¯B = {min Bj : 1 ≤ j ≤ d} ∈ MAX(Sβ), and if (y(j) j=1 is defined as in (118), then for j = 1, . . . , d − 1 with l1( ¯Bj) > 0, where ¯Bj = {min Bi : 1 ≤ i ≤ j}, we have ky(j) Φ,Ek ≤ 1 and ran(y(j) Φ,E) ⊂ (max( ¯Bj), max( ¯Bj+2)). Put max( ¯Bd+1) = ∞. Then E,Φ)d (123) ζ(β(γ + 1), A) XA∈ Eβ(γ+1)(A0∪E) s(β,γ+1,A)Xk=1 x(k) Φ,A = XA∈ Eβ,γ+1(A0∪E) ζ(β(γ + 1), A)x(1) ζ(β, ¯Bj+1)y(j) Φ,E. Φ,A + X1≤j<d l1( ¯Bj )>0 Proof. Observe that (123) immediately follows from (120) and the fact that for 1 ≤ j < d with l1( ¯Bj) > 0, we have ζ(β, ¯Bj) = ζ(β, ¯Bj+1). For 1 ≤ j < d with l1( ¯Bj) > 0 and A ∈ E (j) β,γ+1(A0 ∪ E) note that ∪j i=1Bi ≺ A (cid:22) ∪j+1 i=1 Bi and (124) uA = s(β,γ+1,A)Xk=2 x(k) Φ,A = P(max supp(xCpβ,γ (A,1)),max supp(xA)](xA) = P(max supp(x ),max supp(xA)]xA ∪ j i=1 Bj i.e. kuAk ≤ 1 and by Definition 9.5 (ii) we obtain max( ¯Bj) ≤ max(∪j max supp(uA) ≤ max supp(xA) < min{m ∈ M : m > max(A)} ≤ max( ¯Bj+2) i=1Bi) ≤ max supp(x∪j ) < min supp(uA) and i=1Bi which yields (125) ran(uA) ⊂ (max( ¯Bj), max( ¯Bj+2)). 54 P. MOTAKIS AND TH. SCHLUMPRECHT Furthermore, we have ζ(β(γ+1), A) = ζ(β, ¯Bj+1)ζ(βγ, A\(∪j ζ(β, ¯Bj) (by l1( ¯Bj) > 0), we obtain i=1Bi)), and since ζ(β, ¯Bj+1) = (126) We deduce ζ(β(γ + 1), A) ζ(β, ¯Bj) = ζ(βγ, A \ (∪j i=1Bi)). y(j) Φ,E = XC∈Eβ,γ(Bj+1) ζ(βγ, C)u(∪j i=1Bi)∪C. The above in combination with (125), (126), and the fact that kuAk ≤ 1 yield that ky(j) 1 and ran(y(j) Φ,E) ⊂ (max( ¯Bj), max( ¯Bj+2)). Φ,Ek ≤ (cid:3) Lemma 11.4. Let γ, M ∈ [N ]ω, A0, Φ, and c be as in the statement of Lemma 11.2. Let E ∈ MAX(Sβ(γ+1)(A0)∩[M ]<ω), (Bj)d j=1 be as in the statement of Lemma 11.3 and let Φ(j) E , j = 1, . . . , d − 1 and M (j) ∈ [M ]ω be defined as in (117). For j = 1, . . . , d − 1 with l1( ¯Bj) > ∅ denote by j=1, ¯B, ( ¯Bj)d (cid:18)z(k) E ,C(cid:19)s(β,γ′,C) k=1 Φ(j) !C∈Eβ,γ′ (Bj+1) the block step decomposition of Bj+1 with respect to Φ(j) s(β, γ + 1, (∪j i=1Bi) ∪ C) = s(β, γ, C) + 1 and E . Then for C ∈ Eβ,γ(Bj+1) we have for k = 1, . . . , s(β, γ, C). z(k) Φ(j) E ,C = x(k+1) Φ,(∪j i=1Bi)∪C , Proof. Fix C ∈ Eβ,γ(Bj+1). By (107b), i=1Bi) ∪ C, then s(β, γ + 1, A) = s(β, γ, C) + 1 and Cpβ,γ+1(A, i + 1) = Cpβ,γ(A, i) for i = 1, . . . , s(β, γ, C), and Cpβ,γ+1(A, 1) = ∪j i=1Bi. Fix 1 ≤ k ≤ s(β, γ, C). Let {zC : C ∈ Sβγ ∩ [M (j)]<ω} be the family generating Φ(j) E and n0 = max supp(x∪j ), then by Definition 11.1 and Lemma 9.7 we have i=1Bi if we set A = (∪j Φ,A = PI(A,k+1)(xA) and z(k) x(k+1) Φ(j) with (if k = 1 replace max supp(z∪k−1 = PI(C,k)(zC ) = PI(C,k)(cid:0)P(n0,∞)(xSi≤j Bi∪C)(cid:1), E ,C i=1 Cpβ,γ(C,i) by n0) I(C, k) = (max supp(z∪k−1 = (max supp(P(n0,∞)x(∪j = (max supp(x(∪j i=1Bi)∪(∪k−1 i=1 Cpβ,γ(C,i)), max supp(z∪k i=1Cpβ,γ(C,i))] i=1 Cpβ,γ(C,i))),max supp(P(n0,∞)x(∪j i=1Bi)∪(∪k−1 i=1 Cpβ,γ(C,i))), max supp(x(∪j i=1Bi)∪(∪k i=1Cpβ,γ(C,i)))] i=1Cpβ,γ(C,i)))] = I(A, k+1), i=1Bi)∪(∪k where we used n0 ≤ max supp(x∪j i=1Bi ), which follows from Definition 9.5 (ii). Hence, z(k) Φ(j) E ,C = PI(A,k+1)P(n0,∞)(xA) = PI(A,k+1)(xA) = x(k+1) Φ,A . (cid:3) A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 55 Proposition 11.5. Assume that Sz(X) ≤ ωβ. Then, for every 1 ≥ c > 0 and M ∈ [N]ω, there exists N ∈ [M ]ω with the following property: for every γ ≤ β, L ∈ [N ]ω, and A0 ⊂ N that is either empty or a singleton, every semi-embedding Φ : Sβγ(A0) ∩ [L]<ω → X from Sβγ ∩ [L]<ω into X starting at A0, that is c-refined, every E ∈ MAX(Sβγ(A0) ∩ [L]<ω) and every (β, γ)-special family of convex combinations {r(A, k) : A ∈ Eβ,γ(A0 ∪ E), 1 ≤ k ≤ s(β, γ, A)} we have (127) XA∈Eβ,γ (A0∪E) ζ(βγ, A) s(β,γ,A)Xk=1 r(A, k)kx(k) Φ,Ak ≥ c 3 , where (x(k) Φ,A)A∈Eβ,γ(A0,E)is the block step decomposition of E for Φ (Definition 11.1). Proof. Fix M ∈ [N]ω and 1 ≥ c > 0. Choose ε > 0 so that c/2 − ε > c/3 and then apply Theorem 6.1 to find N ∈ [M ]∞ so that (69) is satisfied for that ε and β, and, moreover, (128) (cid:16) c 2 − ε − 4 min(N )(cid:17) Ym∈N(cid:16)1 − 1 m(cid:17) > c 3 . We claim that this is the desired set. We shall prove by transfinite induction on γ the following statement: if γ ≤ β, L ∈ [N ]ω, A0 ⊂ L that is either empty or a singleton, and Φ : Sβγ(A0) ∩ [L]<ω → X is a semi-embedding of Sβγ ∩ [L]<ω into X starting at A0, that is c-refined, then for any E ∈ MAX(Sβγ(A0) ∩ [L]<ω) and any (β, γ)-special family of convex combinations {r(A, k) : A ∈ Eβ,γ(A0 ∪ E), 1 ≤ k ≤ s(β, γ, A)} we have (129) XA∈Eβ,γ(A0∪E) ζ(βγ, A) s(β,γ,A)Xk=1 r(A, k)kx(k) Φ,Ak ≥(cid:18) c 2 − ε − 4 min(L)(cid:19) ∞Ym∈L(cid:18)1 − 1 m(cid:19) . In conjunction with (128), this will yield the desired result. Let γ = 1, L ⊂ N , A0 be a subset of L that is either empty or a singleton, and Φ : Sβ(A0) ∩ [L]<ω → X be a semi-embedding of Sβ ∩ [L]<ω into X starting at A0 that is c-refined. Let E ∈ MAX(Sβ(A0) ∩ [N ]<ω). By (107a), for each A ∈ Eβ,1(A0 ∪ E) we obtain that the block step decomposition of xA Φ,A) = (xA), and hence if {r(A, k) : A ∈ Eβ,1(A0 ∪ E), 1 ≤ k ≤ s(β, 1, A)} is a is just (x(1) (β, 1)-special family of convex combinations then r(A, 1) = 1. Hence, XA∈Eβ,1(A0∪E) ζ(β, A) r(A, k)kx(k) s(β,1,A)Xk=1 ≥(cid:13)(cid:13)(cid:13) XA(cid:22)A0∪E Φ,Ak = XA∈Eβ,1(A0∪E) ζ(β, A)xA(cid:13)(cid:13)(cid:13) − XA(cid:22)A0∪E 2 A /∈Eβ,1(A0∪E) (by (112)) ≥ kΦ(E)k − ζ(β, A)kxAk ζ(β, A) min(A0 ∪ E) 2 ζ(β, A0) − 3 min(L) (by (33)). c 2 ≥ > c 2 c 2 − − min(A0 ∪ B) (by 9.5 (iv)) To verify the induction step, let first γ < β be an ordinal number for which the conclusion holds. Let L ⊂ N , A0 be a subset of L that is either empty or a singleton, and Φ : 56 P. MOTAKIS AND TH. SCHLUMPRECHT Sβ(γ+1)(A0) ∩ [L]<ω → X be a semi-embedding of Sβ(γ+1) ∩ [L]<ω into X starting at A0 that is c-refined. Let E ∈ MAX(Sβ(γ+1)(A0) ∩ [L]<ω) and {r(A, k) : A ∈ Eβ,γ+1(A0 ∪ E), 1 ≤ k ≤ s(β, γ + 1, A)} be a (β, γ + 1)-special family of convex combinations. Let A0 ∪ E = ∪d j=1Bj, where B1 < · · · < Bd are in MAX(Sβγ ∩ [L]<ω) and ¯B = min{Bj : 1 ≤ j ≤ d} is in MAX(Sβ). By Lemma 11.3 and the choice of the set N , we obtain (130) (cid:13)(cid:13)(cid:13) X1≤j<d l1( ¯Bj )>0 ζ(β, ¯Bj+1)y(j) Φ,E(cid:13)(cid:13)(cid:13) < ε. Combining first Lemmas 11.2 and 11.3, then applying Definition 9.5 (iv), (33), and finally using (130) we deduce (131) ζ(β(γ +1), A)kx(1) Φ,Ak XA∈ Eβ,γ+1(A0∪E) ζ(β(γ + 1), A)x(1) ≥(cid:13)(cid:13)(cid:13) XA∈ Eβ,γ+1(A0∪E) ≥ kΦ(E)k −(cid:13)(cid:13)(cid:13) X1≤j<d l1( ¯Bj )>0 Φ,A(cid:13)(cid:13)(cid:13) Φ,E(cid:13)(cid:13)(cid:13) − 3 ζ(β, ¯Bj+1)y(j) 3 min(A0 ∪ E) ≥ − ζ(β(γ +1), A0) − ε − c 2 c 2 min(A0 ∪ E) ≥ c 2 − ε − 4 min(L) . We distinguish two cases. If r(A, 1) = 1, for all A ∈ Eβ,γ+1(A0 ∪ E), then XA∈Eβ,γ+1(A0∪E) ζ(β(γ + 1), A) r(A, k)kx(k) Φ,Ak s(β,γ+1,A)Xk=1 = XA∈Eβ,γ+1(A0∪E) ζ(β(γ + 1), A)kx(1) ζ(β(γ + 1), A)kx(1) Φ,Ak. Φ,Ak ≥ XA∈ Eβ,γ+1(A0∪E) By (131), the result follows in that case. Otherwise we have r1 = r(A, 1) < 1, for all A ∈ Eβ,γ+1(A0 ∪ E). For 1 ≤ j < d, with E : Sβγ ∩ [L(j)]<ω → X as E is a semi-embedding of Sβγ ∩ [L(j)]<ω into X, starting l1( ¯Bj) > 0, define L(j) = L ∩ (max(A0 ∪ (∪1≤i≤jBi)), ∞), and Φ(j) in (117). By Lemma 9.7, each Φ(j) at ∅, that is c-refined. By Lemma 10.9, the family {r(j)(C, k) : C ∈ Eβ,γ(Bj+1)}, with r(j)(C, k) = 1 1 − r(A(1), 1) r(cid:16)(cid:16)∪j i=1Bi(cid:17) ∪ C, k + 1(cid:17) for k = 1, . . . , s(β, γ, C) = s(β, γ, (∪j i=1Bi) ∪ C) − 1 and some A(1) ∈ Eβ,γ+1(A0 ∪ E), is a (β, γ)-special family of convex combinations. Hence, by the inductive assumption applied to the map Φ(j) E , and Lemma 11.4 we deduce (132) XA∈E (j) β,γ+1(A0∪E) ζ(β(γ + 1), A) s(β,γ+1,A)Xk=2 r(A, k)kx(k) Φ,Ak A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 57 =ζ(β, ¯Bj)(cid:0)1−r(A(1), 1)(cid:1) XC∈Eβ,γ (Bj+1) ≥ζ(β, ¯Bj)(cid:0)1−r(A(1), 1)(cid:1)(cid:16) c − ε − 2 ζ(βγ, C) r(j)(C, k)kz(k) Φ(j) E ,C k s(β,γ,C)Xk=1 4 min(L(j))(cid:17) Ym∈L(j)(cid:16)1 − 3 m(cid:17). We combine (131) with (132): XA∈Eβ,γ+1(A0∪E) ζ(β(γ + 1), A) s(β,γ+1,A)Xk=1 r(A, k)kx(k) Φ,Ak ζ(β(γ + 1), A) s(β,γ+1,A)Xk=1 r(A, k)kx(k) Φ,Ak ζ(β(γ + 1), A)r(A, 1)kx(1) Φ,Ak ζ(β(γ + 1), A) s(β,γ+1,A)Xk=2 r(A, k)kx(k) Φ,Ak Assume now that γ ≤ β is a limit ordinal number and that the claim holds for all γ′ < γ. Let L ∈ [N ]ω, A0 be a subset of L that is either empty or a singleton, and Φ : Sβγ(A0)∩[L]<ω → X be a semi-embedding of Sβγ ∩[L]<ω into X starting at A0 that is c-refined. We distinguish between two cases, namely whether A0 is a singleton or whether it is empty. In the first case, A0 = {a0}, for some a0 ∈ L. By Remark 9.6, the map Ψ with Ψ = Φ can be seen as a semi-embedding of Sβη(γ,a0) ∩ [L]<∞ into X starting at A0, that is c-refined. If E ∈ MAX(Sβγ(A0) ∩ [L]<ω) then E ∈ MAX(Sβη(γ,a0)(A0) ∩ [L]<ω) and by (114) we have ((x(k) Ψ,A)s(β,η(γ,a0),A) k=1 )A∈Eβ,η(γ,a0)(A0∪E) = ((x(k) Φ,A)s(β,γ,A) k=1 )A∈Eβ,γ(A0∪E), whereas if {r(A, k) : A ∈ Eβ,γ(A0 ∪ E), 1 ≤ k ≤ s(β, γ, A)} is a (β, γ)-special family of convex combinations, by the remark following Definition 10.7, it is a (β, η(γ, a0))-special family of − ε − ≥ XA∈ Eβ,γ+1(A0∪E) = XA∈ Eβ,γ+1(A0∪E) + X1≤j<d l1( ¯Bj )>0 XA∈E (j) ≥ r(A(1), 1)(cid:16) c + X1≤j<d ≥(cid:16) X1≤j<d ≥(cid:16)1 − ≥(cid:16) c l1( ¯Bj )>0 1 l1( ¯Bj )>0 − ε − 2 2 β,γ+1(A0∪E) 4 min(L)(cid:17) 2 4 2 − ε − ζ(β, ¯Bj)(1 − r(A(1), 1))(cid:16) c ζ(β, ¯Bj)(cid:17)(cid:16) c min( ¯B)(cid:17)(cid:16) c min(L)(cid:17) Ym∈L(cid:16)1 − min(L)(cid:17) Ym∈L(1)(cid:16)1 − min(L)(cid:17) Ym∈L(1)(cid:16)1 − m(cid:17). − ε − 4 2 1 4 − ε − 4 min(L)(cid:17) Ym∈L(j)(cid:16)1 − 1 m(cid:17) 1 m(cid:17) m(cid:17) (by Lemma 3.7 (v)) 1 58 P. MOTAKIS AND TH. SCHLUMPRECHT convex combinations as well. Applying the inductive assumption for η(γ, a0) < γ yields (133) XA∈Eβ,γ(A0∪E) ζ(βγ, A) s(β,γ,A)Xk=1 XA∈Eβ,η(γ,a0)(A0∪E) ≥(cid:18) c − ε − = 2 Φ,A(cid:13)(cid:13)(cid:13) r(A, k)(cid:13)(cid:13)(cid:13)x(k) ζ(βη(γ, a0), A) 4 min(L)(cid:19) ∞Ym∈L(cid:18)1 − Ψ,A(cid:13)(cid:13)(cid:13) r(A, k)(cid:13)(cid:13)(cid:13)x(k) s(β,η(γ,a0),A)Xk=1 m(cid:19) . 1 In the second case, A0 is empty. Let B ∈ MAX(Sβγ ∩ [L]<ω) and set a0 = min(B). By Remark 9.6, if L′ = L ∩ [a0, ∞), then the map Ψ : Sβη(γ,a0)(A0) ∩ [L′] → X with Ψ(A) = Φ(A0 ∪ A), is a semi-embedding of Sβη(γ,a0) ∩ [L′]<ω into X starting at A0 that is c-refined. By (115) we obtain ((x(k) Ψ,A)s(β,η(γ,a0),A) k=1 )A∈Eβ,η(γ,a0)(B) = ((x(k) Φ,A)s(β,γ,A) k=1 )A∈Eβ,γ(A0∪E) whereas if {r(A, k) : A ∈ Eβ,γ(B), 1 ≤ k ≤ s(β, γ, A)} is (β, γ)-special family of convex combinations, by the Remark following Definition 10.8, it is a (β, η(γ, a0))-special family of convex combinations as well. The result follows in the same manner as in (133). (cid:3) Theorem 11.6. Assume that X is a reflexive and separable Banach space, with the property that Sz(X) ≤ ωβ and Sz(X∗) < β. Then for no L ∈ [N]ω, does there exists a semi-embedding of Sβ2 ∩ [L]<ω into X, starting at ∅. Proof. By the main Theorem of [33] we can embed X into a reflexive space Z with basis, so that Sz(Z) = Sz(X), and Sz(Z ∗) = Sz(X∗). Thus we may assume that X has a basis, which must be shrinking and boundedly complete, since X is reflexive. By re-norming X, we may assume that the bases of X and X∗ are bimonotone. Choose α0 with Sz(X∗) ≤ ωα0 < β (this is possible due to the form of β). Note that CB(Sα0) = ωα0 + 1 < β + 1 = CB(Fβ,β). (134) Towards a contradiction, assume that there exists L ∈ [N]ω and a semi-embedding Ψ of Sβ2 ∩ [L]<ω into X, starting at ∅. By Lemma 9.8 there exist 1 ≥ c > 0, M ∈ [L]ω, and a semi- embedding Φ of Sβ2 ∩ [M ]<ω into X, starting at ∅, that is c-refined. Applying Proposition 11.5, we may pass to a further subset of M , again denoted by M , so that (127) holds. Fix 0 < ε < c/3 and apply Theorem 6.1 to the space X∗ and the ordinal number a0 to find a further subset of M , which we again denote by M , so that (69) is satisfied. By Propositions 2.2 and 2.13 we may pass to a subset of M , again denoted by M , so that Sα0 ∩ [M ]<ω ⊂ Fβ,β. By Lemma 10.2 we obtain that for any B ∈ MAX(Sβ2 ∩ [M ]<ω) and A ∈ Eβ,β(B), there exists A ∈ MAX(Sα0) with (135) A (cid:22)(cid:8)min(Cpβ,β(A, k)) : 1 ≤ k ≤ s(β, β, A)(cid:9) . Choose B ∈ MAX(Sβ2 ∩[M ]<ω). We will define a (β, β)-special family of convex combina- } ∈ tions {r(A, k) : A ∈ Eβ,β(B), 1 ≤ k ≤ s(β, β, A)}. For A ∈ Eβ,β(B) let A = {aA MAX(Sα0 ) be as in (135). For 1 ≤ k ≤ s(β, β, A) set 1 , . . . , aA dA (136) r(A, k) =(cid:26) ζ(α0, Ak) 0 if k ≤ # A otherwise, A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 59 1 , . . . , aA k } for 1 ≤ k ≤ # A. We will show that this family satisfies Definition where Ak = {aA 10.8 (i) and (ii). The first assertion is straightforward, to see the second one let A(1), A(2) ∈ Eβ,β(B), so that if (Dk(A(1)))s(β,β,A) are the maximal chains of Aβ,β(B) provided by Lemma 10.2, then for some k we have Dk(A(1)) = Dk(A(2)). By Lemmas 10.6 and 10.7 we obtain min(Cpβ,β(A(1), m)) = min(Cpβ,β(A(2), m)) for m = 1, . . . , k, which implies A(1) m for m = 1, . . . , min{k, # A(1)}. By (136) it easily follows that r(A(1), k) = r(A(2), k). Since (127) is satisfied, we obtain and (Dk(A(2)))s(β,β,A′) m = A2 k=1 i=1 (137) XA∈Eβ,β(B) ζ(β2, A) dAXk=1 Φ,A(cid:13)(cid:13)(cid:13) ≥ ζ(α0, Ak)(cid:13)(cid:13)(cid:13)x(k) c 3 . For each A ∈ Eβ,β(B) and k = 1, . . . , dA choose f (k) ran(f (k) A ) ⊂ ran(x(k) Φ,A) A in SX ∗, with f (k) A (x(k) A ) = kx(k) A k and i=1 Cpβ,β(A,i)), max supp(x∪k ⊂(cid:0) max supp(x∪k−1 ⊂(cid:0) min(Cpβ,β(A, k − 1)), min(Cpβ,β(A, k + 1))(cid:1) = (max( Ak−1), max( Ak+1)), i=1Cpβ,β(A,i))(cid:3) where the third inclusion follows from Definition 9.5 (ii). As (69) is satisfied, we obtain that for all A ∈ Eβ,β(B) < ε. A (x(k) dAXk=1 ζ(α0, Ak)f (k) ζ(α0, Ak)f (k) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) A (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ζ(α0, Ak)(cid:13)(cid:13)x(k) Φ,A(cid:13)(cid:13) (by (137)) A  s(β,β,A)Xm=1 A (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ζ(α0, Ak)f (k) ζ(α0, Ak)f (k) ζ(α0, Ak)f (k) x(m) A (xA) (by (113)) < ε (by (138)). Φ,A) (by choice of f (k) A ) Φ,A (since ran(f (k) A ) ⊂ ran(x(k) Φ,A) (138) We finally calculate c 3 ≤ XA∈Eβ,β(B) = XA∈Eβ,β(B) = XA∈Eβ,β(B) = XA∈Eβ,β(B) ≤ XA∈Eβ,β(B) ζ(β2, A) ζ(β2, A) ζ(β2, A) dAXk=1 dAXk=1 dAXk=1 dAXk=1 ζ(β2, A)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dAXk=1 ζ(β2, A) This contradiction completes the proof. (cid:3) Before proving Corollary 1.2 we will need the following observation. Proposition 11.7. Let X be a Banach space, α < ω1 and L be an infinite subset of the natural numbers so that there exist numbers 0 < c < C and a map Φ : Sα ∩ [L]<ω → X that 60 P. MOTAKIS AND TH. SCHLUMPRECHT is a c-lower-dα,∞ and C-upper-dα,1 embedding. Then for every β < α there exists n ∈ N and a map Φβ : Sβ ∩ [L ∩ (n, ∞)]<ω → X that is a c-lower-dβ,∞ and C-upper-dβ,1 embedding. Before proving Proposition 11.7 we need some preliminary observation. Lemma 11.8. Let α be an ordinal number and A ⊂ [0, α] satisfying: (i) α ∈ A and (ii) if β ∈ A and γ < β, then there is γ ≤ η < β with η ∈ A. Then A = [0, α]. Proof. If the conclusion is false, let α0 be the minimum ordinal for which there exists A0 ( [0, α0] satisfying (i) and (ii). Fix an arbitrary γ < α0. By the properties of A0, there exists γ ≤ α < α0 with α ∈ A0. If we set A = A0 ∩[0, α], then it easily follows that A and α satisfy (i) and (ii). By the minimality of α0, we conclude A = [0, α] and hence, since γ ∈ [0, α] = A and A ⊂ A0, we have γ ∈ A0. Since γ was chosen arbitrarily we conclude [0, α0] = A0, a contradiction that completes the proof. (cid:3) Lemma 11.9. Let α < ω1 be a limit ordinal number. Then there exists a sequence of successor ordinal numbers (µ(α, n))n satisfying: (i) µ(α, n) < α for all n ∈ N and limn µ(α, n) = α, (ii) Sα = {A ∈ [N]<ω : A ∈ Sµ(α,min(A))} ∪ {∅} and Sµ(α,n) ∩(cid:2)[n, ∞)(cid:3)<ω ⊂ Sα for all n ∈ N, and (iii) for A ∈ Sα \ {∅}, z(α,A) = z(µ(α,min(A)),A). Proof. We define (µ(α, n))n by transfinite recursion on the set of countable limit ordinal numbers. For α = ω we set (µ(ω, n))n = (λ(ω, n))n. If α is a limit ordinal so that for all α′ < α the corresponding sequence has been defined, set for each n ∈ N µ(α, n) =(cid:26) λ(α, n) µ(λ(α, n), n) if λ(α, n) is a successor ordinal number, or otherwise. The fact that (ii), (iii) and the first part of (i) hold is proved easily by transfinite induc- tion using (23) in Corollary 2.7 and the definition of repeated averages. To show that limn µ(α, n) = α, we will show that for arbitrary L ∈ [N]ω we have supn∈L µ(α, n) = α. Fix L ∈ [N]ω and β < α. Then, since CB(Sα ∩ [L]<ω) = ωα + 1 > ωβ + 1, we have ∅ ∈ (Sα ∩ [L]<ω)(ωβ +1) and hence there exists n ∈ L with {n} ∈ (Sα ∩ [L]<ω)(ωβ ). Using (4) we obtain ∅ ∈ (Sα ∩ [L]<ω)(ωβ )({n}) ⊂ S (ωβ ) ({n}) = (Sα({n}))(ωβ ) = (Sµ(α,n)({n}))(ωβ ) = S (ωβ ) µ(α,n)({n}), α which implies {n} ∈ S (ωβ ) µ(α,n), i.e. CB(Sµ(α,n)) > ωβ + 1 which yields µ(α, n) > β. (cid:3) Remark. It is uncertain whether Sµ(α,n) ⊂ Sµ(α,n+1), it is even uncertain whether µ(α, n) ≤ µ(α, n + 1). It is true however that for 2 ≤ n ≤ m we have µ(α, n) ≤ µ(α, m) + 1. Proof of Proposition 11.7. We shall first treat two very specific cases. In the first case, α = β + 1. Fix n0 > 2 with n0 ∈ L and B0 ∈ MAX(Sβ ∩ [L]<ω) with min(B0) = n0. Define n = max(B0) and Φβ : Sβ ∩ [L ∩ (n, ∞)]<ω → X with Φβ(A) = n0Φ(B0 ∪ A). Then Φβ is the desired embedding. In the second case, α is a limit ordinal and for some n0 > 2 with n0 ∈ L we have β + 1 = µ(α, n0). Fix B0 ∈ MAX(Sβ ∩ [L]<ω) with min(B0) = n0, define n = max(B0) and A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 61 Φβ : Sβ ∩ [L ∩ (n, ∞)]<ω → X with Φβ(A) = n0Φ(B0 ∪ A). Then, using the properties of (µ(α, k))k, it can be seen that Φβ is well defined and it is the desired embedding. In the general case, define A to be the set of all β ≤ α for which such an n and Φβ exist. Since α ∈ A, it remains to show that A satisfies (ii) of Lemma 11.8. Indeed, fix β ∈ A and γ < β. If β = η + 1, then by the first case we can deduce that η ∈ A and γ ≤ η < β. Otherwise, β is a limit ordinal. Let Φβ and nβ witness the fact that β ∈ A and by Lemma 11.9 (i) we may choose n ∈ L with n > nβ so that µ(β, n) > γ + 1. If η is the predecessor of µ(β, n), then by the second case we deduce that η ∈ A and γ ≤ η < β. (cid:3) , α cannot be the Szlenk index of some separable Banach space. Proof of Corollary 1.2. We first recall a result by Causey [5, Theorem 6.2] which says that for a countable ordinal ξ it follows that γ = ωξ is the Szlenk index of some separable Banach space X, if and only if ξ is not of the form ξ = ωη, with η being a limit ordinal. Since α = ωωωα "(a)⇒(b)" From (a) and Causey's result we have Sz(X) < ωα and Sz(X∗) < ωα, and thus there exists a ξ < ω1 with β = ωωξ ≤ ωωξ+1 < α, so that Sz(X) < ωβ and Sz(X) < β. It follows therefore from Theorem 11.6 that for no L ∈ [N]ω there exist numbers 0 < c < C and a map Φ : Sβ2 ∩ [L]<ω → X that is a c-lower-dβ2,∞ and C-upper-dβ2,1 embedding. Since β2 ≤ ωωξ+1 "(b)⇒(a)" follows from Theorems 8.1 and 8.3. < α we conclude our claim from Proposition 11.7 (cid:3) In order to proof Corollary 1.3 recall that every separable Banach space is isometrical equivalent to a subspace of C[0, 1], the space of continuous functions on [0, 1]. The set SB of all closed subspaces of C[0, 1] is given the Effros-Borel structure, which is the σ-algebra generated by the sets {F ∈ SB : F ∩ U 6= ∅}, where U ranges over all open subsets of C[0, 1]. Proof of Corollary 1.2. By [26, Theorem D] the set Cα = {X ∈ SB : X reflexive, and max(Sz(X), Sz(X∗)) ≤ α} is analytic. So it is left to show that its complement is also analytic. Since by Corollary 1.3 SB \ Cα = {X ∈ SB : X not reflexive} ∪ {X ∈ SB : Xreflexive and max(Sz(X), Sz(X∗)) > α} = {X ∈ SB : X not reflexive} ∪ {X ∈ SB : (Sα, d1,α) bi-Lipschitzly embeds into X} and since by [10, Corollary 3.3] the set of reflexive spaces in SB is co-analytic, we deduce our claim from the following Lemma. This Lemma seems to be wellknown, but for completeness we include a short proof. (cid:3) Lemma 11.10. Let (M, d) be a separable metric space. Then A(M ) = {X ∈ SB : (M, d) bi-Lipschitzly embeds into X} is analytic in SB. Proof. After passing to a countable dense subset of M we can assume that M is countable , i.e., the set of all sequences in NN as j=1 ∈ NN. By the Kuratowski-Ryll-Nardzewski Selection s = (sn : n ∈ N) with sn = (sn,j)∞ Theorem [17, Theorem 12.13] there are Borel maps dn : SB → C[0, 1], n ∈ N, so that for and we enumerate M into {mn : n ∈ N}. The set(cid:0)NN(cid:1)N is a Polish space with respect to the product topology. We write an element s ∈(cid:0)NN(cid:1)N every X ∈ SB, the sequence(cid:0)dn(X)(cid:1) is in X and dense in X. Put D(X) = {dn(X) : n ∈ N}, for X ∈ SB. We first note that there is a bi-Lipschitz map Ψ : M → X if and only if ∃C > 1 ∀n ∈ N ∃Ψn : M → D(X), so that (Ψn(m)) is Cauchy in X for all m ∈ M and 62 P. MOTAKIS AND TH. SCHLUMPRECHT d(m, m′) < lim n→∞ kΨn(m) − Ψn(m′)k < Cd(m, m′) for all m, m′ ∈ M . 1 C Since the set × SB : ∀k ∈ N ∀l ∈ N ∃n0 ∈ N ∀n1, n2 ≥ n0 ∃C ∈ N ∀k, l ∈ N ∃n0 ∈ N ∀n ≥ n0  (s, X) ∈(cid:0)NN(cid:1)N is Borel in the product space SB×(cid:0)NN(cid:1)N which there is a bi-Lipschitz map Ψ : M → X (consider for each s ∈ (cid:0)NN(cid:1)N , its projection on the second coordinate is analytic. On the other hand the image of this projection consists exactly of the spaces X ∈ SB for , n ∈ N and (cid:3) 1 C d(mk, ml) < kdsn,k (X) − dsn,l(X)k < Cd(mk, ml) X ∈ SB, the map Ψn : M ∋ mk 7→ dsn,k (X) ∈ X). kdsn1,k (X) − dsn2,k (X)k < 1 l  12. Final comments and open questions The proof of Theorem A, yields the following equivalences. "(a)⇒(b) " follows from Proposition 7.2, "(d)⇒(a)" from Theorem 11.6, while "(b)⇒(c)⇒(d)" is trivial. Corollary 12.1. For a separable Banach space X the following statements are equivalent a) X is not reflexive, b) For all α < ω1 there exists for some numbers 0 < c < C, a c-lower d∞,α, C-upper d1,α embedding of Sα into X. c) For all α < ω1 there is a map Ψα : Sα → X and some 0 < c ≤ 1, so that for all A, B, C ∈ Sα, with the property that A (cid:23) C, B (cid:23) C and A \ C < B \ C cd1,α(A, B) ≤ kΨ(A) − Ψ(B)k ≤ d1,α(A, B). d) For all α < ω1, there is an L ∈ [N]<ω and a semi embedding Ψα : Sα ∩ [L]<ω → X. As mentioned before we can consider for α < ω1 and A ∈ Sα the vector zA to be an element in ℓ+ 1 , with kxkℓ1 ≤ 1. We define Tα = {(A, B) ∈ Sα × Sα : ∃C (cid:22) A and C (cid:22) A, with A \ C < B \ C}. We note that d1,α(A, B) = kzA − zBk1, for (A, B) ∈ Tα. Using this notation we deduce the following sharpening of [30, Theorem 3.1] from Corollary 12.1 Corollary 12.2. Let X be a separable Banach space. Then the following are equivalent: a) X is not reflexive. b) For all α < ω1 there is a map Ψα : Sα → X and some 0 < c ≤ 1, so that cd1,α(A, B) ≤ kΨα(A) − Ψα(B)k ≤ d1,α(A, B), whenever (A, B) ∈ Tα. Next we would like to show that the James space J (c.f. [14, Definition 4.43]) has the property that Sz(J) = Sz(J ∗) = ω, and thus, since J is not reflexive, that Theorem C becomes false without the assumption that X is reflexive. X we have CkPn an upper ℓp-estimate, if there is a constant C so that for every block sequence (xk)n If X is a Banach space with an FDD and 1 < p ≤ ∞. The space X is said to satisfy k=1 in k=1 kxkkp)1/p. A lower ℓp-estimate is defined in the obvious way. If X is a Banach space with an FDD that satisfies an upper ℓp-estimate, note that by a standard argument, the FDD of X has to be shrinking and hence X∗ is separable. Therefore the assumption that X has an FDD with an upper ℓp-estimate implies by [25, Theorem 3.4] that Sz(X) = ω. k=1 xkk ≤ (Pn A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 63 Example 12.3. Let J denote James space, the above implies that Sz(J) = Sz(J ∗) = ω. To see that, we recall that J has two bases (di)i and (ei)i, the first one of which is shrinking and satisfies an upper ℓ2-estimate, whereas the second one is boundedly complete and satisfies a lower ℓ2-estimate. The first fact yields that Sz(J) ≤ ω, whereas the second fact and a duality argument yields that the space X spanned by the basis (e∗ i )i satisfies an upper ℓ2-estimate, i.e. Sz(X) ≤ ω. It is well known that X is of co-dimension one in J ∗ and that J ∗ is isomorphic to its hyperplanes and therefore Sz(J ∗) ≤ ω. This Example shows that Theorem C cannot be true if we omit the requirement that X is reflexive. Nevertheless the following variation of Corollary 1.2 holds. Corollary 12.4. Assume that α < ω1 is an ordinal, for which α = ωα. Then the following statements are equivalent for a separable Banach space X. a) X is reflexive and max(Sz(X), Sz(X∗)) ≤ α, b) There is no map Ψ : Sα → X and 0 < c ≤ 1 so that for all A, B, C ∈ Sα with the property that C (cid:22) A, C (cid:22) B, and A \ C < B \ C, we have cd1,α(A, B) ≤ kΨ(A) − Ψ(B)k ≤ d1,α(A, B). Proof. Let Ψ : Sα → X satisfying the condition stated in (b) for some c > 0. Then Ψ =(cid:0)Ψ−Ψ(∅)(cid:1)/2 also has this property for c/2, maps Sα into BX and Ψ(∅) = 0. "(a)⇒(b)" follows from Theorem 11.6, the in the proof of Corollary 1.2 cited result from [5], and the same argument involving Proposition 11.7 in the proof of Corollary 1.2. "(b)⇒ (a)" follows from Proposition 7.2, Theorems 8.1 and 8.3. (cid:3) Remark. The statement of Corollary 12.4 also holds for α = ω. This can be seen from the proof of the main result in [9]. We finish with stating two open problems. Problems 12.5. a) Does there exists a family of metric spaces (Mi, di) which is a family of test spaces for reflexivity in the sense of [25], i.e., for which it is true that a separable Banach space X is reflexive if and only if not all of the Mi uniformly bi-Lipschitzly embed into X? b) Does there exists a countable family of metric spaces (Mi, di) which is a family of test spaces for reflexivity? c) It follows from Theorem B, that if X is a separable Banach space with non separable bidual then (Sα, d1,α) bi-Lipschitzly embeds into X, for all α < ω1. Is the converse true, or in Ostrovskii's language, are the spaces (Sα, d1,α), α < ω1 test spaces for spaces with separable bi-duals? References [1] D. Alspach and S. Argyros,Complexity of weakly null sequences. Dissertationes Math. (Rozprawy Mat.) 321 (1992), 44 pp. [2] D. Alspach, R. Judd, and E. Odell, The Szlenk index and local ℓ1-indices. Positivity 9 (2005) 1 -- 44. [3] S. A. Argyros, V. Kanellopoulos, K. Tyros, Higher order spreading models Fundamenta Mathematicae 221 (2013), 23 -- 68 [4] S. Argyros and I. Gasparis, Unconditional structures of weakly null sequences. Trans. Amer. Math.Soc. 353 (2001), no. 5, 2019 -- 2058 [5] R. Causey, Concerning the Szlenk index, preprint, arXive 1501.06885. [6] K. Ball, The Ribe programme. S´eminaire Bourbaki. Vol. 2011/2012. Expos´es 1043 -- 1058. Astrisque No. 352 (2013) 147 -- 159. 64 P. MOTAKIS AND TH. SCHLUMPRECHT [7] F. Baudier, Metrical characterization of super-reflexivity and linear type of Banach spaces. Arch. Math. (Basel) 89 (2007) no. 5, 419 -- 429. [8] F. Baudier, On the (β)-distortion of countably branching trees, preprint. [9] F. Baudier, N. Kalton, and G. Lancien, A new metric invariant for Banach spaces. Studia Math. 199 (2010), no. 1, 73 -- 94. [10] B. Bossard, An ordinal version of some applications of the classical interpolation theorem, Fund. Math. 152 (1997), no. 1, 55 -- 74. [11] J. Bourgain, A metric interpretation of super reflexivity, Israel J. Math. 56 (1986) 222 -- 230. [12] J. Bourgain, V. Milman, and H. Wolfson, On type of metric spaces. Trans. Amer. Math. Soc. 294 (1986), no. 1, 295 -- 317. [13] S. J. Dilworth, D. Kutzarova, G. Lancien, and N. L. Randrianarivony, Asymptotic geometry of Banach spaces and uniform quotient maps, Proc. Amer. Math. Soc. 142 (2014), no. 8, 2747 -- 2762. [14] M. Fabian, P. Habala, P. Hjek, V. Montesinos, V. Zizler, Banach space theory. The basis for linear and nonlinear analysis. CMS Books in Mathematics/Ouvrages de Mathmatiques de la SMC. Springer, New York, 2011. xiv+820 pp. [15] I. Gasparis, A dichotomy theorem for subsets of the power set of the natural numbers, Proc. Amer. Math. Soc. 129, no. 3, 759 -- 764. [16] R. C. James, Some self-dual properties of normed linear spaces, Symposium on infinite Dimensional Topology (1972) 159 -- 175 [17] A. Kechris, Classical descriptive set theory. Graduate Texts in Mathematics, 156. Springer-Verlag, New York, 1995. xviii+402 pp. [18] W. B. Johnson, and G. Schechtman, Diamond graphs and super-reflexivity. J. Topol. Anal. 1 (2009), no. 2, 177 -- 189. [19] B. Kloeckner, Yet another short proof of Bourgain's distortion estimate for embedding of trees into uniformly convex Banach spaces. Israel J. Math. 200 (2014), no. 1, 419 -- 422. [20] D. Kutzarova, k − β and k-nearly uniform convex Banach spaces. J. Math. Anal. Appl. 162 (1991) 322 -- 338. [21] M Mendel, and A. Naor, Scaled Enflo type is equivalent to Rademacher type. Bull. Lond. Math. Soc. 39 (2007), no. 3, 493 -- 498. [22] M Mendel, and A. Naor, Markov convexity and local rigidity of distorted metrics Computational geom- etry (SCG'08), 49 - 58, ACM, New York, 2008. [23] A. Naor, An introduction to the Ribe program. Jpn. J. Math. 7 (2012), no. 2, 167 -- 233. [24] E. Odell and Th. Schlumprecht. Trees and branches in Banach spaces, Trans. Amer. Math. Soc. 354, (2002) , 4085 -- 4108. [25] E. Odell and Th. Schlumprecht, Embedding into Banach spaces with finite dimensional decompositions. RACSAM. Rev. R. Acad. Cienc. Exactas F´ıs. Nat. Ser. A Mat. 100 (2006), no. 1-2, 295 -- 323. [26] E. Odell, Th. Schlumprecht and A. Zs´ak, A new infinite game in Banach spaces with applications, Banach spaces and their applications in analysis, 147 -- 182, Walter de Gruyter, Berlin, 2007. [27] E. Odell, Th. Schlumprecht and A. Zs´ak, Banach spaces of bounded Szlenk index. Studia Math. 183 (2007), no. 1, 63 -- 97. [28] M. Ostrovskii, Embeddability of locally finite metric spaces into Banach spaces is finitely determined. Proc. Amer. Math. Soc. 140 (2012), no. 8, 2721 -- 2730. [29] M. Ostrovskii, Metric Embeddings. Bilipschitz and coarse embeddings into Banach spaces. De Gruyter Studies in Mathematics, 49. De Gruyter, Berlin (2013) xii+372 pp. ISBN: 978-3-11-026340-4 [30] M. Ostrovskii, On metric characterizations of the RadonNikod´ym and related properties of Banach spaces, Journal of Topology and Analysis, 6,no. 3 (2014) 441 -- 464. [31] A. Proch´azka, Linear properties of Banach spaces and low distortion embeddings of metric graphs, preprint, arXiv:1603.00741. [32] M. Ribe, On uniformly homeomorphic normed spaces. Ark. Mat. 14 (1976), no. 2, 237 -- 244. [33] Th. Schlumprecht, On Zippin's embedding theorem of Banach spaces into Banach spaces with bases. Adv. Math. 274 (2015), 833 -- 880. [34] P. Suppes, Axiomatic set theory, Dover books in Advanced Mathematics. [35] W. Szlenk, The non-existence of a separable reflexive Banach space universal for all separable reflexive Banach spaces, Studia Math. 30 (1968), 53 -- 61. [36] M. Zippin, Banach spaces with separable duals, Trans. Amer. Math. Soc. 310 (1988) no. 1, 371 -- 379. A METRIC INTERPRETATION OF REFLEXIVITY FOR BANACH SPACES 65 Department of Mathematics, Texas A&M University, College Station, TX 77843, USA E-mail address: [email protected] Department of Mathematics, Texas A&M University, College Station, TX 77843, USA and Faculty of Electrical Engineering, Czech Technical University in Prague, Zikova 4, 166 27, Prague E-mail address: [email protected]
1005.0596
1
1005
2010-05-04T18:01:15
Spaceability in Banach and quasi-Banach sequence spaces
[ "math.FA" ]
Let $X$ be a Banach space. We prove that, for a large class of Banach or quasi-Banach spaces $E$ of $X$-valued sequences, the sets $E-\bigcup _{q\in\Gamma}\ell_{q}(X)$, where $\Gamma$ is any subset of $(0,\infty]$, and $E-c_{0}(X)$ contain closed infinite-dimensional subspaces of $E$ (if non-empty, of course). This result is applied in several particular cases and it is also shown that the same technique can be used to improve a result on the existence of spaces formed by norm-attaining linear operators.
math.FA
math
Spaceability in Banach and quasi-Banach sequence spaces G. Botelho∗, D. Diniz, V. V. F´avaro†, D. Pellegrino‡ Abstract Let X be a Banach space. We prove that, for a large class of Banach or quasi-Banach spaces E of X-valued sequences, the sets E − Sq∈Γ ℓq(X), where Γ is any subset of (0,∞], and E − c0(X) contain closed infinite-dimensional subspaces of E (if non-empty, of course). This result is applied in several particular cases and it is also shown that the same technique can be used to improve a result on the existence of spaces formed by norm-attaining linear operators. Introduction A subset A of a Banach or quasi-Banach space E is µ-lineable (space- able) if A ∪ {0} contains a µ-dimensional (closed infinite-dimensional) lin- ear subspace of E. The last few years have witnessed the appearance of lots of papers concerning lineability and spaceability (see, for example, [1, 2, 3, 10, 13]). The aim of this paper is to explore a technique to prove lineability and spaceability that can be applied in several different settings. It is our opinion that this technique was first used in the context of line- ability/spaceability in our preprint [4], of which this paper is an improved version. Let c denote the cardinality of the set of real numbers R. In [10] it is proved that ℓp − ℓq is c-lineable for every p > q ≥ 1. With the help of [7] this result can be substantially improved in the sense that ℓp −S1≤q<p ℓq is spaceable for every p > 1. In this paper we address the following questions: What about the non-locally convex range 0 < p < 1? Can these results be generalized to sequence spaces other than ℓp? ∗Supported by CNPq Grant 306981/2008-4 and INCT-Matem´atica. †Supported by Fapemig Grant CEX-APQ-00208-09. ‡Supported by INCT-Matem´atica, PROCAD-NF-Capes, CNPq Grants 620108/2008- 8 (Ed. Casadinho) and 301237/2009-3. 2010 Mathematics Subject Classification: 46A45, 46A16, 46B45. 1 As to the first question, it is worth recalling that the structure of quasi- Banach spaces (or, more generally, metrizable complete tvs, called F -spaces) is quite different from the structure of Banach spaces. For our purposes, the consequence is that the extension of lineability/spaceability arguments from Banach to quasi-Banach spaces is not straightforward in general. For example, in [14, Section 6] it is essentially proved (with a different termi- nology) that if Y is a closed infinite-codimensional linear subspace of the Banach space X, then X −Y is spaceable. A counterexample due to Kalton [5, Theorem 1.1] shows that this result is not valid for quasi-Banach spaces (there exists a quasi-Banach space K with an 1-dimensional subspace that is contained in all closed infinite-dimensional subspaces of K). Besides, the search for closed infinite-dimensional subspaces of quasi-Banach spaces is a quite delicate issue. Even fundamental facts are unknown, for example the following problem is still open (cf. [6, Problem 3.1]): Does every (infinite- dimensional) quasi-Banach space have a proper closed infinite-dimensional subspace? Nevertheless we solve the first question in the positive: as a par- ticular case of our results we get that ℓp −S1≤q<p ℓq is spaceable for every p > 0 (cf. Corollary 1.7). As to the second question, we identify a large class of vector-valued se- quence spaces, called invariant sequence spaces (cf. Definition 1.1), such that if E is an invariant Banach or quasi-Banach space of X-valued se- quences, where X is a Banach space, then the sets E −Sq∈Γ ℓq(X), where Γ is any subset of (0,∞], and E − c0(X) are spaceable whenever they are non- empty (cf. Theorem 1.3). Several classical sequence spaces are invariant sequence spaces (cf. Example 1.2). In order to make clear that the technique we use can be useful in a variety of other situations, we finish the paper with an application to the c-lineability of sets of norm-attaining linear operators (cf. Proposition 2.1). From now on all Banach and quasi-Banach spaces are considered over a fixed scalar field K which can be either R or C. 1 Sequence spaces In this section we introduce a quite general class of scalar-valued or vector- valued sequence spaces and prove that certain of their remarkable subsets have spaceable complements. Definition 1.1. Let X 6= {0} be a Banach space. (a) Given x ∈ X N, by x0 we mean the zerofree version of x, that is: if x has only finitely many non-zero coordinates, then x0 = 0; otherwise, x0 = (xj)∞j=1 where xj is the j-th non-zero coordinate of x. (b) By an invariant sequence space over X we mean an infinite-dimensional 2 Banach or quasi-Banach space E of X-valued sequences enjoying the fol- lowing conditions: (b1) For x ∈ X N such that x0 6= 0, x ∈ E if and only if x0 ∈ E, and in this case kxk ≤ Kkx0k for some constant K depending only on E. (b2) kxjkX ≤ kxkE for every x = (xj)∞j=1 ∈ E and every j ∈ N. An invariant sequence space is an invariant sequence space over some Ba- nach space X. Several classical sequence spaces are invariant sequence spaces: Example 1.2. (a) Given a Banach space X, it is obvious that for every 0 < p ≤ ∞, ℓp(X) (absolutely p-summable X-valued sequences), ℓu p(X) (unconditionally p-summable X-valued sequences) and ℓw p (X) (weakly p- summable X-valued sequences) are invariant sequence spaces over X with their respective usual norms (p-norms if 0 < p < 1). In particular, ℓp, 0 < p ≤ ∞, are invariant sequence spaces (over K). (b) The Lorentz spaces ℓp,q, 0 < p < ∞, 0 < q < ∞ (see, e.g., [12, 13.9.1]). It is easy to see that these spaces are invariant sequence spaces (over K): indeed, given 0 6= x0 ∈ ℓp,q, the non-increasing rearrangement of x coincides with that of x0. So kxkp,q = kx0kp,q < ∞. (c) The Orlicz sequence spaces (see, e.g., [8, 4.a.1]). Let M be an Orlicz function and ℓM be the corresponding Orlicz sequence space. The condition M(0) = 0 makes clear that ℓM is an invariant sequence space (over K). For the same reason, its closed subspace hM is an invariant sequence space as well. (d) Mixed sequence spaces (see, e.g., [12, 16.4]). Given 0 < p ≤ s ≤ ∞ and a Banach space X, by ℓm(s;p) (X) we mean the Banach (p-Banach if 0 < p < 1) space of all mixed (s, p)-summable sequences on X. It is not difficult to see that ℓm(s;p) (X) is an invariant sequence space over X. Now we can prove our main result. Given an invariant sequence space E over the Banach space X, regarding both E and ℓp(X) as subsets of X N, we can talk about the difference E − ℓp(X) and related ones. Theorem 1.3. Let E be an invariant sequence space over the Banach space X. Then (a) For every Γ ⊆ (0,∞], E −Sq∈Γ ℓq(X) is either empty or spaceable. (b) E − c0(X) is either empty or spaceable. Proof. Put A = Sq∈Γ ℓq(X) in (a) and A = c0(X) in (b). Assume that E− A is non-empty and choose x ∈ E− A. Since E is an invariant sequence space, x0 ∈ E, and obviously x0 /∈ A. Writing x0 = (xj)∞j=1 we have that x0 ∈ E − A and xj 6= 0 for every j. Split N into countably many infinite 3 pairwise disjoint subsets (Ni)∞i=1. For every i ∈ N set Ni = {i1 < i2 < . . .} and define yi = ∞ Xj=1 xjeij ∈ X N. i = x0 for every i. So 0 6= y0 Observe that y0 i ∈ E for every i. Hence each yi ∈ E because E is an invariant sequence space. Let us see that yi /∈ A: in (a) this occurs because kyikr = kx0kr = kxkr for every 0 < r ≤ ∞ and in (b) because kxjk 9 0. Let K be the constant of condition 1.1(b1) and define s = 1 if E is a Banach space and s = s if E is a s-Banach space, 0 < s < 1. For (aj)∞j=1 ∈ ℓs, Xj=1 = K s(cid:13)(cid:13)x0(cid:13)(cid:13) ajs(cid:13)(cid:13)y0 j(cid:13)(cid:13) s(cid:13)(cid:13)(aj)∞j=1(cid:13)(cid:13) ajs = K s(cid:13)(cid:13)x0(cid:13)(cid:13) ajskyjks ≤ K s kajyjks = s ∞ Xj=1 < ∞. Xj=1 Xj=1 ∞ ∞ ∞ s s s Thus P∞j=1 kajyjk < ∞ if E is a Banach space and P∞j=1 kajyjks < ∞ if E is a s-Banach space, 0 < s < 1. In both cases the series P∞j=1 ajyj converges in E, hence the operator T : ℓs −→ E , T (cid:16)(aj)∞ j=1(cid:17) = ∞ Xj=1 ajyj i (cid:17)∞ i=1 ∈ ℓs, k ∈ N, such that z = limk→∞ T (cid:16)(cid:16)a(k) is well defined. It is easy to see that T is linear and injective. Thus T (ℓs) is a closed infinite-dimensional subspace of E. We just have to show that T (ℓs)−{0} ⊆ E − A. Let z = (zn)∞n=1 ∈ T (ℓs), z 6= 0. There are sequences i (cid:17)∞ i=1(cid:17) in E. Note that, (cid:16)a(k) for each k ∈ N, i (cid:17)∞ T (cid:16)(cid:16)a(k) Xi=1 Fix r ∈ N such that zr Nj, there are (unique) m, t ∈ N such that emt = er. Thus, for each k ∈ N, the r-th coordi- nate of T (cid:16)(cid:16)a(k) m xt. Condition 1.1(b2) assures that 6= 0. Since N = S∞j=1 i=1(cid:17) is the number a(k) convergence in E implies coordinatewise convergence, so i=1(cid:17) = i (cid:17)∞ a(k) i xjeij . a(k) i yi = Xi=1 Xj=1 Xi=1 Xj=1 xjeij = a(k) i ∞ ∞ ∞ ∞ ∞ zr = lim k→∞ a(k) m xt = xt · lim k→∞ a(k) m . 4 It follows that xt 6= 0. Hence limk→∞ a(k) i=1(cid:17) is a(k) mj-th coordinate of T (cid:16)(cid:16)a(k) m xjk = lim m kxjk = kxjk · lim i (cid:17)∞ k→∞a(k) k→∞ka(k) lim k→∞a(k) m = kzrk m xj. Defining αm = kzrk kxtk 6= 0. For j, k ∈ N, the kxtk 6= 0, m = αm kxjk for every j ∈ N. On the other hand, coordinatewise convergence gives limk→∞ ka(k) m xjk = kzmjk, so kzmjk = αmkxjk for each j ∈ N. Observe that m, which depends on r, is fixed, so the natural numbers (mj)∞j=1 are pairwise distinct (remember that Nm = {m1 < m2 < . . .}). (a) As x0 /∈ A, we have kx0kq = ∞ for all q ∈ Γ. Assume first that ∞ /∈ Γ. In this case, kzkq q = ∞ Xn=1 kznkq ≥ ∞ Xj=1 (cid:13)(cid:13)zmj(cid:13)(cid:13) q = ∞ Xj=1 m · kxjkq = αq αq q q = ∞, m ·(cid:13)(cid:13)x0(cid:13)(cid:13) for all q ∈ Γ, proving that z /∈ Sq∈Γ ℓq(X). If ∞ ∈ Γ, kzk∞ = sup n kznk ≥ sup j kzmjk = αm · sup j kxjk = αmkx0k∞ = ∞, proving again that z /∈ Sq∈Γ ℓq(X). (b) As x0 /∈ A, we have kxjk 9 0. Since (kzmjk)∞j=1 is a subsequence of (kznk)∞n=1 , kzmjk = αmkxjk for every j and αm 6= 0, it is clear that kznk 9 0. Thus z /∈ c0(X). Therefore z /∈ A in both cases, so T (ℓs) − {0} ⊆ E − A. We list a few consequences. When we write F ⊂ E we mean that E contains F as a linear subspace and E 6= F . We are not asking neither E to contain an isomorphic copy of F nor the inclusion F ֒→ E to be continuous. Corollary 1.4. Let E be an invariant sequence space over K. (a) If 0 < p ≤ ∞ and ℓp ⊂ E, then E − ℓp is spaceable. (b) If c0 ⊂ E, then E − c0 is spaceable. From the results due to Kitson and Timoney [7] we derive that ℓu p(X)− ℓq for p > 1, are spaceable. However, as is ℓp(X) for p ≥ 1, and ℓp − S0<q<p ℓq to the non-locally convex case: made clear in [7, Remark 2.2], their results are restricted to Fr´echet spaces (see the Introduction). Next we extend the spaceability of ℓu p(X) − ℓp(X) and ℓp − S0<q<p Corollary 1.5. ℓm(s;p) (X) − ℓp(X) and ℓu p(X) − ℓp(X) are spaceable for 0 < p ≤ s < ∞ and every infinite-dimensional Banach space X. Hence ℓw p (X) − ℓp(X) is spaceable as well. 5 it is non-empty. Although we think this is folklore, we have not been able to find a reference in the literature. So, for the sake of completeness, we include a short proof, which was kindly communicated to us by M. C. Matos. n=1 1 √n n=1 ∈ ℓr for all r > 2, for each Lemma 1.6. ℓp − S0<q<p Proof. Since (cid:16) 1√n(cid:17)∞ (yn)∞n=1 ∈ ℓq, 0 < q < 2, it follows from Holder's inequality that (cid:12)(cid:12)(cid:12)(cid:12) ℓq 6= ∅ for every p > 0. /∈ ℓ2 and (cid:16) 1√n(cid:17)∞ √n(cid:19)∞ yn(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:13)(cid:13)(cid:13)(cid:13) (cid:18) 1 Supposing that ℓ2 = S0<q<2 < ∞ for every (yn)∞n=1 ∈ ℓ2. So, consider, for each positive integer k, the continuous linear functional on ℓ2 defined by Tk ((yn)∞n=1) = n=1(cid:13)(cid:13)(cid:13)(cid:13)q′ · k(yn)∞n=1kq < ∞. 1√n yn(cid:12)(cid:12)(cid:12) ℓq, we have that ∞ Pn=1(cid:12)(cid:12)(cid:12) ∞ Xn=1 1√nyn. As sup k∈N Tk ((yn)∞n=1) = ∞ Xn=1 k Pn=1 (cid:12)(cid:12)(cid:12)(cid:12) 1 √n yn(cid:12)(cid:12)(cid:12)(cid:12) < ∞ Proof. By [9, Proposition 1.2(1)] we have that ℓm(∞;p) (X) = ℓp(X) ⊆ ℓm(s;p) (X), and by [9, Theorem 2.1], ℓm(s;p) (X) 6= ℓp(X). On the other hand, the identity operator on any infinite-dimensional Banach space fails to be absolutely p-summing for every 0 < p < ∞ (the case 1 ≤ p < ∞ is well known, and the case 0 < p < 1 follows from the fact that p-summing op- erators are q-summing whenever p ≤ q). So ℓu p(X) 6= ℓp(X). As ℓm(s;p) (X) and ℓu p(X) are invariant sequence spaces over X, the first assertion follows from Theorem 1.3. As ℓu p (X), the second assertion follows. ℓq is spaceable we have to check first that p(X) ⊆ ℓw Before proving that ℓp − S0<q<p for each (yn)∞n=1 ∈ ℓ2, by the Banach-Steinhaus Theorem we conclude that T ((yn)∞n=1) = lim k Tk ((yn)∞n=1) = ∞ Xn=1 1 √n yn defines a continuous linear functional on ℓ2 and it follows that (cid:16) 1√n(cid:17)∞ n=1 ∈ ℓ2 p(cid:17)∞ ℓq. So(cid:16)xn - a contradiction which proves that there is x ∈ ℓ2− S0<q<2 n=1 ∈ ℓp − S0<q<p ℓq. 2 6 Corollary 1.7. ℓp − S0<q<p ℓq is spaceable for every p > 0. Proof. We know that ℓp is an invariant sequence space over K and from ℓq 6= ∅. The result follows from Theorem Lemma 1.6 we have ℓp − S0<q<p 1.3. Remark 1.8. Theorem 1.3 can be applied in a variety of other situations. For example, for Lorentz spaces it applies to ℓq,r−ℓp for 0 < p < q and r > 0, and to ℓp,q − ℓp for 0 < p < q. We believe that the usefulness of Theorem 1.3 is well established, so we refrain from giving further applications. Although our results concern spaceability of complements of linear sub- spaces, the same technique gives the spaceability of sets that are not related to linear subspaces at all. Rewriting the proof of Theorem 1.3 we get: Proposition 1.9. Let E be an invariant sequence space over the Banach space X. Let A ⊆ E be such that: (i) For x ∈ E, x ∈ A if and only if x0 ∈ A. (ii) If x = (xj)∞j=1 ∈ A and y = (yj)∞j=1 ∈ E is such that (kyjk)∞j=1 is a multiple of a subsequence of (kxjk)∞j=1, then y ∈ A. (iii) There is x ∈ E − A with x0 6= 0. Then E − A is spaceable. 2 Norm-attaining operators In this section we show that the technique used in the previous section can be used in a completely different context. Specifically, we extend a result from [11] concerning the lineability of the set of norm-attaining operators. Given Banach spaces E and F and x0 ∈ E such that kx0k = 1 (x0 is said to be a norm-one vector ), a continuous linear operator u : E −→ F attains its norm at x0 if ku(x0)k = kuk. By NAx0(E; F ) we mean the set of continuous linear operators from E to F that attain their norms at x0. In [11, Proposition 6] it is proved that if F contains an isometric copy of ℓq for some 1 ≤ q < ∞, then NAx0(E; F ) is ℵ0-lineable. We generalize this result showing that this set is c-lineable: Proposition 2.1. Let E and F be Banach spaces so that F contains an isometric copy of ℓq for some 1 ≤ q < ∞, and let x0 be a norm-one vector in E. Then NAx0(E; F ) is c-lineable. Proof. The beginning of the proof follows the lines of the proof of [11, It suffices to prove the result for F = ℓq. Split N into Proposition 6]. 7 countably many infinite pairwise disjoint subsets (Ak)∞k=1. For each positive integer k, write Ak = {a(k) 2 < . . .} and define 1 < a(k) ℓ(k) q := {x ∈ ℓq : xj = 0 if j /∈ Ak} . Fix a non-zero operator u ∈ NAx0(E; F ) and proceed as in the proof of [11, Proposition 6] to get a sequence (u(k))∞k=1 of operators belonging to NAx0(E; ℓ(k) q ) such that ku(k)(x)k = ku(x)k for every k and every x ∈ E. By composing these operators with the inclusion ℓ(k) ֒→ ℓq we get operators (and we keep the notation u(k) for the sake of simplicity) belonging to NAx0(E; ℓq). For every (ak)∞k=1 ∈ ℓ1, ∞ Xk=1 akku(k)(x0)k = ku(x0)k akku(k)k = kaku(k)k = ak < ∞, ∞ Xk=1 ∞ Xk=1 q ∞ Xk=1 so the map T : ℓ1 −→ L(E; ℓq) , T ((ak)∞k=1) = ∞ Xk=1 aku(k) is well-defined. It is clear that T is linear and injective. Hence T (ℓ1) is a c-dimensional subspace of ℓq. Since the supports of the operators u(k) are pairwise disjoint, T (ℓ1) ⊆ NAx0(E; ℓq). References [1] R. M. Aron, V. I. Gurariy, J. B. Seoane-Sep´ulveda, Lineability and spaceabil- ity of sets of functions on R, Proc. Amer. Math. Soc. 133 (2005) 795 -- 803. [2] R. M. Aron, F. J. Garc´ıa-Pacheco, D. P´erez-Garc´ıa, J. B. Seoane-Sep´ulveda, On dense-lineability of sets of functions on R, Topology 48 (2009) 149 -- 156. [3] L. Bernal-Gonz´alez, Dense-lineability in spaces of continuous functions, Proc. Amer. Math. Soc. 136 (2008) 3163 -- 3169. [4] G. Botelho, D. Diniz, D. Pellegrino and E. Teixeira, A note on lineability, arXiv:0905.2677 (2009). [5] N. J. Kalton, The basic sequence problem, Studia Math. 116 (1995), 167- 187. [6] N. J. Kalton, Quasi-Banach spaces, Handbook of the Geometry of Banach spaces, Vol. 2, North-Holland, Amsterdam, 2003, pp. 1099 -- 1130. 8 [7] D. Kitson and R. Timoney, Some applications of operator ranges, preprint. [8] J. Lindenstrauss and L. Tzafriri, Classical Banach spaces I and II, Springer, 1996. [9] M. C. Matos, Mappings between Banach spaces that send mixed summable sequences into absolutely summable sequences, J. Math. Anal. Appl. 297 (2004), 833-851. [10] G. Munoz-Fernandez, N. Palmberg, D. Puglisi e J. B. Seoane-Sep´ulveda, Lineability in subsets of measure and function spaces, Linear Algebra Appl. 428 (2008), 2805-2812. [11] D. Pellegrino and E. Teixeira, Norm optimization problem for linear opera- tors in classical Banach spaces, Bull. Braz. Math. Soc. 40 (2009), 417-431. [12] A. Pietsch, Operator ideals, North-Holland Publishing Company, 1980. [13] D. Puglisi, J. B. Seoane-Sep´ulveda, Bounded linear non-absolutely summing operators, J. Math. Anal. Appl. 338 (2008), 292-298. [14] A. Wilansky, Semi-Fredholm maps of F K spaces, Math. Z. 144 (1975), 9-12. [Geraldo Botelho and Vin´ıcius V. F´avaro] Faculdade de Matem´atica, Uni- versidade Federal de Uberlandia, 38.400-902 - Uberlandia, Brazil, e-mails: [email protected], [email protected]. [Diogo Diniz] UAME-UFCG, Caixa Postal 10044 - 58.109-970, Campina Grande, Brazil, e-mail: [email protected]. [Daniel Pellegrino] Departamento de Matem´atica, Universidade Federal da Para´ıba, 58.051-900 - Joao Pessoa, Brazil, e-mail: [email protected]. 9
1607.04565
3
1607
2017-10-20T13:38:11
Extension and boundedness of operators on Morrey spaces from extrapolation techniques and embeddings
[ "math.FA" ]
We prove that operators satisfying the hypotheses of the extrapolation theorem for Muckenhoupt weights are bounded on weighted Morrey spaces. As a consequence, we obtain at once a number of results that have been proved individually for many operators. On the other hand, our theorems provide a variety of new results even for the unweighted case because we do not use any representation formula or pointwise bound of the operator as was assumed by previous authors. To extend the operators to Morrey spaces we show different (continuous) embeddings of (weighted) Morrey spaces into appropriate Muckenhoupt $A_1$ weighted $L_p$ spaces, which enable us to define the operators on the considered Morrey spaces by restriction. In this way, we can avoid the delicate problem of the definition of the operator, often ignored by the authors. In dealing with the extension problem through the embeddings (instead of using duality) one is neither restricted in the parameter range of the $p$'s (in particular $p=1$ is admissible and applies to weak-type inequalities) nor the operator has to be linear. Another remarkable consequence of our results is that vector-valued inequalities in Morrey spaces are automatically deduced. On the other hand, we also obtain $A_\infty$-weighted inequalities with Morrey quasinorms.
math.FA
math
EXTENSION AND BOUNDEDNESS OF OPERATORS ON MORREY SPACES FROM EXTRAPOLATION TECHNIQUES AND EMBEDDINGS JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Abstract. We prove that operators satisfying the hypotheses of the ex- trapolation theorem for Muckenhoupt weights are bounded on weighted Morrey spaces. As a consequence, we obtain at once a number of re- sults that have been proved individually for many operators. On the other hand, our theorems provide a variety of new results even for the unweighted case because we do not use any representation formula or pointwise bound of the operator as was assumed by previous authors. To extend the operators to Morrey spaces we show different (continuous) embeddings of (weighted) Morrey spaces into appropriate Muckenhoupt A1 weighted Lp spaces, which enable us to define the operators on the considered Morrey spaces by restriction. In this way, we can avoid the delicate problem of the definition of the operator, often ignored by the authors. In dealing with the extension problem through the embeddings (instead of using duality) one is neither restricted in the parameter range of the p's (in particular p = 1 is admissible and applies to weak-type inequalities) nor the operator has to be linear. Another remarkable consequence of our results is that vector-valued inequalities in Morrey spaces are automatically deduced. On the other hand, we also obtain A∞-weighted inequalities with Morrey quasinorms. 1. Introduction An important tool of modern harmonic analysis is the extrapolation the- orem due to Rubio de Francia. For a given operator T we suppose that for some p0, 1 ≤ p0 < ∞, and for every weight belonging to the Muckenhoupt class Ap0 the inequality (1.1) k T f Lp0,w(Rn)k =(cid:18)ZRn T f (x)p0w(x)dx(cid:19) 1 p0 ≤ c k f Lp0,w(Rn)k 2010 Mathematics Subject Classification. Primary 42B35, 46E30, 42B15, 42B20; Sec- ondary 42B25. Key words and phrases. Morrey spaces, embeddings, Muckenhoupt weights, extrapo- lation, Calder´on-Zygmund operators, Mihlin-Hormander multipliers, rough operators. The first author is supported by the grants MTM2014-53850-P of the Ministerio de Econom´ıa y Competitividad (Spain) and grant IT-641-13 of the Basque Gouvernment. The second author is supported by the German Academic Exchange Service (DAAD). 1 2 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL holds, where the constant c is independent of f and depends on w only in terms of its Ap0-constant, [w]Ap0 (see Definition 2.1). Then for every p, 1 < p < ∞, and every w ∈ Ap, there exists a constant depending on [w]Ap such that (1.2) k T f Lp,w(Rn)k ≤ c k f Lp,w(Rn)k Sw∈Ap0 (cf. [Rub82, Rub84] for the original results, or [CMP11, Duo11] for modern presentations). One fact which makes extrapolation theory so powerful is that one does not need any assumption on T besides having T well-defined on Lp0,w(Rn). (Actually one can extrapolate just inequalities between pairs of functions, not even an operator is needed.) Even if one is not interested in any weighted results this is very interesting by the fact that for example the weighted L2-boudednesses implies in particular unweighted boundednesses in the complete Lp scale. Appropriate assertions hold also for subclasses of Ap0, that is, one assumes (1.1) for a subclass (e.g. A1) and deduces (1.2) for appropriate subclasses (this applies to different Fourier multipliers, for instance). We generalize results of this theory to Morrey spaces and will get even in the classical Morrey spaces new results on the boundednesses of operators in the sense that we do not need any condition on T except that T satisfies (1.1) (for weights of Ap0 or distinguished subclasses) and is well-defined on There are lots of papers dealing with classes of operators in various Morrey type spaces which admit size estimates as Lp0,w(Rn). Sw∈Ap0 (1.3) (T f )(y) ≤ cZRn f (x) y − xn dx for all f ∈ D(Rn) and y /∈ supp (f ), where D(Rn) are the compactly sup- ported smooth functions (cf. [Pee66, CF87, Nak94, Alv96, DYZ98, Sam09, KS09, GAKS11, Gul12, Mus12, SFZ13, RS14, KGS14, PT15, IST15, Nkm16, Wan16] and the references given there). To obtain the boundedness results they are explicitly using and requiring different representation formulas as (1.3) of the considered operators. We are completely avoiding these assump- tions and encapsulating them through the consideration of weighted spaces. Hence, we get even new results on the classical Morrey spaces regarding Mihlin-Hormander type operators, rough operators, pseudodifferential op- erators, square functions, commutators, Fourier integral operators. . . Moreover, literature about (the nonseparable) Morrey spaces (starting from Peetre [Pee66] and many following scholars) does not care about how to extend the considered operators (satisfying a required representation for- mula only on nondense subsets of the considered Morrey spaces). Some forerunners dealing with the extension of the operators are [Alv96, AX12, RT13, Ros13, RT14, Tri14, RS14, Ada15, RS16]. Mainly, they apply dual- ity (with some exceptions as [Tri14, RS14] justifying (1.3) also for Morrey functions), which restricts the boundedness results to linear operators and BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 3 moreover the range of the integration parameter p (in particular, the exten- sion problem was not covered for p = 1 and the corresponding weak-type inequalities). We show different (continuous) embeddings of (weighted) Mor- rey spaces into different Muckenhoupt A1 weighted Lp spaces, which enables us to define the operators on the considered Morrey spaces by restriction. Another possibility to overcome the extension problem is using the Fatou property of the Morrey space. Nevertheless, this method requires at least a bigger space where the operator is also bounded to extend it by restriction (cf. Triebel [Tri16]), which is exactly the type of our embeddings. However, by the fact that we do not require representation formulas of operators it is even sufficient to show a not necessarily continuous embedding of a Morrey type space into a set of the type Sw∈Ap Lp,w(Rn) (where the considered operator is required to be bounded). This can be seen as a powerful model case to extend operators which are a priori only defined on D(Rn) -- as multipliers and (maximally truncated) singular integrals -- to the various generalizations of Morrey spaces given in the literature and to overcome the extension problem (by the fact that we avoid the requirement of justifying (1.3) also for Morrey functions). If one wants to deal with other spaces where D(Rn) is not dense and which are contained in the setSw∈Ap Lp,w(Rn) and where the maximal operator is bounded (or in their appropriate associated spaces), then there are even very good chances that the presented results could be carried over partially. A extrapolation result for Morrey spaces appears also in [RS16]. Its proof uses the boundedness of the Hardy-Littlewood maximal operator on predual Morrey spaces, which are also used to define the operator by duality. We avoid this in our approach and obtain more general results, valid also for end-points and for limited-range type assumptions. We also extrapolate from inequalities with Lp,w (quasi)-norms for 0 < p < ∞ and w ∈ A∞ to Morrey (quasi)-norms. In this way we get new Morrey versions of several interesting inequalities of this type appearing in the literature for pairs of operators. We consider only Muckenhoupt weights because the extrapolation tech- niques are adapted to them and because there are many operators for which the boundedness on Lebesgue spaces with (subclasses of) Muckenhoupt weights are known, so that we can immediately infer Morrey estimates from our general results. Nevertheless, a description of the admissible weighted Morrey estimates is not known even for the Hardy-Littlewood maximal op- p(λ, w, Rn) (see Defi- erator and, in fact, in the case of the Morrey spaces Lr nition 2.3) the class of admissible weights goes beyond the Muckenhoupt Ap class as shown in [Ta15]. A similar situation holds for the Hilbert transform ([Sam09]) and more generally for the Riesz transforms and other singular integrals ([NkSa17]). On the other hand, in these cases the problem of defin- ing the operators in the corresponding weighted spaces, which in our case is solved by the embeddings, should be considered. 4 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL After introducing the notation and some preliminaries in Section 2, Sec- tion 3 is concerned with continuous embeddings of Morrey type spaces into Muckenhoupt weighted A1 Lebesgue spaces. In Section 4 we present the ex- trapolation theory generalized to Morrey type spaces which will be applied to far-reaching boundedness results given in Section 5. 2. Notation, Morrey spaces and preliminaries A weight is a locally integrable nonnegative function. For 0 < p < ∞, Lp,w(Rn) is the complex quasi-Banach space of functions whose p-th power is integrable with respect to the weight w, with the quasinorm given by kf Lp,w(Rn)k =(cid:16)ZRn f (x)p w(x)dx(cid:17)1/p =(cid:16)ZRn f p w(cid:17)1/p . Moreover, we write w(M ) = RM w(x)dx for the measure of the subset M of Rn. We similarly define Lp,w(M ). If w(x) ≡ 1, we simply write Lp(M ), k · Lp(M )k and M . Furthermore, χM denotes the characteristic function of M . For any p ∈ (1, ∞) we denote by p′ the conjugate index, namely, 1/p+ 1/p′ = 1. If p = 1, then p′ = ∞. Moreover, L(Rn) collects all equivalence classes of almost everywhere coinciding measurable complex functions. A 1 (Rn), which collects all locally integrable functions, subset of L(Rn) is Lloc that is, functions in L1(M ) for any bounded measurable set M of Rn. We consider cubes whose sides are parallel to the coordinate axes. For such a cube Q, δQ stands for the concentric cube with side-length δ times of the side-length of Q. The concrete value of the constants may vary from one formula to the next, but remains the same within one chain of (in)equalities. Definition 2.1. Let w ∈ Lloc that w is a Muckenhoupt weight belonging to Ap for 1 < p < ∞ if 1 (Rn) with w > 0 almost everywhere. We say [w]Ap ≡ sup Q w(Q) Q w1−p′ Q !p−1 (Q) < ∞, where the supremum is taken over all cubes Q in Rn. The quantity [w]Ap is the Ap constant of w. We say that w belongs to A1 if, for any cube Q, w(Q) Q ≤ cw(x) for almost all x ∈ Q. The A1 constant of w, denoted by [w]A1 , is the smallest constant c for which the inequality holds. We say that w is in A∞ if w ∈ Ap for some p. Remark 2.2. The Ap weights are doubling. This means that if w ∈ Ap, there exists a constant c such that (2.1) w(2Q) ≤ cw(Q), for every cube Q in Rn (cf. [Duo01, (7.3)]). where the supremum is taken over all cubes Q in Rn. Moreover, with the quasinorm with the quasinorm p(w, Rn)(cid:13)(cid:13) < ∞} n(cid:17) kf Lp,w(Q)k , p(λ, w, Rn)(cid:13)(cid:13) < ∞} n(cid:17) kf Lp,w(Q)k . Lr Q Lr p + r −(cid:16) 1 w(Q) p(w, Rn) ≡ {f ∈ L(Rn) : (cid:13)(cid:13)f Lr (cid:13)(cid:13)f Lr p(w, Rn)(cid:13)(cid:13) ≡ sup p(λ, w, Rn) ≡ {f ∈ L(Rn) : (cid:13)(cid:13)f Lr (cid:13)(cid:13)f Lr p(λ, w, Rn)(cid:13)(cid:13) ≡ sup p(w, Rn)(cid:13)(cid:13) ≡ sup p(λ, w, Rn)(cid:13)(cid:13) ≡ sup sup t>0 sup t>0 w(Q) −(cid:16) 1 −(cid:16) 1 p + r −(cid:16) 1 p + r p + r Q Q Q Q Q (cid:13)(cid:13)f W Lr (cid:13)(cid:13)f W Lr BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 5 We recall that the Hardy-Littlewood maximal operator, which we denote by M , is bounded on Lp,w(Rn) for 1 < p < ∞ if and only if w ∈ Ap and is of weak-type (1, 1) with respect to the measure w(x)dx if and only if w ∈ A1. On the other hand, we will use the following construction of A1 weights: if M f (x) is finite a.e. and 0 ≤ δ < 1, then M f (x)δ is an A1 weight whose A1-constant depends only on δ. Moreover, the factorization theorem says that w ∈ Ap if and only if there exist w0, w1 ∈ A1 such that w = w0w1−p (cf. [Rub84] or [Gra09, Thm. 9.5.1]). 1 We define some Muckenhoupt weighted Morrey spaces. Definition 2.3. For 0 < p < ∞, − n p ≤ r < 0 and a weight w we define We also define their weak versions given by the quasinorms n(cid:17)tw({x ∈ Q : n(cid:17)tw({x ∈ Q : f (x) > t}) f (x) > t}) 1 p , 1 p . Remark 2.4. We observe that L−n/p If w ≡ 1, Lr space Lr spaces for r = −n/p. (λ, w, Rn) = Lp,w(Rn). p(λ, w, Rn) coincide with the unweighted Morrey p(Rn). The weak Morrey spaces coincide also with the weak Lp,w(Rn) (w, Rn) = L−n/p p(w, Rn) and Lr p p Definition 2.5. Let a nonnegative locally integrable function w on Rn be- long to the reverse Holder class RHσ for 1 < σ < ∞ if it satisfies the reverse Holder inequality with exponent σ, i.e. (cid:18) 1 QZQ w(x)σdx(cid:19) 1 σ c QZQ ≤ w(x)dx, where the constant c is universal for all cubes Q ⊂ Rn. The RHσ classes are decreasing, that is, RHσ ⊂ RHτ for 1 < τ < σ < ∞. On the other hand, Gehring's lemma ([Geh73]) says that if w ∈ RHσ, there exists ǫ > 0 such that w ∈ RHσ+ǫ (openness of the reverse Holder classes). 6 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Remark 2.6. Let w ∈ RHσ. For any cube Q and any measurable E ⊂ Q it holds that (2.2) w(E) w(Q) . ≤ c(cid:18) E Q(cid:19)1/σ′ (See [Duo01, Cor. 7.6]). Since w ∈ Ap implies that w ∈ RHσ for some σ (cf. [Duo01, Thm. 7.4]), the inequality holds for each Ap weight for the appropriate σ. Consequences of the reverse Holder inequalities for weights are that if w ∈ Ap, then there exist ǫ > 0 and s > 1 such that w ∈ Ap−ǫ (openness of Ap classes) and ws ∈ Ap. In the paper we use sometimes classes of the form Ap ∩ RHσ. There is a characterization of R. Johnson and C. Neugebauer for them ([JN91]): (2.3) Ap ∩ RHσ = {w : wσ ∈ Aσ(p−1)+1}. This is also valid for p = 1, that is, A1 ∩ RHσ = {w : wσ ∈ A1}, which can be easily proved from the definition of A1. Remark 2.7. Let w(x) = xα or w(x) = (1 + x)α. Then we have the following: (1) It holds that w ∈ Ap if and only if −n < α < n(p − 1) whenever p ∈ (1, ∞), and −n < α ≤ 0 whenever p = 1. (2) For α ∈ (−n, 0) it holds w ∈ RHσ if and only if 1 < σ < −n/α. (3) For α ≥ 0 it holds w ∈ RHσ for all 1 < σ < ∞. The first one is well known. The second one is easily obtained using (2.3) (for p = 1). The third one can be checked directly or derived from (2.3). 3. Embeddings of Muckenhoupt weighted Morrey spaces into Muckenhoupt weighted Lebesgue spaces p(w, Rn) and Lr In this section we establish different continuous embeddings between Mor- rey spaces of the types Lr p(λ, w, Rn) into some Muckenhoupt A1 weighted Lebesgue spaces which will give us the possibility to define operators on these Morrey spaces by restriction. Extension by continuity is usually not working in Morrey type spaces by their nonseparability and also by the fact that the smooth functions are not dense in these spaces (see [RT14, Prop. 3.7] for Lr p(w, Rn) for p(Rn) and [RS16, Prop. 3.2] for Lr the nonseparability, and see [Pic69, p. 22] for the nondensity). In recent literature this difficulty has been treated using duality for linear operators, whenever the integration parameter p satisfies 1 < p < ∞. Dealing with the extension problem through embeddings one is neither restricted in the parameter range of p (in particular, p = 1 is admissible) nor the operator has to be linear. Proposition 3.1. Let 1 ≤ γ < p < ∞, −n/p ≤ r < 0 and w ∈ Ap/γ. Then there exists q0 ∈ (γ, p) such that for each q ∈ [γ, q0], 1 < s ≤ s0(q), each BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 7 cube Q, and h ∈ L(p/q)′,w(Q) with norm 1, the continuous embedding (3.1) p(w, Rn) ֒→ Lq,M (hswsχQ)1/s(Rn) Lr holds with constant independent of h and depending on Q as w(Q)1/p+r/n. Furthermore, with the additional assumptions w ∈ RHσ and r ≤ −n/(pσ), it also holds that (3.2) p(λ, w, Rn) ֒→ Lq,M (hswsχQ)1/s(Rn) Lr with constant independent of h and depending on Q as Q1/p+r/n. As a consequence, under the same conditions on p, r and w, for q ∈ [γ, q0] there is 0 < α < n such that Lr p(w, Rn), Lr (3.3) p(λ, w, Rn) ֒→ Lq,(1+x)−α(Rn). Proof. We prove the result for γ = 1. The case γ > 1 will follow as a consequence. Choose q0 ∈ (1, p) such that w ∈ Ap/q0. For q ∈ [1, q0], w ∈ Ap/q. Fix one such q and set p = p/q. Let 1 < s < p′ such that w1−p′ ∈ Ap′/s. For a function h with khLp′,w(Q)k = 1, the weight M (hswsχQ)1/s is well defined because hswsχQ ∈ L1(Rn), which implies M (hswsχQ) < ∞ almost everywhere (by the fact that M maps L1(Rn) to the weak L1(Rn) space). We check that hswsχQ ∈ L1(Rn) and obtain an estimate that will be needed below. Indeed, (cid:18)ZQ hsws−1w(cid:19) 1 s (3.4) ≤(cid:13)(cid:13) h Lp′,w(Rn)(cid:13)(cid:13)(cid:18)ZQ (Q)− 1 s w1−p′ ≤ c Q 1 w(s−1) p′ s − 1 p′ p′−s +1(cid:19) 1 p′ ≤ c w(Q) 1 p Q− 1 s′ , where the second inequality holds because w1−p′ ∈ Ap′/s (the exponent of w in the integral is the same as (1 − p′)(1 − (p′/s)′)) and in the last one we use Q ≤ w(Q) 1 p w1−p′ (Q) 1 p′ . Let f ∈ Lr p(w, Rn) and assume that f is nonnegative. To estimate the norm of f in Lq,M (hswsχQ)1/s (Rn) we decompose the integral into two parts: first we integrate over 2Q and next over Rn \ 2Q. Over 2Q we use Holder's inequality, (cid:18)Z2Q f qM (hswsχQ) 1 q s(cid:19) 1 ≤(cid:18)Z2Q f pw(cid:19) 1 p(cid:18)Z2Q M (hswsχQ) The first term of the right-hand side is bounded by w(2Q) and using (2.1) we can replace w(2Q) by w(Q). In the second one we use that w1−p′ ∈ Ap′/s and the boundedness of M on Lp′/s,w1− p′ (Rn) to get a constant times the norm of h in Lp′,w(Rn), which is 1. . p′ q p′ s w1−p′(cid:19) 1 n(cid:13)(cid:13) f Lr p(w, Rn)(cid:13)(cid:13) 1 p + r 1 p p 1 p 1 s′ ≤ c2 ≤ c2 f qw w(Q) 2iQ 1 s Q p w− 1 ∞Xi=1Z2i+1Q\2iQ ∞Xi=1 Z2i+1Q\2iQ f pw! q w1−p′(cid:0)2i+1Q(cid:1) 1 ∞Xi=1 p(w, Rn)(cid:13)(cid:13)q ≤ c3(cid:13)(cid:13) f Lr w(cid:0)2i+1Q(cid:1)(cid:16) 1 p w1−p′(cid:0)2i+1Q(cid:1) 1 w(cid:0)2i+1Q(cid:1) 1 ∞Xi=1 w(cid:0)2i+1Q(cid:1) r w(Q) q p 2 p + r n(cid:17)q n q and it is sufficient to show (3.6) (3.7) Since w ∈ Ap, then p′ ≤ c2i+1Q, in s′ ≤ c w (Q)(cid:16) 1 p + r n(cid:17)q . p′ w(Q) 1 p 2in/sQ w1−p′(cid:0)2i+1Q(cid:1) 1 p′ w(Q) 2in/sQ 1 p . 8 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL To handle the integral over Rn \ 2Q, we use that M (hswsχQ)(x) for x ∈ 2i+1Q \ 2iQ (i ≥ 1) is comparable to the average of hswsχQ over 2i+1Q. Then we obtain (3.5) ZRn\2Q f qM (hswsχQ) 1 s ≤ c1 ∞Xi=1Z2i+1Q\2iQ 2iQ ! 1 f q RQ hsws s Taking into account (2.2), there exists δ > 0 such that w(2i+1Q) r n q ≤ c 2iδrqw(Q) r n q. Therefore, (3.1) is satisfied for s sufficiently close to 1. In the case of the spaces Lr p(λ, w, Rn), the part corresponding to the integral over 2Q is obtained in a similar way using (cid:18)Z2Q p f pw(cid:19) 1 ≤ cQ 1 p + r n(cid:13)(cid:13) f Lr p(λ, w, Rn)(cid:13)(cid:13) . For the integral over Rn \ 2Q, we first modify the last line of (3.5) and then use (3.6) to get f qM (hswsχQ) 1 s ZRn\2Q ≤ c1(cid:13)(cid:13) f Lr ≤ c2(cid:13)(cid:13) f Lr p(λ, w, Rn)(cid:13)(cid:13)q p(λ, w, Rn)(cid:13)(cid:13)q ∞Xi=1(cid:12)(cid:12)2i+1Q(cid:12)(cid:12)(cid:16) 1 ∞Xi=1(cid:12)(cid:12)2i+1Q(cid:12)(cid:12)(cid:16) 1 p + r n(cid:17)q p + r n(cid:17)q 1 p 2in/sQ p′ w(Q) w1−p′(cid:0)2i+1Q(cid:1) 1 s′ (cid:2)w(2i+1Q)−1 w(Q)(cid:3) q 2 in p . BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 9 By the latter we obtain (3.2) showing that (3.8) 2inq[ 1 p + r n + 1 s′q ∞Xi=1 p ≤ c. ](cid:2)w(2i+1Q)−1 w(Q)(cid:3) q ≤ c(cid:18) Q 2iQ(cid:19)1/σ′ = c2−i n σ′ . By the fact that w ∈ RHσ, from (2.2) we deduce (3.9) w(Q) w(2iQ) Inserting (3.9) into the left hand-side of (3.8) and taking into account that s is as close to 1 as desired, the geometric series is convergent for r/n + 1/(pσ) < 0. The endpoint r = −n/(pσ) of the statement is attained from the openness property of the reverse Holder classes. Let now 1 < γ < p and f ∈ Lr p(w, Rn) with w ∈ Ap/γ. Since 1 γ , (3.10) f γ ∈ Lrγ the proof. It gives the existence of q∗ q∗ ∈ (1, q∗ p/γ(w, Rn) (with w ∈ Ap/γ) and we can apply the first part of 0 ∈ (1, p/γ) and s > 1 such that for p/γ(w, Rn)(cid:13)(cid:13)(cid:13) p(w, Rn)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) f γ Lrγ (cid:13)(cid:13) f Lr p/γ(w, Rn)(cid:13)(cid:13)(cid:13) ≤ c(cid:13)(cid:13)(cid:13) f γ Lq∗,M (hswsχQ)1/s (Rn)(cid:13)(cid:13)(cid:13) , (cid:13)(cid:13)(cid:13) f γ Lrγ for any cube Q and h ∈ L(p/(γq∗))′ ,w(Q) with norm 1. Define q0 = q∗ q = q∗γ and (3.1) is proved. The case f ∈ Lr treated similarly. 0γ and p(λ, w, Rn) with w ∈ Ap/γ is 0], To obtain (3.3), let Q be the unit cube and h = cw−1, with c chosen such that h has unit norm. Use now that M (χQ)(x) is bounded below by a positive constant if x ∈ 2Q and by cx−n if x /∈ 2Q. Take α = n/s to conclude. (cid:3) Remark 3.2. Since w ∈ Ap implies that w ∈ RHσ for some σ > 1, the range of values of r satisfying the embedding (3.3) for Lr p(λ, w, Rn) is not empty. In the case of the weights xβ and (1 + x)β such range is (− n − n p ≤ r < 0, p ≤ r < β p , if 0 ≤ β < n( p if − n < β < 0. γ − 1); The following proposition shows that the union of the Lq,w(Rn) spaces for fixed q and all w ∈ Aq is invariant for 1 < q < ∞. This result appears in [KMM16] in a more general context. We give here a direct proof. Proposition 3.3. Let be 1 ≤ γ < q < ∞. Then it holds that (3.11) [w∈Aq/γ Lq,w(Rn) = [γ<p<∞ [w∈Ap/γ Lp,w(Rn) ( [w∈A1 Lγ,w(Rn). Hence, the left-hand side is independent of q. 10 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Proof. Again it is enough to prove the case γ = 1. Once this is proved, the general case is deduced from the observation that f ∈ Lp,w(Rn) with w ∈ Ap/γ is equivalent to f γ ∈ Lp/γ,w(Rn) (with w ∈ Ap/γ). Let γ = 1. We start showing the equality in (3.11). We first prove that if f is in Lp,w(Rn) for 1 < p < q and w ∈ Ap, then f ∈Sv∈Aq Lq,v(Rn). For s > 1 sufficient small w ∈ Ap/s holds. Moreover, we have f s ∈ Lp/s,w and therefore M (f s) ∈ Lp/s,w. Hence (M (f s))1/s ∈ A1. Thus, (M (f s))(p−q)/sw ∈ Aq (see [Duo11, Lemma 2.1]). By reason of p < q finally it holds that f qf p−qw and f ∈Sv∈Aq Lq,v(Rn). That means that we have up to now shown Lp,w(Rn). (3.12) f q(M (f s)) ZRn [w∈Aq 1 s (p−q)w ≤ZRn Lq,w(Rn) = [1<p≤q [w∈Ap We assume now f ∈ Lp,w(Rn) for some p > q and w ∈ Ap. By (3.3) (applied to r = −n/p, since ) we obtain f ∈ Lq,(1+x)−α(Rn) for some q < q and so f is in the right-hand side of (3.12). Here, q = 1 is also admitted. It remains to prove that the inclusion of (3.11) is strict. For this it is enough to notice that there exists f ∈ L1(Rn) for which M f is not locally Indeed, such an f cannot be in Lp,w(Rn) for any p ∈ (1, ∞) integrable. because this would imply M f ∈ Lp,w(Rn) ⊂ Lloc 1 (Rn). A typical example of a function as the one required here is f (x) = x−n(ln x)−2 χ{x≤1/2}. (cid:3) The following embedding result on the considered Morrey spaces is a generalization of [Tri14, Prop. 2.10] to weighted situations. Proposition 3.4. Let 1 ≤ p < ∞ and ν ≥ 1. Let w ∈ A1 ∩ RHσ (that is, wσ ∈ A1) for some σ > ν. Then for r ∈ [−n/p, −nσ′/(pν′)] (if ν = 1, the endpoint r = 0 is excluded), there exists s0 > ν such that for ν < s ≤ s0 and for any cube Q, the continuous embedding (3.13) p(w, Rn) ֒→ Lp,M (wsχQ)1/s (Rn) Lr holds with a constant depending on Q as w(Q)1/p+r/n. Moreover, whenever r ∈ [−n/p, −n(1/σ + 1/ν′)/p], it also holds that (3.14) p(λ, w, Rn) ֒→ Lp,M (wsχQ)1/s(Rn), Lr with a constant depending on Q as Q1/p+r/n. As a consequence, under the same conditions on w and r, there exists some 0 < α < n/ν such that p(w, Rn), Lr Lr (3.15) p(λ, w, Rn) ֒→ Lp,(1+x)−α(Rn). Proof. Let s be such that ν < s < σ. The weight M (wsχQ)1/s is in A1 and satisfies M (wsχQ)1/s ≤ M (ws)1/s ≤ Cw a.e. because ws ∈ A1. Moreover, for x ∈ 2i+1Q \ 2iQ, M (wsχQ)(x) is comparable to the average of wsχQ over 2i+1Q. BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 11 Let f ∈ Lr p(w, Rn) with r as stated and assume that f is nonnegative. Then ZRn f pM (wsχQ) 1 s ≤ c1Z2Q f pw + c2 ∞Xi=1Z2i+1Q\2iQ f p(cid:18)ws(Q) 2iQ (cid:19) 1 s . The first term gives immediately the desired estimate. As for the second one, we notice that (cid:18) ws(Q) Q (cid:19)1/s ≤ C w(Q) Q , because w ∈ RHσ ⊂ RHs and also that w ∈ A1 implies w(2i+1Q) 2i+1Q ≤ Cw(x) a.e. x ∈ 2i+1Q. Inserting these in the above expression we obtain (3.16) ∞Xi=1Z2i+1Q\2iQ ≤ c1(cid:13)(cid:13) f Lr ≤ c2(cid:13)(cid:13) f Lr f pw 2i+1Q w(2i+1Q) w(Q) 2in/sQ ∞Xi=1 p(w, Rn)(cid:13)(cid:13)p p(w, Rn)(cid:13)(cid:13)p w(Q)1+ rp n ∞Xi=1 w(2i+1Q) rp n w(Q)2i n s′ 2i rp σ′ 2i n s′ , If rp < −nσ′/ν′, we can choose s > ν using that w ∈ RHσ and (3.9). sufficiently small such that the geometric series converges. To reach the endpoint rp = −nσ′/ν′, use the openness of the reverse Holder classes. In the case of the spaces Lr p(λ, w, Rn), the left-hand side of (3.16) is bounded by C(cid:13)(cid:13) f Lr p(λ, w, Rn)(cid:13)(cid:13)p ∞Xi=1 (2inQ)1+ rp n 2 in s′ w(Q) w(2i+1Q) . Using (3.9) the geometric series converges for rp + n(1/s′ + 1/σ) < 0. This gives (3.14) in the stated range of r (using again the openness of the reverse Holder classes to get the endpoint). For the particular embeddings (3.15) take as Q the unit cube centered at (cid:3) the origin and α = n/s. Remark 3.5. Since we assume wν ∈ A1, we automatically have w ∈ RHν. But each such w will be in a reverse Holder class w ∈ RHσ for some σ > ν and the range of values of r obtained in the proposition is not empty. In the case w(x) = xβ or w(x) = (1 + x)β, we need −n < βν ≤ 0 to fulfill the condition wν ∈ A1 and in such case the range of values of r satisfying the two embeddings of (3.15) are respectively − n p ≤ r < − n p n n + β 1 ν′ and − n p ≤ r < β p − n p 1 ν′ . 12 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Corollary 3.6. For every 1 ≤ γ ≤ p0 < ∞, every γ < p < ∞ (and also p = p0 if γ = p0), −n/p ≤ r < 0 and w ∈ Ap/γ, it holds that (3.17) Lp0,u(Rn). Lr p(w, Rn) ⊂ [u∈Ap0/γ p(λ, w, Rn) ⊂ [u∈Ap0/γ If moreover w ∈ Ap/γ ∩ RHσ, for every −n/p ≤ r < −n/(pσ) it holds that (3.18) Lr Lp0,u(Rn). Proof. If 1 ≤ γ < p0 we apply the embedding for Lr p(w, Rn) given by (3.1) of Proposition 3.1 together with Proposition 3.3. If γ = p0 and p = p0 = γ, (3.17) follows from Proposition 3.4 with ν = 1. Finally, if γ = p0 and p0 = γ < p < ∞ we use the embedding for Lr p(w, Rn) of Proposition 3.1 with q = γ. Analogously, we obtain (3.18). (cid:3) Remark 3.7. Note that as a consequence of this corollary, the left-hand side of (3.11) in Proposition 3.3 also coincides with the (formally larger) union [γ<p<∞ [−n/p≤r<0 [w∈Ap/γ p(w, Rn). Lr (We recall that L−n/p p (w, Rn) = Lp,w(Rn).) Proposition 3.8. Let 1 ≤ γ < b < ∞ and σ(p) = (b/p)′ for γ < p < b. Let w ∈ Ap/γ ∩ RHσ for σ > σ(p). Let r ∈ [−n/p, −nσ′/b]. Then there exist q0 ∈ (γ, p) and for each q ∈ [γ, q0], some s0(q) > σ(q), such that for σ(q) < s ≤ s0(q), each cube Q and h ∈ L(p/q)′,w(Q) with norm 1, the continuous embedding (3.19) p(w, Rn) ֒→ Lq,M (hswsχQ)1/s(Rn) Lr holds with constant independent of h and depending on Q as w(Q)1/p+r/n. Furthermore, with the additional assumption r ∈ [−n/p, −n/(pσ) − n/b], it also holds that (3.20) p(λ, w, Rn) ֒→ Lq,M (hswsχQ)1/s(Rn) Lr with constant independent of h and depending on Q as Q1/p+r/n. As a consequence, under the same conditions on p, r and w, there exists 0 < α < n/σ(q) such that (3.21) Lr p(w, Rn), Lr p(λ, w, Rn) ֒→ Lq,(1+x)−α(Rn). Proof. We can consider γ = 1. Once this is proved, the case γ > 1 is obtained from it by considering f γ as in Proposition 3.1 (see (3.10)). Let f ∈ Lr p(w, Rn) nonnegative with w ∈ Ap ∩ RHσ. Choose q ∈ (1, p) such that w ∈ Ap/q and set p = p/q. By (2.3), w ∈ Ap ∩ RHσ implies wσ ∈ Aσ(p−1)+1, which in turn implies w1−p′ ∈ A1+(p′−1)/σ by the duality BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 13 property of the weights. If 1 + (p′ − 1)/σ ≤ p′/s, then w1−p′ is possible with s > σ(q) if ∈ Ap′/s. This which is the same as 1 + p′ − 1 σ < p′ σ(q) = p′ (b/q)′ , 1 p′ + 1 pσ < 1 (b/q)′ = 1 − q b , that is, σ > σ(p). The proof continues as for Proposition 3.1. In particular, the estimate (3.4) remains true because w1−p′ ∈ Ap′/s, so that we will get (3.19) if (3.7) holds. Since w ∈ RHσ, we use (3.9) and the left-hand side of (3.7) is bounded by c w (Q)(cid:16) 1 p + r n(cid:17)q 2i( rq σ′ + n s′ ), ∞Xi=1 The geometric series converges for r < −nσ′/(qs′). It is possible to have s > σ(q), if r < −nσ′/b. For the endpoint r = −nσ′/b, use the openness property of the reverse Holder classes. In the case of the spaces Lr p(λ, w, Rn), to get (3.20) we follow again the proof of Proposition 3.1 and we are left with (3.8). We insert (3.9) and the geometric series converges if nq p + r + n s′ − nq pσ′ = nq pσ + r + n s′ < 0. We can get s > σ(q) if r < − nq pσ − n σ(q)′ = − nq pσ − nq b . This holds for q close to 1 if r < −n/(pσ) − n/b. Take as Q the unit cube and as h = cw−1 to get the embeddings of (3.21) (cid:3) with α = n/s. Proposition 3.9. Let 1 ≤ γ < b < ∞ and σ(p) = (b/p)′ for γ < p < b. Let γ < q < b. Then [w∈Aq/γ∩RHσ(q) Lq,w(Rn) = [γ<p<b [w∈Ap/γ∩RHσ(p) Lp,w(Rn). Proof. We assume γ = 1 as in the previous proposition. We notice first that using the factorization theorem of Ap weights we have 1− p Ap ∩ RHσ(p) = {w : wσ(p) ∈ Aσ(p)(p−1)+1} = {u b 0 u1−p 1 : u0, u1 ∈ A1}. Let f ∈ Lp,w(Rn) with w = u1−p/b 1 < p < q < b. We have f u−1/b 0 u1−p 1 0 ∈ Lp(u0u1−p 1 for some u0, u1 ∈ A1, and let ∈ Ap, there ). Since u0u1−p 1 14 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL exists s > 1 such that u0u1−p Hence M ((f u−1/b )s)1/s ∈ A1 and 1 0 ∈ Ap/s. Then (f u−1/b 0 )s ∈ Lp/s(u0u1−p 1 ). − 1 b f (x) ≤ M ((f u 0 )s) 1 s (x)u 1 b 0 (x) a.e. We deduce that ZRn The weight of the left side is − 1 f q[M ((f u b 0 )s) 1 s u 1 b 0 ]p−qu 1− p b 0 u1−p 1 ≤ZRn 1− p f pu b 0 u1−p 1 . 1− q b u 0 − 1 b [M ((f u 0 )s) 1 s ]p−qu1−p 1 = u 1− q b 0 [vθu1−θ 1 ]1−q, where v = M ((f u−1/b Aq ∩ RHσ(q), since vθu1−θ 0 )s)1/s and θ = (q − p)/(q − 1) ∈ (0, 1). Thus, it is in 1 ∈ A1. This proves that [1<p≤q [w∈Ap∩RHσ(p) Lp,w(Rn) = Lq,w(Rn). [w∈Aq∩RHσ(q) This is enough to prove the statement because for p > q we know from Proposition 3.8 (taking r = −n/p) that a function in Lp,w(Rn) is in some Lq,(1+x)−α(Rn) with q near 1 and the weight (1 + x)−α in A1 ∩ RHσ(q) ⊂ Aq ∩ RHσ(q). (cid:3) Our embeddings directly apply to similar results with respect to Ap- weighted weak Lp(Rn) spaces (W Lp,w(Rn) is the space W Lr p(w, Rn) of Defi- nition 2.3 for r = −n/p), which are nonseparable. These kind of embeddings allow one to extend operators to a domain which is a weak Lp space via re- striction. The connection with the Morrey spaces Lr p(w, Rn) lies in the use of a different norm for W Lp,w(Rn). For 1 < p < ∞ the space W Lp,w(Rn) can be normed with the norm (3.22) kf W Lp,w(Rn)k∗ ≡ sup M w(M )(cid:16) 1 p − 1 p(cid:17) kf Lp,w(M )k , where 0 < p < p and the supremum is taken over all measurable subsets of Rn with 0 < w(M ) < ∞. This norm is equivalent to the usual quasi-norm given in Definition 2.3 ([Gra08, Exercise 1.1.12]). It is obvious from (3.22) that W Lp,w(Rn) ֒→ Lr p(w, Rn), with r = −n/p > −n/p. Moreover, if w ∈ Ap/γ for some γ ∈ [1, p), p can be chosen such that w ∈ Ap/γ, and Proposition 3.1 can be used to deduce that there exists q0 > γ such that for each q ∈ [γ, q0] there is 0 < α < n for which W Lp,w(Rn) ֒→ Lq,(1+x)−α(Rn). The same type of embeddings can be obtained from Propositions 3.4 and 3.8. BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 15 4. Boundedness on Morrey spaces: general results In this section we obtain the general theorems giving the boundedness on the corresponding Morrey spaces from the assumption of weighted estimates on Lp,w(Rn) spaces. Instead of using operators, we can establish the general theorems in terms of pairs of functions, as it is usual in the recent presen- tations of the extrapolation theorems. This has the advantage of providing immediately several different versions. Theorem 4.1. Let 1 ≤ p0 < ∞ and let F be a collection of nonnegative measurable pairs of functions. Assume that for every (f, g) ∈ F and every w ∈ Ap0 we have (4.1) k g Lp0,w(Rn)k ≤ c1 k f Lp0,w(Rn)k , where c1 does not depend on the pair (f, g) and it depends on w only in terms of [w]Ap0 . Then for every 1 < p < ∞ (and also for p = 1 if p0 = 1), every −n/p ≤ r < 0 and every w ∈ Ap we have Furthermore, for every 1 < p < ∞ (and also for p = 1 if p0 = 1) and w ∈ Ap ∩ RHσ, if −n/p ≤ r ≤ −n/(pσ) we have (4.2) (4.3) p(w, Rn)(cid:13)(cid:13) ≤ c2(cid:13)(cid:13) f Lr (cid:13)(cid:13) g Lr p(λ, w, Rn)(cid:13)(cid:13) ≤ c3(cid:13)(cid:13) f Lr (cid:13)(cid:13) g Lr p(w, Rn)(cid:13)(cid:13) . p(λ, w, Rn)(cid:13)(cid:13) . The constants c2 and c3 in (4.2) and (4.3) do not depend on the pair (f, g) but may depend on w and the involved parameters. If (4.1) is assumed to hold only when the left-hand side is finite, (4.2) and (4.3) still hold assuming that their left-hand side is finite. Remark 4.2. In (4.1) we do not need to require that f and g are in Lp0,w(Rn) for all w ∈ Ap0. In the first part of the statement we assume that (4.1) holds whenever the right-hand side is finite (and in such case, this forces the left-hand side to be finite). Then the conclusion of the theorem is that for every f in Lr p(λ, w, Rn), g is in the same space and the norm inequality is satisfied. When g = T f for some operator T this means that T f is in the same Morrey space as f is, for all the functions in the space. p(w, Rn) or in Lr In the last part of the statement we only assume that (4.1) holds when the left-hand side is finite. That is, it could happen that for a particular weight and a particular pair of functions the left-hand side is infinity and the right-hand side is finite. Under this less restrictive hypothesis, the fact that f is in Lr p(λ, w, Rn) does not imply the same thing for g, but if g also is assumed to be in the corresponding Morrey space, then the norm inequality holds. p(w, Rn) or in Lr In either case the extrapolation theorem in the usual weighted Lebesgue spaces theorem says that if (4.1) holds for a particular p0, it holds for 1 < p < ∞ and Ap weights with the same interpretation of the inequalities. The proof given in [Duo11], for instance, is only valid in the first situation, although 16 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL this is not explicitely indicated in the paper. The proofs in [CMP11] cover the second case (left-hand side finite), and are valid also for the first case. In the applications to operators, the advantage of using the first version is that one proves directly the result for the whole weighted Lebesgue space, without making the extension from a dense subclass. The part of the theorem corresponding to (4.2) in the case p > 1 was proved in [RS16] using a different method, which involves the predual Morrey spaces. Proof. (a) The case p > 1. From the extrapolation theorem we can assume (4.1) for any p ∈ (1, ∞) and the weight class Ap. Let w ∈ Ap. Choose q ∈ (1, p) and s > 1 such that the embedding (3.1) holds. Set p = p/q. Let Q be a fixed cube. Since we fix such a function h and we have q gqhw(cid:19) 1 , (cid:18)ZQ = p gpw(cid:19) 1 (cid:18)ZRn pq g pqw(cid:19) 1 =(cid:18)ZQ gqhwχQ(cid:19) 1 q sup h : khL p′,w(Q)k=1(cid:18)ZQ s(cid:19) 1 s(cid:19) 1 f qM (hswsχQ) gqM (hswsχQ) 1 1 q q , ≤(cid:18)ZRn ≤ c(cid:18)ZRn (4.4) where the second inequality holds because M (hswsχQ)1/s ∈ A1 ⊂ Aq. We checked in the proof of Proposition 3.1 that the weight is well defined. Note that the constant in (4.4) is independent of Q and h because the A1-constant of M (hswsχQ)1/s, which is an upper bound of the Aq-constant, depends only on s (actually, it behaves as (s − 1)−1; see [Gra09, Theorem 9.2.7]). We are now in the situation of the embedding (3.1) and the proof of (4.2) for p > 1 is complete. In the case of the spaces Lr (b) The case p = 1. We assume now that (4.1) holds for p0 = 1 and p(λ, w, Rn), we use the embedding (3.2). w ∈ A1 and prove the theorem for p = 1 and w ∈ A1. We want to apply the embeddings of Proposition 3.4 with ν = 1. Choose s > 1 such that (3.13) and (3.14) hold. Fix a cube Q ⊂ Rn. The weight M (wsχQ)1/s is in A1. Then we have ZQ gw ≤ZQ gM (wsχQ) f M (wsχQ) 1 s . 1 s ≤ c1ZRn Thus, (4.1) and (4.2) follow from (3.13) and (3.14). (c) If the assumption (4.1) holds only when the left-hand side is finite, the proof is still valid assuming that the left-hand side of (4.2) or (4.3) is p(λ, w, Rn), the embeddings finite. Indeed, since g is in Lr (3.1) and (3.2) say that the second term in (4.4) is finite and the proof of p(w, Rn) or in Lr BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 17 part (a) goes on. The same thing is true for part (b) of the proof using the embeddings of Proposition 3.4. (cid:3) There are weighted inequalities between pairs of operators with Lp,w- quasinorms for 0 < p < ∞ and w ∈ A∞. Such inequalities hold when the left-hand side is finite. We can get similar type of estimates with Morrey quasinorms. Corollary 4.3. Let 0 < p0 < ∞ and let F be a collection of nonnegative measurable pairs of functions. Assume that for every (f, g) ∈ F and every w ∈ A∞ we have (4.5) k g Lp0,w(Rn)k ≤ c1 k f Lp0,w(Rn)k , whenever the left-hand side is finite. Then for every 0 < p < ∞, every −n/p ≤ r < 0 and every w ∈ A∞ we have (4.6) (4.7) whenever the left-hand side is finite. Furthermore, if w ∈ RHσ and −n/p ≤ r ≤ −n/(pσ) we have (cid:13)(cid:13) g Lr p(w, Rn)(cid:13)(cid:13) ≤ c2(cid:13)(cid:13) f Lr p(λ, w, Rn)(cid:13)(cid:13) ≤ c3(cid:13)(cid:13) f Lr (cid:13)(cid:13) g Lr p(w, Rn)(cid:13)(cid:13) , p(λ, w, Rn)(cid:13)(cid:13) , whenever the left-hand side is finite. Proof. By extrapolation in Lp,w spaces we can assume that (4.5) holds for all p0 ∈ (0, ∞) (see [CMP04]). Let 0 < p < ∞. Given w ∈ A∞, take q ∈ [1, ∞) such that w ∈ Aq. If p ≥ q, since Aq ⊂ Ap, the result follows directly from Theorem 4.1. If p < q, we have for u ∈ A1, whenever the left-hand side is finite, and we apply Theorem 4.1 to obtain (cid:13)(cid:13)(cid:13) gp/q(cid:12)(cid:12)(cid:12) L1,u(Rn)(cid:13)(cid:13)(cid:13) ≤ c(cid:13)(cid:13)(cid:13) f p/q(cid:12)(cid:12)(cid:12) L1,u(Rn)(cid:13)(cid:13)(cid:13) , (w, Rn)(cid:13)(cid:13)(cid:13) , (w, Rn)(cid:13)(cid:13)(cid:13) ≤ c2(cid:13)(cid:13)(cid:13) f p/q(cid:12)(cid:12)(cid:12) Lrp/q (cid:13)(cid:13)(cid:13) gp/q(cid:12)(cid:12)(cid:12) Lrp/q q q which is the desired result using the scaling (3.10) with γ = p/q. The proof of (4.7) is similar. (cid:3) fromSw∈Ap0/γ Corollary 4.4. Let 1 ≤ γ ≤ p0 < ∞. Assume that T is an operator acting Lp0,w(Rn) into the space of measurable functions satisfying k T f Lp0,w(Rn)k ≤ c1 k f Lp0,w(Rn)k (4.8) for all f ∈ Lp0,w(Rn) and w ∈ Ap0/γ, with a constant depending on [w]Ap0 /γ. Then for every γ < p < ∞ (and also p = γ if p0 = γ), every −n/p ≤ r < 0 p(w, Rn) by restric- and every w ∈ Ap/γ, we have that T is well-defined on Lr tion and, moreover, (4.9) for all f ∈ Lr (cid:13)(cid:13) T f Lr p(w, Rn)(cid:13)(cid:13) ≤ c2(cid:13)(cid:13) f Lr p(w, Rn)(cid:13)(cid:13) , p(w, Rn). 18 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Furthermore, for p as before and every w ∈ Ap/γ ∩ RHσ, if −n/p ≤ r ≤ p(λ, w, Rn) by restriction and, −n/(pσ) we have that T is well-defined on Lr moreover, (4.10) (cid:13)(cid:13) T f Lr p(λ, w, Rn)(cid:13)(cid:13) ≤ c3(cid:13)(cid:13) f Lr p(λ, w, Rn)(cid:13)(cid:13) for all f ∈ Lr p(λ, w, Rn). The constants c2 and c3 in (4.9) and (4.10) do not depend on f but may depend on the weight w and the involved parameters. If T satisfies instead of (4.8) the weak-type assumption k T f W Lp0,w(Rn)k ≤ c1 k f Lp0,w(Rn)k the estimates (4.9) and (4.10) are replaced by their weak-type counterparts, that is, (cid:13)(cid:13) T f W Lr p(w, Rn)(cid:13)(cid:13) and (cid:13)(cid:13) T f W Lr p(λ, w, Rn)(cid:13)(cid:13) appear at the left- hand side of the inequalities. Proof. The operator is well defined by restriction as a consequence of the embeddings of Corollary 3.6. To obtain the estimates (4.9) and (4.10) it is enough to apply Theorem 4.1 to the pairs (f γ, T f γ) with Lp0,w(Rn) and to use the scaling property of the norms in p(w, Rn) and Lr To deal with the weak-type inequalities we apply Theorem 4.1 to the pairs (cid:3) p(λ, w, Rn) (see (3.10)). (f γ, tγχ{T f >t}). f ∈ Sw∈Ap0/γ Lr Remark 4.5. That the operator is well-defined, for instance, on the set ditional condition. Usually one applies the theorem to continuous opera- Sw∈Ap Lp,w(Rn) for some 1 < p < ∞ is for practical reasons not an ad- tors which are defined on an common dense subset ofTw∈Ap Lp,w(Rn), for example singular integrals defined on D(Rn). The singular integrals can be extended to each Lp,w(Rn) by continuity. Such extensions coincide for functions in the intersection of two weighted Lebesgue spaces, because for each such function one can find a sequence in D(Rn) converging to it in both spaces. This implies the well-definedness of the singular integrals on S1<p<∞,w∈Ap Lp,w(Rn). Concerning the spaces Lr p(λ, w, Rn) it arises the question about the ne- cessity of the additional weight condition, that is, whether for fixed r being in some reverse Holder class is necessary. It is known that our result could not hold for all Ap weights (cf. [Sam09]). Remark 4.6. Chiarenza and Frasca in [CF87] noticed that weighted A1 inequalities for singular integrals yields boundedness on unweighted Morrey spaces. With a different approach Adams and Xiao [AX12] revisited this method using weighted inequalities to obtain boundedness in the unweighted Morrey spaces. Corollary 4.7. Assume that we have a sequence of operators {Tj} satisfy- ing the assumptions of Corollary 4.4, and that the constant c1 in (4.8) is independent of j. Then for γ < q, p < ∞ and the same conditions on w and BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 19 r, we have the vector-valued estimates and To prove these vector-valued inequalities in Morrey spaces it is enough to Tjfjq(cid:1)1/q(cid:12)(cid:12)(cid:12)(cid:12)Lr (cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xj fjq(cid:1)1/q(cid:12)(cid:12)(cid:12)(cid:12)Lr Tjfjq(cid:1)1/q(cid:12)(cid:12)(cid:12)(cid:12)Lr (cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xj fjq(cid:1)1/q(cid:12)(cid:12)(cid:12)(cid:12)Lr apply Theorem 4.1 to pairs(cid:16)(cid:0)Pj fjq(cid:1)1/q,(cid:0)Pj Tjfjq(cid:1)1/q(cid:17). p(w, Rn)(cid:13)(cid:13)(cid:13)(cid:13) ≤ c4(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xj p(λ, w, Rn)(cid:13)(cid:13)(cid:13)(cid:13) ≤ c5(cid:13)(cid:13)(cid:13)(cid:13)(cid:0)Xj p(w, Rn)(cid:13)(cid:13)(cid:13)(cid:13), p(λ, w, Rn)(cid:13)(cid:13)(cid:13)(cid:13). Some operators satisfy weighted inequalities for classes of weights of the type Aq ∩ RHσ. This is the case of adjoints of linear operators bounded on Lp,w(Rn) for p > γ and w ∈ Ap/γ. By duality the adjoint operator is bounded on Lp,w(Rn) for 1 < p ≤ γ and {w1−p : w ∈ Ap′/γ}, which is the same as Ap ∩ RHγ ′/(γ ′−p). Also operators bounded in a limited range of values of p satisfy weighted inequalities for weights in classes of such type (see [AM07] and [CDL12]). Similarly, we have sometimes the weak or strong boundedness with weights in a class of the type {w : wν ∈ A1}. Theorem 4.8. Let 1 ≤ p− ≤ p0 ≤ p+ < ∞ and let F be a collection of nonnegative measurable pairs of functions. Let F be a collection of nonneg- ative pairs of function in Lp0,w(Rn). Assume that for every (f, g) ∈ F and every w ∈ Ap0/p− ∩ RH(p+/p0)′ we have (4.11) k g Lp0,w(Rn)k ≤ c1 k f Lp0,w(Rn)k , where c1 depends on the Ap0-constant of the weight w. Let p− < p < p+ (and p = p− if p0 = p−) and w ∈ Ap/p− ∩ RHσ with σ ≥ (p+/p)′. Then for every −n/p ≤ r ≤ −nσ′/p+ we have (4.12) (4.13) Furthermore, if −n/p ≤ r ≤ −n/(pσ) − n/p+ we have (cid:13)(cid:13) g Lr p(w, Rn)(cid:13)(cid:13) ≤ c2(cid:13)(cid:13) f Lr (cid:13)(cid:13) g Lr p(λ, w, Rn)(cid:13)(cid:13) ≤ c3(cid:13)(cid:13) f Lr p(w, Rn)(cid:13)(cid:13) . p(λ, w, Rn)(cid:13)(cid:13) . The constants c2 and c3 in (4.12) and (4.13) do not depend on f but may depend on the weights and the involved parameters. Proof. It is enough to prove the result for p− = 1, which we assume in what follows. To treat the case p− > 1 write (4.11) for the pair (f p−, gp−) on Lp0/p−,w(Rn), and apply the case p− = 1. (a) The case p > p− (that is, p > 1). First we observe that by limited range extrapolation (cf. [AM07, Thm. 4.9] or [CMP11, Thm. 3.31]) (4.11) holds in Lp,v(Rn) with v ∈ Ap ∩ RH(p+/p)′ for 1 < p < p+. Let w ∈ Ap ∩ RHσ. We will apply Proposition 3.8 with b = p+ and γ = 1. We choose q ∈ (1, p) and s > (p+/q)′ such that the embedding (3.19) holds. 20 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Let Q be a fixed cube. As in the proof of Theorem 4.1, we use duality, and for h in Lp′,w(Q) of norm 1 we get (4.4), that is, (cid:18)ZRn gqhwχQ(cid:19) 1 q ≤ c(cid:18)ZRn f qM (hswsχQ) 1 q s(cid:19) 1 , with constant independent of h and Q, because M (hswsχQ)1/s ∈ A1 ∩ RH(p+/q)′, due to the choice of s > (p+/q)′. Under the conditions required to r we can apply Proposition 3.8 to obtain (4.12) and (4.13). (b) The case p = p− (that is, p = 1). Now we are assuming that the + ∈ A1). The inequality holds for p = 1 and w ∈ A1 ∩ RHp′ proof is as in part (b) of the proof of Theorem 4.1 except for the fact that we need to choose s > p′ + to guarantee that M (wsχQ)p′ +/s is in A1. (that is, wp′ + Under the assumptions on r, we are in the conditions of Proposition 3.4 +. Then (4.12) and (4.13) are deduced from (3.13) and (3.14) (cid:3) with ν = p′ with p = 1. Using this theorem and the embeddings given in Propositions 3.4 and 3.8 we can obtain results analogous to those of Corollary 4.4, for the corre- sponding ranges of p and classes of weights. Corollary 4.9. Let 1 ≤ p− ≤ p0 ≤ p+ < ∞. Assume that T is an opera- tor acting fromSw∈Ap0/p− ∩RH(p+/p0)′ Lp0,w(Rn) into the space of measurable functions that satisfies k T f Lp0,w(Rn)k ≤ c1 k f Lp0,w(Rn)k for all f ∈ Lp0,w(Rn) and w ∈ Ap0/p− ∩ RH(p+/p0)′, with a bound depending on the constant of the weight. Then for every p− < p < p+ (and also p = p− if p0 = p−) and w ∈ Ap/p− ∩ RHσ with σ ≥ (p+/p)′, and for every −n/p ≤ r ≤ −nσ′/p+ we have for all f ∈ Lr p(w, Rn). Furthermore, if −n/p ≤ r ≤ −n/(pσ) − n/p+ we have (cid:13)(cid:13) T f Lr p(w, Rn)(cid:13)(cid:13) ≤ c2(cid:13)(cid:13) f Lr p(λ, w, Rn)(cid:13)(cid:13) ≤ c3(cid:13)(cid:13) f Lr (cid:13)(cid:13) T f Lr p(w, Rn)(cid:13)(cid:13) p(λ, w, Rn)(cid:13)(cid:13) for all f ∈ Lr p(λ, w, Rn). Weak-type inequalities like in Corollary 4.4 and vector-valued inequalities like in Corollary 4.7 can be written also in this setting. Remark 4.10. In [CMP11] the authors get very general results with re- spect to extrapolation in Banach function spaces. But in this abstract set- ting they do not get results either at the endpoint (A1 weights) or with respect to limited range extrapolation (cf. Theorems 4.8). They use a kind of duality (associated spaces) which is available for Morrey spaces of type p(w, Rn) (cf. [RS16]), but we completely avoid it. However, their results Lr deliver weighted inequalities which means that for applications to operators BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 21 one still has to deal with the well-definedness of the operators (which we have done via embeddings). 5. Applications to mapping properties of operators There is a plentiful list of operators fulfilling the requirements of one or more of the theorems of the previous section. As we mentioned in the introduction several of those operators have been proved to be bounded in Morrey spaces (unweighted or weighted) using in general the particular form of the operator or some particular pointwise bound. In most cases only the size estimate with the Morrey norm has been proved, without any discussion about the possible definition of the operator in the corresponding Morrey space. All of this is avoided with our approach. In what follows, we present a collection of applications of our general results. Vector-valued Morrey inequalities also hold in all cases, but we do not mention them. In the last subsection we give Morrey estimates with A∞ weights and 0 < p < ∞. 5.1. Calder´on-Zygmund operators and their maximal truncations. It is by now a classical result that Calder´on-Zygmund operators are bounded on Lp,w(Rn) for w ∈ Ap (1 < p < ∞), and that they are of weak-type (1, 1) with respect to A1 weights. Here we can understand the term Calder´on- Zygmund operator in the general sense, that is, when the operator is repre- sented by a two-variable kernel K(x, y) with appropriate size and regularity estimates. The maximal operator associated to the Calder´on-Zygmund op- erators taking the supremum of the truncated integrals also satisfies similar weighted estimates. For a proof the reader can consult [Gra09, Chapter 9]. Then we can apply Corollary 4.4. The regularity assumptions on K(x, y) can be weakened and stated in some integral form called Lr-Hormander condition. In the convolution case it appears implicitely in [KW79], and in a more general setting in [RRT86], In that case the weighted inequalities hold for Lp,w(Rn) for for instance. p ≥ γ and w ∈ Ap/γ, where γ = r′. Then Corollary 4.4 applies. If the estimates are symmetric in the variables x and y of the kernel, duality can be used for the weighted estimates and the results of Corollary 4.9 apply. 5.2. Multipliers. Kurtz and Wheeden studied in [KW79] weighted inequal- ities for classes of multipliers defined as follows (see also [ST89]). Let cT f (ξ) = m(ξ) f (ξ). We say that m ∈ M (s, l) if Dαm(ξ)s!1/s R>0 Rsα−nZR<ξ<2R sup < +∞ for all α ≤ l. The result of Kurtz and Wheeden says that for 1 < s ≤ 2 and n/s < l ≤ n, Tm is bounded on Lp,w(Rn) if n/l < p < ∞ and w ∈ Apl/n; its dual result for 1 < p < (n/l)′; and of weak-type (1, 1) with respect to weights w such that wn/l ∈ A1. Although they do not make it explicit, interpolation gives 22 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL that Tm is bounded on Lp,w(Rn) for wlp/n ∈ A1, when 1 < p ≤ n/l. We can apply Corollaries 4.4 and 4.9, depending on the values of p. A particularly interesting case is M (2, l), which corresponds to the Mihlin- Hormander multipliers. For l = n the full range of results holds applying Corollary 4.4 with γ = 1. Let us particularize some of the results valid for any l > n/2: • for p ≥ 2 and w ∈ Ap/2, we have boundedness on Lr p(w, Rn) with- out restriction on r; for w ∈ Ap/2 ∩ RHσ we have boundedness on Lr p(λ, w, Rn) for −n/p ≤ r ≤ −n/(pσ) (Corollary 4.9 with p− = 2); • for 1 < p < 2 and w ∈ Ap∩RHσ with σ′ < 2/p, we have boundedness p(λ, w, Rn) for − − p(w, Rn) when r ∈ [−n/p, −nσ′/2] and on Lr on Lr n/p ≤ r ≤ −n/(pσ) − n/2 (Corollary 4.9); • for p ≥ 2 and w(x) = xβ (or (1 + x)β) with −n < β < n(p/2 − 1), p(w, Rn) without restriction on r (Corol- we have boundedness on Lr lary 4.4 with γ = 2); • for p ≥ 2 and w(x) = xβ (or (1 + x)β) with −n < β < 0 we have p(λ, w, Rn) for −n/p ≤ r ≤ β/p and if 0 ≤ β < p(λ, w, Rn) for −n/p ≤ r < 0 boundedness on Lr n(p/2 − 1), we have boundedness on Lr (Corollary 4.9 with p− = 2 together with Remark 2.7); • for 1 < p ≤ 2 and w(x) = xβ (or (1 + x)β) with −np/2 < β ≤ 0, p(w, Rn) holds for − n the boundedness on Lr This is valid for p = 1 using the weak Morrey estimate. In the case of Lr p . We apply Corollary 4.9 with p− = p and (p+/p)′ = 2/p, and Remark 2.7. p ≤ r < −n(cid:16) 1 p(λ, w, Rn) the range is − n 2(cid:17) + β p − 1 p ≤ r < − n2 p − 1 n+β(cid:16) 1 2(cid:17). • for 1 < p ≤ 2 and w(x) = xβ (or (1 + x)β) with 0 ≤ β < n(p − 1), we have boundedness on Lr p(λ, w, Rn) when r ∈ [−n/p, −n/2). Here we use Corollary 4.9 with p+ = 2 and the fact that these weights are in RHσ for all σ > 1. p(w, Rn) and on Lr The origin of Morrey spaces is in the study of the smoothness properties of solutions of PDEs. For this reason results on the multipliers M (s, l) related to their smoothness are interesting (cf. [Tri13, Tri14, Ros12]). Another interesting case of multipliers is that of Marcinkiewicz multipliers in one dimension. They are bounded on Lp,w(R) for w ∈ Ap (see [Kur80]) and Corollary 4.4 gives all the Morrey boundedness results for 1 < p < ∞. 5.3. Rough operators. The rough singular integrals are defined in general by TΩ,hf (x) = p.v.ZRn Ω(y′)h(y) yn f (x − y)dy with Ω ∈ L1(Sn−1) (y′ = yy−1) and integral zero, and h defined on (0, ∞). The (weighted) boundedness of the operator is proved using additional as- sumptions on Ω and h. For instance, if both Ω and h are in L∞, it was proved BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 23 in [DR86] that TΩ,h is bounded on Lp,w(Rn) for w ∈ Ap (1 < p < ∞), so that the full range is obtained for Morrey spaces from Theorem 4.1. If for fixed q > 1 we know that Ω ∈ Lq(Sn−1) and h ∈ Lq((0, ∞); dt/t) then the boundedness on Lp,w(Rn) holds for w ∈ Ap/q′ (q′ ≤ p < ∞) and also for some classes of weights obtained by duality for the range 1 < p ≤ q. These classes can further be extended by interpolation. Results of this type are in [Wat90] and [Duo93]. The weighted inequalities are very much like those for the multipliers of the previous section, hence similar Morrey esti- mates are obtained. Due to the openness properties of Ap weights, D. Wat- son noticed that for Ω ∈Tq<∞ Lq(Sn−1) and h ∈Tq<∞ Lq((0, ∞); dt/t) the full range Lp,w(Rn) for w ∈ Ap and 1 < p < ∞ is also obtained. Hence the full range of Morrey estimates is deduced. Notice that in this case the kernel of the operator TΩ,h need not satisfy the size estimate (1.3). Similar weighted estimates are known for the maximal operator defined from the truncated integrals of the rough operators. 5.4. Square functions. There exist several types of square functions for which our theorems apply. The classical Littlewood-Paley operators in one dimension associated to the dyadic intervals are bounded on Lp,w(R) for w ∈ Ap (see [Kur80]). A similar result holds in all dimensions in we consider the square functions of discrete or continuous type built using dilations of a Schwartz function with zero integral (see [Ryc01, Prop. 1.9] and the comments following it). The Lusin area integral, the gλ functions, Stein's n-dimensional extension of the Marcinkiewicz integral and the intrinsic square function of M. Wilson are other examples of square functions for which the weighted inequalities on Lp,w(Rn) for w ∈ Ap (1 < p < ∞) are known to hold (see [TW90] and [Wil08, Thm. 7.2 and notes of Chapter 7]). Of course, all of them are covered by our Corollary 4.4. Rough variants have been considered also for square functions with the introduction of some Ω ∈ Lq(Sn−1) as for the singular integrals mentioned above. The class of weights for which boundedness holds depends on q (see [DFP99] and [DS02]). On the other hand, we can mention Rubio de Francia's results of Littlewood- Paley type. The square function built using comparable cubes in Rn is bounded on Lp,w(Rn) for w ∈ Ap/2 and 2 ≤ p < ∞ (see [Rub83]) and in this case we can apply Corollary 4.4 with γ = 2 to obtain the corresponding Morrey estimates. For the square function built using arbitrary disjoint in- tervals in R we can apply Corollary 4.4 to the result of [Rub85] in which the boundedness of the operator is obtained for Lp,w(Rn) with w ∈ Ap/2 and p > 2 (the endpoint p = 2 remains open). 5.5. Commutators. There are several results on weighted inequalities for commutators to which we can apply our theorems to obtain the correspond- ing Morrey estimates. Let us mention in particular the remarkable result of 24 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL ´Alvarez et al. in [ABKP93], which allows to obtain many weighted inequal- ities for commutators with BMO functions. A good list of applications can be seen in their paper. For each one of them we obtain the corresponding result in the Morrey spaces. 5.6. Other operators. There are many more results in the literature con- cerning weighted inequalities with Ap type weights. We briefly mention here some other remarkable operators for which our theorems apply directly. Al- though in some cases particular proofs of their Morrey estimates do exist, not all of them seem to have been considered before. Oscillatory singular integrals with kernels of the form eiP (x,y)K(x, y) where P (x, y) is a polynomial and K(x, y) is a Calder´on-Zygmund standard kernel or a kernel of rough type have been studied and weighted results appear for instance in [LDY07]. A further variant are the so-called multilinear oscilla- tory singular integrals in which the kernel is modified by inserting a factor x − y−mRm+1(A; x, y) where the second term is the (m + 1)-th order re- mainder of the Taylor series of a function A expanded in x about y. Some weighted inequalities are in [Wu04]. Other types of oscillating kernels are of the form Ka,b+iy(t) = exp(ita)(1 + t)−(b+iy) for which weighted inequal- ities are obtained in [CKS83], and strongly singular convolution operators corresponding to multipliers of the form θ(ξ)eiξb ξ−a, where θ is a smooth radial cut-off function (vanishing in a neighborhood of the origin and which is identically 1 outside a compact set), studied in [Cha84]. Bochner-Riesz operators at the critical index satisfy Ap-weighted esti- mates for 1 < p < ∞ and weak-type (1, 1) estimates with A1 weights ([SS92] and [Var96]). The boundedness on the Morrey spaces are obtained from Corollary 4.4. Below the critical index only partial results can be obtained, but still there are some subclasses of Ap weights (see [CDL12], for instance) for which the corresponding theorems in the previous section can be ap- plied. Some weighted inequalities also hold for the maximal Bochner-Riesz operator. Weighted inequalities for pseudodifferential operators are for instance in [Yab85] and [Sat05], while for Fourier integral operators they can be found in [DSS14]. In both cases, Morrey space estimates can be deduced from our results in Section 4. 5.7. Morrey norm inequalities with A∞ weights. Let 0 < p < ∞ and w ∈ A∞. Then the following inequalities hold whenever the left-hand side is finite: (5.1) (5.2) (5.3) (cid:13)(cid:13) T f Lr (cid:13)(cid:13) M f Lr (cid:13)(cid:13) Iαf Lr p(w, Rn)(cid:13)(cid:13) ≤ c1(cid:13)(cid:13) M f Lr p(w, Rn)(cid:13)(cid:13) ≤ c2(cid:13)(cid:13)(cid:13) M ♯f(cid:12)(cid:12)(cid:12) Lr p(w, Rn)(cid:13)(cid:13) ≤ c3(cid:13)(cid:13) Mαf Lr p(w, Rn)(cid:13)(cid:13) , p(w, Rn)(cid:13)(cid:13)(cid:13) , p(w, Rn)(cid:13)(cid:13) , BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 25 where T is a Calder´on-Zygmund operator in (5.1), M is the Hardy-Littlewood maximal operator in (5.1) and (5.2), M ♯ is the sharp maximal function (see [Duo01, p. 117]) in (5.2), and Iα and Mα are the fractional integral and the fractional maximal operator of order α ∈ (0, n) (see [Duo01, p. 88-89]) in p(w, Rn) if w (5.3). Similar inequalities hold with Lr is assumed to be in RHσ and −n/p ≤ r ≤ −n/(pσ). p(λ, w, Rn) instead of Lr The inequalities are derived from their Lp,w counterparts using Corollary 4.3. A proof of the needed Lp,w estimates using extrapolation can be found in [CMP04], where references concerning the original inequalities (obtained by means of good-λ inequalities) are given. In a similar way, other Lp,w inequalities appearing in [CMP04] can also be written with Morrey norms. In that paper weak-type inequalities and vector-valued inequalities are also given and they can also be transferred to the Morrey setting, but we do not go into this in detail. Weighted inequalities with Morrey norms involving M ♯ with A∞ weights may be found also in [NkSa17, Corollary 2]. References [AX12] [Ada15] [Alv96] Adams, D. R.; Xiao, J.: Morrey spaces in harmonic analysis. Ark. Mat. 50, 201 -- 230 (2012). Adams, D. R.: Morrey spaces. Lecture Notes in Applied and Numerical Har- monic Analysis. Birkhauser/Springer, Cham, 2015. ´Alvarez, J.: Continuity of Calder´on-Zygmund type operators on the predual of a Morrey space. In: Clifford algebras in analysis and related topics, CRC Press, Boca Raton, 309 -- 319 (1996). [CDL12] [ABKP93] ´Alvarez, J.; Bagby, R. J.; Kurtz, D. S.; P´erez, C.: Weighted estimates for commutators of linear operators. Studia Math. 104, no. 2, 195 -- 209 (1993). Auscher, P.; Martell, J. M., Weighted norm inequalities, off-diagonal estimates and elliptic operators. I. General operator theory and weights, Adv. Math. 212 (2007), 225 -- 276. Carro, M. J.; Duoandikoetxea, J.; Lorente, M.: Weighted estimates in a limited range with applications to the Bochner-Riesz operators. Indiana Univ. Math. J. 61, no. 4, 1485 -- 1511 (2012). Chanillo, S.; Kurtz, G.; Sampson, G.: Weighted Lp estimates for oscil- lating kernels, Ark. Mat. 21, no. 2, 233 -- 257 (1983). Chanillo, S.: Weighted norm inequalities for strongly singular convolution operators, Trans. Amer. Math. Soc. 281, no. 1, 77 -- 107 (1984). Chiarenza, F.; Frasca, M.: Morrey spaces and Hardy-Littlewood maximal function. Rend. Mat. Appl. (7) 7, no. 3-4, 273 -- 279 (1987). [CKS83] [AM07] [Cha84] [CF87] [CMP04] Cruz-Uribe, D. V.; Martell, J.M.; P´erez, C.: Extrapolation from A∞ weights and applications. J. Funct. Anal. 213, no. 2, 412-439 (2004). [CMP11] Cruz-Uribe, D. V.; Martell, J.M.; P´erez, C.: Weights, extrapolation and the theory of Rubio de Francia. Operator Theory: Advances and Applications 215. Basel: Birkhauser (2011). Ding, Y.; Fan D.; Pan Y.: Weighted boundedness for a class of rough Marcinkiewicz integrals, Indiana Univ. Math. J. 48, 1037 -- 1055 (1999). Ding, Y.; Yang, D.; Zhou, Z.: Boundedness of sublinear operators and commutators on Morrey spaces. Yokohama Math. J. 46, No.1, 15 -- 27 (1998). [DYZ98] [DFP99] 26 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL [DSS14] [Duo93] [Duo01] [Duo11] [DR86] [DS02] [Geh73] [Gra08] [Gra09] Dos Santos Ferreira, D.; Staubach, W.: Global and local regularity of Fourier integral operators on weighted and unweighted spaces. Mem. Amer. Math. Soc. 229, no. 1074 (2014). Duoandikoetxea, J.: Weighted norm inequalities for homogeneous singular integrals. Trans. Amer. Math. Soc. 336, no. 2, 869 -- 880 (1993). Duoandikoetxea, J.: Fourier analysis. Graduate Studies in Mathematics. 29. Providence, RI: American Mathematical Society (2001). Duoandikoetxea, J.: Extrapolation of weights revisited: new proofs and sharp bounds. J. Funct. Anal. 260, No. 6, 1886 -- 1901 (2011). Duoandikoetxea, J.; Rubio de Francia, J. L.: Maximal and singular integral operators via Fourier transform estimates. Invent. Math. 84, no. 3, 541 -- 561 (1986). Duoandikoetxea, J.; Seijo, E.: Weighted inequalities for rough square functions through extrapolation. Studia Math. 149, no. 3, 239 -- 252 (2002). Gehring, F. W.: The Lp-integrability of the partial derivatives of a quasi- conformal mapping, Acta Math. 130, 265 -- 277 (1973). Grafakos, L.: Classical Fourier analysis. 2nd ed. Grad. Texts in Math. 249. Springer, New York, 2008. Grafakos, L.: Modern Fourier analysis. 2nd ed. Grad. Texts in Math. 250. Springer, New York, 2009. [GAKS11] Guliyev, V. S.; Aliyev, S. S.; Karaman, T.; Shukurov, P. S.: Bounded- ness of Sublinear Operators and Commutators on Generalized Morrey Spaces. Integr. Equ. Oper. Theory 71, 327 -- 355 (2011). Guliyev, V. S.: Generalized weighted Morrey spaces and higher order com- mutators of sublinear operators. Eurasian Math. J. 3, No. 3, 33 -- 61 (2012). Izumi, T.; Sawano, Y.; Tanaka, H.: Littlewood-Paley theory for Morrey spaces and their preduals. Rev. Mat. Complut. 28, No. 2, 411 -- 447 (2015). Johnson, R.; Neugebauer, C. J.: Change of variable results for Ap and reverse Holder RHr-classes, Trans. Amer. Math. Soc. 328 (1991), 639 -- 666. [IST15] [Gul12] [JN91] [KGS14] Karaman, T.; Guliyev, V. S.; Serbetci, A. Boundedness of sublinear operators generated by Calder´on-Zygmund operators on generalized weighted Morrey spaces. An. Stiint. Univ. Al. I. Cuza Iasi, Ser. Noua, Mat. 60, No. 1, 227 -- 244 (2014). [KMM16] Knese, G.; McCarthy, J. E.; Moen, K.: Unions of Lebesgue spaces and [KS09] [Kur80] [KW79] [LDY07] A1 majorants. Pacific J. Math. 280 (2016), no. 2, 411 -- 432. Komori, Y.; Shirai, S.: Weighted Morrey spaces and a singular integral operator. Math. Nachr. 282, No. 2, 219 -- 231 (2009). Kurtz, D. S.: Littlewood-Paley and multiplier theorems on weighted Lp spaces. Trans. Amer. Math. Soc. 259, no. 1, 235 -- 254 (1980). Kurtz, D. S..; Wheeden, R. L.: Results on weighted norm inequalities for multipliers. Trans. Amer. Math. Soc. 255, 343 -- 362 (1979). Lu, S.; Ding, Y.; Yan, D.: Singular integrals and related topics. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2007. [Mus12] Mustafayev, Rza Ch.: On boundedness of sublinear operators in weighted [Nak94] [Nkm16] Morrey spaces. Azerb. J. of Math. 2, 66 -- 79 (2012). Nakai, E.: Hardy-Littlewood maximal operator, singular integral operators and the Riesz potentials on generalized Morrey spaces. Math. Nachr. 166, 95 -- 103 (1994). Nakamura, S.: Generalized weighted Morrey spaces and classical operators. Math. Nachr. 289 (2016), 2235 -- 2262. [NkSa17] Nakamura, S.; Sawano, Y.: The singular integral operator and its commu- tator on weighted Morrey spaces. Collect. Math. 68 (2017), no. 2, 145 -- 174. BOUNDEDNESS ON MORREY SPACES, EXTRAPOLATION AND EMBEDDINGS 27 [Pee66] [Pic69] [PT15] [Ros12] [Ros13] [RS14] [RS16] [RT13] [RT14] [Rub82] [Rub83] [Rub84] [Rub85] [RRT86] [Ryc01] [Sam09] [Sat05] [SFZ13] [SS92] [ST89] [Ta15] [TW90] Peetre, J.: On convolution operators leaving Lp,λ spaces invariant. Ann. Mat. Pura Appl. 72, 295 -- 304 (1966). Piccinini, L. C.: Propriet`a di inclusione e interpolazione tra spazi di Morrey e loro generalizzazioni. Tesi di perfezionamento, Scuola Normale Superiore Pisa, 1 -- 153 (1969). Poelhuis, J.; Torchinsky, A.: Weighted local estimates for singular inte- gral operators. Trans. Am. Math. Soc. 367 (2015), 7957 -- 7998. Rosenthal, M.: Local means, wavelet bases and wavelet isomorphisms in Besov-Morrey and Triebel-Lizorkin-Morrey spaces. Math. Nachr. 286, No. 1, 59 -- 87 (2013). Rosenthal, M.: Mapping properties of operators in Morrey spaces and wavelet isomorphisms in related Morrey smoothness spaces. PhD-Thesis, Jena, 2013. Rosenthal, M.; Schmeisser, H.-J.: On the boundedness of singular inte- grals in Morrey spaces and its preduals, J. Fourier Anal. Appl., 22(2), 462 -- 490, 2016. Erratum, J. Fourier Anal. Appl., 22(2), p. 491, 2016. Rosenthal, M.; Schmeisser, H.-J.: The boundedness of operators in Muck- enhoupt weighted Morrey spaces via extrapolation techniques and duality. Rev. Mat. Complut. 29 (2016), no. 3, 623 -- 657. Rosenthal, M.; Triebel, H.: Calder´on-Zygmund operators in Morrey spaces. Rev. Mat. Complut. 27, 1 -- 11 (2014). Rosenthal, M.; Triebel, H.: Morrey spaces, their duals and preduals. Rev. Mat. Complut. 28, 1 -- 30 (2015). Rubio de Francia, J. L.: Factorization and extrapolation of weights. Bull. Am. Math. Soc., New Ser. 7, 393 -- 395 (1982). Rubio de Francia, J. L.: Estimates for some square functions of Littlewood- Paley type. Publ. Sec. Mat. Univ. Aut`onoma Barcelona 27, no. 2, 81 -- 108 (1983). Rubio de Francia, J. L.: Factorization theory and Ap weights. Amer. J. Math. 106, 533 -- 547 (1984). Rubio de Francia, J. L.: A Littlewood-Paley inequality for arbitrary inter- vals. Rev. Mat. Iberoamericana 1, no. 2, 1 -- 14 (1985). Rubio de Francia, J. L.; Ruiz, F. J.; Torrea, J. L.: Calder´on-Zygmund theory for operator-valued kernels. Adv. in Math. 62, no. 1, 7 -- 48 (1986). Rychkov, V. S.: Littlewood-Paley theory and function spaces with Aloc weights. Math. Nachr. 224, 145 -- 180 (2001). Samko, N.: Weighted Hardy and singular operators in Morrey spaces. J. Math. Anal. Appl. 350, 56 -- 72 (2009). Sato, S.: A note on weighted estimates for certain classes of pseudo- differential operators. Rocky Mountain J. Math. 35, no. 1, 267 -- 284 (2005). Shi, S.; Fu, Z.; Zhao, F.: Estimates for operators on weighted Morrey spaces and their applications to nondivergence elliptic equations. J. Inequal. Appl. 2013, Article ID 390, 16 p., (2013). Shi, X. L.; Sun, Q. Y.: Weighted norm inequalities for Bochner-Riesz op- erators and singular integral operators. Proc. Amer. Math. Soc. 116, no. 3, 665 -- 673 (1992). Stromberg, J.-O.; Torchinsky, A.: Weighted Hardy spaces. Lecture Notes in Mathematics, 1381. Springer-Verlag, Berlin, 1989. Tanaka, H.: Two-weight norm inequalities on Morrey spaces. Ann. Acad. Sci. Fenn. Math. 40, 773 -- 791 (2015). Torchinsky, A.; Wang S.: A note on the Marcinkiewicz integral, Colloq. Math. 60-61, 235 -- 243 (1990). p 28 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL [Tri13] [Tri14] [Tri16] [Var96] Triebel, H.: Local function spaces, heat and Navier-Stokes equations. Zurich: European Mathematical Society, 2013. Triebel, H.: Hybrid function spaces, heat and Navier-Stokes equations. Zurich: European Mathematical Society, 2014. Triebel, H.: The Fatou property in function spaces, heat kernels, admis- sible norms and mapping properties. In: Function Spaces and Inequalities, Springer, Singapore (2017). Vargas, A. M.: Weighted weak type (1,1) bounds for rough operators. J. London Math. Soc. (2) 54, no. 2, 297 -- 310 (1996). [Wan16] Wang, H.: Boundedness of θ-type Calder´on-Zygmund operators and commu- tators in the generalized weighted Morrey spaces, J. Funct. Spaces 2016, Art. ID 1309348, 18 p. [Wat90] Watson, D. K.: Weighted estimates for singular integrals via Fourier trans- [Wil08] [Wu04] [Yab85] form estimates. Duke Math. J. 60, no. 2, 389 -- 399 (1990). Wilson, M.: Weighted Littlewood-Paley theory and exponential-square in- tegrability, Lecture Notes in Mathematics, 1924, Springer, Berlin, 2008. Wu, H.: On multilinear oscillatory singular integrals with rough kernels. J. Math. Anal. Appl. 296, no. 2, 479 -- 494 (2004). Yabuta, K.: Calder´on-Zygmund operators and pseudodifferential operators. Comm. Partial Differential Equations 10, no. 9, 1005 -- 1022 (1985). Universidad del Pa´ıs Vasco/Euskal Herriko Unibertsitatea, Departamento de Matem´aticas/Matematika saila, Apdo. 644, 48080 Bilbao, Spain E-mail address: [email protected], [email protected]
1610.03420
1
1610
2016-10-11T16:42:43
Reproducing pairs of measurable functions and partial inner product spaces
[ "math.FA" ]
We continue the analysis of reproducing pairs of weakly measurable functions, which generalize continuous frames. More precisely, we examine the case where the defining measurable functions take their values in a partial inner product space (PIP spaces). Several examples, both discrete and continuous, are presented.
math.FA
math
Reproducing pairs of measurable functions and partial inner product spaces J-P. Antoine a and C. Trapani b a Institut de Recherche en Math´ematique et Physique, Universit´e catholique de Louvain B-1348 Louvain-la-Neuve, Belgium E-mail address: [email protected] b Dipartimento di Matematica e Informatica, Universit`a di Palermo, I-90123 Palermo, Italy E-mail address: [email protected] Abstract We continue the analysis of reproducing pairs of weakly measurable functions, which gen- eralize continuous frames. More precisely, we examine the case where the defining measurable functions take their values in a partial inner product space (pip-space). Several examples, both discrete and continuous, are presented. AMS classification numbers: 41A99, 46Bxx, 46C50, 46Exx Keywords: Reproducing pairs, continuous frames, upper and lower semi-frames, partial inner product spaces, lattices of Banach spaces 1 Introduction Frames and their relatives are most often considered in the discrete case, for instance in signal processing [10]. However, continuous frames have also been studied and offer interesting mathe- matical problems. They have been introduced originally by Ali, Gazeau and one of us [1, 2] and also, independently, by Kaiser [13]. Since then, several papers dealt with various aspects of the concept, see for instance [11] or [17]. However, there may occur situations where it is impossible to satisfy both frame bounds. Therefore, several generalizations of frames have been introduced. Semi-frames [6, 7], for example, are obtained when functions only satisfy one of the two frame bounds. It turns out that a large portion of frame theory can be extended to this larger framework, in particular the notion of duality. More recently, a new generalization of frames was introduced by Balazs and Speckbacher [20], namely, reproducing pairs. Here, given a measure space (X, µ), one considers a couple of weakly measurable functions (ψ, φ), instead of a single mapping, and one studies the correlation between the two (a precise definition is given below). This definition also includes the original definition of a continuous frame [1, 2] given the choice ψ = φ. The increase of freedom in choosing the mappings ψ and φ, however, leads to the problem of characterizing the range of the analysis operators, which in general need no more be contained in L2(X, dµ), as in the frame case. Therefore, we extend the theory to the case where the weakly measurable functions take their values in a partial inner product space (pip-space). We discuss first the case of a rigged Hilbert space, then we consider a genuine pip-space. We conclude with two natural families of examples, namely, Hilbert scales and several pip-spaces generated by the family {Lp(X, dµ), 1 6 p 6 ∞}. 2 Preliminaries Before proceeding, we list our definitions and conventions. The framework is a (separable) Hilbert space H, with the inner product h··i linear in the first factor. Given an operator A on H, we 1 denote its domain by D(A), its range by Ran (A) and its kernel by Ker (A). GL(H) denotes the set of all invertible bounded operators on H with bounded inverse. Throughout the paper, we will consider weakly measurable functions ψ : X → H, where (X, µ) is a locally compact space with a Radon measure µ, that is, hψxf i is µ−measurable for every f ∈ H. Then the weakly measurable function ψ is a continuous frame if there exist constants m > 0 and M < ∞ (the frame bounds) such that m kf k2 6ZX hf ψxi2 dµ(x) 6 M kf k2 , ∀ f ∈ H. (2.1) (2.2) Given the continuous frame ψ, the analysis operator Cψ : H → L2(X, dµ) 1 is defined as (Cψf )(x) = hf ψxi, f ∈ H, and the corresponding synthesis operator C∗ψ : L2(X, dµ) → H as (the integral being understood in the weak sense, as usual) C∗ψξ =ZX ξ(x) ψx dµ(x), for ξ ∈ L2(X, dµ). (2.3) We set S := C∗ψCψ, which is self-adjoint. More generally, the couple of weakly measurable functions (ψ, φ) is called a reproducing pair if [8] (a) The sesquilinear form Ωψ,φ(f, g) =ZX hf ψxihφxgi dµ(x) (2.4) is well-defined and bounded on H × H, that is, Ωψ,φ(f, g) 6 c kf k kgk, (b) The corresponding bounded (resolution) operator Sψ,φ belongs to GL(H). for some c > 0. Under these hypotheses, one has Sψ,φf =ZX hf ψxiφx dµ(x), ∀ f ∈ H, (2.5) the integral on the r.h.s. being defined in weak sense. If ψ = φ, we recover the notion of continuous frame, so that we have indeed a genuine generalization of the latter. Notice that Sψ,φ is in general neither positive, nor self-adjoint, since S∗ψ,φ = Sφ,ψ . However, if ψ, φ is reproducing pair, then ψ, S−1 ψ,φφ is a dual pair, that is, the corresponding resolution operator is the identity. Therefore, there is no restriction of generality to assume that Sφ,ψ = I [20]. The worst that can happen is to replace some norms by equivalent ones. In [8], it has been shown that each weakly measurable function φ generates an intrinsic pre- Hilbert space Vφ(X, µ) and, moreover, a reproducing pair (ψ, φ) generates two Hilbert spaces, Vψ(X, µ) and Vφ(X, µ), conjugate dual of each other with respect to the L2(X, µ) inner product. Let us briefly sketch that construction, that we will generalize further on. Given a weakly measurable function φ, let us denote by Vφ(X, µ) the space of all measurable functions ξ : X → C such that the integral RX ξ(x)hφxgi dµ(x) exists for every g ∈ H and defines a bounded conjugate linear functional on H, i.e., ∃ c > 0 such that ZX (cid:12)(cid:12)(cid:12)(cid:12) ξ(x)hφxgi dµ(x)(cid:12)(cid:12)(cid:12)(cid:12) 6 c kgk , ∀ g ∈ H. 1As usual, we identify a function ξ with its residue class in L2(X, dµ). (2.6) 2 Clearly, if (ψ, φ) is a reproducing pair, all functions ξ(x) = hf ψxi = (Cψf )(x) belong to Vφ(X, µ). By the Riesz lemma, we can define a linear map Tφ : Vφ(X, µ) → H by the following weak relation hTφξgi =ZX Next, we define the vector space ξ(x)hφxgi dµ(x), ∀ ξ ∈ Vφ(X, µ), g ∈ H. (2.7) and equip it with the norm Vφ(X, µ) = Vφ(X, µ)/Ker Tφ k[ξ]φkφ := sup kgk61(cid:12)(cid:12)(cid:12)(cid:12) ZX ξ(x)hφxgi dµ(x)(cid:12)(cid:12)(cid:12)(cid:12) = sup kgk61 hTφξgi , (2.8) where we have put [ξ]φ = ξ + Ker Tφ for ξ ∈ Vφ(X, µ). Clearly, Vφ(X, µ) is a normed space. However, the norm k·kφ is in fact Hilbertian, that is, it derives from an inner product, as can be seen as follows. First, it turns out that the map bTφ : Vφ(X, µ) → H, bTφ[ξ]φ := Tφξ is a well-defined isometry of Vφ(X, µ) into H. Next, one may define on Vφ(X, µ) an inner product by setting and one shows that the norm defined by h··i(φ) coincides with the norm k · kφ defined in (2.8). One has indeed h[ξ]φ[η]φi(φ) := hbTφ[ξ]φbTφ[η]φi, [ξ]φ, [η]φ∈ Vφ(X, µ), k[ξ]φk(φ) =(cid:13)(cid:13)(cid:13)bTφ[ξ]φ(cid:13)(cid:13)(cid:13) = kTφξk = sup kgk61 hTφξgi = k[ξ]φkφ . Thus Vφ(X, µ) is a pre-Hilbert space. With these notations, the main result of [8] reads as Theorem 2.1 If (ψ, φ) is a reproducing pair, the spaces Vφ(X, µ) and Vψ(X, µ) are both Hilbert spaces, conjugate dual of each other with respect to the sesquilinear form hhξηiiµ :=ZX ξ(x)η(x) dµ(x), (2.9) which coincides with the inner product of L2(X, µ) whenever the latter makes sense. This is true, in particular, for φ = ψ, since then ψ is a continuous frame and Vψ(X, µ) is a closed subspace of L2(X, µ). In this paper, we will consider reproducing pairs in the context of pip-spaces. The motivation is the following. Let (ψ, φ) be a reproducing pair. By definition, hSψ,φf gi =ZX hf ψxihφxgi dµ(x) =ZX Cψf (x) Cφg(x) dµ(x) (2.10) is well defined for all f, g ∈ H. The r.h.s. coincides with the sesquilinear form (2.9), that is, the L2 inner product, but generalized, since in general Cψf, Cφg need not belong to L2(X, dµ). If, follow- ing [20], we make the innocuous assumption that ψ is uniformly bounded, i.e., supx∈X kψxkH 6 c for wavelets or coherent states), then (Cψf )(x) = for some c > 0 (often kψxkH = const., e.g. hf ψxi ∈ L∞(X, dµ). These two facts suggest to take Ran Cψ within some pip-space of measurable functions, possibly related to the Lp spaces. We shall present several possibilities in that direction in Section 6. 3 3 Reproducing pairs and RHS We begin with the simplest example of a pip-space, namely, a rigged Hilbert space (RHS). Let indeed D[t] ⊂ H ⊂ D×[t×] be a RHS with D[t] reflexive (so that t and t× coincide with the respective Mackey topologies). Given a measure space (X, µ), we denote by h·, ·i the sesquilinear form expressing the duality between D and D×. As usual, we suppose that this sesquilinear form extends the inner product of D (and H). This allows to build the triplet above. Let x ∈ X 7→ ψx, x ∈ X 7→ φx be weakly measurable functions from X into D×. Instead of (2.4), we consider he sesquilinear form ΩDψ,φ(f, g) =ZX hf, ψxihφx, gi dµ(x), f, g ∈ D, (3.1) and we assume that it is jointly continuous on D × D, that is ΩD ∈ B(D, D) in the notation of [3, Sec.10.2]. Writing hf, ψxihφx, gi dµ(x), ∀ f, g ∈ D, (3.2) hSψ,φf , gi :=ZX we see that the operator Sψ,φ belongs to L(D, D×), the space of all continuous linear maps from D into D×. 3.1 A Hilbertian approach We first assume that the sesquilinear form ΩD is well-defined and bounded on D × D in the topology of H. Then ΩDψ,φ extends to a bounded sesquilinear form on H × H, denoted by the same symbol. The definition of the space Vφ(X, µ) must be modified as follows. Instead of (2.6), we suppose that the integral below exists and defines a conjugate linear functional on D, bounded in the topology of H, i.e., ZX ξ(x)hφx, gi dµ(x)(cid:12)(cid:12)(cid:12)(cid:12) 6 c kgk , ∀ g ∈ D. ξ(x)hφx, gi dµ(x) = hhφ,ξgi, ∀g ∈ D. (cid:12)(cid:12)(cid:12)(cid:12) ZX Then the functional extends to a bounded conjugate linear functional on H, since D is dense in H. Hence, for every ξ ∈ Vφ(X, µ), there exists a unique vector hφ,ξ ∈ H such that (3.3) It is worth remarking that this interplay between the two topologies on D is similar to the approach of Werner [21], who treats L2 functions as distributions, thus identifies the L2 space as the dual of D = C∞0 with respect to the norm topology. And, of course, this is fully in the spirit of pip-spaces. Then, we can define a linear map Tφ : Vφ(X, µ) → H by Tφξ = hφ,ξ ∈ H, ∀ ξ ∈ Vφ(X, µ), (3.4) in the following weak sense hTφξgi =ZX ξ(x)hφx, gi dµ(x), g ∈ D, ξ ∈ Vφ(X, µ). In other words we are imposing that RX ξ(x)φx dµ(x) converge weakly to an element of H. The rest proceeds as before. We consider the space Vφ(X, µ) = Vφ(X, µ)/Ker Tφ, with the norm k[ξ]φkφ = kTφξk, where, for ξ ∈ Vφ(X, µ), we have put [ξ]φ = ξ + Ker Tφ. Then Vφ(X, µ) is a pre-Hilbert space for that norm. 4 Note that φ was called in [8] µ-independent whenever Ker Tφ = {0}. In that case, of course, Vφ = Vφ. Assume, in addition, that the corresponding bounded operator Sψ,φ is an element of GL(H). Then (ψ, φ) is a reproducing pair and Theorem 3.14 of [8] remains true, that is, Theorem 3.1 If (ψ, φ) is a reproducing pair, the spaces Vφ(X, µ) and Vψ(X, µ) are both Hilbert spaces, conjugate dual of each other with respect to the sesquilinear form h[ξ]φ[η]ψi =ZX ξ(x)η(x) dµ(x), ∀ ξ ∈ Vφ(X, µ), η ∈ Vψ(X, µ). (3.5) Example 3.2 To give a trivial example, consider the Schwartz rigged Hilbert space S(R) ⊂ L2(R, dx) ⊂ S×(R), (X, µ) = (R, dx), ψx(t) = φx(t) = 1√2π transform, so that hf φ(·)i ∈ L2(R, dx). In this case eixt . Then Cφf = bf , the Fourier Ωψ,φ(f, g) =ZR and Vφ(R, dx) = L2(R, dx). 3.2 The general case hf, ψxihφx, gi dx = hbf bgi = hf gi, ∀f, g ∈ S(R), In the general case, we only assume that the form Ω is jointly continuous on D × D, with no other regularity requirement. In that case, the vector space Vφ(X, µ) must be defined differently. Let the topology of D be given by a directed family P of seminorms. Given a weakly measurable function φ, we denote again by Vφ(X, µ) the space of all measurable functions ξ : X → C such that the integral RX ξ(x)hφx, gi dµ(x) exists for every g ∈ D and defines a continuous conjugate linear functional on D, namely, there exists c > 0 and a seminorm p ∈ P such that This in turn determines a linear map Tφ : Vφ(X, µ) → D× by the following relation ξ(x)hφx, gi dµ(x)(cid:12)(cid:12)(cid:12)(cid:12) 6 c p(g). (cid:12)(cid:12)(cid:12)(cid:12) ZX hTφξ, gi =ZX ξ(x)hφx, gi dµ(x), ∀ ξ ∈ Vφ(X, µ), g ∈ D. (3.6) Next, we define as before the vector space Vφ(X, µ) = Vφ(X, µ)/Ker Tφ, and we put again [ξ]φ = ξ + Ker Tφ for ξ ∈ Vφ(X, µ). Now we need to introduce a topology on Vφ(X, µ). We proceed as follows. Let M be a bounded subset of D[t]. Then we define hTφξ, gi . (3.7) That is, we are defining the topology of Vφ(X, µ) by means of the strong dual topology t× of D× which we recall is defined by the seminorms kF kM = sup g∈M hF gi, F ∈ D×, where M runs over the family of bounded subsets of D[t]. As said above, the reflexivity of D entails that t× is equal to the Mackey topology τ (D×, D). More precisely, 5 bpM([ξ]φ) := sup g∈M Lemma 3.3 The map bTφ : Vφ(X, µ) → D×, bTφ[ξ]φ := Tφξ is a well-defined linear map of Vφ(X, µ) into D× and, for every bounded subset M of D[t], one has bpM([ξ]φ) = kTφξkM, ∀ξ ∈ Vφ(X, µ) The latter equality obviously implies the continuity of Tφ. Next we investigate the dual Vφ(X, µ)∗ of the space Vφ(X, µ), that is, the set of continuous linear functionals on Vφ(X, µ). First, we have to choose a topology for Vφ(X, µ)∗. As usual we take the strong dual topology. This is defined by the family of seminorms where R runs over the bounded subsets of Vφ(X, µ). qR(F ) := sup [ξ]φ∈R F ([ξ]φ), Theorem 3.4 Assume that D[t] is a reflexive space and let φ be a weakly measurable function. If F is a continuous linear functional on Vφ(X, µ), then there exists a unique g ∈ D such that F ([ξ]φ) =ZX ξ(x)hφx, gi dµ(x), ∀ ξ ∈ Vφ(X, µ) (3.8) Moreover, every g ∈ H defines a continuous functional F on Vφ(X, µ) with kF kφ∗ 6 kgk, by (3.8). Proof : Let F ∈ Vφ(X, µ)∗. Then, there exists a bounded subset M of D[t] such that F ([ξ]φ) 6bpM([ξ]φ) = kTφξk , ∀ ξ ∈ Vφ(X, µ). M Let Mφ := {Tφξ : ξ ∈ Vφ(X, µ)} = Ran bTφ. Then Mφ is a vector subspace of D×. Let eF be the functional defined on Mφ by We notice that eF is well-defined. Indeed, if Tφξ = Tφξ′, then ξ − ξ′ ∈ Ker Tφ. Hence, [ξ]φ = [ξ′]φ Hence, eF is a continuous linear functional on Mφ which can be extended (by the Hahn-Banach theorem) to a continuous linear functional on D×. Thus, in virtue of the reflexivity of D, there exists a vector g ∈ D such that eF (Tφξ) := F ([ξ]φ), ξ ∈ Vφ(X, µ). and F ([ξ]φ) = F ([ξ′]φ) In conclusion, eF (Tφξ) = hbTφ[ξ]φ, gi = hTφξ, gi =ZX F ([ξ]φ) =ZX ξ(x)hφx, gi dµ(x), ∀ ξ ∈ Vφ(X, µ). ξ(x)hφx, gi dµ(x). Moreover, every g ∈ D obviously defines a continuous linear functional F on Vφ(X, µ) by (3.8). In addition, if R is a bounded subset of Vφ(X, µ), we have qR(F ) = sup [ξ]φ∈R = sup [ξ]φ∈R F ([ξ]φ) = sup [ξ]φ∈R hTφξ, gi 6 sup ξ(x)hφx, gi dµ(x)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ZX [ξ]φ∈RbpM([ξ]φ), 6 for any bounded subset M of D containing g. ✷ In the present context, the analysis operator Cφ is defined in the usual way, given in (2.2). Then, particularizing the discussion of Theorem 3.4 to the functional hCφg, ·i, one can interpret the analysis operator Cφ as a continuous operator from D to Vφ(X, µ)∗. As in the case of frames or semi-frames, one may characterize the synthesis operator in terms of the analysis operator. Proposition 3.5 Let φ be weakly measurable, then bTφ ⊆ C∗φ. If, in addition, Vφ(X, µ) is reflexive, then bT ∗φ = Cφ. Moreover, φ is µ-total (i.e. Ker Cφ = {0}) if and only if Ran bTφ is dense in D×. φ = bTφ since, for every f ∈ D, Proof : Vφ(X, µ)∗∗ → H [19, Sec.IV.7.4]. Let C ♯ [ξ]φ ∈ Vφ(X, µ), As Cφ : D → Vφ(X, µ)∗ is a continuous operator, it has a continuous adjoint C∗φ : ↾Vφ(X, µ). Then C ♯ φ := C∗φ (3.9) hCφf , [ξ]φi =ZX hf, φxiξ(x) dµ(x) = hf, bTφ[ξ]φi. φ = C∗φ = bTφ. If Vφ(X, µ) is reflexive, we have, of course, C ♯ If φ is not µ-total, then there exists f ∈ D, f 6= 0 such that (Cφf )(x) = 0 for a.e. x ∈ X. Hence, f ∈ (Ran bTφ)⊥ := {f ∈ D : hF f i = 0, ∀F ∈ Ran bTφ} by (3.9). Conversely, if φ is µ-total, as (Ran bTφ)⊥ = Ker Cφ = {0}, by the reflexivity of D and D×, it follows that Ran bTφ is dense in D×. ✷ In a way similar to what we have done above, we can define the space Vψ(X, µ), its topology, the residue classes [η]ψ, the operator Tψ, etc, replacing φ by ψ. Then, Vψ(X, µ) is a locally convex space. Theorem 3.6 Under the condition (3.1), every bounded linear functional F on Vφ(X, µ), i.e., F ∈ Vφ(X, µ)∗, can be represented as F ([ξ]φ) =ZX ξ(x)η(x) dµ(x), ∀ [ξ]φ ∈ Vφ(X, µ), (3.10) with η ∈ Vψ(X, µ). The residue class [η]ψ ∈ Vψ(X, µ) is uniquely determined. Proof : By Theorem 3.4, we have the representation F (ξ) =ZX ξ(x)hφx, gi dµ(x). It is easily seen that η(x) = hg, φxi ∈ Vψ(X, µ). It remains to prove uniqueness. Suppose that Then F (ξ) =ZX ZX ξ(x)η′(x) dµ(x). ξ(x)(η′(x) − η(x)) dµ(x) = 0. Now the function ξ(x) is arbitrary. Hence, taking in particular for ξ(x) the functions hf, ψxi, f ∈ D, we get [η]ψ = [η′]ψ. ✷ The lesson of the previous statements is that the map j : F ∈ Vφ(X, µ)∗ 7→ [η]ψ ∈ Vψ(X, µ) (3.11) 7 is well-defined and conjugate linear. On the other hand, j(F ) = j(F ′) implies easily F = F ′. Therefore Vφ(X, µ)∗ can be identified with a closed subspace of Vψ(X, µ) := {[ξ]ψ : ξ ∈ Vψ(X, µ)}, the conjugate space of Vψ(X, µ). Working in the framework of Hilbert spaces, as in Section 3.1, we proved in [8] that the spaces Vφ(X, µ)∗ and Vψ(X, µ) can be identified. The conclusion was that if (ψ, φ) is a reproducing pair, the spaces Vφ(X, µ) and Vψ(X, µ) are both Hilbert spaces, conjugate dual of each other with respect to the sesquilinear form (3.5). And if φ and ψ are also µ-total, then the converse statement holds true. In the present situation, however, a result of this kind cannot be proved with techniques similar to those adopted in [8], which are specific of Hilbert spaces. In particular, the condition (b), Sψ,φ ∈ GL(H), which was essential in the proof of [8, Lemma 3.11], is now missing, and it is not clear by what regularity condition it should replaced. However, assume that Ran bCψ,φ[k·kφ] = Vφ(X, µ)[k · kφ] and Ran bCφ,ψ[k·kψ] = Vψ(X, µ)[k · kψ], where we have defined the operator bCφ,ψ : H → Vψ(X, µ) by bCφ,ψf := [Cφf ]ψ and similarly for bCψ,φ. Then the proof of [8, Theorem 3.14] works and the same result may be obtained. This is, however, a strong and non-intuitive assumption. 4 Reproducing pairs and genuine pip-spaces In this section, we will consider the case where our measurable functions take their values in a genuine pip-space. However, for simplicity, we will restrict ourselves to a lattice of Banach spaces (LBS) or a lattice of Hilbert spaces (LHS) [4]. For the convenience of the reader, we have summarized in the Appendix the basic notions concerning LBSs and LHSs. Let (X, µ) be a locally compact, σ-compact measure space. Let VJ = {Vp, p ∈ J} be a LBS or a LHS of measurable functions with the property ξ ∈ Vp, η ∈ Vp =⇒ ξη ∈ L1(X, µ) and (4.1) (cid:12)(cid:12)(cid:12)(cid:12) ZX ξ(x)η(x) dµ(x)(cid:12)(cid:12)(cid:12)(cid:12) 6 kξkp kηkp. Thus the central Hilbert space is H := Vo = L2(X, µ) and the spaces Vp, Vp are dual of each other with respect to the L2 inner product. The partial inner product, which extends that of L2(X, µ), is denoted again by h··i. As usual we put V = Pp∈J Vp and V # = Tp∈J Vp. Thus ψ : X → V means that ψ : X → Vp for some p ∈ J. Example 4.1 A typical example is the lattice generated by the Lebesgue spaces Lp(R, dx), 1 6 p 6 ∞, with 1 p = 1 [4]. We shall discuss it in detail in Section 6. p + 1 We will envisage two approaches, depending whether the functions ψx themselves belong to V or rather the scalar functions Cψf . 4.1 Vector-valued measurable functions ψx This approach is the exact generalization of the one used in the RHS case. Let x ∈ X 7→ ψx, x ∈ X 7→ φx weakly measurable functions from X into V , where the latter is equipped with the weak topology σ(V, V #). More precisely, assume that ψ : X → Vp for some p ∈ J and φ : X → Vq for some q ∈ J, both weakly measurable. In that case, the analysis of Section 3.1 may be repeated verbatim, simply replacing D by V #, thus defining reproducing pairs. The problem with this approach is that, in fact, it does not exploit the pip-space structure, only the RHS V # ⊂ H ⊂ V ! Clearly, this approach yields no benefit, so we turn to a different strategy. 8 4.2 Scalar measurable functions Cψf Let ψ, φ be weakly measurable functions from X into H. In view of (2.10), (4.1) and the definition of V , we assume that the following condition holds: (p) ∃ p ∈ J such that Cψf = hf ψ·i ∈ Vp and Cφg = hgφ·i ∈ Vp, ∀ f, g ∈ H. We recall that Vp is the conjugate dual of Vp. In this case, then Ωψ,φ(f, g) :=ZX hf ψxihφxgi dµ(x), f, g ∈ H, defines a sesquilinear form on H × H and one has Ωψ,φ(f, g) 6 kCψf kp kCφgkp , ∀ f, g ∈ H. (4.2) If Ωψ,φ is bounded as a form on H × H (this is not automatic, see Corollary 4.4), there exists a bounded operator Sψ,φ in H such that ZX hf ψxihφxgi dµ(x) = hSψ,φf gi, ∀ f, g ∈ H. (4.3) Then (ψ, φ) is a reproducing pair if Sψ,φ ∈ GL(H). Let us suppose that the spaces Vp have the following property (k) If ξn → ξ in Vp, then, for every compact subset K ⊂ X, there exists a subsequence {ξK n } of {ξn} which converges to ξ almost everywhere in K. We note that condition (k) is satisfied by Lp-spaces [18]. As seen before, Cψ : H → V , in general. This means, given f ∈ H, there exists p ∈ J such that Cψf = hf ψ·i ∈ Vp. We define Dr(Cψ) = {f ∈ H : Cψf ∈ Vr}, r ∈ J. In particular, Dr(Cψ) = H means Cψ(H) ⊂ Vr. Proposition 4.2 Assume that (k) holds. Then Cψ : Dr(Cψ) → Vr is a closed linear map. Proof : Let fn → f in H and {Cψfn} be Cauchy in Vr. Since Vr is complete, there exists ξ ∈ Vr such that kCψfn − ξkr → 0. By (k), for every compact subset K ⊂ X, there exists a subsequence {f K n )(x) → ξ(x) a.e. in K. On the other hand, since fn → f in H, we get n } of {fn} such that (Cψf K hfnψxi → hf ψxi, ∀x ∈ X, and the same holds true, of course, for {f K everywhere. Thus, f ∈ Dr(Cψ) and ξ = Cψf . n }. From this we conclude that ξ(x) = hf ψxi almost ✷ By a simple application of the closed graph theorem we obtain Corollary 4.3 Assume that (k) holds. If for some r ∈ J, Cψ(H) ⊂ Vr, then Cψ : H → Vr is continuous. Combining Corollary 4.3 with (4.2), we get 9 Corollary 4.4 Assume that (k) holds. If Cψ(H) ⊂ Vp and Cψ(H) ⊂ Vp, the form Ω is bounded on H × H, that is, Ωψ,φ(f, g) 6 c kf k kgk. Hence, if condition (k) holds, Cψ(H) ⊂ Vr implies that Cψ : H → Vr is continuous. If we don't know whether the condition holds, we will have to assume explicitly that Cψ : H → Vr is continuous. If Cψ : H → Vr continuously, then C∗ψ : Vr → H exists and it is continuous. By definition, if ξ ∈ Vr, hf ψxiξ(x) dµ(x), ∀ f ∈ H. (4.4) hCψf ξi =ZX hf ψxiξ(x) dµ(x) = hf ZX The relation (4.4) then implies that ZX Thus, ψxξ(x) dµ(x)i, ∀ f ∈ H. C∗ψξ =ZX ψxξ(x) dµ(x). Of course, what we have said about Cψ holds in the very same way for Cφ. Assume now that for some p ∈ J, Cψ : H → Vp and Cφ : H → Vp continuously. Then, C∗φ : Vp → H so that C∗φCψ is a well-defined bounded operator in H. As before, we have Hence, C∗φη =ZX η(x)φx dµ(x), ∀ η ∈ Vp. C∗φCψf =ZX hf ψxiφx dµ(x) = Sψ,φf, ∀ f ∈ H, the last equality following also from (4.3) and Corollary 4.4. Of course, this does not yet imply that Sψ,φ ∈ GL(H), thus we don't know whether (ψ, φ) is a reproducing pair. Let us now return to the pre-Hilbert space Vφ(X, µ). First, the defining relation (3.3) of [8] must be written as ξ ∈ Vφ(X, µ) ⇔ (cid:12)(cid:12)(cid:12)(cid:12) ZX ξ(x)(Cφg)(x) dµ(x)(cid:12)(cid:12)(cid:12)(cid:12) 6 c kgk , ∀ g ∈ H. Since Cφ : H → Vp, the integral is well defined for all ξ ∈ Vp. This means, the inner product on is in fact the partial inner product of V , which coincides with the L2 inner product the r.h.s. whenever the latter makes sense. We may rewrite the r.h.s. as where h··i denotes the partial inner product. Next, by (4.1), one has, for ξ ∈ Vp, g ∈ H, hξCφgi 6 c kgk , ∀ g ∈ H, ξ ∈ Vp. hξCφgi 6 kξkp kCφgkp 6 c kξkp kgk , where the last inequality follows from Corollary 4.3 or the assumption of continuity of Cφ. Hence indeed ξ ∈ Vφ(X, µ), so that Vp ⊂ Vφ(X, µ). As for the adjoint operator, we have C∗φ : Vp → H. Then we may write, for ξ ∈ Vp, g ∈ H, hξCφgi = hTφξgi, thus C∗φ is the restriction from Vφ(X, µ) to Vp of the operator Tφ : Vφ → H introduced in Section 2, which reads now as hTφξgi =ZX ξ(x)hφxgi dµ(x), ∀ ξ ∈ Vp, g ∈ H. (4.5) 10 Thus C∗φ ⊂ Tφ. Next, the construction proceeds as in Section 3. The space Vφ(X, µ) = Vφ(X, µ)/Ker Tφ, with the norm k[ξ]φkφ = kTφξk, is a pre-Hilbert space. Then Theorem 3.14 and the other results from Section 3 of [8] remain true. In particular, we have: Theorem 4.5 If (ψ, φ) is a reproducing pair, the spaces Vφ(X, µ) and Vψ(X, µ) are both Hilbert spaces, conjugate dual of each other with respect to the sesquilinear form (2.9), namely, hhξηiiµ :=ZX ξ(x)η(x) dµ(x). Note the form (2.9) coincides with the inner product of L2(X, µ) whenever the latter makes sense. Let (ψ, φ) is a reproducing pair. Assume again that Cφ : H → Vp continuously, which me may write bCφ,ψ : H → Vp/Ker Tψ, where bCφ,ψ : H → Vψ(X, µ) is the operator defined by bCφ,ψf := [Cφf ]ψ, already introduced at the end of Section 3.2. In addition, by [8, Theorem 3.13], one has Ran bCψ,φ[k·kφ] = Vφ(X, µ)[k · kφ] and Ran bCφ,ψ[k·kψ] = Vψ(X, µ)[k · kψ]. Putting everything together, we get Corollary 4.6 Let (ψ, φ) be a reproducing pair. Then, if Cψ : H → Vp and Cφ : H → Vp continuously, one has bCφ,ψ : H → Vp/Ker Tψ = Vψ(X, µ) ≃ Vφ(X, µ)∗, bCψ,φ : H → Vp/Ker Tφ = Vφ(X, µ) ≃ Vψ(X, µ)∗. (4.6) (4.7) In these relations, the equality sign means an isomorphism of vector spaces, whereas ≃ denotes an isomorphism of Hilbert spaces. Proof : On one hand, we have Ran bCφ,ψ = Vψ(X, µ). On the other hand, under the assumption Cφ(H) ⊂ Vp, one has Vp ⊂ Vψ(X, µ), hence Vp/Ker Tψ = {ξ + Ker Tψ, ξ ∈ Vp} ⊂ Vψ(X, µ). Thus we get Vψ(X, µ) = Vp/Ker Tψ as vector spaces. Similarly Vφ(X, µ) = Vp/Ker Tφ. ✷ Notice that, in Condition (p), the index p cannot depend on f, g. We need some uniformity, in the form Cψ(H) ⊂ Vp and Cφ(H) ⊂ Vp . This is fully in line with the philosophy of pip-spaces: the building blocks are the (assaying) subspaces Vp, not individual vectors. 5 The case of a Hilbert triplet or a Hilbert scale 5.1 The general construction We have derived in the previous section the relations Vp ⊂ Vφ(X, µ), Vp ⊂ Vψ(X, µ), and their equivalent ones (4.6), (4.7). Then, since Vψ(X, µ) and Vφ(X, µ) are both Hilbert spaces, it seems natural to take for Vp, Vp Hilbert spaces as well, that is, take for V a LHS. The simplest case is then a Hilbert chain, for instance, the scale (A.3) {Hk, k ∈ Z} built on the powers of a self-adjoint operator A > I . This situation is quite interesting, since in that case one may get results about spectral properties of symmetric operators (in the sense of pip-space operators) [9]. Thus, let (ψ, φ) be a reproducing pair. For simplicity, we assume that Sψ,φ = I, that is, ψ, φ are dual to each other. If ψ and φ are both frames, there is nothing to say, since then Cψ(H), Cφ(H) ⊂ L2(X, µ) = Ho, so that there is no need for a Hilbert scale. Thus we assume that ψ is an upper semi-frame and 11 φ is a lower semi-frame, dual to each other. It follows that Cψ(H) ⊂ L2(X, µ). Hence Condition (p) becomes: There is an index k > 1 such that if Cψ : H → Hk and Cφ : H → Hk continuously, thus Vp ≡ Hk and Vp ≡ Hk. This means we are working in the Hilbert triplet Vp ≡ Hk ⊂ Ho = L2(X, µ) ⊂ Hk ≡ Vp . (5.1) Next, according to Corollary 4.6, we have Vψ(X, µ) = Hk/Ker Tψ and Vφ(X, µ) = Hk/Ker Tφ, as vector spaces. In addition, since φ is a lower semi-frame, [6, Lemma 2.1] tells us that Cφ has closed range in L2(X, µ) and is injective. However its domain D(Cφ) := {f ∈ H :ZX hf φxi2 dν(x) < ∞} need not be dense, it could be {0}. Thus Cφ maps its domain D(Cφ) onto a closed subspace of L2(X, µ), possibly trivial, and the whole of H into the larger space Hk. 5.2 Examples As for concrete examples of such Hilbert scales, we might mention two. First the Sobolev spaces H k(R), k ∈ Z, in H0 = L2(R, dx), which is the scale generated by the powers of the self-adjoint operator A1/2, where A := 1− d2 dx2 . The other one corresponds to the quantum harmonic oscillator, with Hamiltonian Aosc := x2 − d2 dx2 . The spectrum of Aosc is {2n + 1, n = 0, 1, 2, . . .} and it gets diagonalized on the basis of Hermite functions. It follows that A−1 osc, which maps every Hk onto Hk−1, is a Hilbert-Schmidt operator. Therefore, the end space of the scale D∞(Aosc) := Tk Hk, which is simply Schwartz' space S of C∞ functions of fast decrease, is a nuclear space. Actually one may give an explicit example, using a Sobolev-type scale. Let HK be a repro- ducing kernel Hilbert space (RKHS) of (nice) functions on ameasure space (X, µ), with kernel function kx, x ∈ X, that is, f (x) = hf kxiK , ∀f ∈ HK. The corresponding reproducing kernel is K(x, y) = ky(x) = hkykxiK . Choose the weight function m(x) > 1, the analog of the weight (1 + x2) considered in the Sobolev case. Define the Hilbert scale Hk, k ∈ Z, determined by the multiplication operator Af (x) = m(x)f (x), ∀x ∈ X. Hence, for each l > 1, Hl ⊂ H0 ≡ HK ⊂ Hl . Then, for some n > 1, define the measurable functions φx = kxmn(x), ψx = kxm−n(x), so that Cψ : HK → Hn, Cφ : HK → Hn continuously, and they are dual of each other. Thus (ψ, φ) is a reproducing pair, with ψ an upper semi-frame and φ a lower semi-frame. In this case, one can compute the operators Tψ, Tφ explicitly. The definition (4.5) reads as hTφξgiK =ZX ξ(x)hφxgiK dµ(x), ∀ ξ ∈ Hn, g ∈ HK. Now hφxgiK = hkxmn(x)giK = hkxg mn(x)iK = g(x) mn(x) ∈ Hn. Thus hTφξgiK =ZX ξ(x) g(x) mn(x) dµ, that is, (Tφξ)(x) = ξ(x) mn(x) or Tφξ = ξ mn. However, since the weight m(x) > 1 is invertible, g mn runs over the whole of Hn whenever g runs over HK. Hence ξ ∈ Ker Tφ ⊂ Hn means that hTφξgiK = 0, ∀g ∈ HK, which implies ξ = 0, since the duality between Hn and Hn is separating. The same reasoning yields Ker Tψ = {0}. Therefore Vφ(X, µ) = Hn and Vψ(X, µ) = Hn. 12 A more general situation may be derived from the discrete example of Section 6.1.3 of [8]. m with norm Take a sequence of weights m := {mn}n∈N ∈ c0, mn 6= 0, and consider the space ℓ2 kξkℓ2 :=Pn∈N mnξn2. Then we have the following triplet replacing (5.1) m ℓ2 1/m ⊂ ℓ2 ⊂ ℓ2 m. (5.2) (5.3) Next, for each n ∈ N, define ψn = mnθn, where θ is a frame or an orthonormal basis in ℓ2. Then ψ is an upper semi-frame. Moreover, φ := {(1/mn)θn}n∈N is a lower semi-frame, dual to ψ, thus (ψ, φ) is a reproducing pair. Hence, by [8, Theorem 3.13], Vψ ≃ Ran Cφ = M1/m(Vθ(N)) = ℓ2 m and Vφ ≃ Ran Cψ = Mm(Vθ(N)) = ℓ2 1/m (here we take for granted that Ker Tψ = Ker Tφ = {0}). For making contact with the situation of (5.1), consider in ℓ2 the diagonal operator A := diag[n], n ∈ N, that is (Aξ)n = n ξn, n ∈ N, which is obviously self-adjoint and larger than 1. r(k), where (r(k))n = nk (note that 1/r(k) ∈ c0). Then Hk = D(Ak) with norm kξkk = (cid:13)(cid:13)Akξ(cid:13)(cid:13) ≡ ℓ2 Hence we have Hk = ℓ2 r(k) ⊂ Ho = ℓ2 ⊂ Hk = ℓ2 1/r(k) , where (1/r(k))n = n−k. In addition, as in the continuous case discussed above, the end space of the scale, D∞(A) := Tk Hk, is simply Schwartz' space s of fast decreasing sequences, with dual D∞(A) :=Sk Hk = s′, the space of slowly increasing sequences. Here too, this construction shows that the space s is nuclear, since every embedding A−1 : Hk+1 → Hk is a Hilbert-Schmidt operator. However, the construction described above yields a much more general family of examples, since the weight sequences m are not ordered. 6 The case of Lp spaces Following the suggestion made at the end of Section 2, we present now several possibilities of taking Ran Cψ in the context of the Lebesgue spaces Lp(R, dx). As it is well-known, these spaces don't form a chain, since two of them are never comparable. We have only Lp ∩ Lq ⊂ Ls, for all s such that p < s < q. Take the lattice J generated by I = {Lp(R, dx), 1 6 p 6 ∞}, with lattice operations [4, Sec.4.1.2]: • Lp ∧ Lq = Lp ∩ Lq is a Banach space for the projective norm kf kp∧q = kf kp + kf kq • Lp ∨ Lq = Lp + Lq is a Banach space for the inductive norm kf kp∨q = inff =g+h {kgkp + khkq; g ∈ Lp, h ∈ Lq} • For 1 < p, q < ∞, both spaces Lp ∧ Lq and Lp ∨ Lq are reflexive and (Lp ∧ Lq)× = Lp ∨ Lq. Moreover, no additional spaces are obtained by iterating the lattice operations to any finite order. Thus we obtain an involutive lattice and a LBS, denoted by VJ. It is convenient to introduce a unified notation: L(p,q) =(cid:26) Lp ∧ Lq = Lp ∩ Lq, Lp ∨ Lq = Lp + Lq, if p > q, if p 6 q. Following [4, Sec.4.1.2], we represent the space L(p,q) by the point (1/p, 1/q) of the unit square J = [0, 1] × [0, 1]. In this representation, the spaces Lp are on the main diagonal, intersections Lp ∩ Lq above it and sums Lp + Lq below, the duality is [L(p,q)]× = L(p,q), that is, symmetry with 13 1/q ✻ • ❅ L∞ ∩ L1 L∞ ∩ L(t) L∞ ∩ L2 L∞ ∩ L(t) • • • L2 ∩ L1 • L1 • ❅ ❅ ❅ ❅ L(s) • Lq • ❅ ❅ ❅ Lq ∩ Lq ❅ • L(t) • ✁✁ ✁ ✁ ✁ ✁ • ❅ ✁ ❅ ✁ L2 ✁ ✁ ❍❍❍❍❍❍❍❍❍❍❍❍ ❅ ❅ ❅ ❅ • ✁ ✁ • L(t) Lq • ❅ ❅ • ❅ Lq + Lq = (Lq ∩ Lq)× ❅ ❅ ❅ L1 + L2 • L(s) • L∞ • L2 + L∞ ❅ ❅ • L1 + L∞ ✲ 1/p Figure 1: (i) The pair L(s), L(s) for s in the second quadrant; (ii) The pair L(t), L(t) for t in the first quadrant. respect to L2. Hence, L(p,q) ⊂ L(p′,q′) if (1/p, 1/q) is on the left and/or above (1/p′, 1/q′) The extreme spaces are 2 V # J = L(∞,1) = L∞ ∩ L1, and VJ = L(1,∞) = L1 + L∞. For a full picture, see [4, Fig.4.1]. There are three possibilities for using the Lp lattice for controlling reproducing pairs (1) Exploit the full lattice J , that is, find (p, q) such that, ∀f, g ∈ H, Cψf # Cφg in the pip-space VJ, that is, Cψf ∈ L(p,q) and Cφg ∈ L(p,q). (2) Select in VJ a self-dual Banach chain VI, centered around L2, symbolically. . . . L(s) ⊂ . . . ⊂ L2 ⊂ . . . ⊂ L(s) . . . , (6.1) such that Cψf ∈ L(s) and Cφg ∈ L(s) (or vice-versa). Here are three examples of such Banach chains. • The diagonal chain : q = p L∞ ∩ L1 ⊂ . . . ⊂ Lq ∩ Lq ⊂ . . . ⊂ L2 ⊂ . . . ⊂ Lq + Lq = (Lq ∩ Lq)× ⊂ . . . ⊂ L1 + L∞. 2The space L1 + L∞ has been considered by Gould [12]. 14 • The horizontal chain q = 2 : L∞ ∩ L2 ⊂ . . . ⊂ L2 ⊂ . . . ⊂ L1 + L2. • The vertical chain p = 2 : L2 ∩ L1 ⊂ . . . ⊂ L2 ⊂ . . . ⊂ L2 + L∞. All three chains are presented in Figure 1. In this case, the full chain belongs to the second and fourth quadrants (top left and bottom right). A typical point is then s = (p, q) with, 2 6 p 6 ∞, 1 6 q 6 2, so that one has the situation depicted in (6.1), that is, the spaces L(s), L(s) to which Cψf , resp. Cφg, belong, are necessarily comparable to each other and to L2. In particular, one of them is necessarily contained in L2. Note the extreme spaces of that type are L2, L∞ ∩L2, L∞ ∩L1 and L2 ∩ L1 (see Figure 1). (3) Choose a dual pair in the first and third quadrant (top right, bottom left). A typical point is then t = (p′, q′), with 1 < p′, q′ < 2, so that the spaces L(t), L(t) are never comparable to each other, nor to L2. Let us now add the uniform boundedness condition mentioned at the end of Section 2, supx∈X kψxkH 6 c and supx∈X kφxkH 6 c′ for some c, c′ > 0. Then Cψf (x) = hf ψxi ∈ L∞(X, dµ) and Cφf (x) = hf φxi ∈ L∞(X, dµ). Therefore, the third case reduces to the second one, since we have now (in the situation of Figure 1). L∞ ∩ L(t) ⊂ L∞ ∩ L2 ⊂ L∞ ∩ L(t). Following the pattern of Hilbert scales, we choose a (Gel'fand) triplet of Banach spaces. One could have, for instance, a triplet of reflexive Banach spaces such as L(s) ⊂ . . . ⊂ L2 ⊂ . . . ⊂ L(s), (6.2) corresponding to a point s inside of the second quadrant, as shown in Figure 1. according to (4.6) and (4.7), Vψ = L(s)/Ker Tψ and Vφ = L(s)/Ker Tφ. In this case, On the contrary, if we choose a point t in the second quadrant, case (3) above, it seems that no triplet arises. However, if (ψ, φ) is a nontrivial reproducing pair, with Sφ,ψ = I, that is, ψ, φ are dual to each other, one of them, say ψ, is an upper semi-frame and then necessarily φ is a lower semi-frame [6, Prop.2.6]. Therefore Cψ(H) ⊂ L2(X, µ), that is, case (3) cannot be realized. Inserting the boundedness condition, we get a triplet where the extreme spaces are no longer reflexive, such as L∞ ∩ L(t) ⊂ L∞ ∩ L2 ⊂ L∞ ∩ L(t), and then Vψ = (L∞ ∩ L(t))/Ker Tψ and Vφ = (L∞ ∩ L(t))/Ker Tφ. In conclusion, the only acceptable solution is the triplet (6.2), with s strictly inside of the second quadrant, that is, s = (p, q) with, 2 6 p < ∞, 1 < q 6 2. A word of explanation is in order here, concerning the relations Vψ = L(s)/Ker Tψ and Vφ = L(s)/Ker Tφ. On the l.h.s., L(s) and L(s) are reflexive Banach spaces, with their usual norm, and so are the quotients by Tψ, resp. Tφ. On the other hand, Vψ(X, µ)[k · kψ] and Vφ(X, µ)[k · kφ] are Hilbert spaces. However, there is no contradiction, since the equality sign = denotes an isomorphism of vector spaces only, without reference to any topology. Moreover, the two norms, Banach and Hilbert, cannot be comparable, lest they are equivalent [16, Coroll. 1.6.8], which is impossible in the case of Lp, p 6= 2. The same is true for any LBS where the spaces Vp are not Hilbert spaces. 15 Although we don't have an explicit example of a reproducing pair, we indicate a possible x i ∈ x i ∈ . construction towards one. Let θ(1) : R → L2 be a measurable function such that hhθ(1) Lq, ∀ h ∈ L2, 1 < q < 2 and let θ(2) : R → L2 be a measurable function such that hhθ(2) Lq, ∀ h ∈ L2. Define ψx := min(θ(1) Then we have and φx := max(θ(1) x , θ(2) x ) ≡ θ(1) x ∧ θ(2) x x , θ(2) x ) ≡ θ(1) x ∨ θ(2) x (Cψh)(x) = hhψxi ∈ Lq ∩ Lq, ∀ h ∈ L2 (Cφh)(x) = hhφxi ∈ Lq + Lq, ∀ h ∈ L2 and we have indeed Lq ∩ Lq ⊂ L2 ⊂ Lq + Lq. It remains to guarantee that ψ and φ are dual to each other, that is, ZX hf ψxihφxgi dµ(x) =ZX Cψf (x) Cφg(x) dµ(x) = hf gi, ∀ f, g ∈ L2. 7 Outcome We have seen in [8] that the notion of reproducing pair is quite rich. It generates a whole mathematical structure, which ultimately leads to a pair of Hilbert spaces, conjugate dual to each other with respect to the L2(X, µ) inner product. This suggests that one should make more precise the best assumptions on the measurable functions or, more precisely, on the nature of the range of the analysis operators Cψ, Cφ. This in turn suggests to analyze the whole structure in the language of pip-spaces, which is the topic of the present paper. In particular, a natural choice is a scale, or simply a triplet, of Hilbert spaces, the two extreme spaces being conjugate duals of each other with respect to the L2(X, µ) inner product. Another possibility consists of exploiting the lattice of all Lp(R, dx) spaces, or a subset thereof, in particular a (Gel'fand) triplet of Banach spaces. Some examples have been described above, but clearly more work along these lines is in order. Appendix. Lattices of Banach or Hilbert spaces and operators on them A.1 Lattices of Banach or Hilbert spaces For the convenience of the reader, we summarize in this Appendix the basic facts concerning pip-spaces and operators on them. However, we will restrict the discussion to the simpler case of a lattice of Banach (LBS) or Hilbert spaces (LHS). Further information may be found in our monograph [4] or our review paper [5]. Let thus J = {Vp, p ∈ I} be a family of Hilbert spaces or reflexive Banach spaces, partially or- dered by inclusion. Then I generates an involutive lattice J , indexed by J, through the operations (p, q, r ∈ I): Vr ↔ Vr = V ×r , the conjugate dual of Vr . involution: Vp∧q := Vp ∧ Vq = Vp ∩ Vq . infimum: . supremum: Vp∨q := Vp ∨ Vq = Vp + Vq. It turns out that both Vp∧q and Vp∨q are Hilbert spaces, resp. reflexive Banach spaces, under appropriate norms (the so-called projective, resp. inductive norms). Assume that the following conditions are satisfied: (i) I contains a unique self-dual, Hilbert subspace Vo = Vo. 16 (ii) for every Vr ∈ I, the norm k · kr on Vr = V ×r In addition to the family J = {Vr, r ∈ J}, it is convenient to consider the two spaces V # and V defined as is the conjugate of the norm k · kr on Vr. Vq. (A.1) V =Xq∈I Vq, V # = \q∈I These two spaces themselves usually do not belong to I. We say that two vectors f, g ∈ V are compatible if there exists r ∈ J such that f ∈ Vr, g ∈ Vr . Then a partial inner product on V is a Hermitian form h··i defined exactly on compatible pairs of vectors. In particular, the partial inner product h··i coincides with the inner product of Vo on the latter. A partial inner product space (pip-space) is a vector space V equipped with a partial inner product. Clearly LBSs and LHSs are particular cases of pip-spaces. From now on, we will assume that our pip-space (V, h··i) is nondegenerate, that is, hf gi = 0 for all f ∈ V # implies g = 0. As a consequence, (V #, V ) and every couple (Vr, Vr), r ∈ J, are a dual pair in the sense of topological vector spaces [14]. In particular, the original norm topology on Vr coincides with its Mackey topology τ (Vr, Vr), so that indeed its conjugate dual is (Vr)× = Vr, ∀ r ∈ J. Then, r < s implies Vr ⊂ Vs, and the embedding operator Esr : Vr → Vs is continuous and has dense range. In particular, V # is dense in every Vr. In the sequel, we also assume the partial inner product to be positive definite, hf f i > 0 whenever f 6= 0. A standard, albeit trivial, example is that of a Rigged Hilbert space (RHS) Φ ⊂ H ⊂ Φ# (it is trivial because the lattice I contains only three elements). Familiar concrete examples of pip-spaces are sequence spaces, with V = ω the space of all loc(R, dx), the complex sequences x = (xn), and spaces of locally integrable functions with V = L1 space of Lebesgue measurable functions, integrable over compact subsets. Among LBSs, the simplest example is that of a chain of reflexive Banach spaces. The prototype is the chain I = {Lp := Lp([0, 1]; dx), 1 < p < ∞} of Lebesgue spaces over the interval [0, 1]. L∞ ⊂ . . . ⊂ Lq ⊂ Lr ⊂ . . . ⊂ L2 ⊂ . . . ⊂ Lr ⊂ Lq ⊂ . . . ⊂ L1, (A.2) where 1 < q < r < 2 (of course, L∞ and L1 are not reflexive). Here Lq and Lq are dual to each other (1/q + 1/q = 1), and similarly Lr, Lr (1/r + 1/r = 1). As for a LHS, the simplest example is the Hilbert scale generated by a self-adjoint operator A > I in a Hilbert spaceHo. Let Hn be D(An), the domain of An, equipped with the graph norm kf kn = kAnf k, f ∈ D(An), for n ∈ N or n ∈ R+, and Hn := H−n = H×n (conjugate dual): D∞(A) :=\n Hn ⊂ . . . ⊂ H2 ⊂ H1 ⊂ H0 ⊂ H1 ⊂ H2 . . . ⊂ D∞(A) :=[n Hn. (A.3) Note that here the index n may be integer or real, the link between the two cases being established by the spectral theorem for self-adjoint operators. Here again the inner product of H0 extends to each pair Hn, H−n, but on D∞(A) it yields only a partial inner product. A standard example is the scale of Sobolev spaces H s(R), s ∈ Z, in H0 = L2(R, dx). A.2 Operators on LBSs and LHSs Let VJ be a LHS or a LBS. Then an operator on VJ is a map from a subset D(A) ⊂ V into V , such that (i) D(A) =Sq∈d(A) Vq, where d(A) is a nonempty subset of J; (ii) For every q ∈ d(A), there exists p ∈ J such that the restriction of A to Vq is a continuous linear map into Vp (we denote this restriction by Apq); 17 (iii) A has no proper extension satisfying (i) and (ii). We denote by Op(VJ , ) the set of all operators on VJ . The continuous linear operator Apq : Vq → Vp is called a representative of A. The properties of A are conveniently described by the set j(A) of all pairs (q, p) ∈ J × J such that A maps Vq continuously into Vp Thus the operator A may be identified with the collection of its representatives, A ≃ {Apq : Vq → Vp : (q, p) ∈ j(A)}. (A.4) It is important to notice that an operator is uniquely determined by any of its representatives, in virtue of Property (iii): there are no extensions for pip-space operators. We will also need the following sets: d(A) = {q ∈ J : there is a p such that Apq exists}, i(A) = {p ∈ J : there is a q such that Apq exists}. The following properties are immediate: . d(A) is an initial subset of J: if q ∈ d(A) and q′ < q, then q′ ∈ d(A), and Apq′ = ApqEqq′, where Eqq′ is a representative of the unit operator. . i(A) is a final subset of J: if p ∈ i(A) and p′ > p, then p′ ∈ i(A) and Ap′q = Ep′pApq. Although an operator may be identified with a separately continous sesquilinear form on V # × V #, or a conjugate linear continuous map V # into V , it is more useful to keep also the algebraic operations on operators, namely: (i) Adjoint: every A ∈ Op(VJ ) has a unique adjoint A× ∈ OpVJ ), defined by hA×yxi = hyAxi, for x ∈ Vq, q ∈ d(A) and y ∈ Vp, p ∈ i(A), (A.5) that is, (A×)qp = (Apq)′, where (Apq)′ : Vp → Vq is the adjoint map of Apq. Furthermore, one has A×× = A, for every A ∈ Op(VJ ): no extension is allowed, by the maximality condition (iii) of the definition. (ii) Partial multiplication: Let A, B ∈ Op(VJ ). We say that the product BA is defined if and only if there is a r ∈ i(A) ∩ d(B), that is, if and only if there is a continuous factorization through some Vr: Vq A→ Vr B→ Vp, i.e., (BA)pq = BprArq, for some q ∈ d(A), p ∈ i(B). (A.6) Of particular interest are symmetric operators, defined as those operators satisfying the relation A× = A, since these are the ones that could generate self-adjoint operators in the central Hilbert space, for instance by the celebrated KLMN theorem, suitably generalized to the pip-space envi- ronment [4, Section 3.3]. Acknowledgement This work was partly supported by the Istituto Nazionale di Alta Matematica (GNAMPA project "Propriet`a spettrali di quasi *-algebre di operatori"). JPA acknowledges gratefully the hospitality of the Dipartimento di Matematica e Informatica, Universit`a di Palermo, whereas CT acknowl- edges that of the Institut de Recherche en Math´ematique et Physique, Universit´e catholique de Louvain. 18 References [1] S.T. Ali, J-P. Antoine, and J-P. Gazeau, Square integrability of group representations on homogeneous spaces I. Reproducing triples and frames, Ann. Inst. H. Poincar´e 55 (1991) 829 -- 856 [2] S.T. Ali, J-P. Antoine, and J-P. Gazeau, Continuous frames in Hilbert space, Annals of Physics 222 (1993) 1 -- 37 [3] J-P. Antoine, A. Inoue, and C. Trapani, Partial *-Algebras and Their Operator Realizations, Mathematics and Its Applications, vol. 553, Kluwer, Dordrecht, NL, 2002 [4] J-P. Antoine and C. Trapani, Partial Inner Product Spaces: Theory and Applications, Lecture Notes in Mathematics, vol. 1986, Springer-Verlag, Berlin, 2009 [5] J-P. Antoine and C. Trapani The partial inner product space method: A quick overview, Adv. in Math. Phys., Vol. 2010 (2010) 457635 ; Erratum, Ibid. Vol. 2011 (2010) 272703. [6] J-P. Antoine and P. Balazs, Frames and semi-frames, J. Phys. A: Math. Theor. 44 (2011) 205201; Corrigendum, ibid. 44 (2011) 479501 [7] J-P. Antoine and P. Balazs, Frames, semi-frames, and Hilbert scales, Numer. Funct. Anal. Optimiz. 33 (2012) 736 -- 769 [8] J-P. Antoine, M. Speckbacher and C. Trapani, Reproducing pairs of measurable functions, preprint, 2016 [9] J-P. Antoine and C. Trapani, Operators on partial inner product spaces: Towards a spectral analysis, Mediterranean J. Math.Mediterranean J. Math. 13(2016) 323-351 [10] O. Christensen, An Introduction to Frames and Riesz Bases, Birkhauser, Boston, MA, 2003 [11] J-P. Gabardo and D. Han, Frames associated with measurable spaces, Adv. Comput. Math. 18 (2003) 127 -- 147 [12] G.G. Gould, On a class of integration spaces, J. London Math. Soc. 34 (1959), 161 -- 172 [13] G. Kaiser A Friendly Guide to Wavelets, Birkhauser, Boston, 1994 [14] G. Kothe, Topological Vector Spaces, Vol. I , Springer-Verlag, Berlin, 1969, 1979 [15] G. Kothe, Topological Vector Spaces, Vol. II , Springer-Verlag, Berlin, 1979; p.84 [16] R.E. Megginson, An Introduction to Banach Space Theory, Springer-Verlag, New York- Heidelberg-Berlin, 1998 [17] A. Rahimi, A. Najati and Y.N. Dehghan, Continuous frames in Hilbert spaces, Methods Funct. Anal. Topol. 12 (2006) 170 -- 182 [18] W. Rudin, Real and Complex Analysis, Int. Edition, McGraw Hill, New York et al., 1987; p.73, from Ex.18 [19] H.H. Schaefer, Topological Vector Spaces, Springer-Verlag, New York-Heidelberg-Berlin, 1971 19 [20] M. Speckbacher and P. Balazs, Reproducing pairs and the continuous nonstationary Gabor transform on LCA groups, J. Phys. A: Math. Theor., 48 (2015) 395201 [21] P. Werner, A distributional-theoretical approach to certain Lebesgue and Sobolev spaces, J. Math. Anal. Appl. 29 (1970) 18 -- 78 20
1712.06177
2
1712
2019-04-17T00:54:21
Homological dimensions of analytic Ore extensions
[ "math.FA" ]
If $A$ is an algebra with finite right global dimension, then for any automorphism $\alpha$ and $\alpha$-derivation $\delta$ the right global dimension of $A[t; \alpha, \delta]$ satisfies \[ \text{rgld} \, A \le \text{rgld} \, A[t; \alpha, \delta] \le \text{rgld} \, A + 1. \] We extend this result to the case of holomorphic Ore extensions and smooth crossed products by $\mathbb{Z}$ of $\hat{\otimes}$-algebras.
math.FA
math
Homological dimensions of analytic Ore extensions Petr Kosenko April 18, 2019 Abstract If A is an algebra with finite right global dimension, then for any automorphism α and α-derivation δ the right global dimension of A[t; α, δ] satisfies rgld A ≤ rgld A[t; α, δ] ≤ rgld A + 1. We extend this result to the case of holomorphic Ore extensions and smooth crossed products by Z of ⊗-algebras. 1 Introduction We will start this paper by recalling the following well-known theorem: Theorem 1.1 ([Wei94], Theorem 4.3.7). If R is a ring then the following estimate takes place for every n ∈ N: rgld R[x1, . . . xn] = n + rgld R. The importance of this theorem lies in the fact that it immediately yields the Hilbert's syzygy theorem in the case when R is a field (see [[Wei94], Corollary 4.3.8]). This fact can be, indeed, generalized to Ore extensions R[t; α, δ], as shown in [MR01]. It turns out that if the global dimension of R is finite, then the global dimension of R[t; α, δ] either stays the same, or increases by one. In this paper we adapt the arguments used in [MR01, ch. 7.5] to the topological setting in order to obtain the estimates for the right homological dimensions of holomorphic(topological) Ore extensions (see [Pir08a, ch. 4.1]) and smooth crossed products by Z (see [Sch93] and [PS94]). Below we state the result in the purely algebraic situation, which is provided in [MR01] and then we present its topological version. Remark. There is an ambiguity in defining Ore extensions, which will be demonstrated below, so, to state the result in the algebraic setting, we need to fix an appropriate definition of Ore extensions: Definition 1.1. Let A be an algebra, α ∈ End(A) and let δ : A → A be a C-linear map, such that the following relation holds for every a, b ∈ A: δ(ab) = δ(a)b + α(a)δ(b). Let us call such maps α-derivations. Then the Ore extension of A w.r.t α and δ is the vector space A[t; α, δ] =( nXi=0 aiti : ai ∈ A) with the multiplication defined uniquely by the following conditions: (1) The relation ta = α(a)t + δ(a) holds for any a ∈ A. (2) The natural inclusions A ֒→ A[t; α, δ] and C[t] ֒→ A[t; α, δ] are algebra homomorphisms. 1 Also, if δ = 0 and α is invertible, then one can define the Laurent Ore extension of A A[t, t−1; α] =( nXi=−n aiti : ai ∈ A) with the multiplication defined the same way. The thing is, the authors of [MR01] denote slightly different type of algebras by A[t; α, δ]: Definition 1.2 ([MR01], pp. 1.2.1-1.2.6). Let A be an algebra, α ∈ End(A) and let δ : A → A be a C-linear map, such that the following relation holds for every a, b ∈ A: δ(ab) = δ(a)α(b) + aδ(b). Let us call such maps opposite α-derivations. Then the (opposite) Ore extension of A w.r.t α and δ is the vector space Aop[t; α, δ] =( nXi=0 tiai : ai ∈ A) with the multiplication defined uniquely by the following conditions: (1) The relation at = t α(a) + δ(a) holds for any a ∈ A. (2) The natural inclusions A ֒→ Aop[t; α, δ] and C[t] ֒→ Aop[t; α, δ] are algebra homomorphisms. Also, if δ = 0 and α is invertible, then one can define the Laurent Ore extension of A Aop[t, t−1; α] =( nXi=−n tiai : ai ∈ A) with the multiplication defined the same way. It is easily seen that in the case of invertible α, the following algebra isomorphisms take place: A[x; α, δ] ∼= Aop[x, α−1, −δα−1], A[x, x−1; α] ∼= Aop[x, x−1; α−1]. (1) Throughout the paper, we will work with Ore extensions in the sense of Definition 1.1 (if not stated otherwise). Now we are ready to state the result in the purely algebraic case, which is, essentially, contained in the [MR01, Theorem 5.7.3]: Theorem 1.2 ([MR01], Theorem 5.7.3). Let A be an algebra, let σ be an automorphism and let δ be a σ-derivation, in the sense of a Definition 1.2. Denote the right global dimension of a ring R by dgr(R). Then the following estimates hold: (1) dgr A ≤ dgr Aop[t; σ, δ] ≤ dgr A + 1 if dgr R < ∞ (2) dgr A ≤ dgr Aop[t, t−1; σ] ≤ dgr A + 1 (3) dgr Aop[t, σ] = dgr A + 1 (4) dgr A[t, t−1] = dgr A + 1 Remark. In fact, the above theorem still holds if we replace Aop[t; σ, δ] with A[t; σ, δ] due to (1). This paper is organized as follows: in the Section 2 we recall the important notions related to homological properties of topological modules, in particular, we provide definitions of homological dimensions for topological algebras and modules. In the Section 3 we compute the estimates for the homological dimensions of holomorphic Ore extensions; we use the bimodules of relative differentials to construct the required projective resolutions. In the Section 4 we compute the estimates for the smooth crossed products by Z. In the Appendix A we provide the computations of algebraic and topological bimodules of relative differentials for different types of Ore extensions. 2 Acknowledgments I want to thank Alexei Pirkovskii for his assistance in writing this paper, Alexei's comments and ideas were very helpful and greatly improved this paper. 2 Homological dimensions Remark. All algebras in this paper are considered to be complex, unital and associative. Also, we will be working only with unital modules. 2.1 Notation Let us introduce some notation (see [Hel86a] and [Pir12] for more details). For locally convex Hausdorff spaces E, F we denote the completed projective tensor product of locally convex spaces E, F by E ⊗F . Definition 2.1. A Fr´echet space is a complete metrizable locally convex space. In other words, a locally convex space X is a Fr´echet space if and only if the topology on X can be generated by a countable family of seminorms. Denote by LCS, Fr the categories of complete locally convex spaces and Fr´echet spaces, respec- tively. Also we will denote the category of vector spaces by Lin. For a detailed introduction to the theory of locally convex spaces and algebras, and relevant examples, the reader can see [Tr´e70], [Jar81], [Mal86], or [Hel06]. Definition 2.2. Let A be a locally convex space with a multiplication µ : A × A → A, such that (A, µ) is an algebra. (1) If µ is separately continuous, then A is called a locally convex algebra. (2) If A is a complete locally convex space, and µ is jointly continuous, then A is called a ⊗-algebra. If A, B are ⊗-algebras and η : A → B is a continuous unital algebra homomorphism, then the pair (B, η) is called a A- ⊗-algebra. A ⊗-algebra with the underlying locally compact space which is a Fr´echet space is called a Fr´echet algebra. Definition 2.3. A locally convex algebra A is called m-convex if the topology on it can be defined by a family of submultiplicative seminorms. Definition 2.4. A complete locally m-convex algebra is called an Arens-Michael algebra. Definition 2.5. Let A be a ⊗-algebra and let M be a complete locally convex space with a structure of a topological A-module w.r.t. the locally convex topology on M . Also suppose that the natural map A × M → M is jointly continuous. Then we will call M a left A- ⊗-module. In a similar fashion we define right A- ⊗-modules and A-B- ⊗-bimodules. A ⊗-module over a Fr´echet algebra which is itself a Fr´echet space is called a Fr´echet A- ⊗-module. For ⊗-algebras A, B we denote A-mod = the category of (unital) left A- ⊗-modules, mod-A = the category of (unital) right A- ⊗-modules, A-mod-B = the category of (unital) A-B- ⊗-bimodules. More generally, for a fixed category C ⊆ LCS we denote A-mod(C) = the category of left (unital) A- ⊗-modules whose underyling LCS belong to C, mod-A(C) = the category of right (unital) A- ⊗-modules whose underyling LCS belong to C, A-mod-B(C) = the category of (unital) A-B- ⊗-bimodules whose underyling LCS belong to C. 3 Complexes of A- ⊗-modules for a ⊗-algebra A . . . dn+1−−−→ Mn+1 dn−→ Mn dn−1−−−→ Mn−1 dn−2−−−→ . . . are usually denoted by {M, d}. Definition 2.6. Let A be a ⊗-algebra and consider a left A- ⊗-module Y and a right A- ⊗-module X. (1) A bilinear map f : X × Y −→ Z, where Z ∈ LCS, is called A-balanced if f (x ◦ a, y) = f (x, a ◦ y) for every x ∈ X, y ∈ Y, a ∈ A. (2) A pair (X ⊗AY, i), where X ⊗AY ∈ LCS, and i : X ⊗AY −→ X ⊗AY is a continuous A-balanced map, is called the completed projective tensor product of X and Y , if for every Z ∈ LCS and continuous A-balanced map f : X × Y −→ Z there exists a unique continuous linear map f : X ⊗AY −→ Z such that f = f ◦ i. For the proof of the existence and uniqueness of complete projective tensor products of ⊗-modules the reader can see [Hel86a, ch. 2.3-2.4]. In this paper we would like to keep in mind a trivial, but nonetheless useful example: Example 2.1. Let A be a ⊗-algebra, and consider a left A- ⊗-module M . Then A ⊗AM ∼ −→ M, a ⊗ m 7→ a · m, (2) is a topological isomorphism of left A- ⊗-modules. Similar isomorphisms can be constructed for right A- ⊗-modules. 2.2 Projectivity and flatness The following definitions shall be given in the case of left modules; the definitions in the cases of right modules and bimodules are similar, just use the following category isomorphisms: mod-A ≃ Aop-mod; A-mod-B ≃ (A ⊗Bop)-mod. Let us fix a ⊗-algebra A. Definition 2.7. A complex of A- ⊗-modules {M, d} is called admissible ⇐⇒ it splits in the category LCS. A morphism of A- ⊗-modules f : X → Y is called admissible if it is one of the morphisms in an admissible complex. Definition 2.8. An additive functor F : A-mod → Lin is called exact ⇐⇒ for every admissible complex {M, d} the corresponding complex {F (M ), F (d)} in Lin is exact. Definition 2.9. (1) A module P ∈ A-mod is called projective ⇐⇒ the functor HomA(P, −) is exact. (2) A module Y ∈ A-mod is called flat ⇐⇒ the functor (−) ⊗AY : mod-A → Lin is exact. (3) A module X ∈ A-mod is called free ⇐⇒ X is isomorphic to A ⊗E for some E ∈ C. Due to the [Hel86a, Theorem 3.1.27], any free module is projective, now we are going to prove that a free module is flat. Proposition 2.1. Let X be a free left A- ⊗-module. Then for every admissible sequence {M, d} of right A- ⊗-modules the complex {M ⊗AX, d ⊗IdX}, where (M ⊗AX)i = Mi ⊗AX splits in LCS. 4 Proof. Fix an isomorphism X ∼= A ⊗E. If suffices to prove this proposition for short admissible sequences. However, notice that the sequence 0 M1 ⊗AX d⊗IdX M2 ⊗AX d⊗IdX M3 ⊗AX 0 is isomorphic to the sequence 0 M1 ⊗E d⊗IdE M2 ⊗E d⊗IdE M3 ⊗E 0, and if f : M1 → M2 is a coretraction in LCS, then f ⊗ Id : M1 ⊗E → M2 ⊗E is a coretraction, as well, a similar argument holds for the retraction M2 → M3. Definition 2.10. Let A be an algebra and let M be a right A-module (or a A-bimodule, resp.). For any endomorphism α : A → A denote by Mα a right A-module (or an A-bimodule, resp.), which coincides with M as an abelian group (left A-module, resp.), and whose structure of right A-module is defined by m ◦ a = mα(a). In a similar fashion one defines αM for left modules. The proof of the following lemma is very similar to the proof of the [Hel86a, Proposition 4.1.5]. Lemma 2.1. Let R be a ⊗-algebra and let A be a R- ⊗-algebra. (1) For every (projective/flat) left module X ∈ R-mod the left A- ⊗-module A ⊗RX ∈ A-mod is (projective/flat). Moreover, for every projective (flat) right module X ∈ mod-R the right A-module X ⊗RA ∈ mod-A is (projective/flat). (2) For every projective bimodule P ∈ R-mod-R and α ∈ Aut(R) the bimodule Aα ⊗RP ⊗RA ∈ A-mod-A is projective. Proof. (1) The module X ∈ R-mod is projective iff P is a retract of a left free R-module ([Hel86a, Theorem 3.1.27]), in other words, there are E ∈ C and a retraction σ : R ⊗E → X. But then the map IdA ⊗ σ : A ⊗E ∼= A ⊗RR ⊗E → A ⊗RX is a retraction, as well. The statement of the lemma for flat modules is simpler, because for every M ∈ A-mod we have the canonical isomorphism M ⊗AA ⊗RX ∼= M ⊗RX. The proofs for the right modules are the same. (2) First of all, notice that the following isomorphism of A-R- ⊗-bimodules takes place: Aα ⊗RR ∼−→ Aα. In particular, Aα ∼= A as a left A- ⊗-module. Any projective bimodule is a retract of a free bimodule, in other words, there exist E ∈ C and a retraction σ : R ⊗E ⊗R → P . Notice that the map IdAα ⊗ σ ⊗ IdA : A ⊗E ⊗A ∼= Aα ⊗RR ⊗E ⊗R ⊗RA → Aα ⊗RP ⊗RA is a retraction of A- ⊗-bimodules, and A ⊗E ⊗A is a free A- ⊗-bimodule. The following lemma proves that the [MR01, Lemma 7.2.2] is true for topological modules and ⊗-algebras. 5 2.3 Homological dimensions Definition 2.11. Let X ∈ A-mod. Suppose that X can be included in a following admissible complex: 0 ← X ε←− P0 d0←− P1 d1←− . . . dn−1←−−− Pn ← 0 ← 0 ← . . . , where every Pi is a projective module. Then we will call such complex a projective resolution of X of length n. By definition, the length of an unbounded resolution equals ∞. Flat resolutions are defined similarly. This allows us to define the notion of a derived functor in the topological case, for example, k (M, N ) are defined similarly to the purely A(M, N ) and TorA see [Hel86a, ch 3.3]. In particular, Extk algebraic situation. Definition 2.12. Consider an arbitrary module M ∈ A-mod(C) for some fixed category C ⊆ LCS. Then due to [Hel86a, Theorem 3.5.4] following number is well-defined and we have the following chain of equalities: dhC A(M ) := min{n ∈ Z≥0 : Extn+1 A (M, N ) = 0 for every N ∈ A-mod(C)} = = {the length of a shortest projective resolution of M in A-mod(C)} ∈ {−∞} ∪ [0, ∞]. As we can see, this number doesn't depend on the choice of the category C: dhC A(M ) = min{n ∈ Z≥0 : Extn+1 ≤ min{n ∈ Z≥0 : Extn+1 A (M, N ) = 0 for every N ∈ A-mod(C)} ≤ A (M, N ) = 0 for every N ∈ A-mod} = dhLCS A (M ), dhLCS A (M ) = {the length of a shortest projective resolution of M in A-mod} ≤ ≤ {the length of a shortest projective resolution of M in A-mod(C)} = dhC A(M ). So we will denote this invariant by dhA(M ), and we will call it the projective homological dimension of M . If A is a Fr´echet algebra, and M is a Fr´echet module, then we can define the weak homological dimension of M : w.dhA(M ) = min{n ∈ Z≥0 : TorA n+1(N, M ) = 0 and TorA n (N, M ) is Hausdorff for every N ∈ mod-A(Fr)} = = {the length of the shortest flat resolution of M } ∈ {−∞} ∪ [0, ∞]. Definition 2.13. Let A be a ⊗-algebra. Then we can define the following invariants of A: dglC(A) = sup{dhA(M ) : M ∈ A-mod(C)} − the left global dimension of A. dgrC(A) = sup{dhAop(M ) : M ∈ mod-A(C)} − the right global dimension of A. db(A) = dhA ⊗Aop(A) − the bidimension of A. For a Fr´echet algebra A we can consider weak dimensions. w.dg(A) = sup{w.dhA(M ) : M ∈ A-mod(Fr)} = = sup{w.dhA(M ) : M ∈ mod-A(Fr)} − the weak global dimension of A. w.db(A) = w.dhA ⊗Aop (A) − the weak bidimension of A. Unfortunately, it is not known to the author whether global dimensions depend on the choice of C. We will denote dgl(A) := dglLCS(A), dgr(A) := dgrLCS(A). For more details the reader can consult [Hel86b]. The following theorem demonstrates one of the most important properties of homological dimen- sions. 6 Theorem 2.1. [Hel86a, Proposition 3.5.5] Let A be a ⊗-algebra. If 0 → X ′ → X → X ′′ → 0 is an admissible sequence of left A- ⊗-modules, then dhA(X) ≤ max{dhA(X ′), dhA(X ′′)} dhA(X ′) ≤ max{dhA(X), dhA(X ′′) − 1} dhA(X ′′) ≤ max{dhA(X), dhA(X ′) + 1}. In particular, dhA(X) = max{dhA(X ′), dhA(X ′′)} except when dhA(X) < dhA(X ′′) = dhA(X ′) + 1. Moreover, the same estimates hold for weak homological dimensions of Fr´echet modules over Fr´echet algebras, for the proof see [Pir08b]. Proposition 2.2. [Pir08b, Proposition 4.1, Corollary 4.4] Let A be a Fr´echet algebra, then for every M ∈ A-mod(Fr) we have w.dhA(M ) = min{n : Extn+1 A (M, N ∗) = 0 for every N ∈ mod-A(Fr)}, where Y ∗ denotes the strong dual of Y . As a corollary, for every admissible sequence 0 → X ′ → X → X ′′ → 0 of left Fr´echet A- ⊗-modules we have the following estimates: w.dhA(X) ≤ max{w.dhA(X ′), w.dhA(X ′′)} w.dhA(X ′) ≤ max{w.dhA(X), w.dhA(X ′′) − 1} w.dhA(X ′′) ≤ max{w.dhA(X), w.dhA(X ′) + 1}. (3) In particular, w.dhA(X) = max{w.dhA(X ′), w.dhA(X ′′)} except when w.dhA(X) < w.dhA(X ′′) = w.dhA(X ′) + 1. Proposition 2.3. Let R ∈ alg(C), let A be a R- ⊗-algebra, which is free as a left R-module, and let M be a right A- ⊗-module. (1) For every projective resolution of M in mod-R 0 ← M ← P0 ← P1 ← P2 ← . . . the complex 0 ← M ⊗RA ← P0 ⊗RA ← P1 ⊗RA ← P2 ⊗RA ← . . . (4) is a projective resolution of M ⊗RA in the category of right A- ⊗-modules. In particular, dhAop(M ⊗RA) ≤ dhRop (M ). (2) Moreover, if C ⊂ Fr, then for every flat resolution of M in mod-R 0 ← M ← F0 ← F1 ← F2 ← . . . the complex 0 ← M ⊗RA ← F0 ⊗RA ← F1 ⊗RA ← F2 ⊗RA ← . . . (5) is a flat resolution of M ⊗RA in the category of right A- ⊗-modules. In particular, w.dhAop(M ⊗RA) ≤ w.dhRop (M ). Proof. (1) Lemma 2.1 implies that Pi ⊗RA is a projective right A- ⊗-module for all i. The complex (4) is admissible, because the functor (−) ⊗RA for a free A preserves admissibility (due to the Proposition 2.1), so it defines a projective resolution of M ⊗RA in the category mod-A. 7 (2) Due to the Lemma 2.1 the modules Fi ⊗RA are flat right A- ⊗-modules for all i. Then the rest of the proof is the same as in (1). Lemma 2.2. Let R be a ⊗-algebra, and consider an R- ⊗-algebra A. 1. If A is projective as a right R- ⊗-module, and M is projective as a right A- ⊗-module, then M is projective as a right R- ⊗-module. In this case for every X ∈ mod-A we have dhRop (X) ≤ dhAop (X). (6) 2. If A is flat as a right R- ⊗-module, and M is flat as a right A- ⊗-module, then M is flat as a right R- ⊗-module. If R, A are Fr´echet algebras, then for every X ∈ mod-A(Fr) we have w.dhAop(X) ≤ w.dhRop (X). (7) Proof. (1) Due to [Hel86a, Theorem 3.1.27] we can fix an isomorphism of right R- ⊗-modules A⊕S ∼= X ⊗R for some right R- ⊗-module S. Suppose that M is a projective right A- ⊗-module. Then due to [Hel86a, Theorem 3.1.27] there exists a right module N such that M ⊕ N = E ⊗A as right A-modules, but then M ⊕ N ⊕ (E ⊗S) ∼= E ⊗A ⊕ (E ⊗S) ∼= E ⊗(X ⊗R) = (E ⊗X) ⊗R. (2) As in the proof of [MR01, Lemma 7.2.2], we notice that for any right R- ⊗-module N we have and this immediately implies our statement. M ⊗RN ∼= M ⊗A(A ⊗RN ), Lemma 2.3. Let R is a ⊗-algebra, α : R → R is an automorphism and M is a right R-module, then dhRop (Mα) = dhRop (M ) and if R is a Fr´echet algebra, then w.dhRop (Mα) = w.dhRop (M ). Proof. The proof relies on the fact that (·)α : mod-R → mod-R and α(·) : R-mod → R-mod can be viewed as functors between mod-R and R-mod, which preserve admissibility of morphisms, projectivity and flatness of modules. Indeed, if f : M → N is an admissible module homomorphism, then fα : Mα → Nα is admissible, because (cid:3)fα = (cid:3)f , where (cid:3) : mod-R → LCS denotes the forgetful functor. The same goes for α(·). Let P ∈ mod-R be projective. Then for any admissible epimorphism ϕ : X → Y we have the following chain of canonical isomorphisms: Hom(Pα, Y ) ≃ Hom(P, Yα−1 ) ≃ Hom(P, Xα−1 ) ≃ Hom(Pα, X). Let F ∈ mod-R be flat. Then Fα ⊗RX ≃ F ⊗R(α−1X) and we already know that α−1(·) preserves admissibility. 3 Estimates for the bidimension and projective global dimensions of holomorphic Ore extensions 3.1 Bimodules of relative differentials Firstly, let us give several necessary algebraic definitions: Definition 3.1. Let S be an algebra, A be an S-algebra and M be an A-bimodule. Then an S-linear map δ : A → M is called an S-derivation if the following relation holds for every a, b ∈ A: δ(ab) = δ(a)b + aδ(b). 8 Example 3.1. Let α be an endomorphism of an algebra A. Then an α-derivation is precisely an A-derivation δ : A → αA. The following definition is due to J. Cuntz and D. Quillen, see [CQ95]: Definition 3.2. Suppose that S is an algebra and (A, η) is an S-algebra, where η : S → A is an algebra homomorphism. Denote by A = A/Im(η(S)) the S-bimodule quotient. Then we can define the bimodule of relative differential 1-forms Ω1 S(A) are usually denoted by a0 ⊗ a1 = a0da1. The A-bimodule structure on Ω1 S(A) is uniquely defined by the following relations: S(A) = A ⊗S A. The elementary tensors in Ω1 b ◦ (a0da1) = ba0da1, (a0da1) ◦ b = a0d(a1b) − a0a1db. The bimodule of relative differential 1-forms together with the canonical S-derivation dA : A → Ω1 S(A), dA(a) = 1 ⊗ a = da has the following universal property: Proposition 3.1 ([CQ95], Proposition 2.4). For every A-bimodule M and an S-derivation D : A → M there is a unique A-bimodule morphism ϕ : Ω1 S(A) → M such that the following diagram is commutative: Ω1 R(A) ∃! ϕ M dA A D Proposition 3.2 ([CQ95], Proposition 2.5). The following sequence of A-bimodules is exact: 0 j Ω1 SA A ⊗S A m A 0, (8) (9) where j(a0 ⊗ a1) = j(a0da1) = a0a1 ⊗ 1 − a0 ⊗ a1 and m denotes the multiplication. In the appendix of this paper we compute the bimodule of relative differential 1-forms of Ore extensions, see Proposition A.1. 3.2 A topological version of the bimodule of relative differentials The following definition serves as a topological version of the Definition 3.1. Definition 3.3. Let R be a ⊗-algebra, and suppose that (A, η) is a R- ⊗-algebra. Denote by A/Im(η(R)) = A the R- ⊗-bimodule quotient. Then we can define the (topological) bimodule of relative R(A) := A⊗R A. The elementary tensors are usually denoted by a0 ⊗a1 = a0da1. b ◦ (a0da1) = ba0da1, (a0da1) ◦ b = a0d(a1b) − a0a1db for every a0, a1, b ∈ A. R(A) is uniquely defined by the following relations: differential 1-forms bΩ1 The structure of A- ⊗-bimodule on bΩ1 notation bΩ1 R(A), unlike in [Pir08a]. ential 1-forms. Remark. To avoid confusion with the algebraic bimodules of differential 1-forms, here we use the We will need to recall the following propositions related to topological bimodule of relative differ- Theorem 3.1. [Pir08a, p. 99] For every A- ⊗-bimodule M and a continuous R-derivation D : A → M R(A) → M such that the following diagram is there exists a unique A- ⊗-bimodule morphism ϕ : bΩ1 commutative: ∃! ϕ M , (10) where dA(a) = da. A R(A) bΩ1 dA D 9 Proposition 3.3. [Pir08a, Proposition 7.2] The short exact sequence 0 j A ⊗RA m A 0, R(A) bΩ1 where j(a0da1) = a0 ⊗ a1 − a0a1 ⊗ 1 and m(a0 × a1) = a0a1, splits in the categories A-mod-R and R-mod-A. 3.3 Holomorphic Ore extensions To give the definition of the holomorphic Ore extension of a ⊗-algebra, associated with an endomor- phism and derivation, we need to recall the definition of localizable morphisms. Definition 3.4. [Pir08a, Definition 4.1] Let X be a LCS and consider a family F ⊂ L(X) of continuous linear maps X → X. Then a seminorm k·k on X is called F-stable if for every T ∈ F there exists a constant CT > 0, such that Definition 3.5. (1) Let X be a LCS. kT xk ≤ CT kxk for every a ∈ A. A family of continuous linear operators F ⊂ L(X) is called localizable, if the topology on X can be defined by a family of F-stable seminorms. (2) Let A be an Arens-Michael algebra. A family of continuous linear operators F ⊂ L(A) is called m-localizable, if the topology on A can be defined by a family of submultiplicative F-stable seminorms. Now we will state the theorem which proves the existence of certain ⊗-algebras which would be reasonable to call the "holomorphic Ore extensions". Theorem 3.2. [Pir08a, Section 4.1] Let A be a ⊗-algebra and suppose that α : A → A is a localizable endomorphism of A, δ : A → A is a localizable α-derivation of A. Then there exists a unique multiplication on the tensor product A ⊗O(C), such that the following conditions are satisfied: (1) The resulting algebra, which is denoted by O(C, A; α, δ), is an A- ⊗-algebra. (2) The natural inclusion is an algebra homomorphism. A[z; α, δ] ֒→ O(C, A; α, δ) (3) For every Arens-Michael A- ⊗-algebra B the following natural isomorphism takes place: Hom(A[z; α, δ], B) ∼= HomA-alg(O(C, A; α, δ), B). Moreover, let α be invertible, and suppose that the pair (α, α−1) is localizable. Then there exists a unique multiplication on the tensor product A ⊗O(C×), such that the following conditions are satisfied: (1) The resulting algebra, which is denoted by O(C×, A; α), is a ⊗-algebra. (2) The natural inclusion is an algebra homomorphism. A[z; α, α−1] ֒→ O(C×, A; α) (3) For every Arens-Michael A- ⊗-algebra B the following natural isomorphism takes place: Hom(A[z; α, α−1], B) ∼= HomA-alg(O(C×, A; α), B). 10 And if we replace the word "localizable" with "m-localizable" in this theorem, then the resulting algebras will become Arens-Michael algebras. By considering Aop[z; α, δ], we can formulate a version of the Theorem 3.2 for opposite Ore ex- In fact, we need this to prove the following tensions, the proof is basically the same in this case. corollary: Corollary 3.1. Let R be a ⊗-algebra with an endomorphism α : R → R and a α-derivation δ such that the pair (α, δ) is localizable. (1) If A = O(C, R; α, δ), then A is free as a left R- ⊗-module. (2) If α is invertible, and (α, α−1) is a localizable pair, then A = O(C×, R; α) is free as a left and right R- ⊗-module. (3) If α is invertible, and the pair (α, δ) is m-localizable, then A = O(C, R; α, δ) is free as a left and right R- ⊗-module. Proof. 1. This already follows from the fact that O(C, R; α, δ) is isomorphic to A ⊗O(C). 2. This is the immediate corollary of the [Pir08a, Lemma 4.12]: just notice that we can define γ : R ⊗O(C) → O(C) ⊗R, γ(r ⊗ zn) = zn ⊗ α−n(r), and it will be a continuous inverse of τ . To finish the proof, we only need to notice that τ and γ are isomorphisms of left and right R- ⊗-modules, respectively. 3. This follows from the above remark: let us denote the resulting "opposite" holomorphic Ore extensions by Oop(C×, R; α, δ). Then the isomorpisms 1 can be extended via the universal properties to the topological isomor- phisms of R- ⊗-algebras as follows: consider the algebra homomorphisms A[t; α, δ] ∼ −→ A[t; α−1, −δα−1] ֒→ Oop(C×, R; α−1, −δα−1), A[t; α−1, −δα−1] ∼−→ A[t; α, δ] ֒→ O(C, R; α, δ). Now notice that the extensions are continuous and inverse on dense subsets of the holomorphic Ore extensions, therefore, they are actually inverse to each other. However, the opposite Ore extensions are free as right R- ⊗-modules by definition. Notice that to apply the universal properties we need the the morphisms to be m-localizable, so that the holomorphic Ore extensions are Arens-Michael algebras. In this paper we compute the topological bimodules of relative differential 1-forms of holomorphic Ore extensions and smooth crossed products by Z, see Propositions A.2 and A.3. 3.4 Upper estimates for the bidimension Theorem 3.3. Suppose that R is a ⊗-algebra, and A is one of the two ⊗-algebras: (1) A = O(C, R; α, δ), where the pair {α, δ} is localizable. (2) A = O(C×, R; α), where the pair {α, α−1} is localizable. Then we have db(Aop) ≤ db(Rop) + 1. 11 Proof. Due to the Proposition 3.3 and Proposition A.2, we have the following sequence of A- ⊗- bimodules, which splits in the categories R-mod-A and A-mod-R: 0 Aα ⊗RA j A ⊗RA m A 0, where m is the multiplication operator. Let 0 ← R ← P0 ← · · · ← Pn ← 0 (11) (12) be a projective resolution of R in R-mod-R. Notice that (12) splits in R-mod and mod-R, because all objects in the resolution are projective as left and right R- ⊗-modules ([Hel86a, Corollary 3.1.18]). Therefore, we can apply the functors Aα ⊗R(−) and A ⊗R(−) to (12) and the resulting complexes of A-R- ⊗-bimodules are still admissible: 0 ← A ← A ⊗RP0 ← · · · ← A ⊗RPn ← 0 0 ← Aα ← Aα ⊗RP0 ← · · · ← Aα ⊗RPn ← 0. (13) (14) Recall that A is a free left R- ⊗-module due to the Corollary 3.1, so the functor (−) ⊗RA preserves admissibility, due to the Proposition 2.1, therefore the following complexes of A- ⊗-bimodules are admissible: 0 ← A ⊗RA ← A ⊗RP0 ⊗RA ← · · · ← A ⊗RPn ⊗RA ← 0 0 ← Aα ⊗RA ← Aα ⊗RP0 ⊗RA ← · · · ← Aα ⊗RPn ⊗RA ← 0 (15) (16) Lemma 2.1 implies that (15) and (16) define projective resolutions for A ⊗RA and Aα ⊗RA. Now we can apply Theorem 2.1 to (11), so we get db(A) = dhAe(A) ≤ max{dhAe(A ⊗RA), dhAe(Aα ⊗RA) + 1} ≤ n + 1. In other words, we have obtained the desired estimate db(A) ≤ db(R) + 1. 3.5 Upper estimates for the right global and weak global dimensions Now we are prepared to state the theorem: Theorem 3.4. Let R be a ⊗-algebra. Suppose that A is one of the two ⊗-algebras: (1) A = O(C, R; α, δ), where α is invertible, and the pair {α, δ} is localizable. (2) A = O(C×, R; α), where the pair {α, α−1} is localizable. Then the right global dimension of A can be estimated as follows: dgr(A) ≤ dgr(R) + 1, and a similar estimate holds for the weak dimensions if R is a Fr´echet algebra: w.dg(A) ≤ w.dg(R) + 1. Proof. Suppose that M is a right A- ⊗-module. Then we can apply the functor M ⊗A(−) to the sequence (11). Notice that the resulting sequence of right A- ⊗-modules 0 M ⊗AAα ⊗RA IdM ⊗j M ⊗AA ⊗RA IdM ⊗m M ⊗AA 0 12 is isomorphic to the sequence 0 Mα ⊗RA j ′ M ⊗RA m M 0. (17) Since (11) splits in A-mod-R, (17) splits in mod-R, in particular, this is an admissible short exact sequence. Now notice that we can apply Theorem 2.1 to (17), so we get dhAop (M ) ≤ max{dhAop (M ⊗RA), dhAop (Mα ⊗RA) + 1} ≤ dhRop (M ) + 1 due to the Corollary 3.1, Proposition 2.3 and Lemma 2.3. Hence, the following estimate holds: dgr(A) ≤ dgr(R) + 1. For the weak dimensions we apply the second part of the Proposition 2.2 to the sequence (17): w.dhAop(M ) ≤ max{w.dhAop(M ⊗RA), w.dhAop(Mα ⊗RA)} + 1 ≤ w.dhRop (M ) + 1. 3.6 Lower estimates In order to obtain lower estimates, we need to formulate the following lemma: Proposition 3.4. Suppose that R is a ⊗-algebra, and A is a R- ⊗-algebra which is free as a left R- ⊗-module. Also assume that there exists an isomorphism of left R- ⊗-modules ϕ : A → R ⊗E such that ϕ(1) = 1 ⊗ x for some x ∈ E. Then i : M → M ⊗RA, i(m) = m ⊗ 1 is an admissible monomorphism for every M ∈ mod-R. In particular, it is a coretraction between the underlying locally convex spaces. Proof. Look at the following diagram: M i M ⊗RA IdM ⊗ϕ M ⊗RR ⊗E π⊗IdE M ⊗E, m→m⊗x where π : M ⊗RR → M, π(m ⊗ r) = mr. Due to the Hahn-Banach theorem there exists a functional f ∈ E∗ such that f (x) = 1, so the map m → m ⊗ x admits a right inverse, which is uniquely defined by n ⊗ y → f (y)n, therefore i as a mapping of lcs admits a right inverse too, because IdM ⊗ ϕ and π ⊗ IdE are invertible. Proposition 3.5. Let R be a Fr´echet algebra, and assume that dgr(R) < ∞. Suppose that the following conditions hold: (1) Let A be a R- ⊗-algebra which is a free left R- ⊗-module, moreover, we can choose an isomorphism of left R- ⊗-modules ϕ : A → R ⊗E in such a way that ϕ(1) = 1 ⊗ x for some x ∈ E. (2) A is projective as a right R- ⊗-module. Then dgrFr(R) ≤ dgrFr(A) and w.dg(R) ≤ w.dg(A). Proof. Fix a Fr´echet module M ∈ mod-R(Fr) such that dhRop (M ) = dgr(R) = n. The Proposition 3.4 states that the map M → M ⊗RA, m → m ⊗ 1 is an admissible monomorphism, so there exists a short exact sequence 0 → M i ֒−→ M ⊗RA → N → 0, where N ∼= (M ⊗RA)/i(M ), which is precisely the cokernel of i. This sequence is admissible because the projection map is open and i is an admissible monomorphism, we can apply the [Hel86a, Proposition 3.1.8(III)]. 13 Remark. If M is an arbitrary R- ⊗-module, then the map N → (M ⊗RA)/i(M ) → ((M ⊗RA)/i(M ))∼ is not open in general. Notice that dhRop (M ) = dgr(R) and dhRop (N ) ≤ dgr(R), therefore dhRop (M ⊗RA) = dgr(R) due to the Proposition 2.1. Now recall that A is projective as a right R- ⊗-module, therefore, due to the Lemma 2.2, and the Proposition 2.3 we have dgr(R) = dhRop (M ⊗RA) L2.2 ≤ dhAop(M ⊗RA) P 2.3 ≤ dhRop (M ) = dgr(R). This immediately implies dgrFr(R) ≤ dgrFr(A). By replacing X with X ∗ in the above argument and applying the Proposition 2.2 we get that w.dg(R) ≤ w.dg(A). As a quick corollary from the Proposition 3.5 and the Corollary 3.1 we obtain lower estimates for the homological dimensions. Theorem 3.5. Let R be a Fr´echet algebra, and suppose that dgr(Rop) < ∞ and A is one of the two ⊗-algebras: (1) A = O(C, R; α, δ), where α is invertible, and the pair (α, δ) is m-localizable. (2) A = O(C×, R; α), where the pair (α, α−1) is localizable. Then the conditions of the Proposition 3.5 are satisfied. As a corollary, we have the following estimates: dgrFr(R) ≤ dgrFr(A), w.dg(R) ≤ w.dg(A). Proof. We have nothing to prove, because the Corollary 3.1 ensures that A is free as a left and right R- ⊗-module, both conditions follow from this and the construction of holomorphic Ore extensions. 4 Homological dimensions of smooth crossed products by Z First of all, let us recall the definition of the space of rapidly decreasing sequences: Definition 4.1. n∈Z s =(cid:26)(an) ∈ CZ : kakk = sup ∼=((an) ∈ CZ : kak2 k =Xn∈Z ∼=((an) ∈ CZ : kakk =Xn∈Z an(n + 1)k < ∞ ∀k ∈ N(cid:27) = an2(n + 1)2k < ∞ ∀k ∈ N) = an(n + 1)k < ∞ ∀k ∈ N) . This is the Example 29.4 in [MDR97]. Clearly, this is a Fr´echet space. The following definitions and theorems are due to L. Schweitzer, see [Sch93] or [PS94] for more detail. Definition 4.2. Suppose that R is a Fr´echet algebra. Let G be one of the groups R, T or Z and suppose that α : G → Aut(R) is an action of G on R. Then α is called an m-tempered action if the topology on R can be defined by a sequence of submultiplicative seminorms {k·km : m ∈ N} such that for every m there exists a polynomial p satisfying for any x ∈ G and r ∈ R. kαx(r)kλ ≤ p(x) krkλ 14 Definition 4.3. (1) Let E be a Hausdorff topological vector space. For a function f : R −→ E and x ∈ R we denote (vector-valued differentiation). f ′(x) := lim h→0 f (x + h) − f (x) h . (2) Let A be a Fr´echet algebra with a fixed generating system of seminorms {k·kλ , λ ∈ Λ} Then we can define the following locally convex spaces: S (Z, A) =(f = (f (k))k∈Z ∈ AZ : kf kλ,k :=Xn∈Z(cid:13)(cid:13)(cid:13)f (n)(cid:13)(cid:13)(cid:13)λ S (T, A) =(cid:26)f : T → A : kf kλ,k := sup z∈T(cid:13)(cid:13)(cid:13)f (k)(z)(cid:13)(cid:13)(cid:13)λ S (R, A) =(cid:26)f : R → A : kf kλ,k,l := sup x∈R(cid:13)(cid:13)(cid:13)xlf (k)(x)(cid:13)(cid:13)(cid:13)λ (n + 1)k < ∞ for all λ ∈ Λ, k ∈ N) , < ∞ for all λ ∈ Λ, k ≥ 0(cid:27) , < ∞ for all λ ∈ Λ, k, l ≥ 0(cid:27) . Theorem 4.1 ([Sch93], Theorem 3.1.7). Let R be a Fr´echet-Arens-Michael algebra with an m-tempered action of one of the groups R, T or Z. Then the space S (G, R) endowed with the following multiplication: (f ∗ g)(x) =ZG f (y)αy(g(xy−1))dy becomes a Fr´echet-Arens-Michael algebra. This algebra is denoted by S (G, R; α). Remark. The algebras S (G, R; α) are, in general, non-unital for G = R, T. They are always unital for G = Z. Proposition 4.1. Consider the following multiplication on S (G, R): (f ∗′ g)(x) =ZG αy−1(f (xy−1))g(y)dy. Then the following locally convex algebra isomorphism takes place: i : S (G, R; α) → (S (G, R), ∗′), i(f )(x) = αx−1(f (x)). In particular, when G = Z, this is an isomorphism of unital algebras. Proof. The mapping i is, obviously, a topological isomorphism of locally convex spaces. Now notice that (i(f ) ∗′ i(g))(x) =ZG =ZG αy−1(i(f )(xy−1))i(g)(y)dy =ZG αx−1(f (y−1))αy−1x−1(g(yx))dy = αx−1(cid:18)ZG = i(f ∗ g)(x), αx−1(f (xy−1))αy−1 (g(y))dy = f (y−1)αy−1(g(yx))dy(cid:19) = therefore, i is an algebra homomorphism. For example, let us consider a Fr´echet-Arens-Michael algebra R with a m-tempered action of Z and fix a generating family of submultiplicative seminorms {k·kλ λ ∈ Λ} on R, such that for where p is a polynomial. In this case the algebra R is contained in S (Z, R; α): kαn 1 (r)kλ ≤ p(n) krkλ , (r ∈ R, n ∈ Z), where (rei)j := δijr. R ֒→ S (Z, R; α), r 7→ re0, Hence, S (Z, R; α) becomes a unital R- ⊗-algebra, and in the appendix we prove the Proposition R(S (Z, R; α)) is similar to the algebraic and holomorphic cases. This gives us an opportunity to formulate the following theorem: A.3, which states that the structure of bΩ1 15 Theorem 4.2. Let R be a Fr´echet-Arens-Michael algebra with with a m-tempered Z-action α. If we denote A = S (Z, R; α), then we have db(A) ≤ db(R) + 1, dgr(A) ≤ dgr(R) + 1, w.dg(A) ≤ w.dg(R) + 1. The proof of the theorem is very similar to the proofs of Theorems 3.3 and 3.4. As a simple corollary from the Proposition 4.1, we get that S (Z, R; α) is free as a left and right R- ⊗-module, and together with Proposition 3.5 we obtain the lower estimates: Theorem 4.3. Let R be Fr´echet-Arens-Michael algebra with dgr(Rop) < ∞ and with a m-tempered Z-action α. Denote A = S (Z, R; α). Then the conditions of the Proposition 3.5 are satisfied. In particular, we have dgrFr(R) ≤ dgrFr(A), w.dg(R) ≤ w.dg(A). A Relative bimodules of differential 1-forms of Ore extensions Proposition A.1. Let R be a C-algebra. Suppose that (1) A = R[t; α, δ], where α : R → R is an endomorphism and δ : R → R is an α-derivation. (2) A = R[t, t−1; α], where α : R → R is an automorphism. Then Ω1 R(A) is canonically isomorphic as an A-bimodule to Aα ⊗R A. Proof. The first part of the proof works for the both cases. Define the map ϕ : Aα × A → Ω1 follows: R(A) as ϕ(f, g) = f d(tg) − f tdg = f (dt)g, (f, g ∈ A). This map is balanced, because ϕ(f, g) + ϕ(f, g′) = ϕ(f, g + g′), ϕ(f, g) + ϕ(f ′, g) = ϕ(f + f ′, g), (f, f ′, g, g′ ∈ A) and ϕ(f, rg) = f (dt)rg = f (d(tr))g = f d(α(r)t + δ(r))g = f α(r)(dt)g = ϕ(f ◦ r, g). Also we have hϕ(f, g) = ϕ(hf, g), ϕ(f, g)h = ϕ(f, gh). Therefore, ϕ induces a well-defined homomorphism of A-bimodules ϕ : Aα ⊗R A → Ω1 We will use the universal property of Ω1 R(A) to construct the inverse morphism. R(A). (1) Suppose that A = R[t; α, δ]. Consider the following linear mapping: D : A → Aα ⊗R A, D(rtn) = rtk ⊗ tn−k−1. n−1Xk=0 Now we want to show that D is an R-derivation. First of all, notice that for any k=0 rktk ∈ R[t; α, δ] and n ≥ 0 we have f =Pm D(f tn) = mXk=0 rkD(tk+n) = mXk=0 rk(D(tk)tn + tkD(tn)) = D(f )tn + f D(tn). (18) It suffices to show that D(tnr) = D(tn)r, let us prove it by induction w.r.t. n: D(tn+1r) = D(tnδ(r) + tnα(r)t) = D(tn)δ(r) + D(tn)α(r)t + tnα(r) ⊗ 1 = = D(tn)tr + tnα(r) ⊗ 1 = D(tn)tr + tn ⊗ r = (D(tn)t + tnD(t))r = D(tn+1)r. (19) 16 (2) Suppose that A = R[t, t−1; α]. Consider the following linear mapping: D : A → Aα ⊗R A, D(rtn) =(Pn−1 −Pn k=0 rtk ⊗ tn−k−1, k=1 rt−k ⊗ tn+k−1, if n ≥ 0, if n < 0. As in the first case, this map turns out to be an R-derivation. Notice that it suffices to prove the following statements: • D(tnr) = D(tn)r for n ∈ Z, r ∈ R. • D(tmtn) = D(tm)tn + tmD(tn) for n, m ∈ Z. If n < 0, then, to prove the first identity, we can use the argument, similar to (19), so let us concentrate on the proof of the second statement. If m, n have the same sign, then we can repeat the argument in (18). Now suppose that m > 0 and n < 0. In this case we have D(tm)tn + tmD(tn) = tk ⊗ tn+m−k−1 − m−1Xk=0 tm−k ⊗ tn+k−1. nXk=1 If m = n, then everything cancels out and we get 0. If m < n, then after the cancellations we get − nXk=m+1 tm−k ⊗ tn+k−1 = − m−nXj=1 t−j ⊗ tn+m+j−1 = D(tm−n). If m > n, then we get and this finishes the argument. m+n−1Xk=0 tk ⊗ tn+m−k−1 = D(tm+n), The rest of this proof works in the both cases. Notice that ϕ ◦ D = dA, because tk ⊗ tn−k−1! = r (tkd(tn−k) − tk+1d(tn−k−1)) = r(dtn) = dA(rtn) n−1Xk=0 t−k ⊗ tn+k−1 = −r = rd(tn) = dA(rtn). (t−kd(tn+k) − t−k+1d(tn+k−1)) = nXk=1 for every r ∈ A, and for n < 0 we have ϕ ◦ D(rtn) = rϕ n−1Xk=0 ϕ ◦ D(rtn) = −rϕ nXk=1 R(A) that ϕ ◦ eD = IdΩ1 Denote the extension of D by eD, so D = eD ◦ dA. Therefore, we can derive from the universal property R(A). And of Ω1 eD ◦ ϕ(a ⊗ b) = a(eD ◦ ϕ(1 ⊗ 1))b = eD(dt) = a ⊗ b ⇒ eD ◦ ϕ = IdAα⊗A. The following proposition was already proven by A. Yu. Pirkovskii, see [Pir08a, Proposition 7.8], but we present another proof, which is similar to the proof of the Proposition A.1; it even works in the case of localizable morphisms. Moreover, the proof can be carried over to the case of smooth crossed products by Z, as we will see later. Proposition A.2. Let R be an Arens-Michael algebra. Suppose that A is one of the following ⊗-algebras: 17 (1) A = O(C, R; α, δ), where α : R → R is an endomorphism and δ : R → R is a α-derivation, such that the pair (α, δ) is a localizable pair of morphisms. (2) A = O(C×, R; α), where α : R → R is an automorphism, such that the pair (α, α−1) is a localizable pair of morphisms. R(A) is canonically isomorphic to Aα ⊗RA. Then bΩ1 kδ(x)kλ ≤ C kxkλ. Define the map Aα × A →bΩ1 Proof. Fix a generating family of seminorms {k·kλ : λ ∈ Λ} on R such that kα(x)kλ ≤ C kxkλ and R(A) as in the proof of the Proposition 0.1: ϕ(f, g) = f d(zg) − f zdg = f (dt)g, (f, g ∈ A). This is a R-balanced map (the proof is literally the same as in the Proposition A.1), also it is easily seen from the continuity of the multiplication on A that this map is continuous. Therefore, this map (1) Suppose that A = O(C, R; α, δ). Consider the following linear map: induces a continuous A- ⊗-bimodule homomorphism Aα ⊗RA →bΩ1 n−1Xk=0 D : R[z; α, δ] → Aα ⊗RA, D(rzn) = R(A). rzk ⊗ zn−k−1. For now it is defined on the dense subset of A; let us prove that this map is continuous. Fix λ1, λ2 ∈ Λ and ρ1, ρ2 ∈ R≥0. Denote the projective tensor product of k·kλ1,ρ1 by γ. and k·kλ2,ρ2 γ(D(f )) ≤ Then for every f =Pm mXk=1 mXk=1 ≤ k=0 fkzk ∈ R[z; α, δ] ⊂ A we have γ(D(fkzk)) = γ k−1Xl=0 mXk=1 fkzl ⊗ zk−l−1! ≤ mXk=1 kfkkλ1 k−1Xl=0 1ρk−l−1 ρl 2 ≤ kfkkλ1 (2 max{ρ1, ρ2, 1})k = kf kλ1,2 max{ρ1,ρ2,1} . (2) Suppose that A = O(C×, R; α). Consider the following linear map: D : R[t, t−1; α] → Aα ⊗RA, D(rzn) =(Pn−1 −Pn k=0 rzk ⊗ zn−k−1, k=1 rz−k ⊗ zn+k−1, if n ≥ 0, if n < 0. For now it is defined on the dense subset of A; let us prove that this map is continuous. Fix λ1, λ2 ∈ Λ and ρ1, ρ2 ∈ R≥0. Denote the projective tensor product of k·kλ1,ρ1 by γ. Suppose that n ≥ 0. Then we have and k·kλ2,ρ2 If n < 0, then γ(D(rzn)) = γ n−1Xl=0 γ(D(rzn)) = γ nXl=1 Therefore, for every f =Pm mXk=−m rzl ⊗ zn−l−1! ≤ krkλ1 rz−l ⊗ zn+l−1 ≤ krkλ1 γ(cid:16)D(fkzk)(cid:17) ≤ γ(D(f )) = −m fkzk ∈ R[z, z−1; α] ⊂ A we have 1ρn−l−1 ρl 2 ≤ krznkλ1,2 max{ρ1,ρ2,1} . n−1Xl=0 ρ−l 1 ρn+l−1 2 ≤ krznkλ1,2 min{ρ1,ρ2,1} . nXl=1 = kf kλ1,2(ρ1+ρ2+1) . mXk=−m(cid:13)(cid:13)(cid:13)fkzk(cid:13)(cid:13)(cid:13)λ1,2(ρ1+ρ2+1) 18 Therefore, this map can be uniquely extended to the whole algebra A; we will denote the extension by D, as well. Notice D is also an R-derivation: the equality D(ab) − D(a)b − aD(b) = 0 holds for R[z; α, δ] × R[z; α, δ] ⊂ A × A, which is a dense subset of A × A. Therefore, D(ab) = D(a)b + aD(b) for every a, b ∈ A. R(A) → Aα ⊗RA, so Notice that ϕ ◦ D = dA. Denote the extension of D : A → Aα ⊗RA by eD : bΩ1 D = eD ◦ dA. Therefore we can derive from the universal property of bΩ1 R(A) that ϕ ◦ eD = Id. And eD ◦ ϕ(a ⊗ b) = a(eD ◦ ϕ(1 ⊗ 1))b = eD(dt) = a ⊗ b. Therefore, the equality eD ◦ ϕ = Id holds on a dense subset of Aα ⊗RA, but eD ◦ ϕ is continuous, therefore, eD ◦ ϕ = Id holds everywhere on Aα ⊗RA. Z on R. If we denote the algebra S (Z, R; α) by A, then bΩ1 Proposition A.3. Let R be a Fr´echet-Arens-Michael algebra and consider an m-tempered action α of ⊗RA. Proof. Fix a generating system of submultiplicative seminorms {k·kλ : λ ∈ Λ} on R, such that R(A) is canonically isomorphic to Aα1 kαn(x)kλ = kαn 1 (x)kλ ≤ p(n) kxkλ (x ∈ R, λ ∈ Λ). Define the map ϕ : Aα1 × A →bΩ1 R(A) as follows: ϕ(f, g) = f d(e1 ∗ g) − (f ∗ e1)dg = f (de1)g. It is a continuous R-balanced linear map (the proof is literally the same as in the previous propositions). Now consider the following linear map: D : A → Aα1 ⊗RA, D(ren) = n−1Pk=0 nPk=1 − rek ⊗ en−k−1, if n ≥ 0 . re−k ⊗ en+k−1, if n < 0 Let us prove that it is a well-defined and continuous map from A to Aα1 and k·kλ2,k2 k1, k2 ∈ Z≥0. Denote the projective tensor product of k·kλ1,k1 have ⊗RA. Fix λ1, λ2 ∈ Λ and by γ. Let n ≥ 1, then we γ(D(ren)) = γ n−1Xk=0 rek ⊗ en−k−1! ≤ krkλ1 (nk2 + 2k1(n − 1)k2 + . . . nk1) ≤ ≤ krkλ1 ≤ krkλ1 (nmax{k1,k2} + 2max{k1,k2}(n − 1)max{k1,k2} + · · · + nmax{k1,k2}) ≤ n2max{k1,k2}+1 = krenkλ1,2max{k1,k2}+1 . For n < 0 the argument is pretty much the same: γ(D(ren)) = γ nXk=1 re−k ⊗ en+k−1 ≤ krkλ1 (2k1 (n + 1)k2 + 3k1nk2 + . . . (n + 1)k12k2) ≤ ≤ krkλ1 ≤ krkλ1 (2max{k1,k2}(n + 1)max{k1,k2} + · · · + (n + 1)max{k1,k2}2max{k1,k2}) ≤ n(n + 1)2 max{k1,k2} ≤ 2 krkλ1 n4 max{k1,k2}+1 = 2 krenkλ1,4max{k1,k2}+1 . It is easily seen that for every f ∈ R ⊗c00 we have γ(Df ) ≤ Xm∈Z γ(D(f (m)em)) ≤ 2Xm∈Z(cid:13)(cid:13)(cid:13)f (m)em(cid:13)(cid:13)(cid:13)λ1,4max{k1,k2}+1 = 2 kf kλ1,4max{k1,k2}+1 . Then D is a R-derivation which can be uniquely extended to the whole algebra A, the extension eD is the inverse of ϕ, the proof is the same as in the Proposition A.2. 19 References [CQ95] [Hel06] J. Cuntz and D. Quillen. "Algebra Extensions and Nonsingularity". In: Journal of the American Mathematical Society 8.2 (1995). A. Ya. Helemskii. Lectures and Exercises on Functional Analysis. Translations of mathe- matical monographs 233. American Mathematical Society, 2006. isbn: 0821840983, 9780821840986, 0821835521. [Hel86a] A. Ya. Helemskii. The homology of Banach and topological algebras. Mathematics and Its Applications. Kluwer Academic Publishers, 1986. isbn: 978-94-010-7560-2. [Hel86b] A. Ya. Helemskii. The homology of Banach and topological algebras. Mathematics and Its Applications. Kluwer Academic Publishers, 1986. isbn: 978-94-010-7560-2, 978-94-009 -2354-6. [Jar81] [Mal86] Hans Jarchow. Locally Convex Spaces. 1st ed. Mathematische Leitfaden. Vieweg+Teubner Verlag, 1981. isbn: 978-3-322-90561-1,978-3-322-90559-8. A. Mallios. Topological Algebras Selected Topics. North-Holland Mathematics Studies 124. Elsevier Science Ltd, 1986. [MDR97] Reinhold Meise, Vogt Dietmar, and M. S. Ramanujan. Introduction to functional analysis. Oxford graduate texts in mathematics 2. Clarendon Press; Oxford University Press, 1997. isbn: 0-1985-1485-9. [MR01] [Pir08a] [Pir08b] [Pir12] [PS94] [Sch93] [Tr´e70] [Wei94] J.C McConnell and J.C. Robson. Noncommutative Noetherian rings. 2nd ed. Graduate studies in mathematics 30. American Mathematical Society, 2001. isbn: 0-8218-2169-5. A. Yu. Pirkovskii. "Arens-Michael envelopes, homological epimorphisms, and relatively quasi-free algebras". In: Transactions of the Moscow Mathematical Society 69 (2008). A Yu Pirkovskii. "Weak homological dimensions and biflat Kothe algebras". In: Sbornik: Mathematics 199.5 (2008), p. 673. A. Yu. Pirkovskii. "Homological dimensions and Van den Bergh isomorphisms for nuclear Fr´echet algebras". In: Izvestiya: Mathematics 76.4 (2012). N. Christopher Phillips and Larry B. Schweitzer. "Representable K-theory of smooth crossed products by R and Z". In: Transactions of the American Mathematical Society 344.1 (1994), pp. 173 -- 201. Larry B. Schweitzer. "Dense m-convex Fr´echet subalgebras of operator algebra crossed products by Lie groups". In: International Journal of Mathematics 4.04 (1993), pp. 601 -- 673. Francois Tr´eves. Topological Vector Spaces, Distributions And Kernels. Academic Press, 1970. isbn: 9780126994506,0126994501,9780080873374. Charles A. Weibel. An introduction to homological algebra. Cambridge studies in advanced mathematics 38. Cambridge University Press, 1994. isbn: 0-5214-3500-5. P. Kosenko, Department of Mathematics, Higher School of Economics, Moscow, Usacheva str. 6 E-mail address, P. Kosenko: [email protected] 20
1110.4933
1
1110
2011-10-22T01:11:45
Duality, Cohomology, and Geometry of Locally Compact Quantum Groups
[ "math.FA" ]
In this paper we study various convolution-type algebras associated with a locally compact quantum group from cohomological and geometrical points of view. The quantum group duality endows the space of trace class operators over a locally compact quantum group with two products which are operator versions of convolution and pointwise multiplication, respectively; we investigate the relation between these two products, and derive a formula linking them. Furthermore, we define some canonical module structures on these convolution algebras, and prove that certain topological properties of a quantum group, can be completely characterized in terms of cohomological properties of these modules. We also prove a quantum group version of a theorem of Hulanicki characterizing group amenability. Finally, we study the Radon--Nikodym property of the $L^1$-algebra of locally compact quantum groups. In particular, we obtain a criterion that distinguishes discreteness from the Radon--Nikodym property in this setting.
math.FA
math
DUALITY, COHOMOLOGY, AND GEOMETRY OF LOCALLY COMPACT QUANTUM GROUPS MEHRDAD KALANTAR AND MATTHIAS NEUFANG Abstract. In this paper we study various convolution-type algebras associ- ated with a locally compact quantum group from cohomological and geomet- rical points of view. The quantum group duality endows the space of trace class operators over a locally compact quantum group with two products which are operator versions of convolution and pointwise multiplication, respectively; we investigate the relation between these two products, and derive a formula linking them. Furthermore, we define some canonical module structures on these convolution algebras, and prove that certain topological properties of a quantum group, can be completely characterized in terms of cohomologi- cal properties of these modules. We also prove a quantum group version of a theorem of Hulanicki characterizing group amenability. Finally, we study the Radon -- Nikodym property of the L1-algebra of locally compact quantum groups. In particular, we obtain a criterion that distinguishes discreteness from the Radon -- Nikodym property in this setting. 1. Introduction The most fundamental objects in abstract harmonic analysis are algebras of functions on a locally compact group G, endowed with the convolution, respectively, pointwise product, such as the group algebra L1(G) and the Fourier algebra A(G). Despite being dual to each other in a canonical way, these two products cannot be compared and linked to one another in an obvious way, because they live on very different spaces. However, as we shall show in this paper, it is the duality of locally compact quantum groups G that provides a common ground on which these two products can be studied simultaneously on one space, namely the trace class operators T (L2(G)). Our goal in this paper is to study locally compact quantum groups G from cohomological and geometrical points of view. The fact that the co-multiplication of a locally compact quantum group is implemented by its fundamental unitary, enables one to lift the product of L1(G) to T (L2(G)). Therefore, T (L2(G)) can be canonically endowed with two products which arise from G and G; in the classical case of a locally compact group G, these products are indeed operator versions of the convolution and the pointwise products. The paper is organized as follows. The preliminary definitions and results which are needed, are briefly recalled in section 2. In section 3, we first define the quantum version of the convolution and pointwise products on the space T (L2(G)) of trace class operators on the Hilbert space L2(G) of a locally compact quantum group G. We then study the basic properties of these algebras, and use the duality theory of locally compact quantum groups to derive a formula linking the two products associated with G and G. In section 4, we consider various module structures associated with convolution algebras over a locally compact quantum group, and investigate their cohomological properties. We show that topological properties of a locally compact quantum group 1 2 MEHRDAD KALANTAR AND MATTHIAS NEUFANG G are equivalent to cohomological properties of certain convolution algebras over G. In [18], the second-named author introduced and studied the above-mentioned convolution product on T (L2(G)) for a locally compact group G. The correspond- ing results on the equivalence of topological and cohomological properties in this situation were obtained in [21]. We also establish in this section a quantum group version of a theorem of Hulanicki stating that a discrete group is amenable if and only if its left regular representation is an isometry on positive elements of l1(G): indeed, we show that for any co-amenable locally compact quantum group G, the latter condition is equivalent to co-amenability of the dual G, i.e., to G having Reiter's property (P2), as introduced and studied in [4]. In the last section, for a locally compact quantum group G, we study a geomet- ric property of L1(G), namely the Radon -- Nikodym property (RNP). While, for a locally compact group G, the space L1(G) has the RNP if and only if G is discrete, the dual statement, with L1(G) replaced by the Fourier algebra A(G), is not true in general. So, the RNP and discreteness are not equivalent for arbitrary locally compact quantum groups. We characterize the difference between both properties in this general setting in terms of a covariance condition. The results in this paper are based on [12], written under the supervision of the second-named author. 2. Preliminaries We recall from [15] and [28] that a (von Neumann algebraic) locally compact quantum group G is a quadruple (L∞(G), Γ, ϕ, ψ), where L∞(G) is a von Neumann algebra with a co-multiplication Γ : L∞(G) → L∞(G)⊗L∞(G), , and ϕ and ψ are (normal faithful semifinite) left and right Haar weights on L∞(G), respectively. For each locally compact quantum group G, there exist a left funda- mental unitary operator W on L2(G, ϕ) ⊗ L2(G, ϕ) and a right fundamental unitary operator V on L2(G, ψ) ⊗ L2(G, ψ) which satisfy the pentagonal relation (2.1) The co-multiplication Γ on L∞(G) can be expressed as W12W13W23 = W23W12 and V12V13V23 = V23V12. Γ(x) = W ∗(1 ⊗ x)W = V (x ⊗ 1)V ∗ (2.2) We can identify L2(G, ϕ) and L2(G, ψ) (cf. [15, Proposition 2.11]), and we simply use L2(G) for this Hilbert space in the rest of this paper. (x ∈ L∞(G)). Let L1(G) be the predual of L∞(G). Then the pre-adjoint of Γ induces on L1(G) an associative completely contractive multiplication (2.3) ⋆ : f1 ⊗ f2 ∈ L1(G) ⊗L1(G) → f1 ⋆ f2 = (f1 ⊗ f2) ◦ Γ ∈ L1(G). A locally compact quantum group G is called co-amenable if L1(G) has a bounded left (equivalently, right or two-sided) approximate identity (cf. [2, Theorem 3.1]). The left regular representation λ : L1(G) → B(L2(G)) is defined by λ : L1(G) ∋ f 7→ λ(f ) = (f ⊗ ι)(W ) ∈ B(L2(G)), which is an injective and completely contractive algebra homomorphism from L1(G) into B(L2(G)). Then ′′ L∞( G) = {λ(f ) : f ∈ L1(G)} DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 3 is the von Neumann algebra associated with the dual quantum group G. Anal- ogously, we have the right regular representation ρ : L1(G) → B(L2(G)) defined by ρ : L1(G) ∋ f 7→ ρ(f ) = (ι ⊗ f )(V ) ∈ B(L2(G)), which is also an injective and completely contractive algebra homomorphism from L1(G) into B(L2(G)). Then ′′ L∞( G′) = {ρ(f ) : f ∈ L1(G)} is the von Neumann algebra associated with the dual (commutant) quantum group G′. It follows that W ∈ L∞(G) ¯⊗L∞( G) and V ∈ L∞( G′) ¯⊗L∞(G). We obtain the corresponding reduced quantum group C∗-subalgebra C0(G) = {(ι ⊗ f )(W ) : f ∈ L1( G)} k·k = { f ′ ⊗ ι)(V ) : f ′ ∈ L1( G′) k·k of L∞(G) with the co-multiplication Γ : C0(G) → M (C0(G) ⊗ C0(G)), where M (C0(G)⊗C0(G)) is the multiplier algebra of the minimal C∗-algebra tensor product C0(G) ⊗ C0(G). Let M (G) denote the operator dual C0(G)∗ of C0(G). The space M (G) is a completely contractive dual Banach algebra (i.e., the multiplication on M (G) is separately weak∗ continuous), and M (G) contains L1(G) as a norm closed two- sided ideal via the embedding L1(G) → M (G) : f 7→ fC0(G). If G is a locally compact group, then C0(Ga) is the C∗-algebra C0(G) of con- tinuous functions on G vanishing at infinity, and M (Ga) is the measure algebra M (G) of G. Correspondingly, C0( Ga) is the left group C∗-algebra C∗ λ(G) of G, and C0( G′ ρ (G) of G. Hence, we have M ( Ga) = Bλ(G) and M ( G′ a) is the right group C∗-algebra C∗ a) = Bρ(G). We also briefly recall some standard definitions and notations from the cohomol- ogy theory of Banach algebras (cf. [9]). Actually, as one might expect, here in the general setting of locally compact quantum groups, we need to take the quantum (operator space) structure of the underlying Banach spaces into account as well. So we work in the category of operator spaces; we shall define our module structures, and their corresponding objects, in the quantized Banach space category as well. A completely bounded linear map σ : X → Y from an operator space X into an operator space Y is called admissible if it has a completely bounded right inverse. Let A be a Banach algebra and P be a right A-module. P is called projective if for all A-modules X and Y , any admissible morphism σ : X → Y , and any morphism ρ : P → Y , there exists a morphism φ : P → X such that σ ◦ φ = ρ. Denote by X ⊳ A the closed linear span of the set {x ⊳ a : a ∈ A, x ∈ X} ⊆ X. Then X is called essential if X ⊳ A = X. Many categorical statements which hold in the category of Banach spaces, also hold in this setting with an obvious slight categorical modification. In particular, 4 MEHRDAD KALANTAR AND MATTHIAS NEUFANG the following result which is well-known in the classical setting (cf. [8]); we will use it frequently in our work. Theorem 2.1. an essential right A-module X is projective if and only if there exists a morphism ψ : X → Xb⊗A such that m ◦ ψ = ιX , where m : Xb⊗A → X is the canonical module action morphism, and Xb⊗A is regarded as a right A-module, via the action (x ⊗ a) ⊳ b = x ⊗ ab. The case of left modules and bi-modules are analogous. 3. Convolution and Pointwise Products for Locally Compact Quantum Groups In this section we define a quantum analogous of the convolution and pointwise products for a locally compact quantum group, study the basic properties, and state a formula linking them. Let G be a locally compact quantum group, and V ∈ L∞( G′)⊗L∞(G) its right fundamental unitary. We can lift the co-products Γ and Γ to B(L2(G)), still using the same notation, as follows: Γ : B(L2(G)) → B(L2(G))⊗B(L2(G)), Γ : B(L2(G)) → B(L2(G))⊗B(L2(G)), x 7→ V (x ⊗ 1)V ∗; x 7→ V ′(x ⊗ 1) V ′∗. Then the preadjoint maps Γ∗, Γ∗ : T (L2(G))b⊗T (L2(G)) → T (L2(G)) define two different completely contractive products on the space of trace class operators T (L2(G)). We denote these products by ⋆ and • respectively. We also denote by T⋆(G) and T•(G) the (quantized) Banach algebras (T (L2(G)), ⋆) and (T (L2(G)), •), respectively. If G = L∞(G) for a locally compact group G, then T⋆(G) is the convolution algebra introduced by Neufang in [18]. Applied to the classical setting, i.e., the commutative and co-commutative cases, the following lemma justifies why the above products are considered as quantum versions of convolution and point-wise products. Proposition 3.1. The canonical quotient map π : T⋆(G) ։ L1(G) and the trace map tr : T⋆(G) → C are Banach algebra homomorphisms. Proof. First part follows from the fact that Γ(L∞(G)) ⊆ L∞(G)⊗L∞(G), and the second part is an easy consequence of the identity Γ(1) = 1⊗1. (cid:3) The above Proposition allows us to define (right) T⋆(G)-module structures on L1(G) and C as follows: f ⊳ ρ = f ⋆ π(ρ) and λ ⊳ ρ = λtr(ρ), where ρ ∈ T⋆(G), f ∈ L1(G), and λ ∈ C. We will show later that some of the topological properties of G can be deduced from these module structures. But, first we prove some properties of the lifted co-products and their induced products. Proposition 3.2. Let x ∈ B(L2(G)). If Γ(x) = y ⊗ 1 for some y ∈ B(L2(G)) then we have x = y ∈ L∞( G). DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 5 Proof. We have (3.1) ((ι ⊗ ω)V )x = (ι ⊗ ω)(V (x ⊗ 1)) = (ι ⊗ ω)((y ⊗ 1)V ) = y((ι ⊗ ω)V ) for all ω ∈ T (L2(G)). Since L∞( G′) = {(ι ⊗ ω)V : ω ∈ T (L2(G))} w∗ , it follows from (3.1) that a′x = ya′ for all a′ ∈ L∞( G′). a′x = xa′ for all a′ ∈ L∞( G′), we have x ∈ L∞( G). In particular for a′ = 1, it follows that x = y, and since (cid:3) Lemma 3.3. Let x ∈ B(L2(G)). If Γ(x) = 1 ⊗ y, for some y ∈ B(L2(G)), then x ∈ L∞( G′) and y ∈ L∞(G). Proof. Since (1 ⊗ y) = V (x ⊗ 1)V ∗ ∈ B(L2(G))⊗L∞(G), we have y ∈ L∞(G). Similarly, x ⊗ 1 = V ∗(1 ⊗ y)V ∈ L∞( G′)⊗L∞(G) implies that x ∈ L∞( G′). (cid:3) Corollary 3.4. Let x ∈ B(L2(G)). If Γ(x) = 1 ⊗ x, then x ∈ C1. Proof. If Γ(x) = 1 ⊗ x, the Lemma 3.3 implies that x ∈ L∞(G) ∩ L∞( G′) which equals C1. (cid:3) Now we investigate the relation between these two products on T (L2(G)), and find a formula (3.2) linking them. Proposition 3.5. For ρ, ξ and η in T (L2(G)), the following two relations hold: ρ ⋆ (ξ • η) = hη, 1iρ ⋆ ξ; ρ • (ξ ⋆ η) = hη, 1iρ • ξ. Proof. Let x ∈ L∞(G) and x ∈ L∞( G). Then we have hρ ⋆ (ξ • η), xxi = hρ ⊗ (ξ • η), Γ(xx)i = hρ ⊗ ξ ⊗ η, (ι ⊗ Γ)(Γ(xx))i = hρ ⊗ ξ ⊗ η, (Γ(x) ⊗ 1)(x ⊗ 1 ⊗ 1)i = hη, 1ihρ ⊗ ξ, Γ(x)(x ⊗ 1)i = hη, 1ihρ ⋆ ξ, xxi, which, by weak∗ density of the span of the set {xx : x ∈ L∞(G), x ∈ L∞( G)} in B(L2(G)), implies the first formula. The second relation follows along similar lines. (cid:3) Since there are two different multiplications on T (L2(G)) arising from G and G, it is tempting to consider the corresponding two actions at the same time by defining a bi-module structure on T (L2(G)), using these two products. But one can deduce from the above proposition, that multiplication from the left and right via these products, is not associative, and so we cannot turn T (L2(G)) into a T⋆(G) − T•(G) bimodule in this fashion. However, next theorem will provide us with a way of doing so. 6 MEHRDAD KALANTAR AND MATTHIAS NEUFANG Theorem 3.6. For ρ, ξ and η in T (L2(G)), the following relation holds: (3.2) (ρ ⋆ ξ) • η = (ρ • η) ⋆ ξ. Equivalently denoting by m and m the product maps corresponding to ⋆ and •, respectively, we have m ◦ ( m⊗ι) = m ◦ (m⊗ι) ◦ (ι⊗σ) on the triple (operator space) projective tensor product of T (L2(G)) with itself; here σ is the flip map. Remark 3.7. This theorem shows that the dual products on quantum groups "anti- commute": the minus sign of a usual anti-commutation relation in an algebra (with respect to a given product) is replaced by the flip map when comparing two different products. Proof. Let x ∈ L∞(G) and x ∈ L∞( G). Then we have: h(ρ ⋆ ξ) • η, xxi = h(ρ ⋆ ξ) ⊗ η, Γ(xx)i = hρ ⊗ ξ ⊗ η, (Γ ⊗ ι)[(x ⊗ 1)Γ(x)]i = hρ ⊗ ξ ⊗ η, (Γ(x) ⊗ 1)Γ(x)13i = hρ ⊗ η ⊗ ξ, Γ(x)13(Γ(x) ⊗ 1)i = hρ ⊗ η ⊗ ξ, (Γ ⊗ ι)[Γ(x)(x ⊗ 1)]i = h(ρ • η) ⊗ ξ, Γ(xx)i = h(ρ • η) ⋆ ξ, xxi. Hence, theorem follows, again by weak∗ density of the span of the set {xx : x ∈ L∞(G), x ∈ L∞( G)} ⊆ B(L2(G)). (cid:3) Theorem 3.6 has even more significance: the quantum group duality may be encoded by this relation. In fact, one might be able to start from this relation on trace class operators on a Hilbert space, with some extra conditions, to arrive to an equivalent axiomatic definition for locally compact quantum groups. We intend to address this project in a subsequent paper. Proposition 3.8. The space T (L2(G)) becomes a T⋆(G)op − T•(G) bimodule via the actions where ρ ∈ T (L2(G)), η ∈ T⋆(G) and ξ ∈ T•(G). η ⊲ ρ = ρ ⋆ η and ρ ⊳ ξ = ρ • ξ, Proof. We only need to check the associativity of the left-right action. For this, using Theorem 3.6, we obtain (η ⊲ ρ) ⊳ ξ = (ρ ⋆ η) • ξ = (ρ • ξ) ⋆ η = η ⊲ (ρ ⊳ ξ). (cid:3) The following proposition is known and has been stated in many different places. Proposition 3.9. Let G be a locally compact quantum group. Then the following hold: (1) L1(G) has a left (right) identity if and only if G is discrete. DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 7 (2) L1(G) has a bounded left (right) approximate identity if and only if G is co-amenable. In contrast to the last proposition, we have the following. Proposition 3.10. Let G be a locally compact quantum group. Then the following hold: (1) T⋆(G) does not have a left identity, unless G is trivial, and it has a right identity if and only if G is discrete; (2) T⋆(G) does not have a left approximate identity, unless G is trivial, and it has a bounded right approximate identity if and only if G is co-amenable. (1): let ω0 ∈ T (L2(G)) be a non-zero normal functional whose restriction Proof. to L∞(G) is zero. Since we have it follows that Γ(B(L2(G))) ⊆ B(L2(G))⊗L∞(G), hρ ⋆ ω0, xi = hω0, (ρ⊗ι)Γ(x)i = 0, for all ρ ∈ T (L2(G)) and x ∈ B(L2(G)), which obviously implies that there does not exist a left identity, unless G is trivial (equal to C). Now, let G be discrete, e ∈ L1(G) be the unit element, and e ∈ T (L2(G)) be a norm preserving weak∗-extension of e. Then we have hρ ⋆ e, xxi = hρ ⊗ e, Γ(x)(x ⊗ 1)i = hxρ ⊗ e, Γ(x)i = hπ(xρ) ⋆ e, xi = hπ(xρ), xi = hxρ, xi = hρ, xxi for all ρ ∈ T⋆(G), x ∈ L∞(G) and x ∈ L∞( G), where π : T (L2(G)) ։ L1(G) is the canonical quotient map. Since the span of the set {xx : x ∈ L∞(G), x ∈ L∞( G)} is weak∗ dense in B(L2(G)), it follows that e is a right identity for T⋆(L2(G)). Conversely, assume that T⋆(G) has a right identity e. Then, since by Proposition 3.1, the map π : T⋆(G) → L1(G) is a surjective homomorphism, π(e) is clearly a right identity for L1(G), whence G is discrete by Proposition 3.9. (2): Similarly to the first part, one can show that T⋆(G) cannot possess a left approximate identity, unless it is trivial (equal to C). Let G be co-amenable. Then, by [2, Theorem 3.1], there exists a net (ξi) of unit vectors in L2(G) such that kV ∗(η⊗ξi) − η⊗ξik → 0 for all unit vectors η ∈ L2(G). Now, for all x ∈ B(L2(G)) and η ∈ L2(G) with kxk = kηk = 1 we have hωη ⋆ ωξi − ωη, xi = hV (x⊗1)V ∗(η⊗ξi), η⊗ξii − h(x⊗1)(η⊗ξi), η⊗ξii = h(x⊗1)(V ∗(η⊗ξi) − η⊗ξi), V ∗(η⊗ξi)i + h(x⊗1)(η⊗ξi), V ∗(η⊗ξi) − η⊗ξii ≤ 2kV ∗(η⊗ξi) − η⊗ξik → 0. Since the span of the set {ωη : η ∈ L2(G)} is norm dense in T (L2(G)), it follows that (ωξi ) is a right bounded approximate identity for T⋆(G). Conversely, if T⋆(G) has a right bounded approximate identity, then a similar (cid:3) argument to the proof of part (1) shows that G is co-amenable. 8 MEHRDAD KALANTAR AND MATTHIAS NEUFANG Proposition 3.11. A locally compact quantum group G is compact if and only if there exists a state ϕ ∈ T⋆(G) such that (3.3) for all x ∈ L∞(G), x ∈ L∞( G) and ρ ∈ T⋆(G). hρ ⋆ ϕ, xxi = hρ ⋆ ϕ, xxi = hρ, xih ϕ, xi Proof. Suppose that G is compact with normal Haar state ϕ, and ϕ ∈ T⋆(G) is a norm preserving extension of ϕ. Then ϕ is a state (since k ϕk = ϕ(1) = 1), and we have hρ ⋆ ϕ, xxi = hρ ⊗ ϕ, Γ(x)(x ⊗ 1)i = hxρ ⊗ ϕ, Γ(x)i = hπ(xρ) ⊗ π( ϕ), Γ(x)i = hπ(xρ) ⋆ ϕ, xi = hπ(xρ), 1ihϕ, xi = hxρ, 1ihϕ, xi = hρ, xihϕ, xi = hρ, xih ϕ, xi. In a similar way, we can show that hρ ⋆ ϕ, xxi = hρ, xih ϕ, xi. Conversely, suppose such a state ϕ ∈ T⋆(G) exists. Let ϕ = π( ϕ) ∈ L1(G)+ and f ∈ L1(G), and let f ∈ T⋆(G) be a weak∗-extension of f . Then, by putting x = 1 in equation (3.3), we have hf ⋆ ϕ, xi = h f ⋆ ϕ, xi = h f , 1ih ϕ, xi = hf, 1ihϕ, xi for all x ∈ L∞(G). Hence, ϕ is a left invariant state in L1(G), and so G is compact, by [2, Proposition 3.1]. (cid:3) 4. Cohomological Properties of Convolution Algebras The following result was proved in the more general setting of Hopf-von Neumann algebras in [1, Theorem 2.3]. Proposition 4.1. C is a projective T⋆(G)-module if and only if G is compact. In the following, we want to prove a statement similar to Proposition 4.1, for discreteness of G. But the situation is more subtle in this case. There are some technical difficulties which arise when one tries to link the quantum group structure to the quantum Banach space structure. This happens mainly because the latter is essentially defined based on the Banach space structure of these algebras, and do not seem to see all aspects of the quantum group structure. These technical issues appear also in some of the open problems in this theory, and seem to be a major subtle point (c.f. [3]). To avoid such difficulties, in the rest of this section, we assume that the mor- phisms are completely contractive, rather than just completely bounded. Theorem 4.2. Let G be a locally compact quantum group. Then the following are equivalent: (1) there exists a normal conditional expectation E : B(L2(G)) → L∞(G) which satisfies Γ ◦ E = (E ⊗ E)Γ; (2) there exists a normal conditional expectation E : B(L2(G)) → L∞(G) which satisfies Γ ◦ E = (E ⊗ ι)Γ; (3) there exists a normal conditional expectation E : B(L2(G)) → L∞(G) which satisfies E(L∞( G)) ⊆ C1; (4) G is discrete. DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 9 Proof. (1) ⇔ (2) : This follows from the facts that E = ι on L∞(G), and that Γ(B(L2(G))) ⊆ B(L2(G))⊗L∞(G). (1) ⇒ (3) : Let x ∈ L∞( G). We have: Γ(E(x)) = (E ⊗ E)Γ(x) = (E ⊗ E)(x ⊗ 1) = E(x) ⊗ 1, which implies that E(x) ∈ C1. (3) ⇒ (4) : Assumption (3) implies that E ∈ CBσ,L∞( G) L∞(G) (B(L2(G))), and hence it follows by [11, Theorem 4.5] that there exists a right centralizer m ∈ cb(L1( G)) such that C r E = Θr( m). Now, define a complex-valued map f on L∞( G) such that E(x) = f (x)1 for all x ∈ L∞( G). Since E is a unital linear normal positive map, f is a normal state on L∞( G), and for every ω ∈ L1( G) and x ∈ L∞( G) we have h m(ω), xi = hω, Θr( m)(x)i = hω, E(x)i = hω, f (x)1i = hω, 1i f (x). Hence m(ω) = hω, 1i f . Now, fix ω0 ∈ L1( G), then for all ω ∈ L1( G) we have ω ⋆ f = ω ⋆ m(ω0) = m(ω ⋆ ω0) = hω ⋆ ω0, 1i f = hω, 1i f . Hence, f is a normal left invariant state on L∞( G), and therefore G is compact by [2, Proposition 3.1], and (4) follows. (4) ⇒ (2) : Let e be the identity of L1(G), and let e ∈ T (L2(G)) be a norm- preserving extension of e. Define: E : B(L2(G)) → L∞(G); x 7→ (e⊗ι)Γ(x). Then E is normal, unital and completely contractive, since both (e⊗ι) and Γ are, which also implies that kEk = 1. For all x ∈ L∞(G) and f ∈ L1(G) we have hf, E(x)i = hf, (e⊗ι)Γ(x)i = he⊗f, xi = hf, xi, which implies that E2 = E, and E is surjective. Hence, E is a conditional expec- tation on L∞(G). Now, for all x ∈ B(L2(G)), we have Γ(E(x)) = Γ((e⊗ι)Γ(x)) = (e⊗ι⊗ι)((ι⊗Γ)Γ(x)) = (e⊗ι⊗ι)((Γ⊗ι)Γ(x)) = (((e⊗ι)Γ)⊗ι)Γ(x) = (E⊗ι)Γ(x). Hence, Γ ◦ E = (E⊗ι)Γ, and (2) follows. (cid:3) Remark 4.3. One can easily modify the above proof to obtain a right version of Theorem 4.2; then in part (2) we have Γ ◦ E = (ι⊗E)Γ, and in part (3), E(L∞( G′)) ⊆ C1. Corollary 4.4. For a locally compact quantum group G the following are equiva- lent: (1) there exists an isometric algebra homomorphism Φ : L1(G) → T⋆(G) such that π ◦ Φ = ιL1(G); (2) G is discrete. 10 MEHRDAD KALANTAR AND MATTHIAS NEUFANG Proof. If G is discrete, then Φ may be taken to be the pre-adjoint of the map E constructed in the proof of the implication (4) ⇒ (2) in Theorem 4.2. For the converse, note that Φ∗ : B(L2(G)) → L∞(G) is a normal surjective norm- one projection, i.e., a normal conditional expectation. Moreover, for x ∈ B(L2(G)) and ρ, η ∈ T⋆(G) we have hρ ⊗ η, Γ(Φ∗(x))i = hρ ⋆ η, Φ∗(x)i = hΦ(ρ ⋆ η), xi = hΦ(ρ) ⋆ Φ(η), xi = hΦ(ρ) ⊗ Φ(η), Γ(x)i = hρ ⊗ η, (Φ∗ ⊗ Φ∗)Γ(x)i, which implies Γ ◦ Φ∗ = (Φ∗ ⊗ Φ∗)Γ, and hence the theorem follows from Theorem 4.2. (cid:3) As we promised earlier in this section, in the following (Theorem 4.7), we prove that discreteness of a locally compact quantum group G, can also be characterized in terms of projectivity of its convolution algebras. We recall that here the morphisms canonical map associated with the module action. are completely contractive maps, and m : L1(G)b⊗T⋆(G) → L1(G) denotes the Lemma 4.5. If Φ : L1(G) → T⋆(G) is such that π ◦ Φ = ιL1(G), then for all η, ρ ∈ T⋆(G) we have: η ⋆ Φ(π(ρ)) = η ⋆ ρ. Proof. Recall that Γ(x) ∈ B(L2(G))⊗L∞(G) for all x ∈ B(L2(G)). Therefore, we clearly obtain that η ⋆ ρ = η ⋆ π(ρ). Hence, we have η ⋆ Φ(π(ρ)) = η ⋆ π(Φ(π(ρ))) = η ⋆ π(ρ) = η ⋆ ρ. (cid:3) Lemma 4.6. Assume that Ψ : L1(G)b⊗T⋆(G) is a T⋆(G)-module morphism which satisfies m ◦ Ψ = ιL1(G). Then for any x ∈ L∞(G) we have Ψ∗(1⊗x) ∈ C1. Proof. Let x ∈ L∞(G), then we have hf ⋆ π(ρ), Ψ∗(1⊗x)i = hΨ(f ⋆ π(ρ)), 1⊗xi = hΨ(f ⊳ ρ), 1⊗xi = hΨ(f ) ⊳ ρ, 1⊗xi = hΨ(f ) ⋆ ρ, 1⊗xi = hΨ(f )⊗ρ, 1⊗Γ(x)i = hΨ(f )⊗ρ, 1⊗x⊗1i = hΨ(f ), 1⊗xihρ, 1i = hf, Ψ∗(1⊗x)ihρ, 1i for all f ∈ L1(G) and ρ ∈ T⋆(G). Since π : T⋆(G) → L1(G) is surjective, we get hf ⊗g, Γ(Ψ∗(1⊗x))i = hf ⋆ g, Ψ∗(1⊗x)i = hf, Ψ∗(1⊗x)ihg, 1i = hf ⊗g, Ψ∗(1⊗x)⊗1i for all f, g ∈ L1(G). Hence we have Γ(Ψ∗(1⊗x)) = Ψ∗(1⊗x)⊗1, which implies that Ψ∗(1⊗x) ∈ C1. (cid:3) Theorem 4.7. For a locally compact quantum group G, the following are equiva- lent: (1) L1(G) is a projective T⋆(G)-module; DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 11 (2) G is discrete. Proof. (1) ⇒ (2) : Assume that L1(G) is a projective T⋆(G)-module. So, there exists a T⋆(G)-module morphism Ψ : L1(G) → L1(G)b⊗T⋆(G) such that m ◦ Ψ = ιL1(G). Let R be the unitary antipode of G (cf. [14]). Then R(x) = J x J for all x ∈ L∞(G), where J is the modular conjugate associated with the dual Haar weight ϕ. Using the same formula x 7→ J x J (cid:0)x ∈ B(L2(G))(cid:1), we can extend the map R to B(L2(G)). Then it is clear that R(L∞( G′)) = L∞( G). Denote by χ the flip map a⊗b 7→ b⊗a, and define the map T : B(L2(G)) → L∞(G), T := Ψ∗ ◦ χ ◦ (R⊗R) ◦ Γ. We shall prove that the map E := T 2 satisfies the conditions of (the right version of) part (3) of Theorem 4.2 (see the Remark 4.3). First note that T is normal and contractive. Moreover, for x ∈ L∞(G), we have T (x) = Ψ∗ ◦ χ ◦ (R⊗R) ◦ Γ(x) = Ψ∗ ◦ Γ(R(x)) = R(x). This implies that T 2 L∞(G) = ι, and so E2 = E. Hence, E : B(L2(G)) → L∞(G) is a normal conditional expectation on L∞(G). Now, for all x′ ∈ L∞( G′) we have T (x′) = Ψ∗ ◦ χ ◦ (R⊗R) ◦ Γ(x′) = Ψ∗ ◦ χ ◦ (R⊗R)(x′⊗1) = Ψ∗ ◦ χ(R(x′)⊗1) = Ψ∗(1⊗R(x′)). Hence, E(L∞( G′)) ⊆ C1, by Lemma 4.6, and so G is discrete by (the right version of) Theorem 4.2. (2) ⇒ (1) : Let e ∈ L1(G) be the identity element, and Φ : L1(G) → T⋆(G), as in Corollary 4.4. Define the map Ψ : L1(G) → L1(G)b⊗T⋆(G) by Ψ(f ) = e⊗Φ(f ) (f ∈ L1(G)). Since π ◦ Φ = ιL1(G), we have m ◦ Ψ = ιL1(G). Moreover, using Lemma 4.5, we have Ψ(f ⊳ ρ) = e⊗Φ(f ⊳ ρ) = e⊗Φ(f ⋆ π(ρ)) = e⊗(cid:0)Φ(f ) ⋆ Φ(π(ρ))(cid:1) = e⊗(cid:0)Φ(f ) ⋆ ρ(cid:1) = Ψ(f ) ⊳ ρ for all f ∈ L1(G) and ρ ∈ T⋆(G). Therefore Ψ is a morphism, and so L1(G) is projective. (cid:3) The next theorem was proved for the case of Kac algebras in [7, Theorem 6.6.1], but the proof in there is based on the structure theory of discrete Kac algebras. Here we present a different argument for the general case of locally compact quantum groups. Theorem 4.8. If G is both compact and discrete, then G is finite (dimensional). 12 MEHRDAD KALANTAR AND MATTHIAS NEUFANG Proof. If G is compact, then L1(G) is an ideal in L1(G)∗∗ with the left (equivalently, right) Arens product, by [24, Theorem 3.8]. But since G is also discrete, L1(G) is unital, and its unit is obviously also an identity element for the left Arens product of L1(G)∗∗. Being a unital ideal (via the canonical embedding), L1(G) must be equal to L1(G)∗∗. So L1(G) is reflexive, hence L∞(G) is, which implies that L∞(G) is finite-dimensional, by [16, Proposition 1.11.7]. (cid:3) Using Theorem 4.8, we can now follow a similar idea as the proof of [21, Theorem 3.7], to prove a quantum version of the latter. Theorem 4.9. Let G be a co-amenable locally compact quantum group. Then the following are equivalent: (1) T⋆(G) is biprojective; (2) G is finite. Proof. (1) ⇒ (2) : Since G is co-amenable, Proposition 3.10 implies that T⋆(G) has a bounded right approximate identity. Since L1(G) and C are both essential T⋆(G)-modules, they are T⋆(G)-projective by [9, 7.1.60], which implies that G is both compact and discrete, by Proposition 4.1 and Theorem 4.7. Hence, G is finite by Theorem 4.8. (2) ⇒ (1) : Consider the short exact sequence where 0 → I → T⋆(G) π−→ L1(G) → 0, I := {ρ ∈ T (L2(G)) : ρL∞(G) = 0}. Since G is finite, it is in particular a compact Kac algebra, and so L1(G) is operator biprojective. Hence, (1) follows from [21, Lemma 4.2]. (cid:3) We can also define a right L1(G)-module structure on T (L2(G)), as follows: (4.1) ρ ⊳ f := (ρ ⊗ f ) ◦ Γ (ρ ∈ T (L2(G)), f ∈ L1(G)). Theorem 4.10. For a locally compact quantum group G, the following are equiv- alent: (1) there exists an isometric L1(G)-module map Φ : L1(G) → T (L2(G)) such that π ◦ Φ = ιL1(G); (2) G is discrete. Proof. If G is discrete, then the predual of the map E constructed in the proof of the implication (4) ⇒ (2) in Theorem 4.2, is easily seen to satisfy the desired conditions. Conversely, if such a map Φ exists, then it is straightforward to see that the map E := Φ∗ : B(L2(G)) → L∞(G) enjoys the properties in part (2) of Theorem 4.2, and so G is discrete. (cid:3) In the following, we shall consider another important cohomology-type property for the convolution algebras associated with a locally compact quantum group G, namely amenability. Next theorem is in fact a generalization of a result due to Hulanicki who consid- ered the case G = L∞(G) for a discrete group G, to the setting of locally compact quantum groups. This result was proved in the Kac algebra case by Kraus and Ruan in [13, Theorem 7.6]. But their argument is based essentially on the fact that DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 13 in the Kac algebra setting, the left regular representation is a ∗-homomorphism, which does not hold anymore in the general setting of locally compact quantum groups, so it appears that their proof cannot be modified for the latter case. Here we present a different argument, inspired by the proof of [22, Theorem 2.4]. Theorem 4.11. Let G be a co-amenable locally compact quantum group. Then the following are equivalent: (1) the left regular representation λ : L1(G) → L∞( G) is isometric on L1(G)+; (2) G is co-amenable, i.e., G has Reiter's property (P2). Proof. (1) ⇒ (2) : We first show that (1) implies that k1 + λ(f )k = 1 + kf k for all f ∈ L1(G)+. To show this, let (eα) ∈ L1(G)+ be a bounded approximate identity, and g ∈ L1(G)+ with keαk = kgk = 1 for all α. Then we have kf k + 1 = f (1) + 1 = f (1)g(1) + eα(1)g(1) = hf ⋆ g + eα ⋆ g, 1i = kf ⋆ g + eα ⋆ gk → kf ⋆ g + gk = kλ(f ⋆ g) + λ(g)k = k(λ(f ) + 1)λ(g)k ≤ kλ(f ) + 1k kλ(g)k = kλ(f ) + 1k ≤ kλ(f )k + 1 = kf k + 1, which implies our claim. Since G is co-amenable, there exists ε ∈ M (G)+ λ(ε) = 1, by [2, Theorem 3.1]. Let 1 such that F0 = {F ∪ {ε} : F ⊆ L1(G) + 1 , F is finite}. Then for each F ∈ F0 we have So (cid:13)(cid:13)Pf ∈F λ(f )(cid:13)(cid:13) = F , and therefore there exists a sequence (ξn) of unit vectors in L2(G) such that (cid:13)(cid:13) X f ∈F f(cid:13)(cid:13) = hX f ∈F f, 1i = F . lim n (cid:13)(cid:13) X f ∈F λ(f )ξn(cid:13)(cid:13) = F . Now fix f0 ∈ F , and let F ′ = F \{ε, f0}. Then we have but since lim n (cid:13)(cid:13)(λ(f0)ξn + ξn) + X (cid:13)(cid:13) X λ(f )ξn(cid:13)(cid:13) ≤ F − 2 f ∈F ′ f ∈F ′ λ(f )ξn(cid:13)(cid:13) = F , ∀n ∈ N, it follows that limn kλ(f0)ξn + ξnk = 2, which yields lim n kλ(f0)ξn − ξnk2 = 0. Since both f0 ∈ F and F ∈ F0 were arbitrary, there exists a net (ξi) of unit vectors in L2(G) such that for all f ∈ L1(G)+ 1 , and since L∞(G) is standard on L2(G) we have kλ(f )ξi − ξik → 0 kW (η ⊗ ξi) − η ⊗ ξik → 0 14 MEHRDAD KALANTAR AND MATTHIAS NEUFANG for all unit vectors η ∈ L2(G). Hence, G is co-amenable, by [2, Theorem 3.1]. (2) ⇒ (1) : Let χ denote the flip map a⊗b 7→ b⊗a. Since G is co-amenable and χ(W ) is an isometry, [2, Theorem 3.1] ensures the existence of a net (ξi) of unit vectors in L2(G) such that kW (η ⊗ ξi) − η ⊗ ξik2 = 0, ∀η ∈ L2(G). lim i Now, let f ∈ L1(G)+. Since L∞(G) is in standard form in B(L2(G)), we have f = ωζ, for some ζ ∈ L2(G) with kf k = kζk. Assuming kλ(f )k ≤ 1, we obtain: 1 ≥ lim i hλ(f )ξi, ξii = lim i h(f ⊗ ι)W ξi, ξii = lim i hW (ζ ⊗ ξi), ζ ⊗ ξii = lim i kζ ⊗ ξik2 = kζk2 = kf k2. So, kλ(f )k ≤ 1 implies kf k ≤ 1, therefore the conclusion follows. The equivalence between co-amenability of G and Reiter's property (P2) of G is (cid:3) the statement of [4, Theorem 5.4]. Remark 4.12. Note that the assumption of co-amenability of G is not necessary for the implication (2) ⇒ (1). Also, this condition is not necessary if G is a Kac algebra, as an easy modification of our argument shows. 5. The Radon -- Nikodym Property for L1(G) In the last part of this paper we shall investigate a geometric property for the convolution algebra L1(G) of a locally compact quantum group G, namely the Radon -- Nikodym property (in short: RNP). The following are some well-known results concerning the Radon -- Nikodym Prop- erty of Banach spaces (cf. [5]). Proposition 5.1. (1) The RNP is inherited by closed subspaces, and is stable under isomor- phisms. (2) If H is a Hilbert space then T (H) has the RNP. (3) Let G be a locally compact group. Then L1(G) has the RNP if and only if G is discrete. Proposition 5.2. Let G be a locally compact quantum group. If there exists f ∈ L1(G)1 such that the map L1(G) ∋ ω 7→ f ⋆ ω ∈ L1(G) is isometric, then L1(G) has the RNP. Proof. Assume that such f ∈ L1(G) exists. Let f ∈ T (L2(G)) be a norm-preserving weak∗ extension of f . Then the map L1(G) ∋ ω 7→ f ⊳ ω ∈ T⋆(G) is an isometric embedding, where the action ⊳ is defined as 4.1. To see this, let ω ∈ L1(G). Then we have kωk = kf ⋆ ωk = kπ( f ⊳ ω)k ≤ k f ⊳ ωk ≤ k f k kωk = kωk. This implies that L1(G) is isomorphic to a subspace of T (L2(G)), hence the claim follows from parts (1) and (2) of Proposition 5.1. (cid:3) DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 15 Part (3) of Proposition 5.1, at first glance, suggests that one might have a dual version of this statement, saying that the Fourier algebra A(G) has the RNP if and only if G is compact. But in fact, this is not the case. A counter-example is given by the Fell group (see [25, Remark 4.6]) which is non-compact, but its Fourier algebra has the RNP. Analogously to our earlier discussion on cohomological properties of G, one may need to take the operator space structure of L1(G) into account as well. Indeed, there is an operator space version of the RNP, due to Pisier (see [23]), which may be useful in this context. But, in the following we point out another way of looking at this problem. First, we give a general result. Theorem 5.3. Let M ⊆ B(H) be a von Neumann algebra. Then the following are equivalent: (1) M∗ has the RNP; (2) there is a normal conditional expectation from B(H) onto M . Proof. (1) ⇒ (2) : By [25, Theorem 3.5], M is atomic, i.e., M is an l∞-direct sum of B(Hi)'s for some Hilbert spaces Hi. So M = N ∗∗ where N = ⊕∞K(Hi) is an ideal in M . Then, we have (2) by [27, Theorem 5]. (2) ⇒ (1) : The pre-adjoint map of the conditional expectation defines an isometric embedding of M∗ into T (H). Then in view of parts (1) and (2) of Proposition 5.1, we obtain (2). (cid:3) Corollary 5.4. Let G be locally compact quantum group. Then the following are equivalent: (1) there exists a normal conditional expectation E from B(L2(G)) onto L∞(G); (2) L1(G) has the RNP. In Theorem 4.2 we gave a characterization of discreteness of G in terms of exis- tence of a normal and covariant conditional expectation. By comparing that result with Corollary 5.4 above, we see that the covariance accounts precisely for the difference between the RNP and discreteness, for L1-algebras of locally compact quantum groups. Next theorem shows that although discreteness and the RNP are not equivalent in general for L1(G), but with extra conditions on G, that could be the case. Theorem 5.5. Let G be a compact Kac algebra. Then the following are equivalent: (1) L1(G) has the RNP; (2) G is finite. (1) implies, by [25, Theorem 3.5], that L∞(G) is Proof. (2) ⇒ (1) is obvious. atomic, i.e., L∞(G) = ⊕∞B(Hi) for some Hilbert spaces Hi. Since G is a compact Kac algebra, the Haar weight ϕ is a finite faithful trace, hence all Hi's are finite- dimensional and the restriction of ϕ to each B(Hi) is its unique trace. Thus, ϕ = Pi tri, and since ϕ is finite, there can be only finitely many summands. So, G is finite. (cid:3) As a special case we obtain the following which can also be deduced from a result by Lau -- Ulger [17, Theorem 4.3] stating that for an [IN] locally compact group G, the Fourier algebra A(G) has the RNP if and only if G is compact. 16 MEHRDAD KALANTAR AND MATTHIAS NEUFANG Corollary 5.6. Let G be a discrete group. Then the following are equivalent: (1) the Fourier algebra A(G) has the RNP; (2) G is finite. References 1. O. Y. Aristov, Amenability and compact type for Hopf-von Neumann algebras from the homo- logical point of view, Banach algebras and their applications, Contemp. Math., Amer. Math. Soc., Providence, RI. 363 (2004), no. 1, 15 -- 37. 2. E. B´edos and L. Tuset, Amenability and co-amenability for locally compact quantum groups, Internat. J. Math. 14 (2003), 865 -- 884. 3. M. Daws, Operator biprojectivity of compact quantum groups, Proc. Amer. Math. Soc. 138 (2010), no. 4, 1349 -- 1359. 4. M. Daws and V. Runde, Reiter's properties (P1) and (P2) for locally compact quantum groups, J. Math. Anal. Appl. 364 (2010), no. 2, 352 -- 365. 5. J. Diestel and J. Uhl Jr., Vector Measures, American Mathematical Society, Providence, RI., 1977. 6. E. G. Effros and Z.-J. Ruan, Operator Spaces, London Mathematical Society Monographs, New Series, 23, The Clarendon Press, Oxford University Press, New York, 2000. 7. M. Enock and J. M Schwartz, Kac Algebras and Duality of Locally Compact Groups, Springer- Verlag, Berlin, 1992. 8. A. Ya. Helemskii, The Homology of Banach and Topological Algebras, Moscow University Press, 1986 Nauka, Moscow, 1989 (Russian); English transl.: Kluwer Academic Publishers, Dordrecht, 1989. 9. A. Ya. Helemskii, Banach and Polynormed Algebras: General Theory, Representations, Ho- mology, Nauka, Moscow, 1989 (Russian); English transl.: Oxford University Press, 1993. 10. Z. Hu, M. Neufang and Z.-J. Ruan, Multipliers on a new class of Banach algebras, locally compact quantum groups, and topological centres, Proc. London Math. Soc. 100 (2010), 429- 458. 11. M. Junge, M. Neufang and Z.-J. Ruan, A representation theorem for locally compact quantum groups. Internat. J. Math. 20 (2009), no. 3, 377 -- 400. 12. M. Kalantar, Towards Harmonic analysis on Locally compact quantum groups: From Groups to Quantum Groups and Back, Ph.D. thesis, Carleton University, Ottawa, 2010. 13. J. Kraus and Z.-J. Ruan, Multipliers of Kac algebras, Internat. J. Math. 8 (2009), no. 2, 213 -- 248. 14. J. Kustermans and S. Vaes, Locally compact quantum groups, Ann. Sci. Ecole Norm. Sup. 33 (2000), 837 -- 934. 15. J. Kustermans and S. Vaes, Locally compact quantum groups in the von Neumann algebraic setting, Math. Scand. 92 (2003), 68 -- 92. 16. B.-R. Li, Introduction to Operator Algebras, World Scientific Pub Co Inc, 1992. 17. A.T.-M. Lau and A. Ulger, Some geometric properties on the Fourier and Fourier-Stieltjes algebras of locally compact groups, Arens regularity and related problems, Trans. Amer. Math. Soc. 337 (1993), no. 1, 321 -- 359. 18. M. Neufang, Abstrakte harmonische Analyse und Modulhomomorphismen uber von Neumann- Algebren, Ph.D. thesis, Saarbrucken, 2000. 19. M. Neufang, Z.-J. Ruan and N. Spronk, Completely isometric representations of McbA(G) and UCB( G)∗, Trans. Amer. Math. Soc. 360 (2008), no. 3, 1133 -- 1161. 20. M. Neufang and V. Runde, Harmonic operators: the dual perspective, Math. Z. 255 (2007), no. 3, 669 -- 690. 21. A. Y. Pirkovskii, Biprojectivity and biflatness for convolution algebras of nuclear operators, Canad. Math. Bull. 47 (2004), no. 3, 445 -- 455. 22. G. Pisier, Similarity Problems and Completely Bounded Maps, Lecture Notes in Mathematics 1618, Springer-Verlag, Berlin, 1996. 23. G. Pisier, Non-Commutative Vector Valued Lp-Spaces and Completely p-Summing Maps, Ast´erisque, no. 247, 1998. 24. V. Runde, Characterizations of compact and discrete quantum groups through second duals, J. Operator Theory, 60 (2008), no. 2, 415 -- 428. DUALITY, COHOMOLOGY, AND GEOMETRY OF QUANTUM GROUPS 17 25. K. F. Taylor, Geometry of the Fourier algebras and locally compact groups with atomic unitary representations, Math. Ann. 262 (1983), no. 2, 183 -- 190. 26. J. Tomiyama, On the projection of norm one in W ∗-algebras, Proc. Japan Acad. 33 (1957), 608 -- 612. 27. J. Tomiyama, On the projection of norm one in W ∗-algebras. III, Tohoku Math. J. 2 (1959), 11, 125 -- 129. 28. S. Vaes, Locally compact quantum groups, Ph.D. thesis, Katholieke Universitiet Leuven, Leu- ven, 2001. 29. J. G. Wendel, Left centralizers and isomorphisms of group algebras, Pacific J. Math. (1952), 251 -- 261. 30. S. L. Woronowicz, Compact quantum groups, Sym´etries quantiques (Les Houches, 1995), 845 -- 884, North-Holland, Amsterdam, 1998. Mehrdad Kalantar School of Mathematics and Statistics, Carleton University, Ottawa, Ontario, K1S 5B6, Canada E-mail address: [email protected] Matthias Neufang School of Mathematics and Statistics, Carleton University, Ottawa, Ontario, K1S 5B6, Canada E-mail address: [email protected] Universit´e Lille 1 -- Sciences et Technologies, UFR de Math´ematiques, Laboratoire de Math´ematiques Paul Painlev´e (UMR CNRS 8524), 59655 Villeneuve d'Ascq C´edex, France E-mail address: [email protected] The Fields Institute for Research in Mathematical Sciences, Toronto, Ontario, Canada M5T 3J1 E-mail address: [email protected]
1204.5277
2
1204
2013-02-21T21:57:23
Topological convolution algebras
[ "math.FA" ]
In this paper we introduce a new family of topological convolution algebras of the form $\bigcup_{p\in\mathbb N} L_2(S,\mu_p)$, where $S$ is a Borel semi-group in a locally compact group $G$, which carries an inequality of the type $\|f*g\|_p\le A_{p,q}\|f\|_q\|g\|_p$ for $p > q+d$ where $d$ pre-assigned, and $A_{p,q}$ is a constant. We give a sufficient condition on the measures $\mu_p$ for such an inequality to hold. We study the functional calculus and the spectrum of the elements of these algebras, and present two examples, one in the setting of non commutative stochastic distributions, and the other related to Dirichlet series.
math.FA
math
TOPOLOGICAL CONVOLUTION ALGEBRAS DANIEL ALPAY AND GUY SALOMON p counterpart of Banach algebras of the formSp∈N Φ′ Abstract. In this paper we introduce a dual of reflexive Fr´echet p (where the Φ′ are (dual of) Banach spaces with associated norms k · kp), which carry inequalities of the form kabkp ≤ Ap,qkakqkbkp and kbakp ≤ Ap,qkakqkbkp for p > q + d, where d is preassigned and Ap,q is a constant. We study the functional calculus and the spectrum of the elements of these algebras. We then focus on the particular case Φ′ p = L2(S, µp), where S is a Borel semi-group in a locally compact group G, and multiplication is convolution. We give a sufficient condition on the measures µp for such inequalities to hold. Finally we present three examples, one is the algebra of germs of holomorphic functions in zero, the second related to Dirichlet series and the third in the setting of non commutative stochastic distributions. 1. Introduction Let G be a locally compact topological group with a left Haar measure µ. The convolution of two measurable functions f and g is defined by (f ∗ g)(x) =ZG f (y)g(y−1x)dµ(y). It is well known that L1(G, µ) is a Banach algebra with the convolution product, while L2(G, µ) is usually not closed under the convolution. More precisely, Rickert showed in 1968 that L2(G, µ) is closed under convolution if and only if G is compact; see [16]. In this case G is unimodular (i.e. its left and right Haar measure coincide) and it holds that for any f, g ∈ L2(G, µ). kf ∗ gk ≤pµ(G)kf kkgk, 1991 Mathematics Subject Classification. Primary: 16S99, 43A15. Secondary: 13J99, 60H40. Key words and phrases. convolution algebra, topological algebras, non commu- tative stochastic distributions. The authors thank the referee for her/his remarks and in particular for the suggestion to consider the non-unimodular case. D. Alpay thanks the Earl Katz family for endowing the chair which supported his research, and the Binational Science Foundation Grant number 2010117. 1 2 DANIEL ALPAY AND GUY SALOMON Thus, L2(G, µ) is a convolution algebra which is also a Hilbert space. In this paper we introduce and study a new type of convolution alge- bras which behave locally as Hilbert spaces, rather than being Banach spaces, even when the group G is not compact. More precisely, let (µp) be a sequence of measures on G such that µ ≫ µ1 ≫ µ2 ≫ · · · , (where µ is the left Haar measure of G) and let S ⊆ G be a Borel semi- group with the following property: There exists a non negative integer d such that for any p > q + d and any f ∈ L2(S, µq) and g ∈ L2(S, µp), the products f ∗ g and g ∗ f belong to L2(S, µp) and (1.1) kf ∗ gkp ≤ Ap,qkf kqkgkp and kg ∗ f kp ≤ Ap,qkf kqkgkp, where k · kp is the norm associate to L2(S, µp), where Ap,q is a positive constant and where the convolution of two measurable functions with respect to S is defined by (3.1). More generally, we consider dual of reflexive Fr´echet spaces of the form Sp∈N Φ′p, where the Φ′p are (dual of) Banach spaces with associated norms k · kp), which are moreover algebras and carry inequalities of the form kabkp ≤ Ap,qkakqkbkp and kbakp ≤ Ap,qkakqkbkp for p > q + d, where d is preassigned and Ap,q is a constant. We call these spaces strong algebras. They are topological algebras (see Theorem 2.3). Furthermore, the multiplication operators La : x 7→ ax and Ra : x 7→ xa are bounded from the Banach space Φ′p into itself where a ∈ Φ′q and p > q + d. Such a setting allows to consider power series. IfP∞n=0 cnzn converges in the open disk with radius R, then for any a ∈ Φ′q with kakq < , we obtain R Aq+d+1,q ∞Xn=0 ∞Xn=0 cnkankq+d+1 ≤ cn(Ap,qkf kq)n < ∞, and hence P∞n=0 cnan ∈ Φ′q+d+1. In this way we are also able to con- sider the invertible elements of the algebraSp Φ′p. When we return to the case where Φ′p = L2(S, µp), we give a suffi- cient condition on the sequence of measures (µp) such that (1.1) holds. More precisely, we show that if their Radon-Nikodym derivatives with TOPOLOGICAL CONVOLUTION ALGEBRAS 3 respect to the left Haar measure are submultiplicative, and if there ex- ists d such that for any p > q + d, the functions dµp belong to L1 of dµq both the left and right Haar measures, then (1.1) holds (see Theorem 3.4). There is one well known example for such an algebra, namely the germs of holomorphic functions at the origin, with the pointwise multiplica- tion (which is in terms of power series, a convolution). See e.g. [12]. We show that it can be identified as a union of Hardy spaces on decreas- ing sequence of disks, and that with respect to the associated norms it satisfies (1.1) (see Theorem 4.3). This allows to develop analytic calculus of germs, and to conclude easily some topological properties of this space, e.g. that it is nuclear (see Corollary 4.4). Another example for a convolution algebra associated to a semi-group and which satisfies (1.1) is the space P of all functions f : [1, ∞) → R such that there exists p ∈ N such that f (x)2 dx xp+1 < ∞, with the convolution 1 Z ∞ (f ∗ g)(x) =Z x 1 f (y)g(cid:18) x y(cid:19) dy y , ∀x ∈ [1, ∞). Inequalities (1.1) lead then to apparently new inequalities associated to Dirichlet series. Yet another example (which was the motivation for the present work) is given by the Kondratiev space of Gaussian stochastic distributions. It can be defined as Sp∈N L2(ℓ, µp) where ℓ is the free commutative semi-group generated by the natural numbers, and µp(α) = (2N)−αp = (2n)−α(n)·p, ∀α ∈ ℓ. ∞Yn=1 In this case, the convolution is sometimes called Wick product. In 1996, Vage (see [17], [5, p. 118]) showed that the Kondratiev space with the convolution product satisfies (1.1), where Ap,q = Xα∈ℓ (2N)−(p−q)α! 1 2 < ∞. 4 DANIEL ALPAY AND GUY SALOMON This fact plays a key role in stochastic partial differential equations and in stochastic linear systems theory; see [11, 3, 2]. We show here that the non-commutative version of this space still satisfies (1.1). The outline of the paper is as follows: In Section 2 we study topological and spectral properties of strong algebras and of their elements. In Section 3 we consider the case where the multiplication is a convolution. The above mentioned examples are presented in Sections 4, 5 and 6. 2. Strong algebras We now introduce a family of locally convex algebras, more precisely duals of reflexive complete countably normed spaces, which satisfy spe- cial inequalities. We recall that a countably normed space Φ is a locally convex space whose topology is defined using a countable set of com- patible norms (k · kn) i.e. norms such that if a sequence (xn) that is a Cauchy sequence in the norms k · kp and k · kq converges to zero in one of these norms, then it also converges to zero in the other. The sequence of norms (k · kn) can be always assumed to be non-decreasing. Denoting by Φp the completion of Φ with respect to the norm k · kp, we obtain a sequence of Banach spaces Φ1 ⊇ Φ2 ⊇ · · · ⊇ Φp ⊇ · · · . It is a well known result that Φ is complete if and only if Φ =T Φp, and Φ is a Banach space if and only if there exists some p0 such that Φp = Φp0 for all p > p0. Denoting by Φ′ the dual space of Φ and by Φ′p the dual of Φp it is clear that Φ′1 ⊆ Φ′2 ⊆ · · · ⊆ Φ′p ⊆ · · · , kf kp = supkxkp≤1 f (x), and these norms on Φ′ form a decreasing se- quence. For further reading on countably normed spaces and their duals we refer to [8]. and that Φ′ =S Φ′p. A functional in f ∈ Φ′p has the respective norm Definition 2.1. Let A =Sp Φ′p be a dual of a complete reflexive count- ably normed space. We call A a strong algebra if it satisfies the property that there exists a constant d such that for any q and for any p > q + d there exists a positive constant Ap,q such that for any a ∈ Φ′q and b ∈ Φ′p, kabkp ≤ Ap,qkakqkbkp and kbakp ≤ Ap,qkakqkbkp. We topologized A with the strong topology, that is, a neighborhood of zero is defined by means of any bounded set B ⊆ Tp Φp and any TOPOLOGICAL CONVOLUTION ALGEBRAS 5 number ǫ > 0, as the set of all a ∈ A for which a(b) < ǫ. sup b∈B With this topology A is locally convex. Since A is the dual of the re- flexive Fr´echet spaceT Φp, its topology coincides with its topology as the inductive limit of the Banach spaces Φ′p (see the proof of [6, IV.23, Proposition 4]). In particular, it satisfies the universal property of in- ductive limits, i.e. any linear map from A to another locally convex space is continuous if and only if the restriction of the map to the any of the spaces Φ′p is continuous (see [6, II.29]). We now show that the multiplication is continuous in A. Before that, we show it is separately continuous. Proposition 2.2. Let a ∈ A. Then the linear mappings La : x 7→ ax, Ra : x 7→ xa are continuous. r : Φ′r → A be the restriction Proof. Suppose that a ∈ Φ′q, and let LaΦ′ of the map La to Φ′r. If B is a bounded set of Φ′r then in particular we may choose p ≥ q + d such that p ≥ r, so B ⊆ {x ∈ Φ′p : kxkp < λ}. Thus, for any x ∈ B kLaΦ′ p(x)kq ≤ Ap,qλkxkp. p(B) is bounded in Φ′p and hence in A. Thus, for any r, Hence, LaΦ′ LaΦ′ r : Φ′r → A is bounded and hence continuous. Since A =Sp∈N Φ′p is a strong dual of the reflexive Fr´echet spaceT Φp, it is the inductive limit of the Hilbert spaces Φ′p. So by the universal property of inductive limits, La is continuous. The proof for Ra is similar. (cid:3) Theorem 2.3. Let A be a strong algebra. Then the multiplication is a continuous function A × A → A in the strong topology. Hence (A, +, ·) is a topological C-algebra. This follows immediately from Proposition 2.2 together with the fol- lowing theorem, proved in [6, IV.26]. Theorem 2.4. Let E1 and E2 be two reflexive Fr´echet spaces, and let G a locally convex Hausdorff space. For i = 1, 2, let Fi be the strong dual of Ei. Then every separately continuous bilinear mapping u : F1 × F2 → G is continuous. Henceforward, we assume that A is a unital strong algebra. The fol- lowing theorems shows that, as in the Banach algebra case, one can develop an analytic calculus in strong algebras. 6 DANIEL ALPAY AND GUY SALOMON Theorem 2.5. Assuming P∞n=0 cnzn converges in the open disk with radius R, then for any a ∈ A such that there exist p, q with p > q + d and Ap,qkakq < R it holds that cnan ∈ Φ′p ⊆ A. ∞Xn=0 Proof. This follows from cnkankp ≤ ∞Xn=0 ∞Xn=0 cn(Ap,qkakq)n < ∞. (cid:3) Theorem 2.6. Let a ∈ A be such that there exist p, q such that p > q+d and Ap,qkakq < 1 then 1 − a is invertible (from both sides) and it holds that k(1 − a)−1kp ≤ 1 1 − Ap,qkakq , k1 − (1 − a)−1kp ≤ Ap,qkakq 1 − Ap,qkakq . Proof. Due to Theorem 2.5 we have that an! (1 − a) = 1, Moreover, clearly and we have that an ∈ Φ′p ⊆ A. ∞Xn=0 an! = ∞Xn=0 (1 − a) ∞Xn=0 ∞Xn=0 ∞Xn=0 ∞Xn=1 ∞Xn=1 kankp ≤ k(1 − a)−1kp ≤ kankp ≤ (Ap,qkakq)n = 1 1 − Ap,qkakq , and k1 − (1 − a)−1kp ≤ (Ap,qkakq)n = Ap,qkakq 1 − Ap,qkakq . (cid:3) As was mentioned before, since A = Sp Φ′p is the strong dual of a reflexive Fr´echet space, it is also the inductive limit of the Banach spaces Φ′p, i.e. its strong topology coincides is the finest locally-convex topology such that the embeddings Φ′p ֒→ A are continuous. The following theorem refers to the case where the strong topology of A is the finest topology (rather than the finest locally-convex topology) such that these mappings are continuous. There are two important TOPOLOGICAL CONVOLUTION ALGEBRAS 7 cases when this happens: the first is when A is a Banach algebra, and the second is when the embeddings Φ′p ֒→ Φ′p+1 are compact. In particular, the examples described in Sections 4 and 6 pertain to the second case (see Theorem 3.7). Theorem 2.7. If the topology of A is the finest topology such that the embeddings Φ′p ֒→ A are continuous, then the set of invertible elements GL(A) is open, and the function a 7→ a−1 is continuous. Proof. Note that in this case, U is open in A if and only if U ∩ Φ′p is open in Φ′p for every p. Let U be an open set of A, and let p ∈ N. Let Ua be the set of all b ∈ A such that there exists p > q + d for which kbkp < 1 Ap+d,pAp,qka−1kq . Clearly Ua ∩ Φ′p is open in Φ′p for any p, so Ua is open. Moreover, for any b ∈ Ua Ap+d,pka−1bkp ≤ Ap,qka−1kqkbkp < 1. In view of Theorem 2.6, 1 − a−1b is invertible, and therefore a − b = a(1 − a−1b) is invertible too. Thus, a + Ua ⊆ GL(A), and so GL(A) is open. Now, we note that, (a + b)−1 − a−1 =(cid:0)a(1 + a−1b)−1(cid:1) − a−1 =(cid:0)(1 + a−1b)−1 − 1(cid:1) a−1. = (1 + a−1b)−1a−1 − a−1 Therefore, for any b ∈ Ua, k(a + b)−1 − a−1kp+d ≤ Ap+d,qka−1kqk(1 + a−1b)−1 − 1kp+d ≤ Ap+d,qka−1kq Ap+d,pka−1bkp 1 − Ap+d,pka−1bkp ≤ Ap+d,qka−1kq Ap+d,pAp,qka−1kqkbkp 1 − Ap+d,pAp,qka−1kqkbkp Thus, the function u : b 7→ (a + b)−1 − a−1 p is continuous with respect to satisfy u(Ua ∩ Φ′p) ⊆ Φ′p+d, and uUa∩Φ′ the topologies of Ua ∩ Φ′p in the domain and Φ′p+d in the range. Now, let V be an open set of A. Then, u−1(V )∩Φ′p = u−1 (V ∩Φ′p+d) is open in Ua ∩ Φ′p. In particular, u−1(V ) ∩ Φ′p is open in Φ′p. Since p was arbitrary, u−1(V ) is open in A, so u is continuous. Since a was arbitrary, a 7→ a−1 is continuous. (cid:3) Ua∩Φ′ p 8 DANIEL ALPAY AND GUY SALOMON Definition 2.8. The spectrum of an element a ∈ A is defined by σ(a) = {λ ∈ C : a − λ is not invertible}. In [14, p. 167] Naimark defines topological algebras as algebras which are locally convex vector spaces, and for which the product is separately continuous in each variable. Since strong algebras are locally convex as strong dual complete of complete countably normed spaces, we can apply to strong algebras which satisfy the assumption of Theorem 2.7, the results of Naimark [14, §3, p. 169] on topological algebras with continuous inverse. In particular he showed that Theorem 2.9 (Mark Naimark). The spectrum of any element in a locally convex unital algebra with continuous inverse is non-empty and closed. In the next theorem, we get a bound on the spectrum, which is spe- cific to strong algebras, even if they do not satisfy the assumption of Theorem 2.7. Theorem 2.10. For any a ∈ A, σ(a) ⊆ {z ∈ C : z ≤ Proof. For every 0 6= λ ∈ σ(a), 1 − a q and for any p > q + d inf Ap,qkakq}. {(p,q):p>q+d} λ is not invertible and thus for any Thus, λ ≤ ≥ 1. Ap,q(cid:13)(cid:13)(cid:13) a λ(cid:13)(cid:13)(cid:13)q inf {(p,q):p>q+d} Ap,qkakq. (cid:3) 3. The convolution algebra Sp L2(S, µp) Definition 3.1. Let G be a locally compact topological group with a Haar measure µ and let S ⊆ G be a Borel semi-group. The convolution of two measurable functions f, g with respect to S is defined by (3.1) f (y)g(y−1x)dµ(y) (f ∗ g)(x) =ZS∩xS−1 for any x ∈ S such that the integral converges. Definition 3.2. Let G be a locally compact topological group with a left Haar measure µ and let S ⊆ G be a Borel semi-group. Let (µp) be a sequence of measures on G such that µ ≫ µ1 ≫ µ2 ≫ · · · . Sp∈N L2(S, µp) is called a strong convolution algebra if there exists a TOPOLOGICAL CONVOLUTION ALGEBRAS 9 non negative integer d such that for every p > q + d there exists a pos- itive constant Ap,q such that for every f ∈ L2(S, µq) and g ∈ L2(S, µp), kf ∗ gkp ≤ Ap,qkf kqkgkp and kg ∗ f kp ≤ Ap,qkf kqkgkp, where k · kp is the norm associate to L2(S, µp). Remark 3.3. One can easily verify that requiring the measures (µp) ofSp L2(S, µp) to obey assures thatSp L2(S, µp) is a strong algebra. dµ2 dµ dµ1 dµ 1 ≥ ≥ ≥ · · · > 0 µ-a.e. The following theorem allows to define a wide family of strong convo- lution algebras. For a converse theorem in the where G is discrete see Theorem 3.6. Theorem 3.4. Assume that for every x, y ∈ S and for every p ∈ N dµp dµ (xy) ≤ dµp dµ (x) dµp dµ (y), Then, for every choice of f ∈ L2(S, µq) and g ∈ L2(S, µp), (3.2) kf ∗ gkp ≤(cid:18)ZS kg ∗ f kp ≤(cid:18)ZS dµp dµq and dµp dµq 2 2 kf kqkgkp kf kqkgkp, dµ(cid:19) 1 deµ(cid:19) 1 ZS deµ < ∞ p,q = max(cid:16)RS dµp dµq dµp dµq dµp dµq ZS dµ < ∞ and negative integer d such that where eµ is the right Haar measure. In particular, if there exists a non for every positive integers p, q such that p > q + d, thenSp L2(S, µp) is deµ(cid:17)). Before we prove this theorem, we need the following lemma, which is the core of the theorem. a strong convolution algebra (with A2 dµ,RS dµp dµq Lemma 3.5. Let ν, λ be two Borel measures on G such that λ ≪ ν ≪ µ. If for any x, y ∈ S dλ dµ (y) µ-a.e., (xy) ≤ dλ dµ dλ dµ (x) 10 DANIEL ALPAY AND GUY SALOMON then for any f ∈ L2(S, ν) and g ∈ L2(S, λ) 2 (3.3) dλ dν kf ∗ gkL2(S,λ) ≤(cid:18)ZS kg ∗ f kL2(S,λ) ≤(cid:18)ZS dµ(cid:19) 1 deµ(cid:19) 1 where eµ is the right Haar measure. Proof. We denote dλ dν and 2 kf kL2(S,ν)kgkL2(S,λ) kf kL2(S,ν)kgkL2(S,λ). gλ,µ(x) = g(x)s dλ dµ (x). Then, kf ∗ gk2 and dµ (x) fλ,µ(x) = f (x)s dλ L2(S,λ) =ZS(cid:12)(cid:12)(cid:12)(cid:12)ZS∩xS−1 ≤ZS(cid:18)ZS∩xS−1 =ZS ZS∩xS−1 ≤ZS ZS∩xS−1 =ZS(cid:18)ZS∩xS−1 =ZSZS∩xS−1ZS∩xS−1 =ZSZS ≤ZSZS fλ,µ(y)dµ(y)(cid:19)2 =(cid:18)ZS fλ,νdµ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZSr dλ dν 2 kgk2 L2(S,λ) kgk2 L2(S,λ) 2 dλ(x) dλ(x) f (y)g(y−1x)dµ(y)(cid:12)(cid:12)(cid:12)(cid:12) f (y)g(y−1x)dµ(y)(cid:19)2 f (y)g(y−1x)s dλ (x)dµ(y)!2 f (y)sdλ (y)g(y−1x)s dλ fλ,µ(y)gλ,µ(y−1x)dµ(y)(cid:19)2 dµ dµ dµ dµ(x) dµ(x) (y−1x)dµ(y)!2 dµ(x) fλ,µ(y)fλ,µ(y)gλ,µ(y−1x)gλ,µ(y−1x)dµ(y)dµ(y)dµ(x) fλ,µ(y)fλ,µ(y)(cid:18)ZS∩yS∩yS gλ,µ(y−1x)gλ,µ(y−1x)dµ(x)(cid:19) dµ(y)dµ(y) fλ,µ(y)fλ,µ(y) kgλ,µk2 L2(S,µ) dµ(y)dµ(y) TOPOLOGICAL CONVOLUTION ALGEBRAS 11 L2(S,λ) ≤(cid:18)ZS =(cid:18)ZS (g ∗ f )(x) =ZS∩xS−1 dλ dν dλ dν L2(S,ν) kgk2 L2(S,µ) kgk2 dµ(cid:19) kfν,µ(y)k2 dµ(cid:19) kf k2 g(y)f (y−1x)dµ(y) =ZS∩S−1x L2(S,λ) yields the first inequality of (3.3). As for the second inequality, note that f (y)g(xy−1)deµ(y). Thus, replacing the terms µ, S ∩ xS−1 and y−1x in the proof of the first inequality, with the terms µ, S ∩ S−1x and xy−1 respectively, yields a proof for the second inequality. . (cid:3) We are now ready to prove Theorem 3.4. Proof of Theorem 3.4. In view of Lemma 3.5, it holds that kf ∗ gkp ≤(cid:18)ZS kg ∗ f kp ≤(cid:18)ZS dµp dµq and dµp dµq 2 dµ(cid:19) 1 deµ(cid:19) 1 2 kf kqkgkp kf kqkgkp, for any p > q + d. This yields the requested result. (cid:3) In the following theorems we focus on the case where G is a discrete group. In this case, the Haar measure is simply the counting measure. Nonetheless, we keep using the notations of the general case, namely integrals instead of sums, and L2 instead of ℓ2. Theorem 3.6. LetSp L2(S, µp) is a strong convolution algebra, where S is a semi-group in a discrete group G, such that dµp dµ dµ(cid:19)p =(cid:18) dµ1 for every p ∈ N, and such that there exists d for whichRS(cid:16) dµ1 dµ(cid:17)d x, y ∈ S, dµp dµ dµp dµ dµp dµ (xy) ≤ (x) (y) for every p ∈ N. dµ < ∞. Then for any Proof. Denoting by δx the characteristic function of {x}, we have kδx ∗ δykp ≤ ZS(cid:18) dµ1 dµ(cid:19)p−q 2 dµ! 1 kδxkqkδykp. 12 DANIEL ALPAY AND GUY SALOMON Setting q = p − (d + 1) we obtain dµp dµ (xy) ≤ ZS(cid:18) dµ1 (xy) ≤ ZS(cid:18) dµ1 dµ(cid:19)d+1 dµ! 1 dµ(cid:19)d+1 dµ! 1 2p(cid:18) dµ1 dµ or dµ1 dµ 2 dµp−(d+1) dµ (x) dµp dµ (y), (x)(cid:19) p−(d+1) p (cid:18) dµ1 dµ (y)(cid:19) . (cid:3) Taking p → ∞ yields the requested result. dµp dµq embedding dµ < ∞, then the Tq,p : L2(S, µq) ֒→ L2(S, µq) is Hilbert-Schmidt. As a result, if for any p > q + d the above integral Theorem 3.7. If G is discrete and for p > q,RS converges (as in the assumption of Theorem 3.4), thenSp L2(S, µp) is dµ (x)(cid:17)− 1 dµ (x). Thus, e(q) Proof. For any x ∈ S, kδxk2 (x ∈ S) form an orthonormal basis of L2(S, µq), and x = (cid:16) dµq q = dµq nuclear. 2 δx kTq,pk2 kTq,pe(q) x k2 HS =Xx∈S p =ZS dµp dµq dµ, where k · kHS denotes the Hilbert Schmidt norm of the embedding Tq,p. (cid:3) A first example of a strong convolution algebra was given in our previ- ous paper [4]. We briefly discuss it now. Recall first that the Schwartz space of complex tempered distributions can be viewed as the union of the Hilbert spaces s′p :=(a ∈ CN0 : (n + 1)−2pa(n)2 < ∞) . ∞Xn=0 Let G = Z with its discrete topology, µ the counting measure (which is also the Haar measure of G), S = N0 and setting (3.4) dµp(n) = (n + 1)−2p, we conclude that s′p = L2(S, µp), and s′ = Sp∈N s′p = Sp∈N L2(S, µp). The convolution then becomes the standard one sided convolution of sequences. TOPOLOGICAL CONVOLUTION ALGEBRAS 13 One may ask whether s′ is a strong convolution algebra, i.e. if there exists a constant d such that for any p > q + d there exists a positive constant Ap,q such that for any a ∈ s′q and b ∈ s′p, ka ∗ bkp ≤ Ap,qkakqkbkp and kb ∗ akp ≤ Ap,qkakqkbkp. Since, and dµp dµ dµ = ZS(cid:18) dµ1 dµ(cid:19)d ∞Xn=0 dµ(cid:19)p (n) = (n + 1)−2p =(cid:18) dµ1 (n + 1)−2 < ∞ for any p ∈ N, we conclude in view of the last theorem that the answer is negative, that is, s′ =Sp∈N s′p is not a strong convolution algebra. In [4] we re- place the measures (n + 1)−2p by 2−np, and obtain a strong convolution algebra that contains s′, and which can be identified as the dual of a space of entire holomorphic functions that is included in the Schwartz space of complex-valued rapidly decreasing smooth functions. Other examples of strong convolution algebras, which can be con- structed in the manner described in the last theorem, are given in Sections 5 and 6 respectively. To end this section, we make a brief discussion on the case where G is non-unimodular (i.e. its left and right Haar measure do not co- incide). As is clear from the statement of Theorem 3.4 and from the proof of Lemma 3.5, if G is a non-unimodular group, since the integrals RS dµp dµq dµ,RS dµp dµq equalities (3.2) is stricter than the other. Nonetheless, if both integrals are finite, then we may always take A2 and obtain deµ need not be equal, it may happen that one of the in- deµ(cid:17), p,q = max(cid:16)RS dµ,RS dµp dµq dµp dµq kf ∗ gkp ≤ Ap,qkf kqkgkp and kg ∗ f kp ≤ Ap,qkf kqkgkp. As an example, consider the non-unimodular group G = {(a, b) ∈ R2 : a > 0} with the multiplication (a, b) · (c, d) = (ac, ad + b) which can be identified as the subgroup of GL2(R) (cid:26)(cid:18)a b 0 1(cid:19) ∈ GL2(R) : a > 0(cid:27) . This is the so-called ax + b group. One can easily verify that its left Haar measure is given by dµ(a, b) = a−1dadb, and that its right Haar 14 DANIEL ALPAY AND GUY SALOMON measure is given by deµ(a, b) = a−2dadb. Let S be the semigroup and set dµp(a, b) = a−(p+2)e−bpdadb. SoSp L2(S, µp) is the space of all measurable functions S → C satisfy {(a, b) ∈ R2 : a ≥ 1, b ≥ 0} f (a, b)2a−(p+2)e−bpdadb < ∞ for some p ∈ N, with a convolution product 1 Z ∞ 0 Z ∞ (f ∗ g)(a, b) =Z b 0 Z a 1 Since clearly dµp p > q, f (x, y)g(cid:0)x−1a, x−1(b − y)(cid:1) a−1dadb. dµ ((a, b)(a′, b′)) ≤ dµp dµ (a, b) dµp dµ (a′, b′), and since for any 1 and dµp dµq dµ = (2(p − q) + 1)(p − q) ZS By theorem 3.4, Sp L2(S, µp) is a strong convolution algebra with stricter. More precisely, denoting Bp,q = (cid:16) deµ = (2(p−q)+1)(p−q)(cid:17) 1 f ∈ L2(S, µq) and g ∈ L2(S, µp) (p > q) we have, in view of (3.2), on of the inequalities is 2 , for every 1√2(p−q) . However, 2(p − q)2 ZS dµp dµq 1 , Ap,q = 1 kf ∗ gkp ≤ Bp,qkf kqkgkp and kg ∗ f kp ≤ Ap,qkf kqkgkp. A question which we leave open is whether there exists a "one-sided strong convolution algebra" which is not a ("two-sided") strong convo- lution algebra, i.e. an algebra of the formSp∈N L2(S, µp) of which there are d and Ap,q such that for every p > q + d and for every f ∈ L2(S, µq) and g ∈ L2(S, µp), kf ∗ gkp ≤ Ap,qkf kqkgkp, but where the "reflected" inequality, namely kg ∗ f kp ≤ Ap,qkf kqkgkp, does not hold for any choice of d and Ap,q. Clearly, such an example should be over a non- unimodular group. 4. The space of germs of holomorphic functions in zero Let H(C) be the space of entire holomorphic functions. It can be topologized as follows. For any p ∈ N we define the n-norm on H(C) by kf k2 p = f (z)2 dz, 1 2πZz=2p then the topology of H(C) is the associated Fr´echet topology. TOPOLOGICAL CONVOLUTION ALGEBRAS 15 Proposition 4.1. The topology defined above coincides with the usual topology of H(C), that is the topology of uniform convergence on com- pact sets. Proof. Clearly, kf k2 p = f (z)2 dz ≤ 2p sup z∈2p ¯D f (z)2. On the opposite direction, using Cauchy theorem, for any z ∈ 2p ¯D 1 2πZz=2p f (z)2 =(cid:12)(cid:12)(cid:12)(cid:12) ≤ 1 1 2πZz=2p+1 2πZz=2p+1(cid:12)(cid:12)(cid:12)(cid:12) 2πZz=2p+1 f (ω)2 ω − z f (ω)2 dω(cid:12)(cid:12)(cid:12)(cid:12) ω − z(cid:12)(cid:12)(cid:12)(cid:12) dω ≤ 2−p 1 = 2−pkf k2 p+1 f (ω)2 dω Thus, and we conclude that f (z) ≤ 2−p/2kf kp+1, sup z∈2p ¯D kf kp ≤ 2p/2 sup z∈2p ¯D f (z) ≤ kf kp+1. (cid:3) Let H2(2pD) be the Hardy space of the disk 2pD, i.e. H2(2pD) =(cid:26)f is holomorphic on 2pD : sup 0<r<1 1 2πZz=r·2p f (z)2 dz < ∞(cid:27) , then clearly, H2(2pD) is the completion of H(C) with respect to the p-norm, and H2(D) ⊇ H2(2D) ⊇ · · · ⊇ H2(2pD) ⊇ · · · . Therefore, we obtain H(C) = \p∈N H2(2pD), i.e. H(C) is a countably Hilbert space. 16 DANIEL ALPAY AND GUY SALOMON The Hardy space can be viewed also in terms of power series, and with anzn : ∞Xn=0 an222np < ∞) , ∞Xn=0 an222np. 1 f (z)2 dz = and f (z) =P∞n=0 anzn , it holds that H2(2pD) =( ∞Xn=0 2πZz=r·2p ∞Xn=0 H(C) =( ∞Xn=0 Hence we obtain, sup 0<r<1 anzn : with the associated topology. an222np < ∞ for all p ∈ N) Now, let H′(C) denote the dual space of H(C), equipped with the strong topology as dual of countably Hilbert space. Then, H′(C) = [p∈N H2(2−pD), where it holds that H2(D) ⊆ H2(2−1D) ⊆ · · · ⊆ H2(2−pD) ⊆ · · · . Therefore, we conclude Proposition 4.2. H′(C) can be identified with the space of holomor- phic germs at zero O0. The multiplication of two elements in H′(C), f (z) = P∞n=0 anzn and g(z) =P∞n=0 bnzn is defined by (f g)(z) = ∞Xn=0 nXm=0 ambn−m! zn. Theorem 4.3. Let f ∈ H2(2−qD) and g ∈ H2(2−pD) where p > q + d. Then f g ∈ H2(2−pD). Furthermore, kf gk−p ≤ (1 − 4−(p−q))−1kf k−qkgk−p. Proof. This follows directly from Theorem 3.4. (cid:3) In view of Theorems 2.3, 2.7, 2.10 and 3.7, we conclude: Corollary 4.4. The following properties hold: (a) The multiplication is continuous, i.e. (H′(C), +, ·) is a topological algebra. TOPOLOGICAL CONVOLUTION ALGEBRAS 17 (b) GL(H′(C)) is open and the function f 7→ f−1 is continuous. (c) f ∈ H′(C) is invertible if and only if f (0) 6= 0. (d) H′(C) is nuclear. Note that the last item is well known; see for instance [15, pp. 105-106] (since the dual of nuclear space is nuclear) or [9, 10]. Remark 4.5. Let (ξn) denote the Hermite functions, which form an orthonomal basis of L2(C), and consider the isometrically isomorphism G : L2(R) → H2(D) defined by G : ξn 7→ zn. In [4] we showed that the space G of all entire functions f (z) such that ZZC f (z)2 e 1−2−2p 1+2−2p x2− 1+2−2p 1−2−2p y2 dxdy < ∞ for all p ∈ N. can be viewed as the space of all functionsP∞n=0 anξn such that for any p ∈ N an222np < ∞. ∞Xn=0 Thus, the image of G under G is H′(C). We also note that G is included in the Schwartz space of rapidly decreasing smooth complex functions S . Thus, the Schwartz space of tempered distributions S ′ is included in G ′ which can be identified with the space of gems of holomorphic functions in zero. 5. The space P We now present a new example of a convolution algebraSp L2(S, µp). Let P be the space of all functions f : [1, ∞) → R such that there exists p ∈ N with Z ∞ 1 f (x)2 dx xp+1 < ∞. In particular, any restriction of a polynomial function into [1, ∞) be- longs to P. Thinking of [1, ∞) as a group with respect to the multipli- cation of the real numbers, and since the Haar measure of ((0, ∞), ·) is dx x = d(ln(x)), we obtain the following convolution We also note that for any p > q, 1 (f ∗ g)(x) =Z x Z ∞ 1 f (y)g(cid:18) x y(cid:19) dy y , ∀x ∈ [1, ∞). x−(p+1) x−(q+1) dx x = 1 p − q . 18 DANIEL ALPAY AND GUY SALOMON Thus, using Theorem 3.4, we obtain that for any p > q, (5.1) 1 (cid:18)Z x Z ∞ 1 f (y)g(cid:18) x y(cid:19) dy y (cid:19)2 dx xp+1 ≤ 1 p − qZ ∞ 1 f 2(x) dx xq+1Z ∞ 1 g2(x) dx xp+1 . We now present inequalities on Dirichlet series. To that purpose we note that in (5.1) we can assume that p and q are positive real numbers. Definition 5.1. The one-sided Mellin transform is defined by (Mf )(s) =Z ∞ 1 xsf (x) dx x . Thus, (5.1) can be rewritten as (5.2) (M(f ∗ g)2)(−t) ≤ 1 t − r (Mf 2)(−r)(Mg2)(−t) for any t > r. Example 5.2. Consider the Dirichlet series Since for any s in the half-plane of absolute convergence, ∞Xn=1 ann−s = sZ ∞ 1 anXn≤ x y bn ∞Xn=1 1 Z x Z ∞ 1 sXn≤y ann−s and an dy y dx xt+1 ≤ y−sXn≤y y  dy 2 bnn−s. ∞Xn=1 = s MXn≤y ∞Xn=1 t(t − r)r 1 (see [13, Theorem 1.3, p. 13], [7, p. 10]) (5.2) yields an! (−s), ann−r ∞Xn=1 bnn−t. For example taking the zeta function of Riemann, ζ(s) = n−s, ∞Xn=1 which converges for Re s > 1, one obtains 1 s ⌊y⌋(cid:22)x y !2 1 Z x y(cid:23) dy Z ∞ 1 s ⌊y⌋(cid:22)x y !2 1 Z x y(cid:23) dy Z ∞ Hence, dx xt+1 ≤ ζ(t)ζ(r) t(t − r)r for any 1 < r < t. dx xt+1 ≤ inf t(t − r)r(cid:19) r∈(1,t)(cid:18) ζ(t)ζ(r) for any t > 1. TOPOLOGICAL CONVOLUTION ALGEBRAS 19 In a similar way, if we take ϕ(n) to be the phi Euler function (that is ϕ(n) is the number of positive integers less than or equal to n and relatively prime with n), we obtain 1 Z x Z ∞ 1 sXn≤y ϕ(n)Xn≤ x y ϕ(n) dy y  2 dx xt+1 ≤ inf tζ(t)(t − r)rζ(r)(cid:19) . r∈(1,t)(cid:18) ζ(t − 1)ζ(r − 1) 6. The Kondratiev space of stochastic distributions We now present two other examples of strong convolution algebras. First we consider the Kondratiev space S−1 of stochastic distributions. The space S−1 plays an important role in white noise space analysis and in the theory of linear stochastic differential equations and linear stochastic systems; see [11, 3, 2, 4, 1]. Next we introduce its non com- mutative counterpart S−1. Further properties of S−1 and applications will be presented in a forthcoming publication. ℓ = N(N) (6.1) and let S−1 be the space of all functions f : ℓ → C such that there 0 : supp(α) is finite(cid:9) = ⊕n∈NN0en, Let 0 =(cid:8)α ∈ NN exists p ∈ N with Xα∈ℓ Denoting f (α)2(2N)−αp < ∞. µp(α) = (2N)−αp = (2n)−α(n)p ∞Yn=1 we conclude that S−1 is the convolution algebraSp L2(S, µp), with the convolution f (β)g(α − β), ∀α ∈ ℓ. (f ∗ g)(α) =Xβ≤α In stochastic analysis, this convolution is called the Wick product; see [5]. We also note that for every p > q + 1, Xα∈ℓ (2N)−αp (2N)−αq =Xα∈ℓ (2N)−α(p−q) = A(p − q)2 < ∞. The fact that the last term is finite is due to [18]. Another proof can be found in [4]. Thus, due to Theorem 3.4, we obtain that for every p > q + 1, (6.2) kf ∗ gkp ≤ A(p − q)kf kqkgkp. 20 DANIEL ALPAY AND GUY SALOMON This result (in the context of Kondratiev space of stochastic distribu- tions) was first proved by Vage in 1996 (see [17]). We now introduce a construction of the non-commutative version of S−1 as a convolution algebra (in a forthcoming paper we show that this space can also be obtained via a union of full Fock spaces). We replace the free commutative semi-group generated by N, namely ℓ, with the free (non-commutative) semi-group generated by N, which we will denote by ℓ. For α = zα1 in ∈ ℓ (where i1 6= i2 6= · · · 6= in) i1 we define zα2 i2 · · · zαn (2N)α = (2j)(Pk∈{m : im=j} αk). Proposition 6.1. For every p, q integers such that p > q + 1, it holds that (2N)−α(p−q) < ∞. nYk=1 (2ik)αk = Yj∈{i1,...,in} B(p − q)2 =Xα∈ℓ Proof. We note that B(p − q)2 =Xα∈ℓ (2N)−α(p−q) = = ∞Xn=0 Xα∈ℓ,α=n ∞Xn=0 Xα∈ℓ,α=n ∞Yk=1 ∞Yk=1 n! α! (2k)−αk(p−q) (2k)−αk(p−q). Considering now an experiment with N results, where the probability of the result k is pk = c · (2k)−(p−q) (c is chosen such that P pk = 1, and clearly for any p > q + 1, c > 1), the probability that repeating the experiment n times yields that the result k occurs αk times for any k is Thus, k = cn n! pαk α! ∞Yk=1 n! α! ∞Yk=1 Xα∈ℓ,α=n n! α! ∞Yk=1 (2k)−αk(p−q). (2k)−αk(p−q) = c−n, and we obtain the requested result. (cid:3) We conclude that the non-commutative version of the Kondratiev space is also a strong convolution algebra. TOPOLOGICAL CONVOLUTION ALGEBRAS 21 References [1] D. Alpay and H. Attia. An interpolation problem for functions with values in a commutative ring. In A Panorama of Modern Operator Theory and Related Topics, volume 218 of Operator Theory: Advances and Applications, pages 1 -- 17. Birkhauser, 2012. [2] D. Alpay and D. Levanony. Linear stochastic systems: a white noise approach. Acta Appl. Math., 110(2):545 -- 572, 2010. [3] D. Alpay, D. Levanony, and A. Pinhas. Linear stochastic state space theory in the white noise space setting. SIAM Journal of Control and Optimization, 48:5009 -- 5027, 2010. [4] D. Alpay and G. Salomon. New topological C-algebras with applications in linear system theory. Infin. Dimens. Anal. Quantum Probab. Relat. Top., 2012. [5] F. Biagini, Y. Hu, B. Øksendal, and T. Zhang. Stochastic calculus for frac- tional Brownian motion and applications. Probability and its Applications (New York). Springer-Verlag London Ltd., London, 2008. [6] N. Bourbaki. Topological vector spaces. Chapters 1 -- 5. Springer-Verlag, Berlin, 1987. [7] P. D. T. A. Elliott. Duality in analytic number theory, volume 122 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 1997. [8] I.M. Gelfand and G.E. Shilov. Generalized functions. Volume 2. Academic Press, 1968. [9] A. Grothendieck. Sur certains espaces de fonctions holomorphes. I, J. Reine Angew. Math (1953), vol. 192, pp. 35-64 [10] A. Grothendieck. Sur certains espaces de fonctions holomorphes. II, J. Reine Angew. Math (1953), vol. 192, pp. 78-95 [11] H. Holden, B. Øksendal, J. Ubøe, and T. Zhang. Stochastic partial differential equations. Probability and its Applications. Birkhauser Boston Inc., Boston, MA, 1996. [12] D. H. Luecking and L. A. Rubel. Complex analysis. Universitext. Springer- Verlag, New York, 1984. A functional analysis approach. [13] H. L. Montgomery and R. C. Vaughan. Multiplicative number theory. I. Clas- sical theory, volume 97 of Cambridge Studies in Advanced Mathematics. Cam- bridge University Press, Cambridge, 2007. [14] M. A. Naımark. Normed algebras. Wolters-Noordhoff Publishing, Groningen, third edition, 1972. Translated from the second Russian edition by Leo F. Boron, Wolters-Noordhoff Series of Monographs and Textbooks on Pure and Applied Mathematics. [15] A. Pietsch. Nuclear locally convex spaces. Ergebnisse der Mathematik und ihrer Grenzgebiet, Band 66. Springer-Verlag, 1972. [16] N. W. Rickert. Convolution of L2-functions. Colloq. Math., 19:301 -- 303, 1968. [17] G. Vage. Hilbert space methods applied to stochastic partial differential equa- tions. In H. Korezlioglu, B. Øksendal, and A.S. Ustunel, editors, Stochastic analysis and related topics, pages 281 -- 294. Birkauser, Boston, 1996. [18] T Zhang. Characterizations of white noise test functions and Hida distribu- tions. Stochastics, 41:71 -- 87, 1992. 22 DANIEL ALPAY AND GUY SALOMON (DA) Department of Mathematics Ben Gurion University of the Negev P.O.B. 653, Be'er Sheva 84105, ISRAEL E-mail address: [email protected] (GS) Department of Mathematics Ben Gurion University of the Negev P.O.B. 653, Be'er Sheva 84105, ISRAEL E-mail address: [email protected]
1705.07039
1
1705
2017-05-19T15:07:45
Geometric aspects of $p$-angular and skew $p$-angular distances
[ "math.FA" ]
Corresponding to the concept of $p$-angular distance $\alpha_p[x,y]:=\left\lVert\lVert x\rVert^{p-1}x-\lVert y\rVert^{p-1}y\right\rVert$, we first introduce the notion of skew $p$-angular distance $\beta_p[x,y]:=\left\lVert \lVert y\rVert^{p-1}x-\lVert x\rVert^{p-1}y\right\rVert$ for non-zero elements of $x, y$ in a real normed linear space and study some of significant geometric properties of the $p$-angular and the skew $p$-angular distances. We then give some results comparing two different $p$-angular distances with each other. Finally, we present some characterizations of inner product spaces related to the $p$-angular and the skew $p$-angular distances. In particular, we show that if $p>1$ is a real number, then a real normed space $\mathcal{X}$ is an inner product space, if and only if for any $x,y\in \mathcal{X}\smallsetminus{\lbrace 0\rbrace}$, it holds that $\alpha_p[x,y]\geq\beta_p[x,y]$.
math.FA
math
GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES J. ROOIN1, S. HABIBZADEH2 AND M.S. MOSLEHIAN3 Abstract. Corresponding to the concept of p-angular distance αp[x, y] := (cid:13)(cid:13)kxkp−1x − kykp−1y(cid:13)(cid:13), we first introduce the notion of skew p-angular distance βp[x, y] :=(cid:13)(cid:13)kykp−1x − kxkp−1y(cid:13)(cid:13) for non-zero elements of x, y in a real normed linear space and study some of significant geometric properties of the p-angular and the skew p-angular distances. We then give some results comparing two different p-angular distances with each other. Finally, we present some char- acterizations of inner product spaces related to the p-angular and the skew p-angular distances. In particular, we show that if p > 1 is a real number, then a real normed space X is an inner product space, if and only if for any x, y ∈ X r {0}, it holds that αp[x, y] ≥ βp[x, y]. 1. Introduction Throughout this paper, let X denotes an arbitrary non-zero normed linear space over the field of real numbers. Clarkson [3] introduced the concept of angular distance between non-zero ele- ments x and y in X by In [16], Maligranda considered the p-angular distance y . α[x, y] =(cid:13)(cid:13)(cid:13)(cid:13) kxk1−p − x x kyk(cid:13)(cid:13)(cid:13)(cid:13) kxk − kyk1−p(cid:13)(cid:13)(cid:13)(cid:13) y αp[x, y] =(cid:13)(cid:13)(cid:13)(cid:13) (p ∈ R) between non-zero vectors x and y in X as a generalization of the concept of angular distance. Corresponding to the notion of p-angular distance, we define the concept of skew p-angular distance between non-zero vectors x and y in X , as βp[x, y] =(cid:13)(cid:13)(cid:13)(cid:13) x kyk1−p − y kxk1−p(cid:13)(cid:13)(cid:13)(cid:13) (p ∈ R). 2010 Mathematics Subject Classification. 46B20, 46C15, 26D15. Key words and phrases. p-angular distance, skew p-angular distance, inequality, characteri- zation of inner product spaces. 1 2 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN We set β[x, y] for βp[x, y] when p = 0 and call it skew angular distance between non-zero elements x and y in X . Evidently, it holds that βp[x, y] = kxkp−1kykp−1α2−p[x, y]. (1.1) Dunkl and Williams [10] obtained a useful upper bound for the angular dis- tance. They showed that α[x, y] ≤ 4kx − yk kxk + kyk . The following result providing a lower bound for the p-angular distance was stated without a proof by Guraril in [12]: 2−pkx − ykp ≤ αp[x, y], where p ≥ 1 and x, y ∈ X . Finally, we recall the result of Hile [14]: αp[x, y] ≤ kykp − kxkp kyk − kxk kx − yk, (1.2) for p ≥ 1 and x, y ∈ X with kxk 6= kyk. For some recently obtained upper and lower bounds for the p-angular distance the reader is referred to [8, 9] and [16]. Numerous basic characterizations of inner product spaces under various condi- tions were first given by Fr´echet, Jordan and von Neumann; see [4] and references therein. Since then, the problem of finding necessary and sufficient conditions for a normed space to be an inner product space has been investigated by many mathematicians by considering some types of orthogonality or some geometric aspects of underlying spaces; see, e.g., [11, 15]. There is an interesting book by Amir [2] that contains several characterizations of inner product spaces, which are based on norm inequalities, various notions of orthogonality in normed linear spaces and so on. Among significant characterizations of inner product spaces related to p-angular distance, we can mention [1, 4, 5, 6]. The next two theorems due to Lorch and Ficken will be used in this paper. Theorem A(Lorch) [15]. Let (X ,k · k) be a normed space. Then the following statements are mutually equivalent: (i) For each x, y ∈ X if kxk = kyk, then kx + yk ≤ kλx + λ−1yk (for all (ii) For each x, y ∈ X if kx + yk ≤ kλx + λ−1yk (for all λ 6= 0), then λ 6= 0). kxk = kyk. (X ,k · k) is an inner product space. (iii) GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 3 Theorem B (Ficken) [11]. Let (X ,k·k) be a normed space. Then the follow- ing statements are mutually equivalent: (i) For each x, y ∈ X if kxk = kyk, then kαx + βyk = kβx + αyk (for all α, β > 0). (ii) For each x, y ∈ X if kxk = kyk, then kλx + λ−1yk = kλ−1x + λyk for all λ > 0. (X ,k · k) is an inner product space. (iii) In this paper, first we study some topological aspects of p-angular distances such as metrizability, consistency and completeness. Then, we compare two ar- bitrary p-angular and q-angular distances with each other, which generalize the results of Maligranda [16] and Dragomir [8]. Finally, we present two different characterizations of inner product spaces related to the p-angular and the skew p-angular distances. 2. Some initial observations In this section, first we examine some topological facts of the p-angular and the skew p-angular distances. Then we compare the p-angular distance with the skew p-angular distance in inner product spaces and give suitable representations for the p-angular distance, which will be used in the sequel for characterizations of inner product spaces. 2.1. Geometric properties of the p-angular distance. In this subsection, we study the metrizability, the consistency and the completeness concepts regarding to the p-angular and the skew p-angular distances. Theorem 2.1. For p 6= 0, αp[x, y] is a metric on X r {0}, which is consistent with α1[x, y] = kx − yk; they induce the same topology on X r {0}. If p and q are distinct non-zero real numbers, then αp is not equivalent with αq. If p 6= 1, then αp is not translation invariant. Proof. Clearly αp is a metric. Let α1[xn, x] = kxn− xk → 0 as n → ∞ in X r{0}. Thus limn→∞kxnk = kxk, and so αp[xn, x] =(cid:13)(cid:13)kxnkp−1xn − kxkp−1x(cid:13)(cid:13) ≤(cid:13)(cid:13)kxnkp−1xn − kxnkp−1x(cid:13)(cid:13) +(cid:13)(cid:13)kxnkp−1x − kxkp−1x(cid:13)(cid:13) ≤ kxnkp−1kxn − xk + kxk(cid:12)(cid:12)kxnkp−1 − kxkp−1(cid:12)(cid:12) → 0 Therefore the topology of αp is weaker than the topology of α1 on X r {0}. (as n → ∞). 4 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN Now we assume that αp[xn, x] → 0 as n → ∞ in X r {0}. We have xn x (cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:13)(cid:13)(cid:13)(cid:13) kxk1−p(cid:12)(cid:12)(cid:12)(cid:12) kxnk kxnk1−p − kxk → 0 (as n → ∞), and so limn→∞kxnkp = kxkp, which implies that limn→∞kxnk = kxk. Thus, kxk1−p(cid:13)(cid:13)(cid:13)(cid:13) α1[xn, x] = kxn − xk = kxk1−p(cid:13)(cid:13)kxkp−1xn − kxkp−1x(cid:13)(cid:13) kxnk1−p − ≤ kxk1−p((cid:13)(cid:13)kxkp−1xn − kxnkp−1xn(cid:13)(cid:13) +(cid:13)(cid:13)kxnkp−1xn − kxkp−1x(cid:13)(cid:13)) = kxk1−p(kxnk(cid:12)(cid:12)kxkp−1 − kxnkp−1(cid:12)(cid:12) + αp[xn, x]) → 0 (as n → ∞). Therefore the topology of α1 is weaker than the topology of αp on X r{0}. Hence these two metrics are consistent on X r {0}. Next, let p, q ∈ R r {0} such that p < q. By contrary, assume that there exists a number M > 0 such that for every x, y ∈ X r {0}, αp[x, y] ≤ M αq[x, y]. Fix a unit vector a ∈ X r {0}. For each λ, µ > 0, we have αp[λa, µa] ≤ M αq[λa, µa], or λp − µp ≤ Mλq − µq. In particular, if we put λ = 1 n where n = 1, 2, . . . and t > 0, then we have nq−p1 − tp ≤ M1 − tq, or and µ = t M ≥ nq−p(cid:12)(cid:12) 1−tq(cid:12)(cid:12) (t 6= 1). Now letting t → ∞ in the case when p < q < 0, and t → 0 in the case when 0 < p < q, we get M ≥ nq−p (n = 1, 2, . . .), and so M = ∞, which is a contradiction. In the case where p < 0 < q taking µ = 1, we obtain λp−1 ≤ Mλq−1. Now letting λ → 0+, we get M = ∞, a contradiction. Therefore αp is not equivalent to αq. 1−tp n Now, we show that if p 6= 1, then αp is not translation invariant. By contrary, assume that for each x, y, z ∈ X we have αp[x+z, y+z] = αp[x, y], whenever x, y, x+z, y+z 6= 0. Fixing a unit vector a ∈ X r {0}, put x = λa, y = γa and z = µa, where λ, µ, γ ∈ R. In particular, if we put λ = µ = 1 and γ > 0, then we have 2p − (γ + 1)p = 1 − γp. Now letting γ → ∞ in the case where p < 0, and γ → 0 in the case where p > 0, we get a contradiction. In the case where p = 0, we also get a contradiction by taking λ = 1, µ = −2 and γ = −1. This completes the proof. Remark 2.2. It may happen that two metrics d1 and d2 on a set E are consistent and there exists m > 0 such that md2 ≤ d1 but there exists no M > 0 such that d1 ≤ M d2. For a classical example, take E = [1,∞), d1(x, y) = x − y and d2(x, y) = (cid:12)(cid:12) y(cid:12)(cid:12). Two metrics d1 and d2 induce the same topology on E and d2(x, y) ≤ d1(x, y), but since d2 is bounded, there exists no M > 0 such that d1 ≤ M d2. Since in Theorem 2.1, p and q are arbitrary, this case cannot occur. In spite of αp, the following remark shows that when p 6= 1, never βp is a metric on X r {0}. Remark 2.3. Let X be a normed linear space. Take a ∈ X with kak = 1, and put x = ra, y = sa, z = ta, where r, s, t ∈ R. Let p > 1 and take r = 1, s = −1 and x − 1 (cid:3) 1 GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 5 t > 0. We obtain βp[x, y] = 2 > tp−1 − t + tp−1 + t = βp[x, z] + βp[y, z], for small enough t. This shows that βp is not a metric on X r {0} in this case. Now let p < 1, and take r = 2, t = 1 and s > 0. Since for small enough s, (2s)1−pβp[x, y] = 22−p − s2−p > s1−p(22−p − 1) + 21−p(s2−p − 1) = (2s)1−pβp[x, z] + (2s)1−pβp[y, z], βp is not a metric on X r {0}. Now we are going to compare completeness of an arbitrary nonempty subset of X r {0} with respect to αp and αq. To do this, we need some lemmas. Lemma 2.4. Let p 6= 0, A be a nonempty αp-complete subset of X r {0} and {xn} be a Cauchy sequence in X r {0}. Then If p > 0, then A is norm-bounded from below, and if p < 0, then A is norm-bounded from above. (i) (ii) If p > 0, then {xn} is norm-bounded from above, and if p < 0, then {xn} is norm-bounded from below. Proof. (i) Let p > 0. By contrary, assume that there exists a sequence {xn} in A such that limn→∞kxnk = 0. Therefore αp[xm, xn] =(cid:13)(cid:13)(cid:13) xm kxmk1−p − xn kxnk1−p(cid:13)(cid:13)(cid:13) ≤ kxmkp + kxnkp → 0 (m, n → ∞), and so {xn} is a αp-Cauchy sequence. Since A is αp-complete, there exists x ∈ A such that limn→∞ αp[xn, x] = 0. Since kxnkp − kxkp ≤ αp[xn, x], we get kxkp = limn→∞ kxnkp − kxkp ≤ 0, and so x = 0, which is a contradiction. Therefore, A is norm-bounded from below. Now, let p < 0. If A is not norm-bounded from above, then there exists a sequence {xn} in A such that limn→∞kxnk = ∞. By a similar argument we conclude that {xn} is a αp-Cauchy sequence in A and so there exists x ∈ A such that limn→∞ αp[xn, x] = 0. Therefore kxkp = 0, which is impossible. (cid:3) (ii) Obvious. The following lemma comparing αp with αq without any restrictions on p and q, plays an essential role in our study. 6 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN Lemma 2.5. Let p, q ∈ R and q 6= 0. Then for any non-zero elements x, y ∈ X , (2.1) (2.2) p p + p − q min(kxkp−q,kykp−q)αq[x, y] ≤ αp[x, y] ≤ q + p − q q max(kxkp−q,kykp−q)αq[x, y]. In particular if q = 1, then p p + p − 1 min(kxkp−1,kykp−1)kx − yk ≤ αp[x, y] ≤ (1 + p − 1) max(kxkp−1,kykp−1)kx − yk. Proof. We have αp[x, y] =(cid:13)(cid:13)kxkp−1x − kykp−1y(cid:13)(cid:13) ≤(cid:13)(cid:13)kxkp−qkxkq−1x − kxkp−qkykq−1y(cid:13)(cid:13) +(cid:13)(cid:13)kxkp−qkykq−1y − kykp−qkykq−1y(cid:13)(cid:13) = kxkp−qαq[x, y] + kykq(cid:12)(cid:12)kxkp−q − kykp−q(cid:12)(cid:12) . p−q q on the closed interval with endpoints kxkq and Consider the function f (t) = t kykq. By the Mean-Value Theorem, there exists a point η between kxkq and kykq such that Since the function t q (cid:12)(cid:12)kxkp−q − kykp−q(cid:12)(cid:12) = f (kxkq) − f (kykq) =(cid:12)(cid:12) is monotone, we obtain (cid:12)(cid:12)η q ≤ max(kxkp−2q,kykp−2q), p−2q p−2q η whence p − q q p−2q q kxkq − kykq . p − q q αp[x, y] ≤ kxkp−qαq[x, y] +(cid:12)(cid:12) p − q q ≤(cid:0)kxkp−q +(cid:12)(cid:12) αp[x, y] ≤ q + p − q (cid:12)(cid:12) max(kxkp−2qkykq,kykp−q)kxkq − kykq (cid:12)(cid:12) max(kxkp−2qkykq,kykp−q)(cid:1)αq[x, y]. max(kxkp−q,kxkp−2qkykq,kykp−q)αq[x, y]. Thus, By symmetry, we have q αp[x, y] ≤ q + p − q q max(kykp−q,kykp−2qkxkq,kxkp−q)αq[x, y]. (2.3) (2.4) GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 7 For proving (2.1) we can assume that kxk ≤ kyk. If q < 0, then kykq ≤ kxkq and so kxkp−2qkykq ≤ kxkp−q. Now, the right inequality in (2.1) follows from (2.3). Similarly if q > 0, then (2.4) yields the right inequality in (2.1). The left inequality in (2.1) follows from the right one by interchanging the roles of p and q. (cid:3) Theorem 2.6. The following statements hold. (i) (ii) If pq > 0, then for each ∅ 6= A ⊆ X r {0}, the metric space (A, αp) is complete if and only if (A, αq) is complete. If p > 0 and q < 0, then there exist nonempty sets A, B ⊆ X r {0} such that A is αp-complete but not αq-complete and B is αq-complete but not αp-complete. Proof. (i) Let ∅ 6= A ⊆ X r {0} be αp-complete. Assume {xn} is a αq-Cauchy sequence in A. First, suppose that p, q > 0. Since A is αp-complete, A, and as a result, {xn} is norm-bounded from below. On the other hand, since {xn} is αq- Cauchy, {xn} is norm-bounded from above. Thus, {xn} is norm-bounded from be- low and above, and so there exists 0 ≤ M < ∞ such that max(kxmkp−q,kxnkp−q) ≤ M (m, n = 1, 2, . . .). Therefore by the right hand side of inequality (2.1), αp[xm, xn] ≤ q + p − q q M αq[xm, xn] → 0 (m, n → ∞). Hence, {xn} is a αp-Cauchy sequence in A. Since A is αp-complete, there exists x ∈ A such that limn→∞ αp[xn, x] = 0, and by the consistency of αp and αq, we reach limn→∞ αq[xn, x] = 0. So, A is αq-complete. Now, let p, q < 0. Since A is αp-complete, {xn} is norm-bounded from above. On the other hand, since {xn} is αq-Cauchy, {xn} is norm-bounded from below. So, {xn} is again norm-bounded from above and below. Similar to the above argument, there exists x ∈ A such that limn→∞ αq[xn, x] = 0, and therefore A is αq-complete. (ii) Take a unit vector a ∈ X and let A = {λa : λ ≥ 1} and B = {λa : 0 < λ ≤ 1}. It is easily seen that A is αp-complete. Since q < 0 and A is not norm- bounded from above, A is not αq-complete. Similarly B is αq-complete, but not αp-complete. (cid:3) 2.2. p-angular distance in inner product spaces. In this part, we suppose that (X ,h·,·i) is a real inner product space with the induced norm k · k, defined by kxk2 = hx, xi. Proposition 2.7. Let X be an inner product space, x, y ∈ X r {0} and p ∈ R. Then the following properties hold. 8 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN (i) αp[x, y] ≤ βp[x, y] for all p < 1, (ii) αp[x, y] = βp[x, y] for p = 1, (iii) αp[x, y] ≥ βp[x, y] for all p > 1. In each of (i) and (iii) equality holds if and only if kxk = kyk. Proof. It is sufficient to note that αp[x, y]2 − βp[x, y]2 = (kxk2 − kyk2)(kxk2p−2 − kyk2p−2). (cid:3) Proposition 2.8. Let X be an inner product space, p ∈ R and x, y ∈ X r {0}. Then αp[x, y] =p(kxkp+1 − kykp+1)(kxkp−1 − kykp−1) + kxkp−1kykp−1kx − yk2. In particular if p = 0, then α[x, y] =skx − yk2 − (kxk − kyk)2 kxkkyk . Proof. The first identity follows from α2 p[x, y] = kxk2p + kyk2p − 2kxkp−1kykp−1Rehx, yi = kxk2p + kyk2p − kxkp−1kykp−1(kxk2 + kyk2 − kx − yk2) = (kxkp+1 − kykp+1)(kxkp−1 − kykp−1) + kxkp−1kykp−1kx − yk2. (cid:3) Now, we show the relation between the p-angular distance and the errors of the Cauchy-Schwarz inequality; kxkkyk± hx, yi ≥ 0. For this reason we need the following elementary lemma. Lemma 2.9. Let X be an inner product space. If x and y are linearly independent vectors of X and t ∈ R, then kx + tyk = pkxk2kyk2 − hx, yi2 kyk ∞ 2 whenever, −hx, yi −pkxk2kyk2 − hx, yi2 kyk2 tkyk2 + hx, yi Xk=0(cid:18) 1 k(cid:19)(cid:18) pkxk2kyk2 − hx, yi2(cid:19)2k ≤ t ≤ −hx, yi +pkxk2kyk2 − hx, yi2 . kyk2 , (2.5) GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 9 Proof. Employing the binomial series [13], we get kx + tyk = (t2kyk2 + 2thx, yi + kxk2) 1 2 1 2 1 2 kyk (cid:19)2 =(cid:20)(cid:18)tkyk + hx, yi = pkxk2kyk2 − hx, yi2 kyk = pkxk2kyk2 − hx, yi2 + kxk2 − hx, yi2 kyk2 (cid:21) pkxk2kyk2 − hx, yi2(cid:19)2 tkyk2 + hx, yi (cid:20)(cid:18) Xk=0(cid:18) 1 k(cid:19)(cid:18) tkyk2 + hx, yi pkxk2kyk2 − hx, yi2(cid:12)(cid:12)(cid:12)(cid:12) + 1(cid:21) pkxk2kyk2 − hx, yi2(cid:19)2k tkyk2 + hx, yi ≤ 1, kyk (cid:12)(cid:12)(cid:12)(cid:12) ∞ 2 , (cid:3) whenever which is equivalent to (2.5). Theorem 2.10. Let X be an inner product space and p ∈ R. If x and y are linearly independent vectors of X , then αp[x, y] = pkxk2kyk2 − hx, yi2 (2.6) ∞ , pkxk2kyk2 − hx, yi2 (cid:19)2k k(cid:19)(cid:18)kxk1−pkyk1+p − hx, yi 2 Xk=0(cid:18) 1 kxk1−pkyk whenever kykp ≤ √2kxkp and (cid:16) − 1 ≤ (cid:17) kxk−pkykp − p2 − kxk−2pkyk2p 2 ≤ hx, yi kxkkyk ≤ kxk−pkykp + p2 − kxk−2pkyk2p 2 (cid:16) ≤ 1(cid:17). (2.7) Similar expansion holds if we change the roles of x and y with each other. Proof. We have Taking t = − kykp−1 αp[x, y] = kxkp−1pkxk2kyk2 − hx, yi2 . 2 ∞ kyk kxkp−1 in Lemma 2.9, we reach αp[x, y] =(cid:13)(cid:13)kxkp−1x − kykp−1y(cid:13)(cid:13) = kxkp−1(cid:13)(cid:13)(cid:13)(cid:13) x − kykp−1 kxkp−1 y(cid:13)(cid:13)(cid:13)(cid:13) pkxk2kyk2 − hx, yi2 (cid:19)2k Xk=0(cid:18) 1 k(cid:19)(cid:18)kxk1−pkyk1+p − hx, yi kxkp−1 ≤ −hx, yi +pkxk2kyk2 − hx, yi2 ≤ −kykp−1 hx, yi − kxk1−pkyk1+p(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ≤pkxk2kyk2 − hx, yi2, kyk2 kyk2 , 2hx, yi2 − 2kxk1−pkyk1+phx, yi + kxk2kyk2(kxk−2pkyk2p − 1) ≤ 0, provided that, −hx, yi −pkxk2kyk2 − hx, yi2 But, this condition is in turn equivalent to or which is equivalent to kxk2kyk2(2 − kxk−2pkyk2p) ≥ 0 and (2.7). . (cid:3) 10 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN The following corollary shows that α[x, y] is completely expressible by kxk, kyk and the errors of the Cauchy-Schwarz inequality; kxkkyk ± hx, yi ≥ 0. Corollary 2.11. Let X be an inner product space. If x and y are linearly inde- pendent vectors of X , then whenever hx, yi ≥ 0, and kxkkyk α[x, y] = pkxk2kyk2 − hx, yi2 α[x, y] =vuut4 − kxk2kyk2 − hx, yi2 kxk2kyk2 ∞ Xk=0(cid:18) 1 2 kxkkyk + hx, yi(cid:19)k k(cid:19)(cid:18)kxkkyk − hx, yi , (2.8) (cid:20) ∞ Xk=0(cid:18) 1 2 k(cid:19)(cid:18)kxkkyk + hx, yi kxkkyk − hx, yi(cid:19)k(cid:21)2 , (2.9) whenever hx, yi < 0. Proof. The equality (2.8) follows from (2.6) by taking p = 0. If hx, yi < 0, then hx,−yi > 0, and so (2.9) follows from (2.8) and α[x, y] =p4 − α2[x,−y]. 3. Comparison of p-angular and q-angular distances (cid:3) In this section, we compare two quantities αp with αq for arbitrary p, q ∈ R. There are several papers related to comparison of αp with α1; see, e.g., [8]-[10]. The advantage of taking p and q arbitrary is that, whenever we find an inequality involving αp and αq, we can obtain its reverse by changing the roles of p and q with each other, which is as sharp as the first one. 3.1. Generalizations of Maligranda's results. The following theorem is a generalization of Maligranda's inequalities [16]. Theorem 3.1. Let p, q ∈ R, q 6= 0 and x, y ∈ X r {0}. (i) If p q ≥ 1, then p 2p − q max(kxkp−q,kykp−q)αq[x, y] ≤ αp[x, y] ≤ max(kxkp−q,kykp−q)αq[x, y]. p q (3.1) (ii) If 0 ≤ p αq[x, y] q ≤ 1, then p q · max(kxkq−p,kykq−p) ≤ αp[x, y] 2q − p ≤ · q αq[x, y] max(kxkq−p,kykq−p) . (3.2) GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 11 (iii) If p q ≤ 0, then p 2p − q · max(kxkp,kykp) max(kxkq,kykq) ≤ αp[x, y] 2q − p ≤ · q αq[x, y] max(kxkp,kykp) max(kxkq,kykq) αq[x, y]. (3.3) Proof. The left inequalities are obtained from right ones by interchanging the roles of p and q. So, it is sufficient to prove only the right inequalities. Without loss of generality we may assume that kxkq ≤ kykq. By the triangle inequality, we have αp[x, y] ≤(cid:13)(cid:13)kxkp−qkxkq−1x − kykp−qkxkq−1x(cid:13)(cid:13) +(cid:13)(cid:13)kykp−qkxkq−1x − kykp−qkykq−1y(cid:13)(cid:13) = kxkq(cid:12)(cid:12)kxkp−q − kykp−q(cid:12)(cid:12) + kykp−qαq[x, y]. q ≥ 0, we have kxkp−q ≤ kykp−q, and so αp[x, y] ≤ kxkq(kykp−q − kxkp−q) + kykp−qαq[x, y]. (i) Let p q ≥ 1. Since p−q If p q ≥ 2, then since p−2q q ≥ 0, we get Z kykq p − q q kxkq kykp−q − kxkp−q = t p−2q q dt ≤ p − q q kykp−2q(kykq − kxkq), which leads to Whence kxkq(kykp−q − kxkp−q) ≤ ≤ p − q q p − q q kxkqkykp−2q(kykq − kxkq) kykp−qαq[x, y]. p qkykp−qαq[x, y]. If 1 ≤ p q ≤ 2, it follows from p−2q p − q q kykp−q − kxkp−q = αp[x, y] ≤ q ≤ 0 that Z kykq kxkq t p−2q q dt ≤ p − q q kxkp−2q(kykq − kxkq), which gives kxkq(kykp−q − kxkp−q) ≤ p − q q kxkp−q(kykq − kxkq) ≤ p − q q kykp−qαq[x, y], and again αp[x, y] ≤ p qkykp−qαq[x, y]. 12 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN (ii) Let 0 ≤ p q ≤ 1. The inequality q−p q ≥ 0 yields that kxkq−p ≤ kykq−p, and so kxkq(cid:12)(cid:12)kxkp−q − kykp−q(cid:12)(cid:12) = kxkq kykq−p − kxkq−p kxkq−pkykq−p . It follows from kykq−p − kxkq−p = q dt ≤ q − p q kxk−p(kykq − kxkq), that Hence kxkq t− p q − p q Z kykq kxkq kykq−p − kxkq−p kxkq−pkykq−p ≤ · kykq − kxkq q − p kykq−p + q αp[x, y] ≤ . q q − p · kykq − kxkq kykq−p kykq−p ≤(cid:16)2 − αq[x, y] p q(cid:17) αq[x, y] kykq−p . (cid:3) (iii) The same reasoning as in the proof of (ii) yields (iii). Now, taking q = 1 in Theorem 3.1, we obtain the following corollary in which the right inequalities are due to Maligranda [16] and left ones are new suitable reverses to them. Corollary 3.2. Let x, y ∈ X r {0}. (i) (ii) (iii) p If p ≥ 1, then 2p − 1 If 0 ≤ p ≤ 1, then If p ≤ 0, then max(kxk,kyk)p−1kx − yk ≤ αp[x, y] ≤ p max(kxk,kyk)p−1kx − yk. pkx − yk max(kxk,kyk)1−p ≤ αp[x, y] ≤ (2 − p)kx − yk max(kxk,kyk)1−p . p 2p − 1 · max(kxkp,kykp) max(kxk,kyk) kx − yk ≤ αp[x, y] ≤ (2 − p) max(kxkp,kykp) max(kxk,kyk) kx − yk. Corollary 3.3. Let p 6= 2 and x, y ∈ X r {0}. (i) If p 2−p ≥ 1, then p 3p − 2 max(kxkp−1kyk1−p,kykp−1kxk1−p)βp[x, y] ≤ αp[x, y] ≤ max(kxkp−1kyk1−p,kykp−1kxk1−p)βp[x, y]. (ii) If 0 ≤ p p 2 − p 2−p ≤ 1, then p (2 − p) · βp[x, y] max(kxkp−1kyk1−p,kykp−1kxk1−p) ≤ αp[x, y] 4 − 3p 2 − p · ≤ βp[x, y] max(kxkp−1kyk1−p,kykp−1kxk1−p) . GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 13 (iii) If p 2−p ≤ 0, then 3p − 2 · p max(kxkp,kykp) βp[x, y] max(kxkkykp−1,kykkxkp−1) ≤ αp[x, y] 4 − 3p ≤ 2 − p · max(kxkp,kykp) max(kxkkykp−1,kykkxkp−1) In particular, for p = 0 and q = 1, it follows from (ii) that βp[x, y]. α[x, y] ≤ 2 min(cid:26)kxk kyk kxk(cid:27) β[x, y]. , kyk Proof. Take q = 2 − p in Theorem 3.1 and consider (1.1). Remark 3.4. In (3.1), (3.2) and (3.3), the constants 2 − p q in the right inequalities are best possible. In fact, consider X = R2 with the norm of x = (x1, x2) given by kxk = x1 +x2. Take x = (1 + ǫ) q (1, ǫ) and y = (1, 0), where ǫ > 0 is small. If p q and p (cid:3) 1−q q ≥ 1, then αp[x, y] αq[x, y] max(kxkp−q,kykp−q) as ǫ → 0+. In the case 0 ≤ p q ≤ 1, we obtain = p q (1 + ǫ) ǫ −1 − 1 · 1 (1 + ǫ) −1 p q + 1 → p q αp[x, y] αq[x, y] max(kxkq−p,kykq−p) = (1 + ǫ)1− p q − 1 + 1 → 2 − p q ǫ as ǫ → 0+. In the case when p q ≤ 0, the best possibility of the constant 2 − p q in the right inequality of (3.3) is similarly verified. The best possibility of constants and p possibility of constants 2 − p by changing the roles of p and q. 2p−q q in the left inequalities of (3.1), (3.2) and (3.3) are obtained from the best q in the right hand sides of these inequalities q and p p Remark 3.5. Let p, q ∈ R and q 6= 0. It is easily seen that in the case where p q ≥ 1 (resp. 0 ≤ p q ≤ 1), the right (resp. left) hand side of inequality (3.1) (resp. (3.2)) is as the same as the right (resp. left) hand side of inequality (2.1), but the left (resp. right) hand side of inequality (3.1) (resp. (3.2)) gives better estimate than the left (resp. right) hand side of inequality (2.1). In the case when p q ≤ 0, using the fact that min(cid:16) a c , b d(cid:17) ≤ max(a, b) max(c, d) ≤ max(cid:16) a c , b d(cid:17) (a, b, c, d > 0), both sides of (3.3) are better estimates than both corresponding sides of (2.1). 14 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN 3.2. Generalization of Dragomir's results. The following theorem yields the result of Dragomir in [8], if we take q = 1. Theorem 3.6. Let x, y ∈ X r {0}, p, q ∈ R and q 6= 0. (i) If p q ≥ 1, then p q (ii) If p αp[x, y] ≤ q < 1 and x, y are linearly independent, then αq[x, y]Z 1 αq[x, y]Z 1 0 (cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) 0 (cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) 2q − p αp[x, y] ≤ q p q −1 dt. (3.4) −1 p q dt. (3.5) Proof. We suppose that x, y are linearly independent and prove (3.4) and (3.5) by one strike. As one can observe, this proof works also in the case when p q ≥ 1 and x, y are linearly dependent. The function f : [0, 1] → [0,∞) given by and the vector-valued function h : [0, 1] → X given by p −1 , q f (t) =(cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) h(t) =(cid:2)(1 − t)kxkq−1x + tkykq−1y(cid:3), are both absolutely continuous on [0, 1]. Therefore, the function g : [0, 1] → X given by g(t) = f (t)h(t) is absolutely continuous. The function k(t) := k(1 − t)kxkq−1x + tkykq−1yk is convex, and so except than at most a countable number of points, k′(t) exists. It is easily verified that k′(t) ≤ αq[x, y] in each differentiability point t. We have g′(t) = f ′(t)h(t) + f (t)h′(t) p q p q − 1(cid:17)k(t) for almost all t ∈ [0, 1]. Thus, q − 1(cid:12)(cid:12)(cid:12)(cid:12) =(cid:16) p kg′(t)k ≤(cid:18)(cid:12)(cid:12)(cid:12)(cid:12) αp[x, y] =(cid:13)(cid:13)kykp−1y − kxkp−1x(cid:13)(cid:13) = kg(1) − g(0)k =(cid:13)(cid:13)(cid:13)(cid:13) −2k′(t)h(t) + f (t)(cid:2)kykq−1y − kxkq−1x(cid:3) + 1(cid:19)(cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) g ′(t)dt(cid:13)(cid:13)(cid:13)(cid:13) 0 (cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) + 1(cid:19)αq[x, y]Z 1 and so, the proofs of (3.4) and (3.5) are complete. ≤Z 1 Z 1 dt, −1 −1 p q p q 0 for almost all t ∈ [0, 1]. Utilizing the norm inequality for the vector-valued integral, we get 0 kg ′(t)kdt αq[x, y] (cid:3) ≤(cid:18)(cid:12)(cid:12)(cid:12)(cid:12) p q − 1(cid:12)(cid:12)(cid:12)(cid:12) Corollary 3.7. Let x, y ∈ X be linearly independent and p, q ∈ R r {0}. GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 15 (i) If 0 < p q ≤ 1, then αp[x, y] ≥ If p q ≥ 1 or p (ii) αp[x, y] ≥ 2p − q Remark 3.8. (i) If p p p q q p −1 q < 0, then αq[x, y](cid:18)Z 1 0 (cid:13)(cid:13)(1 − t)kxkp−1x + tkykp−1y(cid:13)(cid:13) αq[x, y](cid:18)Z 1 0 (cid:13)(cid:13)(1 − t)kxkp−1x + tkykp−1y(cid:13)(cid:13) q ≥ 1, then, by the triangle inequality, we have −1 ≤ [(1 − t)kxkq + tkykq] p p q . dt(cid:19)−1 dt(cid:19)−1 −1 . −1 p q for any t ∈ [0, 1]. Integrating both sides on [0, 1], we get if kxk 6= kyk, and by (3.4) we obtain the chain of inequalities −1 p q dt ≤ q p(cid:18)kykp − kxkp kykq − kxkq(cid:19) q (cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) Z 1 0 (cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) αp[x, y] ≤ p q αq[x, y]Z 1 ≤ kykp − kxkp kykq − kxkq 0 (cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) αq[x, y], −1 p q dt (3.6) which provides a generalization and refinement of Hile's inequality (1.2). p (ii) If p q ≥ 2, then the function f : [0, 1] → [0,∞) given by f (t) = [(1− t)kxkq + −1 is convex. Employing the Hermite-Hadamard inequality for the convex q tkykq] function f (see [17] and references therein) we obtain q kykq − kxkq(cid:19) =Z 1 p(cid:18)kykp − kxkp 0 [(1 − t)kxkq + tkykq] p q −1dt ≤ kxkp−q + kykp−q 2 ≤ max{kxkp−q,kykp−q}, which by (3.4), implies the following sequence of inequalities 0 (cid:13)(cid:13)(1 − t)kxkq−1x + tkykq−1y(cid:13)(cid:13) [(1 − t)kxkq + tkykq] −1dt p q −1 p q dt p q p q αp[x, y] ≤ αq[x, y]Z 1 αq[x, y]Z 1 ≤ = kykp − kxkp kykq − kxkq αq[x, y]kxkp−q + kykp−q p ≤ q for kxk 6= kyk. αq[x, y] 2 0 p q ≤ αq[x, y] max{kxkp−q,kykp−q} (3.7) 16 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN In particular, inequality (3.7) shows that in the case p q ≥ 2, inequality (3.6) is better than inequality (3.1). Remark 3.9. Let X be an inner product space. a, b ∈ X , b 6= 0, it holds that It is known [7] that for any min t∈R ka + tbk = pkak2kbk2 − ha, bi2 . kbk Hence, if x and y are linearly independent vectors of X , then by taking a = x and b = y − x, we obtain k(1 − t)x + tyk = kx + t(y − x)k ≥ pkxk2kyk2 − hx, yi2 kx − yk (t ∈ R). This implies that Z 1 0 k(1 − t)x + tyk−1dt ≤ Taking p = 0 and q = 1 in (3.5), we get kx − yk pkxk2kyk2 − hx, yi2 . α[x, y] ≤ 2kx − ykZ 1 0 k(1 − t)x + tyk−1dt ≤ 2kx − yk2 . pkxk2kyk2 − hx, yi2 This implies an upper estimation for the error of the Cauchy-Schwarz inequality as follows pkxk2kyk2 − hx, yi2 ≤ 2kxkkykkx − yk2 kkykx − kxkyk (kykx 6= kxky). 4. Characterizations of inner product spaces In this section, corresponding to Propositions 2.7 and 2.8, we give two char- acterizations of inner product spaces regarding to the p-angular and the skew p-angular distances. The following characterization extends a result of Dehghan [6] from p = 0 to an arbitrary real number p 6= 1. Theorem 4.1. Let p > 1 (p < 1 resp.) is a real number. Then a normed space X is an inner product space, if and only if for any x, y ∈ X r {0}, αp[x, y] ≥ βp[x, y] (αp[x, y] ≤ βp[x, y] resp.). (4.1) Proof. If X is an inner product space, the conclusion follows from Proposition 2.7. Now, let X be a normed space satisfying the condition (4.1). Since for arbitrary non-zero elements x and y of X , the inequality αp[x, y] ≤ βp[x, y] is equivalent to β2−p[x, y] ≤ α2−p[x, y], it is sufficient to consider the case when p > 1. GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 17 Let x, y ∈ X ,kxk = kyk and γ 6= 0. From Theorem A it is enough to prove that kγx + γ−1yk ≥ kx + yk. Clearly, we can assume that kxk = kyk = 1 and p y for x and y respectively, we γ > 0. Applying inequality (4.1) to γ obtain p x and −γ− 1 1 Now using the mathematical induction, we get kγx + γ−1yk ≥(cid:13)(cid:13)γ 2−p p x + γ− 2−p p y(cid:13)(cid:13). (4.2) 2−p kγx + γ−1yk ≥(cid:13)(cid:13)(cid:13)(cid:13) p (cid:1)n γ(cid:0) 2−p p (cid:12)(cid:12) < 1, and so n→∞(cid:13)(cid:13)(cid:13)(cid:13) p (cid:1)n γ(cid:0) 2−p kγx + γ−1yk ≥ lim p (cid:1)n x + γ−(cid:0) 2−p y(cid:13)(cid:13)(cid:13)(cid:13) p (cid:1)n x + γ−(cid:0) 2−p (n = 1, 2, . . .). = kx + yk. y(cid:13)(cid:13)(cid:13)(cid:13) (cid:3) Since p > 1, we have (cid:12)(cid:12) This completes the proof. Remark 4.2. If X is not an inner product space, then for each p 6= 1 there exist xi, yi ∈ X r {0} (i = 1, 2), such that αp[x1, y1] < βp[x1, y1] and αp[x2, y2] > βp[x2, y2]. In fact if p > 1, then by Theorem 4.1 there exist x1, y1 ∈ X r {0} such that αp[x1, y1] < βp[x1, y1]. On the other hand, due to an arbitrary one dimensional subspace M = {λe : λ ∈ R} of X with kek = 1 is an inner product space via hλe, µei := λµ, for any x2, y2 ∈ M r {0} with kx2k 6= ky2k, we have αp[x2, y2] > βp[x2, y2]. A similar argument carry out when p < 1. Now we give the second characterization of inner product spaces related to Proposition 2.8. Theorem 4.3. Let p 6= 1. Then a normed space X is an inner product space if and only if for any x, y ∈ X r {0}, αp[x, y] =p(kxkp+1 − kykp+1)(kxkp−1 − kykp−1) + kxkp−1kykp−1kx − yk2. Proof. If X is an inner product space, then identity (4.3) follows from Proposition 2.8. Now, let X be a normed space satisfying condition (4.3). We prove that X is an inner product space by considering the following three cases for p. (4.3) Case 1. Assume that p 6= 0,−1. Let x, y ∈ X ,kxk = kyk and λ 6= 0. From Theorem A it is enough to prove that kx + yk ≤ kλx + λ−1yk. We may assume that kxk = kyk = 1 and λ > 0. Applying identity (4.3) to λ p y for x p x and −λ− 1 1 18 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN and y respectively, we obtain kλx + λ−1yk =q(λ =s λ 2p+2 p+1 p − λ− p+1 p − 1 λ λ · p+1 p 2p−2 p − 1 λ p−1 p p )(λ p−1 p − λ 1−p p ) + kλ 1 p x + λ− 1 p yk2 + kλ 1 p x + λ− 1 p yk2. If p > 1, then 2p + 2 > 2p − 2 > 0, and if p < −1, then 2p − 2 < 2p + 2 < 0. p − 1) ≥ 0. Hence kλx + λ−1yk ≥ For p > 1, we therefore have (λ kλ p yk. It yields that p − 1)(λ p x + λ− 1 2p−2 2p+2 1 (n = 1, 2, . . .). (4.4) kλx + λ−1yk ≥ kλ Thus, 1 pn x + λ− 1 pn yk kλx + λ−1yk ≥ lim n→∞kλ 1 pn x + λ− 1 pn yk = kx + yk. Now if p < 1, then 1 p > 1, and so by substituting p by 1 p in (4.4) we get kλx + λ−1yk ≥ kλpn x + λ−pn yk (n = 1, 2, . . .). Hence, kλx + λ−1yk ≥ lim and so, X is an inner product space. n→∞kλpnx + λ−pnyk = kx + yk, Case 2. Suppose that p = 0. Let x, y ∈ X , kxk = kyk = 1 and λ > 0. Replacing x and y by λx and −λ−1y respectively, in identity (4.3), we get kx + yk2 = kλx + λ−1yk2 − (λ − )2 ≤ kλx + λ−1yk2. 1 λ It follows from Theorem A that X is an inner product space. Case 3. Let p = −1. Assume x, y ∈ X such that kxk = kyk and λ > 0. Applying identity (4.3) to λx and −λ−1y instead of x and y respectively, we obtain kλx + λ−1yk = kλ−1x + λyk. Therefore, Theorem B ensures that X is an inner product space. (cid:3) Remark 4.4. It seems that the characterization of inner product spaces in The- orem 4.1 can be extended in a more general case. For example, the following inequality x 1 + kxk − (cid:13)(cid:13)(cid:13)(cid:13) y 1 + kyk(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13) x 1 + kyk − y 1 + kxk(cid:13)(cid:13)(cid:13)(cid:13) is also a characterization of inner product spaces. In fact, (4.5) holds in any inner prod- uct spaces and conversely, if (4.5) holds in a normed linear space X , then substituting x and y by nx and ny (n = 1, 2, . . .) respectively, we obtain (x, y ∈ X ), (4.5) x 1 n + kxk − (cid:13)(cid:13)(cid:13)(cid:13) 1 y n + kyk(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13) x 1 n + kyk − . 1 y n + kxk(cid:13)(cid:13)(cid:13)(cid:13) GEOMETRIC ASPECTS OF p-ANGULAR AND SKEW p-ANGULAR DISTANCES 19 Now letting n → ∞, we get α[x, y] ≤ β[x, y] (x, y ∈ X r {0}), and so X is an inner product space. Acknowledgement. The authors would like to sincerely thank the anonymous referee for carefully reading the article. The corresponding author, M. S. Mosle- hian, would like to thank The Tusi Mathematical Research Group (TMRG). References 1. A. M. Al-Rashed, Norm inequalities and characterizations of inner product spaces, J. Math. Anal. Appl. 176 (1993) 587–593. 2. D. Amir, Characterizations of inner product spaces, Operator Theory: Advances and Ap- plications, 20, Birkhauser Verlag, Basel, 1986.. 3. J. A. Clarkson, Uniformly convex spaces, Trans. Amer. Math. Soc. 40 (1936), 396–414. 4. F. Dadipour, M. S. Moslehian, A characterization of inner product spaces related to the p-angular distance, J. Math. Anal. Appl. 371 (2010), no. 2, 677–681. 5. F. Dadipour, F. Sadeghi, A. Salemi, Characterizations of inner product spaces involving homogeneity of isosceles orthogonality. Arch. Math. (Basel) 104 (2015), no. 5, 431–439. 6. H. Dehghan, A characterization of inner product spaces related to the skew-angular distance, Math. Notes 93 (2013), no. 4, 556–560. 7. S. S. Dragomir, A survey of some recent inequalities for the norm and numerical radius of operators in Hilbert spaces, Banach J. Math. Anal. 1 (2007), no. 2, 154–175. 8. S. S. Dragomir, New inequalities for the p-Angular distance in normed spaces with applica- tions, Ukrainian Math. J. 67 (2015), no. 1, 19–32. 9. S. S. Dragomir, Upper and lower bounds for the p-angular distance in normed spaces with applications, J. Math. Inequal. 8 (2014), no. 4, 947–961. 10. C. F. Dunkl, K.S. Williams, Mathematical notes: A simple norm inequality, Amer. Math. Monthly 71 (1964), no. 1, 53–54. 11. F. A. Ficken, Note on the existence of scalar products in normed linear spaces, Ann. of Math. (2) 45 (1944), 362–366. 12. V. I. Guraril, Strengthening the Dunkl-Williams inequality on the norm of elements of Banach spaces, Dopov¯ıd¯ı Akad. Nauk Ukraın. RSR (1966), 35–38 (Ukrainian). 13. E. Hewitt, K. Stromberg, Real and abstract analysis, Springer, New York, 1965. 14. G. N. Hile, Entire solutions of linear elliptic equations with Laplacian principal part. Pacific J. Math. 62 (1976), no. 1, 127–140. 15. E. R. Lorch, On certain implications which characterize Hilbert space, Ann. of Math. 49 (1948), no. 3, 523–532. 16. L. Maligranda, Simple norm inequalities, Amer. Math. Monthly. 113 (2006), no. 3, 256–260. 17. M. S. Moslehian, Matrix Hermite–Hadamard type inequalities, Houston J. Math. 39 (2013), no. 1, 177–189. 20 J. ROOIN, S. HABIBZADEH, M.S. MOSLEHIAN 1Department of Mathematics, Institute for Advanced Studies in Basic Sci- ences (IASBS), Zanjan 45137-66731, Iran E-mail address: [email protected] 2Department of Mathematics, Institute for Advanced Studies in Basic Sci- ences (IASBS), Zanjan 45137-66731, Iran; The Tusi Mathematical Research Group (TMRG), Mashhad, Iran. E-mail address: [email protected] 3 Department of Pure Mathematics, Center of Excellence in Analysis on Al- gebraic Structures (CEAAS), Ferdowsi University of Mashhad, P. O. Box 1159, Mashhad 91775, Iran. E-mail address: [email protected], [email protected]
1502.00083
1
1502
2015-01-31T08:41:32
Extension of Euclidean operator radius inequalities
[ "math.FA", "math.OA" ]
To extend the Euclidean operator radius, we define $w_p$ for an $n$-tuples of operators $(T_1,\ldots, T_n)$ in $\mathbb{B}(\mathscr{H})$ by $w_p(T_1,\ldots,T_n):= \sup_{\| x \| =1} \left(\sum_{i=1}^{n}| \langle T_i x, x \rangle |^p \right)^{\frac1p}$ for $p\geq1$. We generalize some inequalities including Euclidean operator radius of two operators to those involving $w_p$. Further we obtain some lower and upper bounds for $w_p$. Our main result states that if $f$ and $g$ are nonnegative continuous functions on $\left[ 0,\infty \right) $ satisfying $f\left( t\right) g\left(t\right) =t$ for all $t\in \left[ 0,\infty \right) $, then \begin{equation*} w_{p}^{rp}\left( A_{1}^{\ast }T_{1}B_{1},\ldots ,A_{n}^{\ast }T_{n}B_{n}\right) \leq \frac{1}{2}\left\Vert \underset{i=1}{\overset{n}{\sum }}\Big( \left[ B_{i}^{\ast }f^{2}\left( \left\vert T_{i}\right\vert \right) B_{i}\right] ^{rp}+\left[ A_{i}^{\ast }g^{2}\left( \left\vert T_{i}^{\ast }\right\vert \right) A_{i}\right] ^{rp}\Big)\right\Vert \end{equation*} for all $p\geq 1$, $r\geq 1$ and operators in $ \mathbb{B}(\mathscr{H})$.
math.FA
math
EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES MOHAMMAD SAL MOSLEHIAN1, MOSTAFA SATTARI1 AND KHALID SHEBRAWI2 Abstract. To extend the Euclidean operator radius, we define wp for an n-tuples of operators (T1, . . . , Tn) in B(H ) by wp(T1, . . . , Tn) := supkxk=1 (Pn i=1 hTix, xip) for p ≥ 1. We generalize some inequalities including Euclidean operator radius of two operators to those involving wp. Further we obtain some lower and upper 1 p bounds for wp. Our main result states that if f and g are nonnegative continuous functions on [0,∞) satisfying f (t) g (t) = t for all t ∈ [0,∞), then +(cid:2)A∗ wrp 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1(cid:16)(cid:2)B∗ for all p ≥ 1, r ≥ 1 and operators in B(H ). i ) Ai(cid:3)rp(cid:17)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) i f 2 (Ti) Bi(cid:3)rp p (A∗ 1T1B1, . . . , A∗ i g2 (T ∗ n 1 nTnBn) ≤ 1. Introduction Let B(H ) be the C ∗-algebra of all bounded linear operators on a Hilbert space (H ,h·,·i). The numerical radius of A ∈ B(H ) is defined by w(A) = sup{hAx, xi : x ∈ H ,kxk = 1}. It is well known that w(·) defines a norm on B(H ), which is equivalent to the usual operator norm k · k. Namely, we have 1 2kAk ≤ w(A) ≤ kAk. for each A ∈ B(H ). It is known that if A ∈ B(H ) is self-adjoint, then w(A) = kAk. An important inequality for w(A) is the power inequality stating that w(An) ≤ wn(A) for n = 1, 2, . . .. There are many inequalities involving numerical radius; see [2, 3, 4, 10, 11, 12] and references therein. The Euclidean operator radius of an n-tuple (T1, . . . , Tn) ∈ B(H )(n) := B(H ) × 2010 Mathematics Subject Classification. 47A12, 47B15, 47A30, 47A63. Key words and phrases. Euclidean operator radius, numerical radius, Cartesian decomposition, self-adjoint operator. 1 The particular cases n = 1 and n = 2 are numerical radius and Euclidean operator radius. Some interesting properties of this radius were obtained in [9]. For example, it is established that n 1 2√n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 TiT ∗ 1 2 i(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n ≤ we(T1,· · · , Tn) ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 i(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 2 . (1.1) TiT ∗ We also observe that if A = B + iC is the Cartesian decomposition of A, then w2 e(B, C) = sup kxk=1{hBx, xi2 + hCx, xi2} = sup kxk=1hAx, xi2 = w2(A). By the above inequality and A∗A + AA∗ = 2(B2 + C 2), we have 1 16kA∗A + AA∗k ≤ w2(A) ≤ 1 2kA∗A + AA∗k. We define wp for n-tuples of operators (T1, . . . , Tn) ∈ B(H )(n) for p ≥ 1 by wp(T1, . . . , Tn) := sup kxk=1 n Xi=1 hTix, xip! 1 p . 2 M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI . . . × B(H ) was defined in [9] by we(T1, . . . , Tn) := sup kxk=1 n Xi=1 hTix, xi2! 1 2 . It follows from Minkowski's inequality for two vectors a = (a1, a2) and b = (b1, b2), namely, (a1 + b1p + a2 + b2p) 1 p ≤ (a1p + a2p) 1 p + (b1p + b2p) 1 p for p > 1 that wp is a norm. Moreover wp, p ≥ 1, for n-tuple of operators (T1, . . . , Tn) ∈ B(H )(n) satisfies the following properties: (i) wp (T1, . . . , Tn) = 0 ⇔ T1 = . . . = Tn = 0. (ii) wp (λT1, . . . , λTn) = λ wp (T1, . . . , Tn) for all λ ∈ C. (iii) wp (T1 + T1´, . . . , Tn + Tn´) ≤ wp (T1, . . . , Tn)+wp (T1´, . . . , Tn´) for (T1´, . . . , Tn´) ∈ B(H )(n). (iv) wp (X ∗T1X, . . . , X ∗TnX) ≤ kXk2 wp (T1, . . . , Tn) for X ∈ B(H ). Dragomir [1] obtained some inequalities for the Euclidean operator radius we(B, C) = 1 supkxk=1 (hBx, xi2 + hCx, xi2) In section 2 of this paper we extend some his results including inequalities for the Euclidean operator radius of linear operators to wp (p ≥ 1). In addition, we apply some known inequalities for getting new inequalities for wp in two operators. 2 of two bounded linear operators in a Hilbert space. EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES 3 In section 3 we prove inequalities for wp for n-tuples of operators. Some of our result in this section, generalize some inequalities in section 2. Further, we find some lower and upper bounds for wp. 2. Inequalities for wp for two operators To prove our generalized numerical radius inequalities, we need several known lemmas. The first lemma is a simple result of the classical Jensen inequality and a generalized mixed Cauchy -- Schwarz inequality [7, 8, 6]. Lemma 2.1. For a, b ≥ 0, 0 ≤ α ≤ 1 and r 6= 0, (a) aαb1−α ≤ αa + (1 − α)b ≤ [αar + (1 − α)br] (b) If A ∈ B(H ), then hAx, yi2 ≤ hA2αx, xihA∗2(1−α)y, yi for all x, y ∈ H , r ≥ 1, for 1 r where A = (A∗A) 1 2 . (c) Let A ∈ B(H ), and f and g be nonnegative continuous functions on [0,∞) satisfying f (t) g (t) = t for all t ∈ [0,∞). Then hAx, yi ≤ kf (A) xk kg (A∗) yk for all x, y ∈ H . Lemma 2.2 (McCarthy inequality [5]). Let A ∈ B(H ), A ≥ 0 and let x ∈ H be any unit vector. Then (a) hAx, xir ≤ hArx, xi (b) hArx, xi ≤ hAx, xir Inequalities of the following lemma were obtained for the first time by Clarkson[7]. r ≥ 1, 0 < r ≤ 1. for for Lemma 2.3. Let X be a normed space and x, y ∈ X. Then for all p ≥ 2 with p + 1 1 q = 1, (a) 2(kxkp + kykp)q−1 ≤ kx + ykq + kx − ykq, (b) 2(kxkp + kykp) ≤ kx + ykp + kx − ykp ≤ 2p−1(kxkp + kykp), (c) kx + ykp + kx − ykp ≤ 2(kxkq + kykq)p−1. If 1 < p ≤ 2 the converse inequalities hold. Making the transformations x → x+y 2 we observe that inequalities (a) and (c) in Lemma 2.3 are equivalent and so are the first and the second inequalities of (b). First of all we obtain a relation between wp and we for p ≥ 1. 2 and y → x−y 4 M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI Proposition 2.4. Let B, C ∈ B(H ). Then wp (B, C) ≤ wq (B, C) ≤ 2 1 q − 1 p wp (B, C) for p ≥ q ≥ 1. In particular wp (B, C) ≤ we (B, C) ≤ 2 1 2 − 1 p wp (B, C) (2.1) for p ≥ 2, and 2 1 2 − 1 p wp (B, C) ≤ we (B, C) ≤ wp (B, C) for 1 ≤ p ≤ 2. Proof. An application of Jensen's inequality says that for a, b > 0 and p ≥ q > 0, we have (ap + bp) 1 p ≤ (aq + bq) 1 q . Let x ∈ H be a unit vector. Choosing a = hBx, xi and b = hCx, xi, we have (cid:16) hBx, xip + hCx, xip(cid:17) p ≤(cid:16)hBx, xiq + hCx, xiq(cid:17) 1 1 q . Now the first inequality follows by taking the supremum over all unit vectors in H . A simple consequence of the classical Jensen's inequality concerning the convexity or the concavity of certain power functions says that for a, b ≥ 0, 0 ≤ α ≤ 1 and p ≥ q, we have (αaq + (1 − α) bq) For α = 1 2, we get 1 q ≤ (αap + (1 − α) bp) 1 p . (aq + bq) 1 q 1 q ≤ 2 − 1 p (ap + bp) 1 p . Again let x ∈ H be a unit vector. Choosing a = hBx, xi and b = hCx, xi we get (cid:16)hBx, xiq + hCx, xiq(cid:17) 1 q ≤ 2 1 q − 1 p(cid:16)hBx, xip + hCx, xip(cid:17) 1 p . Now the second inequality follows by taking the supremum over all unit vectors in H . (cid:3) On making use of inequality (2.1) we find a lower bound for wp (p ≥ 2). Corollary 2.5. If B, C ∈ B(H ), then for p ≥ 2 wp(B, C) ≥ 2 1 p −2kB∗B + C ∗Ck 1 2 . EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES 5 Proof. According to inequalities (1.1) and (2.1) we can write and we(B, C) ≥ 1 2√2kB∗B + C ∗Ck 1 2 wp(B, C) ≥ 2 1 p − 1 2 we(B, C), respectively. We therefore get desired inequality. (cid:3) The next result is concerned with some lower bounds for wp. This consequence has several inequalities as special cases. Our result will be generalized to n-tuples of operators in the next section. Proposition 2.6. Let B, C ∈ B(H ). Then for p ≥ 1 wp(B, C) ≥ 2 1 p −1 max (w(B + C), w(B − C)) . (2.2) This inequality is sharp. Proof. We use convexity of function f (t) = tp (p ≥ 1) as follows: (hBx, xip + hCx, xip) 1 p ≥ 2 ≥ 2 = 2 1 p 1 p 1 p −1(hBx, xi + hCx, xi) −1hBx, xi ± hCx, xi −1h(B ± C)x, xi . Taking supremum over x ∈ H with kxk = 1 yields that −1w(B ± C). wp(B, C) ≥ 2 1 p For sharpness one can obtain the same quantity 2 1 p w(B) on both sides of the in- equality by putting B = C. (cid:3) Corollary 2.7. If A = B + iC is the Cartesian decomposition of A, then for all p ≥ 2 wp(B, C) ≥ 2 and 1 p −1 max (kB + Ck,kB − Ck) , 1 p w(A) ≥ 2 −2 max (k(1 − i)A + (1 + i)A∗k,k(1 + i)A + (1 − i)A∗k) Proof. Obviously by inequality (2.2) we have the first inequality. For the second we use inequality (2.1). (cid:3) 6 M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI Corollary 2.8. If B, C ∈ B(H ), then for p ≥ 1 wp(B, C) ≥ 2 1 p −1 max{w(B), w(C)}. (2.3) In addition, if A = B + iC is the Cartesian decomposition of A, then for p ≥ 2 1 p w(A) ≥ 2 −2 max (kA + A∗k,kA − A∗k) . Proof. By inequality (2.2) and properties of the numerical radius, we have 2wp(B, C) ≥ 2 1 p −1(w(B + C) + w(B − C)) ≥ 2 1 p −1w(B + C + B − C) . So By symmetry we conclude that wp(B, C) ≥ 2 1 p −1w(B) . wp(B, C) ≥ 2 1 p −1 max(w(B), w(C)). While the second inequality follows easily from inequality (2.1). (cid:3) Now we apply part (b) of Lemma 2.3 to find some lower and upper bounds for wp (p > 1). 1 p 1 Proposition 2.9. Let B, C ∈ B(H ). Then for all p ≥ 2, (i) 2 1 p −1wp(B + C, B − C) ≤ wp(B, C) ≤ 2− 1 −1(cid:0)wp(B + C) + wp(B − C)(cid:1) p(cid:0)wp(B + C) + wp(B − C)(cid:1) (ii) 2 If 1 < p ≤ 2 these inequalities hold in the opposite direction. Proof. Let x ∈ H be a unit vector. Part (b) of Lemma 2.3 implies that for any p ≥ 2 p ≤ wp(B, C) ≤ 2− 1 p wp(B + C, B − C); 1 p . 21−p(a + bp + a − bp) ≤ ap + bp ≤ (a + bp + a − bp) . 1 2 Replacing a = hBx, xi and b = hCx, xi in above inequalities we obtain the desired inequalities. (cid:3) Remark 2.10. In inequality (2.3), if we take B + C and B − C instead of B and C, then for p ≥ 1 −1 max{w(B + C), w(B − C)} . By employing the first inequality of part (i) of Proposition 2.9, we get wp(B + C, B − C) ≥ 2 1 p wp(B, C) ≥ 2 2 p −2 max{w(B + C), w(B − C)} EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES 7 for p ≥ 1. Taking B + C and B − C instead of B and C in the second inequality of part (ii) of Proposition 2.9, we reach wp(B + C, B − C) ≤ 21− 1 p (wp(B) + wp(C)) 1 p . for all p ≥ 1. Now by applying the second inequality of part (i) of Proposition 2.9, we infer for p ≥ 1 that wp(B, C) ≤ 21− 2 p (wp(B) + wp(C)) 1 p . So 2 p 2 −2 max{w(B + C), w(B − C)} ≤ wp(B, C) ≤ 21− 2 p (wp(B) + wp(C)) 1 p . Moreover if B and C are self-adjoint, then 2 p 2 −2 max{kB + Ck,kB − Ck} ≤ wp(B, C) ≤ 21− 2 p (kBkp + kCkp) 1 p for all p ≥ 1. In the following result we find another lower bound for wp (p ≥ 1). Theorem 2.11. Let B, C ∈ B(H ). Then for p ≥ 1 wp(B, C) ≥ 2 1 p −1w 1 2 (B2 + C 2). Proof. It follows from (2.2) that 2 p 2 −2w2(B ± C) ≤ w2 p(B, C) . Hence 2w2 p(B, C) ≥ 2 ≥ 2 ≥ 2 2 p 2 p 2 p It follows that −2(cid:2)w2(B + C) + w2(B − C)(cid:3) −2(cid:2)w(cid:0)(B + C)2(cid:1) + w(cid:0)(B − C)2(cid:1)(cid:3) −2(cid:2)w(cid:0)(B + C)2 + (B − C)2(cid:1)(cid:3) = 2 wp(B, C) ≥ 2 2 (B2 + C 2). 1 1 p −1w 2 p −1w(B2 + C 2) . (cid:3) 8 M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI Corollary 2.12. If A = B + iC is the Cartesian decomposition of A , then And wp(B, C) ≥ 2 1 p −1kB2 + C 2k 1 2 . w(A) ≥ 2 1 p − 3 2kA∗A + AA∗k 1 2 . for any p ≥ 2. Proof. The first inequality is obvious. For the second we have A∗A + AA∗ = 2(B2 + C 2). Now by using inequality (2.1) the proof is complete. (cid:3) Corollary 2.13. If B, C ∈ B(H ), then for p ≥ 2 wp(B, C) ≥ 2 2 p − 3 2 w 1 2 (cid:0)B2 + C 2(cid:1) . Proof. By choosing B + C and B − C instead of B and C in Theorem 2.11 and employing part (i) of Proposition 2.9 we conclude that the desired inequality. (cid:3) The following result providing other bound for wp (p > 1) may be stated as follows: Proposition 2.14. Let B, C ∈ B(H ). Then wp(B, C) ≤ wq(cid:18) B + C 2 , 2 (cid:19) . B − C for any p ≥ 2, 1 < q ≤ 2 with 1 Proof. Let x ∈ H be a unit vector. Part (a) of Lemma 2.3 implies that q = 1. If 1 < p ≤ 2, the reverse inequality holds. p + 1 So ap + bp ≤ 2 1 1−q (a + bq + a − bq) 1 q−1 . (ap + bp) 1 p ≤ 2 1 p(1−q) (a + bq + a − bq) 1 p(q−1) . Now replacing a = hBx, xi and b = hCx, xi in the above inequality we conclude that . (2.4) (cid:3) q 1 (cid:28)(cid:18) B + C (hBx, xip + hCx, xip) +(cid:12)(cid:12)(cid:12)(cid:12) (cid:28)(cid:18) B − C By taking supremum over x ∈ H with kxk = 1 we deduce that 2 (cid:19) B − C p ≤(cid:18)(cid:12)(cid:12)(cid:12)(cid:12) 2 (cid:19) x, x(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) wp(B, C) ≤ wq(cid:18) B + C 2 , 1 q q(cid:19) 2 (cid:19) x, x(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) for any p ≥ 2, 1 < q ≤ 2 with 1 p + 1 q = 1. EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES 9 Corollary 2.15. Inequality (2.4) implies that wp(B, C) ≤(cid:18)wq(cid:18) B + C 2 (cid:19) + wq(cid:18) B − C 2 (cid:19)(cid:19) 1 q . for any 1 < q ≤ 2, p ≥ 2 with 1 p + 1 q = 1. Further, if Band C are self-adjoint, then wp(B, C) ≤ 1 2 (kB + Ckq + kB − Ckq) 1 q . If 1 < p ≤ 2, the converse inequalities hold. Corollary 2.16. If B, C ∈ B(H ), then 2 (cid:19) ≤ 2 B − C q = 1. If p ≥ 2, the above inequality is valid in the wq(cid:18) B + C for all 1 < p ≤ 2 with 1 opposite direction. p wp(cid:18) B + C 2 (cid:19) . B − C p + 1 2 2 , , 1 Proof. By Proposition 2.14 we have wq(cid:18) B + C 2 , 2 (cid:19) ≤ wp(B, C). B − C for all 1 < p ≤ 2 with 1 p + 1 q = 1. Proposition 2.9 follows that wp(B, C) ≤ 2 1 p −1wp(B + C, B − C) = 2 1 p wp(cid:18) B + C 2 , 2 (cid:19) . B − C We therefore get the desired inequality. (cid:3) 3. Inequalities of wp for n-tuples of operators In this section, we are going to obtain some numerical radius inequalities for n- tuples of operators. Some generalization of inequalities in the previous section are also established. According to the definition of numerical radius, we immediately get the following double inequality for p ≥ 1 wp (T1, . . . , Tn) ≤ n Xi=1 wp (Ti)! Xi=1 w (Ti) . ≤ 1 p n An application of Holder's inequality gives the next result, which is a generalization of inequality (2.2). 10 M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI n Theorem 3.1. Let (T1, . . . , Tn) ∈ B(H )(n) and 0 ≤ αi ≤ 1, i = 1, . . . n, with Pi=1 αi = 1. Then wp (T1, . . . , Tn) ≥ w(cid:18)α 1− 1 1− 1 p 1 T1 ± α p 2 T2 ± . . . ± α p 1− 1 n Tn(cid:19) for any p > 1. Proof. In the Euclidean space Rn with the standard inner product, Holder's inequal- ity q = 1 and (x1, . . . , xn), n Xi=1 xip! xiyi ≤ n yiq! Xi=1 holds, where p and q are in the open interval (1,∞) with 1 (y1, . . . , yn) ∈ Rn. For (y1, . . . , yn) =(cid:18)α ≤ n Xi=1 xip! , . . . , α 1− 1 1 1− 1 i 1− 1 α n p p p Thus 1 1 q p 1 p + 1 p n Xi=1 n (cid:19) we have p n q! Xi=1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) xi(cid:12)(cid:12)(cid:12)(cid:12) 1− 1 i 1− 1 i α α 1 q . p . Xi=1 (cid:12)(cid:12)(cid:12)(cid:12) xi(cid:12)(cid:12)(cid:12)(cid:12) n Xi=1 1 p xip! ≥ n Xi=1 (cid:12)(cid:12)(cid:12)(cid:12) Choosing xi = hTix, xi , i = 1, . . . n, we get n Xi=1 1 p p n 1− 1 i hTix, xip! Tix, x(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) Xi=1 (cid:12)(cid:12)(cid:12)(cid:12) (cid:28)α ≥ ≥(cid:12)(cid:12)(cid:12)(cid:12) (cid:28)α 1 T1x, x(cid:29) ±(cid:28)α =(cid:12)(cid:12)(cid:12)(cid:12) (cid:28)(cid:18)α 1 T1 ± α 1− 1 1− 1 1− 1 p p p p 1− 1 2 T2x, x(cid:29) ± . . . ±(cid:28)α n Tn(cid:19) x, x(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) 1− 1 p 2 T2 ± . . . ± α p 1− 1 n Tnx, x(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) . Now the result follows by taking the supremum over all unit vectors in H . (cid:3) Now we give another upper bound for the powers of wp. This result has several inequalities as special cases, which considerably generalize the second inequality of (1.1). EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES 11 Theorem 3.2. Let (T1, . . . , Tn) , (A1, . . . , An) , (B1, . . . , Bn) ∈ B(H )(n) and let f and g be nonnegative continuous functions on [0,∞) satisfying f (t) g (t) = t for all t ∈ [0,∞). Then wrp p (A∗ 1T1B1, . . . , A∗ nTnBn) ≤ n 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1(cid:16)(cid:2)B∗ i f 2 (Ti) Bi(cid:3)rp for p ≥ 1 and r ≥ 1. Proof. Let x ∈ H be a unit vector. i g2 (T ∗ +(cid:2)A∗ i ) Ai(cid:3)rp(cid:17)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) hTiBix, Aixip n Xi=1 n n n i TiBix, xip hA∗ Xi=1 Xi=1 Xi=1 Xi=1 (cid:10)B∗ Xi=1 (cid:10)(cid:0)B∗ n n = ≤ = = ≤ kf (Ti) Bixkp kg (T ∗ i ) Aixkp (by Lemma 2.1(c)) p 2 hg (T ∗ i ) Aix, g (T ∗ i ) Aixi p 2 p hf (Ti) Bix, f (Ti) Bixi i f 2 (Ti) Bix, x(cid:11) 2 (cid:10)A∗ i g2 (T ∗ i f 2 (Ti) Bi(cid:1)p 2 (cid:10)(cid:0)A∗ x, x(cid:11) 1 p 2 i ) Aix, x(cid:11) i ) Ai(cid:1)p i g2 (T ∗ 1 2 x, x(cid:11) ≤ ≤ n Xi=1 (cid:18)1 2(cid:0)(cid:10)(cid:0)B∗ i f 2 (Ti) Bi(cid:1)p x, x(cid:11)r +(cid:10)(cid:0)A∗ n Xi=1 (cid:18)1 2(cid:10)(cid:0)(cid:0)B∗ i f 2 (Ti) Bi(cid:1)rp +(cid:0)A∗ i g2 (T ∗ ≤ 1 2* n Xi=1 (cid:0)(cid:0)(cid:0)B∗ i f 2 (Ti) Bi(cid:1)rp +(cid:0)A∗ i g2 (T ∗ (by Lemma 2.2(a)) 1 r x, x(cid:11)r(cid:1)(cid:19) 1 r i g2 (T ∗ (by Lemma 2.1(a)) i ) Ai(cid:1)p i ) Ai(cid:1)rp(cid:1) x, x(cid:11)(cid:19) i ) Ai(cid:1)rp(cid:1)(cid:1) x, x+! (by Lemma 2.2(a)) 1 r M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI 12 Thus n Xi=1 ≤ i TiBix, xip!r hA∗ 2* n Xi=1 (cid:0)(cid:0)B∗ 1 i f 2 (Ti) Bi(cid:1)rp +(cid:0)A∗ i g2 (T ∗ i ) Ai(cid:1)rp(cid:1)! x, x+ Now the result follows by taking the supremum over all unit vectors in H . (cid:3) Choosing A = B = I, we get. Corollary 3.3. Let (T1, . . . , Tn) ∈ B(H )(n) and let f and g be nonnegative contin- uous functions on [0,∞) satisfying f (t) g (t) = t for all t ∈ [0,∞). Then wrp p (T1, . . . , Tn) ≤ for p ≥ 1 and r ≥ 1. 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n i )(cid:1)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 (cid:0)f 2rp (Ti) + g2rp (T ∗ Letting f (t) = g (t) = t 1 2 , we get. Corollary 3.4. Let (T1, . . . , Tn) , (A1, . . . , An) , (B1, . . . , Bn) are in B(H )(n). Then wrp p (A∗ 1T1B1, . . . , A∗ nTnBn) ≤ i Ti Bi)rp + (A∗ i T ∗ n 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1(cid:16) (B∗ i Ai)rp(cid:17)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) for p ≥ 1 and r ≥ 1. Corollary 3.5. Let (A1, . . . , An) , (B1, . . . , Bn) ∈ B(H )(n). Then wrp p (A∗ 1B1, . . . , A∗ nBn) ≤ 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n Xi=1(cid:16)Bi2rp + Ai2rp(cid:17)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) for p ≥ 1 and r ≥ 1. Corollary 3.6. Let (T1, . . . , Tn) ∈ B(H )(n). Then wp p (T1, . . . , Tn) ≤ for 0 ≤ α ≤ 1, and p ≥ 1. In particular. 1 wp p (T1, . . . , Tn) ≤ n 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 (cid:16)Ti2αp + T ∗ 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) i 2(1−α)p(cid:17)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) i p)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (Tip + T ∗ Xi=1 . n EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES 13 Corollary 3.7. Let B, C ∈ B(H ). Then wp p (B, C) ≤ 1 2(cid:13)(cid:13)(cid:13) B2αp + B∗2(1−α)p + C2αp + C ∗2(1−α)p(cid:13)(cid:13)(cid:13) for 0 ≤ α ≤ 1, and p ≥ 1. In particular. wp p (B, C) ≤ 1 2 kBp + B∗p + Cp + C ∗pk . The next results are related to some different upper bounds for wp for n-tuples of operators, which have several inequalities as special cases. Proposition 3.8. Let (T1, . . . , Tn) ∈ B(H )(n). Then 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 (cid:16)Ti2α + T ∗ wp (T1, . . . , Tn) ≤ 1 n 1 p i 2(1−α)(cid:17)p(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) for 0 ≤ α ≤ 1, and p ≥ 1. Proof. By using the arithmetic-geometric mean, for any unit vector x ∈ H we have n Xi=1 hTix, xip ≤ n Xi=1 (cid:18)(cid:10)Ti2α x, x(cid:11) 1 2 DT ∗ i 2(1−α) x, xE 1 2(cid:19)p (by Lemma 2.1(b)) n 1 2p 1 2p 1 2p ≤ = ≤ i 2(1−α) x, xE(cid:17)p n Xi=1 (cid:16)(cid:10)Ti2α x, x(cid:11) +DT ∗ Xi=1 D(cid:16)Ti2α + T ∗ Xi=1 D(cid:16)Ti2α + T ∗ i 2(1−α)(cid:17) x, xEp i 2(1−α)(cid:17)p x, xE n . (by Lemma 2.2(a)) Now the result follows by taking the supremum over all unit vectors in H . (cid:3) Proposition 3.9. Let (T1, . . . , Tn) ∈ B(H )(n). Then n wp (T1, . . . , Tn) ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 for 0 ≤ α ≤ 1, and p ≥ 2. 1 p (αTip + (1 − α)T ∗ i p)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 14 M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI Proof. For every unit vector x ∈ H , we have n Xi=1 p 2 p 2 n n n = ≤ ≤ hTix, xip Xi=1 (cid:0)hTix, xi2(cid:1) Xi=1 (cid:16)(cid:10)Ti2α x, x(cid:11)DT ∗ i 2(1−α) x, xE(cid:17) i (1−α)p x, xE hTiαp x, xiDT ∗ Xi=1 i p x, xi(1−α) Xi=1 hTip x, xiα hT ∗ Xi=1(cid:16)α hTip x, xi + (1 − α) hT ∗ i p(cid:17)x, xE Xi=1 D(cid:16)αTip + (1 − α)T ∗ ≤ =* n i p)! x, x+ . Xi=1 (αTip + (1 − α)T ∗ ≤ ≤ n n n (by Lemma 2.1(b)) (by Lemma 2.2(a)) (by Lemma 2.2(b)) i p x, xi(cid:17)(by Lemma 2.1(a)) Now the result follows by taking the supremum over all unit vectors in H . (cid:3) Remark 3.10. As special cases, (1) For α = 1 2 , we have (Tip + T ∗ (2) For B, C ∈ B(H ), 0 ≤ α ≤ 1, and p ≥ 1, we have wp p (T1, . . . , Tn) ≤ Xi=1 n 1 2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) . i p)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) p (B, C) ≤ kαBp + (1 − α)B∗p + αCp + (1 − α)C ∗pk . wp In particular, wp p (B, C) ≤ 1 2 kBp + B∗p + Cp + C ∗pk . The next result reads as follows. Proposition 3.11. Let (T1, . . . , Tn) ∈ B(H )(n), 0 ≤ α ≤ 1, r ≥ 1 and p ≥ 1. Then wp (T1, . . . , Tn) ≤ n Xi=1 (cid:13)(cid:13)αTi2r + (1 − α) T ∗ i 2r(cid:13)(cid:13) p 2r! 1 p . EXTENSION OF EUCLIDEAN OPERATOR RADIUS INEQUALITIES 15 Proof. Let x ∈ H be a unit vector. n Xi=1 = ≤ ≤ ≤ ≤ ≤ p 2 n n n n p 2 hTix, xip Xi=1 (cid:0)hTix, xi2(cid:1) i 2(1−α) x, xE(cid:17) Xi=1 (cid:16)(cid:10)Ti2α x, x(cid:11)DT ∗ i 2 x, x(cid:11)(1−α)(cid:17) Xi=1 (cid:16)(cid:10)Ti2 x, x(cid:11)α(cid:10)T ∗ Xi=1 (cid:16)α(cid:10)Ti2 x, x(cid:11)r + (1 − α)(cid:10)T ∗ Xi=1(cid:16)α(cid:10)Ti2r x, x(cid:11) + (1 − α)(cid:10)T ∗ Xi=1 (cid:10)(cid:0)αTi2r + (1 − α)T ∗ i 2r(cid:1) x, x(cid:11) n n p 2 i 2 x, x(cid:11)r(cid:17) i 2r x, x(cid:11)(cid:17) p 2r . (by Lemma 2.1(b)) (by Lemma 2.2(b)) p 2r (by Lemma 2.1(a)) p 2r (by Lemma 2.2(a)) Now the result follows by taking the supremum over all unit vectors in H . (cid:3) Remark 3.12. Some special cases can be stated as follows: (1) For α = 1 2 , we have wp (T1, . . . , Tn) ≤ 1 2 p 2r n Xi=1 (cid:13)(cid:13)Ti2r + T ∗ i 2r(cid:13)(cid:13) 1 p . p 2r! (2) For B, C ∈ B(H ), 0 ≤ α ≤ 1, and p ≥ 1, we have wp (B, C) ≤(cid:16)(cid:13)(cid:13)αB2r + (1 − α)B∗2r(cid:13)(cid:13) In particular, p 2r +(cid:13)(cid:13)αC2r + (1 − α)C ∗2r(cid:13)(cid:13) 2r(cid:17) 2r +(cid:13)(cid:13)C2r + C ∗2r(cid:13)(cid:13) p p 1 p . p 2r(cid:17) 1 p . wp (B, C) ≤ 1 2 1 2r (cid:16)(cid:13)(cid:13)B2r + B∗2r(cid:13)(cid:13) 16 M.S. MOSLEHIAN, M. SATTARI, K. SHEBRAWI References 1. S. S. Dragomir, Some inequalities for the Euclidean operator radius of two operators in Hilbert Spaces, Linear Algebra and Its Appl, 419 (1) (2006), 256 -- 264. 2. K. E. Gustufson and D. K. M. Rao, Numerical Range, Springer-Verlag, New york, 1997. 3. O. Hirzallah, F. Kittaneh and K. Shebrawi, Numerical radius inequalities for certain 2 × 2 operator matrices, Integral Equations Operator Theory 71 (2011), no. 1, 129 -- 147. 4. O. Hirzallah and F. Kittaneh, Numerical radius inequalities for several operators, Math. Scand. 114 (2014), no. 1, 110-119. 5. J. Pecari´c, T. Furuta, J. Mi´ci´c Hot and Y. Seo, Mond-Pecari´c Method in Operator Inequalities, Element, Zagreb, 2005. 6. F. Kittaneh, Notes on some inequalities for Hilbert space operators Publ. Res. Inst. Math. Sci. 24 (1988), no. 2, 283 -- 293. 7. D. S. Mitrinovi´c, J. E. Pecari´c and A. M. Fink, Classical and new inequalities in analysis, Kluwer Academic Publishers, 1993. 8. J. I. Fujii, M. Fujii, M. S. Moslehian, J. E. Pecari´c and Y. Seo, Reverses Cauchy -- Schwarz type inequalities in pre-inner product C∗-modules, Hokkaido Math. J. 40 (2011) , 1 -- 17. 9. G. Popescu, Unitary invariants in multivariable operator theory, Mem. Amer. Math. Soc., Vol. 200, no 941, 2009. 10. A. Salemi and A. Sheikhhosseini, Matrix Young numerical radius inequalities, Math. Inequal. Appl. 16 (2013), no. 3, 783 -- 791. 11. T. Yamazaki, On upper and lower bounds for the numerical radius and an equality condition, Studia Math. 178 (2007), no. 1, 83 -- 89. 12. T. Yamazaki, On numerical range of the Aluthge transformation, Linear Algebra Appl. 341 (2002), 111 -- 117. 1 Department of Pure Mathematics, Center Of Excellence in Analysis on Al- gebraic Structures (CEAAS), Ferdowsi University of Mashhad, P. O. Box 1159, Mashhad 91775, Iran E-mail address: [email protected] E-mail address: [email protected] 2 Department of Applied Sciences, Al-Balqa' Applied University Salt, Jordan; Department of Mathematics, College of Science, Qassim University, Qassim, Saudi Arabia E-mail address: [email protected]
1009.4351
1
1009
2010-09-22T13:04:17
Constructing pairs of dual bandlimited frame wavelets in $L^2(\mathbb{R}^n)$
[ "math.FA" ]
Given a real, expansive dilation matrix we prove that any bandlimited function $\psi \in L^2(\mathbb{R}^n)$, for which the dilations of its Fourier transform form a partition of unity, generates a wavelet frame for certain translation lattices. Moreover, there exists a dual wavelet frame generated by a finite linear combination of dilations of $\psi$ with explicitly given coefficients. The result allows a simple construction procedure for pairs of dual wavelet frames whose generators have compact support in the Fourier domain and desired time localization. The construction relies on a technical condition on $\psi$, and we exhibit a general class of function satisfying this condition.
math.FA
math
CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) JAKOB LEMVIG Abstract. Given a real, expansive dilation matrix we prove that any bandlimited function ψ ∈ L2(Rn), for which the dilations of its Fourier transform form a partition of unity, generates a wavelet frame for certain translation lattices. Moreover, there exists a dual wavelet frame generated by a finite linear combination of dilations of ψ with explicitly given coefficients. The result allows a simple construction procedure for pairs of dual wavelet frames whose generators have compact support in the Fourier domain and desired time localization. The construction relies on a technical condition on ψ, and we exhibit a general class of function satisfying this condition. 0 1 0 2 p e S 2 2 ] . A F h t a m [ 1 v 1 5 3 4 . 9 0 0 1 : v i X r a 1. Introduction For A ∈ GLn(R) and y ∈ Rn, we define the dilation operator on L2(Rn) by DAf (x) = det A1/2 f (Ax) and the translation operator by Tyf (x) = f (x− y). Given a n× n real, expansive matrix A and a lattice of the form Γ = P Zn for P ∈ GLn(R), we consider wavelet systems of the form where the Fourier transform of ψ has compact support. Our aim is, for any given real, expansive dilation matrix A, to construct wavelet frames with good regularity properties and with a dual frame generator of the form {DAj Tγψ}j∈Z,γ∈Γ , (1) φ = cjDAj ψ bXj=a for some explicitly given coefficients cj ∈ C and a, b ∈ Z. This will generalize and extend the one-dimensional results on constructions of dual wavelet frames in [15, 18] to higher dimensions. The extension is non-trivial since it is unclear how to determine the translation lattice Γ and how to control the support of the generators in the Fourier domain. This will be done by considering suitable norms in Rn and non-overlapping packing of ellipsoids in lattice arrangements. The construction of redundant wavelet representations in higher dimensions is usually based on extension principles [7, 8, 10 -- 14, 16, 17]. By making use of extension principles one is restricted to considering expansive dilations A with integer coefficients. Our constructions work for any real, expansive dilation. Moreover, in the extension principle Date: November 23, 2018. 2000 Mathematics Subject Classification. 42C40. Key words and phrases. real and expansive dilation matrix, bandlimited wavelets, dual frames, non- tight frames, partition of unity. 1 2 JAKOB LEMVIG the number of generators often increases with the smoothness of the generators. We will construct pairs of dual wavelet frames generated by one smooth function with good time localization. It is a well-known fact that a wavelet frame need not have dual frames with wavelet structure. In [20] frame wavelets with compact support and explicit analytic form are constructed for real dilation matrices. However, no dual frames are presented for these wavelet frames. This can potentially be a problem because it might be difficult or even impossible to find a dual frame with wavelet structure. Since we exhibit pairs of dual wavelet frames, this issue is avoided. The principal importance of having a dual generator of the form (1) is that it will inherit properties from ψ preserved by dilation and linearity, e.g. vanishing moments, good time localization and regularity properties. For a more complete account of such matters we refer to [15]. In the rest of this introduction we review basic definitions. A frame for a separable Hilbert space H is a countable collection of vectors {fj}j∈J for which there are constants 0 < C1 ≤ C2 < ∞ such that If the upper bound holds in the above inequality, then {fj} is said to be a Bessel sequence with Bessel constant C2. For a Bessel sequence {fj} we define the frame operator by for all f ∈ H. C1 kfk2 ≤Xj∈J(cid:12)(cid:12)hf, fji(cid:12)(cid:12)2 ≤ C2 kfk2 Sf =Xj∈J S : H → H, hf, fjifj. This operator is bounded, invertible, and positive. A frame {fj} is said to be tight if we can choose C1 = C2; this is equivalent to S = C1I where I is the identity operator. Two Bessel sequences {fj} and {gj} are said to be dual frames if f =Xj∈J hf, gjifj ∀f ∈ H. It can be shown that two such Bessel sequences are indeed frames. Given a frame {fj}, at least one dual always exists; it is called the canonical dual and is given by {S−1fj}. Only a frame, which is not a basis, has several duals. For f ∈ L1(Rn) the Fourier transform is defined by f (ξ) =RRn f (x) e−2πihξ,xi dx with Sets in Rn are, in general, considered equal if they are equal up to sets of measure zero. The boundary of a set E is denoted by ∂E, the interior by E◦, and the closure by E. Let B ∈ GLn(R). A multiplicative tiling set E for {Bj : j ∈ Z} is a subset of positive measure such that the usual extension to L2(Rn). (cid:12)(cid:12)(cid:12)Rn \[j∈Z Bj(E)(cid:12)(cid:12)(cid:12) = 0 and (cid:12)(cid:12)Bj(E) ∩ Bl(E)(cid:12)(cid:12) = 0 for l 6= j. In this case we say that {Bj(E) : j ∈ Z} is an almost everywhere partition of Rn, or that it tiles Rn. A multiplicative tiling set E is bounded if E is a bounded set and 0 /∈ E. By B-dilative periodicity of a function f : Rn → C we understand f (x) = f (Bx) for (2) CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 3 a.e. x ∈ Rn, and by a B-dilative partition of unity we understand Pj∈Z f (Bjx) = 1; note that the functions in the "partition of unity" are not assumed to be non-negative, but can take any real or complex value. A (full-rank) lattice Γ in Rn is a point set of the form Γ = P Zn for some P ∈ GLn(R). The determinant of Γ is d(Γ) = det P; note that the generating matrix P is not unique, and that d(Γ) is independent of the particular choice of P . 2. The general form of the construction procedure Fix the dimension n ∈ N. We let A ∈ GLn(R) be expansive, i.e., all eigenvalues of A have absolute value greater than one, and denote the transpose matrix by B = At. For any such dilation A, we want to construct a pair of functions that generate dual wavelet frames for some translation lattice. Our construction is based on the following result which is a consequence of the characterizing equations for dual wavelet frames by Chui, Czaja, Maggioni, and Weiss [6, Theorem 4]. Theorem 2.1. Let A ∈ GLn(R) be expansive, let Γ be a lattice in Rn, and let Ψ = {ψ1, . . . , ψL}, Ψ = { ψ1, . . . , ψL} ⊂ L2(Rn). Suppose that the two wavelet systems ψl : j ∈ Z, γ ∈ Γ, l = 1, . . . , L} form {DAj Tγψl : j ∈ Z, γ ∈ Γ, l = 1, . . . , L} and {DAj Tγ Bessel families. Then {DAj Tγψl} and {DAj Tγ ψl} will be dual frames if the following conditions hold ψl(Bjξ) ψl(Bjξ) = d(Γ) a.e. ξ ∈ Rn, (3) (4) LXl=1Xj∈Z LXl=1 ψl(ξ) ψl(ξ + γ) = 0 a.e. ξ ∈ Rn for γ ∈ Γ∗ \ {0}. Proof. By ξ = Bjω for j ∈ Z, condition (4) becomes (5) LXl=1 ψl(Bjω) ψl(Bjω + γ) = 0 a.e. ω ∈ Rn for γ ∈ Γ∗ \ {0}. We use the notation as in [6], thus Λ(A, Γ) = {α ∈ Rn : ∃(j, γ) ∈ Z × Γ∗ : α = B−jγ} and IA,Γ(α) = {(j, γ) ∈ Z × Γ∗ : α = B−jγ}. Since IA,Γ(α) ⊂ Z × (Γ∗ \ {0}) for any α ∈ Λ(A, Γ) \ {0}, equation (5) yields 1 d(Γ) X(j,γ)∈IA,Γ(α) LXl=1 ψl(Bjω) ψl(Bj(ω + B−jγ)) = 0 a.e. ω ∈ Rn for α 6= 0. By IA,Γ(0) = Z × {0}, we can rewrite (3) as 1 d(Γ) X(j,γ)∈IA,Γ(0) LXl=1 ψl(Bjω) ψl(Bj(ω + B−jγ)) = 1 a.e. ω ∈ Rn, 4 JAKOB LEMVIG (7) C1 = 1 d(Γ) using that B−jγ = 0 for all j ∈ Z. Gathering the two equations displayed above yields 1 d(Γ) X(j,γ)∈IA,Γ(α) LXl=1 ψl(Bjω) ψl(Bj(ω + B−jγ)) = δα,0 a.e. ω ∈ Rn, for all α ∈ Λ(A, Γ). The conclusion follows now from [6, Theorem 4]. The following result, Lemma 2.2, gives a sufficient condition for a wavelet system to form a Bessel sequence; it is an extension of [3, Theorem 11.2.3] from L2(R) to L2(Rn). Lemma 2.2. Let A ∈ GLn(R) be expansive, Γ a lattice in Rn, and φ ∈ L2(Rn). Suppose that, for some set M ⊂ Rn satisfying ∪l∈ZBl(M) = Rn, (6) 1 C2 = (cid:3) Then the wavelet system {DAj Tγφ}j∈Z,γ∈Γ is a Bessel sequence with bound C2. Further, if also d(Γ) sup ξ∈MXj∈Z Xγ∈Γ∗(cid:12)(cid:12)(cid:12) φ(Bjξ) φ(Bjξ + γ)(cid:12)(cid:12)(cid:12) < ∞. −Xj∈Z Xγ∈Γ∗\{0}(cid:12)(cid:12)(cid:12) φ(Bjξ) φ(Bjξ + γ)(cid:12)(cid:12)(cid:12) 2 inf ξ∈MXj∈Z(cid:12)(cid:12)(cid:12) φ(Bjξ)(cid:12)(cid:12)(cid:12)  > 0, holds, then {DAj Tγφ}j∈Z,γ∈Γ is a frame for L2(Rn) with frame bounds C1 and C2. Proof. The statement follows directly by applying Theorem 3.1 in [5] on generalized shift invariant systems to wavelet systems. In the general result for generalized shift invariant systems [5, Theorem 3.1], the supremum/infimum is taken over Rn, but because of the B-dilative periodicity of the series in (6) and (7) for wavelet systems, it suffices to take the supremum/infimum over a set M ⊂ Rn that has the property that ∪l∈ZBl(M) = Rn up to sets of measure zero. (cid:3) Theorem 2.1 and Lemma 2.2 are all we need to prove the following result on pairs of dual wavelet frames. Theorem 2.3. Let A ∈ GLn(R) be expansive and ψ ∈ L2(Rn). Suppose that ψ is a bounded, real-valued function with supp ψ ⊂ ∪d j=0B−j(E) for some d ∈ N0 and some bounded multiplicative tiling set E for {Bj : j ∈ Z}, and that for a.e. ξ ∈ Rn. (8) ψ(Bjξ) = 1 Xj∈Z Let bj ∈ C for j = −d, . . . , d and let m = max{j : bj 6= 0} and m = − min{j : bj 6= 0}. Take a lattice Γ in Rn such that (9) (10) and define the function φ by (cid:16) d[j=0 B−j(E) + γ(cid:17) ∩ m+d[j=−m mXj=−m φ(x) = d(Γ) B−j(E) = ∅ for all γ ∈ Γ∗ \ {0}, bj det A−j ψ(A−jx) for x ∈ Rn. CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 5 If b0 = 1 and bj + b−j = 2 for j = 1, 2, . . . , d, then the functions ψ and φ generate dual frames {DAj Tγψ}j∈Z,γ∈Γ and {DAj Tγφ}j∈Z,γ∈Γ for L2(Rn). Proof. On the Fourier side, the definition in (10) becomes φ(ξ) = d(Γ) bj ψ(Bjξ). mXj=−m Since ψ by assumption is compactly supported in a "ringlike" structure bounded away from the origin, this will also be the case for φ. This property implies that ψ and φ will generate wavelet Bessel sequences. The details are as follows. The support of ψ and φ is (11) and, (13) Note that 0 ≤ m, m ≤ d. The sets {Bj(E) : j ∈ Z} tiles Rn, whereby we see that (12) B−j(E). B−j(E), d[j=0 supp ψ ⊂ supp φ ⊂ m+d[j=−m (cid:12)(cid:12)(cid:12) supp ψ(Bj·) ∩ B−d(E)(cid:12)(cid:12)(cid:12) = 0 for j < 0 and j > d, (cid:12)(cid:12)(cid:12) supp φ(Bj·) ∩ B−d(E)(cid:12)(cid:12)(cid:12) = 0 for j < −m and j > m + d. Since m, m ≥ 0, condition (9) implies that ψ(Bjξ) ψ(Bjξ + γ) = 0 for j ≥ 0 and γ ∈ Γ∗ \ {0}. Therefore, using (12), we find that Xj∈Z Xγ∈Γ∗(cid:12)(cid:12)(cid:12) ψ(Bjξ) ψ(Bjξ + γ)(cid:12)(cid:12)(cid:12) = dXj=0(cid:16) ψ(Bjξ)(cid:17)2 < ∞ for ξ ∈ B−d(E). An application of Lemma 2.2 with M = B−d(E) shows that ψ generates a Bessel sequence. Similar calculations using (13) will show that φ generates a Bessel sequence; in this case the sum over γ ∈ Γ∗ will be finite, but it will in general have more than one nonzero term. To conclude that ψ and φ generate dual wavelet frames we will show that condi- tions (3) and (4) in Theorem 2.1 hold. By B-dilation periodicity of the sum in condition (3), it is sufficient to verify this condition on B−d(E). For ξ ∈ B−d(E) we have by (12), 1 d(Γ)Xj∈Z ψ(Bjξ) φ(Bjξ) = 1 d(Γ) ψ(Bjξ) φ(Bjξ) dXj=0 = ψ(ξ)hb0 ψ(ξ) + b1 ψ(Bξ) + · · · + bd ψ(Bdξ)i + ψ(Bξ)hb−1 ψ(ξ) + b0 ψ(Bξ) + · · · + bd−1 ψ(Bdξ)i + · · · + ψ(Bdξ)hb−d ψ(Bdξ)i , ψ(ξ) + · · · + b−1 ψ(Bd−1ξ) + b0 6 JAKOB LEMVIG and further, by an expansion of these terms, = bl−j ψ(Bjξ) ψ(Blξ) dXj,l=0 dXj=0 = b0 ψ(Bjξ)2 + dXj,l=0 j>l (bj−l + bl−j) ψ(Bjξ) ψ(Blξ). Using that b0 = 1 and bj−l + bl−j = 2 for j 6= l and j, l = 0, . . . , d, we arrive at 1 d(Γ)Xj∈Z ψ(Bjξ) φ(Bjξ) = 2 ψ(Bjξ) ψ(Blξ) dXj=0 =(cid:18) dXj=0 ψ(Bjξ)2 + dXj,l=0 ψ(Bjξ)(cid:19)2 =(cid:18)Xj∈Z j>l ψ(Bjξ)(cid:19)2 = 1, exhibiting that ψ and φ satisfy condition (3). By (11) we see that condition (9) implies that the functions φ and ψ(· + γ) will have disjoint support for γ ∈ Γ∗ \ {0}, hence (4) is satisfied. Remark 1. The use of the parameters bj in the definition of the dual generator together with the condition b−j + bj = 2 was first seen in the work of Christensen and Kim [4] on pairs of dual Gabor frames. (cid:3) We can restate Theorem 2.3 for wavelet systems with standard translation lattice Zn and dilation eA = P −1AP , where P ∈ GLn(R) is so that Γ = P Zn. The result follows directly by an application of the relations D eAj DP = DP DAj for j ∈ Z and DP TP k = TkDP for k ∈ Zn, and the fact that DP is unitary as an operator on L2(Rn). Corollary 2.4. Suppose ψ, {bj}, A and Γ are as in Theorem 2.3. Let P ∈ GLn(R) be such that Γ = P Zn, and let eA = P −1AP . Then the functions ψ = DP ψ and φ = DP φ, where φ is defined in (10), generate dual frames {D eAj Tk ψ}j∈Z,k∈Zn and {D eAj Tk φ}j∈Z,k∈Zn for L2(Rn). The following Example 1 is an application of Theorem 2.3 in L2(R2) for the quincunx matrix. In particular, we construct a partition of unity of the form (8) for the quincunx matrix. Example 1. The quincunx matrix is defined as A =(cid:18)1 −1 1 (cid:19) , 1 and its action on R2 corresponds to a counter clockwise rotation of 45 degrees and a dilation by √2I2×2. Define the tent shaped, piecewise linear function g by CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 7 x2 1.00 0.75 0.50 0.25 J5 J3 J1 J4 J2 0.25 0.50 0.75 1.00 x1 Figure 1. Sketch of the triangular domains Ji, i = 1, 2, 3, 4, 5. g(x1, x2) = −1 + 2x1 + 2x2, 2x2, 2x1, 2 − 2x1, 2 − 2x2, 0 for (x1, x2) ∈ J1, for (x1, x2) ∈ J2, for (x1, x2) ∈ J3, for (x1, x2) ∈ J4, for (x1, x2) ∈ J5, otherwise,  where the sets Ji are the triangular domains sketched in Figure 1. Note that the value at "the top of the tent" is g(1/2, 1/2) = 1. Define ψ as a mirroring of g in the x1 axis and the x2 axis: ψ(ξ1, ξ2) = g(ξ1, ξ2) g(ξ1,−ξ2) g(−ξ1, ξ2) g(−ξ1,−ξ2) for (ξ1, ξ2) ∈ [0,∞) × [0,∞) , for (ξ1, ξ2) ∈ [0,∞) × (−∞, 0) , for (ξ1, ξ2) ∈ (−∞, 0) × [0,∞) , for (ξ1, ξ2) ∈ (−∞, 0) × (−∞, 0) . Since the transpose B of the quincunx matrix also corresponds to a rotation of 45 degrees (but clockwise) and a dilation by √2I2×2, we see thatPj∈Z We are now ready to apply Theorem 2.3 with E = [−1, 1]2\ B−1([−1, 1]2) = [−1, 1]2\ I1 and d = 2; the set E is the union of the domians J4 and J5 and their mirrored versions. We choose b−2 = b−1 = 0 and b1 = b2 = 2d(Γ), hence m = 0 and m = 2. Therefore, ψ(Bjξ) = 1. d[j=0 m+d[j=−m B−j(E), B−j(E) ⊂ [−1, 1]2 , that shows that we can take Γ∗ = 2Z2 or Γ = 1/2Z2, since ([−1, 1]2 + γ) ∩ [−1, 1]2 = ∅ whenever 0 6= γ ∈ 2Z2. Defining the dual generator according to (16) yields (14) φ(x) = (1/4)ψ(x) + (1/4)ψ(A−1x) + (1/8)ψ(A−2x); 8 JAKOB LEMVIG using that d(Γ) = 1/4, and we remark that φ is a piecewise linear function since this is the case for ψ. The conclusion from Theorem 2.3 is that ψ and φ generate dual frames {DAj Tk/2ψ}j,k∈Z and {DAj Tk/2φ}j,k∈Z for L2(R2). finite sums on E; for {DAj Tk/2ψ} one finds C1 = 4/3 and C2 = 4. The frame bounds can be found using Lemma 2.2 since the series (6) and (7) are When the result on constructing pairs of dual wavelet frames is written in the gen- erality of Theorem 2.3, it is not always clear how to choose the set E and the lattice Γ. In Example 1 we showed how this can be done for the quincunx dilation matrix and constructed a pair of dual frame wavelets. In Section 3 and Theorem 3.3 we specify how to choose E and Γ for general dilations. The issue of exhibiting functions ψ satisfying the condition (8) is addressed in Section 4. In one dimension, however, it is straightforward to make good choices of E and Γ as is seen by the following corollary of Theorem 2.3. The corollary unifies the construction procedures in Theorem 2 and Proposition 1 from [15] in a general procedure. Corollary 2.5. Let d ∈ N0, a > 1, and ψ ∈ L2(R). Suppose that ψ is a bounded, real-valued function with supp ψ ⊂ [−ac,−ac−d−1] ∪ [ac−d−1, ac] for some c ∈ Z, and that (15) Xj∈Z ψ(ajξ) = 1 for a.e. ξ ∈ R. Let bj ∈ C for j = −d, . . . , d, let m = − min{j : {bj 6= 0}}, and define the function φ by (16) φ(x) = bja−jψ(a−jx) for x ∈ R. dXj=−m Let b ∈ (0, a−c(1 + am)−1]. If b0 = b and bj + b−j = 2b for j = 1, 2, . . . , d, then ψ and φ generate dual frames {Daj Tbkψ}j,k∈Z and {Daj Tbkφ}j,k∈Z for L2(R). Proof. In Theorem 2.3 for n = 1 and A = a we take E = [−ac,−ac−1]∪ [ac−1, ac] as the multiplicative tiling set for {aj : j ∈ Z}. The assumption on the support of ψ becomes supp ψ ⊂ Moreover, since d[j=0 a−j(E) = [−ac,−ac−d−1] ∪ [ac−d−1, ac]. d[j=0 a−j(E) ⊂ [−ac, ac] , 2d[j=−m a−j(E) ⊂(cid:2)−ac+m, ac+m(cid:3) , and ([−ac, ac] + γ) ∩(cid:2)−ac+m, ac+m(cid:3) = ∅ for γ ≥ ac + ac+m = ac(1 + am), the choice Γ∗ = b−1Z for b−1 ≥ ac(1 + am) satisfies equation (9). This corresponds to Γ = bZ for 0 < b ≤ a−c(1 + am)−1. (cid:3) CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 9 The assumptions in Corollary 2.5 imply that m ∈ {0, 1, . . . , d}; we note that in case m = 0, the corollary reduces to [15, Theorem 2]. 3. A special case of the construction procedure We aim for a more automated construction procedure than what we have from The- orem 2.3, in particular, we therefore need to deal with good ways of choosing E and Γ. The basic idea in this automation process will be to choose E as a dilation of the difference between I∗ and B−1(I∗), where I∗ is the unit ball in a norm in which the matrix B = At is expanding "in all directions"; we will make this statement precise in Section 3.1. This idea is instrumental in the proof of Theorem 3.3. 3.1. Some results on expansive matrices. We need the following well-known equiv- alent conditions for a (non-singular) matrix being expansive. Proposition 3.1. For B ∈ GLn(R) the following assertions are equivalent: (i) B is expansive, i.e., all eigenvalues λi of B satisfy λi > 1. (ii) For any norm · on Rn there are constants λ > 1 and c ≥ 1 such that Bjx ≥ 1/cλj x for all j ∈ N0, for any x ∈ Rn. (iii) There is a Hermitian norm · ∗ Bjx∗ ≥ λj x∗ on Rn and a constant λ > 1 such that for all j ∈ N0, for any x ∈ Rn. λ > 1. (iv) E ⊂ λE ⊂ BE for some ellipsoid E = {x ∈ Rn : P x ≤ 1}, P ∈ GLn(R), and By Proposition 3.1 we have that for a given expansive matrix B, there exists a scalar product with the induced norm · ∗ so that Bx∗ ≥ λx∗ for x ∈ Rn, holds for some λ > 1. We say that · ∗ is a norm associated with the expansive matrix B. Note that such a norm is not unique; we will follow the construction as in the proof of [2, Lemma 2.2], so let c and λ be as in (ii) in Proposition 3.1 for the standard Euclidean norm with 1 < λ < λi for i = 1, . . . , n, where λi are the eigenvalues of B. For k ∈ N satisfying k > 2 ln c/ ln λ we introduce the symmetric, positive definite matrix K ∈ GLn(R): K = I + (B−1)tB−1 + · · · + (B−k)tB−k. (17) = xtKy. It might not The scalar product associated with B is then defined by hx, yi∗ be effortless to estimate c and λ for some given B, but it is obvious that we just need to pick k ∈ N such that BtKB − λ2K becomes positive semi-definite for some λ > 1 since this corresponds to hKBx, Bxi ≥ λ2 hKx, xi, that is, Bx2 = K 1/2· associated with We let I∗ denote the unit ball in the Hermitian norm · ∗ B, i.e., for all x ∈ Rn. ∗ ≥ λ2 x2 ∗ (18) I∗ = {x ∈ Rn : x∗ ≤ 1} =(cid:8)x ∈ Rn : K 1/2x ≤ 1(cid:9) =(cid:8)x ∈ Rn : xtKx ≤ 1(cid:9) , 10 JAKOB LEMVIG and we let O∗ denote the annulus O∗ = I∗ \ B−1(I∗). λ > 1, norm, i.e., I∗ ⊂ λI∗ ⊂ B(I∗), The ringlike structure of O∗ is guaranteed by the fact that B is expanding in all direc- tions in the · ∗ (19) which is (iv) in Proposition 3.1. We note that by an orthogonal substitution I∗ takes the form {x ∈ Rn : µ1x2 n ≤ 1}, where µi are the positive eigenvalues of K and x = Qx with the ith column of Q ∈ O(n) comprising of the ith eigenvector of K. The annulus O∗ is a bounded multiplicative tiling set for {Bj : j ∈ Z}. This is a consequence of the following result. Lemma 3.2. Let B ∈ GLn(R) be an expansive matrix. For x 6= 0 there is a unique j ∈ Z so that Bjx ∈ O∗; that is, (20) 1 + · · · + µn x2 with disjoint union. Rn \ {0} =[j∈Z Bj(O∗) Proof. From equation (19) we know that {Bl(I∗)}l∈Z is a nested sequence of subsets of Rn, thus are disjoint sets. Since B−jx∗ ≤ λ−j x∗ also have Bl(I∗) \ Bl−1(I∗) = Bl(O∗), l ∈ Z, and Bjx∗ ≥ λj x∗ for j ≥ 0 and λ > 1, we l[m=−l+1 ⊃(cid:8)x ∈ Rn : λ−l x∗ ≤ 1 and λl x∗ Bm(O∗) = Bl(I∗) \ B−l(I∗) =(cid:8)x ∈ Rn : B−lx∗ ≤ 1 and Blx∗ > 1(cid:9) > 1(cid:9) =(cid:8)x ∈ Rn : λ−l < x∗ ≤ λl(cid:9) . A =(cid:18)3 −3 0 (cid:19) . Taking the limit l → ∞ we get (20). Example 2. Let the following dilation matrix be given (21) 1 Here we are interested in the transpose matrix B = At with eigenvalues µ1,2 = 3/2 ± i√3/2, hence B is an expansive matrix with µ1,2 = √3 > 1. The dilation matrix B is not expanding in the standard norm · 2 in Rn, i.e., I2 6⊂ B(I2), as shown by Figure 2. In order to have B expanding the unit ball we need to use the Hermitian norm from (iii) in Proposition 3.1 associated with B. In (17) we take k = 2 so that the real, symmetric, positive definite matrix K is (cid:3) K = I + (B−1)tB−1 + (B−2)tB−2 =(cid:18)28/9 16/9 8/3(cid:19) , 16/9 and let hx, yi∗ positive definite for λ = 1.03 and thus := xtKy. The choice k = 2 suffices since it makes BtKB − λ2K semi- Bx∗ ≥ λx∗ , x ∈ R2, CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 11 4 2 K4 K3 K2 K1 0 1 2 3 4 K2 K4 Figure 2. Boundaries of the sets I2, B(I2), B2(I2), and B3(I2) marked by solid, long dashed, dashed, and dotted lines, respectively. Note that I2 \ B(I2) is non-empty, and even I2 \ B2(I2) is non-empty. holds for λ = 1.03. Figure 3 and 4 illustrate that B indeed expands the Hermitian norm unit ball I∗ in all directions. We also remark that the Hermitian norm with k = 1 will not make the 6 4 2 K4 K3 K2 K1 0 1 2 3 4 K2 K4 K6 Figure 3. The unit ball I∗ in the Hermitian norm · ∗ associated with B and its dilations B(I∗), B2(I∗), B3(I∗). Only the boundaries are marked. dilation matrix B expanding in Rn; in this case we have a situation similar to Figure 2. 12 JAKOB LEMVIG 1.0 0.5 K1.0 K0.5 0 0.5 1.0 K0.5 K1.0 Figure 4. A zoom of Figure 3. Boundaries of the sets I∗, B(I∗), B2(I∗), and B3(I∗) marked by solid, long dashed, dashed, and dotted lines, re- spectively. 3.2. A crude lattice choice. Let us consider the setup in Theorem 2.3 with the set E = Bc(O∗) for some c ∈ Z, where the norm · ∗ = K 1/2· is associated with B. Let µ of the ellipsoid I∗, i.e., ℓ = maxx∈I∗ x2. Then we can take any lattice Γ = P Zn, where P is a non-singular matrix satisfying be the smallest eigenvalue of K such that ℓ =p1/µ is the largest semi-principal axis (22) kPk2 ≤ 1 ℓkAck2 (1 + kAmk2) , as our translation lattice in Theorem 2.3. To see this, recall that we are looking for a lattice Γ∗ such that, for γ ∈ Γ∗ \ {0}, (23) For our choice of E we find that supp φ ⊂ Bc+m(I∗) and supp ψ ⊂ Bc(I∗). Since supp φ ∩ supp ψ(· ± γ) = ∅. (cid:12)(cid:12)Bc+mx(cid:12)(cid:12)2 ≤(cid:13)(cid:13)Bc+m(cid:13)(cid:13)2 x2 ≤(cid:13)(cid:13)Bc+m(cid:13)(cid:13)2 ℓ for any x ∈ I∗, and similar for Bcx, we have the situation in (23) whenever γ2 ≥ ℓ(kAck2 +kAc+mk2). Here we have used that for the 2-norm kAk2 = kBk2. For z ∈ Zn we have z2 ≤ kP tk2 (P t)−1z2 = kPk2 (P t)−1z2, therefore, by z2 ≥ 1 for z 6= 0, we have 1 kPk2 (cid:12)(cid:12)(P t)−1z(cid:12)(cid:12)2 ≥ for z ∈ Z \ {0}. CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 13 Now, by assuming that P satisfies (22), we have γ2 = (P t)−1z2 ≥ 1/kPk2 ≥ ℓkAck2 (1 + kAmk2) ≥ l(kAck2 +(cid:13)(cid:13)Ac+m(cid:13)(cid:13)2) for 0 6= γ = (P t)−1z ∈ (P t)−1Zn = Γ∗, hence the claim follows. A lattice choice based on (22) can be rather crude, and produces consequently a wavelet system with unnecessarily many translates. From equation (22) it is obvious that any lattice Γ = P Zn with kPk sufficiently small will work as translation lattice for our pair of generators ψ and φ. Hence, the challenging part is to find a sparse translation lattice whereby we understand a lattice Γ with large determinant d(Γ) := det P. In the dual lattice system this corresponds to a dense lattice Γ∗ with small volume d(Γ∗) of the fundamental parallelotope IΓ∗ since d(Γ)d(Γ∗) = 1. In Theorem 3.3 in the next section we make a better choice of the translation lattice compared to what we have from (22). Using a crude lattice approach as above, we can easily transform the translation lattice to the integer lattice if we allow multiple generators. We pick a matrix P that satisfies condition (22) and whose inverse is integer valued, i.e., Q := P −1 ∈ GLn(Z). The conclusion from Theorem 2.3 is that {DAj TQ−1kψ}j∈Z,k∈Zn and {DAj TQ−1kφ}j∈Z,k∈Zn are dual frames. The order of the quotient group Q−1Zn/Zn is det Q, so let {di : i = 1, . . . ,det Q} denote a complete set of representatives of the quotient group, and define Ψ = {Tdiψ : i = 1, . . . ,det Q} , Φ = {Tdiφ : i = 1, . . . ,det Q} . Since {DAj TQ−1kψ}j∈Z,k∈Zn = {DAj Tkψ}j∈Z,k∈Zn,ψ∈Ψ and likewise for the dual frame, the statement follows. = h· , · i1/2 ∗ 3.3. A concrete version of Theorem 2.3. We list some standing assumptions and conventions for this section. General setup. We assume A ∈ GLn(R) is expansive. Let · ∗ be a Her- mitian norm as in (iii) in Proposition 3.1 associated with B = At, let I∗ denote the unit ball in the · ∗ -norm, and let K ∈ GLn(R) be the symmetric, positive definite matrix = ytKx. Let Λ := diag(λ1, . . . , λn), where {λi} are the eigenvalues of such that hx, yi∗ K, and let Q ∈ O(n) be such that the spectral decomposition of K is QtKQ = Λ. The following result is a special case of Theorem 2.3, where we, in particular, specify how to choose the translation lattice Γ. Since we in Theorem 3.3 define Γ, it allows for a more automated construction procedure. Theorem 3.3. Let A, I∗, K, Q, Λ be as in the general setup. Let d ∈ N0 and ψ ∈ L2(Rn). Suppose that ψ is a bounded, real-valued function with supp ψ ⊂ Bc(I∗)\ Bc−d−1(I∗) for some c ∈ Z, and that (8) holds. Take Γ = (1/2)AcQ√ΛZn. Then the function ψ and the function φ defined by φ(x) = d(Γ)"ψ(x) + 2 det A−j ψ(A−jx)# (24) for x ∈ Rn, generate dual frames {DAj Tγψ}j∈Z,γ∈Γ and {DAj Tγφ}j∈Z,γ∈Γ for L2(Rn) Remark 2. Note that d(Γ) = 2−n det Ac (λ1 · · · λn)1/2 and √Λ = diag(√λ1, . . . ,√λn). dXj=0 14 JAKOB LEMVIG Proof. The annulus O∗ is a bounded multiplicative tiling set for the dilations {Bj : j ∈ Z} by Lemma 3.2, hence this is also the case for Bc(O∗) for c ∈ Z. The support of ψ is supp ψ ⊂ Bc(I∗) \ Bc−d−1(I∗) = ∪d j=0Bc−j(O∗). Therefore we can apply Theorem 2.3 with E = Bc(O∗), bj = 2 and b−j = 0 for j = 1, . . . , d so that m = 0 and m = d. The only thing left to justify is the choice of the translation lattice Γ. We need to show that condition (9) with m = 0 and m = d in Theorem 2.3 is satisfied by Γ∗ = 2BcQΛ−1/2Zn. By the orthogonal substitution x = Qx the quadratic form xtKx of equation (18) reduces to where λi > 0, hence in the x = Qtx coordinates I∗ is given by λ1 x2 1 + · · · + λn x2 n, 1/√λ1(cid:19)2 + · · · +(cid:18) xn 1/√λn(cid:19)2 < 1) I∗ =(x ∈ Rn :(cid:18) x1 which is an ellipsoid with semi axes 1√λ1 ( I∗ + γ) ∩ I∗ = ∅ or, in the x coordinates, , . . . , 1√λn . Therefore, in the x coordinates, for 0 6= γ ∈ 2Λ−1/2Zn, By applying Bc to this relation it becomes (I∗ + γ) ∩ I∗ = ∅ for 0 6= γ ∈ 2QΛ−1/2Zn. (25) (cid:0)Bc(I∗) + γ(cid:1) ∩ Bc(I∗) = ∅ for 0 6= γ ∈ Γ∗ = 2BcQΛ−1/2Zn, whereby we see that condition (9) is satisfied with m = 0 and Γ∗ = 2BcQΛ−1/2Zn. The dual lattice of Γ∗ is Γ = 1/2A−cQΛ1/2Zn. It follows from Theorem 2.3 that ψ and φ generate dual frames for this choice of the translation lattice. (cid:3) The frame bounds for the pair of dual frames {DAj Tγψ}j∈Z,γ∈Γ and {DAj Tγφ}j∈Z,γ∈Γ in Theorem 3.3 can be given explicitly as C1 = 1 d(Γ) inf ξ∈Bc−d(O∗) dXj=0(cid:16) ψ(Bjξ)(cid:17)2 , C2 = 1 d(Γ) sup ξ∈Bc−d(O∗) dXj=0(cid:16) ψ(Bjξ)(cid:17)2 , and C1 = 1 d(Γ) inf ξ∈Bc−d(O∗) dXj=−d(cid:16) φ(Bjξ)(cid:17)2 , C2 = 1 d(Γ) sup ξ∈Bc−d(O∗) dXj=−d(cid:16) φ(Bjξ)(cid:17)2 , respectively. The frame bounds do not depend on the specific structure of Γ, but only on the determinant of Γ; in particular, the condition number C2/C1 is independent of Γ. To verify these frame bounds, we note that equation (25) together with the fact supp ψ, supp φ ⊂ Bc(I∗) imply that ψ(ξ) ψ(ξ + γ) = φ(ξ) φ(ξ + γ) = 0 for a.e. ξ ∈ Rn and γ ∈ Γ∗ \ {0}. and 2 Xj∈Z Xγ∈Γ∗(cid:12)(cid:12)(cid:12) ψ(Bjξ) ψ(Bjξ + γ)(cid:12)(cid:12)(cid:12) =Xj∈Z(cid:12)(cid:12)(cid:12) ψ(Bjξ)(cid:12)(cid:12)(cid:12) Xj∈Z Xγ∈Γ∗(cid:12)(cid:12)(cid:12) φ(Bjξ) φ(Bjξ + γ)(cid:12)(cid:12)(cid:12) =Xj∈Z(cid:12)(cid:12)(cid:12) φ(Bjξ)(cid:12)(cid:12)(cid:12) 2 = = dXj=0(cid:16) ψ(Bjξ)(cid:17)2 dXj=−d(cid:16) φ(Bjξ)(cid:17)2 , , CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 15 Therefore, by equations (12) and (13) with E = Bc(O∗), m = 0 and m = d, we have for ξ ∈ Bc−d(O∗). The stated frame bounds follow from Lemma 2.2. Example 3. Let A and K be as in Example 2. The eigenvalues of K are λ1 = (26 + 2√65)/9 ≈ 4.7 and λ2 = (26 − 2√65)/9 ≈ 1.1. Let the normalized (in the standard norm) eigenvectors of K be columns of Q ∈ O(2) and Λ = diag(λ1, λ2), hence QtKQ = Λ. By the orthogonal transformation x = Qx the Hermitian norm unit ball I∗ becomes I∗ =(x ∈ R2 :(cid:18) x1 1/√λ1(cid:19)2 which is an ellipse with semimajor axis 1/√λ2 ≈ 0.95 and semiminor axis 1/√λ1 ≈ 0.46. Since Λ−1/2 = diag(1/√λ1, 1/√λ2), we have (cid:12)(cid:12)(cid:12)( I∗ + γ) ∩ I∗(cid:12)(cid:12)(cid:12) = 0 By the orthogonal substitution back to x coordinates, we get 1/√λ2(cid:19)2 +(cid:18) x2 for 0 6= γ ∈ 2Λ−1/2Z2. < 1) ⊂ I2 (I∗ + γ) ∩ I∗ = 0 for 0 6= γ ∈ 2QΛ−1/2Z2. Suppose that ψ is a bounded, real-valued function with supp ψ ⊂ Bc(I∗)\ Bc−d−1(I∗) for c = 1 that satisfies the B-dilative partition (8). Since c = 1 we need to take Γ∗ = 2B1QΛ−1/2Z2 and Γ = 1/2A−1QΛ1/2Z2, see Figure 5 and 6. 3.4. An alternative lattice choice. Let the setup up and assumptions be as in The- orem 3.3, except for the lattice Γ which we want to choose differently. As in Section 3.2 the dual lattice Γ∗ needs to satisfy (23) for γ ∈ Γ∗\{0}. We want to choose Γ∗ as dense as possible since this will make the translation lattice Γ as sparse as possible and the wavelet system with as few translates as possible. Since supp ψ, supp φ ⊂ Bc(I∗), we are looking for lattices Γ∗ that packs the ellipsoids Bc(I∗) + γ, γ ∈ Γ∗, in a non-overlapping, optimal way. By the coordinate transformation x = Λ−1/2QtB−cx, the ellipsoid Bc(I∗) turns into the standard unit ball I2 in Rn. This calculations are as follows. Bc(I∗) =(cid:8)Bcx : x2 ∗ ≤ 1(cid:9) =(cid:8)x : K 1/2B−cx2 2 ≤ 1o 2 ≤ 1(cid:9) =nx :(cid:12)(cid:12)K 1/2B−cBcQΛ−1/2 x(cid:12)(cid:12)2 =(cid:8)x :(cid:10)x, Λ−1/2QtKQΛ−1/2 x(cid:11)2 ≤ 1(cid:9) =(cid:8)x : x2 2 ≤ 1(cid:9) , and we arrive at a standard sphere packing problem with lattice arrangement of non- overlapping unit n-balls. The proportion of the Euclidean space Rn filled by the balls is called the density of the arrangement, and it is this density we want as high as possible. 16 JAKOB LEMVIG 6 4 2 K6 K4 K2 0 2 4 6 K2 K4 K6 Figure 5. The dual lattice Γ∗ = 2BcQΛ−1/2Z2 for c = 1 is shown by dots, and the boundary of the set Bc(I∗) by a solid line. Boundaries of the set Bc(I∗) translated to several different γ ∈ Γ∗ \ {0} are shown with dashed lines. Recall that supp ψ, supp φ ⊂ Bc(I∗), hence supp φ ∩ supp ψ(· + γ) = ∅ for γ ∈ Γ∗ \ {0}. 6 4 2 K6 K4 K2 0 2 4 6 K2 K4 K6 Figure 6. The translation lattice Γ = (1/2)AcQΛ1/2Z2 for c = 1. Taking Γ as in Theorem 3.3 corresponds to a square packing of the unit n-balls I2 + k by the lattice 2Zn, i.e., k ∈ 2Zn. The density of this packing is Vn2−n, where Vn is the volume of the n-ball: V2n = πn/(n!) and V2n+1 = (22n+1n!πn)/(2n + 1)!. This is not the densest packing of balls in Rn since there exists a lattice with density bigger than 1.68n2−n for each n 6= 1 [9]; a slight improvement of this lower bound was obtained in [1] for n > 5. Moreover, the densest lattice packing of hyperspheres is known up to CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 17 dimension 8, see [19]; it is precisely this dense lattice we want to use in place of 2Zn (at least whenever n ≤ 8). balls in a hexagonal lattice, has the highest density π/√12. Hence using P Z2 with In R2 Lagrange proved that the hexagonal packing, where each ball touches 6 other P =(cid:18)2 1 √3(cid:19) 0 instead of 2Z2 improves the packing by a factor of π/√12 π/22 = 4/√12 = 2/√3. It is easily seen that this factor equals the relation between the area of the fundamental parallelogram of the two lattices det 2I2×2 / det P. In Figure 5 we see that each ellipse only touches 4 other ellipses corresponding to the square packing 2Zn; in the optimal In R3 Gauss proved that the highest density is packing each ellipse touch 6 others. π/√18 obtained by the hexagonal close and face-centered cubic packing; here each ball touches 12 other balls. 4. Dilative partition of unity With Theorem 3.3 at hand the only issue left is to specify how to construct functions satisfying the partition of unity (8) for any given expansive matrix. In the two examples of this section we outline possible ways of achieving this. 4.1. Constructing a partition of unity. As usual we fix the dimension n ∈ N and the expansive matrix B ∈ GLn(R). In the examples in this section we construct functions satisfying the assumptions in Theorem 3.3, that is, a real-valued function g ∈ L2(Rn) with supp g ⊂ Bc(I∗) \ Bc−d−1(I∗) for some c ∈ Z and d ∈ N0 so that the B-dilative partition (26) holds. Xj∈Z g(Bjξ) = 1 for a.e. ξ ∈ Rn, In the construction we will use that the radial coordinate of the surface of the ellipsoid ∂Bj(I∗), j ∈ Z, can be parametrized by the n− 1 angular coordinates θ1, . . . , θn−1. The radial coordinate expression will be of the form h(θ1, . . . , θn−1)−1/2 for some positive, trigonometric function h, where h is bounded away from zero and infinity with the specific form of h depending on the dimension n and the length and orientation of the ellipsoid axes. We illustrate this with the following example in R4. We want to find the radial coordinate r of the ellipsoid (cid:8)x ∈ R4 : (x1/ℓ1)2 + (x2/ℓ2)2 + (x3/ℓ3)2 + (x4/ℓ4)2 = 1(cid:9) , as a function the angular coordinates θ1, θ2 and θ3. We express x = (x1, x2, x3, x4) ∈ R4 in the hyperspherical coordinates (r, θ1, θ2, θ3) ∈ {0} ∪ R+ × [0, π] × [0, π] × [0, 2π) as ℓi > 0, i = 1, 2, 3, 4, 18 follows: JAKOB LEMVIG x1 = r cos θ1, x3 = r sin θ1 sin θ2 cos θ3, x2 = r sin θ1 cos θ2, x4 = r sin θ1 sin θ2 sin θ3. 1 cos2 θ1 + ℓ−2 + ℓ−2 Then we substitute xi, i = 1, . . . , 4, in the expression above and factor out r2 to obtain r2f (θ1, θ2, θ3) = 1, where (27) f (θ1, θ2, θ3) = ℓ−2 2 sin2 θ1 cos2 θ2 4 sin2 θ1 sin2 θ2 sin2 θ3. 3 sin2 θ1 sin2 θ2 cos2 θ3 + ℓ−2 The conclusion is that r = r(θ1, θ2, θ3) = f (θ1, θ2, θ3)−1/2. Example 4. For d = 1 in Theorem 3.3 we want g ∈ C s 0(Rn) for any given s ∈ N ∪ {0}. The choice d = 1 will fix the "size" of the support of g so that supp g ⊂ Bc(I∗)\Bc−2(I∗) for some c ∈ Z. Now let r1 = r1(θ1, . . . , θn−1) and r2 = r2(θ1, . . . , θn−1) denote the radial coordinates of the surface of the ellipsoids ∂Bc−1(I∗) and ∂Bc(I∗) parametrized by n−1 angular coordinates θ1, . . . , θn−1, respectively. Let f be a continuous function on the annulus S = Bc(O∗) satisfying f∂Bc−1(I∗) = 1 and f∂Bc(I∗) = 0. Using the parametrizations r1, r2 of the surfaces of the two ellipsoids and fixing the n−1 angular coordinates we realize that we only have to find a continuous function f : [r1, r2] → R of one variable (the radial coordinate) satisfying f (r1) = 1 and f (r2) = 0. For example the general function f ∈ C 0(S) of d variables can be any of the functions below: f (x) = f (r, θ1, . . . , θn−1) = f (x) = f (r, θ1, . . . , θn−1) = f (x) = f (r, θ1, . . . , θn−1) = 1 r2 − r , r2 − r1 (r2 − r)2 (r2 − r1)3 (2(r − r1) + r2 − r1), 2 + 1 ), (28c) where r = x ∈ [r1, r2], θ1, . . . , θn−2 ∈ [0, π], and θn−1 ∈ [0, 2π); recall that r1 = r1(θ1, . . . , θn−1) and r2 = r2(θ1, . . . , θn−1). In definitions (28b) and (28c) the function f even belongs to C 1(S). Define g ∈ L2(R) by: 2 cos π( r−r1 r2−r1 (28a) (28b) (29) 1 − f (Bx) f (x) 0 for x ∈ Bc−1(I∗) \ Bc−2(I∗), for x ∈ Bc(I∗) \ Bc−1(I∗), otherwise. g(x) = This way g becomes a B-dilative partition of unity with supp g ⊂ Bc(I∗) \ Bc−2(I∗), so we can apply Theorem 3.3 with ψ = g and d = 2. We can simplify the expressions for the radial coordinates r1, r2 of the surface of the ellipsoids ∂Bc−1(I∗) and ∂Bc(I∗) from the previous example by a suitable coordinate change. The idea is to transform the ellipsoid Bc−1(I∗) to the standard unit ball I2 by a first coordinate change x = Λ1/2QtB−c+1x. This will transform the outer ellipsoid Bc(I∗) to another ellipsoid. A second and orthogonal coordinate transform x = Qt x ′ will make the semiaxes of this new ellipsoid parallel to the coordinate axes, leaving CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 19 the standard unit ball I2 unchanged. Here Q′ comes from the spectral decomposition of A−1B−1, i.e., A−1B−1 = Qt In the x coordinates r1 = 1 is a constant and ′ r2 = f−1/2 with f of the form (27) for n = 4 and likewise for n 6= 4. for all d ∈ N; moreover, the constructed function will belong to C∞0 (Rn). Example 5. For sufficiently small δ > 0 define ∆1, ∆2 ⊂ Rn by In the construction in Example 4 we assumed that d = 1. The next example works Λ′Q′. ∆1 = Bc−d−1(I∗) + B(0, δ), ∆2 + B(0, δ) = Bc(I∗). This makes ∆2 \ ∆1 a subset of the annulus Bc(I∗)\ Bc−d−1(I∗); it is exactly the subset, where points less than δ in distance from the boundary have been removed, or in other words ∆2 \ ∆1 + B(0, δ) = Bc(I∗) \ Bc−d−1(I∗). For this to hold, we of course need to take δ > 0 sufficiently small, e.g. such that ∆1 ⊂ r∆1 ⊂ ∆2 holds for some r > 1. hδ = δ−dh(δ−1·). By convoluting the characteristic function on ∆2 \ ∆1 with hδ we obtain a smooth function living on the annulus Bc(I∗)\ Bc−d−1(I∗). So let p ∈ C∞0 (Rn) be defined by Let h ∈ C∞0 (Rn) satisfy supp h = B(0, 1), h ≥ 0, and R h dµ = 1, and define and note that supp p = Bc(I∗) \ Bc−d−1(I∗) since supp hδ = B(0, δ). Normalizing the function p in a proper way will give us the function g we are looking for. We will normalize p by the function w: p = hδ ∗ χ∆2\∆1, w(x) =Xj∈Z p(Bjx). For a fixed x ∈ Rn\{0} this sum has either d or d + 1 nonzero terms, and w is therefore bounded away from 0 and ∞: ∃c, C > 0 : c < w(x) < C for all x ∈ Rn \ {0}, hence we can define a function g ∈ C∞0 (Rn) by (30) g(x) = p(x) w(x) for x ∈ Rn \ {0}, and, g(0) = 0. The function g will be an almost everywhere B-dilative partition of unity as is seen by using the B-dilative periodicity of w: Xj∈Z g(Bjx) =Xj∈Z p(Bjx) w(Bjx) =Xj∈Z p(Bjx) w(x) = 1 w(x)Xj∈Z p(Bjx) = 1. Since p is supported on the annulus Bc(I∗) \ Bc−d−1(I∗), we can simplify the definition in (30) to get rid of the infinite sum in the denominator; this gives us the following 20 expression JAKOB LEMVIG g(x) = p(x)/ p(Bjx) for x ∈ Rn \ {0}. dXj=−d We can obtain a more explicit expression for p by the following approach. Let r1 = r1(θ1, . . . , θn−1) and r2 = r2(θ1, . . . , θn−1) denote the radial coordinates of the surface of the ellipsoids ∂Bc−d−1(I∗) and ∂Bc(I∗) parametrized by n − 1 angular coordinates θ1, . . . , θn−1, respectively. Finally, let p ∈ C∞0 (Rn) be defined by p(x) = η(x − r1) η(r2 − x), with r1 = r1(θ1, . . . , θn−1) and r2 = r2(θ1, . . . , θn−1) where θ1, . . . , θn−1 can be found from x, and η(x) =(e−1/x 0 x > 0, x ≤ 0. References [1] K. Ball, A lower bound for the optimal density of lattice packings, Int. Math. Res. Not. 1992 (1992), 217 -- 221. [2] M. Bownik, Anisotropic Hardy spaces and wavelets, Mem. Amer. Math. Soc. 164 (2003), no. 781. [3] O. Christensen, An introduction to frames and Riesz bases, Birkhäuser Boston Inc., Boston, MA, 2003. [4] O. Christensen, R.Y. Kim, On dual Gabor frame pairs generated by polynomials, J. Fourier Anal. Appl. 16 (2010), 1 -- 16. [5] O. Christensen, A. Rahimi, Frame properties of wave packet systems in L2(Rn), Adv. Comput. Math. 29 (2008), 101 -- 111. [6] C.K. Chui, W. Czaja, M. Maggioni, G. Weiss, Characterization of general tight wavelet frames with matrix dilations and tightness preserving oversampling, J. Fourier Anal. Appl. 8 (2002), 173 -- 200. [7] C.K. Chui, W. He, Construction of multivariate tight frames via Kronecker products, Appl. Com- put. Harmon. Anal. 11 (2001), 305 -- 312. [8] I. Daubechies, B. Han, A. Ron, Z. Shen, Framelets: MRA-based constructions of wavelet frames, Appl. Comput. Harmon. Anal. 14 (2003), 1 -- 46. [9] H. Davenport, C.A. Rogers, Hlawka's theorem in the geometry of numbers, Duke Math. J. 14 (1947), 367 -- 375. [10] M. Ehler, On multivariate compactly supported bi-frames, J. Fourier Anal. Appl. 13 (2007), 511 -- 532. [11] M. Ehler, B. Han, Wavelet bi-frames with few generators from multivariate refinable functions, Appl. Comput. Harmon. Anal. 25 (2008), 407 -- 414. [12] B. Han, Compactly supported tight wavelet frames and orthonormal wavelets of exponential decay with a general dilation matrix, J. Comput. Appl. Math. 155 (2003), 43 -- 67. [13] B. Han, Q. Mo, Multiwavelet frames from refinable function vectors, Adv. Comput. Math. 18 (2003), 211 -- 245. [14] M.-J. Lai, J. Stöckler, Construction of multivariate compactly supported tight wavelet frames, Appl. Comput. Harmon. Anal. 21 (2006), 324 -- 348. [15] J. Lemvig, Constructing pairs of dual bandlimited framelets with desired time localization, Adv. Comput. Math. 30 (2009), 231 -- 247 [16] A. Ron, Z. Shen, Compactly supported tight affine spline frames in L2(Rd), Math. Comp. 67 (1998), 191 -- 207. [17] A. Ron, Z. Shen, Construction of compactly supported affine frames in L2(Rd), Advances in Wavelets (Hong Kong, 1997), 27 -- 49, Springer-Verlag, Singapore, 1999. CONSTRUCTING PAIRS OF DUAL BANDLIMITED FRAME WAVELETS IN L2(Rn) 21 [18] Z. Shang, X. Zhou, Dual generators for weighted irregular wavelet frames and reconstruction error, Appl. Comput. Harmon. Anal. 22 (2007), 356 -- 367. [19] N.J.A. Sloane, The sphere packing problem, Proceedings of the International Congress of Mathe- maticians, Vol. III (Berlin, 1998). Doc. Math. 1998, Extra Vol. III, 387 -- 396. [20] D. Yang, X. Zhou, Z.Z. Yuan, Frame wavelets with compact supports for L2(Rn), Acta Math. Sin. (Engl. Ser.) 23 (2007), 349 -- 356. Institute of Mathematics, University of Osnabrück, 49069 Osnabrück, Germany E-mail address: [email protected]
1512.02591
1
1512
2015-12-08T19:10:20
Developed Matrix inequalities via Positive Multilinear Mappings
[ "math.FA", "math.OA" ]
Utilizing the notion of positive multilinear mappings, we give some matrix inequalities. In particular, Choi--Davis--Jensen and Kantorovich type inequalities including positive multilinear mappings are presented.
math.FA
math
DEVELOPED MATRIX INEQUALITIES VIA POSITIVE MULTILINEAR MAPPINGS MAHDI DEHGHANI, MOHSEN KIAN AND YUKI SEO Abstract. Utilizing the notion of positive multilinear mappings, we give some matrix inequalities. In particular, Choi–Davis–Jensen and Kantorovich type inequalities including positive multilinear mappings are presented. 1. Introduction 5 1 0 2 c e D 8 ] . A F h t a m [ 1 v 1 9 5 2 0 . 2 1 5 1 : v i X r a Throughout the paper assume that Mp(C) := Mp is the C ∗-algebra of all p × p complex matrices with the identity matrix I. A matrix A ∈ Mp is positive (denoted by A ≥ 0) if it is Hermitian and all its eigenvalues are nonnegative. If in addition A is invertible, then it is called strictly positive (denoted by A > 0). A linear mapping Φ : Mn → Mp is called positive if Φ preserves the positivity and Φ is called unital if Φ(I) = I. The notion of positive linear mappings on C ∗-algebras is a well-known tool to study matrix inequalities which have been studied by many mathematicians. If Φ : Mn → Mp is a unital positive linear mapping, then the famous Kadison’s inequality states that Φ(A)2 ≤ Φ(A2) for every Hermitian matrix A. Moreover, Choi’s inequality asserts that Φ(A)−1 ≤ Φ(A−1) for every strictly positive matrix A, see e.g. [4, 9, 10]. A continuous real function f : J → R is said to be matrix convex if f(cid:0) A+B for all Hermitian matrices A, B with eigenvalues in J. Positive linear mappings have been used to characterize 2 (cid:1) ≤ f (A)+f (B) 2 matrix convex and matrix monotone functions. It is well-known that a continuous real function f : J → R is matrix convex if and only if f (Φ(A)) ≤ Φ(f (A)) for every unital positive linear mapping Φ and every Hermitian matrix A with spectrum in J. The later inequality is known as the Choi-Davis-Jensen inequality, see [9, 10, 12]. For more information about matrix inequalities concerning positive linear mappings the reader is referred to [2, 7, 8, 13, 15, 16, 20, 22, 26] and references therein. 2010 Mathematics Subject Classification. Primary 15A69 ; Secondary 47A63,47A64, 47A56. Key words and phrases. Positive multilinear mapping, positive matrix, spectral decompo- sition, matrix convex, matrix mean, Karcher mean. 1 2 M. DEHGHANI, M. KIAN AND Y. SEO It is known that if A = (aij) and B = (bij) are positive matrices, then so is their Hadamard product A ◦ B, which is defined by the entrywise product as A ◦ B = (aijbij). The same is true for their tensor product A ⊗ B, see e.g. [14]. Moreover, the mapping (A, B) → A⊗B is also linear in each of its variables. So if we define Φ : M2 q → Mp by Φ(A, B) = A⊗ B, then Φ is multilinear and positive in the sense that Φ(A, B) is positive whenever A, B are positive. However, the Choi–Davis–Jensen type inequality f (Φ(A, B)) ≤ Φ(f (A), f (B)) does not hold in general for a unital positive multilinear mapping Φ and matrix convex functions f . This can be a motivation to study matrix inequalities via positive multilinear mappings. The main aim of the present work is to inquire matrix inequalities using posi- tive multilinear mappings. Some of these inequalities would be generalization of inequalities for the Hadamard product and the tensor product of matrices. In Section 2, we give basic facts and examples for positive multilinear mappings. In Section 3, Choi–Davis–Jensen type inequalities are presented for positive multi- linear mappings and some of its applications is given. In Section 4, we obtain some convex theorem for functions concerning positive multilinear mappings and then we give some Kantorovich type inequalities. Section 5 is devoted to some re- verses of the Choi–Davis–Jensen inequality. In Section 6, we consider the Karcher mean and present multilinear versions of some known results. 2. Preliminaries We start by definition of a positive multilinear mapping. Definition 2.1. A mapping Φ : Mk q → Mp is said to be multilinear if it is linear in each of its variable. It is called positive if Φ(A1,· · · , Ak) ≥ 0, whenever Ai ≥ 0 (i = 1,· · · , k). It is called strictly positive if Φ(A1,· · · , Ak) > 0 whenever Ai > 0 (i = 1,· · · , k). If Φ(I,· · · , I) = I, then Φ is called unital. Example 2.2. The tensor product of every two positive matrices is positive again. This ensures that the mapping Φ : Mk q → Mqk defined by Φ(A1,· · · , Ak) = A1 ⊗ · · · ⊗ Ak is positive. Moreover, it is multilinear and unital. If Ψ : Mqk → Mp is a unital positive linear mapping, then the map defined by Φ(A1, . . . , Ak) = Ψ(A1 ⊗ · · · ⊗ Ak) POSITIVE MULTILINEAR MAPPINGS 3 is a unital positive multilinear mapping. In particular, the Hadamard product Φ(A1,· · · , Ak) = A1 ◦ · · · ◦ Ak. is positive and multilinear. Example 2.3. Assume that Xi ∈ Mq (i = 1,· · · , k) and Pk q → Mp defined by Φ(A1,· · · , Ak) = Pk mapping Φ : Mk and unital. However, it is not multilinear. q → Mp defined by Φ(A1,· · · , Ak) := Tr(A1) · · · Tr(Ak)I is positive and multilinear, where Tr is the canonical trace. Example 2.4. The mappings Φ : Mk i=1 X ∗ i=1 X ∗ i Xi = I. The i AiXi is positive (2.1) Example 2.5. Define the mapping Φ : M2 p → Mp by Φ(T, S)x = (T x ⊗ Sx)x, where (u ⊗ v)(w) = hw, viu for u, v, w ∈ Cp. Then Φ is multilinear. Moreover, if T and S are positive matrices, then hΦ(T, S)x, xi = h(T x ⊗ Sx)x, xi = hhSx, xiT x, xi = hSx, xihT x, xi ≥ 0 for every x ∈ Cp. Therefore, Φ is a positive multilinear mapping. It would be useful to see that there is a relevant connection between positive lin- ear and multilinear mappings. Corresponding to every positive multilinear map- ping Φ : Mk q → Mp, the map Ψ : Mq → Mp defined by Ψ(X) = Φ(X, I,· · · , I) is positive and linear. Conversely, if Ψ : Mq → Mp is positive and linear, then Φ : Mk q → Mp, is positive and multilinear. Φ(A1,· · · , Ak) = Ψ(A1) ⊗ A2 ⊗ · · · ⊗ Ak We state basic properties for positive multilinear mappings: Lemma 2.6. Every positive multilinear mapping Φ : Mk preserving; i.e., q → Mp is adjoint- Φ(T1, . . . , Tk)∗ = Φ(T ∗ 1 , . . . , T ∗ k ) for all Ti ∈ Mq (i = 1, . . . , k). Proof. It suffices to show that Φ(A1, . . . , Ak) is Hermitian for every Hermitian A1, . . . , Ak. Every Hermitian matrix Ai has a Jordan decomposition Ai = Ai+ − Ai− where Ai± ≥ 0. 4 M. DEHGHANI, M. KIAN AND Y. SEO Then we have Φ(A1, . . . , Ak) = Φ(A1+ − A1−, . . . , Ak+ − Ak−) = Φ(A1+, . . . , Ak+) − Φ(A1−, A2+, . . . , Ak+) + · · · + (−1)kΦ(A1−, . . . , Ak−) and each term is positive and hence Φ(A1, . . . , Ak) is Hermitian. (cid:3) The next Lemma is easy to verify. Lemma 2.7. Every positive multilinear mapping Φ : Mk i.e., q → Mp is monotone, 0 ≤ Ai ≤Bi for all i = 1, . . . , k The norm of a multilinear mapping Φ : Mk k Φ k= kA1k=···=kAkk=1 k Φ(A1, . . . , Ak) k= sup =⇒ Φ(A1, A2, . . . , Ak) ≤ Φ(B1, B2, . . . , Bk). q → Mp is defined by kA1k≤1,··· ,kAkk≤1 k Φ(A1, . . . , Ak) k . sup A version of Russu-Dye theorem (see for example [4, Theorem 2.3.7]) has been proved in [6] as follows: Theorem 2.8. [6] If Φ : Mq → Mp is a unital positive multilinear mapping, then k Φ k= 1. 3. Jensen type inequalities If Φ is a unital positive linear mapping, then the Choi–Davis–Jensen inequality Φ(f (A)) ≥ f (Φ(A)) (3.1) holds for every matrix convex function f on J and every Hermitian matrix A whose spectrum is contained in J. However, this inequality does not hold for a unital positive multilinear mapping in general. For example, consider the matrix convex function f (t) = t2 − t and the unital positive multilinear map- If A = 2I and B = I, then 2I = f (Φ(A, B)) (cid:2) ping Φ(A, B) = A ◦ B. Φ(f (A), f (B)) = 0. Another example is the unital positive multilinear map- ping Ψ(A, B) = 1 p Tr(AB) for A, B ∈ Mp and f (Ψ(A, B)) 6≤ Ψ(f (A), f (B)). We consider an addition hypothesis on f under which this inequality holds. For this end, assume that the positive multilinear mapping Φ : M2 q → Mp is defined by Φ(A, B) = A◦ B. If inequality (3.1) is valid for a matrix convex function f , then POSITIVE MULTILINEAR MAPPINGS 5 f (A ◦ B) ≤ f (A) ◦ f (B). In particular, f (xy) ≤ f (x)f (y) for all x, y in domain of f . This turns out to be a necessary and sufficient condition under which (3.1) holds for every unital positive multilinear mapping. Definition 3.1. Let J be an interval in R. A real valued continuous function f is super-multiplicative on J if f (xy) ≥ f (x)f (y) for all x, y ∈ J, while f is sub-multiplicative if f (xy) ≤ f (x)f (y) for all x, y ∈ J. We present a Choi–Davis–Jensen inequality for positive multilinear mappings. Theorem 3.2. Let Ai ∈ Mq (i = 1, . . . , k) be positive matrices and Φ : Mk q → [0,∞) → R is a sub- Mp be a unital positive multilinear mapping. multiplicative matrix convex function (resp. a super-multiplicative matrix concave If f : function), then f (Φ(A1, . . . , Ak)) ≤ Φ(f (A1), . . . , f (Ak)) (resp. f (Φ(A1, . . . , Ak)) ≥ Φ(f (A1), . . . , f (Ak))). j=1 λijPij (i = 1,· · · , k) is the spectral decomposition j=1 Pij = I. Then Proof. Assume that Ai =Pqi of each Ai for which Pqi Φ(f (A1), f (A2),··· , f (Ak)) = Φ Xj=1  Xj1 Xj2 = q1 q1 q2 ··· qk Xjk f (λ1j)P1j ,··· , qk Xj=1 f (λkj)Pkj  (3.2) f (λ1j1)f (λ2j2)··· f (λkjk)Φ(P1j1 ,··· , Pkjk ), 1 2 we get q1 q2 qk q1 q2 ··· λ1j1 λ2j2 ··· λkjkΦ(P1j1 ,··· , Pkjk ) C(j1,··· , jk)λ1j1 λ2j2 ··· λkjkC(j1,··· , jk)  by multilinearity of Φ. With C(j1,· · · , jk) := Φ(P1j1,· · · , Pkjk) f (Φ(A1,··· , Ak)) = f Xj1 Xj2  = f Xj1 Xj2  Xjk Xj1 Xj2 ··· Xjk Xj1 Xj2 = Φ(f (A1),··· , f (Ak)). Xjk Xjk C(j1,··· , jk)f (λ1j1 λ2j2 ··· λkjk)C(j1,··· , jk) ··· ··· ≤ q1 q2 q1 q2 ≤ qk qk qk  (by multilinearity of Φ) (by matrix convexity of f ) f (λ1j1)f (λ2j2)··· f (λkjk)Φ(P1j1 ,··· , Pkjk ) (by f (xy) ≤ f (x)f (y) on [0,∞)) 6 M. DEHGHANI, M. KIAN AND Y. SEO As an example, the power functions f (t) = tr are sub-multiplicative and so we have the following extension for Kadison and Choi’s inequalities. Corollary 3.3. Suppose that Φ : Mk mapping. q → Mp is a unital positive multilinear (cid:3) (1) If 0 ≤ r ≤ 1, then Φ(Ar 1,· · · , Ar k) ≤ Φ(A1,· · · , Ak)r (Ai ≥ 0), (2) If −1 ≤ r ≤ 0 and 1 ≤ r ≤ 2, then Φ(A1,· · · , Ak)r ≤ Φ(Ar 1,· · · , Ar k) (Ai > 0). q → Mp be a uninal positive multilinear map. If either 2 , then Corollary 3.4. Let Φ : Mk s ≤ t, s 6∈ (−1, 1), t 6∈ (−1, 1) or 1 2 ≤ s ≤ 1 ≤ t or s ≤ −1 ≤ t ≤ − 1 1,· · · , As k) for all positive matrices A1,· · · , Ak. Proof. If 0 ≤ s ≤ t, then r = s Ai now implies that t ≤ 1. Applying Corollary 3.3 with At 1,· · · , At k) s ≤ Φ(At Φ(As 1 t 1 i instead of Φ(As 1,· · · , As k) ≤ Φ(At 1,· · · , At k) s t . Applying the Lowner-Heinz inequality with the power 1 s yields that Φ(As 1,· · · , As k) 1 s ≤ Φ(At 1,· · · , At k) 1 t . (cid:3) Example 3.5. The matrix concave function f (t) = log t defined on (0,∞) is super-multiplicative on [1, e2]. For if x, y ∈ [1, e2], then log x, log y ∈ [0, 2] and so (log x)(log y) ≤ 4. Hence (log x)2(log y)2 ≤ 4(log x)(log y) and so f (x)f (y) = log x log y ≤ 2p(log x)(log y) ≤ log x + log y = log(xy) = f (xy), where we used the arithmetic-geometric mean inequality. Now applying Theorem 3.2 we get log(A1 ⊗ A2) ≥ log(A1) ⊗ log(A2) for all positive matrices Ai with eigenvalues in [1, e2]. However the inequality log(xyz) ≤ log x log y log z does not hold for x, y, x ∈ [1, e2] in general. POSITIVE MULTILINEAR MAPPINGS 7 We recall the theory of matrix means by Kubo–Ando[17]. A key for the theory is that there is a one-to-one correspondence between the matrix mean σ and the nonnegative matrix monotone function f (t) on (0,∞) with f (1) = 1 by the formula f (t) = 1 σ t for all t > 0, or A σ B = A 1 2 f (A− 1 2 BA− 1 2 )A 1 2 for all A, B > 0. We say that f is the representing function for σ. In this case, notice that f (t) is matrix monotone if and only if it is matrix concave [12]. Moreover, every matrix mean σ is subadditive: A σ C + B σ D ≤ (A + B) σ (C + D). By a theorem of Ando [1], if A and B are positive matrices and Φ is a strictly positive linear mapping, then for each α ∈ [0, 1] Φ(A♯αB) ≤ Φ(A)♯αΦ(B), (3.3) where the α-geometric matrix mean is defined by A♯αB = A 2 BA− 1 1 2 A 1 2 (cid:16)A− 1 2(cid:17)α and the geometric matrix mean ♯ is defined by ♯ = ♯1/2, namely A♯B = A 1 2 (cid:16)A− 1 2 BA− 1 1 2 A 1 2 . 2(cid:17) By aid of Theorem 3.2, we show the positive multilinear mapping version of Ando’s inequality (3.3). Proposition 3.6. Let σ be a matrix mean with a super-multiplicative representing function f : [0,∞) → [0,∞), and Φ : Mk q → Mp be a strictly positive multilinear mapping. Then Φ(A1σB1,· · · , AkσBk) ≤ Φ(A1,· · · , Ak)σΦ(B1,· · · , Bk) for all A1,· · · , Ak > 0 and B1,· · · , Bk ≥ 0. Proof. For A1,· · · , Ak > 0 assume that a map Ψ : Mk k(cid:17) Φ(A1,· · · , AK)− 1 Ψ(X1,· · · , Xk) := Φ(A1,· · · , Ak)− 1 1 ,· · · , A Clearly Ψ is unital strictly positive and multilinear. For every i = 1,· · · , k, put (i = 1,· · · k). Then it follows from matrix concavity of f and Yi := A q → Mp is defined by k XkA 2 Φ(cid:16)A 1 X1A 2 2 − 1 − 1 i BiA i 2 2 2 1 1 2 1 1 2 . 8 M. DEHGHANI, M. KIAN AND Y. SEO Theorem 3.2 that Φ(A1,· · · , Ak)− 1 2 2 Φ(A1σB1,· · · , AkσBk)Φ(A1,· · · , Ak)− 1 = Ψ (f (Y1),· · · , f (Yk)) ≤ f(cid:0)Ψ(Y1,· · · , Yk)(cid:1) = f(cid:16)Φ(A1,· · · , Ak)− 1 2(cid:17) , 2 Φ(B1,· · · , Bk)Φ(A1,· · · , Ak)− 1 (cid:3) which completes the proof. For a matrix mean σ with the representing function f , assume that σo, the transpose of σ is defined by A σo B = B σ A. In this case, the representing function of σo is f o defined by f o(t) = tf (1/t). For α ∈ [0, 1], the transpose of geometric mean ♯α turns out to be ♯1−α. Ando [1] showed that if A, B > 0, then A ◦ B ≥ (A ♯α B) ◦ (A ♯1−α B) for every α ∈ [0, 1]. We state a generalization of this result using positive multi- linear mappings. Corollary 3.7. Let σ be a matrix mean with the representing function f which is super-multiplicative and not affine. Let A, B be positive matrices. If Φ : M2 q → Mp is a unital positive multilinear mapping, then Φ(A, B) + Φ(B, A) ≥ Φ(A σ B, A σo B) + Φ(A σo B, A σ B). In particular, for each α ∈ [0, 1] Φ(A, B) + Φ(B, A) ≥ Φ(A ♯α B, A ♯1−α B) + Φ(A ♯1−α B, A ♯α B). Proof. It follows from subadditivity of σ and Theorem 3.6 that Φ(A, B) + Φ(B, A) = [Φ(A, B) + Φ(B, A)] σ [Φ(B, A) + Φ(A, B)] ≥ Φ(A, B)σΦ(B, A) + Φ(B, A)σΦ(A, B) ≥ Φ(AσB, BσA) + Φ(BσA, AσB) = Φ(AσB, AσoB) + Φ(AσoB, AσB). If A is positive, then A ◦ A−1 ≥ I. This inequality is known as Fiedler’s inequality, see [4, 11]. As an application of Theorem 3.6 , we have the following extension of Fiedler’s inequality by Aujla-Vasudeva [3]. (cid:3) POSITIVE MULTILINEAR MAPPINGS 9 q → Mp be a positive multilinear map. If α, β ∈ R Proposition 3.8. Let Φ : M2 and λ ∈ [0, 1], then Φ(Aα, Aβ) + Φ(Aβ, Aα) ≥ Φ(A(1−λ)α+λβ, A(1−λ)β+λα) + Φ(A(1−λ)β+λα, A(1−λ)α+λβ) for all matrices A > 0. Proof. We have Φ(Aα, Aβ) + Φ(Aβ, Aα) =(cid:0)Φ(Aα, Aβ) + Φ(Aβ, Aα)(cid:1) ♯λ(cid:0)Φ(Aβ, Aα) + Φ(Aα, Aβ)(cid:1) ≥ Φ(Aα, Aβ)♯λΦ(Aβ, Aα) + Φ(Aβ, Aα)♯λΦ(Aα, Aβ) ≥ Φ(Aα♯λAβ, Aβ♯λAα) + Φ(Aβ♯λAα, Aα♯λAβ) = Φ(A(1−λ)α+λβ, A(1−λ)β+λα) + Φ(A(1−λ)β+λα, A(1−λ)α+λβ). With α = 1 and β = −1 and λ = 1 2 we get the following result. Corollary 3.9. Let Φ : M2 A > 0, then q → Mp be a strictly positive multilinear map. If (cid:3) Φ(A, A−1) + Φ(A−1, A) ≥ 2Ip. The next lemma will be used in the sequel. Lemma 3.10. [4, Theorem 1.3.3] Let A,B be strictly positive matrices. Then the block matrix " A X X ∗ B # is positive if and only if A ≥ XB−1X ∗. In [9], Choi generalized Kadison’s inequality to normal matrices by showing that if Φ is a unital positive linear mapping, then Φ(A)Φ(A∗) ≤ Φ(A∗A) and Φ(A∗)Φ(A) ≤ Φ(A∗A). for every normal matrix A. A similar result holds true for positive multilinear mappings. Proposition 3.11. Let Φ : Mk 2A2,· · · , A∗ 2A2,· · · , A∗ 1A1, A∗ 1A1, A∗ Φ(A∗ q → Mp be a positive multilinear map. Then kAk) ≥ Φ(A1,· · · , Ak)Φ(A∗ kAk) ≥ Φ(A∗ 1,· · · , A∗ k) k)Φ(A1,· · · , Ak) 1,· · · , A∗ Φ(A∗ for all normal matrices A1, A2,· · · , Ak. 10 M. DEHGHANI, M. KIAN AND Y. SEO Proof. Since the matrix Zµ = µ2 µ 1 ! is positive for all µ ∈ C, it follows ¯µ that λ1j1 · · · λkjk2 λ1j1 · · · λkjk λ1j1 · · · λkjk 1 ! = Zλ1j1 ◦ · · · ◦ Zλkjk j=1 λijPij is the spectral decomposition of each Ai, then is positive. If Ai =Pq 2A2,· · · , A∗ 1A1, A∗ Φ(A∗ λ1j12P1j1,· · · , λkjk2Pkjk! q Xjk=1 kAk) = Φ q Xj1=1 Xj2=1 Xj1=1 = q q · · · q Xjk=1 λ1j1λ2j2 · · · λkjk2Φ(P1j1,· · · , Pkjk) and q q q Φ(A1,· · · , Ak) = Φ(A∗ 1,· · · , A∗ k) = q · · · Xj2=2 Xj1=1 Xj2=2 Xj1=1 · · · λ1j1 · · · λkjkΦ(P1j1,· · · , Pkjk) Xjk=1 Xjk=1 λ1j1 · · · λkjkΦ(P1j1,· · · , Pkjk). q q It follows that Φ(A∗ 1A1,··· , A∗ kAk) Φ(A1,··· , Ak) 1,··· , A∗ Φ(A∗ k) 1 q = ··· Xj1=1 Lemma 3.10 then implies that ! Xjk=1(cid:16)Zλ1j1 ◦ ··· ◦ Zλkjk(cid:17) ⊗ Φ(P1j1 ··· , Pkjk ) ≥ 0. q Φ(A∗ 1A1, A∗A2,· · · , A∗ kAk) ≥ Φ(A1, A2,· · · , Ak)Φ(A∗ 1, A∗ 2,· · · , A∗ k). (cid:3) Ando [1] unified Kadison and Choi inequalities into a single form as follows: Let Φ be a strictly positive linear mapping on Mp. Then Φ(HA−1H) ≥ Φ(H)Φ(A)−1Φ(H) (3.4) whenever H is Hermitian and A > 0. Moreover, it is known a more general version of (3.4) in [4, Proposition 2.7.5]: A ≥ X ∗A−1X =⇒ Φ(A) ≥ Φ(X)∗Φ(A)−1Φ(X) (3.5) for X is arbitrary and A > 0. POSITIVE MULTILINEAR MAPPINGS 11 As another application of Theorem 3.2 we present a multilinear map version of (3.4) and (3.5). Proposition 3.12. Let Φ be a strictly positive multilinear mapping. Then Φ(H1, . . . , Hk)Φ(A1, . . . , Ak)−1Φ(H1, . . . , Hk) ≤ Φ(H1A−1 1 H1, . . . , HkA−1 k Hk) whenever Hi is Hermitian and Ai > 0 (i = 1, . . . , k). Proof. Assume that the positive multilinear mapping Ψ is defined by Ψ(Y1, . . . , Yk) = Φ(A1, . . . , Ak)− 1 2 Φ(A 1 1 1 1 2 1 Y1A 2 1 , . . . , A 2 k YkA 2 k )Φ(A1, . . . , Ak)− 1 2 . By Corollary 3.3 we have (3.6) for every Hermitian Yi. Choose Yi = A Ψ(Y 2 1 , . . . , Y 2 k ) ≥ Ψ(Y1, . . . , Yk)2 to get − 1 − 1 i HiA i 2 2 Ψ(A 2 − 1 1 H1A−1 − 1 1 H1A 1 2 , . . . , A 2 − 1 k HkA−1 k HkA − 1 Multiplying both sides by Φ(A1, . . . , Ak) 2 − 1 − 1 1 H1A 1 2 2 k ) ≥ Ψ(A 2 we conclude the desired inequality. (cid:3) , . . . , A 1 2 − 1 k HkA 2 − 1 k )2. Proposition 3.13. Let Φ be a strictly positive multilinear mapping. If Ai > 0 and Xi ∈ Mq, then i A−1 Φ(A1. . . . , Ak) ≥ Φ(X1, . . . , Xk)∗Φ(A1, . . . , Ak)−1Φ(X1, . . . , Xk). i Xi (i = 1, . . . , k) =⇒ Ai ≥X ∗ Proof. Let Ψ be the positive multilinear mapping defined by (3.6). By Theo- rem 3.11 i Yi ≤ I (i = 1, . . . , k) =⇒ Ψ(Y1, . . . , Yk)∗Ψ(Y1, . . . , Yk) ≤ I. Y ∗ Let Ai ≥ X ∗ − 1 i X ∗ i XiA A i A−1 2 2 i A−1 i Xi and put Yi = A − 1 i ≤ I. Hence − 1 1 A 1 − 1 k X ∗ , . . . , A k A 2 2 2 − 1 1 X ∗ Ψ(A 2 − 1 − 1 i XiA i 2 (i = 1, . . . , k). Then Y ∗ i Yi = 2 − 1 k )Ψ(A 2 − 1 − 1 1 X1A 1 2 , . . . , A 2 − 1 k XkA − 1 2 k ) ≤ I Multiply both sides by Φ(A1, . . . , Ak) 1 2 to obtain the result. (cid:3) 12 M. DEHGHANI, M. KIAN AND Y. SEO 4. Kantorovich inequality and convex theorems Let wi be positive scalars. If m ≤ ai ≤ M for some positive scalars m and M, then the Kantorovich inequality n Xi=1 wiai! n Xi=1 wia−1 i ! ≤ (m + M)2 4mM n Xi=1 wi!2 (4.1) holds. There are several operator version of this inequality, see e.g. [21]. For example if A is strictly positive matrix with 0 < m ≤ A ≤ M for some positive scalars m < M, then hA−1x, xi ≤ (m + M)2 4mM hAx, xi−1 (x ∈ Cp, k x k= 1). (4.2) We would like to refer the reader to [24] to find a recent survey concerning operator Kantorovich inequality. We show a matrix version of the Kantorovich inequality (4.1) including positive multilinear mappings: If t1,· · · , tn are positive scalars and 0 < m ≤ Ai ≤ M (i = 1, . . . , k), then Φ n Xi=1 tiAi, n Xi=1 tiA−1 i ! ≤ m2 + M 2 2mM Φ n Xi=1 ti, ti! . n Xi=1 (4.3) Another particular case is Φ(A, B) = A ◦ B and n = 1 which gives A ◦ A−1 ≤ m2+M 2 2mM , see [23]. First we give a more general form of (4.3). Proposition 4.1. Let Φ : M2 Ai, Bi ∈ Mq are positive matrices with m ≤ Ai, Bi ≤ M (i = 1,· · · , n), then q → Mp be a unital positive multilinear map. If Φ n Xi=1 1 1 X 2 i AiX 2 i , 1 2 i B−1 i Y Y n Xi=1 1 2 i ! + Φ n Xi=1 ≤ m2 + M 2 mM 1 X 2 i A−1 i X n 1 2 i , Xi=1 Φ n Xi=1 2 1 i ! Yi! 1 Y 2 i BiY Xi, n Xi=1 (4.4) for all positive matrices Xi, Yi. Proof. Let Ai and Bj are positive matrices with eigenvalues λi1,· · · , λiq and µj1,· · · , µjq, respectively, so that m ≤ λik ≤ M and m ≤ µjk ≤ M for every k = 1,· · · , q. Hence λikµ−1 jk + λ−1 ik µjk ≤ m2 + M 2 mM . (4.5) POSITIVE MULTILINEAR MAPPINGS 13 ℓ=1 µjℓPjℓ, then 2 i A−1 i X 2 j QjℓY 1 2 i PikX 1 1 1 1 2 2 2 2 i , Y j BjY j (cid:17) j ! + Φ q Xk=1 j (cid:17) j QjℓY i , Y 1 2 1 2 1 2 1 1 2 1 2 1 1 1 2 1 2 2 q i , i , Y λikX µ−1 jℓ Y i PikX j B−1 j Y k=1 λikPik and Bj =Pq j (cid:17) + Φ(cid:16)X Xℓ=1 ik µjℓ)Φ(cid:16)X Φ(cid:16)X jℓ + λ−1 (λikµ−1 i PikX Xk=1 Xℓ=1 q q 1 2 1 2 1 2 i AiX If Ai =Pq Φ(cid:16)X = Φ q Xk=1 Xℓ=1 Xk=1 = q q ≤ = m2 + M 2 mM m2 + M 2 mM Φ(Xi, Yj), i , Y 1 2 j QjℓY 1 2 j (cid:17) λ−1 ik X 1 1 2 i PikX 2 i , 1 µjℓY 2 j QjℓY q Xℓ=1 1 2 j ! where the inequality follows from (4.5). The multilinearity of Φ then can be applied to show (4.4). (cid:3) With the assumption as in Proposition 4.1, if we put B = A and Y = X, then we get 1 2 1 n X Φ n Xi=1 If t1,· · · , tn are positive scalars, then with Xi = ti we get (4.3). i ! ≤ Xi=1 m2 + M 2 i A−1 i AiX 2mM i X i , X 2 1 2 1 2 Xi, Φ n Xi=1 Let Φ : M2 q → Mp be defined as Example 2.5. Assume that n = 1 and X = Y = I. If 0 < m ≤ A, B ≤ M, then as another consequence of Proposition 4.1 we get another operator Kantorovich inequality as follows: Xi! . n Xi=1 hx, B−1xihAx, xi + hx, BxihA−1x, xi ≤ m2 + M 2 mM . Next we prove a convex theorem involving the positive multilinear mappings. Theorem 4.2. If Φ : M2 function q → Mp is a positive multilinear mapping, then the f (t) = Φ(cid:0)X ∗A1+tX, Y ∗B1−tY(cid:1) + Φ(cid:0)X ∗A1−tX, Y ∗B1+tY(cid:1) is convex on R and attains its minimum at t = 0 for all strictly positive matrices A, B ∈ Mq and all X, Y ∈ Mq. Proof. Assume that A, B are positive matrices in Mq with eigenvalues λ1,· · · , λq and µ1,· · · , µq, respectively. Since g(s) = xs + x−s (x > 0) is convex on R, for every i, j = 1,· · · , q and x = λiµ−1 2(λs j we have + λ−(s+t) + λ−(s−t) . (4.6) µs−t µs+t j + λs−t i µ−(s−t) i µ−s j + λ−s i µs i µ−(s+t) j j j i i j) ≤ λs+t 14 M. DEHGHANI, M. KIAN AND Y. SEO Multiply both sides of (4.6) by λiµj to obtain 2(λ1+s i µ1−s j + λ1−s i µ1+s j i ) ≤ λ1+s+t + λ1+s−t µ1−(s+t) j µ1−(s−t) j i + λ1−(s+t) µ1+s+t i j + λ1−(s−t) i µ1+s−t j . (4.7) j=1 µjQj is the spectral decomposition of A, B, respectively, then i=1 λiPi and B =Pq Now if A =Pq 2f (s) = 2(cid:2)Φ(cid:0)X ∗A1+sX, Y ∗B1−sY(cid:1) + Φ(cid:0)X ∗A1−sX, Y ∗B1+sY(cid:1)(cid:3) j (cid:17) Φ (X ∗PiX, Y ∗QjY ) 2(cid:16)λs+1 i µ−s+1 + λ−s+1 µs+1 = q q j i Xj=1 Xi=1 (by multilinearity of Φ) µ1−(s+t) j + λ1−(s+t) i µ1+s+t j + λ1+s−t i µ1−(s−t) j + λ1−(s−t) i µ1+s−t j (cid:17) q q ≤ Xi=1 Xj=1(cid:16)λ1+s+t i Φ (X ∗PiX, Y ∗QjY ) (by (4.7) and the positivity of Φ) = Φ(cid:16)X ∗A1+s+tX, Y ∗B1−(s+t)Y(cid:17) + Φ(cid:16)X ∗A1−(s+t)X, Y ∗B1+(s+t)Y(cid:17) + Φ(cid:16)X ∗A1+s−tX, Y ∗B1−(s−t)Y(cid:17) + Φ(cid:16)X ∗A1−(s−t)X, Y ∗B1+(s−t)Y(cid:17) = f (s + t) + f (s − t). Therefore f is convex on R. Since f (−t) = f (t), this together with the convexity of f implies that f attains its minimum at t = 0. (cid:3) Two special cases of the last theorem reads as follows. Corollary 4.3. Let Φ : M2 0, the function q → Mp be a positive multilinear map. For all A, B ≥ f (t) = Φ(cid:0)A1+t, B1−t(cid:1) + Φ(cid:0)A1−t, B1+t(cid:1) is decreasing on the interval [−1, 0], increasing on the interval [0, 1] and attains its minimum at t = 0. Proof. Follows from Theorem 4.2 with X = Y = I. (cid:3) Note that applying Corollary 4.3 with Φ(A, B) = A ⊗ B implies [25, Theorem 3.2]. Corollary 4.4. Let A, B ≥ 0. mapping, then the function If Φ : M2 q → Mp is a positive multilinear f (t) = Φ(cid:0)At, B1−t(cid:1) + Φ(cid:0)A1−t, Bt(cid:1) POSITIVE MULTILINEAR MAPPINGS is decreasing on [0, 1 2], increasing on [ 1 2 , 1] and attains its minimum at t = 1 2 . Proof. Replace A, B in Theorem 4.2 by A 1 2 , B 1 2 and 1+t 2 by t. 15 (cid:3) As an example, the mapping (X, Y ) → Tr(XY ) is positive and multilinear. Hence, the function f (t) = Tr(cid:0)AtB1−t(cid:1) + Tr(cid:0)A1−tBt(cid:1) 2], increasing on [ 1 2, 1] and has a minimum at t = 1 2. In other is decreasing on [0, 1 words 1 2 B 2Tr(cid:16)A 1 2(cid:17) ≤ Tr(cid:0)AtB1−t(cid:1) + Tr(cid:0)A1−tBt(cid:1) ≤ Tr(A) + Tr(B) for all t ∈ [0, 1] and all positive matrices A and B. 5. Reverse of the Choi–Davis–jensen inequality In this section, we consider reverse type inequalities of the Choi–Davis–Jensen inequality for positive multilinear mappings. The following lemma is well-known. Lemma 5.1. [23] Let Ai be strictly positive matrices with m ≤ Ai ≤ M for some scalars 0 < m < M and let Ci i Ci = I. If a real valued continuous function f on [m, M] is convex (resp. concave) and (i = 1, . . . , k) be matrices with Pk i=1 C ∗ f (t) > 0 on [m, M], then C ∗ k Xi=1 C ∗ i f (Ai)Ci ≤ α(m, M, f )f k Xi=1 i AiCi! ≤ Xi=1 k i AiCi! i f (Ai)Ci,! C ∗ where C ∗ resp. β(m, M, f )f k Xi=1 f (t)(cid:18) f (M ) − f (m) M − m α(m, M, f ) = max(cid:26) 1 (t − m) + f (m)(cid:19) ; m ≤ t ≤ M(cid:27) (5.1) (cid:18)resp. β(m, M, f ) = min(cid:26) 1 f (t)(cid:18) f (M ) − f (m) M − m (t − m) + f (m)(cid:19) ; m ≤ t ≤ M(cid:27)(cid:19) . (5.2) Theorem 5.2. Let Ai be positive matrices with m ≤ Ai ≤ M for some scalars 0 < m < M (i = 1, . . . , k). Let Φ be a unital positive multilinear map. If a 16 M. DEHGHANI, M. KIAN AND Y. SEO real valued continuous function f is super-multiplicative convex on [m, M] (resp. sub-multiplicative concave), then Φ(f (A1), . . . , f (Ak)) ≤ α(mk, M k, f )f (Φ(A1, . . . , Ak)) (resp. β(mk, M k, f )f (Φ(A1, . . . , Ak)) ≤ Φ(f (A1), . . . , f (Ak)), where α(m, M, f ) and β(m, M, f ) are defined by (5.1) and (5.2), respectively. Proof. We only prove the super-multiplicative convex case. Assume that Ai = Pqi j=1 λijPij (i = 1,· · · , k) is the spectral decomposition of each Ai for which Pqi j=1 Pij = I. Then Φ(f (A1), f (A2),··· , f (Ak)) = Φ Xj=1  Xj1 Xj2 f (λ1j1)f (λ2j2)··· f (λkjk)Φ(P1j1 ,··· , Pkjk ), f (λkj)Pkj  f (λ1j)P1j ,··· , Xjk Xj=1 (5.3) ··· = qk qk q1 q1 q2 1 2 we get q1 q2 q1 q2 qk qk = · · · Xjk Xjk Xj2 Xj2 f (λ1j1)f (λ2j2) · · · f (λkjk)Φ(P1j1,· · · , Pkjk) by multilinearity of Φ. With C(j1,· · · , jk) := (Φ(P1j1,· · · , Pkjk)) Φ(f (A1),· · · , f (Ak)) Xj1 Xj1 ≤ ≤ α(mk, M k, f )f q1 Xjk Xj1 = α(mk, M k, f )f q1 Xjk Xj1 = α(mk, M k, f )f (Φ(A1,· · · , Ak)) . C(j1,· · · , jk)f (λ1j1λ2j2 · · · λkjk)C(j1,· · · , jk) Xj2 Xj2 C(j1,· · · , jk)λ1j1λ2j2 · · · λkjkC(j1,· · · , jk)! λ1j1λ2j2 · · · λkjkΦ(P1j1,· · · , Pkjk)! · · · · · · · · · qk qk q2 q2 Corollary 5.3. Let Ai (i = 1, . . . , k) be positive matrices with m ≤ Ai ≤ M for some scalars 0 < m < M. Let Φ be a unital positive multilinear map. If r 6∈ [0, 1], then (cid:3) Φ(Ar 1, . . . , Ar k) ≤ K(mk, M k, r) Φ(A1, . . . , Ak)r. POSITIVE MULTILINEAR MAPPINGS 17 If r ∈ [0, 1], then the inequality is reversed. Here the generalized Kantorovich constant K(m, M, r) is defined by K(m, M, r) = mM r − Mmr (r − 1)(M − m)(cid:18)r − 1 r mM r − Mmr(cid:19)r M r − mr . (5.4) In particular, and Φ(A2 1, . . . , A2 k) ≤ Φ(A−1 1 , . . . , A−1 k ) ≤ (M k + mk)2 4M kmk Φ(A1, . . . , Ak)2 (M k + mk)2 4M kmk Φ(A1, . . . , Ak)−1. Theorem 5.4. Let σ be a matrix mean with the representing function f which is sub-multiplicative and not affine. Let Ai, Bi be positive matrices with m ≤ Ai, Bi ≤ M for some scalars 0 < m < M and i = 1, . . . , k. Let Φ be a strictly positive unital multilinear map. Then Φ(A1σB1, . . . , AkσBk) ≥ β(cid:18) mk M k , M k mk , f(cid:19) Φ(A1, . . . , Ak) σ Φ(B1, . . . , Bk) where β(m, M, f ) is defined by (5.2). In particular, for each α ∈ [0, 1] Φ(A1♯αB1, . . . , Ak♯αBk) ≥ K(cid:18) mk M k , mk , α(cid:19) Φ(A1, . . . , Ak) ♯α Φ(B1, . . . , Bk) M k where K(m, M, α) is defined by (5.4). Proof. Let Ψ be a strictly positive multilinear map defined by Ψ(X1, . . . , Xk) = Φ(A1, . . . , Ak)− 1 2 Φ(A 1 1 1 1 2 1 X1A 2 1 , . . . , A 2 k XkA 2 k )Φ(A1, . . . , Ak)− 1 2 −1 −1 and Yi := A 2 i BiA i 2 (i = 1,· · · k). Then it follows from m/M ≤ Yi ≤ M/m that 2 Φ(B1,· · · , Bk)Φ(A1,· · · , Ak) −1 2 (cid:17) −1 M k M k β(cid:18) mk mk , f(cid:19) f(cid:16)Φ(A1,· · · , Ak) M k , = β(cid:18) mk mk , f(cid:19) f(cid:0)Ψ(Y1,· · · , Yk)(cid:1) M k , ≤ Ψ (f (Y1),· · · , f (Yk)) = Φ(A1,· · · , Ak) −1 (by Theorem 5.2) 2 Φ(A1σB1,· · · , AkσBk)Φ(A1,· · · , Ak) −1 2 . Multiplying both sides by Φ(A1,· · · , Ak) 1 2 we get the desired result. (cid:3) Utilizing Lemma 5.1, we present a reverse additivity inequality for matrix means. 18 M. DEHGHANI, M. KIAN AND Y. SEO Lemma 5.5. Let σ be a matrix mean with the representing function f , and A, B, C, D positive matrices with 0 < m ≤ A, B, C, D ≤ M for some scalars m ≤ M. Then , M m β(cid:18) m , f(cid:19) [(A + B) σ (C + D)] ≤ A σ C + B σ D where β(m, M, f ) is defined by (5.2). In particular, for each α ∈ [0, 1] M K(m/M, M/m, α) [(A + B) ♯α (C + D)] ≤ A ♯α C + B ♯α D and for α = 1/2 2 4√Mm √M + √m [(A + B) ♯ (C + D)] ≤ A ♯ C + B ♯ D Proof. Put X = A1/2(A + B)−1/2, Y = B1/2(A + B)−1/2, V = A−1/2CA−1/2 and W = B−1/2DB−1/2. Since m/M ≤ V, W ≤ M/m and X ∗X + Y ∗Y = I, it follows from Lemma 5.1 that (A + B) σ (C + D) = (A + B)1/2f (X ∗V X + Y ∗W Y )(A + B)1/2 ≤ β(m/M, M/m, f )−1(A + B)1/2 [X ∗f (V )X + Y ∗f (W )Y ] (A + B)1/2 = β(m/M, M/m, f )−1 (A σ C + B σ D) . (cid:3) We show a reverse inequality for positive multilinear maps and matrix means in Corollary 3.7: Theorem 5.6. Let σ be a matrix mean with the representing function f , and A, B, C, D positive matirces with 0 < m ≤ A, B, C, D ≤ M for some scalars m ≤ M. Let Φ be a unital positive multilinear map. Then Φ(A, B) + Φ(B, A) ≤ β(cid:18) m2 M 2 , M 2 m2 , f(cid:19)−2 (cid:2)Φ(A σ B, A σ0 B) + Φ(A σ0 B, A σ B)(cid:3) where β(m, M, f ) is defined by (5.2). POSITIVE MULTILINEAR MAPPINGS 19 Proof. Since m2 ≤ Φ(A, B), Φ(B, A) ≤ M 2, it follows that Φ(A, B) + Φ(B, A) = [Φ(A, B) + Φ(B, A)] σ [Φ(B, A) + Φ(A, B)] [Φ(A, B) σ Φ(B, A) + Φ(B, A) σ Φ(A, B)] (by Lemma 5.5) ≤ β(cid:18) m2 M 2 , ≤ β(cid:18) m2 M 2 , M 2 m2 , f(cid:19)−1 m2 , f(cid:19)−2 M 2 Xi=1 k Xi=1 (cid:2)Φ(A σ B, A σ0 B) + Φ(A σ0 B, A σ B)(cid:3) , where the last inequality follows from Theorem 5.4. (cid:3) 6. Information monotonicity for Karcher mean In Proposition 3.6, we showed the positive multilinear map version of Ando’s inequality (3.3) due to matrix means. In this section, we consider its n-variable positive multilinear map version. For this, we recall the notion of the Karcher mean and matrix power mean: Let A = (A1, . . . , Ak) be a k-tuple of strictly positive matrices and ω = (ω1, . . . , ωk) a weight vector such that ωi ≥ 0 and Pk i=1 ωi = 1. The Karcher mean of A1, . . . , Ak is the unique positive invertible solution of the Karcher equation k ωi log(X −1/2AiX −1/2) = 0 and we denote it by GK(ω; A) = G(ω; A). The matrix power mean of A1, . . . , Ak is the unique positive invertible solution of a non-linear matrix equation X = ωi(X ♯t Ai) for t ∈ (0, 1] and we denote it by Pt(ω; A). For t ∈ [−1, 0), we define Pt(ω; A) := P−t(ω; A−1)−1, where A−1 = (A−1 k ). The matrix power mean satisfies all desirable prop- erties of power arithmetic means of positive real numbers and interpolates be- 1 , . . . , A−1 tween the weighted harmonic and arithmetic means. Moreover, the Karcher mean coincides with the limit of matrix power means as t → 0. The matrix power mean satisfies an information monotonicity: For each t ∈ (0, 1] Φ(Pt(ω; A)) ≤ Pt(ω; Φ(A)) (6.1) for any unital positive linear map Φ. For more details on the Karcher mean and the matrix power mean; see [4, 5, 18, 19]. 20 M. DEHGHANI, M. KIAN AND Y. SEO In the final section, we show a positive multilinear map version of (6.1). Let k weight vectors k ) k-tuples of strictly positive matrices for i = 1, . . . , n. Φ be a unital positive multilinear map and ω(i) = (ω(i) and A(i) = (A(i) Define 1 , . . . , A(i) 1 , . . . , ω(i) Φ(A(1), . . . , A(n)) := (Φ(A(1) k , A(2) 1 , A(2) k , . . . , A(n) Φ(A(1) k )) 1 , . . . , A(n) 1 ), Φ(A(1) 2 , A(2) 2 , . . . , A(n) 2 ), . . . , and ω(1) · · · ω(n) := (ω(1) 1 ω(2) 1 · · · ω(n) 1 , ω(1) 2 ω(2) 2 · · · ω(n) 2 , . . . , ω(1) k ω(2) k · · · ω(n) k ). Theorem 6.1. Let Φ be a unital positive multilinear map. If t ∈ (0, 1], then Φ(Pt(ω(1); A(1)), . . . , Pt(ω(n); A(n))) ≤ Pt(ω(1)ω(2) · · · ω(n); Φ(A(1), A(2), . . . , A(n))). If t ∈ [−1, 0) and 0 < m ≤ A(i) j ≤ M for j = 1, . . . , k and i = 1, . . . , n and some scalars 0 < m < M, then Pt(ω(1)ω(2) · · ·ω(n); Φ(A(1), A(2), . . . , A(n))) (M n + mn)2 ≤ 4M nmn Φ(Pt(ω(1); A(1)), . . . , Pt(ω(n); A(n))). Proof. We only prove the case of n = 2: Let A = (A1, . . . , Ak), B = (B1, . . . , Bk) be k-tuples of positive matrices and ω = (ω1, . . . , ωk), ω′ = (ω′ k) weight 1, . . . , ω′ vectors. Put Xt = Pt(ω; A) and Yt = Pt(ω′; B). Then we have Xt =P ωi(Xt ♯t Ai and Yt =P ω′ i(Yt ♯t Bi). By Proposition 3.6, it follows that k Φ(Xt, Yt) = ≤ Xij=1 Xij=1 k ωiω′ jΦ(Xt ♯t Ai, Yt ♯t Bj) ωiω′ jΦ(Xt, Yt) ♯t Φ(Ai, Bj). (6.2) ij=1 ωiω′ Define a map f (X) = Pk jX ♯t Φ(Ai, Bj) and then it follows from [19, Theorem 3.1] that f n(X) → Pt(ω(2); Φ(A, B)) for all X > 0 as n → ∞. On the other hand, Φ(Xt, Yt) ≤ f (Φ(Xt, Yt)) by (6.2). This follows that Φ(Xt, Yt) ≤ f n(Φ(Xt, Yt)) for any n. Hence as n → ∞ we have the desired inequality for t ∈ (0, 1]. POSITIVE MULTILINEAR MAPPINGS 21 Let t ∈ [−1, 0). By Corollary 5.3, Φ(A−1) ≤ (M k+mk)2 4M kmk Φ(A)−1 for all k-tuple of strictly positive matrices. Therefore, Φ(P−t(ω; A−1), P−t(ω′; B−1)) ≤ P−t(ωω′; Φ(A−1, B−1)) (M 2 + m2)2 ≤ P−t(ωω′; = 4M 2m2 Φ(A, B)−1) 4M 2m2 P−t(ωω′; Φ(A, B)−1) (M 2 + m2)2 and this implies Pt(ωω′; Φ(A, B)) = P−t(ωω′; Φ(A, B)−1)−1 ≤ ≤ = (M 2 + m2)2 (M 2 + m2)2 4M 2m2 Φ(P−t(ω; A−1), P−t(ω′; B−1))−1 4M 2m2 Φ(P−t(ω; A−1)−1, P−t(ω′; B−1) 4M 2m2 Φ(Pt(ω; A), Pt(ω′; B)). (M 2 + m2)2 Theorem 6.2. If Φ is unital positive multilinear map, then (cid:3) Φ(G(ω(1); A(1)), . . . , G(ω(n); A(n))) ≤ G(ω(1) · · · ω(n); Φ(A(1), . . . , A(n))) ≤ Proof. By Theorem 6.1, (M n + mn)2 4M nmn Φ(G(ω(1); A(1)), . . . , G(ω(n); A(n))). Φ(Pt(ω(1); A(1)), . . . , Pt(ω(n); A(n))) ≤ Pt(ω(1) · · · ω(n); Φ(A(1), . . . , A(n))) for all t ∈ (0, 1]. As t → 0, we have Φ(G(ω(1); A(1)), . . . , G(ω(n); A(n))) ≤ G(ω(1) · · · ω(n); Φ(A(1), . . . , A(n))). Also, since (M n + mn)2 4M nmn Φ(Pt(ω(1); A(1)), . . . , Pt(ω(n); A(n))) ≥ Pt(ω(1) · · · ω(n); Φ(A(1), . . . , A(n))) for all t ∈ [−1, 0), as t → 0, we have (M n + mn)2 4M nmn Φ(G(ω(1); A(1)), . . . , G(ω(n); A(n))) ≥ G(ω(1) · · · ω(n); Φ(A(1), . . . , A(n))). (cid:3) 22 M. DEHGHANI, M. KIAN AND Y. SEO Remark 6.3. Let Φ be a unital positive linear map. Then for each t ∈ [−1, 0) Pt(ω; Φ(A)) ≤ (M + m)2 4Mm Φ(Pt(ω; A)) and this is a complementary result of [18, Proposition 3.6 (13)]. References [1] T. Ando, Concavity of certain maps on positive definite matrices and applications to Hadamard products, Linear Algebra Appl., 26 (1979), 203–241. [2] K. M.R. Audenaert, F. Hiai, On matrix inequalities between the power means: Counterex- amples, Linear Algebra Appl., 439 (2013), 1590–1604. [3] J.S.Aujla and H.L.Vasudeva, Inequalities involving Hadamard product and operator means, Math. Japon., 42 (1995), 265–272. [4] R. Bhatia, Positive Definite Matrices, Princeton University Press, Princeton, 2007. [5] R. Bhatia and J. Holbrook, Riemannian geometry and matrix means, Linear Algebra Appl., 413 (2006), 594–618. [6] R. Bhatia, P. Grover, T. Jain, Derivatives of tensor powers and their norms, Electron. J. Linear Algebra 26 (2013), 604–619. [7] R. Bhatia and R. Sharma, Some inequalities for positive linear maps, Linear Algebra Appl. 436 (2012), 1562–1571. [8] J. -C. Bourin and E. Ricard, An asymmetric Kadison’s inequality, Linear Algebra Appl. 433 (2010), 499–510. [9] M. D. Choi, A Schwarz inequality for positive linear maps on C ∗-algebras, Illinois J. Math. 18 (1974), 565–574. [10] C. D`avis, A Schwarz inequality for convex operator functions, Proc. Amer. Math. Soc, 8 (1957), 42–44. [11] M. Fiedler, Uber eine Ungleichung fur positiv definite Matrizen, Math. Nachr., 23 (1961), 197–199. [12] T. Furuta, H. Mi´ci´c, J. Pecari´c and Y. Seo, Mond–Pecaric Method in Operator Inequalities, Zagreb, Element, 2005. [13] T. Furuta, Around Choi inequalities for positive linear maps, Linear Algebra Appl. 434 (2011), 14–17. [14] R. A. Horn, C. R. Johnson, Topics in Matrix Analysis, Cambridge Univ Press, New York, 1991. [15] R. Kaur, M. Singh and J. S. Aujla, Generalized matrix version of reverse Hlder inequality, Linear Algebra Appl. 434 (2011), 636–640. [16] M. Kian and M. S. Moslehian, Refinements of the operator Jensen–Mercer inequality, Elec- tron. J. Linear Algebra 26 (2013), 742–753. [17] F. Kubo and T. Ando, Means of positive linear operators, Math. Ann. 246 (1980), 205–224. POSITIVE MULTILINEAR MAPPINGS 23 [18] J. Lawson and Y. Lim, Karcher means and Karcher equations of positive definite operators, Trans. Amer. Math. Soc., Series B, 1 (2014), 1–22. [19] Y. Lim and M. P´alfia, Matrix power means and the Karcher mean, J. Funct. Anal., 262(2012), 1498–1514. [20] M. Lin, On an operator Kantorovich inequality for positive linear maps, J. Math. Anal. Appl. 402 (2013), 127–132. [21] J. S. Matharu, J. S. Aujla, Hadamard product versions of the Chebyshev and Kantorovich inequalities, J. Inequal. Pure Appl. Math. 10 (2009), 1–6. [22] A. Matkovi´c, J. Pecari´c, I. Peri´c, A variant of Jensen’s inequality of Mercer’s type for operators with applications, Linear Algebra Appl. 418 (2006), 551–564. [23] H. Mi´ci´c, J. Pecari´c and Y. Seo, Complementary inequalities to inequalities of Jensen and Ando based on the Mond–Pecari´c method, Linear Algebra Appl. 318 (2000), 87–107. [24] M. S. Moslehian, Recent developments of the operator Kantorovich inequality, Expo. Math. 30 (2012), 376–388. [25] M. S. Moslehian, J. S. Matharu and J. S. Aujla, Non-commutative Callebaut inequality, Linear Algebra Appl. 436 (2012), 3347–3353. [26] M. Niezgoda, Choi–Davis–Jensen’s inequality and generalized inverses of linear operators, Electron. J. Linear Algebra 26 (2013), 406–416. Mehdi Dehghani: Department of Pure Mathematics, Faculty of Mathematical Sciences, University of Kashan, Kashan, Iran E-mail address: [email protected] Mohsen Kian: Department of Mathematics, Faculty of Basic Sciences, Uni- versity of Bojnord, P. O. Box 1339, Bojnord 94531, Iran E-mail address: [email protected] and [email protected] Yuki Seo: Department of Mathematics Education, Osaka Kyoiku University, Asahigaoka Kashiwara Osaka 582-8582, Japan. E-mail address: [email protected]
1602.00013
4
1602
2016-06-13T07:12:23
Inverse Function Theorems for Generalized Smooth Functions
[ "math.FA" ]
Generalized smooth functions are a possible formalization of the original historical approach followed by Cauchy, Poisson, Kirchhoff, Helmholtz, Kelvin, Heaviside, and Dirac to deal with generalized functions. They are set-theoretical functions defined on a natural non-Archimedean ring, and include Colombeau generalized functions (and hence also Schwartz distributions) as a particular case. One of their key property is the closure with respect to composition. We review the theory of generalized smooth functions and prove both the local and some global inverse function theorems.
math.FA
math
Inverse Function Theorems for Generalized Smooth Functions Paolo Giordano and Michael Kunzinger 1 Introduction Since its inception, category theory has underscored the importance of unrestricted composition of morphisms for many parts of mathematics. The closure of a given space of "arrows" with respect to composition proved to be a key foundational prop- erty. It is therefore clear that the lack of this feature for Schwartz distributions has considerable consequences in the study of differential equations [28, 14], in math- ematical physics [4, 6, 8, 10, 18, 21, 25, 41, 42, 43, 46], and in the calculus of variations [30], to name but a few. On the other hand, Schwartz distributions are so deeply rooted in the linear framework that one can even isomorphically approach them focusing only on this aspect, opting for a completely formal/syntactic viewpoint and without requiring any functional analysis, see [49]. So, Schwartz distributions do not have a notion of pointwise evaluation in general, and do not form a category, although it is well known that certain subclasses of distributions have meaningful notions of pointwise evaluation, see e.g. [35, 36, 47, 45, 16, 15, 51]. This is even more surprising if one takes into account the earlier historical gene- sis of generalized functions dating back to authors like Cauchy, Poisson, Kirchhoff, Helmholtz, Kelvin, Heaviside, and Dirac, see [29, 33, 34, 50]. For them, this "gen- eralization" is simply accomplished by fixing an infinitesimal or infinite parameter in an ordinary smooth function, e.g. an infinitesimal and invertible standard devia- tion in a Gaussian probability density. Therefore, generalized functions are thought of as some kind of smooth set-theoretical functions defined and valued in a suitable 6 1 0 2 n u J 3 1 ] . A F h t a m [ 4 v 3 1 0 0 0 . 2 0 6 1 : v i X r a Paolo Giordano University [email protected] Vienna, of Oskar-Morgenstern-Platz Michael Kunzinger University [email protected] Vienna, of Oskar-Morgenstern-Platz 1, 1090 Wien e-mail: 1, 1090 Wien e-mail: 1 2 Paolo Giordano and Michael Kunzinger non-Archimedean ring of scalars. From this intuitive point of view, they clearly have point values and form a category. This aspect also bears upon the concept of (a generalized) solution of a differ- ential equation. In fact, any theory of generalized functions must have a link with the classical notion of (smooth) solution. However, this classical notion is deeply grounded on the concept of composition of functions and, at the same time, it is often too narrow, as is amply demonstrated e.g. in the study of PDE in the presence of singularities. In our opinion, it is at least not surprising that also the notion of distributional solution did not lead to a satisfying theory of nonlinear PDE (not even of singular ODE). We have hence a wild garden of flourishing equation-dependent techniques and a zoo of counter-examples. The well-known detaching between these techniques and numerical solutions of PDE is another side of the same question. One can say that this situation presents several analogies with the classical compass-and-straightedge solution of geometrical problems, or with the solution of polynomial equations by radicals. The distinction between algebraic and irrational numbers and the advent of Galois theory were essential steps for mathematics to start focusing on a different concept of solution, frequently nearer to applied prob- lems. In the end, these classical problems stimulated more general notions of geo- metrical transformation and numerical solution, which nowadays have superseded their origins. The analogies are even greater when observing that first steps toward a Galois theory of nonlinear PDE are arising, see [5, 7, 38, 39]. Generalized smooth functions (GSF) are a possible formalization of the original historical approach of the aforementioned classical authors. We extend the field of r eRn, Y ⊆ real numbers into a natural non Archimedean ring r eR and we consider the simplest notion of smooth function on the extended ring of scalars r eR. To define a GSF r eRd, we simply require the minimal logical conditions f : X −→ Y , X ⊆ e ⊆ Rn, defines a set- so that a net of ordinary smooth functions fe ∈ C theoretical map X −→ Y which is infinitely differentiable; see below for the details. This freedom in the choice of domains and codomains is a key property to prove that GSF are closed with respect to composition. As a result, GSF share so many properties with ordinary smooth functions that frequently we only have to formally generalize classical proofs to the new context. This allows an easier approach to this new theory of generalized functions. (W e , Rd), W It is important to note that the new framework is richer than the classical one because of the possibility to express non-Archimedean properties. So, e.g., two dif- ferent infinitesimal standard deviations in a Gaussian result in infinitely close Dirac- delta-like functionals but, generally speaking, these two GSF could have different infinite values at infinitesimal points h ∈ are embedded as GSF, but this embedding is not intrinsic and it has to be chosen depending on the physical problem or on the particular differential equation we aim to solve. r eR. For this reason, Schwartz distributions In the present work, we establish several inverse function theorems for GSF. We prove both the classical local and also some global versions of this theorem. It is remarkable to note that the local version is formally very similar to the classical one, but with the sharp topology instead of the standard Euclidean one. We also show ¥ Inverse Function Theorems for Generalized Smooth Functions 3 the relations between our results and the inverse function theorem for Colombeau functions established by using the discontinuous calculus of [2, 3]. The paper is self-contained in the sense that it contains all the statements of results required for the proofs of the new inverse function theorems. If proofs of preliminaries are omitted, we give references to where they can be found. 1.1 Basic notions The ring of generalized scalars In this work, I denotes the interval (0,1] ⊆ R and we will always use the variable e for elements of I; we also denote e -dependent nets x ∈ RI simply by (xe ). By N we denote the set of natural numbers, including zero. We start by defining the non-Archimedean ring of scalars that extends the real field R. For all the proofs of results in this section, see [19, 18]. Definition 1. Let r = (r e ) ∈ RI be a net such that lime →0+ r e = 0+, then (i) I (r ) := {(r −ae (ii) ) a ∈ R>0} is called the asymptotic gauge generated by r . : P(e ) to denote If P(e ) is a property of e ∈ I, we use the notation ∀0e ∃e 0 ∈ I ∀e ∈ (0,e 0] : P(e ). We can read ∀0e as for e small. I (r ) : xe = O(Je ) as e → 0+. (iii) We say that a net (xe ) ∈ RI is r -moderate, and we write (xe ) ∈ Rr (iv) Let (xe ), (ye ) ∈ RI, then we say that (xe ) ∼r (ye ) if ∀(Je ) ∈ I (r ) : xe = ) as e → 0+. This is an equivalence relation on the ring Rr of ye + O(J−1e moderate nets with respect to pointwise operations, and we can hence define if ∃(Je ) ∈ which we call Robinson-Colombeau ring of generalized numbers, [48, 8]. We denote the equivalence class x ∈ r eR := Rr / ∼r , r eR simply by x =: [xe ] := [(xe )]∼ ∈ r eR. := [r e ] ∈ In the following, r will always denote a net as in Def. 1, and we will use the simpler [18]. We also use the notation dr r eR and de := [e ] ∈ (e )eR. class of differential equations we need to solve for the generalized functions we are (ze ) ∈ RI such that (ze ) ∼r 0 (we then say that (ze ) is r -negligible) and xe ≤ ye + ze for e small. Equivalently, we have that x ≤ y if and only if there exist representatives notation eR for the case r e = e . The infinitesimal r can be chosen depending on the going to introduce, see [20]. For motivations concerning the naturality of r eR, see We can also define an order relation on r eR by saying [xe ] ≤ [ye ] if there exists (xe ), (ye ) of x, y such that xe ≤ ye for all e . Clearly, r eR is a partially ordered ring. The usual real numbers r ∈ R are embedded in r eR considering constant nets [r] ∈ r eR. r eR. Our notations for intervals Even if the order ≤ is not total, we still have the possibility to define the infimum [xe ] ∧ [ye ] := [min(xe ,ye )], and analogously the supremum function [xe ] ∨ [ye ] := [max(xe ,ye )] and the absolute value [xe ] := [xe ] ∈ 4 Paolo Giordano and Michael Kunzinger are: [a,b] := {x ∈ [x,y] := {x + r · (y − x) r ∈ [0,1]} ⊆ x ≈ y to denote that x − y is an infinitesimal number, i.e. x − y ≤ r for all r ∈ R>0. This is equivalent to lime →0+ xe − ye = 0 for all representatives (xe ), (ye ) of x, y. r eR a ≤ x ≤ b}, [a,b]R := [a,b]∩ R, and analogously for segments r eRn and [x,y]Rn = [x,y] ∩ Rn. Finally, we write norm, i.e. [xe ] := [xe ] ∈ Topologies on r eRn On the r eR-module r eRn, we can consider the natural extension of the Euclidean r eRn. Even if this generalized norm takes values in r eR, it shares several properties with usual norms, like the triangular inequality or the property y · x = y · x. It is therefore natural to consider on r eRn topologies generated by balls defined by this generalized norm and suitable notions of being "strictly less than a given radius": r eR, where [xe ] ∈ (i) We write x < y if ∃r ∈ (ii) We write x <R y if ∃r ∈ R>0 : r ≤ y − x. Definition 2. Let c ∈ (iii) Br(c) :=nx ∈ r(c) :=nx ∈ (iv) BF (v) BE r eR, then: r eR≥0 : r is invertible, and r ≤ y − x r eRn and x, y ∈ r eRn x − c < ro for each r ∈ r eR>0. r eRn x − c <R ro for each r ∈ R>0. r (c) := {x ∈ Rn x − c < r}, for each r ∈ R>0, denotes an ordinary Eu- clidean ball in Rn. r(c) r ∈ R>0, c ∈ r eR>0, c ∈ The relations <, <R have better topological properties as compared to the usual strict order relation a ≤ b and a 6= b (that we will never use) because both the sets r eRno andnBF of ballsnBr(c) r ∈ r eRno are bases for two topologies on r eRn. The former is called sharp topology, whereas the latter is called Fermat topology. We will call sharply open set any open set in the sharp topology, and large open set any open set in the Fermat topology; clearly, the latter is coarser than the former. The existence of infinitesimal neighborhoods implies that the sharp topology induces the discrete topology on R. This is a necessary result when one has to deal with continuous generalized functions which have infinite derivatives. In fact, if f ′(x0) is infinite, only for x ≈ x0 we can have f (x) ≈ f (x0). The following result is useful to deal with positive and invertible generalized numbers (cf. [24, 40]). Lemma 1. Let x ∈ r eR. Then the following are equivalent: x is invertible and x ≥ 0, i.e. x > 0. (i) (ii) For each representative (xe ) ∈ Rr of x we have ∀0e : xe > 0. (iii) For each representative (xe ) ∈ Rr of x we have ∃m ∈ N ∀0e : xe > r me Inverse Function Theorems for Generalized Smooth Functions 5 Internal and strongly internal sets A natural way to obtain sharply open, closed and bounded sets in r eRn is by using a net (Ae ) of subsets Ae ⊆ Rn. We have two ways of extending the membership relation xe ∈ Ae to generalized points [xe ] ∈ Definition 3. Let (Ae ) be a net of subsets of Rn, then r eR: (i) net (Ae ). See [44] for the introduction and an in-depth study of this notion. [Ae ] :=n[xe ] ∈ (ii) Let (xe ) be a net of points of Rn, then we say that xe ∈e Ae , and we read it as (xe ) strongly belongs to (Ae ), if ∀0e : xe ∈ Ae and if (x′e ) ∼r (xe ), then also x′e ∈ Ae for e small. Moreover, we set hAe i :=n[xe ] ∈ r eRn ∀0e : xe ∈ Aeo is called the internal set generated by the r eRn xe ∈e Aeo, and r eR>0 such that K ⊆ Br(0). Analogously, a net (Ae ) is sharply bounded if (iii) Finally, we say that the internal set K = [Ae ] is sharply bounded if there exists we call it the strongly internal set generated by the net (Ae ). r ∈ there exists r ∈ Therefore, x ∈ [Ae ] if there exists a representative (xe ) of x such that xe ∈ Ae for e small, whereas this membership is independent from the chosen representative in the case of strongly internal sets. Note explicitly that an internal set generated by a constant net Ae = A ⊆ Rn is simply denoted by [A]. r eR>0 such that [Ae ] ⊆ Br(0). The following theorem shows that internal and strongly internal sets have dual topological properties: Theorem 1. For e ∈ I, let Ae ⊆ Rn and let xe ∈ Rn. Then we have (i) [xe ] ∈ [Ae ] if and only if ∀q ∈ R>0 ∀0e : d(xe ,Ae ) ≤ r qe . Therefore [xe ] ∈ [Ae ] if and only if [d(xe ,Ae )] = 0 ∈ [xe ] ∈ hAe i if and only if ∃q ∈ R>0 ∀0e : d(xe ,Ace ) > r qe , where Ace := Rn \ Ae . Therefore, if (d(xe ,Ace )) ∈ Rr , then [xe ] ∈ hAe i if and only if [d(xe ,Ace )] > 0. [Ae ] is sharply closed and hAe i is sharply open. [Ae ] = [cl (Ae )], where cl (S) is the closure of S ⊆ Rn. On the other hand hAe i = hint(Ae )i, where int (S) is the interior of S ⊆ Rn. r eR. (ii) (iii) (iv) We will also use the following: Lemma 2. Let (W net such that [Be ] ⊆ hW e i, then e ) be a net of subsets in Rn for all e , and (Be ) a sharply bounded ∀0e : Be ⊆ W e . Sharply bounded internal sets (which are always sharply closed by Thm. 1 (iii)) serve as compact sets for our generalized functions. For a deeper study of this type of sets in the case r = (e ) see [44, 17]; in the same particular setting, see [19] and references therein for (strongly) internal sets. 6 Paolo Giordano and Michael Kunzinger Generalized smooth functions For the ideas presented in this section, see also e.g. [19, 18]. deviation. If we denote this probability density by f (x,s ), and if we set s = [s e ] ∈ Using the ring r eR, it is easy to consider a Gaussian with an infinitesimal standard r eR>0, where s ≈ 0, we obtain the net of smooth functions ( f (−,s e ))e ∈I. This is r eRd be arbitrary subsets of generalized points. Definition 4. Let X ⊆ Then we say that the basic idea we develop in the following r eRn and Y ⊆ f : X −→ Y is a generalized smooth function if there exists a net of functions fe ∈ C hW The space of GSF from X to Y is denoted by r e i, f ([xe ]) = [ fe (xe )] ∈ Y and (¶ a fe (xe )) ∈ Rd G C (X,Y ). (W e , Rd) defining f in the sense that X ⊆ r for all x = [xe ] ∈ X and all a ∈ Nn. Let us note explicitly that this definition states minimal logical conditions to obtain a set-theoretical map from X into Y and defined by a net of smooth functions. In particular, the following Thm. 2 states that the equality f ([xe ]) = [ fe (xe )] is mean- ingful, i.e. that we have independence from the representatives for all derivatives [xe ] ∈ X 7→ [¶ a fe (xe )] ∈ e , Rd) be a net of smooth functions that defines a generalized smooth r eRd be arbitrary subsets of generalized points. r eRd, a ∈ Nn. r eRn and Y ⊆ (W Theorem 2. Let X ⊆ Let fe ∈ C map of the type X −→ Y , then ∀a ∈ Nn ∀(xe ), (x′e ) ∈ Rnr : [xe ] = [x′e ] ∈ X ⇒ (¶ a ue (xe )) ∼r (¶ a ue (x′e )). (i) (ii) ∀[xe ] ∈ X ∀a ∈ Nn ∃q ∈ R>0 ∀0e : supy∈BEe q (xe ) ¶ a ue (y) ≤ e −q. (iii) For all a ∈ Nn, the GSF g : [xe ] ∈ X 7→ [¶ a in the sharp topology, i.e. each x ∈ X possesses a sharp neighborhood U such that g(x) − g(y) ≤ Lx − y for all x, y ∈ U and some L ∈ fe (xe )] ∈ eRd is locally Lipschitz (X,Y ) is continuous with respect to the sharp topologies in- r eR. (iv) Each f ∈ G C duced on X, Y . r (v) Assume that the GSF f is locally Lipschitz in the Fermat topology and that its Lipschitz constants are always finite: L ∈ R. Then f is continuous in the Fermat topology. f : X −→ Y is a GSF if and only if there exists a net ve ∈ C a generalized smooth map of type X −→ Y such that f = [ve (−)]X . (Rn, Rd) defining (vi) (vii) Subsets S ⊆ r eRs with the trace of the sharp topology, and generalized smooth maps as arrows form a subcategory of the category of topological spaces. We will call this category r , the category of GSF. G C The differential calculus for GSF can be introduced showing existence and unique- ness of another GSF serving as incremental ratio. For its statement, if P(h) is a property of h ∈ P(h) and ∀Fh : P(h) for ∃r ∈ R>0 ∀h ∈ BF r eR, then we write ∀sh : P(h) to denote ∃r ∈ r eR>0 ∀h ∈ Br(0) : r(c) : P(h). ¥ ¥ ¥ ¥ ¥ ¥ Inverse Function Theorems for Generalized Smooth Functions 7 (W r G C (U, Theorem 3. Let U ⊆ f ∈ functions fe ∈ C e , R). Then r eRn be a sharply open set, let v = [ve ] ∈ There exists a sharp neighborhood T of U × {0} and a generalized smooth map r ∈ such that r eRn, and let r eR) be a generalized smooth map generated by the net of smooth r eR), called the generalized incremental ratio of f along v, r eR) is another generalized incremental ratio of f along v de- If ¯r ∈ fined on a sharp neighborhood S of U × {0}, then f (x + hv) = f (x) + h · r(x,h). ∀x ∈ U ∀sh : (i) (ii) (T, (S, r G C r G C ∀x ∈ U ∀sh : r(x,h) = ¯r(x,h). (iii) We have r(x,0) =h ¶ fe ¶ ve (xe )i for every x ∈ U and we can thus define ¶ f ¶ v (x) := r(x,0), so that G C (U, ¶ f ¶ v ∈ r r eR). If U is a large open set, then an analogous statement holds replacing ∀sh by ∀Fh and sharp neighborhoods by large neighborhoods. chain rule. Finally, we note that for each x ∈ U, the map D f (x).v := Using this result we obtain the usual rules of differential calculus, including the Note that this result permits to consider the partial derivative of f with respect to an arbitrary generalized vector v ∈ r eR-linear in v ∈ by L( number defined by the operator norms of the matrices Ae ∈ L(Rn, Rd). r eRn. The set of all the r eR-linear maps r eRn −→ r eRd). For A = [Ae (−)] ∈ L( r eRn which can be, e.g., infinitesimal or infinite. r eRd is r eRd will be denoted r eRd), we set A := [Ae ], the generalized ¶ f ¶ v (x) ∈ r eRn, r eRn, Embedding of Schwartz distributions and Colombeau functions We finally recall two results that give a certain flexibility in constructing embeddings of Schwartz distributions. Note that both the infinitesimal r and the embedding of Schwartz distributions have to be chosen depending on the problem we aim to solve. A trivial example in this direction is the ODE y′ = y/ de , which cannot be solved for r = (e ), but it has a solution for r = (e−1/e ). As another simple example, if we need the property H(0) = 1/2, where H is the Heaviside function, then we have to choose the embedding of distributions accordingly. This corresponds to the philos- ophy followed in [26]. See also [20] for further details. If j ∈ D(Rn), r ∈ R>0 and x ∈ Rn, we use the notations r ⊙ j for the function for the function y ∈ Rn 7→ j (y − x) ∈ R. These x ∈ Rn 7→ 1 notations permit to highlight that ⊙ is a free action of the multiplicative group (R>0, ·,1) on D(Rn) and ⊕ is a free action of the additive group (R>0, +,0) on D(Rn). We also have the distributive property r ⊙ (x ⊕ j ) = rx ⊕ r ⊙ j . Lemma 3. Let b ∈ Rr be a net such that lime →0+ be = +¥ exists a net (y r(cid:1) ∈ R and x ⊕ j e )e ∈I of D(Rn) with the properties: . Let d ∈ (0,1). There rn · j (cid:0) x ¥ ¥ ¥ ¥ ¥ 8 Paolo Giordano and Michael Kunzinger e = 1 for all e ∈ I. e ) ⊆ B1(0) for all e ∈ I. (i) (ii) (iii) ∀a ∈ Nn ∃p ∈ N : supx∈Rn ¶ a y (iv) ∀ j ∈ N ∀0e : 1 ≤ a ≤ j ⇒ ´ xa e ≤ 1 + h . e )e ∈I can be chosen so that ´ 0 supp(y ´ y ∀h ∈ R>0 ∀0e : ´ y If n = 1, then the net (y e satisfies (i) – (vi) then in particular y be e (x) = O(bpe ) as e → 0+. · y (v) (vi) If y := b−1e ⊙ y e (x)dx = 0. −¥ y e = d. e satisfies (ii) - (v). where r fW c := {[xe ] ∈ [W Concerning embeddings of Schwartz distributions, we have the following result, : xe ∈ K} is called the set of compactly supported points in W ⊆ Rn. Theorem 4. Under the assumptions of Lemma 3, let W ⊆ Rn be an open set and let (y be ) be the net defined in Lemma 3. Then the mapping ] ∃K ⋐ W ∀0e i bW : T ∈ E ′(W ) 7→h(cid:16)T ∗ y be (cid:17) (−)i ∈ r G C uniquely extends to a sheaf morphism of real vector spaces ( r fW c, r eR) i b : D ′ −→ r G C and satisfies the following properties: ( r g(−)c, r eR), r ( (i) G C (−) −→ (−) : C (ii) (iii) (iv) r g(−)c, i bW (T )e · j = hT,j i for all j ∈ D(W ) and all T ∈ D ′(W ). If b ≥ dr −a for some a ∈ R>0, then i bC r eR) is a sheaf morphism of algebras. If T ∈ E ′(W ) then supp(T ) = supp(i bW (T )). lime →0+´W i b commutes with partial derivatives, i.e. ¶ a (cid:0)i bW (T )(cid:1) = i bW (¶ a T ) for each T ∈ D ′(W ) and a ∈ N. Concerning the embedding of Colombeau generalized functions, we recall that the special Colombeau algebra on W is defined as the quotient G s(W ) := EM(W )/N s(W ) of moderate nets over negligible nets, where the former is EM(W ) := {(ue ) ∈ C (W )I ∀K ⋐ W ∀a ∈ Nn ∃N ∈ N : sup x∈K ¶ a ue (x) = O(e −N)} and the latter is N s(W ) := {(ue ) ∈ C (W )I ∀K ⋐ W ∀a ∈ Nn ∀m ∈ N : sup x∈K ¶ a ue (x) = O(e m)}. Using r = (e ), we have the following compatibility result: Theorem 5. A Colombeau generalized function u = (ue ) + N s(W )d ∈ G s(W )d de- fines a generalized smooth map u : [xe ] ∈ Lipschitz on the same neighborhood of the Fermat topology for all derivatives. This r fW c −→ [ue (xe )] ∈ eRd which is locally ¥ ¥ ¥ ¥ ¥ ¥ ¥ ( 9 G C Inverse Function Theorems for Generalized Smooth Functions assignment provides a bijection of G s(W )d onto r set W ⊆ Rn. For GSF, suitable generalizations of many classical theorems of differential and integral calculus hold: intermediate value theorem, mean value theorems, Taylor formulas in different forms, a sheaf property for the Fermat topology, and the ex- treme value theorem on internal sharply bounded sets (see [18]). The latter are called r eRd) for every open r fW c, functionally compact subsets of r eRn and serve as compact sets for GSF. A theory of compactly supported GSF has been developed in [17], and it closely resembles the classical theory of LF-spaces of compactly supported smooth functions. It re- sults that for suitable functionally compact subsets, the corresponding space of com- pactly supported GSF contains extensions of all Colombeau generalized functions, and hence also of all Schwartz distributions. Finally, in these spaces it is possible to prove the Banach fixed point theorem and a corresponding Picard-Lindelof theorem, see [37]. 2 Local inverse function theorems As in the case of classical smooth functions, any infinitesimal criterion for the in- vertibility of generalized smooth functions will rely on the invertibility of the cor- responding differential. We therefore note the following analogue of [24, Lemma 1.2.41] (whose proof transfers literally to the present situation): (i) Lemma 4. Let A ∈ (iv) det(A) is invertible. A is nondegenerate, i.e., x ∈ (ii) A : r eRn → (iii) A : r eRn → r eRn, x tAh = 0 ∀h ∈ r eRn×n be a square matrix. The following are equivalent: r eRn implies x = 0. r eRn is injective. r eRn is surjective. r eRn, let f ∈ Theorem 6. Let X ⊆ the sharp interior of X, D f (x0) is invertible in L( sharp neighborhood U ⊆ X of x0 and a sharp neighborhood V of f (x0) such that f : U → V is invertible and f −1 ∈ r eRn) and suppose that for some x0 in r eRn). Then there exists a r eRn, G C (X, r r G C (V,U). nq 1 Proof. Thm. 2.(vi) entails that f can be defined by a globally defined net fe ∈ (Rn, Rn). Hadamard's inequality (cf. [11, Prop. 3.43]) implies D f (x0)−1 ≥ C C det (D f (x0)−1), where C ∈ R>0 is a universal constant that only depends on the dimension n. Thus, by Lemma 4 and Lemma 1, detD f (x0) and consequently also a := D f (x0)−1 is invertible. Next, pick positive invertible numbers b, r ∈ such that ab < 1, B2r(x0) ⊆ X and r eR D f (x0) − D f (x) < b ¥ ¥ ¥ ¥ 10 Paolo Giordano and Michael Kunzinger for all x ∈ B2r(x0). Such a choice of r is possible since every derivative of f is continuous with respect to the sharp topology (see Thm. 2.(iv) and Thm. 3.(iii)). Pick representatives (ae ), (be ) and (re ) of a, b and r such that for all e ∈ I we have be > 0, ae be < 1, and re > 0. Let (x0e ) be a representative of x0. Since [Bre (x0e )] ⊆ B2r(x0), by Lemma 2 we can also assume that Bre (x0e ) ⊆ W e , and D fe (x0e ) − D fe (x) < be for all x ∈ Ue := Bre (x0e ). Now let ce := ae 1−ae be . Then c := [ce ] > 0 and by [13, Th. 6.4] we obtain for each e ∈ I: (a) For all x ∈ Ue := Bre (x0e ), D fe (x) is invertible and D fe (x)−1 ≤ ce . (b) Ve := fe (Bre (x0e )) is open in Rn. (c) (d) setting y0e := fe (x0e ), we have Bre /ce (y0e ) ⊆ fe (Bre (x0e )). The sets U := hUe i = Br(x0) ⊆ X and V := hVe i are sharp neighborhoods of x0 and f (x0), respectively, by (d), and so it remains to prove that [ fe −1 (V,U). We first note that by (a), D fe (x)−1 ≤ ce for all x ∈ Bre (x0e ), which by Hadamard's fe Ue : Ue −→ Ve is a diffeomorphism, and Ue (−)] ∈ G C r inequality implies det(D fe (x)) ≥ 1 C · cne (x ∈ Bre (x0e )). Now for [ye ] ∈ V and 1 ≤ i, j ≤ n we have (see e.g. [11, (3.15)]) ¶ j( f −1e )i(ye ) = 1 det(D fe ( f −1e (ye ))) · Pi j((¶ s f re ( f −1e (ye )))r,s), (1) (2) where Pi j is a polynomial in the entries of the matrix in its argument. Since [ f −1e (ye )] ∈ U ⊆ X, it follows from (1) and the fact that f U ∈ (U, G C r (¶ j( f −1e )i(ye )) ∈ Rn r . r eRn) that Higher order derivatives can be treated analogously, thereby establishing that every derivative of ge := fe −1 is moderate. To prove the claim, it remains to show that Ue [ge (ye )] ∈ U = hUe i for all [ye ] ∈ V = hVe i. Since ge : Ve −→ Ue , we only prove that if (xe ) ∼r (ge (ye )), then also xe ∈ Ue for e small. We can set y′e := fe (xe ) because is defined on the entire Rn. By the mean value theorem applied to fe and the fe moderateness of f ′, we get y′e − ye = fe (xe ) − fe (ge (ye )) ≤ r Ne · xe − ge (ye ). Therefore (y′e ) ∼r (ye ) and hence y′e ∈ Ve and ge (y′e ) = xe ∈ Ue for e small. ⊓⊔ From Thm. 2.(iv), we know that any generalized smooth function is sharply con- tinuous. Thus we obtain: Corollary 1. Let X ⊆ such that D f (x) is invertible for each x ∈ X. Then f is a local homeomorphism with respect to the sharp topology. In particular, it is an open map. r eRn be a sharply open set, and let f ∈ r eRn) be (X, G C r ¥ ¥ ¥ Inverse Function Theorems for Generalized Smooth Functions 11 r Any such map f will therefore be called a local generalized diffeomorphism. If (Y,X), then it is called a global f ∈ generalized diffeomorphism. (X,Y ) possesses an inverse in r G C G C Following the same idea we used in the proof of Thm. 6, we can prove a sufficient condition to have a local generalized diffeomorphism which is defined in a large neighborhood of x0: r G C r eRn, let f ∈ Theorem 7. Let X ⊆ Fermat interior of X, D f (x0) is invertible in L( is finite, i.e. D f (x0)−1 ≤ k for some k ∈ R>0, and D f is Fermat continuous. Then there exists a large neighborhood U ⊆ X of x0 and a large neighborhood V of f (x0) such that f : U → V is invertible and f −1 ∈ r eRn) and suppose that for some x0 in the r eRn). Assume that D f (x0)−1 r eRn, (V,U). (X, G C r Proof. We proceed as above, but now we have re = r ∈ R>0, be = b ∈ R>0 because of our assumptions. Setting ce Therefore, there exists s ∈ R>0 such that s < r that now BF hoods of x0 and f (x0) respectively. c . We can continue as above, noting s(y0) ⊆ Br/c(y0) ⊆ V are large neighbor- ⊓⊔ r eR>0 is finite. 1−ae b , we have that c := [ce ] ∈ r(x0) ⊆ U = Br(x0) ⊆ X and BF := ae Example 1. r ( (i) G C r eR, e ) so that y e (be x)] is, up to sheaf isomorphism, r eR). We can take Thm. 4, for n = 1, shows that d (x) = [be y the Dirac delta. This also shows directly that d ∈ the net (y e (0) = 1 for all e . In this way, H′(0) = d (0) = b is an infinite number. We can thus apply the local inverse function theorem 6 to the Heaviside function H obtaining that H is a generalized diffeomorphism in an infinitesimal neighborhood of 0. This neighborhood cannot be finite because H′(r) = 0 for all r ∈ R6=0. (ii) By the intermediate value theorem for GSF (see [18, Cor. 42]), in the interval [0,1/2] the Dirac delta takes any value in [0,d (0)]. So, let k ∈ [0,1/2] such that d (k) = 1. Then by the mean value theorem for GSF (see [18, Thm. 43]) d (d (1)) − d (d (k)) = d (0) − d (1) = b − 0 = (d ◦ d )′(c) · (1 − k) for some c ∈ [k,1]. Therefore (d ◦ d )′(c) = b d ◦ d for all h ∈ r eR>0, and around c the composition is invertible. Note that (d ◦ d )(r) = b for all r ∈ R6=0, and (d ◦ d )(h) = 0 1−k ∈ r eR such that d (h) is not infinitesimal. r eR>0 be an infinitesimal generalized number, i.e. r ≈ 0. (iii) Let f (x) := r · x for x ∈ Now, let r ∈ (iv) Let f (x) := sin x Bs(rx0) 7→ y/r ∈ imal, so that s ≈ 0. This shows that the assumption in Thm. 7 on D f (x0)−1 being finite is necessary. r fRc. Then f ′(x0) = r ≈ 0 and Thm. 6 yields f −1 : y ∈ r eR>0. But y/r is finite only if y is infinites- r fRc for some s ∈ r cos x r , which is p 2 ≈ 0, k ∈ Z. By always an infinite number e.g. if ∃lime →0+ xe Thm. 6, we know that f is invertible e.g. around x = 0. It is easy to recognize r eR) and f ′(x) = 1 r . We have f ∈ 6= r(2k + 1) r eR, G C ( r that f is injective in the infinitesimal interval(cid:0)− is proved that f is not injective in any large neighborhood of x = 0. Therefore, p 2 r, + p 2 r(cid:1). In [11, Exa. 3.9], it ¥ ¥ ¥ ¥ ¥ ¥ 12 (v) (cid:16) f (− p 2 r,+ p 2 r)(cid:17)−1 Paolo Giordano and Michael Kunzinger is a GSF that cannot be extended to a Colombeau general- ized function. Similarly, f (x) := r sin x, x ∈ extended outside the infinitesimal neighborhood (−r,r). r eR, has an inverse function which cannot be (vi) Thm. 6 cannot be applied to f (x) := x3 at x0 = 0. However, if we restrict to , −r) ∪ (r, +¥ ), then the inverse function f −1(y) = y1/3 is defined in , −r3) ∪ (r3, +¥ ) and has infinite derivative at each infinitesimal point x ∈ (−¥ y ∈ (−¥ in its domain. In [2], Aragona, Fernandez and Juriaans introduced a differential calculus on spaces of Colombeau generalized points based on a specific form of convergence of dif- ference quotients. Moreover, in [3], an inverse function theorem for Colombeau generalized functions in this calculus was established. In the one-dimensional case it was shown in [19] that any GSF is differentiable in the sense of [2, 3], with the same derivative. Below we will show that this compatibility is in fact true in arbi- trary dimensions and that Theorem 6 implies the corresponding result from [3]. In the remaining part of the present section, we therefore restrict our attention to the case r e = e , the gauge that is used in standard Colombeau theory (as well as in First, we recall the definition from [2]: [2, 3]), and hence r eR =eR and r fW c = eW c. Definition 5. A map f from some sharply open subset U of eRn to eRm is called differentiable in x0 ∈ U in the sense of [2] with derivative A ∈ L(eRn,eRm) if f (x) − f (x0) − A(x − x0)e = 0, (3) lim x→x0 x − x0e where v : (xe ) ∈ Rn (e ) 7→ sup{b ∈ R xe = O(e b)} ∈ (−¥ ,¥ ] − e : x ∈eRn 7→ exp(−v(x)) ∈ [0,¥ ). The following result shows compatibility of this notion with the derivative in the sense of GSF. Lemma 5. Let U be sharply open ineRn, let x0 ∈ U and suppose that f ∈ G C Then f is differentiable in the sense of [2] in x0 with derivative D f (x0). (U,eRm). Proof. Without loss of generality we may suppose that m = 1. Let f be defined (Rn, R) for all e . Since (D2 fe (xe )) is moderate, it follows from by the net fe ∈ C Thm. 2.(ii) that there exists some q > 0 such that supy∈BEe q (xe ) D2 fe (y) ≤ e −q for e small. Then by Taylor's theorem we have fe (xe )− fe (x0e ) − D fe (x0e )(xe − x0e ) = = (cid:229) a =2 a a ! 1 0 (1 − t)a −1¶ a a fe (x0e + t(xe − x0e )) dt · (xe − x0e ) . ¥ ¥ Inverse Function Theorems for Generalized Smooth Functions For [xe ] ∈ Bde q(x0) this implies that f (x) − f (x0) − D f (x0)(x − x0)e ≤ eqx − x02 e, thereby establishing (3) with A = D f (x0), as claimed. 13 ⊓⊔ It follows that any f ∈ G C in the sense of [2]. (U,eRm) is in fact even infinitely often differentiable Based on these observations we may now give an alternative proof for [3, Th. 3]: such that f : U → V is a diffeomorphism in the sense of [2]. Theorem 8. Let W ⊆ Rn be open, f ∈ G s(W )n, and x0 ∈ eW c such that detD f (x0) is invertible in eR. Then there are sharply open neighborhoods U of x0 and V of f (x0) (eW c,eRn). Moreover, eW c Proof. By Thm. 5, f can be viewed as an element of G C is sharply open, which together with Lemma 4 shows that all the assumptions of Thm. 6 are satisfied. We conclude that f possesses an inverse f −1 in G C (V,U) for a suitable sharp neighborhood V of f (x0). Finally, by Lemma 5, both f and f −1 are infinitely differentiable in the sense of [2]. ⊓⊔ 3 Global inverse function theorems The aim of the present section is to obtain statements on the global invertibility of generalized smooth functions. For classical smooth functions, a number of criteria for global invertibility are known, and we refer to [9, 31] for an overview. The following auxilliary result will repeatedly be needed below: r Lemma 6. Let f ∈ Assume that /0 6= [Ae ] ⊆ X. Let b : Rd −→ R be a set-theoretical map such that ¯b : [ye ] ∈ Y 7→ [b(ye )] ∈ (X,Y ) be defined by ( fe ), where X ⊆ r eRn and Y ⊆ r eRd. G C r eR is well-defined (e.g., b(x) = x). If f satisfies ∀x ∈ X : ¯b [ f (x)] > 0, (4) then ∃q ∈ R>0 ∀0e ∀x ∈ Ae : b ( fe (x)) > r qe . (i) (ii) For all K ⋐ Rn, if [K] ⊆ X then ∀0e ∀x ∈ K : b ( fe (x)) > 0. Proof. In fact, suppose to the contrary that there was a sequence (e k)k ↓ 0 and a sequence xk ∈ Ae k such that b( fe k (xk)) ≤ r ke k. Let Ae 6= /0 for e ≤ e 0, and pick ae ∈ Ae . Set xe :=(cid:26) xk for e = e k ae otherwise. (5) It follows that x := [xe ] ∈ [Ae ] ⊆ X, and hence ¯b [ f (x)] > 0 by (4). Therefore, b(cid:0) fe k (xk)(cid:1) > r pe k for some p ∈ R>0 by Lemma 1, and this yields a contradiction. The second part follows by setting Ae = K in the first one and by noting that r e > 0. ⊓⊔ ¥ ¥ ¥ ¥ 14 Paolo Giordano and Michael Kunzinger ( ( After these preparations, we now turn to generalizing global inverse function theorems from the smooth setting to GSF. We start with the one-dimensional case. Here it is well-known that a smooth function f : R → R is a diffeomorphism onto its image if and only if f ′(x) > 0 for all x ∈ R. It is a diffeomorphism onto R if in addition there exists some r > 0 with f ′(x) > r for all x ∈ R. Despite the fact that r G C (i) (ii) (iii) r fRc is non-Archimedean, there is a close counterpart of this result in GSF. Theorem 9. Let f ∈ such that f ′(x) > r for all x ∈ f has a defining net ( ¯fe ) consisting of diffeomorphisms ¯fe : R → R. f is a global generalized diffeomorphism in r If r > 0, then f ( r G C r fRc, r fRc) and suppose that there exists some r ∈ R≥0 r fRc. Then r fRc, so f is a global generalized diffeomorphism in Proof. Let ( fe ) be a defining net for f such that fe ∈ C r fRc, Lemma 6 implies that for each n ∈ N 2 (vi)). Since f (x) > 0 for every x ∈ there exists some e n > 0 and some qn > 0 such that for each e ∈ (0,e n] and each x ∈ [−n,n] we have f ′e (x) > r qne . Clearly we may suppose that e n ↓ 0, qn+1 > qn for all n and that r qne < 1. Now for any n ∈ N>0 let j n : R → [0,1] be a smooth cut-off function with j n ≡ 1 on [−(n − 1),n − 1] and suppj n ⊆ [−n,n]. Supposing that f ′e (x) > 0 on [−n,n] (the case f ′e (x) < 0 on [−n,n] can be handled analogously), we set (R, R) for each e (cf. Thm. r fRc) = r fRc, f ( r fRc)). r fRc). r fRc, G C ( vne (x) := f ′e (x)j n(x) + 1 − j n(x) ¯fe (x) := fe (0) + x vne (t)dt 0 (x ∈ R) (x ∈ R, e n+1 < e ≤ e n), := fe for e ∈ (e 0,1]. Then ¯fe ∈ C (R, R) for each e , and for each x ∈ R and ¯fe and each e ∈ (e n+1,e n], we have ¯f ′e (x) = f ′e (x)j n(x) + 1 −j n(x) > r qne if and only if j n(x) · [1 − f ′e (x)] < 1 −r qne . The latter inequality holds if x /∈ [−n,n] or if f ′e (x) ≥ 1. Otherwise, j n(x) ≤ 1 < 1−r qne therefore is a diffeomorphism from R onto R. Also, ¯fe (x) = fe (x) for all x ∈ [−n,n] as soon as e ≤ e n+1. Hence also ( ¯fe ) is a defining net for f . This proves (i). 1− f ′e (x) because 1 > f ′e (x) > r qne For each e ≤ e 0, let ge be the global inverse of ¯fe . We claim that g := [ge ] is . Any such ¯fe r fRc) onto r fRc that is inverse to f . For this it suffices to show that a GSF from f ( (ye )) is r -moderate. whenever y = [(ye )] ∈ f ( To see this, it suffices to observe that for y = f (x), f satisfies the assumptions of the local inverse function theorem (Thm. 6) at x, and so the proof of that result shows that g is a GSF when restricted to a suitable sharp neighborhood of y. But this in particular entails the desired moderateness property at y, establishing (ii). r fRc), then for each k ∈ N, the net (g(k)e Finally, assume that r > 0. The same reasoning as in the proof of (i) now produces a defining net ( ¯fe ) with the property that ¯f ′e (x) > r for all e ≤ e 0 and all x ∈ R. Again, each ¯fe is a diffeomorphism from R onto R, and we denote its inverse by ge : R → R. Due to (ii) it remains to show that f : r fRc → r fRc is onto. ¥ ¥ ¥ ¥ ¥ r Inverse Function Theorems for Generalized Smooth Functions 15 To this end, note first that g′e (y) < 1/r for all e ≤ e 0 and all y ∈ Rn. Also, since r fRc), there exists some real number C > 0 such that fe (0) ≤ C for f ∈ e small. For such e and any [ye ] ∈ r fRc we obtain by the mean value theorem r fRc, G C ( ge (ye ) = ge (ye ) − ge ( fe (0)) ≤ ye − fe (0) ≤ (ye +C), (6) 1 r 1 r so that ge (ye ) remains in a compact set for e small. Based on this observation, the same argument as in (2) shows that, for any y = [ye ] ∈ (ye )) is moderate, so (ge ) defines a GSF r fRc −→ set x := g(y) to obtain f (x) = y. r fRc. Hence given y ∈ r fRc it suffices to ⊓⊔ r fRc and any k ≥ 1, (g(k)e Turning now to the multi-dimensional case, we first consider Hadamard's global inverse function theorem. For its formulation, recall that a map between topological spaces is called proper if the inverse image of any compact subset is again compact. As is easily verified, a continuous map a a (x) → ¥ : Rn → Rm is proper if and only if as x → ¥ (7) . Theorem 10. (Hadamard) A smooth map f : Rn → Rn is a global diffeomorphism if and only if it is proper and its Jacobian determinant never vanishes. For a proof of this result we refer to [22]. The following theorem provides an extension of Thm. 10 to the setting of GSF. Theorem 11. Suppose that Rn −→ Rn such that: r f ∈ G C ( c, r fRn r fRn c) possesses a defining net fe : (i) ∀x ∈ Rn ∀e ∈ I : D fe (x) is invertible in L(Rn, Rn), and for each x ∈ is invertible in L( r fRn (ii) There exists some e ′ ∈ I such that infe ∈(0,e ′] fe (x) → +¥ as x → ¥ Then f is a global generalized diffeomorphism in r r eRn). r eRn, G C . ( c, D f (x) is a global diffeomorphism Rn → Rn for each e ≤ e ′ Proof. By Thm. 10, each fe and we denote by ge : Rn → Rn the global inverse of fe . In order to prove that the net (ge )e ≤e ′ defines a GSF, we first note that, by (ii), the net ( fe )e ≤e ′ is 'uniformly proper' in the following sense: Given any M ∈ R≥0 there exists some M′ ∈ R≥0 such that when x ≥ M′ then ∀e ≤ e ′ : fe (x) ≥ M. Hence, for any K ⋐ Rn, picking M > 0 with K ⊆ BM(0) it follows that ge (K) ⊆ c into itself, i.e. BM′ (0) =: K′ ⋐ Rn for all e ≤ e ′. Thereby, the net (ge )e ≤e ′ maps r fRn ∀[ye ] ∈ c : [ge (ye )] ∈ (8) c). c, r fRn r fRn Moreover, for each K ⋐ Rn, assumption (i), Lemma 4, Lemma 1 and Lemma 6 yield r fRn c. r fRn ∃q ∈ R>0 ∀0e ∀x ∈ K : detD fe (x) > r qe . (9) ¥ ¥ ¥ 16 Paolo Giordano and Michael Kunzinger r fRc and any b ≥ 1, r fRn r fRn c). Finally, that g ⊓⊔ c, From (8) and (9) it follows as in (2) that, for any y = [ye ] ∈ (¶ b ge (ye )) is moderate, so g := [ye ] 7→ [ge (ye )] ∈ G C ( r is the inverse of f on r fRn c follows as in Thm. 9. The next classical inversion theorem we want to adapt to the setting of general- ized smooth functions is the following one: Theorem 12. (Hadamard-L´evy) Let f : X → Y be a local diffeomorphism between Banach spaces. Then f is a diffeomorphism if there exists a continuous non- decreasing function b : R≥0 → R>0 such that 0 1 b (s) ds = +¥ , D f (x)−1 ≤ b (x) ∀x ∈ X. This holds, in particular, if there exist a, b ∈ R>0 with D f (x)−1 ≤ a + bx for all x ∈ X. For a proof, see [9]. We can employ this result to establish the following global inverse function the- orem for GSF. Theorem 13. Suppose that f ∈ r G C (i) f possesses a defining net fe : Rn −→ Rn such that ∀x ∈ Rn ∀e ∈ I : D fe (x) is invertible in L(Rn, Rn), and for each x ∈ r eRn). (ii) There exists a net of continuous non-decreasing functions b e : R≥0 −→ R>0 c, D f (x) is invertible in L( such that ∀0e ∀x ∈ Rn : D fe (x)−1 ≤ b e (x) and r eRn, c) satisfies: ( c, r fRn r fRn r fRn 0 1 b e (s) ds = +¥ . Then f is a global generalized diffeomorphism in r G C If instead of (ii) we make the stronger assumption (iii) ∃C ∈ R>0 : ∀0e ∀x ∈ Rn : D fe (x)−1 ≤ C, then f is a global generalized diffeomorphism in r G C ( r fRn c, f ( c)). r fRn ( c, r fRn c). r fRn In particular, (ii) applies if there exist a, b ∈ r eR>0 that are finite (i.e., ae , be < R for some R ∈ R and e small) with D fe (x)−1 ≤ ae +be x for e small and all x ∈ r fRn c. Proof. From (ii) it follows by an e -wise application of Thm. 12 that there exists some e 0 > 0 such that each fe with e < e 0 is a diffeomorphism: Rn → Rn. We denote by ge its inverse. Using assumption (i), it follows exactly as in the proof of Thm. 9 (ii) that g := [ge ] is an element of r c) that is inverse to f . G C Assuming (iii), for any [ye ] ∈ r fRn r fRn c), c and e small, we have r fRn Dge (ye ) = (D fe (ge (ye )))−1 ≤ C, ( f ( so the mean value theorem yields ¥ ¥ ¥ ¥ ¥ ¥ ¥ Inverse Function Theorems for Generalized Smooth Functions ge (ye ) = ge (ye ) − ge ( fe (0)) ≤ Cye − fe (0), 17 (10) which is uniformly bounded for e small since f ∈ c). We conclude that (ge ) satisfies (8). From here, the proof can be concluded literally as in Thm. 11. ⊓⊔ G C ( r Remark 1. By Thm. 5, for r = (e ), the space r with the special Colombeau algebra G s(Rn)n. In this picture, r c) corresponds to the space of c-bounded Colombeau generalized functions on Rn (cf. [32, 24]). Therefore, under the further assumption that f ( c, theo- rems 11 and 13 can alternatively be viewed as global inverse function theorems for c-bounded Colombeau generalized functions. r fRn r fRn G C c, ( c, r fRn r fRn r eRn) can be identified r fRn r fRn r fRn c) = G C c, ( 4 Conclusions Once again, we want to underscore that the statement of the local inverse function theorem 6 is the natural generalization to GSF of the classical result. Its simplicity relies on the fact that the sharp topology is the natural one for GSF, as explained above. This natural setting permits to include examples in our theory that cannot be incorporated in an approach based purely on Colombeau generalized functions on classical domains (cf. Example 1 and [11]). Moreover, as Thm. 7 shows, the concept of Fermat topology leads, with com- parable simplicity, to sufficient conditions that guarantee solutions defined on large (non-infinitesimal) neighborhoods. Acknowledgement: We are indebted to the referee for several helpful comments that have substantially improved the results of Section 3. P. Giordano has been sup- ported by grants P25116 and P25311 of the Austrian Science Fund FWF. M. Kun- zinger has been supported by grants P23714 and P25326 of the Austrian Science Fund FWF. References 1. Abraham, R., Marsden, J.E., Ratiu, T.: Manifolds, Tensors, Analysis and Applications, second ed., Springer-Verlag, New York (1988) 2. Aragona, J., Fernandez, R., Juriaans, S.O.: A discontinuous Colombeau differential calculus. Monatsh. Math. 144, 13–29 (2005) 3. Aragona, J., Fernandez, R., Juriaans, S.O., Oberguggenberger, M.: Differential calculus and integration of generalized functions over membranes. Monatsh. Math. 166, 1–18 (2012) 4. Bampi, F., Zordan, C.: Higher order shock waves. Journal of Applied Mathematics and Physics, 41, 12–19 (1990) 5. Bertrand, D.: Review of 'Lectures on differential Galois theory', by A. Magid. Bull. Amer. Math. Soc. 33, 289–294 (1996) ¥ ¥ ¥ 18 Paolo Giordano and Michael Kunzinger 6. Bogoliubov, N.N., Shirkov, D.V.: The Theory of Quantized Fields. New York, Interscience (1959) 7. Casale, G.: Morales-Ramis theorems via Malgrange pseudogroup. Ann. Inst. Fourier 59, no. 7, 2593–2610 (2009) 8. Colombeau, J.F.: Multiplication of distributions - A tool in mathematics, numerical engineer- ing and theoretical Physics. Springer-Verlag, Berlin Heidelberg (1992) 9. De Marco, G., Gorni, G., Zampieri, G.: Global inversion of functions: an introduction. NoDEA Nonlinear Differential Equations Appl. 1 no. 3, 229–248 (1994) 10. Dirac, P.A.M.: The physical interpretation of the quantum dynamics. Proc. R. Soc. Lond. A, 113 27, 621–641 (1926–27) 11. Erlacher, E.: Local existence results in algebras of generalized functions. Doctoral thesis, Uni- versity of Vienna (2007) http://www.mat.univie.ac.at/ diana/uploads/publication47.pdf 12. Erlacher, E., Grosser, M.: Inversion of a 'discontinuous coordinate transformation' in General Relativity. Appl. Anal. 90, 1707–1728 (2011) 13. Erlacher, E., Inversion of Colombeau generalized functions. Proc. Edinb. Math. Soc. 56, 469–500 (2013) 14. Erlacher, E., Grosser, M.: Ordinary Differential Equations in Algebras of Generalized Func- tions. In: Pseudo-Differential Operators, Generalized Functions and Asymptotics, S. Mola- hajloo, S. Pilipovi´c, J. Toft, M. W. Wong eds, Operator Theory: Advances and Applications Volume 231, 253–270 (2013) 15. Estrada, R., Vindas, J.: A general integral. Dissertationes Math. 483, 1–49 (2012) 16. Campos Ferreira, J.: On some general notions of superior limit, inferior limit and value of a distribution at a point. Portugaliae Math. 28, 139–158 (2001) 17. Giordano P., Kunzinger M.: A convenient notion of compact set for generalized functions. Proc. Edinb. Math. Soc., in press. See arXiv 1411.7292. 18. Giordano P., Kunzinger M.: Steinbauer R.: A new approach to generalized functions for math- ematical physics. See http://www.mat.univie.ac.- at/giordap7/GenFunMaps.pdf. 19. Giordano, P., Kunzinger, M., Vernaeve, H.: Strongly internal sets and generalized smooth functions. Journal of Mathematical Analysis and Applications 422, issue 1, 56–71 (2015) 20. Giordano, P., Luperi Baglini, L.: Asymptotic gauges: Generalization of Colombeau type alge- bras. Math. Nachr. 289, 2-3, 1–28 (2015) 21. Gonzales Dominguez, A., Scarfiello, R.: Nota sobre la formula v.p. 1 Mat,. Argentina 1, 53–67 (1956) x · d = − 1 2 d ′. Rev. Un. 22. Gordon, W. B.: On the diffeomorphisms of Euclidean space. Amer. Math. Monthly 79, 755–759 (1972) 23. Grant, K.D.E., Mayerhofer, E., Steinbauer, R.: The wave equation on singular space-times. Commun. Math. Phys. 285, 399–420 (2009) 24. Grosser, M., Kunzinger, M., Oberguggenberger, M., Steinbauer, R.: Geometric theory of gen- eralized functions. Kluwer, Dordrecht (2001) 25. Gsponer, A.: A concise introduction to Colombeau generalized functions and their applica- tions in classical electrodynamics. European J. Phys. 30, no. 1, 109–126 (2009) 26. Hairer, M.: A theory of regularity structures. Invent. Math. 198, 269–504 (2014) 27. Hormann, G., Kunzinger, M., Steinbauer, R.: Wave equations on non-smooth space-times. In: Asymptotic Properties of Solutions to Hyperbolic Equations, M. Ruzhansky and J. Wirth (Eds.) Progress in Mathematics 301. 162–186, Birkhauser (2012). 28. Kanwal, R.P.: Generalized functions. Theory and technique. 2nd edition, Birkhauser (1998) 29. Katz, M.G., Tall, D.: A Cauchy-Dirac delta function. Foundations of Science (2012) See http://dx.doi.org/10.1007/s10699-012-9289-4 and http://arxiv.org/abs/1206.0119. 30. Konjik, S., Kunzinger, M., Oberguggenberger, M.: Foundations of the Calculus of Variations in Generalized Function Algebras. Acta Applicandae Mathematicae 103 n. 2, 169–199 (2008) 31. Krantz, S. G., Parks, H. R.: The implicit function theorem. History, theory, and applications. Reprint of the 2003 edition. Modern Birkhauser Classics. Birkhauser/Springer, New York (2013) Inverse Function Theorems for Generalized Smooth Functions 19 32. Kunzinger: M. Generalized functions valued in a smooth manifold. Monatsh. Math., 137 31–49 (2002) 33. Laugwitz, D.: Definite values of infinite sums: aspects of the foundations of infinitesimal analysis around 1820. Arch. Hist. Exact Sci. 39, no. 3, 195–245 (1989) 34. Laugwitz, D.: Early delta functions and the use of infinitesimals in research. Revue d'histoire des sciences, tome 45, 1, 115–128 (1992) 35. Łojasiewicz, S.: Sur la valeur et la limite d'une distribution en un point. Studia Math. 16, 1–36 (1957) 36. Łojasiewicz, S.: Sur la fixation des variables dans une distribution. Studia Math. 17, 1–64 (1958) 37. Luperi Baglini, L., Giordano, P.: A fixed point iteration method for arbitrary generalized ODE. In preparation. 38. Malgrange, B.: On nonlinear differential Galois theory. Dedicated to the memory of Jacques- Louis Lions. Chinese Ann. Math. Ser. B 23, no. 2, pp. 219–226 (2002). 39. Malgrange, B.: Pseudogroupes de Galois et th´eorie de Galois diff´erentielle. Pr´epublications IHES M/10/11 (2010). 40. Mayerhofer, E.: On Lorentz geometry in algebras of generalized functions. Proc. Roy. Soc. Edinburgh Sect. A 138, no. 4, 843–871 (2008) 41. Marsden, J.E.: Generalized Hamiltonian mechanics a mathematical exposition of non-smooth dynamical systems and classical Hamiltonian mechanics. Archive for Rational Mechanics and Analysis 28, n. 5, 323–361 (1968). 42. Marsden, J. E.: Non smooth geodesic flows and classical mechanics. Canad. Math. Bull. 12, 209–212 (1969) 43. Mikusinski, J.: On the square of the Dirac delta distribution. Bull. Acad. Polon. Sci. 14, 511–513 (1966) 44. Oberguggenberger, M., Vernaeve, H.: Internal sets and internal functions in Colombeau theory. J. Math. Anal. Appl. 341, 649–659 (2008) 45. Peetre, J.: On the value of a distribution at a point. Portugaliae Math. 27, 149–159 (1968) 46. Raju, C.K.: Products and compositions with the Dirac delta function. J. Phys. A: Math. Gen. 15, 381–396 (1982) 47. Robinson, A., Non-Standard Analysis. North-Holland, Amsterdam, (1966) 48. Robinson, A.: Function theory on some nonarchimedean fields. Amer. Math. Monthly 80 (6) 87–109; Part II: Papers in the Foundations of Mathematics (1973) 49. Sebastiao e Silva, J.: Sur une construction axiomatique de la theorie des distributions. Rev. Fac. Ciencias Lisboa, 2a Serie A, 4, pp. 79-186, (1954/55) 50. van der Pol, B., Bremmer, H.: Operational calculus (3rd ed.). New York: Chelsea Publishing Co. (1987) 51. Vernaeve, H., Vindas, J.: Characterization of distributions having a value at a point in the sense of Robinson. J. Math. Anal. Appl 396, 371–374 (2012)
1902.05305
1
1902
2019-02-14T11:33:13
Some Properties of Thinness and Fine Topology with Relative Capacity
[ "math.FA", "math.GN" ]
In this paper, we introduce a thinness in sense to a type of relative capacity for weighted variable exponent Sobolev space. Moreover, we reveal some properties of this thinness and consider the relationship with finely open and finely closed sets. We discuss fine topology and compare this topology with Euclidean one. Finally, we give some information about importance of the fine topology in the potential theory.
math.FA
math
SOME PROPERTIES OF THINNESS AND FINE TOPOLOGY WITH RELATIVE CAPACITY CIHAN UNAL AND ISMAIL AYDIN Abstract. In this paper, we introduce a thinness in sense to a type of relative capacity for weighted variable exponent Sobolev space. Moreover, we reveal some properties of this thinness and consider the relationship with finely open and finely closed sets. We discuss fine topology and compare this topology with Euclidean one. Finally, we give some information about importance of the fine topology in the potential theory. 1. Introduction The history of potential theory begins in 17th century. Its development can be traced to such greats as Newton, Euler, Laplace, Lagrange, Fourier, Green, Gauss, Poisson, Dirichlet, Riemann, Weierstrass, Poincar´e. We refer to the book by Kellogg [21] for references to some of the old works. The Sobolev spaces W k,p (Ω) are usually defined for open sets Ω. This makes sometimes difficulties to classical method for nonopen sets. The authors in [22] and [25] present different approach is to investigate Sobolev spaces on finely open sets. This is just a part of fine potential theory in Rd. Kov´acik and R´akosn´ık [24] introduced the variable exponent Lebesgue space Lp(.)(Rd) and the Sobolev space W k,p(.)(cid:0)Rd(cid:1). They present some basic properties of the variable exponent Lebesgue space Lp(.)(Rd) and the Sobolev space W k,p(.)(cid:0)Rd(cid:1) such as reflexivity and Holder inequalities were obtained. For a historical journey, we refer [9], [12], [24], [27] and [28]. The variational capacity has been used extensively in nonlinear potential theory on Rd. Let Ω ⊂ Rd is open and K ⊂ Ω is compact. Then the relative variational p-capacity is defined by capp (K, Ω) = inf f ZΩ ▽f (x)p dx, where the infimum is taken over smooth and zero boundary valued functions f in Ω such that f ≥ 1 in K. The set of admissible functions f can be replaced by the continuous first order Sobolev functions with f ≥ 1 in K. The p-capacity is a Choquet capacity relative to Ω. For more details and historical background, see [19]. Also, Harjulehto et al. [16] defined a relative capacity with variable exponent. They studied properties of the capacity and compare it with the Sobolev capacity. In [29], the authors expanded this relative capacity to weighted variable exponent. Moreover, they investigate properties of this capacity and give some relationship 2000 Mathematics Subject Classification. Primary 31C40, 46E35; Secondary 32U20, 43A15. Key words and phrases. Fine topology, Thinness, Relative capacity, Weighted variable expo- nent Sobolev spaces. 1 2 CIHAN UNAL AND ISMAIL AYDIN between defined capacity in [16] and Sobolev capacity. Besides to these studies, the Riesz capacity which is an another representative for capacity theory has been considered by [30]. In [1] and [8], the authors have explored some properties of the p (.)-Dirichlet energy integral ∇f (x)p(x) dx ZΩ over a bounded domain Ω ⊂ Rd. They have discussed the existence and regularity of energy integral minimizers. As an alternative method the minimizers in one dimensional case have been studied by the authors in [15]. Moreover, Harjulehto et al. [17] considered the Dirichlet energy integral, with boundary values given in the Sobolev sense, has a minimizer provided the variable exponent satisfies a certain jump condition. The fine topology was introduced by Cartan [6] in 1946. Classical fine topology has found many applications such as its connections to the theory of analytic func- tions and probability. For classical treatment we can refer [4], [7], [11], [13] and [20]. Also, Meyers [26] first generalized the fine topology to nonlinear theories. For the historical background and an excellent scientific survey we refer [19] and references therein. In this study, we present (p (.) , ϑ)-thin sets in sense to (p (.) , ϑ)-relative capacity and consider the basic and advanced properties. We discuss some results about (p (.) , ϑ)-relative capacity in (p (.) , ϑ)-thin sets. Moreover, we generalize several properties of fine topology and find new results by Wiener type integral. 2. Notation and Preliminaries In this paper, we will work on Rd with Lebesgue measure dx. The measure µ is doubling if there is a fixed constant cd ≥ 1, called the doubling constant of µ such that µ (B (x0, 2r)) ≤ cdµ (B (x0, r)) for every ball B (x0, r) in Rd. Also, the elements of the space C∞ infinitely differentiable functions with compact support. We denote the family of all measurable functions p (.) : Rd → [1, ∞) (called the variable exponent on Rd) 0 (cid:0)Rd(cid:1) are the by the symbol P(cid:0)Rd(cid:1). In this paper, the function p(.) always denotes a variable exponent. For p (.) ∈ P(cid:0)Rd(cid:1) , put p+ = ess sup p(x), p− = ess inf x∈Rd p(x). x∈Rd A measurable and locally integrable function ϑ : Rd → (0, ∞) is called a weight function. The weighted modular is defined by p(.),ϑ (f ) =ZRd f (x)p(x) ϑ (x) dx. The weighted variable exponent Lebesgue spaces Lp(.) functions f on Rd endowed with the Luxemburg norm ϑ kf kp(.),ϑ = inf  λ > 0 :ZRd(cid:12)(cid:12)(cid:12)(cid:12) f (x) λ (cid:12)(cid:12)(cid:12)(cid:12) p(x) (cid:0)Rd(cid:1) consist of all measurable ϑ (x) dx ≤ 1  . THINNESS AND FINE TOPOLOGY WITH RELATIVE CAPACITY 3 When ϑ (x) = 1, the space Lp(.) The space Lp(.) properties of this space were investigated in [2], [3], [23]. (cid:0)Rd(cid:1) is the variable exponent Lebesgue space. (cid:0)Rd(cid:1) is a Banach space with respect to k.kp(.),ϑ . Also, some basic We set the weighted variable exponent Sobolev spaces W k,p(.) ϑ ϑ W k,p(.) ϑ with the norm (Rd) =nf ∈ Lp(.) ϑ (cid:0)Rd(cid:1) : Dαf ∈ Lp(.) ϑ (cid:0)Rd(cid:1) by ϑ (Rd), 0 ≤ α ≤ ko Now, let 1 < p− ≤ p (.) ≤ p+ < ∞, k ∈ N and ϑ− 1 kf kk,p(.),ϑ = X0≤α≤k kDαf kp(.),ϑ 0 is a multiindex, α = α1 + α2 + ... + αd, and Dα = ∂α α2 x2 ...∂ αd xd α1 x1 ∂ ∂ . It where α ∈ Nd is already known that W k,p(.) ϑ embedding Lp(.) Sobolev spaces W k,p(.) (cid:0)Rd(cid:1) ֒→ L1 ϑ ϑ In particular, the space W 1,p(.) (cid:0)Rd(cid:1) is a reflexive Banach space. loc(cid:0)Rd(cid:1) . Thus, the loc(cid:0)Rd(cid:1) holds and then the weighted variable exponent p(.)−1 ∈ L1 ϑ (cid:0)Rd(cid:1) is well-defined by [[3], Proposition 2.1]. (cid:0)Rd(cid:1) =nf ∈ Lp(.) ϑ (Rd)o . (cid:0)Rd(cid:1) : ∇f ∈ Lp(.) (cid:0)Rd(cid:1) is defined by ϑ W 1,p(.) ϑ ϑ The function ρ1,p(.),ϑ : W 1,p(.) (Rd) −→ [0, ∞) is shown as ρ1,p(.),ϑ (f ) = ρp(.),ϑ (f )+ ρp(.),ϑ (∇f ) . Also, the norm kf k1,p(.),ϑ = kf kp(.),ϑ + k∇f kp(.),ϑ makes the space W 1,p(.) W 1,p(.) of variable exponent spaces can be found in [10],[24]. (cid:0)Rd(cid:1) a Banach space. The local weighted variable exponent Sobolev space ϑ,loc (cid:0)Rd(cid:1) is defined in the classical way. More information on the classic theory Let Ω ⊂ Rd is bounded and ϑ is a weight function. It is known that a function ϑ f ∈ C∞ 0 (Ω) satisfy Poincar´e inequality in L1 ϑ(Ω) if and only if the inequality ZΩ f (x) ϑ (x) dx ≤ c (diam Ω)ZΩ ∇f (x) ϑ (x) dx holds [19]. Unal and Aydın [29] defined an alternative capacity -called relative (p (.) , ϑ)- capacity-for Sobolev capacity in sense to [16]. For this, they recall that C0(Ω) = {f : Ω −→ R : f is continuous and suppf ⊂ Ω is compact} , where suppf is the support of f . Suppose that K is a compact subset of Ω. Also, they denote Rp(.),ϑ (K, Ω) =nf ∈ W 1,p(.) ϑ and define (Ω) ∩ C0 (Ω) : f > 1 on K and f ≥ 0o cap∗ p(.),ϑ (K, Ω) = inf f ∈Rp(.),ϑ(K,Ω)ZΩ ▽f (x)p(x) ϑ (x) dx. In addition, if U ⊂ Ω is open, then capp(.),ϑ (U, Ω) = sup K⊂U compact cap∗ p(.),ϑ (K, Ω) , 4 CIHAN UNAL AND ISMAIL AYDIN and also for an arbitrary set E ⊂ Ω we define capp(.),ϑ (E, Ω) = inf E⊂U⊂Ω U open capp(.),ϑ (U, Ω) . They call capp(.),ϑ (E, Ω) the variational (p (.) , ϑ)-capacity of E relative to Ω, briefly the relative (p (.) , ϑ)-capacity. Also, the relative (p (.) , ϑ)-capacity has the following properties. P1 . capp(.),ϑ (∅, Ω) = 0. P2 . If E1 ⊂ E2 ⊂ Ω2 ⊂ Ω1, then capp(.),ϑ (E1, Ω1) ≤ capp(.),ϑ (E2, Ω2) . P3 . If E is a subset of Ω, then capp(.),ϑ (E, Ω) = inf E⊂U⊂Ω U open capp(.),ϑ (U, Ω) . P4 . If K1 and K2 are compact subsets of Ω, then capp(.),ϑ (K1 ∪ K2, Ω) + capp(.),ϑ (K1 ∩ K2, Ω) ≤ capp(.),ϑ (K1, Ω) +capp(.),ϑ (K2, Ω) . P5 . Let Kn is a decreasing sequence of compact subsets of Ω for n ∈ N. Then P6 . If En is an increasing sequence of subsets of Ω for n ∈ N, then lim n−→∞ lim n−→∞ capp(.),ϑ (Kn, Ω) = capp(.),ϑ(cid:18) ∞ Tn=1 capp(.),ϑ (En, Ω) = capp(.),ϑ(cid:18) ∞ Sn=1 Kn, Ω(cid:19) . En, Ω(cid:19) . P7 . If En ⊂ Ω for n ∈ N, then capp(.),ϑ(cid:18) ∞ Sn=1 En, Ω(cid:19) ≤ ∞ Pn=1 capp(.),ϑ (En, Ω) . Theorem 1. [29]If capp(.),ϑ (B (x0, r) , B (x0, 2r)) ≥ 1 and µϑ is a doubling mea- sure, then there exist positive constants C1, C2 such that C1µϑ (B (x0, r)) ≤ capp(.),ϑ (B (x0, r) , B (x0, 2r)) ≤ C2µϑ (B (x0, r)) where the constants depend on r, p−, p+, constants of doubling measure and Poincar´e inequality. Theorem 2. [29]If E ⊂ B (x0, r) , capp(.),ϑ (E, B (x0, 4r)) ≥ 1 and 0 < r ≤ s ≤ 2r, then the inequality 1 C capp(.),ϑ (E, B (x0, 2r)) ≤ capp(.),ϑ (E, B (x0, 2s)) ≤ capp(.),ϑ (E, B (x0, 2r)) holds where the constants depend on r, p−, p+, constants of doubling measure and Poincar´e inequality. The proofs can be found in [29]. We say that a property holds (p(.), ϑ)-quasieverywhere if it satisfies except in a set of capacity zero. Recall also a function f is (p(.), ϑ)-quasicontinuous in Rd if for each ε > 0 there exists a set A with the capacity of A is less than ε such that f restricted to Rd − A is continuous. If the capacity is an outer capacity, we can suppose that A is open. More detail can be found in [3]. THINNESS AND FINE TOPOLOGY WITH RELATIVE CAPACITY 5 Let Ω ⊂ Rd be an open set. The space W 1,p(.) 0,ϑ (Ω) is denoted as the set of all measurable functions f if there exists a (p(.), ϑ)-q.c. such that f = f ∗ a.e. in Ω and f ∗ = 0 (p(.), ϑ)-q.e. f ∈ W 1,p(.) the trace of f ∗ vanishes. More detail about the space can be seen by [14], [19], [31]. Moreover, A ⋐ B means that A is a compact subset of B. Throughout this (Ω) , if there exist a (p(.), ϑ)-q.c. function f ∗ ∈ W 1,p(.) 0,ϑ function f ∗ ∈ W 1,p(.) in Rd − Ω. In other words, ϑ ϑ (cid:0)Rd(cid:1) (cid:0)Rd(cid:1) such that loc(cid:0)Rd(cid:1) . Also, paper, we assume that 1 < p− ≤ p (.) ≤ p+ < ∞ and ϑ− 1 we will denote p(.)−1 ∈ L1 µϑ (Ω) =ZΩ ϑ (x) dx. 3. The (p (.) , ϑ)-Thinness and Fine Topology Now, we present (p (.) , ϑ)-thinness and consider some properties of this thinness before considering the fine topology. Definition 1. A set E ⊂ Rd is (p (.) , ϑ)-thin at x ∈ Rd if (3.1) 1 Z0 (cid:18) capp(.),ϑ (E ∩ B (x, r) , B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) 1 p(x)−1 dr r < ∞. Also, we say that E is (p (.) , ϑ)-thick at x ∈ Rd if E is not (p (.) , ϑ)-thin at x ∈ Rd. In the definition of (p (.) , ϑ)-thinness we make a convention that the integral is 1 if capp(.),ϑ (B (x, r) , B (x, 2r)) = 0. Also, the integral in (3.1) is usually called the Wiener type integral, briefly Wiener integral, as Wp(.),ϑ (E, x) = 1 Z0 (cid:18) capp(.),ϑ (E ∩ B (x, r) , B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) 1 p(x)−1 dr r . In addition, we denote the Wiener sum W sum p(.),ϑ (E, x) as W sum p(.),ϑ (E, x) = ∞ Xi=0 capp(.),ϑ(cid:0)E ∩ B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! 1 p(x)−1 . Now we give a relationship between these two notions. The proof can be found in [29]. Theorem 3. Assume that the hypotheses of Theorem 1 and Theorem 2 are hold. Then there exist positive constants C1, C2 such that C1Wp(.),ϑ (E, x) ≤ W sum p(.),ϑ (E, x) ≤ C2Wp(.),ϑ (E, x) for every E ⊂ Rd and x0 /∈ E. In particular, Wp(.),ϑ (E, x0) is finite if and only if W sum p(.),ϑ (E, x0) is finite. The previous theorem tell us that the notions Wp(.),ϑ and W sum under some conditions. In some cases, the Wiener sum W sum than the Wiener integral Wp(.),ϑ. p(.),ϑ are equivalent p(.),ϑ is more practical 6 CIHAN UNAL AND ISMAIL AYDIN Definition 2. A set U ⊂ Rd is called (p (.) , ϑ)-finely open if Rd − U is (p (.) , ϑ)- thin at x ∈ U. Equivalently, a set is (p (.) , ϑ)-finely closed if it includes all points where it is not (p (.) , ϑ)-thin. Moreover, the fine interior of A, briefly fine-intA, is the largest (p (.) , ϑ)-finely open set contained in A. In a similar way, the fine closure of F, briefly fine-cloF, is the smallest (p (.) , ϑ)-finely closed set containing F. Theorem 4. The (p (.) , ϑ)-fine topology on Rd is generated by (p (.) , ϑ)-finely open sets. Proof. Firstly, we denote τ F =   =   E ⊂ Rd : E ⊂ Rd : 1 capp(.),ϑ (B (x, r) , B (x, 2r)) capp(.),ϑ(cid:0)(cid:0)Rd − E(cid:1) ∩ B (x, r) , B (x, 2r)(cid:1) Z0 Z0 (cid:18) capp(.),ϑ (B (x, r) − E, B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) p(x)−1 dr r 1 1 ∪ ∅ < ∞  ∪ ∅. 1 p(x)−1 dr r ! < ∞  It is obvious that ∅ ∈ τ F . Since capp(.),ϑ (∅, B (x, 2r)) = 0, we have 1 capp(.),ϑ(cid:0)B (x, r) − Rd, B (x, 2r)(cid:1) capp(.),ϑ (B (x, r) , B (x, 2r)) ! Z0 capp(.),ϑ (B (x, r) , B (x, 2r))(cid:19) Z0 (cid:18) capp(.),ϑ (∅, B (x, 2r)) 1 1 p(x)−1 dr r 1 p(x)−1 dr r < ∞. = This follows that Rd ∈ τ F . Now, we assert that finite intersections of (p (.) , ϑ)-finely n open sets are (p (.) , ϑ)-finely open. Assume that x ∈ Ui where U1, U2, ..., Un are (p (.) , ϑ)-finely open. Thus, if we consider the subadditivity of relative (p (.) , ϑ)- capacity and the cases of the exponent p (.) as 1 < p (.) ≤ 2 and p (.) > 2, then we get Ti=1 1 1 n Ti=1   capp(.),ϑ (B (x, r) , B (x, 2r)) Ui, B (x, 2r)(cid:19) capp(.),ϑ(cid:18)B (x, r) −  Z0  Z0 (cid:18) n capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) Pi=1 Z0 (cid:18) capp(.),ϑ (B (x, r) − Ui, B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) Pi=1 capp(.),ϑ (B (x, r) − Ui, B (x, 2r)) n 1 ≤ ≤ C 1 p(x)−1 dr r 1 p(x)−1 dr r 1 p(x)−1 dr r < ∞ where C > 0 depends on n, p−, p+. Therefore Ui is (p (.) , ϑ)-finely open. Finally, we need to show that arbitrary unions of (p (.) , ϑ)-finely open sets are (p (.) , ϑ)- Ui where Ui, i ∈ I, are (p (.) , ϑ)-finely open sets, and I is n Ti=1 finely open. Let x ∈ Si∈I THINNESS AND FINE TOPOLOGY WITH RELATIVE CAPACITY 7 an index set. Thus, for every i ∈ I, we have (3.2) 1 Z0 (cid:18) capp(.),ϑ (B (x, r) − Ui, B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) 1 p(x)−1 dr r < ∞. Moreover, it is clear that B (x, r)−Si∈I Ui ⊂ B (x, r)−Uj or equivalently Ti∈I B (x, r) − Uj for j ∈ I. If we consider the properties of relative (p (.) , ϑ)-capacity and (3.2), then we get (B (x, r) − Ui) ⊂ 1 1 1 Z0 capp(.),ϑ (B (x, r) , B (x, 2r)) Ui, B (x, 2r)(cid:19) capp(.),ϑ(cid:18)Ti∈I capp(.),ϑ(cid:18)B (x, r) − Si∈I   (B (x, r) − Ui) , B (x, 2r)(cid:19)    Z0  Z0 (cid:18) capp(.),ϑ (B (x, r) − Uj, B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) Ui is (p (.) , ϑ)-thin at x and as x ∈ Si∈I capp(.),ϑ (B (x, r) , B (x, 2r)) 1 p(x)−1 dr r = ≤ 1 p(x)−1 dr r 1 p(x)−1   dr r < ∞. Therefore, Rd − Si∈I is (p (.) , ϑ)-finely open. Ui was arbitrary, Si∈I Ui (cid:3) Corollary 1. Every open set is (p (.) , ϑ)-finely open. Proof. Assume that A is an open set in Rd. For every x ∈ A, by the definition of openness, there exists a t > 0 such that B (x, t) ⊂ A. It is easy to see that B (x, r) ⊂ B (x, t) ⊂ A for small enough r > 0. This follows that 1 Z0 (cid:18) capp(.),ϑ (B (x, r) − A, B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) 1 p(x)−1 dr r < ∞ ,that is, A ⊂ Rd is (p (.) , ϑ)-finely open. (cid:3) Remark 1. By the similar method in Corollary 1, it can be shown that every closed set is (p (.) , ϑ)-finely closed and that finite union of (p (.) , ϑ)-finely closed sets is (p (.) , ϑ)-finely closed again. Corollary 2. The (p (.) , ϑ)-fine topology generated by the (p (.) , ϑ)-finely open sets is finer than Euclidean topology. The opposite claim of Corollary 1 is not true in general. To see this, we give the Lebesgue spine to E =n(x, t) ∈ R2 × R : t > 0 and x < e− 1 as a counter example, see [5, Example 13.4]. Now, we consider the more general case in sense to Corollary 1. 8 CIHAN UNAL AND ISMAIL AYDIN Theorem 5. Assume that A ⊂ Rd is an open or (p (.) , ϑ)-finely open set. More- over, let, the relative (p (.) , ϑ)-capacity of E is zero. Then A − E is (p (.) , ϑ)-finely open. Proof. By the Corollary 1, we can consider that A ⊂ Rd is an open set. Thus, for all y ∈ A, (3.3) 1 Z0 (cid:18) capp(.),ϑ (B (y, r) − A, B (y, 2r)) capp(.),ϑ (B (y, r) , B (y, 2r)) (cid:19) 1 p(y)−1 dr r < ∞. Moreover, if we consider the properties of relative (p (.) , ϑ)-capacity, for all x ∈ A − E and r > 0, we have capp(.),ϑ (B (x, r) − (A − E) , B (x, 2r)) = capp(.),ϑ ((B (x, r) − A) ∪ (B (x, r) ∩ E) , B (x, 2r)) ≤ capp(.),ϑ ((B (x, r) − A) , B (x, 2r)) (3.4) +capp(.),ϑ ((B (x, r) ∩ E) , B (x, 2r)) . Using the (3.3) and (3.4), we get ≤ 1 1 capp(.),ϑ (B (x, r) , B (x, 2r)) Z0 (cid:18) capp(.),ϑ (B (x, r) − (A − E) , B (x, 2r)) Z0 (cid:18) capp(.),ϑ ((B (x, r) − A) , B (x, 2r)) (cid:19) Z0 (cid:18) capp(.),ϑ ((B (x, r) ∩ E) , B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) capp(.),ϑ (B (x, r) , B (x, 2r)) + 1 1 p(x)−1 dr r (cid:19) 1 p(x)−1 dr r 1 p(x)−1 dr r < ∞. This completes the proof. (cid:3) Now, we give that (p (.) , ϑ)-thinness is a local property. Theorem 6. A ⊂ Rd is (p (.) , ϑ)-thin at x ∈ Rd if and only if for any δ > 0, the set A ∩ B (x, δ) is (p (.) , ϑ)-thin at x ∈ Rd. Proof. Let A ⊂ Rd and x ∈ Rd. Assume that A ⊂ Rd is (p (.) , ϑ)-thin at x ∈ Rd. This follows that 1 1 Z0 (cid:18) capp(.),ϑ (A ∩ B (x, r) , B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) p(x)−1 dr r < ∞. By the monotonicity of relative (p (.) , ϑ)-capacity, we have (3.5) 1 Z0 (cid:18) capp(.),ϑ (A ∩ B (x, δ) ∩ B (x, r) , B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) 1 p(x)−1 dr r (cid:19) < ∞ for any δ > 0. This completes the necessary condition part of the proof. Now, we assume that for any δ > 0, the set A∩B (x, δ) is (p (.) , ϑ)-thin at x ∈ Rd. Thus (3.5) THINNESS AND FINE TOPOLOGY WITH RELATIVE CAPACITY 9 is satisfied for all δ > 0, in particular, for 0 < r ≤ δ ≤ 1. Let A is (p (.) , ϑ)-thick at x ∈ Rd. Then we have 1 capp(.),ϑ (B (x, r) , B (x, 2r)) Z0 (cid:18) capp(.),ϑ (A ∩ B (x, δ) ∩ B (x, r) , B (x, 2r)) Z0 (cid:18) capp(.),ϑ (A ∩ B (x, r) , B (x, 2r)) capp(.),ϑ (B (x, r) , B (x, 2r)) (cid:19) p(x)−1 dr r 1 1 1 p(x)−1 dr r (cid:19) = ∞. = This is a contradiction. That is the desired result. (cid:3) Theorem 7. Let the hypotheses of Theorem 3 hold. Moreover, assume that there is a point x ∈ A such that Rd − A is (p (.) , ϑ)-thin at x. Then there exist a point x ∈ A and s > 0 such that capp(.),ϑ (B (x, s) − A, B (x, 2s)) < capp(.),ϑ (B (x, s) , B (x, 2s)) . Proof. Since Rd − A is (p (.) , ϑ)-thin at x ∈ A, by the Theorem 3, we have W sum p(.),ϑ(cid:0)Rd − A, x(cid:1) = This follows that ∞ ∞ = capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) Xi=0 capp(.),ϑ(cid:0)(cid:0)Rd − A(cid:1) ∩ B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) Xi=0 capp(.),ϑ(cid:0)B(cid:0)x, 2−i(cid:1) − A, B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! i−→∞ capp(.),ϑ(cid:0)B(cid:0)x, 2−i(cid:1) − A, B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! = 0. p(x)−1 p(x)−1 lim inf 1 1 p(x)−1 ! < ∞. 1 By the definition of limit, we get the desired result. (cid:3) Remark 2. The proof of the previous theorem can be considered by using Wiener integral Wp(.),ϑ (A, x) with similar method. Here, there is not necessary the condi- tion that the hypotheses of Theorem 3 are hold. Definition 3. Let U be a (p (.) , ϑ)-finely open set. A function f : U −→ R is (p (.) , ϑ)-finely continuous at x0 ∈ U if {x ∈ U : f (x) − f (x0) ≥ ε} is (p (.) , ϑ)- thin at x0 for each ε > 0. Remark 3. Assume that U is a (p (.) , ϑ)-finely open set and f : U −→ R is (p (.) , ϑ)-finely continuous at x0 ∈ U. Then f is continuous function with respect to the (p (.) , ϑ)-fine topology on U. Indeed, if we consider the definition of finely continuous, then the set {x ∈ U : f (x) − f (x0) < ε} is (p (.) , ϑ)-finely open. This follows that f : U −→ R is continuous at x0 ∈ U in sense to the (p (.) , ϑ)-fine topology on U . The converse argument is still an open problem, see [14]. Moreover, this argument for the constant exponent was considered by [25, Teorem 2.136]. It is note that a set A is a (p (.) , ϑ)-fine neighbourhood of a point x if and only if x ∈ A and Rd − A is (p (.) , ϑ)-thin at x, see [19]. Theorem 8. Let A ⊂ B (x0, r) , capp(.),ϑ (A, B (x0, 4r)) ≥ 1 and 0 < r ≤ s ≤ 2r. Assume that A ⊂ Rd is (p (.) , ϑ)-thin at x ∈ A − A. Then there exists an open neighbourhood U of A such that U is (p (.) , ϑ)-thin at x and x /∈ U. 10 CIHAN UNAL AND ISMAIL AYDIN Proof. Using the same methods in the Theorem 2 and Theorem 1, it can be found for r ≤ s ≤ 2r that (3.6) and (3.7) capp(.),ϑ (A ∩ B (x, r) , B (x, 2r)) ≈ capp(.),ϑ (A ∩ B (x, r) , B (x, 2s)) capp(.),ϑ (B (x, r) , B (x, 2r)) ≈ capp(.),ϑ (B (x, s) , B (x, 2s)) where the constants in ≈ depend on r, p−, p+, constants of doubling measure and Poincar´e inequality, see [29]. If we consider the Theorem 2 and the monotonicity of relative (p (.) , ϑ)-capacity, then we have capp(.),ϑ(cid:16)A ∩ B (x, 2−i), B(cid:0)x, 21−i(cid:1)(cid:17) ≤ T capp(.),ϑ(cid:16)A ∩ B (x, 2−i), B(cid:0)x, 22−i(cid:1)(cid:17) ≤ T capp(.),ϑ(cid:0)A ∩ B(cid:0)x, 21−i(cid:1) , B(cid:0)x, 22−i(cid:1)(cid:1) . By the definition of relative (p (.) , ϑ)-capacity, it can be taken open sets Ui ⊃ A ∩ B (x, 2−i) such that ≤   (3.8) Denote 1 p(x)−1 capp(.),ϑ (B (x, 2−i) , B (x, 21−i))! capp(.),ϑ(cid:0)Ui, B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ(cid:16)A ∩ B (x, 2−i), B(cid:0)x, 21−i(cid:1)(cid:17) capp(.),ϑ (B (x, 2−i) , B (x, 21−i))   1 p(x)−1 + 1 2i . U =(cid:0)Rd − B1(cid:1) ∪(cid:0)U1 − B2(cid:1) ∪(cid:0)(U1 ∩ U2) − B3(cid:1) ∪(cid:0)(U1 ∩ U2 ∩ U3) − B4(cid:1) ∪ ... where Bi = B(cid:0)x, 2−i(cid:1) . It is easy to see that U is open, A ⊂ U holds and x /∈ U. Since B(cid:0)x, 2−i(cid:1) ⊂ B (x, 2−i) holds for i ∈ N, it is clear that U ∩ B(cid:0)x, 2−i(cid:1) ⊂ Ui. By (3.8), we have 1 1 p(x)−1 + 1 2i . ≤ ≤   1 p(x)−1 capp(.),ϑ(cid:0)U ∩ B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! capp(.),ϑ (B (x, 2−i) , B (x, 21−i))! capp(.),ϑ(cid:0)Ui, B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ(cid:16)A ∩ B (x, 2−i), B(cid:0)x, 21−i(cid:1)(cid:17) capp(.),ϑ (B (x, 2−i) , B (x, 21−i))   capp(.),ϑ(cid:16)A ∩ B (x, 2−i), B(cid:0)x, 21−i(cid:1)(cid:17) ≤ C1capp(.),ϑ(cid:16)A ∩ B (x, 2−i), B(cid:0)x, 22−i(cid:1)(cid:17) ≤ C2capp(.),ϑ(cid:0)A ∩ B(cid:0)x, 21−i(cid:1) , B(cid:0)x, 22−i(cid:1)(cid:1) . p(x)−1 Moreover, if we consider (3.6), then we get (3.9) THINNESS AND FINE TOPOLOGY WITH RELATIVE CAPACITY 11 By (3.7), the inequality (3.10) ≥ capp(.),ϑ(cid:0)B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ(cid:0)B(cid:0)x, 21−i(cid:1) , B(cid:0)x, 22−i(cid:1)(cid:1) 1 C2 holds. If we combine (3.9) and (3.10), then we have 1 p(x)−1 ∞ capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! Xi=1 capp(.),ϑ(cid:0)U ∩ B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) Xi=1 capp(.),ϑ(cid:0)A ∩ B(cid:0)x, 21−i(cid:1) , B(cid:0)x, 22−i(cid:1)(cid:1) capp(.),ϑ (B (x, 21−i) , B (x, 22−i)) ! ∞ ≤ C 1 p(x)−1 + 1. Since A is (p (.) , ϑ)-thin at x, by considering the definition of Wiener sum W sum we conclude p(.),ϑ, ∞ Xi=1 capp(.),ϑ(cid:0)U ∩ B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! 1 p(x)−1 < ∞. This follows that W sum p(.),ϑ (U, x) 1 p(x)−1 ∞ = Xi=0 capp(.),ϑ(cid:0)U ∩ B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! = (cid:18) capp(.),ϑ (U ∩ B (x, 1) , B (x, 2)) capp(.),ϑ (B (x, 1) , B (x, 2)) (cid:19) Xi=1 capp(.),ϑ(cid:0)U ∩ B(cid:0)x, 2−i(cid:1) , B(cid:0)x, 21−i(cid:1)(cid:1) capp(.),ϑ (B (x, 2−i) , B (x, 21−i)) ! p(x)−1 + ∞ 1 1 p(x)−1 < ∞. Hence U is (p (.) , ϑ)-thin at x. Thus the claim is follows from definition of open neighbourhood. (cid:3) Now, we consider the usage area of (p (.) , ϑ)-fine topology in potential theory. We define (p (.) , ϑ)-Laplace equation as (3.11) − ∆p(.),ϑ = − div(cid:16)ϑ (x) ∇f p(.)−2 ∇f(cid:17) = 0 for every f ∈ W 1,p(.) (Ω) . 0,ϑ Definition 4. ([31]) Let Ω ⊂ Rd for d ≥ 2, be an open set. A function f ∈ W 1,p(.) ϑ,loc (Ω) called a (weak) weighted solution (briefly (p (.) , ϑ)-solution) of (3.11) in Ω, if ∇f (x)p(x)−2 ∇f (x) · ∇g (x) ϑ (x) dx = 0 ZΩ 12 CIHAN UNAL AND ISMAIL AYDIN whenever g ∈ C∞ supersolution (briefly (p (.) , ϑ)-supersolution) of (3.11) in Ω, if 0 (Ω) . Moreover, a function f ∈ W 1,p(.) ϑ,loc (Ω) is a (weak) weighted (3.12) ZΩ ∇f (x)p(x)−2 ∇f (x) · ∇g (x) ϑ (x) dx ≥ 0 whenever g ∈ C∞ if −f is a (p (.) , ϑ)-supersolution in Ω, and a weighted solution in Ω. 0 (Ω) is nonnegative. A function f is a weighted subsolution in Ω Definition 5. ([14], [19]) A function f : Ω −→ (−∞, ∞] is (p (.) , ϑ)-superharmonic in Ω if (i) f is lower semicontinuous, (ii) f is finite almost everywhere, (iii) Assume that D ⋐ Ω is an open set. If g is a (p (.) , ϑ)-solution in D, which is continuous in D, and satisfies f ≥ g on ∂D, then f ≥ g in D. Note that every (p (.) , ϑ)-supersolution in Ω, which satisfies f (x) = ess lim inf y−→x f (y) for all x ∈ Ω, is (p (.) , ϑ)-superharmonic in Ω. On the other hand every locally bounded (p (.) , ϑ)-superharmonic function is a (p (.) , ϑ)-supersolution. The proof can be easily seen by using the similar method in [18], [19]. Let S(cid:0)Rd(cid:1) be the class of all (p (.) , ϑ)-superharmonic functions in Rd. Since (p (.) , ϑ)-superharmonic functions are lower semicontinuous and since S(cid:0)Rd(cid:1) is closed under truncations, (p (.) , ϑ)-fine topology is the coarsest topology on Rd making all locally bounded (p (.) , ϑ)-superharmonic functions continuous, see [19]. 4. Acknowledgment We express our thanks to Professor Jana Bjorn for kind comments and helpful suggestions. References [1] Acerbi, E., Mingione, G. Regularity results for a class of functionals with non-standard growth, Arch. Ration. Mech. and Anal. 156 (2001), 121-140. [2] Aydın, I. On variable exponent amalgam spaces, An. St. Univ. Ovidius Constanta, Ser. Mat. 20(3), (2012), 5-20. [3] Aydın, I. Weighted variable Sobolev spaces and capacity, J. Funct. Spaces Appl. 2012, Article ID 132690, 17 pages, doi:10.1155/2012/132690, (2012). [4] Bjorn, J. Fine continuity on metric spaces, Manuscripta Math. 125 (2008), 369-381. [5] Bjorn, A., Bjorn, J. Nonlinear Potential Theory on Metric Spaces, Zurich: European Math- ematical Society (EMS), 2011. [6] Cartan, H.: The´orie g´en´eral du balayage en potential newtonien, Ann. Univ. Grenoble Math. Phys. 22 (1946), 221-280. [7] Constantinescu, C., Cornea, A. Potential Theory on Harmonic Spaces, Springer-Verlag, Berlin, 1972. [8] Coscia, A., Mingione, G.: Holder continuity of the gradient of p(x)-harmonic mappings, C. R. Acad. Sci. Paris, Ser. 1 Math. 328(4) (1999), 363-368. [9] Diening, L., Hasto, P., Nekvinda, A. Open problems in variable exponent Lebesgue and Sobolev spaces, In Proceedings of the Function Spaces, Differential Operators and Nonlinear Analysis (FSDONA'04), Milovy, Czech, (2004), 38-58. THINNESS AND FINE TOPOLOGY WITH RELATIVE CAPACITY 13 [10] Diening, L., Harjulehto, P., Hasto, P., Ruzicka, M. Lebesgue and Sobolev Spaces with Variable Exponents, Springer-Verlag, Berlin, 2011. [11] Doob, J. L. Classical Potential Theory and Its Probabilistic Counterpart, Springer-Verlag, New-York, 1984. [12] Fan, X., Zhao, D. On the spaces Lp(x) (Ω) and W k,p(x) (Ω) , J. Math. Anal. Appl. 263(2) (2001), 424-446. [13] Fuglede, B. Fine Potential Theory, In Potential Theory-Surveys and Problems, pp. 82-97, Lecture Notes in Mathematics 1344, Springer-Verlag, Berlin, 1988. [14] Harjulehto, P., Latvala, V. Fine topology of variable exponent energy superminimizers, Ann. Acad. Sci. Fenn.-M. 33 (2008), 491-510. [15] Harjulehto, P., Hasto, P., Koskenoja, M. The Dirichlet energy integral on intervals in variable exponent Sobolev spaces, Z. Anal. Anwend. 22(4) (2003), 911-923. [16] Harjulehto, P., Hasto, P., Koskenoja, M. Properties of capacities in variable exponent Sobolev spaces, J. Anal. Appl. 5(2) (2007), 71-92. [17] Harjulehto, P., Hasto, P., Koskenoja, M., Varonen, S. The Dirichlet energy integral and variable exponent Sobolev spaces with zero boundary values, Potential Anal. 25 (2006), 205- 222. [18] Harjulehto, P., Hasto, P., Koskenoja, M., Lukkari, T., Marola, N. Obstacle problem and superharmonic functions with nonstandard growth, Nonlinear Anal. 67 (2007), 3424-3440. [19] Heinonen, J., Kilpelainen, T., Martio, O. Nonlinear Potential Theory of Degenerate Elliptic Equations, Oxford University Press, Oxford, 1993. [20] Helms, L. L. Introduction to Potential Theory, Pure and Applied Mathematics, Vol. XXII, Wiley-Interscience, New-York, 1969. [21] Kellogg, O. D. Foundations of Potential Theory, Springer, 1929. [22] Kilpelainen, T., Mal´y, J. Supersolutions to degenerate elliptic equation on quasi open sets, Commun. Part. Diff. Eq. 17 (1992), 455-464. [23] Kokilashvili, V., Samko, S. Singular integrals in weighted Lebesgue spaces with variable exponent, Georgian Math. J. 10(1) (2003), 145-156. [24] Kov´acik, O., R´akosn´ık, J. On spaces Lp(x) and W k,p(x), Czechoslovak Math. J. 41(116)(4) (1991), 592-618. [25] Mal´y, J., Ziemer, W. P. Fine Regularity of Solutions of Elliptic Partial Differential Equations, Math. Surveys and Monographs, 51, Amer. Math. Soc., Providence, RI, 1997. [26] Meyers, N. G. Continuity properties of potentials, Duke Math. J. 42 (1975), 157-166. [27] Musielak, J. Orlicz Spaces and Modular Spaces, Lecture Notes in Math. 1034. Springer-Verlag 1983. [28] Samko, S. On a progress in the theory of Lebesgue spaces with variable exponent: maximal and singular operators, Integr. Transf. Spec. F. 16(5-6) (2005), 461-482. [29] Unal, C., Aydın, I. On some properties of relative capacity and thinness in weighted variable exponent Sobolev spaces, Preprint, https://arxiv.org/pdf/1902.04777.pdf. [30] Unal, C., Aydın, I. The Riesz capacity in weighted variable exponent Lebesgue spaces, Int. J. Appl. Math. 30(2) (2017), 163-176. [31] Unal, C., Aydın, I. Weighted variable exponent Sobolev spaces with zero boundary values and capacity estimates, Sigma J. Eng. & Nat. Sci. 36(2) (2018), 373-388. Sinop University, Faculty of Arts and Sciences, Department of Mathematics E-mail address: [email protected] Sinop University, Faculty of Arts and Sciences, Department of Mathematics E-mail address: [email protected]
1109.0450
1
1109
2011-09-02T14:13:08
Further development of positive semidefinite solutions of the operator equation $\sum_{j=1}^{n}A^{n-j}XA^{j-1}=B$
[ "math.FA" ]
In \cite{Positive semidefinite solutions}, T. Furuta discusses the existence of positive semidefinite solutions of the operator equation $\sum_{j=1}^{n}A^{n-j}XA^{j-1}=B$. In this paper, we shall apply Grand Furuta inequality to study the operator equation. A generalized special type of $B$ is obtained due to \cite{Positive semidefinite solutions}.
math.FA
math
Further development of positive semidefinite solutions of the operator equationPn Jian Shi1 Zongsheng Gao j=1 An−jXAj−1 = B ∗ LMIB & School of Mathematics and Systems Science, Beihang University, Beijing, 100191, China Abstract In [7], T. Furuta discusses the existence of positive semidefinite solutions j=1 An−jXAj−1 = B. In this paper, we shall apply Grand Furuta inequality to study the operator equation. A generalized special type of B is obtained due to [7]. of the operator equation Pn Keywords: Furuta inequality; Grand Furuta inequality; operator equation; matrix equation; positive semidefinite operator; positive definite operator Mathematics Subject Classification: 15A24, 47A62, 47A63. 1 Introduction A capital letter T means a bounded linear operator on a Hilbert space. T > 0 and T > 0 mean a positive semidefinite operator and a positive definite operator, respectively. In the middle of last century, E. Heinz et al. studied operator theory and obtained the following famous theorem: Theorem LH (Lowner-Heinz inequality, [9] [8]). If A > B > 0, then Aα > Bα for any α ∈ [0, 1]. In 1987, T. Furuta proved the following operator inequality as an extension of The- orem LH: Theorem F (Furuta inequality, [4]). If A > B > 0, then for each r > 0, r 2 ApB r 2 ) 1 q > (B r 2 BpB r 2 ) 1 q (B (1.1) ∗This work is supported by National Natural Science Fund of China (10771011 and 11171013 ). 1Corresponding author. E-mail addresses: [email protected] (J. Shi), [email protected] (Z. Gao). 1 r 2 ApA r 2 ) 1 q > (A r 2 BpA r 2 ) 1 q (A (1.2) hold for p > 0, q > 1 with (1 + r)q > p + r. K. Tanahashi, in [10], proved the conditions p, q in Theorem F are best possible if r > 0. In 1995, T. Furuta showed another operator inequality which interpolates Theorem F: Theorem GF (Grand Furuta inequality, [5]). If A > B > 0 with A > 0, then for each t ∈ [0, 1] and p > 1, A1−t+r > {A holds for s > 1 and r > t. r 2 (A− t 2 BpA− t 2 )sA 1−t+r (p−t)s+r r 2 } (1.3) Afterwards, some nice proof of Grand Furuta inequality are shown, such as [2], [6]. K. Tanahashi, in [11], proved that the outer exponent value of (1.3) is the best possible. Later on, the proof was improved, see [3], [12]. Recently, T. Furuta proved the following theorem by Furuta inequality: Theorem A ([7]). If A is a positive definite operator and B is positive semidefinite operator. Let m and n be nature numbers. There exists positive semidefinite operator solution X of the following operator equation: An−j XAj−1 = A n Xj=1 nr 2(m+r) ( m Xi=1 n(m−i) m+r BA n(i−1) m+r )A A nr 2(m+r) (1.4) for r such that (r > 0, r > m−n n−1 , if n > m; if m > n > 2. In the rest of this short paper, we shall apply Grand Furuta inequality to discuss the j=1 An−jBAj−1 = B, existence of positive semidefinite solution of operator equation Pn and show a generalized special type of B due to Theorem A. 2 Extension of Furuta's result Lemma 2.1 ([1], [7]). Let A be a positive definite operator and B a positive semidef- inite operator. Let m be a positive integer and x > 0. Then d j=1 Am−j BAj−1. Pm 2 dx [(A + xB)m](cid:12)(cid:12)x=0 = Theorem 2.1. Let A be a positive definite operator and B be a positive semidefinite operator. Let m, n, k be positive integers, t ∈ [0, 1]. There exists positive semidefinite operator solution X which satisfies the operator equation: An−jXAj−1 =A 2(m−t)k+2r { nr n(m−t)(k−i) (m−t)k+r [A − t 2 · A n (m−t)k+r · ( n Xj=1 k Xi=1 − t 2 A · n (m−t)k+r ]A n(m−t)(i−1) (m−t)k+r }A nr 2(m−t)k+2r m Xj=1 n(m−j) (m−t)k+r BA A n(j−1) (m−t)k+r )· (2.1) for r such that (r > t, r > max{ (m−t)k−(1−t)n n−1 if (1 − t)n > (m − t)k ; if (m − t)k > (1 − t)n with n > 2 . , t}, Proof. As in the proof of [[7], Theorem 2.1], by A + xB > A > 0 holds for any x > 0, then A−1 > (A + xB)−1 > 0. Replace A by A−1, B by (A + xB)−1, p by m, s by k in (1.3), and take reverse, we have (A r 2 (A− t 2 (A + xB)mA− t 2 )kA r 2 ) 1−t+r (m−t)k+r > A1−t+r. (2.2) For any α ∈ [0, 1], apply Lowner-Heinz inequality to (2.2), and take 1 n = 1−t+r (m−t)k+r · α, the following inequality is obtained: (A r 2 (A− t 2 (A + xB)mA− t 2 )kA r 2 ) 1 n > A (m−t)k+r n . (2.3) By α ∈ [0, 1] and the condition of r in Grand Furuta inequality , we can take r > t if (1 − t)n > (m − t)k, or r > max{ (m−t)k−(1−t)n Take Y (x) = (A r 2 (A− t 2 (A + xB)mA− t n−1 2 )kA r 2 ) , t} if (m − t)k > (1 − t)n with n > 2. 1 n . According to (2.3), Y (x) > Y (0) = (m−t)k+r n A xB)mA− t for any x > 0, then Y ′(0) > 0. Differentiate Y n(x) = A r 2 (A− t 2 (A + 2 )kA r 2 , use Lemma 2.1, then take x = 0, the following equality holds: Y (0)n−j Y ′(0)Y j−1 = n Xj=1 2 (A + xB)mA− t d dx d dx = r [A 2 (A− t [Y n(x)](cid:12)(cid:12)(cid:12)(cid:12)x=0 Xi=1 [ (A− t k r = A 2 { k Xi=1 ·[ (A− t 2 (A + xB)mA− t r 2 )kA 2 ](cid:12)(cid:12)(cid:12)(cid:12)x=0 2 )k−i(cid:12)(cid:12)(cid:12)x=0 2 )i−1(cid:12)(cid:12)(cid:12)x=0 Xj=1 ]}A m r 2 3 [A(m−t)(k−i)(A− t 2 ( Am−j BAj−1)A− t 2 )A(m−t)(i−1)]}A r 2 . r = A 2 { 2 (A + xB)mA− t ] · [ (A− t 2 (A + xB)mA− t ] 2 )′(cid:12)(cid:12)(cid:12)x=0 Replace Y (0) by A (m−t)k+r n , Y ′(0) by X, we have (m−t)k+r n A (n−j)XA (m−t)k+r n (j−1) n Xj=1 k Xi=1 n m Xj=1 r = A 2 { A(m−t)(k−i)[A− t 2 ( Am−j BAj−1)A− t 2 ]A(m−t)(i−1)}A r 2 . Replace A by A (m−t)k+r in above equality, then (2.1) is obtained. (cid:3) Remark 2.1. If take t = 0 and k = 1 in Theorem 2.1, this theorem is just Theorem A, which is the main result of [7]. Example 2.1. We use the same example as [7]: For two l × l matrices A and B, take A = diag(a1, a2, . . . , a2), all entries of B are 1. If m, n, k are positive integers, t ∈ [0, 1], there exists positive semidefinite matrix X which satisfies: (m−t)k+r n A (n−j)XA (m−t)k+r n (j−1) n Xj=1 r 2 { = A Xi=1 for r such that (r > t, k A(m−t)(k−i)[A− t 2 ( m Xj=1 Am−j BAj−1)A− t 2 ]A(m−t)(i−1)}A r 2 r > max{ (m−t)k−(1−t)n n−1 if (1 − t)n > (m − t)k ; if (m − t)k > (1 − t)n with n > 2 . , t}, It is easy to calculate the expression of X: X = a r−t 2 p a r−t 2 q (cid:16)Σk i=1a(m−t)(k−i) p a(m−t)(i−1) q j=1am−j p ((m−t)k+r)(n−j) ((m−t)k+r)(j−1) Σn j=1a p n a q (cid:17)(cid:16)Σm n aj−1 q (cid:17) !p,q=1,2,...,l . (2.4) Remark 2.2. The condition of r in Theorem 2.1 is necessary. If the condition cannot be fulfilled, the solution of the equation may be not positive semidefinite. For example, take A =(cid:18)1 0 0 2(cid:19) , B =(cid:18)1 1 1 1(cid:19) , and m = 2, n = 2, k = 2, t = 1 r (cid:11) max{ (m−t)k−(1−t)n n−1 2 in Example 2.1. If we put r = 1 2 , then , t}. By (2.4), the solution of the following matrix equation 3 4 BA− 1 4 + A− 1 4 BA 3 1 4(cid:16)A 3 2(cid:0)A 5 2 B + A 3 2 BA + ABA 4(cid:1) +(cid:0)A 5 2 3 4 BA− 1 4 + A− 1 4 BA 3 4(cid:1)A 1 4 3 2(cid:17)A 7 4 X + XA 7 4 = A A = A = 4 3 + 6 × 2 1 2 3 + 6 × 2 1 16 × 2 2 ! 3 2 + BA 1 2 4 is 2 3+6×2 1+2×2 1 2 3 4 3+6×2 1+2×2 2 × 2 1 2 3 4 3 4 X =  .   However, X is not a positive semidefinite matrix because its eigenvalues are {5.4007 . . . , −0.0372 . . .}. Remark 2.3. In [1], the authors showed that if A and Y are positive semidefinite matrices in matrix equation An−1X + An−2XA + · · · + AXAn−2 + XAn−1 = Y , then so is X. By Theorem 2.1, in some special cases, if Y can be expressed as the right hand of (2.1), though it is not positive semidefinite, then there still exists positive semidefinite solution satisfies An−1X + An−2XA + · · · + AXAn−2 + XAn−1 = Y . For example, take A =(cid:18)1 0 2 × 2 0 1 3(cid:19) , Y = 4 3 × 2 1 4 + 6 × 2 3 4 3 × 2 1 4 + 6 × 2 3 4 32 ! . Although Y is not a positive semidefinite matrix (because its eigenvalues are {37.5589 . . . , −1.5589 . . .}) , by simple calculation, the solution of the following matrix equation is A2X + AXA + XA2 = Y 4 3 3×2 1+2×2 1 4 +6×2 3 4 1 3 +4×2 3×2 1 4 +6×2 3 4 1+2×2 1 3 +4×2 2 3 2 3 4×2 3 1 3 X =  ,   which is still a definite matrix whose eigenvalues are {2.9013 . . . , 0.1119 . . .}. The critical reason is that Y can be expressed as follows, Y = A 3 8 { 2 Xi=1 9 8 (2−i)[A− 3 16 ( A 2 Xj=1 3 4 (2−j)BA A 3 4 (j−1))A− 3 16 ]A 9 8 (i−1)}A 3 8 , which is the right hand of (2.1) under the condition of m = 2, n = 3, k = 2, t = 1 2 , r = 1. Remark 2.4. The following question remains open: How to investigate the properties of the solution of operator equation X n−1A + X n−2AX + · · · + XAX n−2 + AX n−1 = B? 5 References [1] R. Bhatia, M. Uchiyama, The operator equation Pn 27 (2009), 251-255. i=0 An−iBi = Y , Expo. Math., [2] M. Fujii, E. Kamei, Mean theoretic approach to the grand Furuta inequality, Proc. Amer. Math. Soc., 124 (1996), 2751-2756. [3] M. Fujii, A. Matsumoto, R. Nakamoto, A short proof of the best possibility for the grand Furuta inequality, J. Inequal. Appl., 4 (1999), 339-344. [4] T. Furuta, A > B > 0 assures (BrApBr)1/q > B(p+2r)/q for r > 0, p > 0, q > 1 with (1 + 2r)q > p + 2r, Proc. Amer. Math. Soc., 101 (1987), 85-88. [5] T. Furuta, An extension of the Furuta inequality and Ando-Hiai log majorization, Linear Algebra Appl., 219 (1995), 139-155. [6] T. Furuta, Simplified proof of an order preserving operator inequality, Proc. Japan Acad., 74 (1998), 114. [7] T. Furuta, Positive semidefinite solutions of the operator equation j=1 An−jXAj−1 = B, Linear Algebra Appl., 432 (2010), 949-955. Pn [8] E. Heinz, Beitrage zur Storungsteorie der Spektralzerlegung, Math. Ann., 123 (1951), 415-438. [9] K. Lowner, Uber monotone MatrixFunktionen, Math. Z., 38 (1934), 177-216. [10] K. Tanahashi, Best possibility of the Furuta inequality, Proc. Amer. Math. Soc., 124 (1996), 141-146. [11] K. Tanahashi, The best possibility of the grand Furuta inequality, Proc. Amer. Math. Soc., 128 (2000), 511-519. [12] T. Yamazaki, Simplified proof of Tanahashi's result on the best possibility of gen- eralized Furuta inequality, Math. Inequal. Appl., 2 (1999), 473-477. 6
1409.7565
4
1409
2015-04-17T14:23:08
Weyl Numbers of Embeddings of Tensor Product Besov Spaces
[ "math.FA" ]
In this paper we investigate the asymptotic behaviour of Weyl numbers of embeddings of tensor product Besov spaces into Lebesgue spaces. These results will be compared with the known behaviour of entropy numbers.
math.FA
math
Weyl Numbers of Embeddings of Tensor Product Besov Spaces Van Kien Nguyen & Winfried Sickel Friedrich-Schiller-University Jena, Ernst-Abbe-Platz 2, 07737 Jena, Germany [email protected] & [email protected] July 15, 2018 Abstract In this paper we investigate the asymptotic behaviour of Weyl numbers of em- beddings of tensor product Besov spaces into Lebesgue spaces. These results will be compared with the known behaviour of entropy numbers. 1 Introduction 5 1 0 2 r p A 7 1 ] . A F h t a m [ 4 v 5 6 5 7 . 9 0 4 1 : v i X r a Weyl numbers have been introduced by Pietsch [32]. They are relatives of approximation numbers. Recall, the n-th approximation number of the linear operator T ∈ L(X, Y ) is defined to be an(T ) := inf{kT − Ak : A ∈ L(X, Y ), rank(A) < n} , n ∈ N . (1.1) Here X and Y are quasi-Banach spaces. Now we are in position to define Weyl numbers. The n-th Weyl number of the linear operator T ∈ L(X, Y ) is given by xn(T ) := sup{an(T A) : A ∈ L(ℓ2, X), kAk ≤ 1} , n ∈ N . Approximation and Weyl numbers belong to the family of s-numbers, see Section 4 for more details. The particular interest in Weyl numbers stems from the fact that they are the smallest known s-number satisfying the famous Weyl-type inequalities, i.e., 2n−1 Yk=1 (cid:16) λk(T )(cid:17)1/(2n−1) ≤ √2e(cid:16) n Yk=1 xk(T )(cid:17)1/n (1.2) holds for all n ∈ N, in particular, λ2n−1(T ) ≤ √2e(cid:16) n Yk=1 xk(T )(cid:17)1/n , 1 see Pietsch [32] and Carl, Hinrichs [11]. Here T : X → X is a compact linear operator in a Banach space X and (λn(T ))∞ n=1 denotes the sequence of all non-zero eigenvalues of T , repeated according to algebraic multiplicity and ordered such that λ1(T ) ≥ λ2(T ) ≥ . . . ≥ 0 . Also as a consequence of (1.2) one obtains, for all p ∈ (0,∞), the existence of a constant cp (independent of T ) s.t. ∞ Xn=1 (cid:16) λn(T )p(cid:17)1/p ≤ cp(cid:16) ∞ Xn=1 xp n(T )(cid:17)1/p . Hence, Weyl numbers may be seen as an appropriate tool to control the eigenvalues of T . Many times operators of interest can be written as a composition of an identity between appropriate function spaces and a further bounded operator, see, e.g., the monographs of Konig [27] and of Edmunds, Triebel [18]. This motivates the study of Weyl numbers of identity operators. Pietsch [34], Lubitz [29], Konig [27] and Caetano [8, 9, 10] studied the Weyl numbers of id : Bt p1,q1((0, 1)d) denotes the isotropic Besov spaces. Zhang, Fang and Huang [65] and Gasiorowska and Skrzypczak [22] investigated the case of embeddings of weighted Besov spaces, defined on Rd, into Lebesgue spaces. Here we are interested in the investigation of the asymptotic behaviour p1,q1((0, 1)d) → Lp2((0, 1)d), where Bt of Weyl numbers of the identity id : St p1,p1B((0, 1)d) → Lp2((0, 1)d) , p1,p1B((0, 1)d) denotes a d-fold tensor product of univariate Besov spaces where St Bt p1,p1B((0, 1)d) can be interpreted as a special case of the scale of Besov spaces of dominating mixed p1,p1(0, 1). This notation is chosen in accordance with the fact that St smoothness, see Section 5. The behaviour of xn(id : St p1,p1B((0, 1)d) → Lp2((0, 1)d)), 1 < p2 < ∞, will be discussed in Subsection 3.1. Here, up to some limiting situations, we have the complete picture, i.e., we know the exact asymptotic behaviour of the Weyl numbers. For the extreme cases p2 = ∞ and p2 = 1, see Subsection 3.2 and Subsection 3.3, we are also able to describe In Subsection 3.4 we discuss the behaviour of the behaviour in almost all situations. xn(id : St mix((0, 1)d) denotes a space of Holder- Zygmund type. Finally, in Subsection 3.5, we compare the behaviour of Weyl numbers mix((0, 1)d)), where Z s p1,p1B((0, 1)d) → Z s with that one of entropy numbers. Summarizing we present an almost complete picture of the behaviour of xn(id : p1,p1B((0, 1)d) → Lp2((0, 1)d)), 0 < p1 ≤ ∞, 1 ≤ p2 ≤ ∞, t > max(0, 1 St ). This is a little bit surprising since the corresponding results for approximation numbers are much p1 − 1 p2 less complete, see, e.g., [5]. Let us mention in this context that Weyl numbers have some specific properties not shared by approximation numbers like interpolation properties, see Theorem 4.2, or the inequality (3.3). 2 The paper is organized as follows. Our main results are discussed in Section 3. In Section 4 we recall the definition of s-numbers and discuss some further properties of Weyl numbers. Section 5 is devoted to the function spaces under consideration. In Section 6 we inves- tigate the Weyl numbers of embeddings of certain sequence spaces associated to tensor product Besov spaces and spaces of dominating mixed smoothness. This will be followed by Section 7, where, beside others, Theorem 3.1 (our main result) will be proved. In Ap- p1 → ℓm pendix A we recall the behaviour of the Weyl numbers of embeddings idm p2. Finally, in Appendix B, a few more facts about the Lizorkin-Triebel spaces of dominating p,qF ((0, 1)d) and the Besov spaces of dominating mixed mixed smoothness St smoothness St p,qB((0, 1)d) are collected. p,qF (Rd), St p1,p2 : ℓm p,qB(Rd), St Notation As usual, N denotes the natural numbers, N0 := N ∪ {0}, Z the integers and R the real numbers. For a real number a we put a+ := max(a, 0). By [a] we denote the integer part of a. If ¯j ∈ Nd 0, i.e., if ¯j = (j1, . . . , jd), jℓ ∈ N0, ℓ = 1, . . . , d, then we put ¯j1 := j1 + . . . + jd . By Ω we denote the unit cube in Rd, i.e., Ω := (0, 1)d. If X and Y are two quasi-Banach spaces, then the symbol X ֒→ Y indicates that the embedding is continuous. As usual, the symbol c denotes positive constants which depend only on the fixed parameters t, p, q and probably on auxiliary functions, unless otherwise stated; its value may vary from line to line. Sometimes we will use the symbols "." and "&" instead of "≤" and "≥", respectively. The meaning of A . B is given by: there exists a constant c > 0 such that A ≤ c B. Similarly & is defined. The symbol A ≍ B will be used as an abbreviation of A . B . A. For a discrete set ∇ the symbol ∇ denotes the cardinality of this set. Finally, the symbols id, id∗ will be used for identity operators, id∗ mainly in connection with sequence spaces. The symbol idm p1,p2 refers to the identity idm p1,p2 : ℓm p1 → ℓm p2 . (1.3) Tensor products of Besov and Sobolev spaces are investigated in [50], [48], [49] and Hansen [24]. General information about Besov and Lizorkin-Triebel spaces of dominating mixed p,qF (Rd)). The smoothness can be found, e.g., in [1, 2, 3, 4, 46, 47, 61] (St (Fourier analytic) definitions of these spaces are reviewed in the Appendix B. The reader, p,qB(Rd), St who is interested in more elementary descriptions of these spaces, e.g., by means of differ- ences, is referred to [1, 47] and [60]. 2 Some preparations As a preparation for the main results we recall under which conditions the identity p1,p1B((0, 1)d) ֒→ Lp2((0, 1)d) is compact, see Vybiral [61, Thm. 3.17]. St 3 t > max(cid:16)0, 1 p1 − 1 p2(cid:17) . (2.1) Proposition 2.1. The following assertions are equivalent: (i) The embedding St p1,p1B((0, 1)d) ֒→ Lp2((0, 1)d) is compact; (ii) We have Since we are exclusively interested in compact embeddings the restriction (2.1) will be always present. Also for later use, we recall the Weyl numbers of the embedding id : Bt p1,q1(0, 1) → Lp2(0, 1). Let 0 < p1, q1 ≤ ∞, 1 ≤ p2 ≤ ∞ and t > max(0, 1/p1 − 1/p2). Then, in all cases listed in Prop. 2.2, we have xn(id) = xn(id : Bt p1,q1(0, 1) → Lp2(0, 1)) ≍ n−α , n ∈ N . (2.2) Here the value of α = α(t, p1, p2) is given in the following proposition. Proposition 2.2. The value of α in (2.2) is given by I II III IV ∗ IV∗ V ∗ V∗ α = t α = t + 1 α = t + 1 2 p2 − 1 p2 − 1 p2 − 1 p1 p1 α = t + 1 α = tp1 2 + 1 2 p1 α = t − 1 α = tp1 2 if p1, p2 ≤ 2; if p1 ≤ 2 ≤ p2; if 2 ≤ p1 ≤ p2; if 2 ≤ p2 < p1 and t > 1/p2−1/p1 p1/2−1 ; if 2 ≤ p2 < p1 < ∞ and t < 1/p2−1/p1 p1/2−1 ; if p2 ≤ 2 < p1 and t > 1 if p2 ≤ 2 < p1 < ∞ and t < 1 p1 p1 . ; 1 p2 1 1 2 V IV 0 III 1 2 I II 1 1 p1 Figure 1 The above results indicate a decom- position of the (1/p1, 1/p2)-plane into five parts. In regions IV and V we have a further splitting into the cases of small (IV∗, V∗) and large smoothness (IV ∗, V ∗). Proposition 2.2 has been proved by Pietsch [34] and Lubitz [29, Satz 4.13] in case 1 ≤ p1, q1, p2 ≤ ∞, we refer also to Konig [27] in this context. The proof of the general case, i.e., also for values of p1, q1 less than 1, can be found in the 4 thesis of Caetano [10], see also [9]. Obviously we do not have a dependence on the fine- index q1. This will be different in the dominating mixed case with d > 1. The behaviour of the Weyl numbers in the limiting situations t = 1/p2−1/p1 (see IV∗, IV ∗) and t = 1/p1 p1/2−1 (see V∗, V ∗) is open, in particular it seems to be unknown whether the behaviour is still polynomial in n. 3 The main results It seems to be appropriate to split our considerations into the three cases: (i) 1 < p2 < ∞, (ii) p2 = ∞ and (iii) p2 = 1. 3.1 The Littlewood-Paley case Littlewood-Paley analysis is one of the main tools to understand the behaviour of the Weyl numbers if 1 < p2 < ∞ (i.e., the target space Lp2 allows a Littlewood-Paley-type decomposition). The cases p2 = 1 and p2 = ∞ require different techniques and will be treated in the next subsections. As in the isotropic case the results suggest to work with the same decomposition of the (1/p1, 1/p2)-plane as in Proposition 2.2. So, the symbols I, II, . . ., used below, have the same meaning as in Figure 1 (and therefore as in Prop. 2.2). In addition the regions I ∗ and I∗ are given by p1, p2 ≤ 2. Let 0 < p1 ≤ ∞, 1 < p2 < ∞ and t > max(0, 1/p1 − 1/p2). Then in all cases, listed in Theorem 3.1, we have xn(id) = xn(id : St p1,p1B((0, 1)d) → Lp2((0, 1)d)) ≍ n−α (log n)(d−1) β , n ≥ 2 . (3.1) The values of α = α(t, p1, p2) and β = β(t, p1, p2) are be given in the following theorem. Theorem 3.1. The values of α and β in (3.1) are given by α = t and β = t + 1 α = t and β = 0 if t > 1 p1 − 1 2 ; p1 2 − 1 if t < 1 2 + 1 p2 and β = t + 1 + 1 p2 + 1 p2 and β = t + 1 and β = t + 1 ; if t > 1/p2−1/p1 p1/2−1 ; 2 and β = t + 1 2 − 1 p1 t < 1/p2−1/p1 p1/2−1 ; + 1 2 and β = t + 1 2 and β = t + 1 2 − 1 p1 p1 2 − 1 if if t > 1 p1 ; t < 1 p1 . I ∗ I∗ II III IV ∗ IV∗ V ∗ V∗ p1 α = t − 1 α = t − 1 α = t − 1 α = tp1 p1 p1 α = t − 1 α = tp1 ; p1 p1 − 1 2 ; p2 − 1 p2 − 1 p2 − 1 if p1 p1 5 Thm. 3.1 gives the final answer about the behaviour of the xn in almost all cases. It is interesting to notice that in regions I, IV and V we have a different behaviour for small smoothness compared with large smoothness. Only in the resulting limiting cases we are not able to characterize the behaviour of the xn(id). However, estimates from below and above also for these limiting situations will be given in Subsection 6.3. In essence the proof is standard. Concerning the estimate from above the first step consists p1,p1B((0, 1)d) → Lp2((0, 1)d) to id∗ : st,Ω in a reduction step. By means of wavelet characterizations we switch from the consider- ation of id : St st,Ω (the id∗ xn(id∗ : st,Ω p,q b and µ=0 id∗ µ µ are identities with respect to certain subspaces) which results in an estimate of p,q f are appropriate sequence spaces. Next, this identity is splitted into id∗ =P∞ p1,p1b → s0,Ω p2,2f , where st,Ω p1,p1b → s0,Ω p2,2f ) xn(id∗) ≤ J Xµ=0 xnµ(id∗ µ) + L Xµ=J+1 xnµ(id∗ µ) + ∞ Xµ=L+1 kid∗ µk µ=0(nµ − 1), see (6.2). Till this point we would call the proof standard, compare, e.g., with Vybiral [62]. But now the problem consists in choosing J, L and where n − 1 =PL nµ in a way leading to the desired result. This is the real problem which we solved in Subsection 6.3. In a further reduction step estimates of xnµ(id∗ µ) are traced back to estimates of xnµ(idDµ p1,p2), see (1.3). All what is needed about these number is collected in Appendix A. Concerning the estimate from below one has to figure out appropriate subspaces of St known estimates of xn(idDµ p1,p1b). Then, also in this case, all can be reduced to the p1,p1B((0, 1)d) (st,Ω p1,p2). 3.2 The extreme case p2 = ∞ Let us recall a result of Temlyakov [53], see also [14]. Proposition 3.2. Let t > 1 2 . Then we have xn(id : St 2,2B((0, 1)d) → L∞((0, 1)d)) ≍ (log n)(d−1)t nt− 1 2 , n ≥ 2 . Remark 3.3. (i) In the literature many times the notation H t are used instead of St (ii) In [53] and [14] the authors deal with approximation numbers an(id : St 2,2B((0, 1)d) → L∞((0, 1)d)), see (1.1). However, for Banach spaces Y and Hilbert spaces H we always have mix((0, 1)d) and M W t 2,2B((0, 1)d). 2((0, 1)d) xn(T : H → Y ) = an(T : H → Y ) , see [35, Prop. 2.4.20]. By using specific properties of Weyl numbers we will extent Prop. 3.2 to the following result. 6 Theorem 3.4. Let 0 < p1 ≤ ∞. Then we have xn(id : St p1,p1B((0, 1)d) → L∞((0, 1)d)) (log n) (d−1)(t+ 1 2 − 1 p1 nt− 1 2 (log n) (d−1)(t+ 1 2 − 1 p1 t− 1 p1 n (3.2) ) ) if if 0 < p1 ≤ 2 , t > 1 p1 ; 2 < p1 ≤ ∞ , t > 1 2 + 1 p1 ; ≍   for all n ≥ 2. Remark 3.5. (i) Recall that St p1,p1B((0, 1)d) is compactly embedded into L∞((0, 1)d) if and only if t > 1/p1, see Prop. 2.1. The cases 2 < p1 ≤ ∞ and t ∈ (cid:0) 1 2(cid:3) remain open. (ii) Considering p2 → ∞ in parts II and III of Thm. 3.1 then it turns out that in (3.2) there is an additional log factor, more exactly (log n)(d−1)/2. (iii) To prove Theorem 3.4 we shall employ an inequality due to Pietsch [33]. For any , 1 p1 + 1 p1 linear operator T we have n1/2 xn(T ) ≤ π2(T ) , where π2(T ) refers to the 2-summing norm of T . (iv) There is a small number of cases, where the exact order of sn(id : St p1,q1B((0, 1)d) → L∞((0, 1)d)), p1 6= 2, if n tends to infinity, has been found. Here sn stands for any s-number, see Sect. 4. Beside Prop. 3.2 we refer to Temlyakov [54] where n ∈ N , (3.3) dn(id : St ∞,∞B(T2) → L∞(T2)) ≍ n−t (log n)t+1 , t > 0, is proved for all n ≥ 2. Here dn denotes the n-th Kolmogorov number and T2 refers to the two-dimensional periodic case. For some partial results (i.e., with a gap between the estimates from above and below) with respect to Kolmogorov numbers we refer to Romanyuk [43]. 3.3 The extreme case p2 = 1 Let us recall a result obtained by Romanyuk [44] (again Romanyuk has dealt with ap- proximation numbers, but see Remark 3.3 for this). Proposition 3.6. Let t > 0. Then we have xn(id : St 2,2B((0, 1)d) → L1((0, 1)d)) ≍ (log n)(d−1)t nt , n ≥ 2 . By making use of the embedding S0 1,2F ((0, 1)d) ֒→ L1((0, 1)d) we are able to extend Prop. 3.6 to the following. 7 Theorem 3.7. Let 0 < p1 ≤ ∞ and t > ( 1 xn(id : St p1,p1B((0, 1)d) → L1((0, 1)d)) p1 − 1)+. Then n−t n−t(log n)(d−1)(t+ 1 n−t+ 1 n− tp1 − 1 2 (log n)(d−1)(t+ 1 2 − 1 p1 2 (log n)(d−1)(t+ 1 2 − 1 p1 p1 ) ) 2 − 1 p1 if if if if ) t < 1 t > 1 p1 < 2 , p1 ≤ 2 , 2 < p1 ≤ ∞ , 2 < p1 < ∞ , p1 − 1 2 , p1 − 1 2 , t > 1 p1 t < 1 p1 , , ≍   for all n ≥ 2. Remark 3.8. The most interesting case is given by p1 = 1. It follows that we have xn(id : St 1,1B((0, 1)d) → L1((0, 1)d)) ≍( n−t n−t(log n)(d−1)(t− 1 2 ) if if t < 1 2 , t > 1 2 , for all n ≥ 2. We are not aware of any other result concerning s-numbers (Kol- mogorov numbers, approximation numbers, ...) where the exact order of sn(id : 1,1B((0, 1)d) → L1((0, 1)d)) if n tends to infinity, has been found. A few more results St concerning approximation and Kolmogorov numbers are known in case of the embed- dings id : St 1,1B((0, 1)d) → Lp2((0, 1)d), 1 < p2 < ∞. E.g., in [44] Romanyuk has proved for 2 ≤ p1 < ∞ and t > 0 2 − 1 p1,p1B((0, 1)d) → L1((0, 1)d)) ≍ n−t(log n)(d−1)(t+ 1 p1 p1,p1B((0, 1)d) → L1((0, 1)d), p1 > 1, and id : St an(id : St ) for all n ≥ 2. 3.4 A version of Holder-Zygmund spaces (related to tensor products) as target spaces As a supplement we investigate St p1,p1B((0, 1)d) → Z s Zygmund spaces. Let j ∈ {1, . . . d}. For m ∈ N, hj ∈ R and x ∈ Rd we put mix((0, 1)d), where the spaces Z s : mix((0, 1)d) are versions of Holder- the Weyl numbers of the embeddings id ∆m hj,jf (x) := m (−1)m−ℓ(cid:18)m Xℓ=0 ℓ(cid:19) f (x1, . . . , xj−1, xj + ℓhj, xj+1, . . . , xd) . This is the m-th order difference in direction j. Mixed differences are defined as follows. Let e be a non-trivial subset of {1, . . . d}. For h ∈ Rd we define ∆m h,e :=Yj∈e ∆m hj,j . Of course, here ∆m hj,j · ∆m hℓ,ℓ has to be interpreted as ∆m hj,j ◦ ∆m hℓ,ℓ. Definition 3.9. Let s > 0. Let m ∈ N s.t. m − 1 ≤ s < m. Then f ∈ Z s k f Z s mix((0, 1)d)k := k f C((0, 1)d)k + max hj−s e⊂{1,...,d} mix((0, 1)d) if sup khk∞≤1Yj∈e 8 sup x∈Ωm,e,h(cid:12)(cid:12)(cid:12) ∆m h,ef (x)(cid:12)(cid:12)(cid:12) < ∞ , where Ωm,e,h := {x ∈ (0, 1)d : and (x1 + ε1ℓ1h1, . . . , xd + εdℓdhd) ∈ (0, 1)d ∀¯ℓ ∈ Nd 0 , k ¯ℓk∞ ≤ m} , εj :=( 1 0 if if j ∈ e , j 6∈ e . A few properties of these spaces are obvious: • Let d = 1 and m = 1, i.e., 0 < s < 1. Then Z s Holder-continuous functions of order s. mix(0, 1) is the classical space of • Let d = 1, s = 1 and m = 2. Then Z 1 • If f (x) = f1(x1)· . . .·fd(xd) with fj ∈ Zmix(0, 1), j = 1, . . . , d, then f ∈ Z s mix(0, 1) is the classical Zygmund space. mix((0, 1)d) follows and k f Z s mix((0, 1)d)k ≍ d Yj=1 k fj Z s mix(0, 1)k . • Let f ∈ Z s Then g ∈ Z s • We define Z s mix((0, 1)d) and define g(x′) := f (x′, 0), where x = (x′, xd), x′ ∈ Rd−1. mix((0, 1)d−1) follows. mix(Rd) by replacing (0, 1)d by Rd in the Def. 3.9. Let E : Z s(0, 1) → Z s(R) be a linear and bounded extension operator such that E maps C(0, 1) into itself. Then E⊗. . .⊗E (d-fold tensor product) is well-defined on C((0, 1)d) and maps this space into itself. Observe that ∆m h,ef (x) can be written as the e-fold iteration of a directional difference. As a consequence we obtain that Ed := E ⊗ . . . ⊗ E maps Z s mix((0, 1)d) into itself. Less obvious is the following lemma, see [47, Rem. 2.3.4/3] or [60]. Lemma 3.10. Let s > 0. Then holds in the sense of equivalent norms. mix((0, 1)d) = Ss Z s ∞,∞B((0, 1)d) Essentially by the same methods as used for the proof of Thm. 3.1 one obtains the following. Theorem 3.11. Let s > 0 and t > s + 1 p1 . Then it holds xn(id : St p1,p1B((0, 1)d) → Z s mix((0, 1)d)) ≍ for all n ≥ 2.   (log n) (d−1)(t−s− 1 p1 nt−s− 1 2 (log n) (d−1)(t−s− 1 p1 t−s− 1 p1 n ) ) if if 0 < p1 ≤ 2 , 2 ≤ p1 ≤ ∞ , 9 p1,p1B((0, 1)d) is compactly embedded into Z s Remark 3.12. (i) Recall that St if and only if t > s + 1/p1, see [61]. (ii) Observe, that Thm. 3.11 is not the limit of Thm. 3.4 for s ↓ 0. There, in Thm. 3.4, is an additional factor (log n)(d−1)/2 as many times in this field. (iii) If we replace Z s s > 0 becomes superfluous. ∞,∞B((0, 1)d) in Theorem 3.11, then the restriction mix((0, 1)d)) by Ss mix((0, 1)d) Finally, we wish to mention that these methods also apply in case of approximation numbers. As a result we get the following. Theorem 3.13. Let n ∈ N, n ≥ 2 and s > 0. Then we have an(id : St p1,p1B((0, 1)d) → Z s mix((0, 1)d)) (log n) (d−1)(t−s− 1 p1 t−s− 1 p1 n (log n) (d−1)(t−s− 1 p1 nt−s− 1 2 (d−1)(t−s− 1 p1 (log n) p′ 2 (t−s− 1 1 p1 n ) ) ) ) if 2 ≤ p1 ≤ ∞ , t − s > 1 p1 , if 1 ≤ p1 < 2 , t − s > 1 , if 1 < p1 < 2 , 1 > t − s > 1 p1 , ≍   1 is the conjugate of p1. for all n ≥ 2. Here p′ Remark 3.14. As in Remark 3.12, if we replace Z s Theorem 3.13, then the restriction s > 0 becomes superfluous. mix((0, 1)d)) by Ss ∞,∞B((0, 1)d) in 3.5 A comparison with entropy numbers There are good reasons to compare Weyl numbers with entropy numbers. Both, entropy and Weyl numbers, are tools to control the behaviour of eigenvalues of linear operators. Let us recall the definition of entropy numbers. Definition 3.15. Let T : X → Y be a bounded linear operator between complex quasi- Banach spaces, and let n ∈ N. Then the n-th (dyadic) entropy number of T is defined as en(T : X → Y ) := inf{ε > 0 : T (BX) can be covered by 2n−1 balls in Y of radius ε} , where BX := {x ∈ X : kxkX ≤ 1} denotes the closed unit ball of X. In particular, T : X → Y is compact if and only if limn→∞ en(T ) = 0 . For details and basic properties like multiplicativity, additivity, behaviour under interpolation etc. we refer to the monographs [12, 18, 27, 32]. Most important for us is the Carl-Triebel inequality which states √2 en(T ) , λn(T ) ≤ 10 cf. Carl, Triebel [13] (see also the monographs [12] and [18]). Entropy numbers of embeddings id : St p1,p1B((0, 1)d) ֒→ Lp2((0, 1)d) have been inves- tigated in Vybiral [61]. The picture is less complete than in case of Weyl numbers. Only for sufficiently large smoothness the behaviour is exactly known. For 0 < p1 ≤ ∞ and 1 < p2 < ∞ we have en(id) ≍ n−t(log n)(d−1)(t+ 1 2 − 1 p1 ) t > max(cid:16)0, 1 p1 − 1 2 , 1 p1 − 1 p2(cid:17) , n ≥ 2 . We use Figure 2 to ex- plain the different be- haviour of entropy and Weyl numbers. Weyl numbers are essentially smaller than entropy numbers in regions IV and V, entropy numbers are essentially smaller than Weyl numbers in regions II and III, and they show a similar be- haviour in region I∗. if I 1 p2 1 1 2 V lim n→∞ xn en = 0 IV xn en lim n→∞ = 0 III lim n→∞ en xn = 0 0 I ∗ : en ≍ xn II lim n→∞ en xn = 0 1 2 Figure 2 1 1 p1 Remark 3.16. (i) Further estimates of the decay of entropy numbers related to embed- dings id : St p1W ((0, 1)d) ֒→ Lp2((0, 1)d)) can be p1,q1B((0, 1)d) ֒→ Lp2((0, 1)d) (id : St found in Belinsky [6], Dinh Dung [16], and Temlyakov [51]. (ii) There are many contributions dealing with the behaviour of dn(id : St p1,q1B((0, 1)d) ֒→ Lp2((0, 1)d)) (Kolmogorov numbers) and an(id : St p1,q1B((0, 1)d) ֒→ Lp2((0, 1)d)). How- ever, the picture is much less complete than in case of Weyl numbers. We refer to Bazarkhanov [5] for the most recent publication in this direction. The topic itself has been investigated at various places over the last 30 years, see, e.g., Temylakov [51, 52], Galeev [19, 20, 21] and Romanyuk [36, 37, 38, 39, 40, 41, 42, 43, 44]. 4 Weyl numbers - basic properties Weyl numbers are special s-numbers. For later use we recall this general notion following Pietsch [35, 2.2.1] (note that this differs slightly from earlier definitions in the literature). Let X, Y, X0, Y0 be quasi-Banach spaces. As usual, L(X, Y ) denotes the space of all continuous linear operators from X to Y . Finally, let Y be p-Banach space for some 11 p ∈ (0, 1], i.e., k x + y Y kp ≤ k xY kp + k y Y kp for all x, y ∈ Y . (4.1) An s-function is a map s assigning to every operator T ∈ L(X, Y ) a scalar sequence (sn(T )) such that the following conditions are satisfied: (a) kTk = s1(T ) ≥ s2(T ) ≥ ... ≥ 0 for all T ∈ L(X, Y ); (b) sp n(S) + sp n+m−1(S + T ) ≤ sp m(T ) for S, T ∈ L(X, Y ) and m, n = 1, 2, . . . ; (c) sn(BT A) ≤ kBk · sn(T ) · kAk for A ∈ L(X0, X), T ∈ L(X, Y ), B ∈ L(Y, Y0); (d) sn(T ) = 0 if rank(T ) < n for all n ∈ N; (e) sn(id : ℓn 2 → ℓn 2 ) = 1 for all n ∈ N. We will refer to (a) as monotonicity, to (b) as additivity, to (c) as ideal property, to (d) as the rank property and to (e) as normalization (norm-determining property) of the s- numbers. Sometimes a further property is of some use. Let Z be a quasi-Banach space. An s- function is called multiplicative if (f ) sn+m−1(ST ) ≤ sn(S) sm(T ) for T ∈ L(X, Y ), S ∈ L(Y, Z) and m, n = 1, 2, . . . . Examples The following numbers are s-numbers: (i) Kolmogorov numbers are multiplicative s-numbers, see, e.g., [32, Thm. 11.9.2]. (ii) Approximation numbers are multiplicative s-numbers, see, e.g., [35, 2.3.3]. (iii) The n-th Gelfand number of the linear operator T ∈ L(X, Y ) is defined to be cn(T ) := infnk T J X M k : codim (M ) < no, where J X are multiplicative s-numbers, see, e.g., [35, Prop. 2.4.8]. M : M → X refers to the canonical injection of M into X. Gelfand numbers (iv) Weyl numbers are multiplicative s-numbers, see [35, 2.4.14, 2.4.17]. Entropy numbers do not belong to the class of s-numbers since they do not satisfy (d). Remark 4.1. There is an alternative way to calculate the n-th Weyl number. Indeed, for T ∈ L(X, Y ) it holds xn(T ) := supncn(T A) : A ∈ L(ℓ2, X), kAk ≤ 1o , see Pietsch [33]. 12 Interpolation properties of Weyl numbers For later use we add the following assertion concerning interpolation properties of Weyl numbers. Theorem 4.2. Let 0 < θ < 1. Let X, Y, Y0, Y1 be a quasi-Banach spaces. Further we assume Y0 ∩ Y1 ֒→ Y and the existence of a positive constant C such that kyY k ≤ C kyY0k1−θkyY1kθ for all y ∈ Y0 ∩ Y1. (4.2) Then, if we obtain T ∈ L(X, Y0) ∩ L(X, Y1) ∩ L(X, Y ) xn+m−1(T : X → Y ) ≤ C x1−θ (T : X → Y0) xθ for all n, m ∈ N. Here C is the same constant as in (4.2). Remark 4.3. Interpolation properties of Kolmogorov and Gelfand numbers have been m(T : X → Y1) n studied by Triebel [55]. Theorem 4.2 and Theorem 7.9 below show that Gelfand and Weyl numbers share the same interpolation properties. 5 Tensor product Besov spaces and spaces of dominating mixed smoothness As mentioned before tensor product Besov spaces can be interpreted as special cases of the scale of Besov spaces of dominating mixed smoothness. For us it will be convenient to introduce these classes of dominating mixed smoothness by means of wavelets. In the Appendix B below we recall the probably better known Fourier-analytic definition. In addition we shall introduce Lizorkin-Triebel spaces of dominating mixed smoothness. They will be used in our proofs of the main results for Besov spaces. Let ¯ν = (ν1, . . . , νd) ∈ Nd 0 and ¯m = (m1, . . . , md) ∈ Zd. Then we put 2−¯ν ¯m = (2−ν1m1, ..., 2−νd md) and Q¯ν, ¯m :=nx ∈ Rd : 2−νℓ mℓ < xℓ < 2−νℓ (mℓ + 1) , ℓ = 1, . . . , do . By χ¯ν, ¯m(x) we denote the characteristic function of Q¯ν, ¯m. First we have to introduce some sequence spaces. 0, ¯m ∈ Zd}, then we Definition 5.1. If 0 < p, q ≤ ∞, t ∈ R and λ := {λ¯ν, ¯m ∈ C : ¯ν ∈ Nd define q < ∞o p(cid:17) λ¯ν, ¯mp(cid:1) p,qb :=nλ : kλst p,qbk =(cid:16) X¯ν∈Nd p )q(cid:0) X¯m∈Zd 2¯ν1(t− 1 1 q st 0 13 and, if p < ∞, st p,qf =nλ : kλst p,qfk =(cid:13)(cid:13)(cid:13)(cid:16) X¯ν∈Nd 0 X¯m∈Zd 2¯ν1tλ¯ν, ¯mχ¯ν, ¯m(.)q(cid:17) 1 q(cid:12)(cid:12)(cid:12) Lp(Rd)(cid:13)(cid:13)(cid:13) with the usual modification for p or/and q equal to ∞. Remark 5.2. Let σ ∈ R. For later use we mention that the mapping yields an isomorphism of st Jσ : (λ¯ν, ¯m)¯ν, ¯m 7→ (2σ¯ν1 λ¯ν, ¯m)¯ν, ¯m p,qa onto st−σ p,q a, a ∈ {b, f}. < ∞o (5.1) Now we recall wavelet bases of Besov and Lizorkin-Triebel spaces of dominating mixed smoothness. Let N ∈ N. Then there exists ψ0, ψ1 ∈ C N (R), compactly supported, −∞ tm ψ1(t) dt = 0 , Z ∞ j ∈ N0, m ∈ Z}, where such that {2j/2 ψj,m : m = 0, 1, . . . , N , if if j = 0, m ∈ Z , j ∈ N , m ∈ Z , ψj,m(t) :=( ψ0(t − m) p1/2 ψ1(2j−1t − m) is an orthonormal basis in L2(R), see [63]. We put Then Ψ¯ν, ¯m(x) := ψνℓ,mℓ(xℓ) . d Yℓ=1 Ψ¯ν, ¯m , ¯ν ∈ Nd 0, ¯m ∈ Zd , is a tensor product wavelet basis of L2(Rd). Vybiral [61] has proved the following. Lemma 5.3. Let 0 < p, q ≤ ∞ and t ∈ R. (i) There exists N = N (t, p) ∈ N s.t. the mapping f 7→ (2¯ν1hf, Ψ¯ν, ¯mi)¯ν∈Nd 0 , ¯m∈Zd W : p,qB(Rd) onto st is an isomorphism of St (ii) Let p < ∞. Then there exists N = N (t, p, q) ∈ N s.t. the mapping W is an isomor- phism of St p,qF (Rd) onto st p,qf . p,qb. Spaces on Ω We put Ω := (0, 1)d. For us it will be convenient to define spaces on Ω by restrictions. We shall need the set D′(Ω), consisting of all complex-valued distributions on Ω. 14 Definition 5.4. (i) Let 0 < p, q ≤ ∞ and t ∈ R. Then St f ∈ D′(Ω) such that there exists a distribution g ∈ St endowed with the quotient norm p,qB((0, 1)d) is the space of all p,qB(Rd) satisfying f = gΩ. It is k f St p,qB((0, 1)d)k = infnkgSt p,qB(Rd)k : (ii) Let 0 < p < ∞, 0 < q ≤ ∞ and t ∈ R. Then St f ∈ D′(Ω) such that there exists a distribution g ∈ St endowed with the quotient norm gΩ = fo . p,qF ((0, 1)d) is the space of all p,qF (Rd) satisfying f = gΩ. It is k f St p,qF ((0, 1)d)k = infnkgSt p,qF (Rd)k : gΩ = fo . Several times we shall work with the following consequence of this definition in combi- nation with Lemma 5.3. Let t, p and q be fixed. Let the wavelet basis Ψ¯ν, ¯m be admissible in the sense of Lemma 5.3. We put For given f ∈ St AΩ ¯ν :=n ¯m ∈ Zd : supp Ψ¯ν, ¯m ∩ Ω 6= ∅o , p,qA(Ω), A ∈ {B, F}, let Ef be an element of St kEf St p,qA(Rd)k ≤ 2k f St p,qA(Ω)k and ¯ν ∈ Nd 0 . p,qA(Rd) s.t. (Ef )Ω = f . (5.2) We define g := X¯ν∈Nd 0 X¯m∈AΩ ¯ν 2¯ν1 hEf, Ψ¯ν, ¯mi Ψ¯ν, ¯m . Then it follows that g ∈ St supp g ⊂ {x ∈ Rd : max p,qA(Rd), gΩ = f , j=1,... ,dxj ≤ c1} and k g St p,qA(Rd)k ≤ c2 k f St p,qA(Ω)k . Here c1, c2 are independent of f . Tensor products of Besov spaces Tensor products of Besov spaces have been investigated in [25], [48] and [49]. We recall some results from [48] and [49]. For the basic notions of tensor products used here we refer to [28] and [15]. By αp we denote the p-nuclear norm and by γp the projective tensor p-norm. Theorem 5.5 (Tensor products of Besov spaces on the interval). Let d ≥ 1 and let t ∈ R. (i) Let 1 < p < ∞. Then the following formula Bt p,p(0, 1) ⊗αp St p,pB((0, 1)d) = St = St p,pB((0, 1)d) ⊗αp Bt p,pB((0, 1)d+1) p,p(0, 1) 15 holds true in the sense of equivalent norms. (ii) Let 0 < p ≤ 1. Then the following formula Bt p,p(0, 1) ⊗γp St p,pB((0, 1)d) = St = St p,pB((0, 1)d) ⊗γp Bt p,pB((0, 1)d+1) p,p(0, 1) holds true in the sense of equivalent quasi-norms. Remark 5.6. For easier notation we put γp := αp if 1 < p < ∞. One can iterate the process of taking tensor products. Defining for m > 2 X1 ⊗γp X2 ⊗γp . . . ⊗γp Xm := X1 ⊗γp (cid:16) . . . Xm−2 ⊗γp (Xm−1 ⊗γp Xm)(cid:17) we obtain an interpretation of St univariate Besov spaces, namely p,pB((0, 1)d), 0 < p < ∞, as an iterated tensor product of St p,pB((0, 1)d) = Bt p,p(0, 1) ⊗γp . . . ⊗γp Bt p,p(0, 1) , 0 < p < ∞ . The iterated tensor products, considered in this paper, do not depend on the order of the tuples which are formed during the process of calculating X1 ⊗γp X2 ⊗γp . . . ⊗γp Xm, i.e., (X1 ⊗γp X2) ⊗γp X3 = X1 ⊗γp (X2 ⊗γp X3) . Consequently, if p < ∞, we may deal with St Bt p,p(0, 1). p,pB((0, 1)d) instead of Bt p,p(0, 1) ⊗γp . . . ⊗γp 6 Weyl numbers of embeddings of sequence spaces In this section we will estimate the behavior of Weyl numbers of the identity mapping id∗ : p0,p0b → s0,Ω st,Ω p,2 f . Here we assume that p0 varies in (0,∞] and p in (0,∞). 6.1 Preparations For technical reasons we need a few more sequence spaces. Recall, AΩ ¯ν has been defined in (5.2). Definition 6.1. If 0 < p ≤ ∞, 0 < q ≤ ∞, t ∈ R and λ = {λ¯ν, ¯m ∈ C : ¯ν ∈ Nd 0, ¯m ∈ AΩ ¯ν } , then we define st,Ω p,q b :=nλ : kλst,Ω p,q bk =(cid:16) X¯ν∈Nd 0 2¯ν1(t− 1 p )q(cid:0) X¯m∈AΩ ¯ν λ¯ν, ¯mp(cid:1) q p(cid:17) 1 q < ∞o 16 and, if p < ∞, p,q f :=nλ : st,Ω kλst,Ω p,q fk =(cid:13)(cid:13)(cid:13)(cid:16) X¯ν∈Nd 0 X¯m∈AΩ ¯ν 2¯ν1tλ¯ν, ¯mχ¯ν, ¯m(.)q(cid:17) 1 q(cid:12)(cid:12)(cid:12) Lp(Rd)(cid:13)(cid:13)(cid:13) < ∞o . In addition we need the following sequence of subspaces. Definition 6.2. If 0 < p ≤ ∞, 0 < q ≤ ∞, t ∈ R, µ ∈ N0 and λ = {λ¯ν, ¯m ∈ C : ¯ν ∈ Nd 0, ¯ν1 = µ, ¯m ∈ AΩ ¯ν } , then we define (st,Ω p,q b)µ =nλ : kλ(st,Ω p,q b)µk =(cid:16) X¯ν1=µ and, if p < ∞, 2¯ν1(t− 1 p )q(cid:0) X¯m∈AΩ ¯ν λ¯ν, ¯mp(cid:1) q p(cid:17) 1 q < ∞o 1 (st,Ω To avoid repetitions we shall use st 2¯ν1tλ¯ν, ¯mχ¯ν, ¯m(.)q(cid:17) p,q f )µ =nλ : kλ(st,Ω p,q f )µk =(cid:13)(cid:13)(cid:13)(cid:16) X¯ν1=µ X¯m∈AΩ < ∞o . p,q a)µ with a ∈ {b, f} in case that an assertion holds for both scales simultaneously. Here in this paper we do not deal with the spaces st,Ω or (st,Ω elementary lemmas are taken from [61, Lemma 3.10] and [24, Lemma 6.4.2]. ∞,qf )µ. But we will use the convention that, whenever st,Ω ∞,qa ∞,qb)µ. The two following ∞,qa)µ occur, this has to be interpreted as st,Ω Lp(Rd)(cid:13)(cid:13)(cid:13) ∞,qf and (st,Ω p,qa, st,Ω p,q a, (st,Ω ∞,qb and (st,Ω q(cid:12)(cid:12)(cid:12) ¯ν Lemma 6.3. (i) We have AΩ ¯ν ≍ 2¯ν1, Dµ := X¯ν1=µ with equivalence constants independent of ¯ν ∈ Nd AΩ ¯ν ≍ µd−12µ 0 and µ ∈ N0. (ii) Let 0 < p < ∞ and t ∈ R. Then p,pf = st,Ω st,Ω p,pb and (st,Ω p,pf )µ = (st,Ω p,pb)µ = 2µ(t− 1 p )ℓDµ p , µ ∈ N0 , with the obvious interpretation for the quasi-norms. Lemma 6.4. (i) Let 0 < p0, p ≤ ∞ and 0 < q ≤ ∞. Then p,q a)µk ≍ 2µ( 1 p0,qa)µ → (st,Ω µ : (st,Ω k id∗ p0 − 1 p )+ with equivalence constants independent of µ ∈ N0. 17 (ii) Let 0 < q0, q ≤ ∞ and 0 < p ≤ ∞. Then p,q0a)µ → (st,Ω µ : (st,Ω kid∗ p,q a)µk ≍ µ(d−1)( 1 q − 1 q0 )+ with equivalence constants independent of µ ∈ N0. Corollary 6.5. Let 0 < p0, p, q0, q ≤ ∞ and t ∈ R. Then p,q a)µk . 2µ(cid:0)−t+( 1 − 1 p0,q0a)µ → (s0,Ω µ : (st,Ω k id∗ p0 p )+(cid:1)µ(d−1)( 1 q − 1 q0 )+, with a constant behind . independent of µ. Proof . This is an immediate consequence of Lemma 6.4. (cid:4) Sometimes the previous estimate can be improved. Lemma 6.6. Let 0 < p0 < p < ∞, 0 < q0, q ≤ ∞ and t ∈ R. Then − 1 p ). p0 p0,q0f )µ → (s0,Ω p,q f )µk . 2µ(−t+ 1 µ : (st,Ω kid∗ Proof . This assertion is contained in [24]. Since this phd is not published we give a proof. Let λ be a sequence such that λ¯ν, ¯m = 0 if ¯ν1 6= µ. Since p0 < p the Sobolev-type embedding yields st,Ω p0,q0f ֒→ s see [47, Thm. 2.4.1](d = 2) or [26], we have + 1 p ,Ω t− 1 p0 p,q f , kλ(s0,Ω p,q f )µk = kλs0,Ω . 2µ(−t+ 1 p0 p,q fk = 2µ(−t+ 1 p )kλst,Ω − 1 p0 − 1 p )kλs + 1 p ,Ω t− 1 p0 p,q p0,q0fk = 2µ(−t+ 1 p0 − 1 fk p )kλ(st,Ω p0,q0f )µk . This proves the claim. (cid:4) 6.2 Weyl numbers of embeddings of sequence spaces related to spaces of dominating mixed smoothness - preparations For µ ∈ N0 we define where id∗ µ : st,Ω p0,p0b → s0,Ω p,2 f , (id∗ µλ)¯ν, ¯m :=  λ¯ν, ¯m if ¯ν1 = µ, otherwise. 0 The main idea of our proof is the following splitting of id∗ : st,Ω identities between building blocks p0,p0b → s0,Ω p,2 f into a sum of id∗ = ∞ Xµ=0 id∗ µ = J Xµ=0 id∗ µ + L Xµ=J+1 id∗ µ + ∞ Xµ=L+1 id∗ µ, (6.1) 18 where J and L are at our disposal. These numbers J and L will be chosen in dependence on the parameters. Let us mention that a similar splitting has been used by Vybiral [61] for the estimates of related entropy numbers. The additivity and the monotonicity of the Weyl numbers and the quasi-triangle inequality (4.1) yield J xρ n(id∗) ≤ Xµ=0 where n − 1 = PL brevity we put xρ nµ(id∗ µ) + L Xµ=J+1 xρ nµ(id∗ µ) + µ=0(nµ − 1). Of course, kid∗ α = t −(cid:18) 1 p0 − ρ := min(1, p) , (6.2) µ : (st,Ω p0,p0f )µ → (s0,Ω p,2 f )µk. For µkρ, ∞ kid∗ Xµ=L+1 µk = k id∗ p(cid:19)+ 1 . Then by Corollary 6.5, we have kid∗ µk . 2−µα µ(d−1)( 1 2 − 1 p0 )+, which results in the estimate ∞ Xµ=L+1 kid∗ µkρ . 2−LαρL(d−1)ρ( 1 2 − 1 p0 )+ . Now we choose nµ Then we get nµ := Dµ + 1, µ = 0, 1, ...., J . J Xµ=0 nµ ≍ J Xµ=0 µ(d−1)2µ ≍ J d−12J and xnµ(id∗ µ) = 0, see the rank property of the s-numbers, which implies nµ(id∗ xρ µ) = 0 . J Xµ=0 Summarizing (6.2)-(6.6) we have found (6.3) (6.4) (6.5) (6.6) n(id∗) . xρ L Xµ=J+1 xρ nµ(id∗ µ) + 2−LαρL(d−1)ρ( 1 2 − 1 p0 )+ . (6.7) Now we turn to the problem to reduce the estimates for the Weyl numbers xnµ(id∗ estimates for xn(idm µ) to p0,p). Proposition 6.7. Let 0 < p0 ≤ ∞ and t ∈ R. Then we have the following assertions. (i) If 0 < p ≤ 2, then p + 1 µ(d−1)(− 1 2 )2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . xn(id∗ µ) . 2µ(−t+ 1 p0 − 1 2 ) xn(idDµ p0,2). (6.8) 19 (ii) If 2 ≤ p < ∞, then 2µ(−t+ 1 p0 − 1 2 ) xn(idDµ p0,2) . xn(id∗ µ) . µ(d−1)( 1 2 − 1 p )2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . (6.9) Proof . Step 1. Estimate from above. We define δ := max(p, 2) and consider the following diagram: (st,Ω p0,p0b)µ id∗ µ (s0,Ω p,2 f )µ id2 id1 (s0,Ω δ,δ f )µ Using property (c) of the s-numbers we conclude xn(id∗ µ) ≤ k id1 k xn(id2) . By Corollary 6.5, we have From Lemma 6.3 (ii), we derive kid1k . µ(d−1)( 1 2 − 1 δ ) . xn(id2) . 2µ(−t+ 1 p0 − 1 δ ) xn(idDµ p0,δ) , taking into account property (c) of the s-numbers and the commutative diagram 2µ(t− 1 p0 ) (ℓDµ p0 ) id3 −−−−→ ℓDµ p0 id Dµ p0,δ y ℓDµ δ , id2y 2− µ δ (ℓDµ δ ) ←−−−−id4 i.e., id2 = id4 ◦ idDµ p0,δ ◦ id3, k id3 k = 2−µ(t− 1 p0 ) and k id4 k = 2 µ δ . Altogether this implies xn(id∗ µ) . µ(d−1)( 1 2 − 1 δ ) 2µ(−t+ 1 p0 − 1 δ ) xn(idDµ p0,δ) . Step 2. Now we turn to the estimate from below. We define γ := min(p, 2) and use the following commutative diagram (st,Ω p0,p0b)µ id2 (s0,Ω γ,γ f )µ id∗ µ id1 (s0,Ω p,2 f )µ 20 This time we have xn(id2) ≤ k id1 k xn(id∗ µ) and by Corollary 6.5, we get k id1 k . µ(d−1)( 1 γ − 1 2 ) . Similarly as in Step 1 Lemma 6.3 (ii) yields xn(id2) & 2µ(−t+ 1 p0 − 1 γ ) xn(idDµ p0,γ) . Inserting this in our previous estimate we find xn(id∗ µ) & µ(d−1)( 1 2 − 1 γ ) 2µ(−t+ 1 p0 − 1 γ ) xn(idDµ p0,γ) . The proof is complete. (cid:4) Proposition 6.8. Let 0 < p0 ≤ ∞ and t ∈ R. Then we have the following assertions. (i) If 0 < p ≤ 2, then xn(id∗ µ) . 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . (ii) If 2 ≤ p < ∞, then 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . xn(id∗ µ) . (6.10) (6.11) Proof . Step 1. Proof of (i). We consider the following diagram (st,Ω p0,p0b)µ id∗ µ (s0,Ω p,2 f )µ id2 id1 (s0,Ω p,p f )µ This implies xn(id∗ 6.3 we derive µ) ≤ k id1 k xn(id2). Corollary 6.5 yields kid1k . 1 and from Lemma xn(id2) . 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . Altogether we have found xn(id∗ µ) . 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . Step 2. Proof of (ii). We use the following diagram (st,Ω p0,p0b)µ id2 (s0,Ω p,p f )µ id∗ µ id1 (s0,Ω p,2 f )µ 21 Because of xn(id2) ≤ k id1 k xn(id∗ µ), k id1 k . 1, see Corollary 6.5, and xn(id2) & 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) , (see Lemma 6.3), we obtain xn(id∗ µ) & 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . The proof is complete. We need a few more results of the above type. Lemma 6.9. Let 0 < p0, p < ∞ and 0 < ǫ < p. Then xn(id∗ µ) . 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p−ǫ) . Proof . We consider the following diagram (st,Ω p0,p0b)µ id∗ µ (s0,Ω p,2 f )µ id2 id1 (sr,Ω p−ǫ,p−ǫf )µ Clearly, xn(id∗ µ) ≤ k id1 k xn(id2) and by Lemma 6.6 we have (cid:4) (6.12) k id1 k . 2µ(−r+ 1 p−ǫ − 1 p ) . Further we know xn(id2) = 2µ(r− 1 p−ǫ −t+ 1 p0 ) xn(idDµ p0,p−ǫ) . Inserting the previous inequality in this identity we obtain (6.12). (cid:4) Lemma 6.10. For all µ ∈ N0 and all n ∈ N we have µ) ≤ xn(id∗) . xn(id∗ Proof . We consider the following diagram (6.13) st,Ω p0,p0b id1x (st,Ω p0,p0b)µ id∗ −−−−→ s0,Ω p,2 f id∗ µ−−−−→ (s0,Ω p,2 f )µ . id2 y Here id1 is the canonical embedding and id2 is the canonical projection. Since id∗ id2 ◦ id∗ ◦ id1 the property (c) of the s-numbers yields µ = xn(id∗ µ) ≤ k id1 kk id2 k xn(id∗) = xn(id∗) . This completes the proof. (cid:4) 22 6.3 Weyl numbers of embeddings of sequence spaces related to spaces of dominating mixed smoothness - results Now we are in position to deal with the Weyl numbers of id∗ : st,Ω p,2 f . We have to continue with the proof already started in (6.1)-(6.7). Therefore we need to distinguish p0,p0b → s0,Ω several cases. Always the positions of p0 and p relative to 2 are of importance. 6.3.1 The case 0 < p0 ≤ 2 ≤ p < ∞ Theorem 6.11. Let 0 < p0 ≤ 2 ≤ p < ∞ and t > 1 p (log n)(d−1)(t+ 1 2 − 1 xn(id∗) ≍ n−t+ 1 n ≥ 2 . p . Then p0 − 1 p − 1 ) , p0 Proof . Step 1. Estimate from below. Since p ≥ 2, from (6.13) and (6.11) we derive xn(id∗) & 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . Next we choose n = [ Dµ (a) in Appendix A we get 2 ] (here [ x ] denotes the integer part of x ∈ R). Then from property xn(idDµ p0,p) & (Dµ) which implies 1 2 − 1 p0 ≍ (2µ µd−1) 1 2 − 1 p0 , xn(id∗) & 2µ(−t+ 1 2 − 1 p ) µ(d−1)( 1 2 − 1 p0 ) . Because of 2µ ≍ n logd−1 n we conclude xn(id∗) & n−t+ 1 2 − 1 p (log n)(d−1)(t− 1 p0 + 1 p ) . Step 2. Estimate from above. Let L, J and α as in (6.1)-(6.3). By our assumptions we obviously have 2−αLL(d−1)( 1 2 − 1 p0 )+ = 2L(−t+ 1 p0 − 1 p ) . For given J we choose L > J large enough such that 2L(−t+ 1 p0 − 1 p ) . J (d−1)( 1 2 − 1 p0 ) 2J(−t+ 1 2 − 1 p ) . For the sum in (6.7), we define nµ := [Dµ 2(J−µ)λ] ≤ Dµ 2 , J + 1 ≤ µ ≤ L , where λ > 1 is at our disposal. We choose λ such that which is always possible under the given restrictions. Then 1 2 t − + 1 p > λ(cid:18) 1 p0 − 1 2(cid:19) L Xµ=J+1 nµ ≍ J d−12J 23 (6.14) (6.15) (6.16) follows. If p > 2, we choose ǫ > 0 such that 2 ≤ p − ǫ. From (6.12) we obtain xnµ(id∗ µ) . 2µ(−t+ 1 p0 − 1 p ) xnµ(idDµ p0,p−ǫ). If p = 2, then (6.9) implies xnµ(id∗ µ) . 2µ(−t+ 1 p0 − 1 2 ) xnµ(idDµ p0,2) . Employing property (a) in Appendix A we obtain xnµ(id∗ µ) . 2µ(−t+ 1 = µ(d−1)( 1 p0 2 − 1 p0 − 1 p )(cid:16)µd−1 2µ 2(J−µ)λ(cid:17) p ) 2(J−µ)λ( 1 ) 2µ(−t+ 1 2 − 1 1 2 − 1 p0 2 − 1 p0 ) . (6.17) Our special choice of λ in (6.15) yields xρ nµ(id∗ µ) . J (d−1)ρ( 1 2 − 1 p0 ) 2Jρ(−t+ 1 2 − 1 p ) . L Xµ=J+1 Inserting (6.14) and (6.17) into (6.7) leads to n(id∗) . J (d−1)ρ( 1 xρ 2 − 1 p0 Notice ) 2Jρ(−t+ 1 2 − 1 p ) . n = 1 + L Xµ=0 (nµ − 1) = 1 + Dµ + J Xµ=0 L ([Dµ 2(J−µ)λ] − 1) ≍ J d−1 2J , Xµ=J+1 see (6.4), (6.5) and (6.16). Hence, our proof works for a certain subsequence (nJ )∞ the natural numbers. More exactly, with J=1 of nJ := 1 + Dµ + J Xµ=0 L ([Dµ 2(J−µ)λ] − 1) , Xµ=J+1 J ∈ N , and L = L(J) chosen as the minimal admissible value in (6.14) we find xnJ (id∗) . J (d−1)( 1 2 − 1 p0 ) 2J(−t+ 1 2 − 1 p ) . We already know A J d−1 2J ≤ nJ ≤ B J d−1 2J , J ∈ N , for suitable A, B > 0. Without loss of generality we assume B ∈ N. Then we conclude from the monotonicity of the Weyl numbers xB J d−1 2J (id∗) . log(B J d−1 2J )(d−1)( 1 2 − 1 p0 )(cid:16) B J d−1 2J logd−1(B J d−1 2J )(cid:17)−t+ 1 2 − 1 p . Employing one more times the monotonicity of the Weyl numbers and in addition its polynomial behaviour we can switch from the subsequence (B J d−1 2J )J to n ∈ N in this formula by possibly changing the constant behind .. This finishes our proof. (cid:4) 24 6.3.2 The case 2 ≤ p0 ≤ p < ∞ Theorem 6.12. Let 2 ≤ p0 ≤ p < ∞ and t > 1 xn(id∗) ≍ n−t+ 1 p0 p0 − 1 − 1 p (log n)(d−1)(t+ 1 p . Then p − 1 p0 ) , n ≥ 2 . Proof . Step 1. Estimate from below. We apply the same arguments as in proof of the previous theorem. However, notice that xn(idDµ (a) in Appendix A. With n = [Dµ/2] and xn(idDµ p0,p) has a different behaviour, see property p0,p) & 1 we conclude that xn(id∗) & 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) & 2µ(−t+ 1 p0 − 1 p ) , see (6.13) and (6.11). Because of 2µ ≍ n logd−1 n this results in the estimate xn(id∗) & n−t+ 1 p0 − 1 p (log n)(d−1)(t− 1 p0 + 1 p ) . Step 2. Estimate from above. For J ∈ N and λ ∈ st,Ω Xµ=0 X¯ν1=µ X¯m∈AΩ SJ λ := J ¯ν p0,p0b we put λ¯ν, ¯me¯ν, ¯m , where {e¯ν, ¯m, ¯ν ∈ Nd 0, ¯m ∈ AΩ ¯ν } is the canonical orthonormal basic of s0,Ω 2,2 b. Obviously kid∗ − SJ : st,Ω p0,p0b → s0,Ω p,2 fk ≤ ∞ Xµ=J+1 kid∗ µ : (st,Ω p0,p0b)µ → (s0,Ω p,2 f )µk. Using Lem. 6.6 and (st,Ω p0,p0b)µ = (st,Ω p0,p0f )µ we get kid∗ − SJ : st,Ω p0,p0b → s0,Ω p,2 fk ≤ 2−µ(t− 1 p0 + 1 p ) . 2−J(t− 1 p0 + 1 p ) . ∞ Xµ=J+1 Because of rank(SJ ) ≍ 2J J d−1 we conclude in case n = 2J J d−1 that an(id∗) . 2−J(t− 1 p0 + 1 p ) . Since xn ≤ an we can complete the proof of the estimate from above by arguing as at the end of the proof of Thm. 6.11. (cid:4) 6.3.3 The case 2 ≤ p < p0 ≤ ∞ Theorem 6.13. Let 2 ≤ p < p0 ≤ ∞ and t > 1/p−1/p0 − 1 p (log n)(d−1)(t+ 1 p0 xn(id∗) ≍ n−t+ 1 p0/2−1 . Then p − 1 p0 ) , n ≥ 2 . Proof . Step 1. Estimate from below. Because of p > 2, (6.13) and (6.11) imply xn(id∗) & 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . 25 We choose n = [Dµ/2]. Then property (b)(part(iii)) in Appendix A yields xn(idDµ Hence p0,p) & 1. xn(id∗) & 2µ(−t+ 1 p0 − 1 p ) . n logd−1 n Because of 2µ ≍ this implies the desired estimate. Step 2. Estimate from above. Since 2 ≤ p < p0 we obtain )+ = 2−Lt L(d−1)( 1 2−αL L(d−1)( 1 2 − 1 p0 2 − 1 p0 ) . For given J we choose L large enough such that 2−LtL(d−1)( 1 2 − 1 p0 ) ≤ 2−γJt (6.18) for some γ > 1 (to be chosen later on). We define nµ := [Dµ 2(J−µ)β ] ≤ Dµ, J + 1 ≤ µ ≤ L, where the parameter β > 1 will be also chosen later on. Hence L Xµ=J+1 nµ ≍ 2J J d−1 . The restriction t > 1/p−1/p0 p0/2−1 implies −t + 1 p0 − 1 p + 1/p − 1/p0 1 − 2/p0 < 0 . If p > 2 we choose ǫ > 0 such that 2 ≤ p − ǫ and − t + 1 p0 − 1 p + p0 1 p−ǫ − 1 1 − 2 p0 < 0 . (6.19) In this situation we derive from property (b)(part(i)) in Appendix A p−ǫ − 1 1 − 2 p0,p−ǫ) .(cid:18) Dµ nµ(cid:19) ≍ 2− (J −µ)β xnµ(idDµ 1 r := p0 1 r 1 , r p0 The estimate (6.12) guarantees L Xµ=J+1 nµ(id∗ xρ µ) . 2µρ(−t+ 1 p0 L Xµ=J+1 − 1 p ) xρ nµ(idDµ p0,p−ǫ) . In case p = 2, again property (b)(part(i)) in Appendix A yields xnµ(idDµ p0,2) .(cid:18) Dµ nµ(cid:19) 1 2 ≍ 2− (J −µ)β r , 1 r := 1 2 . From (6.10) we obtain L Xµ=J+1 xρ nµ(id∗ µ) . 2µρ(−t+ 1 p0 L Xµ=J+1 26 − 1 2 ) xρ nµ(idDµ p0,2) . . (6.20) (6.21) Now (6.20) and (6.21) yield L Xµ=J+1 nµ(id∗ xρ µ) . = L Xµ=J+1 Xµ=J+1 L The condition (6.19) can be rewritten as 2µρ(−t+ 1 p0 − 1 p ) 2− (J −µ)βρ r 2µρ(−t+ 1 p0 − 1 p + β r ) 2− J βρ r Now we choose β > 1 such that −t + 1 p + β r < 0. Then 1 p + 1 r < 0 . −t + 1 p0 − p0 − 1 nµ(id∗ xρ µ) . 2Jρ(−t+ 1 p0 L Xµ=J+1 − 1 p + β r ) 2− J βρ r = 2Jρ(−t+ 1 p0 follows. Inserting this and (6.18) into (6.7) we find Choosing then we conclude p ) + 2−γJtρ(cid:17) . p0 − 1 xn(id∗) .(cid:16)2Jρ(−t+ 1 p0 − 1 γ := −t + 1 −t p > 1 . − 1 p ) xn(id∗) . 2J(−t+ 1 p0 − 1 p ) and this is enough to prove the estimate from above, compare with the end of the proof of Thm. 6.11. (cid:4) Theorem 6.14. Let 2 ≤ p < p0 < ∞ and 0 < t < 1/p−1/p0 ) , 2 (log n)(d−1)(t+ 1 2 − 1 p0 xn(id∗) ≍ n− tp0 p0/2−1 . Then n ≥ 2 . Proof . Step 1. Estimate from below. From (6.9) and (6.13) we derive 2µ(−t+ 1 p0 − 1 2 ) xn(idDµ p0,2) . xn(id∗) . Now we choose n = [D 2 p0 µ ]. Then it follows from property (b)(part(ii)) in Appendix A that xn(idDµ 1 2 − 1 p0 p0,2) & D µ & (µd−12µ) 1 2 − 1 p0 . This implies xn(id∗) & µ(d−1)( 1 2 − 1 p0 ) 2−tµ . Rewriting the right-hand side in dependence on n we obtain xn(id∗) & n− tp0 2 (log n)(d−1)(t+ 1 2 − 1 p0 ) . 27 Step 2. Estimate from above. Since 2 ≤ p < p0 we have 2−αL L(d−1)( 1 2 − 1 p0 )+ = 2−tL L(d−1)( 1 2 − 1 p0 ) . For fixed J ∈ N we choose L :=h p0 2 J + (d − 1)( p0 2 − 1) log Ji . 2−Lt = 2−t([ p0 2 J+(d−1)( p0 2 −1) log J]) ≍ 2− p0 2 Jt J (d−1)(t− tp0 2 ) Hence and L(d−1)( 1 2 − 1 p0 ) =(cid:16)h p0 2 J + (d − 1)( p0 2 − 1) log Ji(cid:17)(d−1)( 1 2 − 1 p0 ) . J (d−1)( 1 2 − 1 p0 ) . This results in the estimate 2−Lt L(d−1)( 1 2 − 1 p0 ) . 2− p0 2 Jt J (d−1)(t− tp0 2 + 1 2 − 1 p0 ) . (6.22) We define where β > 0 will be fixed later on. Consequently nµ :=(cid:2)Dµ 2{(µ−L)β+J−µ}(cid:3) ≤ Dµ , J + 1 ≤ µ ≤ L , L Xµ=J+1 nµ . 2J J d−1 . Employing property (b)(part(i)) in Appendix A we get xnµ(idDµ p0,p) .(cid:16) Dµ nµ(cid:17) We continue by applying (6.9) 1 r . 2− (µ−L)β+J −µ r , 1 r := 1/p − 1/p0 1 − 2/p0 . µ(d−1)ρ( 1 2 − 1 p ) 2µρ(−t+ 1 p0 µ(d−1)ρ( 1 2 − 1 p ) 2µρ(−t+ 1 p0 µ(d−1)ρ( 1 2 − 1 p ) 2µρ(−t+ 1 p0 − 1 p ) xρ nµ(idDµ p0,p) − 1 p ) 2− {(µ−L)β+J −µ}ρ r − 1 p + 1 r − β r ) 2 (Lβ−J )ρ r . L Xµ=J+1 xρ nµ(id∗ L L . µ) . Xµ=J+1 Xµ=J+1 Xµ=J+1 1/p − 1/p0 p0/2 − 1 = L Because of t < we can choose β > 0 such that −t + 1 p + 1 r − β ⇐⇒ p0 − 1 + 1 p 1 p0 − −t + r > 0. Then 1 r > 0 L Xµ=J+1 xρ nµ(id∗ µ) . L(d−1)ρ( 1 2 − 1 p ) 2Lρ(−t+ 1 p0 = L(d−1)ρ( 1 2 − 1 p ) 2Lρ(−t+ 1 p0 28 − 1 p + 1 r − β r ) 2 (Lβ−J )ρ r − 1 p + 1 r ) 2− J ρ r (6.23) (6.24) (6.25) follows. Inserting the definition of L we conclude L(d−1)( 1 2 − 1 p ) . J (d−1)( 1 2 − 1 p ) and 2L(−t+ 1 p0 Now (6.25) yields − 1 p + 1 r ) 2− J r . 2(cid:2) p0 . 2− tp0 = 2− tp0 2 J+(d−1)( p0 2 J J (d−1)( p0 2 J J (d−1)(t− tp0 2 −1) log J(cid:3)(cid:2)−t+ 1 2 −1)(−t+ 1 p0 − 1 p0 2 − 1 p0 + 1 p ) . − 1 p + 1 r(cid:3)2− J r p + 1 r ) xρ nµ(id∗ µ) . J (d−1)ρ(t− tp0 2 − 1 p0 + 1 2 ) 2− tp0 2 Jρ . L Xµ=J+1 This, together with (6.22), has to be inserted into (6.7) xn(id∗) . J (d−1)(t− tp0 2 − 1 p0 + 1 2 ) 2− tp0 2 J . The same type of arguments as at the end of the proof of Thm. 6.11 complete the proof. (cid:4) Remark 6.15. Without going into details we mention the following estimate for the limiting case t = 1/p−1/p0 p0/2−1 . For all n ≥ 2 we have n− tp0 2 (log n)(d−1)(t+ 1 2 − 1 p0 ) . xn(id∗) . n− tp0 2 (log n)(d−1)(t+ 1 2 − 1 p0 )(log n) 1 r + 1 ρ , where r is as in (6.24) and ρ = min(1, p). 6.3.4 The case 0 < p0, p ≤ 2 We need some preparations. Lemma 6.16. Let 0 < p0, p ≤ 2 and t > ( 1 n−t(log n)(d−1)(t+ 1 p0 − 1 2 − 1 p0 p )+. Then )+ . xn(id∗) holds for all n ≥ 2. Proof . Step 1. We consider the following commutative diagram st,Ω p0,p0b id1x 2µ(t− 1 p0 )ℓAµ p0 id∗ −−−−→ s0,Ω p,2 f y Iµ−−−−→ 2µ(0− 1 id2 p )ℓAµ p . Here Aµ = AΩ the canonical projection. From property (c) of the s-numbers we derive ¯ν for some ¯ν with ¯ν1 = µ, id1 is the canonical embedding, whereas id2 is xn(Iµ) = xn(id2 ◦ id∗ ◦ id1) ≤ kid1kkid2k xn(id∗) = xn(id∗) . 29 Again the ideal property of the s-numbers guarantees xn(Iµ) = 2µ(−t+ 1 p0 − 1 p ) xn(idAµ p0,p) . We choose n = [Aµ/2]. Then property (a) in Appendix A yields xn(Iµ) ≥ 2µ(−t+ 1 p0 − 1 p ) xn(idAµ p0,p) & 2µ(−t+ 1 p0 − 1 p ) 2µ( 1 p − 1 p0 which implies xn(id∗) & n−t. This proves the lemma if t + 1 Step 2. From (6.13) and (6.8) we have 2 − 1 ) = 2−µt ≍ n−t , p0 ≤ 0. xn(id∗) & µ(d−1)( 1 2 − 1 p ) 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . We choose n := [Dµ/2]. Then property (a) in Appendix A leads to xn(idDµ p − 1 p0 p0,p) & D µ 1 & (µd−12µ) 1 p − 1 p0 , which implies xn(id∗) & µ(d−1)( 1 2 − 1 p0 ) 2−tµ . Because of 2µ ≍ n logd−1 n this yields xn(id∗) & n−t(log n)(d−1)(t+ 1 2 − 1 p0 ) . The proof is complete. (cid:4) Lemma 6.17. If 0 < p0, p ≤ 2 and t > 1 p0 − 1 2 . Then xn(id∗) . n−t(log n)(d−1)(t+ 1 2 − 1 p0 ) holds for all n ≥ 2. Proof . The restriction t > 1 2 implies the following chain of continuous embeddings p0 − 1 p0,p0b ֒→ s0,Ω st,Ω 2,2 f ֒→ s0,Ω p,2 f . Now we consider the commutative diagram st,Ω p0,p0b id∗ s0,Ω p,2 f id1 id2 s0,Ω 2,2 f The ideal property of the s-numbers and Thm. 6.11 (applied with p = 2) yield the claim. (cid:4) 30 Lemma 6.18. Let 0 < p ≤ p0 < 2 and 0 < t < 1 xn(id∗) . n−t p0 − 1 2 . Then holds for all n ≥ 1. Proof . For given J ∈ N we choose L := J + (d − 1) [log J]. Then )+ = 2−Lt ≍ 2−tJ J (d−1)(−t) . 2−Lα L(d−1)( 1 2 − 1 p0 We define for some β > 0. Then (6.23) follows. Property (a) in Appendix A yields nµ :=(cid:2)Dµ 2(µ−L)β+J−µ(cid:3) , J + 1 ≤ µ ≤ L , xnµ(idDµ p0,2) .(cid:16)Dµ2(µ−L)β+J−µ(cid:17) This, in connection with (6.8), leads to 1 2 − 1 p0 . L Xµ=J+1 nµ(id∗ xρ µ) . . 2µρ(−t+ 1 p0 2µρ(−t+ 1 p0 L Xµ=J+1 Xµ=J+1 L − 1 2 )(cid:16)Dµ 2(µ−L)β+J−µ(cid:17)ρ( 1 2 − 1 p0 − 1 2 +( 1 2 − 1 p0 )β)(cid:16)µ(d−1)2−Lβ+J(cid:17)ρ( 1 2 − 1 p0 ) . Because of t < 1 p0 − 1 2 we can select β > 0 such that (6.26) ) Consequently L Xµ=J+1 −t + 1 p0 − 1 2 1 2 − 1 p0(cid:1)β > 0 . +(cid:0) nµ(id∗ xρ µ) . 2Lρ(−t+ 1 p0 = 2Lρ(−t+ 1 p0 . 2Lρ(−t+ 1 p0 − 1 2 +( 1 2 − 1 p0 )β)(cid:16)L(d−1) 2−Lβ+J(cid:17)ρ( 1 2 − 1 p0 ) − 1 − 1 ) ) 2 − 1 p0 2 )(cid:16)L(d−1) 2J(cid:17)ρ( 1 2 )(cid:16)J (d−1) 2J(cid:17)ρ( 1 2 )(cid:16)J (d−1) 2J(cid:17)ρ( 1 2 − 1 p0 − 1 = 2Lρ(−t) 2Lρ( 1 p0 2 − 1 p0 ) . (6.27) Observe 2L( 1 p0 − 1 2 )(cid:16)J (d−1) 2J(cid:17) 1 2 − 1 p0 = 2(J+(d−1)[log J])( 1 p0 − 1 2 )(cid:16)J (d−1) 2J(cid:17) 1 2 − 1 p0 ≍ 1 . Replacing L by J + (d − 1)[log J] in (6.27) we obtain L Xµ=J+1 xρ nµ(id∗ µ) . 2−Lρt . (2J J d−1)−ρt. 31 This inequality, together with (6.26), yield xnJ (id∗) . n−t J , where nJ := 1 + Dµ + J Xµ=0 L Xµ=J+1(cid:16)(cid:2)Dµ 2(µ−L)β+J−µ(cid:3) − 1(cid:17) , Now we can continue as at the end of the proof of Thm. 6.11. J ∈ N . (cid:4) It remains to investigate the following situation: 0 < p0 < p < 2 and 1 p0− 1 2 . The estimates of the Weyl numbers xn(id∗) from above will be the most complicated part within this paper. p < t < 1 p0− 1 Lemma 6.19. Let 0 < p0 < p < 2 and 1 p0 − 1 p < t < 1 p0 − 1 2 . Then xn(id∗) . n−t holds for all n ≥ 1. Proof . Step 1. We need to replace the decomposition of id∗ from (6.1) by a more sophis- ticated one: id∗ = id∗ µ + J Xµ=0 L Xµ=J+1 id∗ µ + K Xµ=L+1 id∗ µ + ∞ Xµ=K+1 id∗ µ with J < L < K . Here J, L and K will be chosen later on. As in (6.2) this decomposition results in the estimate xρ n(id∗) ≤ J Xµ=0 where n − 1 =PK xρ nµ(id∗ µ)+ L K µ)+ xρ nµ(id∗ Xµ=J+1 Xµ=L+1 µ=0(nµ − 1). Cor. 6.5 yields µ k . 2−µα µ(d−1)( 1 k id∗ xρ nµ(id∗ µ)+ ∞ Xµ=K+1 kid∗ µkρ, ρ = min(1, p) , (6.28) 2 − 1 p0 )+ = 2µ(−t+ 1 p0 − 1 p ) and therefore As above we choose see (6.4). Hence ∞ Xµ=K+1 k id∗ µ kρ . 2Kρ(−t+ 1 p0 − 1 p ) . nµ := Dµ + 1, µ = 0, 1, ...., J , J Xµ=0 nµ ≍ J d−1 2J and xρ nµ(id∗ µ) = 0 , J Xµ=0 32 see (6.5) and (6.6). Inserting this into (6.28) we obtain xρ n(id∗) . L Xµ=J+1 nµ(id∗ xρ µ) + K Xµ=L+1 nµ(id∗ xρ µ) + 2Kρ(−t+ 1 p0 − 1 p ) . (6.29) Step 2. For given J we choose K large enough such that 2K(−t+ 1 p0 − 1 p ) ≤ 2−Jt J (d−1)(−t) . Furthermore, we choose L := J + (d − 1)[log J] also in dependence on J. This implies 2−Lt ≍ 2−tJ J (d−1)(−t) . Now we fix our remaining degrees of freedom by defining if J + 1 ≤ µ ≤ L , if L + 1 ≤ µ ≤ K . Here β, γ > 0 will be fixed later. Since γ > 0, applying (6.23), we have nµ := (cid:2)Dµ 2(µ−L)β+J−µ(cid:3) (cid:2)J d−12J 2(L−µ)γ(cid:3)  K Xµ=J+1 nµ ≍ J d−12J . (6.30) Substep 2.1. We estimate the first sum in (6.29). Making use of the same arguments as in proof of Lemma 6.18 we find L Xµ=J+1 nµ(id∗ xρ µ) . 2−Lρt . 2−tJρ J −(d−1)ρt . (6.31) Substep 2.2. Now we estimate the second sum in (6.29). Therefore we consider the following splitting of nµ, L + 1 ≤ µ ≤ K nµ ≍ J d−1 2J 2(L−µ)γ = J d−1 2µ 2L−µ 2−(d−1)[log J] 2(L−µ)γ = 2µ 2(L−µ)(γ+1) , where we used the definition of L. Observe nµ ≤ Dµ/2. The inequality (6.10) and property (a) in Appendix A lead to the estimate xnµ(idµ) . 2µ(−t+ 1 p0 − 1 p0,p) . 2µ(−t+ 1 p ) xn(idDµ p − 1 p0 p0 = 2−µt 2(L−µ)(γ+1)( 1 ) . − 1 p ) (2µ 2(L−µ)(γ+1)) 1 p − 1 p0 This implies K Xµ=L+1 nµ(id∗ xρ µ) . K Xµ=L+1 2−µρt 2(L−µ)(γ+1)( 1 p − 1 p0 )ρ . Choosing γ > 0 such that t > (γ + 1)(cid:16) 1 p0 − 1 p(cid:17) 33 we conclude K Xµ=L+1 xρ nµ(id∗ µ) . 2−Ltρ ≍ 2−tJρJ −(d−1)tρ . Hence, inserting the previous inequality and (6.31) into (6.29), xn(id∗) . 2−tJ J (d−1)(−t) follows. Based on this estimate and (6.30) one can finish the proof as before. (cid:4) As a corollary of Lem. 6.16 - Lem. 6.19 we obtain the main result of this subsection. Theorem 6.20. Let 0 < p0, p ≤ 2 and t >(cid:16) 1 (i) If t > 1 p0 − 1 2 , then . p(cid:17)+ p0 − 1 xn(id∗) ≍ n−t(log n)(d−1)(t+ 1 2 − 1 p0 ) holds for all n ≥ 2. (ii) If t < 1 p0 − 1 2 , then xn(id∗) ≍ n−t holds for all n ≥ 1. Remark 6.21. Again we comment on the limiting situation t = 1 ρ := min(1, p) and t = 1 p0 − 1 2 it follows p0 − 1 n−t . xn(id∗) . n−t(log log n)t+ 1 ρ , n ≥ 3. 2 . For 0 < p0, p < 2, This is the only limiting case where the gap is of order log log n to some power. For that reason we give a few more details. In principal we argue as in Lemma 6.18. For given J ∈ N, J ≥ 4, we choose L := J + (d − 1) [log J] as above. Next we define Then and J + 1 ≤ µ ≤ L . L nµ :=h 2J J d−1 log J i , Xµ=J+1 p0,2) .(cid:16) 2J J d−1 log J (cid:17) nµ ≍ 2J J d−1 xnµ(idDµ 1 2 − 1 p0 follow, see property (a) in Appendix A. Applying (6.8) we find L Xµ=J+1 xρ nµ(id∗ µ) . L p0 − 1 2µρ(−t+ 1 Xµ=J+1 2 )(cid:16) 2J J d−1 . (cid:0)2J J d−1(cid:1)−ρt (log J)1+ρt . log J (cid:17)ρ( 1 2 − 1 p0 ) As in Lemma 6.18 this proves the claim. 34 6.3.5 The case 0 < p ≤ 2 < p0 ≤ ∞ This is the last case we have to consider. Theorem 6.22. Let 0 < p ≤ 2 < p0 ≤ ∞ and t > 1 p0 . Then xn(id∗) ≍ n−t+ 1 p0 − 1 2 (log n)(d−1)(t+ 1 2 − 1 p0 ) holds for all n ≥ 2. Proof . Step 1. Estimate from below. Since p ≤ 2, from (6.13) and (6.8) we derive xn(id∗) & µ(d−1)( 1 2 − 1 p ) 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . We choose n := [Dµ/2] and obtain from property (c)(part(i)) in Appendix A that xn(idDµ p0,p) & (Dµ) 1 p − 1 2 & (µd−1 2µ) 1 p − 1 2 . This implies xn(id∗) & 2µ(−t− 1 2 + 1 p0 ) . Using 2µ ≍ n logd−1 n we conclude xn(id∗) & n−t− 1 2 + 1 p0 (log n)(d−1)(t+ 1 2 − 1 p0 ) . Step 2. Estimate from above. We consider the commutative diagram st,Ω p0,p0b id∗ s0,Ω p,2 f id1 id2 s0,Ω 2,2 f From p < 2 we derive s0,Ω 2,2 f ֒→ s0,Ω p,2 f which implies k id2 k < ∞. The ideal property of the s-numbers in combination with Thm. 6.13 yield xn(id∗) . n−t+ 1 p0 2 (log n)(d−1)(t+ 1 − 1 2 − 1 p0 ) if t > 1/2−1/p0 p0/2−1 = 1 p0 . (cid:4) Theorem 6.23. Let 0 < p ≤ 2 < p0 < ∞ and 0 < t < 1 2 (log n)(d−1)(t+ 1 p0 xn(id∗) ≍ n− tp0 . Then 2 − 1 p0 ) holds for all n ≥ 2. 35 Proof . Step 1. Estimate from below. Since p ≤ 2, from (6.13) and (6.8) we obtain xn(id∗) & µ(d−1)( 1 2 − 1 p ) 2µ(−t+ 1 p0 − 1 p ) xn(idDµ p0,p) . With n := [D 2 p0 µ ] property (c)(part(ii)) in Appendix A yields xn(idDµ p − 1 p0 p0,p) & D µ 1 & (µd−1 2µ) 1 p − 1 p0 . Hence xn(id∗) & µ(d−1)( 1 2 − 1 p0 ) 2−tµ . Since 2µ ≍ p0 2 n logd−1(n p0 2 ) we conclude xn(id∗) & n− tp0 2 (log n)(d−1)(t+ 1 2 − 1 p0 ) . Step 2. Estimate from above. Again we consider the commutative diagram st,Ω p0,p0b id∗ s0,Ω p,2 f id1 id2 s0,Ω 2,2 f In addition we know xn(id1) ≍ n− tp0 2 (log n)(d−1)(t+ 1 2 − 1 p0 ) , n ≥ 2 . if 2 < p0 < ∞ and 0 < t < 1 yields p0 , see Thm. 6.14. Now the ideal property of the s-numbers if 0 < t < 1 p0 . xn(id∗) . n− tp0 2 (log n)(d−1)(t+ 1 2 − 1 p0 ) , n ≥ 2 (cid:4) Remark 6.24. In the limiting situation t = 1 p0 > 0 we have n− 1 2 (log n) (d−1) 2 . xn(id∗) . n− 1 2 (log n) (d−1) 2 (log n) 1 2 + 1 ρ for all n ≥ 2. Here ρ = min(1, p). 7 Proofs Here we will give proofs of the assertions in Section 3. For better readability we continue to work with (p0, p) instead of (p1, p2). 36 7.1 Proof of the main Theorem 3.1 The heart of the matter is the following in principal well-known lemma. Lemma 7.1. Let 0 < p0 ≤ ∞, 0 < p < ∞ and t ∈ R. Then xn(id∗ : st,Ω p0,p0b → s0,Ω p,2 f ) ≍ xn(cid:0)id : St p0,p0B(Ω) → S0 p,2F (Ω)(cid:1) holds for all n ∈ N. Proof . Step 1. Let 0 < p0 < ∞. Let E : Bt p0,p0(R) denote a linear and continuous extension operator. For existence of those operators we refer, e.g., to [57, 3.3.4] p0,p0(0, 1) → Bt or [45]. Without loss of generality we may assume that supp Ef ⊂ [¯ν∈Nd 0, ¯m∈AΩ ¯ν supp Ψ¯ν, ¯m , see Section 5, for all f . Then the d-fold tensor product operator Ed := E ⊗ . . . ⊗ E p0,p0B((0, 1)d) = Bt p0,p0B(Rd) = Bt maps the tensor product space St p0,p0(0, 1)⊗γp0 . . . ⊗γp0 p0,p0(0, 1) into the tensor product space St Bt p0,p0(R), see [48], and is again a linear and continuous extension operator. This follows from the fact that γp0 is an uniform quasi-norm. Hence Ed ∈ L(St Step 2. Let p0 = ∞. We discussed extension operators in this case in Subsection 3.4. Now we can argue as in Step 1. p0,p0(R) ⊗γp0 . . . ⊗γp0 p0,p0B(Ω), St p0,p0B(Rd)). Bt Step 3. We follow [61] and consider the commutative diagram St p0,p0B(Ω) idy S0 p,2F (Ω) Ed−−−−→ St p0,p0B(Rd) RΩ←−−−− S0 p,2F (Rd) The mapping W is defined as W−−−−→ st,Ω p0,p0b id∗ y W ∗ ←−−−− s0,Ω p,2 f Wf := (cid:16)2¯ν1 hf, Ψ¯ν,¯ki(cid:17)¯ν∈Nd . 0, ¯k∈AΩ ¯ν Furthermore, W ∗ is defined as W ∗λ := X¯ν∈Nd 0 X¯k∈AΩ ¯ν λ¯ν,¯k Ψ¯ν,¯k and RΩ means the restriction to Ω. The boundedness of Ed,W,W ∗, RΩ and the ideal property of the s-numbers yield xn(id) . xn(id∗). A similar argument with a slightly modified diagram yields xn(id∗) . xn(id) as well. (cid:4) Next we need to recall an adapted Littlewood-Paley assertion, see Nikol'skij [31, 1.5.6]. 37 Lemma 7.2. Let 1 < p < ∞. Then S0 p,2F (Rd) = Lp(Rd) and S0 p,2F (Ω) = Lp(Ω) in the sense of equivalent norms. Proof of Thm. 3.1. Lemma 7.1 and Lemma 7.2 allow to carry over the results obtained in Section 6 to the level of function spaces. Theorem 3.1 becomes a consequence of Theorems 6.11 - 6.14, Theorem 6.20 and Theorems 6.22, 6.23. (cid:4) 7.2 Proofs of the results in Subsections 3.2 Recall that st,Ω ∞,qa or (st,Ω ∞,qa)µ has to be interpreted as st,Ω ∞,qb and (st,Ω ∞,qb)µ. Lemma 7.3. Let t, r ∈ R and 0 < p, q, p0, q0 ≤ ∞. Then xn(id1 : st,Ω p0,q0a → sr,Ω p,q a) ≍ xn(id2 : st−r,Ω p0,q0 a → s0,Ω p,q a) , n ∈ N . Proof . We consider the commutative diagram st,Ω p0,q0a Jry st−r,Ω p0,q0 a id1 −−−−→ sr,Ω p,q a id2 −−−−→ s0,Ω p,q a. J−r x Here Jr is the isomorphism defined in (5.1). Hence xn(id1) . xn(id2). But st−r,Ω p0,q0 a id2 −−−−→ s0,Ω p,q a J−ry Jr x −−−−→ sr,Ω p,q a yields xn(id2) . xn(id1) as well. The proof is complete. st,Ω p0,q0a id1 (cid:4) Proof of Theorem 3.4. Step 1. Estimate from above. Under the given restrictions there always exists some r > 1 diagram 2 such that t > r +(cid:16) 1 . We consider the commutative 2(cid:17)+ p0 − 1 St p0,p0B((0, 1)d) id1 L∞((0, 1)d) id2 id3 2,2B((0, 1)d) Sr The multiplicativity of the Weyl numbers yields x2n−1(id1) ≤ xn(id2) xn(id3) . 38 From Lemmas 7.2 and 7.3 we have xn(id2) ≍ xn(id : St−r p0,p0B((0, 1)d) → L2((0, 1)d)). (7.1) Prop. 3.2, Thm. 3.1 and (7.1) lead to x2n−1(id1) . (log n)(d−1)r nr− 1 2 (log n) (d−1)(t−r− 1 p0 + 1 2 ) nt−r (log n) (d−1)(t−r− 1 p0 + 1 2 ) t−r− 1 p0 + 1 2 n if if 0 < p0 ≤ 2 , t − r > 1 p0 − 1 2 , 2 ≤ p0 ≤ ∞ , t − r > 1 p0 .   Finally, the monotonicity of the Weyl numbers yields the claim for all n ≥ 2. Step 2. Estimate from below. The claim will follow from the next proposition. Proposition 7.4. Let t > 1 p0 . As estimates from below we get xn(id : St p0,p0B((0, 1)d) → L∞((0, 1)d)) (log n) (d−1)(t+ 1 2 − 1 p0 nt− 1 2 (log n) (d−1)(t+ 1 2 − 1 p0 t− 1 p0 n ) ) if if 0 < p0 ≤ 2 , 2 ≤ p0 ≤ ∞ , &   for all n ≥ 2. Proof . Again we shall use the multiplicativity of the Weyl numbers, but this time in connection with its relation to the 2-summing norm [33, Lemma 8]. Let us recall this notion. An operator T ∈ L(X, Y ) is said to be absolutely 2-summing if there is a constant C > 0 such that for all n ∈ N and x1, . . . , xn ∈ X the inequality x∗∈X ∗,kx∗X ∗k≤1(cid:16) Xj=1 < xj, x∗ > 2(cid:17)1/2 k T xj Y k2(cid:17)1/2 ≤ C (7.2) sup n Xj=1 (cid:0) n holds (see [32, Chapter 17]). The norm π2(T ) is given by the infimum of all C > 0 satisfying (7.2). X ∗ refers to the dual space of X. Pietsch [33] proved the inequality n1/2 xn(S) ≤ π2(S) , n ∈ N , for any linear operator S. Using this inequality with respect S = id we conclude x2n−1(id : St ≤ xn(id : St ≤ xn(id : St = xn(id : St p0,p0B((0, 1)d) → L2((0, 1)d)) p0,p0B((0, 1)d) → L∞((0, 1)d)) xn(id : L∞((0, 1)d) → L2((0, 1)d)) p0,p0B((0, 1)d) → L∞((0, 1)d)) n−1/2 π2(id : L∞((0, 1)d) → L2((0, 1)d)) p0,p0B((0, 1)d) → L∞((0, 1)d)) n−1/2 ; 39 where in the last equality we have used that π2(id : L∞((0, 1)d) −→ L2((0, 1)d)) = kid : L∞((0, 1)d) −→ L2((0, 1)d)k = 1 , see [35, Example 1.3.9]). Since 1 n 2 x2n−1(id : St p0,p0B((0, 1)d) → L2((0, 1)d)) (log n) (d−1)(t+ 1 2 − 1 p0 ) nt− 1 2 (log n) (d−1)(t+ 1 2 − 1 p0 t− 1 p0 n if if ) 0 < p0 ≤ 2 , t > 1 p0 − 1 2 , 2 ≤ p0 ≤ ∞ , t > 1 p0 , ≍   see Thm. 6.20, Thm. 6.13, this proves the claimed estimate from below. (cid:4) 7.3 Proof of the results in Subsection 3.3 As a preparation we need the following counterpart of the classical result F 0 1,2(Rd) ֒→ L1(Rd) in the dominating mixed situation. The following proof we learned from Dachun Yang and Wen Yuan [64]. Lemma 7.5. We have S0 1,2F (Rd) ֒→ L1(Rd). Proof . Let f ∈ S0 density of S(Rd) in S0 defined in (9.2). Let φ0, φ ∈ C ∞ 0 (R) be functions s.t. 1,2F (Rd). We may assume that f is a Schwartz function, due to the 1,2F (Rd). Let (ϕ¯j)¯j be the smooth dyadic decomposition of unity φ0(t) = 1 on supp ϕ0 φ(t) = 1 on supp ϕ1 . We put φj(t) := φ(2−j+1t), j ∈ N, and It follows φ¯j := φj1 ⊗ . . . ⊗ φjd , ¯j = (j1, . . . , jd) ∈ Nd 0 . ϕ¯j(x) · φ¯j(x) = 1 for all x ∈ Rd , X¯j∈Nd 0 see (9.1) and (9.2). This implies f = X¯j∈Nd 0 F −1[ϕ¯j(ξ) φ¯j (ξ)Ff (ξ)] (convergence in S′(Rd)). Let g ∈ L∞(Rd). Holder's inequality yields hf, gi ≤ X¯j∈Nd ≤ (cid:13)(cid:13)(cid:0) X¯j∈Nd 0 hF −1[ϕ¯jFf ],F −1[φ¯j Fg]i F −1[ϕ¯jFf ]2(cid:1) 0 1 2L1(Rd)(cid:13)(cid:13) (cid:13)(cid:13)(cid:0) X¯j∈Nd 0 40 F −1[φ¯jFg]2(cid:1) 1 2L∞(Rd)(cid:13)(cid:13). Next we are going to use the tensor product structure of F −1φ¯j and the fact that F −1φjl, l = 1, . . . , d, are Schwartz functions. For any M > 0, we have F −1[φ¯jFg](x) . ZRd g(y) (F −1φ¯j)(x − y) dy 2¯j1 . kgL∞(Rd)k ZRn Qd l=1(1 + 2jlxl − yl)(1+M ) Yl=1 ZR = kgL∞(Rd)k dy . (1 + 2jlxl − yl)(1+M ) 2jl d dy Some elementary calculations yield ZR 2jl (1 + 2jlxl − yl)(1+M ) dy . 2−jlM with constants independent of jl. Inserting this in our previous estimate we obtain F −1[φ¯jFg](x) . kgL∞(Rd)k 2−¯j1M . Hence (cid:13)(cid:13)(cid:0) X¯j∈Nd 0 F −1[φ¯jFg](x)2(cid:1) 1 2L∞(cid:13)(cid:13) . kgL∞(Rd)k X¯j∈Nd ∞ 0 2−¯j1M . kgL∞(Rd)k . kgL∞(Rd)k . Xµ=0 X¯j1=µ 2−¯j1M Therefore, we obtain kfL1(Rd)k = sup kgL∞(Rd)k=1 hf, gi . k(X¯j∈Nd 0 F −1[ϕ¯jFf ]2) 1 2L1(Rd)k. That completes our proof. (cid:4) Proof of Theorem 3.7. Step 1. Estimate from above. From the chain of embeddings St p0,p0B((0, 1)d) ֒→ S0 1,2F ((0, 1)d) ֒→ L1((0, 1)d), together with Lem. 7.1, Thm. 6.20, Thm. 6.22, Thm. 6.23 and the abstract properties of Weyl numbers, see Section 4, we derive the upper bound. Step 2. We prove the lower bound for the case p0 < 2 and t < 1 under the condition t > max(0, 1 p0 − 1 p0 − 1), the chain of embeddings holds true 2 . First we note that, St p0,p0B((0, 1)d) ֒→ L1((0, 1)d) ֒→ S0 1,∞B((0, 1)d). Then the ideal property of the s-numbers yields xn(id : St p0,p0B((0, 1)d) → S0 1,∞B((0, 1)d)) . xn(St p0,p0B((0, 1)d) → L1((0, 1)d)) . (7.3) 41 Next we consider the commutative diagram Bt+r p0,p0(0, 1) Exty St+r p0,p0B((0, 1)d) id1−−−−→ Br 1,∞(0, 1) Tr x id−−−−→ Sr 1,∞B((0, 1)d) Here the linear operators Ext and Tr are defined as follows. For g ∈ Bt+r p0,p0(0, 1), we put (Extg)(x1, ..., xd) = g(x1), x = (x1, ..., xd) ∈ Rd . In case of f ∈ Sr 1,∞B((0, 1)d) we define (Trf )(x1) = f (x1, 0, ..., 0) , x1 ∈ R . p0,p0(0, 1) continuously into St+r Note that the condition r > 1 guarantees that the operator Tr is well defined, see [47, Thm. 2.4.2]. Furthermore, Ext maps Bt+r p0,p0B((0, 1)d). This follows from the fact that k · St+r p0,p0B((0, 1)d)k is a cross-quasi-norm, see the formula in Rem. 9.4(i). Hence id1 = Tr ◦ id ◦ Ext and p0,p0(0, 1) → Br . xn(id : St+r xn(id1 : Bt+r 1,∞B((0, 1)d)) . 1,∞(0, 1)) (7.4) p0,p0B((0, 1)d) → Sr Making use of a lifting argument, see Lem. 7.3, we conclude that xn(id : St+r p0,p0B((0, 1)d) → Sr ≍ xn(id : St 1,∞B((0, 1)d)) p0,p0B((0, 1)d) → S0 1,∞B((0, 1)d)) . The lower bound is now obtained from (7.3), (7.4), (7.5) and xn(id1 : Bt+r 1,∞(0, 1)) ≍ n−t, p0,p0(0, 1) → Br n ∈ N , p0 − 1), see Lubitz [29] and Caetano [10]. if 0 < p0 ≤ 2 and t > max(0, 1 Step 3. We prove that xn(id : St p0,p0B((0, 1)d) → L1((0, 1)d)) & n−t(log n)(d−1)(t+ 1 2 . There always exists a pair (θ, p) such that 2 − 1 p0 if p0 ≤ 2 and t > 1 p0 − 1 (7.5) ) 0 < θ < 1 , 1 < p < 2 and kfLp((0, 1)d)k ≤ kfL1((0, 1)d)k1−θ kfL2((0, 1)d)kθ for all f ∈ L2((0, 1)d). Next we employ the interpolation property of the Weyl numbers, see Thm. 4.2, and obtain x2n−1(id : St p0,p0B((0, 1)d) → Lp((0, 1)d)) . x1−θ n p0,p0B((0, 1)d) → L1((0, 1)d)) xθ (id : St n(id : St p0,p0B((0, 1)d) → L2((0, 1)d)). 42 Note that 0 < p0 ≤ 2 and t > 1 p0 − 1 2 imply xn(id : St p0,p0B((0, 1)d) → L2((0, 1)d)) ≍ xn(id : St p0,p0B((0, 1)d) → Lp((0, 1)d)) ≍ n−t(log n)(d−1)(t+ 1 2 − 1 p0 ), see Thm. 3.1. This leads to xn(id : St p0,p0B((0, 1)d) → L1((0, 1)d)) & n−t(log n)(d−1)(t+ 1 2 − 1 p0 ). The lower bounds in the remaining cases can be proved similarly. (cid:4) Proof of Theorem 3.11. Define id∗ : st,Ω Cor. 6.5 yields 7.4 Proofs of the results in Subsection 3.4 p0,p0b → s0,Ω µ k . 2µ( 1 k id∗ −t) . p0 ∞,∞b and id∗ µ : (st,Ω p0,p0b)µ → (s0,Ω ∞,∞b)µ. (7.6) Arguing as in proof of Prop. 6.7 one can establish the following. Lemma 7.6. Let 0 < p0 ≤ ∞ and t ∈ R. Then )ℓDµ p0 → ℓDµ µ) ≍ xn(id∗∗ µ : 2µ(t− 1 xn(id∗ p0 ∞ ) ≍ 2µ(−t+ 1 p0 ) xn(idDµ p0,∞) for all n ∈ N. Property (a) in Appendix A yields xn(idDµ p0,∞) ≍( 1 n 1 2 − 1 p0 if if 2 ≤ p0 ≤ ∞ , 0 < p0 ≤ 2 , if 2n ≤ Dµ. Now we may follow the proof of Thm. 6.11. This results in the following useful statement. Theorem 7.7. (i) Let 0 < p0 ≤ 2 and t > 1 ∞,∞b) ≍ n−t+ 1 p0,p0b → s0,Ω xn(id∗ : st,Ω p0 . Then 2 (log n)(d−1)(t− 1 p0 ) , n ≥ 2 . (ii) Let 2 ≤ p0 ≤ ∞ and t > 1 p0 . Then xn(id∗ : st,Ω p0,p0b → s0,Ω ∞,∞b) ≍ n−t+ 1 p0 (log n)(d−1)(t− 1 p0 ) , n ≥ 2 . By making use of a lifting argument, see Lemma 7.3, and the counterpart of Lemma 7.1 for this situation, i.e., xn(id∗ : st,Ω p0,p0b → s0,Ω we immediately get the following corollary. ∞,∞b) ≍ xn(cid:0)id : St p0,p0B(Ω) → S0 ∞,∞B(Ω)(cid:1) , n ∈ N , 43 Corollary 7.8. Let s ∈ R. (i) Let 0 < p0 ≤ 2 and t > s + 1 xn(id : St p0,p0B((0, 1)d) → Ss p0 p0 (ii) Let 2 ≤ p0 ≤ ∞ and t > s + 1 xn(id : St p0,p0B((0, 1)d) → Ss Now Thm. 3.11 follows from Z s . Then . Then ∞,∞B((0, 1)d)) ≍ n−t+s+ 1 2 (log n)(d−1)(t−s− 1 p0 ) , n ≥ 2 . ∞,∞B((0, 1)d)) ≍ n−t+s+ 1 mix((0, 1)d) = Ss p0 (log n)(d−1)(t−s− 1 p0 ) , ∞,∞B((0, 1)d), see Lemma 3.10. n ≥ 2 . (cid:4) Proof of Theorem 3.13. The lower estimate in the case of high smoothness is a direct consequence of xn ≤ an and Theorem 3.11. Step 1. We prove the upper bound of an(id : St p0 > 1. First, recall p0,p0B((0, 1)d) → S0 ∞,∞B((0, 1)d)) in case an(idDµ p0,∞) ≍  1 1− 1 p0 min(1, D µ n− 1 2 ) if if 2 ≤ p0 ≤ ∞ , 1 < p0 < 2 , if 2n ≤ Dµ, see [23, 62]. To avoid nasty calculations by checking this behaviour for p0 ≥ 2 one may use the elementary chain of inequalities xn(idDµ p0,∞) ≤ an(idDµ p0,∞) . 1 in combination with property (a) in Appendix A. Let 2 ≤ p0 ≤ ∞. Because of an(idDµ p0,∞) ≍ xn(idDµ p0,∞) ≍ 1 if 2n ≤ Dµ we may argue as in case of Weyl numbers, see the proof of Thm. 3.11 given above. Now we consider the case 1 < p0 < 2 and t > 1. We define id∗ : st,Ω p0,p0b → s0,Ω ∞,∞b and id∗ µ : (st,Ω p0,p0b)µ → (s0,Ω ∞,∞b)µ . (7.6) and Lemma 7.6 yield and k id∗ µ k . 2µ( 1 p0 −t) an(id∗ µ : 2µ(t− 1 µ) ≍ an(id∗∗ ∞ ) ≍ 2µ(−t+ 1 for all n ∈ N. Now we get as in Subsection 5.2, formula (6.7), µ) + 2L(−t+ 1 )ℓDµ p0 → ℓDµ an(id∗) . anµ(id∗ p0 p0 L ) , p0 Xµ=J+1 since ρ = 1 here. For we define 1 < λ < 1 2 + t 2 nµ := Dµ 2(J−µ)λ , J + 1 ≤ µ ≤ L . 44 ) an(idDµ p0,∞) (7.7) (7.8) (7.9) Then, as above, nµ ≤ Dµ 2 and L Xµ=J+1 nµ ≍ J d−12J follows. From (7.7) and an(idDµ 1− 1 p0 µ n− 1 2 ) we conclude anµ(id∗ p0,∞) ≍ min(1, D µ) . 2µ(−t+ 1 . 2µ(−t+ 1 . 2µ(−t+ 1 p0 p0 1− 1 )D p0 µ 2 )µ(d−1)( 1 ) anµ(idDµ p0,∞) (Dµ.2(J−µ)λ)− 1 2 − 1 2 (J−µ)λ . p0 )2− 1 2 This leads to L Xµ=J+1 anµ(id∗ µ) . 2µ(−t+ 1 2 )µ(d−1)( 1 2 − 1 p0 )2− 1 2 (J−µ)λ L Xµ=J+1 . 2J(−t+ 1 2 )J (d−1)( 1 2 − 1 p0 ), since λ satisfies λ < 1 way. Now we choose L large enough such that 2 + t 2 , see (7.9), guaranteeing the convergence of the series in that 2L(−t+ 1 p0 ) In view of (7.8) this yields . 2J(−t+ 1 2 )J (d−1)( 1 2 − 1 p0 ). an(id∗) . 2J(−t+ 1 2 )J (d−1)( 1 2 − 1 p0 ). This proves the estimate from above. Step 2. Let p0 = 1. Then we use an(idDµ 1,∞) . n− 1 2 if 2n ≤ Dµ, see [62]. This is just the limiting case of Step 1. So we argue as there. Step 3. It remains to consider the following case: 1 < p0 < 2, s = 0 and 1 < t < 1. p0 Substep 3.1. Estimate from above. In this case we define and p′ 0 2 + (d − 1)p′ L :=hJ nµ :=(cid:2)Dµ 2(µ−L)β+J−µ(cid:3) , 0(cid:16) 1 p0 − 1 2(cid:17) log Ji (7.10) J + 1 ≤ µ ≤ L , for some β > 0. Here p′ 0 is the conjugate of p0, i.e., 1 p0 + 1 p′ 0 = 1. Again we have nµ ≤ Dµ 2 and L Xµ=J+1 nµ ≍ J d−12J . 45 From (7.7), (7.8) and an(idDµ 1− 1 p0 µ n− 1 2 ) we get p0,∞) ≍ min(1, D 1− 1 )D p0 µ 2µ(−t+ 1 p0 2(cid:1)2 2µ(cid:0)−t+1− β an(id∗) . . L Xµ=J+1 Xµ=J+1 L 2 + 2L(−t+ 1 p0 ) (cid:2)Dµ 2(µ−L)β+J−µ(cid:3)− 1 2 µ(d−1)(cid:0) 1 p0(cid:1) + 2L(−t+ 1 2 − 1 p0 Lβ−J ) . The condition t < 1 guarantees that we can choose β > 0 such that −t + 1 − β we have 2 > 0. Then Lβ−J an(id∗) . 2L(cid:0)−t+1− β 2(cid:1)2 2 J (d−1)(cid:0) 1 = 2L(−t+1)2− J 2 J (d−1)(cid:0) 1 2 − 1 2 − 1 p0(cid:1) + 2L(−t+ 1 p0 ) p0(cid:1) + 2L(−t+ 1 p0 ) . Now, replacing L by the value in (7.10), a simple calculation yields an(id∗) . 2J p′ 0 2 (cid:0)−t+ 1 p0(cid:1)J (d−1)(cid:2) p′ 0 2 (cid:0)−t+ 1 p0(cid:1)+t− 1 p0(cid:3) . Rewriting this in dependence on n we obtain an(id∗) . n− p′ 2 (t− 1 0 p0 )(log n)(d−1)(t− 1 p0 ). This proves the estimate from above. Substep 3.2. Estimate from below. First of all, notice that we can prove an(id∗) ≥ an(id∗ µ) as in (6.13). We choose n = [D 2 p′ 0 µ ]. By employing again (7.7) and an(idDµ p0,∞) ≍ min(1, D 1− 1 p0 µ n− 1 2 ) we obtain the desired estimate. Finally, by making use of a lifting argument, see Lemma 7.3, and the counterpart of Lemma 7.1 we finish our proof. (cid:4) 7.5 Proof of interpolation properties of Weyl numbers For the basics in interpolation theory we refer to the monographs [7, 30, 56]. To begin with we deal with Gelfand numbers. The n-th Gelfand number is defined as cn(T ) = inf Mn sup kxXk≤1,x∈Mn kT xY k (7.11) where Mn is a subspace of X such that codimMn < n, see also Section 4. Next we recall the interpolation properties of Gelfand numbers, for the case of Banach spaces we refer to Triebel [55]. Theorem 7.9. Let 0 < θ < 1. Let X, Y, X0, Y0 be quasi-Banach spaces. Further we assume Y0 ∩ Y1 ֒→ Y and the existence of a positive constant C with kyY k ≤ C kyY0k1−θ kyY1kθ for all y ∈ Y0 ∩ Y1. (7.12) 46 Then, if it follows T ∈ L(X, Y0) ∩ L(X, Y1) ∩ L(X, Y ) n cn+m−1(T : X → Y ) ≤ C c1−θ (T : X → Y0) cθ for all n, m ∈ N. Here C is the same constant as in (7.12). Proof . We follow the proof in [55]. Let Ln and Lm be subspaces of X such that codimLn < n and codimLm < m respectively. Then codim(Ln ∩ Lm) < m + n − 1. Furthermore, by assumption, for all x ∈ X we have T x ∈ Y0 ∩ Y1. From (7.11) and (7.12) we derive m(T : X → Y1) cm+n−1(T : X → Y ) = inf Ln,Lm sup kxXk≤1 x∈Ln∩Lm kT xY k ≤ C inf Ln,Lm sup kxXk≤1 x∈Ln∩Lm sup kxXk≤1 x∈Ln ≤ C(cid:0) inf Ln = C c1−θ n k T xY0k1−θ k T xY1kθ k T xY0k(cid:1)1−θ(cid:0) inf kxXk≤1 x∈Lm sup Lm k T xY1k(cid:1)θ (T : X → Y0) cθ m(T : X → Y1) . The proof is complete. (cid:4) Remark 7.10. Triebel [55] worked with Gelfand widths. For compact operators Gelfand widths and Gelfand numbers coincide, see also [55]. Hence, if we require T ∈ K(X, Y0) ∩ K(X, Y1) ∩ L(X, Y ) , where K(X, Y ) stands for the subspace of L(X, Y ) formed by the compact operators, then Theorem 7.9 remains true for Gelfand widths. Without extra conditions on T Gelfand widths and Gelfand numbers may not coincide, see Edmunds and Lang [17] for a discussion of this question. Now we ready prove the Thm. 4.2. Proof of Theorem 4.2. Let A ∈ L(ℓ2, X) such that kAk ≤ 1. Then from Thm. 7.9 we conclude cn+m−1(T A : ℓ2 → Y ) ≤ C c1−θ n (T A : ℓ2 → Y0) cθ m(T A : ℓ2 → Y1). Employing Remark 4.1(ii) we obtain cn+m−1(T A : ℓ2 → Y ) ≤ C x1−θ n (T : X → Y0) xθ m(T : X → Y1). Now taking the supremum with respect to A we find xn+m−1(T : X → Y ) ≤ C x1−θ n (T : X → Y0) xθ m(T : X → Y1). The proof is complete. (cid:4) 47 8 Appendix A - Weyl numbers of the embeddings ℓm p0 → ℓm p The Weyl numbers of id : ℓm p have been investigated at various places, we refer to Lubitz [29], Konig [27], Caetano [8, 9] and Zhang, Fang, Huang [65]. We shall need the p0 → ℓm following. (a) ([29, Korollar 2.2] and [65]) Let n, m ∈ N and 2n ≤ m. Then we have xn(idm p0,p) ≍ 1 n n m 1 1 p − 1 p0 2 − 1 p0 p − 1 p0 1 if if if if 2 ≤ p0 ≤ p ≤ ∞ , 0 < p0 ≤ p ≤ 2 , 0 < p0 ≤ 2 ≤ p ≤ ∞ , 0 < p < p0 ≤ 2 .   (b) ([29, Korollare 2.6, 2.8, Satz 2.9]) Let 2 ≤ p < p0 ≤ ∞ and n, m, k ∈ N, k ≥ 2. Then we have (i) xn(idm (ii) xn(idm (iii) xn(idkn p0,p) & m p0,p) ≍ 1. 1 r p0,p) .(cid:18) m n(cid:19) p0 if 1 ≤ n ≤ [m if n ≤ m, p − 1 1 r 1 = 2 p0 ], 1/p − 1/p0 1 − 2/p0 , (c) ([65]) Let 0 < p ≤ 2 < p0 ≤ ∞ and n, m ∈ N. Then (i) xn(idm p0,p) & m (ii) xn(idm p0,p) & m 1 p − 1 1 p − 1 2 if n ≤ m 2 , p0 if n ≤ m 2 p0 . 9 Appendix B - Function spaces of dominating mixed smoothness 9.1 Besov and Lizorkin-Triebel spaces on R Here we recall the definition and a few properties of Besov and Sobolev spaces defined on R. We shall use the Fourier analytic approach, see e.g. [57]. Let ϕ ∈ C ∞ 0 (R) be a function such that ϕ(t) = 1 in an open set containing the origin. Then by means of ϕ0(t) = ϕ(t) , ϕj(t) = ϕ(2−j t) − ϕ(2−j+1t) , t ∈ R , j ∈ N , (9.1) we get a smooth dyadic decomposition of unity, i.e., ∞ Xj=0 ϕj(t) = 1 for all t ∈ R , and supp ϕj is contained in the dyadic annulus {t ∈ R : b < ∞ independent of j ∈ N. a 2j ≤ t ≤ b 2j} with 0 < a < 48 Definition 9.1. Let 0 < p, q ≤ ∞ and s ∈ R. (i) The Besov space Bs such that k f Bs p,q(R) is then the collection of all tempered distributions f ∈ S ′(R) p,q(R)k :=(cid:16) 2jsq kF −1[ϕjFf ](· )Lp(R)kq(cid:17)1/q ∞ Xj=0 p,q(R) is then the collection of all tempered is finite. (ii) Let p < ∞. The Lizorkin-Triebel space F s distributions f ∈ S ′(R) such that k f F s p,q(R)k :=(cid:13)(cid:13)(cid:13)(cid:16) ∞ Xj=0 2jsq F −1[ϕjFf ](· )q(cid:17)1/q(cid:12)(cid:12)(cid:12) Lp(R)(cid:13)(cid:13)(cid:13) is finite. Remark 9.2. (i) There is an extensive literature about Besov and Lizorkin-Triebel spaces, we refer to the monographs [31], [57], [58] and [59]. These quasi-Banach spaces Bs p,q(R) and F s p,q(R) can be characterized in various ways, e.g. by differences and derivatives, whenever s is sufficiently large, i.e., s > max(0, 1/p − 1) in case of Besov spaces and s > max(0, 1/p− 1, 1/q − 1) in case of Lizorkin-Triebel spaces. We refer to [57] for details. (ii) The spaces Bs p,q(R) do not coincide as sets except the case p = q. p,q(R) and F s 9.2 Besov and Lizorkin-Triebel spaces of dominating mixed smoothness Detailed treatments of Besov and Lizorkin-Triebel spaces of dominating mixed smoothness are given at various places, we refer to the monographs [1, 47], the survey [46] as well as to the booklet [61]. If ϕj, j ∈ N0, is a smooth dyadic decomposition of unity as in (9.1), then by means of ϕ¯j := ϕj1 ⊗ . . . ⊗ ϕjd , ¯j = (j1, . . . , jd) ∈ Nd 0 , (9.2) we obtain a smooth decomposition of unity on Rd. Definition 9.3. Let 0 < p, q ≤ ∞ and t ∈ R. (i) The Besov space of dominating mixed smoothness St tempered distributions f ∈ S ′(Rd) such that p,qB(Rd) is the collection of all k f St p,qB(Rd)k :=(cid:16) X¯j∈Nd 0 2¯j1tq kF −1[ϕ¯j Ff ](· )Lp(Rd)kq(cid:17)1/q is finite. (ii) Let 0 < p < ∞. The Lizorkin-Triebel space of dominating mixed smoothness St is the collection of all tempered distributions f ∈ S ′(Rd) such that p,qF (Rd) k f St p,qF (Rd)k :=(cid:13)(cid:13)(cid:13)(cid:16) X¯j∈Nd 0 is finite. 2¯j1tq F −1[ϕ¯j Ff ](· )q(cid:17)1/q(cid:12)(cid:12)(cid:12) Lp(Rd)(cid:13)(cid:13)(cid:13) 49 Remark 9.4. (i) The most interesting property of these classes for us consists in the following: if f (x) = d Yj=1 fj(xj) , x = (x1, . . . , xd) , fj ∈ At p,q(R) , j = 1, . . . , d , then f ∈ St p,qA(Rd) and k f St p,qA(Rd)k = d Yj=1 k fj As p,q(R)k , A ∈ {B, F} . I.e., Lizorkin-Triebel and Besov spaces of dominating mixed smoothness have a cross- quasi-norm. (ii) These classes St p,qF (Rd) are quasi-Banach spaces. If either t > max(0, (1/p) − 1) (B-case) or t > max(0, 1/p − 1, 1/q − 1) (F-case), then they can be characterized by differences, we refer to [47] and [60] for details. (iii) Again the spaces St p = q. p,qF (Rd) do not coincide as sets except the case p,qB(Rd) as well as St p,qB(Rd) and St (iv) For d = 1 we have St p,qA(R) = At p,q(R) , A ∈ {B, F} . Acknowledgement: The authors would like to thank A. Hinrichs for a hint concern- ing a misprint in the phd-thesis of Lubitz [29], V.N. Temlyakov for a hint concerning a misprint in his paper [51], T. Kuhn for an explanation how to use (3.3) and Dachun Yang and Wen Yuan for a nice new proof of the continuous embedding F 0 1,2(Rd) ֒→ L1(Rd). References [1] T.I. Amanov, Spaces of differentiable functions with dominating mixed derivatives. Nauka Kaz. SSR, Alma-Ata, 1976. [2] D.B. Bazarkhanov, Characterizations of Nikol'skij-Besov and Lizorkin-Triebel function spaces of mixed smoothness. Proc. Steklov Inst. 243 (2003), 46-58. [3] D.B. Bazarkhanov, Equivalent (quasi)normings of some function spaces of gen- eralized mixed smoothness. Proc. Steklov Inst. 248 (2005), 21-34. [4] D.B. Bazarkhanov, Wavelet representations and equivalent normings of some function spaces of generalized mixed smoothness. Math. Zh. 5 (2005), 12-16. [5] D.B. Bazarkhanov, Estimates for widths of classes of periodic multivariable functions. Doklady Academii Nauk 436(5) (2011), 583 - 585 (russian), engl. transl. in Doklady Math. 83(1) (2011), 90-92. 50 [6] E.S. Belinsky, Estimates of entropy numbers and Gaussian measures for classes of functions with bounded mixed derivative. JAT 93 (1998), 114-127. [7] J. Bergh and J. Lofstrom, Interpolation Spaces. An Introduction. Springer, New York, 1976. [8] A.M Caetano, Weyl numbers in function spaces. Forum Math. 2(2) (1990), 249- 263. [9] A.M. Caetano, Weyl numbers in function spaces. II. Forum Math. 3(6) (1991), 613-621. [10] A.M. Caetano, Asymptotic distribution of Weyl numbers and eigenvalues. Phd- thesis, University of Sussex, Brighton, 1991. [11] B. Carl and A. Hinrichs, Optimal Weyl-type inequalities for operators in Ba- nach spaces. Positivity 11 (2007), 41-55. [12] B. Carl and I. Stephani, Entropy, compactness and the approximation of oper- ators. Cambridge Univ. Press, Cambridge, 1990. [13] B. Carl and H. Triebel, Inequalities between eigenvalues, entropy numbers and related inequalities of compact operators in Banach spaces. Math. Ann. 251 (1980), 129-133. [14] F. Cobos, T. Kuhn and W. Sickel, Optimal approximation of Sobolev functions in the sup-norm. Preprint, Madrid, Leipzig, Jena 2014. [15] A. Defant and K. Floret, Tensor norms and operator ideals. North Holland, Amsterdam, 1993. [16] Dinh Dung, Non-linear approximations using sets of finite cardinality or finite pseudo-dimension. J. Complexity 17(2) (2001), 467-492. [17] D.E. Edmunds and J. Lang, Gelfand numbers and widths. JAT 166 (2013), 78-84. [18] D.E. Edmunds and H. Triebel, Function spaces, entropy numbers, differential operators. Cambridge Univ. Press, Cambridge, 1996. [19] E.M. Galeev, Approximation of classes of periodic functions of several variables by nuclear operators. Math. Notes 47 (1990), 248-254. [20] E.M. Galeev, Linear widths of Holder-Nikol'skii classes of periodic functions of several variables. Mat. Zametki 59 No. 2 (1996), 189-199 (russian), engl. transl. in Math. Notes 59 (1996), No. 2, 133-140. 51 [21] E.M. Galeev, Widths of the Besov classes Br p,θ(Td). Math. Notes 69, No. 5, (2001), 605-613. [22] A. Gasiorowska and L. Skrzypczak, Some s-numbers of embeddings of func- tion spaces with weights of logarithmic type. Math. Nachr. 286(7) (2013), 644-658. [23] E.D. Gluskin, Norms of random matrices and widths of finite-dimensional sets. Math. USSR Sb. 48 (1984) 173-182. [24] M. Hansen, Nonlinear approximation and function spaces of dominating mixed smoothness. Phd thesis, Friedrich-Schiller-University Jena, 2010. [25] M. Hansen, On tensor products of quasi-Banach spaces. Preprint 63, DFG-SPP 1324, Marburg, 2010. [26] M. Hansen and J. Vybiral, The Jawerth-Franke embedding of spaces of domi- nating mixed smoothness. Georg. J. Math. 16(4) (2009), 667-682. [27] H. Konig, Eigenvalue distribution of compact operators. Birkhauser, Basel, 1986. [28] W.A. Light and E.W. Cheney, Approximation theory in tensor product spaces. Lecture Notes in Math. 1169, Springer, Berlin, 1985. [29] C. Lubitz, Weylzahlen von Diagonaloperatoren und Sobolev-Einbettungen. Bon- ner Math. Schriften 144, phd-thesis, Bonn 1982. [30] A. Lunardi, Interpolation theory. Lect. Notes, Scuola Normale Superiore Pisa, 2009. [31] S. M. Nikol'skij, Approximation of functions of several variables and imbedding theorems. Springer, Berlin, 1975. [32] A. Pietsch, Operator Ideals. North-Holland, Amsterdam, 1980. [33] A. Pietsch, Weyl numbers and eigenvalues of operators in Banach spaces. Math. Ann. 247 (1980), 149-168. [34] A. Pietsch, Eigenvalues of integral operators. I. Math. Ann. 247 (1980), 169-178. [35] A. Pietsch, Eigenvalues and s-numbers. Cambridge University Press, Cambridge, 1987. [36] A.S. Romanyuk, Approximation of the Besov classes of periodic functions of several variables in a space Lq. Ukrainian Math. J. 43(10) (1991), 1297-1306. [37] A.S. Romanyuk, The best trigonometric approximations and the Kolmogorov diameters of the Besov classes of functions of many variables. Ukrainian Math. J. 45 (1993), 724-738. 52 [38] A.S. Romanyuk, On Kolmogorov widths of classes Br p,θ of periodic function of many variables with low smoothness in the space Lq. Ukrainian Math. J. 46 (1994), 915-926. [39] A.S. Romanyuk, On the best approximations and Kolmogorov widths of the Besov classes of periodic functions of many variables. Ukrainian Math. J. 47 (1995), 91- 106. [40] A.S. Romanyuk, Linear widths of the Besov classes of periodic functions of many variables. I. Ukrainian Math. J. 53 (2001), 647-661. [41] A.S. Romanyuk, Linear widths of the Besov classes of periodic functions of many variables. II. Ukrainian Math. J. 53 (2001), 820-829. [42] A.S. Romanyuk, On estimates of the Kolmogorov widths of the classes Br p,θ in the space Lq. Ukrainian Math. J. 53 (2001), 996-1001. [43] A.S. Romanyuk, Kolmogorov widths of the Besov classes Br p,θ in the metric of the space L∞. Ukr. Mat. Visn. 2(2) (2005), 201-218. [44] A.S. Romanyuk, Best approximations and widths of classes of periodic functions of many variables. Math. Sbornik 199 (2008), 253-275. [45] V. S. Rychkov, On restrictions and extensions of the Besov and Triebel-Lizorkin spaces with respect to Lipschitz domains. J. London Math. Soc. 60 (1999), 237-257. [46] H.-J. Schmeisser, Recent developments in the theory of function spaces with dom- inating mixed smoothness. In: Proc. Conf. NAFSA-8, Prague 2006, (ed. J. Rakos- nik), Inst. of Math. Acad. Sci., Czech Republic, Prague, 2007, pp. 145-204. [47] H.-J. Schmeisser, H. Triebel, Topics in Fourier analysis and function spaces. Geest & Portig, Leipzig, 1987 and Wiley, Chichester, 1987. [48] W. Sickel and T. Ullrich, Tensor products of Sobolev-Besov spaces and appli- cations to approximation from the hyperbolic cross. JAT 161 (2009), 748-786. [49] W. Sickel and T. Ullrich, Spline interpolation on sparse grids. Applicable Anal- ysis 90 (2011), 337-383. [50] F. Sprengel, A tool for approximation in bivariate periodic Sobolev spaces. In: Approximation Theory IX, Vol. 2, Vanderbilt Univ. Press, Nashville (1999), 319- 326. [51] V.N. Temlyakov, The estimates of asymptotic characteristics on functional classes with bounded mixed derivative or difference. Trudy Mat. Inst. Steklov. 189 (1989), 138-167. 53 [52] V.N. Temlyakov, Approximation of periodic functions. Nova Science, New York, 1993. [53] V.N. Temlyakov, On approximate recovery of functions with bounded mixed derivative. J. Complexity 9 (1993), 41 -- 59. [54] V.N. Temlyakov, An inequality for trigonometric polynomials and its application for estimating the Kolmogorov widths. East J. on Approximations 2 (1996), 253 -- 262. [55] H. Triebel, Interpolationseigenschaften von Entropie und Durchmesseridealen kompakter Operatoren. Studia Math. 34 (1970), 89-107. [56] H. Triebel, Interpolation Theory, Function Spaces, Differential Operators. North- Holland Publishing Co., Amsterdam-New York, 1978. [57] H. Triebel, Theory of function spaces. Birkhauser, Basel, 1983. [58] H. Triebel, Theory of function spaces II. Birkhauser, Basel, 1992. [59] H. Triebel, Theory of function spaces III. Birkhauser, Basel, 2006. [60] T. Ullrich, Function spaces with dominating mixed smoothness. Charac- terizations by differences. Jenaer Schriften zur Mathematik und Informatik Math/Inf/05/06, Jena, 2006. [61] J. Vybiral, Function spaces with dominating mixed smoothness. Dissertationes Math. 436 (2006). [62] J. Vybiral, Widths of embeddings in function spaces. J. Complexity 24 (2008), 545-570. [63] P. Wojtaszczyk, A mathematical introduction to wavelets. Cambridge Univ. Press, Cambridge, 1997. [64] D. Yang and W. Yuan, Personal communication. Summer 2014. [65] S. Zhang, G. Fang and F. Huang, Some s-numbers of embeddings in function spaces with polynomial weights. J. Complexity 30(4) (2013), 514-532. 54
1506.01503
1
1506
2015-06-04T08:15:32
Hadamard triples generate self-affine spectral measures
[ "math.FA" ]
Let $R$ be an expanding matrix with integer entries and let $B,L$ be finite integer digit sets so that $(R,B,L)$ form a Hadamard triple on ${\br}^d$. We prove that the associated self-affine measure $\mu = \mu(R,B)$ is a spectral measure, which means it admits an orthonormal bases of exponential functions in $L^2(\mu)$. This settles a long-standing conjecture proposed by Jorgensen and Pedersen and studied by many other authors.
math.FA
math
HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI Abstract. Let R be an expanding matrix with integer entries and let B, L be finite integer digit sets so that (R, B, L) form a Hadamard triple on Rd. We prove that the associated self-affine mea- sure µ = µ(R, B) is a spectral measure, which means it admits an orthonormal bases of exponential functions in L2(µ). This settles a long-standing conjecture proposed by Jorgensen and Pedersen and studied by many other authors. Contents 1. Introduction 2. Preliminaries 3. The quasi-product form 4. Proof of the theorem References 1 3 7 12 16 1. Introduction In 1974, Fuglede [Fug74] was studying a question of Segal on the existence of commuting exten- sions of the partial differential operators on domains of Rd. Fuglede proved that the domains Ω for which such extensions exist are exactly those with the property that there exists an orthogonal basis for L2(Ω), with Lebesgue measure, formed with exponential functions {e2πihλ , xi : λ ∈ Λ} where Λ is some discrete subset of Rd. Such sets were later called spectral sets and Λ was called a spectrum for Ω. In the same paper, Fuglede proposed his famous conjecture that claims that the spectral sets are exactly those that tile Rd by some translations. The conjecture was later proved to be false in dimension 5 or higher, by Tao [Tao04] and then in dimension 3 or higher [Mat05, KM06b, KM06a, FMM06]. At this moment, the conjecture is still open in dimensions 1 and 2. In 1998, while working on the Fuglede conjecture, Jorgensen and Pedersen [JP98] asked a related question: what are the measures for which there exist orthogonal bases of exponential functions? Let µ be a compactly supported Borel probability measure on Rd and let h·,·i denote the standard inner product on Rd. The measure µ is called a spectral measure if there exists a countable set Λ ⊂ Rd, called spectrum of the measure µ, such that the collection of exponential functions E(Λ) := 2010 Mathematics Subject Classification. Primary 42B05, 42A85, 28A25. Key words and phrases. Hadamard triples, quasi-product form, self-affine sets, spectral measure. 1 2 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI {e2πihλ,xi : λ ∈ Λ} forms an orthonormal basis for L2(µ). We define the Fourier transform of µ to be bµ(ξ) =Z e−2πihξ,xidµ(x). In [JP98], Jorgensen and Pedersen made a surprising discovery: they constructed the first exam- ple of a singular, non-atomic spectral measure. The measure is the Hausdorff measure associated to a Cantor set, where the scaling factor is 4 and the digits are 0 and 2. They also proved that the usual Middle Third Cantor measure is non-spectral. Also, Strichartz proved in [Str06] that the Fourier se- ries associated to such spectral fractal measures can have much better convergence properties than their classical counterparts on the unit interval: Fourier series of continuous functions converge uniformly, Fourier series of Lp-functions converge in the Lp-norm. Since Jorgensen and Pedersen's discovery, many other examples of singular measures have been constructed, and various classes of fractal measures have been analyzed [JP98, LW02, Str98, Str00, DJ06, Li14, YL15, Li15, DS15, and references therein]. All these constructions have used the central idea of Hadamard matrices and Hadamard triples to construct the spectral singular measures by infinitely many iterations. It has been conjectured since Jorgensen and Pedersen's discovery that all Hadamard triples will generate spectral self-affine measures. Let us recall all the necessary definitions below: Definition 1.1. Let R ∈ Md(Z) be an d × d expansive matrix (all eigenvalues have modulus strictly greater than 1) with integer entries. Let B, L ⊂ Zd be finite sets of integer vectors with N := #B = #L (# denotes the cardinality). We say that the system (R, B, L) forms a Hadamard triple (or (R−1B, L) forms a compatible pair in [ LW02] ) if the matrix (1.1) is unitary, i.e., H ∗H = I. H = 1 √N he2πihR−1b,ℓiiℓ∈L,b∈B Definition 1.2. For a given expansive d × d integer matrix R and a finite set of integer vectors B with #B =: N , we define the affine iterated function system (IFS) τb(x) = R−1(x + b), x ∈ Rd, b ∈ B. The self-affine measure (with equal weights) is the unique probability measure µ = µ(R, B) satisfying (1.2) µ(E) =Xb∈B 1 N µ(τ −1 b (E)), for all Borel subsets E of Rd. This measure is supported on the attractor T (R, B) which is the unique compact set that satisfies T (R, B) = [b∈B τb(T (R, B)). The set T (R, B) is also called the self-affine set associated with the IFS. One can refer to [Hut81] and [Fal97] for a detailed exposition of the theory of iterated function systems. We say that µ = µ(R, B) satisfies the no overlap condition if We say that B is a simple digit set for R if distinct elements of B are not congruent (mod R(Zd)). µ(τb(T (R, B)) ∩ τb′(T (R, B))) = 0, ∀b 6= b′ ∈ B. HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES 3 It is well known that if (R, B, L) forms a Hadamard triple, then B must be a simple digit set for R and L must be a simple digit set for RT . Furthermore, we need only consider equal-weight measures since, in the case when the weights are not equal, the self-affine measures cannot admit any spectrum, by the no overlap condition (see Theorem 1.4 below) and [DL14, Theorem 1.5]. In this paper we prove that the conjecture proposed by Jorgensen and Pedersen is valid: Theorem 1.3. Let (R, B, L) be a Hadamard triple. Then the self-affine measure µ(R, B) is spectral. In dimension 1, Theorem 1.3 was first proved by Laba and Wang [ LW02] and refined in [DJ06]. The situation becomes more complicated when d > 1. Dutkay and Jorgensen showed that the conjecture is true if (R, B, L) satisfies a technical condition called reducibility condition [DJ07]. There are some other additional assumptions proposed by Strichartz guaranteeing Theorem 1.3 is true [Str98, Str00]. Some low-dimensional special cases were also considered by Li [Li14, Li15]. In [DL15], we introduced the following set (1.3) and proved Z = {ξ :bµ(ξ + k) = 0, for all k ∈ Zd} Theorem 1.4. [DL15, Theorem 1.7 and 1.8] Let (R, B, L) be a Hadamard triple and µ = µ(R, B) be the associated equal weight self-affine measure. Then we have: (i) µ has the no-overlap condition. (ii) Suppose furthermore that, Z = ∅, then µ is a spectral measure with a spectrum in Zd. The complete resolution of Theorem 1.3 points towards the case Z 6= ∅. It was found that there exist spectral self-affine measures with Z 6= ∅ [DL15, Example 5.4]. To prove Theorem 1.3, our strategy is to first show that in the case when Z 6= ∅ the digit set B will be reduced a quasi product- form structure (Section 3). This requires an analysis of Z as an invariant set of some dynamical system, and we use the techniques in [CCR96] (Section 2). Our methods are also similar to the ones used in [LW97]. However, as B is not a complete set of representatives (mod R(Zd)) (as it is in [LW97]), several additional adjustments will be needed. From the quasi-product form structure obtained, we construct the spectrum directly by induction on the dimension d (Section 4). 2. Preliminaries We first discuss some preliminary reduction that we can perform in order to prove our main theorem. Definition 2.1. Let R1, R2 be d × d integer matrices, and the finite sets B1, B2, L1, L2 be in Zd. We say that two triples (R1, B1, L1) and (R2, B2, L2) are conjugate (through the matrix M ) if there exists an integer matrix M such that R2 = M R1M −1, B2 = M B1 and L2 = (M T )−1L1. Proposition 2.2. Suppose that (R1, B1, L1) and (R2, B2, L2) are two conjugate triples, through the matrix M . Then (i) If (R1, B1, L1) is a Hadamard triple then so is (R2, B2, L2). (ii)The measure µ(R1, B1) is spectral with spectrum Λ if and only if µ(R2, B2) is spectral with spectrum (M T )−1Λ. 4 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI Proof. The proof follows from some simple computations, see e.g. [DJ07, Proposition 3.4]. (cid:3) Fora given integral expanding matrix R and a simple digit set B for R. We let Bn := B + RB + R2B + ... + Rn−1B = n−1Xj=0 Rjbj : bj ∈ B . We define Z[R, B] to be the smallest R-invariant lattice containing all Bn (invariant means R(Z[R, B])⊂ Z[R, B]). By Proposition 2.3 below, to prove Theorem 1.3, there is no loss of generality if we assume that Z[R, B] = Zd. Proposition 2.3. If the lattice Z[R, B] is not full-rank, then the dimension can be reduced; more precisely, there exists 1 ≤ r < d and a unimodular matrix M ∈ GL(n, Z) such that M (B) ⊂ Zr×{0} and (2.1) (2.2) (2.3) M RM −1 =(cid:20)A1 C 0 A2(cid:21) where A1 ∈ Mr(Z), C ∈ Mr,d−r(Z), A2 ∈ Md−r(Z). In addition, M (T (R, B)) ⊂ Rr × {0} and the Hadamard triple (R, B, L) is conjugate to the Hadamard triple (M RM −1, M B, (M T )−1L), which is a triple of lower dimension. If the lattice Z[R, B] is full rank but not Zd, then the system (R, B, L) is conjugate to one ( R, B, L) for which Z[ R, B] = Zd. Moreover, M is given by Z[R, B] = M (Zd). Proof. See Proposition 4.1 in [DL15] (cid:3) In the following, we introduce the main technique that will be used. We start with the following definition. Definition 2.4. Let u ≥ 0 be an entire function on Rd, i.e., real analytic on Rd. Let L be a simple digit set for RT . Suppose that Xl∈L u((RT )−1(x + l)) > 0, (x ∈ Rd) and all ℓ ∈ L A closed set K in Rd is called u-invariant (with respect to the system (u, RT , L)) if, for all x ∈ K We say that the transition, using ℓ, from x to τℓ(x) is possible, if ℓ ∈ L and u(cid:0)(RT )−1(x + ℓ)(cid:1) > 0. We say that K is Zd-periodic if K + n = K for all n ∈ Zd. u(cid:0)((RT )−1(x + ℓ)(cid:1) > 0 =⇒ (RT )−1(x + ℓ) ∈ K. We say that a subspace W of Rd is a rational subspace if W has a basis of vectors with rational components. The following theorem follows from Proposition 2.5, Theorem 2.8 and Theorem 3.3 in [CCR96]. Theorem 2.5. Let L be a complete set of representatives (mod RT (Zd)). Let u ≥ 0 be an entire function on Rd and let K be a closed u-invariant Zd-periodic set different from Rd. Suppose in addition that g is an entire function which is zero on K. Then (i) there exists a point x0 ∈ Rd, such that (RT )mx0 ≡ x0(mod Zd) for some integer m ≥ 1, and HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES 5 (ii) a proper rational subspace W (may equal {0}) such that RT (W ) = W and the union S = m−1[k=0 ((RT )kx0 + W + Zd) is invariant and g is zero on S. in (RT )k−1x0 + W + Zd. Moreover, all possible transitions from a point in (RT )kx0 + W + Zd, 1 ≤ k ≤ m, lead to a point Let (R, B, L) be a Hadamard triple and we aim to apply Theorem 2.5 to our set Z in (1.3). We define the function Taking the Fourier transform of the invariance equation (1.2), we can compute explicitly the Fourier transform of µ = µ(R, B) as , (x ∈ Rd). 1 N Xb∈B 2 e2πihb , xi(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) uB(x) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) bµ(ξ)2 = uB((RT )−1ξ)bµ((RT )−1(ξ))2, (x ∈ Rd). ∞Yn=1 bµ(x)2 = uB((RT )−nx), (x ∈ Rd), (2.4) (2.5) (2.6) Iterating (2.5), we obtain and the convergence in the product is uniform on compact sets. See e.g. [DJ07]. It is well known that both uB and bµ2 are entire functions on Rd. Proposition 2.6. Suppose that (R, B, L) forms a Hadamard triple and Z[R, B] = Zd. Let L be a complete set of representatives (mod RT (Zd)) containing L. Suppose that the set is non-empty. Then Z :=nξ ∈ Rd :bµ(ξ + k) = 0 for all k ∈ Zdo (i) Z is uB-invariant. (ii) There exist a point x0 ∈ Rd such that (RT )mx0 ≡ x0(mod(RT )Zd), for some integer m ≥ 1. (iii) There exists a proper rational subspace W 6= {0} of Rd such that RT (W ) = W and the union S = is uB-invariant and is contained in Z. ((RT )kx0 + W + Zd) m−1[k=0 Moreover, all possible transitions from a point in (RT )kx0 + W + Zd, 1 ≤ k ≤ m, lead to a point in (RT )k−1x0 + W + Zd. 6 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI Proof. We first prove that Z is uB-invariant. Take x ∈ Z and ℓ ∈ L such that uB((RT )−1(x+ℓ)) > 0. Let k ∈ Zd. We have, with (2.5), 0 = bµ(x + ℓ + RT k)2 = uB((RT )−1(x + ℓ + RT k))bµ((RT )−1(x + ℓ + RT k))2 Therefore, bµ((RT )−1(x + ℓ) + k) = 0 for all k ∈ Zd. So (RT )−1(x + ℓ) is in Z, and this shows that Since (R, B, L) form a Hadamard triple, by the Parseval identity, (see e.g. [ LW02, DJ07]), Z is uB-invariant. = uB((RT )−1(x + ℓ))bµ((RT )−1(x + ℓ) + k)2. Xl∈L Xl∈L uB((RT )−1(x + l)) = 1, uB((RT )−1(x + l)) > 0, (x ∈ Rd), (x ∈ Rd). (2.7) Hence, (2.8) We can apply Theorem 2.5 with u = uB and g = bµ to obtain all other conclusions except the non-triviality of W . We now check that W 6= {0}. Suppose W = {0}. First we show that for 1 ≤ k ≤ m there is a unique ℓ ∈ L such that uB((RT )−1((RT )kx0 + ℓ) > 0. Equation (2.8) shows that there exists at least one such ℓ. Assume that we have two different ℓ and ℓ′ in L with this property. Then the transitions are possible, so (RT )−1((RT )kx0 + ℓ) ≡ (RT )k−1x0 ≡ (RT )−1((RT )kx0 + ℓ′) (mod (Zd)). (2.9) But then ℓ ≡ ℓ′(mod RT (Zd)) and this is impossible since L is a complete set of representatives. By a translation, we can assume 0 ∈ B. From (2.7), and since the elements in L are distinct (mod RT (Zd)), we see that there is exactly one ℓk ∈ L such that uB((RT )−1((RT )kx0 + ℓk)) > 0. Therefore uB((RT )−1((RT )kx0 + ℓk)) = 1. But then, by (2.9), uB((RT )k−1x0) = 1. We have (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xb∈B e2πihb , (RT )k−1x0i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = N. As #B = N and 0 ∈ B, we have equality in the triangle inequality, and we get that e2πihb , (RT )kx0i = 1 for all b ∈ B. Then(cid:10)Rk−1b , x0(cid:11) ∈ Z for all b ∈ B, 1 ≤ k ≤ m. Because (RT )mx0 ≡ x0(mod Zd), we get that(cid:10)Rkb , x0(cid:11) ∈ Z for all k ≥ 0 and thus Since Z[R, B] = Zd, this means that x0 ∈ Zd. But x0 ∈ Z, so 1 =bµ(0) =bµ(x0 − x0) = 0, which is x0 ∈ Z[R, B]⊥ := {x ∈ Rd : hλ , xi ∈ Z for all λ ∈ Z[R, B]}. a contradiction. This shows W 6= {0}. As W 6= {0}, we can conjugate R through some M so that (R, B, L) has a much more regular (cid:3) structure. Proposition 2.7. Suppose that (R, B, L) forms a Hadamard triple and Z[R, B] = Zd and let µ = µ(R, B) be the associated self-affine measure µ = µ(R, B). Suppose that the set Z :=nx ∈ Rd :bµ(x + k) = 0 for all k ∈ Zdo , HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES 7 is non-empty. Then there exists an integer matrix M with det M = 1 such that the following assertions hold: (i) The matrix R := M RM −1 is of the form (2.10) R =(cid:20)R1 C R2(cid:21) , 0 with R1 ∈ Mr(Z), R2 ∈ Md−r(Z) expansive integer matrices and C ∈ M(d−r)×r(Z). (ii) If B = M B and L = (M T )−1L, then ( R, B, L) is a Hadamard triple. (iii) The measure µ(R, B) is spectral with spectrum Λ if and only if the measure µ( R, B) is spectral with spectrum (M T )−1Λ. (iv) There exists y0 ∈ Rd−r such that (RT 2 )my0 ≡ y0(mod(RT 2 )Zd) for some integer m ≥ 1 such that the union is contained in the set S = m−1[k=0 (Rr × {(RT 2 )ky0} + Zd) Z :=nx ∈ Rd :bµ(x + k) = 0 for all k ∈ Zdo , where µ = µ( R, B). The set S is invariant (with respect to the system (u B, RT , L), where L is a complete set of representatives (mod RT Zd). In addition, all possible transitions from 2 )k−1y0} + Zd. a point in Rr × {(RT Proof. We use Proposition 2.6 and we have x0 and a rational subspace W 6= {0} invariant for R with all the mentioned properties. By [Sch86, Theorem 4.1 and Corollary 4.3b], there exists an integer matrix M with determinant 1 such that M V = Rr ×{0}. The rest follows from Proposition 2.6, by conjugation, y0 is the second component of M x0. 2 )ky0} + Zd, 1 ≤ k ≤ m leads to a point in Rr × {(RT (cid:3) 3. The quasi-product form From now on, we assume that (R, B, L) satisfies all the properties of ( R, B, L) in Proposition In this section, we will prove that if Z 6= ∅, the Hadamard triple will be conjugate to a 2.7. quasi-product form structure. We first introduce the following notations. Definition 3.1. For a vector x ∈ Rd, we write it as x = (x(1), x(2))T with x(1) ∈ Rr and x(2) ∈ Rd−r. We denote by π1(x) = x(1), π2(x) = x(2). For a subset A of Rd, and x1 ∈ Rr, x2 ∈ Rd−r, we denote by A2(x1) := {y ∈ Rd−r : (x1, y)T ∈ A}, A1(x2) := {x ∈ Rr : (x, x2)T ∈ A}. We also make a note on the notation. Throughout the rest of the paper, we use A× B to denote the Cartesian product of A and B so that A × B = {(a, b) : a ∈ A, b ∈ B}. Our main theorem in this section is as follows: Theorem 3.2. Suppose that R =(cid:20)R1 C R2(cid:21) , 0 8 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI (R, B, L0) is a Hadamard triple and µ = µ(R, B) is the associated self-affine measure and Z 6= ∅. Then the set B has the following quasi-product form: (3.1) where B =(cid:8)(ui, vi + Qci,j)T : 1 ≤ i ≤ N1, 1 ≤ j ≤ det R2(cid:9) , (i) N1 = N/ det R2, (ii) Q is a (d − r) × (d − r) integer matrix with det Q ≥ 2 and R2Q = Q R2 for some (d − r) × (d − r) integer matrix fR2, (iii) the set {Qci,j : 1 ≤ j ≤ det R2} is a complete set of representatives (mod R2(Zd−r)), for all 1 ≤ i ≤ N1. Moreover, one can find some L ≡ L0(mod RT (Zd)) so that (R, B, L) is a Hadamard triple and (R1, π1(B), L1(ℓ2)) and (R2, B2(b1), π2(L)) are Hadamard triples on Rr and Rd−r respectively. The following lemma allows us to find a representative L with certain injectivity property. (3.2) Lemma 3.3. Suppose that the Hadamard triple (R, B, L) satisfies the properties (i) and (iv) in Proposition 2.7. Then there exists set L′ and a complete set of representatives L such that (R, B, L′) is a Hadamard triple and the following property holds: ′ 2 (Zd−r)(cid:17) =⇒ π2(ℓ) = π2(ℓ′). Proof. We note that if L ≡ L′ (mod RT (Zd)), then (R, B, L′) is a Hadamard triple since(cid:10)R−1b , RT m(cid:11) 2)T be in L (or L) such that ℓ2 ≡ ∈ Z for any m ∈ Zd. Now, ℓ′ 2(mod RT ) and π2(ℓ) ≡ π2(ℓ′) (cid:16)mod RT 2 (Zd−r)). We replace ℓ′ by let ℓ = (ℓ1, ℓ2)T , l′ = (ℓ′ ℓ, ℓ′ ∈ L′ (or ∈ L ′ 1, ℓ′ ℓ′′ = ℓ′ + RT (0, (RT 2 )−1(ℓ2 − ℓ′ 2))T ∈ Zd. Then ℓ′′ ≡ ℓ′(mod RT (Zd)) and the Hadamard property for L is preserved, the new set L is a complete set of representatives (mod RT (Zd)) and π2(ℓ′′) = l2. Repeating this procedure, we obtain our lemma. (cid:3) To simplify the notation, in what follows we relabel L′ by L and L property (3.2). We prove some lemmas for the proof of Theorem 3.2. ′ by L so that L and L possess Lemma 3.4. Suppose that the Hadamard triple (R, B, L) satisfies the properties (i) and (iv) in Proposition 2.7. Then (3.3) (i) For every b1 ∈ π1(B) and b2 6= b′ Xℓ2∈π2(L) 2 in B2(b1), #L1(ℓ2)e2πihR−1 2 (b2−b′ 2) , ℓ2i = 0. Also, for all b1 ∈ π1(B), #B2(b1) ≤ #π2(L) and the elements in B2(b1) are not congruent mod R2(Zd−r). (3.4) (ii) For every ℓ2 ∈ π2(L) and ℓ1 6= ℓ′ Xb1∈π1(B) 1 in L1(ℓ2), #B2(b1)e2πihR−1 1 b1 , (ℓ1−ℓ′ 1)i = 0. HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES 9 Also, for all ℓ2 ∈ π2(L), #L1(ℓ2) ≤ #π1(B) and the elements in L1(ℓ2) are not congruent (iii) The set π2(L) is a complete set of representatives (mod RT π2(L), the set L1(ℓ2) is a complete set of representatives (mod RT 1 (Zr)). 2 (Zd−r)) and, for every ℓ2 ∈ Proof. We first prove (i). Take b1 ∈ π1(B) and b2 6= b′ we have 2 in B2(b1), from the mutual orthogonality, (cid:0)mod RT 1 (Zr)(cid:1). 0 = Xℓ2∈π2(L) Xℓ1∈L1(ℓ2) 2)T , (ℓ1,ℓ2)Ti = Xℓ2∈π2(L) Xℓ1∈L1(ℓ2) #L1(ℓ2)e2πih(RT 2 )−1(b2−b′ 2) , ℓ2i. e2πih(RT 2 )−1(b2−b′ 2) , ℓ2i e2πih(RT )−1(0,b2−b′ = Xℓ2∈π2(L) (cid:16)p#L1(ℓ2)e2πih(RT This shows (3.3) and that the rows of the matrix 2 )−1b2 , ℓ2i(cid:17)ℓ2∈π2(L),b2∈B2(b1) are orthogonal. Therefore #B2(b1) ≤ #π2(L) for all b1 ∈ π1(B). Equation (3.3) implies that the elements in B2(b1) cannot be congruent mod RT 2 (Zd−r) (ii) follows from an analogous computation. For (iii), the elements in π2(L) are not congruent (mod RT 2 (Zd−r)), by the construction in 1 (Zr)) then (ℓ1, ℓ2)T ≡ 1 ∈ L1(l2) are congruent (mod RT 1, as L is a complete set of representatives of RT (Zd). From If ℓ2 ∈ π2(L) and ℓ1, ℓ′ Lemma 3.3. (ℓ′ 1, ℓ2)T (mod RT (Zd)). Thus, ℓ1 = ℓ′ these, we have #π2(L) ≤ det R2 and, for all ℓ2 ∈ π2(L), #L1(ℓ2) ≤ det R1. Since det R = det R1 det R2 ≥ Xℓ2∈π2(L) #L1(ℓ2) = #L = det R, we must have equalities in all inequalities and we get that the sets are indeed complete sets of representatives. (cid:3) Lemma 3.5. Let 1 ≤ j ≤ m. If the the transition from (x, (RT ℓ ∈ L, then π2(ℓ) = 0. Proof. If the transition is possible with digit ℓ = (ℓ1, ℓ2)T , then, by Proposition 2.7, 2 )jy0)T is possible with the digit (RT )−1((x, (RT for some y ∈ Rr, and therefore (RT 3.3, ℓ2 = 0. 2 )jy0)T + (ℓ1, ℓ2)T ) ≡ (y, (RT 2 )−1ℓ2 ≡ 0 (mod Zd−r), so ℓ2 ≡ 0 (mod RT 2 )j−1y0)T (mod Zd), 2 (Zd−r)). By Lemma (cid:3) 2 )jy0, 1 ≤ j ≤ m. Then, for all x ∈ Rr and all ℓ = (ℓ1, ℓ2) ∈ L with Lemma 3.6. Let yj := (RT π2(ℓ) = ℓ2 6= 0, we have that (3.5) Xb2∈B2(b1) e2πihb2 , (RT 2 )−1(yj+ℓ2)i = 0. 10 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI Proof. We have that (RT )−1 is of the form (RT )−1 =(cid:20)(RT 1 )−1 0 D (RT 2 )−1(cid:21) . Then, for all x ∈ Rr and all ℓ = (ℓ1, ℓ2) ∈ L with π2(ℓ) = ℓ2 6= 0, we have that uB((RT )−1((x, yj )T + (ℓ1, ℓ2)T )) = 0, because such transitions are not possible by Lemma 3.5. Then e2πi(hb1 , R−1 1 (x+l1)+D(y1+l2)i+hb2 , (RT 2 )−1(yj +l2)i) = 0. Since x is arbitrary, it follows that Xb1∈π1(B) Xb2∈B2(b1) Xb1∈π1(B) e2πihb1 , xi Xb2∈B2(b1) e2πihb2 , (RT 2 )−1(yj +l2)i = 0 for all x ∈ Rd. Therefore, by linear independence of exponential functions, we obtain (3.5). (cid:3) Lemma 3.7. For every b1 ∈ π1(B), the set B2(b1) is a complete set of representatives (mod R2(Zd−r)). Therefore #B2(b1) = det R2 = #π2(L) and (R2, B2(b1), π2(L)) is a Hadamard triple. Also, for every ℓ2 ∈ π2(L), #L1(ℓ2) = #π1(B) = det R2 =: N1 and (R1, π1(B), L1(ℓ2)) is a Hadamard triple. Proof. Let b1 ∈ π1(B). We know from Lemma 3.4(i) that the elements of B2(b1) are not congruent (mod R2(Zd−r)). We can identify B2(b1) with a subset of the group Zd−r/R2(Zd−r). The dual 2 (Zd−r) which we can identify with π2(L). For a function f on Zd−r/R2(Zd−r), group is Zd−r/RT the Fourier transform is N f (ℓ2) = 1 p det R2 Xb2∈Zd−r/R2(Zd−r) Let 1 ≤ j ≤ m. Consider the function f (b2) =( e−2πihb2 , (RT 0, 2 )−1yji, (3.6) f (b2)e−2πihb2 , (RT 2 )−1ℓ2i, (ℓ2 ∈ π2(L2)). if b2 ∈ B2(b1) if b2 ∈ (Zd−r/R2(Zd−r)) \ B2(b1). Then equation (3.5) shows that f (ℓ2) = 0 for ℓ2 ∈ π2(L), ℓ2 6= 0. Thus f = c· χ0 for some constant c and by f (0) = c, (3.7) c = e−2πihb2 , (RT 2 )−1(yj )i. 1 p det R2 Xb2∈B2(b1) Now we apply the inverse Fourier transform and we get f (b2) = 1 p det R2 Xℓ2∈π2(L) cχ0(ℓ2)e2πihb2 , (RT 2 )−1ℓ2i = c . p det R2 So f (b2) is constant and therefore B2(b1) = Zd−r/R2(Zd−r), which means that B2(b1) is a complete set of representatives and #B2(b1) = det R2. Since the elements in #π2(L) are not congruent Zd−r), we get that #π2(L) ≤ det R2, and with Lemma 3.4 (i), it follows that #π2(L) = (mod RT 2 11 2 (Zd−r)), so HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES #B2(b1) = det R2. In particular, π2(L) is a complete set of representatives (mod RT (R2, B2(b1), π1(L)) form a Hadamard triple. SincePb1∈π1(B) #B2(b1) = N , we get that #π1(B) = N/ det R2. With Lemma 3.4(ii), we have #L1(ℓ2) ≤ #π2(L)π1(B) = N. N = Xℓ2∈π2(L) Therefore we have equality in all inequalities so #L1(ℓ2) = #π1(B) = N/ det R2. Then (3.4) shows that (R1, π1(B), L1(ℓ2)) is a Hadamard triple for all ℓ2 ∈ π2(L). (cid:3) Proof of Theorem 3.2. By Lemma 3.7, we know that B must have the form [b1∈π1(B) {b1} × B2(b1) where #π1(B) = N1 and B2(b1) is a set of complete representative (mod RT ating elements π1(B) = {u1, ..., uN1} and B2(ui) = {di,1, ..., di, det R2}, we can write 2 (Zd−r)). By enumer- B =(cid:8)(ui, di,j)T : 1 ≤ i ≤ N1, 1 ≤ j ≤ det R2(cid:9) . It suffices to show di,j are given by vi +Qci,j where Q has the properties (ii) and (iii) in the theorem. From the equation (3.6) and the fact that f is a constant, we have, for b2 ∈ B2(b1) and b1 ∈ π1(B), e−2πihb2 , (RT 2 )−1yji = f (b2) = c , p det R2 e2πih(b2−b′ 2) , (RT 2 )−1yji = 1. which implies (from (3.7)) that 1 det R2 Xb′ 2∈B2(b1) By applying the triangle inequality to the sum above, we see that we must have e2πihb2−b′ 2 , (RT 2 )−1yji = 1, 2 ∈ B2(b1), b1 ∈ π1(B), 1 ≤ j ≤ m. which means 2 , (RT (3.8) Here we recall that yj = (RT (cid:10)b2 − b′ Define now the lattice 2 )jy0. 2 )−1yj(cid:11) ∈ Z for all b2, b′ Γ := {x ∈ Zd−r :(cid:10)x , (RT 2 )−1yj(cid:11) ∈ Z, ∀ 1 ≤ j ≤ m}. We first claim that the lattice Γ is of full-rank. Indeed, since (RT that all the points (RT all the denominators of all the components of the vectors (RT 2 )myj ≡ yj(mod Zd−r), it follows 2 )−1(yj) have only rational components. Let m be a common multiple for 2 )−1(yj), 1 ≤ j ≤ m. If {ei} are the Next we prove that Γ is a proper sublattice of Zd−r. The vectors yj are not in Zd−r because canonical vectors in Rd−r, then(cid:10) mei , (RT bµ((0, yj)T + k) = 0 for all k ∈ Zd, and that would contradict the fact that bµ(0) = 1. This implies 2 )−1(yj)(cid:11) ∈ Z so mei ∈ Γ, and thus Γ is full-rank. 2 )−1(yj) are not in Zd−r so Γ is a proper sublattice of Zd−r. that the vectors (RT 12 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI Since Γ is a full-rank lattice in Zd−r, there exists an invertible matrix with integer entries Q such that Γ = QZd−r, and since Γ is a proper sublattice, it follows that det Q > 1 so det Q ≥ 2. In addition, we know from (3.8) that, for all ui ∈ π1(B) and di,j, di,j′ ∈ B2(ui), di,j − di,j′ ∈ Γ. Therefore, if we fix an element vi = di,j0 ∈ B2(ai), then all the elements in B2(ai) are of the form di,j = vi + Qci,j for some ci,j ∈ Zd−r. The fact that B2(ai) is a complete set of representatives (mod R2Zd−r) (Lemma 3.7) implies that the set of the corresponding elements Qci,j is also a complete set of representatives (mod R2Zd−r). This shows (iii). It remains to show R2Q = QfR2 for some for some (d − r) × (d − r) integer matrix fR2. Indeed, if x ∈ Γ, and 0 ≤ j ≤ m − 1, then (cid:10)R2x , (RT 2 )my0 ≡ y0 (mod Zd−r). So R2x ∈ Γ. Then, for the canonical vectors ei, there exist ei ∈ Zd−r such that R2Qei = Qeei. Let fR2 be the matrix with columns eei. Then R2Q = QfR2. Finally, by choosing L with the property Lemma 3.3, the Hadamard triple properties of both (R1, π1(B), L1(ℓ2)) and (R2, B2(b1), π1(L)) on Rr and Rd−r respectively are direct consequences of Lemma 3.7. (cid:3) 2 )j+1y0(cid:11) ∈ Z, since (RT 2 )jy0(cid:11) = (cid:10)x , (RT In the last section, we will prove our main theorem. We first need to study the spectral property 4. Proof of the theorem of the quasi-product form. Suppose now the pair (R, B) is in the quasi-product form (4.1) R =(cid:20)R1 C R2(cid:21) 0 (4.2) and {di,j : 1 ≤ j ≤ N2} (di,j = vi + Qci,j as in Theorem 3.2) is a complete set of representatives (mod R2Zd−r). We will show that the measure µ = µ(R, B) has a quasi-product structure. B =(cid:8)(ui, di,j)T : 1 ≤ i ≤ N1, 1 ≤ j ≤ N2 := det R2(cid:9) , Note that we have and, by induction, R−k =(cid:20)R−k For the invariant set T (R, B), we can express it as a set of infinite sums, R−(l+1) 2 CR−(k−l) 1 . 1 0 1 Dk R−k R−1 2 CR−1 −R−1 1 R−1 =(cid:20) 2 (cid:21) , where Dk := − T (R, B) =( ∞Xk=1 ∞Xk=1 ∞Xk=1 1 aik , R−k y = 0 R−1 2 (cid:21) k−1Xl=0 ∞Xk=1 R−kbk : bk ∈ B) . Therefore any element (x, y)T ∈ T (R, B) can be written in the following form x = Dkaik + R−k 2 dik,jk . HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES 13 Let X1 be the attractor (in Rr) associated to the IFS defined by the pair (R1, π1(B) = {ui : 1 ≤ i ≤ N1}) (i.e. X1 = T (R1, π1(B))). Let µ1 be the (equal-weight) invariant measure associated to this pair. For each sequence ω = (i1i2 . . . ) ∈ {1, . . . , N1}N = {1, . . . , N1} × {1, . . . , N1} × ..., define (4.3) x(ω) = R−k 1 uik . ∞Xk=1 As (R1, π1(B)) forms Hadamard triple with some L1(ℓ2) by Lemma 3.7, the measure µ(R1, π1(B)) has the no-overlap property (Theorem 1.4). It implies that for µ1-a.e. x ∈ X1, there is a unique ω such that x(ω) = x. We define this as ω(x). This establishes a bijective correspondence, up to measure zero, between the set Ω1 := {1, . . . , N1}N and X1. The measure µ1 on X1 is the pull-back of the product measure which assigns equal probabilities 1 N1 to each digit. For ω = (i1i2 . . . ) in Ω1, define For ω ∈ Ω1, define g(ω) :=P∞ For x ∈ X1, define Ω2(ω) := {(di1,j1di2,j2 . . . din,jn . . . ) : jk ∈ {1, . . . , N2} for all k ∈ N}. k=1 Dkaik and g(x) := g(ω(x)), for x ∈ X1. Also Ω2(x) := Ω2(ω(x)). X2(x) := X2(ω(x)) :=( ∞Xk=1 2 dik,jk : jk ∈ {1, . . . , N2} for all k ∈ N) . R−k Note that the attractor T (R, B) has the following form T (R, B) = {(x, g(x) + y)T : x ∈ X1, y ∈ X2(x)}. For ω ∈ Ω1, consider the product probability measure µω, on Ω2(ω), which assigns equal ω on X2(ω). Let to each digit dik,jk at level k. Next, we define the measure µ2 probabilities rω : Ω2(ω) → X2(ω), 1 N2 rω(di1,j1di2,j2 . . . ) = Define µ2 x := µ2 ω(x) := µω(x) ◦ r−1 ω(x). R−k 2 dik,jk . ∞Xk=1 Note that the measure µ2 x is the infinite convolution product δR−1 ω(x) = (i1i2 . . . ), B2(ik) := {dik,j : 1 ≤ j ≤ N2} and δA := 1 The following lemmas were proved in [DJ07]. 2 B2(i1) ∗ δR−2 2 B2(i2) ∗ . . . , where #APa∈A δa, for a subset A of Rd−r. Lemma 4.1. [DJ07, Lemma 4.4] For any bounded Borel functions on Rd, Lemma 4.2. [DJ07, Lemma 4.5] If Λ1 is a spectrum for the measure µ1, then f dµ =ZX1ZX2(x) ZT (R,B) bµ(x + λ1, y)2 =ZX1 bµ2 F (y) := Xλ1∈Λ1 f (x, y + g(x)) dµ2 x(y) dµ1(x). s(y)2 dµ1(s), (x ∈ Rr, y ∈ Rd−r). 14 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI We recall also the Jorgensen-Pedersen Lemma for checking in general when a countable set is a spectrum for a measure. Lemma 4.3. [JP98] Let µ be a compactly supported probability measure. Then Λ is a spectrum for L2(µ) if and only if We need the following key proposition. Xλ∈Λ bµ(ξ + λ)2 ≡ 1. Proposition 4.4. For the quasi-product form given in (4.1) and (4.2), there exists a lattice Γ2 such that for µ1-almost every x ∈ X1, the set Γ2 is a spectrum for the measure µ2 x. Proof. First we replace the first component (R1, π1(B)) by a more convenient pair which allows us to use the theory of self-affine tiles from [LW97]. Define R† :=(cid:20)N1 0 R2(cid:21) , 0 We will use the super-script † to refer to the pair (R†, B†). B† =(cid:8)(i, di,j )T : 1 ≤ i ≤ N1 − 1, 1 ≤ j ≤ det R2(cid:9) . As di,j is a complete residue (mod R2(Zd−r)), the set B† is a complete set of representatives (mod R†(Zd−r+1)). By [LW97, Theorem 1.1], µ† is the normalized Lebesgue measure on T (R†, B†) and this tiles Rd−r+1 with some lattice Γ∗ ⊂ Zd−r+1. The attractor X † 1 corresponds to the pair (N1,{0, 1, . . . , N1 − 1}) so X † 1 is the Lebesgue measure on [0, 1]. We need the following claim: 1 is [0, 1] and µ† Claim: the set T (R†, B†) actually tiles Rd−r+1 with a set of the form Z× Γ2, where Γ2 is a lattice in Rd−r. Proof of claim: This claim was established implicitly in the proof of Theorem 1.1, in section 7, p.101 of [LW97], we present it here for completeness. Let Γ∗ be the lattice on Rd−r+1 which is a tiling set of T (R†, B†). We observe that the orthogonal projection of T (R†, B†) to the first coordinate is [0, 1]. Hence, for any γ ∈ Γ∗, the orthogonal projection of T (R†, B†) is [0, 1] + γ1, where γ = (γ1, γ2)T . As Γ∗ ⊂ Zd−r+1. these projections [0, 1] + γ1 are measure disjoint for different γ1's. Therefore, the tiling of T (R†, B†) by Γ∗ naturally divides up into cylinders: Focusing on one of the cylinders, say U (0), this cylinder is tiled by Γ where U (γ1) := ([0, 1] + γ1) × Rd−r. Γ = Γ ∩ ({0} × Zd−r). claim. As Rd−r+1 =Sγ1 U (γ1), this means T (R†, B†) also tiles by Z× Γ2. This completes the proof of the spectrum of the form Z × Γ2, with Γ2 the dual lattice of Γ2. Because of the claim, it follows from the well-known result of Fuglede [Fug74] that µ† has a HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES 15 We prove that Γ2 is an orthogonal set for the measure µ†2 s for µ† 1-almost every s ∈ X † 1. Indeed, for γ2 6= 0 in Γ2, since Z × Γ2 is a spectrum for µ†, we have for all λ1 ∈ Z, with Lemma 4.1, 1(x) 1ZX † e−2πih(λ1,γ2) , (x,y)i dµ†(x, y) =ZX † 0 =ZT (R†,B†) e−2πih(λ1,γ2) , (x,y)i dµ†2 x(y) dµ† 2(x) This implies that that =ZX † 1 e−2πiλ1xZX † ZX † 2(x) e−2πihγ2 , yi dµ†2 x(y) dµ† 1(x). 2(x) e−2πihγ2 , yi dµ†2 x(y) = 0, 1. This means that Γ2 is an orthogonal sequence for µ†2 x for µ† 1-a.e. for all γ2 ∈ Γ2 for µ† x ∈ X † 1 so (4.4) for µ† 1-a.e. x ∈ X † With (4.4), we have (4.5) for µ† 4.3. 1-a.e. s ∈ X † 1. With Lemma 4.4, we have dµ†2 x(y + γ2)2 ≤ 1, 1-a.e. x ∈ X † Xγ2∈Γ2 bµ†(x + λ1, y + γ2)2 =ZX † 1 Xγ2∈Γ2 Xγ2∈Γ2 cµ†2 s(y + γ2)2 = 1, 1 = Xγ2∈Γ2 Xλ1∈Z (y ∈ Rd−r), 1(s). cµ†2 s(y + γ2)2 dµ† (y ∈ Rd−r), 1, which means that Γ2 is a spectrum for almost every measure µ†2 s by Lemma Now, we are switching back to our original pair (R, B). Note that we have the maps x : Ω1 → X1 1, defined by ω 7→ x(ω) as above in (4.3), and analogously for x†. The maps are 1 be the composition ψ = x† ◦ x−1. i.e. and x† : Ω1 → X † measure preserving bijections. Let Ψ : X1 → X † R−j Ψ ∞Xj=1 1 uj = N −j 1 j. ∞Xj=1 Consider the measure ν(E) = µ† condition, we can check easily that ν and µ1 agrees on all the cylinder sets of T (R1, π1(B)). i.e. 1(Ψ(E)) for Borel set E in T (R1, π1(B)). Because of the no-overlap ν(τi1 ◦ ... ◦ τin(T (R1, π1(B)))) = 1 N n = µ1(τi1 ◦ ... ◦ τin(T (R1, π1(B)))) for all i1, ..., in ∈ {0, 1, ..., N1 − 1}. This shows that ν = µ1 and therefore µ1(E) = µ† any Borel set E. Consider the set 1(Ψ(E)) for Then N = {x ∈ T (R1, π1(B)) : Γ2 is not a spectrum for µ2 x} Ψ(N ) = {Ψ(x) ∈ X † 1 : Γ2 is not a spectrum for µ2 x} 16 DORIN ERVIN DUTKAY, JOHN HAUSSERMANN, AND CHUN-KIT LAI Note also that, on the second component, the two pairs (R, B) and (R†, B†) are the same, more precisely X2(x) = X † x = µ†2 2(Ψ(x)) and µ2 Ψ(N ) = {Ψ(x) ∈ X † Ψ(x) for all x ∈ X1. This means that 1 : Γ2 is not a spectrum for µ2 Ψ(x)} which has µ† 0 and this completes the proof. 1-measure 0, by the arguments in the previous paragraph. Hence, µ1(E) = µ† 1(Ψ(E)) = (cid:3) Proof of Theorem 1.3. To prove Theorem 1.3, we use induction on the dimension d. We know from [ LW02, DJ06] that the result is true in dimension one (See also [DL15, Example 5.1] for an independent proof by considering Z). Assume it is true for any dimensions less than d. First, after some conjugation, we can assume that Z[R, B] = Zd, according to Proposition 2.3. Next, if the set in (1.3) Z = ∅, then the result follows from Theorem 1.4 (ii). Suppose now that Z 6= ∅. Then, by Proposition 2.7, we can conjugate with some matrix so that (R, B) are of the quasi-product form given in (4.1) and (4.2). By Theorem 3.2, (R1, π1(B), L1(ℓ2)) forms a Hadamard triple with some L on Rr where 1 ≤ r < d. By induction hypothesis, the measure µ1 is spectral. Let Λ1 be a spectrum for µ1. By Proposition 4.4, there exists Γ2 such that Γ2 is a spectrum for µ2 x for µ1-almost everywhere x. Then we have, with (4.5), and Lemma 4.2, Xγ2∈Γ2 Xλ1∈Λ1 bµ(x + λ1, y + γ2)2 =ZX1 Xγ2∈Γ2 s(y + γ2)2 dµ1(s) =ZX1 cµ2 1dµ1(s) = 1. This means that Λ1 × Γ2 is a spectrum for µ by Lemma 4.3 and this completes the whole proof of Theorem 1.3. (cid:3) Acknowledgements. This work was partially supported by a grant from the Simons Foundation (#228539 to Dorin Dutkay). References [CCR96] D. Cerveau, J.-P. Conze, and A. Raugi. Ensembles invariants pour un op´erateur de transfert dans Rd. Bol. [DJ06] [DJ07] [DL14] [DL15] [DS15] [Fal97] Soc. Brasil. Mat. (N.S.), 27(2):161 -- 186, 1996. Dorin Ervin Dutkay and Palle E. T. Jorgensen. Iterated function systems, Ruelle operators, and invariant projective measures. Math. Comp., 75(256):1931 -- 1970 (electronic), 2006. Dorin Ervin Dutkay and Palle E. T. Jorgensen. Fourier frequencies in affine iterated function systems. J. Funct. Anal., 247(1):110 -- 137, 2007. Dorin Ervin Dutkay and Chun-Kit Lai. Uniformity of measures with Fourier frames. Advances in. Math., 252:684 -- 707, 2014. Dorin Ervin Dutkay and Chun-Kit Lai. Self-affine spectral measures and frame spectral measures on Rd. submitted, 2015. Xin-Rong Dai and Qiyu Sun. Spectral measures with arbitrary Hausdorff dimensions. J. Funct. Anal., 268(8):2464 -- 2477, 2015. Kenneth Falconer. Techniques in Fractal geometry. John Wiley & Sons Ltd., 1997. Mathematical founda- tions and applications. [FMM06] B´alint Farkas, M´at´e Matolcsi, and P´eter M´ora. On Fuglede's conjecture and the existence of universal spectra. J. Fourier Anal. Appl., 12(5):483 -- 494, 2006. HADAMARD TRIPLES GENERATE SELF-AFFINE SPECTRAL MEASURES 17 [Fug74] Bent Fuglede. Commuting self-adjoint partial differential operators and a group theoretic problem. J. Funct. [Hut81] [JP98] Anal., 16:101 -- 121, 1974. John E. Hutchinson. Fractals and self-similarity. Indiana Univ. Math. J., 30(5):713 -- 747, 1981. Palle E. T. Jorgensen and Steen Pedersen. Dense analytic subspaces in fractal L2-spaces. J. Anal. Math., 75:185 -- 228, 1998. [KM06a] Mihail N. Kolountzakis and M´at´e Matolcsi. Complex Hadamard matrices and the spectral set conjecture. Collect. Math., (Vol. Extra):281 -- 291, 2006. [KM06b] Mihail N. Kolountzakis and M´at´e Matolcsi. Tiles with no spectra. Forum Math., 18(3):519 -- 528, 2006. [Li14] [Li15] [LW97] Jian-Lin Li. Analysis of µM , D-orthogonal exponentials for the planar four-element digit sets. Math. Nachr., 287(2-3):297 -- 312, 2014. Jian-Lin Li. Spectral self-affine measures on the spatial Sierpinski gasket. Monatsh. Math., 176(2):293 -- 322, 2015. Jeffrey C. Lagarias and Yang Wang. Integral self-affine tiles in Rn. II. Lattice tilings. J. Fourier Anal. Appl., 3(1):83 -- 102, 1997. Izabella Laba and Yang Wang. On spectral Cantor measures. J. Funct. Anal., 193(2):409 -- 420, 2002. [ LW02] [Mat05] M´at´e Matolcsi. Fuglede's conjecture fails in dimension 4. Proc. Amer. Math. Soc., 133(10):3021 -- 3026 (elec- tronic), 2005. [Sch86] Alexander Schrijver. Theory of linear and integer programming. Wiley-Interscience Series in Discrete Math- [Str98] [Str00] ematics. John Wiley & Sons, Ltd., Chichester, 1986. A Wiley-Interscience Publication. Robert S. Strichartz. Remarks on: "dense analytic subspaces in fractal l2-spaces". J. Anal. Math., 75:229 -- 231, 1998. Robert S. Strichartz. Mock Fourier series and transforms associated with certain Cantor measures. J. Anal. Math., 81:209 -- 238, 2000. Robert S. Strichartz. Convergence of mock Fourier series. J. Anal. Math., 99:333 -- 353, 2006. [Str06] [Tao04] Terence Tao. Fuglede's conjecture is false in 5 and higher dimensions. Math. Res. Lett., 11(2-3):251 -- 258, 2004. [YL15] Ming-Shu Yang and Jian-Lin Li. A class of spectral self-affine measures with four-element digit sets. J. Math. Anal. Appl., 423(1):326 -- 335, 2015. [Dorin Ervin Dutkay] University of Central Florida, Department of Mathematics, 4000 Central Florida Blvd., P.O. Box 161364, Orlando, FL 32816-1364, U.S.A., E-mail address: [email protected] [John Haussermann] University of Central Florida, Department of Mathematics, 4000 Central Florida Blvd., P.O. Box 161364, Orlando, FL 32816-1364, U.S.A., E-mail address: [email protected] [Chun-Kit Lai] Department of Mathematics, San Francisco State University, 1600 Holloway Avenue, San Francisco, CA 94132. E-mail address: [email protected]
1512.03202
1
1512
2015-12-10T10:32:01
Norms supporting the Lebesgue differentiation theorem
[ "math.FA" ]
A version of the Lebesgue differentiation theorem is offered, where the $L^p$ norm is replaced with any rearrangement-invariant norm. Necessary and sufficient conditions for a norm of this kind to support the Lebesgue differentiation theorem are established. In particular, Lorentz, Orlicz and other customary norms for which Lebesgue's theorem holds are characterized.
math.FA
math
NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A Abstract. A version of the Lebesgue differentiation theorem is offered, where the Lp norm is replaced with any rearrangement-invariant norm. Necessary and sufficient conditions for a norm of this kind to support the Lebesgue differentiation theorem are established. In par- ticular, Lorentz, Orlicz and other customary norms for which Lebesgue's theorem holds are characterized. 1. Introduction and main results 5 1 0 2 c e D 0 1 ] . A F h t a m [ 1 v 2 0 2 3 0 . 2 1 5 1 : v i X r a A standard formulation of the classical Lebesgue differentiation theorem asserts that, if u ∈ L1 loc(Rn), n ≥ 1, then 1 u(y) dLn(y) exists and is finite for a.e. x ∈ Rn, (1.1) Ln(Br(x))ZBr(x) lim r→0+ where Ln denotes the Lebesgue measure in Rn, and Br(x) the ball, centered at x, with radius r. Here, and in what follows, "a.e." means "almost every" with respect to Lebesgue measure. In addition to (1.1), one has that (1.2) ku − u(x)k⊘ L1(Br(x)) = 0 for a.e. x ∈ Rn, lim r→0+ L1(Br(x)) stands for the averaged norm in L1(Br(x)) with respect to the normalized where k · k⊘ Lebesgue measure 1 Ln(Br (x)) Ln. Namely, L1(Br (x)) = kuk⊘ 1 Ln(Br(x))ZBr(x) u(y) dLn(y) for u ∈ L1 loc(Rn). A slight extension of this property ensures that an analogous conclusion holds if the L1-norm in (1.2) is replaced with any Lp-norm, with p ∈ [1, ∞). Indeed, if u ∈ Lp loc(Rn), then (1.3) the averaged norm k · k⊘ p = ∞. ku − u(x)k⊘ Lp(Br (x)) = 0 lim r→0+ Lp(Br(x)) being defined accordingly. By contrast, property (1.3) fails when for a.e. x ∈ Rn, 2000 Mathematics Subject Classification. 46E35, 46E30. Key words and phrases. Lebesgue differentiation theorem, rearrangement-invariant spaces, Lorentz spaces, Orlicz spaces, Marcinkiewicz spaces. This research was partly supported by the research project of MIUR (Italian Ministry of Education, Univer- sity and Research) Prin 2012, n. 2012TC7588, "Elliptic and parabolic partial differential equations: geometric aspects, related inequalities, and applications", by GNAMPA of the Italian INdAM (National Institute of High Mathematics, by the grant P201/13/14743S of the Grant Agency of the Czech Republic, and by the research project GA UK No. 62315 of the Charles University in Prague. 1 2 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A The question thus arises of a characterization of those norms, defined on the space L0(Rn) of measurable functions on Rn, for which a version of the Lebesgue differentiation theorem continues to hold. In the present paper we address this issue in the class of all rearrangement-invariant norms, i.e. norms which only depend on the "size" of functions, or, more precisely, on the measure of their level sets. A precise definition of this class of norms, as well as other notions employed hereafter, can be found in Section 2 below, where the necessary background material is collected. Let us just recall here that, if k · kX(Rn) is a rearrangement-invariant norm, then (1.4) kukX(Rn) = kvkX(Rn) whenever u∗ = v∗, where u∗ and v∗ denote the decreasing rearrangements of the functions u, v ∈ L0(Rn). More- over, given any norm of this kind, there exists another rearrangement-invariant function norm k · kX(0,∞) on L0(0, ∞), called the representation norm of k · kX(Rn), such that (1.5) kukX(Rn) = ku∗kX(0,∞) for every u ∈ L0(Rn). By X(Rn) we denote the Banach function space, in the sense of Luxem- burg, of all functions u ∈ L0(Rn) such that kukX(Rn) < ∞. Classical instances of rearrangement- invariant function norms are Lebesgue, Lorentz, Orlicz, and Marcinkiewicz norms. In analogy with (1.3), a rearrangement-invariant norm k · kX(Rn) will be said to satisfy the Lebesgue point property if, for every u ∈ Xloc(Rn), (1.6) Here, k · k⊘ measure 1 ku − u(x)k⊘ X(Br (x)) = 0 for a.e. x ∈ Rn. lim r→0+ X(Br (x)) denotes the norm on X(Br(x)) with respect to the normalized Lebesgue Ln(Br (x)) Ln -- see (2.15), Section 2. We shall exhibit necessary and sufficient conditions for k · kX(Rn) to enjoy the Lebesgue point property. To begin with, a necessary condition for k · kX(Rn) to satisfy the Lebesgue point property is to be locally absolutely continuous (Proposition 3.1, Section 3). This means that, for each function u ∈ Xloc(Rn), one has lim kuχKj kX(Rn) = 0 for every non-increasing sequence j→∞ {Kj} of bounded measurable sets in Rn such that ∩j∈NKj = ∅. The local absolute continuity of k · kX(Rn) is in turn equivalent to the local separability of X(Rn), namely to the separability of each subspace of X(Rn) consisting of all functions which are supported in any given bounded measurable subset of Rn. As will be clear from applications of our results to special instances, this necessary assumption is not yet sufficient. In order to ensure the Lebesgue point property for k · kX(Rn), it has to be complemented with an additional assumption on the functional GX, associated with the representation norm k · kX(0,∞), and defined as (1.7) GX (f ) = kf−1kX(0,∞) for every non-increasing function f : [0, ∞) → [0, ∞]. Here, f−1 : [0, ∞) → [0, ∞] denotes the (generalized) right-continuous inverse of f . Such an assumption amounts to requiring that GX be "almost concave". By this expression, we mean that the functional GX , restricted to the convex set C of all non-increasing functions from [0, ∞) into [0, 1], fulfils the inequality in the definition of concavity possibly up to a multiplicative positive constant c. Namely, (1.8) c kXi=1 λiGX (fi) ≤ GX(cid:18) kXi=1 λifi(cid:19) NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 3 for any numbers λi ∈ (0, 1), i = 1, . . . , k, k ∈ N, such that Pk i=1 λi = 1, and any functions fi ∈ C, i = 1, . . . , k. Clearly, the functional GX is concave on C, in the usual sense, if inequality (1.8) holds with c = 1. Theorem 1.1. A rearrangement-invariant norm k · kX(Rn) satisfies the Lebesgue point property if, and only if, it is locally absolutely continuous and the functional GX is almost concave. Remark 1.2. In order to give an idea of how the functional GX looks like in classical instances, consider the case when k · kX(Rn) = k · kLp(Rn). One has that GLp(f ) =((cid:0)pR ∞0 sp−1f (s) dL1(s)(cid:1) 1 L1({s ∈ [0, ∞) : f (s) > 0}) p if p ∈ [1, ∞), if p = ∞ , for every non-increasing function f : [0, ∞) → [0, ∞]. The functional GLp is concave for every p ∈ [1, ∞]. However, k · kLp(Rn) is locally absolutely continuous only for p < ∞. Remark 1.3. The local absolute continuity of a rearrangement invariant norm k · kX(Rn) and the almost concavity of the functional GX are independent properties. For instance, as noticed in the previous remark, the norm k · kL∞(Rn) is not locally absolutely continuous, although the functional GL∞ is concave. On the other hand, whenever q < ∞, the Lorentz norm k · kLp,q(Rn) is locally absolutely continuous , but GLp,q is almost concave if and only if q ≤ p. The Luxemburg norm k · kLA(Rn) in the Orlicz space LA(Rn) is almost concave for every N -function A, but is locally absolutely continuous if and only if A satisfies the ∆2-condition near infinity. These properties are established in Section 6 below, where the validity of the Lebesgue point property for various classes of norms is discussed. An alternative characterization of the rearrangement-invariant norms satisfying the Lebesgue point property involves a maximal function operator associated with the norms in question. The relevant operator, denoted by MX , is defined, at each u ∈ Xloc(Rn), as (1.9) where B stands for any ball in Rn. MXu(x) = sup B∋x kuk⊘ X(B) for x ∈ Rn, In the case when X(Rn) = L1(Rn), the operator MX coincides with the classical Hardy- Littlewood maximal operator M. It is well known that M is of weak type from L1(Rn) into L1(Rn). Moreover, since ku∗k⊘ L1(0,s) = u∗(t) dL1(t) for s ∈ (0, ∞), 1 sZ s 0 for every u ∈ L1 loc(Rn), the celebrated Riesz-Wiener inequality takes the form (Mu)∗(s) ≤ Cku∗k⊘ L1(0,s) for s ∈ (0, ∞), for some constant C = C(n) [4, Theorem 3.8, Chapter 3]. The validity of the Lebesgue point property for a rearrangement-invariant norm k · kX(Rn) turns out to be intimately connected to a suitable version of these two results for the maximal operator MX defined by (1.9). This is the content of our next result, whose statement makes use of a notion of weak-type operators between local rearrangement-invariant spaces. We say that MX is of weak type from Xloc(Rn) into L1 loc(Rn) if for every bounded measurable set K ⊆ Rn, there exists a constant C = C(K) such that (1.10) Ln({x ∈ K : MX u(x) > t}) ≤ C t kukX(Rn) for t ∈ (0, ∞), for every function u ∈ Xloc(Rn) whose support is contained in K. 4 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A Theorem 1.4. Let k·kX(Rn) be a rearrangement-invariant norm. Then the following statements are equivalent: (i) k · kX(Rn) satisfies the Lebesgue point property; (ii) k · kX(Rn) is locally absolutely continuous, and the Riesz-Wiener type inequality (1.11) (MXu)∗(s) ≤ Cku∗k⊘ X(0,s) for s ∈ (0, ∞), holds for some positive constant C, and for every u ∈ Xloc(Rn); (iii) k·kX(Rn ) is locally absolutely continuous, and the operator MX is of weak type from Xloc(Rn) into L1 loc(Rn). Remark 1.5. The local absolute continuity of the norm k·kX(Rn ) is an indispensable hypothesis in both conditions (ii) and (iii) of Theorem 1.4. Indeed, its necessity is already known from Theorem 1.1, and, on the other hand, it does not follow from the other assumptions in (ii) or (iii). For instance, both these assumptions are fulfilled by the rearrangement-invariant norm k · kL∞(Rn), which, however, is not locally absolutely continuous, and, in fact, does not satisfy the Lebesgue point property. [3, 2, 11, 12]. Remark 1.6. Riesz-Wiener type inequalities for special classes of rearrangement-invariant norms have been investigated in the literature -- see e.g. In particular, in [2] inequality (1.11) is shown to hold when k · kX(Rn) is an Orlicz norm k · kLA(Rn) associated with any Young function A. The case of Lorentz norms k · kLp,q(Rn) is treated in [3], where it is proved that (1.11) holds if, and only if, 1 ≤ q ≤ p. In fact, a different notion of maximal operator is considered in [3], which, however, is equivalent to (1.9) when k · kX(Rn) is a Lorentz norm, as is easily seen from [7, Equation (3.7)]. A simple sufficient condition for the validity of the Riesz-Wiener type inequality for very gen- eral maximal operators is proposed in [12]. In our framework, where maximal operators built upon rearrangement-invariant norms are taken into account, such condition turns out to be also necessary, as will be shown in Proposition 4.2. The approach introduced in [12] leads to alternative proofs of the Riesz-Wiener type inequality for Orlicz and Lorentz norms, and was also used in [14] to prove the validity of (1.11) for further families of rearrangement-invariant norms, including, in particular, all Lorentz endpoint norms k · kΛϕ(Rn). A kind of rearrangement inequality for the maximal operator built upon these Lorentz norms already appears in [11]. Results on weak type boundedness of the maximal operator MX are available in the literature as well [1, 8, 13, 15, 22]. For instance, in [22] it is pointed out that the operator MLp,q is of weak type from Lp,q(Rn) into Lp(Rn), if 1 ≤ q ≤ p, and hence, in particular, it is of weak type from Lp,q loc(Rn) into L1 loc(Rn). Our last main result provides us with necessary and sufficient conditions for the Lebesgue point property of a rearrangement-invariant norm which do not make explicit reference to the local absolute continuity of the relevant norm. Theorem 1.7. Let k·kX(Rn) be a rearrangement-invariant norm. Then the following statements are equivalent: (i) k · kX(Rn) satisfies the Lebesgue point property; (ii) For every function u ∈ X(Rn), supported in a set of finite measure, (iii) For every function u ∈ X(Rn), supported in a set of finite measure, Ln({x ∈ Rn : MXu(x) > 1}) < ∞; (MX u)∗(s) = 0. lim s→∞ NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 5 Theorems 1.1, 1.4 and 1.7 enable us to characterize the validity of the Lebesgue point property in customary classes of rearrangement-invariant norms. The following proposition deals with the case of standard Lorentz norms k · kLp,q(Rn). Proposition 1.8. The Lorentz norm k · kLp,q(Rn) satisfies the Lebesgue point property if, and only if, 1 ≤ q ≤ p < ∞. Since Lp,p(Rn) = Lp(Rn), Proposition 1.8 recovers, in particular, the standard result, men- tioned above, that the norm k · kLp(Rn) enjoys the Lebesgue point property if, and only if, 1 ≤ p < ∞. This fact is also reproduced by the following proposition, which concerns Orlicz norms k·kLA(Rn) built upon a Young function A. Proposition 1.9. The Orlicz norm k · kLA(Rn) satisfies the Lebesgue point property if, and only if, the Young function A satisfies the ∆2-condition near infinity. The last two results concern the so called Lorentz and Marcinkiewicz endpoint norms k · kΛϕ(Rn) and k · kMϕ(Rn), respectively, associated with a (non identically vanishing) concave function ϕ : [0, ∞) → [0, ∞). Proposition 1.10. The Lorentz norm k · kΛϕ(Rn) satisfies the Lebesgue point property if, and only if, ϕ(s) = 0. Proposition 1.11. The Marcinkiewicz norm k · kMϕ(Rn) satisfies the Lebesgue point property if, and only if, ϕ(s) > 0, namely, if and only if, (Mϕ)loc(Rn) = L1 loc(Rn). s When the present paper was almost in final form, it was pointed out to us by A. Gogatishvili that the Lebesgue point property of rearrangement-invariant spaces has also been investigated in [5, 6, 17, 19]. The analysis of those papers is however limited to the case of functions of one variable. Moreover, the characterizations of those norms having Lebesgue point property that are proved there are less explicit, and have a somewhat more technical nature. In this section we recall some definitions and basic properties of decreasing rearrangements and rearrangement-invariant function norms. For more details and proofs, we refer to [4, 16]. 2. Background Let E be a Lebesgue-measurable subset of Rn, n ≥ 1. The Riesz space of measurable functions +(E) = {u ∈ L0(E) : u ≥ 0 a.e. in E}, from E into [−∞, ∞] is denoted by L0(E). We also set L0 0(E) = {u ∈ L0(E) : u is finite a.e. in E}. The distribution function u∗ : [0, ∞) → [0, ∞] and L0 and the decreasing rearrangement u∗ : [0, ∞) → [0, ∞] of a function u ∈ L0(E) are defined by (2.1) for t ∈ [0, ∞), u∗(t) = Ln({y ∈ E : u(y) > t}) lim s→0+ lim s→0+ and by (2.2) (2.3) u∗(s) = inf{t ≥ 0 : u∗(t) ≤ s} for s ∈ [0, ∞), respectively. The Hardy-Littlewood inequality tells us that ZE u(y)v(y) dLn(y) ≤ Z ∞ 0 u∗(s)v∗(s) dL1(s) for every u, v ∈ L0(E). The function u∗∗ : (0, ∞) → [0, ∞], given by (2.4) u∗∗(s) = u∗(t)dL1(t) for s ∈ (0, ∞) , 1 s Z s 0 6 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A is non-increasing and satisfies u∗ ≤ u∗∗. Moreover, (2.5) for every u, v ∈ L0 +(E). (u + v)∗∗ ≤ u∗∗ + v∗∗ A rearrangement-invariant norm is a functional k · kX(E) : L0(E) → [0, ∞] such that (N1): ku + vkX(E) ≤ kukX(E) + kvkX(E) for all u, v ∈ L0 +(E); kλukX(E) = λ kukX(E) kukX(E) > 0 if u does not vanish a.e. in E; for all λ ∈ R, u ∈ L0(E); (N2): kukX(E) ≤ kvkX(E) whenever 0 ≤ u ≤ v a.e. in E; (N3): supk kukkX(E) = kukX(E) if {uk} ⊂ L0 (N4): kχGkX(E) < ∞ for every measurable set G ⊆ E, such that Ln(G) < ∞; (N5): for every measurable set G ⊆ E, with Ln(G) < ∞, there exists a positive constant +(E) with uk ր u a.e. in E; C(G) such that kukL1(G) ≤ C(G) kuχGkX(E) for all u ∈ L0(E); (N6): kukX(E) = kvkX(E) for all u, v ∈ L0(E) such that u∗ = v∗. The functional k · kX(E) is a norm in the standard sense when restricted to the set (2.6) X(E) = {u ∈ L0(E) : kukX(E) < ∞}. The latter is a Banach space endowed with such norm, and is called a rearrangement-invariant Banach function space, briefly, a rearrangement-invariant space. as Given a measurable subset E′ of E and a function u ∈ L0(E′), define the functionbu ∈ L0(E) Then the functional k · kX(E′) given by for u ∈ L0(E′) is a rearrangement-invariant norm. If Ln(E) < ∞, then if x ∈ E′, if x ∈ E \ E′. 0 bu(x) =(u(x) kukX(E′) = kbukX(E) (2.7) L∞(E) → X(E) → L1(E), where → stands for a continuous embedding. The local r.i. space Xloc(E) is defined as Xloc(E) = {u ∈ L0(E) : uχK ∈ X(E) for every bounded measurable set K ⊂ E}. The fundamental function of X(E) is defined by (2.8) ϕX(E)(s) = kχGkX(E) for s ∈ [0, Ln(E)), where G is any measurable subset of E such that Ln(G) = s. It is non-decreasing on [0, Ln(E)), ϕX(E)(0) = 0 and ϕX(E)(s)/s is non-increasing for s ∈ (0, Ln(E)). Hardy's Lemma tells us that, given u, v ∈ L0(E) and any rearrangement-invariant norm k·kX(E), (2.9) if u∗∗ ≤ v∗∗, then kukX(E) ≤ kvkX(E). The associate rearrangement-invariant norm of k · kX(E) is the rearrangement-invariant norm k · kX′(E) defined by (2.10) kvkX′(E) = sup(cid:8) ∫ E u(y)v(y) dLn(y) : u ∈ L0(E), kukX(E) ≤ 1(cid:9). NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 7 The corresponding rearrangement-invariant space X′(E) is called the associate space of X(E). The Holder type inequality (2.11) u(y)v(y) dLn(y) ≤ kukX(E)kvkX′(E) holds for every u ∈ X(E) and v ∈ X′(E). One has that X(E) = X′′(E). ZE The rearrangement-invariant norm, defined as kf kX(0,Ln(E)) = kukX′(E)≤1Z ∞ sup 0 f∗(s)u∗(s) dL1(s) for f ∈ L0(0, Ln(E)), is a representation norm for k · kX(E). It has the property that (2.12) kukX(E) = ku∗kX(0,Ln(E)) for every u ∈ X(E). For customary rearrangement-invariant norms, an expression for k · kX(0,Ln(E)) is immediately derived from that of k · kX(E). X(0, Ln(E)) as The dilation operator Dδ : X(0, Ln(E)) → X(0, Ln(E)) is defined for δ > 0 and f ∈ (Dδf )(s) =(f (sδ) 0 if sδ ∈ (0, Ln(E)), otherwise, and is bounded [4, Chap. 3, Proposition 5.11]. We shall make use of the subspace X 1(0, ∞) of X(0, ∞) defined as (2.14) X 1(0, ∞) = {f ∈ X(0, ∞) : f (s) = 0 for a.e. s > 1}. Now, assume that E is a measurable positive cone in Rn with vertex at 0, namely, a measurable set which is closed under multiplication by positive scalars. In what follows, we shall focus the nontrivial case when Ln(E) does not vanish, and hence Ln(E) = ∞. Let k · kX(E) be a rearrangement-invariant norm, and let G be a measurable subset of E such that 0 < Ln(G) < ∞. We define the functional k · k⊘ (2.13) (2.15) (2.16) for u ∈ L0(E). We call it the averaged norm of k · kX(E) on G, since X(G) as kuk⊘ X(G) =(cid:13)(cid:13)(uχG)( npLn(G) ·)(cid:13)(cid:13)X(E) Ln(G)(cid:1) X(cid:0)G, Ln X(G) = kuχGk kuk⊘ for u ∈ L0(E), where k · k k · kX(G), save that the Lebesgue measure Ln is replaced with the normalized Lebesgue measure Ln Ln(G) . Notice that (2.17) Ln(G)(cid:1) denotes the rearrangement-invariant norm, defined as X(G) = k(uχG)∗(Ln(G) ·)kX (0,∞) X(cid:0)G, Ln kuk⊘ for u ∈ L0(E). Moreover, (2.18) k1k⊘ X(G) is independent of G. The Holder type inequality for averaged norms takes the form: (2.19) u(y)v(y) dLn(y) ≤ kuk⊘ X(G) kvk⊘ X′(G) 1 Ln(G)ZG 8 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A for u, v ∈ L0(E). We conclude this section by recalling the definition of some customary, and less standard, instances of rearrangement-invariant function norms of use in our applications. In what follows, we set p′ = p p−1 for p ∈ (1, ∞), with the usual modifications when p = 1 and p = ∞. We also adopt the convention that 1/∞ = 0. Prototypal examples of rearrangement-invariant function norms are the classical Lebesgue norms. Indeed, kukLp(Rn) = ku∗kLp(0,∞), if p ∈ [1, ∞), and kukL∞ (Rn) = u∗(0). Then the functional defined as Let p, q ∈ [1, ∞]. Assume that either 1 < p < ∞ and 1 ≤ q ≤ ∞, or p = q = 1, or p = q = ∞. (2.20) kukLp,q(Rn) = ks 1 p− 1 q u∗(s)kLq(0,∞) for u ∈ L0(Rn), is equivalent (up to multiplicative constants) to a rearrangement-invariant norm. The corresponding rearrangement-invariant space is called a Lorentz space. Note that k · kLp,q(0,∞) is the representation norm for k · kLp,q(Rn), and Lp,p(Rn) = Lp(Rn). Moreover, Lp,q(Rn) → Lp,r(Rn) if 1 ≤ q < r ≤ ∞. Let A be a Young function, namely a left-continuous convex function from [0, ∞) into [0, ∞], which is neither identically equal to 0, nor to ∞. The Luxemburg rearrangement-invariant norm associated with A is defined as (2.21) kukLA(Rn) = inf(λ > 0 : ZRn A(cid:18) u(x) λ (cid:19) dLn(x) ≤ 1) for u ∈ L0(Rn). Its representation norm is kukLA(0,∞). The space LA(Rn) is called an Orlicz space. In particular, LA(Rn) = Lp(Rn) if A(t) = tp for p ∈ [1, ∞), and LA(Rn) = L∞(Rn) if A(t) = ∞χ(1,∞)(t). Recall that A is said to satisfy the ∆2 -- condition near infinity if it is finite valued and there exist constants C > 0 and t0 ≥ 0 such that (2.22) If A satisfies the ∆2 -- condition near infinity, and u ∈ LA(Rn) has support of finite measure, then for t ∈ [t0, ∞) . A(2t) ≤ CA(t) ZRn A(cu(x))dLn(x) < ∞ for every positive number c. A subclass of Young functions which is often considered in the literature is that of the so called N -functions. A Young function A is said to be an N -function if it is finite-valued, and A(t) t = 0 A(t) t lim t→∞ = ∞ . lim t→0+ Let ϕ : [0, ∞) → [0, ∞) be a concave function which does not vanish identically. Hence, in particular, ϕ is non-decreasing, and ϕ(t) > 0 for t ∈ (0, ∞). The Marcinkiewicz and Lorentz endpoint norm associated with ϕ are defined as (2.23) (2.24) kukMϕ(Rn) = sup u∗∗(s)ϕ(s), s∈(0,∞) kukΛϕ(Rn) = Z ∞ 0 u∗(s)dϕ(s), for u ∈ L0(Rn), respectively. The representation norms are k · kMϕ(0,∞) and k · kΛϕ(0,∞), respec- tively. The spaces Mϕ(Rn) and Λϕ(Rn) are called Marcinkiewicz endpoint space and Lorentz NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 9 endpoint space associated with ϕ. The fundamental functions of Mϕ(Rn) and Λϕ(Rn) coincide with ϕ. In fact, Mϕ(Rn) and Λϕ(Rn) are respectively the largest and the smallest rearrangement- invariant space whose fundamental function is ϕ, and this accounts for the expression "endpoint" which is usually attached to their names. Note the alternative expression kf kΛϕ(0,∞) = f∗(0)ϕ(0+) +Z ∞ 0 f∗(s)ϕ′(s)dL1(s), (2.25) for f ∈ L0(0, ∞), where ϕ(0+) = lim s→0+ ϕ(s). 3. A necessary condition: local absolute continuity In the present section we are mainly concerned with a proof of the following necessary condi- tions for a rearrangement-invariant norm to satisfy the Lebesgue point property. Proposition 3.1. If k · kX(Rn) is a rearrangement-invariant norm satisfying the Lebesgue point property, then: (i) k · kX(Rn) is locally absolutely continuous; (ii) X(Rn) is locally separable. The proof of Proposition 3.1 is split in two steps, which are the content of the next two lemmas. Lemma 3.2. Let k·kX(Rn) be a rearrangement-invariant norm which satisfies the Lebesgue point property. Then: (H) Given any function f ∈ X 1(0, ∞), any sequence {Ik} of pairwise disjoint intervals in (0, 1), and any sequence {ak} of positive numbers such that ak ≥ L1(Ik) and (3.1) one has that k(f χIk )∗k⊘ X(0,ak) > 1, ak < ∞. ∞Xk=1 Let us stress in advance that condition (H) is not only necessary, but also sufficient for a rearrangement-invariant norm to satisfy the Lebesgue point property. This is a consequence of Proposition 5.1, Section 5, and of the following lemma. Lemma 3.3. If a rearrangement-invariant norm k · kX(Rn) fulfills condition (H) of Lemma 3.2, then k · kX(Rn) is locally absolutely continuous. The proof of Lemma 3.2 in turn exploits the following property, which will also be of later use. Lemma 3.4. Let k · kX(Rn) be a rearrangement-invariant norm. Given any function f ∈ X(0, ∞), the function F : (0, ∞) → [0, ∞), defined as (3.2) F (r) = r kf∗k⊘ X(0,r) for r ∈ (0, ∞), is non-decreasing on (0, ∞), and the function F (r) r the function F is continuous on (0, ∞). is non-increasing on (0, ∞). In particular, 10 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A Proof. Let 0 < r1 < r2. An application of (2.10) tells us that F (r1) = r1k(f∗χ(0,r1))(r1 ·)kX(0,∞) = r1 g∗(s)f∗(r1s) dL1(s) kgkX′(0,∞)≤1 Z 1 sup 0 0 = sup kgkX′(0,∞)≤1 Z r1 kgkX′(0,∞)≤1 Z r2 = r2k(f∗χ(0,r2))(r2 ·)kX(0,∞) = F (r2). r1(cid:17) f∗(t) dL1(t) ≤ g∗(cid:16) t r2(cid:17) f∗(t) dL1(t) = r2 g∗(cid:16) t sup ≤ 0 sup kgkX′(0,∞)≤1 Z r2 r1(cid:17) f∗(t) dL1(t) g∗(cid:16) t kgkX′(0,∞)≤1 Z 1 g∗(s)f∗(r2s) dL1(s) sup 0 0 Namely, F is non-decreasing on (0, ∞). The fact that the function F (r) r (0, ∞) is a consequence of property (N2) and of the inequality is non-increasing on f∗(r1 ·)χ(0,r1)(r1 ·) ≥ f∗(r2 ·)χ(0,r2)(r2 ·) if 0 < r1 < r2. Hence, in particular, the function F is continuous on (0, ∞) (see e.g. [10, Chapter 2, p. 49]). (cid:3) Proof of Lemma 3.2. Assume that k · kX(Rn) satisfies the Lebesgue point property. Suppose, by contradiction, that condition (H) fails, namely, there exist a function f ∈ X 1(0, ∞), a sequence {Ik} of pairwise disjoint intervals in (0, 1) and a sequence {ak} of positive numbers, with ak ≥ L1(Ik), fulfilling (3.1) and such thatP∞k=1 ak = ∞. We may assume, without loss of generality, that (3.3) ak = 0. lim k→∞ (3.4) Indeed, if (3.3) fails, then the sequence {ak} can be replaced with another sequence, enjoying the same properties, and also (3.3). To verify this assertion, note that, if (3.3) does not hold, then there exist ε > 0 and a subsequence {akj } of {ak} such that akj ≥ ε for all j. Consider the sequence {bj}, defined as Equation (3.4) immediately tells us that bj ≥ L1(Ikj ), andP∞j=1 bj = ∞. Moreover, Lemma 3.4 and the inequality bj ≤ akj for j ∈ N ensure that (3.1) holds with ak and Ik replaced by bj and Ikj , respectively. Finally, j , L1(Ikj )(cid:9) for j ∈ N. bj = max(cid:8) ε bj = 0, sinceP∞j=1 L1(Ikj ) ≤ 1, and hence lim Moreover, by skipping, if necessary, a finite number of terms in the relevant sequences, we L1(Ikj ) = 0. lim j→∞ j→∞ j=0 aj) for each k ∈ N. We define the function g : may also assume that (3.5) L1(Ik) < 1 . ∞Xk=1 j=0 aj,Pk (f χIk)∗(cid:16)s − ∞Xk=1 k−1Xj=0 Now, set a0 = 0, and Jk = (Pk−1 (0, ∞) → [0, ∞) as g(s) = and the function u : Rn → [0, ∞) as aj(cid:17)χJk(s) for s ∈ (0, ∞), u(y) = sup k∈N g(y1 + k − 1) χ(0,1)n (y) for y = (y1, . . . , yn) ∈ Rn. NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 11 The function u belongs to X(Rn). To verify this fact, note that (3.6) for every t ≥ 0. Indeed, thanks to the equimeasurability of g and f χ∪k∈NIk , Ln({y ∈ Rn : u(y) > t}) ≤ L1({s ∈ (0, ∞) : f χ∪k∈NIk(s) > t}) Ln({y ∈ Rn : u(y) > t}) = L1({s ∈ (0, 1) : sup k∈N = L1(∪k∈N{s ∈ (0, 1) : g(s + k − 1) > t}) ≤ = L1({s ∈ (0, ∞) : g(s) > t}) = L1({s ∈ (0, ∞) : f χ∪k∈NIk(s) > t}). g(s + k − 1) > t}) ∞Xk=1 L1({s ∈ (0, 1) : g(s + k − 1) > t}) From (3.6) it follows that whence u ∈ X(Rn). Next, one has that kukX(Rn) ≤ kf χ∪k∈NIkkX(0,∞) ≤ kf kX(0,∞) < ∞ , (3.7) To prove (3.7), set lim sup r→0+ kuk⊘ X(Br (x)) > 0 for a.e. x ∈ (0, 1)n. Λk = {l ∈ N : Jl ⊆ [k − 1, k]} for k ∈ N. SinceP∞l=0 al = ∞ and lim l→∞ (3.8) al = 0, we have that (Jl − k + 1) = (0, 1), Jl = lim l→∞ lim k→∞ [l∈Λk where Jl denotes the closure of the open interval Jl. Equation (3.8) has to be interpreted in the following set-theoretic sense: fixed any x ∈ (0, 1)n, there exist k0 and an increasing sequence {lk}∞k=k0 in N such that lk ∈ Λk and x1 ∈ (Jlk − k + 1) for all k ∈ N greater than k0. Such a k0 can be chosen so that B√nalk (x) ⊆ (0, 1)n for all k ≥ k0, since lim k→∞ alk = 0. On the other hand, for every k ≥ k0, one also has nYi=1 Consequently, for every y ∈ Rn, B√nalk (x) ⊇ [xi − alk , xi + alk ] ⊇ (Jlk − k + 1) × [xi − alk , xi + alk ]. nYi=2 uχB√nalk (x)(y) ≥ g(y1 + k − 1) χ(Jlk−k+1)(y1) ,xi+alk ](yi) Therefore, thanks to the boundedness on rearrangement-invariant spaces of the dilation operator, defined as in (2.13), one gets kuk⊘ X(B√nalk (x)) = k(uχB√nalk ≥ C k(f χIlk )∗(Ln(B√nalk )∗(alk ·)kX (0,∞) = k(f χIlk (x)) ·)kX (0,∞) )∗k⊘ X(0,alk ) > 1 = (f χIlk )∗(cid:16)y1 + k − 1 − Hence, (3.9) (uχB√nalk (x))∗(s) ≥ (f χIlk χ[xi−alk nYi=2 aj(cid:17) χ(Jlk−k+1)(y1) k−1Xj=0 )∗(cid:0)(2alk )1−ns(cid:1) nYi=2 χ[xi−alk ,xi+alk ](yi). for s ∈ (0, ∞). 12 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A for some positive constant C = C(n). Hence inequality (3.7) follows, since lim k→∞ To conclude, consider the set M = {y ∈ (0, 1)n : u(y) = 0}. This set M has positive measure. alk = 0. Indeed, (3.6) with t = 0 and (3.5) imply Ln(M ) = 1 − Ln({y ∈ (0, 1)n : u(y) > 0}) ≥ 1 − L1({s ∈ (0, ∞) : f χ∪k∈NIk(s) > 0}) ≥ 1 − L1(Ik) > 0. ∞Xk=1 Then, estimate (3.7) tells us that lim sup r→0+ ku − u(x)k⊘ X(Br (x)) = lim sup r→0+ kuk⊘ X(Br (x)) > 0 for a.e. x ∈ M . This contradicts the Lebesgue point property for k · kX(Rn). (cid:3) Proof of Lemma 3.3. Let k · kX(Rn) be a rearrangement-invariant norm satisfying condition (H). We first prove that, if (H) is in force, then (3.10) lim t→0+ kg∗χ(0,t)kX(0,∞) = 0 for every g ∈ X 1(0, ∞). Arguing by contradiction, assume the existence of some g ∈ X 1(0, ∞) for which (3.10) fails. From property (N2) of rearrangement-invariant norms, this means that some ε > 0 exists such that kg∗χ(0,t)kX(0,∞) ≥ ε for every t ∈ (0, 1). Thanks to (N1), we may assume, without loss of generality, that ε = 2. Then, by induction, construct a decreasing sequence {bk}, with 0 < bk ≤ 1, such that (3.11) kg∗χ(bk+1,bk)kX(0,∞) > 1 for every k ∈ N. To this purpose, set b1 = 1, and assume that bk is given for some k ∈ N. Then define hl = g∗χ(cid:0) bk l ,bk(cid:1) for l ∈ N, with l ≥ 2. Since 0 ≤ hl ր g∗χ(0,bk), property (N3) tells us that khlkX(0,∞) ր kg∗χ(0,bk)kX(0,∞). Inasmuch as kg∗χ(0,bk)kX(0,∞) ≥ 2, then there exists an l0, with l0 ≥ 2, such that khl0kX(0,∞) > 1. Defining bk+1 = bk l0 Observe that choosing f = g∗χ(0,1), and ak = 1, Ik = (bk+1, bk) for each k ∈ N provides a contradiction to assumption (H). Indeed, inequality (3.1), which agrees with (3.11) in this case, entails that 0 < bk+1 < bk and kg∗χ(bk+1,bk)kX(0,∞) = khl0 kX(0,∞) > 1, as desired. holds for every k, whereasP∞k=1 ak = ∞. Consequently, (3.10) does hold. Now, take any u ∈ Xloc(Rn) and any non-increasing sequence {Kj} of measurable bounded Ln(Kj) = 0. We may assume sets in Rn such that ∩j∈NKj = ∅. Clearly, uχK1 ∈ X(Rn) and lim j→∞ that Ln(K1) < 1, whence (uχKj )∗ ∈ X 1(0, ∞) for each j ∈ N. From (3.10) it follows that lim j→∞ kuχKj kX(Rn) = lim j→∞ k(uχKj )∗kX(0,∞) ≤ lim j→∞ k(uχK1)∗χ(0,Ln(Kj))kX(0,∞) = 0, namely the local absolute continuity of k · kX(Rn). (cid:3) Proof of Proposition 3.1. Owing to [4, Corollary 5.6, Chap. 1], assertions (i) and (ii) are equiv- alent. Assertion (i) follows from Lemmas 3.2 and 3.3. (cid:3) NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 13 4. The functional GX and the operator MX This section is devoted to a closer analysis of the functional GX and the operator MX asso- ciated with a rearrangement-invariant norm k · kX(Rn). We begin with alternate characterizations of the almost concavity of the functional GX. In what follows, we shall make use of the fact that (4.1) khkX(0,∞) = kh∗kX(0,∞) = k(h∗)∗kX(0,∞) = GX(h∗) for every h ∈ L0(0, ∞). Moreover, by a partition of the interval (0, 1) we shall mean a finite collection {Ik : k = 1, . . . , m}, where Ik = (τk−1, τk) with 0 = τ0 < τ1 < · · · < τm = 1. Proposition 4.1. Let k · kX(Rn) be a rearrangement-invariant norm. Then the following condi- tions are equivalent: (i) the functional GX is almost concave; (ii) a positive constant C exists such that (4.2) mXk=1 L1(Ik) kf k⊘ X(Ik) ≤ Ckf kX(0,∞) for every f ∈ X 1(0, ∞), and for every partition {Ik : k = 1, . . . , m} of (0, 1); (iii) a positive constant C exists such that (4.3) mXk=1 Ln(Bk) kuk⊘ X(Bk ) ≤ C Ln(cid:0) ∪m k=1 Bk(cid:1) kuk⊘ X(∪m k=1Bk) for every u ∈ Xloc(Rn), and for every finite collection {Bk : k = 1, . . . , m} of pairwise disjoint balls in Rn. Proof. (i) ⇒ (ii) Assume that GX is almost concave. Fix any function f ∈ X 1(0, ∞), and any partition {Ik : k = 1, . . . , m} of (0, 1). It is easily verified that (cid:16)(f χIk )∗(L1(Ik)·)(cid:17)∗ = (f χIk)∗ L1(Ik) and (cid:0)f χ∪m k=1Ik(cid:1)∗ mXk=1 = (f χIk)∗ . mXk=1 L1(Ik) GX(cid:16) (f χIk)∗ L1(Ik)(cid:17) Hence, by (4.1) and the almost concavity of GX , there exists a constant C such that L1(Ik) kf k⊘ X(Ik) = mXk=1 L1(Ik) k(f χIk )∗(L1(Ik)·)kX (0,∞) = (f χIk)∗(cid:17) = C GX(cid:16)(f χ∪m mXk=1 ≤ C GX(cid:16) mXk=1 k=1 λk = 1. For each k = 1, . . . , m, write fk = (gk)∗, ak = Pk k=1Ik )∗(cid:17) ≤ C GX(f∗) = C kf kX(0,∞). This yields inequality (4.2). (ii) ⇒ (i) Take any finite collections {gk : k = 1, . . . , m} in C and {λk : k = 1, . . . , m} in (0, 1), i=1 λi, and respectively, with Pm Ik = (ak−1, ak) with a0 = 0. Then define f (t) = fk( t−ak−1 λk )χIk (t) for t ∈ (0, ∞). mXk=1 14 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A k=1 λkgk. Owing to (4.1) and (4.2), one thus obtains Observe that f∗ =Pm k=1 λk(fk)∗ =Pm GX(cid:16) mXk=1 λkgk(cid:17) = GX (f∗) = kf kX(0,∞) ≥ λkkf∗k kX(0,∞) = 1 C = mXk=1 1 C 1 C mXk=1 mXk=1 λkkf k⊘ X(Ik) = 1 C λkGX(cid:0)(fk)∗(cid:1) = mXk=1 mXk=1 1 C λkkf∗k ( 1 λk ·)k⊘ X (0,λk) λkGX (gk), whence the almost concavity of GX follows. (ii) ⇒ (iii) Fix any function u ∈ Xloc(Rn), and any finite collection {Bk : k = 1, . . . , m} of pairwise disjoint balls in Rn. Set ak = Ln(Bk), for k = 1, . . . , m, and a0 = 0. Define aj · − k−1Xj=0 aj(cid:1) χIk(cid:13)(cid:13)(cid:13)X(0,∞) 1 C Ln(∪m k=1Bk) Ln(Bk)kuk⊘ X(Bk ). mXk=1 Thanks to rearrangement-invariance of X(0, ∞), assumption (ii) ensures that kuk⊘ X(∪m for k = 1, . . . , m. j=0 aj j=0 aj j=1 aj Ik = Pk−1 j=1 aj! ,Pk Pm Pm k=1Bk) =(cid:13)(cid:13)(cid:13)(cid:0)uχ∪m ak ·(cid:1)(cid:13)(cid:13)(cid:13)X(0,∞) k=1Bk(cid:1)∗(cid:0) mXk=1 j=1 aj(cid:13)(cid:13)(cid:13)(uχBk )∗(cid:0) mXk=1 mXk=1 akPm mXk=1 CPm mXk=1 (uχBk )∗(cid:0) mXj=1 aj(cid:1) χIk(cid:13)(cid:13)(cid:13) k−1Xj=0 ak k(uχBk )∗(ak ·)kX(0,∞) = =(cid:13)(cid:13)(cid:13) 1 j=1 aj ak · − 1 C ≥ = ⊘ X(Ik) Inequality (4.3) is thus established. (iii) ⇒ (ii) Assume that f ∈ X 1(0, ∞), and that {Ik : k = 1, . . . , m} is a partition of (0, 1). Let {Bk : k = 1, . . . , m} be a family of pairwise disjoint balls in Rn such that Ln(Bk) = L1(Ik), k=1Bk and fulfilling (uχBk )∗ = and let u be a measurable function on Rn vanishing outside of ∪m (f χIk)∗, for k = 1, 2, . . . , m. Assumption (iii) then tells us that mXk=1 L1(Ik)kf k⊘ X(Ik) ≤ C L1(∪m k=1Ik)kf k⊘ k=1Ik) = C kf k⊘ X(∪m X(0,1) = C kf kX(0,∞), namely (4.2). (cid:3) We next focus on the maximal operator MX . Criteria for the validity of the Riesz-Wiener type inequality (1.11) are the content of the following result. Proposition 4.2. Let k · kX(Rn) be a rearrangement-invariant norm. Then the following condi- tions are equivalent: (i) the Riesz-Wiener type inequality (1.11) holds for some positive constant C, and for every u ∈ Xloc(Rn); (ii) a positive constant C1 exists such that (4.4) min k=1,...,m kuk⊘ X(Bk ) ≤ C1kuk⊘ X(∪m k=1Bk) for every u ∈ Xloc(Rn), and for every finite collection {Bk : k = 1, . . . , m} of pairwise disjoint balls in Rn; NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 15 (iii) a positive constant C2 exists such that kf k⊘ min k=1,...,m X(Ik) ≤ C2kf kX(0,∞) for every f ∈ X 1(0, ∞), and for every partition {Ik : k = 1, . . . , m} of (0, 1). Proof. (i) ⇒ (ii) Let u ∈ Xloc(Rn), and let {Bk : k = 1, . . . , m} be a collection of pairwise disjoint balls in Rn. When mink=1,...,m kuk⊘ X(Bk ) = 0, then (4.4) trivially holds. Assume that mink=1,...,m kuk⊘ If x ∈ Bj for some j = 1, . . . , m, then k=1Bk)(cid:1), and any t ∈(cid:0)0, mink=1,...,m kuk⊘ X(Bk ) > 0. Fix any s ∈(cid:0)0, Ln(∪m X(Bk )(cid:1). MX (uχ∪m k=1Bk )(x) ≥ kuχ∪m k=1Bkk⊘ X(Bj ) ≥ min k=1,...,m kuk⊘ X(Bk ) > t. Thus, and, consequently, (4.5) Since the last inequality holds for every t < mink=1,...,m kuk⊘ Ln({x ∈ Rn : MX (uχ∪m k=1Bk )(x) > t}) ≥ Ln(∪m k=1Bk) > s, k=1Bk )(cid:1)∗(s) ≥ t. (cid:0)MX(uχ∪m k=1Bk )(cid:1)∗(s) ≥ min k=1,...,m (cid:0)MX(uχ∪m X(Bk ), one infers that kuk⊘ X(Bk ). On the other hand, an application of assumption (i) with u replaced by uχ∪m (4.6) (cid:0)MX(uχ∪m k=1Bk )(cid:1)∗(s) ≤ C k(uχ∪m Coupling (4.5) with (4.6) implies that k=1Bk )∗k⊘ X(0,s) k=1Bk tells us that for s ∈(cid:0)0, Ln(∪m for s ∈(cid:0)0, Ln(∪m k=1Bk)(cid:1). k=1Bk)(cid:1). k=1Bk )∗k⊘ X(0,s) , which is guaranteed by min k=1,...,m kuk⊘ X(Bk) ≤ C k(uχ∪m k=1Bk )∗k⊘ X(0,s) Thus, owing to the continuity of the function s 7→ k(uχ∪m Lemma 3.4, we deduce that min k=1,...,m kuk⊘ X(Bk ) ≤ Ck(uχ∪m k=1Bk )∗k⊘ X(0,Ln(∪m k=1Bk)) = Ckuk⊘ k=1Bk), X(∪m namely (4.4). (ii) ⇒ (i) By [14, Proposition 3.2], condition (ii) implies the existence of a constant C′ such that (4.7) for every u ∈ Xloc(Rn). By the boundedness of the dilation operator on rearrangement-invariant spaces, there exists a constant C′′, independent of u, such that (4.8) (MX u)∗(s) ≤ C′ku∗(cid:0)3−ns ·(cid:1)χ(0,1)( · )kX (0,∞) ku∗(cid:0)3−ns ·(cid:1)χ(0,1)( · )kX(0,∞) ≤ C′′ku∗(s ·)χ(0,1)(3n · )kX(0,∞) for s ∈ (0, ∞), for s ∈ (0, ∞). ≤ C′′ku∗(s ·)χ(0,1)( · )kX(0,∞) = C′′ku∗k⊘ X(0,s) Inequality (1.11) follows from (4.7) and (4.8). (ii) ⇔ (iii) The proof is completely analogous to that of the equivalence between conditions (ii) and (iii) in Proposition 4.1. We omit the details for brevity. (cid:3) Condition (H) introduced in Lemma 3.2 can be characterized in terms of the maximal operator MX as follows. Proposition 4.3. Let k · kX(Rn) be a rearrangement-invariant norm. Then the following asser- tions are equivalent: 16 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A (i) k · kX(Rn) fulfils condition (H) in Lemma 3.2; (ii) For every function u ∈ X(Rn), supported in a set of finite measure, Ln({x ∈ Rn : MX u(x) > 1}) < ∞. Proof. (i) ⇒ (ii) Let u ∈ X(Rn) be supported in a set of finite measure. Set E = {x ∈ Rn : MX u(x) > 1}. According to (1.9), for any y ∈ E, there exists a ball By in Rn such that y ∈ By and kuk⊘ X(By ) > 1. Define E1 =ny ∈ E : Ln(By) > max{1, Ln({u > 0})}o. (4.9) (4.10) We claim that, if y ∈ E1, then for s ∈ (0, ∞). Indeed, since Ln(By) ≥ 1, by the monotonicity of the decreasing rearrangement (uχBy )∗(cid:0)Ln(By)s(cid:1) ≤ (uχBy )∗(s)χ(cid:0)0, Ln({u>0}) Ln(By ) (cid:1)(s) (uχBy )∗(s) ≥ (uχBy )∗(cid:0)Ln(By)s(cid:1) Thus, inequality (4.10) certainly holds if s ∈ (cid:0)0, Ln({u>0}) (cid:3). On the other hand, Ln(By) ≥ Ln({u > 0}), and since (uχBy )∗(s) = 0 for s ≥ Ln({u > 0}), we have that (uχBy )∗(cid:0)Ln(By)s(cid:1) = 0 if s ∈(cid:0)Ln({u>0}) , ∞(cid:1). Thereby, inequality (4.10) also holds for these values of s. Owing to X(By ) = k(uχBy )∗(Ln(By) ·)kX(0,∞) ≤(cid:13)(cid:13)(cid:13)(uχBy )∗χ(cid:0)0, Ln({u>0}) Since (i) is in force, equation (3.10) holds with g = u∗χ(0,1) ∈ X 1(0, ∞), namely, Ln(By ) (cid:1)(cid:13)(cid:13)(cid:13)X(0,∞) ≤(cid:13)(cid:13)(cid:13)u∗χ(cid:0)0, Ln({u>0}) Ln(By ) (cid:1)(cid:13)(cid:13)(cid:13)X(0,∞) for s ∈ (0, ∞). 1 < kuk⊘ Ln(By) (4.10), (4.11) Ln(By) . This implies the existence of some t0 ∈ (0, 1) such that ku∗χ(0,t)kX(0,∞) < 1 for every t ∈ (0, t0). Thus, (4.11) entails that lim t→0+ ku∗χ(0,t)kX(0,∞) = 0. Ln({u > 0}) Ln(By) ≥ t0 for every y ∈ E1. Hence, by (4.9), (4.12) Ln(By) ≤ maxn1, sup y∈E Ln({u > 0}) t0 o. An application of Vitali's covering lemma, in the form of [21, Lemma 1.6, Chap. 1], ensures that there exists a countable set I ⊆ E such that the family {By : y ∈ I} consists of pairwise disjoint balls, such that E ⊆ ∪y∈I 5By. Here, 5By denotes the ball, with the same center as By, whose radius is 5 times the radius of By. If I is finite, then trivially Ln(E) ≤ 5nPy∈I Ln(By) < ∞. Assume that, instead, I is infinite, and let {yk} be the sequence of its elements. For each k ∈ N, set, for simplicity, Bk = Byk , and (4.13) αk = Ln({y ∈ Bk : u(y) 6= 0}), Ik = Pk−1 i=0 αi α i=1 αi α ! , ,Pk ak = Ln(Bk) α , NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 17 where α = Ln({y ∈ Rn : u(y) 6= 0}) and α0 = 0. Note that {Ik} is a sequence of pairwise disjoint intervals in (0, 1), and ak ≥ L1(Ik) for each k ∈ N. The function f : (0, ∞) → [0, ∞), defined by belongs to X 1(0, ∞), and f (s) = for s ∈ (0, ∞), ∞Xk=1(cid:0)uχ{x∈Bk:u(x)6=0}(cid:1)∗(cid:16)αs − αi(cid:17)χIk (s) X(0,ak) = k(uχ{x∈Bk:u(x)6=0})∗(α ·)k⊘ k−1Xi=1 X(0,ak) k(f χIk )∗k⊘ = k(uχBk )∗(Ln(Bk) ·)kX (0,∞) = kuk⊘ X(Bk ) > 1. By (i), one thus obtains thatP∞k=1 Ln(Bk) = αP∞k=1 ak < ∞. Hence Ln(E) ≤ 5nP∞k=1 Ln(Bk) < ∞, also in this case. (ii) ⇒ (i) Let f ∈ X 1(0, ∞), let {Ik} be any sequence of pairwise disjoint intervals in (0, 1), and let {ak} be a sequence of real numbers, such that ak ≥ L1(Ik), fulfilling (3.1). Consider any sequence {Bk} of pairwise disjoint balls in Rn, such that Ln(Bk) = ak for k ∈ N. For each k ∈ N, choose a function gk : Rn → [0, ∞), supported in Bk, and such that gk is equimeasurable with f χIk. Then, define u = P∞k=1 gk. Note that u ∈ X(Rn), since u∗ = (f χS∞k=1 Ik )∗ ≤ f∗. Furthermore, u is supported in a set of finite measure. Thus, assumption (ii) implies that (4.14) Ln({x ∈ Rn : MXu(x) > 1}) < ∞. If x ∈ Bk for some k ∈ N, then MXu(x) ≥ kuk⊘ (4.15) X(Bk ) = kgkk⊘ X(Bk ) = kg∗kk⊘ X(0,Ln(Bk)) = k(f χIk )∗k⊘ X(0,ak) > 1. Consequently, and ∞Xk=1 ∪∞k=1Bk ⊆ {x ∈ Rn : MXu(x) > 1} ak = Ln(∪∞k=1Bk) ≤ Ln({x ∈ Rn : MX u(x) > 1}) < ∞. Condition (i) is thus fulfilled. (cid:3) 5. Proofs of Theorems 1.1, 1.4 and 1.7 The core of Theorems 1.1, 1.4 and 1.7 is contained in the following statement. Proposition 5.1. Given a rearrangement-invariant norm k · kX(Rn), consider the following properties: (i) k · kX(Rn) satisfies the Lebesgue point property; (ii) k · kX(Rn) fulfills condition (H) of Lemma 3.2; (iii) The functional GX is almost concave; (iv) The Riesz-Wiener type inequality (1.11) holds for some positive constant C, and for every u ∈ Xloc(Rn); (v) The operator MX is of weak type from Xloc(Rn) into L1 loc(Rn). Then: If, in addition, k · kX(Rn) is locally absolutely continuous, then (i) ⇒ (ii) ⇒ (iii) ⇒ (iv) ⇒ (v). 18 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A A proof of Proposition 5.1 requires the next lemma. (v) ⇒ (i). Lemma 5.2. Let k · kX(Rn) be a rearrangement-invariant norm such that (5.1) ϕX(Rn)(s) = 0. lim s→0+ If u : Rn → R is any simple function, then ku − u(x)k⊘ X(Br (x)) = 0 for a.e. x ∈ Rn. lim r→0+ Proof. Let E be any measurable subset of Rn. By the Lebesgue density theorem, (5.2) Ln(Br(x) \ E) Ln(Br(x)) Ln(Br(x) ∩ E) Ln(Br(x)) lim r→0+ lim r→0+ = 0 for a.e. x ∈ E = 0 for a.e. x ∈ Rn \ E. Since (5.3) (5.4) lim lim lim r→0+ kχE − χE(x)k⊘ it follows from (5.1) that X(Br (x)) = r→0+(cid:13)(cid:13)(cid:13)χ(cid:16)0, Ln(Br (x)\E) r→0+(cid:13)(cid:13)(cid:13)χ(cid:16)0, Ln(Br (x)∩E) Hence, if u is any simple function having the form u =Pk disjoint measurable subsets of Rn, and a1, . . . , ak ∈ R, then kχE − χE(x)k⊘ Ln(Br (x)) (cid:17)(cid:13)(cid:13)(cid:13)X(0,∞) Ln(Br (x)) (cid:17)(cid:13)(cid:13)(cid:13)X(0,∞) lim r→0+ X(Br (x)) = 0 for a.e. x ∈ Rn. for a.e. x ∈ E, for a.e. x ∈ Rn \ E, i=1 aiχEi, where E1, . . . , Ek are pairwise (5.5) lim r→0+ ku − u(x)k⊘ X(Br (x)) ≤ lim r→0+ for a.e. x ∈ Rn. ai kχEi − χEi(x)k⊘ X(Br (x)) = 0 (cid:3) kXi=1 Proof of Proposition 5.1. (i) ⇒ (ii) This is just the content of Lemma 3.2 above. (ii) ⇒ (iii) We prove this implication by contradiction. Assume that the functional GX is not almost concave. Owing to Proposition 4.1, this amounts to assuming that, for every k ∈ N, there exist a function fk ∈ X 1(0, ∞) and a partition {Jk,l : l = 1, . . . , mk} of (0, 1) such that (5.6) L1(Jk,l) kfkk⊘ X(Jk,l) > 4kkfkkX(0,∞). Define the function f : (0, ∞) → R as mkXl=1 (5.7) f (t) = for t ∈ (0, ∞). ∞Xk=1 χ(2−k,2−k+1)(t) fk(2kt − 1) 2k(cid:13)(cid:13)χ(2−k,2−k+1) fk(2k · −1)(cid:13)(cid:13)X(0,∞) Since f ∈ L0(0, ∞), f = 0 on (1, ∞) and kf kX(0,∞) ≤P∞k=1 2−k, we have that f ∈ X 1(0, ∞). Let us denote by Λ the set {(k, l) ∈ N2 : l ≤ mk}, ordered according to the lexicographic order, and define the sequence {Ik,l} as (5.8) Ik,l = 1 2k Jk,l + 1 2k for (k, l) ∈ Λ. NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 19 Each element Ik,l is an open subinterval of (0, 1). Moreover, the intervals Ik,l and Ih,j are disjoint if (k, l) 6= (h, j). Actually, if k 6= h, then Ik,l ∩ Ih,j ⊆ (2−k, 21−k) ∩ (2−h, 21−h) = ∅; if, instead, k = h but l 6= j, then the same conclusion immediately follows from the fact that the intervals Jk,l and Jk,j are disjoint. Owing to (5.6) and (5.7), (5.9) L1(Ik,l) kf k⊘ X(k,l)∈Λ L1(Ik,l) k(f χIk,l)∗(L1(Ik,l) ·)kX(0,∞) L1(Jk,l) 2k k(fkχJk,l)∗(L1(Jk,l) ·)kX (0,∞) 2kkχ(2−k,21−k)fk(2k · −1)kX(0,∞) X(Ik,l) = X(k,l)∈Λ = X(k,l)∈Λ ∞Xk=1 ∞Xk=1 ≥ = 1 4kkf∗k (2k ·)kX(0,∞) 4kkfkkX(0,∞) mkXl=1 ∞Xk=1 ≥ 4kkf∗k (2k ·)kX(0,∞) ≤ 2}, and observe that X(Ik,l) ≤ 2 X(k,l)∈M L1(Ik,l) kf k⊘ X(Ik,l) X(k,l)∈M L1(Jk,l) kfkk⊘ X(Jk,l) 4kkfkkX(0,∞) 4kkf∗k kX(0,∞) = ∞Xk=1 1 = ∞. L1(Ik,l) ≤ 2. Set M = {(k, l) ∈ Λ : kf k⊘ (5.10) From (5.9) and (5.10) we thus infer that X(k,l)∈Λ\M L1(Ik,l) kf k⊘ X(Ik,l) = ∞. On the other hand, assumption (ii) implies property (3.10). This property, applied with g = f χIk,l, in turn ensures that, for every (k, l) ∈ Λ, lim t→+∞ k(f χIk,l)∗k⊘ X(0,t) = lim t→+∞ k(f χIk,l)∗(t ·)kX (0,∞) ≤ lim t→+∞ k(f χIk,l)∗ χ(0, 1 t )kX(0,∞) = 0. Note that the inequality holds since the function (f χIk,l)∗ belongs to X 1(0, ∞), and is non- increasing, and hence (f χIk,l)∗(ts) ≤ (f χIk,l)∗(s) χ(0, 1 Thus, owing to Lemma 3.4, if (k, l) ∈ Λ \ M , there exists a number ak,l ≥ L1(Ik,l) such that t )(s) for s ∈ (0, ∞). (5.11) Furthermore, by (5.11), k(f χIk,l)∗k⊘ X(0,ak,l) = 2 . (5.12) X(k,l)∈Λ\M ak,l = = 1 2 X(k,l)∈Λ\M 2 X(k,l)∈Λ\M 1 ak,l k(f χIk,l)∗k⊘ X(0,ak,l) ≥ L1(Ik,l) kf k⊘ X(Ik,l) = ∞. 1 2 X(k,l)∈Λ\M L1(Ik,l) k(f χIk,l)∗k⊘ X(0,L1(Ik,l)) Thanks to (5.12), the function f ∈ X 1(0, ∞), defined by (5.7), the sequence {Ik,l}, defined by (5.8), and the sequence {ak,l} contradict condition (H) in Lemma 3.2, and, thus, assumption (ii). (iii) ⇒ (iv) This implication follows from Propositions 4.1 and 4.2, since condition (ii) of Propo- sition 4.1 trivially implies condition (iii) of Proposition 4.2. 20 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A (iv) ⇒ (v) Let K be a bounded subset of Rn. Fix any function u ∈ Xloc(Rn) whose support is contained in K. Clearly, u = uχK. From (iv), we infer that sup t>0 t Ln({x ∈ K : MX u(x) > t}) = sup t>0 t Ln({x ∈ Rn : χKMXu(x) > t}) = sup s>0 s(cid:0)χK MX u(cid:1)∗(s) ≤ s(cid:0)MXu(cid:1)∗(s) ≤ C X(0,Ln(K)) ≤ C′ku∗kX(0,∞) = C′kukX(Rn), s∈(0,Ln(K)) sup ≤ C Ln(K) ku∗k⊘ sup s∈(0,Ln(K)) sku∗k⊘ X(0,s) for some constants C and C′, where the last but one inequality follows from Lemma 3.4, and the last one from the boundedness of the dilation operator on rearrangement-invariant spaces. Property (v) is thus established. Finally, assume that k · kX(Rn) is locally absolutely continuous and satisfies condition (v). Since Rn is the countable union of balls, in order to prove (i) it suffices to show that, given any u ∈ Xloc(Rn) and any ball B in Rn, Equation (5.13) will in turn follow if we show that, for every t > 0, the set ku − u(x)k⊘ X(Br (x)) = 0 for a.e. x ∈ B. lim r→0+ (5.13) (5.14) At = {x ∈ B : lim sup r→0+ ku − u(x)k⊘ X(Br (x)) > 2t} has measure zero. To prove this, we begin by observing that, since k · kX(Rn) is locally absolutely continuous, [4, Theorem 3.11, Chap. 1] ensures that for any ε > 0 there exists a simple function vε supported on B such that uχB = vε + wε and kwεkX(B) < ε. Clearly, wε is supported on B as well. Moreover, [4, Theorem 5.5, Part (b), Chap. 2] and Lemma 5.2 imply that kvε − vε(x)k⊘ X(Br (x)) = 0 for a.e. x ∈ B. lim r→0+ Fix any ε > 0. Then ku − u(x)k⊘ lim sup r→0+ X(Br (x)) ≤ lim sup r→0+ = lim sup r→0+ kvε − vε(x)k⊘ kwε − wε(x)k⊘ X(Br (x)) + lim sup r→0+ X(Br (x)) ≤ MXwε(x) + wε(x) kχ(0,1)kX(0,∞). kwε − wε(x)k⊘ X(Br (x)) Therefore, (5.15) At ⊆ {x ∈ B : MXwε(x) > t} ∪ {y ∈ B : wε(y) kχ(0,1)kX(0,∞) > t} Owing to (v), for t ∈ (0, ∞). Ln({x ∈ B : MXwε(x) > t}) ≤ C t kwεkX(B) for t ∈ (0, ∞). On the other hand, Ln({y ∈ B : wε(y)kχ(0,1)kX(0,∞) > t}) ≤ kχ(0,1)kX(0,∞)kwεkX(B) for every t ∈ (0, ∞), where C0 is the norm of the embedding X(B) → L1(B). Inasmuch as kwεkX(B) < ε, the last two inequalities, combined with (5.15) and with the subadditivity of the outer Lebesgue measure, imply that the outer Lebesgue measure of At does not exceed kχ(0,1)kX(0,∞) kwεkL1(B) ≤ C0 t 1 t for every t ∈ (0, ∞). Hence, Ln(At) = 0, thanks to the arbitrariness of ε > 0. ε t(cid:0)C + C0kχ(0,1)kX(0,∞)(cid:1) (cid:3) NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM Proof of Theorem 1.1. This is a consequence of Propositions 3.1 and 5.1. Proof of Theorem 1.4. This is a consequence of Propositions 3.1 and 5.1. 21 (cid:3) (cid:3) Proof of Theorem 1.7. The equivalence of conditions (i) and (ii) follows from Proposition 4.3, Proposition 5.1 and Lemma 3.3. In order to verify the equivalence of (ii) and (iii), it suffices to observe that, thanks to the positive homogeneity of the maximal operator MX, one has that Ln({x ∈ Rn : MX u(x) > 1}) < ∞ for every u ∈ X(Rn) supported in a set of finite measure if, and only if, Ln({x ∈ Rn : MXu(x) > t}) < ∞ for every u ∈ X(Rn) supported in a set of finite measure and for every t ∈ (0, ∞). The latter condition is equivalent to (iii). (cid:3) 6. Proofs of Propositions 1.8 -- 1.11 In this last section, we show how our general criteria can be specialized to characterize those rearrangement-invariant norms, from customary families, which satisfy the Lebesgue point prop- erty, as stated in Propositions 1.8 -- 1.11. In fact, these propositions admit diverse proofs, based on the different criteria provided by Theorems 1.1, 1.4 and 1.7. For instance, Propositions 1.8 -- 1.10 can be derived via Theorem 1.4, combined with results on the local absolute continuity of the norms in question and on Riesz-Wiener type inequalities contained in [2] (Orlicz norms), [3] (norms in the Lorentz spaces Lp,q(Rn)), and [14] (norms in the Lorentz endpoint spaces Λϕ(Rn)). Let us also mention that, at least in the one-dimensional case, results from these propositions overlap with those of [5, 6, 19]. Hereafter, we provide alternative, more self-contained proofs of Propositions 1.8 -- 1.11, relying upon our general criteria. Let us begin with Proposition 1.8, whose proof requires the following preliminarily lemmas. Lemma 6.1. Let p, q ∈ [1, ∞] be admissible values in the definition of the Lorentz norm k · kLp,q(Rn). Then q p dL1(s)(cid:17) 1 q 1 p (cid:16)pR ∞0 sq−1(f (s)) s(f (s)) sup s∈(0,∞) L1({s ∈ (0, ∞) : f (s) > 0}) if 1 < p < ∞ and 1 ≤ q < ∞, or p = q = 1; if 1 < p < ∞ and q = ∞; if p = q = ∞ , (6.1) GLp,q (f ) =  for every non-increasing function f : [0, ∞) → [0, ∞]. Hence, the functional GLp,q is concave if 1 ≤ q ≤ p. Proof. Equation (6.1) follows from a well-known expression of Lorentz norms in terms of the distribution function (see, e.g., [9, Proposition 1.4.9]), from equality (4.1) and from the fact that every non-increasing function f : [0, ∞) → [0, ∞] agrees a.e. with the function f = (f∗)∗. The fact that GLp,q is concave if 1 ≤ q ≤ p is an easy consequence of the representation formulas (6.1). In particular, the fact that the function [0, ∞) ∋ t 7→ tα is concave if 0 < α ≤ 1 plays a role here. (cid:3) Lemma 6.2. Suppose that 1 ≤ p < q < ∞. Then there exists a function u ∈ Lp,q(Rn), having support of finite measure, such that (6.2) Ln({x ∈ Rn : MLp,q u(x) > 1}) = ∞. 22 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A Proof. We shall prove that the norm k · kLp,q(Rn) does not satisfy condition (H) from Lemma 3.2, if 1 ≤ p < q < ∞. The conclusion will then follow via Proposition 4.3. To this purpose, define f : (0, ∞) → [0, ∞) as p ∞Xk=1(cid:18) 3k k(cid:19) 1 χ (cid:18) 1 2·3k , 1 2·3k−1 (cid:19) (s) for s ∈ (0, ∞), (6.3) f (s) = c p(cid:1) 1 where c >(cid:0) q q . Observe that f = f∗χ(0,1) a.e., since f is a nonnegative decreasing function in (0, ∞) with support in (0, 1). Moreover, f ∈ Lp,q(0, ∞), since 1 s p q kf kq 2·3k−1 1 2·3k Lp,q(0,∞) = cq ∞Xk=1Z (cid:18) 3k p−1 dL1(s) ≤ cq ∞Xk=1(cid:18) 3k k(cid:19) q k(cid:19) q pZ thanks to the assumption that q > p. For each k ∈ N, set Ik =(cid:0) 1 2·3k , = c Z ∞ p−1 dL1(s)! 1 0 (cid:18) 3k k(cid:19) q k(f χIk )∗k⊘ k(cid:17) s 3k )(cid:16) s Lp,q (0,ak ) χ(0, 1 and hence condition (H) of Lemma 3.2 fails for the norm k · kLp,q(Rn). p q 1 2·3k−1 0 q p−1 dL1(s) < ∞ , s k . Then 1 2·3k−1(cid:1) and ak = 1 = c(cid:18) p q(cid:19) 1 > 1, q q (cid:3) We are now in a position to accomplish the proof of Proposition 1.8. Proof of Proposition 1.8. By Lemma 6.1, the functional GLp,q is concave if 1 ≤ q ≤ p. Moreover, the norm k · kLp,q(Rn) is locally absolutely continuous if and only if q < ∞ -- see e.g. [16, Theorem 8.5.1]. Thereby, an application of Theorem 1.1 tells us that, if 1 ≤ q ≤ p < ∞, then the norm k · kLp,q(Rn) has the Lebesgue point property. On the other hand, coupling Theorem 1.7 with Lemma 6.2 implies that the norm k · kLp,q(Rn) does not have the Lebesgue point property if 1 ≤ p < q < ∞. In the remaining case when q = ∞, the norm k · kLp,q(Rn) is not locally absolutely continuous. Hence, by Theorem 1.1, it does not have the Lebesgue point property. (cid:3) One proof of Proposition 1.9, dealing with Orlicz norms, will follow from Theorem 1.7, via the next lemma. Lemma 6.3. Let A be a Young function satisfying the ∆2-condition near infinity. Then for every u ∈ LA(Rn), supported in a set of finite measure. Ln({x ∈ Rn : MLAu(x) > 1}) < ∞ Proof. Owing to Proposition 4.3, it suffices to show that condition (H) from Lemma 3.2 is fulfilled by the Luxemburg norm. Consider any function f ∈ LA and any sequence {ak} of positive real numbers such that 1 (0, ∞), any sequence {Ik} of pairwise disjoint intervals in (0, 1), ak ≥ L1(Ik) and k(f χIk )∗k⊘ LA(0,ak) > 1 for k ∈ N. Since 0 for every k ∈ N, one has that A((f χIk )∗(s)) dL1(s) =ZIk ak <Z ak A(f (s)) dL1(s) ≤Z 1 ∞Xk=1ZIk ak < 0 ∞Xk=1 A(f (s)) dL1(s) A(f (s)) dL1(s) < ∞. NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 23 Notice that the last inequality holds owing to the assumption that A satisfies the ∆2-condition near infinity, and f has support of finite measure. Altogether, condition (H) is satisfied by the norm k · kLA(Rn). (cid:3) Proof of Proposition 1.9. If A satisfies the ∆2-condition near infinity, then the norm k · kLA(Rn) fulfills the Lebesgue point property, by Lemma 6.3 and Theorem 1.7. Conversely, assume that the norm k · kLA(Rn) fulfills the Lebesgue point property. Then it has to be locally absolutely continuous, by either Theorem 1.1 or Theorem 1.4. Owing to [18, Theorem 14 and Corollary 5, Section 3.4], this implies that A satisfies the ∆2-condition near infinity. (cid:3) In the next proposition, we point out the property, of independent interest, that the functional GLA is almost concave for any N -function A. Such a property, combined with the fact that the norm k · kLA(Rn) is locally absolutely continuous if and only if A satisfies the ∆2-condition near infinity, leads to an alternative proof of Proposition 1.9, at least when A is an N -function, via Theorem 1.1. Proposition 6.4. The functional GLA is almost concave for every N -function A. Proof. The norm k · kLA(Rn) is equivalent, up to multiplicative constants, to the norm k · kLA(Rn) defined as for u ∈ L0(Rn) -- see [18, Section 3.3, Proposition 4 and Theorem 13]. One has that kukLA(Rn) = inf(cid:26) 1 k(cid:18)1 +ZRn 1 k(cid:18)1 +ZRn A(ku(x)) dx(cid:19) = +Z ∞ A(ku(x)) dx(cid:19) : k > 0(cid:27) +Z ∞ A′(kt)f (t) dt(cid:27) A′(kt)u∗(t) dt, 1 k GLA(f ) = inf(cid:26) 1 k 0 where A′ denotes the left-continuous derivative of A. Altogether, we have that 0 for every non-increasing function f : [0, ∞) → [0, ∞). The functional GLA is concave, since it is the infimum of a family of linear functionals, and hence the functional GLA is almost concave. (cid:3) Let us next focus on the case of Lorentz endpoint norms, which is the object of Proposi- tion 1.10. Lemma 6.5. Assume that ϕ : [0, ∞) → [0, ∞) is a (non identically vanishing) concave function. Then (6.4) GΛϕ(f ) =Z L1({f >0}) 0 ϕ(f (t)) dL1(t) for every non-increasing function f : [0, ∞) → [0, ∞]. concave. In particular, the functional GΛϕ is Proof. Take any non-increasing function f : [0, ∞) → [0, ∞]. Set h = f∗, whence f = f∗ = (f∗)∗ = h∗ a.e., and h∗(0) = f∗(0) = L1({f > 0}). From equations (2.25) and (4.1), one has, via 24 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A Fubini's theorem, GΛϕ(f ) = GΛϕ(h∗) = khkΛϕ(0,∞) = h∗(0)ϕ(0+) +Z ∞ 0 h∗(s)ϕ′(s) dL1(s) dL1(t) ϕ′(s) dL1(s) ϕ′(s) dL1(s) dL1(t) 0 Z h∗(s) Z h∗(t) 0 0 0 = h∗(0)ϕ(0+) +Z ∞ = h∗(0)ϕ(0+) +Z h∗(0) = h∗(0)ϕ(0+) +Z h∗(0) =Z h∗(0) 0 0 [ϕ(h∗(t)) − ϕ(0+)] dL1(t) ϕ(h∗(t)) dL1(t) =Z L1({f >0}) 0 ϕ(f (t)) dL1(t). Hence, formula (6.4) follows. In order to verify the concavity of GΛϕ, fix any pair of non-increasing functions f, g : [0, ∞) → [0, ∞] and λ ∈ (0, 1). Observe that {t ∈ [0, ∞) : λf (t) + (1 − λ)g(t) > 0} = {t ∈ [0, ∞) : f (t) > 0} ∪ {t ∈ [0, ∞) : g(t) > 0}. The monotonicity of f and g ensures that the two sets on the right-hand side of the last equation are intervals whose left endpoint is 0. Consequently, (6.5) L1({λf + (1 − λ)g > 0}) = max{L1({f > 0}), L1({g > 0})}. On making use of equations (6.4) and (6.5), and of the concavity of ϕ, one infers that the functional GΛϕ is concave as well. (cid:3) Proof of Proposition 1.10. By Lemma 6.5, the functional GΛϕ is concave for every non identically vanishing concave function ϕ : [0, ∞) → [0, ∞). On the other hand, it is easily verified, via equation (2.25), that the norm k · kΛϕ(Rn) is locally absolutely continuous if, and only if, ϕ(0+) = 0. The conclusion thus follows from Theorem 1.1. (cid:3) We conclude with a proof of Proposition 1.11. s Proof of Proposition 1.11. Assume first that lims→0+ ϕ(s) = 0. Then we claim that the norm k·kMϕ(Rn) is not locally absolutely continuous, and hence, by either Theorem 1.1 or Theorem 1.4, it does not have the Lebesgue point property. To verify this claim, observe that the function (0, ∞) ∋ s 7→ s ϕ(s) is quasiconcave in the sense of [4, Definition 5.6, Chapter 2], and hence, by [4, Chapter 2, Proposition 5.10], there exists a concave function ψ : (0, ∞) → [0, ∞) such that 1 2 ψ(s) ≤ s ϕ(s) ≤ ψ(s) for s ∈ (0, ∞). Let ψ′ denote the right-continuous derivative of ψ, and define u(x) = ψ′(ωnxn) for x ∈ Rn, where ωn is the volume of the unit ball in Rn. Then u∗ = ψ′ in (0, ∞), so that 1 ≤ u∗∗(s)ϕ(s) ≤ 2 for s ∈ (0, ∞). The second inequality in the last equation ensures that u ∈ Mϕ(Rn), whereas the first one tells us that u does not have a locally absolutely continuous norm in Mϕ(Rn). Conversely, assume that lims→0+ on any given subset of Rn with finite measure (see e.g. k · kMϕ(Rn) has the Lebesgue point property, since k · kL1(Rn) has it. loc(Rn), with equivalent norms [20, Theorem 5.3]). Hence, the norm (cid:3) ϕ(s) > 0, then (Mϕ)loc(Rn) = L1 s NORMS SUPPORTING THE LEBESGUE DIFFERENTIATION THEOREM 25 References [1] N. Aıssaoui, Maximal operators, Lebesgue points and quasicontinuity in strongly nonlinear potential theory, Acta Math. Univ. Comenian. (N.S.) 71 (2002), no. 1, 35 -- 50. [2] R. J. Bagby & J. D. Parsons, Orlicz spaces and rearranged maximal functions, Math. Nachr. 132 (1987), 15 -- 27. [3] J. Bastero, M. Milman & F. J. Ruiz, Rearrangement of Hardy-Littlewood maximal functions in Lorentz spaces, Proc. Amer. Math. Soc 128 (2000), 65 -- 74. [4] C. Bennett & R.Sharpley, "Interpolation of operators", Academic Press, Boston, 1988. [5] M. S. Braverman & A. A. Sedaev, L. Kh. Poritskaya's characterization of symmetric spaces. (Russian) Qualitative and approximate methods for investigating operator equations (Russian), 21 -- 33, 167, Yaroslav. Gos. Univ., Yaroslavl', 1982. [6] M. S. Braverman & A. A. Sedaev, On a characteristic of symmetric spaces. (Russian) Funktsional. Anal. i Prilozhen, 15 (1981), no. 2, 65 -- 66. [7] P. Cavaliere & A. Cianchi, Classical and approximate Taylor expansions of weakly differentiable functions, Ann. Acad. Sci. Fenn. Math. 39 (2014), no. 2, 527 -- 544. [8] M. Ciesielski & A. Kaminska, Lebesgue's differentiation theorems in R.I. quasi-Banach spaces and Lorentz spaces Γp,w, J. Funct. Spaces Appl. (2012), Art. ID 682960, 28 pp. [9] L. Grafakos, "Classical Fourier Analysis, Second edition", Grad. Texts Math. Vol. 49 (Springer, Berlin, 2008). [10] S. G. Krein, Yu. I. Petunin & E. M. Semenov, "Interpolation of Linear Operators", American Mathemat- ical Society, Providence, R.I., 1982. [11] M. A. Leckband, A note on maximal operators and reversible weak type inequalities, Proc. Amer. Math. Soc 92 (1984), no. 1, 19 -- 26. [12] A. K. Lerner, A new approach to rearrangements of maximal operators, Bull. London Math. Soc. 37 (2005), no. 5, 771 -- 777. [13] F.E. Levis, Weak inequalities for maximal functions in Orlicz-Lorentz spaces and applications, J. Approx. Theory 162 (2010), no. 2, 239 -- 251. [14] M. Masty lo & C. P´erez, The Hardy-Littlewood maximal type operators between Banach function spaces, Indiana Univ. Math. J. 61 (2012), no. 3, 883 -- 900. [15] C. P´erez, Two weighted inequalities for potential and fractional type maximal operators, Indiana Univ. Math. J. 43 (1994), no. 2, 663 -- 683. [16] L. Pick, A. Kufner, O. John & S. Fuc´ık, "Function spaces. Vol. 1. Second revised and extended edition", de Gruyter & Co., Berlin, 2013. [17] L. H. Poritskaya, Lebesgue - Banach points of functions in symmetric spaces, Mat. Zametki 23 (1978), no. 4, 581 -- 592. [18] M. M. Rao & Z. D. Ren "Theory of Orlicz spaces", Marcel Dekker, Inc., New York, 1991. [19] D. V. Salehov, E. M. Semenov, On Lebesgue - Banach points (Russian), Uspehi Mat. Nauk 20 (1965), no. 6 (126), 151 -- 156. [20] L. Slav´ıkov´a, Almost-compact embeddings, Math. Nachr. 285 (2012), no. 11 -- 12, 1500 -- 1516. [21] E.M. Stein, "Singular integrals and differentiability properties of functions", Princeton University Press, Princeton, N.J. 1970. [22] E.M. Stein, Editor's note: The differentiability of functions in Rn, Ann. of Math. (2) 113 (1981), no. 2, 383 -- 385. Dipartimento di Matematica, Universit`a di Salerno, Via Giovanni Paolo II, 84084 Fisciano (SA), Italy E-mail address: [email protected] Dipartimento di Matematica e Informatica "U. Dini", Universit`a di Firenze, Viale Morgagni 57/A, 50134 Firenze, Italy E-mail address: [email protected] Department of Mathematical Analysis, Faculty of Mathematics and Physics, Charles Univer- sity, Sokolovsk´a 83, 186 75 Praha 8, Czech Republic E-mail address: [email protected] 26 PAOLA CAVALIERE, ANDREA CIANCHI, LUBOS PICK AND LENKA SLAV´IKOV ´A Department of Mathematical Analysis, Faculty of Mathematics and Physics, Charles Univer- sity, Sokolovsk´a 83, 186 75 Praha 8, Czech Republic E-mail address: [email protected]
1601.07298
1
1601
2016-01-27T09:25:52
The joint modulus of variation of metric space valued functions and pointwise selection principles
[ "math.FA" ]
Given $T\subset\mathbb{R}$ and a metric space $M$, we introduce a nondecreasing sequence of pseudometrics $\{\nu_n\}$ on $M^T$ (the set of all functions from $T$ into $M$), called the \emph{joint modulus of variation}. We prove that if two sequences of functions $\{f_j\}$ and $\{g_j\}$ from $M^T$ are such that $\{f_j\}$ is pointwise precompact, $\{g_j\}$ is pointwise convergent, and the limit superior of $\nu_n(f_j,g_j)$ as $j\to\infty$ is $o(n)$ as $n\to\infty$, then $\{f_j\}$ admits a pointwise convergent subsequence whose limit is a conditionally regulated function. We illustrate the sharpness of this result by examples (in particular, the assumption on the $\limsup$ is necessary for uniformly convergent sequences $\{f_j\}$ and $\{g_j\}$, and `almost necessary' when they converge pointwise) and show that most of the known Helly-type pointwise selection theorems are its particular cases.
math.FA
math
The joint modulus of variation of metric space valued functions and pointwise selection principles Vyacheslav V. Chistyakov∗,a, Svetlana A. Chistyakovaa aDepartment of Informatics, Mathematics and Computer Science, National Research University Higher School of Economics, Bol'shaya Pecherskaya Street 25/12, Nizhny Novgorod 603155, Russian Federation Abstract Given T ⊂ R and a metric space M, we introduce a nondecreasing sequence of pseudometrics {νn} on M T (the set of all functions from T into M), called the joint modulus of variation. We prove that if two sequences of functions {fj} and {gj} from M T are such that {fj} is pointwise precompact, {gj} is pointwise convergent, and the limit superior of νn(fj, gj) as j → ∞ is o(n) as n → ∞, then {fj} admits a pointwise convergent subsequence whose limit is a conditionally regulated function. We illustrate the sharpness of this result by examples (in particular, the assumption on the lim sup is necessary for uniformly convergent sequences {fj} and {gj}, and 'almost necessary' when they converge pointwise) and show that most of the known Helly-type pointwise selection theorems are its particular cases. joint modulus of variation, metric space, regulated function, Key words: pointwise convergence, selection principle, generalized variation MSC 2000: 26A45, 28A20, 54C35, 54E50 1. Introduction The purpose of this paper is to present a new sufficient condition (which is almost necessary) on a pointwise precompact sequence {fj} ≡ {fj}∞ j=1 of functions fj mapping a subset T of the real line R into a metric space (M, d), ∗Corresponding author. Email addresses: [email protected], [email protected] (Vyacheslav V. Chistyakov), [email protected] (Svetlana A. Chistyakova) Preprint submitted to arXiv (math.FA) September 19, 2018 under which the sequence admits a pointwise convergent subsequence. The historically first result in this direction is the classical Helly Selection Prin- ciple, in which the assumptions are as follows: T = [a, b] is a closed interval, M = R, and {fj} is uniformly bounded and consists of monotone functions ([29], [31, II.8.9 -- 10], [39, VIII.4.2], and [10, Theorem 1.3] if T ⊂ R is ar- bitrary). Since a real function on T of bounded (Jordan) variation is the difference of two nondecreasing bounded functions, Helly's theorem extends to uniformly bounded sequences of functions, whose Jordan's variations are uniformly bounded. Further generalizations of the latter pointwise selection principle are concerned with replacement of Jordan's variation by more gen- eral notions of variation ([2, 3], [6] -- [10], [15, 16, 22, 25, 26, 32, 33, 38, 41, 43]). In all these papers, the pointwise limit of the extracted subsequence of {fj} is a function of bounded generalized variation (in the corresponding sense), and so, it is a regulated function (with finite one-sided limits at all points of the domain). Note that pointwise selection principles (or the sequential compactness in the topology of pointwise convergence) and regulated func- tions are of importance in real analysis ([28, 31, 39]), stochastic analysis and generalized integration ([37]), optimization ([1, 36]), set-valued analysis ([2, 10, 19, 20, 30]), and other fields. A unified approach to the diverse selection principles mentioned above was proposed in [11, 12]. It is based on the notion of modulus of variation of a function introduced in [4, 5] (see also [28, 11.3.7]) and does not refer to the uniform boundedness of variations of any kind, and so, can be applied to sequences of non-regulated functions. However, the pointwise limit of the ex- tracted subsequence of {fj} is again a regulated function. In order to clarify this situation and expand the amount of sequences having pointwise conver- gent subsequences, we define the notion of the joint modulus of variation for metric space valued functions: this is a certain sequence of pseudometrics {νn} on the product set M T (of all functions from T into M). Making use of {νn}, we obtain a powerful pointwise selection principle (see Theorem 1 in Section 2). Putting gj = c for all j ∈ N, where c : T → M is a constant function, we get the selection principle from [11], which already contains all selection principles alluded to above as particular cases. In contrast to results from [11, 12], the pointwise limit f from Theorem 1 may not be regulated in general -- this depends on the limit function g, namely, since νn(f, g) = o(n), the function f is only conditionally regulated with respect to g (in short, g-regulated ). In particular, if g = c, then f is regulated in the usual sense. 2 Finally, we point out that following the ideas from [13], Theorem 1 may be extended to sequences of functions with values in a uniform space M. The paper is organized as follows. In Section 2, we present necessary def- initions and our main result, Theorem 1. In Section 3, we establish essential properties of the joint modulus of variation, which are needed in the proof of Theorem 1 in Section 5. Section 4 is devoted to the study of g-regulated (and, in particular, regulated) functions. In the final Section 6, we extend the Helly-type selection theorems from [25] and [22, 32] by exploiting Theorem 1. 2. Main result Let ∅ 6= T ⊂ R, (M, d) be a metric space with metric d, and M T denote the set of all functions f : T → M mapping T into M. The letter c stands for a constant function c : T → M (i.e., c(s) = c(t) in M for all s, t ∈ T ). The joint oscillation of two functions f, g ∈ M T is the quantity (f, g)(T ) = sup(cid:8)(f, g)({s, t}) : s, t ∈ T(cid:9) ∈ [0, ∞], where (f, g)({s, t}) = sup z∈M(cid:12)(cid:12)d(f (s), z) + d(g(t), z) − d(f (t), z) − d(g(s), z)(cid:12)(cid:12) (2.1) is the joint increment of f and g on the two-point set {s, t} ⊂ T , for which the following two inequalities hold: (f, g)({s, t}) ≤ d(f (s), f (t)) + d(g(s), g(t)), (f, g)({s, t}) ≤ d(f (s), g(s)) + d(f (t), g(t)). (2.2) (2.3) Since (f, c)({s, t}) = d(f (s), f (t)) (= the increment of f on {s, t} ⊂ T ) is independent of c, the quantity f (T ) = (f, c)(T ) is the usual oscillation of f on T , also known as the diameter of the image f (T ) = {f (t) : t ∈ T } ⊂ M. Clearly, by (2.2), (f, g)(T ) ≤ f (T ) + g(T ). We denote by B(T ; M) = {f ∈ M T : f (T ) < ∞} the family of all bounded functions on T equipped with the uniform metric d∞ given by d∞(f, g) = sup t∈T d(f (t), g(t)) for f, g ∈ B(T ; M) (d∞ is an extended metric on M T , i.e., may assume the value ∞). We have d∞(f, g) ≤ d(f (s), g(s)) + f (T ) + g(T ) for all s ∈ T and, by virtue of (2.3), (f, g)(T ) ≤ 2d∞(f, g). 3 If n ∈ N, we write {Ii}n 1 ≺ T to denote a collection of n two-point subsets Ii = {si, ti} of T (i = 1, . . . , n) such that s1 < t1 ≤ s2 < t2 ≤ · · · ≤ sn−1 < tn−1 ≤ sn < tn (so that the intervals [s1, t1], . . . , [sn, tn] with end-points in T are non-overlapping). We say that a collection {Ii}n 1 ≺ T with Ii = {si, ti} is a partition of T if (setting t0 = s1) si = ti−1 for all i = 1, . . . , n, which is written as {ti}n 0 ≺ T . The joint modulus of variation of two functions f, g ∈ M T is the sequence {νn(f, g)}∞ n=1 ⊂ [0, ∞] defined by νn(f, g) = sup(cid:26) n Xi=1 (f, g)(Ii) : {Ii}n 1 ≺ T(cid:27) for all n ∈ N, (2.4) where (f, g)(Ii) = (f, g)({si, ti}) is the quantity from (2.1) if Ii = {si, ti} (for finite T with the number of elements #(T ) ≥ 2, we make use of (2.4) for n ≤ #(T ) − 1, and set νn(f, g) = ν#(T )−1(f, g) for all n > #(T ) − 1). Note that, given f, g ∈ M T , we have ν1(f, g) = (f, g)(T ) and ν1(f, g) ≤ νn(f, g) ≤ nν1(f, g) for all n ∈ N. (2.5) Further properties of the joint modulus of variation are presented in Section 3. For a sequence of functions {fj} ⊂ M T and f ∈ M T , we write: (a) fj → f on T to denote the pointwise (or everywhere) convergence of {fj} to f (that is, limj→∞ d(fj(t), f (t)) = 0 for all t ∈ T ); (b) fj ⇒ f on T to denote the uni- form convergence of {fj} to f meaning, as usual, that limj→∞ d∞(fj, f ) = 0. The uniform convergence implies the pointwise convergence, but not vice versa. Recall that a sequence {fj} ⊂ M T is said to be pointwise precompact on T if the closure in M of the set {fj(t) : j ∈ N} is compact for all t ∈ T . n=1 ⊂ R, we Making use of E. Landau's notation, given a sequence, {µn}∞ write µn = o(n) to denote the condition limn→∞ µn/n = 0. Our main result, a pointwise selection principle for metric space valued functions in terms of the joint modulus of variation, is as follows. Theorem 1. Let ∅ 6= T ⊂ R and (M, d) be a metric space. Suppose {fj}, {gj} ⊂ M T are two sequences of functions such that (a) {fj} is pointwise precompact on T , (b) {gj} is pointwise convergent on T to a function g ∈ M T , and µn ≡ lim sup νn(fj, gj) = o(n). j→∞ (2.6) 4 Then, there is a subsequence of {fj}, which converges pointwise on T to a function f ∈ M T such that νn(f, g) ≤ µn for all n ∈ N. This theorem will be proved in Section 5. Now, a few remarks are in order. Given f ∈ M T and a constant function c : T → M, the quantity νn(f ) ≡ νn(f, c) = sup(cid:26) n Xi=1 d(f (si), f (ti)) : {Ii}n 1 ≺ T(cid:27) (2.7) (with Ii = {si, ti}) is independent of c, and the sequence {νn(f )}∞ n=1 ⊂ [0, ∞] is known as the modulus of variation of f in the sense of Chanturiya ([4, 5, 11, 12, 13, 28]). It characterizes regulated (or proper) functions on T = [a, b] as follows. We say that f : [a, b] → M is regulated and write f ∈ Reg([a, b]; M) if d(f (s), f (t)) → 0 as s, t → τ − 0 for every a < τ ≤ b, and d(f (s), f (t)) → 0 as s, t → τ ′ + 0 for every a ≤ τ ′ < b (and so, by Cauchy's criterion, one-sided limits f (τ − 0), f (τ ′ + 0) ∈ M exist provided M is complete). We have Reg([a, b]; M) = {f ∈ M [a,b] : νn(f ) = o(n)} (2.8) (more general characterizations for dense subsets T of [a, b] can be found in [12, 13]). A certain relationship between characterizations of regulated functions and pointwise selection principles is exhibited in [18]. 3. The joint modulus of variation We begin by studying the joint increment (2.1), whose properties are gathered in the following lemma. Lemma 1. Given f, g, h ∈ M T and s, t ∈ T , we have: (a) (f, f )({s, t}) = 0; (b) (f, g)({s, t}) = (g, f )({s, t}); (c) (f, g)({s, t}) ≤ (f, h)({s, t}) + (h, g)({s, t}); (d) d(f (s), f (t)) ≤ d(g(s), g(t)) + (f, g)({s, t}); (e) d(f (t), g(t)) ≤ d(f (s), g(s)) + (f, g)({s, t}). Proof. Properties (a), (b), and (c), showing that (f, g) 7→ (f, g)({s, t}) is a pseudometric on M T , are straightforward. To establish (d) and (e), take into account equality d(x, y) = maxz∈M d(x, z) − d(y, z). 5 Remark 1. (a) If (f, g)({s, t}) = 0, then (d), (e), and (b) imply equalities d(f (s), f (t)) = d(g(s), g(t)) and d(f (t), g(t)) = d(f (s), g(s)). In addition to Lemma 1, the function (s, t) 7→ (f, g)({s, t}) is a pseudometric on T . (b) If F (z) denotes the absolute value under the supremum sign in (2.1), then F : M → [0, ∞) and F (z) − F (z0) ≤ 4d(z, z0) for all z, z0 ∈ M. (c) By Lemma 1(d), f (T ) ≤ g(T ) + (f, g)(T ) = g(T ) + ν1(f, g). So, (cid:12)(cid:12)f (T ) − g(T )(cid:12)(cid:12) ≤ (f, g)(T ) ≤ f (T ) + g(T ), Moreover, it follows from Lemma 1(e) that f, g ∈ B(T ; M). d∞(f, g) ≤ d(f (s), g(s)) + (f, g)(T ) ≤ 3d∞(f, g) for all s ∈ T. (d) Suppose the triple (M, d, +) is a metric semigroup ([10, Section 4]), i.e., (M, d) is a metric space, (M, +) is an Abelian semigroup with the oper- ation of addition +, and d(x, y) = d(x + z, y + z) for all x, y, z ∈ M. Then, the joint increment (2.1) may be alternatively replaced by (f, g)({s, t}) = d(f (s) + g(t), f (t) + g(s)). (3.1) The joint modulus of variation (2.4) involving (3.1) was employed in [17]. Furthermore, if (M, k · k) is a normed linear space (over R or C), we may set (f, g)({s, t}) = kf (s) + g(t) −f (t) −g(s)k = k(f −g)(s) −(f −g)(t)k. (3.2) Quantities (3.1) and (3.2) have the same properties as (2.1): see (2.2), (2.3), Lemma 1 and Remark 1(a). In the sequel, we make use of more general quantity (2.1). If f, g ∈ M T , n ∈ N and ∅ 6= E ⊂ T , we set νn(f, g; E) = νn(f E, gE), where f E ∈ M E is the restriction of f to E (i.e., f E(t) = f (t) for all t ∈ E). Accordingly, νn(f, g) = νn(f, g; T ). The following properties of the joint modulus of variation are immediate. The sequence {νn(f, g)}∞ n=1 is nondecreasing, νn+m(f, g) ≤ νn(f, g) + νm(f, g) for all n, m ∈ N, and νn(f, g; E) ≤ νn(f, g; T ) provided n ∈ N and E ⊂ T . It follows from (2.4) and Lemma 1(a) -- (c) that, for every n ∈ N, the function (f, g) 7→ νn(f, g) is a pseudometric on M T (possibly assuming infinite values) and, in particular (cf. (2.4) and (2.7)), νn(f, g) ≤ νn(f ) + νn(g) and νn(f ) ≤ νn(g) + νn(f, g) (3.3) for all n ∈ N and f, g ∈ M T . Furthemore, if f, g ∈ B(T ; M), then, by (2.5), the sequence {νn(f, g)/n}∞ n=1 is bounded in [0, ∞). 6 Essential properties of the joint modulus of variation are presented in Lemma 2. Given n ∈ N, f, g ∈ M T , and ∅ 6= E ⊂ T , we have: (a) (f, g)({s, t}) + νn(f, g; E − s ) ≤ νn+1(f, g; E − τ = (−∞, τ ] ∩ E for τ ∈ E; t ) for all s, t ∈ E with s ≤ t, where E − (b) νn+1(f, g; E) ≤ νn(f, g; E) + νn+1(f, g; E) n + 1 ; (c) if {fj}, {gj} ⊂ M T are such that fj → f and gj → g on E, then νn(f, g; E) ≤ lim inf j→∞ νn(fj, gj; E); (d) if {fj}, {gj} ⊂ M T are such that fj ⇒ f and gj ⇒ g on E, then νn(f, g; E) = limj→∞ νn(fj, gj; E). Proof. (a) We may assume that s < t. Let {Ii}n we find {Ii}n t , and so, 0 ≺ E − 1 ≺ E − s . Setting I0 = {s, t}, (f, g)(I0) + n Xi=1 (f, g)(Ii) ≤ νn+1(f, g; E − t ). The inequality in (a) follows by taking the supremum over all {Ii}n 1 ≺ E − s . (b) We may assume that νn+1(f, g; E) is finite, and apply the idea from [5, Lemma]. Given ε > 0, there is {Ii}n+1 1 ≺ E (depending on ε) such that n+1 Xi=1 (f, g)(Ii) ≤ νn+1(f, g; E) ≤ n+1 Xi=1 (f, g)(Ii) + ε. Setting a0 = min1≤i≤n+1 (f, g)(Ii), the left-hand side inequality implies (n + 1)a0 ≤ νn+1(f, g; E). The right-hand side inequality gives νn+1(f, g; E) ≤ νn(f, g; E) + a0 + ε, from which our inequality follows due to the arbitrariness of ε > 0. (c) First, we note that, given j ∈ N and s, t ∈ T , we have (cid:12)(cid:12)(fj, gj)({s, t}) − (f, g)({s, t})(cid:12)(cid:12) ≤ d(fj(s), f (s)) + d(fj(t), f (t)) + d(gj(s), g(s)) + d(gj(t), g(t)). (3.4) 7 In fact, Lemma 1(c) and inequality (2.3) imply (fj, gj)({s, t}) ≤ (fj, f )({s, t}) + (f, g)({s, t}) + (g, gj)({s, t}) ≤ d(fj(s), f (s)) + d(fj(t), f (t)) + (f, g)({s, t}) + d(g(s), gj(s)) + d(g(t), gj(t)). (3.5) Exchanging fj and f as well as gj and g, we obtain (3.4). From the pointwise convergence of {fj} and {gj} and (3.4), we find lim j→∞ (fj, gj)({s, t}) = (f, g)({s, t}) for all s, t ∈ E. By definition (2.4), given {Ii}n 1 ≺ E, we have n Xi=1 (fj, gj)(Ii) ≤ νn(fj, gj; E) for all j ∈ N. Passing to the limit inferior as j → ∞, we get n Xi=1 (f, g)(Ii) ≤ lim inf j→∞ νn(fj, gj; E). (3.6) Since {Ii}n 1 ≺ E is arbitrary, it remains to take into account (2.4). (d) It follows from (3.5) that, for any s, t ∈ E and j ∈ N, (fj, gj)({s, t}) ≤ 2 sup τ ∈E d(fj(τ ), f (τ ))+(f, g)({s, t})+2 sup τ ∈E d(gj(τ ), g(τ )), and so, definition (2.4) implies νn(fj, gj; E) ≤ 2n sup τ ∈E d(fj(τ ), f (τ ))+νn(f, g; E)+2n sup τ ∈E d(gj(τ ), g(τ )) (3.7) for all j ∈ N. Passing to the limit superior as j → ∞, we get lim sup νn(fj, gj; E) ≤ νn(f, g; E). j→∞ Now, the equality in (d) is a consequence of Lemma 2(c). Remark 2. If the value ν1(f, g; E) = (f, g)(E) (see (2.5)) is finite for an E ⊂ T (e.g., when f, g ∈ B(E; M)), inequality in Lemma 2(b) is equivalent to νn+1(f, g; E) νn(f, g; E) . n n + 1 ≤ Thus, the limit limn→∞ νn(f, g; E)/n always exists in [0, ∞). 8 4. Conditionally regulated functions Since νn = νn(·, ·) is a (extended) pseudometric on M T , we may introduce an equivalence relation ∼ on M T as follows: given f, g ∈ M T , we set f ∼ g if and only if νn(f, g) = o(n). The equivalence class R(g) = {f ∈ M T : f ∼ g} of a function g ∈ M T is called the regularity class of g, and any representative f ∈ R(g) is called a conditionally regulated or, more precisely, g-regulated function. This termi- in the framework of the product set M [a,b], we nology is justified by (2.8): have Reg([a, b]; M) = R(c) for any constant function c : [a, b] → M. Note that, in Theorem 1, condition 'νn(f, g) ≤ µn for all n ∈ N' means that f ∈ R(g), and so, the class R(g) is worth studying in more detail. Theorem 2. Given a function g ∈ M T , we have: (a) g ∈ B(T ; M) if and only if R(g) ⊂ B(T ; M); (b) R(g) is closed with respect to the uniform convergence, but not closed with respect to the pointwise convergence in general ; (c) if (M, d) is a complete metric space, then the pair (R(g), d∞) is also a complete metric space. Proof. (a) The sufficiency is clear, because g ∈ R(g). Now, suppose that g ∈ B(T ; M), so that, by (2.7), ν1(g) = (g, c)(T ) = g(T ) < ∞. Given f ∈ R(g), νn(f, g) = o(n), and so, νn0(f, g) ≤ n0 for some n0 ∈ N. It follows from (3.3) and (2.5) that f (T ) = ν1(f ) ≤ ν1(g) + ν1(f, g) ≤ g(T ) + νn0(f, g) ≤ g(T ) + n0 < ∞, which implies f ∈ B(T ; M). (b) It is to be shown that if {fj} ⊂ R(g) and fj ⇒ f on T with f ∈ M T , then f ∈ R(g). We will prove a little bit more: namely, if {fj}, {gj} ⊂ M T are such that fj ∈ R(gj) for all j ∈ N, fj ⇒ f and gj ⇒ g on T with f, g ∈ M T , then f ∈ R(g) (the previous assertion follows if gj = g for all j ∈ N). In fact, exchanging fj and f , and gj and g in (3.7), we get νn(f, g) n ≤ 2d∞(f, fj) + νn(fj, gj) n 9 + 2d∞(g, gj), n, j ∈ N. By the uniform convergence of {fj} and {gj}, given ε > 0, there is a number j0 = j0(ε) ∈ N such that d∞(f, fj0) ≤ ε and d∞(g, gj0) ≤ ε. Since fj0 is in R(gj0), we have νn(fj0, gj0) = o(n), and so, there exists n0 = n0(ε) ∈ N such that νn(fj0, gj0)/n ≤ ε for all n ≥ n0. The estimate above with j = j0 implies νn(f, g)/n ≤ 5ε, n ≥ n0, which means that νn(f, g) = o(n) and f ∈ R(g). As for the pointwise convergence, consider a sequence of real step func- tions converging pointwise to the Dirichlet function (= the characteristic function of the rationals Q) on T = [0, 1] (see [11, Examples 4, 5] and Ex- ample 2(a) in Section 5). (c) First, we show that d∞(f, f ′) < ∞ for all f, f ′ ∈ R(g). In fact, since f ∼ f ′, we have νn(f, f ′) = o(n), and so, νn0(f, f ′) ≤ n0 for some n0 ∈ N. Given s ∈ T , it follows from Remark 1(c) and (2.5) that d∞(f, f ′) ≤ d(f (s), f ′(s)) + ν1(f, f ′) ≤ d(f (s), f ′(s)) + νn0(f, f ′) < ∞. The metric axioms for d∞ on R(g) are verified in a standard way. In order to prove that R(g) is complete, suppose {fj} ⊂ R(g) is a Cauchy sequence, i.e., d∞(fj, fk) → 0 as j, k → ∞. Since d(fj(t), fk(t)) ≤ d∞(fj, fk) for all t ∈ T and (M, d) is complete, there exists f ∈ M T such that fj → f on T . Noting that fj → fj and fk → f on T as k → ∞ (and arguing as in (3.6)), we get d∞(fj, f ) ≤ lim inf k→∞ d∞(fj, fk) = lim k→∞ d∞(fj, fk) < ∞ for all j ∈ N. Since the sequence {fj} is d∞-Cauchy, we find lim sup j→∞ d∞(fj, f ) ≤ lim j→∞ lim k→∞ d∞(fj, fk) = 0. Thus, limj→∞ d∞(fj, f ) = 0, and so, fj ⇒ f on T . Applying item (b) of this Theorem, we conclude that f ∈ R(g). A traditionally important class of regulated functions is the space of func- tions of bounded Jordan variation, BV(T ; M), which is introduced by means of the joint modulus of variation as follows. Since the sequence {νn(f, g)}∞ n=1 is nondecreasing for all f, g ∈ M T , the quantity (finite or not) V (f, g) = limn→∞ νn(f, g) = supn∈N νn(f, g) is called the joint variation of functions f and g on T . The value V (f ) ≡ V (f, c) 10 is independent of a constant function c : T → M and is the usual Jordan variation of f on T : V (f ) = sup(cid:26) n Xi=1 d(f (ti), f (ti−1)) : n ∈ N and {ti}n 0 ≺ T(cid:27), the supremum being taken over all partitions {ti}n 0 of T (cf. Section 2). The set BV(T ; M) = {f ∈ M T : V (f ) < ∞} is contained in B(T ; M) ∩ R(c) (in fact, f (T ) = ν1(f ) ≤ V (f ) and νn(f, c)/n ≤ V (f )/n for all f ∈ BV(T ; M)). The following notion of ε-variation Vε(f ), due to Frankov´a [25, Section 3], provides an alternative characterization (cf. (2.8)) of regulated functions: given f ∈ M T and ε > 0, set Vε(f ) = inf(cid:8)V (g) : g ∈ BV(T ; M) and d∞(f, g) ≤ ε(cid:9) (inf ∅ = ∞). It was shown in [25, Proposition 3.4] (for T = [a, b]) that (4.1) Reg([a, b]; M) = {f ∈ M [a,b] : Vε(f ) < ∞ for all ε > 0} (4.2) (although it was assumed in [25] that M = RN , the proof of the last assertion carries over to any metric space M, cf. [11, Lemma 3]). The notion of ε-variation will be needed in Section 6. Example 1. Given x, y ∈ M with x 6= y, let f = Dx,y : T = [0, 1] → M be the Dirichlet-type function of the form: Dx,y(t) = (cid:26) x if y if t ∈ [0, 1] is rational, t ∈ [0, 1] is irrational. (4.3) Clearly, f /∈ Reg([0, 1]; M). Moreover (cf. (4.2)), we have Vε(f ) = ∞ if 0 < ε < d(x, y)/2, and Vε(f ) = 0 if ε ≥ d(x, y). (4.4) To see this, first note that, given g ∈ M [0,1], inequality d∞(f, g) ≤ ε is equivalent to the following two conditions: d(x, g(s)) ≤ ε ∀ s ∈ [0, 1] ∩ Q, and d(y, g(t)) ≤ ε ∀ t ∈ [0, 1] \ Q. (4.5) Suppose 0 < ε < d(x, y)/2. To show that d∞(f, g) ≤ ε implies V (g) = ∞, 0 ≺ [0, 1] be a partition of [0, 1] such that points we let n ∈ N, and {ti}2n 11 i=0 are rational and points {t2i−1}n {t2i}n inequality for d and (4.5), we get i=1 are irrational. By the triangle 2n n V (g) ≥ ≥ n Xi=1 d(g(t2i), g(t2i−1)) d(g(ti), g(ti−1)) ≥ Xi=1 Xi=1(cid:0)d(x, y) − d(x, g(t2i)) − d(g(t2i−1), y)(cid:1) ≥ n(cid:0)d(x, y) − 2ε(cid:1). If ε ≥ d(x, y), we set g(t) = x (or g(t) = y) for all t ∈ [0, 1], so that (4.5) is satisfied and V (g) = 0. Thus, Vε(f ) = 0. The second assertion in (4.4) can be refined, provided d(x, y)/2 = max{d(x, z0), d(y, z0)} for some z0 ∈ M. (4.6) In fact, we may set g(t) = z0 for all t ∈ [0, 1], so that (4.5) holds whenever d(x, y)/2 ≤ ε, and V (g) = 0. This implies Vε(f ) = 0 for all ε ≥ d(x, y)/2. A few remarks concerning condition (4.6) are in order. Since d(x, y) ≤ d(x, z) + d(z, y) ≤ 2 max{d(x, z), d(y, z)} for all z ∈ M, condition (4.6) refers to a certain form of 'convexity' of M (which is not restrictive for our purposes). For instance, if (M, k · k) is a normed linear space with d(x, y) = kx − yk, we may set z0 = (x + y)/2. More generally, by Menger's Theorem ([35], [27, Example 2.7]), if a metric space (M, d) is complete and metrically convex (i.e., given x, y ∈ M with x 6= y, there is z ∈ M such that x 6= z 6= y and d(x, y) = d(x, z) + d(z, y)), then, for any x, y ∈ M, there is an isometry ϕ : [0, d(x, y)] → M such that ϕ(0) = x and ϕ(d(x, y)) = y. In this case, we set z0 = ϕ(d(x, y)/2). More examples of metrically convex metric spaces can be found in [21, 24]. Finally, if M = {x, y}, then condition (4.6) is not satisfied, and we have Vε(f ) = ∞ for all 0 < ε < d(x, y), which is a consequence of (4.5). 5. Proof of the main result Proof of Theorem 1. With no loss of generality we may assume that T is uncountable; otherwise, by virtue of assumption (a) and the standard Cantor diagonal procedure, we extract a pointwise convergent subsequence of {fj} and apply Lemma 2(c). Note that µn is finite for all n ∈ N: in fact, µn ≤ n whenever n ≥ n0 for some n0 ∈ N and, since n 7→ νn(fj, gj) is nondecreasing for all j ∈ N, we have µn ≤ µn0 ≤ n0 for all 1 ≤ n < n0. 12 We divide the rest of the proof into four steps for clarity. Step 1. Let us show that there is a subsequence of {j}∞ j=1, again denoted by {j}, and a nondecreasing sequence {αn}∞ n=1 ⊂ [0, ∞) such that lim j→∞ νn(fj, gj) = αn ≤ µn for all n ∈ N. (5.1) j=1 ⊂ N (i.e., a subsequence of {j}∞ We set α1 = µ1. The definition (2.6) of µ1 implies that there is an in- creasing sequence {J1(j)}∞ j=1) such that ν1(fJ1(j), gJ1(j)) → α1 as j → ∞. Setting α2 = lim supj→∞ ν2(fJ1(j), gJ1(j)), we find α2 ≤ µ2, and there is a subsequence {J2(j)}∞ j=1 such that ν2(fJ2(j), gJ2(j)) → α2 as j → ∞. Inductively, if n ≥ 3 and a subse- quence {Jn−1(j)}∞ j=1 is already chosen, we define αn as the limit superior of νn(fJn−1(j), gJn−1(j)) as j → ∞, so that αn ≤ µn. Now, we pick a subsequence {Jn(j)}∞ j=1 such that νn(fJn(j), gJn(j)) → αn as j=n is a subsequence of {Jn(j)}∞ j → ∞. Noting that the sequence {Jj(j)}∞ j=1 (for all n ∈ N) and denoting the diagonal sequences {fJj(j)}∞ j=1 and {gJj(j)}∞ j=1 again by {fj} and {gj}, respectively, we obtain (5.1). j=1 of {Jn−1(j)}∞ j=1 of {J1(j)}∞ j=1 of {j}∞ In the sequel, the set of all nondecreasing bounded functions mapping T into R+ = [0, ∞) is denoted by Mon(T ; R+). Step 2. In this step, we prove that there are subsequences of {fj} and {gj} from (5.1), again denoted by {fj} and {gj}, respectively, and a sequence of functions {βn}∞ n=1 ⊂ Mon(T ; R+) such that lim j→∞ νn(fj, gj; T − t ) = βn(t) for all n ∈ N and t ∈ T , (5.2) where T − t = {s ∈ T : s ≤ t} for t ∈ T . Note that, for each n ∈ N, the function t 7→ νn(fj, gj; T − on T , and νn(fj, gj; T − (5.1), there is a sequence {Cn}∞ n, j ∈ N. In what follows, we apply the diagonal procedure once again. t ) is nondecreasing t ) ≤ νn(fj, gj) for all t ∈ T and n ∈ N. By virtue of n=1 ⊂ R+ such that νn(fj, gj) ≤ Cn for all t )}∞ The sequence {t 7→ ν1(fj, gj; T − j=1 ⊂ Mon(T ; R+) is uniformly bounded by constant C1, and so, by Helly's Selection Principle, there are an increas- ing sequence {K1(j)}∞ j=1) and a function β1 ∈ Mon(T ; R+) such that ν1(fK1(j), gK1(j); T − t ) → β1(t) as j → ∞ for all j=1 ⊂ Mon(T ; R+) is uni- t ∈ T . The sequence {t 7→ ν2(fK1(j), gK1(j); T − t )}∞ formly bounded on T by constant C2, and so, again by Helly's Theorem, there j=1 and a function β2 ∈ Mon(T ; R+) are a subsequence {K2(j)}∞ j=1 ⊂ N (i.e., a subsequence of {j}∞ j=1 of {K1(j)}∞ 13 t ) → β2(t) as j → ∞ for all t ∈ T . such that ν2(fK2(j), gK2(j); T − Induc- tively, if n ≥ 3 and a subsequence {Kn−1(j)}∞ j=1 and a function βn−1 ∈ Mon(T ; R+) are already chosen, we apply the Helly Theorem to j=1 ⊂ Mon(T ; R+), the sequence of functions {t 7→ νn(fKn−1(j), gKn−1(j); T − which is uniformly bounded on T by constant Cn: there are a subsequence j=1 and a function βn ∈ Mon(T ; R+) such that {Kn(j)}∞ νn(fKn(j), gKn(j); T − j=n is a subsequence of {Kn(j)}∞ j=1 (for all n ∈ N), the diagonal sequences {fKj(j)}∞ j=1 and {gKj(j)}∞ j=1, again denoted by {fj} and {gj}, respectively, satisfy condi- tion (5.2). t ) → βn(t) as j → ∞ for all t ∈ T . Since {Kj(j)}∞ j=1 of {Kn−1(j)}∞ j=1 of {j}∞ t )}∞ Step 3. Let Q be an at most countable dense subset of T . Note that Q contains all isolated (= non-limit) points of T (i.e., points t ∈ T such that the intervals (t − δ, t) and (t, t + δ) lie in R \ T for some δ > 0). The set Qn ⊂ T of discontinuity points of nondecreasing function βn is at most countable. n=1 Qn, we find that S is an at most countable dense Setting S = Q ∪ S∞ subset of T and βn is continuous at all points of T \ S for all n ∈ N. (5.3) Since the set {fj(t) : j ∈ N} is precompact in M for all t ∈ T , and S ⊂ T is at most countable, we may assume (applying the diagonal procedure again and passing to a subsequence of {fj} if necessary) that, given s ∈ S, there is a point f (s) ∈ M such that d(fj(s), f (s)) → 0 as j → ∞. In this way, we obtain a function f : S → M. Step 4. Now, we show that, for every t ∈ T \ S, the sequence {fj(t)}∞ j=1 converges in M. For this, we prove that this sequence is Cauchy in M, i.e., d(fj(t), fk(t)) → 0 as j, k → ∞. Let ε > 0 be arbitrarily fixed. By assumption (2.6), µn/n → 0 as n → ∞, so we choose and fix a number n = n(ε) ∈ N such that µn+1 n + 1 ≤ ε. By property (5.1), there is a number j1 = j1(ε, n) ∈ N such that νn+1(fj, gj) ≤ αn+1 + ε ≤ µn+1 + ε for all j ≥ j1. The definition of the set S and (5.3) imply that the point t is a limit point for T and, at the same time, a point of continuity of the function βn. By the density of S in T , there is s = s(ε, n, t) ∈ S such that βn(t) − βn(s) ≤ ε. 14 It follows from (5.2) that there is j2 = j2(ε, n, t, s) ∈ N such that νn(fj, gj; T − t ) − βn(t) ≤ ε and νn(fj, gj; T − s ) − βn(s) ≤ ε ∀ j ≥ j2. Assuming that s < t (the arguments are similar if t < s) and applying Lemma 2(a), (b), we get, for all j ≥ max{j1, j2}: (fj, gj)({s, t}) ≤ νn+1(fj, gj; T − ≤ νn+1(fj, gj; T − + νn(fj, gj; T − + βn(s) − νn(fj, gj; T − νn+1(fj, gj; T − t ) s ) t ) − νn(fj, gj; T − s ) t ) − νn(fj, gj; T − t ) t ) − βn(t) + βn(t) − βn(s) + ε + ε + ε ≤ ≤ ≤ n + 1 νn+1(fj, gj) + 3ε n + 1 µn+1 n + 1 + ε n + 1 + 3ε ≤ 5ε. j=1 and j=1 are Cauchy in M, and so, there is a number j3 = j3(ε, s, t) ∈ N Being convergent (see condition (b)), sequences {fj(s)}∞ {gj(t)}∞ such that, for all j, k ≥ j3, we have j=1, {gj(s)}∞ d(fj(s), fk(s)) ≤ ε, d(gj(s), gk(s)) ≤ ε, and d(gj(t), gk(t)) ≤ ε. By virtue of inequality (2.3), we get (gj, gk)({s, t}) ≤ d(gj(s), gk(s)) + d(gj(t), gk(t)) ≤ 2ε ∀ j, k ≥ j3. Putting j4 = max{j1, j2, j3} and applying Lemma 1(e), (c), (b), we find d(fj(t), fk(t)) ≤ d(fj(s), fk(s)) + (fj, fk)({s, t}) ≤ d(fj(s), fk(s)) + (fj, gj)({s, t}) + (gj, gk)({s, t}) + (gk, fk)({s, t}) ≤ ε + 5ε + 2ε + 5ε = 13ε for all j, k ≥ j4. Since j4 depends only on ε (and t), this proves that {fj(t)}∞ j=1 is a Cauchy se- quence in M, which together with assumption (a) establishes its convergence in M to an element denoted by f (t) ∈ M. 15 Here and at the end of Step 3, we have shown that f : T = S∪(T \S) → M k=1 of the original sequence j=1. Since gjk → g pointwise on T as k → ∞ as well, we conclude from is a pointwise limit on T of a subsequence {fjk}∞ {fj}∞ Lemma 2(c) that νn(f, g) ≤ lim inf k→∞ νn(fjk, gjk) ≤ lim sup j→∞ νn(fj, gj) = µn ∀ n ∈ N, and so, νn(f, g) = o(n), or f ∈ R(g). This completes the proof of Theorem 1. (cid:3) Remark 3. (a) Condition (b) in Theorem 1 may be replaced by the following one: {gj(t)}∞ j=1 is a Cauchy sequence in M for every t ∈ T . However, if (M, d) is not complete, we may no longer infer the property νn(f, g) ≤ µn, n ∈ N, of the pointwise limit f (as there may be no g). (b) Condition (2.6) is necessary for the uniformly convergent sequences {fj} and {gj}: in fact, if fj ⇒ f and gj ⇒ g on T , and νn(f, g) = o(n), then it follows from Lemma 2(d) that limj→∞ νn(fj, gj) = νn(f, g) = o(n). (c) Condition (2.6) is 'almost necessary' in the following sense. Suppose T ⊂ R is a measurable set with Lebesgue measure L(T ) < ∞, {fj}, {gj} ⊂ M T are two sequences of measurable functions, which converge pointwise (or almost everywhere) on T to functions f, g ∈ M T , respectively, such that νn(f, g) = o(n). By Egorov's Theorem, given ε > 0, there exists a measurable set Eε ⊂ T such that L(T \ Eε) ≤ ε, fj ⇒ f and gj ⇒ g on Eε. So, as in the previous remark (b), we have lim j→∞ νn(fj, gj; Eε) = νn(f, g; Eε) ≤ νn(f, g) = o(n). Example 2. (a) Condition (2.6) is not necessary for the pointwise conver- gence even if gj = c for all j ∈ N. To see this, let T = [0, 1] and x, y ∈ M with x 6= y. Given j ∈ N, define fj : T → M by: fj(t) = x if j! · t is integer, and fj(t) = y otherwise, t ∈ [0, 1]. The pointwise precompact sequence {fj} ⊂ M T consists of bounded regulated functions (in fact, νn(fj, c) = o(n), and so, fj ∈ Reg([0, 1]; M) = R(c) for all j ∈ N). It converges pointwise on T to the Dirichlet-type function Dx,y from (4.3). Note that νn(Dx,y, c) = nd(x, y), and so, Dx,y /∈ R(c). Since the usual Jordan variation V (fj) of fj on T = [0, 1] is equal to 2 · j!d(x, y), we find νn(fj, c) = d(x, y) ·(cid:26) n 2 · j! if n < 2 · j!, if n ≥ 2 · j!, n, j ∈ N. Thus, limj→∞ νn(fj, c) = d(x, y) · n, i.e., condition (2.6) does not hold. 16 (b) Under the assumptions of Theorem 1, condition (2.6) does not in general imply lim supj→∞ νn(fj, g) = o(n). To see this, let gj = fj be as in example (a) above, so that g = Dx,y. Given n, j ∈ N, choose a collection {Ii}n 1 ≺ (0, 1/j!) with Ii = {si, ti} such that si is rational and ti is irrational for all i = 1, . . . , n. Noting that, by virtue of (2.1), (fj, g)({si, ti}) = sup z∈M d(y, z) − d(x, z) = d(y, x), we get νn(fj, g) ≥ n Xi=1 (fj, g)(Ii) = nd(y, x) for all n, j ∈ N. (c) The choice of an appropriate sequence {gj} is essential in Theorem 1. Let {xj}, {yj} ⊂ M be two sequences, which converge in M to x, y ∈ M, respectively, x 6= y. Define fj : T = [0, 1] → M by fj = Dxj ,yj , j ∈ N (cf. (4.3)). Clearly, {fj} converges uniformly on T to Dx,y (so, {fj} is pointwise precompact on T ), and νn(fj, c) = nd(xj, yj) for all n, j ∈ N. Since d(xj, yj) − d(x, y) ≤ d(xj, x) + d(yj, y) → 0 as j → ∞, we find limj→∞ νn(fj, c) = nd(x, y), condition (2.6) is not satisfied, and The- orem 1 is inapplicable with gj = c, j ∈ N. On the other hand, set gj = Dx,y for all j ∈ N. Given {s, t} ⊂ T , we have, by virtue of (2.3), (fj, gj)({s, t}) ≤ d(fj(s), gj(s)) + d(fj(t), gj(t)) ≤ 2εj, where εj = max{d(xj, x), d(yj, y)} → 0 as j → ∞. This implies the inequal- ity νn(fj, gj) ≤ 2nεj, and so, condition (2.6) is fulfilled. 6. Extensions of known selection theorems In this section, we consider extensions of two selection theorems from [25] and [22, 32]. The other selection theorems from the references in the Introduction were shown to be particular cases of [11] -- [13] (see Remark 5). By virtue of the inequality νn(fj, gj)/n ≤ V (fj, gj)/n, instead of condi- tion (2.6) in Theorem 1 we may assume that lim supj→∞ V (fj, gj) < ∞ or supj∈N V (fj, gj) < ∞, in which cases the resulting pointwise limit f of a sub- sequence of {fj} satisfies the regularity condition of the form V (f, g) < ∞. 17 Making use of the notion of ε-variation (Section 4), we get the following Theorem 3. Given ∅ 6= T ⊂ R and a metric space (M, d), let {fj} ⊂ M T be a pointwise precompact sequence of functions such that lim sup Vε(fj) < ∞ for all ε > 0. (6.1) j→∞ Then, there is a subsequence {fjk} of {fj}, which converges pointwise on T to a regulated function f ∈ R(c). Proof. Taking into account Theorem 1, it suffices to verify that condition (6.1) implies lim supj→∞ νn(fj, c) = o(n), which is (2.6) with gj = c for all j ∈ N. In fact, by (6.1), for every ε > 0 there is j0 = j0(ε) ∈ N and C(ε) > 0 such that Vε(fj) ≤ C(ε) for all j ≥ j0. Definition (4.1) yields the existence of gj ∈ BV(T ; M) such that d∞(fj, gj) ≤ ε and V (gj) ≤ Vε(fj)+(1/j) ≤ C(ε)+1 for all j ≥ j0. By (3.7) (where we replace gj and g by c, and f -- by gj), νn(fj, c) n ≤ 2d∞(fj, gj) + νn(gj, c) n ≤ 2ε + V (gj) n ≤ 2ε + C(ε) + 1 n for all j ≥ j0 and n ∈ N. Consequently, 1 n lim sup νn(fj, c) ≤ j→∞ 1 n sup j≥j0 νn(fj, c) ≤ 2ε + C(ε) + 1 n ∀ ε > 0, n ∈ N. This implies that the left-hand side in this inequality tends to zero as n → ∞: given η > 0, we set ε = η/4 and choose a number n0 = n0(η) ∈ N such that (C(ε) + 1)/n0 ≤ η/2, which yields 2ε + (C(ε) + 1)/n ≤ η for all n ≥ n0. Remark 4. (a) If M = RN in Theorem 3, we may infer that Vε(f ) does not exceed the limit superior from (6.1): in fact, it follows from [25, Proposi- tion 3.6] that Vε(f ) ≤ lim inf k→∞ Vε(fjk) for all ε > 0. (b) Theorem 3 extends Theorem 3.8 from [25], which has been established for T = [a, b] and M = RN under the assumption that supj∈N Vε(fj) < ∞ for every ε > 0. The last assumption on the uniform boundedness of ε-variations is more restrictive than condition (6.1) as the following example shows. Example 3. Let {xj} and {yj} be two sequences from M such that xj 6= yj for all j ∈ M and, for some x ∈ M, xj → x and yj → x in M as j → ∞. 18 We set fj = Dxj ,yj , j ∈ N, and f (t) = x for all t ∈ T = [0, 1]. The sequence {fj} ⊂ M T converges uniformly on T to the constant function f : d∞(fj, f ) = max{d(xj, x), d(yj, x)} → 0 as j → ∞. Given ε > 0, there is j0 = j0(ε) ∈ N such that d(xj, yj) ≤ ε for all j ≥ j0, and so, by (4.4), Vε(fj) = 0 for all j ≥ j0, which implies condition (6.1): lim sup j→∞ Vε(fj) ≤ sup j≥j0 Vε(fj) = 0. On the other hand, if k ∈ N is fixed and 0 < ε < d(xk, yk)/2, then (4.4) gives Vε(fk) = ∞, and so, supj∈N Vε(fj) = ∞. Now, we are going to present an extension of a Helly-type selection the- orem from [32, Section 4, Theorem 1] and [22, Theorem 2]. Let κ : [0, 1] → [0, 1] be a continuous, increasing and concave function such that κ(0) = 0, κ(1) = 1, and κ(τ )/τ → ∞ as τ → +0 (e.g., κ(τ ) = τ (1 − log τ ), κ(τ ) = τ α with 0 < α < 1, or κ(τ ) = 1/(1 − 1 2 log τ ), see [33]). Let T = [a, b] be a closed interval in R, a < b. We set T = b − a, and if 0 ≺ [a, b] is a partition of T (i.e., a = t0 < t1 < · · · < tn−1 < tn = b), we {ti}n also set Ii = {ti−1, ti} and Ii = ti − ti−1, i = 1, . . . , n. The joint κ-variation of functions f, g ∈ M T = M [a,b] is defined by Vκ(f, g) = sup(cid:26) n Xi=1 (f, g)(Ii)(cid:30) n Xi=1 κ(cid:0)Ii/T (cid:1) : n ∈ N and {ti}n 0 ≺ [a, b](cid:27), where (f, g)(Ii) = (f, g)({ti−1, ti}) is given by (2.1). Since (f, c)(Ii) = d(f (ti−1), f (ti)) is independent of a constant function c : [a, b] → M, the quantity Vκ(f ) ≡ Vκ(f, c) is the Korenblum κ-variation of f ∈ M [a,b], introduced in [32, p. 191] and [33, Section 5] for M = R. The following theorem is a generalization of Theorem 2 from [22], estab- lished for real functions of bounded κ-variation under the assumption that supj∈N Vκ(fj) < ∞ and based on the decomposition of any f ∈ R[a,b] with Vκ(f ) < ∞ into the difference of two real κ-decreasing functions. Theorem 4. Under the assumptions of Theorem 1, suppose that condition (2.6) is replaced by the following one: lim supj→∞ Vκ(fj, gj) < ∞. Then, there is a subsequence of {fj}, which converges pointwise on T = [a, b] to a function f ∈ R(g) such that Vκ(f, g) < ∞. 19 Proof. In order to show that (2.6) is satisfied, let n ∈ N, {Ii}n Ii = {si, ti}, and set I ′ the definition of Vκ(fj, gj) and the concavity of function κ, we have 1 ≺ [a, b] with i = si+1 − ti, i = 1, . . . , n − 1. By i = {ti, si+1} and I ′ n Xi=1 (fj, gj)(Ii) ≤ (fj, gj)({a, s1}) + n−1 n Xi=1 (fj, gj)(Ii) n + Xi=1 T (cid:19)+ ≤(cid:20)κ(cid:18)s1−a ≤ (2n+1)κ(cid:18) + n−1 Xi=1 (fj, gj)(I ′ i) + (fj, gj)({tn, b}) n−1 Xi=1 κ(cid:18)Ii T (cid:19)+ κ(cid:18) I ′ Xi=1 (2n+1)(b − a)(cid:20)(s1−a) + Xi=1 (si+1−ti) + (b − tn)(cid:21)(cid:19)Vκ(fj, gj) 1 n i T (cid:19)+κ(cid:18)b−tn T (cid:19)(cid:21)Vκ(fj, gj) (ti −si) ≤ (2n+1)κ(cid:18) 1 2n+1(cid:19)Vκ(fj, gj). Thus, νn(fj, gj) n ≤ (cid:18)2 + 1 n(cid:19)κ(cid:18) 1 2n + 1(cid:19)Vκ(fj, gj) for all j, n ∈ N, and so, condition (2.6) in Theorem 1 is satisfied. Let f ∈ R(g) be the pointwise limit of a subsequence {fjm} of {fj}. Arguing as in the proof of Lemma 2(c), we get Vκ(f, g) ≤ lim inf m→∞ Vκ(fjm, gjm) ≤ lim sup Vκ(fj, gj) < ∞. (cid:3) j→∞ Remark 5. Since Theorem 1 is an extension of results from [11, 12], it also contains as particular cases all pointwise selection principles based on various notions of generalized variation. These principles may be further generalized in the spirit of Theorem 4 replacing the increment f (Ii) = d(f (si), f (ti)) applied in [11] -- [13] by the joint increment (f, g)(Ii) from (2.1). Finally, it is worth mentioning that the joint modulus of variation (2.4), defined by means of (2.1), plays an important role in the extension of a result from [17] to metric space valued functions: 20 Theorem 5. Given ∅ 6= T ⊂ R and a metric space (M, d), let {fj} ⊂ M T be a pointwise precompact sequence of functions such that lim N→∞ sup j,k≥N νn(fj, fk) = o(n). Then, there is a subsequence of {fj}, which converges pointwise on T . Taking into account Lemmas 1 and 2, the proof of this theorem follows the same lines as the ones given in [17, Theorem 1] (where (M, d, +) is a metric semigroup and νn(fj, fk) is defined via (3.1)), and so, it is omitted. Note that the limit of a pointwise convergent subsequence of {fj} in Theorem 5 may be a non-regulated function. For more details, examples and relations with previously known 'regular' and 'irregular' versions of pointwise selection principles from [23, 40, 42] we refer to [14, Section 5.2], [17, 18, 34]. 7. Conclusions In the context of functions of one real variable taking values in a metric space, we have presented a pointwise selection principle, which can be ap- plied to arbitrary (regulated and non-regulated) sequences of functions. It is based on notions of joint increment and joint modulus of variation, the latter being a nondecreasing sequence of pseudometrics on the appropriate product space. Our selection principle extends the classical Helly Theorem and contains as particular cases many selection theorems based on various notions of generalized variation of functions. In contrast to previously es- tablished selection principles, the main assumption in our theorem on the limit superior is 'almost necessary', and it is shown by examples that in a certain sense this assumption is sharp. The notion of conditionally regulated functions explains the limitations of previously known selection results. References [1] V. Barbu, Th. Precupanu, Convexity and Optimization in Banach Spaces, second ed., Reidel, Dordrecht, 1986. [2] S. A. Belov, V. V. Chistyakov, A selection principle for mappings of bounded variation, J. Math. Anal. Appl. 249 (2) (2000) 351 -- 366; and comments on this paper: J. Math. Anal. Appl. 278 (1) (2003) 250 -- 251. 21 [3] P. C. Bhakta, On functions of bounded variation relative to a set, J. Aus- tral. Math. Soc. 13 (1972) 313 -- 322. [4] Z. A. Chanturiya, The modulus of variation of a function and its appli- cation in the theory of Fourier series, Soviet Math. Dokl. 15 (1974) 67 -- 71. [5] Z. A. Chanturiya, Absolute convergence of Fourier series, Math. Notes 18 (1975) 695 -- 700. [6] V. V. Chistyakov, On mappings of bounded variation, J. Dynam. Control Systems 3 (2) (1997) 261 -- 289. [7] V. V. Chistyakov, On the theory of multivalued mappings of bounded variation of one real variable, Sb. Math. 189 (5 -- 6) (1998) 797 -- 819. [8] V. V. Chistyakov, Mappings of bounded variation with values in a metric space: generalizations, in: Pontryagin Conference, 2, Nonsmooth Anal- ysis and Optimization (Moscow, 1998). J. Math. Sci. (New York) 100 (6) (2000) 2700 -- 2715. [9] V. V. Chistyakov, On multi-valued mappings of finite generalized varia- tion, Math. Notes 71 (3 -- 4) (2002) 556 -- 575. [10] V. V. Chistyakov, Selections of bounded variation, J. Appl. Anal. 10 (1) (2004) 1 -- 82. [11] V. V. Chistyakov, The optimal form of selection principles for functions of a real variable, J. Math. Anal. Appl. 310 (2) (2005) 609 -- 625. [12] V. V. Chistyakov, A selection principle for functions of a real variable, Atti Sem. Mat. Fis. Univ. Modena e Reggio Emilia 53 (1) (2005) 25 -- 43. [13] V. V. Chistyakov, A pointwise selection principle for functions of a real variable with values in a uniform space, Siberian Adv. Math. 16 (3) (2006) 15 -- 41. [14] V. V. Chistyakov, Metric Modular Spaces: Theory and Applications, SpringerBriefs in Mathematics, Springer, Switzerland, 2015. [15] V. V. Chistyakov, O. E. Galkin, On maps of bounded p-variation with p > 1, Positivity 2 (1) (1998) 19 -- 45. [16] V. V. Chistyakov, O. E. Galkin, Mappings of bounded Φ-variation with arbitrary function Φ, J. Dynam. Control Systems 4 (2) (1998) 217 -- 247. [17] V. V. Chistyakov, C. Maniscalco, A pointwise selection principle for met- ric semigroup valued functions, J. Math. Anal. Appl. 341 (1) (2008) 613 -- 625. 22 [18] V. V. Chistyakov, C. Maniscalco, Yu. V. Tretyachenko, Variants of a se- lection principle for sequences of regulated and non-regulated functions, in: L. De Carli, K. Kazarian and M. Milman (Eds.), Topics in Clas- sical Analysis and Applications in Honor of Daniel Waterman, World Scientific Publishing, Hackensack, NJ, 2008, pp. 45 -- 72. [19] V. V. Chistyakov, A. Nowak, Regular Carath´eodory-type selectors under no convexity assumptions, J. Funct. Anal. 225 (2) (2005) 247 -- 262. [20] V. V. Chistyakov, D. Repovs, Selections of bounded variation under the excess restrictions, J. Math. Anal. Appl. 331 (2) (2007) 873 -- 885. [21] V. V. Chistyakov, A. Rychlewicz, On the extension and generation of set-valued mappings of bounded variation, Studia Math. 153 (3) (2002) 235 -- 247. [22] D. S. Cyphert, J. A. Kelingos, The decomposition of functions of bound- ed κ-variation into differences of κ-decreasing functions, Studia Math. 81 (2) (1985) 185 -- 195. [23] L. Di Piazza, C. Maniscalco, Selection theorems, based on generalized variation and oscillation, Rend. Circ. Mat. Palermo, Ser. II, 35 (3) (1986) 386 -- 396. [24] R. Duda, On convex metric spaces. III, Fund. Math. 51 (1962) 23 -- 33. [25] D. Frankov´a, Regulated functions, Math. Bohem. 116 (1) (1991) 20 -- 59. [26] S. Gni lka, On the generalized Helly's theorem, Funct. Approx. Com- ment. Math. 4 (1976) 109 -- 112. [27] K. Goebel, W. A. Kirk, Topics in Metric Fixed Point Theory, Cambridge Stud. Adv. Math., vol. 28, Cambridge Univ. Press, Cambridge, 1990. [28] C. Goffman, T. Nishiura, D. Waterman, Homeomorphisms in Analysis, Math. Surveys Monogr., vol. 54, Amer. Math. Soc., Providence, Rhode Island, 1997. [29] E. Helly, Uber lineare Funktionaloperationen, Sitzungsber. Naturwiss. Kl. Kaiserlichen Akad. Wiss. Wien 121 (1912) 265 -- 297. [30] H. Hermes, On continuous and measurable selections and the existence of solutions of generalized differential equations, Proc. Amer. Math. Soc. 29 (3) (1971) 535 -- 542. [31] T. H. Hildebrandt, Introduction to the Theory of Integration, Academic Press, New York, London, 1963. 23 [32] B. Korenblum, An extension of the Nevanlinna theory, Acta Math. 135 (1) (1975) 187 -- 219. [33] B. Korenblum, A generalization of two classical convergence tests for Fourier series, and some new Banach spaces of functions, Bull. Amer. Math. Soc. (N. S.) 9 (2) (1983) 215 -- 218. [34] C. Maniscalco, A comparison of three recent selection theorems, Math. Bohem. 132 (2) (2007) 177 -- 183. [35] K. Menger, Untersuchungen uber allgemeine Metrik, Math. Ann. 100 (1928) 75 -- 163. [36] B. S. Mordukhovich, Approximation Methods in the Problems of Opti- mization and Control, Nauka, Moscow, 1988 (in Russian). [37] P. Muldowney, A Modern Theory of Random Variation: With Appli- cations in Stochastic Calculus, Financial Mathematics, and Feynman Integration, John Wiley & Sons, Hoboken, New Jersey, 2012. [38] J. Musielak, W. Orlicz, On generalized variations (I), Studia Math. 18 (1959) 11 -- 41. [39] I. P. Natanson, Theory of Functions of a Real Variable, third ed., Nauka, Moscow, 1974 (in Russian). English translation: Ungar, New York, 1965. [40] K. Schrader, A generalization of the Helly selection theorem, Bull. Amer. Math. Soc. 78 (3) (1972) 415 -- 419. [41] M. Schramm, Functions of Φ-bounded variation and Riemann-Stieltjes integration, Trans. Amer. Math. Soc. 287 (1985) 49 -- 63. [42] Yu. V. Tret'yachenko, V. V. Chistyakov, The selection principle for pointwise bounded sequences of functions, Math. Notes 84 (4) (2008) 396 -- 406. [43] D. Waterman, On Λ-bounded variation, Studia Math. 57 (1976) 33 -- 45. 24
1009.3583
1
1009
2010-09-18T21:09:50
A note on Mahler's conjecture
[ "math.FA" ]
Let $K$ be a convex body in $\mathbb{R}^n$ with Santal\'o point at 0\. We show that if $K$ has a point on the boundary with positive generalized Gau{\ss} curvature, then the volume product $|K| |K^\circ|$ is not minimal. This means that a body with minimal volume product has Gau{\ss} curvature equal to 0 almost everywhere and thus suggests strongly that a minimal body is a polytope.
math.FA
math
A note on Mahler's conjecture Shlomo Reisner, Carsten Schutt and Elisabeth M. Werner ∗† Abstract Let K be a convex body in Rn with Santal´o point at 0. We show that if K has a point on the boundary with positive generalized Gauss curvature, then the volume product KK ◦ is not minimal. This means that a body with minimal volume product has Gauss curvature equal to 0 almost everywhere and thus suggests strongly that a minimal body is a polytope. 1 Introduction A convex body K in Rn is a compact, convex set with nonempty interior. The polar body K z with respect to an interior point z of K is K z = {y ∀x ∈ K : hy, x − zi ≤ 1}. There is a unique point z ∈ K such that the volume product KK ◦ is minimal. This point is called the Santal´o point s(K). The Blaschke-Santal´o inequality asserts that the maximum of the volume product KK ◦ is attained for all ellipsoids and only for ellipsoids [2, 16, 11]. Thus the convex body for which the maximum is attained is unique up to affine transforms. On the other hand, it is an open problem for which convex bodies the minimum is attained. It is conjectured that the minimum is attained for the simplex. The class of centrally symmetric convex bodies is of particular importance. Mahler conjectured [8, 9] that the minimum in this class is attained for the cube and its polar body, the cross-polytope. If so, the minimum would also be attained by "mixtures" of the cube and the cross-polytope, sometimes called Hanner-Lima bodies. Those are not affine images of the cube or the cross- polytope. Thus, in the class of centrally symmetric convex bodies the minimum is not attained for a unique convex body (up to affine transforms). The first breakthrough towards Mahler's conjecture is the inequality of Bourgain-Milman [3]. They proved that for centrally symmetric convex bodies (cid:16) c n(cid:17)n ≤ KK ◦. Keywords: convex bodies, volume product, Mahler's conjecture. 2000 Mathematics Subject Classification: 52A20. ∗Part of the work was done while the three authors visited the Renyi Institute of Mathematics of the Hungarian Academy of Sciences, in the summer of 2008 and during a workshop at the American Institute of Mathematics in Palo Alto in August of 2010. They are indebted to both Institutes. †E. Werner has been partially supported by an NSF grant, a FRG-NSF grant and a BSF grant. 1 This inequality has recently been reproved with completely different methods by Kuperberg and Nazarov [7, 12]. Their proofs also give better constants. For special classes like zonoids and unconditional bodies Mahler's conjecture has been verified [13, 5, 15, 10, 14]. The inequality of Bourgain-Milman has many applications in various fields of mathe- matics: geometry of numbers, Banach space theory, convex geometry, theoretical computer science. Despite great efforts, a proof of Mahler's conjecture seems still elusive. It is not even known whether a convex body for which the minimum is attained must be a polytope. A result in this direction has been proved by Stancu [20]. It is shown there that if K is of class C 2 with strictly positive Gauss curvature everywhere, then the volume product of K can not be a local minimum. In this paper we show that a minimal body can not have even a single point with positive generalized curvature. By a result of Alexandrov, Busemann and Feller [1, 4] the generalized curvature exists almost everywhere. Therefore, our result implies that a minimal body has almost everywhere curvature equal to 0 and thus suggests strongly that a minimal body is a polytope. We now introduce the concept of generalized curvature. Let U be a convex, open subset of Rn and let f : U → R be a convex function. df (x) ∈ Rn is called subdifferential at the point x0 ∈ U, if we have for all x ∈ U f (x0) + hdf (x0), x − x0i ≤ f (x). A convex function has a subdifferential at every point and it is differentiable at a point if and only if the subdifferential is unique. Let U be an open, convex subset in Rn and f : U → R a convex function. f is said to be twice differentiable in a generalized sense in x0 ∈ U, if there is a linear map d2f (x0) : Rn → Rn and a neighborhood U(x0) ⊆ U such that we have for all x ∈ U(x0) and for all subdifferentials df (x) kdf (x) − df (x0) − d2f (x0)(x − x0)k ≤ Θ(kx − x0k)kx − x0k. Here, k k is the standard Euclidean norm on Rn and Θ is a monotone function with limt→0 Θ(t) = 0. d2f (x0) is called (generalized) Hesse-matrix. If f (0) = 0 and df (0) = 0 then we call the set {x ∈ Rnxtd2f (0)x = 1} the indicatrix of Dupin at 0. Since f is convex, this set is an ellipsoid or a cylinder with a base that is an ellipsoid of lower dimension. The eigenvalues of d2f (0) are called generalized principal curvatures and their product is called the generalized Gauss-Kronecker curvature κ. It will always be this generalized Gauss curvature that we mean throughout the rest of the paper though we may occasionally just call it Gauss curvature. Geometrically the eigenvalues of d2f (0) that are different from 0 are the lengths of the principal axes of the indicatrix raised to the power (−2). To define the generalized Gauss curvature κ(x) of a convex body K at a boundary point x with unique outer normal NK(x), if it exists, we translate and rotate K so that we may 2 assume that x = 0 and NK(x) = −en. κ(x) is then defined as the Gauss curvature of the function f : Rn−1 → R whose graph in the neighborhood of 0 is ∂K. We further denote by H(x, ξ) the hyperplane through x and orthogonal to ξ. H −(x, ξ) and H +(x, ξ) are the two half spaces determined by H(x, ξ). In particular, for ∆ > 0, a convex body K and x ∈ ∂K, the boundary of K, with a unique outer normal NK(x) is the hyperplane through x − ∆NK(x) with normal NK(x). H + (x − ∆NK(x), NK(x)) de- notes the halfspace determined by H (x − ∆NK(x), NK(x)) that does not contain x. H (x − ∆NK(x), NK(x)) We construct two new bodies, Kx(∆), by cutting off a cap and K x(∆) by Kx(∆) = K ∩ H + (x − ∆NK(x), NK(x)) , K x(∆) = co[K, x + ∆NK(x)]. 2 The main theorem Theorem 1. Let K be a convex body in Rn and suppose that there is a point in the boundary of K where the generalized Gauss curvature exists and is not 0. Then the volume product KK s(K) is not a local minimum. Moreover, if K is centrally symmetric with center 0 then, under the above assumption, the volume product KK ◦ is not a local minimum in the class of 0-symmetric convex bodies. In order to prove Theorem 1, we present the following proposition. Proposition 2. Let K be a convex body in Rn whose Santal´o point is at the origin. Suppose that there is a point x in the boundary of K where the generalized Gauss curvature exists and is not 0. Then there exists ∆ > 0 such that or Kx(∆) (Kx(∆))◦ < KK ◦. K x(∆) (K x(∆))◦ < KK ◦. For the proof of Proposition 2 we need several lemmas from [18] and [19]. We refer to [18] and [19] for the proofs. In particular, part (ii) of this lemma can be found in [19] as Lemma 12. Lemma 3. [19] Let K be a convex body in Rn. Let T : Rn → Rn be a linear, invertible map. (i) The normal at T (x) is (T −1)t(NK(x))k(T −1)t(NK(x))k−1. (ii) Suppose that the generalized Gauss-Kronecker curvature κ exists in x ∈ ∂K. Then the generalized Gauss-Kronecker curvature κ exists in T (x) ∈ ∂T (K) and κ(x) = k(T −1)t(NK(x))kn+1 det(T )2κ(T (x)). 3 The next two lemmas are well known. See e.g. [18]. Lemma 4. [18] Let U be an open, convex subset of Rn and 0 ∈ U. Suppose that f : U → R is twice differentiable in the generalized sense at 0 and that f (0) = 0 and df (0) = 0. Suppose that the indicatrix of Dupin at 0 is an ellipsoid. Then there is a monotone, increasing function ψ : [0, 1] → [1,∞) with lims→0 ψ(s) = 1 such that xtd2f (0)x ≤ 2s ψ(s)(cid:27) (cid:26)(x, s)(cid:12)(cid:12)(cid:12)(cid:12) ⊆ {(x, s)f (x) ≤ s} ⊆ {(x, s)xtd2f (0)x ≤ 2sψ(s)}. Lemma 5. [18] Let K be a convex body in Rn with 0 ∈ ∂K and N(0) = −en. Suppose that the indicatrix of Dupin at 0 is an ellipsoid. Suppose that the principal axes biei of the indicatrix are multiples of the unit vectors ei, i = 1, . . . , n − 1. Let E be the n-dimensional ellipsoid 2 n−1(cid:17)2 ≤ n−1 Yi=1 bi! . 2 n−1  n−1 2 i=1 bi) + (cid:16)xn −(cid:0)Qn−1 i=1 bi(cid:1) (Qn−1 , t(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) xn−1 φ(t) (cid:26)(cid:18) x1 , . . . , x ∈ E, xn = t(cid:27) φ(t) ⊆ K ∩ H((0, . . . , 0, t), N(0)) ⊆ {(φ(t)x1, . . . , φ(t)xn−1, t)x ∈ E, xn = t} . E =  x ∈ Rn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) x2 i b2 i n−1 Xi=1 Then there is an increasing, continuous function φ : [0,∞) → [1,∞) with φ(0) = 1 such that we have for all t We call E the standard approximating ellipsoid . Let us denote the lengths of the principal axes of the indicatrix of Dupin by bi, i = 1, . . . , n − 1. Then the lengths ai, i = 1, . . . , n of the principal axes of the standard approxi- mating ellipsoid E are ai = bi n−1 Yi=1 1 n−1 bi! i = 1, . . . , n − 1 and an = n−1 Yi=1 2 n−1 . bi! (1) This follows immediately from Lemma 5. For the generalized Gauss-Kronecker curvature we get an a2 i . (2) n−1 Yi=1 This follows as the generalized Gauss-Kronecker curvature equals the product of the eigen- values of the generalized Hesse matrix. The eigenvalues are b−2 n−1 Yi=1 b−2 i = n−1 Yi=1 bi!2 n−1 Yi=1 bk!  bi n−1 Yk=1 4 −2 1 n−1  i n−1 , i = 1, . . . , n − 1. Thus Yi=1 an a2 i = . In particular, if the indicatrix of Dupin is a sphere of radius √ρ then the standard approxi- mating ellipsoid is a Euclidean ball of radius ρ. We consider the map T : Rn → Rn bi! a1 n−1 Yi=1 T (x) =  x1 2 n−1 , . . . , 2 n−1 xn−1 an−1 n−1 Yi=1 bi! , xn  . (3) This transforms the standard approximating ellipsoid E into a Euclidean ball T (E) with i=1 bi)2/(n−1). This is obvious since the principal axes of the standard approx- imating ellipsoid are given by (1). The map T is volume preserving. radius r = (Qn−1 Lemma 6. Let K be a convex body in Rn with 0 as an interior point. Suppose that the generalized Gauss curvature of ∂K at x0 exists and that the indicatrix is an ellipsoid. Then there is an invertible linear transformation T such that (i) NT (K)(T (x0)) = T (x0) kT (x0)k (ii) The indicatrix of Dupin at T (x0) is a Euclidean ball. (iii) T (K) = T (K)◦ (iv) kT (x0)k = 1. Proof. (i) We first show that there is a linear map T1 such that NT1(K)(T1(x0)) = T1(x0) Let ei, 1 ≤ i ≤ n be an orthonormal basis of Rn such that en = NK(x0). Define T1 by kT1(x0)k . T1 n−1 Xi=1 tiei + tnx0! = n−1 Xi=1 tiei + tnhx0, enien T1 is well-defined since x0 and e1, . . . , en−1 are linearly independent. Indeed, as hen, x0i > 0, x0 /∈ e⊥ = span{e1, . . . , en−1}. Moreover, T1(x0) = hx0, enien T1(x0) kT1(x0)k = en (4) and thus By Lemma 3, the outer normal at T1(x0) is (T −1 1 )t(NK(x0))k(T −1 1 )t(NK(x0))k−1. Then for all 1 ≤ i ≤ n − 1 h(T −1 1 )t(NK(x0)), eii = hNK(x0), T −1 1 eii = hNK(x0), eii = hen, eii = 0. Hence (T −1 k(T −1 1 )t(NK(x0)) 1 )t(NK(x0))k = ±en 5 Since hx0, eni > 0 and since 0 is an interior point, it is +en. Together with (4), this shows that NT1(K)(T1(x0)) = en = T1(x0) kT1(x0)k . (ii) Put x1 = T1(x0) and K1 = T1(K). By Lemma 3, the curvature κ(x1) at x1 ∈ ∂K1 exists and is positive. For 1 ≤ i ≤ n − 1, let bi be the principal curvatures and ai be the principal axes of the standard approximating ellipsoid in x1. Let T2 : Rn → Rn , ξn  . T2(x) =  an−1 n−1 Yi=1 a1 n−1 Yi=1 transforms the indicatrix of Dupin at x1 into an n − 1-dimensional Euclidean ball and the standard approximating ellipsoid E into a n-dimensional Euclidean ball T2(E) with radius r = (Qn−1 Property (i) of the lemma is preserved: i=1 bi)2/(n−1). bi! bi! , . . . , ξn−1 (5) ξ1 2 n−1 2 n−1 NT2(K1)(T2(x1)) = NK1(x1) = en. Indeed, by Lemma 3 NT2(K1)(T2(x1)) = and thus for all 1 ≤ i ≤ n − 1 (T −1 k(T −1 2 )t(NK1(x1)) 2 )t(NK1(x1))k . h(T −1 2 )t(NK1(x1)), eii = hNK1(x1), T −1 2 (ei)i =*en, a1 n−1 Yi=1 n−1 bi!− 2 ei+ = 0. (iii) It is enough to apply a multiple αI of the identity. (iv) We apply the map T3 with where T3(ξ) = (λξ1, . . . , λξn−1, λ−n+1ξn) λ = (αhx0, eni)1/(n−1) Properties (i) and (ii) of the lemma are preserved and, as det(T3) = 1, Property (iii) as well. Finally, we let T (K) = T3(αT2(K1)). (cid:3) Lemma 7. Let K be a convex body in Rn such that ∂K is twice differentiable in the gener- alized sense at x. Suppose that kxk = 1, NK(x) = x, and the indicatrix of Dupin at x is a Euclidean sphere with radius r. Then x ∈ ∂K ◦ and for all 0 < ǫ < min{r, 1 r} there is ∆ > 0 such that Bn 2 (cid:18)x −(cid:18)1 r − ǫ(cid:19) NK ◦(x), 1 r − ǫ(cid:19) ∩ H −(x − ∆NK ◦(x), NK ◦(x)) 6 ⊆ K ◦ ∩ H −(x − ∆NK ◦(x), NK ◦(x)) + ǫ(cid:19) NK ◦(x), 1 r 2 (cid:18)x −(cid:18)1 ⊆ Bn r + ǫ(cid:19) ∩ H −(x − ∆NK ◦(x), NK ◦(x)) Proof. Without loss of generality we can assume that x = NK(x) = en. Clearly then x ∈ ∂K ◦ and NK ◦(x) = x. Let 0 < ǫ < min{r, 1 r}. By Lemma 5, there exists ∆1 such that for all ∆ ≤ ∆1 Bn 2 ((1 − (r − ǫ))en, r − ǫ) ∩ H − ((1 − ∆)en, en) ⊆ K ∩ H − ((1 − ∆)en, en) ⊆ Bn 2 ((1 − (r + ǫ))en, r + ǫ) ∩ H − ((1 − ∆)en, en) . We construct two new convex bodies. and K1 = co(cid:2)K ∩ H + ((1 − ∆1)en, en) , Bn K2 = co(cid:2)K ∩ H + ((1 − ∆1)en, en) , Bn 2 ((1 − (r − ǫ))en, r − ǫ)(cid:3) 2 ((1 − (r + ǫ))en, r + ǫ)(cid:3) . Then K1 ⊆ K ⊆ K2 and there is ∆2 ≤ ∆1 such that for all ∆ ≤ ∆2 K1 ∩ H − ((1 − ∆)en, en) = Bn 2 ((1 − (r − ǫ))en, r − ǫ) ∩ H − ((1 − ∆)en, en) and K2 ∩ H − ((1 − ∆)en, en) = Bn 2 ((1 − (r + ǫ))en, r − ǫ) ∩ H − ((1 − ∆)en, en) . We now compute K 0 K1. K2 is done similarly. 1 and K 0 2 in a neighborhood of x = en. We show the computations for Let ∆ ≤ ∆2 and η be the normal of y ∈ ∂K1 ∩ H − ((1 − ∆)en, en). Then hy, ηi = r − ǫ + (1 − (r − ǫ)) hx, ηi (cid:28)y, r − ǫ + (1 − (r − ǫ)) hx, ηi(cid:29) = 1 r − ǫ + (1 − (r − ǫ))hx, ηi ∈ ∂K ◦ 1 ≤ ∆2 sufficiently small, we consider now a cap of K ◦ 1 . η η 1 and its base Therefore, and hence For ∆K ◦ K ◦ 1 ∩ H((1 − ∆K ◦ 1 )en, en) We compute the distance ρ of of the base (1 − ∆K ◦ 1 )en . Clearly, by Pythagoras η r−ǫ+(1−(r−ǫ))hx,ηi ∈ ∂K ◦ 1 ∩ H((1 − ∆K ◦ 1 )en, en) from the center ρ =s 1 1 )2 (r − ǫ + (1 − (r − ǫ))hx, ηi)2 − (1 − ∆K ◦ 7 Moreover, Hence Therefore or hx, ηi 1 = r − ǫ + [1 − (r − ǫ)]hx, ηi 1 − ∆K ◦ 1 ) (cid:18)r − ǫ + [1 − (r − ǫ)] hx, ηi(cid:19) = hx, ηi 1 )[1 − (r − ǫ)] − 1(cid:19) = −(r − ǫ)(1 − ∆K ◦ 1 ) 1 ) (r − ǫ)(1 − ∆K ◦ r − ǫ + [1 − (r − ǫ)] ∆K ◦ 1 (1 − ∆K ◦ hx, ηi(cid:18)(1 − ∆K ◦ hx, ηi = ρ =s(cid:0)r − ǫ + [1 − (r − ǫ)]∆K ◦ 1(cid:1)2 [1 − 2(r − ǫ)] (r − ǫ)2 ∆K ◦ 1 + ρ =s 2 r − ǫ (r − ǫ)2 ∆2 K ◦ 1 1 )2 − (1 − ∆K ◦ We compare this radius with the corresponding radius of the ball Bn The corresponding radius is r−2ǫ)en, 2 (cid:0)(1 − 1 1 r−2ǫ(cid:1). Therefore, provided that 1 K ◦ 1 r 2∆K ◦ r − 2ǫ − ∆2 ρ ≤r 2∆K ◦ r − 2ǫ − ∆2 1 K ◦ 1 ∆K ◦ 1 ≤ min(cid:26) 2ǫ (r − 2ǫ)(r − ǫ) 1 we get As K1 ⊆ K and consequently K ◦ ⊆ K ◦ , 2ǫ (r − 2ǫ)(2 + r − ǫ)(cid:27) . 1 2 (cid:18)(cid:18)1 − 1 )en, en) ⊆ Bn K ◦ ∩ H −((1 − ∆K ◦ r − 2ǫ(cid:19) en, Similarly, using K ⊆ K2, one shows (with a new ∆K ◦ r + 2ǫ(cid:19) ∩ H −((1 − ∆K ◦ r + 2ǫ(cid:19) en, 2 (cid:18)(cid:18)1 − Bn 1 1 1 r − 2ǫ(cid:19) ∩ H −((1 − ∆K ◦ 1 )en, en). 1 small enough if needed) that 1 )en, en) ⊆ K ◦ ∩ H −((1 − ∆K ◦ 1 )en, en). (cid:3) Lemma 7 implies that if x ∈ ∂K is a twice differentiable point (in the generalized sense), then the point y ∈ ∂K ◦ with hx, yi = 1, is also a twice differentiable point. Compare also [6]. 8 Corollary 8. Let K be a convex body in Rn with 0 as an interior point. Assume that ∂K is twice differentiable in the generalized sense at x and the indicatrix of Dupin at x is an ellipsoid (and not a cylinder with an ellipsoid as its base). Then ∂K ◦ is twice differentiable at the unique point ξ with hξ, xi = 1. Proof. By Lemma 6 we may assume that the indicatrix is a Euclidean ball and that NK(x) = x. By Lemma 7 the statement follows. (cid:3) The next lemma is also well known (see e.g. [17]). Lemma 9. Let K be a convex body in Rn and suppose that the indicatrix of Dupin at x ∈ ∂K exists and is a Euclidean ball of radius r > 0. Let C(r, ∆) be the cap at x of height ∆. Then r (cid:19) C(r, ∆) = g(cid:18) ∆ n+1 2 1 n + 1 2 where limt→0 g(t) = 1. n+1 2 voln−1(Bn−1 2 )∆ n+1 2 r n−1 2 Remark. The conclusions of Lemma 9 also hold if instead of the existence of the indicatrix, we assume the following: Let x ∈ ∂K and suppose that there is r > 0 such that for all ǫ > 0 there is a ∆ǫ such that for all ∆ with 0 < ∆ ≤ ∆ǫ Bn 2 (x − (r − ǫ) NK(x), r − ǫ) ∩ H −(x − ∆NK(x), NK(x)) ⊆ K ∩ H −(x − ∆NK(x), NK(x)) ⊆ Bn 2 (x − (r + ǫ) NK(x), r + ǫ) ∩ H −(x − ∆NK(x), NK(x)). (6) The next lemma is from [21]. There it was assumed that the indicatrix of Dupin at x ∈ ∂K exists and is a Euclidean ball of radius r > 0. However, what was actually used in the proof, were the assumptions (6) of the above Remark. Lemma 10. [21] Let K be a convex body in Rn. Let x ∈ ∂K and suppose that there is r > 0 such that for all ǫ > 0 there is a ∆ǫ such that for all ∆ with 0 < ∆ ≤ ∆ǫ, (6) holds. Then, if ∆ǫ is small enough, we have for 0 < ∆ < ∆ǫ n+1 2 2 n(n + 1)Bn−1 2 n+1 2 r n−1 2 ∆ n−1 2 ǫ r(cid:17) (1 − c ǫ) − n(cid:16)1 + ǫ r(cid:17) n−1 2 (1 + c ǫ) h(cid:18) ∆ r + ǫ(cid:19)n+1) ×((n + 1)(cid:16)1 − K x(∆) \ K(cid:12)(cid:12)(cid:12)(cid:12) 2 n(n + 1)Bn−1 ∆ n+1 2 2 ≤(cid:12)(cid:12)(cid:12)(cid:12) ≤ n+1 2 r n−1 2 9 ×((n + 1)(cid:16)1 + ǫ r(cid:17) n−1 2 (1 + c ǫ) − n(cid:16)1 − ǫ r(cid:17) n−1 2 (1 − c ǫ) h(cid:18) ∆ r + ǫ(cid:19)n+1) . where c is a constant and limt→0 h(t) = 1. Proof of Theorem 1. By assumption there is a point x ∈ ∂K at which ∂K is twice differentiable in the generalized sense. By Lemma 6 we may assume that kxk = 1 and x = NK(x). Moreover, all principal curvature radii at x are equal to r. By Lemma 7, x ∈ ∂K ◦, K ◦ is twice differentiable at x and all principal curvature radii are equal to 1 r . The dual body to Kx(∆) is K x( ∆ 1−∆ ). By Lemma 9, n+1 2 n+1 2 2 n + 1 voln−1(Bn−1 2 )∆ n+1 2 r n−1 2 . Kx(∆) ≤ K − g(cid:18)∆ r (cid:19) 1 − ∆(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) K x(cid:18) ∆ ×((n + 1) (1 + rǫ) ≤ K ◦ + n−1 By Lemma 10 Kx(∆)◦ =(cid:12)(cid:12)(cid:12)(cid:12) n+1 2 2 n(n + 1)Bn−1 2 2 (1 + c ǫ) − n (1 − rǫ) 2 n+1 2 r− n−1 ∆ 2 (1 − c ǫ) h(cid:18) ∆ n−1 1 r + ǫ(cid:19)n+1) . It follows that Kx(∆)Kx(∆)◦ ≤ KK ◦ + K n+1 2 2 n(n + 1)Bn−1 2 n+1 2 r− n−1 2 ∆ ×((n + 1) (1 + rǫ) −K ◦g(cid:18)∆ r (cid:19) n+1 2 1 n + 1 n−1 2 (1 + c ǫ) − n (1 − rǫ) 2 n+1 2 voln−1(Bn−1 )∆ 2 r + ǫ(cid:19)n+1) 1 n−1 2 (1 − c ǫ) h(cid:18) ∆ 2 (r − ǫ) 2 . n−1 n+1 Therefore we have provided that Kx(∆)Kx(∆)◦ < KK ◦, K((n + 1) (1 + rǫ) r(cid:19) < nK ◦ g(cid:18)∆ n+1 2 n−1 2 (1 + c ǫ) − n (1 − rǫ) rn−1(cid:16)1 − r(cid:17) n−1 ǫ . 2 n−1 2 (1 − c ǫ) h(cid:18) ∆ r + ǫ(cid:19)n+1) 1 (7) Now we interchange the roles of K and K ◦. We cut off a cap from K ◦ and apply the remark following Lemma 9. Then the inequality analogous to (7) will be K ◦((n + 1)(cid:16)1 + ǫ r(cid:17) n−1 2 (1 + c ǫ) − n(cid:16)1 − ǫ r(cid:17) n−1 2 (1 − c ǫ) h(cid:18) ∆ r + ǫ(cid:19)n+1) 10 < nKg (r∆) n+1 2 r−(n−1)(1 − rǫ) n−1 2 . (8) Thus the theorem is proved provided that one of the inequalities (7) or (8) holds. Suppose both inequalities do not hold. Then 2 n+1 n−1 n−1 K((n + 1) (1 + rǫ) r (cid:19) ≥ nK ◦ g(cid:18) ∆ ≥ n2K 2 (1 + c ǫ) − n (1 − rǫ) rn−1(cid:16)1 − r(cid:17) g(cid:0) ∆ 2 (1 − ǫ r − rǫ + ǫ2) r(cid:1) n(n + 1)(cid:0)1 + ǫ 2 (1 + c ǫ) − n(cid:0)1 − ǫ r(cid:1) r(cid:1) n−1 n−1 n+1 ǫ 2 We can choose ǫ so small that (n + 1) (1 + rǫ) n−1 2 (1 + c ǫ) − n (1 − rǫ) n−1 and n−1 2 (n + 1)(cid:16)1 + ǫ r(cid:17) (1 + c ǫ) − n(cid:16)1 − ǫ r(cid:17) Moreover, we can choose ǫ so small that 2 (1 − c ǫ) h(cid:18) ∆ r + ǫ(cid:19)n+1) 1 . r+ǫ(cid:1)n+1o n−1 2 g (r∆) n+1 2 n−1 2 (1 − c ǫ) h(cid:0) ∆ r + ǫ(cid:19)n+1 r + ǫ(cid:19)n+1 2 (1 − c ǫ) h(cid:18) ∆ (1 − c ǫ) h(cid:18) ∆ 1 2 n−1 ≤ 2 ≤ 2. Therefore ǫ r − rǫ + ǫ2(cid:17) (cid:16)1 − r (cid:19) 4 ≥ n2g(cid:18)∆ n−1 2 1 2 ≥ n+1 2 g (r∆) n+1 2 Since limt→0 g(t) = 1, this gives a contradiction. The extension of the proof needed in order to prove the symmetric case is obvious. (cid:3) References [1] A.D. Alexandrov, Almost everywhere existence of the second differential of a con- vex function and some properties of convex surfaces connected with it, Uchen. Zap. Leningrad Gos. Univ. Mat. Ser. 6 (1939), 3-35. [2] W. Blaschke, Uber affine Geometrie VII. Neue Extremeigenschaften von Ellipse und Ellipsoid , Leipz. Ber., 69 (1917), 306-318. [3] J. Bourgain and V. D. Milman, New volume ratio properties for convex symmetric bodies in Rn, Invent. Math. 88 (1987), 319-340. [4] H. Busemann and W. Feller, Krummungseigenschaften konvexer Flachen, Acta Math. 66 (1936), 1-47. 11 [5] Y. Gordon, M. Meyer and S. Reisner, Zonoids with minimal volume product -- a new proof , Proc. Amer. Math. Soc. 104 (1988), 273-276. [6] D. Hug, Curvature Relations and Affine Surface Area for a General Convex Body and its Polar, Results in Mathematics V. 29 (1996), 233-248. [7] G. Kuperberg, From the Mahler Conjecture to Gauss Linking Integrals, Geometric And Functional Analysis 18 (2008), 870-892. [8] K. Mahler, Ein Minimalproblem fur konvexe Polygone, Mathematica (Zutphen) B 7 (1939), 118-127. [9] K. Mahler, Ein Ubertragungsprinzip fur konvexe Korper , Casopis Pest. Mat. Fys. 68 (1939), 93-102. [10] M. Meyer, Une caract´erisation volumique de certains espaces norm´es, Israel J. Math. 55 (1986), 317-326. [11] M. Meyer and A. Pajor, On the Blaschke-Santal´o inequality, Arch. Math. 55 (1990), 82-93. [12] F. Nazarov, The Hormander proof of the Bourgain-Milman theorem, preprint, 2009. [13] S. Reisner, Zonoids with minimal volume-product, Math. Z. 192 (1986), 339-346. [14] S. Reisner, Minimal volume product in Banach spaces with a 1-unconditional basis, J. London Math. Soc. 36 (1987), 126-136. [15] J. Saint-Raymond, Sur le volume des corps convexes sym´etriques, S´eminaire d'Initiation `a l'Analyse, 1980-1981, Universit´e PARIS VI, Paris 1981. [16] L. A. Santal´o, Un invariante afin para los cuerpos convexos del espacio de n dimen- siones, Portugal. Math. 8 (1949), 155-161. [17] C. Schutt and E. Werner, The convex floating body, Math. Scand. 66 (1990), 275-290. [18] C. Schutt and E. Werner, Random polytopes of points chosen from the boundary of a convex body GAFA Seminar Notes, Lecture Notes in Math., 1807, Springer-Verlag, (2002), 241-422. [19] C. Schutt and E. Werner, Surface bodies and p-affine surface area, Advances in Math- ematics 187 (2004), 98-145. [20] A. Stancu, Two volume product inequalities and their applications, Canad. Math. Bull. 52 (2009), 464-472. [21] E. Werner, Illumination bodies and affine surface area, Studia Mathematica 110 (1994), 257-269. 12 Shlomo Reisner Department of Mathematics University of Haifa Haifa, 31905 Israel [email protected] Carsten Schutt Mathematisches Seminar Christian-Albrechts Universitat D-24098 Kiel, Ludewig-Meyn Strasse 4, Germany [email protected] Elisabeth Werner Department of Mathematics Case Western Reserve University Cleveland, Ohio 44106, U. S. A. [email protected] Universit´e de Lille 1 UFR de Math´ematique 59655 Villeneuve d'Ascq, France 13
1108.2221
2
1108
2013-02-19T12:25:39
Linear semigroups with coarsely dense orbits
[ "math.FA" ]
Let $S$ be a finitely generated abelian semigroup of invertible linear operators on a finite dimensional real or complex vector space $V$. We show that every coarsely dense orbit of $S$ is actually dense in $V$. More generally, if the orbit contains a coarsely dense subset of some open cone $C$ in $V$ then the closure of the orbit contains the closure of $C$. In the complex case the orbit is then actually dense in $V$. For the real case we give precise information about the possible cases for the closure of the orbit.
math.FA
math
LINEAR SEMIGROUPS WITH COARSELY DENSE ORBITS H. ABELS AND A. MANOUSSOS Abstract. Let S be a finitely generated abelian semigroup of invertible linear operators on a finite dimensional real or complex vector space V . We show that every coarsely dense orbit of S is actually dense in V . More generally, if the orbit contains a coarsely dense subset of some open cone C in V then the closure of the orbit contains the closure of C. In the complex case the orbit is then actually dense in V . For the real case we give precise information about the possible cases for the closure of the orbit. 1. Introduction Recall that a subset Y of a metric space (X, d) is called D-dense in X if the open balls of radius D with centers at points of Y cover X. The subset Y is called coarsely dense in X if there is a positive number D such that Y is D-dense in X. Theorem 1.1. Let V be a finite dimensional real or complex vector space endowed with a norm. Let S be a subsemigroup of GL(V ) gener- ated by a finite set of commuting elements. Then every coarsely dense orbit of S in V is actually dense. We show a more general result, theorem 3.1, as follows. Suppose an orbit O of S has a subset which is coarsely dense in some open cone C in V . Then we show that the closure of O is a cone which contains C. We also give detailed information about the structure of the orbits of the closure S of S. The topology we refer to without further comment here and throughout the whole paper will be the Euclidean topology. We will sometimes also use the Zariski topology. But then we will be more specific and use expressions like Zariski-dense etc. Under our 2010 Mathematics Subject Classification. Primary 47D03 20G20; Secondary 22E10, 22E15, 37C85, 47A16. Key words and phrases. Coarsely dense orbit, D-dense orbit, hypercyclic semi- group of matrices. During this research the second author was fully supported by SFB 701 "Spek- trale Strukturen und Topologische Methoden in der Mathematik" at the University of Bielefeld, Germany. 1 2 H. ABELS AND A. MANOUSSOS more general hypotheses we prove furthermore the following claim, see theorem 4.1. The closure S of S in GL(V ) is actually a group. There are only finitely many maximal S-invariant vector subspaces of V . Let U be the complement of their union. Then U has a finite number of connected components which are open cones in V . For every v ∈ U the orbit S · v is a union of connected components of U, in particular an open cone. The original motivation for work on this paper came from the theory of linear operators. N.S. Feldman [4] initiated the study of linear op- erators T : X → X on a Banach space X which have a coarsely dense orbit. Such orbits arise naturally when perturbing vectors with dense orbits. For instance, if the orbit of a vector x ∈ X is dense in X and y ∈ X is a vector with a bounded orbit then the orbit O(x + y, T ) is coarsely dense. In the same paper, Feldman showed that it is possible for an orbit to be coarsely dense without being dense in X. He also showed that if there exists a coarsely dense orbit then there exists (a possibly different) vector with a dense orbit; see [4, Theorem 2.1]. This result is based on the fact that in infinite dimensional Banach spaces there exists a sequence of norm-one vectors {xn} and a positive number ε such that kxn − xkk ≥ ε for all n 6= k. Coarse density of orbits of a single operator is a phenomenon that occurs only in infinite dimen- sional vector spaces but coarse density of orbits of finitely generated semigroups may occur also on finite dimensional vector spaces. In con- trast with the case of a single operator on an infinite dimensional space, every coarsely dense orbit of a subsemigroup of GL(V ) generated by a finite set of commuting elements is dense as the main result of the present paper shows. The plan of the paper is the following. In section 2 we start by show- ing that under a quite natural and obviously necessary condition the closure S of a subsemigroup S of a vector group is actually a group (Theorem 2.1). This easily implies the same criterion for subsemigroups of compactly generated abelian locally compact topological groups, by Pontryagin's structure theorem, corollary 2.4. It has as corollaries a criterion for density, corollaries 2.2 and 2.5(3), and for cocompact- ness, corollary 2.5(1). These statements may be regarded as duality statements for subsemigroups of such abelian groups. The general phi- losophy is: Test by applying continuous homomorphisms to R. Thus the separating hyperplane theorem for closed convex cones is used in an essential way. These results of section 2 may be of independent interest. LINEAR SEMIGROUPS WITH COARSELY DENSE ORBITS 3 In section 3 we show our main theorem in the more general form mentioned above, namely for cones with inner points. At a crucial point (lemma 3.7) we use a general result about algebraic actions of algebraic groups. In section 4 we give detailed information about the orbits of S, part of which was described above. Section 5 contains remarks and questions. 2. A density criterion for semigroups Theorem 2.1. Let S be a subsemigroup of Rn with the property that whenever a linear form l : Rn → R has a positive value on S it also has a negative value on S. Then the closure of S is a subgroup of Rn. All closed subgroups A of Rn are of the following form. There is a decomposition Rn = V1 ⊕ V2 ⊕ V3 into three (not necessarily non-zero) vector subspaces V1, V2, V3 of Rn such that A = V1 ⊕ Γ2, where Γ2 is a lattice in V2. Thus Corollary 2.2. Let S be a subsemigroup of Rn such that l(S) has dense image in R for every non-zero linear form l : Rn → R. Then S is dense in Rn. We have good criteria for when a subsemigroup of R is dense, see corollary 2.3 and lemma 3.2. For a more general form of this corollary see corollary 2.5(3). Proof of theorem 2.1. We may assume that S spans Rn, equivalently that l(S) is non-zero for every non-zero linear form l : Rn → R. Oth- erwise we consider the vector subspace of Rn which is the intersection of the kernels of all linear forms l : Rn → R which vanish on S. Let P (S) = {r · s ; r ≥ 0, s ∈ S} be the positive hull of S. We claim that the closure P (S) of P (S) is not only a closed cone but also convex. Namely, since S is a semigroup, if s1, s2 are elements of S all positive rational linear combinations of s1 and s2 are in P (S) and hence P (S) contains the convex cone generated by s1 and s2. This easily implies that P (S) is convex. We next claim that P (S) = Rn. Otherwise there would be a non- zero linear form l : Rn → R such that l(P (S)) ≥ 0, by the separating hyperplane theorem. But this contradicts our hypotheses about S. Let us fix a norm k · k on Rn. Then the set of directions { s ksk ; s ∈ S, s 6= 0} of S is dense in the norm 1 sphere S 1 = {v ∈ Rn ; kvk = 1} since P (S) is dense in Rn. Let x 6= 0 be an element of S. We claim that −x ∈ S. This implies that −S ⊂ S and hence the theorem. Given x, there is a basis e1, . . . , en of Rn consisting of elements of S such that 4 H. ABELS AND A. MANOUSSOS x = α1e1 + . . . + αnen with αi < 0 for i = 1, . . . , n, by the density of the set of directions of S, since if x is a linear combination of a basis such that all the coefficients have negative sign then the same holds for a nearby basis and hence also for one with nearby directions. Let Γ be the lattice in Rn generated by e1, . . . , en. There is a sequence of natural numbers mj → ∞ such that −mjx mod Γ converges to the identity element in the compact group Rn/Γ. So for every ε > 0 there is a large natural number m and integers b1, . . . , bn such that k − mx − b1e1 − . . . − bnenk < ε. Comparing coefficients we see that bi > 0 for i = 1, . . . , n. So the element b1e1 + . . . + bnen + (m − 1)x of S has distance less than ε from −x, which implies our claim. (cid:3) Theorem 2.1 has as special case the following corollary. Corollary 2.3. Let S be a subsemigroup of R which contains both a positive and a negative real number. Then S is either dense in R or a cyclic subgroup of R. We shall use the theorem in the following form. Corollary 2.4. Let G be a compactly generated abelian locally compact topological group, e.g. an abelian Lie group whose group of connected components is finitely generated. Let S be a subsemigroup of G. Sup- pose that every continuous homomorphism f : G → R which has a positive value on S also has a negative value on S. Then the closure of S is a subgroup of G. Proof. The group G has a unique maximal compact subgroup K and G/K is isomorphic to a direct sum of a vector group Rm and a lattice Zl, by Pontryagin's structure theorem. Let π : G → G/K be the natural projection. We can think of Rm ⊕ Zl as a subgroup of Rn with n = m + l. The closure of π(S) is a subgroup of G/K by theorem 2.1. The corollary then follows from the following facts. The mapping π is proper so the image π(S) of the closure S of S is closed, hence a subgroup of G/K. And S ∩ K is a closed subsemigroup of the compact group K, hence a subgroup. (cid:3) A subsemigroup S of an abelian topological group G is said to be cocompact if there is a compact subset K of G such that G = S · K. The subsemigroup S of G is called properly discontinuous if it has no accumulation point in G. And S is called crystallographic if it both cocompact and properly discontinuous. Corollary 2.5. Let G be a compactly generated abelian locally compact topological groups and let S be a subsemigroup of G. LINEAR SEMIGROUPS WITH COARSELY DENSE ORBITS 5 (1) Suppose f (S) is cocompact in R for every non-zero continuous homomorphism f : G → R. Then S and S are cocompact in G and S is a subgroup of G. (2) If S is furthermore properly discontinuous then S is a closed discrete crystallographic subgroup of G. (3) If f (S) is dense in R for every non-zero continuous homomor- phism f : G → R then G/K is a vector group isomorphic to Rm for some m, S is a subgroup of G and S · K = G, where K is the maximal compact subgroup of G. Proof. (1) For every non-zero continuous homomorphism f : G → R the subsemigroup f (S) of R is cocompact, hence has both a positive and a negative value. So S is a subgroup of G. As in the proof of corollary 2.4 we may assume that G = Rm ⊕ Zl ⊂ Rn, with n = m + l, by computing modulo the maximal compact subgroup K of G. A subgroup of G is then cocompact if and only if it spans Rn as a vector space. This shows that S is cocompact in G and hence also S, since if G = S · L for some compact subset L of G then G = S · L1 for some compact neighborhood L1 of L. (2) If S is furthermore properly discontinuous then S is closed in G, hence S = S is a closed crystallographic, in particular discrete, subgroup of G. (3) Finally, if f (S) is dense in R for every non-zero continuous ho- momorphism f : G → R then S is a subgroup of G by (1). Also, the factor Zl in Pontryagin's structure theorem G/K ∼= Rm ⊕ Zl does not occur, so G/K ∼= Rm, and S · K = G by corollary 2.2. (cid:3) 3. Proof of the main theorem We shall prove the following theorem which contains theorem 1.1 as the special case C = V . Note that we always refer to the Euclidean topology, unless we explicitly mention the Zariski topology. Theorem 3.1. Let V be a finite dimensional real or complex vector space. Let S be a subsemigroup of GL(V ) generated by a finite set of commuting elements. Suppose there is an open cone C in V and an orbit O = S · v0 of S such that O ∩ C is coarsely dense in C. Then (1) The closure S of S in GL(V ) is an open subgroup of the Zariski closure of S. In particular, S has a finite number of connected components since every real algebraic group has only a finite number of connected components, see [2, V.24.6(c)(i)]. (2) The map S → V , g 7→ gv0, is an analytic diffeomorphism of S onto an open cone in V . In particular dim S = dimR V . 6 H. ABELS AND A. MANOUSSOS (3) The closure of the orbit O is a cone in V which contains C. We will give more information on the cones S · v0 and S · v0 below; see theorem 4.1. The proof of theorem 3.1 is given in steps, proceeding from special cases to more general cases. We start with Lemma 3.2. Let S be a finitely generated subsemigroup of R. Then either S is dense in R or every bounded interval in R contains only a finite number of elements of S. In the latter case, the number of elements of S ∩ [−n, n] is bounded by a polynomial in n. Proof. If S contains both a positive and a negative number then the lemma follows from corollary 2.3. So suppose S is finitely generated and contained in [0, ∞). Let s0 be a minimal positive generator of S. If S is generated by t elements, there are at most nt elements of S in [0, ns0). (cid:3) This implies the case dimR V = 1 of theorem 3.1. Corollary 3.3. Let R∗ + be the multiplicative group of positive real num- bers. Endow it with the metric induced from the Euclidean metric of R. Then every coarsely dense finitely generated subsemigroup of R∗ + is dense. Proof. Let S be a finitely generated coarsely dense subsemigroup of R∗ +. Then the interval [1, x], x > 1, contains at least c · x elements of S for some positive constant c. Thus the interval [0, log x] contains at least c · x elements of the finitely generated subsemigroup log S of the additive group R, hence is dense in R, by the preceding lemma. (cid:3) Let Rn + = {(x1, . . . , xn) ; xi > 0}. Then Rn under componentwise multiplication. By a cone in Rn in Rn that is contained in Rn is a cone if and only if for every non-zero x ∈ C the open ray R∗ is contained in C. We endow Rn Euclidean metric on Rn. +)n is a group + we mean a cone +. In geometric terms, a subset C of Rn + · x + with the metric induced from the + = (R∗ + Lemma 3.4. Let S be a finitely generated subsemigroup of Rn property that for some a ∈ Rn a coarsely dense subset of some open cone in Rn Rn +. + with the + the orbit S · a = {s · a ; s ∈ S} contains +. Then S is dense in Proof. We will use corollary 2.2 in the following form. dense image in R∗ Rn passing to exponentials since continuous homomorphisms α : Rn If α(S) has + for every non-trivial continuous homomorphism α : +. This follows from corollary 2.2 by + → R∗ + then S is dense in Rn + → R∗ + LINEAR SEMIGROUPS WITH COARSELY DENSE ORBITS 7 correspond bijectively to linear maps l : Rn → R under l 7→ exp ◦l ◦ log. We will show that α(S) has dense image in R∗ + → R∗ with α(x1, . . . , xn) = xα1 for (α1, . . . , αn) 6= 0 ∈ Rn. Let C be an open cone in Rn + which contains a coarsely dense subset of the orbit S · a. There are two cases to consider: + for every α : Rn 1 . . . xαn n + (1) P αi 6= 0. We may assume that P αi > 0, by passing to α−1, if necessary, and that P αi < 1, by passing to a positive multiple r · (α1, . . . , αn), r > 0, of (α1, . . . , αn), which changes the image of α by the automorphism x 7→ xr of R∗ +. Then α is positively homogeneous of = αi · x−1 degree P αi and ∂α · α is positively homogeneous of degree ∂xi P αi − 1 < 0 for i = 1, . . . , n. Hence α grows monotonously to +∞ on every ray R := R∗ +, and there is an open cone C ′ in Rn + containing our ray R such that α is distance decreasing on the set of points of C ′ of sufficiently large Euclidean norm. If C ′ is contained in the cone C it follows that α(S · a) = α(S) · α(a) is coarsely dense in R∗ +, by corollary 3.3. The other case is +, hence also α(S). Thus α(S) is dense in R∗ + · x, x ∈ Rn i + = S 1 ∩ Rn +. By our hypothesis about coarse density the set of points of S 1 (2) P αi = 0. Let S 1 = {x ∈ Rn ; kxk = 1} be the norm 1 sphere in Rn and S 1 +. The function α is constant on every ray in Rn + for which the corresponding ray R∗ + ∩ C. It follows that the closure of α(S · a) = α(S) · α(a) contains the open set α(C). Hence α(S) is dense in R∗ + by applying lemma 3.2 to the finitely generated subsemigroup log α(S) of R or to − log α(S). (cid:3) + · x intersects S · a is dense in S 1 The next case we consider is that S is diagonalizable over the complex numbers. Then V decomposes into a direct sum of S-invariant one- or two-dimensional real vector subspaces V1, . . . , Vr, Vr+1, . . . , Vd with the following properties. Each Vi, i ≤ r, is one-dimensional and S acts by multiplication by scalars. Each Vi for i > r is two-dimensional and can be endowed with the structure of a one-dimensional complex vector space such that S acts by multiplication by complex scalars. We thus have an embedding of S into the following group G. Let G be the group of diagonal d × d-matrices whose first r diagonal entries are real and non-zero and whose last d − r entries are complex non- zero. Let δi : G → R∗, resp. C∗, be the projection to i-th diagonal component. So S acts in the following way on V = V1 ⊕ . . . ⊕ Vd; s(v1, . . . , vd) = (δ1(s)v1, . . . , δd(s)vd) for s ∈ S. Lemma 3.5. Suppose S is diagonalizable over the complex numbers. Suppose furthermore that there is an orbit of S which contains a co- arsely dense subset of some open cone. Then the closure of S is a 8 H. ABELS AND A. MANOUSSOS subgroup of G and contains the connected component G0 of the identity of G. Proof. Let v0 be a point of V and let C be an open cone in V such that S · v0 ∩ C is coarsely dense in C. Let V 6=0 be the set of vectors in V = V1 ⊕ . . . ⊕ Vd all of whose components are non-zero. The group G acts simply transitively on V 6=0. Clearly, v0 ∈ V 6=0. We have an absolute value map abs : V → [0, ∞)d, v = (v1, . . . , vd) 7→ (v1, . . . , vd) if we choose isomorphisms Vi ≃ R for i ≤ r and Vi ≃ C for i > r. We also have an absolute value map on G, · : G → (R∗ +)n. The pair ( · , abs) is equivariant, that is abs(g · v) = gabs(v) for g ∈ G and v ∈ V . The image of C under the map abs is an open cone in (R∗ +)n. The image of S · v0 under the map abs is the orbit of abs(v0) under S and contains a coarsely dense subset of abs(C), since abs is distance non-expanding. Note that the notion of coarse density is independent of the norm on V we choose, since any two norms on V are Lipschitz equivalent. It follows from lemma 3.4 that S is dense in (R∗ +)d. This implies that the closure S of S is a group, since the kernel K of the homomorphism · : G → (R∗ +)d is compact and S ∩ K is a closed subsemigroup of K and hence a subgroup. We now claim that the closure S of S contains an open subset of G and hence is an open subgroup of G which finishes the proof. We will actually show that S · v0 is dense in C which implies our claim, since G → V 6=0, g 7→ gv0, is a diffeomorphism. The group G is the direct product of its subgroup (R∗ +)d and its maximal compact subgroup K. We denote the two components of an element g ∈ G by g and arg(g), respectively, since this decomposition is just the polar decomposition of the diagonal entries in C∗ and R∗, respectively. The image of S under · is closed, since · is a proper map, hence S = (R∗ +)d. So for every positive real number r there is an element g ∈ S such that g is the homothety Mr with Mr(x) = r · x. Taking an appropriate power of g we obtain elements gn with arg(gn) arbitrarily close to the identity. It follows that for every point x ∈ C there are arbitrarily large powers gn of g such that g−nx ∈ C. Then there is a point y ∈ S · v0 at distance at most D from g−nx, where D is the constant in the definition of coarse density. Thus the point gny ∈ S · v0 is at distance rn · D from x, since kgny − xk = kgn(y − g−nx)k = rnk arg(gn)(y − g−nx)k ≤ rn · D. This shows that S · v0 is dense in C if we take r < 1. (cid:3) We now come to the proof of the general case of theorem 3.1. Lemma 3.6. Under the hypotheses of theorem 3.1 the orbit O := S · v0 is Zariski-dense in V . LINEAR SEMIGROUPS WITH COARSELY DENSE ORBITS 9 Proof. The claim is that the zero polynomial is the only polynomial function on V which vanishes on O. Thus let f be a non-zero poly- nomial function which vanishes on O. Let fn be the homogeneous component of f of highest degree. Then there is an open subset U of C such that fn has no zero on U . We may assume that U is compact. Then f (tx) = fn(tx) + (f (tx) − fn(tx)) ≥ tn−1(t · min{fn(x) ; x ∈ U} − max{f (x) − fn(x) ; x ∈ U }) for x ∈ U and t > 0, so f (tx) has no zero for t ≫ 0 and x ∈ U, contradicting the fact that O ∩ C is coarsely dense in C. (cid:3) Lemma 3.7. Let G be the Zariski closure of S in GL(V ). Then G → V , g 7→ gv0, is a diffeomorphism of G onto an open subset of V . Proof. By a theorem on algebraic actions [3, § 3.18] the map G → V , g 7→ gv0, to the orbit Gv0 is a submersion and the orbit Gv0 is an open subset - with respect to the Euclidean topology - in its Zariski closure. Furthermore the isotropy group Gv = {g ∈ G ; gv = v} is independent of the chosen point v of the orbit, since Ggv = gGvg−1 and G is abelian, and hence must be trivial, since the identity is the only element of GL(V ) that fixes every point of an open set. (cid:3) Proof of theorem 3.1. Let G be the Zariski closure of S in GL(V ). Then G is an abelian real linear algebraic group and hence is the direct product of its subgroup Gs of semisimple elements and its subgroup Gu of unipotent elements. Let πs : G → Gs and πu : G → Gu be the corre- sponding projection homomorphisms. Let Vu be the G-invariant vector subspace Pu∈Gu(1 − u)V of V . The quotient space Vs := V /Vu has a natural representation of G whose kernel is Gu and the representation of G on Vs is semisimple. Let ps : V → Vs be the natural projection. Then the point ps(v0) has the orbit S · ps(v0) = πs(S)ps(v0). This orbit intersects the open cone ps(C) in a coarsely dense subset of ps(C). It follows from lemma 3.5 that the closure of πs(S) is a subgroup of Gs and dim πs(S) = dimR Vs. On the other hand dim Gs = dimR Vs by the pre- ceding lemma, applied to the action of Gs on Vs. So G0 s ⊂ πs(S) ⊂ Gs, where H 0 denotes the connected component of the identity in a topo- logical group H. The group Gs was described in the paragraph before lemma 3.5. In particular G0 + · 1 of homotheties Mr with Mr(x) = r · x for r ∈ R∗ +. s contains the group R∗ Next we will show that the closure S of S in G is a subgroup of G and S · K = G where K is the maximal compact subgroup of G. It suffices to show that for every non-zero continuous homomorphism f : G → R the image f (S) is dense in R, by corollary 2.5(3). Note that this corollary also implies that G/K is a vector group, in particular 10 H. ABELS AND A. MANOUSSOS connected, which can also be seen from the facts that Gu is connected and Gs is - considered as a Lie group - a direct product of groups R∗ and C∗. Thus let f : G → R be a non-zero continuous homomorphism. Let ϕ : Gv0 → R be the analytic map defined by ϕ(gv0) = f (g). Note that ϕ(gx) = f (g) + ϕ(x) for g ∈ G and x ∈ Gv0. Recall that Gv0 is open in V . We know that R∗ + · 1 ⊂ G0. There are two cases to consider, namely that the restriction f R∗ + · 1 is the zero map or not. + · 1 of f to R∗ Case 1. Suppose f R∗ + · 1 is the zero map. Then ϕ is constant on every ray R∗ +x in Gv0. Let C be our open cone such that S · v0 ∩ C is coarsely dense in C. Then for every open subset U of C the set of x ∈ U for which the corresponding ray R∗ +x intersects S · v0 is dense in U, so f (S) = ϕ(S · v0) contains a dense subset of the open set ϕ(U). Note that ϕ is an open map, namely the composition of a diffeomorphism Gv0 → G and a non-zero homomorphism G → R. It follows from lemma 3.2 that f (S) is dense in R. Case 2. The other case is that f R∗ + · 1 is non-zero. We may assume that f (et 1) = t for every t ∈ R. Let x ∈ C and let B(x, ε) be the ball of radius ε with center x. Then B(etx, etε) = MetB(x, ε) contains a point of S · v0 if etε ≥ D, where D is the constant in the definition of coarse density. There is a constant c such that ϕ(y) − ϕ(z) ≤ cky − zk for every y, z ∈ B(x, ε) since ϕ is continuously differentiable. It follows that for every t ∈ R ϕ(B(etx, etε)) = ϕ(MetB(x, ε)) = f (Met) + ϕ(B(x, ε)) = t + ϕ(B(x, ε)) is contained in the interval in R with center t and radius c · ε. So for sufficiently large t ∈ R there is an element of ϕ(S · v0) = f (S) of distance at most D · c · e−t from t. It follows from lemma 3.2 that f (S) is dense in R. The group G is the direct product of Gu and Gs and Gs is the direct product of its polar part, a product of R∗ +, and its maximal compact subgroup, a product of groups {±1} and S 1. The same proof as at the end of lemma 3.5 shows that S contains the connected component G0 of G. Now R∗ + · 1 is contained in S, hence the orbit of every point under S is a cone, i.e. R∗ +-invariant. It follows that S · v0 contains a dense subset of C, so the closure of S · v0 contains C. Note that it can happen that C is not contained in the orbit S · v0. E.g. if C = R, v0 6= 0 then LINEAR SEMIGROUPS WITH COARSELY DENSE ORBITS 11 S · v0 does not contain 0. Recall that S is the closure of S in the group GL(V ), hence contained in R∗ for V = R. The general picture can be seen from theorem 4.1. (cid:3) 4. Orbit structure We determine the structure of the open orbits of S. The following theorem will be applied for the group G∗ = S of theorem 3.1. Parts of these results are in [1]. Theorem 4.1. Let G∗ be an abelian subgroup of GL(V ) which has an orbit with an inner point. Let G be the Zariski closure of G∗. Then the following claims hold: (1) The group G∗ is an open subgroup of G and contains the con- nected component G0 of G. (2) There is only a finite set of maximal G0-invariant vector sub- spaces of V . They are also G-invariant. These are of real codi- mension 1 or 2 in V . Let H1, . . . , Hr be those of codimension 1 and Hr+1, . . . , Hd be those of codimension 2. We have r + 2(d − r) ≤ dimR V. (3) Let U be the complement of Sd i=1 Hi in V and let v be a point of U . Then the map G → U , g 7→ gv, is an analytic diffeomor- phism. (4) The quotient group G/G0 is isomorphic to (Z/2Z)r. For v ∈ U the orbit G0v is the connected component of v in U . The closure of G0v is the intersection of r half spaces. Namely, for each i = 1, . . . , r there is exactly one open half space defined by Hi which contains v, say Ci. Then G0v = Tr i=1 Ci. (5) If V has a structure as complex vector space such that every h ∈ G∗ is complex linear, then r = 0, G is connected and hence G0 = G∗ = G. Then U is the only open G∗-orbit and thus it is dense in V . Proof. The group G∗ contains an open subset of G, by lemma 3.7. Thus, being a group, G∗ is open in G and hence a closed subgroup of G. This implies (1). The group G is an abelian real linear algebraic group so it has a finite number of connected components. It is also a direct product of its subgroup Gs of semisimple elements and its subgroup Gu of unipotent elements. For the remaining claims we start with special cases. (a) Gs = R∗ · 1. There is a non-zero linear form l : V → R which is fixed by Gu. So Gu leaves every hyperplane parallel to the kernel H 12 H. ABELS AND A. MANOUSSOS of l invariant. There is a vector v with a somewhere dense orbit, so l(v) 6= 0. Now ϕv : G → Gv, g 7→ gv, is a diffeomorphism onto an open subset of V which maps Gs = R∗ · 1 onto the line Rv with the origin removed and Gu to the affine hyperplane H +v. Comparing dimensions we see that Guv is a neighborhood of v in H + v and hence Guv is an open subset of H + v. On the other hand, the orbit of any vector under a unipotent Zariski closed subgroup of GL(V ) is closed [5, Lemma 5.3]. It follows that Guv = H + v. Now Gu(tv) = tGuv = tv + H, so the orbit of any vector v /∈ H is V \ H. This shows that H is the only G or G0-invariant hyperplane in V . This is the case r = d = 1 of our theorem. (b) V has a complex structure such that G∗ consists of complex linear maps and Gs = C∗ · 1. Thus G is connected and hence G0 = G∗ = G. Exactly the same arguments as above show that Gv = V \ H for every v /∈ H. In particular H is the only maximal G-invariant (real or complex) vector subspace of V . This is the case r = 0, d = 1, of our theorem. (c) The general case. Every isotypic component of the Gs-module V is of the type described in (a) or (b), since the hypothesis that the orbit of G∗ has an inner point carries over to the isotypic components and then to their semisimple quotients V modulo Pg∈Gu(1 − g)V , so the other irreducible algebraic subgroups of C∗ do not occur in the simple quotients of V . Note that if a simple quotient is C with the action of C∗ · 1, considered as a real vector space, then the isotypic component has a (unique up to complex conjugation) structure as a complex vector space and every g ∈ G acts by complex linear automorphisms, since G is commutative. So, back to the general case, V has a decomposition into isotypic modules V1 ⊕ . . . ⊕ Vr ⊕ Vr+1 ⊕ . . . ⊕ Vd where Gs acts on Vi as R∗ · 1 for i ≤ r and for i > r the Gs-module Vi has a structure as complex vector space, Gs acts as C∗ · 1 and G acts by complex linear automorphisms. Let H be a maximal G-invariant real vector subspace of V . Then there is an index i, 1 ≤ i ≤ d, such that H = Lj6=i Vj ⊕ H ∗ is a maximal G-invariant subspace of Vi. We thus are with H ∗ i ⊂ Vi in case (a) or (b). In particular, for a given index i there is exactly one such subspace. Let us call it Hi. The natural representation of G on V / ⊕ H ∗ i = L Vi/H ∗ i has image in Lr i=r+1 (cid:0)C∗ · 1Vi/H ∗ i (cid:1) and has an orbit with an inner point. So the dimension of Gs is r + 2(d − r) = dim V / ⊕ H ∗ i . It follows that ⊕H ∗ i has the same codimension in V as Gu has in G. But Gu leaves ⊕H ∗ i invariant. Now let v0 be a point of V whose G∗-orbit has an inner point. Then v0 /∈ S Hi, in other words, for every component i=1 (cid:0)R∗ · 1Vi/H ∗ i and H ∗ i i (cid:1) ⊕ Ld LINEAR SEMIGROUPS WITH COARSELY DENSE ORBITS 13 0 /∈ H ∗ vi 0 of v0 in Vi we have vi i . The map G → Gv0, g 7→ gv0, is a diffeomorphism of G onto an open subset of V , by lemma 3.7. So, comparing dimensions, we see that the orbit of Gu in v0 ⊕ L H ∗ is i open. But Guv0 is also closed since every orbit of a unipotent algebraic group is closed, so Guv0 = v0 ⊕L H ∗ i . On the other hand G = Gs ×Gu, and Gs is an algebraic subgroup of Lr i=r+1 (C∗ · 1Vi) of the same dimension, hence they are equal. This shows claim (3). i=1 (R∗ · 1Vi) ⊕ Ld If V has a complex structure, then every simple quotient of the G-module V again has an open orbit, so we are in case (b), which implies (5). Part (4) spells out the connected components of G, namely G/G0 ∼= Gs/G0 (cid:3) +)r ∼= (Z/2Z)r. ∼= (R∗)r/(R∗ s 5. Open problems and remarks Question 1. Does theorem 2.1 also hold for subsemigroups of topolog- ical vector spaces? Or at least corollary 2.2? G. Soifer asked us the following question. Let S be a finitely gen- erated commutative linear semigroup. Suppose S has an orbit which is coarsely dense along a spanning set of rays. Here a set A is called coarsely dense along a subset Y of a metric space (X, d) if the open balls of radius D with centers at points of A cover Y for some positive number D. Does this imply that there is an orbit which is dense in some open set? The answer is no to this more general question. Here are two examples. 1) Consider the one-parameter group G of diagonal positive ma- trices in SL2(R). Every dense finitely generated subsemigroup of G is coarsely dense along the two positive axes, but has no somewhere dense orbit. 2) Consider the group G of upper triangular 2 × 2-matrices whose diagonal entries are equal and positive. Let S be a finitely gener- ated subsemigroup of G whose closure are all the matrices of G whose off diagonal entries are integers. This semigroup has an orbit which is (coarsely) dense along infinitely many rays, but has no somewhere dense orbit. The following question is open. Question 2. Can one find a finite number R1, . . . , Rm of rays such that every finitely generated commutative linear semigroup S which has a coarsely dense orbit along Sm i=1 Ri also has a somewhere dense orbit? Acknowledgement. We thank the referee for valuable suggestions which helped us to improve the presentation. 14 H. ABELS AND A. MANOUSSOS References [1] A. Ayadi and H. Marzougui, Dynamic of Abelian subgroups of GL(n, C): a structure theorem, Geometria Dedicata 116 (2005), 111-127. [2] A. Borel, Linear Algebraic Groups, 2nd edition, Graduate Texts in Mathemat- ics, vol. 126, Springer-Verlag, 1991. [3] A. Borel and J. Tits, Homomorphismes "abstraits" de groupes alg´ebriques sim- ples, Ann. of Math. (2) 97 (1973), 499-571. [4] N.S. Feldman, Perturbations of hypercyclic vectors, J. Math. Anal. Appl. 273 (2002), 67-74. [5] C.C. Moore, Distal affine transformation groups, Amer. J. Math. 90 (1968), 733-751. Fakultat fur Mathematik, Universitat Bielefeld, Postfach 100131, D-33501 Bielefeld, Germany E-mail address: [email protected] Fakultat fur Mathematik, SFB 701, Universitat Bielefeld, Post- fach 100131, D-33501 Bielefeld, Germany E-mail address: [email protected] URL: http://www.math.uni-bielefeld.de/~amanouss
1308.3504
2
1308
2013-10-07T13:12:53
Bounds for alpha-Optimal Partitioning of a Measurable Space Based on Several Efficient Partitions
[ "math.FA", "math.OC" ]
We provide a two-sided inequality for the alpha-optimal partition value of a measurable space according to n nonatomic finite measures. The result extends and often improves Legut (1988) since the bounds are obtained considering several partitions that maximize the weighted sum of the partition values with varying weights, instead of a single one.
math.FA
math
Bounds for α−Optimal Partitioning of a Measurable Space Based on Several Efficient Partitions Marco Dall'Aglio LUISS University Rome, Italy Camilla Di Luca LUISS University Rome, Italy [email protected] [email protected] October 7, 2013 Abstract We provide a two-sided inequality for the α−optimal partition value of a measurable space according to n nonatomic finite mea- sures. The result extends and often improves Legut (1988) since the bounds are obtained considering several partitions that maximize the weighted sum of the partition values with varying weights, instead of a single one. Introduction 1 Let (C,C) be a measurable space, N = {1, 2, . . . , n}, n ∈ N and let {µi}i∈N be nonatomic finite measures defined on the same σ−algebra C. Let P stand for the set of all measurable partitions (A1, . . . , An) of C (Ai ∈ C for all i ∈ N , ∪i∈N Ai = C, Ai ∩ Aj = ∅ for all i (cid:54)= j). Let ∆n−1 denote the (n − 1)-dimensional simplex. For this definition and the many others taken from convex analysis, we refer to [10]. Definition 1. A partition (A∗ 1, . . . , A∗ α = (α1, . . . , αn) ∈ int∆n−1, if n) ∈ P is said to be α−optimal, for (cid:26)µi(A∗ i ) (cid:27) (cid:26) (cid:26) µi(Ai) (cid:27) (cid:27) vα := min i∈N αi = sup min i∈N : (A1, . . . , An) ∈ P . (1) αi 1 This problem has a consolidated interpretation in economics. C is a non-homogeneous, infinitely divisible good to be distributed among n agents with idiosyncratic preferences, represented by the measures. A partition (A1, . . . , An) ∈ P describes a possible division of the cake, with slice Ai given to agent i ∈ N . A satisfactory compromise between the conflicting interests of the agents, each having a relative claim αi, i ∈ N , over the cake, is given by the α−optimal partition. It can be shown that the pro- posed solution coincides with the Kalai-Smorodinski solution for bargaining problems (See Kalai and Smorodinski [12] and Kalai [11]). When {µi}i∈N are all probability measures, i.e.. µi(C) = 1 for all i ∈ N , the claim vector α = (1/n, . . . , 1/n) describes a situation of perfect parity among agents. The necessity to consider finite measures stems from game theoretic extensions of the models, such as the one given in Dall'Aglio et al. [5]. When all the µi are probability measures, Dubins and Spanier [8] showed that if µi (cid:54)= µj for some i, j ∈ N , then vα > 1. This bound was improved, together with the definition of an upper bound by Elton et al. [9]. A further improvement for the lower bound was given by Legut [13]. The aim of the present work is to provide further refinements for both bounds. We consider the same geometrical setting employed by Legut [13], i.e. the partition range, also known as Individual Pieces Set (IPS) (see Bar- banel [2] for a thorough review of its properties), defined as R := {(µ1(A1), . . . , µn(An)) : (A1, . . . , An) ∈ P} ⊂ Rn +. Let us consider some of its features. The set R is compact and convex (see Lyapunov [17]). The supremum in (1) is therefore attained. Moreover vα = max{r ∈ R+ : (α1r, α2r, . . . , αnr) ∩ R (cid:54)= ∅}. (2) sup(cid:8)(cid:80) So, the vector (vαα1, . . . , vααn) is the intersection between the Pareto frontier of R and the ray rα = {(rα1, . . . , rαn) : r ≥ 0}. i∈N µi(Ai) : (A1, . . . , An) ∈ P(cid:9) on the partition range. Then, he finds To find both bounds, Legut locates the solution of the maxsum problem the convex hull of this point with the corner points ei = (0, . . . , µi(C), . . . , 0) ∈ Rn (µi(C) is placed on the i-th coordinate) to find a lower bound, and uses a separating hyperplane argument to find the upper bound. We keep the same framework, but consider the solutions of several maxsum problems with weighted coordinates to find better approximations. Fix β = (β1, . . . , βn) ∈ ∆n−1 and consider(cid:88) βiµi(Aβ i ) = sup i∈N (cid:40)(cid:88) i∈N (cid:41) . (3) βiµi(Ai) : (A1, . . . , An) ∈ P 2 is absolutely continuous (for instance we may consider η =(cid:80) Let η be a non-negative finite-valued measure with respect to which each µi i∈N µi). Then, by the Radon-Nikodym theorem for each A ∈ C, (cid:90) µi(A) = fidη ∀ i ∈ N, where fi is the Radon-Nikodym derivative of µi with respect to η. Finding a solution for (3) is rather straightforward: A Proposition 1. (see [8, Theorem 2], [1, Theorem 2] [4, Proposition 4.3]) Let β ∈ ∆n−1 and let Bβ = (Aβ n) be an n−partition of C. If 1 , . . . , Aβ for all h, k ∈ N and for all x ∈ Aβ k , (4) βkfk(x) ≥ βhfh(x) 1 , . . . , Aβ n) is optimal for (3). then (Aβ Definition 2. Given β ∈ ∆n−1, an efficient value vector (EVV) with respect to β, uβ = (uβ 1 , . . . , uβ n), is defined by uβ i = µi(Aβ i ), for each i = 1, . . . , n. The EVV uβ is a point where the hyperplane (cid:88) i∈N (cid:88) i∈N βixi = βiui (5) touches the partition range R, so uβ lies on the Pareto border of R. 2 The main result As we will see later only one EVV is enough to assure a lower bound, we give a general result for the case where several EVVs have already been computed. We derive this approximation result through a convex combination of these easily computable points in R, which lie close to (α1vα, . . . , αnvα). Theorem 1. Consider m ≤ n linearly independent vectors u1, . . . , um, where ui = (ui1, ui2, . . . , uin), i ∈ M := {1, 2, . . . , m}, is the EVV associated to βi, βi = (βi1, . . . , βin) ∈ ∆n−1. Let U = (u1, u2, . . . , um) be an n × m matrix and denote as ¯U an m × m submatrix of U such that det( ¯U ) (cid:54)= 0 . 3 (6) (i) if and only if α ∈ cone(u1, . . . , um) (7) det( ¯U ) det( ¯Uαi) ≥ 0 for all i ∈ M, (8) where ¯Uαi is the m× m matrix obtained by replacing the i−th column of ¯U with ¯α ∈ Rm, obtained from α by selecting the elements corresponding to the rows in ¯U . Moreover, α ∈ ri(cone(u1, . . . , um)) if and only if det( ¯U ) det( ¯Uαi) > 0 for all i ∈ M. (ii) For any choice of u1, . . . , um, vα ≤ min i∈M (cid:80) (cid:80) j∈N βijuij j∈N βijαj . Moreover, if (8) holds, then (cid:80) (cid:80) 1 (cid:2) ¯U−1(cid:3) ij ≤ vα where(cid:2) ¯U−1(cid:3) i∈M j∈M ¯αj ij is the ij-th element of ¯U−1. (9) (10) (11) (12) (13) Proof. To prove (i), suppose (8) holds. We show that conv(u1, . . . , um)∩rα (cid:54)= ∅, and therefore that (7) holds, by verifying that the following system of linear equations in the variables r, t1, t2, . . . , tm  t1u11 + t2u21 + . . . + tmum1 = α1r t1u12 + t2u22 + . . . + tmum2 = α2r ... t1u1n + t2u2n + . . . + tmumn = αnr t1 + t2 + . . . + tm = 1 m) with t∗ i ≥ 0, for i ∈ M . First of all, has a unique solution (r∗, t∗ det( ¯U ) (cid:54)= 0 implies det( ¯Uαi∗) (cid:54)= 0 for at least an i∗ ∈ M , otherwise all the EVVs would lie on the same hyperplane, contradicting the linear indepen- dence of such vectors. This fact and (8) imply that the coefficient matrix has 1, . . . , t∗ 4 rank m + 1 and its unique solution can be obtained by deleting the n − m equations corresponding to the rows not in ¯U . Denote each column of ¯U as ¯ui = (¯ui1, . . . , ¯uim), i ∈ M and denote as ¯α = (¯α1, . . . , ¯αm), the vector obtained from α by selecting the same components as each ¯ui. By Cramer's rule we have for each i ∈ M ,  det ti = . . . . . . ... ¯u21 ¯u22 ... ¯ui−1,1 ¯ui−1,2 ¯um1 −¯α1 ¯u11 ¯um2 −¯α2 ¯u11 ... ... ... ¯u1m ¯u2m . . . ¯ui−1,m 0 ¯ui+1,m . . . ¯umm −¯αm 0 1 ¯ui+1,1 ¯ui+1,2 . . . . . . ... 0 0 ... . . . 1 1 1 1 ... ...  ¯u11 1 . . . ¯um1 −¯α1 ... ... ... ¯u1m . . . ¯umm −¯αm 0 1 . . . ... . . . 1  det  ¯u11    (−1)i+(m+1) det ... . . . . . . ... . . . . . . ... ¯u21 ¯u22 ... ¯ui+1,1 ¯ui+1,2 ¯ui−1,1 ¯ui−1,2 ¯um1 −¯α1 ¯um2 −¯α2 ¯u12 ... ... ... ¯u1m ¯u2m . . . ¯ui−1,m ¯ui+1,m . . . ¯umm −¯αm ¯um1 −¯α1 ... ... ... ¯u1m . . . ¯uj−1,m ¯uj+1,m . . . ¯umm −¯αm  ¯u11 . . . ... . . . ... ¯uj−1,1 ¯uj+1,1 ... ... ... j∈M (−1)(m+1)+j det (−1)2m+2 det( ¯Uαi) j∈M (−1)2m+2 det( ¯Uαj) = det( ¯Uαi) j∈M det( ¯Uαj) ≥ 0 (cid:80) (cid:80) (cid:80) = = since by (8) either a determinant is null or it has the same sign of the other determinants. If (10) holds, then ti > 0 for every i ∈ M and (9) holds. Conversely, each row of U not in ¯U is a linear combination of the rows in ¯U . Therefore, each point of span(u1, . . . , um) is identified by a vector x ∈ Rm whose components correspond to the rows in ¯U , while the other components are obtained by means of the same linear combinations that yield the rows of U outside ¯U . Let ¯U−j denote the m × (m − 1) matrix obtained from the matrix ¯U without ¯uj, j ∈ M . Consider a hyperplane H−j in span(u1, . . . , um) through the origin and m − 1 EVVs H−j := det(x, ¯U−j) = 0, 5 where j ∈ N . through H−j, for all j ∈ N either the ray rα is coplanar to H−j, i.e., If α /∈ cone(u1, . . . , um), when we separate the subspace det(¯α, ¯U−j) = 0, or ¯α and ¯uj lie in the same half-space, i.e., det(¯α, ¯U−j) det(¯uj, ¯U−j) > 0. (14) (15) Moving the first column to the j-th position in all the matrices above, we get (8). In case (9) holds, only the inequalities in (15) are feasible and (10) holds. To prove (ii), consider, for any i ∈ N , the hyperplane (5) that intersects the ray rα at the point (¯riα1, . . . , ¯riαn), with (cid:80) (cid:80) ¯ri = j∈N βijuij j∈N βijαj . Since R is convex, the intersection point is not internal to R. So, ¯ri ≥ vα for i ∈ M , and, therefore, mini∈M ¯ri ≥ vα. We get the lower bound for vα as solution in r of the system (13). By Cramer's rule, det   r∗ = det (cid:80) = i∈M 1 . . . 1 1 . . . . . . ... ¯u21 ¯u22 ... ¯um1 ¯um2 ... ¯u11 0 ¯u12 0 ... ... ¯u1m ¯u2m . . . ¯umm 0 1 ¯um1 −¯α1 ¯u11 ¯um2 −¯α2 ¯u12 ... ... ... ¯u1m ¯u2m . . . ¯umm −¯αm 0 1 1 . . . det( ¯U ) (cid:80) j∈M (−1)i+j ¯αj det( ¯Uij) ¯u21 ¯u22 ... . . . . . . ... 1   (cid:80) = = det  (cid:80) i∈M j∈M ¯αj  det( ¯U ) 1 1 ¯u21 ¯u11 ¯u22 ¯u12 ... ... . . . . . . . . . ... 1 0 −¯α1 ¯um1 −¯α2 ¯um2 ... ... −¯αm ¯u1m ¯u2m . . . ¯umm 1 (cid:2) ¯U−1(cid:3) , ij where det( ¯Uij) is the ij-th minor of ¯U . The second equality derives by suitable exchanges of rows and columns in the denominator matrix: In fact, swapping the first rows and columns of the matrix leaves the determinant unaltered. The last equality derives by dividing each row i of ¯U by αi, i ∈ M . Finally, by (2) we have r∗ ≤ vα. 6 Remark 1. The above result shows that whenever det( ¯Uαi) = 0, then ti = 0. Therefore, the corresponding EVV ui is irrelevant to the formulation of the lower bound and can be discarded. We will therefore keep only those EVV that satisfy (10) and will denote them as the supporting EVVs for the lower bound. In the case m = n we have Corollary 1. Suppose that there are n vectors u1, . . . , un, where ui, i ∈ N , is the EVV associated to βi, βi = (βi1, . . . , βin) ∈ ∆n−1. If U = (u1, . . . , un), det(U ) (cid:54)= 0 and, for all i ∈ N det(U ) · det(Uαi) ≥ 0, where Uαi is the n × n matrix obtained by replacing ui with α in U , then (cid:80) i∈N (cid:80) 1 j∈N αj [U−1]ij ≤ vα ≤ min i∈N j∈N βijuij j∈N βijαj (cid:80) (cid:80) (16) (17) where [U−1]ij is the ij-th element of U−1. We next consider two further corollaries that provide bounds in case only one EVV is available. The first one works with an EVV associated to an arbitrary vector β ∈ ∆n−1. Corollary 2. ([6, Proposition 3.4]) Let µ1, . . . , µn be finite measures and let u = (u1, u2, . . . , un) be the EVV corresponding to β ∈ ∆n−1 such that α−1 j uj = max i∈N α−1 i ui. Then, αj +(cid:80) i(cid:54)=j (cid:2)µ−1 i (C)(αiuj − αjui)(cid:3) ≤ vα ≤ uj (cid:80) (cid:80) i∈N βiui i∈N αiui (18) . (19) Proof. Consider the corner points of the partition range ei = (0, . . . , µi(C), . . . , 0) ∈ Rn , where µi(C) is placed on the i-th coordinate (i ∈ N ), and the matrix U = (e1, . . . , ej−1, u, ej+1, . . . , en), where u occupies the j-th position. Now det(U ) = uj µi(C) > 0 det(Uαj) = αj µi(C) > 0 (cid:89) (cid:89) i∈N\{j} i∈N\{j} 7 and, for all i ∈ N \ {j}, det(Uαi) = (αiuj − αjui) (cid:89) k∈N\{i,j} µk(C) , which is positive by (18). Therefore, U satisfies the hypotheses of Corollary 1. Since U has inverse 1 µ1(C) 0 ... 0 ... 0 0 1 µ1(C) ... 0 ... 0 µ1(C)uj µ2(C)uj ··· − u1 ··· − u2 ... . . . ··· ... . . . ··· − un 1 uj µn(C)uj ··· ··· . . . ··· . . . ··· 0 0 ... 0 ... 1 µn(C)  ,  U−1 = vα ≥ r∗ = the following lower bound is guaranteed for vα: αj +(cid:80) i∈N\{j} (cid:2)µ−1 i (C)(αiuj − αjui)(cid:3). uj The upper bound is a direct consequence of Theorem 1. In case all measures µi, i ∈ N, are normalized to one and the only EVV considered is the one corresponding to β = (1/n, . . . , 1/n), we obtain Legut's result. Corollary 3. ([13, Theorem 3]) Let µ1, . . . , µn be probability measures and let u = (u1, u2, . . . , un) be the EVV corresponding to β = (1/n, . . . , 1/n). Let j ∈ N be such that (18) holds. Then, ≤ vα ≤(cid:88) i∈N ui, (20) uj uj − αj(K − 1) where K =(cid:80) i∈N ui. Proof. Simply apply Corollary 2 with µi(C) = 1, for all i ∈ N and β = (1/n, . . . , 1/n). Then vα ≥ r∗ = αj +(cid:80) uj i(cid:54)=j(αiuj − αjui) = uj uj − αj(K − 1) , 8 where K =(cid:80) (cid:80) 1 n (cid:80) i∈N (cid:0) 1 n ui i∈N αi vα ≤ i∈N ui. Finally, by Theorem 1 we have (cid:1) (cid:88) i∈N = ui. It is important to notice that the lower bound provided by Theorem 1 certainly improves on Legut's lower bound only when one of the EVVs forming the matrix U is the one associated to β = (1/n, . . . , 1/n). Example 1. We consider a [0, 1] good that has to be divided among three agents with equal claims, α = (1/3, 1/3, 1/3), and preferences given as density functions of probability measures f1(x) = 1 f2(x) = 2x f3(x) = 30x(1 − x)4 x ∈ [0, 1] , f3 being the density function of a Beta(2, 5) distribution. The preferences of the players are not concentrated (following Definition 12.9 in Barbanel [2]) and therefore there is only one EVV associated to each β ∈ ∆2 (cfr. [2], Theorem 12.12) Figure 1: The density functions in Example 1. Agent 1: tiny dashing; Agent 2: dashing; Agent 3: continuous line. large The EVV corresponding to βeq = (1/3, 1/3, 1/3) is ueq = (0.0501, 0.75, 0.8594). Consequently, the bounds provided by Legut are 1.3437 ≤ vα ≤ 1.6594. Consider now two other EVVs u1 = (0.2881, 0.5556, 0.7761) u2 = (0.25, 0.9375, 0) 9 0.20.40.60.81.0x0.51.01.52.02.5f corresponding to β1 = (0.4, 0.3, 0.3) and β2 = (0.3, 0.6, 0.1), respectively. The matrix U = (u1, u2, ueq) satisfies the hypotheses of Theorem 1 and the improved bounds are 1.4656 ≤ vα ≤ 1.5443. 3 Improving the bounds The bounds for vα depend on the choice of the EVVs that satisfy the hy- potheses of Theorem 1. Any new EVV yields a new term in the upper bound. Since we consider the minimum of these terms, this addition is never harmful. Improving the lower bound is a more delicate task, since we should modify the set of supporting EVVs for the lower bound, i.e. those EVVs that include the ray rα in their convex hull. When we examine a new EVV we should verify whether replacing an EVV in the old set will bring to an improvement. A brute force method would require us to verify whether conditions (8) are verified with the new EVV in place of α. Only in this case we have a guarantee that the new EVV will not make the bound worst. Then, we should verify (8) again, with the new EVV replacing one of the m EVVs, in order to find the EVV from the old set to replace. However, in the following proposition we propose a more efficient condition for improving the bounds, by which we simultaneously verify that the new EVV belongs to the convex hull of the m EVVs and detect the vector to replace. For any couple j, k ∈ M denote as ¯U∗j−k the m× (m− 1) matrix obtained from ¯U by replacing column j with u∗ and by deleting column k. Theorem 2. Let u∗, u1, u2, . . . , um be m + 1 EVVs with the last m vectors satisfying conditions (6) and (10). If there exists j ∈ M such that for every k ∈ M \ {j} det(¯α, ¯U∗j−k) det(¯uj, ¯U∗j−k) ≤ 0 (21) then u∗ ∈ cone(u1, . . . , um) α ∈ cone({ui}i(cid:54)=j, u∗) (22) (23) Moreover, if all the inequalities in (21) are strict then in both (22) and (23) the vectors belong to the relative interior of the respective cones. Proof. Before proving the individual statements, we sketch a geometric in- terpretation for condition (21). As in Theorem 1 we restrict our analysis to the subspace span(u1, . . . , um). For any k ∈ M \ {j}, the hyperplanes H∗j−k = {x ∈ Rm : det(x, ¯U∗j−k) = 0} 10 should separate uj and α (strictly if all the inequalities in (21) are strict) in the subspace span(u1, . . . , um). To prove (22), argue by contradiction and suppose u∗ /∈ cone(u1, . . . , um). Then, for any j ∈ M, there must exist a k (cid:54)= j such that the hyperplane H∗j−k passes through all the EVV (including u∗) but uj and uk, and supports cone(u1, . . . , um). Therefore, α and uj belong to the same strict halfspace defined by H∗j−k, contradicting (21). To show the existence of such hyperplane consider the hyperplane HM in span(u1, . . . , um) passing through u1, . . . , um and denote with u∗ M the in- tersection of ru∗ with such hyperplane. Restricting our attention to the points in HM , the vectors u1, . . . , um form a simplicial polyhedron with M /∈ conv(u1, . . . , um). There must thus exist a k (cid:54)= j such that the (m− 2)- u∗ M and {ui}i /∈{j,k}, dimensional hyperplane HM−jk in HM , passing through u∗ supports conv(u1, . . . , um) and contains uj (and α) in one of its strict halfs- paces (see Appendix.) If we now consider the hyperplane in span(u1, . . . , um) passing through the origin and HM−jk we obtain the required hyperplane H∗j−k. To prove (23) we need some preliminary results. First of all, under (21), det( ¯U∗j) (cid:54)= 0 (24) Otherwise, u∗ would be coplanar to H−j and any hyperplane H∗j−k, k (cid:54)= j, would coincide with it. In such case the separating conditions (21) would not hold. Moreover, there must exist some other h (cid:54)= j for which det( ¯U∗h) (cid:54)= 0 (25) otherwise, u∗ would coincide with uj, making the result trivial. We also derive an equivalent condition for (21). Let ¯Uaj,bk be the m × m matrix obtained from ¯U by replacing vectors uj and uk by some other vectors, say ua and ub, respectively. If we move the first column to the k-th position, (21) becomes det( ¯U∗j,αk) det( ¯U∗j,jk) ≤ 0 for every k (cid:54)= j Switching positions j and k in the second matrix we get det( ¯U∗j,jk) = − det( ¯U∗k) and therefore det( ¯U∗j,αk) det( ¯U∗k) ≥ 0 for every k (cid:54)= j . (26) From (24), (25) and from part (i) of the present Theorem, we derive det( ¯U∗j) det( ¯U∗h) > 0 11 and therefore, (26) yields det( ¯U∗j,αk) det( ¯U∗j) = det( ¯U∗j,αk) det( ¯U∗h) det( ¯U∗j) det( ¯U∗h)−1 ≥ 0 for any k (cid:54)= j (27) Condition (27) and Theorem 1 allow us to conclude that α ∈ conv({uk}k(cid:54)=j, u∗). Regarding the last statement of the Theorem, we have already shown (24). Moreover, if (21) hold with a strict inequality sign for any k (cid:54)= j, then det( ¯U∗k) (cid:54)= 0 for the same k and u∗ ∈ ri(cone(u1, . . . , um)). Similarly, (27) would hold with strict inequality signs and α ∈ ri(cone({uk}k(cid:54)=j, u∗)). Remark 2. If (21) holds, we get not only that rα intersects the convex hull of the m EVVs {ui}i∈M\{j} ∪ {u∗}, but also that the ray ru∗ intersects the convex hull of the m EVVs u1, . . . , um. We can therefore replace uj with u∗ in the set of supporting EVVs for the lower bound. If the test fails for each j ∈ M , we discard u∗ we keep the current lower bound (with its supporting EVVs). In case (21) holds with an equality sign for some k, conditions (24) and (27) together imply det( ¯U∗j,αk) = 0. Therefore, we could discard uk from the set of supporting EVVs for the lower bound. Example 1 (Continued). We consider a list of 1'000 random vectors in ∆2 and, starting from the identity matrix, we iteratively pick each vector in the list. If this satisfies condition (21), then the matrix U is updated. The update occurs 9 times and the resulting EVVs which generate the matrix U are u1 = (0.5144, 0.5663, 0.3447) u2 = (0.4858, 0.5462, 0.4410) u3 = (0.4816, 0.3910, 0.6551) corresponding, respectively, to β1 = (0.4273, 0.2597, 0.3130) β2 = (0.4524, 0.3357, 0.2119) β3 = (0.4543, 0.3450, 0.2007) Correspondingly, the bounds shrink to 1.4792 ≤ vα ≤ 1.4898. 12 The previous example shows that updating the matrix U of EVVs through a random selection of the new candidates is rather inefficient, since it takes more than 100 new random vectors, on average, to find a valid replacement for vectors in U . A more efficient way method picks the candidate EVVs through some accurate choice of the corresponding values of β. In [6] a subgradient method is considered to find the value of vα up to any specified level of precision. In that algorithm, Legut's lower bound is used, but this can be replaced by the lower bound suggested by Theorem 1. Example 1 (Continued). Considering the improved subgradient algorithm, we obtain the following sharper bounds 1.48768 ≤ vα ≤ 1.48775 after 27 iterations of the algorithm in which, at each repetion, a new EVV is considered. Acknowledgements The authors would like to thank Vincenzo Acciaro and Paola Cellini for their precious help. Appendix The proof of (22) in Theorem 2 is based on the following Lemma. This is probably known and too trivial to appear in a published version of the present work. However, we could not find an explicit reference to cite it. Therefore, we state and prove the result in this appendix Lemma 1. Consider n affinely independent points u1, . . . , un in Rn−1 and u∗ /∈ conv(u1, . . . , un). For each j ∈ N there must exist a k (cid:54)= j such that the hyperplane passing through u∗ and {ui}i /∈{j,k} supports conv(u1, . . . , um) and has uj in one of its strict halfspaces. Proof. Suppose the thesis is not true. Then, for any h, (cid:96) ∈ N , the hyper- plane passing through u∗ and {ui}i∈N\{h,(cid:96)} will strictly separate the remaining points uh and u(cid:96). Fix now j ∈ N and consider H−j, the hyperplane passing through the points {ui}i∈N\{j}. Also denote as uj∗ the intersection between H−j and the line joining u∗ and uj. Clearly uj∗ /∈ conv({ui}i∈N\{j}). Therefore, for any k ∈ N \ {j}, the hyperplane H−jk in H−j passing through {ui}i∈N\{j,k} 13 will strictly separate uk and uj∗. Consequently, uj∗ should simultaneously lie in the halfspace of H−jk not containing conv({ui}i∈N\{j}), and in the cone formed by the other hyperplanes H−jh, h ∈ N \ {j, k} and not containing conv({ui}i∈N\{j}). A contradiction. References [1] Barbanel J (2000), On the structure of Pareto optimal cake partitions, J. Math. Econom. 33, No. 4, 401 -- 424. [2] Barbanel J (2005), The Geometry of Efficient Fair Division, Cambridge University Press. [3] Brams S J and Taylor A D (1996), Fair Division. From Cake-cutting to Dispute Resolution, Cambridge University Press. [4] Dall'Aglio M (2001), The Dubins-Spanier Optimization Problem in Fair Division Theory, J. Comput. Appl. Math., 130, No. 1 -- 2, 17 -- 40. [5] Dall'Aglio M, Branzei R and Tijs S H (2009), Cooperation in Dividing the Cake, TOP, 17, No.2, 417 -- 432. [6] Dall'Aglio M. and Di Luca C., Finding maxmin allocations in coopera- tive and competitive fair division, arXiv:1110.4241. [7] Dall'Aglio M and Hill T P (2003), Maximin Share and Minimax Envy in Fair-division Problems, J. Math. Anal. Appl., 281, 346 -- 361. [8] Dubins L E and Spanier E H (1961), How to cut a cake fairly, Amer. Math. Monthly 68, No.1, 1 -- 17. [9] Elton J, Hill T P and Kertz R P (1986), Optimal-partitioning Inequali- ties for Nonatomic Probability Measures, Trans. Amer. Math. Soc. 296, No.2, 703 -- 725. [10] Hiriart-Urruty J B and Lemar´echal C (2001), Fundamentals of Convex Analysis, Springer [11] Kalai E (1977), Proportional Solutions to Bargaining Situations: Inter- personal Utility Comparisons, Econometrica, 45, No. 77, 1623 -- 1630. [12] Kalai E and Smorodinsky M (1975), Other Solutions to Nash's Bargain- ing Problem, Econometrica, 43, No. 3, 513 -- 518. 14 [13] Legut J (1988), Inequalities for α-Optimal Partitioning of a Measurable Space, Proc. Amer. Math. Soc. 104, No. 4, 1249 -- 1251. [14] Legut J (1990), On Totally Balanced Games Arising from Cooperation in Fair Division, Games Econom. Behav., 2, No. 1, 47 -- 60. [15] Legut J, Potters J A M and Tijs S H (1994), Economies with Land -- A Game Theoretical Approach, Games Econom. Behav. 6, No. 3, 416 -- 430. [16] Legut J and Wilczynski M (1988), Optimal Partitioning of a Measurable Space, Proc. Amer. Math. Soc. 104, No.1, 262 -- 264. [17] Lyapunov A (1940), Sur les Fonctions-vecteurs Complet´ement Addi- tives, Bull. Acad. Sci. (URSS), No.4, 465 -- 478. 15
1004.4251
3
1004
2010-09-26T23:50:35
A Brezis-Browder theorem for SSDB spaces
[ "math.FA" ]
In this paper, we show how the Brezis-Browder theorem for maximally monotone multifunctions with a linear graph on a reflexive Banach space, and a consequence of it due to Yao, can be generalized to SSDB spaces.
math.FA
math
A Brezis -- Browder theorem for SSDB spaces by Stephen Simons Abstract In this paper, we show how the Brezis-Browder theorem for maximally monotone multi- functions with a linear graph on a reflexive Banach space, and a consequence of it due to Yao, can be generalized to SSDB spaces. 1 Introduction This note is about generalizations of the following results of Brezis-Browder and Yao: Let E be a nonzero reflexive Banach space with topological dual E ∗ and A be a norm -- closed monotone linear subspace of E × E ∗. Then A is maximally monotone ⇐⇒ A∗ is monotone ⇐⇒ A∗ is maximally monotone. In Theorem 4.4, we show that these results can be successfully generalized to SSDB space (see Section 2). The original results of Brezis -- Browder and Yao appear in Corollary 4.5. Theorem 4.4 depends on a new transversality result, which appears in Theorem 3.3. All vector spaces in the paper will be real. We will use the standard notation of convex analysis for "Fenchel conjugate" and "subdifferential" in a Banach space. 2 SSDB spaces for all b ∈ B, q(b) := 1 space (SSDB space) if B is a nonzero real vector space, ⌊·, ·⌋: B × B → R is a symmetric bilinear form, a Banach space, and ι is a linear isometry from B onto B ∗ such that, for all b, c ∈ B, Definition 2.1. We will say that (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) is a symmetrically self -- dual Banach ("q" stands for "quadratic"), (cid:0)B, k · k(cid:1) is (cid:10)b, ι(c)(cid:11) = ⌊b, c⌋. It is easily seen that if (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) is a SSDB space then so also is (cid:0)B, −⌊·, ·⌋, −q, k · k, −ι(cid:1). Clearly, for all b, c ∈ B, 2 ⌊b, b⌋ We refer the reader to Examples 2.2 and 2.3 below and [3] for various examples. There is also a construction that can be used for producing more pathological examples in [6, Remark 6.7, p. 20]. Let (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) be an SSDB space and A ⊂ B. We say that A is q -- positive if A 6= ∅ and b, c ∈ A =⇒ q(b − c) ≥ 0. We say that A is q -- negative if A 6= ∅ and b, c ∈ A =⇒ q(b − c) ≤ 0. q -- negativity is equivalent to (−q) -- positivity. We say that A is maximally q -- positive if A is q -- positive and A is not properly contained in any other q -- positive set. In this case, b ∈ B =⇒ inf q(A − b) ≤ 0. (2) Similarly, we say that A is maximally q -- negative if A is q -- negative and A is not properly contained in any other q -- negative set. Maximal q -- negativity is equivalent to maximal (−q) -- positivity. 1 (cid:12)(cid:12)q(b) − q(c)(cid:12)(cid:12) = 1 ≤ 1 2(cid:12)(cid:12)⌊b, b⌋ − ⌊c, c⌋(cid:12)(cid:12) = 1 2 kb − ck(cid:13)(cid:13)ι(b + c)(cid:13)(cid:13) = 1 2(cid:12)(cid:12)⌊b − c, b + c⌋(cid:12)(cid:12) = 1 2 kb − ckkb + ck. 2(cid:12)(cid:12)(cid:10)b − c, ι(b + c)(cid:11)(cid:12)(cid:12) ) (1) A Brezis -- Browder theorem for SSDB spaces Let g0 := 1 2 k · k2 on B. Then, for all c ∈ B, g0 Consequently, for all b, c ∈ B, 2 kι(c)k2 = 1 2 kck2 = g0(c). ∗(cid:0)ι(c)(cid:1) = 1 ι(c) ∈ ∂g0(b) ⇐⇒ g0(b) + g0 We note from (1) with c = 0 that ∗(cid:0)ι(c)(cid:1) =(cid:10)b, ι(c)(cid:11) ⇐⇒ g0(b) + g0(c) = ⌊b, c⌋. g0 + q ≥ 0 on B. (3) (4) Examples 2.2. (a) If B is a Hilbert space with inner product (b, c) 7→ hb, ci and the Hilbert space and every nonempty subset of B is q -- positive. norm then (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) is a SSDB space with ⌊b, c⌋ := hb, ci, q = g0 and ι(c) := c, norm then(cid:0)B, ⌊·, ·⌋, q, k·k, ι(cid:1) is a SSDB space with ⌊b, c⌋ := −hb, ci, q = −g0 and ι(c) := −c, (b) If B is a Hilbert space with inner product (b, c) 7→ hb, ci and the Hilbert space and the q -- positive sets are the singletons. (c) If B = R3 under the Euclidean norm and (cid:4)(b1, b2, b3), (c1, c2, c3)(cid:5) := b1c2 + b2c1 + b3c3, then (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) is a SSDB space, 3 and ι(c1, c2, c3) := (c2, c1, c3). Here, if M is any nonempty monotone subset of R × R (in the obvious sense) then M × R is a q -- positive subset of B. The set R(1, −1, 2) is a q -- positive subset of B which is not contained in a set M × R for any monotone subset of R × R. The helix q(b1, b2, b3) = b1b2 + 1 2 b2 Example 2.3. Let E be a nonzero reflexive Banach space with topological dual E ∗. For all (x, x∗) and (y, y∗) ∈ E × E ∗, let (cid:8)(cos θ, sin θ, θ): θ ∈ R(cid:9) is a q -- positive subset of B, but if 0 < λ < 1 then the helix (cid:8)(cos θ, sin θ, λθ): θ ∈ R(cid:9) is not. E × E ∗ by (cid:13)(cid:13)(x, x∗)(cid:13)(cid:13) := pkxk2 + kx∗k2. Then (cid:0)E × E ∗, k · k(cid:1)∗ under the duality (cid:10)(x, x∗), (y∗, y)(cid:11) := hx, y∗i + hy, x∗i. (cid:4)(x, x∗), (y, y∗)(cid:5) := hx, y∗i + hy, x∗i. We norm = (E ∗ × E, k · k(cid:1), (cid:10)(x, x∗), (y∗, y)(cid:11) =(cid:4)(x, x∗), (y, y∗)(cid:5). Thus (cid:0)E × E ∗, ⌊·, ·⌋, q, k · k, ι(cid:1) is a SSDB q(cid:0)(x, x∗) − (y, y∗)(cid:1). Thus if A ⊂ B then A is q -- positive exactly when A is a nonempty 2(cid:2)hx, x∗i + hx, x∗i(cid:3) = hx, x∗i and ι(y, y∗) = (y∗, y). monotone subset of B in the usual sense, and A is maximally q -- positive exactly when A is a maximally monotone subset of B in the usual sense. tions that space with q(x, x∗) = 1 We now note that if (x, x∗), (y, y∗) ∈ B then hx − y, x∗ − y∗i = q(x − y, x∗ − y∗) = It is clear from these defini- We define the reflection map ρ1: E × E ∗ → E × E ∗ by ρ1(x, x∗) := (−x, x∗). Since q ◦ ρ1 = −q, a subset A of B is q -- positive (resp. maximally q -- positive) if, and only if, ρ1(A) is q -- negative (resp. maximally q -- negative). ρ1 and its companion map ρ2 are used in the discussion of an abstract Hammerstein theorem in [5, Section 30, pp. 123 -- 125] and [7, Section 7, pp. 13 -- 15]. 2 A Brezis -- Browder theorem for SSDB spaces 3 A transversality result If f ∈ PC(B), we write f @ for the Fenchel conjugate of f with respect to the pairing ⌊·, ·⌋, that is to say, for all c ∈ B, Let B = (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) be a SSDB space. We write PC(B) for the set of all proper convex functions from B into ]−∞, ∞] and dom f for the set (cid:8)x ∈ B: f (x) ∈ R(cid:9). f @(c) := supB(cid:2)⌊·, c⌋ − f(cid:3). We note then that, for all c ∈ B, (5) f @(c) = supB(cid:2)h·, ι(c)i − f(cid:3) = f ∗(cid:0)ι(c)(cid:1). We write PCLSC(B) for the set {f ∈ PC(B): f is lower semicontinuous on B}. Lemma 3.1. Let (cid:0)B, ⌊·, ·⌋(cid:1) be a SSDB space, f ∈ PC(B) and c ∈ B. We define PC(B) by fc := f (· + c) − ⌊·, c⌋ − q(c). Then (fc)@ = (f @)c fc ∈ (6) and, writing f @ c for the common value of these two functions, for all b, d ∈ B, fc(b) + f @ c (d) − ⌊b, d⌋ = f (b + c) + f @(d + c) − ⌊b + c, d + c⌋. (7) Proof. For all b ∈ B, fc @(b) = supd∈B(cid:2)⌊d, b⌋ + ⌊d, c⌋ + q(c) − f (d + c)(cid:3) = supe∈B(cid:2)⌊e − c, b + c⌋ + q(c) − f (e)(cid:3) = supe∈B(cid:2)⌊e, b + c⌋ − ⌊c, b⌋ − f (e)(cid:3) − q(c) = f @(b + c) − ⌊c, b⌋ − q(c) = (f @)c(b), which gives (6), and (7) follows since fc(b) + f @ c (d) − ⌊b, d⌋ = f (b + c) − ⌊b, c⌋ − q(c) + f @(d + c) − ⌊d, c⌋ − q(c) − ⌊b, d⌋ = f (b + c) + f @(d + c) − ⌊b, c⌋ − ⌊c, c⌋ − ⌊d, c⌋ − ⌊b, d⌋ = f (b + c) + f @(d + c) − ⌊b + c, d + c⌋. (cid:3) Theorem 3.3 is the main result of this section, and uses the following consequence of Rockafellar's formula for the subdifferential of a sum (cid:0)see [4, Theorem 3, p. 86](cid:1). Lemma 3.2. Let X be a nonzero normed space, f : X 7→ ]−∞, ∞] be convex on X, finite at a point of X, and g: X 7→ R be convex and continuous. Then ∂(f + g) = ∂f + ∂g. Theorem 3.3. Let B =(cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) be a SSDB space, f ∈ PCLSC(B) and, when- ever b, d ∈ B, f (b) + f @(d) = ⌊b, d⌋ =⇒ q(b) + q(d) ≤ ⌊b, d⌋. (8) Let Nq(g0) :=(cid:8)b ∈ B: g0(b) + q(b) = 0(cid:9). Then dom f − Nq(g0) = B. 3 A Brezis -- Browder theorem for SSDB spaces Proof. Let c be an arbitrary element of B. Now fc +g0 ∞ as kbk → ∞, i.e., fc + g0 is "coercive". Since g0 is continuous and convex, fc + g0 ∈ PCLSC(B), hence fc + g0 is w(B, B ∗) -- lower semicontinuous. From [2, Proposition 1.3, p. (cid:0)fc +g0(cid:1)(b) → 892], B is reflexive, and so fc+g0 attains a minimum at some b ∈ B. Thus ∂(cid:0)fc+g0(cid:1)(b) ∋ 0. Since g0 is continuous, Lemma 3.2 implies that ∂fc(b) + ∂g0(b) ∋ 0, and so there exists d∗ ∈ ∂fc(b) such that −d∗ ∈ ∂g0(b). Since ι is surjective, there exists d ∈ B such that ι(d) = d∗. Then, from (3), is convex, and g0(b) + g0(d) = −⌊b, d⌋ (9) fc(b) + f @ c (d) = fc(b) + fc f (b + c) + f @(d + c) = ⌊b + c, d + c⌋. Since this gives q(b + c) + q(d + c) ≤ ⌊b + c, d + c⌋, or equivalently, and, using (5), (7), From (8), Adding this to (9), to say, b ∈ Nq(g0). Thus (g0 + q)(b) + (g0 + q)(d) ≤ 0. From (4), c ∈ dom f − Nq(g0), as required. ∗(d∗) = hb, d∗i = ⌊b, d⌋. Combining this with c ∈ dom f − b. q(b) + q(d) ≤ ⌊b, d⌋. that is (cid:3) (g0 + q)(b) = 0, f (b + c) ∈ R, Remark 3.4. While Theorem 3.3 is adequate for the application in Lemma 4.3, if we think of ∂f as a multifunction from B into B ∗, then the proof above actually establishes that D(∂f ) − Nq(g0) = B. 4 Linear q -- positive sets be an SSDB space and A be a linear subspace of B. Then we write A0 for the linear We now come to the main topic of this paper: linear q -- positive sets. Let(cid:0)B, ⌊·, ·⌋, q, k· k, ι(cid:1) subspace (cid:8)b ∈ B: ⌊A, b⌋ = {0}(cid:9) of B. Lemma 4.1. Let B =(cid:0)B, ⌊·, ·⌋, q, k·k, ι(cid:1) be a SSDB space and A be a maximally q -- positive linear subspace of B. Then A0 is a q -- negative linear subspace. Proof. If p ∈ A0 then inf q(A − p) = inf q(A) + q(p) = q(p), and so (2) gives q(p) ≤ 0. If b, c ∈ A0 then b − c ∈ A0 and so q(b − c) ≤ 0. (cid:3) Lemma 4.2. Let B = (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) be a SSDB space and A be a norm -- closed q -- positive linear subspace of B. Define the function qA: B → ]−∞, ∞] by qA := q on A and qA := ∞ on B \ A. Then qA ∈ PCLSC(B) and qA(b) + qA @(d) = ⌊b, d⌋ =⇒ b − d ∈ A0. (10) Proof. Suppose that a, c ∈ A and λ ∈ ]0, 1[ . Then λq(a) + (1 − λ)q(c) − q(cid:0)λa + (1 − λ)c(cid:1) = λ(1 − λ)q(a − c) ≥ 0. It is easily seen that this implies the convexity of qA. (cid:0)See [5, Lemma 19.7, pp. 80 -- 81].(cid:1) We know from (1) that q is continuous. Since A is norm -- closed in B, qA ∈ PCLSC(B). We now establish (10). Let @(d) = ⌊b, d⌋, a ∈ A and λ ∈ R. Then clearly b ∈ A and b+λa ∈ A, and so ⌊b+λa, d⌋−⌊λa, b⌋−λ2q(a) = q(b)+⌊b+λa, d⌋−q(b+λa) = @(d) = ⌊b, d⌋. Thus λ2q(a) + λ⌊b − d, a⌋ ≥ 0. qA(b) + ⌊b + λa, d⌋ − qA(b + λa) ≤ qA(b) + qA Since this inequality holds for all λ ∈ R, ⌊b − d, a⌋ = 0. Since this equality holds for all a ∈ A, we obtain (10). (cid:3) qA(b) + qA 4 A Brezis -- Browder theorem for SSDB spaces Lemma 4.3. Let B =(cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) be a SSDB space, A be a norm -- closed q -- positive linear subspace of B, and A0 be q -- negative. Then: (a) A − Nq(g0) = B. (b) A is maximally q -- positive. Proof. (a) It is clear from (10) that qA(b) + qA say, qA(b) + qA 3.3 with f = qA. @(d) = ⌊b, d⌋ =⇒ q(b − d) ≤ 0, that is to @(d) = ⌊b, d⌋ =⇒ q(b) + q(d) ≤ ⌊b, d⌋. (a) now follows easily from Theorem (b) Now suppose that c ∈ B and A ∪ {c} is q -- positive. From (a), there exists Since A ∪ {c} from which c = a ∈ A. This (cid:3) a ∈ A such that a − c ∈ Nq(g0). Thus is q -- positive, completes the proof of (b). 2 ka − ck2 + q(a − c) = 0. q(a − c) ≥ 0, and so 1 1 2 ka − ck2 ≤ 0, Our next result is suggested by Yao, [8, Theorem 2.4, p. 3]. Theorem 4.4. Let B = (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1) be a SSDB space and A be a norm -- closed q -- positive linear subspace (cid:0)of (cid:0)B, ⌊·, ·⌋, q, k · k, ι(cid:1)(cid:1). Then: (a) A is maximally q -- positive if, and only if, A0 is q -- negative. (b) A is maximally q -- positive if, and only if, A0 is maximally q -- negative. Proof. (a) is clear from Lemmas 4.1 and 4.3(b). "If" in (b) is immediate from "If" in (a). Conversely, let us suppose that A is maximally q -- positive. We know already from (a) that A0 is q -- negative, and it only remains to prove the maximality. Since A is norm -- closed, it follows from standard functional analysis and the surjectivity of ι that A = (A0)0. Now A0 is (−q) -- positive and norm -- closed. Furthermore, (A0)0 = A is q -- positive and thus If we now apply Lemma 4.3(b), with A replaced by A0 and q replaced (−q) -- negative. by −q, we see that A0 is maximally (−q) -- positive, that is to say, maximally q -- negative, which completes the proof of (b). (cid:3) Now let A be a monotone linear subspace of E × E ∗. Then the linear subspace A∗ of It is clear (x, x∗) ∈ A∗ ⇐⇒ for all (a, a∗) ∈ A, hx, a∗i = ha, x∗i. E × E ∗ is defined by: then that A0 = ρ1(cid:0)A∗(cid:1). Corollary 4.5(a) below appears in Brezis -- Browder [1, Theorem 2, pp. 32 -- 33], and Corollary 4.5(b) in Yao, [8, Theorem 2.4, p. 3]. Corollary 4.5. Let E be a nonzero reflexive Banach space with topological dual E ∗ and A be a norm -- closed monotone linear subspace of E × E ∗. Then: (a) A is maximally monotone if, and only if, A∗ is monotone (b) A is maximally monotone if, and only if, A∗ is maximally monotone. Proof. These results follow from Theorem 4.4 and the comments in Example 2.3. (cid:3) References [1] H. Brezis and F. E. Browder, Linear maximal monotone operators and singular nonlin- ear integral equations of Hammerstein, type, Nonlinear analysis (collection of papers in honor of Erich H. Rothe), pp. 31 -- 42. Academic Press, New York, 1978. [2] J-E. Mart´ınez-Legaz, On maximally q -- positive sets, J. of Convex Anal., 16 (2009), 891 -- 898. 5 A Brezis -- Browder theorem for SSDB spaces [3] J-E. Mart´ınez-Legaz, Private communication. [4] R. T. Rockafellar, Extension of Fenchel's duality theorem for convex functions, Duke Math. J. 33 (1966), 81 -- 89. [5] S. Simons, From Hahn -- Banach to monotonicity, Lecture Notes in Mathematics, 1693, second edition, (2008), Springer -- Verlag. [6] -- -- , Banach SSD spaces and classes of monotone sets, http://arxiv/org/abs/ 0908.0383v3, posted July 5, 2010. [7] -- -- , SSDB spaces and maximal monotonicity, http://arxiv/org/abs/1001.0064v2, posted June 17, 2010. [8] L. Yao, The Brezis -- Browder theorem revisited and properties of Fitzpatrick functions of order n, http://arxiv.org/abs/0905.4056v1, posted May 25, 2009. Department of Mathematics University of California Santa Barbara CA 93106-3080 U. S. A. email: [email protected] 6
1612.05410
1
1612
2016-12-16T10:00:49
Topological aspects of order in $C(X)$
[ "math.FA" ]
In this paper we consider the relationship between order and topology in the vector lattice $C_b(X)$ of all bounded continuous functions on a Hausdorff space $X$. We prove that the restriction of $f\in C_b(X)$ to a closed set $A$ induces an order continuous operator iff $A=\overline{\mathrm{Int} A}.$ This result enables us to easily characterize bands and projection bands in $C_0(X)$ and $C_b(X)$ through the one-point compactification and the Stone-\v{C}ech compactification of $X$, respectively. With these characterizations we describe order complete $C_0(X)$ and $C_b(X)$-spaces in terms of extremally disconnected spaces. Our results serve us to solve an open question on lifting un-convergence in the case of $C_0(X)$ and $C_b(X)$.
math.FA
math
TOPOLOGICAL ASPECTS OF ORDER IN C(X) M. KANDI ´C AND A. VAVPETI C Abstract. In this paper we consider the relationship between or- der and topology in the vector lattice Cb(X) of all bounded contin- uous functions on a Hausdorff space X. We prove that the restric- tion of f ∈ Cb(X) to a closed set A induces an order continuous operator iff A = Int A. This result enables us to easily characterize bands and projection bands in C0(X) and Cb(X) through the one- point compactification and the Stone- Cech compactification of X, respectively. With these characterizations we describe order com- plete C0(X) and Cb(X)-spaces in terms of extremally disconnected spaces. Our results serve us to solve an open question on lifting un-convergence in the case of C0(X) and Cb(X). 1. Introduction The interaction between order and topology on spaces of continuous functions have been studied extensively in the past. Although closed ideals, bands and projection bands in C(X) where X is compact and Hausdorff have nice characterizations through closed sets of X, the lattice C(X) is very rarely order complete. Order completeness of the lattice C(X) is equivalent to the fact that X is extremally disconnected [MN91]. In Section 3 we consider the operator ΦA : C(X) → C(A) which maps continuous functions on X to their restrictions on A ⊆ X. Al- though this operator is a contraction between Cb(X) and Cb(A), in 2010 Mathematics Subject Classification. Primary: 46B42, 46E25, 54D15. Sec- ondary: 46A40, 54D10, 54G05. Key words and phrases. vector lattices, continuous functions, separation axioms, bands and projection bands, order continuity, un-convergence. This research was supported by the Slovenian Research Agency grants P1-0222- 0101, P1-0292-0101, J1-6721-0101 and J1-7025-0101. 1 2 M. KANDI ´C AND A. VAVPETI C general it is not order continuous. We prove that whenever X ∈ T3 1 order continuity of ΦA is equivalent to A ⊆ Int A (see 3.3). 2 Although order properties of C(X) where X is a compact Hausdorff space can be found in the main textbooks on vector and Banach lat- tices, results for Cb(X) where X is "just" Hausdorff seem to be left out of the main literature. Our second goal is to provide an elemen- tary characterization of closed ideals, bands and projection bands of Banach lattices Cb(X) and C0(X) through closed sets of X where X is Hausdorff and locally compact, respectively. Main tools we use are order continuity of the restriction operator ΦA and the embeddings of X into X + or βX when X is locally compact Hausdorff space or T3 1 , respectively. There is also a way how to study Cb(X) where X is just Hausdorff. If X is T3 1 , then bands and projection bands have desired characterizations, and when X is only Hausdorff, desired characteri- zations can be obtained only for projection bands (see 5.6). Certain results of Section 4 and Section 5 are applied in Section 6 to charac- terize order complete vector lattices C0(X) and Cb(X). We prove that , respec- C0(X) or Cb(X), when X is locally compact Hausdorff or T3 1 tively, is order complete iff X is extremally disconnected. This enables us to provide a vector lattice argument that X ∈ T3 1 is extremally disconnected iff its Stone- Cech compactification βX is. 2 2 2 2 In Section 7 we consider the so-called problem of "lifting un-convergence" [KMT, Question 4.7]. An application of results from Section 4 and Sec- tion 5 provides a positive answer in the special case of Banach lattices C0(X) and Cb(X). 2. Preliminaries Let E be a vector lattice. A vector subspace Y of E is an ideal or an order ideal whenever 0 ≤ x ≤ y and y ∈ Y imply x ∈ Y . A band B is an ideal with the property that whenever 0 ≤ xα ր x and xα ∈ B for each α imply x ∈ B. Here, the notation xα ր x means that the net (xα) is increasing towards its supremum x. In the literature bands are sometimes referred to as order closed ideals. For a given set A ⊆ E by Ad we denote the set {x ∈ E : x ∧ a = 0 for all a ∈ A}. The set Ad is called the disjoint complement of A in E. It turns out that TOPOLOGICAL ASPECTS OF ORDER IN C(X) 3 Ad is always a band in E. If a band B satisfies E = B ⊕ Bd, then B is called a projection band . If every band in E is a projection band, then E is said to have the projection property . A vector lattice is said to be order complete whenever nonempty bounded by above sets have suprema. It is well-known that order complete vector lattices have the projection property. The spaces Lp(µ) (0 ≤ p ≤ ∞) are order complete whenever µ is a σ-finite measure. When p 6= 0 or ∞, then Lp(µ) is always order complete. On the other hand, the lattice C(X) is rarely order complete. If (fα) is an increasing net in C(X) that is bounded by above, then f = supα fα exists in the vector lattice of all functions on X. If f is continuous, then fα ր f in C(X). A sublattice Y of a vector lattice E is said to be regular if for every subset A of Y , inf A is the same in E and in Y whenever inf A exists in Y . It is well-known (see e.g. [AB03]) that every ideal is a regular sublattice. The following result characterizes regular sublattices (see e.g. [AB03, Theorem 1.20]). Lemma 2.1. For a sublattice Y of E the following statements are equivalent. (a) Y is regular. (b) If sup A exists in Y then sup A exists in E and the two suprema are equal. The following proposition directly follows from Lemma 2.1. Proposition 2.2. Let Y be an ideal in a vector lattice E. If B ⊆ Y is a band in E, then B is a band in Y . Proof. Since every band is an ideal, B is an ideal in E, and hence B = B ∩ Y is an ideal in Y . Suppose now that a net (xα) satisfies 0 ≤ xα ր x in Y . By Lemma 2.1 we have xα ր x in E. That x ∈ B follows from the fact that B is a band in E. Hence, B is a band in Y . (cid:3) In this paper we are concerned with algebras/lattices of continuous functions on a topological space X. For a set A of X we will denote by IntX A and ClX A the interior and the closure of A in X, respectively. If there is no confusion in which space the interior and the closure are 4 M. KANDI ´C AND A. VAVPETI C taken, we will simply write Int A and A instead of IntA X and ClX A, respectively. If A is an open subset of X and B is a subset of A, then IntA B = IntX B. A topological space is said to be completely regular for every closed set F in X and x ∈ X \ F there exists a continuous function f : X → [0, 1] such that f (x) = 1 and f F ≡ 0. If X is completely regular and Hausdorff, we say that X is T3 1 and we write X ∈ T3 1 It is well-known that locally compact . A space X is said to be regular whenever a Hausdorff spaces are T3 1 point and a closed set which does not contain the point can be separated by disjoint neighborhoods. If disjoint closed sets of X can be separated by disjoint neighborhoods, then X is said to be normal . for short. 2 2 2 If not otherwise stated all functions are assumed to be continuous. Since real function rings (resp. algebras) C(X), Cb(X) and C0(X) are also lattices, we introduce the following notation to distinguish between algebraic and lattice notions of ideals. A ring (resp. algebra) ideal is called an r-ideal (resp. an a-ideal ) and an order ideal is called an o-ideal . Since all constant functions on X are in Cb(X) (resp. C(X)) the class of r-ideals in Cb(X) (resp. C(X)) coincides with the class of a-ideals. The ideal J in C(X) is called fixed whenever all functions from J vanish on a nonempty set. The ideals that are not fixed are called free. Fixed ideals play an important role in the characterization -spaces. For the proof of the of compact Hausdorff spaces among T3 1 following Theorem see [GJ76, Theorem 4.11]. 2 Theorem 2.3. For X ∈ T3 1 2 the following assertions are equivalent. (a) X is compact. (b) Every r-ideal in Cb(X) is fixed. (c) Every maximal r-ideal in Cb(X) is fixed. Special examples of fixed ideals are the ideals of the form JF (X) := {f ∈ C(X) : f F ≡ 0} for some subset F of X. Obviously JF (X) is an a-ideal and an o-ideal which is closed in C(X) in the topology of pointwise convergence. It should be obvious that we always have JF (X) = JF (X). For F ⊆ X we denote the set {f ∈ C0(X) : f F ≡ 0} by J 0 F (X). TOPOLOGICAL ASPECTS OF ORDER IN C(X) 5 If X is a compact Hausdorff space there exists a characterisation of closed ideals, bands and projective bands in C(X). We start with a characterization of closed ideals (see e.g. [MN91, Proposition 2.1.9]). Theorem 2.4. For every closed subspace J of C(X) where X is a compact Hausdorff space the following assertions are equivalent. (a) J is an r-ideal. (b) J is an o-ideal. (c) J = JF (X) for some closed set F ⊆ X. Bands and projection bands are characterized as follows (see e.g. [MN91, Corollary 2.1.10]). Theorem 2.5. For a subspace B of C(X) where X is a compact Haus- dorff space the following assertions hold. (a) B is a band iff B = JF (X) for some set F ⊆ X which is a closure of some open set. (b) B is a projection band iff B = JF (X) for some clopen set F ⊆ X. Corollary 2.6. If X is a locally compact noncompact Hausdorff space, then C0(X) is not a band in C(X +). A closed set which is a closure of some open set is the closure of its interior. Indeed, if F = U where U is an open set, then U ⊆ Int F , so that F = U ⊆ Int F ⊆ F . Thus, a closed ideal J in C(X) where X is a compact Hausdorff space is a band iff J = JF (X) for some closed set F in X with F = Int F . For unexplained facts about vector lattices and operators acting on them we refer the reader to [LZ71]. 3. Order density and completely regular spaces Complete regularity of the space basically means that we can sepa- If we can sep- rate points from closed sets by continuous functions. arate only points by continuous functions, i.e., for different points x, y ∈ X there is a continuous function f : X → [0, 1] with f (x) = 0 and f (y) = 1, then X is called functionally Hausdorff . 6 M. KANDI ´C AND A. VAVPETI C If X is not functionally Hausdorff, then there exist x, y ∈ X such that f (x) = f (y) for all f ∈ C(X). In this case we have J{x}(X) = J{y}(X), so that the mapping x 7→ J{x}(X) is not one-to-one. If X ∈ T3 1 , then for closed sets F and G in X we have JF (X) = JG(X) iff F = G. Indeed, if F 6= G, then there exists x ∈ X such that x ∈ F \ G or x ∈ G \ F . Without any loss of generality we can assume that the former happens. Then one can find f : X → [0, 1] with f (x) = 1 and f ≡ 0 on G. Then f ∈ JG(X) = JF (X) implies f (x) = 0 which is a contradiction. 2 Also, when F is a closed set in X ∈ T3 1 , the disjoint complement of JF (X) satisfies JF (X)d = JX\F (X). Indeed, if f ∈ JX\F (X), then f ≡ 0 on X \ F , so that f ∧ g = 0 for all g ∈ JF (X). Hence f ∈ JF (X)d. If f /∈ JX\F (X), there is x ∈ X \ F such that f (x) 6= 0. Take any g ∈ Cb(X) with g(x) = 1 and g ≡ 0 on F . Then g ∈ JF (X) and f ∧ g 6= 0, i.e., f /∈ JF (X)d. 2 In this section we are interested in the so-called "restriction operator" defined as follows. If A is a subset of X, then the restriction f A of f ∈ C(X) is continuous on A. The mapping ΦA(f ) = f A is called the restriction operator . Also, if f ∈ Cb(X), then f A ∈ Cb(A) and the restriction operator is a contractive lattice homomorphism with respect to the k · k∞ norms on Cb(X) and Cb(A). If X is normal and A is closed in X, then the restriction operator is always surjective. The details follow. Theorem 3.1. The following assertions are equivalent. (a) X is normal. (b) For any closed subset A in X the restriction operator ΦA : Cb(X) → Cb(A) is surjective. (c) For any disjoint closed sets A and B in X there is f ∈ Cb(X) such that f A ≡ 1 and f B ≡ 0. It should be noted that (b) and (c) are precisely the statements of Tietze extension theorem and Urysohn's lemma, respectively. If fα ր f in C(X), then for each set A ⊆ X we have fαA ր and fαA ≤ f A for each α. Pick λ ∈ R. If fαA ≡ λ for each α, then f A ≡ λ. In general we cannot expect fαA ր f A, since ΦA is not order TOPOLOGICAL ASPECTS OF ORDER IN C(X) 7 continuous. We will prove that for a closed set A ⊆ X the operator ΦA : C(X) → C(A) is order continuous iff A = Int A (see 3.3). To prove this result we first need the following lemma which deals with a special case. Proposition 3.2. Let X ∈ T3 1 2 and A ⊆ X open subset. (a) If fα ր f in C(X), then fαA ր f A in C(A). (b) If fα ր f in Cb(X), then fαA ր f A in Cb(A). Proof. (a) Let fα ր f in C(X). Suppose there exists g : A → R such that fαA ≤ g < f A for all α. There exists x0 ∈ A such that g(x0) < f (x0) and because f and g are continuous and X is regular, there exists an open set U ⊆ ClX(U) ⊆ A such that g(x) < f (x) for all x ∈ U. Because X ∈ T3 1 there exists λ : X → [0, 1] such that λ(x0) = 1 and λ(X \ U) = 0. We define a function λg : X → R as 2 (λg)(x) =( λ(x)g(x) 0 : x ∈ A : x ∈ X \ ClX U . Since {A, X \ ClX(U)} is an open cover of X and λ ≡ 0 on X \ ClX U we conclude that λg is continuous on X. Then h : X → R defined by h(x) = (λg)(x) + (1 − λ(x))f (x) is continuous and fα ≤ h < f for all α. This is a contradiction as fα ր f . (b) Suppose fα ր f in Cb(X). Since Cb(X) is an ideal in C(X), Lemma 2.1 implies fα ր f in C(X), so that by (a) we have fαA ր f A in C(X). If fαA ≤ g ≤ f A for some g ∈ Cb(A), then fαA ր f A in C(X) implies g = f A, and since f A ∈ Cb(A) we have fαA ր f A in Cb(A). (cid:3) In 5.7 we will see that there is an example of a functionally Hausdorff space X such that for every open set A in X the restriction operator ΦA : C(X) → C(A) is order continuous. Theorem 3.3. Let X ∈ T3 1 equivalent: 2 and A ⊆ X a closed set. The following is (a) A = Int A. (b) If fα ր f in C(X), then fαA ր f A in C(A). (c) If fα ր f in Cb(X), then fαA ր f A in Cb(A). 8 M. KANDI ´C AND A. VAVPETI C Proof. (a) ⇒ (b) Suppose A = Int A and let fα ր f in C(X). Let g : A → R be such that fαA ≤ g ≤ f A for all α. By 3.2, fαInt A ր f Int A in C(Int A), hence gInt A = f Int A. Because of continuity of f and g (on A) and A = Int A we get g = f A. (b) ⇒ (a) Suppose A * Int A. There exists x0 ∈ A \ Int A. Because X ∈ T3 1 2 there exists eψ : X → [−1, 1] such that eψ(x0) = −1 and eψ(Int A) = 1. Let ψ : X → [0, 1] be defined as ψ = max{eψ, 0}. Then U = eψ−1([−1, 0)) ∩ A is an open set in A. Because A is closed, U ⊆ A and U ⊆ eψ−1([−1, 0]), hence x0 ∈ U ⊆ U ⊆ A \ Int A. Let M = {f : X → R : f U ≡ 0, f Int A ≡ 1}. By the construction we have ψ ∈ M, so that M is nonempty. Claim 1: If f ≤ g for all f ∈ M then g(x) ≥ 1 for all x ∈ X \ U. Indeed, pick x 6∈ U . Because X ∈ T3 1 there exists ϕ : X → [0, 1] such that ϕ ≡ 0 on U and ϕ(x) = 1. Then f := min{ϕ + ψ, 1} ∈ M and f (x) = 1, which implies g(x) ≥ f (x) = 1. 2 Claim 2: The set X \ U is dense in X. Indeed, otherwise there exists an open set V ⊆ X such that V ∩ (X \ U) = ∅. Then V ⊆ U ⊆ A\ Int A. This is a contradiction, since from V ⊆ A and V open follows V ⊆ Int A. Let us show that (f )f ∈M ր 1 in C(X). Suppose some function g : X → R satisfies f ≤ g ≤ 1 for all f ∈ M. Claim 1 yields that g ≥ 1 on X \ U, so that together with g ≤ 1 we conclude gX\U ≡ 1. Since by Claim 2 the set X \ U is dense in X, we have g ≡ 1 on X. Therefore, (f )f ∈M ր 1 in C(X). Let us show that (f A)f ∈M does not converge in order to 1 in C(A). If (f A)f ∈M ր 1, by 3.2 (f U )f ∈M ր 1. But f U ≡ 0 for all f ∈ M. (b) ⇒ (c) Suppose fα ր f in Cb(X). Then fα ր f in C(X), so that fαA ր f A in C(A). Since f A is bounded, we have fαA ր f A in Cb(A). (c) ⇒ (b) Since for each g ∈ C(X) we have g ∧ n1 րn g and since for each n ∈ N we have fαA ∧ n1 րα f A ∧ n1, we conclude f A = sup n∈N = sup α (f A ∧ n1) = sup n∈N sup α (fαA ∧ n1) sup n∈N (fαA ∧ n1) = sup n∈N fαA. TOPOLOGICAL ASPECTS OF ORDER IN C(X) 9 (cid:3) Corollary 3.4. Let A be a closed set in X ∈ T3 1 operator ΦA is order continuous iff A = Int A. 2 . Then the restriction 2 Suppose now that X is compact and Hausdorff; hence normal and . If A is closed in X, then the kernel of ΦA : C(X) → C(A) so X ∈ T3 1 is precisely the fixed ideal JA(X). In this special case, by 2.5 and 3.4 the restriction operator ΦA is order continuous iff JA(X) is a band in C(X). When X is merely normal, the restriction operator ΦA is surjective by 3.1. By [LZ71, Theorem 18.13] ΦA : Cb(X) → Cb(A) is order continuous iff ker ΦA = JA(X) is a band in Cb(X). In 5.3 we will prove that for X ∈ T3 1 bands in Cb(X) are precisely closed ideals in Cb(X) of the form JA(X) where A is a closed set satisfying A = Int A. 2 2 If we only assume X ∈ T3 1 , by 3.1 the restriction operator ΦA is not surjective in general. When A = Int A, then 5.3 implies JA(X) is a band in Cb(X) and ΦA : Cb(X) → Ran ΦA ⊆ Cb(A) is an order con- tinuous lattice homomorphism by [LZ71, Theorem 18.13]. Therefore, if fα ր f in Cb(X), then fαA ր f A in the sublattice Ran ΦA. Since Ran ΦA is order dense in Cb(A) (see 3.5 below), then it is also a regular sublattice of Cb(A) and hence by Lemma 2.1 we have fαA ր f A in Cb(A). Proposition 3.5. Let X ∈ T3 1 restriction operator. Then Ran ΦA is order dense in Cb(A). , A ⊆ X, and ΦA : Cb(X) → Cb(A) the 2 Proof. Pick a nonzero nonnegative function f ∈ C(A). Let a ∈ A be such that f (a) > 0. Because f is continuous there exists an open set U in A such that f (x) > f (a) for all x ∈ U. Let V be open in X such that 2 V ∩ A = U. Because X ∈ T3 1 2 ] such that g(a) = f (a) 2 ≥ g(x) and for x ∈ A \ U we have f (x) ≥ 0 = g(x). Hence 0 ≤ gA ≤ f . (cid:3) 2 and g ≡ 0 on V C. For x ∈ U we have f (x) > f (a) there exists g : X → [0, f (a) 2 4. Bands in C0(X) In this section we are interested in closed ideals of the Banach lattice algebra C0(X) where X is a locally compact Hausdorff space. We 10 M. KANDI ´C AND A. VAVPETI C characterize closed ideals, bands and projection bands of C0(X) (see 4.1 and 4.2). The characterizations are pretty much the same as the characterization of closed ideals, bands and projection bands of the algebra of all continuous functions on a given compact Hausdorff space (see 2.4 and 2.5). Our arguments presented in this section are based on the embedding of X into its one-point compactification X +. Before we proceed to our results of this section we briefly recall basic facts about the topology on X +. Let X be a locally compact Hausdorff space and pick any object ∞ which is not in X. It is well-known that the family τ of all open sets in X together with the sets in X + containing ∞ whose complements are compact in X is a topology on X + and the space (X +, τ ) is a compact Hausdorff space. The space X + is called the one-point compact- ification of X. By [Dug66, 8.3 p.245] the embedding X ֒→ X + is an open mapping; in particular, X is an open subset of X +. If X is noncompact, then X is dense in X +, and if X is compact, then X is clopen in X + and ∞ is isolated in X +. Suppose now that X is not compact and f ∈ C0(X). Then f has the unique extension ef ∈ C(X +) on X + given by : x ∈ X : x = ∞ . ef (x) =( f (x) 0 The mapping f 7→ ef is an algebra and lattice isometric isomorphism be- tween C0(X) and the closed maximal lattice and algebra ideal J{∞}(X +) in C(X +). For brevity purposes we will write J{∞} instead of J{∞}(X +). The closed lattice (resp. algebra) ideals of C0(X) are therefore in bi- jective correspondence with the closed lattice (resp. algebra) ideals of C(X +) that consist of functions that vanish at infinity. As was already mentioned closed ideals are characterized similarly as the closed ideals in C(K)-spaces. Theorem 4.1. Let X be a locally compact Hausdorff space and J a closed subspace in C0(X). The following assertions are equivalent: (a) J is an a-ideal in C0(X). (b) J is an o-ideal in C0(X). TOPOLOGICAL ASPECTS OF ORDER IN C(X) 11 (c) J = J 0 F (X) for some closed set F in X. Proof. When X is compact the conclusion follows from 2.4, so that we may assume X is not compact. By [Kan09, Theorem 1.4.6] (a) and (c) are equivalent. It is trivial that (c) implies (b). To finish the proof we prove that (b) implies (a). Since C0(X) is isometrically lattice isomorphic to J{∞}, J is isometrically lattice iso- morphic to some closed lattice ideal J ′ = JF ′(X +) of C(X +) for some closed subset F ′ ⊆ X + which contains the point ∞. By 2.4 J ′ is also an a-ideal, from where it follows that J = {f X : f ∈ J ′} is an a-ideal in C0(X). (cid:3) We proceed by a characterization of bands and projection bands in C0(X). Our proofs presented here rely on the embedding X ֒→ X +. Theorem 4.2. Let X be a locally compact Hausdorff space and J a closed ideal in C0(X). (a) J is a band iff J = J 0 F (X) for some closed set F in X with F = Int F . (b) J is a projection band iff J = J 0 F (X) for some clopen set F in X. Proof. (a) Let F = ClX(Int F ) and let 0 ≤ fα ր f , where fα ∈ J 0 F (X) F (X). in C(X +). If ClX +(Int F ) = F , then JF (X +) is a band in C(X +), F (X). If ClX +(Int F ) = F ∪ {∞}, then ClX +(Int(F ∪{∞})) = F ∪{∞}, hence JF ∪{∞}(X +) is a band in C(X +), and f ∈ C0(X). Denote by efα : X + → R the extension of fα on X +. Then 0 ≤ efα ր ef in J{∞}, so that by Lemma 2.1 efα ր ef hence ef ∈ JF (X +) and f ∈ J 0 ef ∈ JF ∪{∞}(X +), and f ∈ J 0 If ClX (Int F ) ( F , then there exists x ∈ F \ ClX(Int F ). Let U be an open set such that ∞ ∈ U ⊆ U ⊆ {x}C. We claim that x ∈ F ∪ U \ ClX +(Int(F ∪ U)). Since x /∈ ClX (Int F ) there exists an open neighborhood W1 of x such that W1 ∩ Int F = ∅. For W2 := X + \ U 12 M. KANDI ´C AND A. VAVPETI C we have x ∈ W2 and W1 ∩ W2 ∩ Int(F ∪ U ) ⊆ W1 ∩ W2 ∩ (F ∪ U ) = (W1 ∩ W2 ∩ F ) ∪ (W1 ∩ W2 ∩ U ) = W1 ∩ W2 ∩ F. Since the set W1 ∩ W2 ∩ Int(F ∪ U ) is open, we have W1 ∩ W2 ∩ Int(F ∪ U ) ⊆ Int(W1 ∩ W2 ∩ F ) = W1 ∩ W2 ∩ Int F = (W1 ∩ Int F ) ∩ W2 = ∅. Therefore, W1 ∩ W2 is an open neighborhood of x which does not in- tersect Int(F ∪ U ), hence x /∈ ClX + Int(F ∪ U ). Because F ∪ U is closed in X + and F ∪ U 6= ClX +(Int(F ∪ U)), there exists a net (fα)α∈Λ ⊆ JF ∪U (X +) such that fα ր f ∈ C(X +) \ JF ∪U (X +). Then fαX ∈ J 0 F (X) and fαX ր f X ∈ C0(X). Since fαU ≡ 0, 3.2 implies f U ≡ 0, so that by continuity of f we have f U ≡ 0. Now it follows that f F 6≡ 0, hence J 0 F (X) is not a band in C0(X). F (X), f2 := f χF ∈ J 0 F (X)d. (b) If F is clopen, then X \ F is clopen as well, so that the char- acteristic functions χF and χX\F are continuous on X. Then f1 := f χX\F ∈ J 0 F (X)d and f = f1 + f2 imply C0(X) = F (X) ⊕ J 0 J 0 Suppose now that J is a projection band in C0(X). By (a) we have J = J 0 F (X) for some closed set F in X. We claim X \ F is closed in X. Otherwise there is x0 ∈ F ∩ X \ F . Pick any function f ∈ C0(X) with f (x0) = 1. Since J is a projection band, we have C0(X) = J 0 F (X) ⊕ F (X) and f2 ∈ J 0 J 0 (X), X\F so that f (x0) = 0. This contradiction shows X \ F is closed. Hence F is clopen and the proof is finished. (cid:3) (X). Hence, f = f1 + f2 for some f1 ∈ J 0 X\F Closed ideals in C0(X) which are projection bands in C(X +) are more interesting. Not only that the set F which defines a given closed ideal is clopen, its complement in X is compact. Corollary 4.3. Let X be a locally compact Hausdorff space and J 0 a closed ideal in C0(X). Then J 0 iff F is closed in X and X \ F is compact. F (X) F (X) is a projection band in C(X +) TOPOLOGICAL ASPECTS OF ORDER IN C(X) 13 F (X) is a projection band in C(X +). Then Proof. Suppose first that J 0 F ∪ {∞} is clopen in X +, from where it follows that F is clopen in X. Since X \ F = X + \ (F ∪ {∞}) is also closed in X +, the set X \ F is compact. Suppose now that F is closed in X and X \ F is compact. Since F is closed in X it follows that F ∪ {∞} is closed in X +. Since X \ F is compact, F ∪ {∞} is also open in X +. Therefore, F ∪ {∞} is clopen F (X) ∼= JF ∪{∞}(X +) is a projection band in C(X +) in X +, so that J 0 by 2.5(b). (cid:3) Suppose J is an ideal in C0(X). If J is a band in C(X +), then J is a band in C0(X) by 2.2. This fact can be also proved directly with results of this section without referring to general result on vector lattices. Indeed, if J 0 F (X) in C0(X) is band in C(X +), then 2.5 implies that F ∪ {∞} = ClX +(Int(F ∪ {∞})). Since we have IntX F = IntX + F = IntX +(F ∪ {∞}), we have ClX IntX F = X ∩ ClX + IntX F = X ∩ ClX + IntX +(F ∪ {∞}) = X ∩ (F ∪ {∞}) = F. By 4.2 J 0 F (X) is a band in C0(X). 5. Bands in Cb(X) 2 2 . If for some X ∈ T3 1 In this section we are interested in extending 2.4 and 2.5 to the case of the algebra Cb(X) where X is a Hausdorff space. To have a nice characterization of closed ideals of Cb(X) in terms of closed sets from X is too optimistic even when X ∈ T3 1 every closed ideal in Cb(X) is of the form JF (X) for some closed set F in X, then every closed ideal in Cb(X) is fixed, so that by 2.3 X is compact. Hence, if X is a noncompact space, then there exist closed ideals in Cb(X) which are not of the form JF (X) for some closed set F in X. However, if X ∈ T3 1 then bands and projection bands in Cb(X) have similar characterizations as bands in C(K)-spaces (see 5.3). If X is merely Hausdorff, then only projection bands in Cb(X) have a similar characterization as those in C(K)-spaces (see 5.5). Unfortunately, 5.6 2 14 M. KANDI ´C AND A. VAVPETI C shows that there exists a functionally Hausdorff space X and a band JF (X) in Cb(X) with Int F 6= F . The following topological facts are critical in our study of bands and projection bands in Cb(X). 2 • For every space X ∈ T3 1 there exists a compact Hausdorff space βX with the property that the inclusion △ : X ֒→ βX is an embedding and βX contains X as a dense subset. Furthermore, every function f ∈ Cb(X) can be uniquely extended to the continuous function f β ∈ C(βX). • For every Hausdorff space X there exists X/∼ ∈ T3 1 and a con- tinuous surjective mapping τ : X → X/∼ such that the mapping 2 Φτ : Cb(X/∼) → Cb(X) defined as Φτ (ef ) = ef ◦ τ is an isomet- ric isomorphism between Banach lattice algebras Cb(X/∼) and Cb(X). We will recall the construction later in this section. It is well-known that the compact Hausdorff space βX is called the Stone- Cech compactification of X. By [ABo06, Theorem 3.70] the inclusion △ : X ֒→ βX is an open mapping whenever X is locally compact and Hausdorff; in particular, in this case X is open in βX. The mapping β : Cb(X) → C(βX) defined as β(f ) = f β is an iso- metric isomorphism of Banach algebras. By [GJ76, Theorem 1.6] the mapping β is also a lattice isomorphism. Therefore, J is a closed or- der/algebra ideal (resp. band, projection band) in Cb(X) iff β(J) is a closed ideal (resp. band, projection band) in C(βX). Similarly, given a , and the iso- Hausdorff space X, the aforementioned space X/∼ is T3 1 morphism between Banach algebras Cb(X) and Cb(X/∼) is also a lattice isomorphism, again by [GJ76, Theorem 1.6]. Therefore, starting with a subspace J of Cb(X) where X is Hausdorff we deduce that J is a closed order/algebra ideal in Cb(X) iff Φ−1 τ (J) is a closed order/algebra ideal in Cb(X/∼) iff (β ◦ Φ−1 τ )(J) is a closed order/algebra ideal in C(βX). Similar statements hold for bands and projection bands. 2 Therefore, for any closed ideal J in Cb(X) where X ∈ T3 1 there exists a closed subset F in βX such that J = {f X : f ∈ JF (βX)}. Although this description is sufficient to have some information on closed ideals 2 TOPOLOGICAL ASPECTS OF ORDER IN C(X) 15 of Cb(X), it is not the best one since F is a subset of βX and not of X. The following two results are needed in the proof of 5.3. Although the proof of the first one is quite trivial we include it for the sake of completeness. Lemma 5.1. Let F be a closed set in X ∈ T3 1 JF (βX) where F denotes the closure of F in βX. 2 . Then β(JF (X)) = Proof. If f ∈ JF (X), then f ≡ 0 on F , and hence by continuity f β ≡ 0 on F . For the opposite inclusion, pick f β ∈ JF (βX). Then f ≡ 0 on F , so that f β ∈ β(JF (X)). (cid:3) Proposition 5.2. Let X be a dense subset of a topological space Y . If a closed set F ⊆ Y satisfies F = ClY IntY F , then the closed set F ∩ X in X satisfies F ∩ X = ClX IntX (F ∩ X). Proof. Suppose F = ClY IntY F . Since F ∩ X is closed in X, we obvi- ously have ClX IntX(F ∩ X) ⊆ F ∩ X. To prove the opposite inclusion pick any x ∈ F ∩X. We will prove that an arbitrary open neighborhood in X of x intersects IntX (F ∩ X). Let U be an arbitrary open neighborhood in X of x. There exists an open neighborhood V in Y of x such that U = V ∩ X. From F = ClY IntY F it follows V ∩ IntY F 6= ∅, so that from the density of X in Y we conclude V ∩ IntY F ∩ X 6= ∅. Since V ∩ IntY F ∩ X is open in X and is obviously contained in F ∩ X, we have ∅ 6= U ∩ IntY F = V ∩ IntY F ∩ X ⊆ IntX (F ∩ X). In particular, U intersects IntX (F ∩ X) and the proof is finished. (cid:3) 5.2 can be applied in two interesting situations when X is locally and Y = βX. We compact Hausdorff space and Y = X + or X ∈ T3 1 will use 5.2 for the latter situation in the proof of 5.3. 2 Theorem 5.3. Let X ∈ T3 1 2 and let B be a closed ideal in Cb(X) (a) B is a band in Cb(X) iff B = JF (X) for some closed set F in X with F = Int F . (b) B is a projection band iff B = JF (X) for some clopen subset F of X. 16 M. KANDI ´C AND A. VAVPETI C Proof. (a) Suppose B is a band in Cb(X). By 4.2 there exists a closed set F ⊆ βX such that F = ClβX IntβX F and β(B) = JF (βX). We claim that B = JF ∩X(X) and that F ∩ X = ClX IntX(F ∩ X). Let us denote B′ = JF ∩X(X). Then β(B′) = JClβX (F ∩X)(βX). If we prove ClβX(F ∩ X) = F , then β(B′) = β(B), so that B′ = B, and hence B = JF ∩X(X) as claimed. Since F is closed in βX, we have ClβX (F ∩X) ⊆ F . If F 6= ClβX (F ∩ X), pick x ∈ F \ ClβX (F ∩ X). There exists an open neighborhood U of x such that U ∩ (F ∩ X) = ∅. Since x ∈ F = ClβX IntβX F , the set U ∩ IntβX F is nonempty and open in βX. Density of X in βX implies U ∩IntβX F ∩X 6= ∅ which is in contradiction with U ∩F ∩X = ∅. This shows F = ClβX(F ∩ X). To finish the proof of the forward implication we apply 5.2 for Y = βX to obtain F ∩ X = ClX IntX(F ∩ X). Suppose now that B = JF (X) and F is the closure of its interior If 0 ≤ fα ր f in Cb(X) and fα ∈ JF (X) for each α, then in X. fαInt F ր f Int F , by 3.2. Since fα ≡ 0 on Int F , we have f ≡ 0 on Int F , so that by continuity of f we conclude f ≡ 0 on Int F = F. Hence, f ∈ JF (X) and B = JF (X) is a band in Cb(X). (b) If B is a projection band in Cb(X), by 2.5 there exists a clopen subset F ⊆ βX such that β(B) = JF (βX). By the proof of (a) we know that B = JF ∩X(X). Since F is clopen in βX, the definition of the relative topology on X yields that F ∩ X is clopen in X. If B = JF (X) for some clopen subset F of X, one can repeat the arguments of the proof of 4.2(b) for the appropriately chosen nonneg- ative function f ∈ Cb(X) to see that B is a projection band. (cid:3) The most important ingredient of the proof of 5.3(a) is that the set F has an extremely large interior. If we start with a closed ideal J that is not a band in Cb(X), then again we conclude β(J) = JF (βX) for some closed subset F of βX. However, the set F can have a very small interior in βX. In the worst case F ∩ X can be empty and in this case we obviously cannot have J = JF ∩X(X) unless J = Cb(X). The critical step for investigating bands and projection bands in Cb(X) where X ∈ T3 1 was the construction of βX and the isomorphism between Cb(X) and C(βX). As we already mentioned, for a Hausdorff 2 TOPOLOGICAL ASPECTS OF ORDER IN C(X) 17 2 space X there is a way how to construct a space X/∼ ∈ T3 1 such that Cb(X) ∼= Cb(X/∼) and that the isomorphism in question is induced by a continuous surjective function between τ : X → X/∼. Since in this section X/∼ and τ will be needed in details, we recall the construction. On a Hausdorff space X we define an equivalence relation ∼ as fol- lows: x ∼ y iff f (x) = f (y) for all continuous function f : X → R. By τ : X → X/∼ we denote the mapping which maps x into its equiva- lence class [x] ∈ X/∼. Then for each f ∈ C(X) there exists the unique function ef : X/∼ → R such that the following diagram commutes. f X R τ ef X/∼ Indeed, we define ef as ef ([x]) := f (x). If [x] = [x′], then x ∼ x′, so that f (x) = f (x′) as f is continuous. This proves that ef is well-defined and that f = ef ◦ τ. Let C ′ be the family {ef : f ∈ C(X)}. By endowing X/∼ by the weak topology induced by C ′, [GJ76, Theorem 3.7] implies X/∼ ∈ T3 1 is functionally Hausdorff, then τ is just the identity mapping and the topology on X/∼ is strictly weaker than the topology of X. and that τ : X → X/∼ is continuous. If X /∈ T3 1 2 2 The mapping τ : X → X/∼ induces the isomorphism Φτ : C(X/∼) → C(X) defined by f 7→ f ◦ τ. By [GJ76, Theorem 1.6] Φτ is also a lattice isomorphism. By applying [GJ76, Theorem 1.9] the restriction of Φτ to Cb(X/∼) induces the isomorphism (which we denote again by Φτ ) Φτ : Cb(X/∼) → Cb(X). The following lemma explains how ideals of the form JF (X) are transformed by Φτ and its inverse. Lemma 5.4. Let X be a Hausdorff space and X/∼ as before. Then (a) For F ⊆ X we have Φ−1 (b) For F ⊆ X/∼ we have Φτ (JF (X/∼)) = Jτ −1(F )(X). τ (JF (X)) = Jτ (F )(X/∼). Proof. (a) Take a function f ∈ JF (X). Then ef ◦ τ = f implies that ef ≡ 0 on τ (F ). By continuity we conclude ef ≡ 0 on τ (F ), so that 18 M. KANDI ´C AND A. VAVPETI C Φ−1 τ (f ) = ef ∈ Jτ (F )(X/∼). Conversely, if ef ∈ Jτ (F )(X/∼), then f = ef ◦ τ ≡ 0 on F . (b) If f ∈ Φτ (JF (X/∼)), then f = ef ◦ τ for some ef ∈ JF (X/∼). Pick x ∈ τ −1(F ). Then τ (x) ∈ F and f (x) = ef (τ (x)) = 0; hence f ∈ Jτ −1(F )(X). Conversely, choose f ∈ Jτ −1(F )(X). Then f = ef ◦ τ = Φτ (ef ). Since f τ −1(F ) ≡ 0 and τ (τ −1(F )) = F we have ef ≡ 0 on F , and therefore f ∈ Φτ (JF (X/∼)). (cid:3) Theorem 5.5. Let X be a Hausdorff space. A closed ideal J ⊆ Cb(X) is a projection band iff J = JF (X) for some clopen set F ⊆ X. Proof. Obviously, J is a projection band in Cb(X) iff Φ−1 τ (J) is a pro- jection band in Cb(X/∼). Therefore, by 5.3 J is a projection band in Cb(X) iff there exists a clopen set F ⊆ X/∼ such that Φ−1 τ (J) = JF (X/∼). Therefore J is a projection band iff it is of the form J = Φτ (JF (X/∼)) = Jτ −1(F )(X) for some clopen set F in X/∼. Since τ is continuous, τ −1(F ) is clopen and the proof is finished. (cid:3) The natural question that arises here is whether bands in Cb(X) where X is Hausdorff are of the form JF (X) where F = Int F . In general, the answer is no. Example 5.6. Let τe be the Euclidean topology on R. Let X = R be equipped with topology τ generated by B = τe ∪ {U \ N : U ∈ τe}, 2n−1]. The topology τ is stronger than τe and since R is a functionally Hausdorff space, X is functionally Hausdorff as well. Therefore, we have X/∼ = R as sets. where N = S∞ n=1[ 1 2n , 1 • The topology on X/∼ is stronger than the Euclidean topology. Indeed, let F ⊆ R be closed (in the Euclidean topology). Then for every x 6∈ F there exists a continuous function ϕ : R → R such that ϕ(x) = 1 and ϕ(F ) = 0. Since ϕ : X → R is continuous and x ∈ {y ∈ ϕ(x) − ϕ(y) < 1} ⊆ X \ F , X \ F is open, and hence F is a X : closed subset of X/∼. • The space X is not T3. TOPOLOGICAL ASPECTS OF ORDER IN C(X) 19 Indeed, by definition the set N is closed in X and 0 6∈ N . Let U, V ⊆ X be open sets such that 0 ∈ U and F ⊆ V . For every n ∈ N, 1 2n ∈ V 2n − rn, 1 hence ( 1 2n + rn) ⊆ V for some rn > 0. So there exists xn ∈ 2n+1, 1 2n ) ∩ V . The sequence (xn) lies in R \ N and inf xn = 0, hence ( ∅ 6= U ∩ {xn : n ∈ N} ⊂ U ∩ V . Therefore we can not separate the point 0 from the closed set N. We actually proved 0 ∈ ClX/∼ N and 1 similarly, one can prove 0 ∈ ClX/∼(S∞ Because R \ {0} and X \ {0} are homeomorphic they are also home- omorphic to X/∼ \ {0}. Hence ClR A, ClX A and ClX/∼ A may differ only for the point 0. The same is true for interiors of the given set. n=1( 1 2n , 1 2n−1)). • ClX/∼ IntX/∼ F = F where F = N ∪ {0}. Since F is closed in R and since topologies on X and X/∼ are stronger than the Euclidean topology, F is also closed in both X and X/∼. Therefore we have ClX/∼ IntX/∼ F ⊆ F. Due to the above remark and 2n−1), we have IntX F = IntX/∼ F = n=1( 1 2n , 1 the fact that IntR F =S∞ S∞ n=1( 1 2n , 1 F \ {0} = N = 2n−1 ) from where it follows that [ 1 2n , ∞[n=1 1 2n−1] ⊆ ClX/∼ IntX/∼ F. Because R \ N is open in X and (R \ N) ∩ IntX F = ∅, we have 0 /∈ ClX IntX F , and hence ClX IntX F = F \ {0} ⊆ ClX/∼ IntX/∼ F ⊆ ClR IntR F = F. Since 0 ∈ ClX/∼ IntX∼ F , we finally conclude ClX/∼ IntX/∼ F = F . • JF (X) is a band in Cb(X/∼); yet ClX IntX F = F \ {0}. The set F is closed in X and X/∼. The mapping τ is the identity mapping as X is functionally Hausdorff. Since the closed ideal JF (X) satisfies Φ−1 τ (JF (X)) = Jτ (F )(X/∼) = JF (X/∼) and since ClX/∼ IntX/∼ F = F , we conclude that JF (X/∼) is a band in Cb(X/∼). Therefore, JF (X) is band in Cb(X), yet ClX IntX F = F \ {0}. 20 M. KANDI ´C AND A. VAVPETI C The topological space constructed in 5.6 shows us that there are examples of spaces which satisfy weaker separation axioms than T3 1 and at the same time conclusions of 3.2 still hold. 2 2 1 n=1[ 1 2n , τe}, where N = S∞ Example 5.7. Let τe be the Euclidean topology on R. Let X = R be equipped with topology τ generated by B = τe ∪ {U \ N : U ∈ 2n−1 ]. By 5.6 we already know that X is functionally Hausdorff while it is not T3 1 . Take an arbitrary open set A in X and assume 0 ≤ fα ր f in C(X). We claim that fαA ր f A in C(A). Suppose there is g : A → R such that fαA ≤ g and g < f A. Suppose first that there is 0 6= x ∈ A with g(x) < f (x). Then there exists r > 0 such that (x − r, x + r) ⊆ A and g(y) < f (y) for each y ∈ (x − r, x + r). Take any function ϕ : R → [0, 1] which is continuous with respect to the Euclidean topology τe and satisfies ϕ(x) = 1 and ϕ ≡ 0 on R \ (x − r, x + r). Since τ is stronger than τe, the function ϕ : X → [0, 1] is also continuous. Then g ≤ f A − ϕA(f A − g) and fα ≤ f − ϕ(f − g) < f for each α where ϕ(f − g) is defined as ϕ(f − g)(x) =( ϕ(x)(f (x) − g(x)) 0 : x ∈ A : x ∈ X \ [x − r 2 , x + r 2] . This is in contradiction with fα ր f . If 0 ∈ A and g(0) < f (0), then g(y) < f (y) on some open neighbor- hood U of 0 ∈ X, so that g(y) < f (y) on some open neighborhood of y0 ∈ A \ {0} as well. By the previous paragraph this is impossible, and hence fαA ր f A. Question 5.8. Is there an example of a functionally Hausdorff space X, an open set A in X and a net (fα) in C(X) such that fα ր f in C(X) but fαA ր f A does not hold in C(A)? 2 If X ∈ T3 1 , the mapping F 7→ JF (X) is a one-to-one mapping on the set of all closed subsets of X. If X is merely a Hausdorff space, then it is possible that different closed sets induce the same fixed ideal. However, for a set F ⊆ X there is the largest closed set F ′ such that JF (X) = JF ′(X). Theorem 5.9. Let X be a Hausdorff space and F a subset of X. Then τ −1(τ (F )) is the largest closed set F ′ with the property JF (X) = TOPOLOGICAL ASPECTS OF ORDER IN C(X) 21 JF ′(X). Furthermore, we have JF (X) = Jτ −1(τ (F ))(X) = Jτ −1(τ (F ))(X). Proof. From Jτ (F )(X/∼) = Jτ (F )(X/∼) and Lemma 5.4 we conclude Jτ −1(τ (F ))(X) = Jτ −1(τ (F ))(X) = Jτ −1(τ (F ))(X). We claim JF (X) = Jτ −1(τ (F ))(X). Note first that Jτ −1(τ (F ))(X) ⊆ JF (X) follows from the fact F ⊆ τ −1(τ (F )). For the opposite inclusion, take f ∈ JF (X) and x ∈ τ −1(τ (F )). Then τ (x) ∈ τ (F ), so that there is y ∈ F with [x] = [y]. By definition of ∼ we have f (x) = f (y) = 0. This gives f ≡ 0 on τ −1(τ (F )) which proves the claim. The only thing that remains to be proved is that τ −1(τ (F )) is the largest closed set G with the property JF (X) = JG(X). In order to prove this, let G be an arbitrary closed set with JG(X) = JF (X). Then Jτ (G)(X/∼) = Jτ (F )(X/∼). Pick x ∈ G and assume τ (x) /∈ τ (F ). Since X/∼ ∈ T3 1 , there is ef ∈ Cb(X/∼) such that ef (τ (x)) = 1 and ef ≡ 0 on τ (F ). By Lemma 5.4 we have f = Φτ (ef ) ∈ Jτ −1(τ (F ))(X) = JF (X) = JG(X). Since x ∈ G we have f (x) = 0 which is a contradiction. (cid:3) 2 Corollary 5.10. A Hausdorff space X is T3 1 closed subsets F and G implies F = G. 2 iff JF (X) = JG(X) for Proof. If X ∈ T3 1 graph of Section 3. 2 , then the conclusion follows from the second para- 2 . Suppose now that X /∈ T3 1 If X is functionally Hausdorff, then X/∼ = X as sets and τ is the identity mapping. Since topology of X/∼ is strictly weaker than the topology of X, τ is not a closed mapping. Hence, there exists a closed set F in X which is not closed in X/∼. Since τ (F ) is not closed in X/∼, F is a proper subset of τ −1(τ (F )). By the assumption we have JF (X) 6= Jτ −1(τ (F ))(X) which contradicts 5.9. If X is not functionally Hausdorff, there are two different points x and x′ which cannot be separated by continuous functions. Then J{x}(X) = J{x′}(X). On the other hand, the assumption on the unique- ness implies x = x′. Again a contradiction. (cid:3) We already mentioned that order complete vector lattices posses the projection property. It is well-known that the converse statement does 22 M. KANDI ´C AND A. VAVPETI C Indeed, the set E of all real bounded functions on [0, 1] not hold. assuming only finitely many different values becomes a vector lattice when ordered pointwise. This vector lattice is not order complete, yet it has the projection property (see e.g. [Zaa96, Exercise 12.6.(ii)]). Order complete vector lattices are also uniformly complete (see e.g. [AB03]). Although neither projection property nor uniform completeness imply order completeness of a given vector lattice, together they do [AB03, Theorem 1.59]. Since by [MN91, Proposition 1.1.8] Banach lattices are uniformly complete, a given Banach lattice has the projection property iff it is order complete. If we assume that in a given Banach lattice every closed ideal is a band, then the norm on is order continuous. Actually, there is an equivalence between the last two statements (see e.g. [Zaa96, Theorem 17.17]). In the following result we determine when Cb(X) is order continuous. It turns out that whenever X is Hausdorff then Cb(X) is order continuous iff Cb(X) is finite-dimensional. Corollary 5.11. Let X be a Hausdorff space with the property that every closed ideal in Cb(X) is a band. Then the following assertions hold. (a) X is functionally Hausdorff iff X is finite. (b) Cb(X) is finite-dimensional. Proof. (a) By the assumption and 5.3 every maximal ideal in Cb(X/∼) is of the form JF (X/∼) for some closed subset F of X/∼. Hence, X/∼ is compact by 2.3. Also, for each x ∈ X/∼ the closed ideal J{x}(X/∼) is a band, so that {x} is open in X/∼, again by 5.3. Thus, X/∼ is discrete and since it is also compact, it is finite. Since X = X/∼ as sets, X is finite. The converse statement obviously holds. (b) By (a) we conclude X/∼ is finite and since Cb(X) ∼= Cb(X/∼) it (cid:3) follows that Cb(X) is finite-dimensional. 6. Extremally disconnected spaces and order completeness Although the function spaces Lp(µ) are order complete whenever 0 < p < ∞, the Banach lattice of continuous functions on a compact Hausdorff space is very rarely order complete. This happens only in TOPOLOGICAL ASPECTS OF ORDER IN C(X) 23 the case when the underlying topological space is extremally dis- connected , i.e., the closure of every open set is again open. In regular extremally disconnected spaces every point has a basis consisting of clopen sets. In this section, as an application of 4.2 and 5.3 among locally compact Hausdorff spaces and T3 1 -spaces, respectively, we character- ize order complete C0(X) and Cb(X) spaces, respectively. The char- acterization is the same as in the case of compact Hausdorff spaces. Before we proceed to results, we recall the following characterization from [MN91, Proposition 2.1.4]. 2 Theorem 6.1. For a compact Hausdorff space X the following state- ments are equivalent. (a) C(X) is order complete. (b) C(X) has the projection property. (c) X is extremally disconnected. The following theorem is the analog of 6.1 for C0(X). Among the already mentioned 4.2 and 5.3 we also require the fact that order complete vector lattices are precisely those that are uniformly complete and have the projection property. Theorem 6.2. For a locally compact Hausdorff space X the following statements are equivalent. (a) C0(X) is order complete. (b) C0(X) has the projection property. (c) X is extremally disconnected. Proof. (a) ⇒ (b) is obvious and (b) ⇒ (a) follows from the fact that C0(X) is a Banach lattice. (b) ⇒ (c) Let U be an arbitrary open set in X and let us denote by F the closure of U. Then J 0 F (X) is a band in C0(X) by 4.2. By the assumption J 0 F (X) is a projection band, so that by 4.2 there exists a clopen subset G of X such that J 0 G(X). As an application of Urysohn's lemma for locally compact Hausdorff spaces, one can prove F = G. Hence, F is open and X is extremally disconnected. F (X) = J 0 24 M. KANDI ´C AND A. VAVPETI C (c) ⇒ (b) Suppose now that X is extremally disconnected and take an arbitrary band J in C0(X). Then there exists a closed set F which is the closure of its interior and J = J 0 F (X). Since X is extremally disconnected, F is clopen, so that J 0 F (X) is a projection band by 4.2. Therefore, C0(X) has the projection property. (cid:3) Subspaces of extremally disconnected spaces are not necessary ex- tremally disconnected as the following example shows. Example 6.3. Let (X, τ ) be any topological space that is not ex- tremally disconnected and ∞ an object that is not in X. Define Y = X ∪ {∞} and τ1 = {U ∪ {∞} : U ∈ τ } ∪ {∅}. It is easy to see that τ1 is a topology on Y , and that closed sets in Y are precisely the closed sets of X and whole Y . It should be noted that the space (Y, τ1) ∈ T0 while (Y, τ1) /∈ T1. It is also clear that the (X, τ ) ֒→ (Y, τ1) is continuous and that X is not dense in Y . We claim that Y is extremally disconnected. Indeed, pick an open set U in Y . If U = ∅, then its closure is still empty which is open in Y . Otherwise U = V ∪ {∞} for some open set V in X. Since ∞ ∈ U, the closure of U is Y which is open in Y . This proves the claim. Although 6.3 shows that a subspace of an extremally disconnected space is not necessary extremally disconnected, there are positive re- sults [AP74]. Due to the importance of the following proposition we include its short proof for the sake of completeness. Proposition 6.4. Let X be a subspace of an extremally disconnected space Y . Then X is extremally disconnected in either of the following cases. (a) X is open in Y . (b) X is dense in Y . Proof. (a) Take an arbitrary open subset U ⊆ X. Then U is also open in Y , so that ClY (U) is open in Y. Then by the definition of the relative topology we have ClX (U) = ClY (U) ∩ X; hence the closure of U in X is open. TOPOLOGICAL ASPECTS OF ORDER IN C(X) 25 (b) Take any open set U in X. Then there exists an open set V in Y such that U = V ∩ X. The set ClY V is open in Y . If ClX(V ∩ X) = ClY V ∩X, then ClX U = ClX(V ∩X) is open in X and X is extremally disconnected. To prove ClX (V ∩ X) = ClY V ∩ X observe that ClX (V ∩ X) ⊆ ClY V ∩ X follows from the fact that the set ClY V ∩ X is closed in X and contains V ∩X. For the opposite inclusion, pick x ∈ ClY V ∩X and any open neighborhood W of x in X. Then there exists an open set W ′ in Y such that W = W ′ ∩ X. Since x ∈ ClY V , we have W ′ ∩ V 6= ∅, and since X dense in Y the set W ∩ (V ∩ X) = W ′ ∩ V ∩ X is nonempty as well. This proves x ∈ ClX(V ∩ X). (cid:3) We apply 6.4 in the following two cases: • If X + is extremally disconnected, the dense subspace X of X + is extremally disconnected. • If X ∈ T3 1 2 and βX is extremally disconnected, the dense sub- space X of βX is extremally disconnected. The following result now follows immediately from 6.1 and 6.2. Corollary 6.5. Let X be a locally compact Hausdorff space. If C(X +) is order complete, then C0(X) is order complete. The preceding corollary can be proved directly without involving ex- tremally disconnected spaces and their connection to order complete- ness. This follows immediately from C0(X) ∼= J{∞}(X +) and from the fact that every order ideal in an order complete vector lattice is order complete on its own (see e.g. [AB03, Theorem 1.62]). The following diagram shows the connection between order complete- ness and extreme disconnectedness for the locally compact Hausdorff space and its one-point compactification. C(X +) is order complete ⇔ X + is extremally disconnected ⇓ ⇓ C0(X) is order complete ⇔ X is extremally disconnected None of the arrows above cannot be reversed as the following example shows. 26 M. KANDI ´C AND A. VAVPETI C Example 6.6. Take X = N. If we endow X with a discrete topol- ogy, it becomes a locally compact noncompact Hausdorff space. The space C0(X) is precisely the order complete Banach lattice c0, however C(X +) is isometrically lattice isomorphic to the Banach lattice c of all convergent sequences which is not order complete. We continue this section with the variant of 6.2 for the Banach lattice Cb(X) where X ∈ T3 1 . Since the proof is basically the same as the proof of 6.2 we omit it. We should mention that instead of 4.2 one should use 5.3. 2 Theorem 6.7. For X ∈ T3 1 2 the following statements are equivalent. (a) Cb(X) is order complete. (b) Cb(X) has the projection property. (c) X is extremally disconnected. The following diagram reveals the connection between order com- pleteness and extreme disconnectedness for X ∈ T3 1 and its Stone- Cech compactification βX. Since the algebras C(βX) and Cb(X) are isometrically lattice and algebra isomorphic all the arrows here are also reversible in contrast with the arrows in the diagram involving C0(X) and C(X +). 2 C(βX) is order complete ⇔ βX is extremally disconnected m m Cb(X) is order complete ⇔ X is extremally disconnected If X ∈ T3 1 then X is extremally disconnected iff βX is. The fol- lowing example shows that such equivalence does not hold for a locally compact space and its one-point compactification. 2 Example 6.8. The one-point compactification N+ of N is homeomor- phic to the totally disconnected compact space X = { 1 n : n ∈ N} ∪ {0} which is not extremally disconnected. To see this, let us denote sets 2n : n ∈ N} and { 1 { 1 2n−1 : n ∈ N} by A and B, respectively. Then both A and B are open and neither of them is closed. The closure A of A is A ∪ {0} = X \ B which is not open. TOPOLOGICAL ASPECTS OF ORDER IN C(X) 27 In 6.10 we prove that X + is never extremally disconnected when X is a σ-compact locally compact noncompact Hausdorff space. Before we state and prove the aforementioned result we need the following lemma. Lemma 6.9. A locally compact Hausdorff space X is σ-compact iff the point ∞ in X + has a countable basis of open sets. S∞ Proof. Suppose first that X is σ-compact. Then by [Dug66, Theorem XI.7.2] there exist relatively compact open sets Un in X with X = n=1 Un and Un ⊆ Un+1 for each n ∈ N. If V is an open neighborhood of ∞, then X \ V is compact in X. Since the countable family of open sets (Un) is increasing and covers X there exists n ∈ N such that X \ V ⊆ Un. Then Vn := X + \ Un ⊆ V , and hence the family (Vn) is a countable basis for the point ∞ in X +. For the converse let (Vn) be a countable basis for ∞ in X +. Obviously we can assume that Vn ⊆ Vm whenever m ≤ n. The sets Fn := X \ Vn are compact in X and Fm ⊆ Fn whenever m ≤ n. Since X + is Hausdorff,T∞ n=1 Vn = {∞}, so that This proves that X is σ-compact. Fn = (X \ Vn) = X \ ∞[n=1 ∞[n=1 Vn = X. ∞\n=1 (cid:3) Theorem 6.10. Let X be a σ-compact locally compact Hausdorff space. If X + is extremally disconnected, then X is compact. Proof. Suppose X is not compact. Since X is σ-compact, there exist relatively compact open sets (Un) in X that cover X and Un ⊆ Un+1. Then the family (X +\Un) is a countable basis of ∞. For each n ∈ N we choose xn ∈ Un \ Un−1. Since each basis neighborhood X + \ Un contains all except finitely many terms of the sequence (xn) we have xn → ∞ in X +. Since X + is Hausdorff no subnet of (xn)n≥k converges in X for each k ∈ N. This implies that for each k ∈ N the set Fk := {xn : n ≥ k} is closed in X + and so in X. We claim that for each n ∈ N there exists a clopen neighborhood Wn of xn in X such that Wn ∩ Wm = ∅ whenever n 6= m and Wn ∩ Fm = ∅ 28 M. KANDI ´C AND A. VAVPETI C whenever n < m. We construct the desired sets inductively. Since the set X \ F2 is open and x1 ∈ X \ F2, there is a clopen neighborhood W1 of x1 that is contained in X \ F2. Obviously W1 ∩ Fm = ∅ for m > 1. Suppose now that the sets W1, . . . , Wn with the required properties are already constructed. Since the set (X \ W1 ∪ · · · ∪ Wn) \ Fn+2 is an open neighborhood of xn+1 in X, it contains a clopen neighborhood Wn+1 of xn+1 in X. By construction Wn+1 does not contain xk for k 6= n + 1. n=1 W2n is open in X. Since X is locally compact, X is open in X +, so that W is open in X + as well. Because x2n → ∞ we have ∞ ∈ W . From x2n−1 → ∞ we conclude that every neighborhood U of ∞ contains infinitely many elements x2n−1. On the other hand W2n−1 ∩ W = ∅ implies x2n−1 6∈ W for each n ∈ N. Therefore, for an arbitrary neighborhood U of ∞ we have U 6⊆ W , so that W is not open in X +. This is in contradiction that X + is extremally disconnected. Hence, X is compact. (cid:3) The set W =S∞ Corollary 6.11. Let X be a locally compact σ-compact Hausdorff space. If C(X +) is order complete, then X is compact and C(X) is a projection band of codimension one in C(X +). Proof. If C(X +) is order complete then X + is extremally disconnected by 2.5, so that by 6.10 X is compact. Therefore C(X) = C0(X) and ∞ is isolated in X +. This means that {∞} is clopen in X + and 2.5 implies that C(X) = J{∞}(X +) is a projection band in C(X +). That the codimension of C(X) in C(X +) is one is obvious. (cid:3) There is no hope that 6.10 can be applied to X and its Stone- Cech compactification βX. The reason behind it is that the space βX is just too big and the sequential nature of the construction in the proof of 6.10 simply fails. Example 6.12. The set of all positive integers N endowed with a discrete topology is a σ-compact locally compact discrete space. As a discrete space it is definitely extremally disconnected. Since N ∈ T3 1 , βN is extremally disconnected; yet N is not compact. 2 TOPOLOGICAL ASPECTS OF ORDER IN C(X) 29 7. Lifting un-convergence In this section we apply our results to the so-called problem of "lifting un-convergence". We first recall some basic facts needed throughout this section. If E is a Banach lattice, then a net (xα) is said to un-converge to a vector x ∈ E whenever for each y ∈ E+ we have xα − x ∧ y → 0 un−→ x whenever the net (xα) un-converges to in norm. We write xα x. This mode of convergence was introduced by V.G. Troitsky [Tro04] under the name of d-convergence. In [Tro04, Example 20] it was proved that un-convergence in C0(X) coincides with the uniform con- vergence on compacta of X whenever X is normal. In particular, when X is a compact Hausdorff space, then un-convergence coincides with the uniform convergence. In the case of Lp(µ)-spaces with µ fi- nite he proved that un-convergence coincides with convergence in mea- sure [Tro04, Example 23]. It is a standard fact from measure theory that a sequence (fn) converging in measure to f always has a subse- quence converging to f almost everywhere (see e.g. [Fol99, Theorem 2.30]). In [GTX, Proposition 3.1] the authors proved that a sequence (fn) in L0(µ) converges almost everywhere to f ∈ L0(µ) whenever fn − f ∧ g → 0 in order of L0(µ). The latter mode of convergence goes back to [Nak48, DeM64, Kap97]. Kaplan referred to this convergence as unbounded order convergence or uo-convergence for short. Un- til very recently unbounded order convergence was not studied actively and was left out from the active area of research. The systematic study of this mode of convergence and its properties started with papers of Gao, Troitsky and Xanthos [Gao14, GX14, GTX] and others. The systematic study and properties of un-convergence started in [DOT]. Among other things, authors proved that un-convergence is topologi- cal. On the other hand, uo-convergence is not (see e.g. [Ord66]). In [KMT] the authors initiated the study of un-topology, i.e., the topology given by un-convergence. They posed the following question. Question 7.1. Let B be a band in a Banach lattice E. Suppose that every net in B which is un-null in B is also un-null in E. Does this imply that B is a projection band? 30 M. KANDI ´C AND A. VAVPETI C When the norm of E is order continuous, every band is a projection band. In this case it was proved in [KMT] that a net (xα) which is un-null in a sublattice of E it is also un-null in E. [KMT, Example 4.2] shows that there exists a band B in C[−1, 1] and a un-null net (xα) in B which is not un-null in C[−1, 1]. The following theorem provides a positive answer to 7.1 for a Banach lattice Cb(X) of all bounded continuous functions on a Hausdorff space. Theorem 7.2. Let X be a Hausdorff space and J a closed ideal in Cb(X). Suppose that every un-null net (fα) in J is also un-null in Cb(X). Then J is a projection band in Cb(X). Proof. We first consider the case when X is compact. Since J is a closed ideal in C(X), by 2.4 there exists a closed set F such that J = JF (X). If F is not open, there exists x ∈ F \ Int F. Let Bx = {Wλ}λ∈Λ be the set of all open neighborhoods of the point x. Since X is Hausdorff, [Dug66, VII.I.2] implies {x} = \λ∈Λ Wλ. If Λ is finite, {x} is open, so that x ∈ Int F which is a contradiction. Therefore Λ is infinite. Since x ∈ F \ Int F , for each λ ∈ Λ the set Wλ \ F is nonempty. Pick any xλ ∈ Wλ \ F. Also, let Vλ be an arbitrary open neighborhood of xλ in (X \ F ) ∩ Wλ. By Urysohn's lemma for each λ ∈ Λ there exists a nonnegative continuous function fλ such that fλ(xλ) = 1 and f X\Vλ ≡ 0. We define an ordering on the set Λ by λ ≤ µ iff Wµ ≤ Wλ. With this ordering Λ becomes a directed set. We claim that fλ un−→ 0 in J. Pick an arbitrary function g ∈ J. Then g(x) = 0. Continuity of g implies that for every ǫ > 0 there exists an open neighborhood U of x such that g(y) < ǫ for each y ∈ U. There exists λ0 ∈ Λ such that Wλ0 ⊆ U. If λ ≥ λ0, then Vλ ⊆ Wλ ⊆ Wλ0 ⊆ U. If y ∈ X \Vλ, then fλ(y) = 0 and if y ∈ Vλ, then g(y) < ǫ. We conclude fλ ∧ g < ǫ for all λ ≥ λ0. Therefore, fλ un−→ 0 in J. Now we claim that the net (xλ) converges to x. Indeed, let W be an arbitrary neighborhood of x. Then there exists λ0 such that Wλ0 ⊆ W . If λ ≥ λ0, then xλ ∈ Wλ ⊆ Wλ0 ⊆ W which proves the claim. TOPOLOGICAL ASPECTS OF ORDER IN C(X) 31 The assumption fλ un−→ 0 in J implies that fλ → 0 in C(X) which contradicts the fact that fλ(xλ) = 1 for each λ ∈ Λ. Therefore F = Int F is clopen; hence J is a projection band by 2.5. The general case follows from the first part of the proof and that whenever X is "just" Hausdorff, then Cb(X) and C(βX/∼) are isomet- rically algebra and lattice isomorphic. (cid:3) The following corollary follows immediately from 7.2 and 5.5. Corollary 7.3. Let X be a Hausdorff space and x ∈ X such that un−→ 0 in Cb(X). Then x is isolated in fα X. un−→ 0 in J{x}(X) implies fα We proceed with an application of 5.11 and 7.2. Corollary 7.4. Let X be a topological space with the property that for un−→ 0 in every maximal ideal J in Cb(X) fα Cb(X). un−→ 0 in J implies fα (a) If X is functionally Hausdorff, then X is finite. (b) If X is Hausdorff, then Cb(X) is finite-dimensional. Proof. Since every point in X is isolated by 7.3, X is discrete. Discrete compact spaces are obviously finite. (cid:3) We conclude this paper with C0(X)-analogs of results of this section. Since the proofs are basically the same as in the Cb(X)-case, we omit them. Theorem 7.5. Let X be a locally compact Hausdorff space. (a) Let J be a closed ideal in C0(X). If every un-null net (fα) in J is also un-null in C0(X), then J is a projection band in C0(X). un−→ 0 in C0(X), (b) Pick x ∈ X. If fα un−→ 0 in J 0 {x}(X) implies fα then x is isolated in X. (c) If for every maximal ideal J in C0(X) fα fα un−→ 0 in C0(X), then X is discrete. un−→ 0 in J implies References [AB03] C.D. Aliprantis and O. Burkinshaw, Locally solid Riesz spaces with applications to economics, 2nd ed., AMS, Providence, RI, 2003. 32 [AB06] [ABo06] [AP74] [Con99] [DeM64] [DOT] [Dug66] [Fol99] [Gao14] [GJ76] [GX14] [GTX] [KMT] [Kan09] [Kap97] [LZ71] [MN91] M. KANDI ´C AND A. VAVPETI C C.D. Aliprantis and O. Burkinshaw, Positive operators, 2nd edition, Springer 2006. C.D. Aliprantis and K.C. Border, Infinite dimensional analysis. A hitchhiker's guide, 3th ed., Springer, Berlin, 2006. xxii+703 pp. A. V. Arkhangel'skii, V. I. Ponomarev, Fundamentals of general topol- ogy in problems and exercises, Izdat."Nauka", Moskva, 1974,(in Rus- sian)(English translation: ser. Mathematics and its Applications, D. Reidel Publishing Co., Dordrecht-Boston, Mass., 1984. ) J.B. Conway, A course in functional analysis, 2nd edition, Springer- Verlag, New York, 1990. R. DeMarr, Partially ordered linear spaces and locally convex linear topological spaces, Illinois J. Math. 8, 1964, 601–606. Y. Deng, M. O'Brien, and V.G. Troitsky, Unbounded norm conver- gence in Banach lattices, Positivity, to appear, DOI: 10.1007/s11117- 016-0446-9 J. Dugundji, Topology, Allyn and Bacon, Inc., Boston, Mass. 1966. G.B. Folland, Real analysis: Modern techniques and their applications, 2nd edition, Pure and Applied Mathematics, John Wiley & Sons, Inc., New York, 1999. N. Gao, Unbounded order convergence in dual spaces, J. Math. Anal. Appl., 419, 2014, 347–354. L. Gillman, M. Jerison, Rings of Continuous Functions, Graduate Texts in Math. 43, Springer-Verlag, Berlin-Heidelberg-New York, 1976. N. Gao and F. Xanthos, Unbounded order convergence and application to martingales without probability, J. Math. Anal. Appl., 415 (2014), 931–947. N. Gao, V.G. Troitsky, and F. Xanthos, Uo-convergence and its appli- cations to Ces`aro means in Banach lattices, Israel J. Math., to appear. arXiv:1509.07914 [math.FA]. M. Kandi´c, M.A.A. Marabeh, V.G.Troitsky, Unbounded norm topology in Banach lattices, preprint https://arxiv.org/pdf/1608.05489v1.pdf [math.FA]. E. Kaniuth, A course in commutative Banach algebras, Graduate Texts in Math. 246, Springer, New York, 2009. S. Kaplan, On Unbounded Order Convergence, Real Anal. Exchange 23(1), 1997, 175–184. W.A.J. Luxemburg, A.C. Zaanen, Riesz spaces I, North-Holland Math- ematical Library, 1971. P. Meyer-Nieberg, Banach lattices, Springer-Verlag, Berlin, 1991. TOPOLOGICAL ASPECTS OF ORDER IN C(X) 33 [Nak48] [Ord66] [Tro04] [Zaa96] H. Nakano, Ergodic theorems in semi-ordered linear spaces, Ann. of Math. (2), 49, 1948, 538–556. E. T. Ordman, Convergence almost everywhere is not topological, Amer- ican Math. Monthly 73 (1966), no. 2, 1882–183. V. G. Troitsky, Measures on non-compactness of operators on Banach lattices, Positivity 8 (2004), no. 2, 165–178. A.C. Zaanen, Springer, Berlin - Heidelberg - New York (1996). Introduction to Operator Theory in Riesz Spaces, Fakulteta za Matematiko in Fiziko, Univerza v Ljubljani, Jadranska ulica 19, SI-1000 Ljubljana, Slovenija E-mail address: [email protected] Fakulteta za Matematiko in Fiziko, Univerza v Ljubljani, Jadranska ulica 19, SI-1000 Ljubljana, Slovenija E-mail address: [email protected]
1005.0171
1
1005
2010-05-02T22:19:10
Every State on Interval Effect Algebra is Integral
[ "math.FA" ]
We show that every state on an interval effect algebra is an integral through some regular Borel probability measure defined on the Borel $\sigma$-algebra of a compact Hausdorff simplex. This is true for every effect algebra satisfying (RDP) or for every MV-algebra. In addition, we show that each state on an effect subalgebra of an interval effect algebra $E$ can be extended to a state on $E.$ Our method represents also every state on the set of effect operators of a Hilbert space as an integral
math.FA
math
EVERY STATE ON INTERVAL EFFECT ALGEBRA IS INTEGRAL ANATOLIJ DVURE CENSKIJ 1 Mathematical Institute, Slovak Academy of Sciences Stef´anikova 49, SK-814 73 Bratislava, Slovakia E-mail: [email protected], Abstract. We show that every state on an interval effect algebra is an integral through some regular Borel probability measure defined on the Borel σ-algebra of a compact Hausdorff simplex. This is true for every effect algebra satisfying (RDP) or for every MV-algebra. In addition, we show that each state on an effect subalgebra of an interval effect algebra E can be extended to a state on E. Our method represents also every state on the set of effect operators of a Hilbert space as an integral. 1. Introduction Effect algebras were defined by Foulis and Bennett in [FoBe] in order to axioma- tize some measurements in quantum mechanics that are modeled by POV-measures as partial algebras with a partially defined addition, +, an analogue of join of two orthogonal elements. Effect algebras generalize Boolean algebras, orthomodular lattices and posets, orthoalgebras as well as MV-algebras. Because Hilbert space quantum mechanics "lives" in a Hilbert space, H, the system of all Hermitian op- erators, B(H), the system of all effect operators, E(H), i.e. the set of all Hermitian operators between the zero operator and the identity, as well as the system of all orthogonal projectors, P(H), are crucial algebraic structures: B(H) is a po-group that is an antilattice, [LuZa, Thm 58.4], i.e. only comparable elements have join, E(H) is an effect algebra, and P(H) is a complete orthomodular lattice. They are fundamental examples of quantum structures; more on quantum structures see [DvPu]. We note that effect algebras are equivalent to D-posets defined by Kopka and Chovanec [KoCh], where a primary notion is difference of comparable events. An important subfamily of effect algebras are MV-algebras introduced by Chang [Cha] that model many-valued Lukasiewicz logic. The role of MV-algebras for effect algebras is analogous as that of Boolean algebras for orthomodular posets: any orthomodular poset or a lattice effect algebra can be covered by a system of 1Keywords: Effect algebra; Riesz Decomposition Property; state; unital po-group; simplex; Bauer simplex AMS classification: 81P15, 03G12, 03B50 The author thanks for the support by Center of Excellence SAS - Quantum Technologies -, ERDF OP R&D Projects CE QUTE ITMS 26240120009 and meta-QUTE ITMS 26240120022, the grant VEGA No. 2/0032/09 SAV and by the Slovak Research and Development Agency under the contract No. APVV-0071-06, Bratislava. 1 2 ANATOLIJ DVURE CENSKIJ mutually compatible substructures that in the first case are Boolean subalgebras whilst in the second case are MV-subalgebras. Many useful effect algebras are intervals in some unital po-group groups. Such ex- amples are effect algebras with the Riesz Decomposition Property (RDP for short), or unigroups or MV-algebras. A state is an analogue of a probability measure for quantum structures. Roughly speaking, it is a positive functional s on an effect algebra E that preserves + and s(1) = 1 where 1 is the greatest element in E. In general, an effect algebra may have no state, but effect algebra with (RDP) or an interval one has at least one state. The system of all states is always a convex compact Hausdorff space whose the set of extremal states is not necessarily compact. Anyway, using some analysis of compact convex sets it is possible to show that if E is an interval effect algebra or even if it satisfies (RDP) or an MV-algebra, then a state is an integral through a regular Borel probability measure defined on the Borel σ-algebra over a compact Hausdorff simplex. This result generalizes the analogous results by [Kro, Kro1] and Panti [Pan] proved for states on MV-algebras. We show also conditions when, for every state s on E, there is a unique regular Borel probability measure. In addition, we generalize Horn -- Tarski Theorem that says that every state on a Boolean subalgebra of a Boolean algebra B can be extended to a state on B. We show that every state on an effect subalgebra of an interval effect algebra E can be extended to a state on E. The article is organized as follows. Section 2 presents elements of effect algebras, and states on effect algebras are recalled in Section 3. Section 4 is devoted to affine continuous functions on a convex compact Hausdorff topological space and Section 5 shows importance of Choquet and Bauer simplices for our study. Finally, the main results are formulated and proved in Section 6. In Conclusion, we summarize our results. 2. Elements of Effect Algebras An effect algebra is by [FoBe] a partial algebra E = (E; +, 0, 1) with a partially defined operation + and two constant elements 0 and 1 such that, for all a, b, c ∈ E, (i) a + b is defined in E if and only if b + a is defined, and in such a case a + b = b + a; (ii) a + b and (a + b) + c are defined if and only if b + c and a + (b + c) are defined, and in such a case (a + b) + c = a + (b + c); (iii) for any a ∈ E, there exists a unique element a′ ∈ E such that a + a′ = 1; (iv) if a + 1 is defined in E, then a = 0. We recommend [DvPu] for more details on effect algebras. If we define a ≤ b if and only if there exists an element c ∈ E such that a + c = b, then ≤ is a partial order on E, and we denote c := b − a. Then a′ = 1 − a for any a ∈ E. We will assume that E is not degenerate, i.e. 0 6= 1. Let E and F be two effect algebras. A mapping h : E → F is said to be a homomorphism if (i) h(a + b) = h(a) + h(b) whenever a + b is defined in E, and (ii) h(1) = 1. A bijective homomorphism h such that h−1 is a homomorphism is said to be an isomorphism of E and F . EVERY STATE ON INTERVAL EFFECT ALGEBRA IS INTEGRAL 3 Let E be a system of [0, 1]-valued functions on Ω 6= ∅ such that (i) 1 ∈ E, (ii) f ∈ E implies 1 − f ∈ E, and (iii) if f, g ∈ E and f (ω) ≤ 1 − g(ω) for any ω ∈ Ω, then f + g ∈ E. Then E is an effect algebra of [0, 1]-valued functions, called an effect-clan, that is not necessarily a Boolean algebra nor an MV-algebra. We say that an effect algebra E satisfies the Riesz Decomposition Property, (RDP) for short, if x ≤ y1 + y2 implies that there exist two elements x1, x2 ∈ E with x1 ≤ y1 and x2 ≤ y2 such that x = x1 + x2. We recall that E has (RDP) if and only if, [DvPu, Lem. 1.7.5], x1 + x2 = y1 + y2 implies there exist four elements c11, c12, c21, c22 ∈ E such that x1 = c11 + c12, x2 = c21 + c22, y1 = c11 + c21, and y2 = c12 + c22. A partially ordered Abelian group (G; +, 0) (po-group in short) is said to satisfy interpolation provided given x1, x2, y1, y2 in G such that x1, x2 ≤ y1, y2 there exists z in G such that x1, x2 ≤ z ≤ y1, y2. We say also that G is an interpolation group. A po-group (G, u) with strong unit u (= order unit, i.e. given g ∈ G, there is an integer n ≥ 1 such that g ≤ n) is said to be a unital po-group. If the set Γ(G, u) := {g ∈ G : 0 ≤ g ≤ u} (2.1) is endowed with the restriction of the group addition +, then (Γ(G, u); +, 0, u) is an effect algebra. Ravindran [Rav] ([DvPu, Thm 1.7.17]) showed that every effect algebra with (RDP) is of the form (2.1) for some interpolation unital po-group (G, u): Theorem 2.1. Let E be an effect algebra with (RDP). Then there exists a unique (up to isomorphism of Abelian po-groups with strong unit) interpolation group (G, u) with strong unit such that Γ(G, u) is isomorphic with E. Moreover, the category of effect algebras with (RDP) is categorically equivalent with the category of unital po-groups with interpolation; the categorical equivalence is given by the functor: Γ : (G, u) 7→ Γ(G, u). An effect algebra of the form Γ(G, u) for some element u ≥ 0 or if it is isomorphic with Γ(G, u) is called also an interval effect algebra. For example, the system, E(H), of all Hermitian operators, A, of a Hilbert space H (real, complex or quaternionic) such that O ≤ A ≤ I, where the ordering of Hermitian operators is defined by the property: A ≤ B iff (Aφ, φ) ≤ (Bφ, φ) for all φ ∈ H, is an interval effect algebra such that E(H) = Γ(B(H), I). We note that E(H) is without (RDP). Let u be a positive element of an Abelian po-group G. The element u is said to be generative if every element g ∈ G+ is a group sum of finitely many elements of Γ(G, u), and G = G+ − G+. Such an element is a strong unit [DvPu, Lem 1.4.6] for G and the couple (G, u) is said to be a po-group with generative strong unit. For example, if u is a strong unit of an interpolation po-group G, then u is generative. The same is true for I and Γ(B(H), I). Let E be an effect algebra and H be an Abelian (po-) group. A mapping p : G → H that preserves + is called an H-valued measure on E. Remark 2.2. If E is an interval effect algebra, then there is a po-group G with a generative strong unit u such that E ∼= Γ(G, u) and every H-valued measure p : Γ(G, u) → H can be extended to a group-homomorphism φ from G into H. If H is a po-group, then φ is a po-group-homomorphism. Then φ is unique and (G, u) is also unique up to isomorphism of unital (po-) groups, see [DvPu, Cor 1.4.21]; the element u is said to be a universal strong unit for Γ(G, u) and the couple (G, u) is 4 ANATOLIJ DVURE CENSKIJ said to be a unigroup. In particular, the identity operator I is a universal strong unit for Γ(B(H), I), [DvPu, Cor 1.4.25], similarly for Γ(A, I), where A is a von Neumann algebra on H. Also in this case, the category of interval effect algebras is categorically equivalent with the category of unigroups (G, u) with the functor Γ : (G, u) 7→ Γ(G, u). We notify that also a non-commutative version of effect algebras, called pseudo effect algebras, was recently introduced in [DvVe1, DvVe2]; in some important cases they are also intervals but in unital po-groups that are not necessarily Abelian. A very important family of effect algebras is the family of MV-algebras, which were introduced by Chang [Cha]: An MV-algebra is an algebra (M ; ⊕,∗ , 0) of signature h2, 1, 0i, where (M ; ⊕, 0) is a commutative monoid with neutral element 0, and for all x, y ∈ M (i) (x∗)∗ = x, (ii) x ⊕ 1 = 1, where 1 = 0∗, (iii) x ⊕ (x ⊕ y∗)∗ = y ⊕ (y ⊕ x∗)∗. We define also an additional total operation ⊙ on M via x ⊙ y := (x∗ ⊕ y∗)∗. For more on MV-algebras, see the monograph [CDM]. If we define a partial operation + on an MV-algebra M in such a way that a+b is defined in M if and only if a ≤ b∗ (or equivalently a⊙b = 0) and we set a+b := a⊕b, then (M ; +, 0, 1) is a lattice ordered effect algebra with (RDP), where a′ := a∗. According to [Mun], every MV-algebra is isomorphic with (Γ(G, u); ⊕,∗ , 0), where (G, u) is an ℓ-group (= lattice ordered group) with strong unit and a ⊕ b = (a + b) ∧ u, a∗ = u − a, a, b ∈ Γ(G, u). It is worthy recalling that a lattice ordered effect algebra E with (RDP) can be converted into an MV-algebra (in such a way that both original + and the partial addition induced from the MV-structure coincide); we define x⊕y := x+(y∧(x∧y)′) and x∗ = x′, see [DvPu, Thm 1.8.12]. 3. States on Effect Algebras We present some notions and properties about states on effect algebras. For further reading we recommend [DvPu, Dvu]. A state on an effect algebra E is any mapping s : E → [0, 1] such that (i) s(1) = 1, and (ii) s(a + b) = s(a) + s(b) whenever a + b is defined in E. We denote by S(E) the set of all states on E. It can happen that S(E) is empty, see e.g. [DvPu, Exam. 4.2.4]. In important cases, for example, when E satisfies (RDP) or, more generally, if E is an interval effect, S(E) is nonempty, see below. S(E) is if s1, s2 are states on an effect algebra E and λ ∈ [0, 1], always a convex set, i.e. then s = λs1 + (1 − λ)s2 is a also a state on E. A state s is said to be extremal if s = λs1 + (1 − λ)s2 for λ ∈ (0, 1) implies s = s1 = s2. By ∂eS(E) we denote the set of all extremal states of S(E). We say that a net of states, {sα}, on E weakly converges to a state s on E if sα(a) → s(a) for any a ∈ E. In this topology, S(E) is a compact Hausdorff topological space and every state on E belongs to the weak closure of the convex hull of the extremal states, as the Krein-Mil'man Theorem states, see e.g. [Goo, Thm 5.17]. Let (G, u) be an Abelian po-group with strong unit. By a state on (G, u) we mean any mapping s : G → R such that (i) s(g + h) = s(g) + s(h) for all g, h ∈ G, (ii) s(G+) ⊆ R+, and (iii) s(u) = 1. In other words, a state on (G, u) is any EVERY STATE ON INTERVAL EFFECT ALGEBRA IS INTEGRAL 5 po-group homomorphism from (G, u) into the po-group (R, 1) that preserves fixed strong units. We denote by S(G, u) the set of all states and by ∂eS(G, u) the set of all extremal states, respectively, on (G, u). According to [Goo, Cor. 4.4], S(G, u) 6= ∅ whenever u > 0. In a similar way as for effect algebras, we define the weak convergence of states on (G, u), and analogously, S(G, u) is a compact Hausdorff topological space and every state on (G, u) is a weak limit of a net of convex combinations from ∂eS(G, u). If E = Γ(G, u), where (G, u) is an interpolation po-group or a unigroup, then due to the categorical equivalence, every state on E can be extended to a unique state on (G, u), and conversely, the restriction of any state on (G, u) to E gives a state on E. Therefore, if E satisfies (RDP) or more generally if E is an interval effect algebra, S(E) 6= ∅. In addition, extremal states on E are the restrictions of extremal states on (G, u), the space S(E) is affinely homeomorphic with S(G, u), and the space ∂eS(E) is homeomorphic with ∂eS(G, u). We recall that a mapping from one convex set onto another convex set is affine if it preserves convex combinations. We recall that if E is an MV-algebra, then a state s on E is extremal if and only if s(x ∧ y) = min{s(x), s(y)}, x, y ∈ E, therefore ∂eS(E) is compact (see, e.g. [DvPu, Prop. 6.1.19], [Goo, Thm 12.18]). On the other hand, if E is an effect algebra with (RDP), then ∂eS(E) is not necessarily compact, see for example [Goo, Exam. 6.10]. We say that a system of states, S, on an effect algebra E is order determining if s(a) ≤ s(b) for any s ∈ S yields a ≤ b. Let H be a separable Hilbert space (real, complex or quaternionic) with an inner product (·, ·), and L(H) be the system of all closed subspaces of H. Then L(H) is a complete orthomodular lattice [Dvu]. Given a unit vector φ ∈ H, let pφ(M ) := (PM φ, φ), M ∈ L(H), where PM is the orthogonal projector of H onto M. Then pφ is a σ-additive state on L(H), called a pure state. The system of all pure states is order determining. If T is a positive Hermitian operator of finite trace (i.e. Pi(T φi, φi) < ∞ for any orthonormal basis {φi} of H, and we define tr(T ) :=Pi(T φi, φi)), then sT (M ) := tr(T PM ), M ∈ L(H), (3.3) is a σ-additive state, and according to famous Gleason's Theorem, [Gle], [Dvu, Thm 3.2.24], if dim H ≥ 3, for every σ-additive state s on L(H), there is a unique positive Hermitian operator T of finite trace such that s = sT . If now s is a finitely additive state on L(H), then due to the Aarnes Theorem, [Dvu, Thm 3.2.28], s can be uniquely decomposed to the form s = λs1 + (1 − λ)s2, where s1 is a σ-additive state and s2 is a finitely additive state vanishing on each finite-dimensional subspace of H. Moreover, a pure state pφ is an extreme point of the set of σ-additive states, as well as it is an extremal state of L(H). Other extremal states of L(H) vanish on each finite-dimensional subspace of H. Since every state on L(H) can be extended into a unique state on B(H), see e.g. [Dvu, Thm 3.3.1], the state spaces S(L(H)), S(E(H)) and S(B(H)) are mutually affinely homeomorphic whenever 3 ≤ dim H ≤ ℵ0. It is worthy to recall that L(H) is isomorphic to the set P(H) of all orthogonal projectors of H. 6 ANATOLIJ DVURE CENSKIJ We say that a po-group G is Archimedean if for x, y ∈ G such that nx ≤ y for all positive integers n ≥ 1, then x ≤ 0. It is possible to show that a unital po-group (G, u) is Archimedean iff G+ = {g ∈ G : s(g) ≥ 0 for all s ∈ S(G, u)}, [Goo, Thm 4.14]. We have the following characterization of interval effect algebras coming from Archimedean po-groups. Proposition 3.1. Let E = Γ(G, u), where is (G, u) is a unigroup. The following statements are equivalent: (i) S(E) is order determining. (ii) E is isomorphic to some effect-clan. (iii) G is Archimedean. Proof. (i) ⇒ (ii). Given a ∈ E, let a be a function from S(E) into the real interval [0, 1] such that a(s) := s(a) for any s ∈ S(E), and let E = {a : a ∈ E}. We endow E with pointwise addition, so that E is an effect-clan. Since S(E) is order determining, the mapping a 7→ a is an isomorphism and E is Archimedean. (ii) ⇒ (iii). Let E be any effect-clan isomorphic with E, and let a 7→ a be such an isomorphism. Let G( E) be the po-group generated by E. Then G( E) consists of all functions of the form a1 + · · · + an − b1 − · · · − bm, and 1 is its strong unit. Due to the categorical equivalence, (G, u) and (G( E), 1) are isomorphic. But (G( E), 1) is Archimedean, so is (G, u). (iii) ⇒ (i). Due to [Goo, Thm 4.14], G+ = {g ∈ G : s(g) ≥ 0 for all s ∈ S(G, u)}, which means that S(G, u) is order determining. Hence, the restrictions of all states on (G, u) onto E = Γ(G, u) entail S(E) is order determining. (cid:3) 4. Affine Continuous Functions We present some important notions about affine continuous functions that are important for theory of effect algebras. A good source for this topic are monographs [Goo, Alf]. Let K be a convex subset of a real vector space V. A point x ∈ K is said to be extreme if from x = λx1 + (1 − λ)x2, where x1, x2 ∈ K and 0 < λ < 1 we have x = x1 = x2. By ∂eK we denote the set of extreme points of K, and if K is a compact convex subset of a locally convex Hausdorff topological space, then due to the Krein-Mil'man Theorem [Goo, Thm 5.17], K equals to the closure of the convex hull of ∂eK, i.e. K = (con(∂eK))−, where − and con denote the closure and the convex hull, respectively. A mapping f : K → R is said to be affine if, for all x, y ∈ K and any λ ∈ [0, 1], we have f (λx + (1 − λ)y) = λf (x) + (1 − λ)f (y). Given a compact convex set K in a topological vector space, we denote by Aff(K) the set of all affine continuous functions on K. Then Aff(K) is a po-subgroup of the po-group C(K) of all continuous real-valued functions on K (we recall that, for f, g ∈ C(K), f ≤ g iff f (x) ≤ g(x) for any x ∈ K), hence it is an Archimedean unital po-group with the strong unit 1. In addition, C(K) is an ℓ-group (= lattice ordered group). For example, if E is an effect algebra such that its state space is nonempty, then for any a ∈ E, the function a(s) = s(a), s ∈ S(E), is a continuous affine function belonging to Aff(S(E)). EVERY STATE ON INTERVAL EFFECT ALGEBRA IS INTEGRAL 7 In contrast to C(K), the space Aff(K) is not necessarily a lattice. Indeed, accord- ing to [Goo, Thm 11.21], it is possible to show that if (G, u) is a unital interpolation group, then Aff(S(G, u)) is a lattice if and only if ∂eS(G, u) is compact. Let now K be a compact convex subset of a locally convex Hausdorff space, and let S = S(Aff(K), 1). Then the evaluation mapping ψ : K → S defined by ψ(x)(f ) = f (x) for all f ∈ Aff(K) (x ∈ K) is an affine homeomorphism of K onto S, see [Goo, Thm 7.1]. Let B0(K) be the Baire σ-algebra, i.e. the σ-algebra generated by compact Gδ- sets (a Gδ-set is a countable intersection of open sets) of a compact set K, and the elements of B0(K) are said to be Baire sets, and a σ-additive (signed) measure of B0(K) is a Baire measure. Similarly, B(K) is the Borel σ-algebra of K generated by all open subsets of K and any element of B(K) is said to be a Borel set, and any σ-additive (signed) measure on it is said to be a Borel measure. It is well-known that each Baire measure can be extended to a unique regular Borel measure. Let M(K) denote the set of all finite signed regular Borel measures on B(K) and let M+ 1 (K) denote the set of all probability measures, that is, all positive regular Borel measures µ ∈ M(K) such that µ(K) = 1. We recall that a Borel measure µ is called regular if inf{µ(O) : Y ⊆ O, O open} = µ(Y ) = sup{µ(C) : C ⊆ Y, C closed} for any Y ∈ B(K). Due to the Riesz Representation Theorem, [DuSc], we can identify the set M(K) with the dual of the space C(K) by setting µ(f ) =ZK f dµ (4.1) for all µ ∈ M(K) and all f ∈ C(K). We endow M(K) with the weak∗ topology, i.e., a net µα converges to an element µ ∈ M(K) iff µα(f ) → µ(f ) for all f ∈ C(K). Let x be a fixed point in K. Then there is a unique Borel regular measure µ = µx (depending on x) such that f (x) =ZK f dµ, f ∈ C(K). (4.2) Let δx be the Dirac measure concentrated at the point x ∈ K, i.e., δx(Y ) = 1 iff x ∈ Y , otherwise δx(Y ) = 0; then every Dirac measure is a regular Borel probability measure. In formula (4.2) we have then µx = δx. Moreover, [Goo, Prop 5.24], the mapping gives a homeomorphism of K onto ∂eM+ ǫ : x 7→ δx 1 (K). (4.3) Suppose that K is a nonempty compact subset of a locally convex space V, and 1 (K). A point x in V is said to be represented by µ (or x is the barycenter let µ ∈ M+ of µ) if f (x) =ZK f dµ, f ∈ V ∗, where V ∗ stands for the set of continuous linear functionals f in V. 8 ANATOLIJ DVURE CENSKIJ From the Krein -- Mil'man Theorem, it is possible to derive [Phe, p. 6], that every point x of a compact convex set K of a locally convex space V is the barycenter of a probability measure µ ∈ M+ 1 (K) such that µ((∂eK)−) = 1. We note that according to a theorem by Bauer [Phe, Prop 1.4], if K is a convex compact subset of a locally convex space V, then a point x is an extreme point of K iff the Dirac measure δx is a unique probability measure from M+ 1 (K) whose barycenter is x. 5. Choquet Simplices We show that, for theory of effect algebras, simplices are important, for more about simplices see in [Goo, Alf, Phe]. In what follows, for the reader convenience, we present some necessary notions and results on Choquet and Bauer simplices. We recall that a convex cone in a real linear space V is any subset C of V such that (i) 0 ∈ C, (ii) if x1, x2 ∈ C, then α1x1 + α2x2 ∈ C for any α1, α2 ∈ R+. A strict cone is any convex cone C such that C ∩ −C = {0}, where −C = {−x : x ∈ C}. A base for a convex cone C is any convex subset K of C such that every non-zero element y ∈ C may be uniquely expressed in the form y = αx for some α ∈ R+ and some x ∈ K. We recall that in view of [Goo, Prop 10.2], if K is a non-void convex subset of V, and if we set C = {αx : α ∈ R+, x ∈ K}, then C is a convex cone in V, and K is base of C iff there is a linear functional f on V such that f (K) = 1 iff K is contained in a hyperplane in V which misses the origin. Any strict cone C of V defines a partial order ≤C via x ≤C y iff y − x ∈ C. It is clear that C = {x ∈ V : 0 ≤C x}. A lattice cone is any strict convex cone C in V such that C is a lattice under ≤C . A simplex in a linear space V is any convex subset K of V that is affinely isomorphic to a base for a lattice cone in some real linear space. A simplex K in a locally convex Hausdorff space is said to be (i) Choquet if K is compact, and (ii) Bauer if K and ∂eK are compact. For example, for the important quantum mechanical example, if H is a separable complex Hilbert space, S(E(H)) is not a simplex due to [AlSc, Thm 4.4], where it is shown that the state space of S(E(H)) for the two-dimensional H is isomorphic to the unit three-dimensional real ball, consequently S(E(H)) is not a simplex whenever dim H > 1. Of course, E(H) does not satisfy (RDP): Theorem 5.1. If an effect algebra E satisfies (RDP), then S(E) is a Choquet simplex. Proof. In view of Theorem 2.1, there is a unital po-group (G, u) with interpolation such that E = Γ(G, u). Since S(E) is affinely homeomorphic with S(G, u), we have that S(E) is a Choquet simplex iff so is S(G, u). By [Goo, Thm 10.17], S(G, u) is a Choquet simplex. We recall that S(G, u) is a base for the positive cone of the Dedekind complete lattice-ordered real vector space of all relatively bounded homomorphisms from G to R. (cid:3) Example 5.2. If E is an interval effect algebra that does not satisfies (RDP), then S(E) is not necessarily a Choquet simplex. EVERY STATE ON INTERVAL EFFECT ALGEBRA IS INTEGRAL 9 Indeed, let G = {(x1, x2, x3, x4) ∈ Z4 : x1 + x2 = x3 + x4}, u = (1, 1, 1, 1), and E = Γ(G, u). If si(x1, x2, x3, x4) = xi, then ∂eS(E) = {s1, s2, s3, s4} and S(E) is not a simplex because (s1 + s2)/2 = (s3 + s4)/2, see also [Goo, Cor 10.8]. Another example is S(E(H)) for the effect algebra E(H). The importance of Choquet and Bauer simplices follows from the fact that if K is a convex compact subset of a locally convex Hausdorff space, then K is a Choquet simplex iff (Aff(K), 1) is an interpolation po-group, [Goo, Thm 11.4], and K is a Bauer simplex iff (Aff(K), 1) is an ℓ-group, [Goo, Thm 11.21]. Consequently, there is no MV-algebra whose state space is affinely isomorphic to the closed square or the closed unit circle. For two measures µ and λ we write µ ∼ λ iff µ(f ) = λ(f ), f ∈ Aff(K). If µ and λ are nonnegative regular Borel measures on K, write µ ≺ λ iff µ(f ) ≤ λ(f ), f ∈ Con(K), where Con(K) is the set of all continuous convex functions f on K (that is f (αx1 + (1 − α)x2) ≤ αf (x1) + (1 − α)f (x2) for x1, x2 ∈ K and α ∈ [0, 1]). Then ≺ is a partial order on the cone of nonnegative measures. The fact λ ≺ µ and µ ≺ λ implies λ = µ follows from the fact that Con(K) − Con(K) is dense in C(K). Moreover, for any probability measure (= regular Borel probability measure) λ there is a maximal probability measure µ such that µ ≻ λ, [Phe, Lem 4.1], and every point x ∈ K can be represented by a maximal probability measure. From the Bishop-de Leeuw Theorem, [Alf, Cor I.4.12], see also [Phe, p. 24], it follows that if µ is a maximal probability measure on B(K), then for any Baire set B disjoint with ∂eK, µ(B) = 0. In general, the statement is not true if we suppose that B is Borel instead Baire. An important characterization of Choquet simplices is given by Choquet -- Meyer, [Phe, Thm p. 66], saying that a compact convex subset K of a locally convex space is a Choquet simplex iff, for every point x ∈ K, there is a unique maximal probability measure µx such that µx ∼ δx. That is, f (x) =ZK f dµx, f ∈ Aff(K). (5.1) And the characterization by Bauer, [Alf, Thm II.4.1], says that a Choquet sim- plex K is Bauer iff for any point x ∈ K, there is a unique Baire probability (equiv- alently, regular Borel) measure µx such that µx(∂eK) = 1 and (5.1) holds. 6. Effect Algebras with the Bauer State Property We present the main results of the paper, namely we show that every state on an interval effect algebra is an integral. We show that in some cases every state corresponds to a unique regular Borel probability measure. In addition, we generalize Horn -- Tarski result also for interval effect algebras and its subalgebras. We say that an effect algebra has the Bauer state property ((BSP), for short) if the state space S(E) is a Bauer simplex. For example, every lattice ordered effect algebra with (RDP) has (BSP). The next example is a case when also a non-lattice ordered effect algebra with (RDP) has (BSP). The next example is from [Goo, Ex 4.20]. 10 ANATOLIJ DVURE CENSKIJ −→ × H Example 6.1. Let H = Z be a po-group with the discrete ordering and G = Q be the lexicographic product. Then G has interpolation but G is not a lattice. If we set u = (1, 0), then u is a strong unit in G, and let E = Γ(G, u). If s is a state on G, then due to n(k/n, 0) = u we have s(k/n, 0) = k/n. On the other hand, −u ≤ n(0, b) ≤ u so that −1 ≤ ns(0, b) ≤ 1 which proves s(0, b) = 0. Hence, s(a, b) = a proving that S(G, u) is a singleton and therefore, E has (BSP). In addition, the direct product En has also (BSP) for any integer n ≥ 1. We note that in our example Ker(s) = {(0, b) : b ∈ H}, and then the quotient effect algebra (G/Ker(s), u/Ker(s)) ∼= (Q, 1). Theorem 6.2. Let E be an interval effect algebra and let s be a state on E. Let ψ : E → Aff(S(E)) be defined by ψ(a) := a, a ∈ E, where a is a mapping from S(E) into [0, 1] such that a(s) := s(a), s ∈ S(E). Then there is a unique state s on the unital po-group (Aff(S(E)), 1) such that s(a) = s(a) for any a ∈ E. Moreover, this state can be extended to a state on (C(S(E), 1)). The mapping s 7→ s defines an affine homeomorphism from the state space S(E) onto S(Γ(Aff(S(E)), 1)). If, in addition, E is has an order system of states, then bE := {a : a ∈ E} is an effect-clan that is a sub-effect algebra of Γ(Aff(S(E)), 1) and every state s on E can be extended to a state s on Γ(C(S(E)), 1) as well as to a state on the unital po-group (C(S(E)), 1). Proof. Since E is an interval effect algebra and E = Γ(G, u) for some unigroup (G, u), S(E) is non-void. The mapping ψ can be uniquely extended to a po-group homomorphism ψ : G → Aff(S(E)). Let s be a state on (G, u) that is a unique extension of a state s. Applying now [Goo, Prop 7.20], the statement follows, in particular, we have that there is a unique state s on (Aff(S(E)), 1) such that s(a) = s(a), a ∈ E. Due to [Goo, Cor 4.3], we see that this state can be extended to a state on (C(S(E)), 1) as well as to a state on Γ(C(S(E)), 1). The affine homeomorphism s 7→ s follows from [Goo, Thm 7.1]. If E is Archimedean, the result follows from Proposition 3.1 and the first part (cid:3) of the present proof. Theorem 6.3. Let E be an interval effect algebra such that S(E) is a simplex and let s be a state on E. Then there is a unique maximal regular Borel probability measure µs ∼ δs on B(S(E)) such that s(a) =ZS(E) a(x) dµs(x), a ∈ E. (6.1) Proof. Due to our hypothesis, S(E) is a Choquet simplex. By Theorem 6.2, there is a unique state s on (Aff(S(E)), 1) such that s(a) = s(a), a ∈ A. Applying the result by Choquet -- Meyer, [Phe, Thm p. 66], see (5.1), we have f (s) =ZS(E) f (x) dµs, f ∈ Aff(S(E)). Since a ∈ Aff(S(E)) for any a ∈ E, we have the statement in question. (cid:3) We recall that a state s on E is σ additive if an ր a, then limn s(an) = s(a). In the same manner we define a σ-additive state on (G, u). EVERY STATE ON INTERVAL EFFECT ALGEBRA IS INTEGRAL 11 Theorem 6.4. Let E = Γ(G, u) be an interval effect algebra where (G, u) is a unigroup, and let S(E) be a simplex. If s is σ-additive, then its unique extension, s, on (G, u) is σ-additive. Proof. Let s be a σ-additive state on E and let s be its unique extension onto (G, u). By Theorem 6.3, there is a unique maximal regular Borel probability measure µs ∼ δs such that (6.1) holds. For any g ∈ G, let g denote the function on S(E) such that g(s) = s(g), s ∈ S(E), where s is a unique extension of s onto (G, u). If g ∈ G+, then g = g1 + · · · + gk for some g1, . . . , gk ∈ E, g = g1 + · · · + gk, and s(g) = s(g1) + · · · + s(gk) g1(x) dµs(x) + · · · +ZS(E) gk(x) dµs(x) = ZS(E) = ZS(E) g(x) dµs(x). The σ-additivity of µs entails due to the Lebesgue Convergence Theorem, [Hal], that if gn ր g, then limn s(gn) = s(g) for any g ∈ G+. If now gn ց 0, then limn s(gn) = 0; in fact, there is an integer k ≥ 1 such that gn ≤ g1 ≤ ku for any n ≥ 1. This gives ku − gn ր ku and limn s(ku − gn) = s(ku), so that limn s(gn) = 0. Finally, let g ∈ G be arbitrary and let gn ր g. Then g − gn ց 0 that gives (cid:3) limn s(gn) = s(g) and s is σ-additive on (G, u) as claimed. Theorem 6.5. Let E be an interval effect algebra with (BSP) and let s be a state on E. Then there is a unique regular Borel probability measure, µs, on B(S(E)) such that s(a) =Z∂e S(E) a(x) dµs(x), a ∈ E. (6.2) Proof. Due to Theorem 6.3, we have a unique regular Borel probability measure µs ∼ δs such that (6.1) holds. The characterization of Bauer simplices, [Alf, Thm II.4.1], says that then µs is a unique regular Borel probability measure µs on such that (6.1) holds and µs(∂eS(E)) = 1. Hence, (6.2) holds. (cid:3) We recall that Theorems 6.2 -- 6.5 are applicable to every (i) effect algebra satis- fying (RDP) and (BSP), respectively, (ii) MV-algebra and therefore, Theorem 6.5 generalizes the result of [Pan, Kro]. Since the state space of E(H) is not a simplex, the above theorems are not applicable for this important case. However, the next result shows that also in this case, every state s on E(H) is an integral over some Bauer simplex, but the uniqueness of a regular Borel probability measure µs is not more guaranteed. Theorem 6.6. Let s be a state on an interval effect algebra E. Then there is a regular Borel probability measure, µs, on the Borel σ-algebra B(S(E)) such that s(a) =ZS(E) a(x) dµs(x), a ∈ E. (6.3) 12 ANATOLIJ DVURE CENSKIJ Proof. The set S(E) is non-void and a compact convex Hausdorff topological space. Given a state s ∈ S(E), by Theorem 6.2, s can be uniquely extended to a state s on (Aff(S(E)), 1). Since (Aff(S(E)), 1)) is a unital po-group that is a subgroup of the unital group (C(S(E)), 1), but (C(S(E)), 1) is an ℓ-group. Therefore, this state s can be by [Goo, Cor 4.3] extended to a state s′ on the unital ℓ-group C(S(E), 1). But (C(S(E)), 1) satisfies interpolation and Γ(C(S(E)), 1) has (BSP), therefore by Theorem 6.5, there is a regular Borel probability measure νs on the Bauer simplex Ω = S(C(S(E)), 1) such that f (s′) =Z∂eΩ f (x) dνs(x), f ∈ Γ(C(S(E)), 1). (6.4) Let ǫ : S(E) → ∂eM+ 1 (S(E)) defined by (4.3). Then ǫ is a homeomorphism. For any a ∈ E, by [Goo, Cor 7.5], there is a unique continuous affine mapping a : M+ 1 (S(E)) → [0, 1] such that a = a ◦ ǫ. Therefore, a ∈ Γ(C(S(E)), 1) and s(a) = a(s) = a(δs). The homeomorphism ǫ defines a regular Borel measure µs := νs ◦ ǫ on B(S(E)), and using (6.4) and transformation of integrals, we have s(a) = a(ǫ(s)) = Z∂eΩ = ZS(E) a(y) dνs(y) =Zǫ(S(E)) a(x) dµs(x). a ◦ ǫ−1(y) dνs(y) This implies (6.3). (cid:3) Remark 6.7. We note that reasons why we do not know to guarantee the unique- ness of µs in (6.3) are facts that S(E) is not necessarily a Bauer simplex nor a Choquet one, and Aff(S(E)) is sup-norm closed in C(S(E)) so we cannot extend a state on (Aff(S(E)), 1) by continuity to a state on (C(S(E)), 1). Corollary 6.8. Let E be an interval effect algebra with (BSP). The mapping s 7→ µs, where µs is a unique regular Borel probability measure satisfying (6.2), is an affine homeomorphism between S(E) and the set of regular Borel probability measures on B(S(E)) endowed with the weak∗ topology. A state s on E is extremal if and only if µs in (6.2) is extremal. In such a case, µs = δs. Proof. The fact that that the mapping s 7→ µs is affine and injective follows from Theorem 6.5 and (6.2). The surjectivity is clear because if µ is a regular Borel probability measure, then µ defines via (6.2) some state, s, on E. The continuity follows from [Alf, Thm II.4.1(iii)]. Since every Dirac measure δs is always a regular Borel probability measure, (6.2) entails that s has to be extremal. Conversely, if s is extremal, the uniqueness of µs yields that µs = δs. (cid:3) It is worthy to recall that Corollary 6.8 shows that every (finitely additive) state on an interval effect algebra satisfying (BSP) is in a one-to-one correspondence with (σ-additive) regular Borel probability measure on some Borel σ-algebra trough formula (6.2). This observation is interesting because according to de Finetti, a probability measure is only a finitely additive measure, whilst by Kolmogorov [Kol] a probability measure is assumed to be σ-additive. (6.2) shows that there is a natural coexistence between both approaches. EVERY STATE ON INTERVAL EFFECT ALGEBRA IS INTEGRAL 13 A famous result by Horn and Tarski [HoTa] states that every state on a Boolean subalgebra A of a Boolean algebra B can be extended to a state on B, of course not in unique way. E.g. the two-element Boolean subalgebra, {0, 1} has a unique state, 0-1-valued, and it can be extended to many distinct states on B. This result was generalized for MV-algebras, see [Kro1, Thm 6]. In what follows, we generalize this result also for interval effect algebras, and this gives a variant of the Horn -- Tarski Theorem. Theorem 6.9. Let E be an interval effect algebra and F its sub-effect algebra. Then every state s on F can be extended to a state on E. Proof. Let E be an interval effect algebra and F be its subalgebra. According to [DvPu, Cor 1.4.5], F is also an interval effect algebra. According to Remark 2.2, we can assume that E = Γ(G, u), where u is a universal strong unit for E. Then F ∼= Γ(H ′, u′), where H ′ is a po-group with a universal strong unit u′. Let ιF be the embedding of F into E and let ι be an isomorphism of Γ(H ′, u′) onto F. Assume that s is a state on F. Then s′ = s ◦ ι is a state on Γ(H ′, u′). Applying again Remark 2.2, the state s′ can be extended to a unique state, s′′, on Γ(H ′, u′). On the other hand, the mapping ι can be extended to a unique group homomorphism, bι, from (H ′, u′) into (G, u) such thatbι(u′) = u. Because ι was injective, so isbι, and therefore, bι is an embedding. This entails that on the unital subgroup (bι(H ′), u) we have a state s′′′ = s′′ ◦ (bi)−1 that is an extension of the original state s on F. Applying [Goo, Cor 4.3], we see that s′′′ can be extended to a state s on (G, u) so that, the restriction of s onto E = Γ(G, u) is an extension of a state s on F. (cid:3) Corollary 6.10. Let E be an effect algebra satisfying (RDP) and let F be its subalgebra. Then every state on F can be extended to a state on E. Proof. If E satisfies (RDP), then E is an interval effect algebra. The desired result now follows from Theorem 6.9. (cid:3) 7. Conclusion Using techniques of simplices and affine continuous functions on a convex com- pact Hausdorff space, we have showed that every state on an interval effect algebra, i.e. on an interval [0, u] of a unital po-group (G, u), can be represented by an in- tegral, Theorems 6.3 -- 6.6. A sufficient condition for this is the condition that the state space, S(E), of an interval effect algebra E is a Choquet simplex, Theorem 6.4. In this case it is even a Bauer simplex, Theorem 6.5, and there is a one-to-one correspondence between the set of all states on E and the set of all regular Borel probability measures on the Borel σ-algebra B(S(E)), Corollary 6.8. On the other side, an important example for mathematical foundations of quan- tum mechanics, the state space of the effect algebra E(H), the set of all Hermitian operators on a Hilbert space H that are between the zero operator and the identity, is not a simplex, but also in this case we can represent every state as an integral through a regular Borel probability measure over some Bauer simplex, Theorem 6.6, however in this case, the uniqueness of a Borel probability measure is not more guaranteed. In addition, we have proved a variant of the Horn -- Tarski Theorem that every state on a subalgebra of an unital effect algebra can be extended to a state on the interval effect algebra, Theorem 6.9. 14 ANATOLIJ DVURE CENSKIJ References [Alf] E.M. Alfsen, "Compact Convex Sets and Boundary Integrals", Springer-Verlag, Berlin, 1971. [AlSc] E.M. Alfsen, F.W. Schultz, "State Spaces of Operator Algebras", Birhauser, 2001. [Cha] C.C. Chang, Algebraic analysis of many valued logics, Trans. Amer. Math. Soc. 88 (1958), 467 -- 490. [CDM] R. Cignoli, I.M.L. D'Ottaviano, D. Mundici, "Algebraic Foundations of Many-valued Reasoning", Kluwer Academic Publ., Dordrecht, 2000. [DuSc] N. Dunford, J.T. Schwartz, "Linear Operators, Part I," Interscience, New York-London, 1958. [Dvu] A. Dvurecenskij, "Gleason's Theorem and Its Applications", Kluwer Academic Publisher, Dordrecht/Boston/London, 1993, [DvVe1] A. Dvurecenskij, T. Vetterlein, Pseudoeffect algebras. I. Basic properties, Inter. J. Theor. Phys. 40 (2001), 685 -- 701. [DvVe2] A. Dvurecenskij, T. Vetterlein, Pseudoeffect algebras. II. Group representation, Inter. J. Theor. Phys. 40 (2001), 703 -- 726. [DvPu] A. Dvurecenskij, S. Pulmannov´a, "New Trends in Quantum Structures", Kluwer Acad. Publ., Dordrecht, Ister Science, Bratislava, 2000. [FoBe] D.J. Foulis, M.K. Bennett, Effect algebras and unsharp quantum logics, Found. Phys. 24 [Gle] (1994), 1325 -- 1346. A.M. Gleason, Measures on the closed subspaces of a Hilbert space, J. Math. Mech. 6 (1957), 885 -- 893. [Goo] K.R. Goodearl, "Partially Ordered Abelian Groups with Interpolation", Math. Surveys and Monographs No. 20, Amer. Math. Soc., Providence, Rhode Island, 1986. [Hal] P.R. Halmos, "Measure Theory", Springer-Verlag, New York, Heidelberg, Berlin, 1988. [HoTa] A. Horn, A. Tarski, Measures on Boolean algebras, Trans. Amer. Math. Soc. 64 (1948), 467 -- 497. [Kol] A.N. Kolmogorov, "Grundbegriffe der Wahrscheinlichkeitsrechnung", Berlin, 1933. [KoCh] F. Kopka, F. Chovanec, D-posets, Math. Slovaca 44 (1994), 21 -- 34. [Kro] T. Kroupa, Every state on semisimple MV-algebra is integral. Fuzzy Sets and Systems 157 (2006), 2771 -- 2782. [Kro1] T. Kroupa, Representation and extension of states on MV-algebras. Archive Math. Logic 45 (2006), 381 -- 392. [LuZa] W.A.J. Luxemburg, A.C. Zaanen, "Riesz Spaces, I" North-Holland, Amsterdam, London, 1971. [Mun] D. Mundici, Interpretation of AF C ∗-algebras in Lukasiewicz sentential calculus, J. Funct. Anal. 65 (1986), 15 -- 63. [Pan] G. Panti, Invariant measures in free MV-algebras, Comm. Algebra 36 (2008), 2849 -- 2861. [Phe] R.R. Phelps, "Lectures on Choquet's Theorem", Van Nostrand, Princeton, 1965. [Rav] K. Ravindran, On a structure theory of effect algebras, PhD thesis, Kansas State Univ., Manhattan, Kansas, 1996.
1104.0750
1
1104
2011-04-05T07:34:33
Every maximally monotone operator of Fitzpatrick-Phelps type is actually of dense type
[ "math.FA" ]
We show that every maximally monotone operator of Fitzpatrick-Phelps type defined on a real Banach space must be of dense type. This provides an affirmative answer to a question posed by Stephen Simons in 2001 and implies that various important notions of monotonicity coincide.
math.FA
math
Every maximally monotone operator of Fitzpatrick-Phelps type is actually of dense type Heinz H. Bauschke∗, Jonathan M. Borwein†, Xianfu Wang‡, and Liangjin Yao§ April 4, 2011 Abstract We show that every maximally monotone operator of Fitzpatrick-Phelps type defined on a real Banach space must be of dense type. This provides an affirmative answer to a question posed by Stephen Simons in 2001 and implies that various important notions of monotonicity coincide. 2010 Mathematics Subject Classification: Primary 47H05; Secondary 46B10, 47N10, 90C25 Keywords: Fitzpatrick function, maximally monotone operator, monotone operator, mul- tifunction, operator of type (D), operator of type (FP), operator of type (NI), set-valued operator. 1 Introduction Throughout this note, we assume that X is a real Banach space with norm k · k, that X ∗ is the continuous dual of X, and that X and X ∗ are paired by h·, ·i. Let A : X ⇒ X ∗ be ∗Mathematics, Irving K. Barber School, University of British Columbia, Kelowna, B.C. V1V 1V7, Canada. E-mail: [email protected]. †CARMA, University of Newcastle, Newcastle, New South Wales 2308, Australia. E-mail: [email protected]. ‡Mathematics, Irving K. Barber School, University of British Columbia, Kelowna, B.C. V1V 1V7, Canada. E-mail: [email protected]. §Mathematics, Irving K. Barber School, University of British Columbia, Kelowna, B.C. V1V 1V7, Canada. E-mail: [email protected]. 1 a set-valued operator (also known as multifunction) from X to X ∗, i.e., for every x ∈ X, Ax ⊆ X ∗, and let gra A = (cid:8)(x, x∗) ∈ X × X ∗ x∗ ∈ Ax(cid:9) denote the graph of A. The domain of A is dom A = (cid:8)x ∈ X Ax 6= ∅(cid:9), while ran A = A(X) is the range of A. Recall that A is monotone if (1) hx − y, x∗ − y ∗i ≥ 0, ∀(x, x∗) ∈ gra A ∀(y, y ∗) ∈ gra A, and maximally monotone if A is monotone and A admits no proper monotone extension. It will be convenient to also say that gra A is monotone or maximally monotone respectively in this case. We can then simply say that (x, x∗) ∈ X × X ∗ is monotonically related to gra A if {(x, x∗)} ∪ gra A is monotone. We now recall the three fundamental types of monotonicity. Definition 1.1 Let A : X ⇒ X ∗ be maximally monotone. Then three key types of monotone operators are defined as follows. (i) A is of dense type or type (D) (1971, [11]) if for every (x∗∗, x∗) ∈ X ∗∗ × X ∗ with inf (a,a∗)∈gra A ha − x∗∗, a∗ − x∗i ≥ 0, there exist a bounded net (aα, a∗ verges to (x∗∗, x∗). α)α∈Γ in gra A such that (aα, a∗ α)α∈Γ weak*×strong con- (ii) A is of type negative infimum (NI) (1996, [16]) if sup (cid:0)ha, x∗i + ha∗, x∗∗i − ha, a∗i(cid:1) ≥ hx∗∗, x∗i, ∀(x∗∗, x∗) ∈ X ∗∗ × X ∗. (a,a∗)∈gra A (iii) A is of type Fitzpatrick-Phelps (FP) (1992, [10]) if whenever U is an open convex subset of X ∗ such that U ∩ ran A 6= ∅, x∗ ∈ U, and (x, x∗) ∈ X × X ∗ is monotonically related to gra A ∩ (X × U) it must follow that (x, x∗) ∈ gra A. All three of these properties are known to hold for the subgradient of a closed convex function and for every maximally monotone operator on a reflexive space. These and other relationships known amongst these and other monotonicity notions are described in [6, Chap- ter 8]. Monotone operators are fundamental objects in modern Optimization and Variational Analysis; see, e.g., [3, 4, 5], the books [2, 6, 7, 14, 17, 19, 15, 21] and the references therein. In Theorem 3.1 of this paper, we provide an affirmative to the following question, posed by S. Simons [18, Problem 18, page 406]: Let A : X ⇒ X ∗ be maximally monotone such that A is of type (FP). Is A necessarily of type (D)? 2 In consequence, in Corollary 3.2 we record that the three notions in Definition 1.1 actually coincide. We shall utilize the following notation, in addition to standard notions from convex anal- is It is very convenient to identify X with its canonical image in the ysis: The open unit ball (cid:8)x ∈ X kxk ≤ 1(cid:9). bidual space X ∗∗. Moreover, X × X ∗ and (X × X ∗)∗ = X ∗ × X ∗∗ are paired via in X is UX = (cid:8)x ∈ X kxk < 1(cid:9), and the closed unit ball h(x, x∗), (y ∗, y ∗∗)i = hx, y ∗i + hx∗, y ∗∗i, where (x, x∗) ∈ X ×X ∗ and (y ∗, y ∗∗) ∈ X ∗ ×X ∗∗. We recall the following basic fact regarding the second dual ball: Fact 1.2 (Goldstine) (See [13, Theorem 2.6.26] or [8, Theorem 3.27].) The weak*-closure of BX in X ∗∗ is BX ∗∗. The remainder of this paper is organized as follows. In Section 2, we record auxiliary results for subsequent use. The main result (Theorem 3.1) and the promised corollary (Corollary 3.2) are provided in Section 3. 2 Preliminary monotonicity results A now fundamental tool of modern monotone operator theory originated with Simon Fitz- patrick in 1988. It is reprised next: Fact 2.1 (Fitzpatrick) (See [9, Corollary 3.9].) Let A : X ⇒ X ∗ be maximally monotone, and let us set (2) FA : X ∗∗ × X ∗ → ]−∞, +∞] : (x∗∗, x∗) 7→ sup (cid:0)hx∗∗, a∗i + ha, x∗i − ha, a∗i(cid:1). (a,a∗)∈gra A Then for every (x, x∗) ∈ X × X ∗, the inequality hx, x∗i ≤ FA(x, x∗) is true, and equality holds if and only if (x, x∗) ∈ gra A. The function FAX×X ∗ is the classical Fitzpatrick function associated with A. The first relevant relationship established for (FP) operators is due to Stephen Simons: Fact 2.2 (Simons) (See [18, Theorem 17] or [19, Theorem 37.1].) Let A : X ⇒ X ∗ be maximally monotone and of type (D). Then A is of type (FP). 3 The most powerful current information is captured in the following result. Fact 2.3 (Simons / Marques Alves and Svaiter) (See [16, Lemma 15] or [19, Theo- rem 36.3(a)], and [12, Theorem 4.4].) Let A : X ⇒ X ∗ be maximally monotone. Then A is of type (D) if and only if it is of type (NI). The implication type (NI) implies type (D) -- which we exploit -- is very recently due to Marques Alves and Svaiter [12]. 3 Main result The next theorem is our main result. provides the affirmative answer promised to Simons's problem posed in [18, Problem 18]. In conjunction with the corollary that follows, it Theorem 3.1 Let A : X ⇒ X ∗ be maximally monotone such that A is of type (FP). Then A is of type (NI). Proof. After translating the graph if necessary, we can and do suppose that (0, 0) ∈ gra A. Let (x∗∗ 0) ∈ X ∗∗ × X ∗. We must show that 0 , x∗ (3) and we consider two cases. FA(x∗∗ 0 , x∗ 0) ≥ hx∗∗ 0 , x∗ 0i Case 1 : x∗∗ 0 ∈ X. Then (3) follows directly from Fact 2.1. Case 2 : x∗∗ 0 ∈ X ∗∗ r X. By Fact 1.2, there exists a bounded net (xα)α∈I in X that weak* converges to x∗∗ have 0 . Thus, we (4) and (5) M = sup α∈I kxαk < +∞ hxα, x∗ 0i → hx∗∗ 0 , x∗ 0i. Now we consider two subcases. Subcase 2.1 : There exists α ∈ I, such that (xα, x∗ 0) ∈ gra A. By definition, FA(x∗∗ 0 , x∗ 0) ≥ hxα, x∗ 0i + hx∗∗ 0 , x∗ 0i − hxα, x∗ 0i = hx∗∗ 0 , x∗ 0i. 4 Hence (3) holds. Subcase 2.2 : We have (6) Set (7) (xα, x∗ 0) /∈ gra A, ∀α ∈ I. Uε = [0, x∗ 0] + εUX ∗, where ε > 0. Observe that Uε is open and convex. Since (0, 0) ∈ gra A, we have, by definition 0 ∈ Uε. In view of (6) and because A is of type (FP), there exists of Uε, 0 ∈ ran A ∩ Uε and x∗ a net (aα,ε, a∗ α,ε) in gra A such that a∗ α,ε ∈ Uε and (8) haα,ε, x∗ 0i + hxα, a∗ α,εi − haα,ε, a∗ α,εi > hxα, x∗ 0i, ∀α ∈ I. Now fix α ∈ I. By (8), haα,ε, x∗ 0i + hx∗∗ 0 , a∗ α,εi − haα,ε, a∗ α,εi > hx∗∗ 0 − xα, a∗ α,εi + hxα, x∗ 0i. Hence, (9) FA(x∗∗ 0 , x∗ 0) > hx∗∗ 0 − xα, a∗ α,εi + hxα, x∗ 0i. Since a∗ α,ε ∈ Uε, there exist (10) such that (11) tα,ε ∈ [0, 1] and b∗ α,ε ∈ UX ∗ a∗ α,ε = tα,εx∗ 0 + εb∗ α,ε. Using (9), (11), and (4), we deduce that FA(x∗∗ 0 , x∗ 0) > hx∗∗ 0 − xα, tα,εx∗ 0 − xα, x∗ 0 − xα, x∗ 0 − xα, x∗ = tα,εhx∗∗ ≥ tα,εhx∗∗ ≥ tα,εhx∗∗ 0 + εb∗ 0i + εhx∗∗ 0i − εkx∗∗ 0i − ε(kx∗∗ α,εi + hxα, x∗ 0i α,εi + hxα, x∗ 0 − xα, b∗ 0i 0 − xαk + hxα, x∗ 0i 0 k + M) + hxα, x∗ 0i. (12) In view of (10) and since α ∈ I was chosen arbitrarily, we take the limit in (12) and obtain with the help of (5) that (13) FA(x∗∗ 0 , x∗ 0) ≥ −ε(kx∗∗ 0 k + M) + hx∗∗ 0 , x∗ 0i. Next, letting ε → 0 in (13), we have (14) FA(x∗∗ 0 , x∗ 0) ≥ hx∗∗ 0 , x∗ 0i. Therefore, (3) holds in all cases. We now obtain the promised corollary: 5 (cid:4) Corollary 3.2 Let A : X ⇒ X ∗ be maximally monotone. Then the following are equivalent. (i) A is of type (D). (ii) A is of type (NI). (iii) A is of type (FP). Proof. First (i) implies (iii) is Fact 2.2; next Theorem 3.1 shows (iii) implies (ii); while Fact 2.3 implies concludes the circle with (ii) implies (i). (cid:4) We note that while the result is now quite easy, it remained inaccessible until [12, Theo- rem 4.4] was available. Remark 3.3 Let A : X ⇒ X ∗ be maximally monotone. Corollary 3.2 establishes the equiv- alences of the key types (D), (NI), and (FP), which as noted all hold when X is reflexive or A = ∂f , where f : X → ]−∞, +∞] is convex, lower semicontinuous, and proper (see [6, 17, 19]). Furthermore, these notions are also equivalent to type (ED), see [20]. For a nonlinear operator they also coincide with uniqueness of maximal extensions to X ∗∗ (see [12]). In [6, p. 454] there is discussion of this result and of the linear case. Finally, when A is a linear relation, it has recently been established that all these notions coincide with monotonicity of the adjoint multifunction A∗ (see [1]). Acknowledgments Heinz Bauschke was partially supported by the Natural Sciences and Engineering Research Council of Canada and by the Canada Research Chair Program. Jonathan Borwein was par- tially supported by the Australian Research Council. Xianfu Wang was partially supported by the Natural Sciences and Engineering Research Council of Canada. References [1] H.H. Bauschke, J.M. Borwein, X. Wang and L. Yao, "For maximally monotone linear relations, dense type, negative-infimum type, and Fitzpatrick-Phelps type all coincide with monotonicity of the adjoint", submitted; http://arxiv.org/abs/1103.6239v1, March 2011. 6 [2] H.H. Bauschke and P.L. Combettes, Convex Analysis and Monotone Operator Theory in Hilbert Spaces, Springer-Verlag, 2011. [3] J.M. Borwein, "Maximal monotonicity via convex analysis", Journal of Convex Analysis, vol. 13, pp. 561 -- 586, 2006. [4] J.M. Borwein, "Maximality of sums of two maximal monotone operators in general Banach space", Proceedings of the American Mathematical Society, vol. 135, pp. 3917 -- 3924, 2007. [5] J.M. Borwein, "Fifty years of maximal monotonicity", Optimization Letters, vol. 4, pp. 473 -- 490, 2010. [6] J.M. Borwein and J.D. Vanderwerff, Convex Functions, Cambridge University Press, 2010. [7] R.S. Burachik and A.N. Iusem, Set-Valued Mappings and Enlargements of Monotone Opera- tors, Springer-Verlag, 2008. [8] M. Fabian, P. Habala, P. H´ajek, V. Montesinos Santaluc´ıa, J. Pelant and V. Zizler, Functional Analysis and Infinite-Dimensional Geometry, Springer-Verlag, 2001. [9] S. Fitzpatrick, "Representing monotone operators by convex functions", in Work- shop/Miniconference on Functional Analysis and Optimization (Canberra 1988), Proceedings of the Centre for Mathematical Analysis, Australian National University, vol. 20, Canberra, Australia, pp. 59 -- 65, 1988. [10] S. Fitzpatrick and R.R. Phelps, "Bounded approximants to monotone operators on Banach spaces", Annales de l'Institut Henri Poincar´e. Analyse Non Lin´eaire, vol. 9, pp. 573 -- 595, 1992. [11] J.-P. Gossez, "Op´erateurs monotones non lin´eaires dans les espaces de Banach non r´eflexifs", Journal of Mathematical Analysis and Applications, vol. 34, pp. 371 -- 395, 1971. [12] M. Marques Alves and B.F. Svaiter, "On Gossez type (D) maximal monotone operators", Journal of Convex Analysis, vol. 17, pp. 1077 -- 1088, 2010. [13] R.E. Megginson, An Introduction to Banach Space Theory, Springer-Verlag, 1998. [14] R.R. Phelps, Convex Functions, Monotone Operators and Differentiability, 2nd Edition, Springer-Verlag, 1993. [15] R.T. Rockafellar and R.J-B Wets, Variational Analysis, 3rd Printing, Springer-Verlag, 2009. [16] S. Simons, "The range of a monotone operator", Journal of Mathematical Analysis and Ap- plications, vol. 199, pp. 176 -- 201, 1996. [17] S. Simons, Minimax and Monotonicity, Springer-Verlag, 1998. [18] S. Simons, "Five kinds of maximal monotonicity", Set-Valued and Variational Analysis, vol. 9, pp. 391 -- 409, 2001. 7 [19] S. Simons, From Hahn-Banach to Monotonicity, Springer-Verlag, 2008. [20] S. Simons, "Banach SSD Spaces and classes of monotone sets", Journal of Convex Analysis, vol. 18, pp. 227 -- 258, 2011. [21] C. Zalinescu, Convex Analysis in General Vector Spaces, World Scientific Publishing, 2002. 8
1504.07323
1
1504
2015-04-28T01:22:33
Non-commutative Functional Calculus and Spectral Theory
[ "math.FA", "math.OA" ]
We develop a functional calculus for $d$-tuples of non-commuting elements in a Banach algebra. The functions we apply are free analytic functions, that is nc functions that are bounded on certain polynomial polyhedra.
math.FA
math
Non-commutative Functional Calculus and Spectral Theory Jim Agler ∗ U.C. San Diego La Jolla, CA 92093 John E. McCarthy † Washington University St. Louis, MO 63130 . A F h t a m [ 1 v 3 2 3 7 0 . 4 0 5 1 : v i X r a April 29, 2015 Abstract We develop a functional calculus for d-tuples of non-commuting elements in a Banach algebra. The functions we apply are free analytic functions, that is nc functions that are bounded on certain polynomial polyhedra. 1 Introduction 1.1 Overview The purpose of this note is to develop an approach to functional calculus and spectral theory for d-tuples of elements of a Banach algebra, with no assumption that the elements commute. In [29], J.L. Taylor considered this problem, for d-tuples in L(X), the bounded linear operators on a Banach space X. His idea was to start with the algebra Pd, the algebra of free polynomials1 in d variables over the complex numbers, and consider what he called “satellite algebras”, that is algebras A ∗Partially supported by National Science Foundation Grant DMS 1361720 †Partially supported by National Science Foundation Grant DMS 1300280 1We shall use free polynomial and non-commuting polynomial in d variables interchange- ably to mean an element of the algebra over the free semi-group with d generators. 1 that contained Pd, and with the property that every representation from Pd to L(X) that extends to a representation of A has a unique extension. As a representation of Pd is determined by choosing the images of the generators, i.e. choosing T = (T 1, . . . , T d) ∈ L(X)d, the extension of the representation to A, when it exists, would constitute an A-functional calculus for T . The class of satellite algebras that Taylor considered, which he called free ana- lytic algebras, were intended to be non-commutative generalizations of the algebras O(U), the algebra of holomorphic functions on a domain U in Cd (and indeed he proved in [29, Prop 3.3] that when d = 1, these constitute all the free analytic algebras). Taylor had already developed a successful O(U) functional calculus for d-tuples T of commuting operators on X for which a certain spectrum (now called the Taylor spectrum) is contained in U — see [26, 27] for the original articles, and also the article [21] by M. Putinar showing uniqueness. An excellent treatment is in [7] by R. Curto. However, in the non-commutative case, Taylor’s approach in [28, 29] using homological algebra was only partially successful. What would constitute a successful theory? This is of course subjective, but we would argue that it should contain some of the following ingredients, and one has to make trade-offs between them. The functional calculus should use algebras A that one knows something about — the more the algebras are understood, the more useful the theory. Secondly, the condition for when a given T has an A-functional calculus should be related to the way in which T is presented as simply as possible. Thirdly, the more explicit the map that sends φ in A to φ(T ) in L(X), the easier it is to use the theory. Finally, restricting to the commutative case, one should have a theory which agrees with the normal idea of a functional calculus. One does not need an explicit notion of spectrum in order to have a functional calculus. If one does have a spectrum, it should be a collection of simpler objects than d-tuples in L(X), just as in the commutative case the spectrum is a collection of d-tuples of complex numbers, which say something about a commuting d-tuple on a Banach space. The approach that we advocate in this note is to replace Cd as the uni- versal set by the nc-universe M[d] := ∪∞ n=1 Md n, where Mn denotes the n-by-n matrices over C, with the induced operator norm from ℓ2 n. In other words, we look at d-tuples of n-by-n matrices, but 2 instead of fixing n, we allow all values of n. We shall look at certain special open sets in M[d]. Let δ be a matrix of free polynomials in d variables, and define Gδ = {x ∈ M[d] : kδ(x)k < 1}. (1.1) The algebras with which we shall work are algebras of the form H ∞(Gδ). We shall define H ∞(Gδ) presently, in Definition 1.4. For now, think of it as some sort of non-commutative analogue of the bounded analytic functions defined on Gδ. We shall develop conditions for a d-tuple in L(X) to have an H ∞(Gδ) functional calculus, in other words for a particular T ∈ L(X)d to have the property that there is a unique extension of the polynomial functional calculus to all of H ∞(Gδ). 1.2 Non-commutative functions Md n=1 Let M[d] = ∪∞ function φ with the property that if x ∈ Md y ∈ Md sx (respectively xs) denote the tuple (sx1, . . . , sxd) (resp. (x1s, . . . , xds)). n. A graded function defined on a subset of M[d] is a n and n+m, and if s ∈ Mn we let m, we let x ⊕ y = (x1 ⊕ y1, . . . , xd ⊕ yd) ∈ Md n, then φ(x) ∈ Mn. If x ∈ Md Definition 1.2. An nc-function is a graded function φ defined on a set D ⊆ M[d] such that i) If x, y, x ⊕ y ∈ D, then φ(x ⊕ y) = φ(x) ⊕ φ(y). ii) If s ∈ Mn is invertible and x, s−1xs ∈ D ∩ Md n, then φ(s−1xs) = s−1φ(x)s. Observe that any non-commutative polynomial is an nc-function on all of M[d]. Subject to being locally bounded with respect to an appropriate topology, nc-functions are holomorphic [3, 10, 12], and can be thought of as bearing an analogous relationship to non-commutative polynomials as holomorphic functions do to regular polynomials. Nc-functions have been studied by, among others: G. Popescu [17–20]; J. Ball, G. Groenewald and T. Malakorn [5]; D. Alpay and D. Kaliuzhnyi- Verbovetzkyi [4]; and J.W. Helton, I. Klep and S. McCullough [9, 10] and Helton and McCullough [11]. We refer to the book [12] by Kaliuzhnyi- Verbovetskyi and V. Vinnikov on nc-functions. We shall define matrix or operator valued nc-functions in the natural way, and use upper-case letters to denote them. 3 Definition 1.3. Let K1 and K2 be Hilbert spaces, and D ⊆ M[d]. We say a function F is an L(K1, K2)-valued nc-function on D if ∀n ∀x∈D∩Md n F (x) ∈ L(Cn ⊗ K1, Cn ⊗ K2), ∀x,y,x⊕y∈D F (x ⊕ y) = F (x) ⊕ F (y), and ∀n ∀x∈D∩Md n ∀s∈Mn s−1xs ∈ D =⇒ F (s−1xs) = (s−1 ⊗ idK1)F (x)(s ⊗ idK2). A special case of Gδ in (1.1) is when d = IJ and δ is the I-by-J rectangular matrix whose (i, j) entry is the [(i − 1)J + j]th coordinate function. We shall give this the special name E: E(x1, . . . , xIJ ) = x1 xJ+1 ... x2 xJ+2 ... x(I−1)J+1 x(I−1)J+2   xJ · · · · · · x2J ... . . . · · · xIJ .   We shall denote the set GE by BI×J . ∞ BI×J = [n=1(cid:8)x = (x1, . . . , xIJ ) ∈ MIJ n : kE(x)k < 1(cid:9) . Definition 1.4. We let H ∞(Gδ) denote the bounded nc-functions on Gδ, and H ∞ L(K1,K2)(Gδ) denote the bounded L(K1, K2)-valued nc-functions on D. These functions were studied in [3] and [2]. When K1 = K2 = C, we shall L(K1,K2)(Gδ). By a matrix-valued H ∞(Gδ) function, L(K1,K2)(Gδ) with both K1 and K2 finite di- identify H ∞(Gδ) with H ∞ we mean an element of some H ∞ mensional. 2 Hilbert tensor norms We wish to define norms on matrices of elements of L(X). If X were restricted to be a Hilbert space H, there would be a natural way to do this by thinking of an I-by-J matrix in L(H) as a linear map from the (Hilbert space) tensor product H ⊗ CJ to H ⊗ CI. We would like to do this in general. Note first that although any Banach space can be embedded in an oper- ator space (see e.g. [16, Chap. 3]), which in turn can be realized as a subset 4 of some L(H), we would lose the multiplicative structure of L(X), so that will not work in general for our purpose. Let us recall some definitions from the theory of tensor products on Ba- nach spaces [8, 23]. A reasonable cross norm on the algebraic tensor product X ⊗ Y of two Banach spaces is a norm τ satisfying (i) For every x ∈ X, y ∈ Y , we have τ (x ⊗ y) = kxkkyk. (ii) For every x∗ ∈ X ∗, y∗ ∈ Y ∗, we have kx∗ ⊗ y∗k(X⊗Y,τ )∗ = kx∗kky∗k. A uniform cross norm is an assignment to each pair of Banach spaces X, Y a reasonable cross-norm on X ⊗ Y such that if R : X1 → X2 and S : Y1 → Y2 are bounded linear operators, then kR ⊗ SkX1⊗Y1→X2⊗Y2 ≤ kRkkSk. A uniform cross norm τ is finitely generated if, for every pair of Banach spaces X, Y and every u ∈ X ⊗ Y , we have τ (u; X ⊗ Y ) = inf{τ (u; M ⊗ N), u ∈ M ⊗ N, dim M < ∞, dim N < ∞}. A finitely generated uniform cross norm is called a tensor norm. Both the injective and projective tensor products are tensor norms [8, Prop.’s 1.2.1, 1.3.2], [23, Section 6.1], and there are others [8, 23]. When τ is a reasonable cross norm, we shall write X ⊗τ Y for the Banach space that is the completion of X ⊗ Y with respect to the norm given by τ . Definition 2.1. Let X be a Banach space. A Hilbert tensor norm on X is an assignment of a reasonable cross norm h to X ⊗ K for every Hilbert space K with the property: If R : X → X and S : K1 → K2 are bounded linear operators, and K1 and K2 are Hilbert spaces, then kR ⊗ SkL(X⊗hK1, X⊗hK2) ≤ kRkL(X)kSkL(K1,K2). (2.2) Any uniform cross norm is a Hilbert tensor norm, but there are others. Most importantly, if X is itself a Hilbert space, then the Hilbert space tensor product is a Hilbert tensor norm. In what follows, we shall use ⊗ without a subscript to denote the Hilbert space tensor product of two Hilbert spaces, and ⊗h to denote a Hilbert tensor norm. Let X be a Banach space, and let h be a Hilbert tensor norm on X. Let R = (Rij) be an I-by-J matrix with entries in L(X). Then we can think of 5 R as a linear operator from X ⊗ CJ to X ⊗ CI. We shall use h to define a norm for R. Formally, let Eij : CJ → CI be the matrix with 1 in the (i, j) slot and 0 elsewhere. Let K be a Hilbert space. Then we define Rh,K : X ⊗h (CJ ⊗ K) → X ⊗h (CI ⊗ K) Rh,K = I J Xi=1 Xj=1 Rij ⊗h (Eij ⊗ idK) (2.3) Then we define kRkh = sup{kRh,Kk : K is a Hilbert space}, (2.4) and (borrowing notation from the Irish use of a dot or s´eimhi´u for an “h”) kRk• = inf{kRkh : h is a Hilbert tensor norm}. (2.5) Let us record the following lemma for future use. Lemma 2.6. Let R = (Rij) be an I-by-J matrix with entries in L(X). Then kRk• ≥ max i,j kRijkL(X). (2.7) Proof: Let Bi be the 1-by-I matrix with ith entry idX, and the other entries the 0 element of L(X). Let Cj be the J-by-1 column matrix, with jth entry idX, and the other entries 0. Let h be any Hilbert tensor norm on X. By (2.2), we have kBikh and kCjkh are ≤ 1, and since h is a reasonable cross norm we get that they both exactly equal 1. We have kRijkL(X) = kBiRCjkL(X) ≤ kRkL(X⊗h CJ ,X⊗hCI ) ≤ kRkh. Since this holds for every h, we get (2.7). ✷ 3 Free analytic functions Here are some of the primary results of [3]. When δ is an I-by-J rectangular matrix with entries in Pd, and x ∈ Md n, we shall think of δ(x) as an element of L(Cn ⊗ CJ , Cn ⊗ CI). If M is a Hilbert space, we shall write δM(x) for δ(x) ⊗ idM, and think of it as an element of L(Cn ⊗ MJ , Cn ⊗ MI) = L(Cn ⊗ (CJ ⊗ M), Cn ⊗ (CI ⊗ M)). 6 Theorem 3.1. Let δ be an I-by-J rectangular matrix of free polynomials, and assume Gδ is non-empty. Let K1 and K2 be finite dimensional Hilbert spaces. A function Φ is in H ∞ L(K1,K2)(Gδ) if and only if there is a function F in H ∞ L(K1,K2)(BI×J ), with kF k ≤ kΦk, such that Φ = F ◦ δ. Theorem 3.2. Let K1 and K2 be finite dimensional Hilbert spaces. If F is in H ∞ L(K1,K2)(BI×J ) and kF k ≤ 1, then there exists an auxiliary Hilbert space M and an isometry V = (cid:20)A B C D(cid:21) : K1 ⊕ M(I) → K2 ⊕ M(J) (3.3) so that for x ∈ BI×J ∩ Md n, F (x) = idCn⊗A+(idCn⊗B)EM(x)[idCn⊗idM(J )−(idCn⊗D)EM(x)]−1(idCn⊗C). (3.4) Consequently, F has the series expansion ∞ F (x) = idCn⊗A+ (idCn⊗B)EM(x)[(idCn⊗D)EM(x)]k−1(idCn⊗C), (3.5) Xk=1 which is absolutely convergent on Gδ. If we write CnA for idCn ⊗A, then equations (3.4) and (3.5) have the more easily readable form F = CnA + CnB EM [I − CnD EM]−1 CnC F (x) = CnA + ∞ Xk=1 CnB EM(x) [CnD EM(x)]k−1 CnC. (3.6) (3.7) We call (3.4) a free realization of F . The isometry V is not unique, but each term on the right-hand side of (3.7) is a free matrix-valued polynomial, each of whose non-zero entries is homogeneous of degree k. So we can rewrite (3.7) as F (x) = ∞ Xk=0 Pk(x) (3.8) where each Pk is a homogeneous L(K1, K2)-valued free polynomial, and which satisfies kPk(x)k ≤ kxkk ∀x ∈ BI×J , ∀ k ≥ 1. (3.9) 7 These formulas ((3.6) or (3.8)) allow us to extend the domain of F from d-tuples of matrices to d-tuples in L(X). Let X be a Banach space, with a Hilbert tensor norm h. Let T = (Tij) be an I-by-J matrix of elements of L(X). If kT kh < 1, (3.10) where kT kh is defined by (2.4), then we can replace EM(x) in (3.4) by Pi,j Tij ⊗h (Eij ⊗idM), and get a bounded operator from X ⊗h K1 to X ⊗h K2, provided we tensor with idX. Definition 3.11. Let K1 and K2 be finite dimensional Hilbert spaces, and let F be a matrix-valued nc-function on BI×J , bounded by 1 in norm, with a free realization given by (3.4), and an expansion into homogeneous L(K1, K2)- valued free polynomials given by (3.8). Let T = (Tij)i=I,j=J i=1,j=1 be an I-by-J matrix of bounded operators on a Banach space X. Let h be a Hilbert tensor norm on X. Then we define F ♯ h(T ) ∈ L(X ⊗h K1, X ⊗h K2) to equal F ♯ h(T ) = ∞ Xk=0 Pk(T ), (3.12) provided that the right-hand side converges absolutely. We extend the definition of F ♯ to functions of norm greater than 1 by scaling. The definition of F ♯ h(T ) may seem to depend on the choice of free real- ization, but in fact it does not, since the polynomials Pk do not depend on the free realization. It does depend subtly on the choice of h, as F ♯ h(T ) is a bounded linear map in L(X ⊗h K1, X ⊗h K2), but these are all the same if K1 = K2 = C. We shall write F ♯(T ) for the dim(K2)-by-dim(K1) matrix F ♯(T ) = ∞ Xk=0 Pk(T ), (3.13) which is a matrix of elements of L(X). In the following theorem we shall write XA for idX ⊗h A, and TM for Pi,j Tij ⊗h (Eij ⊗ idM), where we assume that h is understood. Theorem 3.14. Suppose X is a Banach space, and T is an I-by-J matrix of elements of L(X). Suppose F is as in Theorem 3.2, of norm at most one. 8 (i) If h is a Hilbert tensor norm on X and kT kh < 1, then F ♯ h(T ) = XA + (X B)TM[I − (XD)TM]−1 XC, and (ii) If kT k• < 1, then kF ♯ h(T )k ≤ 1 1 − kT kh . kF ♯(T )k• ≤ 1 1 − kT k• . (3.15) (3.16) (3.17) (iii) If X is a Hilbert space and H is the Hilbert space tensor product, and kT kH < 1 then kF ♯ H(T )k ≤ 1. (3.18) Proof: (i) Let kT kh = r < 1. Let us temporarily denote by G(T ) the right-hand side of (3.15). By 2.2, we have kX Dk ≤ 1, and by (2.4), kTMk < 1. Therefore the Neumann series [I − (XD)TM]−1 = ∞ Xk=0 [XD TM]k converges to a bounded linear operator in L(X ⊗h (CJ ⊗ M)) of norm at most 1 1−r . Using 2.2 again, we conclude that kG(T )kL(X⊗hK1,X⊗hK2) ≤ 1 + r 1 − r = 1 1 − r . (3.19) Replacing T by eiθT , and integrating G(eiθT ) against e−ikθ, we get, for k ≥ 1, 1 2π Z 2π 0 G(eiθT )e−ikθdθ = X B TM[XD TM]k−1 XC = Pk(T ), where Pk is the homogeneous polynomial from (3.8). Therefore G(T ) is given k=0 Pk(T ), and hence equals F ♯(T ), by the absolutely convergent series P∞ proving (3.15), and, by (3.19), also proving (3.16). (ii) This follows from the definition (2.5). 9 (iii) Using the fact that (cid:20)A B C D(cid:21) is an isometry, and equation (3.15), some algebraic rearrangements give I − F ♯ H(T )∗F ♯ H (T ) = XC ∗[I − T ∗ X C. (3.20) Since kTMk < 1, the right-hand side of (3.20) is positive, and so the left-hand side is also, which means kF ♯ ✷ M XD]−1[I − T ∗ MTM][I − (X D)TM]−1 H(T )k ≤ 1. Suppose Φ(x1, . . . , xd) is in H ∞ L(K1,K2)(Gδ). By Theorem 3.1, we can write L(K1,K2)(BI×J ). Let T = (T 1, . . . , T d) ∈ L(X)d. Φ = F ◦ δ, for some F in H ∞ Then δ(T ) is an I-by-J matrix with entries in L(X). If kδ(T )k• < 1, then one would like to define Φ♯ by Φ♯(T ) = F ♯(δ(T )). (3.21) As F is not unique, this raises questions about whether Φ♯ is well-defined. We address this in Section 4. 4 Existence of Functional Calculus Throughout this section, X will be a Banach space, and T = (T 1, . . . , T d) will be a d-tuple of bounded linear operators on X. Let δ be an I-by-J matrix of free polynomials in Pd, and let Gδ = ∪∞ n=1{x ∈ Md n : kδ(x)k < 1}. We shall say that Gδ is a spectral set for T if kp(T )kL(X) ≤ sup x∈Gδ kp(x)k ∀ p ∈ Pd. (4.1) When P is an I-by-J matrix of polynomials, then we shall consider P to be an L(CJ, CI) valued nc-function. We shall let M(Pd) denote all (finite) matrices of free polynomials, with the norm of P (x) given as the operator norm in L(Cn ⊗ CJ , Cn ⊗ CI) where x = (xij) is a matrix with each xij ∈ Mn. If (4.1) holds for all matrices of polynomials, i.e. kP (T )k• ≤ sup x∈Gδ kP (x)k ∀ n, ∀ P ∈ M(Pd), (4.2) 10 we shall say that Gδ is a complete spectral set for T . If inequalities (4.1) or (4.2) are true with the right-hand side multiplied by a constant K, we shall say Gδ is a K-spectral set (respectively, complete K-spectral set) for T . Theorem 4.3. The following are equivalent. (i) There exists s < 1 such that Gδ/s is a K-spectral set for T . (ii) There exists r < 1 such that the map π : f ◦ ( 1 r δ) 7→ f ♯( 1 r δ(T )) is a well-defined bounded homomorphism from H ∞(Gδ/r) to L(X) with norm less than or equal to K that extends the polynomial functional calculus on Pd ∩ H ∞(Gδ/r). Moreover, if these conditions hold, then π is the unique extension of the evaluation homomorphism on the polynomials to a bounded homomorphism from H ∞(Gδ/r) to L(X). Proof: (ii) ⇒ (i): Let s = r. Let q ∈ Pd. If kqkGδ/r is infinite, there is nothing to prove. Otherwise, by Theorem 3.1, there exists f ∈ H ∞(BI×J ) such that q = f ◦ 1 r δ on Gδ/r, and kf k ≤ kqkGδ/r. Since π is well-defined and extends the polynomial evaluation, π(q) = q(T ) = f ♯( 1 r δ(T )). Therefore kq(T )k ≤ K kf ◦ 1 r δkGδ/r ≤ K kqkGδ/r. Now, suppose (i) holds. Choose r in (s, 1). Let φ ∈ H ∞(Gδ/r), and assume that there are functions f1 and f2 in H ∞(BI×J ) such that φ(x) = f1 ◦ ( 1 r δ)(x) = f2 ◦ ( 1 r δ)(x) ∀x ∈ Gδ/r. Expand each fl as in (3.8) into a series of homogeneous polynomials, so fl(x) = ∞ Xk=0 pl k(x), l = 1, 2. 11 By (3.9), we have kpl k(x)k ≤ kxkk. So k N Xk=0 p1 k( 1 r δ(x)) − N Xk=0 p2 k( 1 r δ(x))kGδ/s = k ∞ Xk=N +1 Xk=N +1 ∞ p1 k( 1 r δ(x)) − ∞ Xk=N +1 p2 k( 1 r δ(x))kGδ/s ( s r )k = 2 sN +1 rN 1 r − s . ≤ 2 N Therefore k N Xk=0 p1 k( 1 r δ(T )) − k=0 p1 well-defined. Therefore both series P∞ Moreover, since PN k=0 p1 Xk=0 lim sup N→∞ have k N p2 k( 1 r δ(T ))k ≤ 2K Xk=0 k( 1 r δ(T )) converge to the same limit, so π(φ) is r − s (4.4) . 1 sN +1 rN k( 1 r δ(x)) converges uniformly to φ(x) on Gδ/s, we p1 k ◦ ( 1 r δ)kGδ/s ≤ kφkGδ/s ≤ kφkGδ/r. Therefore kπ(φ)kL(X) = lim N→∞ k N Xk=0 p1 k( 1 r δ(T ))k ≤ KkφkGδ/r. r δ) and ψ = g ◦ ( 1 The fact that π is a homomorphism follows from it being well defined, as if φ = f ◦ ( 1 r δ), then φψ = (f g) ◦ ( 1 r δ). Finally, to show that π extends the polynomial functional calculus, suppose q is a free polynomial in H ∞(Gδ/r), so q = f ◦ ( 1 r δ). Expand f (x) = P pk(x) into its homogeneous r δ(x)) converges uniformly to q(x) on Gδ/s, so since k=0 pk( 1 parts. Then PN Gδ/s is a K-spectral set for T , π(q) = lim N→∞ N Xk=0 pk( 1 r δ(T )) = q(T ). This last argument shows that π is the unique continuous extension of the evaluation map on polynomials. ✷ A similar result holds for complete K-spectral sets. 12 Theorem 4.5. The following are equivalent. (i) There exists s < 1 such that Gδ/s is a complete K-spectral set for T . (ii) There exists r < 1 such that the map π : F ◦ ( 1 r δ) 7→ F ♯( 1 r δ(T )) is a well-defined completely bounded homomorphism, satisfying kF ♯( 1 r δ(T ))k• ≤ K kF ◦ ( 1 r δ)kGδ/r that extends the polynomial functional calculus on Pd ∩ H ∞(Gδ/r). Moreover, if these conditions hold, then π is the unique extension of the evaluation homomorphism on the polynomials to a bounded homomorphism from H ∞(Gδ/r) to L(X). The proof is very similar to the proof of Theorem 5.2. The only significant difference is that (4.4) becomes k N Xk=0 P 1 k ( 1 r δ(T )) − N Xk=0 P 2 k ( 1 r δ(T ))k• ≤ 2K sN +1 rN 1 r − s . We apply Lemma 2.6 to conclude that both series converge to the same limit matrix. Definition 4.6. We shall say that T has a contractive (respectively, com- pletely contractive, bounded, completely bounded) Gδ functional calculus if there exists 0 < r < 1 such that Gδ/r is a spectral set (respectively, complete spectral set, K spectral set, complete K spectral set) for T . Remark 4.7 Even in the case d = 1, T ∈ L(H), and δ(x) = x, the ques- tion of when T has an H ∞(D) functional calculus becomes murky without the a priori requirement that kT k < 1. By von Neumann’s inequality [30], T will have a completely contractive Gδ functional calculus if kT k < 1. When kT k = 1, then p 7→ p(T ) extends contractively to H ∞(D) if T does not have a singular unitary summand [25, Thm. III.2.3], but to guarantee uniqueness, the standard extra assumption is continuity in the strong operator topology for functions that converge boundedly almost everywhere on the unit circle [25, Section III.2.2]. By Rota’s theorem [22], if σ(T ) ⊆ (D) then T is similar to an opera- tor which has a completely contractive H ∞(D) functional calculus. Again, 13 the situation becomes more delicate if σ(T ) is not required to lie in D. By Paulsen’s theorem [14], T will have a completely bounded polynomial func- tional calculus if and only if T is similar to a contraction. 5 Complete spectral sets If Φ ∈ H ∞(Gδ), and T is a d-tuple with kδ(T )k• < 1, one wants to define Φ♯(T ) as F ♯(δ(T )). But what if there are two different functions, F and F1, both in H ∞(BI×J ), and satisfying Φ(x) = F ◦ δ(x) = F1 ◦ δ(x) ∀ x ∈ Gδ. How does one know that F ♯(δ(T )) = F ♯ choice? 1(δ(T ))? If it doesn’t, is there a “best” We shall say Gδ is bounded if there exists M such that kxk ≤ M, ∀ x ∈ Gδ. This is the same as requiring that Pd ⊆ H ∞(Gδ). A stronger condition than this is to require that the algebra generated by the δij is all of Pd. Definition 5.1. We shall say that δ is separating if every coordinate function xr, 1 ≤ r ≤ d, is in the algebra generated by the functions {δij : 1 ≤ i ≤ I, 1 ≤ j ≤ J}. Theorem 5.2. Assume kδ(T )k• < 1. Then there exists r < 1 such that Gδ/r is a complete K-spectral set for T if and only if there exists s in the interval (kδ(T )k•, 1) such that whenever F is a matrix-valued H ∞(BI×J ) function, and P is a matrix of free polynomials satisfying then F ◦ ( 1 s δ)(x) = P (x) ∀ x ∈ G(1/s)δ, F ♯( 1 s δ(T )) = P (T ). (5.3) (5.4) If δ is separating, then it suffices to check the condition for the case P = 0. Proof: (⇒) By Theorem 4.5, we get (5.3) implies (5.4) whenever Gδ/r is a complete K-spectral set. 14 (⇐) Suppose kδ(T )k• = t < 1, and that s ∈ (t, 1) has the property that (5.3) implies (5.4). Let r = s; we will show that Gδ/r is a complete K-spectral set for T . Let P be a matrix of polynomials; we wish to show that kP (T )k• ≤ K sup{kP (x)k : x ∈ Md n, kδ(x)k < r}. (5.5) Without loss of generality, assume that the right-hand side of (5.5) is finite. By Theorem 3.1, we can find F , a matrix-valued function on H ∞(BI×J ), such that F ◦ ( 1 r δ) = P on Gδ/r and kF k ≤ sup{kP (x)k : x ∈ Md n, kδ(x)k < r}. By (5.4), we have P (T ) = F ♯( 1 r δ(T )), and so, by Theorem 3.14, (5.5) holds, with K = r r−t in general. Now, suppose that implies F ◦ ( 1 s δ) = 0 on G(1/s)δ F ♯( 1 s δ(T )) = 0. (5.6) (5.7) We wish to show that (5.3) implies (5.4). Since δ is separating, there is a matrix H of free polynomials such that H ◦ ( 1 s δ)(x) = P (x). (F − H) ◦ ( 1 s δ)(x) = 0 ∀ x ∈ Gδ/s, Then so by hypothesis F ♯( 1 s δ((T )) = H ♯( 1 s δ(T )), and since H is a polynomial, H ♯( 1 s δ(T )) = H( 1 s δ(T )) = P (T ), as required. ✷ 15 Remark 5.8 To just check the case P = 0, we don’t need to know that δ is separating, we just need to know that whenever a polynomial is bounded on Gδ/r, then it is expressible as a polynomial in the δij. Here is a checkable condition. Theorem 5.9. Suppose δ(0) = 0, and that T ∈ L(X)d has sup 0≤r≤1 kδ(rT )k• < 1. Then T has a completely bounded Gδ functional calculus. Proof: By Theorem 5.2, it is sufficient to prove that (5.3) implies (5.4). Assume (5.3) holds, i.e. F ◦ ( 1 s δ)(x) = ∞ Xk=0 Pk(( 1 s δ(x)) = P (x) ∀ x ∈ G(1/s)δ. By Theorem 3.2, F ◦ ( 1 s δ) − P has a power series expansion in a ball centered at 0 in M[d]. Since δ(0) = 0, for any m ∈ N, the number of terms in F ◦ ( 1 s δ)(x) − P (x) that are of degree m in x is finite. If one expands Pk(( 1 s δ(x)) one gets O((IJ)k) terms, so if then the series expansion for k 1 s δ(x)k < 1 IJ , ∞ Xk=0 Pk(( 1 s δ(x)) converges absolutely. We conclude therefore, by rearranging absolutely con- vergent series, that if R is any d-tuple in L(X) satisfying kδ(R)k• < s IJ , then ∞ Pk(( 1 s δ(R)) = P (R). Xk=0 (5.10) Since δ(0) = 0, we can apply (5.10) to ζT , for all sufficiently small ζ. Now we analytically continue to ζ = 1, to conclude that (5.10) also holds for T . ✷ 16 6 Hilbert spaces If the d-tuple is in L(H)d, it is natural to work with the Hilbert space tensor product and the Hilbert space norm, instead of the norm k · k•. Throughout this section, we will assume that S = (S1, . . . , Sd) is in L(H)d, and all norms (including those used to define spectral and complete spectral sets) will be Hilbert space norms. Many of our earlier results go through with essentially the same proofs, but, since we can use (3.18) instead of (3.17), we get better constants. A sample result, proved like Theorem 5.2, would be: Theorem 6.1. Let S ∈ L(H)d. Then there exists r < 1 such that kF ♯( 1 r δ(S))k ≤ sup{kF (( 1 r δ(x))k : x ∈ Gδ/r} if and only if (i) kδ(S)k < 1 and (ii) whenever F is a matrix-valued H ∞(BI×J ) function with F ◦ ( 1 s δ)(x) = P (x) ∀ x ∈ G(1/s)δ, then F ♯( 1 s δ(S)) = P (S). Example 8.4 shows that condition (i) does not imply (ii) in Theorem 6.1. For the remainder of this section, fix an orthonormal basis {en}∞ n=1 for H. Then we can naturally identify Mn with the operators on H that map ∨n k=1{ek} to itself, and are zero on the orthogonal complement. In this way, Gδ is a subset of G♯ δ, where G♯ δ := {S ∈ L(H)d : kδ(S)k < 1}. Since multiplication is sequentially continuous in the strong operator topology, to get a functional calculus is is enough to know that S ∈ G♯ δ is the strong operator topology limit of a sequence of d-tuples in Gδ/r. For any set A ∈ B(H)d, let us write sclSOT(A) to mean the set of tuples in B(H)d that are strong operator topology limits of sequences from A. 17 Theorem 6.2. Suppose S ∈ [0<r<1 sclSOT(G 1 r δ ). Then S has a completely contractive Gδ functional calculus. Proof: By hypothesis, there exists a sequence xk ∈ G(1/t)δ that con- verges to S in the strong operator topology, for some t < 1. Therefore δ(xk) converges to δ(S) S.O.T., so kδ(S)k = r ≤ t < 1. Let s ∈ (t, 1). By The- orem 6.1, it is sufficient to prove that (5.3) implies (5.4). As in the proof of Theorem 4.3, we can approximate F uniformly on t BI×J by a sequence s QN , the sum of the first N homogeneous polynomials. So for all ε > 0, there exists N0 such that if N ≥ N0 then kQN ( 1 s δ(xk)) − F ( 1 s δ(S)) − F ♯( 1 s δ(xk))k < ε s δ(S)k < ε. and kQN ( 1 As F ((1/s)δ) = P on G(1/s)δ, inequality (6.3) means ∀ N ≥ N0 kQN ( 1 s δ(xk)) − P (xk)k < ε. (6.3) (6.4) (6.5) Since multiplication is sequentially strong operator continuous, and QN is a matrix of polynomials, S.O.T. lim k→∞ [QN ( 1 s δ(xk)) − P (xk)] = QN ( 1 s δ(S)) − P (S). (6.6) The norm of a strong operator topology sequential limit is less than or equal to the limit of the norms, so by (6.5), we get from (6.6) that ∀ N ≥ N0 kQN ( 1 s δ(S)) − P (S)k ≤ ε. (6.7) Using (6.7) in (6.4), we conclude that kF ♯( 1 s δ(S) − P (S)k ≤ 2ε, Since ε was arbitrary, we conclude that (5.4) holds, i.e. F ♯( 1 ✷ s δ(S) = P (S). Corollary 6.8. Suppose each δij is the sum of a scalar and a homogeneous polynomial of degree 1. Then S has a completely contractive Gδ functional calculus if and only if kδ(S)k < 1. 18 Proof: Let ΠN be the projection from H onto ∨n j=1{ej}. Suppose kδ(S)k ≤ r. Let xN = ΠN SΠN . Then xN converges to S in the strong operator topology. Moreover, δ(xn) = ΠN ⊗ idCI δ(S) ΠN ⊗ idCJ , so kδ(xN )k ≤ kδ(S)k. ✷ For Hilbert spaces, replacing completely bounded by completely contrac- tive only changes things up to similarity. This follows from the following theorem of V. Paulsen [13]: Theorem 6.9. Let H and K be Hilbert spaces, and let A be a unital subal- gebra of L(K). Let ρ : A → L(H) be a completely bounded homomorphism. Then there exists an invertible operator a on H, with kakka−1k = kρkcb, such that a−1ρ(·)a is a completely contractive homomorphism. As a consequence, we get the following. Theorem 6.10. Let S be a d-tuple of operators on H. Then S has a com- pletely bounded Gδ functional calculus if and only if there exists an invertible operator a on H such that R = a−1Sa has a completely contractive Gδ func- tional calculus. Proof: Sufficiency is clear. For necessity, suppose 0 < r < 1, and the map H ∞(Gδ/r) ∈ Φ 7→ Φ(S) is a completely bounded map, with c.b. norm K, that extends polynomial evaluations for polynomials that are bounded on Gδ/r. Then in particular, Gδ/r is a complete K-spectral set for S. Let {xk}∞ k=1 be a countable dense set in Gδ/r, and let X = ⊕xk. Then for any matrix valued function P , we have kP kGδ/r = sup{kP (x)k : x ∈ Gδ/r} = kP (X)k. By hypothesis, the map ρ : P (X) 7→ P (S) is completely bounded, with kρkcb ≤ K. By Paulsen’s theorem 6.9, we have there exists a in L(H) such that P (X) 7→ P (a−1Sa) 19 is completely contractive. Therefore Gδ/r is a complete spectral set for a−1Sa. ✷ Remark: We don’t need K to be separable, so we could have taken X to be the direct sum over all of Gδ/r. Indeed, we could sum over all Gδ/r which are complete K spectral sets, and get one similarity that works for all of them. 7 Spectrum There are several plausible ways to define a spectrum for T ∈ L(X)d. Definition 7.1. Let σcc(T ) = {x ∈ M[d] : the map p(T ) 7→ p(x) is completely contractive} σb(T ) = {x ∈ M[d] : the map p(T ) 7→ p(x) is bounded}. By a theorem of R. Smith [24]; [15, Prop 8.11], every bounded map from an operator algebra into a finite dimensional algebra is completely bounded. So we conclude that if S ∈ L(H)d, then σb(S) is the set of all x that are similar to an element of σcc(S). Definition 7.2. Let ∆cc(T ) = {δ : T has a completely contractive Gδ functional calculus} ∆cb(T ) = {δ : T has a completely bounded Gδ functional calculus} Speccc(T ) = \δ∈∆cc(T ) Speccb(T ) = {x ∈ Md Gδ n : ∀ δ ∈ ∆cb(T ), ∃ y ∈ Gδ, with y similar to x}. Proposition 7.3. For every T ∈ L(X)d, we have σcc(T ) ⊆ Speccc(T ). (7.4) The set σcc(T ) is bounded, and for all n, we have σcc(T ) ∩ Md n is compact. If S ∈ L(H)d, then and Speccc(S) is bounded. σb(S) ⊆ Speccb(S), (7.5) 20 Proof: To prove (7.4), observe that if x ∈ σcc(T ), then kδ(x)k ≤ kδ(T )k•, so kδ(x)k < 1 whenever T has a completely contractive Gδ func- tional calculus. Boundedness follows since if x ∈ σcc(T ), then kxrk ≤ kT rk for every r, and continuity shows that the set is closed at each level. For (7.5), if x ∈ σb(S), then by Theorem 6.9, x is similar to some y in σcc(S). By (7.4), x is similar to an element of Speccc(S), so must lie in Speccb(S). To see that Speccc(S) is bounded, we can use Corollary 6.8 to see that if x ∈ Speccc(S), then kγ(x)k ≤ kγ(S)k for every matrix γ such that each term is of total degree at most one; in particular choosing γ(x) = (xr − λr)/kSr − λrk we conclude that kxr − λrk ≤ kSr − λrk ∀ r, λr. ✷ Question 7.6 When is (7.4) an equality? It is possible for σb(T ) to be empty - see Example 8.7. 8 Examples Example 8.1. Suppose δ(x) = (x1 . . . xd), a 1-by-d matrix. Then H ∞(Gδ) is the algebra of all bounded nc functions defined on the row contractions. Functions on the row contractions were studied by Popescu in [17]. Note that a function in H ∞(Gδ) need not have an absolutely convergent power series. When we expand f ∈ H ∞(Gδ) as in (3.7) or (3.8), we get ∞ f (x) = pk(x), Xk=0 where each pk is a homogeneous polynomial of degree k, having dk terms. Knowing merely that all the coefficients are bounded, one would need kxjk < 1 d for each j to conclude that the series converged absolutely. However we do know that ∞ kpk(x)k Xk=0 21 converges for all x in Gδ. By Theorem 5.9 or Theorem 3.14, if T ∈ L(X)d satisfies kδ(T )k• < 1, then the functional calculus F 7→ F ♯(T ) is a completely bounded homomorphism from H ∞(Gδ) to L(X), with com- pletely bounded norm at most 1 1 − kδ(T )k• . Any function in the multiplier algebra of the Drury-Arveson space can be ex- tended without increase of norm to a function in H ∞(Gδ) [2], so in particular one can then apply these functions to T . Example 8.2. This is a similar example to 8.1. This time, let δ be the d-by- d diagonal matrix with the coordinate functions written down the diagonal. Then H ∞(Gδ) will be the free analytic functions defined on d-tuples x with max kxjk < 1. Again, any function that is bounded on the commuting contractive d-tuples can be extended to all of Gδ without increasing its norm [2]. Let T = (T 1, . . . , T d) ∈ L(X)d. We can calculate kδ(T )k• by observing that one gets a Hilbert tensor norm on X ⊗ ℓ2 m if one defines k(x1, x2, . . . , xm)k = qX kxjk2 X . It follows that kδ(T )k• ≤ max kT jk, and since this is easily seen to be a lower bound, we conclude kδ(T )k• = max 1≤j≤d kT jkL(X). (8.3) So, one gets an H ∞(Gδ) functional calculus whenever (8.3) is less than 1. Let us reiterate that if f ∈ H ∞(Gδ) and we expand it in a power series, we have no guarantee that the series will converge absolutely whenever the norm of each T j is less than one; we need to group the terms as in (3.13). Example 8.4. Here is an example of a polynomial that has a different norm on Gδ and G♯ δ, and condition (i) in Theo- rem 6.1 does not imply (ii). δ. Consequently, sclSOT(Gδ) 6= G♯ 22 Let 0 < ε < 0.2. For ease of reading, we shall use (x, y) instead of (x1, x2) to denote coordinates. Let δ(x, y) =   1 ε (yx − I) 0 0 0 1 1+εx 0 0 0 1 1+ε y   . Let p(x) = xy − I. Claim: kpkGδ ≤ ε + 4ε2 kpkG♯ ≥ 1. δ (8.5) (8.6) Proof: Let x ∈ Gδ. Then kyk < 1 + ε, and since yx is bounded below 1+ε . By this, we mean by 1 − ε, we conclude that x is bounded below by 1−ε that for all vectors v, we have So x has an inverse z, and kxvk ≥ 1 − ε 1 + ε kvk. kzk ≤ 1 + ε 1 − ε . Let e = yx − I. Then kek < ε, and y = z + ez. Therefore so yielding (8.5). p(x) = xz + xez − I = xez, kp(x)k ≤ ε(1 + ε)2 1 − ε ≤ ε + 4ε2, For the second inequality, let T = (S, S ∗), where S is the unilateral shift. ✷ 1+ε < 1, and kp(S, S ∗)k = 1, yielding (8.6). Then kδ(S, S ∗)k = 1 Example 8.7. It is easy for σb(T ) to be empty. For example, suppose q(x) = x1x2 − x2x1 − I, and choose T ∈ L(H)2 so that kq(T )k = 1 2 . (This can be done, since by [6] any operator in L(H) that is not a non-zero scalar plus a compact is a commutator). Then for any x ∈ M[2], we have kq(x)k ≥ 1, so x /∈ σcc(T ). Consequently, σb(T ) is also empty. Note that in this example, Gq is empty, though T ∈ G♯ q. 23 Example 8.8. This is an example of our non-commutative approach applied to a single matrix. Let U = {z ∈ C : z < 1, and z − 1 < 1}. Let X be a finite dimensional Banach space, and T ∈ L(X) have σ(T ) ⊂ U. Let 0 δ(x) = (cid:18)x 0 x − 1(cid:19) . Then H ∞(Gδ) will be a space of analytic functions on U, but the norm will not be the sup-norm; it will be the larger norm given by kφk := sup{kφ(S)k : S ∈ L(H), kδ(S)k < 1}. Indeed, by Theorem 3.1, the norm can obtained as kφk = inf{kgkH∞(D2) g(z, z − 1) = φ(z) ∀ z ∈ U}. (It is sufficient to calculate the norm of g in the commutative case, since it always has an extension of the same norm to the non-commutative space, by [2]). By [1, Thm. 4.9], every function analytic on a neighborhood of U is in H ∞(Gδ). Since X is finite dimensional, T is similar to an operator on a Hilbert space, and by the results of Smith and Paulsen, this can be taken to have U as a complete spectral set. Putting all this together, we can write T as a−1Sa, where S is a Hilbert space operator with kδ(S)k < 1. For any φ in H ∞(Gδ), we find a g of minimal norm in H ∞(D2) such that g(z, z − 1) = φ(z) ∀ z ∈ U. Finally, we get the estimate kφ(T )kL(X) ≤ ka−1kkakkgkH∞(D2). If we know max(kT k, kT − 1k) = r < 1, we have the estimate (which works even if X is infinite dimensional) kφ(T )kL(X) ≤ 1 1 − r kgkH∞(D2). 24 References [1] J. Agler and J.E. McCarthy, Operator theory and the Oka extension theorem, Hi- roshima Math. J. To appear. ↑24 [2] [3] , Pick interpolation for free holomorphic functions. Amer. J. Math, to appear. ↑4, 22, 24 , Global holomorphic functions in several non-commuting variables, Canad. J. Math. 67 (2015), no. 2, 241–285. ↑3, 4, 6 [4] D. Alpay and D. S. Kalyuzhnyi-Verbovetzkii, Matrix-J -unitary non-commutative ra- tional formal power series, The state space method generalizations and applications, 2006, pp. 49–113. ↑3 [5] Joseph A. Ball, Gilbert Groenewald, and Tanit Malakorn, Conservative structured noncommutative multidimensional linear systems, The state space method general- izations and applications, 2006, pp. 179–223. ↑3 [6] Arlen Brown and Carl Pearcy, Structure of commutators of operators, Ann. of Math. (2) 82 (1965), 112–127. MR0178354 (31 #2612) ↑23 [7] R.E. Curto, Applications of several complex variables to multiparameter spectral the- ory, Surveys of some recent results in operator theory, 1988, pp. 25–90. ↑2 [8] Joe Diestel, Jan H. Fourie, and Johan Swart, The metric theory of tensor products, American Mathematical Society, Providence, RI, 2008. Grothendieck’s r´esum´e revis- ited. ↑5 [9] J. William Helton, Igor Klep, and Scott McCullough, Analytic mappings between noncommutative pencil balls, J. Math. Anal. Appl. 376 (2011), no. 2, 407–428. ↑3 [10] , Proper analytic free maps, J. Funct. Anal. 260 (2011), no. 5, 1476–1490. ↑3 [11] J. William Helton and Scott McCullough, Every convex free basic semi-algebraic set has an LMI representation, Ann. of Math. (2) 176 (2012), no. 2, 979–1013. ↑3 [12] Dmitry S. Kaliuzhnyi-Verbovetskyi and Victor Vinnikov, Foundations of free non- commutative function theory, AMS, Providence, 2014. ↑3 [13] Vern I. Paulsen, Completely bounded homomorphisms of operator algebras, Proc. Amer. Math. Soc. 92 (1984), no. 2, 225–228. MR754708 (85m:47049) ↑19 [14] V.I. Paulsen, Every completely polynomially bounded operator is similar to a contrac- tion, J. Funct. Anal. 55 (1984), 1–17. ↑14 [15] , Completely bounded maps and operator algebras, Cambridge University Press, Cambridge, 2002. ↑20 [16] Gilles Pisier, Introduction to operator space theory, London Mathematical Society Lecture Note Series, vol. 294, Cambridge University Press, Cambridge, 2003. ↑4 [17] Gelu Popescu, Free holomorphic functions on the unit ball of B(H)n, J. Funct. Anal. 241 (2006), no. 1, 268–333. ↑3, 21 25 [18] [19] [20] , Free holomorphic functions and interpolation, Math. Ann. 342 (2008), no. 1, 1–30. ↑3 , Free holomorphic automorphisms of the unit ball of B(H)n, J. Reine Angew. Math. 638 (2010), 119–168. ↑3 , Free biholomorphic classification of noncommutative domains, Int. Math. Res. Not. IMRN 4 (2011), 784–850. ↑3 [21] M. Putinar, Uniqueness of Taylor’s functional calculus, Proc. Amer. Math. Soc. 89 (1983), 647–650. ↑2 [22] Gian-Carlo Rota, On models for linear operators, Comm. Pure Appl. Math. 13 (1960), 469–472. ↑13 [23] Raymond A. Ryan, Introduction to tensor products of Banach spaces, Springer Mono- graphs in Mathematics, Springer-Verlag London, Ltd., London, 2002. ↑5 [24] R.R. Smith, Completely bounded maps between C*-algebras, J. Lond. Math. Soc. 27 (1983), 157–166. ↑20 [25] B. Szokefalvi-Nagy and C. Foia¸s, Harmonic analysis of operators on Hilbert space, second edition, Springer, New York, 2010. ↑13 [26] J.L. Taylor, The analytic functional calculus for several commuting operators, Acta Math. 125 (1970), 1–38. ↑2 [27] [28] [29] , A joint spectrum for several commuting operators, J. Funct. Anal. 6 (1970), 172–191. ↑2 , A general framework for a multi-operator functional calculus, Advances in Math. 9 (1972), 183–252. MR0328625 (48 #6967) ↑2 , Functions of several non-commuting variables, Bull. Amer. Math. Soc. 79 (1973), 1–34. ↑1, 2 [30] J. von Neumann, Eine Spektraltheorie fur allgemeine Operatoren eines unitaren Raumes, Math. Nachr. 4 (1951), 258–281. ↑13 26
1006.2995
2
1006
2010-09-28T16:18:25
Noncompactness and noncompleteness in isometries of Lipschitz spaces
[ "math.FA" ]
We solve the following three questions concerning surjective linear isometries between spaces of Lipschitz functions $\mathrm{Lip}(X,E)$ and $\mathrm{Lip}(Y,F)$, for strictly convex normed spaces $E$ and $F$ and metric spaces $X$ and $Y$: \begin{enumerate} \item Characterize those base spaces $X$ and $Y$ for which all isometries are weighted composition maps. \item Give a condition independent of base spaces under which all isometries are weighted composition maps. \item Provide the general form of an isometry, both when it is a weighted composition map and when it is not. \end{enumerate} In particular, we prove that requirements of completeness on $X$ and $Y$ are not necessary when $E$ and $F$ are not complete, which is in sharp contrast with results known in the scalar context.
math.FA
math
NONCOMPACTNESS AND NONCOMPLETENESS IN ISOMETRIES OF LIPSCHITZ SPACES JES ´US ARAUJO AND LUIS DUBARBIE Abstract. We solve the following three questions concerning sur- jective linear isometries between spaces of Lipschitz functions Lip(X, E) and Lip(Y, F ), for strictly convex normed spaces E and F and met- ric spaces X and Y : (i) Characterize those base spaces X and Y for which all isome- tries are weighted composition maps. (ii) Give a condition independent of base spaces under which all isometries are weighted composition maps. (iii) Provide the general form of an isometry, both when it is a weighted composition map and when it is not. In particular, we prove that requirements of completeness on X and Y are not necessary when E and F are not complete, which is in sharp contrast with results known in the scalar context. 1. Introduction It is well known that not all surjective (linear) isometries between spaces of Lipschitz functions on general metric spaces X and Y can be written as weighted composition maps (see for instance [22, p. 61]). Attempts to identify the isometries which can be described in that way have been done in three ways, each trying to provide an answer to one of the following questions: (i) Characterize those base spaces X and Y for which all isome- tries are weighted composition maps. (ii) Give a condition independent of base spaces under which all isometries are weighted composition maps. (iii) Provide the general form of an isometry, both when it is a weighted composition map and when it is not. 2010 Mathematics Subject Classification. Primary 47B33; Secondary 46B04, 46E15, 46E40, 47B38. Key words and phrases. Linear isometry, Banach-Stone theorem, vector-valued Lipschitz function, biseparating map. Research partially supported by the Spanish Ministry of Science and Education (MTM2006-14786). The second author was partially supported by a predoctoral grant from the University of Cantabria and the Government of Cantabria. 1 2 JES ´US ARAUJO AND LUIS DUBARBIE The first question was studied by Weaver for a general metric in the scalar-valued setting (see [21] or [22, Section 2.6]), and the second one has been recently treated by Jim´enez-Vargas and Villegas-Vallecillos in the more general setting of vector-valued functions (see [14]). In the latter, the Banach spaces where the functions take values are assumed to be strictly convex. This is certainly not a heavy restriction, as this type of results is known not to hold for general Banach spaces. Strict convexity is actually a very common and reasonable assumption, even if, at least in other contexts, it is not the unique possible (see for instance [2, 4, 6, 12]). As for the third question, an answer was given by Mayer-Wolf for compact base spaces in the scalar context, not for a general metric d, but for powers dα with 0 < α < 1. Weaver proved that completeness and 1-connectedness of X and Y are sufficient conditions, and that the weighted composition isometries must have a very special form. More concretely, given complete 1- connected metric spaces X and Y with diameter at most 2, a linear bijection T : Lip(X) → Lip(Y ) is an isometry if and only if T f = α · f ◦ h for every f , where α ∈ K, α = 1, and h : Y → X is an isometry. Requirements of 1-connectedness on both X and Y (that is, they cannot be decomposed into two nonempty disjoint sets whose distance is greater than or equal to 1) cannot be dropped in general. of completeness cannot be dropped either. And, obviously, Lip(X) and Lip(bX) are linearly isometric when X is not complete (where bX denotes the completion of X), so requirements On the other hand, Jim´enez-Vargas and Villegas-Vallecillos gave a general representation in the spirit of the classical Banach-Stone Theo- rem (along with related results for isometries not necessarily surjective). Assumptions include compactness of base metric spaces and the fact that the isometry fixes a (nonzero) constant function. The conclusion in the surjective case is that the isometry T : Lip(X, E) → Lip(Y, E) is of the form T f (y) = Jy(f (h(y))) for all f ∈ Lip(X, E) and y ∈ Y , where h is a bi-Lipschitz homeomorphism (that is, h and h−1 are Lip- schitz) from Y onto X and J is a Lipschitz map from Y into the set I(E, E) of all surjective linear isometries on the strictly convex Banach space E. They also proved that this result can be sharpened under stronger hypotheses, but the above assumptions remain basically the same, so that the results do not provide an "if and only if" description. Finally, Mayer-Wolf not only characterized the family of compact spaces for which the associated Lipschitz spaces admitted isometries that were not composition operators, but also gave their general form. In principle, it is not clear whether or not his results can be extended to NONCOMPACTNESS AND NONCOMPLETENESS 3 spaces endowed with a metric not of the form dα. In fact, the answer, as we will see here, is not completely positive. The aim of this paper is to give, in the vector-valued setting, a com- plete answer to questions (i),(ii) and (iii) (just assuming strict convexity of E and F ). The general answer is not known even in the scalar set- ting, which can be included here as a special case. We also prove, on the one hand, that conditions of compactness can be replaced with just completeness on base spaces and, on the other hand, that even com- pleteness can be dropped when the normed spaces E and F are not complete (which is in sharp contrast with the behaviour in the scalar case). To solve (ii), we show that the condition on the preservation of a constant function (as given in [14]) can be replaced with a milder one (see Theorem 3.1). We use it to solve (iii) (see Theorem 3.4, and more in general Theorem 3.1, Corollary 3.3 and Remark 3.5). Our answer also applies to results on metrics dα in [18], and a key to understand the generalization is Proposition 3.9. An answer to (i) is given as a direct consequence of the results concerning (ii) and (iii) (see Corollaries 3.6 and 3.7). As a special case we provide the natural counterpart of the description given in [21] (see Corollary 5.3 and Remark 5.4). We finally mention that we do not use the same techniques as in [21] nor as in [14]; instead we study surjective linear isometries through biseparating maps, which has proven successful in various contexts (see for instance [2, 10] for recent references). Other papers where related operators have been recently studied in similar contexts are [1, 8, 9, 13, 17] (see also [5, 11, 16, 18, 19, 20]). 2. Preliminaries and notation Recall that, given metric spaces (X, d1) and (Y, d2), a map f : X → Y is said to be Lipschitz if there exists a constant k ≥ 0 such that d2(f (x), f (y)) ≤ k d1(x, y) for each x, y ∈ X, and that the Lipschitz number of f is L(f ) := sup(cid:26) d2(f (x), f (y)) d1(x, y) : x, y ∈ X, x 6= y(cid:27) . Given a normed space E (over K = R or C), we denote by Lip(X, E) the space of all bounded E-valued Lipschitz functions on X. We en- dow Lip(X, E) with the norm k·kL := max {k·k∞ , L(·)} (where k·k∞ denotes the usual supremum norm). As a particular case, we can consider in X a power d1 α of the metric d1, 0 < α < 1. The corresponding space of all bounded E-valued Lips- α is then denoted by Lipα(X, E). chitz functions on X with respect to d1 4 JES ´US ARAUJO AND LUIS DUBARBIE Recall also that a normed space E is said to be strictly convex if ke1 + e2k < 2 whenever e1, e2 are different vectors of norm 1 in E or, equivalently, that ke1 + e2k = ke1k + ke2k (e1, e2 6= 0) implies e1 = αe2 for some α > 0 (see [15, pp. 332 -- 336]). From this, it follows that, given e1, e2 ∈ E \ {0}, (2.1) ke1k , ke2k < max{ke1 + e2k , ke1 − e2k}, which is an inequality we will often use. The fact that a normed space is strictly convex does not imply that its completion is. Indeed every infinite-dimensional separable Banach space can be renormed to be not stricly convex and to contain a strictly convex dense subspace of codimension one (see [7]). From now on, unless otherwise stated, we assume that E and F are strictly convex normed spaces (including the cases E = K, F = K). As we mentioned above, on our way to Theorem 3.1 we will deal with biseparating maps. Recall that separating maps are those preserving disjointness of cozero sets (where the cozero set of a function f : X → E is defined as c(f ) := {x ∈ X : f (x) 6= 0}). More concretely, we will say that a linear map T : Lip(X, E) → Lip(Y, F ) is separating if c(T f ) ∩ c(T g) = ∅ whenever f, g ∈ Lip(X, E) satisfy c(f ) ∩ c(g) = ∅. Moreover, T is said to be biseparating if it is bijective and both T and its inverse are separating maps. Obviously, if f : X → E is Lipschitz and bounded, then so is the map kf k : X → R defined by kf k (x) := kf (x)k for every x ∈ X. It is also clear that kf k can be continuously extended to a Lipschitz E is complete and X is not, f admits a continuous extension to the complete, every surjective linear isometry T : Lip(X, E) → Lip(Y, F ) function dkf k : bX → R defined on the completion bX of X. More in general, if x ∈ bX \ X, we say that f admits an extension to x if it can be continuously extended to a map bf : X ∪ {x} → E. Clearly, when whole bX, and the extension bf : bX → E is a Lipschitz function with (cid:13)(cid:13)(cid:13)bf(cid:13)(cid:13)(cid:13)∞ = kf k∞ and L(cid:16)bf(cid:17) = L (f ). For this reason, when E and F are can be associated in a canonical way to another one bT : Lip(cid:16)bX, E(cid:17) → Lip(cid:16)bY , F(cid:17) (which coincides with T only if X and Y are complete). Given R > 0, we define in X the following equivalence relation: we put x ∼R y if there exist x1, . . . , xn ∈ X with x = x1, y = xn, and d(xi, xi+1) < R for i = 1, . . . , n − 1. We call R-component each of the equivalence classes of X by ∼R. The set of all R-components in X is denoted by CompR(X). NONCOMPACTNESS AND NONCOMPLETENESS 5 We say that a bijective map h : Y → X preserves distances less than 2 if d1(h(y), h(y′)) = d2(y, y′) whenever d2(y, y′) < 2. We denote by iso<2(Y, X) the set of all maps h : Y → X such that both h and h−1 preserve distances less than 2. Notice that every h ∈ iso<2(Y, X) is a homeomorphism and that, when X is bounded, then it is also a Lipschitz map (see also Remark 3.2). Definition 2.1. Let I(E, F ) be the set of all linear isometries from E onto F . We say that a map T : Lip(X, E) → Lip(Y, F ) is a standard isometry if there exist h ∈ iso<2(Y, X) and a map J : Y → I(E, F ) constant on each 2-component of Y such that for all f ∈ Lip(X, E) and y ∈ Y . T f (y) = Jy(f (h(y))) Remark 2.2. Notice that a standard isometry is indeed a surjective linear isometry. Theorem 3.1 gives a condition under which both classes of operators coincide. Also, when Y is 2-connected, the map J is constant, so there exists a surjective linear isometry J : E → F such that T f (y) = J(f (h(y))) for all f ∈ Lip(X, E) and y ∈ Y . In particular, Corollary 5.3 roughly says that this is the only way to obtain an isometry when one of the base spaces is 1-connected (see also Remark 5.4). In the definition of standard isometry, we see that X and Y are very much related. In particular, one is complete if and only if the other is. There are interesting cases which are almost standard in some sense. For instance, when E is complete, the natural inclusion On the other hand, when E and F are complete, every T can be we immediately obtain a standard isometry from it in a natural way. written as T = iY between this kind of isometries and nonstandard isometries. iX : Lip(X, E) → Lip(cid:16)bX, E(cid:17) is not standard if X is not complete, but −1 ◦bT ◦ iX. In the following definition, we distinguish Lip(Y, F ) is nonstandard if T and bT (if it can be defined, that is, if E Definition 2.3. We say that a surjective linear isometry T : Lip(X, E) → and F are complete) are not standard. Just a special family of spaces allows defining properly nonstandard isometries. We call them spaces of type A. Definition 2.4. We say that a metric space X is of type A if there are a partition of X into two subsets A, B, and a map ϕ : A → B with the following properties: 6 JES ´US ARAUJO AND LUIS DUBARBIE (i) d(x, z) = 1 + d(ϕ(x), z) whenever x ∈ A and z ∈ B satisfy d(ϕ(x), z) < 1, (ii) d(x, z) ≥ 2 whenever x ∈ A and z ∈ B satisfy d(ϕ(x), z) ≥ 1, (iii) d(x1, x2) ≥ 2 whenever x1, x2 ∈ A and ϕ(x1) 6= ϕ(x2). For each E, the operator Sϕ : Lip(X, E) → Lip(X, E) defined, for each f ∈ Lip(X, E), by Sϕf (x) :=(cid:26) f (x) f (ϕ(x)) − f (x) if x ∈ B if x ∈ A is said to be the purely nonstandard map associated to ϕ. Remark 2.5. It is easy to check that Sϕ is linear and bijective. Also −1 = Sϕ, kSϕ(f )k ≤ 1 whenever kf k ≤ 1. Taking into account that Sϕ this implies that Sϕ is indeed a nonstandard isometry. Theorem 3.4 and Remark 3.5 basically say that every nonstandard isometry is the composition of a standard and a purely nonstandard one. Throughout, for each e ∈ E, the constant function from X into E taking the value e will be denoted by e. Also, given a set A, χA stands for the characteristic function on A. As usual, if there is no confusion both the metric of X and that of Y will be denoted by d. Given a surjective linear isometry T : Lip(X, E) → Lip(Y, F ), we denote and A(T ) := {y ∈ Y : T e(y) = 0 ∀e ∈ E} B(T ) := A(T )c. The partition of Y into these two subsets will be very much used in Sections 5 and 6, and the fact that A(T ) is empty will turn out to be basically equivalent to T being standard. This property will receive a special name. We define Property P as follows: P: For each y ∈ Y , there exists e ∈ E with T e(y) 6= 0. 3. Main results We first give some results ensuring that an isometry is standard, and then characterize spaces and describe the isometries when this is not the case. Theorem 3.1 and Corollary 3.3 are proved in Section 4, and Theorem 3.4 in Section 6. It is obvious that, by definition, if T is not nonstandard, then it satisfies Property P. The converse is given by Theorem 3.1 and Corol- lary 3.3. NONCOMPACTNESS AND NONCOMPLETENESS 7 Theorem 3.1. Let T : Lip(X, E) → Lip(Y, F ) be a surjective linear isometry satisfying Property P. Then E and F are linearly isometric. Furthermore, if we are in any of the following two cases: (i) X and Y are complete, (ii) E (or F ) is not complete, then T is standard. Remark 3.2. In Theorem 3.1, we cannot in general ensure that the map h is an isometry or that it preserves distances equal to 2. Indeed, following the same ideas as in [22, Proposition 1.7.1], if (Z, d) is a metric space with diameter diam(Z, d) > 2, then there is a new metric d′(·, ·) := min {2, d(·, ·)} on Z with diam(Z, d′) = 2 such that Lip(Z, E) with respect to d and Lip(Z, E) with respect to d′ are linearly isometric. On the other hand, notice also that, if d′ 2 are defined in a similar way, then the map h : (Y, d2) → (X, d1) belongs to iso<2(Y, X) if and only if h : (Y, d′ 1 and d′ 1) is an isometry. 2) → (X, d′ In Theorem 3.1, when (i) and (ii) do not hold, E and F are complete and X (or Y ) is not. In this case, it is easy to see that in general T is not standard. Nevertheless, we have the following result. Corollary 3.3. Suppose that E and F are complete and X or Y is not. If T : Lip(X, E) → Lip(Y, F ) is a surjective linear isometry satisfying Property P, then bT : Lip(cid:16)bX, E(cid:17) → Lip(cid:16)bY , F(cid:17) is standard. We next give the general form that a nonstandard isometry (or, equivalently, an isometry not satisfying Property P) must take. Theorem 3.4. Assume that we are in any of the following two cases: (i) X and Y are complete, (ii) E (or F ) is not complete, Then there exists a nonstandard isometry T : Lip(X, E) → Lip(Y, F ) if and only if the following three conditions hold simultaneously: (i) X and Y are of type A, (ii) there exists h ∈ iso<2(Y, X), (iii) E and F are linearly isometric. In this case, T = Sϕ◦T ′, where T ′ : Lip(X, E) → Lip(Y, F ) is a stan- dard isometry and Sϕ : Lip(Y, F ) → Lip(Y, F ) is a purely nonstandard isometry. Remark 3.5. In the case when E and F are complete and X or Y is not, if T : Lip(X, E) → Lip(Y, F ) is a nonstandard isometry, then In so is bT , and the description given in Theorem 3.4 applies to bT . 8 JES ´US ARAUJO AND LUIS DUBARBIE particular, T = iY −1 ◦ Sϕ ◦ T ′ ◦ iX where T ′ : Lip(cid:16)bX, E(cid:17) → Lip(cid:16)bY , F(cid:17) is a standard isometry and Sϕ : Lip(cid:16)bY , F(cid:17) → Lip(cid:16)bY , F(cid:17) is purely nonstandard. A direct consequence (and easy to check) of Theorem 3.4 and Re- mark 3.5 is the following. Corollary 3.6. If E is not complete, then there exists a nonstandard isometry from Lip(X, E) onto itself if and only if X is of type A. If E is complete, then there exists a nonstandard isometry from Lip(X, E) onto itself if and only if bX is of type A. Theorem 3.4 says that, under some assumptions, when two spaces of Lipschitz functions are linearly isometric, there exists in fact a standard isometry between them. The following result is a simple consequence of Theorems 3.1 and 3.4, Corollary 3.3 and Remark 3.5. Corollary 3.7. Lip(X, E) and Lip(Y, F ) are linearly isometric if and only if E and F are linearly isometric and • iso<2(Y, X) is nonempty (when E and F are not complete). • iso<2(cid:16)bY , bX(cid:17) is nonempty (when E and F are complete). We finally adapt the above results to the special case of metrics dα, 0 < α < 1. Even if in this case we just deal with metrics and, conse- quently, the general form of the isometries between spaces Lipα(X, E) is completely given by Theorems 3.1 and 3.4, Corollary 3.3 and Re- mark 3.5, it is interesting to see how the condition of being of type A can be translated to metrics dα. This turns out to be more restrictive, and constitutes a generalization of the scalar case on compact spaces given in [18, Theorem 3.3]. Definition 3.8. Let 0 < α < 1. We say that a metric space (X, d) is of type Aα if there are a partition of X into two subsets A, B, and a map ϕ : A → B with the following properties: (i) d(x, ϕ(x)) = 1 for every x ∈ A, (ii) dα(x, z) ≥ 2 whenever x ∈ A and z ∈ B, z 6= ϕ(x), (iii) dα(x1, x2) ≥ 2 whenever x1, x2 ∈ A and ϕ(x1) 6= ϕ(x2). Proposition 3.9. Let 0 < α < 1 and let (X, d) be a metric space. The following two statements are equivalent: (i) (X, dα) is of type A, (ii) (X, d) is of type Aα. NONCOMPACTNESS AND NONCOMPLETENESS 9 Proposition 3.9 will be proved in Section 6. It is clear that, since X is of type Aα if and only if its completion is, the statement of Corollary 3.6 is even simpler when dealing with Lipα(X, E). Also, it is immediate to see that, if (X, d) is of type Aα, then it is of type Aβ for α < β ≤ 1. Consequently, by Theorem 3.4 and Remark 3.5, we conclude the following. Corollary 3.10. Let 0 < α < 1. If there exists a nonstandard isometry between Lipα(X, E) and Lipα(Y, F ), then there exists a nonstandard isometry between Lipβ(X, E) and Lipβ(Y, F ) whenever α < β ≤ 1. Obviously, the converse of Corollary 3.10 is not true in general. The following example shows somehow the differences between cases. Example 3.11. Let X := {−1} ∪ (0, 1) ⊂ R. X is not of type A, but for 0 < α < 1. Consequently, we have its completion bX = {−1} ∪ [0, 1] is. Neither X nor bX are of type Aα • If E is not complete, then all linear isometries from Lip(X, E) onto itself are standard, but there are nonstandard isometries • If E is complete, then there are nonstandard isometries from Lip(X, E) onto itself. Obviously, by definition of nonstandard from Lip(cid:16)bX, E(cid:17) onto itself. isometry, the same holds for Lip(cid:16)bX, E(cid:17). onto itself are standard. The same holds for Lipα(cid:16)bX, E(cid:17). In the case when E is complete and X is not, this is due to the special form of X. • For every E and α ∈ (0, 1), all linear isometries from Lipα(X, E) 4. The case when T satisfies Property P In this section, unless otherwise stated, we assume that T is a linear isometry from Lip(X, E) onto Lip(Y, F ) satisfying Property P. Our first goal consists of showing that T is indeed an isometry with respect to the norm k·k∞. The following two lemmas will be the key tools used to prove it. Lemma 4.1. Let f ∈ Lip(X, E) and x0 ∈ X be such that f (x0) 6= 0. Then there exists g ∈ Lip(X, E) with kg(x0)k = kgk∞ > L(g) such that kg(x0)k + kf (x0)k = k(g + f )(x0)k = kg + f k∞ > L(g + f ). Proof. We put e := f (x0) and assume without loss of generality that kek = 1. We then consider l ∈ Lip(X, E) defined by l(x) := max{0, 2 − 10 JES ´US ARAUJO AND LUIS DUBARBIE d(x, x0)} · e for each x ∈ X. Clearly l satisfies klkL = klk∞ = kl(x0)k = 2 and L(l) ≤ 1, and also kl(x)k < 2 for all x ∈ X, x 6= x0. Take n ∈ N with n > kf kL. Firstly, it is easy to check that k(nl + f )(x0)k = 2n + 1. On the other hand, we also have that L(nl + f ) ≤ nL(l) + L(f ) < 2n and, for x ∈ X \ {x0} with d(x, x0) < 2, k(nl + f )(x)k ≤ n kl(x)k + kf (x)k ≤ 2n − nd(x, x0) + kf (x0)k + L(f )d(x, x0) < 2n + 1, whereas if d(x, x0) ≥ 2, then l(x) = 0 and k(nl + f )(x)k ≤ kf kL < n. (cid:3) Consequently, if we define g := nl, the lemma is proved. Remark 4.2. It is easy to see that if f1, f2 ∈ Lip(X, E) satisfy kf1(x0)k , kf2(x0)k < kf (x0)k, then the proof of Lemma 4.1 can be slightly modified (by tak- ing n > kf kL , kf1kL , kf2kL) so that kg + fik∞ < kg + f k∞ for i = 1, 2. Lemma 4.3. If f ∈ Lip(X, E) satisfies kT f (y0)k = kT f k∞ > L(T f ) for some y0 ∈ Y , then L(f ) ≤ kf k∞. Proof. Suppose that kf k∞ < L(f ). Then, for each e ∈ E, there exists M > 0 such that kf k∞ + M kek < L(f ), so kf ± Mek∞ < L(f ) = L(f ± Me). Therefore, kf ± MekL = L(f ± Me) = L(f ) = kf kL . Since T is an isometry, kT f ± MT ekL = kT f kL = kT f (y0)k, which implies in particular that kT f (y0) ± MT e(y0)k ≤ kT f (y0)k and, by Inequality (2.1), that T e(y0) = 0 for every e ∈ E, which goes against our hypotheses. (cid:3) Remark 4.4. Notice that, in the proof of Lemma 4.3, we just use the fact that there exists e ∈ E with T e(y0) 6= 0, and not the general assumption that Property P holds. Corollary 4.5. T is an isometry with respect to the supremum norm. Proof. Assume that kf k∞ < kT f k∞, and pick ǫ > 0 and y0 ∈ Y such that kf k∞ + ǫ < kT f (y0)k. Next, by Lemma 4.1, we can take g ∈ Lip(Y, F ) with kg(y0)k = kgk∞ > L(g) and such that kg(y0)k + kT f (y0)k = k(g + T f )(y0)k = kg + T f k∞ > L(g + T f ). Applying Lemma 4.3, we conclude both that L(T −1g) ≤(cid:13)(cid:13)T −1g(cid:13)(cid:13)∞ = kg(y0)k and that L(T −1g + f ) ≤(cid:13)(cid:13)T −1g + f(cid:13)(cid:13)∞ = kg + T f kL = k(g + T f )(y0)k . NONCOMPACTNESS AND NONCOMPLETENESS 11 But this is impossible because We conclude that kT f k∞ ≤ kf k∞ for every f . We next see that T −1 also satisfies Property P, which is enough to prove the equality. By = kfk. Also, (cid:13)(cid:13)T −1g + f(cid:13)(cid:13)∞ < kg(y0)k + kT f (y0)k − ǫ < k(g + T f )(y0)k . the above, given a nonzero f ∈ F we have (cid:13)(cid:13)(cid:13)T −1f(cid:13)(cid:13)(cid:13)∞ if(cid:16)T −1f(cid:17) (x0) = 0 for some x0 ∈ X, then there exists k ∈ Lip(X, E), k 6= 0, with kk(x)k +(cid:13)(cid:13)(cid:13)(cid:16)T −1f(cid:17) (x)(cid:13)(cid:13)(cid:13) ≤ kfk for every x ∈ X. By Inequal- ity (2.1), kfk <(cid:13)(cid:13)(cid:13)f + T k(cid:13)(cid:13)(cid:13)∞ or kfk <(cid:13)(cid:13)(cid:13)f − T k(cid:13)(cid:13)(cid:13)∞ Remark 4.6. Notice that, in the proof of Corollary 4.5, we have seen that T −1 also satisfies Property P. paragraph above. , which contradicts the (cid:3) We are now ready to see that, under the assumptions we make in this section, every surjective linear isometry is biseparating. Proposition 4.7. T is biseparating. Proof. We prove that T is separating. Suppose that it is not, so there exist f, g ∈ Lip(X, E) such that c(f )∩c(g) = ∅ but T f (y0) = f1 6= 0 and T g(y0) = f2 6= 0 for some y0 ∈ Y . Taking into account Inequality (2.1), we can assume without loss of generality that kf2k ≤ kf1k < kf1 + f2k. Now, by Lemma 4.1 and Remark 4.2, there exists k ∈ Lip(Y, F ) such that kk + T f k∞ , kk + T gk∞ < kk + T f + T gk∞. On the other hand, since f and g have disjoint cozeros, (cid:13)(cid:13)T −1k + f + g(cid:13)(cid:13)∞ = max(cid:8)(cid:13)(cid:13)T −1k + f(cid:13)(cid:13)∞ ,(cid:13)(cid:13)T −1k + g(cid:13)(cid:13)∞(cid:9) , and consequently kk + T f + T gk∞ = max {kk + T f k∞ , kk + T gk∞}, which is a contradiction. By Remark 4.6, T −1 is also separating. (cid:3) Remark 4.8. In [3, Theorem 3.1] (see also comments after it) a de- scription of biseparating maps S : Lip(X, E) → Lip(Y, F ) is given, but we cannot use it here because assumptions of completeness on X and Y are made in [3]. Under some circumstances automatic continuity of such S can be achieved and, in that case, the description goes as follows (where L(E, F ) denotes the normed space of all linear and continuous operators from E to F ): There exist a homeomorphism k : Y → X and a map K : Y → L(E, F ) (which is easily seen to be also Lipschitz with L(K) ≤ kSk) such that Sf (y) = Ky(f (k(y))) for all f ∈ Lip(X, E) and y ∈ Y . Also, if both X and Y are bounded, then the map k is bi-Lipschitz. 12 JES ´US ARAUJO AND LUIS DUBARBIE Proposition 4.9. Given e ∈ E, T e is constant on each 1-component of Y and kT e(y)k = kek for all y ∈ Y . Proof. Suppose that this is not the case, but there exist e ∈ E, kek = 1, and y1, y2 ∈ Y with y1 ∼1 y2, such that f1 := T e(y1) and f2 := T e(y2) are different. Of course, we may assume without loss of generality that D := d(y1, y2) < 1 and that f1 6= 0. Now, if we consider g ∈ Lip(Y, F ) defined by g(y) := max(cid:26)0, 1 − d(y, y1) D (cid:27) · f1 for all y ∈ Y , then obviously kgkL = L(g) = kf1k /D > kgk∞. As a consequence, using Corollary 4.5, kT −1gkL > kT −1gk∞, and we can take M > 0 such that kT −1g ± MekL = kT −1gkL = kf1k /D. Notice also that, as F is strictly convex, either kf1 + M (f1 − f2)k > kf1k or kf1 − M (f1 − f2)k > kf1k, that is, either k(g + MT e) (y1) − (g + MT e) (y2)k d(y1, y2) or k(g − MT e) (y1) − (g − MT e) (y2)k d(y1, y2) > kf1k D > kf1k D , which implies that either kg + MT ekL > kf1k /D or kg − MT ekL > kf1k /D, yielding a contradiction. Finally suppose that T e(y) = f for all y in B ∈ Comp1(Y ). Since by Proposition 4.7 T −1 is separating, c (T −1 (χB · T e))∩c(cid:0)T −1(cid:0)χY \B · T e(cid:1)(cid:1) = ∅. This implies that e is the only nonzero value taken by T −1 (χB · T e) and, since T −1 is an isometry with respect to k·k∞, we have that kek = kfk. (cid:3) Lemma 4.10. There exists a bijection H : Comp1(X) → Comp1(Y ) and, for each A ∈ Comp1(X), a surjective linear isometry JA : E → F with the property that T (χA · e) = χH(A) ·^ JA(e) = χH(A) · T e for every e ∈ E. Proof. Fix A ∈ Comp1(X) and e ∈ E with kek = 1, and take f := χA · e, g := χX\A · e in Lip(X, E). We have that c(f ) ∩ c(g) = ∅, so by Proposition 4.7 T f and T g have disjoint cozeros. Then, by Proposition 4.9, T f (y), T g(y) ∈ {0, T e(y)} for all y ∈ Y . Now, suppose that y ∼1 y′, and that T f (y) 6= 0 and T f (y′) = 0. We can assume that d(y, y′) < 1. Since kT f (y)k = 1, we deduce L(T f ) ≥ 1/d(y, y′) > 1, which is impossible because kf kL = 1. Reasoning similarly with T −1, T f = χB · f for some 1-component B in Y and some norm-one vector f ∈ F . The conclusion is now easy. (cid:3) NONCOMPACTNESS AND NONCOMPLETENESS 13 Lemma 4.11. Given A, B ∈ Comp1(X), if min {d(A, B), d(H(A), H(B))} < 2, then d(A, B) = d(H(A), H(B)) and JA = JB. Proof. Put D1 := d(A, B), D2 := d(H(A), H(B)). Due to the sym- metric roles of H and H −1 with respect to T and T −1, we can assume without loss of generality that D1 ≤ D2. Pick e ∈ E with kek = 1, and define f := (χA − χB) · e ∈ Lip(X, E). We easily see that kf kL = L(f ) = 2/D1 and, since L(T f ) = kJA(e) + JB(e)k /D2 ≤ 2/D2, we necessarily have D1 = D2 and kJA(e) + JB(e)k = 2, so JA(e) = JB(e) because F is strictly convex. (cid:3) Corollary 4.12. There exists a map J : Y → I(E, F ) which is con- stant on each 2-component of Y and such that T e(y) = Jy(e) for all e ∈ E and y ∈ Y . Proof. We define Jy := JA if y ∈ H(A) and A ∈ Comp1(X). Applying Lemma 4.11, the result follows. (cid:3) Lemma 4.13. Let (yn) be a sequence in Y which is not a Cauchy sequence and such that all yn are pairwise different. Then there exist infinite subsets A1 and A2 of {yn : n ∈ N} with d (A1, A2) > 0. Proof. Taking a subsequence if necessary, we have that there exists ǫ > 0 such that d(y2n, y2n−1) ≥ 3ǫ for all n ∈ N. Let A := {yn : n ∈ N}. Now we have two possibilities: either there exists n0 such that B (yn0, ǫ) contains infinitely many yn or A ∩ B (yk, ǫ) is finite for every k. In the first case, it is clear that A1 := A∩B (yn0, ǫ) and A2 := {y2n : y2n−1 ∈ A1}∪ {y2n−1 : y2n ∈ A1} satisfy d(A1, A2) ≥ ǫ. In the second case, we can find a subsequence (ynk) with d (ynk, ynl) > ǫ when k 6= l, and the result follows easily. (cid:3) In Lemma 4.14 and Corollary 4.15 we do not necessarily assume that d(x,x0)≥ǫ Proof. Fix e ∈ E with kek = 1, and let base spaces are not complete, so it could be the case that bX = X and bY = Y . Lemma 4.14. Given x0 ∈ bX, there exists y0 ∈ bY such that cT f (y0) = 0 whenever f ∈ Lip(X, E) satisfies bf (x0) = 0. f (x) < 1) . A :=(f · e : f ∈ Lip(bX), f (x0) = 1, ∀ǫ > 0 sup We will see that there exists a unique point y0 ∈ bY such that [kT f k(y0) = 1 for every f ∈ A. ǫ ∈ A with c(f ′ (cid:3) 14 JES ´US ARAUJO AND LUIS DUBARBIE Fix f0 ∈ A. By Corollary 4.5, taking into account that kf0k∞ = 1, there exists a sequence (yn) in Y such that kT f0(yn)k ≥ 1−1/n for each n ∈ N. Let us see that it is a Cauchy sequence. Suppose that this is not the case. Either if all yn are pairwise different (by using Lemma 4.13) or not, we see that there exist subsets A1, A2 of {yn : n ∈ N} such that d(A1, A2) > 0 and supyn∈Ai kT f0(yn)k = 1, i = 1, 2. Then we take g1, g2 ∈ Lip(bY ) with 0 ≤ g1, g2 ≤ 1 such that g1(A1) = 1, g2(A2) = 1, and g1g2 ≡ 0. It is immediate that kT f0 + giT f0k∞ = 2 for i = 1, 2. Since, again by Corollary 4.5, kT −1 (giT f0)k∞ = 1, we deduce that \kT −1 (giT f0)k(x0) = 1 for i = 1, 2, which goes against the fact that T −1 is separating. Consequently (yn) is a Cauchy sequence and converges continuous extension to h(y). Proof. Let x0 and y0 be as in Lemma 4.14. Since T −1 is also bisepa- straightforward to see that [kT f k(y0) = 1 for every f ∈ A. Corollary 4.5), we conclude that [kT f k(y0) ≤ ǫ, and we are done. of x0 and kf − fǫk∞ ≤ ǫ. We can take f ′ ǫ)∩c(fǫ) = ∅, and we deduce from the paragraph above that \kT fǫk ≡ 0 on a neighborhood to a point y0 ∈ bY , which obviously satisfies \kT f0k(y0) = 1. Now it is Next suppose that f ∈ Lip(X, E) satisfies bf (x0) = 0. Then, given ǫ > 0, there exists fǫ ∈ Lip(X, E) such that bfǫ ≡ 0 on a neighborhood of y0 in bY ; in particular \kT fǫk(y0) = 0. Since kT f − T fǫk∞ ≤ ǫ (by Corollary 4.15. There exists a bijective map h : bY → bX such that T f (y) = Jy(bf(h(y))) whenever y ∈ Y and f ∈ Lip(X, E) admits a rating, there exists x1 ∈ bX such that bf (x1) = 0 whenever cT f (y0) = 0 and, in particular, whenever bf (x0) = 0. Now, as Lip(bX, E) separates points in bX, we deduce that x1 = x0. As a consequence, it is straight- forward to see that Lemma 4.14 gives us a bijective map between bX and bY , which we denote by h : bY → bX, satisfying cT f (y) = 0 if and only if bf (h(y)) = 0. Finally, if f ∈ Lip(X, E) can be continuously extended to h(y), say bf (h(y)) = e ∈ E, then \(f − e)(h(y)) = 0, and Remark 4.16. As in the proof of Corollary 4.15, the bijection k : bX → bY associated to T −1 satisfies [T −1g(x) = 0 if and only if bg(k(x)) = 0, in E with kenk ≤ 1/4n such thatP∞ the representation follows from Corollary 4.12. (cid:3) g ∈ Lip(Y, F ). This implies that k = h−1. Lemma 4.17. If E is not complete, then there exists a sequence (en) n=1 en does not converge in E. NONCOMPACTNESS AND NONCOMPLETENESS 15 Proof. Clearly, there exists a nonconvergent sequence (un) in E satis- fying kun − un+1k ≤ 1/4n for every n ∈ N. It is then easy to check that it is enough to define en := un − un+1 for each n. (cid:3) Corollary 4.18. If E is not complete, then the map h given in Corol- lary 4.15 is a bijection from Y onto X. Proof. We will prove first that h(y) ∈ X whenever y ∈ Y . If this is not Corollary 4.5, for all x ∈ X. fn(x) := max {0, 1 − 2nd (x, h(y))} It is clear that each fn belongs to Lip(X) and that the case, then take y ∈ Y with h(y) ∈ bX \ X. For each n ∈ N, let L(fn) ≤ 2n. It is easy to see that, since Lip(X,bE) is complete, if we take (en) in E as in Lemma 4.17, then f := P∞ n=1 fn · een belongs to Lip(X,bE), and since all values are taken in E, to Lip(X, E). Thus, by T (fn · een)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞ Finally, by Corollary 4.15, this implies that T f (y) = P∞ which belongs to bF \F , and T f takes values outside F , which is absurd. We deduce from Remark 4.16 that h(Y ) = X. Proof of Theorem 3.1. Taking into account Corollaries 4.12, 4.15 and 4.18, it is enough to show that h ∈ iso<2(Y, X). Let y1, y2 ∈ Y be such that d(y1, y2) < 2. We are going to see that D := d(h(y1), h(y2)) ≤ d(y1, y2). lim k→∞(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) T f − kXn=1 n=1 Jy(en), (cid:3) = 0. Pick e ∈ E with kek = 1 and define g ∈ Lip(X, E) by g(x) := max(cid:26)−1, 1 − 2 d(x, h(y1)) D (cid:27) · e for every x ∈ X. We have that kgk∞ = 1, L(g) = 2/D, g(h(y1)) = e, and g(h(y2)) = −e. Obviously, by Corollary 4.12, Jy1 = Jy2, and 1 < 2 d(y1, y2) = kJy1(e) − Jy2(−e)k d(y1, y2) = kT g(y1) − T g(y2)k d(y1, y2) ≤ kT gkL , which implies that kgkL > 1, and then kgkL = L(g) = 2/D. This means that kT gkL = 2/D, and consequently 2/d(y1, y2) ≤ 2/D. The other inequality can be seen in a similar way working with T −1 (see Remark 4.16). (cid:3) Proof of Corollary 3.3. The fact that bT satisfies Property P follows easily from Proposition 4.9. The conclusion is then immediate by The- orem 3.1. (cid:3) 16 JES ´US ARAUJO AND LUIS DUBARBIE 5. The distance between A(T ) and B(T ) Propositions 5.1 and 5.2 will be used in Section 6. Proposition 5.1. Let T : Lip(X, E) → Lip(Y, F ) be a surjective linear isometry. If A(T ) 6= ∅, then d (A(T ), B(T )) ≥ 1. Proof. Obviously B(T ) 6= ∅. Suppose first that d(A(T ), B(T )) < 1, and take y0 ∈ B(T ) and ǫ > 0 with d(y0, A(T )) < 1 − 2ǫ. We then select f ∈ F , kfk = 1, and define l ∈ Lip(Y, F ) by l(y) := max{0, 2 − d(y, y0)} · f for every y ∈ Y . We have that klkL = klk∞ = kl(y0)k = 2, L(l) ≤ 1, and kl(y)k < 2 for all y ∈ Y \ {y0}. Now, by Lemma 4.3 (see also Remark 4.4), we have that L(T −1l) ≤ kT −1lk∞. Consequently kT −1lkL = kT −1lk∞, and then kT −1lk∞ = 2. Therefore, there is a point x0 in X such that kT −1l(x0)k > 2 − ǫ, that is, T −1l(x0) = αe for some e ∈ E, kek = 1, and α ∈ R, α > 2 − ǫ. Next, obviously (cid:13)(cid:13)e + T −1l(cid:13)(cid:13)L ≥(cid:13)(cid:13)e + T −1l(x0)(cid:13)(cid:13) = k(1 + α) ek > 3 − ǫ, so kT e + lkL > 3 − ǫ. Since L(T e + l) ≤ L(T e) + L(l) ≤ 2, this implies that kT e + lk∞ > 3−ǫ, and hence the set B := {y ∈ Y : k(T e + l) (y)k > 3 − ǫ} is nonempty. Notice that, since kT ek∞ ≤ 1, all points y ∈ B must satisty kl(y)k > 2 − ǫ, which is equivalent to d(y, y0) < ǫ. Thus, for some y1 with d(y1, y0) < ǫ, we have kT e(y1) + l(y1)k > 3 − ǫ, which implies that kT e(y1)k > 1 − ǫ. On the other hand, taking into account that d(y0, A(T )) < 1 − 2ǫ, there exists y2 ∈ A(T ) with d(y0, y2) ≤ 1 − 2ǫ. Finally, observe that kT e(y1) − T e(y2)k d(y1, y2) = kT e(y1)k d(y1, y2) > 1 − ǫ 1 − 2ǫ + ǫ = 1, which allows us to conclude that L(T e) > 1, in contradiction with the fact that kek = 1 and T is an isometry. (cid:3) Proposition 5.2. Let T : Lip(X, E) → Lip(Y, F ) be a surjective linear isometry. If y0 ∈ A(T ), then d (y0, B(T )) = 1. Proof. Suppose on the contrary that there exists s ∈ (0, 1) such that d (B(y0, s), B(T )) > 1 + s. Take f ∈ Lip(Y ) with c(f ) ⊂ B(y0, s) and such that 0 ≤ f ≤ s, f (y0) = s, and L(f ) ≤ 1. Let e ∈ E and f ∈ F have norm 1. It is easy to check that kf · f ± T ek ≤ 1, whereas, since T −1 (f · f) 6= 0, Inequality (2.1) implies that (cid:13)(cid:13)T −1 (f · f) + e(cid:13)(cid:13)∞ > 1 NONCOMPACTNESS AND NONCOMPLETENESS 17 or (cid:13)(cid:13)T −1 (f · f) − e(cid:13)(cid:13)∞ > 1, contradicting the fact that T is an isometry. (cid:3) We next see that Property P holds when Y is 1-connected. Obvi- ously, the same result holds if X is 1-connected (see Remark 4.6). Corollary 5.3. Let Y be 1-connected and suppose that Lip(X, E) and Lip(Y, F ) are linearly isometric. Then X is also 1-connected and every surjective linear isometry T : Lip(X, E) → Lip(Y, F ) satisfies Property P. Proof. By Proposition 5.1, Property P holds when Y is 1-connected. The fact that X is 1-connected can be easily deduced from the rep- resentation of T in Theorem 3.1 or that of bT in Corollary 3.3 (taking into account that a metric space is 1-connected if and only if so is its completion). (cid:3) Remark 5.4. An immediate consequence of Corollary 5.3 is that, when X (or Y ) is 1-connected, every surjective linear isometry T : Lip(X, E) → Lip(Y, F ) is standard in any of the cases (i), (ii), given in Theorem 3.1. 6. The case when T does not satisfy Property P In this section, unless otherwise stated, we assume that T is a linear isometry from Lip(X, E) onto Lip(Y, F ) that does not satisfy Property P (that is, A(T ) 6= ∅). We will make use of Theorem 3.1, so we also assume that we are in any of the following two cases: (i) X and Y are complete, (ii) E (or F ) is not complete. It is then clear by Proposition 5.1 that X is complete if and only if both A (T −1) and B (T −1) are complete. We will introduce two isometries on spaces of Lipschitz functions defined on A (T −1) and B (T −1). The fact that these new isometries turn out to be standard will allow us to obtain a description of T . Lemma 6.1. Suppose that f ∈ Lip(X, E) satisfies f ≡ 0 on B(T −1). Then T f ≡ 0 on B(T ). Proof. Suppose on the contrary that there exists y0 ∈ B(T ) with T f (y0) 6= 0. By Lemma 4.1, we can find g ∈ Lip(Y, F ) with kg(y0)k = kgk∞ > L(g) such that kg(y0)k + kT f (y0)k = k(g + T f )(y0)k = kg + T f k∞ > L(g + T f ). 18 JES ´US ARAUJO AND LUIS DUBARBIE We see that sup x∈B(T −1)(cid:13)(cid:13)T −1g(x) + f (x)(cid:13)(cid:13) = sup x∈B(T −1)(cid:13)(cid:13)T −1g(x)(cid:13)(cid:13) ≤ kgk < k(g + T f )(y0)k . On the other hand, if we put f := (g + T f )(y0), since T −1f ≡ 0 on A(T −1), there exists n ∈ N such that sup x∈A(T −1)(cid:13)(cid:13)(cid:13)T −1g(x) + f (x) + nT −1f(x)(cid:13)(cid:13)(cid:13) < so if we denote k := T −1(cid:16)g + nf(cid:17) + f , then we see that kkk∞ < kT kk. x∈B(T −1)(cid:13)(cid:13)(cid:13)T −1g(x) + f (x) + nT −1f(x)(cid:13)(cid:13)(cid:13) Consequently, kkk∞ < L(k) and there exists e ∈ E with T e(y0) 6= 0 such that kk ± ek∞ < L(k) = L(k ± e). < (n + 1) kfk , sup Also L(T k) = L(g + T f ) < kfk, so if we assume n big enough, then kT kk = kT kk∞. Therefore, kT (k ± e)kL = kk ± ekL = L(k) = kT kk∞ = (n + 1) kfk . This implies that k(n + 1) f ± T e(y0)k = kT (k ± e) (y0)k ≤ (n + 1) kfk , which goes against Inequality (2.1). (cid:3) Using Proposition 5.1, we see that the subspace is isometrically isomorphic to Lip(B (T −1) , E), via the restriction map. In the same way, LipB(X, E) :=(cid:8)f ∈ Lip(X, E) : f(cid:0)A(cid:0)T −1(cid:1)(cid:1) ≡ 0(cid:9) LipA(X, E) :=(cid:8)f ∈ Lip(X, E) : f(cid:0)B(cid:0)T −1(cid:1)(cid:1) ≡ 0(cid:9) and Lip(A (T −1) , E) are isometrically isomorphic. Let denote by IB(T −1) : Lip(B (T −1) , E) → LipB(X, E) and IA(T −1) : Lip(A (T −1) , E) → LipA(X, E), respectively, the corresponding natural isometries. In particular we can write in a natural way Lip(X, E) = LipA(X, E)⊕LipB(X, E) = Lip(A(cid:0)T −1(cid:1) , E)⊕Lip(B(cid:0)T −1(cid:1) , E), where this equality has to be seen as a direct sum just in the linear sense. Next, let RB(T ) : Lip(Y, F ) → Lip(B(T ), F ) be the operator sending each function to its restriction. NONCOMPACTNESS AND NONCOMPLETENESS 19 Lemma 6.2. The map TB := RB(T ) ◦ T ◦ IB(T −1) : Lip(B(cid:0)T −1(cid:1) , E) → Lip(B(T ), F ) is a surjective linear isometry. Proof. Notice first that if f ∈ LipB(X, E) and g ∈ Lip(X, E) satisfy f ≡ g on B (T −1), then kf k ≤ kgk. TB is linear and, by Lemma 6.1, it is easy to check that it is surjec- tive. We next see that it is an isometry. Of course this is equivalent to show that (cid:13)(cid:13)RB(T ) ◦ T (f )(cid:13)(cid:13) = kT (f )k for every f ∈ LipB(X, E), and it is clear that(cid:13)(cid:13)RB(T )(T (f ))(cid:13)(cid:13) ≤ kT (f )k. Since(cid:13)(cid:13)RB(T )(T (f ))(cid:13)(cid:13) = (cid:13)(cid:13)IB(T )(cid:0)RB(T )(T (f ))(cid:1)(cid:13)(cid:13), the fact that(cid:13)(cid:13)RB(T )(T (f ))(cid:13)(cid:13) < kT (f )k is equiv- alent to that (cid:13)(cid:13)T −1(cid:0)IB(T )(cid:0)RB(T ) (T (f ))(cid:1)(cid:1)(cid:13)(cid:13) < kf k , which goes against the first comment in this proof. (cid:3) −1◦T ◦IA(T −1) : Lip(A (T −1) , E) → Lip(A(T ), F ) Lemma 6.3. TA := IA(T ) is standard. Proof. Suppose that this is not the case. Since A(T ) = A(TA) ∪ B(TA), we are in fact saying that A(TA) 6= ∅. For e ∈ E with kek = 1, we have A(TA), and consequently, since A ) · e(cid:17)(cid:17) ⊂ there are sequences (yn) in A(TA) and (zn) in B(T ) with A ) · e(cid:17)(cid:13)(cid:13)(cid:13) = 2. Notice that both T e ≡ 0 and T(cid:0)χA(T −1) · e(cid:1) ≡ 0 on A(TA), so T(cid:0)χB(T −1) · e(cid:1) ≡ 0 on A(TA). On the other hand, by Lemma 6.1, c(cid:16)T(cid:16)χA(T −1 A ) · e(cid:17)(cid:13)(cid:13)(cid:13)∞ A ) · e(cid:17) (zn)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)T(cid:16)e + χA(T −1 A ) · e(cid:17)(cid:13)(cid:13)(cid:13) = 1 =(cid:13)(cid:13)(cid:13)T(cid:16)e + χA(T −1 A ) · e(cid:17) (yn) − T(cid:16)e + χA(T −1 A ) · e(cid:17) (yn) − T(cid:0)χB(T −1) · e(cid:1) (zn)(cid:13)(cid:13)(cid:13) kT ek =(cid:13)(cid:13)(cid:13)T(cid:16)χA(T −1 n→∞(cid:13)(cid:13)(cid:13)T(cid:16)e + χA(T −1 n→∞(cid:13)(cid:13)(cid:13)T(cid:16)χA(T −1 n→∞(cid:13)(cid:13)(cid:13)T(cid:16)χB(T −1) · e + χA(T −1 ≤ (cid:13)(cid:13)(cid:13)T(cid:16)(cid:16)χB(T −1) + χA(T −1 A )(cid:17) · e(cid:17)(cid:13)(cid:13)(cid:13) A ) · e(cid:17) (yn) − T(cid:16)χB(T −1) · e + χA(T −1 2 = lim = lim = lim ≤ 1. d (yn, zn) , d (yn, zn) d (yn, zn) A ) · e(cid:17) (zn)(cid:13)(cid:13)(cid:13) 20 JES ´US ARAUJO AND LUIS DUBARBIE We conclude that A(TA) is empty. (cid:3) It is easy to check that TB satisfies Property P, so by Theorem 3.1, it is standard. We deduce the following result, which allows us to give the values on B(T ) and on A(T ) of the images of all functions in Lip(X, E) and LipA(X, E), respectively. Corollary 6.4. There exist (i) hB ∈ iso<2 (B(T ), B (T −1)) and hA ∈ iso<2 (A(T ), A (T −1)), (ii) and maps JB : B(T ) → I(E, F ) and JA : A(T ) → I(E, F ) constant on each 2-component of B(T ) and A(T ), respectively, such that (i) T f (y) = JBy(f (hB(y))) for all f ∈ Lip(X, E) and y ∈ B(T ), and (ii) T f (y) = JAy(f (hA(y))) for all f ∈ LipA(X, E) and y ∈ A(T ). Lemma 6.5. Let y0 ∈ A(T ) and A ⊂ B (T −1) be such that d(hA(y0), A) = 1. If f ∈ LipB(X, E) satisfies f (A) ≡ e ∈ E, then T f (y0) = −JAy0 (e) . Proof. Notice first that, since y0 ∈ A(T ), T e(y0) = 0, and conse- quently, by Corollary 6.4, T(cid:0)χB(T −1) · e(cid:1) (y0) = −T(cid:0)χA(T −1) · e(cid:1) (y0) = −JAy0 (e). Next we prove the result through several steps. We denote a := JAy0 (e) for short. Step 1. Assume that kek = 1 = kf k. Consider k′ ∈ Lip(X) defined by k′(x) := max {0, 1 − d(x, hA(y0))} for every x ∈ X, and k ∈ LipA(X, E) defined by k := −k′ · e. It is easy to see that (k + f )(hA(y0)) = −e and that (k + f )(x) = e for every x ∈ A. As a consequence, kk + f k = 2. Suppose now that T f (y0) = f 6= −a. By Corollary 6.4, T k(y0) = −a and, since kf k∞ = 1, we can take M < 2 such that kT (k + f )(y0)k = k−a + fk < M. Consequently there exists 0 < r < 1 such that kT (k + f ) (y)k < M for every y ∈ B(y0, r). On the other hand, for y ∈ A(T ) with d(y, y0) ≥ r, kT k(y)k = kT (−k′ · e)(y)k = kmax {0, 1 − d(hA(y), hA(y0))} · JAy(e)k ≤ 1 − r, so kT (k + f )(y)k ≤ 2 − r. Since kT (k + f )(y)k = kT f (y)k ≤ 1 for every y ∈ B(T ), we deduce that kT (k + f )k∞ < 2 = kT (k + f )k. NONCOMPACTNESS AND NONCOMPLETENESS 21 Let M > 0 with M + kT (k + f )k∞ < 2, and y ∈ B(T ) such that hB(y) ∈ A and d(hB(y), hA(y0)) < 1 + M/2. Define b := MT e(y) ∈ F . By Corollary 6.4, T −1b(hB(y)) = Me, and consequently 2d(hB(y), hA(y0)) < 2 + M = (cid:13)(cid:13)(cid:13)(cid:16)k + f + T −1b(cid:17) (hB (y)) −(cid:16)k + f + T −1b(cid:17) (hA (y0))(cid:13)(cid:13)(cid:13) , against the fact that(cid:13)(cid:13)(cid:13)k + f + T −1b(cid:13)(cid:13)(cid:13) = 2. It is easy to check that if n ≥ L(f ), then(cid:13)(cid:13)nχB(T −1) · e(cid:13)(cid:13)∞ =(cid:13)(cid:13)nχB(T −1) · e(cid:13)(cid:13) = Step 2. Assume that kek = 1 = kf k∞. n and that (cid:13)(cid:13)f + nχB(T −1) · e(cid:13)(cid:13)∞ =(cid:13)(cid:13)f + nχB(T −1) · e(cid:13)(cid:13) = n + 1. Using Step 1, T (nχB(T −1) · e)(y0) = −na and T (f + nχB(T −1) · e)(y0) = −(n + 1)a. The conclusion is easy. Step 3. Assume that e = 0. Of course we must prove that T f (y0) = 0. Fix d ∈ E with norm 1. Consider m ∈ LipB(X, E) defined by m(x) := max{0, 1 − d(x, A)} · d for each x ∈ X. We easily check that kmk∞ = 1 = kdk, and if we assume that kf k ≤ 1, then kf (x)k ≤ d(x, A) for every x. As in the proof of Lemma 4.1, we see that km + f k∞ = 1 = kdk. The conclusion follows immediately from Step 2. The rest of the proof is easy. (cid:3) Corollary 6.6. Suppose that A1, A2 ⊂ B (T −1) satisfy d(hA(y0), Ai) = 1 for i = 1, 2. Then d(A1, A2) = 0. Proof. Just assume that d(A1, A2) > 0 and apply Lemma 6.5 to any f ∈ LipB(X, E) such that f (Ai) ≡ (−1)ie 6= 0 for i = 1, 2. This leads to two different values for T f (y0). (cid:3) Corollary 6.7. Let y0 ∈ A(T ). Then there exists exactly one point ϕ(y0) in B(T ) such that d (hB (ϕ(y0)) , hA (y0)) = 1. Also, T f (y0) = −JAy0(f (hB (ϕ (y0)))) for every f ∈ LipB(X, E). Proof. By Lemma 4.13 and Corollary 6.6, we deduce that if (xn) is a sequence in X such that d(hA(y0), xn) ≤ 1 + 1/n for each n ∈ N, then it is a Cauchy sequence, so there is a limit x0 in bX, which necessarily belongs to \B (T −1). Obviously the point x0 does not depend on the sequence we take. 22 JES ´US ARAUJO AND LUIS DUBARBIE If this is not the case, for each n ∈ N, let We next assume that bX is not complete and prove that x0 ∈ B (T −1). fn(x) := max {0, 1 − d (x, B (x0, 1/n))} for all x ∈ X. It is clear that each fn belongs to Lip(X). Since in E, to Lip(X, E), and indeed to LipB(X, E). Thus, since f = Lip(X,bE) is complete, if we take (en) in E as in Lemma 4.17, then f :=P∞ n=1 fn · een belongs to Lip(X,bE), and since all values are taken limk→∞Pk n=1 fn · een, we deduce from Lemma 6.5 that kXn=1 T (fn · een(y0)) kXn=1 T f (y0) = lim k→∞ = − lim k→∞ JAy0 (en) = − ∞Xn=1 which belongs to bF \ F . This is absurd. If we define ϕ(y0) := hB JAy0 (en) , −1(x0) ∈ B(T ), then we are done. (cid:3) Proposition 6.8. For every y ∈ A(T ), JAy = −JBϕ(y). Proof. Fix y ∈ A(T ) and e ∈ E, and let f := T e(ϕ(y)). Then T −1f(hB(ϕ(y))) = e and T −1f ≡ 0 on A (T −1). We conclude from Corollary 6.4 that JBϕ(y)(e) = f, and from Corollary 6.7 that −JAy(e) = f. (cid:3) Next result follows now easily from Corollaries 6.4 and 6.7, and Proposition 6.8. Corollary 6.9. For y ∈ A(T ) and f ∈ Lip(X, E), T f (y) = −JAy(f (hB(ϕ(y)))) + JAy(f (hA(y))) = JBϕ(y)(f (hB(ϕ(y)))) − JBϕ(y)(f (hA(y))). Corollary 6.10. Let y0 ∈ A(T ). If y ∈ B(T ) is such that d(y, ϕ(y0)) ≥ 2, then d(y, y0) ≥ 2. Proof. Let e1, e2 ∈ E be vectors with norm 1 and such that JBϕ(y0)(e1) = JBy(e2). Define f := f1 − f2 ∈ LipB(X, E), where f1(x) := max {0, 1 − d(x, hB (ϕ(y0)))} · e1 and f2(x) := max {0, 1 − d(x, hB (y))} · e2 NONCOMPACTNESS AND NONCOMPLETENESS 23 for every x ∈ X. Obviously, kf1k = 1 = kf2k, so to show that kf k = 1, it is enough to see that, if d(x, hB (ϕ(y0))), d(z, hB (y)) < 1, then kf1(x)k + kf2(z)k ≤ d(x, z). Taking into account that 2 ≤ d(hB (y) , hB (ϕ(y0))) ≤ d(z, hB (y)) + d(x, z) + d(x, hB (ϕ(y0))), it follows that kf1(x)k + kf2(z)k = (1 − d(x, hB (ϕ(y0)))) + (1 − d(z, hB (y))) ≤ d(x, z). On the other hand, by Corollary 6.9, T f (y0) = JBϕ(y0)(e1) = T f (ϕ(y0)), and by the way we have taken e1 and e2, we have T f (y) = −JBy(e2) = −JBϕ(y0)(e1). We conclude that, since kT f k = 1, 2 = kT f (y) − T f (y0)k ≤ d(y, y0). (cid:3) Corollary 6.11. Let y0 ∈ A(T ). Given y ∈ B(T ), if 0 ≤ d(y, ϕ(y0)) < 1, then and if 1 ≤ d(y, ϕ(y0)) < 2, then d(y, y0) = 1 + d(y, ϕ(y0)), d(y, y0) ≥ 2. Proof. Fix e ∈ E with norm 1 and let f (x) := min {1, d(x, hB(ϕ(y0)))}· e for every x ∈ X. Let y ∈ B(T ) with d(y, ϕ(y0)) < 2. Taking into account that kf k = 1, Corollaries 6.4 and 6.9 give d(y, y0) ≥ kT f (y) − T f (y0)k = kmin {1, d(hB(y), hB(ϕ(y0)))} · JBy(e) + JBy(e)k ≥ min {2, d(hB(y), hB(ϕ(y0))) + 1} . The conclusion is immediate. (cid:3) Corollary 6.12. If y1, y2 ∈ A(T ) satisfy ϕ(y1) 6= ϕ(y2), then d(y1, y2) ≥ 2. Proof. Suppose that M := d(y1, y2)/2 < 1, so by Corollary 6.4 JAy1 = JAy2. Put N := d(hB(ϕ(y1)), hB(ϕ(y2))) and, for a fixed e ∈ E with norm 1, let f (x) := max {−M, M − d(x, hA(y1))} · e and g(x) := max {0, N − d(x, hB(ϕ(y1)))} · e for every x ∈ X. If we take A > 0 such that M + AN < 1, then k := χA(T −1)f −AχB(T −1)g has norm 1. Also, by Corollary 6.9, T k(y1) = 24 JES ´US ARAUJO AND LUIS DUBARBIE MJAy1(e)+ANJAy1(e) and T k(y2) = −MJAy2(e) = −MJAy1(e). Con- sequently T k(y1) − T k(y2) d(y1, y2) (cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13) > 1, which is impossible. (cid:3) Proof of Theorem 3.4. Corollaries 6.10, 6.11 and 6.12 show that Y is of type A. We consider the associated purely nonstandard map Sϕ : Lip(Y, F ) → Lip(Y, F ) and see that, given e ∈ E, e 6= 0, the composi- tion Sϕ◦T : Lip(X, E) → Lip(Y, F ) satisfies Sϕ◦T (e) = JBϕ(y)(e) 6= 0 if y ∈ A(T ) and Sϕ ◦ T (e) = JBy(e) 6= 0 if y ∈ B(T ), that is, the com- position satisfies Property P. This implies that Sϕ ◦ T is standard. Since Sϕ = Sϕ T = Sϕ ◦ (Sϕ ◦ T ), and we are done. −1, we have that (cid:3) Remark 6.13. It is easy to check that, if h : Y → X and J : Y → I(E, F ) are the associated maps to Sϕ ◦ T in the proof of Theorem 3.4, then h ≡ hA on A(T ) and h ≡ hB on B(T ). In the same way, J ≡ −JA on A(T ) and J ≡ JB on B(T ). Finally, it is also apparent that, given A ∈ Comp2(B (T −1)), A = A′ ∩ B (T −1), where A′ ∈ Comp2(X), and that a similar fact does not necessarily hold for the elements in Comp2(A (T −1)). Proof of Proposition 3.9. Suppose that (X, dα) is of type A. Then dα(z, y) = 1 + dα(z, ϕ(y)) whenever y ∈ A(T ) and z ∈ B(T ) satisfy 0 < dα(z, ϕ(y)) < 1. In such case, dα(z, y) < d(z, y) and d(z, ϕ(y)) < dα(z, ϕ(y)), and this implies d(z, y) > 1 + d(z, ϕ(y)) = d(y, ϕ(y)) + d(z, ϕ(y)), which is impossible. We deduce that, if d(z, ϕ(y)) < 1, then z = ϕ(y). The rest of the proof is easy. (cid:3) 7. Acknowledgements The authors wish to thank the referee for his/her valuable remarks, which helped them to improve the first version of the paper. References [1] J. Alaminos, J. Extremera, and A. R. Villena, Zero product preserving maps on Banach algebras of Lipschitz functions, J. Math. Anal. Appl. 369 (2010), 94 -- 100. [2] J. Araujo, The noncompact Banach-Stone theorem, J. Operator Theory 55 (2006), 285 -- 294. NONCOMPACTNESS AND NONCOMPLETENESS 25 [3] J. Araujo and L. Dubarbie, Biseparating maps between Lipschitz function spaces, J. Math. Anal. Appl. 357 (2009), 191 -- 200. [4] E. Behrends, M -structure and the Banach-Stone theorem, Lecture Notes in Mathematics 736, Springer, Berlin, 1979. [5] R. J. Fleming and J. E. Jamison, Isometries on Banach spaces: function spaces, Chapman & Hall/CRC, Boca Raton, FL, 2003. [6] R. J. Fleming and J. E. Jamison, Isometries on Banach spaces: vector-valued function spaces, Chapman & Hall/CRC, Boca Raton, FL, 2008. [7] F. J. Garc´ıa-Pacheco and B. Zheng, Geometric properties on non-complete spaces, preprint. [8] M. I. Garrido and J. A. Jaramillo, Homomorphisms on function lattices, Monatsh. Math. 141 (2004), 127 -- 146. [9] M. I. Garrido and J. A. Jaramillo, Lipschitz-type functions on metric spaces, J. Math. Anal. Appl. 340 (2008), 282 -- 290. [10] S. Hern´andez, E. Beckenstein, and L. Narici, Banach-Stone theorems and separating maps, Manuscripta Math. 86 (1995), 409 -- 416. [11] K. Jarosz and V. D. Pathak, Isometries between function spaces, Trans. Amer. Math. Soc. 305 (1988), 193 -- 206. [12] M. Jerison, The space of bounded maps into a Banach space, Ann. of Math. 52 (1950), 309 -- 327. [13] A. Jim´enez-Vargas and M. Villegas-Vallecillos, Into linear isometries between spaces of Lipschitz functions, Houston J. Math. 34 (2008), 1165 -- 1184. [14] A. Jim´enez-Vargas and M. Villegas-Vallecillos, Linear isometries between spaces of vector-valued Lipschitz functions, Proc. Amer. Math. Soc. 137 (2009), 1381 -- 1388. [15] E. Kreyszig, Introductory functional analysis with applications, John Wiley & Sons, New York-London-Sydney (1978). [16] K. de Leeuw, Banach spaces of Lipschitz functions, Studia Math. 21 (1961/62), 55 -- 66. [17] D. H. Leung, Biseparating maps on generalized Lipschitz spaces, Studia Math. 196 (2010), 23 -- 40. [18] E. Mayer-Wolf, Isometries between Banach spaces of Lipschitz functions, Is- rael J. Math. 38 (1981), 58 -- 74. [19] A. K. Roy, Extreme points and linear isometries of the Banach space of Lip- schitz functions, Canad. J. Math. 20 (1968), 1150 -- 1164. [20] M. H. Vasavada, Closed ideals and linear isometries of certain function spaces, Ph.D. Thesis, Wisconsin University, 1969. [21] N. Weaver, Isometries of noncompact Lipschitz spaces, Canad. Math. Bull. 38 (1995), 242 -- 249. [22] N. Weaver, Lipschitz algebras, World Scientific Publishing, River Edge, NJ, 1999. Departamento de Matem´aticas, Estad´ıstica y Computaci´on, Facul- tad de Ciencias, Universidad de Cantabria, Avenida de los Castros s/n, E-39071, Santander, Spain. E-mail address: [email protected] E-mail address: [email protected]
1609.04179
3
1609
2017-01-03T12:05:33
More on functional and quantitative versions of the isoperimetric inequality
[ "math.FA" ]
This paper deals with the famous isoperimetric inequality. In a first part, we give some new functional form of the isoperimetric inequality, and in a second part, we give a quantitative form with a remainder term involving Wasserstein distance of the classical isopemetric inequality. In both parts, we use optimal transportation. Finally, we use our refined isoperimetric inequality in some classical cases arising in convex geometry.
math.FA
math
More on functional and quantitative versions of the isoperimetric inequality Erik Thomas May 21, 2018 Abstract The goal of the present paper is to discuss new functional extensions of the isoperimetric inequality, and quantitative versions involving the Wasser- stein's distances. 1 Introduction We shall work on the Euclidean space (Rn,·,·). The sharp (anisotropic) isoperi- metric inequality can be stated as follow: given a convex body K ⊂ Rn (having zero in its interior), if we denote by n′ = n n − 1 the Lebesgue conjugate to n, we have for every Borel set E ⊂ Rn that pK (E) ≥ nK n E 1 1 n′ , (1) with equality if E = λK + a for some λ > 0 and a ∈ Rn. Here pK(E) = lim inf ε→0 E + εK − E ε . Equivalently, if E has a regular enough boundary ∂E, then where pK (E) =Z∂E hK (−ν(x)) dHn−1(x), hK (z) := sup y∈K y · z, ∀z ∈ Rn. 1 is the support function of the body K, ν(x) is the outer unit normalto ∂E at x ∈ ∂E, and Hn−1 stands for the (n − 1)-dimensional Hausdorff measure on ∂E. The classical Euclidean isoperimetric inequality corresponds to the case when K = Bn 2 = { · ≤ 1}, the Euclidean unit ball. We want to analyse functional versions of (1). Replacing E by a locally Lip- schitz function f : Rn → R+ is standard. We have decided to work with non- negative functions recalling that for f with values in R we can apply the result to f and use the fact that for f ∈ W 1,1 loc (Rn) , we have, almost-everywhere, ∇f = ±∇f . The functionnal inequality takes the same form n (cid:18)ZRn pK (f ) ≥ nK f n′(cid:19)1/n′ (2) , 1 with equality if f = 1E (provided the gradient term below is understood as a capacity of the bounded variation function 1E). Here, pK(f ) =ZRn hK(−∇f (x)) dx. The inequality (2) can be proven directly using a mass transportation method, as observed by Gromov, see the appendix of [M-S]. In the case of the Euclidean ball, K = Bn 2 , we recover pBn 2 (f ) =(cid:13)(cid:13)(cid:13)∇f(cid:13)(cid:13)(cid:13)L1(Rn) . Extending the convex body K to a (convex) function or measure is less obvious. First, one needs to have a proper extension of the notion of support function hK for a convex function V . Actually, the integral term R hK (−∇f ) needs a proper interpretation, so that non only a convex function will enter the game, but also some "convex measure" (in the terminology of Borell [Bor1] and [Bor2]) associated to it. This has been studied recently in several papers. In particular, in [Kl] corresponding extensions of the isoperimetric inequality (2) have been proposed. See also [Co-Fr] and [M-R]. Here we will establish a new inequality that has the advantage to contain the geometric versions (1) and (2). We will do that by picking a good category of convex measures. First, we need to introduce some notation. Let V : Rn → R ∪ {+∞} be a nonnegative convex function such that ZV := ZRn(cid:18)1 + dx < +∞. We associate to V the probability ZV (cid:16)1 + 1 V (x)(cid:19)−n V (x)(cid:17)−n Our generalization for pK (f ) is as follows, for f : Rn → R+ locally Lipschitz we put dµV (x) = n − 1 n − 1 measure dx. (3) 1 1 2 pV (f ) :=ZRn V ∗ −∇f n−1! f n f n n−1 +(cid:18)ZRn V dµV(cid:19)(cid:18)ZRn f n′(cid:19) where V ∗ is the Legendre's transform of V , V ∗ (y) = sup x∈Rn x · y − V (x) , ∀y ∈ Rn. In particular, we have following inequality (known as Young's inequality): ∀x, y ∈ Rn, x · y ≤ V (x) + V ∗ (y) , (4) with equality when y = ∇V (x). Note that when on K V = 1∞ K := 0 +∞ outside K V = 0 µV almost-everywhere and V ∗ = hK . The general isoperimetric-Sobolev inequality, we can get is then as follow. is the "indicatrix" of a convex set K, then pV (f ) =RRn hK (−∇f ) = pK (f ) since Theorem 1. Let V be a nonnegative convex function with ZV =RRn(1 + V n−1 )−n < +∞ and µV the associated probability measure (3). Then, for every nonnegative locally Lipschitz function f on Rn we have pV (f ) ≥(cid:20)n Z 1 n VZRn (1 + V )dµV(cid:21) kfkLn′ (Rn), 1 n − 1 and, when V is finite, with equality when f (x) = (cid:16)1 + 1 a ∈ Rn. n−1 V (x − a)(cid:17)−(n−1) (5) with Thanks to the remark prior to the theorem, we see that when V = 1∞ K , in- equality (5) becomes exactly (2). The second topic of the present paper is mainly independent of what we dis- cussed so far, although based again on mass transport methods. We aim at pre- senting some quantitative forms of the geometric isoperimetric inequality (1) that involve a Kantorovich-Rubinstein (or Wassertein) distance cost to an extremizer. For u : Rn → Rn a Borel map and µ a measure in Rn, we write u♯µ for the measure defined by for all Borel sets M ⊆ Rn. It is called the push-forward of µ through u. u♯µ (M ) := µ(cid:16)u−1 (M )(cid:17) 3 For µ and ν two probability measures in Rn, and a (cost) function c : Rn × Rn → R+, we define the Kantorovich-Rubinstein or Wasserstein transportation cost Wc (µ, ν) by Wc (µ, ν) = inf T :Rn→Rn:T♯µ=νZRn π ZZRn×Rn = inf c(x, y)dπ(x, y) c(x, T (x))dµ (x) where the infimum is taken over probability measures π on Rn × Rn that have µ and ν as marginals, respectively. For 1 ≤ p ≤ +∞ and c(x, y) = y − xp, the p-th power classical p-Kantorovich-Rubinstein distance, W p p , is recovered. We refer to [V] for details. Our measures will be uniform measures on the sets E and K. Given a Borel set E, we will denote by λE the Lebesgue measure restricted to E and normalized to be a probability measure, dλE(x) = 1E (x) E dx where 1E is the indicator function of the set E. Given a Borel set E, we will denote by eE the homothetic of volume one of the set E, namely eE = 1 E 1 n E. Note that for u ∈ GLn (R) we have λu(E) = u♯λE. And with some abuse of notation, we denote for t > 0 by t♯µ the image of µ under the dilation by t, we have t♯λE = λtE and λeE = 1 E ♯ λE. 1 n Our cost function will depend on the set E. Recall that for a probability measure µ on Rn, its Cheeger constant DChe (µ) is the best (i.e. largest) constant such that the following inequality holds for all Borel sets A: µ+ (A) ≥ DChe (µ) min {µ (A) , 1 − µ (A)} where µ+ denotes the measure of the perimeter (or Minkowski content) associated to µ. It can be defined by: µ+ (A) := lim inf ǫ→0 µ (Aǫ) − µ (A) ǫ , 4 where Aǫ = {x ∈ Rn : dist (x, A) < ǫ} . Equivalently, if we denote by hp,q(µ) the best nonnegative constant for which the inequality (cid:18)Z ∇f (x)q dµ(x)(cid:19) 1 q ≥ hp,q(µ)(cid:18)Z (cid:12)(cid:12)(cid:12)(cid:12)f (x) −Z f dµ(cid:12)(cid:12)(cid:12)(cid:12) p p dµ(x)(cid:19) 1 holds for all f ∈ W 1,1 convex, increasing function defined on R+ by loc ∩ Lp (µ), then h1,1(µ) ≤ DChe(µ) ≤ 2h1,1(µ). Let F be the F (t) := t − log(1 + t). The function F behaves like t2 for t small and like t for t large, and minnt2, to ≤ F (t) ≤ 2 minnt2, to , ∀t ≥ 0. This function appears in several mass transport proofs to give a remainder term, to instance in [F-M-P], [B-K], [CE-Go]. Given a probability measure µ, we will use the following cost c : Rn × Rn → R+ that is also used in [CE]: cµ(x, y) := F (DChe (µ) y − x) (6) which behaves like DChe(µ)2 y − x2 for small distances, and like DChe(µ) y − x for large ones. Our main result is the following extension of the isoperimetric inequality. Theorem 2. Let K be a convex body on Rn. Given a Borel set E ⊂ Rn with xdx, we have locally Lipschitz boundary and ReE xdx =ReK R(E, K) := (7) (8) 1 pK (E) n E nK n−1 − 1 ≥ n c nWcλeE(cid:16)λeE, λeK(cid:17), for some universal constant c > 0, and as a consequence R(E, K) ≥ c nF (DChe(λE)W1(λE, λK)) We emphasize here a weakness of this result: the remainder term depends on will not be too wild : it will belong to a family of sets for which we have a good xdx can always be achieved by translating E, we can drop this assumption provided E (on the Cheeger constant of eE, precisely). But in some geometric problems, eE xdx = ReK control on DChe(λeE), as we will see later. Since the condition ReE (τvλeE, λeK) where τvν is the the transportation term is replaced by minv∈Rn WcλeE 5 image of the measure ν by the translation by v in Rn. Note that we still have equality if E = λK for some λ > 0. When E is convex, it is known that DChe(λE ) > 0. In this case, we also know that up to numerical constants, DChe(λE) is the same as h2,2 (λE), the Poincar´e constant associated to E (or the inverse of the spectral gap). Let us compare our results to existing quantitative Sobolev and isoperimetric inequalities, obtained by Figalli-Maggi-Pratelli. In [F-M-P], there is a quantitative isoperimetric inequality (the numerical constant we use are the improved ones obtained by Segal [S]): 1 1 n′ (cid:18)1 + C n7 AK (E)2(cid:19) , pK (E) ≥ nK where AK (E) := infn E ∆(x0+rK) constant. E n E : x0 ∈ Rn, rn K = Eo and C is a numerical (9) This result of Figalli-Maggi-Pratelli is much deeper and in general stronger than ours, since it is universal (the bound does not depend on geometry of E, as n7 AK (E)2 decreases to 0 in our case). We can note however that the quantity C when the dimension n goes to +∞. Actually, there are some particular cases in which our result might give a better bound, both in fixed dimension and when the dimension grows. The reason is that the transportation cost term can be rather large. For instance, we will give examples where our remainder, F (DChe(λE)W1(λE, λK)) decreases slower than 1 n7 of the inequality (9). The rest of the paper is organized as follows. In the next section, we collect some results on optimal transportation theory. Then, we will prove our two The- orems above. In a final section, we will compute our reminder term in several situation of interest arising in convex geometry. I would like to thank my Professor Dario Cordero-Erausquin for his encourage- ments, his careful reviews and his many useful discussions. 2 Proof of Theorem 1 We first give some background about optimal transportation. 2.1 Background on optimal transportation The following Theorem, due to Brenier [Br] and refined then by McCann [Mc1], is the main result in optimal transportation. 6 Theorem 3. If µ and ν are two probability measures on Rn and µ absolutely continuous with respect to Lebesgue measure, then there exists a convex function φ such that T = ∇φ transports µ onto ν. Moreover, T is uniquely determined µ almost-everywhere. That means that for every nonnegative Borel function b : Rn → R+, ZRn b (y) dν (y) =ZRn b (T (x)) dµ (x) . (10) If µ and ν have densities, say F and G, (10) becomes ZRn b (y) G (y) dy =ZRn If φ is C2 the change of variables y = ∇φ(x) in (11) gives the Monge-Ampère equation, for F (x) dx almost-every x ∈ Rn: b (∇φ (x)) F (x) dx. (11) F (x) = G (∇φ (x)) det(cid:16)D2φ (x)(cid:17) , (12) where D2φ is the hessian matrix of φ. Remark 1. When T is the Brenier map between λE and λK with E and K two convex bodies with same volume, (12) is simpler: det(cid:16)D2φ (x)(cid:17) = 1, for λE almost-every x ∈ E. The question of regularity of φ can be asked because, in the previous equal- ity (12), φ seemed to be required C2. In fact, this is not the case, as it was established by McCann [Mc2] that we can give an almost-everywhere sense to (12) by rather standard arguments from measure theory. This almost-everywhere the- ory is sufficient for most applications, including the one in the present paper but it requires some further arguments that will be discussed later. 2.2 Proof of Theorem 1: the inequality Let us recall the frame. Let V : Rn → R+ ∪{+∞} a nonnegative convex function such that ZV = RRn(cid:16)1 + 1 dx < +∞. So we define the probability measure µV by V (x)(cid:19)−n Let f : Rn → R+ a Borel function such that 0 < RRn f define the probability measure µ by dµV (x) = n − 1 1 dx. n n−1 < +∞. So we can n−1 V (x)(cid:17)−n ZV (cid:18)1 + 1 7 dµ (x) = dx. f n n−1 (x) n n−1 RRn f Let T = ∇ϕ the Brenier between µ and µV and we start by studying the regularity of ϕ. It is sufficient to prove the Theorem for measures µ and µV whose support is Rn. We also can assume that f is the convolution of a function compactly supported and a mollifier, so that f is smooth and converges rapidly to 0 at +∞. Then, it is known that prove that ϕ ∈ W 2,1 loc (Rn) and the following equality ZRn f ∆ϕ = −ZRn ∇f · ∇ϕ (13) is valid. Let us prove now the first part of Theorem 1. Proof. We first need the following Fact. Fact 4. [CE-N-V] Let dµ (x) = F (x) dx and dν (y) = G (y) dy two probabil- ity measures on Rn. Let T = ∇ϕ the Brenier map between µ and ν. Then, the following inequality holds: ZRn G1− 1 n ≤ 1 nZRn F 1− 1 n ∆ϕ. Let us give the proof this Fact for completeness. Proof. We start with Monge-Amp`ere equation, for µ almost-every x ∈ Rn, we have: F (x) = G (∇ϕ (x)) det(cid:16)D2ϕ (x)(cid:17) . Then, for µ almost-every x ∈ Rn and thanks to arithmetic-geometric inequality: n An integration with respect to dµ (x) = F (x) dx gives: G− 1 n (∇ϕ (x)) ≤ F − 1 n (x) ∆ϕ (x) . (14) G− 1 n (∇ϕ (x)) dx n (x) dx. G1− 1 1 nZRn F 1− 1 n (x) ∆ϕ (x) dx ≥ ZRn ={z}(11) ZRn 8 If we apply this Fact to our situation, it gives: ZRn (cid:16)1 + 1 n−1 V(cid:17)−n+1 1 n′ V Z 1 nZRn ≤ f kfkLn′ (Rn) ∆ϕ, and nkfkLn′ (Rn) ZRn (cid:16)1 + 1 n−1 V(cid:17)−n+1 1 n′ V Z As nkfkLn′ (Rn) RRn have the following lines: 1 n′ V Z (1+ 1 n−1 V )−n+1 = nkfkLn′ (Rn)Z f ∆ϕ. ≤ZRn V RRn(cid:16)1 + 1 1 n n−1 V(cid:17) dµV , we now nkfkLn′ (Rn)Z 1 n V ZRn(cid:18)1 + 1 n − 1 2.3 Case of equality f ∆ϕ V(cid:19) dµV ≤ ZRn ={z}(13) ZRn (−∇f ) · ∇ϕ n−1 · ∇ϕ! f = ZRn −∇f V ∗ −∇f n−1! f ZRn ≤{z}(4) n−1! f V ∗ −∇f = ZRn V ∗ −∇f n−1! f = ZRn f f f f n n n n n n−1 n n n−1 +ZRn n−1 +(cid:18)ZRn n−1 +(cid:18)ZRn n n n−1 V (∇ϕ) f V ◦ T dµ(cid:19)(cid:18)ZRn V dµV(cid:19)(cid:18)ZRn f n′(cid:19) f n′(cid:19) . n−1 V (x)(cid:17)−(n−1) In this subsection, we establish that the inequality (5) becomes an equality when f (x) = (cid:16)1 + 1 with V : Rn → R a finite convex function. Note that in this case the Brenier map T = ∇ϕ is T (x) = x so D2ϕ = I and Monge- Amp`ere equation is, for µ almost-every x ∈ Rn, V (T (x))(cid:19)−n n−1 (x) =(cid:18)1 + (15) 1 f . n n − 1 det(cid:16)D2ϕ (x)(cid:17) } {z =1 9 If we come back to the proof of the inequality (5), we remark that we use only two inequalities: the inequality in Fact 4 (which is an arithmetic-geometric inequality) and Young's inequality. We note that in our case case, the inequality in Fact 4 in an equality since D2ϕ = I so det(cid:16)D2ϕ (x)(cid:17) = ∆ϕ(x) Let us treat now Young's inequality. We have an equality in Young's inequality if (and only if) for µ almost-every x ∈ Rn n . −∇f (x) n n−1 (x) f = ∇V (T (x)) = ∇V (x). (16) To get this equality, let us take the − 1 n power in (15) to get 1 f n−1 (x) = 1 + 1 n − 1 If we compute the gradient of the previous line, we find (16). n − 1 V (T (x)) = 1 + V (x) . 1 Remark 2. One can prove that (5) is an equality if and only if f (x) =(cid:16)1 + 1 with a ∈ Rn. We decided not to prove this because it is technical. Let us speak about this. If we have an equality in (5), we have an equality in the inequality in Fact 4 and an equality in Young's inequality. An equality in Fact 4 means that det(cid:16)D2ϕ (x)(cid:17) = ∆ϕ(x) so the matrix D2ϕ (x) has only one eigenvalue, say λ (x) and D2ϕ (x) = λ (x) I. The main difficulty is to show that the function λ is constant, which is the case when ϕ is C2 smooth (the details are analyzed in [?]). If we assume that, it is easy to conclude that, up to translations, µ = µV . n n−1 V (x − a)(cid:17)−(n−1) 3 Proof of Theorem 2 Here we establish Theorem 2. It was noted by Figalli, Maggi and Pratelli [F-M-P] (and Segal [S]) that for this kind of result, the general situation follows from the case the two bodies have same volume, equal to one. For completeness, let us recall the argument. Let E Borel set and K a convex body in Rn. The following Lemma establishes a link between R (E, K) = pK (E) 1 n E et K K n′ − 1 and R(cid:16)eE,fK(cid:17) = n′ − 1 where eE and fK are respectively E nK E 1 n 1 n . 1 1 1 peK(eE) n(cid:12)(cid:12)(cid:12)eK(cid:12)(cid:12)(cid:12) n(cid:12)(cid:12)(cid:12)eE(cid:12)(cid:12)(cid:12) 10 Lemma 5. With the previous notations, R (E, K) = 1 pK (E) n E nK 1 n′ − 1 = Proof. Let us note that for all ǫ > 0, we have E + ǫK − E ǫ = E 1 n K 1 1 peK(cid:16)eE(cid:17) n(cid:12)(cid:12)(cid:12)fK(cid:12)(cid:12)(cid:12) n (cid:12)(cid:12)(cid:12)eE(cid:12)(cid:12)(cid:12) n′ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)eE + ǫ K n′ − 1 = R(cid:16)eE,fK(cid:17) . n fK(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) −(cid:12)(cid:12)(cid:12)eE(cid:12)(cid:12)(cid:12) E 1 n 1 n . 1 1 ǫ K E 1 n By taking the limit, we get the equality. Therefore, if we have established Theorem 2 for two sets of volume one, we have the general statement by applying it to eE and fK. So in the rest of this section, E is a Borel set with smooth boundary and K a convex body, both with volume one, E = K = 1. As in [F-M-P] and [S], the argument to establish Theorem 2 starts with optimal transportation. The following Lemma gives a first minimization for the deficit R (E, K) = pK (E) 1 n E n′ − 1. nK 1 Lemma 6. [F-M-P] Let E and K two convex bodies in Rn with same measure 1. Let T = ∇φ the Brenier map between the measures λE and λK. We note by 0 < λ1 ≤ · · · ≤ λn the eigenvalues of D2φ. Then, we have the following inequality (17) R (E, K) ≥ZRn (λA − λG) dλE, where λA = λ1+···+λn n and λG = Πn i=1λ 1 n i . Before briefly recalling the proof of this Lemma, let us speak about the reg- It is known, see [Ca], that when T = ∇φ is ularity of the optimal transport. the brenier map between dµ (x) = f (x) dx and dν (y) = g (y) dy two probability measures supported on two open bounded sets, respectively E and K, with f and g are α-Holder, bounded and with 1 g bounded too, then φ ∈ C2,β (E) for all 0 < β < α. f and 1 Proof. As E = K = 1, we can write 1 nK n E 1 n′ =ZRn n(cid:16)det(cid:16)D2φ(cid:17)(cid:17) 1 n dλE =ZE n(cid:16)det(cid:16)D2φ(cid:17)(cid:17) 1 n , 11 because det(cid:16)D2φ(cid:17) = 1 thanks to the Remark 1. The arithmetic-geometric in- equality gives The divergence theorem provides nK n E 1 1 n′ ≤ZE div T (x) dx. ZE div T (x) dx =Z∂E T (x) · νE (x) dHn−1 (x) , where Hn−1 is the (n − 1)-dimensional Haussdorf measure. By definition of the support function hK , of K, and since T (E) ⊆ K we therefore get 1 n′ ≤ZE divT (x) dx ≤Z∂E hK (νE (x)) dHn−1 (x) = pK (E) . 1 nK n E Thus pK (E) n′ E 1 nK n − 1 ≥ZRn divT n − 1! dλE ≥ZRn 1 (λA − λG) dλE. To go on, we need a quantitative version of the arithmetic-geometric inequality. The following result is due to Alzer [A]. Lemma 7. [A] Let 0 < λ1 ≤ · · · ≤ λn. Let λA = λ1+···+λn We have n and λG = Πn i=1λ 1 n i . nXi=1 (λi − λG)2 ≤ 2nλn (λA − λG) . (18) We can now complete the proof of Theorem 2. If T = ∇ϕ is the Brenier map between λE and λK for two bodies E and K of volume 1. With the previous notations, if we use (17) and (18), we get HS 1 2n kD2φ − Idk2 tr(cid:16)F(cid:16)kD2θkHS(cid:17)(cid:17) , (λA − λG) ≥ where θ (x) = φ (x) − x2 2 and k·kHS refers to the Hilbert-Schmidt norm of a n× n matrix. Let us remark that λG = 1, thanks to, once again, Remark 1. So we have kD2φ − Idk2 1 + kD2φ − IdkHS ≥ 1 2n c n ≥ λn HS R (E, K) ≥ c nZRn tr(cid:16)F(cid:16)kD2θkHS(cid:17)(cid:17) . (19) The treatment of this term is stated in the next Lemma and we refer to [CE]. 12 Lemma 8. [CE] Let µ a probability measure on Rn absolutely continuous with respect to the Lebesgue measure and θ ∈ W 2,1 loc (Rn) with D2θ + Id ≥ 0 almost- everywhere. We assume ∇θ ∈ L1 (µ) and RRn ∇θdµ = 0. Then, tr(cid:16)F(cid:16)D2θ(cid:17)(cid:17) dµ ≥ cZRn F (DChe (µ) ∇θ) dµ, ZRn for some numerical constant c > 0. Note that our assumptionRE xdx =RK xdx rewrites asRE ∇θ = 0, so if we use the previous Lemma with µ = λE, in (19) we find R (E, K) ≥ ≥ 4 Some examples c nZRn F (DChe ∇θ) dλE c nWcλE (λE, λK) . Here we give some examples where our result (i.e. Theorem 2) gives good bounds for the remainder term, better thant the one in [F-M-P]. We will give an example in dimension 2 where our remainder term, depending on a parameter, can be as large as we want and an example in dimension n. We recall that the remainder term in (9) is bounded by 1 when E and K have for measure 1 and decreases to 0 with 1 n7 when the dimension n grows. 4.1 In dimension 2 In this section, we give a toy example in dimension 2. Let, for α > 0, Eα = 2αi and Kα = (cid:20)− α2 2 , α2 2 (cid:21) ×h− 1 2α2 , 1 2α2i . We will prove the fol- h− α 2 , α lowing: 2i ×h− 1 2α , 1 Proposition 9. With the previous notations, we have: lim α→+∞ DChe (λEα) W1 (λEα, λKα) = +∞. Proof. As W1 (λEα, λKα) = W1 (λKα, λEα) , we give an estimation of the last term. Let T = ∇φ the Brenier map which transports the measure λKα onto the measure λEα (by Monge-Amp`ere equation, it verifies det D2φ = 1, in particular it preserves the volume). Let K′ α =(cid:20) α2 4 , α2 2 (cid:21) ×h− 1 2α2 , 1 2α2i . Then, we have: W1 (λEα, λKα) =ZRn T (x) − x dλKα (x) ≥ZK ′ α T (x) − x dx ≥(cid:12)(cid:12)(cid:12)K′ α(cid:12)(cid:12)(cid:12) dist(cid:16)K′ α, Eα(cid:17) . 13 So, we have W1 (λEα, λKα) ≥ 1 4 α2 4 − α 2! . We need an estimation of DChe (λEα) . This constant could be computed ex- plicitly but it is rather easier to compare this constant to the Poincar´e constant h2,2 (λEα) . Indeed, it is known that, up to numerical constants, that DChe (λEα) is the same as h2,2 (λEα), see [Le] and [M]. As λEα = λ[− α 2α ], then 2α , 1 2 ] ⊗ λ[− 1 2 , α h2,2 (Eα) ≥ min(cid:26)h2,2(cid:18)λ[− α h2,2(cid:16)λ[−a,a](cid:17) = π 2 , α a for a > 0,,it follows with Theorem 2 that: 2α , 1 2 ](cid:19) , h2,2(cid:18)λ[− 1 4F cπ 2α ](cid:19)(cid:27) , see [Bo-H] and [Bo2]. Since 2α α2 4 − 2!! , λ 1 R (Eα, Kα) ≥ for some numerical constant c > 0. In particular the remainder term R (Eα, Kα) is not bounded when α grows to +∞ whereas the remainder in (9) remains bounded. 4.2 Estimation of W1 (λK, λL) for K and L isotropic convex bodies Here K and L are two convex bodies with measure 1. We say that a convex body K is in isotropic position if K = 1, it is centered and there exists α > 0, such that ZK x · y2 dx = αy2 , ∀y ∈ Rn. For an isotropic convex body K, we define its isotropic constant LK (= √α) by: We also define for any convex isotropic body K L2 K = M (K) = 1 nZK x2 dx. √nZK x dx. 1 Using Holder inequality and Borell deviation inequality [Bor1, Bor2], we have: for some numerical constant c > 0. For backgrounds, we refer to [Br-Gi-Va-Vr]. Our goal is here is to prove the following Proposition. cL (K) ≤ M (K) ≤ L (K) , 14 Proposition 10. Let K and L two convex bodies of volume 1 in isotropic position. Then, the following estimation for W1 (λK, λL) holds: √nM (K) − M (L) ≤ W1 (λK, λL) ≤ c (LK + LL) √n + 8, (20) for some numerical constant c > 0. In connection with some isoperimetric esti- mates, we are mainly interested with lower bounds. We are mainly interested in the lower bound provided by this Proposition. The upper bound (which is far from sharp when K are L are closed to each other) is stated only to emphasize that the generic expected order of magnitude is √n on both sides. Proof. We first work on the left hand-side of (20). Let us recall the dual of W1 (λK, λL) , known as Kantorovich's duality: W1 (µ, ν) = sup φ 1−Lip(cid:26)ZRn φ dµ −ZRn φ dν(cid:27) . (21) If we take in (21), φ (x) = x or −x , we have: W1 (λK , λL) ≥ (cid:12)(cid:12)(cid:12)(cid:12)ZRn x dλK (x) −ZRn x dλL (x)(cid:12)(cid:12)(cid:12)(cid:12) = M (K) − M (L) . Let us treat now the right hand-side of (20). Let T be a transport map (in particular, it verifies det (∇T ) = 1) which transports the measure λK onto the measure λL, so W1 (λK , λL) ≤RRn T (x) − x dλK (x) =RK T (x) − x dx. Let us recall a deep result of Paouris, see [P]. Theorem 11. [P] There exists a numerical constant c > 0 such that if K is an isotropic convex body in Rn, then (cid:12)(cid:12)(cid:12)nx ∈ K : x ≥ c√nLK to(cid:12)(cid:12)(cid:12) ≤ exp(cid:16)−√nt(cid:17) , Let t ≥ 1 such that exp (−√nt) ≤ 1 n (note that t = 1 works, we will set this value for t) so the following sets K1 = {x ∈ K : x < c√nLK} and L1 = {x ∈ L : x < c√nLL} have their volumes bigger than 1 − 1 n . Finally, let K2 = T −1 (L1) . Since det (∇T ) = 1, we have K2 = L1 . We can now conclude thanks to the following inequality: ∀t ≥ 1. (22) W1 (λK, λL) ≤ZK T (x) − x dx =ZK1∩K2 T (x) − x dx +ZK\(K1∩K2) T (x) − x dx. (23) 15 Let us estimate the two remaining integrals. Thanks to the definitions of K1 and K2, we have: RK1∩K2 T (x) − x dx ≤ c (LK + LL) √n. For the second, we need the fact that K, L ⊆ B(cid:16)0,qn (n + 2)(cid:17) ⊆ B (0, 2n) , see [K-L-S]. Then, for x ∈ K\ (K1 ∩ K2) , we have T (x) − x ≤ 4n and K\ (K1 ∩ K2) ≤ 2 RK\(K1∩K2) T (x) − x dx ≤ 8. Going back to (23), we finally have: W1 (λK, λL) ≤ZK T (x) − x dx ≤ c (LK + LL) √n + 8. n , that gives References [A] H. Alzer, A new rafinement of the arithmetic mean-geometric mean inequality. Rocky Mountain J. Math. 27, no. 3, pp. 663-667 (1997). [B-K] F. Barthe, A. V. Kolesnikov, Mass transport and variants of the logarithmic Sobolev inequality. J. Geom. Anal. 18, no. 4, pp. 921-979 (2008). [Bo1] S.G. Bobkov, On Isopemetric Constants for Log-Concave Probability Dis- tributions. Geometric aspects of functional analysis, pp. 81-88, Lecture Notes in Math., 1910, Springer, Berlin (2007). [Bo2] S.G. Bobkov, On the isoperimetric constants for product measures. J. Math. Sci. (N.Y.), vol. 159, no. 1, pp. 47-53. Translated from: Problems in Math. Analysis, 40, pp. 49-56 (2009). [Bo-H] S.G. Bobkov, C. Houdré, Isoperimetric constants for product probability measures. Ann. Probab. 25, no. 1, pp. 184-205 (1997). [Bor1] C. Borell, Convex measures on locally convex spaces. Ark. Mat. 12, pp. 239-253 (1974). [Bor2] C. Borell, Convex set functions in d-space. Period. Math. Hungar. 6, pp. 111-136 (1975). [Br-Gi-Va-Vr] S. Brazitikos, A. Giannopoulos, P. Valettas, B.-H. Vritsiou, Geom- etry of isotropic convex bodies. Mathematical Surveys and Monographs, 196. AMS, Providence, RI (2014). [Br] Y. Brenier, Polar factorization and monotone rearrangement of vector-valued functions. Comm. Pure Appl. Math. 44, no.4, pp. 375-417 (1991). 16 [Ca] L. Caffarelli, The regularity of mappings with a convex potential. J. Amer. Math. Soc.(1), pp. 99-104 (1996). [Co-Fr] A. Colesanti, I. Fragal`a, The first variation of the total mass of log-concave functions and related inequalities. Advances in Mathematics 244, pp. 708-749 (2013). [CE] D. Cordero-Erausquin, Transport inequalities for log-concave measures, quantitative forms and applications. ArXiv version (2016). [CE-Ga-H] D. Cordero-Erausquin, W. Gangbo, C. Houdré, Inequalities for gen- eralized entropy and optimal transportation. Recent Advances in the Theory and Applications of Mass Transport, Contemp. Math. 353, A.M.S., Providence, R.I., (2004). [CE-Go] D. Cordero-Erausquin, N. Gozlan, Transport proofs of weighted Poincar´e inequalities for log-concave distributions. Bernoulli - to appear (2015). [CE-N-V] D. Cordero-Erausquin, B. Nazaret, C. Villani, A mass transportation approach to sharp Sobolev and Gagliardo-Niremberg inequalities. Advances in Math. 182, pp. 307-332 (2004). [El] R. Eldan, Thin shell implies spectral gap up to polylog via a stochastic lo- calization scheme. Geometric and Functional Analysis. April 2013, Volume 23, Issue 2, pp 532-569 (2013). [F-M-P] A. Figalli, F. Maggi, A. Pratelli, A refined Brunn-Minkowski inequality for convex sets. Ann. Inst. H. Poincar´e Anal. Non Lin´eaires 26, pp. 2511-2519 (2009). [GM] O. Guédon, E. Milman, Interpolating thin-shell and sharp large-deviation estimates for isotropic log-concave measures. Geometric And Functional Anal- ysis 21, no. 5, pp. 1043-1068 (2011). [K-L-S] R. Kannan, L. Lovasz, M. Simonovits, Isoperimetric Problems for Convex Bodies and a Localization Lemma. Discrete Comput. Geom., 13, no. 3-4, pp. 541-559 (1995). [Kl] B. Klartag, Marginals of Geometric Inequalities. Geometric Aspects of Func- tional Analysis, Lecture Notes in Math. 1910, Springer, pp. 133-166 (2007). [Le] M. Ledoux, Spectral gap, logarithmic Sobolev constant, and geometric bounds. Surveys in differential geometry. Vol. 9, pp. 219-240. Int. Press, Somerville, MA (2004). 17 [Mc1] R.J. McCann, Existence and uniqueness of monotone measure-preserving maps. Duke Math. J., Vol. 80, pp. 309-323 (1995). [Mc2] R.J. McCann, A convex principle for interacting gases. Adv. Math., Vol. 128, pp. 153-179 (1997). [M] E. Milman, On the role of Convexity in Isoperimetric, Spectral Gap and Con- centration. Invent. Math. 177, no. 1, pp. 1-43 (2009). [M-R] V.D. Milman, L. Rotem, Mixed integrals and related inequalities. J. Funct. Anal. 264, no. 2, pp. 570-604 (2013). [M-S] V.D. Milman, G. Schechtman, Asymptotic Theory of Finite-Dimensional Normed Spaces. Springer-Verlag, Lecture Notes in Mathematics 1200, 156pp. (1986); 2nd edition (2002). [P] G. Paouris, Concentration of mass on convex bodies. Geometric and Functional Analysis 16, pp. 1021-1049 (2006). [R] R. T. Rockafellar, Convex Analysis. Princeton Mathematical Series, no. 28, Princeton University Press, Princeton, NJ (1970). [S] A. Segal, Remark on Stability of Brunn-Minkowski and Isoperimetric Inequali- ties for Convex Bodies. Geometric Aspects of Functional Analysis Volume 2050 of the series Lecture Notes in Mathematics, pp. 381-391 (2012). [V] C. Villani, Optimal Transportation. Old and new. Springer-Verlag, Berlin (2009). Erik Thomas Institut de Math´ematiques de Jussieu, Universit´e Pierre et Marie Curie - Paris 6, 75252 Paris Cedex 05, France [email protected] 18
1912.10722
1
1912
2019-12-23T10:39:25
Convergence of the summation-integral type operators via statistically
[ "math.FA" ]
Our main aim is to investigate the approximation properties for the summation integral type operators in a statistical sense. In this regard, we prove the statistical convergence theorem using well known Korovkin theorem and the degree of approximation is determined. Also using weight function, the weighted statistical convergence theorem with the help of Korovkin theorem is obtained. The statistical rate of convergence in the terms of modulus of continuity and function belonging to the Lipschitz class is obtained. To support the convergence results of the proposed operators to the function, graphical representations take place and a comparison is shown with Sz\'asz-Mirakjan-Kantorovich operators through examples. The last section deals with, a bivariate extension of the proposed operators to study the rate of convergence for the function of two variables, additionally, the convergence of the bivariate operators is shown graphically.
math.FA
math
Convergence of the summation-integral type operators via statistically Rishikesh Yadav1,†, Ramakanta Meher1,⋆, Vishnu Narayan Mishra2,⊛ 1Applied Mathematics and Humanities Department, Sardar Vallabhbhai National Institute of Technology 2Department of Mathematics, Indira Gandhi National Tribal University, Lalpur, Amarkantak-484 887, Surat, Surat-395 007 (Gujarat), India. †[email protected], ⋆meher [email protected], ⊛[email protected] Anuppur, Madhya Pradesh, India Abstract. Our main aim is to investigate the approximation properties for the summation integral type operators in statistical sense. In this regard, we prove the statistical convergence theorem using well known Korovkin theorem and the degree of approximation is determined. Also using weight function, the weighted statistical convergence theorem with the help of Korovkin theorem is obtained. The statistical rate of convergence in the terms of modulus of continuity and function belonging to the Lipschitz class are obtained. To support the convergence results of the proposed operators to the function, graphical representations take place and a comparison is shown with Sz´asz-Mirakjan-Kantorovich operators through examples. The last section deals with, a bivariate extension of the proposed operators to study the rate of convergence for the function of two variables, additionally, convergence of the bivariate operators is shown with graphically. MSC 2010: 41A25, 41A35, 41A36. Keywords: Sz´asz-Mirakjan operators, Sz´asz-Mirakjan-Kantorovich operators, Korovkin-type approxi- mation results, statistical convergence, modulus of continuity, weighted modulus of continuity. 1. Introduction First of all, Fast [10] introduced statistical convergence and further investigated by Steinhaus [23], in that order Schoenberg [22] reintroduced as well as gave some basics properties and studied the summability theory of the statistical convergence. Nowadays, statistical convergence has become an area, which is broad and also very active, even though it has been introduced over fifty years ago and this is being used very frequently in many areas, we refer to some citations as [4 -- 6, 8, 16, 25]. Also, this area is being concerned to investigate the approximation properties of quantum calculus. For instance, statistical approximation properties are used for studying in various papers [1,3,9,14,17,20,21]. These mentioned operators were investigated via statistical convergence as well. In 2003, Duman [7], studied the A-statistical convergence of the linear positive operators for function belonging to the space of all 2π-periodic and continuous functions on whole real line, where A represents non-negative regular summability matrix. Since the statistical type of convergence has not been examined so far in the theory of approximation. So the main objective of this paper is to investigate the theory of approximation by considering the statistical convergence. Here the Korovkin theory is considered to deals with the approximation of function f by the operators { S∗ n,a(f ; x)} [19] for the summation integral type operators, which are as follows. k+1 (1.1) S∗ n,a(f ; x) = n sa n(x) ∞ Xk=0 n Zk n f (u) du, f ∈ C[0, ∞), ∀ x ∈ [0, ∞), n ∈ N. where, f ∈ C[0, ∞), for all [0, ∞) and n ∈ N, and studied the Sz´asz-Mirakjan- Kantorovich type operators along with its rate of convergence in the sense of local approximation results with the help of modulus of smoothness, second order modulus of continuity, Peetres K-functional and functions belonging to the 1 2 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY Lipschitz class. Further, for computing the order of approximation of the operators, we discuss the weighted approximation properties by using the weighted modulus of continuity and prove the theorem. Here it is considered some Lemmas those will be useful in the further study of theorem. Consider the function ei = xi, i = 0 · · · 4, we yield a lemma. Lemma 1.1. [19] For each x ∈ [0, ∞) and a > 1 fixed, we have 1. S∗ 2. S∗ n,a(e0; x) = 1, 1 2n n,a(e1; x) = + 3. S∗ n,a(e2; x) = 4. S∗ n,a(e3; x) = 1 3n2 + 1 4n3 + x log a , + 1 1 2x log a (cid:16)−1 + a (cid:16)−1 + a (cid:16)−1 + a n(cid:17) n n(cid:17) n2 n(cid:17) n3 x log a 7 2 1 x2(log a)2 , 1 n2 x2(log a)2 n(cid:17)2 (cid:16)−1 + a n(cid:17)2 (cid:16)−1 + a 9 2 1 + To find central moments for the defined operators, we have Lemma 1.2. [19] For each x ≥ 0, we have 1. S∗ n,a(ξx(u); x) = − (−1 + 2nx) 2n + x log a n(−1 + a 1 n ) , + n3 x3(log a)3 (cid:16)−1 + a 1 n(cid:17)3 . n3 2. S∗ n,a(ξ2 x(u); x) = (1 − 3nx + 3n2x2) 3n2 − 2(−1 + a 1 n )(−1 + nx)x log a 3. S∗ n,a(ξ3 x(u); x) = − 4n3 (−1 + 4nx − 6n2x2 + 4n3x3) x(7 − 12nx + 6n2x2) log a + x2(log a)2 (cid:16)−1 + a 1 n(cid:17)2 , n2 (cid:16)−1 + a + 1 n2 n(cid:17)2 2(cid:16)−1 + a , 1 n(cid:17) n3 4. S∗ n,a(ξ4 x(u); x) = (1 − 5nx + 10n2x2 − 10n3x3 + 5n4x4) 3x2(−3 + 2nx)(log a)2 + 4x3(log a)3 − 1 1 1 1 n3 n(cid:17)4 n(cid:17)2 2(cid:16)−1 + a n4 (cid:16)−1 + a n(cid:17)4 5(cid:16)−1 + a n(cid:17)3 −10(cid:16)−1 + a n(cid:17)2 +15(cid:16)−1 + a n(cid:17) x3(−2 + nx)(log a)3 + 5x4(log a)4!. −20(cid:16)−1 + a x(−3 + 7nx − 6n2x2 + 2n3x3) log a x2(5 − 6nx + 2n2x2)(log a)2 1 1 1 where, ξx(t) = (u − x)i, i = 1, 2, 3 · · · 2. Korovkin and Weierstrass type statistical theorem If {On(f ; x)} is a sequence of linear positive operators such that the sequence {On(1; x)}, {On(t; x)}, {On(t2; x)} converge uniformly to 1, x, x2 respectively in the defined interval [a, b], then it implies that the sequence {On(f ; x)} converges to the function f uniformly provided f is bounded and continuous in the interval [a, b]. Before proceeding to statistical convergence, here a brief concept of statistical convergence is considered. Definition 2.1. Consider a set K ⊆ N, such that Kn = {k ∈ K : k ≤ n}, where n ∈ N. Then the natural density d(K) of a set K is defined as (2.1) lim n→∞ 1 n Kn, provided the limit exists. Here Kn represents the cardinality of the set Kn. CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 3 Definition 2.2. Let p ∈ R, a sequence {xn} is said to be convergent statistically to p, if for each ǫ > 0, we have (2.2) i.e. (2.3) d({n ≤ k : xn − p ≥ ǫ}) = 0 1 n lim n {n ≤ k : xn − p ≥ ǫ} = 0, or, xn − p < ǫ, for almost all k. By using the properties of statistical convergence, here we shall prove the Korovkin theorem and the Weierstrass type approximation theorem. In [15], it is proved that the classical Korovkin theorem, according to that theorem, let An(f ; x) be linear positive operators defined on the set of all continuous and bounded function CB[a, b] to C[a, b], be set of all continuous function defined on [a, b] for which, the condition kAn(ei; x) − eikC[a,b] = 0, where ei = xi, i = 0, 1, 2, lim n satisfy, then for any function f ∈ C[a, b], lim n kSn(f ; x) − f (x)kC[a,b] = 0, as n → ∞. Theorem 2.3. [11] Let Pn be a sequence of positive linear operators defined on CB[a, b] to C[a, b] and if it satisfies the conditions st − lim n kPn(ei; x) − eikC[a,b] = 0, where i = 0, 1, 2, then for each function f ∈ CB[a, b], we have st − lim n kPn(f ; x) − f (x)kC[a,b] = 0. With these correlation, we have Theorem 2.4. Let { S∗ n,a} be the sequence defined by (1.1), then for every f ∈ CB[0, l], l > 0 such that st − lim n k S∗ n,a(f ; x) − f (x)k = 0, where CB[0, l] is the space of all continuous and bounded function defined on [0, l] with the norm kf k = sup 0≤x≤l f (x) Proof. By (1) of Lemma 1.1, we easily get k S∗ (2.4) st − lim n n,a(e0; x) − e0k = 0 Now by (2) of Lemma (1.1), we have 1 2n Sn(e1; x) − xk = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ (cid:12)(cid:12)(cid:12)(cid:12) 2n(cid:12)(cid:12)(cid:12)(cid:12) 1 define the sets, for any ǫ > 0 as: + 1 x log a − x(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n(cid:17) n (cid:16)−1 + a log a − 1  +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:16)−1 + a n(cid:17) n   1 1 , l(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) and O = {n : k S∗ n,a(e1; x) − xk ≥ ǫ} 1 2n O′ = (cid:26)n : O′′ = (cid:26)n :  ≥ ǫ 2(cid:27) log a (cid:16)−1 + a 1 n(cid:17) n − 1  l ≥ ǫ 2(cid:27), 4 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY and O ⊆ O′ ∪ O′′ and it can be expressed as dnn ≤ k : k S ∗ n,a(e1; x) − xk ≥ ǫ o ≤ d(cid:26)n ≤ k : 1 2n ≥ ǫ 2 (cid:27) n ≤ k :  +d   log a 1 (cid:16)−1 + a n(cid:17) n l ≥ − 1  ǫ 2  (2.5) But since st − lim n (cid:18) 1 n  2n(cid:19) = 0 and st − lim  log a (cid:16)−1 + a 1 n(cid:17) n − 1  = 0, hence, by Inequality 2.5, it follows (2.6) Similarly, st − lim n kSn(e1; x) − xk = 0. k S∗ 1 3n2 + n,a(e2; x) − e2k = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 3n2(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:12)(cid:12)(cid:12)(cid:12) ≤ m2  1 1 (cid:16)−1 + a +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:16)−1 + a 1 3n2 + 2x log a 1 1 1 + n2 n2 (log a)2 x2(log a)2 l +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:16)−1 + a   +(cid:16) (cid:16)−1 + a − x2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n(cid:17)2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) − 1  n(cid:17)2 (cid:16)−1 + a − 1(cid:17)  n(cid:17)2  (log a)2 − 1   n(cid:17)2 (log a)2 (cid:16)−1 + a 2 log a 2 log a n(cid:17) n2 n(cid:17) n2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n(cid:17) n2 (cid:16)−1 + a +  n(cid:17) n2 n2 n2 1 , 1 1 1 l2 . where m2 = max{1, l, l2}, i.e. (2.7) k S∗ n,a(e2; x) − e2k ≤ m2  1 3n2 + 2 log a 1 (cid:16)−1 + a So again by defining the following sets and for any ǫ > 0, one can find (2.8) (2.9) (2.10) (2.11) where P ⊆ P1 ∪ P2 ∪ P3, it gives 1 2 ≥ log a 2 log a ǫ 3m2 } P2 = {n : P1 = {n : P = {n : k S∗ n,a(e2; x) − e2k ≥ ǫ} 1 3n2 ≥ (cid:16)−1 + a n :    ǫ 3m2 }, n(cid:17) n2 n(cid:17) n  (cid:16)−1 + a − 1 P3 =    n,a(e2; x) − e2k ≥ ǫ} ≤ d(cid:26)n ≤ k : +d  (cid:16)−1 + a n(cid:17) n2 n ≤ k : 2 log a ≥ 1 1 ≥ d{n ≤ k : k S∗ ǫ 3m2  . ǫ 3m2  1 3n2 ≥ ǫ 3m2(cid:27) CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 5 (2.12) Hence (2.13) where, +d n ≤ k :     log a (cid:16)−1 + a 1 n(cid:17) n  ≥ 2 − 1  ǫ 3m2  st − lim αn = 0 = st − lim βn = st − lim γn, αn = 1 3n2 , βn = So by 2.12 and 2.13, we have (2.14) Hence proved. 2 log a (cid:16)−1 + a 1 n(cid:17) n2 , γn =    log a (cid:16)−1 + a 1 n(cid:17) n  2 . − 1  st − lim k S∗ n,a(e2; x) − e2k = 0 (cid:3) Now there is an example which satisfy Theorem 2.4, but not the classical Korovkin theorem. Example 2.1. Consider a sequence of linear positive operators Tn(f ; x) which are defined on CB[0, l] by n,a be the sequence positive linear operators and un is unbounded statistically n,a, where S∗ Tn(f ; x) = (1 + un) S∗ convergent sequence. Since S∗ n,a is statistically convergent and also un is statistically convergent but not convergent so one can observe that the sequence Tn satisfies the Theorem 2.4, but not the classical Korovkin theorem. Definition 2.5. Let ξn be a sequence that is converges statistically to ξ, having degree β ∈ (0, 1), if for each ǫ > 0, we have In this case, we can write {n ≤ k : ξn − ξ ≥ ǫ} k1−β lim n = 0 ξn − ξ = st − o(k−β), k → ∞. Theorem 2.6. Let { S∗ n,a} be a sequence defined by (1.1) that satisfy the conditions (2.15) (2.16) (2.17) st − lim n→∞ st − lim n→∞ st − lim n→∞ k S∗ n,a(e0; x) − e0k = st − o(n−ζ1 ), k S∗ n,a(e1; x) − e1k = st − o(n−ζ2 ), k S∗ n,a(e2; x) − e2k = st − o(n−ζ3 ), as n → ∞. Then for each f ∈ CB[0, l], we have st − lim n→∞ S∗ n,a(f ; x) − f (x)k = st − o(n−ζ ), as n → ∞, where ζ = min{ζ1, ζ2, ζ3}. Proof. One can write the inequality (2.12) of Theorem 2.4 as: {n ≤ p : k S∗ n,a(e2; x) − e2k ≥ ǫ} p1−ζ 1 ≥ ǫ 2 log a 1 p1−ζ1 p1−ζ ≤ (cid:12)(cid:12)(cid:8)n ≤ p : (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (n ≤ p :  (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  n ≤ p :  3n2 ≥ ǫ p1−ζ1 3m2(cid:9)(cid:12)(cid:12) 3m2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:16)−1+a n(cid:17)n2 3m2 − 1 n(cid:17)n!2 log a  ≥ e (cid:16)−1+a  p1−ζ2 p1−ζ p1−ζ2 p1−ζ3 1 + + By letting ζ = min{ζ1, ζ2, ζ3} and as n → ∞ the desired result can be achieved. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) p1−ζ3 p1−ζ . (cid:3) 6 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 3. Weighted Statistical Convergence Next, we introduce the convergence properties of the given operators (1.1) using Korovkin type theorem , recall from [12, 13], here the weight function is w(x) = 1 + γ2(x), where γ : R+ → R+ is an unbounded strictly increasing continuous function for which, there exist M > 0 and α ∈ (0, 1] such that: Consider w(x) = 1 + x2 be a weighted function and let Bw[0, ∞) be the space defined by: x − y ≤ M γ(x) − γ(y), ∀ x, y ≥ 0. Also the spaces Bw[0, ∞) =(cid:26)f : [0, ∞) → R kf kw = sup x≥0 f (x) w(x) < +∞(cid:27) . Cw[0, ∞) = {f ∈ Bw[0, ∞), f is continuous}, C k w[0, ∞) = (cid:26)f ∈ Cw[0, ∞), lim x→∞ f (x) w(x) = kf < +∞(cid:27) . We can move towards the main theorem by considering the functions from the above defined spaces via statistically. Theorem 3.1. Let S∗ n,af be a linear positive operators defined by (1.1), then for each f ∈ C k w[0, ∞), we have st − lim n→∞ k S∗ n,a(f ; x) − f (x)kw = 0. Proof. Using the Lemma 1.1, we have S∗ n,a(e0; x) = 1, then it is obvious that k S∗ n,a(e0; x) − 1kw = 0. Now we have k S∗ n,a(e1; x) − e1)kw = sup Now for any ǫ > 0, on defining the following sets: 1 1 log a − 1 x   1 + x2 − − 1  1 1 1 + x2 x 1 + x2  1 + x2  x log a 1 1 1 2n 1 1 1 2n 1 2n x log a = sup = sup (cid:16)−1 + a + x≥0  n(cid:17) n  x≥0 1 + x2 +  (cid:16)−1 + a n(cid:17) n 1 + x2 + x≥0   n(cid:17) n (cid:16)−1 + a − 1  +    (cid:16)−1 + a n(cid:17) n n,a(e1; x) − e1)kw ≥ ǫo P = nn : k S∗ P ′ = (cid:26)n : P ′′ =  n :   − 1  ≥ ≤   (cid:16)−1 + a n(cid:17) n 2(cid:27) 1 2n ǫ ≥ log a 1 log a 1 1 2n , ǫ 2  2(cid:27) ǫ where P ⊆ P ′ ∪ P ′′, it follows dnn ≤ m : k S∗ n,a(e1; x) − e1)kw ≥ ǫo ≤ d(cid:26)n ≤ m : 1 2n ≥ CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 7 Right hand side of the above inequality (3.1) is statistically convergent, hence (3.2) Similarly, st − lim n k S∗ n,a(e1; x) − e1)kw = 0 +d n ≤ m :   log a (cid:16)−1 + a 1 n(cid:17) n − 1  ≥ . ǫ 2  k S∗ n,a(e2; x) − e2kw = sup x≥0 1 3n2 +   1 3n2 + 2x log a + n−1 + a 2 log a 1 1 no n2 +  x2(log a)2 1 no2 n−1 + a n(cid:17)2 (log a)2 1 n2 1 1 + x2 n2 − x2   − 1   . Similarly for any ǫ > 0, by defining following define sets (3.1) (3.3) (3.4) (3.5) ≤   1 ǫ ≥ n : 2 log a ǫ 3(cid:27) n(cid:17) n2 (cid:16)−1 + a (cid:16)−1 + a n,a(e2; x) − e2kw ≥ ǫo H = nn : k S∗ H ′ = (cid:26)n : 1 3n2 ≥ H ′′ =  3   H ′′′ =  − 1   d  n ≤ m :  (cid:16)−1 + a n :  (cid:16)−1 + a n(cid:17)2 n(cid:17) n2 − 1  n(cid:17) n2 n(cid:17)2 (cid:16)−1 + a (cid:16)−1 + a d{n ≤ m : k S∗ (log a)2 (log a)2 n ≤ m : 2 log a n2 n2 ≥ ≥ ǫ 3  3  ǫ 1 1 1 d  n,a(e2; x) − e2kw = 0. ≥ , ǫ 3  = 0 = 0. where H ⊆ H ′ ∪ H ′′ ∪ H ′′′, it follows n,a(e2; x) − e2kw ≥ ǫ} = 0 By the above relations (3.3,3.4,3.5), we yield as: k S∗ (3.6) st − lim n→∞ Hence, (3.7) S∗ n,a(f ; x) − f (x)kw ≤ k S∗ n,a(e0; x) − e0)kw + k S∗ n,a(e1; x) − e1)kw + k S∗ n,a(e2; x) − e2)kw, we get, (3.8) st − lim n→∞ k S∗ n,a(f ; x) − f (x)kw ≤ st − lim n→∞ k S∗ n,a(e0; x) − e0)kw + st − lim n→∞ k S∗ n,a(e1; x) − e1)kw which implies that Hence proved. + st − lim n→∞ k S∗ n,a(e2; x) − e2)kw, st − lim n→∞ k S∗ n,a(f ; x) − f (x)kw = 0. (cid:3) 8 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 4. Rate of Statistical Convergence In this section, we shall introduce the order of approximation of the operators by means of the modulus of continuity and function belonging to the Lipschitz class. Let f ∈ CB[0, ∞), the space of all continuous and bounded functions defined on the interval [0, ∞) and for any x ≥ 0, the modulus of continuity of f is defined to be ω(f ; δ) = sup f (u) − f (x), u ∈ [0, ∞). u−x≤δ And for any δ > 0 and each x, u ∈ [0, ∞), we have (4.1) f (u) − f (x) ≤ ω(f ; δ)(cid:18) u − x δ + 1(cid:19) Next theorem deals with error estimation using the modulus of continuity: Theorem 4.1. Let f ∈ CB[0, ∞) be a non-decreasing function then we have where S∗ n,a(f ; x) − f (x) ≤ 2ω(cid:16)f ;pδn,a(cid:17) , x ≥ 0, δn,a = (1 − 3nx + 3n2x2) 3n2 − 2(−1 + a 1 n )(−1 + nx)x log a (cid:16)−1 + a 1 n(cid:17)2 n2 + x2(log a)2 (cid:16)−1 + a 1 n(cid:17)2 n2!. Proof. With the linearity and positivity properties of the defined operators (1.1), it can be expressed as (4.2) S∗ n,a(f ; x) − f (x) ≤ S∗ n,a(f (u) − f (x); x) k+1 = n sa n(x) ∞ Xk=0 n Zk n f (u) − f (x) du By using the inequality (4.1), the above inequality can be written as: k+1 S∗ n,a(f ; x) − f (x) ≤ n sa n(x) ∞ Xk=0 n Zk n ω(f ; δ)(cid:18) u − x δ + 1(cid:19) du k+1 = ω(f ; δ)n1 + n δ ∞ Xk=0 sa n(x) n Zk n u − x duo k+1 ≤ ω(f ; δ)n1 + = ω(f ; δ)n1 + 1 2 (u − x)2 du(cid:17) ∞ n δ(cid:16)(cid:16) Xk=0 δq S∗ 1 n sa n(x) Zk x(u); x)o n n,a(ξ2 By choosing, δ = δn,a, where (4.3) δn,a = (1 − 3nx + 3n2x2) 3n2 − Hence, the required result can be obtained. 2(−1 + a 1 n )(−1 + nx)x log a (cid:16)−1 + a 1 n(cid:17)2 n2 + x2(log a)2 (cid:16)−1 + a 1 n(cid:17)2 n2!. (cid:3) CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 9 Remark 4.1. By using the above equation (4.3), one can find that (4.4) st − lim n δn,a = 0, By (4.1), similarly one can get (4.5) st − lim n ω(f ; δn,a) = 0 and hence the pointwise rate of convergence of the operators S∗ n,a(f ; x) can be obtained. Theorem 4.2. [19] Let f ∈ CB[0, ∞) and if f ∈ LipM(α), α ∈ (0, 1] holds that is the inequality f (u) − f (x) ≤ Mu − xα, u, x ∈ [0, ∞), where M is a positive constant then for every x ≥ 0, we have S∗ n,a(f ; x) − f (x) ≤ Mδ α 2 n,a, where δn,a = S∗ n,a((u − x)2; x). Proof. Since we have f ∈ CB[0, ∞) ∩ LipM(α), so S∗ n,a(f ; x) − f (x) ≤ S∗ n,a(f (u) − f (x); x) k+1 ≤ M S∗ n,a(u − xα; x) = M(cid:16)n sa n(x) ∞ Xk=0 Now by applying Holder inequality with p = 2 α and q = 2 2−α , we have n Zk n u − xα du(cid:17). S∗ n,a(f ; x) − f (x) ≤ M  α 2 n k+1 ∞ Xk=0 sa n(x)n n Zk n (u − x)2 du,o = Mδ n,a. Hence proved. Remark 4.2. Similarly by equation (4.3), we can justify (4.6) st − lim n δn,a = 0, ≤ M( S∗ n,a(ξ2 x(u); x)) α 2  α 2 (cid:3) and it can be seen that the rate of statistical convergence of the operators 1.1 to f (x) are estimated by means of function belonging to the Lipschitz class. 10 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 5. Graphical approach Based upon the defined operator (1.1), now we will show the convergence of the operators, for different functions for particular the values of n. Example 5.1. Let the function f (x) = e−2x and choose the values of n = 5, 10 for which the correspond- ing operators are S∗ 10,a(f ; x) respectively. The approach of the operators is faster to the function as in increment in the value of n (when we choose n = 500, 1000) and we can observe, the error is able to obtain as small as we please as the the value of n is large, which can be seen in the given figures (1). 5,a(f ; x), S∗ 1.0 0.8 0.6 0.4 0.2 Ž * S f x = Exp-2 x n,a f ; x for n = 5 n,a f ; x for n = 10 Ž * S a = 1.5 Ž * S f x = Exp-2 x n,a f ; x for n = 500 n,a f ; x for n = 1000 Ž * S a = 1.5 1.0 0.8 0.6 0.4 0.2 0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0 Figure 1. Convergence of the operators S∗ n,a(exp(−2u); x) to f (x) In fact, in Figure 1, as the value of n increases, the operator S∗ n,a(f ; x) approaches towards the function f (x) = exp(−2x) keeping a = 1.5 fixed. Example 5.2. Here, we consider the function f (x) = x and choose the value of n = 100, 500,the convergence can be seen by graphical representation, given by figure 2. Ž * S f x = x n,a f ; x for n = 100 n,a f ; x for n = 500 Ž * S a = 1.5 1.0 0.8 0.6 0.4 0.2 0.0 0.0 0.2 0.4 0.6 0.8 1.0 Figure 2. Convergence of the operators S∗ n,a(u; x) to f (x) Example 5.3. For the convergence of the proposed operators (1.1) to the function f (x) =(cid:0)x − 1 choose n = 15, 30, 500, 1000 and the errors can be observed by the given figures (3). 2(cid:1)(cid:0)x − 1 3(cid:1)(cid:0)x − 1 4(cid:1), CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 11 0.02 0.01 0.00 -0.01 -0.02 -0.03 -0.04 0.00 -0.01 -0.02 -0.03 -0.04 f x = x - 1 2 x - 1 3 x - 1 4 Ž * S Ž * S n,a f ; x for n = 15 n,a f ; x for n = 30 a = 1.5 1 3 x - 1 4 1 2 x - f x = x - n,a f ; x for n = 500 n,a f ; x for n = 1000 S* S* a = 1.5 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 Figure 3. Convergence of the operators S∗ n,a(f ; x) to f (x) Concluding Remark: Since, we have seen the approach of the proposed operators by all the above figures, we can observe that as the degree of the operators is large, the approximation will be better. Moreover, one can observe by Figure 2, as the value of n is increased, the operators S∗ n,a(f ; x) converge to the function f (x) = x and in Figure 3, it can be seen that the operator S∗ n,a(t; x) converges to the function f (x) =(cid:0)x − 1 2(cid:1)(cid:0)x − 1 3(cid:1)(cid:0)x − 1 4(cid:1) for large value of n. 5.1. A comparison to the Sz´asz-Mirakjan-Kantorovich operators. In 1983, V. Totik [24] In- troduced the Kantorovich variant of the Sz´asz-Mirakjan operators in Lp-spaces for p > 1, which are as follows: k+1 (5.1) Kn(f ; x) = ne−nx (nx)k k! ∞ Xk=0 n Zk n f (u) du. Now we shall show a comparison with the above operators (5.1) to the proposed operators defined by (1.1) by graphical representation and convergence behavior will also be seen to the given function. In Figure 4, one can see that the said operators have a better rate of convergence as compared to the above operators (5.1), but both operators converges to the function f (x). Example 5.4. For the same degree of approximation of the operators S∗ n,a(f ; x) and Kn(f ; x) to the function f (x) = x3, the comparison shown by Figure 4. 0.15 0.10 0.05 0.00 0.0 f x = x3 Ž * n,a f ; x for n = 25 S Kn f ; x for n = 25 0.1 0.2 0.3 0.4 0.5 Figure 4. The comparison of the operators S∗ n,a(f ; x) and Kn(f ; x) 12 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY Example 5.5. Comparison of convergence can be seen for the operators S∗ n,a(f ; x) and Kn(f ; x) to the function in the given Figures 5 by choosing n = 4, 20. 1.00 0.95 0.90 0.85 0.80 0.75 0.70 0.65 f x = 1 x+1 Ž * S n,a f ; x for n = 4, a = 1.5 Kn f ; x for n = 4 1.00 0.95 0.90 0.85 0.80 0.75 0.70 f x = 1 x+1 Ž * S n,a f ; x for n = 20, a = 1.5 Kn f ; x for n = 20 0.0 0.1 0.2 0.3 0.4 0.5 0.0 0.1 0.2 0.3 0.4 0.5 Figure 5. The comparison of the operators S∗ n,a(f ; x) and Kn(f ; x) So, from Figure 5, it can be observed that the summation-integral-type operators (1.1) are approaching more faster than Sz´asz-Mirakjan-Kantorovich operators, but for large value of n, both operators converge to the function f (x) = 1 1+x . Example 5.6. Consider n = 5 and function f (x) = cos πx then the approaching of the operators S∗ n,a(f ; x) is faster than Kn(f ; x). * Sn,a  f ; x for n = 5, a = 1.5 f x = cosΠx Kn f ; x for n = 5 1.0 0.5 0.0 -0.5 -1.0 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 Figure 6. The comparison of the operators S∗ n,a(f ; x) and Kn(f ; x) By observing to Figure 6, it can be found that the better rate of convergence is taking place in the n,a(f ; x) rather than Sz´asz-Mirakjan-Kantorovich operators but ulti- summation-integral-type operators S∗ mately they converge to the given function f (x) = cos(πx). 6. An extension in the sense of bivariate operators To study the approximation properties for the function of two variables, we generalize as an extension to the above operators 1.1 into bivariate operators in the space of integral functions to investigate the rate of convergence with the help of the statistical convergence. Let f : C[0, ∞) × C[0, ∞) → C[0, ∞) × C[0, ∞), and we define the operators with one parameter as follows: (6.1) Y ∗ m,m,a(f ; x, y) = m2 m,m(x, y) = a −x−y −1+a where sa ∞ ∞ Xk2=0 Xk1=0 m ! xk1 yk2 (log a)k1 +k2 1 (−1+a 1 m )k1+k2 k1!k2! sa m,m(x, y) k2 +1 k1+1 m Zk2 m Zk1 m m f (u, v) du dv, = sa m(x) × sa m(y). CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 13 Let us define a function ei,j = xiyj, for all x, y ≥ 0, where i, j ∈ N ∪ {0}. Lemma 6.1. For all x, y ≥ 0, the bivariate operators 6.1, satisfy the following equalities: 1. Y ∗ m,m,a(e00; x, y) = 1 2. Y ∗ m,m,a(e11; x, y) = 3. Y ∗ m,m,a(e22; x, y) = 4. Y ∗ m,m,a(e33; x, y) = 1 n + 2y log a(cid:17)o m )x log +3x2(log)2(cid:19) 1 1 1 1 1 1 1 1 9(−1 + a + 6(−1 + a + 6(−1 + a 4(−1 + a 1 m )m2n(cid:16)−1 + a m )4m4n(cid:18)(cid:16)−1 + a (cid:18)(cid:16)−1 + a m(cid:17)2 m(cid:17)6 16(cid:16)−1 + a +4x3(log a)3(cid:17)(cid:16)(cid:16)−1 + a +4y3(log a)3(cid:17)o. m6n(cid:16)(cid:16)−1 + a m(cid:17)3 n + 2x log a(cid:17)(cid:16)−1 + a m(cid:17)2 m )y log +3y2(log)2(cid:19)o m(cid:17)3 + 14(cid:16)−1 + a m(cid:17)2 + 14(cid:16)−1 + a 1 1 1 1 1 1 x log a + 18(cid:16)−1 + a m(cid:17)2 y log a + 18(cid:16)−1 + a 1 m(cid:17) y2(log a)2 1 m(cid:17) x2(log a)2 Proof. To prove (1) of the above Lemma 6.1, we put f (x, y) = e00 = 1 in the above bivariate operators (6.1) and we have ∞ ∞ 1. Y ∗ m,m,a(e00; x, y) = m2 sa m,m(x, y) Xk1=0 = ∞ Xk1=0 Xk2=0 m(x)! ∞ Xk2=0 sa sa = 1. k2 +1 k1 +1 du dv m Zk1 m m m Zk2 m(y)! 2. Y ∗ m,m,a(e11; x, y) = m2 k2 +1 k1 +1 m m ∞ ∞ m2 m ∞ uv du dv sa m,m(x, y) Xk1=0 Xk2=0 m(x)(cid:18) 1 4 (cid:16) Xk1=0 m2 + m )m2n(cid:16)−1 + a Zk1 Zk2 m (cid:19)(cid:17)(cid:16) m(y)(cid:18) 1 m + 2x log a(cid:17)(cid:16)−1 + a Xk2=0 2k1 sa sa ∞ 1 m 1 1 4(−1 + a = = 2k2 m (cid:19)(cid:17) m2 + m + 2y log a(cid:17)o 1 (cid:3) , m2 Similarly, it can be proved for the other equalities. Lemma 6.2. For all x, y ≥ 0 and m ∈ N, we have 1. Y ∗ m,m,a((u − x); x, y) = − 2. Y ∗ m,m,a((v − y); x, y) = − (−1 + 2mx) 2m + x log a m(−1 + a 1 m ) , (−1 + 2ny) 2n + y log a n(−1 + a 1 n ) , 3. Y ∗ m,m,a((u − x)2; x, y) = (1 − 3mx + 3m2x2) 3m2 − 4. Y ∗ m,m,a((v − y)2; x, y) = (1 − 3ny + 3n2y2) 3n2 − 2(−1 + a 1 m )(−1 + mx)x log a x2(log a)2 2(−1 + a n )(−1 + ny)y log a m2 1 1 m(cid:17)2 (cid:16)−1 + a n(cid:17)2 (cid:16)−1 + a 1 n2 + 1 y2(log a)2 m(cid:17)2 (cid:16)−1 + a n(cid:17)2 (cid:16)−1 + a n2 1 + Proof. One can prove the above all equalities with the help of equalities, which are proved in [19]. So (cid:3) we omit the proof. 14 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 7. Rate of convergence of bivariate operators In this section, we find rate of convergence of the biivariate operators 6.1, for function of two variables. Now, we define the supremum norm, by letting X = [0, ∞) × [0, ∞), we have kf k = sup x,y∈X f (x, y), f ∈ CB(X). Consider the modulus of continuity ω(f ; δ1, δ2) for the bivariate operators 6.1, where δ1, δ2 > 0 and is defined by: (7.1) ω(f ; δ1, δ2) = {sup f (u, v) − f (x, y) : (u, v), (x, y) ∈ X, and u − x ≤ δ1, v − y ≤ δ2}. Lemma 7.1. Let f ∈ CB(X), then for δ1, δ2 > 0, we have the following properties of modulus of continuity: 1. For given function f , ω(f ; δ1, δ2) → 0 as δ1, δ2 → 0. 2. f (u, v) − f (x, y) ≤ ω(f ; δ1, δ2)(cid:16)1 + u−x Fore more details, see [2]. δ1 (cid:17)(cid:16)1 + v−y δ2 (cid:17). Theorem 7.1. If f ∈ CB(X) and x, y ∈ [0, ∞), then we have Y ∗ m,m,a(f ; x, y) − f (x, y) ≤ 4ω(f ;pδm,a,qδ′ m,a), 2(−1 + a 1 m )(−1 + mx)x log a (7.2) where (7.3) (7.4) 3m2 δm,a = (1 − 3mx + 3m2x2) m,a = (1 − 3my + 3m2y2) 3m2 δ′ − − 2(−1 + a m )(−1 + my)y log a 1 (cid:16)−1 + a (cid:16)−1 + a 1 m(cid:17)2 m(cid:17)2 1 m2 m2 + + x2(log a)2 y2(log a)2 (cid:16)−1 + a (cid:16)−1 + a 1 m(cid:17)2 m(cid:17)2 1 m2!, m2!. Proof. By using the linearity and the positivity of the defined operators Y ∗ m,m,a(f ; x, y) (6.1) and applying on (2) of the Lemma (7.1), then for any δ1, δ2 > 0, we have Y ∗ m,m,a(f ; x, y) − f (x, y) ≤ Y ∗ m,m,a(f (t, s) − f (x, y); x, y) ≤ ω(f ; δ1, δ2)(cid:18)1 + ≤ ω(f ; δ1, δ2)(cid:18)1 + 1 δ1 1 Y ∗ m,m,a(u − x; x, y)(cid:19) ×(cid:18)1 + m,m,a((u − x)2; x, y)(cid:1) 1 2(cid:19) ×(cid:18)1 + m,m,a(v − y; x, y)(cid:19) δ2 (cid:0)Y ∗ m,m,a((v − y)2; x, y)(cid:1) 1 (using the Cauchy-Schwarz inequality). 1 δ2 Y ∗ 1 2(cid:19) δ1 (cid:0)Y ∗ Next one step will complete the proof. (cid:3) At last, we shall see the rate of convergence of the bivariate operators (6.1) in the sense of functions belonging to the Lipschitz class LipM(α1, α2), where α1, α2 ∈ (0, 1] and M ≥ 0 is any constant and is defined by: (7.5) f (u, v) − f (x, y) ≤ Mu − xα1 v − yα2 , ∀ x, y, u, v ∈ [0, ∞). Our next approach to prove the theorem for finding the rate of convergence, when the function is belonging to the Lipschitz class. Theorem 7.2. Let f ∈ LipM(α1, α2) then for each f ∈ CB(X), we have where δm,a and δ′ m,m,a(f ; x, y) − f (x, y) ≤ Mδ m,a are defined by (7.3) and (7.4) respectively. Y ∗ α1 m,aδ 2 ′ α2 m,a, 2 Proof. Since defined bivariate operators Y ∗ m,m,a(f ; x, y) are linear positive and also f ∈ LipM(α1, α2), where α1, α2 ∈ (0, 1], then we have Y ∗ m,m,a(f ; x, y) − f (x, y) ≤ Y ∗ m,m,a(f (t, s) − f (x, y); x, y) m,m,a(Mu − xα1 v − yα2; x, y) ≤ Y ∗ CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 15 = MY ∗ m,m,a(u − xα1 ; x, y) × Y ∗ m,m,a(v − yα2; x, y) Applying the Holder inequality with p ′ = 2 α1 , q ′ = 2 2−α1 and p ′′ = 2 α2 , q ′′ = 2 2−α2 we have Y ∗ m,m,a(f ; x, y) − f (x, y) ≤ M(Y ∗ m,m,a(u − x)2; x, y)) ′ α2 m,a. α1 m,aδ 2 = Mδ 2 α1 2 × (Y ∗ m,m,a(v − y)2; x, y)) α2 2 Thus the proof is completed. (cid:3) 7.1. Graphical approach of bivariate operators. Now we shall see that, the convergence of the bivaiate operators defined by (6.1) to the function f (x, y) will be presented by graphical representation. Example 7.1. Let f ∈ C(X) and choose m = 5 10, 20, a = 3 (fixed), the convergence of Y ∗ m,m,a(f ; x, y) to the function f (x, y) (blue) takes place and is illustrated in Figure 7. For the different values of m, the corresponding operators Y ∗ 20,20,a(f ; x, y) represent red, green and magenta colors respectively. 10,10,a(f ; x, y) and Y ∗ 5,5,a(f ; x, y), Y ∗ Figure 7. Convergence of the operators Y ∗ m,m,a(f ; x) to f (x, y) 16 CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY Same here for, same function f (x, y), but for large value of m = 100, 500 and then corresponding 500,500,a(f ; x, y) (green) almost overlap to the function f (x, y) 100,100,a(f ; x, y) (red) and Y ∗ operators are Y ∗ (blue), which is illustrated in Figure 8. Figure 8. Convergence of the operators Y ∗ m,m,a(f ; x) to f (x, y) Concluding remark: The convergence of the bivariate operators Y ∗ m,m,a(f ; x) to the function f (x, y) is taking place as if we increase the value of m, i.e. for the large value of m, the bivariate operators converge to the function. Conclusion: Convergence of the proposed operators 1.1 via statistical sense and order of approximation have been determined, moreover the weighted statistical convergence properties and the rate of statistical convergence have been investigated in some sense of local approximation results with the help of modulus of continuity. To support the approximation results, the graphical representations took place and along with to stable of the proposed operators (1.1), a comparison has been shown and obtained the better rate of convergence. An extension is got for study of the rate of convergence in bivariate sense and graphical analysis has been taken place. References [1] Aral A, Dogru O. Bleimann, Butzer, and Hahn operators based on the-integers. Journal of Inequalities and Applications. 2008 Dec 1;2007(1):079410. [2] Anastassiou GA, Gal SG. Approximation Theory. Moduli of Continuity and Global Smoothness Preservation, Birkha user, Boston, 2000. [3] Aktuglu H, Ozarslan MA, Duman O. Matrix summability methods on the approximation of multivariate q-MKZ operators. Bull. Malays. Math. Sci. Soc.(2). 2011 Jan 1;34(3):465-74. [4] Connor J, Kline J. On statistical limit points and the consistency of statistical convergence. Journal of mathematical analysis and applications. 1996 Jan 1;197(2):392-9. [5] Connor J, Swardson MA. Strong integral summability and the Stone- Cech compactification of the half-line. Pacific Journal of Mathematics. 1993 Feb 1;157(2):201-24. [6] Connor J, Ganichev M, Kadets V. A characterization of Banach spaces with separable duals via weak statistical convergence. Journal of Mathematical Analysis and Applications. 2000 Apr 1;244(1):251-61. [7] Duman O. Statistical approximation for periodic functions. Demonstratio Mathematica. 2003 Oct 1;36(4):873-8. [8] Duman O, Khan MK, Orhan C. A-statistical convergence of approximating operators. Mathematical Inequalities and applications. 2003 Oct 1;6:689-700. [9] Ersan S, Dogru O. Statistical approximation properties of q-Bleimann, Butzer and Hahn operators. Mathematical and Computer Modelling. 2009 Apr 1;49(7-8):1595-606. [10] Fast H. Sur la convergence statistique. InColloquium Mathematicae 1951 (Vol. 2, No. 3-4, pp. 241-244). [11] Gadjiev AD, Orhan C. Some approximation theorems via statistical convergence. The Rocky Mountain Journal of Math- ematics. 2002 Apr 1:129-38. [12] Gadzhiev AD. The convergence problem for a sequence of positive linear operators on unbounded sets, and theorems analogous to that of PP Korovkin. InDoklady Akademii Nauk 1974 (Vol. 218, No. 5, pp. 1001-1004). Russian Academy of Sciences. [13] Gadjiev AD, On PP. Korovkin type theorems, Math. Zametki 20 (1976), no. 5, 781786. Math. Notes. 1976;20:5-6. [14] Gupta V, Radu C. Statistical approximation properties of q-Baskakov-Kantorovich operators. Open Mathematics. 2009 Dec 1;7(4):809-18. [15] Korovkin PP. Linear operators and the theory of approximation. India, Delhi. 1960. CONVERGENCE OF THE SUMMATION-INTEGRAL TYPE OPERATORS VIA STATISTICALLY 17 [16] Maddox IJ. Statistical convergence in a locally convex space. In Mathematical Proceedings of the Cambridge Philosophical Society 1988 Jul (Vol. 104, No. 1, pp. 141-145). Cambridge University Press. [17] Mahmudov N, Sabancigil P. A q-analogue of the Meyer-K¯onig and Zeller operators. Bull. Malays. Math. Sci. Soc.(2). 2012 Jan 1;35(1):39-51. [18] Miller HI. A measure theoretical subsequence characterization of statistical convergence. Transactions of the American Mathematical Society. 1995;347(5):1811-9. [19] Mishra VN, Yadav R. Some estimations of summation-integral-type operators. Tbilisi Mathematical Journal. 2018;11(3):175-91. [20] Orkcu M, Dogru O. Weighted statistical approximation by Kantorovich type q-Sz´asz-Mirakjan operators. Applied Mathe- matics and Computation. 2011 Jun 15;217(20):7913-9. [21] Radu C. Statistical approximation properties of Kantorovich operators based on q-integers, Creat. Math. Inform. 2008;17(2):75-84. [22] Schoenberg IJ. The integrability of certain functions and related summability methods. The American Mathematical Monthly. 1959 May 1;66(5):361-775. [23] Steinhaus H. Sur la convergence ordinaire et la convergence asymptotique. In Colloq. Math 1951 (Vol. 2, No. 1, pp. 73-74). [24] Totik V. Approximation by Sz´asz-Mirakjan-Kantorovich operators in Lp, (p > 1). Analysis Mathematica. 1983 Jun 1;9(2):147-67. [25] Zygmund A. Trigonometric Series. Vol. I, II. Reprint of the 1979 edition. Cambridge Mathematical Library. Cambridge Univ. Press, Cambridge. 1988;1:95.
1103.2349
1
1103
2011-03-11T19:20:30
A non-type (D) operator in c0
[ "math.FA" ]
Previous examples of non-type (D) maximal monotone operators were restricted to $\ell^1$, $L^1$, and Banach spaces containing isometriccopies of these spaces. This fact led to the conjecture that non-type (D) operators were restricted to this class of Banach spaces. We present a linear non-type (D) operator in $c_0$.
math.FA
math
A non-type (D) operator in c0 Orestes Bueno∗ B. F. Svaiter† November 9, 2018 Abstract Previous examples of non-type (D) maximal monotone operators were restricted to ℓ1, L1, and Banach spaces containing isometric copies of these spaces. This fact led to the conjecture that non-type (D) operators were restricted to this class of Banach spaces. We present a linear non-type (D) operator in c0. keywords: maximal monotone, type (D), Banach space, extension, bidual. 1 Introduction Let U, V arbitrary sets. A point-to-set (or multivalued) operator T : U ⇒ V is a map T : U → P(V ), where P(V ) is the power set of V . Given T : U ⇒ V , the graph of T is the set Gr(T ) := {(u, v) ∈ U × V v ∈ T (u)}, the domain and the range of T are, respectively, dom(T ) := {u ∈ U T (u) 6= ∅}, R(T ) := {v ∈ V ∃u ∈ U, v ∈ T (u)} and the inverse of T is the point-to-set operator T −1 : V ⇒ U, T −1(v) = {u ∈ U v ∈ T (u)}. A point-to-set operator T : U ⇒ V is called point-to-point if for every u ∈ dom(T ), T (u) has only one element. Trivially, a point-to-point operator is injective if, and only if, its inverse is also point-to-point. ∗Instituto de Mat´ematica Pura e Aplicada (IMPA), Estrada Dona Castorina 110, Rio de Janeiro, RJ, CEP 22460-320, Brazil, [email protected]. The work of this author was partially supported by CAPES †Instituto de Mat´ematica Pura e Aplicada, (IMPA), Estrada Dona Castorina 110, Rio de Janeiro, RJ, CEP 22460-320, Brazil, [email protected]. The work of this author was partially supported by CNPq grants no. 474944/2010-7, 303583/2008-8 and FAPERJ grant E-26/110.821/2008. 1 Let X be a real Banach space. We use the notation X ∗ for the topological dual of X. From now on X is identified with its canonical injection into X ∗∗ = (X ∗)∗ and the duality product in X × X ∗ will be denoted by h·, ·i, hx, x∗i = hx∗, xi = x∗(x), x ∈ X, x∗ ∈ X ∗. A point-to-set operator T : X ⇒ X ∗ (respectively T : X ∗∗ ⇒ X ∗) is monotone, if hx − y, x∗ − y ∗i ≥ 0, ∀(x, x∗), (y, y ∗) ∈ Gr(T ), (resp. hx∗ − y ∗, x∗∗ − y ∗∗i ≥ 0, ∀(x∗∗, x∗), (y ∗∗, y ∗) ∈ Gr(T )), and it is maximal monotone if it is monotone and maximal in the family of monotone operators in X × X ∗ (resp. X ∗∗ × X ∗) with respect to the order of inclusion of the graphs. We denote c0 as the space of real sequences converging to 0 and ℓ∞ as the space of real bounded sequences, both endowed with the sup-norm and ℓ1 as the space of absolutely summable real sequences, endowed with the 1-norm, k(xk)kk∞ = sup k∈N xk, The dual of c0 is identified with ℓ1 in the following sense: for y ∈ ℓ1 k(xk)kk1 = xk. ∞Xk=1 y(x) = hx, yi = xiyi, ∀x ∈ c0. ∞Xi=1 Likewise, the dual of ℓ1 is identified with ℓ∞. It is well known that c0 (as well as ℓ1, ℓ∞, etc.) is a non-reflexive Banach space. Let X be a non-reflexive real Banach space and T : X ⇒ X ∗ be maximal monotone. Since X ⊂ X ∗∗, the point-to-set operator T can also be regarded as an operator from X ∗∗ to X ∗. We Gr(bT ) = Gr(T ). denote bT : X ∗∗ ⇒ X ∗ as the operator such that If T : X ⇒ X ∗ is maximal monotone then bT is (still) trivially monotone but, in general, not maximal monotone. Direct use of the Zorn's Lemma shows that bT has a maximal monotone tone operator T : X ⇒ X ∗, is the point-to-set operator eT : X ∗∗ ⇒ X ∗ whose graph Gr(cid:16)eT(cid:17) is extension. So it is natural to ask if such maximal monotone extension to the bidual is unique. Gossez [4, 5, 6, 7] gave a sufficient condition for uniqueness of such an extension. Definition 1.1 ([4]). Gossez's monotone closure (with respect to X ∗∗ × X ∗) of a maximal mono- given by Gr(cid:16)eT(cid:17) = {(x∗∗, x∗) ∈ X ∗∗ × X ∗ hx∗ − y ∗, x∗∗ − yi ≥ 0, ∀(y, y ∗) ∈ T }. 2 A maximal monotone operator T : X ⇒ X ∗, is of Gossez type (D) if for any (x∗∗, x∗) ∈ in Gr(T ) which converges to (x∗∗, x∗) in the σ(X ∗∗, X ∗)×strong topology of X ∗∗ × X ∗. i )(cid:19)i∈I Gr(cid:16)eT(cid:17) , there exists a bounded net (cid:18)(xi, x∗ maximal monotone extension to the bidual, namely, its Gossez's monotone closure eT : X ∗∗ ⇒ X ∗. Beside this fact, maximal monotone operators of type (D) share many properties with maximal monotone operators defined in reflexive Banach spaces as, for example, convexity of the closure of the domain and convexity of the closure of the range [4]. Gossez proved [7] that a maximal monotone operator T : X ⇒ X ∗ of of type (D) has unique Gossez gave an example of a non-type (D) operator on ℓ1 [5]. Later, Fitzpatrick and Phelps gave an example of a non-type (D) on L1[0, 1] [3]. These examples led Professor J. M. Borwein to define Banach spaces of type (D) as those Banach spaces where every maximal monotone operator is of type (D), and to formulate the following most interesting conjecture [1, §4, question 3]: • Are any nonreflexive spaces X of type (D)? That is, are there nonreflexive spaces on which all maximal monotones on X are type (D). I conjecture 'weakly' that if X contains no copy of ℓ1(N) then X is type (D) as would hold in X = c0. In this work, we answer negatively such conjecture by giving an example of a non-type (D) operator on c0 and proving that, for every space which contains a isometric copy of c0, a non- type (D) operator can be defined. 2 A non-type (D) operator on c0 Gossez's operator [5] G : ℓ1 → ℓ∞ is defined as G(x) = y, yn = xi − ∞Xi=n+1 xi, n−1Xi=1 (1) eq:G which is linear, continuous, anti-symmetric and, therefore, maximal monotone. Gossez's operator is also injective (see the proof of [2, Proposition 3.2]). This operator will be used to define a non-type (D) maximal monotone operator in c0. lem:1 Lemma 2.1. The operator T : c0 ⇒ ℓ1, T (x) = {y ∈ ℓ1 − G(y) = x} (2) eq:op is point-to-point in its domain, is maximal monotone and its range is yi = 0) . ∞Xi=1 R(T ) =(y ∈ ℓ1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 3 (3) eq:range Proof. Since G is injective, T is point-to-point in its domain. Moreover, direct use of (1) shows that G is linear and hy, G(y)i = 0 for any y ∈ ℓ1, which proves that T is also linear and monotone. Using (1) we conclude that for any y ∈ ℓ1, lim n→∞ (G(y))i = − yi ∞Xi=1 which proves (3). Suppose that x ∈ c0, y ∈ ℓ1 and hx − x′, y − y ′i ≥ 0, ∀(x′, y ′) ∈ Gr(T ). (4) eq:mon Define, u1 = (−1, 1, 0, 0, . . . ), u2 = (0, −1, 1, 0, 0), . . . , that is i = 1, 2, . . . (5) eq:um and let i = m i = m + 1 otherwise −1, 1, 0, (um)i = (vm)i =(1, i = m or i = m + 1 0, otherwise vm = G(um), i = 1, 2, . . . (6) eq:vm where the expression of (vm)i follows from (1) and (5). Direct use of (5), (6) and (2) shows that T (−λvm) = λum for λ ∈ R and m = 1, 2, . . . . Therefore, for any λ ∈ R, m = 1, 2, . . . which is equivalent to hx + λvm, y − λumi ≥ 0 hx, yi + λ[hvm, yi − hx, umi] ≥ 0. Since the above inequality holds for any λ, it holds that hx, umi = hvm, yi, m = 1, 2, . . . which, in view of (5), (6) is equivalent to xm+1 − xm = ym+1 + ym, m = 1, 2, . . . (7) eq:ufa Adding the above equality for m = i, i + 1, . . . , j we conclude that yk + yj+1, i < j. xj+1 = xi + yi + 2 jXk=i+1 4 Using the assumptions x ∈ c0, y ∈ ℓ1 and taking the limit j → ∞ in the above equation, we conclude that Also, using the above relation between x and y in (4) with x′ = 0, y ′ = 0 we obtain xi = −"yi + 2 yk# . ∞Xk=1 yk# = −"G(y)i + ∞Xk=i+1 −" ∞Xk=1 yk#2 ≥ 0. Combining the two above equations we conclude that x = −G(y). Hence (x, y) ∈ Gr(T ), which proves the maximal monotonicity of T . pro:nond Proposition 2.2. The operator T : c0 ⇒ ℓ1 defined in Lemma 2.1 has infinitely many maximal monotone extensions to ℓ∞ ⇒ ℓ1. In particular, T is non-type (D). Proof. Let We claim that e = (1, 1, 1, . . . ) h−G(y) + αe − x′, y − y ′i = αhy, ei, ∀(x′, y ′) ∈ Gr(T ), y ∈ ℓ1. (8) eq:mp To prove this claim, first use (2) and (1) to conclude that x′ = −G(y ′) and h−G(y) − x′, y − y ′i = hG(y ′ − y), y − y ′i = 0 As y ′ ∈ R(T ), using (3) we have he, y ′i = 0, which combined with the above equation yields (8). Takeey ∈ ℓ1 such that hey, ei > 0 and define xτ = −G(τey) + ( xτ , τey ) ∈ eT , 1 τ In view of (8), e, 0 < τ < ∞. 0 < τ < ∞. Therefore, for each τ ∈ (0, ∞) there exists a maximal monotone extension Tτ : ℓ∞ ⇒ ℓ1 of T such that However, these extensions are distinct because if τ, τ ′ ∈ (0, ∞) and τ 6= τ ′ then (xτ , τey) ∈ G(Tτ ). hxτ − xτ ′ , τey − τ ′eyi = (τ − τ ′)(1/τ − 1/τ ′)hey, ei < 0. teo:nond Theorem 2.3. Let X be a Banach space such that there exists a non-type (D) maximal monotone operator T : X ⇒ X ∗, and let Ω be another Banach space which contains an isometric copy of X. Then there exists a non-type (D) maximal monotone operator S : Ω ⇒ Ω∗. 5 Proof. We can identify X as a closed subspace of Ω. Define S : Ω ⇒ Ω∗ as (w, w∗) ∈ Gr(S) ⇐⇒ w ∈ X and (w, w∗X ) ∈ Gr(T ). This implies that for every v∗ ∈ Ω∗ such that v∗X ≡ 0 then (w, w∗ + tv∗) ∈ Gr(S), ∀ (w, w∗) ∈ Gr(S), t ∈ R, (9) eq:nrm Take (w, w∗), (z, z∗) ∈ Gr(S), then w, z ∈ X and w − z ∈ X. Thus, since T is monotone and (w, w∗X), (z, z∗X) ∈ Gr(T ), hw − z, w∗ − z∗i = hw − z, w∗X − z∗Xi ≥ 0. Now we prove that S is maximal monotone. Let (z, z∗) ∈ Ω × Ω∗ such that hw − z, w∗ − z∗i ≥ 0, ∀ (w, w∗) ∈ Gr(S), (10) eq:mnrl and assume that z /∈ X. Then, by the Hahn-Banach Theorem, there exists v∗ ∈ Ω∗ such that vX ≡ 0 and hz, v∗i = 1. Using (9), we obtain hw − z, w∗ + tv∗ − z∗i ≥ 0, ∀ (w, w∗) ∈ Gr(S), t ∈ R, which is equivalent to hw − z, w∗ − z∗i ≥ thz − w, v∗i = t, ∀ (w, w∗) ∈ Gr(S), t ∈ R, which leads to a contradiction if we let t → +∞. Hence z ∈ X. Now, using (10) and the maximal monotonicity of T , we conclude that (z, z∗X) ∈ Gr(T ). Therefore (z, z∗) ∈ Gr(S) and S is maximal monotone. Finally, we use [8, eq. (5) and Theorem 4.4, item 2] to prove that S is non-type (D). As T is non-type (D) on X × X ∗ then there exists (x∗, x∗∗) ∈ X ∗ × X ∗∗ such that sup hy, x∗ − y ∗i + hy ∗, x∗∗i < hx∗, x∗∗i. (y,y∗)∈Gr(T ) Define z∗∗ : X ∗ → R as hz∗, z∗∗i = hz∗X, x∗∗i and let z∗ be any continuous extension of x∗ from X to Ω. Take (w, w∗) ∈ Gr(S), then w ∈ X, (w, w∗X) ∈ Gr(T ) and, hence, hw, z∗ − w∗i = hw, z∗X − w∗Xi = hw, x∗ − w∗Xi, hw∗, z∗∗i = hw∗X, x∗∗i, hz∗, z∗∗i = hz∗X, x∗∗i = hx∗, x∗∗i. Therefore, sup hw, z∗i + hw∗, z∗∗i − hw, w∗i < hz∗, z∗∗i, and S is non-type (D) on Ω × Ω∗. (w,w∗)∈Gr(S) Using Proposition 2.2 and Theorem 2.3, we have the following Corollary. Corollary 2.4. Any real Banach space X which contains an isometric copy of c0 has a non-type (D) maximal monotone operator. 6 References [1] J. M. Borwein. Fifty years of maximal monotonicity. Optim. Lett., 4(4):473 -- 490, 2010. [2] O. Bueno and B. F. Svaiter. A maximal monotone operator of type (D) which maximal monotone extension to the bidual is not of type (D), Mar. 2011. arXiv: 1103.0545. [3] S. Fitzpatrick and R. R. Phelps. Some properties of maximal monotone operators on nonre- flexive Banach spaces. Set-Valued Anal., 3(1):51 -- 69, 1995. [4] J.-P. Gossez. Op´erateurs monotones non lin´eaires dans les espaces de Banach non r´eflexifs. J. Math. Anal. Appl., 34:371 -- 395, 1971. [5] J.-P. Gossez. On the range of a coercive maximal monotone operator in a nonreflexive Banach space. Proc. Amer. Math. Soc., 35:88 -- 92, 1972. [6] J.-P. Gossez. On a convexity property of the range of a maximal monotone operator. Proc. Amer. Math. Soc., 55(2):359 -- 360, 1976. [7] J.-P. Gossez. On the extensions to the bidual of a maximal monotone operator. Proc. Amer. Math. Soc., 62(1):67 -- 71 (1977), 1976. [8] M. Marques Alves and B. F. Svaiter. On Gossez type (D) maximal monotone operators. J. Convex Anal., 17(3), 2010. 7
1607.07344
2
1607
2019-04-16T14:49:41
Newton and Bouligand derivatives of the scalar play and stop operator
[ "math.FA" ]
We prove that the play and the stop operator possess Newton and Bouligand derivatives, and exhibit formulas for those derivatives. The remainder estimate is given in a strenghtened form, and a corresponding chain rule is developed. The construction of the Newton derivative ensures that the mappings involved are measurable.
math.FA
math
NEWTON AND BOULIGAND DERIVATIVES OF THE SCALAR PLAY AND STOP OPERATOR MARTIN BROKATE Abstract. We prove that the play and the stop operator possess New- ton and Bouligand derivatives, and exhibit formulas for those deriva- tives. The remainder estimate is given in a strengthened form, and a corresponding chain rule is developed. The construction of the Newton derivative ensures that the mappings involved are measurable. 1. Introduction. The aim of this paper is to show that the play and the stop operator pos- sess Newton as well as Bouligand derivatives, and to compute those deriva- tives. Newton derivatives are needed when one wants to solve equations F (u) = 0 for nonsmooth operators F by Newton's method with a better than lin- ear convergence rate. Bouligand derivatives are closely related to Newton derivatives, and can be used to provide sensitivity results as well as opti- mality conditions for problems involving nonsmooth operators. The scalar play operator and its twin, the scalar stop operator, act on functions u : [a, b] → R and yield functions w = Pr[u; z0] and z = Sr[u; z0] from [a, b] to R. The number z0 plays the role of an initial condition. Their formal definition, in the spirit of [11], is given below in Section 6; alterna- tively, they arise as solution operators of the evolution variational inequality w(t) · (ζ − z(t)) ≤ 0 , for all ζ ∈ [−r, r], z(t) ∈ [−r, r] , z(a) = z0 ∈ [−r, r] , w(t) + z(t) = u(t) . (1a) (1b) (1c) The play and the stop operator are rate-independent; in fact, they constitute the simplest nontrivial examples of rate-independent operators [18, 4, 12, 14] if one disregards relays whose nature is inherently discontinuous. Due to (1c), their mathematical properties are closely related. A lot is known about the play and the stop. Viewed as operators between function spaces, their typical regularity is Lipschitz (or less). In particular, they are not differentiable in the classical sense. The question whether weaker derivatives (e.g., directional derivatives) exist was addressed, to the author's knowledge, for the first time in [3] where it was shown that the 2010 Mathematics Subject Classification. 47H30, 47J40, 49J52, 49M15, 58C20. Key words and phrases. rate independence, hysteresis operator, Newton derivative, Bouligand derivative, play, stop, sensitivity, maximum functional, variational inequality, measurable selector, semismooth, chain rule. 1 2 M. BROKATE play and the stop are directionally differentiable from C[a, b] to Lp(a, b) for p < ∞. (This is not to be confused with the existence and form of time derivatives of functions like t 7→ Pr[u; z0](t), for which there are many results available.) The results below serve to narrow the gap between differentiability and non-differentiability of rate-independent operators. Their proofs given here are based on the same idea as used in [3], namely, to locally represent the play as a composition of operators whose main ingredient is the cumulated maximum. It is natural to ask whether it is possible to prove weak differentiability of the play and the stop operator in the framework of the variational for- mulation (1). Indeed, for elliptic variational inequalities, a large body of literature is available, going back to [13]. In that case, the solution operator is closely linked to the metric projection onto convex sets whose differen- tiability properties also have been analyzed for a long time. For evolution variational inequalities of parabolic type, we refer to the recent contribution [5] and the literature cited there. For rate independent variational inequali- ties, corresponding results do not seem to exist, not even for the ODE case given in (1). Our main results are given in Theorem 7.20 for Newton differentiability and Theorem 8.2 for Bouligand differentiability of the play. They are based on corresponding results for the maximum functional (Proposition 3.4) and the cumulated maximum operator (Proposition 4.8). The extension to the parametric play is given in Proposition 9.5. When attempting to prove Newton differentiability of the play, some is- sues arise which complicate matters and are, at least in part, responsible for the length of this paper. First, the construction of the Newton derivative of the play leads to a set-valued derivative in a natural manner. Its elements L should have the property that the first order approximations δw = Lδu are measurable functions. Since Newton derivatives are not obtained as limits, and we are dealing with operators between function spaces, measurability becomes an issue. Second, with regard to the form of the remainder, we aim at a somewhat stronger result than standard Newton differentiability, hav- ing in mind applications to partial differential equations. Third, we want to treat not only a single play operator, but also a parametric family of play operators, having in mind problems where play operators e.g. are dis- tributed continuously over space. Again, the problem of measurability has to be solved. The proofs of Newton and of Bouligand differentiability are rather similar; for Bouligand derivatives, some of the problems mentioned above do not even arise. Nevertheless, we have chosen to elaborate the proofs for both cases to some extent; the details are somewhat cumbersome and should not be placed too much as a burden on the reader. 2. Notions of derivatives. We collect some established notions of derivatives for mappings F : U → Y , U ⊂ X , NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 3 where X and Y are normed spaces, and U is an open subset of X. These notions are classical, but the terminology is not uniform in the literature. Definition 2.1. (i) The limit, if it exists, F ′(u; h) := lim λ↓0 F (u + λh) − F (u) λ , u ∈ U , h ∈ X , (2) is called the directional derivative of F at u in the direction h. It is an element of Y . (ii) If the directional derivative satisfies F ′(u; h) = lim λ↓0 F (u + λh + r(λ)) − F (u) λ (3) for all functions r : (0, λ0) → X with r(λ)/λ → 0 as λ → 0, it is called the Hadamard derivative of F at u in the direction h. (iii) If the directional derivative exists for all h ∈ X and satisfies kF (u + h) − F (u) − F ′(u; h)k khk lim h→0 = 0 , (4) it is called the Bouligand derivative of F at u in the direction h. (iv) If the Bouligand derivative has the form F ′(u; h) = Lh for some linear continuous mapping L : X → Y , then L is called the Fr´echet derivative of F at u and denoted as DF (u). (v) The mapping F is called directionally (Hadamard, Bouligand, Fr´echet, resp.) differentiable at u (in U , resp.), if the corresponding derivative exists at u (for all u ∈ U , resp.) for all directions h ∈ X. ✷ In the definition above, it is tacitly understood that the limits are taken in the sense "not equal 0". We have F ′(u; λh) = λF ′(u; h) if λ ≥ 0. This as well as the following well-known facts are elementary consequences of the above definitions. Proposition 2.2. Let F be directionally differentiable and locally Lipschitz continuous at u ∈ U . Then F is Hadamard differentiable at u. Moreover, if ℓu is a local Lipschitz constant for F at u, kF ′(u; h1) − F ′(u; h2)k ≤ ℓukh1 − h2k ∀ h1, h2 ∈ X . Consequently, if ℓ is a global Lipschitz constant for F , kF (u + h) − F (u) − F ′(u; h)k ≤ 2ℓkhk ∀ h ∈ X . (5) (6) ✷ Corollary 2.3. If F is locally Lipschitz, then directional and Hadamard differentiability at u ∈ U are equivalent, and are implied by Bouligand dif- ferentiability at u. ✷ In terms of a remainder function, the definition (4) of Bouligand differ- entiability at u is equivalent to kF (u + h) − F (u) − F ′(u; h)k ≤ ρu(khk) · khk , (7) where ρu(δ) ↓ 0 for δ ↓ 0. In view of (6), we may assume that ρu is globally bounded, ρu ≤ 2ℓ ∀ u ∈ U , (8) 4 M. BROKATE if ℓ is a global Lipschitz constant for F . The notion of a Newton derivative is more recent. A mapping G : U → L(X, Y ), the space of all linear and continuous mappings from X to Y , is called a Newton derivative of F in U , if lim h→0 kF (u + h) − F (u) − G(u + h)hk khk = 0 (9) holds for all u ∈ U . It is never unique; for example, modifying G at a single point does not affect the validity of (9) in U . It has turned out to be natural to allow Newton derivatives to be set- valued. For set-valued mappings we write "f : X ⇒ Y " instead of "f : X → P(Y ) \ ∅". Definition 2.4. A mapping G : U ⇒ L(X, Y ) is called a Newton deriv- ative of F in U , if lim h→0 sup L∈G(u+h) kF (u + h) − F (u) − Lhk khk = 0 (10) holds for all u ∈ U . G is called locally bounded if for every u ∈ U the sets {kLk : L ∈ G(v), kv − uk ≤ δ} are bounded for some suitable δ = δ(u). G is called globally bounded if these bounds can be chosen independently from u. It is well known that if F is continuously Fr´echet differentiable in U , then G(u) = {DF (u)} is a single-valued Newton derivative of F in U . We write (10) in remainder form, sup kF (u + h) − F (u) − Lhk ≤ ρu(khk) · khk , (11) L∈G(u+h) where ρu(δ) ↓ 0 as δ ↓ 0. If ℓ is a global Lipschitz constant for F and cG is a global bound for the norms kLk of the elements L ∈ G(U ), we may assume that ρu is globally bounded, ρu ≤ ℓ + cG ∀ u ∈ U , (12) as in the case of the Bouligand derivative. If G : U ⇒ L(X, Y ) is a Newton derivative of F in U , then so is every G : U ⇒ L(X, Y ) satisfying G(u) ⊂ G(u) for all u ∈ U . In particular, every selector S : U → L(X, Y ) of G, that is, S(u) ∈ G(u) for all u ∈ U , yields a single-valued Newton derivative of F in U . We now consider the following situation. The domain of definition U of F can be represented as Un , (13) U = [n∈N where Un ⊂ U are open sets with Un ⊂ Un+1 for all n, and U0 = ∅. We want to obtain a Newton derivative of F on U from Newton derivatives of F on Un. This can be done in the following setting. Let Vn ⊂ U be open sets with V n ⊂ Un ∩ Vn+1 for all n ∈ N, [n∈N Vn = U . (14) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 5 Proposition 2.5. Let Gn be a Newton derivative of F on Un, n ∈ N, with the remainder ρn,u according to (11). Then in the situation just described above, the definition G(u) = Gn(u) , if u ∈ V n \ V n−1, (15) yields a Newton derivative G : U ⇒ L(X; Y ) of F on U with the remainder ρu = max{ρn,u, ρn+1,u} if u ∈ V n \ V n−1. (16) Proof. By construction, V n \ V n−1 , U = [n∈N the union being disjoint. Let u ∈ U , assume that u ∈ V n \ V n−1. We choose δ > 0 such that Bδ(u) = {v : kv − uk < δ} satisfies, see (14), Bδ(u) ∩ V n−1 = ∅ , Bδ(u) ⊂ Un ∩ Vn+1 . Let h ∈ X, khk < δ, let L ∈ G(u+h). If u+h ∈ V n, then u+h ∈ V n \V n−1, u + h ∈ Un and L ∈ Gn(u + h), so kF (u + h) − F (u) − Lhk ≤ ρn,u(khk)khk . If u + h /∈ V n, then u + h ∈ V n+1 \ V n ⊂ Un+1 and L ∈ Gn+1(u + h), so kF (u + h) − F (u) − Lhk ≤ ρn+1,u(khk)khk . This proves the assertions. (cid:3) Remark 2.6. If we have Gn(u) ⊂ Gn+1(u) for all n and u, we may dispense with the sets Vn and simply define a Newton derivative G of F on U by G(u) = Gn(u) , if u ∈ Un \ Un−1. However, in the construction of the Newton derivative of the play given below, this property is not satisfied. ✷ The following result (Lemma 8.11 in [10]) shows that Bouligand and New- ton derivatives are closely related. Proposition 2.7. Let F : U → Y possess the single-valued Newton deriv- ative DN F : U → L(X, Y ). Then F is Bouligand differentiable at u ∈ U if and only if the limit limλ↓0 DN F (u + λh)h exists uniformly w.r.t. h ∈ X with khk = 1. In this case, F ′(u; h) = lim λ↓0 DN F (u + λh)h . 3. The maximum functional We consider ϕ : C[a, b] → R, ϕ(u) = max s∈[a,b] u(s) . (17) ✷ (18) The functional ϕ is convex, positively 1-homogeneous and globally Lipschitz continuous with Lipschitz constant 1, w.r.t. the maximum norm on C[a, b]. By convex analysis, it is directionally (and thus, Hadamard) differentiable. 6 M. BROKATE An explicit formula for the directional derivative is given by (see e.g. [6] for a direct proof) where ϕ′(u; h) = max s∈M (u) h(s) , M (u) = {τ ∈ [a, b], u(τ ) = ϕ(u)} (19) (20) is the set where u attains its maximum. Let us denote the dual of C[a, b] by C[a, b]∗; it consists of all signed regular Borel measures on [a, b]. The subdifferential of ϕ is defined as usual as the set-valued mapping ∂ϕ : C[a, b] ⇒ C[a, b]∗ given by ∂ϕ(u) = {µ : µ ∈ C[a, b]∗, ϕ(v) − ϕ(u) ≥ hµ, v − ui for all v ∈ C[a, b]} . (21) It is not difficult to check that ∂ϕ(u) = {µ : µ ∈ C[a, b]∗, supp(µ) ⊂ M (u), µ ≥ 0, kµk = 1} . (22) In particular, if u has a unique maximum at r ∈ [a, b], that is, M (u) = {r}, then ∂ϕ(u) = {δr}, where δr denotes the Dirac delta at r. A side remark (we will not use this): the directional derivative is linked to the subdifferential by the "max formula" (see [1], Theorem 17.19, for the Hilbert space case) ϕ′(u; h) = max µ∈∂ϕ(u) hµ, hi . The subdifferential is a natural candidate for a Newton derivative of a convex functional. However, the subdifferential of ϕ : C[a, b] → R is not a Newton derivative of ϕ, and ϕ is not Bouligand differentiable. The following example shows that this is true even if we restrict ϕ to W 1,1(a, b). Here and in the sequel we use the norm , 1 ≤ p < ∞ . kukW 1,p = u(a) + ku′kp = u(a) +(cid:16)Z b a u′(s)p ds(cid:17)1/p Example 3.1. Consider u : [0, 1] → R defined by u(s) = 1 − s. We have ϕ(u) = 1 and M (u) = {0}. Define hλ : [0, 1] → R for λ > 0 by hλ(s) =(2s , 2λ , s ≤ λ , s > λ . (23) Then the function u + hλ attains its maximum at s = λ, and khλk1,1 = 2λ , ϕ(u + hλ) = 1 + λ , ϕ′(u; hλ) = max s∈M (u) hλ(s) = hλ(0) = 0 . Consequently, khλk1,1 → 0 but ϕ(u + hλ) − ϕ(u) − ϕ′(u; hλ) khλk1,1 = λ 2λ = 1 2 . (24) Thus, ϕ is not Bouligand differentiable at u on X = W 1,1(a, b). Moreover, setting Φ = (∂ϕ)X we obtain M (u + hλ) = {λ} , Φ(u + hλ) = {δλ} , Φ(u + hλ)hλ = hλ(λ) = 2λ , so ϕ(u + hλ) − ϕ(u) − Φ(u + hλ)hλ khλk1,1 = λ 2λ = 1 2 . (25) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 7 Thus, Φ is not a Newton derivative of ϕ on W 1,1(0, 1). As khλk∞ = khλk1,1 (or due to the embedding W 1,1 → C), the same is true on C[0, 1]. ✷ We will show that Φ is a Newton derivative of ϕ on C 0,α[a, b] for every α > 0, endowed with the norm kukC0,α = u(a) + uC0,α , We set Bε = (−ε, ε). uC0,α = sup t,s∈[a,b] s6=t u(t) − u(s) t − s . (26) Lemma 3.2. The mapping M : C[a, b] ⇒ [a, b] is upper semicontinuous, that is, for every u ∈ C[a, b] and every ε > 0 there exists δ > 0 such that for every h ∈ C[a, b] khk∞ < δ ⇒ M (u + h) ⊂ M (u) + Bε . (27) Proof. By contradiction. Assume that u ∈ C[a, b] and ε > 0 are such that for all n ∈ N there exist hn ∈ C[a, b] with khnk∞ < 1 n and M (u + hn) 6⊂ M (u) + Bε. Let tn ∈ M (u + hn) with d(tn, M (u)) ≥ ε. Passing to a subsequence we get tn → t ∈ [a, b], t /∈ M (u). On the other hand, u(tn)+hn(tn) = ϕ(u+hn). Letting n → ∞ yields u(t) = ϕ(u), so t ∈ M (u), a contradiction. (cid:3) For a function f : I → R, I being an interval, we denote its oscillation on I by (f ) = sup{f (t) − f (s) : t, s ∈ I} , (28) osc I and its modulus of continuity by ωI (f ; ε) = sup{f (t) − f (s) : t, s ∈ I , t − s ≤ ε} . (29) When I = [a, b], we simply write osc(f ) and ω(f ; ε). Lemma 3.3. Let u, h ∈ C[a, b], µ ∈ ∂ϕ(u + h). Then ϕ′(u; h) ≤ ϕ(u + h) − ϕ(u) ≤ hµ, hi . Let moreover be ε > 0 such that M (u + h) ⊂ M (u) + Bε . Then we have hµ, hi − ϕ′(u; h) ≤ sup h(r) − h(s) = ω(h; ε) . s−r≤ε (30) (31) (32) Proof. The first inequality in (30) holds since ϕ is convex; as ϕ(u) − ϕ(u + h) ≥ hµ, −hi, the second inequality follows. Now assume that (31) holds. Recalling (22), given r ∈ supp(µ) ⊂ M (u + h) we find an sr ∈ M (u) with r − sr < ε, so h(r) − ϕ′(u; h) = h(r) − max s∈M (u) h(s) ≤ h(r) − h(sr) ≤ ω(h; ε) . Integrating both sides of this inequality over r ∈ [a, b] with respect to µ yields (32). (cid:3) For the modulus of continuity, we have ω(h; ε) ≤ hC0,α εα , ω(h; ε) ≤ kh′kLpε1−1/p . (33) 8 M. BROKATE Proposition 3.4. Let X = C 0,α[a, b] or X = W 1,p(a, b), with 0 < α ≤ 1 resp. 1 < p ≤ ∞. Then the set-valued mapping Φ = (∂ϕ)X given in (22) is a globally bounded Newton derivative of the maximum functional ϕ on X. In particular, for every u ∈ X there exists a nondecreasing and bounded ρu : R+ → R+ such that ρu(δ) → 0 as δ → 0, ρu is bounded independently from u, and ϕ(u + h) − ϕ(u) − Lh ≤(ρu(khk∞)hC0,α ρu(khk∞)kh′kLp respectively, for every h ∈ X and every L ∈ Φ(u + h). Moreover, ϕ is Bouligand differentiable on X, and for every u ∈ X ϕ(u + h) − ϕ(u) − ϕ′(u; h) ≤(ρu(khk∞))hC0,α ρu(khk∞)kh′kLp (34) (35) respectively, for every h ∈ X. Proof. We consider the case X = C 0,α[a, b]. Let u ∈ X be given, let εu(δ) = inf{ε : M (u + Bδ) ⊂ M (u) + Bε} for δ > 0. Then εu is increasing. As M is upper semicontinous by Lemma 3.2, we have 0 < εu(δ) → 0 as δ → 0, According to (30) and (32), for h ∈ X and L = µ ∈ Φ(u + h) we get ϕ(u + h) − ϕ(u) − Lh ≤ ω(h; εu(khk∞) ≤ εu(khk∞)α · hC0,α . Setting ρu(δ) = εu(δ)α, (34) follows for the Holder case. Since kLkC→R = 1, we have kLkC0,α→R ≤ cα and ϕ(u + h) − ϕ(u) − Lh ≤ 2khk∞ ≤ 2cαkhk∞ , where cα denotes the norm of the embedding C 0,α → C. Thus, cα is a global bound for Φ, and 2cα furnishes a global bound for ρu. The proof for the case X = W 1,p(a, b) is analogous. (One might also refer to Morrey's embedding theorem which implies that W 1,p(a, b) is continu- ously embedded into C 0,α[a, b] for α ≤ 1 − 1/p.) (cid:3) Note that the estimates (34) and (35) are slightly stronger than required for Newton and Bouligand differentiability (the factor ρu(khkX ) instead of ρu(khk∞), as well as the norms instead of the seminorms, would suffice). This strenghtening is motivated by applications to partial differential equa- tions. 4. The cumulated maximum We define the cumulated maximum of a function u ∈ C[a, b] as ϕt(u) = max s∈[a,t] u(s) , t ∈ [a, b] . Setting we obtain an operator (F u)(t) = ϕt(u) F : C[a, b] → C[a, b] . (36) (37) (38) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 9 The function F u is nondecreasing for every u ∈ C[a, b]. Since ϕt(u) − ϕt(v) ≤ max s∈[a,t] u(s) − v(s) , for all u, v ∈ C[a, b], we have kF u − F vk∞,t ≤ ku − vk∞,t , for all u, v ∈ C[a, b], t ∈ [a, b]. (39) Here and in the following we use the notation kuk∞,t = sup s≤t u(s) . (40) For any fixed t ∈ [a, b], the directional derivative of ϕt : C[a, b] → R given in (19) yields that, for all u, h ∈ C[a, b], F P D(u; h)(t) := lim λ↓0 (F (u + λh))(t) − (F u)(t) λ = ϕ′ t(u; h) = max s∈M (u,t) where M (u, t) = {τ : τ ∈ [a, t], u(τ ) = ϕt(u)} h(s) , (41) (42) is the set where u attains its maximum on [a, t]. As in [3], we call pointwise directional derivative of F the function F P D(u; h) : [a, b] → R obtained in this manner. Example 4.3 in [3] shows that the function F P D(u; h) : [a, b] → R does not need to be continuous even though u and h are; so F : C[a, b] → C[a, b] is not directionally differentiable. When this happens, the difference quotients F (u + λh) − F u λ do not converge uniformly to F P D(u; h). They do, on the other hand, con- verge in Lr(a, b) for every r < ∞, as they are uniformly bounded by khk∞. As a consequence, F : C[a, b] → Lr(a, b) is Hadamard differentiable ([3]). In order to obtain Bouligand or Newton differentiability, as in the case of the maximum functional one has to strengthen the norm in the domain space. Indeed, the functions from Example 3.1 can be used to show that F is not Bouligand differentiable on C[a, b]. Bouligand differentiability of the cumulated maximum. Let again X stand for C 0,α[a, b] with 0 < α ≤ 1, or for W 1,p(a, b) with 1 < p ≤ ∞. We want to prove that F : X → Lq(a, b) is Bouligand differentiable for 1 ≤ q < ∞ with the improved remainder estimate as in Proposition 3.4. For this, we have to show that ρF u (δ) := sup khk∞≤δ kF (u + h) − F (u) − F ′(u; h)kLq khkX → 0 as δ → 0. (43) Proposition 4.1. The cumulated maximum F : X → Lq(a, b) is Bouligand differentiable for every q < ∞, and F ′ = F P D. Moreover, kF (u + h) − F (u) − F ′(u; h)kLq ≤ ρF u (khk∞) · khkX , (44) and ρF u (δ) → 0 as δ → 0. In addition, ρF u is bounded uniformly in u. 10 M. BROKATE Proof. Assume that (43) does not hold. Then there exists ε > 0 and a sequence {hn} in X with khnk∞ → 0 and εkhnkX ≤ kF (u + hn) − F (u) − F P D(u; hn)kLq =(cid:18)Z b a where 1/q dn(t)q dt(cid:19) , (45) dn(t) = ϕt(u + hn) − ϕt(u) − ϕ′ t(u; hn) . Setting ρn = dn/khnkX we have ρn(t) → 0 pointwise, because ϕt : X → R is Bouligand differentiable for every t by Proposition 3.4, with the remainder estimate (35). Since moreover {ρn} is uniformly bounded, by dominated convergence kρnkLq → 0 which contradicts (45). Therefore F is Bouligand differentiable and F ′ = F P D. The global bound on ρF u follows from the esti- mate kF (u+h)−F (u)−F ′(u; h)k∞ ≤ 2khk∞ combined with the embedding constants. (cid:3) Newton differentiability of the cumulated maximum. A Newton derivative of the cumulated maximum is constructed from the Newton deriv- ative of the maximum functional given in the previous section. Its elements L will have the form (Lh)(t) = hµt, hi, where µt belongs to the Newton derivative Φt of ϕt. In order that Lh becomes a measurable function, the measures µt are constructed from measurable selectors of the family {Φt}. We first analyze the mapping M : C[a, b] × [a, b] ⇒ [a, b] M (u, t) = {τ : τ ∈ [a, t], u(τ ) = ϕt(u)} . (46) The sets M (u, t) are compact nonempty subsets of [a, b], and M (u, a) = {a}. Lemma 4.2. The set-valued mapping M is upper semicontinuous and mea- surable. Proof. To prove that M is upper semicontinuous according to Definition 10.1, let A ⊂ [a, b] be closed, and let (un, tn) be a sequence in M −1(A) with un → u ∈ C[a, b] and tn → t ∈ [a, b]. In order to show that (u, t) ∈ M −1(A), let τn ∈ A such that τn ∈ M (un, tn), thus un(τn) = ϕtn (un). Passing to a subsequence we have τn → τ ∈ A since A is closed. Moreover, τ ≤ t, un(τn) → u(τ ) and ϕtn (un) = (ϕtn (un) − ϕtn (u)) + ϕtn (u) → ϕt(u) by (39) and since t 7→ ϕt(u) is continuous. Therefore u(τ ) = ϕt(u) and τ ∈ M (u, t). Thus M is upper semicontinuous. It now follows from Proposition 6.2.3 in [15] that M is measurable. (cid:3) The set-valued mapping M possesses a dense sequence of measurable selectors. Proposition 4.3. There exists a sequence {fn} of measurable selectors of M such that M (u, t) = {fn(u, t) : n ∈ N} , for all u ∈ C[a, b], t ∈ [a, b] . (47) In particular max M (u, t) = supn fn(u, t) and min M (u, t) = infn fn(u, t) are measurable selectors of M . NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 11 Proof. This is a consequence of Theorem 6.3.18 in [15], as [a, b] is a complete separable metric space. (cid:3) We consider the mapping Φ : C[a, b] × [a, b] ⇒ C[a, b]∗, Φ(u, t) = {ν ∈ C[a, b]∗ : supp(ν) ⊂ M (u, t), ν ≥ 0, kνk = 1} . (48) The following facts are well known. The closed unit ball K in C[a, b]∗, en- dowed with the weak star topology, is compact (hence complete), metrizable and separable. The sets Φ(u, t) are nonempty convex and weak star compact subsets of K (note that for ν ≥ 0 we have kνk = hν, 1i). Moreover, Φ(u, a) = {δa} , M (u + c, t) = M (u, t) , Φ(u + c, t) = Φ(u, t) for all c ∈ R, (Φ(u, t))(c) = {c} for all c ∈ R. (49) (50) (51) Lemma 4.4. Let {un}, {tn}, {νn} be sequences in C[a, b], [a, b] and C[a, b]∗ ∗ respectively, with un → u, tn → t and νn ⇀ ν, let supp(νn) ⊂ M (un, tn) for all n ∈ N. Then supp(ν) ⊂ M (u, t). Proof. Let f ∈ C ∞ 0 (R \ M (u, t)). We have to show that hν, f i = 0. Let ε = inf{s − τ : s ∈ supp(f ), τ ∈ M (u, t)} . We have ε > 0 because the sets supp(f ) and M (u, t) are disjoint and com- pact. Since M is upper semicontinuous by Proposition 4.2, we may choose N ∈ N such that M (un, tn) ⊂ M (u, t) + Bε/2 holds for all n ≥ N . Then supp(f ) ∩ M (un, tn) = ∅ and thus hνn, f i = 0 for all n ≥ N . Passing to the limit n → ∞ we arrive at hν, f i = 0. (cid:3) Proposition 4.5. The mapping Φ : C[a, b]× [a, b] ⇒ C[a, b]∗ defined in (48) is upper semicontinuous, thus measurable. Proof. Let A ⊂ C[a, b]∗ be weak star closed. We have to show that Φ−1(A) is closed. To this end, let {(un, tn)} be a sequence in Φ−1(A) with un → u in C[a, b] and tn → t in [a, b]. Let νn ∈ Φ(un, tn), so νn ∈ A as well as νn ≥ 0, kνnk = 1 and supp(νn) ⊂ M (un, tn) for all n ∈ N. For some subsequence, ∗ ⇀ ν with ν ≥ 0, kνk = 1 and ν ∈ A. By Proposition 4.4, we have νnk supp(ν) ⊂ M (u, t). Thus, (u, t) ∈ Φ−1(A) and the proof is complete. (cid:3) Proposition 4.6. There exists a sequence {µn} of measurable selectors of Φ such that Φ(u, t) = {µn(u, t) : n ∈ N} , for all u ∈ C[a, b], t ∈ [a, b] , (52) the closure being taken w.r.t. the weak star topology. Proof. This follows from Theorem 6.3.18 in [15], as the unit ball in C[a, b]∗ is a complete separable metrizable space w.r.t. the weak star topology. (cid:3) Lemma 4.7. Let µ be a measurable selector of Φ. Then defines an element L ∈ L(C[a, b]; L∞(a, b)) with kLk = 1 and (Lh)(t) = hµ(u, t), hi kLhk∞,t ≤ khk∞,t . (53) 12 M. BROKATE Proof. For every u, h ∈ C[a, b], the mapping t 7→ hµ(u, t), hi is measurable and satisfies hµ(u, s), hi ≤ khk∞,t for all s ≤ t, since µ(u, s) has support in [a, s]. Thus, L is well-defined, kLk ≤ 1 and (53) holds. As µ ≥ 0 and L(1) = 1, we have kLk = 1. (cid:3) Let X again denote any one of the spaces C 0,α[a, b] for 0 < α ≤ 1 or W 1,p(a, b) for 1 < p ≤ ∞. Proposition 4.8. Let SΦ be the set of all measurable selectors of Φ, let q ∈ [1, ∞). The set-valued mapping G : X ⇒ L(X, Lq(a, b)), G(u) = {L : (Lh)(t) = hµ(u, t), hi, µ ∈ SΦ} (54) defines a Newton derivative of the cumulated maximum F : X → Lq(a, b) with kF (u + h) − F (u) − LhkLq ≤ ρG u (khk∞) · khkX , (55) for all L ∈ G(u + h). Here, ρG ρG u (δ) → 0 as δ → 0, bounded independently from u. u : R+ → R+ is an increasing function with Proof. Fix u ∈ X. For h ∈ X we define d(h, t) = sup µt∈Φ(u+h,t) ϕt(u + h) − ϕt(u) − hµt, hi . Let {µk} be a sequence of measurable selectors of Φ according to Proposition 4.6, set dk(h, t) = ϕt(u + h) − ϕt(u) − hµk(u + h, t), hi Then d(h, t) = supk dk(h, t) by (52), and therefore the mapping t 7→ d(h, t) is measurable. Moreover, sup L∈G(u+h) kF (u + h) − F (u) − LhkLq =(cid:18)Z b a d(h, t)q dt(cid:19) 1/q =: dG(h) . The remainder of the proof is analogous to that of Proposition 4.1. We define ρG u (δ) = sup khk∞≤δ dG(h) khkX . (56) Assume that limδ→0 ρG sequence {hn} in X with khnk∞ → 0 and u (δ) = 0 does not hold. Then there exist ε > 0 and a εkhnkX ≤(cid:18)Z b a d(hn, t)q dt(cid:19) 1/q . (57) Since Φ(·, t) is a Newton derivative of ϕt, we have ρn(t) = d(hn, t)/khnkX → 0 pointwise in t as n → ∞. Moreover, ρn is uniformly bounded. Applying dominated convergence, we arrive at a contradiction to (57). The global boundedness of ρG u follows from the estimate kF (u + h) − F (u) − Lhk∞ ≤ 2khk∞. (cid:3) Proposition 4.10 below shows that the set SΦ is large enough to approxi- mate the whole range of Φ. NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 13 Lemma 4.9. Let f : C[a, b] × [a, b] → [a, b] be a measurable selector of M . Then µ(u, t) = δf (u,t) (58) defines a measurable selector µ : C[a, b] × [a, b] → C[a, b]∗ of Φ. Proof. For each v ∈ C[a, b], the mapping s 7→ v(s) = hδs, vi is continuous from [a, b] to R. Thus, the mapping s 7→ δs is weak star continuous from [a, b] to C[a, b]∗, and consequently (59) defines a measurable mapping. (cid:3) Proposition 4.10. Let {fn} be a sequence of measurable selectors of M such that M (u, t) = {fn(u, t) : n ∈ N} , for all u ∈ C[a, b], t ∈ [a, b] . (59) Taking all rational convex combinations of the mappings (u, t) 7→ δfn(u,t) we obtain a sequence {µn} of measurable selectors of Φ such that Φ(u, t) = {µn(u, t) : n ∈ N} , for all u ∈ C[a, b], t ∈ [a, b] , (60) the closure being taken w.r.t. the weak star topology. Proof. Let u ∈ C[a, b] and t ∈ [a, b] be given. The set D = {fn(u, t) : n ∈ N} is a countable dense subset of M (u, t). The set of all convex combinations with rational coefficients of elements of the set {δτ : τ ∈ D} then is dense in Φ(u, t) w.r.t. the weak star topology. (cid:3) 5. The chain rule In the following sections we will see that the play operator can be rep- resented as a finite composition of cumulated maxima and positive part mappings. The Newton differentiability of these mappings will imply New- ton differentiability of the play, by virtue of the chain rule. It is a standard result that the chain rule is valid for Newton derivatives, see Proposition A.1 in [8] for the single-valued and Proposition 3.8 in [17] for the set-valued case. As a result of investigating the maximum and the cumulated maximum, we have seen above that these operators satisfy a slightly stronger version of Newton and Bouligand differentiability. For the cumulated maximum F : X → Y with X = W 1,p(a, b) or C 0,α[a, b] and Y = Lr(a, b), we have constructed a Newton derivative G : X ⇒ L(X; Y ) with a remainder esti- mate sup L∈G(u+h) kF (u + h) − F (u) − LhkY ≤ ρu(khk X ) · khkX , (61) where X = C[a, b], endowed with the maximum norm. The purpose of this section is to extend the chain rule to this situation, for Newton as well as for Bouligand derivatives. We consider the following setting. Assumption 5.1. (i) X, Y, Z are normed spaces, U ⊂ X and V ⊂ Y are open. F1 : U → Y and F2 : V → Z with F1(U ) ⊂ V are locally Lipschitz. (ii) X and Y are normed spaces with continuous embeddings X ⊂ X and Y ⊂ Y . 14 M. BROKATE (iii) G1 : U ⇒ L(X; Y ) and G2 : V ⇒ L(Y ; Z) satisfy, for every u ∈ U and v ∈ V , sup L1∈G1(u+h) kF1(u + h) − F1(u) − L1hkY ≤ ρ1,u(khk X ) · khkX (62) for every h ∈ X with u + h ∈ U , sup L2∈G2(v+k) kF2(v + k) − F2(v) − L2kkZ ≤ ρ2,v(kkk Y ) · kkkY (63) for every k ∈ Y with v + k ∈ V , with functions ρ1,u, ρ2,v : R+ → R+ satisfying ρ1,u(δ) ↓ 0 and ρ2,v(δ) ↓ 0 for δ ↓ 0. (iv) F1 : (U, k · k X ) → (V, k · k Y ) is continuous. (v) G2 is locally bounded on (V, k · kY ). Since ρ1,u(khk X ) ≤ ρ1,u(ckhkX ) for some constant c, part (iii) of the assumption implies that G1 and G2 are Newton derivatives for F1 in U and F2 in V , respectively. Note also that the assumption "G2 locally bounded" already implies that F2 is locally Lipschitz. In the special case X = X and Y = Y , (62) and (63) reduce to the standard remainder form (11), and part (iv) of the assumption is implied by part (i); the following result then reduces to the standard chain rule for Newton derivatives. Proposition 5.2 (Refined Chain Rule, Newton Derivative). Let Assumption 5.1 hold. Then G : U ⇒ L(X; Z) G(u) = {L2 ◦ L1 : L1 ∈ G1(u), L2 ∈ G2(F1(u))} (64) is a Newton derivative of F = F2 ◦ F1 in U which satisfies, for every u ∈ U , sup L∈G(u+h) kF (u + h) − F (u) − LhkZ ≤ ρu(khk X ) · khkX (65) for every h ∈ X with u + h ∈ U , where ρu : R+ → R+ is a function with ρu(δ) ↓ 0 for δ ↓ 0. Proof. Let u ∈ U , h ∈ X with u + h ∈ U , set k = F1(u + h) − F1(u). Let L1 ∈ G1(u + h), L2 ∈ G2(F1(u + h)) = G2(F1(u) + k). By the triangle inequality, k(F2 ◦ F1)(u + h) − (F2 ◦ F1)(u) − (L2 ◦ L1)hkZ ≤ kF2(F1(u) + k) − F2(F1(u)) − L2kkZ + kL2(k − L1h)kZ (66) Since G2 is locally bounded, there exists a C > 0 such that for sufficiently small khkX we have kL2k ≤ C for all L2 ∈ G2(F1(u + h)). Consequently, for all such h and L2, and for all L1 ∈ G1(u + h) we have by (62) kL2(k−L1h)kZ ≤ CkF1(u+h)−F1(u)−L1hkY ≤ Cρ1,u(khk X )·khkX . (67) Moreover, by (63) kF2(F1(u) + k) − F2(F1(u)) − L2kkZ ≤ ρ2,F1(u)(kkk Y ) · kkkY . (68) Since F1 is locally Lipschitz, kkkY ≤ C1khkX for small enough khkX . Now let us define ρu : R+ → R+ by ρu(λ) = sup{kF1(u + h) − F1(u)k Y : khk X ≤ λ} . (69) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 15 By part (iv) of Assumption 5.1, ρu(λ) → 0 as λ → 0. Putting together the estimates obtained so far, we get k(F2 ◦ F1)(u + h) − (F2 ◦ F1)(u) − (L2 ◦ L1)hkZ ≤(cid:0)C1ρ2,F1(u)(ρu(khk X )) + Cρ1,u(khk X )(cid:1) · khkX independent from the choice of L1 and L2, as long as khkX is sufficiently small. Setting (70) ρu(λ) = C1ρ2,F1(u)(ρu(λ)) + Cρ1,u(λ) we have ρu(λ) → 0 as λ → 0. Thus, it follows from (70) that (65) holds. (cid:3) In order to obtain the refined chain rule for Bouligand derivatives, we replace Assumption 5.1(iii) by F1 and F2 are Bouligand differentiable in U and V , respec- tively. For every u ∈ U , v ∈ V we have kF1(u + h) − F1(u) − F ′ kF2(v + k) − F2(v) − F ′ 1(u; h)kY ≤ ρ1,u(khk X ) · khkX 2(v; k)kZ ≤ ρ2,v(kkk Y ) · kkkY (71) for every h ∈ X with u + h ∈ U and every k ∈ Y with v + k ∈ V . Lemma 5.3. If F1 and F2 are Hadamard differentiable at u resp. F1(u), then F2 ◦ F1 is Hadamard differentiable at u, and the chain rule (F2 ◦ F1)′(u; h) = F ′ 2(F1(u); F ′ 1(u; h)) holds for all h ∈ X. Proof. See e.g. [2], Proposition 2.47. (72) ✷ (cid:3) Proposition 5.4 (Refined Chain Rule, Bouligand Derivative). Let (i) - (iv) of Assumption 5.1 hold, with (iii) replaced by (71). Then F = F2 ◦ F1 is Bouligand differentiable in U , and 1(u; h)) . Moreover, for every u ∈ U and h ∈ X with u + h ∈ U F ′(u; h) = F ′ 2(F1(u); F ′ kF (u + h) − F (u) − F ′ 2(F1(u); F ′ 1(u; h))kZ ≤ ρu(khk X )khkX (73) (74) for some ρu : R+ → R+ with ρu(δ) ↓ 0 as δ ↓ 0. Proof. By Lemma 5.3, F is Hadamard differentiable and the chain rule holds. It remains to show (74) for the remainder. Let u ∈ U , h ∈ X with u+h ∈ U , set k = F1(u + h) − F (u). We have F2(F1(u + h)) − F2(F1(u)) − F ′ 2(F1(u); F ′ 1(u; h)) = (cid:0)F2(F1(u) + k) − F2(F1(u)) − F ′ +(cid:0)F ′ 2(F1(u); F ′ Let Ci be local Lipschitz constants for Fi. The inequality 1(u; h))kX ≤ C2ρ1,u(khk X )khkX 2(F1(u); k)) − F ′ C2kk − F ′ 2(F1(u); k))(cid:1) 1(u; h))(cid:1) (75) yields an estimate for the second term on the right side of (75); the first term is estimated by ρ2,F1(u)(kkk Y ) · kkkY . 16 M. BROKATE Since kkkY ≤ C1khkX , we argue as in the proof of Proposition 5.2 and obtain, with ρu defined as in (69), k(F2 ◦ F1)(u + h) − (F2 ◦ F1)(u) − F ′ 2(F1(u); F ′ 1(u; h))kZ From this, the claim readily follows. ≤(cid:0)C1ρ2,F1(u)(ρu(khk X )) + C2ρ1,u(khk X )(cid:1) · khkX . (cid:3) 6. The scalar play and stop operators The original construction of the play and the stop operators in [11] is based on piecewise monotone input functions. A continuous function u : [a, b] → R is called piecewise monotone, if the restriction of u to each interval [ti, ti+1] of a suitably chosen partition ∆pm = {ti}, a = t0 < t1 < · · · < tN = b, called a monotonicity partition of u, is either nondecreasing or nonincreasing. By Cpm[a, b] we denote the space of all such functions. For arbitrary r ≥ 0, the play operator Pr and the stop operator Sr are constructed as follows. (For more details, we refer to Section 2.3 of [4].) Given a function u ∈ Cpm[a, b] and an initial value z0 ∈ [−r, r], we define functions w, z : [a, b] → R successively on the intervals [ti, ti+1], 0 ≤ i < N , of a monotonicity partition ∆pm of u by z(a) = πr(z0) := max{−r, min{r, z0}} , w(a) = u(a) − z(a) , (76) and w(t) = max{u(t) − r , min{u(t) + r, w(ti)}} , z(t) = v(t) − w(t) , ti < t ≤ ti+1 . (77) In this manner, we obtain operators w = Pr[u; z0] , z = Sr[u; z0] , Pr, Sr : Cpm[a, b] × R → Cpm[a, b] . By construction, u = w + z = Pr[u; z0] + Sr[u; z0] . (78) The play operator satisfies kPr[u; z0] − Pr[v; y0]k ≤ max{ku − vk , z0 − y0} (79) for all u, v ∈ Cpm[a, b] and all z0, y0 ∈ R. Therefore, Pr and Sr can be uniquely extended to Lipschitz continuous operators Pr, Sr : C[a, b] × R → C[a, b] which satisfy (79) for all u, v ∈ C[a, b] and all z0, y0 ∈ R. In [3], Hadamard derivatives of Pr and of Sr have been obtained. We recall some of the terminology used there, as it is also relevant for the present paper. Let (u, z0) ∈ C[a, b] × R be given, let w = Pr[u; z0], z = Sr[u; z0] with r > 0. (For r = 0, Pr[u; z0] = u.) The trajectories {(u(t), w(t)) : t ∈ [a, b]} lie within the subset A = {u − w ≤ r} of the plane R2 bounded by the straight lines u−w = ±r. They consist of parts which belong to the interior, NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 17 the right or the left boundary of A. Correspondingly, the time interval [a, b] decomposes into the three disjoint sets I0 = {t ∈ [a, b] : u(t) − w(t) = z(t) < r} , I∂+ = {t ∈ [a, b] : u(t) − w(t) = z(t) = r} , I∂− = {t ∈ [a, b] : u(t) − w(t) = z(t) = −r} . The set I0 is an open subset of [a, b], the sets I∂± are compact. As I∂+ and I∂− are disjoint, δI := min{τ − σ : τ ∈ I∂+ , σ ∈ I∂−} > 0 . (80) Because of this, there exists a finite partition ∆(u, z0) = {tk} of [a, b] such that on each partition interval Ik = [tk−1, tk] we have z(t) > −r for all t ∈ Ik or z(t) < r for all t ∈ Ik, or both. In the former case, Ik is called a plus interval; on Ik the trajectory stays away from the left boundary of A, and Ik ⊂ I0 ∪ I∂+. In the latter case, Ik is called a minus interval; the trajectory stays away from the right boundary of A, and Ik ⊂ I0 ∪ I∂−. Note that if Ik ⊂ I0, then Ik is a plus as well as a minus interval. It has been proved in [3], Lemma 5.1, that on such intervals the play op- erator behaves like an cumulated maximum resp. minimum. More precisely, on a plus interval Ik, w(t) = Pr[u; z0](t) = max{w(tk−1) , max (u(s) − r)} s∈[tk−1,t] (81) holds, no matter whether u is monotone on Ik or not. On a minus interval, (82) w(t) = Pr[u; z0](t) = min{w(tk−1) , min (u(s) + r)} . s∈[tk−1,t] In particular, w(t) = w(tk−1) if Ik ⊂ I0. Due to (80) and the continuity of Pr, in this manner the play and the stop operator can locally be represented by a finite composition of operators arising from the cumulated maximum resp. minimum. The following result has been proved in [3], Lemma 5.2. Proposition 6.1. For every (u, z0) ∈ C[a, b] × R there exists a partition ∆(u, z0) = {tk}0≤k≤N of [a, b] and a δ > 0 such that every partition interval [tk−1, tk] of ∆ is a plus interval for all (v, y0) ∈ Uδ × R, or it is a minus interval for all (v, y0) ∈ Uδ × R. Here, Uδ := {(v, y0) : kv − uk∞ < δ, y0 − z0 < δ, v ∈ C[a, b], y0 ∈ R} (83) is the δ-neighbourhood of (u, z0) w.r.t the maximum norm. ✷ As a consequence, invoking the chain rule for Hadamard derivatives, it has been proved in [3] that Pr and Sr are Hadamard differentiable on C[a, b]×R, if Lq(a, b) with q < ∞ is chosen as the range space. 7. Newton derivative of the play and the stop We want to use the approach outlined in the previous section in order to obtain a Newton derivative of Pr, based on the Newton derivative of the cumulated maximum. We want to construct the Newton derivative such that its dependence upon (u, z0) becomes measurable in a suitable manner; for this, the local 18 M. BROKATE representation of the play obtained from Proposition 6.1 seems to be of very limited value. Instead, we employ properties of the set-valued mappings involved when constructing above the Newton derivative of the cumulated maximum. To this purpose, we turn around the approach of Proposition 6.1. Instead of finding a suitable partition ∆ for a given (u, z0), for a given par- tition ∆ we consider sets of (u, z0) for which the play can be "decomposed" by ∆. Throughout the following, the space X stands for C 0,α[a, b] or W 1,p(a, b). Let ∆ = {tk} be a partition of [a, b], a = t0 < · · · < tN = b for some N ∈ N. We set Ik = [tk−1, tk] , ∆ = max 1≤k≤N Ik = max 1≤k≤N (tk − tk−1) . We define C ∆ = {u : u ∈ C[a, b], z0 ∈ R, osc Ik (u) < r for all k} , X ∆ = X ∩ C ∆ , Z ∆ = C ∆ × R = {(u, z0) : u ∈ C ∆, z0 ∈ R} . (84) The sets C ∆, X ∆ and Z ∆ are open subsets of C[a, b], X and C[a, b] × R, respectively. The dynamics on an interval for small input oscillation. It turns out below in Proposition 7.5 that an interval I ⊂ [a, b] is a plus or a minus interval for the play if the oscillation of u on I is less than r. This and some other auxiliary results are developed up to Proposition 7.7. Let I = [t∗, t∗] ⊂ [a, b], u ∈ C(I). We denote the cumulated maximum of u on I and the sets where it is attained by (F I u)(t) = max s∈I,s≤t u(s) , t ∈ I , M I (u, t) = {s : s ∈ I, s ≤ t, u(s) = (F I u)(t)} . (85) As above, F I : C(I) → C(I), F I u is nondecreasing and F I (u+c) = F I (u)+c if c is a constant. Moreover, osc I (F I u) = u(t∗) − u(t∗) ≤ osc I (u) , 0 ≤ F I (u) − u ≤ osc I u on I, and consequently osc I (F I (u) − u) ≤ osc I u . The cumulated minimum of u on I can be written as min s∈I,s≤t u(s) = −(F I (−u))(t) , t ∈ I . (86) (87) (88) (89) The corresponding sets of minima are given by M (−u, t). For u ∈ C(I), p ∈ R and r > 0 we define the functions (here and in the following, the max and the min are taken pointwise in t) w+ = max{p, F I (u − r)} , w− = min{p, −F I (−u − r)} , z+ = u − w+ , z− = u − w− . (90) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 19 This corresponds to the operations in (81) and (82). We have w+, w−, z+, z− ∈ C(I). Obviously w− ≤ w+, z+ ≤ z−. Since p ≤ w+ = u − z+ and p ≥ w− = u − z−, we have z+ ≤ u − p ≤ z− . (91) Lemma 7.1. Let u ∈ C(I), p ∈ R, r > 0. (i) We have z+ ≤ r on I. (F I u)(t) ≥ p + r. (ii) We have z− ≥ −r on I. If z−(t) = −r for some t ∈ I, then u(t) = −(F I (−u))(t) ≤ p − r. If z+(t) = r for some t ∈ I, then u(t) = Proof. To obtain (i), we use the estimate z+ = u − w+ = u − max{p, F I (u − r)} = u − p − max{0, F I (u − r − p)} ≤ u − p − F I (u − r − p) = u − F I (u) + r ≤ r . If z+(t) = r, equality holds everywhere, so u(t) = (F I u)(t) and F I (u − r − p)(t) ≥ 0. The proof of (ii) is analogous. (cid:3) We consider inputs in C(I) whose oscillation is smaller than r. Z I Z I = {(u, p) : u ∈ C(I), p ∈ R, osc I + = {(u, p) : u ∈ C(I), p ∈ R, osc I − = {(u, p) : u ∈ C(I), p ∈ R, osc I Z I u < r} u < r, z+ > −r on I} (92) u < r, z− < r on I} The sets Z I they correspond to plus and minus intervals for the play. − are open subsets of Z I in C(I) × R; we will see that + and Z I Lemma 7.2. (i) If (u, p) ∈ Z I (ii) If (u, p) ∈ Z I (iii) If (u, p) ∈ Z I − then F I (u − r − p) < 0 and w+ = p on I. + then F I (−u − r + p) < 0 and w− = p on I. − ∩ Z I + then w+ = w− = p and z+ = z− = u − p on I. Proof. If (u, p) ∈ Z I so w+ = p. F I (−u − r) + p < 0, so w− = p. Lemma 7.3. Let u ∈ C(I), oscI (u) < r, p ∈ R. Then − then u−p−r ≤ z− −r < 0 by (91), so F I (u−r)−p < 0, + then −u + p − r ≤ −z+ − r < 0 by (91), so (cid:3) If (u, p) ∈ Z I min{u − p, 0} ≤ z+ ≤ z− ≤ max{u − p, 0} . (93) Proof. We have −z+ = w+ − u = max{p − u, F I (u) − r − u} ≤ max{p − u, 0} , since F I u − u ≤ oscI u < r by (87). Analogously, −z− = w− − u = min{p − u, −F I (−u − r) − u} ≥ min{p − u, 0} , since F I (−u) − (−u) ≤ oscI (−u) < r by (87). Lemma 7.4. We have Z I = Z I + ∪ Z I −. (cid:3) Proof. Let (u, p) ∈ Z I , assume that (u, p) /∈ Z I +. Then z+(t) ≤ −r for some t ∈ I. By (93), u(t) − p ≤ −r. As oscI (u) < r, we have u − p ≤ 0 on I. By (93), z− ≤ 0 on I, so (u, p) ∈ Z I −. (cid:3) 20 M. BROKATE We define P I − : Z I + : Z I + → C(I) and P I P I +(u, p) = p + max{0, F I (u − r − p)} , P I −(u, p) = p − max{0, F I (−u − r + p)} . − → C(I) by Therefore, in view of (90), u − P I u − P I +(u, p) = u − w+ = z+ > −r −(u, p) = u − w− = z− < r − both expressions simplify to P I + ∩ Z I On Z I Therefore, (94) (95) (96) on I ⇔ (u, p) ∈ Z I +, on I ⇔ (u, p) ∈ Z I − . ±(u, p) = p by Lemma 7.2. P I (u, p) = P I ±(u, p) , if (u, p) ∈ Z I ± (97) yields a well-defined mapping P I : Z I → C(I). The next result states that for u ∈ C ∆ the intervals Ik yield a decompo- sition of the play operator. This is the analogue of Proposition 6.1. Proposition 7.5. Let u ∈ C ∆ and z0 ∈ R, set p = Pr[u; z0](tk−1), k ≥ 1. Then w(t) = Pr[u; z0](t) = P Ik(u, p)(t) , for all t ∈ Ik. (98) Moreover, Ik is a plus interval ⇔ (u, p) ∈ Z Ik + , Ik is a minus interval ⇔ (u, p) ∈ Z Ik − . (99) Proof. Let {un} be a sequence in C[a, b] such that the functions un coincide with u on [0, tk−1], are piecewise linear on Ik and satisfy un → u uniformly. For n large enough we have un ∈ C ∆, so zn > −r on Ik if (u, p) ∈ Z Ik + and zn < r if (u, p) ∈ Z Ik − . It follows that wn = Pr[un; z0] = P Ik (un, p) on Ik, by the definition of the play on Cpm(Ik), see (77). Passing to the limit n → ∞ yields (98). To prove the first equivalence in (99), let Ik be a plus interval. On Ik we then have u − w > −r and, by (81), w = P Ik + (u, p), so (u, p) ∈ Z Ik + , we have u − P Ik (u, p) = u − P Ik + (u, p) > −r on Ik by (95), so u − w > −r by (98). The proof of the second equivalence is analogous. (cid:3) + by (95). Conversely, if (u, p) ∈ Z Ik We specify some properties of points in the "boundary sets" I∂±. Proposition 7.6. Let u ∈ C ∆ and z0 ∈ R, set p = Pr[u; z0](tk−1), k ≥ 1 and w = P Ik (u, p). Then the following holds. (i) Let t ∈ I∂+ ∩ Ik. Then (u, p) ∈ Z Ik + and u(t) = (F Ik u)(t) ≥ p + r . (100) (ii) Let τ ∈ I∂+ ∩ Ik, τ ≤ t ≤ tk. Then M Ik (u, t) ∩ [τ, t] ⊂ I∂+. (iii) Let (u, p) ∈ Z Ik τ ∈ [tk−1, t] with τ ∈ I∂+. (iv) Let (u, p) ∈ Z Ik M Ik (u, t) ⊂ (s, t]. + , let t ∈ Ik with (F Ik u)(t) ≥ p + r. Then there exists + , let s, t ∈ Ik with s < t and w(s) < w(t). Then NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 21 Proof. (i) Let t ∈ I∂+ ∩ Ik, so u(t) − w(t) = r. We have (u, p) ∈ Z Ik otherwise by (98), (97) and (96) w(t) = P Ik (u, p)(t) = P Ik − (u, p)(t) = u(t) − z−(t) > u(t) − r , + since a contradiction. Since z+(t) = r, the remaining assertions are a direct consequence of Lemma 7.1. (ii) Let s ∈ M Ik (u, t) with s ≥ τ . As F Ik (u − p − r) is nondecreasing and F Ik (u − p − r)(τ ) ≥ 0 by (100), we have w(s) − p = max{0, F Ik (u − p − r)(s)} = F Ik (u − p − r)(s) ≤ F Ik (u − p − r)(t) = u(s) − p − r . As u(s) − w(s) ≤ r it follows that u(s) − w(s) = r and therefore s ∈ I∂+. (iii) As F Ik (u − p − r)(tk−1) = u(tk−1) − w(tk−1) − r ≤ 0, we find σ ∈ [tk−1, t] with F Ik (u − p − r)(σ) = 0. As F Ik (u − p − r) is nondecreasing, we have F Ik (u − p − r) ≤ 0 on [tk−1, σ] and therefore w = P Ik + (u, p) = p = p + (F Ik (u − p − r))(σ) = (F Ik u)(σ) − r on [tk−1, σ]. We choose τ ∈ [tk−1, σ] with u(τ ) = (F Ik u)(σ). Then w(τ ) = u(τ ) − r, so τ ∈ I∂+. (iv) By (98), max{p, F Ik (u − r)(s)} = w(s) < w(t) = max{p, F Ik (u − r)(t)}, so (F Ik u)(s) < (F Ik u)(t) and therefore M Ik (u, t) ⊂ (s, t]. (cid:3) For minus intervals, the corresponding results read as follows. Their proofs are analogous to those of Proposition 7.6. Proposition 7.7. Let u ∈ C ∆ and z0 ∈ R, set p = Pr[u; z0](tk−1), k ≥ 1 and w = P Ik (u, p). Then the following holds. (i) Let t ∈ I∂− ∩ Ik. Then (u, p) ∈ Z Ik − and u(t) = −(F Ik (−u))(t) ≤ p − r . (101) (ii) Let τ ∈ I∂− ∩ Ik, τ ≤ t ≤ tk. Then M Ik (u, t) ∩ [τ, t] ⊂ I∂−. (iii) Let (u, p) ∈ Z Ik exists τ ∈ [tk−1, t] with τ ∈ I∂−. (iv) Let (u, p) ∈ Z Ik M Ik (−u, t) ⊂ (s, t]. − , let s, t ∈ Ik with s < t and w(s) > w(t). Then ✷ − , let t ∈ Ik with −(F Ik (−u))(t) ≤ p − r. Then there A Newton derivative on an interval of small input oscillation. We want to obtain a Newton derivative for P I : Z I ∩ (X × R) → Lq(I), where I = [t∗, t∗] ⊂ [a, b]. The mapping P I + decomposes into +(u, p) = p + Fpp( F I (u, p)) . P I Here, F I : Z I ∩ (X × R) → C(I) is defined as F I (u, p) = F I (u − p − r) , and Fpp denotes the positive part mapping (Fppu)(t) = max{0, u(t)} . (102) (103) (104) 22 M. BROKATE We first analyze the mapping F I . We expect to obtain a Newton derivative GI of F I if we choose elements L ∈ GI (u, p) of the form (L(h, η))(t) = hµI (u, t) , h − ηi , (105) where µI (u, t) are probability measures arising from the derivative of the cumulated maximum on I. We define T : C(I)×R → C(I) by T (u, p) = u−p and consider the set-valued mapping ΨI : Z I × I ⇒ (C(I) × R)∗ ΨI (u, p, t) = ΦI (T (u, p) − r, t) ◦ T . (106) ΦI is the mapping defined in (48) with [a, b] and M replaced with I and M I from (85). We compute ΨI , using (50) and (51), ΨI (u, p, t) = ΦI (u, t) ◦ T = ΦI (u, t) ◦ (π1 − j ◦ π2) = ΦI (u, t) ◦ π1 − π2 . (107) Here, π1, π2 : C(I) × R → R denote the projections π1(h, η) = h and π2(h, η) = η, and j maps real numbers to the corresponding constant func- tions. We see that ΨI(u, p, t) actually does not depend on p. Let SI Φ be the set of all measurable selectors of ΦI. Proposition 7.8. The mapping ΨI is usc and has w∗-compact values. Moreover, SI Ψ = {µ : µ(u, p, t) = µI (u, t) ◦ π1 − π2, µI ∈ SI (108) is a set of measurable selectors of ΨI . For F I : Z I ∩ (X × R) → Lq(I) with q < ∞, a Newton derivative GI is given by Φ} GI : Z I ∩ (X × R) ⇒ L(X × R, Lq(I)) GI (u, p) = {L : L has the form (105) with µI ∈ SI Φ} . The elements L of GI (u, p) satisfy kL(h, η)k∞,t ≤ khk∞,t + η for all h ∈ C(I), η ∈ R, t ∈ I. Moreover, the remainder estimate k F I (u + h, p + η) − F I (u, p) − L(h, η)kL q (I) sup L∈ GI (u+p,h+η) (109) (110) (111) ≤ ρ(u,p)(khk∞ + η)k(h, η)kX×R holds. The remainder term ρ(u,p) satisfies ρ(u,p)(δ) ↓ 0 as δ ↓ 0 and is uniformly bounded in (u, p). Proof. Let A : C(I)∗ → (C(I) × R)∗, A(ν) = ν ◦ π1 − π2. Then A is linear and w∗-w∗-continuous, and ΨI = A ◦ ΦI . Since ΦI is usc according to Proposition 4.5 and has w∗-compact values, using Lemma 10.5 we see that the same is true for ΨI . The elements of SI Ψ are measurable as compositions of measurable func- tions. As F I has a Newton derivative given by Proposition 4.8 and F I (u, p) = F I (u − p − r), setting X = Y = C(I) and Z = Lq(I) we check that the as- sumptions of the refined chain rule, Proposition 5.2, are satisfied. Therefore, GI is a Newton derivative of F I and (111) holds. (110) is a consequence NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 23 of (105) and (53). Since F is globally Lipschitz w.r.t. the maximum norm, together with (110) the final assertion follows. (cid:3) Since ΦI (v, t∗) = {δt∗ } for all v, by (108) we have for all µ ∈ SI Ψ hµ(u, p, t∗), (h, η)i = h(t∗) − η (112) for all (u, p) ∈ Z I and all (h, η) ∈ C(I) × I. For functions u : I → R, we consider the positive part mapping Fpp defined by (113) which maps Lq(I) as well as C(I) into itself. Let H : R ⇒ R be the set- valued Heaviside function (Fppu)(t) = max{0, u(t)} , H(x) =  The mapping H is usc. By 0 , x < 0 , [0, 1] , x = 0 , 1 , x > 0 . (114) SH = {λH : λH selector of H, λH(0) ∈ Q} (115) we define a countable family of measurable selectors of H whose values are dense in the range of H. We then define Gpp(u) : Lq(I) ⇒ L(Lq(I), Lq(I)) Gpp(u) = {L : L(h) = (λH ◦ u) · h, λH ∈ SH} . (116) Lemma 7.9. The mapping Gpp is a Newton derivative of Fpp : Lq(I) → Lq(I) for 1 ≤ q < q ≤ ∞. Proof. This is a well-known result, see Proposition 3.49 in [17] or Example 8.14 in [10]. (cid:3) With the composition Fpp ◦ F I we associate the set-valued mapping ΨI pp : C(I) × R × I ⇒ R pp(u, p, t) = H( F I (u, p)(t)) . ΨI By the definition of H, ΨI pp(u, p, t) = {0} , if F I (u, p)(t) < 0. (117) (118) Lemma 7.10. The mapping ΨI values. A set of measurable selectors is given by pp defined in (118) is usc and has compact pp = {λ : λ(u, p, t) = λH( F I (u, p)(t)), λH ∈ SH} . SI (119) Proof. The mapping F I : C(I) × R → C(I) as well as the mapping (v, t) 7→ v(t) are continuous on C(I) × R, and H is usc and has compact values. Therefore, ΨI pp is usc by Lemma 10.5, has compact values, and the elements of SI (cid:3) pp are measurable functions. 24 M. BROKATE We have now all ingredients to define a Newton derivative GI + of P I +. Its elements LI +(u, p) are expected to have the form + ∈ GI +(h, η)(t) = η + λH( F I (u, p)(t)) · hµI (u, t), h − ηi LI (120) with functions λH ∈ SH and measures µI ∈ SI mapping is given by Φ. The associated set-valued + : C(I) × R × I ⇒ (C(I) × R)∗ ΨI +(u, p, t) = π2 + ΨI ΨI pp(u, p, t) · ΨI (u, p, t) , (121) where π2 denotes the projection π2 : C(I) × R → R, π2(h, η) = η. (The elementwise multiplication ΨI pp takes values in R.) We define pp · ΨI makes sense since ΨI + = {ν : ν(u, p, t) = π2 + λ(u, p, t)µ(u, p, t), µ ∈ SI SI Ψ, λ ∈ SI pp} . (122) Proposition 7.11. The mapping ΨI values. The set SI The mapping + in (121) is usc and has w∗-compact + given in (122) consists of measurable selectors of ΨI +. GI + : Z I +(u, p) = {LI + ∩ (X × R) ⇒ L(X × R, Lq(I)) + : LI + given by (120) with λH ∈ SH, µI ∈ SI Φ} GI (123) is a Newton derivative of P I elements LI + of GI + : Z I + ∩ (X × R) → Lq(I) for every q < ∞. The +(u, p) satisfy, for all (u, p) ∈ Z I + ∩ (X × R), kLI +(h, η)k∞,t ≤ max{khk∞,t, η} (124) for all h ∈ C(I), η ∈ R. Moreover, for all such (u, p) the remainder estimate sup +(u+h,p+η) LI +∈GI kP I +(u + h, p + η) − P I +(u, p) − LI +(h, η)kLq (I) (125) ≤ ρ(u,p)(khk∞ + η)k(h, η)kX×R holds for all h ∈ X with u + h ∈ Z I ρ(u,p) satisfies ρ(u,p)(δ ↓ 0 as δ ↓ 0 and is uniformly bounded in (u, p). + and all η ∈ R. The remainder term Proof. Due to Proposition 7.8 and Lemma 7.10, we may apply Proposi- tion 10.8 with F1(u, p, t) = F I (u, p)(t), Ψ1 = ΨI and Ψ2(u, p, t) = π2 + ΨI pp(u, p, t)π3, that is, the elements of Ψ2 have the form pp ∈ ΨI L2(h, η, y) = η + Lt + is usc and has w∗-compact values. pp · y , Lt This shows that ΨI pp . Due to Proposition 7.8 and Lemma 7.9, the assumptions of the refined chain rule, Proposition 5.2, are satisfied with X = C(I), Y = Y = Lq(I) for some ∞ > q > q, Z = Lq(I). This proves (125). The estimate (124) follows from (120) and (53), as λH takes values in [0, 1] and, setting λt = λH( F I (u, p)(t)), LI +(h, η)(t) = (1 − λt)η + λthµI (u, t), hi . Since P I assertion, too, follows in view of (124). + is global Lipschitz continuous w.r.t. the maximum norm, the final (cid:3) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 25 We also need a variant of the preceding proposition. For I = [t∗, t∗] we define P I +,∗ : Z I + → R , P I +,∗(u, p) = P I +(u, p)(t∗) . According to (120), setting LI +,∗(h, y) = LI yields a well-defined element LI +(h, y)(t∗) , LI +,∗ ∈ (C(I) × R)∗. + ∈ GI +(u, p) , Proposition 7.12. The mapping (126) (127) + ∩ (X × R) ⇒ (X × R)∗ +,∗ : Z I GI +,∗(u, p) = {LI GI is a Newton derivative of P I GI +,∗ : Z I +,∗(u, p) satisfy, for all (u, p) ∈ Z I +,∗ given by (127)} +,∗ : LI + ∩ (X × R) → R. The elements LI + ∩ (X × R), (128) +,∗ of LI +(h, η) ≤ max{khk∞,t∗ , η} (129) for all h ∈ C(I), η ∈ R. Moreover, for all such (u, p) the remainder estimate sup +,∗(u+h,p+η) LI +,∗∈GI P I +,∗(u + h, p + η) − P I +,∗(u, p) − LI +,∗(h, η) (130) ≤ ρ(u,p)(khk∞ + η)k(h, η)kX×R holds for all h ∈ X with u + h ∈ Z I ρ(u,p) satisfies ρ(u,p)(δ) ↓ 0 as δ ↓ 0 and is uniformly bounded in (u, p). + and all η ∈ R. The remainder term Proof. We proceed in a manner analogous to the proof of Proposition 7.11. We apply Proposition 5.2 to the decomposition P I +,∗(u, p) = p + max{0, max (u − p − r)} I The Newton derivative of the inner maximum satisfies the refined remainder estimate given in Proposition 3.4. The outer maximum is just the positive part mapping on R. (cid:3) A Newton derivative GI − of the mapping −(u, p) = p − Fpp( F I (−u, −p)) P I (131) is obtained with analogous computations. Its elements LI the form −(h, η)(t) = η − λH( F I (−u, −p)(t)) · hµI (−u, t), −h + ηi LI − ∈ GI −(u, p) have (132) with functions λH ∈ SH and measures µI ∈ SI mapping ΨI − and a set SI − of measurable selectors is given by Φ. The associated set-valued ΨI − : C(I) × R × I ⇒ (C(I) × R)∗ ΨI −(u, p, t) = π2 − ΨI pp(−u, −p, t) · (− ΨI(−u, −p, t)) , − = {ν : ν(u, p, t) = π2 + λ(−u, −p, t)µ(−u, −p, t), µ ∈ SI SI Ψ, λ ∈ SI pp} . (133) where π2 again denotes the projection π2 : C(I) × R → R, π2(h, η) = η. The analogue of Proposition 7.12 also holds on minus intervals. 26 M. BROKATE We combine GI ± and ΨI ± into mappings GI and ΨI . Indeed, on Z I + ∩ Z I −, we have F I (u, p) < 0 and F I (−u, −p) < 0 by Lemma 7.2. Consequently, ΨI pp(u, p, t) = {0} , ΨI ±(u, p, t) = {π2} + ∩ Z I for all (u, p) ∈ Z I −, t ∈ I. The argument of λH in the representations (120) and (132) is negative, therefore LI ±(h, η)(t) = η on I. As the sets Z I ± are open subsets of Z I, from Proposition 7.11 and the corresponding result for P I − we get the following result. Proposition 7.13. The mapping ΨI : Z I × I ⇒ (C(I) × R)∗ defined by ± is well-defined and usc and has w∗-compact values. The ΨI = ΨI set ± on Z I ±, (u, p, t) ∈ Z I SI = {ν : ν(u, p, t) = ν±(u, p, t) with ν± ∈ SI (134) consists of measurable selectors of ΨI. The mapping GI : Z I ∩ (X × R) ⇒ L(X × R, Lq(I)) given by GI = GI ± ∩ (X × R) is well-defined and is a Newton derivative of P I : Z I ∩ (X × R) → Lq(I). The estimates (124) and (125) hold with GI , P I , Z I and LI in place of GI + and LI +, respectively. The remainder term ρ(u,p) satisfies ρ(u,p)(δ ↓ 0 as δ ↓ 0 and is uniformly bounded in (u, p). ✷ ± on Z I +, P I +, Z I ± × I} The initial value. According to (76), the initial value of the play is given by w0(u, z0) = u(a) − πr(z0) = u(a) − max{−r, min{r, z0}} . (135) It is well known that the mapping R : R ⇒ R, 0 , [0, 1] , 1 , x > r , x = r , x < r (136) R(x) =  is a Newton derivative of πr and that R is usc. Then SR = {λ0 : λ0 selector of R, λ0(±r) ∈ Q} (137) defines a countable family of measurable selectors of R. We set Ψ0 : C[a, b] × R ⇒ (C[a, b] × R)∗ Ψ0(u, z0) = {δa} × (−R(z0)) . (138) Lemma 7.14. A Newton derivative of w0 : C[a, b] × R → R is given by G0 : C[a, b] × R ⇒ (C[a, b] × R)∗ , G0(u, z0) = {L : L(h, y) = h(a) − λ0(z0)y , λ0 ∈ SR} . We have L(h, y) ≤ khk∞ + y for all L ∈ G0(u, z0) and all (u, z0) ∈ C[a, b] × R. (139) (140) Proof. Let L ∈ G0(u, z0). Then for all (h, y) ∈ C[a, b] × R we have w0(u + h, z0 + y) − w0(u, z0) − L(h, y) = (πr(z0 + y) − πr(z0) − λ0(z0)y) (141) ≤ ρ(y)y with some ρ(δ) ↓ 0 as δ ↓ 0, since R is a Newton derivative of πr. (cid:3) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 27 A Newton derivative on a partition for small input oscillations. Let ∆ = {tk}0≤k≤N be a partition of [a, b]. According to Lemma 7.5, on the set Z ∆ of small input oscillations, see (84), the play can be written as a composition of the mappings P Ik which belong to the partition intervals Ik = [tk−1, tk]. Consequently, we obtain a Newton derivative of the play on Z ∆ as a composition of the Newton derivatives of P Ik as follows. k : Z ∆ → C(Ik), setting w∆ 0 = w0 and We define w∆ 0 = Ψ0 from (135) and (138), and for k ≥ 1 k : Z ∆ → R and P ∆ Ψ∆ k (u, z0) = P Ik (u, w∆ w∆ k (u, z0)(t) = P Ik (u, w∆ P ∆ k−1(u, z0))(tk) , k−1(u, z0))(t) , t ∈ Ik . (142) Using Lemma 7.5 successively we see that P ∆ k (u, z0) = Pr[u; z0] on Ik. We define Ψ∆ k : Z ∆ × Ik ⇒ (C[a, b] × R)∗ by Ψ∆ 0 = Ψ0 and, for k ≥ 1, Ψ∆ k (u, z0, t) = {L ◦ (π1, Lk−1) : L ∈ ΨIk(u, w∆ k−1(u, z0), t), Lk−1 ∈ Ψ∆ k−1(u, z0, tk−1)} (143) Here π1 denotes the projection π1 : C[a, b] × R → C(Ik), π1(u, z0) = uIk. The mappings ΨIk have been constructed in Proposition 7.13. The elements Lk of Ψ∆ k (u, z0, t) have the form Lk(h, y) = L(h, Lk−1(h, y)) , h ∈ C[a, b], y ∈ R . (144) The sets S∆ 0 = SR, k : µ∆ k = {µ∆ S∆ k (u, z0, t) = ν(u, p, t) ◦ (π1, µ∆ ν ∈ SIk , p = w∆ k−1(u, z0), µ∆ k−1(u, z0, tk−1)), k−1 ∈ S∆ k−1} , k ≥ 1 , (145) (146) (147) (148) (149) (150) consist of measurable selectors of Ψ∆ k . We define W ∆ W ∆ 0 = G0 and inductively for k ≥ 1 k : Z ∆ ⇒ (C[a, b] × R)∗ , k : Lw W ∆ k (u, z0) = {Lw k ∈ W ∆ k (u, z0) satisfy Lw k (h, y) = LIk(h, Lw k = µ∆ k−1(h, y))(tk) , k (u, z0, tk) with µ∆ k ∈ S∆ k } . The elements Lw LIk ∈ GIk (u, wk−1(u, z0)) , Lw k−1 ∈ W ∆ k−1(u, z0) . We define G∆ 0 = G0 and inductively for k ≥ 1 G∆ k : Z ∆ ⇒ L(C[a, b] × R, L∞(Ik)) , G∆ k satisfies (149) for some µ∆ k : L∆ k (u, z0) = {L∆ k (h, y)(t) = hµ∆ L∆ k (u, z0, t) , (h, y)i . k ∈ S∆ k } , The mappings L∆ k satisfy k (h, y)(t) = LIk(h, Lw L∆ LIk ∈ GIk (u, wk−1(u, z0)) , Lw k−1(h, y))(t) , k−1 ∈ W ∆ k−1(u, z0) . 28 M. BROKATE Proposition 7.15. Let 0 ≤ k ≤ N , 1 ≤ q < ∞. (i) The mapping Ψ∆ values, S∆ (ii) The mapping W ∆ of w∆ k is a set of measurable selectors of Ψ∆ k . k : Z ∆ ∩ (X × R) → R. The elements Lw k : Z ∆∩(X ×R) ⇒ (C[a, b]×R)∗ is a Newton derivative k satisfy the estimate k of W ∆ k : Z ∆ × Ik ⇒ (C[a, b] × R)∗ is usc and has w∗-compact Lw k (h, y) ≤ khk∞,tk + y (151) for all h ∈ C[a, b] and y ∈ R, uniformly in (u, z0). Moreover, for all such (u, z0) the remainder estimate sup k (u+h,z0+y) Lw k ∈W ∆ w∆ k (u + h, z0 + y) − w∆ k (u, z0) − Lw k (h, y) (152) ≤ ρ(u,z0)(khk∞ + y)k(h, y)kX×R holds for all h ∈ X with u + h ∈ Z I ρ(u,z0) satisfies ρ(u,z0)(δ) ↓ 0 as δ ↓ 0 and is uniformly bounded in (u, z0). (iii) The mapping G∆ derivative of P ∆ k : Z ∆ ∩ (X × R) ⇒ L(X × R, Lq(Ik)) is a Newton + and all y ∈ R. The remainder term k (u, z0) satisfy the estimate k . The elements L∆ k of G∆ kL∆ k (h, y)k∞,t ≤ khk∞,t + y (153) for all h ∈ C[a, b] and y ∈ R, uniformly in (u, z0). Moreover, for all such (u, z0) the remainder estimate sup k (u+h,z0+y) Lw k ∈W ∆ P ∆ k (u + h, z0 + y) − P ∆ k (u, z0) − L∆ k (h, y) (154) ≤ ρ(u,z0)(khk∞ + y)k(h, y)kX×R holds for all h ∈ X with u + h ∈ Z I ρ(u,z0) satisfies ρ(u,z0)(δ) ↓ 0 as δ ↓ 0 and is uniformly bounded in (u, z0). + and all y ∈ R. The remainder term Proof. We proceed by induction over k. The case k = 0 is treated in Lemma 7.14. Now assume the result is proved for k − 1. (i) We apply Proposition 10.8, setting there J = [a, b], I = Ik, U = C ∆ and F1(u, z0, t) = w∆ k−1(u, z0) , F1 : C ∆ × R × Ik → R , F2(u, z0, p, t) = P Ik (u, p)(t) , F2 : C ∆ × R × R × Ik → R , k−1(u, z0, tk−1) , Ψ2(u, z0, p, t) = ΨIk (u, p, t) , Ψ = Ψ∆ k . Ψ1(u, z0, t) = Ψ∆ (ii) We apply Proposition 5.2 to the decomposition k−1(u, z0)) 7→ w∆ (u, z0) 7→ (u, w∆ k (u, z0) given by the first equation in (142). Its assumptions are satisfied by the in- duction hypothesis and by Proposition 7.12. (151) follows from the estimate Lw k (h, y) ≤ kLIk (h, Lw k−1(h, y))k∞,tk ≤ max{khk∞,tk , khk∞,tk + y} = khk∞,tk + y . (iii) This follows as in (ii), using Proposition 7.13 instead of Proposition 7.12, as well as the estimate kL∆ k (h, y)k∞,t = kLIk (h, Lw k−1(h, y))k∞,t ≤ max{khk∞,t, khk∞,t + y} = khk∞,t + y . NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 29 (cid:3) A Newton derivative of the play on the whole space X × R. Let {∆n} be a sequence of partitions of [a, b] such that ∆n → 0 as n → ∞ and that ∆n+1 is obtained from ∆n by adding a single point t /∈ ∆n, starting from ∆1 = {a, b}. We have Z ∆n ⊂ Z ∆n+1 , C[a, b] × R = [n∈N Z ∆n (155) and consequently X × R = [n∈N (Z ∆n ∩ (X × R)) = [n∈N (X ∆n × R) . (156) of P ∆n We construct a Newton derivative G of the play on X × R from the Newton derivatives G∆n obtained in Proposition 7.15. This is done in two steps. In the first step, we define a Newton derivative G∆n of the play on Z ∆n ∩ (X × R); in the second step we glue together these derivatives using Proposition 2.5. k k Since the elements of G∆n set-valued mappings Ψ∆n order to do this, we consider the following situation. k are obtained as measurable selectors of the In k , we first combine those to a mapping Ψ∆n. Let ∆ = {tj}, a = t0 < · · · < tN = b, be a partition of [a, b], let Ij = [tj−1, tj] for 1 ≤ j ≤ N , where N ≥ 1. If N > 1, let ∆′ be the partition which results from ∆ when we remove a single point tk, k ∈ {1, . . . , N − 1}. Proposition 7.16. (i) If N > 1, Ψ∆ j (u, z0, t) = Ψ∆′ (u, z0, t) for all (u, z0) ∈ Z ∆′ (ii) The mapping Ψ∆ : Z ∆ × [a, b] ⇒ (C[a, b] × R)∗ defined by , t ∈ Ij, 1 ≤ j ≤ N . Ψ∆(u, z0, t) = Ψ∆ j (u, z0, t) , if t ∈ Ij, is well-defined, usc and has w∗-compact values. (iii) The set S∆ = {µ∆ : µ∆ selector of Ψ∆, µ∆ = µ∆ 1 on Z ∆ × [t0, t1], µ∆ = µ∆ k on Z ∆ × (tk−1, tk] for k > 1, µ∆ k ∈ S∆ k } (157) (158) (159) consists of measurable selectors of Ψ∆. Proof. The proof proceeds by induction on N , the number of partition points of ∆ being equal to N + 1. For N = 1 we have Ψ∆ = Ψ∆ 1 , (i) is empty, and (ii) has been proved already in Proposition 7.15. For the induction step N − 1 → N we assume that Ψ∆′ is well-defined. In order to prove (157), it suffices to show that Ψ∆ k+1 = Ψ∆′ (160) Indeed, as the refinement by adjoining {tk} to ∆′ does not change the par- j = Ψ∆′ tition intervals Ij contained in [a, tk−1] and in [tk+1, b], we have Ψ∆ j on Z ∆′ × Ij for all j ≤ k, and (once we have shown that Ψ∆ k on k+1 = Ψ∆′ × Ik+1. k on Z ∆′ 30 M. BROKATE j+1 = Ψ∆′ Z ∆′ × {tk+1}) also Ψ∆ be proved in Lemma 7.17 below. j on Z ∆′ × Ij+1 for all k < j < N . (160) will We now prove (ii). For (u, z0) ∈ Z ∆′ , (158) holds by (i). Let (u, z0) ∈ Z ∆. We define u ∈ C[a, b] by u(t) =(u(t) , u(tN −1) , t ∈ [a, tN −1] , t ∈ IN = [tN −1, b] . Then (u, z0) ∈ Z ∆′ . By (i), we have j (u, z0, t) = Ψ∆ j (u, z0, t) = Ψ∆′ holds for all t ∈ Ij], 1 ≤ j ≤ N − 1. In particular, Ψ∆ (u, z0, t) Ψ∆ j+1(u, z0, tj) = Ψ∆ j (u, z0, tj) for all 1 ≤ j ≤ N − 1. This shows that Ψ∆ is well-defined by (158) if (u, z0) ∈ Z ∆. That Ψ∆ is usc and has w∗-compact values now follows from Lemma 10.4. To prove (iii) it suffices to observe that the functions µ∆ k are measurable. (cid:3) The proof of Proposition 7.16 is based on Lemma 7.17 which in turn is based on Lemma 7.18. The proof of Lemma 7.18 only uses results derived before and up to Proposition 7.15. Lemma 7.17. We have Ψ∆ k+1(u, z0, t) = Ψ∆′ k (u, z0, t) (161) and all t ∈ Ik+1. for all (u, z0) ∈ Z ∆′ Proof. Let (u, z0) ∈ Z ∆′ and t ∈ Ik+1 = [tk, tk+1] be given, set I ′ = [tk−1, t]. Moreover, set w = Pr[u; z0], pk−1 = w(tk−1) and pk = w(tk). We assume that (u, pk−1) ∈ Z I ′ + , that is, I ′ is a plus interval by (99). (The case of a minus interval is treated analogously.) By (143), the elements of Ψ∆′ k (u, z0, t) have the form L′ ◦ (π1, Lk−1) with L′ ∈ ΨI ′ +(u, pk−1, t) =: Ψ′ , Lk−1 ∈ Ψ∆′ k−1(u, z0, tk−1) . (162) For the same reason, the elements of Ψ∆ with k+1(u, z0, t) have the form L◦(π1, Lk) L ∈ ΨIk+1 + (u, pk, t) =: Ψ , Lk ∈ Ψ∆ k (u, z0, tk) , (163) and the elements Lk of Ψ∆ k (u, z0, tk) have the form L′′ ◦ (π1, Lk−1) with L′′ ∈ ΨIk + (u, pk, tk) =: Ψ′′ , Lk−1 ∈ Ψ∆ (164) k−1 on Z ∆′ × Ik−1, in order to prove (161) it suffices to prove k−1(u, z0, tk−1) . Since Ψ∆ that k−1 = Ψ∆′ Ψ′ = {L ◦ (π1, L′′) : L ∈ Ψ, L′′ ∈ Ψ′′} . This is done in Lemma 7.18 below. (165) (cid:3) Let A1, A2, A3 be sets, let Fi be sets of mappings from Ai to Ai+1, i = 1, 2. We define the elementwise composition F2 ◦ F1 = {f2 ◦ f1 : f1 ∈ F1, f2 ∈ F2} . (166) NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 31 We also provide a more explicit representation of ΨI (117) into (121) yields +. Inserting (107) and ΨI +(u, p, t) = π2 + H( F I (u, p)(t)) · (ΦI (u, t) ◦ π1 − π2) . (167) Here, ΦI (u, t) ◦ π1 := {L ◦ π1 : L ∈ ΦI (u, t)}. Setting σ = F I (u, p)(t) we get ΨI +(u, p, t) =  {π2} , σ < 0 , ΦI (u, t) ◦ π1 , σ > 0 , co({π2} ∪ (ΦI (u, t) ◦ π1)) , σ = 0 . (168) (170) (171) Lemma 7.18. Let Ψ, Ψ′ and Ψ′′ be the subsets of (C[a, b] × R)∗ defined in (162) -- (164). Then Ψ′ = Ψ ◦ (π1, Ψ′′) . (169) Proof. We continue to use the notations from the proof of Lemma 7.17; again, the case of a minus interval is treated analogously. We define and obtain from (168) σ′ = F I ′ (u − pk−1 − r)(t) {π2} , ΦI ′ (u, t) ◦ π1 , co({π2} ∪ (ΦI ′ σ′ < 0 , σ′ > 0 , (u, t) ◦ π1)) , σ = 0 . Ψ′ =  For Ψ and Ψ′′ corresponding formulas hold; we replace σ′ with σ′′ = F Ik(u − pk−1 − r)(tk) , σ = F Ik+1(u − pk − r)(t) , (172) (u, t) by ΦIk+1(u, t) resp. ΦIk (u, tk). From the definitions we imme- and ΦI ′ diately see that σ′′ ≤ σ′ and pk = pk−1 ⇒ σ′ = max{σ′′, σ} , as well as σ ≤ 0 ⇒ w = pk σ′ ≤ 0 ⇒ w = pk−1 = pk σ′′ ≤ 0 ⇒ w = pk−1 on [tk, t], on Ik. on I ′, (173) (174) (175) (176) In order to prove (169), we have to distinguish several cases. Case 1: σ′ < 0. Then σ < 0 and σ′′ < 0 by (175) and (173), so Ψ = Ψ′ = Ψ′′ = {π2}, and (169) holds. Case 2: σ′ = 0. As in Case 1, we have w = pk−1 = pk (177) Subcase 2a: σ′′ < 0. Then σ′′ < 0 = σ, so (F Ik u)(tk) < (F Ik+1u)(t) since pk−1 = pk. Therefore, 0 = σ′ = max{σ′′, σ} . on I ′, M I ′ (u, t) = M Ik+1(u, t) , ΦI ′ (u, t) = ΦIk+1(u, t) , Ψ′ = Ψ . As Ψ′′ = {π2}, (169) holds. Subcase 2b: σ′′ = 0. By Proposition 7.6(iii), there exists τ ′′ ∈ Ik such that (178) u(τ ′′) − w(τ ′′) = r = u(τ ′′) − pk . 32 M. BROKATE Subsubcase 2b1: σ < 0. Then on [tk, t] we have u ≤ F Ik+1(u) < r + pk = u(τ ′′), so M I ′ (u, t) = M Ik (u, tk) , ΦI ′ (u, t) = ΦIk(u, tk) , Ψ′ = Ψ′′ . As Ψ = {π2}, (169) holds. Subsubcase 2b2: σ = 0. By Proposition 7.6(iii), there exists τ ∈ [tk, t] such that u(τ ) − w(τ ) = r. Since w is constant on I ′, we have τ ∈ M I ′ For the same reason, (u, t) , M I ′ (u, t) ∩ Ik+1 = M Ik+1(u, t) . τ ′′ ∈ M I ′ (u, t) , M I ′ (u, t) ∩ Ik = M Ik (u, tk) . This gives Setting M I ′ (u, t) = M Ik (u, tk) ∪ M Ik+1(u, t) . (179) Φ = ΦIk+1(u, t) , Φ′ = ΦI ′ (u, t) , Φ′′ = ΦIk(u, tk) , it follows from (179) that Φ′(u, t) = co ( Φ′′ ∪ Φ) . Since the sets involved are convex and the mappings involved are linear, we can compute Ψ ◦ (π1, Ψ′′) = co ({π2} ∪ ( Φ ◦ π1)) ◦ (π1, Ψ′′) = co (Ψ′′ ∪ ( Φ ◦ π1)) = co ({π2} ∪ ( Φ′′ ◦ π1) ∪ ( Φ ◦ π1)) = co ({π2} ∪ ( Φ′ ◦ π1)) = Ψ′ . Case 3: σ′ > 0. We have pk−1 + σ′ = w(t) = pk + max{0, σ} . (180) Therefore w(t) > pk−1, and by Proposition 7.6(iii), there exists τ ′ ∈ I ′ such that u(τ ′) − w(τ ′) = r. Subcase 3a: σ > 0. Then w(t) > pk = w(tk), so M I ′ therefore (u, t) = M Ik+1(u, t) and Ψ = ΦIk+1(u, t) ◦ π1 = ΦI ′ (u, t) ◦ π1 = Ψ′ . Since moreover Ψ ◦ (π1, Ψ′′) = Ψ, (169) holds. Subcase 3b: σ ≤ 0. Then pk−1 < w(t) = pk = pk−1 + max{0, σ′′}, so σ′′ > 0 and Ψ′′ = ΦIk (u, tk) ◦ π1. Subsubcase 3b1: σ < 0. On [tk, t] we have u − w ≤ (F Ik+1u) − w < r. Since u(τ ′) − w(τ ′) = r, it follows that M I ′ (u, t) = M Ik (u, tk). As Ψ = {π2}, Ψ ◦ (π1, Ψ′′) = Ψ′′ = ΦIk (u, tk) ◦ π1 = ΦI ′ (u, t) ◦ π1 = Ψ′ . Subsubcase 3b2: σ = 0. By Proposition 7.6(iii), there exists τ ∈ [tk, t] such that r = u(τ ) − w(τ ) = u(τ ) − pk. On I ′ we have u ≤ w + r ≤ w(τ ) + r = u(τ ) , since w is nondecreasing on I ′ and constant on [tk, t]. This implies, since pk−1 < pk and σ = 0, pk = w(tk) = max{pk−1, F Ik (u − r)(tk)} = F Ik (u − r)(tk) = u(τ ) − r ≤ F Ik+1(u − r)(t) = pk , NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 33 so Consequently, (F Ik+1u)(t) = (F Ik u)(tk) = (F I ′ u)(t) . M I ′ (u, t) = M Ik (u, tk) ∪ M Ik+1(u, t) . The proof of (169) now proceeds as in Subsubcase 2b2, the final computation being modified to Ψ ◦ (π1, Ψ′′) = co ({π2} ∪ ( Φ ◦ π1)) ◦ (π1, Ψ′′) = co (Ψ′′ ∪ ( Φ ◦ π1)) = co (( Φ′′ ◦ π1) ∪ ( Φ ◦ π1)) = Φ′ ◦ π1 = Ψ′ . (cid:3) In order to define a Newton derivative G∆ of the play on Z ∆ ∩ (X × R), we set G∆ : Z ∆ ⇒ L(C[a, b] × R, L∞(a, b)) , G∆(u, z0) = {L∆ : L∆(h, y)(t) = hµ∆(u, z0, t), (h, y)i (181) on [a, b] for some µ∆ ∈ S∆} . Proposition 7.19. Let 1 ≤ q < ∞. The mapping G∆ : Z ∆∩(X ×R) ⇒ L(X ×R, Lq(a, b)) is a Newton derivative of the play Pr : Z ∆ ∩ (X × R) → Lq(a, b). The elements L∆ of G∆(u, z0) satisfy the estimate kL∆(h, y)k∞,t ≤ khk∞,t + y (182) for all h ∈ C[a, b] and y ∈ R, uniformly in (u, z0). Moreover, for all such (u, z0) the remainder estimate sup L∆∈G∆(u+h,z0+y) kPr[u + h; z0 + y] − Pr[u; z0] − L∆(h, y)kLq (a,b) (183) ≤ ρ(u,z0)(khk∞ + y)k(h, y)kX×R + and all y ∈ R, where ρ(u,z0)(δ) ↓ 0 as holds for all h ∈ X with u + h ∈ Z I δ ↓ 0 and ρ(u,z0) is uniformly bounded in (u, z0). Proof. Let (u, z0) ∈ Z ∆ ∩ (X × R), (h, y) ∈ X × R with khkX small enough, and L∆ ∈ G∆(u + h, z0 + y). Since Pr[u; z0] = P ∆ k (u, z0) and Pr[u + h; z0 + k (u + h, z0 + y) on Ik, we have in view of the definition of S∆, S∆ y] = P ∆ k and G∆ k kPr[u + h; z0 + y] − Pr[u; z0] − L∆(h, y)kq Lq (a,b) N kP ∆ k (u + h, z0 + y) − P ∆ k (u, z0) − L∆ k (h, y)kq Lq (Ik) = Xk=1 k ∈ G∆ for some L∆ Proposition 7.15, (153) holds for L∆ claim follows. k (u + h, z0 + y). As G∆ k is a Newton derivative of P ∆ k by k , and (154) holds for the remainder, the (cid:3) On the basis of Proposition 2.5, we now construct a Newton derivative of the play on X × R. We set U = X × R , Un = X ∆n × R . 34 M. BROKATE Let 0 < r1 < r2 < . . . be an increasing sequence of positive numbers with rn < r for all n. Let {In,k} be the partition intervals of ∆n. We define Vn = {(u, z0) : u ∈ X, z0 ∈ R, osc In,k u < rn for all k} . (184) Since rn < rn+1 < r, we have V n ⊂ Un ∩ Vn+1. Moreover, by (156) Vn = X × R = U , because ∆n → 0 as n → ∞. [n Thus, all assumptions of Proposition 2.5 are satisfied. We finally arrive at the main result. Theorem 7.20. Let 1 ≤ q < ∞. The mapping GPr : X × R ⇒ L(X × R, Lq(a, b)) defined by GPr (u, z0) = G∆n(u, z0) (185) is a Newton derivative of the play Pr : X × R → Lq(a, b) with the remainder estimate if (u, z0) ∈ V n \ V n−1 , sup LPr ∈G∆(u+h,z0+y) kPr[u + h; z0 + y] − Pr[u; z0] − LPr (h, y)kLq (a,b) (186) ≤ ρ(u,z0)(khk∞ + y)k(h, y)kX×R where ρ(u,z0)(δ) ↓ 0 as δ ↓ 0. The elements LPr of Gr(u, z0) satisfy the estimate kLPr (h, y)k∞,t ≤ khk∞,t + y for all h ∈ X and y ∈ R, uniformly in (u, z0). They have the form LPr (h, y)(t) = hµPr (u, z0, t), (h, y)i , t ∈ (a, b) , (187) (188) with µPr (u, z0, t) = µ∆n(u, z0, t) (189) The functions µPr : C[a, b] × R × [a, b] → (C[a, b] × R)∗ are measurable. ✷ if (u, z0) ∈ V n \ V n−1 , µ∆n ∈ S∆n . Proof. This follows from Proposition 2.5 and Proposition 7.19 when we choose ρu,z0 = max{ρn,u,z0, ρn+1,u,z0} if (u, z0) ∈ V n \ V n−1 with the remainder terms ρn,u,z0 belonging to G∆n. (cid:3) Since the stop operator is related to the play operator by the formula Sr[u; z0] = u − Pr[u; z0], it also has a Newton derivative. Corollary 7.21. The stop operator Sr : X × R → Lq(a, b) , 1 ≤ q < ∞ , has a Newton derivative given by GSr (u, z0) = π1 − GPr (u, z0) (190) (191) with elements (192) Here, GPr and LPr have the form and properties described in Theorem 7.20, and π1 denotes the projection π1(h, y) = h. ✷ LSr (h, y) = h − LPr (h, y) . NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 35 8. Bouligand derivative of the play and the stop We intend to prove that the play and the stop operator are Bouligand differentiable from X × R to Lq, 1 ≤ q < ∞. It suffices to show that Pr, Sr : X ∆ → Lq(a, b) are Bouligand differentiable for arbitrary partitions ∆, as the sets X ∆ ⊂ X × R are open and their union covers X × R. In the previous section we explained how, on X ∆, the play can be rep- resented as a finite composition of the positive part Fpp, the cumulated maximum F I and continuous linear mappings. By virtue of the chain rule, it therefore suffices to show that Fpp and F I are Bouligand differentiable, and that the function spaces involved in the composition are fitting. The positive part mapping β : R → R, β(x) = max{x, 0} has the direc- tional (in fact, Bouligand) derivative β′(x; y) =(0 , x < 0 or x = 0 , y ≤ 0 , y , x > 0 or x = 0 , y > 0 . (193) Lemma 8.1. Let I ⊂ [a, b] be a closed interval, 1 ≤ q < q ≤ ∞. The map- ping Fpp : Lq(I) → Lq(I), Fpp(u) = max{u, 0}, is Bouligand differentiable, and F ′ pp(u; h)(t) = β′(u(t); h(t)) . (194) (cid:3) Proof. See Examples 8.12 and 8.14 in [10]. It has already be proved in Proposition 4.1 that the cumulated maximum F I : X → Lq(I) is Bouligand differentiable for every q < ∞, and that (F I )′(u; h)(t) = max s∈M I (u,t) h(s) . (195) By the chain rule, the mapping P I + : Z I ∩ (X × R) → Lq(I), +(u, p) = p + Fpp(F I (u − p − r)) P I has the Bouligand derivative (P I +)′((u, p); (h, η))(t) = η + β′( max s∈I,s≤t (u(s) − r − p); max (h(s) − η)) . s∈M I (u,t) (196) An analogous formula holds for the Bouligand derivative of P I −. Applying the chain rule to (142), we obtain the Bouligand derivative of the play recursively as k )′((u, z0); (h, y)) = (P I (w∆ P ′ r[[u; z0]; [h; y]](t) = (P I k )′((u, wk−1(u, z0)); (h, (w∆ k )′((u, wk−1(u, z0)); (h, (w∆ k−1)′((u, z0); (h, y))))(tk ) , k−1)′((u, z0); (h, y))))(t) . (197) We also obtain the refined remainder estimate. Theorem 8.2. The Bouligand derivative of the play operator Pr given in (197) satisfies, for all (u, z0) ∈ X × R, the remainder estimate kPr[u + h; z0 + y] − Pr[u; z0] − P ′ r[[u; z0]; [h; y]]kLq (a,b) ≤ ρu,z0(khk∞ + y)k(h, y)kX×R (198) for all h ∈ X, y ∈ R. Here, ρ(u,z0)(δ) ↓ 0 as δ ↓ 0 and ρ(u,z0) is uniformly bounded in (u, z0). 36 M. BROKATE Proof. The proof is analogous to that for the Newton derivative, using Proposition 5.4 instead of Proposition 5.2. (cid:3) 9. The parametric play operator Instead of a single play operator acting on a function u = u(t), we now want to consider a family of play operators acting on a function u = u(x, t), where x plays the role of a parameter. This has been developed in [18] in order to solve boundary value problems for partial differential equations with hysteresis. Here, we are concerned with parametrizing the Newton derivative of the play. For a given measurable space Ω (that is, a set Ω equipped with a sigma algebra), we want to define the parametric play operator P Ω r by P Ω r [u; z0](x, t) = Pr[u(x, ·); z0(x)](t) (199) for functions u : Ω × [a, b] → R, z0 : Ω → R. The parametric play operator thus represents a parametric family of play operators. For a given metric space X, equipped with the Borel sigma algebra, let M(Ω; X) denote the space of all measurable functions from Ω to X. Lemma 9.1. Formula (199) defines an operator P Ω r : M(Ω; C[a, b]) × M(Ω; R) → M(Ω; C[a, b]) . (200) Proof. The assignment x 7→ (u(x, ·), z0(x)) 7→ Pr[u(x, ·), z0(x)] defines a mapping Ω → C[a, b] × R → C[a, b] which is measurable since Pr is contin- uous. (cid:3) We define the parametric cumulated maximum (that is, the para- metric family of cumulated maxima) F Ω for functions u : Ω → C[a, b]) by (F Ωu)(x) = F (u(x, ·)) , x ∈ Ω . Lemma 9.2. We have F Ω : M(Ω; C[a, b]) → M(Ω; C[a, b]) , F Ω : Lp(Ω; C[a, b]) → Lp(Ω; C[a, b]) , 1 ≤ p ≤ ∞ . (201) (202) Proof. If u : Ω → C[a, b] is measurable, the composition x 7→ u(x, ·) 7→ F (u(x, ·)) defines a measurable mapping since F : C[a, b] → C[a, b] is con- tinuous. As k(F Ωu)(x)k∞ ≤ ku(x, ·)k∞ and because u(x, ·) = v(x, ·) a.e. in x implies that F Ωu = F Ωv a.e. in x, the second assertion in (202) fol- lows. (cid:3) The corresponding set-valued mappings M Ω and ΦΩ are given by M Ω(u, t, x) = M (u(x, ·), t) , ΦΩ(u, t, x) = Φ(u(x, ·), t) . (203) For any given function u : Ω → C[a, b], these formulas define set-valued mappings (x, t) 7→ M (u(x, ·), t) = M Ω(u, t, x) , Ω × [a, b] ⇒ [a, b] , (x, t) 7→ Φ(u(x, ·), t) = ΦΩ(u, t, x) , Ω × [a, b] ⇒ C[a, b]∗ . (204) Lemma 9.3. Let u ∈ M(Ω; C[a, b]). Then the mappings defined in (204) are measurable. NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 37 Proof. The mappings arise as compositions (x, t) 7→ (u(x, ·), t) 7→ M (u(x, ·), t) , Ω × [a, b] → C[a, b] × [a, b] ⇒ [a, b] , (x, t) 7→ (u(x, ·), t) 7→ Φ(u(x, ·), t) , Ω × [a, b] → C[a, b] × [a, b] ⇒ C[a, b]∗ . Due to Propositions 4.2 and 4.5, the assertion follows. (cid:3) In Proposition 4.8, a Newton derivative G of the cumulated maximum F has been constructed from measurable selectors µ of Φ. Any such µ ∈ SΦ defines an element of G(u(x, ·)). More precisely, given u ∈ M(Ω; C[a, b]) and x ∈ Ω we set [(LΩ(x))v](t) = hµ(u(x, ·), t), vi , v ∈ C[a, b] . (205) Proposition 9.4. Let µ be a measurable selector of Φ, let u ∈ M(Ω; C[a, b]). Then (205) defines a mapping LΩ : Ω → L(C[a, b]; L∞(a, b)) with the property LΩ(x) ∈ G(u(x, ·)) for all x ∈ Ω. Let moreover h ∈ M(Ω; C[a, b]). Then (206) (207) (x, t) 7→ [(LΩ(x))h(x, ·)](t) = hµ(u(x, ·), t), h(x, ·)i (208) defines a measurable function from Ω × [a, b] to R. Proof. Proposition 4.8 yields (206) and (207). The composition (x, t) 7→ (u(x, ·), t) 7→ µ(u(x, ·), t) defines a measurable mapping from Ω × [a, b] to C[a, b]∗, since µ : C[a, b] × [a, b] → C[a, b]∗ is measurable. As the mapping x 7→ h(x, ·) is measurable and the mapping (ν, v) 7→ hν, vi is continuous, (208) defines a measurable function. (cid:3) We define GΩ : M(Ω; C[a, b]) ⇒ Map(Ω; L(C[a, b]; L∞(a, b))) GΩ(u) = {LΩ : LΩ satisfies (205) and (206) for some µ ∈ SΦ} , (209) a parametric family of Newton derivatives of the parametric family of cumu- lated maxima F Ω. It is not a Newton derivative of F Ω. (Here, Map(A; B) stands for the set of all mappings from a set A to a set B.) For the parametric play P Ω r we proceed in the same manner. According to Theorem 7.20, the Newton derivative GPr of Pr constructed there has, when evaluated at (u, z0), elements of the form LPr (h, y)(t) = hµPr (u, z0, t), (h, y)i , t ∈ (a, b) , (210) for some µPr as given in (189). We define r : Ω → L(C[a, b] × R; L∞(a, b)) LΩ [(LΩ r (x))(v, y0)](t) = hµPr (u(x, ·), z0(x), t), (v, y0)i (211) for (v, y0) ∈ C[a, b] × R. 38 M. BROKATE Proposition 9.5. Let µPr be as given in (189), let u ∈ M(Ω; C[a, b]) and z0 ∈ M(Ω; R). Then LΩ r as given in (211) satisfies LΩ r (x) ∈ GPr (u(x, ·), z0(x)) for all x ∈ Ω. (212) Let moreover h ∈ M(Ω; C[a, b]), y ∈ M(Ω; R). Then (x, t) 7→ [(LΩ r (x))(h(x, ·), y(x)](t) = hµPr (u(x, ·), z0(x), t), (h(x, ·), y(x))i defines a measurable function from Ω × [a, b] to R. Proof. Analogous to that of Proposition 9.4. (213) (cid:3) We may define GΩ r and view GΩ r (u, z0) as the set of all such mappings LΩ r as a parametric Newton derivative of the parametric play P Ω r . 10. Appendix: set-valued mappings In this section, we recall some standard results from set-valued analysis, given e.g. in [15], and derive some consequences needed in this paper. Let Ψ : X ⇒ Y . We generally assume that Ψ(u) 6= ∅ for every u ∈ X. Definition 10.1. Let X, Y be Hausdorff topological spaces, let Ψ : X ⇒ Y . We say that Ψ is upper semicontinuous (or usc for short), if Ψ−1(A) := {u : u ∈ X, Ψ(u) ∩ A 6= ∅} (214) is closed for every closed subset A of Y . We say that Ψ is measurable if Ψ−1(V ) is measurable for all open V ⊂ Y . A mapping ψ : X → Y is called a measurable selector of Ψ if ψ is measurable and ψ(u) ∈ Ψ(u) for every u ∈ X. Lemma 10.2. Let X, Y be Hausdorff topological spaces. A mapping Ψ : X ⇒ Y is usc if and only if for every u ∈ X and every open set V with Ψ(u) ⊂ V ⊂ Y there exists an open set U ⊂ X with u ∈ U and Ψ(U ) ⊂ V . Proof. See Proposition 6.1.3 in [15]. (cid:3) Obviously, a single-valued mapping is continuous in the usual sense if and only if it is usc in the sense above. The following two lemmas are immediate consequences of Lemma 10.2. Lemma 10.3. Let X, Y be Hausdorff topological spaces, X0 ⊂ X. Let Ψ : X ⇒ Y be usc. Then ΨX0 : X0 ⇒ Y is usc. ✷ Lemma 10.4. Let X, Y be Hausdorff topological spaces, X = X1 ∪ X2 with X1, X2 open. Let Ψj : Xj ⇒ Y be usc for j = 1, 2 such that Ψ1(X1 ∩ X2) = Ψ2(X1 ∩ X2). Then Ψ : X ⇒ Y defined by Ψ(u) = Ψj(u) if u ∈ Xj is usc. ✷ The composition Ψ2 ◦ Ψ1 of two set-valued mappings Ψ1 : X ⇒ Y and Ψ2 : Y ⇒ Z is defined as (Ψ2 ◦ Ψ1)(u) = [v∈Ψ1(u) Ψ2(v) . (215) Lemma 10.5. Let X, Y, Z be Hausdorff topological spaces, let Ψ1 : X ⇒ Y and Ψ2 : Y ⇒ Z be usc. Then Ψ2 ◦ Ψ1 is usc. NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 39 Proof. This is again straightforward, using Lemma 10.2. See Proposition 2.56 in [9]. (cid:3) We will use Lemma 10.5 mainly for the special cases Ψ◦f and f ◦Ψ where Ψ is usc and f is single-valued and continuous. Lemma 10.6. Let X, Y, Z be normed spaces, U ⊂ X and V ⊂ Y open. Let f : U → V , g1 : U → L(X, Y ) and g2 : V → L(Y, Z) be measurable. Then g : U → L(X, Z) defined by g(u) = g2(f (u)) ◦ g1(u) is measurable. Proof. The composition is a continuous mapping from L(X, Y ) × L(Y, Z) to L(X, Z). (cid:3) Proposition 10.7. Let X, Y be Hausdorff topological spaces, let Ψ : X ⇒ Y . We assume that Ψ has compact values, that is, Ψ(u) is compact for all u ∈ X. (i) If Ψ is usc, then the graph of Ψ, Gr Ψ = {(u, v) : u ∈ X, v ∈ Ψ(u)} (216) is closed in X × Y . (ii) If the graph of Ψ is closed in X × Y and if Ψ(X) is relatively compact in Y , then Ψ is usc. Proof. See Proposition 6.1.8, Remark 6.1.9 and Proposition 6.1.10 in [15]. (cid:3) Above we consider compositions of the form F (u, p, t) = F2(u, p, F1(u, p, t), t) (217) for mappings F1 : U × R × I → R and F2 : U × R × R × I → R, where U ⊂ C(J). Here we are concerned with the upper semicontinuity of a corresponding composition of mappings Ψ1, Ψ2 arising in the construction of Newton derivatives. Proposition 10.8. Let I, J ⊂ R be compact intervals, U ⊂ C(J) open. Let F1 : U × R × I → R be continuous. Let Ψ1 : U × R × I ⇒ (C(J) × R)∗ and Ψ2 : U × R × R × I ⇒ (C(J) × R × R)∗ be usc, with w∗-compact values, and locally bounded. Let Ψ : U × R × I ⇒ (C(J) × R)∗ be defined by Ψ(u, p, t) = {L2 ◦ (id, L1) : L1 ∈ Ψ1(u, p, t), L2 ∈ Ψ2(u, p, F1(u, p, t), t)} , (218) where id denotes the identity on C(J) × R. Then Ψ is usc, has w∗-compact values and is locally bounded. Proof. As Ψ1 and Ψ2 are locally bounded, we see from (218) and the conti- nuity of F1 that Ψ is locally bounded. Next, let (Ln 2 ) be arbitrary sequences in (C(J) × R)∗ and (C(J) × R × R)∗ respectively. We claim that 1 ) and (Ln Ln 1 ∗ ⇀ L1 , Ln 2 ∗ ⇀ L2 ⇒ Ln 2 ◦ (id, Ln 1 ) ∗ ⇀ L2 ◦ (id, L1) . (219) Indeed, for any (h, q) ∈ C(J) × R we have hLn 2 , (h, q, Ln 1 (h, q))i → hL2 , (h, q, L1(h, q))i . To prove that Ψ has w∗-compact values, let Ln = Ln 1 ) be a se- quence in Ψ(u, p, t). By assumption, passing to suitable subsequences we 2 ◦ (id, Ln 40 M. BROKATE ∗⇀ L1 ∈ Ψ1(u, p, t) and Ln 2 ∗⇀ L2 ∈ Ψ2(u, p, F1(u, p, t), t). By have Ln 1 (219), Ln ∗⇀ L2 ◦ (id, L1) ∈ Ψ(u, p, t). It remains to prove that Ψ is usc. Let A ⊂ (C(J) × R)∗ be w∗-closed; it suffices to show that Ψ−1(A) is closed. Let (un, pn, tn) ∈ Ψ−1(A) and (un, pn, tn) → (u, p, t). Let Ln ∈ Ψ(un, pn, tn) ∩ A. We have Ln = Ln 1 ∈ Ψ1(un, pn, tn) and Ln 2 ∈ Ψ2(un, pn, F1(un, pn, tn), tn). Since Ψ1 and Ψ2 are locally bounded, ∗⇀ L2. As the graphs passing to a subsequence we get Ln 1 of Ψ1 and Ψ2 are w∗-closed by Proposition 10.7, L1 ∈ Ψ1(u, p, t) and L2 ∈ Ψ2(u, p, F1(u, p, t), t). By (219), Ln ∗ ⇀ L2 ◦ (id, L1) =: L ∈ Ψ(u, p, t). As A is w∗-closed, it follows that L ∈ A. Thus, Ψ−1(A) is closed. (cid:3) ∗⇀ L1, Ln 2 2 ◦ (id, Ln 1 ) for some Ln Acknowledgements. The author thanks Michael Ulbrich in particular for pointing out the line of argument used in the proof of Propositions 4.1 and 4.8, and him as well as Constantin Christof, Michael Hintermuller, Pavel Krejc´ı, Karl Kunisch and Gerd Wachsmuth for valuable discussions. References [1] H. Bauschke, P. Combettes: Convex analysis and monotone operator theory in Hilbert spaces, Springer 2011. [2] J.F. Bonnans, A. Shapiro, Perturbation Analysis of Optimization Problems, Springer, New York, 2000. [3] M. Brokate, P. Krejci, Weak Differentiability of Scalar Hysteresis Operators, Discrete Contin. Dyn. Syst. 35 (2015), 2405-2421. [4] M. Brokate, J. Sprekels, Hysteresis and Phase Transitions, Springer, New York, 1996. [5] C. Christof, Sensitivity analysis and optimal control of obstacle-type evolution varia- tional inequalities, SIAM J. Control Opt. 57 (2019), 192 -- 218. [6] I.V. Girsanov, Lectures on Mathematical Theory of Extremum Problems, Springer, Berlin, 1972. [7] M. Hintermuller, K. Ito, K. Kunisch, The primal-dual active set strategy as a semis- mooth Newton method, SIAM J. Opt. 13 (2003), 865-888. [8] M. Hintermuller, K. Kunisch, PDE-constrained optimization subject to pointwise con- straints on the control, the state, and its derivatives, SIAM J. Opt. 20 (2009), 1133-1156. [9] S. Hu, N.S. Papageorgiu, Handbook of multivalued analysis, volume I: Theory, Kluwer 1997. [10] K. Ito, K. Kunisch, Lagrange Multiplier Approach to Variational Problems and Ap- plications, SIAM Series Advances in Design and Control, SIAM, Philadelphia, 2008. [11] M. A. Krasnosel'skiı, B. M. Darinskiı, I. V. Emelin, P. P. Zabrejko, E. A. Lifshits and A. V. Pokrovskiı, An operator-hysterant, Dokl. Akad. Nauk SSSR 190 (1970), 34-37; Soviet Math. Dokl. 11 (1970), 29-33. [12] P. Krejc´ı, Hysteresis, Convexity and Dissipation in Hyperbolic Equations, Gakk¯otosho, Tokyo, 1996. [13] F. Mignot, Controle dans les in´equations variationelles elliptiques, J. Funct. Anal. 22 (1976), 130 -- 185. [14] A. Mielke, T. Roub´ıcek, Rate-Independent Systems, Springer 2015. [15] N.S. Papageorgiu, S.T. Kyritsi-Yiallourou, Handbook of applied analysis, Springer 2009. [16] M. Ulbrich, Semismooth Newton methods for operator equations in function spaces, SIAM J. Optim. 13 (2003), 805-841. [17] M. Ulbrich, Semismooth Newton Methods for Variational Inequalities and Constrained Optimization Problems in Function Spaces, SIAM, Philadelphia, 2011. [18] A. Visintin, Differential Models of Hysteresis, Springer, Berlin, 1994. NEWTON AND BOULIGAND DERIVATIVES OF THE PLAY AND THE STOP 41 Martin Brokate Dept. of Mathematics, Technical University of Munich, Boltzmannstr. 3, D-85747 Garching, Germany E-mail address: [email protected]
1208.5530
1
1208
2012-08-27T23:41:38
Generalized resolvents of symmetric and isometric operators: the Shtraus approach
[ "math.FA" ]
This exposition paper is devoted to the theory of Abram Vilgelmovich Shtraus and his disciples and followers. This theory studies the so-called generalized resolvents of symmetric and isometric operators in a Hilbert space and provides alternative formulas to the well-known Krein-type formulas. We think that it would be convenient to have an exposition of the corresponding theory with full proofs and necessary corrections.
math.FA
math
Generalized resolvents of symmetric and isometric operators: the Shtraus approach S.M. Zagorodnyuk Contents 1 Introduction. 2 Generalized resolvents of isometric and densely-defined sym- metric operators. 2.1 Properties of a generalized resolvent of an isometric operator. . 2.2 Chumakin's formula for the generalized resolvents of an iso- 1 2 7 7 2.3 metric operator. . . . . . . . . . . . . . . . . . . . . . . . . . . 20 Inin's formula for the generalized resolvents of an isometric operator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 2.4 The generalized resolvents of a symmetric operator. A con- nection with the generalized resolvents of the Cayley transfor- mation of the operator. . . . . . . . . . . . . . . . . . . . . . . 29 2.5 Shtraus's formula for the generalized resolvents of a symmetric densely defined operator. . . . . . . . . . . . . . . . . . . . . . 32 3 Generalized resolvents of non-densely defined symmetric op- erators. 36 3.1 The forbidden operator. . . . . . . . . . . . . . . . . . . . . . 36 3.2 Admissible operators. . . . . . . . . . . . . . . . . . . . . . . . 39 3.3 Properties of dissipative and accumulative operators. . . . . . 45 . . . . . . . . . . . . . . . . 49 3.4 Generalized Neumann's formulas. 3.5 Extensions of a symmetric operator with an exit out of the space. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 3.6 The operator-valued function Bλ generated by an extension of a symmetric operator with an exit out of the space. . . . . 59 3.7 The characteristic function of a symmetric operator. . . . . . . 63 3.8 A connection of the operator-valued function F(λ) with the characteristic function. . . . . . . . . . . . . . . . . . . . . . . 68 3.9 Boundary properties of the operator-valued function F(λ): the case of a densely defined symmetric operator A. . . . . . . . . 71 3.10 A connection of the operator Φ∞ and the operator-valued func- tion F(λ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 3.11 Shtraus's formula for the generalized resolvents of symmetric operator. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90 4 Generalized resolvents of isometric and symmetric operators with gaps in their spectrum. 4.1 Spectral functions of an isometric operator having a constant 2 95 value on an arc of the circle. . . . . . . . . . . . . . . . . . . . 95 4.2 Some decompositions of a Hilbert space in a direct sum of subspaces. Isometric operators with lacunas in a spectrum. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 . . . . . . . . 117 4.3 4.4 Spectral functions of a symmetric operator having a constant value on an open real interval. . . . . . . . . . . . . . . . . . . 125 5 Formal credits. 134 1 Introduction. This exposition paper is devoted to the theory of Abram Vilgelmovich Shtraus and his disciples and followers. This theory studies the so-called generalized resolvents of symmetric and isometric operators in a Hilbert space. The object, which now is called a generalized resolvent of a symmetric operator, appeared for the first time in papers of Neumark and Krein in 40 -- th of the previous century. At that time it was already known, due to the result of Carleman, that each densely defined symmetric operator has a spectral function (the term "spectral function" is due to Neumark). Neumark showed that each spectral function of a (densely defined) symmetric operator is generated by the orthogonal spectral function of a self-adjoint extension of the given symmetric operator in a possibly larger Hilbert space (Neumark's dilation theorem) [31]. In particular, this fact allowed him to describe all solutions of the Hamburger moment problem in terms of spectral functions of the operator defined by the Jacobi matrix [31, pp. 303-305]. However, as was stated by Neumark himself in his paper [32, p. 285], despite of its theoretical generality this result did not give practical tools for finding spectral functions in various concrete cases. For operators with the deficiency index (1, 1) Neumark proposed a de- scription of the generalized resolvents which was convenient for practical ap- plications. As one such an application Nevanlinna's formula for all solutions of the Hamburger moment problem was derived [32, pp. 292-294]. In 1944, independently from Neumark, another description of the generalized resol- vents for operators with the deficiency index (1, 1) was given by Krein [17]. Krein also noticed a possibility of an application of his results to a study of the moment problem and to the problem of the interpolation of bounded analytic functions. In 1946 appeared the famous paper of Krein which gave 3 an analytic description of the generalized resolvents of a symmetric operator with arbitrary equal finite defect numbers. Also there appeared a description of all Π-resolvents in the case that the operator is non-negative (a Π-resolvent -- a generalized resolvent generated by a non-negative self-adjoint extension of the given operator). This paper of Krein became a starting point for a series of papers devoted to constructions of Krein-type formulas for various classes of operators and relations in a Hilbert space, in the Pontriagin space, in the Krein space. Among other mathematicians we may mention the fol- lowing authors here (in the alphabetical order): Behrndt, Derkach, de Snoo, Hassi, Krein, Kreusler, Langer, Malamud, Mogilevskii, Ovcharenko, Sorjo- nen, see, e.g.: [19], [20], [23], [21], [10], [27], [11], [8], [28], [30], [3], [9], and references therein. Starting from the paper of Krein [17] the theory of the resolvent matrix has been developed intensively, see., e.g., papers of Krein and Saakyan [18], Krein and Ovcharenko[22], Langer and Textorius [24], Derkach [7] and papers cited therein. Krein-type formulas are close to the linear fractional transfor- mation and this fact suggested to apply them to interpolation problems, where such form appears by the description of solutions. On the other hand, in 1954 , in his remarkable work [37], Shtraus ob- tained another analytic description of the generalized resolvents of a (densely defined) symmetric operator with arbitrary defect numbers. Characteristic features of Shtraus's formula, and of Shtraus-type formulas in general, are the following: 1) An analytic function-parameter is bounded; 2) The minimal number of all parameters in the formula: an analytic function-parameter and the given operator. These features allow to solve various matrix interpolation problems, in the non-degenerate and degenerate cases. Moreover, these features give possibil- ity to solve not only one-dimensional, but also two-dimensional interpolation problems and to obtain analytic descriptions of these problems. The Shtraus formula set the beginning of a series of papers devoted to the derivation of such type formulas for another classes of operators and their applications. The content of papers devoted to isometric and symmetric op- erators in a Hilbert space is used in preparation of this survey. Formal refer- ences and list of papers will be given below. Besides that, we should mention papers for the generalized resolvents of operators related to the conjugation, see., e.g., papers of Kalinina [16], Makarova [25], [26], and references therein. We should also mention papers on the generalized resolvents in spaces with 4 an indefenite metric, see, e.g., papers of Gluhov, Nikonov [13], [34], [33], Etkin [12], and papers cited therein. Notice, that Shtraus-type formulas can be used to obtain Krein-type for- mulas, see e.g. [2]. Thus, in the theory of generalized resolvents one can see two directions: the derivation of Krein-type formulas and the derivation of Shtraus-type formulas. The Krein formulas are partially described in books of Akhiezer and Glazman [1], M.L. Gorbachuk and V.I. Gorbachuk [14]. The Shtraus- type formulas can be found only in papers. We think that it would be convenient to have an exposition of the corresponding theory with full proofs and necessary corrections. We omit references in the following text using only known facts from classical books. Formal references to all appeared results will be given after- wards, in the section "Formal credits". We hope that this exposition paper will be useful to all those who want to study this remarkable theory. Notations. R C N Z Z+ C+ C− D T De Te Re Rd k ∈ 0, ρ N ×N CK×N C≥ IN C T C ∗ Πλ Πε λ B(M) P card(M) S(D; N, N ′) 5 z < 1} z = 1} z > 1} z 6= 1} the set of all real numbers the set of all complex numbers the set of all positive integers the set of all integers the set of all non-negative integers {z ∈ C : Im z > 0} {z ∈ C : Im z < 0} {z ∈ C : {z ∈ C : {z ∈ C : {z ∈ C : {z ∈ C : Im z 6= 0} the real Euclidean d-dimentional space, d ∈ N means that k ∈ Z+, k ≤ ρ, if ρ < ∞; or k ∈ Z+, if ρ = ∞ the set of all complex matrices of size (K×N), K, N ∈ N the set of all non-negative matrices from CN ×N , N ∈ N the identity matrix of size (N × N), N ∈ N the transpose of the matrix C ∈ (K × N), K, N ∈ N the adjoint of the matrix C ∈ (K × N), K, N ∈ N the half-plane C+ or C−, which contains a point λ ∈ Re 2 , λ ∈ Re {z ∈ Πλ : ε < arg z < π − ε}, the set of all Borel subsets of a set M, which belongs C or Rd, d ∈ N the set of all scalar algebraic polynomials with complex coefficients the quantity of elements of a set M with a finite number of elements a class of all analytic in a domain D ⊆ C operator-valued functions F (z), which values are linear non-expanding operators mapping the whole N into N ′, where N and N ′ are some Hilbert spaces. 0 < ε < π All Hilbert spaces will be assumed to be separable, and operators in them are supposed to be linear. If H is a Hilbert space then 6 (·, ·)H k · kH M Lin M span M EH OH oH PH1 = P H H1 the scalar product in H the norm in H the closure of a set M ⊆ H in the norm of H the linear span of a set M ⊆ H the closed linear span of a set M ⊆ H the identity operator in H, i.e. EHx = x, ∀x ∈ H the null operator in H, i.e. 0Hx = 0, ∀x ∈ H the operator in H with D(oH) = {0}, oH0 = 0 the operator of the orthogonal projection on a subspace H1 in H Indices may be omitted in obvious cases. If A is a linear operator in H, the domain of A the range of A {x ∈ H : Ax = 0} (the kernel of A) the adjoint, if it exists the inverse, if it exists the closure, if A admits the closure the restriction of A to the set M ⊆ H a set of all points of the regular type of A the norm of A, if it is bounded the limit in the sense of the weak operator topology the limit in the sense of the strong operator topology the limit in the sense of the uniform operator topology D(A) R(A) Ker A A∗ A−1 A AM ρr(A) kAk w. − lim s. − lim u. − lim If A is a closed isometric operator then Mζ = Mζ(A) Nζ = Nζ(A) M∞ = M∞(A) N∞ = N∞(A) Rζ = Rζ(V ) If A is a closed symmetric operator (not necessarily densely defined) then Mz = Mz(A) Nz = Nz(A) Rz = Rz(A) then (EH − ζA)D(A), where ζ ∈ C H ⊖ Mζ, where ζ ∈ C R(A) H ⊖ R(A) (EH − ζV )−1, ζ ∈ C\T (A − zEH )D(A), where z ∈ C H ⊖ Mz, where z ∈ C (A − zEH )−1, z ∈ C\R 7 2 Generalized resolvents of isometric and densely- defined symmetric operators. 2.1 Properties of a generalized resolvent of an isomet- ric operator. Consider an arbitrary closed isometric operator V in a Hilbert space H. As it is well known, for V there always exists an unitary extension U, which acts in a Hilbert space eH ⊇ H. Define an operator-valued function Rζ in the following way: Rζ = Rζ(V ) = Ru;ζ(V ) = P ζ ∈ Te. eH H (cid:0)E eH − ζU(cid:1)−1 H, The function Rζ is said to be the generalized resolvent of an isometric operator V (corresponding to the extension U). Let {Ft}t∈[0,2π] be a left-continuous orthogonal resolution of the unity of the operator U. The operator-valued function Ft = P eH H Ft, t ∈ [0, 2π], is said to be a (left-continuous) spectral function of the isometric operator V (corresponding to the extension U). Let F (δ), δ ∈ B(T), be the orthogonal spectral measure of the unitary operator U. The following function F(δ) = P eH H F (δ), δ ∈ B(T), is said to be a spectral measure of the isometric operator V (corresponding to the extension U). Of course, the spectral function and the spectral measure are related by the following equality: F(δt) = Ft, δt = {z = eiϕ : 0 ≤ ϕ < t}, t ∈ [0, 2π], what follows from the analogous property of orthogonal spectral measures. Moreover, generalized resolvents and spectral functions (measures) are con- nected by the following relation: (Rzh, g)H =ZT 1 1 − zζ d(F(·)h, g)H =Z 2π 0 1 1 − zeit d(Fth, g)H, ∀h, g ∈ H, (1) which follows directly from their definitions. This relation allows to talk about the one-to-one correspondence between generalized resolvents and spec- tral measures, in the accordance with the well-known inversion formula for such integrals. 8 The generalized resolvents, as it is not surprizing from their definition, have much of the properties of the Fredholm resolvent (E eH −ζU)−1 of the uni- tary operator U. The aim of this subsection is to investigate these properties and to find out whether they are characteristic. Theorem 2.1 Let V be a closed isometric operator in a Hilbert space H, and Rζ be an arbitrary generalized resolvent of V . The following relations hold: 1) (zRz − ζRζ)f = (z − ζ)RzRζf , for arbitrary z, ζ ∈ Te, f ∈ Mζ(V ); 2) R0 = EH ; 3) For an arbitrary h ∈ H the following inequalities hold: Re(Rζh, h)H ≥ Re(Rζh, h)H ≤ 1 2 khk2 H, 1 2 khk2 H, ζ ∈ D; ζ ∈ De; 4) Rζ is an analytic operator-valued function of a parameter ζ in Te; 5) For an arbitrary ζ ∈ Te\{0} it holds: R∗ ζ ζ = EH − R 1 . Proof. corresponds to Rζ(V ). For the Fredholm resolvent Rζ = Rζ(U) = (E eH − ζU)−1 of the unitary operator U the following relation holds Let U be a unitary operator in a Hilbert space eH ⊇ H, which zRz − ζRζ = (z − ζ)RzRζ, z, ζ ∈ Te, which follows easily from the spectral decomposition of the unitary operator. By applying the operator P eH H to the latter relation, we get zRz − ζRζ = (z − ζ)RzRζ, z, ζ ∈ Te. Taking into account that for an arbitrary element f ∈ Mζ, f = (EH −ζV )gV , gV ∈ D(V ), holds Rζf = (E eH − ζU)−1(EH − ζV )gV = gV = P eH H gV = P eH H (E eH − ζU)−1(E eH − ζV )gV = P eH H (E eH − ζU)−1f = Rζf, 9 we conclude that the first property of Rz is true. The second property follows directly from the definition of a generalized resolvent. Denote by {Ft}t∈[0,2π] the left-continuous orthogonal resolution of the unity of the operator U. For an arbitrary h ∈ H holds Re(Rζh, h)H = Re(Rζh, h) eH = ReZ 2π 0 1 1 − ζeit d(Fth, h) eH 2 − ζe−it − ζeit 21 − ζeit2 d(Fth, h) eH , ζ ∈ Te. =Z 2π 0 =Z 2π 0 On the other hand, we may write (1 − ζ2)(R∗ ζh, R∗ ζh) eH + (h, h)H =Z 2π 0 1 − ζ2 1 − ζeit2 d(Fth, h) eH + (h, h)H 2 − ζe−it − ζeit 1 − ζeit2 d(Fth, h) eH , ζ ∈ Te. Consequently, we have Re(Rζh, h)H = ζhk2 eH + khk2 ζ ∈ Te, (2) 1 2(cid:0)(1 − ζ2)kR∗ H(cid:1) , and the third property of the generalized resolvent of an isometric operator follows. The fourth property follows directly from the same property of the resolvent of the unitary operator U. Using one more time the spectral resolution of the unitary operator U, we easily check that R∗ ζ + R 1 ζ = E eH, ζ ∈ Te\{0}. For arbitrary elements f, g from H we may write (Rζf, g)H = (Rζf, g) eH = (f, R∗ ζg) eH =(cid:16)f,(cid:16)E eH − R 1 ζ(cid:17) g(cid:17) eH =(cid:16)f,(cid:16)EH − R 1 ζ(cid:17) g(cid:17)H , ζ ∈ Te\{0}. From the latter relation it follows the fifth property of the generalized resol- vent. ✷ Theorem 2.2 Let an operator-valued function Rζ in a Hilbert space H be given, which depends on a complex parameter ζ ∈ Te and which values are linear operators defined on the whole H. This function is a generalized resol- vent of a closed isometric operator in H if an only if the following conditions are satisfied: 10 1) There exists a number ζ0 ∈ D\{0} and a subspace L ⊆ H such that (ζRζ − ζ0Rζ0)f = (ζ − ζ0)RζRζ0f, for arbitrary ζ ∈ Te and f ∈ L; 2) The operator R0 is bounded and R0h = h, for all h ∈ H ⊖ Rζ0L; 3) For an arbitrary h ∈ H the following inequality holds: Re(Rζh, h)H ≥ 1 2 khk2 H, ζ ∈ D; 4) For an arbitrary h ∈ H Rζh is an analytic vector-valued function of a parameter ζ in D; 5) For an arbitrary ζ ∈ D\{0} holds: R∗ ζ ζ = EH − R 1 . Proof. The necessity of the properties 1)-5) follows immediately from the previous theorem. Suppose that properties 1)-5) are true. At first we check that prop- In fact, for an arbitrary element h ∈ H, erties 1),2) imply R0 = EH. h = h1 + h2, h1 ∈ Rζ0L, h2 ∈ H ⊖ Rζ0L, using 2) we may write R0h = R0h1 + R0h2 = R0h1 + h2. There exists a sequence h1,n (n ∈ N) of elements of Rζ0L, tending to h1 as n → ∞. Since h1,n = Rζ0f1,n, f1,n ∈ L (n ∈ N), by condition 1) with ζ = 0 and f = f1,n we get −ζ0Rζ0f1,n = −ζ0R0Rζ0f1,n, n ∈ N. Dividing by −ζ0 and passing to the limit as n → ∞ we obtain that R0h1 = h1. Here we used the fact that R0 is continuous. Therefore we get R0h = h. By property 4) the following function is analytic and F (ζ) := (Rζh, h)H − Im F (0) = Im 1 2 1 2 khk2 H, ζ ∈ D, khk2 H = 0. 11 From property 3) it follows that Re F (ζ) = Re(Rζh, h)H − 1 2 khk2 H ≥ 0, ζ ∈ D. This means that the function F (ζ) belongs to the Carath´eodory class of all analytic in D functions satisfying the condition Re F (ζ) ≥ 0. Functions of this class admit the Riesz-Herglotz integral representation. Using this representation we get (Rζh, h)H = F (ζ) + 1 2 khk2 H =Z 2π 0 1 1 − ζeit dσ(t; h, h), ζ ∈ D, where σ(t; h, h) is a left-continuous non-decreasing function on the interval [0, 2π], σ(0; h, h) = 0. The function σ(t; h, h) with such normalization is defined uniquely by the Riesz-Herglotz integral representation. Since Z 2π 0 1 1 − ζeit dσ(t; h, h) = (Rζh, h) = (h, R∗ ζ h) = (h, h)−(h, R 1 h), ζ ∈ D\{0}, ζ where we used property 5), we get (R 1 ζ Therefore 0 h, h) = (h, h) −Z 2π (Rζh, h)H =Z 2π 0 1 1 − ζe−it dσ(t; h, h) =Z 2π 0 1 1 − 1 ζ eit dσ(t; h, h). 1 1 − ζeit dσ(t; h, h), ζ ∈ Te, h ∈ H. For arbitrary h, g from H we set σ(t; h, g) := 1 4 σ(t; h + g, h + g) − 1 4 σ(t; h − g, h − g) + i 4 σ(t; h + ig, h + ig) − i 4 σ(t; h − g, h − g). Then Z 2π 0 1 1 − ζeit dσ(t; h, g) = (Rζh, g)H, ζ ∈ Te, h, g ∈ H. (3) The function σ(t; h, g) is a left-continuous complex-valued function of bounded variation with the normalization σ(0; h, g) = 0. If for the function (Rζh, g)H there would exist a representation of a type (3) with another left-continuous 12 1 complex-valued function of bounded variationeσ(t; h, g) with the normaliza- tioneσ(0; h, g) = 0, then Z 2π where µ(t; h, g) :=eσ(t; h, g) − σ(t; h, g). The function µ is defined uniquely by its trigonometric moments ck = R 2π 0 eiktdµ(t; h, g), k ∈ Z. For ζ < 1, writing the expression under the integral sign in (4) as a sum of the geometric progression and integrating we get ck = 0, k ∈ Z+. For ζ > 1, the expression under the integral sign we can write as 1 − ζeit dµ(t; h, g) = 0, ζ ∈ Te, (4) 0 1 1 − ζeit = 1 − 1 1 − 1 ζ e−it = 1 − 1 ζ k e−ikt, ∞Xk=0 and integrating we obtain that c−k = 0, k ∈ N. Thus, the representation (3) for (Rζh, g)H defines the function σ(t; h, g) with the above-mentioned properties uniquely. For an arbitrary ζ ∈ Te and g, h ∈ H, taking into account property 5), we may write Z 2π 0 = (h, g) − (R 1 ζ dσ(t; g, h) 1 − ζeit = (Rζg, h) = (g, R∗ ζh) = (g, h − R 1 h) ζ dσ(t; h, g) −Z 2π 0 dσ(t; h, g) 1 − 1 ζ e−it 0 h, g) =Z 2π =Z 2π 0 dσ(t; h, g) 1 − ζeit . By the uniqueness of the representation (3) it follows that σ(t; g, h) = σ(t; h, g), h, g ∈ H, t ∈ [0, 2π]. From representation (3) and the linearity of the resolvent it follows that for arbitrary α1, α2 ∈ C and h1, h2, g ∈ H holds σ(t; α1h1 + α2h2, g) = α1σ(t; h1, g) + α2σ(t; h2, g), t ∈ [0, 2π]. Moreover, since the function σ(t; h, h) is non-decreasing, the following esti- mate is true: σ(t; h, h) ≤ σ(2π; h, h) =Z 2π 0 dσ(τ ; h, h) = (h, h)H, t ∈ [0, 2π], h ∈ H. (5) 13 This means that σ(t; h, g) for each fixed t from the interval [0, 2π] is a bounded bilinear functional in H, with the norm less or equal to 1. Consequently, it admits the following representation σ(t; h, g) = (Eth, g)H, t ∈ [0, 2π], where {Et}t∈[0,2π] is an operator-valued function of a parameter t, which values are linear non-expanding operators, defined on the whole H . Let us study the properties of this operator-valued function. Since (Eth, g) = σ(t; h, g) = σ(t; g, h) = (Etg, h) = (h, Etg), for arbitrary h, g ∈ H, the operators Et are all self-adjoint. The function (Eth, h) = σ(t; h, h) is non-decreasing of t on the segment [0, 2π], for an arbitrary h ∈ H. Since σ(0; h, g) = 0, then E0 = EH , and from (5) it is seen that E2π = EH . Taking into account that for arbitrary t ∈ (0, 2π] and h, g ∈ H holds lim s→t−0 (Esh, g)H = lim s→t−0 σ(s; h, g) = σ(t; h, g) = (Eth, g)H, we conclude that Et is left-continuous on the segment [0, 2π] in the weak operator topology sense. Also we have kEth − Eshk2 H = ((Et − Es)2h, h)H ≤ ((Et − Es)h, h)H → 0, as s → t − 0. Here we used the fact that for a self-adjoint operator (Et − Es) with the spectrum in [0, 1] holds the inequality from the latter relation. This fact follows directly from the spectral resolution of the self-adjoint operator (Et − Es). Therefore the function Et is left-continuous on the segment [0, 2π] in the strong operator topology sense, as well. Consequently, {Et}t∈[0,2π] is a generalized resolution of the identity. By the well-known Neumark's dilation theorem there exists an orthogonal reso- lution of unity {eEt}t∈[0,2π] in a Hilbert space eH ⊇ H, such that t ∈ [0, 2π], h ∈ H. Eth = P eH H eEth, From (3) and the definition of Et it follows that Rζ =Z 2π 0 1 1 − ζeit dEt, ζ ∈ Te, h, g ∈ H. (6) Here the convergence of the integral in the right-hand side is understood in the sense of the strong convergence of the integral sums. The existence of the 14 integral follows from the corresponding property of the orthogonal resolution of the identity eEt. For arbitrary ζ ∈ Te: ζ 6= ζ0; f ∈ L and g ∈ H, we may write: ((ζRζ − ζ0Rζ0)f, g)H = ζZ 2π 0 d(Etf, g)H 1 − ζeit − ζ0Z 2π 0 d(Etf, g)H 1 − ζ0eit 1 1 − ζeit 1 1 − ζ0eit d(Etf, g)H = (ζ − ζ0)Z 2π = (ζ − ζ0)Z 2π 0 0 0 1 1 1 − ζeit dZ t ((ζ − ζ0)RζRζ0f, g)H = (ζ − ζ0)Z 2π 1 − ζ0eis d(Esf, g)H =Z 2π 1 − ζeit dZ t 1 1 0 0 0 Z 2π 0 By the first condition of the theorem we conclude that 1 − ζ0eis d(Esf, g)H; 1 1 − ζeit d (EtRζ0f, g)H . 1 1 − ζeit d (EtRζ0f, g)H . By the continuity from the left of the function Et and the initial condition E0 = 0, the last relation inply that Z t 0 1 1 − ζ0eis d(Esf, g)H = (EtRζ0f, g)H , Therefore Z t 0 1 1 − ζ0eis dEsf = EtRζ0f, t ∈ [0, 2π], f ∈ L, g ∈ H. (7) t ∈ [0, 2π], f ∈ L. (8) From condition 3) of the theorem it follows that the equality Rζh = 0, for some h ∈ H and ζ ∈ D, implies the equality h = 0. Thus, the operator Rζ is invertible for all ζ ∈ D. Define the following operator V g = 1 ζ0(cid:1) g, ζ0(cid:0)EH − R−1 g ∈ Rζ0L, with the domain D(V ) = Rζ0L. For an arbitrary element g ∈ Rζ0L, g = Rζ0f , f ∈ L, and an arbitrary h ∈ H, we may write: (V g, h)H =(cid:18) 1 ζ0 (Rζ0 − EH )f, h(cid:19)H = 1 ζ0 ((Rζ0f, h)H − (f, h)H)H 0 1 1 ζ0(cid:18)Z 2π 1 − ζ0eit d(Etf, h)H =Z 2π 1 − ζ0eit d(Etf, h)H −Z 2π eitdZ t eit 0 0 0 = =Z 2π 0 d(Etf, h)H(cid:19) 1 1 − ζ0eis d(Esf, h)H eitd(Etg, h)H, where we have used (7) in the last equality. Therefore eitdEtg, g ∈ D(V ). 0 =Z 2π V g =Z 2π 0 15 (9) Using the latter equality we shall check that the operator V is isometric. In fact, for arbitrary g1, g2 ∈ D(V ) holds (V g1, V g2)H =Z 2π 0 eitd(Etg1, V g2)H =Z 2π 0 eitd(g1, EtV g2)H. (10) On the other hand, for g2 = Rζ0f2, f2 ∈ L, using the definition of the operator V and relation (8) we write EtV g2 = 1 ζ0 Et(Rζ0 − Et)f2 = 1 ζ0 (EtRζ0f2 − Etf2) dEsf2(cid:19) =Z t 0 0 eis 1 − ζ0eis dEsf2. = It follows that 0 1 1 1 − ζ0eis dEsf2 −Z t ζ0(cid:18)Z t (g1, EtV g2)H = (EtV g2, g1)H =Z t =Z t eisdZ s 0 0 0 1 1 − ζ0eir d(Erf2, g1)H =Z t =Z t e−is(Esg1, g2)H, 0 0 eis 1 − ζ0eis d(Esf2, g1)H eisd(Esg2, g1)H where we used (7) to obtain the last equality. Substituting the obtained relation in (10), we get (V g1, V g2)H = (g1, g2)H. Thus, the operator V is isometric. From the definition of the operator V it follows that (EH − ζ0V )g = R−1 ζ0 g, g ∈ D(V ) = Rζ0L, and therefore (EH − ζ0V )D(V ) = L. 16 Consequently, the bounded operator (EH −ζ0V )−1 is defined on the subspace L, and therefore it is closed. Hence, V is closwd, as well. Consider a unitary operator U in eH, which corresponds to the resolution {eEt}t∈[0,2π]: 0 U =Z 2π eitd(Etg, h)H =Z 2π 0 eitdeEt. eitd(eEtg, h) eH = (Ug, h) eH = (P Let us show that U ⊇ V . For arbitrary elements g ∈ D(V ) and h ∈ H holds eH H Ug, h)H, (V g, h)H =Z 2π 0 where we used relation (9). Therefore V g = P eH H Ug, g ∈ D(V ). (11) Moreover, we have kgkH = kV gkH = kP eH H UgkH ≤ kUgk eH = kgkH. Therefore kP eH H UgkH = kUgk eH, and P eH H Ug = Ug, g ∈ D(V ). Comparing the latter equality with (11) we conclude that U ⊇ V . From (3) it follows that (Rζh, g)H =Z 2π 0 1 1 − ζeit d(Eth, g)H =Z 2π 0 1 1 − ζeit d(eEth, g) eH = ((E eH − ζU)−1h, g) eH = (P eH H (E eH − ζU)−1h, g)H, ζ ∈ Te, h, g ∈ H. Consequently, Rζ, ζ ∈ Te, is a generalized resolvent of a closed isometric operator V . ✷ Let us give conditions for the given operator-valued function Rζ to be a generalized resolvent of a concrete prescribed closed isometric operator. Theorem 2.3 Let an operator-valued function Rζ in a Hilbert space H be given, which depends on a complex parameter ζ ∈ Te and which values are linear operators defined on the whole H. Let a closed isometric operator V in H be given. The function Rζ, ζ ∈ Te, is a generalized resolvent of the operator V if an only if the following conditions hold: 17 1) For all ζ ∈ Te and for all g ∈ D(V ) the following equality holds: Rζ(EH − ζV )g = g; 2) The operator R0 is bounded and R0h = h, for all h ∈ H ⊖ D(V ); 3) For an arbitrary h ∈ H the following inequality holds: Re(Rζh, h)H ≥ 1 2 khk2 H, ζ ∈ D; 4) For an arbitrary h ∈ H Rζh is an analytic vector-valued function of a parameter ζ in D; 5) For an arbitrary ζ ∈ D\{0} the following equality is true: R∗ ζ = EH − R 1 . ζ Proof. The necessity of properties 3)-5) follows from the previous theorem. Let Rζ be a generalized resolvent of V , corresponding to a unitary operator U ⊇ V in a Hilbert space eH ⊇ H. Then Rζ(EH − ζV )g = P eH H (cid:0)E eH − ζU(cid:1)−1 (EH − ζV )g g ∈ D(V ). = P eH H (cid:0)E eH − ζU(cid:1)−1 (EH − ζU)g = g, Therefore condition 1) holds, as well. The second property follows from Theorem 2.1, property 2). Suppose that conditions 1)-5) from the statement of the theorem are satisfied. Let us check that for the function Rζ conditions 1),2) from the previous theorem are true. Choose an arbitrary ζ0 ∈ D\{0}. Let L := (EH − ζ0V )D(V ). By the first condition of the theorem we may write Rζ0(EH − ζ0V )g = (EH − ζ0V )−1(EH − ζ0V )g, g ∈ D(V ), and therefore Rζ0f = (EH − ζ0V )−1f ∈ D(V ), f ∈ L; f = (EH − ζ0V )Rζ0f, f ∈ L. Rζf = Rζ(EH − ζ0V )Rζ0f, f ∈ L, ζ ∈ Te. (12) (13) Let us use condition 1) of the theorem with the vector g = Rζ0f ∈ D(V ): Rζ0f = Rζ(EH − ζV )Rζ0f, f ∈ L. 18 Assuming that ζ 6= 0, we divide relation (13) by ζ0 and subtract the last relation divided by ζ from it. Multiplying the obtained relation by ζζ0, we get the required equality from condition 1) of Theorem 2.1. In the case ζ = 0, the required equality takes the following form: Rζ0f = R0Rζ0f , f ∈ L, and it follows directly from the first condition of the theorem with ζ = 0, g = Rζ0f ∈ D(V ). By relation (12) and the definition of L it follows that Rζ0L = (EH − ζ0V )−1L = D(V ). (14) Consequently, the second condition of the theorem 2.1 follows from the second condition of this theorem. Observe that conditions 3)-5) of Theorem 2.1 coincide with the corresponding conditions of this theorem. Thus, applying Theorem 2.1 we conclude that the function Rζ, ζ ∈ Te, is a generalized resolvent of a closed isometric operator in H. Moreover, in the proof of Theorem 2.1 such an operator V1 was constructed explicitly. Recall that its domain was Rζ0L, and V1g = 1 ζ0 (EH − R−1 ζ0 )g, g ∈ Rζ0L. Let us check that V = V1. By (14) we see that the domains of the operators V and V1 coincide. For an arbitrary g ∈ D(V ) = D(V1), using condition 1) we write V1g = 1 ζ0 (EH − R−1 ζ0 )g = 1 ζ0 (EH − (EH − ζ0V ))g = V g. Thus, V1 = V , and therefore the function Rζ, ζ ∈ Te, is a generalized resolvent of the operator V . ✷ In conclusion of this subsection, we establish some propositions which will be used in what follows. Proposition 2.1 Let F (λ) be an operator-valued analytic function in a do- main D ⊆ C, which values are linear bounded operators in a Hilbert space H, defined on the whole H. Suppose that at each point of the domain D there exists the inverse F −1(λ), defined on the whole H. Assume that for each λ ∈ D there exists an open neighborhood of this point in which kF −1(λ)k is bounded, as a function of λ. Then the operator-valued function F −1(λ) is analytic in D, as well. Proof. Consider an arbitrary point µ ∈ D. We may write: F −1(λ) − F −1(µ) = −F −1(λ)(F (λ) − F (µ))F −1(µ), λ ∈ D, (15) 19 and therefore (cid:13)(cid:13)F −1(λ) − F −1(µ)(cid:13)(cid:13) ≤(cid:13)(cid:13)F −1(λ)(cid:13)(cid:13) kF (λ) − F (µ)k(cid:13)(cid:13)F −1(µ)(cid:13)(cid:13) → 0, as λ → µ. Thus, F −1(λ) is continuous in the domain D in the uniform operator topology. From equality (15) it follows that 1 λ − µ (cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13)−F −1(λ)(cid:18) 1 =(cid:13)(cid:13)(cid:13)(cid:13)−F −1(λ)(cid:18) 1 ≤(cid:20)(cid:13)(cid:13)F −1(λ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (F −1(λ) − F −1(µ)) + F −1(µ)F ′(µ)F −1(µ)(cid:13)(cid:13)(cid:13)(cid:13) (F (λ) − F (µ))(cid:19) + F −1(µ)F ′(µ)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F −1(µ)(cid:13)(cid:13) (F (λ) − F (µ)) − F ′(µ)(cid:19) + (F −1(µ) − F −1(λ))F ′(µ)(cid:13)(cid:13)(cid:13)(cid:13) ∗(cid:13)(cid:13)F −1(µ)(cid:13)(cid:13) (F (λ) − F (µ)) − F ′(µ)(cid:13)(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)F −1(µ) − F −1(λ)(cid:13)(cid:13) kF ′(µ)k(cid:21) ∗(cid:13)(cid:13)F −1(µ)(cid:13)(cid:13) → 0, as λ → µ. ✷ λ − µ λ − µ 1 λ − µ Proposition 2.2 Let W be a linear non-expanding operator in a Hilbert space H, and f be an arbitrary element from D(W ) such that: kW f kH = kf kH. Then (W f, W g)H = (f, g)H, ∀g ∈ D(W ). Proof. Since W is non-expanding, we may write kζf + gk2 H − kW (ζf + g)k2 H ≥ 0, g ∈ D(W ), ζ ∈ C. Then (ζf + g, ζf + g) − (ζW f + W g, ζW f + W g) = ζ2kf k2 + 2 Re(ζ(f, g)) +kgk2 − ζ2kW f k2 − 2 Re(ζ(W f, W g)) − kW gk2 = kgk2 − kW gk2 + 2 Re(ζ[(f, g) − (W f, W g)]). 20 Thus, we obtain the inequality kgk2 − kW gk2 + 2 Re(ζ[(f, g) − (W f, W g)]) ≥ 0, g ∈ D(W ), ζ ∈ C. Set ζ = −k((f, g) − (W f, W g)), k ∈ N. Then we get kgk2 − kW gk2 − 2k(f, g) − (W f, W g)2 ≥ 0, k ∈ N. If k tends to the infinity we get (W f, W g)H = (f, g)H. ✷ Proposition 2.3 Let W be a linear non-expanding operator in a Hilbert space H, and V be a closed isometric operator in H. Then the following conditions are equivalent: (i) W ⊇ V ; (ii) W = V ⊕ T , where T is a linear non-expanding operator in H, such that D(T ) ⊆ N0(V ), R(T ) ⊆ N∞(V ). (i)⇒(ii). Let W ⊇ V . For an arbitrary element h ∈ D(V ) holds Proof. W h = V h, kW hk = kV hk = khk. Applying the previous proposition we obtain that (W h, W g) = (h, g), h ∈ D(V ), g ∈ D(W ). Denote M = D(W ) ∩ N0(V ) ⊆ N0(V ). Then (W h, W g) = 0, h ∈ D(V ), g ∈ M. Therefore W M ⊆ N∞(V ). Set T = W M . The operator T is a linear non- expanding operator and D(T ) ⊆ N0(V ), R(T ) ⊆ N∞(V ). Choose an arbitrary element f ∈ D(W ). It can be decomposed as a sum: f = f1 + f2, where f1 ∈ M0(V ) = D(V ), f2 ∈ N0(V ). Since f and f1 belong to D(W ), then f2 belongs, as well. Therefore f2 ∈ M. Thus, we obtain: D(W ) = D(V ) ⊕ D(T ), where the manifold D(T ) is not supposed to be closed. The implication (ii)⇒(i) is obvious. ✷ 2.2 Chumakin's formula for the generalized resolvents of an isometric operator. The construction of the generalized resolvents using the definition of the generalized resolvent is hard. It requires a construction of extensions of the given isometric operator in larger spaces. Moreover, it may happen that different extensions produce the same generalized resolvent. A convenient description of generalized resolvents in terms of an analytic class of operator- valued functions is provided by the following theorem. 21 Theorem 2.4 An arbitrary generalized resolvent Rζ of a closed isometric operator V , acting in a Hilbert space H, has the following form: Rζ = [EH − ζ(V ⊕ Fζ)]−1 , ζ ∈ D, (16) where Fζ is a function from S(D; N0(V ), N∞(V )). Conversely, an arbi- trary function Fζ ∈ S(D; N0(V ), N∞(V )) defines by relation (16) a gener- alized resolvent Rζ of the operator V . Moreover, to different functions from S(D; N0(V ), N∞(V )) there correspond different generalized resolvents of the operator V . Proof. Let Rζ be an arbitrary generalized resolvent of the operator V from the statement of the theorem, which corresponds to a unitary extension U ⊇ V in a Hilbert space eH ⊇ H. For an arbitrary ζ ∈ Te and h ∈ Mζ(V ), h = (EH − ζV )f , f ∈ D(V ), we may write: Rζh = P eH H (E eH − ζU)−1h = P eH H (E eH − ζU)−1(E eH − ζU)f = f = (EH − ζV )−1h; Rζ ⊇ (EH − ζV )−1. By property 3) of Theorem 2.1, the operator Rζ is invertible for ζ ∈ D. Then R−1 ζ ⊇ (EH − ζV ), ζ ∈ D. Set Tζ = 1 ζ (cid:1) , ζ(cid:0)EH − R−1 ζ ∈ D\{0}. The operator Tζ is an extension of the operator V . Suppose that a vector g ∈ H is orthogonal to the domain of the operator Tζ, i.e. 0 = (Rζh, g)H, ∀h ∈ H. Then R∗ ζg = 0. Using property 3) of Theorem 2.1 we write 0 = (g, R∗ ζg)H = (Rζg, g)H = Re(Rζg, g)H ≥ 1 2 (g, g)H ≥ 0. Therefore g = 0. Thus, we see that RζH = H. Using property 3) with arbitrary h ∈ H we may write khk2 H = (h, h)H ≤ 2 Re(Rζh, h)H ≤ 2 (Rζh, h)H ≤ 2kRζhkHkhkH; khkH ≤ 2kRζhkH, ζ ∈ D. Consequently, the operator R−1 ζ is bounded and kR−1 ζ k ≤ 2, ζ ∈ D. 22 (17) Since Rζ is closed, the operator R−1 is closed, as well. Hence, it is defined ζ on the whole space H. Thus, we conclude that the operator Tζ (ζ ∈ D\{0}) is defined on the whole H. Set Bζ = EH − R−1 ζ , ζ ∈ D, and we see that for ζ ∈ D\{0} we have Bζ = ζTζ. Using this fact and the equality R0 = EH, we obtain that Bζ is defined on the whole space H, for all ζ ∈ D. For arbitrary ζ ∈ D and f ∈ D(Bζ), f = Rζh, h ∈ H, we write kBζf k2 H = (Rζh − h, Rζh − h)H = kRζhk2 H − 2 Re(Rζh, h)H + khk2 H. Using one more time property 3) of Theorem 2.1 we conclude that kBζf kH ≤ kf kH, and therefore kBζk ≤ 1. Taking into account inequality (17) and using Proposition 2.1 we conclude that the function R−1 is analytic in D. Therefore the function Bζ is also analytic in D, and B0 = 0. By the Schwarz lemma for operator-valued functions the following inequality holds: kBζk ≤ ζ, ζ ∈ D. Therefore kTζk ≤ 1, ζ ∈ D\{0}. ζ The function Bζ, as an analytic operator-valued function, can be ex- panded into the Maclaurin series: Bζ = ∞Xk=1 Ckζ k, ζ ∈ D, where Ck are some linear bounded operators defined on the whole H. Divid- ing this representation by ζ for ζ 6= 0, we see that the function Tζ coincides with a Maclaurin series. Defining Tζ at zero by continuity, we get an analytic function in D. Applying to Tζ Proposition 2.3, we conclude that Tζ = V ⊕ Fζ, ζ ∈ D, where Fζ is a linear non-expanding operator, defined on the whole N0(V ), with values in N∞(V ). Since Tζ is analytic in D, then Fζ is analytic, as well. Therefore Fζ ∈ S(D; N0(V ), N∞(V )). By the definitions of the functions Tζ and Bζ, from the latter relation we easily get formula (16) with ζ ∈ D\{0}. For ζ = 0 the validity of (16) is obvious. 23 Let us check the second statement of the theorem. Consider an arbitrary function Fζ ∈ S(D; N0(V ), N∞(V )). Set Tζ = V ⊕ Fζ, ζ ∈ D. The function Tζ is an analytic. Observe that there exists the inverse (EH − ζTζ)−1, which is bounded and defined on the whole H, since kζTζk ≤ ζ < 1. Moreover, we have k(EH − ζTζ)hkH ≥ khkH − ζkTζhkH ≥ (1 − ζ)khkH, h ∈ H, ζ ∈ D. Therefore k(EH − ζTζ)−1k ≤ 1 1 − ζ , ζ ∈ D. By Proposition 2.1, the function (EH − ζTζ)−1 is analytic in D. Set and Rζ = (EH − ζTζ)−1, ζ ∈ D, Rζ = EH − R∗ 1 ζ , ζ ∈ De. For the function Rζ conditions 4) and 5) of Theorem 2.3 hold. Let us check the validity of the rest of conditions of this Theorem. Condition 1) for ζ ∈ D follows from the definition of the function Rζ. Choose arbitrary ζ ∈ De and g ∈ D(V ). Then 1 ζ T ∗ 1 ζ(cid:19)−1 1 ζ 1 ζ T 1 T ∗ = − = EH − (cid:18)EH − ζ(cid:19)−1!∗ = EH −(cid:18)EH − ζ(cid:19)−1 Rζ(EH − ζV )g = −ζRζ(cid:18)EH − ζ(cid:19)−1(cid:18)EH − ζ(cid:18)EH − V −1(cid:19) V g V −1(cid:19) V g = g, ζ(cid:18)EH − 1 T ∗ 1 ζ T ∗ 1 1 ; 1 ζ 1 ζ 1 ζ Rζ = EH − R∗ 1 ζ = T ∗ 1 since T ∗ 1 ζ ⊇ V −1. Thus, condition 1) of Theorem 2.3 is true. Since R0 = EH, then condition 2) of Theorem 2.3 is true, as well. Choose arbitrary h ∈ H and ζ ∈ D. Notice that Re(Rζh, h) = 1 2 ((Rζh, h) + (h, Rζh)) 24 = 1 2(cid:8)((EH − ζTζ)−1h, h) + (h, (EH − ζTζ)−1h)(cid:9) . Set f = (EH − ζTζ)−1h. Then Re(Rζh, h) = 1 2 {(f, (EH − ζTζ)f ) + ((EH − ζTζ)f, f )} = 1 2 {(f, f ) − (f, ζTζf ) + (f, f ) − (ζTζf, f )} . On the other hand, the following equality is true: 1 2 (h, h) = 1 2 ((EH − ζTζ)f, (EH − ζTζ)f ) = 1 2(cid:8)(f, f ) − (f, ζTζf ) − (ζTζf, f ) + ζ2kTζf k2(cid:9) . Comparing last relations and taking into account the inequality ζ2kTζf k2 ≤ kf k2, we get: Re(Rζh, h) ≥ 1 2 (h, h). Consequently, condition 3) of Theo- rem 2.3 is true, as well. Therefore the function Rζ is a generalized resolvent of the operator V . Let us check the last assertion of the theorem. Consider an arbitrary generalized resolvent Rζ of the operator V . Suppose that Rζ admits two representations of a type (16) with some functions Fζ and eFζ. Then ζ ∈ D\{0}. (EH − R−1 ζ )N0(V ), 1 ζ Fζ = eFζ = By continuity these functions coincide at zero, as well. ✷ Formula (16) is said to be the Chumakin formula for the generalized resolvents of a closed isometric operator. This formula contains the minimal number of parameters: the operator and the function-parameter Fζ. Due to this fact, it will be not hard to use the Chumakin formula by solving interpolation problems. 2.3 Inin's formula for the generalized resolvents of an isometric operator. Consider a closed isometric operator V in a Hilbert space H. For the oper- ator V we shall obtain another description of the generalized resolvents -- Inin's formula, see formula (25) below. It has a less transparent structure, but instead of this it is more general and coincides with Chumakin's formula in the case z0 = 0. Inin's formula turns out to be very useful in the investi- gation of the generalized resolvents of isometric operators with gaps in their 25 spectrum (by a gap we mean an open arc of the unit circle D, which consists of points of the regular type of an isometric operator). An important role in the sequel will be played by the following operator: Vz = (V − zEH)(EH − zV )−1, z ∈ D. (18) Notice that D(Vz) = Mz and R(Vz) = M 1 operator Vz is isometric and z . It is readily checked that the V = (Vz + zEH)(EH + zVz)−1 = (Vz)−z . (19) Moreover, if V is unitary, then Vz is unitary, as well, and vice versa (this follows from (19)). Let bVz ⊇ Vz be a unitary extension of the operator Vz, acting in a Hilbert space bH ⊇ H. We can define the following operator bV = (bVz + zE bH)(E bH + zbVz)−1, which will be a unitary extension of the operator V in bH. Formula (20) establishes a one-to-one correspondence between all unitary extensions bVz of the operator Vz in a Hilbert space bH ⊇ H, and all unitary extensions bV of the operator V in bH. Fix an arbitrary point z0 ∈ D. Consider an arbitrary linear non-expanding . Denote operator C with the domain D(C) = Nz0 and the range R(C) ⊆ N 1 z0 (20) V + z0;C = Vz0 ⊕ C; VC = VC;z0 = (V + z0;C + z0EH)(EH + z0V + z0;C)−1. If z0 6= 0, we may also write: VC = VC;z0 = 1 z0 EH + z02 − 1 z0 (EH + z0V + z0;C)−1; V + z0;C = − 1 z0 EH + 1 − z02 z0 (EH − z0VC;z0)−1. (21) (22) (23) (24) The operator VC is said to be an orthogonal extension of the closed iso- metric operator V , defined by the operator C. Theorem 2.5 Let V be a closed isometric operator in a Hilbert space H. Fix an arbitrary point z0 ∈ D. An arbitrary generalized resolvent Rζ of the operator V has the following representation: Rζ =(cid:2)E − ζVC(ζ;z0)(cid:3)−1 , ζ ∈ D, (25) 26 where C(ζ) = C(ζ; z0) is a functionfrom S(D; Nz0, N 1 ). Conversely, an arbi- z0 traryfunction C(ζ; z0) ∈ S(D; Nz0, N 1 ) generates by relation (25) a general- z0 ized resolvent Rζ of the operator V . To different functions from S(D; Nz0, N 1 z0 there correspond different generalized resolvents of the operator V . ) Proof. Let V be a closed isometric operator in a Hilbert space H, and z0 ∈ D\{0} be a fixed point. Consider the following linear fractional trans- formation: t = t(u) = , (26) u − z0 1 − z0u which maps the unit circle T on T, and D on D. Let bVz0 be an arbitrary unitary extension of the operator Vz0, acting in a Hilbert space bH ⊇ H, and bV be the corresponding unitary extension of the operator V , defined by relation (20). Choose an arbitrary number u ∈ Te\{0, z0, 1 }. Moreover, the following z0 conditions are equivalent: }. Then t = t(u) ∈ Te\{0, −z0, − 1 z0 u ∈ Te\{0, z0, 1 z0 } ⇔ t ∈ Te\{0, −z0, − 1 z0 }. (27) We may write: − u − z0 1 − z0u (bVz0 − tE bH )−1 =(cid:16)(bV − z0E bH)(E bH − z0bV )−1 (E bH − z0bV )(E bH − z0bV )−1(cid:19)−1 (bV − uE bH)(E bH − z0bV )−1(cid:19)−1 =(cid:18)1 − z02 1 − z02 (E bH − z0bV )(bV − uE bH)−1 1 − z02 (bV − uE bH)−1. z0(1 − z0u) 1 − z02 E bH + (1 − z0u)2 1 − z0u 1 − z0u = . = − Therefore − 1 t (E bH − (E bH − 1 tbVz0)−1 = − ubV )−1 = − 1 z0(1 − z0u) 1 − z02 E bH − (1 − z0u)2 u(1 − z02) (E bH − z0u 1 − z0u E bH + u(1 − z02) (1 − z0u)2t (E bH − 1 1 ubV )−1; tbVz0)−1 27 = − z0u 1 − z0u E bH + u(1 − z02) (1 − z0u)(u − z0) (E bH − 1 tbVz0)−1. z0 E bH + eu − z0 eu(1 − z02) (eu − z0)(1 − z0eu) Moreover, we have Set eu = 1 u , et = 1 z0 t . Observe that eu ∈ Te\{0, 1 eu ∈ Te\{0, , z0} ⇔et ∈ Te\{0, −z0, − 1 z0 1 z0 , z0}, et ∈ Te\{0, −z0, − 1 z0 }. }, (28) and the latter conditions are equivalent to the conditions from relation (27). Then (E bH −eubV )−1 = − Applying the projection operator P we come to the following equality: eu(1 − z02) (eu − z0)(1 − z0eu) (E bH −etbVz0)−1. bH H to the both sides of the last relation z0 eu − z0 1 z0 , z0}, eu ∈ Te\{0, Reu(V ) = − EH + R eu−z0 1−z0 eu (Vz0), (29) where Reu(V ), Ret(Vz0), are the generalized resolvents of the operators V and Vz0, respectively. Since Reu(V ) is analytic in Te, it is uniquely defined by the generalized re- solvent Ret(Vz0), according to relation (29). The same relation (29) uniquely defines the generalized resolvent Ret(Vz0) by the generalized resolvent Reu(V ). Thus, relation (29) establishes a one-to-one correspondence between all gen- eralized resolvents of the operator Vz0, and all generalized resolvents of the operator V . Let us apply Chumakin's formula (16) to the operator Vz0: , ). z0 (30) et ∈ D, a one-to-one correspondence between generalized resolvents. Using (29),(30) we get Ret(Vz0) =(cid:2)EH −et(Vz0 ⊕ F (et))(cid:3)−1 where F (et) is a function of the class S(D; Nz0, N 1 Consider relation (29) for pointseu ∈ D\{0, z0}, what is equivalent to the conditionet ∈ D\{0, −z0}. In this restricted case relation (29) also establishes (eu − z0)(1 − z0eu)(cid:20)EH − eu − z0 + eu(1 − z02) 1 − z0eu(cid:18)Vz0 ⊕ F(cid:18) eu − z0 1 − z0eu(cid:19)(cid:19)(cid:21)−1 , eu ∈ D\{0, z0}, eu − z0 Reu(V ) = − (31) EH z0 28 z0 ). z0 z0 write: V + z0;C(eu) ). We may = (EH − z0VC(eu);z0)(EH − z0VC(eu);z0)−1 ), and all generalized resolvents of the operator V . Relation (31) establishes a one-to-one correspondence between all functions where F (et) ∈ S(D; Nz0, N 1 F (et) from S(D; Nz0, N 1 Set C(eu) = F ( eu−z0 1−z0 eu), u ∈ D. Notice that C(eu) ∈ S(Nz0; N 1 1 − z0eu(cid:18)Vz0 ⊕ F(cid:18)eu − z0 1 − z0eu(cid:19)(cid:19) = EH − eu − z0 EH − eu − z0 1 − z0eu − eu − z0 1 − z0eu 1 − z0eu Reu(V ) = (EH −euVC(eu);z0)−1, Of course, in the caseeu = 0 relation (32) is true, as well. It remains to check the validity of relation (32) in the caseeu = z0. (EH −euVC(eu);z0)(EH − z0VC(eu);z0)−1 eu ∈ D\{0, z0}. By Chumakin's formula for Reu(V ) we see that (Reu(V ))−1 is an analytic operator-valued function in D. Using (32) we may write: By substitution of the last relation into (31), after elementary calculations we get (VC(eu);z0 − z0EH)(EH − z0VC(eu);z0)−1 1 − z02 = (32) (Rz0(V ))−1 = u. − lim eu→z0 (Reu(V ))−1 = EH − u. − lim (33) where the limits are understood in the sense of the uniform operator topology. The operator-valued function V + values are non-expanding operators in H. Then z0;C(eu) = Vz0 ⊕ C(eu) is analytic in D, and its k(EH + z0V + z0;C(eu))hk ≥ khk − z0kV + z0;C(eu))hk ≥ (1 − z0)khk, h ∈ H; eu→z0euVC(eu);z0, k(EH + z0V + z0;C(eu))−1k ≤ 1 1 − z0 , By Proposition 2.1 we obtain that the operator-valued function (EH+z0V + z0;C(eu) + z0EH)(EH + z0V + is analytic in D. Therefore VC(eu);z0 = (V + analytic in D, as well. Passing to the limit in relation (33) we get z0;C(eu))−1 is z0;C(eu))−1 (34) eu ∈ D. (Rz0(V ))−1 = EH − z0VC(z0);z0. Thus, relation (32) holds in the caseeu = z0, as well. ✷ 2.4 The generalized resolvents of a symmetric opera- tor. A connection with the generalized resolvents of the Cayley transformation of the operator. 29 Consider an arbitrary closed symmetric operator A in a Hilbert space H. The domain of A is not supposed to be necessarily dense in H. It is well known that for the operator A there always exists a self-adjoint extension eA, acting in a Hilbert space eH ⊇ H. Define an operator-valued function Rλ in the following way: Rλ = Rλ(A) = Rs;λ(A) = P H, λ ∈ Re. eH H (cid:16)eA − λE eH(cid:17)−1 The function Rλ is said to be a generalized resolvent of the symmetric necessary in those case, where there can appear a muddle with the generalized resolvent of an isometric operator. operator A (corresponding to the extension eA). The additional index s is Let {eEt}t∈R be a left-continuous orthogonal resolution of the identity of the operator eA. The operator-valued function Et = P t ∈ R, eH is said to be a (left-continuous) spectral function of a symmetric operator A (corresponding to the extension eA). Let eE(δ), δ ∈ B(R), be the orthogonal spectral measure of the self-adjoint operator eA. The function δ ∈ B(R), E(δ) = P eH H eEt, H eE(δ), is said to be a spectral measure of a symmetric operator A (corresponding to the extension eA). The spectral function and the spectral measure of the operator A are connected by the following relation: E([0, t)) = Et, t ∈ R, what follows from the analogous property of the orthogonal spectral mea- sures. Moreover, generalized resolvents and spectral functions (measures) of the operator A are connected by the following equality: (Rλh, g)H =ZR 1 t − λ d(E(·)h, g)H =ZR 1 t − λ d(Eth, g)H, ∀h, g ∈ H, (35) which follows directly from their definitions. By the Stieltjes-Perron inversion formula this means that between generalized resolvents and spectral measures there exists a one-to-one correspondence. For arbitrary elements h, g ∈ H we may write: 30 (Rλh, g)H = (Rλ(eA)h, g) eH = (h, R∗ = (h, R∗ λg)H, λ(eA)g) eH = (h, Rλ(eA)g) eH λ ∈ Re, where Rλ(eA) = (eA − λE eH )−1 is the resolvent of the self-adjoint operator eA in a Hilbert space eH, corresponding to Rλ. Therefore s;λ(A) = Rs;λ(A), λ ∈ Re. (36) R∗ Choose and fix an arbitrary number z ∈ Re. Consider the Cayley transfor- mation of the operator A: Uz = Uz(A) = (A − zEH )(A − zEH )−1 = EH + (z − z)(A − zEH )−1. (37) The operator Uz is closed and D(Uz) = Mz, R(Uz) = Mz. It is readily checked that Uz is isometric. Since Uz − EH = (z − z)(A − zEH )−1, the operator Uz has no non-zero fixed points. The operator A is expressed by Uz in the following way: A = (zUz − zEH)(Uz − EH)−1 = zEH + (z − z)(Uz − EH)−1. (38) fixed points. Conversely, if there is a unitary extension Wz of the operator Uz, acting Wz = (bA − zE bH)(bA − zE bH )−1 = E bH + (z − z)(bA − zE bH )−1, Let bA ⊇ A be a self-adjoint extension of the operator A, acting in a Hilbert space bH ⊇ H. Then the following operator: is a unitary extension of the operator Uz, acting in bH, and having no non-zero in a Hilbert space bH and having no non-zero fixed points, then the operator will be a self-adjoint extension of the operator A, acting in bH. Thus, between self-adjoint extensions bA of the operator A and unitary extensions Wz of the operator Uz, acting in bH and having no non-zero fixed points, there exists a bA = (zWz − zE bH)(Wz − E bH)−1 = zE bH + (z − z)(Wz − E bH)−1, one-to-one correspondence according to formulas (35),(38). Theorem 2.6 Let A be a closed symmetric operator in a Hilbert space H, which domain is not necessarily dense in H. Let z ∈ Re be an arbitrary fixed point, and Uz be the Cayley transformation of A. The following equality: (39) Ru; λ−z λ−z (Uz) = λ − z z − z EH + (λ − z)(λ − z) z − z Rs;λ(A), λ ∈ Re\{z, z}, (40) 31 establishes a one-to-one correspondence between all generalized resolvents Rs;λ(A) of the operator A and those generalized resolvents Ru;ζ(Uz) of the closed isometric operator Uz which are generated by extensions of Uz without non-zero fixed points. In the case D(A) = H, equality (40) establishes a one-to-one correspon- dence between all generalized resolvents Rs;λ(A) of the operator A and all generalized resolvents Ru;ζ(Uz) of the operator Uz. Choose and fix a point z ∈ Re. Let Rs;λ(A) be an arbitrary Proof. generalized resolventof a closed symmetric operator A. It is generated by a above, the operator Wz is a unitary extension of the operator Uz, acting self-adjoint operator bA ⊇ A in a Hilbert space bH ⊇ H. Consider the Cay- leytransformation Wz of the operator bA, given by (38). As it was mentioned in bH, and having no non-zero fixed elements. Consider the following linear fractional transformation: ζ = , λ = λ − z λ − z zζ − z ζ − 1 . Supposing that λ ∈ Re\{z, z}, what is equivalent to the inclusion ζ ∈ Te\{0}, we write: (cid:0)E bH − ζWz(cid:1)−1 =(cid:18)E bH − λ − z λ − z(cid:16)E bH + (z − z)(bA − zE bH )−1(cid:17)(cid:19)−1 z − z bH H to the first and to the last parts of this Applying the projection operator P relation we get relation (40). The function Ru;ζ(Uz) is a generalized resolvent of Uz of the required class. Conversely, suppose that Ru;ζ(Uz) is a generalized resolvent of Uz, gener- ated by a unitary extension Wz ⊇ Uz, acting in a Hilbert space bH ⊇ H and having no non-zero fixed elements. Define the operator bA by equality (39). As it was said above, the operator bA is a unitary extension of A in bH. The operator Wz is its Cayley's transformation. Repeating for the operator bA considerations at the beginning of the proof we shall come to relation (40). = = λ − z λ − z z − z z − z(cid:16)(bA − λE bH)(bA − zE bH )−1(cid:17)−1 (bA − zE bH )(bA − λE bH )−1 (bA − λE bH)−1. (λ − z)(λ − z) E bH + = λ − z z − z The bijectivity oof the correspondence (40) is obvious. In fact, rela- tion (40) connects all values of two generalized resolvents except two points where they are defined by the continuity. Consider the case D(A) = H. Let Ru;ζ(Uz) be an arbitrary generalized 32 resolvent of the operator Uz, which is generated by a unitary extensionfWz ⊇ Uz, acting in a Hilbert space eH ⊇ H. Let h ∈ eH be a fixed point of the operatorfWz: fWzh = h. For arbitrary element g ∈ eH we may write: (h,fWzg) eH = (fWzh,fWzg) eH = (h, g) eH; g ∈ eH. (h, (fWz − E eH )g) eH = 0, In particular, this implies that h ⊥ (Uz − EH )D(Uz) = D(A). Therefore h ∈ eH ⊖ H. Denote by M a set of all fixed elements of the operator fWz. The set M is a subspace in eH which is orthogonal to H. Thus, we have eH = H ⊕ M ⊕ H1, where H1 is a subspace in eH. Consider an operator Wz =fWzH⊕H1. The operator Wz has no non-zero fixed points and it is a unitary extension of Uz, if it is considered as an operator in H ⊕ H1. Since for an arbitrary f ∈ H we have: (E eH − ζfWz)−1f = (EH⊕H1 − ζWz)−1f, then the operator Wz also generates the generalized resolvent Ru;ζ(Uz). Thus, the set of those generalized resolvents of the operator Uz, which are generated by unitary extensions without non-zero fixed elements, coincides with the set of all generalized resolvents of the operator Uz. ✷ 2.5 Shtraus's formula for the generalized resolvents of a symmetric densely defined operator. Consider a closed symmetric operator A in a Hilbert space H, assuming that its domain is dense in H, D(A) = H. Fix an arbitrary point z ∈ Re, and consider the Cayley transformation Uz of the operator A from (37). Let F be an arbitrary linear bounded operator with the domain D(F ) = Nz(A) and range R(F ) ⊆ Nz(A). Set Uz;F = Uz;F (A) = Uz ⊕ F. Thus, the operator Uz;F (A) is a linear bounded operator defined on the whole H. This operator has no non-zero fixed elements. In fact, let for an element h ∈ H, h = h1 + h2, h1 ∈ Mz(A), h2 ∈ Nz(A), we have 0 = (Uz;F (A) − EH)h = (Uz − EH)h1 + F h2 − h2. 33 The first summand in the right-hand side of the last equality belongs to D(A), the second summand belongs to Nz(A), and the third belongs to Nz(A). But in the case D(A) = H the linear manifolds D(A), Nz(A) and Nz(A) are linearly independent, and therefore we get an iquality h1 = h2 = 0. Thus, we get h = 0. Define the following linear operator in H: AF = AF,z = (zUz;F − zEH)(Uz;F − EH )−1 = zEH + (z − z)(Uz;F − EH )−1. Notice that the operator AF,z is an extension of the operator A. The opera- tor AF = AF,z is said to be quasi-self-adjoint extension of a symmetric operator A, defined by the operator F . We outline the following prop- erty: A∗ F,z = zEH +(z−z)(U ∗ z;F −EH )−1 = zEH +(z−z)(U −1 z ⊕F ∗−EH )−1 = AF ∗,z. (41) Theorem 2.7 Let A be a closed symmetric operator in a Hilbert space H with the dense domain: D(A) = H. Fix an arbitrary point λ0 ∈ Re. An arbitrary generalized resolvent Rs;λ of the operator A has the following form: Rs;λ = (cid:0)AF (λ) − λEH(cid:1)−1 , (cid:16)AF ∗(λ) − λEH(cid:17)−1 λ ∈ Πλ0 , λ ∈ Πλ0 , (42) where F (λ) is a function from S(Πλ0; Nλ0, Nλ0). Conversely, an arbitrary function F (λ) ∈ S(Πλ0; Nλ0, Nλ0) defines by relation (42) a generalized re- solvent Rs;λ of the operator A. To different functions from S(Πλ0; Nλ0, Nλ0) there correspond different generalized resolvents of the operator A. Here Πλ0 is that half-plane of C+ and C−, which contains the point λ0. Proof. Let A be a closed symmetric operator in a Hilbert space H, D(A) = H, and λ0 ∈ Re be a fixed point. Consider an arbitrary generalized resolvent Rs;λ(A) of the operator A. Let us use Theorem 2.6 with z = λ0. From equality (40) we express Rs;λ(A): Rs;λ(A) = λ0 − λ0 (λ − λ0)(λ − λ0)(cid:18)Ru; λ−λ0 λ−λ0 (Uλ0) − λ − λ0 λ0 − λ0 EH(cid:19) , λ ∈ Re\{λ0, λ0}. Observe that the linear fractional transformation ζ = λ−λ0 maps the half- λ−λ0 plane Πλ0 on D (and R on T). Thus, we can restrict the last relation and 34 consider it only for λ ∈ Πλ0\{λ0}. In this case we may apply to the general- ized resolvent Ru; λ−λ0 (Uλ0) Chumakin's formula (16). Then λ−λ0 Rs;λ(A) = λ0 − λ0 (λ − λ0)(λ − λ0) (cid:20)EH − λ − λ0 λ − λ0 (Uλ0 ⊕ Φ λ−λ0 λ−λ0 )(cid:21)−1 − λ − λ0 λ0 − λ0 EH(cid:19) , λ ∈ Πλ0\{λ0}, (43) where Φζ is a function from S(D; Nλ0(A), Nλ0(A)). Denote F (λ) = Φ λ−λ0 λ−λ0 , λ ∈ Πλ0. Notice that F (λ) ∈ S(Πλ0; Nλ0(A), Nλ0(A)). Then Rs;λ(A) = − λ − λ0 λ0 − λ0(cid:20)EH − = 1 λ − λ0 λ − λ0 λ − λ0 λ0 − λ0 (λ − λ0)(λ − λ0) (cid:20)EH − (Uλ0 ⊕ F (λ))(cid:21)(cid:20)EH − (−EH + (Uλ0 ⊕ F (λ)))(cid:20)EH − λ − λ0 λ − λ0 λ ∈ Πλ0\{λ0}. (Uλ0 ⊕ F (λ))(cid:21)−1 (Uλ0 ⊕ F (λ))(cid:21)−1! (Uλ0 ⊕ F (λ))(cid:21)−1 , λ − λ0 λ − λ0 λ − λ0 λ − λ0 (44) In the half-plane Πλ0 the following inequality holds: Consider the open neighborhood of the point λ0: (cid:12)(cid:12)(cid:12)(cid:12) λ − λ0 λ − λ0(cid:12)(cid:12)(cid:12)(cid:12) < 1, D(λ0) =(cid:26)z ∈ Πλ0 : λ ∈ Πλ0. z − λ0 < In the closed neighborhood D(λ0) the function(cid:12)(cid:12)(cid:12) λ−λ0 tains its maximal value q < 1. Therefore (cid:12)(cid:12)(cid:12)(cid:12) λ − λ0 λ − λ0(cid:12)(cid:12)(cid:12)(cid:12) ≤ q < 1, λ ∈ D(λ0). 1 2 Im λ0(cid:27) . λ−λ0(cid:12)(cid:12)(cid:12) is continuous and at- For an arbitrary element h ∈ H we may write: and therefore λ − λ0 λ − λ0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:20)EH − (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)EH − ≥ (1 − q)khk, (Uλ0 ⊕ F (λ))(cid:21) h(cid:13)(cid:13)(cid:13)(cid:13) ≥(cid:12)(cid:12)(cid:12)(cid:12)khk −(cid:12)(cid:12)(cid:12)(cid:12) (Uλ0 ⊕ F (λ))(cid:21)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ λ − λ0 λ − λ0 35 λ − λ0 λ − λ0(cid:12)(cid:12)(cid:12)(cid:12) k(Uλ0 ⊕ F (λ))hk(cid:12)(cid:12)(cid:12)(cid:12) 1 1 − q , λ ∈ D(λ0). Applying Proposition 2.1 we conclude that the functionhEH − λ−λ0 is analytic in D(λ0). Passing to the limit in relation (44) as λ → λ0 we obtain: (Uλ0 ⊕ F (λ))i−1 λ−λ0 Rs;λ0(A) = 1 λ0 − λ0 (−EH + (Uλ0 ⊕ F (λ0))) . Thus, relation (44) holds for λ = λ0, as well. In our notations, maid at the beginning of this subsection, we have Uλ0 ⊕ F (λ) = Uλ0;F (λ). According to our above remarks Uλ0;F (λ) has no non-zero fixed elements. Therefore there exists the inverse R−1 λ − λ0 λ − λ0 s;λ(A) = (λ − λ0)(cid:20)EH − Uλ0;F (λ)(cid:21)(cid:0)Uλ0;F (λ) − EH(cid:1)−1 =(cid:2)λEH − λ0EH − λUλ0;F (λ) + λ0Uλ0;F (λ)(cid:3)(cid:0)Uλ0;F (λ) − EH(cid:1)−1 = −λEH +(cid:2)λ0Uλ0;F (λ) − λ0EH(cid:3)(cid:0)Uλ0;F (λ) − EH(cid:1)−1 = −λEH + AF (λ);λ0, λ ∈ Πλ0. From this relation we conclude that (42) holds for λ ∈ Πλ0. Suppose now that λ is such that λ ∈ Πλ0. Applying the proved part of the formula for λ we get Using property (36) of the generalized resolvent and using relation (41) we write . Rs;λ =(cid:16)AF (λ) − λEH(cid:17)−1 F (λ) − λEH(cid:17)−1 s;λ =(cid:16)A∗ =(cid:16)AF ∗(λ) − λEH(cid:17)−1 . Rs;λ = R∗ Consequently, relation (42) for λ ∈ Πλ0 is true, as well. Conversely, let an operator-valued function F (λ) ∈ S(Πλ0; Nλ0, Nλ0) be 36 given. Set Φ(ζ) = F(cid:18)λ0ζ − λ0 ζ − 1 (cid:19) , ζ ∈ D. The function Φ(ζ) belongs to S(D; Nλ0(A), Nλ0(A)). By Chumakin's for- mula (16), to this function there corresponds a generalized resolvent Ru;ζ(Uλ0). Define a generalized resolvent Rs;λ(A) of the operator A by relation (43). Repeating for the generalized resolvent Rs;λ(A) considerations after (43) we come to relation (42). Therefore the function F (λ) generates by relation (42) a generalized resolvent of the operator A. Suppose that two operator-valued functions F1(λ) and F2(λ), which be- long to S(Πλ0; Nλ0, Nλ0), generate by relation (42) the same generalized re- solvent. In this case we have (cid:0)AF1(λ) − λEH(cid:1)−1 and therefore =(cid:0)AF2(λ) − λEH(cid:1)−1 , AF1(λ) = AF2(λ), λ ∈ Πλ0. λ ∈ Πλ0, From the last relation it follows that Uλ0;F1(λ) = Uλ0;F2(λ), and we get the equality F1(λ) = F2(λ). ✷ Formula (42) is said to be the Shtraus formula for the generalized resol- vents of a symmetric operator with a dense domain. 3 Generalized resolvents of non-densely de- fined symmetric operators. 3.1 The forbidden operator. Throughout this subsection we shall consider a closed symmetric operator A in a Hilbert space H, which domain can be non-dense in H. Proposition 3.1 Let A be a closed symmetric operator in a Hilbert space H and z ∈ Re. Let elements ψ ∈ Nz(A) and ϕ ∈ Nz(A) be such that ϕ − ψ ∈ D(A), i.e. they are comparable by modulus D(A). Then kϕkH = kψkH. Proof. g ∈ Mz(A) such that Since D(A) = (Uz − EH )Mz(A), then there exists an element ϕ − ψ = Uzg − g. Then ϕ + Uzg = ψ + g. (45) Using the orthogonality of summands in the left and right sides we conclude that 37 kψk2 H + kgk2 H = kϕk2 H + kUzgk2 H = kϕk2 H + kgk2 H, and the required equality follows. ✷ Corollary 3.1 Let A be a closed symmetric operator in a Hilbert space H. Then D(A) ∩ Nz(A) = {0}, ∀z ∈ Re. Proof. and ϕ = 0, we get ψ = 0. ✷ If ψ ∈ D(A) ∩ Nz(A), then applying Proposition 3.1 to elements ψ Let z ∈ Re be an arbitrary number. Consider the following operator: Xzψ = Xz(A)ψ = ϕ, ψ ∈ Nz(A) ∩ (Nz(A) ∔ D(A)) , where ϕ ∈ Nz(A): ψ − ϕ ∈ D(A). This definition will be correct if an element ϕ ∈ Nz such that ψ − ϕ ∈ D(A), is unique (its existence is obvious, since ψ ∈ (Nz(A) ∔ D(A))). Let ϕ1 ∈ Nz: ψ − ϕ1 ∈ D(A). Then ϕ − ϕ1 ∈ D(A). Since ϕ − ϕ1 ∈ Nz, then by applying Corollary 3.1 we get ϕ = ϕ1. From Proposition 3.1 it follows that the operator Xz is isometric. Observe that D(A) = Nz(A)∩(Nz(A) ∔ D(A)) and R(A) = Nz(A)∩(Nz(A) ∔ D(A)). The operator Xz = Xz(A) is said to be forbidden with respect to the symmetric operator A. Now we shall obtain another representation for the operator Xz. Before this we shall prove some auxilliary results: Proposition 3.2 Let A be a closed symmetric operator in a Hilbert space H and z ∈ Re. The following two conditions are equivalent: (i) Elements ψ ∈ Nz(A) and ϕ ∈ Nz(A) are comparable by modulus D(A); (ii) Elements ψ ∈ Nz(A) and ϕ ∈ Nz(A) admit the following representa- tion: ψ = P H Nz(A)h, ϕ = P H Nz(A)h, where h ∈ H is such that Mz(A)h = UzP H P H Mz(A)h. (46) (47) If these conditions are satisfied, the element h is defined uniquely and admits the following representation: h = 1 z − z (A(ψ − ϕ) − zψ + zϕ) . (48) 38 Proof. derive formula (45), where g ∈ Mz(A). Set h = g + ψ. Applying P H this equality we obtain the first equality in (46). From (45) it follows that (i) ⇒ (ii). As at the beginning of the proof of Proposition 3.1, we Nz(A) to ϕ + Uz(h − ψ) = h. The summands in the left-hand side belong to orthogonal subspaces. Ap- plying the projection operators P H Mz(A), we obtain the following equalities: Nz(A) and P H ϕ = P H Nz(A)h, Uz(h − ψ) = P H Mz(A)h, and therefore the second equality in (46) is satisfied. Using the latter equality and the equality P H Mz g = g, we may write Mz (g + ψ) = P H Mz h = P H P H Mz(A)h = Uzg = UzP H Mz(A)h. Consequently, relation (47) is true, as well. (ii) ⇒ (i). Set g = P H Mz h. Then h = P H Mz h + P H Mz h + P H h = P H Nz h = g + ψ, Nzh = Uzg + ϕ, and subtracting last equalities we obtain that ψ − ϕ = Uzg − g ∈ D(A). Moreover, we have: h = g + ψ = (Uz − EH)−1(ψ − ϕ) + ψ = 1 z − z (A − zEH )(ψ − ϕ) + ψ, and relation (48) follows. ✷ Proposition 3.3 Let A be a closed symmetric operator in a Hilbert space H, and z ∈ Re. Then H ⊖ D(A) =(cid:8)h ∈ H : P H Mz(A)h = UzP H Mz(A)h(cid:9) . Proof. For arbitrary elements g ∈ Mz(A) and h ∈ H we may write: (Uzg, h)H = (Uzg, P H Mz(A)h)H; (g, h)H = (g, P H Mz(A)h)H = (Uzg, UzP H Mz(A)h)H, and therefore If h ⊥ D(A), then the left-hand side of the last equality is equal to zero, ((Uz − EH )g, h)H =(cid:0)Uzg,(cid:0)P H since (Uz − EH )g ∈ D(A). Then(cid:0)P H Conversely, if h ∈ H is such that (cid:0)P H Mz − UzP H ((Uz − EH)g, h)H = 0, g ∈ Mz(A), i.e. h ⊥ D(A). ✷ Mz(cid:1) h = 0. Mz(cid:1) h(cid:1)H . Mz(cid:1) h = 0, then we get Mz − UzP H Mz − UzP H 39 Corollary 3.2 Let A be a closed symmetric operator in a Hilbert space H. Then(cid:16)H ⊖ D(A)(cid:17) ∩ Mz(A) = {0}, ∀z ∈ Re. Proof. following equality holds: Let h ∈(cid:16)H ⊖ D(A)(cid:17) ∩ Mz(A), z ∈ Re. By Proposition 3.3 the P H Mz(A)h = UzP H Mz(A)h. Nz(A)h = 0 and ϕ := Nz(A)h are comparable by modulo D(A). By Proposition 3.1 we get ϕ = 0. By Proposition 3.2 we obtain that elements ψ := P H P H From (48) it follows that h = 0. ✷ Let us check that D(Xz) = P H Nz(A)(H ⊖ D(A)), XzP H Nz(A)h = P H Nz(A)h, h ∈ H ⊖ D(A). (49) (50) In fact, if ψ ∈ D(Xz), then ψ ∈ Nz and there exists ϕ ∈ Nz such that ψ −ϕ ∈ D(A). By Propositions 3.2 and 3.3 we see that there exists h ∈ H ⊖ D(A): ψ = P H Nz(A)h. Therefore D(Xz) ⊆ P H Nz(A)(H ⊖ D(A)). Nz(A)(H ⊖ D(A)), ψ = P H Conversely, let ψ ∈ P H Nz(A)h, h ∈ H ⊖ D(A). By Propositions 3.2 and 3.3 we obtain that ψ and ϕ := P H Nz(A)h are compa- rable by modulus D(A). Consequently, we have ψ ∈ D(Xz) and therefore P H Nz(A)(H ⊖ D(A)) ⊆ D(Xz). Thus, equality (49) is true. For an arbitrary element h ∈ H ⊖D(A) we set ψ = P H Nz(A)h. By Propositions 3.2 and 3.3, we have ψ − ϕ ∈ D(A). Therefore ψ ∈ D(Xz) and Xzψ = ϕ, and equality (50) follows. Nz(A)h, ϕ = P H 3.2 Admissible operators. We continue considering of a closed symmetric operator A in a Hilbert space H. Fix an arbitrary point z ∈ Re. As before, Xz will denote the forbidden operator with respect to A. Let V be an arbitrary operator with the domain D(V ) ⊆ Nz(A) and the range R(V ) ⊆ Nz(A). The operator V is said to be z-admissible (or admissible) with respect to symmetric operator A, if the operator V − Xz is invertible. In other words, V is admissible with respect to the operator A, if the equality V ψ = Xzψ can happen only if ψ = 0. Using the definition of the forbidden operator Xz, the latter definition can be formulated in the following way: V is admissible with respect to the operator A, if ψ ∈ D(V ), V ψ − ψ ∈ D(A), implies ψ = 0. 40 If the domain of the operator A is dense in H, then by (49) we get D(Xz) = {0}. Consequently, in the case D(A) = H, an arbitrary operator with the domain D(V ) ⊆ Nz(A) and the range R(V ) ⊆ Nz(A) is admissible with respect to A. Let B be a symmetric extension of the operator A in the space H. Then its Cayley's transformation Wz = (B − zEH)(B − zEH )−1 = EH + (z − z)(B − zEH )−1, (51) is an isometric extension of the operator Uz(A) in H, which has no non-zero fixed elements. Moreover, we have: B = (zWz − zEH)(Wz − EH)−1 = zEH + (z − z)(Wz − EH)−1. (52) Conversely, for an arbitrary isometric operator Wz ⊇ Uz(A), having no non- zero fixed elements, by (52) one can define a symmetric operator B ⊇ A. Thus, formula (51) establishes a one-to-one correspondence between all sym- metric extensions B ⊇ A in H, and all isometric extensions Wz ⊇ Uz(A) in H, having no non-zero fixed elements. By Proposition 2.3, all such extensions Wz have the following form: Wz = Uz(A) ⊕ T, (53) where T is a isometric operator with the domain D(T ) ⊆ Nz(A) and the range R(T ) ⊆ Nz(A). Conversely, if we have an arbitrary isometric operator T with the domain D(T ) ⊆ Nz(A) and the range R(T ) ⊆ Nz(A), then by (53) we define an isometric extension Wz of the operator Uz(A). Question: what additional property should have the operator T which garantees that the operator Wz has no non-zero fixed elements? An answer on this question is provided by the following theorem. Theorem 3.1 Let A be a closed symmetric operator in a Hilbert space H, z ∈ R be a fixed point, and Uz be the Cayley transformation of the operator A. Let V be an arbitrary operator with the domain D(V ) ⊆ Nz(A) and the range R(V ) ⊆ Nz(A). The operator Uz ⊕V having no non-zero fixed elements if and only if the operator V is z-admissible with respect to A. Necessity. To the contrary, suppose that Uz ⊕ V has no non- Proof. zero fixed elements but V is not admissible with respect to A. This means that there exists a non-zero element ψ ∈ D(V ) ∩ D(Xz) such that V ψ = Xzψ, where Xz denotes the forbidden operator. By (49),(50) there exists an element h ∈ H ⊖ D(A): ψ = P H Nz(A)h and 41 V P H Nz(A)h = XzP H Nz(A)h = P H Nz(A)h. By Proposition 3.3 we obtain the following equality: Mz(A)h = UzP H P H Mz(A)h. Then (Uz ⊕ V )h = UzP H Mz(A)h + V P H Nz(A)h = P H Mz(A)h + P H Nz(A)h = h, i.e. h is a fixed element of the operator Uz ⊕ V . Since we assumed that Uz ⊕ V has no non-zero fixed elements, then h = 0, and therefore ψ = 0. We obtained a contradiction. Sufficiency. A, but there exists a non-zero element h ∈ H such that Suppose to the contrary that V is admissible with respect to (Uz ⊕ V )h = UzP H Mz(A)h + V P H Nz(A)h = h = P H Mz(A)h + P H Nz(A)h. Then the following equalities hold: UzP H Mz(A)h = P H Mz(A)h, V P H Nz(A)h = P H Nz(A)h. By Proposition 3.3 and the first of these equalities we get h ∈ H ⊖ D(A). By relations (49),(50) we see that ψ := P H Nz(A)h belongs to D(Xz), and XzP H Nz(A)h = P H Nz(A)h = V P H Nz(A)h, i.e. V ψ = Xzψ. Since V is admissible with respect to A, then ψ = 0. By the definition of the operator Xz, the element ϕ := Xzψ is comparable with ψ by modulus D(A). By Proposition 3.1 we conclude that ϕ = 0. By Proposition 3.2 and formula (48) we get h = 0. The obtained contradiction completes the proof. ✷ Remark 3.1 The last Theorem shows that formulas (51),(52) and (53) es- tablish a one-to-one correspondence between all symmetric extensions B of the operator A in H, and all isometric operators T with the domain D(T ) ⊆ Nz(A) and the range R(T ) ⊆ Nz(A), which are admissible with respect to A. A symmetric operator A is said to be maximal if it has no proper (i.e. different from A) symmetric extensions in H. By the above remark the maximality depends on the number of isometric operators T with D(T ) ⊆ Nz(A), R(T ) ⊆ Nz(A), which are admissible with respect to A. 42 Consider arbitrary subspaces of the same dimension: N 0 z ∈ Nz(A) and z ∈ Nz(A). Let us check that there always exists an isometric operator V , z , and which is admissible with respect to the operator N 0 which maps N 0 A. In order to do that we shall need the following simple proposition. z on N 0 Proposition 3.4 Let H1 and H2 are subspaces of the same dimension in a Hilbert space H. Then there exists an isometric operator V with the domain D(V ) = H1 and the range R(V ) = H2, which has no non-zero fixed elements. Proof. Choose orthonormal bases {fk}d H2, respectively. The operator k=0 and {gk}d k=0, d ≤ ∞, in H1 and U dXk=0 αkfk = dXk=0 αkgk, αk ∈ C, maps isometrically H1 on H2. Denote H0 = {h ∈ H1 ∩ H2 : Uh = h}, i.e. H0 is a set of the fixed elements of the operator U. Notice that H0 is a subspace in H1. Choose and fix a number α ∈ (0, 2π). Set V h = eiαP H1 H0 h + UP H1 H1⊖H0h, h ∈ H1. Suppose that g ∈ H1 is a fixed element V , i.e. V g = eiαP H1 H0 g + UP H1 H1⊖H0g = g = P H1 H0 g + P H1 H1⊖H0g. Since (UP H1 H1⊖H0g, h) = (UP H1 H1⊖H0g, Uh) = (P H1 H1⊖H0g, h) = 0, ∀h ∈ H0, then UP H1 get (eiα − 1)P H1 and P H1 H1⊖H0g ⊥ H0, and equating summands in the previous equality we H1⊖H0g ∈ H0, H1⊖H0g. Therefore P H1 H0 g = 0 and UP H1 H1⊖H0g = P H1 H1⊖H0g = 0. Thus, we get g = 0. ✷ Consider the following linear manifold: M =(cid:8)ψ ∈ D(Xz) ∩ N 0 z : Xzψ ∈ N 0 z(cid:9) . Set Denote Xz;0 = XzM . 43 N ′ z = M ⊆ N 0 z , N ′ z = R(Xz;0) ⊆ N 0 z . Since Xz is isometric, we get dim N ′ z = dim N ′ z. Consider an arbitrary isometric operator W , which maps the subspace N 0 z on the subspace N 0 z of the same dimention. Set z ⊆ N 0 z . K := W N ′ Since W is isometric we get dim K = dim N ′ z = dim N ′ z, dim(N 0 z ⊖ K) = dim(N 0 z ⊖ N ′ z). (54) We shall say that an isometric operator S corrects Xz;0 in N 0 z , if the following three conditions are satisfied: 1) D(S) = N ′ z, R(S) ⊆ N 0 z ; 2) dim(N 0 z ⊖ R(S)) = dim(N 0 z ⊖ N ′ z); 3) S has no non-zero fixed elements in R(Xz;0). By Proposition 3.4 and the first equality in (54), there exists an isometric operator, which maps N ′ z on K, and has no non-zero fixed elements. By the second equality in (54) condition 2) for this operator is satisfied, as well. Therefore an isometric operator correcting Xz;0 in N 0 z always exists. Theorem 3.2 Let A be a closed symmetric operator in a Hilbert space H, z ∈ R be a fixed point, N 0 z ∈ Nz(A) be subspaces of the same dimension. z ∈ Nz(A) and N 0 An arbitrary admissible with respect to A isometric operator V , which maps N 0 z on N 0 z , has the following form: V = SXz;0 ⊕ T, (55) where S is an isometric operator, which corrects Xz;0 in N 0 isometric operator with the domain D(T ) = N 0 N 0 z ⊖ N ′ z , and T is an z and the range R(T ) = z ⊖ R(S). Conversely, an arbitrary isometric operator S, which corrects Xz;0 in N 0 z , z and z ⊖ R(S), generate by relation (55) an admissible with and an arbitrary isometric operator T with the domain D(T ) = N 0 the range R(T ) = N 0 respect to A isometric operator V , which maps N 0 z ⊖ N ′ z on N 0 z . 44 Proof. admissible with respect to A isometric operator, which maps N 0 Denote Let us check the first assertion of the Theorem. Let V be an z on N 0 z . and Then Q = V N ′ z, T = V N 0 z , z ⊖N ′ S = Q(Xz;0)−1 : N ′ z → V N ′ z. V = SXz;0 ⊕ T. The operator S is isometric. Let us check that it corrects Xz;0 in N 0 z . Since dim(N 0 z ⊖ R(S)) = dim(N 0 z ⊖ (V N ′ z)) = dim(V (N 0 z ⊖ N ′ z)) = dim(N 0 z ⊖ N ′ z), then the second condition from the definition of the correcting operator is satisfied. Let ϕ ∈ R(Xz;0) is a fixed element of the operator S: Sϕ = Q(Xz;0)−1ϕ = ϕ. Denote ψ = (Xz;0)−1ϕ = X −1 z;0 ϕ ∈ M. Then V ψ = Qψ = Xz;0ψ. Since V is admissible with respect to A isometric operator then ψ = 0, and ϕ = Xz;0ψ = 0. Thus, the operator S corrects Xz;0 in N 0 z . z ⊖ N ′ z and the range R(T ) = N 0 Conversely, let S be an arbitrary isometric operator which corrects Xz;0 in N 0 z , and T be an arbitrary isometric operator with the domain D(T ) = N 0 z ⊖ R(S). Define the operator V by relation (55). The operator V is isometric and maps N 0 z . Let us check that V is admissible with respect to A. Consider an arbitrary element ψ ∈ D(V ) ∩ D(Xz) such that V ψ = Xzψ. z on N 0 At first, we consider the case ψ ∈ N ′ z. In this case we have SXz;0ψ = Xzψ. By the definition of the closure of an operator and by the continuity of Xz, we can assert that there exists a sequence ψn ∈ D(Xz;0), n ∈ N, such that ψn → ψ, and Xzψn = Xz;0ψn → Xz;0ψ = Xzψ, as n → ∞. Therefore SXzψ = Xzψ. By the definition of the correcting operator, S can not have non-zero fixed elements in R(Xz;0). Two cases are posiible: 1) Xzψ ∈ R(Xz;0) Xzψ = 0. By the invertibility of Xz this means that ψ = 0. 2) Xzψ /∈ R(Xz;0). This means that ψ /∈ D(Xz;0). Since ψ ∈ D(Xz) ∩ N ′ z, then it is possible only in the case Xzψ /∈ N 0 z . But this is impossible since Xzψ = V ψ ∈ N 0 z . Thus, this case should be excluded. Consider the case ψ /∈ N ′ z. In this case ψ /∈ D(Xz;0) and ψ ∈ N 0 z = D(V ). z . But then the z . Consequently, the z is impossible. Thus, we obtain that the operator V is admissible By the definition of Xz;0, it is possible only if Xzψ /∈ N 0 equality V ψ = Xzψ will be impossible, since V ψ ∈ N 0 case ψ /∈ N ′ with respect to A. ✷ 45 Corollary 3.3 Let A be a closed symmetric operator in a Hilbert space H. The following assertions hold: (i) The operator A is maximal if and only if (at least) one of its defect numbers is equal to zero. (ii) If A is maximal then D(A) = H. z ⊆ Nz(A) and N 0 z on N 0 Proof. (i): Necessity. Suppose to the contrary that A is maximal but its both defect numbers are non-zero. Choose and fix an arbitrary number z ∈ Re. Let N 0 z ⊆ Nz(A) be some one-dimensional subspaces. By Theorem 3.2, there exists an operator V , which is admissible with respect to A and which maps N 0 z . By Remark 3.1 it follows that to the operator V there corresponds a symmetric extension B of the operator A, according to relations (51)-(53). This extension is proper since to A by (51)-(53) there corresponds the null isometric operator T . We obtained a contradiction. (ii): Sufficiency. If one of the defect numbers is equal to zero, then the Cayley transformation Uz(A) does not have proper extensions. Consequently, taking into account considerations after (52), the operator A has no proper extensions. (ii). Let A be maximal. From the proved assertion (i) it follows that one of the defect numbers is equal to zero, i.e. Nz(A) = {0}, for some z ∈ Re. By Corollary 3.2 we may write: Fix an arbitrary number z ∈ Re. H ⊖ D(A) =(cid:16)H ⊖ D(A)(cid:17) ∩ Mz(A) = {0}, D(A) = H. ✷ 3.3 Properties of dissipative and accumulative opera- tors. A linear operator A in a Hilbert space H is said to be dissipative (accu- mulative), if Im(Ah, h)H ≥ 0 (respectively Im(Ah, h)H ≤ 0) for all elements h from D(A). A dissipative (accumulative) operator A is said to be max- imal, if it has no proper (i.e. different from A) dissipative (respectively accumulative) extensions in H. We establish some properties of dissipative and accumulative operators, which will be used later. Theorem 3.3 Let A be a closed dissipative (accumulative) operator in a Hilbert space H. The following assertions are true: 1) Points of C− (respectively C+) are points of the regular type of the operator A. The following inequality holds: 46 (cid:13)(cid:13)(A − ξEH)−1(cid:13)(cid:13) ≤ 1 Im ξ , ξ ∈ C− (C+). (56) 2) Choose and fix a number z from C− (respectively from C−). Let B be a dissipative (respectively accumulative) extension of the operator A in the space H. Consider the following operator Wz = (B − zEH)(B − zEH )−1 = EH + (z − z)(B − zEH )−1. (57) The operator Wz is a non-expanding extension of the operator Uz(A) = (A − zEH )(A − zEH )−1 H, having no non-zero fixed elements. More- over, we have: B = (zWz − zEH)(Wz − EH)−1 = zEH + (z − z)(Wz − EH)−1. (58) Conversely, for an arbitrary non-expanding operator Wz ⊇ Uz(A), hav- ing no non-zero fixed elements, by (58) one defines a dissipative (re- spectively accumulative) operator B ⊇ A. Formula (57) establishes a one-to-one correspondence between all dissipative (respectively accu- mulative) extensions B ⊇ A in H, and all non-expanding extensions Wz ⊇ Uz(A) H, having no non-zero fixed elements. 3) If A is maximal, then (A − zEH )D(A) = H, for all points z from C− (respectively C+). Conversely, if there exists a point z0 from C− (respectively C+) such that (A − z0EH )D(A) = H, then the operator A is maximal; 4) If A is maximal, then D(A) = H. 1) Consider the case of a dissipative operator A. Choose an Proof. arbitrary point z from C−, z = x + iy, x ∈ R, y < 0. For an arbitrary element f ∈ D(A) we may write: k(A − zEH )f k2 H = ((A − xEH )f − iyf, (A − xEH )f − iyf )H = k(A − xEH)f k2 H − 2y Im(Af, f )H + y2kf k2 H ≥ y2kf k2 H. Therefore the operator A − zEH has a bounded inverse and inequality (56) holds. In the case of an accumulative operator A, the operator −A is dissipative and we may apply to it the proved part of the assertion. 47 2) Let z from C− (respectively from C+) and B be a dissipative (respectively accumulative) extension of the operator A in the space H. By the proved assertion 1), the operator Wz is correctly defined and bounded. For an arbi- trary element g ∈ D(Wz) = (B − zEH )D(B), g = (B − zEH )f , f ∈ D(B), we may write: kWzgk2 H = (Wzg, Wzg)H = ((B − zEH)f, (B − zEH)f )H = kBf k2 kgk2 = kBf k2 H − z(Bf, f )H − z(f, Bf )H + z2kf k2 H; H = ((B − zEH )f, (B − zEH )f )H H − z(Bf, f )H − z(f, Bf )H + z2kf k2 H, and therefore kWzgk2 H − kgk2 H = (z − z)((Bf, f )H − (f, Bf )H) = 4(Im z) Im(Bf, f )H ≤ 0. Since Wz −EH = (z−z)(B−zEH )−1, then Wz has no non-zero fixed elements. Formula (58) follows directly from (57). Conversely, let Wz ⊇ Uz(A) be a non-expanding operator having no non- zero fixed elements. Define an operator B by equality (58). For an arbitrary element f ∈ D(B) = (Wz − EH)D(Wz), f = (Wz − EH)g, g ∈ D(Wz), we write: (Bf, f )H = ((zWz − zEH )g, (Wz − EH )g)H = zkWzgk2 H − z(Wzg, g)H Im(Bf, f )H = Im(z)kWzgk2 H + Im(z)kgk2 −z(g, Wzg)H + zkgk2 H; H = Im(z)(cid:0)kWzgk2 H − kgk2 H(cid:1) . Consequently, the operator B is dissipative (respectively accumulative). Since Wz ⊇ Uz(A), then B ⊇ A. The correspondence, established by formu- las (57),(58), is obviously one-to-one. 3) Suppose to the contrary that the operator A is maximal but there exists a number z from C− (respectively C−) such that (A − zEH )D(A) 6= H. Consider the following operator Uz = Uz(A) = (A − zEH)(A − zEH )−1 By the proved assertion 2), the operator Uz is non-expanding and has no non- zero fixed elements. Moreover, Uz is closed, since A is closed by assumption. Therefore Uz is defined on a subspace H1 = (A − zEH )D(A) 6= H. Set H2 = H ⊖ H1, dim H2 ≥ 1. Consider the following operator Wz := Uz ⊕ 0H2. The operator Wz is a non-expanding extension of the operator Uz. Let us check that the operator Wz has no non-zero fixed elements. Suppose to the 48 contrary that there exists a non-zero element h ∈ H such that Wzh = h. Then h /∈ D(Uz), since Uz has no non-zero fixed elements. Let h = h1 + h2, h1 ∈ H1, h2 ∈ H2, and h2 6= 0. Then Wzh = (Uz ⊕ 0H2)(h1 + h2) = Uzh1 = h = h1 + h2. Therefore kh1k2 H + kh2k2 H, H ≥ kUzh1k2 H = kh1k2 and we get h2 = 0. The obtained contradiction shows that Wz has no non- zero fixed elements. By the proved assertion 2) to the operator Wz 6= Uz there corresponds a dissipative (respectively accumulative) extension B 6= A of the operator A. This contradicts to the maximality of the operator A, and the first part of assertion 3) is proved. Suppose that we have a closed dissipative (respectively accumulative) operator A in a Hilbert space H, and there exists a point z0 from C− (re- spectively C+) such that (A − z0EH)D(A) = H. Suppose to the contrary that A is not maximal. Let B be a proper dissipative (respectively accu- mulative) extension of the operator A. By the proved assertion 2), for the operator B there corresponds a non-expanding operator Wz0 from (57) with z = z0, having no non-zero fixed elements. Moreover, we have D(Wz0) = (B−z0EH)D(B) ⊇ (A−z0EH)D(A) = H. Therefore D(Wz0) = H = D(Uz0), and we get Wz0 = Uz0, B = A. We obtained a contradiction, since B is a proper extension. 4) Consider an arbitrary element h ∈ H: h ⊥ D(B). Choose and fix an arbitrary number z from C− (respectively C+). By the proved assertion 3), the following equality holds: (A − zEH )D(A) = H. Therefore there exists an element f ∈ D(A) such that (A − zEH )f = h. Then 0 = ((A − zEH )f, f )H = (Af, f )H − zkf k2 H; Im(Af, f )H = Im(z)kf k2 H. Since the left-hand side of the last equality is non-negative (respectively non- positive), and the right-hand side is non-positive (respectively non-negative), then f = 0. ✷ Remark 3.2 Consider a closed symmetric operator A in a Hilbert space H. Choose and fix a number z from C− (C−). Let B be a dissipative (respectively accumulative) extension of the operator A in H. By asser- tion 2) of Theorem 3.3 for it there corresponds a non-expanding operator Wz from (57), having no non-zero fixed elements. By Proposition 2.3, such extensions Wz ⊇ Uz(A) have the following form: Wz = Uz(A) ⊕ T, (59) 49 where T is a linear nonexpanding operator with the domain D(T ) ⊆ Nz(A) and the range R(T ) ⊆ Nz(A). By Theorem 3.1 it follows that T is admissible with respect to the operator A. Conversely, if we have an arbitrary non-expanding operator T , with the domain D(T ) ⊆ Nz(A), the range R(T ) ⊆ Nz(A), admissible with respect to A, then by (59) one defines a non-expanding extension Wz of the operator Uz(A). By Theorem 3.1 this extension has no non-zero fixed elements. Thus, there is a one-to-one correspondence between admissible with respect to A non-expanding operators T and non-expanding operators Wz ⊇ Uz(A) without non-zero fixed elements. Using the proved theorem we conclude that formulas (57)-(59) estab- lish a one-to-one correspondence between admissible with respect to A non- expanding operators T , D(T ) ⊆ Nz(A), R(T ) ⊆ Nz(A), and dissipative (re- spectively accumulative) extensions of the closed symmetric operator A. 3.4 Generalized Neumann's formulas. The classical Neumann's formulas describe all symmetric extensions of a given closed symmetric operator A in a Hilbert space H in the case D(A) = H. The following theorem provides a description such extensions without the assumption D(A) = H, and it describes dissipative and accumulative extensions of A, as well. Theorem 3.4 Let A be a closed symmetric operator in a Hilbert space H, and z from C− (C+) be a fixed number. The following formulas D(B) = D(A) ∔ (T − EH)D(T ), B(f + T ψ − ψ) = Af + zT ψ − zψ, f ∈ D(A), ψ ∈ D(T ), (60) (61) establish a one-to-one correspondence between all admissible with respect to A isometric operators T , D(T ) ⊆ Nz(A), R(T ) ⊆ Nz(A), and all symmetric extensions B of the operator A. Moreover, we have D(T ) = Nz(A) ∩ R(B − zEH ), T ⊆ (B − zEH)(B − zEH )−1. (62) (63) A symmetric operator B is: closed / closed and maximal / self-adjoint, if and only if respectively: D(T ) is: a subspace / D(T ) = Nz(A) or R(T ) = Nz(A) / D(T ) = Nz(A) and R(T ) = Nz(A). Formulas (60),(61) define a one-to-one correspondence between all admis- sible with respect to A non-expanding operators T , D(T ) ⊆ Nz(A), R(T ) ⊆ 50 Nz(A), and all dissipative (respectively accumulative) extensions B of the operator A. Also relations (62),(63) hold. A dissipative (respectively accu- mulative) operator B: closed / closed and maximal, if and only if respectively: D(T ) is: a subspace / D(T ) = Nz(A). Proof. Let A be a closed symmetric operator in a Hilbert space H, and z be from C− (C+). By Remark 3.1, formulas (51),(52) and (53) establish a one- to-one correspondence between all symmetric extensions B of the operator A in H, and all isometric operators T with the domain D(T ) ⊆ Nz(A) and the range R(T ) ⊆ Nz(A), which are admissible with respect to A. Let us check that formulas (52),(53) define the same operator B, as formulas (60),(61). Let the operator B be defined by formulas (52),(53). The domain of B is D(B) = (Wz − EH )D(Wz) = (Uz ⊕ T − EH)(D(Uz) + D(T )) = (Uz − EH)D(Uz) + (T − EH )D(T ). If f ∈ (Uz − EH)D(Uz) ∩ (T − EH)D(T ), then f = (Uz − EH )g, g ∈ D(Uz), and f = (T − EH )h, h ∈ D(T ). Then (Uz ⊕ T − EH)(g − h) = f − f = 0, i.e. g − h is a fixed element of the operator Wz. Since Wz has no non-zero fixed elements, we get g = h. Since g ⊥ h, then g = 0 and f = 0. Thus, for the operator B holds (60). Consider arbitrary elements f ∈ D(A) and ψ ∈ D(V ). Then B(f + V ψ − ψ) = Bf + B(V − EH )ψ = Af + B(Wz − EH)ψ = Af + (zWz − zEH )ψ = Af + zV ψ − zψ, and therefore formula (61) holds, as well. Since Nz(A) ∩ (B − zE)D(B) = Nz(A) ∩ D(Wz) = Nz(A) ∩ (Mz(A) + D(T )) = Nz(A) ∩ D(T ) = D(T ), then relation (62) holds. Relation (63) means that T ⊆ Wz. If the operator B is closed, then Wz is closed, as well. Since Wz is bounded, it is defined on a subspace. Therefore D(Wz) = R(B − zEH ) is closed, D(T ) = Nz(A) ∩ R(B − zEH ) is closed and it is a subspace. Conversely, if D(T ) is a subspace then the direct sum Mz(A) ⊕ D(T ) = D(Wz) is closed. Therefore the operators Wz and B are closed. 51 By Corollary 3.3 a closed symmetric operator B in a Hilbert space H is maximal if and only if (at least) one of its defect numbers is equal to zero. This is equivqlent to the following condition: D(T ) = Nz(A) or R(T ) = Nz(A). If B if self-adjoint then its Cayley's transformation Wz is unitary, i.e. D(T ) = Nz(A) and R(T ) = Nz(A). Conversely, the last conditions imply that Wz is unitary, and therefore B is self-adjoint. By Remark 3.2 formulas (57)-(59) establish a one-to-one correspondence between admissible with respect to A non-expanding operators T , D(T ) ⊆ Nz(A), R(T ) ⊆ Nz(A), and dissipative (respectively accumulative) exten- sions of a closed symmetric operator A. Formulas (60),(61) define the same operator B as formulas (57)-(59). It is checked similar as it was done above for the case of symmetric extensions. The proof of relations (62),(63), and the proof of equivalence: (B is closed) ⇔ (B is a subspace) are the same. Let B be closed and maximal. By assertion 3) of Theorem 3.3 we obtain that R(B − zEH ) = (B − zEH )D(B) = H. From (62) it follows that D(T ) = Nz(A). Conversely, let D(T ) = Nz(A). Then (B − zEH )D(B) = D(Wz) = Mz(A) ⊕ D(T ) = Mz(A) ⊕ Nz(A) = H. By assertion 3) of Theorem the operator B is maximal. ✷ Let A be a closed symmetric operator in a Hilbert space H, and z ∈ Re be a fixed number. Let T be an admissible with respect to A non-expanding operator with D(T ) ⊆ Nz(A), R(T ) ⊆ Nz(A). An extension B of the opera- tor A which corresponds to T in formulas (60),(61) we denote by AT = AT,z and call a quasi-self-adjoint extension of a symmetric operator A defined by the operator T . This definition agrees with the above given definition for the case of a densely defined symmetric operator A. 3.5 Extensions of a symmetric operator with an exit out of the space. Let A be a closed symmetric operator in a Hilbert space H. In order to construct a generalized resolvent of A, according to its definition, one needs to construct self-adjoint extensions eA of A in larger Hilbert spaces eH ⊇ H. Choose and fix a Hilbert space eH ⊇ H. Decompose it as a direct sum: where He = eH ⊖ H. The operator A can be identified with the operator A ⊕ oHe in eH, where oHe is an operator in He with D(oHe) = {0}, oHe0 = 0. eH = H ⊕ He, (64) Keeping in mind this important case, we shall now assume that in a Hilbert space eH of a type (64), where H, He are some Hilbert spaces, there is given a symmetric operator A of the following form: A = A ⊕ Ae, 52 (65) where A, Ae are symmetric operators in Hilbert spaces H,He, respectively. Let us investigate a possibility for the construction of symmetric and self- adjoint extensions of the operator A and their properties. In particular, we can use the generalized Neumann formulas. Fix an arbitrary number z ∈ Re. Since D(A) = D(A) ⊕ D(Ae), then it can be directly verified that Mz(A) = Mz(A) ⊕ Mz(Ae), Nz(A) = Nz(A) ⊕ Nz(Ae), Xz(A) = Xz(A) ⊕ Xz(Ae), (66) (67) (68) where Xz(·) denotes the forbidden operator. Let T be an arbitrary non-expanding operator in eH with the domain D(T ) = Nz(A) and the range R(T ) ⊆ Nz(A). With respect to the decompo- sition (67) (for z and z) the operator T has the following block representation: T21 T22 (cid:19) , T =(cid:18) T11 T12 (69) where T11 = PNz(A)T PNz(A), T12 = PNz(A)T PNz(Ae), T21 = PNz(Ae)T PNz(A), T22 = PNz(Ae)T PNz(Ae). By the definition of the admissible operator the operator T is admissible with respect to operator A if and only if relation ψ ∈ D(Xz(A)), Xz(A)ψ = T ψ, implies ψ = 0. Using decompositions (68),(69) we see that the latter condi- tion is equivalent to the following one: relation ψ1 ∈ D(Xz(A)), ψ2 ∈ D(Xz(Ae)), T11ψ1 + T12ψ2 = Xz(A)ψ1, T21ψ1 + T22ψ2 = Xz(Ae)ψ2, (70) implies ψ1 = ψ2 = 0. Theorem 3.5 Let A be a closed symmetric operator of the form (65), where A, Ae are symmetric operators in some Hilbert spaces H and He, respectively. Let z ∈ Re be a fixed number and T be a non-expanding operator in eH = H ⊕ He, with the domain D(T ) = Nz(A) and the range R(T ) ⊆ Nz(A). The following assertions are true: 53 1) If T is z-admissible with respect to the operator A, then the opera- tors T11 and T22 are z-admissible with respect to operators A and Ae, respectively. 2) If D(A) = H and T22 is z-admissible with respect to the operator Ae, then the operator T is z-admissible with respect to the operator A. Here operators T11 and T22 are the operators from the block representation (69) for T . Proof. Let us check assertion 1). Let ψ1 be an element from D(Xz(A)) such that T11ψ1 = Xz(A)ψ1. Since T is non-expanding and the forbidden operator is isometric, we may write: kψ1k2 H ≥ kT ψ1k2 = kT11ψ1k2 H + kT21ψ1k2 H = kXz(A)ψ1k2 H + kT21ψ1k2 H = kψ1k2 H + kT21ψ1k2 H. Therefore T21ψ1 = 0. Set ψ2 = 0. We see that condition (70) is satisfied. By the condition of the theorem T is admissible with respect to A, and therefore ψ1 = 0. This means that the operator T11 is admissible with respect to A. In a similar way, if ψ2 is an element from D(Xz(Ae)) such that T22ψ2 = Xz(Ae)ψ2, then kψ2k2 H ≥ kT ψ2k2 = kT12ψ2k2 H + kT22ψ2k2 H = kT12ψ2k2 H + kXz(Ae)ψ2k2 H = kT12ψ2k2 H + kψ2k2 H. Therefore T12ψ2 = 0. Set ψ1 = 0. We conclude that condition (70) holds. Consequently, we get ψ2 = 0. Let us check assertion 2). Let ψ1 ∈ D(Xz(A)), ψ2 ∈ D(Xz(Ae)), and relation (70) holds. Since D(A) = H, then by (49) the domain of the operator Xz(A) consists of the null element and ψ1 = 0. The second equality in (70) may be written as T22ψ2 = Xz(Ae)ψ2. Since T22 is admissible with respect to Ae, we get ψ2 = 0. Consequently, T is admissible with respect to A. ✷ Let us continue our considerations started before the formulation of the last theorem. Let ψ1 ∈ Nz(A) and ψ2 ∈ D(Xz(Ae)) be such elements (for example, null elements) that Then T21ψ1 + T22ψ2 = Xz(Ae)ψ2. kT11ψ1 + T12ψ2kH ≤ kψ1kH. (71) (72) In fact, since T is non-expanding, we have 54 kT (ψ1 + ψ2)k2 H H = kT11ψ1 + T12ψ2k2 H + kT21ψ1 + T22ψ2k2 ≤ kψ1k2 H + kψ2k2 H. If (71) holds, then since Xz(Ae) is isometric we get (72). Moreover, if an equality kT (ψ1 +ψ2)kH = kψ1 +ψ2kH holds, then in (72) we have an equality, and vice versa. Define the following operator: Φψ1 = Φ(z; A, T )ψ1 = T11ψ1 + T12ψ2, where ψ1 ∈ Nz(A) is such an element, for which there exists ψ2 ∈ D(Xz(Ae)): T21ψ1 + T22ψ2 = Xz(Ae)ψ2. Let us check that this definition is correct. If for ψ1 ∈ Nz(A) there exists another element eψ2 ∈ D(Xz(Ae)): then subtracting last relations we obtain that T21ψ1 + T22eψ2 = Xz(Ae)eψ2, T22(ψ2 − eψ2) = Xz(Ae)(ψ2 − eψ2). Set ψ1 = 0. We see that equality (71) holds, and therefore inequality (72) is true and it takes the following form: kT12(ψ2 − eψ2)kH ≤ 0, i.e. T12ψ2 = T12eψ2. Thus, the definition of the operator Φ does not depend on the choice of ψ2 and it is correct. By (72) the operator Φ is non-expanding. Moreover, the operator Φ is isometric, if T is isometric (what follows from the considerations before the definition of Φ). Remark 3.3 If it is additionally known that the operator T22 is admissi- ble with respect to the operator Ae, then it exists (Xz(Ae) − T22)−1. Con- sider an arbitrary element ψ1 ∈ D(Φ). By the definition of Φ, for the corresponding to ψ1 the element ψ2 eqality (71) holds, and therefore ψ2 = (Xz(Ae) − T22)−1T21ψ1. Substitute this expression in the definition of the operator Φ to obtain the following representation: Φψ1 = T11ψ1 + T12(Xz(Ae) − T22)−1T21ψ1, ψ1 ∈ D(Φ). (73) 55 Theorem 3.6 Let A be a closed symmetric operator of the form (65), where A, Ae are symmetric operators in some Hilbert spaces H and He, respec- tively. Let z ∈ Re be a fixed number and T be a non-expanding operator in eH = H ⊕ He, with the domain D(T ) = Nz(A) and the range R(T ) ⊆ Nz(A). The operator T is z-admissible with respect to A if and only if the operators Φ(z; A, T ) and T22 are z-admissible with respect to the operators A and Ae, respectively. Here the operators T22 is an operator from the block representa- tion (69) for T . Necessity. Let an operator T be admissible with respect to A. Proof. By Theorem 3.5 the operator T22 is admissible with respect to Ae. Consider an element ψ1 ∈ D(Φ(z; A, T )) such that Φ(z; A, T )ψ1 = Xz(A)ψ1. By the definition of the operator Φ(z; A, T ), the element ψ1 belongs to Nz(A) and there exists ψ2 ∈ D(Xz(Ae)): T21ψ1 + T22ψ2 = Xz(Ae)ψ2. Moreover, we have Φ(z; A, T )ψ1 = T11ψ1 + T12ψ2. Thus, conditions (70) are true. Since T is admissible with respect to A, then these inequalities imply ψ1 = ψ2 = 0. Therefore the operator Φ is admissible with respect to A. Sufficiency. Suppose that Φ(z; A, T ) and T22 are admissible with respect to A and Ae, respectively. Consider some elements ψ1 ∈ D(Xz(A)), ψ2 ∈ D(Xz(Ae)), such that (70) holds. This relation means that ψ1 ∈ D(Φ) and Φψ1 = Xz(A)ψ1. Since Φ is admissible with respect to A, then ψ1 = 0. From the second equality in (70) it follows that T22ψ2 = Xz(Ae)ψ2. Since T22 is admissible with respect to Ae, we get ψ2 = 0. Consequently, the operator T is admissible with respect to the operator A. ✷ Now we shall obtain a more explicit expression for the domain of Φ and for its action. Theorem 3.7 Let A be a closed symmetric operator of the form (65), where A, Ae are symmetric operators in some Hilbert spaces H and He, respectively. Let z ∈ Re be a fixed number and T be a non-expanding operator in eH = H ⊕He, with the domain D(T ) = Nz(A), and with the range R(T ) ⊆ Nz(A). Then (74) (75) (76) h ∈ eΘz, Heho . eH D(Φ(z; A, T )) = P eH Nz(A)eΘz, Φ(z; A, T )P eH Nz(A)h = P eH Nz(A)(Uz(A) ⊕ T )h, where eΘz =nh ∈ eH : P eH He(Uz(A) ⊕ T )h = P Proof. We shall need the next lemma on the structure of the set eΘz. 56 Lemma 3.1 In conditions of Theorem 3.7, the set eΘz consists of elements h ∈ eH, where g1 ∈ Mz(A), ψ1 ∈ Nz(A), g2 ∈ Mz(Ae), ψ2 ∈ Nz(Ae), such that h = g1 + ψ1 + g2 + ψ2, (77) ψ2 ∈ D(Xz(Ae)), Xz(Ae)ψ2 = T21ψ1 + T22ψ2, g2 = 1 z − z (Ae − zEHe)(Xz(Ae) − EH2)ψ2. (78) (79) Proof of Lemma. Consider an arbitrary element h ∈ eΘz. As each element of eH, the element h has a (unique) decomposition (77). Since eH Heh = g2 + ψ2, P P eH He(Uz(A) ⊕ T )h = P eH He (Uz(A)(g1 + g2) + T (ψ1 + ψ2)) = Uz(Ae)g2 + T21ψ1 + T22ψ2, then by the definition of the set eΘz we obtain: Uz(Ae)g2 + T21ψ1 + T22ψ2 = g2 + ψ2. (80) Then T21ψ1 + T22ψ2 − ψ2 = (EHe − Uz(Ae))g2 = (z − z)(Ae − zEHe)−1g2 ∈ D(Ae). Since T21ψ1 + T22ψ2 ∈ Nz(Ae), then ψ2 ∈ D(Xz(Ae)) Xz(Ae)ψ2 = T21ψ1 + T22ψ2. From the latter relations it follows (78),(79). Conversely, if for an element h ∈ H of form (78) holds (78) and (79), then (EHe − Uz(Ae))g2 = (z − z)(Ae − zEHe)−1g2 = (Xz(Ae) − EHe)ψ2 = T21ψ1 + T22ψ2 − ψ2. proof of Lemma). By the proved Lemma it follows that if ψ1 ∈ P Therefore equality (80) holds, which means that h ∈ D(eΘz). ✷ (end of the Nz(A)eΘz, then ψ1 ∈ Nz(A), and there exists element ψ2 ∈ D(Xz(Ae) such that Xz(Ae)ψ2 = T21ψ1+T22ψ2. Conversely, if ψ1 ∈ Nz(A), and there exists an element ψ2 ∈ D(Xz(Ae)) such that Xz(Ae)ψ2 = T21ψ1 + T22ψ2, then Xz(Ae)ψ2 − ψ2 ∈ D(Ae). Define g2 by (79), and as g1 we take an arbitrary element from Mz(A). Then eH h := ψ1 + g1 + ψ2 + g2 ∈ eΘz, and therefore ψ1 ∈ P eH Nz(A)eΘz. Comparing this with the definition of the operator Φ we conclude that 57 relation (74) holds. Choose an arbitrary element h ∈ P representations (77), which elements satisfy relations (78),(79). Then eH Nz(A)eΘz. By the proved Lemma, it has P eH Nz(A)(Uz(A) ⊕ T )h = P eH Nz(A)T (ψ1 + ψ2) = T11ψ1 + T12ψ2; Φ(z; A, T )P eH Nz(A)h = Φ(z; A, T )ψ1 = T11ψ1 + T12ψ2, and the required equality (75) follows. ✷ Corollary 3.4 In conditions of Theorem 3.7 consider the following operator Wz = Uz(A) ⊕ Φ(z; A, T ). Then D(Wz) = P eH HeΘz, (81) (82) WzP eH H h = P eH H (Uz(A) ⊕ T )h, h ∈ eΘz. Proof. By Lemma 3.1 we get Mz(A) ⊆ eΘz. Therefore Let h be an arbitrary element from eΘz, having representation (77) which elements satisfy, according to Lemma 3.1, relations (78),(79). Using (75), we may write Nz(A)eΘz = Mz(A) + D(Φ) = D(Wz). HeΘz = Mz(A) + P eH P eH WzP eH H h = (Uz(A) ⊕ Φ(z; A, T ))(g1 + ψ1) = Uz(A)g1 + P eH Nz(A)h = P eH Mz(A)(Uz(A) ⊕ T )h + P Therefore equality (82) holds. ✷ eH Nz(A)(Uz(A) ⊕ T )h = P eH H (Uz(A) ⊕ T )h. Consider our considerations interrupted by formulations of the last The- orem and its Corollary. In the sequel we assume that the operator T is admissible with respect to the symmetric operator A. In this case, by Theo- rem 3.6 and Remark 3.3 the non-expanding operator Φ(z; A, T ) is admissible with respect to A and has a representation (73). By the generalized Neumann formulas (Theorem 3.4), for the operator T it corresponds a quasi-self-adjoint extension AT of the symmetric operator A. Define the operator B on the manifold D(AT ) ∩ H in H in the following way: Bh = B(z; A, T )h = P eH H AT h, h ∈ D(AT ) ∩ H. (83) 58 Theorem 3.8 Let A be a closed symmetric operator of the form (65), where A, Ae are symmetric operators in some Hilbert spaces H and He, respectively. Let z ∈ Re be a fixed number and T be a non-expanding operator in eH = H ⊕ He, with D(T ) = Nz(A), R(T ) ⊆ Nz(A), which is admissible with respect to A. The operator B, defined on the manifold D(AT ) ∩ H in H by equality (83), admits the following representation: B(z; A, T ) = AΦ(z;A,T ). Proof. Consider a set eΘz, defined by (76). This definition we can rewrite in another form: (84) eΘz =nh ∈ eH : (cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) h ∈ Ho . Therefore Since R(cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) = D(AT ), then (cid:0)(Uz(A) ⊕ T ) − E eH(cid:1)eΘz ⊆ H. (cid:0)(Uz(A) ⊕ T ) − E eH(cid:1)eΘz ⊆ H ∩ D(AT ). Conversely, an arbitrary element g ∈ H ∩ D(AT ) has the following form g =(cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) f , f ∈ eH. By (84) the element f belongs to the set eΘz. Therefore (cid:0)(Uz(A) ⊕ T ) − E eH(cid:1)eΘz = H ∩ D(AT ) = D(B), where the operator B is defined by equality (83). Consider the operator Wz from the Corollary 3.4. By (82) we may write (85) (Wz−EH )P eH H h = P Therefore eH H (cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) h =(cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) h, Taking into account relation (81) we get (Wz − EH)P eH HeΘz = D(B). h ∈ eΘz. (86) (Wz − EH)D(Wz) = D(B). (87) Choose an arbitrary element u ∈ D(B). By (85) the element u has the following form: u =(cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) h, h ∈ eΘz. Then Bu = P eH H AT u 59 = P eH H (cid:0)z(Uz(A) ⊕ T ) − zE eH(cid:1)(cid:0)(Uz(A) ⊕ T ) − E eH(cid:1)−1(cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) h eH H (Uz(A) ⊕ T )h − zP eH H h = P eH H (cid:0)z(Uz(A) ⊕ T ) − zE eH(cid:1) h = zP eH H h − zP = zWzP eH H h = (zWz − zEH )P eH H h, where we used (82). Since the operator Φ is admissible with respect to the operator A, then Wz has no non-zero fixed elements. By formula (86) we get Bu = (zWz − zEH)(Wz − EH)−1(cid:0)(Uz(A) ⊕ T ) − E eH(cid:1) h = (zWz − zEH )(Wz − EH )−1u = AΦ(z;A,T )u, u ∈ D(B). Equality (87) shows that D(B) = D(AΦ(z;A,T )). Therefore B = AΦ(z;A,T ). ✷ 3.6 The operator-valued function Bλ generated by an extension of a symmetric operator with an exit out of the space. Consider a closed symmetric operator A in a Hilbert space H. Let eA be a self-adjoint extension of A, acting in a Hilbert space eH ⊇ H. Denote λ ∈ C; (88) eLλ =nh ∈ D(eA) : (eA − λE eH)h ∈ Ho , HeLλ, L∞ = D(eA) ∩ H. Lλ = P λ ∈ C; eH (89) (90) Notice that the sets eLλ, Lλ and L∞ contain D(A). Since for non-real λ for the self-adjoint operator eA holds (eA − λE eH)D(eA) = eH, then λ ∈ Re. (91) Observe that λ ∈ Re. (92) (eA − λE eH)eLλ = H, eLλ ∩ (eH ⊖ H) = {0}, In fact, if h ∈ eLλ ∩ (eH ⊖ H), then ((eA − λE eH)h, h) eH = 0. Therefore 0 = Im(eAh, h) eH = (Im λ)(h, h) eH, h = 0. Define an operator-valued function Bλ = Bλ(A, eA) for λ ∈ Re, which values are operators in H with the domain: D(Bλ) = Lλ, (93) and BλP eH H h = P eH H eAh, h ∈eLλ. (94) 60 Let us check that such definition of the operator Bλ is correct. If an element g ∈ Lλ admits two representations: g = P eH H h1 = P eH h1 − h2 = P eH H h1 − P eH H h2 + P eH eH⊖Hh1 − P H h2, h1, h2 ∈eLλ, then eH eH⊖H h2 = P eH eH⊖Hh1 − P eH eH⊖Hh2 = P eH eH⊖H(h1 − h2). The operator Bλ for each λ ∈ Re is an extension of A. Define an operator Therefore h1 − h2 ∈eLλ ∩ (eH ⊖ H). By (92) we obtain that h1 = h2. B∞ = B∞(A, eA) in H in the following way: H eAL∞. The operator B∞ is an extension of A, as well. Let us check that B∞ is symmetric. In fact, for arbitrary elements f, g ∈ L∞ we may write B∞ = P (95) eH (B∞f, g)H = (P = (f, P eH eH H eAf, g)H = (eAf, g) eH = (f, eAg) eH H eAg)H = (f, B∞g)H. We emphasize that the operator B∞ is not necessarily closed. The operator- valued function Bλ and the operator B∞ will play an important role in a description of the generalized resolvents of A. Notice that (eA − λE eH)h = (Bλ − λEH )P eH H h, h ∈eLλ, λ ∈ Re. In fact, by the definitions of the sets eLλ and Bλ, for an arbitrary element h from eLλ we may write: eH H h eH eH (eA − λE eH)h = P = BλP H (eA − λE eH )h = P eH H h = (Bλ − λEH )P H eAh − λP eH H h. eH H h − λP (96) Let us check that for λ from C− (C+) the operator Bλ is a maximal closed dissipative (respectively accumulative) extension of A. In fact, using rela- tion (96) for an arbitrary element h from eLλ we may write: H h, h)H = (eAh − λ(E eH − P eH H h)H = (BλP eH H h, P (BλP eH eH eH⊖Hhk2 H . eH H )h, h)H = (eAh, h)H − λkP eH H h, P eH H h)H = −kP Im(BλP eH eH⊖Hhk2 H Im λ, Then 61 and therefore Bλ is a dissipative (respectively accumulative) extension of A. By (91),(96) we get R(Bλ −λEH ) = (Bλ −λEH )P Using formula (62) and the generalized Neumann formulas for the operator Bλ it corresponds a non-expanding operator T with D(T ) = Nz(A). There- fore Bλ is closed and maximal. By the fourth assertion of Theorem 3.3 the operator Bλ is densely defined, λ ∈ Re. HeLλ = (eA−λE eH )eLλ = H. eH Consider a generalized resolvent of A which corresponds to the self-adjoint extension eA: Let us check that Rλ = P eH H (eA − λE eH)−1, λ ∈ Re. Rλ = (Bλ − λEH)−1, λ ∈ Re. (97) Consider an arbitrary element g ∈ H and denote h = (eA − λE eH )−1g, where λ ∈ Re. Since h ∈ D(eA) and (eA − λE eH)h ∈ H, then h ∈eLλ. Moreover, we eH H h = Rλg. By (96) we may write: have P g = (eA − λE eH)h = (Bλ − λEH )P eH H h = (Bλ − λEH)Rλg. By the properties of maximal dissipative and accumulative operators (see The- orem 3.3), there exists the bounded inverse (Bλ − λEH)−1 defined on the whole H. Applying this operator to the latter equality we get the required relation (97). By (97) it follows that the generalized resolvent Rλ of the symmetric op- λ is densely defined. erator A is invertible for all λ ∈ Re, and the operator R−1 The operator Bλ admits another definition by the generalized resolvent: Bλ = R−1 λ + λEH, λ ∈ Re. Therefore B∗ λ = (R∗ λ)−1 + λEH = R−1 λ λ ∈ Re. = Bλ, + λEH (98) (99) where we used the property (36) of the generalized resolvent. Choose and fix a number λ0 ∈ Re. By Theorem 3.4, for the operator Bλ for λ ∈ Πλ0 it corresponds an admissible with respect to A non-expanding operator, which we denote by F(λ), with the domain Nλ0(A) and the range in Nλ0(A). Namely, we set F(λ) = F(λ; λ0, A, eA) = (Bλ − λ0EH )(Bλ − λ0EH )−1(cid:12)(cid:12)Nλ0 (A) , λ ∈ Πλ0. (100) Notice that Bλ = AF(λ),λ0, λ ∈ Πλ0. 62 (101) i.e. Bλ is a quasi-self-adjoint extension of the symmetric operator A, defined by F(λ). By the generalized Neumann formulas for the operator B∞ there corre- sponds an admissible with respect to A isometric operator Φ∞ = Φ∞(λ0; A, eA) with the domain D(Φ∞) ⊆ Nλ0(A) and the range R(Φ∞) ⊆ Nλ0(A). This operator will be used later. Proposition 3.5 Let A be a closed symmetric operator in a Hilbert space H, z ∈ Re be an arbitrary fixed number, and T be a linear non-expanding operator with D(T ) = Nz(A) and R(T ) ⊆ Nz(A), which is z-admissible with respect to A. Then the operator T ∗ is z-admissible with respect to A and (AT,z)∗ = AT ∗,z. Proof. By Theorem 3.4 the operator AT,z is maximal dissipative or accu- mulative, since D(T ) = Nz(A). By Theorem 3.3 we conclude that AT,z is densely defined and therefore there exists its adjoint. By (58) we may write: (AT,z)∗ =(cid:0)zEH + (z − z)(Uz(A) ⊕ T − EH)−1(cid:1)∗ = zEH + (z − z)(Uz(A) ⊕ T ∗ − EH )−1. Consequently, the operator Uz(A) ⊕ T ∗ has no non-zero fixed elements. By Theorem 3.1 the operator T ∗ is admissible with respect to A, and the last equality gives the required formula. ✷ By the proved Proposition and formula (99) we may write: Bλ = B∗ λ = (AF(λ),λ0)∗ = AF∗(λ),λ0, λ ∈ Πλ0. Let us check that the operator-valued function F(λ) is an analytic function of λ in a half-plane Πλ0. By (97) we may write: Bλ − λ0EH = (Bλ − λ0EH )Rλ(Bλ − λEH) = ((Bλ − λEH) + (λ − λ0)EH ) Rλ(Bλ − λEH) = (EH + (λ − λ0)Rλ)(Bλ − λEH ), λ ∈ Re. By the maximality of the dissipative or accumulative operator Bλ the opera- tor (Bλ − λEH)−1 exists and it is defined on the whole H. If we additionally 63 assume that λ ∈ Πλ0, then the operator (Bλ − λ0EH)−1 exists and is defined on the whole H, as well. Then the operator EH + (λ − λ0)Rλ = (Bλ − λ0EH )(Bλ − λEH)−1, has a bounded inverse: (EH +(λ−λ0)Rλ)−1 = (Bλ−λEH )(Bλ−λ0EH)−1 = EH+(λ0−λ)(Bλ−λ0EH)−1, defined on the whole H, for λ ∈ Πλ0. By property (56) we get (cid:13)(cid:13)(EH + (λ − λ0)Rλ)−1(cid:13)(cid:13) ≤ 1 + λ0 − λ Im λ0 , λ ∈ Πλ0. By Proposition 2.1 the function (EH + (λ − λ0)Rλ)−1 is analytic in the half- plane Πλ0. Using relation (97) we may write that (Bλ − λ0EH)(Bλ − λ0EH )−1 = EH + (λ0 − λ0)(Bλ − λ0EH)−1 = EH + (λ0 − λ0)(Bλ − λEH)−1(Bλ − λEH)(Bλ − λ0EH )−1 = EH + (λ0 − λ0)Rλ(EH + (λ − λ0)Rλ)−1, is an analytic function in Πλ0. Thenthe function F(λ) is analytic in Πλ0, as well. From our considerations we conclude that the operator-valued function Bλ admits the following representation: Bλ =(cid:26) AF(λ),λ0, λ ∈ Πλ0 AF∗(λ),λ0, λ ∈ Πλ0 , (102) where F(λ) is an analytic function of λ in the half-plane Πλ0, which values are non-expanding operators with D(F(λ)) = Nλ0(A) and R(F(λ)) ⊆ Nλ0(A), admissible with respect to the operator A. Our next aim will be derivation of a representation for the operator-valued function F(λ), which connects it with the so-called characteristic function of a symmetric operator A. The next subsection will be decoted to the definition of the characteristic function and some its properties. 3.7 The characteristic function of a symmetric opera- tor. Let A be a closed symmetric operator in a Hilbert space H. Choose an arbitrary number λ ∈ Re. Notice that the operator T = 0Nλ is admissible 64 with respect to A. In fact, suppose that for some h ∈ H holds (Uλ ⊕T )h = h. Let h = h1 + h2, h1 ∈ Mλ, h2 ∈ Nλ, then (Uλ ⊕ T )h = Uλh1 = h = h1 + h2, and therefore kh1k2 H = kUλh1k2 H = kh1k2 H + kh2k2 H. Consequently, we get h2 = 0. Since the operator Uλ has no non-zero fixed elements, then h1 = h = 0. By Theorem 3.1 we obtain that T is admissible with respect to A. Consider a quasi-self-adjoint extension Aλ := A0N ,λ of the operator A. By the generalized Neumann formulas for λ from C+ (C−) the operator Aλ is maximal closed dissipative (respectively accumulative) extension of A. By Proposition 3.5 it follows that λ A∗ λ = A∗ 0N λ ,λ = A0Nλ ,λ = Aλ. Moreover, by (60),(61) we may write: D(Aλ) = D(A) ∔ Nλ(A), (103) and Aλ(f + ψ) = Af + λψ, f ∈ D(A), ψ ∈ Nλ(A). (104) Choose and fix an arbitrary number λ0 ∈ Re and consider the half-plane Πλ0 containing the point λ0. If we additionally suppose that λ ∈ Πλ0, for the quasi-self-adjoint extension Aλ by the generalized Neumann formulas it corresponds a non-expanding operator C(λ) with D(C(λ)) = Nλ0(A), R(C(λ)) ⊆ Nλ0(A). Namely, the operator C(λ) is given by the following equality: C(λ) = C(λ; λ0, A) = (Aλ − λ0EH)(Aλ − λ0EH)(cid:12)(cid:12)Nλ0 This operator-valued function C(λ) in the half-plane Πλ0 is said to be the characteristic function of the symmetric operator A. Observe that (A) , λ ∈ Πλ0. Aλ = AC(λ),λ0. (105) Proposition 3.6 Let A be a closed symmetric operator in a Hilbert space H, and λ, z ∈ Re be arbitrary numbers: λ ∈ Πz. The space H can be represented as a direct sum: H = Mλ(A) ∔ Nz(A). 65 The projection operator Pλ,z in H on the subspace Nz(A) parallel to the subspace Mλ(A) (i.e. the operator which to arbitrary vector h ∈ H, h = h1 + h2, h1 ∈ Mλ(A), h2 ∈ Nz(A), put into correspondence a vector Pλ,zh = h2) has the following form: Pλ,z = λ − z z − z Nz(A)(Az − zEH )(Az − λEH )−1. P H (106) Proof. Since Az for z from C+ (C−) is a maximal dissipative (respectively accumulative) extension of A, then there exist the inverses (Az − λEH)−1 and (Az − zEH )−1, defined on the whole H (see Theorem 3.3). The following operator: Sλ,z := (Az − λEH)(Az − zEH )−1 = EH + (z − λ)(Az − zEH )−1, maps bijectively the whole space H on the whole H, and Sλ,z := (Az − zEH )(Az − λEH)−1. By (104) it follows that (Az −zEH )ψ = (z −z)ψ, for an arbitrary ψ ∈ Nz(A). Therefore (Az − zEH )−1 = 1 z−z ENz(A). For an arbitrary g ∈ Nz(A) we may write: Sλ,zg = g + (z − λ)(Az − zEH )−1g = g + z − λ z − z g = z − λ z − z g, (107) to get For an arbitrary vector f ∈ D(A) holds: Sλ,zNz(A) = Nz(A). Sλ,z(A − zEH )f = (A − λEH)f, and therefore Sλ,zMz(A) = Mλ(A). Choose an arbitrary element h ∈ H, and set f = S−1 f1 ∈ Mz(A), f2 ∈ Nz(A), then λ,zh. Let f = f1 + f2, h = Sλ,zf1 + Sλ,zf2 ∈ Mλ(A) + Nz(A). Therefore H = Mλ(A) + Nz(A). Suppose that v ∈ Mλ(A) ∩ Nz(A). Then there exist elements u ∈ Mz(A) and w ∈ Nz(A) such that Sλ,zu = v and Sλ,zw = v. Therefore Sλ,z(u − w) = 0, and by the invertibility of Sλ,z we get u = w. Since elements u and w are orthogonal, then u = w = 0 and v = 0. Thus, the required decomposition of the space H is proved. Moreover, that for the element h holds 66 Pλ,zh = Sλ,zf2 = Sλ,zP H Nz(A)f = Sλ,zP H Nz(A)S−1 λ,zh = z − λ z − z Nz(A)S−1 P H λ,zh = z − λ z − z Nz(A)(Az − zEH )(Az − λEH)−1h, P H where we used relation (107). Thus, equality (106) is proved. ✷ Continue our considerations started before the formulation of the last proposition. Define the following two operator-valued functions: Q(λ) = Pλ,λ0, K(λ) = Pλ,λ0Nλ0 , λ ∈ Πλ0. By representation (106) it follows that Q(λ) and K(λ) are analytic in the half-plane Πλ0. In fact, since the operator Aλ0 is dissipative or accumulative, then estimate (56) holds. By Proposition 2.1 we obtain that the operator- valued function (Az − λEH)−1 is analytic in Πλ0. Then the same is true for Q(λ) and P (λ). Let us check that for each λ ∈ Πλ0 the operator K(λ) is non-expanding. In fact, for arbitrary λ ∈ Πλ0 and ψ ∈ Nλ0 we can assert that ψ−K(λ)ψ belongs to Mλ(A), and therefore ψ − K(λ)ψ = (A − λEH )f , where f ∈ D(A). Let f = (Uλ0(A) − EH )h, where h ∈ Mλ0(A). Then Af = (λ0Uλ0(A) − λ0EH)h. Substituting expressions for f and Af in the expression for ψ − K(λ)ψ we get ψ − K(λ)ψ = (λ0Uλ0(A) − λ0EH )h − λ(Uλ0(A) − EH )h. Therefore ψ + (λ − λ0)Uλ0(A)h = K(λ)ψ + (λ − λ0)h. Using the orthogonality of the summands in the left-hand side and also in the right-hand side we get kψk2 H + λ − λ02kUλ0(A)hk2 H = kK(λ)ψk2 H + λ − λ02khk2 H. Therefore H = δλ,λ0khk2 H, δλ,λ0 := λ − λ02 − λ − λ02. It remains to notice that H − kK(λ)ψk2 kψk2 δλ,λ0 = Re λ − Re λ02 + Im λ − Im λ02 − Re λ − Re λ02 − Im λ − Im λ02 = Im λ + Im λ02 − Im λ − Im λ02 > 0, (108) since Im λ and Im λ0 have the same sign. 67 Let us find a relation between the function K(λ) and the characteristic function C(λ). Choose an arbitrary λ ∈ Πλ0 and an element ψ ∈ Nλ0. By the Neumann formula (61) we may write: AC(λ),λ0(−C(λ)ψ + ψ) = −λ0C(λ)ψ + λ0ψ. On the other hand, by (105),(103) and (104) we get AC(λ),λ0(−C(λ)ψ + ψ) = Aλ(−C(λ)ψ + ψ) = Aλ(f + ψ1) = Af + λψ1, where f ∈ D(A) and ψ1 ∈ Nλ(A) are some elements: Therefore −C(λ)ψ + ψ = f + ψ1 ∈ D(Aλ). −λ0C(λ)ψ + λ0ψ = Af + λψ1. Subtracting fron th last equality the previpus one, multiplied by λ, we obtain that (λ0 − λ)ψ − (λ0 − λ)C(λ)ψ = (A − λEH)f. Thus, the following inclusion holds: (λ − λ0)ψ − (λ − λ0)C(λ)ψ ∈ Mλ(A). (109) Divide the both sides of the last equality by λ − λ0 and apply the operator Pλ,λ0: λ − λ0 λ − λ0 Pλ,λ0ψ − C(λ)ψ = 0, i.e. C(λ)ψ = λ−λ0 λ−λ0 Pλ,λ0ψ. Therefore C(λ) = λ − λ0 λ − λ0 K(λ), λ ∈ Πλ0. Since K(λ) is analytic in Πλ0 and non-expanding, then the characteristic function C(λ) is analytic and kC(λ)k ≤(cid:12)(cid:12)(cid:12)(cid:12) λ − λ0 λ − λ0(cid:12)(cid:12)(cid:12)(cid:12) , λ ∈ Πλ0. (110) Now we shall obtain another formula for the characteristic function C(λ). Let λ ∈ Πλ0 be an arbitrary number. Let us check that elements ψ ∈ Nλ0(A) and bψ ∈ Nλ0(A) are connected by the following relation (λ − λ0)ψ − (λ − λ0)bψ ∈ Mλ(A), (111) 68 if and only if The sufficiency of this assertion follows from relation (109). Suppose that relation (111) holds. Apply the operator Pλ,λ0 to the both sides of this equality: bψ = C(λ)ψ. and therefore what we needed to prove. (λ − λ0)Pλ,λ0ψ − (λ − λ0)bψ = 0, bψ = Pλ,λ0ψ = C(λ)ψ, λ − λ0 λ − λ0 3.8 A connection of the operator-valued function F(λ) with the characteristic function. As it was done above, the space H we represent as (64), where He = Consider an operator A of the form (65), where Ae is a symmetric operator Consider a closed symmetric operator A in a Hilbert space H. Let eA be a self- adjoint extension of A acting in a Hilbert space eH ⊇ H. Choose and fix an arbitrary number λ0 from Re. For the extension eA there correspond operator- valued functions Bλ = Bλ(A, eA) and F(λ) = F(λ; λ0, A, eA) (see (94),(100)). eH ⊖ H, and the operator A can be identified with the operator A ⊕ oHe. in a Hilbert space He, such that A ⊕ Ae ⊆ eA. In particular, we can take as Ae the operator oHe. Thus, the operator eA is a self-adjoint extension of the By the generalized Neumann formulas, for the self-adjoint extension eA of A there corresponds an isometric operator T with the domain D(T ) = Nλ0(A) and the range R(T ) = Nλ0(A), haning block representation (69), where T11 = PNλ0 T22 = PNλ0 Denote Ce(λ) := C(λ; λ0, Ae), i.e. Ce(λ) is a characteristic function of the operator Ae in He. (Ae)T PNλ0 (Ae). operator A. (A)T PNλ0 (A), T12 = PNλ0 (A)T PNλ0 (Ae), T21 = PNλ0 (Ae)T PNλ0(A), By (60) the domain of eA consists of elements ef of the following form: ef = f1 + f2 + T (ψ1 + ψ2) − ψ1 − ψ2 = [f1 + T11ψ1 + T12ψ2 − ψ1] + [f2 + T21ψ1 + T22ψ2 − ψ2], (112) where f1 ∈ D(A), f2 ∈ D(Ae), ψ1 ∈ Nλ0(A), ψ2 ∈ Nλ0(Ae). Notice that an expression in the first square brackets belongs to H, and in the second square brackets belongs to He. Formula (61) gives the following relation for 69 (113) Here also an expression in the first square brackets belongs to H, and in the second square brackets belongs to He. Recall that the manifold Lλ (λ ∈ C) the operator eA: eAef = [Af1 + λ0(T11ψ1 + T12ψ2) − λ0ψ1] + [Aef2 + λ0(T21ψ1 + T22ψ2) − λ0ψ2]. consists of those elements ef ∈ D(eA), for which (eA − λE eH)ef belongs to H. the element ef of form (112) belongs to Lλ if and only if Subtracting from relation (113) relation (112), multiplied by λ, we see that (Ae − λEHe)f2 + (λ0 − λ)(T21ψ1 + T22ψ2) − (λ0 − λ)ψ2 = 0. (114) Suppose now that λ ∈ Πλ0. If relation (114) is satisfied then by the property of the characteristic function, see the text with formula (111), we get Moreover, by (114) we get ψ2 = Ce(λ)(T21ψ1 + T22ψ2). (115) (116) f2 = (Ae − λEHe)−1(cid:0)(λ − λ0)(T21ψ1 + T22ψ2) − (λ − λ0)ψ2(cid:1) . Since λ and λ0 lie in the same half-plane, then the number δλ,λ0 = λ − λ02 − λ − λ02 is positive, see (108). By the property of the characteristic function (110) we obtain that kCe(λ)k ≤(cid:12)(cid:12)(cid:12)(cid:12) λ − λ0 λ − λ0(cid:12)(cid:12)(cid:12)(cid:12) < 1, λ ∈ Πλ0. The operator Ce(λ)T22 may be considered as an operator in a Hilbert space Nλ0(Ae), defined on the whole Nλ0(Ae). Since T22 is non-expanding, then Ce(λ)T22 is contractive and there exists the inverse (ENλ0 (Ae) − Ce(λ)T22)−1, defined on the whole Nλ0(Ae). By (115) we get ψ2 = (ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21ψ1. (117) Conversely, if elements f1 ∈ D(A) and ψ1 ∈ Nλ0(A) we choose arbitrarily, the element ψ2 we define by (117), and f2 we define by (116) (this is correct since (117) implies (115)) and therefore(cid:0)(λ − λ0)(T21ψ1 + T22ψ2) − (λ − λ0)ψ2(cid:1) belongs to Mλ(Ae)), then the element ef of form (112) belongs to Lλ. Thus, the set Lλ (λ ∈ Πλ0) consists of elements ef of form (112), where f1 ∈ D(A) and ψ1 ∈ Nλ0(A) are arbitrary, ψ2 and f2 are defined by (117),(116). 70 Since the domain of Bλ is P D(Bλ) consists of elements of form: eH H Lλ, and its action is defined by (94), then f = f1 + T11ψ1 + T12ψ2 − ψ1, where elements f1 ∈ D(A) and ψ1 ∈ Nλ0(A) are arbitrary, and ψ2 is defined by (117). Moreover, we have Bλf = Af1 + λ0(T11ψ1 + T12ψ2) − λ0ψ1. Substituting an expression for ψ2 from (117) in these expressions we obtain that D(Bλ) consists of elements of the following form: (118) f = f1 +(cid:16)T11 + T12(ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21(cid:17) ψ1 − ψ1, where elements f1 ∈ D(A) and ψ1 ∈ Nλ0(A) are arbitrary. The action of the operator Bλ has the following form: Bλf = Af1 + λ0(cid:16)T11 + T12(ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21(cid:17) − λ0ψ1. (119) expression for F(λ) = F(λ; λ0, A, eA). Since F(λ) has the following form Using the obtained expression for the operator Bλ (λ ∈ Πλ0), we find (see (100)): F(λ) = ENλ0 (A) + (λ0 − λ0) (Bλ − λ0EH )−1(cid:12)(cid:12)Nλ0 (A) , then we need to define the action of the operator (Bλ − λ0EH)−1 on the set Nλ0(A). By (119), for an arbitrary element f ∈ D(Bλ) of the form (118) we may write: (120) (Bλ − λ0EH )f = (A − λ0EH)f1 + (λ0 − λ0)ψ1. Applying the operator (Bλ − λ0EH )−1 to the both sides of the latter equality we get (Bλ − λ0EH)−1(cid:0)(A − λ0EH)f1 + (λ0 − λ0)ψ1(cid:1) = f = f1 +(cid:16)T11 + T12(ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21(cid:17) ψ1 − ψ1. In particular, choosing f1 = 0 we get (Bλ − λ0EH)−1ψ1 = 1 λ0 − λ0(cid:16)T11 + T12(ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21(cid:17) ψ1 1 − λ0 − λ0 ψ1, ψ1 ∈ Nλ0(A). Substituting the obtained relation for the operator (Bλ − λ0EH)−1Nλ0(A) into (120), we obtain that 71 F(λ; λ0, A, eA) = T11 + T12(ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21, (121) This representation will play an important role for the analytic description of the generalized resolvents of a symmetric operator A. λ ∈ Πλ0. 3.9 Boundary properties of the operator-valued func- tion F(λ): the case of a densely defined symmetric operator A. Consider an arbitrary closed symmetric operator A in a Hilbert space H, (122) eH H h, eH P eH (Af, P H eAh = A∗P having a dense domain: D(A) = H. Let eA be a self-adjoint extension of the operator A, acting in a Hilbert space eH ⊇ H. Let us check that h ∈ D(eA). H D(eA) ⊆ D(A∗), P In fact, for arbitrary elements h ∈ D(eA) and f ∈ D(A) we write: H h)H = (Af, h) eH = (eAf, h) eH = (f, eAh) eH = (f, P h ∈ D(eA). Consider the sets eLλ, Lλ and L∞, defined by equalities (88)-(90) for λ ∈ C. By the definition of eLλ, for an arbitrary element g from eLλ it holds: eAh = A∗P (eA − λE eH )g = P and the required relation follows. Using (122) we may write: eH H h + P eH⊖HeAh, eH⊖HeAg − λP H eAh)H, eH eH⊖H eH eH⊖H 0 = P eH eH eH g, i.e. (123) eH λ ∈ C. (126) Therefore It follows that eH P eH eH⊖Hg, eH H g + λP eH⊖HeAg = λP eAg = A∗P (eA − λE eH)g = (A∗ − λEH)P eH eH⊖H g, λ ∈ C. λ ∈ C. g ∈eLλ, g ∈eLλ, g ∈eLλ, eH H g, Proposition 3.7 Let A be a closed symmetric operator in a Hilbert space H, D(A) = H, and eA be a self-adjoint extension of A, acting in a Hilbert (124) (125) 72 space eH ⊇ H. Let λ0 ∈ Re be an arbitrary number. For arbitrary complex numbers λ1 and λ2, λ1 6= λ2, and arbitrary elements h1 ∈eLλ1, h2 ∈eLλ2, set eH H hk = fk + ϕk + ψk, k = 1, 2, gk := P (127) where fk ∈ D(A), ϕk ∈ Nλ0(A), ψk ∈ Nλ0(A). Then (h1, h2) eH = (g1, g2)H + λ0 − λ0 λ1 − λ2 ((ψ1, ψ2)H − (ϕ1, ϕ2)H) . (128) Proof. Notice that 1 λ1 − λ2h(h1, (eA − λ2E eH)h2) eH − ((eA − λ1E eH)h1, h2) eHi λ1 − λ2h(h1, eAh2) eH − λ2(h1, h2) eH − (eAh1, h2) eH + λ1(h1, h2) eHi = (h1, h2) eH . 1 = Using relation (126) to a transformation of the left-hand side of the last expression we get (h1, h2) eH = 1 λ1 − λ2h(h1, (A∗ − λ2EH)P eH H h2) eH − ((A∗ − λ1EH)P eH H h1, h2) eHi = 1 λ1 − λ2 [(g1, (A∗ − λ2EH)g2)H − ((A∗ − λ1EH)g1, g2)H ] = (g1, g2)H + 1 λ1 − λ2 [(g1, A∗g2)H − (A∗g1, g2)H] . Substituting here the expression for gk from (127), and taking into account that A∗ϕk = λ0ϕk, A∗ψk = λ0ψk, k = 1, 2, we obtain the required equal- ity (128). ✷ Consider the operator-valued function Bλ = Bλ(A, eA), λ ∈ Re, and the operator B∞ = B∞(A, eA) (see (93), (94), (95)). As it was said above, the operators Bλ (λ ∈ Re) and B∞ are extensions of the operator A in H. Therefore B∞ = A∗L∞, Bλ = A∗Lλ, λ ∈ Re. (129) Choose and fix an arbitrary number λ0 ∈ Re. Consider the operator- valued function F(λ) = F(λ; λ0, A, eA), λ ∈ Πλ0, defined by (100). Choose arbitrary two points λ′, λ′′ ∈ Πλ0, and arbitrary elements h′ ∈ eLλ′, h′′ ∈ 73 eH H h′′ belong to D(Bλ′) and D(Bλ′′), respectively. By (101) and the generalized Neumann formulas we may write: eH H h′ and P eLλ′′. Then the elements P P eH H h′ = f ′ + F(λ′)ψ′ − ψ′, P eH H h′′ = f ′′ + F(λ′′)ψ′′ − ψ′′, where f ′, f ′′ ∈ D(A), ψ′, ψ′′ ∈ Nλ0(A). By Proposition 3.7 we get (h′, h′′) eH = (P eH H h′, P eH H h′′)H + λ0 − λ0 λ′ − λ′′ On the other hand, we may write: ((ψ′, ψ′′)H − (F(λ′)ψ′, F(λ′′)ψ′′)H) . (h′, h′′) eH = (P eH H h′ + P eH eH⊖Hh′, P eH H h′′ + P eH eH⊖Hh′′) eH = (P eH H h′, P eH H h′′)H + (P eH eH⊖Hh′, P eH eH⊖Hh′′)H. Therefore (P eH eH⊖Hh′, P eH eH⊖Hh′′)H = λ0 − λ0 λ′ − λ′′ ((ψ′, ψ′′)H − (F(λ′)ψ′, F(λ′′)ψ′′)H) . Multiplying the both sides of the last equality by λ′λ′′ and taking into ac- count (124) we get (P eH eH⊖HeAh′, P eH eH⊖HeAh′′)H = (λ0−λ0) (130) Equality (130) will be used in the sequel. Also we shall need the following proposition. λ′λ′′ λ′ − λ′′ ((ψ′, ψ′′)H − (F(λ′)ψ′, F(λ′′)ψ′′)H ) . Proposition 3.8 If a sequence of elements of a Hilbert space is bounded and all its weakly convergent subsequences have the same weak limit, then the sequence converges weakly. Let H = {hn}∞ n=1, hn ∈ H, be a bounded sequence of elements Proof. of a Hilbert space H. Suppose to the contrary that all weakly convergent subsequences of H converge to an element h ∈ H, but the sequence H does In this case there exists an element g ∈ H such that not converge to h. (hn, g)H does not converge to (h, g)H, as n → ∞. Therefore there exists a number ε > 0, and a subsequence {hnk}∞ k=1, such that (hnk, g)H − (h, g)H = (hnk − h, g)H ≥ ε, k ∈ N. (131) Since the sequence {hnk}∞ k=1 is bounded, it contains a weakly convergent subsequence. This subsequence, by assumption, should converge weakly to to an element h. But this is impossible according to (131). The obtained contradiction completes the proof. ✷ 74 Let continue our considerations started before the statement of the last ily closed) symmetric extension of the operator A in H. By the Neumann proposition. Recall that the operator B∞ = B∞(A, eA) is a (not necessar- formulas for it there corresponds an isometric operator Φ∞ = Φ∞(λ0; A, eA) with the domain D(Φ∞) ⊆ Nλ0(A) and the range R(Φ∞) ⊆ Nλ0(A). The operator Φ∞ was already defined for the general case of a not necessarily densely defined operator A (below (101)). Theorem 3.9 Let A be a closed symmetric operator in a Hilbert space H, D(A) = H, and eA be a self-adjoint extension of A, acting in a Hilbert space eH ⊇ H. Let λ0 ∈ Re be an arbitrary number, {λn}∞ n=1, λn ∈ Πλ0, be a n=1, ψn ∈ Nλ0(A), be a sequence, weakly number sequence, tending to ∞, {ψn}∞ converging to an element ψ, and sup n∈N(cid:20) λn2 Im λn(cid:16)kψnkH − kF(λn; λ0, A, eA)ψnkH(cid:17)(cid:21) < ∞. (132) Then ψ ∈ D(Φ∞(λ0; A, eA)), and the sequence {F(λn)ψn}∞ to an element Φ∞ψ. Moreover, if the sequence {ψn}∞ strong sence, then n=1 converges weakly n=1 converges to ψ in a lim n→∞ F(λn)ψn = lim n→∞ F(λn)ψ = Φ∞ψ. (133) Proof. Since elements λn belong to Πλ0, then by relation (101) we , for every n ∈ N. Since ψn ∈ Nλ0(A) = D(F(λn)), then by the generalized Neumann formulas elements F(λn)ψn − get Bλn(A, eA) = AF(λn;λ0,A, eA),λ0 ψn ∈ D(B(λn)) = Lλn, n ∈ N. Consider a sequence of elements gn ∈ eLλn, n ∈ N, such that P eH H gn = F(λn)ψn − ψn, n ∈ N. By (123) we may write: eH eH⊖HeAgn n ∈ N. = λ0F(λn)ψn − λ0ψn + P eH eAgn = A∗(F(λn)ψn − ψn) + P eH⊖HeAgn, H eAgn = λ0F(λn)ψn − λ0ψn, eH Therefore P n ∈ N. (134) 75 n ∈ N, Observe that eH kP H eAgnkH ≤ λ0kF(λn)kkψnkH + λ0kψnkH ≤ K, where K is a constant. Here we used the fact that F(λ) is non-expanding and a weakly convergent sequence is bounded. Let us use formula (130), with λ′ = λ′′ = λn, h′ = h′′ = gn: λn2 Im λn ((ψn, ψn)H − (F(λn)ψn, F(λn)ψn)H) (P eH eH⊖HeAgn, P eH eH⊖HeAgn)H = Im λ0 λn2 Im λn = Im λ0 (kψnkH − kF(λn)ψnkH) (kψnkH + kF(λn)ψnkH) ≤ Im λ0 λn2 Im λn (kψnkH − kF(λn)ψnkH) 2kψnkH. eH equality holds: From (132) it follows that kP eH⊖HeAgnkH ≤ K2, where K2 is a constant. Thus, we conclude that the sequence {eAgn}∞ eH⊖HeAgn = P n=1 is bounded. By (124) the following and therefore eH eH⊖H gn, n ∈ N, 1 λn P eH P eH eH⊖Hgn → 0, n → ∞. (135) Since a bounded set in a Hilbert space is weakly compact, then from the n=1 we can select a weakly convergent subsequence. Let k=1 be an arbitrary subsequence which converges weakly to an ele- sequence {eAgn}∞ {eAgnk}∞ ment h ∈ H: eH H eAgnk ⇀ P Then P eH H h, as k → ∞. From (134) it follows that eAgnk ⇀ h, k → ∞. (136) F(λnk)ψnk ⇀ 1 λ0 P eH H h + λ0 λ0 ψ, k → ∞. (137) Denote ϕ := 1 ψ. By the weak completeness of a Hilbert space the λ0 element ϕ belongs to Nλ0(A). By the definition of elements gn we may write: eH H h + λ0 P λ0 gn = F(λn)ψn − ψn + P eH eH⊖Hgn, n ∈ N. By (135),(137) we get gnk ⇀ ϕ − ψ, k → ∞. 76 and Since eA is closed, and the weak closeness of an operator is equivalent to its closeness, then from (136) and the last relation it follows that ϕ − ψ ∈ D(eA) Since elements ϕ and ψ belong to the space H, then ϕ − ψ ∈ D(eA) ∩ H = L∞ = D(B∞). As it was mentioned before the statement of the theorem, by the Neumann formulas for the operator B∞ there corresponds an isometric operator Φ∞. By (60) the following equality holds: eA(ϕ − ψ) = h. (138) D(B∞) = D(A) ∔ (Φ∞ − EH )D(Φ∞). Therefore ϕ − ψ = f + Φ∞u − u, f ∈ D(A), u ∈ D(Φ∞). Since ϕ ∈ Nλ0(A), ψ ∈ Nλ0(A), by the linear independence of the manifolds D(A), Nλ0(A) and Nλ0(A), we obtain that f = 0, u = ψ Φ∞u = ϕ. Thus, the element ψ belongs to D(Φ∞) and Φ∞ψ = ϕ. Therefore relation (138) takes the following form: Then relation (136) becomes eA(Φ∞ψ − ψ) = h. eAgnk ⇀ eA(Φ∞ψ − ψ), k → ∞. n=1 has the same limit. By Proposition 3.8 we conclude that the sequence Thus, an arbitrary weakly convergent subsequence of a sequence {eAgn}∞ n=1 is a weakly convergent and its limit is equal to eA(Φ∞ψ − ψ). Then {eAgn}∞ n → ∞. P eH By relation (123) we may write: eH H eAgn ⇀ P H eA(Φ∞ψ − ψ), P eH H eA(Φ∞ψ − ψ) = A∗P By (134) and two last relations we get eH H (Φ∞ψ − ψ) = λ0Φ∞ψ − λ0ψ. λ0F(λn)ψn − λ0ψn ⇀ λ0Φ∞ψ − λ0ψ, n → ∞, i.e. F(λn)ψn ⇀ Φ∞ψ, n → ∞, Thus, the first assertion of the theorem is proved. Suppose that the sequence {ψn}∞ n=1 converges strongle to the element ψ. It is known that if a sequence F = {fn}∞ n=1, fn ∈ H, of a Hilbert space H converges weakly to an element f then kf kH ≤ limn→∞kfnkH. (139) In fact, we may write kf k2 H = lim n→∞ (fn, f )H. 77 If {kfnkkH}∞ k=1 is a convergent subsequence of a sequence {kfnkH}∞ n=1, then kf k2 H = lim k→∞ (fnk, f )H ≤ lim k→∞ kfnkkHkf kH. From the last relation it follows the required estimate (139). Applying estimate (139) to the sequence {F(λn)ψn}∞ n=1 we obtain an es- timate: kΦ∞ψkH ≤ limn→∞kF(λn)ψnkH. On the other hand, since F(λ) is non-expanding, then limn→∞kF(λn)ψnkH ≤ lim n→∞ kψnkH = kψkH = kΦ∞ψkH. Therefore limn→∞ kF(λn)ψnkH = kΦ∞ψkH, and a sequence {F(λn)ψn}∞ converges strongly to an element Φ∞ψ. It remains to notice that n=1 kF(λn)ψ − Φ∞ψkH = kF(λn)ψ − F(λn)ψn + F(λn)ψn − Φ∞ψkH ≤ kF(λn)ψ − F(λn)ψnkH + kF(λn)ψn − Φ∞ψkH ≤ kψ − ψnkH + kF(λn)ψn − Φ∞ψkH → 0, n → ∞. Thus, relation (133) is proved. ✷ Corollary 3.5 In conditions of Theorem 3.9, if one additionally assumes that the sequence {λn}∞ n=1 belongs to a set Πε λ0 = {λ ∈ Πλ0 : ε < arg λ < π − ε}, 0 < ε < π 2 , then condition (132) can be replaced by the following condition: sup n∈Nhλn(cid:16)kψnkH − kF(λn; λ0, A, eA)ψnk(cid:17)i < ∞. n=1 belonfs to the set Πε In fact, if the sequence {λn}∞ (140) λ0, then Proof. λn Im λn ≤ M, n ∈ N, where M is a constant, hot depending of n. Therefore from (140) it fol- lows (132). ✷ Notice that from relation (91) it follows that where Rλ(A) is the generalized resolvent of A, corresponding to the self- eLλ = (eA − λE eH)−1H, Lλ = Rλ(A)H, λ ∈ Re, (141) adjoint extension eA. 78 Proposition 3.9 Let A be a closed symmetric operator in a Hilbert space H, and eA be a self-adjoint extension of A, acting in a Hilbert space eH ⊇ H. Let Rλ and Rλ, λ ∈ Re, denote the resolvent of eA and the generalized resolvent of A, corresponding to the extension eA, respectively, and λ0 ∈ Re, 0 < ε < π be arbitrary numbers. Then the following relations hold: 2 λ∈Πε , λ→∞ (−λRλh) = h, λ∈Πε , λ→∞ (−λRλg) = g, lim λ0 lim λ0 h ∈ eH; g ∈ H. Proof. Choose an arbitrary sequence {λn}∞ n=1, λn ∈ Πε λ0, tending to ∞, and an arbitrary element h ∈ eH. Let us check that λnRλnh → −h, n → ∞. (142) Let λn = σn + iτn, σn, τn ∈ R, n ∈ N. Notice that since numbers λn belong to the set Πε λ0, it follows an estimate: (cid:12)(cid:12)(cid:12)(cid:12) σn τn(cid:12)(cid:12)(cid:12)(cid:12) < M1, n ∈ N, (143) where M1 is a constant, not depending on n. Using the functional equation for the resolvent we write Rλn − Rλ0 = (λn − λ0)RλnRλ0, λnRλnRλ0 = −Rλ0 + Rλn + λ0RλnRλ0, n ∈ N. (144) Since λn = τn(cid:12)(cid:12)(cid:12)(cid:12)i + σn τn(cid:12)(cid:12)(cid:12)(cid:12) ≤ τn(cid:18)1 +(cid:12)(cid:12)(cid:12)(cid:12) σn τn(cid:12)(cid:12)(cid:12)(cid:12)(cid:19) ≤ τnM1, then τn → ∞ n → ∞. On the other hand, by the functional equation for the resolvent we may write: kRλf k2 H = (R∗ λRλf, f )H = (RλRλf, f )H = 1 λ − λ ((Rλ − Rλ)f, f )H 1 = λ − λ Therefore (Rλf, f )H − (R∗ λf, f )H) = 1 Im λ Im(Rλf, f )H, λ ∈ Re, f ∈ eH. kRλnf k2 H ≤ 1 τn (Rλnf, f )H ≤ 1 τn kRλnf kHkf kH; kRλnf kH ≤ 1 τn kf kH, kRλnk ≤ 1 τn , n ∈ N, f ∈ eH; n ∈ N. It follows that By (144) we get kRλnkH → 0, n → ∞. u. − lim n→∞ λnRλnRλ0 = −Rλ0. In particular, we obtain that λnRλnu = −u, By (145) and (143) the following inequality holds: u ∈ Rλ0eH = D(eA). τn(cid:12)(cid:12)(cid:12)(cid:12) + 1 < M1 + 1, σn (cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12) kλnRλnk ≤(cid:12)(cid:12)(cid:12)(cid:12) σn + iτn τn 79 (145) n ∈ N. Since the sequence of linear operators {λnRλn}∞ n=1 converges (in the strong operator topology) on a dense set and the norms of the operators are uni- formly bounded, by the Banach-Steinhaus Theorem the sequence {λnRλn}∞ Theorem 3.10 Let A be a closed symmetric operator in a Hilbert space H, this limit operator coincides with −E eH . Thus, relation (142) is proved, and the required relations in the statement of the theorem follow. ✷ converges on the whole eH to a continuous linear operator. By the continuity, D(A) = H, and eA be a self-adjoint extension of A, acting in a Hilbert space eH ⊇ H. Let λ0 ∈ Re be an arbitrary number, ψ be an arbitrary number from D(Φ∞(λ0; A, eA)), and a vector-valued function ψ(λ), λ ∈ Πλ0, with values in Nλ0(A), is defined by the following formula: n=1 Nλ0 (A)(A∗ − λ0EH)Rλ(ψ − Φ∞ψ), P H (146) ψ(λ) = λ λ0 − λ0 where Rλ is a generalized resolvent of A, corresponding to the self-adjoint extension eA. Then , λ→∞ λ∈Πε lim λ0 ψ(λ) = ψ, and there exists a finite limit , λ,ζ→∞(cid:26) λζ lim λ0 λ − ζ λ,ζ∈Πε Here 0 < ε < π 2 . λ∈Πε , λ→∞ lim λ0 F(λ; λ0, A, eA)ψ(λ) = Φ∞ψ, [(ψ(λ), ψ(ζ))H − (F(λ)ψ(λ), F(ζ)ψ(ζ))H](cid:27) . (147) (148) 80 Proof. Denote g = Φ∞(λ0; A, eA)ψ − ψ ∈ D(B∞) = L∞ = D(eA) ∩ H, and by Rλ and Rλ, λ ∈ Re, we denote the resolvent of eA and the generalized resolvent of A, corresponding to the extension eA, respectively. Consider the following two vector-valued functions: g(λ) = P λ ∈ Πλ0. eg(λ) = −λRλg, 1 λ0 − λ0 Then the function ψ(λ) from the statement of the theorem takes the following form: ψ(λ) = Nλ0 (A)(A∗ − λ0EH )g(λ), P H λ ∈ Πλ0. Notice that by (141), the element g(λ) belongs to Lλ, and therefore by (129) the definition of the function ψ(λ) is correct. By Proposition 3.9 we get: eH Heg(λ) = −λRλg, lim λ0 λ∈Πε , λ→∞ g(λ) = g, 0 < ε < π 2 . (149) λ∈Πε lim λ0 , λ→∞eg(λ) = g, λ∈Πε , λ→∞ lim λ0 Moreover, by Proposition 3.9 with h = eAg we obtain that , λ→∞ eAeg(λ). (−λeARλg) = eAg = By (141) it follows that the element eg(λ) belongs to eLλ ⊆ D(eA). By (122) (−λRλeAg) = we write lim λ0 lim λ0 (150) , λ→∞ λ∈Πε λ∈Πε eH g(λ) = P λ ∈ Πλ0. From the last expression and (150) we obtain that Heg(λ) ∈ D(A∗), P H eAeg(λ) = lim λ0 , λ→∞ P eH λ∈Πε P eH H eAg = eH H eAeg(λ) = A∗g(λ), or, taking into account property (122), we write: lim λ0 λ∈Πε , λ→∞ A∗g(λ), λ ∈ Πλ0, lim λ0 λ∈Πε , λ→∞ A∗g(λ) = A∗g, λ ∈ Πλ0. (151) Passing to the limit in the above expression for the function ψ(λ), using (149) and (151), we get: lim λ0 λ∈Πε , λ→∞ ψ(λ) = 1 λ0 − λ0 Nλ0 (A)(A∗ − λ0EH)g. P H By the definition of the element g the following equality holds: (A∗−λ0EH )g = (λ0 − λ0)ψ. Substituting it into the previous relation we obtain the first re- lation in (147). 81 As it was already noticed, the element g(λ) belongs to Lλ = D(Bλ(A, eA)), and therefore by the definition of the function F(λ; λ0, A, eA) and the gener- alized Neumann formulas we write: g(λ) = f (λ) + F(λ)ψ1(λ) − ψ1(λ), λ ∈ Πλ0, where f (λ) ∈ D(A), ψ1(λ) ∈ Nλ0(A). Using the above expression for ψ(λ) we directly calculate: ψ(λ) = 1 λ0 − λ0 Nλ0 (A)(A∗ − λ0EH)g(λ) = ψ1(λ), P H λ ∈ Πλ0. Therefore Observe that g(λ) = f (λ) + F(λ)ψ(λ) − ψ(λ), λ ∈ Πλ0. 1 λ0 − λ0 P H Nλ0 (A∗ − λ0EH)g(λ) = F(λ)ψ(λ), λ ∈ Πλ0. Consequently, by (149),(151) and the definition of g, we get: lim λ0 λ∈Πε , λ→∞ F(λ)ψ(λ) = 1 λ0 − λ0 lim λ0 λ∈Πε , λ→∞ P H Nλ0 (A∗ − λ0EH)g(λ) = 1 λ0 − λ0 P H Nλ0 (A∗g − λ0g) = Φ∞ψ, i.e. the second relation in (147) is satisfied, as well. P eH H h′ = g(λ), P Let us use relation (130) for λ′ = λ, λ′′ = ζ, h′ = eg(λ), h′′ = eg(ζ), eH H h′′ = g(ζ): ((ψ(λ), ψ(ζ))H (P eH eH⊖HeAeg(λ), P eH eH⊖HeAeg(ζ))H = (λ0 − λ0) −(F(λ)ψ(λ), F(ζ)ψ(ζ))H). λ − ζ λζ By relation (150), the left-hand side has a finite limit as λ, ζ → ∞, λ, ζ ∈ Πε and the last assertion of the theorem follows. ✷ λ0, Proposition 3.10 Let A be a closed symmetric operator in a Hilbert space symmetric operators. Let λ0 ∈ Re be an arbitrary number, ψ be an arbitrary H, D(A) = H, and eA be a self-adjoint extension of A, acting in a Hilbert space eH ⊇ H. Suppose that the operator A is a direct sum of two maximal element from D(Φ∞(λ0; A, eA)). Then there exists a finite limit λ − ζh(ψ, ψ)H − (F(λ; λ0, A, eA)ψ, F(ζ; λ0, A, eA)ψ)Hi(cid:27) . , λ,ζ→∞(cid:26) λζ lim λ0 (152) λ,ζ∈Πε Here 0 < ε < π 2 . 82 The operator A has the following form: A = A1 ⊕ A2, where Proof. Ak is a maximal symmetric operator in a Hilbert space Hk, k = 1, 2, and H = H1 ⊕ H2. Suppose at first that the domain of the regularity of A2 is a half-plane Πλ0, and the domain of the regularity of A1 is Π−λ0. Recalling a similar situation in (67) we may write: Nλ(A) = Nλ(A1), Nλ(A) = Nλ(A2), λ ∈ Πλ0. Calculate the function ψ(λ), defined in (146). Let ψ′ be an arbitrary element from Nλ0(A), and λ ∈ Πλ0. From (141) it follows that the element Rλψ′ Using (97) we write: belongs to the manifold Lλ = D(Bλ(A, eA)), which by (129) lies in D(A∗). (A∗ − λEH )Rλψ′ = (Bλ − λEH)Rλψ′ = (Bλ − λEH)(Bλ − λEH )−1ψ′ = ψ′. Conversely, g = (Bλ − λEH)−1ψ′ = Rλψ′. if an element g belongs to Lλ and (A∗ − λEH)g = ψ′, then and the generalized Neumann formulas we write: Since Rλψ′ ∈ Lλ = D(Bλ), then by definition of the function F(λ; λ0, A, eA) Rλψ′ = f + F(λ)ψ1 − ψ1, where f ∈ D(A), ψ1 ∈ Nλ0(A). On the other hand, consider an element g′ = λ − λ0 λ − λ0 (A2 − λEH2)−1F(λ)ψ′ + (F(λ) − EH ) 1 λ − λ0 ψ′. Observe that F(λ)ψ ∈ Nλ0(A) = Nλ(A2), and the point λ belongs to the domain of the regularity of A2, and therefore the first summand in the right- hand side is defined correctly. By the generalized Neumann formulas the element g′ belongs to D(Bλ) = Lλ. Moreover, we have: (A∗ − λEH)g′ = λ − λ0 λ − λ0 (A − λEH)(A2 − λEH2)−1F(λ)ψ′ +(λ0 − λ)F(λ) 1 λ − λ0 ψ′ − (λ0 − λ) 1 λ − λ0 ψ′ = ψ′. By the above conciderations we conclude that Rλψ′ = g′. Notice that Nλ0 (A)(A∗ − λ0EH )Rλψ′ = P H P H Nλ0 (A)(A∗ − λ0EH)g′ = λ0 − λ0 λ − λ0 ψ′. (153) Consider an arbitrary element ϕ ∈ Nλ0(A) = Nλ0(A2) ⊆ H2. Since λ belongs to the domain of the regularity of A2, we may write: Rλϕ = P eH H (eA − λEH)−1ϕ = (A2 − λEH)−1ϕ. 83 It follows that Nλ0 (A)(A∗ − λ0EH)Rλϕ = P H P H Nλ0 (A)(A∗ − λ0EH )(A2 − λEH)−1ϕ = P H Nλ0 (A)(A − λ0EH)(A − λEH)−1ϕ = 0. (154) Let ψ be an arbitrary element from D(Φ∞(λ0; A, eA)) ⊆ Nλ0(A), and λ ∈ Πλ0. Using (153) with ψ′ = ψ and (154), calculate the function ψ(λ) from (146): ψ(λ) = λ λ0 − λ0 Nλ0 (A)(A∗ − λ0EH )Rλ(ψ − Φ∞ψ) = P H λ λ − λ0 ψ, λ ∈ Πλ0. By Theorem 3.10 we conclude that there exists a limit (152). In the case when the domain of the regularity of A2 is a half-plane Π−λ0, and the domain of the regularity of A1 is Πλ0, we can reassign the operators A1 and A2. In the case when the domain of the regularity of the operators A1 and A2 is the same half-plane (Πλ0 or Π−λ0), the operator A is itself a maximal symmetric operator. In this case we set H1 = H, H2 = {0}, A1 = A and A2 = 0H2. For the self-adjoint operator A2 the both half-planes Πλ0 and Π−λ0 are the domains of the regularity and we can apply the proved part of the theorem. ✷ Theorem 3.11 Let λ0 ∈ Re be an arbitrary number and F (λ) be an an- alytic in the half-plane Πλ0 operator-valued function, which values are lin- ear non-expanding operators with the domain D(F (λ)) = N1 and the range R(F (λ)) ⊆ N2, where N1 and N2 are some Hilbert spaces. Let {λn}∞ n=1, λn ∈ Πλ0, be a sequence of numbers, converging to ∞, {gn}∞ n=1, gn ∈ N1, be a sequence of elements, weakly converging to an element g, and sup Im λn n∈N(cid:20) λn2 , λ,ζ→∞(cid:26) λζ (kgnkN1 − kF (λn)gnkN2)(cid:21) < ∞. [(g, g)N1 − (F (λ)g, F (ζ)g)N2](cid:27) . Then there exists a finite limit λ,ζ∈Πε lim λ0 λ − ζ (155) (156) Here 0 < ε < π 2 . Proof. Consider the operator A of the following form: A = A1 ⊕ A2, where Ak is a maximal densely defined symmetric operator in a Hilbert space Hk, 84 k = 1, 2, and define H = H1 ⊕ H2. Choose the operators A1 and A2 such that the domain of the regularity of the operator A2 will be Πλ0, the domain of the regularity of A1 will be Π−λ0, and it holds: dim Nλ0(A1) = dim N1, dim Nλ0(A2) = dim N2. The required operators A1 and A2 is not hard to construct by using the Neumann formulas. As in (67) and in the proof of the previous proposition we may write: Nλ(A) = Nλ(A1) = dim N1, Nλ(A) = Nλ(A2) = dim N2, λ ∈ Πλ0. Consider arbitrary isometric operators U and W , mapping respectively N1 on Nλ0(A1), and N2 on Nλ0(A2). Consider the following operator-valued function: F1(λ) := W F (λ)U −1, λ ∈ Πλ0. The function F1(λ) is analytic in the half-plane Πλ0, and its values are lin- ear non-expanding operators from Nλ0(A) into Nλ0(A). Since A is densely defined, we can apply the Shtraus formula (42) Namely, the following formula Rλ =(cid:0)AF1(λ) − λEH(cid:1)−1 , λ ∈ Πλ0, defines a generalized resolvent Rλ of A. It follows that AF1(λ) = R−1 λ + λEH , λ ∈ Πλ0. The generalized resolvent Rλ is generated by a (not necessarly unique) self- fornula (98) we conclude that adjoint extension eA in a Hilbert space eH ⊇ H. Consider the operator-valued function Bλ(A, eA), λ ∈ Re. Comparing the above expression for AF1(λ) with Recalling the definition of the function F(λ; λ0, A, eA) we see that Bλ = AF1(λ), λ ∈ Πλ0. F(λ; λ0, A, eA) = F1(λ), ψn := Ugn, ψ := Ug, λ ∈ Πλ0. n ∈ N. Denote By the condition of the theorem we get that the sequence ψn weakly converges to the element ψ. From (155) it follows that sup Im λn(cid:16)kψnkH − kF(λ; λ0, A, eA)ψnkH(cid:17)(cid:21) < ∞. n∈N(cid:20) λn2 85 Proposition 3.10, and (152) follows. From that relation it follows the required relation (156). ✷ By Theorem 3.11 we conclude that ψ ∈ D(Φ∞(λ0; A, eA)). We may apply Now we can state a theorem connecting the operator Φ∞(λ0; A, eA) with the operator-valued function F(λ; λ0, A, eA). D(A) = H, and eA be a self-adjoint extension of A, acting in a Hilbert space eH ⊇ H. Let λ0 ∈ Re, 0 < ε < π Theorem 3.12 Let A be a closed symmetric operator in a Hilbert space H, 2 , be arbitrary numbers. Then the following relations hold: limλ∈Πε λ0 D(Φ∞(λ0; A, eA)) = {ψ ∈ Nλ0(A) : ,λ→∞hλ(kψkH − kF(λ; λ0, A, eA)ψkH)i < +∞o , λ∈Πε ,λ→∞ lim λ0 F(λ; λ0, A, eA)ψ, Φ∞(λ0; A, eA)ψ = ψ ∈ D(Φ∞(λ0; A, eA)). Suppose that ψ ∈ D(Φ∞(λ0; A, eA)) ⊆ Nλ0(A). Check the in- If ψ = 0, then the inequality is obvious. Assume that λ0, converging to ∞. Proof. equality in (157). ψ 6= 0. Consider an arbitrary sequence {λn}∞ By (146) we define the function ψ(λ), λ ∈ Πλ0. Set n=1, λn ∈ Πε (158) (157) gn := ψ(λn) ∈ Nλ0(A), n ∈ N. By Theorem 3.10 the following relation hold: lim n→∞ gn = ψ, λn − λn lim n→∞(cid:26) λn2 n→∞(cid:26) λn2 Im λn lim lim n→∞ F(λn; λ0, A, eA)gn = Φ∞ψ, [(gn, gn)H − (F(λn)gn, F(λn)gn)H](cid:27) [kgnkH − kF(λn)gnkH] [kgnkH + kF(λn)gnkH](cid:27) . and there exists a finite limit = 1 2i Notice that kgnkH + kF(λn)gnkH → kψkH + kΦ∞ψkH = 2kψkH, n → ∞, since Φ∞ is isometric. Since ψ 6= 0, it follows that there exists a finite limit n→∞(cid:26) λn2 Im λn lim [kgnkH − kF(λn)gnkH](cid:27) . 86 We can apply Theorem 3.11 to the sequences {λn}∞ n=1, {gn}∞ n=1, which con- verges to g, and the function F(λ; λ0, A, eA), mapping Nλ0(A) in Nλ0(A), and obtain that there exists a finite limit lim λ0 λ∈Πε H − kF(λ)ψk2 Im λ(cid:2)kψk2 ,λ→∞(cid:26) λ2 [kψkH − kF(λ)ψkH] [kψkH + kF(λ)ψkH](cid:27) . H(cid:3)(cid:27) ,λ→∞(cid:26) λ2 Im λ lim λ0 λ∈Πε (159) Since the operator F(λ) is non-expanding, then kF(λn)ψ − Φ∞ψkH = kF(λn)ψ − F(λn)gn + F(λn)gn − Φ∞ψkH ≤ kkF(λn)ψ − F(λn)gnkH + kF(λn)gn − Φ∞ψkH ≤ kkψ − gnkH + kF(λn)gn − Φ∞ψkH → 0, n → ∞. Since the sequence {λn}∞ to ∞, then n=1 was an arbitrary sequence from Πε λ0, converging lim λ0 λ∈Πε , λ→∞ F(λ)ψ = Φ∞ψ. Thus, relation (158) is proved. We may write: lim λ0 λ∈Πε , λ→∞ (kF(λ)ψkH + kψkH) = kΦ∞ψkH + kψkH = 2kψkH. From the latter relation and (159) it follows that there exists a finite limit: Therefore there exists a finite limit Im λ lim λ0 ,λ→∞(cid:26) λ2 ,λ→∞(cid:26) λ2 Im λ [kψkH − kF(λ)ψkH](cid:27) =: l(ψ). [kψkH − kF(λ)ψkH](cid:27) = l(ψ). λ∈Πε lim λ0 λ∈Πε In particular, there exists a finite limit n→∞(cid:26) λn2 lim Im λn [kψkH − kF(λn)ψkH](cid:27) = l(ψ). Now suppose that the sequence {λn}∞ infinite) limit n=1 is such that there exists a (finite or lim n→∞ {λn [kψkH − kF(λn)ψkH]} . 87 Since(cid:12)(cid:12)(cid:12) Im λn λn (cid:12)(cid:12)(cid:12) ≤ 1, n ∈ N, then n→∞(cid:26) λn2 lim n→∞ = lim The required estimate in (157) follows. Conversely, suppose that ψ ∈ Nλ0(A) and {λn [kψkH − kF(λn)ψkH]} Im λn Im λ [kψkH − kF(λn)ψkH](cid:27)(cid:12)(cid:12)(cid:12)(cid:12) λn (cid:12)(cid:12)(cid:12)(cid:12) ≤ l(ψ). ,λ→∞hλ(kψkH − kF(λ; λ0, A, eA)ψkH)i < +∞. limλ∈Πε λ0 This means that there exists a sequence of numbers {λn}∞ converging to ∞, and such that there exists a finite limit n=1, λn ∈ Πε λ0, lim n→∞ {λn [kψkH − kF(λn)ψkH]} . In particular, this means that {λn [kψkH − kF(λn)ψkH]} < ∞. sup n∈N By Corollary 3.5 with ψn = ψ, n ∈ N, we get ψ ∈ D(Φ∞(λ0; A, eA)). ✷ 3.10 A connection of the operator Φ∞ and the operator- valued function F(λ). Our aim here will be to prove the following theorem. Theorem 3.13 Theorem 3.12 remains valid, if one removes the condition D(A) = H. Proof. In the case D(A) = H the theorem is already proved. Suppose there corresponds an admissible with respect to A isometric operator Φ∞ = now that D(A) 6= H. For the extension eA there correspond an operator B∞ = B∞(A, eA) and operator-valued functions Bλ = Bλ(A, eA) and F(λ) = F(λ; λ0, A, eA) (see (95), (93), (94), (100)). Recall that for the operator B∞ Φ∞(λ0; A, eA) with the domain D(Φ∞) ⊆ Nλ0(A) and the range R(Φ∞) ⊆ As before, the space H decompose as in (64), where He = eH ⊖ H. The Nλ0(A) (see a text below (101)). Observe that in the case D(A) = H the operator Φ∞ was intensively used in the previous subsection. operator A may be identified with an operator A := A⊕Ae, Ae = oHe. Thus, 88 the operator eA is a self-adjoint extension of A. By the generalized Neumann formulas for the self-adjoint extension eA of A there corresponds an admissible (Ae)T PNλ0 (Ae). Moreover, we have eA = AT . Notice that the operator with respect to A isometric operator T with the domain D(T ) = Nλ0(A) and the range R(T ) = Nλ0(A), having the block representation (69), where T11 = (Ae)T PNλ0 (A), T22 = PNλ0 PNλ0 B = B(λ0; A, T ), defined in (83), coincides with B∞. By Theorem 3.8 we have: B(λ0; A, T ) = AΦ(λ0;A,T ). Since the Neumann formulas define a one- to-one correspondence, then (A)T PNλ0 (Ae), T21 = PNλ0 (A)T PNλ0 (A), T12 = PNλ0 Φ(λ0; A, T ) = Φ∞(λ0; A, eA). Consider an arbitrary closed symmetric operator A1 in a Hilbert space H1, D(A1) = H1, which has the same defect numbers as A. Consider arbitrary isometric operators U and W , mapping respectively Nλ0(A1) on Nλ0(A), and Nλ0(A1) on Nλ0(A). Then the following operator T21U T22 (cid:19) V :=(cid:18) W −1T11U W −1T12 0 E (cid:19) , T21 T22 (cid:19)(cid:18) U 0 E (cid:19)(cid:18) T11 T12 0 =(cid:18) W −1 0 maps isometrically all the subspace Nλ0(A1) ⊕ Nλ0(Ae) on all the subspace Nλ0(A1) ⊕ Nλ0(Ae). By the first assertion of Theorem 3.5, the operator T22 is adnissible with respect to Ae. By the second assertion of Theorem 3.5 we get that V is admissible with respect to A1 := A1 ⊕ Ae, acting in a Hilbert By (121) we may write: Consider the operator-valued functions B′ space eH1 := H1 ⊕He. By the generalized Neumann formulas for the operator V there corresponds a self-adjoint operator eA1 ⊇ A1 in a Hilbert space eH1: eA1 = (A1)V . F(λ; λ0, A1, eA1) = V11 + V12(ENλ0 (Ae) − Ce(λ)V22)−1Ce(λ)V21, where Ce(λ) = C(λ; λ0, Ae) is the characteristic function of the operator Ae in He. Then λ = Bλ(A1, eA1) F′(λ) = F(λ; λ0, A1, eA1). λ ∈ Πλ0, F′(λ) = W −1T11U +W −1T12(ENλ0 (Ae) −Ce(λ)T22)−1Ce(λ)T21U = W −1F(λ)U, λ ∈ Πλ0, (160) what follows from (121). 89 Consider the following operators B′ Also consider the operator B′ = B(λ0; A1, V ), defined in (83). It coincides ∞. By Theorem 3.8 we get B′ = AΦ(λ0;A1,V ). Since the with the operator B′ Neumann formulas define a one-to-one correspondence then ∞ = B∞(A1, eA1) and Φ′ ∞ = Φ∞(λ0; A1, eA1). Φ′ := Φ(λ0; A, T ) = Φ′ ∞. By Remark 3.3 and formula (73) we may write: Φ′ψ′ 1 = V11ψ′ 1 + V12(Xz(Ae) − V22)−1V21ψ′ 1, ψ′ 1 ∈ D(Φ′), (161) and Φψ1 = T11ψ1 + T12(Xz(Ae) − T22)−1T21ψ1, ψ1 ∈ D(Φ). By the definition of the operator Φ, D(Φ) consists of elements of ψ1 ∈ Nλ0(A) such that there exists ψ2 ∈ D(Xλ0(Ae)): T21ψ1 + T22ψ2 = Xλ0(Ae)ψ2, and D(Φ′) consists of elements ψ′ D(Xλ0(Ae)): 1 ∈ Nλ0(A1) such that there exists ψ′ 2 ∈ V21ψ′ 1 + V22ψ′ 2 = T21Uψ′ 1 + T22ψ′ 2 = Xλ0(Ae)ψ′ 2. Therefore By (161) we get D(Φ) = UD(Φ′). (162) Φ′ψ′ 1 = W −1T11Uψ′ 1 + W −1T12(Xz(Ae) − T22)−1T21Uψ′ 1 = W −1ΦUψ′ 1, Therefore ψ′ 1 ∈ D(Φ′). Φ′ = W −1ΦU. (163) By Theorem 3.12 for the operator A1 in a Hilbert space H1 and its self-adjoint extension eA1 in eH1 we get ∞) =nψ′ ∈ Nλ0(A1) : D(Φ′ limλ∈Πε λ0 Φ′ ∞ψ′ = lim λ0 λ∈Πε ,λ→∞ ,λ→∞ [λ(kψ′kH1 − kF′(λ)ψ′kH)] < +∞o , F′(λ)ψ′, ψ′ ∈ D(Φ′ ∞). By relations (162), (163) and (160) we obtain the required relations (157) and (158). ✷ 90 3.11 Shtraus's formula for the generalized resolvents of symmetric operator. Consider a closed symmetric operator A in a Hilbert space H. Choose and fix an arbitrary point λ0 ∈ Re. A function F (λ) ∈ S(Πλ0; Nλ0(A), Nλ0(A)) is said to be λ0-admissible (admissible) with respect to the operator A, if the validity of lim λ0 λ∈Πε , λ→∞ F (λ)ψ = Xλ0ψ, , λ→∞ [λ(kψkH − kF (λ)ψkH)] < +∞, (164) (165) limλ∈Πε λ0 for some ε: 0 < ε < π 2 , implies ψ = 0. A set of all operator-valued functions F (λ) ∈ S(Πλ0; Nλ0(A), Nλ0(A)), which are λ0-admissible with respect to the operator A, we shall denote by Sa;λ0(Πλ0; Nλ0(A), Nλ0(A)) = Sa(Πλ0; Nλ0(A), Nλ0(A)). In the case D(A) = H, we have D(Xλ0) = {0}. Therefore in this case an arbitrary function from S(Πλ0; Nλ0(A), Nλ0(A)) is admissible with respect to A. Thus, if D(A) = H, then Sa;λ0(Πλ0; Nλ0(A), Nλ0(A)) = S(Πλ0; Nλ0(A), Nλ0(A)). Proposition 3.11 Let A be a closed symmetric operator in a Hilbert space H, and λ0 ∈ Re be an arbitrary number. If F (λ) ∈ S(Πλ0; Nλ0(A), Nλ0(A)) is λ0-admissible with respect to A, then F (ζ) is a λ0-admissible operator with respect to A, for all ζ from Πλ0. Proof. Choose an arbitrary point ζ from Πλ0. Suppose that for an element ψ ∈ D(Xλ0(A)) we have the equality: Since the forbidden ooperator is isometric, we get F (ζ)ψ = Xλ0(A)ψ. kF (ζ)ψkH = kψkH. On the other hand, since the operator F (λ) is non-expanding, we may write: kF (λ)ψkH ≤ kψkH, λ ∈ Πλ0. By the maximum principle for analytic vector-valued functions we get F (λ)ψ = Xλ0(A)ψ, λ ∈ Πλ0. Therefore kF (λ)ψkH = kψkH, λ ∈ Πλ0, and relations (164),(165) hold. Since F (λ) is admissible with respect to A, then ψ = 0. Thus, the operator F (ζ) is admissible with respect to A. ✷ Theorem 3.14 Let A be a closed symmetric operator in a Hilbert space H, and λ0 ∈ Re be an arbitrary point. An arbitrary generalized resolvent Rs;λ of the operator A has the following form: 91 Rs;λ = (cid:0)AF (λ) − λEH(cid:1)−1 (cid:16)AF ∗(λ) − λEH(cid:17)−1 , λ ∈ Πλ0 , λ ∈ Πλ0 , (166) where F (λ) is a function from Sa;λ0(Πλ0; Nλ0, Nλ0). Conversely, an arbitrary function F (λ) ∈ Sa;λ0(Πλ0; Nλ0, Nλ0) defines by relation (166) a general- ized resolvent Rs;λ of the operator A. Moreover, for different functions from Sa;λ0(Πλ0; Nλ0, Nλ0) there correspond different generalized resolvents of the operator A. Proof. Let A be a closed symmetric operator in a Hilbert space H, and λ0 ∈ Re be a fixed number. Consider an arbitrary generalized resolvent Rs;λ(A) of A. It is generated by a self-adjoint extension eA of A in a Hilbert space eH ⊇ H. Consider the function Bλ(A, eA) (see (93),(94)) and the function F(λ; λ0, A, eA) (see (100)). From relations (97) and (102) it follows (166), Nλ0(A) hold (164) and (165). By Theorem 3.13 we get ψ ∈ D(Φ∞(λ0; A, eA)), where F (λ) = F(λ) is a function from S(Πλ0; Nλ0, Nλ0), which values are admissible with respect to A operators. Let us show that the function F (λ) is admissible with respect to A. Suppose that for an element ψ ∈ D(Xλ0(A)) ⊆ and Φ∞(λ0; A, eA)ψ = Xλ0(A)ψ. Since Φ∞ is admissible with respect to A, then ψ = 0. Therefore the function F (λ) is admissible with respect to A. Conversely, let F (λ) be an arbitrary function from Sa;λ0(Πλ0; Nλ0, Nλ0). Consider an arbitrary closed symmetric operator A1 in a Hilbert space H1, D(A1) = H1, which has the same defect numbers as A. Consider arbitrary isometric operators U and W , whic map respectively Nλ0(A1) on Nλ0(A), and Nλ0(A1) on Nλ0(A). Set F1(λ) = W −1F (λ)U, λ ∈ Πλ0. (167) F1(λ) S(Πλ0; Nλ0(A1), Nλ0(A1)). Since in the case of the densely defined operators the theorem was proved, we can assert that F1(λ) generates a generalized resolvent Rλ(A1) of the operator A1 in H1: Rλ(A1) = (cid:0)(A1)F1(λ) − λEH(cid:1)−1 , λ ∈ Πλ0 (cid:16)(A1)F ∗(λ) − λEH(cid:17)−1 , λ ∈ Πλ0 . (168) 92 The operator A1 may be considered as an operator A1 := A1 ⊕Ae in a Hilbert The generalized resolvent Rλ(A1) is generated by a self-adjoint extension eA1 in a Hilbert space eH1 ⊇ H1. Set space eH1 = H1 ⊕ He, where Ae = oHe. By the generalized Neumann formulas for the self-adjoint extension eA1 of A1 there correspond an admissible with respect to A1 isometric opera- tor T with the domain D(T ) = Nλ0(A1) and the range R(T ) = Nλ0(A1), (A1)T PNλ0 (A1), T12 = having the block representation (69), where T11 = PNλ0 PNλ0 (Ae)T PNλ0 (Ae). More- (Ae)T PNλ0 (A1), T22 = PNλ0 He := eH1 ⊖ H1. (A1)T PNλ0 (Ae), T21 = PNλ0 Notice that the operator B = B(λ0; A1, T ), defined by (83), coincides AΦ(λ0;A1,T ). Since the Neumann formulas establish a one-to-one correspon- dence, then over, we have eA1 = (A1)T . with the operator B∞(A1, eA1). By Theorem 3.8 we get: B(λ0; A1, T ) = Φ′ := Φ(λ0; A1, T ) = Φ∞(λ0; A1, eA1). For the extension eA1 there correspond functions B′ λ = Bλ(A1, eA1) and F′(λ) = F(λ; λ0, A1, eA1). By (121) we obtain the following equality: F′(λ) = T11 + T12(ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21, λ ∈ Πλ0, (169) where Ce(λ) = C(λ; λ0, Ae) is the characteristic function of the operator Ae in He. Since T is admissible with respect to A1, by Theorem 3.5 the operator T22 is admissible with respect to Ae. By Remark 3.3 the operator Φ′ = Φ(λ0; A1, T ) admits the following representation: Φ′ψ′ 1 = T11ψ′ 1 + T12(Xλ0(Ae) − T22)−1T21ψ′ 1, ψ′ 1 ∈ D(Φ′). (170) By Theorem 3.13 we get D(Φ′) =nψ′ ∈ Nλ0(A1) : limλ∈Πε λ0 Φ′ψ′ = lim λ0 λ∈Πε ,λ→∞ ,λ→∞ [λ(kψ′kH1 − kF′(λ)ψ′kH1)] < +∞o , F′(λ)ψ′, ψ′ ∈ D(Φ′), (171) (172) where 0 < ε < π 2 . Comparing relations (168) and (97), taking into account (101) we see that F′(λ) = F1(λ), λ ∈ Πλ0. (173) Consider the following operator 93 V =(cid:18) V11 V12 =(cid:18) W 0 T22 (cid:19) V21 V22 (cid:19) :=(cid:18) W T11U −1 W T12 E (cid:19) , 0 E (cid:19)(cid:18) T11 T12 T21 T22 (cid:19)(cid:18) U −1 T21U −1 0 0 whic maps isometrically all the subspace Nλ0(A) ⊕ Nλ0(Ae) on the whole subspace Nλ0(A) ⊕ Nλ0(Ae). A = A⊕Ae is admissible with respect to Ae. By Remark 3.3 we obtain that the operator Φ = Φ(λ0; A, V ) admits the followig representation: eH := H ⊕He. As it was already noticed, the operator T22 Φψ1 = V11ψ1 + V12(Xλ0(Ae) − V22)−1V21ψ1, ψ1 ∈ D(Φ). By the definition of the operator Φ, D(Φ) consists of elements ψ1 ∈ Nλ0(A) such that there exists ψ2 ∈ D(Xλ0(Ae)): V21ψ1 + V22ψ2 = Xλ0(Ae)ψ2, and D(Φ′) consists of elements ψ′ D(Xλ0(Ae)): 1 ∈ Nλ0(A1) such that there exists ψ′ 2 ∈ T21ψ′ 1 + T22ψ′ 2 = V21Uψ′ 1 + T22ψ′ 2 = Xλ0(Ae)ψ′ 2. It follows that From (170) it follows that D(Φ) = UD(Φ′). (174) Φ′ψ′ 1 = T11ψ′ 1 + T12(Xλ0(Ae) − T22)−1T21ψ′ 1 = W −1V11Uψ′ 1 + W −1V12(Xλ0(Ae) − V22)−1V21Uψ′ 1 = W −1ΦUψ′ 1, ψ′ 1 ∈ D(Φ′). Therefore Φ′ = W −1ΦU. (175) From (171),(172) it follows that D(Φ) =nψ ∈ Nλ0(A) : limλ∈Πε λ0 Φψ = lim λ0 λ∈Πε ,λ→∞ F (λ)ψ, ψ ∈ D(Φ), ,λ→∞ [λ(kψkH − kF (λ)ψkH)] < +∞o , (176) (177) 94 where 0 < ε < π 2 . Let us check that the operator Φ is admissible with respect to A. Consider an arbitrary element ψ ∈ D(Φ) ∩ D(Xλ0(A)) such that Using (177) we get Φψ = Xλ0(A)ψ. lim λ0 λ∈Πε ,λ→∞ F (λ)ψ = Xλ0(A)ψ. From (176) it follows that limλ∈Πε λ0 ,λ→∞ [λ(kψkH − kF (λ)ψkH)] < +∞. by Since the function F (λ) is λ0-admissible with respect to A, then ψ = 0. Therefore Φ is admissible with respect to A. Moreover, V22 = T22 is ad- missible with respect to Ae. By Theorem 3.6 the operator V is admissible with respect to A = A ⊕ Ae. By the generalized Neumann formulas for the operator V there corresponds a self-adjoint extension eA := AV A. Denote eA. the generalized resolvent of A, which corresponds to the self-adjoint extension From (167),(169) and (173) it follows that eH H (eA − λE eH)−1, F (λ) = W F1(λ)U −1 = W (T11 + T12(ENλ0 (Ae) − Ce(λ)T22)−1Ce(λ)T21)U −1 Rλ = P λ ∈ Re, = V11 + V12(ENλ0 (Ae) − Ce(λ)V22)−1Ce(λ)V21, λ ∈ Πλ0. By (121) we conclude that F (λ) = F(λ; λ0, A, eA), λ ∈ Πλ0. From (97) and (102) it follows (166) for the constructed generalized resolvent Rλ and for the given function F (λ). Let us check the last assertion of the theorem. Suppose that two functions F1(λ), F1(λ) ∈ Sa;λ0(Πλ0; Nλ0, Nλ0) generate by (166) the same generalized resolvent of A. Then AF1(λ) = AF2(λ), λ ∈ Πλ0. By the generalized Neumann formulas we obtain that F1(λ) = F2(λ), λ ∈ Πλ0. ✷ 95 4 Generalized resolvents of isometric and sym- metric operators with gaps in their spec- trum. 4.1 Spectral functions of an isometric operator having a constant value on an arc of the circle. In the investigation of interpolation problems we shall use spectral functions of the operator related to a problem. There will appear problems with spec- tral functions, which are constant on the prescribed arcs of the unit circle. Such spectral functions and the corresponding generalized resolvents will be studied in this subsection. Proposition 4.1 Let V be a closed isometric operator in a Hilbert space H, and F(δ), δ ∈ B(T), be its spectral measure. The following two conditions are equivalent: (i) F(∆) = 0, for an open arc ∆ of the unit circle T; (ii) The generalized resolvent Rz(V ), corresponding to the spectral measure F(δ), admits analytic continuation on the set D ∪ De ∪ ∆, where ∆ = {z ∈ C : z ∈ ∆}, for an open arc ∆ of T. Proof. (i)⇒(ii). In this case relation (1) takes the following form: (Rzh, g)H =ZT\∆ 1 1 − zζ d(F(·)h, g)H, ∀h, g ∈ H. (178) Choose an arbitrary number z0 ∈ ∆. Since the function and bounded on T\∆, then there exists an integral 1 1−z0ζ is continuous 1 1 − z0ζ Iz0(h, g) :=ZT\∆ (Rzh, h)H − Iz0(h, h) = z − z0(cid:12)(cid:12)(cid:12)(cid:12)ZT\∆ ≤ z − z0ZT\∆ 1 − zζ1 − z0ζ ζ d(F(·)h, g)H. ζ (1 − zζ)(1 − z0ζ) d(F(·)h, h)H, d(F(·)h, h)H(cid:12)(cid:12)(cid:12)(cid:12) z ∈ Te. There exists a neighborhood U(z0) of z0 such that z − ζ ≥ M1 > 0, ∀ζ ∈ T\∆, ∀z ∈ U(z0). Thus, the integral in the last relation is bounded in the neighborhood U(z0). Therefore we get (Rzh, h)H → Iz0(h, h), z ∈ Te, z → z0, ∀h ∈ H. Using the properties of sesquilinear forms we get (Rzh, g)H → Iz0(h, g), z ∈ Te, z → z0, ∀h, g ∈ H. 96 Set Rez := w. − lim z∈Te, z→ez Rz, ∀ez ∈ ∆, where the limit is understood in a sense of the weak operator topology. We may write (cid:18) 1 z − z0 (Rz − Rz0)h, h(cid:19)H =ZT\∆ ζ (1 − zζ)(1 − z0ζ) d(F(·)h, h)H, z ∈ U(z0), h ∈ H. The function under the integral sign is bounded in U(z0), and it tends to (1−z0ζ)2 . By the Lebesgue theorem on a limit we obtain that ζ and therefore z − z0 lim z→z0(cid:18) 1 z→z0(cid:18) 1 lim z − z0 (Rz − Rz0)h, h(cid:19)H (Rz − Rz0)h, g(cid:19)H =ZT\∆ =ZT\∆ ζ (1 − z0ζ)2 d(F(·)h, h)H; ζ (1 − z0ζ)2 d(F(·)h, g)H, for h, g ∈ H. Thus, there exists the derivative of the function Rz at z = z0. (ii)⇒(i). Choose an arbitrary element h ∈ H, and consider the following function σh(t) := (Fth, h)H, t ∈ [0, 2π), where Ft is a left-continuous spec- tral function of V , corresponding to a spectral measure F(δ). Consider the following function: 1 + eitz 1 − eitz dσh(t) =Z 2π 0 1 1 − eitz dσh(t) − 1 2Z 2π 0 dσh(t) 1 2 Choose arbitrary numbers t1, t2, 0 ≤ t1 < t2 ≤ 2π, such that H = (Rzh, h)H − dσh(t) − 1 − eitz khk2 1 2 1 khk2 H. (179) l = l(t1, t2) = {z = eit : t1 ≤ t ≤ t2} ⊂ ∆. (180) We assume that t1 and t2 are points of the continuity of the function Ft. By the inversion formula we may write: σh(t2) − σh(t1) = lim r→1−0Z t2 t1 Re(cid:8)fh(re−iτ )(cid:9) dτ. fh(z) = 0 1 2Z 2π =Z 2π 0 Observe that = 1 2 Re(cid:8)fh(re−iτ )(cid:9) = Re {(Rre−iτ h, h)H} − 97 1 2 khk2 H (((Rre−iτ + R∗ re−iτ )h, h)H) − khk2 H, t1 ≤ τ ≤ t2. (181) 1 2 By (180) we obtain that e−iτ belongs to ∆, for t1 ≤ τ ≤ t2. Therefore lim r→1−0 ((Rre−iτ + R∗ re−iτ )h, h)H = ((Re−iτ + R∗ e−iτ )h, h)H . (182) By Theorem 2.1 the generalized resolvents of an isometric operator have the following property: R∗ z = EH − R 1 z , z ∈ Te. Passing to a limit in (183) as z tends to e−iτ , we get R∗ e−iτ = EH − Re−iτ , t1 ≤ τ ≤ t2. By (181),(182) and (184) we get (183) (184) lim r→1−0 Re(cid:8)fh(re−iτ )(cid:9) = 0, Consider the followinf sector: t1 ≤ τ ≤ t2. (185) L(t1, t2) = {z = re−it : t1 ≤ t ≤ t2, 0 ≤ r ≤ 1}. The generalized resolvent is analytic in each point of the closed sector L(t1, t2). Consequently, the function Re(Rzh, h) is continuous and bounded in L(t1, t2). By the Lebesgue theorem on a limit we conclude that σh(t1) = σh(t2). If 1 /∈ ∆ the required result follows easily. In the case 1 ∈ ∆, we may write ∆ = ∆1 ∪ {1} ∪ ∆2, where the open arcs ∆1 and ∆2 do not contain 1. There- fore σh(t) is constant in intervals, corresponding to ∆1 and ∆2. Suppose that there exists a non-zero jump of σh(t) at t = 0. By (1) we may write: 1 dσh(t) =Z 2π 0 (Rzh, h)H =Z 2π where bσh(t) = σh(t) + σh(+0) − σh(0), t ∈ [0, 2π]. In a neighborhood of 1 the left-hand side and the first summand in the right-hand side are bounded. We obtained a contradiction. ✷ dbσh(t) + 1 − eitz 1 − eitz 1 − z a, a > 0, 0 1 1 98 Theorem 4.1 Let V be a closed isometric operator in a Hilbert space H, and Rz(V ) be an arbitrary generalized resolvent of the operator V . Let {λk}∞ be a sequence of numbers from D such that λk → bλ, as k → ∞; bλ ∈ T. Suppose that for a number z0 ∈ D\{0}, the function C(λ; z0), corresponding to Rz(V ) in Inin's formula, satisfies the following relation: k=1 (186) (187) ∃u. − lim k→∞ C(λk; z0) =: C(bλ; z0). Then for an arbitrary z′ 0), corresponding to the generalized resolvent Rz(V ) by Inin's formula (25) satisfies the following relation: 0 ∈ D\{0}, the function C(λ; z′ ∃u. − lim k→∞ C(λk; z′ 0). 0) =: C(bλ; z′ 0) is a linear non-expanding operator, which maps the whole does not , and the corresponding orthogonal extension VC(bλ;z′ 0);z′ 0 Moreover, C(bλ; z′ 0 into N 1 z′ 0 Nz′ depend on the choice of the point z′ 0 ∈ D\{0}. Proof. Suppose that relation (186) holds for a point z0 ∈ D\{0}. Choose an arbitrary point z′ 0 ∈ D\{0}. Comparing the Inin formula for z0 and for z′ 0, we see that VC(λ;z0);z0 = VC(λ;z′ 0, 0);z′ λ ∈ D. (188) λ ∈ D. (189) Substituting in (188) similar relations for z0 and z′ we get 0, and multiplying by z0z′ 0 By (23) we may write VC(λ;z0);z0 = 1 z0 EH + z′ 0EH + z′ = z0EH + z0(z′ , z0 z02 − 1 z0;C(λ;z0)(cid:17)−1 z0;C(λ;z0)(cid:17)−1 0)(cid:17)−1 (cid:16)EH + z0V + 0(z02 − 1)(cid:16)EH + z0V + 02 − 1)(cid:16)EH + z′ 0)(cid:17)−1 z0;C(λ;z0)(cid:17)−1(cid:19) z0;C(λ;z0)(cid:19)(cid:16)EH + z0V + 1 02 − 1) 0V + z′ 0;C(λ;z′ 0(z02 − 1)(cid:16)EH + z0V + z0 − z′ 0 1 − z′ 0z0 z0(z′ V + = 0V + 0;C(λ;z′ z′ . ((z′ 0 − z0)EH Then = (cid:16)EH + z′ +z′ 1 − z′ 1 − z′ 0z0 02(cid:18)EH + z0;C(λ;z0)(cid:17)−1 , λ ∈ D. (190) Lemma 4.1 Let z0, z′ 0 ∈ D. Then (cid:12)(cid:12)(cid:12)(cid:12) z0 − z′ 0 1 − z′ 0z0(cid:12)(cid:12)(cid:12)(cid:12) < 1. Proof. Consider a linear fractional transformation: w = w(u) = z0−u 1−z0u . If u = 1, then 1 − z0u = u(u − z0) = u − z0. Moreover, we have w(z0) = 0. Therefore w maps D on D. ✷ Using (190) and (191) we may write: EH + z′ 0V + z′ 0;C(λ;z′ 0) = 1 − z′ 1 − z′ 02 0z0(cid:16)EH + z0V + z0;C(λ;z0)(cid:17)(cid:18)EH + z0 − z′ 0 1 − z′ 0z0 V + z0;C(λ;z0)(cid:19)−1 , λ ∈ D. (192) Using (21) we write: V + z0;C(λ;z0) = Vz0 ⊕ C(λ; z0), λ ∈ D. By the statement of the theorem we easily obtain that C(bλ; z0) is a non- expanding operator and 99 (191) ∃u. − lim k→∞ We may write (cid:18)EH + z0 − z′ 0 1 − z′ 0z0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ∗(cid:12)(cid:12)(cid:12)(cid:12) z0 − z′ 0 1 − z′ 0z0(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)V + Since then (cid:13)(cid:13)(cid:13)(cid:13)(cid:18)EH + z0;C(bλ;z0) . (193) V + z0;C(bλ;z0)(cid:19)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0;C(bλ;z0)(cid:19)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) V + . (194) V + V + V + z0 − z′ 0 1 − z′ 0z0 z0 − z′ 0 1 − z′ 0z0 z0;C(λk;z0) = Vz0 ⊕ C(bλ; z0) = V + z0;C(λk;z0)(cid:19)−1 −(cid:18)EH + z0;C(λk;z0)(cid:19)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:18)EH + z0;C(bλ;z0)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:18)EH + 0z0(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)V + (cid:12)(cid:12)(cid:12)(cid:12) z0;C(λk;z0)(cid:13)(cid:13)(cid:13) ≤ δ < 1, z0;C(λk;z0)(cid:19) h(cid:13)(cid:13)(cid:13)(cid:13) ≥(cid:12)(cid:12)(cid:12)(cid:12)khk −(cid:13)(cid:13)(cid:13)(cid:13) z0 − z′ 0 1 − z′ z0 − z′ 0 1 − z′ 0z0 z0 − z′ 0 1 − z′ 0z0 z0;C(λk;z0) − V + z0 − z′ 0 1 − z′ 0z0 V + V + z0;C(λk;z0)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Passing to a limit in (194) we get (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:18)EH + k→∞(cid:18)EH + z0 − z′ 0 1 − z′ 0z0 u. − lim ≥ (1 − δ)khk; z0 − z′ 0 1 − z′ 0z0 V + ≤ z0;C(λk;z0)(cid:19)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:18)EH + V + z0;C(λk;z0)(cid:19)−1 100 1 1 − δ . z0 − z′ 0 1 − z′ 0z0 V + z0;C(bλ;z0)(cid:19)−1 . (195) By (195),(193),(192) we obtain that there exists a limit u. − lim k→∞ such that V + z′ 0;C(λk;z′ 0) = u. − lim k→∞ Vz′ 0 ⊕ C(λk; z′ 0) =: V ′, EH + z′ 0V ′ = 1 − z′ 1 − z′ 02 0z0(cid:16)EH + z0V + z0;C(bλ;z0)(cid:17)(cid:18)EH + Relation (196) shows that V ′Mz′ = Vz′ 0. Set 0 z0 − z′ 0 1 − z′ 0z0 V + z0;C(bλ;z0)(cid:19)−1 . (196) (197) C(bλ; z′ By (196) we obtain that C(bλ; z′ . Thus, we get 0 in N 1 z′ 0 maps Nz′ 0) = V ′Nz′ . 0 0) is a linear non-expanding operator, which V ′ = Vz′ 0) = V + 0;C(bλ;z′ z′ 0) . (198) From (196),(198) it easily follows that u. − lim k→∞ and relation (187) is proved. By (197),(198) we get 0 ⊕ C(bλ; z′ C(λk; z′ 0), 0) = C(bλ; z′ 0)(cid:17)−1 (cid:16)EH + z′ 0V + 0;C(bλ;z′ z′ (199) ; (200) = 1 − z′ 1 − z′ 0z0 02(cid:18)EH + z0 − z′ 0 1 − z′ 0z0 V + z0;C(bλ;z0)(cid:19)(cid:16)EH + z0V + 0)(cid:17)−1 0V + 0;C(bλ;z′ z′ 02)(cid:16)EH + z′ (1 − z′ z0;C(bλ;z0)(cid:17)−1 Subtracting EH from the both sides of the last relation and dividing by −z′ 0 we get: =(cid:16)(1 − z′ 0z0)EH + (z0 − z′ 0)V + 1 z′ 0 EH + z′ 02 − 1 z′ 0 z0;C(bλ;z0)(cid:17)(cid:16)EH + z0V + z0;C(bλ;z0)(cid:17)−1 0)(cid:17)−1 (cid:16)EH + z′ 0V + 0;C(bλ;z′ z′ 101 ; )(cid:17) = − 0z0)EH + (z0 − z′ 0)V + z0;C(bλ;z0) − (EH + z0V + z0;C(bλ;z0) 1 z′ 0(cid:16)(1 − z′ =(cid:16)z0EH + V + z0;C(bλ;z0)(cid:17)−1 ∗(cid:16)EH + z0V + z0;C(bλ;z0)(cid:17)(cid:16)EH + z0V + z0;C(bλ;z0)(cid:17)−1 . From (22) and (23) we get VC(bλ;z′ 0);z′ 0 = VC(bλ;z0);z0 , ∀z′ 0 ∈ D. (201) ✷ Theorem 4.2 Let V be a closed isometric operator in a Hilbert space H, and z0 ∈ D\{0} be a fixed point. Let Rz = Rz(V ) be an arbitrary generalized resolvent of the operator V , and C(λ; z0) ∈ S(Nz0; N 1 ) corresponds to Rz(V ) z0 by Inin's formula (25). The generalized resolvent Rz(V ) admits an analytic continuation on a set D ∪ De ∪ ∆, where ∆ is an open arc of T, if and only if the following conditions hold: 1) The function C(λ; z0) admits a continuation on a set D ∪ ∆, which is continuous in the uniform operator topology; 2) The continued function C(λ; z0) maps isometriclly Nz0 on the whole N 1 z0 , for all points λ ∈ ∆; 3) The operator (EH − λVC(λ;z0);z0)−1 exists and it is defined on the whole H, for all points λ ∈ ∆. Proof. Necessity. Choose an arbitrary point bλ ∈ ∆. Let z ∈ D\{0} be an arbitrary point, and C(λ; z) ∈ S(Nz; N 1 ) correspond to the generalized resolvent Rz(V ) by the Inin formula. Using the Inin formula we may write: z EH − zVC(λ;z);z = = z λ R−1 λ +(cid:16)1 − z λ(cid:17) EH = z λ (EH − λVC(λ;z);z) +(cid:16)1 − λ(cid:20)EH +(cid:18)λ − 1(cid:19) Rλ(cid:21) R−1 z z z λ(cid:17) EH λ , ∀λ ∈ D\{0}. Therefore the following operator 102 − 1(cid:19) Rλ(cid:21) = λ z (EH − zVC(λ;z);z)Rλ, has a bounded inverse, defined on the whole H: z (cid:20)EH +(cid:18)λ − 1(cid:19) Rλ(cid:21)−1 (cid:20)EH +(cid:18)λ z Then = z λ R−1 λ (EH − zVC(λ;z);z)−1, λ ∈ D\{0}. (202) − 1(cid:19) Rλ(cid:21)−1 z , λ ∈ D\{0}. (203) Choose an arbitrary δ: 0 < δ < 1. Assume that z ∈ D\{0} satisfies the following additional condition: λ z δ ; or ε > kRbλk < δ. (EH − zVC(λ;z);z)−1 = condition (204) will take the following form: Rλ(cid:20)EH +(cid:18)λ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)bλ − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) − 1(cid:12)(cid:12)(cid:12)(cid:12) < (cid:12)(cid:12)(cid:12)(cid:12) In this case, there exists an inversehEH +(cid:16) bλ − 1(cid:12)(cid:12)(cid:12)(cid:12) kRλk < δ, (cid:12)(cid:12)(cid:12)(cid:12) (cid:20)EH +(cid:18)λ − 1(cid:19) Rλ(cid:21)−1 − 1(cid:19) Rλ(cid:19) h(cid:13)(cid:13)(cid:13)(cid:13) ≥(cid:12)(cid:12)(cid:12)(cid:12)khk −(cid:12)(cid:12)(cid:12)(cid:12) (cid:13)(cid:13)(cid:13)(cid:13)(cid:18)EH +(cid:18)λ 1+ δ kRbλ kRbλk 1 . k z z 1 ε λ z z , in an open neighborhood U(bλ) ofbλ. Therefore there exists the inverse which is bounded and defined on the whole H. We may write ∀λ ∈ U(bλ), − 1(cid:12)(cid:12)(cid:12)(cid:12) kRλhk(cid:12)(cid:12)(cid:12)(cid:12) λ z Let us check that such points exist. If kRbλk = 0, then it is obvious. In the opposite case we seek z in the following form: z = εbλ, 0 < ε < 1. Then , which is bounded and defined on the whole H. Moreover, by the continuity the following inequality holds: z − 1(cid:17) Rbλi−1 (204) (205) (206) Therefore ≥ (1 − δ)khk, h ∈ H. (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)EH +(cid:18)λ z − 1(cid:19) Rλ(cid:21)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ 1 1 − δ , λ ∈ U(bλ). z . z z z z We write Choose an arbitrary sequence {λk}∞ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)EH +(cid:18) λk ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)EH +(cid:18) λk k=1 of points from D such that λk →bλ, as k → ∞. There exists a number k0 ∈ N such that λk ∈ U(bλ) ∩ D, k ≥ k0. − 1! Rbλ#−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) −"EH + bλ − 1(cid:19) Rλk(cid:21)−1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) − 1(cid:19) Rλk(cid:21)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) − 1! Rbλ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) − 1(cid:19) Rλk − bλ (cid:18)λk ∗(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) − 1! Rbλ#−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) "EH + bλ ="EH + bλ λ∈D, λ→bλ(cid:20)EH +(cid:18) λ Rbλ"EH + bλ The first factor on the right in the last equality is uniformly bounded by (207). Thus, we get − 1! Rbλ#−1 − 1! Rbλ#−1 − 1(cid:19) Rλ(cid:21)−1 From relations (203),(208) it follows that u. − lim u. − lim . (208) . (209) z z λ∈D, λ→bλ z z Thus, there exists the following limit: 103 (207) u. − lim λ∈D, λ→bλ V + z;C(λ;z) = u. − lim z (EH − zVC(λ;z);z)−1 = bλ λ∈D, λ→bλ(cid:18)− bλRbλ"EH + bλ 1 − z2 1 z z2 z = − 1 z EH + EH + 1 − z2 z (EH − zVC(λ;z);z)−1(cid:19) − 1! Rbλ#−1 =: V ′ z , (210) where we used (24). Observe that V + V ′ which maps Nz into N 1 z;C(λ;z) = Vz ⊕ C(λ; z). Set C(bλ; z) = z Nz . By (210) we conclude that C(bλ; z) is a linear non-expanding operator z Mz = Vz, and therefore . Moreover, we have V ′ z Using (210) we easily obtain that V ′ z = Vz ⊕ C(bλ; z) = V + . z;C(bλ;z) u. − lim λ∈D, λ→bλ C(λ; z) = C(bλ; z). 104 (211) By 4.1 we conclude that the last relation holds for z0, where C(bλ; z0) is a linear non-expanding operator which maps Nz0 into N 1 z0 Comparing relation (210) for V ′ and relation (24) we get , and VC(bλ;z0);z0 = VC(bλ;z);z. z = V + z;C(bλ;z) Rbλ"EH + bλ z bλ z − 1! Rbλ#−1 =(cid:16)EH − zVC(bλ;z);z(cid:17)−1 , (212) for the prescribed choice of z. Thus, we continued by the continuity the function C(λ; z0) on a set D∪∆. Let us check that this continuation is continuous in the uniform operator (213) topology. It remains to check that for an arbitrarybλ ∈ ∆ it holds u. − Choose a number z ∈ D\{0}, which satisfies (204) and construct a neigh- borhood U(bλ) as it was done before. For an arbitrary λ ∈ U(bλ) we may write: z z lim λ∈D∪∆, λ→bλ C(λ; z0) = C(bλ; z0). − 1! Rbλ#−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) −"EH + bλ − 1(cid:19) Rλ(cid:21)−1 (cid:20)EH +(cid:18)λ ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) − 1(cid:19) Rλ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) − 1(cid:19) Rλ(cid:21)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) bλ − 1! Rbλ −(cid:18) λ (cid:20)EH +(cid:18)λ ∗(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) − 1! Rbλ#−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) "EH + bλ ="EH + bλ λ→bλ(cid:20)EH +(cid:18)λ − 1(cid:19) Rλ(cid:21)−1 − 1! Rbλ#−1 z z z z z z . Using (207) we get u. − lim . (214) By (212) we get 105 z λ z − 1(cid:19) Rλ(cid:21)−1 =(cid:0)EH − zVC(λ;z);z(cid:1)−1 , Rλ(cid:20)EH +(cid:18)λ ∀λ ∈ (U(bλ)∩D)\{0}, for the prescribed choice of z. In fact, for an arbitraryeλ ∈ ∆ ∩ U(bλ), there exists a neighborhood eU(eλ) ⊂ U(bλ), where inequality (205) is satisfied for the same choice of z. Repeating the arguments below (205) witheλ instead of bλ, we obtain that (215) holds foreλ. For points inside D we can apply (203). From (214),(215) it follows that (215) u. − lim λ∈D∪∆, λ→bλ(cid:0)EH − zVC(λ;z);z(cid:1)−1 =(cid:16)EH − zVC(bλ;z);z(cid:17)−1 Since it was proved that VC(bλ;z);z does not depend on the choice of z ∈ D\{0} (and for z ∈ D this follows from the Inin formula), then the last relation holds for all z ∈ D\{0}. Using relation (24) we get u. − lim λ∈D∪∆, λ→bλ V + z;C(λ;z) = V + z;C(bλ;z) , ∀z ∈ D\{0}, (217) and therefore relation (213) holds. Thus, condition 1) from the statement of the theorem is satisfied. From (212) it is seen that . (216) for the above prescribed choice of z. Therefore the operator Rbλ has a bounded inverse, defined on the whole space H. Then From the latter relation we get Rbλ = EH = z z z − 1! Rbλ# , bλ(cid:16)EH − zVC(bλ;z);z(cid:17)−1"EH + bλ bλ(cid:16)EH − zVC(bλ;z);z(cid:17)−1"EH + bλ − 1! Rbλ# R−1 bλ"EH + bλ − 1! Rbλ# R−1 bλ(cid:19) EH. +(cid:18)1 − bλ Rbλ =(cid:16)EH −bλVC(bλ;z);z(cid:17)−1 R−1 bλ = z z z z bλ bλ . EH − zVC(bλ;z);z = z (218) ; (219) Since VC(bλ;z);z does not depend on the choice of z, then the last relation holds for all z ∈ D\{0}. Consequently, condition 3) from the statement of the theorem holds. Using property (183), and passing to the limit as z →bλ, we get bλ = EH − Rbλ. R∗ On the other hand, from (219) we get: 106 (220) (221) (222) . R∗ From (219)-(221) it is seen that C(bλ;z);z(cid:17)−1 C(bλ;z);z(cid:17)−1 bλ =(cid:16)EH −bλV ∗ EH =(cid:16)EH −bλV ∗ +(cid:16)EH −bλVC(bλ;z);z(cid:17)−1 Multiplying the both sides of the last relation by (EH −bλV ∗ left and by (EH −bλVC(bλ;z);z) from the right we get (EH −bλV ∗ C(bλ;z);z)(EH −bλVC(bλ;z);z) = EH −bλVC(bλ;z);z + EH −bλV ∗ ∀z ∈ D\{0}. C(bλ;z);z , After the multiplication in the left-hand side and simplifying the expression we obtain that ) from the C(bλ;z);z. V ∗ C(bλ;z);zVC(bλ;z);z = EH , ∀z ∈ D\{0}. On the other hand, by (222) we may write: ; 1 VC(bλ;z);z = − (cid:16)EH −bλV ∗ = EH −(cid:16)EH −bλVC(bλ;z);z(cid:17)−1 C(bλ;z);z(cid:17)−1(cid:16)EH −bλVC(bλ;z);z(cid:17) . C(bλ;z);z(cid:17)−1 = −bλVC(bλ;z);z(cid:16)EH −bλVC(bλ;z);z(cid:17)−1 bλ(cid:16)EH −bλV ∗ Since the operator (cid:16)EH −bλVC(bλ;z);z(cid:17)−1 Vz ⊕ C(bλ; z) is unitary, as well. In particular, this means that the operator C(bλ; z) is isometric and it maps Nz on the whole N 1 is defined on the whole H and it is bounded, then we conclude that R(VC(bλ;z);z) = H. Thus, the operator VC(bλ;z);z is unitary in H. Therefore the corresponding operator V + = . Since z is an arbitrary z;C(bλ;z) z point from D\{0}, we obtain that condition 2) from the statement of the theorem is satisfied. 107 Sufficiency. Let conditions 1)-3) be satisfied. Choose an arbitrary point k=1 points from D ∪ ∆ such that (EH + z0V + z0;C(λk;z0))−1 1 z0 = V + z02 − 1 z0 − λk 1 − λkz0 bλ ∈ ∆, and an arbitrary sequence {λk}∞ λk →bλ, as k → ∞. Using (23) we write: z0(cid:19) EH − EH − λkVC(λk;z0);z0 =(cid:18)1 − = (1 − λkz0)(cid:20)EH + = (1 −bλz0)"EH + z0 −bλ 1 −bλz0 z0;C(λk;z0)(cid:21) (EH + z0V + z0;C(bλ;z0)# (EH + z0V + EH −bλVC(bλ;z0);z0 z0 −bλ 1 −bλz0 1 −bλz0(cid:16)EH −bλVC(bλ;z0);z0(cid:17)(cid:16)EH + z0V + Since the operator VC(bλ;z0);z0 Using (223) write V + z0;C(bλ;z0) z0;C(bλ;z0) EH + 1 = V + z0;C(λk;z0))−1; exists the inverse (EH −bλVC(bλ;z0);z0 bounded. Therefore there exists [EH + z0−bλ 1−bλz0 and defined on the whole H. By (223) we get is closed, by condition 3) it follows that there )−1, which is defined on the whole H and ]−1, which is bounded V + z0;C(bλ;z0) )−1. (223) z0;C(bλ;z0)(cid:17) . z0;C(bλ;z0) )"EH + (cid:16)EH −bλVC(bλ;z0);z0(cid:17)−1 z0 −bλ 1 −bλz0 (cid:0)EH − λkVC(λk;z0);z0(cid:1)−1 z0;C(λk;z0))(cid:20)EH + z0 − λk 1 − λkz0 = (EH + z0V + 1 1 −bλz0 = 1 1 − λkz0 (EH + z0V + For points λk, which belong to ∆, we can apply the same argument. For points λk from D we may apply Lemma 4.1. We shall obtain an analogous representation: z0;C(bλ;z0)#−1 V + . (224) V + z0;C(λk;z0)(cid:21)−1 , k ∈ N. (225) By condition 1) we get 108 z0;C(λk;z0) − V + (cid:13)(cid:13)(cid:13)V + sup h∈H, khk=1(cid:13)(cid:13)(cid:13)(C(λk; z0) − C(bλ; z0))PNz0 h(cid:13)(cid:13)(cid:13) z0;C(bλ;z0)(cid:13)(cid:13)(cid:13) = ≤(cid:13)(cid:13)(cid:13)C(λk; z0) − C(bλ; z0)(cid:13)(cid:13)(cid:13) → 0, V + z0;C(λk;z0) = V + u. − lim k→∞ k → ∞; z0;C(bλ;z0) . (226) V + ≤ K, (227) 1 K kgk, K > 0 such that z0 − λk 1 − λkz0 V + z0 − λk 1 − λkz0 The last condition is equivalent to the following condition: ∀λk : λk ∈ U1(bλ). ∀g ∈ H, ∀λk : λk ∈ U1(bλ). (228) Suppose to the contrary that condition (228) is not satisfied. Choose an n; and set Kn = n, n ∈ N. Then for each n ∈ N, there exists an element gn ∈ H, and Let us check that there exists an open neighborhood U1(bλ) ofbλ, and a number (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0;C(λk;z0)(cid:21)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)EH + z0;C(λk;z0)(cid:21) g(cid:13)(cid:13)(cid:13)(cid:13) ≥ (cid:13)(cid:13)(cid:13)(cid:13)(cid:20)EH + arbitrary sequence of open discs U n(bλ) with centres atbλ and radii 1 λkn ∈ U n(bλ), kn ∈ N, such that: z0;C(λkn ;z0)(cid:21) gn(cid:13)(cid:13)(cid:13)(cid:13) < (cid:13)(cid:13)(cid:13)(cid:13)(cid:20)EH + It is clear that gn are all non-zero. Denotebgn = gn z0;C(λkn ;z0)(cid:21)bgn(cid:13)(cid:13)(cid:13)(cid:13) < (cid:13)(cid:13)(cid:13)(cid:13)(cid:20)EH + Since λkn −bλ < 1 n, then limn→∞ λkn =bλ. We may write >(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0;C(bλ;z0)#bgn z0 −bλ 1 −bλz0 z0;C(bλ;z0)!bgn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) + z0 − λkn z0 −bλ 1 −bλz0 z0;C(bλ;z0)#bgn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≥(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) "EH + z0 −bλ 1 −bλz0 "EH + z0 − λkn 1 − λknz0 z0 − λkn 1 − λknz0 V + z0;C(λkn ;z0) − 1 n kgnk. , n ∈ N. Then kgnkH (229) (230) V + V + V + V + 1 n . 1 n 1 − λknz0 V + 109 z0 − λkn 1 − λknz0 V + z0;C(λkn ;z0) − −(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) V + z0;C(bλ;z0)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0 −bλ 1 −bλz0 z0;C(bλ;z0)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0 −bλ 1 −bλz0 V + z0;C(bλ;z0) V + V + V + z0;C(bλ;z0) − , L > 0, (231) V + z0;C(λkn ;z0) − z0 − λk 1 − λkz0 V + z0 − λk 1 − λkz0 V + z0 − λk 1 − λkz0 V + z0 − λkn 1 − λknz0 such that inequality (227) holds. We may write: for sufficiently large n, since the operator [EH + z0−bλ ] has a bounded 1−bλz0 inverse, defined on the whole H, and the norm in the right-hand side tends to zero. Passing to the limit in (231) as n → ∞, we obtain a contradiction. ≥ L −(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Thus, there exists an open neighborhood U1(bλ) of bλ, and a number K > 0 z0;C(bλ;z0)#−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)EH + ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0;C(λk;z0)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:20)EH + −"EH + z0;C(λk;z0)(cid:21)−1 z0 −bλ 1 −bλz0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0;C(λk;z0)(cid:21)−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0 −bλ 1 −bλz0 ∗(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) z0;C(bλ;z0)#−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) "EH + z0 −bλ 1 −bλz0 ="EH + z0;C(λk;z0)(cid:21)−1 z0 −bλ 1 −bλz0 k→∞(cid:0)EH − λkVC(λk;z0);z0(cid:1)−1 =(cid:16)EH −bλVC(bλ;z0);z0(cid:17)−1 λ∈D∪∆, λ→bλ(cid:0)EH − λVC(λ;z0);z0(cid:1)−1 =(cid:16)EH −bλVC(bλ;z0);z0(cid:17)−1 (232) (233) . (234) By the Inin formula for λ ∈ D we have: (cid:0)EH − λVC(λ;z0);z0(cid:1)−1 = Rλ. Thus, relation (234) shows that the function Rλ, λ ∈ D, admits a continuation on a set D ∪ ∆, and this continuation is continuous in the uniform operator topology. By (224),(225),(226),(232) we conclude that k→∞(cid:20)EH + z0;C(bλ;z0)#−1 V + Using (227) we conclude that z0 − λk 1 − λkz0 V + V + . u. − lim and therefore u. − lim u. − lim . , 110 Choose an arbitrary element h ∈ H and consider the following analytic function: f (λ) = fh(λ) = (Rλh, h), λ ∈ D. (235) The function f (λ) admits a continuous continuation on D ∪ ∆, which has the following form: Let us check that f (λ) =(cid:16)(cid:0)EH − λVC(λ;z0);z0(cid:1)−1 h, h(cid:17) , (cid:0)EH − λVC(λ;z0);z0(cid:1)−1 = EH −(cid:0)EH − λV ∗ C(λ;z0);z0(cid:1)−1 λ ∈ D ∪ ∆. Choose an arbitrary number λ ∈ ∆. By condition 2) we conclude that VC(λ;z0);z0 is unitary. Then , ∀λ ∈ ∆. (236) (EH − λV ∗ C(λ;z);z)(EH − λVC(λ;z);z) = EH − λVC(λ;z);z + EH − λV ∗ C(λ;z);z. (237) and from the right by To verify the last relation, it is enough to perform the multiplication in the left-hand side and simplify the obtained expression. Multiplying re- lation (237) from the left by (cid:16)EH − λV ∗ (cid:0)EH − λVC(λ;z0);z0(cid:1)−1 we easily get (236). C(λ;z0);z0(cid:17)−1 f (λ) =(cid:16)h,(cid:0)EH − λVC(λ;z0);z0(cid:1)−1 h(cid:17) =(cid:16)(cid:0)EH − λV ∗ We may write: = (h, h) −(cid:16)(cid:0)EH − λVC(λ;z0);z0(cid:1)−1 Ref (λ) = (h, h), 1 2 C(λ;z0);z0(cid:1)−1 h, h(cid:17) = (h, h) − f (λ); λ ∈ ∆. h, h(cid:17) (238) Denote g(λ) = if (λ) − i 2(h, h), λ ∈ D ∪ ∆. Then Img(λ) = 0. Thus, by the D ∪ ∆ ∪ De. Moreover, the following relation holds: Schwarc principle, g(λ) has an analytic continuation eg(λ) =egh(λ) on a set λ ∈ De. (239) Then the following function eg(λ) =eg(cid:18) 1 λ(cid:19), ieg(λ) + 1 2 1 ef (λ) = efh(λ) := (h, h), λ ∈ D ∪ ∆ ∪ De, is an analytic continuation of f (λ). Using (239) we get λ ∈ De. + (h, h)H = − (h, (EH − R∗ λ(cid:19) + (h, h), ef (λ) = −ef(cid:18) 1 h, h(cid:17)H 4(cid:16)efh+g(λ) − efh−g(λ) + iefh+ig(λ) − iefh−ig(λ)(cid:17) , λ ∈ D ∪ ∆ ∪ De. = (Rλh, h)H, h, g ∈ H, λ ∈ De. λ)h)H + (h, h)H 111 (240) (241) (242) By (183) we may write ef (λ) = −(cid:16)R 1 λ Set Rλ(h, g) = 1 Notice that Rλ(h, g) is an analytic function of λ in D∪∆∪De. From (235),(241) we conclude that Rλ(h, g) = (Rλh, g)H, h, g ∈ H, λ ∈ D ∪ De. (243) From (234) it is seen that Rbλ(h, g) = lim (Rλh, g)H λ∈D, λ→bλ = lim λ∈D, λ→bλ(cid:16)(cid:0)EH − λVC(λ;z0);z0(cid:1)−1 h, g(cid:17)H =(cid:18)(cid:16)EH −bλVC(bλ;z0);z0(cid:17)−1 h, g(cid:19)H Therefore the following operator-valued function bλ ∈ ∆. , h, g ∈ H, (244) Tλ =(cid:26) Rλ, (cid:0)EH − λVC(λ;z0);z0(cid:1)−1 λ ∈ D ∪ De λ ∈ ∆ , is a continuation of Rλ, and it is analytic with respect to the weak operator topology, and therefore with respect to the uniform operator topology, as well. ✷ Corollary 4.1 Theorem 4.2 remains valid for the choice z0 = 0. 112 Proof. Let V be a closed isometric operator in a Hilbert space H, and Rz(V ) be an arbitrary generalized resolvent of V . Let F (λ) = C(λ; 0) ∈ S(N0; N∞) correspond to Rz(V ) by Inin's formula (25) for z0 = 0, whic in this case becomes the Chumakin formula (16). Consider an arbitrary open arc ∆ of T. Choose an arbitrary point z0 ∈ D\{0}. Consider the following isometric operator: V = (V + z0EH )(EH + z0V )−1, D(V) = (EH + z0V )D(V ). Then V = (V − z0EH )(EH − z0V)−1 = Vz0. Recall that the generalized resolvents V and Vz0 are related by (29) and this correspondence is one-to-one. Let Rz(V) be the generalized resolvent which by (29) corresponds to the generalized resolvent Rz(Vz0) = Rz(V ). From (29) weee that Ret(Vz0) has a limit aset →et0 ∈ ∆, if and only if Reu(V) has a limit aseu →eu0 ∈ ∆1, where ∆1 =(eu : eu = et + z0 1 + z0et Thus, Ret(Vz0) admits a continuation by the continuity on Te ∪ ∆, if and only if Reu(V) admits a continuation by the continuity on Te ∪ ∆1. The limit values are connected by (29), as well. By (29) we see that the continuation Ret(Vz0) is analytic if and only if when the continuation Reu(V) is analytic. Thus, Ret(V ) = Ret(Vz0) admits an analytic continuation on Te ∪ ∆, if and only if Reu(V) admits an analytic continuation on Te ∪ ∆1. , et ∈ ∆) . By Theorem 4.2, Reu(V) admits an analytic continuation on Te ∪ ∆1 if and only if 1) C(λ; z0) has a continuation on a set D ∪ ∆1, which is continuous in the uniform operator topology; 2) The continued function C(λ; z0) maps isometrically Nz0 on the whole N 1 z0 , for all points λ ∈ ∆1; 3) The operator (EH − λVC(λ;z0);z0)−1 exists and it is defined on the whole H, for all points λ ∈ ∆1, where C(λ; z0) ∈ S(Nz0; N 1 z0 ) corresponds to Rz(V) by Inin's formula. Recall that C(λ; z0) is connected with F (et) in the following way: u ∈ Te. 1 − z0eu(cid:19) , C(eu) = F(cid:18) eu − z0 Using this relation we easily obtain that condition 1) is equivalent to the following condition: 113 uniform operator topology; and condition 2) is equivalent to the following condition: 1') F (et) has a continuation on a set D ∪ ∆, which is continuous in the 2') The continued function F (et) maps isometrically N0 on the whole N∞, for allet ∈ ∆. (cid:20)EH − From (223) it follows that the operator (EH − λVC(λ;z0);z0)−1 exists and it is defined on the whole H, for all λ ∈ ∆1, if and only if the operator (Vz0 ⊕ C(λ; z0))(cid:21)−1 =(cid:2)EH −et(cid:0)Vz0 ⊕ F (et)(cid:1)(cid:3)−1 , λ − z0 1 − λz0 exists and it is defined on the whole H, for allet ∈ ∆. ✷ 4.2 Some decompositions of a Hilbert space in a direct sum of subspaces. In the sequel we shall use some decompositions of a Hilbert space related to a given isometric operator. Theorem 4.3 Let V be a closed isometric operator in a Hilbert space H. Let ζ ∈ T, and ζ −1 is a point of the regular type of V . Then D(V ) ∔ Nζ(V ) = H; R(V ) ∔ Nζ(V ) = H. Proof. At first, assume that ζ = 1. Check that kV f + gkH = kf + gkH, f ∈ D(V ), g ∈ N1. (245) In fact, we may write: kV f + gk2 H = (V f, V f )H + (V f, g)H + (g, V f )H + (g, g)H = kf k2 H + (V f, g)H + (g, V f )H + kgk2 H. Since 0 = ((E − V )f, g)H = (f, g)H − (V f, g)H, we get (V f, g) = (f, g), f ∈ D(V ), g ∈ N1, (246) and therefore kV f + gk2 H = kf k2 H + (f, g)H + (g, f )H + kgk2 H = kf + gk2 H. 114 Consider the following operator: U(f + g) = V f + g, f ∈ D(V ), g ∈ N1. Let us check that this operator is well-defined on D(U) = D(V ) + N1. Sup- pose that an element h ∈ D(U) admits two representations: h = f1 + g1 = f2 + g2, f1, f2 ∈ D(V ), g1, g2 ∈ N1. By (245) we may write: kV f1+g1−(V f2+g2)k2 H = kV (f1−f2)+(g1−g2)k2 H = kf1−f2+g1−g2k2 H = 0. Thus, U is well-defined. Moreover, it is easily seen that U is linear. Us- ing (246) we write: (U(f +g), U(ef +eg)) = (V f +g, V ef +eg) = (V f, V ef )+(V f,eg)+(g, V ef )+(g,eg) = (f,ef ) + (f,eg) + (g,ef ) + (g,eg) = (f + g,ef +eg), for arbitrary f,ef ∈ D(V ), g,eg ∈ N1. Therefore U is isometric. Suppose that there exists an element h ∈ H, h 6= 0, h ∈ D(V )∩N1. Then 0 = U0 = U(h + (−h)) = V h − h = (V − EH )h, what contradicts to the fact that the point 1 is a point of the regular type of V . Therefore D(V ) ∩ N1 = {0}. (247) Observe that we a priori do not know, if D(U) is a closed manifold. Consider the following operator: W = US, where the manifold S is given by the following equality: S = {h ∈ D(U) : h ⊥ N1} = D(U) ∩ M1. Thus, W is an isometric operator with the domain D(W ) = D(U) ∩ M1. Choose an arbitrary element g ∈ D(U). Let g = gM1 + gN1, gM1 ∈ M1, gN1 ∈ N1 ⊆ D(U). Then gM1 = P H M1g ∈ D(U), gM1 ⊥ N1. Therefore gM1 ∈ D(W ); P H M1D(U) ⊆ D(W ). On the other hand, choose an arbitrary element h ∈ D(W ). Then h ∈ D(U) ∩ M1, and therefore h = P H M1D(U). Consequently, we get M1h ∈ P H D(W ) = P H M1D(U) = P H M1(D(V ) + N1) = P H M1D(V ) ⊆ M1. 115 Choose an arbitrary element f ∈ D(V ). Let f = fM1 + fN1, fM1 ∈ M1, fN1 ∈ N1. Then f − fN1 ∈ D(U), and f − fN1 ⊥ N1, i.e. f − fN1 ∈ D(W ). We may write (W − EH )(f − fN1) = (U − EH)(f − fN1) = U(f − fN1) − f + fN1 = V f − f ; (W − EH )D(W ) ⊇ (V − EH)D(V ) = M1. (248) On the other hand, choose an arbitrary element w ∈ D(W ), w = wD(V )+wN1, wD(V ) ∈ D(V ), wN1 ∈ N1. Since w ⊥ N1, then w = P H M1w = P H M1wD(V ). (249) We may write (W − EH)w = Uw − w = V wD(V ) + wN1 − wD(V ) − wN1 = (V − EH)wD(V ); (W − EH )D(W ) ⊆ (V − EH)D(V ) = M1. From the latter relation and relation (248) it follows that (W − EH )D(W ) = (V − EH )D(V ) = M1. (250) Moreover, if (W − EH)w = 0, then (V − EH )wD(V ) = 0; and therefore wD(V ) = 0. By (249) it implies w = 0. Consequently, there exists the inverse (W − EH)−1. Using (250) we get D(W ) = (W − EH)−1M1. Since D(W ) ⊆ M1, then by (250) we may write: W w = (W − EH )w + w ∈ M1; D(W ) ⊆ M1, W D(W ) ⊆ M1. (251) Consider the closure W of W with the domain D(W ). By (251) we see that D(W ) ⊆ M1, W D(W ) ⊆ M1. Then On the other hand, by (250) we get (W − EH)D(W ) ⊆ M1. (W − EH)D(W ) ⊇ (W − EH )D(W ) = (W − EH)D(W ) = M1. We conclude that (W − EH)D(W ) = M1. 116 (252) Let us check that there exists the inverse (W − EH)−1. Suppose to the contrary that there exists an element h ∈ D(W ), h 6= 0, such that (W − EH )h = 0. By the definition of the closure there exists a sequence of elements hn ∈ D(W ), n ∈ N, tending to h, and W hn → W h, as n → ∞. Then (W − EH)hn → W h − h = 0, as n → ∞. Let hn = h1;n + h2;n, where h1;n ∈ D(V ), h2;n ∈ N1, n ∈ N. Then (W − EH )hn = Uhn − hn = V h1;n + h2;n − h1;n − h2;n = (V − EH)h1;n. Therefore (V − EH)h1;n → 0, as n → ∞. Since (V − EH) has a bounded inverse, then h1;n → 0, as n → ∞. Then hn = P H M1h1;n → 0, as n → ∞. Consequently, we get h = 0, what contradicts to our assumption. M1hn = P H Thus, there exists the inverse (W −EH )−1 ⊇ (W −EH )−1. By (252),(250), we get and therefore (W − EH )−1 = (W − EH)−1, W = W. hus, W may be considered as a closed isometric operator in a Hilbert space M1. The operator (W − EH )−1 is closed and it is defined on the whole M1. Therefore (W − EH )−1 is bounded. This means that the point 1 is a regular point of W . Therefore W is a unitary operator in M1. In particular, this implies that D(W ) = R(W ) = M1. By the definition of W we get D(W ) = M1 ⊆ D(U), and UM1 = W . On the other hand, we have UN1 = EN1. Therefore D(U) = H and U = W ⊕ EN1. So, U is a unitary operator. Therefore D(U) = R(U) = H, what implies D(V ) + N1 = H, R(V ) + N1 = H. (253) The first sum is direct by (247). Suppose that h ∈ R(V )∩N1. Then h = V f , f ∈ D(V ), and we may write: 0 = V f + (−h) = U(f + (−h)). Since U is unitary, we get f = h = V f , (V − EH )f = 0, and therefore f = 0, and h = 0. Thus, the second sum in (253) is direct, as well. So, we proved the theorem in the case ζ = 1. In the general case we can apply the proved part of the theorem to the operator bV := ζV . ✷ 117 Corollary 4.2 In the conditions of Theorem 4.3 the following decomposi- tions hold: (H ⊖ D(V )) ∔ Mζ = H; (H ⊖ R(V )) ∔ Mζ = H. Proof. The proof is based on the following lemma. Lemma 4.2 Let M1 and M2 be two subspaces of a Hilbert space H such that M1 ∩ M2 = {0}, and Then M1 ∔ M2 = H. (H ⊖ M1) ∔ (H ⊖ M2) = H. Proof. Suppose that an element h ∈ H is such that h ∈ ((H ⊖ M1) ∩ (H ⊖ M2)). Then h ⊥ M1, h ⊥ M2, and therefore h ⊥ (M1 + M2), h ⊥ H, h = 0. Suppose that an element g ∈ H satisfies the following condition: g ⊥ ((H ⊖ M1) ∔ (H ⊖ M2)). Then g ∈ M1, g ∈ M2, and therefore g = 0. ✷ Applying the last lemma with M1 = D(V ), M2 = Nζ, and M1 = R(V ), M2 = Nζ, we complete the proof of the corollary. ✷ 4.3 Isometric operators with lacunas in a spectrum. Let V be a closed isometric operator in a Hilbert space H. An open arc ∆ of T is said to be a gap in the spectrum of the isometric operator V , if all points of ∆ are points of the regular type of V . Above we considered conditions when for a prescribed open arc ∆ ⊆ T and a spectrak measure F(δ) there holds: F(∆) = 0. We shall see later that a necessary condition of the existence at least one such spectral function is that ∆ is a gap in the spectrum of V . Theorem 4.2 provides conditions, extracting parameters C(λ; z0) in the Inin formula, which generate generalized resolvents (and therefore the correspond- ing spectral functions), which have an analytic continuation on D ∪ De ∪ ∆ (we can apply this theorem with ∆ instesd of ∆). By Proposition 4.1 for these generalized resolvents there corresond spectral functions F(δ) such that: F(∆) = 0. Moreover, the extracted class of parameters C(λ; z0) by condi- tion 3) of Theorem 4.2, depends on the operator V . Our aim here will be to find instead of condition 3) other conditions, which will not use V . Lemma 4.3 Let V be a closed isometric operator in a Hilbert space H, and ζ ∈ T. Then V P H M0(V )f = ζ −1P H M∞(V )f, ∀f ∈ Nζ(V ), (254) 118 (255) (256) and therefore ∀f ∈ Nζ(V ); ∀f ∈ Nζ(V ). (cid:13)(cid:13)P H M0(V )f(cid:13)(cid:13) =(cid:13)(cid:13)P H (cid:13)(cid:13)P H N0(V )f(cid:13)(cid:13) =(cid:13)(cid:13)P H D(V )f, V u(cid:1)H (cid:0)ζ −1f − V P H M∞(V )f(cid:13)(cid:13) , N∞(V )f(cid:13)(cid:13) , = ζ −1(f, V u)H −(cid:0)P H = (f, ζV u)H − (f, u)H = (f, (ζV − EH)u)H = 0. D(V )f, u(cid:1)H Proof. u ∈ D(V ) = M0 we may write: Choose an arbitrary element f ∈ Nζ. For an arbitrary element Therefore (ζ −1f − V P H M∞ to this element we get (254). Relation (255) is obvious, since V is isometric. Then it easily follows (256). ✷ M0f ) ⊥ M∞. Applying P H Lemma 4.4 Let V be a closed isometric operator in a Hilbert space H, and C be a linear bounded operator in H, D(C) = N0(V ), R(C) ⊆ N∞(V ). Let ζ ∈ T be such number that ζ −1 be an eigenvalue of the operator V + 0;C = V ⊕C. If f ∈ H, f 6= 0, is an eigenvector of V + 0;C, corresponding to ζ −1, then f ∈ Nζ(V ), and CP H N0(V )f = ζ −1P H N∞(V )f. (257) Proof. ζ −1 ∈ T: Let f be an eigenvalue of V + 0;C, corresponding to the eigenvalue (V ⊕ C)f = V P H M0f + CP H M∞f + P H N0f = ζ −1(cid:0)P H N∞f(cid:1) . By the orthogonality of summands, the last relation is equivalent to the following two conditions: V P H M0f = ζ −1P H M∞f ; (258) N0f = ζ −1P H and (257) follows. Relation (258) implies P H V P H CP H M0f ) ⊥ M∞. For an arbitrary element u ∈ D(V ) we may write: N∞f, M∞(ζ −1f − V P H (259) M0f ) = 0; (ζ −1f − 0 =(cid:0)ζ −1f − V P H M0f, V u(cid:1)H = ζ −1(f, V u)H −(cid:0)P H = (f, ζV u)H − (f, u)H = (f, (ζV − EH)u)H. M0f, u(cid:1)H Therefore f ∈ Nζ. ✷ For an arbitrary number ζ ∈ T, define an operator Wζ in the following way: 119 WζP H N0f = ζ −1P H N∞f, f ∈ Nζ, (260) with the domain D(Wζ) = P H If an element g ∈ D(Wζ) admits two representations: g = P H f1, f2 ∈ Nζ, then P H and therefore the definition is correct. Observe that Wζ is linear and N0Nζ. Let us check that this definition is correct. N0f2, N∞(f1 − f2) = 0, N0(f1 − f2) = 0. By (256) this implies P H N0f1 = P H (cid:13)(cid:13)WζP H N0f(cid:13)(cid:13)H =(cid:13)(cid:13)P H N∞f(cid:13)(cid:13)H =(cid:13)(cid:13)P H N0f(cid:13)(cid:13)H . Thus, the operator Wζ is isometric. Notice that R(Wζ) = P H Set N∞Nζ. S = P H N0Nζ , Q = P H N∞Nζ . In what follows, we shall assume that ζ −1 is a point of the regular type of V . Let us check that in this case the operators S and Q are invertible. Suppose to the contrary that there exists an element f ∈ Nζ, f 6= 0: Sf = P H N0f = 0. Then f = P H M0f 6= 0. By Theorem 4.3 we get f ∈ M0 ∩ Nζ = {0}. We come to a contradiction. In a similar manner, suppose that there exists an element g ∈ Nζ, g 6= 0: Qf = P H M∞g 6= 0. By Theorem 4.3 we obtain that g ∈ M∞ ∩ Nζ = {0}. We come to a contradiction, as well. By Theorem 4.3 we get N∞f = 0. Then g = P H P H N0Nζ = N0; P H N∞Nζ = N∞. Thus, the operators S−1 and Q−1 are closed and they are defined on subspaces N0 and N∞, respectively. Therefore S−1 and Q−1 are bounded. From (260) we see that D(Wζ) = N0, R(Wζ) = N∞, and Wζ = ζ −1QS−1. (261) Theorem 4.4 Let V be a closed isometric operator in a Hilbert space H, and C be a linear bounded operator in H, D(C) = N0(V ), R(C) ⊆ N∞(V ). Let ζ ∈ T, and ζ −1 be a point of the regular type of V . The point ζ −1 is an eigenvalue of V + 0;C = V ⊕ C, if and only if the following condition holds: (C − Wζ)g = 0, g ∈ N0(V ), g 6= 0. (262) Proof. Necessity. Since ζ −1 is an eigenvalue of V + we obtain that there exists f ∈ Nζ, f 6= 0, such that 0;C = V ⊕ C, then by 4.4 CP H N0f = ζ −1P H N∞f. (263) 120 N0f = Sf . Since S is invertible, then g 6= 0. Comparing the last relation with the definition of Wζ we see that CP H WζP H N0f . Set g = P H Sufficiency. From (262) we get (263) with f := S−1g. By Lemma 4.3 we see that relations (258),(259) hold. This is equivalent, as we have seen before (258), that ζ −1 is an eigenvalue of V + 0;C = V ⊕ C, corresponding to eigenvector f . ✷ N0f = Theorem 4.5 Let V be a closed isometric operator in a Hilbert space H, and C be a linear bounded operator in H, D(C) = N0(V ), R(C) ⊆ N∞(V ). Let ζ ∈ T, and ζ −1 is a point of the regular type of V . The following relation holds if and only if the following relations hold: R(cid:0)V + 0;C − ζ −1EH(cid:1) = H, (C − Wζ) N0(V ) = N∞(V ); P H M∞Mζ(V ) = M∞(V ). (264) (265) (266) Proof. Necessity. Choose an arbitrary element h ∈ N∞. By (264) there exists an element x ∈ H such that (cid:0)V + 0;C − ζ −1EH(cid:1) x = (V ⊕ C)x − ζ −1x = h. For an arbitrary u ∈ D(V ) we may write: (267) (x, (EH − ζV )u)H = (x, (V −1 − ζEH)V u)H = (x, ((V + 0;C)∗ − ζEH)V u)H 0;C − ζ −1EH)x, V u)H = (h, V u)H = 0, and therefore x ∈ Nζ. Set g = Sx ∈ N0. Using (261) write: = ((V + (C − Wζ)g = CSx − WζSx = CSx − ζ −1Qx. Since h ∈ N∞, we apply P H N∞ to equality (267) and get CP H N0x − ζ −1P H N∞x = h; Therefore and relation (265) holds. CSx − ζ −1Qx = h. (C − Wζ)g = h, 121 that Choose an arbitrary elementbh ∈ M∞. By (264) there existsbx ∈ H such (cid:0)V + 0;C − ζ −1EH(cid:1)bx = (V ⊕ C)bx − ζ −1bx =bh. The last equality is equivalent to the following two equalities, obtained by an application of projectors P H M∞ and P H N∞: By Theorem 4.3 we may write: V P H CP H M0bx − ζ −1P H N0bx − ζ −1P H M∞bx =bh; N∞bx = 0. (268) (269) xD(V ) ∈ D(V ), xNζ ∈ Nζ. Substituting this decomposition in (269) we get bx = xD(V ) + xNζ , CP H N0xNζ − ζ −1P H N∞xNζ − ζ −1P H N∞xD(V ) = 0; (C − Wζ)P H N0xNζ = ζ −1P H N∞xD(V ). On the other hand, substituting the decomposition in (268) we obtain that V xD(V ) + V P H M0xNζ − ζ −1P H M∞xD(V ) − ζ −1P H M∞xNζ =bh; where we used Lemma 4.3. Then V xD(V ) − ζ −1P H M∞xD(V ) =bh, M∞(V − ζ −1EH )xD(V ) =bh, P H and relation (266) follows directly. Sufficiency. Choose an arbitrary element h ∈ H, h = h1 + h2, h1 ∈ M∞, h2 ∈ N∞. By (265) there exists an element g ∈ N0 such that (C − Wζ) g = Cg − Wζg = h2. Set x = S−1g ∈ Nζ. Then CSx − ζ −1Qx = h2; N∞(V ⊕ C)P H P H N0x − ζ −1P H N∞x = h2; P H N∞(cid:0)V + 0;Cx − ζ −1x(cid:1) = h2. By Lemma 4.3 we may write: P H M∞(cid:0)V + 0;Cx − ζ −1x(cid:1) = V P H M0x − ζ −1P H M∞x = 0. 122 (270) (271) (272) (273) (274) (275) M∞exNζ By (266) there exists an element w ∈ Mζ such that (cid:0)V + 0;C − ζ −1EH(cid:1) x = h2. P H M∞w = h1. Therefore Let Then w = (V − ζ −1EH)exD(V ), VexD(V ) − ζ −1P H (C − Wζ) r = ζ −1P H By (265) there exists r ∈ N0 such that SetexNζ := S−1r ∈ Nζ. Then (C − Wζ)P H CP H N0exNζ − ζ −1P H Setex =exD(V ) +exNζ . By (272) we get CP H Using relation (271) and Lemma 4.3 we write: h1 = VexD(V ) − ζ −1P H = V P H Adding relations (273) and (274) we get exD(V ) ∈ D(V ). M∞exD(V ) = h1. N∞exD(V ). N0exNζ = ζ −1P H N∞exD(V ); N∞exNζ − ζ −1P H N∞exD(V ) = 0. N0ex − ζ −1P H N∞ex = 0. M0exNζ − ζ −1P H M∞exD(V ) + V P H M0ex − ζ −1P H M∞ex. 0;C − ζ −1EH(cid:1)ex = h1. (V ⊕ C)ex − ζ −1ex =(cid:0)V + Adding relations (270) and (275) we conclude that relation (264) holds. ✷ Theorem 4.6 Let V be a closed isometric operator in a Hilbert space H, and ∆ be an open arc of T such that ζ −1 is a point of the regular type of V , ∀ζ ∈ ∆. Suppose that the following condition holds: P H M∞(V )Mζ(V ) = M∞(V ), ∀ζ ∈ ∆. (276) Let Rz = Rz(V ) be an arbitrary generalized resolvent of V , and C(λ; 0) ∈ S(D; N0, N∞) corresponds to Rz(V ) by Inin's formula (25). The operator- valued function Rz(V ) has an analytic continuation on a set D ∪ De ∪ ∆ if and only if the following conditions hold: 123 1) C(λ; 0) admits a continuation on a set D ∪ ∆ and this continuation is continuous in the unform operator topology; 2) The continued function C(λ; 0) maps isometrically N0(V ) on the whole N∞(V ), for all λ ∈ ∆; 3) The operator C(λ; 0) − Wλ is invertible for all λ ∈ ∆, and (C(λ; 0) − Wλ)N0(V ) = N∞(V ), ∀λ ∈ ∆. (277) Proof. Necessity. Suppose that Rz(V ) has an analytic continuation on a set D ∪ De ∪ ∆. By Corollary 4.1 we conclude that conditions 1) and 2) hold, and the operator (EH − λVC(λ;0);0)−1 = − 1 λEH )−1 exists and it is defined on the whole H, for all λ ∈ ∆. By Theorem 4.4 we obtain that the operator C(λ; 0) − Wλ is invertible for all λ ∈ ∆. By Theorem 4.5 we obtain that relation (277) holds. λ((V ⊕ C(λ; 0)) − 1 Sufficiency. Suppose that conditions 1)-3) hold. By Theorem 4.5 we obtain that R((V ⊕ C(λ; 0)) − 1 λEH) = R(EH − λVC(λ;0);0) = H. By The- orem 4.4 we see that the operator (V ⊕ C(λ; 0)) − 1 λEH is invertible. By Corollary 4.1 we obtain that Rz(V ) admits an analytic continuation on a set D ∪ De ∪ ∆. ✷ Remark 4.1 By Corollary 4.1, if Rz(V ) has an analytic continuation on a set D ∪ De ∪ ∆, then ((V ⊕ C(λ; 0)) − 1 λEH)−1 exists and it is bounded. Therefore the point λ−1, λ ∈ ∆, are points of the regular type of V . On the other hand, by Theorem 4.5, in this case condition (276) holds. Thus, by Proposition 4.1 condition (276) and a condition that points λ−1, λ ∈ ∆, are points of the regular type of V , are necessary for the existence of a spectral function F of V such that F(∆) = 0. Consequently, these conditions do not imply on the generality of Theorem 4.6. Now we shall obtain an analogous result but the corresponding conditions will be put on the parameter C(λ; z0) of Inin's formula for an arbitrary z0 ∈ D. Before to do that, we shall prove the followng simple proposition: Proposition 4.2 Let V be a closed isometric operator in a Hilbert space H, and z0 ∈ D be a fixed number. For an arbitrary ζ ∈ T the following two conditions are equivalent: (i) ζ −1 ∈ ρr(V ); (ii) 1−ζz0 ζ−z0 ∈ ρr(Vz0). 124 Proof. (i) ⇒ (ii). We may write: Vz0− 1 − ζz0 ζ − z0 EH = (V −z0EH)(EH −z0V )−1− 1 − ζz0 ζ − z0 (EH −z0V )(EH −z0V )−1 = ζ(1 − z02) ζ − z0 (V − 1 ζ EH)(EH − z0V )−1. The operator in the right-hand side has a bounded inverse, defined on (V − ζ −1EH)D(V ). (ii) ⇒ (i). We write: V − 1 ζ EH = (Vz0 + z0EH)(EH + z0Vz0)−1 − (EH + z0Vz0)(EH + z0Vz0)−1 1 ζ = ζ − z0 ζ (cid:18)Vz0 − 1 − ζz0 ζ − z0 EH(cid:19) (EH + z0Vz0)−1, and therefore the operator on the right has a bounded inverse, defined on (Vz0 − 1−ζz0 ζ−z0 EH)D(Vz0). ✷ Theorem 4.7 Let V be a closed isometric operator in a Hilbert space H, and ∆ be an open arc of T such that ζ −1 is a point of the regular type of V , ∀ζ ∈ ∆. Let z0 ∈ D be an arbitrary fixed point, and suppose that the following condition holds: P H M 1 z0 (V )M z0+ζ 1+ζz0 (V ) = M 1 z0 (V ), ∀ζ ∈ ∆. (278) Let Rz = Rz(V ) be an arbitrary generalized resolvent of V , and C(λ; z0) ∈ S(D; Nz0, N 1 ) corresponds to Rz(V ) by the Inin formula (25). The operator- z0 valued function Rz(V ) has an analytic continuation on a set D ∪ De ∪ ∆ if and only if the following conditions hold: 1) C(λ; z0) admits a continuation on a set D ∪ ∆, which is continuous in the uniform operator topology; 2) The continued function C(λ; z0) maps isometrically Nz0(V ) on the whole N 1 z0 (V ), for all λ ∈ ∆; 3) The operator C(λ; z0) − Wλ;z0 is invertible for all λ ∈ ∆, and (C(λ; z0) − Wλ;z0)Nz0(V ) = N 1 z0 (V ), ∀λ ∈ ∆. Here the operator-valued function Wλ;z0 is defined by the following equal- ity: Wλ;z0P H Nz0 (V )f = 1 − z0λ λ − z0 P H N 1 z0 (V )f, f ∈ Nλ(V ), λ ∈ T. Proof. At first, we notice that in the case z0 = 0 this theorem coincides with Theorem 4.6. Thus, we can now assume that z0 ∈ D\{0}. 125 Suppose that Rz(V ) admits an analytic continuation on a set D ∪ De ∪ ∆. Recall that the generalized resolvent Rz(V ) is connected with the general- ized resolvent Rz(Vz0) of Vz0 by (29) and this correspondence is one-to-one. Therefore Rz(Vz0) admits an analytic continuation on a set Te ∪ ∆1, where ∆1 =(cid:26)et : et = λ − z0 1 − z0λ , λ ∈ ∆(cid:27) . By Proposition 4.2, pointset−1,et ∈ ∆1, are points of the regular type of the operator Vz0. Moreover, relation (276), written for Vz0 with ζ ∈ ∆1, coincides with relation (278). We can apply Theorem 4.6 to Vz0 and an open arc ∆1. Then if we rewrite conditions 1)-3) of that theorem in terms of C(λ; z0), using the one-to-one correspondence between C(λ; z0) for V , and C(λ; 0) for Vz0, we easily obtain conditions 1)-3) of the theorem. On the other hand, let conditions 1)-3) be satisfied. Then conditions of Theorem 4.6 for Vz0 hold. Therefore the following function Rz(Vz0) admits an analytic continuation on Te ∪ ∆1. Consequently, the function Rz(V ) admits an analytic continuation on D ∪ De ∪ ∆. ✷ 4.4 Spectral functions of a symmetric operator having a constant value on an open real interval. Now we shall consider symmetric (not necessarily densely defined) operators in a Hilbert space and obtain simlar results for them as for the isometric operators above. The next proposition is an analogue of Proposition 4.1. Proposition 4.3 Let A be a closed symmetric operator in a Hilbert space H, and E(δ), δ ∈ B(R), be its spectral measure. The following two conditions are equivalent: (i) E(∆) = 0, for an open (finite or infinite) interval ∆ ⊆ R; (ii) The generalized resolvent Rz(A), corresponding to the spectral measure E(δ), admits an analytic continuation on a set Re ∪ ∆, for an open (finite or infinite) real interval ∆ ⊆ R. Proof. At first, suppose that the interval ∆ is finite. (i)⇒(ii). In this case relation (35) has the following form: (Rλh, g)H =ZR 1 t − λ d(E(·)h, g)H =ZR 1 t − λ d(Eth, g)H, ∀h, g ∈ H. (279) 126 Choose an arbitrary number z0 ∈ ∆. Since the function 1 t−z0 and bounded on R\∆, then there exists an integral is continuous Iz0(h, g) :=ZR\∆ 1 t − z0 d(E(·)h, g)H. Then (Rzh, h)H − Iz0(h, h) = z − z0(cid:12)(cid:12)(cid:12)(cid:12)ZR\∆ 1 ≤ z − z0ZT\∆ t − zt − z0 1 (t − z)(t − z0) d(E(·)h, h)H(cid:12)(cid:12)(cid:12)(cid:12) d(E(·)h, h)H, z ∈ Re. There exists a neighborhood U(z0) of the point z0 such that z −t ≥ M1 > 0, ∀t ∈ R\∆, ∀z ∈ U(z0). Therefore the integral in the latter relation is bounded in the neighborhood U(z0). We obtain that (Rzh, h)H → Iz0(h, h), z ∈ Re, z → z0, ∀h ∈ H. Using the properties of sesquilinear forms we conclude that (Rzh, g)H → Iz0(h, g), z ∈ Re, z → z0, ∀h, g ∈ H. Set We may write Rez := w. − lim z∈Re, z→ez Rz, (cid:18) 1 z − z0 (Rz − Rz0)h, h(cid:19)H =ZR\∆ ∀ez ∈ ∆. 1 (t − z)(t − z0) d(E(·)h, h)H, z ∈ U(z0), h ∈ H. The function under the integral sign is bounded in U(z0), and it tends to (t−z0)2 . By the Lebesgue theorem we get 1 and therefore z − z0 lim z→z0(cid:18) 1 z→z0(cid:18) 1 lim z − z0 (Rz − Rz0)h, h(cid:19)H (Rz − Rz0)h, g(cid:19)H =ZR\∆ =ZR\∆ 1 (t − z0)2 d(E(·)h, h)H; 1 (t − z0)2 d(E(·)h, g)H, for h, g ∈ H. Consequently, there exists the derivative of Rz at z = z0. 127 (ii)⇒(i). Choose an arbitrary element h ∈ H, and consider the following function σh(t) := (Eth, h)H, t ∈ R, where Et is a left-continuous spectral function of A, corresponding to the spectral measure E(δ). Consider an arbitrary interval [t1, t2] ⊂ ∆. Assume that t1 and t2 are points of the continuity of the function Et. By the Stieltjes-Perron inversion formula : σh(t2) − σh(t1) = lim y→+0Z t2 t1 Im {(Rx+iyh, h)H} dx. If {zn}∞ x ∈ [t1, t2], then n=1, zn ∈ Re, is an arbitrary sequence, tending to some real point zn → x = x, n → ∞. By the analyticity of the generalized resolvent in a neighborhood of x, we may write: Rzk → Rx, R∗ zk → R∗ x, k → ∞. By property (36) we get Therefore R∗ zk = Rzk → Rx, k → ∞. R∗ x = Rx, x ∈ [t1, t2]. Using the last equality we write: Im {(Rx+iyh, h)H} → Im {(Rxh, h)H} = 0, y → +0, x ∈ [t1, t2].. Since the generalized resolvent is analytic in [t1, t2], then the function Im {(Rx+iyh, h)H} is continuous and bounded on [t1, t2]. By the Lebesgue theorem we get σh(t1) = σh(t2). Since points were arbitrary it follows that the function σh(t) = (Eth, h)H is constant on ∆. Since h was arbitrary,using properties of sesquilinear forms we obtain that Et is constant. Consider the case of an infinite ∆. In this case we can represent ∆ as a countable union (not necessarily disjunct) finite open real intervals. Applying the proved part of the proposition for the finite intervals we easily obtain the required statements. We shall also need to use the σ-additivity of the orthogonal spectral mesures. ✷ The following auxilliary peoposition is an analog of Proposition 4.2. Proposition 4.4 Let A be a closed symmetric operator in a Hilbert space H, and λ0 ∈ Re be a fixed number. For an arbitrary λ ∈ R the following two conditions are equivalent: 128 (i) λ ∈ ρr(A); (ii) λ−λ0 λ−λ0 ∈ ρr(Uλ0(A)), Uλ0(A) - A. Proof. (i) ⇒ (ii). We may write: Uλ0(A) − λ − λ0 λ − λ0 EH = EH + (λ0 − λ0)(A − λ0EH )−1 − λ − λ0 λ − λ0 EH = − λ0 − λ0 λ − λ0 (A − λEH )(A − λ0EH)−1. The operator on the right-hand side has the bounded inverse: − λ − λ0 λ0 − λ0 (A−λ0EH )(A−λEH)−1 = − defined on Mλ(A). (ii) ⇒ (i). The following relation holds: λ − λ0 λ0 − λ0(cid:0)EH + (λ − λ0)(A − λEH )−1(cid:1) , A − λEH = λ0EH + (λ0 − λ0)(Uλ0(A) − EH )−1 − λEH = (λ0 − λ)(Uλ0(A) − λ − λ0 λ − λ0 EH)(Uλ0(A) − EH )−1. The operator on the right has the bounded inverse: 1 λ0 − λ (Uλ0(A) − EH )(Uλ0(A) − λ − λ0 λ − λ0 EH )−1, defined on (Uλ0(A) − λ−λ0 λ−λ0 EH )D(Uλ0(A)). ✷ We shall need one more auxilliary proposition which is an addition to Theorem 2.6 Proposition 4.5 Let A be a closed symmetric operator in a Hilbert space H, and z ∈ Re be a fixed point. Let ∆ ⊆ R be a (finite or infinite) open interval, and Rs;λ(A) be a generalized resolvent of A. The following two conditions are equivalent: (i) The generalized resolvent Rs;λ(A) admits an analytic continuation on a set Re ∪ ∆; (ii) The generalized resolvent Ru;ζ(Uz(A)), corresponding to the generalized resolvent Rs;λ(A) by the one-to-one correspondence (40) from Theo- rem 2.6, has an analytic continuation on Te ∪e∆, where , λ ∈ ∆(cid:27) . e∆ =(cid:26)ζ ∈ T : ζ = λ − z λ − z (280) If condition (i) is satisfied, the the following statements are true: the interval ∆ consists of points of the regular type of A; P H Mz(A)Mλ(A) = Mz(A), ∀λ ∈ ∆, Moreover, condition (281) is equivalent to the following condition: the interval e∆ consists of points ζ such that ζ −1 is a point of the regular type of Uz(A); and condition (282) is equivalent to the following condition: P H M∞(Uz(A))Mζ(Uz(A)) = M∞(Uz(A)), ∀ζ ∈ e∆. 129 (281) (282) (283) (284) Proof. (i)⇒(ii). Notice that the linear fractional transformation ζ = λ−z λ−z maps the real line on the unit circle. Relation (40) in terms of ζ reads Ru;ζ(Uz) = 1 1 − ζ EH + (z − z) ζ (1 − ζ)2 Rs; z−zζ 1−ζ (A), ζ ∈ Te\{0}. (285) 1−bζ ∈ ∆. Consider an arbi- trary sequence {ζk}∞ λk := z−zζk 1−ζk of this sequence and passing to the limit we conclude that the generalized Choose an arbitrary point bζ ∈ e∆. Set bλ = z−z bζ k=1, ζk ∈ Te\{0}, tending to bζ. The sequence {λk}∞ ∈ Re\{z, z}, tends to bλ. Writing relation (285) for elements resolvent Ru;ζ(Uz) can be continued by the continuity on e∆. For the limit values relation (285) holds. From the analyticity of thr right-hand side of this relation it follows the analyticity of the continued generalized resolvent Ru;ζ(Uz). k=1, (ii)⇒(i). Let us express from (40) the generalized resolvent of A: Rs;λ(A) = z − z (λ − z)(λ − z)(cid:18)Ru; λ−z λ−z (Uz) − λ − z z − z EH(cid:19) , λ ∈ Re\{z, z}. (286) Proceeding in a similar manner as inthe proof of the previous assertion we show that Rs;λ(A) admits a continuation by the continuity on ∆, and this continuation is analytic. Let condition (i) be satisfied. By the proved part condition (ii) holds, as well. As it was noticed in Remark 4.1, in this case there hold conditions (283) and (284). It is enough to check the equivalence of relations (281) and (283), and also of relations (282) and (284). The first equivalence follows from the 130 definition of e∆ and by Proposition 4.4. The second equivalence follows from relation: M∞(Uz(A)) = Mz(A), Mζ(Uz(A)) = M z−zζ (A), 1−ζ ζ ∈ e∆. ✷ Proposition 4.6 Let A be a closed symmetric operator in a Hilbert space H, and z ∈ Re be an arbitrary fixed point. Let ∆ ⊆ R be a (finite or infinite) open interval, and conditions (281) and (282) hold. Consider an arbitrary generalized resolvent Rs;λ(A) of A. Let Ru;ζ(Uz(A)) be generalized resolvent, corresponding to Rs;λ(A) by relation (40) from Theorem 2.6, and C(λ; 0) ∈ S(D; N0(Uz(A)), N∞(Uz(A))) corresponds to Ru;ζ(Uz(A)) by Inin's formula (25). The generalized resolvent Rs;λ(A) admits an analytic continuation on Re∪ ∆ if and only if the following conditions hold: tinuous in the uniform operator topology; 2) The continued function C(λ; 0) maps isometrically N0(Uz(A)) on the 1) C(λ; 0) admits a continuation on D ∪e∆ and this continuation is con- whole N∞(Uz(A)), for all λ ∈ e∆; 3) The operator C(λ; 0) − Wλ is invertible for all λ ∈ e∆, and ∀λ ∈ e∆, where the operator-valued function Wλ is defined for Uz(A) by (260). (C(λ; 0) − Wλ)N0(Uz(A)) = N∞(Uz(A)), (287) Proof. By Proposition 4.5 a possibility of an analytic continuation of Rs;λ(A) on Re∪∆ is equivalent to a possibility of an analytic continuation of Ru;ζ(Uz(A)) Here e∆ is from (280). on Te ∪ e∆. Using the equivalence of conditions (281),(282) and condi- tions (283),(284) we can apply Theorem 4.6 for Uz(A) and e∆, and the re- Consider an arbitrary symmetric operator A in a Hilbert space H. Fix an arbitrary number z ∈ Re. Consider an arbitrary generalized resolvent Rs;λ(A) of A and the corresponding to it by (286) the generalized resolvent Ru;ζ(Uz) of Uz(A). The linear fractional transformation ζ = λ−z λ−z maps Πz onD. Restrict relation (286) on Πz: quired result follows. ✷ Rs;λ(A) = z − z (λ − z)(λ − z)(cid:18)Ru; λ−z λ−z (Uz) − λ − z z − z EH(cid:19) , λ ∈ Πz\{z}. For the generalized resolvent Ru; λ−z (Uz) we can use the Chumakin formula, and for the generalized resolvent Rs;λ(A) we can apply the Shtraus formula: λ−z 131 z − z (λ − z)(λ − z) (cid:20)EH − = (cid:0)AF (λ),z − λEH(cid:1)−1 λ − z λ − z (Uz ⊕ Φ λ−z λ−z )(cid:21)−1 − λ − z z − z EH! , λ ∈ Πz\{z}. where F (λ) is a function from Sa;z(Πz; Nz(A), Nz(A)), Φζ is a function from S(D; N0(Uz(A)), N∞(Uz(A))) = S(D; Nz(A), Nz(A)). Observe that Φζ = C(ζ; 0), where C(ζ; 0) is a parameter from the Inin formula, corresponding to Ru; λ−z (Uz). After some transformations of the right-hand side we get λ−z (cid:0)AF (λ),z − λEH(cid:1)−1 1 λ−z , λ−z λ−z = (288) (Uz ⊕ Φ λ−z λ − z λ − z λ ∈ Πz\{z}. )(cid:21)−1 λ−z (Uz ⊕ Φ λ−z ) − EH(cid:17)(cid:20)EH − λ − z(cid:16)(Uz ⊕ Φ λ−z If for some h ∈ H, h 6= 0, we had (cid:16)(Uz ⊕ Φ λ−z would set g =hEH − λ−z (cid:0)AF (λ),z − λEH(cid:1)−1 g = 0, what is impossible. Therefore there exists the inverse(cid:16)(Uz ⊕ Φ λ−z = (λ − z)(cid:20)EH − ) − EH(cid:17) h = 0, then we we )i h 6= 0, and from (288) we would get ) − EH(cid:17)−1 ) − EH(cid:17)−1 AF (λ),z − λEH λ ∈ Πz\{z}. Then λ−z λ−z . Uz ⊕ F (λ) = Uz ⊕ Φ λ−z , λ ∈ Πz\{z}. λ−z After elementary transformations we get λ−z λ−z (Uz ⊕ Φ λ−z λ − z λ − z AF (λ),z = zEH + (z − z)(cid:16)(Uz ⊕ Φ λ−z )(cid:21)(cid:16)(Uz ⊕ Φ λ−z ) − EH(cid:17)−1 ((Uz ⊕ F (λ)) − EH)−1 =(cid:16)(Uz ⊕ Φ λ−z λ−z λ−z ) − EH(cid:17)−1 , , , λ ∈ Πz\{z}. Recalling the definition of the quasi-self-adjoint extension AF (λ),z we write: 132 Therefore F (λ) = Φ λ−z λ−z = C(cid:18) λ − z λ − z ; 0(cid:19) , λ ∈ Πz, (289) where the equality for λ = z follows from the continuity of F (λ) and Φζ. Theorem 4.8 Let A be a closed symmetric operator in a Hilbert space H, and z ∈ Re be an arbitrary fixed point. Let ∆ ⊆ R be a (finite or infinite) open interval, and conditions (281) and (282) hold. Consider an arbitrary generalized resolvent Rs;λ(A) of A. Let F (λ) ∈ Sa;z(Πz; Nz(A), Nz(A)) cor- responds to Rs;λ(A) by the Shtraus formula (166). The generalized resolvent Rs;λ(A) admits an analytic continuation on Re∪∆ if and only if the following conditions hold: 1) F (λ) admits a continuation onRe ∪ ∆ and this continuation is contin- uous in the uniform operator topology; 2) The continued function F (λ) maps isometrically Nz(A) on the whole Nz(A), for all λ ∈ ∆; 3) The operator F (λ) − Wλ is invertible for all λ ∈ ∆, and (F (λ) − Wλ)Nz(A) = Nz(A), ∀λ ∈ ∆, (290) where Wλ = W λ−z λ−z , λ ∈ ∆, and Wζ is defined for Uz(A) by (260). Proof. For the generalized resolvent Rs;λ(A) by (286) it corresponds a gen- eralized resolvent of the Cayley transformation Ru;ζ(Uz(A)). Let C(ζ; 0) ∈ S(D; N0(Uz(A)), N∞(Uz(A))) be a parameter from Inin's formula, correspond- Necessity. Let the generalized resolvent Rs;λ(A) admits an analytic contin- uation on Re ∪ ∆. By 4.6 for the parameter C(ζ; 0) it holds: ing to Ru;ζ(Uz(A)). Consider an arc e∆ from (280). (a) C(ζ; 0) admits a continuation on D ∪ e∆ and this continuation is con- on the whole N∞(Uz(A)) = Nz(A), for all ζ ∈ e∆; (c) The operator C(ζ; 0) − Wζ is invertible for all ζ ∈ e∆, and ∀ζ ∈ e∆, (b) The continued function C(ζ; 0) maps isometrically N0(Uz(A)) = Nz(A) (C(ζ; 0) − Wζ)N0(Uz(A)) = N∞(Uz(A)), (291) tinuous in the uniform operator topology; where Wζ is defined for Uz(A) by (260). 133 Define a function F+(λ) in the following way: F+(λ) = C(cid:18) λ − z λ − z ; 0(cid:19) , λ ∈ Πz ∪ ∆. (292) By (289) this function is an extension of the parameter F (λ), corresponding to the generalized resolvent by the Shtraus formula. From condition (a) it follows that this continuation is continuous in the uniform operator topology. From condition (b) it follows that F+(λ) maps isometrically Nz(A) on the whole Nz(A), for all λ ∈ ∆. Consider a function Wλ = W λ−z , λ ∈ ∆. From condition (c) it follows that F+(λ; 0) − Wλ is invertible for all λ ∈ ∆, and λ−z (F+(λ) − Wλ)Nz(A) = Nz(A), ∀λ ∈ ∆. Sufficiency. Suppose that for the parameter F (λ), corresponding to the generalized resolvent Rs;λ(A) by the Shtraus formula, conditions 1)-3) hold. Define a function C+(ζ; 0) in the following way: 1 − ζ (cid:19) , C+(ζ; 0) = F(cid:18)z − zζ ζ ∈ D ∪e∆. (293) By (289) this function is a continuation of the parameter C(ζ; 0). From condition 1) it follows that this continuation is continuous in the uniform operator topology. From condition 2) it follows that C+(ζ; 0) maps isometri- cally N0(Uz(A)) = Nz(A) on the whole N∞(Uz(A)) = Nz(A), for all ζ ∈ e∆. Finally, condition 3) means that C+(ζ; 0) − Wζ is invertible for all ζ ∈ e∆, (C+(ζ; 0) − Wζ)N0(Uz(A)) = N∞(Uz(A)), and By Proposition 4.6 we conclude that the generalized resolvent Rs;λ(A) admits an analytic continuation on Re ∪ ∆. ✷ Notice that by (260) the function Wλ fron the formulation of the last theorem has the following form: WλP H Nz(A)f = λ − z λ − z P H Nz(A)f, f ∈ Nλ(A), λ ∈ ∆. Moreover, we emphasize that conditions (281) and (282) are necessary for the existence at least one generalized resolvent of A, which admits an analytic continuation on Re ∪ ∆. Consequently, these conditions does not imply on the generality of a description such generalized resolvents in Theorem 4.8. ∀ζ ∈ e∆. 134 5 Formal credits. Section 2. §2.1. Results of this subsection, except for Propositions, belong to Chumakin, see [4], [6]. Propositions 2.1-2.3 belong to Shtraus, see [37, footnote on page 83], [40, Lemma 1.1, Lemma 1.2]. §2.2. Chumakin's formula appeared in papers of Chumakin [4] and [6]. §2.3. Inin's formula was obtained in the paper [15, Theorem]. Our proof dif- fers from the original proof and it is based directly on the use of Chumakin's formula for the generalized resolvents and a method which is similar to the metod of Chumakin in the paper [5]. §2.4. The definition of the generalized resolvent of a symmetric, not nec- essarily densely defined operator in a Hilbert space belongs to Shtraus and he established its basic properties, see [37], [41]. A connection between the generalized resolvents of a symmetric operator and the generalized resolvent its Cayley transformation (Theorem 2.6) ina somewhat less general form was established by Chumakin in [5]. §2.5. The Shtraus formula was obtained in [37]. Our proof differs form the original one and we follow the idea of Chumakin, proposed in [5]. Section 3. This section is based on results of Shtraus from papers [37], [38], [39], [40], [41], see also references in these papers. §4.1. Proposition 4.1 is an analog of Theorem 3.1 (A),(B) in a Section 4. paper of McKelvey [29], see also [43, Lemma 1.1]. Theorem 4.1 is an analog of Theorem 2.1 (A) in a paper of McKelvey [29]. Theorem 4.2 is an analog of Theorem 2.1 (B) in a paper of McKelvey [29]. §4.2. Theorem 4.3 appeared in [35, Lemma 1]. However her proof was based on a lemma of Shmulyan [36, Lemma 3]. As far as we know, no correct proof of this lemma was published. An attempt to proof the lemma of Shmulyan was performed by L.A. Shtraus in [42, Lemma]. Unfortunately, the proof was not complete. We used the idea of L.A. Shtraus for the proof a weaker result: Theorem 4.3. §4.3. Results of this subsection mostly belong to Ryabtseva, in some cases with our filling of gaps in proofs, corrections and a generalization. For Lem- mas 4.3, 4.4 see Lemma 2 and its corollary, and also Lemma 3 in [35]. Theo- rem 4.4 was obtained in [35, Theorem 1]. Theorem 4.5 is a corrected version of Theorem 2 in [35]. Theorem 4.6 is a little corrected version of Theorem 4 in [35]. For Proposition 4.2 see [15, p. 34]. §4.4. Proposition 4.3 is contained in Theorem 3.1 (A),(B) in a paper of McKelvey [29], see also [43, Lemma 1.1]. Theorem 4.8 is close to results of Varlamova-Luks, see [45, Theorem], [44, Theorem 2], [43, Theorem 3.1]. 135 References [1] Akhiezer N. I., Glazman I. M. The theory of linear operators in a Hilbert space. -- Moscow: "Nauka", 1966. (Russian). [2] Aleksandrov E. L., Ilmushkin G. M. Generalized resolvents of isometric and symmetric operators // Izvestiya vyssh. uchebn. zavedeniy. -- 1977. -- 1 (176). -- P. 14 -- 23. (Russian). [3] Behrndt J., Kreusler H.-Ch. Boundary Relations and Generalized Re- solvents of Symmetric Operators in Krein Spaces // Integral equations and operator theory. -- 2007. -- 59, no. 3. -- P. 309 -- 327. [4] Chumakin M. E. On the generalized resolvents of an isometric operator // DAN SSSR. -- 1964. -- 154, no. 4. -- P. 791 -- 794. (Russian). [5] Chumakin M. E. On a class of generalized resolvents of an isometric operator // Uchen. zapiski Ulyanovskogo pedinstituta. -- 1966. -- 20, issue 4. -- P. 361 -- 373. (Russian). [6] Chumakin M. E. Generalized resolvents of isometric operators // Sibirskiy matem. zhurnal. -- 1967. -- VIII, no. 4. -- P. 876 -- 892. (Rus- sian). [7] Derkach V. A. On Indefinite Moment Problem and Resolvent Matrices of Hermitian Operators in Krein Spaces // Math. Nachrichten. -- 1997. -- 184. -- P. 135-166. [8] Derkach V. A. On generalized resolvents of Hermitian relations in Krein spaces // Journal of Mathematical Sciences. -- 1999. -- 97, no. 5. -- P. 4420 -- 4460. [9] V. Derkach, S. Hassi, M. Malamud, H. de Snoo. Boundary relations and generalized resolvents of symmetric operators // Russian J. of Math. Physics. -- 2009. -- 16, No. 1. -- P. 17 -- 60. [10] Derkach V. A., Malamud M. M. Generalized resolvents and the bound- ary value problems for Hermitian operators with gaps // Journal of Functional Analysis. -- 1991. -- 95, no. 1. -- P. 1 -- 95. [11] Derkach V. A., Malamud M. M. The extension theory of Hermitian op- erators and the moment problem // Journal of Mathematical Sciences. -- 1995. -- 73, no. 2. -- P. 141 -- 242. Translated from Itogi Nauki i Tekhniki, Seriya Sovremennaya Matematika i Ee Prilozheniya. Tematicheskie Ob- zory Vol. 5, Analiz-3, 1993. 136 [12] Etkin A. B. Generalized resolvents of contractions in the Pontryagin space // Funktsionalniy analiz (Ulyanovsk). -- 1988. -- 28. -- P. 132 -- 141. (Russian). [13] Gluhov V. P., Nikonov A. A. On extensions and generalized resolvents of J-symmetric and J-dissipative relations // Funktsionalniy analiz (Ulyanovsk). -- 1976. -- 7. -- P. 53 -- 59. (Russian). [14] Gorbachuk M. L., Gorbachuk V. I. M.G. Krein's lectures on entire op- erators. -- Basel; Boston: Birkha"user Verlag, 1997. [15] Inin O. T. Resolvents of isometric operators with arbitrary defect num- bers // Uchen. zapiski Ulyanovskogo pedinstituta. -- 1970. -- 24, issue 3. -- part 1. -- P. 33 -- 45. (Russian). [16] Kalinina T.B. The generalized resolvents of an operator which is skew- symmetric with respect to antilinear transformation of a Hilbert space // Funktsionalniy analiz (Ulyanovsk). -- 1983. -- 20. -- P. 60 -- 73. (Russian). [17] Krein M. On Hermitian operators with defect indices equal to a unit // DAN SSSR. -- 1944. -- XLIII, no. 8. -- P. 339 -- 342. (Russian). [18] Krein M. G., Saakyan Sh. N. On some new results in the theory of resolvents of Hermitian operators // DAN SSSR. -- 1966. -- 169, no. 6. -- P. 1269 -- 1272. (Russian). [19] Krein M. G., Langer H. K. On the defect subspaces and generalized resolvents of a Hermitian operator in the space Πχ // Funkts. analiz i ego prilozheniya. -- 1971. -- 5, issue. 2. -- P. 59 -- 71. (Russian). [20] Krein M. G., Langer H. K. On the defect subspaces and generalized resolvents of a Hermitian operator in the space Πχ // Funkts. analiz i ego prilozheniya. -- 1971. -- 5, issue. 3. -- P. 54 -- 69. (Russian). [21] Krein M. G., Ovcharenko I. E. On Q-functions and SC-resolvents of non- densely defined Hermitian contractions // Sibirskiy matem. zhurnal. -- 1977. -- XVIII, no. 5. -- P. 1032 -- 1056. (Russian). [22] Krein M. G., Ovcharenko I. E. Inverse problems for Q-functions and resolvent matrices of positive Hermitian operators // DAN SSSR. -- 1978. -- 242, no. 3. -- P. 521 -- 524. (Russian). [23] Langer H, Sorjonen P. Verallgemeinerte resolventen Hermitescher und isometrischer operatoren im Pontrjaginraum // Annales Academiae Sci- entiarum Fennicae, Series A. -- 1974. -- 561. -- P. 1 -- 45. 137 [24] Langer H., Textorius B. L -resolvent matrices of symmetric linear rela- tions with equal defect numbers; applications to canonical differential relations // Integral equations and operator theory. -- 1982. -- 5, no. 1. -- P. 208 -- 243. [25] Makarova A. D. On a generalized resolvent of a J-symmetric operator // Funktsionalniy analiz (Ulyanovsk). -- 1980. -- 14. -- P. 99 -- 104. (Russian). [26] Makarova A. D. Extensions and generalized resolvents of an opera- tor which is commuting with an involution // Funktsionalniy analiz (Ulyanovsk). -- 1983. -- 20. -- P. 83 -- 92. (Russian). [27] Malamud M. M. On a formula of the generalized resolvents of a non- densely defined Hermitian operator // Ukrainian mathematical journal. -- 1992. -- 44, no. 12. -- P. 1658 -- 1688. (Russian). [28] Malamud M. M., Mogilevskii V. I. The generalized resolvents of an iso- metric operator // Matem. zametki. -- 2003, 73, issue. 3. -- P. 460-465. (Russian). [29] McKelvey R. The spectra of minimal self-adjoint extensions of a sym- metric operator // Pacific J. Math. -- 1962. -- 12, no. 3. -- P. 1003 -- 1022. [30] Mogilevskii V. Boundary triplets and Krein type resolvent formula for symmetric operators with unequal defect numbers // Methods of Func- tional Analysis and Topology. -- 2006. -- 12, no. 3. -- P. 258 -- 280. [31] Neumark M. A. Spectral functions of a symmetric operator // Izvestiya AN SSSR. -- 1940. -- 4. -- P. 277 -- 318. (Russian). [32] Neumark M. A. On spectral functions of a symmetric operator // Izvestiya AN SSSR. -- 1943. -- 7. -- P. 285 -- 296. (Russian). [33] Nikonov A. A. Generalized resolvents of a non-densely defined closed J- symmetric operator in a J-space // Funktsionalniy analiz (Ulyanovsk). -- 1975. -- 5. -- P. 116 -- 124. (Russian). [34] Nikonov A. A. Generalized resolvents of closed J-dissipative operators in a J-space // Funktsionalniy analiz (Ulyanovsk). -- 1975. -- 5. -- P. 125 -- 133. (Russian). [35] Ryabtseva V. M. On the resolvent set of a spectral function of an iso- metric operator with infinite defect indices // Izvestiya vyssh. uchebn. zavedeniy. -- 1968. -- 70, 3. -- P. 75 -- 79. (Russian). 138 [36] Shmulyan Yu. L. Isometric operators with infinite defect indices and their orthogonal extensions // DAN SSSR. -- 1952. -- LXXXVII, no. 1. -- P. 11 -- 14. (Russian). [37] Shtraus A. V. Generalized resolvents of symmetric operators // Izvestiya AN SSSR. -- 1954. -- 18. -- P. 51 -- 86. (Russian). [38] Shtraus A. V. On extensions of a symmetric operator depending on a parameter // Izvestiya AN SSSR. -- 1965. -- 29. -- P. 1389 -- 1416. (Rus- sian). [39] Shtraus A. V. On one-parameter families of extensions of a symmetric operator // Izvestiya AN SSSR. -- 1966. -- 30. -- P. 1325 -- 1352. (Russian). [40] Shtraus A. V. On extensions and the characteristic function of a sym- metric operator // Izvestiya AN SSSR. -- 1968. -- 32. -- P. 186 -- 207. (Russian). [41] Shtraus A. V. Extensions and generalized resolvents of a non-densely defined symmetric operator // Izvestiya AN SSSR. -- 1970. -- 34. -- P. 175 -- 202. (Russian). [42] Shtraus L. A. On orthogonal extensions and the characteristic function of an isometric operator // Funktsionalniy analiz (Ulyanovsk). -- 1980. -- issue 14. -- P. 190 -- 197. (Russian). [43] Varlamova-Luks L. N. On positive spectral functions of symmetric oper- ators with finite defect numbers // Uchen. zapiski Ulyanovskogo pedin- stituta. -- 1966. -- 20, issue. 4. -- P. 335 -- 350. (Russian). [44] Varlamova-Luks L. N. On positive spectral functions of a Hermitian operator with finite defect numbers // Volzhskiy matem. sbornik. -- 1966. -- issue. 5. -- P. 86 -- 91. (Russian). [45] Varlamova-Luks L. N. Positive spectral functions of Hermitian operators with infinite defect numbers // Uchen. zapiski Ulyanovskogo pedinsti- tuta. -- 1967. -- 21, issue 8. -- P. 227 -- 230. (Russian).
1505.06727
3
1505
2015-10-31T22:17:31
Nontrivial twisted sums of $c_0$ and $C(K)$
[ "math.FA" ]
We obtain a new class of compact Hausdorff spaces $K$ for which $c_0$ can be nontrivially twisted with $C(K)$.
math.FA
math
NONTRIVIAL TWISTED SUMS OF c0 AND C(K) CLAUDIA CORREA AND DANIEL V. TAUSK Abstract. We obtain a new class of compact Hausdorff spaces K for which c0 can be nontrivially twisted with C(K). 1. Introduction In this article, we present a broad new class of compact Hausdorff spaces K such that there exists a nontrivial twisted sum of c0 and C(K), where C(K) denotes the Banach space of continuous real-valued functions on K endowed with the supremum norm. By a twisted sum of the Banach spaces Y and X we mean a short exact sequence 0 → Y → Z → X → 0, where Z is a Banach space and the maps are bounded linear operators. This twisted sum is called trivial if the exact sequence splits, i.e., if the map Y → Z admits a bounded linear left inverse (equivalently, if the map Z → X admits a bounded linear right inverse). In other words, the twisted sum is trivial if the range of the map Y → Z is complemented in Z; in this case, Z ∼= X ⊕ Y . As in [7], we denote by Ext(X, Y ) the set of equivalence classes of twisted sums of Y and X and we write Ext(X, Y ) = 0 if every such twisted sum is trivial. Many problems in Banach space theory are related to the quest for condi- tions under which Ext(X, Y ) = 0. For instance, an equivalent statement for the classical Theorem of Sobczyk ([5, 13]) is that if X is a separable Banach space, then Ext(X, c0) = 0 ([3, Proposition 3.2]). The converse of the latter since ℓ1(I) is a projective Banach space. However, the following question space? This problem was stated in [4, 5] and further studied in the recent article [6], in which the author proves that, under the continuum hypothe- statement clearly does not hold in general: for example, Ext(cid:0)ℓ1(I), c0(cid:1) = 0, remains open: is it true that Ext(cid:0)C(K), c0(cid:1) 6= 0 for any nonseparable C(K) sis (CH), the space Ext(cid:0)C(K), c0(cid:1) is nonzero for a nonmetrizable compact namely that Ext(cid:0)C(K), c0(cid:1) is nonzero for a C(K) space under any one of Hausdorff space K of finite height. In addition to this result, everything else that is known about the problem is summarized in [6, Proposition 2], the following assumptions: Date: October 22nd, 2015. 2010 Mathematics Subject Classification. 46B20,46E15. Key words and phrases. Banach spaces of continuous functions; twisted sums of Banach spaces; Valdivia compacta. The first author is sponsored by FAPESP (Process no. 2014/00848-2). 1 NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 2 • K is a nonmetrizable Eberlein compact space; • K is a Valdivia compact space which does not satisfy the countable chain condition (ccc); • the weight of K is equal to ω1 and the dual space of C(K) is not weak*-separable; • K has the extension property ([8]) and it does not have ccc; • C(K) contains an isomorphic copy of ℓ∞. Note also that if Ext(Y, c0) 6= 0 and X contains a complemented isomorphic copy of Y , then Ext(X, c0) 6= 0. Here is an overview of the main results of this article. Theorem 2.3 gives a condition involving biorthogonal systems in a Banach space X which implies that Ext(X, c0) 6= 0. In the rest of Section 2, we discuss some of its implications when X is of the form C(K). It is proven that if K contains a homeomorphic copy of [0, ω] × [0, c] or of 2c, then Ext(cid:0)C(K), c0(cid:1) is nonzero, where c denotes the cardinality of the continuum. In Sections 3 and 4, we investigate the consequences of the results of Section 2 for Valdivia and Corson compacta. Recall that Valdivia compact spaces constitute a large superclass of Corson compact spaces closed under arbitrary products; moreover, every Eberlein compact is a Corson compact (see [11] for a survey on Valdivia compacta). Section 3 is devoted to the proof that, under CH, it holds that Ext(cid:0)C(K), c0(cid:1) 6= 0 for every nonmetrizable Corson compact space K. The question of whether Ext(cid:0)C(K), c0(cid:1) 6= 0 for an arbitrary nonmetrizable Valdivia compact space K remains open (even under CH), but in Section 4 we solve some particular cases of this problem. 2. General results Throughout the paper, the weight and the density character of a topolog- ical space X are denoted, respectively, by w(X ) and dens(X ). Moreover, we always denote by χA the characteristic function of a set A and by A the cardinality of A. We start with a technical lemma which is the heart of the proof of Theorem 2.3. A family of sets (Ai)i∈I is said to be almost disjoint if each Ai is infinite and Ai ∩ Aj is finite, for all i, j ∈ I with i 6= j. Lemma 2.1. There exists an almost disjoint family (An,α)n∈ω,α∈c of subsets n,α)n∈ω,α∈c with of ω satisfying the following property: each A′ n,α ⊂ An,α cofinite in An,α, it holds that supp∈ω Mp = +∞, where: for every family (A′ Mp = (cid:8)n ∈ ω : p ∈ Sα∈c A′ n,α(cid:9). Proof. We will obtain an almost disjoint family (An,α)n∈ω,α∈c of subsets of 2<ω with the desired property, where 2<ω = Sk∈ω 2k denotes the set of finite sequences in 2 = {0, 1}. For each ǫ ∈ 2ω, we set: Aǫ = (cid:8)ǫk : k ∈ ω(cid:9), so that (Aǫ)ǫ∈2ω is an almost disjoint family of subsets of 2<ω. Let (Bα)α∈c be an enumeration of the uncountable Borel subsets of 2ω. Recalling that NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 3 Bα = c for all α ∈ c ([12, Theorem 13.6]), one easily obtains by transfinite recursion a family (ǫn,α)n∈ω,α∈c of pairwise distinct elements of 2ω such that ǫn,α ∈ Bα, for all n ∈ ω, α ∈ c. Set An,α = Aǫn,α and let (A′ n,α)n∈ω,α∈c be as in the statement of the lemma. For n ∈ ω, denote by Dn the set of those ǫ ∈ 2ω such that n ∈ Mp for all but finitely many p ∈ Aǫ. Note that: Dn = [ \ [(cid:8)Cp : p ∈ 2k with n ∈ Mp(cid:9), k0∈ω k≥k0 where Cp denotes the clopen subset of 2ω consisting of the extensions of p. The above equality implies that Dn is an Fσ (and, in particular, a Borel) subset of 2ω. We claim that the complement of Dn in 2ω is countable. Namely, if it were uncountable, there would exist α ∈ c with Bα = 2ω \ Dn. But, since n ∈ Mp for all p ∈ A′ n,α, we have that ǫn,α ∈ Dn, contradicting the fact that ǫn,α ∈ Bα and proving the claim. To conclude the proof of the lemma, note that for each n ≥ 1 the intersection Ti<n Di is nonempty; for ǫ ∈ Ti<n Di, we have that {i : i < n} ⊂ Mp, for all but finitely many (cid:3) p ∈ Aǫ. Let X be a Banach space. Recall that a biorthogonal system in X is a family (xi, γi)i∈I with xi ∈ X, γi ∈ X ∗, γi(xi) = 1 and γi(xj) = 0 for i 6= j. The cardinality of the biorthogonal system (xi, γi)i∈I is defined as the cardinality of I. Definition 2.2. Let (xi, γi)i∈I be a biorthogonal system in a Banach space X. We call (xi, γi)i∈I bounded if supi∈I kxik < +∞ and supi∈I kγik < +∞; weak*-null if (γi)i∈I is a weak*-null family, i.e., if (cid:0)γi(x)(cid:1)i∈I is in c0(I), for all x ∈ X. Theorem 2.3. Let X be a Banach space. Assume that there exist a weak*- null biorthogonal system (xn,α, γn,α)n∈ω,α∈c in X and a constant C ≥ 0 such that: (cid:13)(cid:13)(cid:13) kX i=1 xni,αi(cid:13)(cid:13)(cid:13) ≤ C, for all n1, . . . , nk ∈ ω pairwise distinct, all α1, . . . , αk ∈ c, and all k ≥ 1. Then Ext(X, c0) 6= 0. Proof. By [7, Proposition 1.4.f], we have that Ext(X, c0) = 0 if and only if every bounded operator T : X → ℓ∞/c0 admits a lifting 1, i.e., a bounded operator bT : X → ℓ∞ with T (x) = bT (x)+c0, for all x ∈ X. Let us then show that there exists an operator T : X → ℓ∞/c0 that does not admit a lifting. To this aim, let (An,α)n∈ω,α∈c be an almost disjoint family as in Lemma 2.1 and consider the unique isometric embedding S : c0(ω × c) → ℓ∞/c0 such + c0, where (en,α)n∈ω,α∈c denotes the canonical basis of that S(en,α) = χAn,α 1More concretely, a nontrivial twisted sum of c0 and X is obtained by considering the pull-back of the short exact sequence 0 → c0 → ℓ∞ → ℓ∞/c0 → 0 by an operator T : X → ℓ∞/c0 that does not admit a lifting. NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 4 c0(ω×c). Denote by Γ : X → c0(ω×c) the bounded operator with coordinate functionals (γn,α)n∈ω,α∈c and set T = S ◦ Γ : X → ℓ∞/c0. Assuming by contradiction that there exists a lifting bT of T and denoting by (µp)p∈ω the sequence of coordinate functionals of bT , we have that the set: 2(cid:9) n,α = (cid:8)p ∈ An,α : µp(xn,α) ≥ 1 is cofinite in An,α. n1, . . . , nk ∈ ω pairwise distinct, and α1, . . . , αk ∈ c such that p ∈ A′ i = 1, . . . , k. Hence: A′ It follows that for each k ≥ 1, there exist p ∈ ω, ni,αi, for k 2 ≤ µp(cid:16) kX i=1 xni,αi(cid:17) ≤ CkbT k, which yields a contradiction. (cid:3) Corollary 2.4. Let K be a compact Hausdorff space. Assume that there exists a bounded weak*-null biorthogonal system (fn,α, γn,α)n∈ω,α∈c in C(K) such that fn,αfm,β = 0, for all n, m ∈ ω with n 6= m and all α, β ∈ c. Then (cid:3) Ext(cid:0)C(K), c0(cid:1) 6= 0. Definition 2.5. We say that a compact Hausdorff space K satisfies property (∗) if there exist a sequence (Fn)n∈ω of closed subsets of K and a bounded weak*-null biorthogonal system (fn,α, γn,α)n∈ω,α∈c in C(K) such that: (1) Fn ∩ [ m6=n Fm = ∅ and supp fn,α ⊂ Fn, for all n ∈ ω and all α ∈ c, where supp fn,α denotes the support of fn,α. In what follows, we denote by M (K) the space of finite countably-additive signed regular Borel measures on K, endowed with the total variation norm. We identify as usual the dual space of C(K) with M (K). Lemma 2.6. Let K be a compact Hausdorff space and L be a closed subspace of K. If L satisfies property (∗), then so does K. Proof. Consider, as in Definition 2.5, a sequence (Fn)n∈ω of closed subsets of L and a bounded weak*-null biorthogonal system (fn,α, γn,α)n∈ω,α∈c in C(L). By recursion on n, one easily obtains a sequence (Un)n∈ω of pairwise disjoint open subsets of K with each Un containing Fn. Let Vn be an open subset of K with Fn ⊂ Vn ⊂ Vn ⊂ Un. Using Tietze's Extension Theorem and Urysohn's Lemma, we get a continuous extension fn,α of fn,α to K with support contained in Vn and having the same norm as fn,α. To conclude the proof, let γn,α ∈ M (K) be the extension of γn,α ∈ M (L) that vanishes identically outside of L and observe that ( fn,α, γn,α)n∈ω,α∈c is a bounded weak*-null biorthogonal system in C(K). (cid:3) As an immediate consequence of Lemma 2.6 and Corollary 2.4, we obtain the following result. NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 5 Theorem 2.7. If a compact Hausdorff space L satisfies property (∗), then every compact Hausdorff space K containing a homeomorphic copy of L (cid:3) satisfies Ext(cid:0)C(K), c0(cid:1) 6= 0. We now establish a few results which give sufficient conditions for a space K to satisfy property (∗). Recall that, given a closed subset F of a compact Hausdorff space K, an extension operator for F in K is a bounded operator E : C(F ) → C(K) which is a right inverse for the restriction operator C(K) ∋ f 7→ f F ∈ C(F ). Note that F admits an extension operator in K if and only if the kernel C(KF ) = (cid:8)f ∈ C(K) : f F = 0(cid:9) of the restriction operator is complemented in C(K). A point x of a topo- logical space X is called a cluster point of a sequence (Sn)n∈ω of subsets of X if every neighborhood of x intersects Sn for infinitely many n ∈ ω. Lemma 2.8. Let K be a compact Hausdorff space. Assume that there exist a sequence (Fn)n∈ω of pairwise disjoint closed subsets of K and a closed subset F of K satisfying the following conditions: (a) F admits an extension operator in K; (b) every cluster point of (Fn)n∈ω is in F and Fn ∩ F = ∅, for all n ∈ ω; (c) there exists a family (fn,α, γn,α)n∈ω,α∈c, where (fn,α, γn,α)α∈c is a weak*-null biorthogonal system in C(Fn) for each n ∈ ω and sup n∈ω,α∈c kfn,αk < +∞, sup n∈ω,α∈c kγn,αk < +∞. Then K satisfies property (∗). Proof. From (b) and the fact that the Fn are pairwise disjoint it follows that (1) holds. Let (Un)n∈ω, (Vn)n∈ω, and ( fn,α)n∈ω,α∈c be as in the proof of Lemma 2.6; we assume also that V n ∩F = ∅, for all n ∈ ω. Let γn,α ∈ M (K) be the extension of γn,α ∈ M (Fn) that vanishes identically outside of Fn. We have that ( fn,α, γn,α)n∈ω,α∈c is a bounded biorthogonal system in C(K) and that (γn,α)α∈c is weak*-null for each n, though it is not true in general that the entire family (γn,α)n∈ω,α∈c is weak*-null. In order to take care of this problem, let P : C(K) → C(KF ) be a bounded projection and set γn,α = γn,α◦P . Since all fn,α are in C(KF ), we have that ( fn,α, γn,α)n∈ω,α∈c is biorthogonal. To prove that (γn,α)n∈ω,α∈c is weak*-null, note that (b) implies that limn→+∞ kf Fnk = 0, for all f ∈ C(KF ). (cid:3) Corollary 2.9. Let K be a compact Hausdorff space. If C(K) admits a bounded weak*-null biorthogonal system of cardinality c, then the space [0, ω] × K satisfies property (∗). In particular, L × K satisfies property (∗) for every compact Hausdorff space L containing a nontrivial convergent sequence. Proof. Take Fn = {n} × K, F = {ω} × K, and use the fact that F is a retract of [0, ω] × K and thus admits an extension operator in [0, ω] × K. (cid:3) NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 6 Corollary 2.10. The spaces [0, ω] × [0, c] and 2c satisfy property (∗). In particular, a product of at least c compact Hausdorff spaces with more than one point satisfies property (∗). Proof. The family (χ[0,α], δα−δα+1)α∈c is a bounded weak*-null biorthogonal system in C(cid:0)[0, c](cid:1), where δα ∈ M(cid:0)[0, c](cid:1) denotes the probability measure with support {α}. It follows from Corollary 2.9 that [0, ω] × [0, c] satisfies property (∗). To see that 2c also does, note that the map [0, c] ∋ α 7→ χα ∈ 2c embeds [0, c] into 2c, so that 2c ∼= 2ω × 2c contains a homeomorphic copy of [0, ω] × [0, c]. (cid:3) Recall that a projectional resolution of the identity (PRI) of a Banach space X is a family (Pα)ω≤α≤dens(X) of projection operators Pα : X → X satisfying the following conditions: • kPαk = 1, for ω ≤ α ≤ dens(X); • Pdens(X) is the identity of X; • Pα[X] ⊂ Pβ[X] and Ker(Pβ) ⊂ Ker(Pα), for ω ≤ α ≤ β ≤ dens(X); • Pα(x) = limβ<α Pβ(x), for all x ∈ X, if ω < α ≤ dens(X) is a limit ordinal; • dens(cid:0)Pα[X](cid:1) ≤ α, for ω ≤ α ≤ dens(X). We call the PRI strictly increasing if Pα[X] is a proper subspace of Pβ[X], for ω ≤ α < β ≤ dens(X). Corollary 2.11. Let K and L be compact Hausdorff spaces such that L contains a nontrivial convergent sequence and w(K) ≥ c. If C(K) admits a strictly increasing PRI, then the space L × K satisfies property (∗). Proof. This follows from Corollary 2.9 by observing that if a Banach space X admits a strictly increasing PRI, then X admits a weak*-null biorthogonal system (xα, γα)ω≤α<dens(X) with kxαk = 1 and kγαk ≤ 2, for all α. Namely, pick a unit vector xα in Pα+1[X] ∩ Ker(Pα) and set γα = φα ◦ (Pα+1 − Pα), where φα ∈ X ∗ is a norm-one functional satisfying φα(xα) = 1. (cid:3) 3. Nontrivial twisted sums for Corson compacta Let us recall some standard definitions. Given an index set I, we write Σ(I) = (cid:8)x ∈ RI : supp x is countable(cid:9), where the support supp x of x is defined by supp x = (cid:8)i ∈ I : xi 6= 0(cid:9). Given a compact Hausdorff space K, we call A a Σ-subset of K if there exist an index set I and a continuous injection ϕ : K → RI such that A = ϕ−1[Σ(I)]. The space K is called a Valdivia compactum if it admits a dense Σ-subset and it is called a Corson compactum if K is a Σ-subset of itself. This section will be dedicated to the proof of the following result. Theorem 3.1. If K is a Corson compact space with w(K) ≥ c, then Ext(cid:0)C(K), c0(cid:1) 6= 0. In particular, under CH, we have Ext(cid:0)C(K), c0(cid:1) 6= 0 for any nonmetrizable Corson compact space K. NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 7 The fact that Ext(cid:0)C(K), c0(cid:1) 6= 0 for a Valdivia compact space K which does not have ccc is already known ([6, Proposition 2]). Our strategy for the proof of Theorem 3.1 is to use Lemma 2.8 to show that if K is a Corson compact space with w(K) ≥ c having ccc, then K satisfies property (∗). We start with a lemma that will be used as a tool for verifying the assumptions of Lemma 2.8. Recall that a closed subset of a topological space is called regular if it is the closure of an open set (equivalently, if it is the closure of its own interior). Obviously, a closed subset of a Corson compact space is again Corson and a regular closed subset of a Valdivia compact space is again Valdivia. Lemma 3.2. Let K be a compact Hausdorff space and F be a closed non- open Gδ subset of K. Then there exists a sequence (Fn)n∈ω of nonempty pairwise disjoint regular closed subsets of K such that condition (b) in the statement of Lemma 2.8 holds. Proof. We can write F = Tn∈ω Vn, with each Vn open in K and Vn+1 prop- erly contained in Vn. Set Un = Vn\Vn+1, so that all cluster points of (Un)n∈ω are in F . To conclude the proof, let Fn be a nonempty regular closed set contained in Un. (cid:3) Once we get the closed sets (Fn)n∈ω from Lemma 3.2, we still have to verify the rest of the conditions in the statement of Lemma 2.8. First, we need an assumption ensuring that w(Fn) ≥ c, for all n. To this aim, given a point x of a topological space X , we define the weight of x in X by: w(x, X ) = min(cid:8)w(V ) : V neighborhood of x in X(cid:9). Recall that if K is a Valdivia compact space, then C(K) admits a PRI ([15, Theorem 2]). Moreover, a trivial adaptation of the proof in [15] shows in fact that C(K) admits a strictly increasing PRI. Thus, by the argument in the proof of Corollary 2.11, C(K) admits a weak*-null biorthogonal system (fα, γα)ω≤α<w(K) such that kfαk ≤ 1 and kγαk ≤ 2, for all α. The following result is now immediately obtained. Corollary 3.3. Let K be a Valdivia compact space such that w(x, K) ≥ c, for all x ∈ K. Assume that there exists a closed non-open Gδ subset F admitting an extension operator in K. Then K satisfies property (∗). (cid:3) Assuming that K has ccc, the next lemma allows us to reduce the proof of Theorem 3.1 to the case when w(x, K) ≥ c, for all x ∈ K. Lemma 3.4. Let K be a ccc Valdivia compact space and set: H = (cid:8)x ∈ K : w(x, K) ≥ c(cid:9). Then: (a) H 6= ∅, if w(K) ≥ c; (b) w(cid:0)K \ int(H)(cid:1) < c, where int(H) denotes the interior of H; (c) H is a regular closed subset of K; NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 8 (d) w(x, H) ≥ c, for all x ∈ H. Proof. If H = ∅, then K can be covered by a finite number of open sets with weight less than c, so that w(K) < c. This proves (a). To prove (b), let (Ui)i∈I be maximal among antichains of open subsets of K with weight less than c. Since I is countable and c has uncountable cofinality, we have that U = Si∈I Ui has weight less than c. From the maximality of (Ui)i∈I , it follows that K \ H ⊂ U ; then K \ int(H) = K \ H ⊂ U . To conclude the proof of (b), let us show that w(U ) < c. Let A be a dense Σ-subset of K and let D be a dense subset of A ∩ U with D ≤ w(U ). Then D is homeomorphic to a subspace of Rw(U ), so that w(U ) = w(D) ≤ w(U ) < c. To prove (c), note that H is clearly closed; moreover, by (b), the open set K \int(H) has weight less than c and hence it is contained in K \H. Finally, to prove (d), let V be a closed neighborhood in K of some x ∈ H. By (b), we have w(V \ H) < c. Recall from [9, p. 26] that if a compact Hausdorff space is the union of not more than κ subsets of weight not greater than κ, then the weight of the space is not greater than κ. Since w(V ) ≥ c, it follows from such result that w(V ∩ H) ≥ c. (cid:3) Proof of Theorem 3.1. By Lemma 3.4, it suffices to prove that if K is a nonempty Corson compact space such that w(x, K) ≥ c for all x ∈ K, then K satisfies property (∗). Since a nonempty Corson compact space K admits a Gδ point x ([11, Theorem 3.3]), this fact follows from Corollary 3.3 with F = {x}. (cid:3) Remark 3.5. It is known that under Martin's Axiom (MA) and the negation of CH, every ccc Corson compact is metrizable ([1]). Thus, Theorem 3.1 implies that Ext(cid:0)C(K), c0(cid:1) 6= 0 for every nonmetrizable Corson compact space K under MA. 4. Towards the general Valdivia case In this section we prove that Ext(cid:0)C(K), c0(cid:1) 6= 0 for certain classes of nonmetrizable Valdivia compact spaces K and we propose a strategy for dealing with the general problem. First, let us state some results which are immediate consequences of what we have done so far. Proposition 4.1. If K is a Valdivia compact space with w(K) ≥ c and L is a compact Hausdorff space containing a nontrivial convergent sequence, then L × K satisfies property (∗). Proof. As we have observed in Section 3, if K is a Valdivia compact space, then C(K) admits a strictly increasing PRI. The conclusion follows from Corollary 2.11. (cid:3) Proposition 4.2. Let K be a Valdivia compact space admitting a Gδ point x with w(x, K) ≥ c. Then Ext(cid:0)C(K), c0(cid:1) 6= 0 and, if K has ccc, then K satisfies property (∗). NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 9 Proof. As mentioned before, the non-ccc case is already known. Assuming that K has ccc, define H as in Lemma 3.4 and conclude that H satisfies property (∗) using Corollary 3.3 with F = {x}. (cid:3) Corollary 4.3. Let K be a Valdivia compact space with w(K) ≥ c ad- mitting a dense Σ-subset A such that K \ A is of first category. Then Ext(cid:0)C(K), c0(cid:1) 6= 0 and, if K has ccc, then K satisfies property (∗). Proof. By [11, Theorem 3.3], K has a dense subset of Gδ points. Assuming that K has ccc and defining H as in Lemma 3.4, we obtain that H con- tains a Gδ point of K, which implies that K satisfies the assumptions of Proposition 4.2. (cid:3) Now we investigate conditions under which a Valdivia compact space K contains a homeomorphic copy of [0, ω] × [0, c]. Given an index set I and a subset J of I, we denote by rJ : RI → RI the map defined by setting rJ (x)J = xJ and rJ (x)I\J ≡ 0, for all x ∈ RI . Following [2], given a subset K of RI , we say that J ⊂ I is K-good if rJ [K] ⊂ K. In [2, Lemma 1.2], it is proven that if K is a compact subset of RI and Σ(I) ∩ K is dense in K, then every infinite subset J of I is contained in a K-good set J ′ with J = J ′. Proposition 4.4. Let K be a Valdivia compact space admitting a dense Σ- subset A such that some point of K \A is the limit of a nontrivial sequence in K. Then K contains a homeomorphic copy of [0, ω] × [0, ω1]. In particular, assuming CH, we have that K satisfies property (∗). Proof. We can obviously assume that K is a compact subset of some RI and that A = Σ(I) ∩ K. Since A is sequentially closed, our hypothesis implies that there exists a continuous injective map [0, ω] ∋ n 7→ xn ∈ K \ A. Let J be a countable subset of I such that xnJ 6= xmJ , for all n, m ∈ [0, ω] with n 6= m. Using [2, Lemma 1.2] and transfinite recursion, one easily obtains a family (Jα)α≤ω1 of K-good subsets of I satisfying the following conditions: • Jα is countable, for α < ω1; • J ⊂ J0; • Jα ⊂ Jβ, for 0 ≤ α ≤ β ≤ ω1; • Jα = Sβ<α Jβ, for limit α ∈ [0, ω1]; • for all n ∈ [0, ω], the map [0, ω1] ∋ α 7→ Jα ∩ supp xn is injective. Given these conditions, it is readily checked that the map [0, ω] × [0, ω1] ∋ (n, α) 7−→ rJα(xn) ∈ K is continuous and injective. (cid:3) Remark 4.5. The following converse of Proposition 4.4 also holds: if K is a Valdivia compact space containing a homeomorphic copy of [0, ω] × [0, ω1], then K \A contains a nontrivial convergent sequence, for any dense Σ-subset A of K. Namely, if φ : [0, ω] × [0, ω1] → K is a continuous injection, then φ(ω, α) ∈ K \ A for some α ∈ [0, ω1], since [0, ω1] is not Corson ([11, Ex- ample 1.10 (i)]). Moreover, the nontrivial sequence (cid:0)φ(n, α)(cid:1)n∈ω converges NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 10 to φ(ω, α). One consequence of this observation is that if K \ A contains a nontrivial convergent sequence for some dense Σ-subset A of K, then K \ A contains a nontrivial convergent sequence for any dense Σ-subset A of K. Remark 4.6. If a Valdivia compact space K admits two distinct dense Σ- subsets, then the assumption of Proposition 4.4 holds for K. Namely, given dense Σ-subsets A and B of K and a point x ∈ A \ B, then x is not isolated, since B is dense. Moreover, x is not isolated in A, because A is dense. Finally, since A is a Fr´echet -- Urysohn space ([11, Lemma 1.6 (ii)]), x is the limit of a sequence in A \ {x}. Finally, we observe that the validity of the following conjecture would imply, under CH, that Ext(cid:0)C(K), c0(cid:1) 6= 0 for any nonmetrizable Valdivia compact space K. Conjecture. If K is a nonempty Valdivia compact space having ccc, then either K has a Gδ point or K admits a nontrivial convergent sequence in the complement of a dense Σ-subset. To see that the conjecture implies the desired result, use Lemma 3.4 and Propositions 4.2 and 4.4, keeping in mind that a regular closed subset of a ccc space has ccc as well. The conjecture remains open, but in what follows we present an example showing that it is false if the assumption that K has ccc is removed. Recall that a tree is a partially ordered set (T, ≤) such that, for all t ∈ T , the set (·, t) = (cid:8)s ∈ T : s < t(cid:9) is well-ordered. As in [14, p. 288], we define a compact Hausdorff space from a tree T by considering the subspace P (T ) of 2T consisting of all characteristic functions of paths of T ; by a path of T we mean a totally ordered subset A of T such that (·, t) ⊂ A, for all t ∈ A. It is easy to see that P (T ) is closed in 2T ; we call it the path space of T . Denote by S(ω1) the set of countable successor ordinals and consider the 1 , partially ordered by inclusion. The path space P (T ) tree T = Sα∈S(ω1) ωα is the image of the injective map Λ ∋ λ 7→ χA(λ) ∈ 2T , where Λ = Sα≤ω1 ωα and A(λ) = (cid:8)t ∈ T : t ⊂ λ(cid:9). 1 Proposition 4.7. If the tree T is defined as above, then its path space P (T ) is a compact subspace of RT satisfying the following conditions: (a) P (T ) ∩ Σ(T ) is dense in P (T ), so that P (T ) is Valdivia; (b) P (T ) has no Gδ points; (c) no point of P (T )\Σ(T ) is the limit of a nontrivial sequence in P (T ). Proof. To prove (a), note that χA(λ) = limα<ω1 χA(λα) for all λ ∈ ωω1 1 . Let us prove (b). Since P (T ) is Valdivia, every Gδ point of P (T ) must be in Σ(T ) ([10, Proposition 2.2 (3)]), i.e., it must be of the form χA(λ), with λ ∈ ωα 1 , α < ω1. To see that χA(λ) cannot be a Gδ point of P (T ), it suffices to check that for any countable subset E of T , there exists µ ∈ Λ, µ 6= λ, such that χA(λ) and χA(µ) are identical on E. To this aim, simply take NONTRIVIAL TWISTED SUMS OF c0 AND C(K) 11 P (T ) converging to some ǫ ∈ P (T ) and note that the support of ǫ must be (cid:3) µ = λ ∪ {(α, β)}, with β ∈ ω1 \ (cid:8)t(α) : t ∈ E and α ∈ dom(t)(cid:9). Finally, to prove (c), let (cid:0)χA(λn)(cid:1)n≥1 be a sequence of pairwise distinct elements of contained in the countable set Sn6=m(cid:0)A(λn) ∩ A(λm)(cid:1). have ccc. Namely, setting Ut = (cid:8)ǫ ∈ P (T ) : ǫ(t) = 1(cid:9) for t ∈ T , we have It is easy to see that, for T defined as above, the space P (T ) does not that Ut is a nonempty open subset of P (T ) and that Ut ∩ Us = ∅, when t, s ∈ T are incomparable. References [1] S. A. Argyros, S. Mercourakis, S. Negrepontis, Analytic properties of Corson compact spaces, Proc. Fifth Prague Topological Symposium, Heldermann Verlag, Berlin, 1982, 12 -- 23. [2] S. A. Argyros, S. Mercourakis, S. Negrepontis, Functional-analytic properties of Cor- son compact spaces, Studia Math. 89 (1988) 197 -- 229. [3] A. Avil´es, F. Cabello, J. M. F. Castillo, M. Gonz´alez, Y. Moreno, On separably injective Banach spaces, Advances Math. 234 (2013) 192 -- 216. [4] F. Cabello, J. M. F. Castillo, N. J. Kalton, D. T. Yost, Twisted sums with C(K) spaces, Trans. Amer. Math. Soc. 355 (11) (2003) 4523 -- 4541. [5] F. Cabello, J. M. F. Castillo, D. Yost, Sobczyk's Theorem from A to B, Extracta Math. 15 (2) (2000) 391 -- 420. [6] J. M. F. Castillo, Nonseparable C(K)-spaces can be twisted when K is a finite height compact, to appear in Topology Appl. [7] J. M. F. Castillo, M. Gonz´alez, Three-space problems in Banach space theory, Lecture Notes in Mathematics, vol. 1667, Springer, Berlin, 1997. [8] C. Correa, D. V. Tausk, Extension property and complementation of isometric copies of continuous function spaces, Results Math. 67 (2015) 445 -- 455. [9] R. Hodel, Cardinal functions I, in: K. Kunen, J. E. Vaughan (Eds.), Handbook of set-theoretic topology, North -- Holand, Amsterdam, 1984. [10] O. Kalenda, Continuous images and other topological properties of Valdivia com- pacta, Fundamenta Math. 162 (1999) 181 -- 192. [11] O. Kalenda, Valdivia compact spaces in topology and Banach space theory, Extracta Math. 15 (1) (2000) 1 -- 85. [12] A. S. Kechris, Classical descriptive set theory, Graduate Texts in Mathematics, vol. 156, Springer Verlag, New York, 1995. [13] A. Sobczyk, Projection of the space (m) on its subspace (c0), Bull. Amer. Math. Soc. 47 (1941) 938 -- 947. [14] S. Todorcevi´c, Trees and linearly ordered sets, in: K. Kunen, J. E. Vaughan (Eds.), Handbook of set-theoretic topology, North -- Holand, Amsterdam, 1984. [15] M. Valdivia, Projective resolution of identity in C(K) spaces, Arch. Math. 54 (1990) 493 -- 498. Departamento de Matem´atica, Universidade de Sao Paulo, Brazil E-mail address: [email protected] Departamento de Matem´atica, Universidade de Sao Paulo, Brazil E-mail address: [email protected] URL: http://www.ime.usp.br/~tausk
1007.0430
1
1007
2010-07-02T18:48:56
Duality in reconstruction systems
[ "math.FA" ]
We consider the notion of finite dimensional reconstructions systems (RS's), which includes the fusion frames as projective RS's. We study erasures, some geometrical properties of these spaces, the spectral picture of the set of all dual systems of a fixed RS, the spectral picture of the set of RS operators for the projective systems with fixed weights and the structure of the minimizers of the joint potential in this setting. We give several examples.
math.FA
math
Duality in reconstruction systems ∗† P. G. Massey, M. A. Ruiz and D. Stojanoff Depto. de Matem´atica, FCE-UNLP, La Plata, Argentina and IAM-CONICET 0 1 0 2 l u J 2 ] Abstract We consider the notion of finite dimensional reconstructions systems (RS's), which includes the fusion frames as projective RS's. We study erasures, some geometrical properties of these spaces, the spectral picture of the set of all dual systems of a fixed RS, the spectral picture of the set of RS operators for the projective systems with fixed weights and the structure of the minimizers of the joint potential in this setting. We give several examples. . A F h t a m [ 1 v 0 3 4 0 . 7 0 0 1 : v i X r a Contents 1 Introduction 2 Basic framework of reconstruction systems 3 Erasures and lower bounds. 4 Geometric presentation of RS. 4.1 General Reconstruction systems . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Projective RS's with fixed weights . . . . . . . . . . . . . . . . . . . . . . . . 5 Spectral pictures 5.1 The set of dual RS's 5.2 RS operators of projective systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6 Joint potential of projective RS's 7 Examples and conclusions 1 Introduction 1 4 6 7 8 9 12 12 15 16 20 In this paper we study the notion of finite dimensional reconstruction systems, which gives a new framework for fusion and vector frames. Fusion frames (briefly FF's) were introduced under the name of "frame of subspaces" in [12]. They arise naturally as a generalization of ∗Keywords: Fusion frames, reconstruction systems, dual reconstruction systems. †Mathematics subject classification (2000): 42C15, 15A60. 1 the usual frames of vectors for a Hilbert space H. Several applications of FF's have been studied, for example, sensor networks [15], neurology [26], coding theory [6], [7] , [22], among others. We refer the reader to [14] and the references therein for a detailed treatment of the FF theory. Further developments can be found in [9], [13] and [27]. Given m ∈ N we denote by Im = {1, . . . , m} ⊆ N. In the finite dimensional setting, a FF is a sequence Nw = (wi , Ni)i∈Im where each wi ∈ R>0 and the Ni ⊆ Cd are subspaces that generate Cd. The synthesis operator of Nw is usually defined as given by TNw(xi)i∈Im =Pi∈Im wi xi . TNw : KNw def= Li∈Im Ni → Cd Its adjoint, the so-called analysis operator of Nw , is given by T ∗ Nwy = (wi PNi y)i∈Im for y ∈ Cd, where PNi denotes the orthogonal projection onto Ni . The frame Nw induces a linear encoding-decoding scheme that can be described in terms of these operators. The previous setting for the theory of FF's presents some technical difficulties. For ex- ample the domain of TNw relies strongly on the subspaces of the fusion frame. In particular, any change on the subspaces modifies the domain of the operators preventing smooth per- turbations of these objects. Moreover, this kind of rigidity on the definitions implies that the notion of a dual FF is not clear. An alternative approach to the fusion frame (FF) theory comes from the theory of pro- tocols introduced in [6] and the theory of reconstruction systems considered in [25] and [23]. In this context, we fix the dimensions dim Ni = ki and consider a universal space K = Km , k def= Mi∈ Im Cki , where k = (k1 , . . . , km) ∈ Nm . A reconstruction system (RS) is a sequence V = {Vi}i∈ Im such that Vi ∈ L(Cd , Cki) for every i ∈ Im , which allows the construction of an encoding-decoding algorithm (see Definition 2.1 for details). We denote by RS = RS(m, k, d) the set of all RS's with these fixed parameters. Observe that, if Nw = (wi , Ni)i∈Im is a FF, it can be modeled as a system V = {Vi}i∈ Im ∈ RS such that V ∗i Vi = w2 i PNi for every i ∈ Im . These systems are called projective RS's. On the other hand, a general RS arise from a usual vector frame by grouping together the elements of the frame. Thus, the coefficients involved in the encoding-decoding scheme of RS are vector valued, and they lie in the space K. The main advantage of the RS framework with respect to the fusion frame formalism is that each (projective) RS has many RS's that are dual systems. In particular, the canonical dual RS remains being a RS (for details and definitions see Section 2). In contrast, it is easy to give examples of a FF such that its canonical dual is not a fusion frame. There exists a notion of duality among fusion frames defined by Gavruta (see [19]), where the reconstruction formula of a fixed V involves the FF operator SV of V. Nevertheless, in the context of RS's, we show that the notion of dual systems can be described and characterized in a quite natural way. On the other hand, the RS framework (see Section 2 for a detailed description) allows to make not only a metric but also a differential geometric study of the set of RS's, which will be developed in Section 4 of this paper. Let us fix the parameters (m, k, d) and the sequence v = (vi)i∈Im ∈ Rm >0 of weights. In this work we study some properties of the sets RS = RS(m, k, d) of RS's and PRSv = PRSv (m, k, d) of projective systems with fixed weights v. In section 3 we study erasures in this context. We show conditions which guarantee that, after erasing some of its components, the system keeps being a RS, and we exhibit adequate bounds for it. In section 4 we present a geometrical description of both sets RS and PRSv , and give a sufficient condition (the 2 notion of irreducible systems) in order that the operation of taking RS operators PRSv ∋ V 7→ SV (see Definition 2.1) has smooth local cross sections. In section 5 we study the spectral picture of the set D(V) of all dual systems for a fixed V ∈ RS, and the set OPv of the RS operators of all systems in PRSv . Finally, in section 6 we focus on the main problem of the paper, which needs the results def= (cid:8) (V, W) ∈ PRSv × RS : W ∈ D(V)(cid:9). We look of the previous sections: Let DPv for pairs (V , W) which have the best minimality properties. If there exist tight systems in PRSv (systems whose RS operator is a multiple of the identity) then the pair (V , V #) is minimal, where V # is the canonical dual of V (see Defintion 2.3). Nevertheless, this is not always the case (see [10] or [25]). Therefore we define a joint RS potential given by DPv ∋ (V , W) 7→ RSP (V , W) = tr S2 W ∈ R>0 , which is similar to the potential used in [11] for vector frames. The minimizers of RSP are those pairs which are the best analogue of a tight pair. The main results in this direction are that: V + tr S2 • There exist λv = λv(m, k, d) ∈ Rd >0 such that a pair (V , W) ∈ DPv is a minimizer for the RSP if and only if W = V # and the vector of eigenvalues λ(SV) = λv . • Every such V can be decomposed as a orthogonal sum of tight projective RS's, where the quantity of components and their tight constants are the same for every minimizer. In section 7 we give some examples of these problems, showing sets of parameters for which the vector λv and all minimizers V ∈ PRSv can be explicitly computed. We also present a conjecture which suggest an easy way to compute the vector λv , as the minimal element in the spectral picture of OPv with respect to the majorization (see Conjecture 7.4). General notations. In this case, U∗ is called a coisometry. Given m ∈ N we denote by Im = {1, . . . , m} ⊆ N and 1 = 1m ∈ Rm denotes the vector with all its entries equal to 1. For a vector x ∈ Rm we denote by x↓ the rearrangement of x in a decreasing order, and (Rm)↓ = {x ∈ Rm : x = x↓} the set of ordered vectors. Given H ∼= Cd and K ∼= Cn, we denote by L(H, K) the space of linear operators T : H → K. Given an operator T ∈ L(H, K), R(T ) ⊆ K denotes the image of T , ker T ⊆ H the null space of T and T ∗ ∈ L(K, H) the adjoint of T . If d ≤ n we say that U ∈ L(H, K) is an isometry if U∗U = IH . If K = H we denote by L(H) = L(H , H), by Gl (H) the group of all invertible operators in L(H), by L(H)+ the cone of positive operators and by Gl (H)+ = Gl (H)∩L(H)+. If T ∈ L(H), we denote by σ(T ) the spectrum of T , by rk T the rank of T , and by tr T the trace of T . By fixing orthonormal basis (onb) of the Hilbert spaces involved, we shall identify operators with matrices, using the following notations: By Mn,d(C) ∼= L(Cd , Cn) we denote the space of complex n × d matrices. If n = d we write Mn(C) = Mn,n(C). H(n) is the R-subspace of selfadjoint matrices, Gl (n) the group of all invertible elements of Mn(C), U(n) the group of unitary matrices, Mn(C)+ the set of positive semidefinite matrices, and Gl (n)+ = Mn(C)+ ∩ Gl (n). If d ≤ n, we denote by I(d , n) ⊆ Mn , d(C) the set of isometries, i.e. those U ∈ Mn , d(C) such that U∗U = Id . If W ⊆ H is a subspace we denote by PW ∈ L(H)+ the orthogonal projection onto W , i.e. R(PW ) = W and ker PW = W ⊥. For vectors on Cn we shall use the euclidean norm. On the other hand, for matrices T ∈ Mn(C) we shall use both 3 1. The spectral norm kT k = kT ksp = max kxk=1 kT xk. 2. The Frobenius norm kT k2 = (tr T ∗T )1/2 =(cid:0) Pi,j∈In the inner product hA, Bi = tr B∗A , for A, B ∈ Mn(C). Tij2(cid:1)1/2. This norm is induced by 2 Basic framework of reconstruction systems In what follows we consider (m, k, d)-reconstruction systems, which are more general linear systems than those considered in [4], [6], [7], [8], [20] and [24], that also have an associated reconstruction algorithm. Definition 2.1. Let m, d ∈ N and k = (k1 , . . . , km) ∈ Nm. 1. We shall abbreviate the above description by saying that (m, k, d) is a set of parameters. We denote by n = tr k def= Pi∈ Im ki and assume that n ≥ d. 2. We denote by K = Km , k Cki ∼= Cn. We shall often write each direct summand by Ki = Cki . 3. Given a space H ∼= Cd we denote by def= Li∈ Im L(m, k, d) def= Mi∈Im L(H , Ki) ∼= L(H, K) ∼= Mi∈ Im Mki , d(C) ∼= Mn,d(C) . A typical element of L(m, k, d) is a system V = {Vi}i∈ Im such that each Vi ∈ L(H , Ki). 4. A family V = {Vi}i∈ Im ∈ L(m, k, d) is an (m, k, d)-reconstruction system (RS) for H if SV def= Xi∈ Im V ∗i Vi ∈ Gl (H)+ , (1) i.e., if SV is invertible. SV is called the RS operator of V. In this case, the m-tuple k = (k1 , . . . , km) ∈ Nm satisfies that n = tr k ≥ d. We shall denote by RS = RS(m, k, d) the set of all (m, k, d)-RS's for H ∼= Cd. 5. The system V is said to be projective if there exists a sequence v = (vi)i∈Im ∈ Rm + of positive numbers, the weights of V, such that Vi V ∗i = v2 i IKi , for every i ∈ Im . In this case, the following properties hold: (a) The weights can be computed directly, since each vi = kViksp . (b) Each Vi = viUi for a coisometry Ui ∈ L(H , Ki). Thus V ∗i Vi = v2 i PR(V ∗ i ) ∈ L(H)+ for every i ∈ Im . (c) SV =Pi∈ Im v2 i PR(V ∗ i ) as in fusion frame theory. We shall denote by PRS = PRS(m, k, d) the set of all projective elements of RS. 4 6. The analysis operator of the system V is defined by TV : H → K = Mi∈ Im Ki given by TV x = (V1 x , . . . , Vm x) , for x ∈ H . 7. Its adjoint T ∗ T ∗ V is called the synthesis operator of the system V, and it satisfies that V : K = Mi∈ Im V(cid:0) (yi)i∈ Im(cid:1) = Xi∈ Im Using the previous notations and definitions we have that SV = T ∗ V TV . Ki → H is given by V ∗i yi . T ∗ 8. The frame constants in this context are the following: V is a RS if and only if AV kxk2 ≤ hSV x , xi = Xi∈Im kVi xk2 ≤ BV kxk2 (2) for every x ∈ H, where 0 < AV = λmin(SV ) = kS−1 V 9. As usual, we say that V is tight if AV = BV . k−1 ≤ λmax(SV) = kSVk = BV . In other words, the system V ∈ RS(m, k, d) is tight if and only if SV = τ 10. The Gram matrix of V is GV = TV T ∗ d IH , where τ =Pi∈Im v2 V ∈ L(K)+ ∼= Mn(C)+, where the size of GV i ki . viewed as a matrix is n = tr k =Pi∈Im ki = dim K. 11. Given U ∈ Gl (d), we define V · U def= {Vi U}i∈ Im ∈ RS(m, k, d). Remark 2.2. Let V = {Vi}i∈ Im ∈ RS such that every Vi 6= 0. In case that k = 1m , then V can be identified with a vector frame, since each Vi : Cd → C is in fact a vector 0 6= fi ∈ Cd. In the same manner, the projective RS's can be seen as fusion frames. Here the identification △ is given by Vi ≃(cid:0) kVik , R(V ∗i )(cid:1) for every i ∈ Im . Definition 2.3. For every V = {Vi}i∈ Im ∈ RS(m, k, d), we define the system △ = {Vi S−1 V called the canonical dual RS associated to V. V # def= V · S−1 V }i∈ Im ∈ RS(m, k, d) , Remark 2.4. Given V = {Vi}i∈ Im ∈ RS with SV =Pi∈ Im V ∗i Vi , then V ∗i Vi SV −1 = IH . SV −1 V ∗i Vi = IH , and Xi∈ Im Xi∈ Im △ (3) Therefore, we obtain the reconstruction formulas x = Xi∈ Im S−1 V V ∗i (Vi x) = Xi∈ Im V ∗i Vi(S−1 V x) for every x ∈ H . (4) Observe that, by Eq. (3), we see that the canonical dual V # satisfies that T ∗ V # TV = Xi∈ Im SV −1 V ∗i Vi = IH and SV # = Xi∈ Im S−1 V V ∗i Vi S−1 V = S−1 V . Next we generalize the notion of dual RS's : (5) △ 5 Definition 2.5. Let V = {Vi}i∈ Im and W = {Wi}i∈ Im ∈ RS. We say that W is a dual RS for V if T ∗ W TV = IH , or equivalently if x =Pi∈ Im W ∗i Vi x for every x ∈ H. W TV = IH . Observe that the map RS ∋ W 7→ T ∗ We denote the set of all dual RS's for a fixed V ∈ RS by D(V) def= {W ∈ RS : T ∗ Observe that D(V) 6= ∅ since V # ∈ D(V). W TV = IH } . △ Remark 2.6. Let V ∈ L(m, k, d). Then V ∈ RS ⇐⇒ T ∗ V is surjective. In this case, a system W ∈ D(V) if and only if its synthesis operator T ∗ W is a pseudo-inverse of TV . Indeed, W ∈ D(V) ⇐⇒ T ∗ W is one to one. Thus, in the context of RS's each (m, k, d)-RS has many duals that are (m, k, d)-RS's. This is one of the advantages of the RS's setting. Moreover, the synthesis operator T ∗ V # of the canonical dual V # corresponds to the Moore- Penrose pseudo-inverse of TV . Indeed, notice that TV T ∗ ∈ L(K)+, so that it is an orthogonal projection. From this point of view, the canonical dual V # has some optimal properties that come from the theory of pseudo-inverses. On the other hand the map L(m, k, d) ∋ W 7→ T ∗ W ∈ L(K, H) is R-linear. Then, for every V ∈ RS, the set D(V) of dual systems is convex in L(m, k, d), because the set of △ pseudoinverses of TV is convex in L(K, H). V # = TV S−1 V T ∗ V 3 Erasures and lower bounds. It is a known result in frame theory that, for a given frame F = {fi}i∈I, the set F′ = {fi}i∈I, i6=j is either a frame or a incomplete set for H. In [13] P. Casazza and G. Kutyniok give examples where this situation does not occur in the fusion frame setting. Considering fusion frames as a particular case of reconstruction systems we can rephrase their result in the following way: Theorem 3.1 (Casazza and Kutyniok). Let V = {Vi}i∈Im ∈ PRS with bounds AV , BV. If def= {Vi}i∈Im\J is a projective RS for H ∼= Cd with Pi∈J kVik2 < AV then the sequence VJ bounds AVJ ≥ AV −Pi∈J kVik2 and BVJ ≤ BV . As they notice in [13] with an example, this is not a necessary condition. On the other side, in [3], M. G. Asgari proves that, under certain conditions, a single element can be erased from the original fusion frame (in our setting, a projective RS), and he obtains different lower bounds for the resulting reconstruction system: Theorem 3.2 (Asgari). Let V = {Vi}i∈Im ∈ PRS with bounds AV , BV . Suppose that there ∈ Gl (d), then V j = {Vi}i6=j is a projective RS exists j ∈ Im such that Mj for H ∼= Cd with bounds AVJ ≥ def= Id − V ∗j VjS−1 V j k2 and BVJ ≤ BV . AV +kVjk2kM −1 A2 V Actually, Asgari's result can be generalized to any subset J of Im and general RS's. In the following statement we shall give necessary and sufficient conditions which guarantee that the erasure of {Vi}i∈J of a non necessary projective V = {Vi}i∈ Im ∈ RS provides another RS. Recall that the sharp bounds for V are given by AV = kS−1 V Theorem 3.3. Let V = {Vi}i∈ Im ∈ RS(m, k, d) with bounds AV , BV. Fix a subset J ⊂ Im and consider the matrix MJ k−1 and BV = kSVk. ∈ Md(C). Then, def= Id −Pi∈J V ∗i Vi S−1 V VJ = (Vi)i∈Im\J is a RS for H ∼= Cd ⇐⇒ MJ ∈ Gl (d) . (6) 6 In this case SVJ = MJ SV and VJ has bounds AVJ ≥ AV kM −1 J k Proof. The equality SVJ = MJ SV follows from the following fact: =Xi /∈J MJ = Id −Xi∈J −Xi∈J = SV S−1 V V ∗i Vi S−1 V V ∗i Vi S−1 V and BVJ ≤ BV . V ∗i Vi S−1 V = SVJ S−1 V . This implies the equivalence of Eq. (6). On the other hand, AV kM−1 J k = kS−1 V k−1 kM−1 J k−1 ≤ k(MJ SV)−1k−1 = kS−1 VJ k−1 = AVJ . The fact that 0 < SVJ ≤ SV assures that BVJ ≤ BV . In the case J = {j}, the lower bound in Theorem 3.3 is greater than that obtained in [3]: (cid:3) Proposition 3.4. Let V, and MJ be as in Theorem 3.3, with J = {j}. Then A2 V AV + kVjk2kM−1 J k2 ≤ AV kM−1 J k . Proof. We can suppose kM−1 J k ≥ 1, since otherwise (7) is evident. Note that J k2 ≤ AV kM −1 J k J k2 ≤ AV kM −1 AV +kVjk2kM −1 A2 A2 V . kVjk2 ≥ AV =⇒ J k2 ≤ But if kVjk2 < AV , then kId − MJ k = kV ∗j Vj S−1 V V kVjk2kM −1 k ≤ kVjk2 AV < 1. Therefore kM−1 J k ≤ AV AV−kVjk2 =⇒ AV kM−1 J k ≤ AV + kVjk2kM−1 J k ≤ AV + kVjk2kM−1 J k2 , which clearly implies the inequality of Eq. (7). (7) (cid:3) Remark 3.5. Let J ⊆ Im , V ∈ RS, and MJ be as in Theorem 3.3. Assume that kPi∈J V ∗i Vik < AV (compare with the hypothesis Pi∈J kVik2 < AV of Theorem 3.1). Then, as in the proof of Proposition 3.4, it can be shown that under this assumption it holds that kId − MJ k < 1 =⇒ MJ ∈ Gl (d) and that the lower bounds satisfy AV −Pi∈J kVik2 ≤ AV − kPi∈J V ∗i Vik ≤ AV kM −1 J k ≤ AVJ . Hence Theorem 3.3 generalizes Theorem 3.1 to general RS's with better bounds. The matrix MJ also serves to compute the canonical dual system (VJ )#: If we denote V # = {Wi}i∈Im and V # J = {Wi}i /∈J , then the formula SVJ = MJ SV of Theorem 3.3 gives the equality V # J · M−1 J def= {Wi M−1 J }i /∈J = {Vi S−1 V M−1 J }i /∈J = {Vi S−1 VJ }i /∈J = (VJ )# . That is, (VJ )# is the truncation of the canonical dual V # modified with M−1 J . △ 4 Geometric presentation of RS. In this section we shall study several objects related with the sets RS from a geometrical point of view. On one hand, this study is of independent interest. On the other hand, some geometrical results of this section will be necessary in order to characterize the minimizers of the joint potential, a problem that we shall consider in Section 6. 7 4.1 General Reconstruction systems 4.1. Observe that we can use on L(m, k, d) the natural metric kVk2 =(cid:0) Pi∈Im kVik2(cid:1)1/2 for V = {Vi}i∈ Im ∈ L(m, k, d). Note that kVk2 2 = Xi∈Im kVik2 2 = kTVk2 2 (cid:0) in the space L(H, K) (cid:1) . With this metric it is easy to see that in RS ⊆ L(m, k, d) the following conditions hold: 1. The space RS is open in L(m, k, d), since the map RSO : L(m, k, d) → L(H) given by RSO(V) = SV = T ∗ V TV (for V ∈ L(m, k, d) ) is continuous. because the map L(m, k, d) ∋ W 7→ T ∗ equality T ∗ 2. On the other hand, if we fix V ∈ RS, then the set D(V) is closed in L(m, k, d), W TV ∈ L(H) is continuous. Observe that the △ 4.2. Given a surjective A ∈ L(K , H ), let us consider PKi A∗ ∈ L(H , Ki) for every i ∈ Im . Then A produces a system WA = (PKi A∗)i∈Im ∈ RS such that W is surjective, so that W ∈ RS. W TV = IH =⇒ T ∗ T ∗ WA = A and SWA = AA∗ ∈ Gl (H)+ . Therefore, given a fixed V = {Vi}i∈ Im ∈ RS, we can parametrize RS = {U · V def= (PKi U TV)i∈Im : U ∈ Gl (K)} . In other words, the Lie group Gl (K) acts transitively on RS, where the action is given by the formula U · V = (PKi U TV)i∈Im . Indeed, for every x = (xi)i∈Im ∈ K, V U∗ PKi x = T ∗ T ∗ T ∗U·V x = Pi∈Im V U∗ ∈ L(K, H), which is surjective for every U ∈ Gl (K), so that Therefore T ∗U·V = T ∗ U · V ∈ RS. Hence TU·V = U TV , which shows that this is indeed an action. On the other hand, for every W ∈ RS, since both T ∗ V are surjective, then there exists U ∈ Gl (K) such that T ∗ Fix V = {Vi}i∈ Im ∈ RS. Then we can define a continuous surjective map V U∗. Therefore we have that W = U · V. V U∗ Pi∈Im PKi x = T ∗ W and T ∗ V U∗ x . W = T ∗ (8) πV : Gl (K) → RS given by πV (U) = U · V for U ∈ Gl (K) . The isotropy subgroup of this action is IV = π−1 V Indeed, looking at Eq. (8) we see that U · V = V ⇐⇒ T ∗ V ⇐⇒ U TV = TV . In [16] it is proved that these facts are sufficient to assure that RS is a smooth submanifold of L(m, k, d) (actually it is an open subset) such that the map πV : Gl (K) → RS becomes a smooth submersion. On the other hand, we can parametrize D(V) in two different ways : (V) = {U ∈ Gl (K) : U(cid:12)(cid:12)R(TV ) = Id }. V U∗ = T ∗ D(V) =nW ∈ L(m, k, d) : T ∗ W = T ∗ V # + G , G ∈ L(K, H) and G(cid:12)(cid:12)R(TV ) ≡ 0o Indeed, just observe that ker T ∗ =nU · V # : U ∈ Gl (K) and P U∗ P = Po , where P = PR(TV ) . T ∗U·V # TV = T ∗ V # U∗ TV = IH ⇐⇒ U∗ x ∈ x + ker T ∗ V = R(TV )⊥ = ker P . Therefore V # = ker S−1 V # for every x ∈ R(P ) , V T ∗ which means exactly that P U∗P = P . 8 (9) △ Remark 4.3. This geometric presentation is similar to the presentation of vector frames done in [16]. The relationship is based on the following fact: The space RS can be seen as an agrupation in packets of vector frames. Recall that n = } the canonical ONB of each Ki = Cki, tr k =Pi∈Im ki . Let us denote by Ei = {e(i) and the set E = Si∈Im Ei , which is a reenumeration of the canonical ONB of the space K = Li∈ Im ∼= Cn. Then, there is a natural one to one correspondence 1 , . . . , e(i) Ki ki RS ∋ V = {Vi}i∈ Im ←→(cid:16) ( V ∗i e(i) j )j∈Iki(cid:17)i∈Im =(cid:0) T ∗ V e(cid:1)e∈ E ∈ Hn , (10) where the right term is a general n-vector frame for H. On the other hand, fixed the ONB E of K, the set of n-vector frames for H can be also identified with the space E(K , H) def= {A ∈ L(K, H) : A is surjective }, via the map A ←→(cid:0) A e(cid:1)e∈E The geometrical representation of RS given before is the natural geometry of the space of epimorphisms E(K , H) under the (right) action of Gl (K). Through all these identifications we get the correspondence RS ∋ V ←→ T ∗ Observe that, in terms of Eq. (10), a system V = {Vi}i∈ Im ∈ RS satisfy that V ∈ PRS ⇐⇒ each subsystem ( V ∗i e(i) △ is a multiple of an orthonormal system in H. V ∈ E(K , H). . j )j∈Iki 4.2 Projective RS's with fixed weights Given a fixed sequence of weights v = (vi)i∈Im ∈ Rm with fixed set of weights v: >0 , we define the set of projective RS's PRSv def= n V = {Vi}i∈ Im ∈ PRS : kViksp = vi for every i ∈ Imo . i ki . Observe that tr SV = Pi∈Im v2 tr V ∗i Vi = τ for every V ∈ PRSv . In (11) Denote by τ = Pi∈Im what follows we shall denote by Md(C)+ τ def= {A ∈ Md(C)+ : tr A = τ } and Gl (d)+ τ def= Md(C)+ τ ∩ Gl (d) . the set of d × d positive and positive invertible operators with fixed trace τ , endowed with the metric and geometric structure induced by those of Gl (d). In this section we look for conditions which assure that the smooth map RSO : PRSv → Gl (d)+ τ given by RSO(V) = SV = Xi∈Im V ∗i Vi , (12) for every V = {Vi}i∈ Im ∈ PRSv , has smooth local cross sections. Before giving these conditions and the proof of their sufficiency, we need some notations and two geometrical lemmas: Fix d ∈ N. For every k ∈ Id , we denote by I(k , d) = {U ∈ L(Ck , Cd) : U∗U = Ik} the set of isometries. Given an m-tuple k = (ki)i∈Im ∈ Im d ⊆ Nm, we denote by I(k , d) def= Mi∈Im I(ki , d) ⊆ Mi∈Im L(Ki , H) ∼= L(K , H) , 9 endowed with the product (differential, metric) structure (see [1] for a description of the geometrical structure). Similarly, let Gr(k, d) denote the Grassmann manifold of orthogonal projections of rank k in Cd and let Gr (k , d) def= Mi∈Im Gr (ki , d) ⊆ L(H)m , with the product smooth structure (see [17]). Lemma 4.4. Consider the smooth map Φ : I(k , d) → Gr (k , d) given by Φ(W) = (W1 W ∗1 , . . . , Wm W ∗m) for every W = {Wi}i∈ Im ∈ I(k , d) . Then Φ has smooth local cross sections around any point P = (Pi)i∈Im ∈ Gr (k , d) toward every W ∈ I(k , d) such that Φ(W) = P. In particular, Φ is open and surjective. Proof. Since both spaces have a product structure, it suffices to consider the case m = 1. It is clear that the map Φ is surjective. For every P ∈ Gr (k , d), the C∞ map πP : U(d) → Gr (k , d) given by πP (U) = UP U∗ for U ∈ U(d) is a submersion with a smooth local cross section (see [17]) hP : UP def= {Q ∈ Gr (k , d) : kQ − P k < 1} → U(d) such that hP (P ) = Id . For completeness we recall that, for every Q ∈ UP , the matrix hP (Q) is the unitary part in the polar decomposition of the invertible matrix QP + (Id − Q)(Id − P ). Then, fixed W ∈ I(k , d) such that Φ(W ) = P , we can define the following smooth local cross section for Φ : sP , W : UP → I(k , d) given by sP , W (Q) = hP (Q) W , for every Q ∈ UP . (cid:3) We shall need the following result from [25]. notions and introduce some notations: In order to state it we recall the following 1. Fix v = (vi)i∈Im ∈ Rm >0 . We shall consider the smooth map Ψv : I(k , d) → Md(C)+ given by Ψv(U) = Xi∈Im i Ui U∗i v2 (13) for every U = {Ui}i∈ Im ∈ I(k , d). 2. Given a set P = {Pj : j ∈ Im} ⊆ Md(C)+, we denote by P′ = {Pj : j ∈ Im}′ = {A ∈ Md(C) : APj = Pj A for every j ∈ Im} . (14) Note that P′ is a (closed) unital selfadjoint subalgebra of Md(C). Therefore, P′ 6= C Id ⇐⇒ there exists a non-trivial orthogonal projection Q ∈ P′ . (15) Lemma 4.5 ([25]). Let v = (vi)i∈Im ∈ Rm τ =Pi∈Im v2 i ki . Then the map Sv : Gr (k , d) → Md(C)+ τ given by >0 and P = {Pi}i∈Im ∈ Gr (k , d). Denote by Sv(Q) = Xi∈Im v2 i Qi for Q = {Qi}i∈Im ∈ Gr (k , d) (16) is smooth and, if P satisfies that P′ = C Id , then 10 1. The matrix Sv(P) ∈ Gl (d)+ τ . 2. The image of Sv contains an open neighborhood of Sv(P) in Md(C)+ τ . 3. Moreover, Sv has a smooth local cross section around Sv(P) towards P. (cid:3) 4.6. The set I0(k , d) = {W ∈ I(k , d) : Sv ◦ Φ(W) ∈ Gl (d)+} is open in I(k , d). Observe that its definition does not depend on the sequence v = (vi)i∈Im ∈ Rm >0 of weights. Moreover, the map γ : I0(k , d) → PRSv given by γ(W) = {vi W ∗i }i∈Im ∈ PRSv for every W = {Wi}i∈ Im ∈ I0(k , d) , (17) is a homeomorphism. Hence, using this map γ we can endow PRSv with the differential structure which makes γ a diffeomorphism. With this structure, each space PRSv becomes a submanifold of RS. It is in this sense in which the map RSO : PRSv → Gl (d)+ τ defined in Eq. (12) is smooth. Indeed, we have that RSO = Sv ◦ Φ ◦ γ−1 , (18) where Φ : I(k , d) → Gr (k , d) is the smooth map defined in Lemma 4.4. Now we can give △ an answer to the problem posed in the beginning of this section. Definition 4.7. Let v = (vi)i∈Im ∈ Rm >0 and V = {Vi}i∈ Im ∈ PRSv (m, k, d). We say that △ the system V is irreducible if CV In Section 7 we show examples of reducible and irreducible systems. See also Remark 6.5. Theorem 4.8. Let v = (vi)i∈Im ∈ Rm >0 and τ = Pi∈Im V ∈ PRSv (m, k, d), then the map RSO : PRSv → Gl (d)+ local cross section around SV which sends SV to V. Proof. We have to prove that there exists an open neighborhood A of SV in Gl (d)+ smooth map ρ : A → PRSv such that RSO ( ρ(S) ) = S for every S ∈ A and ρ(SV) = V. Denote by Pi = PR(V ∗ v2 i ki . If we fix an irreducible system i ) for every i ∈ Im , and consider the system def= {V ∗i Vi : i ∈ Im}′ = C Id . τ defined in Eq. (12) has a smooth τ and a γ−1(V) = U = {Ui}i∈ Im ∈ I(k , d) given by Ui = v−1 i V ∗i ∈ I(ki , d) i ∈ Im . Observe that Φ(U) = P = {Pi}i∈Im ∈ Gr (k , d) and Sv(P) = SV . By our hypothesis, we know that P′ = {V ∗i Vi : i ∈ Im}′ = C Id . Let α : A → Gr (k , d) be the smooth section for the map Sv : Gr (k , d) → Md(C)+ τ given by Lemma 4.5. Hence A is an open neighborhood of SV = Sv(P) in Gl (d)+ Take now the cross section β : B → I(k , d) for the map Φ : I(k , d) → Gr (k , d) given by Lemma 4.4, such that B is an open neighborhood of P in Gr (k , d), and that β(P) = U. τ , and α(SV) = P. Finally we recall the diffeomorphism γ : I0(k , d) → PRSv defined in Eq. (17), where I0(k , d) = {W ∈ I(k , d) : Sv ◦ Φ(W) ∈ Gl (d)+} is an open subset of I(k , d) such that U ∈ I0(k , d). Note that γ(U) = V. Changing the first neighborhood A by some smaller open set, we can define the announced smooth cross section for the map RSO by ρ = γ ◦ β ◦ α : A ⊆ Gl (d)+ τ → PRSv . Following our previous steps, we see that ρ(SV) = V and that RSO (18) = Sv ◦ Φ ◦ γ−1 =⇒ RSO(ρ(S) ) = S for every S ∈ A . (cid:3) 11 Remark 4.9. In order to compute "local" minimizers for different functions defined on RS or some of its subsets, we shall consider two different (pseudo) metrics: Given V = {Vi}i∈Im and W = {Wi}i∈Im ∈ RS, we recall the (punctual) metric defined in 4.1: dP (V, W) = Xi∈Im kVi − Wik2 2!1/2 = kTV − TWk2 = kT ∗ V − T ∗ Wk2 . We consider also a pseudo-metric defined by dS(V, W) = kSV − SWk . Let A ⊆ RS and f : A → R a continuous map. Fix V = {Vi}i∈ Im ∈ A. Since the map V 7→ SV is continuous, it is easy to see that if V is a local dS minimizer of f over A, then V is also a local dP minimizer. The converse needs not to be true. Nevertheless, it is true under some assumptions: Theorem 4.8 shows that if V is a local dP minimizer of f : PRSv → R, in order to assure that V is also a local dS minimizer it suffices to assume that {V ∗i Vi : i ∈ Im}′ = C Id , i.e. that V is irreducible. △ 5 Spectral pictures +)↓ is the set of vectors µ ∈ Rd Recall that (Rd the entries are positive (i.e., if µd > 0), we write µ ∈ (Rd λ(S) ∈ (Rd by S† the Moore-Penrose pseudo-inverse of S. We shall also use the following notations: + with non negative and decreasing entries. If all >0)↓. Given S ∈ Md(C)+, we write +)↓ the decreasing vector of eigenvalues of S, counting multiplicities. We denote 1. Given x ∈ Cn then D(x) ∈ Md(C) denotes the diagonal matrix with main diagonal x. 2. If d ≤ n and y ∈ Cd, we write (y , 0n−d) ∈ Cn, where 0n−d is the zero vector of Cn−d. In this case, we denote by Dn(y) = D(cid:0) (y , 0n−d)(cid:1) ∈ Cn. Given A ⊆ Md(C)+ we consider its spectral picture: Λ(A) = {λ(A) : A ∈ A} ⊆ (Rd +)↓ , We say that Λ(A) determines A whenever A ∈ A if and only if λ(A) ∈ Λ(A). It is easy to see that this happens if and only if the set A is saturated with respect to unitary equivalence. 5.1 The set of dual RS's Definition 5.1. Let V ∈ RS. We denote by Λ(D(V) ) = {λ(SW) : W ∈ D(V)} ⊆ (Rd >0)↓ , that is, the spectral picture of the set of all dual RS's for V. (19) △ The following result gives a characterization of Λ(D(V)). Theorem 5.2. Let V = {Vi}i∈ Im ∈ RS and µ ∈ (Rd following conditions are equivalent: >0)↓. We denote by n = tr k. Then the 1. The vector µ ∈ Λ(D(V)). 12 2. There exists an orthogonal projection P ∈ Mn(C) such that rk P = d and λ (P Dn(µ) P ) =(cid:0) λ(S−1 Proof. Let W ∈ D(V) with λ(SW ) = µ. Then T ∗ where GV = TV T ∗ V V ∈ Mn(C)+ is the Gram matrix of V. W TV = I and ), 0n−d(cid:1) = λ(G† V ) , (20) GV GW GV = TV (T ∗ V TW ) (T ∗ V = GV =⇒ Q GW Q = G† V = PR(TV ) . Note that rk Q = rk TV = d, since V is a RS. Also V = TV T ∗ W TV ) T ∗ where Q = GV G† V λ(GW) = λ(TW T ∗ W ) = (λ(T ∗ Then there exists U ∈ U(n) such that W TW ), 0n−d) = (λ(SW), 0n−d) = (µ , 0n−d) . , (21) (22) U∗ D(µ , 0n−d) U = U∗ Dn(µ) U = U∗ Dn(cid:0) λ(SW )(cid:1) U = TW T ∗ Let P = U Q U∗. Note that rk P = rk Q = d. Using (21) and (22) we get the item 2 : W = GW . λ (P Dn(µ) P ) = λ(U Q U∗ Dn(µ) U Q U∗) (22) = λ(Q GW Q) (21) = λ(G† V ) = (λ(S−1 V ) , 0n−d) . Conversely, assume that there exists the projection P ∈ Mn(C)+ of item 2. Observe that there always exists U ∈ RS such that λ(SU ) = λ(T ∗ λ(GU ) = λ(TU T ∗ U ) = (µ , 0n−d) ∈ (Rn U TU ) = µ. Then +)↓ . Let V ∈ U(n) such that V ∗ GU V = Dn(µ). Denote by Q = V P V ∗. Then we get that λ(Q GU Q) = λ(P V ∗ GU V P ) = λ (P Dn(µ) P ) Then there exists W ∈ U(n) such that W ∗ (Q GU Q) W = G† V (20) ), 0n−d) = λ(G† = (λ(S−1 V V . Observe that ) . (23) rk Q = d and W ∗(R(Q) ) ⊇ R(G† V ) = R(GV) = R(TV ) =⇒ W ∗QW = PR(TV ) . Moreover, GV G† V GV = GV G† V = G† V GV = PR(GV ) = PR(TV ) = W ∗QW . Then GV = GV (W ∗ Q GU Q W ) GV = GV PR(GV ) (W ∗ GU W ) PR(GV ) GV = GV (W ∗ GU W ) GV . We can rewrite this fact as TV(cid:0) T ∗ V W ∗TU ) (T ∗ Finally, take W = {PKi W TU Vd}i∈Im ∈ L(m, k, d). Observe that U W TV) = IH =⇒ Vd = T ∗ U W TV(cid:1) T ∗ V W ∗TU T ∗ V = TV T ∗ V . Since T ∗ U W TV ∈ U(d) . (T ∗ V is surjective, (24) V ∗d T ∗ U W ∗ PKi W TU Vd = V ∗d T ∗ U TU Vd = V ∗d SU Vd ∈ Gl (d)+ . Then W ∈ RS and λ(SW ) = λ(SU ) = µ. Similarly, TW = W TU Vd . By Eq. (24), we deduce that T ∗ U W TV = V ∗d Vd = IH , so that W ∈ D(V). (cid:3) SW = Xi∈Im W TV = V ∗d T ∗ 13 Remark 5.3. Let V ∈ RS and µ ∈ (Rd (20) can be characterized in terms of interlacing inequalities. >0)↓ as in Theorem 5.2. It turns out that condition More explicitly, let us denote by >0)↑ , for every i ∈ Id . Similarly, we denote by ρ = λ(S−1 >0)↑. K. Fan and G. Pall showed V that the existence of a projection P satisfying (20) is equivalent to the following inequalities: so that )↑ = (λi(SV)−1)i∈Im ∈ (Rd γi = µd−i+1 γ = µ↑ ∈ (Rd 1. µd−i+1 = γi ≥ ρi = λi(SV)−1 for every i ∈ Id . 2. If n = tr k < 2 d and we denote r = 2 d − n ∈ N, then γi ≤ ρi+n−d = λi+n−d(SV)−1 = λ2 d−n−i+1(S−1 V ) if 1 ≤ i ≤ r . This fact together with Theorem 5.2 give a complete description of the spectral picture of △ the RS operators SW for every W ∈ D(V), which we write as follows. >0)↓. Then, the set Corollary 5.4. Let V = {Vi}i∈ Im ∈ RS, n = tr k and fix µ ∈ (Rd Λ(D(V)) can be characterized as follows: 1. If n ≥ 2 d, we have that µ ∈ Λ(D(V)) ⇐⇒ µj ≥ λj(S−1 V ) = λd−j+1(SV)−1 for every j ∈ Id . (25) 2. If n < 2 d, then µ ∈ Λ(D(V)) ⇐⇒ µ satisfies (25) and also the following conditions: µ↑i = µd−i+1 ≤ λi+n−d(SV )−1 = λ2 d−n−i+1(S−1 V ) for every i ≤ 2 d − n . (26) Proof. It is a direct consequence of Theorem 5.2 and the Fan-Pall inequalities described in Remark 5.3. (cid:3) Corollary 5.5. Let V ∈ RS. Then Λ(D(V)) is a convex set. Proof. It is clear that the inequalities given in Eqs. (25) and (26) are preserved by convex combinations. Observe that also the set (Rd (cid:3) Corollary 5.6. Let V = {Vi}i∈ Im ∈ RS. If W ∈ D(V) then >0)↓ is convex. d RSP (W) def= tr S2 W ≥ tr S−2 V = λ(SV)−2 i = RSP (V #) . (27) Xi=1 Moreover, V # is the unique element of D(V) which attains the lower bound in (27). Proof. The inequality given in Eq. (27) is a direct consequence of (25). With respect to the uniqueness of V #, fix another W ∈ D(V). Then the equalities T ∗ V # TV = I imply that T ∗ V # + A, for some A ∈ L(K, H) that satisfies A TV = 0. With respect to V #, note that R(TV #) = R(TV S−1 V W TV = T ∗ W = T ∗ ) = R(TV) ⊆ ker A, so that also A TV # = 0. Thus, V #TV #(cid:1) + tr (cid:0)AA∗(cid:1) + 2 Re tr (cid:0)A TV #(cid:1) 2 = tr (cid:0)T ∗ tr SW = kT ∗ V # + Ak2 = tr SV # + kAk2 2 . (28) On the other hand, if the lower bound in Eq. (27) is attained W, using (25) we can deduce that λ(SW) = λ(SV #). Then also tr SW = tr SV # . But the previous equality forces that in this case A = 0 and hence W = V #. (cid:3) 14 5.2 RS operators of projective systems In this section we shall fix the parameters (m, k, d) and the sequence v = (vi)i∈Im ∈ Rm >0 of weights. Now we give some new notations: First, recall that the set of projective RS's with fixed set of weights v is PRSv = PRSv (m, k, d) =n {Vi}i∈Im ∈ PRS : kViksp = vi for every i ∈ Imo . We consider the set of operators SV for V ∈ PRSv and its spectral picture: OPv def= {SV : V ∈ PRSv } and Λ(OPv) def= {λ(S) : S ∈ OPv} ⊆ (Rd >0)↓ . (29) We shall give a characterization of the set Λ(OPv) in terms of the Horn-Klyachko's theory of sums of hermitian matrices. In order to do this we shall describe briefly the basic facts about the spectral characterization obtained by Klyachko [21] and Fulton [18]. Let Kr d =(cid:8)(j1, . . . , jr) ∈ (Id) r : j1 < j2 . . . < jr(cid:9) . For J = (j1, . . . , jr) ∈ Kr d , define the associated partition λ(J) = (jr − r, . . . , j1 − 1) . For d)m+1, such that the r ∈ Id−1 denote by LR r Littlewood-Richardson coefficient of the associated partitions λ(J0), . . . , λ(Jm) is positive, i.e. one can generate the Young diagram of λ(J0) from those of λ(J1), . . . , λ(Jm) according to the Littlewood-Richardson rule (see [18]). d (m) the set of (m + 1)-tuples (J0, . . . , Jm) ∈ (Kr The theorem of Klyachko gives a characterization of the spectral picture of the set of all sums of m matrices in H(d) with fixed given spectra, in terms on a series of inequalities involving the (m + 1)-tuples in LR r d (m) (see [21] for a detailed formulation). We give a description of this result in the particular case where these m matrices are multiples of projections: Lemma 5.7. Fix the parameters (m, k, d) and v ∈ Rm >0 µ ∈ (Rm + )↓ . Then there exists a sequence {Pi}i∈Im ∈ Gr (k , d) such that µ = λ(cid:0) Pi∈Im v2 for every r ∈ Id−1 and every (m + 1)-tuple (J0, . . . , Jm) ∈ LR r Proposition 5.8. Fix the parameters (m, k, d) and the vector v ∈ Rm a positive matrix S ∈ Gl (d)+. Then, tr µ = Pi∈Im v2 µi ≤ Pi∈ Im v2 and Pi∈J0 d (m). i Pi(cid:1) if and only if i ki i Ji ∩ Iki , (30) (cid:3) >0 of weights. Fix also S ∈ OPv ⇐⇒ λ(S) ∈ Λ(OPv) ⇐⇒ λ(S) satisfies Eq. (30) . Proof. The set OPv ⊆ Gl (d)+ is saturated by unitary equivalence. Indeed, if V ∈ PRSv and U ∈ U(d), then V · U def= {Vi U}i∈Im ∈ PRSv and U∗SVU = SV·U ∈ OPv . This shows the first equivalence. On the other hand, using Lemma 4.4 and Eq. (17), we can assure that an ordered vector µ ∈ Λ(OPv) if and only if µd > 0 and there exists a sequence of projections P = {Pi}i∈Im ∈ Gr (k , d) such that µ = λ(Sv(P) ) = λ(cid:0) Pi∈Im v2 the second equivalence follows from Lemma 5.7. Corollary 5.9. For every set (m, k, d) of parameters and every vector v ∈ Rm i Pi(cid:1). Hence, (cid:3) >0 of weights, 1. The set Λ(OPv) is convex. 15 2. Its closure Λ(OPv) is compact. 3. A vector µ ∈ Λ(OPv) \ Λ(OPv) ⇐⇒ µd = 0. In other words, Λ(OPv) ∩ Rm >0 = Λ(OPv) . (31) Proof. Denote by M the set of vectors λ ∈ (Rd M is compact and convex. But Proposition 5.8 assures that Λ(OPv) = M ∩ Rd This proves items 2 and 3. Item 1 follows by the fact that also Rd +)↓ which satisfies Eq. (30). It is clear that >0 ⊆ M . (cid:3) >0 is convex. Remark 5.10. With the notations of Corollary 5.9, actually Λ(OPv) = M. This fact is not obvious from the inequalities of Eq. (30), but can be deduced using Lemma 5.7. Indeed, it is clear that if P ∈ Gr (k , d) and Sv(P) /∈ Gl (d)+, then Sv(P) can be approximated by matrices Sv(Q) for sequences Q ∈ Gr (k , d) such that Sv(Q) > 0. Using Lemma 4.4 and Eq. (17), this means that these matrices Sv(Q) ∈ OPv . △ 6 Joint potential of projective RS's Fix the parameters (m, k, d). We consider the set of dual pairs associated to PRSv : DPv = DPv (m, k, d) def= n (V, W) ∈ PRSv × RS : W ∈ D(V)o . We consider on DPv the joint potential: Given (V , W) ∈ DPv , let RSP (V, W) def= RSP (V) + RSP (W) = tr S2 V + tr S2 W ∈ R>0 . (32) We shall describe the structure of the minimizers of the joint potential both from a spectral and a geometrical point of view. We will denote by pv = pv(m, k, d) def= inf { RSP (V, W) : (V, W) ∈ DPv } . (33) Proposition 6.1. For every set (m, k, d) of parameters, the following properties hold: 1. The infimum pv in Eq. (33) is actually a minimum. 2. Let τ =Pi∈Im v2 i ki . For every pair (V, W) ∈ DPv we have that RSP (V, W) ≥ pv ≥ τ 4 + d4 d τ 2 , (34) 3. This lower bound is attained if and only if V is tight (SV = τ τ V = V #. Proof. Given (V, W) ∈ DPv , Corollary 5.6 asserts that RSP (V, V #) ≤ RSP (V, W) and also that equality holds only if W = V #. Thus d Id) and W = d pv = inf V∈PRSv RSP (V, V #) d (35) Consider the strongly convex map F : Rd x ∈ Rd , for >0 . Observe that RSP (V, V #) = F (λ(SV) ) for every V ∈ PRSv . By Corollary 5.9 i=1 x2 i + x−2 i (5) = inf V∈PRSv Xi=1 λi(SV)2 + λi(SV)−2 . >0 → R>0 given by F (x) = Pd 16 >0)↓ , and it becomes also compact under a we know that Λ(OPv) is convex subset of (Rd restriction of the type λd ≥ ε (for any ε > 0). Since a strongly convex function defined in a compact convex set attains its local (and therefore global) minima at a unique point, it follows that there exists a unique λv = λv(m, k, d) ∈ Λ(OPv) such that F ( λv ) = min λ∈Λv(m,k,d) F (λ) = pv . (36) This proves item 1. Moreover, using Lagrange multipliers it is easy to see that the restriction of F to the set (Rd d · 1. Since Λ(OPv) ⊂ (Rd >0 : tr(x) = τ } reaches its minimum in x = τ >0)τ := {x ∈ Rd >0)τ we get that RSP (V, V #) = F (λ(SV) ) ≥ F ( τ d · 1) = τ 4 + d4 d τ 2 for every V ∈ PRSv , d Id , and therefore V # = d and this lower bound is attained if and only if λ(SV) = τ SV = τ Recall that we use in RS the metric dP (V, W) = ( Pi∈Im k2 and the pseudometric dS(V, W) = kSV − SWk for pairs V = {Vi}i∈Im and W = {Wi}i∈Im ∈ RS. Lemma 6.2. If a pair (V, W) ∈ DPv is local dP -minimizer of the joint potential in DPv , then W = V #. d · 1d . Note that in this case (cid:3) 2 )1/2 = kT ∗ V kVi − Wik2 τ V. − T ∗ W Proof. We have shown in Eq. (9) that, since W ∈ D(V), then T ∗ V # + A, for some A ∈ L(K, H) such that A TV = A TV # = 0 ∈ L(H). Recall from Remark 2.6 that the set D(V) is convex. Then the line segment Wt = tW + (1 − t)V # ∈ D(V) satisfies that Wt = T ∗ T ∗ V # + tA for every t ∈ [0, 1]. Then, as in Eq. (28), SWt = SV # + t2 AA∗ and K(t) def= RSP (V , Wt) = RSP (V , V #) + t4 tr (AA∗)2 + 2 t2 tr TV #AA∗T ∗ V # , W = T ∗ for every t ∈ [0, 1]. Observe that K(1) = RSP (V , W). But taking one derivative of K, one gets that if A 6= 0 then K is strictly increasing near t = 1, which contradicts the local dP -minimality for (V , W). Therefore T ∗ (cid:3) Theorem 6.3. For every set (m, k, d) of parameters there exists λv = λv(m, k, d) ∈ (Rd such that the following conditions are equivalent for pair (V, W) ∈ DPv : V # and W = V #. Wt = T ∗ >0)↓ 1. (V, W) is local dS-minimizer of the joint potential in DPv . 2. (V, W) is global minimizer of the joint potential in DPv . 3. It holds that λ(SV) = λv and W = V #. (36). Proof. Take the vector λv defined in Eq. In the proof of Proposition 6.1 we have already seen that a pair (V, W) ∈ DPv is a global minimizer for RSP ⇐⇒ W = V # and λ(SV) = λv . This means that 2 ⇐⇒ 3. Suppose now that (V, W) ∈ DPv is a local dS-minimizer. By Remark 4.9 we know that it is also a local dP -minimizer and by Lemma 6.2 we have that W = V #. In this case, denote λ = λ(SV) and take U ∈ U(d) such that U∗DλU = SV . Consider the segment line h(t) = t λv + (1 − t) λ for every t ∈ [0, 1] . 17 Then h(t) ∈ Λ(OPv) for every t ∈ [0, 1], since Λ(OPv) is a convex set (Corollary 5.9). Consider the continuous curve St = U∗Dh(t)U in OPv and a (not necessarily continuous) curve Vt ∈ PRSv such that S0 = SV , V0 = V and SVt = St for every t ∈ [0, 1]. Nevertheless, since the curve St is continuous, we can assure that the map t 7→ Vt is dS-continuous. Finally, we can consider the map G : [0, 1] → R given by d G(t) = RSP (Vt , V # t ) = tr S2 t + tr S−2 t = hi(t)2 + hi(t)−2 = F (h(t) ) Xi=1 for t ∈ [0, 1], where F is the map defined after Eq. (35). Observe that G(0) = RSP (V , V #) and G(1) = pv , by Eq. (36). Then G has local minima at t = 0 and t = 1. By computing the second derivative of G in terms of the Hessian of F , we deduce that G must be constant, because otherwise it would be strictly convex. From this fact we can see that the map h is also constant, so that λv = λ. Therefore (V, W) = (V , V #) is a global minimizer. (cid:3) Recall that a system V = {Vi}i∈ Im ∈ PRSv is irreducible if CV = {V ∗i Vi : i ∈ Im}′ = C Id . Lemma 6.4. Fix the set (m, k, d) of parameters and the weights v = (vi)i∈Im ∈ Rm >0 . Assume that V ∈ PRSv is irreducible. Then the following conditions are equivalent: 1. The pair (V , V #) is local dP -minimizer of the joint potential in DPv . 2. The pair (V , V #) is global minimizer of the joint potential in DPv . 3. The system V is tight, i.e. SV = τ d Id . Therefore in this case the vector λv of Theorem 6.3 is λv = τ d Proof. Since CV = C Id , we can apply Theorem 4.8. Then the map RSO : PRSv → Gl (d)+ defined in Eq. (12) has a smooth local cross section around SV which sends SV to V. Assume that there exists no σ ∈ R>0 such that SV = σ Id . In this case there exist α, β ∈ σ(SV) such that β > α > 0. Consider the map g : [0, β−α 1d . 2 ] → R>0 given by τ g(t) = (α + t)2 + (α + t)−2 + (β − t)2 + (β − t)−2 . Then g′(0) = 2(α − β) − 2( 1 curve M : [ 0 , ε ] → Gl (d)+ β − 1 τ such that M(0) = SV and α) < 0, which shows that we can construct a continuous tr M(t)2 + tr M(t)−2 < tr S2 V + tr S−2 V = RSP (V , V #) for every t ∈ (0 , ε ] . Hence, using the continuous local cross section mentioned before, we can construct a dP - continuous curve M : [0 , δ ] → PRSv such that RSO ◦ M = M, M(0) = V and RSP (M(t) , M(t)#) = tr M(t)2 + tr M(t)−2 < RSP (V , V #) for t ∈ (0 , δ ] . This shows that (V , V #) is not a local dP -minimizer of the joint potential in DPv . We have proved that 1 =⇒ 3. Note that 3 =⇒ 2 follows from (34) and 2 =⇒ 1 is trivial. (cid:3) Remark 6.5. It is easy to see that, if the parameters (m , k , d) allow the existence of at least one irreducible projective RS, then the set of irreducible systems becomes open and dense in PRSv (m , k , d). Nevertheless, it is not usual that the minimizers are irreducible, even if they are tight (see Remark 6.7 and Examples 7.1 and 7.2). On the other hand, if the system V ∈ PRSv is reducible, there exists a system Q = {Qj}j∈Ip of minimal projections of the unital C∗-algebra CV (with p > 1). This means that 18 j = Q∗j = Qj . • Each Qj ∈ CV , and Q2 • Q is a system of projections: Qj Qk = 0 if j 6= k and Pj∈Ip Qj = IH . • Minimality: The algebra CV has no proper sub projection of any Qj . By compressing the system V to each subspace Hj = R(Qj) in the obvious way, it can be shown that every V ∈ PRSv is an "orthogonal sum" of irreducible subsystems. Another system of projections associated with V are the spectral projections of SV : σ(SV) = {σ1 , . . . , σr}, we denote these projections by If Pσj = Pσj (SV ) def= Pker (S−σj Id) ∈ Md(C)+ , for j ∈ Ir . Recall that SV Pσj = σj Pσj and Pr Theorem 6.6. Fix v = (vi)i∈Im ∈ Rm joint potential in DPv with V = {Vi}i∈Im . Then j=1 Pσj = Id , so that SV =Pr >0 . Let (V, W) ∈ DPv be a dP -local minimizer of the j=1 σj Pσj . △ 1. The RS operator SV ∈ CV = {V ∗i Vi : i ∈ Im}′. 2. If σ(SV) = {σ1, . . . , σr}, then also Pσi = Pσi(SV) ∈ CV for every i ∈ Ir . i W ∗i Proof. Recall that V ∈ PRSv ⊆ RS and hence 0 /∈ σ(SV ). On the other hand, we have already seen in Lemma 6.2 that W must be V #. Let Q = {Qj}j∈Ip be a system of minimal projections of the unital C∗-algebra CV , as in Remark 6.5. Fix j ∈ Ip and denote by Sj = R(Qj). For every i ∈ Im put Ti = Vi(Sj) ⊆ Ki , ti = dim Ti and Wi = ViQj ∈ L(Hj , Ti) . Since Qj ∈ CV then each matrix v−1 is an isometry, so that the compression of V given by W = {Wi}i∈ Im ∈ PRSv (m , t , sj), where t = (t1 , . . . , tm) and sj = dim Sj . Recall that SV commutes with Qj . This implies that W # is the same type of compression to RSv (m , t , sj) of the system V #. A straightforward computation shows that the pair (W, W #) ∈ DPv (m , t , sj) is still a dP - local minimizer of the joint potential in DPv (m , t , sj). Indeed, the key argument is that one can "complete" other systems in PRSv (m , t , sj) near W (and acting in Sj) with the fixed orthogonal complement {Vi(Id − Qj)}i∈Im , getting systems in PRSv (m , k , d) near V. It is easy to see that all the computations involved in the joint potential work independently on each orthogonal subsystem. This shows the minimality of (W, W #). Observe that W ∗i Wi = QjV ∗i ViQj = V ∗i ViQj for every i ∈ Im . Therefore, the minimality of Qj in CV shows that the system W satisfies that CW = C ISj . Hence, we can apply Lemma 6.4 on Sj , and get that SW = αj ISj for some αj > 0. But when we return to L(H), we get that SV Qj =Pi∈Im V ∗i ViQj =Pi∈Im W ∗i Wi = SW = αj Qj . In particular, αj ∈ σ(SV). We have proved that for every j ∈ Ip there exists αj ∈ σ(SV) such that SV Qj = αj Qj and hence each projector Qj ≤ Pαj = Pαj (SV) . Using that Pj∈Ip Qj = Id we see that each where Jk = {j ∈ Ip : αj = σk} . (37) Pσk = Xj∈Jk Qj ∈ CV , Therefore also SV =Pk∈Ir σk Pσk ∈ CV . 19 (cid:3) Remark 6.7. Theorem 6.6 assures that if (V , V #) is a dP -local minimizer of the joint potential in DPv , then V is an orthogonal sum of tight systems in the following sense: If σ(SV) = {σ1, . . . , σr}, and we denote Hj = R(Pσj ) = ker (S − σj Id) for every j ∈ Ir , then H =Lj∈Ir Hj . By Theorem 6.6 each Pσj ∈ CV . Then, putting dj = dim Hj , and kj = (k1 , j , . . . , km , j) , Ki , j = Vi(Hj) ⊆ Ki , ki , j = dim Ki , j for every i ∈ Im and j ∈ Ir , we can define the the tight compression of V to each Hj : V j = {Vi Pσj }i∈Im ∈ PRSv (m , kj , dj) for j ∈ Ir . Indeed, since Pσj ∈ CV then V j is projective. Also SV j = SV Pσj = σj Pσj , which means that V j is σj - tight. Observe that the decomposition of each V j into irreducible tight systems (as in Remark 6.5) follows from the orthogonal decomposition of Hj given in Eq. (37). In particular, every V ∈ PRSv such that λ(SV ) = λv (the unique vector of Theorem 6.3) must have this structure, because in this case (V , V #) is a dS (hence also dP ) local minimizer of the joint potential in DPv . Observe that the structure of all global minimizers V is almost the same: Since λ(SV) = λv , the number r of tight components, the sizes dj and the tight constants σj for each space Hj coincide for every such minimizer V. Note that, if such a V is not tight, then it can not be irreducible. On the other hand, its dual V # can only be projective if Vi Pσj = 0 or Vi for every i ∈ Im and j ∈ Ir . △ 7 Examples and conclusions The following two examples are about irreducible systems. Example 7.1. Let d = k1 + k2 and k = (k1 , k2). Assume that k1 > k2 . We shall see that, in this case, there is no irreducible (Riesz) systems in PRS(2 , k , d). Observe that the situation is the same whatever the weights (v1 , v2) are. Indeed, if V = (V1 , V2) ∈ PRS1(2 , k , d), let Si = R(V ∗i ) and Pi = PSi = V ∗i Vi for i = 1, 2. Then Cd = S1 ⊕ S2 (not necessarily orthogonal). Observe that dim S1 = dim S⊥2 = k1 and 2 k1 > d. Hence T = S1 ∩ S⊥2 6= {0}. Since P = PT ≤ P1 and P ≤ Id − P2 , then P ∈ CV and 0 6= P 6= Id . Therefore CV 6= C Id . In particular, if the decomposition Cd = S1 ⊕ S2 is orthogonal, then SV = P1 + P2 = Id . So, △ in this case V is tight and reducible. Example 7.2. If m ≥ d and k = 1m , then PRS(m , k , d) is the set of m-vector frames for the space Cd. In this case F = {fi}i∈ Im ∈ PRS is reducible ⇐⇒ there exists J ⊆ Im such that ∅ 6= J 6= Im and the subspaces span{fi : i ∈ J} and span{fj : j /∈ J} are orthogonal. Indeed, if A = A∗, then A ∈ CF ⇐⇒ every fi is an eigenvector of A. But different eigenvalues of A must have orthogonal subspaces of eigenvectors. Observe that in this case the set of irreducible systems is an open and dense subset of PRSv , since it is the intersection of 2m − 2 open dense sets (one for each fixed nontrivial J ⊆ Im). △ 7.3. Minimizers and majorization: Theorem 6.3 states that there exists a vector λv = >0)↓ such that a system V ∈ PRSv (m, k, d) satisfies that (V, V #) is a global λv(m, k, d) ∈ (Rd minimizer of the joint potential in DPv if and only if λ(SV) = λv . This vector is found as the unique minimizer of the map F (λ) =Pd 20 i=1 λ2 i + λ−2 i on the convex set Λ(OPv) . In all the examples where λv could be explicitly computed, it satisfied a stronger condition, in terms of majorization (see [5, Cap. II] for definitions and basic properties). We shall see that in these examples there is a vector λ ∈ Λ(OPv ) such that λ ≺ λ(SV) for every V ∈ PRSv (the symbol ≺ means majorization) . (38) Observe that such a vector λ ∈ Λ(OPv ) must be the unique minimizer for F on Λ(OPv ), since the map F is permutation invariant and convex. Hence λ = λv . Moreover, those cases where λv satisfies Eq. (38) have some interesting properties regarding the structure of minimizers of the joint potential. For example, that λ tv(m, k, d) = t2λv(m, k, d) for t > 0, △ a fact that is not evident at all from the properties of these vectors. Conjecture 7.4. For every set of parameters (m, k, d) and v ∈ Rd of Theorem 6.3 satisfies the majorization minimality of Eq. (38) on Λ(OPv) . Example 7.5. Given v = v↓ ∈ Rm >0 and d ≤ m, the d-irregularity of v is the index >0 , the vector λv(m, k, d) △ r = rd(v) def= max (cid:8) j ∈ Id−1 : (d − j) v2 j > Pm i=j+1 v2 i (cid:9) , or r = 0 if this set is empty. In [24, Prop. 2.3] (see also [2, Prop. 4.5]) it is shown that for any set of parameters (m , 1m , d) and every v = v↓ ∈ Rm >0 , there is c ∈ R such that λv(m, d) def= (v2 1 , . . . , v2 r , c 1d−r) ∈ Λ(OPv(m, 1m , d) ) and it satisfies Eq. (38). Therefore λv(m, d) = λv(m, 1m , d) by 7.3. Thus, in the case of △ vector frames, Conjecture 7.4 is known to be true. In the following examples we shall compute explicitly the the vector λv and the global minimizers of the joint potential in PRSv . Since we shall use Eq. (38) as our main tool (showing Conjecture 7.4 in these cases), we need a technical result about majorization, similar to [23, Lemma 2.2]. Recall that the symbol ≺w means weak majorization. Lemma 7.6. Let α , γ ∈ Rn, β ∈ Rm and b ∈ R such that b ≤ mink∈In γk . Then, if γ ≺w α =⇒ (γ , b 1m) ≺w (α , β) . tr (γ , b 1m) ≤ tr (α , β) and Observe that we are not assuming that (α , β) = (α , β)↓. Proof. Let h = tr β and ρ = h m 1m . Then it is easy to see that Pi∈Ik (γ↓ , b 1m)i ≤ Pi∈Ik (α↓ , ρ)i ≤ Pi∈Ik Since (γ↓ , b 1m) = (γ , b 1m)↓, we can conclude that (γ , b 1m) ≺w (α , β). (α↓ , β↓)i for every k ∈ In+m . (cid:3) Example 7.7. Assume that tr k = d. Then the elements of PRSv(m , k , d) are Riesz systems. Assume that the weights are ordered in such a way that v = v↓. We shall see that the vector λ = (v2 1k1 , . . . , v2 1km) ≺ λ(SV) for every V ∈ PRSv(m , k , d). Hence λ 1 m satisfies Eq. (38), and λv = λ by 7.3. Indeed, given V = {Vi}i∈ Im ∈ PRSv , consider the projections Pi = v−2 Si = R(Pi) for every i ∈ Im . Then SV = Pi∈Im v2 , P = PS i Pi and Cd = Li∈Im Si where the direct and Q = Id − P = PS ⊥ . sum is not necessarily orthogonal. Let i V ∗i Vi and denote by Si ⊆ Cd S =Li∈Im−1 21 Consider the restriction A = v2 i Pi ∈ L(S)+. It is well known that the pinching matrix m−1 Pi=1 M = P SV P + Q SV Q =(cid:20) A + v2 m P P2P 0 0 m QPmQ (cid:21) S S⊥ v2 γ = (v2 1 1k1 , . . . , v2 satisfies that λ(M) ≺ λ(SV). Using an inductive argument on m (the case m = 1 is trivial), for the Riesz system V0 = {Vi(cid:12)(cid:12)S }i∈Im−1 (for S) such that SV0 = A, we can assure that in Rd−km . 1km−1) ≺ λ(A) ≺w λ(cid:0) A + v2 Since vm ≤ vm−1 , Lemma 7.6 assures that λ = (γ , v2 β = λ(v2 Recall a system V ∈ PRSv is a minimizer if and only if λ(SV) = λv = λ. Now, it is easy △ to see that λ(SV) = λv if and only if the projections Pi are mutually orthogonal. Example 7.8. Assume that the parameters (m , k , d) satisfy that m QPmQ) ∈ Rkm. Hence, we have proved that λ ≺ λ(SV). m P P2P (cid:1) = α 1km) ≺ (α , β) = λ(M), where m−1 m m = 2 and tr k = k1 + k2 > d , but k1 6= d 6= k2 . Fix v = (v1 , v2) with v1 ≥ v2. For the space PRSv (2 , k , d) the vector λv of Theorem 6.3 and all the global minimizers of the joint potential can be computed: Denote by r0 = k1 + k2 − d , r1 = k1 − r0 and r2 = k2 − r0 . We shall see that the vector µ = ( (v2 1r2) satisfies Eq. (38), so that λv(2 , k , d) = µ by 7.3. Moreover, the minimizers are those systems V = (V1 , V2) ∈ PRSv such that the two projections Pi = v−2 Indeed, if Si = R(Pi) = R(V ∗i ) for i = 1, 2, then M0 = S1 ∩ S2 has dim M0 = r0 . Also Mi = Si ⊖ M0 have dim Mi = ri for i = 1, 2. Hence Cd = M0 ⊥ (M1 ⊕ M2) and i V ∗i Vi (for i = 1, 2) commute. 2) 1r0 , v2 1 1r1 , v2 2 1 + v2 SV = v2 1 P1 + v2 1 P2 = (v2 1 + v2 2) PM0 + v2 1 PM1 + v2 1 PM2 . Note that M1 ⊥ M2 ⇐⇒ P1 P2 = P2 P1 = PM0 . In this case λ(SV) = µ. Otherwise, 2) IM0 and SV (M1 ⊕ M2) = M1 ⊕ M2 . Hence, if we denote by 2 and ∈ Gl(M1 ⊕ M2)+ , then kT ksp ≤ v2 1 + v2 = (v2 1 + v2 = (v2 still SV(cid:12)(cid:12)M0 T = SV(cid:12)(cid:12)M1⊕M2 SV =(cid:20) (v2 1 PM1 + v2 2) Ir0 1 PM2)(cid:12)(cid:12)M1⊕M2 with 0 T (cid:21) M0 M⊥0 1 + v2 0 λ(SV) = ( (v2 1 + v2 2) 1r0 , λ(T ) ) ∈ (Rd >0)↓ . Using Example 7.7 for the space M1 ⊕ M2 , we can deduce that (v2 1 Therefore also µ = ( (v2 1r2) ≺ ( (v2 2) 1r0 , v2 1 1r1 , v2 2 1 + v2 1 + v2 1r1 , v2 2 1r2) ≺ λ(T ). △ 2) 1r0 , λ(T ) ) = λ(SV). Example 7.9. Let m = 3, d = 4, k = (3 , 2 , 2) and v = 13 . Denote by E = {ei : i ∈ I4} the canonical basis of C4. Then λ1(3 , k , 4) = (2 , 2 , 3 2 ) and a minimizer is given by any system V = {Vi}i∈ I3 ∈ PRS1 such that the subspaces Si = R(V ∗i ) for i ∈ I3 are 2 , 3 S1 = span{e1 , e2 , e3} , S2 = span(cid:8) e1 , w2(cid:9) √3 e4 √3 e4 and S3 = span(cid:8) e2 , w3(cid:9) , 2 + where w2 = −e3 . The fact that λ(SV ) = (2 , 2 , 3 2 ) for such a system V is a direct computation. On the other hand, if W = {Wi}i∈ I3 ∈ PRS1(3 , k , 4) , then there exist unit vectors x2 ∈ R(W ∗1 ) ∩ R(W ∗2 ) and x3 ∈ R(W ∗1 ) ∩ R(W ∗3 ). and w3 = −e3 2 , 3 2 − 2 2 22 Denote by T = span{x2 , x3}. If dim T = 1 then λ1(SW ) ≥ hSW x2 , x2i = 3 and λ2(SW ) ≥ 1. If dim T = 2, using that T ⊆ R(W ∗1 ) and xi ∈ R(W ∗i ) for i = 2, 3, we get tr(cid:0)PT W ∗i Wi PT(cid:1) ≥ tr PT + tr Pspan{x2} + tr Pspan{x3} = 4 . λ1(SW) + λ2(SW) ≥ Pi∈ I3 In any case, we have shown that (2 , 2) ≺w α = (λ1(SW) , λ2(SW ) ). Therefore, using Lemma 7.6 we get that (2 , 2 , 3 2 ) ≺ λ(SW) . Now, apply 7.3. The minimizers V ∈ PRSv such that λ(SV) = (2 , 2 , 3 2 ) have some interestig properties. For example they are the sum of two tight systems, V # is not projective, and the involved projections do not commute. More precisely, the cosine of the Friedrich angles of their images are c(Si , Sj) = 1 △ 2 , 3 2 , 3 2 for every i 6= j. References [1] E. Andruchow, G. Corach, Differential geometry of partial isometries and partial unitaries, Illinois J. Math. 48 (2004), 97-120. [2] J. Antezana, P. Massey, M. Ruiz and D. Stojanoff, The Schur-Horn theorem for operators and frames with prescribed norms and frame operator, Illinois J. Math., 51 (2007), 537-560. [3] M. S. Asgari, New characterizations of fusion frames (frames of subspaces), Proc. Indian Acad. Sci. (Math. Sci.) Vol. 119, (2009), 369-382. [4] J.J. Benedetto, M. Fickus, Finite normalized tight frames, Adv. Comput. Math. 18, No. 2-4 (2003), 357-385 . [5] R. Bhatia, Matrix Analysis, Berlin-Heildelberg-New York, Springer 1997. [6] B.G. Bodmann, Optimal linear transmission by loss-insensitive packet encoding, Appl. Comput. Harmon. Anal. 22, no. 3, (2007) 274-285. [7] B.G. Bodmann, D.W. Kribs, V.I. Paulsen Decoherence-Insensitive Quantum Communication by Optimal C ∗- Encoding, IEEE Transactions on Information Theory 53 (2007) 4738-4749. [8] B.G. Bodmann, V.I. Paulsen, Frames, graphs and erasures, Linear Algebra Appl. 404 (2005) 118-146. [9] P.G. Casazza, M. Fickus, Minimizing fusion frame potential, Acta Appl. Math. 107, No. 1-3 (2009), 7-24. [10] P.G. Casazza, M. Fickus, D.G. Mixon, Y. Wang, Z. Zhou, Constructing tight fusion frames, Appl. Comput. Harmon. Anal. (2010), doi:10.1016/j.acha.2010.05.002. [11] P.G. Casazza, M. Fickus, J. Kovacevic, M. T. Leon,J. C. Tremain, A physical interpretation of tight frames, Harmonic analysis and applications, 51-76, Appl. Numer. Harmon. Anal., Birkhauser Boston, MA, 2006. [12] P.G. Casazza, G. Kutyniok, Frames of subspaces, Contemp. Math. 345 (2004), 87-113. [13] P.G. Casazza, G. Kutyniok, Robustness of Fusion Frames under Erasures of Subspaces and of Local Frame Vectors, Contemp. Math. 464 (2008), 149-160 . [14] P.G. Casazza, G. Kutyniok, S. Li, Fusion Frames and Distributed Processing, Appl. Comput. Harmon. Anal. 25, No. 1 (2008), 114-132 . [15] P.G. Casazza, G. Kutyniok, S. Li, C.J. Rozell, Modeling Sensor Networks with Fusion Frames, Proc. SPIE 6701 (2007), 67011M-1-67011M-11. [16] G. Corach, M. Pacheco, D. Stojanoff, Geometry of epimorphisms and frames, Proc. Amer. Math. Soc. 132 (2004), 2039-2049. [17] G. Corach, H. Porta, L. Recht, The geometry of spaces of projections in C ∗-algebras, Adv. Math. 101 (1993), 59-77. [18] Fulton, W., Young tableaux. With applications to representation theory and geometry, London Mathematical Society Student Texts, 35. Cambridge University Press, Cambridge (1997). [19] P. Gavruta, On the duality of fusion frames, J. Math. Anal. Appl. 333 (2007), no. 2, 871-879. [20] R.B. Holmes, V.I. Paulsen, Optimal frames for erasures, Linear Algebra Appl. 377 (2004) 31-51. [21] A. Klyachko, Stable bundles, representation theory and Hermitian operators, Selecta Math. (N.S.) 4 (1998), no. 3, 419-445. [22] G. Kutyniok, A. Pezeshki, R. Calderbank, T. Liu, Robust Dimension Reduction, Fusion Frames, and Grassman- nian Packings, Appl. Comput. Harmon. Anal. 26, (2009) 64-76. [23] P. Massey, Optimal reconstruction systems for erasures and for the q-potential, Linear Algebra Appl. 431 (2009) 1302-1316. [24] P. Massey and M. Ruiz, Minimization of convex functionals over frame operators, Adv Comput Math 32 (2010), 131-153. [25] P. Massey, M. Ruiz and D. Stojanoff, The structure of minimizers of the frame potential on fusion frames, J Fourier Anal Appl, doi:10.1007/s00041-009-9098-5. [26] C.J. Rozell, D.H. Johnson, Analyzing the robustness of redundant population codes in sensory and feature ex- traction systems, Neurocomputing 69 (2006), 1215-1218 . [27] M. Ruiz, D. Stojanoff, Some properties of frames of subspaces obtained by operator theory methods, J. Math. Anal. Appl. 343, No. 1 (2008), 366-378. Pedro Massey, Mariano Ruiz and Demetrio Stojanoff Depto. de Matem´atica, FCE-UNLP, La Plata, Argentina and IAM-CONICET e-mail: [email protected] , [email protected], [email protected] 23
1101.2645
1
1101
2011-01-13T19:53:41
Classical limit of the d-bar operators on quantum domains
[ "math.FA", "math-ph", "math-ph" ]
We study one parameter families $D_t$, $0<t<1$ of non-commutative analogs of the d-bar operator $D_0 = \frac{\d}{\d\bar{z}}$ on disks and annuli in complex plane and show that, under suitable conditions, they converge in the classical limit to their commutative counterpart. More precisely, we endow the corresponding families of Hilbert spaces with the structures of continuous fields over the interval $[0,1)$ and we show that the inverses of the operators $D_t$ subject to APS boundary conditions form morphisms of those continuous fields of Hilbert spaces.
math.FA
math
CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS SLAWOMIR KLIMEK AND MATT MCBRIDE Abstract. We study one parameter families Dt, 0 < t < 1 of non-commutative analogs of the d-bar operator D0 = ∂ ∂z on disks and annuli in complex plane and show that, under suitable conditions, they converge in the classical limit to their commutative counterpart. More precisely, we endow the corresponding families of Hilbert spaces with the structures of continuous fields over the interval [0, 1) and we show that the inverses of the operators Dt subject to APS boundary conditions form morphisms of those continuous fields of Hilbert spaces. 1 1 0 2 n a J 3 1 ] . A F h t a m [ 1 v 5 4 6 2 . 1 0 1 1 : v i X r a 1. Introduction According to the broadest and the most flexible definition, a quantum space is simply a noncommutative algebra. Noncommutative geometry [4] studies what could be considered "geometric properties" of such quantum spaces. One of the most basic examples of a quantum space is the quantum unit disk C(Dt) of [8]. It is defined as the universal unital C ∗-algebra with the generators zt and zt which are adjoint to each other, and satisfy the following commutation relation: [zt, zt] = t(I − ztzt)(I − ztzt), for a continuous parameter 0 < t < 1. It was proved in [8] that C(Dt) has a more concrete representation as the C ∗-algebra generated by the unilateral weighted shift with the weights given by the formula: wt(k) =s (k + 1)t 1 + (k + 1)t . (1.1) In fact, as a C ∗-algebra, C(Dt) is isomorphic to the Toeplitz algebra. Moreover the family C(Dt) is a deformation, and even deformation - quantization of the algebra of continuous functions on the disk C(D) obtained in the limit as t → 0, called the classical limit. The quantum unit disk is one of the simplest examples of a quantum manifold with boundary. It is also an example of a quantum complex domain, with zt playing the role of a quantum complex coordinate. Additionally, biholomorphisms of the unit disk naturally lift to automorphisms of C(Dt), see [8]. In view of this complex analytic interpretation of the quantum unit disk, there is a natural need to define analogs of complex partial derivatives as some kind of unbounded operators on C(Dt) and its various Hilbert space completions. Such constructions have been described in several places in the literature, see for example [2], [3], [7], [9], [10],[11]. In this paper we are primarily concerned with one such choice, the so-called balanced d and d-bar operators of [9] which we describe below. Date: November 10, 2018. 1 2 SLAWOMIR KLIMEK AND MATT MCBRIDE One notices that St := [zt, zt] is an invertible trace class operator (with an unbounded inverse) and defines and Dta = S−1/2 t [a, zt]S−1/2 t Dta = S−1/2 t [zt, a]S−1/2 t , for appropriate a ∈ C(Dt). These two operators have the following easily seen properties Dt(1) = 0, Dt(zt) = 0, Dt(zt) = 1 Dt(1) = 0, Dt(zt) = 1, Dt(zt) = 0 which makes them plausible candidates for quantum complex partial derivatives. To make an even better case of their suitability, one would like to know that in some kind of interpretation of the limit as t → 0, they indeed become the classical partial derivatives. This problem was posed at the end of [7] and it is the subject of the present paper. In fact we consider here a broader classical limit problem by studying quite general families of unilateral weights wt(k), and not just those given by (1.1). Like in [9] such unilateral shifts are still considered coordinates of quantum disks. Additionally we also consider bilateral shifts and the C ∗-algebras they generate. They are quantum analogs of annuli and can be analyzed very similarly to the quantum disks. We start with giving a concrete meaning to the classical limit t → 0, which involves two important steps. The first step is to consider certain bounded functions of the quantum d and d-bar operators to properly manage their unboundedness. In this paper we choose to work with the inverses of the operators Dt subject to APS boundary conditions [1] since they are easy to describe and we can use the results of [3], [9]. The second step of our approach to the classical limit is the choice of framework for studying limits of objects living in different spaces. Such a natural framework is provided by the language of continuous fields, in our case of continuous fields of Hilbert spaces, see [5]. Following [3] and [9] we define, using operators St, weighted Hilbert space completions Ht, 0 < t < 1, of the above quantum domains, while H0 is the classical L2 space. We then equip that family of Hilbert spaces with a natural structure of continuous field, namely the structure generated by the polynomials in complex quantum and classical coordinates. In this setup the study of the classical limit becomes a question of continuity, a property embedded in the definition of the continuous field. Consequently, inverses of the operators Dt subject to APS boundary conditions, are considered as morphisms of the continuous fields of Hilbert spaces. The main result of this paper is that in such a sense the limit of Dt is indeed ∂ ∂z . The paper is organized as follows. In section 2 we review the definitions and properties of continuous fields of Hilbert spaces and their morphisms. In Section 3 we describe the constructions of the quantum disk, the quantum annulus, Hilbert spaces of L2 "functions" on those quantum spaces, d-bar operators and their inverses subject to APS conditions. We state the conditions on weights wt(k) and provide example of such weights. We construct the generating subspace Λ needed for the construction of the continuous field of Hilbert spaces. The main results of this paper are also formulated at the end of that section. Finally, section 4 contains the proofs of the results. CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 3 2. Continuous Fields of Hilbert Spaces In this section we review some aspects of the theory of continuous fields of Hilbert spaces. The main reference here is Dixmier's book [5]. Definition 2.1. A continuous field of Hilbert spaces is a triple, denoted (Ω,H, Γ), where Ω is a locally compact topological space, H = {H(ω) : ω ∈ Ω} is a family of Hilbert spaces, and Γ is a linear subspace ofQω∈Ω H(ω), such that the following conditions hold (3) let x ∈Qω∈Ω H(ω); if for every ω0 ∈ Ω and every ε > 0, there exists x′ ∈ Γ such that (1) for every ω ∈ Ω, the set of x(ω), x ∈ Γ, is dense in H(ω), (2) for every x ∈ Γ, the function ω 7→ kx(ω)k is continuous, kx(ω) − x′(ω)k ≤ ε for every ω in some neighborhood (depending on ε) of ω0, then x ∈ Γ. The point of this definition is to describe a continuous arrangement of a family of different Hilbert spaces. If they are all the same, then the space Γ of continuous functions on Ω with values in that Hilbert space clearly satisfies all the conditions. Below we will use the following terminology. We say that a section x ∈ Qω∈Ω H(ω) is approximable by Γ at ω0 if for every ε > 0, there exists an x′ ∈ Γ and a neighborhood of ω0 such that kx(ω)−x′(ω)k ≤ ε for every ω in that neighborhood. In this terminology condition 3 of the above definition says that if a section x is approximable by Γ at every ω ∈ Ω, then x ∈ Γ. The above definition is a little cumbersome to work with, namely, trying to describe Γ in full detail is usually very difficult since the third condition isn't easy to verify. The following proposition, proved in [5], makes it easier to construct continuous fields. Proposition 2.2. Let Ω be a locally compact topological space, and let H = {H(ω) : ω ∈ Ω} (1) for every ω ∈ Ω, the set of x(ω), x ∈ Λ, is dense in H(ω), (2) for every x ∈ Λ, the function ω 7→ kx(ω)k is continuous, be a family of Hilbert spaces. If Λ is a linear subspace ofQω∈Ω H(ω) such that then Λ extends uniquely to a linear subspace Γ ⊂Qω∈Ω H(w) such that (Ω,H, Γ) is a con- Here one says that if a linear subspace Λ ofQω∈Ω H(ω) satisfies the two conditions above structed as a local completion of Λ, i.e. Γ consists of all those sections x ∈Qω∈Ω H(ω) which then Λ generates the continuous field of Hilbert spaces (Ω,H, Γ). In fact, Γ is simply con- are approximable by Λ at every ω ∈ Ω. following definition. Definition 2.3. Let (Ω,H, Γ) be a continuous field of Hilbert spaces and let T (ω) : H(ω) → H(ω) be a collection of operators acting on the Hilbert spaces H(ω). Define T =Qω∈Ω T (ω) : Qω∈Ω H(ω) →Qω∈Ω H(ω). We say that {T (ω)} is a continuous family of bounded operators Next we consider morphisms of continuous fields of Hilbert spaces. For this we have the tinuous field of Hilbert spaces. in (Ω,H, Γ) if (1) T (ω) is bounded for each ω, (2) sup ω∈ΩkT (ω)k < ∞, (3) T maps Γ into Γ. 4 SLAWOMIR KLIMEK AND MATT MCBRIDE The proposition below contains an alternative description of the third condition above, so it is more manageable. Proposition 2.4. With the notation of the above definition, the following three conditions are equivalent: (1) T maps Γ into Γ, (2) T maps Λ into Γ, (3) for every x ∈ Λ and for every ω ∈ Ω, T (ω)x(ω) is approximable by Λ at ω. Proof. The items above are arranged from stronger to weaker. The proof that (2) is equiva- lent to (3) is a simple consequence of the way that Γ is obtained from Λ described in the para- graph following Proposition 2.2. Condition (2) implies condition (1) because sup ω∈Ω kT (ω)k < ∞ and so, if x(ω) and y(ω) are locally close to each other, so are T (ω)x(ω) and T (ω)y(ω). (cid:3) 3. D-Bar operators on non-commutative domains In this section we review a variety of constructions needed to formulate and prove the results of this paper. Those constructions include the definitions of the quantum disk, the quantum annulus, Hilbert spaces of L2 "functions" on those quantum spaces, and d-bar operators that were discussed in [9]. Other items discussed in this section are APS boundary conditions, inverses of d-bar operators subject to APS conditions, conditions on weights, and a construction of the generating subspace Λ of the continuous field of Hilbert spaces. The main results are stated at the end of this section. In the following formulas we let S be either N or Z. Let t ∈ (0, 1) be a parameter. Let {ek}, k ∈ S be the canonical basis for ℓ2(S). Given a t-dependent, bounded sequence of numbers {wt(k)}, called weights, the weighted shift Uwt is an operator in ℓ2(S) defined by: Uwtek = wt(k)ek+1. The usual shift operator U satisfies Uek = ek+1. If S = N then the shift Uwt is called a unilateral shift and it will be used to define a quantum disk. If S = Z then the shift Uwt is called a bilateral shift and we will use it to define a quantum annulus (also called a quantum cylinder). For the choice of weights 1.1 the shifts Uwt are the quantum complex coordinates zt of the introduction. We require the following condition on the one-parameter family of weights wt(k). Condition 1. The weights wt(k) form a positive, bounded, strictly increasing sequence in k such that the limits w± := lim k→±∞ wt(k) exist, are positive, and independent of t. Consider the commutator St = U ∗ wtUwt − UwtU ∗ wt. It is a diagonal operator Stek = St(k) ek, where Moreover St is a trace class operator with easily computable trace: St(k) := wt(k)2 − wt(k − 1)2. tr(St) =Xk∈S St(k) = (w+)2 − (w−)2 (3.1) in the bilateral case, and tr(St) = (w+)2 in the unilateral case. Additionally St is invertible with unbounded inverse. CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 5 We assume further conditions on the wt(k)'s and the St(k)'s. Those conditions were simply extracted from the proofs in the next section to make the estimates work. They are possibly not optimal, but they cover our motivating example described in the introduction. Condition 2. The function t 7→ wt(k) is continuous for every k, and for every ε > 0, wt(k) converges to w± as k → ±∞ uniformly on the interval t ≥ ε. Condition 3. If h1(t) := sup k∈S St(k) then h1(t) → 0 as t → 0+. k∈S (cid:12)(cid:12)(cid:12)1 − St(k+1) t∈[0,1)(cid:12)(cid:12)(cid:12)1 − wt(k−1) St(k) (cid:12)(cid:12)(cid:12) exists, and is a bounded function wt(k) (cid:12)(cid:12)(cid:12) exists for every k, and h3(k) → 0 Condition 4. The supremum h2(t) := sup of t, and h2(t) → 0 as t → 0+. Condition 5. The supremum h3(k) := sup as k → ±∞. Notice that the last condition implies that wt(k) ≤ const wt(k − 1) (3.2) where the const above does not depend on t and k. This observation will be used in the proofs in the next section. Before moving on, we verify that the weight sequence 1.1 in the example in the introduction satisfies all of the conditions. First we compute: St(k) = t (1 + kt)(1 + (k + 1)t) . Conditions 1 and 2 are all easily seen to be true with w+ = 1. For conditions 3, 4, and 5 simple computations give h1(t) = t/(1 + t), h2(t) = 2t/(1 + 2t) = O(t), and h3(k) = (k + 1 + √k2 + k)−1 = O(1/k), and so, by inspection, these weights meet all the required conditions. Examples of bilateral shifts satisfying the above conditions are: w2 t (k) = α + β tk . 1 + tk For this example h1(t) = βt/(1 + t) and h2(t) = O(t), h3(k) = O(1/k), w2 w2 − = α − β. Another similar example is w2 Next we proceed to the definition of the continuous field of Hilbert spaces over the interval I = [0, 1). Let C ∗(Uwt) be the C ∗-algebra generated by Uwt. Then, in the unilateral case, the algebra C ∗(Uwt) is called the non-commutative disk. There is a canonical map: t (k) = α + β tan−1(tk). + = α + β, C ∗(Uwt) called the restriction to the boundary map. r−→ C(S1) In the bilateral case the algebra C ∗(Uwt) is called the non-commutative cylinder, and we also have restriction to the boundary maps: We then define the Hilbert space Ht, for t > 0, to be the completion of C ∗(Uwt) with respect to the inner product given by: C ∗(Uwt) r=r+⊕r−−→ C(S1) ⊕ C(S1). 6 SLAWOMIR KLIMEK AND MATT MCBRIDE t = tr(cid:16)S1/2 kak2 t aS1/2 t a∗(cid:17) . For t = 0 we set H0 = L2(Dw+) in the unilateral/disk case and H0 = L2(Aw−,w+) in the bilateral/annulus case where Dw+ := {z ∈ C : z ≤ w+} is the disk of radius w+, and Aw−,w+ := {z ∈ C : w− ≤ z ≤ w+} is the annulus with inner radius w− and outer radius w+. In what follows we usually skip the norm subscript as it will be clear from other terms subscript which Hilbert space norm or operator norm is used. Also notice that setting w− = 0 reduces most annulus formulas below to the disk case. So far we have the space of parameters I, and for every t ∈ I we defined the Hilbert space Ht. We also have distinguished elements of Ht, namely quantum complex coordinates Uwt. We use them to generate the continuous field of Hilbert spaces. More precisely we define the generating linear space Λ ⊂Qt∈I Ht to consists of all those x = {x(t)} such that there exists N > 0, (depending on x), and for every n ≤ N there are functions fn, gn ∈ C([(w−)2, (w+)2]), such that for t > 0: gn(cid:0)wt(k)2(cid:1) (U ∗)n , gn(r2)e−inϕ. (3.3) (3.4) and for t = 0: xt(k) =Xn≤N x0(r, ϕ) =Xn≤N U nfn(cid:0)wt(k)2(cid:1) +Xn≤N fn(r2)einϕ +Xn≤N Now we proceed to the definitions of the quantum d-bar operators. The operator Dt in Ht is given by the following expression: Dta = S−1/2 t [a, Uwt] S−1/2 t for t > 0, and for t = 0, D0 = ∂/∂z. Of course we need to specify the domain of Dt since it is an unbounded operator. For reasons indicated in the introduction, in this paper we consider the operators subject to the APS boundary conditions. Let P ± be the spectral projections in L2(S1) of the boundary operators ± 1 ∂θ onto the interval (−∞, 0]. The domain of Dt is then defined to be: ∂ i for the disk. For the annulus we set: dom(Dt) = {a ∈ Ht : kDtak < ∞, r(a) ∈ Ran P +} dom(Dt) = {a ∈ Ht : kDtak < ∞, r+(a) ∈ Ran P +, r−(a) ∈ Ran P −}. Here the maps r, r± are the restriction to the boundary maps, that by the results of [9], continue to make sense for those a ∈ Ht for which kDtak < ∞. If t = 0 the domain of D0 consists of all those first Sobolev class functions f on the disk or the annulus for which the APS condition holds i.e. either r(f ) ∈ Ran P + or r+(f ) ∈ Ran P +, r−(f ) ∈ Ran P −, depending on the case. Here, by slight notational abuse, the symbols r, r± are the classical restriction to the boundary maps. It was verified in [9] that the above defined operators Dt are invertible, with bounded, and even compact inverses Qt. Using [9] we can write down the formulas for Qt. If x ∈ Λ we have the following for t > 0: CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 7 Qtxt(k) = − + U n Xi≥k NXn=0 NXn=1 Xi≤k fn+1(wt(i)2)! gn−1(wt(i)2)! (U ∗)n . wt(k + 1) · · · wt(k + n) wt(i + 1) · · · wt(i + n) · St(i)1/2St(i + n + 1)1/2 wt(k + n) wt(i) · · · wt(i + n − 1) wt(k) · · · wt(k + n − 1) · St(i)1/2St(i + n − 1)1/2 wt(i + n − 1) For the disk the second sum is from 0 to k, while for the annulus it is from −∞ to k. For t = 0 we have D0x0 = NXn=0 ei(n+1)θ 2 (cid:16)2rf ′ n(r2) − n r fn(r2)(cid:17) + NXn=1(cid:16)2rg′ n(r2) + n r gn(r2)(cid:17) e−i(n−1)θ 2 . for both the disk and annulus. From this we can compute the inverse Q0 of D0. A straight- forward calculation gives the following result: Q0x0 = − einθZ (w+)2 r2 fn+1(ρ2) rn−1 ρn d(ρ2) + NXn=0 e−inθZ r2 (w−)2 gn−1(ρ2) ρn−1 rn d(ρ2), NXn=1 for the annulus, and the same formula with w− replaced by 0 for the disk. We are now in the position to state the main results of the paper. They are summarized in the following two theorems: : defined above and let Λ be the linear subspace ofQt∈I Ht defined by 3.3 and 3.4. Also let the t ∈ I} be the family of Hilbert spaces Theorem 3.1. Given I = [0, 1), let H = {Ht conditions on wt(k) and St(k) hold. Then Λ generates a continuous field of Hilbert spaces denoted below by (I,H, Γ). Theorem 3.2. Let Qt : Ht → Ht be the collection of operators for t ∈ [0, 1) defined above. Then {Qt} is a continuous family of bounded operators in the continuous field (I,H, Γ). We finish this section by shortly indicating that the above results are also valid for families of d-bar operators studied in [3]. Let us quickly review the differences. The Hilbert space Ht studied in [3] is the completion of C ∗(Uwt) with respect to the following inner product: The quantum d-bar operator Dt of [3], acting in Ht, is given by the following formula: t = tr(Staa∗). kak2 Dta = S−1 t [a, Uwt]. It turns out that Theorems 3.1 and 3.2 are also true for the above spaces and operators. In fact the proofs are even easier in this case and Condition 4, designed to handle expressions like St(k + n)1/2St(k)1/2 is not even needed. The next section will contain all the analysis needed to prove the two theorems. 8 SLAWOMIR KLIMEK AND MATT MCBRIDE 4. Continuity and the classical limit We will prove the two theorems from the above section by a series of steps that verify the assumptions in the definitions of the continuous field of Hilbert spaces and the continuous family of bounded operators. We concentrate mainly on the annulus case. The disk case is in some respects simpler. Most of the formulas for the annulus are true also in the disk case with a modification: replacing w− by zero. The summation index in the annulus case extends to −∞ and in couple of places the corresponding sums need to be estimated. This is not the issue in the disk case where the summation starts at zero. However the major difficulty in the disk case are the wt terms in the denominator in the formula for the parametrix, since they go to zero as t goes to zero. The proofs we describe below work in both cases, but much shorter arguments are possible in the annulus case. We first verify that Λ generates a continuous field of Hilbert spaces. To this end we need to check two things: the density in Ht of x(t), x ∈ Λ, and the continuity of the norm. The density is immediate, since, for example, the canonical basis elements of Ht, see the proof of Lemma 5.1 in [9], come from Λ. The verification of the continuity of the norm is done in two steps: continuity at t = 0, and at t > 0. If x ∈ Λ, i.e. x is given by formulas 3.3 and 3.4, then the norm of xt in Ht is, for t > 0, given by (4.1) (4.2) while for t = 0 the norm of x0 is kx0k2 = (w−)2 Lemma 4.1. For n ≥ 1 we have kxtk2 = + , St(k + n)1/2St(k)1/2(cid:12)(cid:12)fn(cid:0)wt(k)2(cid:1)(cid:12)(cid:12)2 St(k + n)1/2St(k)1/2(cid:12)(cid:12)gn(cid:0)wt(k)2(cid:1)(cid:12)(cid:12)2 NXn=1Z (w+)2 (cid:12)(cid:12)fn(cid:0)r2(cid:1)(cid:12)(cid:12)2 d(cid:0)r2(cid:1) + St(k) − 1(cid:12)(cid:12)(cid:12)(cid:12) ≤ (2 + h2(t))n−1h2(t), St(k + n) (cid:18) St(k + n) St(k) − 1(cid:19) + NXn=0Xk∈S NXn=1Xk∈S NXn=0Z (w+)2 k∈S (cid:12)(cid:12)(cid:12)(cid:12) − 1(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12) St(k + n + 1) (w−)2 St(k + n) sup (cid:12)(cid:12)gn(cid:0)r2(cid:1)(cid:12)(cid:12)2 d(cid:0)r2(cid:1) . Now we are ready to discuss the continuity of norms 4.1, 4.2 as t → 0+. The next lemma is needed to handle the product of S terms with different arguments. where h2(t) is the function defined in Condition 4. Proof. The proof is by induction. For n = 1 we get Condition 4. The inductive step is St(k + n + 1) St(k) (cid:12)(cid:12)(cid:12)(cid:12) and the lemma is proved. ≤ (1 + h2(t)) (2 + h2(t))n−1h2(t) + h2(t) ≤ (2 + h2(t))nh2(t) St(k + n + 1) St(k + n) − 1(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:3) CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 9 Proposition 4.2. If xt is in Λ then t→0+ kxtk = kx0k lim Proof. Without loss of generality we may assume that xt(k) = U nfn (wt(k)2) and x0(r, ϕ) = fn(r2)einϕ , as the proof is identical for the g terms, and the elements of Λ are finite sums of such x's. We have (w−)2 (w−)2 St(k + n)1/2St(k)1/2(cid:12)(cid:12)fn(wt(k)2)(cid:12)(cid:12)2 St(k)(cid:12)(cid:12)fn(wt(k)2)(cid:12)(cid:12)2 −Z (w+)2 (cid:12)(cid:12)kxtk2 − kx0k2(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xk∈S −Z (w+)2 (cid:12)(cid:12)fn(r2)(cid:12)(cid:12)2 d(r2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xk∈S +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xk∈S(cid:0)St(k + n)1/2St(k)1/2 − St(k)(cid:1)(cid:12)(cid:12)fn(wt(k)2)(cid:12)(cid:12)2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)fn(r2)(cid:12)(cid:12)2 −Z (w+)2 (cid:12)(cid:12)kxtk2 − kx0k2(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xk∈S + const(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xk∈S St(k)(cid:12)(cid:12)fn(wt(k)2)(cid:12)(cid:12)2 − 1#(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) St(k)"(cid:18)St(k + n) St(k) (cid:19)1/2 Since fn is continuous and hence bounded, we can estimate: (w−)2 . (cid:12)(cid:12)fn(r2)(cid:12)(cid:12)2 + d(r2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . + d(r2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Using St(k) = wt(k)2 − wt(k − 1)2 and Condition 3, we see that the first term inside of the absolute value is a difference of a Riemann sum and the integral to which it converges as t → 0+. Hence this term is zero in the limit. As for the second term, since by 3.1, by Condition 4, h2(t) → 0 as t → 0+. We can now prove the first theorem. Pk∈S St(k) = (w+)2 − (w−)2 = const, Lemma 4.1 shows that it also goes to zero, because, (cid:3) Proof. (of Theorem 3.1) We have already verified that Λ satisfies some of the properties of Proposition 2.2. What remains is the proof of the continuity of the norm for t > 0. Notice that by Condition 2 all the terms in formula 4.1 are continuous in t, t > 0. Thus we need to show that the series 4.1 converges uniformly in t (away from t = 0). Assuming again that xt(k) = U nfn (wt(k)2), and using the boundedness of fn we have: St(k + n)1/2St(k)1/2(cid:12)(cid:12)fn(wt(k)2)(cid:12)(cid:12)2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)kxtk2 − M −1Xk=L+1 ≤ const Xk≥M St(k + n)1/2St(k)1/2 + constXk≤L St(k + n)1/2St(k)1/2. 10 SLAWOMIR KLIMEK AND MATT MCBRIDE We use the Cauchy-Schwarz inequality to estimate the first term: Xk≥M St(k + n)1/2St(k)1/2 ≤ Xk≥M ∞Xk=M ≤ St(k + n)!1/2 Xk≥M St(k)!1/2 ≤ (4.3) St(k) = w2 + − w2 t (M). The second term is only present in the annulus case and can be estimated in an analogous way. By Condition 2 again, the difference w2 t (M) is small for large M, uniformly in t on the intervals t ≥ ε > 0, and so, for t > 0, kxtk is (locally) the uniform limit of continuous functions and hence continuous. Therefore Λ generates a continuous field of Hilbert spaces (I,H, Γ). (cid:3) + − w2 Our next concern is with the parametrices Qt(k). To verify that they form a continuous family of bounded operators in (I,H, Γ) we need to check that they are uniformly bounded and that Q maps Γ into itself. We start with the former assertion. Proposition 4.3. The norm of Qt is uniformly bounded in t. Proof. First we write Qtxt(k) in a more compact form: where T (1,n) t Qtxt(k) = − NXn=0 f (k) =Xi≥k g(k) =Xi≤k t U nT (1,n) fn+1(wt(k)2) + NXn=1 wt(k + 1) · · · wt(k + n) wt(i + 1) · · · wt(i + n) · wt(i) · · · wt(i + n − 1) wt(k) · · · wt(k + n − 1) · T (2,n) t T (2,n) t gn−1(wt(k)2) (U ∗)n St(i)1/2St(i + n + 1)1/2 wt(k + n) St(i)1/2St(i + n − 1)1/2 wt(i + n − 1) f (i) g(i). are integral operators between weighted l2 spaces, namely: Here the operators T (1,n) T (1,n) t n and T (2,n) n+1 7→ l2 : l2 t t and T (2,n) t n−1 7→ l2 : l2 l2 n := {f : Xk∈S n where St(k + n)1/2St(k)1/2f (k)2 < ∞} The main technique used to estimate the norms will be the Schur-Young inequality: T : L2(Y ) −→ L2(X) is an integral operator T f (x) =R K(x, y)f (y)dy, then one has kTk2 ≤(cid:18)sup x∈XZY K(x, y)dy(cid:19)(cid:18)sup y∈YZX K(x, y)dx(cid:19) . if The details can be found in [6]. We will also use two integral estimates, with t independent right hand sides: St(k) wt(k) ≤Z (w+)2 wt(i)2 Xi<k dx √x = 2(w+ − wt(i)) ≤ 2(w+ − w−), (4.4) CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 11 St(k) wt(k) ≤Z wt(i)2 (w−)2 dx √x = 2(wt(i) − w−) ≤ 2(w+ − w−). (4.5) Xk≤i Such estimates were described and used in [10] and are simply obtained by estimating the area under the graph of x−1/2, like in the integral test for series. First we estimate the norm of T (1,n) t . Repeatedly using the monotonicity of wt(i) and the Cauchy-Schwarz inequality, we have, like in [10]: St(i)1/2St(i + n + 1)1/2 t k2 ≤ sup kT (1,n) ≤"sup k∈S Xi≥k k∈S Xi≥k wt(i)! Xi≥k St(i) wt(i + n) ! sup i∈S Xk≤i wt(i + n) !#1/2"sup i∈S Xk≤i St(i + n + 1) Using 3.2, 4.4, and 4.5 we see that the norm of T (1,n) estimate on T (2,n) is essentially the same. Therefore one has t t St(k)1/2St(k + n)1/2 ! wt(i + n) St(k) wt(k)! Xk≤i St(k + n) wt(k + n)!#1/2 . (4.6) is bounded uniformly in n and t. The and this completes the proof. (cid:3) kQtk ≤ sup n∈N kT (1,n) t k + sup n∈N kT (2,n) t k ≤ const Next we need to prove that Q maps Γ into itself. This requires checking condition (3) of Proposition 2.4. Thus we need to show that, given x ∈ Λ, Qx is approximable by Λ at every t ∈ I. The hardest part is to show that this is true around t = 0, which we will do now. Let x ∈ Λ be given by formulas 3.3 and 3.4, and define ρn−1 rn d(ρ2), gn−1(ρ2) Notice that one has y ∈ Λ since clearly efn(r2), egn(r2) are in C([(w−)2, (w+)2]), and also we have obvious Q0x0 = y0 which was the motivating property of the above construction of y. We will show that x ∈ Λ is approximable by y ∈ Λ at t = 0. This is stronger than proving that x is approximable by Λ at t = 0. Proposition 4.4. With the above notation the following is true: t→0+ kQtxt − ytk = 0. lim and similarly and set and for t = 0: (w−)2 fn+1(ρ2) egn(r2) :=Z r2 efn(r2) :=Z (w+)2 yt(k) :=Xn≤N U nefn(cid:0)wt(k)2(cid:1) +Xn≤N egn(cid:0)wt(k)2(cid:1) (U ∗)n , y0(r, ϕ) :=Xn≤N efn(r2)einϕ +Xn≤N egn(r2)e−inϕ. rn−1 ρn d(ρ2), r2 12 SLAWOMIR KLIMEK AND MATT MCBRIDE Proof. We show the details for a single gn term in the finite sum. We first obtain a pointwise estimate. Adding and subtracting we get: wt(i + n − 1) wt(i)n−1 St(i)1/2St(i + n − 1)1/2 gn−1(cid:0)wt(i)2(cid:1) − egn(cid:0)wt(k)2(cid:1)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xi≤k wt(i) · · · wt(i + n − 1) wt(k) · · · wt(k + n − 1) ≤Xi≤k(cid:12)(cid:12)(cid:12)(cid:12) wt(i) · · · wt(i + n − 2) wt(k) · · · wt(k + n − 1) − +Xi≤k wt(k) · · · wt(k + n − 1)(cid:12)(cid:12)St(i)1/2St(i + n − 1)1/2 − St(i)(cid:12)(cid:12)(cid:12)(cid:12)gn−1(cid:0)wt(i)2(cid:1)(cid:12)(cid:12) + wt(i) · · · wt(i + n − 2) +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xi≤k wt(k)n gn−1(cid:0)wt(i)2(cid:1) St(i) −Z wt(k)2 wt(k)n (cid:12)(cid:12)(cid:12)(cid:12) St(i)(cid:12)(cid:12)gn−1(cid:0)wt(i)2(cid:1)(cid:12)(cid:12) + wt(k)n gn−1(ρ2) d(ρ2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) wt(i)n−1 ρn−1 (w−)2 := I + II + III. Let us discuss the structure of the above terms. The expression inside the absolute value in term I unfortunately in general does not go to zero as t goes to zero. To go around it we show that the expression is small for large k which then lets us use the smallness of St(i) to get the desired limit. This term is the trickiest to handle. Term II is the most straightforward to estimate along the lines of the proof of Proposition 4.2. Finally expression III is a difference between an integral and its Riemann sum, but because of the small denominator it has to be estimated carefully. To handle term I we need the following observation. Lemma 4.5. With the above notation we have: (cid:12)(cid:12)(cid:12)(cid:12)1 − wt(k)n−1 wt(k + 1) · · · wt(k + n − 1)(cid:12)(cid:12)(cid:12)(cid:12) ≤ where h3(k) is the sequence of Condition 5. n−1Xj=0 jh3(k + n − j), Proof. To prove the statement we write wt(k)n−1 wt(k + 1) · · · wt(k + n − 1) and use an elementary inequality: = wt(k) wt(k)wt(k + 1) wt(k + 1) wt(k + 1)wt(k + 2) wt(k) · · · wt(k + n − 2) wt(k + 1) · · · wt(k + n − 1) if xk ≤ 1. 1 − x1 · · · xn ≤ 1 − x1 + . . . + 1 − xn (cid:3) We concentrate on the expression inside the absolute value in term I: wt(i) · · · wt(i + n − 2) wt(k) · · · wt(k + n − 1) − wt(i) · · · wt(i + n − 2) wt(k) · · · wt(k + n − 1) − wt(k) · · · wt(k + n − 1) − wt(i)n−1 J :=(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12) wt(i)n−1 wt(i)n−1 wt(k)n (cid:12)(cid:12)(cid:12)(cid:12) ≤ wt(k) · · · wt(k + n − 1)(cid:12)(cid:12)(cid:12)(cid:12) + wt(k)n (cid:12)(cid:12)(cid:12)(cid:12) . wt(i)n−1 CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 13 Factoring we get: Using lemma 4.5 yields: 1 1 wt(i)n−2 wt(i + 1) · · · wt(i + n − 2)(cid:12)(cid:12)(cid:12)(cid:12) + wt(k)n−1 wt(k + n − 1)(cid:12)(cid:12)(cid:12)(cid:12)1 − wt(k)(cid:12)(cid:12)(cid:12)(cid:12)1 − n−2Xj=0 wt(k + 1) · · · wt(k + n − 1)(cid:12)(cid:12)(cid:12)(cid:12) . n−1Xj=0 jh3(i + n − 1 − j) + wt(i) 1 J ≤ + 1 1 h4(i) + h5(k). 1 wt(i) J ≤ =: wt(k + n − 1) wt(k + n − 1) jh3(k + n − j) = The functions h4(k) and h5(k) above are t independent and go to zero as k → ±∞. Conse- quently: I(k) ≤ const ≤ const 1 wt(k + n − 1)Xi≤k wt(k + n − 1)Xi≤k 1 St(i)h4(i) + const h5(k)Xi≤k St(i) wt(i) ≤ St(i)h4(i) + const h5(k) =: I1 + I2. We use the following lemma to handle both I1 and I2. This is the tricky part of the argument. Lemma 4.6. If h(k) → 0 as k → ±∞ then and first choose N such that sup k>N h(k) ≤ ε/2 and then choose δ > 0 such that Pk≤N ε/2 far all t ≤ δ. The last inequality is possible because of Condition 3. As a corollary we also have: lim t→0+Xk∈S St(k + n)1/2St(k)1/2h(k) = 0, obtained by estimating like in 4.3: St(k + n)1/2St(k)1/2h(k) ≤ Xk∈S ≤ Xk∈S St(k + n)!1/2 Xk∈S St(k)h(k)2!1/2 ≤ const Xk∈S St(k)h(k)2!1/2 . St(k) ≤ (cid:3) (4.7) Proof. We split the sum: lim t→0+Xk∈S St(k)h(k) = Xk≤N St(k)h(k) = 0. St(k)h(k) + Xk>N St(k)h(k) ≤ St(k) + const sup k>N h(k) Xk∈S ≤ const Xk≤N 14 SLAWOMIR KLIMEK AND MATT MCBRIDE We proceed to show that the norms I1 and I2 are small for small t. This is more straight- 5(k) forward with the I2 term. Namely we have kI2k2 ≤ constPk∈S St(k + n)1/2St(k)1/2h2 which by 4.7 goes to zero as t goes to zero. To estimate I1 we notice first that I1(k) ≤ constXi≤k by 4.5. Consequently we have: St(i) wt(i) kI1k2 =Xk∈S St(i) wt(i) ≤ const h4(i) ≤ constXi≤k 1 (k) ≤ constXk∈S St(k + n)1/2St(k)1/2I 2 St(k + n)1/2St(k)1/2I1(k) ≤ ≤ const Xi,k∈S St(k + n)1/2St(k)1/2 wt(k + n) St(i)h4(i) ≤ constXi∈S St(i)h4(i). The sum over k above is estimated as in 4.6, and we can use Lemma 4.6 again to conclude that kI1k2 goes to zero as t goes to zero. Now we estimate term II. This is done analogously to the way we treated the second term in Proposition 4.2. Using the boundedness of gn−1, the definition of h2(t), and 4.5, we have: II(k) ≤Xi≤k (cid:12)(cid:12)St(i + n − 1)1/2St(i)1/2 − St(i)(cid:12)(cid:12) wt(i) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)St(i + n − 1) (cid:19)1/2 ≤ constXi≤k Consequently kIIk2 ≤ const h2 Finally we estimate III(k). It is clear that this expression is small for small t and a fixed k, as a difference between an integral and its Riemann sum. However this is not enough in the disk case when wt(k)n in the denominator is small for small t. To overcome this difficulty we first replace gn−1 by its step function approximation and then deal directly with the remaining integral of ρn−1. 2(t) which goes to zero by Condition 4. (cid:12)(cid:12)gn−1(wt(i)2)(cid:12)(cid:12) ≤ − 1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ const h2(t). wt(i + n − 1) St(i) St(i) With this strategy in mind we estimate: (w−)2 ρn−1 wt(i)n−1 III(k) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xi≤k ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xi≤k wt(i)n−1 +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xi≤kZ wt(i)2 wt(k)n gn−1(cid:0)wt(i)2(cid:1) St(i) −Z wt(k)2 wt(k)n gn−1(cid:0)wt(i)2(cid:1) St(i) −Z wt(i)2 wt(k)n(cid:0)gn−1(cid:0)wt(i)2(cid:1) − gn−1(ρ2)(cid:1) d(ρ2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) wt(k)n gn−1(ρ2) d(ρ2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ wt(k)n gn−1(cid:0)wt(i)2(cid:1) d(ρ2)!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ρn−1 ρn−1 wt(i−1)2 wt(i−1)2 =: III1(k) + III2(k). + Since continuous functions on a closed interval are uniformly continuous, the function h6(t) := sup i∈S sup ρ2∈[(wt(i−1)2,(wt(i))2](cid:12)(cid:12)gn−1(cid:0)wt(i)2(cid:1) − gn−1(cid:0)ρ2(cid:1)(cid:12)(cid:12) CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 15 goes to zero as t → 0+. Consequently, using the definition of h6(t), term III2 can be estimated as follows: ρn−1 wt(k)n d(ρ2) ≤ h6(t)wt(k)Z 1 0 un−1 d(u2) ≤ const h6(t). (w−)2 III2(k) ≤ h6(t)Z wt(k)2 III1(k) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Xi≤kZ wt(i)2 ≤ constXi≤kZ wt(i)2 This means that kIII2k goes to zero as t → 0+. When estimating III1 we first eliminate gn−1 using its boundedness: wt(i−1)2(cid:18)wt(i)n−1 wt(i−1)2(cid:18)wt(i)n−1 wt(k)n gn−1(cid:0)wt(i)2(cid:1) − wt(k)n(cid:19) d(ρ2). wt(k)n − ρn−1 ρn−1 wt(k)n gn−1(cid:0)wt(i)2(cid:1)(cid:19) d(ρ2)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ What is left is the difference between the integral of ρn−1 and its upper sum which we handle like in the error estimate of the integral test for series. This is summarized in the following sequence of inequalities. wt(k)n − III1(k) ≤ constXi≤k(cid:18)wt(i)n−1 ≤ const Xi≤k ≤ const Xi≤k−1 wt(k)n St(i) − Xi≤k−1 wt(k)n St(i)(cid:18)1 − wt(i)n−1 wt(i)n−1 wt(i − 1)n−1 wt(i)n−1 wt(k)n (cid:19) St(i) ≤ wt(k)n St(i + 1)! ≤ St(i) (cid:19) + const St(k) wt(k) St(i + 1) . Notice that St(k)2 wt(k)2 = St(k) wt(k)2 − wt(k − 1)2 wt(k)2 Hence, using the monotonicity of wt(i) we have ≤ St(k) ≤ h1(t). III1(k) ≤ const h2(t) Xi≤k−1 wt(i)n wt(k)n St(i) wt(i) + constph1(t) ≤ const(cid:16)h2(t) +ph1(t)(cid:17) , and again kIII1k goes to zero as t → 0+. The proof of the proposition is complete. (cid:3) To proceed further we need a better understanding of Γ, the space of continuous sections of our continuous field. We have the following useful result. Lemma 4.7. For t > 0 consider the following expression xt(k) =Xn≤N U nFn(t, k) +Xn≤N Gn(t, k)(U ∗)n such that the functions t 7→ Fn(t, k) and t 7→ Gn(t, k) are continuous for every k, and such that Fn(t, k) and Gn(t, k) are bounded (in both variables). Then xt is approximable by Λ at every t > 0. 16 SLAWOMIR KLIMEK AND MATT MCBRIDE Proof. Without a loss of generality we may assume that xt(k) = U nFn(t, k) as the proof is identical for the G terms, and it will extend to finite sums of such x's. Given t0 ∈ I and ǫ > 0, let y ∈ Λ be such that for t > 0 yt(k) := U nfn(wt(k)2), where we choose fn ∈ C([(w−)2, (w+)2]) such that kFn(t0,·) − fn (wt0(·)2)k ≤ ε 2. This is always possible since the space of sequences of the form k → fn(wt0(k)2), where fn ∈ C([(w−)2, (w+)2]), is a dense subspace in the Hilbert space l2 n. We want to show that kxt − ytk ≤ ε for all t sufficiently close to t0. By the construction of fn this is true at t = t0. We will prove that t → kxt − ytk is continuous for t > 0 which will imply the above inequality. But the inequality means that x is approximable by Λ at t = t0, which is exactly what we want to achieve. The proof that t → kxt − ytk is continuous is analogous to the last part of the proof of Indeed, by the Theorem 3.1, that the norm is continuous for elements of Λ and t > 0. continuity assumptions, kxt − ytk2 is an infinite sum of continuous functions: . kxt − ytk2 =Xk∈S St(k + n)1/2St(k)1/2(cid:12)(cid:12)Fn(t, k) − fn(wt(k)2)(cid:12)(cid:12)2 The series converges uniformly around t0 because, by the boundedness assumptions, we can estimate the remainder as follows: Xk≥M St(k + n)1/2St(k)1/2(cid:12)(cid:12)Fn(t, k) − fn(wt(k)2)(cid:12)(cid:12)2 ≤ const Xk≥M For large M this is small by 4.3. In the annulus case there is also a remainder at −∞ which also goes to zero by an analogous estimate. As a consequence t → kxt − ytk is indeed continuous for t > 0 and the lemma is proved. (cid:3) St(k + n)1/2St(k)1/2. We now have all the tools to finish the proof the second theorem. Proof. (of Theorem 3.2) What remains is to show that Qtxt is approximable by Λ for t > 0 since Propositions 4.3 and 4.4 establish the other properties of {Qt} needed to conclude that they form a continuous family of bounded operators in (I,H, Γ). To prove that Qtxt is approximable by Λ for t > 0 we use Lemma 4.7 with Fn(t, k) =Xi≥k Gn(t, k) =Xi≤k Fn(t, i) :=Xi≥k Gn(t, i) :=Xi≤k wt(k + 1) · · · wt(k + n) wt(i + 1) · · · wt(i + n) · wt(i) · · · wt(i + n − 1) wt(k) · · · wt(k + n − 1) · St(i)1/2St(i + n + 1)1/2 wt(k + n) fn+1(i) St(i)1/2St(i + n − 1)1/2 wt(i + n − 1) gn−1(i). Thus we need to show that Fn(t, k) and Gn(t, k) are continuous and bounded functions of t, for t > 0. This will be done for the Fn(t, k) term only as the argument is analogous for the Gn(t, k) term. In fact, in the disk case the Gn(t, k) is only a finite sum, so the continuity for t > 0 follows immediately from Condition 2. CLASSICAL LIMIT OF THE D-BAR OPERATORS ON QUANTUM DOMAINS 17 Each Fn(t, i) is continuous on the intervals t ≥ ε > 0 by Condition 2, so we must show that for each k, the series defining Fn(t, k) converges uniformly in t. To estimate the tail end of the series we use 3.2, 4.3, and the boundedness of fn+1(i) to get (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∞Xi=M ≤ const wt(k + n) ∞Xi=M fn+1(i)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ t (M)(cid:1) , wt(k + n)(cid:0)w2 + − w2 const wt(k + 1) · · · wt(k + n) wt(i + 1) · · · wt(i + n) · St(i)1/2St(i + n + 1)1/2 wt(k + n) St(i)1/2St(i + n + 1)1/2 ≤ which goes to zero uniformly on the intervals t ≥ ε > 0 as M goes to infinity by Condition 2. Hence Fn(t, k) is a uniform limit of continuous functions and consequently it is continuous for t > 0 and for each k. Next we show that Fn(t, k) and Gn(t, k) are bounded. We have: St(i)1/2St(i + n + 1)1/2 ≤ Fn(t, k) ≤ constXi≥k ≤ const Xi≥k wt(i + n) St(i) wt(i)!1/2 Xi≥k St(i + n + 1) wt(i + n) !1/2 ≤ const, where we used 3.2, 4.4, and 4.5. Similar argument works also for estimating Gn(t, k). Thus the assumptions of Lemma 4.7 are satisfied and Qtxt is approximable by Λ at every t. Hence the collection {Qt} is a continuous family of bounded operators. This finishes the proof. (cid:3) References [1] Atiyah, M. F., Patodi, V. K. and Singer I. M., Spectral asymmetry and Riemannian geometry I, II, III, Math. Proc. Camb. Phil. Soc. 77(1975) 43-69, 78(1975) 43-432, 79(1976) 71-99. [2] Borthwick, D., Klimek, S., Lesniewski, A., and Rinaldi, M.: Supersymmetry and Fredholm modules over quantized spaces, Comm. Math. Phys., 166, 397 -- 415 (1994) [3] Carey, A., Klimek, S. and Wojciechowski, K. P., Dirac operators on noncommutative manifolds with boundary. arXiv:0901.0123v1 [4] Connes, A., Non-Commutative Differential Geometry, Academic Press, 1994 [5] Dixmier, J. C ∗ − Algebras, North Holland Publishing Company 1997 [6] Halmos, P.R. and Sunder V.S., Bounded Integral Operators on L2 Spaces, Springer-Verlag, 1978. [7] Klimek, S.: A note on noncommutative holomorphic and harmonic functions on the unit disk. in Anal- ysis, Geometry and Topology of Elliptic Operators, Papers in Honour of Krzysztof P. Wojciechowski's 50th birthday, edited by Bernhelm BooBavnbek, Slawomir Klimek, Matthias Lesch and Weiping Zhang, World Scientific Publishing Company, 383 -- 400 (2006) [8] Klimek, S. and Lesniewski, A., Quantum Riemann surfaces, I. The unit disk, Comm. Math. Phys., 146, 103-122(1992) [9] Klimek, S. and McBride, M., D-bar Operators on Quantum Domains. Math Phys Anal Geom (2010) 13:357390, DOI 10.1007/s11040-010-9084-9 [10] Klimek, S. and McBride, M., Dirac Operators on the Punctured Disk. SIGMA 6 (2010), 056, 12 pages [11] Vaksman, L.: Quantum Bounded Symmetric Domains, AMS, 2010 18 SLAWOMIR KLIMEK AND MATT MCBRIDE Department of Mathematical Sciences, Indiana University-Purdue University Indianapo- lis, 402 N. Blackford St., Indianapolis, IN 46202, U.S.A. E-mail address: [email protected] Department of Mathematical Sciences, Indiana University-Purdue University Indianapo- lis, 402 N. Blackford St., Indianapolis, IN 46202, U.S.A. E-mail address: [email protected]
1704.07095
1
1704
2017-04-24T08:57:20
Quantitative version of the Bishop-Phelps-Bollob\'as theorem for operators with values in a space with the property $\beta$
[ "math.FA" ]
The Bishop-Phelps-Bollob\'as property for operators deals with simultaneous approximation of an operator $T$ and a vector $x$ at which $T: X\rightarrow Y$ nearly attains its norm by an operator $F$ and a vector $z$, respectively, such that $F$ attains its norm at $z$. We study the possible estimates from above and from below for parameters that measure the rate of approximation in the Bishop-Phelps-Bollob\'as property for operators for the case of $Y$ having the property $\beta$ of Lindenstrauss.
math.FA
math
QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS WITH VALUES IN A SPACE WITH THE PROPERTY β VLADIMIR KADETS AND MARIIA SOLOVIOVA ABSTRACT. The Bishop-Phelps-Bollob´as property for operators deals with simultaneous approximation of an operator T and a vector x at which T : X → Y nearly attains its norm by an operator F and a vector z, respectively, such that F attains its norm at z. We study the possible estimates from above and from below for parameters that measure the rate of approximation in the Bishop-Phelps-Bollob´as property for operators for the case of Y having the property β of Lindenstrauss. 1. INTRODUCTION In this paper X, Y are real Banach spaces, L(X, Y ) is the space of all bounded linear operators T : X → Y , L(X) = L(X, X), X∗ = L(X, R), BX and SX denote the closed unit ball and the unit sphere of X, respectively. A functional x∗ ∈ X∗ attains its norm, if there is x ∈ SX with x∗(x) = (cid:107)x∗(cid:107). The Bishop-Phelps theorem [3] (see also [8, Chapter 1, p. 3]) says that the set of norm-attaining function- als is always dense in X∗. In [4] B. Bollob´as remarked that in fact the Bishop-Phelps construction allows to approximate at the same time a functional and a vector in which it almost attains the norm. Nowadays this very useful fact is called the Bishop-Phelps-Bollob´as theorem. Recently, two moduli have been in- troduced [5] which measure, for a given Banach space, what is the best possible Bishop-Phelps-Bollob´as theorem in that space. We will use the following notation: Π(X) :=(cid:8)(x, x∗) ∈ X × X∗ : (cid:107)x(cid:107) = (cid:107)x∗(cid:107) = x∗(x) = 1(cid:9). Definition 1.1 (Bishop-Phelps-Bollob´as moduli, [5]). Let X be a Banach space. The Bishop-Phelps- Bollob´as modulus of X is the function ΦX : (0, 2) −→ R+ such that given ε ∈ (0, 2), ΦX (ε) is the infimum of those δ > 0 satisfying that for every (x, x∗) ∈ BX × BX∗ with x∗(x) > 1 − ε, there is (y, y∗) ∈ Π(X) with (cid:107)x − y(cid:107) < δ and (cid:107)x∗ − y∗(cid:107) < δ. Substituting (x, x∗) ∈ SX × SX∗ instead of (x, x∗) ∈ BX × BX∗ in the above sentence, we obtain the definition of the spherical Bishop-Phelps- Bollob´as modulus ΦS X (ε). Evidently, ΦS X (ε) (cid:54) ΦX (ε). There is a common upper bound for ΦX (·) (and so for ΦS X (·)) for all Banach spaces which is actually sharp. Namely [5], for every Banach space X and every ε ∈ (0, 2) one has ΦX (ε) (cid:54) √ 2ε. In other words, this leads to the following improved version of the Bishop-Phelps- Bollob´as theorem. Proposition 1.2 ([5, Corollary 2.4]). Let X be a Banach space and 0 < ε < 2. Suppose that x ∈ BX and x∗ ∈ BX∗ satisfy x∗(x) > 1 − ε. Then, there exists (y, y∗) ∈ Π(X) such that (cid:107)x − y(cid:107) < 2ε and (cid:107)x∗ − y∗(cid:107) < √ √ 2ε. The sharpness of this version is demonstrated in [5, Example 2.5] by just considering X = (cid:96)(2) √ 1 , the 2ε for two-dimensional real (cid:96)1 space. For a uniformly non-square Banach space X one has ΦX (ε) < Date: 2017. 2010 Mathematics Subject Classification. Primary 46B04; Secondary 46B20, 46B22, 47A30. Key words and phrases. Bishop-Phelps-Bollob´as theorem; norm-attaining operators; property β of Lindenstrauss. 1 2 all ε ∈ (0, 2) ([5, Theorem 5.9], [7, Theorem 2.3]). A quantifcation of this inequality in terms of a parameter that measures the uniform non-squareness of X was given in [6, Theorem 3.3]. KADETS AND SOLOVIOVA Lindenstrauss in [12] examined the extension of the Bishop -- Phelps theorem on denseness of the family of norm-attaining scalar-valued functionals on a Banach space, to vector-valued linear operators. He introduced the property β, which is possessed by polyhedral finite-dimensional spaces, and by any subspace of (cid:96)∞ that contains c0. Definition 1.3. A Banach space Y is said to have the property β if there are two sets {yα : α ∈ Λ} ⊂ SY , {y∗ Y and 0 (cid:54) ρ < 1 such that the following conditions hold α : α ∈ Λ} ⊂ S∗ α(yα) = 1, α(yγ) (cid:54) ρ if α (cid:54)= γ, (i) y∗ (ii) y∗ (iii) (cid:107)y(cid:107) = sup{y∗ Denote for short by β(Y ) (cid:54) ρ that a Banach space Y has the property β with parameter ρ ∈ (0, 1). Obviously, if ρ1 (cid:54) ρ2 < 1 and β(Y ) (cid:54) ρ1 , then β(Y ) (cid:54) ρ2. If Y has the property β with parameter ρ = 0, we will write β(Y ) = 0. α(y) : α ∈ Λ}, for all y ∈ Y . Lindenstrauss proved that if a Banach space Y has the property β, then for any Banach space X the set of norm attaining operators is dense in L(X, Y ). It was proved later by J. Partington [10] that every Banach space can be equivalently renormed to have the property β. In 2008, Acosta, Aron, Garc´ıa and Maestre in [1] introduced the following Bishop-Phelps-Bollob´as property as an extension of the Bishop-Phelps-Bollob´as theorem to the vector-valued case. Definition 1.4. A couple of Banach spaces (X, Y ) is said to have the Bishop-Phelps-Bollob´as property for operators if for any δ > 0 there exists a ε(δ) > 0, such that for every operator T ∈ SL(X,Y ), if x ∈ SX and (cid:107)T (x)(cid:107) > 1 − ε(δ), then there exist z ∈ SX and F ∈ SL(X,Y ) satisfying (cid:107)F (z)(cid:107) = 1,(cid:107)x − z(cid:107) < δ and (cid:107)T − F(cid:107) < δ. In [1, Theorem 2.2] it was proved that if Y has the property β, then for any Banach space X the pair (X, Y ) has the Bishop-Phelps-Bollob´as property for operators. In this article we introduce an analogue of the Bishop-Phelps-Bollob´as moduli for the vector-valued case. Definition 1.5. Let X, Y be Banach spaces. The Bishop-Phelps-Bollob´as modulus (spherical Bishop- Phelps-Bollob´as modulus) of a pair (X, Y ) is the function Φ(X, Y,·) : (0, 1) −→ R+ (ΦS(X, Y,·) : (0, 1) −→ R+) whose value in point ε ∈ (0, 1) is defined as the infimum of those δ > 0 such that for every (x, T ) ∈ BX × BL(X,Y ) ((x, T ) ∈ SX × SL(X,Y ) respectively) with (cid:107)T (x)(cid:107) > 1 − ε, there is (z, F ) ∈ SX × SL(X,Y ) with (cid:107)x − z(cid:107) < δ and (cid:107)T − F(cid:107) < δ. Under the notation Πε(X, Y ) = {(x, T ) ∈ X × L(X, Y ) : (cid:107)x(cid:107) (cid:54) 1,(cid:107)T(cid:107) (cid:54) 1, (cid:107)T (x)(cid:107) > 1 − ε} , ε (X, Y ) = {(x, T ) ∈ X × L(X, Y ) : (cid:107)x(cid:107) = (cid:107)T(cid:107) = 1, (cid:107)T (x)(cid:107) > 1 − ε} , ΠS Π(X, Y ) = {(x, T ) ∈ X × L(X, Y ) : (cid:107)x(cid:107) = 1,(cid:107)T(cid:107) = 1, (cid:107)T (x)(cid:107) = 1} , the definition can be rewritten as follows: Φ(X, Y, ε) = sup (x,T )∈Πε(X,Y ) inf (z,F )∈Π(X,Y ) ΦS(X, Y, ε) = sup (x,T )∈ΠS ε (X,Y ) inf (z,F )∈Π(X,Y ) max{(cid:107)x − z(cid:107),(cid:107)T − F(cid:107)}, max{(cid:107)x − z(cid:107),(cid:107)T − F(cid:107)}. QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS 3 Evidently, ΦS(X, Y, ε) (cid:54) Φ(X, Y, ε), so any estimation from above for Φ(X, Y,·) is also valid for ΦS(X, Y,·) and any estimation from below for ΦS(X, Y,·) is applicable to Φ(X, Y,·). Also the follow- ing result is immediate. Remark 1.6. Let X, Y be Banach spaces, ε1, ε2 > 0 with ε1 < ε2. Then Πε1(X, Y ) ⊂ Πε2(X, Y ) and ε1(X, Y ) ⊂ ΠS ΠS Notice that a couple (X, Y ) has the Bishop-Phelps-Bollob´as property for operators if and only if ε2(X, Y ). Therefore, Φ(X, Y, ε) and ΦS(X, Y, ε) do not decrease as ε increases. Φ(X, Y, ε) −−−→ ε→0 0. The aim of our paper is to estimate the Bishop-Phelps-Bollob´as modulus for operators which act to a Banach space with the property β. This paper is organized as follows. After the Introduction, in Section 2 we will provide an estimation from above for Φ(X, Y, ε) for Y possessing the property β of Lindenstrauss (Theorem 2.1) and an improvement for the case of X being uniformly non-square (Theorem 2.6). Section 3 is devoted to estimations of Φ(X, Y, ε) from below and related problems. As a bi-product of these estimations we obtain an interesting effect (Theorem 3.7) that Φ(X, Y, ε) is not continuous with respect to the variable Y . In Section 4 we consider a modification of the above moduli which appear if one approximates by pairs (y, F ) with (cid:107)F(cid:107) = (cid:107)F y(cid:107) without requiring (cid:107)F(cid:107) = 1. Finally, in a very short Section 5 we speak about a natural question which we did not succeed to solve. 2. ESTIMATION FROM ABOVE Our first result is the upper bound of the Bishop-Phelps-Bollob´as moduli for the case when the range space has the property β of Lindenstrauss. Theorem 2.1. Let X and Y be Banach spaces such that β(Y ) (cid:54) ρ. Then for every ε ∈ (0, 1) (2.1) ΦS(X, Y, ε) (cid:54) Φ(X, Y, ε) (cid:54) min 2ε , 2 . (cid:26)√ (cid:114) 1 + ρ 1 − ρ (cid:27) The above result is a quantification of [1, Theorem 2.2] which states that if Y has the property β, then for any Banach space X the pair (X, Y ) has the Bishop-Phelps-Bollob´as property for operators. The construction is borrowed from the demonstration of [1, Theorem 2.2], but in order to obtain (2.1) we have to take care about details and need some additional work. At first, we have to modify a little bit the original results of Phelps about approximation of a functional x∗ and a vector x. Proposition 2.2 ([13], Corollary 2.2). Let X be a real Banach space, x ∈ BX, x∗ ∈ SX∗, η > 0 and x∗(x) > 1 − η. Then for any k ∈ (0, 1) there exist ζ∗ ∈ X∗ and y ∈ SX such that (cid:107)x∗ − ζ∗(cid:107) < k. For our purposes we need an improvement which allows to take x∗ ∈ BX∗. ζ∗(y) = (cid:107)ζ∗(cid:107) , (cid:107)x − y(cid:107) < η k , Lemma 2.3. Let X be a real Banach space, x ∈ BX, x∗ ∈ BX∗, ε > 0 and x∗(x) > 1 − ε. Then for any k ∈ (0, 1) there exist y∗ ∈ X∗ and z ∈ SX such that 1 − 1−ε(cid:107)x∗(cid:107) y∗(z) = (cid:107)y∗(cid:107) , (cid:107)x − z(cid:107) < (2.2) Moreover, for any k ∈ [ε/2, 1) there exist z∗ ∈ SX∗ and z ∈ SX such that (2.3) (cid:107)x − z(cid:107) < z∗(z) = 1, (cid:107)x∗ − z∗(cid:107) < 2k. k , (cid:107)x∗ − y∗(cid:107) < k (cid:107)x∗(cid:107) . ε k , KADETS AND SOLOVIOVA 4 Proof. We have that x∗ k ∈ (0, 1) there exist ζ∗ ∈ X∗ and z ∈ SX such that (cid:107)x − z(cid:107) < ζ∗(z) = (cid:107)ζ∗(cid:107) , (cid:107)x∗(cid:107) (x) > 1 − η for η = 1 − 1−ε(cid:107)x∗(cid:107) and we can apply Proposition 2.2. So, for any In order to get (2.2) it remains to introduce y∗ = (cid:107)x∗(cid:107) · ζ∗. This functional also attains its norm at z and , η k (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗(cid:107) − ζ∗(cid:13)(cid:13)(cid:13)(cid:13) < k. (cid:107)x∗(cid:107) − ζ∗(cid:13)(cid:13)(cid:13)(cid:13) < k (cid:107)x∗(cid:107) . (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗ − y∗(cid:107) = (cid:107)x∗(cid:107) · In order to demonstrate the "moreover" part, take k((cid:107)x∗(cid:107) − (1 − ε)) ε(cid:107)x∗(cid:107) . k = (cid:16) 1 ε − (1−ε) ε(cid:107)x∗(cid:107) (cid:17) (cid:54) (cid:16) 1 ε − (1−ε) The inequality (cid:107)x∗(cid:107) (cid:62) x∗(x) > 1 − ε implies that k > 0. On the other hand, k = k k Denote z∗ = y∗ (cid:17) = k < 1, so for this k we can find y∗ ∈ X∗ and z ∈ SX such that (2.2) holds true. (cid:107)y∗(cid:107). Then (cid:107)x − z(cid:107) < ε/k and ε (cid:107)x∗ − z∗(cid:107) (cid:54) (cid:107)x∗ − y∗(cid:107) + (cid:107)y∗ − z∗(cid:107) (cid:54) (cid:107)x∗ − y∗(cid:107) + 1 − (cid:107)y∗(cid:107) (cid:54) (cid:107)x∗ − y∗(cid:107) + 1 − (cid:107)x∗(cid:107) + (cid:107)x∗(cid:107) − (cid:107)y∗(cid:107) (cid:54) 2(cid:107)x∗ − y∗(cid:107) + 1 − (cid:107)x∗(cid:107) . So, we have (cid:107)x∗ − z∗(cid:107) < 2k (cid:107)x∗(cid:107) + 1 − (cid:107)x∗(cid:107) = 2k · ((cid:107)x∗(cid:107) − (1 − ε)) + 1 − (cid:107)x∗(cid:107) (cid:54) 2k. ε + 1− t with t ∈ (1− ε, 1), is increasing (cid:3) ε The last inequality holds, since the function f (t) = 2k·(t−(1−ε)) when k (cid:62) ε/2, so max f = f (1) = 2k. Remark 2.4. One can easily see that for k < ε 2 the "moreover" part with (2.3) is trivially true (and is not sharp) because in this case the inequality (cid:107)x − z(cid:107) (cid:54) ε is weaker than the triangle inequality k (cid:107)x − z(cid:107) (cid:54) 2, so one can just use the density of the set of norm-attaining functionals in order to get the desired (z, z∗) ∈ Π(X) with (cid:107)x∗ − z∗(cid:107) < 2k. Proof of Theorem 2.1. We will use the notations {yα : α ∈ Λ} ⊂ SY and {y∗ Definition 1.3 of the property β. Y from Consider T ∈ BL(X,Y ) and x ∈ BX such that (cid:107)T x(cid:107) > 1 − ε. According to (iii) of Definition 1.3, 2 , 1) and for any δ > 0 there there is α0 ∈ Λ such that y∗ exist z∗ ∈ SX∗ and z ∈ SX such that z∗(z) = 1, (cid:107)z − x(cid:107) < ε/k and (cid:107)z∗ − T ∗(y∗ α0(T x) > 1 − ε. By Lemma 2.3, for any k ∈ [ ε α : α ∈ Λ} ⊂ S∗ α0)(cid:107) < 2k. For η = 2k ρ 1−ρ let us introduce the following operator S ∈ L(X, Y ) (2.4) S(v) = T (v) + [(1 + η)z∗(v) − (T ∗y∗ α0)(v)]yα0. Remark, that for all y∗ ∈ Y ∗ S∗(y∗) = T ∗(y∗) + [(1 + η)z∗ − T ∗y∗ α0]y∗(yα0). According to (iii) of Definition 1.3 the set {y∗ supα (cid:107)S∗y∗ α(cid:107). Let us calculate the norm of S. (cid:107)S(cid:107) (cid:62)(cid:13)(cid:13)S∗(y∗ α0)(cid:13)(cid:13) = (1 + η)(cid:107)z∗(cid:107) = 1 + η. α : α ∈ Λ} is norming for Y , consequently (cid:107)S(cid:107) = QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS 5 On the other hand for α (cid:54)= α0 we obtain (cid:107)S∗(y∗ α)(cid:107) (cid:54) 1 + ρ((cid:13)(cid:13)z∗ − T ∗(y∗ (cid:107)S(cid:107) =(cid:13)(cid:13)S∗(y∗ α0)(cid:13)(cid:13) = (1 + η)(cid:107)z∗(cid:107) = y∗ α0)(cid:13)(cid:13) + η (cid:107)z∗(cid:107)) < 1 + ρ(2k + η) = 1 + η. Therefore, So, we have (cid:107)S(cid:107) = (cid:107)S(z)(cid:107) = 1 + η. Also, (cid:107)S − T(cid:107) (cid:54) η + (cid:107)z∗ − T ∗(y∗ α0(S(z)) (cid:54) (cid:107)S(z)(cid:107) (cid:54) (cid:107)S(cid:107) . α0)(cid:107) < η + 2k. Define F := S(cid:107)S(cid:107). Then (cid:107)F(cid:107) = (cid:107)F (z)(cid:107) = 1 and (cid:107)S − F(cid:107) = (cid:107)S(cid:107)(1 − 1 1+η ) = η. So, (cid:107)T − F(cid:107) < 2k + 2η. Therefore, we have that Let us substitute k = . Then we obtain (cid:113) ε 2 · 1−ρ 1+ρ (cid:16) (cid:107)z − x(cid:107) < ε/k and (cid:107)T − F(cid:107) < 2k here we need ε (cid:54) 2(1−ρ) √ max{(cid:107)z − x(cid:107),(cid:107)T − F(cid:107)} < 2ε . 1 + ρ 1 − ρ (cid:17) 1+ρ to have k ∈ [ε/2, 1) (cid:114) 1 + ρ 1 − ρ . Finally, if ε > 2(1−ρ) F(cid:107)} (cid:54) 2. 1+ρ , we can use the triangle inequality to get the evident estimate max{(cid:107)z − x(cid:107),(cid:107)T − (cid:3) Our next goal is to give an improvement for a uniformly non-square domain space X. We recall that uniformly non-square spaces were introduced by James [9] as those spaces whose two-dimensional subspaces are uniformly separated (in the sense of Banach-Mazur distance) from (cid:96)(2) 1 . A Banach space X is uniformly non-square if and only if there is α > 0 such that for all x, y ∈ BX. The parameter of uniform non-squareness of X, which we denote α(X), is the best possible value of α in the above inequality. In other words, ((cid:107)x + y(cid:107) + (cid:107)x − y(cid:107)) (cid:54) 2 − α 1 2 (cid:26) 1 2 α(X) := 2 − sup x,y∈BX ((cid:107)x + y(cid:107) + (cid:107)x − y(cid:107)) . In [6, Theorem 3.3] it was proved that for a uniformly non-square space X with the parameter of uniform non-squareness α(X) > α0 > 0 (cid:114) X (ε) (cid:54) √ ΦS 1 2 To obtain this fact the authors proved the following technical result. Lemma 2.5. Let X be a Banach space with α(X) > α0. Then for every x ∈ SX , y ∈ X and every k ∈ (0, 1 1 − 1 3 − 1 6 ε ∈ for α0 α0 2ε 0, . (cid:18) (cid:19) (cid:27) (cid:19) 2 ] if (cid:107)x − y(cid:107) (cid:54) k then (cid:110) (cid:13)(cid:13)(cid:13)(cid:13)x − y (cid:107)y(cid:107) (cid:18) 1 − 1 3 (cid:13)(cid:13)(cid:13)(cid:13) (cid:54) 2k (cid:115) 1−ρ (1 − 1/3α0) (cid:18) 1 − 1 3 α0 . α0 (cid:111) (cid:19)(cid:114) 1 + ρ 1 − ρ . (2.5) ΦS(X, Y, ε) (cid:54) 2ε Theorem 2.6. Let X and Y be Banach spaces such that β(Y ) (cid:54) ρ, X is uniformly non-square with α(X) > α0, and ε0 = min . Then for any 0 < ε < ε0 1+ρ 2 1−ρ 1+ρ , 1 2 (1−1/3α0) 6 KADETS AND SOLOVIOVA Before proving the theorem, we need a preliminary result. Lemma 2.7. Let X be a Banach space with α(X) > α0. Then for every 0 < ε < 1 and for every (x, x∗) ∈ SX × BX∗ with x∗(x) > 1 − ε, and for every k ∈ [ 2 ] there is (y, y∗) ∈ Π(X) (cid:19) such that 2(1−1/3α0) , 1 (cid:18) ε (cid:107)x − y(cid:107) < and (cid:107)x∗ − y∗(cid:107) < 2k ε k 1 − 1 3 α0 . Proof. The reasoning is almost the same as in Lemma 2.3. We have that x∗ and we can apply Proposition 2.2 for every k ∈ (0, 1/2]. Let us take (cid:107)x∗(cid:107) (x) > 1−η for η = 1− 1−ε(cid:107)x∗(cid:107) k = k((cid:107)x∗(cid:107) − (1 − ε)) ε(cid:107)x∗(cid:107) . (cid:16) 1 ε − (1−ε) ε (cid:17) The inequality (cid:107)x∗(cid:107) (cid:62) x∗(x) > 1 − ε implies that k > 0. On the other hand, k = k k = k < 1/2, so for this k we can find ζ∗ ∈ X∗ and z ∈ SX such that (cid:16) 1 ε − (1−ε) ε(cid:107)x∗(cid:107) (cid:17) (cid:54) ζ∗(z) = (cid:107)ζ∗(cid:107) , Consider z∗ = ζ∗ (cid:107)ζ∗(cid:107). According to Lemma 2.5 , η k (cid:107)x − z(cid:107) < (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗(cid:107) − ζ∗(cid:13)(cid:13)(cid:13)(cid:13) < k. (cid:107)x∗(cid:107) − z∗(cid:13)(cid:13)(cid:13)(cid:13) < 2k (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:18) (cid:19) (cid:107)x∗(cid:107) − z∗(cid:13)(cid:13)(cid:13)(cid:13) + (cid:13)(cid:13)(cid:13)(cid:13) (cid:54) (cid:107)x∗(cid:107) (cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:18)(cid:13)(cid:13)(cid:13)(cid:13) x∗ (cid:107)x∗(cid:107) − z∗ (cid:12)(cid:12)(cid:12)(cid:12)(cid:19) (cid:12)(cid:12)(cid:12)(cid:12)1 − 1 (cid:18) (cid:107)x∗(cid:107) 2k(1 − 1/3α0) + 1 − 1 3 (cid:107)x∗(cid:107) α0 . (cid:13)(cid:13)(cid:13)(cid:13)z∗ − z∗ (cid:107)x∗(cid:107) (cid:13)(cid:13)(cid:13)(cid:13)(cid:19) (1 − 1/3α0) + 1 − (cid:107)x∗(cid:107) (cid:54) 2k(1 − 1/3α0). = (cid:107)x∗(cid:107) = 2 k((cid:107)x∗(cid:107) − (1 − ε)) ε Then (cid:107)x − z(cid:107) < ε/k and (cid:107)x∗ − z∗(cid:107) = (cid:107)x∗(cid:107) · The last inequality holds, because if we consider the function with t ∈ (1 − ε, 1], then f(cid:48) (cid:62) 0 if k (cid:62) f (t) = 2k(1 − 1/3α0) · (t − (1 − ε)) 2(1−1/3α0), so max f = f (1) = 2k(cid:0)1 − 1 + 1 − t ε ε 3 α0 (cid:1). (cid:3) Proof of Theorem 2.6. The proof is a minor modification of the one given for Theorem 2.1. In order to get (2.5) for ε < ε0 we consider T ∈ SL(X,Y ) and x ∈ SX such that (cid:107)T (x)(cid:107) > 1 − ε. α0(T (x)) > 1 − ε. By Lemma 2.7, for 2 ] and for any ε > 0 there exist z∗ ∈ SX∗ and z ∈ SX such that z∗(z) = 1, Since Y has the property β, there is α0 ∈ Λ such that y∗ any k ∈ [ (cid:107)z − x(cid:107) < ε/k and (cid:107)z∗ − T ∗(y∗ α0)(cid:107) < 2k(1 − 1/3α0). 2(1−1/3α0) , 1 ε For η = 2k(1 − 1/3α0) ρ 1−ρ we define S ∈ L(X, Y ) by the formula (2.4) and take F := S(cid:107)S(cid:107). By the same argumentation as before, we have that (cid:107)x − z(cid:107) < ε/k and (cid:107)T − F(cid:107) < 2k (cid:18) (cid:19) 1 + ρ 1 − ρ . 1 − 1 3 α0 Let us substitute k = 1+ρ (here we need ε < ε0). Then we obtain that QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS (cid:113) ε 2(1−1/3α0) · 1−ρ max{(cid:107)z − x(cid:107),(cid:107)T − F(cid:107)} < (cid:115) (cid:18) (cid:19)(cid:114) 1 + ρ 1 − ρ . 2ε 1 − 1 3 α0 7 (cid:3) 3. ESTIMATION FROM BELOW 3.1. Improvement for Φ((cid:96)(2) 1 , Y, ε). We tried our best, but unfortunately we could not find an example demonstrating the sharpness of (2.1) in Theorem 2.1. So, our goal is less ambitious. We are going to present examples of pairs (X, Y ) in which the estimation of Φ(X, Y, ε) from below is reasonably close to the estimation from above given in (2.1). Theorem 2.6 shows that in order to check the sharpness of Theorem 2.1 one has to try those domain spaces X that are not uniformly non-square. The simplest of them is X = (cid:96)(2) 1 . In [5, Example 2.5] this space worked perfectly for the Bishop-Phelps-Bollob´as modulus for functionals. Nevertheless, this is not so when one deals with the Bishop-Phelps-Bollob´as modulus for operators. Namely, the following theorem demonstrates that for X = (cid:96)(2) 1 Theorem 3.1. Let Y be Banach spaces and β(Y ) (cid:54) ρ. Then the estimation given in Theorem 2.1 can be improved. (cid:40)√ (cid:112)1 − ρ2 + ε 2 ρ2 + ρ(cid:112) ε 2 1 + ρ 2ε (cid:41) , 1 . (3.1) ΦS((cid:96)(2) 1 , Y, ε) (cid:54) Φ((cid:96)(2) 1 , Y, ε) (cid:54) min To prove this theorem we need a preliminary result. Lemma 3.2. Let Y be Banach space such that β(Y ) (cid:54) ρ, y ∈ BY , {yα : α ∈ Λ} ⊂ SY , {y∗ Λ} ⊂ S∗ Then there is z ∈ SY such that Y be the sets from Definition 1.3. For given r ∈ (0, 1), α0 ∈ Λ suppose that y∗ α : α ∈ α0(y) (cid:62) 1 − r. α0(yα0) = 1. Let Proof. Suppose that y∗ us check the properties (i)-(iii) for α0(z) = 1; α(z) (cid:54) 1 for all α ∈ Λ; (i) y∗ (ii) y∗ (iii) (cid:107)y − z(cid:107) (cid:54) r(1+ρ) 1−ρ+ρr . α0(y) = 1 − r0, r0 ∈ [0, r]. According to (i) of Definition 1.3 y∗ (cid:16) 1 − ρ + ρr0 (cid:18) (cid:19) 1 − z := r0ρ r0 + 1−ρ+ρr0 α0(z) = 1 − r0ρ (i) y∗ (ii) For every α (cid:54)= α0 we have y∗ (iii) As {y∗ α : α ∈ Λ} ⊂ S∗ α0(y − z) (cid:54) r, and for every α (cid:54)= α0 we have y∗ y. yα0 + 1 − ρ + ρr0 (cid:17) 1−ρ+ρr0 α(z) (cid:54) = 1; Y is a 1-norming subset, so (cid:107)y − z(cid:107) = sup α∈Λ (1 − r0) = 1; · ρ + 1 − r0ρ 1−ρ+ρr0 1−ρ+ρr0 (cid:16) r0 r0 (cid:17) (cid:12)(cid:12)(cid:12)(cid:12) (cid:54) r0(1 + ρ) 1 − ρ + ρr0 α(y) − y∗ r0 1 − ρ + ρr0 y∗ α(yα0) r0 (cid:111) 1 − ρ + ρr0 r, r(1+ρ) 1−ρ+ρr = r(1+ρ) 1−ρ+ρr . (cid:12)(cid:12)(cid:12)(cid:12) (cid:110) α(y − z) = y∗ So, (cid:107)y − z(cid:107) (cid:54) max y∗ α(y − z). Notice that (cid:54) r(1 + ρ) 1 − ρ + ρr . 8 KADETS AND SOLOVIOVA Finally, (i) and (ii) imply that z ∈ SY . Remark 3.3. For every operator T ∈ L((cid:96)(2) 1 , Y ) Moreover, if the operator T ∈ L((cid:96)(2) which does not coincide neither with ±e1, nor with ±e2, then either the segment [T (e1), T (e2)], or [T (e1),−T (e2)] has to lie on the sphere (cid:107)T(cid:107)SY . (cid:96)(2) 1 (cid:107)T(cid:107) = max{(cid:107)T (e1)(cid:107) ,(cid:107)T (e1)(cid:107)}. 1 , Y ) attains its norm in some point x ∈ S (cid:3) According to (3.2) we can apply Lemma 3.2 for the points T (e1) and T (e2) with r = ε(cid:48) are z1, z2 ∈ SY such that y∗ δ < 1. So, there α0(z2) = 1 and α0(z1) = y∗ max{(cid:107)T (e1) − z1(cid:107) ,(cid:107)T (e2) − z2(cid:107)} (cid:54) ε(cid:48) δ (1 + ρ) 1 − ρ + ρ ε(cid:48) . δ Denote y := x/t ∈ S (cid:96)(2) 1 and define F as follows: F (e1) := z1, F (e2) := z2. Proof of Theorem 3.1. Let us denote A(ρ, ε) := √ as a function of ρ, in particular 2ε = A(0, ε) (cid:54) A(ρ, ε) (cid:54) A(1, ε) = 2. 2 We are going to demonstrate that for every pair (x, T ) ∈ Πε((cid:96)(2) 1 , Y ) with Π((cid:96)(2) max{(cid:107)x − y(cid:107) ,(cid:107)T − F(cid:107)} (cid:54) min{A(ρ, ε), 1}. √ √ 2ε 1+ρ 1−ρ2+ ε 2 ρ2+ρ √ ε . Notice that A(ρ, ε) is increasing 1 , Y ) there exists a pair (y, F ) ∈ Without loss of generality suppose that x = (t(1 − δ), tδ), δ ∈ [0, 1/2], t ∈ [1 − ε, 1]. Evidently, (cid:107)x(cid:107) = t. First, we make sure that Φ((cid:96)(2) 1 , Y, ε) (cid:54) 1. Indeed, we can always approximate (x, T ) by the pair y := e1 and F determined by formula F (ei) := T (ei)/(cid:107)T (ei)(cid:107). Then (cid:107)x − e1(cid:107) = 2tδ + 1 − t (cid:54) 1 and (cid:107)T − F(cid:107) (cid:54) 1. 1 , Y, ε) (cid:54) A(ρ, ε), when A(ρ, ε) < 1. As A(ρ, ε) (cid:62) √ ∈ (0, ε). Therefore, It remains to show that Φ((cid:96)(2) consider ε ∈ (0, 1/2). Since Y has the property β, we can select an α0 such that y∗ Without loss of generality we can assume y∗ ε(cid:48) = ε−(1−t) (cid:0)T(cid:0) x α0(T (e2)) > 1 − ε(cid:48) y∗ (3.2) We are searching for an approximation of (x, T ) by a pair (y, F ) ∈ Π((cid:96)(2) cases: α0(T (e1)) > 1 − ε(cid:48) y∗ 1 − δ α0(T (x)) > 1 − ε. Then y∗ 1 , Y ). Let us consider two Case I: 2tδ + 1 − t (cid:54) A(ρ, ε). In this case we approximate (x, T ) by the vector y := e1 and the 2ε we must α0(T (x)) > 1 − ε. (cid:1)(cid:1) > 1 − ε(cid:48), where and . δ α0 t t operator F such that F (e1) := T (e1) (cid:107)T (e1)(cid:107) , F (e2) := T (e2). Then (cid:107)x − y(cid:107) (cid:54) 2tδ + 1 − t (cid:54) A(ρ, ε) and (cid:107)T − F(cid:107) (cid:54) 1 − (cid:107)T (e1)(cid:107) (cid:54) ε 1 − δ √ Case II: 2tδ + 1 − t > A(ρ, ε). Remark, that in this case 2tδ + 1 − t > we use that A(ρ, ε) (cid:62) √ 2ε, ε ∈ (0, 1/2) and t ∈ (0, 1]), (cid:54) 2ε (cid:54) A(ρ, ε). 2ε, and consequently (here √ 2ε − (1 − t) δ > 2t (cid:62) ε(cid:48). QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS 9 Then (cid:107)F(cid:107) = 1, (cid:107)F (y)(cid:107) (cid:62) y∗ α0(F y) = 1, so F attains its norm in y and (cid:107)T − F(cid:107) (cid:54) ε(cid:48) δ (1 + ρ) 1 − ρ + ρ ε(cid:48) . δ So, in this case (cid:107)x − y(cid:107) (cid:54) ε (cid:54) A(ρ, ε) and (cid:107)T − F(cid:107) (cid:54) (1 + ρ) ε−1+t 1 − ρ + ρ ε−1+t tδ tδ . To prove our statement we must show that if 2tδ + 1 − t > A(ρ, ε), then (1+ρ) ε−1+t 1−ρ+ρ ε−1+t tδ (cid:54) A(ρ, ε). Let us denote f (t, δ) = 2tδ + 1 − t and g(t, δ) = demonstrate that (1+ρ) ε−1+t 1−ρ+ρ ε−1+t tδ tδ = (1+ρ)(ε−1+t) (1−ρ)tδ+ρ(ε−1+t). So, we need to tδ (3.3) min{f (t, δ), g(t, δ)} (cid:54) A(ρ, ε) for all δ ∈ [0, 1/2] and for all t ∈ [1 − ε, 1]. Notice that for every fixed t ∈ [1 − ε, 1] the function f (t, δ) is increasing as δ increases and g(t, δ) is decreasing as δ increases. So, if we find δ0 such that f (t, δ) = g(t, δ), then min{f (t, δ), g(t, δ)} (cid:54) f (t, δ0). If we denote u = f (t, δ) = 2tδ + 1 − t the equation f (t, δ) = g(t, δ) transforms to (3.4) u = 2 − 2(1 − ρ)(u − ε) (t − 1 + ε)(1 + ρ) + (u − ε)(1 − ρ) . The right-hand side of this equation is increasing as t increases, so the positive solution of the equation (3.4) ut is also increasing. This means that we obtain the greatest possible solution, if we substitute t = 1. Then we get the equation From here u = A(ρ, ε), and so, the inequality (3.3) holds. u2 1 + ρ 2 + uρε − ε(1 + ρ) = 0. (cid:3) 1 , Y, ε). So, if X = (cid:96)(2) 3.2. Estimation from below for ΦS((cid:96)(2) 1 , the estimation for the Bishop-Phelps- Bollob´as modulus is somehow better than in Theorem 2.1. Nevertheless, considering (cid:96)(2) 1 we can obtain some interesting estimations from below for ΦS((cid:96)(2) 1 , Y, ε). Notice that the estimations (2.1) and (3.1) give the same asymptotic behavior when ε is convergent to 0. Our next proposition gives the estimation for ΦS((cid:96)(2) Theorem 3.4. For every Banach space Y 1 , Y, ε) from below, when β(Y ) = 0. In particular, ΦS((cid:96)(2) 1 , Y, ε) = min{√ 1 , Y, ε) (cid:62) min{ ΦS((cid:96)(2) 2ε, 1} if β(Y ) = 0. √ 2ε, 1}. Proof. To prove our statement we must show that ΦS((cid:96)(2) inequality ΦS((cid:96)(2) every ε ∈ (0, 1/2) and for every δ > 0 we are looking for a pair (x, T ) ∈ ΠS 1 , Y, ε) (cid:62) 1 for ε > 1/2 will follow from the monotonicity of ΦS((cid:96)(2) 2ε for ε ∈ (0, 1/2). The remaining 1 , Y,·). So, for ε ((cid:96)(2) 1 , Y ) such that 1 , Y, ε) (cid:62) √ max{(cid:107)x − y(cid:107) ,(cid:107)T − F(cid:107)} (cid:62) √ 2ε − δ. 10 for every pair (y, F ) ∈ Π((cid:96)(2) following operator T ∈ S L((cid:96)(2) : 1 ,Y ) KADETS AND SOLOVIOVA 1 , Y ). Fix ξ ∈ SY and ε0 < ε such that √ 2ε0 > √ 2ε − δ. Consider the T (z1, z2) = (z1 + (1 − √ 2ε0)z2)ξ and take x = (1 −(cid:112)ε0/2,(cid:112)ε0/2) ∈ S . Then (cid:107)T (x)(cid:107) = 1 − ε0 > 1 − ε. To approximate the pair (x, T ) by a pair (y, F ) ∈ Π((cid:96)(2) or F attains its norm in a point that belongs to conv{e1, e2}, and so attains its norm in both points e1, e2. In √ √ 2ε − δ. In the second case the first case we are forced to have y = (1, 0), and then (cid:107)x − y(cid:107) = √ √ 2ε0 > we have (cid:107)F − T(cid:107) (cid:62) (cid:107)F (e2) − T (e2)(cid:107) (cid:62) (cid:107)F (e2)(cid:107) − (cid:107)T (e2)(cid:107) = 2ε − δ. (cid:3) 2ε0 > 1 , Y ) we have two possibilities: either y is an extreme point of B (cid:96)(2) 1 (cid:96)(2) 1 Our next goal is to estimate the spherical Bishop-Phelps-Bollob´as modulus from below for the values 2 , 1) and denote Yρ the linear space R2 equipped with the of parameter ρ between 1/2 and 1. Fix a ρ ∈ [ 1 norm (cid:26) (cid:18) (cid:19) (cid:18) (cid:19) x1 + 2 − 1 ρ x2,x2 + 2 − 1 ρ x1,x1 − x2 . (cid:27) (3.5) (cid:107)x(cid:107)ρ = max In other words,  (cid:16) (cid:16) x1 − x2, x1 + x2 + 2 − 1 2 − 1 ρ ρ (cid:17) (cid:17) x2, x1, if x1x2 (cid:54) 0; if x1x2 > 0 and x1 > x2; if x1x2 > 0 and x1 (cid:54) x2. (cid:107)(x1, x2)(cid:107) = and the unit ball Bρ of Xρ is the hexagon absdef, where a = (1, 0); b = ( (−1, 0); e = (− ρ 3ρ−1 ); and f = (0,−1). 3ρ−1 , ρ ρ 3ρ−1 , ρ 3ρ−1 ); c = (0, 1); d = The dual space to Yρ is R2 equipped with the polar to Bρ as its unit ball. So, the norm on Y ∗ = Y ∗ ρ is given by the formula (cid:26) (cid:27) x1 + x2 , x1,x2, ρ 3ρ − 1 (cid:107)x(cid:107)∗ ρ = (cid:107)(x1, x2)(cid:107)∗ = max (cid:17) (cid:17) ); e∗ = (−(cid:16) 2 − 1 ρ ρ and the unit ball B∗ (−1, 1); d∗ = (−1,−(cid:16) ρ of Y ∗ ρ is the hexagon a∗b∗c∗d∗e∗f∗, where a∗ = (1, 2 − 1 2 − 1 ; c∗ = ,−1); and f∗ = (1,−1). The corresponding spheres ρ ); b∗ = 2 − 1 ρ , 1 Sρ and S∗ ρ are shown on Figures 1 and 2 respectively. (cid:16) (cid:17) QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS 11 6 c y2 b y1 a - y3 d c∗ = y∗ 3 6 b∗ = y∗ 2 e f d∗ e∗ Figure 1 Figure 2 Proposition 3.5. In the space Y = Yρ β(Y ) (cid:54) ρ. a∗ = y∗ 1 - f∗ (cid:19) (cid:18) (cid:19)(cid:27) Proof. Consider two sets: (cid:26) (cid:18) 2ρ2 (cid:18) ρ − ρ2 , (cid:19) ρ , 1), y∗ , − 1 2 , 1 2 , y3 = ⊂ SYρ 3ρ − 1 and {y∗ y1 = 1 = (1, 2 − 1 ρ − ρ2 2ρ2 3ρ − 1 3ρ − 1 3ρ − 1 , y2 = 2 = (2 − 1 3 = (−1, 1)} ⊂ SY ∗ ρ ), y∗ ρ . n(yn) = 1 for n = 1, 2, 3 and y∗ n(y) : n = 1, 2, 3} for all y ∈ Yρ, y∗ Then (cid:107)y(cid:107) = sup{y∗ i (yj) (cid:54) ρ 1(y1) = 2ρ2+2ρ−2ρ2−1+ρ 1(y2) = ρ−ρ2+4ρ2−2ρ for all i (cid:54)= j. Indeed, y∗ y∗ y∗ = 1; 1(y3) = 2ρ = − 1−ρ 2 + 1 − 1 (cid:62) −ρ, consequently y∗ 1(y3) (cid:54) ρ (here appears the restriction ρ (cid:62) 1/2); −1 2ρ 2(y3) = −y∗ y∗ y∗ y∗ 2(y2) = y∗ 2(y1) = y∗ 1(y3) (cid:54) ρ; 1(y2) = ρ; y∗ 3(y1) = 3(y1) = ρ; and finally y∗ (cid:3) (cid:27) Theorem 3.6. Let ρ ∈ [1/2, 1), 0 < ε < 1. Then, in the space Y = Yρ (cid:12)(cid:12)(cid:12)−2ρ2+ρ−ρ2 (cid:12)(cid:12)(cid:12) = ρ; y∗ 3(y2) = −y∗ 1(y1) = 1; 3(y3) = 1 2 + 1 2 = 1. = ρ; 3ρ−1 3ρ−1 3ρ−1 ΦS((cid:96)(2) 1 , Y, ε) (cid:62) min Proof. To prove our statement we must show that ΦS((cid:96)(2) remaining inequality ΦS((cid:96)(2) (cid:17) 1 , Y, ε) (cid:62) 1 for ε (cid:62) 1−ρ and for every δ > 0 we are looking for a pair (x, T ) ∈ ΠS 0, 1−ρ 2ρ will follow from the monotonicity of ΦS((cid:96)(2) So, for every ε ∈(cid:16) 0, 1−ρ 2ρ ε ((cid:96)(2) 2ρ 1−ρ for ε ∈ (cid:16) (cid:17) . The 1 , Y,·). 1 , Y ) such that . , 1 1 − ρ (cid:26)(cid:114) 2ρε 1 , Y, ε) (cid:62) (cid:113) 2ρε (cid:114) 2ρε − δ 1 − ρ max{(cid:107)x − y(cid:107) ,(cid:107)T − F(cid:107)} (cid:62) for every pair (y, F ) ∈ Π((cid:96)(2) 1 , Y ). 12 Fix an ε0 < ε such that (cid:113) 2ρε0 1−ρ > x = and T ∈ L((cid:96)(2) 1 , Y ) such that KADETS AND SOLOVIOVA (cid:113) 2ρε (cid:19) (cid:18) 1−ρ − δ. Consider the point ∈ S 1 − (cid:114) 2ρε0 (cid:114) 2ρε0 √ √ 2ρε0 1 − ρ √ √ 2ρε0 1 − ρ (cid:18) 2 2 , (cid:19) 1 − (cid:96)(2) 1 · b, 1 − ρ T (ei) = 1 − ρ ei + (cid:16) ρ (cid:17) where b = (cid:107)T (e2)(cid:107) = 1 and (cid:107)T (x)(cid:107) = 1 − ε0 > 1 − ε. 3ρ−1 , 3ρ−1 ρ is the extreme point of SY from Figure 1. Notice that (cid:107)T(cid:107) = (cid:107)T (e1)(cid:107) = (cid:113) 2ρε0 (cid:96)(2) 1 The part of S consisting of points that have a distance to x less than or equal to 1−ρ lies on the segment [e1, e2). Consequently, in order to approximate the pair (x, T ) we have two options: to approximate the point x by e1, and then we can take F := T ; or as F choose an operator attaining its norm in some point of (e1, e2) (and hence in all points of [e1, e2]), and then we can take y := x. (cid:113) 2ρε0 1−ρ > (cid:113) 2ρε 1−ρ − δ. In the second case let us (cid:114) 2ρε0 In the first case we have (cid:107)T − F(cid:107) = 0 and (cid:107)x − y(cid:107) = demonstrate that (cid:107)T − F(cid:107) = max i If it is not so, then for both values of i = 1, 2 (cid:107)T (ei) − F (ei)(cid:107) (cid:62) 1 − ρ . (cid:107)T (ei) − F (ei)(cid:107) < = (cid:107)T (ei) − b(cid:107). (cid:114) 2ρε0 1 − ρ Since F attains its norm in all points of [e1, e2], the line segment F ([e1, e2]) should lie on a line segment of SY , but the previous inequality makes this impossible, because T (e1) and T (e2) lie on different line (cid:3) segments of SY with b being their only common point. 3.3. Non-continuity of the Bishop-Phelps-Bollob´as modulus for operators. It is known [7, Theorem 3.3] that both (usual and spherical) Bishop-Phelps-Bollob´as moduli for functionals are continuous with respect to X. As a consequence of Theorem 3.6 we will obtain that the Bishop-Phelps-Bollob´as moduli of a pair (X, Y ) as a function of Y are not continuous in the sense of Banach-Mazur distance. Let X and Y be isomorphic. Recall that the Banach-Mazur distance between X and Y is the following quantity d(X, Y ) = inf{(cid:107)T(cid:107)(cid:13)(cid:13)T −1(cid:13)(cid:13) : T : X → Y isomorphism.} A sequence Zn of Banach spaces is said to be convergent to a Banach space Z if d(Zn, Z) −→ Notice, that Yρ −→ ρ→1 (cid:96)(2) 1 . n→∞ 1. Theorem 3.7. Let ρ ∈ [1/2, 1) and Yρ be the spaces defined in (3.5). Then for every ε ∈ (0, 1 2 ) Φ((cid:96)(2) 1 , Yρ, ε) (cid:54)−→ ρ→1 Φ((cid:96)(2) 1 , (cid:96)(2) 1 , ε), and ΦS((cid:96)(2) 1 , Yρ, ε) (cid:54)−→ ρ→1 ΦS((cid:96)(2) 1 , (cid:96)(2) 1 , ε). Proof. On the one hand, from the Theorem 2.1 with ρ = 0 we get for ε ∈ (0, 1 2 ) ΦS((cid:96)(2) 1 , (cid:96)(2) 1 , ε) (cid:54) Φ((cid:96)(2) 1 , (cid:96)(2) 1 , ε) (cid:54) √ 2ε < 1. QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS On the other hand, Theorem 3.6 gives Φ((cid:96)(2) 1 , Yρ, ε) (cid:62) ΦS((cid:96)(2) 1 , Yρ, ε) (cid:62) min (cid:110)(cid:113) 2ρε 1−ρ , 1 (cid:111) −−−→ ρ→1 13 1. (cid:3) 3.4. Behavior of the ΦS(X, Y, ε) when ε → 0. In subsection 3.2 using two-dimensional spaces Y we were able to give the estimation only for ρ ∈ [1/2, 1). This is not surprising, because in every n- dimensional Banach space with the property β we have either ρ = 0, or ρ (cid:62) 1 n. We did not find any mentioning of this in literature, so we give the proof of this fact. Proposition 3.8. Let Y (n) be a Banach space of dimension n with β(Y (n)) (cid:54) ρ < 1 isometric to (cid:96)(n)∞ , i.e. β(Y (n)) = 0. n. Then Y (n) is We need one preliminary result. α : α ∈ Λ} ⊂ S∗ Y be the sets from Definition 1.3. Then Λ = n. Lemma 3.9. Let Y (n) be a Banach space of dimension n with β(Y (n)) (cid:54) ρ < 1 SY , {y∗ Proof. Λ (cid:62) n, because {y∗ demonstrate that every subset of {yα : α ∈ Λ} consisting of n + 1 elements is linearly independent. α : α ∈ Λ} is a 1-norming subset. Assume that Λ > n. We are going to i=1 ⊂ {yα : α ∈ Λ}. Consider the corresponding aiyi (cid:54)= 0. Let j (cid:54) n + 1 be a Without loss of generality we can take a subset {yi}n+1 aiyi with maxai = 1 and let us check that n and {yα : α ∈ Λ} ⊂ linear combination number such that aj = 1. Then we can estimate aiyi aiyi j (yj) + aiy∗ j (yi) aiρ > 0. i=1 n+1(cid:80) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:62) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)n+1(cid:88) i=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)y∗ j (cid:32)n+1(cid:88) i=1 (cid:33)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)ajy∗ n+1(cid:88) i=1 i(cid:54)=j i=1 n+1(cid:80) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:62) 1 − n+1(cid:88) i=1 i(cid:54)=j It follows that Y (n) contains n + 1 linearly independent elements. This contradiction completes the (cid:3) proof. Proof of Proposition 3.8. According to Definition 1.3 together with Lemma 3.9 there are two sets {yi}n i=1 ⊂ i }n i=1 ⊂ S∗ SY (n), {y∗ y∗ i (yi) = 1, y∗ i (yj) < 1/n if i (cid:54)= j, (cid:107)y(cid:107) = sup{y∗ i (y) : i = 1..n}, for all y ∈ Y. Y (n) such that Let us define the operator U : Y (n) → (cid:96)(n)∞ by the formula: U (y) := (y∗ 1(y), y∗ 2(y), ..., y∗ n(y)). Obviously, (cid:107)U (y)(cid:107) = (cid:107)y(cid:107) for all y ∈ Y (n), so, U is isometry. Since dim Y (n) = dim (cid:96)(n)∞ , the operator U is bijective. This means that Y (n) is isometric to (cid:96)(n)∞ , and since β((cid:96)(n)∞ ) = 0, we have that β(Y (n)) = (cid:3) 0. So, in order to obtain all possible values of parameter ρ we must consider spaces of higher dimensions. the linear space Rn equipped with the n , 1) and denote Z = Z(n) For every fixed dimension n fix a ρ ∈ [ 1 norm ρ (cid:40) (3.6) (cid:107)x(cid:107) = max x1,x2, ...,xn, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n(cid:88) i=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:41) xi . 1 ρn 14 KADETS AND SOLOVIOVA Proposition 3.10. Let Z = Z(n) ρ with n (cid:62) 2 and ρ ∈ [ 1 β(Z) (cid:54) ρ. n , 1). Then Proof. Consider two sets:yj = − n(cid:88) i=1 i(cid:54)=j 1 n − 1 + ρn (cid:40) y∗ j = ej, z∗ =  ⊂ SZ n(cid:88) i=1 ei ei + ej, z = ρ (cid:41) n(cid:88) i=1, z∗(cid:111) i=1 ei 1 ρn (cid:110){y∗ ⊂ SZ∗. j}n j (z) = ρ, z∗(z) = 1, y∗ is 1-norming. Also, z∗(yi) = √ It follows directly from (3.6) that the subset n−1+ρn (cid:54) ρ, y∗ i (yi) = 1, j (yi) = − y∗ 1 1 n−1+ρn (cid:54) ρ. (cid:3) Remark that in all our estimations of ΦS(X, Y, ε) appears the multiplier 2ε. So, in order to measure the behavior of ΦS(X, Y, ε) in 0, it is natural to introduce the following quantity Also define Ψ(X, Y ) := lim sup ε→0 ΦS(X, Y, ε) √ 2ε . Ψ(ρ) := sup Y :β(Y )=ρ sup X lim sup Ψ(X, Y ) ε→0 which measures the worst possible behavior in 0 of ΦS(X, Y, ε) when β(Y ) (cid:54) ρ. From Theorem 2.1 we know that (cid:114) 1 + ρ 1 − ρ . Ψ(ρ) (cid:54) Now we will estimate Ψ(ρ) from below. Theorem 3.11. Ψ(ρ) (cid:62) min (cid:26)(cid:114) 2ρ 1 − ρ (cid:27) , 1 for all values of ρ ∈ (0, 1). Proof. From Theorem 3.4 we know that Ψ(ρ) (cid:62) 1. So, we have to check that Ψ(ρ) (cid:62) (cid:113) 2ρ 1−ρ. In ρ ). Denote ei and Γ = {x ∈ SZ : z∗(x) = 1}. Consider the point x = (1 − δ, δ) and the operator T order to estimate Ψ(ρ) from below for small ε we consider the couple of spaces ((cid:96)(2) z∗ = 1 ρn such that 1 , Z(n) n(cid:80) i=1 n(cid:88) k(cid:88) n(cid:88) ei, T (e1) = ρ ei and T (e2) = t ei + i=1 i=1 i=k+1 QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS 15 with k = 1 2 n(1 − ρ) + 1 + θ ∈ N being the nearest natural to 1 4 + 4θ − 2nρ ε0 n − nρ + 2 + 2θ t = −1 + δ (3.7) where δ > 0 will be defined later and ε0 < ε. Then z∗(T (x)) = 1 − ε0 > 1 − ε, so (x, T ) ∈ 1 , Z(n) ε ((cid:96)(2) ρ ). ΠS As usual, we have two options: ρ ). Now we are searching for the best approximation of (x, T ) by a pair (y, F ) ∈ Π((cid:96)(2) 1 , Z(n) , 2 n(1 − ρ) + 1 (so, θ (cid:54) 1/2) and I. We can approximate the point x by e1 and then we can take F = T . In this case we get (3.8) (cid:107)x − y(cid:107) = 2δ. II. We can choose F which attains its norm in all points of the segment [e1, e2], and then we can take y = x. In this case F (e1) and F (e2) must lie in the same face. Besides, if F (e1) (cid:54)∈ Γ, we have (cid:107)T (e1) − F (e1)(cid:107) = 1 − ρ > 1−ρ for ε sufficiently small. To obtain better estimation we must have F (e1) ∈ Γ and, so, F (e2) ∈ Γ. Then √ 2ε (cid:113) 2ρ (cid:107)T − F(cid:107) (cid:62) (cid:107)T (e2) − F (e2)(cid:107) (cid:62) inf h∈Γ (cid:107)T (e2) − h(cid:107) . Let us estimate the distance from T (e2) to the face Γ. hi ∈ Γ, then hi (cid:54) 1 and z∗(h) = 1 ρn hi = 1. So, n(cid:80) i=1 k(cid:80) i=1 hi (cid:62) ρn − (n − k), and n(cid:80) If h = i=1 Therefore, (3.9) max hi (cid:62) 1 k (ρn − (n − k)) = −1 + 4 + 4θ n(1 − ρ) + 2 + 2θ . (cid:107)T (e2) − h(cid:107) (cid:62) max 1(cid:54)i(cid:54)k t − hi (cid:62) t − max hi = 2nρ ε0 δ n(1 − ρ) + 2 + 2θ . Now let us define δ as a positive solution of the equation: (cid:113) √ 2nρ ε0 δ 2δ = n(1 − ρ) + 2 + 2θ . (cid:113) √ 2ε Then δ = 1 2 So, with this δ the estimation (3.8) gives us 1−ρ+(2+θ)/n. Denote C(ε, ρ, n, θ) := 2ε0 2ρ 2ρ 1−ρ+(2+θ)/n . and C0 = C(ε0, ρ, n, θ). and the estimation (3.9) gives us (cid:107)x − y(cid:107) = 2δ = C0, (cid:107)T − F(cid:107) (cid:62) 2nρ ε0 δ In that way, we have shown that ΦS((cid:96)(2) we obtain that ΦS((cid:96)(2) Ψ((cid:96)(2) ρ ) (cid:62)(cid:113) 1 , Z(n) 1−ρ+(2+θ)/n 1 , Z(n) 2ρ = C0. n(1 − ρ) + 2 + 2θ ρ , ε) (cid:62) C0. As ε0 can be chosen arbitrarily close to ε ρ , ε) (cid:62) C(ε, ρ, n, θ) with θ ∈ [−1/2, 1/2]. Consequently, we have that (cid:3) . When n → ∞, we obtain the desired estimation Ψ(ρ) (cid:62)(cid:113) 2ρ 1 , Z(n) 1−ρ. 16 KADETS AND SOLOVIOVA 4. MODIFIED BISHOP-PHELPS-BOLLOB ´AS MODULI FOR OPERATORS The following modification of the Bishop-Phelps-Bollob´as theorem can be easily deduced from Propo- sition 2.2 just by substituting η = ε, k = Theorem 4.1 (Modified Bishop-Phelps-Bollob´as theorem). Let X be a Banach space. Suppose x ∈ BX and x∗ ∈ BX∗ satisfy x∗(x) (cid:62) 1−ε (ε ∈ (0, 2)). Then there exists (y, y∗) ∈ SX×X∗ with (cid:107)y∗(cid:107) = y∗(y) such that ε. √ max{(cid:107)x − y(cid:107),(cid:107)x∗ − y∗(cid:107)} (cid:54) √ ε. The improvement in this estimate comparing to the original version appears because we do not demand (cid:107)y∗(cid:107) = 1. It was shown in [11] that this theorem is sharp in a number of two-dimensional spaces, which makes a big difference with the original Bishop-Phelps-Bollob´as theorem, where the only (up to isometry) two-dimensional space, in which the theorem is sharp, is (cid:96)(2) 1 . Bearing in mind this theorem it is natural to introduce the following quantities. Definition 4.2. The modified Phelps-Bollob´as modulus of a pair (X, Y ) is the function, which is deter- mined by the following formula: (cid:101)Φ(X, Y, ε) = inf{δ > 0 : for every T ∈ BL(X,Y ), if x ∈ BX and (cid:107)T (x)(cid:107) > 1 − ε, then there exist y ∈ SX and F ∈ L(X, Y ) satisfying (cid:107)F (y)(cid:107) = (cid:107)F(cid:107),(cid:107)x − y(cid:107) < δ and (cid:107)T − F(cid:107) < δ}. The modified spherical Bishop-Phelps-Bollob´as modulus of a pair (X, Y ) is the function, which is determined by the following formula: (cid:102)ΦS(X, Y, ε) = inf{δ > 0 : for every T ∈ SL(X,Y ), if x ∈ SX and (cid:107)T (x)(cid:107) > 1 − ε, then there exist y ∈ SX and F ∈ L(X, Y ) satisfying (cid:107)T (y)(cid:107) = (cid:107)T(cid:107) ,(cid:107)x − y(cid:107) < δ and (cid:107)T − F(cid:107) < δ}. By analogy with Theorem 2.1 we prove the next result. Theorem 4.3. Let X and Y be Banach spaces such that Y has the property β with parameter ρ. Then the pair (X, Y ) has the Bishop-Phelps-Bollob´as property for operators and for any ε ∈ (0, 1) (cid:102)ΦS(X, Y, ε) (cid:54)(cid:101)Φ(X, Y, ε) (cid:54) min (cid:26)√ ε (cid:114) 1 + ρ 1 − ρ (cid:27) , 1 . (4.1) clearness. The proof is similar to Theorem 2.1 but it has some modifications and we give it here for the sake of Proof. Consider T ∈ BL(X,Y ) and x ∈ BX such that (cid:107)T (x)(cid:107) > 1 − ε with ε ∈ (cid:16) . Since Y has the property β, there is α0 ∈ Λ such that y∗ α0, we have x ∈ BX , x∗ ∈ BX∗ with x∗(x) > 1 − ε. We can apply the formula (2.2) from Lemma 2.3, for any k ∈ (0, 1). For every k ∈ [ε, 1) let us take α0(T (x)) > 1 − ε. So, if we denote x∗ = T ∗y∗ 0, 1−ρ (cid:17) 1+ρ k((cid:107)x∗(cid:107) − (1 − ε)) ε(cid:107)x∗(cid:107) . k = (cid:16) 1 ε − (1−ε) ε (cid:17) The inequality (cid:107)x∗(cid:107) (cid:62) x∗(x) > 1 − ε implies that k > 0. On the other hand, k = k k and z ∈ SX such that z∗(z) = (cid:107)z∗(cid:107) and = k < 1, so for this k we can find ζ∗ ∈ X∗ and z ∈ SX such that there exist z∗ ∈ X∗ (cid:16) 1 ε − (1−ε) ε(cid:107)x∗(cid:107) (cid:17) (cid:54) (cid:107)x − z(cid:107) < and (cid:107)z∗ − x∗(cid:107) < k (cid:107)x∗(cid:107) . 1 − 1−ε(cid:107)x∗(cid:107) k 17 we define the operator S ∈ L(X, Y ) by the QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS For a real number η satisfying η > ρ(k(cid:107)x∗(cid:107)+(cid:107)x∗(cid:107)·1−(cid:107)z∗(cid:107)) (cid:107)z∗(cid:107)(1−ρ) formula Let us estimate the norm of S. Recall that we denote x∗ = T ∗y∗ S(t) = (cid:107)z∗(cid:107) T (t) + [(1 + η)z∗(t) − (cid:107)z∗(cid:107) T ∗(y∗ α0)(t)]yα0. α0. Thus for all y∗ ∈ Y ∗, Since the set {y∗ S∗(y∗) = (cid:107)z∗(cid:107) T ∗(y∗) + [(1 + η)z∗ − (cid:107)z∗(cid:107) x∗]y∗(yα0). α : α ∈ Λ} is norming for Y it follows that (cid:107)S(cid:107) = supα (cid:107)S∗y∗ α(cid:107). α0)(cid:13)(cid:13) = (1 + η)(cid:107)z∗(cid:107) . On the other hand for α (cid:54)= α0 we obtain (cid:107)S∗(y∗ α)(cid:107) (cid:54) (cid:107)z∗(cid:107) + ρ[(cid:107)z∗ − x∗(cid:107) + (cid:107)x∗(cid:107) · 1 − (cid:107)z∗(cid:107) + η(cid:107)z∗(cid:107)] (cid:54) (1 + η)(cid:107)z∗(cid:107) . (cid:107)S(cid:107) (cid:62)(cid:13)(cid:13)S∗(y∗ α0)(cid:13)(cid:13) = (cid:107)z∗(cid:107) = y∗ (cid:107)S(cid:107) =(cid:13)(cid:13)S∗(y∗ α0(S(z)) (cid:54) (cid:107)S(z)(cid:107) (cid:54) (cid:107)S(cid:107) . Therefore, So, we have (cid:107)S(cid:107) = (cid:107)S(z)(cid:107) = (1 + η)(cid:107)z∗(cid:107). Let us estimate (cid:107)S − T(cid:107). (cid:107)S − T(cid:107) = sup α (cid:107)S∗y∗ α − T ∗y∗ α(cid:107) . Notice also that For α = α0 we get (cid:13)(cid:13)S∗y∗ α0 − T ∗y∗ α0 1 − (cid:107)z∗(cid:107) (cid:54) (cid:107)x∗ − z∗(cid:107) + 1 − (cid:107)x∗(cid:107) < k(cid:107)x∗(cid:107) + 1 − (cid:107)x∗(cid:107) . (cid:13)(cid:13) = (cid:107)(1 + η)z∗ − x∗(cid:107) (cid:54) (cid:107)z∗ − x∗(cid:107) + η(cid:107)z∗(cid:107) k(cid:107)x∗(cid:107)(1 + ρ(cid:107)x∗(cid:107)) + ρ(cid:107)x∗(cid:107)(1 − (cid:107)x∗(cid:107)) < 1 − ρ . Then (cid:107)x − z(cid:107) < ε k and (cid:13)(cid:13)S∗y∗ α0 − T ∗y∗ α0 (cid:13)(cid:13) < (cid:107)x∗(cid:107)−(1−ε) ε k (1 + ρ(cid:107)x∗(cid:107)) + ρ(cid:107)x∗(cid:107)(1 − (cid:107)x∗(cid:107)) 1 − ρ (cid:54) k(1 + ρ) 1 − ρ . The last inequality holds, because if we consider the function with t ∈ (1 − ε, 1), then f(cid:48) (cid:62) 0, so max f = f (1) = k t−(1−ε) ε (1 + ρt) + ρt(1 − t) f (t) = 1 − ρ 1−ρ . For α (cid:54)= α0 we obtain k(1+ρ) α(cid:107) (cid:54) 1 − (cid:107)z∗(cid:107) + ρ((cid:107)z∗ − x∗(cid:107) + (cid:107)x∗(cid:107) · 1 − (cid:107)z∗(cid:107) + η(cid:107)z∗(cid:107)) (cid:107)S∗y∗ α − T ∗y∗ Substituting the value of k we get α(cid:107) < α − T ∗y∗ (cid:107)S∗y∗ < k(cid:107)x∗(cid:107) + 1 − (cid:107)x∗(cid:107) + − ρ(cid:107)x∗(cid:107) · 1 − (cid:107)z∗(cid:107) + ρk(cid:107)x∗(cid:107) + ρ(cid:107)x∗(cid:107) · 1 − (cid:107)z∗(cid:107)] (cid:54) k(cid:107)x∗(cid:107) + 1 − (cid:107)x∗(cid:107) + ρ 1 − ρ [k(cid:107)x∗(cid:107) − ρk(cid:107)x∗(cid:107) + (cid:107)x∗(cid:107) · 1 − (cid:107)z∗(cid:107) [k(cid:107)x∗(cid:107) + (cid:107)x∗(cid:107) · (k(cid:107)x∗(cid:107) + 1 − (cid:107)x∗(cid:107))]. ρ 1 − ρ (cid:20) 1 − ρ(1 − (cid:107)x∗(cid:107)) (cid:107)x∗(cid:107) − (1 − ε) k 1 − ρ (cid:107)x∗(cid:107) − (1 − ε) ε ε (cid:54) k(1 + ρ) 1 − ρ . + ρk 1 − ρ (cid:21) + 1 − (cid:107)x∗(cid:107) 18 KADETS AND SOLOVIOVA To get the last inequality we again use the fact that the function (cid:20) k f1(t) = 1 − ρ(1 − t) 1 − ρ (cid:113) ε(1−ρ) 1+ρ is increasing if k (cid:62) ε, so max f1 = f1(1) = Let us substitute k = 1+ρ). Then we obtain ε < 1−ρ (here we need ε < 1+ρ (cid:21) t − (1 − ε) ε + 1 − t ρk 1 − ρ 1−ρ . So, (cid:107)T − S(cid:107) (cid:54) k(1+ρ) 1−ρ . + k(1+ρ) t − (1 − ε) ε 1−ρ which holds for any ε ∈ (0, 1) and also (cid:115) max{(cid:107)z − x(cid:107),(cid:107)T − S(cid:107)} (cid:54) ε(1 + ρ) 1 − ρ . Finally, if ε (cid:62) 1−ρ max{(cid:107)z − x(cid:107),(cid:107)T − S(cid:107)} (cid:54) 1. 1+ρ, we can always approximate (x, T ) by the same point and an zero operator, so (cid:3) The above theorem implies that if β(Y ) = 0, then (cid:102)ΦS(X, Y, ε) (cid:54)(cid:101)Φ(X, Y, ε) (cid:54) √ Theorem 4.4. (cid:102)ΦS((cid:96)(2) 1 , Y = R. demonstrate that this estimation is sharp for X = (cid:96)(2) √ ε, ε ∈ (0, 1). 1 , R, ε) =(cid:101)Φ((cid:96)(2) 1 , R, ε) = ε. We are going to (4.2) Proof. We must show that for every 0 < ε < 1 and for every δ > 0 there is a pair (x, x∗) from ε ((cid:96)(2) ΠS 1 , R) such that for every pair (y, y∗) ∈ S × (cid:96)(2)∞ with y∗(y) = (cid:107)y∗(cid:107) (cid:96)(2) 1 max{(cid:107)x − y(cid:107) ,(cid:107)x∗ − y∗(cid:107)} (cid:62) √ (cid:16) √ 1 − √ ε0)z2. Notice that x∗(x) = 1 − ε0 > 1 − ε. ε0 > √ ε − δ. (cid:17) Fix an ε0 < ε such that √ x∗(z) := z1 + (1 − 2 ε− δ. Take a point x := √ ε0. U is the intersection of SX with the open ε0, so, U ⊂]e1, e2[, ε0, and (cid:107)x − e2(cid:107) = 2−√ e2, and a functional e1 + ε0 2 ε0 2 √ Consider the set U of those y ∈ SX that (cid:107)x − y(cid:107) < ε0 centered in x. As (cid:107)x − e1(cid:107) = √ ball of radius and for every y = ae1 + be2 ∈ U a > 0 and b > 0. Assume that y∗(y) = (cid:107)y∗(cid:107) for some y ∈ U and (cid:107)x∗ − y∗(cid:107) (cid:54) √ y∗ = (y∗(e1), y∗(e2)), where y∗(e1) = y∗(e2) and y∗(e1) · y∗(e2) (cid:62) 0. Notice that ε ⇒ y∗(e1) (cid:62) 1 − √ x∗(e1) − y∗(e1) = 1 − y∗(e1) (cid:54) (cid:107)x∗ − y∗(cid:107) (cid:54) √ ε0 (cid:62) √ ε0, ε0. Then we are forced to have (cid:16)√ (cid:17) x∗(e2) − y∗(e2) = 1 − 2 Then y∗ = (1 − √ as desired. ε0, 1 − √ ε0), so (cid:107)x∗ − y∗(cid:107) = √ ε0 > √ √ ε0 − y∗(e2) (cid:54) (cid:107)x∗ − y∗(cid:107) (cid:54) √ ε0 ⇒ y∗(e2) (cid:54) 1 − √ ε0. ε − δ. It follows that inequality (4.2) holds, (cid:3) Also with the same space Y = Yρ equipped with the norm (3.5) we have an estimation from below which almost coincides with the estimation (4.1) from above for values of ρ close to 1. Theorem 4.5. Let ρ ∈ [1/2, 1), 0 < ε < 1. Then, in the space Y = Yρ (cid:102)ΦS((cid:96)(2) 1 , Y, ε) (cid:62) min (cid:26)√ (cid:114) 2ρ (cid:27) ε 1 − ρ , 1 . QUANTITATIVE VERSION OF THE BISHOP-PHELPS-BOLLOB ´AS THEOREM FOR OPERATORS 19 Problem 5.1. Is it true, that ΦS(X, R, ε) (cid:54) min{√ 2ε, 1} for all real Banach spaces X? 5. AN OPEN PROBLEM In order to explain what do we mean, recall that for the original Bishop-Phelps-Bollob´as modulus the estimation X (ε) (cid:54) √ ΦS (5.1) holds true for all X. In other words, for every (x, x∗) ∈ SX × SX∗ with x∗(x) > 1 − ε, there is (y, y∗) ∈ SX × SX∗ with y∗(y) = 1 such that max{(cid:107)x − y(cid:107) < 2ε. When we take Y = R in the definition of ΦS(X, Y, ε), the only difference with ΦS X (ε) is that by attaining norm we mean y∗(y) = 1, instead of y∗(y) = 1. So, in the case of ΦS(X, R, ε) we have more possibilities to approximate (x, x∗) ∈ SX × SX∗ with x∗(x) > 1 − ε: 2ε,(cid:107)x∗ − y∗(cid:107)} < √ √ 2ε (y, y∗) ∈ SX × SX∗ with y∗(y) = 1 or y∗(y) = −1. Estimation (5.1) is sharp for the two-dimensional real (cid:96)1 space: (5.2) (ε) = √ ΦS 2ε, (cid:96)(2) 1 but, as we have shown in Theorem 3.4 (5.3) √ 1 , R, ε) = min{ ΦS((cid:96)(2) 2ε, 1}. Estimations (5.2) and (5.3) coincide for ε ∈ (0, 1/2), but for bigger values of ε there is a significant difference. We do not know whether the inequality ΦS(X, R, ε) (cid:54) min{√ (cid:27) Moreover, in all examples that we considered we always were able to estimate ΦS(X, Y, ε) from above by 1. So, we don't know whether the statement of Theorem 2.1 can be improved to 2ε, 1} holds true for all X. (cid:26)√ (cid:114) 1 + ρ ΦS(X, Y, ε) (cid:54) min 2ε 1 − ρ , 1 . REFERENCES [1] M. D. ACOSTA, R. M. ARON, D. GARC´IA AND M. MAESTRE, The Bishop-Phelps-Bollob´as Theorem for operators, J. Funct. Anal. 254 (2008), 2780 -- 2799. [2] M. D. ACOSTA, J. BECERRA GUERRERO, D. GARC´IA, S. K. KIM, AND M. MAESTRE, Bishop-Phelps-Bollob´as prop- erty for certain spaces of operators, J. Math. Anal. Appl. 414 (2014), No. 2, 532 -- 545. [3] BISHOP, E., PHELPS, R.R.: A proof that every Banach space is subreflexive, Bull. Amer. Math. Soc 67 (1961), 97 -- 98. [4] BOLLOB ´AS, B.: An extension to the theorem of Bishop and Phelps, Bull. London Math. Soc. 2 (1970), 181 -- 182. [5] CHICA, M., KADETS, V., MART´IN, M., MORENO-PULIDO, S., RAMBLA-BARRENO, F.: Bishop-Phelps-Bollob´as moduli of a Banach space, J. Math. Anal. Appl. 412 (2014), no. 2, 697 -- 719. [6] CHICA, M., KADETS, V., MART´IN, M., MER´I, J., SOLOVIOVA, M.: Two refinements of the Bishop-Phelps-Bollob´as modulus, Banach J. Math. Anal. 9 (2015), no. 4, 296 -- 315. [7] CHICA, M., KADETS, V., MART´IN, M., MER´I, J.: Further Properties of the Bishop-Phelps-Bollob´as Moduli, Mediter- ranean Journal of Mathematics 13(5):3173-3183 (2016) [8] DIESTEL, J.: Geometry of Banach spaces, Lecture notes in Math. 485, Springer-Verlag, Berlin, 1975. [9] R. C. JAMES Uniformly non-square Banach spaces, Ann. Math. (2) 80 (1964), 542 -- 550. [10] PARTINGTON J.: Norm attaining operators, Israel J. Math. 43 (1982), 273-276. [11] V. KADETS, M. SOLOVIOVA: A modified Bishop-Phelps-Bollob´as theorem and its sharpness, Mat. Stud. 44 (2015), 84 -- 88. [12] J. LINDENSTRAUSS: On operators which attain their norm, Israel J. Math. 1 (1963), 139 -- 148. [13] PHELPS, R. R.: Support Cones in Banach Spaces and Their Applications, Adv. Math. 13 (1974), 1 -- 19. 20 KADETS AND SOLOVIOVA (Kadets) DEPARTMENT OF MATHEMATICS AND INFORMATICS, KHARKIV V. N. KARAZIN NATIONAL UNIVERSITY, PL. SVOBODY 4, 61022 KHARKIV, UKRAINE ORCID: 0000-0002-5606-2679 E-mail address: [email protected] (Soloviova) DEPARTMENT OF MATHEMATICS AND INFORMATICS, KHARKIV V. N. KARAZIN NATIONAL UNIVERSITY, PL. SVOBODY 4, 61022 KHARKIV, UKRAINE ORCID: 0000-0002-3777-5286 E-mail address: [email protected]
1009.1061
1
1009
2010-09-06T14:00:57
Tight embedding of subspaces of $L_p$ in $\ell_p^n$ for even $p$
[ "math.FA" ]
Using a recent result of Batson, Spielman and Srivastava, We obtain a tight estimate on the dimension of $\ell_p^n$, $p$ an even integer, needed to almost isometrically contain all $k$-dimensional subspaces of $L_p$.
math.FA
math
Tight embedding of subspaces of Lp in ℓn p for even p ∗ Gideon Schechtman† September 5, 2010 Abstract Using a recent result of Batson, Spielman and Srivastava, We ob- p , p an even integer, needed tain a tight estimate on the dimension of ℓn to almost isometrically contain all k-dimensional subspaces of Lp. In a recent paper [BSS] Batson, Spielman and Srivastava introduced a new method for sparsification of graphs which already proved to have functional analytical applications. Here we bring one more such application. Improving over a result of [BLM] (or see [JS] for a survey on this and related results), we show that for even p and for k of order n2/p any k dimensional subspace of Lp nicely embeds into ℓn p . This removes a log factor from the previously known estimate. The result in Theorem 2 is actually sharper than stated here and gives the best possible result in several respects, in particular in the dependence of k on n. The theorem of [BSS] we shall use is not specifically stated in [BSS], but is stated as Theorem 1.6 in Srivastava's thesis [Sr]: Theorem 1 [BSS] Suppose 0 < ε < 1 and A = Pm i are given, with vi column vectors in Rk. Then there are nonnegative weights {si}m i=1, at most ⌈k/ε2⌉ of which are nonzero, such that, putting A = Pm i=1 sivivT i , (1 − ε)−2xT Ax ≤ xT Ax ≤ (1 + ε)2xT Ax i=1 vivT for all x ∈ Rk. ∗AMS subject classification: 46B07 †Supported by the Israel Science Foundation 1 Corollary 1 Let X be a k-dimensional subspace of ℓm 2 and let 0 < ε < 1. Then there is a set σ ⊂ {1, 2, . . . , m} of cardinality n ≤ Cε−2k (C an absolute constant) and positive weights {si}i∈σ such that kxk2 ≤ (X six2(i))1/2 ≤ (1 + ε)kxk2 i∈σ for all x = (x(1), x(2), . . . , x(m)) ∈ X. Proof: Let u1, u2, . . . , uk be an orthonormal basis for X; uj = (u1,j, u2,j, . . . , um,j), j = 1, . . . , k. Put vT i = Ik, the k × k identity matrix. Let si be the weights given by Theorem 1 (and σ their support). Then, for all x = Pk i = (ui,1, ui,2, . . . , ui,k), i = 1, . . . , m. Then Pm i=1 vivT i=1 aiui ∈ X, (1 − ε)−2kxk2 = aT m X i=1 vivT i a ≤ aT m X i=1 sivivT i a ≤ (1 + ε)2kxk2. Finally, notice that, for each i = 1, . . . , m, aT vivT i-th coordinate of x. Thus, i a = x(i)2, the square of the aT m X i=1 sivivT i a = m X i=1 six(i)2. We first prove a simpler version of the main result. Proposition 1 Let X is a k dimensional subspace of Lp for some even p and let 0 < ε < 1. Then X (1 + ε)-embeds in ℓn p for n = O((εp)−2kp/2). Proof: Assume as we may that X is a k dimensional subspace of ℓm p for some finite m. Consider the set of all vectors which are coordinatewise products of p/2 vectors from X; i.e, of the form (x1(1)x2(1) . . . xp/2(1), x1(2)x2(2) . . . xp/2(2), . . . , x1(m)x2(m) . . . xp/2(m)) where xj = (xj(1), xj(2), . . . , xj(m)), j = 1, 2, . . . , p/2, are elements of X. We shall denote the vector above as x1 · x2 · · · · · xp/2. The span of this set in Rm, which we denote by X p/2, is clearly a linear space of dimension at 2 most kp/2. Consequently, by Corollary 1, there is a set σ ⊂ {1, 2, . . . , m} of cardinality at most C(εp)−2kp/2 and weights {si}i∈σ such that kyk2 ≤ (X siy2(i))1/2 ≤ (1 + i∈σ εp 4 )kyk2 for all y ∈ X p/2. Restricting to y-s of the form y = (x(1)p/2, x(2)p/2, . . . , x(m)p/2) with x = (x(1), x(2), . . . , x(m)) ∈ X, we get kxkp/2 p ≤ (X sixp(i))1/2 ≤ (1 + i∈σ εp 4 )kxkp/2 p . Raising these inequalities to the power 2/p gives the result. We now state and prove the main result. Theorem 2 Let X be a k dimensional subspace of Lp for some even p ≤ k p )p/2). and let 0 < ε < 1. Then X (1 + ε)-embeds in ℓn Equivalently, for some universal c > 0, for any n and any k ≤ cε4/ppn2/p, any k-dimensional subspace of Lp (1 + ε)-embeds in ℓn p . p for n = O(ε−2( 10k Proof: The only change from the previous proof is a better estimate of the dimension of the auxiliary subspace involved. An examination of the proof above shows that it is enough to apply Corollary 1 to any subspace containing all the vectors xp/2 = x · x · · · ·· x (p/2 times), x = (x(1), . . . , x(m)) ∈ X. The smallest such subspace is the space of degree p/2 homogeneous polynomials in elements of X. Its basis is the set of monomials of the form up1 j2 ·· · ··upl jl with p1 + · · · + pl = p/2, where u1, . . . , uk is a basis for X. The dimension of this space, which is the number of such monomials, is (cid:0)k+p/2−1 (cid:1) ≤ ( 10k j1 ·up2 p/2 p )p/2. Next we remark on the estimate k ≤ cε4/ppn2/p. This estimate improves (unfortunately, only for even p) over the known es- timates (the best of which is in [BLM]) by removing a log n factor that was presented in the best estimate till now. Also, the p in front of the n2/p is a nice feature. It is known that the dependence of k on p and n in this esti- mate is best possible even if one restrict to subspaces of Lp isometric to ℓk (see [BDGJN]). Actually the result above indicates that ℓk 2 are the "worse" subspaces. 2 3 As for the dependence on ε, the published proofs gives at best quadratic dependence while here we get a quadratic dependence for p = 4 and better ones as p grows. For the special case of X = ℓk 2 and p = 4 a better result is 2 isometrically into ℓ4n2 known: One can embed ℓk ([Ko]). But for p = 6, 8, . . . we get better result here even for this special case than what was previously known (The best I knew was a linear dependence on ε - this is unpublished. Here we get ε2/3 for p = 6 and better for larger p-s.) It is not clear that this is the best possible dependence on ε, but note that for some combinations of ε and p and in particular for every ε and large enough p (p > c log 1/ε) the dependence on ε becomes a constant and, up to universal constants, we get the best possible result with respect to all parameters. 4 References [BSS] Batson, J. D., Spielman, D. A., and Srivastava, N., Twice-Ramanujan sparsifiers. STOC 09: Proceedings of the 41st annual ACM symposium on Theory of computing (New York, NY, USA, 2009), ACM, pp. 255262. [BDGJN] Bennett, G., Dor, L. E., Goodman, V., Johnson, W. B., Newman, C. M., On uncomplemented subspaces of Lp, 1 < p < 2. Israel J. Math. 26 (1977), no. 2, 178 -- 187. [BLM] Bourgain, J., Lindenstrauss, J., and Milman, V., Approximation of zonoids by zonotopes. Acta Math. 162 (1989), no. 1-2, 73 -- 141. [JS] Johnson, William B., Schechtman, Gideon, Finite dimensional sub- spaces of Lp. Handbook of the geometry of Banach spaces, Vol. I, 837 -- 870, North-Holland, Amsterdam, 2001. [Ko] Konig, Hermann, Isometric imbeddings of Euclidean spaces into finite- dimensional lp-spaces. Panoramas of mathematics (Warsaw, 1992/1994), 79 -- 87, Banach Center Publ., 34, Polish Acad. Sci., Warsaw, 1995. [Sr] N. Srivastava, PhD dissertation, Yale 2010, http://www.cs.yale.edu/homes/srivastava/dissertation.pdf 4 Gideon Schechtman Department of Mathematics Weizmann Institute of Science Rehovot, Israel E-mail: [email protected] 5
1605.08790
2
1605
2016-12-25T12:00:45
A simple characterization of homogeneous Young measures and weak $L^1$ convergence of their densities
[ "math.FA" ]
We formulate a simple characterization of homogeneous Young measures associated with measurable functions. It is based on the notion of the quasi-Young measure introduced in the previous article published in this Journal. First, homogeneous Young measures associated with the measurable functions are recognized as the constant mappings defined on the domain of the underlying function with values in the space of probability measures on the range of these functions. Then the characterization of homogeneous Young measures via image measures is formulated. Finally, we investigate the connections between weak convergence of the homogeneous Young measures understood as elements of the Banach space of scalar valued measures and the weak* L1 sequential convergence of their densities. A scalar case of the smooth functions and their Young measures being Lebesgue-Stieltjes measures is also analyzed.
math.FA
math
A simple characterization of homogeneous Young measures and weak L1 convergence of their densities Piotr Pucha la Institute of Mathematics, Czestochowa University of Technology, Armii Krajowej 21, 42-200 Cz¸estochowa, Poland Email: [email protected], [email protected] Abstract We formulate a simple characterization of homogeneous Young mea- sures associated with measurable functions. It is based on the notion of the quasi-Young measure introduced in the previous article published in this Journal. First, homogeneous Young measures associated with the measurable functions are recognized as the constant mappings defined on the domain of the underlying function with values in the space of probabil- ity measures on the range of these functions. Then the characterization of homogeneous Young measures via image measures is formulated. Finally, we investigate the connections between weak convergence of the homo- geneous Young measures understood as elements of the Banach space of scalar valued measures and the weak∗ L1 sequential convergence of their densities. A scalar case of the smooth functions and their Young measures being Lebesgue-Stieltjes measures is also analyzed. keywords: homogeneous Young measures; weak convergence of mea- sures; weak convergence of functions; optimization AMS Subject Classification: 46N10; 49M30; 74N15 1 Introduction Young measures is an abstract measure theoretic tool which has arisen in the context of minimizing bounded from below integral functionals not attaining their infima. The often used direct method relays on constructing minimizing considered functional J sequence (un) of functions. This sequence is converg- ing to some function u0 in an appropriate (usually weak∗) topology, while the sequence J (un) of real numbers converges to inf J . However, in general it is not true that the sequence (f (x, un(x), ∇un(x))) of compositions of the inte- grand f of J with elements of (un) converges to f (x, u0(x), ∇u0(x))). This is connected with the lack of quasiconvexity of the integrand f with respect to its third variable. This situation is often taking place in elasticity theory (here f is the density of the internal energy of the displaced body). See for example [4, 5, 7] and references cited there for more details. It is Laurence Chisholm Young who has first realized that there is need to enlarge the space of admissible functions to the space of more general objects. In [10], the very first article devoted to the subject, these objects are called by 1 the author 'generalized curves'. Today we call them 'Young measures'. In fact, "Young measure' is not a single measure, but a family of probability measures. More precisely, let there be given: • Ω -- a nonempty, bounded open subset of Rd with Lebesgue measure M > 0; • a compact set K ⊂ Rl; • (un) -- a sequence of measurable functions from Ω to K, convergent to some function u0 weakly∗ in an appropriate function space; • f -- an arbitrary continuous real valued function on Rd. The sequence (f (un)) has a subsequence weakly∗ convergent to some function g. However, in general g 6= f (u0). It can be proved that there exists a subsequence of (f (un)), not relabelled, and a family (νx)x∈Ω of probability measures with supports suppνx ⊆ K, such that ∀ f ∈ C(Rd) ∀w ∈ L1(Ω) there holds n→∞Z lim f (un(x))w(x)dx = Z Z f (s)νx(ds)w(x)dx := Z Ω Ω K Ω f (x)w(x)dx. This family of probability measures is today called a Young measure associated with the sequence (un). It often happens, both in theory and applications, that the Young measure is a 'one element family', that is it does not depend on x ∈ Ω. We call such Young measure a homogeneous one. Homogeneous Young measures are the first and often the only examples of the Young measures associated with specific sequences of functions. They are also the main object of interest in this article. Calculating an explicit form of particular Young measure is an important task. In elasticity theory, for example, Young measures can be regarded as means of the limits of minimizing sequences of the energy functional. These means summarize the spatial oscillatory properties of minimizing sequences, thus conserving some of that information; information that is entirely lost when quasiconvexification of the original functional is used as the minimiza- tion method. To quote from [4], page 121: The Young measure describes the local phase proportions in an infinitesimally fine mixture (modelled mathemati- cally by a sequence that develops finer and finer oscillations), and a page earlier: The Young measure captures the essential feature of minimizing sequences. It is worth noting, that the Young measures calculated in [4], in the context of elas- ticity theory or micromagnetism, are homogeneous ones. This indicates their role. This is also a case in many other books or articles devoted to the subject. Unfortunately, there is still no any general method of calculating their explicit form (homogeneous or not) when associated with specific, in particular mini- mizing, sequence. Some of the existing methods relay on periodic extensions of the elements of the sequence and application the generalized version of the Riemann-Lebesgue lemma or calculating weak∗ limits of function sequences. In the article [6] there has been proposed an elementary method of calcu- It is based on the notion of the lation an explicit form of Young measures. 2 'quasi-Young measure' introduced there and the approach to Young measures as in [9]. This approach enables us to look at them as at objects associated with any measurable function defined on Ω with values in K. Using only the change of variable theorem the explicit form of the quasi-Young measures for functions that are piecewise constant or piecewise invertible with differentiable inverses has been calculated. The result extends for sequences of oscillating functions of that shape. Finally, it has been proved that calculated quasi-Young measures are the Young measures associated with considered (sequences of) functions. Significantly, all of them are homogeneous. This article can be viewed as the continuation of [6]. First, we recognize quasi-Young measures as homogeneous Young measures which are, in turn, con- stant mappings on Ω with values in the space of probability measures on K. Then we provide the characterization of homogeneous Young measures asso- ciated with measurable functions form Ω to K as images of the normalized Lebesgue measure on Ω with respect to the underlying functions. The proofs of these facts are really very simple, but it seems that such characterization of ho- mogeneous Young measures has not been formulated yet. Then we show that the weak convergence of the densities of the homogeneous Young measures is equiv- alent to the weak convergence of these measures themselves. Here measures are understood as vectors in the Banach space of scalar valued measures with the total variation norm. The proof of this fact is also simple but relies on the non- trivial characterization of the weak sequential L1 convergence of the sequence of functions. In a special one-dimensional case the Young measure associated with function having differentiable inverse is a Lebesgue-Stieltjes measure and we can state the result for mappings having inverses of the C∞ class. 2 Preliminaries and notation We gather now some necessary information about Young measures. As it has been mentioned above, our approach is as in [9], where the material is pre- sented in great detail. It is sketched in [6]. See also the references cited in aforementioned articles. Let Ω be an open subset of Rd with Lebesgue measure M > 0, dµ(x) := 1 M dx, where dx is the d-dimensional Lebesgue measure on Ω and let K ⊂ Rl be compact. We will denote: • rca(K) -- the space of regular, countably additive scalar measures on K, equipped with the norm kρkrca(K) := ρ(Ω), where · stands in this case for the total variation of the measure ρ. With this norm rca(K) is a Banach space, so we can consider a weak convergence of its elements; • rca1(K) -- the subset of rca(K) with elements being probability measures on K; • L∞ w∗(Ω, rca(K)) -- the set of the weakly∗ measurable mappings ν : Ω ∋ x → ν(x) ∈ rca(K). 3 (It follows that homogeneous Young measures are those from the above mappings, that are constant a.e. in Ω) We equip this set with the norm kνkL∞ w∗ (Ω,rca(K)) := ess sup(cid:8)kν(x)krca(K) : x ∈ Ω(cid:9). Usually by a Young measure we understand a family (νx)x∈Ω of regular probability Borel measures on K, indexed by elements x of Ω . However, we should remember that the Young measure is in fact a weakly∗ measurable map- ping defined on an open set Ω ⊂ Rd having positive Lebesgue measure with values in the set rca1(K). The set of the Young measures on the compact set K ⊂ Rl will be denoted by Y(Ω, K): Y(Ω, K) := (cid:8)ν = (ν(x)) ∈ L∞ w∗(Ω, rca(K)) : νx ∈ rca1(K) for a.a x ∈ Ω(cid:9). Remark 2.1. In [9] Young measures are defined as weakly measurable map- pings, but it seems to be an innacuracy, since rca(K) is in fact a conjugate space. Compare footnote 80 on page 36 in [9] with, for example, definition 2.1.1 (d) on page109 in [3]. However, when considering weak convergence of the Young measures, we will look at the rca(K) as at the normed space itself with total variation norm. Definition 2.1. We say that a family (νx)x∈Ω is a quasi-Young measure associa- ted with the measurable function u : Rd ⊃ Ω → K ⊂ Rl if for every integrable function β : K → R there holds Z K β(k)dνx(k) = Z Ω β(u(x))dµ(x). We will write νu to indicate that the (quasi-)Young measure ν is associated with the function u. Remark 2.2. This definition is slightly more general than the one formulated in [6]. However, when dealing with Young measures, we must restrict ourselves to continuous functions β to be able to use Riesz representation theorems. See [9] for details. 3 Homogeneous Young measures We first prove that quasi-Young measures are constant mappings. Proposition 3.1. Let νu be the quasi-Young measure associated with the Borel function u. Then νu, regarded as a mapping from Ω with values in rca(K), is constant. 4 Proof. Suppose that νu is not a constant function. Then there exist x1, x2 ∈ Ω, x1 6= x2, such that νu x2. Then for any integrable β : K → R there holds x1 6= νu Z K β(k)dνu x1 (k) = Z β(u(x))dµ(x) = Z Ω K β(k)dνu x2 (k), a contradiction. Corollary 3.1. Let β : K → R be a continuous function. Then the quasi-Young measure νu associated with the function u is a Young measure, that is it is an element of the set Y(Ω, K). Proof. Take β ≡ 1. Then Z K dνu(k) = Z β(k)dνu(k) = Z β(u(x))dµ(x) = Z K Ω Ω dµ(x) = 1, which means that νu ∈ rca1(K). As a constant function it is weakly measurable, so that νu ∈ L∞ w∗(Ω, rca(K)). Corollary 3.2. Quasi-Young measures associated with measurable functions u : Ω → K are precisely the homogeneous Young measures associated with them. Due to the fact that homogeneous Young measures are constant functions, we will write 'ν' instead of 'ν = (νx)x∈Ω'. Now we will prove that homogeneous Young measures are images of Borel measures with respect to the underlying functions. Recall that u : Rd ⊃ Ω ∋ x → u(x) ∈ K ⊂ Rl and that µ is normed to unity Lebesgue measure on Ω. Theorem 3.1. A homogeneous Young measure ν is associated with Borel func- tion u if and only if ν is an image of µ under u. Proof. The theorem follows from the equalities Z K β(k)dν(k) = Z β(u(x))dµ(x) = Z Ω K β(k)dµu−1. 4 Weak convergence 4.1 Certain facts on weak convergence of functions and measures Let (X, A, ρ) be a measure space and consider a sequence (vn) of scalar func- tions defined on X and integrable with respect to the measure ρ (that is, 5 ∀n ∈ N vn ∈ L1 weakly sequentially to v if ρ(X)) and a function v ∈ L1 ρ(X). Recall that (vn) converges ∀g ∈ L∞ ρ (X) n→∞Z lim vngdρ = Z X X vgdρ. Next theorem characterizes weak sequential L1 convergence of functions and weak convergence of measures. We refer the reader to [1], chapter 6 or [2], chapter VII. Theorem 4.1. (a) let X be a locally compact Hausdorff space and (X, A, ρ) ρ(X) converges -- a measure space with ρ regular. A sequence (vn) ⊂ L1 weakly to some v ∈ L1 ρ(X) if and only if ∀A ∈ A the limit n→∞Z lim A vndρ exists and is finite; (b) let X be a locally compact Hausdorff space and denote by B(X) the σ- algebra of Borel subsets of X. A sequence (ρn) of scalar measures on B(X) converges weakly to some scalar measure ρ on B(X) if and only if ∀A ∈ B(X) the limit exists and is finite. lim n→∞ ρn(A) 4.2 Weak convergence of homogeneous Young measures and their densities Consider the function u : Ω → K of the form u := n Xi=1 uiχΩi , where: • the open sets Ωi, i = 1, 2, . . . , n, are pairwise disjoint and Sn denote this partition of Ω by {Ω}; i=1 Ωi = Ω; • the functions ui : Ωi → K, i = 1, 2, . . . , n, have continuously differentiable inverses u−1 i with Jacobians Ju−1 ; i • for any i = 1, 2, . . . , n ui(Ωi) = K. Theorem 4.2. ([6]) The Young measure νu associated with u is a homogeneous Young measure that is abolutely continuous with respect to the Lebesgue measure dy on K. Its density is of the form g(y) = 1 M n Xi=1 Ju−1 i (y). 6 Consider now a family {Ωl}, l ∈ N, with elements Ωl and a sequence (ul) of functions of the form ul := Pn(l) i=1 ul i, of open partitions of Ω satisfying, for fixed l, the above respective conditions. According to Theorem 4.2 the Young measure associated with ul is absolutely continuous with respect to the Lebesgue measure dy on K with density iχΩl i gl(y) = 1 M n(l) Xi=1 J(ul i)−1 (y). Theorem 4.3. Let (ul) be the sequence of functions described above and denote by νl the Young measure with density gl associated with the function ul. Then the sequence (gl) is weakly convergent in L1(K) to some function h if and only if the sequence (νl) is weakly convergent to some measure η. Proof. . (⇒) Since (gl) is weakly convergent in L1(K), then for any measurable A ⊆ K the limit l→∞Z lim gldy A exists and is finite. This in turn is equivalent to the fact that the sequence (νl) of Young measures is weakly convergent to some measure η. (⇐) We proceed as above, but start the reasoning from the weak convergence of the sequence (νl). Let I be an open interval in R with Lebesgue measure M > 0 and u -- a strictly monotonic differentiable real valued function on I. We can assume that u is strictly increasing. Then the set K := u(I) is compact and for any β ∈ C(K, R) we have Z K β(y) 1 M (u−1)′(y)dy = Z β(y)d 1 M u−1(y). K This means that the (homogeneous) Young measure associated with u is a Lebesgue-Stieltjes measure on K. This observation together with standard in- ductive argument lead to the following one-dimensional corollary to theorem 4.3. Corollary 4.1. Let (un) be a sequence of real valued functions from a bounded, nondegenerate interval I ⊂ R such that: (i) for any n ∈ N un is a C∞-diffeomorphism; (ii) u(l) n (I) := Kl is compact, l ∈ N ∪ {0}; n is strictly positive or negative, n ∈ N. (iii) u(l) Then for any fixed l ∈ N the weak L1 convergence of the sequence (cid:0)(u−1 Stieltjes measures associated with the elements of the sequence (cid:0)(u−1 n )(l)(cid:1). is equivalent to the weak convergence of the sequence of the Young-Lebesgue- n )(l)(cid:1) 7 Example 4.1. (see [8]) Let I := (a, b), a < b, K := [c, d], c < d and let (un) be a sequence of strictly monotonic functions from I with values in K such that: (i) ∀ n ∈ N un(I) = K; (ii) ∀ n ∈ N un has continuously differentiable inverse (u−1 n )′; (iii) the sequence (u−1 n )′ is nondecreasing. By monotonicity of the sequence of derivatives for any Borel subset A ⊆ K and any natural numbers m ≤ n we have Z A m )′(y)dy ≤ Z (u−1 A (u−1 n )′(y)dy. Further, ∀ n ∈ N the Young measure associated with function (u−1 geneous one. Since it is probability measure on K, the sequence (cid:16)RA is monotonic and bounded. Thus for any Borel subset A ⊆ K the limit n )′ is a homo- n )′(y)dy(cid:17) (u−1 n→∞Z lim (u−1 n→∞Z n )′(y)dy = lim du−1 n (y) A A exists and is finite, which by theorem 4.1 yields the weak L1 convergence of the n )′(cid:1). This, by theorem 4.3, is equivalent to the weak convergence of the sequence of Young measures, whose elements are associated with respective sequence (cid:0)(u−1 elements of (cid:0)(u−1 n )′(cid:1). Acknowledgements. Author would like to thank Professor Agnieszka Ka lamajska for discussion that took place during Dynamical Systems and Ap- plications IV conference in June 2016, L´od´z, Poland. References [1] J. J. Benedetto and W. Czaja, Integration and Modern Analysis., Birkhauser, Boston, Basel, Berlin, 2009. [2] J. Diestel, Sequences and Series in Banach Spaces., Springer-Verlag, New York, Berlin, Heidelberg, Tokyo, 1984. [3] L. Gasi´nski, N. S. Papageorgiou, Nonlinear analysis, in Series in Mathe- matical Analysis and Applications, vol. 9, R. P. Agarwal, D. O'Regan, eds, Chapman & Hall/CRC, Taylor & Francis Group, Boca Raton, London, New York, Singapore, 2006. [4] S. Muller, Variational Models for Microstructure and Phase Transitions., in Calculus of variations and geometric evolution problems, S. Hildebrandt, and M. Struwe, eds, Lecture Notes in Mathematics Volume 1713, Springer Verlag, Berlin, Heidelberg, Germany, 1999. 8 [5] P. Pedregal, Variational Methods in Nonlinear Elasticity., SIAM, Philadel- phia, USA, 2000. [6] P. Pucha la, An elementary method of calculating an explicit form of Young measures in some special cases, Optimization Vol. 63, No.9, (2014), pp. 1419 -- 1430. [7] P. Pucha la, Minimization of functional with integrand expressed as min- imum of quasiconvex functions -- general and special cases, in Calculus of Variations and PDEs, T. Adamowicz, A. Ka lamajska, S. Mig´orski, A. Ochal, eds, Banach Center Publications Volume 101, Warszawa 2014. [8] P. Pucha la, Weak convergence in L1 of the sequences of monotonic func- tions, J. Appl. Math. Comput. Mech., Vol.13, No.3, (2014), pp. 195 -- 199. [9] T. Roub´ıcek, Relaxation in Optimization Theory and Variational Calculus, Walter de Gruyter, Berlin, New York, 1997. [10] L.C. Young, Generalized curves and the existence of an attained absolute minimum in the calculus of variations, Comptes Rendus de la Soci´et´e des Sciences et des Lettres de Varsovie, classe III 30 (1937), pp. 212 -- 234. 9
1907.08420
1
1907
2019-07-19T09:23:55
Hausdorff operators on Bergman spaces of the upper half plane
[ "math.FA" ]
In this paper we study Hausdorff operators on the Bergman spaces $A^{p}(\mathbb{U})$ of the upper half plane.
math.FA
math
HAUSDORFF OPERATORS ON BERGMAN SPACES OF THE UPPER HALF PLANE GEORGIOS STYLOGIANNIS Abstract. In this paper we study Hausdorff operators on the Bergman spaces Ap(U) of the upper half plane. 1. introduction Given a σ-finite positive Borel measure µ on (0,∞), the associated Hausdorff operator Hµ, for suitable functions f is given by (1) Hµ(f )(z) :=Z ∞ 0 1 t t(cid:17) dµ(t), f(cid:16)z where U = {z ∈ C : Im z > 0} is the upper half plane. Its formal adjoint, the quasi-Hausdorff operator H∗µ in the case of real Hardy spaces H p(R) is z ∈ U (2) H∗µ(f )(z) :=Z ∞ 0 f (tz) dµ(t). Moreover for appropriate functions f and measures µ they satisfy the fundamental identity: 0 x ∈ R, (3) \Hµ(f )(x) =Z ∞ where bf denotes the Fourier transform of f . bf (tx) dµ(t) = H∗µ(bf )(x), The theory of Hausdorff summability of Fourier series started with the paper of Hausdorff [Ha21] in 1921. Much later Hausdorff summa- bility of power series of analytic functions was considered in [Si87] and [Si90] on composition operators and the Ces´aro means in Hardy H p spaces. General Hausdorff matrices were considered in [GaSi01] and [GaPa06]. In [GaPa06] the authors studied Hausdorff matrices on a large class of analytic function spaces such as Hardy spaces, Bergman spaces, BMOA, Bloch etc. They characterized those Hausdorff matri- ces which induce bounded operators on these spaces. Results on Hausdorff operators on spaces of analytic functions were extended in the Fourier transform setting on the real line, starting with [LiMo00] and [Ka01]. There are many classical operators in analysis which are special cases of the Hausdorff operator if one chooses suitable 1 2 GEORGIOS STYLOGIANNIS measures µ such as the classical Hardy operator, its adjoint operator, the Ces´aro type operators and the Riemann-Liouville fractional inte- gral operator. See the survey article [Li013] and the references there in. In recent years, there is an increasing interest on the study of boundedness of the Hausdorff operator on the real Hardy spaces and Lebesque spaces (see for example [An03], [BaGo19], [FaLi14], [LiMo01] and [HuKyQu18]). Motivated by the paper of Hung et al. [HuKyQu18] we describe the measures µ that will induce bounded operators on the Bergman spaces Ap(U) of the upper half-plane. Next Theorem summarizes the main results (see Theorems 3.5 and 3.7 ): Theorem 1.1. Let 1 ≤ p < ∞ and µ be an σ-finite positive measure on (0,∞). The Hausdorff operator Hµ is bounded on Ap(U) if and only if Moreover Z ∞ 0 1 t1− 2 p dµ(t) < ∞. HµAp(U)→Ap(U) =Z ∞ 1 t1− 2 2. preliminaries 0 p dµ(t). To define single-valued functions, the principal value of the argument it is chosen to be in the interval (−π, π]. For 1 ≤ p < ∞, we denote by Lp(dA) the Banach space of all measurable functions on U such that fLp(dAa) :=(cid:18) 1 πZU fpdA(cid:19)1/p < ∞, where dA is the area measure. The Bergman space Ap(U) consists of all holomorphic functions f on U that belong to Lp(dA). Sub-harmonicity yields a constant C > 0 such that (4) f (z)p ≤ C (Im(z))2fp Ap(U), z ∈ U, for f ∈ Ap(U) and (Im(z))2f (z)p = 0 lim z→∂ bU for functions Ap(U), where bU := U ∪ {∞} (see [ChKoSm17]). In par- ticular, this shows that each point evaluation is a continuous linear functional on Ap(U). The duality properties of Bergman spaces are well known in literature see[Zh90] and [BaBoMiMi16]. It is proved that for 1 < p < ∞, 1 q = 1, the dual space of the Bergman space Ap(U) is (Ap(U))∗ ∼ Aq(U) under the duality pairing, p + 1 hf, gi = 1 πZU f (z)g(z) dA(z). HAUSDORFF OPERATORS ON BERGMAN SPACES OF THE UPPER HALF PLANE3 3. Main results In what follows, unless otherwise stated, µ is a positive σ-finite mea- sure on (0,∞). We start by giving a condition under which Hµ is well defined. then Lemma 3.1. Let 1 ≤ p < ∞ and f ∈ Ap(U). If R ∞ t(cid:17) dµ(t) f(cid:16)z Hµ(f )(z) =Z ∞ is a well defined holomorphic function on U. 1 t 0 0 Proof. For f ∈ Ap(U) using (4) we have 1 1− 2 p t dµ(t) < ∞, Hµ(f )(z) ≤Z ∞ 0 1 t(cid:12)(cid:12)(cid:12)f(cid:16) z t(cid:17)(cid:12)(cid:12)(cid:12) dµ(t) p Z ∞ 1 t1− 2 0 p ≤ C fAp(U) Im(z) 2 dµ(t) < ∞. Thus Hµf is well defined, and is given by an absolutely convergent integral, so it is holomorphic. (cid:3) 2+λ p , then Proof. Using polar coordinates for the integral over U we find Lemma 3.2. Let λ > 0 and δ > 0. If gλ,δ(z) = z + δi− 1 λδλ Lp(dA) ≤ 2 2+λ · · 2 1 λδλ ≤ gλ,δp 2(cid:19)2+λ (cid:18)1 πZU(cid:12)(cid:12)(cid:12)(cid:12) z + δi(cid:12)(cid:12)(cid:12)(cid:12) r2 + δ2 + 2rδ sin(θ)(cid:19) 2+λ 0 (cid:18) πZ π 0 Z ∞ dA(z) 2+λ = 1 1 1 1 2 gλ,δp Lp(dA) = r drdθ. Denote by I the last double integral. Then I ≤Z ∞ rdr 2 r2 + δ2(cid:19) 2+λ 0 (cid:18) 1 r + δ(cid:19)2+λ 0 (cid:18) 1 2 Z ∞ 2+λ ≤ 2 = 2 2+λ 2 1 λδλ . (r + δ)dr 4 GEORGIOS STYLOGIANNIS On the other hand, 0 (cid:18) 1 r + δ(cid:19)2+λ I ≥Z ∞ δ (cid:18) 1 r + δ(cid:19)2+λ ≥Z ∞ ≥(cid:18) 1 2(cid:19)2+λZ ∞ δ (cid:18)1 2(cid:19)2+λ 1 =(cid:18)1 λδλ , rdr rdr r(cid:19)2+λ and the assertion follows. rdr (cid:3) 3.1. Test functions. We now consider the test functions which are defined as follows. Let z = x + iy ∈ U and and fε(z) = ϕε(z) = z + εi z + εi = x − i(y + ε) px2 + (y + ε)2 1 , 2 p +ε (z + εi) with ε > 0 small enough. Note that, fε ≡ gpε,ε with respect to the notation of Lemma 3.2, and that ϕε(z) lies on the unit circle with −π < arg(ϕε(z)) < 0, and the following identity holds fε(z) = ϕε(z) 2 p +εfε(z). Let a, b ∈ (−π, π] and set A[a,b] = {z ∈ U : a ≤ arg(z) ≤ b}, with obvious modifications in the case of A(a,b), A(a,b] and A[a,b). Lemma 3.3. The following holds: (i) If 2 < p < ∞ and 2 p + ε ≤ 1, then Re fε(z) ≥ Re ϕε(z)fε(z) for every z ∈ A(0, π 2 ]. (ii) If 1 < p ≤ 2 and 1 < 2 p + ε < 2, then for every z ∈ A[ π 4 , π (ii) If p = 1, 0 < θ0 < π Im fε(z) > C(p)Im ϕε(z)fε(z) 2 ]. 16 and (2 + ε)( π 2 + θ0) < 5π Re fε(z) > Re ϕε(z)fε(z) 4 , then for every z ∈ A[ π 2 , π 2 +θ0] HAUSDORFF OPERATORS ON BERGMAN SPACES OF THE UPPER HALF PLANE5 Proof. Taking real and imaginary parts we have Re fε(z) = fε(z)Re ϕε(z) 2 p +ε = fε(z) cos(( 2 p + ε)θ) and Im fε(z) = fε(z)Im ϕε(z) 2 p +ε = fε(z) sin(( 2 p + ε)θ), where θ := θ(z, ε) = arg ϕε(z). (i): It easy is see that Re fε(z) = fε(z) cos(( 2 p + ε)θ) ≥ fε(z) cos(θ) = Re ϕε(z)fε(z). (ii): Let a = a(p) > 0 such that 1 < 2 simple geometric arguments imply that θ ∈ [− π − aπ p + ε)θ < − π 2 < ( 2 4 . This implies that sin(( 2 p + ε < a < 2. Since z ∈ A[ π 4 , π 2 ] 4 ). Moreover −π < 2 ,− π (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < √2 √2 2 }fε(z) sin(θ) min{sin( aπ 2 ), p + ε)θ) sin(θ) √2 2 } <(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2 ]. We calculate for every z ∈ A[ π 4 , π Im fε(z) = fε(z) sin(( 2 p + ε)θ) > min{sin( aπ 2 ), √2 2 }Im ϕε(z)fε(z). aπ 2 ), = min{sin( This proves (ii) with C(p) = min{sin( a(p)π 2 +θ0] we have that θ ∈ (− π (iii): Since z ∈ A[ π 2 , π π + θ0)(2 + ε) < (2 + ε)θ ≤ − < −( 2 √2 2 }. 2 − θ0,− π π 2 5π 4 − ), 2 2 ]. Thus (2 + ε) < −π. This implies that cos((2 + ε)θ) > cos(θ) and therefore Re fε(z) > Re ϕε(z)fε(z). (cid:3) 3.2. Growth estimates. Let a, b ∈ (−π, π] and set S[a,b] = {z ∈ U : a ≤ arg(z) ≤ b, z ≥ 1}, be a truncated sector with obvious modifications in the case of S(a,b), S(a,b] and S[a,b). Since µ is positive ReHµ(fε) = Hµ(Re fε) and ImHµ(fε) = Hµ(Im fε). Note that if Re fε or Im fε have constant sign on some sector A, then and Hµ(Im fε)(z) = Hµ(Im fε)(z) Hµ(Re fε)(z) = Hµ(Re fε)(z) 6 GEORGIOS STYLOGIANNIS for every z ∈ A. Lemma 3.4. Let 1 ≤ p < ∞ and suppose that Hµ is bounded on Ap(U). Then there are ε(p) and k(p) positive constants such that Hµ(fε)p Ap(U) ≥ k(p) Z 1 0 ε dµ(t)!p 1 pε , 1 t1− 2 p−ε for every ε in (0, ε(p)]. Proof. We will consider three cases for the range of p. Note that if z is in a truncated sector S then z/t belongs to the corresponding sector A for every t > 0. Case I. Let 2 < p < ∞ and ε(p) such that 2 every ε in (0, ε(p)] p + ε(p) < 1. Then for Hµ(fε)p Ap(U) ≥ ReHµ(fε)p Ap(U) 1 2 ](cid:12)(cid:12)(cid:12)(cid:12)Z ∞ πZS(0, π 2 ](cid:18)Z ∞ πZS(0, π 1 0 0 p Ap(U) = Hµ(Re fε)p 1 t Re fε(z/t) dµ(t)(cid:12)(cid:12)(cid:12)(cid:12) tRe fε(z/t) dµ(t)(cid:19)p 1 dA(z) ≥ = dA(z). Denote by I the last integral on S(0, π 2 ]. By (i) of Lemma 3.3 we have I ≥ = 1 πZS(0, π πZS(0, π 1 2 ] ε 0 1 2 ] Z 1 tRe ϕε(z/t) fε(z/t) dµ(t)!p Z 1 p−ε px2 + (y + tε)2! 2 0 dµ(t) t1− 2 p +ε+1 1 ε p dA(z) xpdxdy. HAUSDORFF OPERATORS ON BERGMAN SPACES OF THE UPPER HALF PLANE7 Using polar coordinates and noting that tε < z for z ≥ 1 and t ≤ ε−1, we have rp+1(cos(θ))p dθ π dr p dµ(t) t1− 2 p−ε rp+1(cos(θ))p dθ π dr p−ε!p p +ε+1 ε ε ε 2 2 1 1 t1− 2 p +ε+1 dµ(t) 0 Z 1 pr2 + 2rtε sin(θ) + t2ε2! 2 0 I ≥Z ∞ 1 Z π 0 Z 1 √r2 + 2rtε + t2ε2(cid:19) 2 0 (cid:18) ≥Z ∞ 1 Z π dµ(t)!pZ ∞ ≥ k(p) Z 1 dµ(t)!p = k(p) Z 1 where k(p) = 2−p(ε+1)R π Case II. Let 2 < p < ∞ and ε(p) such that 1 < 2 for every ε in (0, ε(p)] 0 (cos(θ))p dθ 4π . 1 r1+pε dr 1 t1− 2 p−ε 1 t1− 2 p−ε 1 pε , 0 0 1 2 ε p + ε(p) < 2. Then Hµ(fε)p Ap(U) ≥ = 1 πZS[ π πZS[ π 1 0 2 ](cid:12)(cid:12)(cid:12)(cid:12)Z ∞ 2 ](cid:18)Z ∞ 0 1 t p Im fε(z/t) dµ(t)(cid:12)(cid:12)(cid:12)(cid:12) tIm fε(z/t) dµ(t)(cid:19)p 1 4 , π 4 , π dA(z) dA(z). Denote by I the last integral on S[ π 2 ]. By (ii) of Lemma 3.3 we have 4 , π dA(z) p ε ε 0 1 4 , π ≥ C(p) C(p) I ≥ π ZS[ π π ZS[ π tIm ϕε(z/t) fε(z/t) dµ(t)!p 2 ] Z 1 p−ε Z 1 0 px2 + (y + tε)2! 2 sired conclusion with constant k(p) = C(p)2−p(ε+1)R π dµ(t) t1− 2 p +ε+1 4 , π 2 ] 2 π 4 1 (sin(θ))p dθ 4π . Using polar coordinates and working as is Case I, we arrive at the de- ypdxdy. 8 GEORGIOS STYLOGIANNIS Case III. Let p = 1 and θ0 as in Lemma 3.3. Let ε(1) such that (2 + ε(1))( π 4 . Then for every ε in (0, ε(1)] 2 + θ0) < 5π Hµ(fε)A1(U) ≥ = 1 πZS[ π πZS[ π 1 2 , π 2 , π 0 2 +θ0](cid:12)(cid:12)(cid:12)(cid:12)Z ∞ 2 +θ0]Z ∞ 0 1 t Re fε(z/t) dµ(t)(cid:12)(cid:12)(cid:12)(cid:12) dA(z) 1 tRe fε(z/t) dµ(t) dA(z). Denote by I the last integral on S[ π have 2 +θ0]. By (iii) of Lemma 3.3 we 2 , π 1 tRe ϕε(z/t) fε(z/t) dµ(t) dA(z) I ≥ ε 1 2 , π πZS[ π πZS[ π 2 +θ0]Z 1 0 2 +θ0]Z 1 0 ε 1 1 = 2 , π px2 + (y + tε)2!3+ε desired conclusion with constant k(1) = −2−(ε+1)R π dµ(t) t−1−ε (−x)dxdy. Using polar coordinates and working as is Case I, we arrive at the 2 +θ0 π 2 cos(θ) dθ 4π . (cid:3) Theorem 3.5. Let 1 ≤ p < ∞. The operator Hµ is bounded on Ap(U) if and only if then Lemma 3.1 implies that Hµ(f ) is well defined and holomorphic in U. An easy computation involving the Minkowski inequality shows that for all 1 ≤ p < ∞ t(cid:17) dµ(t)(cid:12)(cid:12)(cid:12)(cid:12) Hµ(f )Ap(U) =(cid:18)ZU(cid:12)(cid:12)(cid:12)(cid:12)Z ∞ f(cid:16) z ≤ fAp(U)Z ∞ 1 t1− 2 1 t 0 0 p dµ(t) < ∞. p dA(z)(cid:19)1/p Thus Hµ is bounded on Ap(U). Conversely, suppose that Hµ is bounded. Let fε(z) = (z + εi)−( 2 p +ε) with ε > 0 small enough. By Lemma 3.2 1 pεεpε . Ap(U) ∼ fεp (5) Proof. Suppose that 0 Z ∞ Z ∞ 0 1 t1− 2 p 1 t1− 2 p dµ(t) < ∞. dµ(t) < ∞, HAUSDORFF OPERATORS ON BERGMAN SPACES OF THE UPPER HALF PLANE9 Moreover Lemma 3.4 implies that there is a constant k = k(p) > 0 such that fεp Ap(U)HµAp(U)→Ap(U) ≥ Hµ(fε)p Ap(U) ≥ k Z 1 0 ε dµ(t)!p 1 pε . 1 t1− 2 p−ε Thus by letting ε → 0, we have in comparison to (5) Z ∞ 0 1 t1− 2 p dµ(t) < ∞. (cid:3) In our way to computing the norm of Hµ we will firstly compute the norm of the truncated Hausdorff operator Hδ µ given by : 0 1 t µ(f )(z) :=Z ∞ Hδ 1 t Proposition 3.6. Let 1 ≤ p < ∞ and 0 < δ < 1. If t(cid:17) X[δ,1/δ](t) dµ(t) =Z 1 f(cid:16)z Z ∞ dµ(t) < ∞, 1 t1− 2 p 0 δ δ t(cid:17) dµ(t). f(cid:16) z Proof. As in Theorem 3.5 an application of Minkowski inequality gives then Hδ µ is bounded with δ µAp(U)→Ap(U) =Z 1 Hδ µAp(U)→Ap(U) ≤Z 1 Hδ Let fε(z) = (z + i)− 2 (6) δ δ δ 1 t1− 2 p dµ(t). 1 t1− 2 p dµ(t). p−ε with ε > 0 small enough. We calculate µ(fε)(z) − fε(z)Z 1 Hδ δ δ δ =Z 1 δ 1 t1− 2 p 1 t1− 2 p dµ(t) (ϕε,z(t) − ϕε,z(1)) dµ(t), where ϕε,z(t) = tε (z + ti) . 2 p +ε 10 GEORGIOS STYLOGIANNIS For any t ∈ [δ, 1/δ], calculus gives ϕε,z(t) − ϕε,z(1) ≤ t − 1 sup{ϕ′ε,z(s) : s ∈ [δ, 1/δ]} ( 2 p + ε)(1/δ)ε z + iδ + ε)(1/δ)ε+1gp(ε+1),δ(z). δ εδε−1 p +ε+1! = εδε−2gpε,δ(z) + ( z + iδ ≤ 2 p +ε + 1 2 2 p Where above we followed the notation of Lemma 3.2. Thus by an easy application of Minkowski inequality followed by the triangular inequality we have µ(fε)(z) − fε(z)Z 1 Hδ δ δ 1 t1− 2 p dµ(t)Ap(U) δ δ ≤Z 1 ≤Z 1 δ δ 1 t1− 2 1 t1− p (ϕε,z(t) − ϕε,z(1))Ap(U)dµ(t) 2 p dµ(t)(cid:18)εδε−2gpε,δLp(dA) + ( 2 p + ε)(1/δ)ε+1gp(ε+1),δLp(dA)(cid:19) . This, together with Lemma 3.2 (recall that fε = gpε,ε), yields 1 1− t 2 p dµ(t)Ap(U) δ δ Hδ µ(fε)(z) − fε(z)R 1 ≤Z 1 dµ(t) × 1 t1− 2 fεAp(U) δ p δ εδε−2gpε,δLp(dA) + ( 2 p + ε)(1/δ)ε+1gp(ε+1),δLp(dA) fεAp(U) → 0 as ε → 0. This and (6) imply that δ Z 1 δ 1 t1− 2 p dµ(t) = Hδ µAp(U)→Ap(U). (cid:3) Theorem 3.7. Let 1 ≤ p < ∞. If 1 t1− 2 Z ∞ 0 p dµ(t) < ∞ then Proof. By Theorem 3.5 we have that HµAp(U)→Ap(U) =Z ∞ HµAp(U)→Ap(U) ≤Z ∞ 0 0 1 t1− 2 p 1 t1− 2 p dµ(t). dµ(t). HAUSDORFF OPERATORS ON BERGMAN SPACES OF THE UPPER HALF PLANE11 Minkowski inequality implies that (7) Hµ − Hδ µAp(U)→Ap(U) ≤Z(0,δ)∪( 1 δ ,∞) 1 t1− 2 p dµ(t). By Proposition 3.6 µAp(U)→Ap(U). This, combined with (7), allows us to conclude that δ p 1 t1− 2 δ Z 1 HµAp(U)→Ap(U) ≥Z ∞ 0 dµ(t) = Hδ dµ(t)−2Z(0,δ)∪(1/δ,∞) p 1 t1− 2 1 t1− 2 p as δ → 0. Hence, HµAp(U)→Ap(U) =Z ∞ 0 1 t1− 2 p dµ(t). dµ(t) →Z ∞ 0 1 t1− 2 p dµ(t) (cid:3) 3.3. The quasi-Hausdorff operator. Let f, g ∈ A2(U) and assume that Hµ is bounded on A2(U). Thus Z ∞ 0 1 t1− 2 p dµ(t) < ∞. We have ZUZ ∞ 0 1 t(cid:17)(cid:12)(cid:12)(cid:12)g(z) dµ(t)dA(z) ≤ ZU(cid:18)Z ∞ t(cid:12)(cid:12)(cid:12)f(cid:16)z ≤(cid:18)Z ∞ 1 t1− 2 p 0 0 1 t(cid:17)(cid:12)(cid:12)(cid:12) dµ(t)(cid:19)2 t(cid:12)(cid:12)(cid:12)f(cid:16)z dµ(t)(cid:19)1/2 fA2(U)gA2(U) < ∞, dA(z)!1/2 gA2(U) where we applied the Cauchy-Schwarz and Minkowski inequalities. There- fore hHµ(f ), gi =ZU Hµ(f )(z)g(z)dA(z) = = = z t z t 0 1 1 f ( 1 t 1 t πZUZ ∞ πZ ∞ 0 ZU πZU f (z)Z ∞ f ( 1 0 ) dµ(t)g(z)dA(z) )g(z) dA(z) dµ(t) tg(tz) dµ(t) dA(z), where we applied a change of variables and Fubini's Theorem twice. This means that the adjoint H∗µ of Hµ on A2(U) is: 12 GEORGIOS STYLOGIANNIS H∗µ(f )(z) =Z ∞ 0 tf (tz) dµ(t). We will consider H∗µ on Ap(U) and suppose for a moment that it is well defined for functions in Ap(U). Let λ(t) = t−1, t > 0, then λ maps (0,∞) onto (0,∞) and is measurable. Set f (tz) = fz(t) then 0 H∗µ(f )(z) =Z ∞ =Z ∞ =Z ∞ =Z ∞ 0 0 0 tf (tz) dµ(t) tfz(t) dµ(t) 1 λ(t)(cid:19) dµ(t) fz(cid:18) 1 t(cid:17) dν(t), f(cid:16)z λ(t) 1 t = Hν(f )(z) ∗ ∗ (µ)(t) and λ where dν = dλ (µ) is the push-forward measure of µ with respect to λ. We can now apply the results of the first part of the paper to have: Theorem 3.8. Let 1 ≤ p < ∞. The quasi-Hausdorff operator H∗µ is bounded on Ap(U) if and only if Moreover Z ∞ 0 t1− 2 p dµ(t) < ∞. H∗µAp(U)→Ap(U) =Z ∞ 0 t1− 2 p dµ(t). References [An03] K. F. Andersen, Boundedness of Hausdorff operators on Lp(Rn), H 1(Rn), and BM O(Rn). Acta Sci. Math. (Szeged) 69 (2003), no. 1-2, 409-418. [BaGo19] R. Bandaliyev, and P. G´orka, Hausdorff operator in Lebesgue spaces. Math. Inequal. Appl. 22 (2019), no. 2, 657-676. [BaBoMiMi16] S. Ballamoole, J. O. Bonyo, T. L. Miller and V. G. Miller, Ces´aro-like operators on the Hardy and Bergman spaces of the half plane. Complex Anal. Oper. Theory 10 (2016), no. 1, 187-203. [ChDaFa18] J. Chen, J. Dai, D. Fan and X. Zhu, Boundedness of Hausdorff operators on Lebesgue spaces and Hardy spaces. Sci. China Math. 61 (2018), no. 9, 1647-1664. [ChFaZh16] J. Chen, D. Fan and X. Zhu, The Hausdorff operator on the Hardy space H 1(R1). Acta Math. Hungar. 150 (2016), no. 1, 142-152. HAUSDORFF OPERATORS ON BERGMAN SPACES OF THE UPPER HALF PLANE13 [ChKoSm17] Choe B. R., Koo H. and Smith W. Difference of composition opera- tors over the half-plane. Trans. Amer. Math. Soc. 369 (2017), no. 5, 3173-3205. [FaLi14] D. Fan and X. Lin, Hausdorff operator on real Hardy spaces. Analysis (Berlin) 34 (2014), no. 4, 319-337. [GaSi01] P. Galanopoulos and A. G. Siskakis, Hausdorff matrices and composition operators. Illinois J. Math. 45 (2001), no. 3, 757-773. [GaPa06] P. Galanopoulos and M. Papadimitrakis, Hausdorff and quasi-Hausdorff matrices on spaces of analytic functions. Canad. J. Math. 58 (2006), no. 3, 548-579. [Ga07] J. B. Garnett, Bounded analytic functions. Revised first edition. Graduate Texts in Mathematics, 236. Springer, New York, 2007. [Ge92] C. Georgakis, The Hausdorff mean of a Fourier-Stieltjes transform. Proc. Amer. Math. Soc. 116 (1992), no. 2, 465-471. [GiMo95] D. V. Giang and F. M´oricz, The Ces´aro operator is bounded on the Hardy space H 1. Acta Sci. Math. (Szeged) 61 (1995), no. 1-4, 535-544. [Go59] R. R. Goldberg, Averages of Fourier coefficients. Pacific J. Math. 9 (1959), 695-699. [Go98] B. I. Golubov, Boundedness of the Hardy and Hardy-Littlewood operators in the spaces ReH 1 and BMO (in Russian). Mat. Sb. 188(1997), 93-106. - English transl. in Russian Acad. Sci. Sb. Math. 86 (1998). [Ha21] F. Hausdorff, Summationsmethoden und Momentfolgen I. Math. Z. 9(1921), 74-109. [HuSi17] W. A. Hurwitz and L. L. Silverman, The consistency and equivalence of certain definitions of summability. Trans. Amer. Math. Soc. 18(1917), 1-20. [HuKyQu18] H. D. Hung, L. D. Ky, and T. T. Quang, Norm of the Hausdorff operator on the real Hardy space H 1(R). Complex Anal. Oper. Theory 12 (2018), no. 1, 235-245. [Ka01] Y. Kanjin, The Hausdorff operators on the real Hardy spaces Hp(R). Studia Math. 148 (2001), no. 1, 37-45. [Ko80] P. Koosis, Introduction to H p spaces. With an appendix on Wolff's proof of the corona theorem. London Mathematical Society Lecture Note Series, 40. Cambridge University Press, Cambridge-New York, 1980. [Li013] E. Liflyand, Hausdorff operators on Hardy spaces. Eurasian Math. J. 4 (2013), no. 4, 101-141. [LiMi09] E. Liflyand and A. Miyachi, Boundedness of the Hausdorff operators in H p spaces, 0 < p < 1, Boundedness of the Hausdorff operators in H p spaces, 14 GEORGIOS STYLOGIANNIS 0 < p < 1. Studia Math. 194 (2009), no. 3, 279-292. [LiMo02] E. Liflyand, F. M´oricz, Commuting relations for Hausdorff operators and Hilbert transforms on real Hardy spaces. Acta Math. Hungar. 97 (2002), no. 1-2, 133-143. [LiMo01] E. Liflyand, F. M´oricz, The multi-parameter Hausdorff operator is bounded on the product Hardy space H 11(R × R). Analysis (Munich) 21 (2001), no. 2, 107-118. [LiMo00] E. Liflyand and F. Moricz, The Hausdorff operator is bounded on the real Hardy space H 1(R), Proc. Amer. Math. Soc. 128 (2000), no. 5, 1391-1396. [Mo05] F. M´oricz, Multivariate Hausdorff operators on the spaces H 1(Rn) and BM O(Rn). Anal. Math. 31 (2005), no. 1, 31-41. [RuFa16] J. Ruan and D. Fan, Hausdorff operators on the power weighted Hardy spaces. J. Math. Anal. Appl. 433 (2016), no. 1, 31-48. [Si87] A. G. Siskakis, Composition operators and the Ces´aro operator on H p. J. London Math. Soc. (2) 36 (1987), no. 1, 153-164. [Si90] A. G. Siskakis, The Ces´aro operator is bounded on H 1. Proc. Amer. Math. Soc. 110 (1990), no. 2, 461-462. [Zh90] K. H. Zhu, Operator theory in function spaces. Monographs and Textbooks in Pure and Applied Mathematics, 139. Marcel Dekker, Inc., New York, 1990.
1604.06341
1
1604
2016-04-21T14:53:19
Bochner integrals in ordered vector spaces
[ "math.FA" ]
We present a natural way to cover an Archimedean directed ordered vector space $E$ by Banach spaces and extend the notion of Bochner integrability to functions with values in $E$. The resulting set of integrable functions is an Archimedean directed ordered vector space and the integral is an order preserving map.
math.FA
math
Bochner integrals in ordered vector spaces A.C.M. van Rooij I W.B. van Zuijlen II 28th February 2018 Abstract We present a natural way to cover an Archimedean directed ordered vector space E by Banach spaces and extend the notion of Bochner integrability to functions with values in E. The resulting set of integrable functions is an Archimedean directed ordered vector space and the integral is an order preserving map. Mathematics Subject Classification (2010). 28B05, 28B15. Keywords: Bochner integral, ordered vector space, ordered Banach space, closed cone, generating cone. 1 Introduction We extend the notion of Bochner integrability to functions with values in a vector space E that may not itself be a Banach space but is the union of a collection B of Banach spaces. The idea is the following. We call a function f , defined on a measure space X and with values in E, "integrable" if for some D in B all values of f lie in D and f is Bochner integrable as a function X → D. Of course, one wants a certain consistency: the "integral" of such an f should be independent of the choice of D. In [17], Thomas obtains this consistency by assuming a Hausdorff locally convex topology on E, entailing many continuous linear functions E → R. Their restrictions to the Banach spaces that constitute B enable one to apply Pettis integration, which leads to the desired uniqueness. Our approach is different, following a direct-limit-like construction. We assume E to be an ordered vector space with some simple regularity properties (Archimedean, directed) and show that E is the union of a certain increasing system B of Banach spaces with closed, generating positive cones (under the ordering of E). Uniqueness of the integral follows from properties of such ordered Banach spaces. Moreover, the integrable functions form a vector space and the integral is linear and order preserving. IRadboud University Nijmegen, Department of Mathematics, P.O. Box 9010, 6500 GL Nijmegen, the Netherlands. IILeiden University, Mathematical Institute, P.O. Box 9512, 2300 RA, Leiden, the Netherlands. 1 In Section 3 we study ordered Banach spaces with closed generating cones. We give certain properties which can be used to give an alternative proof of a classical theorem which states that every order preserving linear map is continuous, and generalise it to order bounded linear maps. In Section 4 we study Bochner integrable functions with values in an ordered Banach space with closed generating cone. In Section 5 we present the definition of a Banach cover and the definition of the extension of the Bochner integral to functions with values in a vector space that admits a Banach cover. In Section 6 we show that an Archimedean ordered vector space possesses a Banach cover consisting of ordered Banach spaces spaces with closed generating cones. In Section 7 we study integrable functions with values in Archimedean ordered vector spaces. In Section 8 we compare the integral with integrals considered in [15]. In Section 9 we present an application to view the convolution as an integral. 2 Notation N is {1, 2, . . . }. We write "for all n" instead of "for all n ∈ N". To avoid confusion: • An "order" is a "partial order". • We call an ordered vector space Archimedean (see Peressini [14]) if for all a, b ∈ E if na ≤ b for all n ∈ N, then a ≤ 0. (In some places, e.g., the following holds: Birkhoff [4], such spaces are said to be 'integrally closed'.) As is common in literature, our notations do not distinguish between a function on a measure space and the class of that function. 3 Ordered Banach spaces with closed generating cones In this section we describe properties of ordered Banach spaces with closed generating cones. Using these properties we prove in Theorem 3.11 that an order bounded map between ordered Banach spaces with closed generating cones is continuous. 3.1 Definition A normed ordered vector space is a normed vector space with an order that makes it an ordered vector space. An ordered Banach space is a Banach space that is a normed ordered vector space. A priori there is no connection between the ordering and the norm of a normed ordered vector space. One reasonable and useful connection is the assumption that the (positive) cone be closed. 3.2 Theorem Let E be a normed ordered vector space. E+ is closed if and only if x ≤ y ⇐⇒ α(x) ≤ α(y) for all α ∈ (E′)+. (1) Consequently, whenever E+ is closed then (E′)+ separates the points of E and E is Archimedean. 2 Proof. Since E+ is convex, E+ is closed if and only if it is weakly closed (i.e., σ(E, E′)- closed); see [7, Theorem V.1.4]. The rest follows by [1, Theorem 2.13(3&4)]. The following theorem is due to Andô [2]. See also [1, Corollary 2.12]. 3.3 Theorem Let D be an ordered Banach space with a closed generating1 cone D+. There exists a C > 0 such that Ckxk ≥ inf{kak : a ∈ D+, −a ≤ x ≤ a} (x ∈ E). (2) 3.4 Definition Let D be a directed2 ordered Banach space. If C > 0 is such that (2) holds, then we say that the norm k · k is C-absolutely dominating.3 We say that a norm k · k is absolutely dominating if it is C-absolutely dominating for some C > 0. On a Banach lattice the norm is 1-absolutely dominating. Actually for Banach lattices there is equality in (2). 3.5 We refer the reader to Appendix A for the following facts: If k · k is C-absolutely dominating on a directed ordered Banach space D, then C ≥ 1. Whenever there exists a absolutely dominating norm, then for all ε > 0 there exists an equivalent (1 + ε)- absolutely dominating norm. All norms on a directed ordered vector space D that make D complete and D+ closed are equivalent (see 6.2). 3.6 Let D be a directed ordered Banach space. Then k · k is absolutely dominating if and only if the (convex) set [−a, a] [a∈D+,kak≤1 (3) is a neighbourhood of 0. 3.7 Definition (See [12, §16]) Let E be an ordered vector space. We say that a sequence u−→ x) whenever (xn)n∈N in E converges uniformly to an element x ∈ E (notation: xn there exist a ∈ E+, εn ∈ (0, ∞) with εn → 0 and −εna ≤ xn − x ≤ εna (n ∈ N). (4) Note that one may replace "εn → 0" by "εn ↓ 0". If E is Archimedean then the x as above is unique. We will only consider such convergence in Archimedean spaces. We say that a sequence (xn)n∈N in E is a uniformly Cauchy sequence if there exists an a ∈ E+ such that for all ε > 0 there exists an N such that −εa ≤ xn − xm ≤ εa for all n, m ≥ N . E is called uniformly complete whenever it is Archimedean and all uniformly Cauchy sequences converge uniformly. 1D+ is generating if D = D+ 2D is directed if D = D+ 3Batty and Robinson [3] call the cone D+ approximately C-absolutely dominating, and, Messer- schmidt [13] calls D approximately C-absolutely conormal, if the norm on D is C-absolutely dominating. − D+, i.e., if D+ is generating. − D+. 3 3.8 Lemma Let D1, D2 be ordered vector spaces and T : D1 → D2. If T is linear and order bounded, then T preserves uniform convergence. Proof. Suppose xn ∈ D1, εn ∈ (0, 1) and a ∈ D+ and εn ↓ 0. Let b ∈ D+ −εnb ≤ T xn ≤ εnb for all n. 2 be such that T [−a, a] ⊂ [−b, b]. Then ε−1 1 are such that −εna ≤ xn ≤ εna n T xn ∈ [−b, b], i.e., 3.9 Theorem Let D be an ordered Banach space. Consider the following conditions. (i) k · k is absolutely dominating. εn → 0 such that an ≤ εna for all n.4 (ii) If a1, a2, · · · ∈ D+, Pn∈N kank < ∞, then there exist a ∈ D+, εn ∈ (0, ∞) with (iii) If x1, x2, · · · ∈ D,Pn∈N kxnk < ∞, then xn If D+ is closed, then D satisfies (ii). Suppose D is Archimedean and directed. Then the following are equivalent. u−→ 0. (a) D+ is closed. (b) D satisfies (i) and (ii). (c) D satisfies (iii). exists. Because D+ is closed a ≥ ε−1 (a)=⇒(b). (ii) is implied by the above argument. By Theorem 3.3 we have (i). Proof. Suppose D+ is closed. Let a1, a2, · · · ∈ D+, Pn∈N kank < ∞. Choose εn ∈ (0, ∞), εn → 0 such thatPn∈N ε−1 n kank < ∞. By norm completeness, a :=Pn∈N ε−1 (b)=⇒(c). Let x1, x2, · · · ∈ D, Pn∈N kxnk < ∞. Using (i) let a1, a2, · · · ∈ D+ with Pn∈N kank < ∞ be such that −an ≤ xn ≤ an. By (ii) it then follows that xn (c)=⇒ (a). Take b in the closure of D+. For n ∈ N, choose xn ∈ D+,Pn∈N kxn−bk < ∞. By (iii) there exist a ∈ D+, εn ∈ (0, ∞) with εn → 0 and −εna ≤ xn − b ≤ εna, so that b ≥ xn − εna ≥ −εna. Then b ≥ 0 because D is Archimedean. n an for all n. u−→ 0. n an 3.10 Lemma Let D be an ordered Banach space for which D+ is closed. Then x1, x2, · · · ∈ D, a ∈ D, b ∈ D, kxn − ak → 0, xn u−→ b =⇒ a = b. (5) Proof. Assume b = 0. There exist c ∈ D+, εn ∈ (0, ∞) with εn ↓ 0 and −εnc ≤ xn ≤ εnc. Then −εN c ≤ xn ≤ εN c whenever n ≥ N . Since D+ is closed, −εN c ≤ a ≤ εN c for all N , and so, as D is Archimedean by Theorem 3.2, a = 0. With this we can easily prove the following theorem. In a recently published book by Aliprantis and Tourky [1]5 but also in older Russian papers by Wulich [5] one can find the proof that an order preserving linear map is continuous (see [1, Theorem 2.32] 4For normed Riesz spaces property (ii) is equivalent to what is called the weak Riesz-Fischer condition (see Zaanen [18, Ch. 14, §101]). 5For complete metrisable ordered vector spaces. 4 or combine [5, Theorem III.2.2] (which states the result for D2 = R) and [6, Theorem VI.2.1]). Theorem 3.11 is more general in the sense that it states that linear order bounded maps are continuous. 3.11 Theorem Let D1, D2 be ordered Banach spaces. Suppose D+ erating and D+ continuous. Consequently, if T is an order isomorphism then it is a homeomorphism. 1 is closed and gen- 2 is closed. Let T : D1 → D2 be linear and order bounded. Then T is Proof. Let x1, x2, · · · ∈ D1, xn → 0 and suppose T xn → c for some c ∈ D2. If from this we can prove c = 0, then by the Closed Graph Theorem T will be continuous. We u−→ 0 in D2 u−→ 0 in D1 (Theorem 3.9), so T xn may assume Pn∈N kxnk < ∞. Then xn (Lemma 3.8). Hence, c = 0 according to Lemma 3.10. We present a consequence as has been done for order preserving linear maps in [6, Theorem VI.2.2]. 3.12 Corollary Let D1 be an ordered Banach space with closed generating cone. Let D2 be an ordered normed vector space with a normal cone. Let T : D1 → D2 be linear and order bounded. Then T is continuous. 2 = D′ 2 is normal, we have D∼ Proof. As D+ 2 (see [1, Corollary 2.27]). Let ∆ be the unit ball in D1. Then for all φ ∈ D′ 2, the map φ ◦ T is an order bounded functional, whence continuous by Theorem 3.11. Thus φ ◦ T (∆) is bounded for all φ ∈ D′ 2. By an application of the Principle of Uniform Boundedness (see [7, Corollary III.4.3]), T (∆) is norm bounded. 4 The Bochner integral on ordered Banach spaces In this section (X, A, µ) is a complete σ-finite measure space, with µ 6= 0. For more assumptions see 4.3 and 4.7. We define the integral of simple functions in Definition 4.1 and recall the definition and facts on Bochner integrability in Definition 4.2 and 4.4. After that, we consider an ordered Banach space D. In Theorems 4.5 and 4.6 we describe the order structure of the space BD of Bochner integrable functions. In 4.8 we summarise the results of 4.9 – 4.16, in which we compare closedness and generatingness of the positive cones of D and BD. 4.1 Definition Let E be a vector space. We say that a function f : X → E is simple if there exist N ∈ N, a1, . . . , aN ∈ E, A1, . . . , AN ∈ A with µ(A1), . . . , µ(AN ) < ∞ for which f = N Xn=1 an1An. 5 (6) The simple functions form a linear subspace S of EX , which is a Riesz subspace of EX in case E is a Riesz space. We define ϕ : S → E by ϕ(f ) = N Xn=1 µ(An)an, (7) where f, N, An, an are as in (6). ϕ(f ) is called the integral of f . We write SR for the linear space of simple functions X → R. Definition 4.2 and the facts in 4.4 can be found in Chapter III in the book by E. Hille and R.S. Phillips, [9]. 4.2 Definition Let (D, k · k) be a Banach space. A function f : X → D is called Bochner integrable whenever there exists a sequence of simple functions (sn)n∈N such that Z kf (x) − sn(x)k dµ(x) → 0. (8) Then the sequence (ϕ(sn))n∈N converges. Its limit is independent of the choice of the sequence (sn)n∈N and is called the Bochner integral of f . 4.3 For the rest of this section, D is a Banach space (with norm k · k), and, we write B (or BD) for the Banach space of classes of Bochner integrable functions X → D, with norm k · kB (see 4.4(b)). We write b (or bD) for the Bochner integral on B. 4.4 Some facts on the Bochner integrable functions: (a) [9, Theorem 3.7.4] & [16, Proposition 2.15] If f : X → D is Borel measurable, f (X) is separable and R kf k dµ < ∞, then f is Bochner integrable. (b) [9, Theorem 3.7.8] B is a Banach space under the norm k · kB : B → [0, ∞) given by kf kB =Z kf k dµ. (9) (c) [9, Theorem 3.7.12] Let E be a Banach space and T : D → E be linear and continuous. If f ∈ BD, then T ◦ f ∈ BE and T (bD(f )) = bE(T ◦ f ). (10) (d) [9, Theorem 3.7.9] Let f1, f2, . . . be in B, f : X → D and h ∈ L1(µ)+. Suppose that fn(x) → f (x) and kfn(x)k ≤ h(x) for µ-almost all x ∈ X. Then f ∈ B and b(fn) → b(f ). (11) 6 (e) If D is an ordered Banach space then B is an ordered Banach space under the ordering given by f ≤ g (in B) ⇐⇒ f ≤ g µ − a.e.. (12) 4.5 Theorem Let D be a Banach lattice. Then BD is a Banach lattice and b is linear and order preserving. Proof. The Bochner integrable functions form a Riesz space because of the inequality kx − yk ≤ kx − yk. 4.6 Theorem Let D be an ordered Banach space for which D+ is closed. Then b is order preserving. Proof. Let f ∈ B and f ≥ 0. Then α(b(f )) =Z α ◦ f dµ ≥ 0 (α ∈ (D′)+). (13) Whence b(f ) ≥ 0, by Theorem 3.2. 4.7 For the rest of this section, D is an ordered Banach space. 4.8 The following is a list of results presented in 4.9 – 4.16. (a) B+ is closed if and only if D+ is (4.9). (b) If B is directed, then so is D (straightforward, see also Theorem 4.12). (c) Let C > 0 and D+ be closed and generating. If k · k is C-absolutely dominating, then so is k · kB (Lemma 4.11). (d) Let C > 0. If B is directed and k · kB is C-absolutely dominating, then so is k · k (Theorem 4.12). (e) B+ is closed and generating if and only if D+ is closed and generating (Theorem 4.16). (f) If there exist disjoint A1, A2, . . . in A with 0 < µ(A) < ∞ for all n: If B+ is generating, then so is D+ and k · k is absolutely dominating (Corollary 4.14). (g) If no such A1, A2, . . . exist: B+ is generating if and only if D+ is generating (4.15). 4.9 Whenever f ∈ B, fn ∈ B+ withR kf − fnk dµ → 0, then there exist Bochner integ- rable gn ≥ 0 with Pn∈NR kf − gnk dµ < ∞; this implies gn → f µ-almost everywhere. So whenever D+ is closed this implies f ≥ 0 µ-almost everywhere. We infer that B+ is closed whenever D+ is. On the other hand, if B+ is closed then so is D+. Indeed, let A ∈ A, 0 < µ(A) < ∞. If an ∈ D+ and an → a, then ka1A − an1AkB → 0. Therefore a ∈ D+. 7 4.10 Lemma Suppose D+ is generating and C > 0 is such that k · k is C-absolutely dominating. Let f : X → D be simple, let ε > 0. Then there exists a simple g : X → D+ with −g ≤ f ≤ g and R kgk dµ ≤ CR kf k dµ + ε. Proof. Write f =PN x1, . . . , xN ∈ D. Let κ = µ(SN R kgk dµ =PN n=1 kankµ(An) ≤PN −an ≤ xn ≤ an, kank ≤ Ckxnk + ε κ . Put g = PN n=1(Ckxnk + ε n=1 xn1An with disjoint sets A1, . . . , AN ∈ A of finite measure and n=1 An) and assume κ > 0. For each n, choose an ∈ D+, n=1 an1An. Then −g ≤ f ≤ g and κ )µ(An) = CR kf k dµ + ε. 4.11 Lemma Suppose D+ is closed and generating and C > 0. Then B+ is closed and generating. If k · k is C-absolutely dominating, then so is k · kB. Proof. B+ is closed by 4.9. Assume that k · k is C-absolutely dominating. Let f : X → D be Bochner integrable and let ε > 0. We prove there exists a Bochner integrable g : X → D+ with −g ≤ f ≤ g µ-a.e. and R kgk dµ ≤ C(R kf k dµ + ε). Choose simple functions s1, s2, . . . with R kf − snk dµ < ε2−n−1. Then sn → f pointwise outside a µ-null set Y . Define f1 = s1, fn = sn − sn−1 for n ∈ {2, 3, . . . }. Observe Z kf1k dµ <Z kf k dµ + ε2−2, Z kfnk dµ < ε2−n−1 + ε2−n−2 < ε2−n (n ∈ {2, 3, . . . }). (14) For each n, choose a simple gn : X → D+ with −gn ≤ fn ≤ gn such that Z kg1k dµ < C(cid:18)Z kf k dµ + ε2−2(cid:19) , Z kgnk dµ < Cε2−n (n ∈ {2, 3, . . . }). As Pn∈NR kgnk dµ < ∞ there is a µ-null set Z ⊂ X for which kgn(x)k < ∞ (x ∈ X \ Z). Xn∈N Put X0 := X \ (Y ∪ Z). Define g : X → D+ by g(x) =(Pn∈N gn(x) x ∈ X0, x /∈ X0. 0 (15) (16) (17) Then g is Bochner integrable and R kgk dµ ≤Pn∈NR kgnk dµ ≤ C(R kf k dµ + ε). Moreover, for x ∈ X0 we have −g(x) ≤ PN n=1 fn(x) = sN (x) ≤ g(x) for all N , whereas sN (x) → f (x) since x /∈ Y . From the closedness of D+ it follows that −g ≤ f ≤ g on X0. 8 In the following theorems (4.12, 4.13, 4.14 and 4.15) we derive properties of D from properties of B. In Theorem 4.16 we show that D has a closed and generating cone if and only if B does. 4.12 Theorem Assume B is directed. Let C > 0 and suppose k · kB is C-absolutely dominating. Then D is directed and k · k is C-absolutely dominating. Proof. Let x ∈ D and x 6= 0. Let A ∈ A with 0 < µ(A) < ∞. Then x1A ∈ B, kx1AkB = µ(A)kxk. Let C ′ > C. There is a g ∈ B with −g ≤ x1A ≤ g, R kg(t)k dµ(t) ≤ C ′kx1AkB. Then RA kg(t)k dµ(t) ≤ C ′µ(A)kxk, so µ{t ∈ A : kg(t)k > C ′kxk} < µ(A). In particular, there is a t ∈ A with kg(t)k ≤ C ′kxk and −g(t) ≤ x ≤ g(t). 4.13 Theorem Let D be an ordered Banach space such that the Bochner integrable functions N → D form a directed space. Then D is directed and k · k is absolutely dominating. Proof. D is directed. In case k·k is not absolutely dominating, there exist x1, x2, · · · ∈ D such that for every n 2nkxnk ≤ inf{kak : a ∈ D+, −a ≤ xn ≤ a} and kxnk = 2−n. Then n 7→ xn is Bochner integrable, so, by our assumption, there exist an ∈ D+ with −an ≤ xn ≤ an for all n and Pn∈N kank < ∞ which is false. 4.14 Corollary Suppose there exist disjoint A1, A2, . . . in A with 0 < µ(An) < ∞ for all n. Suppose B is directed. Then D is directed and k · k is absolutely dominating. Proof. This follows from Theorem 4.13 since f 7→ Pn∈N f (n)1An forms an isometric order preserving isomorphism from the Bochner integrable functions N → D into B. 4.15 Whenever there do not exist A1, A2, . . . as in Corollary 4.14, then X = A1 ∪ · · · ∪ AN , where A1, . . . , AN are disjoint atoms. Let αn = µ(An) ∈ (0, ∞). Define a norm n=1 xn1An) = (α1x1, . . . , αN xN ), is an isometric order preserving isomorphism. Therefore, k · kN on DN by kxkN =PN n=1 kxnk. Then T : B → DN defined by T (PN D+ is generating ⇐⇒ (DN )+ is generating ⇐⇒ B+ is generating. (18) 4.16 Theorem D+ is closed and generating if and only if B+ is. Moreover, if D+ is closed and generating and C > 0, then k · k is C-absolutely dominating if and only if k · kB is. Proof. This follows from Lemma 4.11 and Theorem 4.12. 9 5 An extension of the Bochner integral We present the definition of a Banach cover (Definition 5.2), some examples, and use this notion to extend the Bochner integral to functions with values in such a vector space (5.6). The next result follows by definition of the Bochner integral or by 4.4(c). 5.1 Theorem Let D1 and D be Banach spaces and suppose D1 ⊂ D and the inclusion map is continuous. Suppose f : X → D has values in D1 and is Bochner integrable as map X → D1 with integral I. Then f is Bochner integrable as a map X → D with integral I. 5.2 Definition Let E be a vector space. Suppose B is a collection of Banach spaces whose underlying vector spaces are linear subspaces of E. B is called a Banach cover of E if S B = E and for all D1 and D2 in B there exists a D ∈ B with D1, D2 ⊂ D such that both inclusion maps D1 → D and D2 → D are continuous. For a D ∈ B, we write k · kD for its norm if not indicated otherwise. If E is an ordered vector space, then an ordered Banach cover is a Banach cover whose elements are seen as ordered subspaces of E. 5.3 A bit of pedantry: strictly speaking, a Banach space is a couple (D, k · k) consisting of a vector space D and a norm k · k. One usually talks about the "Banach space D", the norm being understood. Mostly, we adopt that convention but not always. In the context of the above definitions one has to be careful. A Banach cover may contain several Banach spaces with the same underlying vector space, so that a formula like "D1, D2 ⊂ D" really is ambiguous. What we mean is only an inclusion relation between the vector spaces and no connection between the norms is assumed a proiri. However, suppose (D1, k · k1) and (D2, k · k2) are elements of a Banach cover B and D1 = D2. There is a Banach space (D, k·kD) in B with D1, D2 ⊂ D and with continuous inclusion maps. If a sequence (xn)n∈N in D1(= D2) is k · k1-convergent to a and k · k2- convergent to b, then it is k · kD-convergent to a and b, so a = b. Hence, by the Closed Graph Theorem the norms k · k1 and k · k2 are equivalent. Similarly, if (D1, k · k1) and (D2, k · k2) are elements of a Banach cover and D1 ⊂ D2, then the inclusion map D1 → D2 automatically is k · k1-k · k2-continuous. 5.4 Example Let E be a uniformly complete Riesz space. The set of principal ideals B = {(Eu, k · ku) : u ∈ E+, u 6= 0} is an ordered Banach cover of E: for u, v ∈ E+ with u, v > 0 one has Eu, Ev ⊂ Eu+v and k · ku+v ≤ k · ku on Eu. 5.5 Example Let (Y, B, ν) be a complete σ-finite measure space. Let M be the space of classes of measurable functions Y → R. A function ρ : M → [0, ∞] is called an function norm if (i) ρ(f ) = 0 ⇐⇒ f = 0 a.e., (ii) ρ(αf ) = αρ(f ) (where 0 · ∞ = 0), (iii) ρ(f ) = ρ(f ), (iv) ρ(f + g) ≤ ρ(f ) + ρ(g), (v) 0 ≤ f ≤ g a.e. implies ρ(f ) ≤ ρ(g) for f, g ∈ M and α ∈ R. For such function norm ρ the set Lρ = {f ∈ M : ρ(f ) < ∞} (19) 10 is a normed Riesz space called a Köthe space (see [11, Ch. III §18] or [12, Ch. 1 §9]). A Köthe space Lρ is complete, i.e., a Banach lattice if and only if for all un ∈ L+ ρ with Pn∈N ρ(un) < ∞, Pn∈N un is an element of Lρ ([11, Theorem 19.3]). Examples of complete Köthe spaces are Orlicz spaces ([11, §20]). In particular, the Banach spaces Lp(ν) for p ≥ 1 are Köthe spaces. We introduce other examples: For w ∈ M + with w > 0 a.e. define ρw : M → [0, ∞] by ρw(f ) :=Z f w dν. (20) Then ρw is a function norm. Both the set {Lρw : w ∈ M +, w > 0 a.e.} and the set of all complete Köthe spaces are Banach covers of M (see Appendix B). 5.6 Let E be a vector space with a Banach cover B. Let (X, A, µ) be a complete σ-finite measure space, µ 6= 0. (1) For D ⊂ B denote by BD the vector space of all Bochner integrable functions X → D, and, by bD the Bochner integral BD → D. (2) Let D1, D2 ∈ B, f1 ∈ BD1, f2 ∈ BD2, f1 = f2 µ-a.e. Then bD1(f1) = bD2(f2). Proof. Choose D ∈ B as in Definition 5.2. Then f1, f2 ∈ BD and bD1(f1) = bD(f1) = bD(f2) = bD2(f2). (3) If D1, D2 ∈ B have the same underlying vector space, then BD1 = BD2 since the identity map D1 → D2 is a homeomorphism (see 5.3). (4) We call a function f : X → E B-integrable if there is a D ∈ B such that f is µ-a.e. equal to some element of BD. (5) By U we indicate the vector space of all µ-equivalence classes of B-integrable func- tions. For D ∈ B we have a natural map TD : BD → U, assigning to every element of BD its µ-equivalence class. The space TD(BD) is a Banach space, which we indicate by BD.6 We write bD for the map BD → D determined by bD(TD(f )) = bD(f ) (f ∈ BD). U is the union of the sets BD. By (2) there is a unique u : U → E determined by u(f ) = bD(f ) (D ∈ B, f ∈ BD). The above leads to the following theorem. (21) (22) 5.7 Theorem U is a vector space, u is linear and {BD : D ∈ B} is a Banach cover of U. 5.8 6Even though we use the same notation as in Section 4, see 4.3, the meaning of BD is slightly different. 11 (1) For any vector space the finite dimensional linear subspaces form a Banach cover. (2) If E is a Banach space, then {E} is a Banach cover. (3) If E is a Banach space and B is a Banach cover with E ∈ B, then U is just the space of (classes of) Bochner integrable functions X → E and u is the Bochner integral. (4) A special case of (3): Let E be a uniformly complete Riesz space with a unit e and let B be the Banach cover of principal ideals as in Example 5.4. Then (E, k·ke) ∈ B. 5.9 Let E be a vector space and B1 and B2 be Banach covers of E. Suppose that for all D1 ∈ B1 there exists a D2 ∈ B2 with D1 ⊂ D2 such that the inclusion map is continuous. Write Ui, ui for the set of Bi-integrable functions and the Bi-integral, for i ∈ {1, 2}. Then U1 ⊂ U2 and u1 = u2 on U1. 5.10 Example (Different covers and different integrals) Let (E, k·k) be an infin- ite dimensional Banach space. Let T : E → E a linear bijection that is not continuous; say, there exist xn ∈ E with kxnk = 2−n, T xn = xn, T (Pn∈N xn) 6= Pn∈N xn. Define kxkT = kT xk for x ∈ E. Then (E, k · kT ) is a Banach space. The map f : N → E given by f (n) = xn is Bochner integrable in (E, k · k) and in (E, k · kT ), but the integrals do not agree. Whence with B1 = {(E, k · k)} and B2 = {(E, k · kT )} we have f ∈ U1 ∩ U2 but u1(f ) 6= u2(f ). As an immediate consequence of Theorem 4.5 we obtain the following theorem. 5.11 Theorem Suppose E is a Riesz space and B is a Banach cover of E that consists of Banach lattices that are Riesz subspaces of E. Then U is a Riesz space and u is order preserving. Proof. This is a consequence of Theorem 4.5. 5.12 Whenever E is an ordered vector space and B an ordered Banach cover of E, then U is an ordered vector space. In order for u to be order preserving, one needs a condition on B. This and other matters will be treated in §7. A sufficient condition turns out to be closedness of D+ for every D ∈ B (see Theorem 4.6 and Theorem 7.1). First we will see in §6 that all Archimedean directed ordered vector spaces admit such ordered Banach covers. (The Archimedean property is necessary as follows easily from Theorem 3.2). 5.13 Whenever E is a vector space and B is a Banach cover of E, then the set {A ⊂ E : there exists a D ∈ B such that A is bounded in D} (23) forms a bornology on E (we refer to the book of Hogbe-Nlend [10] for the theory of bornologies). 12 6 Covers of ordered Banach spaces with closed generating cones In this section E is an Archimedean directed ordered vector space. B is the collection of all ordered Banach spaces that ordered linear subspaces of E whose cones are closed and generating. We intend to prove that B is a Banach cover of E (Theorem 6.5). 6.1 Lemma Let (D1, k · k1), (D2, k · k2) be in B. Let z1, z2, · · · ∈ D1 ∩ D2, a ∈ D1, b ∈ D2, kzn − ak1 → 0, kzn − bk2 → 0. Then a = b. Proof. We may assume Pn∈N kzn − ak1 < ∞ and Pn∈N kzn − bk2 < ∞. Then zn u−→ b in D2 by Theorem 3.9. Then zn u−→ a u−→ b in E. Because E is u−→ a and zn in D1 and zn Archimedean, a = b. 6.2 If D is an ordered Banach space with closed generating cone D+, under each of two norms k · k1 and k · k2, then these norms are equivalent. Indeed, the identity map (D, k · k1) → (D, k · k2) has a closed graph by Lemma 6.1. 6.3 Theorem Let (D1, k · k1), (D2, k · k2) be in B. D1 + D2 is an ordered Banach space with closed generating cone under the norm k · k : D1 + D2 → [0, ∞) defined by kzk := inf{kxk1 + kyk2 : x ∈ D1, y ∈ D2, z = x + y}. (24) Moreover, if C > 0 and k · k1 and k · k2 are C-absolutely dominating, then so is k · k. Proof. D1 × D2 is a Banach space under the norm (x, y) 7→ kxk1 + kyk2. From Lemma 6.1 it follows that ∆ := {(a, b) ∈ D1 × D2 : a = −b} is closed in D1 × D2. Then D1 × D2/∆ is a Banach space under the quotient norm. This means that D1 + D2 is a Banach space under k · k. (In particular, k · k is a norm.) Since D+ 2 ⊂ (D1 + D2)+, the latter is generating. 1 + D+ prove un u−→ 0 in D1 + D2. By Theorem 3.9 it follows that (D1 + D2)+ is closed. We prove that (D1 + D2)+ is closed. Let u1, u2, · · · ∈ D1 + D2, Pn∈N kunk < ∞; we u−→ 0. Choose xn ∈ D1, yn ∈ D2 with un = xn + yn and Pn∈N kxnk1 < ∞, Pn∈N kynk2 < ∞. Then, see Theorem 3.9, xn that un Suppose C > 0 is such that k·k1 and k·k2 are C-absolutely dominating. Let z ∈ D1 +D2, ε > 0. Choose x ∈ D1, y ∈ D2 with z = x + y, kxk1 + kyk2 ≤ kzk + ε 3 . Choose a ∈ D+ with −a ≤ x ≤ a, kak1 < Ckxk1 + ε 2 with −b ≤ y ≤ b, kbk2 < Ckyk2 + ε 3 . Set c = a + b. Then c ∈ (D1 + D2)+, −c ≤ z ≤ c and kck ≤ kak1 + kbk2 < Ckxk1 + Ckyk2 + 2 ε u−→ 0 in D2. It follows u−→ 0 in D1 and yn 3 and b ∈ D+ 1 3 < Ckzk + ε. 6.4 Let x ∈ E. (We make a D ∈ B with x ∈ D.) Choose a ∈ E such that −a ≤ x ≤ a. Let D = R(a − x) + R(a + x) = Ra + Rx. D is a directed ordered vector space. Define k · k : D → [0, ∞) by kyk := inf{s ≥ 0 : −sa ≤ y ≤ sa}. (25) 13 Then k · k is a norm on D, −kyka ≤ y ≤ kyka for all y ∈ D and kak = 1. Thus (D, k · k) is a directed ordered Banach space. Moreover D+ is closed: Let y ∈ D, y1, y2, · · · ∈ D+, ky − ynk < 1 n . Then y ≥ yn − 1 k·k is 1-absolutely dominating: kak = 1, so inf{kck : c ∈ D+, −c ≤ y ≤ c} ≤ inf{s ≥ n a ≥ − 1 n a, so y ≥ 0. 0 : −sa ≤ y ≤ sa} = kyk. Even kyk = inf{kck : c ∈ D+, −c ≤ y ≤ c}: For c ∈ D+ with −c ≤ y ≤ c and s ≥ 0 such that c ≤ sa we have −sa ≤ −c ≤ y ≤ c ≤ sa and so kyk ≤ kck. 6.5 Theorem B is a Banach cover of E. Moreover, {D ∈ B : k · kD is 1-absolutely dominating } (26) is a Banach cover of E. Proof. By 6.4 each element of E is contained in an ordered Banach space with closed gen- erating cone (with a 1-absolutely dominating norm). By Theorem 6.3 and by definition of the norm, B forms a Banach cover of E. 6.6 It is reasonable to ask if an analogue of Theorem 6.5 holds in the world of Riesz spaces: does every Archimedean Riesz space have a Banach cover consisting of Riesz spaces? The answer is negative. Let E be the Riesz space of all functions f on N for which there exist N ∈ N and r, s ∈ R such that f (n) = sn + r for n ≥ N . Suppose E has a Banach cover B consisting of Riesz subspaces of E. There is a D ∈ B that contains the constant function 1 and the identity map i : N → N. For every n ∈ N, 1{1,...,n} = 1 ∨ (n + 1)1 − i ∨ n1 ∈ D. (27) It follows that D = E, so E is a Banach space under some norm. But E is the union of an increasing sequence D1 ⊂ D2 ⊂ · · · of finite dimensional -hence, closed- linear subspaces: Dn = R1 + Ri + R1{1} + · · · + R1{n}. (28) By Baire's Category Theorem, some Dn has nonempty interior in E. Then E = Dn and we have a contradiction. 6.7 In Theorem 6.5 we single out one particular Banach cover B. If we consider only Banach covers consisting of directed spaces with closed cones, this B is the largest and gives us the largest collection of integrable functions. Without directedness there may not be a largest Banach cover. For instance, consider Example 5.10. Impose on E the trivial ordering (x ≤ y if and only if x = y). Then E+ = {0}, and both B1 and B2 consist of Banach spaces with closed (but not generating) cones. 14 7 The integral for an Archimedean ordered vector space As a consequence of Theorem 4.6 we obtain the following extension of Theorem 5.11. 7.1 Theorem Let E be an ordered vector space with an ordered Banach cover B so that D+ is closed for all D ∈ B. Then u is order preserving. Moreover, E and U are Archimedean. 7.2 Lemma Let D be an ordered Banach space with a closed generating cone D+. Let T be a linear order preserving map of D into an Archimedean ordered vector space H. Then ker T is closed and T (D) equipped with the norm k · kq given by kzkq = inf{kxk : x ∈ D, T x = z}, (29) has a closed generating cone T (D)+. Proof. (I) Let x1, x2, · · · ∈ ker T , x ∈ D, xn → x; we prove x ∈ ker T . We assume u−→ T x, so T x = 0. (II) D/ ker T is a Banach space under the quotient norm k · kQ. The formula x + ker T 7→ T x describes a linear bijection D/ ker T → T (D) and Pn∈N kx − xnk < ∞. By Theorem 3.9 xn u−→ x. By Lemma 3.8 T xn kx + ker T kQ = kT xkq (x ∈ D). (30) It follows that k · kq is indeed a norm, turning T (D) into a Banach space. T (D+) ⊂ T (D)+, whence T (D) is directed. We prove that T (D) satisfies (iii) of Theorem 3.9: Let z1, z2, · · · ∈ T (D) be such that Pn∈N kznkq < ∞. Choose xn ∈ D such that T xn = zn, Pn∈N kxnk < ∞. Using u−→ 0. Then zn = T xn (iii) for D, xn u−→ 0 by Lemma 3.8. 7.3 In the proof of Lemma 7.2 we mentioned the inclusion T (D+) ⊂ T (D)+. This inclusion can be strict. Take D = H = R2, T (x, y) = (x, x + y). Then T (D+) 6= T (D)+. From Theorem 3.11, 4.4(c) and Lemma 7.2 we get: 7.4 Theorem Let E1, E2 be ordered vector spaces, Ei endowed with the Banach cover Bi consisting of the ordered Banach spaces with closed generating cones. Let T : E1 → E2 be linear and order preserving. If f : X → E1 is B1-integrable, then T ◦ f : X → E2 is B2-integrable, and u2(T ◦ f ) = T (u1(f )). 7.5 In view of Theorem 3.11 the reader may wonder why in Theorem 7.4 T is required to be order preserving and not just order bounded, the more so because of the following considerations. Let D and H be as in Lemma 7.2 and T be a linear order bounded map of D into H. As the implication (c) =⇒ (a) of Theorem 3.9 is valid for Archimedean (but not necessarily directed) D, following the lines of the proof of Lemma 7.2 ker T is closed and T (D) equipped with the norm as in (29) has a closed cone T (D)+. However, we also need T (D) to be directed and order boundedness of T is no guarantee for that. 15 An alternative approach might be to drop the directedness condition on the spaces that constitute B. However, the ordered Banach spaces with closed cones may not form a Banach cover. For an example, let E be ℓ∞ and let B be the collection of all ordered Banach spaces that are subspaces of ℓ∞ and have closed cones. We make D1, D2 ∈ B. For D1 we take ℓ∞ with the usual norm k · k∞. Choose a linear bijection T : ℓ∞ → ℓ∞ that is not continuous. For a ∈ ℓ∞ put a′ = (a1, −a1, a2, −a2, . . . ). For D2 we take the vector space {a′ : a ∈ ℓ∞} with the norm k · kT given by ka′kT = kT ak∞. Then D2 is a Banach space and D+ 2 , begin {0}, is closed in D2. Suppose B is a Banach cover. Let D be as in Definition 5.2. By the Open Mapping Theorem the identity map D1 → D is a homeomorphism. By the continuity of the inclusion map D2 → D there exists a number c such that ka′kT ≤ cka′k∞ for all a ∈ ℓ∞. Then kT ak∞ ≤ cka′k∞ ≤ ckak∞ for a ∈ ℓ∞, so T is continuous. Contradiction. 7.6 Theorem Let E be an ordered vector space such that E∼ separates the points of E.7 Assume B is a Banach cover consisting of ordered Banach spaces with closed generating cones. Let f ∈ U. Then α ◦ f ∈ L1(µ) for all α ∈ E∼. Moreover, I ∈ E is such that α(I) =Z α ◦ f dµ for all α ∈ E∼, (31) if and only if I = u(f ). Proof. Let f ∈ U and let I = u(f ). By Theorem 7.4 α ◦ f ∈ L1(µ) for all α ∈ E∼ and (31) holds. I = u(f ) is the only element of E for which (31) holds because E∼ separates the points of E. 7.7 Remark Functions with values in a Banach space that are Bochner integrable are also Pettis integrable. To some extent the statement of Theorem 7.6 is similar. Indeed, the definition of Pettis integrability could be generalised for vector spaces V which are equipped with a set S of linear maps V → R that separates the points of V , in the sense that one calls a function f : X → V Pettis integrable if α ◦ f ∈ L1(µ) for all α ∈ S and there exists a I ∈ V such that α(I) = R α ◦ f dµ for all α ∈ S. Then Theorem 7.6 implies that every f ∈ U is Pettis integrable when considering V = E and S = E∼. Observe, however, that even for a Riesz space E, E∼ may be trivial (see, e.g., [11, 5.A]). 8 Comparison with other integrals In this section (X, A, µ) is a complete σ-finite measure space and E is a dir- ected ordered vector space with an ordered Banach cover B so that D+ is closed for each D ∈ B. 7We write E∼ for the space of order bounded linear maps E → R. 16 In 5.6 we have introduced an integral u on a space U of B-integrable functions X → E.8 In [15], starting from a natural integral ϕ on the space S of all simple functions X → E we have made integrals ϕV , ϕL, ϕLV , . . . on spaces SV , SL, SLV , . . . . There is an elementary connection: S is part of U and u coincides with ϕ on S. (Indeed, let f ∈ S. Being a finite set, f (X) is contained in D for some D ∈ B. Then f is Bochner integrable as a map X → D.) In general, SV and SL are not subsets of U, but we can prove that u coincides with ϕV on SV ∩ U and with ϕL on SL ∩ U. Better than that: u is "compatible" with ϕV in the sense that u and ϕV have a common order preserving linear extension SV + U → E. Similarly, u is "compatible with ϕL, ϕLV , . . . ". 8.1 Lemma (a) Let f ∈ U, g ∈ SV , f ≤ g. Then u(f ) ≤ ϕV (g). (b) Let f ∈ U, g ∈ SL, f ≤ g. Then u(f ) ≤ ϕL(g). Proof. (a) By the definition of ϕV and by the text preceding this lemma we have ϕV (g) = inf{ϕ(h) : h ∈ S, h ≥ g} = inf{u(h) : h ∈ S, h ≥ g}. As g ≥ f and u is order preserving (Theorem 7.1), it follows that ϕV (g) ≥ u(f ). (b) Let g ∈ SL and assume f ≤ g. Let g1, g2 ∈ S+ (Bi)i∈N be a ϕ-partition for both g1 and g2. Write An = Sn f 1An ≤ g1An, thus by (a) (and Theorem 7.1) L be such that g = g1 − g2. Let i=1 Bi for n ∈ N. Then u(f 1An) ≤ u(g1An) = ϕ(g1An ) = ϕ(g1 1An) − ϕ(g21An) ≤ ϕL(g1) − ϕ(g2 1Ak ) (k ∈ N, k < n). (32) Which implies u(f 1An) + ϕ(g2 1Ak ) ≤ ϕL(g1) for all k < n. Then letting n tend to ∞ (apply 4.4(d): f (x)1An(x) → f (x) for all x ∈ X) we obtain u(f ) ≤ ϕL(g1) − ϕ(g21Ak ) (k ∈ N), (33) from which we conclude u(f ) ≤ ϕL(g). 8.2 Theorem (a) If g ∈ SLV and f ≤ g, then u(f ) ≤ ϕLV (g). (b) If SV is stable, g ∈ SV LV and f ≤ g, then u(f ) ≤ ϕV LV (g). Proof. Follow the lines of the proof of the lemma with SV , SL or SV L instead of S. 8.3 (Comments on Theorem 8.2) (1) The theorem supersedes the lemma because SV + SL ⊂ SLV . (2) As a consequence, u = ϕLV on U ∩ SLV , and u = ϕV LV on U ∩ SV LV if SV is stable. (3) Recall that stability of SV is necessary for the existence of SV LV . 8In this section we close an eye for the difference between a function and its equivalence class. There will be no danger of confusion. 17 8.4 Theorem Let E be a uniformly complete Riesz space and B be the Banach cover of principal ideals (see Example 5.4). U is a linear subspace of SLV and u = ϕLV on U. Proof. Let u ∈ E+ and let f : X → Eu be Bochner integrable. We prove f ∈ SLV . For simplicity of notation, put D = Eu. Let SD be the space of simple functions X → D. By [15, Corollary 9.8] we have f ∈ (SD)LV and u(f ) = ϕLV (f ). Since D is a Riesz ideal in E, the identity map D → E is order continuous. Then [15, Theorem 8.14] implies f ∈ SLV . Contrary to Theorem 8.4, in [15, Example 9.9(II)] U is not a linear subspace of SLV . Then next example shows, in the context of Theorem 8.4, that the inclusion may be strict. 8.5 Example (U ( SLV ) For X = N, A = P(N) and µ the counting measure and E = c. As is mentioned in [15, Examples 9.9(I)], the function n 7→ 1{n} is an element of SLV but not Bochner integrable. With B the Banach cover of principal ideals, the function n 7→ 1{n} is not B-integrable (see 5.8(4)). 9 An example: Convolution To illustrate the B-integral as an extension of the Bochner integral we consider the following situation. (This introduction requires some knowledge of harmonic analysis on locally compact groups, the balance of this section does not.) Let G be a locally compact group. For f : G → R and x ∈ G we let Lxf : G → R be the function y 7→ f (x−1y). For a finite measure µ on G and f in L1(G) one defines their convolution product to be the element µ ∗ f of L1(G) given for almost every y ∈ G by (µ ∗ f )(y) =Z f (x−1y) dµ(x) =Z (Lxf )(y) dµ(x). (34) The map x 7→ Lxf of G into L1(G) is continuous and bounded, hence Bochner integrable with respect to µ. It is not very difficult to prove that µ ∗ f =Z Lxf dµ(x). (35) Similar statements are true for other spaces of functions instead of L1(G), such as Lp(G), with 1 < p < ∞, and C0(G), the space of continuous functions that vanish at infinity. But consider the space C(G) of all continuous functions on G. The integrals they do if µ has compact support. Thus, one can reasonably define µ ∗ f for f ∈ C(G) and compactly supported µ. However, there is no natural norm on C(G) (except, of R f (x−1y) dµ(x) will not exist for all f ∈ C(G), y ∈ G and all finite measures µ, but course, if G is compact), so we cannot speak of R Lxf dµ(x) as a Bochner integral. We will see that, at least for σ-compact G, it is a B-integral where B is the Banach cover of C(G) that consists of the principal (Riesz) ideals. 18 9.1 Theorem Let G be a σ-compact locally compact group. For every f ∈ C(G) there exists a w ∈ C(G)+ such that Every Lxf x 7→ Lxf is continuous relative to k · kw. (x ∈ G) lies in the principal ideal C(G)w; (36) Proof. Choose compact K1 ⊂ K2 ⊂ . . . such that K1 is a neighbourhood of e; Kn = K −1 For x ∈ G, define [x] to be the smallest n with x ∈ Kn. Then [x] = [x−1], [xy] ≤ n ; KnKn ⊂ Kn+1, Sn∈N Kn = G. 1 + [x] ∨ [y] for all x, y ∈ G by definition. Let f ∈ C(G). Define u, v : G → [1, ∞) as follows: u(x) := 1 + sup f (Kn+1), v(x) = [x]u(x) if [x] = n. (37) (1) If x, y ∈ G, [x] ≤ [y] = n, then x, y ∈ Kn, so f (xy) ≤ sup f (Kn+1) ≤ u(y). (38) (2) Hence, for all x, y ∈ G: f (x−1y) ≤ u(x) ∨ u(y) ≤ u(x)v(y), i.e., Lxf ≤ u(x)v. Let a ∈ G, ε > 0. In (5), using (3) and (4), we show the existence of a neighbourhood U of e with x ∈ aU =⇒ Lxf − Laf ≤ εv. Choose a p with a ∈ Kp. (3) K1 contains an open set V containing e. We make a q ∈ N with x ∈ aV =⇒ Lxf − Laf ≤ εv on G \ Kq. Let x ∈ aV . Then x, a ∈ Kp+1, so [x], [a] ≤ p + 1. By (1): (39) (40) (Lxf − Laf )(y) ≤ f (x−1y) + f (a−1y) ≤ 2u(y) if [y] ≥ p + 1. (41) Moreover, εv(y) = ε[y]u(y) ≥ 2u(y) if [y] ≥ 2 For y ∈ G \ Kq we have [y] > q, so (Lxf − Laf )(y) ≤ εv(y). (4) We show there exists an open set W containing e with ε . Take q ∈ N with q ≥ p + 1 and q ≥ 2 ε . x ∈ aW =⇒ Lxf − Laf ≤ εv on Kq. (42) The function (x, y) 7→ (Lxf − Laf )(y) on G × G is continuous and Kq is compact. Hence by [15, Theorem 8.15] the function G 7→ [0, ∞) x 7→ sup y∈Kq (Lxf − Laf )(y) is continuous. Its value at a is 0, so there exists an open set W containing e with (Lxf − Laf )(y) ≤ ε (x ∈ aW ). sup y∈Kq 19 (43) (44) As v(y) ≥ 1 for all y we obtain (42). (5) With U = V ∩ W we have x ∈ aU =⇒ Lxf − Laf ≤ εv on G. (45) Therefore, to prove the theorem it is sufficient to show there exists a continuous function w ≥ v: (6) Set αn = n(1 + sup f (Kn+1)) for all n; then [x] = n =⇒ v(x) = αn. (46) Put K0 = ∅. For all n ∈ N we have V Kn−1 ⊂ K1Kn−1 ⊂ Kn, so Kn−1 is a subset of K ◦ n, the interior of Kn. By Urysohn [8, Theorem VII.4.1 and Theorem XI.1.2] for all n there is a continuous gn : G → [0, 1] with gn = 0 on Kn−1, gn = 1 on G \ K ◦ n ⊃ G \ Kn. (47) Let b ∈ G. There is a l with b ∈ Kl. bV is an open set containing b. As bV ⊂ Kl+1 and gn = 0 on Kn−1 we have: gn = 0 on bV as soon as n ≥ l + 2. Hence, w := Pn∈N αn+1gn is a continuous function G → [0, ∞). For every x ∈ G there is an n with [x] = n; then x /∈ Kn−1, gn−1(x) = 1 and w(x) ≥ αn = v(x). 9.2 Theorem Let G be a σ-compact locally compact group. Let µ be a finite measure on the Borel σ-algebra of G with a compact support. Let B be the Banach cover of C(G) consisting of the principal ideals as in Example 5.4. Then for every f ∈ C(G) the function x 7→ Lxf is B-integrable and its integral is the "convolution product" µ ∗ f : (µ ∗ f )(y) =Z f (x−1y) dµ(x) (y ∈ G). (48) Proof. By Theorem 9.1 there exists a w ∈ C(G)+ such that (36) holds. This implies that the map x 7→ Lxf is Borel measurable and {Lxf : x ∈ G} is separable in C(G)w. As x 7→ kLxf kw is continuous and thus bounded on the support of µ, the map x 7→ Lxf is B-integrable (see 4.4(a)). That the integral is equal to µ ∗ f follows by 4.4(c). 9.3 Remark Theorem 9.2 compares to [15, Example 8.17] in the sense that in both situations the convolution is equal to an integral of the translation. Though the situation is slightly different in the sense that in Theorem 9.2 we consider σ-compact locally compact groups, while in [15, Example 8.17] we considered metric locally compact groups (the fact that in [15, Example 8.17] Lxf (t) = f (tx−1) is reminiscent for its statement). Acknowledgements W.B. van Zuijlen is supported by the ERC Advanced Grant VARIS-267356 of Frank den Hollander. 20 A Appendix: Absolutely dominating norms In this section, C > 0 and D is an ordered Banach space with closed generating cone and with a C-absolutely dominating norm k · k. As is mentioned in 3.5, we show that C ≥ 1 (A.1) and that for every ε > 0 there exists an equivalent (1 + ε)-absolutely norm (Theorem A.3). Furthermore, we discuss (in A.4 – A.10) whenever there exists an equivalent norm k · k1 for which kxk1 = inf{kak1 : a ∈ D+, −a ≤ x ≤ a} (x ∈ D). (49) This is done by means of the norm N introduced in A.2. Example A.11 illustrates that the existence of such equivalent norm may fail. A.1 (C has to be ≥ 1) Suppose that C < 1. Choose C ′ > 0 such that C < C ′ < 1. For all a ∈ D+ with a 6= 0 there exists a b ∈ D+ with a ≤ b and kbk ≤ C ′kak. Let a ∈ D+ with kak = 1. Iteratively one obtains a sequence a ≤ a1 ≤ a2 ≤ · · · with u−→ 0 by ka1k ≤ C ′ and kan+1k ≤ C ′kank for all n. Then Pn∈N kank < ∞ and thus an Theorem 3.9, which contradicts 0 < a ≤ an. A.2 Define N : D → [0, ∞) by N (x) = inf{kak : a ∈ D+, −a ≤ x ≤ a}. (50) N is a seminorm, and actually a norm because (see Theorem 3.9) N (x) = 0 ⇐⇒ there is an a∗ ∈ D+ with − 1 n a∗ ≤ x ≤ 1 n a∗ (n ∈ N). (51) Because k · k is C-absolutely dominating one has N ≤ Ck · k. A.3 Theorem For all ε > 0 there exists an equivalent norm ρ on D, for which (1 + ε)2ρ(x) ≥ inf{ρ(a) : a ∈ D+, −a ≤ x ≤ a} (x ∈ D). (52) Proof. Define ρ := εN + k · k. ρ is a norm which is equivalent to k · k, since N ≤ Ck · k. Let x ∈ D, x 6= 0. Choose a ∈ D+ with −a ≤ x ≤ a such that kak ≤ (1 + ε)N (x). Note that by definition of N we have N (a) ≤ kak. Whence ρ(a) ρ(x) = N (a) + εkak N (x) + εkxk ≤ N (a) + εkak N (x) ≤ kak + εkak N (x) ≤ (1 + ε) kak N (x) ≤ (1 + ε)2. (53) A.4 Suppose k · k1 is a norm equivalent to k · k for which there exists a C > 0 such that Ckxk1 = inf{kak1 : a ∈ D+, −a ≤ x ≤ a} (x ∈ D). (54) Then it is straightforward to show that N is equivalent to k · k. 21 A.5 N is equivalent to k · k if and only if there exists a c > 0 such that N ≥ ck · k. The latter is true if and only if − 1 n a∗ ≤ xn ≤ 1 n a∗ (n ∈ N) =⇒ lim n→∞ kxnk = 0. A.6 Theorem N (x) = inf{N (a) : a ∈ D+, −a ≤ x ≤ a}. (55) (56) Proof. (≤) Take a ∈ D+, −a ≤ x ≤ a; we prove N (x) ≤ N (a). For all b ≥ a, −b ≤ x ≤ b, whence N (x) ≤ kbk. Thus N (x) ≤ inf{kbk : a ≤ b} = N (a), the latter by definition of N (a). (≥) Take ε > 0. Choose a ∈ D+, −a ≤ x ≤ a, kak ≤ N (x) + ε. Then N (a) ≤ kak ≤ N (x) + ε. A.7 Theorem [1, Theorem 2.38] For an ordered normed vector space E the following are equivalent. (a) The cone E+ is normal. (b) The normed space E admits an equivalent monotone norm. (c) There is a c > 0 such that 0 ≤ x ≤ y implies kxk ≤ ckyk. A.8 Suppose D is an ordered Banach space with closed generating cone and suppose there exists a c > 0 such that kxk ≤ c inf{kak : a ∈ D+, −a ≤ x ≤ a} = cN (x). (57) Then for x ∈ D+, a ∈ D+ with x ≤ a one has kxk ≤ ckak. A.9 Suppose D is an ordered Banach space with closed generating cone and k · k is monotone. Let x ∈ D and a ∈ D+ be such that −a ≤ x ≤ a. Then 0 ≤ x + a ≤ 2a and whence kxk ≤ kx + ak + kak ≤ 3kak.Thus kxk ≤ 3 inf{kak : a ∈ D+, −a ≤ x ≤ a} = 3N (x). (58) (59) We conclude: A.10 Theorem Let D be an ordered Banach space with closed generating cone. The following are equivalent (a) There exist a norm k · k1 that is equivalent to k · k for which kxk1 = inf{kak1 : a ∈ D+, −a ≤ x ≤ a} (x ∈ D). (60) 22 (b) There exists a c > 0 such that N ≥ ck · k. (c) There exists a monotone norm that is equivalent to k · k. (d) E+ is normal. In the following we give an example of an ordered Banach space with closed gener- ating cone D for which none of (a) – (d) of Theorem A.10 holds. A.11 Example Let D be ℓ1 with its natural norm. Define T : ℓ1 → RN by T x = (x1, x1 + x2, x1 + x2 + x3, . . . ). As T is linear, D is an ordered vector space under the relation (cid:22) x (cid:22) y ⇐⇒ T x ≤ T y. (61) (62) The positive cone of ℓ1 is included in D+, whence D is directed. Moreover, D+ is closed and so k · k is absolutely dominating. With xn = (1, −1, 1, −1, . . . , ±1, 0, 0 . . . ) and a = (1, 0, 0, . . . ) we have −a ≤ xn ≤ a and kxnk = n, kak = 1. B Appendix: Banach cover of Köthe spaces In this section (Y, B, ν) is a complete σ-finite measure space and M is the space of classes of measurable functions Y → R as in Example 5.5. B.1 Lemma {Lρw : w ∈ M +, w > 0 a.e.} is a Banach cover of M . Proof. Lρw is complete since f 7→ f w is an isometric bijection Mw → L1(ν). Let f ∈ M . We show there exists a w ∈ M +, w > 0 a.e., with f ∈ Lρw . By the σ-finiteness of ν there is a u ∈ L1(ν), u > 0 a.e. Put w = (f + 1)−1u. Then w ∈ M +, w > 0 a.e., and f ∈ Lρw because R f w dν ≤R u dν < ∞. Lρv and ρw∧v ≤ ρw on Lρw and ρw∧v ≤ ρv on Lρv . If w, v ∈ M +, w > 0, v > 0 a.e., then w ∧ v > 0 a.e., Lρw∧v is a subset of Lρw and B.2 Theorem The complete Köthe spaces form a Banach cover of M . Proof. Let ρ be a function norm and Lρ be complete. If fn ∈ M + for n ∈ N and will use this fact below. Pn∈N ρ(fn) < ∞, then Pn∈N fn ∈ Lρ, so Pn∈N fn < ∞ a.e. and fn → 0 a.e. We By Lemma B.1 it suffices to prove the following. Let ρ1, ρ2 be function norms, Lρ1 and Lρ2 complete. We make a function norm ρ such that Lρ is complete and ρ ≤ ρ1, ρ ≤ ρ2. (Then Lρ1, Lρ2 ⊂ Lρ and we are done.) Define ρ : M → [0, ∞] by ρ(f ) = inf{ρ1(g) + ρ2(h) : g, h ∈ M +, g + h ≥ f }. (63) 23 If ρ(f ) = 0, choose gn, hn with gn + hn ≥ f , ρ1(gn) + ρ2(hn) ≤ 2−n. Then (by the above), gn → 0 a.e., hn → 0 a.e. Hence, f = 0 a.e. It follows easily that ρ is a function norm. Obviously, ρ ≤ ρ1, ρ ≤ ρ2. For the completeness of Lρ: Let u1, u2, · · · ∈ L+ ρ , Pn∈N ρ(un) < ∞. Choose gn, hn ∈ M +, gn+hn ≥ un, ρ1(gn)+ρ2(hn) < ρ(un)+2−n. ThenPn∈N ρ1(gn) < ∞, soPn∈N gn ∈ Lρ1, ρ1(Pn∈N gn) < ∞. Similarly ρ2(Pn∈N hn) < ∞. Then Pn∈N un ∈ Lρ. References [1] Charalambos D. Aliprantis and Rabee Tourky. Cones and Duality, volume 84 of Graduate Studies in Mathematics. American Mathematical Society, 2007. [2] T. Andô. On fundamental properties of a Banach space with a cone. Pacific J. Math., 12:1163–1169, 1962. [3] Charles J. K. Batty and Derek W. Robinson. Positive One-Parameter Semigroups on Ordered Banach Spaces. Acta Applicandae Mathematicae, 2(3):221–296, 1984. [4] G. Birkhoff. Lattice Theory, volume XXV of Colloquium publications. American Mathematical Society, third edition, 1967. [5] Б. З. Вулих. Введение в теорию конусов в нормированных пространствах. Изд- во Калинин. ун-та, pages 1–84, 1977. B.Z. Wulich. Einführung in die Theorie der Kegel in normierten Räumen. (Russisch). Izd. Staatl. Universität Kalinin. 1977 (handschriftl. deutsche Übersetzung: M.R.Weber). [6] Б. З. Вулих. Специальные вопросы геометрии конусов в нормированных про- странствах. Изд-во Калинин. ун-та, pages 1–84, 1978. B.Z. Wulich. Spezielle Probleme der Geometrie von Kegeln in normierten Räumen. (Russisch). Izd. Staatl. Universität Kalinin. 1978 (handschriftl. deutsche Übersetzung: M.R.Weber). [7] John B. Conway. A Course in Functional Analysis. Springer, second edition, 2007. [8] James Dugundji. Topology. Allyn and Bacon series in Advanced Mathematics. Allyn and Bacon, Inc., 1966. [9] E. Hille and R.S. Phillips. Functional Analysis and Semi-Groups, volume 31 of Colloquium publications. American Mathematical Society, 1957. [10] H. Hogbe-Nlend. Bornologies and Functional Analysis: Introductory Course on the Theory of Duality Topology-bornology and Its Use in Functional Analysis, volume 62 of Notas de matemática. North-Holland Publishing Company, 1977. [11] E. de Jonge and A.C.M. van Rooij. Introduction to Riesz spaces. Mathematisch Centrum, 1977. 24 [12] W. A. J. Luxemburg and A. C. Zaanen. Riesz spaces. Vol. I. North-Holland Publishing Co., Amsterdam-London; American Elsevier Publishing Co., New York, 1971. North-Holland Mathematical Library. [13] Miek Messerschmidt. Normality of spaces of operators and quasi-lattices. Positivity, 19(4):695–724, 2015. [14] Anthony L. Peressini. Ordered Topological Vector Spaces. Harper & Row, Publishers, 1967. [15] Arnoud van Rooij and Willem van Zuijlen. Integrals for functions with values in partially ordered vector spaces. To appear in Positivity, preprint available at http://arxiv.org/abs/1505.05997. [16] Raymond A. Ryan. Introduction to tensor products of Banach spaces. Springer Monographs in Mathematics. Springer-Verlag London, Ltd., London, 2002. [17] G. Erik F. Thomas. Integration of Functions With Values in Locally Convex Suslin Spaces. Trans. Amer. Math. Soc., 212:61–81, 1975. [18] A.C. Zaanen. Riesz Spaces II. North-Holland Publishing Company, 1983. 25
1801.07953
1
1801
2018-01-24T12:19:55
The applications of Cauchy-Schwartz inequality for Hilbert modules to elementary operators and i.p.t.i. transformers
[ "math.FA", "math.OA" ]
We apply the inequality $\left|\left<x,y\right>\right|\le\|x\|\,\left<y,y\right>^{1/2}$ to give an easy and elementary proof of many operator inequalities for elementary operators and inner type product integral transformers obtained during last two decades, which also generalizes all of them.
math.FA
math
THE APPLICATIONS OF CAUCHY-SCHWARTZ INEQUALITY FOR HILBERT MODULES TO ELEMENTARY OPERATORS AND I.P.T.I. TRANSFORMERS DRAGOLJUB J. KEČKIĆ Abstract. We apply the inequality hx, yi ≤ x hy, yi1/2 to give an easy and elementary proof of many operator inequalities for elementary operators and inner type product integral transformers obtained during last two decades, which also generalizes all of them. 1. Introduction Let A be a Banach algebra, and let aj, bj ∈ A. Elementary operators, introduced by Lummer and Rosenblum in [13] are mappings from A to A of the form n (1.1) x 7→ ajxbj. Xj=1 Finite sum may be replaced by infinite sum provided some convergence condition. A similar mapping, called inner product type integral transformer (i.p.t.i. trans- formers in further), considered in [6], is defined by (1.2) X 7→ZΩ AtXBtdµ(t), where (Ω, µ) is a measure space, and t 7→ At, Bt are fields of operators in B(H). During last two decades, there were obtained a number of inequalities involving elementary operators on B(H) as well as i.p.t.i. type transformers. The aim of this paper is two give an easy and elementary proof of those proved in [4, 5, 6, 7, 8, 9, 10] and [15] using the Cauchy Schwartz inequality for Hilbert C∗-modules – the inequality stated in the abstract, which also generalizes all of them. 2. Preliminaries Throughout this paper A will always denote a semifinite von Neumann algebra, and τ will denote a semifinite trace on A. Lp(A; τ ) will denote the non-commutative Lp space, Lp(A; τ ) = {a ∈ A kakp = τ (ap)1/p < +∞}. dualities are realized by It is well known that L1(A; τ )∗ ∼= A, Lp(A; τ )∗ ∼= Lq(A; τ ), 1/p + 1/q = 1. Both Lp(A; τ ) ∋ a 7→ τ (ab) ∈ C, b ∈ Lq(A; τ ) or b ∈ A. 2010 Mathematics Subject Classification. Primary: 47A63, Secondary: 47B47, 47B10, 47B49, 46L08. Key words and phrases. Cauchy Schwartz inequality, unitarily invariant norm, elementary operator, inner product type transformers. The author was supported in part by the Ministry of education and science, Republic of Serbia, Grant #174034. 1 2 DRAGOLJUB J. KEČKIĆ For more details on von Neumann algebras the reader is referred to [18], and for details on Lp(A, τ ) to [2]. Let M be a right Hilbert W ∗-module over A. (Since M is right we assume that A-valued inner product is A-linear in second variable, and adjoint A-linear in the first.) We assume, also, that there is a faithful left action of A on M , that is, an embedding (and hence an isometry) of A into Ba(M ) the algebra of all adjointable bounded A-linear operators on M . Hence, for x, y ∈ M and a, b ∈ A we have hx, yi a = hx, yai , hxa, yi = a∗ hx, yi , hx, ayi = ha∗x, yi . For more details on Hilbert modules, the reader is referred to [12] or [14]. We quote the basic property of A-valued inner product, a variant od Cauchy- Schwartz inequality. Proposition 2.1. Let M be a Hilbert C∗-module over A. For any x, y ∈ M we have (2.1) hx, yi2 ≤ kxk2 hy, yi , hx, yi ≤ kxk hy, yi1/2 , in the ordering of A. The proof can be found in [12, page 3] or [14, page 3]. Notice: 1◦ the left inequality implies the right one, since t 7→ t1/2 is operator increasing function; 2◦ Both inequalities holds for A-valued semi-inner product, i.e. even if h·,·i may be degenerate. Finally, we need a counterpart of Tomita modular conjugation. Definition 2.1. Let M be a Hilbert W ∗-module over a semifinite von Neumann algebra A, and let there is a left action of A on M . A (possibly unbounded) mapping J, defined on some M0 ⊆ M with values (i) J(axb) = b∗J(x)a∗; (ii) in M , we call modular conjugation if it satisfies: τ (hJ(y), J(x)i) = τ (hx, yi) whenever hx, xi, hy, yi, hJ(x), J(x)i, hJ(y), J(y)i ∈ L1(A, τ ). In what follows, we shall use simpler notation x instead of J(x). Thus, the determining equalities become (2.2) axb = b∗xa∗, τ (hy, xi) = τ (hx, yi). The module M together with left action of A and the modular conjugation J we shall call conjugated W ∗-module. Definition 2.2. Let M be a conjugated W ∗-module over A. We say that x ∈ M0 is normal, if (i) hx, xi x = xhx, xi, (ii) hx, xi = hx, xi. Remark 2.1. It might be a nontrivial question, whether J can be defined on an arbitrary Hilbert W ∗-module in a way similar to the construction of Tomita's mod- ular conjugation (see [17]). However for our purpose, the preceding definition is enough. Examples of conjugated modules are following. Example 2.1. Let A be a semifinite von Neumann algebra, and let M = An. For x = (x1, . . . , xn), y = (y1, . . . , yn) ∈ M , a ∈ A, define right multiplication, left action of A, the A-valued inner product and modular conjugation by (2.3) xa = (x1a, . . . , xna), ax = (ax1, . . . , axn); THE APPLICATIONS OF CAUCHY-SCHWARTZ INEQUALITY 3 (2.4) hx, yi = x∗ 1y1 + ··· + x∗ nyn, x = (x∗ 1, . . . , x∗ n). All required properties are easily verified. The element x = (x1, . . . , xn) is normal whenever all xj are normal and mutually commute. We have hx, ayi = which is the term of the form (1.1). x∗ j ayj, n Xj=1 There are two important modules with infinite number of summands. Example 2.2. Let A be a semifinite von Neumann algebra. We consider the standard Hilbert module l2(A) over A and its dual module l2(A)′ defined by +∞ a∗ l2(A) =n(x1, . . . , xn, . . . ) (cid:12)(cid:12)(cid:12) Xk=1 l2(A)′ =n(x1, . . . , xn, . . . ) (cid:12)(cid:12)(cid:12) (cid:13)(cid:13)(cid:13) kak converges in norm of Ao. kak(cid:13)(cid:13)(cid:13) ≤ M < +∞o. Xk=1 a∗ n kxk weakly converges.) (It is clear that x ∈ l2(A)′ if and only if the series P x∗ The basic operation on these modules are given by (2.3) and (2.4) with infinite number of entries. x = (x∗ 1, . . . , x∗ The main difference between l2(A) (l2(A)′ respectively) and An is the fact that n, . . . ) is defined only on the subset of l2(A) consisting of those x ∈ M for which P xkx∗ The element x = (x1, . . . , xn, . . . ) ∈ M0 is normal whenever all xj are normal k converges in the norm of A. and mutually commute. Remark 2.2. The notation l2(A)′ comes from the fact that l2(A)′ is isomorphic to the module of all adjointable bounded A-linear functionals Λ : M → A. For more details on l2(A) or l2(A)′ see [14, §1.4 and §2.5]. Example 2.3. Let A be a semifinite von Neumann algebra and let (Ω, µ) be a measure space. Consider the space L2(Ω, A) consisting of all weakly-∗ measurable functions such that RΩ kxk2dµ < +∞. The weak-∗ measurability is reduced to the measurability of functions ϕ(x(t)) for all normal states ϕ, since the latter generate the predual of A. Basic operations are given by a · x(t) = ax(t), x(t) · a = x(t)a, All required properties are easily verified. Mapping x 7→ x is again defined on a proper subset of L2(Ω, A). The element x is normal if x(t) is normal for almost all t, and x(t)x(s) = x(s)x(t) for almost all (s, t). x(t)∗y(t)dµ(t), x(t) = x(t)∗. hx, yi =ZΩ Again, for a ∈ A we have hx, ayi =ZΩ which is the term of the form (1.2). x(t)∗ay(t)dµ(t), 4 DRAGOLJUB J. KEČKIĆ Thus, norm estimates of elementary operators (1.1), or i.t.p.i. transformers (1.2) are estimates of the term hx, ayi. In section 4 we need two more examples. Example 2.4. Let M1 and M2 be conjugated W ∗-modules over a semifinite von Neumann algebra A. Consider the interior product of Hilbert modules M1 and M2 constructed as follows. The linear span of x1 ⊗ x2, x1 ∈ M1, x2 ∈ M2 subject to the relations a(x1 ⊗ x2) = ax1 ⊗ x2, x1a ⊗ x2 = x1 ⊗ ax2, (x1 ⊗ x2)a = x1 ⊗ x2a, and usual bi-linearity of x1⊗x2, can be equipped by an A-valued semi-inner product (2.5) hx1 ⊗ x2, y1 ⊗ y2i = hx2,hx1, y1i y2i . The completion of the quotient of this linear span by the kernel of (2.5) is denoted by M1 ⊗ M2 and called interior tensor product of M1 and M2. For more details on tensor products, see [12, Chapter 4]. If M1 = M2 = M , M ⊗ M can be endowed with a modular conjugation by x1 ⊗ x2 = x2 ⊗ x1. All properties are easily verified. Also, x normal implies x ⊗ x is normal and hx ⊗ x, x ⊗ xi = hx, xi2. Example 2.5. Let Mn, n ∈ N be conjugated modules. Their infinite direct sum L+∞ n=1 Mn is the module consisting of those sequences (xn), xn ∈ Mn such that P+∞ n=1 hxn, xni weakly converges, with the A-valued inner product +∞ h(xn), (yn)i = hxn, yni . Xn=1 The modular conjugation can be given by (xn) = (xn). Specially, we need the full Fock module +∞ F = Mn=0 where M ⊗0 = A, M ⊗1 = M , M ⊗2 = M ⊗ M , M ⊗3 = M ⊗ M ⊗ M , etc. For x ∈ M , kxk < 1 the elementP+∞ n=0 x⊗n ∈ F (where x⊗0 := 1) is well defined. It is normal whenever x is normal. Also, for normal x, we have M ⊗n, (2.6) x⊗n, *+∞ Xn=0 x⊗n+ = +∞ Xn=0 +∞ Xn=0(cid:10)x⊗n, x⊗n(cid:11) = +∞ Xn=0 hx, xin = (1 − hx, xi)−1. We shall deal with unitarily invariant norms on the algebra B(H) of all bounded Hilbert space operators. For more details, the reader is referred to [3, Chapter III]. We use the following facts. For any unitarily invariant norm ·, we have A = A∗ = A = U AV = A for all unitaries U and V , as well as kAk ≤ A ≤ kAk1. The latter allows the following interpolation Lemma. Lemma 2.2. Let T and S be linear mappings defined on the space C∞ of all compact operators on Hilbert space H. If kT xk ≤ kSxk for all x ∈ C∞, kT xk1 ≤ kSxk1 for all x ∈ C1 then T x ≤ Sx THE APPLICATIONS OF CAUCHY-SCHWARTZ INEQUALITY 5 for all unitarily invariant norms. Proof. The norms k · k and k · k1 are dual to each other, in the sense kxk = sup kyk1=1 tr(xy), kxk1 = sup kyk=1 tr(xy). Hence, kT ∗xk ≤ kS∗xk, kT ∗xk1 ≤ kS∗xk1. Consider the Ky Fan norm k·k(k). Its dual norm is k·k♯ (k) = max{k·k, (1/k)k·k1}. Thus, by duality, kT xk(k) ≤ kSxk(k) and the result follows by Ky Fan dominance property, [3, §3.4]. (cid:3) 3. Cauchy-Schwartz inequalities Cauchy-Schwartz inequality for k · k follows from (2.1), for k · k1 by duality and for other norms by interpolation. Theorem 3.1. Let A be a semifinite von Neumann algebra, let M be a conjugated W ∗-module over A and let a ∈ A. Then: (3.1) k hx, ayik ≤ kxkkykkak, k hx, ayik1 ≤ k hx, xi1/2 ahy, yi1/2 k1; and k hx, ayik2 ≤ kykk hx, xi1/2 ak2. (3.2) k hx, ayik2 ≤ kxkkahy, yi1/2 k2, Specially, if A = B(H), τ = tr and x, y normal, then (3.3) hx, ayi ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)hx, xi1/2 ahy, yi1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) for all unitarily invariant norms ·. Proof. By (2.1), we have k hx, ayik ≤ kxkkayk ≤ kxkkykkak, which proves the first inequality in (3.1). For the proof of the second, note that by (2.2), for all a ∈ L1(A; τ ) we have τ (b hx, ayi) = τ (hxb∗, ayi) = τ (hya∗, bxi) = τ (ahy, bxi). Hence for hx, xi, hy, yi ≤ 1 kbk=1τ (b hx, ayi) = sup k hx, ayik1 = sup In the general case, let ε > 0 be arbitrary, and let x1 = (hx, xi + ε)−1/2x and y1 = (hy, yi + ε)−1/2y. Then x1 = x(hx, xi + ε)−1/2 and y1 = y(hy, yi + ε)−1/2 (by (2.2)). Thus kbk=1τ (ahy, bxi) ≤ kak1k hy, bxik ≤ kak1, hx1, x1i = (hx, xi + ε)−1/2 hx, xi (hx, xi + ε)−1/2 ≤ 1, by continuous functional calculus. Hence (3.4) k hx, ayik1 =(cid:13)(cid:13)(cid:13)D(hx, xi + ε)1/2x1, a(hy, yi + ε)1/2y1E(cid:13)(cid:13)(cid:13)1 =(cid:13)(cid:13)(cid:13)Dx1, (hx, xi + ε)1/2a(hy, yi + ε)1/2y1E(cid:13)(cid:13)(cid:13)1 ≤ (hx, xi + ε)1/2a(hy, yi + ε)1/2(cid:13)(cid:13)(cid:13)1 =(cid:13)(cid:13)(cid:13) = , and let ε → 0. (Note k(hx, xi + ε)1/2 − hx, xi1/2 k ≤ ε1/2.) To prove (3.2), by (2.1) we have hx, ayi2 ≤ kxk2 hay, ayi = kxk2 hy, a∗ayi . (3.5) Apply k · k1 to the previous inequality. By (3.1) we obtain 2 a∗ahy, yi k hx, ayi k2 2 ≤ kxk2k hy, a∗ayik1 ≤ kxk2k hy, yi 1 1 2 k1 = kxk2kahy, yi 1 2 k2 2. 6 DRAGOLJUB J. KEČKIĆ This proves the first inequality in (3.2). The second follows from duality k hx, ayik2 = k hy, a∗xi k2 ≤ kykka∗ hx, xi1/2 k2 = kykk hx, xi1/2 ak2. Finally, if A = B(H), τ = tr and x, y normal. Then (3.3) holds for k·k1 by (3.1). For the operator norm, it follows by normality. Namely then xhx, xi = hx, xi x and we can repeat argument from (3.4). Now, the general result follows from Lemma 2.2 (cid:3) Corollary 3.2. If A = B(H) and M = l2(A)′ (Example 2.2), then (3.3) is [4, Theorem 2.2] (the first formula from the abstract). If M = L2(Ω, A), A = B(H), (3.3) is [6, Theorem 3.2] (the second formula from the abstract). Remark 3.1. The inequality (3.5) for M = B(H)n is proved in [9] using complicated identities and it plays an important role in this paper. Using three line theorem (which is a standard procedure), we can interpolate results of Theorem 3.1 to Lp(A, τ ) spaces. Theorem 3.3. Let A be a semifinite von Neumann algebra, and let M be a con- jugated W ∗-module over A. For all p, q, r > 1 such that 1/q + 1/r = 2/p, we have (3.6) k hx, ayikp ≤(cid:13)(cid:13)(cid:13)(cid:13)Dhx, xiq−1 x, xE1/2q aDhy, yir−1 y, yE1/2r(cid:13)(cid:13)(cid:13)(cid:13)p Proof. Let u, v ∈ M0 and let b ∈ A. For 0 ≤ Re λ, Re µ ≤ 1 consider the function 2 E . 2 u(hu, ui + ε) This is an analytic function (obviously). f (λ, µ) =D(hu, ui + ε)− λ 2 , b(hv, vi + ε)− µ 2 v(hv, vi + ε) λ−1 µ−1 . On the boundaries of the strips, we estimate. For Re λ = Re µ = 0 2 u(hu, ui + ε)− 1 f (it, is) =D(hu, ui + ε)− it 2 E . Since (hu, ui+ε)−it/2, (hu, ui+ε)it/2, (hv, vi+ε)−is/2 and (hv, vi+ε)is/2 are unitary operators, and since the norm of u(hu, ui + ε)−1/2, v(hv, vi + ε)−1/2 does not exceed 1, by (3.1) we have 2 , b(hv, vi + ε)− is 2 v(hv, vi + ε)− 1 2 + is 2 + it (3.7) For Re λ = Re µ = 1 kf (it, is)k ≤ kbk. f (1+it, 1+is) =D(hu, ui + ε)− 1 By a similar argument, by (3.1) we obtain 2 − it 2 u(hu, ui + ε) it 2 , b(hv, vi + ε)− 1 2 − is 2 v(hv, vi + ε) is 2E . (3.8) kf (1 + it, 1 + is)k1 ≤ kbk1. For Re λ = 0, Re µ = 1, by (3.2) we have 2 u(hu, ui + ε)− 1 f (it, 1+is) =D(hu, ui + ε)− it and hence 2 + it 2 , b(hv, vi + ε)− 1 2 − is 2 v(hv, vi + ε) is 2 E , (3.9) Similarly (3.10) kf (it, 1 + is)k2 ≤ kbk2. kf (1 + it, is)k2 ≤ kbk2. THE APPLICATIONS OF CAUCHY-SCHWARTZ INEQUALITY 7 Let us interpolate between (3.7) and (3.9). Let r > 1. Then 1/r = θ·1+(1−θ)·0 for θ = 1/r. Then (θ · (1/2) + (1 − θ) · 0)−1 = 2r and hence, by three line theorem (see [11] and [2]) we obtain (3.11) kf (0 + it, 1/r)k2r ≤ kbk2r Similarly, interpolating between (3.10) and (3.8) we get (3.12) kf (1 + it, 1/r)k 2r r+1 ≤ kbk 2r r+1 since 1/r = θ·1+(1−θ)·0 for same θ = 1/r and (θ·1+(1−θ)·(1/2))−1 = 2r/(r +1) Finally, interpolate between (3.11) and (3.12). Then 1/q = θ · 1 + (1 − θ) · 0 for θ = 1/q and therefore (θ · r+1 2qr · (r + 1 + q − 1) = p. Thus 2r + (1 − θ) 1 2r )−1 = 1 i.e. kf (1/q, 1/r)kp ≤ kbkp. (cid:13)(cid:13)(cid:13)D(hu, ui + ε)− 1 After substitutions 2q u(hu, ui + ε) 1−q 2q , b(hv, vi + ε)− 1 2r v(hv, vi + ε) 1−q 2q E(cid:13)(cid:13)(cid:13)p ≤ kbkp. u = xhx, xi(q−1)/2 , v = hy, yi(r−1)/2 , b = (hu, ui + ε)1/2qa(hv, vi + ε)1/2r, we obtain (cid:13)(cid:13)(cid:13)Dxhx, xi(q−1)/2 (hx, xiq + ε)(1−q)/2q, ay hy, yi(q−1)/2 (hy, yiq + ε)(1−q)/2qE(cid:13)(cid:13)(cid:13)p ≤ a(cid:16)Dhy, yir−1 y, yE + ε(cid:17)1/2r(cid:13)(cid:13)(cid:13)(cid:13)p ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)Dhx, xiq−1 x, xE + ε(cid:17)1/2q which after ε → 0 yields (3.6), using the argument similar to that in [14, Lemma 1.3.9]. (cid:3) Remark 3.2. In a special case A = B(H), τ = tr, M = l2(A)′, r = q = p formula (3.6) becomes [5, Theorem 2.1] (the main result). Also, for A = B(H), τ = tr, M = L2(Ω, A) formula (3.6) becomes [6, Theorem 3.3] (the first displayed formula from the abstract), there proved with an additional assumption that Ω is σ-finite. In the next two section we derive some inequalities that regularly arise from Cauchy-Schwartz inequality. 4. Inequalities of the type 1 − hx, yi ≥ (1 − kxk2)1/2(1 − kyk2)1/2 The basic inequality can be proved as +∞ +∞ Xn=0 Xn=0 (1 − hx, yi)−1 ≤ hx, yin ≤ kxk2n!1/2 +∞ +∞ Xn=0 Xn=0 Following this method we prove: kxknkykn ≤ kyk2n!1/2 = (1 − kxk2)−1/2(1 − kyk2)−1/2. 8 DRAGOLJUB J. KEČKIĆ Theorem 4.1. Let M be a conjugated W ∗-module over A = B(H), let x, y ∈ M0 be normal, and let hx, xi, hy, yi ≤ 1. Then (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (1 − hx, xi)1/2a(1 − hy, yi)1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ a − hx, ayi in any unitarily invariant norm. Proof. We use examples 2.4 and 2.5. Denote T a = hx, ayi. We have T 2a = hx,hx, ayi yi = hx ⊗ x, ay ⊗ yi and by induction T ka =(cid:10)x⊗k, ay⊗k(cid:11). Suppose kxk, kyk ≤ δ < 1. Then kx⊗kk, ky⊗kk ≤ δk and hence (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)T k(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ δ2k. Then +∞ ≤ y⊗k+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Put b = (I − T )−1a. Then (4.1) (4.2) (4.3) +∞ +∞ T k. (I − T )−1 = Xn=0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=0(cid:10)x⊗k, ay⊗k(cid:11)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) x⊗k+1/2 a*+∞ Xk=0 y⊗k, x⊗k, a (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) *+∞ Xk=0 Xk=0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) y⊗k+1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=0 (1 − hx, xi)−1/2a(1 − hy, yi)−1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) +∞ . . +∞ +∞ x⊗k, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) T ka(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) b =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) *+∞ Xk=0 Xk=0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(I − T )−1a(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) by (3.3) and normality of x and y. Invoking (2.6), inequality (4.2) becomes Finally, note that the mappings I − T and a 7→ (1 − hx, xi)−1/2a(1 − hy, yi)−1/2 commute (by normality of x and y) and put (1−hx, xi)−1/2(a− T a)(1−hy, yi)−1/2 in place of a, to obtain the conclusion. If kxk, kyk = 1 then put δx instead of x and let δ → 1−. (cid:3) Remark 4.1. If M = L2(Ω, A) this is [6, Theorem 4.1] (the last formula from the abstract). If M = B(H) × B(H), x = (I, A), y = (I, B) then it is [4, Theorem 2.3] (the last formula from the abstract). Remark 4.2. Instead of t 7→ 1 − t we may consider any other function f such that 1/f is well defined on some [0, c) and has Taylor expansion with positive coefficients, say cn. Then distribute √cn on both arguments in inner product in (4.2) and after few steps we get For instance, for t 7→ (1 − t)α, α > 0 we have (1 − t)−α = P cntn, where cn = Γ(n + α)/(Γ(α)n!) > 0 and we get (4.4) (f (x∗x))1/2a(f (y∗y))1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ f (T ) . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (1 − hx, xi)α/2a(1 − hy, yi)α/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ (I − T )αa (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n(cid:19)x∗nayn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (1 − x∗x)α/2a(1 − y∗y)α/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xn=0 (−1)n(cid:18)a +∞ , which is the main result of [15]. Varying f , we may obtain many similar inequalities. in any unitarily invariant norm. For M = A = B(H), (4.4) reduces to THE APPLICATIONS OF CAUCHY-SCHWARTZ INEQUALITY 9 Finally, if normality condition on x and y is dropped, we can use (3.6) to obtain some inequalities in Lp(A; τ ) spaces. (4.5) Theorem 4.2. Let M be a conjugated W ∗-module over a semifinite von Neumann algebra A, let x, y ∈ M0, kxk, kyk < 1 and let z⊗n+−1/2 Xn=0 kp ≤ k∆−1/q ∆z =*+∞ Xn=0 k∆1−1/q (a − hx, ayi)∆−1/r for z ∈ {x, y, x, y}. a∆1−1/r z⊗n, Then kp, +∞ x x y y , for all p, q, r > 1 such that 1/q + 1/r = 2/p. Proof. Let b = (I − T )a. We have a = (I − T )−1b and hence Xn=0(cid:10)x⊗n, by⊗n(cid:11) ∆1−1/r k∆1−1/q a∆1−1/r ∆1−1/q +∞ x x y +∞ Xn=0Dx⊗n∆1−1/q x , by⊗n∆1−1/r y y kp =(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) Xn=0Dx⊗n∆1−1/q x by (3.6), where u = +∞ Xn=0*+∞ = (cid:13)(cid:13)(cid:13)p E(cid:13)(cid:13)(cid:13)p ≤ kubvkp, x⊗n+1/2q x . , x⊗n∆1−1/q x ∆1−1/q x x⊗n, ∆1−1/q Eq−1 After a straightforward calculation, we obtain u = ∆−1/q and the conclusion follows. x and similarly v = ∆−1/r (cid:3) y Remark 4.3. When A = B(H), τ = tr, this is the main result of [10], from which we adapted the proof for our purpose. However, unaware of Fock module technique, the authors of [10] produced significantly more robust formulae. Also, in [10], the assumptions are relaxed to r(Tx,x), r(Ty,y) ≤ 1, where r stands for the spectral radius and Tx,y(a) = hx, ayi. This easily implies r(Tx,y) ≤ 1. First, it is easy to see that kTz,zk = kzk2. Indeed, by (3.1) we have kTz,zk ≤ kzk2. On the other hand, choosing a = 1 we obtain kTz,zk ≥ kTz,z(1)k = k hz, zik = kzk2. Again, by (3.1), we have kTx,yk ≤ kxkkyk = pkTx,xkkTy,yk. Apply this to x⊗n x,yk ≤ qkT n and y⊗n instead of x and y and we get kT n y,yk from which we easily conclude r(Tx,y)2 ≤ r(Tx,x)r(Ty,y) by virtue of spectral radius formula. (In a similar way, we can conclude r(Tx,x) = kxk for normal x.) If some of r(Tx,x), r(Ty,y) = 1 then define ∆x = limδ→0 ∆δx = inf 0<δ<1 ∆δx, etc, and the result follows, provided that series that defines ∆x and ∆y are weakly convergent. Thus, if both r(Tx,x), r(Ty,y) < 1, the series in (4.5) converge. x,xkkT n 5. Grüss type inequalities For classical Grüss inequality, see [16, §2.13]. We give a generalization to Hilbert modules following very simple approach from [1] in the case of Hilbert spaces. Theorem 5.1. Let M be a conjugated W ∗-module over B(H), and let e ∈ M be such that he, ei = 1. The mapping Φ : M × M → B(H), Φ(x, y) = hx, yi − hx, eihe, yi. is a semi-inner product. 10 DRAGOLJUB J. KEČKIĆ If, moreover x, y ∈ M0 are normal with respect to Φ. Then (5.1) hx, ayi − hx, eihe, ayi ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) in any unitarily invariant norm. (hx, xi − hx, ei2)1/2a(hy, yi − hy, ei2)1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Finally, if x, y belongs to balls with diameters [me, M e] and [pe, P e] (m, M , p, P ∈ R), respectively, then (5.2) hx, ayi − hx, eihe, ayi ≤ 1 4 aM − mP − p. (Here, x belongs to the ball with diameter [y, z] iff kx − y+z 2 k ≤ k z−y 2 k.) Proof. The mapping Φ is obviously linear in y and conjugate linear in x. Moreover, by inequality (2.1) hx, eihe, xi = he, xi2 ≤ kek2 hx, xi = hx, xi , i.e. Φ(x, x) ≥ 0. Hence Φ is an A-valued (semi)inner product. If x, y are normal, then, by (3.3) we obtain (5.3) Φ(x, ay) ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Φ(x, x)1/2aΦ(y, y)1/2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) in any unitarily invariant norm. Write down exact form of Φ and we obtain (5.1). Finally, for the last conclusion, note that Φ(x, x) = Φ(x − ec, x − ec) for any c ∈ C (direct verification), and hence Φ(x, x) ≤ hx − ec, x − eci, which implies kΦ(x, x)1/2k ≤ kx − eck. Choosing c = (M + m)/2, we obtain kΦ(x, x)1/2k ≤ (M − m)/2. Similarly, kΦ(y, y)1/2k ≤ (P − p)/2. Thus (5.3) implies (5.2). Remark 5.1. Choose M = L2(Ω, µ), µ(Ω) = 1 and choose e to be the function identically equal to 1. Then (cid:3) Φ(x, ay) =ZΩ x(t)∗ay(t)dµ(t) −ZΩ x(t)∗dµ(t)ZΩ ay(t)dµ(t), and from (5.1) and (5.2) we obtain main results of [8]. Remark 5.2. Applying other inequalities from section 3, we can derive other results from [8]. Also, applying inequality hx, ayi2 ≤ kxk2 hay, ayi to the mapping Φ instead of h·,·i we obtain the key result of [7], there proved by complicated identities. 6. Concluding remarks Both, elementary operators and i.p.t.i. transformers on B(H) are special case of (6.1) hx, T yi where x, y are vectors from some Hilbert W ∗-module M over B(H) and T : M → M is given by left action of B(H). Although there are many results independent of the representation (6.1), a lot of inequalities related to elementary operators and i.p.t.i. transformers can be reduced to elementary properties of the B(H)-valued inner product. THE APPLICATIONS OF CAUCHY-SCHWARTZ INEQUALITY 11 References [1] S. S. Dragomir. Generalization of Grüss's inequality in inner product spaces and applications. J Math. Anal. Appl., 237(1):74–82, 1999. [2] T. Fack and H. Kosaki. Generalized s-numbers of τ-measurable operators. Pacific J Math., 123(2):269–300, 1986. [3] I. C. Gohberg and M. G. Kreın. Itroduction to the theory of linear nonselfadjoint operators, volume 018 of Translations of Mathematical Monographs. American Mathematical Society, Providence RI, 1st edition, 1969. [4] D. R. Jocić. Cauchy-Schwarz and means inequalities for elementary operators into norm ideals. Proc. Amer. Math. Soc., 126(9):2705–2711, 1998. [5] D. R. Jocić. The Cauchy-Schwarz norm inequality for elementary operators in Schatten ideals. J. London Math. Soc., 60(3):925–934, 1999. [6] D. R. Jocić. Cauchy-Schwarz norm inequalities for weak∗-integrals of operator valued func- tions. J. Funct. Anal., 218(2):318–346, 2005. [7] D. R. Jocić, Ð. Krtinić, M. Lazarević, P. Melentijević, and S. Milošević. Refinements of inequalities related to Landau-Grüss inequalities for elementary operators acting on ideals associated to p-modified unitarily invariant norms. Complex Anal. Oper. Theory, 12(1):195– 205, 2018. [8] D. R. Jocić, Ð. Krtinić, and M. S. Moslehian. Landau and Grüss type inequalities for inner product type integral transformers in norm ideals. Math. Inequal. Appl., 16(1):109–125, 2013. [9] D. R. Jocić and S. Milošević. Refinements of operator Cauchy-Schwarz and Minkowski inequalities for p-modified norms and related norm inequalities. Linear Algebra Appl., 488(1):284–301, 2016. [10] D. R. Jocić, S. Milošević, and V. Ðurić. Norm inequalities for elementary operators and other inner product type integral transformers with the spectra contained in the unit disc. Filomat, 31(2):197–206, 2017. [11] H. Kosaki. Applications of the complex interpolation method to a von Neumann algebra (non-commutative Lp-spaces). J. Funct. Anal., 56(1):29–78, 1984. [12] E. C. Lance. Hilbert C ∗-Modules: A toolkit for operator algebraists. Cambridge University Press, 1995. [13] G. Lumer and M. Rosenblum. Linear operator equations. Proc. Amer. Math. Soc., 10(1):32– 41, 1959. [14] V. M. Manuilov and E. V. Troitsky. Hilbert C ∗-modules, volume 226 of Translations of mathematical monographs. AMS, Providence, Rhode Island, 2005. [15] S. Milošević. Norm inequalities for elementary operators related to contractions and operators with spectra contained in the unit disk in norm ideals. Adv. Oper. Theory, 1(2):147–159, 2016. [16] D. S. Mitrinović. Analytic inequalities, volume 165 of Die Grundlehren der Matematischen Wissenschaften in Einzeldarstellungen Band. Springer, 1970. [17] M. Takesaki. Tomita's theory of modular Hilbert algebras and its applications, volume 128 of Lecture Notes Math. Springer, 1970. [18] M. Takesaki. Theory of Operator Algebras I, volume 124 of Encyclopaedea of Mathematical Sciences. Springer, Berlin, Heidelberg, etc., 2001. University of Belgrade, Faculty of Mathematics, Studentski trg 16-18, 11000 Beograd, Serbia E-mail address: [email protected]
1603.02471
1
1603
2016-03-08T10:42:19
The best constant in the Khintchine inequality of the Orlicz space $L_{\psi_2}$ for equidistributed random variables on spheres
[ "math.FA" ]
We compute the best constant in the Khintchine inequality for equidistributed random variables on the $N$-sphere in the Orlicz space $L_{\psi_2}$.
math.FA
math
THE BEST CONSTANT IN THE KHINTCHINE INEQUALITY OF THE ORLICZ SPACE Lψ2 FOR EQUIDISTRIBUTED RANDOM VARIABLES ON SPHERES HAUKE DIRKSEN Abstract. We compute the best constant in the Khintchine inequality for equidistributed random variables on the N -sphere in the Orlicz space Lψ2 . 1. Introduction The classical Khintchine inequality compares the Lp-norm of a sum of Rade- macher variables with the ℓ2-norm of the coefficients of the sum. The computation of the best possible constants has attracted a lot of interest. For the classical case, Haagerup found the best constants for general p ∈ (1,∞) in [1]. Also Khintchine inequalities for different kinds of random variables were investigated, for example rotationally invariant random vectors in [3]. A second variation of the problem changes the underlying space. The Khintchine inequality in Orlicz spaces has been considered in various cases, the first example is a paper by Rodin and Semyonov [7]. we denote the norm of the Orlicz space Lψq (Ω, Σ, µ). This is given by Let q > 0 and ψq(x) := exp(xq) − 1 for x ∈ R. By k·kψq c (cid:19)(cid:21) ≤ 1}, := inf{c > 0 E(cid:20)ψq(cid:18)kXk kXkψq for X ∈ Lψq . By k·k we denote the Euclidean norm. For q ≤ 2 one can still compare the Lψq -norm and the ℓ2-norm, see [4]. For q > 2, Pisier proved that the Lorentz sequence spaces ℓq′,∞ (1/q + 1/q′ = 1), instead of ℓ2 come into play, see [6]. This fact was already mentioned by Rodin and Semyonov [7]. Here we compute the best constant for the Orlicz space Lψ2 and equidistributed variables on N -dimensional spheres. We apply the technique from [5]. Peskir reduces the case of the Orlicz space to the classical Khintchine inequality in Lq. The optimality of the constants from Lq carries over to Lψ2. The same reduction technique can be used for variables on spheres. Konig and Kwapien computed the optimal constants in [3]. Again the optimality carries over. In this paper we prove the following result. Date: March 8, 2016. 2010 Mathematics Subject Classification. 41A44, 46B15, 60G50. Key words and phrases. Khintchine inequality, best constant, Orlicz space, random variables on the sphere. 1 2 H. DIRKSEN Theorem 1.1. Let Xj , j = 1, . . . , n be an i.i.d. sequence of equidistributed random variables on the N -sphere SN−1. For all a = a1, . . . , an ∈ R we have ajXj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)ψ2 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 Nr 1 where the constant b(N ) :=q 2 1−( 1 2 ) ≤ b(N )(cid:18) nXj=1 a2 2 j(cid:19) 1 , is optimal. 2 N 1√ln 2 Note that b(N ) decreases to for N → ∞. In Section 2 we prove that the inequality is true. Therefore we consider the series expansion of the exponential function. Then we apply the Khintchine inequality from [3]. In Section 3 we show 1√n Xj that the constant b(N ) can not be smaller. We show that with Yn :=Pn we get asymptotic equality in Theorem 1.1 for n → ∞. 2. Proof of the inequality j=1 Let C > 0. Applying Beppo-Levi we may interchange the limit and the expected value. 2 1 1 = 1 k! 1 k! (2.1) Eexp =E ∞Xk=0 ∞Xk=0  j=1 kajk2 j=1 ajXj(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)Pn C2Pn j=1 kajk2(cid:17)k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 C2k(cid:16)Pn E (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 C2k(cid:16)Pn j=1 kajk2(cid:17)k 2 2k ajXj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2k ajXj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) sphere and use the constants for p = 2k, which gives(cid:16)eb(2k)(cid:17)2k N(cid:1)k(cid:16) Γ(k+ N =(cid:0) 2 E (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) kajk2 =eb(2k)2k nXj=1 This holds for all k ∈ N and therefore for every summand in (2.1). Note thateb(2k) Now we apply Konig's and Kwapie´n's Khintchine inequality for variables on the see [3, Theorem 3]. We obtain does not depend on n. Therefore we get nXj=1 (cid:17), Γ N 2 ) 2k k . 2 1 C2k(cid:18) 2 N(cid:19)k Γ(k + N 2 ) ! . Γ( N 2 ) 1 k! 1 nXj=1 kajk2 2k ≤eb(2k) ajXj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)  ≤ j=1 kajk2 Eexp j=1 ajXj(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)Pn ∞Xk=0 C2Pn 2 ) = Γ(cid:0) N 2(cid:1)Qk 2(cid:1). l=1(cid:0)k − l + N N x)− N 2 (2.2) Note that Γ(k + N Consider the function f (x) := (1 − 2 (2.2) is the Taylor expansion of the function f at the point x = 1 2 . The right-hand side of inequality C 2 . KHINTCHINE INEQUALITY FOR SN −1-VARIABLES IN Lψ2 3 So we get 2  ≤ f(cid:18) 1 j=1 kajk2 Eexp (cid:13)(cid:13)(cid:13)Pn j=1 ajXj(cid:13)(cid:13)(cid:13) C2(cid:19) . C2Pn Nr 1 Now let C := b(N ) =q 2 . Then f(cid:0) 1 C 2(cid:1) = 2 and this proves that the inequality from Theorem 1.1 holds true. 1−( 1 2 ) 2 N 3. Proof of the optimality family of equidistributed random . Then the family of random variables 2 N j=1 1−( 1 2 ) 1√n Xj. In this section let Xj, j ∈ N be an i.i.d. variables on the sphere SN−1. Denote Yn :=Pn Nr 1 Lemma 3.1. Let C ≥q 2 2 , n ∈ N exp 1√n Xj(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)Pn  C (cid:19)2!p# < ∞. E" exp(cid:18)kYnk I(p) := sup n∈N is uniformly integrable. j=1 C Proof. According to [2, Theorem 6.19] it suffices to prove that for some p > 1, First note that for a N -dimensional Gaussian variable Z we have EhkXjk2ki = 1 ≤ EhkZk2ki. Using a theorem of Zolotarev [8, Theorem 3] this implies P (kYnk > t) ≤ exp(−N q(t)), where q(t) = 1 to 1, say γ ∈ ( 1 2 (t2 − ln t− 1). For large t we have t2 − ln t− 1 > γt2 for some γ close 2 , 1). Therefore we find I(p) = sup P exp pkYnk2 P(cid:18)kYnk > C2 ! > t! dt √ppln(t)(cid:19) dt C 1 2 C γ p dt. 0 n∈NZ ∞ n∈NZ ∞ ≤ 1 +Z ∞ = 1 + sup t− N 1 2 kZkψ2 = √2 . 2 N q1 − ( 1 2 ) So we can choose p ∈ (1, N Lemma 3.2. Let Z be a N -dimensional Gaussian variable. Then we have 2 C2γ) such that the latter integral is finite. (cid:3) 4 H. DIRKSEN Proof. Let C > √2. We compute E"exp kZk2 C2 !# = 2 ! dx C2 ! exp −kxk2 exp kxk2 C2(cid:19) dx exp− j(cid:18) 1 NXj=1 exp(cid:18)− C2 (cid:19)(cid:19) dt t2(cid:18) C2 − 2 2 − x2 1 1 1 = (2π)N/2ZRN (2π)N/2ZRN √2πZR NYj=1 =(cid:18) C2 C2 − 2(cid:19) N ≤ 2 if and only if C ≥r 2 1 2 = 2 . 1 1−( 1 2 ) . This proves the lemma. 2 N (cid:3) Now we have(cid:16) C 2 C 2−2(cid:17) N 2 Lemma 3.3. Let Z be a N -dimensional Gaussian variable. Then we have lim n→∞(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nXj=1 1 √n Xj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)ψ2 = kZkψ2 . Proof. Assume lim supn→∞ kYnkψ2 > kZkψ2. Then there exists a subsequence nk, k ∈ N and some ǫ > 0 such that kYnkkψ2 > kZkψ2 + ǫ. According to Lemma 3.1 the family exp(cid:18) kYnk kZkψ2 grable. Also Gn := exp(cid:18) kYnk kZkψ2 + ǫ(cid:19)2 − exp(cid:18) kZk +ǫ(cid:19)2!, n ∈ N is uniformly inte- kZkψ2 + ǫ(cid:19)2 , n ∈ N is uniformly integrable. For M > 0 we have Z Gn dP ≤Z{Gn≤M} Gn dP + sup n∈NZ{Gn>M} Gn dP. For every fixed M > 0, the first integral tends to 0 for n → ∞ by the central limit theorem. The second integral tends to 0 for M → ∞ due to the uniform integrability. Therefore n→∞Z exp kYnk kZkψ2 + ǫ!2 lim dP =Z exp kZk kZkψ2 + ǫ!2 dP. KHINTCHINE INEQUALITY FOR SN −1-VARIABLES IN Lψ2 5 This implies 2 ≥Z exp kZk kZkψ2! dP kZkψ2 + ǫ! dP >Z exp kZk n→∞Z exp kYnk k→∞Z exp kYnkk kZkψ2 + ǫ! dP kZkψ2 + ǫ! dP = lim = lim ≥ 2, which is a contradiction. Therefore lim supn→∞ kYnkψ2 ≤ kZkψ2. In the same way we show lim inf n→∞ kYnkψ2 ≥ kZkψ2 . This finishes the proof our Theorem. (cid:3) Acknowledgment. I thank Prof. H. Konig for his support and advice during my PhD studies. Part of my research was funded by DFG project KO 962/10-1. References [1] Uffe Haagerup, The best constants in the Khintchine inequality, Stud. Math. 70 (1982), 231 -- 283. [2] Achim Klenke, Probability theory. A comprehensive course, London: Springer, 2008. [3] Hermann Konig and Stanislaw Kwapie´n, Best Khintchine type inequalities for sums of inde- pendent, rotationally invariant random vectors, Positivity 5 (2001), no. 2, 115 -- 152. [4] Michel Ledoux and Michel Talagrand, Probability in Banach spaces: Isoperimetry and pro- cesses, Berlin: Springer, 1991. [5] Goran Peskir, Best constants in Kahane-Khintchine inequalities in Orlicz spaces, J. Multi- variate Anal. 45 (1993), no. 2, 183 -- 216. [6] Gilles Pisier, De nouvelles caract´erisations des ensembles de sidon, Adv. Math., Suppl. Stud. 7B (1981), 685-726. [7] V.A. Rodin and E.M. Semyonov, Rademacher series in symmetric spaces, Anal. Math. 1 (1975), 207 -- 222. [8] V. M. Zolotarev, Some remarks on multidimensional Bernstein-Kolmogorov-type inequalities, Theory Probab. Appl. 13 (1968), 281 -- 286, Translation by B. Seckler. Department of Mathematics, Kiel University E-mail address: [email protected]
1705.03643
1
1705
2017-05-10T07:59:20
Operateurs absolument continus et interpolation
[ "math.FA" ]
In the first part of this work, we study the absolutely continuous operators which are defined on fuction spaces with wide sense. In the second part, we show some results concerning the absoltely continuous operators when the function spaces (with wide sense) are interpolation spaces.
math.FA
math
OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION DAHER MOHAMMAD Abstract. Dans la deuxi`eme partie, on ´etablit quelques r´esultats concernant les op´erateurs absolument continus lorsque les espaces des fonctions (au sens large) sont des espaces d'interpolation. Abstract. Abstract. In the first part of this work, we study the absolutely continuous operators which are defined on fuction spaces with wide sense. In the second part, we show some results concerning the ab- soltely continuous operators when the function spaces (with wide sense) are interpolation spaces. Mots Cl´es:absolument continu Introduction. Soit X, Y deux espaces de Banach. D´esignons par L(X, Y ) les op´erateurs born´es de X `a valeurs dans Y et K(X, Y ) le sous-espace de L(X, Y ) form´es des op´erateurs compacts. Nous introduisons dans la premi`ere partie de ce travail, les op´erateurs absolument continus dans L(X, Y ) lorsque X est un espace de fonctions au sens large, cette d´efinition coincide avec celle de [Ben-Sh], si X est un espace de fonctions mesurables. Dans la suite, nous donnons des conditions n´ecessaires pour q'un op´erateur dans K(X, Y ) soit absolu- ment continu. Dans la deuxi`eme partie, nous ´etudions les op´erateurs absolument continus sur les espaces d'interpolation, comme des espaces de fonc- tions au sens large. Finalement, nous montrons que B∗ 1)θ isom´etriquement au sens large, pour tout couple d'interpolation (B0, B1) tel que B0 ∩ B1 est dense dans B0 et B1 et tout θ ∈ ]0, 1[ . θ =, (B∗ 0, B∗ 1. Op´erateurs absolument continus Soit Y un espace de Banach complexe, Y ∗ son dual. Pour y ∈ Y et y∗ ∈ X ∗ on note hy, y∗i = y∗(y). 1991 Mathematics Subject Classification. 46A32, 47L05, 46B70. 1 2 DAHER MOHAMMAD Dfinition 1. Soient Y un espace de Banach, (Ω, Σ, µ) un espace mesur´e et ∆Y : Σ × Y → Y une application. On dit que Y est un espace des fonctions au sens large sur (Ω, Σ, µ, ∆Y ) si ∆Y v´erifie les conditions suivantes: I) ∆Y (A ∩ B, f ) = ∆Y (A, ∆Y (f, B)), ∀A, B ∈ Σ et ∀f ∈ Y. II) ∆Y (A, αf + βg) = α∆Y (A, f ) + β∆Y (A, g), ∀A ∈ Σ, ∀f, g ∈ Y et ∀α, β ∈ C. III) ∆Y (A ∪ B, f ) = ∆Y (A, f ) + ∆Y (B, f ), ∀f ∈ Y et ∀A, B ∈ Σ tel que A ∩ B = ∅ . IV) Si µ(A) = 0 (A ∈ Σ), alors ∆Y (A, f ) = 0 et ∆Y (Ω, f ) = f , ∀f ∈ Y. V) Il existe une constante C > 0 telle que k∆Y (A, f )k ≤ C kf k , ∀f ∈ Y et ∀A ∈ Σ. Exemple 1. Soit Y un espace de fonctions mesurables sur [0, ∞[ (cf.[Ben-Sh]). On d´efinit ∆Y par ∆Y (A, f ) = f XA, (A, f ) ∈ Σ × Y. Il est facile de voir que ∆Y v´erifie les conditions I),II),III),IV),V). Pour (A, f ) ∈ Σ × Y on note ∆Y (A, f ) = f XA. Soit Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ). Pour f ∈ Y notons N(f ) = sup {kf XAk ; A ∈ Σ} . Il est ´evident que N(.) est une norme ´equivalente sur Y et N(f XA) ≤ N(f ) pour tout (A, f ) ∈ Σ × Y. On peut donc supposer que C = 1 dans V). Soit Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ). Pour Y ∗(A, f ∗)i = Y ∗(A, f ∗) ∈ Y ∗ par hf, ∆∗ (A, f ∗) ∈ Σ × Y ∗, on d´efinit ∆∗ h∆Y (A, f ), f ∗i pour tout f ∈ Y. Nous allons la proposition ´evidente suivante: Proposition 1. Soit Y un espace de fonctions au sens large sur (Ω , Σ, µ, ∆Y ). Alors Y ∗ est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆∗ Y ∗). Soient Y un espace de Banach et Z un sous-espace ferm´e de X. On note [x] l'image de x dans l'espace quotient X/Y. Dfinition 2. Soient Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et Z un sous-espace de Banach de Y . On dit que Z est stable par ∆Y , si ∆Y (A, f ) ∈ Z pour tout (A, f ) ∈ Σ × Z. Consid´erons Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et Z un sous-espace stable par ∆Y . On d´efinit ∆X/Y : Σ×Y /Z → Y /Z, par ∆Y /Z(A, [f ]) = [f XA] pour tout A ∈ Σ et tout f ∈ Y. Proposition 2. Soient Y est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et Z un sous-espace de Banach de Y stable par ∆Y . Alors Y /Z est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y /Z). OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 3 D´emonstration. Il est facile de voir que ∆Y /Z v´erifie I), II), III),IV). Montrons que ∆Y /Z v´erifie la condition V). Pour tout (A, f ) ∈ Σ × Y et tout g ∈ [f ] on a(cid:13)(cid:13)∆Y /Z(A, [f ])(cid:13)(cid:13)Y /Z = k[∆(A, f ])kY /Z ≤ k∆(A, g)kY ≤ kgk . Par cons´equent(cid:13)(cid:13)∆Y /Z(A, [f ])(cid:13)(cid:13)Y /Z ≤ k[f ]kY /Z.(cid:4) Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X),(Ω, Σ, , µ, ∆Y ) respectivement et T : X → Y est un op´erateur born´e. L'identification T (∆X) = ∆Y signifie que T [∆X (A, f )] = ∆Y (A, T (f )) pour tout (A, f ) ∈ Σ × X. Dfinition 3. Soit Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) : a) Soit f ∈ Y . On dit que f est absolument continu, si pour tout ε > 0, il existe δ > 0 tel que si µ(A) < δ, alors kf XAkY < ε. b) Une partie Z de Y est dite absolument continue, si pour tout f ∈ Z, f est absolument continu. c) Une partie Z de Y est dite uniform´ement absolument continue, si pour tout ε > 0, il existe δ > 0, tel que si µ(A) < δ, alors kf XAk < ε pour tout f ∈ Z. Soient Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et Z une partie de Y. Posons V = ℓ∞(Z, Y ). On d´efinit ∆V (A, (fi)i∈Z) = (fiXA)i∈Z, A ∈ Σ, (fi)i∈Z ∈ V. Il est clair que V est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆V ). Supposons que Z est uni- form´ement born´e. Il est facile de voir que Z est uniform´ement ab- solument continue, si et seulement si l'´el´ement (f )f ∈Z est absolument continu dans V. Remarque 1. Soient Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et Z une partie absolument continue dans Y. Alors l'adh´erence de Z dans Y est absolument continue. Pour tout f ∈ Y notons ν f (A) = (f XA), A ∈ Σ, la mesure ν f est finiment additive `a valeurs dans Y. Dfinition 4. Soient Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et f ∈ Y. On dit que la mesure ν f est µ−d´enombrablement addi- tive, si pour toute suite (Ak)k≥0 dans Σ, deux-`a-deux disjoints v´erifiant µ( ∪ k≥0 Ak) < +∞, alors νf ( ∪ k≥0 νf (Ak). Ak) =Xk≥0 Lemme 1. Soient Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et f ∈ Y. Les assertions suivantes sont ´equivalentes. 1) f est absolument continu. 4 DAHER MOHAMMAD 2) νf est µ−d´enombrablement additive. 3) Pour toute suite d´ecroissante (An)n≥0 dans Σ telle que µ(An) → n→∞ 0 dans Y. 0, alors νf (An) → n→∞ D´emonstration. 1) =⇒ 2). Fixons ε > 0. Il existe δ > 0 tel que si µ(A) < δ, k(f XA)kY < ε. Soient (Ak)k≥0 une suite dans Σ deux-`a-deux disjoints telle que µ( ∪ k≥0 µ(Ak) < δ pour tout n ≥ n0. ∞. Il existe n0 ≥ 1 tel que µ( ∪ k≥n Ak) =Xk≥n Ak) < Donc νf (Ak)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Ak)(cid:13)(cid:13)(cid:13)(cid:13) Ak) − νf ( ∪ k≤n k≥0 νf ( ∪ k≥0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Ak) −Xk≤n = (cid:13)(cid:13)(cid:13)(cid:13)νf ( ∪ k>n (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13)νf (∪Ak) )(cid:13)(cid:13)(cid:13)(cid:13) < ε. = (cid:13)(cid:13)(cid:13)(cid:13)(f X ∪Ak Ak ) =Xk≥0 k>n ν f (Ak).(cid:4) Par cons´equent νf ( ∪ k≥0 j≥k+1 2) =⇒ 3). Soit (An)n≥0 une suite d´ecroissante dans Σ telle que µ(An) → 0. Il existe une suite (Ank)k≥0 telle que µ(Ank ) < 2−k, ∀k ∈ N. Notons pour tout k ∈ N Bk = Ank − Ank+1. Il est clair que µ( ∪ Bk) < +∞, donc k≥0 n→+∞ νf ( ∪ k≥0 IV), ν f ( ∩ νf (Bk). Comme µ( ∩ Anj ) = 0 d'apr`es la condition Bk) =Xk≥0 Anj(cid:21) , ceci implique que ν f (Ank) = ν f ( ∪ Anj ) = 0 pour tout k. D'autre part, Ank = (cid:20) ∪ Bj(cid:21) ∪ (cid:20) ∩ Bj) =Xj≥k Bk)(cid:13)(cid:13)(cid:13)(cid:13) < )(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)ν f ( ∪ tout k. Fixons ε > 0, il existe k0 ∈ N v´erifiant(cid:13)(cid:13)(cid:13)νf (Ank0 ε. Il en r´esulte que kνf (An)k ≤(cid:13)(cid:13)(cid:13)νf (Akk0 )(cid:13)(cid:13)(cid:13) < ε pour tout n ≥ nk0.(cid:4) νf (Bj) pour j≥k+1 j≥k+1 k≥k0 j≥k j≥k OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 5 3) =⇒ 1). Supposons qu'il existe ε0 tel que pour tout k ∈ N, il existe Bk ∈ Σ, Bk) ≤ v´erifiant µ(Bk) < 2−k et kνf (Bk)k > ε0. Observons que µ( ∪ k≥n µ(Bk) < +∞. Notons pour tout n ∈ N An = ∪ k≥n Bk. Nous avons Xk≥n alors µ(An) → 0, donc il existe n0 tel que kνf (An0)k < ε0. n→+∞ D'autre par, Bn0 ⊂ An0, par cons´equent kνf (Bn0)k ≤ kν f (An0)k < ε0, ce qui est impossible. Il en r´esulte que f est absolument continu.(cid:4) Remarque 2. Dans le lemme 1, on a toujours 1) =⇒ 2) sans la con- dition V ) de la d´efinition 1. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X) et Y un espace de Banach. D´efinissons ∆L(X,Y )(A, T )(f ) = T XA(f ) = T (f XA), T ∈ L(X, Y ), A ∈ Σ, f ∈ X. Il est clair que ∆L(Y,X) v´erifie les conditions (I),(II),(III),IV),V), par cons´equent L(Y, X) est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆L(X,Y )). Soient X un espace de Banach et Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ). D´efinissons ∆′ L′(X,Y )(A, T )(f ) = T XA(f ) = (T (f ))XA = ∆Y (A, T (f )), T ∈ L(X, Y ), A ∈ Σ, f ∈ X. On v´erifie facilement que ∆′ L(Y,X) v´erifie les conditions (I),(II),(III),IV),V), donc L(Y, X) est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆′ L(X,Y )). Dfinition 5. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), Y un espace de Banach et T : X → Y un op´erateur born´e. On dit que T est absolument continu, si T est absolument continu par rapport `a (Ω, Σ, µ, ∆L(X,Y )). Dfinition 6. Soient X un espace de Banach, Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et T : X → Y un op´erateur born´e. On dit que T est absolument continu `a gauche, si T est absolument continu par rapport `a (Ω, Σ, µ, ∆′ L(X,Y )). Il est facile de montrer le lemme suivant: Lemme 2. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X) et T : X → Y un op´erateur born´e. Alors T est absolument continu si et seulment si T ∗ : Y ∗ → X ∗ est absolument continu `a gauche. Remarque 3. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X ), (Ω, Σ, µ, ∆Y ) respectivement et T : X → Y un op´erateur born´e. Si T (∆X ) = ∆Y , alors T est absolument continu si et seulement si T est absolument continu `a gauche. 6 DAHER MOHAMMAD Remarque 4. Soient Y un espace de fonctions au sens large sur (Ω, Σ, µ, ∆Y ) et K un compact dans Y pour la norme. Si K est ab- solument continue, alors K est uniform´ement absolument continue. Preuve. En effet, Consid´erons (Bn)n≥0 une suite d´ecroissante dans Σ telle que µ(Bn) → n→∞ 0. Soit f ∈ K. Il existe m = m(f ) tel que kf XBmkY < ε. Notons pour tout m ∈ N Om = {f ∈ K; kf XBmkY < ε}. D'apr`es ce qui pr´ec`ede (Om)m≥0 est un recouvrement ouvert de K, comme K est un com- pact, il existe m1, ..., mm ∈ N verifiant K ⊂ Om1 ∪ ... ∪ Omm. Noter m0 = max(m1, ..., mm). Montrons que kf XBnkY < ε pour tout f ∈ K et pour tout n ≥ m0. Soient n ≥ m0 et f ∈ K. Il existe mj tel que (cid:13)(cid:13)ν f (Bmj )(cid:13)(cid:13)Y < ε. On a alors kν f (Bn)kY ≤ (cid:13)(cid:13)νf (Bmj )(cid:13)(cid:13)Y < ε, car Bn ⊂ Bmj . Donc supf ∈K kνf (Bn)kY < ε. D'apr`es le lemme 1, (f )f ∈K est absolument continu dans ℓ∞(K, X). Finalement d'apr`es la remarque 3, K est uni- form´ement absolument continue.(cid:4) (cid:4) Dfinition 7. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X) et T : X → Y un op´erateur born´e. On dit que T est ponctuellement faiblement absolument continu (par rapport `a (Ω, Σ, µ, ∆X )), si pour tout f ∈ X, tout y∗ ∈ Y ∗ et tout ε > 0, il existe δ > 0 tels que si µ(A) < δ, alors hT (f XA), y∗i < ε. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X) et T : X → Y un op´erateur born´e. Pour tout f ∈ X on d´efinit la mesure νT (f ) par νT (f )(A) = T (f XA). Lemme 3. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X) et T : X → Y un op´erateur born´e. Supposons que T soit ponctuelle- ment faiblement absolument continu. Alors pour tout f ∈ X νT (f ) est µ−d´enombrablement additive. D´emonstration. Soit (Ak)k≥0 une suite de sous-ensembles mesurables deux-`a-deux µ(Ak) < +∞. Comme T est ponctuellement Ak).(cid:4) k≥0 faiblement absolument continu, d'apr`es la remarque 2 pour tout M ⊂ ). T (f XAk) converge faiblement dans Y vers T (f X ∪Ak k≥0 disjoints telle que Xk≥0 N, la serie Xk∈M D'apr`es [DU, Coroll.4,Chap.I-4], la s´erieXk≥0 tionnellement dans Y vers T (f X ∪ T (f XAk) converge incondi- OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 7 Dfinition 8. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), Y un espace de Banach et T : X → Y un op´erateur born´e. On dit que T ∗∗ est ponctuellement pr´efaiblement absolument continu, si la mesure A ∈ Σ → hy∗, T ∗∗(f ∗∗XA)i =hy∗, T ∗∗(∆∗∗ est µ−d´enombrablement additive, pour tout f ∗∗ ∈ X ∗∗. X ∗∗(A, f ∗∗))i = hy∗, ν T ∗∗(A)(f ∗∗)i Pour tout Banach X on note BX la boule unit´e ferm´ee de X. Lemme 4. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), Y un espace de Banach et T : X → Y un op´erateur born´e. Alors X ∗∗(A, f ∗∗))i = h(T XA)∗y∗, f ∗∗i pour tout (A, y∗) ∈ Σ × Y ∗ hy∗, T ∗∗(∆∗∗ et tout f ∗∗ ∈ X ∗∗. X ∗∗(A, , f ∗∗)i = h∆∗ X ∗∗(A, f ∗∗))i = X ∗∗(A, f ))i = hy∗, T (∆X(A, f ))i=h(T XA)∗y∗, f i, D´emonstration. Soit (A, y∗) ∈ Σ×Y ∗. Il est clair que l'application f ∗∗ → hy∗, T ∗∗(∆∗∗ X ∗(A, T ∗(y∗)), f ∗∗)i est pr´efaiblement con- hT ∗y∗, ∆∗∗ tinue : BX ∗∗ → C et l'application f ∗∗ → h(T XA)∗y∗, f ∗∗i est pr´efaiblement continue:BX ∗∗ → C. D'autre part, hy∗, T ∗∗(∆∗∗ pour tout f ∈ BX. Comme BX est pr´efaiblement dense dans BX ∗∗, alors hy∗, T ∗∗(∆∗∗ X ∗∗(A, f ∗∗))i = h(T XA)∗y∗, f ∗∗i , pour tout f ∗∗ ∈ X ∗∗.(cid:4) Lemme 5. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X) , Y un espace de Banach et T : X → Y un op´erateur ponctuelle- ment faiblement absolument continu. Supposons que µ est une mesure born´ee, X est s´eparable et ne contient pas ℓ1 isomorphiquement. Alors T ∗∗ est ponctuellement pr´efaiblement absolument continu. D´emonstration. Soient f ∗∗ ∈ X ∗∗ et y∗ ∈ Y ∗. Comme X ne contient pas ℓ1 isomor- phiquement, d'apr`es [Ros], il existe une suite born´ee (fn)n≥0 dnas X f ∗∗ pr´efaiblement dans X ∗∗. Pour tout n on d´efinit telle que fn → n→∞ la mesure νn par νn(A) = hT (fnXA), y∗i , A ∈ Σ. Remarquons d'apr`es hy∗, T ∗∗(f ∗∗XA)i pour tout A ∈ Σ, donc le lemme 4 que ν n(A) → n→∞ d'apr`es [DU, Cor.6,Chap.1-5], limµ(A)→0 νn(A) existe uniform´ement en n. Par cons´equent limµ(A)→0 hy∗, T ∗∗(f ∗∗XA)i = 0.(cid:4) Lemme 6. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), Y un espace de Banach et T : X → Y un op´erateur ponctuellement faiblement absolument continu. Supposons que X est un espace de Grothendieck. Alors T ∗∗ est ponctuellement pr´efiablement absolument continu. D´emonstration. Soient y∗ ∈ Y ∗ et (Ak)k≥0 une suite dans Σ telle que µ(Ak) → n→∞ 0. L'op´erateur T est ponctuellement faiblement continu, donc T ∗(y∗)XAn → n→∞ 8 DAHER MOHAMMAD 0 pr´efaiblement dans X ∗, comme X est un espace de Grothendieck, 0 faiblement [DU, p.179] dans X ∗, d'apr`es la remarque T ∗(y∗)XAn → n→∞ 2, T ∗∗ est ponctuellement pr´efiablement absolument continu.(cid:4) Proposition 3. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), Y un espace de Banach et T : X → Y un op´erateur compact. Supposons que T ∗∗ est ponctuellement pr´efaiblement absolu- ment continu. Alors T est absolument continu. D´emonstration. Il suffit de montrer que ν T est µ−d´enombrablement additive, d'apr`es le lemme 1. Consid´erons (Ak)k≥0 une suite dans Σ, deux-`a-deux disjoints. T ∗∗ est k∈M ponctuellement pr´efaiblement absolument continu, doncXk∈M Ak)(f ∗∗)(cid:29) , pour tout M ⊂ N. (cid:28)y∗, νT ∗∗ (∪ D'autre part, d'apr`es le lemme 4,Xk∈M Ak)(f ∗∗)(cid:29) =(cid:28)y∗, (T X ∪Ak et(cid:28)y∗, νT ∗∗ (∪ hy∗, (T XAk)∗∗(f ∗∗)i = (cid:28)y∗, (T X ∪Ak Xk∈M hνT (Ak), u∗i = (cid:28)ν T (∪ Coroll.3], Xk∈M que la s´erie Xk≥0 hy∗, ν T ∗∗(Ak)(f ∗∗)i =Xk∈M )∗∗(f ∗∗)(cid:29) . Il en r´esulte que )∗∗(f ∗∗)(cid:29) . Comme pour tout Ak), u∗(cid:29) pour tout M ⊂ N et A ∈ Σ, νT (A) = T XA est un op´erateur compact , alors d'apr`es [Kalt, tout u∗ ∈ [K(X, Y )]∗ . Ceci entraıne d'apr`es [DU, Chap.I-4,Coroll..4] Ak) νT (Ak) converge inconditionnellement vers νT ( ∪ k≥0 k∈M k∈M k∈M k∈M hy∗, νT ∗∗(Ak)(f ∗∗)i = hy∗, (T XAk)∗∗(f ∗∗)i dans L(X,Y).(cid:4) Corollaire 1. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X), (Ω, Σ, µ, ∆Y ) respectivement tel que T (∆X) = ∆Y et T : X → Y un op´erateur compact. Supposons que T soit ponctuellement faiblement absolument continu. Alors T est absolument continu. D´emonstration. D'apr`es la proposition 3, il suffit de montrer que T ∗∗ est ponctuelle- ment pr´efaiblement absolument continu. Etape 1: Soient y∗ ∈ Y ∗ et (An)n≥0 une suite de sous-ensembles Y ∗(Ak, y∗)] → 0. 0. Montrons que T ∗ [∆∗ mesurables telle que µ(An) → n→∞ OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 9 En ffet, pour tout f ∈ X on a hf, T ∗ [∆∗ h∆Y (Ak, T (f )), y∗i = hT (∆X(Ak, f )), y∗i . Y ∗(Ak, y∗)]i = hT (f ), ∆∗ Y ∗(Ak, y∗)i = Comme T est ponctuellement faiblement absolument continu, hT (∆X (Ak, f )), y∗i → k→∞ 0, c'est-`adire que T ∗ [∆∗ Y ∗(Ak, y∗)] → k→∞ 0 pr´efaiblement. D'autre part, T ∗ est un op´erateur compact et la suite (∆∗ Y ∗(Ak, y∗))k≥0 est born´ee, donc T ∗ [∆∗ Y ∗(Ak, y∗)] → k→∞ 0 en norme dans X ∗. Etape 2: Montrons que T ∗∗ est ponctuellement pr´efaiblement absol- ument continu. Pour cela, soit f ∗∗ ∈ X ∗∗. Remarquons que pour tout y∗ ∈ Y ∗ on a hy∗, T ∗∗(∆∗∗ X ∗∗(Ak, f ∗∗))i = hT ∗(y∗), ∆∗∗ X ∗∗(Ak, f ∗∗)i = h∆∗ X ∗(Ak, T ∗(y∗)), f ∗∗i = hT ∗ [∆∗ Y ∗(Ak, y∗)] , f ∗∗i → k→∞ 0, car T ∗(∆∗ Y ∗) = ∆∗ absolument continu.(cid:4) X ∗. Donc T ∗∗ est ponctuellement pr´efaiblement Remarque 5. Soit X un espace de Banach au sens large sur (Ω, Σ, µ, ∆X ), Y un espace de Banach, X ∗ est absolument continu et T : X → Y un op´erateur born´e. Alors T ∗∗ est ponctuellement pr´efaiblement absolu- ment continu Preuve. En effet, consid´erons A ∈ Σ, f ∗∗ ∈ X ∗∗ et y∗ ∈ Y ∗. Remarquons que hy∗, T ∗∗(f ∗∗XA)i = X ∗(A, T ∗(y∗)), f ∗∗i , donc la mesure A → hy∗, T ∗∗(f ∗∗XA))i est µ−d´enombrablement h∆∗ additive, car X ∗ est absolument continu.(cid:4) Corollaire 2. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X). Alors K(X, Y ) est absolument continu si et seulement si X ∗ est absolument continu. D´emonstration. Supposons que K(X, Y ) est absolument continu. Pour f ∗ ∈ X ∗, on d´efinit Uf ∗ : X → Y, par Uf ∗(f ) = hf, f ∗i y0, f ∈ X, o`u ky0kY = 1. Fixons f ∗ ∈ X ∗. Comme Uf ∗ est absolument continu, pour tout ε > 0, il existe δ > 0 telle que sup(cid:8)kUf ∗(f XA)kY ; f ∈ BX(cid:9) < ε, si µ(A) < δ. Choisissons y∗ ∈ Y ∗ tel que hy0, y∗i = 1. Pour tout f ∈ BX , on a kUf ∗(f XA)kY ≥ hUf ∗(f XA), y∗i = hf XA), f ∗i hy0, y∗i = hf, ∆∗ X ∗(A, f ∗)i , par cons´equent k∆X ∗(A, f ∗)k < ε, si µ(A) < δ; 10 DAHER MOHAMMAD Inversement, supposons que X ∗ est absolument continu. Soit T ∈ K(X, Y ). D'apr`es la remarque 5, T est ponctuellement pr´efaiblement absolument continu. En appliquant la proposition 3, on voit que T est absolument continu.(cid:4) Remarque 6. Dans [Ni], on introduit les op´erateurs absolument con- tinus, pour cette d´efinition tout op´erateur compact est absolument con- tinu, si X ne contient pas ℓ1 isomorphiquement. Corollaire 3. Soient X un espace de fonctions sur (Ω, Σ, µ, ∆X), Y un espace de Banach et T : X → Y un op´erateur compact et ponctuelle- ment faiblement absolument continu. Si Y a la propri´et´e de l'approximation born´ee, alors T est absolument continu. D´emonstration. Comme Y a la propri´et´e de l'approximation born´ee, il existe une suite g´en´eralis´ee (Ui)i∈I d'op´erateurs du rang finis:Y → Y , telle que supi∈I kUik < +∞ et Ui → IY uniform´ement sur tout compact de Y. Pour tout i ∈ I, notons Ti = Ui ◦ T. Comme Ti → T dans L(X, Y ), il suffit de montrer que Ti est absolument continu, d'apr`es la proposi- tion 3, il suffit de montrer que (Ti)∗∗ est ponctuellement pr´efaiblement absolument continu. Fixons i ∈ I. L'op´erateur Ti est du rang fini, il existe donc une suite k )k≤n dans X ∗ et une suite de vecteurs (yk)k≤n dans Y de vecteurs (f ∗ Consid´erons (Ak)k≥0 une suite de sous-ensembles mesurables deux- `a-deux disjoints et y∗ ∈ Y ∗. Comme Ti est ponctuellement faiblement hf, (TiXAk)∗(y∗)i = hf, f ∗ k i yk, f ∈ X. tels que Ti(f ) =Xk≤n absolument continu, d'apr`es la remarque 2, Xk≥0 hTi(f XAk), y∗i = (cid:28)Ti(f X ∪ Xk≥0 r´esulte que la s´erieXk≥0 k≥0 dans X ∗. D'autre part, l'op´erateur (TiXA)∗ est `a valeurs dans un sous- espace de dimension finie F0 de X ∗ ind´ependant de A, car TiXA est `a valeurs dans un sous espace de Y ind´ependant de A, par cons´equent la Ak)∗y∗ dans X ∗. (TiXAk)∗y∗ converge fortement vers (TiX ∪ k≥0 s´erieXk≥0 Soit f ∗∗ ∈ X ∗∗. D'apr`es le lemme 4,Xk≥0 i (XAkf ∗∗)i =Xk≥0 est ponctuellement pr´efaiblement absolument continu.(cid:4) donc T ∗∗ hy∗, T ∗∗ i Ak), y∗(cid:29) = (cid:28)f, (TiX∪Ak k≥0 )∗y∗(cid:29) . Il en (TiXAk)∗y∗ converge pr´efaiblement vers (TiX ∪ k≥0 Ak)∗y∗ h(TiXAk )∗y∗, f ∗∗i, OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 11 Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X ), Y un espace de Banach et T : X → Y un op´erateur born´e. Notons ET ={T XA; A ∈ Σ} et FT le sous-espace ferm´e engendr´e par ET dans L(X, Y ). Proposition 4. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), µ une mesure born´ee, Y un espace de Banach et T : X → Y un op´erateur born´e ponctuellement faiblement absolument continu. Supposons que pour tout ξ ∈ (FT )∗, il existe une suite (ξn)n≥0 dans X ⊗ Y ∗ telle que ξn → ξ pr´efaiblement dans (FT )∗. Alors T est n→∞ absolument continu. D´emonstration. Soit ξ ∈ (FT )∗. Il existe une suite (ξn)n≥0 dans X ⊗Y ∗ telle ξn → n→∞ ξ pr´efaiblement. Fixons n ∈ N. On d´efinit la mesure νn par ν n(A) = hT XA, ξni . D'apr`es l'hypoth`ese, limµ(A)→0 νn(A) = 0 pout tout n ∈ N. D'autre part, limn→∞ ν n(A) = hT XA, ξi pour tout A ∈ Σ, d'apr`es[DU, limµ(A)→0 ν n(A) = 0, uniform´ement en n ∈ N. Il Cor.6,Chap.1-5], en r´esulte que limµ(A)→0 hT XA, ξi = 0. Soit (Ak)k≥0 une suite dans Σ deux-`a-deux disjoint. D'apr`es la remarque 2, pour tout M ⊂ N Ak, ξE , pour tout ξ ∈ (FT )∗. Il en r´esulte T XAk converge inconditionnellement vers T X∪Ak dans k≥0 hT XAk, ξi = DT X ∪ Xk∈M que la s´erie Xk≥0 k∈M L(X, Y ) [DU, Corol.6,Chap.1-4].(cid:4) Corollaire 4. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), µ une mesure born´ee, Y un espace de Banach et T : X → Y ∗ un op´erateur born´e ponctuellement faiblement absolument ∧ ⊗ Y ne contient continu. Supposons que X, Y sont s´eparables et que X pas ℓ1 isomorphiquement. Alors T est absolument continu. D´emonstration. D'apr`es [DU, Chap.VIII-2,Coroll.2]-[Sch], (X ∧ ξ ∈ [L(X, Y ∗)]∗ = (X ⊗Y )∗∗. Comme X ∧ ⊗Y telle ξn → [Ros], il existe une suite (ξn)n0 dans X n→∞ Pour conclure, il suffit d'appliquer la proposition 4.(cid:4) ∧ ⊗Y )∗ = L(X, Y ∗). Soit ∧ ⊗Y ne contient pas ℓ1, d'apr`es ξ pr´efaiblement. Proposition 5. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), µ une mesure born´ee, Y un espace de Banach et T : X → Y un op´erateur absolument continu. Supposons que FT ne contient pas c0. Alors FT est un espace W CG. 12 DAHER MOHAMMAD D´emonstration. Comme T est absolument continu, d'apr`es le lemme 1, ν T est µ−d´enombrablement additive. D'autre part, FT ne contient pas c0 et ν T est born´ee, d'apr`es [DU, Chap.1-4,Th.2], ν T est fortement additive. En appliquant le r´esultat de [DU, Chap.I-5,Coroll.3], on voit que ET est faiblement com- pact, donc FT est un espace W CG.(cid:4) Proposition 6. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), µ une mesure born´ee, Y un espace de Banach et T : X → Y un op´erateur born´e . Supposons que FT est un espace W CG et T est ponctuellement faiblement absolument continu. Alors T est absolument continu. D´emonstration. Comme FT est un espace W CG, il existe un compact faible K dans FT tel que l'espace ferm´e engendr´e par K est ´egale `a FT . Soit J : B(FT )∗ → C(K) l'injection canonique. Consid´erons maintenant ξ dans la boule unit´e de (FT )∗. D'apr`es le th´eor`eme de Hahn-Banach, il existe ∧ η dans la boule unit´e de [L(X, Y ∗∗)]∗ qui prolonge ξ. Comme (X ⊗ Y ∗)∗ est l'espace L(X, Y ∗∗) [DU, Chap.VIII-2,Coroll.2]-[Sch], il existe ∧ ⊗ Y ∗ telle que une suite g´en´eralis´es (ηi)i∈I dans la boule unit´e de X ηi → η σ([L(X, Y ∗∗)]∗ , L(X, Y ∗∗)). D´esignons pour tout i ∈ I par ξi la restriction de ηi `a FT . Nous avons alors ξi → ξ σ((FT )∗, FT ), J(ξi) → J(ξ) pour la topologie de la convergence simple dans C(K). D'apr`es [Groth], il existe une suite (ξin)n≥0 telle que J(ξin) → J(ξ) n→∞ ξ, σ((FT )∗, FT ), d'apr`es la dans (C(K), τ p), ceci implique que ξin → n→∞ proposition 4, T est absolument continu.(cid:4) Remarque 7. Dans la proposition 6 on peur remplacer FT est un espace W CG par l'existence d'un espace W CG de L(X, Y ) qui contient FT . Proposition 7. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), Y un espace de Banach et T : X → Y une application lin´eaire. Supposons que T soit p−sommant pour un p ∈ [1, ∞ [ et X ∗ est absolument continu. Alors T est absolument continu. D´emonstration. Il suffit de montrer que νT est µ−d´enomrablement additive, d'apr`es le lemme 1. Soit (Ak)k≥0 une suite dans Σ deux-`a-deux disjoints. Comme T est p−sommant, il existe une mesure positive G sur la boule unit´e de X ∗ toute f ∗ ∈ BX ∗. En appliquant le th´eor`eme de convergence domin´ee, dG(f ∗) (1.1) dG(f ∗). → 0 pour m,n→∞ dG(f ∗). → 0 (Obser- m,n→∞ X ∗( m ∪ k=n (Ak, f ∗)(cid:13)(cid:13)(cid:13)X ∗ ≤ → 0, c'est-`a-dire que OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 13 hf, f ∗idG(f ∗). Donc pour tout m ≥ n telle que kT (f )k ≤ C ZBX∗ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ν T (Ak)(f )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) mXk=n ∆∗ ∆∗ dG(f ∗) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) hf XAk), f ∗i(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ C ZBX∗ mXk=n (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) X ∗(Ak, f ∗)+(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) *f, = C ZBX∗ mXk=n X ∗(Ak, f ∗)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X ∗ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ C kf k ZBX∗ mXk=n X ∗(Ak, f ∗)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X ∗ Comme X ∗ est absolument continu,(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) mXk=n X ∗(Ak, f ∗)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X ∗ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) nous d´eduisons que ZBX∗ mXk=n X ∗(Ak, f ∗)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X ∗ vons que pour tout m ≥ n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)∆∗ mXk=n ν T (Ak)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1). Il en r´esulte d'apr`es (1.1) que(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) mXk=n la s´erieXk≥0 νT (Ak) converge en norme dans L(X, Y ).(cid:4) m,n→∞ ∆∗ ∆∗ ∆∗ Soit X, Y deux espaces de Banach. D´esignons par Π(X, Y ) l'espaces des op´erateurs T : X → Y tel que T est p−sommant pour un p ∈ [1, +∞[ et par L0(X, Y ) l'adh´erence de Π(X, Y ) dans L(X, Y ). Corollaire 5. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X) et Y un espace de Banach. Supposons que X ∗ soit absol- ument continu. Alors L0(X, Y ) est absolument continu. 2. Espaces de fonctions au sens large pour les espaces d'interpolation Soient B = (B0, B1) un couple d'interpolation au sens de [Ber-Lof, chap. II], et θ ∈ ]0, 1[ . 14 DAHER MOHAMMAD Soit S = {z ∈ C; 0 ≤ Re(z) ≤ 1} et S0 = {z ∈ C; 0 < Re(z) < 1} . On d´esigne par F (B) l'espace des fonctions F `a valeurs dans B0 + B1, continues born´ees sur S, holomorphes sur S0, telles que, pour j ∈ {0, 1}, l'application τ → F (j + iτ ) est continue `a valeurs dans Bj et kF (j + iτ )kBj →τ →∞ 0. On le munit de la norme kF kF (B) = max{supτ ∈R kF (iτ )kB0 , supτ ∈R kF (1 + iτ )kB1}. Notons F0(B0, B1) =( nXk=0 Fk ⊗ bk; bk ∈ B0 ∩ B1, Fk ∈ F (C), n ∈ N) . L'espace (B0, B1)θ = Bθ =(cid:8)F (θ); F ∈ F (B)(cid:9) est de Banach [Ber-Lof, th.4.1.2], pour la norme d´efinie par kakBθ = infnkF kF (B) ; F (θ) = ao . Toute F ∈ F (B) est repr´esent´ee `a partir de ses valeurs au bord en utilisant la mesure harmonique, de densit´e Q0(z, iτ ) et Q1(z, 1 + iτ ), z ∈ S0, τ ∈ R, [Ber-Lof, section 4.5]: F (z) =ZR F (iτ )Q0(z, τ )dτ +ZR F (1 + iτ )Q1(z, τ )dτ . (2.1) On note G(B) l'espace des fonctions g `a valeurs dans B0 + B1 , continues sur S, holomorphes `a l'interieur de S, telles que kg(z)kB0+B1 supz∈S (C) (C ′) g(j + iτ ) − g(j + iτ ′) ∈ Bj, ∀τ , τ ′ ∈ R, j ∈ {0, 1} et la quantit´e (1+z) < ∞. suivante est finie: supτ 6=τ ′∈R(k(g(iτ ) − g(iτ ′))/τ − τ ′kB0), kgkQG(B) = max(cid:20) supτ 6=τ ′∈R(k(g(1 + iτ ) − g(1 + iτ ′))/τ − τ ′kB1) (cid:21) . L'espace (B0, B1)θ = Bθ =(cid:8) g′(θ); g ∈ G(B)(cid:9) est de Banach [Ber-Lof, Ceci d´efinit bien une norme sur QG(B) le quotient de G(B) (par les constante `a valeurs dans B0 + B1). th.4.1.4] pour la norme d´efinie par kakBθ = infnkgkQG(B) ; g′(θ) = ao . Soit p ∈ [1, +∞[. L'espace d'interpolation Bθ,p est d´efini par Bθ,p =(a ∈ B0 + B1; kakBθ,p =(cid:20)ZR+ (K(a, t)/tθ)pdt/t(cid:21)1/p < +∞) o`u OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 15 ) est un espace de Banach [Ber-Lof, th 3.4.2]. K(a, t) = inf(cid:8)ka0kB0 + t ka1kB1 ; a = a0 + a1, aj ∈ Bj, j ∈ {0, 1}(cid:9) . (Bθ,p, k.kBθ,p Dfinition 9. Soit (B0, B1) un couple d'interpolation. Supposons que j ∈ Bj est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆j), {0, 1} . On dit que (B0, B1) est un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)) si ∆0(A, f ) = ∆1(A, f ), pour tout A ∈ Σ et tout f ∈ B0 ∩ B1. Il est facile de montrer la proposition suivante: Proposition 8. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)) et fj, gj ∈ Bj, j ∈ {0, 1} . Supposons que f0 + f1 = g0 + g1. Alors ∆0(A, f0) + ∆1(A, f1) = ∆0(A, g0) + ∆1(A, g1) pour tout A ∈ Σ. On d´efinit (∆0 + ∆1)(A, f ) = ∆0(A, f0) + ∆1(A, f1), o`u A ∈ Σ et f = f0 + f1 ∈ B0 + B1. On remarque que B0 + B1 est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆0 + ∆1). Exemple 2. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X),(Ω, Σ, , µ, ∆Y ) respectivement et i : X → Y une injection continue. Supposons que i(∆X ) = ∆Y . Alors (X, Y ) est un couple d'interpolation compatible avec (Ω, Σ, µ, (∆X , ∆Y )). Proposition 9. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)) et θ ∈ ]0, 1[ . Alors Aθ est un espace de fonc- tions au sens large sur (Ω, Σ, µ, ∆θ), o`u ∆θ(A, f ) = (∆0 + ∆1)(A, f ), (A,f ) ∈ Σ × Bθ. D´emonstration. Soit (A, f ) ∈ Σ×Bθ. Montrons que ∆θ(A, f ) ∈ Bθ et k∆θ(A, f )kBθ ≤ kf kBθ . Il existe F ∈ F (B0, B1) telle que F (θ) = f. L'op´erateur u → (∆0 + ∆1)(A, u) est born´e, donc l'application FA : z ∈ S → (∆0 + ∆1)(A, F (z)) ∈ B0+B1 est holomorphe sur S0. Comme (∆0+∆1)(A, (F (j+ i.)) = ∆j(A, F (j + i.)), j ∈ {0, 1} , FA ∈ F (B0, B1). Il est clair que kFAkF (B0,B1) ≤ kF kF (B0,B1) . Donc k∆θ(A, f )kBθ ≤ kf kBθ .(cid:4) Proposition 10. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)), θ ∈ ]0, 1[ et p ∈ ]1, +∞[ . Alors Aθ,p est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆θ,p), o`u ∆θ,p(A, f ) = (∆0 + ∆1)(A, f ), (A, f ) ∈ Σ ×Bθ,p. 16 DAHER MOHAMMAD D´emonstration. Soient (A, f ) ∈ Σ × Bθ,p et ε, t > 0. Il existe f0 ∈ B0 et f1 ∈ B1 tel que f = f0 + f1, et K(f, t) + ε > kf0kB0 + t kf1kB1 . D'autre part, ∆θ,p(A, f ) = ∆0(A, f0) + ∆1(A, f1), donc K [∆θ,p(A, f ), t] k≤ ∆0(A, f0)kB0 + t k∆1(A, f1)kB1 ≤ kf0kB0 + t kf kB1 ≤ K(f, t) + ε. Il en r´esulte que ∆θ,p(A, f ) ∈ Bθ et k∆θ,p(A, f )kBθ,p ≤ kf kBθ,p .(cid:4) Remarque 8. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X),(Ω, Σ, , µ, ∆Y ) respectivement, i : X → Y une injec- tion continue, θ ∈ ]0, 1[ et p ∈ ]1, +∞[ . Supposons que i(∆X ) = ∆Y . D'apr`es la proposition 9, (X, Y )θ est un espace de fonctions au sens large sur (Ω, Σ, µ, ∆θ). Nous remarquons que pour tout 0 < θ < β < 1, i(∆θ) = ∆β, o`u i : (X, Y )θ → (X, Y )β l'injection canonique. Proposition 11. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X ),(Ω, Σ, µ, ∆Y ) respectivement et i : X → Y une injec- tion absolument continu, d'image dense. Si i(∆X ) = ∆Y , alors pour tout 0 < θ < β < 1, i : (X, Y )θ → (X, Y )β est absolument continu. D´emonstration. Etape 1: Soit β ∈ ]0, 1[ . Montrons que i : (X, Y )β → Y est absolu- ment continu. Soit ε > 0. Comme i : X → Y est absolument continu, il existe δ > 0 tel que si µ(A) < δ, alors kνi(A))kL(X,Y ) < ε 1 1−θ . Consid´erons f ∈ X tel que kf k(X,Y )θ < 1. Il existe F ∈ F0(X, Y ) v´erifiant F (θ) = f et kF kF (X,Y ) < 1. Fixons un ensemble mesurable A de Ω tel µ(A) < δ. La fonction z ∈ S → F (z)XA = (∆0 + ∆1)(A, F (z)) ≤ ZR ≤ ZR ZR kνi(A))kL(X,Y ) × kF (iτ )kX Q(θ, iτ ) 1 − θ kF (1 + iτ kY Q(θ, 1 + iτ ) θ ≤ ε. θ dτ OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 17 est dans F (X, Y ) ⊂ F (Y ), donc d'apr`es [Ber-Lof, Lemme 4.3.2] ki [F (θ)XA]kY = k∆Y (A, iF (θ))kY kF (iτ )XAkY Q(θ, iτ ) 1 − θ 1−θ dτ ×ZR θ )dτ kF (1 + iτ )XAkY Q(θ, 1 + iτ θ 1−θ dτ × Etape 2: Soit β ∈ ]0, 1[ . Montrons que i : X → (X, Y )β est absolu- ment continu. D'apr`es le lemme 2 et la remarque 3, i∗ : Y ∗ → X ∗ est absolument continu, donc d'apr`es l'´etape 1 i∗ : (X ∗, Y ∗)β → X ∗ est absolument continu. En r´eappliquant le lemme 2 et la remarque 3, on voit que i∗∗ : X ∗∗ → [(X ∗, Y ∗)β]∗ est absolument continu. D'autre part, d'apr`es [Da], la restriction de i∗∗ `a X est `a valeurs dans (X, Y )β et (X, Y )β est un sous-espace ferm´e de [(X ∗, Y ∗)β]∗. Donc i : X → (X, Y )β est absolument continu. Etape 3: Soit θ, β ∈ ]0, 1[ tel que θ < β. Montrons que i : (X, Y )θ → (X, Y )β est absolument continu. D'apr`es l'´etape 1, i : (X, Y )θ → Y est absolument continu, en appli- quant l'´etape 2, on voit que i : (X, Y )θ → [(X, Y )θ, Y ]η est absolument continu, pour tout η ∈ ]0, 1[ . Choisissons (1−η)θ = β. D'apr`es le th´eor`eme de r´eit´eration [Ber-Lof, Th.4.6.1], [(X, Y )θ, Y ]η = (X, Y )β. On en d´eduit que i : (X, Y )θ → (X, Y )β est ponctuellement absolument continu.(cid:4) Lemme 7. Supposons que i: X → Y soit une injection continue d'image dense. Alors pour tout 0 < θ < 1 et tout 1 < p < +∞ (X, Y )θ,p est un sous-espace ferm´e de [(X ∗, Y ∗)θ,p′]∗, ou p′ est le con- jugu´e de p. D´emonstration. Il est ´evident que l'injection :((X, Y )θ,p → [(X ∗, Y ∗)θ,p′]∗ est continue. D'autre part, d'apr`es [Ber-Lof, Th.3.7.1], il existe une constante C > 0, telle que kf ∗k(X ∗,Y ∗)θ,p′ ≤ C kf ∗k[(X,Y )θ,p]∗ pour tout f ∗ ∈ [(X, Y )θ,p]∗ . Soient f ∈ (X, Y )θ,p et ε > 0. Il existe f ∗ dans la boule unit´e de 18 DAHER MOHAMMAD [(X, Y )θ,p]∗ tel que kf k(X,Y )θ,p f ∗ ∈ (X ∗, Y ∗)θ,p′ et kf ∗k(X ∗,Y ∗)θ,p′ ≤ C, d'o`u le lemme.(cid:4) ≤ hf, f ∗i + ε. D'apr`es ce qui pr´ec`ede, Par un argument analogue `a celui de la proposition 11 (en utilisant le lemme 7) on montre: Proposition 12. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X ),(Ω, Σ, µ, ∆Y ) respectivement et i : X → Y une injec- tion absolument continu d'image dense. Supposons que i(∆X ) = ∆Y . Alors pour tout 0 < θ < β < 1 et tout 1 < p < +∞, i : (X, Y )θ,p → (X, Y )β,p est absolument continu. Lemme 8. Soient X un espace de fonctions au sens large sur (Ω, Σ, µ, ∆X), Y un espace de Banach et i : X → Y une injection ponctuellement faiblement absolument continu, θ ∈ ]0, 1[ et p ∈ ]1, +∞[. Alors i : X → (X, Y )θ,p est ponctuellement faiblement absolument continu. D´emonstration. Soient f ∈ X, x∗ ∈ [(X, Y )θ,p]∗ et ε, ε′ > 0 tel que kik kf k ε′ + ε′ ≤ ε . D'apr`es [Ber-Lof, Th.3.7.1] [(X, Y )θ,p]∗ = (X ∗, Y ∗)θ,p′. D'autre part, le th´eor`eme 3.4.2 de [Ber-Lof] nous montre que Y ∗ est dense dans [(X, Y )θ,p]∗ , par cons´equent il existe z∗ ∈ Y ∗ tel que kx∗ − z∗k [(X,Y )θ,p]∗ < ε′. D'apr`es l'hypoth`ese, il existe δ > 0 tel que si µ(A) < δ, hi(f XA), z∗)i < ε′. Soit A ∈ Σ tel que µ(A) < δ. On a alors hi(f XA), x∗i = hi(f XA), x∗ − z∗ + z∗i ≤ hi(f XA), x∗ − z∗i + hi(f XA), z∗i ≤ × kx∗ − z∗k[(X,Y )θ,p]∗ + ε′ ≤ ki(f XA)k(X,Y )θ,p kik kf k ε′ + ε′ ≤ ε. Il en r´esulte que i : X → (X, Y )θ,p est faiblement absolument continu.(cid:4) Lemme 9. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X), (Ω, Σ, µ, ∆Y ) respectivement, θ ∈ ]0, 1[ , i : X → Y une injection ponctuellement faiblement absolument continu et p ∈ ]1, +∞[ . Supposons que i(∆X ) = ∆Y . Alors i : (X, Y )θ,p → Y est ponctuelle- ment faiblement absolument continu. D´emonstration. Soient f ∈ (X, Y )θ,p, y∗ ∈ Y ∗ et ε, ε′ > 0 tel que ε′ + ky∗k kik ε′ ≤ ε. < ε′. D'apr`es l'hypoth`ese il Il existe f1 ∈ X tel que kf − f1k(X,Y )θ,p existe δ > 0, tel que si µ(A) < δ hi(f1XA), y∗i < ε′, . Choisissons A ∈ Σ tel que µ(A) < δ. On a alors hi(f XA), y∗i ≤ hi(f1XA), y∗i + ky∗k ki(f − f1)AkY < ε′ + kik ky∗k ε′ ≤ ε.(cid:4) OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 19 Par un argument analogue `a celui de la proposition 11 (en utilisant le th´eor`eme de r´eit´eration [Ber-Lof, Th.3.5.3] et les lemmes 8, 9) on tire le corollaire suivant: Corollaire 6. Soient X, Y deux espaces de fonctions au sens large sur (Ω, Σ, µ, ∆X), (Ω, Σ, µ, ∆Y ) respectivement, 0 < θ < β < 1, p ∈ ]1, +∞[ et i : X → Y une injection ponctuellement faiblement absol- ument continu. Supposons que i(∆X ) = ∆Y . Alors i : (X, Y )θ,p → (X, Y )β,p est ponctuellement faiblement absolument continu. Proposition 13. Soient (B0, B1), (C0, C1) deux couples d'interpolation compatibles avec (Ω, Σ, , µ, (∆0, ∆1)), (Ω, Σ, µ, (∆2,∆3)) respectivement, θ ∈ ]0, 1[ et T : Bj → Cj un op´erateur born´e, j ∈ {0, 1} (avec ). Supposons que T : B0 → C0 est absolument T0B0∩B1 continu. Alors T : Bθ → Cθ est absolument continu. = T1B0∩B1 D´emonstration. Soit A ∈ Σ. Observons que T XA : Bj → Cj est born´e, d'apr`es [Ber-Lof, Th.4.1.2] kT XAkBθ→Cθ Donc T : Bθ → Cθ est absolument continu.(cid:4) ≤(cid:2)kT XAkB0→C0(cid:3)1−θ×(cid:2)kT XAkB1→C1(cid:3)θ . Proposition 14. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)), θ ∈ ]0, 1[. Supposons que B0 est absolument continu. Alors Bθ est absolument continu. D´emonstration. D'apr`es la remarque 1, il suffit de montrer que B0∩B1 est absolument continue dans Bθ. Pour cela, soient f ∈ B0 ∩ B1 et F ∈ F0(B0, B1) tels que F (θ) = f. Posons FA(z) = (∆0 + ∆1)(A, F (z)), A ∈ Σ, z ∈ S, . Il est clair que FA ∈ F (B0, B1), d'apr`es [Ber-Lof, Lemme.4.3.2], pour tout A ∈ Σ nous avons kFA(θ)kBθ k∆0(A, F (iτ )kB0 ≤ ZR ×ZR Q(θ, iτ ) 1 − θ 1−θ dτ k∆1(A, F (1 + iτ )kB1 Q(θ, 1 + iτ ) θ θ dτ .(2.2) 20 DAHER MOHAMMAD Soit (An)n≥0 une suite dans Σ telle que µ(An) → n→∞ est absolument continu pour tout τ ∈ R ∆0(An, F (iτ )) → n→∞ quant le th´eor`eme de convergence domin´ee, on voit queZR 0. Il en r´esulte d'apr`es (2.2) que ∆θ(An, f ) = FAn(θ) → n→∞ 0. Comme B0 0, en appli- k∆0(An, F (iτ )kB0 0.(cid:4) Q(θ,iτ ) 1−θ dτ → n→∞ Dfinition 10. Soient X, Y deux espaces de fonctions sur (Ω, Σ, µ, ∆X), (Ω, Σ, µ, ∆Y ) respectivement. On dit que X, Y sont isom´etriques au sens large, s'il existe un op´erateur isom´etrie surjectif T : X → Y v´erifiant T (∆X) = ∆Y . Remarque 9. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)), θ, α, β, η ∈ ]0, 1[ tels que θ = (1 − η)α + ηβ. Alors Aθ = (Aα, Aβ)η isom´etriquement au sens large. Preuve. En effet, Il est facile de voir que (Aα, Aβ) est compatible avec (Ω, Σ, µ, (∆α, ∆β)). par rapport `a (Ω, Σ, µ, (∆α, ∆β)), f ∈ B0 ∩ B1 et A ∈ Σ. Consid´erons e∆η l'application qui d´efinie l'espace de fonctions (Bα, Bβ)η clair que e∆η(A, f ) = ∆α(A, f ) = ∆β(A, f ) = ∆0(A, f ) = ∆θ(A, f ). isom´etriques et e∆η(A, f ) = ∆θ(A, f ) pour tout A ∈ Σ et tout f ∈ Bθ, D'autre part, d'apr`es le th´eor`eme de r´eit´eration Bθ et (Bα, Bβ)η sont Pour tout g ∈ G(B0, B1) et tout A ∈ Σ notons gA(z) = (∆0 + car B0 ∩ B1 est dense dans (Bα, Bβ)η = Aθ.(cid:4) Il est ∆1)(A, g(z)), z ∈ S. Lemme 10. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)), θ ∈ ]0, 1[ , A ∈ Σ et g ∈ G(B0, B1). Alors gA ∈ G(B0, B1) et kgAkG ≤ kgkG . D´emonstration. Il est clair qu gA est holomorphe sur S0 et continue sur S `a valeurs dans B0 + B1 , car (∆0 + ∆1)(A, .) est un op´erateur born´e sur B0 + B1. D'autre part, pour tout τ , τ ′ ∈ R on a gA(j + iτ ) − gA(j + iτ ′) = ∆j(A, g(j + iτ )) − ∆j(A, g(j + iτ ′) = ∆j(A, g(j + iτ )) − g(j + iτ ′)) ∈ Bj, (2.3) j ∈ {0, 1} . La relation (2.3) montre que gA v´erifie les condtions C, C ′ et kgAkG ≤ kgkG .(cid:4) Remarque 10. Pour tout θ ∈ ]0, 1[ et tout g ∈ G(B0, B1) on a gA(A, .)′(θ) = (∆0 + ∆1)(A, g′(θ)). OPR ´ERATEUR ABSOLUMENT CONTINUES ET INTERPOLATION 21 D'apr`es le lemme 10 et la remarque 10, on a la proposition suivante: Proposition 15. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)) et θ ∈ ]0, 1[ . Alors Bθ est un espace de fonction au sens large sur (Ω, Σ, µ, ∆θ), o`u ∆θ(A, f ) = (∆0+∆1)(A, f ), (A, f ) ∈ Σ × Bθ. Lemme 11. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)) et θ ∈ ]0, 1[ . Si B0 ∩ B1 est dense dans B0 et B1, alors (B∗ 1) est compactible avec (Ω, Σ, µ, (∆∗ 0, B∗ 0, ∆∗ 1)). D´emonstration. Soient f ∗ ∈ B∗ 0 ∩ B∗ 0(A, f ∗)i = 1(A, f ∗)i pour tout f ∈ B0 ∩ B1. D'autre part, d'apr`es [Ber-Lof, 1(A, f ∗) dans 1 et A ∈ Σ. Remarquons que hf, ∆∗ 0(A, f ∗) = ∆∗ 1, donc ∆∗ 0 + B∗ hf, ∆∗ Th.2.7.1] (B0 ∩ B1)∗ = B∗ B∗ 0 + B∗ 1.(cid:4) Proposition 16. Soient (B0, B1) un couple d'interpolation compatible avec (Ω, Σ, µ, (∆0, ∆1)) et θ ∈ ]0, 1[ Supposons que B0 ∩ B1 soit dense dans B0 et B1. Alors B∗ 1)θ isom´etriquement au sens large. θ = (B∗ 0, B∗ D´emonstration. D'apr`es le th´eor`eme de dualit´e [Ber-Lof, Th.4.5.1], on a isom´etriquement θ = (B∗ θ , f ∈ B0 ∩ B1 et ∆θ l'application 0, B∗ 1) tel 1)θ. Il existe g ∈ G(B∗ 0, B∗ 0, B∗ B∗ 1)θ. Consid´erons f ∗ ∈ B∗ qui d´efinie l'espace de fonctions (B∗ que g′(θ) = f ∗. Pour tout A ∈ Σ on a hf, ∆∗ θ(A, f ∗)i = h∆θ(A, f ), g′(θ)i . D'autre part, ∆θ(A, g′(θ)) = (∆∗ 0 + ∆∗ 1)(A, g′(θ)) et il existe b∗ j ∈ B∗ j telle que g′(θ) = b∗ 0 + b∗ 1. Ceci implique que (cid:10)f, ∆θ(A, g′(θ)(cid:11) = hf, (∆∗ 0 + ∆∗ 0(A, b∗ = hf, ∆∗ = h∆θ(A, f ), (b∗ = hf, ∆∗ θ(A, g′(θ))i 1)(A, g′(θ))i 1)i = h∆0(A, f ), b∗ 1(A, b∗ 0) + ∆∗ 0 + b∗ 1i = h∆θ(A, f ), (g′(θ)i = 0i + h∆1(A, f ), b∗ 1i Il en r´esulte que hf, ∆∗ et A ∈ Σ, donc ∆θ =∆∗ B∗ θ = (B∗ 0, B∗ θ(A, f ∗i =(cid:10)f, ∆θ(A, g′(θ)(cid:11) pour f ∈ B0 ∩ B1 1 = (B0 ∩ B1)∗, c'est-`a-dire que θ dans B∗ 0 + B∗ 1)θ isom´etriquement au sens large.(cid:4) References [Ber] J. Bergh, On the relation between the two complex methods of interpo- lation, Indiana Univ. Math. J. 28, 775-777, (1979). [Ben-Sh] C. Bennet, R. Sharpley, Interpolation of operators, Academie Press, (1988). 22 DAHER MOHAMMAD [Ber-Lof] J. Bergh, J. Lofstrom, Interpolation spaces an introduction, Springer- [Da] [Cal] [DU] Verlag-Berlin Heidelberg New York, (1976). M. Daher, Une remarque sur les espaces d'interpolation faiblement lo- calement uniform´ement convexes, arXiv:1206.4848. A. P. Calder´on, Intermediate spaces and interpolation, the complex method, Studia Math. 24, 113-190, (1964). J. Diestel, J. J. Uhl, Vector measures, Math. Surveys 15 A.M.S, (1977). [Groth] A. Grothendieck, Crit`eres de compacit´e dans les espaces fonctionnels [Kalt] [Ni] [Ros] [Sch] g´en´eraux, Amer. J. Math. 168-186, (1952).) N. J. Kalton, Spaces of compact operators, Math. Ann. 208, 267-278, (1974). Niculescu P. Consantin, Absolute in Banach spaces theory, Rev. Roum. Pure Appl. Vol. 24, 413-422, (1979). H. P. Rosenthal, A characterzation of spaces containing ℓ1, Proc. Nat. Sci. (U. S. A), Vol. 71, No. 2, 411-2413, (1974). R. Schatten, A theory of cross spaces, Princetton Univ. Press, (1950). E-mail address: [email protected]
1303.3711
2
1303
2013-03-31T13:18:33
Correction of a proof in "Connes' embedding conjecture and sums of hermitian squares"
[ "math.FA" ]
We show that Connes' embedding conjecture (CEC) is equivalent to a real version of the same (RCEC). Moreover, we show that RCEC is equivalent to a real, purely algebraic statement concerning trace positive polynomials. This purely algebraic reformulation of CEC had previously been given in both a real and a complex version in a paper of the last two authors. The second author discovered a gap in this earlier proof of the equivalence of CEC to the real algebraic reformulation (the proof of the complex algebraic reformulation being correct). In this note, we show that this gap can be filled with help of the theory of real von Neumann algebras.
math.FA
math
CORRECTION OF A PROOF IN "CONNES' EMBEDDING CONJECTURE AND SUMS OF HERMITIAN SQUARES" SABINE BURGDORF†, KEN DYKEMA∗, IGOR KLEP, AND MARKUS SCHWEIGHOFER Abstract. We show that Connes' embedding conjecture (CEC) is equivalent to a real version of the same (RCEC). Moreover, we show that RCEC is equivalent to a real, purely algebraic statement concerning trace positive polynomials. This purely algebraic reformulation of CEC had previously been given in both a real and a complex version in a paper of the last two authors. The second author discovered a gap in this earlier proof of the equivalence of CEC to the real algebraic reformulation (the proof of the complex algebraic reformulation being correct). In this note, we show that this gap can be filled with help of the theory of real von Neumann algebras. 3 1 0 2 r a M 1 3 ] . A F h t a m [ 2 v 1 1 7 3 . 3 0 3 1 : v i X r a 1. Introduction and erratum for [4] Alain Connes stated in 1976 the following conjecture [2, Section V]. Conjecture 1.1 (CEC). If ω is a free ultrafilter on N and F is a II1-factor with separable predual, then F can be embedded into an ultrapower Rω of the hyperfinite II1-factor R. The last two authors gave a purely algebraic statement which is equivalent to Conjecture 1.1, cf. statements (i) and (ii) in [4, Thm. 3.18]. Before stating this we recall some notation used in [4]. For k ∈ {R, C}, khXi denotes the polynomial ring in n non-commuting self- adjoint variables X = (X1, . . . , Xn) over k, which is equipped with the natural involution f 7→ f ∗, i.e. it is the natural involution on k, fixes each Xi and reverses the order of words. Then Mk denotes the quadratic module in khXi generated by {1 − X 2 i i = 1, . . . , n}. Two cyc ∼ g) if f − g is a sum of polynomials f, g ∈ khXi are said to be cyclically equivalent (f commutators in khXi. Theorem 1.2 (Klep, Schweighofer). The following statements are equivalent: (a) CEC is true. (b) If f ∈ ChXi and if tr(f (A)) ≥ 0 for all tuples A of self-adjoint contractions in Cs×s for all s ∈ N, then for every ε ∈ R>0, f + ε is cyclically equivalent to an element of MC. In the same paper the authors gave the following real version of the purely algebraic reformulation of Connes' embedding conjecture. Theorem 1.3. The following statements are equivalent: (a) CEC is true. (b) If f = f ∗ ∈ RhXi and if tr(f (A)) ≥ 0 for all tuples A of symmetric contractions in Rs×s for all s ∈ N, then for every ε ∈ R>0, f + ε is cyclically equivalent to an element of MR. Date: March 13, 2013. 2000 Mathematics Subject Classification. 46L10 (11E25, 13J30). Key words and phrases. Connes' embedding problem, real von Neumann algebras. †Research partially supported by ERC. ∗Research supported in part by NSF grant DMS-1202660. 1 2 BURGDORF, DYKEMA, KLEP, AND SCHWEIGHOFER However, their proof of the implication (b) =⇒ (a) in Theorem 1.3 is not correct. The incorrect part of that argument is Proposition 2.3 of [4]; in fact, the polynomial f = i(X1X2X3 − X3X2X1) provides a counter-example to the statement of that proposition. In this note, we present a proof of Theorem 1.3 that uses real von Neumann algebras as well as techniques and results from [4]. An inspection of the proof of [4, Proposition 2.3] shows that a weaker version of it remains true, namely the version where R is replaced by C in its formulation. This weaker version is enough for the first application of [4, Proposition 2.3], namely, in the proof of [4, Theorem 3.12]. It is the second application, namely, in the proof of [4, Theorem 3.18], that is illegitimate and will be circumvented by this note. In particular, it will follow that the statements of all results in [4] are correct with the exception of [4, Proposition 2.3]. We note that Narutaka Ozawa [6] provides a proof that uses similar (though not precisely identical) methods to the one presented here; our proofs were begun independently and, initially, completed independently; however, after release of a first version of this paper, Ozawa noticed a problem with our proof and provided a result (Proposition 2.7 below) that fixed it, which he kindly allows us to print here. 2. Real von Neumann algebras Real von Neumann algebras were first systematically studied in the 1960s by E. Størmer (see [8], [7]). They are closely related to von Neumann algebras with involutory ∗-antiauto- morphisms. Through this relation much of the structure theory of von Neumann algebras (e.g. the type classification and the integral decomposition into factors) can be transferred to the real case. See, for example, [1] (or [5], where the definition is slightly different but easily seen to be equivalent). Definition 2.1. A real von Neumann algebra Mr is a unital, weakly closed, real, self- adjoint subalgebra of the (real) algebra bounded linear operators on a complex Hilbert space, with the property Mr ∩ iMr = {0}. Remark 2.2. A basic fact of real von Neumann algebras (see, e.g., the introduction of [1], or references cited therein) is that they correspond to (complex) von Neumann algebras with involutory ∗-antiautomorphism. An involutory ∗-antiautomorphism on a von Neumann algebra M is a complex linear map α : M → M satisfying α(x∗) = α(x)∗, α(xy) = α(y)α(x) and α2(x) = x for all x, y ∈ M. This correspondence works as follows. (i) Let Mr be a real von Neumann algebra. Then the (complex) von Neumann algebra U (Mr) generated by Mr is equal to the complexification of Mr [8, Thm. 2.4], i.e. U (Mr) = M′′ r = Mr + iMr. Moreover, the involution ∗ on Mr generates a natural involutory ∗-antiautomorphism α on U (Mr) by α(x + iy) = x∗ + iy∗. (ii) Let M be a von Neumann algebra with involutory ∗-antiautomorphism α, the ∗- subalgebra Mα = {x ∈ M α(x) = x∗} is then a real von Neumann algebra. In fact, let x = iy ∈ Mα ∩ iMα with x, y ∈ Mα, then x∗ = α(x) = iα(y) = (−y)∗ = −x∗ and thus x∗ = 0, which implies Mα ∩ iMα = {0}. Since ∗ and α are continuous in the weak topology, Mα is weakly closed. One then easily sees that U (Mα) = M and U (Mr)α = Mr. CONNES' EMBEDDING CONJECTURE 3 Definition 2.3. Let Mr be a real von Neumann algebra. (i) Mr is called a real factor if its center Z(Mr) consists of only the real scalar operators. (ii) Mr is said to be hyperfinite if there exists an increasing sequence of finite dimensional real von Neumann subalgebras of Mr such that its union is weakly dense in Mr. (iii) Mr of type In, I∞, II1, etc. if its complexification U (Mr) is of the corresponding type. We immediately see that a real von Neumann algebra Mr is a factor if and only if U (Mr) is a factor. If Mr is a real hyperfinite factor, then U (Mr) is a hyperfinite factor [1, Prop. 2.5.10]. But the converse implication does not hold in general, see e.g. [1, Ch. 2.5]. The situation is different for the hyperfinite II1-factors. There is (up to isomorphism) a unique real hyperfinite II1-factor [9, Thm. 2.1] (see also [3]). Theorem 2.4 (Størmer). Let M be a type II1-factor and Mr a real factor such that M = Mr + iMr. Then the following conditions are equivalent. (i) Mr is hyperfinite. (ii) Mr is the weak closure of the union of an increasing sequence {Rn} of real unital factors such that Rn is isomorphic to R2n ×2n . (iii) Mr is countably generated, and given x1, . . . , xn ∈ Mr and ε > 0 there exist a finite dimensional real von Neumann subalgebra Nr of Mr and y1, . . . , yn ∈ Nr such that kyk − xkk2 < ε for all k = 1, . . . , n. From now on we will denote the unique real hyperfinite II1-factor by Rr. Remark 2.5. The correspondence between R and Rr in the hyperfinite case of Remark 2.2 is given by the involutory ∗-antiautomorphism on R which is induced by the matrix transpose. To be more specific, with R the closure of the infinite tensor product of matrix algebras C2×2 and letting t2 be the matrix transpose on C2×2, then we define α to be the 1 t2; see the involutory ∗-antiautomoprhism on R that when restricted to N∞ also [9, Cor. 2.10]. Then Rα = Rr and hence R = Rr + iRr. C2×2 is ⊗∞ 1 The construction of the ultrapower of the real hyperfinite II1-factor Rr with trace τ works as in the complex case, see e.g. [9]. Let ω be a free ultrafilter on N. Parallel to ℓ∞(R) in the complex case consider the real C ∗-algebra ℓ∞(Rr) = {(rk)k∈N ∈ RN r supk∈N krkk < ∞}. krk)1/2 = 0}. Then Jω is a closed maximal Further let Jω = {(rk)k ∈ ℓ∞(Rr) limk→ω τ (r∗ ideal in ℓ∞(Rr). The quotient C ∗-algebra Rω r := ℓ∞(Rr)/Jω is called the ultrapower of Rω r and is a finite real von Neumann algebra. Since Iω = Jω + iJω, where Iω is the closed ideal in ℓ∞(R) used for the construction of Rω, we have Rω = (Rr + iRr)ω = Rω r + iRω r . A final topic in this section is related to generating sets of real von Neumann algebras. For Mr a real von Neumann algebra, an element x of Mr is said to be symmetric if x∗ = x and antisymmetric if x∗ = −x. Clearly, writing an arbitrary x ∈ M as x = x + x∗ 2 + x − x∗ 2 , (1) every element of Mr is the sum of of a symmetric and an antisymmetric element. Lemma 2.6. Let Mr be any real von Neumann algebra and consider the real von Neumann algebra M2(Mr) endowed with the usual adjoint operation: c d(cid:19)∗ (cid:18)a b c∗ = (cid:18)a∗ b∗ d∗(cid:19) . (2) 4 BURGDORF, DYKEMA, KLEP, AND SCHWEIGHOFER Then M2(Mr) has a generating set consisting of symmetric elements. Proof. Let S ⊆ Mr be a generating set for Mr. Using the trick (1), we may without loss of generality assume S = Ssym ∪ Santisym, where Ssym consists of symmetric elements and Santisym of antisymmetric elements. Then (cid:26)(cid:18)1 0 0 0(cid:19) , (cid:18)0 1 0 0(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) 1 0(cid:19)(cid:27) ∪(cid:26)(cid:18)x 0 is a generating set for M2(Mr) consisting of symmetric elements. x ∈ Ssym(cid:27) ∪(cid:26)(cid:18) 0 x ∈ Santisym(cid:27) x −x 0(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) We thank Narutaka Ozawa for providing the proof of the following proposition and for allowing us to present it here. Proposition 2.7. Every real von Neumann algebra of type II1 is isomorphic to M2(Mr), for some real von Neumann algebra Mr, endowed with the usual adjoint operation (2). Proof. Let N be a (complex) von Neumann algebra of type II1 endowed with an involutory ∗-antiautomorphism α and let Nr = {x ∈ N α(x) = x∗} be the corresponding real von Neumann algebra. It will suffice to find a partial isometry v ∈ Nr such that v∗v + vv∗ = 1, for this will imply that Nr has the desired 2×2 matrix structure. By a maximality argument, it will suffice to find a nonzero partial isometry w ∈ Nr so that w∗w ⊥ ww∗. Let E be the center-valued trace on N . Let p ∈ N be any nonzero projection such that E(p) ≤ 1 3 and let q = p ∨ α(p). Then q ∈ Nr and E(q) ≤ 2 3 . There is a nonzero element x ∈ N such that x∗x ≤ q and xx∗ ≤ 1 − q. Multiplying by the imaginary number i, if necessary, we may without loss of generality assume y = x + α(x∗) is nonzero. But y ∈ Nr and y = (1 − q)yq. In the polar decomposition y = wy of y, we have that w ∈ Nr is a nonzero partial isometry, w∗w ≤ q and ww∗ ≤ (1 − q). Combining the previous proposition and lemma, we get: Corollary 2.8. Every real von Neumann algebra of type II1 is generated by a set of sym- metric elements. 3. The real version of Connes' embedding conjecture and proof of the real algebraic reformulation Here is the real version of Connes' embedding conjecture: Conjecture 3.1 (RCEC). If ω is a free ultrafilter on N and Fr is a real II1-factor with separable predual, then Fr can be embedded into an ultrapower Rω r of the real hyperfinite II1-factor. We will prove that RCEC is equivalent to CEC, in the course of proving Theorem 1.3. We will use two preliminary results. The first of these is the following real analogue of [4, Prop. 3.17]. Proposition 3.2. For every real II1-factor Fr with separable predual and faithful trace τ , the following statements are equivalent: (i) For every free ultrafilter ω on N, Fr is embeddable in Rω r ; (ii) There is an ultrafilter ω on N such that Fr is embeddable in Rω r ; (iii) For each n ∈ N and f = f ∗ ∈ RhXi, having positive trace on all symmetric con- tractions in Rs×s, s ∈ N, implies that τ (f (A)) ≥ 0 for all symmetric contractions A ∈ F n r ; CONNES' EMBEDDING CONJECTURE 5 (iv) For all ε ∈ R>0, n, k ∈ N and symmetric contractions A1, . . . , An ∈ Fr, there is an s ∈ N and symmetric contractions B1, . . . , Bn ∈ Rs×s such that for all w ∈ hXik: τ (w(A1, . . . , An)) − Tr(w(B1, . . . , Bn)) < ε, where Tr is the normalized trace on s × s matrices. Proof. The implication (i) =⇒ (ii) is obvious. The proof of (ii) =⇒ (iii) =⇒ (iv) works as in the complex case (see the proof of [4, Prop. 3.17]), where for (ii) =⇒ (iii) one uses the fact (see Theorem 2.4) that Rr is generated by a union of an increasing sequence of real matrix algebras. For the implication (iv) =⇒ (i), by Corollary 2.8, Fr has a generating set A1, A2, . . . that is a sequence of symmetric contractions. The rest of the proof then works as in the complex case, cf. also Theorem 2.4(iii). For a von Neumann algebra M, we let Mop denote the opposite von Neumannn algebra of M. This is the algebra that is equal to M as a set and with the same ∗-operation, but with multiplication operation · defined by a · b = ba. It is well known that Mop is also a von Neumann algebra, (and it is very easy to see this in the case that M has a normal faithful tracial state). A ∗-antiautomorphism of M is precisely an isomorphism M → Mop of von Neumann algebras. The following proposition is true for arbitrary von Neumann algebras, but since here we need it only for finite von Neumann algebras, for convenience and ease of proof we state it only in this case. Proposition 3.3. Let M be a von Neumann algebra with a normal, faithful, tracial state τ . Then the von Neumann algebra tensor product M⊗Mop has an involutory ∗- antiautomorphism. Proof. Let N = M⊗Mop and let α : M ⊗ Mop → N be the linear map defined on the algebraic tensor product that satisfies α(a ⊗ b) = b ⊗ a. Then α is ∗-preserving, and antimultiplicative. Indeed, we have α ((a ⊗ b)(c ⊗ d)) = α(ac ⊗ db) = db ⊗ ac = (d ⊗ c)(b ⊗ a) = α(c ⊗ d) α(a ⊗ b). Since α is trace-preserving, it extends to an isomorphism N → N op of von Neumann algebras, i.e., a ∗-antiautomorphism of N . Moreover, it is clear that α2 = id. Now we are ready to prove Theorem 1.3. For convenience, the two conditions in that theorem are restated here along with a third, which we will show are all equivalent. Theorem 3.4. The following statements are equivalent: (a) CEC is true. (b) If f = f ∗ ∈ RhXi and if tr(f (A)) ≥ 0 for all tuples A of symmetric contractions in Rs×s for all s ∈ N, then for every ε ∈ R>0, f + ε is cyclically equivalent to an element of MR. (c) RCEC is true. Proof. As remarked after the statement of Theorem 1.3, (a) =⇒ (b) was proved in [4]. The implication (b) =⇒ (c) follows from Proposition 3.2. Indeed, if Fr is a real II1-factor with separable predual, then (b) implies that condition (iii) of Proposition 3.2 holds; by (i) of that proposition, it follows that Fr embeds in Rω r . For (c) =⇒ (a), we assume that RCEC is true and we will show that every II1-factor F with separable predual can be embedded into Rω. Suppose first that F has an involutory 6 BURGDORF, DYKEMA, KLEP, AND SCHWEIGHOFER ∗-antiautomorphism α. Then F can be written as F = Fr + iFr, where Fr is the real II1- factor inside F corresponding to α as in Remark 2.2. By RCEC there exists an embedding ι of Fr into Rω r . This implies by C-linear extension of ι that F = Fr + iFr embeds into r = (Rr + iRr)ω = Rω. Now if F is any II1-factor, by the above case and Rω Proposition 3.3, the II1-factor F⊗F op embeds in Rω, and from the identification of F with F ⊗ 1 ⊆ F⊗F op, we get that F embeds in Rω. r + iRω References [1] S. Ayupov, A. Rakhimov, and S. Usmanov, Jordan, real and Lie structures in operator algebras, Math- ematics and its Applications, vol. 418, Kluwer Academic Publishers Group, Dordrecht, 1997. [2] A. Connes, Classification of injective factors. Cases II1, II∞, IIIλ, λ 6= 1, Ann. of Math. (2) 104 (1976), 73 -- 115. [3] T. Giordano, Antiautomorphismes involutifs des facteurs de von Neumann injectifs. I, J. Operator The- ory 10 (1983), 251 -- 287. [4] I. Klep and M. Schweighofer, Connes' embedding conjecture and sums of Hermitian squares, Adv. Math. 217 (2008), 1816 -- 1837. [5] B. Li, Real operator algebras, World Scientific Publishing Co. Inc., River Edge, NJ, 2003. [6] N. Ozawa, About conjecture -- algebraic approaches, the Connes embedding available at http://arxiv.org/abs/1212.1703. [7] E. Størmer, On anti-automorphisms of von Neumann algebras, Pacific J. Math. 21 (1967), 349 -- 370. [8] , Irreducible Jordan algebras of self-adjoint operators, Trans. Amer. Math. Soc. 130 (1968), 153 -- 166. [9] , Real structure in the hyperfinite factor, Duke Math. J. 47 (1980), 145 -- 153. S.B., EPFL, SB-MATHGEOM-EGG, Station 8, 1015 Lausanne, Switzerland E-mail address: [email protected] K.D., Department of Mathematics, Texas A&M University, College Station, TX 77843-3368, USA E-mail address: [email protected] I.K., Department of Mathematics, The University of Auckland, Private Bag 92019, Auck- land 1142, New Zealand E-mail address: [email protected] M.S., FB Mathematik und Statistik, Universitat Konstanz, D-78457 Konstanz, Germany E-mail address: [email protected]
1510.08241
4
1510
2016-07-06T16:20:24
A Geometrical Stability Condition for Compressed Sensing
[ "math.FA" ]
During the last decade, the paradigm of compressed sensing has gained significant importance in the signal processing community. While the original idea was to utilize sparsity assumptions to design powerful recovery algorithms of vectors $x \in \mathbb{R}^d$, the concept has been extended to cover many other types of problems. A noteable example is low-rank matrix recovery. Many methods used for recovery rely on solving convex programs. A particularly nice trait of compressed sensing is its geometrical intuition. In recent papers, a classical optimality condition has been used together with tools from convex geometry and probability theory to prove beautiful results concerning the recovery of signals from Gaussian measurements. In this paper, we aim to formulate a geometrical condition for stability and robustness, i.e. for the recovery of approximately structured signals from noisy measurements. We will investigate the connection between the new condition with the notion of restricted singular values, classical stability and robustness conditions in compressed sensing, and also to important geometrical concepts from complexity theory. We will also prove the maybe somewhat surprising fact that for many convex programs, exact recovery of a signal $x_0$ immediately implies some stability and robustness when recovering signals close to $x_0$.
math.FA
math
A Geometrical Stability Condition for Compressed Sensing ∗ Axel Flinth† Institut fur Mathematik Technische Universitat Berlin May 31, 2021 Abstract During the last decade, the paradigm of compressed sensing has gained significant importance in the signal processing community. While the original idea was to utilize sparsity assumptions to design powerful recovery algorithms of vectors x ∈ Rd, the concept has been extended to cover many other types of problems. A noteable example is low-rank matrix recovery. Many methods used for recovery rely on solving convex programs. A particularly nice trait of compressed sensing is its geometrical intuition. In recent papers, a classical optimality condition has been used together with tools from convex geometry and probability theory to prove beautiful results concerning the recovery of signals from Gaussian measurements. In this paper, we aim to formulate a geometrical condition for stability and robustness, i.e. for the recovery of approximately structured signals from noisy measurements. We will investigate the connection between the new condition with the notion of restricted singular values, classical stability and robustness conditions in compressed sensing, and also to important geometrical con- cepts from complexity theory. We will also prove the maybe somewhat surprising fact that for many convex programs, exact recovery of a signal x0 immediately implies some stability and robustness when recovering signals close to x0. Keywords: Compressed Sensing, Convex Geometry, Grassmannian Condition Number, Sparse Recovery. MSC(2010): Primary: 52A20, 90C25. Secondary: 94A12. 1 Introduction Suppose that we are given linear measurements b ∈ Rm of a signal x0 ∈ Rd, i.e. b = Ax0 for some matrix A ∈ Rm,d, and are asked to recover the signal from them. If d > m, this will not be trivial, since the map x0 (cid:55)→ b in that case won't be injective. If, however, one assumes that x0 in some sense is sparse, e.g., that many of x0's entries vanish, we can still recover the signal, e.g. with the help of (cid:96)1-minimization [8]: 6 1 0 2 l u J 6 ] . A F h t a m [ 4 v 1 4 2 8 0 . 0 1 5 1 : v i X r a min(cid:107)x(cid:107)1 subject to Ax = b (P1) This is the philosophy of compressed sensing, an area of mathematics which has achieved major attention over the last decade. It has become a standard technique to choose A at random, and then to ask the question how large the number of measurements m has to be in order for (P1) to be successful with high probability. A popular assumption is that A has the Gaussian distribution, i.e., that the entries are i.i.d. standard normally distributed. A widely used criterion to ensure that (P1) is successful is the RIP -property. Put a bit informally, a matrix A is said to possess the RIP -property if its RIP -constants; δk = min δ > 0 : ∀x k-sparse : (1 − δ)(cid:107)x(cid:107)2 2 ≤ (cid:107)Ax(cid:107)2 2 ≤ (1 + δ)(cid:107)x(cid:107)2 ∗This article has been accepted for publication in Linear Algebra and its Applications. †E-mail: [email protected] 10.1016/j.laa.2016.04.017 It can be found via its DOI 1 (cid:110) (cid:111) , 2 are small. The idea of using convex programs like (P1) to recover structured signals has come to be used in a much wider sense than the one above. Some examples of structure assumptions that have been considered in the literature are dictionary sparsity [10], block sparsity [20], sparsity with prior information [13, 16, 17], saturated vectors (i.e. vectors with x(i) = (cid:107)x(cid:107)∞ for many i) [14] and low-rank assumptions for matrix completion [7]. Although these problems may seem very different at first sight, they can all be solved with the help of a convex program of the form min f (x) subject to Ax = b, (Pf ) where f is some convex function defined on an appropriate space. In all of the mentioned examples above, f is chosen to be a norm, but this is not per se necessary. The connection between the different convex program approaches was thoroughly investigated in [9], in which the very general case of f being an atomic norm was investigated. An atomic norm (cid:107)·(cid:107)A is thereby the gauge function of the convex hull of a compact set A; (cid:107)x(cid:107)A = sup{t > 0 tx ∈ conv A} . The authors derive bounds on how many Gaussian measurements are required for a convex optimization problem of the form (Pf ) to be able to recover a signal x0 with high probability. Their arguments have a geometrical flavor, since they utilize a well known optimality condition regarding the descent cone D((cid:107)·(cid:107)A , x0) (see Definition 1 and Lemma 3 below) of (cid:107)·(cid:107)A. Other important theoretical tools are Gaussian widths and Gordon's escape through a mesh lemma [15], which will be discussed in Section 3.3 of this article. The authors of [3] make a similar analysis for even more general functions f , only assuming that they are convex. They use the so called statistical dimension of the descent cone for determining the threshold value of measurements. The Problem of Stability and Robustness In applications, the linear measurements b are often contaminated with noise. This means that we are actually given data b = Ax + n, where n is a noise vector. A popular assumption is that n is bounded in (cid:96)2-norm, i.e. (cid:107)n(cid:107)2 ≤ . Moreover, it is not often not entirely realistic to assume that the signal x0 is exactly sparse (or more generally does not exactly have the structure assumed by the model), but rather that the distance to the set of sparse (structured) signals C is small, i.e. (1) where (cid:107)·(cid:107)∗ is some norm. If the above quantity is small for a vector x0, we will call it approximately structured. There are several approaches to approximately recover approximately structured signals from noisy measure- dist(cid:107)·(cid:107)∗ (x0,C) = inf c∈C (cid:107)x0 − c(cid:107)∗ , ments -- the one we will consider is the following regularized convex optimization problem: (P  f ) This approach was investigated already in the earliest works on compressed sensing, where f = (cid:107)·(cid:107)1 and C = Sk = {k-sparse signals} [8]. In short, it turns out that (somewhat stronger) assumptions on the RIP -constants suffice to prove that the solution x∗ of (P  min f (x) subject to (cid:107)Ax − b(cid:107)2 ≤ . f ) in this case satisfies a bound of the form (cid:107)x − x∗(cid:107)2 ≤ κ1 dist(cid:107)·(cid:107)1 (x0,Sk) + κ2, where κ1 and κ2 depend on the RIP -constants of A. In this paper, we will formulate and investigate a geometrical criterion for general convex programs (P  f ) to satisfy such a bound, at least when f is a norm. We will call such programs robust (with respect to noise) and stable (with respect to distance of x0 to C). The criterion, which we will call the Angular Separation Criterion or ASC, will only depend on the relative positions of the kernel of A 2 and the descent cone D(f, x0). More specifically, we will prove that if the mentioned sets have a positive angular separation, (P  f ) will be robust and, in the case of f being a norm, stable. We will furthermore relate the ASC to known criteria for stability and robustness from the literature. We will also prove the somewhat remarkable fact that for a very large class of norms, the ASC is in fact implied by the ability of (Pf ) to recover signals exactly from noiseless measurements. The idea originates from a part of the Master Thesis [11] of the author. Related Work and Contributions of This Paper Some research towards a geometrical understanding of the stability of compressed sensing has already been conducted. Here we list a few of the approaches that have been considered. First, we would like to mention the so called RIP -N SP -condition from the recent paper [5]. This paper only deals with the classical compressed sensing setting, i.e., that the signal of interest is approximately sparse and (cid:96)1-minimization is used to recover it. A matrix A is said to satisfy the RIP -N SP -condition if there exists another matrix B, which has the RIP -property, such that ker A = ker B. The authors of [5] prove that this is enough to secure stability and robustness of (P  1). This is intriguing, as it shows that stability and robustness can be secured by only considering the kernel of the measuring matrix. Another line of research is the so called Robust Width Property, which was developed in [6]. The authors of said article define compressed sensing spaces, a general framework that covers both different types of sparsity in Rd as well as the case of low rank matrices. The main part of the definition of a compressed sensing space is a norm decomposability condition; if (cid:107)·(cid:107) is the norm induced by the inner product of the Hilbert space H, and (cid:107)·(cid:107)∗ is another norm on H, (H,C,(cid:107)·(cid:107)∗) is said to be a compressed sensing space with bound L if for every a in the subset C ⊆ H and z ∈ H, there exists a decomposition z = z1 + z2 such that (cid:107)a + z(cid:107)∗ = (cid:107)a(cid:107)∗ + (cid:107)z1(cid:107)∗ , (cid:107)z2(cid:107)∗ ≤ L(cid:107)z(cid:107) . A matrix is then said to satisfy the (ρ, α)-robust width property if for every z ∈ H with (cid:107)Az(cid:107) ≤ α(cid:107)z(cid:107), we have (cid:107)z(cid:107) ≤ ρ(cid:107)z(cid:107)∗. This robust width property is in fact equivalent with the stability and robustness of (P  f ) with f (x) = (cid:107)x(cid:107)∗. One large difference between this approach and ours is that we do not require any norm decomposability conditions. For instance, as was pointed out in [5], the (cid:96)∞-norm in Rd is not well suited to be included in this framework. As for the problem of stability of recovering almost sparse vectors, we want to mention the paper [21]. The authors of that article carry out an asymptotic analysis of the threshold amount of Gaussian measurements needed for the classical technique of (cid:96)1-minimization for sparse recovery to be stable. The analysis heavily relies on the theory of so called Grassmannian angles of polytopes, which is a purely geometrical concept. This approach has connections to, but is still relatively far away from, the one in this work. In particular, it is by no means straight-forward, if at all possible, to generalize it other problems than (cid:96)1-minimization. The condition presented in this paper is highly related to several other geometric stability measures: so called restricted singular values [2] of matrices on cones, as also Renegar's condition number [4, 18] as well as the Grassmannian condition number [1] (see Section 2.1). In fact, already in the previously mentioned paper [9], it is proven that if the smallest singular value of A restricted to the descent cone of the functional f does not vanish, we will have stability. During the final review of this paper, the author was made aware of the fact that also the connection between Renegar's Condition number and robustness of compressed sensing has recently been investigated in [19]. This work provides a new, more elementary, perspective to the above mentioned notions, and in particular establishes its relations to classical criteria for stability in compressed sensing. Another contribution of this paper is the observation that if f is a norm, the criterion also implies stability, an observation which, to the best of the knowledge of the author, has not been done before. 3 Notation Throughout the whole paper, H will denote a general, finite-dimensional Hilbert space. The corresponding inner product will be denoted (cid:104)·,·(cid:105), and the induced norm by (cid:107)·(cid:107). For a subspace U ⊆ H, we will write Note that due to the finite-dimensionality of H, this infimum is in fact attained. (cid:107)x + U(cid:107) = inf u∈U (cid:107)x + u(cid:107) . In the final part of the paper, we will deal with Gaussian vectors and Gaussian matrices. A random vector, or linear map, is said to be Gaussian if its representation in an orthonormal basis, or in a pair of such, has i.i.d. standard normally distributed entries. The entries of a vector x in Rd will be denoted x(1), x(2), . . . , x(d). Rd is equipped with the standard scalar product whose induced norm is the (cid:96)2-norm: d(cid:88) (cid:104)x, y(cid:105) = x(i)y(i), (cid:107)x(cid:107)2 = x(i)2. (cid:118)(cid:117)(cid:117)(cid:116) d(cid:88) 2 A Geometrical Robustness Condition i=1 i=1 Let us begin by considering the most classical compressed sensing setting: that is, the problem of retrieving a signal x0 ∈ Rd with few non-zero entries from exact measurements b = Ax0 using (cid:96)1-minimization (i.e. (P1)). One of the most well known criteria for (P1) to be successful is the so called Null Space Property, N SP . A matrix A satisfies the N SP with respect to the index set S0 ⊆ {1, . . . d} if for every η ∈ ker A, we have η(i) < η(i) . (2) (cid:88) i∈S0 (cid:88) i /∈S0 The N SP with respect to S is in fact equivalent to (P1) recovering a signal x0 with support S [12]. The N SP has a geometrical meaning. In order to explain it, let us first define descent cones of a function f . Definition 1. Let H be a finite-dimensional Hilbert space, and f : H → R. The descent cone D(f, x0) of f at the point x0 ∈ H is the cone generated by the descent directions of f at x0, i.e. D(f, x0) = {η ∃τ > 0 : f (x0 + τ η) ≤ f (x0)} . Example 2. The descent cone of the (cid:96)1-norm at a vector x0 supported on the set S0 is given by Proof: Since the conditions for η to belong to both the left hand and the right hand set, respectively, is invariant under scaling, we may assume that η has small norm. Then we have D((cid:107)·(cid:107)1 , x0) = (cid:111) sgn(x0(i))ηi . i /∈S0 (cid:110) ηi ≤ −(cid:88) η (cid:88) (cid:40)x0(i) + sgn(x0(i))ηi i∈S0 i ∈ S0 i /∈ S0 , x0(i) + η(i) = which implies (cid:107)x0 + η(cid:107)1 − (cid:107)x0(cid:107) =(cid:80) when(cid:80) ηi ≤ −(cid:80) i /∈S0 sgn(x0(i))ηi i /∈S0 i∈S0 ηi +(cid:80) η(i) i∈S0 sgn(x0(i))ηi. This is smaller than or equal to zero exactly With the last example in mind, it is not hard to convince oneself that Equation (2) exactly states that the vector η does not lie in the descent cone of any signal supported on the set S0. I.e., the N SP actually reads ∀x0 : supp x0 = S0 : D((cid:107)·(cid:107)1 , x0) ∩ ker A = {0} . This observation can be generalized to more general situations, as the following well-known lemma shows. 4 Figure 1: The impact of an angular separation of D(f, x0) and ker A. Note that locally, the sets x0 + D(f, x0) and {xf (x) ≤ f (x0)} have the same structure. [9, Proposition 2.1].) Let f : H → R be convex, A : H → Rm be linear and consider the Lemma 3. (E.g. program (Pf ), with b = Ax0 for noiseless recovery of the signal x0. The solution of (Pf ) is equal to x0 if and only if D(f, x0) ∩ ker A = {0} . (3) f )? The main difference between the program (Pf ) and (P  Can we use the previous lemma to develop a geometrical intuition of what we have to assume in order to prove stability and robustness of the recovery using (P  f ) is that the former is only allowed to search for a solution in the set x0 +ker A, while the latter can search in a tubular neighborhood of the same set. Figure 1 suggests that if the descent cone D(f, x0) and ker A do not only trivially intersect each other, but also have an angular separation, it should be possible to prove that the intersection of the mentioned tubular neighborhood and the set {xf (x) ≤ f (x0)} is not large. This should in turn imply robustness. (In fact, this intuition was used already in [8] when proving robustness in the original compressed sensing setting.) To provide a precise formulation of angular separation, we first define the θ-expansion of a cone. Definition 4. Let H be a finite-dimensional Hilbert space with norm (cid:107)·(cid:107) and scalar product (cid:104)·,·(cid:105). Let further C ⊆ H be a convex cone, i.e. a convex set with τ C = C for every τ > 0. Then for θ ∈ [0, π], we define the θ-expansion C∧θ as the set C∧θ =(cid:8)x ∈ H(cid:12)(cid:12) ∃y ∈ C : (cid:104)x, y(cid:105) ≥ cos(θ)(cid:107)x(cid:107)(cid:107)y(cid:107)(cid:9) . For an illustration of the relation between a cone C and its θ-expansion C∧θ, see Figure 2. Before moving on, let us make some remarks. (i) C = C∧0 ⊆ C∧θ for every θ > 0. Remark 5. (ii) If C is closed, C∧θ can alternatively defined as the set C∧θ = cone({x ∈ H(cid:12)(cid:12) (cid:107)x(cid:107) = 1, (cid:104)x, y(cid:105) ≥ cos(θ)}). sup y∈C,(cid:107)y(cid:107)=1 (iii) C∧θ is always a cone, but not always convex. As a concrete counterexample, consider the closed, convex cone K ⊆ R3 K = cone(cid:0)(cid:8)x ∈ R3 (cid:12)(cid:12) x(1) = 1, x(2) = 0,x(3) ≤ 1(cid:9)(cid:1) . 5 Figure 2: A convex cone C and its θ-expansion C∧θ. Using the previous remark, we can calculate K∧θ exactly. We have for x ∈ R3 sup y∈K,(cid:107)x(cid:107)=1 (cid:104)x, y(cid:105) = sup t∈[−1,1] x(1) + tx(3) √ 1 + t2 = (cid:40)(cid:112)x(1)2 + x(3)2 x(1)+x(3) √ 2 if x(3) ≤ x(1) else, where the last equality can be proven using elementary calculus. The last remark now tells us that K∧ π given by 6 is (cid:40) y ∈ R3 (cid:12)(cid:12) √ (cid:113) 2 (cid:16) 3(cid:107)y(cid:107)2 2 (cid:17) ≤ (cid:40)(cid:112)y(1)2 + y(3)2 (cid:113) 2 (cid:17) y(1)+y(3) (cid:16) √ 2 if y(3) ≤ y(1) else. (cid:41) (cid:113) 2 (cid:17) 3 ,−1 . (cid:16) This set is not convex; for instance, the points y1 = 1, 3 , 1 and y2 = 1, are both contained in the set, whereas 1 2 y1 + 1 2 y2 = 1, 3 , 0 is not. See also Figure 3. Remark 6. It is not hard to convince oneself that δ(x, y) = arccos((cid:104)x, y(cid:105)) defines a metric on Sd−1. In particular, the triangle inequality holds: arccos((cid:104)x, y(cid:105)) ≤ arccos((cid:104)x, z(cid:105)) + arccos((cid:104)z, y(cid:105)). After these preparations, we may state our robustness condition. Proposition 7. Let x0 ∈ H and A : H → Rm be linear. Consider the program (P  and b = Ax0 + n with (cid:107)n(cid:107) ≤ . If there exists a θ > 0 such that f ), where f : H → R is convex D∧θ(f, x0) ∩ ker A = {0} , (4) then there exists a constant κ > 0 so that the solution x of (P  (cid:107)x − x0(cid:107) ≤ κ. f ) obeys κ depends on µ and the smallest non-zero singular value σmin(A) of A. 6 Figure 3: The convex cone K and its non-convex π 6 -extension K∧ π 6 For simplicity, we will call (4) the θ-angular separation condition, or θ-ASC. We will in fact not prove this proposition directly. Instead, we first establish a connection to so called restricted singular values of the matrix A, which then implies Proposition 7 as a corollary. of a linear map A : H → Rm restricted to the cones C ⊆ H and D ⊆ Rm was defined as by The concept of restricted singular values was extensively studied in the paper [2]. There, the singular value σC→D(A) = min x∈C,(cid:107)x(cid:107)=1 (cid:107)ΠDAx(cid:107)2 , where ΠD denotes the Euclidean (metric, orthogonal ) projection (or nearest point map) of the convex set D: ΠD(x) = argminy∈D (cid:107)x − y(cid:107)2 . (See also, for instance, [3].) If D = Rm, one also speaks of the minimal gain [9]. Interesting for us is that if σD(f,x0)→Rm(A) > 0, (P  f ) will be robust. Lemma 8. (See [2], [9, Proposition 2.2].) Let f : H → R be convex and A : H → Rm be linear with m < dimH. If σD(f,x0)→Rm (A) > 0, any solution x∗ of (P  f ) obeys (cid:107)x∗ − x0(cid:107)H ≤ 2 σD(f,x0)→Rm . Remark 9. (i) Note that in the references we cited, the lemma is proved under slightly less general conditions (e.g., the function f is assumed to be a norm). The proof does however work, line for line, also in our case, and we therefore omit it, for the sake of brevity. (ii) As is indicated by the formulation of Lemma 8, the solution vector x∗ by no means has to be unique, even for arbitrarily small  > 0. We can construct an example of such a situation as follows: Consider a sparse vector x0 ∈ Rd supported on the set S and a matrix A ∈ Rm,d such that σD((cid:107)·(cid:107)1,x0)→Rm > 0 - i.e. in particular that x0 is recovered from its exact measurements by P1. Suppose that the solution x∗ of P  situation). For any i ∈ S(cid:48)\S, consider the matrix (cid:101)A ∈ Rm,d+1 formed by concatenating A with a copy of the 1 is unique for some  > 0 and is supported on a set S(cid:48) which is larger than S (this is arguably the most common It is then clear that x0 still is recovered from the exact measurements b = (cid:101)Ax0 via the exact program P1, i:th column of A, and the vectors x∗ and x0 formed by concatenating x∗ and x0 with a zero, respectively. and hence (see Section 3.1) σD((cid:107)·(cid:107),x0)→Rm ( A) > 0, but that any vector for θ ∈ [0, 1] solves P  1. Hence, the solution of the relaxed problem P  θ = θx∗ + (1 − θ)ed+1 x∗ 1 for (cid:101)A is not unique. 7 We now prove that the ASC to some extent is equivalent to σD(f,x0)→Rm(A) > 0. Lemma 10. Let C ⊆ H be a non-empty convex cone and A : H → Rd be a linear map. Then the following are equivalent (1) There exists a θ > 0 such that C∧θ ∩ ker A = {0}. (2) σC→Rm(A) > 0. In particular, if σmin / max(A) denotes the smallest/largest non-vanishing singular value of A, respectively, we have for every θ with C∧θ ∩ ker A = {0} that (5) Proof. (1) ⇒ (2). Suppose that C∧θ ∩ ker A = {0} and let x ∈ C have unit norm. Then we have for every y ∈ ker A sin(θ)σmin(A) ≤ σC→Rm (A) ≤ sin (θ) σmax(A). (cid:107)x − y(cid:107)2 2 = 1 + (cid:107)y(cid:107)2 − 2(cid:104)x, y(cid:105) ≥ 1 + (cid:107)y(cid:107)2 − 2 cos(θ)(cid:107)y(cid:107) = 1 − cos2(θ) + ((cid:107)y(cid:107) − cos(θ))2 ≥ sin2(θ), since due to C∧θ ∩ ker A = {0}, there must be (cid:104)x, y(cid:105) ≤ cos(θ)(cid:107)y(cid:107). Since y ∈ ker A was arbitrary, we obtain (cid:107)x + ker A(cid:107) ≥ sin(θ). This has the consequence (cid:107)Ax(cid:107) ≥ σmin(A)(cid:107)x + ker A(cid:107) ≥ σmin(A) sin(θ). It follows that σC→Rm(A) ≥ σmin(A) sin(θ) > 0, since σmin(A) > 0 due to ker A (cid:54)= H, which follows from C∧θ ∩ ker A = {0}, and θ > 0. (2) ⇒ (1) Suppose that σC→Rm(A) > 0 and let x ∈ C and y ∈ ker A have unit norm. Define θ through (cid:104)x, y(cid:105) = cos(θ). Our goal is to prove that θ has to be larger than some number θ0 > 0. Since x ∈ C, we have (cid:107)Ax(cid:107)2 ≥ σC→Rm(A). We also have for every t ∈ R, due to y ∈ ker A, σC→Rm (A) ≤ (cid:107)Ax(cid:107)2 = (cid:107)A(x − ty)(cid:107)2 ≤ σmax(A)(cid:107)x − ty(cid:107)2 = σmax(A)(cid:112)1 + t2 − 2t cos(θ). Choosing t = cos(θ) yields sin (θ) =(cid:112)1 − cos2(θ) ≥ σC→Rm(A) > 0, σmax(A) where we in the last step used σC→Rm(A) > 0. This proves the claim. Now Proposition 7 easily follows from combining Lemma 8 with Lemma 10. We will return to the connection between restricted singular values and the ASC in the next section. Remark 11. The ASC is not necessary for robust recovery. Consider for example (cid:96)2-minimization in H = Rd: (P  2) min(cid:107)x(cid:107)2 subject to (cid:107)Ax − b(cid:107) ≤ . given by(cid:8)v ∈ Rd (cid:104)x0, v(cid:105) < 0(cid:9). We claim that this implies that for each θ > 0, D∧θ((cid:107)·(cid:107)2 , x0) ∩ ker A (cid:54)= {0}. In order for (cid:96)2-minimization to exactly recover a signal x0 (which of course is necessary for robust recovery, choose  = 0), we need to have x0 ⊥ ker A, since the solution of (P2) is given by Πker A⊥ x0. Furthermore, D((cid:107)·(cid:107)2 , x0) is To see why, let v ∈ D((cid:107)·(cid:107)2 , x0) and η be a nonzero element of ker A. Such an element necessarily exists as soon as d > m. For any λ > 0, v + λη ∈ D((cid:107)·(cid:107)2 , x0). This since we have (cid:104)x0, η(cid:105) = 0 due to x0 ⊥ ker A (cid:51) η. Consequently, (cid:104)v + λη, x0(cid:105) = (cid:104)v, x0(cid:105) < 0, i.e. v + λη ∈ D((cid:107)·(cid:107)2 , x0). Now consider the quotient (cid:104)v + λη, η(cid:105) (cid:107)v + λη(cid:107)2 (cid:107)η(cid:107)2 ≥ (cid:104)v, η(cid:105) + λ(cid:107)η(cid:107)2 ((cid:107)v(cid:107)2 + λ(cid:107)η(cid:107)2)(cid:107)η(cid:107)2 2 . 8 By letting λ → ∞, this quotient can be made arbitrarily close to 1. Since v + λη ∈ D((cid:107)·(cid:107)2 , x0), this means that ker A (cid:51) η ∈ D∧θ((cid:107)·(cid:107)2 , x0) for every θ > 0. Hence, we have a non-trivial intersection between the θ-expansion of the descent cone and the kernel for every θ > 0. 2) necessarily lies in ker A⊥ (any part in ker A can be removed without affecting (cid:107)Ax − b(cid:107)2, and at the same time making (cid:107)x(cid:107)2 smaller). We already argued that x0 also has this property. This has the immediate consequence that It is, however, not hard to convince oneself that the solution x of (P  (cid:107)x − x0(cid:107)2 ≤ σmin(A)−1 (cid:107)Ax − Ax0(cid:107)2 ≤ 2σmin(A)−1, i.e. we have robustness. Now we prove that under the assumption that f is a norm on H, the ASC in fact also implies stability. Theorem 12. Let (cid:107)·(cid:107)∗ be a norm on H and consider the convex program min(cid:107)x(cid:107)∗ subject to (cid:107)Ax − b(cid:107) ≤ , (6) where b = Ax0 + n with (cid:107)n(cid:107) ≤ . Suppose that there exists a θ > 0 so that A fulfills the θ-ASC for every x0 in some subset C ⊆ H. Then the following is true for (6): There exist constants κ1 and κ2 such that any solution x∗ fulfills (cid:107)x∗ − x0(cid:107) ≤ κ1 + κ2 dist(cid:107)·(cid:107)∗ (x0, cone(C)). (7) Here, cone(C) denotes the cone generated by C, i.e., the set {λx0, λ > 0 and x0 ∈ C}. The first constant κ1 depends on σmin(A), (cid:107)·(cid:107)∗ and θ. The second constant κ2 depends on θ, (cid:107)·(cid:107)∗ and the condition number σmax(A)/σmin(A) of A. Proof. (See Fig. 4 for a graphical depiction of the proof.) Let x0 be a, not necessarily unique, vector in cone(C) with (cid:107)x0 − x0(cid:107)∗ = δ := dist(cid:107)·(cid:107)∗ (x0, cone(C)). Due to the homogenity of (cid:107)·(cid:107)∗, we have D((cid:107)·(cid:107)∗ , x0) = D((cid:107)·(cid:107)∗ , λx0) for every λ > 0. Therefore we may without loss of generality scale the problem such that (cid:107)x0(cid:107)∗ = 1. Also, since all norms on the finite dimensional space H are equivalent, there exists γ > 0 so that for each x ∈ H, γ−1 (cid:107)x(cid:107)∗ ≤ (cid:107)x(cid:107) ≤ γ (cid:107)x(cid:107)∗. These two facts have the consequence that (cid:107)Ax0 − Ax0(cid:107) ≤ σmax(A)(cid:107)x0 − x0(cid:107) ≤ γσmax(A)δ. Since (cid:107)Ax∗ − Ax0(cid:107) ≤ (cid:107)Ax∗ − b(cid:107) + (cid:107)b − Ax0(cid:107) ≤ 2, this implies that (cid:107)Ax∗ − Ax0(cid:107) ≤ (cid:107)Ax∗ − Ax0(cid:107) + (cid:107)Ax0 − Ax0(cid:107) ≤ 2 + γσmax(A)δ, i.e. (cid:107)x0 − x + ker A(cid:107) ≤ σmin(A)−1(2 + γσmax(A)δ). Let h be the vector in ker A so that (cid:107)x0 + h − x(cid:107) = (cid:107)x0 − x + ker A(cid:107). Now since x∗ is a solution of (6), there must be (cid:107)x∗(cid:107)∗ ≤ (cid:107)x0(cid:107)∗ ≤ (cid:107)x0(cid:107)∗ + δ = (cid:107)(1 + δ)x0(cid:107)∗, i.e. h := x∗ − (1 + δ)x0 ∈ D((cid:107)·(cid:107)∗ , (1 + δ)x0) = D((cid:107)·(cid:107)∗ , x0), where the latter is due to the homogeneity of (cid:107)·(cid:107)∗. Now we have (cid:107)x∗ − x0(cid:107) ≤ (cid:107)x∗ − (1 + δ)x0(cid:107) + (cid:107)δx0(cid:107) ≤ (cid:107)h(cid:107) + γδ. Due to h ∈ ker A and h ∈ D((cid:107)·(cid:107) , x0), (4) implies that (cid:104)h, h(cid:105) ≤ cos(θ)(cid:107)h(cid:107)(cid:107)h(cid:107). This has the consequence that (cid:107)h − h(cid:107)2 ≥ (cid:107)h(cid:107)2 + (cid:107)h(cid:107)2 − 2 cos(θ)(cid:107)h(cid:107)(cid:107)h(cid:107) = ((cid:107)h(cid:107) − cos(θ)(cid:107)h(cid:107))2 + (1 − cos2(θ))(cid:107)h(cid:107)2 ≥ sin2(θ)(cid:107)h(cid:107)2, which implies that (cid:107)h(cid:107)2 ≤ sin−1(θ)(cid:107)h − h(cid:107). Since (cid:107)h − h(cid:107) = (cid:107)h − x∗ + (1 + δ)x0(cid:107) ≤ (cid:107)x0 + h − x∗(cid:107) + δ (cid:107)x0(cid:107) ≤ σmin(A)−1(2 + γσmax(A)δ) + γδ, we have (cid:107)h(cid:107) ≤ (sin(θ)σmin(A)) −1 (2 + γσmax(A)δ) + sin−1(θ)γδ. Finally, we estimate (cid:107)x∗ − x0(cid:107)∗ ≤ (cid:107)x∗ − x0(cid:107)∗ + (cid:107)x0 − x0(cid:107)∗ ≤ γ (cid:107)x∗ − x0(cid:107) + δ ≤ γ((cid:107)h(cid:107) + γδ) + δ ≤ sin−1(θ)(cid:0)γσmin(A)−1(2 + γσmax(A)δ) + δ(cid:1) + 2δ =: κ1 + κ2 dist(cid:107)·(cid:107)(x0, cone(C)), where κ1 = 2γ(sin(θ)σmin(A))−1 and κ2 = sin θ−1(γ2σmax(A)/σmin(A) + 1) + 2, which is what we wanted to prove. 9 Figure 4: The proof of Proposition 12. 2.1 ASC and Two Other Geometrical Notions of Stability. Let us end this section by briefly discussing the connection between the ASC-condition and two other mea- sures for stability of a linear embedding, namely the so-called Renegar's condition number [4, 18] CR(A) and the Grassmannian condition number [1] C(A) of a matrix. They were originally introduced to study the stability of the homogeneous convex feasibility problem: Given a closed convex cone K ⊆ Rm with non-empty interior not containing a subspace (a regular cone), for which A ∈ Rm,d does there exist a z with polar C∗ =(cid:8)x ∈ Rd ∀y ∈ C : (cid:104)x, y(cid:105) ≤ 0(cid:9) is regular, it is well known that if the range of the transposed matrix Let us call matrices such feasible. The connection to our problem follows from duality: given a cone C whose A∗ intersects the int C∗ , the kernel of a matrix A can't intersect C non-trivially (see for instance [4]). Az ∈ int K? the following sets of subspaces : Since the range of A∗ always is equal to the orthogonal complement ker A⊥ of ker A, it makes sense to define Pm(K) =(cid:8)U ∈ G(d, m) U⊥ ∩ K∗ (cid:54)= {0}(cid:9) , Dm(C) = {U ∈ G(d, m) U ∩ K (cid:54)= {0}} G(d, m) denotes the Grassmannian manifold of m-dimensional subspaces of Rd. It can be proven that both Pm(K) and Dm(K) are closed in G(d, m) and that they share a common boundary Σm(K). The Grassmannian condition number of a matrix A with respect to a regular cone C is defined as the inverse of the distance of ker A to Σm(C), i.e. C(A) = 1 dist(ker A, Σm(C)) . The distance is thereby calculated with respect to the canonical metric on G(d, m): if U and V are m-dimensional subspaces of Rd and ΠU and ΠV are the orthogonal projections onto them, we define d(U, V ) = (cid:107)ΠU − ΠV (cid:107) . Here, (cid:107)·(cid:107) denotes the operator norm. C(A) is very closely related to the ASC-condition: if A is feasible, the largest angle θ∗ so that A satisfies the θ-ASC satisfies [1, Proposition 1.6] sin(θ∗) = 1 C(A) . 10 (8) The Grassmannian condition number is itself closely related to the so-called Renegar's condition number CR(A). Given a feasible matrix A (in the same sense as above), it is defined as the distance from A to the set of infeasible matrices, i.e. CR(A) = min{(cid:107)∆A(cid:107) A + ∆A is infeasible} . In our setting, it can be proven that [19, Lemma 2.2] CR(A) = σmax(A) σC→Rm (A) . (9) The article [19] contains some more details and interesting results concerning the connection between CR(A) and the robustness properties of compressed sensing problems. In particular, they prove the following version of Lemma 8 of this article: if we assume that the noise level is below (cid:107)A(cid:107)  (which in particular is interesting when the measurement matrix A only is known up to some error ∆A), every solution x∗ of the program P  f obeys (cid:107)x∗ − x0(cid:107)2 ≤ 2CR(A). Let us end this section by noting that one can prove Lemma 10 by using the following inequality from [1, Theorem 1.4] : C(A) ≤ CR(A) ≤ σmax(A) σmin(A) C(A). Using (8) and (9), this can be rewritten as sin(θ∗)σmin(A) ≤ σC→Rm(A) ≤ sin(θ∗)σmax(A), which is exactly the inequality (5). 3 When is the ASC satisfied? Having established that the ASC implies stability and robustness for signal recovery using the convex program (6), it is of course interesting to ask for which matrices this condition is satisfied. In this section, we will first prove the maybe somewhat remarkable fact that, for many reasonable norms, the weak N SP -like condition (3) in fact implies that the ASC is satisfied for some θ > 0. However, the above reasoning only yields the existence of a θ > 0 with (4), and does not give any control of the size of θ. Therefore, we will also briefly discuss the relation between already known stability conditions for compressed sensing, and that the ASC can be secured with high probability using random Gaussian matrices. Using the concept of Gaussian widths, we will argue that if one needs m0 measurements to secure that (Pf ) recovers a signal x0 with high probability from noiseless measurements, we need m0 + O(cid:0)sin(cid:0) θ (cid:1) d(cid:1) to secure the 2 ASC for θ > 0. 3.1 Exact Recovery Implies Some Stability and Some Robustness. The first result of this subsection was essentially already proven in [2]. Let us state it, and for completeness also give a proof, and then discuss its implications and limitations. Theorem 13. Let C ⊆ H be a closed convex cone, and A : H → Rm be linear. Then if C ∩ ker A = {0}, there exists a θ > 0 such that C∧θ ∩ ker A = {0}, i.e., the θ-ASC holds. 11 Proof. Under the assumption that C and D are closed, the restricted singular value σC→D(A) vanishes if and only if either AC ∩ D∗ (cid:54)= {0} or C ∩ ker A (cid:54)= {0} [2, Proposition 2.2]. Since D = Rm in our case, D∗ = {0}, and we hence by contraposition have the equivalence σC→Rm(A) > 0 ⇔ C ∩ ker A = 0. Since by Lemma 10, σC→Rm(A) > 0 is equivalent to the existence of a θ > 0 such that C∧θ ∩ ker A = {0}, the claim is proven. On a theoretical level, the last proposition implies that as soon as the recovery of some class of signals C from exact measurements with the help of a convex program is guaranteed, we also have stability and robustness for the recovery of signals close to C from noisy measurements. As simple and beautiful the result is, it has its flaws. In particular, we have no control whatsoever over the size of the parameter θ, which in turn implies that we have no control over the constants in (7). In the case that the norm (cid:107)·(cid:107)∗ has a unit ball which is a polytope, we can do a bit better. Although we still cannot provide any general bound on the size of θ, we can prove that it will have the same size for all points lying in the same face of the unit ball. Before stating and proving the result, let us note that the assumption that the unit ball of (cid:107)·(cid:107)∗ is a polytope is not far-fetched. In particular, it is true for both (cid:96)1-minimization (and its many variants, i.e., also for weighted norms etc.), and for (cid:96)∞-minimization -- or in general any atomic norm generated by a finite set of atoms A. Let us now formulate the main part of the argument in the following lemma. Lemma 14. Let P ⊆ H be a closed polytope and C ⊆ P be a union of faces of P . Suppose that the linear subspace U ⊆ H has the property that for each x0 ∈ C, x0 + U intersects P only in x0. Then for each x0 ∈ C, there exists a µ < 1 such that ∀x ∈ P, z ∈ U : (cid:104)x − x0, z(cid:105) ≤ µ(cid:107)x − x0(cid:107)2 (cid:107)z(cid:107)2 . The size of µ is only dependent on which face x0 lies in. Although the proof of this lemma is elementary, it is relatively long. Therefore, we postpone it to Appendix A. Instead, we use it to prove the aforementioned result about stability and robustness for recovery using convex programs involving norms with polytope unit balls. Corollary 15. Let A : H → Rm be given. Suppose that (cid:107)·(cid:107)∗ is a norm whose unit ball is a polytope, and C be a union of faces of that polytope. If the program (P(cid:107)·(cid:107)∗ ) recovers x0 from the noiseless measurements Ax0 for every x0 ∈ C, all signals x0 close to the cone generated by C will be stably and robustly recovered by (P (cid:107)·(cid:107)∗ ) in the sense of (7). The constants κ1 and κ2 will only depend on which face the normalized version x0/(cid:107)x0(cid:107)∗ of x0 lies closest to. Proof. Since each x0 in C is recovered exactly by (P(cid:107)·(cid:107)∗ ) by noiseless measurements, we will by Lemma 3 have D((cid:107)·(cid:107)∗ , x0) ∩ ker A = {0} for each x0 ∈ C. Since the descent cone of (cid:107)·(cid:107)∗ at x0 is generated by the vectors x − x0, where x ∈ P = {y(cid:107)y(cid:107)∗ ≤ 1}, the conditions of Lemma 14 are satisfied. Said lemma therefore implies that D∧θ((cid:107)·(cid:107)∗ , x0) ∩ ker A = {0} for x0 ∈ C, where θ > 0 only depends on which face x0 lies in. This together with Theorem 12 implies the claim. 3.2 ASC Compared to Classical Stability and Robustness Conditions in Com- pressed Sensing. In the following, we will relate the ASC to two well-known criteria for stability and robustness of (cid:96)1-minimization from the literature: the RIP and the RN SP . We will begin by considering the RIP . It is a well-known fact ) will recover any s-sparse vector in a that if the restricted isometry constant δ2s is small, the program (P (cid:107)·(cid:107)1 robust and stable manner. E.g. in [12, Theorem 6.12], it is proved that if δ2s < 4/ 12 √ 41, (7) will be satisfied for some constants κ1, κ2 only dependent on δ2s. Having this in mind, it is of course interesting to ask oneself if it is possible to directly prove that a small δ2s will imply the ASC for some θ > 0. The next proposition gives a positive answer to that question, and it furthermore provides the control of the size of θ > 0 we lacked in the previous section. Proposition 16. Suppose that δs and δ2s of the matrix A satisfies (cid:32) (cid:114) √ (cid:33) d s Then D∧θ((cid:107)·(cid:107)1 , x0) ∩ ker A = {0} for every s-sparse x0. 1 − δs δs + 1 5 + 1 δ2s 1 + δs ≤ cos(θ). For the proof of this claim, we need two lemmata. The first one, we cite from the book [12]. Lemma 17. (See [12, Proposition 6.3]). Let u and v be s-sparse vectors with disjoint support. Then we have (cid:104)Au, Av(cid:105) ≤ δ2s (cid:107)u(cid:107)2 (cid:107)v(cid:107)2 , where δ2s is the (2s)-th restricted isometry constant of A. The next one is about the structure of vectors in the descent cone of the (cid:96)1-norm at a sparse vector. Lemma 18. Let x0 ∈ Rd be supported on the set S0 with S0 = s, and u ∈ D((cid:107)·(cid:107)1 , x0). Let furthermore (Si)n be a partition of the set {1, . . . , d}\S0 with the properties i=1 u(i) ≥ max i∈Sk+1 u(i) and Sk ≤ s (10) (cid:107)uSi(cid:107)2 < √ 5(cid:107)u(cid:107)2 . n(cid:88) (cid:13)(cid:13)2 ≤ 1√ i=0 min i∈Sk for every k = 1, . . . n − 1. Then we have Proof. First, (10) implies that(cid:13)(cid:13)uSk+1 following estimate due to (cid:107)v(cid:107)2 ≤ (cid:107)v(cid:107)1 for v ∈ Rd: (cid:107)uSk(cid:107)2 ≤ (cid:107)uS0(cid:107)2 + (cid:107)uS1(cid:107)2 + 1√ s Now, since u ∈ D((cid:107)·(cid:107)1 , x0), we have (see Example 2) k=0 n(cid:88) (cid:107)1 ≤ −(cid:88) i∈S0 (cid:107)uSc 0 s (cid:107)uSk(cid:107)1 for k = 1, . . . n [12, Lemma 6.10]. Therefore, we have the n(cid:88) k=2 (cid:13)(cid:13)uSk−1 (cid:13)(cid:13)1 ≤ (cid:107)uS0(cid:107)2 + (cid:107)uSc 0 (cid:107)2 + 1√ s (cid:107)uSc 0 (cid:107)1 sgn(x0(i))ui ≤ (cid:107)uS0(cid:107)1 ≤ √ s(cid:107)uS0(cid:107)2 , where the last inequality is due to the fact that uS0 is supported on S0. This implies (cid:107)uS0(cid:107)2 + (cid:107)uSc 0 (cid:107)2 + 1√ s (cid:107)uSc 0 (cid:107)1 ≤ 2(cid:107)uS0(cid:107)2 + (cid:107)uSc 0 (cid:107)2 = (2, 1) · ((cid:107)uS0(cid:107)2 ,(cid:107)uSc 0 (cid:107)2) (cid:113)(cid:107)uS0(cid:107)2 √ 5 ≤ √ 2 + (cid:107)uSc 0 (cid:107)2 2 = 5(cid:107)u(cid:107)2 , where we used the Cauchy-Schwarz inequality in the third step. With these two lemmas, we may prove Proposition 16. 13 Proof of Proposition 16. Let u ∈ D((cid:107)·(cid:107)1 , x0) and v ∈ ker A, where without loss of generality (cid:107)u(cid:107)2 = (cid:107)v(cid:107)2 = 1. Our goal is to prove that necessarily (cid:104)u, v(cid:105) < cos(θ). Let us begin by partitioning the set {1, . . . , d}\S0 into n sets S1, . . . Sn so that Si ∩ Sj = ∅ for i (cid:54)= j, Si = s for i = 1, . . . n − 1, Sn ≤ s, and the monotonicity criterion (10) is met for u. Then n ≤ d s , and we have (cid:33) (cid:107)uSi + vSi(cid:107)2 2 − (cid:107)uSi − vSi(cid:107)2 2 n(cid:88) i=0 (cid:104)u, v(cid:105) = (cid:104)uSi, vSi(cid:105) = 1 4 ≤ 1 4 (cid:32) n(cid:88) (cid:32) n(cid:88) (cid:32) n(cid:88) i=0 i=0 i=0 1 1 − δs 2δs 1 − δ2 s (cid:107)AuSi + AvSi(cid:107)2 2 − 1 1 + δs (cid:107)AuSi − AvSi(cid:107)2 = 1 4 ((cid:107)AuSi(cid:107)2 2 + (cid:107)AvSi(cid:107)2 2) + 4 1 − δ2 s (cid:33) (cid:33) (cid:104)AuSi, AvSi(cid:105) 2 , where we in the second to last step used that uSi ± vSi for each i is s-sparse, and the definition of δs. Now since v ∈ ker A, we have AvSi = −(cid:80)n (cid:104)AuSi, AvSi(cid:105) = − n(cid:88) k=1,k(cid:54)=i AvSk , and consequently, (cid:104)AuSi, AvSk(cid:105) ≤ n(cid:88) δ2s (cid:107)uSi(cid:107)2 (cid:107)vSk(cid:107)2 ≤ n(cid:88) k=1 k(cid:54)=i k=1 k(cid:54)=i k=1 δ2s (cid:107)uSi(cid:107)2 (cid:107)vSk(cid:107)2 , where we in the second to last step applied Lemma 17. Again using the definition of δs, we conclude (cid:107)Aui(cid:107)2 (1 + δs)(cid:107)ui(cid:107)2 2. Combining all of the previous estimates, we obtain 2 ≤ (1 + δs)(cid:107)vi(cid:107)2 2 and (cid:107)Avi(cid:107)2 2 ≤ (cid:104)u, v(cid:105) ≤ n(cid:88) Now we use (cid:107)u(cid:107) ,(cid:107)v(cid:107) = 1, Lemma 18, the inequality(cid:80)n 2(1 − δs) i=0 ((cid:107)uSi(cid:107)2 2 + (cid:107)vSi(cid:107)2 δs (cid:104)u, v(cid:105) < δs 1 − δs √ 5 + which is what we wanted to prove. s 2) + δ2s 1 − δ2 i=0 xi ≤ √ (cid:114) δ2s 1 − δ2 s d s (cid:33)(cid:32) n(cid:88) i=0 (cid:32) n(cid:88) n + 1(cid:112)(cid:80)n i=0 (cid:107)uSi(cid:107)2 (cid:33) (cid:107)vSi(cid:107)2 i=0 x2 i and n ≤ d s to conclude + 1 ≤ cos(θ), Now let us turn to another criterion for robust and stable recovery using (cid:96)1-minimization: the Robust Null- Space Property or RN SP . Let 0 ≤ γ < 1 and τ > 0. A matrix A ∈ Rm,d satisfies the (γ, τ )-RN SP with respect to the index set T ⊆ {1, 2, . . . , d} if for every x ∈ Rd, we have (cid:107)xT(cid:107)1 ≤ γ (cid:107)xT c(cid:107)1 + τ (cid:107)Ax(cid:107)2 . (11) In fact, it turns out that the RN SP with respect to an index set T is equivalent to that the ASC is fulfilled for any x0 supported on T , in the sense specified by the following theorem. Theorem 19. Let A ∈ Rm,d. (1) If the (γ, τ )-RN SP is satisfied, then the ASC-condition is satisfied uniformly for all x0 supported on T , i.e. there exists a θ > 0 with for every x0 supported on T . D∧θ((cid:107)·(cid:107)1 , x0) ∩ ker A = ∅ (12) (2) If there exists a θ > 0 such that (12) holds for every x0 supported on T , there exists 0 ≤ γ < 1 and τ > 0 such that the (γ, τ )-RN SP holds with respect to T . 14 Proof. (1). Due to Lemma 10, it suffices to prove that σD((cid:107)·(cid:107)1,x0)(A) > σ0 > 0 for every x0 supported on T . To this end, note that the RN SP implies that (cid:107)Ax(cid:107)2 ≥ 1 τ x∈D((cid:107)·(cid:107)1,x0) min (cid:107)x(cid:107)2=1 x∈D((cid:107)·(cid:107)1,x0) min (cid:107)x(cid:107)2=1 ((cid:107)xT(cid:107)1 − γ(cid:107)xT c(cid:107)1) ≥ 1 − γ ≥ 1 − γ τ τ x∈D((cid:107)·(cid:107)1,x0) min (cid:107)x(cid:107)2=1 x∈D((cid:107)·(cid:107)1,x0) min (cid:107)x(cid:107)2=1 (cid:107)xT(cid:107)1 (cid:107)x(cid:107)1 ≥ (1 − γ) . 1 2 2τ We used that since x ∈ D((cid:107)·(cid:107)1 , x0), (cid:107)xT(cid:107)1 ≥ (cid:107)xT c(cid:107)1. This in particular implies that (cid:107)xT(cid:107)1 ≥ 1 has been proven. (2). Suppose that there exists a θ with (12) for every x0 supported on T . Let us begin by arguing that this implies that there exists a 0 ≤ γ < 1 with the following property: If (cid:107)xT(cid:107)1 ≥ γ(cid:107)xT c(cid:107)1, there exists an x0 supported on T with x ∈ D∧ θ Consider an x with (cid:107)xT(cid:107)1 ≥ γ(cid:107)xT c(cid:107)1 (where the value of γ is yet to be determined) and consider x0 := −xT . Then we have x ∈ D((cid:107)·(cid:107)1 , x0), where 2 (cid:107)x(cid:107)1 The claim 2 ((cid:107)·(cid:107)1 , x0). (cid:40) x(i) = x(i) γx(i) i ∈ T i ∈ T c, (cid:88) x(i) ≤(cid:88) x(i) = −(cid:88) since i∈T since (cid:107)xT(cid:107)1 ≥ γ(cid:107)xT c(cid:107)1 (see also Example 2). Now we have i∈T c sgn x0(i)x(i), i∈T x(i) = γ (cid:88) i∈T x(i)2 + γ(cid:80) i∈T c (cid:107)x(cid:107)2 (cid:107)x(cid:107)2 (cid:80) (cid:104)x, x(cid:105) (cid:107)x(cid:107)2 (cid:107)x(cid:107)2 = i∈T c x(i)2 = (cid:107)xT(cid:107)2 + γ(1 − (cid:107)xT(cid:107)2) (cid:113)(cid:107)xT(cid:107)2 + γ2(1 − (cid:107)xT(cid:107)2) a + γ(1 − a) (cid:112)a + γ2(1 − a) ≥ inf 0≤a≤1 where we in the second to last step without loss of generality assumed that (cid:107)x(cid:107)2 = 1. Since 0 ≤ a, γ ≤ 1, we have (cid:112)a + γ2(1 − a) ≤(cid:112)a + γ(1 − a), and consequently ≥(cid:112)a + γ(1 − a). Hence, we have √ a+γ(1−a) a+γ2(1−a) (cid:112)a + γ(1 − a) = √ γ ≥ cos (cid:18) θ (cid:19) , 2 a + γ(1 − a) (cid:112)a + γ2(1 − a) inf 0≤a≤1 ≥ inf 0≤a≤1 2 ((cid:107)·(cid:107)1 , x0). if we choose γ close enough to 1. This proves that x ∈ D∧ θ Now we claim that there exist a τ > 0 such that the (γ, τ )-RN SP with respect to T is satisfied. To see this, first note that (11) is trivial for x with (cid:107)xT(cid:107)1 ≤ γ(cid:107)xT c(cid:107)1. For the other x, which we without loss of generality may assumed to be (cid:96)2-normalized, we know by the above argument that x ∈ D∧ θ 2 ((cid:107)·(cid:107)1 , x0) for some x0 supported on T . This implies that x /∈ (ker A)∧ θ 2 , there would exist unit norm vectors y ∈ ker A and z ∈ D((cid:107)·(cid:107)1 , x0) such that arccos(cid:104)x, y(cid:105) ≤ θ 2 and arccos(cid:104)x, z(cid:105) ≤ θ 2 . Remark 6 would then imply that arccos(cid:104)y, z(cid:105) ≤ θ, i.e., (cid:104)x, y(cid:105) ≥ cos(θ), which is a contradiction to (12). (cid:18) y ∈ ker A arbitrary, we have (cid:1), which is seen as in the proof of Theorem 12; for (cid:18) θ 2 implies that (cid:107)x + ker A(cid:107)2 ≥ sin(cid:0) θ 2 , since if indeed x ∈ (ker A)∧ θ But, x /∈ (ker A)∧ θ (cid:19)(cid:19)2 (cid:18) θ (cid:18) θ (cid:18) θ (cid:19) (cid:19) (cid:19) 2 (cid:107)x − y(cid:107)2 2 ≥ 1 + (cid:107)y(cid:107)2 2 − 2 cos (cid:107)y(cid:107) = (cid:107)y(cid:107)2 − cos + 1 − cos2 ≥ sin2 2 . 2 2 2 15 Therefore, we have for normalized x with (cid:107)xT(cid:107)1 ≥ γ(cid:107)xT c(cid:107)1 (cid:107)Ax(cid:107)2 ≥ σmin(A)(cid:107)x + ker A(cid:107)2 ≥ σmin(A) sin Therefore, we can choose τ =(cid:112)T(σmin(A) sin θ)−1, since then (cid:18) θ T (cid:107)x(cid:107)2 ≤(cid:112)T (cid:107)xT(cid:107)1 ≤ σmin(A) sin (cid:18) √ (cid:19)(cid:19)−1 (cid:107)Ax(cid:107)2 ≤ γ(cid:107)xT c(cid:107)1 + τ (cid:107)Ax(cid:107)2 , 2 (cid:18) θ (cid:19) . 2 i.e. (11) is satisfied for any x ∈ Rd. 3.3 How Many Gaussian Measurements Are Needed to Secure the ASC for a Given θ? In this final section we will assume that the linear map A : H → Rm is Gaussian, and ask how large m has to be such that θ-ASC for some given θ is satisfied with high probability. Typically in compressed sensing, one can prove that as long as the parameter m is larger than a certain threshold, depending on the dimension of H and the type of structured signals, a program of the form (P  f ) will exhibit stability and robustness. One way of proving such results is to use Gordon's "Escape through a mesh"- lemma [15]. This lemma relates the so called Gaussian width w(T ) to the dimension that a uniformly distributed random subspace has to have in order to miss a set T contained in the unit sphere S(H) of H (which clearly is equivalent to missing cone(T )). Let us make this idea more precise. If g is a Gaussian vector in H, it is defined by (cid:18) (cid:19) . w(T ) = E (cid:104)x, g(cid:105) sup x∈T There are several ways to state the "Escape through a mesh"-lemma. The following one from [12, Th. 9.21] let (cid:96)m denote the expected length of a Gaussian vector in Rm, i.e. is probably the most convenient for us: (cid:96)m = E ((cid:107)g(cid:107)2) for g ∈ Rm Gaussian, and T a subset of S(H). Then we have This particularly implies that if (cid:96)m > w(T ) + 2(cid:112)η−1, inf x∈T P (cid:107)Ax(cid:107)2 ≤ (cid:96)m − w(T ) − t (cid:18) (cid:19) (cid:19) ≤ exp (cid:18)−t2 (cid:112) (cid:107)Ax(cid:107)2 ≥ (cid:96)m − w(T ) − 2 2 η−1 (cid:18) inf x∈T P (ker A ∩ T = ∅) > P . (cid:19) > 1 − η. (13) (14) Since (cid:96)m ≈ √ m, the estimate (14) qualitatively tells us that if m > w(T )2, ker A will probably miss the set cone(T ). If we choose T such that cone(T ) = D(f, x0) ∩ S(H), we can relate this to exact recovery through the program Pf . (13) suggests that it is possible to directly relate the Gaussian width of the cone C to the threshold amount of measurements needed to establish σC→Rm(A) > 0, or equivalently the ASC for some θ > 0, with high probability. This was discussed in detail in [2, 9], and we refer to those articles for more information. We can however also compare the Gaussian width of C∧θ ∩ S(H) to the one of C ∩ S(H) for a convex cone C. This is also interesting in its own right, since Gaussian widths in general are hard to estimate. We have the following result. Proposition 20. Let C ⊆ H be a convex cone, and θ ∈ [0, π where (cid:96)H := E ((cid:107)g(cid:107)) is the expected length of a Gaussian vector in H. In particular, w(C ∩ S(H)) ≤ w(cid:0)C∧θ ∩ S(H)(cid:1) ≤ cos(θ)w(C ∩ S(H)) + sin(θ)(cid:96)H w(C ∩ S(H))2 ≤ w(cid:0)C∧θ ∩ S(H)(cid:1)2 ≤ w(C ∩ S(H))2 + O (sin(θ) dimH) . 2 ]. Then we have 16 Proof. Since C ⊆ C∧θ, we trivially have w(C ∩ S(H)) ≤ w(C∧θ ∩ S(H)). To prove the upper bound, notice that if a unit vector x lies in C∧θ, there exists a unit vector y ∈ C with (cid:104)x, y(cid:105) ≥ cos(θ), or (see Remark 6) δ(x, y) ≤ θ. The triangle inequality for the sphere metric δ now implies for every g ∈ H Using the monotonicity properties of cos, this implies δ y, g(cid:107)g(cid:107) =≤ δ x, g(cid:107)g(cid:107) + δ(x, y) ≤ δ x, g(cid:107)g(cid:107) (cid:17) (cid:16) (cid:16) (cid:16) (cid:17) (cid:17)(cid:17) ≤ cos (cid:16) pos (cid:16) (cid:16) (cid:16) x, = cos δ x, g(cid:107)g(cid:107) δ y, g(cid:107)g(cid:107) (cid:28) where pos(t) = max(t, 0) denotes the positive part of a real number t. Using the cosine addition formula, this yields + θ. (cid:17) (cid:17) − θ (cid:16) if δ (cid:17)(cid:17) , (cid:17) (cid:16) g (cid:107)g(cid:107) (cid:29) 1 (cid:68) (cid:68) x, g(cid:107)g(cid:107) (cid:69) ≤ y, g(cid:107)g(cid:107) cos(θ) + sin ≤ cos(θ) y, g(cid:107)g(cid:107) (cid:69) (cid:16) (cid:68) (cid:16) (cid:68) (cid:16) (cid:16) (cid:17)(cid:17) y, g(cid:107)g(cid:107) < θ δ else sin(θ) y, g(cid:107)g(cid:107) + sin(θ), (cid:69) (cid:17)(cid:17) ≤ 1, sin(θ) ∈ [0, 1], cos(θ) ≥ 0 (since θ ∈ [0, π (cid:69) + sin(θ) ≥ cos2(θ) + sin2(θ) = 1 where the last step follows from sin δ y, g(cid:107)g(cid:107) (cid:16) for y with δ cos(θ) y, g(cid:107)g(cid:107) y, g(cid:107)g(cid:107) (cid:17) w(cid:0)C∧θ ∩ S(H)(cid:1) = E (cid:32) < θ. Consequently sup x∈C∧θ,(cid:107)x(cid:107)=1 (cid:33) (cid:104)x, g(cid:105) (cid:32) sup cos(θ)(cid:104)y, g(cid:105) + (cid:107)g(cid:107) sin(θ) ≤ E = cos(θ)w(C ∩ S(H)) + sin(θ)(cid:96)H. y∈C,(cid:107)y(cid:107)=1 2 ]) and (cid:33) For the second inequality, we simply need to square the first one: w(cid:0)C∧θ ∩ S(H)(cid:1)2 ≤ cos2(θ)w(C ∩ S(H))2 + 2 cos(θ) sin(θ)w(C ∩ S(H)(cid:96)H + sin2(θ)(cid:96)2H √ ≤ w(C ∩ S(H))2 + (2 cos(θ) + sin(θ)) sin(θ)(cid:96)2H ≤ w(C ∩ S(H))2 + 5 sin(θ)(cid:96)2H, √ where we utilized that w(C ∩ S(H)) ≤ E ((cid:107)g(cid:107)) = (cid:96)H = O( dimH). The ASC states that if D∧θ(f, x0)∩ ker A = {0}, the program (P  f ) will stably recover x0. According to above discussion, this will be the case provided m is larger than w(D∧θ(f, x0) ∩ S(H))2. The last proposition therefore shows that in order to achieve the ASC for a given θ, we need O(sin (θ) dimH) more Gaussian measurements than we would need to ensure (3). Let us end by considering an example which shows that this claim in general cannot be substantially improved. For this, we will use an alternative tool to calculate thresholds as described above: the statistical dimension δ(C) of a convex cone C. It was introduced in [3], where the authors also proved that if m > δ(C), the probability that ker A ∩ C = {0} is very high. δ(C) is always close to w(C ∩ S(H)) -- in fact, we have [3, Prop 10. 2] w(C)2 ≤ δ(C) ≤ w(C)2 + 1. Example 21. Consider the circular cone Circd(α) ⊆ Rd, Circd(α) =(cid:8)x ∈ Rdx(1) ≥ cos(α)(cid:107)x(cid:107)2 (cid:9) 17 According to [3, Proposition 3.4], the statistical dimension of Circd(α) is equal to d sin2(α)+O(1). It is furthermore clear that Circ∧θ d (α) = Circd(α + θ) (in particular again a convex cone). Hence, Circ∧θ d (α) δ = d sin2(α + θ) + O(1) = d(sin2(α) cos2(θ) + 1 2 sin(2α) sin(2θ) + cos2(α) sin2(θ)) + O(1). (cid:16) (cid:17) If we assume that θ is small, we have cos2(θ) ≈ 1, sin2(θ) ≈ 0 and sin(2θ) ≈ 4 sin (θ). Hence, we obtain for such θ (cid:16) Circ∧θ d (α) (cid:17) ≈ δ(Circd(α)) + O (d sin (θ)) + O(1). δ Acknowledgement The author acknowledges support by Deutsche Forschungsgemeinschaft (DFG) Grant KU 1446/18 - 1, as well as Grant SPP 1798, as well as the Deutscher Akademischer Austausch Dienst (DAAD), which supported him with a scholarship for his master studies, during which the research was completed. He also likes to thank Gitta Kutyniok for the supervision of the Master Thesis leading up to this article, Martin Genzel for careful proofreading and many useful suggestions for increasing the readability of the paper and Dae Gwan Lee for pointing out a few errors, and suggesting ways to improve some statements in the paper. He also wishes to thank the anonymous reviewers, whose comments, criticisms and suggestions greatly improved the final version of this article, in particular regarding Sections 2.1 and 3.3. References [1] D. Amelunxen and P. Burgisser. A coordinate-free condition number for convex programming. SIAM J. Optim, 22(3):1029 -- 1041, 2012. [2] D. Amelunxen and M. Lotz. Gordon's inequality and condition numbers in conic optimization. Preprint. arXiv:1408.3016, 2014. [3] D. Amelunxen, M. Lotz, M. B. McCoy, and J. A. Tropp. Living on the edge: phase transitions in convex programs with random data. Information and Inference, 3:224 -- 294, 2014. [4] A. Belloni and R. M. Freund. A geometric analysis of renegar's condition number and its interplay with conic curvature. Math. Program., 119:95 -- 107, 2007. [5] J. Cahill, X. Chen, and R. Wang. The gap between the null space property and the restricted isometry property. Preprint. arXiv:1506.03040, 2015. [6] J. Cahill and D. G. Mixon. Robust width: A characterization of uniformly stable and robust compressed sensing. Preprint. arXiv:1408.4409, 2014. [7] E. J. Candes and Y. Plan. Matrix completion with noise. P. IEEE, 98(6):925 -- 936, 2010. [8] E. J. Cand`es, J. K. Romberg, and T. Tao. Stable signal recovery from incomplete and inaccurate measure- ments. Comm. Pure Appl. Math., 59(8):1207 -- 1223, 2006. [9] V. Chandrasekaran, B. Recht, P. A. Parrilo, and A. S. Willsky. The convex geometry of linear inverse problems. Found. of Comput. Math., 12(6):805 -- 849, 2012. [10] M. Davenport, D. Needell, M. B. Wakin, et al. Signal space CoSaMP for sparse recovery with redundant dictionaries. IEEE Trans. Inf. Theory, 59(10):6820 -- 6829, 2013. [11] A. Flinth. Discrete compressed sensing and geometry. Master's thesis, Technische Universitat Berlin, 2015. 18 [12] S. Foucart and H. Rauhut. A mathematical introduction to Compressed Sensing. Birkhauser, 2013. [13] M. Friedlander, H. Mansour, R. Saab, and O. Yilmaz. Recovering compressively sampled signals using partial support information. IEEE Trans. Inf. Theory, 58(2):1122 -- 1134, 2012. [14] J.-J. Fuchs. Spread representations. In Asilomar Conference on Signals, Systems, and Computers, 2011. [15] Y. Gordon. On Milman's inequality and random subspaces which escape through a mesh in Rn. In J. Lin- denstrauss and V. D. Milman, editors, Geometric Aspects of Functional Analysis, volume 1317 of Lecture Notes in Mathematics, pages 84 -- 106. Springer Berlin Heidelberg, 1988. [16] M. A. Khajehnejad, W. Xu, A. S. Avestimehr, and B. Hassibi. Weighted (cid:96)1 minimization for sparse signals with prior information. In IEEE Inter. Symp. Inf. Theory, 2009., pages 483 -- 487, 2009. [17] S. Oymak, M. Khajehnejad, and B. Hassibi. Recovery threshold for optimal weight (cid:96)1-minimization. In IEEE Inter. Symp. Inf. Theory, 2012., pages 2032 -- 2036, 2012. [18] J. Renegar. Some perturbation theory for linear programming. Math. Program., 65(1):73 -- 91, 1994. [19] V. Roulet, N. Boumal, and A. D'Aspremont. Renegar's condition number and compressed sensing perfor- mance. Preprint. arXiv:1506.03295, 2015. [20] M. Stojnic, F. Parvaresh, and B. Hassibi. On the reconstruction of block-sparse signals with an optimal number of measurements. IEEE Trans. Signal. Process., 57(8):3075 -- 3085, 2009. [21] W. Xu and B. Hassibi. Precise stability phase transitions for minimization: A unified geometric framework. IEEE Trans. Inf. Theory, 57(10):6894 -- 6919, 2011. A Proof of Lemma 14 Here, as promised, we present the proof of Lemma 14. Proof of Lemma 14. If we define for x0 ∈ C µ(x0) = sup x∈P\{x−x0} sup z∈U,(cid:107)z(cid:107)=1 (cid:28) x − x0 (cid:107)x − x0(cid:107) , z (cid:29) , we need to prove that µ(x0) < 1 for each x0 ∈ C, and further more that µ(x1) = µ(x2) if x1 and x2 lie in the relative interior of the same face F . To do this, let us begin by remarking that for each x ∈ P and z ∈ U , (cid:107)z(cid:107) = 1, there must be < 1 (15) Otherwise, the equality statement of the Cauchy-Schwarz inequality would imply that x − x0 = λz for some λ ∈ R, which means that x ∈ x0 + U , which is ruled out by assumption. (15) implies that we only need to prove that the supremum defining µ(x0) is attained in some points x ∈ P\{x0} and z ∈ U ∩ SH, where we introduced the notation SH = {z ∈ H (cid:107)z(cid:107) = 1}. Towards this end, let (xn) and (zn) be such that (cid:28) x − x0 (cid:107)x − x0(cid:107) , z (cid:29) (cid:28) x − x0 (cid:107)x − x0(cid:107) , z (cid:29) → µ(x0). (16) Since H is finite-dimensional, and P and U ∩ SH are closed, there will exist x ∈ P and z ∈ U ∩ SH such that, possibly after going over to subsequences, xn → x, zn → z. It is only left to prove that we can choose xn in such a way so that x (cid:54)= x0. 19 k(cid:92) H(ai, αi) ∩ (cid:96)(cid:92) i=1 j=1 F = H−(bj, βj), say, with P = (cid:84)k i=1 H−(ai, αi) ∩(cid:84)(cid:96) So suppose xn → x0. By the assumption of the theorem, x0 is a member of some face F of P , which by definition is defined by a set of linear equalities and inequalities. If we write H−(c, γ) = {x ∈ H : (cid:104)c, x(cid:105) ≤ γ} and H(c, γ) = {x ∈ H : (cid:104)c, x(cid:105) = γ}, we have j=1 H−(bj, βj). Hereby, we may without loss of generality assume that all vectors (ai) and (bj) are normalized, and furthermore that (cid:104)bj, x0(cid:105) < βj for all j (if not, we may rename some of the bj to ai's). If we define xn = x0 + λnvn for some vn = xn − x0 and some λn > 0, it is not hard to convince oneself that also xn is a sequence that satisfies (16). If we can prove that λn can be chosen in such a manner that (cid:107)xn − x0(cid:107) ≥ δ > 0, but still xn ∈ P for every n, we are done. Due to the definition of P , we know that xn ∈ P , if and only if (cid:104)ai, x0 + λnvn(cid:105) ≤ αi and (cid:104)bj, x0 + λvn(cid:105) ≤ βj for each i and j. Since (cid:104)ai, x0(cid:105) = αi and (cid:104)ai, xn(cid:105) ≤ αi, the conditions involving ai's will always be fulfilled. As for the ones involving bj's, notice that (cid:104)bj, λn(xn − x0)(cid:105) ≤ βj − (cid:104)bj, x0(cid:105) for each j. We assumed that βj − (cid:104)bj, x0(cid:105) > 0 for each j. Therefore, if (cid:104)bj, vn(cid:105) ≤ 0 for all j, (cid:104)bj, x0 + λnvn(cid:105) ≤ βj for each possible λn, in particular for λn so large that (cid:107)xn − x0(cid:107)2 ≥ . Otherwise we choose λn in such a way so that for some j0, (cid:104)bj0, λnvn(cid:105) = βj0 − (cid:104)bj0 , x0(cid:105). This has the consequence that (cid:107)λn(xn − x0)(cid:107) ≥ (cid:104)bj0 , λn(xn − x0)(cid:105) = βj0 − (cid:104)bj0, x0(cid:105) ≥ min (βj − (cid:104)bj, x0(cid:105)) =: δ > 0. For this choice of λn, we have xn ∈ P and (cid:107)xn − x0(cid:107) ≥ δ > 0 for all n, which we set out to prove. j Now we will prove that µ(x0) only depends on in which face F of P x0 lies. Let us assume that x0 and x0 both lie in the relative interior of the face F , which is described in the same way as above, i.e. that (cid:104)ai, x0(cid:105) = (cid:104)ai, x0(cid:105) = αi (cid:104)bj, x0(cid:105) < βj,(cid:104)bj, x0(cid:105) < βj. Now let x and z be such that (cid:104)(x − x0)/(cid:107)x − x0(cid:107)2 , z(cid:105) = µ(x0). It is clear that (cid:104)ai, x(cid:105) ≤ αi and (cid:104)bj, x(cid:105) ≤ βj for each i and j, since x ∈ P . We may in fact assume that (cid:104)bj, x(cid:105) < βj for each j: If there exists a j with (cid:104)bj, x(cid:105) = βj, we may, by considering the segment between x and x0 (on which the function x → (cid:104)(x − x0)/(cid:107)x − x0(cid:107) , z(cid:105) is constant), find a vector x with (cid:28) x − x0 (cid:29) (cid:28) x − x0 (cid:29) µ(x0) = (cid:107)x − x0(cid:107) , z = (cid:107)x − x0(cid:107) , z , but (cid:104)bj, x(cid:105) < βj, since (cid:104)bj, x0(cid:105) < βj. since (cid:104)bj, x(cid:105) < βj. Hence, x + (x0 − x0) ∈ P and therefore Now, (cid:104)ai, x + (x0 − x0)(cid:105) = (cid:104)ai, x(cid:105) ≤ αi. Furthermore, if (cid:107)x0 − x0(cid:107) is small, we will have (cid:104)bj, x + (x0 − x0)(cid:105) < βj, (cid:28) x + (x0 − x0) − x0 (cid:29) (cid:28) x − x0 (cid:29) (cid:107)x + (x0 − x0) − x0(cid:107) , z = (cid:107)x − x0(cid:107) , z = µ(x0). µ(x0) ≥ A symmetric argument proves that µ(x0) ≥ µ(x0). Hence, around each x0 ∈ relint F , there exists an open ball Ux0 so that µ is constant in F ∩ Ux0 . Since relint F is a connected set, this proves that µ is constant over the whole of relint F . 20
1501.01580
2
1501
2015-08-12T15:14:37
A relaxation result in BVxL^p for integral functionals depending on elastic strain and chemical composition
[ "math.FA" ]
An integral representation result is obtained for the relaxation of a class of energy functionals depending on two vector fields with different behaviors which appear in the context of thermochemical equilibria and are related to image decomposition models and directors theory in nonlinear elasticity.
math.FA
math
A relaxation result in BV × Lp for integral functionals depending on chemical composition and elastic strain. Grac¸a Carita∗, Elvira Zappale† July 10, 2018 Abstract An integral representation result is obtained for the relaxation of a class of energy functionals de- pending on two vector fields with different behaviors which appear in the context of thermochemical equilibria and are related to image decomposition models and directors theory in nonlinear elasticity. Keywords: relaxation, convexity-quasiconvexity. MSC2000 classification: 49J45, 74Q05 1 Introduction In this paper we consider energies depending on two vector fields with different behaviors: u ∈ W 1,1 (Ω; Rn) and v ∈ Lp (Ω; Rm) , Ω being a bounded open subset of RN . Let 1 < p ≤ ∞ and for every (u, v) ∈ W 1,1(Ω; Rn) × Lp(Ω; Rm) define the functional J (u, v) :=ZΩ f (v, ∇u) dx (1.1) where f : Rm × Rn×N → [0, ∞) is a continuous function. Minimization of energies depending on two independent vector fields have been introduced to model several phenomena. For instance the case of thermochemical equilibria among multiphase multicomponent solids and Cosserat theories in the context of elasticity: we refer to [9, 8] and the references therein for a detailed explanation about this kind of applications. In the Sobolev setting, after the pioneer works [8, 9], relaxation with a Carath´eodory density f ≡ f (x, u, ∇u, v) , and homogenization for density of the type f(cid:0) x respectively. ε , ∇u, v(cid:1) have been considered in [5] and [4], In the present paper we are interested in studying the lower semicontinuity and relaxation of (1.1) with respect to the L1−strong×Lp−weak convergence. Clearly, bounded sequences {uh} ⊂ W 1,1(Ω; Rn) may converge in L1, up to a subsequence, to a BV function. In the BV -setting this question has been already addressed in [7], only when the density f is convex- quasiconvex (see (2.2)) and the vector field v ∈ L∞(Ω; Rm). Here we allow v to be in Lp(Ω; Rm), p > 1 and f is not necessarily convex-quasiconvex. We provide an argument alternative to the one in [7], devoted to clarify some points in the lower semicontinuity result therein. We also emphasize that under specific restrictions on the density f , i.e. f (x, u, v, ∇u) ≡ W (x, u, ∇u) + ϕ(x, u, v), such analysis was considered already in [10] in order to describe image decomposition models. In [11] a general f was taken into account when the target u is in W 1,1(Ω; Rn). In this manuscript we consider f ≡ f (v, ∇u) and u ∈ BV (Ω; Rn). ∗CIMA-UE, Departamento de Matem´atica, Universidade de ´Evora, Rua Romao Ramalho, 59 7000 671 ´Evora, Portugal e-mail: [email protected] †D.I.In., Universita' degli Studi di Salerno, Via Giovanni Paolo II 132, 84084 Fisciano (SA) Italy e-mail:[email protected] 1 We study separately the cases 1 < p < ∞ and p = ∞. To this end, we introduce for 1 < p < ∞ the functional J p (u, v) := inf(cid:26)lim inf h→∞ J (uh, vh) : uh ∈ W 1,1 (Ω; Rn) , vh ∈ Lp (Ω; Rm) , uh → u in L1, vh ⇀ v in Lp(cid:27) , (1.2) for any pair (u, v) ∈ BV (Ω; Rn) × Lp(Ω; Rm) and, for p = ∞ the functional J ∞(u, v) := inf(cid:26)lim inf h→∞ J (uh, vh) : uh ∈ W 1,1 (Ω; Rn) , vh ∈ L∞ (Ω; Rm) , uh → u in L1, vh ∗ ⇀ v in L∞(cid:27) , (1.3) for any pair (u, v) ∈ BV (Ω; Rn) × L∞ (Ω; Rm) . Since bounded sequences {uh} in W 1,1(Ω; Rn) converge in L1 to a BV function u and bounded sequences {vh} in Lp(Ω; Rm) if 1 < p < ∞, (in L∞(Ω; Rm) if p = ∞) weakly converge to a function v ∈ Lp(Ω; Rm), (weakly ∗ in L∞), the relaxed functionals J p and J ∞ will be composed by an absolutely continuous part and a singular one with respect to the Lebesgue measure (see (2.12)). On the other hand, as already emphasized in [7], it is crucial to observe that v, regarded as a measure, is absolutely continuous with respect to the Lebesgue one, besides it is not defined on the singular sets of u. Namely in those sets where the singular part with respect the Lebesgue measure of the distributional gradient of u, Dsu, is concentrated. Thus specific features of the density f will come into play to ensure a proper integral representation. The integral representation of (1.2) will be achieved in Theorem 1.1 under the following hypotheses: (H1)p There exists C > 0 such that 1 C (bp + ξ) − C ≤ f (b, ξ) ≤ C (1 + bp + ξ) , for (b, ξ) ∈ Rm × Rn×N . (H2)p There exists C′ > 0, L > 0, 0 < τ ≤ 1 such that t > 0, ξ ∈ Rn×N , with t ξ > L =⇒(cid:12)(cid:12)(cid:12)(cid:12) f (b, tξ) t − f ∞ (b, ξ)(cid:12)(cid:12)(cid:12)(cid:12) where f ∞ is the recession function of f defined for every b ∈ Rm as ≤ C′ bp + 1 t + ξ1−τ tτ ! , f ∞(b, ξ) := lim sup t→∞ f (b, tξ) t . (1.4) In order to characterize the functional J ∞ introduced in (1.3) we will replace assumptions (H1)p and (H2)p by the following ones: (H1)∞ Given M > 0, there exists CM > 0 such that, if v ≤ M then 1 CM ξ − CM ≤ f (b, ξ) ≤ CM (1 + ξ) , for every ξ ∈ Rn×N . (H2)∞ Given M > 0, there exist C′ M > 0, L > 0, 0 < τ ≤ 1 such that b ≤ M, t > 0, ξ ∈ Rn×N , with t ξ > L =⇒(cid:12)(cid:12)(cid:12)(cid:12) f (b, tξ) t − f ∞ (b, ξ)(cid:12)(cid:12)(cid:12)(cid:12) ξ1−τ ′ ≤ C M . tτ Section 2 is devoted to notations, preliminaries about measure theory and some properties of the energy densities. In particular, we stress that a series of results is presented in order to show all the properties and relations among the relaxed energy densities involved in the integral representation and that can be of further use for the interested readers since they often appear in the integral representation context. Section 3 contains the arguments necessary to prove the main results stated below. 2 Theorem 1.1 Let J be given by (1.1), with f satisfying (H1)p and (H2)p and let J p be given by (1.2) then J p(u, v) =ZΩ CQf (v, ∇u) dx +ZΩ (CQf )∞(cid:18)0, dDsu dDsu(cid:19) dDsu, for every (u, v) ∈ BV (Ω; Rn) × Lp(Ω; Rm). We denote by CQf the convex-quasiconvex envelope of f in (2.5) and (CQf )∞ represents the recession function of CQf , defined according to (1.4), which coincides, under suitable assumptions, (see assumptions (2.6), (2.7), Proposition 2.12 and Remark 2.13), with the convex-quasiconvex envelope of f ∞, CQ(f ∞), and this allows us to remove the parenthesis. For the case p = ∞ we have the following. Theorem 1.2 Let J be given by (1.1), with f satisfying (H1)∞ and (H2)∞ and let J ∞ be given by (1.3) then J ∞(u, v) =ZΩ CQf (v, ∇u) dx +ZΩ (CQf )∞(cid:18)0, dDsu dDsu(cid:19) dDsu, for every (u, v) ∈ BV (Ω; Rn) × L∞(Ω; Rm). For the case 1 < p < ∞, the proof of the lower bound is presented in Theorem 3.1 while the upper bound is in Theorem 3.2, both under the extra hypothesis (H0) f is convex-quasiconvex. The case p = ∞ is discussed in subsection 3.2. Furthermore, we observe that Proposition 2.14 in subsection 2.3 is devoted to remove the convexity-quasiconvexity assumption on f . 2 Notations preliminaries and properties of the energy densities In this section, we start by establishing notations, recalling some preliminary results on measure theory that will be useful through the paper and finally we recall the space of functions of bounded variation. Then we deduce the main properties of convex-quasiconvex functions, recession functions and related envelopes. If ν ∈ SN −1 and {ν, ν2, . . . , νN } is an orthonormal basis of RN , Qν denotes the unit cube centered at the origin with its faces either parallel or orthogonal to ν, ν2, . . . , νN . If x ∈ RN and ρ > 0, we set Q(x, ρ) := x + ρ Q and Qν(x, ρ) := x + ρ Qν, Q is the cube (cid:0)− 1 Let Ω be a generic open subset of RN , we denote by M(Ω) the space of all signed Radon measures in Ω with bounded total variation. By the Riesz Representation Theorem, M(Ω) can be identified to the dual of the separable space C0(Ω) of continuous functions on Ω vanishing on the boundary ∂Ω. The N -dimensional Lebesgue measure in RN is designated as LN . 2 , 1 . 2(cid:1)N If µ ∈ M(Ω) and λ ∈ M(Ω) is a nonnegative Radon measure, we denote by dµ dλ the Radon-Nikod´ym derivative of µ with respect to λ. By a generalization of the Besicovich Differentiation Theorem (see [1, Proposition 2.2]), it can be proved that there exists a Borel set E ⊂ Ω such that λ(E) = 0 and dµ dλ (x) = lim ρ→0+ µ(x + ρ C) λ(x + ρ C) for all x ∈ Supp λ \ E and any open bounded convex set C containing the origin. We recall that the exceptional set E above does not depend on C. An immediate corollary is the generalization of Lebesgue-Besicovitch Differentiation Theorem given below. Theorem 2.1 If µ is a nonnegative Radon measure and if f ∈ L1 loc(RN , µ) then lim ε→0+ 1 µ(x + εC)Zx+εC f (y) − f (x)dµ(y) = 0 for µ- a.e. x ∈ RN and for every, bounded, convex, open set C containing the origin. 3 Definition 2.2 A function u ∈ L1(Ω; Rn) is said to be of bounded variation, and we write u ∈ BV (Ω; Rn), if all its first distributional derivatives, Djui, belong to M(Ω) for 1 ≤ i ≤ n and 1 ≤ j ≤ N . The matrix-valued measure whose entries are Djui is denoted by Du and Du stands for its total variation. We observe that if u ∈ BV (Ω; Rn) then u 7→ Du(Ω) is lower semicontinuous in BV (Ω; Rn) with respect to the L1 loc(Ω; Rn) topology. By the Lebesgue Decomposition Theorem we can split Du into the sum of two mutually singular measures Dau and Dsu, where Dau is the absolutely continuous part and Dsu is the singular part of Du with respect to the Lebesgue measure LN . By ∇u we denote the Radon-Nikod´ym derivative of Dau with respect to the Lebesgue measure so that we can write Proposition 2.3 If u ∈ BV (Ω; Rn) then for LN −a.e. x0 ∈ Ω Du = ∇uLN + Dsu. lim ε→0+ 1 ε ( 1 εN ZQ(x0,ε) u(x) − u(x0) − ∇u (x0) · (x − x0) N −1 N N N −1 dx) = 0. (2.1) For more details regarding functions of bounded variation we refer to [2]. 2.1 Convex-quasiconvex functions We start by recalling the notion of convex-quasiconvex function, presented in [7] (see also [8] and [9]). Definition 2.4 A Borel measurable function f : Rm × Rn×N → R is said to be convex-quasiconvex if, for every (b, ξ) ∈ Rm × Rn×N , there exists a bounded open set D of RN such that f (b, ξ) ≤ 1 DZD f (b + η(x), ξ + ∇ϕ(x)) dx, (2.2) for every η ∈ L∞(D; Rm), with RD η(x) dx = 0, and for every ϕ ∈ W 1,∞ Remark 2.5 (D; Rn). 0 i) It can be easily seen that, if f is convex-quasiconvex then condition (2.2) is true for any bounded open set D ⊂ RN . ii) A convex-quasiconvex function is separately convex. iii) The growth condition from above in (H1)p, ii) and [4, Proposition 2.11], entail that there exists γ > 0 such that f (b, ξ) − f (b′, ξ′) ≤ γ(cid:16)ξ − ξ′ +(cid:16)1 + bp−1 + b′p−1 + ξ 1 p′ + ξ′ for every b, b′ ∈ Rm, ξ, ξ′ ∈ Rn×N , where p > 1 and p′ its conjugate exponent. 1 p′(cid:17) b − b′(cid:17) (2.3) iv) In case of growth conditions expressed by (H1)∞ (see [11, Proposition 4]), ii) entails that, given M > 0 there exists a constant β (M, n, m, N ) such that f (b, ξ) − f (b′, ξ′) ≤ β (1 + ξ + ξ′) b − b′ + β ξ − ξ′ (2.4) for every b, b′ ∈ Rm, such that b ≤ M and b′ ≤ M, for every ξ, ξ′ ∈ Rn×N . We introduce the notion of convex-quasiconvex envelope of a function, which is crucial to deal with the relaxation procedure. Definition 2.6 Let f : Rm × Rn×N → R be a Borel measurable function bounded from below. The convex- quasiconvex envelope is the largest convex-quasiconvex function below f, i.e., CQf (b, ξ) := sup {g (b, ξ) : g ≤ f, g convex-quasiconvex} . 4 By Theorem 4.16 in [9], the convex-quasiconvex envelope coincides with the so called convex-quasiconvexification CQf (b, ξ) = inf(cid:26) 1 DZD f (b + η(x), ξ + ∇ϕ(x)) dx : η ∈ L∞ (D; Rm), ZD (2.5) η(x)dx = 0, ϕ ∈ W 1,∞ 0 (D; Rn)o . As for convexity-quasiconvexity, condition (2.5) can be stated for any bounded open set D ⊂ RN . It can also be showed that if f satisfies a growth condition of type (H1)p then in (2.2) and (2.5) the spaces L∞ and W 1,∞ can be replaced by Lp and W 1,1 , respectively. 0 0 The following proposition, that will be exploited in the sequel, can be found in [11, Proposition 5]. The proof is omitted since it is very similar to [10, Proposition 2.1]. Proposition 2.7 Let f : Rm × Rn×N → [0, ∞) be a continuous function satisfying (H1)p. Then CQf is continuous and satisfies (H1)p . Consequently, CQf satisfies (2.3). In order to deal with v ∈ L∞(Ω; Rm) and to compare with the result in BV × Lp, 1 < p < ∞, one can consider a different setting of assumptions on the energy density f . Namely, following [11, Proposition 6 and Remark 7], if α : [0, ∞) → [0, ∞) is a convex and increasing function, such that α(0) = 0 and if f : Rm × Rn×N → [0, ∞) is a continuous function satisfying 1 C (α(b) + ξ) − C ≤ f (b, ξ) ≤ C(1 + α(b) + ξ) (2.6) for every (b, ξ) ∈ Rm × Rn×N , then CQf satisfies a condition analogous to (2.6). Moreover, CQf is a continuous function. Analogously, one can assume that f satisfies the following variant of (H2)∞: there exist c′ > 0, L > 0, 0 < τ ≤ 1 such that t > 0, ξ ∈ Rn×N , with t ξ > L =⇒(cid:12)(cid:12)(cid:12)(cid:12) f (b, tξ) t − f ∞ (b, ξ)(cid:12)(cid:12)(cid:12)(cid:12) ≤ c′ α(b) + 1 t + ξ1−τ tτ ! . (2.7) We observe that, if from one hand (2.6) and (2.7) generalize (H1)p and (H2)p respectively, from the other hand they can be regarded also as a stronger version of (H1)∞ and (H2)∞, respectively. 2.2 The recession function Let f : Rm × Rn×N → [0, ∞[, and let f ∞ : Rm × Rn×N → [0, ∞[ be its recession function, defined in (1.4). The following properties are an easy consequence of the definition of recession function and conditions (H0), (H1)p and (H2)p, when 1 < p < ∞. Proposition 2.8 Provided f satisfies (H0), (H1)p and (H2)p, then 1. f ∞ is convex-quasiconvex; 2. there exists C > 0 such that 1 C ξ ≤ f ∞ (b, ξ) ≤ C ξ ; (2.8) 3. f ∞(b, ξ) is constant with respect to b for every ξ ∈ Rn×N ; 4. f ∞ is continuous. Remark 2.9 We emphasize that not all the assumptions (H1)p and (H2)p in Proposition 2.8 are necessary to prove items above. In particular, one has that: i) The proof of 2. uses only the fact that f satisfies (H1)p. ii) To prove 3. it is necessary to require that f satisfies only (H0) and (H1)p. In fact, under the assumptions that f satisfies (2.3) one can avoid to require (H0). 5 Proof. 1. The convexity-quasiconvexity of f ∞ can be proven exactly as in [7, Lemma 2.1]. 2. By definition (1.4) we may find a subsequence {tk} such that f ∞ (b, ξ) = lim tk→∞ f (b, tkξ) tk . By (H1)p one has f ∞ (b, ξ) ≤ lim tk→∞ C (1 + bp + tkξ) tk = C ξ andf ∞ (b, ξ) ≥ lim tk→∞ 1 C (bp + tkξ) − C tk ≥ 1 C ξ . Hence (H1)p holds for f ∞. 3. We start by observing that (2.8) and 1. guarantee that f ∞ satisfies (2.3). Let ξ ∈ Rn×N , and let b, b′ ∈ Rm, up to a subsequence, by (1.4) and (2.3) it results that, f ∞(b, ξ) − f ∞(b′, ξ) ≤ lim tk→∞ f (b, tkξ) − f (b′, tkξ) tk ≤ lim tk→∞ γ(1 + bp−1 + b′p−1 + tkξ 1 p′ )b − b′ tk = 0. By interchanging the role of b and b′, it follows that f ∞(·, ξ) is constant and this concludes the proof. Remark 2.10 Under assumptions (H0), (H1)∞ and (H2)∞, f ∞ satisfies properties analogous to those at the beginning of subsection ??? In particular in [7, Lemma 2.1 and Lemma 2.2] it has been proved that i) f ∞ is convex-quasiconvex; ii) 1 CM ξ ≤ f ∞(b, ξ) ≤ CM ξ, for every b, with b ≤ M ; iii) If rankξ ≤ 1, then f ∞(b, ξ) is constant with respect to b. Remark 2.11 We observe that, if f : Rm × Rn×N → [0, ∞) is a continuous function satisfying (H1)p and (H2)p, then the function (CQf )∞ : Rm ×Rn×N → [0, ∞[, obtained first taking the convex-quasiconvexification in (2.5) of f and then its recession through formula (1.4) applied to CQf , satisfies the following properties: 1. (CQf )∞ is convex-quasiconvex; 2. there exists C > 0 such that 1 C ξ ≤ (CQf )∞(b, ξ) ≤ Cξ, for every (b, ξ) ∈ Rm × Rn×N ; 3. for every ξ ∈ Rn×N , (CQf )∞(·, ξ) is constant, i.e. (CQf )∞ is independent on v; 4. (CQf )∞ is Lipschitz continuous in ξ. Under the same set of assumptions on f , one can prove that the convex-quasiconvexification of f ∞, CQ(f ∞), satisfies the following conditions: 5. CQ(f ∞) is convex-quasiconvex; 6. there exists C > 0 such that 1 C ξ ≤ CQ(f ∞)(b, ξ) ≤ Cξ, for every (b, ξ) ∈ Rm × Rn×N ; 7. for every ξ ∈ Rn×N , and assuming that f satisfies (2.3), CQ(f ∞)(·, ξ) is constant, i.e. CQ(f ∞) is independent on b; 8. CQ(f ∞) is Lipschitz continuous in ξ. The above properties are immediate consequences of Propositions 2.7, 2.8 and (2.3). In particular 8. follows from 3. of Proposition 2.8, without requiring (H2)p. On the other hand, Proposition 2.12 below entails that CQ(f )∞ is independent on b, without requiring that f is Lipschitz continuous, but replacing this assumption with (H2)p. We also observe that (CQf )∞ and CQ(f ∞) are only quasiconvex functions, since they are independent of b. In particular, in our setting, these functions coincide as it is stated below. 6 Proposition 2.12 Let f : Rm × Rn×N → [0, ∞) be a continuous function satisfying (H1)p and (H2)p. Then CQ (f ∞) (b, ξ) = (CQf )∞ (b, ξ) for every (b, ξ) ∈ Rm × Rn×N . Proof. The proof will be achieved by double inequality. For every (b, ξ) ∈ Rm × Rn×N the inequality follows by Definition 2.6, and the fact that CQf (b, ξ) ≤ f (b, ξ). In fact, (1.4) entails that the same inequality holds when, passing to (·)∞. Finally, 1. in Proposition 2.8, guarantees (2.9). (CQf )∞(b, ξ) ≤ CQ(f ∞)(b, ξ) (2.9) In order to prove the opposite inequality, fix (b, ξ) ∈ Rm × Rn×N and, for every t > 1, take ηt ∈ L∞(Q; Rm), with 0 average, and ϕt ∈ W 1,∞ (Q; Rn) such that 0 f (b + ηt, tξ + ∇ϕt(y)) dy ≤ CQf (b, tξ) + 1. (2.10) ZQ By (H1)p and Proposition 2.7, we have that kb + ηtkLp(Q), (cid:13)(cid:13)∇( 1 t ϕt, one has ψt ∈ W 1,∞ on t. Defining ψt := 1 t ϕt)(cid:13)(cid:13)L1(Q) ≤ C for a constant independent (Q; Rn) and thus 0 CQ(f ∞)(b, ξ) ≤ZQ f ∞(b + ηt, ξ + ∇ψt(y)) dy. Let L be the constant appearing in condition (H2)p. We split the cube Q in the set {y ∈ Q : tξ + ∇ψt(y) ≤ L} and its complement in Q. Then we apply condition (H2)p and (2.8) to get CQ(f ∞)(b, ξ) ≤ ZQ(cid:18)C 1 + b + ηtp t + C ξ + ∇ψt1−τ tτ + f (b + ηt, tξ + ∇ϕt) t + C L t(cid:19) dy. Applying Holder inequality and (2.10), we get CQ(f ∞)(b, ξ) ≤ C tτ (cid:18)ZQ ξ + ∇ψt dy(cid:19)1−τ + CQf (b, tξ) + 1 t + C L t + C′ t , and the desired inequality follows by definition of (CQf )∞ and using the fact that ∇ψt has bounded L1 norm, letting t go to ∞. Remark 2.13 It is worth to observe that inequality (CQf ∞) (b, ξ) ≤ CQ (f ∞) (b, ξ) for every (b, ξ) ∈ Rm × Rn×N , has been proven without requiring neither (H1)p and (H2)p on f , nor (H1)∞ and (H2)∞. Furthermore, we emphasize that the proof of Proposition 2.12 cannot be performed in the same way in the case p = ∞, with assumptions (H1)p and (H2)p replaced by (H1)∞ and (H2)∞. Indeed, an L∞ bound on b + ηt analogous to the one in Lp cannot be obtained from (H1)∞. On the other hand, it is possible to deduce the equality between CQf ∞ and (CQf )∞, when f satisfies (2.6) and (2.7). 2.3 Auxiliary results Here we prove that assumption (H0) on f is not necessary to provide an integral representation for J p in (1.2). Indeed, we can assume that f : Rm × Rn×N → [0, ∞[ is a continuous function and satisfies assumptions (H1)p and (H2)p, (p ∈ (1, ∞]). First we extend, with an abuse of notation, the functional J in (1.1), to L1(Ω; Rn) × Lp(Ω; Rm), p ∈ (1, ∞], as ZΩ ∞ J(u, v) :=  f (v, ∇u)dx if (u, v) ∈ W 1,1(Ω; Rn) × Lp(Ω; Rm), (2.11) otherwise. 7 Then we define, according to Definition 2.6 the convex-quasiconvex envelope of f , CQf , and introduce, in analogy with (2.11) and (1.2), the functional ZΩ ∞ CQf (v, ∇u) dx if (u, v) ∈ W 1,1(Ω; Rn) × Lp(Ω; Rm), JCQf (u, v) :=  JCQf (uh, vh) : uh ∈ W 1,1 (Ω; Rn) , vh ∈ Lp (Ω; Rm) , uh → u in L1, vh ⇀ v in Lp(cid:27) , otherwise, (p ∈ (1, ∞]) and, JCQf p (u, v) := inf(cid:26)lim inf h→∞ for any pair (u, v) ∈ BV (Ω; Rn) × Lp (Ω; Rm) , p ∈ (1, ∞). Analogously, one can consider JCQf ∞ (u, v) := inf(cid:26)lim inf h→∞ JCQf (uh, vh) : uh ∈ W 1,1 (Ω; Rn) , vh ∈ Lp (Ω; Rm) , uh → u in L1, vh ∗ ⇀ v in L∞(cid:27) , for any pair (u, v) ∈ BV (Ω; Rn) × L∞ (Ω; Rm) . Clearly, it results that for every (u, v) ∈ BV (Ω; Rn) × Lp (Ω; Rm), JCQf p (u, v) ≤ J p(u, v), but, as in [11, Lemma 8 and Remark 9], the following proposition can be proven. Proposition 2.14 Let p ∈ (1, ∞] and consider the functionals J and JCQf and their corresponding relaxed functionals J p and JCQf p. If f satisfies conditions (H1)p and (H2)p if p ∈ (1, ∞), and both f and CQf satisfy (H1)∞ and (H2)∞ if p = ∞, then for every (u, v) ∈ BV (Ω; Rn) × Lp(Ω; Rm), p ∈ (1, ∞]. J p(u, v) = JCQf p(u, v) Remark 2.15 The argument has not been shown since it is already contained in [11, Lemma 8 and Remark 9]. In [11] it is not required that f satisfies (H2)p, (p ∈ (1, ∞]). Indeed, the coincidence between the two functionals J p and JCQf p holds independently on this assumption on f , but in order to remove hypothesis (H0) from the representation theorem we need to assume that CQf inherits the same properties as f , which is the case as it has been observed in Proposition 2.7. It is also worth to observe that, when p = ∞, (2.7) is equivalent to f ∞(b, ξ) − f (b, ξ) ≤ C(1 + α(b) + ξ) for every (b, ξ) ∈ Rm × Rn×N , and this latter property is inherited by CQf and CQf ∞ as it can be easily verified arguing as in [10, Proposition 2.3]. Thus Proposition 2.14 holds when p = ∞ just requiring that f satisfies (2.6) and (2.7). The following result can be deduced in full analogy with [11, Theorem 12], where it has been proven for J ∞. Proposition 2.16 Let Ω be a bounded and open set of RN and let f : Rm × Rn×N → R be a continuous function satisfying (H1)p and (H2)p, 1 < p ≤ ∞. Let J be the functional defined in (1.1), then J p in (1.2) (1 < p < ∞), (1.3) (p = ∞) is a variational functional. By virtue of this result, it turns out that for every (u, v) ∈ BV (Ω; Rn)×Lp(Ω; Rm), J p(u, v; ·), (p ∈ (1, ∞]) is the restriction to the open subsets in Ω of a Radon measure on Ω, thus it can be decomposed as the sum of two terms J p(u, v; ·) = J a p(u, v; ·) + J s p(u, v; ·), (2.12) a p(u, v; ·) and J where J the Lebesgue measure, respectively. Next proposition deals with the scaling properties of J p. s p(u, v; ·) denote the absolutely continuous part and the singular part with respect to 8 Proposition 2.17 Let f : Rm × Rn×N → R be a continuous and convex-quasiconvex function, let J and J p be the functionals defined respectively by (1.1) and (1.2) when p ∈ (1, ∞], respectively ( (1.3), when p = ∞). Then the following scaling properties are satisfied J p(u + η, v; Ω) = J p(u, v; Ω) for every η ∈ Rn, J p (u(· − x0), v(· − x0); x0 + Ω) = J p(u(·), v(·); Ω) for every x0 ∈ RN , (2.13) J p(cid:18)u, v; Ω − x0 (cid:19) = −N J p(u, v; Ω), where u(y) := u(x0+y)−u(x0) and v(y) := v(x0 + y), for y ∈ Ω−x0 . The following result will be exploited in the sequel. The proof is omitted since it develops along the lines of [2, Lemma 5.50], the only differences being the presence of v and the convexity-quasiconvexity of f . Lemma 2.18 Let f : Rm × RN ×n → R be a continuous and convex-quasiconvex function, and let J and J p be the functionals defined respectively by (1.1) and (1.2). Let ν ∈ SN −1, η ∈ Sn−1 and ψ : R → R, bounded and increasing. Denoted by Q the cube Qν, let u ∈ BV (Q; Rn) be representable in Q as and let w ∈ BV (Q; Rn) be such that supp(w − u) ⊂⊂ Q. Let v ∈ Lp(Q; Rm). Then u(y) = ηψ(y · ν), J p(w, v; Q) ≥ f(cid:18)ZQ vdy, Dw(Q)(cid:19) . 3 Main Results This section is devoted to deduce the results stated in Theorems 1.1 and 1.2. We start by proving the lower bound in the case 1 < p < ∞. For what concerns the upper bound we present, for the reader's convenience, a self contained proof in Theorem 3.2. For the sake of completeness we observe that the upper bound, in the case 1 < p < ∞, could be deduced as a corollary from the case p = ∞ (see Theorem 1.2), which, in turn, under slightly different assumptions, is contained in [7]. 3.1 Lower semicontinuity in BV × Lp, 1 < p < ∞ Theorem 3.1 Let Ω be a bounded open set of RN , let f : Rm × Rn×N → [0, ∞) be a continuous function satisfying (H0), (H1)p and (H2)p, and let J p be the functional defined in (1.2). Then J p(u, v; Ω) ≥ZΩ f (v, ∇u)dx +ZΩ f ∞(cid:18)0, dDsu d Dsu(cid:19) d Dsu for any (u, v) ∈ BV (Ω; Rn) × Lp(Ω; Rm). Proof. The proof will be achieved, in two steps, namely by showing that lim →0+ Jp(u, v; Q(x0; )) LN (Q(x0, )) ≥ f (v(x0), ∇u(x0)), for LN − a.e. x0 ∈ Ω, lim →0+ Jp (u, v; Q(x0, )) Du(Q(x0, )) ≥ f ∞(cid:18)0, dDsu d Dsu (x0)(cid:19) , for Dsu − a.e. x0 ∈ Ω. (3.1) (3.2) (3.3) (3.4) Indeed, if (3.2) and (3.4) hold then, by virtue of (2.12), and [2, Theorem 2.56], (3.1) follows immediately. 9 Step 1. exploiting [11, Theorem 1.1]. Inequality (3.2) is obtained through an argument entirely similar to [2, Proposition 5.53] and For LN −a.e. x0 ∈ Ω it results that u is approximately differentiable (see (2.1)) and lim →0+ 1 LN (Q(x0, ))ZQ(x0,) v(x) − v(x0)dx = 0. Consequently, given > 0, and defined u and v as in Proposition 2.17, it results that u → u0 in L1(Ω; Rn), where u0 := ∇u(x0)x and v → v(x0) in Lp(Ω; Rm). Then the scaling properties (2.13), and the lower semicontinuity of J p entail that lim inf →0+ J p(u, v; Q(x0, )) N = lim inf →0+ J p(u, v; Q) ≥ J p(u0, v(x0); Q). (3.5) Then the lower semicontinuity result proven in [11, Theorem 11], when u is in W 1,1(Ω; Rn) and v ∈ Lp(Ω; Rm), allows us to estimate the last term in (3.5) as follows J p(u0, v(x0); Q) ≥ f (v(x0), ∇u(x0)), and that provides (3.2). Step 2. Here we present the proof of (3.4). To this end we exploit techniques very similar to [1] (see [2, Proposition 5.53]). Let Du = zDu be the polar decomposition of Du (see [2, Corollary 1.29]), for z ∈ SN ×n−1, and recall that for Dsu- a.e. x0, z(x0) admits the representation η(x0) ⊗ ν(x0), with η(x0) ∈ Sn−1 and ν(x0) ∈ SN −1, (see [2, Theorem 3.94]). In the following, we will denote the cube Qν(x0, 1) by Q. To achieve (3.4) it is enough to show that lim →0+ J p(u, v; Q(x0, )) Du(Q(x0, )) ≥ f ∞(0, z(x0)) at any Lebesgue point x0 of z relative to Du such that the limit on the left hand side exists and z(x0) = η(x0) ⊗ ν(x0), lim →0+ Du(Q(x0, )) N = ∞, 0 = lim →0+ RQ(x0,) vpdx Du(Q(x0, )) = lim →0+ RQ(x0,) vdx Du(Q(x0, )) . (3.6) (3.7) The above requirements are, indeed, satisfied at Dsu-a.e. x0 ∈ Ω, by Besicovitch's derivation theorem and Alberti's rank-one theorem (see [2, Theorem 3.94]). Set η ≡ η(x0) and ν ≡ ν(x0), for < N − 1 2 dist(x0, ∂Ω), define u(x0 + y) − u N u(y) := Du(Q(x0, )) , y ∈ Q, where u is the average of u in Q(x0, ). Analogously define, as in Proposition 2.17, v(y) := v(x0 + y), y ∈ Q. Let us fix t ∈ (0, 1). By [2, formula (2.32)], there exists a sequence {h} converging to 0 such that lim h→∞ Du(Q(x0, th)) Du(Q(x0, h)) ≥ tN . (3.8) (3.9) Denote uh by uh, then Duh(Q) = 1 and, passing to a not relabelled subsequence, {uh} converges in L1(Q; Rn) to a BV function u. Correspondingly, denote vh by vh. Then, arguing as in [2, Proof of Proposition 5.53] we have Du(Q) ≤ 1 and Du(Qt) ≥ tN , (3.10) 10 where Qt := tQ. It results that u(y) = ηψ(y · ν), for some bounded increasing function ψ in (cid:0)− 1 c (Q) such that ϕ = 1 on Qt and 0 ≤ ϕ ≤ 1, and let us define wh := ϕuh + (1 − ϕ)u. The functions wh ϕ ∈ C1 converge to u in L1(Q; Rn) and moreover we have 2(cid:1). Take 2 , 1 D(wh − uh)(Q) ≤ D(uh − u)(Q \ Qt) +ZQ ≤ Duh(Q \ Qt) + Du(Q \ Qt) +ZQ ∇ϕuh − udy. ∇ϕuh − udy Therefore, by (3.9) and (3.10), one has lim sup h→∞ D(wh − uh)(Q) ≤ 2(1 − tN ). Similarly, consequently Dwh(Q \ Qt) ≤ Duh(Q \ Qt) + Du(Q \ Qt) +ZQ ∇ϕuh − udy, lim sup h→∞ Dwh(Q \ Qt) ≤ 2(1 − tN ). (3.11) (3.12) Setting ch := Du(Q(x0,h)) N h conditions (H1)p, we have , by the scaling properties of J p in Proposition 2.17 and by the growth J p(u, v; Q(x0, h)) Du(Q(x0, h)) = J p(chuh, vh; Q) ch ≥ J p(chwh, vh; Qt) ch ≥ J p(chuh, vh; Q) ch − C(c−1 h Q \ Qt + Dwh(Q \ Qt) + c−1 h ZQ\Qt vhpdy). By (3.6), ch → ∞, moreover taking into account (3.8) and (3.7), by (3.12), it results that lim →0+ J p(u, v; Q(x0, )) Du(Q(x0, )) ≥ lim sup h→∞ J p(chuh, vh; Q) ch − 2C(1 − tN ). On the other hand, Lemma 2.18 entails that, for every h ∈ N, J p(chwh, vh; Q) ≥ f(cid:18)ZQ vhdy, chDwh(Q)(cid:19) ≥ f(cid:18)ZQ vhdy, chDuh(Q)(cid:19) − chγD(wh − uh)(Q), where γ is the constant appearing in (2.3). Then by (3.11), we have that lim →0+ J p(u, v; Q(x0, )) Du(Q(x0, )) ≥ lim sup h→∞ f(cid:16)RQ vhdy, chDuh(Q)(cid:17) ch − 2(C + γ)(1 − tN ). By the definition of uh, Duh(Q) = Du(Q(x0,h)) Now, taking into account (2.3) and (H2)p , we have Du(Q(x0,h)) , hence Duh(Q) → z(x0), since x0 is a Lebesgue point of z. lim sup h→∞ ch f(cid:16)RQ vhdy, chDuh(Q)(cid:17) f ∞(cid:18)ZQ = lim h→∞   = f ∞(z(x0)), Lp norm in the unit cube with respect to p (i.e. RQ vhdyp ≤RQ vhpdy), and (3.7). vhdy, z(x0)(cid:19) + C(cid:12)(cid:12)(cid:12)RQ vhdy(cid:12)(cid:12)(cid:12) + 1 ch p where it has been exploited the fact that ch → ∞, 3. of Proposition 2.8, the nondecreasing behaviour of the f(cid:16)RQ vhdy, chz(x0)(cid:17) ch = lim h→∞ 11 3.2 Relaxation We start by observing that Theorem 1.2 is contained in [7] under a uniform coercivity assumption. We do not propose the proof in our setting, since it develops along the lines of Theorems 3.1 and 3.2. On the other hand, several observations about Theorem 1.2 are mandatory: i) If f satisfies (H1)p and (H2)p then J p(u, v) ≤ J ∞(u, v) for every (u, v) ∈ BV (Ω; Rn) × L∞(Ω; Rm). ii) For the reader's convenience we observe that the proof of the lower bound in Theorem 1.2 develops exactly as that of Theorem 3.1, using the L∞ bound on v to deduce (3.7) and the uniform bound on v in (3.8), (H2)∞ and (2.4) in order to estimate lim sup h→∞ f(cid:16)RQ vhdy, chDuh(Q)(cid:17) ch . Regarding the upper bound, the bulk part follows from [11, Theorems 12 and 14], while for the singular part we can argue exactly as proposed in the proof of the upper bound in [7] just considering conditions (H1)∞ and (H2)∞ in place of (H1)p and (H2)p. iii) The above arguments remain true under assumptions (2.6) and (2.7). We are now in position to prove the upper bound for the case BV × Lp, for 1 < p < ∞. We emphasize that an alternative proof could be obtained via a truncation argument from the case p = ∞ as the one presented in [11, Theorem 12], but we prefer the self contained argument below. Theorem 3.2 Let Ω be a bounded open set of RN and let f : Rm ×Rn×N → [0, ∞) be a continuous function. Then, assuming that f satisfies (H0), (H1)p and (H2)p, J p (u, v) ≤ZΩ f (v, ∇u) dx +ZΩ f ∞(cid:18)0, dDsu d Dsu (x)(cid:19) d Dsu (x) , for every (u, v) ∈ BV (Ω; Rn) × Lp(Ω; Rm). Proof. First we observe that Proposition 2.16 entails that J p is a variational functional. Thus the inequality can be proved analogously to [2, Proposition 5.49]. For what concerns the bulk part, it is enough to observe that given u ∈ BV (Ω; Rn) and v ∈ Lp(Ω; Rm), taking a sequence of standard mollifiers {εk }, where εk → 0, it results that ∇uk = ∇u ∗ εk + Dsu ∗ εk , where uk := u ∗ εk . The local Lipschitz behaviour of f in (2.3) gives ZA f (v, ∇uk)dx ≤ZA f (v, ∇u ∗ εk )dx + γDsu(Iεk (A)) for every k ∈ N, where Iεk (A) denotes the εk neighborhood of A. Then if Dsu(∂A) = 0, letting εk → 0, we obtain for every open subset A of Ω. Thus we can conclude that J p(u, v; A) ≤ZA f (v, ∇u)dx + γDsu(A), J a p(u, v; B) ≤ZB f (v(x), ∇u(x))dx for every (u, v) ∈ BV (Ω; Rn) × Lp(Ω; Rm) and B Borel subset of Ω. To achieve the result, it will be enough to show that J s p(u, v; B) ≤ZB f ∞(cid:18)0, dDsu dDsu(cid:19) dDsu for every B Borel subset of Ω. For every ξ ∈ Rn×N and b ∈ Rm, define the function g(b, ξ) := sup t≥0 f (t 1 p b, tξ) − f (0, 0) t . 12 It is easily seen that g is (p, 1)-positively homogeneous, i.e. tg(b, ξ) = g(t Rm × Rn×N , g is continuous and, since f satisfies (2.3), g inherits the same property. 1 p b, tξ) for every t > 0, (b, ξ) ∈ Moreover, the monotonicity property of difference quotients of convex functions ensures that, whenever rank ξ ≤ 1, g(b, ξ) = f ∞ p (b, ξ), where the latter is defined as f ∞ p (b, ξ) := lim sup t→∞ f (t 1 p b, tξ) t . In particular g(0, ξ) = f ∞(0, ξ) = f ∞ Then for every open set A ⊂⊂ Ω such that Du(A) = 0, defining for every h ∈ N, uh := u ∗ εh and vh := v where {εh } is a sequence of standard mollifiers and εh → 0. Then uh → u in L1. Also [2, Theorem 2.2] entails that Duh → Du weakly ∗ in A and Duh(A) → Du(A). Thus p (0, ξ), whenever rank ξ ≤ 1. J p(u, v; A) ≤ lim inf h→∞ ZA f (v, ∇uh)dx ≤ lim sup h→∞ ZA f (v, 0)dx + lim inf h→∞ ZA g(v, ∇uh)dx. (3.13) For what concerns the first term in the right hand side, we have that it is bounded by RA(1 + vp)dx, thus taking the Radon-Nikod´ym derivative with respect to Dsu we obtain 0. Regarding the second term in the right hand side of (3.13), we have g (v(x), Du ∗ h) dx lim inf h→∞ ZA h→∞ ZA ≤ lim sup g (0, Du ∗ h) dx + CZA v(x)pdx +ZA v(x)Du ∗ h 1 p′ dx. Taking the Radon-Nikod´ym derivative, the last two terms disappear, since Du ∗ h → Du, vpLN is singular with respect to Dcu and the Holder inequality can be applied, i.e. v(x)Du ∗ h ZA 1 p′ dx ≤(cid:18)ZA v(x)pdx(cid:19) 1 p (cid:18)ZA Du ∗ ndx(cid:19) 1 p′ . Then the thesis is achieved via the same arguments as in [2, Proposition 5.49]. Remark 3.3 It is worth to observe that an alternative argument to the one presented above, concerning the upper bound inequality for the singular part, can be provided by means of approximation. In fact, one can prove that J s p(u, v; B) ≤ZB f ∞(cid:18)0, dDsu dDsu(cid:19) dDsu for every B Borel subset of Ω, when u ∈ BV (Ω; Rn) and v ∈ C(Ω; Rm), and then via a standard approximation argument via mollification allows to reach every v ∈ Lp(Ω; Rm). For what concerns the case v ∈ C(Ω; Rm) it is enough to consider the function g(b, ξ) := sup t≥0 f (b, tξ) − f (b, 0) t , exploit its properties of positive 1-homogeneity in the second variable, i.e. tg(b, ξ) = g(b, tξ), for every t > 0, (b, ξ) ∈ Rm × Rn×N , (2.4), and the fact that when rank ξ ≤ 1, then g(b, ξ) is constant with respect to b and f ∞(b, ξ) = g(b, ξ) = f ∞(0, ξ). To conclude it is enough to apply Reshetnyak continuity theorem. Proof of Theorem 1.1. remove assumption (H0). The result follows from Theorems 3.1 and 3.2, applying Proposition 2.14 to Acknowledgements The research of the authors has been partially supported by Funda¸cao para a Ciencia e Tecnologia (Portuguese Foundation for Science and Technology) through UTA-CMU/MAT/0005/2009 and CIMA-UE. 13 References [1] L. Ambrosio & G. Dal Maso, On the Relaxation in BV (Ω; Rm) of Quasi-convex Integrals. Journal of Functional Analysis, 109, (1992), 76-97. [2] L. Ambrosio, N. Fusco & D. Pallara, Functions of Bounded Variation and Free Discontinuity Problems. Clarendon Press, Oxford, (2000). [3] L. Ambrosio, S. Mortola & V. M. Tortorelli, Functionals with linear growth defined on vector valued BV functions. J. Math. Pures Appl., IX. S´er. 70, No.3, (1991), 269-323. [4] G. Carita, A.M. Ribeiro & E. Zappale, An homogenization result in W 1,p × Lq.J. Convex Anal. 18, n. 4, (2011), 1093-1126. [5] G. Carita, A.M. Ribeiro & E. Zappale, Relaxation for some integral functionals in W 1,p w × Lq w. Bol. Soc. Port. Mat.,(2010), Special Issue, 47-53. [6] B. Dacorogna, Direct Methods in the Calculus of Variations. Second Edition, Springer, (2008). [7] I. Fonseca, D. Kinderlehrer & P. Pedregal, Relaxation in BV × L∞ of functionals depending on strain and composition. Lions, Jacques-Louis (ed.) et al., Boundary value problems for partial differential equations and applications. Dedicated to Enrico Magenes on the occasion of his 70th birthday. Paris: Masson. Res. Notes Appl. Math., 29, (1993), 113-152. [8] I. Fonseca, D. Kinderlehrer & P. Pedregal, Energy functionals depending on elastic strain and chemical composition. Calc. Var. Partial Differential Equations, 2, (1994), 283-313. [9] H. Le Dret & A. Raoult, Variational convergence for nonlinear shell models with directors and related semicontinuity and relaxation results. Arch. Ration. Mech. Anal. 154, No. 2, (2000), 101-134. [10] A. M. Ribeiro & E. Zappale, Relaxation of certain integral functionals depending on strain and chemical composition. Chinese Annals of Mathematics Series B, 34B(4) (2013), 491-514. [11] A. M. Ribeiro & E. Zappale, Lower semicontinuous envelopes in W 1,1 × Lp. Banach Center Publ. 101 (2014), 187-206. 14
1807.07059
2
1807
2018-10-01T13:22:36
Discrepancy for convex bodies with isolated flat points
[ "math.FA" ]
We consider the discrepancy of the integer lattice with respect to the collection of all translated copies of a dilated convex body having a finite number of flat, possibly non-smooth, points in its boundary. We estimate the $L^{p}$ norm of the discrepancy with respect to the translation variable as the dilation parameter goes to infinity. If there is a single flat point with normal in a rational direction we obtain an asymptotic expansion for this norm. Anomalies may appear when two flat points have opposite normals. When all the flat points have normals in generic irrational directions, we obtain a smaller discrepancy. Our proofs depend on careful estimates for the Fourier transform of the characteristic function of the convex body.
math.FA
math
DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS LUCA BRANDOLINI, LEONARDO COLZANI, BIANCA GARIBOLDI, GIACOMO GIGANTE, AND GIANCARLO TRAVAGLINI Abstract. We consider the discrepancy of the integer lattice with respect to the collection of all translated copies of a dilated convex body having a finite number of flat, possibly non-smooth, points in its boundary. We estimate the Lp norm of the discrepancy with respect to the translation variable as the dilation parameter goes to infinity. If there is a single flat point with normal in a rational direction we obtain an asymptotic expansion for this norm. Anomalies may appear when two flat points have opposite normals. When all the flat points have normals in generic irrational directions, we obtain a smaller discrepancy. Our proofs depend on careful estimates for the Fourier transform of the characteristic function of the convex body. Let B be a convex body in Rd, that is a convex bounded set with nonempty 1. Introduction interior, and for every R > 1 and z ∈ Rd let DR (z) = −Rd B + Xm∈Zd χRB (z + m) be the discrepancy between the number of integer points inside a dilated and trans- lated copy of B and its volume. The function z 7→DR (z) is periodic and a straight- forward computation shows that it has the Fourier expansion (1) (3) RdbχB (Rm) e2πim·z Xm∈Zd\{0} bχB (ζ) =ZB e−2πiζ·zdz. as ζ → +∞. For example, if the boundary of B is smooth and has everywhere where bχB (ζ) denotes the Fourier transform of χB (z), that is The size of DR (z) as R → +∞ is therefore closely connected to the decay ofbχB (ζ) positive Gaussian curvature then bχB (ζ) has the decay bχB (ζ) 6 cζ− d+1 6 (see [28, Chapter 8]), and it can be shown that this rate of decay is optimal. Under the assumption (2), in [2, Corollary 3] the authors proved the following estimates for the Lp norm of the discrepancy function log 2d (R) p = 2d/ (d − 1) , (d+1) (1− 1 p ) p > 2d/ (d − 1) . (cid:18)ZTd DR (z)p dz(cid:19)1/p 1 6 p < 2d/ (d − 1) , cR cR cR d−1 2 2 d−1 (2) d−1 d(d−1) 2 1991 Mathematics Subject Classification. 11H06, 42B05. Key words and phrases. Discrepancy, Integer points, Fourier analysis. 1 On the other hand, when B is symmetric about a point and d ≡ 1 (mod 4), (cid:18)ZTd DR (z)p dz(cid:19)1/p 2 (cid:18)ZTd DR (z)p dz(cid:19)1/p 2 (cid:18)ZTd DR (z)p dz(cid:19)1/p lim sup R→+∞ lim inf R→+∞ R− d−1 R− d−1 > 0 for every p > 1, = 0 for every p < 2d d − 1 . 2 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI In [2, Theorem 5] it has also been shown that the above estimates are sharp in the range 1 6 p < 2d/ (d − 1). More precisely, using the asymptotic expansion d 6≡ 1 (mod 4) one has, for every p > 1, for bχB (ζ), it has been proved that when B is not symmetric about a point or > cR d−1 2 . Up to now we have considered the case of positive Gaussian curvature. When the Gaussian curvature of the boundary of B vanishes at some point the estimate (2) fails and the rate of decay depends on the direction. More precisely the decay of the Fourier transform (2) holds in a given direction Θ if the Gaussian curvature does not vanish at the points on the boundary of B where the normal is ±Θ. When the curvature vanishes the rate of decay ofbχB (ρΘ) can be significantly smaller. We will see that in this case the behavior of the Lp norms of the discrepancy function may differ from the case of positive Gaussian curvature. To the authors' knowledge the discrepancy for convex bodies with vanishing Gaussian curvature has been considered only for specific classes of convex bodies and only for L∞ estimates. See e.g. [8], [10], [14], [15], [20], [21], [22], [26], [27]. See also [21] for an estimate from below of the L2 discrepancy associated to the curve x2 + y4 = 1. Throughout the paper we will use bold symbols only for d-dimensional points and non-bold symbol for lower dimensional points. Moreover when we write a point z = (x, t) or ζ = (ξ, s) we agree that x, ξ ∈ Rd−1 and t, s ∈ R. We are happy to thank Gabriele Bianchi for some interesting remarks on the geometric properties of the convex bodies considered in this paper (see [1]). 2. Statements of the results In this paper we study the Lp norms of the discrepancy function associated to a convex body whose boundary has a finite number of isolated flat points. The relevant example is a convex body B such that ∂B has everywhere positive Gaussian curvature except at the origin and such that, in a neighborhood of the origin, ∂B is the graph of the function t = xγ, with x ∈ Rd−1 and some γ > 2. This function is smooth at the origin only when γ is a positive even integer, and the geometric control of the Fourier transform in [7] does not apply directly. We are actually interested in a larger class of convex bodies and this is why we introduce the following definition. Definition 1. Let U be a bounded open neighborhood of the origin in Rd−1, let Φ ∈ C∞ (U \ {0}) and let γ > 1. For every x ∈ U \ {0} let µ1 (x) , . . . , µd−1 (x) be the eigenvalues of the Hessian matrix of Φ. We say that Φ ∈ Sγ (U ) if for j = 1, . . . , d − 1, 0 < inf x∈U\{0}x2−γ µj (x) (4) DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 3 and, for every multi-index α, sup x∈U\{0}xα−γ(cid:12)(cid:12)(cid:12)(cid:12) ∂αΦ ∂xα (x)(cid:12)(cid:12)(cid:12)(cid:12) < +∞. Observe that if Φ ∈ Sγ (U ) then for some c1, c2 > 0, c1 xγ−2 6 µj (x) 6 c2 xγ−2 . (5) Moreover, since γ > 1, we have Φ ∈ C1 (U ). Definition 2. Let B be a convex body in Rd and let z ∈ ∂B and let γ > 1. We say that z is an isolated flat point of order γ if, in a neighborhood of z and in a suitable Cartesian coordinate system with the origin in z, ∂B is the graph of a function Φ ∈ Sγ (U ), as in the previous definition. Convex bodies with flat points can be easily constructed by taking powers of strictly convex functions. Proposition 3. Let U be a bounded open neighborhood of the origin in Rd−1, let H ∈ C∞ (U ) such that H (0) = 0, ∇H (x) = 0 and assume that its Hessian matrix is positive definite at the origin. Let γ > 1. Then the function Φ (x) = [H (x)]γ/2 ∈ Sγ (U ). We have already observed that some of the results in this paper for the singularity xγ with γ even integer are not new. However observe that x2n is analytic, while the above definition does not imply that the boundary is smooth. For example, in dimension 2 consider a singularity of the kind Φ′′ (x) = 2 + sin (log (x)). Interestingly, in the following Proposition 11 concerning te decay of the Fourier transform of a convex body with a flat point of order γ, the case γ = 2 with non smooth flat points requires some extra care. The discrepancy for convex bodies with flat points in the above class is described by the following theorem. Theorem 4. Let B be a bounded convex body in Rd. Assume that ∂B is smooth with everywhere positive Gaussian curvature except for a finite number of isolated flat points of order at most γ. 1) For 1 < γ 6 2 we have (cid:18)ZTd DR (z)p dz(cid:19)1/p 1 6 p < 2d/ (d − 1) , 2 d(d−1) log 2d (R) p = 2d/ (d − 1) , p ) (d+1) (1− 1 p > 2d/ (d − 1) . 1 6 p 6 (2d) / (d + 1 − γ) p > (2d) / (d + 1 − γ) d−1 2 d−1 d−1 cR cR cR 6 6( cR(d−1)(1− 1 γ ) d+1 (1− 2 γp ) cR d(d−1) 2) For 2 < γ 6 d + 1 we have (cid:18)ZTd DR (z)p dz(cid:19)1/p 3) For γ > d + 1 and every p > 1 we have (cid:18)ZTd DR (z)p dz(cid:19)1/p 6 cR(d−1)(1− 1 γ ). The picture summarizes our estimates for the discrepancy. 4 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI 1/p 1 d−1 2d d−1 2 R (d−1)(cid:16)1− 1 γ (cid:17) R d(d−1) d+1 (cid:16)1− 1 p (cid:17) R R d(d−1) d+1 (cid:16)1− 2 γp (cid:17) 1 2 d + 1 γ The proof of the above theorem relies on precise estimate for the Fourier trans- form of χB (z). See Proposition 11 below. The estimates in point 1) are the same as in [2] for the case of positive Gaussian curvature and are independent of γ. On the contrary, as we will see from the proof, in the cases 2) and 3) the flat points give the main contribution. In the case 2) if p = +∞ then d (d − 1) . d + 1 d (d − 1) d + 1 (cid:18)1 − 2 pγ(cid:19) = Hence, when γ 6 d + 1 the estimates for the L∞ discrepancy of B match Landau's estimates for the L∞ discrepancy of the ball. See e.g. [24]. For p ≥ 2d/ (d − 1) the above result extends a theorem of Colin de Verdiere [8]. In the next theorem we consider convex bodies with a flat point with normal pointing in a rational direction. In this case, some of the previous estimates can be improved to asymptotic estimates. Theorem 5. Let B be a bounded convex body in Rd. Assume that ∂B is smooth with everywhere positive Gaussian curvature except at most at two points P and Q with outward unit normals −Θ and Θ which are flat of order γP and γQ respectively. Let S (t) = {z ∈ B : z · Θ = t} be (d − 1)-dimensional measure of the slices of B that are orthogonal to Θ. The function S (t) is supported in P · Θ 6 t 6 Q · Θ and is known to be smooth in P · Θ < t < Q· Θ. Assume that there exist two smooth functions GP (r) and GQ (r) with GP (0) 6= 0 and GQ (0) 6= 0 such that, for u > 0 sufficiently small S (P · Θ + u) = u and S (Q · Θ − u) = u d−1 γP GP(cid:16)u1/γP(cid:17) γQ GQ(cid:16)u1/γQ(cid:17) . d−1 Finally, assume that the direction Θ is rational, that is αΘ ∈ Zd for some α, and denote by m0 the first non-zero integer point in the direction Θ. Define AP (z) = 2GP (0) Γ(cid:16) d−1 γP d−1 γP (2π m0) + 1(cid:17) +1 k−1− d−1 γP sin(cid:18)2πkm0 · z − π 2 d − 1 γP (cid:19) , +∞Xk=1 DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 5 and AQ (z) = − 2GQ (0) Γ(cid:16) d−1 γQ d−1 γQ (2π m0) + 1(cid:17) +1 k −1− d−1 γQ sin(cid:18)2πkm0 · z + π 2 d − 1 γQ (cid:19) . +∞Xk=1 1) Let γP > γQ > 2 and assume that one of the two alternatives holds: 2 < γP 6 d + 1 and p < (2d) / (d + 1 − γP ) , or Then there exist constants δ > 0 and c > 0 such that for every R > 1, γP > d + 1 and p 6 +∞. In particular, as R → +∞, we have the following asymptotic (cid:18)ZTd(cid:12)(cid:12)(cid:12)DR (z) − R(d−1)(1−1/γP )AP (z−RP )(cid:12)(cid:12)(cid:12) (cid:18)ZTd DR (z)p dz(cid:19)1/p p dz(cid:19)1/p ∼ R(d−1)(1−1/γP )(cid:18)ZTd AP (z)p dz(cid:19)1/p . 2) Let γP = γQ = γ and assume that one of the two alternatives holds: 6 cR(d−1)(1−1/γP )−δ. 2 < γ 6 d + 1 and p < (2d) / (d + 1 − γ) , or Then there exist constants δ > 0 and c > 0 such that for every R > 1, γ > d + 1 and p 6 +∞. (cid:18)ZTd(cid:12)(cid:12)(cid:12)DR (z) − R(d−1)(1−1/γ) (AP (z−RP ) + AQ (z − RQ))(cid:12)(cid:12)(cid:12) 6 cR(d−1)(1−1/γ)−δ. p dz(cid:19)1/p Note that the series that define AP (z) and AQ (z) converge uniformly and ab- solutely. In particular these functions are bounded and continuous. Observe that the asymptotic estimate of point 1) includes the case of a single flat point, that is γP > γQ = 2. In point 2) it is not excluded that for particular values of P , Q, GP (0), GQ (0) and R, the terms AP (z−RP ) and AQ (z − RQ) may cancel each other and the discrepancy gets smaller. Corollary 6. Under the assumptions in point 2 in the previous theorem assume furthermore that GP (0) = GQ (0) and that (d − 1) /γ is an even integer. Then for every R such that Rm0 · (P − Q) is an integer we have In particular with this choice of the parameters AP (z−RP ) + AQ (z − RQ) = 0. (cid:18)ZTd DR (z)p dz(cid:19)1/p 6 cR(d−1)(1−1/γ)−δ. The case γ = 2 is not covered by the above corollary, however observe that for γ = (d − 1) / (2k) = 2, that is d ≡ 1 (mod 4), and p < 2d/ (d − 1), one formally would obtain R→+∞(R lim inf d−1 2 (cid:18)ZTd DR (z)p dz(cid:19)1/p) = 0. Actually this is true, even if the proof is more delicate. The case of a ball and p = 2 has been proved by L. Parnovski and A. Sobolev in [23]. Moreover, in [2] it is shown that this phenomenon also occurs for convex smooth domains with positive Gaussian curvature and p < 2d/ (d − 1) if and only if the domains are symmetric and d ≡ 1 (mod 4). 6 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI As remarked by Kendal the above Lp estimates for the discrepancy can be turned into almost everywhere pointwise estimates using a Borel-Cantelli type argument. See [21, §3] for the proof. Proposition 7. Assume that for some β > 0 let λ (t) be an increasing function and let Rn → +∞ such that Then for almost every z ∈ Td there exists c > 0 such that 6 κRβ, (cid:18)ZTd DR (z)p(cid:19)1/p +∞Xn=1 DRn (z) < cRβ λ (Rn)−p < +∞. nλ (Rn) . If the flat points on the boundary of the domain B have "irrational" normals then the discrepancy can be smaller than the one described in the above theorems. In particular, we have the following result that applies to every convex body, without curvature or smoothness assumption. Theorem 8. Let B be a bounded convex body in Rd and for σ ∈ SO (d) denote by DR,σ the discrepancy associated to the rotated body σB. Then we have the following mixed norm inequalities. 1) If 1 6 p 6 2, we have 2) If 2 6 p < 2d/ (d − 1), we have ZSO(d)(cid:18)ZTd DR,σ (z)p dz(cid:19)2/p ZSO(d)(cid:18)ZTd DR,σ (z)p dz(cid:19)1/(p−1) dσ!1/2 6 cR d−1 2 . p dσ! p−1 6 cR d−1 2 . For the planar case d = 2 we can state a slightly more precise result. Theorem 9. Let B be a bounded convex body in R2. Assume that ∂B is smooth with everywhere positive curvature except a single flat point of order γ > 2. Let (α, β) be the unit outward normal at the flat point and assume the Diophantine property that for some δ < 2/ (γ − 2) there exists c > 0 such that for every n ∈ Z Here kxk denotes the distance of x from the closest integer. Then (cid:13)(cid:13)(cid:13)(cid:13)n α β(cid:13)(cid:13)(cid:13)(cid:13) > c n1+δ . (cid:18)ZTd DR (z)2 dz(cid:19)1/2 6 cR 1 2 . By a classical result of Jarnik (see e.g. [12, §10.3]) the set of real numbers ω that are (2 + δ)-well approximable, that is knωk 6 n1−(2+δ) = n−1−δ for infinitely many n, has Hausdorff dimension 2 2+δ . In particular the exceptional set in the above theorem, where the discrepancy may be larger than R1/2 has Hausdorff dimension at most γ−2 γ−1 . DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 7 3. Estimates for the Fourier transforms The main ingredient in the proof of our results on the discrepancy comes from integrals. suitable estimates of the decay of bχB (ζ). We start studying a family of oscillatory As usual we write d-dimensional points through the notation z = (x, t) and ζ = (ξ, s) (see the Introduction). Lemma 10. Let U ⊂ Rd−1 be an open ball about the origin of radius b, let Φ ∈ Sγ (U ) for some γ > 1, let ψ be a smooth function supported in {0 < a 6 x 6 b}, for every positive integer k let Φk (x) = 2kγΦ(cid:0)2−kx(cid:1) and let Then there exist constants c, c1, c2 > 0 and, for every M > 0, a constant cM such that for every k > 0 Ik (ξ, s) =ZRd−1(cid:0)∇Φ(cid:0)2−kx(cid:1) ,−1(cid:1) e−2πi(ξ,s)·(x,Φk(x))ψ (x) dx. Ik (ξ, s) 6 c (1 + s + ξ)− d−1 cM (1 + s)−M cM (1 + ξ)−M for every (ξ, s) if ξ 6 c1 s , if c2 s 6 ξ . 2 Proof. The behaviour of the oscillatory integral Ik (ξ, s) depends on the points where the amplitude ψ (x) is not zero and the phase (ξ, s)· (x, Φk (x)) is stationary. This happens only when ξ ≈ s and in this case, since the phase is non degener- ate, one obtains the classical estimate c(ξ, s)−(d−1)/2. In all other directions the oscillatory integral has a fast decay. In particular, when ξ 6 c1 s one obtains the decay cM (1 + s)−M , and when ξ > c2 s, one obtains the decay cM (1 + ξ)−M . For the sake of completeness we include the full details of the proof. By the definition of the class Sγ we have In particular when x belongs to the support of ψ (x) we have (cid:12)(cid:12)(cid:12)(cid:12) ∂αΦk 6 cα xγ−α . ∂xα (x)(cid:12)(cid:12)(cid:12)(cid:12) = 2kγ2−kα(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ∂αΦ ∂xα (cid:0)2−kx(cid:1)(cid:12)(cid:12)(cid:12)(cid:12) 6 2kγ2−kαcα(cid:12)(cid:12)2−kx(cid:12)(cid:12)γ−α ∂xα (x)(cid:12)(cid:12)(cid:12)(cid:12) 6 cα. ∂αΦk Moreover, the Hessian matrix of Φk (x) satisfies and it follows that the eigenvalues µ(k) values µj (x) of Hess Φ (x) by the identity j (x) of Hess Φk (x) are related to the eigen- Hess Φk (x) = 2k(γ−2) Hess Φ(cid:0)2−kx(cid:1) (x) = 2k(γ−2)µj(cid:0)2−kx(cid:1) . µj(cid:0)2−kx(cid:1) 2−kxγ−2 xγ−2 > cxγ−2 . µ(k) j (x) = µ(k) j (6) (7) By (4) If x belongs to the support of ψ (x) we have µ(k) j (x) > c > 0. 8 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI Since ∇Φ(cid:0)2−kx(cid:1) = 2−k(γ−1)∇Φk (x) The phase in the integral Ik (ξ, s) is stationary when by (6) all the derivatives of ∇Φ(cid:0)2−kx(cid:1) are uniformly bounded. By (6) there exits c2 > 0 such that ∇Φk (x) 6 c2 ξ > c2 s we have ∇ (ξ · x + sΦk (x)) = ξ + s∇Φk (x) = 0. ξ + s∇Φk (x) > ξ − s∇Φk (x) > 1 2 ξ . 2 for every k. It follows that for Integrating by parts M times gives (see e.g. Proposition 4, p. 341, in [28]) Ik (ξ, s) 6 cM (1 + ξ)−M . Let now ξ 6 c1 s where c1 is a constant which will be determined later on. Let us consider the function with t ∈ [0, 1]. Then, for t ∈ (0, 1] F (t) = ∇Φk (tx) · x F ′ (t) = xT Hess Φk (tx) x and by (7) the eigenvalues of Hess Φk (tx) are bounded from below by tγ−2 xγ−2. Then F (1) >Z 1 0 xT Hess Φk (tx) xdt > cZ 1 ∇Φk (x) > ∇Φk (x) · x 0 x and therefore It follows that tγ−2 xγ−2 x2 dt > cxγ > cxγ−1 > 2c1 > 0. Integrating by parts M times gives ξ + s∇Φk (x) > s∇Φk (x) − ξ > c1 s . Ik (ξ, s) 6 cM (1 + s)−M . Finally, for every (ξ, s), Theorem 1, p. 348, in [28] gives Ik (ξ, s) 6 c (1 + ξ + s)− d−1 2 . (cid:3) Proposition 11. Let γ > 1 and let B be a bounded convex body in Rd with every- where positive Gaussian curvature with the exception of a single flat point of order γ. Let Θ be the outward unit normal to ∂B at the flat point and for every ζ ∈ Rd write ζ = ξ + sΘ, with s = ζ · Θ and ξ · Θ = 0. Then, if 1 < γ 6 2 If γ > 2 the following three upper bounds hold: 2 . bχB (ζ) 6 cζ− d+1 cs−1− d−1 cξ−(d−1) γ−2 cξ− d+1 , . γ 2 2(γ−1) s− d−1 2(γ−1) −1 , bχB (ζ) 6 The particular case where the boundary in a neighborhood of the flat point has equation t = xγ with γ > 2 and d = 2 has been already considered in [6]. The same case with γ > 2 and d > 2 has been considered in [4], but we acknowledge that the proof of the rate of decay in the horizontal directions was not correctly justified. (8) (9) DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 9 Proof. Choose a smooth function η (z) supported in a neighborhood of the flat point and such that η (z) = 1 in a smaller neighborhood. For every z ∈ ∂B let ν (z) be its outward unit normal. Applying the divergence theorem we decompose the Fourier transform as bχB (ζ) =ZB = −1 e−2πiζ·zdz = −1 4π2 ζ2Z∂B ∇(cid:0)e−2πiζ·z(cid:1) · ν (z) dσ (z) ζ · ν (z) e−2πiζ·zdσ (z) 2πi ζ2Z∂B 2πi ζ2Z∂B 2πi ζ2Z∂B = −1 + −1 = K1 (ζ) + K2 (ζ) . ζ · ν (z) e−2πiζ·zη (z) dσ (z) ζ · ν (z) e−2πiζ·z [1 − η (z)] dσ (z) (10) Since in the support of the function 1 − η (z) the Gaussian curvature is bounded away from zero, the method of stationary phase gives the classical estimate (see Theorem 1, p. 348, in [28]) K2 (ζ) 6 cζ− d+1 2 . (11) By a suitable choice of coordinates we can assume that z = (x, t), the flat point is the point (0, 0), its outward normal is (0,−1) and that the relevant part of the surface ∂B is described by the equation t = Φ (x) with Φ ∈ Sγ. Hence Write ϕ (x) = η (x, Φ (x)) and ψ (x) = ϕ (x) − ϕ (2x) so that for every x 6= 0 ν (x, Φ (x)) = . (∇Φ (x) ,−1) q1 + ∇Φ (x)2 +∞Xk=0 ψ(cid:0)2kx(cid:1) . ϕ (x) = Observe that ϕ (x) is smooth and a suitable choice of η (z) guarantees ϕ (x) = 1 if x 6 ε/2 and ϕ (x) = 0 if x > ε for some ε > 0. With the above choice of coordinate we can also write ζ = (ξ,−s) so that, following the notation of the previous lemma, we have K1 (ζ) = −1 ζ · = −ζ (∇Φ (x) ,−1) e−2πiζ·(x,Φ(x))ϕ (x)q1 + ∇Φ (x)2dx q1 + ∇Φ (x)2 (∇Φ (x) ,−1) e−2πiζ·(x,Φ(x))ϕ (x) dx (∇Φ (x) ,−1) e−2πiζ·(x,Φ(x))ψ(cid:0)2kx(cid:1) dx 2πi ζ2ZRd−1 2πi ζ2 ·ZRd−1 +∞Xk=0ZRd−1 = −ζ 2πi ζ2 · +∞Xk=0 ·ZRd−1(cid:0)∇Φ(cid:0)2−ky(cid:1) ,−1(cid:1) e−2πi(2−kξ,−2−kγ s)·(y,2kγ Φ(2−ky))ψ (y) dy +∞Xk=0 2πi ζ2 · Ik(cid:0)2−kξ,−2−kγs(cid:1) . −2−k(d−1)ζ 2πi ζ2 = 2−k(d−1) −ζ = (12) 10 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI By the previous lemma Hence, K1 (ζ) 6 2 . c (cid:12)(cid:12)Ik(cid:0)2−kξ,−2−kγs(cid:1)(cid:12)(cid:12) 6 c(cid:0)1 +(cid:12)(cid:12)2−kγs(cid:12)(cid:12) +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)− d−1 2−k(d−1)(cid:0)1 +(cid:12)(cid:12)2−kγs(cid:12)(cid:12) +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)− d−1 2−k(d−1)(cid:12)(cid:12)2−kξ(cid:12)(cid:12)− d−1 +∞Xk=0 +∞Xk=0 2 6 cξ− d+1 K1 (ζ) 6 (ξ, s) c ξ . 2 2 . In particular, for every (ξ, s) ∈ Rd we have Assume now γ 6= 2. Our second estimate for K1 (ζ) is as follows. For every (ξ, s) ∈ Rd we have K1 (ζ) 6 2 c 2−k(d−1) + 2−k(d−1)(cid:0)1 +(cid:12)(cid:12)2−kγs(cid:12)(cid:12)(cid:1)− d−1 s X2k<s1/γ 2 X2k<s1/γ γ + cs−1− d−1 6 c s c +∞Xk=0 s X2k>s1/γ 6 cs−1− d−1 6( cs− d+1 cs−1− d−1 γ 2 γ < 2, γ > 2. 2−k(d−1)(cid:12)(cid:12)2−kγs(cid:12)(cid:12)− d−1 2 2k(d−1)( γ 2 −1) Note that when γ = 2 the previous computation gives cs− d+1 log (2 + s). How- ever when Φ (y) is smooth it is well known that the correct estimate is cs− d+1 2 . With a more careful analysis we show that this is the case also in our setting, even if we do not assume smoothness at the flat point. Indeed notice that condition (5) allows higher derivatives to blow up at the flat point. Let γ = 2. Then 2 K1 (ζ) 6 (ξ, s) By the previous lemma we have so that (cid:12)(cid:12)Ik(cid:0)2−kξ,−2−2ks(cid:1)(cid:12)(cid:12) 6 +∞Xk=0 2−k(d−1)(cid:12)(cid:12)Ik(cid:0)2−kξ,−2−2ks(cid:1)(cid:12)(cid:12) 6 c X2k6c1s/ξ Xc1s/ξ<2k<c2s/ξ + c = S1 + S2 + S3. (13) c 2−k(d−1)(cid:12)(cid:12)Ik(cid:0)2−kξ,−2−kγs(cid:1)(cid:12)(cid:12) +∞Xk=0 c(cid:0)1 +(cid:12)(cid:12)2−2ks(cid:12)(cid:12) +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)− d−1 cM(cid:0)1 +(cid:12)(cid:12)2−2ks(cid:12)(cid:12)(cid:1)−M cM(cid:0)1 +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)−M 2 for every (ξ, s) if 2k 6 c1 if c2 s ξ , 6 2k. s ξ 2−k(d−1)(cid:0)1 +(cid:12)(cid:12)2−2ks(cid:12)(cid:12)(cid:1)−M + c Xc2s/ξ62k 2−k(d−1)(cid:0)1 +(cid:12)(cid:12)2−2ks(cid:12)(cid:12) +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)− d−1 2 2−k(d−1)(cid:0)1 +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)−M DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 11 We have 2−k(d−1)(cid:0)1 +(cid:12)(cid:12)2−2ks(cid:12)(cid:12)(cid:1)−M 2−k(d−1) + cs−M Xs1/2>2k 22kM 2−k(d−1) S1 6 c +∞Xk=0 6 c Xs1/262k 6 cs− d−1 S2 6 c Xc2s/ξ62k 2 and Xmax(c2s/ξ,ξ)62k 6 c Observe that Xmax(c2s/ξ,ξ)62k Also, the series is non void only if c2 cξ−M Xc2s/ξ62k<ξ 2Mk−k(d−1). 2 , s d−1 ξd−1! 6 cs− d−1 2 . 2−k(d−1)(cid:0)1 +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)−M 2−k(d−1) + cξ−M Xc2s/ξ62k<ξ ,ξ−(d−1)! d−1 s(cid:19)d−1 2 min ξd−1 2−k(d−1) 6 c min (cid:18)ξ 6 cs− d−1 ξ−M Xc2s/ξ62k<ξ 2Mk−k(d−1) 6 cξ−M Xξ>2k s 2 2Mk−k(d−1) s ξ < ξ, that is cs1/2 < ξ. In this case we have 2Mk−k(d−1) 6 cξ−(d−1) 2 . 6 cs− d−1 Xc1s/ξ<2k<c2s/ξ bχB (ζ) 6 cs− d+1 2 We now turm to S3 that contains a sum with a finite number of terms. We have, S3 6 cs− d−1 2 1 6 cs− d−1 2 . Substituting into (13) gives the estimate when γ = 2. It remains to prove the second row in (9). We have K1 (ζ) 6 6 c s c s 2 +∞Xk=0 2−k(d−1)(cid:0)(cid:12)(cid:12)2−kγs(cid:12)(cid:12) +(cid:12)(cid:12)2−kξ(cid:12)(cid:12)(cid:1)− d−1 s− d−1 2k(d−1)( γ X2k 6(s/ξ) s− d−1 2(γ−1) −1 . γ−1 1 2 6 cξ− (d−1)(γ−2) 2(γ−1) 2 −1) + ξ− d−1 2 X2k>(s/ξ) 1 γ−1 2−k d−1 2  (cid:3) 12 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI Remark 12. Let z ∈ ∂B, let Tz be the tangent hyperplane to ∂B in z and let S (z, δ) = {w ∈ ∂B : dist (w, Tz) < δ} . In [7] it is proved that when the boundary of B is smooth and of finite type (every one dimensional tangent line to ∂B makes finite order of contact with ∂B), then bχB (ζ) 6 cζ−1hσ(cid:16)S(cid:16)z+,ζ−1(cid:17)(cid:17) + σ(cid:16)S(cid:16)z−,ζ−1(cid:17)(cid:17)i where z+ and z− are the two points on ∂B with outer normal parallel to ζ and σ is the surface measure. In our case ∂B is not necessarily smooth, but the above result in fact holds. Indeed, let B as in Proposition 11, and choose coordinates such that z = (x, t), the flat point is the point (0, 0), its outward normal is (0,−1) and that the relevant part of the surface ∂B is described by the equation t = Φ (x) with Φ ∈ Sγ. Fix z = (x, t) ∈ ∂B. Elementary geometric observations lead to σ (S (z, δ)) > cσ (S (0, c1δ)) ≈(cid:16)δ1/γ(cid:17)d−1 σ (S (z, δ)) > c(cid:18)δ(cid:16)xγ−2(cid:17)−1(cid:19) d−1 2 for δ 6 cxγ . for δ > cxγ , (14) (15) . It follows that for a given The unit normal to ∂B in (x, Φ (x)) is ζ = (ξ, s) the point (x, Φ (x)) in ∂B with normal in the direcion ζ satisfies ξ / s = ∇Φ (x) ≈ xγ−1. √∇Φ(x)2+1 a) If s > ξγ then s1−γ > (ξ / s)γ ≈ x(γ−1)γ. Hence (∇Φ(x),−1) 1 ζ ≈ 1 s > cxγ so that, by (14), σ(cid:16)S(cid:16)z,ζ−1(cid:17)(cid:17) > cσ(cid:16)S(cid:16)0, c1 ζ−1(cid:17)(cid:17) ≈ ζ−(d−1)/γ ≈ s−(d−1)/γ . b) If ξ 6 s 6 ξγ, then as before 1 and by (15) σ(cid:16)S(cid:16)z,ζ−1(cid:17)(cid:17) > c 1 c) If ξ > s, then x ≈ 1and by (15) The geometric estimate 1 s 2 6 cxγ γ−1! d−1 ζ ≈ s(cid:19) 2−γ s(cid:18)ξ σ(cid:16)S(cid:16)z,ζ−1(cid:17)(cid:17) > ξ− d−1 bχB (ζ) 6 cζ−1 σ(cid:16)S(cid:16)z,ζ−1(cid:17)(cid:17) . . 2 = cξ−(d−1) γ−2 2(γ−1) s− d−1 2(γ−1) now follows from Proposition 11. As said before, the above proposition is the main ingredient in the estimate of the discrepancy associated to the convex body B. In particular it follows that the directions where the Fourier transform has the slowest rate of decay play a relevant role in the estimates of the discrepancy. Actually the Fourier transform in a given direction depends on the two points in ∂B have normals in that direction. The interplay between the contribution of these points is exploited in the following proposition. DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 13 Proposition 13. Let B be a bounded convex body in Rd. Assume that ∂B is smooth with everywhere positive Gaussian curvature except at most at two points P and Q which are flat of order γP and γQ respectively and have outward unit normals −Θ and Θ. Let S (t) = {z ∈ B : z · Θ = t} be (d − 1)-dimensional measures of the slices of B that are orthogonal to Θ. The function S (t) is supported in P · Θ 6 t 6 Q · Θ and is known to be smooth in P · Θ < t < Q· Θ. Assume that there exist two smooth functions GP (r) and GQ (r) with GP (0) 6= 0 and GQ (0) 6= 0 such that, for u > 0 sufficiently small and Then, as s → +∞, d−1 d−1 γP GP(cid:16)u1/γP(cid:17) γQ GQ(cid:16)u1/γQ(cid:17) . + 1(cid:17) + 1(cid:17) 2 (cid:16) d−1 i π γQ 2 (cid:16) d−1 γP e−i π +1 +1 e S (P · Θ + u) = u S (Q · Θ − u) = u γP (2π) Γ(cid:16) d−1 Γ(cid:16) d−1 max(γP ,γQ)! . (2π) γQ d d−1 γP d−1 γQ bχB (sΘ) = e−2πisΘ·P GP (0) + e−2πisΘ·QGQ (0) + O s −1− +1(cid:17) sgn(s) s−1− d−1 γP +1(cid:17) sgn(s) −1− d−1 γQ s Observe that when in a neighborhood of the points P and Q the boundary of B is smooth with positive Gaussian curvature K (P ) and K (Q) then we have γP = γQ = γ = 2, GP (0) = (2π)(d−1)/2 K −1/2 (Q). Γ( d+1 2 ) Hence we obtain the classical formula K −1/2 (P ), and GQ (0) = (2π)(d−1)/2 Γ( d+1 2 ) bχB (sΘ) = + 1 2π 1 2π e−2πisΘ·P K −1/2 (P ) e−i π 2 ( d−1 e−2πisΘ·QK −1/2 (Q) ei π 2 ( d−1 2 2 +1) sgn(s) s− d+1 2 +1) sgn(s) s− d+1 2 2 (cid:17) . + O(cid:16)s− d+2 See [16] and [17]. See also [18, Corollary 7.7.15]. Proof. Let η (t) be a smooth cutoff function with η (t) = 1 if t 6 ε and η (t) = 0 if t > 2ε with ε small. Since S (t) is smooth inside P · Θ < t < Q · Θ, see [1], for every N > 0 we have χB (z) e−2πiz·sΘdz =Z +∞ −∞ Z{z·Θ=t} χB (z) dz! e−2πistdt S (t) e−2πistdt P ·Θ bχB (sΘ) =ZRd =Z Q·Θ =Z +∞ +Z +∞ = e−2πisP ·ΘZ +∞ 0 0 0 η (u) S (P · Θ + u) e−2πis(P ·Θ+u)du η (u) S (Q · Θ − u) e−2πis(Q·Θ−u)du + O(cid:16)s−N(cid:17) u d−1 γP GP(cid:16)u1/γP(cid:17) η (u) e−2πisudu 14 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI It is enough to consider Since G (r) is smooth, for every N > 0 we can write the Taylor expansion 0 u + e−2πisQ·ΘZ +∞ K (s) =Z +∞ N −1Xk=0 +Z +∞ K (s) = u 0 0 G(k) (0) k! d−1+N (cid:12)(cid:12)(cid:12)(cid:12)Z +∞ 0 η (u) GN(cid:16)u1/γ(cid:17) u Z +∞ 0 d−1 u γ u d−1 d−1+k γQ GQ(cid:16)u1/γQ(cid:17) η (u) e2πisudu + O(cid:16)s−N(cid:17) . γ G(cid:16)u1/γ(cid:17) η (u) e−2πisudu. Z +∞ γ GN(cid:16)u1/γ(cid:17) η (u) e−2πisudu. γ GN(cid:0)u1/γ(cid:1) η (u) has enough bounded e−2πisudu(cid:12)(cid:12)(cid:12)(cid:12) 6 cs−1− d η (u) e−2πisudu d−1+N d−1+N γ . 0 γ η (u) uαe−2πisudu For N large enough, the function u derivatives so that a repeated integration by parts gives Finally, all other terms in the above sum have the form and can be estimated by the following lemma. (cid:3) Lemma 14. If η is as above then, for every α > −1 and s 6= 0, we have Z +∞ 0 tαe−2πistη (t) dt = Γ (α + 1) (2π s)α+1 e−i π 2 (α+1) sgn(s) + O(cid:16)s−N(cid:17) . The above result is not surprising since, in the sense of distributions, Z +∞ 0 tαe−2πistdt = 2 (α+1) sgn(s). Γ (α + 1) (2π s)α+1 e−i π See e.g. [13]. The following is a direct proof. Proof. Assume first s > 0. An integration by parts gives e−2πist d dt tαe−2πistη (t) dt = 1 2πisZ +∞ 0 [tαη (t)] dt e−2πisttα−1η (t) dt + e−2πisttαη′ (t) dt. Z +∞ 2πisZ +∞ 0 α 0 = 1 2πisZ +∞ 0 Since supp η′ ⊂ (ε, 2ε) the term tαη′ (t) is smooth so that 1 2πisZ +∞ 0 e−2πisttαη′ (t) dt = O(cid:16)s−N(cid:17) . Repeating the integration by parts k times, with k 6 α gives Z +∞ 0 Z +∞ 0 tαe−2πistη (t) dt α (α − 1)··· (α − k + 1) = (2πis)k tαe−2πistη (t) dt = 0 Z +∞ (2πis)αZ +∞ e−2πisttα−kη (t) dt + O(cid:16)s−N(cid:17) . e−2πistη (t) dt + O(cid:16)s−N(cid:17) α! 0 Assume first that α is an integer and take k = α. Then DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 15 = = α! (2πis)α+1 + α! (2πis)α+1Z +∞ 0 α! (2πis)α+1 + O(cid:16)s−N(cid:17) . If α is not an integer we take k = [α] + 1. Then e−2πistη′ (t) dt Z +∞ 0 tαe−2πistη (t) dt α (α − 1)··· (α − [α]) = (2πis)[α]+1 Z +∞ 0 e−2πisttλ−1η (t) dt + O(cid:16)s−N(cid:17) , where λ = α − [α]. By [11, (4) pag. 48] we have e2πisttλ−1η (t) dt e−πi(n+λ−2)/2η(n) (0) (2πs)−n−λ + O(cid:0)s−N(cid:1) and since η(n) (0) = 0 for every n > 0 and η (0) = 1 we obtain = − 0 0 Γ (n + λ) Z +∞ e−2πisttλ−1η (t) dt =Z +∞ N −1Xn=0 Z +∞ tαe−2πistη (t) dt n! 0 α (α − 1)··· (α − [α]) (2πis)[α]+1 Z +∞ 2 i(α+1) + O(cid:0)s−N(cid:1) 0 = = Γ (α + 1) (2πs)α+1 e− π e−2πisttλ−1η (t) dt + O(cid:16)s−N(cid:17) also in this case. Let now s < 0. Then Z +∞ 0 tαe−2πistη (t) dt =Z +∞ tαe−2πi(−s)tη (t) dt = 0 Γ (α + 1) (2π s)α+1 ei π 2 (α+1) + O(cid:16)s−N(cid:17) (cid:3) In the next proposition we show that assumptions of Proposition 13 are satisfied when the flat points are as in Proposition 3. Proposition 15. Let γ > 1 and let B be a bounded convex body in Rd. Let U be a bounded open neighborhood of the origin in Rd−1 and let H (x) ∈ C∞ (U ) such that H (0) = 0, ∇H (0) = 0 and Hess H (0) positive definite (see Proposition 3). Assume there exists a neighborhood of the origin W ⊂ Rd such that, in suitable coordinates, ∂B ∩ W =n(x, t) ∈ Rd : t = (H (x))γ/2o ∩ W. As before, let Then, there exists a smooth function G (r) such that for t > 0 sufficiently small we have S (t) =(cid:12)(cid:12)(cid:8)x ∈ Rd−1 : (x, t) ∈ B(cid:9)(cid:12)(cid:12) . S (t) = t d−1 γ G(cid:16)t1/γ(cid:17) 16 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI with G (0) equal to the (d − 1)-dimensional measure of the ellipsoid Proof. For t small enough we have x ∈ Rd−1 : 1 2 d−1Xj,k=1 ∂2H ∂xj∂xk  (0) xjxk 6 1 . S (t) =Z{x∈Rd−1:t>(H(x))γ/2} dx. By Morse's lemma (see [28, p. 346]), there exists a diffeomorphism Ψ (y) between two small neighborhoods of the origin in Rd−1 such that Then, where H (Ψ (y)) = y2 . S (t) =Z{x∈Rd−1:t>(H(x))γ/2} dx =Z{y6t1/γ} d−1 = t JΦ(cid:16)t1/γu(cid:17) du = t γ Z{u61} G (r) =Z{z61} JΦ (ru) du. JΨ (y) dy d−1 γ G(cid:16)t1/γ(cid:17) Finally observe that G (0) = lim JΦ (ru) du = lim r→0 1 rd−1Z{w6r} JΦ (w) dw 1 r→0Z{u61} rd−1Z{w26r2} rd−1Z{H(y)6r2} d−1Xj,k=1 x ∈ Rd−1 : 1 2 1 = lim r→0 = lim r→0  =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) JΦ (w) dw dy = lim ∂2H ∂xj∂xk dx r2 61} r→0Z{ H(rx) (0) xjxk 6 1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . (cid:3) 4. Proofs of the results Proof of Proposition 3. Let α be a multi-index. It is not difficult to prove by in- duction on α that ∂αΦ ∂xα (x) is a finite sum of terms of the form c [H (x)]γ/2−k ∂β1H ∂xβ1 (x) × ··· × ∂βkH ∂xβk (x) with k 6 α and multi-indices β1, . . . , βk such that β1 + ··· + βk = α. Since Hess H (0) is positive definite there are positive constants c1 and c2 such that in a neighborhood of the origin and c1 x2 6 H (x) 6 c2 x2 ∂H ∂xj (cid:12)(cid:12)(cid:12)(cid:12) (x)(cid:12)(cid:12)(cid:12)(cid:12) 6 c2 x . DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 17 Moreover, since H (x) is smooth It follows that ∂βj H ∂xβj (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)[H (x)]γ/2−k ∂β1H 6 c(cid:16)x2(cid:17)γ/2−k 6 cxγ−α . (x)(cid:12)(cid:12)(cid:12)(cid:12) 6 cxmax(2−βj ,0) 6 cx2−βj . (x) × ··· × ∂βkH ∂xβk (x)(cid:12)(cid:12)(cid:12)(cid:12) ∂xβ1 x2−β1 × ··· × x2−βk 6 cxγ−2k x2k−(β1+···+βk) This proves (5). To prove (4) let us write ∂2Φ ∂xj∂xk (x) = = γ 2 γ 2 so that (γ/2 − 1) [H (x)]γ/2−2 ∂H [H (x)]γ/2−1 ∂2H ∂xj∂xk ∂xj (x) + (γ/2 − 1) (x) ∂H ∂xk (x) + γ 2 ∂H ∂xj [H (x)]γ/2−1 ∂2H ∂xj∂xk (x) (x) (x) ∂H ∂xk H (x) ! Hess Φ (x) = γ 2 [H (x)]γ/2−1 Hess M (x) where M (x) is the matrix with entries ∂2H ∂xj∂xk (x) + (γ/2 − 1) ∂H ∂xj (x) ∂H ∂xk H (x) (x) . Let A = Hess H (0), since H (x) = 1 2 xT Ax + O(cid:16)x3(cid:17) , ∇H (x) = Ax + O(cid:16)x2(cid:17) Hess H (x) = A + O (x) we have M (x) = A + O (x) + (γ/2 − 1)(cid:16)Ax + O(cid:16)x2(cid:17)(cid:17)(cid:16)Ax + O(cid:16)x2(cid:17)(cid:17)T 1 2 xT Ax + O(cid:16)x3(cid:17) = A + (γ − 2) Let us show that the matrix Ax (Ax)T xT Ax + O (x) . A + (γ − 2) Ax (Ax)T xT Ax is positive definite. Indeed, for all y ∈ Rd−1 we have yT A + (γ − 2) Ax (Ax)T xT Ax ! y = yT Ay + (γ − 2)(cid:0)yT Ax(cid:1)2 xT Ax . When γ > 2 we easily obtain yT A + (γ − 2) Ax (Ax)T xT Ax ! y > yT Ay > λ1 y2 18 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI where λ1 is the smallest eigenvalue of A. For 1 < γ < 2, by Cauchy-Schwarz inequality for the inner product defined by hy, xi = yT Ax we have Hence (cid:0)yT Ax(cid:1)2 xT Ax xT Ax = yT Ay. 6 (cid:0)yT Ay(cid:1)(cid:0)xT Ax(cid:1) xT Ax ! y > yT Ay + (γ − 2) yT Ay Ax (Ax)T yT A + (γ − 2) = (γ − 1) yT Ay > (γ − 1) λ1 y2 . Let µ1 (x) be the smallest eigenvalue of Hess Φ (x) . We want to show that µ1 (x) > cxγ−2. This is equivalent to show that We have yT Hess Φ (x) y = γ 2 yT Hess Φ (x) y > cxγ−2 y2 . [H (x)]γ/2−1 yT A + (γ − 2) y2 − c2(cid:16)x2(cid:17)γ/2−1 + [H (x)]γ/2−1 yT O (x) y > c1(cid:16)x2(cid:17)γ/2−1 = c1 xγ−2 y2 − c2 xγ−2 x y2 > cxγ−2 y2 xT Ax ! y xy2 Ax (Ax)T for x small enough. (cid:3) To prove the theorems and the corollary it is convenient to introduce a mollified discrepancy. Lemma 16. Let ϕ (z) be a compactly supported smooth function in Rd with integral 1. Then, if the support of ϕ (z) is sufficiently small, for every 0 < ε < 1 and R > 1 we have ε−dϕ(cid:0)ε−1·(cid:1) ∗ χ(R−ε)B (z) 6 χRB (z) 6 ε−dϕ(cid:0)ε−1·(cid:1) ∗ χ(R+ε)B (z) . In particular, B(cid:16)(R − ε)d − Rd(cid:17) + Dε,R−ε (z) 6 DR (z) 6 B(cid:16)(R + ε)d − Rd(cid:17) + Dε,R+ε (z) , where Dε,R (z) = Rd X06=m∈Zdbϕ (εm)bχB (Rm) e2πim·z. The above lemma is well known. See e.g. [3, pag. 195] for a proof. Also the following result is well known, the following is elementary proof. Lemma 17. For every integer M > 0 and every neighborhood U of the origin in Rd there exists a smooth function ϕ (z) supported in U such that and for every multi-index α, with 0 < α 6 M bϕ (0) = 1 ∂αbϕ ∂ζα (0) = 0. We want to find constants c0, c1, . . . , cM such that the function satisfies the lemma. We have so that and for every multi-index α ZRd ψ (z) dz = 1. ck ϕ (z) = MXk=0 2kdckψ(cid:0)2kz(cid:1) MXk=0 ckbψ(cid:0)2−kζ(cid:1) bϕ (ζ) = MXk=0 MXk=0 ckbψ (0) = bϕ (0) = MXk=0 ck2−kα ∂αbψ ∂αbϕ c0 +(cid:0)2−1(cid:1)1 c0 +(cid:0)2−M(cid:1)1 c1 + ··· +(cid:0)2−1(cid:1)M c1 + ··· +(cid:0)2−M(cid:1)M ∂ζα (0) = ∂ζ α (0) . c0 + c2 + ··· + cM = 1 ... (cid:0)2−1(cid:1)0 (cid:0)2−M(cid:1)0  Hence the coefficients ck are the solution of the non singular linear system DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 19 Proof. Let ψ (z) be a smooth function supported in U such that cM = 0 cM = 0 (cid:3) The following lemma collects the main estimates that we will use later. Lemma 18. Assume the inequalities bχB (ζ) 6 γ , cs−1− d−1 cξ−(d−1) γ−2 cξ− d+1 2 . 2(γ−1) s− d−1 2(γ−1) −1 , proved in Proposition 11, where ζ = ξ + sΘ, with s = ζ · Θ and ξ · Θ=0 for some Θ ∈ Rd with Θ = 1 and γ > 2 and let ϕ (z) as in the previous lemma. 1) For every τ > 0 and p > 2d/ (d − 1) there exists c such that for every ε > 0 and R > 1, 2) For every τ > 0 there exists c such that for every ε > 0, R > 1 and z ∈ Td, 1/p p Rd Rd Xm−(m·Θ)Θ>τ bϕ (εm)bχB (Rm) e2πim·z(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZTd(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dz (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 06=m∈Zd, m−(m·Θ)Θ<τ bϕ (εm)bχB (Rm) e2πim·z(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) bϕ (ζ) 6 cM (1 + ζ)−M . X 6 cR d−1 2 ε− d−1 2 + d p . 6 cR(d−1)(1− 1 γ ). Proof. Let us prove 1). For every m ∈ Zd, write m = m1 + m2 with m1 = m− (m · Θ) Θ and m2 = (m · Θ) Θ. Also observe that for every M > 0, 20 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI Since p > 2, by the Hausdorff-Young inequality with 1/p + 1/q = 1 and the as- p q/p dz sumption on bχB (ζ) we have Rd Xm1>τ bϕ (εm)bχB (Rm) e2πim·z(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZTd(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 6 Rdq Xm1>τ bϕ (εm)q bχB (Rm)q 2 Xτ 6m16m2 Xm1>max(m2,τ ) 6 cRq d−1 + cRq d−1 2 = A + B. (1 + ε m1)−M m1− d+1 2 q (1 + ε m2)−M m1−q(d−1) γ−2 2(γ−1) m2−q d−1 2(γ−1) −q Since in the series in A the quantities m1 and m2 are bounded away from zero we can control the series with an integral, Xτ <m16m2 6 cZZ{τ <ξ6s} 6 cZ +∞ 6 cZ +∞ τ τ (1 + ε m2)−M m1−q(d−1) γ−2 2(γ−1) m2−q d−1 2(γ−1) −q (1 + ε s)−M ξ−q(d−1) γ−2 2(γ−1) s−q d−1 2(γ−1) −q dξds 2(γ−1) −q"Z{ξ6s} ξ−q(d−1) γ−2 2(γ−1) dξ# ds (1 + ε s)−M s−q d−1 (1 + ε s)−M sd−1−q d+1 2 ds 6 cε−(d−q d+1 2 ) (note that since p > 2d/ (d − 1) we have q < 2d/ (d + 1)). Similarly, for the series in B, This proves point 1) in the statement. Similarly, to prove point 2) observe that, by (1 + ξ)−M ξ1− d+1 2 q dξ = cε d+1 2 q−d. (1 + ε m1)−M m1− d+1 2 q (1 + ε ξ)−M ξ− d+1 2 q dξds d+1 = cε Xm1>max(m2,τ ) 6 cZZ{ξ>s} 2 q−dZRd−1 the assumption on bχB (ζ), we have (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Rd X06=m∈Zd,m1<τ bϕ (εm)bχB (Rm) e2πim·z(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 6 Rd X06=m∈Zd,m1<τ 6 cR(d−1)(1− 1 bχB (Rm) γ ) X06=m∈Zd,m1<τ m2−1− d−1 γ . Note that the last series is essentially one dimensional and it is convergent. (cid:3) Proof of Theorem 4. The discrepancy will be estimated using the size of bχB (ζ). Since the main contribution to the size of this Fourier transform comes from the flat points on ∂B and since with a suitable partition of unity we can isolate such flat points, without loss of generality we can assume the existence of only a single flat point of order γ. DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 21 The case 1 < γ 6 2 follows from the argument used in [2] for the smooth case. This essentially reduces to the Hausdorff-Young inequality and follows from the estimate 2 bχB (ζ) 6 cζ− d+1 that holds true also in our case by (8). Let us now prove point 2) and point 3) in the theorem. To prove point 2) we observe that the case p 6 2d/ (d + 1 − γ) follows from the case p = 2d/ (d + 1 − γ), and the case 2d/ (d + 1 − γ) 6 p 6 +∞ follows by interpolation between p = 2d/ (d + 1 − γ) and p = +∞. Hence to prove point 2) it suffices to consider only the cases p = 2d/ (d + 1 − γ) and p = +∞. Similarly to prove point 3) it suffices to consider only the case p = +∞. Observe that since γ > 2, all these values of p are greater than 2d/ (d − 1). By Lemma 16 we have Replacing R ± ε with R for simplicity, Lemma 18, with a fixed τ > 0, gives 6 cRd−1ε + max ± kDε,R±εkLp(Td) ± kDε,R±εkLp(Td) . kDRkLp(Td) 6 B max ± (cid:12)(cid:12)(cid:12)(R ± ε)d − Rd(cid:12)(cid:12)(cid:12) + max kDε,RkLp(Td) =ZTd(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Rd X06=m∈Zdbϕ (εm)bχB (Rm) e2πim·z(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dz Rd Xm−(m·Θ)Θ>τ bϕ (εm)bχB (Rm) e2πim·z(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 6ZTd(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dz +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 06=m∈Zd, m−(m·Θ)Θ<τ bϕ (εm)bχB (Rm) e2πim·z(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) X γ ) p + cR(d−1)(1− 1 2 ε− d−1 2 + d 6 cR p 1/p p 1/p Rd d−1 The choice ε = R− d−1 d+1−2d/p then gives kDRkLp(Td) 6 cRd(d−1) p−2 p−2d+dp + cR(d−1)(1− 1 γ ). For p = 2d/ (d + 1 − γ) and 2 < γ 6 d + 1, or p = +∞ and γ > d + 1 we obtain γ ). kDRkLp(Td) 6 cR(d−1)(1− 1 For 2 < γ < d + 1 and p = +∞ we obtain kDRkL∞(Td) 6 cR d(d−1) d+1 . (cid:3) Proof of Theorem 5. The idea of the proof is simple. For every m ∈ Zd, write m = m1 + m2 with m1 = m− (m · Θ) Θ and m2 = (m · Θ) Θ, and split the Fourier expansion of the discrepancy as DR (z) = Rd Xm1=0, m26=0bχB (Rm) e2πim·z + Rd Xm16=0bχB (Rm) e2πim·z. We will see that the main term is the first one and it follows from Proposition 13 that (16) Rd Xm1=0, m26=0bχB (Rm) e2πim·z 22 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI ∼ R(d−1)(1−1/γP )AP (z−RP ) + R(d−1)(1−1/γQ)AQ (z−RQ) . The details are as follows. Let Dε,R (z) be the mollified discrepancy as in the proof of Theorem 4 and let Y (z,R) = R(d−1)(1−1/γP )AP (z−RP ) + R(d−1)(1−1/γQ)AQ (z−RQ) . From Lemma 16 with a cut-off function as in Lemma 17 we have DR (z) − Y (z,R) 6 B max + Y (z,R ± ε) − Y (z,R) ± (cid:12)(cid:12)(cid:12)(R ± ε)d − Rd(cid:12)(cid:12)(cid:12) + max ± Dε,R±ε (z) − Y (z,R ± ε) (17) The first term in the right-hand side is bounded by cRd−1ε. For the third term we have Y (z,R ± ε) − Y (z,R) The two terms are similar, let us consider only the first one. Then 6 (R ± ε)(d−1)(1−1/γP ) AP (z − (R ± ε) P ) − AP (z − RP ) 6(cid:12)(cid:12)(cid:12)(R ± ε)(d−1)(1−1/γP ) AP (z − (R ± ε) P ) − R(d−1)(1−1/γP )AP (z − RP )(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(R ± ε)(d−1)(1−1/γQ) AQ (z − (R ± ε) Q) − R(d−1)(1−1/γQ)AQ (z − RQ)(cid:12)(cid:12)(cid:12) . (cid:12)(cid:12)(cid:12)(R ± ε)(d−1)(1−1/γP ) AP (z − (R ± ε) P ) − R(d−1)(1−1/γP )AP (z − RP )(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(R ± ε)(d−1)(1−1/γP ) − R(d−1)(1−1/γP )(cid:12)(cid:12)(cid:12) AP (z − RP ) +∞Xk=1 ×(cid:12)(cid:12)(cid:12)(cid:12)sin(cid:18)2πkm0 · (z − (R ± ε) P ) − +∞Xk=1 γP (cid:19) − sin(cid:18)2πkm0 · (z − RP ) − Since AP (z − (R ± ε) P ) − AP (z − RP ) γP sin (επkm0 · P ) k−1− d−1 k−1− d−1 6 c d − 1 π 2 6 c γP π 2 d − 1 γP (cid:19)(cid:12)(cid:12)(cid:12)(cid:12) d−1 γP . 6 cε and AP (z − RP ) 6 c we have (cid:12)(cid:12)(cid:12)(R ± ε)(d−1)(1−1/γP ) AP (z − (R ± ε) P ) − R(d−1)(1−1/γP )AP (z − RP )(cid:12)(cid:12)(cid:12) γP + R(d−1)(1−1/γP )−1ε. 6 cR(d−1)(1−1/γP )ε d−1 It remains to estimate the second term in (17) and for simplicity we replace R ± ε with R. We have Dε,R (z) − Y (z, R) 26=0bϕ (εm)bχB (Rm) e2πim·z − Y (z, R) =Rd Xm1=0,m +Rd Xm16=0bϕ (εm)bχB (Rm) e2πim·z (18) DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 23 = I (z) + II (z) . For I (z) we have a pointwise estimate, I (z) 6 RdXs6=0 +(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) bϕ (εsm0) − 1bχB (Rsm0) RdXs6=0bχB (Rsm0) e2πism0·z − Y (z, R)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . Our choice of the function ϕ (z) yields dζα (0) = 0 dαbϕ for every multi-index α with α 6 M . Hence and by Proposition 11 (recall that γP > γQ) RdXs6=0 bϕ (εsm0) − 1bχB (Rsm0) 6 cR(d−1)(cid:16)1− 1 bϕ (ζ) − 1 6 cM ζM , min(cid:16)εM sM , 1(cid:17)s−1− d−1 εM sM s−1− d−1 γP γP (cid:17)Xs6=0 γP (cid:17) Xεs61 6 cR(d−1)(cid:16)1− 1 6 cR(d−1)(cid:16)1− 1 γP (cid:17)ε d−1 γP . γP + cR(d−1)(cid:16)1− 1 γP (cid:17) Xεs>1 s−1− d−1 γP By our assumption on the direction Θ and by Proposition 13 a long but direct computation gives = R(d−1)(1−1/γP )AP (z−RP ) + R(d−1)(1−1/γQ)AQ (z−RQ) Rd Xm1=0,m6=0bχB (Rm) e2πim·z = RdXs6=0bχB (Rsm0) e2πism0·z γP(cid:17) + O(cid:16)Rd−1− d = Y (z, R) + O(cid:16)Rd−1− d γP(cid:17) I (z) 6 cR(d−1)(cid:16)1− 1 γP + Rd−1− d γP (cid:17)ε d−1 Hence, we have the pointwise estimate (19) The assumption that αΘ ∈ Zd for some α implies that the requirement m1 6= 0 is equivalent to m1 > τ for some τ > 0. By Lemma 18 we therefore have γP . kIIkp 6 cR d−1 2 ε− d−1 2 + d p . Collecting the estimates (18), (19) and (20) we have kDR (z) − Y (z, R)kLp(Td) 6 cR(d−1)(1−1/γP )ε d−1 γP + cRd−1ε + cRd−1− d γP + cR (20) d−1 2 ε− d−1 2 +d/p. The choice ε = R− d−1 d+1−2d/p gives kDR (z) − Y (z, R)kLp(Td) 6 cR(d−1)(cid:16)1− 1 γP (cid:17)(cid:16)R− d−1 d+1−2d/p d−1 γP + R d−1 γP − d−1 d+1−2d/p + R− 1 γP(cid:17) . 24 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI Since our assumption implies 1/p > (d + 1 − γP ) / (2d), all the exponents of R in the parenthesis are negative and therefore kDR (z) − Y (z, R)kLp(Td) 6 cR(d−1)(cid:16)1− 1 γP (cid:17)−δ for some δ > 0. This proves immediately point 2). It also prove point 1) as long as one notices that if γP > γQ (cid:13)(cid:13)(cid:13)R(d−1)(1−1/γQ)AQ (z−RQ)(cid:13)(cid:13)(cid:13)Lp(Td) for a suitable δ > 0. 6 cR(d−1)(cid:16)1− 1 γP (cid:17)−δ (cid:3) Proof of Corollary 6. Because of our assumptions the constants in front of the two series that define AP (z−RP ) and AQ (z − RQ) are the same. A simple computa- tion gives AP (z−RP ) + AQ (z − RQ) = 4GP (0) Γ(cid:16) d−1 γ + 1(cid:17) × cos(cid:18)2πkm0 ·(cid:18)z−R (2π m0) d−1 γ +1 k−1− d−1 +∞Xk=1 2 (cid:19)(cid:19) . P + Q γ sin(cid:18)π(cid:18)km0 · R (Q − P ) − d − 1 2γ (cid:19)(cid:19) and since m0 · R (Q − P ) and d−1 AP (z−RP ) + AQ (z − RQ) = 0 2γ are integers. (cid:3) Proof of Theorem 8. By Theorem 1.1 in [5] we have the following estimate for the L2 average decay of the Fourier transform ZSO(d) bχB (Rσm)2 dσ!1/2 6 c (R m)− d+1 2 . Hence, applying the Hausdorff-Young inequality to (1) with 2 6 p < 2d/ (d − 1) and 1/p + 1/q = 1, we obtain ZSO(d)(cid:18)ZTd DR,σ (z)p dz(cid:19)q/p dσ 6 Rdq X06=m∈ZdZSO(d) bχB (Rσm)q dσ 6 Rdq X06=m∈Zd ZSO(d) bχB (Rσm)2 dσ!q/2 2 6 cRq d−1 2 . 6 cRq d−1 2 X06=m∈Zd m−q d+1 In a similar way if 1 6 p 6 2 we have, ZSO(d)(cid:18)ZTd DR,σ (z)p dz(cid:19)2/p dσ 6ZSO(d)ZTd DR,σ (z)2 dzdσ = R2d X06=m∈ZdZSO(d) bχB (Rσm)2 dσ 6 cRd−1 X06=m∈Zd m−(d+1) 6 cRd−1. DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 25 (cid:3) Proof of Theorem 9. Without loss of generality we can assume α 6 β. By Propo- sition 11 we have γ , bχB (m, n) 6 cαm + βn−1− 1 c−βm + αn− γ−2 c−βm + αn− 3 ZTd DR (z)2 dz = R4 X(m,n)6=(0,0) 2 . We have 2(γ−1) αm + βn− 1 2(γ−1) −1 , 6 R4 + R4 + R4 bχB (Rm, Rn)2 X0<−βm+αn<1/2 X X 1/26−βm+αn6αm+βn 0<αm+βn<−βm+αn bχB (Rm, Rn)2 bχB (Rm, Rn)2 bχB (Rm, Rn)2 = I + II + III. Using the above estimate for bχB (Rm, Rn) we have III 6 cR (m, n)−3 6 cR X(m,n)6=(0,0) 0<αm+βn<−βm+αn X (m, n)−3 6 cR. In the term II the quantity −βm + αn and αm + βn are bounded away from zero so that, arguing as in the proof of Lemma 18, we can replace the series with the corresponding integral, II 6 cR X 1/26−βm+αn<αm+βn 6 cRZ{1/26ξ<s} ξ− γ−2 (γ−1) s− 1 (γ−1) −2 dsdξ 6 cR. −βm + αn− γ−2 (γ−1) αm + βn− 1 (γ−1) −2 In the term I observe that −βm + αn < 1/2 implies αm + βn ≈ n. Then γ−1 αm + βn− 1 γ−1 −2 α β − γ−2 γ−1 −βm + αn− γ−2 I 6 cR X−βm+αn<1/2 6 cR X−βm+αn<1/2(cid:13)(cid:13)(cid:13)(cid:13) n− 1 6 cR X−βm+αn<1/2(cid:16)n−1−δ(cid:17)− γ−2 +∞Xn=1 γ−1 −2 6 cR. n(cid:13)(cid:13)(cid:13)(cid:13) n(1+δ) γ−2 γ−1 n− 1 6 cR γ−1 γ−1 −2 γ−1 −2 n− 1 In the last inequality we used the assumption δ < 2/(γ − 2). (cid:3) 26 L. BRANDOLINI, L. COLZANI, B. GARIBOLDI, G. GIGANTE, AND G. TRAVAGLINI References [1] G. Bianchi, The covariogram and Fourier-Laplace transform in C n. Proc. Lond. Math. Soc. 113 (2016), 1 -- 23. [2] L. Brandolini, L. Colzani, G. Gigante, G. Travaglini, Lp and Weak-Lp estimates for the number of integer points in translated domains, Math. Proc. Cambridge Philos. Soc., 159 (2015), 471 -- 480. [3] L. Brandolini, G. Gigante, G. Travaglini, Irregularities of distribution and average decay of Fourier transforms. In A panorama of discrepancy theory, 159 -- 220, Lecture Notes in Math., 2107, Springer 2014 [4] L. Brandolini, A. Greenleaf, G. Travaglini, Lp − Lp′ estimates for overdetermined Radon transforms, Trans. Amer. Math. Soc. 359 (2007), 2559 -- 2575. [5] L. Brandolini, S. Hofmann, A. Iosevich, Sharp rate of average decay of the Fourier transform of a bounded set. Geom. Funct. Anal. 13 (2003), 671 -- 680. [6] L. Brandolini, M. Rigoli, G. Travaglini, Average decay of Fourier transforms and geometry of convex sets, Revista Matematica Iberoamericana 14 (1998), 519 -- 560. [7] J. Bruna, A. Nagel, S. Wainger, Convex hypersurfaces and Fourier transforms, Ann. of Math. 127, (1988), 333 -- 365. [8] Y. Colin de Verdi`ere, Nombre de points entiers dans une famille homoth`etique de domains de Rn. Ann. Sci. Ecole Norm. Sup. 10 (1977), 559 -- 575. [9] F. Dell'Oro, E. Laeng, V. Pata, A quantitative Riemann-Lebesgue lemma with application to equations with memory. Proc. Amer. Math. Soc. 145 (2017), 2909 -- 2915. [10] J. Guo, A note on lattice points in model domains of finite type in Rd. Arch. Math. 108 (2017), 45 -- 53. [11] A. Erd´elyi, Asymptotic expansions. Dover Publications, Inc., New York, 1956 [12] K. Falconer, Fractal geometry. Mathematical foundations and applications, John Wiley & Sons, Ltd., Chichester, 2014. [13] I. M. Gel'fand, G. E. Shilov, Generalized functions. Vol. 1, AMS Chelsea Publishing, Provi- dence, RI, 2016 [14] J. Guo, Lattice points in rotated convex domains, Rev. Mat. Iberoam. 31 (2015), 411 -- 438. [15] J. Guo, Lattice points in large convex planar domains of finite type. Illinois J. Math. 56 (2012), 731 -- 757. [16] C. S. Herz, Fourier transforms related to convex sets. Ann. of Math.75 (1962), 81 -- 92. [17] E. Hlawka, Uber Integrale auf convexen Korpen, I, II, Monatshefte fur Mathematik 54 (1950), 1 -- 36, 81 -- 99. [18] L. Hormander, The analysis of linear partial differential operators I, Springer-Verlag, Berlin, 2003. [19] M. N. Huxley, A fourth power discrepancy mean, Monatshefte fur Mathematik 73 (2014), 231 -- 238. [20] A. Iosevich, E. Sawyer, A. Seeger, Two problems associated with convex finite type domains. Publ. Mat. 46 (2002), 153 -- 177. [21] D. G. Kendall, On the number of lattice points inside a random oval, Quarterly Journal of Mathematics Oxford, 19 (1948), 1 -- 26. [22] E. Kratzel, Mittlere Darstellungen naturlicher Zahlen als Summen von n k-ten Potenzen, Czechoslovak Math. J. 23 (1973), 57 -- 73. [23] L. Parnovski, A. Sobolev, On the Bethe-Sommerfeld conjecture for the polyharmonic opera- tor, Duke Math. J. 107 (2001), 209 -- 238. [24] M. Pinsky, N. Stanton, P. E. Trapa, Fourier series of radial functions in several variables, J. Funct. Anal., 116 (1993), 111 -- 132. [25] A. N. Podkorytov, On the asymptotics of the Fourier transform on a convex curve, Vestnik Leningrad University Mathematics 24 (1991), 57 -- 65. [26] B. Randol, A lattice-point problem. Trans. Amer. Math. Soc. 121 (1966), 257 -- 268. [27] B. Randol, A lattice-point problem. II. Trans. Amer. Math. Soc. 125 (1966), 101 -- 113. [28] E. M. Stein, Harmonic analysis: real-variable methods, orthogonality, and oscillatory inte- grals. Princeton University Press, 1993. DISCREPANCY FOR CONVEX BODIES WITH ISOLATED FLAT POINTS 27 Dipartimento di Ingegneria Gestionale, dell'Informazione e della Produzione, Uni- versit`a degli Studi di Bergamo, Viale Marconi 5, Dalmine BG, Italy E-mail address: [email protected] Dipartimento di Matematica e Applicazioni, Universit`a di Milano-Bicocca, Via Cozzi 55, Milano, Italy E-mail address: [email protected] Dipartimento di Ingegneria Gestionale, dell'Informazione e della Produzione, Uni- versit`a degli Studi di Bergamo, Viale Marconi 5, Dalmine BG, Italy E-mail address: [email protected] Dipartimento di Ingegneria Gestionale, dell'Informazione e della Produzione, Uni- versit`a degli Studi di Bergamo, Viale Marconi 5, Dalmine BG, Italy E-mail address: [email protected] Dipartimento di Matematica e Applicazioni, Universit`a di Milano-Bicocca, Via Cozzi 55, Milano, Italy E-mail address: [email protected]
1511.07167
3
1511
2016-09-08T09:51:40
Sharp Gaussian estimates for Schr\"odinger heat kernels: $L^p$ integrability conditions
[ "math.FA", "math.AP" ]
We give new sufficient conditions for comparability of the fundamental solution of the Schr\"odinger equation $\partial_t=\Delta+V$ with the Gauss-Weierstrass kernel and show that local $L^p$ integrability of $V$ for $p> 1$ is not necessary for the comparability.
math.FA
math
SHARP GAUSSIAN ESTIMATES FOR SCHRÖDINGER HEAT KERNELS: Lp INTEGRABILITY CONDITIONS KRZYSZTOF BOGDAN, JACEK DZIUBAŃSKI, AND KAROL SZCZYPKOWSKI Abstract. We give new sufficient conditions for comparability of the fundamental solution of the Schrödinger equation ∂t = ∆ + V with the Gauss-Weierstrass kernel and show that local Lp integrability of V for p > 1 is not necessary for the comparability. 1. Introduction and Preliminaries Let d = 1, 2, . . .. We consider the Gauss-Weierstrass kernel, g(t, x, y) = (4πt)−d/2e−y−x2/(4t), t > 0, x, y ∈ Rd. It is well known that g is a time-homogeneous probability transition density. For a function V we let G be the Schrödinger perturbation of g by V , i.e., the fundamental solution of ∂t = ∆ + V , determined by the following Duhamel or perturbation formula for t > 0, x, y ∈ Rd, G(t, x, y) = g(t, x, y) +Z t 0 ZRd G(s, x, z)V (z)g(t − s, z, y)dzds. Under appropriate assumptions on V , the definition of G may be given by the Feynmann-Kac formula [5, Section 6], the Trotter formula [27, p. 467], the perturbation series [5], or by means of quadratic forms on L2 spaces [10, Sec- tion 4]. In particular the assumption V ∈ Lp(Rd) with p > d/2 was used by Aronson [2], Zhang [27, Remark 1.1(b)] and by Dziubański and Zienkiewicz [11]. Aizenman and Simon [1, 21] proposed functions V (z) from the Kato class, which contains Lp(Rd) for every p > d/2 [1, Chapter 4], [9, Chapter 3, Example 2]. An enlarged Kato class was used by Voigt [23] in the study of Schrödinger semigroups on L1 [23, Proposition 5.1]. For time-dependent perturbations V (u, z), Zhang [24, 26] introduced the so-called parabolic Kato condition. It was then generalized and employed by Schnaubelt and Voigt [20], Liskevich and Semenov [17], Liskevich, Vogt and Voigt [18], and Gulisas- hvili and van Casteren [13]. We say that G has sharp Gaussian estimates if G is comparable with g, at least in bounded time (see below for details). A sufficient condition for the sharp Gaussian estimates was given by Zhang in [27]. As noted in [27, Remark 1.2(c)], the condition may be stated in terms of the bridges of g. Bogdan, Jakubowski and Hansen [5, Section 6] and Bogdan, Jakubowski and Sydor [6] gave analogous conditions for general transition densities. 2010 Mathematics Subject Classification. Primary 47D06, 47D08; Secondary 35A08, 35B25. Key words and phrases. Schrödinger perturbation, sharp Gaussian estimates. 1 SHARP GAUSSIAN ESTIMATES 2 Given a real-valued Borel measurable function V on Rd we ask if there are c1 ≤ G(t, x, y) g(t, x, y) ≤ c2, numbers 0 < c1 ≤ c2 < ∞ such that the following two-sided bounds hold, (1) t > 0, x, y ∈ Rd. We also ponder a weaker property -- if for a given T ∈ (0,∞), 0 < t ≤ T, x, y ∈ Rd. (2) G(t, x, y) g(t, x, y) ≤ c2 , c1 ≤ We call (1) and (2) sharp Gaussian estimates or bounds, respectively global and local in time. We observe that the inequalities in (1) and (2) are stronger than plain Gaussian estimates: c1 (4πt)−d/2e− y−x2 4tε1 ≤ G(t, x, y) ≤ c2 (4πt)−d/2e− y−x2 where 0 < ε1, c1 ≤ 1 ≤ ε2, c2 < ∞. We note in passing that Berenstein proved the Gaussian estimates for V ∈ Lp with p > d/2 (see [16]), Simon [21, Theorem B.7.1] resolved them for V in the Kato class, Zhang used sub- parabolic Kato class for the same end in [26] and the so-called 4G inequality was used by Bogdan and Szczypkowski in [7]. For further discussion we refer the reader to [17], [18], [19], [27], and the Introduction in [7]. 4tε2 , It is difficult to explicitely characterize (1) and (2), especially for those V that take on positive values. Arsen'ev proved (2) for V ∈ Lp + L∞ with p > d/2, d ≥ 3, and van Casteren [22] proved it for V in the intersection of the Kato class and Ld/2 + L∞ for d ≥ 3 (see [19]). Arsen'ev also obtained (1) for V ∈ Lp with p > d/2 under additional smoothness assumptions (see [16]). Zhang [27, Theorem 1.1] and Milman and Semenov [19, Theorem 1C, Remark (2)] gave sufficient supremum-integral conditions for (2) and (1) for signed V and characterized (1) for V ≤ 0. Their results left open certain natural questions about the class of admissible functions V , especially for dimensions d ≥ 4. We were particularly motivated by the question of Liske- vich and Semenov about the connection of the sharp Gaussian estimates, the potential-boundedness and the Ld/2 integrability condition, cf. [16, Remark (3), p. 602]. In this work we use potential-boundedness (7) and bridges potential-boundedness of V to study the connection of the sharp Gauss- ian estimates and the Lp integrability, disregarding the Kato condition. In Theorem 2.9 below we give new sufficient conditions for the sharp Gaussian estimates, which help verify that Lp integrability is not necessary for (1) or (2). Namely, for d ≥ 3 we present in Corollary 3.4 functions V such that (1) holds but V /∈ L1(Rd)∪Sp>1 Lp loc(Rd). Our examples are highly anisotropic because they are constructed from tensor products, and to study them we crucially use factorization of the Gauss-Weierstrass kernel. Before discussing the present results we should mention our more recent paper [4], where we give a new characterization of (1) and resolve the question of Liskevich and Semenov. Both papers grew out from our work on this question but contain different observations. In fact [4] uses the preliminary results stated in this Introduction, apart from which the two papers have no overlap. The structure of the remainder of the paper is the following. Below in this section we give definitions and preliminaries, and organize the relevant In particular in Lemma 1.1 we present results existing in the literature. SHARP GAUSSIAN ESTIMATES 3 characterizations of (1) and (2) for V ≤ 0. In Theorem 2.9 of Section 2 we propose new sufficient conditions for (1) and (2), with emphasis on those functions V which factorize as tensor products. In Section 3 we prove Corol- lary 3.4 and give examples which illustrate and comment our results. In Section 4 we give supplementary details. Let N = {1, 2, . . .}, f + = max{0, f} and f − = max{0,−f}. All the Here is a quantity to characterize (1) and (2): considered functions V : Rd → [−∞,∞] are assumed Borel measurable. S(V, t, x, y) =Z t t > 0, x, y ∈ Rd. In what follows we often abbreviate S(V ). The motivation for using S(V ) comes from [27, Lemma 3.1 and Lemma 3.2] and from [5, (1)]. g(s, x, z)g(t − s, z, y) V (z) dzds, 0 ZRd g(t, x, y) In the next two results we compile [27, Theorem 1.1] and observations from [5] and [6] to give conditions for the sharp Gaussian estimates. For completeness, the proofs are given in Section 4. Lemma 1.1. If V ≤ 0, then (1) is equivalent to (3) sup t>0, x,y∈Rd S(V, t, x, y) < ∞. If V ≤ 0, then for every T ∈ (0,∞), (2) is equivalent to (4) sup S(V, t, x, y) < ∞. 0<t≤T, x,y∈Rd It is appropriate to say that V satisfying (3) or (4) has bounded poten- tial for bridges (is bridges potential-bounded), globally or locally in time, respectively, cf. Section 2. Lemma 1.2. If for some h > 0 and 0 ≤ η < 1 we have sup 0<t≤h, x,y∈Rd S(V +, t, x, y) ≤ η, and S(V −) is bounded on bounded subsets of (0,∞) × Rd × Rd then (5) G(t, x, y) , g(t, x, y) ≤(cid:18) 1 1 − η(cid:19)1+t/h e−S(V −,t,x,y) ≤ t > 0, x, y ∈ Rd . We record the following observations on integrability and on the potential- boundedness (7) of functions V which are bridges potential-bounded. Lemma 1.3. If S(V, t, x, y) < ∞ for some t > 0, x, y ∈ Rd, then V ∈ L1 loc(Rd). If (4) holds, then (6) g(s, x, z)V (z) dzds < ∞ . sup x∈RdZ T 0 ZRd x∈RdZ ∞ 0 ZRd If (3) even holds, then V has bounded Newtonian potential: (7) sup g(s, x, z)V (z) dzds < ∞ . If V ≥ 0, then (1) implies (3) and (2) implies (4). If d = 3 and V ≤ 0, then (7) is equivalent to (1). SHARP GAUSSIAN ESTIMATES 4 Proof. The first statement follows, because g(t, x, y) is locally bounded from below on (0,∞) × Rd × Rd (see [12, Lemma 3.7] for a quantitative general result). Since RRd S(V, t, x, y)g(t, x, y) dy = R t 0 RRd g(s, x, z)V (z) dzds, we see that (4) implies (6) and (3) implies (7). The next to the last sentence easily follows from Duhamel formula and the fact that G ≥ g in this case. The last statement follows from [19, Remark (2) and (3) on p. 4]. (cid:3) We note that (7) and thus also (3) fail for all nonzero V in dimensions d = 1 and d = 2, because then R ∞ 0 g(s, x, z)ds ≡ ∞. Consequently, (1) fails for nontrivial V ≤ 0 if d = 1 or 2. For d ≥ 3 we let Cd = Γ(d/2−1) . The Newtonian potential of nonnegative function f and x ∈ Rd will be denoted −∆−1f (x) :=Z ∞ z − xd−2 f (z) dz . Thus, (7) reads k∆−1V k∞ < ∞. We also note that by Theorem 1C (2), Re- mark (3) on p. 4, and the comments before Theorem 1B in [19], k∆−1V k∞ < 1 suffices for (1) when d = 3 and V ≥ 0. g(s, x, z)f (z) dzds = CdZRd 0 ZRd 4πd/2 1 2. Sufficient conditions for the sharp Gaussian estimates Recall from [8, (2.5)] that for p ∈ [1,∞], kPtfk∞ ≤ C(d, p) t−d/(2p)kfkp , t > 0 , where Ptf (x) =RRd g(t, x, z)f (z)dz, f ∈ Lp(Rd) and (4π)−d/(2p)(1 − p−1)(1−p−1)d/2, C(d, p) =((4π)−d/2, if p = 1 , if p ∈ (1,∞] . We will give an analogue for the bridges T t,y s T t,y s f (x) =ZRd Clearly, g(s, x, z) g(t − s, z, y) g(t, x, y) . Here t > 0, y ∈ Rd, and f (z) dz , 0 < s < t, x ∈ Rd . (8) s f (x) = T t,x T t,y t−sf (y), 0 < s < t, x, y ∈ Rd . By the Chapman-Kolmogorov equations (the semigroup property) for the kernel g, we have T t,y s 1 = 1. We also note that S(V ) is related to the potential (0-resolvent) operator of T as follows, V (x) ds . Lemma 2.1. For p ∈ [1,∞] and f ∈ Lp(Rd) we have S(V, t, x, y) =Z t T t,y s 0 s fk∞ ≤ C(d, p) (cid:20) (t − s)s kT t,y t Proof. We note that g(s, x, z) g(t − s, z, y) (4π)−d/2g(t, x, y) =(cid:20) (t − s)s t (cid:21)−d/(2p) (cid:21)− d 2 kfkp , 0 < s < t, y ∈ Rd . exp(cid:20)−z − x2 4s − y − z2 4(t − s) + y − x2 4t (cid:21) . SHARP GAUSSIAN ESTIMATES 5 As in [25, (3.4)], we have (9) z − x2 4s y − x ≤ √sz − x√s . 4t + y − z2 4(t − s) ≥ y − x2 √t(cid:18)z − x2 s + √t − sy − z √t − s ≤ t − s (cid:19)1/2 + y − z2 . Indeed, (9) obtains from by the triangle and Cauchy-Schwarz inequalities: For p = 1, the assertion of the lemma results from (9). For p ∈ (1,∞), we let p′ = p/(p − 1), apply Hölder's inequality and the semigroup property, and by the first case we obtain s f (x) ≤ g(t, x, y)−1hZ g(s, x, z)p′ T t,y i−d/(2p) = C(d, p)h s(t − s) t g(t − s, z, y)p′ kfkp. dzi1/p′ kfkp Here we also use the identity g(s, x, z)p′ For p = ∞, the assertion follows from the identity T t,y = g(s/p′, x, z)(4πs)(1−p′)d/2(p′)−d/2. (cid:3) s 1 = 1. Zhang [27, Proposition 2.1] showed that (1) and (2) hold for V in specific Lp spaces (see also [27, Theorem 1.1 and Remark 1.1]). We can reprove his result as follows. Proposition 2.2. Let V : Rd → R and p, q ∈ [1,∞]. (a) If V ∈ Lp(Rd), p > d/2 and c = C(d, p) [Γ(1−d/(2p))]2 Γ(2−d/p) kV kp, then sup x,y∈Rd S(V, t, x, y) ≤ c t1−d/(2p) , t > 0 . (b) If V ∈ Lp(Rd) ∩ Lq(Rd) and q < d/2 < p, then (3) holds. Proof. Part (a) follows from Lemma 2.1, so we proceed to (b). For t > 2, Estimating the first term of the sum, by Lemma 2.1 we obtain (10) Z t/2 0 T t,y s (11) 0 0 T t,y s T t,y s V (x) ds =Z t/2 Z t V (x) ds ≤ ckV kpZ 1 ≤ c′ kV kpZ 1 T t,x s V (x) ds +Z t/2 (cid:21)−d/(2p) ds + ckV kqZ t/2 0 (cid:20) (t − s)s s−d/(2p)ds + c′ kV kqZ ∞ t 0 1 1 0 V (y) ds . (cid:20) (t − s)s t (cid:21)−d/(2q) ds s−d/(2q)ds. By (8), the second term has the same bound. For t ∈ (0, 2] we use (a). (cid:3) By Lemma 1.1 and 1.2 we get the following conclusion. Corollary 2.3. Under the assumptions of Proposition 2.2(a), G satisfies the sharp local Gaussian bounds (2). If V ≤ 0 and the assumptions of Proposition 2.2(b) hold, then G has the sharp global Gaussian bounds (1). Recall that [16, Theorem 2] and [19, Remark (1) and (4) on p. 4] yield (1) for d ≥ 4 if k∆−1V −k∞ and kV −kd/2 are finite, k∆−1V +k∞ < 1 and kV −kd/2 is small. We can reduce Proposition 2.2(b) to this result as follows. SHARP GAUSSIAN ESTIMATES 6 Lemma 2.4. The assumptions of Proposition 2.2(b) necessitate that d ≥ 3, V ∈ Ld/2(Rd) and k∆−1V k∞ < ∞. Proof. Plainly, the assumptions of Proposition 2.2(b) imply d > 2 and V ∈ Ld/2(Rd). We now verify that k∆−1V k∞ < ∞. By Hölder's inequality, sup x∈RdZB(0,1) x∈RdZBc(0,1) sup V (z + x) zd−2 V (z + x) zd−2 dz ≤ kz2−d1B(0,1)(z)kp′ kV kp < ∞ , dz ≤ kz2−d1Bc(0,1)(z)kq′ kV kq < ∞ , where p′, q′ are the exponents conjugate to p, q, respectively. (cid:3) In what follows, we propose suitable sufficient conditions for (1) and (2). We let d1, d2 ∈ N and d = d1 + d2. Remark 2.5. The Gauss-Weierstrass kernel g(t, x) in Rd can be represented as a tensor product: g(t, x) = (4πt)−d1/2e−x12/(4t) (4πt)−d2/2e−x22/(4t) , where x1 ∈ Rd1, x2 ∈ Rd2 and x = (x1, x2). The kernels of the bridges factorize accordingly: g(s, x, z) g(t − s, z, y) g(t, x, y) (4πs)−d1/2e−z1−x12/(4s)(4π(t − s))−d1/2e−y1−z12/(4(t−s)) (4πt)−d1/2e−y1−x12/(4t) (4πs)−d2/2e−z2−x22/(4s)(4π(t − s))−d2/2e−y2−z22/(4(t−s)) (4πt)−d2/2e−y2−x22/(4t) . = × Corollary 2.6. Let V1 : Rd1 → R, V2 : Rd2 → R, and V (x) = V1(x1)V2(x2), where x = (x1, x2) ∈ Rd, x1 ∈ Rd1 and x2 ∈ Rd2. Assume that V1 ∈ L∞(Rd1) and supt>0, x2,y2∈Rd2 S(V2, t, x2, y2) < ∞. Then (3) holds. Proof. In estimaing S(V, t, x, y) we first use the factorization of the bridges and the boundedness of V1, and then the Chapman-Kolmogorov equations and the boundedness of S(V2). (cid:3) Let p, p1, p2 ∈ [1,∞]. Definition 2.7. We write f ∈ Lp1(Rd1)×Lp2(Rd2) if there are f1 ∈ Lp1(Rd1) and f2 ∈ Lp2(Rd2), such that f (x1, x2) = f1(x1)f2(x2) , x1 ∈ Rd1, x2 ∈ Rd2. Clearly, Lp(Rd1) × Lp(Rd2) ⊂ Lp(Rd1+d2), in fact kfkp = kf1kpkf2kp if f (x1, x2) = f1(x1)f2(x2). Lemma 2.8. For f (x1, x2) = f1(x1)f2(x2) ∈ Lp1(Rd1)×Lp2(Rd2), 0 < s < t and y ∈ Rd, we have s fk∞ ≤ C(d1, p1) C(d2, p2)(cid:20) (t − s)s kT t,y (cid:21)−d1/(2p1)−d2/(2p2) kf1kp1kf2kp2 . t SHARP GAUSSIAN ESTIMATES Proof. We proceed as in the proof of Lemma 2.1, using Remark 2.5. We extend Proposition 2.2 as follows. 7 (cid:3) Theorem 2.9. Let d1, d2 ∈ N, d = d1 + d2, V : Rd → R, p1, p2 ∈ [1,∞] and d1 2p1 + d2 2p2 = 1 . (a) If r ∈ (p1,∞] and V ∈ Lr(Rd1) × Lp2(Rd2), then S(V, t, x, y) ≤ c t1−d1/(2r)−d2/(2p2) , where c = C(d1, r)C(d2, p2) [Γ(1−d1/(2r)−d2/(2p2))]2 sup x,y∈Rd Γ(2−d1/r−d2/p2) (b) If 1 ≤ q < p1 < r ≤ ∞ and V ∈ (cid:2)Lq(Rd1) ∩ Lr(Rd1)(cid:3)× Lp2(Rd2), then (3) holds. kV1krkV2kp2 . Proof. We follow the proof of Proposition 2.2, replacing Lemma 2.1 by Lemma 2.8. (cid:3) By Lemma 1.1 and 1.2 we get the following conclusion. Corollary 2.10. Under the assumptions of Theorem 2.9(a), G satisfies the sharp local Gaussian bounds (2). If V ≤ 0 and the assumptions of Theo- rem 2.9(b) hold, then G has the sharp global Gaussian bounds (1). Clearly, if U ≤ V , then S(U ) ≤ S(V ). This may be used to extend the conclusions of Theorem 2.9 and Corollary 2.10 beyond tensor products V (x1, x2) = V1(x1)V2(x2). 3. Examples Let 1A denote the indicator function of A. In what follows, G in (1) is the Schrödinger perturbation of g by V . Example 3.1. Let d ≥ 3 and 1 < p < ∞. For x1 ∈ R, x2 ∈ Rd−1 we let V (x1, x2) = −x1−1/p1x1<11x2<1. Then (1) holds but V /∈ Lp Indeed, V (x1, x2) = V1(x1)V2(x2), where loc(Rd). V1(x1) = −x1−1/p1x1<1, V2(x2) = 1x2<1, x2 ∈ Rd−1. x1 ∈ R, Let and 1 ≤ q < p1 < r < p, p2 = d − 1 2 p1 . p1 − 1/2 Since d ≥ 3, p2 > 1. In the notation of Theorem 2.9 we have d1 = 1, d2 = d − 1, and indeed d1/(2p1) + d2/(2p2) = 1. Since V1 ∈ Lr(R) ∩ Lq(R) and V2 ∈ Lp2(Rd−1), the assumptions of Theorem 2.9(b) are satisfied, and (1) follows by Corollary 2.10. Clearly, V /∈ Lp loc(Rd). SHARP GAUSSIAN ESTIMATES 8 Example 3.2. For d ≥ 3, n = 2, 3, . . ., let Vn(x) = x1−1+1/n1x1<11x2<1, where x = (x1, x2), x1 ∈ R, x2 ∈ Rd−1. Let an = supt>0, x,y∈Rd S(Vn, t, x, y), ∞ Vn(x) , 1 n2 an Xn=2 V (x) = − Then (1) holds but V /∈Sp>1 Lp Indeed, 0 < an < ∞ by Example 3.1, and so Xn=2 S(V, t, x, y) ≤ loc(Rd). t>0, x,y∈Rd sup ∞ x ∈ Rd. 1 n2 < ∞ . p−1⌉, m p ≥ 1. Then, m2am(cid:19)pZx2<1Zx1<1 x1− m−1 This yields the global sharp Gaussian bounds. For p > 1 we let m = ⌈ p and we have m ≥ 2, m−1 ZB(0,2) V (x)p dx ≥(cid:18) 1 Example 3.3. Let d ≥ 3 and V (x1, x2) = −1 Then (1) holds but V /∈ L1(Rd). Indeed, we denote V2(x2) = arrangement inequality [15, Chapter 3], in dimension d = 3 we have (x2+1)3 , x2 ∈ Rd, and by the symmetric re- (x2+1)3 for x1 ∈ Rd−3, x2 ∈ R3. m p dx1dx2 = +∞ . −1 0 ≤ ∆−1V2 ≤ C3ZR3 By Lemma 1.3 and Lemma 1.1 1 z(z + 1)3 dz < ∞ . sup t>0, x2,y2∈R3 S(V2, t, x2, y2) < ∞. loc(Rd). By Corollary 2.6 and Lemma 1.1 we see that (1) holds. Clearly, V /∈ L1(Rd). Corollary 3.4. For every d ≥ 3 there is a function V such that (1) holds but V /∈ L1(Rd) ∪Sp>1 Lp Proof. Take the sum of the functions from Example 3.2 and Example 3.3. (cid:3) We can have nonnegative examples, too. Namely, let V ≤ 0 be as in Corollary 3.4. Then M = supt>0,x,y∈Rd S(V, t, x, y) < ∞. We let V = V /(M + 1). Then V ≥ 0, V /∈ L1(Rd) ∪Sp>1 Lp Therefore (5) holds for V with h = ∞ and η = M/(M + 1), which yields (1). Let d1, d2 ∈ N, d = d1 + d2, V1 : Rd1 → R, V2 : Rd2 → R, and V (x1, x2) = V1(x1)+V2(x2), where x1 ∈ Rd1 and x2 ∈ Rd2. Let G1(t, x1, y1), G2(t, x2, y2) be the Schrödinger perturbations of the Gauss-Weierstrass kernels on Rd1 and Rd2 by V1 and V2, respectively. Then G(t, (x1, x2), (y1, y2)) := G1(t, x1, y1) G2(t, x2, y2) is the Schrödinger perturbation of the Gauss-Weierstrass kernel on Rd by V . Clearly, if the sharp Gaussian estimates hold for G1 and G2, then they hold for G. Our next example is aimed to show that such trivial conclusions are invalid for tensor products V (x1, x2) = V1(x1)V2(x2). S( V , t, x, y) = M/(M + 1) < 1 . loc(Rd) and t>0, x,y∈Rd sup SHARP GAUSSIAN ESTIMATES 9 Example 3.5. Let ε ∈ [0, 1). For x1, x2 ∈ R3 let V (x1, x2) = V1(x1)V2(x2), where V1(x) = V2(x) = − 1 − ε 2 x−1−ε 1x<1. Then the fundamental solutions in R3 of ∂t = ∆ + V1 and ∂t = ∆ + V2 satisfy (1) and (2), but that of ∂t = ∆ + V in R6 satisfies neither (1) nor (2). Indeed, by the symmetric rearrangement inequality [15, Chapter 3], 0 ≤ −∆−1V1(x) ≤ −∆−1V1(0) = zz−1−ε dz = 1/2, for all x ∈ R3. Thus, k∆−1V1k∞ = k∆−1V2k∞ < ∞. By Lemma 1.3 we get (1) for the fundamental solutions in R3 of ∂t = ∆ + V1 and ∂t = ∆ + V2. However, the fundamental solution in R6 of ∂t = ∆ + V fails (2). Indeed, if 0 p(s, 0, a)ds, then by [9, Lemma 3.5], 1 − ε 8π Z{z∈R3:z<1} 1 g(s, 0, x)V (x) dxds ≥Z{x∈R6:x2≤T }Z T we let T ≤ 1, a ∈ R6, a = 1, and c =R 1 Z T 0 ZR6 ≥ cZ{x∈R6:x2≤T } ≥ cZ{x1∈R3:x12<T /2} V1(x1)Z{x2∈R3:x22<T /2} c(1 − ε) ≥ x4V (x) dx 1 2 0 Z{x1∈R3:x12<T /2} V1(x1)Z{x2∈R3:x22<T /2} Z{x1∈R3:x12<T /2} V1(x1) x12(T /2 + x12) πT dx1 √T /2 r−1−ε T /2 + r2 dr = ∞ . = c(1 − ε) 2 = π2c T (1 − ε)2Z 0 g(s, 0, x)dsV (x) dx V2(x2) (x12 + x22)2 dx2dx1 x2−1 (x12 + x22)2 dx2dx1 By Lemma 1.3, (4) fails, and so does (2), cf. Lemma 1.1. Thus, the sharp Gaussian estimates may hold for the Schrödinger perturbations of the Gauss- Weierstrass kernels by V1 and V2 but fail for the Schrödinger perturbation of the Gauss-Weierstrass kernel by V (x1, x2) = V1(x1)V2(x2). Considering −V1 and −V2 above by the last two sentences of Section 1, we can have a similar example for nonnegative perturbations, because 1/2 < 1. Let us also remark that the sharp global Gaussian estimates may hold for V (x1, x2) = V1(x1)V2(x2) but fail for V1 or V2. Indeed, it suffices to consider V1(x1) = −1x1<1 on R3 and V2 ≡ 1 on R, and to apply Theorem 2.9. We see that it is the combined effect of the factors V1 and V2 that matters -- as captured in Section 2. 4. Appendix Following [5, 7] we study and use the following functions f (t) = sup x,y∈Rd S(V, t, x, y) , t ∈ (0,∞), F (t) = sup 0<s<t f (s) = sup x,y∈Rd 0<s<t S(V, s, x, y) , t ∈ (0,∞] . SHARP GAUSSIAN ESTIMATES 10 We fix V and x, y ∈ Rd. For 0 < ε < t, we consider S(V, t − ε, x, y) =Z t g(s, x, z)g(t − ε − s, z, y) g(t − ε, x, y) 0 ZRd By Fatou's lemma we get V (z) 1[0,t−ε](u) dzds. S(V, t, x, y) ≤ lim inf ε→0 S(V, t − ǫ, x, y), meaning that (0,∞) ∋ t 7→ S(V, t, x, y) is lower semicontinuous on the left. It follows that f is lower semi-continuous on the left, too. In consequence, f (t) ≤ F (t) and F (t) = sup0<s≤t f (s) for 0 < t < ∞. We next claim that f is sub-additive, that is, (12) f (t1 + t2) ≤ f (t1) + f (t2) , t1, t2 > 0 . This follows from the Chapman-Kolmogorov equations for g. have S(V, t1 + t2, x, y) = I1 + I2, where g(s, x, z)g(t1 + t2 − s, z, y) V (z) dzds Indeed, we g(s, x, z)g(t1 − s, z, w)g(t2, w, y)g(t1, x, w) g(t1 + t2, x, y)g(t1, x, w) V (z) dwdzds S(V, t1, x, w) dw ≤ f (t1) , and I2 equals g(t1 + t2, x, y) g(t1 + t2, x, y) g(t2, w, y)g(t1, x, w) 0 ZRd 0 ZRdZRd I1 =Z t1 =Z t1 ≤ZRd Z t1+t2 g(s, x, z)g(t1 + t2 − s, z, y) ZRd =Z t1+t2 ZRdZRd ≤ZRd g(t1, x, w)g(t2, w, y) g(t1 + t2, x, y) t1 t1 V (z) dzds g(t1 + t2, x, y) g(t1, x, w)g(s − t1, w, z)g(t2 − (s − t1), z, y)g(t2, w, y) g(t1 + t2, x, y)g(t2, w, y) V (z) dwdzds S(V, t2, w, y) dw ≤ f (t2) . This yields (12). Lemma 4.1. For all t, h > 0 we have f (t) ≤ F (h) + t f (h)/h. Proof. Let k ∈ N be such that (k − 1)h < t ≤ kh, and let θ = t − (k − 1)h. Then t = θ + (k − 1)h, and by (12) we get f (t) ≤ f (θ) + t f (h)/h ≤ F (h) + t f (h)/h , (cid:3) since 0 < θ ≤ h. Corollary 4.2. F (t) ≤ F (h) + t F (h)/h and F (2t) ≤ 2F (t) for t, h > 0. Proof of Lemma 1.2. The left-hand side of (5) follows from [27, pp. 467-469] and Lemma 1.3, or we can use [5, (41)], which follows therein from Jensen's inequality and the second displayed formula on page 252 of [5]. We now prove the right hand side of (5). Since G is increasing in V , we may assume that V ≥ 0. For 0 < s < t, x, y ∈ Rd, we let p0(s, x, t, y) = g(t − s, x, y) and pn(s, x, t, y) = R t s RRd pn−1(s, x, u, z)V (z)p0(u, z, t, y) dz du, n ∈ N. Let SHARP GAUSSIAN ESTIMATES 11 Q : R × R → [0,∞) satisfy Q(u, r) + Q(r, v) ≤ Q(u, v). By [14, Theorem 1] (see also [6, Theorem 3]) if there is 0 < η < 1 such that p1(s, x, t, y) ≤ [η + Q(s, t)]p0(s, x, t, y), (13) then (14) p(s, x, t, y) := ∞ Xn=0 pn(s, x, t, y) ≤(cid:16) 1 1 − η(cid:17)1+ Q(s,t) η p0(s, x, t, y) . Corollary 4.2 and the assumptions of the lemma imply that (13) is satis- fied with η = F (h) < 1 and Q(s, t) = (t − s)F (h)/h. Since G(t, x, y) = p(0, x, t, y), the proof of (5) is complete (see also [5, (17)]). Proof of Lemma 1.1. Let V ≤ 0. By the proof of Theorem 1.1(a) at the bottom of p. 468 in [27], the boundedness of S(V, t, x, y) is necessary and sufficient for (1). In particular, by the displayed formula proceeding [27, (3.1)] the boundedness of S(V, t, x, y) is sufficient for (1). Alternatively we can apply Jensen inequality to the second displayed formula on p. 252 in [5]. The first part of Lemma 1.1 is proved. The second part is obtained in the same way, by restricting the considerations, and the transition kernel, to bounded time interval. (cid:3) (cid:3) As a consequence of Corollary 4.2 we obtain the following result. Corollary 4.3. Let V ≤ 0 and T > 0. Then (2) holds if and only if (15) for some constants C > 0, c ≥ 0. In fact we can take Ce−ctg(t, x, y) ≤ G(t, x, y) , t > 0, x, y ∈ Rd , ln C = − sup x,y∈Rd 0<t≤T S(V, t, x, y) and c = 1 T sup x,y∈Rd S(V, T, x, y) . Proof. Obviously, (15) implies (2) for every fixed T > 0. Conversely, if (2) holds for fixed T > 0, then by Lemma 1.2 and 4.1 we have G(t, x, y) g(t, x, y) ≥ e−S(V,t,x,y) ≥ e−f (t) ≥ e−F (T )e−tf (T )/T . (cid:3) We note in passing that the above proof shows that (2) is determined by the behavior of supx,y∈Rd S(V, t, x, y) for small t > 0. We end our discussion by recalling the connection of G to ∆ + V aforementioned in Abstract. As it is well known, and can be directly checked by using the Fourier transform or by arguments of the semigroup theory [3, Section 4], Z ∞ s ZRd g(u − s, x, z)h∂uφ(u, z) + ∆φ(u, z)idzdu = −φ(s, x) , c (R × Rd), the smooth compactly for all s ∈ R, x ∈ Rd and for all φ ∈ C ∞ supported test functions on space-time. Similarly, if V satisfies the assump- tions of Lemma 1.2, then by [27, Theorem 1.1] for all s ∈ R, x ∈ Rd, c (R × Rd), φ ∈ C ∞ Z ∞ s ZRd G(u − s, x, z)h∂uφ(u, z) + ∆φ(u, z) + V (z)φ(u, z)idzdu = −φ(s, x) . SHARP GAUSSIAN ESTIMATES 12 We refer to [6, Lemma 4] for a general approach to such identities. Acknowledgement Jacek Dziubański was supported by the Polish National Science Cen- ter (Narodowe Centrum Nauki) grant DEC-2012/05/B/ST1/00672. Karol Szczypkowski was partially supported by IP2012 018472 and by the German Science Foundation (SFB 701). References [1] M. Aizenman and B. Simon. Brownian motion and Harnack inequality for Schrödinger operators. Comm. Pure Appl. Math., 35(2):209 -- 273, 1982. [2] D. G. Aronson. Non-negative solutions of linear parabolic equations. Ann. Scuola Norm. Sup. Pisa (3), 22:607 -- 694, 1968. [3] K. Bogdan, Y. Butko, and K. Szczypkowski. Majorization, 4G Theorem and Schrödinger perturbations. Journal of Evolution Equations, pages 1 -- 20, 2015. [4] K. Bogdan, J. Dziubański, and K. Szczypkowski. Characterization of sharp global Gaussian estimates for Schr\"odinger heat kernels. ArXiv e-prints, June 2016. [5] K. Bogdan, W. Hansen, and T. Jakubowski. Time-dependent Schrödinger perturba- tions of transition densities. Studia Math., 189(3):235 -- 254, 2008. [6] K. Bogdan, T. Jakubowski, and S. Sydor. Estimates of perturbation series for kernels. J. Evol. Equ., 12(4):973 -- 984, 2012. [7] K. Bogdan and K. Szczypkowski. Gaussian estimates for Schrödinger perturbations. Studia Math., 221(2):151 -- 173, 2014. [8] R. Carmona. Regularity properties of Schrödinger and Dirichlet semigroups. J. Funct. Anal., 33(3):259 -- 296, 1979. [9] K. L. Chung and Z. X. Zhao. From Brownian motion to Schrödinger's equation, volume 312 of Grundlehren der Mathematischen Wissenschaften [Fundamental Prin- ciples of Mathematical Sciences]. Springer-Verlag, Berlin, 1995. [10] E. B. Davies. One-parameter semigroups, volume 15 of London Mathematical Society Monographs. Academic Press Inc. [Harcourt Brace Jovanovich Publishers], London, 1980. [11] J. Dziubański and J. Zienkiewicz. Hardy spaces H 1 for Schrödinger operators with compactly supported potentials. Ann. Mat. Pura Appl. (4), 184(3):315 -- 326, 2005. [12] T. Grzywny and K. Szczypkowski. Kato classes for Lévy processes. ArXiv e-prints, Mar. 2015. [13] A. Gulisashvili and J. A. van Casteren. Non-autonomous Kato classes and Feynman- Kac propagators. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2006. [14] T. Jakubowski. On combinatorics of Schrödinger perturbations. Potential Anal., 31(1):45 -- 55, 2009. [15] E. H. Lieb and M. Loss. Analysis, volume 14 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, second edition, 2001. [16] V. Liskevich and Y. Semenov. Two-sided estimates of the heat kernel of the Schrödinger operator. Bull. London Math. Soc., 30(6):596 -- 602, 1998. [17] V. Liskevich and Y. Semenov. Estimates for fundamental solutions of second-order parabolic equations. J. London Math. Soc. (2), 62(2):521 -- 543, 2000. [18] V. Liskevich, H. Vogt, and J. Voigt. Gaussian bounds for propagators perturbed by potentials. J. Funct. Anal., 238(1):245 -- 277, 2006. [19] P. D. Milman and Y. A. Semenov. Heat kernel bounds and desingularizing weights. J. Funct. Anal., 202(1):1 -- 24, 2003. [20] R. Schnaubelt and J. Voigt. The non-autonomous Kato class. Arch. Math. (Basel), 72(6):454 -- 460, 1999. [21] B. Simon. Schrödinger semigroups. Bull. Amer. Math. Soc. (N.S.), 7(3):447 -- 526, 1982. SHARP GAUSSIAN ESTIMATES 13 [22] J. A. van Casteren. Pointwise inequalities for Schrödinger semigroups. In Semigroup theory and applications (Trieste, 1987), volume 116 of Lecture Notes in Pure and Appl. Math., pages 67 -- 94. Dekker, New York, 1989. [23] J. Voigt. Absorption semigroups, their generators, and Schrödinger semigroups. J. Funct. Anal., 67(2):167 -- 205, 1986. [24] Q. S. Zhang. On a parabolic equation with a singular lower order term. Trans. Amer. Math. Soc., 348(7):2811 -- 2844, 1996. [25] Q. S. Zhang. Gaussian bounds for the fundamental solutions of ∇(A∇u)+B∇u−ut = 0. Manuscripta Math., 93(3):381 -- 390, 1997. [26] Q. S. Zhang. On a parabolic equation with a singular lower order term. II. The Gaussian bounds. Indiana Univ. Math. J., 46(3):989 -- 1020, 1997. [27] Q. S. Zhang. A sharp comparison result concerning Schrödinger heat kernels. Bull. London Math. Soc., 35(4):461 -- 472, 2003. Wrocław University of Science and Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland E-mail address: [email protected] University of Wrocław, Pl. Grunwaldzki 2/4, 50-384 Wrocław, Poland E-mail address: [email protected] Universität Bielefeld, Postfach 10 01 31, D-33501 Bielefeld, Germany and Wrocław University of Science and Technology, Wybrzeże Wyspiańskiego 27, 50-370 Wrocław, Poland E-mail address: [email protected], [email protected]
1309.3694
2
1309
2013-09-25T02:51:26
Isomorphism, nonisomorphism, and amenability of L^p UHF algebras
[ "math.FA" ]
In a previous paper, we introduced L^p UHF algebras for p in [1, \infty). We concentrated on the spatial L^p UHF algebras, which are classified up to isometric isomorphism by p and the scaled ordered K_0-group. In this paper, we concentrate on a larger class, the L^p UHF algebras of tensor product type constructed using diagonal similarities. Such an algebra is still simple and has the same K-theory as the corresponding spatial L^p UHF algebra. For each choice of p and the K-theory, we provide uncountably many nonisomorphic such algebras. We further characterize the spatial algebras among them. In particular, if A is one of these algebras, then A is isomorphic (not necessarily isometrically) to a spatial L^p UHF algebra if and only if A is amenable as a Banach algebra, also if and only if A has approximately inner tensor flip. These conditions are also equivalent to a natural numerical condition defined in terms of the ingredients used to construct the algebra.
math.FA
math
ISOMORPHISM, NONISOMORPHISM, AND AMENABILITY OF Lp UHF ALGEBRAS N. CHRISTOPHER PHILLIPS Abstract. In a previous paper, we introduced Lp UHF algebras for p ∈ [1, ∞). We concentrated on the spatial Lp UHF algebras, which are classified up to isometric isomorphism by p and the scaled ordered K0-group. In this paper, we concentrate on a larger class, the Lp UHF algebras of tensor product type constructed using diagonal similarities. Such an algebra is still simple and has the same K-theory as the corresponding spatial Lp UHF algebra. For each choice of p and the K-theory, we provide uncountably many nonisomorphic such algebras. We further characterize the spatial algebras among them. In particular, if A is one of these algebras, the following are equivalent: • A is isomorphic (not necessarily isometrically) to a spatial Lp UHF al- gebra. • A is amenable as a Banach algebra. • A has approximately inner tensor flip. These conditions are also equivalent to a natural numerical condition defined in terms of the ingredients used to construct the algebra. In [14], we introduced analogs of UHF C*-algebras which act on Lp spaces instead of on Hilbert space. The purpose of this paper is to prove further results about such algebras. For each p ∈ [1, ∞) and choice of K-theory, we prove the existence of uncountably many such algebras with the given K-theory and which are not isomorphic, even when the isomorphisms are not required to be isometric. We give several conditions which characterize the "standard" examples of [14] within a larger class, including amenability and approximate innerness of the tensor flip. In [14], we were primarily interested in a particular family of examples, the spatial Lp UHF algebras, which were needed for the proof of simplicity of the analogs of Cuntz algebras acting on Lp spaces. For p = 2, a spatial Lp UHF algebra is a UHF C*-algebra. For general p ∈ [1, ∞), we showed that p, together with the scaled ordered K0-group (equivalently, the supernatural number), is a complete invariant for both isomorphism and isometric isomorphism of spatial Lp UHF algebras. We also proved that spatial Lp UHF algebras are amenable Banach algebras. For a larger class of Lp UHF algebras, we proved simplicity and existence of a unique normalized trace. Here, we consider Lp UHF algebras A of tensor product type, a subclass of the class for which we proved simplicity in [14]. Tensor product type means that there are a sequence d = (d(1), d(2), . . .) in {2, 3, 4, . . .}, probability spaces (Xn, Bn, µn), and unital representations ρn : Md(n) → L(Lp(Xn, µn)) such that, with (X, B, µ) being the product measure space, A is the closed subalgebra of the bounded oper- ators on Lp(X, µ) =N∞ n=1 Lp(Xn, µn) generated by all operators of the form ρ1(a1) ⊗ ρ2(a2) ⊗ · · · ⊗ ρm(am) ⊗ 1 ⊗ 1 ⊗ · · · Date: 24 September 2013. 2010 Mathematics Subject Classification. Primary 46H20; Secondary 46H05, 47L10. This material is based upon work supported by the US National Science Foundation under Grant DMS-1101742. 1 2 with N. CHRISTOPHER PHILLIPS m ∈ Z≥0, a1 ∈ Md(1), a2 ∈ Md(2), · · · , am ∈ Md(m). For our strongest results, we further require that the representations ρn be direct sums of finitely or countably many representations similar to the identity repre- sentation. That is, identifying Md(m) with the algebra of bounded operators on a probability space with d(m) points, they are direct sums of maps a 7→ sas−1 for invertible elements s ∈ Md(m). We mostly further require that the matrices s all be diagonal. The K-theory of A is determined by the formal product of the numbers d(n), more precisely, by the formal prime factorization in which the exponent of a prime p is the sum of its exponents in the prime factorizations of the numbers d(n). To avoid the intricacies of K-theory, we work in terms of this invariant, called a "supernatural number" (Definition 1.1 below), instead. Our convention is that isomorphisms of Banach algebras are required to be con- tinuous and have continuous inverses, but they are not required to be isometric. To specify the more restrictive version, we refer to an isometric isomorphism, or say that two Banach algebras are isometrically isomorphic. For the rest of the introduction, fix p ∈ [1, ∞) and a supernatural number N. For an Lp UHF algebra A ⊂ L(Lp(X, µ)) of tensor product type constructed using diag- onal similarities and with the given supernatural number N, we give (Theorem 4.10) a number of equivalent conditions for isomorphism with the spatial Lp UHF algebra with supernatural number N. One of these is approximate innerness of the tensor flip on the Lp operator tensor product: there is a bounded net (vλ)λ∈Λ in the Lp operator tensor product A ⊗p A such that (v−1 λ )λ∈Λ is also bounded and such that limλ∈Λ vλ(a ⊗ b)v−1 λ = b ⊗ a for all a, b ∈ A. (The C*-algebra version of this condition has been studied in [5]. It is rare.) Another condition is numerical: with ρn as in the description above, it is that n=1(kρnk−1) < ∞. Moreover, kρnk is easily computed in terms of the similarities Two further equivalent conditions are amenability and symmetric amenability as a Banach algebra. See Proposition 2.4 and the preceding discussion for more on these conditions. Recall that a C*-algebra is amenable if and only if it is nuclear. (See Corollary 2 of [3] and Theorem 3.1 of [10].) In Theorem 2 of [12], it is shown that a unital nuclear C*-algebra is symmetrically amenable if and only if every nonzero quotient of A has a tracial state. A fifth condition is that A be similar to a spatial Lp UHF algebra, that is, that there be an invertible element v ∈ L(Lp(X, µ)) such that vAv−1 is a spatial Lp UHF subalgebra of L(Lp(X, µ)). When p = 2, our results imply that if A is amenable then it is in fact similar to a C*-algebra. In this case, we can omit the requirement that the similarities in the construction of our algebra be diagonal. A question open for some time (see Problem 30 in the "Open Problems" chapter of [16]) asks whether an amenable closed subalgebra of the bounded operators on a Hilbert space is similar to a C*-algebra. Little seems to be known. Two main positive results are for subalgebras of K(H) (see the last paragraph of [9]) and for commutative subalgebras of finite factors (see [2]). A claimed result for singly generated operator algebras [7] has been retracted [8]. If A ⊂ L(H) is a closed amenable subalgebra, then A is already a C*-algebra if A is generated by elements which are normal in L(H) ([4]) or if A is unital and 1-amenable (Theorem 7.4.18(2) of [1]). The answer is negative in the inseparable case [6]. The class of examples for which we give a positive answer, the L2 UHF algebras of tensor product type constructed using similarities, is small (although it includes uncountably many isomorphism types for every supernatural number N ), but is quite different from any other classes of P∞ used in the construction of A. AMENABILITY 3 examples for which the question was previously known to have a positive answer. The fact that the theorem holds for any p ∈ [1, ∞) when the similarities are diagonal suggests that there might be an interesting Lp analog of this question. We have partial results in the same direction for general Lp UHF algebras of tensor product type. We have no counterexamples to show that the results de- scribed above fail in this broader class. However, we also do not know how to deduce anything from amenability of a general Lp UHF algebras of tensor product type, or even from the stronger condition of symmetric amenability. We do have a condition related to approximate innerness of the tensor flip, and a numerical condition involving norms of representations, which for algebras in this class imply isomorphism to the corresponding spatial algebra. We further prove (Theorem 5.14) that for every p ∈ [1, ∞) and supernatural number N, there are uncountably many pairwise nonisomorphic Lp UHF algebras of tensor product type constructed using diagonal similarities. In particular, there are uncountably many nonisomorphic such algebras which are not amenable. This paper is organized as follows. In Section 1, we recall some material from [14], in particular, the construction of Lp UHF algebras of tensor product type. In Section 2, we discuss amenability, symmetric amenability, approximate innerness of the tensor flip, and approximate innerness of the related tensor half flip. For Lp UHF algebras of tensor product type, we prove the implications we can between these conditions, several related conditions, and isomorphism to the spatial Lp UHF algebra. In Section 3, we specialize to Lp UHF algebras of tensor product type constructed using systems of similarities. The new feature is that the norms of the representa- tions involved are more computable. We prove a number of lemmas which will be used in the remaining two sections. Although we do not formally state it, at the end we discuss a strengthening of the main result of Section 2. In Section 4 we further specialize to diagonal similarities. One can then get good information on the norms of matrix units. We prove the equivalence of a number of conditions (some of them described above) to isomorphism with the corresponding spatial Lp UHF algebra. Section 5 gives the construction of uncountably many nonisomorphic algebras in this class. We are grateful to Volker Runde for useful email discussions and for providing references, and to Narutaka Ozawa for suggesting that we consider amenability of our algebras. Some of this work was done during a visit to Tokyo University during December 2012. We are grateful to that institution for its hospitality. 1. Lp UHF algebras of tensor product type In this section, we recall the construction of Lp UHF algebras of (infinite) tensor product type (Example 3.8 of [14]), and several other results and definitions of [14]. We start by recalling supernatural numbers, which are the elementary form of the isomorphism invariant for spatial Lp UHF algebras. Definition 1.1 (Definition 3.3 of [14]). Let P be the set of prime numbers. A supernatural number is a function N : P → Z>0 ∪ {∞} such thatPt∈P N (t) = ∞. Let d = (d(1), d(2), . . .) be a sequence in {2, 3, 4, . . .}. We define rd(n) = d(1)d(2) · · · d(n) for n ∈ Z≥0. (Thus, rd(0) = 1.) We then define the supernatural number associated with d to be the function Nd : P → Z≥0 ∪ {∞} given by Nd(t) = sup(cid:0)(cid:8)k ∈ Z≥0 : there is n ∈ Z≥0 such that tk divides rd(n)(cid:9)(cid:1). 4 N. CHRISTOPHER PHILLIPS Definition 1.2 (Definition 3.4 of [14]). Let (X, B, µ) be a σ-finite measure space, let p ∈ [1, ∞), and let A ⊂ L(Lp(X, µ)) be a unital subalgebra. Let N be a supernatural number. We say that A is an Lp UHF algebra of type N if there exist a sequence d as in Definition 1.1 with Nd = N, unital subalgebras D0 ⊂ D1 ⊂ · · · ⊂ A, and algebraic isomorphisms σn : Mrd(n) → Dn, such that A =S∞ Notation 1.3. For d ∈ Z>0 and p ∈ [1, ∞], we let lp normalized counting measure on {1, 2, . . . , d}, that is, the total mass is 1. We further let M p with the algebra Md of d × d complex matrices in the standard way. d = lp(cid:0){1, 2, . . . , d}(cid:1), using d(cid:1) with the usual operator norm, and we algebraically identify M p d = L(cid:0)lp n=0 Dn. d One gets the same normed algebra M p d if one uses the usual counting measure. (We make a generalization explicit in Lemma 2.11 below.) Many articles on Banach spaces use Lp(X, µ) rather than Lp(X, µ), and use ld p for what we call lp d. Our convention is chosen to avoid conflict with the standard notation for Leavitt algebras, which appear briefly here and play a major role in the related papers [13], [14], and [15]. We now recall the construction of Lp UHF algebras of tensor product type (Ex- ample 3.8 of [14]). See [14] for further details and for justification of the statements made here, in particular, for the tensor product decomposition of Lp of a product of measure spaces. Example 1.4 (Example 3.8 of [14]). Let p ∈ [1, ∞). We take N = Z>0. For each n ∈ N, let (Xn, Bn, µn) be a probability space, let d(n) ∈ {2, 3, . . .}, and let ρn : Md(n) → L(Lp(Xn, µn)) be a representation (unital, by our conventions). For every subset S ⊂ N, let (XS, BS, µS) be the product measure space of the spaces (Xn, Bn, µn) for n ∈ S. We let 1S denote the identity operator on Lp(XS, µS). For S ⊂ N and n ∈ Z≥0 we take S≤n = S ∩ {1, 2, . . . , n} and S>n = S ∩ {n + 1, n + 2, . . .}. We make the identification (1.1) Lp(XS, µS) = Lp(XS≤n, µS≤n ) ⊗p Lp(XS>n, µS>n ). Suppose now that S is finite. Set l(S) = card(S), and write S = {mS,1, mS,2, . . . , mS,l(S)} with mS,1 < mS,2 < · · · < mS,l(S). We make the identification Lp(XS, µS) = Lp(XmS,1, µmS,1)⊗p Lp(XmS,2, µmS,2)⊗p · · ·⊗p Lp(XmS,l(S) , µmS,l(S)). Set d(S) = d(mS,j) and MS = Md(mS,j) ⊂ Lp(XS, µS). l(S)Yj=1 l(S)Oj=1 We take d(∅) = 1 and M∅ = C. Then MS ∼= Md(S). Further let ρS : MS → L(Lp(XS, µS)) be the unique representation such that for a1 ∈ Md(mS,1), a2 ∈ Md(mS,2), · · · , al(S) ∈ Md(mS,l(S)), we have ρS(a1 ⊗ a2 ⊗ · · · ⊗ al(S)) = ρd(mS,1)(a1) ⊗ ρd(mS,2)(a2) ⊗ · · · ⊗ ρd(mS,l(S))(al(S)). For finite sets S ⊂ T ⊂ N, there is an obvious homomorphism ϕT,S : MS → MT , obtained by filling in a tensor factor of 1 for every element of T \ S. We then define ρT,S = ρT ◦ ϕT,S : MS → L(Lp(XT , µT )). When S is finite but T is not, we define a representation of MS on Lp(XT , µT ) as follows. Choose some n ≥ sup(S), and, following (1.1) and Theorem 2.16(3) of [13], for a ∈ MS set ρT,S(a) = ρT≤n,S(a) ⊗ 1T>n ∈ L(Lp(XT , µT )). AMENABILITY 5 We equip MS with the norm kak = kρS(a)k for all a ∈ MS. Then the maps ϕT,S, for S ⊂ T ⊂ N finite, and ρT,S, for S ⊂ T ⊂ N with S finite, are all isometric. For S ⊂ N finite, we now define AS ⊂ L(Lp(X, µ)) by AS = ρN,S(MS), and for S ⊂ N n=0 AS≤n . In a similar way, we define AT,S ⊂ L(Lp(XT , µT )), infinite, we set AS =S∞ for arbitrary S, T with S ⊂ T ⊂ N. The algebra A = AN is an Lp UHF algebra of type Nd in the sense of Def- inition 1.2. When the ingredients used in its construction need to be specified, we set d = (d(1), d(2), . . .) and ρ = (ρ1, ρ2, . . .), and write A(d, ρ). We also take σn : Mrd(n) → A to be σn = ρN,N≤n. The basic representations of M p d are the spatial representations. For the purposes of this paper, we use the following condition for a representation to be spatial. Proposition 1.5. Let p ∈ [1, ∞), let d ∈ Z>0, let (X, B, µ) be a σ-finite measure space, and let ρ : M p d → L(Lp(X, µ)) be a unital homomorphism. Let (ej,k)j,k=1,2,...,d be the standard system of matrix units for M p d . Then ρ is a spatial representa- j=1 Xj such that for j, k = 1, 2, . . . , d the operator ρ(ej,k) is a spatial partial isometry, in the sense of Definition 6.4 of [13], with domain support Xk and range support Xj. tion if and only if there is a measurable partition X = `d Proof. For p 6= 2, this statement easily follows by combining several parts of The- orem 7.2 of [13]. For p = 2 (and in fact for any p ∈ [1, ∞)), it is also easily proved directly. (cid:3) k → L(Lp(X, µ)) and σ : M p Corollary 1.6. Let p ∈ [1, ∞), let k, l ∈ Z>0, let (X, B, µ) and (Y, C, ν) be σ- finite measure spaces, and let ρ : M p l → L(Lp(Y, ν)) be spatial representations. Identify Lp(X, µ) ⊗p Lp(Y, ν) with Lp(X × Y, µ × ν) as in Theorem 2.16 of [13] and M p kl as in Corollary 1.13 of [14]. Then ρ ⊗ σ : M p kl → L(cid:0)Lp(X × Y, µ × ν)(cid:1) is a spatial representation. Proof. By Lemma 6.20 of [13], the tensor product of spatial partial isometries is again a spatial partial isometry, with domain and range supports given as the products of the domain and range supports of the tensor factors. The result is then immediate from Proposition 1.5. (cid:3) k ⊗p M p l with M p For p ∈ [1, ∞) \ {2}, a representation of Md on an Lp space is spatial if and only if it is isometric (see Theorem 7.2(3) of [13]), but this is not true for p = 2. We briefly recall that a representation of Md has enough isometries (Definition 2.7(1) of [14]) if there is an irreducibly acting subgroup of its invertible group inv(Md) whose images are isometries, that it locally has enough isometries (Defi- nition 2.7(2) of [14]) if it is a direct sum (in the sense of Definition 3.6 below) of representations with enough isometries (presumably using different subgroups for the different summands), and that it dominates the spatial representation (Defini- tion 2.7(3) of [14]) if in addition one of the summands is spatial. Definition 1.7 (Definition 3.9 of [14]). Let (X, B, µ) be a σ-finite measure space, let p ∈ [1, ∞), and let A ⊂ L(Lp(X, µ)) be a unital subalgebra. (1) We say that A is an Lp UHF algebra of tensor product type if there ex- ist d and ρ = (ρ1, ρ2, . . .) as in Example 1.4 such that A is isometrically isomorphic to A(d, ρ). (2) We say that A is a spatial Lp UHF algebra if in addition it is possible to choose each representation ρn to be spatial in the sense of Definition 7.1 of [13]. (3) We say that A locally has enough isometries if it is possible to choose d and ρ as in (1) such that, in addition, ρn locally has enough isometries, in the sense of Definition 2.7(2) of [14], for all n ∈ Z>0. 6 N. CHRISTOPHER PHILLIPS (4) If A locally has enough isometries, we further say that A dominates the spatial representation if it is possible to choose d and ρ as in (3) such that, in addition, for all n ∈ Z>0 the representation ρn dominates the spatial representation in the sense of Definition 2.7(3) of [14]. In particular, if we take the infinite tensor product of the algebras M p d(n), repre- sented as in Notation 1.3, the resulting Lp UHF algebra of tensor product type is spatial. Theorem 1.8. Let p ∈ [1, ∞) and let N be a supernatural number. Then there exists a spatial Lp UHF algebra A of type N. It is unique up to isometric isomor- phism. It is of tensor product type, in fact isometrically isomorphic to A(d, ρ) as in Example 1.4 for any sequence d with Nd = N and any sequence ρ consisting of spatial representations. Moreover, for any σ-finite measure space (Z, D, λ) and any closed unital subalgebra D ⊂ L(Lp(Z, λ)), any two choices of spatial direct system of type N (as in Definition 3.5 of [14]), with any unital isometric representations of the direct limits on Lp spaces, give isometrically isomorphic Lp tensor products A ⊗p D. Proof. For all but the last two sentences, see Theorem 3.10 of [14]. For the second last sentence, the only additional fact needed is that A(d, ρ) is a spatial Lp UHF algebra. This follows from Definition 3.5 of [14] and Corollary 1.6. We prove the last statement. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces. Let A ⊂ L(Lp(X, µ)) and B ⊂ L(Lp(Y, ν)) be spatial Lp UHF algebras of type N. Let A0 ⊂ A1 ⊂ · · · ⊂ A and B0 ⊂ B1 ⊂ · · · ⊂ B be subalgebras such that, with ϕn,m : Am → An and ψn,m : Bm → Bn being the inclusion maps for m ≤ n, the systems(cid:0)(An)n∈Z≥0, (ϕn,m)m≤n(cid:1) and(cid:0)(Bn)n∈Z≥0, (ψn,m)m≤n(cid:1) are n=0 An and B =S∞ spatial direct systems of type N, and such that A =S∞ ∞[n=0 An⊗pD and B⊗pD = Bn ⊗p D ∼= lim −→ Bn⊗pD, n=0 Bn. with the isomorphisms being isometric. Theorem 3.7 of [14] shows that the direct limits are isometrically isomorphic. (cid:3) Then A⊗pD = An ⊗p D ∼= lim −→ ∞[n=0 2. Spatial Lp UHF algebras In this section, we consider Lp UHF algebras of tensor product type for a fixed value of p and a fixed supernatural number. We give some conditions which im- ply that such an algebra is isomorphic to the corresponding spatial algebra. We also prove that the spatial Lp UHF algebras are symmetrically amenable, hence amenable in the sense of Banach algebras, and that they have approximately inner tensor flip. The machinery we develop will be used for the more special classes considered in later sections. We will need the projective tensor product of Banach spaces. Many of its prop- erties are given in Section B.2.2 of Appendix B in [16]. There are few proofs, but the omitted proofs are easy. tensor product of Definition B.2.10 of [16], and similarly for more than two factors. When the norm on this space must be made explicit to avoid confusion, we denote it by k · kπ. (It is given on an element µ in the algebraic tensor product E ⊗alg k=1 kξkk · kηkk for which ξ1, ξ2, . . . , ξn ∈ E and k=1 ξk ⊗ ηk = µ.) If E1, E2, F1, and F2 are Banach Notation 2.1. Let E and F be Banach spaces. We denote by Eb⊗F the projective F by the infimum of all sums Pn η1, η2, . . . , ηn ∈ F satisfy Pn AMENABILITY 7 The projective tensor product is the maximal tensor product, in the sense that other cross norm. (See Exercise B.2.9 of [16].) The projective tensor product of Banach algebras A and B is a Banach algebra (Exercise B.2.14 of [16]), in which, spaces, a1 ∈ L(cid:0)E1, F1), and a2 ∈ L(E2, F2(cid:1), we write ab⊗b for the tensor product map in L(cid:0)E1b⊗E2, F1b⊗F2(cid:1). there is a contractive linear map from Eb⊗F to the completion of E ⊗alg F in any by the definition of a cross norm,(cid:13)(cid:13)ab⊗b(cid:13)(cid:13) = kak · kbk for a ∈ A and b ∈ B. linear maps ∆A : Ab⊗A → A and ∆op A : Ab⊗A → A such that ∆A(a ⊗ b) = ab and Lemma 2.2. Let A be a Banach algebra. Then there are unique contractive The main purpose of the following lemma is to establish notation for the maps Proof. This is immediate from the standard properties of the projective tensor product. (cid:3) A (a ⊗ b) = ba for all a, b ∈ A. in the statement. ∆op Lemma 2.3. Let A and B be Banach algebras, and let ϕ : A → B be a continuous homomorphism. Then Proof. Use the relation ϕ(xy) = ϕ(x)ϕ(y) for x, y ∈ A. ∆B ◦(cid:0)ϕb⊗ϕ(cid:1) = ϕ ◦ ∆A and ∆op B ◦(cid:0)ϕb⊗ϕ(cid:1) = ϕ ◦ ∆op A . (cid:3) The definition of an amenable Banach algebra is given in Definition 2.1.9 of [16]; see Theorem 2.2.4 of [16] for two standard equivalent conditions. We will also use symmetric amenability (introduced in [11]). The following characterizations for unital Banach algebras are convenient here. Proposition 2.4. Let A be a unital Banach algebra. Then: (1) A is amenable if and only if there is a bounded net (mλ)λ∈Λ in Ab⊗A such that lim λ∈Λ(cid:0)(a ⊗ 1)mλ − mλ(1 ⊗ a)(cid:1) = 0 for all a ∈ A and such that limλ∈Λ ∆A(mλ) = 1. (2) A is symmetrically amenable if and only if there is a bounded net (mλ)λ∈Λ in Ab⊗A such that lim λ∈Λ(cid:0)(a ⊗ b)mλ − mλ(b ⊗ a)(cid:1) = 0 for all a, b ∈ A and such that limλ∈Λ ∆A(mλ) = limλ∈Λ ∆op A (mλ) = 1. Proof. Part (1) is Theorem 2.2.4 of [16]. We have taken advantage of the fact that A is unital to rewrite the definition of the module structures on Ab⊗A used there in terms of the multiplication in Ab⊗A and to simplify the condition involving ∆A. With the same modifications, Part (2) follows from Proposition 2.2 of [11]. (In [11], the commutation relation in (2) is stated separately for a ⊗ 1 and 1 ⊗ b, but the combination of those is clearly equivalent to our condition.) (cid:3) We will need the diagonal in Md from Example 2.2.3 of [16] (used in Example 2.3.16 of [16]), and called m there. Lemma 2.5. Let d ∈ Z>0, let (ej,k)j,k=1,2,...,d be the standard system of matrix units for Md, and let yd ∈ Md ⊗ Md be given by yd = 1 r,s=1 er,s ⊗ es,r. Then: dPd (1) y2 d = d−2 · 1. (2) yd(a ⊗ b)y−1 (3) ∆Md (yd) = ∆op Md d = b ⊗ a for all a, b ∈ Md. (yd) = 1. 8 N. CHRISTOPHER PHILLIPS (4) For any finite subgroup G ⊂ inv(Md) whose natural action on Cd is irre- ducible, we have yd = 1 card(G)Xg∈G g ⊗ g−1. (5) If x ∈ Md satisfies x(a ⊗ b) = (b ⊗ a)x for all a, b ∈ Md and ∆Md (x) = 1, then x = yd. (6) Let ϕ : Md → Md be an automorphism. Then (ϕ ⊗ ϕ)(yd) = yd. (7) For every p ∈ [1, ∞), we have kydkπ = 1 in M p (8) For every p ∈ [1, ∞), we have kydkp = d−1 in M p (9) Let d1, d2 ∈ Z>0. Let db⊗M p d ⊗p M p d . d . ϕ : (Md1 ⊗ Md2) ⊗ (Md1 ⊗ Md2) → (Md1 ⊗ Md1) ⊗ (Md2 ⊗ Md2) be the homomorphism determined by ϕ(a1 ⊗ a2 ⊗ b1 ⊗ b2) = a1 ⊗ b1 ⊗ a2 ⊗ b2 for a1, b1 ∈ Md1 and a2, b2 ∈ Md2. Then ϕ(yd1d2) = yd1 ⊗ yd2. Proof. Parts (1) and (3) are computations. Part (1) shows that yd is invertible. Part (2) is then also a computation, best done by taking both a and b to be standard matrix units. For (5), let x be as there. Then y−1 d x commutes with every element of Md ⊗ Md by (2). Since this algebra has trivial center, it follows that there is λ ∈ C such that x = λyd. Then λ = λ∆Md (yd) = ∆Md (x) = 1. To prove (6), check that (ϕ ⊗ ϕ)(yd) satisfies the conditions on x in (5). (For the second condition, one will need Lemma 2.3.) Now let G be as in part (4). Set x = Pg∈G g ⊗ g−1. Then, using (2) at the second step and the fact that G is a group at the third step, we have (2.1) ydx =Xg∈G yd(g ⊗ g−1) =Xg∈G (g−1 ⊗ g)yd = xyd. We regard G × G as a subgroup of inv(Md ⊗ Md) via (h, k) 7→ h ⊗ k for h, k ∈ G. Let h, k ∈ G. Then, using the change of variables g 7→ kg−1h−1 at the third step and (2.1) at the last step, we have ydx(h ⊗ k) =Xg∈G =Xg∈G yd(gh ⊗ g−1k) (g−1k ⊗ gh)yd =Xg∈G (hg ⊗ kg−1)yd = (h ⊗ k)xyd = (h ⊗ k)ydx. The natural action of G × G on Cd ⊗ Cd is irreducible by Corollary 2.9 of [14]. Therefore ydx is a scalar. So x = yd(ydx) is a scalar multiple of yd. Computing ∆(x) = card(G) · 1, we get x = card(G)yd. We now prove part (7). Since ∆Md (yd) = 1 and k∆Md k ≤ 1, it is clear that kydkπ ≥ 1. For the reverse inequality, let G ⊂ inv(Md) be the group of signed permutation matrices. Using Lemma 2.11 of [14] and part (4), we get kydkπ =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 card(G) Xg∈G g ⊗ g−1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)π ≤ 1 card(G)Xg∈G kgk · kg−1k = 1. This proves (7). For (8), we identify M p of [14]. Then d · yd becomes a permutation matrix, so that kd · ydk = 1. d ⊗p M p d = M p d2 following Corollary 1.13 Part (9) follows easily by verifying the conditions in part (5). (cid:3) Definition 2.6. Let (X, B, µ) be a σ-finite measure space, let p ∈ [1, ∞), and let A ⊂ L(Lp(X, µ)) be a unital subalgebra. AMENABILITY 9 for all a ∈ A. (1) We say that A has approximately inner Lp-tensor half flip (with con- stant M ) if there exists a net (vλ)λ∈Λ in inv(A ⊗p A) such that kvλk ≤ M λ = 1 ⊗ a and(cid:13)(cid:13)v−1 λ (cid:13)(cid:13) ≤ M for all λ ∈ Λ, and such that limλ∈Λ vλ(a ⊗ 1)v−1 (cid:13)(cid:13)v−1 λ (cid:13)(cid:13) ≤ M for all λ ∈ Λ, and such that limλ∈Λ vλ(a ⊗ b)v−1 (2) We say that A has approximately inner Lp-tensor flip (with constant M ) if there exists a net (vλ)λ∈Λ in inv(A ⊗p A) such that kvλk ≤ M and λ = b ⊗ a for all a, b ∈ A. We caution that, except for C*-algebras in the case p = 2, the conditions in Def- inition 2.6 presumably depend on how the algebra A is represented on an Lp-space, and are not intrinsic to A. It may well turn out that if they hold for some algebra A, and if ϕ : A → B is an isomorphism such that ϕ and ϕ−1 are completely bounded in a suitable Lp operator sense, then they hold for B. (Also see Proposition 2.9 be- low.) Pursuing this idea is beyond the scope of this paper. As an easily accessible substitute for completely bounded isomorphism, we use the condition that there be an isomorphism ψ : A ⊗p A → B ⊗p A such that ψ(a1 ⊗ a2) = ϕ(a1) ⊗ a2 for all a1, a2 ∈ A. The conditions in Definition 2.6 are very strong. See [5] for the severe restrictions that having an approximately inner C* tensor flip places on a C*-algebra; the machinery used there has been greatly extended since. We are interested these conditions because of the following consequence. The hypotheses are strong because we do not yet have a proper general theory of tensor products of algebras on Lp spaces. Theorem 2.7. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞), and let A ⊂ L(Lp(X, µ)) and B ⊂ L(Lp(Y, ν)) be closed unital subalgebras such that A has approximately inner Lp-tensor half flip. Let ϕ : A → B be a unital homomorphism, and suppose that there is a continuous homomorphism ψ : A ⊗p A → B ⊗p A such that ψ(a1 ⊗ a2) = ϕ(a1) ⊗ a2 for all a1, a2 ∈ A. Then ϕ is injective and bounded below, the range of ϕ is closed, and ϕ is a Banach algebra isomorphism from A to its range. Proof. It suffices to find C > 0 such that kϕ(a)k ≥ Ckak for all a ∈ A. By hypothe- sis, there are M ∈ [1, ∞) and a net (vλ)λ∈Λ in inv(A ⊗p A) satisfying the conditions in Definition 2.6(1). Set C = M −2kψk−2. For all a ∈ A, we have lim λ∈Λ ψ(vλ)(ϕ(a) ⊗ 1A)ψ(cid:0)v−1 λ (cid:1) = ψ(cid:18)lim λ∈Λ Therefore vλ(a ⊗ 1A)v−1 λ (cid:19) = ψ(1A ⊗ a) = 1B ⊗ a. kak = k1B ⊗ ak ≤ (kψkM ) · kϕ(a) ⊗ 1k · (kψkM ) = M 2kψk2kϕ(a)k. That is, kϕ(a)k ≥ Ckak. (cid:3) We give two permanence properties for approximately inner Lp-tensor (half) flip. Proposition 2.8. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞), and let A ⊂ L(Lp(X, µ)) and B ⊂ L(Lp(Y, ν)) be closed unital subalgebras. Suppose that A and B have approximately inner Lp-tensor (half) flip. Then the same is true of A ⊗p B. Proof. Let (vλ)λ∈Λ1 and (wλ)λ∈Λ2 be nets as in the appropriate part of Defini- tion 2.6 for A and B. Using Theorem 2.6(5) of [13], one checks that the net (vλ1 ⊗ wλ2 )(λ1,λ2)∈Λ1×Λ2 satisfies the condition in the appropriate part of Defi- nition 2.6 for A ⊗ B. (cid:3) 10 N. CHRISTOPHER PHILLIPS Proposition 2.9. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞), and let A ⊂ L(Lp(X, µ)) and B ⊂ L(Lp(Y, ν)) be closed unital subalgebras. Suppose that there is a continuous homomorphism ϕ : A → B with dense range such that the algebraic tensor product of two copies of ϕ extends to a continuous homomorphism ϕ ⊗p ϕ : A ⊗p A → B ⊗p B. If A has approximately inner Lp-tensor (half) flip, then so does B. Proof. We give the proof for the tensor flip; the proof for the tensor half flip is the same. Let (vλ)λ∈Λ be a net as in Definition 2.6(2). For λ ∈ Λ, set wλ = (ϕ⊗pϕ)(vλ). Then (wλ)λ∈Λ and (w−1 λ → b⊗a for all a, b ∈ ϕ(A). Density of ϕ(A) and boundedness of (wλ)λ∈Λ and (w−1 λ )λ∈Λ allows us to use an ε λ → b ⊗ a for all a, b ∈ B. (cid:3) λ )λ∈Λ are bounded nets in B ⊗p B. We have wλ(a⊗b)w−1 3 argument to conclude that wλ(a ⊗ b)w−1 A is symmetrically amenable. let σn : Mrd(n) → A be a unital homomorphism whose range is Dn. (2) Let p ∈ [1, ∞), let (X, B, µ) be a σ-finite measure space, and suppose that Proposition 2.10. Let d be a sequence in {2, 3, 4, . . .}. Let rd be as in Defini- tion 1.1. Let A be a unital Banach algebra, let D0 ⊂ D1 ⊂ · · · ⊂ A be an n=0 Dn, and for n ∈ Z>0 increasing sequence of unital subalgebras such that A =S∞ (1) Suppose that, in Ab⊗A, we have supn∈Z>0(cid:13)(cid:13)(cid:0)σnb⊗σn(cid:1)(yrd(n))(cid:13)(cid:13)π < ∞. Then A ⊂ L(Lp(X, µ)) and that, in A ⊗p A, we have supn∈Z>0 rd(n)(cid:13)(cid:13)(σn ⊗p σn)(yrd(n))(cid:13)(cid:13) < ∞. Then A has approximately inner Lp-tensor flip. (cid:0)σnb⊗σn(cid:1)(yrd(n)) for n ∈ Z>0. The hypotheses imply that (zn)n∈Z>0 is bounded in Ab⊗A. Lemma 2.5(3) and Lemma 2.3 imply that ∆A(zn) = ∆op n→∞(cid:2)zn(a ⊗ b) − (b ⊗ a)zn(cid:3) = 0 n=0 Dn. Since (zn)n∈Z>0 is bounded andS∞ for all a, b ∈S∞ Proof. We prove (1), by verifying the conditions in Proposition 2.4(2). Set zn = It follows from Lemma 2.5(2) that zn(a ⊗ b) = (b ⊗ a)zn for all a, b ∈ Dn. follows that (2.2) holds for all a, b ∈ A. This completes the proof. For (2), we verify the condition in Definition 2.6(2). Set Therefore (2.2) lim A (zn) = 1. n=0 Dn is dense in A, it vn = rd(n)(σn ⊗p σn)(yrd(n)) n = vn for all n ∈ Z>0, so (v−1 for n ∈ Z>0. By hypothesis, (vn)n∈Z>0 is bounded in A ⊗p A. Lemma 2.5(1) implies that v−1 It follows from Lemma 2.5(2) that vn(a ⊗ b)v−1 n = b ⊗ a for all a, b ∈ Dn. Arguing as at the end of the proof of (1), we get limn→∞ vn(a ⊗ b)v−1 (cid:3) n )n∈Z>0 is also bounded. n = b ⊗ a for all a, b ∈ A. The construction in Example 1.4 requires probability measures, so that infinite products make sense. This causes problem when passing to subspaces. We state for reference the solution we usually use. Lemma 2.11. Let (X, B, µ) be a measure space, let p ∈ [1, ∞), and let α ∈ (0, ∞). Then Lp(X, µ) and Lp(X, αµ) are equal as vector spaces, but with kξkLp(X,αµ) = α1/pkξkLp(X,µ) for ξ ∈ Lp(X, µ). Moreover, the identity map from L(Lp(X, µ)) to L(Lp(X, αµ)) is an isometric isomorphism. Proof. All statements in the lemma are immediate. (cid:3) Definition 2.12. Let the notation be as in Example 1.4. A subsystem of (d, ρ) consists of subspaces Zn ⊂ Xn such that µn(Zn) > 0 and Lp(Zn, µn) is an invariant subspace for ρn for all n ∈ Z>0. AMENABILITY 11 For n ∈ Z>0, set λn = µn(Zn)−1µnZn and let ιn : L(Lp(Zn, µn)) → L(Lp(Zn, λn)) be the isometric isomorphism of Lemma 2.11. The Lp UHF algebra of the subsystem (Zn)n∈Z>0 is the one constructed as in Example 1.4 using the same sequence d and the sequence γ of representations given by γn(x) = ιn(cid:0)ρn(x)Lp(Zn,µn)(cid:1) for n ∈ Z>0 and x ∈ Md(n). Lemma 2.13. Let the notation be as in Definition 2.12 and Example 1.4. Set A = A(d, ρ) and B = A(d, γ). Let σn : Mrd(n) → A be as at the end of Example 1.4, and let τn : Mrd(n) → B be the analogous map for (d, γ). Then there is a unique contractive unital homomorphism κ : A → B such that κ ◦ σn = τn for all n ∈ Z>0. Moreover, κ has the following properties: (1) κ has dense range. (2) Whenever (Y, C, ν) is a σ-finite measure space and D ⊂ L(Lp(Y, C, ν)) is a closed unital subalgebra, there is a contractive homomorphism κD : A ⊗p D → B ⊗p D such that κD(a ⊗ x) = κ(a) ⊗ x for all a ∈ A and x ∈ D. To get the map κ, we really only need kτn(x)k ≤ kσn(x)k for all n ∈ Z≥0 and all x ∈ Mrd(n). We need the subsystem condition to get (2). Proof of Lemma 2.13. We construct the homomorphism κD in (2) for arbitrary D. We then get κ by taking D = C. We continue to follow the notation of Definition 2.12 and Example 1.4. In particular, for S ⊂ N finite, we let XS and its measure µS be as in Example 1.4. We abbreviate AN≤n to An, and set αn = ρN≤n,N≤n+1 : An → An+1, so that αn is isometric and A is isometrically isomorphic to lim An. Analogously, define ZS = −→n Qk∈S Zk, let λS be the corresponding product measure, set Bn = BN≤n , and set Bn. Set cS =Qk∈S µk(Zk)−1, so that λS × ν = βn = γN≤n,N≤n+1, giving B = lim −→n cSµS × ν. Let ιS : L(Lp(ZS × Y, µS × ν)) → L(Lp(ZS × Y, λS × ν)) be the identification of Lemma 2.11, and define κD,n : An ⊗p D → Bn ⊗p D by for a ∈ An ⊗p D. κD,n(a) = ιN≤n(cid:0)aLp(ZN≤n ×Y, µN≤n ×ν)(cid:1) One checks immediately that κD,n is contractive for all n ∈ Z≥0. Since An is finite dimensional, κD,n is also bijective. Moreover, the following diagram (not including the dotted arrow) now clearly commutes: A0 ⊗p D κD,0 B0 ⊗p D α0⊗pidD β0⊗pidD A1 ⊗p D κD,1 / B1 ⊗p D α1⊗pidD A2 ⊗p D · · · / A ⊗p D κD,2 κD β1⊗pidD / B2 ⊗p D / · · · / B ⊗p D. For n ∈ Z>0, there is a homomorphism An ⊗p D → A ⊗p D given by a ⊗ x 7→ a ⊗ 1N≤n ⊗ x for a ∈ An and x ∈ D. Except for the fact that the domain is now taken to be normed, this homomorphism is just σn ⊗ idD. It is isometric by Theorem 2.16(5) of [13]. Thus we can identify A ⊗p D with lim An ⊗p D. Similarly −→n Bn ⊗p D, using the maps which become, after we can identify B ⊗p D with lim −→n forgetting the norms, τn ⊗ idD. Commutativity of the part of diagram with solid arrows, contractivity of κD,n, and the fact that κD,n becomes the identity when the norms are forgotten, therefore imply the existence of a contractive homomorphism κD : A ⊗p D → B ⊗p D such that κD ◦ (σn ⊗ idD) = τn ⊗ idD for all n ∈ Z>0. Its (cid:3) range is dense because it containsS∞ n=0 Bn ⊗ D. / /   / /   / /   /   ✤ ✤ ✤ / / / / 12 N. CHRISTOPHER PHILLIPS Corollary 2.14. Let p ∈ [1, ∞), and let A be an Lp UHF algebra of tensor product type which dominates the spatial representation in the sense of Definition 1.7(4). Let B be the spatial Lp UHF algebra whose supernatural number N is the same as that of A, as in Theorem 1.8. Then there exists a contractive unital homomorphism κ : A → B with the following properties: (1) κ has dense range. (2) Let σn : Mrd(n) → A be as at the end of Example 1.4, and identify Mrd(n) with M p rd(n). Then κ ◦ σn is isometric for all n ∈ Z≥0. (3) Whenever (Y, C, ν) is a σ-finite measure space and D ⊂ L(Lp(Y, C, ν)) is a closed unital subalgebra, there is a continuous homomorphism κD : A ⊗p D → B ⊗p D such that κD(a ⊗ x) = κ(a) ⊗ x for all a ∈ A and x ∈ D. Proof. Adopt the notation of Example 1.4. We apply Lemma 2.13, for n ∈ Z>0 tak- ing Zn ⊂ Xn to be a subspace, whose existence is guaranteed by Definition 1.7(4), such that Lp(Zn, µn) is invariant under ρn and a 7→ ρn(a)Lp(Zn,µn) is spatial. Ev- erything except (2) is immediate. In part (2), we have to identify the algebra B⊗pD which appears here with the algebra B ⊗p D in Lemma 2.13. We first observe that κ ◦ σn is spatial for all n ∈ Z≥0. Thus, the conclusion is correct for some unital isometric representation of B on an Lp space. By the last part of Theorem 1.8, the tensor product B ⊗p D is the same for any unital isometric representation of B on an Lp space. (cid:3) Lemma 2.15. Let (αn)n∈Z>0 be a sequence in [1, ∞). ThenP∞ and only ifQ∞ n=1 αn < ∞. Proof. If supn∈Z>0 αn = ∞, then both statements fail. Otherwise, set βn = αn − 1 for n ∈ Z>0, and set M = supn∈Z>0 βn. Then for all n ∈ Z>0 we have 0 ≤ βn ≤ M. Using concavity of β 7→ log(1 + β) on the interval [0, M ] for the first inequality, we get n=1(αn − 1) < ∞ if M −1 log(M + 1)βn ≤ log(1 + βn) ≤ βn. n=1 log(1+βn) < ∞, which clearly happens (cid:3) ThereforeP∞ if and only ifQ∞ n=1 βn < ∞ if and only ifP∞ n=1(1 + βn) < ∞. Theorem 2.16. Let p ∈ [1, ∞). Let A be an Lp UHF algebra of tensor product type (Definition 1.7(1)), and let d, ρ = (ρ1, ρ2, . . .), and σn : Mrd(n) → A be as there. Assume that A dominates the spatial representation (Definition 1.7(4)); in particular, that A locally has enough isometries (Definition 1.7(3)). Let B be the spatial Lp UHF algebra whose supernatural number is the same as that of A, as in Theorem 1.8. Then the following are equivalent: (1) There exists an isomorphism ϕ : A → B such that the algebraic tensor product of two copies of ϕ extends to an isomorphism ϕ ⊗p ϕ : A ⊗p A → B ⊗p B. (2) A ⊗p A has approximately inner tensor half flip. (3) A has approximately inner Lp-tensor flip. (4) A has approximately inner Lp-tensor half flip. (5) There is a uniform bound on the norms of the homomorphisms σn ⊗p σn : M p rd(n)2 → A ⊗p A. (6) The homomorphisms ρn ⊗p ρn : M p d(n)2 → A ⊗p A satisfy (kρn ⊗p ρnk − 1) < ∞. ∞Xn=1 Moreover, these conditions imply the following: AMENABILITY 13 (7) A is simple. (8) P∞ n=1(kρnk − 1) < ∞. (9) A is symmetrically amenable. (10) A is amenable. Proof. It is clear that (3) implies (4). That (4) implies (2) is Proposition 2.8. We show that (2) implies both (1) and (5). Let κ : A → B be as in Corollary 2.14. Corollary 2.14(3) provides continuous homomorphisms γ1 : A ⊗p A → B ⊗p A and γ2 : B ⊗p A → B ⊗p B such that γ1(a ⊗ b) = κ(a) ⊗ b for all a, b ∈ A and γ2(b ⊗ a) = b ⊗ κ(a) for all a ∈ A and b ∈ B. Then κ = γ2 ◦ γ1 : A ⊗p A → B ⊗p B is a continuous homomorphism such that κ(a ⊗ b) = κ(a) ⊗ κ(b) for all a, b ∈ A. We next claim that whenever (Y, C, ν) is a σ-finite measure space and D ⊂ L(Lp(Y, C, ν)) is a closed unital subalgebra, there is a continuous homomorphism β : A ⊗p A ⊗p D → B ⊗p B ⊗p D such that β(a ⊗ b ⊗ x) = κ(a) ⊗ κ(b) ⊗ x for all a, b ∈ A and x ∈ D. To prove this, apply Corollary 2.14(3) with A ⊗p D in place of D, and apply Corollary 2.14(3), after changing the order of the tensor factors, with B ⊗p D in place of D, to get continuous homomorphisms β1 : A ⊗p A ⊗p D → B ⊗p A ⊗p D and β2 : B ⊗p A ⊗p D → B ⊗p B ⊗p D, such that β1(a ⊗ b ⊗ x) = κ(a) ⊗ b ⊗ x for all a, b ∈ A and x ∈ D, and such that β2(b ⊗ a ⊗ x) = b ⊗ κ(a) ⊗ x for all a ∈ A, and b ∈ B, and x ∈ D. Then set β = β2 ◦ β1. This proves the claim. Apply the claim with D = A ⊗p A, obtaining β : A ⊗p A ⊗p A ⊗p A → B ⊗p B ⊗p A ⊗p A such that β(a ⊗ x) = κ(a) ⊗ x for all a, x ∈ A ⊗p A. Theorem 2.7 provides C > 0 such that kκ(a)k ≥ Ckak for all a ∈ A ⊗p A. The map κ clearly has dense range. Therefore κ is an isomorphism. Moreover, for a ∈ A we have kκ(a)k = kκ(a) ⊗ 1Bk = kκ(a ⊗ 1A)k ≥ Cka ⊗ 1Ak = Ckak. Since κ has dense range, it follows that κ is also an isomorphism. Part (1) is proved. To get (5), observe that the choice of κ and Corollary 2.14(2) imply that κ ◦ σn is isometric for all n ∈ Z≥0. Corollary 1.13 of [14] now implies that κ ◦ (σn ⊗ σn) = (κ ◦ σn) ⊗ (κ ◦ σn) is isometric for all n ∈ Z≥0. Therefore supn∈Z≥0 kσn ⊗p σnk ≤ C−1, which is (5). Using Corollary 1.13 of [14] to see that the maps M p rd(n) → B ⊗p B are isometric, it follows from Proposition 2.10(2) and Lemma 2.5(8) that B has approximately inner tensor flip. The implication from (1) to (3) is then Proposi- tion 2.9. rd(n) ⊗p M p Next assume (5); we prove (6). For n ∈ Z>0, by suitably permuting tensor factors and using Lemma 1.11 of [14] and Corollary 1.13 of [14], we can make the identification M p rd(n)2 = M p d(1)2 ⊗p M p d(2)2 ⊗p · · · ⊗p M p d(n)2. Then nYk=1 by Lemma 1.14 of [14]. ThusQ∞ by Lemma 2.15. kρk ⊗p ρkk ≤ kσn ⊗p σnk k=1 kρk ⊗p ρkk < ∞. SoP∞ n=1(cid:0)kρn ⊗p ρnk − 1(cid:1) < ∞ 14 N. CHRISTOPHER PHILLIPS Next, we show that (6) implies (3). First,Q∞ k=1 kρk ⊗p ρkk < ∞ by Lemma 2.15. By Lemma 2.5(9), up to a permutation of tensor factors (which is isometric by Lemma 1.11 of [14]), for d1, d2 ∈ Z>0 we have yd1 ⊗ yd2 = yd1d2. Therefore, using the tensor product decompositions in Example 1.4 on L2(X × X, µ × µ), we get rd(n)(σn ⊗p σn)(yrd(n)) = d(1)(ρ1 ⊗p ρ1)(yd(1)) ⊗ d(2)(ρ2 ⊗p ρ2)(yd(2)) ⊗ · · · ⊗ d(n)(ρn ⊗p ρn)(yd(n)) ⊗ (1N≤n ⊗ 1N≤n ). So Theorem 2.16(5) of [13] and Lemma 2.5(8) imply that sup n∈Z≥0 rd(n)k(σn ⊗p σn)(yrd(n))k ≤ kρk ⊗p ρkk < ∞. Thus A has approximately inner Lp-tensor flip by Proposition 2.10(2). ∞Yk=1 We have now completed the proof of the equivalence of conditions (1), (2), (3), It remains to show that these conditions imply (7), (8), (9), (4), (5), and (6). and (10). That (1) implies (7) is Theorem 3.13 of [14]. Assume (6); we prove (8). We have kρn ⊗p ρnk ≥ kρnk2 by Lemma 1.13 of [14]. n=1 kρnk2 < ∞, whence n=1 kρn ⊗p ρnk < ∞, so Q∞ n=1(kρnk − 1) < ∞ by Lemma 2.15. Assume (1); we prove (9). Since symmetric amenability only depends on the isomorphism class of a Banach algebra, it is enough to prove symmetric amenability when A = B. The maps σn : M p rd(n) → B are then isometric, so that the maps Lemma 2.15 implies that Q∞ Q∞ n=1 kρnk < ∞. SoP∞ σnb⊗σn : M p rd(n)b⊗M p and Lemma 2.5(7). Condition (10) follows from (9). rd(n) → Ab⊗A are contractive. Now apply Proposition 2.10(1) (cid:3) 3. Lp UHF algebras constructed from similarities In this section, we describe and derive the properties of a special class of Lp UHF algebras of tensor product type. In Example 1.4, for n ∈ Z>0 we will require that the representation ρn : Md(n) → L(Lp(Xn, µn)) be a finite or countable direct sum of representations of the form x 7→ sxs−1 for invertible elements s ∈ M p d(n). The results of Section 4 and Section 5 are for algebras in this class. One of the main purposes of this section is to set up notation. The definition of a system of d-similarities is intended only for local use, to simplify the description of our construction. Notation 3.1. For d ∈ Z>0, we let (ej,k)j,k=1,2,...,d be the standard system of matrix units for Md. Definition 3.2. Let d ∈ Z>0. A system of d-similarities is a triple S = (I, s, f ) consisting of a countable (possibly finite) index set I, a function s : I → inv(Md), and a function f : I → [0, 1] such that: (1) 1 ∈ ran(s). (2) ran(s) ⊂ inv(Md). (3) ran(s) is compact. (4) f (i) > 0 for all i ∈ I. (5) Pi∈I f (i) = 1. We say that s is diagonal if, in addition: (6) s(i) is a diagonal matrix for all i ∈ I. The basic system of d-similarities is the system S0 = (I0, s0, f0) given by I0 = {0}, s0(0) = 1, and f0(0) = 1. AMENABILITY 15 We will associate representations to systems of d-similarities, but we first need some lemmas. Lemma 3.3. Let E be a Banach space, and let s ∈ L(E) be invertible. Define α : L(E) → L(E) by α(a) = sas−1 for all a ∈ L(E). Then kαk = ksk · ks−1k. Proof. It is clear that kαk ≤ ksk ·ks−1k. We prove the reverse inequality. Let ε > 0. Choose δ > 0 such that (ksk − δ)(ks−1k − δ) > ksk · ks−1k − ε. Choose η ∈ E such that kηk = 1 and ksηk > ksk−δ. Choose µ ∈ E such that kµk = 1 and ks−1µk > ks−1k − δ. Use the Hahn-Banach Theorem to find a continuous linear functional ω on E such that kωk = 1 and ω(s−1µ) = ks−1µk. Define a ∈ L(E) by aξ = ω(ξ)η for all ξ ∈ E. Then kak = 1. Also, sas−1µ = s(ω(s−1µ))η = ks−1µk · sη. Therefore ksas−1µk = ks−1µk · ksηk > (ksk − δ)(ks−1k − δ) > ksk · ks−1k − ε. So kα(a)k ≥ ksas−1k > ksk · ks−1k − ε, whence kαk > ksk · ks−1k − ε. Since ε > 0 is arbitrary, it follows that kαk ≥ ksk · ks−1k. (cid:3) Corollary 3.4. Let d ∈ Z>0, let (I, s, f ) be a system of d-similarities, and let p ∈ [1, ∞). For i ∈ I, let ψi : Md → M p d be the representation ψi(a) = s(i)as(i)−1 for a ∈ Md. Then supi∈I kψi(a)kp is finite for all a ∈ Md. Proof. Set M = sup(cid:0)(cid:8)kvkpkv−1kp : v ∈ ran(s)(cid:9)(cid:1). Then M < ∞ because ran(s) is a compact subset of inv(Md). Identify Md with M p (cid:3) d . Then Lemma 3.3 implies supi∈I kψik ≤ M. We need infinite direct sums of representations of algebras on Lp spaces. (The finite case is given in Lemma 2.14 of [13], and the infinite case is discussed in Remark 2.15 of [13] and implicitly used in Example 2.14 of [14].) Since we work with σ-finite measure spaces, we must restrict to countable direct sums. We must require p 6= ∞ since the infinite direct sum of L∞ spaces is usually not L∞ of the disjoint union. Lemma 3.5. Let A be a unital complex algebra, and let p ∈ [1, ∞). Let I be a countable index set, and for i ∈ I let (Xi, Bi, µi) be a σ-finite measure space and let πi : A → L(Lp(Xi, µi)) be a unital homomorphism. Assume that supi∈I kπi(a)k is finite for every a ∈ A. Let E0 be the algebraic direct sum over i ∈ I of the spaces Lp(Xi, µi). Equip E0 with the norm k(ξi)i∈I k =(cid:16)Xi∈I kξikp p(cid:17)1/p , π(a)(cid:0)(ξi)i∈I(cid:1) =(cid:0)πi(a)ξi(cid:1)i∈I and let E be the completion of E0 in this norm. Then E ∼= Lp(cid:0)`i∈I Xi(cid:1) , using the obvious disjoint union measure, and there is a unique representation π : A → L(E) such that for a ∈ A and (ξi)i∈I ∈ E0. Moreover, kπ(a)k = supi∈I kπi(a)k for all a ∈ A. Proof. This is immediate. (cid:3) Definition 3.6. Let the notation and hypotheses be as in Lemma 3.5. We call E the Lp direct sum of the spaces Lp(Xi, µi), and we call π the Lp direct sum of the representations πi for i ∈ I. We write E =Li∈I Lp(Xi, µi) and π =Li∈I πi, letting p be understood. 16 N. CHRISTOPHER PHILLIPS Remark 3.7. We can form direct sums of bounded families of elements as well as of representations. (Just take A to be the polynomial rang C[x].) Then, for example, if ai ∈ L(Lp(Xi, µi)) is invertible for all i ∈ I, and supi∈I kaik and supi∈I ka−1 i k are . Moreover, both finite, then a =Li∈I ai is invertible with inverse a−1 =Li∈I a−1 kak = supi∈I kaik and ka−1k = supi∈I ka−1 i k. (Take A = C[x, x−1].) i We need to know that the sum of orthogonal spatial partial isometries is again a spatial partial isometry. This result is related to those of Section 6 of [13], but does not appear there. We refer to Section 6 of [13] for the terminology. Here we only need the application to direct sums (Corollary 3.9), but the more general statement will be needed elsewhere. Lemma 3.8. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞], domain support Ej ⊂ X, with range support Fj ⊂ Y, with spatial realization Sj, with phase factor gj, and with reverse tj. Suppose that the sets Ej are disjoint up to sets of measure zero, and that the sets Fj are disjoint up to sets of measure zero. and for j ∈ Z>0 let sj ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be a spatial partial isometry with Then there exists a unique spatial partial isometry s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) with domain support E =S∞ j=1 Ej and with range support F =S∞ sLp(Ej ,µ) = sjLp(Ej ,µ) for all j ∈ Z>0. Its spatial realization S is given by j=1 Fj , and such that S(B) = Sj(B ∩ Ej ) ∞[j=1 for B ∈ B with B ⊂ E. Its phase factor g is given almost everywhere by g(y) = gj(y) for y ∈ Fj. Its reverse t is obtained in the same manner as s using the tj in place j=1 sjξ, with weak* of the sj. Moreover, for all ξ ∈ Lp(X, µ), we have sξ = P∞ convergence for p = ∞ and norm convergence for p ∈ [1, ∞). Proof. Everything is easy to check directly. (cid:3) Corollary 3.9. Let p ∈ [1, ∞], and let I be a countable set. For i ∈ I let (Xi, Bi, µi) be a spatial partial isometry with domain support Ei ⊂ Xi, with range support Fi ⊂ Yi, with spatial realization Si, with phase factor gi, and with reverse ti. Then and (Yi, Ci, νi) be σ-finite measure spaces, and let si ∈ L(cid:0)Lp(Xi, µi), Lp(Yi, νi)(cid:1) s =Li∈I si (as in Remark 3.7) is a spatial partial isometry with domain support `i∈I Ei, with range support `i∈I Fi, and with reverseLi∈I ti. The phase factor g is determined by gEi = gi for i ∈ I, and the spatial realization S is determined by S(B) = Si(B) for B ∈ Bi with B ⊂ Ei. Proof. Apply Lemma 3.8. The disjointness condition is trivial. (cid:3) Lemma 3.10. Let the notation and hypotheses be as in Lemma 3.5. If πi is spatial for every i ∈ I, thenLi∈I πi is spatial. The following definition is Example 2.15 of [14]. Proof. Use Corollary 3.9 to verify the criterion in Proposition 1.5. (cid:3) Definition 3.11. Let d ∈ Z>0, let S = (I, s, f ) be a system of d-similarities, and let p ∈ [1, ∞). We define XS = XI,s,f = {1, 2, . . . , d} × I, with the σ-algebra BS consisting of all subsets of XS and the atomic measure µS = µI,s,f determined by µI,s,f ({(j, i)}) = d−1f (i) for j = 1, 2, . . . , d and i ∈ I. For i ∈ I, we make the obvious identification (Lemma 2.11) (3.1) We then define L(cid:0)Lp({1, 2, . . . , d} × {i}, µS)(cid:1) = M p ψi : Md → L(cid:0)Lp({1, 2, . . . , d} × {i}, µS)(cid:1) d . AMENABILITY 17 by ψi(a) = s(i)as(i)−1 for a ∈ Md. The representation associated with (p, S) is then the unital homomorphism ψp,S = ψp,I,s,f : Md → L(cid:0)Lp(XS, µS)(cid:1) obtained as the Lp direct sum over i ∈ I, as in Definition 3.6, of the representations ψi, and whose existence is ensured by Corollary 3.4 and Lemma 3.5. Define M p,S d = M p,I,s,f d = ψp,S(Md) ⊂ L(cid:0)Lp(XS, µS)(cid:1). We further define the p-bound of S to be Rp,S = Rp,I,s,f = sup i∈I ks(i)kpks(i)−1kp. Lemma 3.12. Let d ∈ Z>0, let S be a system of d-similarities, and let p ∈ [1, ∞). Further let (Y, C, ν) be a σ-finite measure space, and let B ⊂ L(Lp(Y, ν)) be a unital closed subalgebra. Then the assignment x ⊗ b 7→ ψp,S(x) ⊗ b defines a bijective homomorphism ψp,S B from the algebraic tensor product Md ⊗alg B to Proof. The statement is immediate from the fact that Md is finite dimensional. (cid:3) M p,S d ⊗p B ⊂ L(cid:0)Lp(XS × Y, µI,s,f × ν)(cid:1). Notation 3.13. Let the notation be as in Lemma 3.12. For x ∈ Md ⊗alg B, we define kxkp,S = kxkp,I,s,f = (cid:13)(cid:13)ψp,I,s,f defined kxkp,S for x ∈ Md. B (x)(cid:13)(cid:13). Taking B = C, we have in particular Notation 3.14. Let d ∈ Z>0, let S1 and S2 be systems of d-similarities, and let p ∈ [1, ∞). Further let (Y, C, ν) be a σ-finite measure space, and let B ⊂ L(Lp(Y, ν)) be a unital closed subalgebra. Using the notation of Lemma 3.12, we define κp,S2,S1 B by : M p,S1 d ⊗p B → M p,S2 d ⊗p B κp,S2,S1 B = ψp,S2 B ◦(cid:0)ψp,S1 B (cid:1)−1 . Lemma 3.15. Let d ∈ Z>0, and let p ∈ [1, ∞). Let S = (I, s, f ), S1 = (I1, s1, f1), S2 = (I2, s2, f2), and S3 = (I3, s3, f3) be systems of d-similarities, and let S0 = (I0, s0, f0) be the basic system of d-similarities (with s(0) = 1). Further let (Y, C, ν) be a σ-finite measure space, and let B ⊂ L(Lp(Y, ν)) be a unital closed subalgebra. Then: B B B (x)(cid:13)(cid:13) = supi∈I(cid:13)(cid:13)(s(i) ⊗ 1)ψp,S0 (1) (cid:13)(cid:13)ψp,S (2) (cid:13)(cid:13)κp,S,S0 (3) (cid:13)(cid:13)κp,S0,S (5) If ran(s2) ⊂ ran(s1) then(cid:13)(cid:13)κp,S2,S1 (cid:13)(cid:13) = Rp,S. (cid:13)(cid:13) = 1. (6) If ran(s1) = ran(s2) then κp,S2,S1 B (x)(s(i)−1 ⊗ 1)(cid:13)(cid:13) for all x ∈ Md ⊗alg B. (cid:13)(cid:13) = 1. is an isometric bijection. (4) κp,S3,S2 = κp,S3,S1 ◦ κp,S2,S1 B B B B B . Proof. Part (4) is immediate. Identify Md ⊗alg B with M p (The identification map is just ψp,S0 be the identity.) d ⊗p B, and write k · kp for the corresponding norm. to B . Thus, in the following, we are taking ψp,S0 B For i ∈ I, define a unital homomorphism ψi,B : M p d ⊗p B → L(cid:0)Lp(cid:0){1, 2, . . . , d} × {i} × Y, µS × ν(cid:1)(cid:1) by making the identification (3.1) and setting ψi,B(x) = (s(i) ⊗ 1)x(s(i)−1 ⊗ 1) for x ∈ M p d , we get kψi,Bk = ks(i)kpks(i)−1kp. d ⊗p B. Using Lemma 3.3 and ky ⊗ 1kp = kykp for all y ∈ M p 18 N. CHRISTOPHER PHILLIPS Parts (1) and (2) now both follow from the fact that ψp,S B (x) is the operator on Lp ai∈I {1, 2, . . . , d} × {i} × Y, µS × ν! therefore supi∈I kψi,B(x)kp. Moreover, we also have which acts on Lp(cid:0){1, 2, . . . , d} × {i} × Y, µS × ν(cid:1) as ψi,B(x), and whose norm is (cid:13)(cid:13) ≤ 1 in part (5) is now immediate, as for all x ∈ Md ⊗alg B. The inequality(cid:13)(cid:13)κp,S2,S1 is part (6). The inequality(cid:13)(cid:13)κp,S2,S1 B (x)(cid:13)(cid:13)p,S = sup(cid:0)(cid:8)k(v ⊗ 1)x(v−1 ⊗ 1)kp : v ∈ ran(s)(cid:9)(cid:1) (cid:13)(cid:13)ψp,S (cid:13)(cid:13) ≥ 1 in part (5) follows from κp,S2,S1 By Definition 3.2(1), part (3) is a special case of part (5). (1) = 1. (cid:3) B B B We get some additional properties for diagonal systems of similarities. Lemma 3.16. Let d ∈ Z>0, let S = (I, s, f ) be a diagonal system of d-similarities, let S0 be the basic system of d-similarities (Definition 3.2), and let p ∈ [1, ∞). For i ∈ I, define αi,1, αi,2, . . . , αi,d ∈ C \ {0} by s(i) = diag(αi,1, αi,2, . . . , αi,d). For j, k = 1, 2, . . . , d, define rj,k = supi∈I αi,j · αi,k−1. Let (Y, C, ν) be a σ-finite measure space, and let B ⊂ L(Lp(Y, ν)) be a unital closed subalgebra. Then: (1) (s(i) ⊗ 1)(ej,k ⊗ b)(s(i)−1 ⊗ 1) = αi,j α−1 i,k ej,k ⊗ b for i ∈ I, b ∈ B, and j, k = 1, 2, . . . , d. (2) 1 ≤ rj,k < ∞ for j, k = 1, 2, . . . , d. (3) rj,j = 1 for j = 1, 2, . . . , d. (4) kej,k ⊗ bkp,S = rj,kkbk for j, k = 1, 2, . . . , d and b ∈ B. (5) kej,j ⊗ bkp,S = kbk for j = 1, 2, . . . , d and b ∈ B. B (7) Let b1, b2, . . . , bd ∈ B. Then (cid:13)(cid:13) = Rp,S = sup(cid:0)(cid:8)rj,k : j, k ∈ {1, 2, . . . , d}(cid:1). (6) (cid:13)(cid:13)κp,S,S0 (cid:13)(cid:13)e1,1 ⊗ b1 + e2,2 ⊗ b2 + · · · + ed,d ⊗ ad(cid:13)(cid:13)p,S = max(cid:0)kb1k, kb2k, . . . , kbdk(cid:1). Proof. Part (1) is a calculation. In part (2), finiteness of rj,k follows from compact- ness of ran(s). The inequality rj,k ≥ 1 follows from max(cid:0)αi,j · αi,k−1, αi,j −1 · αi,k(cid:1) ≥ 1. Part (3) is clear. Part (4) follows from part (1), Lemma 3.15(1), and the fact that kej,kkp = 1 in M p d . Part (5) follows from part (4) and part (3). For part (6), by Lemma 3.15(2) it suffices to prove that This statement follows from the computation Rp,S = sup(cid:0)(cid:8)rj,k : j, k ∈ {1, 2, . . . , d}(cid:1). αi,k−1(cid:19) = sup αi,j!(cid:18) sup 1≤k≤d 1≤j≤d 1≤j,k≤d ks(i)kpks(i)−1kp = sup αi,j · αi,k−1 by taking the supremum over i ∈ I. We prove part (7). Set b = e1,1 ⊗ b1 + e2,2 ⊗ b2 + · · · + ed,d ⊗ bd. Regarded as an element of M p d ⊗p B, the element b is the Lp direct sum, over k ∈ {1, 2, . . . , d}, of the operators bk acting on Lp({k}×Y ), so kbkp = supk=1,2,...,d kbkk. Since S is diagonal, s(i) ⊗ 1 commutes with b for all i ∈ I. The result now follows from Lemma 3.15(1). (cid:3) AMENABILITY 19 Lemma 3.17. Let d1, d2 ∈ Z>0, let S1 = (I1, s1, f1) be a system of d1-similarities, let S2 = (I2, s2, f2) be a system of d2-similarities, and let p ∈ [1, ∞). Identify Cd1 ⊗ Cd2 with Cd1d2 via an isomorphism which sends tensor products of standard basis vectors to standard basis vectors, and use this isomorphism to identify Md1 ⊗Md2 = L(Cd1 ⊗ Cd2) with Md1d2 = L(Cd1d2). Set I = I1 × I2, and define s : I → Md1d2 and f : I → (0, 1] by s(i1, i2) = s1(i1) ⊗ s2(i2) and f (i1, i2) = f1(i1)f2(i2) for i1 ∈ I2 and i2 ∈ I2. Then S = (I, s, f ) is a system of d1d2-similarities and the same identification as already used becomes an isometric isomorphism M p,S1 ⊗p M p,S2 . We have Rp,S = Rp,S1 Rp,S2. Moreover, if S1 and S2 are diagonal, then so is S. → M p,S d1d2 d1 d2 Proof. The proof is straightforward, and is omitted, except that to prove that Rp,S = Rp,S1 Rp,S2, we need to know that kv1 ⊗ v2k = kv1k · kv2k for v1 ∈ L(Lp(XS1, µS1)) and v2 ∈ L(Lp(XS2 , µS2)). For this, we use Theorem 2.16(5) of [13]. (cid:3) Definition 3.18. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence of integers such that d(n) ≥ 2 for all n ∈ Z>0, and let I = (I1, I2, . . .), s = (s1, s2, . . .), and f = (f1, f2, . . .) be sequences such that Sn = (In, sn, fn) is a system of d(n)-similarities for all n ∈ Z>0. We define the (p, d, I, s, f )-UHF algebra Dp,d,I,s,f and associated sub- algebras by applying the construction of Example 1.4 using the representations ρn = ψp,Sn : Md(n) → L(Lp(XSn, µSn )). We let (XI,s,f , BI,s,f , µI,s,f ) be the prod- uct measure space (X, B, µ) from Example 1.4, that is, (XI,s,f , BI,s,f , µI,s,f ) = For m, n ∈ Z≥0 with m ≤ n, we define (XSn , BSn, µSn). ∞Yn=1 D(m,n) p,d,I,s,f = M p,Sm+1 d(m+1) ⊗p M p,Sm+2 d(m+2) ⊗p · · · ⊗p M p,Sn d(n) . (This algebra is called AN≤n,N>m in Example 1.4.) We then define D(m,∞) the direct limit p,d,I,s,f to be D(m,∞) p,d,I,s,f = lim −→ n D(m,n) p,d,I,s,f using, for n1, n2 ∈ Z≥0 with n2 ≥ n1 ≥ m, the maps ψ(m) given by ψ(m) p,d,I,s,f → D(m,n2) p,d,I,s,f p,d,I,s,f . (This algebra is called AN,N>m in n2,n1 (x) = x ⊗ 1 for x ∈ D(m,n1) n2,n1 : D(m,n1) Example 1.4.) We regard it as a subalgebra of L(cid:0)Lp(XI,s,f , µI,s,f )(cid:1). For n ≥ m p,d,I,s,f be the map associated with the direct limit. p,d,I,s,f . (This algebra is called A(d, ρ) in Example 1.4.) we let ψ(m) p,d,I,s,f → D(m,∞) Finally, we set Dp,d,I,s,f = D(0,∞) ∞,n : D(m,n) The algebra D(n,n) p,d,I,s,f is the empty tensor product, which we take to be C. Theorem 3.19. Let the hypotheses and notation be as in Definition 3.18. Then: (1) For m, n ∈ Z≥0 with m ≤ n, there is a system T of d(m+1)d(m+2) · · · d(n)- similarities such that there is an isometric isomorphism D(m,n) p,d,I,s,f ∼= M p,T d(m+1)d(m+2)···d(n) which sends tensor products of standard matrix units to standard matrix l=m+1 Rp,Sl and, if Sl is diagonal for l = m + units. Moreover, Rp,T =Qn 1, m + 2, . . . , n, then T is diagonal. 20 N. CHRISTOPHER PHILLIPS (2) For every m ∈ Z≥0, the algebra D(m,∞) p,d,I,s,f is an Lp UHF algebra of infinite tensor product type which locally has enough isometries in the sense of Definition 2.7(2) of [14]. (3) For every m ∈ Z≥0, the algebra D(m,∞) p,d,I,s,f is simple and has a unique con- tinuous normalized trace. (4) For every m ∈ Z≥0, using part (1) and Definition 3.21 for the definition of the domain, there is a unique isometric isomorphism such that ϕ(x ⊗ y) = ψ(m) ϕ : D(m,n) p,d,I,s,f ⊗p D(n,∞) p,d,I,s,f → D(m,∞) p,d,I,s,f ∞,n(x)(1 ⊗ y) for x ∈ D(m,n) p,d,I,s,f and y ∈ D(n,∞) p,d,I,s,f . Proof. Part (1) follows from Lemma 3.17 by induction. p,d,I,s,f . is isometrically isomorphic to D(0,∞) We prove parts (2) and (4). By renumbering, without loss of generality m = 0. In Example 1.4, for n ∈ Z>0 we set Xn = XSn, µn = µSn, and ρn = ψp,Sn . It is n=1 Xn) easy to see that the algebra constructed there (and represented on Lp (Q∞ tries. We verify Definition 2.7(2) of [14] by using the partition Xn = `i∈In For part (2), it remains to check that this algebra locally has enough isome- Yi with Yi = {1, 2, . . . , d(n)} × {i} for i ∈ In. Take G0 to be the group of signed permutation matrices (Definition 2.9 of [14]), and, for i ∈ In, in the application of Definition 2.7(1) of [14] to ρn(·)Lp(Yv,µn) we take the finite subgroup G to be G = sn(i)−1G0sn(i). Using Lemma 2.10 of [14], we easily see that G acts irreducibly and we easily get kρn(g)Lp(Yi,µSn )k = 1 for all g ∈ G. with product measure ν. Applying part (2) with n in place of m, we obtain D(n,∞) For part (4), set X =Qn as a closed unital subalgebra of L(Lp(Y, ν)). Set J =Qn m=1 XSn , with product measure µ, and Y =Q∞ p,d,I,s,f m=1 Im, and define g : J → (0, 1] and t : J → Md(1)d(2)···d(n) by m=n+1 XSn , g(i1, i2, . . . ⊗ in) = f1(i1)f2(i2) · · · fn(in) and t(i1, i2, . . . ⊗ in) = s1(i1) ⊗ s2(i2) ⊗ · · · ⊗ sn(in) for ik ∈ Ik for k = 1, 2, . . . , n. Then it is easy to see (compare with part (1)) that we we can identify D(m,n) p,d,I,s,f ⊂ d(m+1)d(m+2)···d(n), and that D(0,∞) p,d,I,s,f with M p,J,t,g L (Lp (Q∞ n=1 Xn)) is the closed linear span of all x ⊗ y with x ∈ D(0,n) and y ∈ D(n,∞) p,d,I,s,f ⊂ L(Lp(X, µ)) p,d,I,s,f ⊂ L(Lp(Y, ν)). Part (4) is now immediate. Part (3) follows from part (2) and from Corollary 3.12 and Theorem 3.13 of [14]. (cid:3) Proposition 3.20. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence of integers such that d(n) ≥ 2 for all n ∈ Z>0, and let I = (I1, I2, . . .), s = (s1, s2, . . .), t = (t1, t2, . . .), and f = (f1, f2, . . .) be sequences such that Sn = (In, sn, fn) and Tn = (In, tn, fn) are systems of d(n)- similarities for all n ∈ Z>0. Assume one of the following: (1) For every n ∈ Z>0 and every i ∈ In, there is γn(i) ∈ C \ {0} such that tn(i) = γn(i)sn(i). (2) For every n ∈ Z>0 and every i ∈ In, there is vn(i) ∈ inv(Md(n)) such that kvn(i)kp = kvn(i)−1kp = 1 and tn(i) = vn(i)sn(i). (3) For every n ∈ Z>0 there is wn ∈ inv(Md(n)) such that kwnkp = kw−1 n kp = 1 and for all i ∈ In we have tn(i) = wnsn(i)w−1 n . AMENABILITY 21 Then Rp,Tn = Rp,Sn for all n ∈ Z>0, and there is an isometric isomorphism Dp,d,I,t,f ∼= Dp,d,I,s,f . In part (3), we could take tn(i) = sn(i)w−1 n . We would then no longer need wn to be isometric, merely invertible, and the proof would be a bit simpler. But this operation rarely gives 1 ∈ ran(tn). Proof of Proposition 3.20. We use the notation of Definition 3.18 and Example 1.4, with the following modifications. For n ∈ Z≥0, set Xn = XIn,sn,fn , which is the same space as XIn,tn,fn , and set µn = µIn,sn,fn , which is equal to µIn,tn,fn . Set k=n+1 Xn, and call the product measure λn. Thus Z0 = XI,s,f , and we abbreviate this space to X and call the measure on it µ. We set k=1 Xn, and call the product measure νn. Set Zn =Q∞ Yn =Qn An,s = D(0,n) p,d,I,s,f ⊗ 1Lp(Zn,λn) and An,t = D(0,n) p,d,I,t,f ⊗ 1Lp(Zn,λn), both of which are subsets of L(Lp(X, µ)). We set As = Dp,d,I,s,f and At = Dp,d,I,t,f . Thus As = An,s and At = An,t. ∞[n=0 ∞[n=0 We prove the case (1). For d ∈ Z>0, s ∈ inv(Md), and γ ∈ C \ {0}, we have (γs)a(γs)−1 = sas−1 for all a ∈ Md, and also kγskp · k(γs)−1kp = kskp · ks−1kp. It follows that ψp,Tn = ψp,Sn for all n ∈ Z>0. Therefore As and At are actually equal as subsets of L(Lp(X, µ)). It is also immediate that Rp,Tn = Rp,Sn for all n ∈ Z>0. We next prove the case (2). (We will refer to this argument in the proof of the case (3) as well.) Let n ∈ Z>0. For all i ∈ In, we clearly have kvn(i)sn(i)kp = ksn(i)kp and k(vn(i)sn(i))−1kp = ksn(i)−1kp. So Rp,Tn = Rp,Sn. For i ∈ In, we interpret vn(i) as an element of L(cid:0)Lp({1, 2, . . . , d(n)} × {i}, µn)(cid:1). By Lemma 2.11, we still have kvn(i)kp = kvn(i)−1kp = 1. Following Remark 3.7, set vn =Li∈I vn(i), so kvnk = n k = 1. For n ∈ Z≥0, define zn ∈ L(Lp(X, µ)) with respect to the decomposition Lp(X, µ) = Lp(X1, µ1) ⊗ Lp(X2, µ2) ⊗ · · · ⊗ Lp(Xn, µn) ⊗ Lp(Zn, λn) kv−1 by zn = v1 ⊗ v2 ⊗ · · · ⊗ vn ⊗ 1Lp(Zn,λn). Using Theorem 2.16(5) of [13], we get kznk = 1. An analogous tensor prod- n k = 1. The formula a 7→ znaz−1 uct decomposition gives kz−1 n defines a bijec- tion ϕn : An,s → An,t, which is isometric because kznk = kz−1 n k = 1. Clearly ϕn+1An,s = ϕn. Therefore there exists an isometric homomorphism ϕ : As → At such that ϕAn,s = ϕn for all n ∈ Z≥0. Clearly ϕ has dense range. Therefore ϕ is surjective. This completes the proof of the case (2). We now prove the case (3). Let n ∈ Z>0. We clearly have kwnsn(i)w−1 n kp = ksn(i)kp and k(wnsn(i)w−1 n )−1kp = ksn(i)−1kp. So Rp,Tn = Rp,Sn . Let σn,s : Mrd(n) → An,s and σn,t : Mrd(n) → An,t be the maps analogous to σn in Example 1.4, except that the codomains are taken to be An,s instead of As and An,t instead of At. In the proof of the case (2), take vn(i) = wn for n ∈ Z>0 and i ∈ In, and then for n ∈ Z>0 let vn be as there and for n ∈ Z≥0 let zn be as there. Thus zn is a bijective isometry. Further set yn = w1 ⊗ w2 ⊗ · · · ⊗ wn ∈ Mrd(n). Since σn,s and σn,t are bijections, there is a unique bijection βn : An,s → An,t such that βn(σn,s(x)) = σn,t(ynxy−1 n ) for all x ∈ Mrd(n). Since σn+1, s(x ⊗ 1) = σn,s(x) and σn+1, t(x ⊗ 1) = σn,t(x) for all x ∈ Mrd(n), we get βn+1An,s = βn for all n ∈ Z≥0. 22 N. CHRISTOPHER PHILLIPS We now show that βn is isometric for all n ∈ Z≥0. Doing so finishes the proof, in the same way as at the end of the proof of the case (2). Fix n ∈ Z≥0. Let x ∈ M p rd(n), and interpret yn as an element of M p rd(n). For i = (i1, i2, . . . , in) ∈ I1 × I2 × · · · × In, make the abbreviations s(i) = s1(i1) ⊗ s2(i2) ⊗ · · · ⊗ sn(in) and t(i) = t1(i1) ⊗ t2(i2) ⊗ · · · ⊗ tn(in). Then t(i) = yns(i)y−1 n . By Lemma 3.17 and Lemma 3.15(1), and similarly with t in place of s. Theorem 2.16(5) of [13], applied to both yn and y−1 n , shows that yn is isometric. For i ∈ I1 × I2 × · · · × In, we use this fact at the second step and t(i) = yns(i)y−1 kσn,s(x)k = sup(cid:0)(cid:8)ks(i)xs(i)−1kp : i ∈ I1 × I2 × · · · × In(cid:9)(cid:1), (cid:13)(cid:13)t(i)ynxy−1 n (cid:13)(cid:13)p = ks(i)xs(i)−1kp. n t(i)−1(cid:13)(cid:13)p =(cid:13)(cid:13)yns(i)xs(i)−1y−1 Taking the supremum over i ∈ I1 ×I2 ×· · ·×In, we get kσn,t(ynxy−1 as desired. n )k = kσn,s(x)k, (cid:3) n at the first step to get Since the Banach algebras M p,I,s,f and M p,I,s,f ⊗p B and only depend on ran(s), d d we can make the following definition, based on Example 2.15 of [14]. Definition 3.21. Let d ∈ Z>0, let K ⊂ inv(Md) be a compact set with 1 ∈ K, and let p ∈ [1, ∞). Choose any system S = (I, s, f ) of d-similarities such that ran(s) = K. (It is obvious that there is such a system of d-similarities.) We then define M p,K as a Banach algebra. For any σ-finite measure space (Y, C, ν) and any unital closed subalgebra B ⊂ L(Lp(Y, ν)), we define M p,K d ⊗p B as a Banach algebra. We define d ⊗p B = M p,S d = M p,S d k · kp,K = k · kp,S, ψp,K B = ψp,S B , and Rp,K = Rp,S. If K = {1}, taking S to be the basic system S0 of d-similarities (Definition 3.2), we get M p,{1} B . If K1, K2 ⊂ inv(Md) are compact sets which contain 1, then we choose systems S1 = (I1, s1, f1) and S2 = (I2, s2, f2) of d-similarities such that ran(s1) = K1 and ran(s2) = K2, and define d and ψp,{1} = ψp,S0 = M p B d κp,K2,K1 B = κp,S2,S1 B : M p,K1 d ⊗p B → M p,K2 d ⊗p B. We say that K is diagonal if K is contained in the diagonal matrices in Md. Definition 3.22. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence of integers such that d(n) ≥ 2 for all n ∈ Z>0, and let K = (K1, K2, . . .) be a sequence of compact subsets Kn ⊂ inv(Md(n)) with 1 ∈ Kn for n ∈ Z>0. We define the (p, d, K)- UHF algebra Dp,d,K and associated subalgebras as follows. Choose any sequences I = (I1, I2, . . .), s = (s1, s2, . . .), and f = (f1, f2, . . .) such that for all n ∈ Z>0, the triple Sn = (In, sn, fn) is a system of d(n)-similarities with ran(sn) = Kn. Then, following Definition 3.18, define D(m,n) p,d,I,s,f for m, n ∈ Z≥0 ∪ {∞} with m ≤ n and m 6= ∞, define ψ(m) p,d,K as in Definition 3.18, and define Dp,d,K = D(0,∞) p,d,K → D(m,∞) p,d,K = D(m,n) ∞,n : D(m,n) p,d,I,s,f . One can prove that for algebras of the form Dp,d,K as in Definition 3.22, the conditions (1), (2), (3), (4), (5), and (6) in Theorem 2.16 are also equivalent to the following: (11) There is a uniform bound on the norms of the maps σn : M p rd(n) → A. (12) P∞ n=1(kρnk − 1) < ∞. AMENABILITY 23 The new feature (which will be made explicit in a more restrictive context in n of Mdn and an n (x)w−1 for Theorem 4.10 below) is that there are a spatial representation ρ(0) invertible operator w with kwk · kw−1k = kρnk such that ρn(x) = wρ(0) all x ∈ Md(n). Since ρ(0) n is again spatial (Lemma 1.12 of [14]) and n ⊗ ρ(0) (ρn ⊗p ρn)(x) = (w ⊗ w)(cid:0)ρ(0) n ⊗ ρ(0) n (cid:1)(x)(w ⊗ w)−1 for x ∈ Md(n) ⊗p Md(n), we can use Lemma 3.3 to get kρn ⊗p ρnk = kρnk2, rather than merely kρn ⊗p ρnk ≥ kρnk2. Similarly, we get kσn ⊗p σnk = kσnk2. 4. Amenability of Lp UHF algebras constructed from diagonal similarities Let A ⊂ L(Lp(X, µ)) be an Lp UHF algebra of tensor product type constructed using a system of diagonal similarities. The main result of this section (Theo- rem 4.10) is that a number of conditions, of which the most interesting is probably amenability, are equivalent to A being isomorphic to the spatial Lp UHF algebra B with the same supernatural number. If there is an isomorphism, we can in fact realize B as a subalgebra of L(Lp(X, µ)), in such a way that the isomorphism can be taken to be given by conjugation by an invertible element in L(Lp(X, µ)). When p = 2, the conditions can be relaxed: we do not need to assume that the similarities are diagonal. The key technical ideas are information about the form of an approximate di- agonal, and an estimate on its norm based on information about the norms of off diagonal matrix units after conjugating by an invertible element. We only know sufficiently good estimates for conjugation by diagonal matrices, which is why we restrict to diagonal similarities. Lemma 4.1. Let M ∈ [1, ∞). Let B be a unital Banach algebra which has an approximate diagonal with norm at most M. Let G ⊂ inv(B) be a finite subgroup. and (g ⊗ 1)z = z(1 ⊗ g) for all g ∈ G. Then for every ε > 0 there exists z ∈ Bb⊗B such that kzkπ < M + ε, ∆B(z) = 1, The proof is easily modified to show that we can also require that, for all a in a given finite set F ⊂ A, we have k(a ⊗ 1)z − z(1 ⊗ a)k < ε. We don't need this refinement here. Proof of Lemma 4.1. Using the definition of ∆B (see Lemma 2.2), for a, b ∈ B and x ∈ Bb⊗B, we get (4.1) Set r = supg∈G kgk. Set ∆B(cid:0)(a ⊗ 1)x(1 ⊗ b)(cid:1) = a∆B(x)b. 8r2(cid:19) . δ = min(cid:18) 1 ε 4M ε 4r 2 ε , , , Since δ ≤ 1 2 , the element ∆B(z0) is invertible in B and By hypothesis, there exists z0 ∈ Bb⊗B such that kz0kπ ≤ M, k∆B(z0) − 1k < δ, and(cid:13)(cid:13)(g ⊗ 1)z0 − z0(1 ⊗ g)(cid:13)(cid:13)π < δ for all g ∈ G. (cid:13)(cid:13)∆B(z0)−1 − 1(cid:13)(cid:13) < Set z1 = (∆B(z0)−1 ⊗ 1)z0. Then ∆B(z1) = 1 by (4.1). Further, ≤ 2δ. 1 − δ δ kz1 − z0k ≤ k∆B(z0)−1 ⊗ 1k · kz0k < 2δM. 24 N. CHRISTOPHER PHILLIPS (4.2) Since 2δM ≤ ε 2 , we get kz1k < M + ε 2 . Also, for g ∈ G we have (cid:13)(cid:13)(g ⊗ 1)z1 − z1(1 ⊗ g)(cid:13)(cid:13)π ≤(cid:13)(cid:13)(g ⊗ 1)z0 − z0(1 ⊗ g)(cid:13)(cid:13)π + 2kgk · kz1 − z0k . < δ + 2r(cid:16) ε 8r2(cid:17) ≤ ε 2r Now define z = 1 card(G) Xh∈G (h ⊗ 1)z0(1 ⊗ h)−1. From (4.2), for g ∈ G we get (cid:13)(cid:13)(g ⊗ 1)z1(1 ⊗ g)−1 − z1(cid:13)(cid:13)π ≤(cid:13)(cid:13)(g ⊗ 1)z1 − z1(1 ⊗ g)(cid:13)(cid:13)π · k(1 ⊗ g)−1kπ <(cid:16) ε It follows that kz − z1k < ε 2 , whence kzk < M + ε. 2r(cid:17) r ≤ For g ∈ G we also get (g ⊗ 1)z = = 1 card(G) Xh∈G card(G) Xh∈G 1 (gh ⊗ 1)z0(1 ⊗ h)−1 (h ⊗ 1)z0(1 ⊗ g−1h)−1 = z(1 ⊗ g). Finally, using ∆B(z1) = 1 and (4.1), we have ∆B(z) = This completes the proof. 1 card(G) Xh∈G h∆B(z1)h−1 = 1. ε 2 . (cid:3) Lemma 4.2. Let A1 and A2 be unital Banach algebras, let d1, d2 ∈ Z>0, equip Md1 and Md2 with any algebra norms, and equip B1 = Md1 ⊗ A1 and B2 = Md2 ⊗ A2 with any algebra tensor norm. Then B1 and B2 are complete, and there is a unique algebra bijection such that ϕ : Md1 ⊗alg Md2 ⊗alg(cid:0)A1b⊗A2(cid:1) → B1b⊗B2 ϕ(cid:0)x1 ⊗ x2 ⊗(cid:0)a1b⊗a2(cid:1)(cid:1) = (x1 ⊗ a1)b⊗(x2 ⊗ a2) for all x1 ∈ Md1, x2 ∈ Md2, a1 ∈ A1, and a2 ∈ A2. Notation 4.3. We will often implicitly use the isomorphism ϕ of Lemma 4.2 to d1Xj,k=1 d2Xl,m=1 ej,k ⊗ el,m ⊗ aj,k,l,m write particular elements of B1b⊗B2 as there in the form x1 ⊗ x2 ⊗ a with x1 ∈ Md1, x2 ∈ Md2, and a ∈ Ab⊗A, or (using Notation 3.1) a general element of B1b⊗B2 as with aj,k,l,m ∈ A1b⊗A2 for j, k = 1, 2, . . . , d1 and l, m = 1, 2, . . . , d2. on B1 are equivalent. Thus, we may take B1 = Md1b⊗A1. Similarly, we may take B2 = Md2b⊗A2, and make the identification of algebras Proof of Lemma 4.2. Completeness of B1 and B2 follows from finite dimensionality of Md1 and Md2. Also by finite dimensionality of Md1, any two algebra tensor norms Now ϕ is just the permutation of projective tensor factors isomorphism Md1 ⊗alg Md2 ⊗alg(cid:0)A1b⊗A2(cid:1) = Md1b⊗Md2b⊗A1b⊗A2. ϕ : Md1b⊗Md2b⊗A1b⊗A2 → Md1b⊗A1b⊗Md2b⊗A2. This completes the proof. The following lemma was suggested by the last paragraph in Section 7.5 of [1]. (cid:3) AMENABILITY 25 Lemma 4.4. Let A be a unital Banach algebra, let d ∈ Z>0, equip Md with any algebra norm, and equip B = Md ⊗ A with any algebra tensor norm. Suppose for all x ∈ Md. Then, rearranging tensor factors as in Notation 4.3, there exist (x ⊗ 1A ⊗ 1B)z = z(1B ⊗ x ⊗ 1A) (4.3) z ∈ Bb⊗B satisfies ∆B(z) = 1 and elements zj,k ∈ Ab⊗A for j, k = 1, 2, . . . , d such that dXj=1 dXj,k,l=1 ej,k ⊗ el,j ⊗ zl,k and z = Proof. Use Lemma 4.2 and Notation 4.3 to write z = dXj,k,l,m=1 ej,k ⊗ el,m ⊗ aj,k,l,m ∆A(zj,j) = 1. with aj,k,l,m ∈ Ab⊗A for j, k, l, m = 1, 2, . . . , d. Let r, s ∈ {1, 2, . . . , d}. In (4.3) put x = es,r, getting es,k ⊗ el,m ⊗ ar,k,l,m − ej,k ⊗ el,r ⊗ aj,k,l,s = 0. Let p ∈ {1, 2, . . . , d} and multiply on the right by ep,p ⊗ er,r ⊗ 1A b⊗A, getting (4.4) es,p ⊗ el,r ⊗ ar,p,l,r − ej,p ⊗ el,r ⊗ aj,p,l,s = 0. dXk,l,m=1 dXl=1 dXj,k,l=1 dXj,l=1 Let q ∈ {1, 2, . . . , d}. First, assume r 6= s and multiply (4.4) on the left by er,r ⊗ eq,q ⊗ 1A b⊗A. The first sum is annihilated, and we get −er,p ⊗ eq,r ⊗ ar,p,q,s = 0. Thus (by injectivity in Lemma 4.2), we have ar,p,q,s = 0 whenever p, q, r, s ∈ {1, 2, . . . , d} satisfy r 6= s. For arbitrary r, s ∈ {1, 2, . . . , d}, multiply (4.4) on the left by es,s ⊗ eq,q ⊗ 1A b⊗A. This gives es,p ⊗ eq,r ⊗ ar,p,q,r − es,p ⊗ eq,r ⊗ as,p,q,s = 0. Therefore ar,p,q,r = as,p,q,s for all p, q, r, s ∈ {1, 2, . . . , d}. With zl,k = a1,k,l,1 for k, l = 1, 2, . . . , d, we therefore get z = ej,k ⊗ el,j ⊗ zl,k. dXj,k,l=1 Apply ∆B to this equation, using ej,kel,j = 0 when k 6= l and ej,kek,j = ej,j, to get 1 = ∆B(z) = ej,j ⊗ ∆A(zk,k). dXj,k=1 k=1 ∆A(zk,k) = 1. Lemma 4.5. Let d ∈ Z>0, let α1, α2, . . . , αd ∈ C \ {0}, and set SoPd β = min(cid:0)α1, α2, . . . , αd(cid:1) and γ = max(cid:0)α1, α2, . . . , αd(cid:1). Let E be a normed vector space, and let ξ1, ξ2, . . . , ξd ∈ E satisfy(cid:13)(cid:13)(cid:13)Pd ζ1, ζ2, . . . , ζd ∈ S1 = {ζ ∈ C : ζ = 1} and j0 ∈ {1, 2, . . . , d} Then there exist j=1 ξj(cid:13)(cid:13)(cid:13) = 1. (cid:3) 26 N. CHRISTOPHER PHILLIPS such that (4.5) or (4.6) (ζj0 αj0 )(ζkαk)−1ξk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (ζj0 αj0 )−1(ζkαk)ξk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1 ≥ β−1/2γ1/2 ≥ β−1/2γ1/2. 2 , . . . , ζ(0) Proof. We claim that it suffices to consider the case αj > 0 for j = 1, 2, . . . , d. To see this, assume that the result has been proved for α1, α2, . . . , αd, yielding 1 , ζ(0) ζ(0) d ∈ S1. Then the result for α1, α2, . . . , αd follows by taking ζj = sgn(αj )ζ(0) for j = 1, 2, . . . , d. We therefore assume that αj > 0 for j = 1, 2, . . . , d. Both (4.5) and (4.6) are unchanged if we choose any ρ > 0 and replace αj by ραj for j = 1, 2, . . . , d. We may therefore assume that β = 1. Reordering the αj and the ξj , we may assume that 1 = α1 ≤ α2 ≤ · · · ≤ αd = γ. j By the Hahn-Banach Theorem, there is a linear functional ω : E → C such that kωk = 1 and ω(ξj) = 1. dXj=1 Applying ω at the first step in both the following calculations, and using α1 = 1 1 j=1 λj ≥ 1. (4.7) = σ−1 1 ≥ = αkλk and (4.8) in the first and αd = γ in the second, we get (σ1α1)−1(σkαk)ξk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ξk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k λk! = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1 γ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1(cid:0)σdαd(cid:1)(cid:0)σkαk(cid:1)−1 dXk=1 Set σj = sgn(ω(ξj )) and λj = ω(ξj) for j = 1, 2, . . . , d. ThenPd (σ1α1)−1(σkαk)ω(ξk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≥(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dXk=1 1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) αkσkω(ξk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dXk=1 dXk=1 ω(ξk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) γ(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dXk=1(cid:0)σdαd(cid:1)(cid:0)σkαk(cid:1)−1 = γ−1σdαd(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) k σkω(ξk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dXk=1 dXk=1Xj6=k k−1Xj=1(cid:0)αkα−1 dXk=1 dXk=1 k αj(cid:1)λj λk λk!2 2λjλk = dXk=1 k−1Xj=1 dXk=1 dXk=1 ξk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) γ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1(cid:0)σdαd(cid:1)(cid:0)σkαk(cid:1)−1 (σ1α1)−1(σkαk)ξk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Now, using the inequality α + α−1 ≥ 2 for α > 0 at the third step, we get αkλk! dXk=1 Combining this result with (4.7) and (4.8), we get (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1 αkλk · α−1 k λk + ≥ γ1/2 or j + α−1 dXk=1 dXk=1 λ2 k + λ2 k + = ≥ α−1 αkλk · α−1 j λj ≥ 1. ≥ γ−1/2. 1 α−1 = α−1 k λk. AMENABILITY 27 In the first case, we get the conclusion of the lemma by choosing j0 = 1 and ζj = σj for j = 1, 2, . . . , d. In the second case, we get the conclusion by choosing j0 = d and ζj = σj for j = 1, 2, . . . , d. (cid:3) Lemma 4.6. Let d ∈ Z>0, let S = (I, s, f ) be a diagonal system of d-similarities, and let p ∈ [1, ∞). Let (Y, C, ν) be a σ-finite measure space, and let A ⊂ L(Lp(Y, ν)) be a unital closed subalgebra. Set B = M p,S ∆B(z) = 1 and d ⊗p A. Suppose z ∈ Bb⊗B satisfies (x ⊗ 1A ⊗ 1B)z = z(1B ⊗ x ⊗ 1A) for all x ∈ M p,S d . Then kzkS,π ≥ R1/2 p,S . Proof. Set B0 = M p Let S0 be the basic system of d-similarities (Definition 3.2). d ⊗p A. Let i ∈ I; we show that kzkS,π ≥ ks(i)k1/2 p ks(i)−1k1/2 p . such that Adopt Notation 4.3, and use Lemma 4.4 to find zj,k ∈ Ab⊗A for j, k = 1, 2, . . . , d ej,k ⊗ el,j ⊗ zl,k and ∆A(zj,j) = 1. z = dXj,k,l=1 dXj=1 There are α1, α2, . . . , αd ∈ C \ {0} such that s(i) = diag(α1, α2, . . . , αd). Apply Lemma 4.5 with E = A, with α1, α2, . . . , αd as above, and with ξj = ∆(zj,j) for j = 1, 2, . . . , d. Then β = ks(i)−1k−1 p and γ = ks(i)kp. We obtain ζ1, ζ2, . . . , ζd ∈ S1 and j0 ∈ {1, 2, . . . , d} such that Define u = diag(ζ1, ζ2, . . . , ζd) and w = us(i). Then kukp = ku−1kp = 1, so ku ⊗ 1kp = ku−1 ⊗ 1kp = 1. Since B0 = M p d ⊗p A and B = M p,S d ⊗p A are just Md ⊗alg A as algebras (by Lemma 3.12), the formula b 7→ (s(i) ⊗ 1)b(s(i)−1 ⊗ 1) defines a homomorphism ϕ0 : B → B0. It follows from Lemma 3.15(1) that kϕ0(b)kp,S0 ≤ kbkp,S for all b ∈ B. Therefore also the formula ϕ(b) = (w ⊗ 1)b(w−1 ⊗ 1) = (u ⊗ 1)ϕ0(b)(u−1 ⊗ 1) defines a contractive homomorphism ϕ : B → B0. Also, b 7→ b defines a contrac- tive homomorphism from B to B0, namely the map κp,S0,S of Notation 3.14 and Lemma 3.15(3). The projective tensor product of contractive linear maps is con- B tractive, and ∆B0 : B0b⊗B0 → B0 is contractive. So there are contractive linear maps ∆1, ∆2 : Bb⊗B → B0 such that for all b1, b2 ∈ B we have the following formu- las (in each case, the first one justifies contractivity and the second one is in terms of what happens in Md ⊗alg B): and B ∆1(b1 ⊗ b2) = ∆B0(cid:0)ϕ0(b1) ⊗ κp,S0,S ∆2(b1 ⊗ b2) = ∆B0(cid:0)κp,S0,S wej,kw−1el,j =(0 B (b2)(cid:1) = (w ⊗ 1)b1(w−1 ⊗ 1)b2 (b1) ⊗ ϕ0(b2)(cid:1) = b1(w ⊗ 1)b2(w−1 ⊗ 1). (ζjαj)(ζkαk)−1ej,j l 6= k l = k. We evaluate ∆1(z) and ∆2(z). For j, k, l = 1, 2, . . . , d, we have (4.9) or (4.10) (ζj0 αj0 )(ζkαk)−1∆(zk,k)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1 (ζj0 αj0 )−1(ζkαk)∆(zk,k)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1 ≥ ks(i)k1/2 p ks(i)−1k1/2 p ≥ ks(i)k1/2 p ks(i)−1k1/2 p . 28 N. CHRISTOPHER PHILLIPS Therefore ∆1(z) = dXj,k,l=1 wej,kw−1el,j ⊗ ∆A(zl,k) = ej,j ⊗ dXj=1 dXk=1 (ζj αj)(ζkαk)−1∆A(zk,k). Similarly, one gets ∆2(z) = (ζjαj)−1(ζkαk)∆A(zk,k). If (4.9) holds, then we use (4.9) and kej0,j0 ⊗ 1kp = kej0,j0 kp = 1 to get kzkS,π ≥ k∆1(z)k ≥(cid:13)(cid:13)(ej0,j0 ⊗ 1)∆1(z)(ej0,j0 ⊗ 1)(cid:13)(cid:13) ≥ ks(i)k1/2 p ks(i)−1k1/2 p . ej,j ⊗ dXk=1 dXj=1 (ζj0 αj0 )(ζkαk)−1∆(zk,k)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dXk=1 If instead (4.10) holds, we similarly get kzkS,π ≥ k∆2(z)k ≥(cid:13)(cid:13)(ej0,j0 ⊗ 1)∆1(z)(ej0,j0 ⊗ 1)(cid:13)(cid:13) ≥ ks(i)k1/2 This completes the proof. p ks(i)−1k1/2 p . (cid:3) Theorem 4.7. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence of integers such that d(n) ≥ 2 for all n ∈ Z>0, and let I = (I1, I2, . . .), s = (s1, s2, . . .), and f = (f1, f2, . . .) be sequences such that Sn = (In, sn, fn) is a system of d(n)-similarities for all n ∈ Z>0. Suppose the algebra Dp,d,I,s,f of Definition 3.22 is amenable. Then n=1 Rp,Sn < ∞. Q∞ Qn Rp,T = Qn G0 ⊂ inv(cid:0)M p,L of [14]), and set r Proof. The hypothesis means that there is M ∈ [1, ∞) such that Dp,d,I,s,f has an approximate diagonal with norm at most M. Let n ∈ Z>0. We show that l=1 Rp,Sl ≤ (M + 1)2. This will prove the theorem. Set r = d(1)d(2) · · · d(n). Parts (1) and (4) of Theorem 3.19 provide a diago- nal system T = (J, t, g) of r-similarities, a σ-finite measure space (Y, C, ν), and a unital closed subalgebra A ⊂ L(Lp(Y, ν)) (namely A = Dn,∞ p,d,I,s,f ), such that r ⊗p A. Let l=1 Rp,Sl and Dp,d,I,s,f is isometrically isomorphic to M p,T Apply Lemma 4.1 with B = M p,L such that kzkπ < M + 1, ∆M p,L The signed permutation matrices span M p,L r ⊗p A, with G as given, and with ε = 1, getting (cid:1) be the group of signed permutation matrices (Definition 2.9 G = {g ⊗ 1A : g ∈ G0} ⊂ inv(cid:0)M p,L r ⊗p A(cid:1). r ⊗p A(cid:1) r ⊗p A(cid:1) ⊗alg(cid:0)M p,L z ∈(cid:0)M p,L r ⊗pA ⊗ x ⊗ 1A(cid:1) r ⊗pA(cid:1)z = z(cid:0)1M p,L (cid:0)x ⊗ 1A ⊗ 1M p,L nYl=1 r ⊗pA(z) = 1, and (g ⊗ 1)z = z(1 ⊗ g) for all g ∈ G. . Therefore Lemma 4.6 applies, and we conclude M + 1 > kzkL,π ≥ R1/2 by Lemma 2.11 of [14], so R1/2 p,Sl p,L = . r (cid:3) for all x ∈ M p,L r l=1 Rp,Sl ≤ (M + 1)2, as desired. SoQn AMENABILITY 29 Lemma 4.8. Let d ∈ Z>0, let S = (I, s, f ) be a diagonal system of d-similarities, and let p ∈ [1, ∞). Then, following the notation of Definition 3.11, there exist a spatial representation τ : Md → L(cid:0)Lp(XS, µS)(cid:1) and w ∈ inv(cid:0)L(cid:0)Lp(XS, µS)(cid:1)(cid:1) such that (4.11) kwk = Rp,S, kw − 1k = Rp,S − 1, kw−1k = 1, kw−1 − 1k = 1 − and ψp,S(x) = wτ (x)w−1 for all x ∈ Md. Proof. Let i ∈ I. There are α1, α2, . . . , αd ∈ C \ {0} such that 1 , Rp,S s(i) = diag(α1, α2, . . . , αd) ∈ M p d . Set βi = min(cid:0)α1, α2, . . . , αd(cid:1), and define ui = diag(cid:0)sgn(α1), sgn(α2), . . . , sgn(αd)(cid:1) Then s(i) = βiwiui, ui is isometric, and wi = β−1 i diag(cid:0)α1, α2, . . . , αd(cid:1). kwik = ks(i)k · ks(i)−1k, kwi − 1k = ks(i)k · ks(i)−1k − 1, kw−1 i k = 1, kw−1 i − 1k = 1 − ks(i)k · ks(i)−1k for all x ∈ Md, and x 7→ uixu−1 i , is a spatial repre- 1 (4.12) (4.13) s(i)xs(i)−1 = wi(uixu−1 sentation of M p d . i Define a representation )w−1 i τi : Md → L(cid:0)Lp({1, 2, . . . , d} × {i}, µS)(cid:1) by using the identification (3.1) in Definition 3.11 on the representation x 7→ uixu−1 for x ∈ Md. In the notation of Definition 3.11, we then have ψi(x) = wiτi(x)w−1 for all x ∈ Md. of the representations τi. Then τ is spatial by Lemma 3.10. Further take w to be the Lp direct sum of the operators wi for i ∈ I. Since Rp,S = supi∈I ks(i)k · ks(i)−1k, it Let τ : Md → L(cid:0)Lp(XS, µS)(cid:1) be the Lp direct sum, as in Definition 3.6, over i ∈ I follows from (4.12) and (4.13) that w is in fact in L(cid:0)Lp(XS, µS)(cid:1) and satisfies (4.11). It is clear that ψp,S(x) = wτ (x)w−1 for all x ∈ Md. (cid:3) i i Lemma 4.9. Adopt the notation of Example 1.4, but suppose that, for each n ∈ N, instead of ρn : Md(n) → L(Lp(Xn, µn)) we have two representations ρ(1) n , ρ(2) n : Md(n) → L(Lp(Xn, µn)). Let A(1) and A(2) be the corresponding Lp UHF algebras of tensor product type. Suppose that for every n ∈ Z>0 there is wn ∈ inv(cid:0)L(Lp(Xn, µn))(cid:1) such that n (x) for all x ∈ Md(n). Suppose further that n = ρ(2) n (x)w−1 wnρ(1) ∞Xn=1 kwn − 1k < ∞ and n − 1(cid:13)(cid:13) < ∞. Then there is y ∈ inv(cid:0)L(Lp(X, µ))(cid:1) such that yA(1)y−1 = A(2). ∞Xn=1(cid:13)(cid:13)w−1 Proof. For j = 1, 2, we adapt the notation of Example 1.4 by letting σ(j) A(j) be the analog, derived from the representations ρ(j) A at the end of Example 1.4. We further set A(j) n : Mrd(n) → n , of the map σn : Mrd(n) → n (Mrd(n)) ⊂ L(Lp(X, µ)), n = σ(j) so that A(j) = S∞ n y−1 = A(2) yA(1) n for all n ∈ Z≥0. n=0 A(j) n . It suffices to find y ∈ inv(cid:0)L(Lp(X, µ))(cid:1) such that 30 N. CHRISTOPHER PHILLIPS We claim that M1 = sup n∈Z≥0 nYk=1 kwnk < ∞ and M2 = sup n∈Z≥0 nYk=1(cid:13)(cid:13)w−1 n (cid:13)(cid:13) < ∞. The proofs are the same for both, so we prove only the first. For n ∈ Z>0, we observe that kwnk ≤ 1 + kwn − 1k, so max(kwnk, 1) − 1 ≤ kwn − 1k. Therefore Using Lemma 2.15 at the second step, we therefore get ∞Xn=1 ∞Xn=1(cid:2) max(kwnk, 1) − 1(cid:3) ≤ ∞Yn=1 nYk=1 sup n∈Z≥0 kwnk ≤ kwn − 1k < ∞. max(kwnk, 1) < ∞. The claim is proved. For n ∈ Z≥0, set yn = w1 ⊗ w2 ⊗ · · · ⊗ wn ⊗ 1N>n ∈ L(Lp(X, µ)). Then Therefore kyn − yn−1k =(cid:13)(cid:13)w1 ⊗ w2 ⊗ · · · ⊗ wn−1 ⊗ (wn − 1) ⊗ 1N>n(cid:13)(cid:13) kwkk ≤ M1kwn − 1k. = kwn − 1k n−1Yk=1 kwn − 1k < ∞. kyn − yn−1k ≤ M1 ∞Xn=1 n−1(cid:13)(cid:13) =(cid:13)(cid:13)w−1 ∞Xn=1 n−1Yk=1(cid:13)(cid:13)w−1 (cid:13)(cid:13)y−1 n−1(cid:13)(cid:13) < ∞, whence z = limn→∞ y−1 n=1(cid:13)(cid:13)y−1 It follows that y = limn→∞ yn ∈ L(cid:0)Lp(X, µ)(cid:1) exists. We also get k (cid:13)(cid:13) ≤ M2(cid:13)(cid:13)w−1 n − 1(cid:13)(cid:13), n ∈ L(cid:0)Lp(XI,s,f , µI,s,f )(cid:1) exists. soP∞ Clearly yz = zy = 1, so y is invertible with inverse z. Since ylA(1) all l ∈ Z>0 with l ≥ n, it follows that yA(1) n − 1(cid:13)(cid:13) n y−1 = A(2) n . l = A(2) n − y−1 n y−1 n for (cid:3) n −y−1 Theorem 4.10. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence in {2, 3, . . .}, and let I = (I1, I2, . . .), s = (s1, s2, . . .), and f = (f1, f2, . . .) be sequences such that Sn = (In, sn, fn) is a diagonal system of d(n)-similarities for all n ∈ Z>0. Set A = Dp,d,I,s,f ⊂ L(cid:0)Lp(XI,s,f , µI,s,f )(cid:1) as in Definition 3.18 with the given choices of I, s, and f. For n ∈ Z≥0, following the notation of Example 1.4, let σn : Mrd(n) → A be σn = ρN,N≤n , and following the notation of Definition 3.18, set ρn = ψp,Sn. (The notation ρn is used in Example 1.4.) Let B be the spatial Lp UHF algebra whose supernatural number is the same as that of A. Then the following are equivalent: (1) A ∼= B. (2) There exists an isomorphism ϕ : A → B such that the algebraic tensor product of two copies of ϕ extends to an isomorphism ϕ ⊗p ϕ : A ⊗p A → B ⊗p B. (3) A is symmetrically amenable. (4) A is amenable. AMENABILITY 31 (5) Whenever (Y, C, ν) is a σ-finite measure space, C ⊂ L(Lp(Y, ν)) is a closed unital subalgebra, and ϕ : A → C is a unital continuous homomorphism such that ϕ ⊗p idA : A ⊗p A → C ⊗p A is bounded, then ϕ is bounded below. (6) A ⊗p A has approximately inner Lp-tensor half flip. (7) A has approximately inner Lp-tensor flip. (8) A has approximately inner Lp-tensor half flip. (9) There is a uniform bound on the norms of the maps σn : M p (10) There is a uniform bound on the norms of the maps rd(n) → A. σn ⊗p σn : M p rd(n)2 → A ⊗p A. n=1(kρnk − 1) < ∞. (11) P∞ (12) With Rp,Sn as in Definition 3.11 for n ∈ Z>0, we haveP∞ (13) The homomorphisms ρn ⊗p ρn : M p d(n)2 → A ⊗p A satisfy ∞. n=1(Rp,Sn − 1) < (kρn ⊗p ρnk − 1) < ∞. ∞Xn=1 (14) There exists a spatial Lp UHF algebra C ⊂ L(cid:0)Lp(XI,s,f , µI,s,f )(cid:1) which is isometrically isomorphic to B, and v ∈ inv(cid:0)L(cid:0)Lp(XI,s,f , µI,s,f )(cid:1)(cid:1), such that vAv−1 = C. Some of the conditions in Theorem 4.10 do not involve the triple (I, s, f ). It follows that these all hold, or all fail to hold, for any choice of (I, s, f ) giving an isomorphic algebra. The algebra B ⊗p B in (2) is a spatial Lp UHF algebra. In particular, using Theorem 3.10(5) of [14], one can check that, up to isometric isomorphism, it does not depend on how B is represented on an Lp space. Proof of Theorem 4.10. We simplify the notation, following Example 1.4, by setting Xn = XSn and µn = µSn for all n ∈ Z>0. We freely use other notation from Example 1.4. The equivalence of (2), (6), (7), (8), (10), and (13) follows from Theorem 2.16. That (2) implies (1) is trivial. Assume (1); we prove (3). Since symmetric amenability only depends on the isomorphism class of a Banach algebra, it is enough to prove symmetric amenability when A = B. The maps σn : M p rd(n) → B are then isometric, so that the maps σnb⊗σn : M p apply Proposition 2.10(1) and Lemma 2.5(7). The implication from (3) to (4) is trivial. rd(n) → Bb⊗B are contractive. Now rd(n)b⊗M p The implication from (4) to (12) is immediate from Theorem 4.7 and Lemma 2.15. Now assume (12); we prove (11). Let Tn be the basic system of d(n)-similarities (called S0 in Definition 3.2). When one identifies Md(n) with M p d(n), the map ρn = ψp,Sn : Md(n) → M p,Sn d(n) of Nota- tion 3.14. Then Lemma 3.15(2) implies that kρnk = Rp,Sn for n ∈ Z>0. The desired implication is now clear. d(n) becomes the map κp,Sn,Tn d(n) → M p,Sn : M p C Assume (11); we prove (14). Apply Lemma 4.8 to Sn for all n ∈ Z>0, obtaining spatial representations τn : Md(n) → L(Lp(Xn, µn)) and invertible elements wn ∈ L(Lp(Xn, µn)), satisfying the estimates given there. We have Rp,Sn = kρnk as in the proof of the implication from (12) to (11). Therefore the estimates become kwnk = kρnk, kwn − 1k = kρnk − 1, kw−1 n k = 1, and kw−1 n − 1k = 1 − 1 kρnk . 32 N. CHRISTOPHER PHILLIPS isometrically isomorphic to B. Applying the construction of Example 1.4 to d and (τ1, τ2, . . .), we obtain a spatial ral number is the same as that of A. Lemma 4.9 provides an invertible element Lp UHF algebra C = A(d, τ ) ⊂ L(cid:0)Lp(XI,s,f , µI,s,f )(cid:1). Obviously its supernatu- y ∈ L(cid:0)Lp(XI,s,f , µI,s,f )(cid:1) such that yAy−1 = C. Theorem 1.8 implies that C is We now show that (14) implies (2). By hypothesis, the formula ϕ(a) = vav−1 for a ∈ A defines an isomorphism ϕ : A → B. Then a 7→ (v ⊗ v)a(v ⊗ v)−1 is an isomorphism from A ⊗p A to B ⊗p B which sends a1 ⊗ a2 to ϕ(a1) ⊗ ϕ(a2) for all a1, a2 ∈ A. It remains only to prove that (5) and (9) are equivalent to the other conditions. We prove that (9) is equivalent to (10). It suffices to prove that kσn⊗σnk = kσnk2 for all n ∈ Z>0. Repeated application of Lemma 3.17 shows that there is a diagonal system T0 = (J0, t0, g0) of rd(n)-similarities such that σn = ψp,T0 . The same lemma further shows that setting J = J0 × J0 and setting t(j, k) = t0(j) ⊗ t0(k) and g(j, k) = g0(j)g0(k) gives a diagonal system T = (J, t, g) of rd(n)2-similarities such that σn ⊗ σn = ψp,T C . The relation kσn ⊗ σnk = kσnk2 now follows from Lemma 3.15(1), Lemma 3.3, and the relation kv ⊗ wk = kvk · kwk for v, w ∈ M p C rd(n). That (8) implies (5) follows from Theorem 2.7. We show that (5) implies (1). Let κ : A → B be as in Corollary 2.14. By Corollary 2.14(3), there is a continuous homomorphism γ : A ⊗p A → B ⊗p A such that γ(a1 ⊗ a2) = κ(a1) ⊗ a2 for all a1, a2 ∈ A. Condition (5) implies that κ has closed range. Since κ has dense range by Corollary 2.14(1), we conclude that κ is an isomorphism. (cid:3) For p = 2, we can do a little better: we don't need to require the systems of similarities to be diagonal. Theorem 4.11. Let d = (d(1), d(2), . . .) be a sequence in {2, 3, . . .}, and let I = (I1, I2, . . .), s = (s1, s2, . . .), and f = (f1, f2, . . .) be sequences such that Sn = (In, sn, fn) is a system of d(n)-similarities for all n ∈ Z>0. Set A = D2,d,I,s,f ⊂ L(cid:0)L2(XI,s,f , µI,s,f )(cid:1) as in Definition 3.22 with the given choices of I, s, and f. If A is amenable, then there exists v ∈ inv(cid:0)L(cid:0)L2(XI,s,f , µI,s,f )(cid:1)(cid:1) such that vAv−1 is a C*-algebra. Problem 30 in the "Open Problems" chapter of [16] asks whether an amenable closed subalgebra of the bounded operators on a Hilbert space is similar to a C*- algebra. This problem has been open for some time, and little seems to be known. (See the discussion in the introduction.) Theorem 4.11 shows that the answer is yes for a class of algebras which, as far as we know, is quite different from any other class considered in this context. Proof of Theorem 4.11. Let Rp,S be as in Definition 3.11. By Proposition 3.20(1), for all n ∈ Z>0 and i ∈ In, we can replace sn(i) by ksn(i)−1ksn(i). Therefore we may assume that ksn(i)−1k = 1. (This change preserves the requirement 1 ∈ ran(sn).) n=1(R2,Sn − 1) < ∞. This is the main part of the proof. We claim thatP∞ n =(cid:0)Jn, t(0) To prove the claim, for all n ∈ Z>0 choose in, jn ∈ In such that sn(in) = 1 and ksn(jn)k · ksn(jn)−1k > R2,Sn − 2−n. Set Jn = {in, jn}. We define systems T (0) one element, we set t(0) n , gn(cid:1) of d(n)-similarities as follows. If in = jn, so that Jn has only n (in) = 1 and gn(in) = 1. If in 6= jn, we set t(0) n = snJn , gn(in) = fn(in) fn(in) + fn(jn) , and gn(jn) = fn(jn) fn(in) + fn(jn) . By construction, we have R2,T (0) n > R2,Sn − 2−n. Set AMENABILITY 33 J = (J1, J2, . . .), and g = (g1, g2, . . .). Define a subsystem of (d, ρ) as in Definition 2.12 by taking 1 , t(0) t(0) =(cid:0)t(0) 2 , . . .(cid:1), Zn = {1, 2, . . . , d(n)} × Jn = Xd(n),T (0) n . Let B(0) for n ∈ Z>0. Let λn and γn be as in Definition 2.12. Then λn = µd(n),T (0) be the Lp UHF algebra of the subsystem (Zn)n∈Z>0 as in Definition 2.12, which is equal to D2,d,J,t(0),g. Let κ : A → B(0) be the homomorphism of Lemma 2.13. Since A is amenable and κ has dense range, it follows from Proposition 2.3.1 of [16] that B(0) is amenable. n For n ∈ Z>0, define a function zn : Jn → Md(n) by zn(in) = 1 and, if jn 6= in, using polar decomposition to choose a unitary zn(jn) such that zn(jn)sn(jn) = [sn(jn)∗sn(jn)]1/2. Define t(1) n : Jn → inv(Md(n)) by t(1) n (i) = zn(i)sn(i) for i ∈ Jn. Then set T (1) n = 1 , t(1) (cid:0)Jn, t(1) (cid:0)t(1) n , gn(cid:1), which is a system of d(n)-similarities. Define a sequence t(1) by t(1) = 2 , . . .(cid:1), and set B(1) = D2,d,J,t(1),g. Proposition 3.20(2) implies that B(1) is isometrically isomorphic to B(0), so that B(1) is amenable, and that R2,T (1) R2,T (0) for all n ∈ Z>0. = n n For n ∈ Z>0, further choose a unitary vn ∈ Md(n) such that vnt(1) diagonal. Define tn : Jn → inv(Md(n)) by tn(i) = vn(i)t(1) Then set Tn =(cid:0)Jn, tn, gn(cid:1), which is a system of diagonal d(n)-similarities. Define a sequence t by t =(cid:0)t1, t2, . . .(cid:1), and set B = D2,d,J,t,g. Proposition 3.20(3) implies that B is isometrically isomorphic to B(1), so that B is amenable, and that R2,Tn = R2,T (0) n=1(R2,Tn −1) < ∞. Therefore for all n ∈ Z>0. From Theorem 4.10 we getP∞ n n is n (i)vn(i)∗ for i ∈ In. n (jn)v∗ (R2,Sn − 1) < ∞Xn=1 ∞Xn=1(cid:18)R2,Tn + 1 2n − 1(cid:19) < ∞. The claim is proved. For n ∈ Z>0 and i ∈ In, use the polar decomposition of sn(i)∗ to find a selfadjoint element cn(i) ∈ Md(n) and a unitary un(i) ∈ Md(n) such that sn(i) = cn(i)un(i). Then (Sn, un, fn) is a system of d(n)-similarities. We have kcn(i)k = ksn(i)k and kcn(i)−1k = ksn(i)−1k = 1, so 1 ≤ cn(i) ≤ ksn(i)k · 1. Therefore kcn(i) − 1k ≤ ksn(i)k − 1 ≤ R2,Sn − 1. Also kcn(i)−1 − 1k ≤ kcn(i)−1k · k1 − cn(i)k = k1 − cn(i)k ≤ R2,Sn − 1. Recalling that XTn = {1, 2, . . . , d(n)} × In, and applying Lemma 2.11 and Re- mark 3.7, let cn ∈ inv(L(L2(XTn , µTn )) be the direct sum over i ∈ In of the elements cn(i). Thus kcn − 1k ≤ R2,Sn − 1 and kc−1 Set u = (u1, u2, . . .). SinceP∞ and P∞ n=1 kcn − 1k < ∞ n − 1k < ∞. So Lemma 4.9 provides an invertible element v ∈ L(L2(XI,s,f , µI,s,f ) such that vAv−1 = D2,d,I,u,f . Since un(i) is unitary for all n ∈ Z>0 and i ∈ In, it is immediate that D2,d,I,u,f is a C*-subalgebra of L(L2(XI,s,f , µI,s,f ). n=1(R2,Sn − 1) < ∞, we haveP∞ n − 1k ≤ R2,Sn − 1. n=1 kc−1 Question 4.12. Let p ∈ [1, ∞). Let A be an amenable Lp UHF algebra, but not of the type to which Theorem 4.10 applies. Does it follow that A is isomorphic to a spatial Lp UHF algebra? Does it follow that A is similar to a spatial Lp UHF (cid:3) 34 N. CHRISTOPHER PHILLIPS algebra? What if we assume, say, that A is an Lp UHF algebra of tensor product type? What if we assume that A is in fact symmetrically amenable? Amenability of spatial Lp UHF algebras is used in the proof of Corollary 5.18 of [14] to prove that spatial representations of the Leavitt algebra Ld generate amenable Banach algebras. d is amenable for p ∈ [1, ∞) and d ∈ {2, 3, . . .}.) It is clear from Theorem 4.10 (and will be much more obvious from Theorem 5.14 below) that there are Lp UHF algebras of tensor product type which are not amenable. However, we do not know how to use known examples to con- struct nonamenable versions of Op d. (That is, Op Question 4.13. Let p ∈ [1, ∞) and let d ∈ {2, 3, . . .}. Does there exist a σ-finite measure space (X, B, µ) and a representation ρ of the Leavitt algebra Ld on Lp(X, µ) such that kρ(sj)k = 1 and kρ(tj)k = 1 for j = 1, 2, . . . , d, but such that ρ(Ld) is not amenable? The situation is a bit different from what we have considered here. If kρ(sj)k = 1 and kρ(tj)k = 1 for j = 1, 2, . . . , d, then the standard matrix units in the analog for ρ(Ld) of the UHF core of Od all have norm 1. We do not know whether there are nonamenable Lp UHF algebras in which all the standard matrix units have norm 1. 5. Many nonisomorphic Lp UHF algebras In this section, for fixed p ∈ (1, ∞) and a fixed supernatural number N, we prove that there are uncountably many mutually nonisomorphic Lp UHF algebras of infinite tensor product type with the same supernatural number N. We rule out not just isometric isomorphism but isomorphisms which convert the norm to an equivalent norm. Our algebras are all obtained using diagonal similarities, so are covered by Theorem 4.10. In particular, all except possibly one of them is not amenable. We do not use any explicit invariant. Rather, we give lower bounds on the norms of nonzero homomorphisms from matrix algebras into the algebras we consider, which increase with the size of the matrix algebra. We use these to prove the nonexistence of continuous homomorphisms between certain pairs of our algebras. Our construction uses a class of diagonal similarities which is easy to deal with, and which we now introduce. Notation 5.1. Let p ∈ [1, ∞), let d ∈ Z>0, and let γ ∈ [1, ∞). We define Kd,γ ⊂ inv(Md) to be the compact set consisting of all diagonal matrices in Md whose diagonal entries are all in [1, γ]. When we need a system Sd,γ = (Id,γ, sd,γ, fd,γ) of d-similarities such that ran(sd,γ) = Kd,γ, we take Id,γ to consist of all diagonal matrices in Md whose diagonal entries are all in [1, γ] ∩ Q and sd,γ to be the identity map. For each d and γ, we choose once and for all an arbitrary function fd,γ : Id,γ → (0, 1] satisfyingPi∈Id,γ f (i) = 1. Let (Y, C, ν) be a σ-finite measure space, and let A ⊂ L(Lp(Y, ν)) be a closed subalgebra. In Definition 3.21, we make the following abbreviations. d = M p,Kd,γ (1) M p,γ (2) k · kp,γ = k · kp,Kd,γ . (3) κp,γ2,γ1 d . d In particular, M p,1 = κp,Kd,γ2 ,Kd,γ1 and κp,γ2,γ1 d = M p d . d,A = κ p,Kd,γ2 ,Kd,γ1 A for γ1, γ2 ∈ [1, ∞). We adapt Lemma 3.15 and Lemma 3.16 to our current situation. Lemma 5.2. Adopt Notation 5.1. Also let (Y, C, ν) be a σ-finite measure space, and let A ⊂ L(Lp(Y, ν)) be a closed unital subalgebra. Then: AMENABILITY 35 (1) kκp,γ,1 d,A k = γ. (3) For j, k = 1, 2, . . . , d and a ∈ A, in M p,γ (2) For γ2 ≥ γ1 ≥ 1, we have(cid:13)(cid:13)κp,γ1,γ2 (cid:13)(cid:13) = 1. kej,k ⊗ akp,γ =(kak d,A j = k γkak j 6= k. d ⊗p A we have (4) For any x ∈ Md ⊗ A, we have kxkp,γ = sup v∈Kd,γ k(v ⊗ 1)x(v−1 ⊗ 1)kp. (5) Let a1, a2, . . . , ad ∈ A. Then in M p,γ d ⊗p A, we have (cid:13)(cid:13)e1,1 ⊗ a1 + e2,2 ⊗ a2 + · · · + ed,d ⊗ ad(cid:13)(cid:13)p,γ = max(cid:0)ka1k, ka2k, . . . , kadk(cid:1). The estimates in (1) and (2) mean that kxk ≤ kκp,γ,1 (x)k ≤ γkxk for all x ∈ M p d . The reason for our definition of Kd,γ is to ensure that β ≥ γ implies Kd,γ ⊂ Kd,β. If we used diagonal matrices with diagonal entries in {1, γ}, our (easy) proof that kxkp,β ≥ kxkp,γ would break down, and we do not know whether this inequality would still be true. d Proof of Lemma 5.2. For j, k ∈ {1, 2, . . . , d}, let rj,k be as in Lemma 3.16 with K = Kd,γ. Then rj,j = 1 for j = 1, 2, . . . , d by Lemma 3.16(3), and it is easy to check that rj,k = γ for j, k ∈ {1, 2, . . . , d} with j 6= k. Part (1) is now immediate from Lemma 3.16(6), and part (2) is immediate from Lemma 3.15(5). Using the computation of rj,k above, part (3) follows from Lemma 3.16(4) and (5). Part (4) is Lemma 3.15(1), and part (5) follows from Lemma 3.16(7). (cid:3) Notation 5.3. Let d = (d(1), d(2), . . .) be a sequence in {2, 3, . . .}, and let γ = (γ(1), γ(2), . . .) be a sequence in [1, ∞). Using the choices of Notation 5.1, define Id,γ = (Id(1),γ(1), Id(2),γ(2), . . .), sd,γ = (sd(1),γ(1), sd(2),γ(2), . . .), and fd,γ = (fd(1),γ(1), fd(2),γ(2), . . .). Following Definition 3.18, we now set Bp,γ d = Dp,d,Id,γ ,sd,γ ,fd,γ . When d, p, and γ are understood, we just call this algebra B. For n ∈ Z≥0, we then further set Bn = D(0,n) . Thus p,d,Id,γ ,sd,γ ,fd,γ Bn = M p,γ(1) d(1) ⊗p M p,γ(2) d(2) ⊗p · · · ⊗p M p,γ(n) d(n) . We can isometrically identify B0, B1, . . . with subalgebras of B in such a way that B0 ⊂ B1 ⊂ · · · and B =S∞ Thus, Bp,γ d n=0 Bn. is the Lp spatial infinite tensor product of the algebras M p,γ(n) . d(n) Lemma 5.4. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence in {2, 3, . . .}, and let β = (β(1), β(2), . . .) and γ = (γ(1), γ(2), . . .) be sequences in [1, ∞). Let l ∈ Z>0, and suppose that γ(j) ≥ β(j) for j = l + 1, l + 2, . . . . Then the obvious map on the algebraic direct limits extends to a continuous unital homomorphism from Bp,γ d which has dense range. to Bp,β d Proof. We follow Notation 5.3 and the notation of Definition 3.18. We apply part (2) of the conclusion of Lemma 2.13, with D = D(0,l) , to get a contractive unital homomorphism d → D(0,l) ⊗p D(l,∞) p,d,Id,γ ,sd,γ ,fd,γ τ : Bp,γ p,d,Id,β,sd,β ,fd,β p,d,Id,γ ,sd,γ ,fd,γ 36 N. CHRISTOPHER PHILLIPS which has dense range. As algebras, we have D(0,l) p,d,Id,γ ,sd,γ ,fd,γ = D(0,l) p,d,Id,β ,sd,β,fd,β = Mrd(l). Since Mrd(l) is finite dimensional, the identity map on Mrd(l) ⊗alg D(l,∞) extends to a continuous bijection D(0,l) → D(0,l) ⊗p D(l,∞) ⊗p D(l,∞) p,d,Id,β,sd,β ,fd,β p,d,Id,γ ,sd,γ ,fd,γ p,d,Id,β,sd,β ,fd,β p,d,Id,β ,sd,β ,fd,β p,d,Id,β,sd,β ,fd,β Compose this map with τ to complete the proof of the lemma. = Bp,β d . (cid:3) We now give some results on perturbations of homomorphisms from direct sums of matrix algebras. We adopt some notation from the beginning of Section 1 of [11]. Definition 5.5. Let A and D be Banach algebras, and let T : A → D be a bounded linear map. We define a bounded bilinear map T ∨ : A × A → D by T ∨(x, y) = T (xy) − T (x)T (y) for x, y ∈ A. We further define d(T ) to be the distance from T to the set of bounded homomorphisms from A to D, that is, d(T ) = inf(cid:0)(cid:8)kT − ϕk : ϕ : A → D is a continuous homomorphism(cid:9)(cid:1). Following [11], in the notation of Definition 5.5 it is easy to check that d(T ) = 0 if and only if T is a homomorphism. We recall that if E, F, and G are Banach spaces, and b : E × F → G is a bilinear map, then kbk is the least constant M such that kb(ξ, η)k ≤ M kξk ·kηk for all ξ ∈ E and η ∈ F. We next recall the following estimate. Lemma 5.6 (Proposition 1.1 of [11]). Adopt the notation of Definition 5.5. Then kT ∨k ≤(cid:0)1 + d(T ) + 2kT k(cid:1)d(T ). The following result, but with δ depending on D, is Corollary 3.2 of [11]. Lemma 5.7. Let A be a Banach algebra which is isomorphic, as a complex algebra, to a finite direct sum of full matrix algebras. The for every ε > 0 and M ∈ [0, ∞) there is δ > 0 such that, whenever D is a Banach algebra and T : A → D is a linear map with kT k ≤ M and kT ∨k < δ, then there is a homomorphism ϕ : A → D such that kϕ − T k < ε. Example 1.5 of [11] shows that it is not possible to take δ to be independent of M, even for A = C (with nonunital homomorphisms). Proof of Lemma 5.7. Let A be as in the hypotheses, let ε > 0, and let M ∈ [0, ∞). Suppose that the conclusion fails. Then for every n ∈ Z>0 there is a Banach algebra Dn and a linear map Tn : A → Dn such that kTnk ≤ M and kT ∨k < 1 n , but no homomorphism ϕ : A → Dn such that kϕ − T k < ε. Let D be the Banach algebra of all bounded sequences b = (bn)n∈Z>0 ∈Q∞ n=1 Dn, with the pointwise operations and the norm kbk = supn∈Z>0 kbnk. For n ∈ Z>0, let πn : D → Dn be given by πn(b) = bn. Corollary 3.2 of [11] provides δ > 0 such that whenever T : A → D is a linear map with kT k ≤ M and kT ∨k < δ, then there is a homomorphism ϕ : A → D such that kϕ − T k < ε. Choose n ∈ Z>0 such that 1 n < δ. Define T : A → D by T (x) = (0, 0, . . . , 0, Tn(x), Tn+1(x), . . .) for x ∈ A. Then kT k ≤ M and kT ∨k < 1 n < δ. So there is a homomorphism ϕ : A → D such that kϕ−T k < ε. The map πn◦ϕ : A → Dn is a homomorphism such that kπn ◦ ϕ − Tnk < ε, contradicting the assumption that no such homomorphism exists. (cid:3) AMENABILITY 37 The following simple lemma will be used several times. Lemma 5.8. Let A be a unital Banach algebra, let D be a Banach algebra, and let ϕ, ψ : A → D be homomorphisms such that kϕ − ψk < 1. If ϕ is nonzero then so is ψ, and if D and ϕ are unital then so is ψ. Proof. For the first statement, ϕ(1) is a nonzero idempotent in D, whence kϕ(1)k ≥ 1. Therefore ψ(1) 6= 0. For the second, ψ(1) is an idempotent in D with kψ(1)−1k < 1. Therefore ψ(1) is an invertible idempotent, so ψ(1) = 1. (cid:3) Lemma 5.9. Let A be a Banach algebra which is isomorphic, as a complex algebra, to a finite direct sum of full matrix algebras. Let ε > 0, and let M ∈ [1, ∞). Then there is δA,ε,M > 0 such that whenever D is a Banach algebra, C ⊂ D is a subalgebra, ϕ : A → D is a homomorphism such that kϕk ≤ M, and S : A → C is a linear map such that kSk ≤ M and kS−ϕk < δA,ε,M , then there is a homomorphism ψ : A → C such that kψ − ϕk < ε. If D, C, and ϕ are unital, then we may require that ψ be unital. Proof. Without loss of generality, ε < 1. Apply Lemma 5.7 with ε with A and M as given, obtaining δ0 > 0. Set 2 in place of ε and δ = min(cid:18)1, ε 2 , δ0 2(1 + M )(cid:19) . Let ϕ and S be as in the hypotheses, with δA,ε,M = δ. Lemma 5.6 implies that kS∨k < δ(1 + δ + 2M ) ≤ δ(2 + 2M ) ≤ δ0. Therefore there exists a homomorphism ψ : A → C such that kψ − Sk < ε kS − ϕk < δ ≤ ε 2 , it follows that kϕ − ψk < ε. 2 . Since It remains to prove, under the conditions in the last sentence, that ψ is unital. (cid:3) Since ε < 1, this follows from Lemma 5.8. Lemma 5.10. Let A be a Banach algebra which is isomorphic, as a complex algebra, to a finite direct sum of full matrix algebras. Let D be a Banach algebra, and let D0 ⊂ D1 ⊂ · · · ⊂ D be an increasing sequence of subalgebras such that n=0 Dn. Then for every homomorphism ϕ : A → D and every ε > 0, there is n ∈ Z≥0 and a homomorphism ψ : A → D such that kψ − ϕk < ε and ψ(A) ⊂ Dn. Moreover, if D and ϕ are unital, and Dn is unital for n ∈ Z≥0, then ψ can be chosen to be unital. D =S∞ Proof. Apply Lemma 5.9 with A as given, with kϕk + 1 in place of M, and with ε as given, obtaining δ = δA,ε,1+kϕk > 0. Set N = dim(A). Choose a basis (xk)k=1,2,...,N for A consisting of elements xk ∈ A such that kxkk = 1 for k = 1, 2, . . . , N. Define a bijection S : l1({1, 2, . . . , N }) → A by identifying l1({1, 2, . . . , N }) with CN and setting S(α1, α2, . . . , αN ) = αkxk NXk=1 for α1, α2, . . . , αN ∈ C. Set δ0 = 1 2 kS−1k−1 min(1, δ). Choose n ∈ Z≥0 such that there are b1, b2, . . . , bN ∈ Dn with kbk−ϕ(xk)k < δ0 for k = 1, 2, . . . , N. Let T : A → Dn be the unique linear map such that T (xk) = bk for k = 1, 2, . . . , N. We claim that kT − ϕk < 2kS−1kδ0. To see this, let x ∈ A. Choose k=1 αkxk. Then S−1(x) = (α1, α2, . . . , αN ), so α1, α2, . . . , αN ∈ C such that x =PN k(α1, α2, . . . , αN )k1 ≤ kS−1k · kxk. Now kT (x) − ϕ(x)k ≤ αk · kT (xk) − ϕ(xk)k ≤ NXk=1 αkδ0 ≤ kS−1k · δ0 · kxk. NXk=1 38 N. CHRISTOPHER PHILLIPS So kT − ϕk ≤ kS−1kδ0 < 2kS−1kδ0. Since 2kS−1kδ0 ≤ 1, the claim implies that kT k < kϕk + 1. Since 2kS−1kδ0 ≤ δ, we get kT −ϕk < δ, so the choice of δ, with Dn in place of C in Lemma 5.9, provides a homomorphism ψ : A → D such that kψ − ϕk < ε and ψ(A) ⊂ Dn, which is unital under the conditions in the last sentence. (cid:3) The following lemma is the key step of our argument. For any M, γ0, d0, d, → whose range is in and p, if γ is sufficiently large and there is a nonzero homomorphism ϕ : M p,γ0 M p,γ C · 1 ⊗p A and whose norm is nearly the same as that of ϕ. d ⊗p A, then there is a nonzero homomorphism from M p,γ0 d0 d0 Lemma 5.11. Let p ∈ [1, ∞), let d0, d ∈ Z>0, and let γ0 ∈ [1, ∞). Let M ∈ [1, ∞) and let ε > 0. Then there is R ∈ [0, ∞) such that whenever γ ∈ [R, ∞), the following holds. Let (Y, C, ν) be a σ-finite measure space, and let A ⊂ L(Lp(Y, ν)) be a closed subalgebra. Let ϕ : M p,γ0 d ⊗p A be a nonzero homomorphism such that kϕk ≤ M. Then there is a nonzero homomorphism ψ : M p,γ0 → A such that kψk ≤ M + ε. → M p,γ d0 d0 Proof. We use Notation 5.1 throughout. Without loss of generality, ε < 1. Let δ > 0 be the constant of Lemma 5.9 obtained using ε and M as given and with M p in place of A. Set R = 2d2M γ0/δ. Now let γ, A, and ϕ be as in the statement d0 of the lemma. Define C = M p,γ d ⊗p A and C0 = M p d ⊗p A = M p,1 d ⊗p A, with norms k·kp,γ and k·kp,1. We will also need the maps, defined as in Notation 5.1, κp,1,γ d,A : C → C0, Define T : M p,γ0 d0 : M p,γ0 κp,1,γ0 d0 d0 → M p,γ d ⊗p A by → M p d0 , and κp,γ0,1 d0 : M p d0 → M p,γ0 d0 . T (x) = (el,l ⊗ 1)ϕ(x)(el,l ⊗ 1) dXl=1 for x ∈ M p,γ0 d0 . For l = 1, 2, . . . , d, there is a linear map Tl : M p,γ0 → A such that d0 (el,l ⊗ 1)ϕ(x)(el,l ⊗ 1) = el,l ⊗ Tl(x) d0 Since kel,lkp,γ = 1, we have kTlk ≤ kϕk for l = 1, 2, . . . , d, so for all x ∈ M . Lemma 5.2(5) implies that kT k ≤ kϕk. d,A ◦T ◦κp,γ0,1 We now claim that(cid:13)(cid:13)κp,1,γ let x ∈ M p d0 such that d0 satisfy kxkp ≤ 1. There are elements al,m ∈ A for l, m = 1, 2, . . . , d −κp,1,γ d,A ◦ϕ◦κp,γ0,1 d0 (cid:13)(cid:13) < δ. To prove the claim, (cid:0)ϕ ◦ κp,γ0,1 d0 (cid:1)(x) = dXl,m=1 el,m ⊗ al,m. For l 6= m, using Lemma 5.2(3) at the first step, kel,l ⊗ 1kp,γ = kem,m ⊗ 1kp,γ = 1 at the third step, and Lemma 5.2(1) at the fifth step, we get γkal,mk = kel,m ⊗ al,mkp,γ = k(el,l ⊗ 1)ϕ(x)(em,m ⊗ 1)kp,γ ≤ kϕ(x)kp,γ ≤ kϕk · kxkp,γ0 ≤ kϕkγ0kxkp ≤ M γ0. kal,mk ≤ M γ0γ−1 and kel,m ⊗ al,mkp,1 = kal,mk ≤ M γ0γ−1. So Now d,A ◦ T ◦ κp,γ0,1 d0 (cid:13)(cid:13)(cid:0)κp,1,γ (cid:1)(x) −(cid:0)κp,1,γ d,A ◦ ϕ ◦ κp,γ0,1 d0 (cid:1)(x)(cid:13)(cid:13)p,1 ≤ dXl=1Xm6=l ≤ d2M γ0γ−1. kel,m ⊗ al,mkp,1 Thus AMENABILITY 39 = δ 2 < δ. d ⊗p A be the subalgebra consisting of all diagonal matrices l=1 A via the map (a1, a2, . . . , ad) 7→ d0 d0 γ ≤ R d2M γ0 d2M γ0 The claim is proved. d,A ◦ T ◦ κp,γ0,1 l=1 A with the norm − κp,1,γ d,A ◦ ϕ ◦ κp,γ0,1 Let D ⊂ C0 = M p then isometric. By construction, the range of κp,1,γ claim above and the choice of δ provide a homomorphism ψ0 : M p d0 (cid:13)(cid:13) ≤ (cid:13)(cid:13)κp,1,γ in Md(A). We algebraically identify D withLd l=1 A to C0. EquipLd l=1 el,l ⊗ al fromLd Pd k(a1, a2, . . . , ad)k = max(cid:0)ka1k, ka2k, . . . , kadk(cid:1) for a1, a2, . . . , ad ∈ A. By Lemma 5.2(5), the identification of Ld (cid:13)(cid:13)ψ0 − κp,1,γ (cid:13)(cid:13)κp,1,γ0 So, using(cid:13)(cid:13)κp,1,γ For l = 1, 2, . . . , d and a = Pd (cid:13)(cid:13) = 1 (from Lemma 5.2(2)) at the second step, we get (cid:13)(cid:13) ·(cid:13)(cid:13)κp,1,γ0 (cid:13)(cid:13)ψ1 − κp,1,γ d,A (cid:13)(cid:13) = 1 (from Lemma 5.2(2)), we get d,A ◦ ϕ(cid:13)(cid:13) + ε ≤ kϕk + ε. (cid:13)(cid:13) < ε. Since κp,1,γ d,A ◦ ϕ(cid:13)(cid:13) ≤(cid:13)(cid:13)ψ0 − κp,1,γ kψ1k <(cid:13)(cid:13)κp,1,γ and ε < 1, it follows from Lemma 5.8 that ψ0 is nonzero. Set ψ1 = ψ0 ◦ κp,1,γ0 d,A ◦ ϕ ◦ κp,γ0,1 d,A ◦ ϕκp,γ0,1 d0 d,A ◦ ϕ ◦ κp,γ0,1 d0 = idM p,γ0 d0 . Using κp,γ0,1 d0 ◦ κp,1,γ0 d0 d0 d0 d0 d0 (cid:13)(cid:13) < ε. l=1 A with D is d,A ◦ T is contained in D. The → D such that is a nonzero homomorphism at the first step, and k=1 ek,k ⊗ ak ∈ D, define πl(a) = al. The formula defines a contractive homomorphism πl : D → A. Choose l such that (πl◦ψ1)(1) 6= 0. Then ψ = πl ◦ ψ1 : M p → A is a nonzero homomorphism such that kψk < kϕk + ε. d0 This completes the proof. (cid:3) Recall that for a sequence d = (d(1), d(2), . . .) in {2, 3, . . .} and n ∈ Z≥0, we defined rd(n) = d(1)d(2) · · · d(n). Lemma 5.12. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence in {2, 3, . . .}, and let α ∈ [1, ∞). Then for every M ∈ [1, ∞) and l ∈ Z>0, there is a nondecreasing sequence β = (β(1), β(2), . . .) in [1, ∞) such that, whenever γ = (γ(1), γ(2), . . .) is a nondecreasing sequence in [1, ∞) such that γ(j) ≥ β(j) for j = l, l + 1, . . . , whenever Bp,γ is a nonzero homomorphism, then kϕk > M. is as in Notation 5.3, and whenever ϕ : M p,α rd(l) → Bp,γ d d Proof. For m = l, l + 1, . . . , apply Lemma 5.11 with rd(l) in place of d0, with d(m) in place of d, with α in place of γ0, with 2−(m−l+1) in place of ε, and with M + 1 + 2−(m−l+1) in place of M. Let β0(m) be the resulting value of R. For m = 1, 2, . . . , l − 1 set β(m) = 1, and for m = l, l + 1, . . . set β(m) = max(cid:0)β0(l), β0(l + 1), . . . , β0(m)(cid:1). Now let γ be a nondecreasing sequence in [1, ∞) such that γ(j) ≥ β(j) for rd(l) → Bp,γ j = l, l + 1, . . . . Suppose that there is a nonzero homomorphism ϕ : M p,α such that kϕk ≤ M. Let B = Bp,γ d , B0, B1, . . . be as in Notation 5.3. Lemma 5.10 provides n ∈ Z≥0 and a homomorphism ψ : M p,α rd(l) → Bn such that kψ − ϕk < 1. Then ψ 6= 0 by Lemma 5.8. Therefore n ≥ l. Also, kψk ≤ M +1 ≤ M +1+2−(n−l+1). Set ψn = ψ. Using choice of β0(n), the inequality γ(n) ≥ β(n) ≥ β0(n), and the , we get a nonzero homomor- tensor product decomposition Bn = Bn−1 ⊗p M p,γ(n) phism ψn−1 : M p,α rd(l) → Bn−1 such that d(n) d kψn−1k ≤(cid:0)M + 1 + 2−(n−l+1)(cid:1) + 2−(n−l+1) = M + 1 + 2−(n−l). 40 N. CHRISTOPHER PHILLIPS Similarly, there is now a nonzero homomorphism ψn−2 : M p,α rd(l) → Bn−2 such that kψn−2k ≤ M + 1 + 2−(n−l−1). Proceed inductively. We eventually find a nonzero homomorphism ψl−1 : M p,α Bl−1 such that rd(l) → Since Bl−1 ∼= Mrd(l−1) and rd(l − 1) < rd(l), this is a contradiction. (cid:3) kψl−1k ≤ M + 1 + 1. Lemma 5.13. Let p ∈ [1, ∞), let d = (d(1), d(2), . . .) be a sequence in {2, 3, . . .}, and let α = (α(1), α(2), . . .) be a nondecreasing sequence in [1, ∞). Then there is a nondecreasing sequence β = (β(1), β(2), . . .) in [1, ∞) such that, whenever l ∈ Z>0 and γ = (γ(1), γ(2), . . .) is a nondecreasing sequence in [1, ∞) with γ(j) ≥ β(j) for j = l, l + 1, . . . , and Bp,α are as in Notation 5.3, then there is no nonzero continuous homomorphism from Bp,α and Bp,γ d d to Bp,γ d . d Proof. For each m ∈ Z>0, apply Lemma 5.12 with d as given, with m in place k=1 α(k) in place of α, and with M = m. Call the resulting sequence of l, with Qm βm =(cid:0)βm(1), βm(2), . . .(cid:1). Define β(m) = max(cid:0)βm(1), βm(2), . . . , βm(m)(cid:1) for m ∈ Z>0. Clearly β = (β(1), β(2), . . .) is a nondecreasing sequence in [1, ∞). Now let l ∈ Z>0 and let γ = (γ(1), γ(2), . . .) be a nondecreasing sequence in [1, ∞) such that γ(j) ≥ β(j) for j = l, l + 1, . . . . Suppose that ϕ : Bp,α is a nonzero continuous homomorphism. Choose m ∈ Z>0 such that m > max(l, kϕk). d → Bp,γ d Set η =Qm k=1 α(k). Let ψ : Mrd(m) → Md(1) ⊗ Md(2) ⊗ · · · ⊗ Md(m) be an isomorphism which sends standard matrix units to tensor products of stan- dard matrix units. In particular, the image of the diagonal subalgebra of Mrd(m) is the tensor product of the diagonal subalgebras of the algebras Md(k) for k = 1, 2, . . . , m. We claim that ψ is a contractive homomorphism ψ : M p,η rd(m) → M p,α(1) d(1) ⊗p M p,α(2) d(2) ⊗p · · · ⊗p M p,α(m) d(m) . To see this, apply Lemma 3.17 repeatedly to form the tensor product S = (I, s, f ) of the systems Sd(k),α(k) of Notation 5.1 for k = 1, 2, . . . , m. Also let Srd(m),η be k=1 Id(k),α(k), the matrix s(i) is diagonal and as in Notation 5.1. For i ∈ I = Qm satisfies ks(i)k =(cid:13)(cid:13)sd(1),α(1)(i1) ⊗ sd(2),α(2)(i2) ⊗ · · · ⊗ sd(m),α(m)(im)(cid:13)(cid:13) ksd(k),α(k)(ik)k ≤ α(k) = η. = mYk=1 mYk=1 Moreover, all its diagonal entries are real and at least 1. Therefore ran(s) ⊂ Krd(m),η = ran(srd(m),η), so ψ is contractive by Lemma 3.15(5). As in Notation 5.3, we have an isometric inclusion M p,α(1) d(1) ⊗p M p,α(2) d(2) ⊗p · · · ⊗p M p,α(m) d(m) → Bp,α d . Call it ι. Since kψk ≤ 1, the map σ = ϕ ◦ ι ◦ ψ : M p,η rd(m) → Bp,γ d AMENABILITY 41 is a nonzero homomorphism such that kσk ≤ kϕk. Since m ≥ l, for k = m, m+1, . . . we have γ(k) ≥ β(k) ≥ βm(k). The choice of βm therefore implies that kσk ≥ m. Since m > kϕk, this is a contradiction. (cid:3) Theorem 5.14. Let p ∈ [1, ∞), and let d = (d(1), d(2), . . .) be a sequence in {2, 3, . . .}. Let Ω be the set of countable ordinals. Then there exists a family (γκ)κ∈Ω of nondecreasing sequences in [1, ∞) such that whenever κ, λ ∈ Ω satisfy κ < λ, then there is no nonzero continuous homomorphism from Bp,γκ , but there is a unital continuous homomorphism from Bp,γλ d with dense range. to Bp,γκ to Bp,γλ d d d Proof. We construct (γκ)κ∈Ω by transfinite induction on κ. We start by taking γ0 = (1, 1, . . .). Suppose now κ ∈ Ω and we have the sequences γµ for µ < κ. First suppose that κ is a successor ordinal, that is, there is λ ∈ Ω such that κ = λ+1. Apply Lemma 5.13 with p and d as given and with γλ in place of α, obtaining a nondecreasing se- quence β = (β(1), β(2), . . .) in [1, ∞). Set γκ(n) = max(β(n), γλ(n)) for n ∈ Z>0. Then there is no nonzero continuous homomorphism from Bp,γλ . How- d d → Bp,γλ ever, Lemma 5.4 provides a continuous unital homomorphism ψ : Bp,γκ with dense range. For any ordinal µ < λ, the induction hypothesis provides a continuous unital homomorphism from Bp,γλ with dense range, giving a continuous unital homomorphism from Bp,γκ , again with dense range. Suppose now µ ∈ Ω satisfies µ < λ and that there is a nonzero continuous homomor- phism ϕ : Bp,γµ . Then ψ ◦ ϕ is a nonzero continuous homomorphism from Bp,γµ to Bp,γλ . We have contradicted the induction hypothesis, and the successor ordinal case of the induction step is complete. to Bp,γµ to Bp,γµ d → Bp,γκ to Bp,γκ d d d d d d d d d Now suppose that κ is a limit ordinal. Let (λn)n∈Z>0 be an enumeration of {λ ∈ Ω : λ < κ} (in arbitrary order). For n ∈ Z>0, apply Lemma 5.13 with p and d as given and with γλn in place of α, obtaining a nondecreasing sequence βn = (βn(1), βn(2), . . .) in [1, ∞). Now recursively define γκ(1) = max(β1(1), γλ1 (1)) and γκ(n) = max(cid:0)γκ(n − 1), βn(1), βn(2), . . . , βn(n), γλ1 (n), γλ2 (n), . . . , γλn (n)(cid:1) for n ≥ 2. We verify that γκ satisfies the required conditions. So let λ ∈ Ω satisfy λ < κ. Choose n ∈ Z>0 such that λn = λ. Then γκ(k) ≥ γλ(k) for k = n, n + 1, . . . , so Lemma 5.4 provides a continuous unital homomorphism Bp,γκ with dense range. Also, γκ(k) ≥ βn(k) for k = n, n + 1, . . . , so the choice of βn using Lemma 5.13 ensures that there is no nonzero continuous homomorphism from Bp,γκ to Bp,γλ (cid:3) . This completes the proof. d → Bp,γλ d d d Problem 5.15. Let p ∈ [1, ∞). For a given supernatural number N, give invariants which classify up to isomorphism some reasonable class of nonspatial Lp UHF alge- bras of tensor product type, such as those constructed using diagonal similarities. References [1] D. P. Blecher and C. Le Merdy, Operator algebras and their modules -- an operator space approach, London Mathematical Society Monographs, New Series, no. 30, Oxford Science Publications, The Clarendon Press, Oxford University Press, Oxford, 2004. [2] Y. Choi, On commutative, operator amenable subalgebras of finite von Neumann algebras, J. reine angew. Math. 678(2013), 201 -- 222. [3] A. Connes, On the cohomology of operator algebras, J. Funct. Anal. 28(1978), 248 -- 253. [4] P. C. Curtis, Jr. and R. J. Loy, A note on amenable algebras of operators, Bull. Austral. Math. Soc. 52(1995), 327 -- 329. [5] E. G. Effros and J. Rosenberg, C*-algebras with approximately inner flip, Pacific J. Math. 77(1978), 417 -- 443. [6] I. Farah and N. Ozawa, A nonseparable amenable operator algebra which is not isomorphic to a C*-algebra, preprint (arXiv: 1309.2145v3 [math.OA]). 42 N. CHRISTOPHER PHILLIPS [7] D. R. Farenick, B. E. Forrest, and L. W. Marcoux, Amenable operators on Hilbert spaces, J. reine angew. Math. 582(2005), 201 -- 228. [8] D. R. Farenick, B. E. Forrest, and L. W. Marcoux, Erratum: "Amenable operators on Hilbert spaces" [J. reine angew. Math. 582(2005), 201 -- 228] , J. reine angew. Math. 602(2007), 235. [9] J. A. Gifford, Operator algebras with a reduction property, J. Australian Math. Soc. 80(2006), 297 -- 315. [10] U. Haagerup, All nuclear C*-algebras are amenable, Invent. Math. 74(1983), 305 -- 319. [11] B. E. Johnson, Approximately multiplicative maps between Banach algebras, J. London Math. Soc. (2) 37(1988), 294 -- 316. [12] N. Ozawa, Dixmier approximation and symmetric amenability for C*-algebras, preprint (arXiv:1304.3523v1 [math.OA]). [13] N. C. Phillips, Analogs of Cuntz algebras on Lp spaces, preprint (arXiv: 1201.4196v1 [math.FA]). [14] N. C. Phillips, Simplicity of UHF and Cuntz algebras on Lp spaces, preprint (arXiv: 1309.0115v2 [math.FA]). [15] N. C. Phillips, Crossed products of Lp operator algebras and the K-theory of Cuntz algebras on Lp spaces, in preparation. [16] V. Runde, Lectures on Amenability, Springer-Verlag Lecture Notes in Math. no. 1774, Springer-Verlag, Berlin, 2002. Department of Mathematics, University of Oregon, Eugene OR 97403-1222, USA. E-mail address: [email protected]
1909.12683
1
1909
2019-09-27T13:38:44
A study of symmetric points in Banach spaces
[ "math.FA" ]
We completely characterize the left-symmetric points, the right-symmetric points, and, the symmetric points in the sense of Birkhoff-James, in a Banach space. We obtain a complete characterization of the left-symmetric (right-symmetric) points in the infinity sum of two Banach spaces, in terms of the left-symmetric (right-symmetric) points of the constituent spaces. As an application of this characterization, we explicitly identify the left-symmetric (right-symmetric) points of some well-known three-dimensional polyhedral Banach spaces.
math.FA
math
A STUDY OF SYMMETRIC POINTS IN BANACH SPACES DEBMALYA SAIN, SAIKAT ROY, SATYA BAGCHI, VITOR BALESTRO Abstract. We completely characterize the left-symmetric points, the right- symmetric points, and, the symmetric points in the sense of Birkhoff-James, in a Banach space. We obtain a complete characterization of the left-symmetric (right-symmetric) points in the infinity sum of two Banach spaces, in terms of the left-symmetric (right-symmetric) points of the constituent spaces. As an application of this characterization, we explicitly identify the left-symmetric (right-symmetric) points of some well-known three-dimensional polyhedral Ba- nach spaces. 1. Introduction. The purpose of this short note is to study the left-symmetric points and the right-symmetric points of a Banach space. We also explore the left-symmetric points and right-symmetric points of the infinity sum of two Banach spaces and ap- ply the obtained results to certain polyhedral Banach spaces. Let us first establish the relevant notations and the terminologies to be used throughout the article. Throughout the text, we use the symbols X, Y, Z to denote Banach spaces. In Y denote the in- this article we work with only real Banach spaces. Let X L∞ finity sum of X and Y, endowed with the norm k(x, y)k = max{kxk, kyk} for any Y. Since there is no chance of confusion, we use the same symbol k k (x, y) ∈ X L∞ Y. Similarly, we use the same symbol to denote the norm function in X, Y and X L∞ 0 to denote the zero vector of any Banach space. Given x ∈ X and r > 0, let B(x, r) denote the open ball with center at x and radius r. Let BX = {x ∈ X : kxk ≤ 1} and SX = {x ∈ X : kxk = 1} be the unit ball and the unit sphere of X respectively. We say that X is polyhedral if BX have only finitely many extreme points. Let EX denote the set of all extreme points of the unit ball BX. A vector x ∈ X is said to be Birkhoff-James orthogonal to a vector y ∈ X, written as x ⊥B y, if kx + λyk ≥ kxk for all λ ∈ R. (see [1] and [3]) Clearly, this definition is applicable in any Banach space, in particular, in the infinity sum of two given Banach spaces. It is easy to see that Birkhoff-James orthogonality is homoge- neous, i.e., x ⊥B y implies that αx ⊥B βy, for any two scalars α, β ∈ R. However, since Birkhoff-James orthogonality is not symmetric in general, i.e., x ⊥B y does not necessarily imply that y ⊥B x, the notions of left-symmetric points and the 2010 Mathematics Subject Classification. Primary 46B20, Secondary 52A21. Key words and phrases. Birkhoff-James orthogonality; left-symmetric point; right-symmetric point; polyhedral Banach spaces. The research of Dr. Debmalya Sain is sponsored by Dr. D. S. Kothari Postdoctoral Fellowship under the mentorship of Professor Gadadhar Misra. The research of Mr. Saikat Roy is supported by CSIR MHRD in terms of Junior research Fellowship. 1 2 SAIN, ROY, BAGCHI AND BALESTRO right-symmetric points were introduced in [12]. Let us recall the relevant defini- tions here. An element x ∈ X is said to be a left-symmetric point in X if for any y ∈ X, x ⊥B y implies that y ⊥B x. On the other hand, x ∈ X is said to be a right-symmetric point in X if for any y ∈ X, y ⊥B x implies that x ⊥B y. x ∈ X is said to be a symmetric point in X if x is both left-symmetric and right-symmetric. In this article, we characterize the left-symmetric points and the right-symmetric points in a given Banach space. In order to obtain the required characterization, we require the following two notions introduced in [12]. Given any vector x ∈ X, we define the sets x+ and x− by the following: we say that y ∈ x+ if kx+λyk ≥ kxk for all λ ≥ 0, and we have that y ∈ x− if kx + λyk ≥ kxk for all λ ≤ 0. Basic properties of these two notions, their connection with Birkhoff-James orthogonality and some applications of it have been explored in [11]. In particular, it is easy to see that the set of Birkhoff-James orthogonal vectors of x is given in terms of x+ and x− by x⊥ := {y ∈ X : x ⊥B y} = x+ ∩ x−. As an application of the results obtained by us, we explicitly identify the left-symmetric points and the right-symmetric points of a three-dimensional polyhedral Banach space whose unit ball is a right prism whose base is a regular 2n-gon. In this direction, the first step is to characterize the left-symmetric points and the right-symmetric points of the analogous two- dimensional space, whose unit ball is a regular 2n-gon. We would like to remark that in the two-dimensional case, the notions of left-symmetric points and right- symmetric points may be viewed as a local generalization of Radon planes [10], i.e., two-dimensional normed planes where Birkhoff-James orthogonality is symmetric. We refer the readers to [2, 5, 8] for some of the extensive studies on Radon planes. It has been proved in [5] that if X is a two-dimensional polyhedral Banach space whose unit ball is a regular 2n-gon then X is a Radon plane if and only if n is odd. Therefore, it is perhaps not surprising that in the three-dimensional case of the unit ball being a right prism with regular 2n-gon as its base, the description of the left-symmetric points and the right-symmetric points depends on whether n is even or odd. 2. Main Results We begin with a complete characterization of the left-symmetric points of a Banach space. Theorem 2.1. Let X be a real Banach space. Then x ∈ X is a left-symmetric point in X if and only if given any u ∈ X, the following two conditions hold true: (i) u ∈ x− implies that x ∈ u−, (ii) u ∈ x+ implies that x ∈ u+. Proof. Let us first prove the sufficient part of the theorem. Let u ∈ X be such that x ⊥B u. Equivalently, u ∈ x+ ∩ x−. Suppose by contradiction that u 6⊥B x. Then either x /∈ u+ or x /∈ u−. In either case, we arrive at a contradiction to our initial assumption. This completes the proof of the sufficient part of the theorem. We now prove the necessary part of the theorem. We prove (i) and note that the proof of (ii) can be completed similarly. Let us assume that u ∈ x−. Without loss of generality, let kuk = 1. Now, there are two possible cases: A STUDY OF SYMMETRIC POINTS IN BANACH SPACES 3 Case I: Let u ∈ x+. In that case, we have that x ⊥B u. Since x is a left-symmetric point, we have u ⊥B x. In particular, x ∈ u−. Case II: Let u /∈ x+. Then it follows from Proposition 2.1 of [11] that u ∈ x− \ x⊥. Let A = {αx + βu : kαx + βuk = 1, α ≥ 0, β ≥ 0}. In other words, A is the closed shorter arc on SX ∩ span{u, x} between u and x. Let B = {β : ∃ α ∈ R such that αx + βu ∈ A and (αx + βu) ∈ x− \ x⊥}. Clearly, 1 ∈ B and B is a bounded subset of R. Therefore, inf B exists. Let inf B = β0. Then from definition of infimum, there exists a sequence {αnx + βnu} ∈ A such that {αnx + βnu} ∈ x− \ x⊥ and βn −→ β0. Since {αn} is clearly bounded, we get that it has a convergent subsequence. Hence, passing to such a subsequence if necessary we may assume that αn −→ α0 ≥ 0. Then {αnx + βnu} −→ α0x + β0u. Moreover, since A is closed, α0x + β0u ∈ A. Now, it follows from the continuity of the norm function that if z ∈ SX is such that y ∈ z− \ z⊥ then there exists δ > 0 so that for every w ∈ B(z, δ), we have that y ∈ w− \ w⊥. Therefore, we obtain that α0x + β0u /∈ x− \ x⊥, as otherwise the definition of infimum will be violated from the previous observation. On the other hand, since (αnx + βnu) ∈ x− \ x⊥, once again it follows from the continuity of the norm function that (α0x + β0u) ∈ x−. Combining these two observations, it is now easy to deduce that (α0x + β0u) ∈ x⊥. We note that 0 < β0 ≤ 1. Suppose by contradiction that x /∈ u−. We claim that it follows from x /∈ u− that (α0x + β0u) 6⊥B x. In order to prove our claim, we note that (α0x + β0u) ⊥B x implies that k(α0x + β0u) + λxk ≥ k(α0x + β0u)k, for all λ ∈ R. Therefore, k (α0+λ) ≥ kuk for all λ ∈ R (here, we recall that α0x + β0u ∈ A, meaning it is a unit vector, and we also remember that we are assuming that kuk = 1). Taking (α0+λ) = µ, we have that kµx + uk ≥ kuk for all µ ∈ R, i.e., u ⊥B x. In other words, x ∈ u+ ∩ u−, which is in contradiction to x /∈ u−. It follows that we have indeed that (α0x + β0u) 6⊥B x. However, since x ⊥B (α0x+β0u), this contradicts the fact that x is a left-symmetric point in X. Consequently, we must have that x ∈ u−. This completes the proof of (cid:3) the necessary part of the theorem and thereby establishes it completely. k(α0x + β0u)k = 1 β0 β0 x + uk ≥ 1 β0 β0 Our next aim is to completely characterize the right-symmetric points of a Ba- nach space. Theorem 2.2. Let X be a real Banach space. Then x ∈ X is a right-symmetric point in X if and only if given any u ∈ X, the following two conditions hold true: (i) x ∈ u− implies that u ∈ x−, (ii) x ∈ u+ implies that u ∈ x+. Proof. Let us first prove the sufficient part of the theorem. Let u ∈ X be such that u ⊥B x. Equivalently, x ∈ u+ ∩ u−. Suppose by contradiction that x 6⊥B u. Then either u /∈ x+ or u /∈ x−. In either case, we arrive at a contradiction to our initial assumption. This completes the proof of the sufficient part of the theorem. We now prove the necessary part of the theorem. We prove (i) and note that the proof of (ii) can be completed similarly. Let us assume that x ∈ u−. Now, there are two possible cases: Case I: Let x ∈ u+. In that case, we have that u ⊥B x. Since x is a right- symmetric point in X, it follows that x ⊥B u. In particular, u ∈ x−. 4 SAIN, ROY, BAGCHI AND BALESTRO Case II: Let x /∈ u+. Then it follows from Proposition 2.1 of [11] that x ∈ u− \ u⊥. Suppose by contradiction that u /∈ x−. As in the proof of the previous theorem, let A = {αu + βx : kαu + βxk = 1, α ≥ 0, β ≥ 0}, and recall also that A is the closed shorter arc on SX ∩ span{u, x} between u and x. Let B = {β : ∃ α ∈ R such that αu + βx ∈ A and x ∈ (αu + βx)− \ (αu + βx)⊥}. It is easy to check that B is a bounded subset of R and 0 ∈ B. Let β0 = sup B. It follows from the definition of supremum that there exists a sequence {αnu + βnx} in A such that x ∈ (αnu + βnx)− \ (αnu + βnx)⊥ and βn −→ β0. Since {αn} is clearly bounded, we get that it has a convergent subsequence. Hence, passing to such a subsequence if necessary we may assume that αn −→ α0 ≥ 0. Then {αnu + βnx} −→ α0u + β0x. Moreover, it is easy to see that α0u + β0x ∈ A. Now, it follows from the continuity of the norm function that if z ∈ SX is such that x ∈ z− \ z⊥ then there exists δ > 0 so that for every w ∈ B(z, δ), we have that x ∈ w− \ w⊥. Therefore, by using the definition of supremum, we obtain that x /∈ (α0u + β0x)− \ (α0u + β0x)⊥. On the other hand, since x ∈ (αnu + βnx)−, once again it follows from the continuity of the norm function that x ∈ (α0u + β0x)−. Combining these two observations, it is now easy to deduce that x ∈ (α0u + β0x)⊥. We note that α0, β0 > 0. In Theorem 2.1 of [14], taking ǫ = 0 it is easy to see that x⊥ = x+ ∩ x− is a normal cone in X. We recall that a subset K of X is said to be a normal cone in X if (1) K + K ⊂ K, (2) αK ⊂ K for all α ≥ 0 and (3) K ∩ (−K) = {0}. Since u ∈ x+ \ x−, it follows that if v ∈ SX is such that x ⊥B v, then v must be of the form v = γ0u + δ0x, where either of the following is true: (a) γ0 > 0 and δ0 < 0, (b) γ0 < 0 and δ0 > 0. In particular, it follows that x 6⊥B (α0u + β0x), since α0, β0 > 0. However, since (α0u + β0x) ⊥B x, this contradicts our assumption that x is a right-symmetric point in X. Therefore, it must be true that u ∈ x−. This completes the proof of the (cid:3) necessary part of the theorem and thereby establishes it completely. We note that as a consequence of Theorem 2.1 and Theorem 2.2, it is possible to completely characterize symmetric points of a Banach space. We record this useful observation in the next theorem. The proof of the theorem is omitted as it is now obvious. Theorem 2.3. Let X be a real Banach space. Then x ∈ X is a symmetric point in X if and only if given any u ∈ X, the following four conditions hold true: (i) u ∈ x− implies that x ∈ u−, (ii) u ∈ x+ implies that x ∈ u+, (iii) x ∈ u− implies that u ∈ x−, (iv) x ∈ u+ implies that u ∈ x+. We now turn our attention towards obtaining an explicit characterization of the left-symmetric points and the right-symmetric points of the infinity sum of two given Banach spaces. Since Birkhoff-James orthogonality is homogeneous, it A STUDY OF SYMMETRIC POINTS IN BANACH SPACES 5 is clearly sufficient to describe the left-symmetric points and the right-symmetric points on the unit sphere of the concerned space. First we characterize the left- symmetric points of the infinity sum of two Banach spaces. Theorem 2.4. Let X, Y be real Banach spaces and let Z = X L∞ SZ is a left-symmetric point in Z if and only if either of the following is true: (i) x ∈ SX is a left-symmetric point in X and y = 0, (ii) x = 0 and y ∈ SY is a left-symmetric point in Y. Y. Then (x, y) ∈ Proof. Let us first prove the sufficient part of the theorem. We first assume that (i) Y, i.e., max{kx + λuk, kλvk} ≥ kxk = 1. holds true. Let (x, 0) ⊥B (u, v) in X L∞ This shows that for any λ with sufficiently small absolute value, kx + λuk ≥ kxk. Using the convexity of the norm function, it is now easy to verify that kx+λuk ≥ kxk for any λ, i.e., x ⊥B u. Since x is a left-symmetric point in X, it follows that u ⊥B x. Now, for any λ ∈ R, we have the following: k(u, v) + λ(x, 0)k = max{ku + λxk, kvk} ≥ max{kuk, kvk} = k(u, v)k. In other words, (u, v) ⊥B (x, 0) whenever (x, 0) ⊥B (u, v). This proves that (x, 0) is a left-symmetric point in Z. Similarly, assuming that (ii) holds true, we can show that (0, y) is a left-symmetric point in Z. This completes the proof of the sufficient part of the theorem. Let us now prove the necessary part of the theorem. Let (x, y) ∈ SZ be a left-symmetric point in Z. Clearly, either kxk = 1 or kyk = 1. Without loss of generality, let us assume that kxk = 1. We note that kyk ≤ 1. We claim that y = 0. Suppose by contradiction that y 6= 0. Since max{kxk, ky + λyk} ≥ kxk = k(x, y)k = 1, it follows that (x, y) ⊥B (0, y). However, for any λ < 0 with sufficiently small absolute value, we have that k(0, y) + λ(x, y)k = max{kλxk, ky + λyk} < kyk = k(0, y)k. In other words, (0, y) 6⊥B (x, y). Clearly, this contradicts our assumption that (x, y) is a left-symmetric point in Z. So, we conclude that if (x, y) ∈ SZ is a left-symmetric point in Z with kxk = 1 then y = 0. In that case, we also claim that x is a left-symmetric point in X. Suppose by contradiction that there exists u ∈ SX such that x ⊥B u but u 6⊥B x. It is easy to see that either x /∈ u+ or x /∈ u−. Without loss of generality, let us assume that x /∈ u+. Then ku + λxk < kuk = 1, for some λ > 0. Using this fact, it is easy to verify that (x, 0) ⊥B (u, 0) but (u, 0) 6⊥B (x, 0). Once again, this contradicts our assumption that (x, 0) is a left-symmetric point in Z. This contradiction completes the proof of our claim that x is a left-symmetric point in X. Thus we see that all the conditions stated in (i) are satisfied. Similarly, assuming that kyk = 1, we can show that all the conditions stated in (ii) hold true. This completes the proof of the necessary (cid:3) part of the theorem and establishes it completely. We next characterize the right-symmetric points of the infinity sum of two Ba- nach spaces. We require the following lemma to serve our purpose. Y. Let (u, v) ∈ SZ Lemma 2.1. Let X, Y be real Banach spaces and let Z = X L∞ be such that kuk = kvk = 1. If (x, y) ∈ Z is such that (u, v) ⊥B (x, y) then either of the following is true: (i) x ∈ u+ and y ∈ v−, (ii) x ∈ u− and y ∈ v+. Proof. It follows from Proposition 2.1 of [11] that either x ∈ u+ or x ∈ u−. Without loss of generality, let us assume that x ∈ u+. Let us consider the following two cases: 6 SAIN, ROY, BAGCHI AND BALESTRO Case I: Let x ∈ u−. By the same argument, either y ∈ v+ or y ∈ v−. In each of these two sub-cases, it is easy to see that the statement of the lemma holds true. Case II: Let x /∈ u−. Then ku + λxk < kuk = 1, whenever λ < 0 is of sufficiently small absolute value. Since (u, v) ⊥B (x, y), we have that k(u, v) + λ(x, y)k = max{ku + λxk, kv + λyk} ≥ k(u, v)k = 1 for any λ ∈ R. Therefore, whenever λ < 0 is of sufficiently small absolute value, we must have that kv + λyk ≥ kvk = 1. Using the convexity of the norm function, it is now easy to deduce that kv+λyk ≥ kvk = 1 for all λ < 0. In other words, y ∈ v−. This completes the proof of the lemma. (cid:3) Now the promised characterization: Y. Then (x, y) ∈ Theorem 2.5. Let X, Y be real Banach spaces and let Z = X L∞ SZ is a right-symmetric point in Z if and only if kxk = kyk = 1 and x, y are right-symmetric points in X and Y respectively. Proof. Let us first prove the sufficient part of the theorem. Let (u, v) ∈ Z be such that (u, v) ⊥B (x, y). Let us consider the following two cases: Case I: Let kuk 6= kvk. Without loss of generality, we assume that kuk > kvk. Since (u, v) ⊥B (x, y), we have that max{ku + λxk, kv + λyk} ≥ k(u, v)k = kuk for all λ ∈ R. Therefore, whenever λ ∈ R is of sufficiently small absolute value, it is easy to see that ku + λxk ≥ kuk. As argued before, using the convexity of the norm function, it can be proved that ku + λxk ≥ kuk for all λ ∈ R, i.e., u ⊥B x. Since x is a right-symmetric point in X, we have that x ⊥B u. Thus we can deduce the following for any λ ∈ R : k(x, y) + λ(u, v)k = max{kx + λuk, ky + λvk} ≥ kx + λuk ≥ kxk = k(x, y)k = 1. This proves that (x, y) ⊥B (u, v). Case II: Let kuk = kvk. If kuk = kvk = 0 then it is immediate that (x, y) ⊥B (0, 0) = (u, v). Without loss of generality, let us assume that kuk = kvk = 1. Applying Lemma 2.1, we observe that either of the following must hold true: (i) x ∈ u+ and y ∈ v−, (ii) x ∈ u− and y ∈ v+. Let us first assume that (i) holds true. Since x, y are right-symmetric points in X and Y respectively, applying Theorem 2.2, it follows that u ∈ x+ and v ∈ y−. Now, for any λ ≥ 0, the following holds true: k(x, y) + λ(u, v)k = max{kx + λuk, ky + λvk} ≥ kx + λuk ≥ kxk = k(x, y)k = 1. On the other hand, for any λ ≤ 0, the following holds true: k(x, y) + λ(u, v)k = max{kx + λuk, ky + λvk} ≥ ky + λvk ≥ kyk = k(x, y)k = 1. This proves that (x, y) ⊥B (u, v). Similarly, assuming that (ii) holds true, we can deduce the same conclusion. Therefore, (x, y) ⊥B (u, v), whenever (u, v) ⊥B (x, y). In other words, (x, y) is a right-symmetric point in Z. This completes the proof of the sufficient part of the theorem. Let us now prove the necessary part of the theorem. Let (x, y) ∈ SZ be a right-symmetric point in Z. Clearly, kyk ≤ 1. We claim that kyk = 1. Suppose by contradiction that kyk < 1. We note that under this assumption, we have that kxk = 1. Choose w ∈ SY such that y ∈ w+. Now, for any λ ≥ 0, we have that A STUDY OF SYMMETRIC POINTS IN BANACH SPACES 7 k(−x, w) + λ(x, y)k ≥ kw + λyk ≥ kwk = 1 = k(−x, w)k. On the other hand, for any λ < 0, we have that k(−x, w)+λ(x, y)k ≥ k−x+λxk = 1−λ > k(−x, w)k. This proves that (−x, w) ⊥B (x, y). However, when λ > 0 is of sufficiently small absolute value, we have that k(x, y) + λ(−x, w)k = max{kx − λxk, ky + λwk} = k(1 − λ)xk = 1 − λ < 1 = k(x, y)k. In particular, this implies that (x, y) 6⊥B (−x, w). This contradicts our assumption that (x, y) is a right-symmetric point in Z. Therefore, it follows that kyk = 1. Similarly, we can show that kxk = 1. Our next claim is that x is a right-symmetric point in X. Suppose by contradiction that x is not a right-symmetric point in X. Then there exists u ∈ SX such that u ⊥B x but x 6⊥B u. It follows from Proposition 2.1 of [11] that either u ∈ x+ \ x− or u ∈ x− \ x+. Without loss of generality, let us assume that u ∈ x+ \ x−. Choose v ∈ SY such that v ∈ y+ \ y−. Then there exists λ0 < 0 such that kx + λ0uk < kxk = 1 and ky + λ0vk < kyk = 1. This shows that (x, y) 6⊥B (u, v). On the other hand, since u ⊥B x, it is easy to check that (u, v) ⊥B (x, y). Clearly, this contradicts our assumption that (x, y) is a right-symmetric point in Z. Therefore, x must be a right-symmetric point in X. Similarly, we can show that y is a right-symmetric point in Y. This completes the proof of the necessary part of the theorem and establishes (cid:3) it completely. We would like to apply Theorem 2.4 and Theorem 2.5 to identify the left- symmetric points and the right-symmetric points of some three-dimensional poly- hedral Banach spaces. However, first we need to consider the analogous two- dimensional polyhedral Banach spaces. It was observed by Heil [5] that if X is a two-dimensional real polyhedral Banach space such that SX is a regular polygon with 2n sides, then Birkhoff-James orthogonality is symmetric in X if and only if n is odd. In that case, every element of X is a left-symmetric point as well as a right- symmetric point in X. In the next two theorems, we characterize the left-symmetric points and the right-symmetric points of a two-dimensional real polyhedral Banach space X, when SX is a regular polygon with 2n sides, where n is even. In this context, let us mention the following notations which are relevant to our next two theorems. Let N = {0, 1, . . . , 4n − 1}. We equip N with a binary operation " ∔ ", defined by n1 ∔ n2 = n1 + n2 if n1 + n2 < 4n, and n1 ∔ n2 = n1 + n2 − 4n, if n1 + n2 ≥ 4n, for all n1, n2 ∈ N . For any j ∈ N , we will denote j ∔ n by p, j ∔ 2n by q, j ∔ 3n by p and j ∔ 4n by q. It is trivial to see that q = j but we use this notation only to make our theorems look more convenient. Let us recall that a polyhedron Q is said to be a face of the polyhedron P if either Q = P or if we can write Q = P ∩ δM , where M is a closed half-space in X containing P, and δM denotes the boundary of M . If dim Q = i, then Q is called an i-face of P . (n − 1)-faces of P are called facets of P and 1-faces of P are called edges of P. Let us now characterize the left-symmetric points and the right symmetric points of a two-dimensional polyhedral Banach space whose unit sphere is a regular polygon with 4n sides. Theorem 2.6. Let X be a two-dimensional real polyhedral Banach space such that SX is a regular polygon with 4n sides, where n ∈ N. Then x ∈ SX is a left-symmetric point in X if and only if x is the midpoint of some edge of SX. Proof. Up to an isomorphism, we may assume that the vertices of BX are vj = (cos 2πj 4n ), j ∈ N . Let Fj be the facet of SX containing the vertices vj and vj∔1. Let us first prove the necessary part of the theorem. Suppose, u ∈ SX is a 4n , sin 2πj 8 SAIN, ROY, BAGCHI AND BALESTRO left-symmetric point in X. Without loss of generality, we may assume that u ∈ Fj, for some j ∈ N . Let fj be the supporting functional corresponding to the facet Fj. Then fj(x, y) = {x cos π(2j+1) 4n + y sin π(2j+1) 4n } sec π 4n for all (x, y) ∈ X. , 2 2 , 2 vp+vp∔1 vp+vp∔1 vp+vp∔1 2 vp+vp∔1 } ⊆ u⊥ ∩ SX. Since It can be easily checked that kerfj ∩ SX = { vp+vp∔1 }. Therefore, { vp+vp∔1 is the middle point of the edge 2 Fp, is smooth. Now, using analogous techniques as above, we obtain 2 )⊥ ∩ SX = { vq+vq∔1 vp+vp∔1 }. Since we have as- ( sumed that u ∈ Fj is a left-symmetric point, we get that u must be of the form vj +vj∔1 . In other words, u must be the midpoint of the edge Fj. We now prove the sufficient part of the theorem. Let u be the midpoint of some edge Fj , j ∈ N . Then u⊥ ∩ SX = { vp+vp∔1 are the middle points of the edges Fp and Fp respectively. Once again, using above arguments, it follows that u ∈ ( vp+vp∔1 (cid:3) )⊥. This completes the proof of the theorem. } = { vq+vq∔1 )⊥ ∩ ( vp+vp∔1 }. Clearly, vp+vp∔1 vp+vp∔1 vq+vq∔1 vp+vp∔1 vj +vj∔1 2 2 , 2 , 2 2 2 , 2 2 2 2 , 2 2 Theorem 2.7. Let X be a two-dimensional real polyhedral Banach space such that SX is a regular polygon with 4n sides, where n ∈ N. Then x ∈ SX is a right- symmetric point in X if and only if x is an extreme point of BX. , vp+vp∔1 4n , sin 2πj Proof. Up to an isomorphism, we may assume that the vertices of BX are vj = (cos 2πj 4n ), j ∈ N . Let Fj be the facet of SX containing the vertices vj and vj∔1. Let us first prove the necessary part of the theorem. In order to prove so, we will show that any smooth point of SX is not a right-symmetric point in X. Let u ∈ SX be a smooth point. Without loss of generality, suppose u ∈ Fj for some j ∈ N . Then by the previous theorem, u⊥ ∩ SX = { vp+vp∔1 }, namely, the middle points of the edges Fp and Fp respectively. Clearly, Fp ∩ (u+ \ u⊥) and Fp ∩ (u+ \ u⊥) are nonempty and contain smooth points of BX. Let w and w′ be two smooth points of BX lying in Fp ∩ (u+ \ u⊥) and Fp ∩ (u+ \ u⊥) respectively. Then it is easy to see that either, u ∈ w−, or u ∈ w′−. In either occasion, w and w′ are contained in u+ \ u⊥. Hence, by Theorem 2.2, u is not right-symmetric. Therefore, if u ∈ SX is right-symmetric point in X then u must be an extreme point of BX. We now prove the sufficient part of the theorem. Suppose u ∈ SX is an extreme point of BX. Without loss of generality, we may assume that u = v0. Clearly, v0 ∈ F0 ∩ F4n−1. Let f0 and f4n−1 be the supporting functionals corre- sponding to the facets F0 and F4n−1 respectively. It is straightforward to check that ker f0 ∩ SX = { vn+vn∔1 }. Let x1 = vn+vn−1 } and ker f4n−1 ∩ SX = { vn+vn−1 and let x2 = v3n+v3n∔1 . Consider the sets , v3n+v3n∔1 , v3n+v3n−1 2 2 2 2 2 2 2 2 S0 = {αx1 + βx2 : kαx1 + βx2k = 1, α > 0, β > 0}, D0 = SX \ S0. Now, it is easy to see that S0 = (v+ 0 \ v− 0 ) ∩ SX = {z ∈ SX : f0(z) > 0} ∩ {z ∈ SX : f4n−1(z) > 0}, A STUDY OF SYMMETRIC POINTS IN BANACH SPACES 9 D0 = v− 0 ∩ SX = {z ∈ SX : f0(z) ≤ 0} ∪ {z ∈ SX : f4n−1(z) ≤ 0}. r \ v− 0 \ v− In other words, (v+ 0 ) ∩ SX is the shorter arc between x1 and x2 excluding x1 and x2 on SX, and v− 0 ∩ SX is the closed longer arc between x1 and x2 on SX. In similar spirit, for any extreme point vr of BX, we define Sr = (v+ r ) ∩ SX and Dr = v− r ∩ SX . It is easy to see that S0 ∩ EX = {v3n∔1, v3n∔2, . . . , v0, v1, . . . , vn−1} ({v0} for n = 1) and D0 ∩EX = {vn, vn∔1, . . . , v3n}. We shall prove that v0 /∈ w− for all w ∈ S0. We observe that for any smooth point κ ∈ BX lying in Fj, κ− is properly contained in both v− j∔1. Therefore, to prove v0 /∈ w− for every w ∈ S0, it is sufficient to prove v0 /∈ v− for every vj ∈ S0. It is evident that v0 ∈ v− r if and only j if Dr ∩ EX contains v0. Now, it is easily revealed that Dr ∩ EX = {vr∔s}3n s=n. We observe that (r ∔ s) = 0 if and only if r ∈ {n, n + 1, . . . , 3n} i.e., vr ∈ D0. Since (S0 ∩ EX) ∩ (D0 ∩ EX) = ∅, we get that v0 /∈ w− for all w ∈ S0. In other words, v0 ∈ w− implies w ∈ v− 0 . In similar fashion, it can be shown that v0 ∈ w+ implies w ∈ v+ 0 . Therefore, by Theorem 2.2, v0 is right-symmetric and this completes the proof of the theorem. j and v− (cid:3) Let M = {0, 1, . . . , 2n−1}. We equip M with a binary operation "∔", defined by n1 ∔n2 = n1 +n2, if n1 +n2 < 2n, and n1 ∔n2 = n1+n2−2n, if n1 +n2 ≥ 2n, for all n1, n2 ∈ M. Let Z be a three-dimensional real polyhedral Banach space such that its unit sphere SZ is a right prism having a regular 2n-gon as its base.Up to an iso- morphism, we may assume that the vertices of BX are v±j = (cos 2jπ 2n , ±1), j ∈ M. Clearly, the points (vj +v−j ) , j ∈ M, are co-planer. Let us consider the plane containing (vj +v−j ) , j ∈ M. We observe that this plane is uniquely deter- mined and we name it as the X-plane. Let us denote any rectangular facet of SZ containing the verices v±j and v±(j∔1) by Vj . The plane z = 1 intersects SZ in two 2n-gonal facets. Let us denote these facets by ±U . In particular, SZ contains two types of facets, 2n , sin 2jπ 2 2 (a) Facets (Vj)2n−1 j=0 which intersect the X-plane. (b) Facets ±U which are parallel to the X-plane. In our context the facets of type-(a) will be called vertical facets and the facets of type-(b) will be called horizontal facets. The edges of rectangular facets that intersect the X-plane will be called vertical edges and the edges of rectangular facets that are parallel to the X-plane will be called horizontal edges. Therefore, each vertical edge is contained in the supporting planes of vertical facets Vj and Vj∔1 for some j ∈ M and each horizontal edge is contained in the supporting planes of horizontal facet U (or −U ) and vertical facet Vj for some j ∈ M. Now, X = {(x, y, 0) ∈ Z}. Let Y = {(0, 0, z) ∈ Z}. Let us equip X and Y with such norms that SX is a regular 2n-gon and extreme points of BX are (cos 2jπ 2n ), j ∈ M Y. The points ±1 in SY are both and BY = [−1, 1]. It is easy to see that Z = X L∞ left-symmetric and right-symmetric in Y. When n is even, applying Theorem 2.4 and Theorem 2.6, we observe that a point in SZ is left-symmetric in Z if and only if it is the centroid of some vertical facet or it is the centroid of some horizontal facet in SZ. Similarly, applying Theorem 2.5 and Theorem 2.7, we observe that a point in SZ is a right-symmetric point in Z if and only if it is an extreme point of 2n , sin 2jπ 10 SAIN, ROY, BAGCHI AND BALESTRO BZ. Whenever n is odd, X becomes a Radon plane. Now, applying Theorem 2.4, we observe that a point in SZ is a left-symmetric point in Z if and only if it is the centroid of some horizontal facet in SZ or it lies in SX. Similarly, applying Theorem 2.5, we observe that a point in SZ is a right-symmetric point in Z if and only if it lies in the horizontal edges of SZ. We rewrite these simple facts in form of two separate theorems below to end the present article. Theorem 2.8. Suppose Z is a three-dimensional real polyhedral Banach space such that its unit sphere is a right prism having a regular 4n-gon as its base. Then z ∈ SZ is a left-symmetric point in Z if and only if it is either a centroid of a vertical facet in SZ or a centroid of a horizontal facet in SZ. Also, z ∈ SZ is a right-symmetric point in Z if and only if it is an extreme point of BZ. Theorem 2.9. Suppose Z is a three-dimensional real polyhedral Banach space such that its unit sphere is a right prism having a regular (4n + 2)-gon as its base. Then z ∈ SZ is a left-symmetric point in Z if and only if it is either a centroid of a horizontal facet in SZ or of the form (x, y, 0). Also, z ∈ SZ is a right-symmetric point in Z if and only if it lies on a horizontal edge of SZ. References [1] Alonso, J., Martini, H., Wu, S., On Birkhoff orthogonality and isosceles orthogonality in normed linear spaces, Aequationes Math., 83 (2012), No.1-2, 153-189. [2] Balestro, V., Martini, H., Teixeira, R., A new construction of Radon curves and related topics, Aequationes Math., 90 (2016), No.5, 1013-1024. [3] Birkhoff, G., Orthogonality in linear metric spaces, Duke Math. J., 1 (1935), 169-172. [4] Day, M.M., Some characterization of inner product spaces, Trans. Amer. Math. Soc., 62 (1947), 320-337. [5] Heil, E., Abschatzungen fur einige Affininvarianten konvexer Kurven, Monatsh. Math., 71 (1967), 405-423. [6] James, R.C., Inner product in normed linear spaces, Bull. Amer. Math. Soc., 53 (1947), 559-566. [7] James, R.C., Orthogonality and linear functionals in normed linear spaces, Trans. Amer. Math. Soc., 61 (1947), 265-292. [8] Martini, H., Swanepoel, K. J., Antinorms and Radon curves, Aequationes Math, 72 (2006), No.1-2, 110-138. [9] Martini, H., Swanepoel, K. J., Weiss, G., The geometry of Minkowski spaces-a survey. I., Expo. Math., 19 (2001), No.2, 97-142. [10] Radon, J., U ber eine besondere Art ebener konvexer Kurven, Leipziger Berichte, Math. Phys. Klasse, 68 (1916), 23-28. [11] Sain, D., Birkhoff-James orthogonality of linear operators on finite dimensional Banach spaces, J. Math. Anal. Appl., 447, Issue 2, (2017), 860-866. [12] Sain, D., On the norm attainment set of a bounded linear operator, J. Math. Anal. Appl., 457, Issue 1, (2018), 67-76. [13] Sain, D., Paul, K., Bhunia, P. and Bag, S., On the numerical index of polyhedral Banach spaces, arXiv:1809.04778v1 [math.FA]. [14] Sain, D., Paul, K. and Mal, A., On Approximate Birkhoff-James Orthogonality and Normal Cones in a Normed Space, J. Convex Anal., 26 (2019), No.1, 341-351. (Sain) Department of Mathematics, Indian Institute of Science, Bengaluru 560012, Karnataka, INDIA, E-mail address: [email protected] (Roy) Department of Mathematics, National Institute of Technology Durgapur, Durgapur 713209, West Bengal, INDIA E-mail address: [email protected] A STUDY OF SYMMETRIC POINTS IN BANACH SPACES 11 (Bagchi) Department of Mathematics, National Institute of Technology Durgapur, Durgapur 713209, West Bengal, INDIA E-mail address: [email protected] (Balestro) Instituto de Mathem ´Atica e Estat ´Istica, Universidade Federal Fluminense, 24210201 Niter ´Oi, BRAZIL E-mail address: [email protected]
1702.00602
3
1702
2019-12-07T08:11:27
Unions of cubes in $\mathbb{R}^{n}$, combinatorics in $\mathbb{Z}^{n}$ and the John-Nirenberg and John-Str\"omberg inequalities
[ "math.FA", "math.MG" ]
Suppose that the $d$-dimensional unit cube $Q$ is the union of three disjoint "simple" sets $E$, $F$ and $G$ and that the volumes of $E$ and $F$ are both greater than half the volume of $G$. Does this imply that, for some cube $W$ contained in $Q$. the volumes of $E\cap W$ and $F\cap W$ both exceed $s$ times the volume of $W$ for some absolute positive constant $s$? Here, by "simple" we mean a set which is a union of finitely many dyadic cubes. We prove that an affirmative answer to this question would have deep consequences for the important space $BMO$ of functions of bounded mean oscillation introduced by John and Nirenberg. The notion of a John-Str\"omberg pair is closely related to the above question, and the above mentioned result is obtained as a consequence of a general result about these pairs. We also present a number of additional results about these pairs. (The second and third versions present the same results as the first version. The bibliography has been updated. The presentation is more detailed and hopefully more reader-friendly. Some misprints and some small errors in a couple of the proofs have been corrected.)
math.FA
math
UNIONS OF CUBES IN Rn, COMBINATORICS IN Zn AND THE JOHN-NIRENBERG AND JOHN-STRÖMBERG INEQUALITIES MICHAEL CWIKEL Abstract. Suppose that the d-dimensional unit cube Q is the union of three disjoint "simple" sets E, F and G and that the volumes of E and F are both greater than half the volume of G. Does this imply that, for some cube W contained in Q. the volumes of E ∩ W and F ∩ W both exceed s times the volume of W for some absolute positive constant s? Here, by "simple" we mean a set which is a union of finitely many dyadic cubes. We prove that an affirmative answer to this question would have deep consequences for the important space BM O of functions of bounded mean oscillation introduced by John and Nirenberg. We recall and use the notion of a John-Strömberg pair which is closely related to the above question. The above mentioned result is obtained as a consequence of a general result about these pairs. We also present a number of additional results about these pairs. 9 1 0 2 c e D 7 ] . A F h t a m [ 3 v 2 0 6 0 0 . 2 0 7 1 : v i X r a 1. Introduction We are going to pose and discuss a rather simply formulated geometrical and almost combinatorical question. We shall refer to it as Question A(1/2). Although it may seem to have no connection with functional analysis, this question in fact has its origins in the study of a very important space of functions. The space BMO, or BMO(D), which consists of all functions of bounded mean oscillation on a suitable subset D of Rd, plays a surprisingly wide range of different important roles in several branches of analysis. (For some details about those roles see, e.g., [3, p. 132] or [1, p. 4].) One of our aims here is to warmly invite readers, including those with interests quite unrelated to spaces of functions or even to analysis in general, to study our Question A(1/2), and hopefully even resolve it. In particular (as our title seeks to indicate, and as confirmed by [4]) those readers with expertise in geometry in Rn or combinatorical problems in Zn may well have some valuable insights. Those who choose to respond to this invitation will be able to effectively do so, without needing any familiarity whatsoever with the space BMO, nor with the proofs in this paper, nor with the contents of the earlier papers which gave rise to it. Before formulating Question A(1/2) we shall mention another question which motivates it and then fix some necessary (but rather simple) terminology and notation. This paper is a sequel to [3] and its preliminary more detailed version [1]. The main motivation for those papers, and also for this one, is the wish for a better understanding of the celebrated John-Nirenberg inequality [5]. This inequality is satisfied by every function in the above mentioned space BMO(D). In particular, it is hoped to ultimately find an answer to: Key words and phrases. BMO, John-Strömberg pairs, John-Nirenberg inequality. This research was supported by funding from the Martin and Sima Jelin Chair in Mathematics, by the Technion V.P.R. Fund and by the Fund for Promotion of Research at the Technion. 1 2 MICHAEL CWIKEL Question J-N. Can the constants in the John-Nirenberg inequality for BMO functions of d variables be chosen to be independent of d? We refer to [6], [7], [8], [9], [10] and [11], and also to references in these papers. for some rather recent results concerning best constants in certain versions of the John-Nirenberg inequality. These deal mainly with the case where d = 1. The results of [10] apply to all values of d. Some remarks on pp. 7 -- 8 of [7] recall some reasons for being interested in the sizes of these constants, also for d > 1. We are grateful to Andrei Lerner, Pavel Shvartsman and Leonid Slavin for information about these and related results. As will be seen in a moment, our Question A(1/2) is quite easy to formulate, and is expressed in terms readily accessible to a general mathematical audience. Its importance lies in the fact that an affirmative answer to it would imply an affirmative answer to Question J-N. It would seem to be at least slightly easier to answer Question A(1/2) than to answer certain very similar questions which have an analogous role and which are posed in [1] and [3], and are also discussed in the brief survey article [2]. In the formulation of Question A(1/2), and indeed throughout this paper, we shall use the following notation and terminology (much of which is standard, and most of which was also used in [1] and [3]). Definition 1.1. We shall understand that (i) d is a positive integer, and that (ii) the word cube means a closed cube in Rd with sides parallel to the axes, i.e., the cartesian product of d bounded closed intervals, all of the same length. Special roles will be played by the d-dimensional unit cube [0, 1]d and its dyadic subcubes (i.e, the 1 ≤ nj ≤ 2k). It will sometimes be convenient to use the standard notation Q(x, r) for cubes of the formQd the cube of side length 2r centred at x ∈ Rd, i.e., Q(x, r) =ny ∈ Rd : ky − xkℓ∞ j=1(cid:2)(nj − 1)2−k, nj2−k(cid:3), where k ∈ N and the integers nj all satisfy d ≤ ro. Furthermore, (iii) we shall let Q(Rd) denote the collection of all cubes in Rd, and (iv) the word d-multi-cube will mean a non-empty subset of [0, 1]d which is the union of finitely many dyadic subcubes of [0, 1]d, and (v) we will denote the d-dimensional Lebesgue measure of any Lebesgue measurable subset E of Rd by λ(E). (But, as explained below, for Question A(1/2) you do not need to know anything at all about Lebesgue measure.) (The values of d in (ii) and in (v) will be clear from the contexts in which they appear.) In the formulation of Question A(1/2) we will not need any subtle properties of Lebesgue measure. In fact we will only be dealing with the simpler notion of d-dimensional volume. This is because the only Lebesgue measurable subsets E of Rd that we will encounter in Question A(1/2) are cubes or d-multi-cubes and their intersections with other cubes. All m=1 Em of non-overlapping sets Em, where each Em is the cartesian product of d closed intervals. So we simply have that each λ(Em) is the product of these are finite unions E =SM of the lengths of those d intervals, and that λ(E) =PM With some small adjustments, we can replace consideration of λ(E) for all the above mentioned sets E by consideration of the cardinality of E ∩ 2−kZd for a suitably large choice of k ∈ N. (This possibility was already very briefly hinted at in [2].) This is one reason for our suggestion above, in the opening paragraph of this paper, that Question A(1/2) might perhaps also be approachable via combinatorical considerations. m=1 λ(Em). After these preparations, here at last is the promised question: CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 3 Question A(1/2). Does there exist an absolute constant s > 0 which has the following property? For every positive integer d, whenever E+ and E− are two disjoint d-multi- cubes which satisfy (1.1) min {λ(E+), λ(E−)} > 1 2 (1 − λ(E+) − λ(E−)) , then there exists a cube W which is contained in [0, 1]d and for which (1.2) min{λ(W ∩ E+), λ(W ∩ E−)} ≥ sλ(W ) . Most of the work which is required to show that an affirmative answer to Question A(1/2) implies an affirmative answer to Question J-N has already been done in [1] and [3]. Because of that we will not have any need at all here to deal with any details concerning any of the versions of the function space BMO, nor to even recall their definitions. It was shown in [1] and [3] that an affirmative answer to Question J-N would be a consequence of an affirmative answer to the following question which was formulated at the beginnings of both of those papers, and is clearly quite similar to Question A(1/2). the following property? Question A. Do there exist two absolute constants τ ∈ (0, 1/2) and s > 0 which have For every positive integer d and for every cube Q in Rd, whenever E+ and E− are two disjoint measurable subsets of Q whose d-dimensional Lebesgue measures satisfy then there exists a cube W which is contained in Q and for which min{λ(E+), λ(E−)} > τ λ(Q \ E+ \ E−) , min{λ(W ∩ E+), λ(W ∩ E−)} ≥ sλ(W ) . It is obvious that an affirmative answer to Question A would imply an affirmative answer to Question A(1/2). Our task here is to show that the reverse implication also holds, namely that affirmatively answering the apparently at least slightly less demanding Question A(1/2) would suffice to also affirmatively answer Question A and therefore also Question J-N. In the next section we shall discuss some known results related to Question A(1/2) including some limitations which can be anticipated in any attempts to solve it. Then, after recalling some more notions and providing some preliminary results in Section 3. we will obtain the above mentioned reverse implication in Section 4, as an immediate consequence (see Corollary 4.2 below) of the main result of this paper, Theorem 4.1, which is formulated in a slightly more general context than Questions A and A(1/2). We note that Theorem 4.1 could also conceivably be used to deduce an affirmative answer for Question J-N from certain variants of Question A, which might perhaps be easier to answer than Question A(1/2). Some explorations of the possibility of such options, including some relevant numerical experiments, will perhaps be discussed in a future sequel to this paper. The reader may perhaps care to take note of some issues raised in Section 10 of [1] which might turn out to be relevant for resolving Question A(1/2) or its variants. The above mentioned more general context in which Theorem 4.1 is formulated revolves around the notion of John-Strömberg pairs, whose definition we will recall in Section 3. 4 MICHAEL CWIKEL In Section 5 we shall present a number of additional results about these pairs, beyond those which will be needed for our main result. Finally, the extremely brief Section 6 will make some minor comments about the papers [1] and [3]. 2. Some constraints and some known results related to Question A(1/2) As soon as we plunge into any attempt to answer Question A(1/2) we can almost immediately identify our real "enemies". We can of course always assume that λ(E−) ≤ λ(E+). If we restrict our attention to the case where r0λ(E+) ≤ λ(E−) ≤ λ(E+) for some fixed r0 ∈ (0, 1] then an obvious calculation shows that, by simply choosing W = [0, 1]d, we can obtain (1.2) for s = r0/ (3r0 + 1). This shows that our above mentioned "enemies" are the cases where the ratio λ(E−)/λ(E+) is arbitrarily small. It is known that if Question A(1/2) has an affirmative answer, then the positive number s for which that answer holds must satisfy s ≤ (2.1) √5 − 2. This follows from the third of three results which have been obtained by Ron Holzman [4] in connection with Question A(1/2). It is a pleasure to describe these results. The first of them is an affirmative answer to a non-trivial special case of Question A(1/2): Holzman has shown, for every d ∈ N, that whenever E+ and E− are d-multi-cubes which are each finite unions of dyadic cubes, all of side length 1/2, and they satisfy (1.1), then there exists a cube W in [0, 1]d which satisfies (1.2) for s = 1/4. In fact, as will be explained below in Remark 5.8, our Lemma 5.7 shows that this value of s cannot be improved in the following sense: If the general form of Question A(1/2) has an affirmative answer, then the positive number s which appears in that answer must satisfy s ≤ 1/4. So this already comes quite close to establishing (2.1). The second of Holzman's recent results is that this largest possible value 1/4 for s is indeed attained in another special case of Question A(1/2), where d is restricted to take only one value, namely d = 1, but where there is no restriction on the form of the disjoint measurable subsets E+ and E− of [0, 1]. (See also Remark 5.11.) Holzman's third result (see Remark 5.8 for details) uses a particular example when d = 2 to show that the analogue of his second result, when the single chosen value for d is greater than 1, does not hold. This example shows that for each d ≥ 2, as already stated above, the relevant value of s cannot be greater than √5 − 2. 3. Some further definitions and preliminary results The following definition (which is effectively the same as Definition 7.10 of [1, p. 29] and Definition 7.9 of [3, pp. 153 -- 154]) recalls a notion which plays a central role in [1] and [3]. and which is of course closely related to Question A. Definition 3.1. Let d be a positive integer and let E be a non-empty collection of Lebesgue measurable subsets of Rd. Suppose that each E ∈ E satisfies 0 < λ(E) < ∞. Let τ and s be positive numbers with the following property: (*) Let Q be an arbitrary set in E and let E+ and E− be arbitrary disjoint measurable subsets of Q. Suppose that (3.1) Then there exists a set W ⊂ Q which is also in E and for which (3.2) min {λ(E+), λ(E−)} > τ λ(Q \ E+ \ E−) . min{λ(E+ ∩ W ), λ(E− ∩ W )} ≥ sλ(W ) . CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 5 Then we will say that (τ, s) is a John-Strömberg pair for E. Remark 3.2. In the context of the preceding definition, if E+, E− and W are measurable subsets of Q which satisfy E+ ∩ E− = ∅ and λ(W ) > 0 and (3.2) for some s > 0, then λ(W ) ≥ λ(E+ ∩ W ) + λ(E− ∩ W ) ≥ 2 min{λ(E+ ∩ W ), λ(E− ∩ W )} ≥ 2sλ(W ) > 0 and therefore s ≤ 1/2. Consequently, for any choice of E, any pair (τ, s) which is a John-Strömberg pair for E must satisfy 0 < s ≤ 1/2. (Cf. Remark 7.16 of [1, p. 30].) It is known that (cid:0)√2 − 1, 2−d(3 − 2√2)(cid:1) is a John-Strömberg pair for Q(Rd). Several "natural" choices of the collection E are mentioned on pp. 4 -- 5 of [1] and pp. 132 -- 133 of [3]. These are relevant for studying a number of variants of the function space BMO which have been considered in the literature. The most "classical" of these choices, motivated by the context of [5] and by Question J-N, is when we take E to be the collection Q(Rd) of all cubes in Rd (as introduced in Definition 1.1(iii)) for some fixed d ∈ N. (See Example 7.12(i) on p. 30 of [1].) Indeed in this paper we will need to use Definition 3.1 only in the special case where E is the collection Q(Rd). But we have indicated the possibility of considering other choices of the collection E, since there may turn out to be variants of our main result in this paper, and these may turn out to be relevant for future research in the contexts of at least some of those other choices. (We refer to [1] and [3] for various results, in particular Theorem 9.1 of [1, p. 41] and [3, p. 164], which deal with John-Strömberg pairs and their interactions with versions of the John-Nirenberg inequality for variants of the space BMO corresponding to other choices of E.) Definition 3.3. It will sometimes be convenient, for each d ∈ N, to let JS(d) denote the set of all ordered pairs (τ, s) of positive real numbers which are John-Strömberg pairs for Q(Rd). Obviously, an equivalent reformulation of Question A is: Question A(τ, s). Do there exist two absolute constants τ ∈ (0, 1/2) and s > 0 for This makes it very relevant to use the following known result. (Note that below in which (τ, s) ∈ JS(d) for every d ∈ N? Subsection 5.1 we will present a result containing some new slight variants of it.) Theorem 3.4. Let d be a positive integer. The ordered pair of positive numbers (τ, s) is a John-Strömberg pair for Q(Rd) if and only if it has the following property: (∗) Whenever F+ and F− are disjoint d-multi-cubes which satisfy min{λ(F+), λ(F−)} > τ λ([0, 1]d \ F+ \ F−) , (3.3) then there exists a cube W contained in [0, 1]d for which min{λ(W ∩ F+), λ(W ∩ F−)} ≥ sλ(W ) . (3.4) Proof. Obviously every John-Strömberg pair (τ, s) for Q(Rd) must have the property (∗). The fact that property (∗) implies that (τ, s) is a John-Strömberg pair for Q(Rd) is exactly the content of Theorem 10.2 of [1, p. 43], and is proved on pp. 44 -- 47 of [1]. However, some other small issues remain to be clarified: It should be mentioned that there are a few minor misprints in the proof of Theorem 10.2 of [1]. But they do not effect its validity. On page 46, in the third and fourth lines above the inequality (10.16)) the notation Fε should of course be changed to F♭ (in two places) and Ωε should be changed to Ω♭. Then ∗ should be (2r∗)d . Similarly, four factors on line 11 of page 47 rd n should be (2rn)d and rd 6 MICHAEL CWIKEL of 2d have been (harmlessly) omitted on line 5 of page 63 of [1] in the proof of Lemma 10.1, which is an ingredient in the proof of Theorem 10.2. That line should be = (2Rn)d − (2r∗)d + (2Rn)d − (2rn)d . It should also be mentioned that, in the formulation of Theorem 10.2 of [1], it is stated that the number τ must satisfy 0 < τ < 1/2. But the proof is valid for all τ > 0. (The requirement that τ < 1/2 was imposed only because this is relevant in Theorem 9.1 of [1, p. 41] and [3, p. 164].) The clarification of these small issues completes the proof of Theorem 3.4. (cid:3) We can now present our main result, and its obvious corollary for dealing with Question A(1/2), 4. The main result 0 < s ≤ 1/2 Theorem 4.1. Let d be a positive integer. Suppose that (τ, s) is a John-Strömberg pair for Q(Rd). Then (4.1) and, for each θ ∈ (0, s/(1 − s)), the pair ((1 − θ)τ, s − θ(1 − s)) is also a John-Strömberg pair for Q(Rd). Corollary 4.2. Suppose that the answer to Question A(1/2) is affirmative, i.e., suppose that (1/2, s) is a John-Strömberg pair for Q(Rd) for some positive number s which does not 2−2s , s depend on d. Then, if we choose θ = s/(2−2s) in Theorem 4.1, we obtain that(cid:0) 1 2(cid:1) is a John-Strömberg pair for Q(cid:0)Rd(cid:1) for all d ∈ N, which gives us an affirmative answer to Question A(τ, s) and therefore also to Question A and Question J-N. 2 · 2−3s Proof of Theorem 4.1. Let us explicitly choose particular (necessarily positive) numbers τ , s and θ which satisfy the hypotheses of the theorem. The inequalities (4.1), as observed in Remark 3.2, follow from the fact that (τ, s) is a John-Strömberg pair for Q(Rd). Note that (4.1) and the conditions imposed on θ ensure that 0 < θ < 1 and also that 0 < θ(1 − s) < s. Consequently, the numbers (1 − θ) τ and s − θ(1 − s) are both strictly positive (as indeed they must be if they are to form a John-Strömberg pair). Throughout this proof we will let Q denote the d-dimensional unit cube, Q = [0, 1]d. Let F+ and F− be two arbitrary disjoint subsets of Q which are both d-multi-cubes and satisfy (4.2) In view of Theorem 3.4, it will suffice to show that there exists a subcube W of [0, 1]d such that min{λ(F+), λ(F−)} > (1 − θ)τ λ (Q \ F+ \ F−) (4.3) min{λ(W ∩ F+), λ(W ∩ F−)} ≥ (s − θ(1 − s)) λ(W ) . For some sufficiently large integer N, both of the sets F+ and F− are finite unions of dyadic cubes all of the same side length 2−N , as of course is the whole unit cube Q. Since the distance between F+ and F− must be positive, there must be some dyadic subcubes of Q of side length 2−N which are not contained in F+ ∪ F−, and thus their interiors are all contained in Q\ F+ \ F−. Therefore the set Q\ F+ \ F− is also, at least to within some set of measure zero, a finite union of dyadic cubes of side length 2−N . Let F+ denote the collection of all dyadic cubes of side length 2−N which are contained j=1[aj, aj + 2−N ] in the collection in F+. For each δ ∈ (0, 2−N −1) and for each cube W =Qn δ of F+ which is "slightly smaller" than F+. It is H + δ H(W, δ). := [W ∈F+ With the help of this set, we will now choose a particular suitable value for δ which must remain unchanged for the rest of this proof. Obviously λ(F+) − λ(H + δ ) for each δ ∈(cid:0)0, 2−N −1(cid:1) and limδց0 λ(F+ \ H + δ ) = 0. Therefore, since the inequality in (4.2) is strict, we can and will choose our fixed value of δ to be sufficiently small to ensure that δ(cid:1) . δ ), λ(F−)(cid:9) > (1 − θ)τ λ (Q \ F+ \ F−) + τ λ(cid:0)F+ \ H + (4.5) Let G denote the collection of all dyadic cubes of side length 2−N whose interiors are contained in Q \ F+ \ F−. Given an arbitrary positive integer k, we divide each cube in the collection G into 2dk dyadic subcubes of side length 2−N −k. Let Gk denote the collection of all dyadic cubes obtained in this way. In other words, Gk is simply the collection of all dyadic cubes of side length 2−N −k whose interiors are contained in Q \ F+ \ F−. For each W ∈ Gk we let U(W, k) be the (closed) cube concentric with W whose volume satisfies (4.6) min(cid:8)λ(H + δ ) = λ(F+ \ H + λ (U(W, k)) = θλ(W ). We now introduce two more sets which will play particularly useful roles for us. First we define Vk := [W ∈Gk U(W, k) CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 7 side length is 2−N − 2δ, i.e., H(W, δ) =Qn F+, let H(W, δ) denote the cube concentric with W contained in the interior of W whose j=1[aj + δ, aj + 2−N − δ]. We will later use the following obvious fact: If V is a cube which intersects with H(W, δ) and has side length less than δ, then (4.4) We now introduce a special subset H + defined by V ⊂ W . and then use Vk to define the disjoint union which contains and will be "controllably" larger than F−. In accordance with usual very standard notation, we will denote the interior and the H − k := F− ∪ Vk boundary of any given cube W by W ◦ and ∂W respectively. We of course have (4.7) W ◦! ∪ Z for some set Z ⊂SW ∈Gk ∂W . Since λ(Z) = 0, this implies that Q \ F+ \ F− = [W ∈Gk λ (Q \ F+ \ F−) = XW ∈Gk λ(W ) . (4.8) In the following calculation we shall use (4.7) in the second line, and then, in the third line, the fact that, for every cube W ∈ Gk, the set Z of zero measure introduced in (4.7) is disjoint from the cube W ◦ and therefore also from U(W, k). The final line of the calculation will use the facts that, for each pair of distinct cubes W and W ′ in Gk, we have U(W ′, k) ⊂ (W ′)◦ and also W ◦ ∩ (W ′)◦ = ∅ and therefore W ◦ ∩ U(W ′, k) = ∅. Thus we obtain that 8 MICHAEL CWIKEL Q \ F+ \ H − U(W, k)! U(W, k)!! U(W, k)! W ◦!! \ [W ∈Gk W ◦! \ [W ∈Gk (W ◦ \ U(W, k))! . k = Q \ F+ \ F− \ [W ∈Gk = Z ∪ [W ∈Gk = Z ∪ [W ∈Gk = Z ∪ [W ∈Gk λ (W ◦ \ U(W, k)) = (1 − θ) XW ∈Gk k ) = XW ∈Gk = (1 − θ)λ(Q \ F+ \ F−) . δ \ H − k which is not in F+ must be in Q \ F+ \ H − k ⊂(cid:0)Q \ F+ \ H − k(cid:1) ≤ λ(cid:0)Q \ F+ \ H − δ(cid:1) . k(cid:1) ∪(cid:0)F+ \ H + δ(cid:1) . k(cid:1) + λ(cid:0)F+ \ H + δ \ H − δ \ H − λ(W ) k and then invoke (4.5) followed by (4.9) and k . This Since the two sets in parentheses on the right side of (4.10) are disjoint, it follows that Since λ(Z) = 0, the preceding equalities, together with (4.6) and (4.8), imply that Any point in the set Q \ H + shows that (4.9) (4.10) λ(Q \ F+ \ H − Q \ H + λ(cid:0)Q \ H + (4.11) Now we can first use the fact that F− ⊂ H − then (4.11), to obtain that min(cid:8)λ(H + k )(cid:9) δ ), λ(H − ≥ min(cid:8)λ(H + δ ), λ(F−)(cid:9) δ(cid:1) > (1 − θ)τ λ (Q \ F+ \ F−) + τ λ(cid:0)F+ \ H + δ(cid:1) k ) + τ λ(cid:0)F+ \ H + = τ λ(Q \ F+ \ H − ≥ τ λ(Q \ H + δ \ H − k ) . (4.12) Since Vk ⊂ Q \ F+ \ F− and F+ ∩ F− = ∅, we have (4.13) Furthermore H + k are disjoint measurable subsets of Q. According to our hypotheses, (τ, s) is a John-Strömberg pair for Q(Rd). Therefore, for each k ∈ N, since the quantity in the first line of (4.12) is strictly larger than the quantity in its last line, we deduce that there must exist some subcube Wk of Q which satisfies H − k ⊂ F− ∪ (Q \ F+ \ F−) = Q \ F+ . δ ⊂ F+ and so H + δ and H − (4.14) We now claim that the side length of Wk, which we can conveniently write as (λ(Wk))1/d, must satisfy δ ), λ(Wk ∩ H − min(cid:8)λ(Wk ∩ H + k )(cid:9) ≥ sλ(Wk) . (4.15) To show this we first observe that, since the cube Wk intersects with H + δ , it must intersect with the cube H(W, δ) for at least one cube W in the collection F+. If (4.15) does not (λ(Wk))1/d ≥ δ . CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 9 hold, i.e., if Wk has side length less than δ, then (cf. the discussion immediately preceding (4.4)) Wk must be completely contained in that particular cube W and therefore also in F+. Consequently Wk cannot intersect with the set H − k . (Here we have used (4.13) once more.) This contradicts (4.14) and shows that (4.15) does hold. contains all dyadic cubes of side length 2−N −k which intersect with Wk, and the side j=1 [aj, bj], then we choosefWk to beQd Let fWk be a new cube containing Wk and concentric with Wk. More precisely, if j=1(cid:2)aj − 2−N −k, bj + 2−N −k(cid:3). Clearly fWk Wk =Qd length(cid:16)λ(fWk)(cid:17)1/d = (λ(Wk))1/d + 21−N −k . This, together with (4.15), gives us that offWk satisfies (cid:16)λ(fWk)(cid:17)1/d λ(cid:16)fWk(cid:17) λ (Wk) = 1/d  (λ (Wk))1/d + 21−N −k (λ (Wk))1/d = 1 + 21−N −k (λ (Wk))1/d ≤ 1 + 21−N −k δ . It follows that (4.16) λ(cid:16)fWk(cid:17) λ(cid:16)fWk(cid:17) − λ(Wk) = λ(Wk) λ (Wk) − 1 ≤ λ(Wk) (cid:18)1 + δ (cid:19)d 21−N −k − 1! . Let Hk be the collection of all cubes in Gk which intersect with Wk. Clearly (4.17) Our definitions of Vk and of Hk also immediately give us that W ◦ ⊂ (Q \ F+ \ F−) ∩fWk. U(W, k)! U(W, k)! [W ∈Hk Wk ∩ Vk = Wk ∩ [W ∈Gk = Wk ∩ [W ∈Hk ⊂ [W ∈Hk U(W, k) . This and then (4.6) will enable us to obtain the first line in the following calculation. Its second line will use the fact that the interiors of the cubes in Gk and therefore also in Hk, are pairwise disjoint. Its fourth line will use the fact that H + The justifications of all other steps should be evident. (We wonder, casually, whether it might somehow be possible to replace the simple-minded transition from the sixth to the Its sixth line will follow from the obvious inclusion Wk ⊂ fWk. Its third line will use (4.17). δ ⊂ F+. 10 MICHAEL CWIKEL seventh line by a sharper estimate, which might then lead to a (slightly) stronger version of Theorem 4.1.) λ(W ) W ◦! λ(U(W, k)) = θ XW ∈Hk λ(Wk ∩ Vk) ≤ XW ∈Hk λ(W ◦) = θλ [W ∈Hk = θ XW ∈Hk ≤ θλ(cid:16)(Q \ F+ \ F−) ∩fWk(cid:17) δ \ F−) ∩fWk(cid:17) ≤ θλ(cid:16)(Q \ H + δ \ F−) ∩ Wk(cid:1) + θλ(cid:16)fWk \ Wk(cid:17) ≤ θλ(cid:0)(Q \ H + δ \ F−) ∩ Wk(cid:1) + θ(cid:16)λ(cid:16)fWk(cid:17) − λ (Wk)(cid:17) = θλ(cid:0)(Q \ H + δ ) ∩ Wk(cid:1) + θ(cid:16)λ(cid:16)fWk(cid:17) − λ (Wk)(cid:17) ≤ θλ(cid:0)(Q \ H + δ(cid:1) + θ(cid:16)λ(cid:16)fWk(cid:17) − λ (Wk)(cid:17) . = θλ(cid:0)Wk \ H + δ(cid:1) + εkλ(Wk) λ(Wk ∩ Vk) ≤ θλ(cid:0)Wk \ H + (cid:17)d − 1(cid:19) and we will later use the obvious fact that lim k→∞ εk = 0 . (4.18) where εk := θ(cid:18)(cid:16)1 + 21−N −k δ (4.19) Combining the result of this calculation with (4.16), we deduce that, for each k ∈ N, In view of (4.14), we have that λ(Wk \ H + (4.20) We now have the ingredients needed to estimate λ (Wk ∩ F−) from below. Since λ(Wk ∩ H − k ) = λ(Wk ∩ F−) + λ(Wk ∩ Vk) we can use (4.14) and (4.18) and then (4.20) to obtain that δ ) = λ(Wk) − λ(Wk ∩ H + δ ) ≤ (1 − s)λ(Wk) . λ(Wk ∩ F−) = λ(Wk ∩ H − k ) − λ(Wk ∩ Vk) ≥ sλ(Wk) − θλ(cid:0)Wk \ H + ≥ sλ(Wk) − θ (1 − s) λ(Wk) − εkλ(Wk) = (s − θ(1 − s) − εk)λ(Wk) . δ(cid:1) − εkλ(Wk) (4.21) Furthermore, again with the help of (4.14), we also obviously have that (4.22) λ(Wk ∩ F+) ≥ λ(Wk ∩ H + δ ) ≥ sλ(Wk) . As already observed at the beginning of this proof, s − θ(1 − s) and θ(1 − s) are both strictly positive. So, in view of (4.19), for all sufficiently large k, we will also have s > (s − θ(1 − s) − εk) and (s − θ(1 − s) − εk) > 0. Therefore, we can now deduce from (4.2) and (4.21) and (4.22) that ((1 − θ)τ, s − θ(1 − s) − εk) is a John-Strömberg pair for Q(Rd) for all sufficiently large k. This seems to be very close to our required result. Indeed it will only require a little more effort to obtain that result. For each k ∈ N, let xk be the centre of the cube Wk and let rk be half its side length. I.e., we can set Wk = Q(xk, rk) in the standard notation recalled above in Definition 1.1(ii). CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 11 Of course xk ∈ Q and, by (4.15) and the fact that Wk ⊂ Q, we also have δ/2 ≤ rk ≤ 1/2. Therefore there exists a strictly increasing sequence {nk}k∈N of positive integers such that the sequences {xnk}k∈N and {rnk}k∈N converge, respectively, to a point x ∈ Q and a number r ∈ [δ/2, 1/2]. Let W be the cube W = Q(x, r). Then, by Lemma 10.1 of [1, p. 43] and (4.21) and (4.19), it follows that W ⊂ Q and λ(Wnk ∩ F−) (s − θ(1 − s) − εnk)λ(Wnk) λ(W ∩ F−) = lim ≥ lim = (s − θ(1 − s))λ(W ) . Similarly, by replacing k by nk in (4.22) and then passing to the limit as k tends to ∞, we obtain that k→∞ k→∞ λ(W ∩ F+) ≥ sλ(W ) ≥ (s − θ(1 − s))λ(W ) . All this shows that the subcube W of Q satisfies (4.3) and therefore completes the proof of the theorem. (cid:3) Remark 4.3. It seems quite possible that a more elaborate version of the preceding proof might show that the hypotheses of Theorem 4.1 can yield a stronger result, namely that (τ ′, s′) is a John-Strömberg pair for some positive number τ ′ smaller than (1 − θ) τ and/or for some number s′ greater than s − θ(1 − s). Perhaps a strategy for proving this might involve separately considering the two cases where λ(F+) > cλ(F−) and where λ(F−) ≤ λ(F+) ≤ cλ(F−) for some suitably chosen constant c > 1. 5. Some further results about John-Strömberg pairs 5.1. Other characterizations of John-Strömberg pairs. It will be convenient to begin by introducing some "technical" terminology. Definition 5.1. Let F+ and F− be two disjoint measurable subsets of [0, 1]d which both have positive measure and which also satisfy one of the following two conditions: (i) λ(cid:0)[0, 1]d \ F+ \ F−(cid:1) = 0. (ii) There exists a non-empty open subset Ω of [0, 1]d for which (5.1) max {λ(F+ ∩ Ω), λ(F− ∩ Ω)} < λ(Ω) and min{λ(F+ ∩ Ω), λ(F− ∩ Ω)} = 0. Then we shall say that (F+, F−) is a tame couple in [0, 1]d. Remark 5.2. Obviously the two conditions (i) and (ii) are mutually exclusive. Further- more, the following simple argument shows that the condition (ii) is equivalent to (ii)' There exists a dyadic cube V contained in [0, 1]d for which (5.2) max{λ(F+ ∩ V ), λ(F− ∩ V )} < λ(V ) and min{λ(F+ ∩ V ), λ(F− ∩ V )} = 0. The Lebesgue measures of each of the sets appearing in (5.2) remain unchanged if the dyadic cube V in them is replaced by its interior V ◦. Therefore condition (ii)' implies condition (ii). Conversely, suppose that some open set Ω satisfies (5.1).We can suppose that 0 = λ(F− ∩ Ω) ≤ λ(F+ ∩ Ω) < λ(Ω), (Otherwise simply reverse the roles of F+ and F− in what is to follow.) By Theorem 1.11 of [12, p. 8], Ω can be expressed as the union n=1 λ(F+ ∩ Vn) < n=1 λ(Vn) and consequently λ(F+ ∩ Vn) < λ(Vn) for at least one n which of course also satisfies λ(F− ∩ Vn) = 0. So, for that choice of n, the dyadic cube Vn satisfies (5.2). Therefore condition (ii) implies condition (ii)' and so (ii) and (ii)' are indeed equivalent. of a sequence of non-overlapping dyadic cubes Ω = Sn∈N Vn. So P∞ P∞ 12 MICHAEL CWIKEL We can now present the new variant of Theorem 3.4 to which we referred in the preamble to that theorem. It will specify two more properties of a pair (τ, s), which we shall label as (∗∗) and (∗ ∗ ∗), and which are each equivalent to the property that (τ, s) ∈ JS(d). Note that the only difference between the statement of property (∗∗) in Theorem 5.3, and the statement of property (∗) in Theorem 3.4 is that the strict equality ">" which appears in (3.3) in the statement of (∗) has been replaced by "≥" in (5.3) in the statement of (∗∗). Property (∗ ∗ ∗) of Theorem 5.3 is more elaborate, and requires the terminology of Definition 5.1. Theorem 5.3. Let d be a positive integer and let (τ, s) be an ordered pair of positive numbers. Then each of the following two properties is equivalent to the property that (τ, s) is a John-Strömberg pair for Q(Rd). (∗∗) Whenever F+ and F− are disjoint d-multi-cubes which satisfy min {λ(F+), λ(F−)} ≥ τ λ([0, 1]d \ F+ \ F−) , (5.3) then there exists a cube W contained in [0, 1]d for which min{λ(W ∩ F+), λ(W ∩ F−)} ≥ sλ(W ) . (5.4) (∗∗∗) Whenever (F+, F−) is a tame couple of measurable subsets of [0, 1]d which satisfies (5.3), then there exists a cube W contained in [0, 1]d for which (5.4) holds. Proof. Let us first show that property (∗∗∗) implies property (∗∗). Suppose that F+ and F− are arbitrary disjoint d-multi-cubes. By the same simple reasoning as was given in the paragraph immediately following (4.3) in the proof of Theorem 4.1, we see that the set [0, 1]d \ F+ \ F− must contain the interior V ◦ of at least one dyadic subcube V of [0, 1]d. The set Ω = V ◦ of course satisfies the condition (ii) in Definition 5.1. Consequently, (F+, F−) is a tame couple in [0.1]d. It immediately follows that property (∗ ∗ ∗) indeed does imply property (∗∗). Obviously, if the pair (τ, s) of positive numbers has property (∗∗), then it also has property (∗) of Theorem 3.4 and therefore (τ, s) is a John-Strömberg pair for Q(Rd). Now suppose that (τ, s) is a John-Strömberg pair for Q(Rd). Then, by letting {τn}n∈N and {sn}n∈N be the special constant sequences τn = τ and sn = s, we see that (τ, s) satisfies the hypotheses of Lemma 5.5 which we will formulate and prove below. In view of part (b) of that lemma, (τ, s) has property (∗ ∗ ∗) . This completes the proof of the theorem. (cid:3) Remark 5.4. It is very natural to wonder whether property (∗ ∗ ∗) is equivalent to a stronger and more simply expressed variant of that property in which the same implication is required to hold for all pairs of disjoint measurable subsets F+, F− of [0, 1]d which both have positive measure, thus omitting the requirement that F+ and F− should form a tame couple. (In other words, as in Remark 7.17 of [1, p. 30], we are essentially wondering whether in Definition 3.1, at least in the case where E = Q(Rd), it would be equivalent to replace ">" by "≥" in (3.1), of course then with the necessary proviso that λ(E+) and λ(E−) are both positive.) We are unable to answer this question, but Lemma 5.5 shows that if its answer is negative, then the sets F+ and F−which provide a counterexample, must both have very intricate structure (and it would not be inappropriate to refer to them as forming a "wild" couple). Lemma 5.5. Let d be a positive integer. Let {τn}n∈N and {sn}n∈N be two sequences of positive numbers which converge to positive limits, τ and s respectively. Suppose that (τn, sn) is a John-Strömberg pair for Q(Rd) for every n ∈ N. Then CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 13 (a) the limiting pair (τ, s) has the property (∗∗) of Theorem 5.3, and, (b) in the special case where sn ≥ s for every n ∈ N, the limiting pair (τ, s) also has the property (∗ ∗ ∗) of Theorem 5.3. Proof. In this proof it will be convenient to use a trivial generalization of the standard notation for cubes recalled in Definition 1.1(ii) and to introduce some additional obvious terminology as follows: Definition 5.6. We permit ourselves to extend the notation Q(x, r) for a cube of side length r centred at x, for each x ∈ Rd and r > 0, also to the case where r = 0, by letting Q(x, 0) denote the singleton {x} . Let {xk}k∈N be a convergent sequence in Rd and let {rk}k∈N be a convergent sequence of positive numbers. For each k ∈ N let Wk be the cube Q(xk, rk). Then we refer to {Wk}k∈N as a convergent sequence of cubes, and define its limit limk→∞ Wk to be the cube or singleton Q (limk→∞ xk, limk→∞ rk). Let us fix an arbitrary d ∈ N, and arbitrary sequences {τn}n∈N and {sn}n∈N of positive numbers which tend respectively to the positive numbers τ and s, and have the property that (τn, sn) ∈ JS(d) for each n ∈ N. In order to deduce that (τ, s) has property (∗∗) and/or property (∗ ∗ ∗), we begin by fixing two arbitrary disjoint measurable sets F+ and F− which are contained in of [0, 1]d, which form a tame couple in [0, 1]d, and which also satisfy (5.3). To obtain part (a) of the lemma we have the task of proving, for these choices of F+, F−, {τn}n∈N, {sn}n∈N, τ and s, that there exists a subcube W of [0, 1]d which satisfies (5.4) . If needed at any stage of our proof of that fact, we may make the additional assumption that F+ and F− are both d-multi-cubes. An analogous task is required to obtain part (b) of the lemma, with the difference that in our proof this time, instead of being able to assume that F+ and F− are d-multi-cubes, we may make the additional assumption, if needed, that sn ≥ s for all n. There is quite a lot of overlap in the ingredients which will be used for performing these two tasks, and it may help avoid some confusion if we give some general description, in advance, of each of the four steps which we shall use to accomplish both of them almost simultaneously. We stress that neither of the two above mentioned additional assumptions will be needed in the first two of these four steps. In Step 1 of the proof, we shall use the given measurable sets F+ and F− and the given sequences {τn}n∈N and {sn}n∈N to construct a special sequence {W (nk)}k∈N of subcubes of [0, 1]d which converges either to a cube or to a singleton set, in the sense of Definition 5.6. We shall perform this construction in two different (and unexplained and sometimes complicated) ways, depending on which of the two conditions of Definition 5.1 is satisfied by (F+, F−). It is only in later steps of the proof that we will be able to properly see the usefulness of the particular features of these constructions. converges to a cube W , then that cube is contained in [0, 1]d and satisfies (5.4). In Step 2, we shall see that if the sequence {W (nk)}k∈N obtained in the preceding step In Step 3, we shall see that whenever the disjoint measurable sets F+ and F− are both required to also be d-multi-cubes, then the sequence {W (nk)}k∈N can be always be constructed so that it converges to a subcube rather than a singleton. In view of Step 2, this will complete the proof of part (a) of the lemma. Finally, in Step 4, in order to complete the proof of part (b), it will remain (again in view of Step 2) only to deal with the case where the sequence {W (nk)}k∈N converges to a singleton. We will do this by showing that, in this case, when we also impose the 14 MICHAEL CWIKEL requirement that the sequence {sn}n∈N satisfies sn ≥ s for each n, then there exists a integer k0 for which the subcube W = W (nk0) is contained in [0, 1]d and satisfies (5.4). STEP 1: Construction of the special convergent sequence {W (nk)}k∈N. Since (F+, F−) is a tame couple, it must satisfy either condition (i) or condition (ii) of Definition 5.1. Our construction of {W (nk)}k∈N will be quite simple in the case where condition (i) holds, i.e., when (5.5) λ(cid:0)[0, 1]d \ F+ \ F−(cid:1) = 0. The strict inequalities 0 < λ(F+) < λ(cid:16)[0, 1]d(cid:17) enable us to use Lemma 7.1 of [1, p. 25] or of [3, p. 150] (whose simple proof is a special case of the proof of Lemma 7.5 on pages 26-27 of [1]) to ensure the existence of a cube W = Q(x, r) ⊂ [0, 1]d (of course with r > 0) for which (5.6) λ(W ∩ F+) = λ(W \ F+) = 1 2 λ(W ). The condition (5.5) is of course equivalent to λ(F+ ∪ F−) = 1 and to the fact that the sets F− and [0, 1]d \ F+ coincide to with sets of measure zero. Therefore we also have (5.7) In this case we will contruct our required special convergent sequence {W (nk)}k∈N by simply setting xk = x, rk = r, W (k) = W and nk = k for each k ∈ N. Then the sequence {W (nk)}k∈N of course converges, not to a singleton, but to the cube W which is its constant value. It remains to consider the more complicated case when F+ and F− satisfy condition (ii) λ (W ∩ F−) = λ (W \ F+) . or equivalently condition (ii)' of Remark 5.2. In that case, since the roles of F+ and F− are interchangeable in properties (∗∗) and (∗ ∗ ∗) and in Definition 5.1 and in (5.4), we can assume, without loss of generality, that, by Remark 5.2, there exists a dyadic cube V contained in [0, 1]d for which (5.8) 0 = λ(F− ∩ V ) ≤ λ(F+ ∩ V ) < λ(V ). Since the interior V ◦ of V satisfies λ(V ◦ \ F+) = λ(V \ F+) = λ(V ) − λ(V ∩ F+) > 0, λ(Q(z,r)∩(V ◦\F+)) the Lebesgue differentiation theorem guarantees the existence of a point z ∈ V ◦ \ F+ for = 1. Since Q(z, r) ⊂ V ◦ for all sufficiently small r, we also which limrց0 have limrց0 λ(Q(z,r)) = 0. These properties of V ◦ and z enable us to assert the existence of a sequence {ρk}k∈N of positive numbers tending to monotonically to zero, such that for each k we have λ(Q(z,r)) = 1 and, consequently, limrց0 λ(Q(z,r)) λ(Q(z,r)\F+)) λ(Q(z,r)∩F+) (5.9) and (5.10) and and therefore also λ (Q(z, ρk)) < λ(V ◦ \ F+) Q(z, ρk) ⊂ V ◦ λ (Q(z, ρk) ∩ F+) < λ (Q(z, ρk)) (5.11) for each k ∈ N. (5.12) (5.13) Let us now define a sequence {F+(k)}k∈N of measurable sets by setting F+(k) = F+ ∪ Q(z, ρk) for each k ∈ N. Note that, in view of (5.11), λ (F+(k)) = λ(F+) + λ (Q(z, ρk) \ F+) > λ(F+). CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 15 λ (Q(z, ρk) \ F+) > 0. We also introduce the set (5.14) Note that (5.15) since in fact G := F− \ V ◦. F+(k) ∩ G = ∅ for each k ∈ N F+(k) ∩ G = (F+ ∪ Q(z, ρk)) ∩ G = (F+ ∩ G) ∪ (Q(z, ρk) ∩ G) We also claim that ⊂ (F+ ∩ F−) ∪ (Q(z, ρk) \ V ◦) = ∅ ∪ ∅. λ(cid:16)[0, 1]d \ F+(k) \ G(cid:17) > 0 for all k ∈ N. (5.16) To show this we first observe that, obviously, V ◦ \ F+(k) = V ◦ \ F+(k)\ G. Consequently, λ(cid:16)[0, 1]d \ F+(k) \ G(cid:17) ≥ λ (V ◦ \ F+(k) \ G) = λ (V ◦ \ F+(k)) = λ (V ◦ \ F+ \ Q(z, ρk)) = λ (V ◦ \ F+) − λ ((V ◦ \ F+) ∩ Q(z, ρk)) ≥ λ (V ◦ \ F+) − λ (Q(z, ρk)) . By (5.9) this last expression is strictly positive for each k, which completes our proof of (5.16). We next observe that, by the first part of (5.8) (and of course also (5.14)), (5.17) λ(F−) = λ(F− ∩ V ◦) + λ(F− \ V ◦) = λ(F− \ V ◦) = λ(G). We use (5.16) then (5.15) then (5.13) and (5.17), and finally the disjointness of F+ and F− to obtain that Since τ > 0, this implies that 0 < λ(cid:0)[0, 1]d \ F+(k) \ G(cid:1) = 1 − λ(F+(k)) − λ(G) < 1 − λ (F+) − λ(F−) = λ(cid:16)[0, 1]d \ F+ \ F−(cid:17) . 0 < τ λ(cid:0)[0, 1]d \ F+(k) \ G(cid:1) < τ λ(cid:0)[0, 1]d \ F+ \ F−(cid:1) . (5.18) Then, since min{λ(F+(k)), λ(G)} ≥ min{λ(F+), λ(F−)}, it follows from 5.18 and the fact that F+ and F− satisfy (5.3) that min{λ(F+(k)), λ(G)} > τ λ([0, 1]d \ F+(k) \ G) > 0 for each k ∈ N. We use these strict inequalities to construct a sequence {j(k)}k∈N of positive integers such that j(k) ≥ k and j(k) is sufficiently large to ensure that τj(k) is sufficiently close to τ to imply that min{λ(F+(k)), λ(G))} > τj(k)λ([0, 1]d \ F+(k) \ G) for each k ∈ N. 16 MICHAEL CWIKEL Since(cid:0)τj(k), sj(k)(cid:1) ∈ JS(d) and since F+(k) and G are disjoint, we see, in accordance with Definition 3.1 (for E = Q(Rd)), that there exists a subcube W (k) of [0, 1]d which satisfies (5.19) The fact that j(k) ≥ k for each k ensures that (5.20) min{λ(W (k) ∩ F+(k)), λ(W (k) ∩ G)} ≥ sj(k)λ(W (k)) . lim k→∞ sj(k) = s. Since W (k)∩ F+ ⊂ W (k)∩ F+(k) = (W (k)∩ F+)∪ (W (k) ∩ Q(z, rk)) ⊂ (W (k)∩ F+)∪ Q(z, rk) and λ (Q(z, rk)) = (2rk)d, we obtain that (5.21) λ (W (k) ∩ F+) ≤ λ (W (k) ∩ F+(k)) ≤ λ(W (k) ∩ F+) + (2ρk)d for all k ∈ N. A slight variant of the simple reasoning in (5.17), again using the first part of (5.8) and (5.14), gives us that λ (W (k) ∩ F−) = λ (W (k) ∩ F− ∩ V ◦) + λ (W (k) ∩ F− \ V ◦) (5.22) = λ (W (k) ∩ F− \ V ◦) = λ (W (k) ∩ G) for all k ∈ N. For each k ∈ N we let the point xk in [0, 1]d and the number rk ∈ (0, 1/2] be the centre and half-side length respectively of the subcube W (k). I.e., we have W (k) = Q(xk, rk). There exists a strictly increasing sequence {nk}k∈N of positive integers such that the sequences {xnk}k∈N and {rnk}k∈N converge, respectively, to a point x ∈ [0, 1]d and a number r ∈ [0, 1/2]. We can now declare the sequence {W (nk)}k∈N = {Q (xnk, rnk)}k∈N to be the special convergent sequence which we set out to construct in this step of the proof in the case where (F+, F−) satisfies condition (ii) of Definition 5.1. Since we have already specified our construction of {W (nk)}k∈N when (F+, F−) satisfies condition (i) of Definition 5.1, this completes Step 1 of our proof. STEP 2: A proof that whenever the limit of {W (nk)}k∈N is a cube, then that cube has the two properties required to immediately complete the proof of the theorem. Suppose that the limit of the sequence {W (nk)}k∈N = {Q (xnk , rnk)}k∈N which was constructed in the previous step, is indeed a cube W = Q(x, r), i.e., that (5.23) r := lim k→∞ rnk > 0. Let us now prove that this implies that W is contained in [0, 1]d and that it satisfies (5.4). (We defer treatment of the case where limk→∞ rnk = 0 to Step 4.) In the case which was dealt with at the beginning of the previous step, where F+ and F− satisfy condition (i) of Definition 5.1, and where indeed we always have r > 0, it is already known that the cube W is contained in [0, 1]d . We note that, by the reasoning in Remark 3.2, we have sn ≤ 1/2 for each n ∈ N and therefore also s ≤ 1/2. So (5.4) follows immediately from (5.6) and (5.7). We turn to the remaining case, where F+ and F− satisfy condition (ii) of Definition 5.1, and therefore the sequence {W (nk)}k∈N is constructed in the more elaborate way described in the second and much longer part of the previous step. Here the positivity of the limit r permits us to apply Lemma 10.1 of [1, p. 43] (whose proof was briefly discussed above near the end of the first paragraph of the proof of Theorem 3.4) to the sequence CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 17 {W (nk)}k∈N = {Q(xnk , rnk)}k∈N to obtain the first required conclusion, that W ⊂ [0, 1]d, and to also obtain that (5.24) λ(W ∩ F+) = lim k→∞ λ (W (nk) ∩ F+) and (here also using (5.22)) that (5.25) λ(W ∩ F−) = lim k→∞ λ (W (nk) ∩ F−) = lim k→∞ λ (W (nk) ∩ G) = λ(W ∩ G), and also that (5.26) λ(W ) = lim k→∞ λ (W (nk)) . From (5.24) and (5.21) we see that (5.27) λ(W ∩ F+) = lim k→∞ λ (W (nk) ∩ F+(nk)) , and from (5.26) and (5.20) we see that (5.28) sλ(W ) = lim k→∞ sj(nk)λ (W (nk)) . In view of (5.19) we have that λ(W (nk) ∩ F+(nk)) ≥ sj(nk)λ(W (nk)) and λ(W (nk) ∩ G) ≥ sj(nk)λ(W (nk)) for each k ∈ N. If we take the limit as k tends to ∞ in each of these two inequalities, and apply (5.27), (5.25) and (5.28), then the two resulting inequalities can be rewritten as the required single inequality (5.4). Thus we have shown that if (5.23) holds, then in both cases, i.e., whether it is condition (i) or condition (ii) of Definition 5.1 which applies to F+ and F−, it follows that the limiting cube W indeed has both the properties required to complete the proof of part (a) and also part (b) of the lemma. It remains to explain, as we shall do in the remaining two steps of the proof, how we can sometimes guarantee that (5.23) does hold, and how we can proceed when it does not hold. STEP 3: Completion of the proof of part (a) of the lemma. Our preceding treatment of the case where (5.23) holds, will now enable us to complete the proof of part (a) of the lemma in full generality. Part (a) refers to the property (∗∗), so in our proof of it we can and must assume that the disjoint sets F+ and F− are both d-multi-cubes. We once again refer (as we did at the beginning of the proof of Theorem 5.3) to the simple reasoning after (4.3) in the proof of Theorem 4.1 which shows that there exists a dyadic cube V whose interior is contained in [0, 1]d \ F+ \ F− and which therefore must satisfy (5.8). So we can construct the sequence of cubes {W (nk)}k∈N for this particular choice of V in exactly the way that was done in the second part of Step 1 of this proof. This choice of V implies that the set G introduced in (5.14) must satisfy G = F− \ V ◦ = F−. Since they are d-multi-cubes, F+ and F− = G are compact and so of course is Q(z, ρk). Consequently, the set F+(k) (defined by (5.12)) is also compact. In view of (5.10), it is disjoint from G. So the distance, which we denote by dist∞ (F+(k), G), between F+(k) and G with respect to the ℓ∞ metric on Rd, must be positive. In fact, since F+(k) ⊂ F+(1) (because ρk ≤ ρ1), we have that (5.29) dist∞ (F+(k), G) ≥ dist∞ (F+(1), G) > 0 for all k ∈ N. Now we proceed more or less similarly to the last steps of the proof of Theorem 4.1. By (5.19) the cube W (k) = Q(xk, rk) intersects both of the sets F+(k) and G. So its side 18 MICHAEL CWIKEL length 2rk cannot be smaller than dist∞ (F+(k), G)). Consequently, also using (5.29), we see that 1 2 ≥ rk ≥ δ0 := 1 2 dist∞ (F+(1), G) > 0 for all k ∈ N. Consequently r ≥ δ0 and so (5.23) holds, permitting us to use the reasoning of Step 2 to ensure the existence of cube W which has the properties required to complete the proof of part (a). STEP 4: Completion of the proof of part (b) of the lemma. In view of Step 2, if the sequence of cubes {W (nk)}k∈N constructed in Step 1 converges to the cube W , then that cube is contained in [0, 1]d and satisfies (5.4) and no further reasoning is required to complete the proof of part (b). Thus it remains only to deal with the case where r = limk→∞ rnk = 0. It is clear from the first part of Step 1 that this cannot happen if (F+, F−) satisfies condition (i) of Definition 5.1. So the sequence {W (nk)}k∈N has necessarily been constructed as in the second part of Step 1, via the sets F (k) and other sets introduced there. Let us first see that, in this case, the point x = limk→∞ xnk cannot coincide with the point z which appears in the definition (5.12) of the sets F+(k). If x = z and is therefore in the interior V ◦ of the dyadic cube V used in the construction, then there exists some k which is sufficiently large to ensure that rnk + kxnk − xk∞ < dist∞ (x, ∂V ). This means that every point y in the cube W (nk) = Q(xnk , rnk) satisfies ky − xk∞ ≤ ky − xnkk∞ + kxnk − xk∞ < dist∞ (x, ∂V ). Consequently W (nk) ⊂ V ◦ and therefore W (nk) ∩ G = ∅. (Recall that G is defined by (5.14)). Since this contradicts (5.19), we have indeed shown that x 6= z. In view of this fact, there exists an integer k0 which is large enough to ensure that is a element of the sequence {ρk}k∈N with limit 0 which is used in the definition where ρnk0 (5.12) of the sets F+(k). If y is a point in the intersection of the two cubes W (nk0) = + ρnk0 + rnk0 < kx − zk∞ , (cid:13)(cid:13)(cid:13)x − xnk0(cid:13)(cid:13)(cid:13)∞ , rnk0(cid:17) and Q(cid:16)z, ρnk0(cid:17), then (5.30) Q(cid:16)xnk0 kx − zk∞ ≤ (cid:13)(cid:13)(cid:13)x − xnk0(cid:13)(cid:13)(cid:13)∞ ≤ (cid:13)(cid:13)(cid:13)x − xnk0(cid:13)(cid:13)(cid:13)∞ +(cid:13)(cid:13)(cid:13)xnk0 − y(cid:13)(cid:13)(cid:13)∞ + ρnk0 + rnk0 . + ky − zk∞ But this contradicts (5.30) and enables us to conclude that W (nk0) and Q(cid:16)z, ρnk0(cid:17) must be disjoint and therefore that We apply this, together with (5.22) for k = nk0 and then (5.19) for k = nk0, to obtain that the cube W := W (nk0) satisfies W (nk0) ∩ F+ (nk0) = W (nk0) ∩ F+. min{λ (W ∩ F+) , λ (W ∩ F−)} = min{λ (W ∩ F+ (nk0)) , λ (W ∩ G)} ≥ sj(nk0 )λ(W ). (5.31) The cube W , like all other cubes in the sequence {W (nk)}k∈N is contained in [0, 1]d. Finally we have to recall that in the statement of part (b) of the lemma, the sequence {sn}n∈N is required to satisfy sn ≥ s for each n. So (5.31) shows that, also in this last remaining case, we have obtained a subcube W of [0, 1]d which satisfies (5.4) for the given sets F+ and F−. This completes the proof of the lemma. (cid:3) CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 19 5.2. Some pairs which are not John-Strömberg pairs. The following result is a more elaborate variant of Remark 3.2. Lemma 5.7. For each d ∈ N and each τ > 0, and for every s > 1/(2 + 1/τ ), the pair (τ, s) is not in JS(d). Proof. If s > 1/2 then the result follows from Remark 3.2. So we can assume that 1/(2 + 1/τ ) < s ≤ 1/2. Let us choose some number a ∈ (1/(2 + 1/τ ), s) and then let (Obviously when d = 1 we have to interpret the previous definition to mean that E− = [0, a] and E+ = [1 − a, 1].) Since a < 1/2, these two sets are disjoint. Furthermore E− = (cid:8)(x, t) : x ∈ [0, 1]d−1, 0 ≤ t ≤ a(cid:9) and E+ = (cid:8)(x, t) : x ∈ [0, 1]d−1, 1 − a ≤ t ≤ 1(cid:9). λ(cid:16)[0, 1]d \ E+ \ E−(cid:17) = 1 − 2a = 1 − 2a a min{λ(E+), λ (E−)} or, equivalently, Since a > 1/(2 + 1/τ ) it follows that (5.32) 1 1 a − 2 min {λ(E+), λ (E−)} = λ(cid:16)[0, 1]d \ E+ \ E−(cid:17) . min{λ(E+), λ (E−)} > τ λ(cid:16)[0, 1]d \ E+ \ E−(cid:17) . a −2 > τ and so 1 1 We will complete the proof of this lemma by showing that, although the two disjoint measurable subsets E+ and E− of [0, 1]d satisfy (5.32), there does not exist any subcube W of [0, 1]d which satisfies min{λ(E+ ∩ W ), λ (E− ∩ W )} ≥ sλ(W ). This will follow from the inequality (5.33) min{λ(E+ ∩ W ), λ (E− ∩ W )} λ(W ) ≤ a which will be seen to hold for every subcube W of [0, 1]d. This inequality seems intuitively quite obvious, but let us nevertheless give a detailed (and perhaps not optimally elegant) proof. Let W be an arbitrary subcube of [0, 1]d and let θ be its side length. Of course θ ∈ (0, 1] and W must be of the form W = {(x, t) : x ∈ W0, t ∈ [β, β + θ]} where W0 is some subcube of [0, 1]d−1 of side length θ and [β, β + θ] ⊂ [0, 1]. (If d = 1 then W is simply the interval [β, β + θ].) We only need to consider the case where [β, β + θ] has a non-empty intersection with each one of the intervals [0, a] and [1 − a, 1], since otherwise at least one of the two sets E+∩ W and E−∩ W is empty and (5.33) is a triviality. The non-emptiness of the above mentioned two intersections implies that β ≤ a and β + θ ≥ 1 − a. When θ and β satisfy all the above mentioned conditions we have that λ(E− ∩ W ) = θd−1(a − β) and λ (E+ ∩ W ) = θd−1 (β + θ − (1 − a)). Therefore (5.34) min{λ(E+ ∩ W ), λ (E− ∩ W )} ≤ θd−1 sup [min{a − β, β + θ − (1 − a)}] . β∈R For each fixed choice of θ, the expression in the square brackets on the right side of (5.34) is the minimum of a strictly decreasing function of β and a strictly increasing function of β. Therefore its supremum and thus its maximum is attained when β takes the unique 20 MICHAEL CWIKEL value for which these two functions are equal, namely when β = (1 − θ) /2. This shows that (5.35) min{λ(E+ ∩ W ), λ (E− ∩ W )} λ(W ) ≤ θd−1 θd (cid:18)a − 1 − θ 2 (cid:19) = 1 2 − 1 θ(cid:18) 1 2 − a(cid:19) . Finally, the facts that θ ∈ (0, 1] and 1 bounded above by 2 − a > 0 imply that the right side of (5.35) is which establishes (5.33) and so completes the proof of the lemma. (cid:3) 1 2 −(cid:18)1 2 − a(cid:19) = a Remark 5.8. In particular, when τ = 1/2, Lemma 5.7 shows that (5.36) + ε(cid:19) /∈ JS(d) for every d ∈ N and every ε > 0. , 2 1 4 (cid:18)1 4(cid:1) ∈ JS(1), shows that s = 1 2, 1 Thus the second of the three results obtained recently by Ron Holzman (see Section 2), 4 is the largest possible value of s for which namely that(cid:0) 1 2, s(cid:1) ∈ JS(1). In the light of (5.36), the first of his three results can also be considered, in (cid:0) 1 some sense, to be best possible. At least in the non-trivial special case that he considered there he obtained that there exists a cube W in [0, 1]d which satisfies (1.2) for s = 1/4. (Apparently one cannot exclude the possibility that, at least in that special case, one might be able to also obtain a cube W in [0, 1]d which satisfies (1.2) for some s > 1/4.) These results tempt one to wonder whether perhaps the property(cid:0) 1 hold for all d ∈ N, If so, that would of course also answer Question A(1/2) affirmatively, and in an optimally strong, even "dramatically strong" way. However, in the third of his results relating to this question, Holtzman has analysed the following example suggested by the author and has shown that this is an "impossible dream". This property fails to hold already for d = 2. Therefore (cf. Theorem 5.14 below) it also does not hold for any other d > 1. 4(cid:1) ∈ JS(d) might 2, 1 Let F+ be the rectangle F+ =[0, 1]× [2/3, 1] and let F− be the union of the two squares [0, 1/3] × [0, 1/3] and [2/3, 1] × [0, 1/3]. Then (F+, F−) is a tame couple in [0, 1]2. The areas of the three disjoint sets F+, F− and [0, 1]2 \ F+ \ F− are respectively 1/3, 2/9 and 4/9, and this ensures that F+ and F− satisfy (5.3) for τ = 1/2 and d = 2. (In this example λ will of course always denote two-dimensional Lebesgue measure.) Let W denote the collection of all closed squares with sides parallel to the axes which are contained in [0.1]2. For each W ∈ W let f (W ) = min{λ(W ∩ F+), λ(W ∩ F−)} . λ(W ) The third of Ron Holzman's results is that, for these choices of F+ and F−, sup{f (W ) : W ∈ W} = max{f (W ) : W ∈ W} = √5 − 2. In view of part (∗ ∗ ∗) of Theorem 5.3, this shows that, indeed, (1/2, 1/4) /∈ JS(2) and, furthermore (again recalling Theorem 5.14), that (1/2, s) /∈ JS(d) for every s > √5 − 2 and d ≥ 2. It is tempting to wonder whether a sequence of appropriate variants of this example for subsets of [0, 1]d, where d tends to ∞, might lead to a negative answer to Question A(1/2). Initial attempts to find such a sequence have not yielded anything decisive. CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 21 5.3. Some further properties of the set JS(d). Fact 5.9. For each d ∈ N and each τ > 0 there exists some s > 0 such that (τ, s) ∈ JS(d). Proof. This is a consequence of Theorem 7.8 of [1, pp. 28 -- 29] a.k.a. Theorem 7.7 of [3, pp. 152 -- 153] which, for each τ > 0, provides a positive number s depending on τ and d such that (τ, s) ∈ JS(d). (The explicit formula for s will be recalled and used in Theorem 5.12 below. (cid:3) The preceding result ensures that the supremum in the following definition is taken over a non-empty set. Definition 5.10. For each d ∈ N and τ > 0 let σ(τ, d) := sup{s > 0 : (τ, s) ∈ JS(d)} . Remark 5.11. Thus the second and third results of Ron Holzman mentioned in Section 2 (cf. also Remark 5.8) can be written as 1 4 σ(cid:18) 1 2 and σ(cid:18) 1 , 1(cid:19) = We can now readily establish several properties of the set JS(d) and the function σ(τ, d). Theorem 5.12. For each fixed d ∈ N, (i) (τ, σ(τ, d)] ∈ JS(d) for each τ > 0. (ii) JS(d) = {(x, y) : x > 0, 0 < y ≤ σ(x, d)} (iii) The function x 7→ σ(x, d) is non-decreasing and continuous and satisfies , 2(cid:19) ≤ √5 − 2. 2 Proof. For part (i) let us fix an arbitrary τ > 0 and first note the obvious fact that (5.39) which implies that (τ, s′) ∈ JS(d) for every s′ ∈ (0, σ(τ, d)) . Therefore the sequences {τn}n∈N and {sn}n∈N which we define by τn := τ and sn := (1 − 2−n) σ(τ, d) must satisfy (τn, sn) ∈ JS(d) for every n ∈ N. For the proof of part (i) we now simply apply part (a) of Lemma 5.5 to the sequences {τn}n∈N and {sn}n∈N and then apply Theorem 5.3. Part (ii) then follows immediately from part (i) and (5.39). For part (iii) we first use another obvious fact, namely that (τ, s) ∈ JS(d) ⇒ (τ ′, s) ∈ JS(d) for all τ ′ > τ to immediately deduce that the function x 7→ σ(x, d) is non-decreasing. This latter property means that, in order to show that this function is continuous, it will suffice to show that (5.40) lim n→∞ σ ((1 + 1/n)τ, d) ≤ σ(τ, d) and lim n→∞ σ ((1 − 1/n)τ, d) ≥ σ(τ, d) for each τ > 0. We obtain the first of these inequalities by again using Theorem 5.3 together with part (a) of Lemma 5.5 to show that the limit of the sequence {((1 + 1/n)τ, σ ((1 + 1/n)τ, d))}n∈N of points in JS(d) must also be a point in JS(d). We next remark that, since (τ, σ(τ, d)) ∈ (5.37) where (5.38) ϕ(x, d) ≤ σ(x, d) ≤ 1 2 + 1 x for all x > 0, ϕ(x, d) = (cid:26) 2−d(x − x2)/(1 + x) 2−d(cid:0)3 − 2√2(cid:1) (τ, s) ∈ JS(d) ⇒ (τ, s′) ∈ JS(d) for all s′ ∈ (0, s), , 0 < x ≤ √2 − 1 x ≥ √2 − 1. , 22 MICHAEL CWIKEL JS(d), it follows from Theorem 4.1, that ((1 − 1/n)τ, σ(τ, d) − (1 − σ(τ, d)) /n) ∈ JS(d) for all sufficiently large n ∈ N. Therefore σ ((1 − 1/n)τ, d) ≥ σ(τ, d)− (1 − σ(τ, d)) /n for these same values of n. This suffices to prove the second inequality in (5.40) and complete the proof of continuity. The formula (5.38) for ϕ(x, d) in the estimate from below in (5.37) is obtained by once more appealing to Theorem 7.8 of [1, pp. 28 -- 29] a.k.a. Theorem 7.7 of [3, pp. 152 -- 153], and using the formulae appearing at the end of the statement of that theorem in the particular case where the collection of sets E is chosen to be Q(Rd) the collection of all cubes in Rd. We can replace M in those formulae by 2d using the fact that, in the terminology introduced just before the statement of that theorem, Q(Rd) is 2d-decomposable. We can also replace δ there by 1/2 using the fact (see Definition 7.4 of [1, p. 26] or [3, p. 151] and the remark immediately following it) that 1/2 is a bi-density constant for Q(Rd). The estimate from above in (5.37) follows from Lemma 5.7. This completes the proof of part (iii) and therefore of the whole theorem. (cid:3) Remark 5.13. There is another different kind of lower bound for σ(τ, d), in terms of values of σ(τ ′, d) for appropriate numbers τ ′ greater than τ , which can be obtained from Theorem 4.1. We have not bothered to explicitly state it here. The following result seems intuitively completely obvious. But we shall provide an explicit proof. Theorem 5.14. The inclusion JS(d + 1) ⊂ JS(d) and consequently also the inequality σ(τ, d + 1) ≤ σ(τ, d) both hold for every d ∈ N and τ > 0. Proof. This is the one place in this paper where we need to use the more explicit notation λd instead of λ to denote d-dimensional Lebesgue measure. We will use Theorem 3.4. Suppose that (τ, s) ∈ JS(d + 1). Let F+ and F− be two arbitrarily chosen disjoint d-multi-cubes which satisfy (3.3), i.e., the inequality min{λd(F+), λd(F−)} > τ λd([0, 1]d \ F+ \ F−) , (5.41) for this given value of τ . We define two subsets H+ and H− of [0, 1]d+1 as the cartesian products H+ = F+ × [0, 1] and H− = F− × [0, 1]. They are disjoint, since F+ and F−are disjoint. For each dyadic subcube E of [0, 1]d, the cartesian product E × [0, 1] is the union of 2k dyadic subcubes of [0, 1]d+1, where k is such that the side length of E is 2−k. It follows that H+ and H− are both (d + 1)-multi-cubes. subset of Rd+1 whenever E is a Lebesgue measurable subset of Rd, and that Now, and also again later, we shall use the very standard facts that (i) for each bounded closed interval [a, b], the set E × [a, b] is a Lebesgue measurable (ii) λd+1(E × [a, b]) = (b − a)λd(E) for each such E. These two facts together with the fact that [0, 1]d+1\H+\H− = ([0, 1]d\F+\F−)×[0, 1], and together with our assumption that F+ and F− satisfy (5.41), imply that min{λd+1(H+), λd+1(H−)} > τ λd+1([0, 1]d+1 \ H+ \ H−). Consequently, by Theorem 3.4, our assumption that (τ, s) ∈ JS(d + 1) guarantees the existence of a subcube V of [0, 1]d+1 for which (5.42) We can write V as the cartesian product V = W × [a, b], where W is a subcube of [0, 1]d and [a, b] is a subinterval of [0, 1] whose length b − a of course equals the side length of V and of W . Obviously V ∩ H+ = (W ∩ F+)× [a, b] and V ∩ H− = (W ∩ F−)× [a, b]. So we min{λd+1(V ∩ H+), λd+1(V ∩ H−)} ≥ sλd+1(V ). CONSTANTS IN THE JOHN-NIRENBERG INEQUALITY 23 can again apply the standard facts (i) and (ii), which were recalled in an earlier step of this proof, to obtain the formulae λd+1(E) = (b − a)λd(E) in the three cases where E is W ∩ F+ or W ∩ F− or W . When we substitute these formulae in (5.42) and divide both sides of the inequality by b − a, we obtain that the subcube W of [0, 1]d satisfies (3.4) of Theorem 3.4, i.e., that min{λd(W ∩ F+), λd(W ∩ F−)} ≥ sλd(W ) . Since F+ and F− were chosen arbitrarily, we can once more apply Theorem 3.4 to deduce that (τ, s) ∈ JS(d), and so complete the proof of the present theorem. (cid:3) 6. Some comments and minor corrections for the papers [1, 3]. In the first paragraph of the proof of Theorem 7.8 [1, p. 30] a.k.a. Theorem 7.7 [3, pp. 154 -- 155] it is shown that it suffices to consider the case where the two sets E+ and E− are both compact. The justification of this is a little clearer if in the formula on the third line of the proof one replaces G by Q \ H+ \ H−, which is obviously permissible. other small corrections and clarifications of some small issues in [1]. We refer to the remarks made above in the course of the proof of Theorem 3.4 for some References [1] M. Cwikel, Y. Sagher and P. Shvartsman, A new look at the John-Nirenberg and John-Strömberg theorems for BMO. Lecture Notes. arXiv:1011.0766v1 [math.FA] (Posted on 2 Nov 2010 ). [2] M. Cwikel, Y. Sagher and P. Shvartsman, A geometrical/combinatorical question with implications for the John-Nirenberg inequality for BMO functions, Banach Center Publ. 95 (2011), 45 -- 53 . [3] M. Cwikel, Y. Sagher and P. Shvartsman, A new look at the John-Nirenberg and John-Strömberg theorems for BMO, J. Funct. Anal. 263 (2012), 129 -- 166. [4] R. Holzman, Private Communication. [5] F. John and L. Nirenberg, On functions of bounded mean oscillation, Comm. Pure Appl. Math. 14 (1961), 415 -- 426. [6] A. K. Lerner, The John-Nirenberg inequality with sharp constants, C. R. Math. Acad. Sci. Paris 351 (2013), 463 -- 466. [7] L. Slavin, The John -- Nirenberg constant of BMOp, 1 ≤ p ≤ 2. arXiv:1506.04969v1 [8] L Slavin and V. Vasyunin, Inequalities for BMO on α-Trees. IMRN (On line as of 1 Oct 2015 ). [math.CA](Posted on 16 Jun 201 5). http://imrn.oxfordjournals.org/content/early/2015/09/30/imrn.rnv258.full.pdf?keytype=ref&ijkey=tvZXuAfzcBuzGVl [9] L. Slavin and V. Vasyunin, The John -- Nirenberg constant of BM Op, p > 2. arXiv:1601.03848 [math.CA] (Posted on 15 Jan 2016 ). [10] L. Slavin and P. Zatitskii, Dimension-free estimates for semigroup BM O and Ap arXiv:1908.02602 [math.CA] (Posted on 7 Aug 2019 and revised on 22 Aug 2019 ) [11] D. Stolyarov and P. Zatitskiy, Sharp transference principle for BM O and Ap, arXiv:1908.09497 [math.CA] (Posted on 26 Aug 2019 ) [12] R. Wheeden and A. Zygmund, Measure and integral. An introduction to real analysis. Pure and Applied Mathematics, Vol. 43. Marcel Dekker, Inc., New York-Basel, 1977. Department of Mathematics, Technion - Israel Institute of Technology, Haifa 32000, Israel E-mail address: [email protected]
1809.09465
2
1809
2019-05-08T13:33:04
Characterizations of woven frames
[ "math.FA" ]
In a separable Hilbert space $\mathcal H$, two frames $\{f_i\}_{i \in I}$ and $\{g_i\}_{i \in I}$ are said to be woven if there are constants $0<A \leq B$ so that for every $\sigma \subset I$, $\{f_i\}_{i \in \sigma} \cup \{g_i\}_{i \in \sigma ^c}$ forms a frame for $\mathcal H$ with the universal bounds $A, B$. This article provides methods of constructing woven frames. In particular, bounded linear operators are used to create woven frames from a given frame. Several examples are discussed to validate the results. Moreover, the notion of woven frame sequences is introduced and characterized.
math.FA
math
Characterizations of woven frames A. Bhandaria, S. Mukherjeea,1,∗ aDept. of Mathematics, NIT Meghalaya, Shillong 793003, India Abstract In a separable Hilbert space H, two frames {fi}i∈I and {gi}i∈I are said to be woven if there are constants 0 < A ≤ B so that for every σ ⊂ I, {fi}i∈σ ∪ {gi}i∈σc forms a frame for H with the universal bounds A, B. This article provides methods of constructing woven frames. In particular, bounded linear operators are used to create woven frames from a given frame. Several examples are discussed to validate the results. Moreover, the notion of woven frame sequences is introduced and characterized. Keywords: Frames, Woven Frames, Gap, Angle 2010 MSC: Primary 42C15; Secondary 46C07, 97H60 1. Introduction Hilbert space frame was first initiated by D. Gabor [1] in 1946 to recon- struct signals using fourier co-efficients. Later, in 1986, frame theory was reintroduced and popularized by Daubechies, Grossman and Meyer [2]. Frame theory literature became richer through several generalizations, namely, G-frame (generalized frames) [3], K-frame (frames for operators ∗Corresponding author Email addresses: [email protected] (A. Bhandari), [email protected] (S. Mukherjee) 1Supported by DST-SERB project MTR/2017/000797. Preprint submitted to Elsevier May 9, 2019 (atomic systems)) [4], fusion frame (frames of subspaces) ([5, 6]), K-fusion frame (atomic subspaces) [7], etc. and some spin-off applications by means of Gabor analysis in ([8, 9]), dynamical system in mathematical physics in [10], nature of shift invariant spaces on the Heisenberg group in [11], characterizations of discrete wavelet frames in CN in [12], extensions of dual wavelet frames in [13], constructions of disc wavelets in [14], orthogonality of frames on locally compact abelian groups in [15] and many more. Let us consider a scenario: suppose in a sensor network system, there are sensors A1, A2,··· , An which capture data to produce certain results. These sensors can be characterized by frames. In case one of these sensors, say Ak, fails to operate due to some technical reason, then the results obtained from these sensors may contain errors. Now assume that there are another set of sensors B1, B2,··· , Bn which does play similar role as Ai's. In addition, in the case of Ak fails, Bk can substitute so that obtained results are error free. Such an intertwinedness between two sets of sensors, or in general between two frames, leads to the idea of weaving frames. Weaving frames or woven frames were recently introduced by Bemrose et. al. in [16]. After that Deepshikha et. al. produced a generalized form of weaving frames in [17], they also studied the weaving properties of generalized continuous frames in [18], vector-valued (super) weaving frames in [19]. This article focuses on study, characterize and explore several properties of woven frames. The outline of this article is organized as follows. Section 2 is devoted to the basic definitions and results related to various kinds of frames, angle and gap between subspaces. Moreover, the characterizations of woven frames are analyzed in Section 3. Finally, woven frame sequences are established in Section 4. Throughout the paper, H is a separable Hilbert space. We denote by 2 Hn an n-dimensinal Hilbert space, {ei}i∈[n], ei = (δi,k)k≥1 is an orthonormal basis in Hn, L(Hn) to be a collection of all bounded, linear operators on Hn, R(T ) is denoted as the range of the operator T , by δ(M, N ) we denote the gap between two closed subspaces M and N of a Hilbert space H, c0(M, N ) is denoted as the cosine of the minimal angle between M and N , [n] = {1, 2,··· , n} and the index set I is either finite or countably infinite. Given J ⊂ I, a Bessel sequence {fi}i∈I and {ci} ∈ l2(I), we define TJf ({ci}) = Pi∈J cifi. It is to be noted that TJf is well-defined. 2. Preliminaries In this section we recall basic definitions and results needed in this paper. For more details we refer the books written by Casazza and Kutyniok [20] and Ole Christensen [21]. 2.1. Frame A collection {fi}i∈I in H is called a frame for H if there exist constants A, B > 0 such that Akfk2 ≤ Xi∈I hf, fii2 ≤ Bkfk2, (1) for all f ∈ H. The numbers A, B are called frame bounds. The operator S : H → H, defined by Sf = Pi∈Ihf, fiifi is called the frame operator for {fi}i∈I . positive, self-adjoint and invertible. It is well-known that the frame operator is linear, bounded, Definition 2.1. Let {fi}i∈I and {gi}i∈I be two frames for H. f ∈ H, f = Pi∈Ihf, giifi, then {gi}i∈I is called a dual frame of {fi}i∈I . If S is the frame operator of {fi}i∈I , then {S−1fi}i∈I is said to be the canonical dual frame of {fi}i∈I . If for all 3 Proposition 2.2. ([21, 22]) A finite family {fi}i∈[m] in Hn, forms a frame for Hn if and only if span{fi}i∈[m] = Hn. 2.2. Woven and full spark frame In a Hilbert space H, a family of frames {fij}i∈N,j∈[M ] is said to be weakly woven if for any partition {σj}j∈[M ] of N, {fij}i∈σj ,j∈[M ] forms a frame for H. Also, in H, two frames F = {fi}i∈I and G = {gi}i∈I are said to be woven if for every σ ⊆ I, {fi}i∈σ ∪{gi}i∈σc also forms a frame for H and the associated frame operator for every weaving is defined as, SFGf = Xi∈σ hf, fiifi + Xi∈σchf, giigi, for all f ∈ H. Theorem 2.3. [16] In H, two frames are weakly woven if and only if they are woven. Moreover, a frame with m elements in Hn, is said to be a full spark frame if every subset of the frame, with cardinality n, is also a frame for Hn. For example, if {ei}i∈[2], ei = (δi,k)k≥1, is an orthonormal basis in R2, then {e1, e2, e1 + e2} is a full spark frame for R2. Furthermore, if every element of a finite frame can be represented as a linear combination of the remaining others, then the frame is called a weak full spark frame. For example, if {ei}i∈[3] is an orthonormal basis of R3, {e1, e1, e2, e2, e3, e3} is a weak full spark frame but not a full spark frame. In this context, it is a fortuitous evident that every nontrivial (other than exact) full spark frame is also a weak full spark frame. Proposition 2.4. [16] Let {fij}i∈I be a collection of Bessel sequences in H with bounds Bj's for every j ∈ [m], then every weaving forms a Bessel 4 sequence with bound Pj∈[m] is bounded by r Pj∈[m] Bj. Bj and norm of corresponding synthesis operator Proposition 2.5. Every frame sequence is a Bessel sequence. Proof. Let {fi}i∈I be a frame sequence in H, then it forms a frame for F = span{fi}i∈I . Therefore H = F ⊕ F ⊥ and hence for every f ∈ H we have f = fF + fF ⊥. Therefore for some B > 0 we obtain, Xi∈I hf, fii2 = Xi∈I hfF , fii2 ≤ BkfFk2 ≤ Bkfk2. Remark 2.6. The converse implication of the above proposition is not nec- essarily true, which is evident from the following fact: Consider {ei}i∈N as an orthonormal basis for H and let us define fi = ei + ei+1, i ∈ N. Then {fi}i∈N forms a Bessel sequence in H but not a frame sequence. For detail discussion regarding the same we refer the Example 5.1.10 in [21]. Lemma 2.7. Let {fi}i∈I and {gi}i∈I be frames for H with bounds A1, B1 and A2, B2, respectively. Then the following results are equivalent: 1. {fi}i∈I and {gi}i∈I are woven. 2. For every σ ⊂ I, if Sσ is the associated frame operator for the corre- sponding weaving, then for every f ∈ H we have kSσfk ≥ kkfk for some k > 0, independent of σ. Proof. (1 =⇒ 2) 5 Let {fi}i∈I and {gi}i∈I be woven with universal frame bounds A, B. Therefore for every σ ⊂ I and for every f ∈ H we have, Akfk2 ≤ Xi∈σ hf, fii2 + Xi∈σc hf, gii2 ≤ Bkfk2. If Sσ is the associated frame operator for the corresponding weaving, then from the above inequality we have, Akfk2 ≤ hSσf, fi ≤ Bkfk2. Therefore, kSσfk = supkgk=1hSσf, gi ≥ (cid:28)Sσf, f kfk(cid:29) ≥ Akfk. (2 =⇒ 1) For every f ∈ H and σ ⊂ I, Sσf = TσT ∗σ f , where Tσ, T ∗σ are the corresponding synthesis and analysis operators, respectively; and we have k2kfk2 ≤ kSσfk2 = kTσT ∗σ fk ≤ kTσk2kT ∗σ fk2 and hence we obtain, hf, fii2 + Xi∈σc hf, gii2. B1 + B2kfk2 ≤ kT ∗σ fk2 = Xi∈σ k2 The universal upper frame bound for the weaving will be achieved by Propo- sition 2.4. Theorem 2.8. [16] In Hn, two frames {fi}i∈[m] and {gi}i∈[m] are woven if and only if for every σ ⊆ [m], span({fi}i∈σ ∪ {gi}i∈σc ) = Hn. 2.3. Gap and angle between subspaces Let M and N be two closed subspaces of a Hilbert space H. Then the gap between M and N is given by, δ(M, N ) = max{δ(M, N ), δ(N, M )}, where dist(x, N ), SM is the unit sphere in M and dist(x, N ) is δ(M, N ) = sup x∈SM the distance from x to N . 6 Again the cosine of the angle between two closed subspaces M and N of a Hilbert space H is given by, c(M, N ) = sup{hx, yi : x ∈ M∩(M∩N )⊥,kxk ≤ 1, y ∈ N∩(M∩N )⊥,kyk ≤ 1} and the cosine of the minimal angle of the same is given by, c0(M, N ) = sup{hx, yi : x ∈ M,kxk ≤ 1, y ∈ N,kyk ≤ 1}. For the extensive discussion regarding the gap and the angle between two subspaces, we refer ([23, 24, 25]). Remark 2.9. [25] Let M and N be two closed subspaces of a Hilbert space H. Then the followings are satisfied: 1. δ(M, N ) = 0 if and only if M ⊂ N . 2. δ(M, N ) = 0 if and only if M = N . Lemma 2.10. [24] Let M and N be two closed subspaces of a Hilbert space H. Then c0(M, N ) = 0 if and only if M ⊥ N . Theorem 2.11. [24] Let M and N be two closed subspaces of a Hilbert space H. Then the following statements are equivalent: 1. c0(M, N ) < 1 . 2. M ∩ N = {0} and M + N is closed. Theorem 2.12. (Douglas' factorization theorem [26]) Let H1,H2, and H be Hilbert spaces and S ∈ L(H1,H), T ∈ L(H2,H). Then the following results are equivalent: 1. R(S) ⊆ R(T ). 2. SS∗ ≤ αT T ∗ for some α > 0. 3. S = T L for some L ∈ L(H1,H2). 7 3. Characterization of Woven Frames In this section, we characterize woven frames, mainly through construc- tions of frames from given frames. The proposed constructions are based on the images of a given frame by means of bounded linear operators. Be- fore diving into the main results, we start the discussion with the following Proposition. Proposition 3.1. Let {fi}i∈[m] be a frame for Hn. Suppose fm+1 = 0, then {(fi − fi+1)}i∈[m] is also a frame for Hn and these two frames are woven. Proof. The proof will be followed from elementary row operations. Remark 3.2. In the above Proposition instead of fm+1 = 0, if fm+1 = f1, then {(fi − fi+1)}i∈[m] may not be a frame for Hn. For example, let {ei}i∈[3] be an orthonormal basis in R3, then {−e1 + e2, e1 + e2,−2e1 + e2 − e3, e1 + e2 + e3} is a frame for R3. But clearly, {(fi − fi+1)}i∈[4] = {−2e1, 3e1 + e3,−3e1 − 2e3, 2e1 + e3} is not a frame for R3. But if this is a frame, then they must be woven, which is evident from the fact that f = Pi∈[m] bi (fi − fi+1) can be written as f = (b1 − bj)f1 + (b2 − b1)f2 + ...(bj − bj−1)fj + (bj+1 − bj)(fj+1 − fj+2) + ...(bm−1 − bj)(fm−1 − fm) + (bm − bj)(fm − fm+1). Remark 3.3. If {fi}i∈[m] is a frame for Hn and suppose fm+1 = 0, then {(αfi + βfi+1)}i∈[m], α, β 6= 0, is also a frame for Hn and they are woven. In the following result, we present conditions under which image of a given frame under an idempotent operator is woven with the said frame. 8 Lemma 3.4. Let F (6= 0) ∈ L(H) be a closed range, idempotent operator with R(F ) = R(F ∗). Suppose {fi}i∈I is a frame for R(F ∗), then {F fi}i∈I is also a frame for R(F ∗) and they are woven. Proof. Let {fi}i∈I be a frame for R(F ∗) with bounds A, B. Then for every f ∈ R(F ∗) we have, Xi∈I hf, F fii2 = Xi∈I hF ∗f, fii2 ≤ BkF ∗fk2 ≤ BkFk2kfk2. Again since f ∈ R(F ∗) = R(F ), kfk2 = k(F ∗)†F ∗fk2 ≤ k(F ∗)†k2kF ∗fk2 and hence kfk2 k(F ∗)†k2 ≤ kF ∗fk2. Therefore for every f ∈ R(F ∗) we obtain, hf, F fii2 = Xi∈I hF ∗f, fii2 ≥ AkF ∗fk2 ≥ A k(F ∗)†k2kfk2. Xi∈I Consequently, {F fi}i∈I forms a frame for R(F ∗). Moreover, for every f ∈ R(F ∗), there exists g ∈ H such that F ∗g = f and since F 2 = F , for every σ ⊂ I and for all f ∈ R(F ∗) we obtain, Xi∈σ hf, fii2 + Xi∈σc hf, F fii2 = Xi∈σ = Xi∈σ = Xi∈I = Xi∈I hF ∗g, fii2 + Xi∈σc hF ∗g, F fii2 hg, F fii2 + Xi∈σc hg, F fii2 hg, F fii2 hf, fii2. Therefore, our assertion is tenable. Remark 3.5. It is to be noted that, if one of the conditions of F 2 = F and R(F ) = R(F ∗) fails, then the conclusion of the above Lemma may not hold. This is evident from the following two examples. 9 Example 3.6. Consider an idempotent operator F on R2 so that F e1 = e1 + 2e2, F e2 = 0, then R(F ) = span {e1 + 2e2} 6= span {e1} = R(F ∗). Now for the frame F = {e1, e2, e1 +e2} for R2, F (F) = {e1 +2e2, 0, e1 +2e2} is not a frame for R(F ∗). Example 3.7. Consider an operator F on R3 so that F e1 = e1 + e2, F e2 = −e1 + e2, F e3 = 0, then F 2 6= F but R(F ) = R(F ∗). Now let us choose a frame {fi}i∈[3] = {e1, e1 − e2, 2e1} for R(F ∗), then {F fi}i∈[3] = {e1 + e2, 2e1, 2e1 + 2e2} is also a frame for R(F ∗), but they are not woven, which can be verified for σ = {1, 3}. In the following outcomes we study woven-ness of frames and their im- ages under invertible operators. Remark 3.8. The image of a frame under invertible operators is not nec- essarily woven with the frame, which is evident from the following example: F = {e2, e1 + e2, 2e2} is a frame for R2. Let us consider an invertible op- erator T so that T e1 = e1 − e2, T e2 = −e1 − e2. Then TF = {−(e1 + e2),−2e2,−2(e1 + e2)} is also a frame for R2, however they are not woven, which can be verified by considering σ = {1, 3}. Proposition 3.9. The image of woven frames under invertible operator preserves their woven-ness. Proof. Let {fi}i∈I and {gi}i∈I be woven in H with universal bounds A, B and suppose T ∈ L(H) is an invertible operator. Then {T fi}i∈I and {T gi}i∈I are also frames for H. 10 For every σ ⊂ I and for every f ∈ H we obtain, Xi∈σ hf, T fii2 + Xi∈σc hf, T gii2 = Xi∈σ hT ∗f, fii2 + Xi∈σc hT ∗f, gii2 ≥ AkT ∗fk2 ≥ kT −1k2kfk2. A The upper frame bound of the respective weaving will be achieved from the Proposition 2.4. In the following theorem we discuss a necessary and sufficient condition of woven frames. Theorem 3.10. Let F = {fi}i∈I and G = {gi}i∈I be two frames for H. Then they are woven if and only if R(TFG) = H, where TFG is the associated synthesis operator of the respective weaving. Proof. Since F and G are frames for H, by Proposition 2.4 every weaving is a Bessel sequence and hence TFG is well-defined. Let R(TFG ) = H = R(IH). Therefore using Theorem 2.12, there exists fk2 ≥ Akfk2 and hence for every A > 0 and for every f ∈ H we have kT ∗ FG σ ⊂ I we obtain, Xi∈σ hf, fii2 + Xi∈σc hf, gii2 ≥ Akfk2. The converse implication will be followed from the definition of woven frame. Example 3.11. F = {e1 + e2, e1 + 2e2, e1 − e2} is a frame for R2. If S is its frame operator, then SF = {5e1 + 8e2, 7e1 + 14e2, e1 − 4e2}. It is to be noted that the associated synthesis operators of every weaving are onto. 11 Since invertible operators preserve linear independency of vectors, so it is a natural intuition that a finite frame is woven with its image under the associated frame operator. Problem 1: If {fi}i∈I is a frame for H with the associated frame operator S, Can {fi}i∈I and {Sfi}i∈I woven? At this moment, we are impotent to deliver an affirmative response, although we strongly believe that the same can be executed in this context. If so, then using Proposition 3.9, it is evident that {fi}i∈I and {S−1fi}i∈I are woven. Problem 2: Whether a frame is woven with its dual? 4. Woven Frame Sequence A family {fi}i∈I in H is said to be a frame sequence if it forms a frame for its closed, linear span. It is to be noted that {fi}i∈I is not necessarily In this section we explore the possibilities of two frame a frame for H. sequences together, through the concept of woven frames, form a frame for H. Definition 4.1. Two frame sequences F = {fi}i∈I and G = {gi}i∈I in H, are said to be woven frame sequences, if for every σ ⊂ I, {fi}i∈σ ∪ {gi}i∈σc forms a frame for H. Example 4.2. For example, In R3, the frame sequences {e1 + 2e3, e1 − e3,−e1 + 2e3, e1 + 3e3} and {e1 − e2, e1 + 2e2,−e1 + 3e2, e1 − 2e2} are woven whereas {e1 + 2e3, e1 − e3,−e1 + 2e3, e1 + 3e3} and {e1 − e2, e1 + 2e2,−e1 + 3e2, e1} are not. The notion of woven frame sequences is beneficial for its practical im- portance, because instead of two given frames, if we consider two frame 12 sequences, then less restriction is there in our primary assumption and due to this fact, it is cost-effective. Theorem 4.3. Let F = {fi}i∈[m] and G = {gi}i∈[m] be two frame sequences in Hn. Then the following statements are satisfied : 1. F and G are not woven if there exists a non-trivial σ ⊂ [m] so that c0{span(Fσ ∪ Gσc ), span(Fσc ∪ Gσ)} < 1 . 2. If for every non-trivial σ ⊂ [m], δ{span(Fσ∪Gσc ), span(Fσc∪Gσ)} = 0 and c0{(span(Fσ ∪ Gσc )), (span(Fσc ∪ Gσ))c} = 0 = c0{(span(Fσc ∪ Gσ)), (span(Fσ ∪ Gσc))c, then F and G are woven. Proof. Using Lemma 2.10 and Theorem 2.11 , our assertions are quickly plausible. Theorem 4.4. In Hn, if F = {fi}i∈[m] and G = {gi}i∈[m] are two woven frame sequences, then for every non-trivial σ ⊂ [m], δ{span(Fσ ∪ Gσc ), span(Fσc ∪ Gσ)} = 0 . Proof. If F and G are woven, then for every non-trivial σ ⊂ [m], both Fσ ∪ Gσc and Fσc ∪ Gσ constitute frames for Hn. Hence the conclusion directly follows from the Remark 2.9 . Remark 4.5. It is to be noted that, the two foregoing outcomes also hold for characterizing woven frames. In the following results we explore sufficient conditions for woven-ness be- tween frame and frame sequence. The following theorem shows that woven- ness is preserved under perturbation. 13 Theorem 4.6. Let F = {fi}i∈I , G = {gi}i∈I be two woven frames for H with the universal frame bounds A, B and TG be the corresponding synthesis operator of G. If H = {hi}i∈I is a frame sequence in H with the associated synthesis operator TH so that (kTGk +kTHk)kTG − THk < A, then F and H are woven. Proof. For every σ ⊂ I, let Pσ be the orthogonal projection on span{ei}i∈σ and therefore Tσf = TF Pσ. Since F and G are woven with the universal bounds A, B, then using Lemma 2.7, for every σ ⊂ I and every f ∈ H we have, Akfk ≤ kXi∈σ hf, fiifi + Xi∈σchf, giigik. (2) The proof will be completed with the following steps. Step 1: For all f ∈ H and σ ⊂ I, k Xi∈σchf, giigi − Xi∈σchf, hiihik ≤ (kTGk + kTHk)kTG − THkkfk. proof of Step 1: Using Proposition 2.5 and utilizing the properties of the respective synthesis operators, for every f ∈ H and σ ⊂ I we have, k Xi∈σchf, giigi − Xi∈σchf, hiihik ≤ kTGkkT ∗ G − T ∗Hkkfk + kTG − THkkT ∗Hkkfk = (kTGk + kTHk)kTG − THkkfk. Step 2: For every weaving, universal lower frame bound is [A−kTG−THk(kTGk+kTHk)]2 B+B1 , where B1 is an upper frame bound for H. proof of Step 2: Applying equation (2) and step 1, for every f ∈ H we obtain, hf, fiifi + Xi∈σchf, hiihik hf, fiifi + Xi∈σchf, giigik − kXi∈σchf, giigi − Xi∈σchf, hiihik k Xi∈σ ≥ kXi∈σ ≥ [A − kTG − THk(kTGk + kTHk)]kfk. 14 Therefore the conclusion follows from Lemma 2.7. The universal upper frame bound of the weaving will be achieved from the Proposition 2.4. Remark 4.7. If the frames F, G are woven in H and G, H are woven in H, then F and H are not necessarily woven in H, which is evident from the following example. Example 4.8. Let F = {e1, e2, 2e1}, G = {2e1,−e2,−2e2} and H = {e1,−e1, 2e2}. Then F and G are woven as well as G and H are woven , but F and H are not woven in R2, as if we consider σ = {3} then the associated weaving is {e1,−e1, 2e1}. Corollary 4.9. Let F = {fi}i∈I be a frame for H with lower frame bound A and G = {gi}i∈I be a frame sequence in H. Let TF and TG be corresponding synthesis operators, respectively. Then F and G are woven if (kTF − TGk)(kTFk + kTGk) < A. Theorem 4.10. Let F = {fi}i∈I be a frame for H with frame bounds A1, B1 and G = {gi}i∈I be a frame sequence in H with bounds A2, B2. Suppose 2 = λ2 < 1 so that (λ1√B1 + 0 < (Pi∈I kfik2) λ2√B2) < A1. Then F and G are woven. 2 = λ1 < 1 and 0 < (Pi∈I kgik2) 1 1 Proof. The proof will be completed with the following steps. Step 1: For every σ ⊂ I and every f ∈ H, k Xi∈σchf, fiifi − Xi∈σchf, giigik ≤ (λ1√B1 + λ2√B2)kfk. 15 proof of Step 1: Using Proposition 2.5, for all f ∈ H and σ ⊂ I we have, k Xi∈σchf, fiifi − Xi∈σchf, giigik ≤ k Xi∈σchf, fiifik + k Xi∈σchf, giigik ≤ Xi∈σc khf, fiifik + Xi∈σc khf, giigik ≤ Xi∈I khf, giigik ≤ (Xi∈I kfik2) + (Xi∈I kgik2) ≤ (λ1√B1 + λ2√B2)kfk. khf, fiifik +Xi∈I 2 (Xi∈I hf, fii2) 2 (Xi∈I hf, gii2) 1 1 2 1 1 2 Step 2: For every weaving the lower frame bound is [A1−(λ1√B1+λ2√B2)]2 proof of Step 2: Applying Step 1, for every f ∈ H we obtain, kXi∈σ hf, fiifi + Xi∈σchf, giigik ≥ kXi∈I hf, fiifik − k Xi∈σchf, fiifi − Xi∈σchf, giigik B1+B2 . ≥ [A1 − (λ1√B1 + λ2√B2]kfk. Therefore applying Lemma 2.7, our goal is executed. Furthermore, universal upper frame bound of the weaving will be accom- plished by utilizing the Proposition 2.4. Acknowledgments The first author is highly indebted to the fiscal support of MHRD, Govern- ment of India. He also extends his massive gratitude to Dr. Manideepa Saha (NIT Meghalaya, India) and Professor Kallol Paul (Jadavpur University, In- dia) for their useful suggestions and comments to improve this article. The second author is supported by DST-SERB project MTR/2017/000797. 16 References [1] D. Gabor, Theory of communication, J.I.E.E. 93 (1946) 429 -- 459. [2] I. Daubechies, A. Grossmann, Y. Mayer, Painless nonorthogonal ex- pansions, Journal of Mathematical Physics 27 (5) (1986) 1271 -- 1283. [3] W. Sun, G-frames and G-riesz bases, J. Math. Anal. Appl. 322 (1) (2006) 437 -- 452. [4] L. Gavrut¸a, Frames for operators, Applied and Computational Har- monic Analysis 32 (1) (2012) 139 -- 144. [5] P. Casazza, G. Kutyniok, Frames of subspaces, Contemporary Math, AMS 345 (2004) 87 -- 114. [6] A. Khosravi, M. S. Asgari, Frames of subspaces and approximation of the inverse frame operator, Houston Journal of Mathematics 33 (3) (2007) 907 -- 920. [7] A. Bhandari, S. Mukherjee, Atomic subspaces for operators, Submitted, arXiv:1705.06042. [8] T. C. Easwaran Nambudiri, K. Parthasarathy, Generalised WeylHeisen- berg frame operators, Bulletin des Sciences Math´ematiques 136 (1) (2012) 44 -- 53. [9] T. C. Easwaran Nambudiri, K. Parthasarathy, A characterisation of WeylHeisenberg frame operators, Bulletin des Sciences Math´ematiques 137 (3) (2013) 322 -- 324. [10] S. Mishra, S. Chakraborty, Dynamical system analysis of quintom dark energy model, The European Physical Journal C 78:917. 17 [11] S. Arati, R. Radha, Frames and riesz bases for shift invariant spaces on the abstract Heisenberg group, Indagationes Mathematicae 30 (1) (2019) 106 -- 127. [12] Deepshikha, L. Vashisht, Necessary and sufficient conditions for discrete wavelet frames in CN, Journal of Geometry and Physics 117 (2017) 134 -- 143. [13] T. C. Easwaran Nambudiri, K. Parthasarathy, Bessel sequences, wavelet frames, duals and extensions, Indagationes Mathematicae 29 (3) (2018) 907 -- 915. [14] L. D. Abreu, J. E. Gilbert, Wavelet-type frames for an interval, Expo- sitiones Mathematicae 32 (3) (2014) 274 -- 283. [15] A. Gamber, N. K. Shukla, Pairwise orthogonal frames generated by reg- ular representations of lca groups, Bulletin des Sciences Math´ematiques 152 (2019) 40 -- 60. [16] T. Bemrose, P. Casazza, K. Grochenig, M. Lammers, R. Lynch, Weav- ing frames, Operators and Matrices 10 (4) (2016) 1093 -- 1116. [17] Deepshikha, L. Vashisht, G. Verma, Generalized weaving frames for operators in Hilbert spaces, Results in Mathematics 72 (3) (2017) 1369 -- 1391. [18] L. Vashisht, Deepshikha, Weaving properties of generalized continuous frames generated by an iterated function system, Journal of Geometry and Physics 110 (2016) 282 -- 295. [19] Deepshikha, L. Vashisht, Vector-valued (super) weaving frames, Jour- nal of Geometry and Physics 134 (2018) 48 -- 57. 18 [20] P. Casazza, G. Kutyniok, Finite Frames: Theory and Applications, Applied and Numerical Harmonic Analysis, Birkhauser Boston, 2012. [21] O. Christensen, Frames and Bases-An Introductory Course, Birkhauser, Boston, 2008. [22] D. Han, K. Kornelson, D. R. Larson, E. Weber, Frames for Undergrad- uates, AMS, 2007. [23] O. M. Baksalary, G. Trenkler, On angles and distances between sub- spaces, Linear Algebra and its Applications 431 (2009) 2243 -- 2260. [24] F. Deutsch, The angle between subspaces of a Hilbert space, Approxi- mation Theory, Wavelets and Applications 454 (1995) 107 -- 130. [25] T. Kato, Perturbation Theory for Linear Operators, Springer, New York, 1980. [26] R. G. Douglas, On majorization, factorization and range inclusion of operators on Hilbert space, Proc. Amer Math. Society 17 (2) (1966) 413 -- 415. 19
1201.4196
1
1201
2012-01-20T01:19:25
Analogs of Cuntz algebras on $L^p$ spaces
[ "math.FA", "math.OA" ]
For $d = 2, 3, \ldots$ and $p \in [1, \infty),$ we define a class of representations $\rho$ of the Leavitt algebra $L_d$ on spaces of the form $L^p (X, \mu),$ which we call the spatial representations. We prove that for fixed $d$ and $p,$ the Banach algebra ${{\mathcal{O}}_{d}^{p}}$ obtained as the closure of the image of $L_d$ under the representation $\rho$ is the same for all spatial representations $\rho.$ When $p = 2,$ we recover the usual Cuntz algebra ${\mathcal{O}}_{d}.$ We give a number of equivalent conditions for a representation to be spatial. We show that for distinct $p_1$ and $p_2$ in $[1, \infty)$ and arbitrary $d_1$ and $d_2$ in $\{ 2, 3, \ldots \},$ there is no nonzero continuous homomorphism from ${\mathcal{O}}_{d_1}^{p_1}$ to ${\mathcal{O}}_{d_2}^{p_2}.$
math.FA
math
ANALOGS OF CUNTZ ALGEBRAS ON Lp SPACES N. CHRISTOPHER PHILLIPS Abstract. For d = 2, 3, . . . and p ∈ [1, ∞), we define a class of representations ρ of the Leavitt algebra Ld on spaces of the form Lp(X, µ), which we call the spatial representations. We prove that for fixed d and p, the Banach algebra Op d = ρ(Ld) is the same for all spatial representations ρ. When p = 2, we recover the usual Cuntz algebra Od. We give a number of equivalent conditions for a representation to be spatial. We show that for distinct p1, p2 ∈ [1, ∞) and arbitrary d1, d2 ∈ {2, 3, . . .}, there is no nonzero continuous homomorphism from Op1 d1 to Op2 d2 . The algebras that we call the Leavitt algebras Ld (see Definition 1.1) are a special case of algebras introduced in characteristic 2 by Leavitt [20], and generalized to arbitrary ground fields in [21]. The Cuntz algebra Od, introduced in [9], can be defined as the norm closure of the range of a *-representation of Ld on a Hilbert space. (Cuntz did not define Od this way.) The Cuntz algebras have turned out to be one of the most fundamental families of examples of C*-algebras. Leavitt algebras were long obscure, but they have recently attracted renewed attention. In this paper we study the analogs of Cuntz algebras on Lp spaces. That is, we consider the norm closure of the range of a representation of Ld on a space of the form Lp(X, µ). It turns out that there is a rich theory of such algebras, of which we exhibit the beginning in this paper. Our main results are as follows. There are many possible Lp analogs of Cuntz algebras (although we mostly do not know for sure that they really are essentially different), but there is a natural class of such algebras, namely those that come from what we call spatial representations (Definition 7.4(2)). Spatial representa- tions are those for which the standard generators form an Lp analog of what is called a row isometry in multivariable operator theory on Hilbert space. (The sur- vey article [11] emphasizes the more general row contractions on Hilbert spaces, but row isometries also play a significant role. See especially Section 6.2 of [11].) We give a number of rather different equivalent conditions for a representation to be spatial-further evidence that this is a natural class. For fixed p ∈ [1, ∞) \ {2}, we prove a uniqueness theorem: the Banach algebras coming from any two spa- tial representations are isometrically isomorphic via an isomorphism which sends the standard generators to the standard generators. We call the Banach algebra obtained this way Op d. We further obtain a strong dependence on p. Specifically, for p1 6= p2 and any d1, d2 ∈ {2, 3, . . .}, there is no nonzero continuous homomorphism from Op1 d1 d. The usual Cuntz algebra is O2 to Op2 d2 . Date: 19 January 2012. 2000 Mathematics Subject Classification. Primary 46H05, 47L10; Secondary 46H35. This material is based upon work supported by the US National Science Foundation under Grants DMS-0701076 and DMS-1101742. It was also partially supported by the Centre de Recerca Matem`atica (Barcelona) through a research visit conducted during 2011. 1 2 N. CHRISTOPHER PHILLIPS Some of our results are valid, with the same proofs, for the infinite Leavitt al- gebra L∞ (which gives Lp analogs of O∞) and for the Cohn algebras (which give Lp analogs of Cuntz's algebras Ed). In such cases, we include the corresponding re- sults. However, in many cases, the results, or at least the proofs, must be modified. We do not go in that direction in this paper. The methods here have little in common with C*-algebra methods. Indeed, the results on spatial representations have no analog for C*-algebras, and the result on nonexistence of nonzero continuous homomorphisms does not make sense if one only considers C*-algebras. Uniqueness of Op d is of course true when p = 2, but, as far as we can tell, our proof for p 6= 2 does not work when p = 2. This paper is organized as follows. In Section 1, we define Leavitt algebras and Cohn algebras, and give some basic facts about them which will be needed in the rest of the paper. In Section 2, we discuss representations on Banach spaces. We define several natural conditions on representations (weaker than being spatial). We describe ways to get new representations from old ones, some of which work in general and some of which are special to Leavitt and Cohn algebras. Section 3 contains a large collection of examples of representations on Lp spaces. In Section 6 we develop the theory of (semi)spatial partial isometries on Lp spaces associated to σ-finite measure spaces. The results here are the basic technical tools needed to prove our main results. Roughly speaking, a semispatial partial isometry from Lp(X, µ) to Lp(Y, ν) comes from a map from a subset of Y to a subset of X. Lamperti's Theorem [18], which plays a key role, asserts that for p ∈ (0, ∞) \ {2}, every isometry from Lp(X, µ) to Lp(Y, ν) is semispatial. It is unfortunately not really true that semispatial partial isometries from Lp(X, µ) to Lp(Y, ν) come from point maps. Instead, they come from suitable homomor- phisms of σ-algebras. For our theory of spatial partial isometries, we need a much more extensive theory of these than we have been able to find in the literature. In Section 4 we recall some standard facts about Boolean σ-algebras, and in Section 5 we discuss the maps on functions and measures induced by a suitable homomor- phism of the Boolean σ-algebras of measurable sets mod null sets. Sections 7, 8, and 9 contain our main results. In Section 7, we give equivalent conditions for representations of Leavitt algebras on Lp spaces to be spatial. Along the way, we define spatial representations of the algebra Md of d × d matrices, and we give a number of equivalent conditions for a representation of Md to be spatial. In Section 8, we prove that and two spatial representations of the Leavitt algebra Ld on Lp spaces give the same norm on Ld, and thus lead to isometrically isomorphic Banach algebras. In Section 9, we prove the nonexistence of nonzero continuous homomorphisms between the resulting algebras for different values of p. In [22], we will show that Op d is an amenable purely infinite simple Banach algebra, and in [23], we will show that its topological K-theory is the same as for the ordinary Cuntz algebra Od. The methods in these papers are much closer to C*-algebra methods. Scalars will always be C. Much of what we do also works for real scalars. We use complex scalars for our proof of the equivalence of several of the conditions in Theorem 7.2 for a representation of Md to be spatial. We do not know whether complex scalars are really necessary. We have tried to make this paper accessible to operator algebraists who are not familiar with operators on spaces of the form Lp(X, µ). CUNTZ ALGEBRAS ON Lp SPACES 3 We are grateful to Joe Diestel, Ilijas Farah, Coenraad Labuschagne, and Volker Runde for useful email discussions and for providing references. We are especially grateful to Bill Johnson for extensive discussions about Banach spaces, and to Guillermo Cortinas and Mar´ıa Eugenia Rodr´ıguez, who carefully read an early draft and whose comments led to numerous corrections and improvements. Some of this work was carried out during a visit to the Instituto Superior T´ecnico, Universidade T´ecnica de Lisboa, and during an extended research visit to the Centre de Recerca Matem`atica (Barcelona). I also thank the Research Institute for Mathematical Sciences of Kyoto University for its support through a visiting professorship. I am grateful to all these institutions for their hospitality. 1. Leavitt and Cohn algebras In this section, we define Leavitt algebras and some of their relatives. We describe a grading, a linear involution, and a conjugate linear involution. We give some useful computational lemmas. Definition 1.1. Let d ∈ {2, 3, 4, . . .}. We define the Leavitt algebra Ld to be the universal complex associative algebra on generators s1, s2, . . . , sd, t1, t2, . . . , td sat- isfying the relations (1.1) (1.2) and (1.3) tjsj = 1 for j ∈ {1, 2, . . . , d}, tjsk = 0 for j, k ∈ {1, 2, . . . , d} with j 6= k, sjtj = 1. dXj=1 These algebras were introduced in Section 3 of [20] (except that the base field there is Z/2Z), and they are simple (Theorem 2 of [21], with an arbitrary choice of the field). Definition 1.2. Let d ∈ {2, 3, 4, . . .}. We define the Cohn algebra Cd to be the universal complex associative algebra on generators s1, s2, . . . , sd, t1, t2, . . . , td sat- isfying the relations (1.1) and (1.2) (but not (1.3)). These algebras are a special case of algebras introduced at the beginning of Section 5 of [8]. What we have called Cd is called U1,d in [8], and also in [8] the field is allowed to be arbitrary. Our notation, and the name "Cohn algebra", follow Definition 1.1 of [4], except that we specifically take the field to be C and suppress it in the notation. Definition 1.3. Let d ∈ {2, 3, 4, . . .}. We define the (infinite) Leavitt algebra L∞ to be the universal complex associative algebra on generators s1, s2, . . . , t1, t2, . . . satisfying the relations (1.4) and (1.5) tjsj = 1 for j ∈ Z>0 tjsk = 0 for j, k ∈ Z>0 with j 6= k. When it is necessary to distinguish the generators of L∞ from those of Ld and Cd, we denote them by s(∞) , . . . , t(∞) , s(∞) , t(∞) 2 , . . . . 1 2 1 Cd1 ιd1,d2−−−−→ Ld2 Ed1 −−−−→ Od2. y y 4 N. CHRISTOPHER PHILLIPS This algebra is simple, by Example 3.1(ii) of [3]. Remark 1.4. The algebras Ld, Cd, and L∞ are all examples of Leavitt path algebras. For Ld see Example 1.4(iii) of [2], for Cd see Section 1.5 of [1], and for L∞ see Example 3.1(ii) of [3]. (Warning: There are two possible conventions for the choice of the direction of the arrows in the graph, and both are in common use.) Lemma 1.5. Let Cd be as in Definition 1.2 and let Ld be as in Definition 1.1, with the generators named as there (using the same names in both kinds of alge- bras). For d1, d2 ∈ {2, 3, 4, . . . , ∞} with d1 < d2, there is a unique homomorphism ιd1,d2 : Cd1 → Ld2 such that ιd1,d2(sj) = sj and ιd1,d2(tj ) = tj for j ∈ {1, 2, . . . , d1}. Moreover, ιd1,d2 is injective. Proof. Existence and uniqueness of ιd1,d2 are immediate from the definitions of the algebras as universal algebras on generators and relations. We prove injectivity. We presume that there is a purely algebraic proof, but one can easily see this by comparing with the C*-algebras, following Remark 2.9 below. Let Ed1 be the extended Cuntz algebra, as in Remark 2.9. There is a commutative diagram The left vertical map is injective by Theorem 7.3 of [25] and Remark 1.4, and the bottom horizontal map is well known to be injective. Therefore ιd1,d2 is injective. (cid:3) Lemma 1.6. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). (1) There exists a unique conjugate linear antimultiplicative involution a 7→ a∗ on A such that s∗ j = tj and t∗ j = sj for all j. (2) There exists a unique linear antimultiplicative involution a 7→ a′ on A such that s′ j = tj and t′ j = sj for all j. The properties of a 7→ a∗ are just the algebraic properties of the adjoint of a complex matrix: (a + b)∗ = a∗ + b∗, (λa)∗ = λa∗, (ab)∗ = b∗a∗, and (a∗)∗ = a for all a, b ∈ A and λ ∈ C. The properties of a 7→ a′ are the same, except that it is linear: (λa)′ = λa′ for all a ∈ A and λ ∈ C. Proof of Lemma 1.6. See Remark 3.4 of [25], where explicit formulas, valid for any graph algebra, are given, and Remark 1.4. Both parts may also be easily obtained using the universal properties of algebras on generators and relations: a 7→ a is a homomorphism from A to its opposite algebra, and a 7→ a∗ is the composition of a 7→ a with a homomorphism from A to its complex conjugate algebra. (cid:3) One can get Lemma 1.6(1) by using the fact (Theorem 7.3 of [25] and Remark 1.4) that there are injective maps from Ld, Cd, and L∞ to C*-algebras which preserve the intended involution. (For Ld and L∞, injectivity is automatic because the algebras are simple. See Remark 2.9 for definitions of *-representations on Hilbert spaces.) CUNTZ ALGEBRAS ON Lp SPACES 5 Proposition 1.7. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Then there is a unique Z-grading on A determined by deg(sj) = 1 and deg(tj) = −1 for all j. Proof. The proof is easy. (See after Definition 3.12 in [25].) (cid:3) We will need some of the finer algebraic structure of Ld, and associated notation. We roughly follow the beginning of Section 1 of [9], starting with 1.1 of [9]. n = {1, 2, . . . , d}n, and we define W ∞ Notation 1.8. Let d ∈ {2, 3, 4, . . . , ∞}, and let n ∈ Z≥0. For d < ∞, we define W d n is the set of all sequences α = (cid:0)α(1), α(2), . . . , α(n)(cid:1) with α(l) ∈ {1, 2, . . . , d} (or α(l) ∈ Z>0 if d = ∞) for l = 1, 2, . . . , n. We set n = (Z>0)n. Thus, W d W d ∞ = W d n . ∞an=0 ∞ words (on {1, 2, . . . , d} or Z>0 as appropriate). We call the elements of W d α ∈ W d α ∈ W d which we write as ∅. For α ∈ W d a word in W d If ∞, the length of α, written l(α), is the unique number n ∈ Z≥0 such that n . Note that there is a unique word of length zero, namely the empty word, n , we denote by αβ the concatenation, m and β ∈ W d m+n. Notation 1.9. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ n . If n ≥ 1, (Definition 1.3). Let n ∈ Z≥0, and let α =(cid:0)α(1), α(2), . . . , α(n)(cid:1) ∈ W d we define sα, tα ∈ A by sα = sα(1)sα(2) · · · sα(n−1)sα(n) and tα = tα(n)tα(n−1) · · · tα(2)tα(1). We take s∅ = t∅ = 1. For emphasis: order . We do this to get convenient formulas in Lemma 1.10(3). when working with Cuntz algebras, one simply uses s∗ to have s∗ in the definition of tα, we take the generators tα(l) in reverse In particular, j in place of tj, and we want α = tα. Lemma 1.10. Let the notation be as in Notation 1.8 and Notation 1.9. (1) Let α, β ∈ W d (2) In the Z-grading on A of Proposition 1.7, we have deg(sα) = l(α) and ∞. Then sαβ = sαsβ and tαβ = tβtα. deg(tα) = −l(α) for all α ∈ W d ∞. (3) Let α ∈ W d ∞. Then the involutions of Lemma 1.6 satisfy s′ α = s∗ α = tα and t′ α = t∗ α = sα. (4) Let a1, a2, . . . , an ∈ {s1, s2, . . .} ∪ {t1, t2, . . .}. Suppose a1a2 · · · an 6= 0. Then there exist unique α, β ∈ W d a1a2 · · · an = sαtβ. ∞ such that (5) Let α, β ∈ W d ∞ satisfy l(α) = l(β). Then tβsα = 1 if α = β, and tβsα = 0 otherwise. Proof. Parts (1), (2), (3), and (5) are obvious. (Part (5) is also in Lemma 1.2(b) of [9].) Using Part (3), we see that Part (4) is Lemma 1.3 of [9]. (cid:3) 6 N. CHRISTOPHER PHILLIPS Lemma 1.11. Let d ∈ {2, 3, 4, . . .}, let Ld be as in Definition 1.1, and let m ∈ Z≥0. Then the collection (sαtβ)α,β∈W d is a system of matrix units for a unital subalgebra of Ld isomorphic to Mdm. That is, identifying Mdm with the linear maps on a vector space with basis W d m, there is a unique homomorphism ϕm : Mdm → Ld such that ϕm(eα,β) = sαtβ. m, with matrix units eα,β for α, β ∈ W d m sαtα = 1, by induction on m. The case m = 1 is relation (1.3) in Definition 1.1. Assuming the result holds for m, use this and the case m = 1 at the last step to get m Proof. We prove that Pα∈W d dXj=1 sαtα = Xβ∈W d Xα∈W d m+1 m sβsjtjtβ = Xβ∈W d m sβ(cid:18)Xd j=1 sjtj(cid:19) tβ = 1. The statement of the lemma now follows from Lemma 1.10(5), or is Proposi- (cid:3) tion 1.4 of [9]. Lemma 1.12. Let d ∈ {2, 3, 4, . . .}, let m ∈ Z>0, and let a1, a2, . . . , am ∈ Ld. Then there exist n ∈ Z≥0, a finite set F ⊂ W d ∞, and numbers λk,α,β ∈ C for k = 1, 2, . . . , m, α ∈ F, and β ∈ W d n , such that (1.6) for k = 1, 2, . . . , m. ak = Xα∈F Xβ∈W d n λk,α,β sαtβ Proof. Since a1, a2, . . . , am are linear combinations of products of the sj and tj, it suffices to prove the statement when a1, a2, . . . , am are products of the sj and tj. By Lemma 1.10(4), we may assume ak = sαk tβk with αk, βk ∈ W d ∞. Set For k = 1, 2, . . . , m, set lk = n − l(βk). Take n = max(cid:0)l(β1), l(β2), . . . , l(βm)(cid:1). Lemma 1.11 and Lemma 1.10(1) imply F = m[k=1(cid:8)αkα : α ∈ W d lk(cid:9). sαk sαtαtβk = Xα∈W d lk lk ak = sαk tβk = Xα∈W d sαkαtβkα. This expression has the form in (1.6). (cid:3) Definition 1.13. Let A be any of Ld, Cd, or L∞. Let λ = (λ1, λ2, . . . , λd) ∈ Cd. j=1 C, and take λ = (λ1, λ2, . . .).) Define (For A = L∞, take d = ∞, take Cd =L∞ sλ, tλ ∈ A by sλ = λj sj and tλ = dXj=1 λj tj. dXj=1 In principle, this notation conflicts with Notation 1.9, but no confusion should arise. Lemma 1.14. Let the notation be as in Definition 1.13. Let λ, γ ∈ Cd. Then tλsγ =(cid:16)Xj λj γj(cid:17) · 1. CUNTZ ALGEBRAS ON Lp SPACES 7 Proof. This is immediate from the relations tjsk = 1 for j = k and tjsk = 0 for j 6= k. (cid:3) 2. Representations on Banach spaces In this section, we discuss representations of Leavitt and Cohn algebras on Ba- nach spaces. Much of what we say makes sense for representations on general Banach spaces, but some only works for representations on spaces of the form Lp(X, µ). Some of the constructions work for general algebras, but some are special to representations of Leavitt and Cohn algebras. Some of what we do is intended primarily to establish notation and conventions. (For example, representations are required to be unital, and isomorphisms are required to be surjective.) All Banach spaces in this article will be over C. Notation 2.1. Let E and F be Banach spaces. We denote by L(E, F ) the Banach space of all bounded linear operators from E to F, and by K(E, F ) ⊂ L(E, F ) the closed subspace of all compact linear operators from E to F. When E = F, we get the Banach algebra L(E) and the closed ideal K(E) ⊂ L(E). The following definition summarizes terminology for Banach spaces that we use. We will need both isometries and isomorphisms of Banach spaces. Definition 2.2. If E and F are Banach spaces, we say that a ∈ L(E, F ) is an iso- morphism if a is bijective. (By the Open Mapping Theorem, a−1 is also bounded.) If an isomorphism exists, we say E and F are isomorphic. We say that a ∈ L(E, F ) is an isometry if kaξk = kξk for every ξ ∈ E. (We do not require that a be surjective.) If there is a surjective isometry from E to F, we say that E and F are isometrically isomorphic. If A and B are Banach algebras, we say that a homomorphism ϕ : A → B is an isomorphism if it is continuous and bijective. (By the Open Mapping Theorem, ϕ−1 is also continuous.) If an isomorphism exists, we say A and B are isomorphic. If in addition ϕ is isometric, we call it an isometric isomorphism. If such a map exists, we say A and B are isometrically isomorphic. For emphasis (because some authors do not use this convention): isomorphisms are required to be surjective. The following notation for duals is intended to avoid conflict with the notation for adjoints. Notation 2.3. Let E be a Banach space. We denote by E′ its dual Banach space L(E, C), consisting of all bounded linear functionals on E. If F is another Banach space and a ∈ L(E, F ), we denote by a′ the element of L(F ′, E′) defined by a′(ω)(ξ) = ω(aξ) for ξ ∈ E and ω ∈ F ′. We will also need some notation for specific spaces. Notation 2.4. For any set S, we give lp(S) the usual meaning (using counting measure on S), and we take (as usual) lp = lp(Z>0). For d ∈ Z>0 and p ∈ [1, ∞], we let lp norm, and we algebraically identify M p matrices in the standard way. d = lp(cid:0){1, 2, . . . , d}(cid:1). We further let M p d with the algebra Md of d × d complex d = L(cid:0)lp d(cid:1) with the usual operator 8 N. CHRISTOPHER PHILLIPS We warn of a notational conflict. Many articles on Banach spaces use Lp(X, µ) rather than Lp(X, µ), and use ld d. Our convention is chosen to avoid conflict with the standard notation for the Leavitt algebra Ld of Definition 1.1. d are of course equivalent, but they are For fixed d, the norms on the various M p p for what we call lp not equal. The following example illustrates this. Example 2.5. Let d ∈ Z>0. Let η, µ ∈ Cd, and let ωη : Cd → C be the linear j=1 ηj ξj. Let a ∈ Md be the rank one operator given by q = 1, and regard a as an element of M p d . functional ωη(ξ) = Pd aξ = ωη(ξ)µ. Let p, q ∈ [1, ∞] satisfy 1 Then one can calculate that kak = kµkpkηkq. p + 1 The following terminology and related observation will be used many times. Definition 2.6. Let (X, B, µ) be a measure space and let p ∈ [1, ∞]. For a function ξ ∈ Lp(X, µ) (or, more generally, any measurable function on X) and a subset E ⊂ X, we will say that ξ is supported in E if ξ(x) = 0 for almost all x ∈ X \ E. If I is a countable set, and (ξi)i∈I is a family of elements of Lp(X, µ) or measurable functions on X, we say that the ξi have disjoint supports if there are disjoint subsets Ei ⊂ X such that ξi is supported in Ei for all i ∈ I. Remark 2.7. Let (X, B, µ) be a measure space, let p ∈ [1, ∞), let I be a countable set, and let ξi ∈ Lp(X, µ), for i ∈ I, have disjoint supports. Then p p (cid:13)(cid:13)(cid:13)Xi∈I ξi(cid:13)(cid:13)(cid:13) =Xi∈I kξikp p. Definition 2.8. Let A be any unital complex algebra. Let E be a nonzero Banach space. A representation of A on E is a unital algebra homomorphism from A to L(E). We do not say anything about continuity. We will mostly be interested in repre- sentations of Ld, Cd, and L∞, for which we do not use a topology on the algebra, or of Md, for which all representations are continuous by finite dimensionality. Remark 2.9. The well known representations of Ld, Cd, and L∞ are those on a Hilbert space H. Choose any d isometries w1, w2, . . . , wd ∈ L(H) (or, for the case of L∞, isometries w1, w2, . . . ∈ L(H)) with orthogonal ranges. Then we obtain a representation ρ : Cd → L(H) or ρ : L∞ → L(H) by setting ρ(sj) = wj and ρ(tj) = w∗ j = 1, we get a representation of Ld. These representations are even *-representations: making A a *-algebra as in Lemma 1.6(1), we have ρ(a∗) = ρ(a)∗ for all a ∈ A. j for all j. If d < ∞ and Pd j=1 wjw∗ ρ(cid:16)1 −Pd The closures ρ(A) do not depend on the choice of ρ (in case A = Cd, provided j(cid:17) 6= 0), and are the usual Cuntz algebra Od when A = Ld (in- cluding the case d = ∞), and the extended Cuntz algebras Ed when A = Cd. See Theorem 1.12 of [9] for Ld and L∞, and see Lemma 3.1 of [10] for Cd. j=1 sjs∗ Further examples of representations of Ld, Cd, and L∞ will be given in Section 3. Representations of Ld, Cd, and L∞ have a kind of rigidity property. It is stronger for Ld than for the others: a representation is determined by the images of the sj or by the images of the tj. Lemma 2.10. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let B be a unital algebra over C, and let ϕ, ψ : Ld → B be unital CUNTZ ALGEBRAS ON Lp SPACES 9 homomorphisms such that for all j we have ϕ(sj) = ψ(sj) and ϕ(sj tj) = ψ(sjtj). Then ϕ = ψ. The same conclusion holds if we replace ϕ(sj) = ψ(sj) with ϕ(tj ) = ψ(tj). Proof. Assume ϕ(sj) = ψ(sj) for all j. Using the relations tjsjtj = tj at the first step and ϕ(tj)ϕ(sj ) = 1 at the last step, we calculate: ϕ(tj ) = ϕ(tj )ϕ(sj tj) = ϕ(tj)ψ(sj tj) = ϕ(tj)ψ(sj )ψ(tj) = ϕ(tj )ϕ(sj )ψ(tj) = ψ(tj ). The first statement follows. (using sjtjsj = sj) gives ϕ(sj) = ϕ(sj tj)ϕ(sj) If instead ϕ(tj ) = ψ(tj ) for all j, similar reasoning = ψ(sjtj)ϕ(sj ) = ψ(sj)ψ(tj )ϕ(sj ) = ψ(sj)ϕ(tj )ϕ(sj ) = ψ(sj). This completes the proof. (cid:3) Lemma 2.11. Let d ∈ {2, 3, 4, . . .}, let B be a unital algebra over C, and let ϕ, ψ : Ld → B be unital homomorphisms such that ϕ(sj) = ψ(sj ) for j ∈ {1, 2, . . . , d}. Then ϕ = ψ. The same conclusion holds if we replace ϕ(sj) = ψ(sj) with ϕ(tj ) = ψ(tj). Proof. For j ∈ {1, 2, . . . , d}, define idempotents ej, fj ∈ B by ej = ϕ(sjtj) and fj = ψ(sjtj). By Lemma 2.10, it suffices to show that ej = fj for all j. First assume that ϕ(sj) = ψ(sj) for all j. Using this statement at the first and k=1 ek = 1 at the last step, we have third steps, tjsk = 0 for j 6= k at the second step, and Pd dXk=1 fjej = ψ(sj tjsj)ϕ(tj ) = ψ(sjtjsk)ϕ(tk) = dXk=1 (2.1) fjek = fj. If now j 6= k, then This equation, together withPd (2.2) fkej = fkekej = 0. k=1 fk = 1, gives ej = fkej = fjej. dXk=1 dXk=1 The proof is completed by combining (2.1) and (2.2). Now assume that ϕ(tj ) = ψ(tj) for all j. With similar justifications, we get fjej = ψ(sj )ϕ(tjsjtj) = ψ(sk)ϕ(tksjtj) = fkej = ej. dXk=1 Pd So for j 6= k we have fjek = fjfkek = 0. Combining these results gives fj = (cid:3) k=1 fjek = fjej, whence ej = fj. The analog of Lemma 2.11 for L∞ and Cd is false. See Example 3.5. The following definition gives several natural conditions to ask of a representation of Ld, Cd, or L∞ on a Banach space E. The condition in (3) is motivated by the following property of a *-representation ρ of Ld or Cd on a Hilbert space H (as in Remark 2.9): for λ1, λ2, . . . , λd ∈ C and ξ ∈ H, we have (2.3) (cid:13)(cid:13)(cid:13)(cid:13)ρ(cid:18)Xd j=1 λj sj(cid:19) ξ(cid:13)(cid:13)(cid:13)(cid:13) = k(λ1, λ2, . . . , λd)k2kξk. 10 N. CHRISTOPHER PHILLIPS In Definition 7.4 and Definition 7.6, we will see further conditions on representations which are natural when E = Lp(X, µ). Definition 2.12. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2) or L∞ (Definition 1.3). Let E be a nonzero Banach space, and let ρ : A → L(E) be a representation. (1) We say that ρ is contractive on generators if for every j, we have kρ(sj)k ≤ 1 and kρ(tj)k ≤ 1. (2) We say that ρ is forward isometric if ρ(sj) is an isometry for every j. (3) We say that ρ is strongly forward isometric if ρ is forward isometric and (following Definition 1.13) for every λ ∈ Cd, the element ρ(sλ) is a scalar multiple of an isometry. Remark 2.13. A representation which is contractive on generators is clearly for- ward isometric. A representation of Ld which is contractive on generators need not be strongly forward isometric. See Example 3.11 below. We will see in Example 3.5 below that a strongly forward isometric representation of L∞ need not be contractive on generators. We do not know whether this can happen for Ld with d finite. We now describe several ways to make new representations from old ones. The first two (direct sums and tensoring with the identity on some other Banach space) work for representations of general algebras. They also work for more general choices of norms on the direct sum and tensor product than we consider here. For simplicity, we restrict to specific choices which are suitable for representations on spaces of the form Lp(X, µ). Lemma 2.14. Let A be a unital complex algebra, and let p ∈ [1, ∞]. Let n ∈ Z>0, and for l = 1, 2, . . . , n let (Xl, Bl, µl) be a σ-finite measure space and let ρl : A → l=1 Lp(Xl, µl) with the norm L(Lp(Xl, µl)) be a representation. Equip E =Ln k(ξ1, ξ2, . . . , ξn)k =(cid:16)Xn l=1 kξlkp p(cid:17)1/p . Then there is a unique representation ρ : A → L(E) such that ρ(a)(ξ1, ξ2, . . . , ξn) =(cid:0)ρ1(a)ξ1, ρ2(a)ξ2, . . . , ρn(a)ξn(cid:1) for a ∈ Ld and ξl ∈ Lp(Xl, µl) for l = 1, 2, . . . , n. If A is any of Ld, Cd, or L∞, and each ρl is contractive on generators or forward isometric, then so is ρ. Proof. This is immediate. (cid:3) Remark 2.15. The norm used in Lemma 2.14 identifies E with Lp (`n using the obvious measure. We write this space as Lp(X1, µ1) ⊕p Lp(X2, µ2) ⊕p · · · ⊕p Lp(Xn, µn). l=1 Xl) , We write the representation ρ as and call it the p-direct sum of ρ1, ρ2, . . . , ρn. ρ = ρ1 ⊕p ρ2 ⊕p · · · ⊕p ρn, Example 3.11 below shows that if ρ1, ρ2, . . . , ρn are strongly forward isometric, it does not follow that ρ is strongly forward isometric. One can form a p-direct sum Li∈I ρi over an infinite index set I provided supi∈I kρi(a)k < ∞ for all a ∈ A. CUNTZ ALGEBRAS ON Lp SPACES 11 We now consider tensoring with the identity on some other Banach space. This requires the theory of tensor products of Banach spaces and of operators on them. We will consider only a very special case. Fix p ∈ [1, ∞). We need a tensor product, defined on pairs of Banach spaces both of the form Lp(X, µ) for σ-finite measures µ, for which one has a canonical isometric identification Lp(X, µ) ⊗ Lp(Y, ν) = Lp(X × Y, µ × ν). (One can't reasonably expect something like this for p = ∞.) The tensor product described in Chapter 7 of [12] will serve our purpose. To simplify the notation, we simply write Lp(X, µ) ⊗p Lp(Y, ν). Lp(X, µ)e⊗∆p Lp(Y, ν) We note that there is a more general construction, the M -norm of [7], defined before (4) on page 3 of [7]. This norm is defined for the tensor product of a Banach lattice (this includes all spaces Lp(X, µ) for all p) and a Banach space. Theorem 3.2(1) of [7] shows that whenever p ∈ [1, ∞) and (X, B, µ) is a finite measure space, then the completion of E ⊗alg Lp(X, µ) in this norm is isometrically isomorphic to the space of Lp functions on X with values in E. In particular, regardless of the value of p, this norm gives the properties in Theorem 2.16. Theorem 2.16. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces. Let p ∈ [1, ∞). Write Lp(X, µ) ⊗p Lp(Y, ν) for the Banach space completed tensor prod- isomorphism uct Lp(X, µ)e⊗∆p Lp(Y, ν) defined in 7.1 of [12]. Then there is a unique isometric Lp(X, µ) ⊗p Lp(Y, ν) ∼= Lp(X × Y, µ × ν) which identifies, for every ξ ∈ Lp(X, µ) and η ∈ Lp(Y, ν), the element ξ ⊗ η with the function (x, y) 7→ ξ(x)η(y) on X × Y. Moreover: (1) Under the identification above, the linear span of all ξ ⊗ η, for ξ ∈ Lp(X, µ) and η ∈ Lp(Y, ν), is dense in Lp(X × Y, µ × ν). (2) kξ ⊗ ηkp = kξkpkηkp for all ξ ∈ Lp(X, µ) and η ∈ Lp(Y, ν). (3) The tensor product ⊗p is commutative and associative. (4) Let (X1, B1, µ1), (X2, B2, µ2), (Y1, C1, ν1), and (Y2, C2, ν2) be σ-finite measure spaces. Let a ∈ L(cid:0)Lp(X1, µ1), Lp(X2, µ2)(cid:1) and b ∈ L(cid:0)Lp(Y1, ν1), Lp(Y2, ν2)(cid:1). Then there exists a unique c ∈ L(cid:0)Lp(X1 × Y1, µ1 × ν1), Lp(X2 × Y2, µ2 × ν2)(cid:1) such that, making the identification above, c(ξ ⊗ η) = aξ ⊗ bη for all ξ ∈ Lp(X1, µ1) and η ∈ Lp(Y1, ν1). We call this operator a ⊗ b. (5) The operator a ⊗ b of (4) satisfies ka ⊗ bk = kak · kbk. (6) The tensor product of operators defined in (4) is associative, bilinear, and satisfies (when the domains are appropriate) (a1 ⊗b1)(a2 ⊗b2) = a1a2 ⊗b1b2. Proof. The identification of Lp(X, µ)e⊗∆p Lp(Y, ν) is in 7.2 of [12]. Part (1) is part of the definition of a tensor product of Banach spaces. Part (2) is in 7.1 of [12]. Part (3) follows from the corresponding properties of products of measure spaces. 12 N. CHRISTOPHER PHILLIPS Parts (4) and (5) are a special case of 7.9 of [12], or of Theorem 1.1 of [14]. Part (6) follows from part (4) and part (1) by examining what happens on elements of the form ξ ⊗ η. (cid:3) In fact, the statements about tensor products of operators in Theorem 2.16(4) and Theorem 2.16(5) are valid in considerably greater generality; see, for example, Theorem 1.1 of [14]. We need a slightly more general statement in the proof of the following lemma. Lemma 2.17. Let A be a unital complex algebra, let p ∈ [1, ∞), let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, and let ρ : A → L(Lp(X, µ)) be a representa- tion. Then there is a unique representation ρ ⊗p 1 : A → Lp(X × Y, µ × ν) such that, following Theorem 2.16(4), we have (ρ ⊗p 1)(a) = ρ(a) ⊗ 1 for all a ∈ A. This representation satisfies k(ρ ⊗p 1)(a)k = kρ(a)k for all a ∈ A. If A is any of Ld, Cd, or L∞, and ρ is any of contractive on generators, forward isometric, or strongly forward isometric, then so is ρ ⊗p 1. Proof. Existence of ρ ⊗p 1 and k(ρ ⊗p 1)(a)k = kρ(a)k follow from parts (4), (5), and (6) of Theorem 2.16. If A is one of Ld, Cd, or L∞ and ρ is contractive on gener- ators, then the norm equation implies that ρ ⊗p 1 is also contractive on generators. To prove that ρ ⊗p 1 is forward isometric or strongly forward isometric when ρ is, it suffices to prove that if s ∈ L(Lp(X, µ)) is isometric (not necessarily surjective), then so is s ⊗ 1 ∈ L(cid:0)Lp(X × Y, µ × ν)(cid:1). Let E ⊂ Lp(X, µ) be the range of s. Let t : E → Lp(X, µ) be the inverse of the corestriction of s. Then ktk = 1. Let F ⊂ Lp(X × Y, µ × ν) be the closed linear span of all ξ ⊗ η with ξ ∈ E and η ∈ Lp(Y, ν). Then Theorem 1.1 of [14] implies that t ⊗ 1 : F → Lp(X × Y, µ × ν) is defined and satisfies kt ⊗ 1k = 1. Since ks ⊗ 1k = 1 and (t ⊗ 1)(s ⊗ 1) = 1, this implies that s ⊗ 1 is isometric. (cid:3) In the proof of Lemma 2.17, we could also have used Lamperti's Theorem (The- orem 6.9 below) instead of Theorem 1.1 of [14] to show that kt ⊗ 1k = 1. Finally, we present constructions of new representations that are special to the kinds of algebras we consider. They will play important technical roles. Lemma 2.18. Let d ∈ {2, 3, 4, . . .}, let E be a nonzero Banach space, and let ρ : Ld → L(E) be a representation. Let u ∈ L(E) be invertible. Then there is a unique representation ρu : Ld → L(E) such that for j = 1, 2, . . . , d we have ρu(sj) = uρ(sj) and ρu(tj) = ρ(tj)u−1. Assume further that u is isometric. isometric, or strongly forward isometric (Definition 2.12), then so is ρu. If ρ is contractive on generators, forward Proof. For the first part, we check the relations (1.1), (1.2), and (1.3) in Defini- tion 1.1. For j ∈ {1, 2, . . . , d} we have [ρ(tj)u−1][uρ(sj)] = ρ(tj)ρ(sj) = 1, for distinct j, k ∈ {1, 2, . . . , d} we have [ρ(tj)u−1][uρ(sk)] = ρ(tj)ρ(sk) = 0, and we have dXj=1 [uρ(sj)][ρ(tj )u−1] = u(cid:18)Xd j=1 sjtj(cid:19) u−1 = u · 1 · u−1 = 1. CUNTZ ALGEBRAS ON Lp SPACES 13 The second part follows directly from the definitions of the conditions on the (cid:3) representation. For representations of Cd and L∞, we only need one sided invertibility. Lemma 2.19. Let d ∈ {2, 3, 4, . . .}, let E be a nonzero Banach space, and let ρ : Cd → L(E) be a representation. Let u, v ∈ L(E) satisfy vu = 1. Then there is a unique representation ρu,v : Cd → L(E) such that for j = 1, 2, . . . , d we have ρu,v(sj) = uρ(sj) and ρu,v(tj) = ρ(tj)v. If u is an isometry and ρ is forward isometric or strongly forward isometric, then so is ρu,v. If kuk, kvk ≤ 1 and ρ is contractive on generators, then so is ρu,v. Proof. The proof is essentially the same as that of Lemma 2.18, using the relations (1.1) and (1.2) in Definition 1.1 (see Definition 1.2). Since no relation involves sjtj, we do not need to have uv = 1. (cid:3) Lemma 2.20. Let ρ : L∞ → L(E) be a representation. Let u, v ∈ L(E) satisfy vu = 1. Then there is a unique representation ρu,v : L∞ → L(E) such that for j = 1, 2, . . . we have ρu,v(sj) = uρ(sj) and ρu,v(tj) = ρ(tj)v. If u is an isometry and ρ is forward isometric or strongly forward isometric, then so is ρu,v. If kuk, kvk ≤ 1 and ρ is contractive on generators, then so is ρu,v. Proof. The proof is the same as that of Lemma 2.19, using the relations (1.4) and (1.5) in Definition 1.3. (cid:3) Examples 3.4 and 3.5 illustrate this construction. Lemma 2.21. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let E be a nonzero Banach space, and let ρ : A → L(E) be a representation. Using Notation 2.3 and the notation of Lemma 1.6(2), define a function ρ′ : A → L(E′) by ρ′(a) = ρ(a′)′. Then ρ′ is a representation of A, called the dual of ρ. Proof. This follows from Lemma 1.6(2). (cid:3) 3. Examples of representations In this section, we give a number of examples of representations of Leavitt alge- bras on Banach spaces. These examples illustrate some of the possible behavior of representations. We begin with basic examples of representations on lp. Example 3.1. Fix p ∈ [1, ∞]. We take lp = lp(Z>0). Let d ∈ {2, 3, 4, . . .}. Define functions f1, f2, . . . , fd : Z>0 → Z>0 by fd,j(n) = d(n − 1) + j for n ∈ Z>0. The functions fj are injective and have disjoint ranges whose union is Z>0. For j = 1, 2, . . . , d, define vd,j, wd,j ∈ L(lp) by, for ξ =(cid:0)ξ(1), ξ(2), . . .(cid:1) ∈ lp (vd,jξ)(n) =(ξ(cid:0)f −1 d,j (n)(cid:1) n ∈ ran(fd,j) and (wd,jξ)(n) = ξ(fd,j(n)). and n ∈ Z>0, 0 n 6∈ ran(fd,j) 14 N. CHRISTOPHER PHILLIPS Then vd,j is isometric, wd,j is contractive, and there is a unique representation ρ : Ld → L(lp) such that ρ(sj) = vd,j and ρ(tj) = wd,j for j = 1, 2, . . . , d. If we let δn ∈ lp be the sequence given by δn(n) = 1 and δn(m) = 0 for m 6= n, then vd,jδn = δfd,j (n) and wd,jδn =(δf −1 0 d,j (n) n ∈ ran(fd,j) n 6∈ ran(fd,j) for all n ∈ Z>0. For p ∈ [1, ∞), these formulas determine vd,j, wd,j ∈ L(lp) uniquely. The representation ρ is clearly contractive on generators (Definition 2.12(1)). One can check directly that ρ is strongly forward isometric (Definition 2.12(3)), with (Or use Lemma 7.7 below.) (cid:13)(cid:13)(cid:13)(cid:13)ρ(cid:18)Xd j=1 λjsj(cid:19) ξ(cid:13)(cid:13)(cid:13)(cid:13) = k(λ1, λ2, . . . , λd)kpkξk. Example 3.2. Let the notation be as in Example 3.1. Then there is a unique representation π : Cd → L(lp) such that π(sj) = vd+1, j and π(tj) = wd+1, j for j = 1, 2, . . . , d. Since Pd scend to a representation of Ld. j=1 wd+1, jvd+1, j 6= 1, this representation does not de- Example 3.3. Let p ∈ [1, ∞]. We obtain a representation ρ of L∞ on lp by the same method as in Example 3.1, taking v∞,j and w∞,j to come from the functions f∞,j(n) = 2jn − 2j−1 for j, n ∈ Z>0. Again, this representation is contractive on generators and strongly forward isometric. In Example 3.3, the ranges of the ρ(sj) span a dense subspace of lp, except when p = ∞. The following example shows that this need not be the case. Example 3.4. Let p ∈ [1, ∞]. Let ρ : L∞ → L(lp) be as in Example 3.3. Then there exists a unique representation π : L∞ → L(lp) such that for π(sj) = ρ(sj+1) and π(tj) = ρ(tj+1) for all j ∈ Z>0. Like ρ, this representation is contractive on generators and strongly forward isometric. This example can be obtained from ρ by the method of Lemma 2.20. Define u, v ∈ L(lp) by (uξ)(n) =(ξ(n/2) n is even n is odd 0 and (vξ)(n) = ξ(2n) for ξ ∈ lp and n ∈ Z>0. Using the notation of Example 3.1, for p 6= ∞ these are determined by uδn = δ2n and Then π = ρu,v. vδn =(δn/2 n is even n is odd. 0 CUNTZ ALGEBRAS ON Lp SPACES 15 Let ρ be as in Example 3.3, and let π be as in Example 3.4. We do not know whether the Banach algebras ρ(L∞) and π(L∞) are isometrically isomorphic, or even whether they are isomorphic. The representations of Example 3.3 and Example 3.4 are both very strongly tied to the structure of lp as a space of functions. (They are spatial in the sense of Definition 7.4(2) below.) The following example is less regular. Example 3.5. Let p ∈ [1, ∞]. Let the notation be as in Example 3.3 and Exam- ple 3.4. Define y ∈ L(lp) by (yξ)(n) =(ξ(2n) n is even ξ(2n) + ξ(n) n is odd. Thus, in the notation of Example 3.1, we have yδn =(δn/2 n is even n is odd. δn We have yu = 1, so, using Lemma 2.20, we obtain a representation σ = ρu,y of L∞. We have σ(sj ) = π(sj) for all j, but σ(t1) 6= π(t1). Indeed, π(t1)(δ1 + δ2) = δ1, but y(δ1 + δ2) = 2δ1, so σ(t1)(δ1 + δ2) = 2δ1. Thus the analog of Lemma 2.11 for L∞ is false. By restriction, it also fails for Cd. The representation σ is strongly forward isometric, since π is. For p = 1, it is easy to check that kyk ≤ 1, from which it follows that σ is contractive on generators. Now suppose p 6= 1. We saw above that σ(t1)(δ1 + δ2) = 2δ1. Taking 1 p = 0 when p = ∞, we have kσ(t1)(δ1 + δ2)k = 2 > 21/p = kδ1 + δ2k. So σ is not contractive on generators. In particular, a strongly forward isometric representation of L∞ need not be contractive on generators. We do not know whether this can happen for representations of Ld when d is finite. Example 3.6. The restrictions of the representations of Examples 3.3, 3.4, and 3.5 to the standard copy of Cd in L∞ (see Lemma 1.5) are representations of Cd on lp. In particular, strongly forward isometric does not imply contractive on generators for representations of Cd. Example 3.7. Let ρ : Ld → L(lp) be as in Example 3.1. One immediately checks that there is a representation σ : Ld → L(lp) such that for j = 1, 2, . . . , d, σ(sj ) = 1 2 ρ(sj) and σ(tj ) = 2ρ(tj). Then σ has the property (part of the definition of being strongly forward isometric, Definition 2.12(3)) that for every λ ∈ Cd, the element σ(sλ) is a scalar multiple of an isometry. Moreover, kσ(sj)k ≤ 1 for j = 1, 2, . . . , d. Also, the values of σ on sjtj ∈ Ld are the same as for the strongly forward isometric representation ρ. However, σ is not strongly forward isometric. The Banach algebras ρ(Ld) and σ(Ld) are isometrically isomorphic–in fact, they are equal. Taking u = 1 2 , and following Lemma 2.18, we can write σ = ρu. Example 3.8. Let ρ : Ld → L(lp) be as in Example 3.1, and let σ : Ld → L(lp) be as in Example 3.7. Let π = ρ ⊕p σ be as in Remark 2.15. Then kπ(sj)k = 1 for j = 1, 2, . . . , d, but π is not forward isometric and not contractive on generators. 16 N. CHRISTOPHER PHILLIPS It seems unlikely that the Banach algebras ρ(Ld) and π(Ld) are isomorphic. Taking v = diag(cid:0)1, 1 π = (ρ ⊕p ρ)v. 2(cid:1) and using the notation of Lemma 2.18, we can write Example 3.8 was obtained using Lemma 2.18 with a diagonal scalar matrix. In the following example, we instead use a nondiagonalizable scalar matrix. We do not know how the Banach algebras obtained as the closures of the ranges differ. Example 3.9. Let ρ : Ld → L(lp) be as in Example 3.1. Let ρ ⊕p ρ be as in Remark 2.15. Set Let τ be the representation τ = (ρ ⊕p ρ)w of Lemma 2.18. We have 0 1(cid:19) ∈ L(lp ⊕p lp). w =(cid:18)1 1 τ (sj ) =(cid:18)ρ(sj) ρ(sj) ρ(sj)(cid:19) and τ (tj ) =(cid:18)ρ(tj) −ρ(tj) ρ(tj) (cid:19) 0 0 for j = 1, 2, . . . , d. We now turn to a different kind of modification of our basic example, using automorphisms of Ld rather than Lemma 2.18. Example 3.10. Let d ∈ {2, 3, 4, . . .} and let p ∈ [1, ∞]. Set ω = exp(2πi/d). Let q be the conjugate exponent, that is, 1 q = 1. For k = 1, 2, . . . , d, define elements of Ld by p + 1 vk = d−1/p ωjksj and wk = d−1/q ω−jktj. dXj=1 dXj=1 (We take d−1/p = 1 when p = ∞ and d−1/q = 1 when p = 1.) It follows from Lemma 1.14 that wj vk = 1 for j = k and wjvk = 0 for j 6= k. A computation shows k=1 vkwk = 1. Therefore there is an endomorphism ϕ : Ld → Ld such that ϕ(sk) = vk and ϕ(tk) = wk for k = 1, 2, . . . , d. A computation shows that that Pd sk = d−1/q Therefore ϕ is bijective. ω−jkvj and tk = d−1/p dXj=1 ωjkwj . dXj=1 It follows that whenever ρ is a representation of Ld on a Banach space E, then ρ ◦ ϕ is also a representation. Moreover, (ρ ◦ ϕ)(Ld) = ρ(Ld). Now take ρ to be as in Example 3.1. We claim that ρ ◦ ϕ is contractive on gen- erators and strongly forward isometric. To prove this, it is convenient to introduce the matrix u =(cid:0)uj,k(cid:1)d j,k=1 = d−1/p ω ω2 ... ω2 ω4 ... ωd−1 ωd−2 1 1 . . . ωd−1 . . . ωd−2 . . . . . . . . . ... ω 1 1 1 ... 1 1 .  We then have ϕ(sk) = uj,ksj. dXj=1 CUNTZ ALGEBRAS ON Lp SPACES 17 Therefore, for λ = (λ1, λ2, . . . , λd) ∈ Cd, and following Definition 1.13, ϕ(sλ) = dXk=1 dXj=1 λkuj,ksj = dXj=1 (uλ)j sj = suλ. So, for λ ∈ Cd and ξ ∈ lp, Example 3.1 implies that k(ρ ◦ ϕ)(sλ)ξkp = kuλkpkξkp. In particular, taking λ = δj, for p 6= ∞ we get k(ρ ◦ ϕ)(sj )ξkp = kuδjkpkξkp =(cid:18)d−1Xd k=1(cid:12)(cid:12)ωjk(cid:12)(cid:12)p(cid:19)1/p kξkp = kξkp. One also checks that kuδjkp = 1 when p = ∞. We conclude that ρ ◦ ϕ is strongly forward isometric. It remains to prove that ρ ◦ ϕ is contractive on generators. Assume p 6= 1, ∞. Let ξ = (ξ(1), ξ(2), . . .) ∈ lp. Let wd,j be as in Example 3.1. Then = ξ(m)p = kξkp p. kwd,jξkp p = dXj=1 dXj=1 ∞Xn=1(cid:12)(cid:12)ξ(d(n − 1) + j)(cid:12)(cid:12)p m=1(cid:12)(cid:12)(cid:12)(cid:12)Xd p = d−p/qX∞ m=1(cid:18)Xd ≤ d−p/qX∞ dXj=1 ∞Xm=1 ω−jk(wd,jξ)(m)(cid:12)(cid:12)(cid:12)(cid:12) ω−jkq(cid:19)p/q(cid:18)Xd = d−p/qdp/q p = kξkp p. kwd,jξkp j=1 j=1 p Now, for k ∈ {1, 2, . . . , d}, we get, using Holder's inequality at the second step and ω−jk = 1 at the third step, k(ρ ◦ ϕ)(tk)ξkp j=1(cid:12)(cid:12)(wd,jξ)(m)(cid:12)(cid:12)p(cid:19) Thus ρ is contractive on generators. Easier calculations show that ρ is contractive on generators when p = 1 and p = ∞ as well. For p ∈ [1, ∞) \ {2}, we claim that, unlike the representation ρ of Example 3.18, we have in general k(ρ ◦ ϕ)(sλ)ξkp 6= kλkpkξkp. Since we have already seen that k(ρ◦ϕ)(sλ)ξkp = kuλkpkξkp, it suffices to find some λ ∈ Cd such that kuλkp 6= kλkp. For an explicit easily checked example, take d = 2, p = 3, and λ = (1, 2). Then kλkp p = 1 + 2p = 9, uλ = 2−1/3(1, 3), and kuλkp p = 1 2 (1 + 3p) = 14. For arbitrary p ∈ [1, ∞) \ {2} and arbitrary d, define σ : M p σ(a) = uau−1 for a ∈ M p d be the usual matrix unit. Then one checks that σ(e1,1) is not multiplication by any characteristic function. This vio- lates condition (5) in Theorem 7.2 below. Therefore σ is not isometric. So u is not isometric. d . Let ej,k ∈ M p d → L(cid:0)lp d(cid:1) by In the following examples derived from Example 3.10, we exclude p = 2 and p = ∞. When p = 2, we do not get new behavior for the closure of the image of Ld. We have not checked what happens when p = ∞. 18 N. CHRISTOPHER PHILLIPS Example 3.11. Let d ∈ {2, 3, 4, . . .} and let p ∈ [1, ∞) \ {2}. Let ρ be as in Example 3.1, and let ϕ and ρ ◦ ϕ be as in Example 3.10. Let π = ρ ⊕p (ρ ◦ ϕ), as in Remark 2.15. Then π is forward isometric and contractive on generators because ρ and ρ ◦ ϕ are. However, for p 6= 2, Example 3.8 and Example 3.1 show that there is λ ∈ Cd and ξ ∈ lp such that kρ(sλ)ξk 6= k(ρ ◦ ϕ)(sλ)ξk. Therefore π(sλ) is not a scalar multiple of an isometry, so π is not strongly forward isometric. We do not know whether the Banach algebras ρ(Ld) and π(Ld) are isomorphic. There are more complicated versions of Example 3.10 which also give represen- tations which are strongly forward isometric and contractive on generators. For example, one might split the generators into families and treat each family sepa- rately in the manner of Example 3.10. We give three special cases which are easy to write down and which we want for specific purposes. Example 3.12. Let d = 3, let p ∈ [1, ∞) \ {2}, and let ρ be as in Example 3.1 with this choice of d. There is a unique automorphism ψ of L3 such that ψ(s1) = s1, ψ(s2) = 2−1/p(s2 + s3), ψ(s3) = 2−1/p(s2 − s3), and ψ(t1) = t1, ψ(t2) = 2−1/q(t2 + t3), ψ(t3) = 2−1/q(t2 − t3). Then ρ ◦ ψ is strongly forward isometric, with k(ρ ◦ ψ)(sλ)ξkp =(cid:0)λ1p + 1 2 λ2 + λ3p + 1 2 λ2 − λ3p(cid:1)1/p kξkp, and contractive on generators. Moreover, σ = ρ ⊕p (ρ ◦ ψ), as in Remark 2.15, is forward isometric and contractive on generators, but not strongly forward isometric. Letting π be as in Example 3.11 with d = 3, we do not know whether the Banach algebras σ(L3) and π(L3) are isomorphic. We do note that π(L3) has an isometric automorphism which cyclically permutes the elements π(sj ), but there is no isometric automorphism of σ(L3) which does this. Possibly there isn't even any continuous automorphism of σ(L3) which does this. Example 3.13. Let p ∈ [1, ∞) \ {2}. Let d0, n ∈ {2, 3, . . .}. Set d = nd0. In Ld, call the standard generators sj,m and tj,m for j = 1, 2, . . . , d0 and m = 1, 2, . . . , n. By reasoning similar to that of Example 3.10, there is an automorphism α of Ld such that, for j = 1, 2, . . . , d0 and l = 1, 2, . . . , n, we have (3.1) α(sk,m) = d−1/p ωjksj,m and α(tk,m) = d−1/q ω−jktj,m. dXj=1 dXj=1 Take ρ to be as in Example 3.1. Then ρ ◦ α is a representation of Ld which is strongly forward isometric and contractive on generators, and for which one has (ρ ◦ α)(Ld) = ρ(Ld) but, in general, k(ρ ◦ α)(sλ)ξkp 6= kλkpkξkp. We presume that ρ ⊕p (ρ ◦ α) is essentially different from both ρ and the repre- sentation ρ ⊕p (ρ ◦ ϕ) of Example 3.11. However, we do not have a proof of anything like this. Example 3.14. Let the notation be as in Example 3.13. Define a homomorphism β : Cd0 → Ld by β(sj) = sj,1 and β(tj ) = tj,1 for j = 1, 2, . . . , d0. Then ρ ◦ β is a representation of Cd0 which is strongly forward isometric and contractive on generators, and for which one has k(ρ ◦ β)(sλ)ξkp = kλkpkξkp for λ ∈ Cd0 and ξ ∈ lp. CUNTZ ALGEBRAS ON Lp SPACES 19 The next example is the analog of Example 3.13 for L∞. Example 3.15. Let p ∈ [1, ∞) \ {2} and let d0 ∈ {2, 3, . . .}. In L∞, call the standard generators sj,m and tj,m for j = 1, 2, . . . , d0 and m ∈ Z>0. Then there is an automorphism α of L∞ defined by the formula (3.1), but now for j = 1, 2, . . . , d0 and m ∈ Z>0. Take ρ to be as in Example 3.3. Then ρ ◦ α is a representation of L∞ which is strongly forward isometric and contractive on generators, and for which one has (ρ ◦ α)(L∞) = ρ(L∞) but, in general, k(ρ ◦ α)(sλ)ξkp 6= kλkpkξkp. We can then form the direct sum representation π = ρ⊕p(ρ◦α) as in Remark 2.15. We do not know whether π(L∞) is isomorphic to ρ(L∞) as a Banach algebra. Example 3.16. Let the notation be as in Example 3.15. Define a homomorphism β from L∞ (with conventionally named generators) to L∞ (with generators named as in Example 3.15) by β(sj) = sj,1 and β(tj ) = tj,1 for j = 1, 2, . . . , d0. Then ρ ◦ β is a representation of L∞ which is strongly forward isometric and contractive on generators, and for which one has k(ρ ◦ β)(sλ)ξkp = kλkpkξkp for λ ∈ C∞ and ξ ∈ lp. We do not know whether (ρ ◦ β)(L∞) is isomorphic to ρ(L∞) as a Banach algebra. Remark 3.17. Example 3.10 is based on the Fourier transform from functions on Zd to functions on Zd. We have not investigated the possibility of using other finite abelian groups. We finish this section with basic examples of representations on Lp([0, 1]). Example 3.18. Let d ∈ {2, 3, 4, . . .} and let p ∈ [1, ∞]. For j = 1, 2, . . . , d, define a function gj : [0, 1] →(cid:20) j − 1 d , j d(cid:21) by gj(x) = d−1(j + x − 1). Then we claim that there is a unique representation ρ : Ld → L(Lp([0, 1])) such that for j = 1, 2, . . . , d, x ∈ [0, 1], and ξ ∈ Lp([0, 1]) we have and j (cid:0)ρ(sj)ξ(cid:1)(x) =(d1/pξ(g−1 (x)) x ∈(cid:2) j−1 (cid:0)ρ(tj)ξ(cid:1)(x) = d−1/pξ(gj(x)). otherwise 0 d , j d(cid:3) The proof of the claim is a straightforward verification that the proposed oper- ators ρ(sj) and ρ(tj ) satisfy the defining relations for Ld in Definition 1.1. It is easy to check that ρ is contractive on generators and is forward isometric. In fact, ρ is strongly forward isometric. This follows from general theory. (See Lemma 7.9 below.) But it is also easily checked, using Remark 2.7, that for λ ∈ Cd and ξ ∈ Lp([0, 1]), we have kρ(sλ)ξkp = kλkpkξkp. Example 3.19. Let p ∈ [1, ∞]. We give an example of a representation of L∞ on Lp([0, 1]). In the following, if p = ∞ then expressions with p in the denominator are taken to be zero. For j ∈ Z>0, define a function fj : [0, 1] →(cid:20) 1 2j , 1 2j−1(cid:21) and j 0 (cid:0)ρ(sj)ξ(cid:1)(x) =(2j/pξ(cid:0)f −1 (x)(cid:1) x ∈(cid:2) 1 (cid:0)ρ(tj)ξ(cid:1)(x) = 2−j/pξ(fj(x)). 2j , otherwise 1 2j−1(cid:3) 20 N. CHRISTOPHER PHILLIPS by fj(x) = 2−j(1+x). Then we claim that there is a unique representation ρ : L∞ → L(Lp([0, 1])) such that for j ∈ Z>0, x ∈ [0, 1], and ξ ∈ Lp([0, 1]) we have The proof of the claim is a straightforward verification that the proposed operators ρ(sj) and ρ(tj) satisfy the defining relations for L∞ in Definition 1.3. One can check that ρ is contractive on generators and forward isometric. (We will see in Lemma 7.9 below that it is in fact strongly forward isometric.) 4. Boolean σ-algebras In this section and the next, we describe the background for the characterization, due to Lamperti, of isometries on Lp(X, µ). Parts of this material can be found in Section 1 of Chapter X of [13] and in Lamperti's paper [18], but these references contain only enough to state and prove the characterization theorem, not enough to make serious use of it. A somewhat more systematic presentation can be found in Chapter 15 of [24], especially Section 2, and we use the terminology from there, but we need more than is there, and the form in which it is presented there makes citation of specific results difficult. For this section, on abstract Boolean σ-algebras, we follow [16] and just state the basic results. (However, we use notation more suggestive of unions and intersections than the notation of [16]: our E ∨ F is E + F there, our E ∧ F is E · F there, and our E′ is −E there.) Definition 4.1 (Definition 1.1 of [16]). A Boolean algebra is a set B with two commutative associative binary operations (E, F ) 7→ E ∨ F and (E, F ) 7→ E ∧ F, a unary operation E 7→ E′, and distinguished elements 0 and 1, satisfying the following for all E, F, G ∈ B: (1) E ∨ (E ∧ F ) = E and E ∧ (E ∨ F ) = E. (2) E ∧ (F ∨ G) = (E ∧ F ) ∨ (E ∧ G) and E ∨ (F ∧ G) = (E ∨ F ) ∧ (E ∨ G). (3) E ∨ E′ = 1 and E ∧ E′ = 0. Subalgebras and homomorphisms of Boolean algebras have the obvious mean- ings. (See Definitions 1.7 and 1.3 of [16].) Example 4.2. The standard example is the power set P(X) of a set X, with E ∨ F = E ∪ F, E ∧ F = E ∩ F, E′ = X \ E, 0 = ∅. and 1 = X. We therefore refer to the operations in a Boolean algebra as union, intersection, and complementation. Proofs that the axioms imply the other expected properties can be found in Section 1.5 of [16]. However, for the purposes of finite algebraic manipulations, it is easier to rely on the following theorem, which implies that all finite identities which hold among sets also hold in any Boolean algebra. Theorem 4.3 (Theorem 2.1 of [16]). Let B be a Boolean algebra. Then there exists a set X and an isomorphism of B with a Boolean subalgebra of P(X). CUNTZ ALGEBRAS ON Lp SPACES 21 Definition 4.4. Let B be a Boolean algebra, and let E, F ∈ B. We define E ≤ F to mean E ∧ F = E. We say that E and F are disjoint if E ∧ F = 0. We define the symmetric difference of E and F to be E △ F = (E ∧ F ′) ∨ (E′ ∧ F ). In P(X), disjointness means that the intersection is empty, E ≤ F means E ⊂ F, and symmetric difference has its usual meaning. Thus, by Theorem 4.3, the relation of Definition 4.4 is a partial order on B, in which 0 is the least element and 1 is the greatest element. In particular, if E ≤ F and F ≤ E, then E = F. Definition 4.5 (Section 15.2 of [24]). A Boolean σ-algebra is a Boolean algebra B in which whenever E1, E2, . . . ∈ B, then there is a least element E ∈ B such that En ≤ E for all n ∈ Z>0. In Definition 1.28 of [16], a Boolean σ-algebra is called a σ-complete Boolean algebra. (It is also required that greatest lower bounds of countable collections exist, but this follows from Definition 4.5 by complementation.) Definition 4.6. Let B be a Boolean σ-algebra. The element E in Definition 4.6 n=1 En = n=1 En. We call it the union of the En. We further defineV∞ n)′ , and call it the intersection of the En. n=1 E′ is denotedW∞ (W∞ The operationsW∞ in [24], but is proved in [16].) n=1 En andV∞ n=1 En behave as expected. (This is not proved Lemma 4.7. Let B be a Boolean σ-algebra. F ∨ En = (F ∨ En) (F ∧ En), F ∨ En = (F ∨ En) (F ∧ En). (1) Let E1, E2, . . . ∈ B. Then En!′ = E′ n ∞^n=1 (2) Let E1, E2, . . . , F ∈ B. Then ∞_n=1 ∞_n=1 ∞^n=1 ∞_n=1 and ∞_m=1 ∞_n=1 ∞^n=1 ∞_n=1 ∞_m=1 (3) Let Em,n ∈ B for m, n ∈ Z>0. Then En!′ = E′ n. and F ∧ and ∞^n=1 ∞_n=1 ∞^n=1 ∞^n=1 ∞^m=1 and F ∧ En = En = ∞_n=1 ∞_n=1 ∞^n=1 ∞^m=1 ∞^n=1 Em,n = En,m and Em,n = En,m. Proof. These statements all follow easily from the various parts of Lemma 1.33 of [16], or from their duals (stated afterwards). (cid:3) Example 4.8. Let X be a set. Then a σ-algebra of subsets of X is a Boolean σ-algebra. We introduce σ-homomorphisms and σ-ideals. In [16], they are called σ-complete homomorphisms (Definition 5.1 of [16]) and σ-complete ideals (Definition 5.19 of [16]). 22 N. CHRISTOPHER PHILLIPS Definition 4.9. Let B and C be Boolean σ-algebras. A σ-homomorphism from B to C is a function S : B → C which is a homomorphism of Boolean algebras in the obvious sense, and moreover such that whenever E1, E2, . . . ∈ B then S ∞_n=1 En! = ∞_n=1 S(En). Lemma 4.10. The σ-homomorphisms of Definition 4.9 have the following proper- ties. (1) The composition of two σ-homomorphisms is a σ-homomorphism. (2) A σ-homomorphism preserves order and disjointness (as in Definition 4.4). (3) A σ-homomorphism S is injective if and only if S(E) = 0 implies E = 0. Proof. The first part is obvious, and the second is easy. The nontrivial direction of the third part is proved by considering symmetric differences. (See Lemma 5.3 of [16].) (cid:3) Definition 4.11. Let B be a Boolean σ-algebra. A σ-ideal in B is a subset N ⊂ B which is closed under countable unions, such that 0 ∈ N , and such that whenever E ∈ B and F ∈ N satisfy E ⊂ F, then E ∈ N . The standard example is as follows. Example 4.12. Let X be a set, and let B be a σ-algebra on X. Let µ be a measure with domain B. Then N (µ) = {E ∈ B : µ(E) = 0} is a σ-ideal in B. Definition 4.13. Let B be a Boolean σ-algebra. If N ⊂ B is a σ-ideal, we define E ∼ F to mean E △ F ∈ N . We define B/N to be the quotient set of B by ∼ . If E ∈ B, we write [E] for its image in B/N . Lemma 4.14. Let the notation be as in Definition 4.13. Then ∼ is an equivalence relation, the obvious induced operations on B/N are well defined and make B/N a Boolean σ-algebra, and E 7→ [E] is a surjective σ-homomorphism. Proof. This is outlined in Section 15.2 of [24], and is contained in Lemma 5.22 of [16] and the remark after its proof. (cid:3) We finish this section with a lemma which will be needed later. 1 , E(0) Lemma 4.15. Let B be a Boolean σ-algebra, and let N ⊂ B be a σ-ideal. Suppose E1, E2, . . . ∈ B/N are pairwise disjoint (Definition 4.4). Then there exist disjoint elements E(0) 2 , . . . ∈ B such that(cid:2)E(0) n (cid:3) = En for all n ∈ Z>0. Proof. We first show that if E, F ∈ B/N are disjoint, then there exist disjoint E0, F0 ∈ B such that [E0] = E and F0 = F. Choose any E0 ∈ B such that [E0] = E and any K ∈ B such that [K] = F. Then [K ∧ E0] = 0, so we can take F0 = K ∧ E′ 0. Now we prove the statement. For m 6= n choose disjoint elements Sm,n, Tm,n ∈ B such that [Sm,n] = Em and [Tm,n] = En. Then take E(0) n = ^k∈Z>0\{n} (Sn,k ∧ Tk,n). This completes the proof. (cid:3) CUNTZ ALGEBRAS ON Lp SPACES 23 5. Measurable set transformations In this section, we consider measure spaces (X, B, µ) and (Y, C, ν), and a suitable σ-homomorphism S : B/N (µ) → C/N (ν). We describe how to use S to produce maps on measurable functions mod equality almost everywhere and on measures. The idea is not new; it can be found in [18] and (for functions, under stronger hypotheses) in Chapter X of [13]. However, we need much more than can be found in these references. First, we give some definitions and notation which will be frequently used later. Definition 5.1. Let (X, B, µ) be a measure space. We denote by L0(X, µ) the vector space of all complex valued measurable functions on X, mod the functions which vanish almost everywhere. We follow the usual convention of treating ele- ments of L0(X, µ) as functions when convenient. If E ∈ B, we denote by χE the characteristic function of E, which is a well defined element of L0(X, µ). Definition 5.2. Let (X, B, µ) be a measure space, and let N (µ) be as in Exam- ple 4.12. For ξ ∈ L0(X, µ), we define the support of ξ to be the element of B/N (µ) given as follows. Choose any actual function ξ0 : X → C whose class in L0(X, µ) is ξ, and set Further, for E ∈ B/N (µ), define χE ∈ L0(X, µ) to be the class of χE0 for any supp(ξ) =(cid:2)(cid:8)x ∈ X : ξ0(x) 6= 0(cid:9)(cid:3). E0 ∈ B with [E0] = E. Definition 5.2 is a natural kind of strengthening of Definition 2.6. Remark 5.3. In Definition 5.2, it is easy to check that supp(ξ) and χE are well defined. Clearly ξ is supported in E, in the sense of Definition 2.6, if and only if [E] ≥ supp(ξ). The following definition and constructions based on it are adapted from the beginning of Section 1 in Chapter X of [13]. Definition 5.4. Let (X, B, µ) and (Y, C, ν) be measure spaces, and let N (µ) and N (ν) be as in Example 4.12. A measurable set transformation from (X, B, µ) to (Y, C, ν) is a σ-homomorphism (in the sense of Definition 4.9) S : B/N (µ) → C/N (ν). By abuse of notation, we write S : (X, B, µ) → (Y, C, ν). For E ∈ B, we also write, by abuse of notation, S(E) for some choice of F ∈ C such that [F ] = S([E]). Injectivity and surjectivity always refer to properties of the map S : B/N (µ) → C/N (ν). We denote by ranY (S) the collection of all subsets F ∈ C such that [F ] = S([E]) for some E ∈ B. In [13], a set transformation is taken to be a multivalued map from B to C, with possible values differing only up to a set of measure zero, and which preserves the appropriate set operations, that is, which defines a σ-homomorphism from B/N (µ) to C/N (ν). Also, (X, B, µ) = (Y, C, ν), and the map is required to be measure preserving. In our situation, if S is injective, then at least the type of transformation in [13] sends sets of nonzero measure to sets of nonzero measure. At the beginning of Section 3 of [18], what we call an injective σ-homomorphism is called a regular set isomorphism. The definition specifies a map of sets modulo null sets, although without formally defining an appropriate domain. It omits the requirement (implicit above) that S(X) = Y, but this is easily restored by replacing Y by S(X). We do not use the term "regular set isomorphism" because 24 N. CHRISTOPHER PHILLIPS such maps need not be surjective, and we need to consider cases in which they are not. Also, we are more careful with the formalism because we need to use such maps systematically. Lemma 5.5. Let the notation be as in Definition 5.4. Then ranY (S) is a sub-σ- algebra of B. Proof. The proof is easy, and is omitted. (cid:3) Proposition 5.6. Let (X, B, µ) and (Y, C, ν) be measure spaces. Let S : (X, B, µ) → (Y, C, ν) be a measurable set transformation (Definition 5.4). Then (following the notation of Definition 5.1) there is a unique linear map S∗ : L0(X, µ) → L0(Y, ν) such that: (1) S∗(χE) = χS(E) for all E ∈ B/N (µ). (2) Whenever (ξn)n∈Z>0 is a sequence of measurable functions on X which con- verges pointwise almost everywhere [µ] to ξ, then S∗(ξn) → S∗(ξ) pointwise almost everywhere [ν]. Moreover: (3) Let n ∈ Z>0, let f : Cn → C be continuous, and let ξ1, ξ2, . . . , ξn ∈ L0(X, µ). Set η(x) = f(cid:0)ξ1(x), ξ2(x), . . . , ξn(x)(cid:1) for x ∈ X. Then S∗(η)(y) = f(cid:0)S∗(ξ1)(y), S∗(ξ2)(y), . . . , S∗(ξn)(y)(cid:1) for almost all y ∈ Y. (In particular, S∗ preserves products and preserves arbitrary positive powers of the absolute value of a function.) (5) S∗ is injective if and only if S is injective. (4) The range of S∗ is L0(cid:0)Y, νran(S)(cid:1). (6) If ξ ∈ L0(X, µ) and B ⊂ C is a Borel set, then S([ξ−1(B)]) =(cid:2)S∗(ξ)−1(B)(cid:3). (7) Let (Z, D, λ) be another measure space, and let T be a measurable set transformation from (Y, C, ν) to (Z, D, λ). Then (T ◦ S)∗ = T∗ ◦ S∗. For the proof, we use the following well known lemma. Lemma 5.7. Let (X, B, µ) be a measure space, and let (Eα)α∈Q be a family of measurable sets such that Eα ⊂ Eβ whenever α ≤ β, and such that \α∈Q Eα = ∅ and [α∈Q Eα = X. Then there exists a unique measurable function ξ : X → R such that for all α ∈ Q we have {x ∈ X : ξ(x) < α} ⊂ Eα ⊂ {x ∈ X : ξ(x) ≤ α}. Moreover, for α ∈ R we have and {x ∈ X : ξ(x) < α} = [β∈Q∩(−∞,α) {x ∈ X : ξ(x) > α} = [β∈Q∩(α,∞) Eβ (X \ Eβ). Proof. The first part of the statement is Lemma 9 in Chapter 11 of [24]. The last two equations follow easily. (cid:3) (5.3) (5.2) Set α (cid:3) = S([Eα]). (cid:2)F (0) α (cid:1) ∪ \α β ∩(cid:0)Y \ F (0) D0 =[α<β α ! ∪ Y \[α Lemma 5.7 with(cid:0)Y \ D0, CY \D0, νY \D0(cid:1) in place of (X, B, µ). Then ν(D0) = 0 and the sets F (0) F (0) F (0) Choose any set D ∈ C such that D0 ⊂ D and ν(D) = 0. For α ∈ Q define α ! . F (0) α ∩ (Y \ D0) ⊂ Y satisfy the the hypotheses of CUNTZ ALGEBRAS ON Lp SPACES 25 Proof of Proposition 5.6. We prove uniqueness. Let T1, T2 : L0(X, µ) → L0(Y, ν) be linear and satisfy (1) and (2). It follows from linearity and (1) that whenever ξ ∈ L0(X, µ) is a simple function, we have T1(ξ) = T2(ξ) almost everywhere [ν]. Since every measurable function is a pointwise limit of simple functions, this conclusion in fact holds for all ξ ∈ L0(X, µ). We now prove existence. We follow the construction on page 454 of [13], which is done under stronger hypotheses. The verification of linearity will be done as a special case of (3). First let ξ ∈ L0(X, µ) be real valued. For α ∈ Q define (5.1) Eα = {x ∈ X : ξ(x) < α}. Then the sets Eα satisfy the hypotheses of Lemma 5.7. For each α ∈ Q, choose a set F (0) α ∈ C such that (5.4) Fα =(F (0) α ∪ D α > 0 F (0) α ∩ (Y \ D) α ≤ 0. Then the sets Fα ⊂ Y satisfy the the hypotheses of Lemma 5.7 with (Y, C, ν) in place of (X, B, µ). So Lemma 5.7 provides a measurable function η : Y → R such that for all α ∈ Q, we have (5.5) {y ∈ Y : η(y) < α} ⊂ Fα ⊂ {y ∈ Y : η(y) ≤ α}. We define S∗(ξ) = η. We claim that, up to equality almost everywhere [ν], the function η does not α that F (0) α in place of F (0) our construction, with the choice α ∈ C such that (cid:2)eF (0) depend on the choices made in its construction. First, if we replace D with eD, and call the new function eη, then eη = η off D ∪ eD, and ν(cid:0)D ∪ eD(cid:1) = 0. Now suppose α (cid:3) = S([Eα]). is replaced by some other set eF (0) Then ν(cid:0)F (0) α (cid:1) = 0 for all α ∈ Q. Let D0 be as in (5.3), and define eD0 α △ eF (0) analogously, using eF (0) α . Let η andeη be the functions resulting from D = eD = D0 ∪ eD0 ∪ [α∈Q(cid:0)F (0) α ∩ (Y \ D) = eF (0) α ∪ D = eF (0) for all α ∈ Q. It follows that η =eη. This completes the proof of the claim. α (cid:1). α △ eF (0) for all α ∈ Q, so also α ∩ (Y \ D) α ∪ D We have F (0) F (0) nXk=1 nXk=1 26 N. CHRISTOPHER PHILLIPS We next claim that if ξ = Pn the abuse of notation from Definition 5.4) we have S∗(ξ) = γkχS(Bk). k=1 γkχBk , with B1, B2, . . . , Bn ∈ B, then (using To prove this, first suppose that B1, B2, . . . , Bn are disjoint. Using Lemma 4.15, choose disjoint sets C1, C2, . . . , Cn ∈ C such that [Ck] = S([Bk]) for k = 1, 2, . . . , n. For α ∈ Q choose F (0) empty. Taking D = ∅ gives α = Sγk<α Ck. Then the set D0 in the construction of η is S∗(ξ) = γkχCk , as desired. The general case is easily reduced to the disjoint case by using instead of the sets Bk all possible nonempty intersections E1 ∩ E2 ∩ · · · ∩ En in which Ek is either Bk or X \ Bk. This proves the claim. It follows immediately that (3) holds when ξ1, ξ2, . . . , ξn are real simple functions and f is a continuous function from Rn to R. We now prove (2). Let (ξn)n∈Z>0 be a sequence of real measurable functions on X, and suppose that ξn(x) → ξ(x) almost everywhere [µ]. Changing the ξn and ξ on a set of measure zero, we may assume that ξn(x) → ξ(x) for all x ∈ X. For n ∈ Z>0 and α ∈ Q, let En,α, F (0) n be the sets of (5.1), (5.2), and (5.3) for the construction of S∗(ξn), and let Eα, F (0) α , and D0 be the corresponding sets for ξ. Let α, β ∈ Q satisfy α > β. Since ξ(x) < α implies lim supn→∞ ξn(x) < α and ξ(x) ≥ α implies lim inf n→∞ ξn(x) > β, we get n,α, and D(0) Set and Set Eα ⊂ ∞[n=1 ∞\m=n Em,α Bα = F (0) Cα,β =(cid:0)Y \ F (0) D = D0 ∪ ∞[n=1 (X \ Em,β). F (0) and X \ Eα ⊂ ∞\m=n ∞[n=1 m,α! α ∩ Y \ ∞\m=n ∞[n=1 α (cid:1) ∩ Y \ m,β(cid:1)! . ∞\m=n(cid:0)Y \ F (0) ∞[n=1 n ! ∪[α∈Q Bα ∪[α>β Cα,β . D(0) Since S is a σ-homomorphism, ν(Bα) = 0 and ν(Cα,β) = 0. Then ν(D) = 0. Using this choice of D, define sets Fn,α as in (5.4) for ξn and Fα for ξ, and let ηn and η be the representatives the construction gives for S∗(ξn) and S∗(ξ). Let y ∈ Y \ D. We claim that ηn(y) → η(y). This will complete the proof of (2) for real functions. To prove the claim, let ε > 0, and choose α, β ∈ Q such that η(y) − ε < β < η(y) < α < η(y) + ε. CUNTZ ALGEBRAS ON Lp SPACES 27 Then y ∈ Fα by (5.5). Since y 6∈ Bα, we have y ∈ Fm,α, ∞[n=1 ∞\m=n so (5.5) for ηn implies that there is n1 ∈ Z>0 such that for all m ≥ n1 we have ηm(y) ≤ α. Then also ηm(y) < η(y) + ε. Furthermore, (5.5) implies y 6∈ Fβ, and y 6∈ Cα,β, so y ∈ (Y \ Fm,β). ∞[n=1 ∞\m=n Using (5.5) for ηn, we get n2 ∈ Z>0 such that for all m ≥ n2 we have ηm(y) ≥ β > η(y) − ε. We have thus shown that ηn(y) → η(y), as desired. It now follows that (3) holds whenever ξ1, ξ2, . . . , ξn are real measurable functions and f is a continuous function from Rn to R, because ξ1, ξ2, . . . , ξn are pointwise limits of real simple functions. We define S∗ on complex functions ξ by S∗(ξ) = S∗(Re(ξ))+iS∗(Im(ξ)). It is easy to check that S∗ satisfies (1) and (2). For (3), first let f : Cn → R be continuous. Define g : R2n → R by g(s1, t1, s2, t2, . . . , sn, tn) = f (s1 + it1, s2 + it2, . . . , sn + itn) Then one checks that for s1, t1, s2, t2, . . . , sn, tn ∈ R. Set η(x) = f(cid:0)ξ1(x), ξ2(x), . . . , ξn(x)(cid:1) for x ∈ X. S∗(η)(y) = g(cid:0)S∗(Re(ξ1))(y), S∗(Im(ξ1))(y), . . . , S∗(Re(ξn))(y), S∗(Im(ξn))(y)(cid:1) = f(cid:0)S∗(ξ1)(y), S∗(ξ2)(y), . . . , S∗(ξn)(y)(cid:1) for y ∈ Y, as desired. The extension to continuous functions f : Cn → C is now easy. We now prove (4). It suffices to show that a real valued measurable function Choose E(0) Gα = {y ∈ Y : λ(y) < α} ∈ ran(S). λ ∈ L0(cid:0)Y, νran(S)(cid:1) is in the range of S∗. For α ∈ Q, define α  , α ∈ B such that S(cid:0)(cid:2)E(0) D =X \ [α∈Q Eα =(D ∪Sβ<α E(0) α (cid:3)(cid:1) = [Gα]. Define α  ∪\α∈Q (X \ D) ∩Sβ<α E(0) α > 0 α ≤ 0. and set E(0) E(0) β β These sets satisfy the hypotheses of Lemma 5.7. Let ξ be the function obtained second part of Lemma 5.7 implies that Eα = {x ∈ X : ξ(x) < α} for all α ∈ Q. from Lemma 5.7. One easily checks that Eα = Sβ<α Eβ for all α ∈ Q, so the Since S is a σ-homomorphism and Gα =Sβ<α Gβ for all α ≤ β, one easily checks that S([Eα]) = [Gα] for all α ∈ Q. In the construction of S(ξ) at the beginning of the proof, we may therefore take F (0) α = Gα for all α ∈ Q, giving D0 = ∅, and then we may take D = ∅. Thus Fα = Gα for all α ∈ Q. So η = S∗(ξ) satisfies (5.5) for all α ∈ Q. Such a function is unique by Lemma 5.7, and λ is such a function, so S∗(ξ) = λ. This completes the proof of (4). and set Gα ∪\α∈Q H =Y \ [α∈Q α =(H ∪Sβ<α Gβ Gα , (Y \ H) ∩Sβ<α Gβ α ≤ 0. α > 0 F (0) 28 N. CHRISTOPHER PHILLIPS We next prove (6). Suppose first that ξ is real valued. For α ∈ Q, let Eα = {x ∈ X : ξ(x) < α} (as in (5.1)). Choose Gα ⊂ Y such that [Gα] = S([Eα]) (as in (5.2), where the set is called F (0) α ). Set in the construction of α . One easily checks that for all α ∈ Q. The second part of Lemma 5.7 therefore implies S([Eα]) for all α ∈ Q. Therefore we may use the sets F (0) α S∗(ξ). This gives D0 = ∅. Take D = ∅. So Fα = F (0) F (0) that Fα = {y ∈ Y : S∗(ξ)(y) < α} for all α ∈ Q. We have verified that Since Eα =Sβ<α Eβ for all α ∈ Q and S is a σ-homomorphism, we have(cid:2)F (0) α (cid:3) = α =Sβ<α F (0) for all α ∈ Q. Since the collection(cid:8)(−∞, α) : α ∈ Q(cid:9) generates the Borel σ-algebra S(cid:0)(cid:2){x ∈ X : ξ(x) < α}(cid:3)(cid:1) =(cid:2){y ∈ Y : S∗(ξ)(y) < α}(cid:3) and S is a σ-homomorphism, we have (6) for real valued ξ and all Borel subsets of R. β By considering real and imaginary parts separately, one sees that (6) holds for complex ξ whenever B is the product of two Borel subsets of R. Such products generate the Borel subsets of C, so (6) holds for arbitrary B. For (5), first suppose that S is not injective. Then there is E ∈ B such that µ(E) 6= 0 but S([E]) = [∅]. So χE 6= 0 but S∗(χE) = 0. On the other hand, suppose S∗ is not injective. Then there is a nonzero ξ ∈ L0(X, µ) such that S∗(ξ) = 0. By considering the positive or negative part of the real or imaginary part of ξ, we may assume that ξ is nonnegative. Since ξ 6= 0, there is ε > 0 such that the set E = {x ∈ X : ξ(x) > ε} satisfies µ(E) 6= 0. Using part (6) at the first step, we get Therefore S is not injective. [S(E)] =(cid:2){y ∈ Y : S∗(ξ) > ε}(cid:3) = [∅]. Part (7) follows from uniqueness, since T∗ ◦ S∗ and (T ◦ S)∗ are both linear and (cid:3) satisfy (1) and (2). Corollary 5.8. Let the notation be as in Proposition 5.6, and assume in addition that µ and ν are σ-finite. Then the following are equivalent: (1) S is surjective. (2) S∗ : L0(X, µ) → L0(Y, ν) is surjective. (4) For every F ∈ C there are disjoint sets E1, E2, . . . ∈ B with finite measure (3) The range S∗(cid:0)L0(X, µ)(cid:1) contains χF for every F ∈ C such that ν(F ) < ∞. such that, with E =S∞ Proof. That (1) implies (2) is Proposition 5.6(4). That (2) implies (3) is trivial. n=1 En, we have S∗(χE) = χF . We prove that (3) implies (4). Since ν is σ-finite and S∗ preserves pointwise almost everywhere limits (Proposition 5.6(2)), it suffices to consider sets F with CUNTZ ALGEBRAS ON Lp SPACES 29 ν(F ) < ∞. Choose ξ ∈ L0(X, µ) such that S∗(ξ) = χF . Set E = {x ∈ X : ξ(x) = 1}. Then S([E]) = [F ] by Proposition 5.6(6). Now use σ-finiteness of µ to write n=1 with µ(En) < ∞ for all n. Finally, we prove that (4) implies (1). Let F ∈ C. Then (4) implies that there is ξ ∈ L0(X, µ) such that S∗(ξ) = χF . Set E = {x ∈ X : ξ(x) = 1}. Then E ∈ B and S([E]) = [F ] by Proposition 5.6(6). (cid:3) E =`∞ Lemma 5.9. Let the notation be as in Definition 5.4. Let λ be a measure on C such that λ << ν. Then there exists a unique measure S∗(λ) on B such that whenever E ∈ B and F ∈ C satisfy [F ] = S([E]), then S∗(λ)(E) = λ(F ). Moreover: (1) S∗(λ) << µ. (2) If σ is another measure on C and σ << λ, then S∗(σ) << S∗(λ). (3) If S∗(λ) is σ-finite, then λ is σ-finite. (4) For every nonnegative function ξ ∈ L0(X, µ), and for every function ξ ∈ L1(X, S∗(λ)), we have ZX ξ d(S∗(λ)) =ZY S∗(ξ) dλ. (5) If (X, B, µ) = (Y, C, ν) and S is the identity map, then S∗(λ) = λ. (6) Let (Z, D, ρ) be another measure space, and let T be a measurable set transformation from (Y, C, ν) to (Z, D, ρ). Suppose σ is a measure on D such that σ << ρ. Then (T ◦ S)∗(σ) = S∗(T ∗(σ)). Proof. Uniqueness of S∗(λ) is obvious. For existence, we have to prove that S∗(λ) is well defined, satisfies S∗(∅) = 0, and is countably additive. The first follows from λ << ν, the second is immediate, and the third follows from the first together with Lemma 4.15 and Lemma 4.10(2). Parts (1), (2), (5), and (6) are clear. To prove (3), write X =S∞ Fn ∈ C such that S([En]) = [Fn]. Then λ(Fn) < ∞. Therefore F =S∞ n=1 En with S∗(λ)(En) < ∞ for all n ∈ Z>0. Choose n=1 Fn is a subset of Y on which λ is σ-finite. Also [Y \ F ] = [∅]. Since λ << ν, this implies that λ(Y \ F ) = 0. Therefore λ is σ-finite. It remains to verify (4). It suffices to do so when ξ is nonnegative. The result is immediate for simple functions. Choose a sequence (ξn)n∈Z>0 of nonnegative simple functions such that ξn → ξ pointwise and ξ1 ≤ ξ2 ≤ · · · pointwise. Propo- sition 5.6(2) implies that S∗(ξn) → S∗(ξ) pointwise almost everywhere [ν], and Proposition 5.6(6) implies that S∗(ξ1) ≤ S∗(ξ2) ≤ · · · pointwise almost every- where [ν]. Using λ << ν, S∗(λ) << µ, and the Monotone Convergence Theorem (twice), we get ZX n→∞ZX ξ dS∗(λ) = lim ξn dS∗(λ) = lim n→∞ZY S∗(ξn) dλ =ZY S∗(ξ) dλ. (cid:3) This completes the proof. The converse to Lemma 5.9(3) is false. Take X = Z, take µ to be counting measure, take Y = Z × Z, take ν to be counting measure, take λ = ν, and take S(E) = E × Z for E ⊂ X. We really need to push measures forwards rather than pull them back. Definition 5.10. Let the notation be as in Definition 5.4, and assume that S is injective. Let λ be a measure defined on B which is absolutely continuous with 30 N. CHRISTOPHER PHILLIPS respect to µ. Then we define S∗(λ) to be the measure (S−1)∗(λ) on the σ-algebra ran(S). The measure S∗(µ) is called µ∗ in the statement of Theorem 3.1 of [18]. We need some notation. Notation 5.11. Let (X, B, µ) be a measure space, and let E ∈ B. We denote by BE the σ-algebra on E consisting of all F ∈ B such that F ⊂ E. We call it the restriction of B. If λ is a measure defined on a σ-algebra B, and C ⊂ B is a σ-algebra on a set E ∈ B, we write λC for the restriction of λ. If C = BE, we just write λE. When no confusion can arise, we often just write λ. For example, for a measure space (X, B, µ), we often write Lp(E, µ) rather than Lp(E, µE). Also, we identify without comment Lp(E, µ) as the subspace of Lp(X, µ) consisting of those functions which vanish off E. Lemma 5.12. In the situation of Definition 5.10, the measures S∗(µ) and νran(S) are mutually absolutely continuous. Proof. We have S∗(µ) << νran(S) by Lemma 5.9(1). Also, S∗(S∗(νran(S))) = νran(S) by Parts (5) and (6) of Lemma 5.9, and S∗(νran(S)) << µ by Part (1) of Lemma 5.9, so νran(S) << S∗(µ) by Part (2) of Lemma 5.9. (cid:3) We summarize some standard computations. Corollary 5.13. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, and let S be a bijective measurable set transformation from (X, B, µ) to (Y, C, ν). Set h =(cid:20) d(S∗(µ)) dν (cid:21) ∈ L0(Y, ν). (cid:20) dS∗(σ) dS∗(λ)(cid:21) = S∗(cid:18)(cid:20) dσ dλ(cid:21)(cid:19) almost everywhere [S∗(λ)]. Then h has values in (0, ∞) almost everywhere [ν], and (cid:20) d((S−1)∗(ν)) dµ (cid:21) = 1 (S−1)∗(h) = (S−1)∗(cid:18) 1 h(cid:19) . Moreover, whenever ξ ∈ L0(X, µ) is nonnegative, or one of the integrals in the following exists (in which case they all do), we have ZX ξ dµ =ZY S∗(ξ) d(S∗(µ)) =ZY S∗(ξ)h dν. Proof. This follows from Lemma 5.9(4) for S and S−1, combined with standard properties of Radon-Nikodym derivatives and Proposition 5.6(3). (cid:3) Corollary 5.14. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, and let S be a bijective measurable set transformation from (X, B, µ) to (Y, C, ν). (1) Let λ be a σ-finite measure on (X, B) such that λ << µ. Then S∗(λ) is σ-finite and S∗(λ) << ν. (2) Let λ and σ be σ-finite measures on (X, B) such that σ, λ, and µ are all mutually absolutely continuous. Then S∗(σ), S∗(λ), and S∗(µ) are all mutually absolutely continuous, and CUNTZ ALGEBRAS ON Lp SPACES 31 Proof. For (1), set τ = S∗(λ). Then τ = (S−1)∗(λ), so S∗(τ ) = λ by Lemma 5.9(6). Therefore Lemma 5.9(3) implies that τ is σ-finite, and it follows from Lemma 5.9(1), applied to S−1, that S∗(λ) << µ. We prove (2). Mutual absolute continuity follows from Lemma 5.9(2), applied to S−1. In particular, we can use Corollary 5.13 with σ and λ in place of µ. Next, since the measures are σ-finite by (1), the required Radon-Nikodym derivatives exist. We then prove the result by showing that ZY η · S∗(cid:18)(cid:20) dσ dλ(cid:21)(cid:19) dS∗(λ) =ZY η dS∗(σ) for all nonnegative η ∈ L0(Y, ν). By Corollary 5.8(2), we may assume that η = S∗(ξ) for some ξ ∈ L0(X, µ). We may take ξ to be nonnegative by Proposition 5.6(6). Using Corollary 5.13 for (X, B, σ) and Proposition 5.6(3) at the first step, and Corollary 5.13 for (X, B, λ) at the third step, we have ZY S∗(ξ)S∗(cid:18)(cid:20) dσ dλ(cid:21)(cid:19) dS∗(λ) =ZX ξ(cid:20) dσ dλ(cid:21) dλ =ZX ξ dσ =ZY S∗(ξ) dS∗(σ). This completes the proof. (cid:3) There is a more general version of Corollary 5.14(2), in which we only assume σ << λ << µ. Since it has a more complicated statement and we don't need it, we omit it. 6. Spatial partial isometries and Lamperti's Theorem We will need a systematic theory of isometries and partial isometries on Lp spaces. The main result is Lamperti's Theorem [18], according to which, for p ∈ (0, ∞)\{2}, every isometry between Lp spaces is "semispatial" in a sense which we describe be- low. The material we need in order to to make effective use of Lamperti's Theorem seems not to be in the literature, so we describe it here. There are two choices of how to formulate the definitions, giving different results on L2(X, µ). We adopt the stricter choice. We will need the following three computations. Lemma 6.1. Let (X, B, µ) and (Y, C, ν) be measure spaces, let p ∈ [1, ∞], let S be an injective measurable set transformation from (X, B, µ) to (Y, C, ν) such that νran(S) is σ-finite, and let g be a measurable function on Y such that g(y) = 1 for almost all y ∈ Y. (1) The Radon-Nikodym derivative h =(cid:20) dS∗(µ) d(νran(S))(cid:21) ∈ L0(Y, νran(S)) exists and satisfies h(y) ∈ (0, ∞) for almost all y ∈ Y with respect to νran(S). (2) Let h be as in (1). Let ξ be a measurable function on X. Define a measurable function η on Y by Then kηkp = kξkp. η = gh1/pS∗(ξ). 32 N. CHRISTOPHER PHILLIPS (3) Let ξ be a measurable function on X, let η be as in (2), and suppose that S is bijective. Then for almost all x ∈ X, we have ξ = (S−1)∗(cid:18) 1 g(cid:19)(cid:20) d(S−1)∗(ν) dµ (cid:21)1/p (S−1)∗(η). Remark 6.2. Lemma 6.1(2) says that ξ 7→ gh1/pS∗(ξ) defines an isometry in measure νran(S) need not be σ-finite, and in this case we do not get an element of L(cid:0)Lp(X, µ), Lp(Y, nu)(cid:1). (This will be made explicit in Lemma 6.5 below.) The L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1). Example: take X to consist of one point x, take Y = Z>0, take µ and ν to be counting measure, take S(∅) = ∅ and S(X) = Y, and take g = 1. Proof of Lemma 6.1. We prove (1). The measures S∗(µ) and νran(S) are mutually absolutely continuous by Lemma 5.12. The measure νran(S) is σ-finite by hypoth- esis. n=1 Xn with µ(Xn) < ∞. Choose Yn ∈ C such that S([Xn]) = [Yn]. Then S∗(µ)(Yn) = µ(Xn) < ∞. Since n=1 Yn) = 0. The claim follows, as We claim that S∗(µ) is σ-finite. Write X = S∞ [Y \S∞ n=1 Yn] = S([∅]), we have S∗(µ) (Y \S∞ For the remaining two parts, we present the proof for p 6= ∞. (The case p = ∞ is simpler.) For (2), we have, using g = 1 almost everywhere at the first step and Corollary 5.13 at the third step, does (1). kηkp p =ZY hS∗(ξ)p dν =ZY S∗(ξ)p dS∗(µ) =ZX ξp dµ = kξkp p. For (3), use Corollary 5.13 at the first step, Proposition 5.6(3) at the second step, and Proposition 5.6(7) at the third step, to get, with equalities almost every- where [µ], (S−1)∗(cid:18) 1 g(cid:19)(cid:20) d(S−1)∗(ν) = (S−1)∗(cid:18) 1 (cid:21)1/p g(cid:19) (S−1)∗(cid:18) 1 h(cid:19)1/p dµ (S−1)∗(η) This completes the proof. (S−1)∗(η) = (S−1)∗(S∗(ξ)) = ξ. (cid:3) Definition 6.3. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces. A semis- patial system for (X, B, µ) and (Y, C, ν) is a quadruple (E, F, S, g) in which E ∈ B, in which F ∈ C, in which S is an injective measurable set transformation from C-measurable function on F such that g(y) = 1 for almost all y ∈ F. We say that (E, F, S, g) is a spatial system if, in addition, S is bijective. (E, BE, µE) to (cid:0)F, C0F , νF(cid:1) such that νran(S) is σ-finite, and in which g is a p ∈ [1, ∞]. A linear map s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) is called a semispatial partial isometry if there exists a semispatial system (E, F, S, g) such that, for every ξ ∈ Lp(X, µ), we have Definition 6.4. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, and let g(y)(cid:16)h dS∗(µE ) d(νran(S))i (y)(cid:17)1/p 0 S∗(ξE)(y) y ∈ F y 6∈ F. (sξ)(y) = CUNTZ ALGEBRAS ON Lp SPACES 33 (When p = ∞, we take to be the constant function 1.) d(νran(S))(cid:21)1/p (cid:20) dS∗(µE) We call s a spatial partial isometry if (E, F, S, g) is in fact a spatial system. We call (E, F, S, g) the (semi)spatial system of s. (We will see in Lemma 6.6 below that it is essentially unique.) We call E and F the domain support and the range support of s. The sub-σ-algebra ran(S) ⊂ CF is called the range σ-algebra, the measurable set transformation S is called the (semi)spatial realization of s, and g is called the phase factor . If µ(X \ E) = 0, we call s a (semi)spatial isometry. Lemma 6.5. Let (X, B, µ) and (Y, C, ν) be measure spaces, let p ∈ [1, ∞], and let (E, F, S, g) be a semispatial system for (X, B, µ) and (Y, C, ν). (1) There exists a unique semispatial partial isometry s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) (2) Let s be as in (1). Then ksξkp = kξEkp for every ξ ∈ Lp(X, µ), and whose semispatial system is (E, F, S, g). ksk ≤ 1. (3) Let s be as in (1). Then for any set B ∈ C, the range of s is contained in Lp(Y, νB) if and only if B contains F up to a set of measure zero. (4) Suppose s as in (1) is a semispatial isometry. Then s is isometric as a linear map, that is, ksξkp = kξkp for every ξ ∈ Lp(X, µ). Proof. Lemma 6.1(2), applied with E in place of X and F in place of Y, implies existence of s in (1). Uniqueness is obvious. Part (2) follows from Lemma 6.1(2). Part (4) is a special case of (2). For (3), it is obvious that if B contains F up to a set of measure zero, then ran(s) ⊂ Lp(Y, νB). Now suppose ran(s) ⊂ Lp(Y, νB). Choose ξ ∈ Lp(E, µ) such that ξ(x) 6= 0 for all x ∈ X. It follows from Proposition 5.6(6) that S∗(ξE) is nonzero almost everywhere on F, from the hypotheses that g is nonzero almost everywhere on F, and from Lemma 6.1(1) that d(νran(S))(cid:21) (cid:20) dS∗(µE) is nonzero almost everywhere on F. Therefore sξ is nonzero almost everywhere on F, whence B contains F up to a set of measure zero. (cid:3) Lemma 6.6. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞], and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be a semispatial partial isometry. Then its spatial system is unique up to changes in the domain and range supports by sets of measure zero and up to equality almost everywhere [ν] for the phase factors. Remark 6.7. In Lemma 6.6, we identify BE/N (µE) with a subset of B/N (µ). This subset does not change if E is modified by a set of measure zero. We treat CF /N (νF ) similarly. With these identifications, it makes sense to say that the measurable set transformation component is actually unique. In particular, he part about the domain and range supports says they are uniquely determined as elements of B/N (µ) and C/N (ν). Proof of Lemma 6.6. Suppose that for j = 1, 2, the sets Ej ⊂ X and Fj ⊂ Y are measurable, Sj is an injective measurable set transformation from (Ej, BEj , µEj ) to (Fj, CFj , νFj ), and gj is a measurable function on Fj with gj(y) = 1 for 34 N. CHRISTOPHER PHILLIPS almost all y ∈ Fj. Let s1 and s2 be the semispatial partial isometries obtained from Definition 6.4. We have to prove that if s1 = s2, then [E1] = [E2], [F1] = [F2], S1 = S2, and g1 = g2 almost everywhere [ν]. It follows from Lemma 6.5(3) that [F1] = [F2]. It is clear from Lemma 6.5(2) that we must have [E1] = [E2]. For j = 1, 2, set d(νran(Sj))(cid:21) . hj =(cid:20) d(Sj)∗(µE) Let B ∈ BE1 satisfy µ(B) < ∞. Then χB ∈ Lp(X, µ). From the formula for s1(χB), and since h1 and g1 are nonzero almost everywhere on F, it follows that S1([B]) = [C] if and only if s1(χB) is nonzero almost everywhere on C and zero almost everywhere on Y \ C. Of course, the same applies to s2 and S2. Since µ is σ-finite and S1 and S2 are σ-homomorphisms, it follows that S1 = S2, and also that h1 = h2. It remains to prove that g1 = g2 almost everywhere [ν]. Choose ξ ∈ Lp(E, µ) such that ξ(x) 6= 0 for all x ∈ X. It follows from Proposition 5.6(6) that (S1)∗(ξE) is nonzero almost everywhere on F1, and from Lemma 6.1(1) that h1 is nonzero almost everywhere on F1. From s1ξ = s2ξ almost everywhere, we therefore get g1 = g2 almost everywhere on F1, as desired. (cid:3) Remark 6.8. Let (X, B, µ) and (Y, C, ν) be measure spaces, let p ∈ [1, ∞], and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be a semispatial partial isometry with domain support E ⊂ X, range support F ⊂ Y, semispatial realization S, and range σ-algebra C0. Then s is spatial if and only if C0 = CF . If this is the case, then s is a spatial isometry if and only if µ(X \ E) = 0. Using the terminology we have introduced, we can now state Lamperti's Theorem as follows. Theorem 6.9 (Lamperti). Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞) \ {2}, and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be an isometric (not nec- essarily surjective) linear map. Then s is a semispatial isometry in the sense of Definition 6.4. Proof. The proof is the same as that of Theorem 3.1 of [18]. In [18] it is assumed that (X, B, µ) = (Y, C, ν), but this is never used in the proof. (cid:3) Remark 6.10. Let p ∈ [1, ∞) \ {2}, and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be an isometry. Theorem 6.9 states that there exists a semispatial system (E, F, S, g), with E = X, such that the construction of Definition 6.4 yields s. The function g was only required to be measurable with respect to CF , not with respect to ran(S). In general, the stronger condition fails. Take X to consist of one point x and Y to consist of two points y1 and y2. Let µ and ν be the counting measures. Identify Lp(X, µ) with C and Lp(Y, ν) with C2 in the obvious way. Define s by s(λ) = 2−1/p(λ, −λ). The semispatial system must be (X, Y, S, g), with S(∅) = ∅ and S(X) = Y, and with g(y1) = 1 and g(y2) = −1. The function g is then not ran(S)-measurable. We are primarily interested in partial isometries which are spatial rather than merely semispatial. We will need notation for multiplication operators on Lp(X, µ). CUNTZ ALGEBRAS ON Lp SPACES 35 Notation 6.11. Let (X, B, µ) be a measure space, and let p ∈ [1, ∞]. We let mX,µ,p (or, when no confusion should arise, just mX or m) be the homomorphism mX,µ,p : L∞(X, µ) → L(Lp(X, µ)) defined by for f ∈ L∞(X, µ), ξ ∈ Lp(X, µ) and x ∈ X. (cid:0)mX,µ,p(f )ξ(cid:1)(x) = f (x)ξ(x) Lemma 6.12. Let (X, B, µ) and (Y, C, ν) be measure spaces, let p ∈ [1, ∞], and let (E, F, S, g) be a spatial system for (X, B, µ) and (Y, C, ν) (in the sense of Defi- nition 6.3). whose spatial system is (E, F, S, g). (2) Let s be as in (1). Then the range of s is Lp(F, ν) ⊂ Lp(Y, ν). (3) Let s be as in (1). There exists a unique spatial partial isometry t ∈ (1) There exists a unique spatial partial isometry s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) L(cid:0)Lp(Y, ν), Lp(X, µ)(cid:1) whose semispatial system is(cid:0)F, E, S−1, (S−1)∗(g)−1(cid:1). (4) Let s be as in (1) and let t be as in (3). Let u ∈ L(cid:0)Lp(Y, ν), Lp(X, µ)(cid:1) Moreover, using Notation 6.11, we have ts = m(χE) and st = m(χF ). satisfy us = m(χE) and uLp(Y \F, ν) = 0. Then u = t. Proof. Part (1) follows from Lemma 6.5(1) and Definition 6.4. The existence of t in (3) follows from part (1), and the formulas for ts and st in (3) follow from Lemma 6.1(3). Lemma 6.5(4) and the formula for st imply (2). For (4), let η ∈ Lp(Y, ν). Write η = η1 + η2 with η1 ∈ Lp(F, ν) and η2 ∈ Lp(Y \ F, ν). We show that uη1 = tη1 and uη2 = tη2. Clearly uη2 and tη2 are both zero. Also, using ran(t) = Lp(E, µ) at the last step, we have uη1 = um(χF )η1 = ustη1 = m(χE)tη1 = tη1. This completes the proof. (cid:3) Definition 6.13. Let (X, B, µ) and (Y, C, ν) be measure spaces, let p ∈ [1, ∞], and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be a spatial partial isometry. The spatial partial isometry t of Lemma 6.12(3) is called the reverse of s. Remark 6.14. When p = 2, the reverse of s is of course s∗. However, we can't define it this way when p 6= 2. Lemma 6.15. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞], and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be a semispatial partial isometry with domain support E ⊂ X, range support F ⊂ Y, and range σ-algebra C0. Then the following are equivalent: (1) s is spatial. (2) ran(s) = Lp(F, µ). (3) C0 = CF . Proof. That (3) implies (1) is clear from the definitions, and (1) implies (2) by Lemma 6.12(2). So assume (2). Let S be the semispatial realization of s. Thus S is an injective measurable set transformation from (E, BE, µE) to (F, CF , νF ) such that ran(S∗) contains χB for every B ∈ CF with ν(B) < ∞. It follows from Corollary 5.8 that S is surjective, which is (3). (cid:3) Lemma 6.16. Let p ∈ [1, ∞) \ {2}, let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let E ⊂ X and F ⊂ Y be measurable subsets, and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) satisfy the following conditions: 36 N. CHRISTOPHER PHILLIPS (1) The range of s is Lp(F, ν) ⊂ Lp(Y, ν). (2) sLp(E,µ) is isometric. (3) sLp(X\E, µ) = 0. Then s is a spatial partial isometry in the sense of Definition 6.4, and has domain support E and range support F. Proof. Apply Theorem 6.9 with E in place of X, to conclude that s is a semispatial partial isometry. Then s is spatial by Lemma 6.15. (cid:3) In the rest of this section, we describe some operations which give new spatial partial isometries from old ones. Most of the statements have analogs for semispatial partial isometries, but we don't need them and don't prove them. We begin by showing that the product of two spatial partial isometries is again a spatial partial isometry. On a Hilbert space, the product of two partial isometries is usually not a partial isometry, unless the range projection of the second com- mutes with the domain projection of the first. With spatial partial isometries, this commutation relation is automatic. Lemma 6.17. Let (X1, B1, µ1), (X2, B2, µ2), and (X3, B3, µ3) be σ-finite measure spaces. Let p ∈ [1, ∞]. Let s ∈ L(cid:0)Lp(X1, µ1), Lp(X2, µ2)(cid:1) and u ∈ L(cid:0)Lp(X2, µ2), Lp(X3, µ3)(cid:1) be spatial partial isometries, with reverses t and w, and with spatial systems (E1, E2, S, g) and (F2, F3, V, h). Then vs is a spatial partial isometry. Its domain support is E = S−1(E2 ∩ F2), its range support is F = V (E2 ∩ F2), and its re- verse is tw. Its spatial realization is the composite V0 ◦ S0 of the restriction S0 of S to B1S−1(E2∩F2) and the restriction V0 of V to B2E2∩F2 . Its phase factor is k = (V0)∗(gE2∩F2)(hV (E2∩F2)). Proof. It is immediate that (E, F, V0 ◦ S0, k) is a spatial system for (X1, B1, µ1) and (X3, B3, µ3). Let y ∈ L(cid:0)Lp(X1, µ1), Lp(X3, µ3)(cid:1) be the corresponding spatial partial isometry. We claim that y = vs, and it is enough to prove that yξ = vsξ for ξ in each of the three spaces Lp(X1 \ E1, µ1), Lp(E1 \ E, µ1), and Lp(E, µ1). In the first case, yξ = 0 and sξ = 0. In the second case, yξ = 0. Also sξ ∈ Lp(X2 \ F2, µ2), so vsξ = 0. So let ξ ∈ Lp(E, µ1). Then yξ = (V0)∗(gE2∩F2)(hV (E2∩F2))(cid:20) d(V0 ◦ S0)∗(µ1E) and vsξ = (hF )(cid:20) d(V0)∗(µ2E2∩F2) d(µ3F ) (cid:21)1/p (V0 ◦ S0)∗(ξ) d(µ3F ) (cid:21)1/p (V0)∗ (gE2∩F2 )(cid:20) d(S0)∗(µ1E) d(µ2E2∩F2) (cid:21)1/p (S0)∗(ξ)! . One sees that these are equal by combining Corollary 5.14 with several applications of Proposition 5.6(3) and standard properties of Radon-Nikodym derivatives. This completes the proof that y = vs, and thus the proof that vs is as claimed. It remains to identify the reverse. We use Lemma 6.12(4). First observe that wη = 0 for η ∈ Lp(X3 \ F3, µ3) and twη = 0 for η ∈ Lp(F3 \ F, µ3). Also, for ξ ∈ Lp(X1, µ1), the first paragraph of the proof shows that vsξ = vsm(χE)ξ. It is easy to check that sm(χE) = m(χE2∩F2)s. Therefore twvs = t(wv)sm(χE ) = tm(χF2 )m(χE2∩F2 )s = tsm(χE) = m(χE). CUNTZ ALGEBRAS ON Lp SPACES This completes the proof. 37 (cid:3) Lemma 6.18. Let (X, B, µ) be a σ-finite measure space, let p ∈ [1, ∞], and let e ∈ L(Lp(X, µ)). Then the following are equivalent: (1) e is an idempotent spatial partial isometry. (2) e is a spatial partial isometry, and there is E ∈ B such that the spatial system of e is (E, E, idBE , χE). (3) There is E ∈ B such that e = m(χE). Proof. It is clear that (2) and (3) are equivalent, and that (2) implies (1). So assume (1). Let the spatial system of e be (E, F, S, g), and let f be the reverse of e (Definition 6.13). Lemma 6.17 implies that f 2 = f. Therefore (6.1) m(χF )m(χE) = (ef )(f e) = ef e = em(χE) = e. Multiplying (6.1) on the left by m(χE) gives (6.2) m(χE)m(χF )m(χE) = m(χE)e = (f e)e = f e = m(χE). The left hand sides of (6.1) and (6.2) are clearly equal, so e = m(χE). This is (3). (cid:3) We now consider the tensor product of two spatial partial isometries. We need a particular case of the product of two σ-homomorphisms, which we can get from Lamperti's Theorem. Quite possibly something more general is true, but we don't need it. Lemma 6.19. Let (X1, B1, µ1), (X2, B2, µ2), (Y1, C1, ν1), and (Y2, C2, ν2) be σ-finite measure spaces. Let S : B1/N (µ1) → B2/N (µ2) and V : C1/N (ν1) → C2/N (ν2) be bijective σ-homomorphisms. Let B1×C1 and B2×C2 be the product σ-algebras on X1 ×Y1 and X2 ×Y2, or their completions. Then there is a unique σ-homomorphism S × V : (B1 × C1)/N (µ1 × ν1) → (B2 × C2)/N (µ2 × ν2) such that, whenever E1 ∈ B1, E2 ∈ B2, F1 ∈ C1, and F2 ∈ C2 satisfy S([E1]) = [E2] and V ([F1]) = [F2], we have (6.3) (S × V )([E1 × F1]) = [E2 × F2]. Proof. It does not matter whether we use the product σ-algebras or their comple- tions, because the Boolean σ-algebras (B1 × C1)/N (µ1 × ν1) and (B2 × C2)/N (µ2 × ν2) are the same with either choice. Uniqueness of S×V follows from the fact that the measurable rectangles generate the product σ-algebra. We prove existence. Fix any p ∈ [1, ∞)\{2}. Let s and v be the spatial isometries with spatial systems (X1, X2, S, 1) and (Y1, Y2, V, 1). Then s and v are isometric bijections. Applying Theorem 2.16(5) to s and v, and to their inverses, and applying Theorem 2.16(6), we see that s ⊗ v : Lp(X1 × Y1, µ1 × ν1) → Lp(X2 × Y2, µ2 × ν2) is an isometric bijection. Lemma 6.16 implies that it is spatial. We take S × V to be its spatial realization. The relation (6.3) is easily deduced from (s ⊗ v)(ξ ⊗ η) = sξ ⊗ vη, Fubini's Theorem, and σ-finiteness of all the measures involved. (cid:3) 38 N. CHRISTOPHER PHILLIPS In the next lemma, we exclude p = ∞ because we use Theorem 2.16. Lemma 6.20. Let (X1, B1, µ1), (X2, B2, µ2), (Y1, C1, ν1), and (Y2, C2, ν2) be σ-finite measure spaces. Let p ∈ [1, ∞). Let s ∈ L(cid:0)Lp(X1, µ1), Lp(X2, µ2)(cid:1) and v ∈ L(cid:0)Lp(Y1, ν1), Lp(Y2, ν2)(cid:1). be spatial partial isometries, with reverses t and w, and with spatial systems (E1, E2, S, g) and (F2, F3, V, h). Then s ⊗ v ∈ L(cid:0)Lp(X1 × Y1, µ1 × ν1), Lp(X2 × Y2, µ2 × ν2)(cid:1) (as in Theorem 2.16) is a spatial partial isometry. With S × V as in Lemma 6.19 and g1 ⊗ g2 as in Theorem 2.16, its spatial system is (6.4) and its reverse is t ⊗ w. (cid:0)E1 × F1, E2 × F2, S × V, g1 ⊗ g2(cid:1), Proof. Let y be the spatial partial isometry with spatial system given by (6.4), and let z be its reverse. It is clear that S∗(µ1) × V∗(ν1) and (S × V )∗(µ1 × ν1) agree on measurable rectangles. The measures S∗(µ1) and V∗(ν1) are σ-finite by Corollary 5.14(1). The product of σ-finite measures is uniquely determined by its values on mea- surable rectangles. (In Chapter 12 of [24], see Theorem 8 and the discussion after Lemma 14.) Therefore S∗(µ1) × V∗(ν1) = (S × V )∗(µ1 × ν1). Set k =(cid:20) d(S × V )∗(µ1E1 × ν1F1) d(µ2E2 × ν2F2 ) (cid:21) and l =(cid:20) dS∗(µ1E1) d(µ2E2 ) (cid:21) ⊗(cid:20) dV∗(ν1F1) d(ν2F2 ) (cid:21) . Then for every measurable rectangle R ⊂ E2 × F2, we have ZR k d(µ2E2 × ν2F2 ) =ZR l d(µ2E2 × ν2F2 ). Since G 7→ZG l d(µ2E2 × ν2F2) is a product of σ-finite measures on E2 × F2, it follows from uniqueness of product measures (as above) that k = l almost everywhere [µ2E2 × ν2F2 ]. One now checks that s⊗v = y and t⊗w = z by showing, in each case, that they agree on elementary tensors. (cid:3) Lemma 6.21. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, let p ∈ [1, ∞), (E, F, S, g) and with reverse t. Let q ∈ (1, ∞] satisfy 1 q = 1, and for any σ- finite measure space (Z, D, λ), identify Lp(Z, λ)′ with Lq(Z, λ) in the usual way. and let s ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1) be a spatial partial isometry with spatial system Then s′ ∈ L(cid:0)Lq(Y, ν), Lq(X, µ)(cid:1) is a spatial partial isometry with spatial system (cid:0)F, E, S−1, (S−1)∗(g)(cid:1) and with reverse t′. Proof. To simplify the notation, let p + 1 h =(cid:20) dS∗(µ) dν (cid:21) and k =(cid:20) d(S−1)∗(ν) dµ (cid:21) . CUNTZ ALGEBRAS ON Lp SPACES 39 Then h = S∗(k)−1 by Corollary 5.13. Let u ∈ L(cid:0)Lq(Y, ν), Lq(X, µ)(cid:1) be the spatial partial isometry with the spatial system specified for s′. Then for ξ ∈ Lp(X, µ) and η ∈ Lq(Y, ν), we have, using Corollary 5.13 at the second step, ZX ξ · uη dµ =ZX =ZY ξ(S−1)∗(g)k1/q(S−1)∗(η) dµ =ZY S∗(ξ)gh−1/qηh dν =ZY S∗(ξ)gh1/pηh dν =ZY S∗(ξ)gS∗(k)1/qηh dν sξ · η dν. Thus s′ = u. To identify the reverse of s′, we calculate: t′s′ = (st)′ = mY,ν,p(χF )′ = mY,ν,q(χF ) and Now apply Lemma 6.12(4). t′mX,µ,q(χX\E) =(cid:2)mX,µ,p(χX\E)t(cid:3)′ = 0. (cid:3) We finish this section with a lemma on homotopies that will be needed later. We do not know whether surjectivity is necessary in the hypotheses. Lemma 6.22. Let p ∈ [1, ∞) \ {2}. Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, and let λ 7→ sλ ∈ L(cid:0)Lp(X, µ), Lp(Y, ν)(cid:1), for λ ∈ [0, 1], be a norm continuous path of surjective isometries. Let Sλ be the spatial realization of sλ. Then S0 = S1. Proof. We prove that if v0 and v1 are surjective spatial isometries whose spatial realizations V0 and V1 are distinct, then kv0 −v1k ≥ 1. (It follows that the spatial re- alization must be constant along a homotopy.) The measurable set transformations V0 and V1 are bijective by Lemma 6.15. Choose a set E ∈ B such that V0([E]) 6= V1([E]). Without loss of generality we may assume that V0([E]) and V1([E]) have representatives F0, F1 ∈ C such that F0 does not contain F1 up to sets of measure zero. That is, there is F ⊂ Y such that ν(F ) > 0, F ⊂ F0, and F ∩ F1 = ∅. Since V0 is bijective, the set Q = V −1 (F ) satisfies µ(Q) > 0. Since µ is σ-finite, replacing F by a suitable subset allows us to also assume that µ(Q) < ∞. Correcting by a set of measure zero, we may further assume that Q ⊂ E. 0 Define ξ = µ(Q)−1/pχQ ∈ Lp(X, µ). Then kξkp = 1. Moreover, v0ξ is supported in F and, since Q ⊂ E, the function v1ξ is supported in F1. Therefore kv0ξ − v1ξkp p = kv0ξkp p + kv1ξkp p = 2, so kv0 − v1k ≥ 21/p ≥ 1. (cid:3) 7. Spatial representations In this section, we define and characterize spatial representations on spaces of the form Lp(X, µ), first of Md and then of Ld for finite d. In each case, we give a number of equivalent conditions for a representation to be spatial, some of them quite different from each other. In particular, some characterizations are primarily in terms of how the representation interacts with X, while others make sense for a representation on any Banach space. We consider Md first because we use the results about Md in the theorem for Ld. We will see in Theorem 7.2 that, for fixed p, any two spatial representations of Md determine the same norm on Md, and we will see in Theorem 8.7 (in the 40 N. CHRISTOPHER PHILLIPS next section) that, for for fixed p and when d < ∞, any two spatial representations of Ld determine the same norm on Ld. Definition 7.1. Let d ∈ Z>0, and let (ej,k)d j,k=1 be the standard system of ma- trix units in Md. Let p ∈ [1, ∞], let (X, B, µ) be a σ-finite measure space, and let ρ : Md → L(Lp(X, µ)) be a representation. (Recall that, by convention, represen- tations are unital. See Definition 2.8.) We say that ρ is spatial if ρ(ej,k) is a spatial partial isometry, in the sense of Definition 6.4, for j, k = 1, 2, . . . , d. Theorem 7.2. Let d ∈ Z>0, let p ∈ [1, ∞) \ {2}, let (X, B, µ) be a σ-finite measure space, and let ρ : Md → L(Lp(X, µ)) be a representation. Let (ej,k)d j,k=1 be the standard system of matrix units in Md. Then the following are equivalent: (1) ρ is spatial. (2) For j, k = 1, 2, . . . , d, the operator ρ(ej,k) is a spatial partial isometry with reverse ρ(ek,j). (3) ρ is isometric as a map from M p (4) ρ is contractive as a map from M p (5) kρ(ej,k)k ≤ 1 for j, k = 1, 2, . . . , d, and there exists a measurable parti- j=1 Xj such that for j = 1, 2, . . . , d, the matrix unit ej,j acts d (as in Notation 2.4) to L(Lp(X, µ)). d to L(Lp(X, µ)). (following Notation 6.11) as ρ(ej,j) = m(χXj ). tion X =`d (6) There exists a measurable partition X = `d (7) There exists a measurable partition X = `d j=1 Xj such that for j = 1, 2, . . . , d the operator ρ(ej,1) is a spatial partial isometry with domain support X1 and range support Xj. j=1 Xj such that for j, k = 1, 2, . . . , d the operator ρ(ej,k) is zero on Lp(X \ Xk, µ) and restricts to an isometric isomorphism from Lp(Xk, µ) to Lp(Xj, µ). (8) With γ being counting measure on Nd = {1, 2, . . . , d}, there exists a σ-finite measure space (Y, C, ν) and a bijective isometry u : Lp(Nd × Y, γ × ν) → Lp(X, µ) such that, following the notation of Theorem 2.16(4), for all a ∈ Md we have ρ(a) = u(a ⊗ 1)u−1. d(cid:1) is In the notation of Lemma 2.17, Theorem 7.2(8) says that if ρ0 : Md → L(cid:0)lp the standard representation, then ρ is similar, via an isometry, to ρ0 ⊗p 1. We specifically use complex scalars in the proof that (3) and (4) imply the other conditions. We don't know whether complex scalars are necessary. When p = 2, conditions (1) and (3) are certainly not equivalent. We have not investigated what happens when p = ∞, but Lamperti's Theorem is not available in this case. It is essential that the representation be unital. Many of the conditions of Theo- rem 7.2 are never satisfied for nonunital representations, but (3) and (4) do occur. We show by example that they do not imply that the representation is spatial. Example 7.3. We adopt the notation of Example 2.5, with p ∈ [1, ∞) \ {2}. Set ξ = 2−1/p(1, 1) and η = 2−1/q(1, 1). (If p = 1, take η = (1, 1).) Let e ∈ M p Example 2.5. Then 2 be the rank one operator called a in e = 1 2(cid:18)1 1 1 1(cid:19) , CUNTZ ALGEBRAS ON Lp SPACES 41 which is an idempotent. Since kξkp = kηkq = 1, we get kek = 1. Now take (X0, B0, µ0) to be any σ-finite measure space with a spatial represen- tation ρ0 : Md → L(cid:0)Lp(X0, µ0)(cid:1). Set X = X0 ∐ X0, and equip it with the obvious σ-algebra B and with the measure µ whose restriction to each copy of X0 is µ0. Identify Lp(X, µ) with lp 2 ⊗p Lp(X0, µ0) as in Theorem 2.16. Define a nonunital representation ρ : Md → L(Lp(X, µ)) by ρ(a) = e ⊗ ρ0(a) for a ∈ Md. Then Theo- rem 2.16(5) implies that, regarded as a map M p d → L(Lp(X, µ)), the representation ρ is isometric. But it is not spatial, not even in a sense suitable for nonunital representations. Proof of Theorem 7.2. We first prove the equivalence of (1), (5), (6), and (7), be- ginning with (1) implies (5). So assume (1). We have kρ(ej,k)k ≤ 1 because this is true for all spatial partial isometries. For each j and k there is a spatial system for ρ(ej,k), say (Ej,k, Fj,k, Sj,k, gj,k). Apply Lemma 6.18 to ρ(ej,j). We obtain sets Xj ⊂ X such that (Ej,j, Fj,j, Sj,j, gj,j) = (Xj, Xj, idBXj , χXj ) and ρ(ej,j) = m(χXj ) now gives the rest of (5). j=1 ρ(ej,j) = 1, the sets Xj are essentially disjoint and, j=1 Xj = X. Modification by sets of measure zero for j = 1, 2, . . . , d. SincePd up to a set of measure zero, Sd We next prove (5) implies (7). We take the partition X = `d in (5). Let j, k ∈ {1, 2, . . . , d}. The equation ρ(ej,k)(cid:0)1 − ρ(ek,k)(cid:1) = 0 translates to ρ(ej,k)ξ = 0 for all ξ ∈ Lp(X \ Xk, µ), which is the first part of (7). For ξ ∈ Lp(Xk, µ), use j=1 Xj to be as kρ(ej,k)k ≤ 1, kρ(ek,j)k ≤ 1, and ρ(ek,j)ρ(ej,k)ξ = ρ(ek,k)ξ = ξ to get kρ(ej,k)ξk = kξk. So ρ(ej,k) is isometric on Lp(Xk, µ). Since ρ(ej,k)ρ(ek,j)ξ = ξ for ξ ∈ Lp(Xj, µ), we have ran(ρ(ej,k)) ⊃ Lp(Xj, µ), while the equation m(χXj )ρ(ej,k)ξ = ρ(ej,j)ρ(ej,k)ξ = ρ(ej,k)ξ for ξ ∈ Lp(Xk, µ) implies ran(ρ(ej,k)) ⊂ Lp(Xj, µ). This completes the proof of (7). That (7) implies (6) is immediate from Lemma 6.16. We prove that (6) implies (1). From (6), we see that ran(ρ(ej,1)) = Lp(Xj, µ) for j = 1, 2, . . . , d. Moreover, ρ(e1,1) = m(χX1 ) by Lemma 6.18. Now fix j, and consider ρ(e1,j). For k 6= j we have ρ(e1,j)Lp(Xk,µ) = 0 since e1,jek,1 = 0. Also ρ(e1,j)ρ(ej,1) = ρ(e1,1) = m(χX1 ). So Lemma 6.12(4) implies that ρ(e1,j) is the reverse of ρ(ej,1). In particular, ρ(e1,j) is a spatial partial isometry. For j, k = 1, 2, . . . , d, it now follows from Lemma 6.17 that ρ(ej,k) = ρ(ej,1)ρ(e1,k) is a spatial partial isometry, which is (1). We next prove that (2) and (8) are equivalent to the conditions we have already j=1 Xj be as in (7). Define Y = X1 and ν = µX1 . Identify Lp(Nd × Y, γ × ν) with the space of sequences (η1, η2, . . . , ηd) ∈ Lp(Y, ν)d, with the norm considered. We start with (7) implies (8). Let the partition X =`d p(cid:1)1/p k(η1, η2, . . . , ηd)k =(cid:0)kη1kp Define u : Lp(Nd × Y, γ × ν) → Lp(X, µ) by . p + kη2kp p + · · · + kηdkp u(η1, η2, . . . , ηd) = η1 + ρ(e2,1)η2 + ρ(e3,1)η3 + · · · + ρ(ed,1)ηd. 42 N. CHRISTOPHER PHILLIPS (Note that η1 = ρ(e1,1)η1.) Then u is isometric because the summands ρ(ej,1)ηj are supported in disjoint subsets of X. It is easy to check that u is bijective and that ρ(ej,k) = u(ej,k ⊗ 1)u−1 for j, k = 1, 2, . . . , d. This proves (8). Now assume (8); we prove (2). It is trivial that the standard representation of Md on Lp(Nd) satisfies (2). It follows from Lemma 6.20 that the representation ρ0(a) = a⊗ 1 also satisfies (2). The operator u is a spatial isometry by Lemma 6.16, and Lemma 6.12(4) implies that its reverse is u−1. Lemma 6.17 now implies that ρ satisfies (2). That (2) implies (1) is trivial. We finish by proving that (3) and (4) are equivalent to the conditions we have already considered. That (8) implies (3) follows from the norm relation ka ⊗ 1k = kak. (See Theorem 2.16(5).) That (3) implies (4) is trivial. Assume (4); we prove (5). For j = 1, 2, . . . , d and ζ ∈ S1, set tj,ζ = 1 − ej,j + ζej,j ∈ M p d . One checks immediately that ktj,ζk = 1, and that t−1 j,ζ = tj,ζ−1. Since ρ is con- tractive, it follows that ρ(tj,ζ) is a bijective isometry for all j and ζ. So ρ(tj,ζ) is spatial by Lemma 6.16. Since ρ(tj,ζ) is bijective, its spatial system has the form (X, X, Sj,ζ, gj,ζ). Clearly Sj,1 = idB. So Lemma 6.22 implies that Sj, −1 = idB. Therefore ρ(tj, −1) = m(gj, −1). It follows that there is a unital algebra homomor- phism ϕ : l∞(cid:0){1, 2, . . . , d}(cid:1) → L∞(X, µ) such that ϕ(χ{j}) = 1 2 (1 − gj, −1) for j = 1, 2, . . . , d. So there is a partition X = j=1 Xj such that for j = 1, 2, . . . , d we have gj, −1 = χXj almost everywhere [µ]. (cid:3) Condition (5) now follows. `d We now turn to representations of Ld and Cd. In Definition 2.12, we defined what it means for a representation to be contractive on generators, forward isometric (on the sj), and strongly forward isometric (in addition, the linear combinations of the sj are sent to scalar multiples of isometries). We now introduce several further properties of a representation. Definition 7.4. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let p ∈ [1, ∞], let (X, B, µ) be a σ-finite measure space, and let ρ : A → L(Lp(X, µ)) be a representation. (1) We say that ρ is disjoint if there exist disjoint sets X1, X2, . . . , Xd ⊂ X (if A = L∞, disjoint sets X1, X2, . . . ⊂ X) such that ran(ρ(sj )) ⊂ Lp(Xj, µ) for all j. (2) We say that ρ is spatial if for each j, the operators ρ(sj) and ρ(tj) are spatial partial isometries, with ρ(tj) being the reverse of ρ(sj) in the sense of Definition 6.13. The definition of a spatial representation is quite strong. For p 6= 2, ∞, we will see that representations satisfying some much weaker conditions are necessarily spatial. The representations of Ld in Examples 3.1 and 3.18 are spatial and disjoint. The representations in Examples 3.7 and 3.8 are disjoint, but not spatial; in fact, they are neither contractive on generators nor forward isometric. The representations in Examples 3.10, 3.11, 3.12, and 3.13 are contractive on generators but not disjoint CUNTZ ALGEBRAS ON Lp SPACES 43 and not spatial. The representations of L∞ in Examples 3.3, 3.4, and 3.19 are spatial and disjoint. The one in Example 3.5 is disjoint, but it is not spatial. (For j ∈ Z>0, the operator ρ(sj) is a spatial partial isometry, but ρ(tj ) is not.) The representation of L∞ of Example 3.16 is disjoint but not spatial, since the images of neither the sj nor the tj are spatial. Lemma 7.5. Let A be any of Ld, Cd, or L∞. Let p ∈ [1, ∞]. Then there is an injective spatial representation of A on lp(Z>0). Proof. For Ld (including d = ∞), we can use any spatial representation, say that of Example 3.1 for d < ∞ and that of Example 3.3 for d = ∞, because Ld is simple (Theorem 2 of [21] for d < ∞ and Example 3.1(ii) of [3] for d = ∞). For Cd, we use the representation π of Example 3.2. We check injectivity. With ρ as in Example 3.1 (for d + 1 in place of d) and ιd, d+1 as in Lemma 1.5, we have π = ρ ◦ ιd, d+1. Moreover, ιd, d+1 is injective by Lemma 1.5 and we saw above that ρ is injective. (cid:3) We give two further conditions, also motivated by Equation (2.3) (before Defi- nition 2.12) for the C*-algebras and by the analogous equation for the adjoints. Definition 7.6. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2) or L∞ (Definition 1.3). Let E be a nonzero Banach space, and let ρ : A → L(E) be a representation. Let p ∈ [1, ∞]. (1) We say that ρ is p-standard on span(s1, s2, . . . , sd) if, following Defini- tion 1.13, the map λ 7→ ρ(sλ) from Cd to L(E) is isometric from lp(cid:0){1, 2, . . . , d}(cid:1) to L(E). (In the case A = L∞ and p 6= ∞, we say that ρ is p-standard on span(s1, s2, . . .) if the map λ 7→ ρ(sλ) from C∞ to L(E) extends to an iso- metric map from lp(Z>0) to L(E). For p = ∞, we use C0(Z>0) in place of l∞.) (2) Let q ∈ [1, ∞] be the conjugate exponent, that is, 1 p + 1 q = 1. We say that ρ is p-standard on span(t1, t2, . . . , td) if, following Definition 1.13, the map λ 7→ ρ(tλ) from Cd to L(E) is isometric from lq(cid:0){1, 2, . . . , d}(cid:1) to L(E). (In the case A = L∞ and p 6= 1, we say that ρ is p-standard on span(t1, t2, . . .) if the map λ 7→ ρ(tλ) from C∞ to L(E) extends to an isometric map from lq(Z>0) to L(E). For p = 1, we use C0(Z>0) in place of l∞.) We will see in Theorem 7.7 below that a spatial representation of Ld on a space of the form Lp(X, µ) is necessarily p-standard on both span(s1, s2, . . . , sd) and span(t1, t2, . . . , td). Example 3.8 shows that a representation which is p-standard on span(s1, s2, . . . , sd) need not be spatial, or even contractive on generators. If in that example one instead defines π(sj) = ρ(sj) ⊕p 2ρ(sj) and π(tj ) = ρ(tj) ⊕p 1 2 ρ(tj), the resulting representation is p-standard on span(t1, t2, . . . , td) but not spatial. One could fix the difficulty with Example 3.8 by incorporating strongly forward isometric in the definition of p-standard on span(s1, s2, . . . , sd), but we don't know what the analogous fix for p-standard on span(t1, t2, . . . , td) should be. Theorem 7.7. Let d ∈ Z>0, let Ld be as in Definition 1.1, let p ∈ (1, ∞) \ {2}, let (X, B, µ) be a σ-finite measure space, and let ρ : Ld → L(Lp(X, µ)) be a representation. Then the following are equivalent: (1) ρ is spatial. 44 N. CHRISTOPHER PHILLIPS (2) For j = 1, 2, . . . , d, the operator ρ(sj) is a spatial partial isometry. (3) For j = 1, 2, . . . , d, the operator ρ(tj) is a spatial partial isometry. (see Lemma 1.11) is a spatial representation of Md in the sense of Defini- tion 7.1. (4) ρ is forward isometric and the restriction of ρ to span(cid:0)(sj tk)d (5) ρ is contractive on generators and the restriction of ρ to span(cid:0)(sjtk)d j,k=1(cid:1) ∼= Md j,k=1(cid:1) is a spatial representation of Md. (6) ρ is forward isometric and disjoint. (7) ρ is contractive on generators and disjoint. (8) ρ is strongly forward isometric and disjoint. (9) ρ is p-standard on span(s1, s2, . . . , sd) and is strongly forward isometric. (10) ρ is p-standard on span(t1, t2, . . . , td) and the representation ρ′ of Lemma 2.21 is strongly forward isometric. (11) (cid:0)ρ(s1) ρ(s2) · · · ρ(sd)(cid:1) is a row isometry, in the sense that, using the notation of Remark 2.15, it defines an isometric linear map Lp(X, µ) ⊕p Lp(X, µ) ⊕p · · · ⊕p Lp(X, µ) → Lp(X, µ). (12) ρ is forward isometric and for j = 1, 2, . . . , d there is Xj ⊂ X such that ran(ρ(sj )) = Lp(Xj, µ). (13) ρ is contractive on generators and for j = 1, 2, . . . , d there is Xj ⊂ X such that ran(ρ(sj )) = Lp(Xj, µ). (14) The representation ρ′ of Lemma 2.21 is spatial. For p = 1, all the conditions except (3), (10), and (14) are equivalent, and the other conditions imply these three. Various remarks are in order. First, when p = 1, we do not know whether the conditions (3), (10), and (14) imply the representation is spatial. (The last two of these are the ones which for p = 1 involve representations on L∞(X, µ).) Second, some of the equivalences fail for representations of L∞ and Cd. Assume p 6= 1. Then the representation of L∞ of Example 3.5 satisfies (2), (6), (9), (11), and (12), but not (3) and hence not (1). The same is true for the restriction to Cd, using the map ιd,∞ of Lemma 1.5. For p ∈ (1, ∞) \ {2}, the representation of L∞ of Example 3.16, and its dual, both satisfy (6), (7), (8), (9), (10), and (11), but none of (2), (3), (12), (13). or (14). The same is true for the representation of Cd0 of Example 3.14. Third, various other implications one might hope for fail. Example 3.10 shows that contractive on generators does not imply that the restriction to Md is spatial. Example 3.7 shows that the restriction to Md being spatial does not imply that the whole representation is spatial. If ρ is p-standard on span(s1, s2, . . . , sd), it does not follow that ρ is p-standard on span(t1, t2, . . . , td), or that ρ′ is q-standard on span(s1, s2, . . . , sd), by Example 3.8. A spatial representation of Ld must be strongly forward isometric, since that is part of Theorem 7.7(8). The converse is not true; Example 3.10 is a counterexample. Next, we recall from Theorem 7.2 that conditions (4) and (5) actually have many equivalent formulations. Similarly, condition (14) is equivalent to analogs on Lq(X, µ) of all the other conditions of Theorem 7.7. If p > 2 then p-standard on span(s1, s2, . . . , sd) can be weakened to kρ(sλ)k ≤ kλkp in (9). (See Lemma 7.10(1).) If p < 2, then p-standard on span(t1, t2, . . . , td) can be weakened to kρ(tλ)k ≤ kλkq in (10). (See Corollary 7.11(2).) CUNTZ ALGEBRAS ON Lp SPACES 45 Finally, it is again essential that the representation be unital. Many of the conditions in Theorem 7.7 actually do make sense for nonunital representations, but the following example shows that they do not imply that the representation is spatial. Example 7.8. Let the notation be as in Example 7.3, except take ρ0 to be a spatial representation of Ld on Lp(X0, µ0). Set ρ(a) = e ⊗ ρ0(a) for a ∈ Ld. Then ρ is not spatial, not even in a sense suitable for nonunital representations. However, it is disjoint, strongly forward isometric, contractive on generators, p-standard on span(s1, s2, . . . , sd), and isometric (although not spatial) on span(cid:0)(sjtk)d It is convenient to break the proof of Theorem 7.7 into several lemmas. Some of them hold in greater generality than needed. The first, in effect, shows that (1) implies all the other conditions. j,k=1(cid:1). Lemma 7.9. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let p ∈ [1, ∞], let (X, B, µ) be a σ-finite measure space, and let ρ : Ld → L(Lp(X, µ)) be a spatial representation. Then: (1) ρ is contractive on generators. (2) ρ is strongly forward isometric. (3) ρ is disjoint. (4) ρ is p-standard on span(s1, s2, . . . , sd) (Definition 7.6(1)). (5) ρ is p-standard on span(t1, t2, . . . , td) (Definition 7.6(2)). rem 7.7(11). (7) For each j there is Xj ⊂ X such that ran(ρ(sj )) = Lp(Xj, µ). (6) (cid:0)ρ(s1) ρ(s2) · · · ρ(sd)(cid:1) is a row isometry, in the sense described in Theo- j,k=1(cid:1) ∼= Md (see (8) If A = Ld with d 6= ∞, the restriction of ρ to span(cid:0)(sjtk)d Lemma 1.11) is a spatial representation of Md in the sense of Definition 7.1. (9) If p 6= ∞, the representation ρ′ of Lemma 2.21 is spatial. Proof. Part (1) follows from Lemma 6.5(2), part (7) follows from Lemma 6.15, and part (9) follows from Lemma 6.21. Part (8) follows from the fact (Lemma 6.17) that the product of spatial partial isometries is again a spatial partial isometry. To prove part (3), let Xj be as in part (7). Suppose µ(Xj ∩ Xk) 6= 0. Then ρ(tj)ρ(sk) is a nonzero spatial partial isometry by Lemma 6.17, contradicting the relation (1.2) or (1.5), as appropriate, in the definition of A. Part (3) follows. We prove (2), (4), and (6) together. Let s =(cid:0)ρ(s1) ρ(s2) · · · ρ(sd)(cid:1) ∈ L(cid:0)Lp(X, µ) ⊕p Lp(X, µ) ⊕p · · · ⊕p Lp(X, µ), Lp(X, µ)(cid:1). Thus s(ξ1, ξ2, . . . , ξd) = ρ(sj )ξj dXj=1 for ξ1, ξ2, . . . , ξd ∈ Lp(X, µ). We already proved that ρ is disjoint, so ρ(s1)ξ1, ρ(s2)ξ2, . . . , ρ(sd)ξd have disjoint supports. Since kρ(sj)ξj k = kξjk by Lemma 6.5(4), it follows from Remark 2.7 that (7.1) ks(ξ1, ξ2, . . . , ξd)kp kρ(sj)ξjkp kξkp p = k(ξ1, ξ2, . . . , ξd)kp p. p =Xj p =Xj 46 N. CHRISTOPHER PHILLIPS This proves (6). Now let λ ∈ Cd and let ξ ∈ Lp(X, µ). In (7.1), take ξj = λj ξ for j = 1, 2, . . . , d, getting kρ(sλ)ξkp λjpkξkp p = kλkp pkξkp p. p =Xj This equation implies both (2) and (4). It remains to prove (5). Let q ∈ [1, ∞] satisfy 1 q = 1. For γ ∈ Cd (possibly d = ∞, in which case we follow the convention of Definition 1.13), define ωγ : Cd → p + 1 C by ωγ(λ) =Pd j=1 γjλj. We show that for γ ∈ Cd, we have kρ(tγ)k ≥ kγkq. Let ε > 0, and choose (using the usual pairing between lp and lq) an element λ ∈ Cd such that kλkp = 1 and ωγ(λ) > kγkq − ε. By part (4) (already proved), we have kρ(sλ)k = 1. Using Lemma 1.14 at the third step, we then get kρ(tγ)k = kρ(tγ)k · kρ(sλ)k ≥ kρ(tγ)ρ(sλ)k = kωγ(λ) · 1k > kγkq − ε. Since ε > 0 is arbitrary, the desired conclusion follows. We now show the reverse. Let γ ∈ Cd and let ξ ∈ Lp(X, µ). Let the disjoint sets Xj be as in (3) (already proved). Set ξj = ξXj . Since ρ(tj) is a spatial partial isometry with domain support Xj, Lemma 6.5(2) gives kρ(tj)ξkp = kξjkp. Now, using Holder's inequality at the third step and Remark 2.7 at the last step, we get kρ(tγ)ξkp ≤ = So kρ(tγ)k ≤ kγkq. dXj=1 dXj=1 γj · kρ(tj)ξkp γj · kξjkp ≤ kγkq(cid:18)Xd j=1 kξjkp p(cid:19)1/p = kγkqkξkp. (cid:3) Lemma 7.10. Let A be any of Ld, Cd, or L∞. Let p ∈ [1, ∞) \ {2}, let (X, B, µ) be a σ-finite measure space, and let ρ : Ld → L(Lp(X, µ)) be a representation. Adopt the notational conventions of Definition 7.6 in case A = L∞. (1) Suppose p > 2. If ρ is forward isometric and for all λ ∈ Cd we have kρ(sλ)k ≤ kλkp, then ρ is disjoint. (2) Suppose p < 2. If ρ is strongly forward isometric and is p-standard on span(s1, s2, . . . , sd), then ρ is disjoint. Proof. For each j, Theorem 6.9 implies that ρ(sj) is semispatial. Let Sj be its semispatial realization and let Xj be its range support. Let ξ ∈ Lp(X, µ). We now claim that, under either set of hypotheses, ρ(sj)ξ and ρ(sk)ξ have essentially disjoint supports for j 6= k. Assume the hypotheses of (1). Let δ1, δ2, . . . , δd be the standard basis vectors in Cd. Use the hypotheses with the choices λ = δj + δk and λ = δj − δk at the first step, and the fact that ρ(sj ) and ρ(sk) are isometric at the second step, to get (7.2) kρ(sj)ξ + ρ(sk)ξkp p + kρ(sj)ξ − ρ(sk)ξkp p ≤ 2kξkp p + 2kξkp p By Corollary 2.1 of [18], we must have in fact p + kρ(sj)ξ − ρ(sk)ξkp kρ(sj)ξ + ρ(sk)ξkp p = 2kρ(sj)ξkp p + 2kρ(sk)ξkp p, = 2kρ(sj)ξkp p + 2kρ(sk)ξkp p. CUNTZ ALGEBRAS ON Lp SPACES 47 and it then follows, by the condition for equality in Corollary 2.1 of [18], that ρ(sj)ξ and ρ(sk)ξ have essentially disjoint supports. Now assume instead the hypotheses of (2). In this case, kρ(sj)ξ + ρ(sk)ξkp p + kρ(sj)ξ − ρ(sk)ξkp p = kδj + δkkp pkξkp p + kδj − δkkp pkξkp p, so we have equality at the first step in (7.2). The condition for equality in Corol- lary 2.1 of [18] therefore implies that ρ(sj)ξ and ρ(sk)ξ have essentially disjoint supports. The claim is proved. Since µ is σ-finite, there is ξ ∈ Lp(X, µ) such that ξ(x) > 0 for all x ∈ X. It follows from the definition of a semispatial partial isometry, Proposition 5.6(6), and Lemma 6.1(1) that ρ(sj)ξ is nonzero almost everywhere on Xj, and similarly that ρ(sk)ξ is nonzero almost everywhere on Xk. Therefore Xj and Xk are essentially disjoint. The conclusion of the lemma follows. (cid:3) Corollary 7.11. Let A be any of Ld, Cd, or L∞. Let p ∈ (1, ∞)\{2}, let (X, B, µ) be a σ-finite measure space, and let ρ : Ld → L(Lp(X, µ)) be a representation. Adopt the notational conventions of Definition 7.6 in case A = L∞. Let q ∈ (1, ∞) \ {2} satisfy 1 q = 1, and let ρ′ : A → L(Lq(X, µ)) be as in Lemma 2.21. p + 1 (1) Suppose p > 2. If ρ is p-standard on span(t1, t2, . . . , td) and ρ′ is strongly forward isometric, then ρ′ is disjoint. (2) Suppose p < 2. If for all λ ∈ Cd we have kρ(tλ)k ≤ kλkq, and ρ′ is forward isometric, then ρ′ is disjoint. Proof. Under the hypotheses of (1), we have q < 2. Also, by duality, for λ ∈ Cd we have kρ′(sλ)k = kρ(tλ)k = kλkq. Therefore ρ′ satisfies the hypotheses of Lemma 7.10(2), with q in place of p, so is disjoint. A similar argument shows that if ρ satisfies the hypotheses of (2), then ρ′ is (cid:3) disjoint by Lemma 7.10(1). Lemma 7.12. Let d ∈ Z>0, let p ∈ [1, ∞) \ {2}, let (X, B, µ) be a σ-finite mea- sure space, and let ρ : Ld → L(Lp(X, µ)) be a representation which is disjoint (Definition 7.4(1)) and forward isometric (Definition 2.12(2)). Then ρ is spatial (Definition 7.4(2)). Proof. Let X1, X2, . . . , Xd ⊂ X be the sets of Definition 7.4(1), so that ran(ρ(sj)) ⊂ j=1 ρ(sj)ρ(tj) = 1, the closed linear span of the ranges of the ρ(sj) is all of Lp(X, µ), so (up to a set of measure zero, which we may ignore) j=1 Xj = X and ran(ρ(sj)) = Lp(Xj, µ). It follows from Lemma 6.16 that ρ(sj) is a spatial isometry for j = 1, 2, . . . , d. Using Lemma 6.12(3), Definition 6.13, and j=1 Xj = X, we see that there is a representation σ : Ld → L(Lp(X, µ)) such that, for j = 1, 2, . . . , d, we have σ(sj ) = ρ(sj) and σ(tj ) is the reverse of ρ(sj). Clearly σ is spatial. Lemma 2.11 implies that ρ = σ. (cid:3) Lp(Xj , µ) for all j. SincePd Sd `d Proof of Theorem 7.7. We begin by observing that (1) implies all the other condi- tions. For (2) and (3), this is trivial. For all the others except (10), this follows from the various conclusions of Lemma 7.9. To get (10), we must also use Lemma 7.9(9) to see that ρ′ is spatial, and then apply Lemma 7.9(2) to ρ′. 48 N. CHRISTOPHER PHILLIPS We now prove that all the other conditions imply (1) (omitting (3), (10), and (14) when p = 1), mostly via one of (2), (6), or (14). That (6) implies (1) is Lemma 7.12. To see that (14) implies (1) for p 6= 1, we apply Lemma 7.9(9) to ρ′ and use (ρ′)′ = ρ. We prove that (2) implies (6) (and hence also implies (1)). For j = 1, 2, . . . , d, let Xj ⊂ X be the range support of ρ(sj). Then ran(ρ(sj )) = Lp(Xj, µ) because ρ(sj) is spatial. Therefore ran(ρ(sjtj)) = Lp(Xj, µ). If now j 6= k, then ρ(sjtj) and ρ(sktk) are idempotents whose product is zero, so Lp(Xj, µ) ∩ Lp(Xk, µ) = {0}. Disjointness of ρ follows, and the other past of (6) is immediate. For p 6= 1, we prove that (3) implies (14). It follow from Lemma 6.21 that ρ′ satisfies (2). We have already proved that (2) implies (1), so we conclude that ρ′ is spatial, which is (14). To prove that (4) and (5) imply (6), we use the implication from (1) to (5) If we start with (5), we also use in Theorem 7.2 to conclude that ρ is disjoint. Remark 2.13. That (7) implies (6) is Remark 2.13, and that (8) implies (6) is clear. That (9) implies (6) follows from Lemma 7.10. We prove that if p 6= 1 then (10) implies (14). It follows from Corollary 7.11 that ρ′ is disjoint. We have already proved that (6) implies (1), so we conclude that ρ′ is spatial, which is (14). We now prove that (11) implies (9). Since we already proved that (9) implies (6), this will show that (11) implies (6). Let λ ∈ Cd and let ξ ∈ Lp(X, µ). Using the assumption at the second step, we get kρ(sλ)ξkp =(cid:13)(cid:13)(cid:0)ρ(s1) ρ(s2) · · · ρ(sd)(cid:1)(cid:0)λ1ξ, λ2ξ, . . . , λdξ(cid:1)(cid:13)(cid:13)p =(cid:13)(cid:13)(cid:0)λ1ξ, λ2ξ, . . . , λdξ(cid:1)(cid:13)(cid:13)p = kλkpkξkp. This equation implies both parts of condition (9). That (12) implies (2) is immediate from Lemma 6.16. To see that (13) implies (2), (cid:3) use in addition Remark 2.13. 8. Spatial representations give isometric algebras In this section, we prove that for d < ∞, any two spatial representations of Ld (in the sense of Definition 7.4(2)) on Lp(X, µ), for the same value of p, give isometrically isomorphic Banach algebras. The main technical tools are the notion of a free representation (Definition 8.1) and the spatial realizations of spatial isometries. Definition 8.1. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let (X, B, µ) be a σ-finite measure space, let p ∈ [1, ∞], and let ρ : A → L(Lp(X, µ)) be a representation. (1) We say that ρ is free if there is a partition X =`m∈Z Em such that for all m ∈ Z and all j, we have ρ(sj )(Lp(Em, µ)) ⊂ Lp(Em+1, µ) and ρ(tj )(Lp(Em, µ)) ⊂ Lp(Em−1, µ). (2) We say that ρ is approximately free if for every N ∈ Z>0 there is n ≥ N m=0 Em such that for m = 0, 1, . . . , n − 1 and all j, and a partition X =`n−1 taking En = E0 and E−1 = En−1, we have ρ(sj )(Lp(Em, µ)) ⊂ Lp(Em+1, µ) and ρ(tj )(Lp(Em, µ)) ⊂ Lp(Em−1, µ). When dealing with approximately free representations, we always take the index m in Em mod n. CUNTZ ALGEBRAS ON Lp SPACES 49 To produce free representations, we follow Lemmas 2.18, 2.19, and 2.20, produc- ing representations of the form (ρ(·) ⊗ 1)1⊗u for suitable u. To simplify the notation and avoid conflict, we will abbreviate this representation to ρu. Lemma 8.2. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let p ∈ [1, ∞). Let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces. Let ρ : A → L(Lp(X, µ)) be a representation, and let u ∈ L(Lp(Y, ν)) be invertible. Then there exists a unique representation ρu : A → L(cid:0)Lp(X × Y, µ × ν)(cid:1) such that for all j we have (using the notation of Theorem 2.16(4)) ρu(sj ) = ρ(sj) ⊗ u and ρu(tj) = ρ(tj) ⊗ u−1. This construction has the following properties: (1) If a ∈ A is homogeneous of degree k (with respect to the Z-grading of Proposition 1.7), then ρu(a) = ρ(a) ⊗ uk. (2) If u is isometric, p 6= 2, and ρ is spatial (Definition 7.4(2)), then ρu is spatial. (3) If there is a partition Y =`m∈Z Fm such that u(Lp(Fm, ν)) = Lp(Fm+1, ν) for all m ∈ Z, then ρu is free in the sense of Definition 8.1(1). Proof. Existence and uniqueness of ρu follow from Lemma 2.17, combined with the appropriate one of Lemmas 2.18, 2.19, and 2.20, taking v = u−1 if A = Cd or L∞. It suffices to prove part (1) when a is a product of generators sj and tj. By ∞ such that a = sαtβ. Lemma 1.10(4), we may assume that there are words α, β ∈ W d Then deg(a) = l(α) − l(β) (by Lemma 1.10(2)), and one checks directly that ρu(sαtβ) = ρ(sαtβ) ⊗ ul(α)−l(β). We prove (2). Lemma 6.16 implies that u is spatial, and Lemma 6.12(4) implies that u−1 is the reverse of u. Now apply Lemma 6.20 to conclude that ρ(sj ⊗ u) is spatial with reverse ρ(tj ) ⊗ u−1. To prove (3), we use the partition X × Y = am∈Z X × Ym. The sets X × Ym clearly have the required properties. (cid:3) The construction of Lemma 8.2 preserves various other properties of representa- tions, but we will not need this information. Proposition 8.3. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let p ∈ [1, ∞), let (X, B, µ) be a σ-finite measure space, and let ρ : A → L(Lp(X, µ)) be a representation. Let u ∈ L(lp(Z)) be the bilateral shift, (uη)(m) = η(m − 1) for η ∈ lp(Z). Let ρu be as in Lemma 8.2. Then for every a ∈ A we have kρu(a)k ≥ kρ(a)k. Proof. Let a ∈ A. Let ε > 0; we show that kρu(a)k ≥ kρ(a)k − ε. Recall the Z-grading of Proposition 1.7. Choose N0 ∈ Z>0 such that there are homogeneous elements a−N0, a−N0+1, . . . , aN0−1, aN0 ∈ A such that deg(ak) = k for all k and a = PN0 that k=−N0 kζkp = 1 and kρ(a)ζk > kρ(a)k − 1 2 ε. ak. Choose ζ ∈ Lp(X, µ) such 50 N. CHRISTOPHER PHILLIPS Set r = kρ(a)ζk. If r ≤ 1 N ∈ Z>0 such that 2 ε, then kρ(a)k < ε, and we are done. Otherwise, choose Let ν be counting measure on Z. We identify Lp(X ×Z, µ×ν) with the space of all p < sequences ξ = (ξm)m∈Z with ξm ∈ Lp(X, µ) for all m and such thatPm∈Z kξmkp ∞, with N > N0rp rp −(cid:0)r − 1 2 ε(cid:1)p . kξmkp p!1/p . kξkp = Xm∈Z ξm =(0 (2N + 1)−1/pζ n > N n ≤ N. Now define ξ ∈ Lp(X × Z, µ × ν) by Then kξkp p = 1. Set η = ρu(a)ξ. We have(cid:2)(1 ⊗ u)ξ(cid:3)m = ξm−1 for all m ∈ Z. Using Lemma 8.2(1), for all m ∈ Z with −N + N0 ≤ m ≤ N − N0 we get ηm = N0Xk=−N0 (2N + 1)−1/pρ(ak)ξm−k = (2N + 1)−1/p N0Xk=−N0 ρ(ak)ζ = (2N + 1)−1/pρ(a)ζ. There are 2N − 2N0 + 1 such values of m. It follows that kηkp p ≥ (2N − 2N0 + 1)(2N + 1)−1kρ(ak)ζkp >(cid:18)1 − N0 N (cid:19) rp > 1 − So p =(cid:18) 2N − 2N0 + 1 ! rp =(cid:0)r − 1 rp −(cid:0)r − 1 2 ε(cid:1)p 2 ε(cid:1)p 2N + 1 rp . (cid:19) rp kηk > r − 1 2 ε = kρ(a)ζk − 1 2 ε > kρ(a)k − ε. This shows that kρu(a)k > kρ(a)k − ε. (cid:3) Corollary 8.4. Let A be any of Ld, Cd, or L∞. Let p ∈ [1, ∞) \ {2}. Then there exists an injective free spatial representation of A on lp(Z>0). Proof. By Lemma 7.5, there is an injective spatial representation ρ of A on lp(Z>0). Let ρu be the representation of A on lp(Z>0) ⊗p lp(Z) ∼= lp(Z>0) of Proposition 8.3. The inequality kρu(a)k ≥ kρ(a)k for all a ∈ A implies that ρu is also injective. Moreover, ρu is spatial by Lemma 8.2(2) and free by Lemma 8.2(3). (cid:3) We want to prove an inequality in the opposite direction from that of Proposi- tion 8.3. We need a lemma. Lemma 8.5. Let (X, B, µ) be a σ-finite measure space, with µ 6= 0. Let d ∈ {2, 3, 4, . . . , ∞}, let X1, X2, . . . , Xd ⊂ X (X1, X2, . . . ⊂ X if d = ∞) be disjoint measurable sets, and for each j let Sj be an injective measurable set transformation CUNTZ ALGEBRAS ON Lp SPACES 51 (Definition 5.4) from (X, B, µ) to (cid:0)Xj, BXj , µXj(cid:1). Then for every n ∈ Z>0 there exists E ∈ B with µ(E) 6= 0 such that the elements Sα(1) ◦ Sα(2) ◦ · · · ◦ Sα(m)([E]) ∈ B/N (µ), for all m ∈ {0, 1, 2, . . . , d} and all words α = (α(1), α(2), . . . α(m)) ∈ W d Notation 1.8) are disjoint in the sense of Definition 4.4. m (see Proof. In this proof, we will write expressions like Sj(E) for measurable subsets E ⊂ X, meaning that Sj(E) is taken to be some measurable subset of X whose image in B/N (µ) is Sj([E]) in the sense of Definition 4.13. We remember that such a set is only defined up to sets of measure zero. Also, disjointness and containment of subsets of X will only be up to sets of measure zero. No problem will arise, because we only deal with countably many subsets of X, and we can therefore make the conclusion exact at the end by deleting a set of measure zero. By analogy with Notation 1.9, for a word α = (α(1), α(2), . . . , α(m)) ∈ W d m, we define Sα = Sα(1) ◦ Sα(2) ◦ · · · ◦ Sα(m), which is a measurable set transformation from (X, B, µ) to a suitable subset of X. We take S∅ = idB/N (µ). As with products of generators of Ld, if α, β ∈ W d ∞ and αβ is their concatenation, then Sα ◦ Sβ = Sαβ. We first claim that for all m ∈ Z>0 and all α, β ∈ W d m, we have Sα(X)∩Sβ(X) = ∅. To see this, let k be the least integer such that α(k) 6= β(k). Set α0 = (α(k), α(k + 1), . . . , α(m)), β0 = (β(k), β(k + 1), . . . , β(m)), and γ = (α(1), α(2), . . . , α(k − 1)). Then γα0 = α and γβ0 = β. The sets Sα0 (X) and Sβ0(X) are disjoint be- cause they are contained in the disjoint sets Xα(k) and Xβ(k). It now follows from Lemma 4.10(2) that (Sγ ◦ Sα0 )(X) and (Sγ ◦ Sβ0)(X) are disjoint, which implies the claim. Our second claim is that if D ⊂ X and n ∈ Z>0 satisfy µ(D) > 0 and D ∩ 1 (D) = ∅, then there exists a subset F ⊂ D such that µ(F ) 6= 0 and such that 1 (F ) = ∅. To prove the 1 (F ) = ∅, then for every 1 (G) = ∅. Thus, if we prove the claim for just Sn for all m ∈ Z>0 such that m divides n, we have F ∩ Sm claim, first observe that if for some fixed m we have F ∩ Sm subset G ⊂ F we also have G ∩ Sm one divisor m of n, an induction argument will yield the claim as stated. Define F = D \ (D ∩ Sm 1 (F ) = ∅. We need only show that µ(F ) > 0. Suppose not. Then (as usual, up to a set of measure zero) we have D ⊂ Sm (D) for all k ∈ Z>0. In particular, D ⊂ Sn 1 (D) = ∅ and µ(D) > 0. The claim is proved. 1 (D). By induction, we get D ⊂ Skm 1 (D), which contradicts D ∩ Sn 1 (D)). Clearly F ∩ Sm 1 X \Sn is an injective measurable set transformation. Since Sn 1 (X) contains X \S1(X), which contains S2(X), and µ(S2(X)) > 0 because S2 1 is an injective measurable Our third claim is that for all n ∈ Z>0, we have µ(cid:0)Sn(X) \ S2n(X)(cid:1) 6= 0. Indeed, set transformation, it follows that µ(cid:0)Sn 1 (X)(cid:1) 6= 0. This proves the claim. (X). Then µ(E0) > 0 by the third claim, 1 (E0) = ∅. The second claim therefore provides a subset E ⊂ E0 Set N = n!. Define E0 = SN 1 (X \ Sn 1 1 (X) \ S2N and also E0 ∩ SN such that µ(E) 6= 0 and such that E ∩ Sm 1 (E) = ∅ for m = 1, 2, . . . , n. 52 N. CHRISTOPHER PHILLIPS We show that E satisfies the conclusion of the lemma. So let α and β be distinct words with length at most n. We have to prove that Sα(E) ∩ Sβ(E) = ∅. Without loss of generality l(α) ≥ l(β). Suppose l(α) = l(β). Then, using the first claim at the second step, Sα(E) ∩ Sβ(E) ⊂ Sα(X) ∩ Sβ(X) = ∅. Suppose now l(α) > l(β). Set m = l(α) and r = l(β). Define a new word γ with l(γ) = m by Thus Sγ = Sβ ◦ Sm−r 1 γ = (β(1), β(2), . . . , β(r), 1, 1, . . . , 1). . We consider two cases, the first of which is γ 6= α. Then, since m − r ≤ n ≤ N, Sα(E) ⊂ Sα(X) and Sβ(E) ⊂ Sβ(SN 1 (X)) ⊂ Sγ(X). So Sα(E) ∩ Sβ(E) = ∅ by the first claim. It remains to consider the case γ = α. Then Sα(E) ∩ Sβ(E) = Sβ(Sm−r 1 (E)) ∩ Sβ(E). Since Sm−r ∅. 1 (E) ∩ E = ∅, it follows from Lemma 4.10(2) that Sα(E) ∩ Sβ(E) = (cid:3) Proposition 8.6. Let d ∈ {2, 3, 4, . . .}, and let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces. Let p ∈ [1, ∞)\{2}. Let ρ : Ld → L(Lp(X, µ)) be an approximately free spatial representation, and let ϕ : Ld → L(Lp(Y, ν)) be a spatial representation. Then for all a ∈ Ld, we have kρ(a)k ≤ kϕ(a)k. Proof. We adopt the same conventions with regard to measurable set transforma- tions as described at the beginning of the proof of Lemma 8.5. In particular, set operations and relations are only up to sets of measure zero. Also, as there, for the measurable set transformations Sj and Rj defined below, we write Sα = Sα(1) ◦ Sα(2) ◦ · · · ◦ Sα(m), and define Rα similarly. By definition, the operators ϕ(sj ) are spatial isometries. Therefore they have spatial systems (Y, Yj, Sj, gj), in which Sj is a bijective measurable set transforma- tion. In particular, ϕ(sj)(Lp(E, µ)) = Lp(Sj(E), µ). By Lemma 7.9(3), the sets Yj are disjoint. Since d < ∞, we get Y =`d Similarly, the operators ρ(sj) have range supports, say, Xj, and X =`d j=1 Xj. Moreover, the spatial realization Rj of ρ(sj) is a bijective measurable set transfor- mation from X to Xj. j=1 Yj. Let a ∈ Ld. We want to prove that kρ(a)k ≤ kϕ(a)k. By scaling, without loss of generality kρ(a)k = 1. Apply Lemma 1.12, obtaining N0 ∈ Z≥0, a finite set F0 ⊂ W d α,β ∈ C for α ∈ F0 and β ∈ W d ∞, and numbers λ(0) N0, such that a = Xα∈F0 Xβ∈W d N0 λ(0) α,βsαtβ. Set N1 = max(cid:0){l(α) : α ∈ F0}(cid:1). Let τ be any fixed word of length N0. Set b = sτ a. Set F = {τ α0 : α0 ∈ F0}. Thus, for all α ∈ F, we have N0 ≤ l(α) ≤ N0 + N1. For CUNTZ ALGEBRAS ON Lp SPACES 53 α ∈ F and β ∈ W d N0 , write α = τ α0 with α0 ∈ F0, and set λα,β = λ(0) α0,β. Then b = Xα∈F Xβ∈W d N0 λα,βsαtβ. Since ρ and ϕ are both contractive on generators (by Lemma 7.9(1)), we have kρ(a)k = kρ(tτ )ρ(sτ a)k ≤ kρ(tτ )k · kρ(sτ a)k ≤ kρ(b)k ≤ kρ(sτ )k · kρ(a)k ≤ kρ(a)k, so kρ(b)k = kρ(a)k = 1, and similarly kϕ(b)k = kϕ(a)k. It therefore suffices to prove that kρ(b)k ≤ kϕ(b)k. Let ε > 0. We prove that kϕ(b)k > 1 − ε. If N0 = N1 = 0, then b is a scalar, and this inequality is immediate. Otherwise, N0 + N1 > 0. Choose r ∈ Z>0 such that r > (N0 + N1)(cid:18) 2 ε(cid:19)p . Choose ξ(0) ∈ Lp(X, µ) such that (cid:13)(cid:13)ξ(0)(cid:13)(cid:13)p = 1 and (cid:13)(cid:13)ρ(b)ξ(0)(cid:13)(cid:13)p > 1 − 1 2 ε. Since ρ is approximately free, there is N ≥ (N0 + N1)r and a partition X = m=0 Dm such that for m = 0, 1, . . . , N − 1 and all j, taking DN = D0 and `N −1 D−1 = DN −1, we have ρ(sj)(Lp(Dm, µ)) ⊂ Lp(Dm+1, µ) and ρ(tj )(Lp(Dm, µ)) ⊂ Lp(Dm−1, µ). Write ξ(0) = with ξ(0) m ∈ Lp(Dm, µ) for m = 0, 1, . . . , N − 1. We claim that there is a set T of N0 + N1 consecutive values of m such that Since the sets Dm are disjoint, Remark 2.7 gives ξ(0) . < ε 2 ξ(0) m N −1Xm=0 m (cid:13)(cid:13)(cid:13)p = Xm∈T(cid:13)(cid:13)ξ(0) m (cid:13)(cid:13)p N0 + N1(cid:19)1/p 1 . p. p ξ(0) (cid:13)(cid:13)(cid:13)Xm∈T (cid:13)(cid:13)(cid:13)Xm∈T m (cid:13)(cid:13)(cid:13) 2(cid:18) (cid:13)(cid:13)ξ(0) m (cid:13)(cid:13)p < ε p It is therefore enough to prove that there is a set T of N0 + N1 consecutive values of m such that for all n ∈ T we have Suppose that there is no such set T. Then, in particular, for k = 0, 1, . . . , r − 1 there is such that nk ∈(cid:2)k(N0 + N1), (k + 1)(N0 + N1)(cid:1) ∩ Z ε 2(cid:18) 1 N0 + N1(cid:19)1/p . (cid:13)(cid:13)ξ(0) nk(cid:13)(cid:13)p ≥ 54 Then contradicting(cid:13)(cid:13)ξ(0) N. CHRISTOPHER PHILLIPS p ≥ 2(cid:17)p(cid:18) p ≥ r(cid:16) ε r−1Xk=0(cid:13)(cid:13)ξ(0) (cid:13)(cid:13)ξ(0)(cid:13)(cid:13)p nk(cid:13)(cid:13)p nk(cid:13)(cid:13)p = 1. This proves the claim. (cid:13)(cid:13)(cid:13)(cid:13)XN0−1 m +XN −1 m=N −N1 ξ(0) m=0 1 N0 + N1(cid:19) > 1, m (cid:13)(cid:13)(cid:13)(cid:13) < ε 2 . ξ(0) By cyclically permuting the indices of the sets Dm, we may assume that Define and 0 ≤ m ≤ N0 − 1 and N − N1 ≤ m ≤ N − 1 m N0 ≤ m ≤ N − N1 − 1 0 ξm =(ξ(0) N −1Xm=0 2 ε, so kρ(b)ξk > 1 − ε. Also clearly kξk ≤(cid:13)(cid:13)ξ(0)(cid:13)(cid:13) = 1. Then(cid:13)(cid:13)ξ − ξ(0)(cid:13)(cid:13) < 1 sentation ψ : Ld → L(cid:0)Lp(X × Y, µ × ν)(cid:1) by ψ(c) = 1 ⊗ ϕ(c) for all c ∈ Ld. It follows from Lemma 2.17 that kψ(b)k = kϕ(b)k. We are now going to define an isometry (not necessarily surjective) Following Lemma 2.17, except with the factors in the other order, define a repre- N −1Xm=N −N1 N0−1Xm=0 ξm = ξ(0) − ξ(0) m − ξ(0) m . ξ = which will partially intertwine ρ and ψ. u : Lp(X, µ) → Lp(X × Y, µ × ν) The bijective measurable set transformations Rj at the beginning of the proof j=1 Xj, and preserve disjointness and finite intersections and unions. Since X =`d identifying, as usual, sets with their images in B/N (µ), we get for m = 0, 1, . . . , N − 2. Define Set Dγ = Rγ(D0) for γ ∈ W. Then Dm+1 = Rj(Dm) W = daj=1 N −1[m=0 X = aγ∈W W d m. Dγ. For γ ∈ W, define eγ = m(χDγ ) ∈ L(Lp(X, µ)) (following Notation 6.11). Then the eγ are idempotents of norm 1 andPγ∈W eγ = 1. Apply Lemma 8.5 to the injective measurable set transformations Sj at the beginning of the proof, obtaining a set E ⊂ Y with ν(E) 6= 0 such that the sets Eγ = Sγ(E), for γ ∈ W, are disjoint. Then ϕ(sγ)(Lp(E, ν)) = Lp(Eγ, ν) for all γ. Choose η0 ∈ Lp(E, ν) such that kη0kp = 1. As in Theorem 2.16, identify Lp(X × Y, µ × ν) with Lp(X, µ) ⊗p Lp(Y, ν). For any ξ ∈ Lp(X, µ) (not just the specific element ξ considered above), define uξ = Xγ∈W ρ(tγ)eγξ ⊗ ϕ(sγ)η0. CUNTZ ALGEBRAS ON Lp SPACES 55 We claim that u is isometric. Let ξ ∈ Lp(X, µ) be arbitrary. Since the sets Dγ are disjoint, Remark 2.7 gives Then u ∈ L(cid:0)Lp(X, µ), Lp(X × Y, µ × ν)(cid:1). p = Xγ∈W kξkp keγξkp p. On the other hand, for γ ∈ W, we have Lp(Dγ, µ) ⊂ ran(ρ(sγ)), so kρ(tγ)eγξkp = keγξkp. Also, the elements ρ(tγ)eγξ ⊗ ϕ(sγ)η0 are supported in the disjoint sets X × Eγ (in fact, in D∅ × Eγ), so (using kϕ(sγ)η0kp = kη0kp = 1 at the third step) kuξkp ρ(tγ)eγξ ⊗ ϕ(sγ)η0(cid:13)(cid:13)(cid:13) p · kϕ(sγ)η0kp p p kρ(tγ)eγξkp p =(cid:13)(cid:13)(cid:13)Xγ∈W = Xγ∈W p = Xγ∈W keγξkp p = kξkp p. This completes the proof that u is isometric. Now set G0 = W d m and G = N1[m=0 N0+N1[m=N0 W d N0(cid:9). m =(cid:8)α0α1 : α0 ∈ G0 and α1 ∈ W d Thus, G is the set of all words with lengths from N0 through N0 + N1, and F ⊂ G. We can therefore write (8.1) b = Xα∈G Xβ∈W d N0 λα,βsαtβ, by taking λα,β = 0 for α ∈ G \ F. We claim that for any ξ ∈ Lp(X, µ) which is supported in N −N1−1[m=N0 [γ∈W d m Dγ, and any b ∈ Ld of the form (8.1) (not just the particular b used above), we have (8.2) uρ(b)ξ = ψ(b)uξ. By linearity, it suffices to prove that for each γ ∈ W with N0 ≤ l(γ) ≤ N −N1−1, the claim holds for all ξ which are supported in Dγ. Write γ = γ0γ1 with γ0 ∈ W d N0 and 0 ≤ l(γ1) ≤ N − N0 − N1 − 1. Then ξ ∈ ran(ρ(sγ)) ⊂ ran(ρ(sγ0 )), so ρ(sγ0)ρ(tγ0 )ξ = ξ. For β ∈ W d tβsγ0 = δβ,γ0 · 1, so N0, we have ρ(b)ξ = Xα∈G Xβ∈W d N0 λα,βρ(sα)ρ(tβ)ρ(sγ0 )ρ(tγ0 )ξ = Xα∈G λα,γ0 ρ(sα)ρ(tγ0)ξ. Since γ = γ0γ1, the element ρ(tγ0)ξ is supported in Dγ1. Let α ∈ G. Since l(α) + l(γ1) ≤ N −1, it follows that ρ(sα)ρ(tγ0 )ξ is supported in Dαγ1 . Therefore (recalling that Notation 1.9 gives tαγ1 = tγ1 tα), uρ(sα)ρ(tγ0 )ξ = ρ(tαγ1 )ρ(sα)ρ(tγ0 )ξ ⊗ ϕ(sαγ1 )η0 = ρ(tγ1 tαsαtγ0 )ξ ⊗ ϕ(sαγ1 )η0 = ρ(tγ)ξ ⊗ ϕ(sαγ1 )η0. 56 Thus N. CHRISTOPHER PHILLIPS uρ(b)ξ = Xα∈G λα,γ0 ρ(tγ)ξ ⊗ ϕ(sαγ1 )η0. On the other hand, using tβsγ0 = δβ,γ0 for β ∈ W d N0 at the third step, ψ(b)uξ = (1 ⊗ ϕ(b))(cid:0)ρ(tγ)ξ ⊗ ϕ(sγ)η0(cid:1) λα,βρ(tγ)ξ ⊗ ϕ(sα)ϕ(tβ)ϕ(sγ0 )ϕ(sγ1 )η0 = Xα∈G Xβ∈W d = Xα∈G N0 λα,γ0 ρ(tγ)ξ ⊗ ϕ(sα)ϕ(sγ1 )η0 = uρ(b)ξ. This completes the proof of the claim. We now return to our specific choices of ξ and b. They satisfy the hypotheses in the claim, so we have, using (8.2) in the second calculation, kuξkp = kξkp ≤ 1 and kψ(b)uξkp = kuρ(b)ξkp = kρ(b)ξkp > 1 − ε. Therefore kϕ(b)k = kψ(b)k > 1 − ε. This completes the proof. (cid:3) Theorem 8.7. Let d ∈ {2, 3, 4, . . .}, let (X, B, µ) and (Y, C, ν) be σ-finite measure spaces, and let ρ : Ld → L(Lp(X, µ)) and ϕ : Ld → L(Lp(Y, ν)) be spatial repre- sentations (Definition 7.4(2)). Then the map ρ(sj) 7→ ϕ(sj) and ρ(tj) 7→ ϕ(tj ), for j = 1, 2, . . . , d, extends to an isometric isomorphism ρ(Ld) → ϕ(Ld). Proof. The statement is symmetric in ρ and ϕ, so it suffices to prove that for all a ∈ Ld, we have kϕ(a)k ≤ kρ(a)k. Let u ∈ L(lp(Z)) be the bilateral shift, and let ϕu : Ld → L(cid:0)Lp(Y, ν) ⊗ lp(Z)(cid:1) be as in Lemma 8.2. Proposition 8.3 implies that kϕ(a)k ≤ kϕu(a)k. Since ϕu is free (by Lemma 8.2(3)), it is clearly essentially free. Since ϕu is spatial (by Lemma 8.2(2)), Proposition 8.6 therefore implies that kϕu(a)k ≤ kρ(a)k. (cid:3) Theorem 8.7 justifies the following definition. Definition 8.8. Let d ∈ {2, 3, 4, . . .} and let p ∈ [1, ∞). We define Op d to be the completion of Ld in the norm a 7→ kρ(a)k for any spatial representation ρ of Ld on a space of the form Lp(X, µ) for a σ-finite measure space (X, B, µ). We write sj and tj for the elements in Op d obtained as the images of the elements with the same names in Ld, as in Definition 1.1. When p = 2, we get the usual Cuntz algebra Od. Indeed, if ρ is a spatial represen- tation on L2(X, µ), then, by Remark 6.14, we have ρ(tj ) = ρ(sj)∗ for j = 1, 2, . . . , d. Thus, ρ is a *-representation. Proposition 8.9. Let d ∈ {2, 3, 4, . . .}, let p ∈ [1, ∞) \ {2}, let (X, B, µ) be a σ-finite measure space, and let ρ : Op d → L(Lp(X, µ)) be a unital contractive homo- morphism. Then ρ is isometric. Proof. Let ρ0 be the composition of ρ with the obvious map Ld → Op d. By The- orem 8.7, it suffices to prove that ρ0 is spatial. We prove this by verifying condi- tion (5) of Theorem 7.7. (By the last part of Theorem 7.7, this suffices when p = 1 as well as when p ∈ (1, ∞)\{2}.) That ρ0 is contractive on generators is immediate. Also, the obvious map sending the standard matrix unit ej,k ∈ Md to sjtk ∈ Op d is isometric from M p d, by Lemma 7.9(8) and the equivalence of conditions d to Op CUNTZ ALGEBRAS ON Lp SPACES 57 (1) and (3) in Theorem 7.2. Therefore the restriction of ρ0 to span(cid:0)(sjtk)d is contractive as a map from M p d to L(Lp(X, µ)). The equivalence of conditions (1) and (4) in Theorem 7.2 therefore shows that this restriction is spatial. This completes the verification of condition (5) of Theorem 7.7. (cid:3) j,k=1(cid:1) Corollary 8.10. Let d ∈ {2, 3, 4, . . .}, let p ∈ [1, ∞)\{2}, let (X, B, µ) be a σ-finite measure space such that Lp(X, µ) is separable, and let ρ : Op d → L(Lp(X, µ)) be a not necessarily unital contractive homomorphism. Then ρ is isometric. Proof. It is clear that e = ρ(1) is an idempotent in L(Lp(X, µ)). Set E = ran(e). Then ρ defines a contractive unital homomorphism from Op d to L(E). The hypotheses imply that kρ(1)k = 1. It follows from Theorem 3 in Section 17 of [17] that there is a measure space (Y, C, ν) such that E is isometrically isomorphic to Lp(Y, ν). Since Lp(X, µ) is separable, so is E, and therefore we may take ν to be σ-finite. Now apply Proposition 8.9. (cid:3) Unfortunately, unlike the case of C*-algebras (p = 2), we know of no result which d from Proposition 8.9 or Corollary 8.10, or the d in [22], using methods much allows us to deduce simplicity of Op other way around. We will prove simplicity of Op closer to those used in the C*-algebra case in [9]. 9. Nonisomorphism of algebras generated by spatial representations for different p To what extent do the various algebras Op d differ from each other? Taking d1 6∼= p = 2, the K-theory computation in [10] shows that, for d1 6= d2, we have O2 O2 . We will show in [23] that the K-theory is the same for p 6= 2 as for p = 2, d2 giving the analogous nonisomorphism result. Here, we address what happens when one instead varies p. Here, K-theory is of no help. Instead, we show by more direct methods that for p1 6= p2 and for d1 and d2 arbitrary, there is no nonzero continuous homomorphism from Op1 . In fact, there is no nonzero continuous d1 homomorphism from Op1 d1 to L(lp2(Z>0)). to Op2 d2 This result gives a different proof of the fact (Corollary 6.15 of [6]) that for dis- tinct p1, p2 ∈ (1, ∞), there is no nonzero continuous homomorphism from L(lp1(Z>0)) to L(lp2(Z>0)). (We are grateful to Volker Runde for providing this reference. In this connection, we note that Corollary 2.18 of [6] implies that if E and F are Banach spaces such that L(E) is isomorphic to L(F ) as Banach algebras, then E is isomorphic to F as Banach spaces.) Lemma 9.1. Let p ∈ [1, ∞). Let (X, B, µ) be a σ-finite measure space. Let ρ : L∞ → L(lp(X, µ)) be a spatial representation. Let E be a Banach space. Suppose there is a nonzero continuous homomorphism ϕ : ρ(L∞) → L(E). Then lp(Z>0) is isomorphic as a Banach space (recall the conventions in Definition 2.2) to a subspace of E. Proof. The element ϕ(1) ∈ L(E) is an idempotent. Replacing E by ϕ(1)E, we may assume that ϕ is unital (and E is nonzero). p + 1 q = 1. Let q ∈ (1, ∞] satisfy 1 Let λ 7→ sλ and λ 7→ tλ be as in Definition 1.13 for L∞. For p ∈ (1, ∞), it follows from Lemma 7.9(4) and Lemma 7.9(5) that the maps λ 7→ ρ(sλ) and λ 7→ ρ(tλ) ex- tend to isometric maps sρ : lp(Z>0) → L(Lp(X, µ)) and tρ : lq(Z>0) → L(Lp(X, µ)). For p = 1, we similarly get sρ as above and tρ : C0(Z>0) → L(Lp(X, µ)). 58 N. CHRISTOPHER PHILLIPS Fix η0 ∈ E with kη0k = 1. Define v : lp(Z>0) → E as follows. For λ = (λ1, λ2, . . .) ∈ lp(Z>0), set v(λ) = ϕ(sρ(λ))η0. Then v is bounded, because kv(λ)k ≤ kϕk · ksρ(λ)k · kη0k = kϕk · kλkp. We claim that for all λ ∈ lp(Z>0), we have kv(λ)k ≥ kϕk−1kλkp. The claim will imply that v is an isomorphism of lp(Z>0) with some closed subspace of E, completing the proof of the lemma. We prove the claim. Let λ ∈ lp(Z>0). Without loss of generality λ 6= 0. First suppose p 6= 1. It is well known that there exists γ ∈ lq(Z>0) \ {0} such that (9.1) and (9.2) γjλj = 1 ∞Xj=1 kγkq = kλk−1 p . (The method of proof can be found, for example, at the beginning of Section 6.5 of [24].) By Lemma 1.14, for every n ∈ Z>0 we have (cid:16)Xn j=1 γjρ(tj )(cid:17)(cid:16)Xn j=1 λjρ(sj)(cid:17) =(cid:16)Xn j=1 γjλj(cid:17) · 1. Letting n → ∞ and applying (9.1), we get tρ(γ)sρ(λ) = 1. Therefore η0 = ϕ(tρ(γ))ϕ(sρ(λ))η0 = ϕ(tρ(γ))v(λ). So, using (9.2) at the third step, 1 = kη0k ≤ kϕk · ktρ(γ)k · kv(λ)k = kϕk · kλk−1 p · kv(λ)k. It follows that kv(λ)k ≥ kϕk−1kλkp, as desired. Now suppose p = 1. Let ε > 0. Choose γ ∈ c0(Z>0) with finite support and such that γjλj ≥ 0 for j ∈ Z>0, γjλj > 1 − ε, and kγkq = kλk−1 p . ∞Xj=1 Then tρ(γ)sρ(λ) = α · 1 with α > 1 − ε. We get αη0 = ϕ(tρ(γ))v(λ) and 1 − ε ≤ kϕk · kλk−1 p · kv(λ)k. Since ε > 0 is arbitrary, we again get kv(λ)k ≥ kϕk−1kλkp, as desired. (cid:3) Theorem 9.2. Let p1, p2 ∈ [1, ∞) be distinct. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let ρ : A → L(lp1(Z>0)) be a spatial representation. Then there is no nonzero continuous homomorphism from ρ(A) to L(lp2(Z>0)). Proof. Suppose that ϕ : ρ(A1) → L(lp2(Z>0)) is a continuous homomorphism. Regardless of what A1 is, there is a unital homomorphism ψ : L∞ → A1 such that (in the notation of Definition 1.3) ψ(cid:0)s(∞) j (cid:1) = sj 2s1 and ψ(cid:0)t(∞) j (cid:1) = t1tj 2 for all j ∈ Z>0. Since ρ is spatial, one easily checks that ρ ◦ ψ is a spatial represen- tation of L∞ on lp1(Z>0). Set B = (ρ ◦ ψ)(L∞). Then Lemma 9.1, applied to ϕB, CUNTZ ALGEBRAS ON Lp SPACES 59 provides an isomorphism of lp1(Z>0) with a subspace of lp2(Z>0). The remark after Proposition 2.a.2 of [19] (on page 54) therefore implies that p1 = p2. (cid:3) Corollary 9.3. Let p1, p2 ∈ [1, ∞) be distinct. Let A1 and A2 be any two of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3). Let ρ1 : A1 → L(lp1(Z>0)) be a spatial representation (Definition 7.4(2)), and let ρ2 : A2 → L(lp2(Z>0)) be an arbitrary representation. Then there is no nonzero continuous homomorphism from ρ(A1) to ρ(A2). Proof. Combine Theorem 9.2 and Lemma 7.5. (cid:3) In particular, there is no nonzero continuous homomorphism from Op1 d1 We recover part of Corollary 6.15 of [6]. to Op2 d2 . Corollary 9.4. Let p1, p2 ∈ [1, ∞) be distinct. Then there is no nonzero continuous homomorphism from L(lp1(Z>0)) to L(lp2(Z>0)). Proof. Suppose ϕ : L(lp1(Z>0)) → L(lp2(Z>0)) is a nonzero continuous homomor- phism. Use Lemma 7.5 to choose an injective spatial representation ρ : L∞ → L(lp1(Z>0)). Then 1 ∈ ρ(L∞), so ϕρ(L∞) is nonzero, contradicting Theorem 9.2. (cid:3) Corollary 9.3 does not rule out isomorphism as Banach spaces. In fact, we have the following result. Proposition 9.5. Let p ∈ [1, ∞), and suppose 1 q = 1. Let A be any of Ld (Definition 1.1), Cd (Definition 1.2), or L∞ (Definition 1.3), let (X, B, µ) be a σ- finite measure space, and let ρ : A → L(Lp(X, µ)) be any representation. Then ρ(A) is isometrically antiisomorphic to a subalgebra of L(Lq(X, µ)), namely ρ′(A) with ρ′ as in Lemma 2.21. p + 1 Proof. Use Lemma 2.21 and the fact that (following Notation 2.3) one always has ka′k = kak. (cid:3) In particular, when ρ is spatial, ρ(A) and ρ′(A) are isometrically isomorphic as Banach spaces, even though (for p 6= 2) they are not even isomorphic as Banach algebras. References [1] G. Abrams, P. Ara, and M. Siles Molina, Leavitt Path Algebras, Springer-Verlag, to appear. [2] G. Abrams and G. Aranda Pino, The Leavitt path algebra of a graph, J. Algebra 293(2005), 319–334. [3] G. Abrams and G. Aranda Pino, The Leavitt path algebras of arbitrary graphs, Houston J. Math. 34(2008), 423–442. [4] G. Abrams and D. Funk-Neubauer, On the simplicity of Lie algebras associated to Leavitt algebras, preprint (arXiv:0908.2179v2 [math.RA]). [5] A. Arias and J. D. Farmer, On the structure of tensor products of lp-spaces, Pacific J. Math. 175(1996), 13–37. [6] E. Berkson and H. Porta, Representations of B(X), J. Funct. Anal. 3(1969), 1–34. [7] J. Chaney, Banach lattices of compact maps, Math. Z. 129(1972), 1–19. [8] P. M. Cohn, Some remarks on the invariant basis property, Topology 5(1966), 215–228. [9] J. Cuntz, Simple C*-algebras generated by isometries, Commun. Math. Phys. 57(1977), 173– 185. [10] J. Cuntz, K-theory for certain C*-algebras, Ann. Math. 113(1981), 181–197. 60 N. CHRISTOPHER PHILLIPS [11] K. R. Davidson, Free semigroup algebras. A survey, pages 209–240 in: Systems, Approxima- tion, Singular Integral Operators, and Related Topics (Bordeaux, 2000), Oper. Theory Adv. Appl. 129, Birkhauser, Basel, 20014. [12] A. Defant and K. Floret, Tensor Norms and Operator Ideals, North-Holland Mathematics Studies no. 176, North-Holland Publishing Co., Amsterdam, 1993. [13] J. L. Doob, Stochastic Processes, John Wiley & Sons, Inc., New York, 1953. [14] T. Figiel, T. Iwaniec, and A. Pe lczy´nski, Computing norms and critical exponents of some operators in Lp-spaces, Studia Math. 79(1984), 227–274. [15] W. B. Johnson and J. Lindenstrauss, Basic concepts in the geometry of Banach spaces, pages 1–84 in: Handbook of the Geometry of Banach Spaces, Vol. 1, W. B. Johnson and J. Lindenstrauss (eds.), Elsevier, Amsterdam, London, New York, Oxford, Paris, Shannon, Tokyo, 2001. [16] S. Koppelberg, Handbook of Boolean algebras. Vol. 1 , J. D. Monk and R. Bonnet (eds), North-Holland, Amsterdam, 1989. [17] H. E. Lacey, The Isometric Theory of Classical Banach Spaces, Grundlehren der mathema- tischen Wissenschaften no. 208, Springer-Verlag, Berlin, Heidelberg, New York, 1974. [18] J. Lamperti, On the isometries of certain function-spaces, Pacific J. Math. 8(1958), 459–466. [19] J. Lindenstrauss and L. Tzafriri, Classical Banach spaces. I. Sequence spaces, Ergebnisse der Mathematik und ihrer Grenzgebiete, vol. 92, Springer-Verlag, Berlin, New York, 1977. [20] W. G. Leavitt, The module type of a ring, Trans. Amer. Math. Soc. 103(1962), 113–130. [21] W. G. Leavitt, The module type of homomorphic images, Duke Math. J. 32(1965), 305–311. [22] N. C. Phillips, Simplicity of UHF and Cuntz algebras on Lp spaces, in preparation. [23] N. C. Phillips, Crossed products of Lp operator algebras and the K-theory of Cuntz algebras on Lp spaces, in preparation. [24] H. L. Royden, Real Analysis, 3rd. ed., Macmillan, New York, 1988. [25] M. Tomforde, Uniqueness theorems and ideal structure for Leavitt path algebras, J. Algebra 318(2007), 270–299. Department of Mathematics, University of Oregon, Eugene OR 97403-1222, USA, and Research Institute for Mathematical Sciences, Kyoto University, Kitashirakawa- Oiwakecho, Sakyo-ku, Kyoto 606-8502, Japan. E-mail address: [email protected]
1211.6048
2
1211
2013-10-19T16:19:11
Local sampling and approximation of operators with bandlimited Kohn-Nirenberg symbols
[ "math.FA", "cs.IT", "math.CA", "cs.IT" ]
Recent sampling theorems allow for the recovery of operators with bandlimited Kohn-Nirenberg symbols from their response to a single discretely supported identifier signal. The available results are inherently non-local. For example, we show that in order to recover a bandlimited operator precisely, the identifier cannot decay in time nor in frequency. Moreover, a concept of local and discrete representation is missing from the theory. In this paper, we develop tools that address these shortcomings. We show that to obtain a local approximation of an operator, it is sufficient to test the operator on a truncated and mollified delta train, that is, on a compactly supported Schwarz class function. To compute the operator numerically, discrete measurements can be obtained from the response function which are localized in the sense that a local selection of the values yields a local approximation of the operator. Central to our analysis is to conceptualize the meaning of localization for operators with bandlimited Kohn-Nirenberg symbol.
math.FA
math
LOCAL SAMPLING AND APPROXIMATION OF OPERATORS WITH BANDLIMITED KOHN-NIRENBERG SYMBOLS FELIX KRAHMER AND G OTZ E. PFANDER Abstract. Recent sampling theorems allow for the recovery of operators with bandlimited Kohn-Nirenberg symbols from their response to a single discretely supported identifier signal. The available results are inherently non-local. For example, we show that in order to recover a bandlimited operator precisely, the identifier cannot decay in time nor in frequency. Moreover, a concept of local and discrete representation is missing from the theory. In this paper, we develop tools that address these shortcomings. We show that to obtain a local approximation of an operator, it is sufficient to test the operator on a truncated and mollified delta train, that is, on a compactly supported Schwarz class function. To compute the operator numerically, discrete measurements can be obtained from the response function which are localized in the sense that a local selection of the values yields a local approximation of the operator. Central to our analysis is to conceptualize the meaning of localization for operators with bandlimited Kohn-Nirenberg symbol. Keywords. Operator identification, pseudodifferential operators, Kohn-Nirenberg symbol, time frequency localization, local approximation, tight Gabor frames. 2010 Mathematics Subject Classification. 47G30, 94A20. Primary 41A35, 94A20; Secondary 42B35, 47B35, 1. Introduction In communications engineering, the effect of a slowly time-varying communication channel is commonly modeled as superposition of translations (time shifts due to multipath propagation) and modulations (frequency shifts caused by Doppler effects). In order to recover transmitted signals from their channel outputs, precise knowledge of the nature of the channel is required. A common procedure for channel identification in this sense is to periodically send short duration test signals. The resulting outputs are then used to estimate channel parameters which allow for an inversion of the operator [15, 2, 16, 27, 1, 14]. Kailath [15] and Bello [2] analyzed the identifiability of such channels. In mathematical terms, the channels considered are characterized by bandlimited Kohn-Nirenberg symbols and the channel identification problem becomes an operator identification problem: can an operator with bandlim- ited Kohn-Nirenberg symbol be identified from the output corresponding to a given test input signal? Kozek and Pfander [16], and Pfander and Walnut [27] gave mathematical proof of the assertions by Kailath and Bello that there exists a suitable test signal as long as the band support of the symbol of the operator has outer Jordan content less than one. The suggested test signals are periodically weighted regularly spaced Dirac-delta distributions as introduced in [27]. In [24], Pfander coined the term operator sampling as the resulting theory has many direct parallels to the sampling theory for bandlimited functions. For example, an operator sampling reconstruction Felix Krahmer is with the University of Gottingen, Institute for Numerical and Applied Mathematics, Lotzestrasse 16-18, 37083 Gottingen, Germany, Tel.: +49 551 39 10584, Fax: +49 551 39 3944, [email protected]. Gotz E. Pfander is with Jacobs University Bremen, School of Engineering and Science, Campus Ring 12, 28759 Bremen, Germany. Tel.: +49 421 200 3211, Fax: +49 421 200 49 3211, [email protected]. Date: June 4, 2018. 1 formula was established which generalizes the reconstruction formula in the classical sampling theorem for bandlimited functions (see [24] and Theorem 2.2 below). As the test signals which appear in the results of [16, 27, 24, 26] decay neither in time nor in frequency, they cannot be realized in practice. In this paper, we show that indeed, for stable identification of operator classes defined by a bandlimitation of the Kohn-Nirenberg symbol, test signals that lack decay in time and frequency are necessary. When seeking to recover only the operator's action on a time-frequency localized subspace, however, this ideal but impractical signal can be replaced with a mollified and truncated copy; the test signal can thereby be chosen to be a compactly supported Schwartz function as shown below. Furthermore, an important difference to the sampling theory for bandlimited functions is that the response to a test signal in operator sampling is a square-integrable function rather than a discrete set of sample values. While there are many ways to discretely represent the response function, the question remains which of the multitude of commonly considered representations allow to recover the operator most efficiently. In the case of a bandlimited function, one feature that distinguishes the representation by samples is locality: a sample is the function value at a given location; due to the smoothness of bandlimited functions it represents the function in the neighborhood of the sampling point. A key consequence of this feature is that it allows to approximate the function in a given region using only samples taken in a fixed-size neighborhood of it. In this paper we develop discrete representations of operators with bandlimited Kohn-Nirenberg symbols that, on the one hand, can be computed in a direct and simple way from the output corresponding to a test signal and, on the other hand, have locality properties analogous to those we appreciate in the classical sampling theory. We work with the same concept of locality as in the localized sampling results mentioned above, namely, locality will be defined through the action of the operator on time-frequency localized functions. Combining the two parts, we obtain that time-frequency measurements of the output corresponding to a truncated and mollified weighted sum of Dirac delta distributions yield a local discrete representation of a bandlimited operator. The paper is organized as follows. In Section 2 we recall operator sampling terminology in some detail and discuss previous results. We then summarize our main results in Section 3. Section 4 provides results on local approximations of operators; in Section 5 we discuss identification using smooth and finite duration test signals, and Section 6 uses Gabor frames to derive our novel discretization scheme for operators with bandlimited Kohn-Nirenberg symbols. 2. Background 2.1. Symbolic calculus. The Schwartz kernel theorem states that every continuous linear oper- ator H : S(R) → S′(R) is of the form Hf (x) =Z κ(x, t)f (t)dt for a unique kernel κ ∈ S′(R2), where S(Rd) is the Schwartz space, and S′(Rd) is its dual, the space of tempered distributions [10]. This integral representation is understood in the weak sense, that is, hHf, gi = hκ, f⊗gi for all f, g ∈ S(R), where f⊗g(x, y) = f (x)g(y) and h· , ·i is the sesquilinear pairing between S(R) and S′(R), and as S(R) continuously embeds into L2(R), Schwartz kernel representations exist in particular for bounded operators on L2(Rd). Each such operator has consequently a spreading function representation (2.1) Hf (x) =ZZ η(t, γ)e2πiγx f (x − t) dt dγ, 2 and a Kohn-Nirenberg symbol representation (2.2) where the Fourier transform bf is normalized by for integrable f . Hf (x) =Z h(x, t)f (x − t)dt, Hf (x) =Z σ(x, ξ)bf (ξ) e2πixξ dξ, F f (ξ) = bf (ξ) =Z f (t) e−2πitξ dt a time-varying impulse response representation We write Hσ and σH , ηH , κH when it is necessary to emphasize the correspondence between H and σ, η, κ. The symbols σ and η are related via the symplectic Fourier transform Fs which is defined densely by Fsσ(t, γ) =ZZ σ(x, ξ)e−2πi(xγ−tξ) dx dξ , AkσkL∞(R2) ≤ kσkL∞(R2) ≤ BkσkL∞(R2) that is, σ = Fsη. For convenience, we use the notation ηH (t, γ) = e2πiγtηH (t, γ) and σH for its symplectic Fourier transform. A straightforward computation shows σH ∗ = σH , where H ∗ denotes the adjoint of H. In our proofs, we shall frequently transition from σ to σ. This does not cause a problem in our analysis since inequality (2.4) below combined with kHkL(L2(R)) = kH ∗kL(L2(R)) shows that for M ⊆ R2 compact there exist A, B > 0 with (2.3) for all Hσ ∈ OP W (M ). 2.2. Sampling in operator Paley Wiener spaces. Following [15, 2, 16, 26, 24], the operators considered in this paper are assumed to have strictly bandlimited Kohn-Nirenberg symbols, that is, they have compactly supported spreading function. Slowly time-varying mobile communications channels may violate this assumption [29]; a more refined model is that the spreading function exhibits rapid decay. Still this suffices to guarantee that truncating the spreading function intro- duces a global error that can be controlled. For example, applying Theorem 2.2 to an operator H whose spreading function has L2 distance at most ǫ to a function supported on a rectangle of area one results in an operator which differs from H differs by at most 2ǫ in operator norm (see [17] for further details). This justifies to restrict to the simpler model of strictly bandlimited symbols. The space of bounded operators whose Kohn-Nirenberg symbols are bandlimited to a given set M -- we will also use the shorthand terminology bandlimited operators -- is called operator Paley-Wiener space1; it is denoted by OP W (M ) = {H ∈ L(L2(R)) : suppFsσH ⊆ M}. The Kohn-Nirenberg symbol of an L2-bounded operator with suppFsσH compact is bounded. In fact, for some A, B > 0 we have, (2.4) where kHkL(L2(R)) is the operator norm of H (Proposition 4.1 below). AkσHkL∞(R) ≤ kHkL(L2(R)) ≤ BkσHkL∞(R) , H ∈ OP W (M ), Certainly, if we have direct access to σH , then some of our approximation theoretic goals can be accomplished using classical two-dimensional sampling results applied to σH . In the model considered here, however, we do not have access to any of the values of the symbol σH of the 1In general terms, operator Paley-Wiener spaces are defined by requiring its members to have bandlimited Kohn- Nirenberg symbols which are in a prescribed weighted and mixed Lp space [24]. For example, to restrict the attention to bandlimited Hilbert-Schmidt operators, we would consider only operators with square integrable symbols. These form a subset of the operators considered in this paper. 3 operator H directly, but we must rely on the operator output Hw which results from applying H to a single test input w. Due to stability consideration, we say that the linear space OP W (M ) is identifiable by w if for A, B > 0 we have (2.5) for all H ∈ OP W (M ) [16]. "Sampling" the operator means that the identifier w in (2.5) is a weighted sequence of Dirac delta distributions, that is, AkHkL(L2(R)) ≤ kHwkL2(R) ≤ BkHkL(L2(R)) , w =Xk∈Z ckδkT , where ck is an appropriately chosen periodic sequence [20, 27, 24]. A guiding paradigm in the sampling theory of operators is the direct analogy to sampling of bandlimited functions. To illustrate this analogy, we compare the classical sampling theorem (often credited to Cauchy, Kotelnikov, Shannon, and Whittaker, among others), Theorem 2.1, with the corresponding result for operators, Theorem 2.2 [24]. Note that Theorem 2.1 formally follows from Theorem 2.2 by choosing the operator H in Theorem 2.2 to be the pointwise multiplication operator f 7→ σ · f [24]. The engineering intuition underlying sampling theorems is that reducing a function to periodic samples at a rate of 1/T samples per unit interval corresponds to a periodization with shift 1/T in frequency space [22]. Thus, as long as T Ω ≤ 1, a function bandlimited to [− Ω 2 ] can be recovered via a convolution with a low-pass kernel, that is, a function φ that satisfies 2 , Ω (2.6) bφ(ξ) =(1/Ω, 0, if ξ ≤ Ω 2 , if ξ ≥ 1 2T . If T Ω = 1, the only such function is the sinc kernel φ(t) = sinc(πt/T ) = sin(πt/T ) . For T Ω < 1, there are many such functions; in particular φ in the Schwartz class is possible. With this notion, the classical sampling theorem reads as follows. Theorem 2.1. For g ∈ L2(R) with suppF g ⊆ [− Ω (2.7) 2 ] and T Ω ≤ 1, we have 2 , Ω πt/T g(x) =Xn∈Z g(nT ) φ(x − nT ) with uniform convergence and convergence in L2(R). Here, φ is any low-pass kernel satisfying (2.6). Recall that every operator H on L2(R) is in one-to-one correspondence with its kernel κH, that is, for a unique tempered distribution κH , we have Hf (x) = R κH (x, y) f (y) dy weakly. In the following, χA denotes the characteristic function of a set A. Theorem 2.2. [24] For H : L2(R) −→ L2(R) with σH ∈ L2(R2), suppFsσH ⊆ [0, T ]×[− Ω and T Ω ≤ 1, we have (2.8) 2 , Ω 2 ], κH (x + t, x) = χ[0,T ](t)Xk∈Z(cid:0)HXn∈Z δnT(cid:1)(t + kT ) φ(x − kT ), with convergence in L2(R2) and uniform convergence in x for each t. Again, φ is any low-pass kernel satisfying (2.6). We point to an important difference between the applicability of Theorems 2.1 and 2.2: Theorem 2.1, a bandlimitation to a large set [− Ω other side, Theorem 2.2 is not applicable if the bandlimiting set [0, T ]×[− Ω than one. Indeed, the the following is known. Pn∈Z cnδnT if M is compact with measure less than 1. If M is open and has area greater than 1, [27, 25] OP W (M ) is identifiable in the sense of (2.5) with appropriate w = in 2 ] can be resolved by choosing a small T ; on the 2 ] has area greater then there exists no tempered distribution w identifying OP W (M ). Theorem 2.3. 2 , Ω 2 , Ω 4 Hence, it is necessary to restrict ourselves to operator Paley-Wiener spaces defined by compact sets M with Lebesgue measure one. For such spaces, one can extend Theorem 2.2 to the following. Theorem 2.4. [26] Let M be compact with Lebesgue measure less than one. Then there exists T, Ω > 0 with T Ω = 1 L , L prime, δ > 0, and L-periodic sequences {cn}n, {bjq}q, j = 0, 1, . . . , n− 1, so that (2.9) κH (x + t, x) = LT L−1Xj=0 r(t − kj T )e2πinjΩxXq∈Z (cid:0)HXn bjq cnδnT(cid:1)(t − (kj − q)T ) φ(x + (kj − q)T ) , H ∈ OP W (M ), where r, φ are Schwartz class functions that satisfy r(t)bφ(γ) = 0 if (t, γ) /∈ (−δ, T + δ) × (−δ, Ω + δ), Xk∈Z r(t − kT ) ≡ 1 ≡Xn∈Zbφ(γ − nΩ) . 3. Main Results Moreover, (2.9) converges in L2(R) and uniformly in x for each t. and (2.10) (3.1) 3.1. Local representations of operators. In classical as well as in operator sampling, working with Schwartz class kernels r,φ is of advantage. Indeed, in the classical sampling theorem, the slow decay of the sinc kernel in (2.7) implies that a small perturbation of just a few coefficients g(nT ) can lead to significant deviations of all values g(t) outside of the sampling grid T Z; this includes values achieved at locations far from the sampling points nT . Hence to approximately recover the function values locally, that is, on an compact interval, it does not suffice to know the function samples in a constant size neighborhood of that interval. When working with Schwartz class kernels, in contrast, such a local approximate reconstruction is possible; one can achieve (cid:12)(cid:12)g(x) − XnT ∈[a,b] g(nT ) φ(x − nT )(cid:12)(cid:12) < ǫ, for all x ∈ [a + d(ǫ), b − d(ǫ)] where the neighborhood size d(ǫ) depends on the approximation level ǫ but not on the interval [a, b]. A corresponding possibility of using local information for local reconstruction is not given in Theorem 2.2. Moreover, the identifier w =Pn∈Z δnT neither decays in time or in frequency, clearly showing that in practice this input signal is not usable. However, in the framework of Theorem 2.2, this is unavoidable, as we show in the following theorem. Theorem 3.1. If the tempered distribution w identifies OP W ([0, T ]×[−Ω/2, Ω/2]), T Ω > 0, then w decays weakly neither in time nor in frequency, that is, we have neither for all Schwartz class functions ϕ. hw, ϕ(· − x)i x→±∞−→ 0 nor ξ→±∞ −→ 0 hbw, ϕ(· − ξ)i We address this problem by developing a concept of "local recovery" of an operator, in analogy to the local recovery of a function in (3.1). Indeed, the key to most results presented in this paper is to aim only for the recovery of the operator restricted to a set of functions "localized" on a prescribed set S in the time-frequency plane. This is indeed reasonable in communications where band and time constraints on transmitted signals are frequently present. In [14], for example, operators that map bandlimited input signals to finite duration output signals are considered. Bivariate Fourier series expansions of such an operator's compactly supported Kohn-Nirenberg symbol allow the authors to discretize the a-priori continuous input-output relations (2.2) and (2.1). 5 Our definition of function localization in time and frequency is based on Gabor frames. It involves translation and modulation operators, These operators are unitary on L2(R) and isomorphisms on all function and distribution spaces considered in this paper. Ttf : f 7→ f (· − t) and Mν : f 7→ e2πiν(·)f. For any g ∈ L2(R) and a, b > 0, we say that the Gabor system (g, aZ × bZ) = {TkaMℓbg}k,ℓ∈Z is a tight frame for L2(Rd) if for some A > 0, the so-called frame bound, we have f = A Xk,ℓ∈Z hf,Tka Mℓbgi Tka Mℓbg for all f ∈ L2(Rd). Each coefficient in this expansion can be interpreted to reflect the local behavior of the function near the indexing point in time-frequency space. Hence, a natural way to define time-frequency localized functions is that all but certain expansion coefficients are small. Definition 3.2. Let (g, aZ × bZ), g ∈ S(R), be a tight frame for L2(R) with frame bound 1. We say that f ∈ L2(R) is ǫ -- time-frequency localized on the set S if X(ka,ℓb)∈S hf,MℓbTka gi2 ≥ (1 − ǫ2) X(ka,ℓb)∈R2 hf,MℓbTka gi2 . for S ⊆ R2. Our next result states that a sufficient condition for two operators to approximately agree on functions ǫ -- time-frequency localized on a set S is that their Kohn-Nirenberg symbols almost agree on a neighborhood of S. Below, B(r) denotes the Euclidean unit ball with radius r and center 0; the dimension will always be clear from the context. For brevity of notation, we set kσHkL∞(R2), kσ eHkL∞(R2) ≤ µ and kσH − σ eHkL∞(S) ≤ ǫ µ S − B(r) =(cid:0)Sc + B(r)(cid:1)c Theorem 3.3. Fix M compact and let (g, aZ × bZ), g ∈ S(R), be a tight frame for L2(R) with frame bound 1. Then any pair of operators H, eH ∈ OP W (M ) for which one has on a set S ⊆ R2 satisfy for all f ∈ L2(R) that are ǫ -- time-frequency localized on S − B(cid:0)d(ǫ)(cid:1) in the sense of Definition 3.2. Here C > 0 is an absolute constant and d : (0, 1) −→ R+ satisfies d(ǫ) = o( kp1/ǫ) for all k ∈ N as Our next main result concerns truncated and mollified versions of the identifierPn cnδnT and provides localized versions of Theorems 2.2 and 2.4. For S = R2, it reduces to Theorems 2.2 and 2.4. Theorem 3.4. Fix M compact with Lebesgue measure µ(M ) < 1 and let (g, aZ × bZ), g ∈ S(R), be a tight frame for L2(R) with frame bound 1. Let S ⊆ I1× I2 ⊆ R2, where I1 and I2 may coincide kHf − eHfkL2(R) ≤ C ǫ µkfkL2(R) A generalization of Theorem 3.3 -- labeled Theorem 4.2 -- is proven in Section 4. ǫ → 0. with R. Furthermore, choose the tempered distribution ϕ such that ϕ ≥ 0 and bϕ ≡ 1 on I2 and let Then for any H ∈ OP W (M ) with kσHkL∞(R2) ≤ µ and eH ∈ OP W (M ) defined via cnϕ(· − nT ). (3.2) κ eH(x + t, x) = LT ew = XnT ∈I1 r(t − kjT )(cid:16)Xq∈Z bjqHew(t − (kj − q)T )φ(x + (kj − q)T )(cid:17) e2πinj Ωx, 6 L−1Xj=0 one has kHf − eHfkL2(R) ≤ C ǫ µkfkL2(R) Here C > 0 is an absolute constant, r and φ are Schwartz class functions defined as in Theorem 2.4, but for δ > 0 such that µ(M + [−3δ, 3δ]2) < 1, and d : (0, 1) −→ R+ is a function independent of for all f ∈ L2(R) that are ǫ -- time-frequency localized on S − B(cid:0)d(ǫ)(cid:1) in the sense of Definition 3.2. S which satisfies d(ǫ) = o( kp1/ǫ) for all k ∈ N as ǫ → 0. For rectangular bandlimitation domains M = [0, T ]×[− Ω PnT ∈I1 ϕ(· − nT ) and define eH via the formula κ eH (x + t, x) = TXn∈Z(cid:0)H XnT ∈I1 ϕ(· − nT )(cid:1)(t + nT ) φ(x − nT ) . 2 ] one can choose the identifier 2 , Ω Note that this theorem is completely analogous to the condition (3.1) for localized function sampling. Due to the two-dimensional nature of the operator, however, localization is an issue in both time (restricting to a finite number of deltas) and frequency (replacing the deltas by approximate identities). If one is interested in localization only in time or only in frequency, one can choose one of the Ii to be R and thus consider ew = XnT ∈I1 ew =Xn cnδnT or cnϕ(· − nT ), again with (cn) ≡ 1 in case of rectangular domains M . 3.2. Local sampling of operators. An additional important structural difference between classi- cal sampling and operator sampling remains: in Theorems 2.2 and 3.4, the reconstruction formulas (2.8) and (2.9) involve as "coefficients" functions, not scalars. Among the many possibilities to discretely represent the operator's response to the identifier w, we consider Gabor representations of this sample function. A time-frequency localized subset of the coefficients will then yield a cor- responding local approximation of the operator. Theorem 3.5 below establishes a reconstruction formula based on Gabor coefficients that allows for the exact recovery of the operator; Theorem 3.6 shows that a local subset of the coefficients yields a local approximation of the operator. Again, one obtains considerably simpler formulas for rectangular domains, but for reasons of brevity, we focus on the comprehensive setup of arbitrary domains. For a Schwartz class function φ and a tempered distribution f on R we call Vφf (x, ξ) = hf,MξTxφi, x, ξ ∈ R, the short-time Fourier transform of f with respect to the window function φ. Throughout this paper, all pairings h·,·i are taken to be linear in the first component and antilinear in the second. Theorem 3.5. For M compact with Lebesgue measure µ(M ) < 1 there exists L prime, δ > 0, T, Ω > 0 with T Ω = 1/L, and L-periodic sequences {cn}n, {bjq}q, j = 0, . . . , L − 1, so that for H ∈ OP W (M ), (3.3) σH (x, ξ) = LT β1β2 e−2πi(xnjΩ+ξkj T )e2πinj Ωkj T L−1Xj=0 Xm,ℓ∈Z where σ(j) m,ℓ =Xq∈Z β1 σ(j) m,ℓ Vφr(cid:0)x−(cid:0) mL bjq φ(cid:0)(−q − kj − mL/β1)T(cid:1) hHXn 7 + kj(cid:1)T, ξ−(cid:0) ℓL β2 + nj(cid:1)Ω(cid:1), cnδnT , TqTMℓΩL/β2 ri, (3.4) and r, φ are Schwartz class functions such that r and bφ are real valued and satisfy2 bφ(γ) = 0 if γ /∈ (−δ − Ω/2, δ + Ω/2), r(t) = 0 if t /∈ (−δ, δ + T ), and (3.5) Xk∈Z r(t + kT )2 ≡ 1 ≡Xn∈Zbφ(γ + nΩ)2, Observe that the reconstruction formulas given in Theorems 2.4 and 3.4 require the functions r with oversampling rates β2 ≥ 1 + 2δ/T and β1 ≥ 1 + 2δ/Ω.3 and bφ to generate partitions of unity (3.5), while (2.10) above requires that the functions obtained by taking the square of the modulus form partitions of unity. Theorem 3.6. Fix M compact with µ(M ) < 1, let T, Ω, L and w, r, φ be defined in Theorem 2.4, and let (g, aZ × bZ), g ∈ S(R), be a tight frame for L2(R) with frame bound 1. Then H ∈ OP W (M ) with kσHkL∞(R2) ≤ µ, and eH ∈ OP W (M ) defined via its symbol +kj(cid:1)T, ξ−(cid:0) ℓL eσ(x, ξ) = m,ℓ Vφr(x−(cid:0) mL LT β1β2 L−1Xj=0 e−2πi(xnj Ω+ξkj T )e2πinj Ωkj T X(mLT /β1,ℓLΩ/β2)∈Seσ(j) m,ℓ =Xq∈Z eσ(j) bjq φ(cid:0)(−q − kj − mL/β1)T(cid:1) hHew, TqTMℓΩL/β2 ri, kHf − eHfkL2(R) ≤ C ε µkfkL2(R) for all f ∈ L2(R) which are ǫ -- time-frequency localized on S − B(d(ǫ)) with respect to (g, aZ × bZ) in the sense of Definition 3.2. Again S ⊆ I1 × I2 ⊂ R2 is given, ϕ and ew are defined as in Theorem 3.4, C > 0, and d can again be chosen independent of S with d(ǫ) = o( kp1/ǫ) for all k ∈ N as ǫ → 0. β2 +nj(cid:1)Ω(cid:1), where satisfy β1 The discrete representations introduced in Theorems 3.5 and 3.6 resolve a fundamental concep- tual difference between classical sampling and operator sampling. In contrast to classical sampling, where the sampling values can be extracted individually, the contributions of the different Dirac- deltas in the operator sampling formula are combined in a single function and cannot easily be separated. Hence, while choosing a higher sampling rate in the function case yields more informa- tion, in the operator case, this additional information is mixed in an inseparable way. These aliasing effects [16] make it impossible to obtain redundant representations merely by oversampling in The- orem 2.2 or Theorem 2.4. In reconstruction formula (3.3), however, the oversampling parameters βi can be chosen arbitrarily, allowing for representations of arbitrarily large redundancy. This interplay of large redundancy and good local representation properties of the discrete coef- ficients can be exploited to coarsely quantize bandlimited operators, i.e., to represent these samples by values from a finite alphabet which allow for approximate recovery via the same reconstruction formulas as in Theorems 3.5 and 3.6. For such methods, as they have been studied in the math- ematical literature for frame expansions over Rn [3, 4, 19] or the space of bounded bandlimited functions on R [6, 13, 7], the possibility to oversample is of crucial importance. We will, however, leave this to future work. 2For example, we can choose r = χ[0,T ) ∗ϕδ , where ϕδ is an approximate identity, that is, a non-negative function with ϕδ ∈ S(R), supp ϕδ ⊆ [−δ/2, δ/2], and R ϕδ = 1. 3Then the Gabor systems {rk,l = TkT Mℓ/β2T r}k,ℓ∈Z, {TnΩMm/β1Ω bφ}m,n∈Z, and {Φm,−n,l,−k = T(mT L/β1,ℓLΩ)M(nΩ,/β2,kT )}m,n,k,ℓ∈Z are tight Gabor frames with A = β2/T , A = β1/Ω, and A = β1β2/(T Ω) = β1β2L, respectively, whenever β2 ≥ 1 + 2δ/T and β1 ≥ 1 + 2δ/Ω. 8 (4.2) if 1 ≤ p < ∞ and Akfkp M p(Rd) ≤Xλ∈Λ hf, π(λ)gip ≤ Bkfkp M p(Rd), f ∈ M p(Rd) 4. Local approximation of bandlimited operators In this section we show that a local approximation of an operator's symbol always yields a local approximation of the operator in the sense of Definition 3.2. The given results are of general interest and will be stated in more general terms than other results in this paper. This does not increase the difficulty of proof, but necessitates to recall additional terminology from time-frequency analysis. s(Λ) denotes the set For that, recall that for any full rank lattice Λ = AZ2d ⊆ R2d, det A 6= 0, ℓp of sequences (cλ)λ∈Λ for which kckℓp s (Λ) =(cid:16)Xλ∈Λ(cid:12)(cid:12)(kλks ∞ + 1) cλ(cid:12)(cid:12)p(cid:17)1/p < ∞. A time-frequency shift by λ = (t, ν) ∈ Λ is denoted by π(λ) = MνTt and in the following we will consider Gabor systems of the form (g, Λ) = {π(λ)g}λ∈Λ. Among the many equivalent definitions of modulation spaces, we choose the following. Let g0(x) = e−kxk, 1 ≤ p ≤ ∞ and s ∈ R. Then (4.1) M p s (Rd) = k(hf, π(λ)g0i)λkℓp s ( 1 2 Z2d) < ∞}, s (Rd) = {f ∈ S′(Rd) : kfkM p where we generally omit the subscript s = 0. For details on modulations paces, see, for example, [11, 8]. In the following we shall use the fact that whenever (g, Λ) is a tight L2-Gabor frame (see below for a precise definition) with g ∈ M 1(Rd) then replacing the L2-Gabor frame (g0, 1 Z2d) in (4.1) with (g, Λ) leads to an equivalent norm on M p(Rd) [11]. That is, there exist positive constants A and B with 2 AkfkM ∞(Rd) ≤ sup λ∈Λ hf, π(λ)gi ≤ BkfkM ∞(Rd), f ∈ M ∞(Rd) if p = ∞. In either case, we call (g, Λ) an ℓp-frame with lower frame bound A and upper frame bound B. If we can choose A = B in case of p = 2 then we call (g, Λ) a tight Gabor frame. The norm equivalence (2.4) follows from the following result since M 2(R) = L2(R). Theorem 4.1. Let 1 ≤ p ≤ ∞ and M compact. Then there exist positiv constants A = A(M, p) and B = B(M, p) with AkσHkL∞(R2) ≤ kHkL(M p(R)) ≤ B kσHkL∞(R2), H ∈ OP W (M ). Proof. Theorem 2.7 in [24] (see for example the proof of Theorem 3.3 in [24]) provides C = C(M, p) with kHfkM p(R) ≤ C kσHkL∞(R2) kfkM p(R) In addition, we shall use the following facts. for all H ∈ OP W (M ). This establishes the existence of B = B(M, p) above. In [9, 11] it is shown that the operator norm of an operator mapping the modulation space M 1(R) into its dual M ∞(R) is equivalent to the M ∞(R2) norm of its kernel κ, which can easily shown to be equivalent to the M ∞(R2) norm of the time-varying impulse response h. Moreover, we use the fact that M ∞(R2) is invariant under Fourier transforms (in some or all variables) and that the M ∞(R2) norm can be replaced by the L∞(R2) norm if we restrict ourselves to functions bandlimited to a fixed set M [21, 24]. Last but not least, we use that the identity map embedding M p(R) into M q(R), p ≤ q, is bounded. Writing . to express that A ≤ CB for some constant C depending only on the support M and A ≍ B to denote equivalence in norms, i.e., A . B and B . A, we obtain for all H ∈ OP W (M ) kσHkL∞(R2) ≍ kσHkM ∞(R2) ≍ khHkM ∞(R2) ≍ kκHkM ∞(R2) ≍ kHkL(M 1(R),M ∞(R)) . kHkL(M p(R)) and the result follows. (cid:3) 9 We proceed to prove the following generalization of Theorem 3.3. Indeed, the earlier stated result follows again from the fact that L2(R) = M 2(R) and g ∈ S(R) implies g ∈ M 1 s (R) for all s ≥ 1. We focus on the case of arbitrary domains; a simpler proof for rectangular domains can be obtained using Theorem 2.2 instead of Theorem 2.4. Theorem 4.2. Fix M compact and p ∈ [1,∞]. Let (g, Λ), g ∈ M 1 L2(R) with frame bound 1.Then any H ∈ OP W (M ) with s (R), s ≥ 1, be a tight frame for kσHkL∞(R2) ≤ µ and kσHkL∞(S) ≤ ǫ µ, satisfies for all f ∈ M p(R) time-frequency localized on S − B(cid:0)d(ǫ)(cid:1) =(cid:16)Sc + B(cid:0)d(ǫ)(cid:1)(cid:17)c for p < ∞, kHfkM p(R) ≤ C ǫ µkfkM p(R) Xλ∈Λ∩(S−B(d(ǫ))) hf, π(λ)gip ≥ (1 − ǫp) Xλ∈Λ hf, π(λ)gip , in the sense that, or, for p = ∞, sup(cid:8)hf, π(λ)gi, λ ∈ Λ ∩(cid:0)S − B(d(ǫ))(cid:1)} ≥ (1 − ǫ) sup(cid:8)hf, π(λ)gi, λ ∈ Λ(cid:9) . Proof. Step 1. Preliminary observations and choice of auxiliary objects. Here C > 0 is an absolute constant and d : (0, 1) −→ R+ is a function independent of S which satisfies d(ǫ) = o(ǫ−1/s) as ǫ → 0. φ ∈ S(R2) withR φ(x) dx = 1 and supp φ ⊆ [− 1 is called dual lattice of the lattice Λ in R2. Let eΛ be a lattice containing Λ with the property that there exists a compact and convex fundamental domain D ofeΛ⊥ which contains M + [− 1 e2πihµ,λi = 1 for all λ ∈ Λ} Λ⊥ = {µ ∈ R2 : L2(R2) F (χD ∗ φ) 2 ]2. Recall that Choose a nonnegative 2 , 1 2 ]2. Set 2 , 1 and, using the sampling theorem for lattices in Rn [23, 11], we obtain for all H ∈ OP W (M ) σP = kχD ∗ φk−1 σH =Xλ∈ eΛ H =Xλ∈ eΛ σH (λ) TλσP σH (λ) π(λ)P π(λ)∗. and hence (4.3) (4.5) As explained above, the fact that (g, Λ) is a Gabor frame in L2(R) with g ∈ M 1(R), implies that it is also an ℓp-frame for M p(R) and there exists C1, C2 > 0 with kfkM p(R) ≤ C1 k{hf, π(λ)gi}λ∈Λkℓp(Λ) ≤ C1C2 kfkM p(R), (4.4) As the synthesis map is the adjoint of the analysis map, we also have f ∈ M p(R). (cid:13)(cid:13)Xλ∈Λ cλπ(λ)g(cid:13)(cid:13)M p(R) ≤ C2 k{cλ}λ∈Λkℓp(Λ). ℓ=1(Λ + µℓ) for some µ1, µ2, . . . , µn. Here n is finite, as otherwise the set would be dense, hence not a discrete lattice, and depends only on M and (g, Λ). It is easily seen that (g, Λ + µℓ), ℓ = 1, . . . , n, also satisfies (4.4) and (4.5). Setting Since Λ is a subgroup of eΛ, we have eΛ = Sn eg = n−1/2 g ∈ M 1 2 − 1 1 frame bounds equal 1 and an ℓp-frame with for M p(R) with s (R), we conclude that the Gabor system (eg,eΛ) is a tight frame for L2(R) with kfkM p(R) ≤ C1 n p k{hf, π(eλ)egi}eλ∈ eΛkℓp(eΛ) 10 1 2 − 1 p C2 n 1 p − 1 2 kfkM p(R) = C1C2 kfkM p(R), f ∈ M p(R). ≤ C1 n We claim that (4.6) nhP π(λ)eg, π(eλ)egio ∈ ℓ1 s(eΛ ×eΛ) . s (R2) To see this, recall that σP ∈ S(R2) ⊆ M 1 as e2πixξ is a Fourier multiplier and hence also a time multiplier for M 1 s (R2) (Lemma 2.1 in [12], related results can be found in [5, 30, 31, 32]). A direct computation implies that for λ = (t, ν) s (R2), and, hence,eσP given by σP (x, ξ) e2πixξ is in M 1 andeλ = (et,eν) we have Equation (4.2) implies that the right hand side defines an ℓ1 hP π(t, ν)eg, π(et,eν)egi =(cid:12)(cid:12)(cid:12)ZZ σP (x, ξ) e2πixξ \MνTteg(ξ)MeνTeteg(x) dξ dx(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)ZZ σP (x, ξ) e2πixξ M−tTνbeg(ξ)MeνTeteg(x) dξ dx(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)heσP ,M(eν,t)T(et,ν)eg⊗begi(cid:12)(cid:12). and (g⊗beg, eΛ ×eΛ) is a Gabor frame with window eg⊗beg in M 1 Sk = Xk(λ,eλ)k∞=k hP π(λ)eg, π(eλ)egi, we have {(k + 1)sSk} ∈ ℓ1(N). That is, {Sk} = o(k−(s+1)) and for some C > 0 we haveP∞ {Sk}k∈N0 defined by C K −s, K ∈ N. For ǫ > 0 set d(ǫ) = (C/ǫ)1/s and observe that then s (R2) s (R2). Hence, (4.6) holds and for s(eΛ ×eΛ) sequence since fσP ∈ M 1 k=K Sk ≤ Xeλ∈ eΛ Xλ∈ eΛ∩B(d(ǫ))c(cid:12)(cid:12)(cid:12)hP π(λ)eg, π(eλ)egi(cid:12)(cid:12)(cid:12) ≤ ∞Xk=d(ǫ) Sk ≤ C(cid:0)(C/ǫ)1/s(cid:1)−s = ǫ. Now, set A(λ,eλ) = hP π(λ)eg, π(eλ)egi if λ ∈eΛ ∩ B(d(ǫ))c and 0 else. Step 2. Decomposing Hf as Hf = Hinfin + Houtfin + Hfout. We set and Λin = Λ ∩(cid:0)S − B(d(ǫ))(cid:1), fin = Xλ∈Λin eΛin =eΛ ∩(cid:0)S − B(d(ǫ))(cid:1), Λout = Λ \ Λin, hf, π(λ)gi π(λ)g = Xλ∈ eΛin fout = f − fin, cλ π(λ)eg, where cλ = √nhf, π(λ)gi if λ ∈ Λ and 0 else. Similarly, inspired by (4.3), we set for H ∈ OP W (M ) eΛout =eΛ \eΛin. Hin = Xλ∈ eΛ∩S σH (λ) π(λ)P π(λ)∗, Hout = H − Hin and note that Hin, Hout ∈ OP W (D + [− 1 Step 3. Bounding kHoutfinkM p(R). We use the separation ofeΛin andeΛ∩ Sc by d(ǫ) to compute 2 ]2). 2 , 1 hHoutfin, π(eλ)egi =(cid:12)(cid:12)(cid:12)h Xν∈ eΛ∩Sc ≤ Xν∈ eΛ∩Sc ≤ Xν∈ eΛ∩Sc σH (ν)π(ν)P π(ν)∗ Xλ∈ eΛin σH (ν) Xλ∈ eΛin σH (ν) Xλ∈ eΛin cλπ(λ)eg, π(eλ)egi(cid:12)(cid:12)(cid:12) cλ(cid:12)(cid:12)(cid:12)hπ(ν)P π(ν)∗π(λ)eg, π(eλ)egi(cid:12)(cid:12)(cid:12) cλ(cid:12)(cid:12)(cid:12)hP π(λ − ν)eg, π(eλ − ν)egi(cid:12)(cid:12)(cid:12) 11 σH (ν) Xλ∈ eΛin ≤ Xν∈ eΛ∩Sc ≤ kσHkL∞(R2)Xν∈ eΛ Xλ∈ eΛ cλ A(λ − ν,eλ − ν) cλ A(λ − ν,eλ − ν). (cid:12)(cid:12)h{hHoutfin, π(eλ)egi}eλ∈ eΛ,{deλ}eλ∈ eΛi(cid:12)(cid:12) For every sequence {dλ} ∈ ℓq(eΛ), 1/p + 1/q = 1, we conclude ≤ kσHkL∞(R2)Xeλ∈ eΛXν∈ eΛ Xλ∈ eΛ cλ A(λ − ν,eλ − ν)deλ = kσHkL∞(R2)Xλ∈ eΛXeλ∈ eΛXν∈ eΛ cλ+ν A(λ,eλ)deλ+ν ≤ kσHkL∞(R2)k{cλ}kℓp(eΛ)k{dλ}kℓq(eΛ)Xλ∈ eΛXeλ∈ eΛ p C1k{hHoutfin, π(eλ)egi}kℓp(eΛ) p C1kσHkL∞(R2)k{cλ}kℓp(eΛ)Xλ∈ eΛXeλ∈ eΛ 2 hf, π(λ)gi}kℓp(Λ) ǫ ≤ n1− 1 kHoutfinkM p(R) ≤ n ≤ n A(λ,eλ) p C1µk{n ≤ n and 1 2 − 1 1 2 − 1 1 2 − 1 1 A(λ,eλ) p C1 C2 ǫ µkfkM p(R). Step 4. Bounding kHfoutkM p(R). By Proposition 4.1 we have kHfoutkM p(R) ≤ B(M, p)kσHkL∞(R2)kfoutkM p(R). By hypothesis, for p < ∞ we have hf, π(λ)giπ(λ)gkp hf, π(λ)gip 2 Xλ∈Λout M p (R) ≤ Cp 2 ǫpkfkp ≤ C2p hf, π(λ)gip M p(R) , kfoutkp M p(R) = k Xλ∈Λout 2 ǫpXλ∈Λ ≤ Cp kfoutkM ∞(R) = k Xλ∈Λout and for p = ∞ we have hf, π(λ)giπ(λ)gkM ∞ (R) ≤ C2 khf, π(λ)gikℓ∞(Λout) ≤ C2 ǫkhf, π(λ)gikℓ∞(Λ) ≤ C2 2 ǫkfkM ∞(R) . We conclude kHfoutkM p(R) ≤ B(M, p) C2 ǫ kσHkL∞kfkM p(R) ≤ B(M, p) C2 ǫ µkfkM p(R). Step 5. Bounding kHinfinkM p(R). ℓ∞(Λ) → L∞(R2), Since σP ∈ S(R2), the operator cλ TλσP {cλ} 7→Xλ∈ eΛ is bounded, say with operator norm bound C3. Then, Proposition 4.1 implies 2 ]2, p)kσHinkL∞(R2)kfinkM p(R) kHinfinkM p(R) ≤ B(D+[− 1 2 , 1 2 ]2, p) C3 k{σH(λ)}kℓ∞(eΛ∩S)(1 + ǫ)kfkM p(R) 2 , 1 ≤ B(D+[− 1 2 ]2, p) C3 ǫ µkfkM p(R). ≤ 2 B(D+[− 1 2 , 1 12 Since all constants are independent of ǫ, µ, H, and f , we summarize kHfkM p(R) = kHinfin + Houtfin + HfoutkM p(R) ≤ Cǫ µkfkM p(R) . 5. Operator identification using localized identifiers (cid:3) This section analyzes identifiers that are localized in time and frequency. Theorem 3.1 shows that such functions cannot serve as an identifier for the entire Paley-Wiener. Proof of Theorem 3.1. Let r 6= 0 be a Schwartz function with supp r ⊆ [0, T ] and φ 6= 0 be a Schwartz function with suppbφ ⊆ [−Ω/2, Ω/2]. Let Hn be defined via its kernel κn(x, y) = φ(x − n)r(x − y), so hn(x, t) = φ(x − n)r(t) and ηn(t, ν) =R hn(x, t)e−2πixν dx = r(t)e2πinνbφ(ν), so Hn ∈ OP W ([0, T ]×[−Ω/2, Ω/2]) with kσHnkL∞(R2) = kbrkL∞(R) kφkL∞(R). If w identifies OP W ([0, T ]×[−Ω/2, Ω/2]), then by definition Hnw ∈ L2(R). Then Z Hnw(x)2 dx =Z hκn(x, y), w(y)iy2 dx =Z φ(x − n)2 hr(x − y), w(y)iy2 dx. Clearly, hr(x − y), w(y)iy x→±∞−→ 0 would imply kHnwkL2(R) n→±∞−→ 0 and contradict identifiability (2.5) since by (2.4) we have kHnkL(L2(R)) ≥ AkσHnkL∞(R2) = AkbrkL∞(R) kφkL∞(R) for all n ∈ Z. 2 , Ω To show that an identifier w cannot decay in frequency, we choose Hn ∈ OP W ([0, T ]×[− Ω 2 ]) to have spreading functions ηn(t, ν) = r(t)e2πintbφ(ν)e−2πitν . Let g be a Schwartz function and compute using Fubini's Theorem and, for notational simplicity, using bilinear pairings in place of sesquilinear ones, and we can conclude as above. (cid:3) We proceed by showing that local identification of operators is possible with identifiers localized both in time and frequency, Theorem 3.4. Proof of Theorem 3.4. The proof proceeds in two steps. First we show that replacing each Dirac-delta by a suitable smoothed out version locally introduces only a small error and identifi- cation using the resulting smooth identifier can be interpreted as sampling a modified bandlimited operator. Second we show that reducing to a finite number of samples also locally yields only a small error. Applying this to the modified operator arising in the first part proves that both reductions together also yield only a small error. For the first part, choose ϕ ∈ S with supp ϕ ⊆ [−δ, δ], kbϕkL∞(R) = 1, and bϕ(ξ) − 1 ≤ ǫ for ξ ∈ I2. Define Cϕ : f 7→ f ∗ ϕ and set HC = H ◦ Cϕ. Observe that HC f (x) =ZZ ηH (t, ν)e2πixν f ∗ ϕ(x − t) dt dν =ZZZ ηH (t, ν)e2πixν f (x − t − y)ϕ(y) dy dt dν =ZZZ ηH (t − y, ν)e2πixνf (x − t)ϕ(y) dy dt dν =ZZ (cid:16)Z ηH (t − y, ν)ϕ(y) dy(cid:17) e2πixνf (x − t)dt dν , 13 hHnw(x), g(x)ix =(cid:10)ηn(t, ν),he2πixν w(x − t), g(x)ix(cid:11)t,ν =(cid:10)r(t)e2πintbφ(ν),he2πi(x−t)ν w(x − t), g(x)ix(cid:11)t,ν =(cid:10)r(t)e2πint w(x − t) g(x), hbφ(ν), e2πi(x−t)νiν(cid:11)t,x =(cid:10)hr(t)e2πint, w(x − t) φ(x − t)it, g(x)(cid:11)x =(cid:10)hbr(ξ − n), e−2πixξ bw ∗bφ(ξ)iξ, g(x)(cid:11)x =(cid:10)br(ξ − n)bw ∗bφ(ξ),bg(ξ)(cid:11)ξ . L2(R) =Z br(ξ − n)2 hbw(ξ − ν), bφ(ν)iν2 dξ , kHnwk2 L2(R) = k[Hnwk2 Hence, that is, ηHC (t, ν) = ηH (·, ν) ∗ ϕ(t) and supp ηHC ⊆ supp ηH + [−δ, δ]×{0}. We can apply Theorem 2.4 for the operator HC with M1 := M + [−δ, δ]×{0} in place of M . As by assumption M1 + [−δ, δ]2 still has measure less than one, this can be done with δ, r and φ as given in the theorem. Defining w1 := ϕ ∗ w, we obtain (5.1) κHC (x + t, x) = LT Observe that bjqHw1(t − (kj − q)T )φ(x + (kj − q)T )(cid:17) e2πinj Ωx. r(t − kjT )(cid:16)Xq∈Z L−1Xj=0 σHC (x, ξ) = FsηHC (x, ξ) = σH (x, ξ)bϕ(ξ), and, by hypothesis, we have kσHCkL∞(R2) ≤ kσHkL∞(R2) ≤ µ and kσH − σHCkL∞(S) ≤ ǫµ. Note that for I1 = R, (5.1) agrees with (3.2) and we have HC = eH, so this establishes the result. nonnegative and satisfy Pn ψ(x − nT ) = 1 and suppbψ ⊂ [−1/T, 1/T ]. Such a function can be Set PA(x) =PnT ∈A ψ(x−nT ), so P[−N,N ] → 1 and P[−N,N ]c → 0 uniformly on compact subsets For the second part, let us assume S ⊆ I1 × R and M1 ⊂ [c, d] × R. Let ψ ∈ S(R) be obtained by choosing an arbitrary bandlimited, nonnegative ψ0 ∈ S with kψ0kL1 = 1 and defining ψ = χ[0,T ] ∗ ψ0. as N → ∞. Moreover, PA(x) ≤ 1 for all A. Choose N (ǫ) so that PI1+[−N (ǫ),N (ǫ)](x) − 1 ≤ ǫ for x ∈ I1 + [c, d] and choose R(ǫ) with (5.2) kP[I1+[−N (ǫ),N (ǫ)](x) Vφ∗ r(x − q, ξ)kL1(R2) < ǫ(1 − ǫ)D. XqT /∈I1+[−R(ǫ),R(ǫ)] where the nature of D is derived by the computations below. The existence of such R(ǫ) follows from the fact that PI1+[−N (ǫ),N (ǫ)](x) and Vφ∗ r decay faster than any polynomial. Furthermore, as ψ ∈ S, a similar argument to the one given in the proof of 4.2 shows that for both R(ǫ) and N (ǫ), the growth rate is again bounded by o( kp1/ǫ for arbitrarily large kN. Let w2 = PkT ∈I1+[−R(ǫ),R(ǫ)]+[−δ,T +δ] ckδkT and observe that eH as defined in the theorem bjqHC w2(t − (kj − q)T )φ(x + (kj − q)T )(cid:17) e2πinj Ωx. r(t − kjT )(cid:16)Xq∈Z h eH (x + t, t) = κ eH (x + t, x) L−1Xj=0 satisfies = LT ckδkT (x) = HC Since M1 ⊂ [c, d] × R, we have supp HC δy ⊆ [c + y, d + y], and therefore, kT ∈I1+[−R(ǫ),R(ǫ)]+[−T −δ,δ]+[c,d] HCw(x) = HCXk∈Z Note that eH ∈ OP W (M2), where M2 = M1 + [−δ, δ]2 (for details, see, for example, [26]). As M2 + [−δ, δ]2 still has measure less than one, this implies that we can apply Theorem 2.4 again with the same δ. We obtain hHC (x + t, t) − h eH (x + t, t) x ∈ K ≡I1 + [−R(ǫ), R(ǫ)] + [−T − δ, δ] . ckδkT (x) = HC w2(x), X =LT =LT L−1Xj=0 L−1Xj=0 r(t − kjT )(cid:16)Xq∈Z r(t − kjT )(cid:16) XqT /∈K−(t−kj T ) bjqHC(cid:0)w − w2(cid:1)(t − (kj − q)T ) φ(x + (kj − q)T )(cid:17) e2πinj Ωx bjqHC(cid:0)w − w2(cid:1)(t − (kj − q)T ) φ(x + (kj − q)T )(cid:17) e2πinj Ωx 14 =LT L−1Xj=0 r(t − kjT )(cid:16) XqT /∈I1+[−R(ǫ),R(ǫ)] bjqHC(cid:0)w − w2(cid:1)(t − (kj − q)T ) φ(x + (kj − q)T )(cid:17) e2πinj Ωx. Setting eK = K c + [−δ, T + δ] and using that (σHC (x, ξ) − σ eH (x, ξ)) PI1+[−N (ǫ),N (ǫ)](x) is ban- dlimited to M + {0}×[−1/T, 1/T ]), we compute kσHC − σ eHkL∞(S) ≤ 1/(1 − ǫ)k(σHC (x, ξ) − σ eH (x, ξ)) PI1+[−N (ǫ),N (ǫ)](x)kL∞(R2) ≍ 1/(1 − ǫ)k(σHC (x, ξ) − σ eH (x, ξ)) PI1 +[−N (ǫ),N (ǫ)](x)kM ∞(R2) ≍ 1/(1 − ǫ)k(hHC (x, t) − h eH (x, t)) PI1+[−N (ǫ),N (ǫ)](x)kM ∞(R2) r(t − kjT ) e2πinjΩ(x−t) ≍ 1/(1 − ǫ)(cid:13)(cid:13)(cid:13)LT PI1+[−N (ǫ),N (ǫ)](x) L−1Xj=0 ≤ LT /(1 − ǫ) ≤ LT /(1 − ǫ) XqT /∈I1+[−R(ǫ),R(ǫ)] L−1Xj=0 bjqHC (w − w2)(t − (kj − q)T ) φ(x − t + (kj − q)T )(cid:13)(cid:13)(cid:13)M ∞(R2) XqT /∈I1+[−R(ǫ),R(ǫ)](cid:13)(cid:13)(cid:13) PI1+[−N (ǫ),N (ǫ)](x) r(t − kjT ) e2πinjΩ(x−t) bjqHC(w − w2)(t − (kj − q)T ) φ(x − t + (kj − q)T )(cid:13)(cid:13)(cid:13)M ∞(R2) XqT /∈I1+[−R(ǫ),R(ǫ)](cid:13)(cid:13)(cid:13)HC (w − w2)(t − (kj − q)T )(cid:13)(cid:13)(cid:13)M ∞(R2) L−1Xj=0 (cid:13)(cid:13)(cid:13) PI1+[−N (ǫ),N (ǫ)](x) r(t − kj T ) e2πinjΩ(x−t) bjq φ(x − t + (kj − q)T )(cid:13)(cid:13)(cid:13)M 1(R2) bjq(cid:13)(cid:13)(cid:13)PI1+[−N (ǫ),N (ǫ)](x) r(t) φ(x − t − qT )(cid:13)(cid:13)(cid:13)M 1(R2) L−1Xj=0 XqT /∈I1+[−R(ǫ),R(ǫ)] LT 1 − ǫ , ≤ kHCkL(M ∞(R)) kw − w2kM ∞(R) where we used the invariance of the M ∞ and M 1 norm under translation and modulation and, for the last inequality, Theorem 4.1 -- noting that, for functions constant in one of the coordinate directions, the M ∞(R) and M ∞(R2) norms agree. The second to last inequality is based on M 1(R2) being a Banach algebra, namely on kg1g2kM 1(R2) ≤ kg1kM 1(R2)kg2kM 1(R2) for g1, g2 ∈ M 1(R2). Indeed, for f ∈ M ∞(R2) and g ∈ M 1(R2), we have kf gkM ∞(R2) = ≤ sup sup k ef kM 1 (R2)=1hf g,efi = k ef kM 1 (R2)=1kfkM ∞(R)kefkM 1(R2)kgkM 1(R2) = kfkM ∞(R2)kgkM 1(R2) . k ef kM 1(R2 )=1hf,ef gi ≤ k ef kM 1(R2 )=1kfkM ∞(R)kef gkM 1(R2) sup sup Note that with φ∗(t) = φ(−t), we have Z r(t)φ(x − t)e−2πitξdt = Vφ∗ r(x, ξ), which is a bandlimited function since ZZ Vφ∗ r(x, ξ)e2πitξ−xν dx dξ =Z r(t)φ(x − t)e−2πixνdx = r(t)bϕ(ν) e−2πitν . Using that the M 1-norm is invariant under partial Fourier transforms and the equivalence between the M 1 and L1 norms which is implied by the bandlimitation of PI1+[−N (ǫ),N (ǫ)](x + q) Vφ∗ r(x, ξ) 15 to (−1/T, 1/T )×{0} + (−δ, Ω + δ)×(−δ, T + δ), we obtain (cid:13)(cid:13)(cid:13)PI1+[−N (ǫ),N (ǫ)](x + q) r(t) φ(x − t)(cid:13)(cid:13)(cid:13)M 1(R2) ≍(cid:13)(cid:13)(cid:13)PI1+[−N (ǫ),N (ǫ)](x + q) Vφ∗ r(x, ξ)(cid:13)(cid:13)(cid:13)M 1(R2) ≍(cid:13)(cid:13)(cid:13)PI1+[−N (ǫ),N (ǫ)](x + q) Vφ∗ r(x, ξ)(cid:13)(cid:13)(cid:13)L1(R2) Fix g ∈ S(R) and observe that kVgfkLp(R2) defines a norm on M p(R) equivalent to the M p(R) norm given in (4.1) [11]. For any A ⊂ R we obtain the uniform bound . k XnT ∈A cnδnTkM ∞(R) ≍ kVg XnT ∈A ≤ k XnT ∈A cnδnTkL∞(R) = k XnT ∈A cng(nT − t)kL∞(R) ≤ kXn∈Zcng(nT − t)kL∞(R) < ∞. cng(nT − t)e2πiνnTkL∞(R) The first norm inequality stems from the fact that for all g ∈ M 1(R), kVgfkLp(R2) defines a norm on M p(R) equivalent to the M p(R) norm given in (4.1). Combining this upper bound on kw − w2kM ∞(R) with the above estimate for kσHC − σ eHkL∞(S) and (5.2), we conclude kσHC − σ eHkL∞(S) . DkHCkL(M ∞(R)) L2T 1 − ǫ kbjqkℓ∞ XqT /∈I1+[−R(ǫ),R(ǫ)](cid:13)(cid:13)(cid:13)PI1+[−N (ǫ),N (ǫ)](x + q) Vφ∗ r(x, ξ)(cid:13)(cid:13)(cid:13)L1(R2) ≤ DǫkHCkL(M ∞(R)) ≍ DǫkσHCkL∞(R2) ≤ DǫkσHkL∞(R2) ≤ Dǫµ. Choosing R(ǫ) above large to yield D small enough to compensate all the multiplicative constants, we obtain As a meaningful statement is only obtained for ǫ < 1, this bound directly implies that kσHC − σ eHkL∞(S) ≤ ǫµ. Combining this with the bound kσ eHkL∞(R2) ≤ 2µ. kσH − σ eHkL∞(R2) ≤ kσH − σHCkL∞(R2) + kσHC − σ eHkL∞(R2) ≤ 2ǫµ, Theorem 3.3 directly yields the result with a constant of twice the size as in Theorem 3.3. (cid:3) 6. Reconstruction of bandlimited operators from discrete measurements This section concerns the discrete representation given in Theorem 3.5. First, we prove this theorem, hence establishing that indeed this representation is globally exact. Proof of Theorem 3.5: The proof is similar to the proof of Theorem 2.4 given in [26]. The main idea is to use a Jordan domain argument to cover a fixed compact set M of size less than one by shifts of a rectangle that still have combined area less than one and then to combine identifiability results for each of them to obtain identifiability for the whole set. Indeed, there exist L prime and T, Ω > 0 with T Ω = 1 L such that supp(η) ⊆ L−1[j=0 R + (kjT, njΩ) ⊆ [−(L − 1)T /2, (L + 1)T /2] × [−LΩ/2, LΩ/2] where R = [0, T )× [−Ω/2, Ω/2), and the sequence (kj, nj) ∈ Z2 consists of distinct pairs. For δ > 0 small enough (and possibly slightly smaller T, Ω, and a larger prime L), one can even achieve = [−1/(2Ω) + T /2, 1/(2Ω) + T /2] × [−1/(2T ), 1/(2T )] Mδ ⊆ L−1[j=0 R + (kjT, njΩ) ⊆ [−(L − 1)T /2, (L + 1)T /2] × [−LΩ/2, LΩ/2] 16 where Mδ is the δ-neighborhood of M . Fix such δ and let r, φ ∈ S(R) satisfy (3.4) and (3.5) for this δ. Clearly, a fact that we shall use below. (6.1) (k, n) 6= (kj , nj) for all j implies Sδ∩(cid:16)R+(kT, nΩ)(cid:17) = ∅ and η(t, γ)r(t−kT )bφ(γ−nΩ) = 0, Define the identifier w =Pn∈Z cn δnT , where {cn} is L-periodic and observe that Hw(x) =ZZ η(t, γ) e2πiγxw(x − t) dt dγ =ZZ η(t, γ) e2πiγ(x−t)Xk∈Z ckZ η(x − kT, γ) e2πiγkT dγ L−1Xk=0 for any p ∈ Z. We shall use the non-normalized Zak transform ZLT : L2(R) −→ L2(cid:0)[0, LT ) × [−Ω/2, Ω/2)(cid:1) defined by ck+pZ η(x − (mL + k + p)T, γ) e2πiγ(mL+k+p)T dγ =Xk∈Z = Xm∈Z ckδkT (x − t) dt dγ f (t − nLT ) e2πinLT γ . ZLT f (t, γ) =Xn∈Z We compute using the Poisson summation formula and the fact that Ω = 1/LT = e2πiT nLν (ZLT ◦ H)w(t, ν) =Xn∈Z Hw(t − nLT ) e2πinLT ν ck+pZ η(t − (nL + mL + k + p)T, γ) e2πiγ(mL+k+p)T dγ L−1Xk=0 e2πiT nLνZ η(t − (mL + k + p)T, γ) e2πiγT ((m−n)L+k+p)dγ = Xm,n∈Z L−1Xk=0 ck+p Xm,n∈Z ck+pXm∈ZZ η(t − (mL + k + p)T, γ) e2πiγ(mL+k+p)TXn∈Z L−1Xk=0 ck+pXm∈ZZ η(t − (mL + k + p)T, γ) e2πiγ(mL+k+p)T 1 L−1Xk=0 LT Xn∈Z L−1Xk=0 ck+p Xm,n∈Z η(t − (mL + k + p)T, ν + nΩ) e2πi(ν+Ωn)(mL+k+p)T e2πinL(ν−γ)T dγ = Ω = = δn/LT (ν − γ)dγ By (6.1) we get for p = 0, . . . , L − 1, = Ω L−1Xk=0 L−1Xj=0 r(t)bφ(ν)η(t − (mL + k)T, ν + nΩ)e2πiT (ν+nΩ)(mL+k+p) r(t)bφ(ν)(ZLT ◦ H)w(t + pT, ν) ck+p Xm,n∈Z cp+kj r(t)bφ(ν)η(t + kjT, ν + njΩ) e2πi(ν+nj Ω)T (p+kj ). (T kj M nj c)p(cid:16)e2πiνkj T r(t)bφ(ν) η(t + kjT, ν + njΩ)(cid:17) , L−1Xj=0 = Ωe2πiνpT = Ω 17 where here and in the following, T : (c0, c1, . . . , cL−2, cL−1) 7→ (cL−1, c0, . . . , cL−3, cL−2) and M : (c0, c1, . . . , cL−2, cL−1) 7→ (e2πi0/Lc0, e2πi1/Lc1, . . . , e2πi(L−2)/LcL−2, e2πi(L−1)/LcL−1), that is, (T kj M nj c)p = e2πi cp+kj . Equivalently, we obtain the matrix equation nj (p+kj ) L (6.2) [e−2πiνpT r(t)bφ(ν)(ZLT ◦ H)w(t + pT, ν)]L−1 = ΩA[e2πiνkj T r(t)bφ(ν)η(t + kjT, ν + njΩ)]L−1 p=0 j=0 where A is a L × L matrix, whose jth column is T kj M nj c ∈ CL. A is a submatrix of the L × L2 marix G, whose columns are {T kM lc}L−1 k,l=0. It was shown in [18] that if L is prime, then we can choose c ∈ CL such that every L × L submatrix of G is invertible. In fact, the set of such c ∈ CL is a dense open subset of CL [18]. Hence we can apply the matrix A−1 =: [bjp]L j,p=1 on both sides of Equation (6.2) to obtain (6.3) e2πiνkj T r(t)bφ(ν)η(t + kjT, ν + njΩ) = LT bjpe−2πiνpT r(t)bφ(ν)(ZLT ◦ H)w(t + pT, ν) for every j = 0, 1, . . . , L − 1. In fact, until this point the proof agrees with the proof of (2.9) in Theorem 2.4. Indeed, if we extend {bjp}p to a L-periodic sequence by setting bj,p+mL = bjp, replace the so far unused property (3.5) by (2.10) then further computations [24] give L−1Xp=0 h(x, t) = LT L−1Xj=0 r(t − kjT )(cid:16)Xq∈Z bjqHw(t − (kj + q)T )φ(x − t + (kj + q)T )(cid:17)e2πinj Ω(x−t). Observe that (3.5) implies that (r, T Z × ΩL whenever β2 ≥ 1+2δ/T as, in this case, (r, β2 and the Ron-Shen criterion applies [11, 28]. The same arguments imply that (bφ, ΩZ × LT Z) = {TkTMℓLΩ/β2r}k,ℓ∈Z is a tight Gabor frame Z× 1 Z) = (r, β2T Z×ΩLZ) is an orthogonal sequence Z) is a tight Gabor frame. Using a simple tensor argument, we obtain that {Ψm,n,l,k}m,n,l,k∈Z forms a tight Gabor frame where ΩL β2 β1 T Ψm,n,l,k(t, ν) = T(kT,nΩ)ML(ℓΩβ2T,T /β1) r⊗bφ(t, ν) = e2πiL( mT (ν−nΩ) + ℓΩ(t−kT ) β1 β2 ) r(t − kT ) bφ(ν − nΩ) . The frame bound is T ΩL2T Ω/(β1β2) = 1/(β1β2). We set Φm,−n,l,−k = FsΨm,n,l,k. Clearly, as Fs is unitary, we have that {Φm,n,l,k}m,n,l,k∈Z forms a tight frame with frame bound 1/(β1β2), in fact, a tight Gabor frame as Φm,n,l,k(x, ξ) = FsΨm,−n,l,−k(x, ξ) = (FT−kT MℓLΩ/β2r)(ξ) (F −1T−nΩMmT L/β1bφ)(x) = (MkTTℓLΩ/β2br)(ξ) (MnΩTmT L/β1φ)(x) = e2πi(nm+kl)/λ(TℓLΩ/β2MkTbr)(ξ) (TmT L/β1MnΩφ)(x). Note that (6.1) together with the fact that the symplectic Fourier transform is unitary implies that the coefficients in the Gabor frame expansion of σ satisfy hσ, Φm,−nj ,l,−kji = hη, Ψm,n,l,ki = 0 unless (n, k) = (nj, kj) for some j. Hence we need to estimate σ(j) m,ℓ = hσ, Φm,−nj,l,−kji for j = 0, 1, . . . , L − 1. We obtain by (6.3) σ(j) m,ℓ = hσ, Φm,−nj ,l,−kji = hη, Ψm,nj,l,kji + β1 ℓΩ(t−kj T ) T m(ν−nj Ω) =ZZ η(t, ν)e−2πiL( =ZZ r(t) bφ(ν)η(t + kjT, ν + njΩ)e2πiνkj T e−2πi(L( mνT ) r(t − kjT ) bφ(ν − njΩ)dtdν + ℓtΩ β2 18 β2 β1 )+νkj T dtdν = LT = LT =ZZ LT L−1Xp=0 L−1Xp=0 L−1Xp=0 = LTXq∈Z = LTXq∈Z = LTXq∈Z = LT + ℓtΩ β2 )+νkj T )dtdν + ℓtΩ β2 )+νkj T )dtdν β1 β1 L−1Xp=0 bjpe−2πiνpT r(t)bφ(ν)(ZLT ◦ H)w(t + pT, ν) e−2πi(L( mνT bjpZZ r(t)bφ(ν) e−2πiνpT (ZLT ◦ H)w(t + pT, ν) e−2πi(L( mνT bjpZZ r(t)bφ(ν) e−2πiνpTXq∈Z bjpXq∈Z(cid:16)Z r(t)Hw(t + pT − qLT )e−2πiL ℓtΩ bjq(cid:16)Z r(t)Hw(t + qT )e−2πiL ℓtΩ bjqφ(T (−q − kj − mL/β1))(cid:16)Z Hw(t)e−2πiL ℓΩ(t−qT ) β2 β2 dt(cid:17)(cid:16)Z bφ(ν) e2πiνT (−q−kj −mL/β1) dν(cid:17) Hw(t + pT − qLT ) e2πiνqLT e−2πi(L( mνT β1 r(t − qT )dt(cid:17) bjqφ(T (−q − kj − mL/β1)) hHw,TqT MℓLΩ/β2 ri, β2 dt(cid:17)(cid:16)Z bφ(ν) e2πiνT (qL−p−kj −mL/β1) dν(cid:17) + ℓtΩ β2 )+νkj T dtdν where bjq = bjq′ for q = mL + q′ with q′ = 0, 1, . . . , L − 1. We can hence set Cq,l(Hw) = hHw,TqTMℓLΩ/β2 ri. To sum up, hσ, Φm,−nj,l,−kjiΦm,−nj ,l,−kj (x, ξ) σ(x, ξ) = = (6.4) where 1 β1β2 LT β1β2 L−1Xj=0 Xm,ℓ∈Z L−1Xj=0 e−2πi(xnjΩ+ξkj T ) Xm,ℓ∈Z m,ℓ =Xq∈Z σ(j) σ(j) m,ℓ br(cid:0)ξ − ℓLΩ β2 (cid:1) φ(cid:0)x − mT L β1 (cid:1) , bjq φ(a(−q − kj − mL/β1)) Cq,l(Hw). Applying the symplectic Fourier transform to (6.4) yields η(t, ν) = e−2πiνt η(t, ν) = e−2πiνt LT β1β2 L−1Xj=0 Xm,ℓ∈Z σ(j) m,ℓ Fs (cid:16)M(−njΩ,−kj T ) T( mT L β1 , ℓLΩ β2 )φ⊗br(cid:17)(t, ν) = e−2πiνt LT β1β2 L−1Xj=0 Xm,ℓ∈Z σ(j) m,ℓ T(kj T,−nj Ω) M( ℓLΩ β2 ,− mT L β1 ) r⊗bφ (t, ν) β1 β2 = = ,− mT L LT β1β2 LT β1β2 σ(j) m,ℓ T(kj T,−nj Ω) M( ℓLΩ σ(j) m,ℓ e2πinj Ωkj T T(kj T,−nj Ω) M( ℓLΩ L−1Xj=0 Xm,ℓ∈Z L−1Xj=0 Xm,ℓ∈Z FsU (x, ξ) =ZZ r(t)bφ(ν)e−2πiνte−2πi(ξt−νx) dν dt For U (t, ν) = r⊗bφ (t, ν) e−2πiνt, we have 19 ) (cid:16)r⊗bφ (t, ν) e−2πi(ν+nj Ω)(t−kj T )(cid:17), −nj Ω, kj T − mT L β1 β2 ) (cid:16)r⊗bφ (t, ν) e−2πiνt(cid:17). =Z r(t)φ(x − t)e−2πiξt dt =Z r(t)φ(t − x)e−2πiξt dt = Vφr(x, ξ), where we used that bφ real valued implies φ(y) = φ(−y). Now, we compute σ(x, ξ) = Fsη (x, ξ) = LT β1β2 = LT β1β2 (6.5) = LT β1β2 σ(j) m,ℓ e2πinj Ωkj T Fs (cid:16)T(kj T,−nj Ω) M( ℓLΩ σ(j) m,ℓ e2πinj Ωkj T M(−nj Ω,−kj T ) T( mT L L−1Xj=0 Xm,ℓ∈Z L−1Xj=0 Xm,ℓ∈Z L−1Xj=0 e−2πi(xnj Ω+ξkj T )e2πinj Ωkj T Xm,ℓ∈Z β1 β2 −nj Ω, kj T − mT L β1 ) U(cid:17) (x, ξ), −kj T, ℓLΩ β2 −nj Ω) Vφr (x, ξ), σ(j) m,ℓ Vφr(x − mT L β1 + kj T, ξ − ℓLΩ + njΩ β2 (cid:1). The convergence in (6.4) and (6.5) is defined in the weak sense, but can be shown to converge absolutely and uniformly on compact subsets. (cid:3) Next we prove Theorem 3.6, that is, the direct local correspondence between the discretization values and the operator action. Proof of Theorem 3.6. We intend to apply Theorems 3.3 and 3.5. We assume that the set M as well as its enclosing rectangular grid are fixed, hence also the parameters T , Ω, and L. The dependence of the constants, auxiliary functions, etc., in the following derivations on these parameters will be suppressed for notational convenience; this should be seen as analogue to the one-dimensional scenario where the arising constants also depend on the shape and not just the size of the frequency support. Furthermore, set Q = max(LT, LΩ). We can bound using (2.3) (6.6) σ(j) m,ℓ = hσ, Φm,−nj ,ℓ,−kji ≤ kσk∞kΦm,−nj,ℓ,−kjk1 ≤ Bkσk∞kr ⊗ φk1 ≤ Bµ For the second inequality, we used that the L1-norm is invariant under translations and modula- tions. Furthermore, note that Vφr ∈ S(R2), so there is a decreasing positive function ρ ∈ S([0,∞)) Now observe that, as ρ is decreasing, such that for eρ(x, ξ) = ρ(cid:16)x(cid:17)ρ(cid:16)ξ(cid:17) one has Vφr ≤ 1 αjZα(j−1) ∞Xj=1 αρ(αj) ≤ ρ(0) + ∞Xj=0 8 CTeρ pointwise. ρ(t)dt = kρk1 + kρk∞. We use this estimate to bound for arbitrary (x, ξ) σ(x, ξ) LT β1β2 =(cid:12)(cid:12)(cid:12) ≤ LT β1β2 LT β1β2 ≤ µ ≤ 8LΩT β1 σ(j) e−2πi(xnj Ω+ξkj T )e2πinj Ωkj T X(mLT /β1,ℓLΩ/β2)∈S L−1Xj=0 X(mLT /β1,ℓLΩ/β2)∈S σ(j) L−1Xj=0 X(mLT /β1,ℓLΩ/β2)∈S ρ(cid:16) mL ∞Xm,ℓ=0 L−1Xj=0 m,ℓVφr(x −(cid:0) mL + kj(cid:1)T, ξ −(cid:0) ℓL m,ℓ(cid:12)(cid:12)(cid:12)Vφr(x −(cid:0) mL + nj(cid:1)Ω(cid:1)(cid:12)(cid:12)(cid:12) + kj(cid:1)T, ξ −(cid:0) ℓL + kj(cid:1)T(cid:12)(cid:12)(cid:12)(cid:17) ρ(cid:16)(cid:12)(cid:12)(cid:12)ξ −(cid:0) ℓL ρ(cid:16)(cid:12)(cid:12)(cid:12)x −(cid:0) mL 2(cid:16)kρk1 + kρk∞(cid:17)2 T(cid:17) LΩ Ω(cid:17) ≤ ρ(cid:16) ℓL + nj(cid:1)Ω(cid:1)(cid:12)(cid:12)(cid:12)(cid:17) Cµ 8 CT LT β1 β2 β2 β2 β1 β1 β2 β1 β2 µ 4 20 + nj(cid:1)Ω(cid:1)(cid:12)(cid:12)(cid:12) and hence kσ − σk∞ ≤ kσk∞ + kσk∞ ≤ µ + µ 2(cid:16)kρk1 + kρk∞(cid:17)2 =: C1µ. By the definition of S, for every δ > 0, there is a constant C(δ) such that for any fixed 0 ≤ j < L, (6.7) and hence, for (x, ξ) ∈ S − B(C(δ)), (6.8) δ ≥ L2 β1β2 δ ≥8kρkL1(R+)kρkL1[C(δ)−2Q,∞) ≥ 8keρkL1(([−C(δ)+Q,C(δ)−Q]2)c), + kj(cid:17)T(cid:12)(cid:12)(cid:12)(cid:17)ρ(cid:16)(cid:12)(cid:12)(cid:12)ξ −(cid:16) ℓL ρ(cid:16)(cid:12)(cid:12)(cid:12)x −(cid:16) mL Xℓ,m∈Z Ω(cid:17) /∈[−C(δ),C(δ)]2 (cid:16)x− mL T,ξ− ℓL β2 β2 β1 β1 + nj(cid:17)Ω(cid:12)(cid:12)(cid:12)(cid:17). To obtain (6.8) from (6.7), the boundary term in the discretization of the integral and the shifts by kj and nj, respectively, are each compensated by increasing the dimensions of the integra- tion/summation domain by LT and LΩ in time and frequency, respectively, both of which are bounded by Q. Note furthermore that, as (x, ξ) ∈ S − B(C(δ)), a necessary condition for (cid:16)x − mL β1 T, ξ − ℓL β2 Ω(cid:17) /∈ [−C(δ), C(δ)]2 Thus, using (6.6) and the triangle inequality, we can bound (6.8) from below obtaining (mLT /β1, ℓLΩ/β2) /∈ S. δµ ≥(cid:12)(cid:12)(cid:12) L2T β1β2 X(mLT /β1,ℓLΩ/β2) /∈S σ(j) m,ℓVφr(cid:16)x −(cid:16) mL β1 + kj(cid:17)T, ξ −(cid:16) ℓL β2 Hence forming a weighted average (with complex weighting factors of modulus one) of Equa- tion (6.9) over the L choices of j, we obtain is that (6.9) δµ ≥(cid:12)(cid:12)(cid:12) + nj(cid:17)Ω(cid:17)(cid:12)(cid:12)(cid:12). +kj(cid:17)T, ξ −(cid:16) ℓL β2 +nj(cid:17)Ω(cid:17)(cid:12)(cid:12)(cid:12) LT β1β2 L−1Xj=0 e−2πi(xnj Ω+ξkj T )e2πinj Ωkj T X( mLT β1 , ℓLΩ β2 σ(j) m,ℓVφr(cid:16)x −(cid:16) mL β1 ) /∈S =σ(x, ξ) − σ(x, ξ). This yields kσ − σkL∞(S−B(C(δ))) ≤ δµ. Hence by Theorem 3.3, we conclude that kHf − Hfk2 ≤ C δ C1 µ for all functions f which are δ C1 follows by choosing δ = min(cid:16) C1 -time-frequency-localized to S − B(C(δ)) − B(d(ǫ)). The result (cid:3) C ǫ, C1ǫ(cid:17) and D(ǫ) = C(δ) + d(ǫ). Acknowledgments. The authors thank the anonymous referee for the constructive comments, which greatly improved the paper, and Onur Oktay, who participated in initial discussions on the project. Part of this research was carried out during a sabbatical of G.E.P. and a stay of F.K. at the Department of Mathematics and the Research Laboratory for Electronics at the Mas- sachusetts Institute of Technology. Both are grateful for the support and the stimulating research environment. F.K. acknowledges support by the Hausdorff Center for Mathematics, Bonn. G.E.P. acknowledges funding by the German Science Foundation (DFG) under Grant 50292 DFG PF-4, Sampling Operators. 21 References 1. W.U. Bajwa, K. Gedalyahu, and Y.C. Eldar, Identification of parametric underspread linear systems and super- resolution radar, IEEE Trans. Signal Process. 59 (2011), no. 6, 2548 -- 2561. 2. P.A. Bello, Measurement of random time-variant linear channels, IEEE Trans. Comm. 15 (1969), 469 -- 475. 3. J. J. Benedetto, A. M. Powell, and O. Yılmaz, Sigma-Delta quantization and finite frames, IEEE Trans. Inform. Theory 52 (2006), 1990 -- 2005. 4. J.J. Benedetto and O. Oktay, Pointwise comparison of PCM and Σ∆ quantization, Constr. Approx. 32 (2010), no. 1, 131 -- 158. 5. ´A. B´enyi, K. Grochenig, K. A. Okoudjou, and L. G. Rogers, Unimodular Fourier multipliers for modulation spaces, J. Funct. Anal. 246 (2007), no. 2, 366 -- 384. 6. I. Daubechies and R. DeVore, Reconstructing a bandlimited function from very coarsely quantized data: A family of stable sigma-delta modulators of arbitrary order, Ann. Math. 158 (2003), 679 -- 710. 7. P. Deift, C. S. Gunturk, and F. Krahmer, An optimal family of exponentially accurate one-bit sigma-delta quantization schemes, Comm. Pure Appl. Math. 64 (2011), no. 7, 883 -- 919. 8. H.G. Feichtinger, Atomic characterizations of modulation spaces through Gabor-type representations, Rocky Mountain J. Math. 19 (1989), 113 -- 126. 9. H.G. Feichtinger and K. Grochenig, Gabor wavelets and the Heisenberg group: Gabor expansions and short time Fourier transform from the group theoretical point of view, Wavelets, Wavelet Anal. Appl., vol. 2, Academic Press, Boston, MA, 1992, pp. 359 -- 397. 10. G.B. Folland, Harmonic analysis in phase space, Annals of mathematics studies, vol. 122, Princeton University Press, 1989. 11. K. Grochenig, Foundations of Time-Frequency Analysis, Birkhauser, Boston, 2001. 12. K. Grochenig and C. Heil, Modulation spaces and pseudodifferential operators, Integral Equations Operator Theory 34 (1999), no. 4, 439 -- 457. MR 1702232 (2001a:47051) 13. C. S. Gunturk, One-bit Sigma-Delta quantization with exponential accuracy, Comm. Pure Appl. Math. 56 (2003), 1608 -- 1630. 14. R. Heckel and H. Boelcskei, Identification of sparse linear operators, preprint. 15. T. Kailath, Measurements on time-variant communication channels., IEEE Trans. Inform. Theory 8 (1962), no. 5, 229 -- 236. 16. W. Kozek and G.E. Pfander, Identification of operators with bandlimited symbols, SIAM J. Math. Anal. 37 (2005), no. 3, 867 -- 888. 17. F. Krahmer and G.E. Pfander, Sampling and quantization of approximately bandlimited operators, in prepara- tion, 2013. 18. F. Krahmer, G.E. Pfander, and P. Rashkov, Uncertainty principles for timefrequency representations on finite Abelian groups, Appl. Comput. Harmon. Anal. 25 (2008), 209 -- 225. 19. M. Lammers, A.M. Powell, and O Yılmaz, Alternative dual frames for digital-to-analog conversion in sigma- delta quantization, Adv. Comput. Math. 32 (2010), no. 1, 73 -- 102. 20. J. Lawrence, G.E. Pfander, and D. Walnut, Linear independence of Gabor systems in finite dimensional vector spaces, J. Fourier Anal. Appl. 11 (2005), no. 6, 715 -- 726. 21. K.A. Okoudjou, A Beurling-Helson type theorem for modulation spaces, J. Funct. Spaces Appl. 7 (2009), no. 1, 33 -- 41. 22. A.V. Oppenheim, R.W. Schafer, and J.R. Buck, Discrete-time signal processing, 2nd ed., Prentice-Hall signal processing, Prentice-Hall, Upper Saddle River, NJ, 1999. 23. D. P. Petersen and D. Middleton, Sampling and reconstruction of wave-number-limited functions in N - dimensional Euclidean spaces, Information and Control 5 (1962), 279 -- 323. MR 0151331 (27 #1317) 24. G.E. Pfander, Sampling of operators, to appear in J. Four. Anal. Appl. 25. , Measurement of time -- varying Multiple -- Input Multiple -- Output channels, Appl. Comp. Harm. Anal. 24 (2008), 393 -- 401. 26. G.E. Pfander and D. Walnut, Sampling and reconstruction of operators, preprint. 27. G.E. Pfander and D.F. Walnut, Measurement of time-variant linear channels, IEEE Trans. Inform. Theory 52 (2006), no. 11, 4808 -- 4820. 28. A. Ron and Z. Shen, Frames and stable bases for shift -- invariant subspaces L2(Rd), Canadian Journal of Mathematics 47 (1995), no. 5, 1051 -- 1094. 29. T. Strohmer, Pseudodifferential operators and Banach algebras in mobile communications, Appl. Comput. Harmon. Anal. 20 (2006), no. 2, 237 -- 249. 30. J. Toft, Continuity properties for modulation spaces, with applications to pseudo-differential calculus -- I, Journal of Functional Analysis 207 (2004), no. 2, 399 -- 429. 31. 32. , Continuity properties for modulation spaces, with applications to pseudo-differential calculus -- II, Annals of Global Analysis and Geometry 26 (2004), no. 1, 73 -- 106. , Continuity and schatten properties for pseudo-differential operators on modulation spaces, Modern trends in pseudo-differential operators, Springer, 2007, pp. 173 -- 206. 22
1105.2476
1
1105
2011-05-12T13:49:26
Compact Inverses of The Multipoint Normal Diferential Operators For First Order
[ "math.FA" ]
In this work, firstly all normal extensions of a multipoint minimal operator generated by linear multipoint diferential-operator expression for first order in the Hilbert space of vector functions in terms of boundary values at the endpoints of the infinitely many separated subintervals are described. Finally, a compactness properties of the inverses of such extensions has been investigated.
math.FA
math
Compact Inverses of The Multipoint Normal Differential Operators For First Order Erdal UNLUYOL 1∗ , Elif OTKUN C¸ EVIK 1, Zameddin ISMAILOV 1 1 Karadeniz Technical University, Faculty of Science, Department of Mathematics, 61080 Trabzon, Turkey [email protected], e−[email protected], [email protected] Abstract In this work, firstly all normal extensions of a multipoint minimal operator generated by linear multipoint differential-operator expression for first order in the Hilbert space of vector functions in terms of boundary values at the endpoints of the infinitely many separated subintervals are described. Finally, a compactness properties of the inverses of such extensions has been investigated. 2010 AMS Subject Classification: 47A05, 47A20 Keywords: Direct sum of Hilbert spaces and operators; multipoint selfadjoint operator; formally normal operator; normal operator; extension. 1 Introduction It is known that traditional infinite direct sum of Hilbert spaces Hn, n ≥ 1 and infinite direct sum of operators An in Hn, n ≥ 1 are defined as H = Hn =(u = (un) : un ∈ Hn, n ≥ 1 and ∞ ⊕ n=1 kunk2 Hn ∞Xn=1 < +∞) , A = ∞ ⊕ n=1 An, D(A) = {u = (un) ∈ H : un ∈ D(An), n ≥ 1 and Au = (Anun) ∈ H} . A linear space H is a Hilbert space with norm induced by the inner product (u, v)H = ∞Xn=1 (un, vn)Hn , u, v ∈ H [1]. The general theory of linear closed operators in Hilbert spaces and its applications to physical problems have been investigated by many researches (for example, see [1], [2]). However, many physical problems requires the study of the theory of linear operators in direct sums in Hilbert spaces ( for example, see [3]-[8] and references therein). We note that a detail analysis of normal subspaces and operators in Hilbert spaces have been 1 ∗ corresponding author: [email protected] (Erdal UNLUYOL), Fax: +90 (462) 325 31 95 1 studied in [9] (see references in it ). This study contains three section except introduction. In section 2, the multipoint minimal and maximal operators for the first order differential-operator expression are determined. In sec- tion 3, all normal extensions of multipoint formally normal operators are described in terms of boundary values in the endpoints of the infinitely many separated subintervals. Finally In section 4, compactness properties of the inverses of such extensions have been established. 2 The Minimal and Maximal Operators Throughout this work (an) and (bn) will be sequences of real numbers such that −∞ < an < bn < an+1 < · · · < +∞, Hn is any Hilbert space, ∆n = (an, bn) , L2 n = L2 (Hn, ∆n) , L2 = 0 W 1 2 = 0 W 1 2 (Hn, ∆n) , H = ∞ ⊕ n=1 ∞ ⊕ n=1 L2 (Hn, ∆n) , n ≥ 1, (bn − an) < +∞, W 1 sup n≥1 of the operator T . l (·) is a linear multipoint differential-operator expression for first order in L2 in the following form Hn, cl (T )-closure 2 (Hn, ∆n) , 2 = W 1 ∞ ⊕ n=1 ∞ ⊕ n=1 and for each n ≥ 1 l (u) = (ln (un)) ln (un) = u′ n + Anun, (2.1) (2.2) where An : D (An) ⊂ Hn → Hn is a linear positive defined selfadjoint operator in Hn. It is clear that formally adjoint expression to (2.2) in the Hilbert space L2 n is in the form n (vn) = −v′ l+ n + Anvn, n ≥ 1. (2.3) We define an operator L′ n0 on the dense manifold of vector functions D′ n0 in L2 n as ′ n0 :=(un ∈ L2 n : un = D φkfk, φk ∈ C∞ 0 (∆n), fk ∈ D(An), k = 1, 2, · · · , m; m ∈ N) mXk=1 n0un := ln (un) , n ≥ 1. with L′ Since the operator An > 0, n ≥ 1, then from the relation Re (L′ n0un, un)L2 n = 2 (Anun, un)L2 n ≥ 0, un ∈ D′ n0 it implies that L′ The closure cl (L′ n0 is an accretive in L2 n0) of the operator L′ n, n ≥ 1. Hence the operator L′ n, n ≥ 1. n0 is called the minimal operator generated by differential- n0 has a closure in L2 2 operator expression (2.2) and is denoted by Ln0 in L2 n, n ≥ 1. The operator L0 is defined by D (L0) :=(u = (un) : un ∈ D (Ln0) , n ≥ 1, kLn0unk2 L2 n ∞Xn=1 < +∞) with L0u := (Ln0un) , u ∈ D (L0) , L0 : D (L0) ⊂ L2 → L2 is called a minimal operator ( multipoint ) generated by differential-operator expression (2.1) in Hilbert space L2 and denoted by L0 = Ln0. ∞ ⊕ n=1 In a similar way the minimal operator for twopoints denoted by L+ n0 in L2 n, n ≥ 1 for the formally adjoint linear differential-operator expression (2.3) can be constructed. In this case the operator L+ 0 defined by 0(cid:1) :=(v := (vn) : vn ∈ D(cid:0)L+ D(cid:0)L+ 0 v := (cid:0)L+ n0vn(cid:1) , v ∈ D(cid:0)L+ 0 : D(cid:0)L+ with L+ (multipoint) generated by l+ (v) = (l+ 0(cid:1) , L+ ∞Xn=1(cid:13)(cid:13)L+ n0(cid:1) , n ≥ 1, 0(cid:1) ⊂ L2 → L2 is called a minimal operator n0vn(cid:13)(cid:13)2 0 = L+ n0. L2 n n (vn)) in the Hilbert space L2 and denoted by L+ < +∞) ∞ ⊕ n=1 We now state the following relevant result. Theorem 2.1. The minimal operators L0 and L+ The following defined operators in L2 L :=(cid:0)L+ 0(cid:1)∗ = ∞ ⊕ n=1 0 are densely defined closed operators in L2. Ln and L+ := (L0)∗ = ∞ ⊕ n=1 L+ n are called maximal operators (multipoint) for the differential-operator expression l (·) and l+ (·) respectively.It is clear that Lu = (ln (un)) , u ∈ D (L), D (L) :=(u = (un) ∈ L2 : un ∈ D (Ln) , n ≥ 1 , ∞Xn=1 n (vn)(cid:1) , v ∈ D(cid:0)L+(cid:1) , D(cid:0)L+(cid:1) :=(v = (vn) ∈ L2 : vn ∈ D(cid:0)L+ ∞Xn=1(cid:13)(cid:13)L+ n vn(cid:13)(cid:13)2 n(cid:1) , n ≥ 1 , L+v =(cid:0)l+ kLnunk2 L2 n < ∞) , < ∞) L2 n and L0 ⊂ L, L+ 0 ⊂ L+. Furthermore, the validity of following proposition is clear. Theorem 2.2. The domain of the operators L and L0 are D (L) =(cid:8)u = (un) ∈ L2 : (1) for each n ≥ 1 vector function un ∈ L2 is absolutely continuous in interval ∆n; (2) ln (un) ∈ L2 n, n ≥ 1; (3) l (u) = (ln (un)) ∈ L2(cid:9) =(cid:8)u = (un) ∈ L2 : un ∈ D (Ln) , n ≥ 1 and l (u) = (ln (un)) ∈ L2(cid:9) , n, un kAnk ≤ c < +∞, then for any u = (un) ∈ L2 we D(L0) = {u = (un) ∈ D (L) : un (an) = un (bn) = 0, n ≥ 1}. Remark 2.3. If An ∈ B (H) , n ≥ 1 and sup n≥1 have (Au) = (Anun) ∈ L2. 3 Now the following results can be proved . Theorem 2.4. AD (L0) ⊂ L2. If a minimal operator L0 is formally normal in L2, then D (L0) ⊂ 0 W 1 2 and Theorem 2.5. If A1/2W 1 2 ⊂ W 1 2 , then minimal operator L0 is formally normal in L2. Proof: In this case from the following relations L+ 0 u = L0u − 2Au, u ∈ D (L0) , it implies that D (L0) = D(cid:0)L+ D (L+). On the other hand for any u ∈ D (L0) 0(cid:1). Since D(cid:0)L+ L0u = L+ 0(cid:1) 0 u + 2Au, u ∈ D(cid:0)L+ 0(cid:1) ⊂ D (L∗ 0) = D (L+), it is obtained that D (L0) ⊂ kL0uk2 L2 = (u′ + Au, u′ + Au)L2 = ku′k2 = ku′k2 L2 + kAuk2 L2 L2 + [(u′, Au)L2 + (Au, u′)L2] + kAuk2 L2 and kL+uk2 L2 = (−u′ + Au, −u′ + Au)L2 = ku′k2 = ku′k2 L2 + kAuk2 L2 . L2 − [(u′, Au)L2 + (Au, u′)L2] + kAuk2 L2 Thus, it is established that operator L0 is formally normal in L2. Remark 2.6. If An ∈ B (H) , n ≥ 1 and sup n≥1 D (L) = D (L+) . Remark 2.7. If AW 1 2 ⊂ L2, then D (L0) = D(cid:0)L+ kAnk ≤ c < +∞, then D (L0) = D(cid:0)L+ 0(cid:1) and D (L) = D (L+). 0(cid:1) and 3 Description of Normal Extensions of the Minimal Oper- ator In this section the main purpose is to describe all normal extensions of the minimal operator L0 in L2 in terms in the boundary values of the endpoints of the subintervals . First, we will show that there exists normal extension of the minimal operator L0. Consider the following extension of the minimal operator L0  eLu := u′ + Au, AW 1 D(cid:16)eL(cid:17) =(cid:8)u = (un) ∈ W 1 2 ⊂ W 1 2 , 2 : un (an) = un (bn) , n ≥ 1(cid:9) . 4 Under the condition on the coefficient A we have = (u′, v)L2 + (Au, v)L2 = (u, v) ′ L2 +(cid:16)u, −v ′ + Av(cid:17)L2 ′ [(un(bn), vn(bn))Hn − (un(an), vn(an))Hn ] + (u, −v + Av)L2 From this it is obtained In this case it is clear that D(cid:16)eL(cid:17) = D(cid:16)eL∗(cid:17). On the other hand, since for each u ∈ D(cid:16)eL(cid:17) 2 2 : vn (an) = vn (bn) , n ≥ 1(cid:9) . L2 + [(u′, Au)L2 + (Au, u′)L2] + kAuk2 L2 , L2 − [(u′, Au)L2 + (Au, u′)L2 ] + kAuk2 L2 = (cid:16)eLu, v(cid:17)L2 ∞Pn=1  eL∗v := −v′ + Av, D(cid:16)eL∗(cid:17) =(cid:8)v = (vn) ∈ W 1 (cid:13)(cid:13)(cid:13)eLu(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)eL∗u(cid:13)(cid:13)(cid:13) = ku′k2 = ku′k2 L2 2 L2 and ∞Pn=1 = (cid:13)(cid:13)(cid:13)eL∗u(cid:13)(cid:13)(cid:13)L2 Then (cid:13)(cid:13)(cid:13)eLu(cid:13)(cid:13)(cid:13)L2 minimal operator L0. (u′, Au)L2 + (Au, u′)L2 = (u, Au) = ′ L2 [(un(bn), Anun(bn))Hn − (un(an), Anun(an))Hn ] = 0. for every u ∈ D(cid:16)eL(cid:17). Consequently, eL is a normal extension of the The following result establishes the relationship between normal extensions of L0 and normal extensions of Ln0, n ≥ 1. Theorem 3.1. The extension eL = only if for any n ≥ 1, fLn is so in L2 n . ∞ ⊕ n=1fLn of the minimal operator L0 in L2 is a normal if and Now using the Theorem 3.1 and [10] we can formulate the following main result of this section, where it is given a description of all normal extension of the minimal operator L0 in L2 in terms of boundary values of vector functions at the endpoints of subintervals. Theorem 3.2. Let A1/2W 1 L0 in L2, then it is generated by differential-operator expression (2.1) with boundary conditions n=1fLn is a normal extension of the minimal operator 2 . If eL = 2 ⊂ W 1 ∞ ⊕ un(bn) = Wnun(an), un ∈ D(Ln), (3.1) ∞ ⊕ n=1 Wn in H = where Wn is a unitary operator in Hn and WnA−1 W = ∞ ⊕ n=1 On the contrary, the restriction of the maximal operator L to the linear manifold u ∈ D(L) Wn in H with property W A−1 = Hn is determined uniquely by the extension eL, that is eL = LW . ∞ satisfying the condition (3.1) with any unitary operator W = ⊕ n=1 A−1W is a normal extension of the minimal operator L0 in L2. n Wn, n ≥ 1. The unitary operator n = A−1 5 4 Some Compactness Properties of The Normal Extensions The following two proposition can be easily proved in general case. Theorem 4.1. For the point spectrum of A = spaces Hn, n ≥ 1 t is true that ∞ ⊕ n=1 An in the direct sum H = ∞ ⊕ n=1 Hn of Hilbert σp(A) = σp(An) ∞[n=1 ∞ Theorem 4.2. Let An ∈ B(Hn), n ≥ 1, A = ⊕ n=1 necessary and sufficient condition is that the sup n≥1 ∞ ⊕ n=1 An and H = kAnk be finite. In this case kAk = sup n≥1 Hn. In order for A ∈ B(H) the kAnk. Let C∞(·) and Cp(·), 1 ≤ p < ∞ denote the class of compact operators and the Schatten-von Neumann subclasses of compact operators in corresponding spaces respectively. Definition 4.3.[11] Let T be a linear closed and densely defined operator in any Hilbert space H. If ρ(T ) 6= Ø and for λ ∈ ρ(T ) the resolvent operator Rλ(T ) ∈ C∞(H), then operator T : D(T ) ⊂ H → H is called an operator with discrete spectrum In first note that the following results are true. Theorem 4.4. If the operator A = An as an operator with discrete spectrum in H = ∞ ⊕ n=1 Hn, then for every n ≥ 1 the operator An is so in Hn. ∞ ⊕ n=1 Remark 4.5. Unfortunately, the converse of the Theorem 4.4 is not true in general case. Indeed, consider the following sequence of operators Anun = un, 0 < dimHn = dn < ∞, n ≥ 1. In this case for every n ≥ 1 operator An is an operator with discrete spectrum. But an inverse of the Hn, because dimH = ∞ and direct sum operator A = An is not compact operator in H = ∞ ⊕ n=1 ∞ ⊕ n=1 A is an identity operator in H. Theorem 4.6. If A = ∞ ⊕ n=1 An, An is an operator with discrete spectrum in Hn, n ≥ 1, Ø and lim n→∞ kRλ(An)k = 0, then A is an operator with discrete spectrum in H. ρ(An) 6= ∞Tn=1 Proof. For each λ ∈ ρ(An) we have Rλ(An) ∈ C∞(Hn), n ≥ 1. Now define the following operators Km : H → H, m ≥ 1 as ∞Tn=1 Km := {Rλ(A1)u1, Rλ(A2)u2, . . . , Rλ(Am)um, 0, 0, . . . }, u = (un) ∈ H. The convergence of the operators Km to the operator K in operator norm will be investigated. 6 For the u = (un) ∈ H we have kKmu − Kuk2 H = kRλ(An)unk2 Hn ≤ kRλ(An)k2kkunk2 Hn ∞Xn=m+1 ≤ (cid:18) sup n≥m+1 ∞Xn=m+1 kunk2 kRλ(An)k(cid:19)2 ∞Xn=1 Hn =(cid:18) sup n≥m+1 kRλ(An)k(cid:19)2 kuk2 H thus we get kKmu − Kuk ≤ sup kRλ(An)k, m ≥ 1. n≥m+1 This means that sequence of operators (Km) converges in operator norm to the operator K. Then by the important theorem of the theory of compact operators it is implies that K ∈ C∞(H) [1], because for any m ≥ 1, Km ∈ C∞(H). Finally, using the Theorem 4.6 can be proved the following result. Theorem 4.7. If A−1 n ∈ C∞(Hn), n ≥ 1, sup n≥1 (bn − an) < ∞ and the sequence of first minimal eigenvalues λ1(An) of the operators An, n ≥ 1 is satisfy the condition λ1(An) → ∞ as n → ∞, ∞ ⊕ n=1 Ln is an operator with discrete spectrum in L2. then the extension eL = Theorem 4.8. Let H = ∞ ⊕ n=1 Hn, A = ∞ ⊕ n=1 An and An ∈ Cp(Hn), n ≥ 1, 1 ≤ p < ∞. In order for A ∈ Cp(H) the necessary and sufficient condition is that the series Now we will dedicate an application the last theorem. µp k(An) be convergent. ∞Pn=1 ∞Pk=1 For all n ≥ 1, Hn is a Hilbert space, ∆n = (an, bn), −∞ < an < bn < an+1 < · · · < n Wn = 2 (Hn, ∆n), Hn = L2(Hn, ∆n), D(LWn ) = {un ∈ Hn ∞, An : D(An) ⊂ Hn → Hn, An = A∗ WnA−1 W 1 n ≥ E, Wn : Hn → Hn is unitary operator, A−1 2 (Hn, ∆n) : un(bn) = Wnun(an)}, LWn : Hn → Hn, W = n + Anun, AnW 1 2 (Hn, ∆n) ⊂ W 1 n , LWn un = u′ Wn, LW = LWn, H = ∞ ⊕ n=1 ∞ ⊕ n=1 ∞ ⊕ n=1 and h = sup n≥1 (bn − an) < ∞. Since for all n ≥ 1 Wn is a unitary operator in Hn, then LWn is normal operator in Hn [10]. W LW is true, i.e. LW is a normal n ∈ Cp(Hn), for p > 1, then L−1 ∈ Cp(Hn), p > 1 for Wn n ∈ C∞(Hn), n ≥ 1, then eigenvalues λq(LWn ), q ≥ 1 of Also for LW : D(LW ) ⊂ H → H, the relation LW L∗ It is known that, if A−1 operator in H. all n ≥ 1[10]. On the other hand, if A−1 operator LWn is in form W = L∗ λq(LWn ) = λm(An) + i an − bn(cid:16)argλm(W ∗ n e(−An(bn−an))) + 2kπ(cid:17) , m ≥ 1, k ∈ Z, n ≥ 1, where q = q(m, k) ∈ N, m ≥ 1, k ∈ Z. Therefore we have the following corollary. 7 Theorem 4.9. If A = ∞ ⊕ n=1 An, H = ∞ ⊕ n=1 Hn and A−1 ∈ Cp/2(H), 2 < p < ∞, then L−1 W ∈ Cp(H). Proof. The operator LW is a normal in H. Consequently, for the characteristic numbers of normal operator L−1 W ) , q ≥ 1 holds [1]. Now we search for convergence of W an equality µq(L−1 W ) = λq(L−1 q=1 µp q(L−1 W ), 2 < p < ∞. the seriesP∞ ∞Xq=1 ∞Xn=1 q (L−1 µp Wn m(An) + m(An) + ) = ≤ ≤ ∞Xn=1 ∞Xn=1 ∞Xn=1 ∞Xk=−∞ ∞Xk=−∞ ∞Xm=1(cid:0)λ2 ∞Xm=1(cid:18)λ2 ∞Xm=1(cid:18)λ2 m(An)(cid:1)−p/2 + 2 1 4k2π2 (bn − an)2 (δ(m, n) + 2kπ)2(cid:19)−p/2 (bn − an)2(cid:19)−p/2 ∞Xm=1(cid:18)λ2 ∞Xn=1 ∞Xk=1 m(An) + 4k2π2 (bn − an)2(cid:19)−p/2 where δ(m, n) = argλm(W ∗ all t, s ∈ R \ {0} and last equation we have the inequality n e(−An(bn−an))), n ≥ 1, m ≥ 1. Then from the inequality ts t2+s2 ≤ 1 2 for ∞Xn=1 ∞Xk=1 ∞Xm=1(cid:18)λ2 m(An) + 1 λm(An)(cid:12)(cid:12)(cid:12)(cid:12) p/2 ∞Xk=1(cid:12)(cid:12)(cid:12)(cid:12) 1 p/2! k(cid:12)(cid:12)(cid:12)(cid:12) Since A−1 ∈ Cp/2(H), then the series λm(An)−p/2 is convergent. Thus the series 4k2π2 ≤ 2−pπ−p/2hp/2 ∞Xn=1 (bn − an)2(cid:19)−p/2 ∞Pn=1 ∞Pm=1 ∞Xm=1(cid:18)λ2 ∞Xm=1(cid:12)(cid:12)(cid:12)(cid:12) (bn − an)2(cid:19)−p/2 ∞Xn=1 ∞Xk=1 m(An) + 4k2π2 λm(An)−p ≤ ∞Xn=1 ∞Xm=1 λm(An)−p/2 ∞Xn=1 ∞Xm=1 ∞Pn=1 ∞Pm=1 ∞Pq=1 is also convergent. Then from the relation and the convergence of the series λm(An)−p/2 we get that the series λm(An)−p is convergent too. Consequently the series q (L−1 µp W ), 2 < p < ∞ is convergent and thus L−1 W ∈ ∞Pn=1 ∞Pm=1 Cp(H), 2 < p < ∞. The Theorem 4.8 and 4.9. can be can generalized. Corollary 4.10. Let for all n ≥ 1 An ∈ Cpn (Hn), 1 ≤ pn < ∞ and p = sup n≥1 pn < ∞. For A ∈ Cp(H) the necessary and sufficient condition is that the series µp k(An) be convergent. ∞Pn=1 ∞Pk=1 pn < ∞, then L−1 W ∈ Cp(H). Theorem 4.11. If A−1 n ∈ Cpn/2(Hn), 2 ≤ pn < ∞, p = sup n≥1 8 References [1] N. Dunford, J.T. Schwartz, Linear Operators I; II, Second ed., Interscience, New York, 1958; 1963. [2] F.S. Rofe-Beketov , A.M. Kholkin, Spectral Analysis of Differential Operators, First ed., World Scientific Mono- graph Series in Mathematics, Vol.7, New Jersey, 2005. [3] S. Timoshenko, Theory of Elastic Stability, second ed., McGraw-Hill, New York, 1961. [4] F.R. Gantmakher, M.G. Krein, Oscillating Matrices and Kernels and Small Oscillations of Mechanical Systems, First ed., Gostekhteorizdat, Moscow, 1950, (in Russian). [5] A. Zettl, Sturm-Liouville Theory, First ed., Amer. Math. Soc., Math. Survey and Monographs vol. 121, USA, 2005. [6] A.N. Kochubei, Symmetric Operators and Nonclassical Spectral Problems, Mat. Zametki, 25, 3 (1979), 425-434. [7] Z.I. Ismailov, Multipoint Normal Differential Operators for First Order, Opusc. Math., 29, 4, (2009), 399-414. [8] M.S. Sokolov, An Abstract Approach to Some Spectral Problems of Direct Sum Differential Operators, Electron. J. Differential Equations, vol. 2003(2003), No.75, pp. 1-6. [9] E.A. Coddington, Extension theory of formally normal and symmetric subspaces, Mem. Amer. Math. Soc., 134 (1973) 1-80. [10] Z.I. Ismailov, Compact inverses of first-order normal differential operators, J. Math. Anal. Appl., 320 (2006) 266-278. [11] V. I. Gorbachuk, M. L. Gorbachuk, Boundary value problems for operator-differential equations, First ed., Kluwer Academic Publisher, Dordrecht, 1991. 9
1911.10329
1
1911
2019-11-23T08:55:39
On a compact trace embedding theorem in Musielak-Sobolev spaces
[ "math.FA" ]
By a stronger compact boundary embedding theorem in Musielak-Orlicz-Sobolev space developed in the paper, variational method is employed to deal with the nonlinear elliptic equation with the nonlinear Neumann boundary condition in the framework of Musielak-Orlicz-Sobolev space.
math.FA
math
ON A COMPACT TRACE EMBEDDING THEOREM IN MUSIELAK-SOBOLEV SPACES LI WANG AND DUCHAO LIU Abstract. By a stronger compact boundary embedding theorem in Musielak- Orlicz-Sobolev space developed in the paper, variational method is employed to deal with the nonlinear elliptic equation with the nonlinear Neumann boundary condition in the framework of Musielak-Orlicz-Sobolev space. 1. Introduction In this paper we develop a stronger compact trace embedding of Sobolev type in Musielak-Orlicz-Sobolev (or Musielak-Sobolev) spaces than the existed results in mathematics literature [11]. And as an application of this stronger compact trace embedding theorem, the existence and multiplicity of weak solutions in Musielak- Orlicz-Sobolev spaces for the following equation are considered (−div(a1(x, ∇u)∇u) + a0(x, u) = 0 a1(x, ∇u) ∂u ∂n = g(x, u) in Ω, on ∂Ω, where Ω is a bounded smooth domain in RN , a0, a1 : Ω × R → R are Carath´eodory functions and ∂u ∂n is the outward unit normal derivative on ∂Ω. In the study of nonlinear differential equations, it is well known that more general functional space can handle differential equations with more complex nonlinearities. If we want to study a general form of differential equations, it is very important to find a proper functional space in which the solutions may exist. Musielak-Sobolev space is such a general kind of Sobolev space that the classical Sobolev spaces, Orlicz-Sobolev spaces and variable exponent Sobolev spaces can be interpreted as its special cases. Differential equations in Orlicz-Sobolev spaces have been studied extensively in recent years (see [3, 4, 5, 6, 9, 12]). To the best of our knowledge, however, differen- tial equations in Musielak-Sobolev spaces have been studied little. In [2], Benkirane and Sidi give an embedding theorem in Musielak-Sobolev spaces and an existence result for nonlinear elliptic equations. In [8] and [7], Fan gives two embedding theorems and some properties of differential operators in Musielak-Sobolev spaces. In the research of the paper [10], Liu and Zhao give existence the results for non- linear elliptic equations with homogeneous Dirichlet boundary. In a much recent paper [11], Liu, Wang and Zhao give a trace embedding theorem and a compact ∗Research supported by the National Natural Science Foundation of China (NSFC 11501268 and NSFC 11471147) and the Fundamental Research Funds for the Central Universities (lzujbky- 2016-101). 1991 Mathematics Subject Classification. 35B38, 35D05, 35J20. Key words and phrases. Musielak-Sobolev space; Nonlinear boundary condition; Mountain pass type solution; Genus theory. 1 2 WANG AND LIU trace embedding theorem in bounded domain about this kind of functional space. Motivated by [11], we find that the results in that paper can be better improved and it is possible to make an application of the result to get some existence results for nonlinear elliptic equations with nonlinear Neumann boundary condition. By developing a stronger trace embedding theorem in Musielak-Sobolev spaces, our aim in this paper is to study the existence of solutions to a kind of elliptic differen- tial equation with Neumann boundary condition under the functional frame work of Musielak-Sobolev spaces, which is a more general case in the variable exponent Sobolev spaces and Orlicz-Sobolev spaces. As an example of Musielak-Sobolev spaces we claim that not only variable ex- ponent Sobolev spaces satisfy the conditions in our theorem (see Example 2.1 in Section 2), but some more complex case can also work (see Example 2.2 in Section 2). And the current existed theories in mathematical literature can not cover the case in Example 2.2. The paper is organized as follows. In Section 2, we develop a stronger compact trace embedding theorem and for the readers' convenience recall some definitions and properties about Musielak-Orlicz-Sobolev spaces. In Section 3, we give some propositions of differential operators in Musielak-Sobolev spaces. In Section 4, we give the existence results of weak solutions to the equations. 2. A stronger compact trace embedding theorem in Musielak-Orlicz-Sobolev spaces In this section, we will give a stronger compact trace embedding theorem to proceed. And we will list some basic definitions and propositions about Musielak- Orlicz-Sobolev spaces. Firstly, we give the definition of N-function and generalized N-function as follows. Definition 2.1. A function A : R → [0, +∞) is called an N -function, denoted by A ∈ N , lim t→0+ A(t) t = 0 and lim t→+∞ A(t) t = +∞. A function A : Ω × R → [0, +∞) is called a generalized N -function, denoted by A ∈ N (Ω), if for each t ∈ [0, +∞), the function A(·, t) is measurable, and for a.e. x ∈ Ω, A(x, ·) ∈ N . Let A ∈ N (Ω), the Musielak-Orlicz space LA(Ω) is defined by LA(Ω) := {u : u is a measurable real function, and ∃λ > 0 such that ZΩ A(cid:18)x, kukLA(Ω) = kukA := inf(cid:26)λ > 0 :ZΩ with the (Luxemburg) norm λ (cid:19) dx < +∞} u(x) A(cid:18)x, λ (cid:19) dx ≤ 1(cid:27). u(x) The Musielak-Sobolev space W 1,A(Ω) can be defined by W 1,A(Ω) := {u ∈ LA(Ω) : ∇u ∈ LA(Ω)} with the norm kuk := kukW 1,A(Ω) := kukA + k∇ukA, ON A COMPACT TRACE EMBEDDING THEOREM IN MUSIELAK-SOBOLEV SPACES ∗ 3 where k∇ukA := k∇ukA. And kukρ := inf{λ > 0,RΩ is an equivalent norm on W 1,A(Ω), see in [8]. A(x, u λ ) + A(x, ∇u λ (cid:1) dx ≤ 1} We say that A ∈ N (Ω) satisfies Condition ∆2(Ω), if there exists a positive constant K > 0 and a nonnegative function h ∈ L1(Ω) such that A(x, 2t) ≤ KA(x, t) + h(x) for x ∈ Ω and t ∈ R. A is called locally integrable if A(·, t0) ∈ L1 loc(Ω) for every t0 > 0. For x ∈ Ω and t ≥ 0, we denote by a(x, t) the right-hand derivative of A(x, t) at t, at the same 0 a(x, s) ds for x ∈ Ω and t ∈ R. time define a(x, t) = −a(x, −t). Then A(x, t) =R t Define eA : Ω × R → [0, +∞) by eA(x, s) = sup eA is called the complementary function to A in the sense of Young. It is well known that eA ∈ N (Ω) and A is also the complementary function to eA. + (x, s) the right-hand derivative of eA(x, ·) at s, at the same time define a−1 for x ∈ Ω and s ≥ 0, we have For x ∈ Ω and s ≥ 0, we denote by a−1 + (x, −s) for x ∈ Ω and s ≤ 0. Then (st − A(x, t)) for x ∈ Ω and s ∈ R. + (x, s) = −a−1 t∈R a−1 + (x, s) = sup{t ≥ 0 : a(x, t) ≤ s} = inf{t > 0 : a(x, t) > s}. Proposition 2.1. (See [10].) Let A ∈ N (Ω). Then the following assertions hold. (1) A(x, t) ≤ a(x, t)t ≤ A(x, 2t) for x ∈ Ω and t ∈ R; (2) A and eA satisfy the Young inequality and the equality holds if s = a(x, t) or t = a−1 st ≤ A(x, t) + eA(x, s) for x ∈ Ω and s, t ∈ R + (x, s). Let A, B ∈ N (Ω). We say that A is weaker than B, denoted by A 4 B, if there exist positive constants K1, K2 and a nonnegative function h ∈ L1(Ω) such that A(x, t) ≤ K1B(x, K2t) + h(x) for x ∈ Ω and t ∈ [0, +∞). eB(Ω). eA(Ω) ֒→ L where u ∈ LA(Ω); LB(Ω) ֒→ LA(Ω) and L Proposition 2.3. (See [10].) Let A ∈ N (Ω) satisfy ∆2(Ω). Then the following assertions hold, Proposition 2.2. (See [10].) Let A, B ∈ N (Ω), and A 4 B. Then eB 4 eA, (1) LA(Ω) = {u : u is a measurable function, and RΩ A(x, u(x)) dx < +∞}; (2) RΩ A(x, u(x)) dx < 1 (resp. = 1; > 1) ⇔ kukA < 1 (resp. = 1; > 1), (3) RΩ A(x, un(x)) dx → 0 (resp. 1; +∞) ⇔ kunkA → 0 (resp. 1; +∞), (4) un → u in LA(Ω) ⇒RΩ A(x, un(x)) − A(x, u(x)) dx → 0 as n → ∞; (5) If eA also satisfies ∆2(Ω), then u(x)v(x) dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ 2kukAkvk eA, eA(Ω) for every u ∈ LA(Ω). where {un} ⊂ LA(Ω); ∀ u ∈ LA(Ω), v ∈ L (cid:12)(cid:12)(cid:12)(cid:12)ZΩ (6) a(·, u(·)) ∈ L eA(Ω); hJ(v), wi =ZΩ v(x)w(x) dx, ∀ v ∈ L eA(Ω), w ∈ LA(Ω) 4 WANG AND LIU Proposition 2.4. (See [10].) Let A ∈ N (Ω) be locally integrable. Then (LA(Ω), k · k) is a separable Banach space, W 1,A(Ω) is reflexive provided LA(Ω) is reflexive. The following assumptions on A ∈ N (Ω) will be used. (A1) inf x∈Ω A(x, 1) = c1 > 0; (A2) For every t0 > 0, there exists c = c(t0) > 0 such that A(x, t) t ≥ c and eA(x, t) t ≥ c, ∀ t ≥ t0, x ∈ Ω. Remark 2.1. It is easy to see that (A2) ⇒ (A1). Proposition 2.5. (See [10].) If A ∈ N (Ω) satisfies (A1), then LA(Ω) ֒→ L1(Ω) and W 1,A(Ω) ֒→ W 1,1(Ω). Proposition 2.6. (See [8].) Let A ∈ N (Ω), both A and eA be locally integrable and satisfy ∆2(Ω) and (A2). Then LA(Ω) is reflexive, and the mapping J : L (LA(Ω))∗ defined by eA(Ω) → is a linear isomorphism and kJ(v)k(LA(Ω))∗ ≤ 2kvkL eA(Ω). To give the embedding theorems, the following assumptions are introduced. (P1) Ω ⊂ RN (N ≥ 2) is a bounded domain with the cone property, and A ∈ N (Ω); (P2) A : Ω × R → [0, +∞) is continuous and A(x, t) ∈ (0, +∞) for x ∈ Ω and t ∈ (0, +∞). (P3) Suppose A ∈ N (Ω) satisfies (2.1) Z 1 0 A−1(x, t) n+1 n t dt < +∞, ∀x ∈ Ω, replacing, if necessary, A by another N (Ω)-function equivalent to A near infinity. Under assumption (P1), (P2) and (P3), for each x ∈ Ω we can define A∗(x, t) for x ∈ Ω and t ∈ RN as in [11]. We call A∗ the Sobolev conjugate function of A. Set (2.2) T (x) := lim s→+∞ A−1 ∗ (x, s). Then A∗(x, ·) ∈ C1(0, T (x)). Furthermore for A ∈ N (Ω) and T (x) = +∞ for any x ∈ Ω, it is known that A∗ ∈ N (Ω) (see in [1]). Let X be a metric space and f : X → (−∞, +∞] be an extended real-valued function. For x ∈ X with f (x) ∈ R, the continuity of f at x is well defined. For x ∈ X with f (x) = +∞, we say that f is continuous at x if given any M > 0, there exists a neighborhood U of x such that f (y) > M for all y ∈ U . We say that f : X → (−∞, +∞] is continuous on X if f is continuous at every x ∈ X. Define Dom(f ) = {x ∈ X : f (x) ∈ R} and denote by C1−0(X) the set of all locally Lipschitz continuous real-valued functions defined on X. The following assumptions will also be used. (P4) T : Ω → [0, +∞] is continuous on Ω and T ∈ C1−0(Dom(T )); ON A COMPACT TRACE EMBEDDING THEOREM IN MUSIELAK-SOBOLEV SPACES ∗ 5 (P5) A∗ ∈ C1−0(Dom(A∗)) and there exist positive constants δ0 < 1 N , C0 and T (x) such that 0 < t0 < min x∈Ω ∇xA∗(x, t) ≤ C0(A∗(x, t))1+δ0 , j = 1, · · · , N, for x ∈ Ω and t ∈ [t0, T (x)) provided ∇xA∗(x, t) exists. Let A, B ∈ N (Ω), we say that A ≪ B if, for any k > 0, lim t→+∞ A(x, kt) B(x, t) = 0 uniformly for x ∈ Ω. The following theorems are embedding theorems for Musielak-Sobolev spaces developed by Fan in [7] and Liu and Zhao in [11]. Theorem 2.7. (See [7].) Let (P1)-(P5) holds. Then (1) There is a continuous embedding W 1,A(Ω) ֒→ LA∗(Ω); (2) Suppose that B ∈ N (Ω), B : Ω × [0, +∞) → [0, +∞) is continuous, and B(x, t) ∈ (0, +∞) for x ∈ Ω and t > 0. If B ≪ A∗, then there is a compact embedding W 1,A(Ω) ֒→֒→ LB(Ω). To give the boundary trace embedding theorem, we need the following assump- tion: (P 1 ∂ ) A is an N (Ω)-function such that T (x) defined in Equation (2.2) satisfies (2.3) T (x) = +∞, ∀x ∈ Ω. The following theorem is trace embedding theorem in [11]. Theorem 2.8. (See [11], Theorem 4.2.) Let Ω ⊂ RN be an open bounded domain with Lipschitz boundary. Suppose that A ∈ N (Ω) satisfies (P 1)-(P 5) and (P 1 ∂ ). N −1 Then there is a continuous boundary trace embedding W 1,A(Ω) ֒→ LA ∗ N (∂Ω), in N −1 which A ∗ N (x, t) = [A∗(x, t)] N −1 N , ∀x ∈ ∂Ω, t ∈ R+. The following theorem is a compact trace embedding theorem in [11]. Theorem 2.9. (See [11], Theorem 4.4.) Let Ω ⊂ RN be an open bounded domain with Lipschitz boundary. Suppose A ∈ N (Ω) satisfies (P 1)-(P 5) and (P 1 ∂ ). There ∈ ∈ ∆2(Ωδ) and A ∗ exist two constants ǫ > 0 and δ > 0 such that A ∗ N (Ωδ). Then for any θ ∈ N (∂Ω) satisfies N −1 N − ǫ 2 N −1 N −ǫ (2.4) the boundary trace embedding W 1,A(Ω) ֒→ Lθ(∂Ω) is compact. (x, t) for any t > 1 and x ∈ ∂Ω, θ(x, t) ≤ A ∗ N −1 N −ǫ We claim that Condition (2.4) is strong for the compact embedding in Theorem 2.9. In fact, some more weaker assumptions are needed instead in the theorem. We give a stronger result of the compact trace embedding theorem in Musielak-Sobolev sapces as follows. Theorem 2.10. Suppose Ω ⊂ RN is a bounded domain with Lipschitz boundary. A ∈ N (Ω) satisfies (A2), (P 1)-(P 5) and (P 1 ∂ ). Then for any continuous θ ∈ N (∂Ω) satisfying θ(x, t) ∈ (0, +∞) for any x ∈ ∂Ω, t > 0 and θ ≪ A , the boundary ∗ trace embedding W 1,A(Ω) ֒→ Lθ(∂Ω) is compact. N −1 N Before we give a proof of Theorem 2.10, we need a lemma in [7] or [1]. 6 WANG AND LIU Lemma 2.1. (See Lemma 3.9 in [7].) Let A, B ∈ N (Ω). Suppose that B : Ω × [0, +∞) → [0, +∞) is continuous, B ≪ A and for x ∈ Ω, t ∈ (0, +∞), B(x, t) > 0. If a sequence {un} is bounded in LA(Ω), and convergent in measure on Ω, then it is convergent in norm in LB(Ω). The proof of Theorem 2.10. By Propositon 2.5, we know W 1,A(Ω) ֒→ W 1,1(Ω). And by the boundary compact embedding in Lebesgue-Sobolev space, we have W 1,A(Ω) ֒→ W 1,1(Ω) ֒→֒→ L1(∂Ω). Then for any bounded sequence in W 1,A(Ω), there exists a subsequence such that it is convergent in L1(∂Ω). Then by the Lebesgue theorem, the subsequence is convergent in measure on ∂Ω. At the same time by Theorem 2.8 the subsequence we have found is bounded in LA∗ Then by Lemma 2.1, this subsequence is convergent in Lθ(∂Ω). N (∂Ω). N −1 Remark 2.2. It is clear that θ(x, t) ≤ A ∗ N −1 N −ǫ (x, t) for any t > 1 and x ∈ ∂Ω N −1 implies θ ≪ A ∗ 2.9. N . Then it is clear that Theorem 2.10 is stronger than Theorem Next we claim that not only variable exponent Sobolev spaces satisfy the con- ditions in Theorem 2.10 (see the following Example 2.1), but also some more complex case also satisfies conditions in Theorem 2.10 (see the following Example 2.2). And the current existing theories in mathematical literature can not cover the case in Example 2.2. Example 2.1. Let p ∈ C1−0(Ω) and 1 < q ≤ p(x) ≤ p+ := supx∈Ωp(x) < n (q ∈ R) for x ∈ Ω. Define A : Ω × R → [0, +∞) by A(x, t) = tp(x). Then it is readily checked that A satisfies (A2), (P 1), (P 2) and (P 3). It is easy to see that p ∈ C1−0(Ω) means A ∈ C1−0(Ω) and for s > 0 (2.5) A−1 ∗ (x, s) = Then T (x) = +∞, which implies that (P 1 In additional for x ∈ Ω np(x) n−p(x) np(x) . s n − p(x) ∂ ) is satisfied (⇒ (P 4) is satisfied). ∇xA(x, t) = tp(x)(ln t)∇p(x). Since for any ǫ > 0, ln t δ1 < 1 n , c1 and t1 such that tǫ → 0 as t → +∞, we conclude that there exist constants ∂A(x, t) ∂xj (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ≤ c1A1+δ1 (x, t) for all x ∈ Ω and t ≥ t1. Combining by A ∈ ∆2(Ω), from Proposition 3.1 in [7] it is easy to see that Condition (P 5) is satisfied. By now conditions in Theorem 2.10 are verified. By (2.5), we can see that A∗(x, t) =(cid:18) n − p(x) np(x) n−p(x) t(cid:19) np(x) . n−1 n Then A ∗ n −ǫ n−1 that A ∗ ∈ ∆2(Ω) and since p(x) ≤ p+ < n we can see there exists a ǫ > 0 such ∈ N (Ωδ). Then all conditions in Theorem 2.10 can be satisfied. ON A COMPACT TRACE EMBEDDING THEOREM IN MUSIELAK-SOBOLEV SPACES ∗ 7 Example 2.2. Let p ∈ C1−0(Ω) satisfy 1 < p− ≤ p(x) ≤ p+ := supx∈Ωp(x) < n−1. Define A : Ω × R → [0, +∞) by A(x, t) = tp(x) log(1 + t), It is obvious that A satisfies (A2), (P 1), (P 2) and (P 3). Pick ǫ > 0 small enough such that p+ + ǫ < n. Then for t > 0 big enough, A(x, t) ≤ ctp++ǫ, which implies that T (x) = +∞ for all x ∈ Ω. Thus (P 1 ∂ ) is satisfied (⇒ (P4) is satisfied). Since p ∈ C1−0(Ω) and A ∈ C1−0(Ω × R), by Proposition 3.1 in [7], A∗ ∈ C1−0(Ω × R). Combining A ∈ ∆2(Ω), it is easy to see that condition (P 5) is satisfied. Then conditions in Theorem 2.10 are verified. 3. Variational structures in Musielak-Sobolev spaces Firstly, we give some assumptions of the differential operators in Musielak- Sobolev spaces. We say that A : Ω × [0, +∞) → [0, +∞) satisfies Condition (A ), if both A and (P 1 eA ∈ N (Ω) are locally integrable and satisfy ∆2(Ω), A satisfies (A2), (P1)-(P5), ∂ ) in Section 2 and (A∞) below. (A∞) There exists a continous function A∞ : [0, +∞) → [0, +∞), such that A∞(α) → +∞ as α → +∞ and A(x, αt) ≥ A∞(α)αA(x, t) for any x ∈ Ω, t ∈ [0, +∞) and α > 0. We always assume that Condition (A ) holds in this paper. And by Proposition 2.6, the Musielak-Sobolev space W 1,A(Ω) is a separable and reflexive Banach space under our assumptions in (A ). We say that a function a1 : Ω × R → R satisfies (A1), if the following conditions (A11) − (A14) are satisfied: (A11) a1(x, t) : Ω × R → R is a Carath´eodory function; (A12) There exists a positive constant b1 such that a1(x, t)t2 ≤ b1A(x, t) for x ∈ Ω and t ∈ R; (A13) There exists a positive constant b2 such that a1(x, t)t ≥ b2a(x, t) for x ∈ Ω and t ∈ R; (A14) (a1(x, ξ)ξ − a1(x, η)η)(ξ − η) > 0 for x ∈ Ω and ξ, η ∈ RN . We say that a function a0 : Ω × R → R satisfies (A0), if the following conditions (A01)-(A04) are satisfied: (A01) a0(x, t) : Ω × R → R is a Carath´eodory function; (A02) There exists a positive constant b1 such that a0(x, t)t ≤ b1A(x, t) for x ∈ Ω and t ∈ R; (A03) There exists a positive constant b2 such that a0(x, t) ≥ b2a(x, t) for x ∈ Ω and t ∈ R; (A04) (a0(x, t1) − a0(x, t2))(t1 − t2) > 0 for x ∈ Ω and t1, t2 ∈ R. Assume X is a real reflexive Banach space. We call the map T : X → X ∗ is of type S+, if for any weakly convergent sequence un ⇀ u0 in X satisfying lim n→∞ hT (un), un − u0i ≤ 0, we have un → u0 in X. 8 WANG AND LIU Define A′ : W 1,A(Ω) → (W 1,A(Ω))∗ by hA′(u), vi =ZΩ a1(x, ∇u)∇u∇v + a0(x, u)v dx. Remark 3.1. By Theorem 2.1 and Theorem 2.2 in [8], A′ is a bounded, continuous, coercive, strictly monotone homeomorphic and S+ type operator. 4. Existence results of weak solutions In this section, we consider the existence of solutions to the following Neumann boundary value problem (4.1) (−div(a1(x, ∇u)∇u) + a0(x, u) = 0, a1(x, ∇u) ∂u ∂n = g(x, u) in Ω, on ∂Ω, where Ω is a bounded smooth domain in RN , a0, a1 : Ω × R → R are functions satisfying (A0) and (A1), the A(x, t) in (A0) and (A1) satisfies (A ), and ∂u ∂n is the outward unit normal derivative on ∂Ω. We say that u is a weak solution for (4.1) if ZΩ a1(x, ∇u)∇u∇ϕ + a0(x, u)ϕ dx =Z∂Ω g(x, u)ϕ dσ, ∀ϕ ∈ W 1,A(Ω). Theorem 4.1. If g(x, u) = g(x) and g ∈ L weak solution. ^ N −1 A N ∗ (∂Ω), then (4.1) has a unique Theorem 4.1 follows from Remark 3.1 and Proposition 2.6 immediately. Next, we assume the following conditions on g, (G1) Ψ ∈ N (∂Ω) ∩ ∆2(∂Ω) with Ψ ≪ A ∗ exists a constant K1 > 0 such that N N −1 , ψ(x, s) := ∂Ψ(x,s) ∂s exists, and there g(x, t) ≤ K1ψ(x, t) + h(x), x ∈ ∂Ω, t ∈ R, where 0 ≤ h ∈ L eΨ(∂Ω). (G2) There exist θ > 0, C > 0, and M > 0, such that for s ≥ M , 0 < θG(x, s) ≤ sg(x, s), ∀x ∈ ∂Ω, where θ > Ca := max{ lim s→∞ a1(x,s)s2 A1(x,s) , lim s→∞ a0(x,s)s A0(x,s) }. Remark 4.1. In fact, by assumptions (A02), (A03), (A12) and (A13), we have Ca ∈ [1, b1 ]. The Condition (G2) is the generalized Ambrosetti-Rabinowitz condition for b2 the nonlinear elliptic operator. Lemma 4.1. Let (G1) hold and set G(u) := R∂Ω weakly continuous in W 1,A(Ω) and G′ : W 1,A(Ω) → (W 1,A(Ω))∗ is a completely continuous operator. G(x, u) dσ. Then G is sequentially Proof. Suppose un ⇀ u weakly in W 1,A(Ω), to prove G is sequentially weakly continuous, it is sufficient to prove that n→∞Z∂Ω lim G(x, un) − G(x, u) dσ = 0. ON A COMPACT TRACE EMBEDDING THEOREM IN MUSIELAK-SOBOLEV SPACES ∗ 9 In fact, (cid:12)(cid:12)(cid:12)(cid:12)Z∂Ω G(x, un) − G(x, u) dσ(cid:12)(cid:12)(cid:12)(cid:12) ≤Z∂Ω ≤ K1Z∂Ω G(x, un) − G(x, u) dσ Ψ(x, un) − Ψ(x, u) dσ + 2kun − ukLΨ(∂Ω)khkL eΨ(∂Ω). From assumption (G1), the compact embedding W 1,A(Ω) ֒→ LΨ(∂Ω) and Propo- sition 2.3 in [10], we know that the right hand side of the above inequality tends to 0. g(x, u)v dσ, by the compact embedding W 1,A(Ω) ֒→ LΨ(∂Ω), Since hG′(u), vi = R∂Ω G′ is a completely continuous operator. (cid:3) Under assumption (A∞), the following assumptions on A and Ψ will be used. (Ψ∞) Ψ(x, αt) ≤ Ψ∞(α)αΨ(x, t), in which Ψ∞ : R+ → R+ is a nondecreasing function and satisfies and (4.2) Ψ∞(α) → ∞, as α → ∞ lim t→∞ Ψ∞(Ct) A∞(t) = 0, ∀ C > 0; (Ψ0) Ψ(x, αt) ≤ Ψ0(α)αΨ(x, t), in which Ψ0 : R+ → R+ is a nondecreasing function and satisfies A∞(α) Ψ0(Cα) → ∞, as α → 0, ∀C > 0; (A0) There exists a function A0 : (0, +∞) → (0, +∞) such that A(x, αt) ≤ A0(α)αA(x, t); (G3) g(x, t) = −g(x, −t), for x ∈ ∂Ω, t ∈ R; (G4) There is a function G0 : (0, +∞) → (0, +∞) satisfying G(x, αt) ≥ G0(α)αG(x, t) and G0(α) A0(α) → ∞, as α → 0 for x ∈ ∂Ω, t ∈ R, α > 0. Theorem 4.2. If g(x, u) satisfies Condition (G1) and (Ψ∞), then (4.1) has a weak solution. Define the functional J(u) corresponding (4.1) by A1(x, ∇u) + A0(x, u) dx −Z∂Ω G(x, u) dσ, J(u) =ZΩ in which A1(x, t) =R t 0 g(x, s) ds. Under the assumptions in Theorem 4.2, J(u) is well defined in W 1,A(Ω) and J ∈ C1. 0 a1(x, s)s ds, A0(x, t) =R t 0 a0(x, s) ds, G(x, t) =R t Remark 4.2. Under assumption (G1), by Lemma 4.1 and Remark 3.1, we know that J ′ is of type S+. 10 WANG AND LIU The proof of Theorem 4.2. By (A03), (A13), (G1) and (Ψ∞), we obtain A1(x, ∇u) + A0(x, u) dx −Z∂Ω A(x, ∇u) + A(x, u) dx −Z∂Ω J(u) =ZΩ ≥ b2ZΩ ≥ b2A∞(kukρ)kukρZΩ A(x, ∇u kukρ ) + A(x, u kukρ ) dx G(x, u) dσ Ψ(x, u) + u(x)h(x) dσ − K1Ψ∞(kukΨ)kukΨZ∂Ω Ψ(x, u kukΨ ) dσ − kukΨkhkΨ ≥ b2A∞(kukρ)kukρ − C1Ψ∞(ckukρ)kukρ − CkukρkhkΨ = (b2A∞(kukρ) − C1Ψ∞(ckukρ) − CkhkΨ)kukρ ≥ b2 2 A∞(kukρ)kukρ → ∞, as kukρ → ∞. The last inequality sign follows from (4.2). By Lemma 4.1 and Remark 3.1, we know that J is weakly lower semicontinuous in W 1,A(Ω). Then J admits a minimum in W 1,A(Ω) which is a weak solution for (4.1). The proof is completed. Lemma 4.2. Under the assumptions (G1) and (G2), J satisfies the P.S. condition. Proof. Let {un}n∈N be a P.S. sequence of J, i. e., J(un) → C, J ′(un) → 0. Fixing β ∈ ( 1 Ca+ε ) with ε > 0 small enough, by the definition of Ca, there exists Cε > 0, θ , such that 1 C + 1 + kunk ≥J(un) − βhJ ′(un), uni =ZΩ A1(x, ∇un) + A0(x, un) dx − βZΩ + a0(x, un)un dx +Z∂Ω a1(x, ∇un)∇un2 βg(x, un)un − G(x, un) dσ A1(x, ∇un) + A0(x, un) dx+ ≥(1 − β(Ca + ε))ZΩ Z∂Ω βg(x, un)un − G(x, un) dσ − Cε ≥(1 − β(Ca + ε))b2A∞(kunk)kunk − Cε. From the above inequality, it is easy to see that {un} is bounded in W 1,A(Ω). Then there exists a subsequence of {un} (still denoted by {un}) such that un ⇀ u0 in W 1,A(Ω). So we conclude that hJ ′(un) − J ′(u0), un − u0i → 0. Since J ′ is a S+ type operator, we can conclude that un → u0. (cid:3) Theorem 4.3. Under the assumptions (G1), (G2), (Ψ0) and h ≡ 0 in (G1), (4.1) has a nontrivial solution. Proof. It is sufficient to verify that J satisfies conditions in Mountain Pass Lemma (see [14]). ON A COMPACT TRACE EMBEDDING THEOREM IN MUSIELAK-SOBOLEV SPACES ∗ 11 By (G1) and (Ψ0), we obtian J(u) ≥b2A∞(kuk)kuk − K1Z∂Ω Ψ(x, u kukΨ kukΨ) dσ ≥b2A∞(kuk)kuk − K1Ψ0(kukΨ)kukΨ ≥(cid:2)b2A∞(kuk) − C1Ψ0(Ckuk)(cid:3)kuk. By the assumption (Ψ0), there exists a small γ > 0 such that b2A∞(kuk) − C1Ψ0(Ckuk) > 0 for 0 < kuk ≤ γ. Hence, for kuk = γ, there exists a δ > 0 such that J(u) ≥ δ > 0. Next we will show that there exists a w ∈ W 1,A(Ω) with kwk > ρ such that J(w) < 0. In fact, for any w ∈ W 1,A(Ω) ∩ W 1,Ca+ε(Ω) with ε small enough such that Ca + ε < θ, we can get J(tw) =ZΩ A1(x, t∇w) + A0(x, tw) dx −Z∂Ω G(x, tw) dσ ≤C1tCa+εZΩ ∇wCa+ε + wCa+ε dx − C2tθZ∂Ω wθ dσ + Cε. Let t → ∞, we have J(tw) → −∞, which completes the proof. (cid:3) Denote X := W 1,A(Ω). Then under our assumptions in (A ), there exists {ej}∞ j=1 ⊂ X and {e∗ j }∞ j=1 ⊂ X ∗ such that i , eji =( 1, 0, he∗ i = j, i 6= j; and X = span{ej, j = 1, 2, · · · }, X ∗ = span{e∗ j , j = 1, 2, · · · }. For k = 1, 2, · · · , denote Xj = span{ej}, Yk = Xj, Zk = kMj=1 Xj. ∞Mj=k Lemma 4.3. Let Ψ ≪ A ∗ N , and denote N −1 βk = sup(cid:26)Z∂Ω Ψ(x, u) + h(x)u dσ : kuk ≤ ϑ, u ∈ Zk(cid:27), in which ϑ is any given positive number. Then lim k→∞ βk = 0. Ψ(x, uk) + h(x)uk dσ − β < 1 that 0 ≤ R∂Ω Proof. It is obvious that 0 < βk+1 ≤ βk. Then there exists a β ≥ 0 such that βk → β as k → ∞. By the definition of βk, there exist uk ∈ Zk and kukk ≤ ϑ such k . Then there exists a subsequence of {uk} (still denoted by {uk}) satisfying uk ⇀ u in W 1,A(Ω) and uk → u in LΨ(∂Ω). By the definition of Zk we obtain j , ui = lim k→∞ ∀j = 1, 2, · · · . j , uki = 0, he∗ he∗ Then u = 0, which means uk → 0 in LΨ(∂Ω). βk → 0 follows by Proposition 2.3 in [10] and the assumption h ∈ L (cid:3) eΨ(∂Ω). 12 WANG AND LIU Theorem 4.4. Let (G1), (G3), (G4) and (Ψ∞) hold. Then (4.1) has a sequence of solutions {un}n∈N with negative energy and I(un) → 0. Proof. By Condition (G1) and (Ψ∞), the functional J(u) is coercive and satisfies the P.S. condition. By (G3), J(u) is an even functional. Denote by γ(U ) the genus of U ∈ Σ := {U ⊂ W 1,A(Ω)\{0} : U is compact and U = −U } (see [13, 14]). Set Σk := {U ∈ Σ : γ(U ) ≥ k}, k = 1, 2, · · · , ck := inf U∈Σk I(u), k = 1, 2, · · · . sup u∈U Then it is clear that −∞ < c1 ≤ c2 ≤ · · · ≤ ck ≤ ck+1 ≤ · · · . Next we will show that ck < 0 for every k ∈ N. In fact, selecting Ek be the k-dimensional subspaces of W 1,A(Ω) with Ek ⊂ C∞(Ω) such that u∂Ω 6≡ 0 for all u ∈ Ek\{0}, by (G4) and (A0), we have for u ∈ Ek, kuk = 1, J(tu) =ZΩ A1(x, t∇u) + A0(x, tu) dx −Z∂Ω G(x, tu) dσ ≤A0(t)tZΩ A1(x, ∇u) + A0(x, u) dx − G0(t)tZ∂Ω G(x, u) dσ, where :=A0(t)tak − G0(t)tgk, ak = sup(cid:26)ZΩ gk = inf(cid:26)Z∂Ω A1(x, ∇u) + A0(x, u) dx : u ∈ Ek, kuk = 1(cid:27), G(x, u) dσ : u ∈ Ek, kuk = 1(cid:27). Observing that 0 < ak < +∞ and gk > 0 since Ek is of finite dimensional, by the assumption (G4), there exists positive constants ρ and ε such that J(ρu) < −ε for u ∈ Ek, kuk = 1. Therefore, if we set Sρ,k = {u ∈ Ek : kuk = ρ}, then Sρ,k ⊂ J −ε. Hence, by the monotonicity of the genus γ(J −ε) ≥ γ(Sρ,k) = k, By the minimax argument, each ck is a critical value of J (see [13]), hence there is a sequence of weak solutions {±uk, k = 1, 2, · · · } of the problem (4.1) such that J(uk) = ck < 0. Next we will show that ck → 0. In fact, by (G1) and Lemma 4.3, for u ∈ Zk, kuk ≤ ϑ, we obtain J(u) ≥ZΩ ≥ −Z∂Ω A1(x, ∇u) + A0(x, u) dx −Z∂Ω Ψ(x, u) + h(x)u dσ ≥ −βk. Ψ(x, u) + h(x)u dσ ON A COMPACT TRACE EMBEDDING THEOREM IN MUSIELAK-SOBOLEV SPACES ∗ 13 It is clear that for U ∈ Σk, we have γ(U ) ≥ k. Then according to the properties of genus, we conclude U ∩ Zk 6= ∅ for U ∈ Σk. Then sup J(u) ≥ −βk. u∈U∈Σk By the definition of ck, it is easy to see that 0 ≥ ck ≥ −βk, which implies ck → 0 by βk → 0 from Lemma 4.3. (cid:3) 5. Acknowledgments The research is partly supported by the National Natural Science Foundation of China (NSFC 11501268 and 11471147) and the Fundamental Research Funds for the Central Universities (lzujbky-2016-101). The authors express thanks to professor Peihao Zhao for his valuable suggestions on the paper. References [1] R. Adams, J. Fournier, Sobolev Spaces, second ed., Acad. Press, New York, 2003. [2] A. Benkirane, M. Sidi EI Vally, An existence result for nonlinear elliptic equations in Musielak- Orlicz-Sobolev spaces, Bull. Belg. Math. Soc., 20(1), 2013, 57-75. [3] S. Byun, J. Ok, and L. Wang, W 1,p(·)-Regularity for elliptic eqautions with measurable coefficients in nonsmooth domains. Communications in Mathematical Physics, 329, 2014, 937 -- 958. [4] T. K. Donaldson and N. S. Trudinger, Orlicz-sobolev spaces and imbedding theorems. J. Func. Anal., 8, 1971, 52 -- 75. [5] Michela Eleuteri, Paolo Marcellini and Elvira Mascolo, Lipschitz continuity for energy inte- grals with variable exponents. Atti Accad. Naz. Lincei Cl. Sci. Fis. Mat. Natur., 27 (2016), 2016, 61-87. [6] X. Fan, Boundary trace embedding theorems for variable exponent Sobolev spaces. J. Math. Anal. Appl., 339, 2008, 1395 -- 1412. [7] X. Fan, An embedding theorem for Musielak-Sobolv spaces, Nolinear Anal., 75, 2012, 1959- 1971. [8] X. Fan, Differential equations of divergence form in Musielak-Sobolev spaces and a sub- supersolution method, J. Math. Anal. Appl., 386, 2012, 593-604. [9] X. L. Fan and Q. H. Zhang, Existence of solutions for p(x)-Laplacian Dirichlet problem. Nonlinear Analysis, 52, 2003, 1843 -- 1852. [10] D. Liu, P. Zhao, Solutions for a quasilinear elliptic equation in Musielak-Sobolev spaces, Nonlinear Analysis: Real World Applications, 26, 2015, 315-329. [11] D. Liu, B. Wang and P. Zhao, On the trace regularity results of Musielak-Orlicz-Sbolev spaces in a bounded domain. Communications on Pure and Applied Analysis, 15(5), 2016, 1643 -- 1659. [12] M. Mihailescu, P. Pucci and V. Radulescu, Eigenvalue problems for anisotropicquasilinear elliptic equations with variable exponent. J. Math. Anal. Appl., 340 (2008), 2008, 687 -- 698. [13] M. Struwe, Variational Methods, Third Edition, Springer, Berlin, 1996. [14] M. Willem, Minimax Theorems, Birkhauser, Bostom, 1996. (Li Wang) School of Mathematics and Statistics, Lanzhou University, Lanzhou 730000, P. R. China E-mail address: [email protected] (Duchao Liu) School of Mathematics and Statistics, Lanzhou University, Lanzhou 730000, P. R. China E-mail address: [email protected], Tel.: +8613893289235, fax: +8609318912481
1805.03646
4
1805
2019-10-09T14:26:11
Pre-Markov Operators
[ "math.FA", "math.OA" ]
A positive linear operator $T$ between two unital $f$-algebras, with point separating order duals, $A$ and $B$ is called a Markov operator for which $% T\left( e_{1}\right) =e_{2}$ where $e_{1},e_{2}$ are the identities of $A$ and $B$ respectively. Let $A$ and $B$ be semiprime $f$-algebras with point separating order duals such that their second order duals $A^{\sim \sim }$ and $B^{\sim \sim }$ are unital $f$-algebras. In this case, we will call a positive linear operator $T:A\rightarrow B$ \ to be a Pre-Markov operator, if the second adjoint operator of $T$ is a Markov operator. A positive linear operator $T$ between two semiprime $f$-algebras, with point separating order duals, $A$ and $B$ is said to be contractive if $Ta\in B\cap \left[ 0,I_{B}\right] $ whenever $a\in A\cap \left[ 0,I_{A}\right] $, where $I_{A}$ and $I_{B}$ are the identity operators on $A$ and $B$ respectively. In this paper we characterize pre-Markov algebra homomorphisms. In this regard, we show that a pre-Markov operator is an algebra homomorphism if and only if its second adjoint operator is an extreme point in the collection of all Markov operators from $A^{\sim \sim }$ to $B^{\sim \sim }$. Moreover we characterize extreme points of contractive mappings from $A$ to $B$. In addition, we give a condition that makes an order bounded algebra homomorphism is a lattice homomorphism.
math.FA
math
Pre-Markov Operators H¯ulya Duru and Serkan Ilter Abstract. A positive linear operator T between two unital f -algebras, with point separating order duals, A and B is called a Markov operator for which T (e1) = e2 where e1, e2 are the identities of A and B respectively. Let A and B be semiprime f -algebras with point separating order duals such that their second order duals A∼∼ and B∼∼ are unital f -algebras. In this case, we will call a positive linear operator T : A → B to be a pre-Markov operator, if the second adjoint operator of T is a Markov operator. A positive linear operator T between two semiprime f -algebras, with point separating order duals, A and B is said to be contractive if T a ∈ B ∩ [0, IB] whenever a ∈ A ∩ [0, IA], where IA and IB are the identity operators on A and B respectively. In this paper we characterize pre-Markov algebra homomorphisms. In this regard, we show that a pre-Markov operator is an algebra homomorphism if and only if its second adjoint operator is an extreme point in the collection of all Markov operators from A∼∼ to B∼∼. Moreover we characterize extreme points of contractive operators from A to B. In addition, we give a condition that makes an order bounded algebra homomorphism is a lattice homomorphism. 1. Introduction A positive linear operator T between two unital f -algebras, with point sepa- rating order duals, A and B is called a Markov operator for which T (e1) = e2 where e1, e2 are the identities of A and B respectively. Let A and B be semiprime f -algebras with point separating order duals such that their second order duals A∼∼ and B∼∼ are unital f -algebras. In this case, we will call a positive linear operator T : A → B to be a pre-Markov operator, if the second adjoint operator of T is a Markov operator. A positive linear operator T between two semiprime f -algebras, with point separating order duals, A and B is said to be contractive if T a ∈ B ∩ [0, IB] whenever a ∈ A ∩ [0, IA], where IA and IB are the identity operators on A and B respectively. The collection of all pre-Markov operators is a convex set. In this paper, first of all, we characterize pre-Markov algebra homomorphisms. In this regard, we show that a pre-Markov operator is an algebra homomorphism if and only if its second adjoint operator is an extreme point in the collection of all Markov operators from A∼∼ to B∼∼ (Theorem 1). In addition, we characterize the extreme points of all contractive operators T : A → B whenever A and B are Archimedean semiprime 2010 Mathematics Subject Classification. Primary 47B38, 46B42. Key words and phrases. Markov Operator, f-algebra, algebra homomorphism, lattice homo- morphism,contractive operator, Arens Multiplication. 1 2 H ¯ULYA DURU AND SERKAN ILTER f-algebras provided B is relatively uniformly complete (Proposition 3). For the second aim, let A and B be Archimedean semiprime f-algebras and T : A → B a linear operator. Huijsman and de Pagter proved in [8] the following: (1) If T is a positive algebra homomorphism then it s a lattice homomorphism. (2) In addition, if the domain A is relatively uniformly complete and T is an algebra homomorphism then it s a lattice homomorphism and the assumption that the domain A of T is relatively uniformly complete is not reduntant (Theorem 5.1 and Example 5.2.). (3) In addition, if the domain A has a unit element and T is an order bounded algebra homomorphism then it s a lattice homomorphism (Theorem 5.3). We prove that any order bounded algebra homomorphism T : A → B is a lattice homomorphism, if the region B is relatively uniformly complete (Corollary 3). In this regard, first we give an alternate proof of Lemma 6 in [10] for order bounded operators with the relatively uniformly complete region instead of positive operators with Dedekind complete region (Proposition 4 and 5). In the last part, we give a necessary and sufficient condition for a positive operator to be a lattice homomorphism (Proposition 7). 2. Preliminaries For unexplained terminology and the basic results on vector lattices and semiprime f -algebras we refer to [1, 11, 13, 15]. The real algebra A is called a Riesz algebra or lattice-ordered algebra if A is a Riesz space such that ab ∈ A whenever a, b are positive elements in A. The Riesz algebra is called an f -algebra if A satisfies the condition that a ∧ b = 0 implies ac ∧ b = ca ∧ b = 0 for all 0 ≤ c ∈ A. In an Archimedean f -algebra A, all nilpotent elements have index 2. Indeed, assume that a3 = 0 for some 0 ≤ a ∈ A. Since the equality (cid:0)a2 − na(cid:1) ∧ (cid:0)a − na2(cid:1) = 0 implies (cid:0)a2 − na(cid:1) ∧ a2 = (cid:0)a2 − na(cid:1) = 0 we get a2 = 0 as A is Archimedean. The same argument is true for all n ≥ 3. Throughout this paper A will show an Archimedean semiprime f -algebra with point separating order dual A∼ [15]. By definition, if zero is the unique nilpotent element of A, that is, a2 = 0 implies a = 0, A is called semiprime f-algebra. It is well known that every f -algebra with unit element is semiprime. Let A be a lattice ordered algebra. If A is a lattice ordered space, then the first order dual space A∼ of A is defined as the collection of all order bounded linear functionals on A and A∼ is a Dedekind complete Riesz space. The second order dual space of A is denoted by A∼∼. Let a ∈ A, f ∈ A∼ and F, G ∈ A∼∼. Define f · a ∈ A∼, by and F · f ∈ A∼, by and F · G ∈ A∼∼, by (f · a) (b) = f (ab) (F · f ) (a) = F (f · a) (F · G) (f ) = F (G · f ) The last equality is called the Arens multiplication in A∼∼ [2]. The second order dual space A∼∼ of a semiprime f -algebra A is again an f - algebra with respect to the Arens multiplication [4]. In the literature, there are PRE-MARKOV 3 several studies, for example [5, 6, 7, 9], that respond the question "Under what conditions does the f -algebra A∼∼ have a unit element?". Let A and B be semiprime f -algebras with point separating order duals such that their second order duals A∼∼ and B∼∼ have unit elements E1 and E2 respec- tively. Let T : A → B be an order bounded operator. We denote the second adjoint operator of T by T ∗∗. Since A and B have point separating order duals, the linear operator J1 : A → A∼∼, which assigns to a ∈ A the linear functional ba defined on A∼ by ba (f ) = f (a) for all a ∈ A, is an injective algebra homomorphism. Therefore we will identify A with J1 (A), and B with J2 (B) in the similar sense. Definition 1. Let A and B be semiprime f -algebras with point separating order duals such that their second order duals A∼∼ and B∼∼ are unital f -algebras. In this case, we call a positive linear operator T : A → B to be a pre-Markov operator, if the second adjoint operator of T is a Markov operator. That is, the second adjoint operator T ∗∗ : A∼∼ → B∼∼ of T is a positive linear and T ∗∗ (E1) = E2, where E1 and E2 are the unitals of A and B respectively. Recall that a positive operator T : A → B satisfying 0 ≤ T (a) ≤ E2 whenever 0 ≤ a ≤ E1 is called a contractive operator. In this point we remark that , if A and B are semiprime f -algebras with point separating order duals and T : A → B is a positive linear operator, then T ∗∗ is positive. Indeed, let 0 ≤ F ∈ A∼∼ and 0 ≤ g ∈ B∼. Then 0 ≤ g ◦ T ∈ A∼ and therefore F (g ◦ T ) = T ∗∗ (F ) ≥ 0. Proposition 1. Let A and B be semiprime f -algebras with point separating order duals such that their second order duals A∼∼ and B∼∼ have unit elements E1 and E2 respectively. T : A → B is contractive if and only if T ∗∗ is contractive. Proof. Suppose that T is contractive. Then T ∗∗ is positive. Let F ∈ [0, E1] ∩ A∼∼. In order to prove that T ∗∗ is contractive we shall show that T ∗∗ (E1) ≤ E2. Due to [9], E1 (f ) = sup f (A ∩ [0, E1]) E2 (g) = sup g (B ∩ [0, E2]) for all f ∈ A∼ and g ∈ B∼. Let a ∈ A ∩ [0, E1] and 0 ≤ g ∈ B∼. Since T is contractive, T (a) ∈ B ∩ [0, E2] so g (T (a)) ≤ E2 (g) which implies that T ∗∗E1 (g) = E1 (g ◦ T ) ≤ E2 (g). Thus T ∗∗ (E1) ≤ E2. Conversely, assume that T ∗∗ is contractive. Let a ∈ A ∩ [0, E1] and 0 ≤ g ∈ B∼. Then cT a (g) = g (T a) ≤ T ∗∗E1 (g) ≤ E2 (g) Thus 0 ≤ T a = cT a ≤ E2. Corollary 1. Let A and B be semiprime f -algebras with point separating order duals such that their second order duals A∼∼ and B∼∼ have unit elements E1 and E2 respectively. If T : A → B is a pre-Markov operator then T is contractive. (cid:3) Proof. Since T ∗∗ (E1) = E2 and T ∗∗ is positive, T ∗∗ is contractive. By Propo- (cid:3) sition 1 we have the conclusion. For a ∈ A, the mapping πa : A → OrthA, defined by πa (b) = a.b is an orthomorphism on A. Since A is a Archimedean semiprime f -algebra, the mapping π : A → OrthA, defined by π (a) = πa is an injective f -algebra homomorphism. Hence we shall identify A with π (A). 4 H ¯ULYA DURU AND SERKAN ILTER 3. Main Results Theorem 1. Let A and B be semiprime f -algebras with point separating order duals such that their second order duals A∼∼ and B∼∼ have unit elements E1 and E2 respectively. A pre-Markov operator T : A → B is an algebra homomorphism if and only if its second adjoint operator T ∗∗ is an algebra homomorphism. Proof. Suppose that the pre-Markov operator T is an algebra homomorphism. Since T ∗∗ is a Markov operator, due to [8], it is enough to show that it is a lattice homomorphism. Let F, G ∈ A∼∼ such that F ∧ G = 0. Since A∼∼ and B∼∼ are semiprime f - algebras, F · G = 0. We shall show that T ∗∗ (F ) · T ∗∗ (G) = 0. Let a, b ∈ A and f ∈ B∼. Then it follows from the following equations ((f · T a) ◦ T ) (b) = (f · T a) (T b) = f (T aT b) = f (T (ab)) = (f ◦ T ) (ab) = ((f ◦ T ) · a) (b) that (3.1) (f · T a) ◦ T = (f ◦ T ) · a. On the other hand, the following equations ((G ◦ T ∗) · f ) ◦ T ) (a) = ((G ◦ T ∗) · f ) (T a) = (G ◦ T ∗) (f · T a) = G ((f · T a) ◦ T ) hold. Thus ((G ◦ T ∗) · f ) ◦ T ) (a) = G ((f · T a) ◦ T ). From here, by setting the equation (3.1), we conclude that ((G ◦ T ∗) · f ) ◦ T ) (a) = G ((f ◦ T ) · a) = (G · (f ◦ T )) (a) which implies (3.2) ((G ◦ T ∗) · f ) ◦ T ) = (G · (f ◦ T )) . Taking into account the equation (3.2), we get (T ∗∗ (F ) · T ∗∗ (G)) (f ) = T ∗∗ (F ) ((T ∗∗ (G) · f )) = (F ◦ T ∗) ((G ◦ T ∗) · f ) = F ((G ◦ T ∗) · f ) ◦ T ) = F (G · (f ◦ T )) thus we have (T ∗∗ (F ) · T ∗∗ (G)) (f ) = (F · G) (f ◦ T ) = 0 as desired. Conversely suppose that T ∗∗ is an algebra homomorphism. Let a, b ∈ A. It follows from T (ab) = \T (ab) = T ∗∗(cid:16)bab(cid:17) = T ∗∗(cid:16)ba ·bb(cid:17) = T ∗∗ (ba) · T ∗∗(cid:16)bb(cid:17) = cT a · cT b = T a.T b that T is an algebra homomorphism. (cid:3) In the proof of Theorem 1 we proved the following corollary as well; Corollary 2. Let A, B and their second order duals A∼∼ and B∼∼ be semiprime f -algebras and T : A → B a positive algebra homomorphism. Then T ∗∗ is a lattice homomorphism. Theorem 2. Let A and B be semiprime f -algebras with point separating order duals and T : A → B a positive linear operator. If the second order duals A∼∼ and B∼∼ have unit elements and T is an algebra homomorphism, then T is an extreme point of the contractive operators from A to B. PRE-MARKOV 5 Proof. Suppose that T is a positive algebra homomorphism. Then due to [14, Theorem 4.3], T is a contractive operator. Let 2T = T1 + T2 for some contractive operators T1, T2 from A to B. In this case, 2T ∗∗ = T ∗∗ 2 . By Proposition 1, T ∗∗, 2 are contractive and by Corollary 2, T ∗∗ is a lattice homomorphism. T ∗∗ Taking into account [3, Theorem 3.3], we derive that T ∗∗ is an extreme point in the collection of all contractive operators from A∼∼ to B∼∼. Thus T ∗∗ = T ∗∗ 1 = T ∗∗ 2 and therefore T = T1 = T2. (cid:3) 1 and T ∗∗ 1 +T ∗∗ At this point, we recall the definition of uniform completion of an Archimedean Riesz space. If A is an Archimedean Riesz space and bA is the Dedekind completion of A, then A, the closure of A in bA with respect to the relatively uniform topology [11], is so called that relatively uniformly completion of A [12]. If A is an semiprime f -algebra then the multiplication in A can be extended in a unique way into a lattice ordered algebra multiplication on A such that A becomes a sub-algebra of A and A is an relatively uniformly complete semiprime f -algebra. In [14, Theorem 3.4] it is shown that a positive operator T from a Riesz space A to a uniformly complete space B, has a unique positive linear extension T : A → B to the relatively uniformly completion A of A, defined by, T (x) = sup {T (a) : 0 ≤ a ≤ x} for 0 ≤ x ∈ A. We also recall that A satisfies the Stone condition (that is, x ∧ nI ∗ ∈ A, for all x ∈ A, where I denotes the identity on A of OrthA) due to Theorem 2.5 in [7]. For the completeness we give the easy proof of the following proposition. Proposition 2. Let A and B be Archimedean semiprime f -algebras such that B is relatively uniformly complete. In this case, T : A → B is contractive if and only if T is contractive. Proof. Suppose that T is contractive. Let x ∈ A ∩ (cid:2)0, I(cid:3), here I is the unique extension to A of the identity mapping I : A → A. Since T is contractive, a ∈ A ∩ [0, x] implies that I is an upper bound for the set {T (a) : a ≤ x, a ∈ A}, so T (x) ≤ I. Therefore T is contractive. The converse implication is trivial, since T is the extension of T , we get 0 ≤ T (a) = T (a) ≤ I whenever a ∈ A ∩ [0, I]. (cid:3) Proposition 3. Let A and B be Archimedean semiprime f - algebras such that B is relatively uniformly complete and let T : A → B be a contractive operator. Then T is an extreme point in the collection of all contractive operators from A to B if and only if T is an extreme point of all contractive operators from A to B. Proof. Suppose that T is an extreme point in the set of all contractive op- erators from A to B. We shall show that for arbitrary ε > 0 and contractive operator S from A to B satisfying εT − S ≥ 0 implies that T = S. Let 0 ≤ x ∈ A. Then there exists a positive sequence (an)n in A converging relatively uniformly to x. Since T and S are relatively uniformly continuos, the sequence εT (an) −S (an) = εT (an) − S (an) converges to εT (x) −S (x). Therefore, since (an)n is positive sequence and εT − S ≥ 0, we get εT −S ≥ 0. Since T is an extreme point, we have T = S, so that T = S. Conversely assume that T is an extreme point in the set of all contractive operators from A to B. Let ε > 0 be any number and let S be any contractive operator from A to B satisfying εT −S ≥ 0. Let U be the restriction of S to A. Since S is contractive, by Proposition 2, S A= U 6 H ¯ULYA DURU AND SERKAN ILTER is contractive and by the uniquness of the extension, we infer that S = U . Hence (cid:0)εT − S(cid:1) A= εT − U ≥ 0. Thus T = S, which shows that T is an extreme (cid:3) point. After proving the following Propositions 4 and 5 for order bounded operators with the relatively uniformly complete region, we remarked that both were proved in [10] for the positive operators with Dedekind complete region. They might be regarded as the alternate proofs. Proposition 4. Let T : A → B be an order bounded operator where A and B are Archimedean f -algebras and B is, in addition, relatively uniformly complete. Then T is an algebra homomorphism iff T is an algebra homomorphism. Proof. Suppose that T : A → B is an algebra homomorphism and x, y be positive elements in A. By [14], since and xy = sup {Ry (a) : 0 ≤ a ≤ x, a ∈ A} Ry (a) = sup {ab : 0 ≤ b ≤ y, b ∈ A} . Now as T is relatively uniformly continuous, we get, T (Ry (a)) = sup(cid:8)T (ab) = T (ab) = T (a) T (b) : 0 ≤ b ≤ y, b ∈ A(cid:9) = T (a) sup {T (b) : 0 ≤ b ≤ y, b ∈ A} and then = T (a) T (y) T (xy) = sup(cid:8)T (Ry (a)) : 0 ≤ a ≤ x, a ∈ A(cid:9) = sup(cid:8)T (a) T (y) : 0 ≤ a ≤ x, a ∈ A(cid:9) = T (y) sup {T (a) : 0 ≤ a ≤ x, a ∈ A} = T (x) T (y) Hence T is an algebra homomorphism. The converse is trivial. (cid:3) In [8], both were proved that an algebra homomorphism T : A → B need not be a lattice homomorphism if the domain A is not relatively uniformly complete (Example 5.2) and an order bounded algebra homomorphism T : A → B is a lat- tice homomorphism whenever the domain A has a unit element. We remarked that Proposition 4 yields that the second result also holds for an order bounded alge- bra homomorphism without unitary domain but the region is relatively uniformly complete. Corollary 3. Let A be an Archimedean semiprime f -algebra and B a rela- tively uniformly complete Archimedean f -algebra. Then any order bounded algebra homomorphism T : A → B is a lattice homomorphism. Proof. By Proposition 4, T is an algebra homomorphism and since A is rel- atively uniformly complete, T is a lattice homomorphism [8]. Thus T is a lattice homomorphism. (cid:3) Proposition 5. Let A be an Archimedean f -algebra and let B be a relatively uniformly complete semiprime f - algebra. Then the operator T : A → B is a lattice homomorphism iff T is a lattice homomorphism. PRE-MARKOV 7 Proof. Suppose that T is a lattice homomorphism. Let x ∈ A. Let a ∈ [0, x+] ∩ A and b ∈ [0, x−] ∩ A. Since T is a lattice homomorphism, we have T (a ∧ b) = T (a) ∧ T (b) = 0. On the other hand, it follows from the equality T (a) ∧ T (cid:0)x−(cid:1) = sup(cid:8)T (a) ∧ T (b) : 0 ≤ b ≤ x−, b ∈ A(cid:9) that T (cid:0)x+(cid:1) T (cid:0)x−(cid:1) = sup(cid:8)T (a) ∧ T (cid:0)x−(cid:1) : 0 ≤ a ≤ x+, a ∈ A(cid:9) = 0 which its turn is equivalent to T is a lattice homomorphism, as B is semiprime. Converse is trivial. (cid:3) In this point, we remark that Lemma 3.1 and Theorem 3.3 in [3] are also true for Archimedean semiprime f -algebras without the Stone condition on the domain A whenever B is relatively uniformly complete. Proposition 6. Let A and B be Archimedean semiprime f -algebras, B rela- tively uniformly complete and T : A → B a contractive operator. Assume that A has unit element. For y ∈ A, define Hx (y) = T (xy) − T (x) T (y). Then T + Hx are contractive mappings for all x ∈ A ∩ [0, I]. Proof. By Proposition 2, T is contractive. Since A satisfies the Stone condi- (cid:3) tion, due to [3, Lemma 3.1], we have the conclusion. Corollary 4. Let A and B be Archimedean semiprime f -algebras such that B is relatively uniformly complete and let T : A → B be a contractive operator. If A has unit element, then T +Ta are contractive for all a ∈ A ∩ [0, I], here Ta (b) = T (ab) − T (a) T (b). Proof. By Proposition 6, T +Hx are contractive mappings for all x ∈ A∩[0, I]. Let a ∈ A ∩ [0, I] and 0 ≤ b ∈ A. Then 0 ≤ (cid:16)T + Ha(cid:17) (b) = T (b) + Ta (b) holds. Thus T + Ta is positive. Let b ∈ A ∩ [0, I]. It follows from 0 ≤ (cid:16)T + Ha(cid:17) (b) = T (b) + Ta (b) ≤ I that T +Ta are contractive. (cid:3) Proposition 7. Let A and B be Archimedean semiprime f -algebras such that B is relatively uniformly complete and let T : A → B be a positive linear operator. T is contractive and it is an extreme point in the collection of all contractive operators from A to B if and only if T is an algebra homomorphism. Proof. Let T be an extreme point in the collection of all contractive operators from A to B. Then by Proposition 3, T is an extreme point of all contractive operators from A to B. It follows from [3, Theorem 3.3] that T is an algebra homomorphism. By Proposition 4, T is an algebra homomorphism. Conversely, if T is an algebra homomorphism, then due to [14, Theorem 4.3], T is a contractive operator. By Proposition 2, T is contractive and by Proposition 4, T is an algebra homomorphism. Thus T is an extreme point in the set of all contractions from A to B due to [3, Theorem 3.3]. By using Proposition 3, we have the conclusion. (cid:3) 8 H ¯ULYA DURU AND SERKAN ILTER References [1] C. D. Aliprantis, O. Burkinshaw, Positive Operators, Academic Press, Orlando, 1985 [2] R. Arens, The adjoint of bilinear operation, Proc. Am. Math. Soc. 2 (1951) 839-848 [3] M. A. Ben Amor, K. Boulabiar, C. El Adeb, Extreme contractive operators on Stone f - algebras, Indagationes Mathematicae 25 (2014) 93-103 [4] S. J. Bernau, C.B. Huijsmans, The order bidual of almost f-algebras and d-algebras, Trans.Amer.Math.Soc. 347 (1995) 4259-4275 [5] K. Boulabiar, J. Jaber, The order bidual of f-algebras revisited, Positivity 15 (2011) 271-279 [6] C.B. Huijsmans, The order bidual of lattice-ordered algebras II, J.Operator Theory 22 (1989) 277-290 [7] C. B. Huijsmans, B. de Pagter, The order bidual of lattice-ordered algebras, J.Funct.Analysis 59 (1984) 41-64 [8] C. B. Huijsmans, B. de Pagter, Subalgebras and Riesz subspaces of an f-algebra, Proc.Lond.Math.Soc. 48 (1984) 161-174 [9] J. Jaber, f-algebras with σ−bounded approximate unit, Positivity, 18 (2014) 161-170 [10] J. Jaber, Contractive operators on semiprime f -algebras, Indagationes Mathematicae, 28 (2017) 1067-1075 [11] W. A. J. Luxemburg, A. C. Zaanen, Riesz Spaces I, North Holland, Amsterdam, 1971 [12] J. Quinn, Intermediate Riesz spaces, Pac.J.Math. 56 (1975) 255-263 [13] H. H. Schaefer, Banach Lattices and Positive Operators, Springer, Berlin, 1974 [14] A. Triki, Algebra homomorphisms in complex almost f-algebras, Comment Math.Univ.Carol. 43 (2002) 23-31 [15] A. C. Zaanen, Riesz Spaces II, North- Holland, Amsterdam, 1983 Istanbul University, Faculty of Science, Mathematics Department, Vezneciler- Istanbul, 34134, Turkey E-mail address: [email protected] Current address: Istanbul University, Faculty of Science, Mathematics Department, Vezneciler- Istanbul, 34134, Turkey E-mail address: [email protected]
1605.05724
1
1605
2016-05-18T08:00:03
Skew-symmetric operators and reflexivity
[ "math.FA", "math.OA" ]
In contrast to the subspaces of all $C$-symmetric operators, we show that the subspaces of all skew-C symmetric operators are reflexive and even hyperreflexive with the constant $\kappa(\C^s)\leqslant 3$.
math.FA
math
SKEW-SYMMETRIC OPERATORS AND REFLEXIVITY CHAFIQ BENHIDA*, KAMILA KLI´S-GARLICKA**, AND MAREK PTAK**,*** Abstract. In contrast to the subspaces of all C-symmetric operators, we show that the subspaces of all skew-C symmetric operators are reflexive and even hyperreflexive with the constant κ(C s) 6 3. 1. Introduction and Preliminaries Let H be a complex Hilbert space with an inner product h., .i and let B(H) be the Banach algebra of all bounded linear operators on H. Recall that the space of trace class operators τ c is predual to B(H) with the dual action hT, f i = tr(T f ), for T ∈ B(H) and f ∈ τ c. The trace norm in τ c will be denoted by k · k1. By Fk we denote the set of all operators of rank at most k. Often rank-one operators are written as x ⊗ y, for x, y ∈ H, and (x ⊗ y)z = hz, yix for z ∈ H. Moreover, tr(T (x ⊗ y)) = hT x, yi. Let S ⊂ B(H) be a closed subspace. Denote by S⊥ the preanihilator of S, i.e., S⊥ = {t ∈ τ c : tr(St) = 0 for all S ∈ S}. A weak∗ closed subspace S is k-reflexive iff rank-k operators are linearly dense in S⊥, i.e., S⊥ = [S⊥ ∩ Fk] (see [8]). k-hyperreflexivity introduced in [1, 6] is a stronger property than k- reflexivity, i.e., each k-hyperreflexive subspace is k-reflexive. A subspace S is called k-hyperreflexive if there is a constant c > 0 such that dist(T, S) ≤ c · sup{tr(T t) : t ∈ Fk ∩ S⊥, ktk1 ≤ 1}, (1) for all T ∈ B(H). Note that dist(T, S) is the infimum distance. The supremum on the right hand side of (1) will be denoted by αk(T, S). The smallest constant for which inequality (1) is satisfied is called the k-hyperreflexivity constant and is denoted κk(S). If k = 1, the letter k will be omitted. Recall that C is a conjugation on H if C : H −→ H is an antilinear, isometric involution, i.e., hCx, Cyi = hy, xi for all x, y ∈ H and C 2 = I. An operator T in B(H) is said to be C -- symmetric if CT C = T ∗. C -- symmetric operators have been intensively studied by many authors in the last decade (see [2], [3], [4], 2010 Mathematics Subject Classification. Primary 47A15, Secondary 47L05. Key words and phrases. skew -- C symmetry, C -- symmetry, reflexivity, hyperreflexivity. The first named author was partially supported by Labex CEMPI (ANR-11-LABX-0007- 01). The research of the second and the third author was financed by the Ministry of Science and Higher Education of the Republic of Poland. 1 2 CHAFIQ BENHIDA, KAMILA KLI´S-GARLICKA, AND MAREK PTAK [5]). It is a wide class of operators including Jordan blocs, truncated Toeplitz operators and Hankel operators. Recently, in [5], the authors considered the problem of reflexivity and hyper- reflexivity of the subspace C = {T ∈ B(H) : CT C = T ∗}. They have shown that C is transitive and 2-hyperreflexive. Recall that T ∈ B(H) is a skew -- C sym- metric iff CT C = −T ∗. In this paper, C s = {T ∈ B(H) : CT C = −T ∗} -- the subspace of all skew -- C symmetric operators will be investigated from the reflex- ivity and hyperreflexivity point of view. It follows directly from the definition that C and C s are weak∗ closed. We emphasize that the notion of skew symmetry is linked to many problems in physics and that any operator T ∈ B(H) can be written as a sum of a C -- symmetric operator and a skew -- C symmetric operator. Indeed, T = A + B, where A = 1 2 (T + CT ∗C) and B = 1 The aim of this paper is to show that C s is reflexive and even hyperreflexive. 2 (T − CT ∗C). Easy calculations show the following. 2. Preanihilator Lemma 2.1. Let C be a conjugation in a complex Hilbert space H and h, g ∈ H. Then (1) C(h ⊗ g)C = Ch ⊗ Cg, (2) h ⊗ g − Cg ⊗ Ch ∈ C s . In [2, Lemma 2] it was shown that C ∩ F1 = {α · h ⊗ Ch : h ∈ H, α ∈ C}. We will show that it is also a description of the rank-one operators in the pre- anihilator of C s. Proposition 2.2. Let C be a conjugation in a complex Hilbert space H. Then C s ⊥ ∩ F1 = C ∩ F1 = {α · h ⊗ Ch : h ∈ H, α ∈ C}. Proof. To prove "⊃" let us take T ∈ C s and h ⊗ Ch ∈ C ∩ F1. Then hT, h ⊗ Chi = hT h, Chi = hh, CT hi = hh, −T ∗Chi = = −hT h, Chi = −hT, h ⊗ Chi. Hence hT, h ⊗ Chi = 0 and h ⊗ Ch ∈ C s ⊥ ∩ F1. For the converse inclusion let us take a rank-one operator h ⊗ Cg ∈ C s ⊥. Since Cg ⊗ h − Ch ⊗ g ∈ C s, by Lemma 2.1 we have 0 = hCg ⊗ h − Ch ⊗ g, h ⊗ Cgi = h(Cg ⊗ h)h, Cgi − h(Ch ⊗ g)h, Cgi = = khk2 · kCgk2 − hh, gihCh, Cgi = khk2 · kgk2 − hh, gi2. SKEW-SYMMETRIC OPERATORS AND REFLEXIVITY 3 Hence hh, gi = khk kgk, i.e., there is equality in Cauchy-Schwartz inequality. (cid:3) Thus h, g are linearly dependent and the proof in finished. Lemma 2.3. Let C be a conjugation in a complex Hilbert space H. Then C s ⊥ ∩ F2 ⊃ {h ⊗ g + Cg ⊗ Ch : h, g ∈ H}. Example 2.4. Note that for different conjugations we obtain different sub- spaces. Let C1(x1, x2, x3) = (¯x3, ¯x2, ¯x1) be a conjugation on C3. Then 0 b a 0 −b c 0 −c −a C s 1 =     C1 =      : a, b, c ∈ C    : a, b, c ∈ C   . ∗ a b ∗ b c ∗ c a 1)⊥ are of the form α(x1, x2, x3) ⊗ (¯x3, ¯x2, ¯x1) Rank-one operators in C1 and in (C s for α ∈ C. If we now consider another conjugation C2(x1, x2, x3) = (¯x2, ¯x1, ¯x3) on C3, and then and C s 2 =     C2 =     0 a b 0 −a c −c −b 0 a ∗ b ∗ a c ∗ c b ,  : a, b, c ∈ C    : a, b, c ∈ C   . Similarly, rank-one operators in C2 and in (C s (¯x2, ¯x1, ¯x3). 2)⊥ are of the form α(x1, x2, x3) ⊗ Example 2.5. Let C be a conjugation in H. Consider C = (cid:18) 0 C C 0 (cid:19) the conjugation in H⊕ H (see [7]). An operator T ∈ B(H⊕ H) is skew- C symmetric, if and only if T = (cid:18) A D −CA∗C (cid:19), where A, B, D ∈ B(H) and B, D are skew- C symmetric. Moreover, rank-one operators in C s (Cg ⊕ Cf ) for f, g ∈ H and α ∈ C. ⊥ are of the form α(f ⊕ g) ⊗ B Example 2.6. Let us consider the classical Hardy space H 2 and let α be a nonconstant inner function. Define K 2 α = H 2 ⊖ αH 2 and Cαh(z) = αzh(z). Then Cα is a conjugation on K 2 α denote the compressions of the unilateral shift S and the backward shift S ∗ to K 2 α, respectively. Recall after [9] that the kernel functions in K 2 α for λ ∈ C are projections of appropriate kernel α. By Sα and S ∗ 4 CHAFIQ BENHIDA, KAMILA KLI´S-GARLICKA, AND MAREK PTAK functions kλ onto K 2 Sα and S ∗ A ∈ B(K 2 λ . Since α are Cα -- symmetric (see [2]), for a skew -- Cα symmetric operator α) we have λ = kλ − α(λ)αkλ. Denote by kα α, namely kα λ = Cαkα hASn αkα λ , (S ∗ α)mkα λ i = hCα(S ∗ − hSm α)mkα α Cαkα λ , CαASn αkα λ , A∗CαSn λ i = αkα λ i = −hASm α kα λ , (S ∗ α)nkα λ i, for all n, m ∈ N. Note that if n = m, then hASn α kα λ , (S ∗ α)nkα λ i = 0. (2) (3) In particular, we may consider the special case α = zk, k > 1. Then the equality (3) implies that a skew -- Cα symmetric operator A ∈ B(K 2 zk ) has the matrix representation in the canonical basis with 0 on the diagonal orthogonal to the main diagonal. Indeed, let A ∈ B(K 2 zk ) have the matrix (aij )i,j=0,...,k−1 with respect to the canonical basis. Note that Czk f = zk−1 ¯f , kzk 0 = zk−1. Hence for 0 ≤ n ≤ k − 1 we have 0 = 1, kzk 0 = hASn α1, (S ∗ α)nzk−1i = hAzn, zk−n−1i = an,k−n−1. Moreover, from the equality (2) we can obtain that hAzn, zk−m−1i = −hAzm, zk−n−1i, which implies that an,k−m−1 = −am,k−n−1 for 0 ≤ m, n ≤ k − 1. 3. Reflexivity The following theorem can be obtained as a corollary of Theorem 1. However, we think that the proof presented here is also interesting. Theorem 3.1. Let C be a conjugation in a complex Hilbert space H. The subspace C s of all skew -- C symmetric operators on H is reflexive. Proof. By Proposition 2.2 it is necessary to show that if hT, h⊗Chi = hT h, Chi = 0 for any h ∈ H, then CT C = −T ∗. Recall after [2, Lemma 1] that H can be decomposed into its real and imagi- nary parts H = HR + i HI. Recall also that we can write h = hR + ihI ∈ H with 2 (I + C)h ∈ HR and hI = 1 hR = 1 2i (I − C)h ∈ HI . Then ChR = hR, ChI = hI and Ch = C(hR + ihI ) = hR − ihI . Z (cid:21), where Let T ∈ B(H). The operator T can be represented as (cid:20) W X Y W : HR → HR, Z : HI → HI , X : HI → HR, Y : HR → HI and W, X, Y, Z are real linear. The condition CT C = −T ∗ is equivalent to the following: W = −W ∗, Z = −Z ∗, Y = X ∗. SKEW-SYMMETRIC OPERATORS AND REFLEXIVITY 5 On the other hand, the condition hT h, Chi = 0 for any h = hR + ihI is equivalent to hW hR, hRi + hXhI, hRi − hY hR, hIi − hZhI , hI i = 0 (4) for any hR ∈ HR, hI ∈ HI . In particular, hW hR, hRi = 0 for any hR ∈ HR. Let h ′ R, h ′′ R ∈ HR. Then hW h ′ R, h ′ Ri = 0, hW h ′′ R, h ′′ Ri = 0 and 0 = hW (h ′ R + h ′′ R), h ′ R + h ′′ Ri = hW h ′ R, h ′′ Ri + hW h ′′ R, h ′ Ri. Hence hW h ′ R, h ′′ Ri = hh ′ R, −W h ′′ Ri and finally W ∗ = −W . Since, by (4), in particular hZhI , hI i = 0 for any hI ∈ HI we can also get Z ∗ = −Z. Because hW hR, hRi = 0 = hZhI , hI i for any hR ∈ HR, hI ∈ HI , hence by (4) we get hXhI , hRi − hY hR, hI i = 0. Thus Y = X ∗ and the proof is finished. (cid:3) Recall that a single operator T ∈ B(H) is called reflexive if the weakly closed algebra generated by T and the identity is reflexive. In [7] authors character- ized normal skew symmetric operators and by [10] we know that every normal operator is reflexive. Hence one may wonder, if all skew -- C symmetric operators are reflexive. The following simple example shows that it is not true. Example 3.2. Consider the space C2 and a conjugation C(x, y) = (¯x, ¯y). Note that operator T = (cid:18) 0 A(T ) generated by T consists of operators of the form (cid:18) a A(T )⊥ = (cid:26)(cid:18) t −1 0 (cid:19) is skew-C symmetric. The weakly closed algebra −b a (cid:19). Hence s −t (cid:19) : t, s ∈ C(cid:27). It is easy to see, that A(T )⊥ ∩ F1 = {0}, b 1 s which implies that T is not reflexive. 4. Hyperreflexivity Theorem 4.1. Let C be a conjugation in a complex Hilbert space H. Then the subspace C s of all skew -- C symmetric operators is hyperreflexive with the constant κ(C s) 6 3 and 2-hyperreflexive with κ2(C s) = 1. Proof. Let A ∈ B(H). Firstly, similarly as in the proof of Theorem 4.2 [5] it can be shown that A − CA∗C ∈ C s. It is also shown there that (CAC)∗ = CA∗C. 6 CHAFIQ BENHIDA, KAMILA KLI´S-GARLICKA, AND MAREK PTAK Hence we have 2 (A − CA∗C)k = 1 2 kA + CA∗Ck d(A, C s) 6 kA − 1 = 1 = 1 = 1 = 1 6 1 2 sup{hh, (A∗ + CAC)gi : khk, kgk 6 1} 2 kA∗ + CACk = 1 2 sup{hA, h ⊗ g + Cg ⊗ Chi : khk, kgk 6 1} 2 sup{hA, h ⊗ Cg + g ⊗ Chi : khk, kgk 6 1} = α2(A, C s) 2 sup{hA, (h + g) ⊗ C(h + g) − h ⊗ Ch − g ⊗ Cgi : khk, kgk 6 1} 2 sup{hA, h ⊗ Chi : khk 6 1} + 1 2 sup{hA, g ⊗ Cgi : kgk 6 1}+ 2 sup{4hA, 1 2 (h + g))i : khk, kgk 6 1} + 1 6 3 α(A, C s). 2 (h + g) ⊗ C( 1 We have used the characterization given in Proposition 2.2. (cid:3) References [1] W. T. Arveson: Interpolation problems in nest algebras, J. Funct. Anal. 20 (1975), 208 -- 233. [2] S. R. Garcia and M. Putinar: Complex symmetric operators and applications, Trans. Amer. Math. Soc. 358 (2006), 1285 -- 1315. [3] : Complex symmetric operators and applications II, Trans. Amer. Math. Soc. 359 (2007), 3913 -- 3931. [4] S. R. Garcia and W. R. Wogen: Some new classes of complex symmetric operators, Trans. Amer. Math. Soc. 362 (2010), 6065 -- 6077. [5] K. Kli´s-Garlicka and M. Ptak: C-symmetric operators and reflexivity, Operators and Matrices 9 no. 1 (2015), 225 -- 232. [6] K. Kli´s, M. Ptak: k-hyperreflexive subspaces, Houston J. Math. 32 (2006), 299 -- 313. [7] C. G. Li and S. Zhu: Skew symmetric normal operators, Proc. Amer. Math. Soc. 141 no. 8 (2013), 2755 -- 2762. [8] A.I. Loginov and V.S. Shul'man: Hereditary and intermediate reflexivity of W*-algebras, Izv. Akad. Nauk. SSSR, 39 (1975), 1260 -- 1273; Math. USSR-Izv. 9 (1975), 1189 -- 1201 [9] D. Sarason: Algebraic properties of truncated Toeplitz operators, Operators and Matrices 1 no. 4 (2007), 491 -- 526. [10] : Invariant subspaces and unstarred operator algebras, Pacific J. Math. 17 no. 3 (1966), 511 -- 517. [11] S. Zhu: Skew symmetric weighted shifts, Banach J. Math. Anal. 9 no. 1 (2015), 253 -- 272. [12] : Approximate unitary equivalence to skew symmetric operators, Complex Anal. Oper. Theory 8 no. 7 (2014), 1565 -- 1580. * Universite Lille 1 Laboratoire Paul Painlev´e 59655 Villeneuve d'Ascq France E-mail address: [email protected] SKEW-SYMMETRIC OPERATORS AND REFLEXIVITY 7 ** Department of Applied Mathematics University of Agriculture Balicka 253c 30-198 Krakow Poland E-mail address: [email protected] *** Institut of Mathematics Pedagogical University ul. Podchora¸ zych 2 30-084 Krak´ow Poland E-mail address: [email protected]
1208.0979
1
1208
2012-08-05T04:23:13
A New Fixed Point Theorem for Non-expansive Mappings and Its Application
[ "math.FA" ]
We use $KKM$ theorem to prove the existence of a new fixed point theorem for non-expansive mapping:Let M be a bounded closed convex subset of Hilbert space H, and $A:M\rightarrow M$ be a non-expansive mapping, then exists a fixed point of A in M, we also apply this Theorem to study the solution for an integral equation,we can weak some conditions comparing with Banach's contraction principe.
math.FA
math
A New Fixed Point Theorem for Non-expansive Mappings and Its Application Department of Mathematics, Sichuan University, Chengdu 610064,P.R.China Chunyan Yang e-mail: [email protected] Feb,12,2012 2 1 0 2 g u A 5 ] . A F h t a m [ 1 v 9 7 9 0 . 8 0 2 1 : v i X r a Abstract. We use KKM theorem to prove the existence of a new fixed point theorem for non-expansive mapping:Let M be a bounded closed convex subset of Hilbert space H, and A : M → M be a non-expansive mapping, then exists a fixed point of A in M, we also apply this Theorem to study the solution for an integral equation,we can weak some conditions comparing with Banach's contraction mapping principle. Key words: Bounded closed convex subset, non-expansive mapping, fixed point,integral equation 2000 AMS Subject Classification: 47H10,47J25,47J05 1 Introduction and Main Results It is well-known that Banach, Schauder,Brouwer presented three kinds of fixed point theorems in 1910's-1930's. Banach′s fixed point theorem is also called contraction mapping principle , and Schauder′s fixed point theorem is a generalization of Brouwer′s from finite dimension into infinite dimension. Specifically, for a compact operator A, if its domain is a bounded non-empty closed convex subset, then there is at least one fixed point u : A · u = u. All the above three fixed point theorems need good operators, for example, the compactness for the operator can make the unit closed sphere in infinite dimensional space be a compact set. Here, a nature question is whether the fixed point theorem of other good operator or mapping could be established. A problem is whether there is a fixed point theorem for A when k = 1 (where A is called the non-expansive mapping). For solving this problem, the main tool we will use is KKM theorem.It's well known that the Polish mathematician Knaster, Kuratowski, M azurkiewicz got KKM theorem in 1929, they also applied it to the proof of Brouwer fixed point theorem. 1 Theorem1.1.( Banach fixed point theorem([1])): Let (X, d) be a complete metric space, and A : X → X be a contraction mapping, that is , for any x,y ∈ X, there holds d(Ax, Ay) ≤ kd(x, y), (0 ≤ k < 1). Then there is a unique fixed point x∗ for A, i.e. Ax∗ = x∗. Banach's fixed point theorem can be applied to many fields in mathematics,especially to integral equation: u(x) = λZ b a F (x, y, u(y))dy + f (x), a ≤ x ≤ b (1.1) Theorem1.2.([2]) Assume (a) f : [a, b] → R is continuous (b) F (x, y, u(y)) = K(x, y)u(y),and K : [a, b] × [a, b] → R is continuous. Let Γ = maxa≤x,y≤b K(x, y),and (c) real number λ be given such that (b − a) · λ · Γ < 1. Then the problem (1.1) has a unique solution u ∈ X = C[a, b]. In this paper,we get: Theorem1.3. Let M be a bounded closed convex subset of Hilbert space H, and A : M → M be a non-expansive mapping, then exists a fixed point of A in M. Notice that our fixed point theorem is in Hilbert Space , since any linear continuous functional in Hilbert space can be represented by the inner product, which is called Riesz representation theorem, while in the Banach space no such theorem. We apply the above theorem to linear integral equation,we obtain Theorem1.4. For the integral equation(1.1) with F (x, y, u(y)) = K(x, y)u(y), assume (1)The function f : [a, b] → R is in L2[a, b]. (2)K(x, y) ∈ L2([a, b] × [a, b]),K(x, ·) ∈ L2[a, b],for any given x ∈ [a, b] (3)Let the real number λ be given such that, λR b a R b a K(x, y)2dxdy ≤ 1, if f ≡ 0; and, λZ b a Z b a K(x, y)2dxdy < 1, if f 6= 0. Then (1.1) has at least one solution u ∈ L2[a, b]. 2 Related Definitions and Well-known Theorems Definition 2.1. (KKM mapping([3])): The mapping G : X → 2Rn , where X is a non-empty subset of the vector space Rn , is called KM M mapping. if ∀{x0, . . . , xn} ⊂ X, we have Co({x0, . . . , xn}) ⊂ ∪n i=1G(xi), where Co({x0, . . . , xn}) is the closed convex hull of x0, . . . , xn. 2 Theorem 2.1. (KKM Theorem ([4])) Let ∆n be n dimensional simplex which has vertices {e0, . . . , en} and is in Rn+1. Let M0, . . . , Mn be n dimensional closed subsets in Rn+1, which satisfy ∀{ei0, . . . , eik} ⊂ {e0, . . . , en}, Co({ei0, . . . , eik}) ⊂ ∪k j=0Mij. Then ∩Mi is not empty. Theorem 2.2. (F KKM ([5])): Let F n be Hausdorff linear topological space, X be a non-empty subset of F n. Let G : X −→ 2F n be a KKM mapping, and G(x) is weakly compact for any x ∈ X . Then ∩x∈X G(x) is non-empty. Definition 2.2.( Semi-continuous([6])): Let X be a Banach space, X ∗ be its dual space. Let T : X → X ∗. If ∀y ∈ X, ∀tn ≥ 0and x0 + tn · y ∈ X, we have limx0+tn·y→x0T (x0 + tn · y) → T (x0), we call T is semi − continuous at point x0. Fur- thermore if T is semi − continuous at every point x in X, T is called semi − continuous on X. Definition 2.3.( Monotonicity([7])): Let X be a Banach space, X ∗ be its dual space. We call T :→ 2X ∗ monotonous , if < u−v, x−y >≥ 0 ,for ∀x, y ∈ X, u ∈ T (x), v ∈ T (y). 3 Some Lemmas for the Proof of Our Theorem 1.3. We require the following Lemmas for proving the fixed point theorem of the non- expansive mapping using the KKM theorem . Lemma 3.1: Let M be a bounded closed convex subset of Hilbert space H, L : M → H be monotonous semi-continuous mapping. If x0 ∈ M, then if and only if < L(x0), y − x0 >≥ 0, ∀y ∈ M, < L(y), y − x0 >≥ 0, ∀y ∈ M. Proof Let x0 ∈ M such that : < L(x0), y − x0 >≥ 0, ∀y ∈ M. Since the monotonicity of L, we have < L(y) − L(x0), y − x0 >≥ 0 ⇔< L(y), y − x0 > − < L(x0), y − x0 >≥ 0, ∀y ∈ M. Thus , < L(y), y − x0 >≥< L(x0), y − x0 >≥ 0, ∀y ∈ M. On the other hand,if x0 ∈ M then < L(y), y − x0 >≥ 0, ∀y ∈ M. We let y = h · µ + (1 − h)x0, ∀h ∈ (0, 1], ∀µ ∈ M, 3 since M is convex, then y ∈ M. Thus, < L(y), y − x0 >=< L(x0 + h · (µ − x0)), h · (µ − x0) >≥ 0(h > 0). We let h → 0, then x0 + h · (µ − x0) → x0, since L is semi-continuous , so < L(x0), (µ − x0) >≥ 0, ∀µ ∈ M. Lemma 3.2([8]): Let M be a bounded closed convex subset of Hilbert space H, andL : M → H, P : M → 2M , and P (y) = {x ∈ M, < L(x), x − y >≤ 0}, ∀y ∈ M, then P is a KKM mapping. Proof: If P is not a KKM mapping. By the definition of KKM mapping, ∃{x0 . . . xn} = w, w ⊂ M, let ¯x = P hi · xi ∈ Co(w), (P hi = 1, hi ≥ 0), then ¯x 6∈ ∪n i=0P (xi), that is to say, < L(¯x), ¯x − y >> 0. As the arbitrariness of y, let y = P hi · xi, we have< L(¯x), ¯x − xi >> 0, then < L(¯x), hi · (¯x − xi) >> 0(∃hi > 0), thus, < L(¯x),X hi · (¯x − xi) >=< L(¯x),X hi · ¯x −X hi · xi >=< L(¯x), ¯x − y >= 0. This is a contradiction with < L(¯x), ¯x − y >> 0, ∀y ∈ M. Hence P is a KKM mapping. Lemma 3.3: Let M be a bounded closed convex subset of Hilbert space H, L : M → H be a semi-continuous monotonous mapping, then ∃x0 ∈ M s.t. < L(x0), x0 − y >≤ 0, ∀y ∈ M Proof: Let P : M → 2M and P (y) = {x ∈ M, < L(x), x−y >≤ 0}, ∀y ∈ M. According to Lemma 3.2,P is a KKM mapping. Mean-while, let mappingJ : M → 2M and J(y) = {x ∈ M, < L(y), x − y >≤ 0}, ∀y ∈ M. By the monotonicity of the operator L and Lemma 3.1, P (y) ⊂ J(y), ∀y ∈ M. Since P is a KKM mapping, Co({x0 . . . xn}) ⊂ ∪n KKM mapping. i=0P (xi) ⊂ ∪n i=0J(xi). hence J is Furthermore, it can be easy to prove that J(y) is weakly compact. In fact, Since M is a bounded closed convex subset, it can be convinced that M is weakly compact. But J(y) ⊂ M, thus we only need to prove J(y) is weakly closed. In order to do that let yn ∈ J(y) ⊂ M, s.t. yn ⇀ y0, so, yn − y ⇀ y0 − y, ∀y ∈ M. According to the semi-continuity of L,< L(y), yn − y >→< L(y), y0 − y >. By yn ∈ J(y),we have < L(y), yn − y >≤ 0, so < L(y), y0 − y >≤ 0. Hence y0 ∈ J(y), that is to say, J(y) is weakly closed. From the above proof, J is a KKM mapping, and ∀y, J(y) is weakly compact. According to F KKM theorem, ∩y∈M J(y) 6= ∅. According to Lemma 3.1,∩y∈M J(y) = ∩y∈M P (y) 6= ∅,thus ∃x0 ∈ ∩y∈M P (y), s.t., < L(x0), x0 − y >≤ 0, ∀y ∈ M. 4 4 The Proof of Fixed Point Theorem1.3 for the Non- Expansive Mapping Proof: Let L(x) = x − A(x), ∀x ∈ M, then L : M → H.By the non-expansive property for the operator A ,we will prove L= I - A is a monotonous operator: < L(x) − L(y), x − y >> 0, ∀x, y ∈ M. In fact, < L(x) − L(y), x − y > = < x − y − (Ax − Ay), x − y > = kx − yk2− < Ax − Ay, x − y > ≥ kx − yk2 − kAx − Ayk2 · kx − yk2 ≥ kx − yk2 − kx − yk2 = 0. Since A is non-expansive mapping :∀x, y ∈ M ,kAx − Ayk ≤ kx − yk,it can be derived that L is semi-continuous mapping. According to Lemma 3.3, ∃x0 ∈ M ,s.t. < L(x0), x0 − x >≤ 0, ∀x ∈ M Specially we let x = A(x0) ∈ M and substitute it into (4.1) ,we get: < L(x0), x0 − A(x0) >≤ 0 (4.1) (4.2) Since L(x0) = x0 − A(x0), so from (4.2) we have kx0 − A(x0)k ≤ 0. It's clearly that norm kx0 − A(x0)k ≥ 0, hence kx0 − A(x0)k = 0, that is to say, A(x0) = x0. 5 The Proof of Theorem1.4 Proof: Let (Au)(x) = λR b a F (x, y, u(y))dy + f (x), a a kAu − Avk2 = Z b ≤ λ2Z b = [λ2Z b λZ b [Z b a Z b a K(x, y)((u(y) − v(y))dy2dx K2dyZ b a u(y) − v(y)2dy]dx K(x, y)2dxdy] · ku − vk2 L2. a a If [λ2R b a R b a K(x, y)2dxdy] ≤ 1, then A is non-expansive mapping. 5 Furthermore,we restrict A on a bounded closed convex subset M of L2 such that AM ⊆ M.Given r > o,letM = {u ∈ L2kukL2 ≤ r}, a a a kAu(x)kL2 ≤ kλZ b ≤ λ[Z b Z b ≤ λ{Z b [Z b ≤ λ{Z b [Z b = λ[Z b a Z b ≤ λ[Z b a Z b a a a a a a F (x, y, u(y))dykL2 + kf kL2 F dy2dx]1/2 + kf kL2 K(x, y)u(y)dy]2dx}1/2 + kf kL2 K2dyZ b K(x, y)2dxdy]1/2 · kukL2 + kf kL2 u(y)2dy]dx}1/2 a K(x, y)2dxdy]1/2 · r + kf kL2 ≤ r. (1).If f ≡ 0,and λ[Z b a Z b a Then we can choose any given r > 0. (2).If f 6= 0,and λ[Z b a Z b a K(x, y)2dxdy]1/2 ≤ 1. K(x, y)2dxdy]1/2 < 1. Then choose r such that, r ≥ kf kL2 1 − λ[R b a R b a K(x, y)2dxdy]1/2 . Acknowledgements. I would like to thank Professor Zhang Shiqing for his lectures on Functional Analysis ,the paper was written when I attended his lectures, I would like to thank Professor Shiqing Zhang for his many helpful discussions,encouragements and suggestions and corrections,this paper is supported partially by NSF of China. References [1] E.Zeidler,Applied functional analysis:main principles and their applications ,Springer,1995. [2] E.Zeidler,Applied functional analysis:application to mathematical physics,Springer,1995,P.22-23. [3] K.C.Border,Fixed point theorems with applications to ecomomics and game the- ory,Cambridge University Press,2009,P.26. 6 [4] B.Knaster,C.Kuratowski and S.Muzurkiewicz.,Ein Beweis des Fixepunktsatzes fur n-dimensional simplexe,Fund.Math.14(1929),P.132-137. [5] K.Fan,A of ,Ann.Math.142(1961),P.303-310. generalization Tychonoff's fixed point theorm [6] W.Hang, X.Cheng,An introduction to variational inequalities :elementary the- ory,numerical analysis and applications,Academic Press,2003,P.92. [7] B.Meng, X.Cao,A course in opreater and game theory,The MIT Press,1998,P.36. [8] S.Itoh,W.Takahashi,K.Yanagi,Varicational problems,J.Math.Soc.Japan,1978,P.17-55. Inequalities and complementarity 7
1307.7958
1
1307
2013-07-30T13:03:44
Banach spaces with no proximinal subspaces of codimension 2
[ "math.FA" ]
The classical theorem of Bishop-Phelps asserts that, for a Banach space X, the norm-achieving functionals in X* are dense in X*. Bela Bollobas's extension of the theorem gives a quantitative description of just how dense the norm-achieving functionals have to be: if (x,f) is in X x X* with ||x||=||f||=1 and |1-f(x)|< h^2/4 then there are (x',f') in X x X* with ||x'||= ||f'||=1, ||x-x'||, ||f-f'||< h and f'(x')=1. This means that there are always "proximinal" hyperplanes H in X (a nonempty subset E of a metric space is said to be "proximinal" if, for x not in E, the distance d(x,E) is always achieved - there is always an e in E with d(x,E)=d(x,e)); for if H= ker f (f in X*) then it is easy to see that H is proximinal if and only if f is norm-achieving. Indeed the set of proximinal hyperplanes H is, in the appropriate sense, dense in the set of all closed hyperplanes H in X. Quite a long time ago [Problem 2.1 in his monograph "The Theory of Best approximation and Functional Analysis" Regional Conference series in Applied Mathematics, SIAM, 1974], Ivan Singer asked if this result generalized to closed subspaces of finite codimension - if every Banach space has a proximinal subspace of codimension 2, for example. In this paper I show that there is a Banach space X such that X has no proximinal subspace of finite codimension n>1. So we have a converse to Bishop-Phelps-Bollobas: a dense set of proximinal hyperplanes can always be found, but proximinal subspaces of larger, finite codimension need not be.
math.FA
math
Banach spaces with no proximinal subspaces of codimension 2 C.J.Read Department of Pure Mathematics, University of Leeds, LS2 9JT, UK [email protected]; http://solocavediver.com/maths Abstract The classical theorem of Bishop-Phelps asserts that, for a Banach space X, the norm-achieving functionals in X ∗ are dense in X ∗. B´ela Bollob´as's extension of the theorem gives a quantitative description of just how dense the norm-achieving functionals have to be: if (x, ϕ) ∈ X × X ∗ with kxk = kϕk = 1 and 1 − ϕ(x) < ε2/4 then there are (x′, ϕ′) ∈ X × X ∗ with kx′k = kϕ′k = 1, kx − x′k ∨ kϕ − ϕ′k < ε and ϕ′(x′) = 1. This means that there are always "proximinal" hyperplanes H ⊂ X (a nonempty subset E of a metric space is said to be "proximinal" if, for x /∈ E, the distance d(x, E) is always achieved - there is always an e ∈ E with d(x, E) = d(x, e)); for if H = ker ϕ (ϕ ∈ X ∗) then it is easy to see that H is proximinal if and only if ϕ is norm-achieving. Indeed the set of proximinal hyperplanes H is, in the appropriate sense, dense in the set of all closed hyperplanes H ⊂ X. Quite a long time ago [Problem 2.1 in his monograph "The Theory of Best approximation and Functional Analysis" Regional Conference series in Applied Mathematics, SIAM, 1974], Ivan Singer asked if this result generalized to closed subspaces of finite codimension - if every Banach space has a proximinal subspace of codimension 2, for example. In this paper I will show that there is a Banach space X such that X has no proximinal subspace of finite codimension n ≥ 2. So we have a converse to Bishop-Phelps-Bollob´as: a dense set of proximinal hyperplanes can always be found, but proximinal subspaces of larger, finite codimension need not be. 1 1 Introduction. I'm grateful to David Blecher for awakening me to the joys of prox- iminality in the context of operator algebras (norm-closed subalgebras of B(H)), and to Gilles Godefroy for alerting me to this particular problem. The original Bishop-Phelps theorem is [1], and Bollob´as' improved version of the theorem is [3]. The place where the problem solved in this paper was originally posed is in Ivan Singer [6]. Gilles Godefroy's exhaustive survey article on isometric preduals in Banach spaces, which discusses this problem among many others, is [5]. Our work with David Blecher involving proximinality of ideals in operator algebras is [2]. This is a successful attempt to generalize, to a noncommutative setting, the classical Glicksberg peak set theorem in uniform algebras (Theorem 12.7 in Gamelin [4]). All the Banach spaces in this paper are over the real field. At risk of stating the obvious, a proximinal subset is necessarily closed; so we lose no generality later on by assuming that a (hypothetical) proximinal subspace of finite codimension is the intersection of the kernels of finitely many continuous linear functionals. Let c00(Q) denote the terminating sequences with rational coeffi- cients (a much-loved countable set), and let (uk)∞ k=1 be a sequence of elements of c00(Q) which lists every element infinitely many times. For x ∈ c00(Q), write u−1{x} for the infinite set {k ∈ N : uk = x}. Let (ak)∞ k=1 be a strictly increasing sequence of positive integers. We impose a growth condition: if uk 6= 0, we demand that ak > max supp uk, and ak ≥ kukk1, (1) where supp u denotes the (finite) support of u ∈ c00(Q), and kuk1 denotes the l1 norm. For E ⊂ N we write AE for the set {ak : k ∈ E}; for x ∈ c00(Q) we write Ax for Au−1{x}. Ax is an infinite set, and in view of (1), for each x ∈ c00(Q) \ {0} we have min Ax > max supp x, min Ax ≥ kxk1. (2) Given sequences (uk), (ak) as described above, we define a new norm k·k on c0 as follows: ∞ kxk = kxk0 + 2−a2 k hx, uk − eak i. (3) Xk=1 Here kxk0 = supn xn is the usual norm on c0; (ej) are the unit vectors; and the duality hx, uk − eak i is the hc0, l1i duality. Now in view of (1), k kuk − eak k1 ≤ k=12−a2 k=12−a2 we have kuk − eak k1 = 1+kukk1 ≤ 1+ak, soP∞ P∞ n=1(1 + n) · 2−n2 kxk0 ≤ kxk ≤ 3kxk0 k (1 + ak) ≤P∞ < 2. Accordingly, we have (4) for all x ∈ c0. For our main theorem in this paper, we shall show: Theorem 1.1 The Banach space (c0, k·k) has no proximinal subspace H of finite codimension n ≥ 2. 2 2 Gateaux derivatives Recall that if X is a real vectorspace, u, x ∈ X and f : X → R, then the Gateaux derivative (of f , at x, in direction u) is defined as df (x; u) = lim h→0 f (x + hu) − f (x) h , (5) when that limit exists. We will make use of the one-sided forms of this derivative: f (x + hu) − f (x) df+(x; u) = lim h→0+ h and df−(x; u) = lim h→0− f (x + hu) − f (x) h , . (6) (7) Obviously df−(x; u) = −df+(x; −u) for all f, x and u such that either derivative exists. Of particular interest to us is when X = c0 and f (x) = kxk as defined in (3) (the "usual" norm on c0 will always be referred to as k·k0 in this paper). The derivatives d±f (x; u) for this function f will be written d±kx; uk. Now it is a fact that the derivative d±kx; uk exists everywhere. To see this, let us prove some small lemmas. Lemma 2.1 If kxk0 denotes the c0-norm, the derivative d+kx; uk0 exists at all points x, u ∈ c0. In fact, if x = 0 then the derivative is kuk0; whereas if x 6= 0, we may write E+ = {n ∈ N : xn = kxk0, unxn > 0} and E− = {n ∈ N : xn = kxk0, unxn ≤ 0}, and we have d+kx; uk0 =(max{un : n ∈ E+}, − min{un : n ∈ E−}, if E+ 6= ∅; if E+ = ∅, E− 6= ∅. (8) Proof. This is an easy calculation which we omit (note that E+ and E− cannot both be empty!). Lemma 2.2 Let X be a Banach space and ϕ ∈ X ∗. Then the Gateaux derivative of f (x) = ϕ(x) exists at all points (x; u) ∈ X × X. We have d+f (x; u) = f (u)σ(ϕ(u)ϕ(x)), where the sign σ(t) =(+1, −1, if t ≥ 0; if t < 0. (9) Proof. This is an even simpler calculation, which we also omit. Definition 2.3 For a real normed space X and a function f : X → R, define the Lipschitz constant Lip1f = sup{ f (x) − f (y) kx − yk : x, y ∈ X, x 6= y}. (10) 3 d+fn exists everywhere also. Lemma 2.4 Let X be a real normed space, and (fn)∞ n=0 a sequence of functions from X to R, such that d+fn(x; u) exists at each (x; u) ∈ n=0fn(0) converges. Then n=0 n=0Lip1fn < ∞, and P∞ n=0fn exists everywhere on X, and d+f = P∞ n=0fn(x) converges becauseP∞ X × X. Suppose P∞ the function f = P∞ Proof. The sumP∞ and fn(x)−fn(0) ≤ kxk·Lip1fn soP∞ Since d+fn(x; u) ≤ kuk·Lip1fn, we find that the sumP∞ ε > 0, we can choose N so large that kuk ·P∞ n=0fn(0) converges, n=0fn(x)−fn(0) converges also. n=0d+fn(x; u) converges; we claim the sum is d+f (x; u). For given x, u 6= 0, and n=N +1Lip1fn < ε/3, so ∞ d+fn(x; u) < ε/3 Xn=N +1 (11) (12) and for every h > 0, ∞ (fn(x + hu) − fn(x))/h ≤ kuk · Xn=N +1 ∞ Lip1fn < ε/3 Xn=N +1 n=1fn(x))/h → n=1d+fn(x; u), so we can choose δ > 0 such that whenever 0 < h < n=1fn(x + hu) −PN also. As h → 0+, we know (PN PN δ, we have N ( Xn=1 N N fn(x + hu) − fn(x))/h − Xn=1 d+fn(x; u) < ε/3. Xn=1 (13) Adding up (11), (12) and (13), we find that whenever 0 < h < δ, we n=1d+fn(x; u) < ε. This completes (cid:3) have (f (x + hu) − f (x))/h −P∞ the proof. Corollary 2.5 The new norm k·k on c0 has a one-sided derivative d+kx; uk everywhere. Furthermore, d+kx; uk = d+kx; uk0 + ∞ Xk=1 2−a2 k σkhu, uk − eak i, (14) where σk = σk(x; u) = σ(hu, uk − eak ihx, uk − eak i), and the function σ is as in (9). Proof. If we write f0(x) = kxk0 and fk(x) = 2−a2 k hx, uk − eak i, then the Lipschitz constants for fk are 1 (if k = 0) or 2−a2 k kuk − eak k1 ≤ (1 + ak) · 2−a2 k=0Lip1fk < ∞, and the derivatives d+fk are given by Lemma 2.1 and Lemma 2.2. We have n=0d+fn(x; u) by Lemma 2.4. (cid:3) The key link between Gateaux derivatives and proximinality is as k for k > 0. Accordingly P∞ k=0fk(x) so d+kx; uk = P∞ This sum works out to expression (14). kxk = P∞ follows: 4 Lemma 2.6 Suppose (X, k·k) is a Banach space, H ⊂ X a subspace, and suppose that for some x ∈ X \ H, and v ∈ H, the Gateaux deriva- tives d±kx; vk both exist, are nonzero, and have the same sign. Then kxk 6= inf{kyk : y ∈ x + H}. x is not a closest point to zero in the coset x + H. Proof. We may consider y = x + hv for small nonzero h ∈ R. De- pending on the sign of h, the norm kyk is roughly kxk + h · d±kx; vk. But the signs of d±kx; vk are the same, so if h is chosen correctly, we get kyk < kxk. (cid:3) Corollary 2.7 Suppose H ⊂ X as in Lemma 2.6, and there is an x ∈ X \ H such that for every z ∈ H, there is a v ∈ H such that the Gateaux derivatives d±kx + z; vk exist, are nonzero, and have the same sign. Then H is not proximinal in X. Proof. For in this case, there is no element x + z ∈ x + H which achieves the minimum distance from that coset to zero. Equivalently, there is no element z ∈ H which achieves the minimum distance from H to −x. H is not proximinal. (cid:3) 3 Approximate linearity of d± It is a feature of the Gateaux derivative df (x; u) that it does not have to be linear in u. This is of course also true of the single-sided derivatives df±. So, in this section we develope a result asserting "approximate linearity" of d±kx; vk for x, v ∈ c0. Definition 3.1 Let f : c0 → R be such that d+f (x; v) exists for all x, v ∈ c0. Let x ∈ c0, and let γ ∈ l1 be such that the support E = {i : hei, γi 6= 0} is infinite. We shall say d+f (x) is approximately linear on E (and approximately equal to γ) if there is an "error sequence" (εi)i∈E with εi > 0, εi → 0, such that for all v ∈ c0 with supp v ⊂ E, we have d+f (x; v) − hv, γi ≤Xi∈E εiviγi. (15) Note that if v is chosen so that hv, γi > P εiviγi, then (15) im- plies that d+f (x; v) is nonzero, and has the same sign as d−f (x; v) = −d+f (x; −v). Lemma 3.2 Let x ∈ c0 be given, and z1, . . . , zm ∈ c00 such that hx, zj i 6= 0 for any j = 1, . . . , m. Let Azi = Au−1{zi} as in §1, and let A = ∪m i=1Azi . Then d+kxk is approximately linear on a cofinite subset i=1γie∗ i , A0 ⊂ A, the derivative being approximately equal to γ = P∞ where γi =(−2−a2 0, k σ(hx, zj i) if i = ak ∈ A0 ∩ Azj otherwise. (16) 5 The error sequence (εi)i∈A0 can be taken to be εi = 2a2 k · ∞ Xl=k+1 2−a2 l , i = ak ∈ A0. (17) Proof. Let v be any vector supported on A. The error δ = d+kx; vk− hv, γi is given by (14) and (16); we have 2−a2 k σkhv, uk − eak i δ = d+kx; vk0 + Xk∈N\∪m Xj=1 Xk∈u−1{zj } 2−a2 + m j=1u−1{zj } k σkhv, uk − eak i + vak σ(hx, zj i) (18) where σk = σ(hv, uk − eak ihx, uk − eak i). Now by Lemma 2.1, the derivative d+kx; vk0 is zero unless vi 6= 0 for some i ∈ E = {n : xn = kxk0}. This set E is finite, and v will be supported on A0; so if we choose our cofinite set A0 ⊂ A so that A0 ∩ E = ∅, we have d+kx; vk0 = 0. If we also ensure that A0 ∩ supp zj = ∅ for each j = 1, . . . , m, we find that when v is supported on A0, and k ∈ u−1{zj}, we have hv, uk − eak i = hv, zj − eak i = −hv, eak i = −vak . (19) So if we choose A0 so that A0 ∩ (E∪m (18) simplifies somewhat to j=1supp zj) = ∅, the expression 2−a2 k σkhv, uk − eak i j=1u−1{zj } δ = Xk∈N\∪m Xj=1 Xk∈u−1{zj } + m (σkvak + vak σ(hx, zj i)); (20) and σk itself simplifies to σk = σ(−hx, zj − eak i · vak ). Now Fj = {k ∈ N : hx, eak i ≥ hx, zj i} is a finite set; we may thus also assume that A0 does not meet any Fj. In that case, hx, zj − eak i is nonzero and has the same sign as hx, zj i, so σk = −vak σ(hx, zj i). So the second term in (20) disappears, and we have δ = Xk∈N\∪m j=1u−1{zj } 2−a2 k σkhv, uk − eak i. (21) Even better, v is supported on A = a · ∪m hv, eak i are zero in (21), and we have j=1u−1{zj}, so all terms 2−a2 l σlhv, uli; 2−a2 l vi · hei, uli. (22) j=1u−1{zj } δ = Xl∈N\∪m δ ≤ Xi∈A0 Xl∈N\∪m j=1u−1{zj } 6 Now in every case when i ∈ A0 we have i = ak for some k ∈ u−1{zj}, j = 1, . . . , m. If heak , uli 6= 0 then the support supp uk is not contained in [0, ak). But if l ≤ k then the support of ul is contained in [0, ak) by (1). So k < l in every case when heak , uli 6= 0. Accordingly, δ ≤ Xi=ak∈A0Xl>k 2−a2 l vi · hei, uli ≤ Xi=ak∈A0Xl>k 2−a2 l · al · vi because kulk1 ≤ al by (1) again. Now for i = ak ∈ A0, we have γi = 2−a2 l as in (17), we have εi → 0 and k by (16); so writing εi = 2a2 k ·P∞ δ = d+kx; vk − hv, γi ≤ Xi∈A0 l=k+12−a2 εiviγi exactly as in (15). So d+kxk is approximately linear on a cofinite subset A0 ⊂ A, with the derivative γ ∈ l1 given by (16), and the error sequence (εi) given by (17). (cid:3) 4 Using the Hahn-Banach Theorem. Lemma 4.1 Let H ⊂ c0 be a closed subspace of finite codimension, i=1 ker ϕi, where each ϕi ∈ l1. Let x /∈ H be an element of say H = ∩N minimum norm in the coset x+H, and let z1, . . . , zm ∈ c00 be such that hx, zj i 6= 0 for any j = 1, . . . , m. Let A0 ⊂ A = ∪m j=1Azj be a cofinite subset satisfying the conditions of Lemma 3.2, and let γ ∈ l1 be the approximate derivative as in Lemma 3.2, (εi)i∈A0 the error sequence as in (17). Then there is a ϕ ∈ lin{ϕj : i ≤ j ≤ m} such that for every i ∈ A0, we have hei, ϕi − γi ≤ εiγi. (23) Proof. We consider the weak-* topology on l1 with respect to its usual predual, c0. The set G = {ϕ ∈ l1 : hei, ϕi − γi ≤ εiγi (all i ∈ A0), and hei, ϕi = 0 (all i /∈ A0)} is a weak-* compact convex set. The set Φ = lin{ϕi, i = 1, . . . , N} + lin{ej : j ∈ N \ A0} ⊂ l1 is a weak-* closed subspace, because it is {ϕ ∈ l1 : ϕ(u) = 0 for every u ∈ c0 supported on A0, such that ϕi(u) = 0 (i = 1, . . . , N )}. If Φ ∩ G 6= ∅, then the assertion of the Lemma is satisfied. If Φ ∩ G = ∅, then the Hahn-Banach Separation Lemma tells us that there is a weak-* continuous v ∈ l∞ separating them; of course the weak-* continuity means that v ∈ c0. We may assume hϕ, vi = 0 for ϕ ∈ Φ, but hϕ, vi ≥ 1 whenever ϕ ∈ G. Since v annihilates Φ, the support of v is contained in A0. By approximate linearity of d+kxk, from (15) we have d+kx; vk − hv, γi ≤ Xi∈A0 εiviγi; (24) and the same is true with d+ replaced by d−. We cannot have d+kx; vk and d−kx; vk the same sign, or Lemma 2.6 would tell us x does not 7 have minimum norm in the coset x + H. So, as observed after (15), we must have εiviγi. (25) hv, γi ≤ Xi∈A0 Let us write η = hv, γi/Pi∈A0 εiviγi ∈ [−1, 1] (noting that the denominator cannot be zero since εi, γi are never zero for i ∈ A0, and v 6= 0 is supported on A0). Define a new ϕ ∈ l1 by hei, ϕi =(γi(1 − ηεiσ(viγi)), 0, if i ∈ A0; otherwise. (26) We then have hei, ϕi − γi ≤ εiγi (i ∈ A0), so ϕ ∈ G, yet hv, ϕi = hv, γi − η · Xi∈A0 εiviγiσ(viγi) = 0. (27) This contradicts the Hahn-Banach separation of v, which asserts that for such ϕ we should have hv, ϕi ≥ 1. Thus the Lemma is proved. (cid:3) Let us now begin to use our information to investigate proximinal subspaces. If i = al for some l ∈ N, we shall write αi = 2−a2 l . Theorem 4.2 Let H ⊂ (c0, k·k) be a proximinal subspace of finite i=1 ker ϕi, where ϕi ∈ l1. Let zj ∈ c00 codimension, say H = ∩N (j = 1, . . . , m), and A = ∪m i=1Azi . Write Φ = lin{ϕi : i = 1, . . . , N}, and let hei, ϕi : i ∈ A} < ∞}. (28) Φ0 = {ϕ ∈ Φ : sup{α−1 i Let θ0 : Φ0 → l∞(A) be the linear map such that (θ0ϕ)i = α−1 i hei, ϕi (i ∈ A); (29) and let q : l∞(A) → l∞(A)/c0(A) be the quotient map. Write θ = qθ0. Let x ∈ H be an element such that kxk is minimal in the coset x + H, and suppose hx, zii 6= 0 for any i = 1, . . . , m. Then the image θΦ0 includes the vector σx + c0(A) ∈ l∞(A)/c0(A), where (σx)i = σ(hx, zj i) if i ∈ Azj , j = 1, . . . , m. (30) Proof. By Lemma 4.1, there is a ϕ ∈ Φ such that for all but finitely many i ∈ A, we have hei, ϕi − γi ≤ εiγi, where for i = al ∈ Azj we define γi = −2−a2 l σ(hx, zj i) = −αiσ(hx, zj i). Since εi → 0 it is clear that sup{α−1 hei, ϕi} < ∞, so ϕ ∈ Φ0, and the image θϕ is the vector −σx + c0(A), because for i = al ∈ Azj we have (θ0ϕ)i = α−1 hei, ϕi ∈ −σ(hx, zj i) + [−εi, εi]. So, θ(−ϕ) = σx + c0(A). (cid:3) i i 5 Proof of Theorem 1.1 Suppose towards a contradiction that H ⊂ (c0, k·k) is a proximinal subspace of finite codimension N ≥ 2. Any proximinal subspace must i=1 ker ϕi, where the ϕi ∈ l1 are linearly be closed, so let us say H = ∩N independent. For r = 0, . . . , N + 1, let us write βr = rπ/(2N + 2), and 8 for r = 1, . . . , N + 1 let us pick x(r) ∈ c0 such that hx(r), ϕ1i = cos βr, and hx(r), ϕ2i = sin βr. Perturbing each x(r) by an element of H as necessary, we can assume that each (cid:13)(cid:13)x(r)(cid:13)(cid:13) is minimal in the coset x(r) + H. Writing ζr = (βr + βr−1)/2 (r = 1, . . . , N + 1), we define the linear functional ψr = sin ζr · ϕ1 − cos ζr · ϕ2, so hx(r), ψsi = sin ζs cos βr − cos ζs sin βr = sin(ζs − βr); thus hx(r), ψsi > 0 if s > r, but hx(r), ψsi < 0 if s ≤ r. Pick a finite sequence (zr)N +1 r=1 ⊂ c00 with kzr − ψrk1 sufficiently small, and we will also find that hx(r), zsi > 0 if s > r, but hx(r), zsi < 0 if s ≤ r. We find that the sequence (σ(hx(r), zsi))N +1 s=1 ∈ RN +1 is the vector yr = (−1, −1, . . . , −1, 1, 1, . . . 1), where there are r entries −1 followed by N + 1 − r entries +1. It is a fact that the yr span RN +1 - they are linearly independent. We can apply Theorem 4.2 with the sequence z1, . . . , zN +1, and x can be any of the vectors x(1), . . . , x(N +1). The map θ is the same for each x(r) (because the sequence αi doesn't change, only the signs σ(hx(r), zsi)). Writing A = ∪N +1 j=1 Azj , we find that the image θΦ0 must contain, for each r = 1, . . . , N + 1, the vector σx(r) + c0(A) with (σx(r) )i = σ(hx(r), zji) = hyr, eji for all i ∈ Azj . (31) (where here (ej)N +1 j=1 denote the unit vector basis of RN +1). Because the vectors yr are independent, the dimension of θΦ0 must be at least N + 1. However Φ0 ⊂ Φ, and dim Φ = N . This contradiction implies that H is not proximinal. (cid:3) References [1] Bishop, E. and Phelps, R. R., A proof that every Banach space is subreflexive, Bull. Amer. Math. Soc. 67 (1961), 97-98. [2] Blecher, David P. and Read, Charles John, Operator algebras with contractive approximate identities, II. J. Funct. Anal. 264 (2013), no. 4, 1049-1067. [3] Bollob´as, B., An extension to the theorem of Bishop and Phelps, Bull. London Math. Soc. 2 (1970) 181-182. [4] Gamelin, T.W., Uniform Algebras, second edition, Chelsea, New York, 1984. [5] Godefroy, Gilles, Existence and Uniqueness of isometric preduals: a survey. Contemporary Mathematics 85 (1989) 131-193. [6] Singer, Ivan The theory of best approximation and functional analysis. Conference Board of the Mathematical Sciences Regional Conference Series in Applied Mathematics 13. Society for Indus- trial and Applied Mathematics, Philadelphia, Pa., 1974. 9
1611.04185
1
1611
2016-11-13T20:38:28
Positive definite kernels and boundary spaces
[ "math.FA", "math-ph", "math-ph" ]
We consider a kernel based harmonic analysis of "boundary," and boundary representations. Our setting is general: certain classes of positive definite kernels. Our theorems extend (and are motivated by) results and notions from classical harmonic analysis on the disk. Our positive definite kernels include those defined on infinite discrete sets, for example sets of vertices in electrical networks, or discrete sets which arise from sampling operations performed on positive definite kernels in a continuous setting. Below we give a summary of main conclusions in the paper: Starting with a given positive definite kernel $K$ we make precise generalized boundaries for $K$. They are measure theoretic "boundaries." Using the theory of Gaussian processes, we show that there is always such a generalized boundary for any positive definite kernel.
math.FA
math
POSITIVE DEFINITE KERNELS AND BOUNDARY SPACES PALLE JORGENSEN AND FENG TIAN Abstract. We consider a kernel based harmonic analysis of "boundary," and boundary representations. Our setting is general: certain classes of positive definite kernels. Our theorems extend (and are motivated by) results and notions from classical harmonic analysis on the disk. Our positive definite ker- nels include those defined on infinite discrete sets, for example sets of vertices in electrical networks, or discrete sets which arise from sampling operations performed on positive definite kernels in a continuous setting. Below we give a summary of main conclusions in the paper: Starting with a given positive definite kernel K we make precise generalized boundaries for K. They are measure theoretic "boundaries." Using the theory of Gaussian processes, we show that there is always such a generalized boundary for any positive definite kernel. Contents Introduction 1. 2. Generalized boundary spaces for positive definite kernels 3. Boundary theory References 1 3 7 9 1. Introduction Our purpose is to make precise a variety of notions of "boundary" and boundary representation for general classes of positive definite kernels. And to prove theorems which allow us to carry over results and notions from classical harmonic analysis on the disk to this wider context (see [JP98a, JP98b, Str98]). We stress that our positive definite kernels include those defined on infinite discrete sets, for example sets of vertices in electrical networks, or discrete sets which arise from sampling operations performed on positive definite kernels in a continuous setting, and with the sampling then referring to suitable discrete subsets. See, e.g., [JS13, ZS16, HJY11]. Below we give a summary of main conclusions in the paper: Starting with a given positive definite kernel K on S × S, we introduce generalized boundaries for the set S that carries K. It is a measure theoretic "boundary" in the form of a probability space, but it is not unique. The set of measure boundaries will 2000 Mathematics Subject Classification. Primary 47L60, 46N30, 46N50, 42C15, 65R10, 31C20, 62D05, 94A20, 39A12; Secondary 46N20, 22E70, 31A15, 58J65. Key words and phrases. Computational harmonic analysis, Hilbert space, reproducing kernel Hilbert space, discrete analysis, interpolation, reconstruction, Gaussian free fields, distribution of point-masses, Green's function, non-uniform sampling, transforms, optimization, covariance. 1 be denoted M (K). We show that there is always such a generalized boundary probability space associated to any positive definite kernel. For example, as an element in M (K), we can take a "measure" boundary to be the Gaussian process having K as its covariance kernel. This exists by Kolmogorov's consistency theorem. Definition 1.1. By a probability space, we mean a triple (B, F , µ) where: 2 • B is a set, • F is a σ-algebra of subsets of B, and • µ is a probability measure defined on F , i.e., µ (∅) = 0, µ (B) = 1, µ (F ) ≥ 0 ∀F ∈ F , and if {Fi}i∈N ⊂ F , Fi ∩ Fj = ∅, i 6= j in N, then µ (∪iFi) = Pi µ (Fi). Conclusions, a summary: (1) For every positive definite kernel K, we define a "measure theoretic bound- ary space" M (K). Set M (K) :=(cid:8) (B, F , µ) a measure space which yields a factorization for K, see Definition 2.4(cid:9). This set M (K) generalizes other notions of "boundary" used in the lit- erature for networks, and for more general positive definite kernels, and their associated reproducing kernel Hilbert spaces (RKHSs). See, e.g., [JT15b, AJ15, JN15, AJV14]. (2) For any positive definite kernel K, the corresponding M (K) is always non- empty. The natural Gaussian process path-space with covariance kernel K, and Wiener measure µ is in M (K). (3) Given K, let H (K) be the associated RKHS. Then for every µ ∈ M (K) there is a canonical isometry Wµ mapping H (K) into L2 (µ). For details, see Theorem 2.10. (4) The isometry Wµ in (3) generally does not map onto L2 (µ). It does however for the 1 4 -Cantor example, i.e., the restriction of Haus- 2 to the standard 1 dorff measure of dimension 1 4 -Cantor set. In this case, we have a positive definite kernel on D × D, where D is the unit disk in the complex plane; and we can take the circle as boundary for D. For µ, we take the corresponding 1 4 - Cantor measure. But in general, for positive definite functions K, a "measure theoretic boundary space" is much "bigger" than probability spaces on the metric boundary for K. (5) Using the isometries from (3), we can turn M (K) into a partially ordered set; see Definition 3.2. Then, using Zorn's lemma, one shows that M (K) always contains minimal elements. The minimal elements are not unique. (6) And even if µ is chosen minimal in M (K), the corresponding isometry Wµ still generally does not map onto L2 (µ). A case in point: the Szegö kernel, and µ = Lebesgue measure on a period interval. Remark 1.2. The Cantor examples in (4) are special cases of affine-selfsimilarity limit (fractal) contractions. See, e.g., [LLST16, CCEMR16, DJ15, DJ12, BJ11, DHJ09, DJ08, DJ06]. 3 The general role for the fractal dimension in these cases is as follows: dimf ractal = ln s ln d = logd (s) , where s = the number of translations in each iteration, and d = the linear scale. For example, the middle-third Cantor fractal has dimF = ln 2 ln 3 = log3 (2). The Sierpinski-gasket has dimF = ln 3 ln 2 < 2. For the Sierpinski construction in R3, we have dimF = ln 4 ln 2 = 2 < 3. 2. Generalized boundary spaces for positive definite kernels Definition 2.1. Let S be any set. A function K : S × S → C is positive definite iff (Def.) cicjK (si, sj) ≥ 0, (2.1) Xi Xj for all {si}n i=1 ⊂ S, and all (ci)n i=1 ∈ Cn. Remark 2.2. (i) Given a positive definite kernel K on S × S, there is then an associated mapping ES : S → {Functions on S} given by ES (t) = K (t, ·) , (2.2) where the dot "·" in (2.2) indicates the independent variable; so S ∋ s −→ K (t, s) ∈ C. (ii) We shall assume that ES is 1-1, i.e., if s1, s2 ∈ S, and k (s1, t) = k (s2, t), ∀t ∈ S, then it follows that s1 = s2. This is not a strong limiting condition on K. Notation 2.3. We shall view the Cartesian product BS :=YS C = CS (2.3) as the set of all functions S → C. It follows from assumption (ii) that ES : S → BS is an injection, i.e., with ES, we may identity S as a "subset" of BS. For v ∈ S, set πv : BS −→ C, πv (x) = x (v) , ∀x ∈ BS; (2.4) i.e., πv is the coordinate mapping at v. The topology on BS shall be the product topology; and similarly the σ-algebra FS will be the the one generated by {πv}v∈S, i.e., generated by the family of subsets π−1 v (M ) , v ∈ S, and M ⊂ C a Borel set. (2.5) Definition 2.4. Fix a positive definite kernel K : S × S → C. Let M (K) be the set of all probability spaces (see Definition 1.1), so that (B, F , µ) ∈ M (K) iff (Def.) there exists an extension ZB for all (s1, s2) ∈ S × S. K B : S × B −→ C, and K B (s1, b)K B (s2, b) dµ (b) = K (s1, s2) , (2.6) Remark 2.5. In Examples 2.15-3.1, we discuss the case where 4 S = D = {z ∈ C z < 1} B = ∂D =(cid:8)z ∈ C z = 1, or z = eix, x ∈ (−π, π](cid:9) ; but in the definition of M (K), we allow all possible measure spaces (B, F , µ) as long as the factorization (2.6) holds. Questions: (1) Given (2.1) what are the solutions (B, F , µ) to (2.6)? (2) Are there extensions K B : S × B → C such that B is a boundary with respect to the metric on S? That is, distK (s1, s2) = kKs1 − Ks2kH (2.7) and lim K B (·, b) = limi→∞ K (·, si). (3) Find the subsets S0 ⊂ S such that the following sampling property holds for all f ∈ C (B) (or for a subspace of C (B)): f (b) = Xsi∈S0 f (si) K B (si, b) , ∀b ∈ B. (2.8) Example 2.6 (Shannon). Let BL be the space of band-limited functions on R, where We have BL =nf ∈ L2 (R) f (ξ) = 0, ξ ∈ R\(cid:2)− 1 f (t) =Xn∈Z sin π (t − n) π (t − n) f (n) 2 , 1 2(cid:3)o . , ∀t ∈ R, ∀f ∈ BL. (2.9) Definition 2.7. We say (B, F , µ) ∈ GC, generalized Carleson measures, iff (Def.) there exists a constant Cµ such that ZB ef (b) 2dµ (b) ≤ Cµ kf k2 where ef in (2.10) is defined via the extension ef (b) := hK B b , f iH (K), Set (GC)1 := generalized Carleson measures with Cµ = 1. H (K) , ∀f ∈ H (K) , b ∈ B, f ∈ H (K) . (2.10) (2.11) Note. The case Cµ = 1 is of special interest. For classical theory on Carleson measures, we refer to [Tre84, Coh86, Kan11, Zhu12, CIJ15, BFG+15]. Definition 2.8. Let Hi, i = 1, 2 be Hilbert spaces. We say that H1 is boundedly contained in H2 iff (Def.) H1 ⊂ H2 (as a subset), and if the inclusion map H1 → H2, h 7→ h, is bounded. That is, there exits C < ∞ such that for all h ∈ H1, khkH2 ≤ C khkH1. (2.12) Remark 2.9. Note that if (B, F , µ) is a measure space, K : S × S → C is a positive definite kernel, then (B, F , µ) ∈ GC if and only if H (K) is boundedly contained in L2 (B, F , µ); see (2.10). We stress that with the inclusion H (K) "⊂" L2 (µ) we can make the implicit 5 identification f ∼ ef where ef (b) = heKb, f iH (K), and (2.13) is to be understood for a.a. b w.r.t. (F , µ). ∀f ∈ H (K) , b ∈ B; (2.13) In [JT15a], we showed that for all positive definite kernel K (s, t), (s, t) ∈ S × S, we have M (K) 6= ∅. Moreover, Theorem 2.10. Fix a positive definite kernel K : S × S → C, then If (B, F , µ) ∈ M (K), then the mapping M (K) ⊂ (GC)1 . H (K) ∋ K(s, ) −→ K B(s, · ↑ on S · ↑ on B ) ∈ L2 (B, µ) extends by linearity and closure to an isometry (see Definition 2.7) (2.14) (2.15) WB : H (K) −→ L2 (B, µ) , However, WB is generally not onto L2 (B, µ). More specifically, we have or equivalently, (cid:13)(cid:13)(cid:13)Xj Xj1 Xj2 2 H (K) =(cid:13)(cid:13)(cid:13)Xj cjK (sj, ·)(cid:13)(cid:13)(cid:13) cj1 cj2 K (sj1 , sj2) =ZB(cid:12)(cid:12)(cid:12)Xj 2 f −→ ef . cjK B (sj, ·)(cid:13)(cid:13)(cid:13) cjK B (sj, b)(cid:12)(cid:12)(cid:12) , (2.16) L2(B,µ) 2 dµ (b) (2.17) for all finite sums, where {sj}, {cj} ⊂ Cn, ∀n ∈ N. Proof. Suppose (B, F , µ) ∈ M (K), i.e., assume (B, F , µ) is a measure space such (2.7). (2.11): that (2.6) holds. Set K B = eK, refer to the extension eK : S × B → C introduced in We claim that then (2.10) holds for all f ∈ H (K). Here ef is defined via eK; see ef (b) := heKb, f iH (K), ∀f ∈ H (K) , ∀b ∈ B.. Claim: f 7→ ef is isometric from H (K) into L2 (µ), i.e., kefkL2(B,µ) = kf kH (K) , ∀f ∈ H (K) . Proof of (2.18). It is enough to consider the case where f =Pi ciKsi (finite sum), see (2.1); so that ef =Pi cieKsi on B, and (2.18) L2(B,µ) =Xi Xj kefk2 =Xi Xj =Xi Xj cicjheKsieKsj iL2(B,µ) cicjZB eKsi (b)eKsj (b) dµ (b) cicjK (si, sj) (see (2.6) , use µ ∈ M (K)) = kf k2 H (K) , by (2.1) and the defn. of H (K) . (cid:3) Corollary 2.11. Suppose H (K) ∋ f is a Carleson measure, then the adjoint operator WB−−−→ ef ∈ L2 (B, µ) is bounded, i.e., that µ W ∗ B : L2 (B, µ) −→ H (K) 6 is given by W ∗ B (F ) (s) =ZB eK (s, b)F (b) dµ (b) , ∀F ∈ L2 (B, µ) . Proof. For all F ∈ L2 (B, µ), and all s ∈ S, we have hKs, W ∗ BF iH (K) = (W ∗ BF ) (s) (reprod prop., and W ∗ B F ∈ H (K)) (by duality) = hWBKs, F iL2(µ) =ZB eK (s, b)F (b) dµ (b) which is the desired conclusion (2.19). We now turn to the Gaussian measure boundary: (2.19) (cid:3) Corollary 2.12. Suppose K : S × S → C is a given positive definite kernel, and that there is a measure space (F , µ) where F is a σ-algebra of subsets of S, such that the RKHS H (K) satisfies H (K) ⊂ L2 (S, F , µ) (isometric inclusion), then (S, F , µ) ∈ M (K) iff K (s, x)K (t, x) dµ (x) , ∀ (s, t) ∈ S × S. (2.20) K (s, t) =ZS Example 2.13. The condition in (2.20) is satisfied for Bargmann's Hilbert space H of entire analytic functions on C (see [DG88, Bar67]) subject to kf k2 H = = 1 2πZZR2 2πZC 1 f (x + i y)2 e− x2+y2 2 dx dy (2.21) f (z)2 e− z2 2 dx dy < ∞. The following kernel (Bargmann's kernel) is positive definite on C × C: K (z, w) = exp zw 2 − z2 + w2 4 ! . (2.22) It is known that K in (2.22) satisfies (2.20) with respect to the measure µ on C, given by dµ (z) = dA (z) = dx dy. (2.23) 1 2π Theorem 2.14. Let (K, S) be a positive definite kernel (Definition 2.1) such that the associated mapping ES : S → BS is 1-1 (see (2.2)). Then there is a probability space (BS, FS, µS) which satisfies the condition (2.1) in Definition 2.1. Proof. This argument is essentially the Kolmogorov inductive limit construction. For every n ∈ N, ∀ {s1, · · · , sn} ⊂ S, we associate a measure µ{s1,··· ,sn} on BS as follows: Let µ{s1,··· ,sn} be the measure on BS which has (πs1 , · · · , πsn ) as an n vector valued random variable with Gaussian the specific distribution: mean zero, and joint covariance matrix {K (si, sj)}n i,j=1. By a standard argument, one checks that then µ{s1,··· ,sn} is a consistent system of measures on BS; and (by Kolmogorov) that there is a unique probability measure µS on the measure space (BS , FS) such that, for all (s1, · · · , sn), the marginal distribution of µS coincides with µ{s1,··· ,sn}. (cid:3) Example 2.15 (WB is onto). Let 7 V = D = {z ∈ C z < 1} B = ∂D =(cid:8)z ∈ C z = 1, or z = eix, x ∈ (−π, π](cid:9) . K (z, w) = (z, w) ∈ D × D, (2.24) (2.25) Set and K B (z, x) = (z, x) ∈ D × B. Then (2.6) holds for the case when µ 1 = the 1 4 4 -Cantor measure on B; see [DJ06]. ∞Yl=0(cid:16)1 + (zw)4l(cid:17) , ∞Yl=0(cid:18)1 +(cid:0)zei2πx(cid:1)4l(cid:19) , Proof. (Sketch) Set then The desired conclusion = {0, 1, 4, 5, 16, 17, 20, 21, 64, 65, · · ·} bi4i bi ∈ {0, 1} , n ∈ N) Λ4 =( nXi=0 ∞Yl=0(cid:16)1 + t4l(cid:17) =Xλ∈Λ tλ, t < 1. K (z, w) =ZC 1 4 KC 1 (z, x)KC 1 4 4 (w, x) dµ 1 4 (x) (2.26) (2.27) (2.28) follows from the fact that {eλ λ ∈ Λ4} is an ONB in L2(cid:0)C1/4, µ1/4(cid:1) by [JP98a]. (cid:3) 3. Boundary theory We now turn to the details regarding boundary theory. To connect it to the classical theory of kernel spaces of analytic functions on the disk, we begin with an example, and we then turn to the case of the most general positive definite kernels; but not necessarily restricting the domain of the kernels to be considered. Example 3.1 (WB is not onto). Let K (z, w) = K B (z, x) = 1 1 − zw 1 1 − zei2πx , (Szegö kernel) , and (3.1) µ = restriction of Lebesgue measure to [0, 1] . Let H2 be the Hardy space on D. Then WB : H2 −→ L2 ([0, 1] , µLeb) is isometric, but not onto. Indeed, WB (H2) = spanL2(0,1) {en (x) n ∈ N0 = {0} ∪ N} . Returning to the general case, we show below that there is always a minimal element in M (K); see Definition 2.4. Definition 3.2. Suppose (Bi, Fi, µi) ∈ M (K), i = 1, 2. We say that 8 if ∃ϕ : B2 −→ B1, s.t. (B1, F1, µ1) ≤ (B2, F2, µ2) µ2 ◦ ϕ−1 = µ1, and ϕ−1 (F1) = F2. Lemma 3.3. M (K) has minimal elements. Proof. If (3.3)-(3.4) hold, then L2 (B1, µ1) ∋ f W21 −−−−→ f ◦ ϕ ∈ L2 (B2, µ2) is isometric, i.e., and ZB2 W21f {z} f ◦ ϕ 2 dµ2 =ZB1 f 2 dµ1, WB2 = W21WB1 on H (K) , (3.2) (3.3) (3.4) (3.5) (3.6) i.e., the diagram commutes: WB1 L2 (B1, µ1) W21 H (K) / L2 (B2, µ2) WB2 We can then use Zorn's lemma to prove that ∀K, M (K) has minimal elements (B, F , µ). But even if (B, F , µ) is minimal, WB : H (K) → L2 (µ) may not be onto. (cid:3) In the next result, we shall refer to the partial order "≤" from (3.2) when con- sidering minimal elements in M (K). And, in referring to M (K), we have in mind a fixed positive definite function K : S × S → C, specified at the outset; see Definitions 2.1 and 2.4. Theorem 3.4. Let (K, S) be a fixed positive definite kernel, and let M (K) be the corresponding boundary space from Definition 2.4. Then, for every (X, λ) ∈ M (K), there is a (M, ν) ∈ M (K) such that (M, ν) ≤ (X, λ) , and (M, ν) is minimal in the following sense: Suppose (B, µ) ∈ M (K) and then it follows that (B, µ) ≃ (M, ν), i.e., we also have (M, ν) ≤ (B, µ). (B, µ) ≤ (M, ν) , (3.7) (3.8)   0 0 / 9 Proof. We shall use Zorn's lemma, and the argument from Lemma 3.3. Let L = {(B, µ)} be a linearly ordered subset of M (K) s.t. (B, µ) ≤ (X, λ) , ∀ (B, µ) ∈ L; (3.9) and such that, for every pair (Bi, µi), i = 1, 2, in L, one of the following two cases must hold: (3.10) To apply Zorn's lemma, we must show that there is a (BL, µL) ∈ M (K) such that (B1, µ1) ≤ (B2, µ2) , or (B2, µ2) ≤ (B1, µ1) . (BL, µL) ≤ (B, µ) , ∀ (B, µ) ∈ L. (3.11) Now, using (3.9)-(3.10), we conclude that the measure spaces {(B, µ)}L have an inductive limit, i.e., the existence of: µL := ind limit BL B−−→ L µB . (3.12) In other words, we may apply Kolmogorov's consistency to the family L of measure spaces in order to justify the inductive limit construction in (3.12). We have proved that every linearly ordered subset L (as specified) has a "lower bound" in the sense of (3.11). Hence Zorn's lemma applies, and the desired conclu- sion follows, i.e., there is a pair (M, ν) ∈ M (K) which satisfies the condition (3.8) from the theorem. (cid:3) Acknowledgement. The co-authors thank the following colleagues for helpful and en- lightening discussions: Professors Sergii Bezuglyi, Ilwoo Cho, Paul Muhly, Myung- Sin Song, Wayne Polyzou, and members in the Math Physics seminar at The Uni- versity of Iowa. References [AJ15] [AJV14] [Bar67] Daniel Alpay and Palle Jorgensen, Spectral theory for Gaussian processes: reproduc- ing kernels, boundaries, and L2-wavelet generators with fractional scales, Numer. Funct. Anal. Optim. 36 (2015), no. 10, 1239–1285. MR 3402823 Daniel Alpay, Palle Jorgensen, and Dan Volok, Relative reproducing kernel Hilbert spaces, Proc. Amer. Math. Soc. 142 (2014), no. 11, 3889–3895. MR 3251728 V. Bargmann, On a Hilbert space of analytic functions and an associated integral transform. Part II. A family of related function spaces. Application to distribution theory, Comm. Pure Appl. Math. 20 (1967), 1–101. MR 0201959 (34 #1836) [BFG+15] Alain Blandignères, Emmanuel Fricain, Frédéric Gaunard, Andreas Hartmann, and William T. Ross, Direct and reverse Carleson measures for H(b) spaces, Indiana Univ. Math. J. 64 (2015), no. 4, 1027–1057. MR 3385785 Jana Bohnstengel and Palle Jorgensen, Geometry of spectral pairs, Anal. Math. Phys. 1 (2011), no. 1, 69–99. MR 2817339 [BJ11] [CIJ15] [DG88] [Coh86] [CCEMR16] F. Calabrò, A. Corbo Esposito, G. Mantica, and T. Radice, Refinable functions, functionals, and iterated function systems, Appl. Math. Comput. 272 (2016), no. part 1, 199–207. MR 3418124 Hong Rae Cho, Joshua Isralowitz, and Jae-Cheon Joo, Toeplitz operators on Fock- Sobolev type spaces, Integral Equations Operator Theory 82 (2015), no. 1, 1–32. MR 3335506 William S. Cohn, Carleson measures and operators on star-invariant subspaces, J. Operator Theory 15 (1986), no. 1, 181–202. MR 816238 Ingrid Daubechies and A. Grossmann, Frames in the Bargmann space of entire func- tions, Comm. Pure Appl. Math. 41 (1988), no. 2, 151–164. MR 924682 (89e:46028) Dorin Ervin Dutkay, Deguang Han, and Palle E. T. Jorgensen, Orthogonal expo- nentials, translations, and Bohr completions, J. Funct. Anal. 257 (2009), no. 9, 2999–3019. MR 2559724 [DHJ09] 10 [DJ06] [DJ08] [DJ12] [DJ15] [HJY11] [JN15] [JP98a] [JP98b] [JS13] [JT15a] [JT15b] [Kan11] [LLST16] [Str98] [Tre84] [Zhu12] [ZS16] Dorin Ervin Dutkay and Palle E. T. Jorgensen, Iterated function systems, Ruelle operators, and invariant projective measures, Math. Comp. 75 (2006), no. 256, 1931– 1970 (electronic). MR 2240643 , Fourier series on fractals: a parallel with wavelet theory, Radon transforms, geometry, and wavelets, Contemp. Math., vol. 464, Amer. Math. Soc., Providence, RI, 2008, pp. 75–101. MR 2440130 , Fourier duality for fractal measures with affine scales, Math. Comp. 81 (2012), no. 280, 2253–2273. MR 2945155 , Spectra of measures and wandering vectors, Proc. Amer. Math. Soc. 143 (2015), no. 6, 2403–2410. MR 3326023 Takaki Hayashi, Jean Jacod, and Nakahiro Yoshida, Irregular sampling and central limit theorems for power variations: the continuous case, Ann. Inst. Henri Poincaré Probab. Stat. 47 (2011), no. 4, 1197–1218. MR 2884231 Palle E. T. Jorgensen and Robert Niedzialomski, Extension of positive definite func- tions, J. Math. Anal. Appl. 422 (2015), no. 1, 712–740. MR 3263485 Palle E. T. Jorgensen and Steen Pedersen, Dense analytic subspaces in fractal L2- spaces, J. Anal. Math. 75 (1998), 185–228. MR 1655831 , Local harmonic analysis for domains in Rn of finite measure, Analysis and topology, World Sci. Publ., River Edge, NJ, 1998, pp. 377–410. MR 1667822 Palle E. T. Jorgensen and Myung-Sin Song, Compactification of infinite graphs and sampling, Sampl. Theory Signal Image Process. 12 (2013), no. 2-3, 139–158. MR 3285408 P. Jorgensen and F. Tian, Infinite weighted graphs with bounded resistance metric, ArXiv e-prints (2015). Palle Jorgensen and Feng Tian, Discrete reproducing kernel Hilbert spaces: sam- pling and distribution of Dirac-masses, J. Mach. Learn. Res. 16 (2015), 3079–3114. MR 3450534 Si Ho Kang, Some Toeplitz operators on weighted Bergman spaces, Bull. Korean Math. Soc. 48 (2011), no. 1, 141–149. MR 2778503 Benoît Loridant, Jun Luo, Tarek Sellami, and Jörg M. Thuswaldner, On cut sets of attractors of iterated function systems, Proc. Amer. Math. Soc. 144 (2016), no. 10, 4341–4356. MR 3531184 Robert S. Strichartz, Remarks on: "Dense analytic subspaces in fractal L2-spaces" [J. Anal. Math. 75 (1998), 185–228; MR1655831 (2000a:46045)] by P. E. T. Jor- gensen and S. Pedersen, J. Anal. Math. 75 (1998), 229–231. MR 1655832 Tavan T. Trent, Carleson measure inequalities and kernel functions in H 2(µ), J. Operator Theory 11 (1984), no. 1, 157–169. MR 739800 Kehe Zhu, Analysis on Fock spaces, Graduate Texts in Mathematics, vol. 263, Springer, New York, 2012. MR 2934601 Li-kai Zhou and Zhong-gen Su, Discretization error of irregular sampling approxi- mations of stochastic integrals, Appl. Math. J. Chinese Univ. Ser. B 31 (2016), no. 3, 296–306. MR 3541255 (Palle E.T. Jorgensen) Department of Mathematics, The University of Iowa, Iowa City, IA 52242-1419, U.S.A. E-mail address: [email protected] URL: http://www.math.uiowa.edu/~jorgen/ (Feng Tian) Department of Mathematics, Hampton University, Hampton, VA 23668, U.S.A. E-mail address: [email protected]
1911.07377
2
1911
2019-12-01T21:09:23
Some Results and a K-theory Problem about Threshold Commutants mod Normed Ideals
[ "math.FA" ]
We extend to the case of a threshold ideal our result with J. Bourgain about the essential centre of the commutant mod a diagonalization ideal for a n-tuple of commuting Hermitian operators . We also compute the $K_0$-group of the commutant mod trace-class of a unitary operator with spectrum equal to its essential spectrum. We present the problem of computing the $K_1$-group for a commutant mod trace-class in its simplest case.
math.FA
math
Some Results and a K-Theory Problem About Threshold Commutants mod Normed Ideals Dan-Virgil Voiculescu Dedicated to Ciprian Foias on the occasion of his 85th birthday. Abstract. We extend to the case of a threshold ideal our result with J. Bourgain about the essential centre of the commutant mod a diagonalization ideal for a n-tuple of commuting Hermitian op- erators. We also compute the K0-group of the commutant mod trace-class of a unitary operator with spectrum equal to its essen- tial spectrum. We present the problem of computing the K1-group for a commutant mod trace-class in its simplest case. 1. Introduction The commutant modulo a normed ideal of a n-tuple of Hermit- ian operators is a Banach algebra with involution, which is not a C ∗- algebra. However, if the normed ideal is sufficiently large, then the quotient of this Banach algebra by its ideal of compact operators is a C ∗-algebra. An interesting situation arises for certain threshold normed ideals, when the quotient by the compact ideal is not yet a C ∗-algebra, but is isomorphic as a Banach algebra with involution to a C ∗-algebra, that is the quotient norm is equivalent to a C ∗-norm. In technical terms of the quasicentral modulus kJ(τ ) of the n-tuple τ with respect to the normed ideal J the first situation occurs when kJ(τ ) vanishes, while the situation of the threshold ideal is when kJ(τ ) is finite but non-zero (for details and references see our survey paper [12]). When n = 1, the 2010 Mathematics Subject Classification. Primary: 47A55; Secondary: 47A40, 46L80, 47L20. Key words and phrases. absolutely continuous spectrum, K-groups, essential centre, threshold commutant mod normed ideal. Research supported in part by NSF Grant DMS-1665534. 1 2 DAN-VIRGIL VOICULESCU case of one Hermitian operator T , the threshold situation occurs when the normed ideal is the trace-class C1 and T has absolutely continuous spectrum of bounded multiplicity (a sufficient condition). Note that in this setting the Kato -- Rosenblum theorem and trace-class scattering theory can be used. The present paper continues the study of threshold commutants mod normed ideals for n-tuples of commuting Hermitian operators. We deal with three questions. One result is about the centre of the quotient by the compact ideal, which we show, roughly, is generated by the classes of components of τ . This extends to the threshold, the result of [3] in the case of vanishing kJ(τ ). We should say that this doesn't mean that the quotient algebra is isomorphic to the C ∗-algebra of a continuous field of C ∗-algebras over the essential spectrum of τ (this is pointed out in the discussion in section 5). A second result is about the K0-group of the commutant mod trace-class of a unitary operator. This is an extension of the result for one Hermitian operator in [11], with a new feature that roughly the K0-class views the Hardy-subspace as absolutely continuous with multiplicity 1/2. The third question we bring up is the computation of the K1-group in the simplest case. The remarks we make about this problem suggest that techniques from trace-class scattering theory may be relevant. The paper, besides the introduction and references has four more sections. Section 2 contains preliminaries. Section 3 proves the result about the essential centre. Section 4 deals with K0 in the case of a unitary operator. Section 5 is a discussion around K1 in the simplest case. 2. Preliminaries Throughout this paper H will be a complex separable infinite di- mensional Hilbert space and B(H), K(H), B/K(H), or simply B, K, B/K will denote the bounded operators, the compact operators and the Calkin algebra. The canonical homomorphism B → B/K will be denoted by p. If τ = (Tj)1≤j≤n is a n-tuple of bounded Hermitian operators and (J, J) is a normed ideal (see [4], [9]) we denote by E(τ ; J) = {X ∈ B [Tj, X] ∈ J, 1 ≤ j ≤ n} the commutant mod J of τ , which is a Banach algebra with involution when endowed with the norm kXk = kXk + max 1≤j≤n [X, Tj]J. SOME RESULTS AND A K-THEORY PROBLEM 3 We also denote by K(τ ; J) = E(τ ; J) ∩ K and by E/K(τ ; J) = E(τ ; J)/K(τ ; J) its compact ideal and the quotient Banach algebra. The notation also makes sense when we drop the condition that the Tj be Hermitian. The algebra is still involutive when we require only that {T1, . . . , Tn} = {T ∗ n } and also in case n = 1 when T1 = U is a unitary operator. 1 , . . . , T ∗ Note also that E/K(τ ; J) is algebraically isomorphic to p(E(τ ; J)) ⊂ B/K and that the map E/K(τ ; J) → p(E(τ ; J)) is contractive. The quasicentral modulus of τ with respect to J is the number kJ(τ ) = lim inf A∈R+ 1 max 1≤j≤n [A, Tj]J where R+ with the natural order (see [12]). 1 is the set of finite rank positive contractions on H endowed If kJ(τ ) = 0 and the finite rank operators are dense in J, then E/K(τ ; J) is a C ∗-algebra, while if only kJ(τ ) < ∞, E/K(τ ; J) is a Ba- nach algebra with involution isomorphic to a C ∗-algebra and p(E(τ ; J)) is a C ∗-subalgebra of B/K (see [12] and references therein). By Mn we denote the n × n matrices. We have the identification MnE(τ ; J) ∼ E(τ ⊗ In; J). Here τ ⊗ In is viewed as acting on H ⊗ Cn ≃ Hn. The above identifica- tion is up to equivalent norms, but this will not matter in the questions we will consider. By (Cp, p) we denote the Schatten -- von Neumann p-classes. The idempotents in E(τ ; J) will be denoted by PE(τ ; J) and the Hermitian idempotents will be denoted by PhE(τ ; J). If τ is a n-tuple of commuting normed operators we denote by σ(τ ) the joint spectrum and by σe(τ ) = σ(p(τ )) the joint essential spectrum. 3. Commutation mod K The commutation mod K result we prove in this section is a stronger result which has as a corollary the essential centre result for threshold commutants mod normed ideals which extends to the threshold the result of [3]. 4 DAN-VIRGIL VOICULESCU Theorem 3.1. Let τ = (Tj)1≤j≤n be a n-tuple of commuting Hermitian operators, the joint spectrum of which K = σ(τ ) ⊂ Rn is a perfect set. If X is so that then we have X ∈ C ∗(τ ) + K. [X, E(τ ; C1)] ⊂ K Proof. Clearly it suffices to prove the assertion when X = X ∗. Remark also that τ can be replaced by another n-tuple of commut- ing Hermitian operators which has the same spectrum as τ and which is a trace-class perturbation of τ . Indeed this will not change E(τ ; C1) and C ∗(τ ) + K. Using for instance the adaptation of the Voiculescu theorem to normed ideals [10], we can find such a modified τ which is unitarily equivalent to τ ⊕ δ where δ is a given n-tuple of commut- ing Hermitian operators with σ(δ) ⊂ K and which is diagonalizable in some orthonormal basis. Based on this we can assume that for every k ∈ K the joint eigenspace E(τ ; {k})H is either zero or infinite dimen- sional. Indeed, for this we choose δ to be an infinite ampliation of the restriction of τ to the direct sum of the E(τ ; {k})H where k ∈ K. Since (τ )′ ⊂ E(τ ; C1) we have [X, (τ )′] ⊂ K and hence by [5] it follows that X ∈ (τ )′′ + K. This means there is a bounded Borel function f : K → R so that X ∈ f (τ ) + K. Thus the proof reduces to the case when X = f (τ ). Identifying Rn with ℓ∞({1, . . . , n}) we shall use on Rn the met- ric arising from the ℓ∞-norm. If ω ⊂ K is a Borel set, let ∆(ω) = diam(σ(X E(τ ; ω)H) that is the diameter of the spectrum of the restriction of X to the spectral subspace of τ for ω. We shall prove that if diam ωj → 0 for a sequence of Borel sets ωj ⊂ K then we have ∆(ωj) → 0. Suppose the contrary that diam ωj → 0 and ∆(ωj) 9 0 for some sequence. Passing to a subsequence we can assume there is k0 ∈ K so that rj = d(ωj, k0) ↓ 0 and that all ∆(ωj) > ε for some ε > 0. On the other hand, observe that if Ωp ↑ Ω then ∆(Ωp) ↑ ∆(Ω). In particular, we have ∆(ωj\(B(k0, rn)\{k0})) ↑ ∆(ωj) and ∆(ωj\B(k0, rn)) ↑ ∆(ωj\{k0} as n → ∞, where B(k0, rn) denotes the ball in the ℓ∞ metric. In particular, passing to a subsequence and replacing then ωj by ωj\(B(k0; rj+1)\{k0}) we may assume the ωj\{k0} are pairwise disjoint. Remark also that the spectrum of X E(τ ; ωj)H equals σ(X E(τ ; ωj\{k0})H or σ(X E(τ ; ωj\{k0})H) ∪ {f (k0)} depending on whether E(τ ; {k0}) is = 0 or 6= 0. We consider now two cases. SOME RESULTS AND A K-THEORY PROBLEM 5 a) If lim inf j→∞ ∆(ωj\{k0}) = 0, passing to a subsequence so that ∆(ωj\{k0}) → 0, we must have k0 ∈ ωj and E(τ ; {k0}) 6= 0. Moreover since ∆(ωj) > ε > 0, there is λj ∈ σ(X E(τ ; ωj\{k0})H) so that λj − f (k0) > ε/2 for each j. Recall also that by our additional assumptions on τ we have that E[τ ; {k0}]H is infinite dimensional. We can then choose for each j a number µj ∈ {λj, f (k0)} so that µj −µj+1 > ε/4 for all j ∈ N. We also choose an orthonormal sequence (hj)j∈N of vectors so that hj ∈ E(τ ; ωj)H ∪ E(τ ; {k0})H kXhj − µjhjk < 10−j. Here hj ∈ E(τ ; ωj\{k0})H if µj = λj and hj ∈ E(τ ; {k0})H if µj = f (k0). Finding such an orthonormal sequence is possible because the ωj\{k0} are disjoint and E(τ ; {k0})H is infinite dimensional. Consider then the shift operator on the sequence (hj)j∈N: We have that Y = X h·, hjihj+1. j k[Tk, h·, hjihj+1]k = = kh·, Tkhjihj+1 − h·, hjiTkhj+1k ≤ ≤ 10(rj + rj+1). Passing again if necessary to a subsequence we may assume the sequence of rj's is summable and hence [Y, Tk] ∈ C1, 1 ≤ k ≤ n so that Y ∈ E(τ ; C1). On the other hand [X, Y ] − X j (h·, hjiµj+1hj+1 − h·, µjhjihj+1) is compact since k[X, h·hjihj+1] − (µj+1 − µj)h·hj, hj+1ik ≤ kXhj − µjkjk + kXhj+1 − µj+1hj+1k ≤ 2 · 10−j. Since (µj+1 − µj)h·hj, hj+1i X j 6 DAN-VIRGIL VOICULESCU is not compact, because µj+1 − µj ≥ ε/4, we have arrived at a con- tradiction having [Y, X] /∈ K, while Y ∈ E(τ ; C1). b) The other case being when lim inf j→∞ ∆(ωj\{k0}) > 0 which means inf j∈N ∆(ωj\{k0}) > 0, we can find αj, βj ∈ σ(X E(τ ; ωj\{k0})H) so that αj − βj > η for some η > 0 for all j ∈ N. Then we can choose µj ∈ {αj, βj} so that µj − µj−1 > η/2. We can also find unit vectors hj ∈ E(τ ; ωj\{k0})H so that kXhj − µjhjk < 10−j. The ωj\{k0} being disjoint, this sequence will be orthonormal and we can proceed along the same lines as in case a). We consider the shift operator Y = X h·, hjihj+1 j and like in case a) we arrive at the contradiction that [Y, Tk] ∈ C1, 1 ≤ k ≤ n, while [Y, X] /∈ K. Concluding we have proved that diam(ωj) → 0 implies ∆(ωj) → 0. We shall use this to reach the desired conclusion that X = f (τ ) ∈ C ∗(τ ). Thus for every ε > 0 there is ϕ(ε) > 0 so that diam(ω) < ϕ(ε) ⇒ ∆(ω) < ε. Then given an integer m > 0 we construct a continuous function gm : K → R as follows. Let 100δ = ϕ(2−m) and let (ωm,p)1≤p≤N be the partition of K into the Borel sets which are the nonempty K ∩(x+δ[0, 1)n) where x ∈ δZn. Observe that ωm,p ∩ ωm,q 6= ∅ ⇒ d(ωm,p, ωm,q) ≤ 2δ. Let further (θp)1≤p≤N be a continuous partition of unity θp : K → [0, 1] so that d(ωm,p, supp θp) ≤ 3δ. For each 1 ≤ p ≤ N let also sp ∈ σ(X E(τ ; ωm,p)H). We define gm = P1≤p≤N spθp. Then, if θq ωm,p 6= 0 we must have d(ωm,p, ωm,q) < 10δ and hence sp − sq ≤ 2−m. This gives for k ∈ ωm,p gm(k) − sp ≤ X θq(k)sp − sq ≤ 2−m. q Since (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1≤p≤N X spE(τ ; ωm,p) − X ≤ 2−m, (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) we infer that kgm(τ ) − Xk ≤ 2−m+1. On the other hand gm(τ ) ∈ C ∗(τ ) so that X ∈ C ∗(τ ). (cid:3) Theorem 3.2. Let τ be a n-tuple of commuting Hermitian operators the spectrum σ(τ ) of which is a perfect set and let J be a normed ideal SOME RESULTS AND A K-THEORY PROBLEM 7 in which the finite rank operators are dense so that kJ(τ ) < ∞. Then the centre of the C ∗-algebra p(E(τ ; J)) is generated by p(τ ), so that its spectrum is σ(τ ). Proof. The fact that p(E(τ ; J)) is a C ∗-algebra because of kJ(τ ) < ∞ and the fact that [Tk, E(τ ; J)] ⊂ K, 1 ≤ k ≤ n, implies that the C ∗- algebra generated by p(τ ) is contained in the centre of p(E(τ ; J)). The opposite inclusion follows from X ∈ E(τ ; J)). The opposite inclusion follows from X ∈ E(τ ; J) + K and [X, E(τ ; J)] ⊂ K then [X, E(τ ; C1)] ⊂ K and Theorem 3.1 gives X ∈ C ∗(τ ) + K. Note also that since σ(τ ) is a perfect set we have σ(τ ) = σ(p(τ )). (cid:3) 4. The case of a unitary operator In this section we compute the ordered group K0(E(U; C1)) where U is a unitary operator with σ(U) = σe(U). Our previous result in [11] for a Hermitian operator is a particular case. Indeed if T = T ∗ and U = exp(iαT ) where α > 0, αkT k < π then E(U; C1) = E(T ; C1). Conversely if σ(U) 6= T then there is T = T ∗ so that E(U; C1) = E(U, U ∗; C1). Thus the novelty will appear in case σ(U) = T. Remark that E(U; C!) = E(U, U ∗; C1) and also the norms are equal. The equivalence relation P ∼ Q in PE(U; C1), P = XY , Q = Y X with X, Y ∈ E(U; C1), like in [11] is easily seen to be equivalent to PP H = V ∗V, PQH = V V ∗ where PX is the orthogonal projection onto X and V is a partial isome- try in E(U; C1). Moreover P ∈ PE(U; C1) implies PP H ∈ E(U; C1). We also have P ∼ PP H. Thus we can work with Hermitian projections. Also, as pointed out in [11], if P, Q ∈ Ph(E(U; C1)) then P ∼ Q iff there is a partial isometry V ∈ B(H) so that V V ∗ = P , V ∗V = Q and V QUQ − P UP V ∈ C1 and this then also implies that V ∈ E(U; C1). The absolutely continuous and singular subspaces for U will be denoted by Hac(U) and Hsin g(U) and the corresponding projections Eac(U) and Esin g(U). The multiplicity function of UHac(U) will be denoted by mac(U) and is a Borel function defined on T up to Haar null sets and taking values in {0, 1, 2, . . . , ∞}. We shall also denote by ω(U) the set mac−1(∞), defined up to a null-set. When dealing with matrices we pass to U ⊗ In and Hn as outlined in the preliminaries. It will be convenient to consider an extension of the multiplicity function to deal with unitaries mod trace-class, that is the set U C1(H) = {F ∈ B(H) F ∗F − I ∈ C1, F F ∗ − I ∈ C1}. 8 DAN-VIRGIL VOICULESCU Remark that if F ∈ UC1(H) then F is a Fredholm operator and if ind F = 0 then there is a unitary operator F1 so that F − F1 ∈ C1 and the multiplicity function of the absolutely continuous part of F1 does not depend on the choice of F1. Indeed if F2 is another unitary operator so that F − F1 ∈ C1 then F1 − F2 ∈ C1 and by [1] we have F1 Hac(F1) and F2 Hac(F2) are unitarily equivalent. The existence of F1 can be seen as follows, first replace F by F so that F is invertible and F − F is finite rank and then F1 = F ( F ∗ F )−1/2 will do. Thus on the set Ω = {F ∈ UC1(H) ind F = 0} there is a well-defined essential absolutely continuous multiplicity func- tion emac(F ) so that if F1, F2 ∈ Ω and F1 − F2 ∈ C1 then emac(F1) = emac(F1) Haar a.e. and if F ∈ Ω is unitary, then mac(F ) = emac(F ) Haar a.e. Remark also that emac(F1 ⊕F2) = emac(F1)+emac(F2). To extend emac to all of UC1(H) let J be a conjugate-linear antiunitary involution of H and define emac(F ) = 1 2 emac(F ⊕ JF ∗J) for any F ∈ UC1(H). Indeed ind(F ⊕ JF ∗J) = 0 so the right-hand side has been defined. Clearly the definition does not depend on the choice of J and on the choice of F mod C1. Remark also that if F ∈ UC1(H) and pol(F ) is the partial isometry from F ∗H to F H then F − pol(F ) ∈ C1 so that emac(F ) = emac(pol(F )). We can now proceed as follows to see what emac(F ) is. We can extend pol(F ) to V which is an isometry or co-isometry depending on whether the index is ≤ 0 or ≥ 0. Then emac(pol(F )) = emac(V ). The Wold decomposition of V shows that V is unitarily equivalent to W or W ⊕ Sn or W ⊕ S∗n where W is a unitary operator and S is a unilateral shift of multiplicity one and then emac(F ) = emac(V ) = mac(W ) + 1 2 ind V Indeed mac(W ) = mac(JW ∗J) while since n = ind V = ind F . emac(Sn ⊕ S∗n) = n since Sn ⊕ S∗n is a finite rank perturbation of a bilateral shift of multiplicity n for which mac is n. We record the results of this discussion as the next Lemma. Lemma 4.1. The essential multiplicity function emac(F ) of a mod trace-class unitary operator is defined Haar a.e. on T and takes values in {0, 1/2, 1, 3/2, 2, . . . ∞} and has the following properties a) If F is unitary then emac(F ) = mac(F ). b) If F1, f2 are mod trace-class unitary and W is a unitary operator so that W F1 − F2W ∈ C1 then emac(F1) = emac(F2). SOME RESULTS AND A K-THEORY PROBLEM 9 c) emac(F1 ⊕ F2) = emac(F1) + emac(F2). d) We have emac(F ∗)(eiθ) = emac(F )(e−iθ) and if J is a conjugate-linear antiunitary operator the also emac(JF J)(eiθ) = emac(F )(e−iθ). e) If ind F ≡ 0(mod 2) then emac(F ) takes values in {0, 1, 2, . . . , ∞} and if ind F ≡ 1(mod 2) then emac(F ) takes values in {1/2, 3/2, 5/2, . . . , ∞}. Moreover emac(F ) ≥ 1/2ind F . f) If S is the unilateral shift operator in ℓ2(N) then emac(Sn) = emac(S∗n) = n/2. g) emac(F ) = emac(pol(F )). We will also need to prove a second lemma. Lemma 4.2. Let F1, F2 ∈ UC1(H) be so that σ(Fj) ⊃ T, j = 1, 2, ind F1 = ind F2 and emac(F1) = emac(F2). Then there is a unitary operator W so that W F1 − F2W ∈ C1. Proof. To begin observe that the statement for the Fj's is equivalent to that for the F ∗ j 's, so we may assume ind Fj ≤ 0. Observe also that the spectrum condition is equivalent to σe(Fj) = T. To prove the lemma we replace successively the Fj's by others which are unitarily equivalent mod C1 to them, and hence satisfy the same assumptions, till we get to an obvious assertion. First we pass from Fj to pol(Fj) and then after a finite rank perturbation, the index being ≤ 0, we can assume Fj is an isometry. Using then, for instance, the adapted Voiculescu theorem [10] we can replace Fj by Fj ⊕ Gj where Gj is a unitary operator with singular spectrum. Passing to the Wold decomposition of the isometry Fj we arrive at Ej ⊕ Sn ⊕ Gj where S is the unilateral shift and E1, E2 are unitary operators with equal mac(Ej) = emac(Fj) − n/2, j = 1, 2. Choosing G1 = E2 Hsing(E2), G2 = E1 Hsing(E1) we arrive at unitary equivalence. (cid:3) Like in the case of a Hermitian operator T in [11], where we de- fined an ordered group F(T ) and constructed an isomorphism with K0(E(T ; C1)), we shall define an ordered group F(U) and construct an isomorphism with K0(E(U; C1)), only F(U) will be slightly more com- plicated to describe than F(T ). The elements of F(U) are pairs (f, u) 10 DAN-VIRGIL VOICULESCU where f : T\ω(U) → 1/2Z is a measurable function up to almost ev- erywhere equivalence, n ∈ Z and so that n + f (z) ≤ C mac(U)(z) for some constant C and 2f (z) ≡ n(mod 2) for almost all z ∈ T\ω(U). The operation on F(U) is componentwise addition. Moreover F(U) is an ordered group, the semigroup of elements which are ≥ 0 being F+(U) = {(f, n) ∈ F(U) f (z) ≥ n/2 for z ∈ T\ω(U)a.e.} Remark that unless mac(U) ≥ 1, we must have n = 0 and F(U) looks like the F(T ). Theorem 4.2. Let U be a unitary operator with σ(U) = σe(U). If P ∈ Ph(E(U ⊗ In; C1)) then P (U ⊗ In) P Hn ∈ UC1(P Hn). There exists a unique isomorphism f emac(U) : K0(E(U; C1)) → F(U) so that if P ∈ Pn(E(U ⊗ In; C1)) then f emac(U)([P ]0) = (emac(P (U⊗In) P Hn) T\ω(U), ind P (U⊗In) P Hn). Proof. If σ(U) 6= T the remarks preceding the theorem show that the statement is equivalent to the result in [11]. Thus we may assume σ(U) = T. Since [P, U ⊗ In] ∈ C1 it follows that P (U ⊗ In) P Hn ∈ UC1(P Hn). To check that f emac(U)([P ]0) is well-defined let us first see that the formula depends only on [P ]0. Indeed if [P ]0 = [Q]0 then P ⊕ I ⊗ Im ∼ Q ⊕ I ⊗ Im for some m. Then (P ⊕ I ⊗ Im)(U ⊗ In+m) (P Hn ⊕ Hm) and (Q ⊕ I ⊗ Im)(U ⊗ In+m) (QHn ⊕ Hm) are unitarily equivalent mod C1 and this gives emac((P ⊕ I ⊗ Im)(U ⊗ In+m) (P Hn ⊕ Hm)) = emac((Q ⊕ I ⊗ Im)(U ⊗ In+m) (QHn ⊕ Hm)) which gives emac(P U P Hn) + memac(U) = emac(QU QHn) + memac(U) and hence emac(P U P Hn) (T\ω(U)) = emac(QU QHn) (T\ω(U)). We also have clearly ind((P ⊕ I ⊗ Im)(U ⊗ In+m) (P Hn ⊕ Hm)) = ind((Q ⊕ I ⊗ Im)(U ⊗ In+m) (QHn ⊕ Hm) which means that ind(P (U ⊗ In) P Hn) = ind(Q(U ⊗ In) QHn). SOME RESULTS AND A K-THEORY PROBLEM 11 Next we need also to check that emac(P (U ⊗ In) P Hn) (T − ω(U)), ind(P (U ⊗ Im) P Hn)) ∈ F(U). We have emac(P (U ⊗ In) P Hn) ≥ 1/2ind(P (U ⊗ In) P Hn) and 2emac(P (U ⊗ In) P Hn) ≡ ind(P (U ⊗ In) P Hn)(mod 2) by Lemma 4.1 e) applied to F = P (U ⊗ In) P Hn. If G = (I ⊗ In − P )(U ⊗ In) (I ⊗ In − P )Hn), then F ⊕ G is unitarily equivalent mod C1 to U ⊗ In so that by Lemma 4.1 c), b) and a) we have emac(F ) ≤ emac(F ⊕ G) = emac(U ⊗ In) = nmac(U). Thus f emac(U)([P ]0) is well-defined. Also since for P ∈ Ph(E(U ⊗ In; C1)), Q ∈ Ph(E(U ⊗ In; C1) we have (P ⊕Q)(U⊗In+m) (P ⊕Q)Hn+m−(P (U⊗In)P Hn)⊕(Q(U⊗Im) QHm) ∈ C1 using Lemma 4.1 we get f emac(U)([P ]0) + f emac(U)([Q]0) = f emac(U)([P ⊕ Q]0). This implies that f emac(U) extends to a homomorphism, which is also unique, of K0(E(U; C1)) into F(U). We shall denote in the rest of the proof this extension still by f emac(U). The next step is to check that f emac(U) is one-to-one and onto. To check that f emac(U) is onto, it suffices to show that for every (f, n) ∈ F+(U) there is P ∈ Ph(E(U ⊗ Im; C1)) for some m ∈ N, so that Indeed, (g, n) ∈ F(U) implies (g, n) ∈ f emac(U)([P ]0) = (f, n). F(U) so (g, n) = (g, n)−(g−g, n−n) ∈ F+(U)−F+(U). If (f, n) ∈ F+(U) let f : T → {0, 1, 2, . . . } be defined as f (z) = f (z)−n/2 for z ∈ T\ω(U) and f (z) = 0 for z ∈ ω(U). Since f ≤ C mac(U) we can find a projection Q ∈ (U ⊗Im)′ for some m so that macQ(U ⊗Im) QHm = f . Thus ( f , 0) = f emac(U)([Q]0) and we must find a projection P so that (n/2, n) = f emac(U)([T ]0). If n 6= 0, (n/2, n) ∈ F(U) implies that mac(U) ≥ 1 on T. Then there is a projection P ∈ Ph(E(U; C1)), P ∈ (U)′ so that mac( P U P H) = 1. Thus P U P H is unitarily equivalent to the bilateral shift in ℓ2(Z) and hence we can find a projection ≈ P ∈ Ph(E(U; C1)) so that ≈ P U ≈ P H is unitarily equivalent to the unilateral shift of multiplicity 1 if n > 0. It follows that if P = have f emac(U)([P ]0) = (n/2, n). Note that this also proves that F+(U) ⊂ f emac(U)((K0(E(U; C1))+). ≈ P ⊗ In we will 12 DAN-VIRGIL VOICULESCU To show f emac(U) is one-to-one it suffices to show that if P, Q ∈ Ph(E(U ⊗ In, C1)) are so that f emac(U)([P ]0) = f emac(U)([Q]0) then I ⊕ P ∼ I ⊕ Q. The equality of f emac(U) for [P ]0 and [Q]0 gives that ind(P (U ⊗ In) P Hn) = ind(Q(U ⊗ In) QHn) and emac(P (U ⊗In) P Hn) T\ω(U) = emac(Q(U ⊗In) QHn) T\ω(U). Note that we can replace P, Q by P ⊕ I, Q ⊕ I and n by n + 1 and the Fredholm indices don't change, but the equality of the emac will extend to all of T. Then and F1 = (P ⊕ I)(U ⊗ In+1) (P ⊕ I)Hn+1 F2 = (Q ⊕ I)(U ⊗ In+1) (Q ⊕ I)Hn+1 satisfy the assumptions of Lemma 4.2. Then the unitary equivalence mod C1 of F1 and F2 which the lemma asserts, gives P ⊕ I ∼ Q ⊕ I. Thus f emac(U) is a bijection and to conclude the proof we need only to show that f emac(U)([P ]0) ∈ F+(U). This follows from Lemma 4.1 e) applied to P (U ⊗ In) P Hn. (cid:3) 5. A K-theory problem We point out in this section the simplest case of the problem of computing the K1-group, accompanied by a few remarks. Problem. Let T be the Hermitian operator of multiplication by the coordinate function in L2([0, 1], dλ), (T f )(x) = xf (x), dλ Lebesgue measure. What is K1(E(T, C1)? Is it non-trivial? An example of a unitary operator U in E(T, C1) about the triviality of the K1-class of which one may wonder is the following. Using for instance the absorption version of the Voiculescu theorem [10], there is a unitary operator V : L2([0, 1], dλ) → L2([0, 1], dλ) ⊕ ℓ2(N) ⊕ ℓ2(N) so that V T V ∗ − T ⊕ αI ⊕ βI ∈ C1 where 0 ≤ α < β ≤ 1. Clearly this shows that E(T ; C1) and E(T ⊕αI ⊕ βI) are isomorphic. Let S be the unilateral shift operator in ℓ2(N) and let U = R + I ⊕ S ⊕ S∗ SOME RESULTS AND A K-THEORY PROBLEM 13 where R is a rank one partial isometry from 0⊕0⊕ker S∗ to 0⊕ker S∗⊕0 so that I ⊕ S ⊕ S∗ + R is unitarily equivalent to I ⊕ W where W is a bilateral shift operator i.e. a unitary operator. Is [U]1 trivial or not? A general remark one may find useful is that by [11] we have K0(K(T, C1)) = K0(K) = Z and K1(K(T, C1)) = K1(K) = 0 so that the 6-terms K-theory exact sequence for 0 → K(T, C1) → E(T, C1) → E/K(T, C1) → 0 gives that K1(E(T, C1)) is isomorphic to ker ∂ ⊂ K1(E/K(T, C1)) where ∂ is the connecting map given by the Fredholm index from K1(E/K(T ; C1)) → K0(E(T ; C1)) (which is easily seen to be surjective). Recall also that E/K(T ; C1) is isomorphic to a C ∗-algebra with center generated by the class of T in B/K i.e. T + K. The joint spectrum of the classes of T and U in E/K(T ; C1) is then ([0, 1] × {1}) ∪ ({α, β} × T) ⊂ C × C where T = {z ∈ C z = 1}. This also shows that E/K(T ; C1) is far from being the C ∗-algebra of a continuous field of C ∗-algebras over [0, 1] the spectrum of its center. We recall that if B and A are Hermitian operators without singular spectrum and B − A ∈ C1, then the Kato -- Rosenblum theorem ([6], [7], [8]) implies that the strong limit W +(B, A) = s − lim t→+∞ eitBe−itA exists and is a unitary operator intertwining B and A BW +(B, A) = W +(B, A)A. A consequence of this is the following result about K1(E(T ; C1)). Proposition 5.1. Let U ∈ E(T ⊗ In; C1) ≃ Mn(E(T ; C1)) be a unitary operator. Then in K1(E(T ; C1)) we have [U]1 = [W +(U(T ⊗ In)U ∗, T ⊗ In)]1. Proof. Since U(T ⊗In)U ∗ = W +(U(T ⊗In)U ∗, T ⊗In)(T ⊗In)(W +(U(T ⊗In)U ∗, T ⊗In))∗ we have U ∗W +(U(T ⊗ In)U ∗, T ⊗ In) ∈ (T ⊗ In)′. The commutant (T ⊗In)′ is a von Neumann algebra so K1((T ⊗In)′) = 0 and this implies [U ∗W +(U(T ⊗ In)U ∗, T ⊗ In)]1 = 0 which is the desired result. (cid:3) 14 DAN-VIRGIL VOICULESCU The preceding proposition used only the fact that the spectrum of T is absolutely continuous. Also, the proposition holds with W + replaced by W −(B, A) = W +(−B, −A). The interest of Proposition 5.1 for the K1-problem is that the K1- classes are precisely the classes of W +(A, T ⊗In) where A−T ⊗In ∈ C1 is so that W +(A, T ⊗In) is unitary. Thus the problem can be rephrased in terms of special trace-class perturbations of the T ⊗ In. References [1] Birman, M. S., and Krein, M. G., On the theory of wave operators and scatter- ing operators, Dokl. Akad. Nauk SSSR 144, 475 -- 478 (1962) (Sov. Math. Dokl. 3, 740 -- 747 (1962). [2] Blackadar, B., K-theory for operator algebras, MSRI Publications, Vol. 5, Springer Verlag, 1986. [3] Bourgain, J., and Voiculescu, D. V., The essential centre of the mod-a- diagonalization ideal commutant of an n-tuple of commuting Hermitian op- erators, Non-commutative analysis, operator theory and applications, 77 -- 80, Oper. Theory Adv. Appl. 252, Linear Oper. Linear Syst., Birkhauser/Springer (2016). [4] Gohberg, I. C., and Krein, M. G, Introduction to the theory of linear non- selfadjoint operators, Translations of Mathematical Monographs, Vol. 18, AMS, Providence, RI (1969). [5] Johnson, B. E., and Parrott, S. K., Operators commuting modulo the set of compact operators with a von Neumann algebra, J. Funct. Anal. 11 (1972), 39 -- 61. [6] Kato, T., "Perturbation Theory for Linear Operators", Classics of Mathemat- ics, Springer Verlag, 1995. [7] Putnam, C. R., "Commutation Properties of Hilbert Space Operators and Related Topics", Springer Verlag, Berlin -- Heidelberg -- New York, 1967. [8] Reed, M., and Simon, B., Methods of modern physics, Vol. III: Scattering theory, Academic Press, 1979. [9] Simon, B., Trace ideals and their applications, 2nd Ed., Mathematical Surveys and Monographs, Vol. 120, AMS, Providence, RI, 2005. [10] Voiculescu, D. V., Some results on norm-ideal perturbations of Hilbert space operators I, J. Operator Theory 1 (1979), 3 -- 37. [11] Voiculescu, D. V., K-theory and perturbations of absolutely continuous spectra, Comm. Math. Phys. 365 (2019), No. 1, 363 -- 373. [12] Voiculescu, D. V., Commutants mod normed ideals, arXiv:1810.12497. D.V. Voiculescu, Department of Mathematics, University of Cali- fornia at Berkeley, Berkeley, CA 94720-3840
1010.1848
1
1010
2010-10-09T13:47:22
Mean convergence of Fourier-Dunkl series
[ "math.FA", "math.CA" ]
In the context of the Dunkl transform a complete orthogonal system arises in a very natural way. This paper studies the weighted norm convergence of the Fourier series expansion associated to this system. We establish conditions on the weights, in terms of the $A_p$ classes of Muckenhoupt, which ensure the convergence. Necessary conditions are also proved, which for a wide class of weights coincide with the sufficient conditions.
math.FA
math
Mean convergence of Fourier-Dunkl series ´Oscar Ciaurria,1, Mario P´erezb,1,2,∗, Juan Manuel Reyesc, Juan Luis Varonaa,1 aCIME and Departamento de Matem´aticas y Computaci´on, Universidad de La Rioja, 26004 Logrono, Spain bIUMA and Departamento de Matem´aticas, Universidad de Zaragoza, 50009 Zaragoza, Spain cDepartament de Tecnologia, Universitat Pompeu Fabra, 08003 Barcelona, Spain 0 1 0 2 t c O 9 ] Abstract In the context of the Dunkl transform a complete orthogonal system arises in a very natural way. This paper studies the weighted norm convergence of the Fourier series expansion associated to this system. We establish conditions on the weights, in terms of the Ap classes of Muckenhoupt, which ensure the convergence. Necessary conditions are also proved, which for a wide class of weights coincide with the sufficient conditions. Keywords: Dunkl transform, Fourier-Dunkl series, orthogonal system, mean convergence 2000 MSC: Primary 42C10; Secondary 33C10 . A F h t a m [ 1 v 8 4 8 1 . 0 1 0 1 : v i X r a 1. Introduction For α > −1, let Jα denote the Bessel function of order α: Jα(x) =(cid:16) x 2(cid:17)α ∞Xn=0 (−1)n(x/2)2n n! Γ(α + n + 1) (a classical reference on Bessel functions is [17]). Throughout this paper, by Jα(z) function zα we denote the even , z ∈ C. (1) In this way, for complex values of z, let 1 2α ∞Xn=0 (−1)n(z/2)2n n! Γ(α + n + 1) Iα(z) = 2αΓ(α + 1) Jα(iz) (iz)α = Γ(α + 1) (z/2)2n n! Γ(n + α + 1) ; ∞Xn=0 the function Iα is a small variation of the so-called modified Bessel function of the first kind and order α, usually denoted by Iα. Also, let us take Eα(z) = Iα(z) + z 2(α + 1) Iα+1(z), z ∈ C. These functions are related with the so-called Dunkl transform on the real line (see [6] and [7] for details), which is a generalization of the Fourier transform. In particular, E−1/2(x) = ex and the Dunkl transform ∗Corresponding author Email addresses: [email protected] ( ´Oscar Ciaurri), [email protected] (Mario P´erez), [email protected] (Juan Manuel Reyes), [email protected] (Juan Luis Varona) 1Supported by grant MTM2009-12740-C03-03, Ministerio de Ciencia e Innovaci´on, Spain 2Supported by grant E-64, Gobierno de Arag´on, Spain This paper has been published in J. Math. Anal. Appl. 372 (2010), 470 -- 485; doi:10.1016/j.jmaa.2010.07.029 of order α = −1/2 becomes the Fourier transform. Very recently, many authors have been investigating the behaviour of the Dunkl transform with respect to several problems already studied for the Fourier transform; for instance, Paley-Wiener theorems [1], multipliers [4], uncertainty [16], Cowling-Price's theorem [11], transplantation [14], Riesz transforms [15], and so on. The aim of this paper is to pose and analyse in this new context the weighted Lp convergence of the associated Fourier series in the spirit of the classical scheme which, for the trigonometric Fourier series, can be seen in Hunt, Muckenhoupt and Wheeden's paper [10]. The function Iα is even, and Eα(ix) can be expressed as Eα(ix) = 2αΓ(α + 1)(cid:18) Jα(x) xα + Jα+1(x) xα+1 xi(cid:19) . x Let {sj}j≥1 be the increasing sequence of positive zeros of Jα+1. The real-valued function Im Eα(ix) = 2(α+1) Iα+1(ix) is odd and its zeros are {sj}j∈Z where s−j = −sj and s0 = 0. In connection with the Dunkl transform on the real line, two of the authors introduced the functions ej, j ∈ Z, as follows: e0(x) = 2(α+1)/2Γ(α + 2)1/2, ej(x) = 2α/2Γ(α + 1)1/2 Iα(isj) Eα(isjx), j ∈ Z \ {0}. The case α = −1/2 corresponds to the classical trigonometric Fourier setting: I−1/2(z) = cos(iz), I1/2(z) = sin(iz) , sj = πj, E−1/2(isjx) = eiπjx, and {ej}j∈Z is the trigonometric system with the appropriate mul- tiplicative constant so that it is orthonormal on (−1, 1) with respect to the normalized Lebesgue measure (2π)−1/2 dx. iz For all values of α > −1, in [5] the sequence {ej}j∈Z was proved to be a complete orthonormal system in L2((−1, 1), dµα), dµα(x) = (2α+1Γ(α + 1))−1x2α+1 dx. That is to say and for each f ∈ L2((−1, 1), dµα) the series Z 1 −1 ej(x)ek(x) dµα(x) = δjk ∞Xj=−∞(cid:18)Z 1 −1 f (y)ej(y) dµα(y)(cid:19) ej(x), which we will refer to as Fourier-Dunkl series, converges to f in the norm of L2((−1, 1), dµα). The next step is to ask for which p ∈ (1, ∞), p 6= 2, the convergence holds in Lp((−1, 1), dµα). The problem is equivalent, by the Banach-Steinhauss theorem, to the uniform boundedness on Lp((−1, 1), dµα) of the partial sum operators Snf given by Snf (x) =Z 1 −1 f (y)Kn(x, y) dµα(y), where Kn(x, y) = nXj=−n ej(x)ej (y). We are interested in weighted norm estimates of the form kSn(f )U kLp((−1,1),dµα) ≤ Ckf V kLp((−1,1),dµα), where C is a constant independent of n and f , and U , V are nonnegative functions on (−1, 1). Before stating our results, let us fix some notation. The conjugate exponent of p ∈ (1, ∞) is denoted by p′. That is, 1 p + 1 p′ = 1, or p′ = p p − 1 . 2 For an interval (a, b) ⊆ R, the Muckenhoupt class Ap(a, b) consists of those pairs of nonnegative functions (u, v) on (a, b) such that (cid:18) 1 IZI u(x) dx(cid:19)(cid:18) 1 IZI v(x)− 1 p−1 dx(cid:19)p−1 ≤ C, for every interval I ⊆ (a, b), with some constant C > 0 independent of I. The smallest constant satisfying this property is called the Ap constant of the pair (u, v). We say that (u, v) ∈ Aδ p(a, b) (where δ > 1) if (uδ, vδ) ∈ Ap(a, b). It follows from Holder's inequality that Aδ p(a, b) ⊆ Ap(a, b). If u ≡ 0 or v ≡ ∞, it is trivial that (u, v) ∈ Ap(a, b) for any interval (a, b). Otherwise, for a bounded interval (a, b), if (u, v) ∈ Ap(a, b) then the functions u and v− 1 p−1 are integrable on (a, b). Throughout this paper, C denotes a positive constant which may be different in each occurrence. 2. Main results We state here some Ap conditions which ensure the weighted Lp boundedness of these Fourier-Dunkl orthogonal expansions. For simplicity, we separate the general result corresponding to arbitrary weights in two theorems, the first one for α ≥ −1/2 and the second one for −1 < α < −1/2. Theorem 1. Let α ≥ −1/2 and 1 < p < ∞. Let U , V be weights on (−1, 1). Assume that (cid:16)U (x)px(α+ 1 2 )(2−p), V (x)px(α+ 1 2 )(2−p)(cid:17) ∈ Aδ p(−1, 1) (2) for some δ > 1 (or δ = 1 if U = V ). Then there exists a constant C independent of n and f such that kSn(f )U kLp((−1,1),dµα) ≤ Ckf V kLp((−1,1),dµα). Theorem 2. Let −1 < α < −1/2 and 1 < p < ∞. Let U , V be weights on (−1, 1). Let us suppose that U , V satisfy the conditions (cid:16)U (x)px(2α+1)(1−p), V (x)px(2α+1)(1−p)(cid:17) ∈ Aδ (cid:0)U (x)px2α+1, V (x)px2α+1(cid:1) ∈ Aδ p(−1, 1), p(−1, 1) for some δ > 1 (or δ = 1 if U = V ). Then there exists a constant C independent of n and f such that (3) (4) kSn(f )U kLp((−1,1),dµα) ≤ Ckf V kLp((−1,1),dµα). As we mentioned in the introduction, the case α = −1/2 corresponds to the classical trigonometric case. Accordingly, (2) reduces then to (U p, V p) ∈ Aδ p(−1, 1). It should be noted also that taking real and imaginary parts in these Fourier-Dunkl series we would obtain the so-called Fourier-Bessel series on (0, 1) (see [18, 2, 3, 9]), but the known results for Fourier-Bessel series do not give a proof of the above theorems. Also in connection with Fourier-Bessel series on (0, 1), Lemma 3 below can be used to improve some results of [9]. Theorems 1 and 2 establish some sufficient conditions for the Lp boundedness. Our next result presents some necessary conditions. To avoid unnecessary subtleties, we exclude the trivial cases U ≡ 0 and V ≡ ∞. Theorem 3. Let −1 < α, 1 < p < ∞, and U , V weights on (−1, 1), neither U ≡ 0 nor V ≡ ∞. If there exists some constant C such that, for every n and every f , kSn(f )U kLp((−1,1),dµα) ≤ Ckf V kLp((−1,1),dµα), 3 then U ≤ CV almost everywhere on (−1, 1), and U (x)px(α+ 1 2 )(2−p) ∈ L1((−1, 1), dx), = V (x)−p′ x(α+ 1 2 )(2−p′) ∈ L1((−1, 1), dx), U (x)px2α+1 ∈ L1((−1, 1), dx), = V (x)−p′ x2α+1 ∈ L1((−1, 1), dx). (cid:16)V (x)px(α+ 1 2 )(2−p)(cid:17)− 1 p−1 (cid:16)V (x)px(2α+1)(1−p)(cid:17)− 1 p−1 Notice that the first two integrability conditions imply the other two if α ≥ −1/2, while the last two imply the other if −1 < α < −1/2. When U , V are power-like weights, it is easy to check that the conditions of Theorem 3 are equivalent to the Ap conditions (2), (3), (4). By power-like weights we mean finite products of the form x − tγ, for some constants t, γ. For these weights, therefore, Theorems 1, 2 and 3 characterize the boundedness of the Fourier-Dunkl expansions. For instance, we have the following particular case: Corollary. Let b, A, B ∈ R, 1 < p < ∞, and U (x) = xb(1 − x)A(1 + x)B. Then, there exists some constant C such that for every f and n if and only if −1 < Ap < p − 1, −1 < Bp < p − 1 and kU Snf kLp((−1,1),dµα) ≤ CkU f kLp((−1,1),dµα) −1 + p(cid:16)α + 1 2(cid:17)+ < bp + 2α + 1 < p − 1 + p(2α + 1) − p(cid:16)α + 1 2(cid:17)+ , where (α + 1 2 )+ = max{α + 1 2 , 0}. In the unweighted case (U = V = 1) the boundedness of the partial sum operators Sn, or in other words the convergence of the Fourier-Dunkl series, holds if and only if 4(α + 1) 2α + 3 < p < 4(α + 1) 2α + 1 in the case α ≥ −1/2, and for the whole range 1 < p < ∞ in the case −1 < α < −1/2. Remark. These conditions for the unweighted case are exactly the same as in the Fourier-Bessel case when the orthonormal functions are 21/2Jα+1(sn)−1Jα(snx)x−α and the orthogonality measure is x2α+1 dx on the interval (0, 1). Other variants of Bessel orthogonal systems exist in the literature, see [2, 3, 18]. For instance, one can take the functions 21/2Jα+1(sn)−1Jα(snx), which are orthonormal with respect to the measure x dx on the interval (0, 1). The conditions for the boundedness of these Fourier-Bessel series, as can be seen in [3], correspond to taking A = B = 0 and b = α − 2α+1 in our corollary. Another usual case is to take the functions (2x)1/2Jα+1(sn)−1Jα(snx), which are orthonormal with respect to the measure dx on (0, 1). Passing from one orthogonality to another consists basically in changing the weights. Then, from the weighted Lp boundedness of any of these systems we easily deduce a corresponding weighted Lp boundedness for any of the other systems. p In the case of the Fourier-Dunkl series on (−1, 1) we feel, however, that the natural setting is to start from Jα(z)z−α, since these functions, defined by (1), are holomorphic on C; in particular, they are well defined on the interval (−1, 1). 4 3. Auxiliary results We will need to control some basic operator in weighted Lp spaces on (−1, 1). For a function g : (0, 2) → R, the Calder´on operator is defined by Ag(x) = 1 xZ x 0 g(y) dy +Z 2 x g(y) y dy, that is, the sum of the Hardy operator and its adjoint. The weighted norm inequality kAgkLp((0,2),u) ≤ CkgkLp((0,2),v) holds for every g ∈ Lp((0, 2), v), provided that (u, v) ∈ Aδ (see [12, 13]). Let us consider now the operator J defined by p(0, 2) for some δ > 1, and δ = 1 is enough if u = v for x ∈ (−1, 1) and suitable functions f . With the notation f1(t) = f (1 − t), we have Jf (x) =Z 1 −1 f (y) 2 − x − y dy and a simple change of variables proves that the weighted norm inequality Jf (x) =(cid:12)(cid:12)(cid:12)(cid:12)Z 2 0 f (1 − t) 1 − x + t dt(cid:12)(cid:12)(cid:12)(cid:12) ≤ A(f1)(1 − x) holds for every f ∈ Lp((−1, 1), v), provided that (u, v) ∈ Aδ p(−1, 1) for some δ > 1 (or δ = 1 if u = v). The Hilbert transform on the interval (−1, 1) is defined as kJf kLp((−1,1),u) ≤ Ckf kLp((−1,1),v) Hg(x) =Z 1 −1 g(y) x − y dy. The above weighted norm inequality holds also for the Hilbert transform with the same Aδ (see [10, 13]). In both cases, the norm inequalities hold with a constant C depending only on the Aδ of the pair (u, v). p(−1, 1) condition p constant Our first objective is to obtain a suitable estimate for the kernel Kn(x, y). With this aim, we will use some well-known properties of Bessel (and related) functions, that can be found on [17]. For the Bessel functions we have the asymptotics if z < 1, arg(z) ≤ π; and Jν (z) = zν 2νΓ(ν + 1) + O(zν+2), Jν (z) =r 2 πzhcos(cid:16)z − νπ 2 − π 4(cid:17) + O(eIm(z)z−1)i , if z ≥ 1, arg(z) ≤ π − θ. The Hankel function of the first kind, denoted by H (1) ν , is defined as (5) (6) where Yν denotes the Weber function, given by H (1) ν (z) = Jν(z) + iYν(z), Yν(z) = Jν(z) cos νπ − J−ν(z) sin νπ , if ν /∈ Z, Yn(z) = lim ν→n Jν(z) cos νπ − J−ν(z) sin νπ 5 , if n ∈ Z. From these definitions, we have H (1) ν (z) = J−ν(z) − e−νπiJν(z) i sin νπ , if ν /∈ Z, H (1) n (z) = lim ν→n J−ν (z) − e−νπiJν (z) i sin νπ , if n ∈ Z. For the function H (1) ν , the asymptotic H (1) ν (z) =r 2 πz ei(z−νπ/2−π/4)[C + O(z−1)] (7) holds for z > 1, −π < arg(z) < 2π, with some constant C. As usual for the Lp convergence of orthogonal expansions, the results are consequences of suitable estimates for the kernel Kn(x, y). The next lemma contains an estimate for the difference between the kernel Kn(x, y) and an integral containing the product of two Eα functions. This integral can be evaluated using Lemma 1 in [5]. Next, to obtain the estimate we consider an appropriate function in the complex plane having poles in the points sj and integrate this function along a suitable path. Lemma 1. Let α > −1. Then, there exists some constant C > 0 such that for each n ≥ 1 and x, y ∈ (−1, 1), Kn(x, y) −Z Mn −Mn (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) where Mn = (sn + sn+1)/2. Kn(x, y) = 2α+1Γ(α + 2) + ≤ C(cid:18) xy−(α+1/2) 2 − x − y + 1(cid:19) , Eα(izx)Eα(izy) dµα(z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) nXj=1 2α+1Γ(α + 1) (xy)α Proof. Using elementary algebraic manipulations, the kernel Kn(x, y) can be written as Jα(sjx)Jα(sjy) + Jα+1(sjx)Jα+1(sjy) . (8) Jα(sj )2 Let us find a function whose residues at the points sj are the terms in the series, so that this series can be expressed as an integral. The identities −J ′ α+1(z)H (1) α+1(z) + Jα+1(z)(H (1) α+1)′(z) = 2i πz (see [19, p. 76]), and zJ ′ α+1(z) + (α + 1)Jα+1(z) = −zJα(z), give and −J ′ α+1(sj)H (1) α+1(sj) = 2i πsj for every j ∈ N. Then, J ′ α+1(sj) = −Jα(sj ) − 2i π xy1/2 Jα(sj x)Jα(sjy) + Jα+1(sjx)Jα+1(sjy) Jα(sj)2 = − 2i π xy1/2 Jα(sj x)Jα(sjy) + Jα+1(sjx)Jα+1(sjy) J ′ α+1(sj )2 = xy1/2sjH (1) α+1(sj) Jα(sjx)Jα(sjy) + Jα+1(sjx)Jα+1(sjy) J ′ α+1(sj ) = lim z→sj (z − sj)Hx,y(z) = Res(Hx,y, sj), 6 where we define Hx,y(z) = xy1/2 zH (1) α+1(z) Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy) Jα+1(z) (the factor xy1/2 is taken for convenience). The fact that Jν(−z) = eνπiJν(z) gives Res(Hx,y, sj) = Res(Hx,y, −sj). Since the definition of H (1) α+1(z) differs in case α ∈ Z, for the rest of the proof we will assume that α /∈ Z; the other case can be deduced by considering the limit. The function Hx,y(z) is analytic in C \ ((−∞, −Mn] ∪ [Mn, ∞) ∪ {±sj : j = 1, 2, . . . }). Moreover, the points ±sj are simple poles. So, we have ZS∪I(ε) Hx,y(z) dz = 0, (9) where I(ε) is the interval [−Mn, Mn] warped with upper half circles of radius ε centered in ±sj, with j = 1, . . . , n and S is the path of integration given by the interval Mn + i[0, ∞) in the direction of increasing imaginary part and the interval −Mn + i[0, ∞) in the opposite direction. The existence of the integral is clear for the path I(ε); for S this fact can be checked by using (5), (6) and (7). Indeed, on S we obtain that H(1) (cid:12)(cid:12)(cid:12) α+1(z) Jα+1(z)(cid:12)(cid:12)(cid:12) ≤ Ce−2 Im(z). Similarly, on S one has where for −1 < α < −1/2, and for α ≥ −1/2. Thus (cid:12)(cid:12)(cid:12)xy1/2zJα(zx)Jα(zy)(cid:12)(cid:12)(cid:12) ≤ CeIm(z)(x+y)hα hα x,y(z) = max{xzα+1/2, 1} max{yzα+1/2, 1} x,y(z) hα x,y(z) = 1 Hx,y(z) ≤ C(cid:0)hα x,y(z) + hα+1 x,y (z)(cid:1) e− Im(z)(2−x−y), and the integral on S is well defined. From the definition of Hx,y(z), we have ZI(ε) Hx,y(z) dz =ZI(ε) xy1/2zJ−α−1(z) i sin(α + 1)π · (10) dz Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy) Jα+1(z) − xy1/2 e−(α+1)πi i sin(α + 1)πZI(ε) z (Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy)) dz. The function in the first integral is odd, and the function in the second integral has no poles at the points sj. Then, the first integral equals the integral over the symmetric path −I(ε) = {z : −z ∈ I(ε)}. Putting z − sj = ε for the positively oriented circle, this gives ε→0ZI(ε) lim Hx,y(z) dz = lim ε→0 −1 2 Xsj <MnZz−sj =ε xy1/2zJ−α−1(z) i sin(α + 1)π · Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy) Jα+1(z) dz z (Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy)) dz − xy1/2 e−(α+1)πi i sin(α + 1)πZ Mn −Mn Res(Hx,y, sj) = −πi Xsj<Mn − xy1/2 e−(α+1)πi i sin(α + 1)π (1 − e2πiα)Z Mn 0 7 z (Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy)) dz = −4xy1/2 nXj=1 + 2xy1/2Z Mn 0 Jα(sjx)Jα(sj y) + Jα+1(sj x)Jα+1(sjy) Jα(sj)2 z (Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy)) dz. This, together with (9), gives nXj=1 Jα(sjx)Jα(sjy) + Jα+1(sjx)Jα+1(sjy) Jα(sj)2 = 1 4xy1/2ZS Hx,y(z) dz + 1 2Z Mn 0 z (Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy)) dz. Then, it follows from (8) that Kn(x, y) = 2α+1Γ(α + 2) + 2α−1Γ(α + 1) (xy)αxy1/2 ZS Hx,y(z) dz Now, it is easy to check the identity + 2αΓ(α + 1) (xy)α Z Mn 0 z(Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy)) dz. so that 0 (xy)α 2αΓ(α + 1) Z Mn Kn(x, y) −Z Mn −Mn (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) We conclude showing that 2α−1Γ(α + 1) xyα+1/2 −Mn ≤ 2α+1Γ(α + 2) + z(Jα(zx)Jα(zy) + Jα+1(zx)Jα+1(zy)) dz =Z Mn Eα(izx)Eα(izy) dµα(z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)ZS Hx,y(z) dz(cid:12)(cid:12)(cid:12)(cid:12) ≤ C(cid:18) (cid:12)(cid:12)(cid:12)(cid:12)ZS Hx,y(z) dz(cid:12)(cid:12)(cid:12)(cid:12) ≤ CZ ∞ e−t(2−x−y) dt = 2 − x − y 1 0 C + xyα+1/2(cid:19) , 2 − x − y . Eα(izx)Eα(izy) dµα(z), (cid:12)(cid:12)(cid:12)(cid:12)ZS Hx,y(z) dz(cid:12)(cid:12)(cid:12)(cid:12) . (11) for −1 < x, y < 1. For α ≥ −1/2, the bound (11) follows from (10). Indeed, in this case For −1 < α < −1/2, we have Hx,y(z) ≤ Cxyα+1/2e− Im(z)(2−x−y) if z ∈ S. With this inequality we obtain (11) as follows: From the previous lemma and the identity (see [5]) (cid:12)(cid:12)(cid:12)(cid:12)ZS Hx,y(z) dz(cid:12)(cid:12)(cid:12)(cid:12) ≤ Cxyα+1/2Z ∞ Z 1 Eα(ixz)Eα(iyz) dµα(z) = −1 0 e−t(2−x−y) dt = C xyα+1/2 2 − x − y ≤ C(cid:18)xyα+1/2 + 1 2 − x − y(cid:19) . 1 xIα+1(ix)Iα(iy) − yIα+1(iy)Iα(ix) 2α+1Γ(α + 2) x − y , which holds for α > −1, x, y ∈ C, and x 6= y, we obtain that Kn(x, y) − B(Mn, x, y) − B(Mn, y, x) ≤ C(cid:18) xy−(α+1/2) 2 − x − y 8 + 1(cid:19) (12) with B(Mn, x, y) = M 2(α+1) n 2α+1Γ(α + 2) xIα+1(iMnx)Iα(iMny) x − y or, by the definition of Iα and the fact that Jα(z) zα is even, B(Mn, x, y) = 2αΓ(α + 1) MnxJα+1(Mnx)Jα(Mny) xα+1yα(x − y) . 4. Proof of Theorem 1 We can split the partial sum operator Sn into three terms suitable to apply (12): Snf (x) =Z 1 −1 f (y)B(Mn, x, y) dµα(y) +Z 1 +Z 1 −1 f (y)hKn(x, y) − B(Mn, x, y) − B(Mn, y, x)i dµα(y) f (y)B(Mn, y, x) dµα(y) −1 =: T1,nf (x) + T2,nf (x) + T3,nf (x). (13) With this decomposition, the theorem will be proved if we see that kU Tj,nf kp Lp((−1,1),dµα) ≤ CkV f kp Lp((−1,1),dµα), j = 1, 2, 3, for a constant C independent of n and f . 4.1. The first term We have T1,nf (x) = = 1 2α+1Γ(α + 1)Z 1 −1 M 1/2 n xJα+1(Mnx) 2xα+1 f (y)B(Mn, x, y)y2α+1 dy Z 1 −1 f (y)M 1/2 n Jα(Mny)yα+1 x − y dy. According to (5) and (6) and the assumption that α ≥ −1/2, we have Jα(z) ≤ Cz−1/2, Jα+1(z) ≤ Cz−1/2, for every z > 0. Using these inequalities and the boundedness of the Hilbert transform under the Ap condition (2) gives kU T1,nf kp Lp((−1,1),dµα) −1 −1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = CZ 1 Z 1 −1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CZ 1 Z 1 ≤ CZ 1 −1(cid:12)(cid:12)(cid:12)f (x)M 1/2 ≤ CZ 1 −1 −1 dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 9 f (y)M 1/2 n Jα(Mny)yα+1 f (y)M 1/2 n Jα(Mny)yα+1 x − y x − y p p U (x)pM p/2 n Jα+1(Mnx)px2α+1−αp dx U (x)px(α+ 1 2 )(2−p) dx p n Jα(Mnx)xα+1(cid:12)(cid:12)(cid:12) f (x)pV (x)px2α+1 dx = CkV f kp Lp((−1,1),dµα). V (x)px(α+ 1 2 )(2−p) dx 4.2. The second term This term is given by T2,nf (x) = = 1 2α+1Γ(α + 1)Z 1 Z 1 n Jα(Mnx) M 1/2 2xα −1 −1 f (y)B(Mn, y, x)y2α+1 dy f (y)yM 1/2 n Jα+1(Mny)yα y − x dy and everything goes as with the first term. 4.3. The third term According to (12), T3,nf (x) ≤ Cx−(α+1/2)Z 1 −1 f (y)yα+1/2 2 − x − y dy + CZ 1 −1 f (y) y2α+1 dy so it is enough to have both and bounded by −1 −1(cid:12)(cid:12)(cid:12)(cid:12)Z 1 Z 1 (cid:12)(cid:12)(cid:12)(cid:12)Z 1 −1 p 2 − x − y f (y)yα+1/2 dy(cid:12)(cid:12)(cid:12)(cid:12) f (x) x2α+1 dx(cid:12)(cid:12)(cid:12)(cid:12) CZ 1 −1 U (x)px2α+1−p(α+1/2) dx pZ 1 −1 U (x)px2α+1 dx (14) (15) f (x)pV (x)px2α+1 dx. For the boundedness of (14) it suffices to impose but this is exactly (2). By duality, the boundedness of (15) is equivalent to (cid:16)U (x)px2α+1−p(α+1/2), V (x)px2α+1−p(α+1/2)(cid:17) ∈ Aδ (cid:18)Z 1 V (x)−p/(p−1)x2α+1 dx(cid:19)p−1 U (x)px2α+1 dx(cid:19)(cid:18)Z 1 −1 −1 p(−1, 1), < ∞. Now, it is easy to check that (cid:18)Z 1 −1 U (x)px2α+1 dx(cid:19)(cid:18)Z 1 ≤(cid:18)Z 1 V (x)−p/(p−1)x2α+1 dx(cid:19)p−1 2 )(2−p) dx(cid:19)(cid:18)Z 1 U (x)px(α+ 1 −1 −1 −1(cid:16)V (x)px(α+ 1 2 )(2−p)(cid:17)− 1 p−1 dx(cid:19)p−1 ≤ C, the last inequality following from the Ap condition (2). 10 5. Proof of Theorem 2 We begin with a simple lemma on Ap weights. Lemma 2. Let 1 < p < ∞, (u, v) ∈ Ap(−1, 1), (u1, v1) ∈ Ap(−1, 1). Let w, ζ be weights on (−1, 1) such that either w ≤ C(u + u1) and ζ ≥ C1(v + v1) or w−1 ≥ C(u−1 + u−1 1 ) and ζ−1 ≤ C1(v−1 + v−1 1 ) for some constants C, C1. Then (w, ζ) ∈ Ap(−1, 1) with a constant depending only on C, C1 and the Ap constants of (u, v) and (u1, v1). Proof. Assume that w ≤ C(u + u1) and ζ ≥ C1(v + v1). For any interval I ⊆ (−1, 1), (cid:18) 1 IZI ζ− 1 p−1(cid:19)p−1 ≤ 1 C1 min((cid:18) 1 IZI v− 1 p−1(cid:19)p−1 ,(cid:18) 1 IZI − 1 v 1 p−1 (cid:19)p−1) . Therefore, (cid:18) 1 IZI w(cid:19)(cid:18) 1 IZI ζ− 1 p−1(cid:19)p−1 ≤ C C1(cid:18) 1 IZI u(cid:19)(cid:18) 1 IZI v− 1 p−1(cid:19)p−1 + C C1(cid:18) 1 IZI u1(cid:19)(cid:18) 1 IZI − 1 v 1 p−1 (cid:19)p−1 . This proves that (w, ζ) ∈ Ap(−1, 1) with a constant depending on C, C1 and the Ap constants of (u, v) and (u1, v1). Assume now that w−1 ≥ C(u−1 + u−1 1 ) and ζ−1 ≤ C1(v−1 + v−1 1 ). Then 1 IZI w ≤ 1 C min(cid:26) 1 IZI u, 1 IZI u1(cid:27) for any interval I ⊆ (−1, 1). On the other hand, the inequality 1 2 (aλ + bλ) ≤ (a + b)λ ≤ 2λ(aλ + bλ), a, b ≥ 0, λ > 0 (16) (17) gives and ζ− 1 p−1 p−1 ≤ C 1 1 (cid:18) 1 IZI ζ− 1 p−1(cid:19)p−1 1 (cid:1) 1 (cid:0)v−1 + v−1 ≤ 2pC1(cid:18) 1 IZI 1 p−1 p−1 ≤ C 1 2 v− 1 p−1(cid:19)p−1 (cid:1), − 1 p−1 + v 1 p−1 1 p−1(cid:0)v− 1 + 2pC1(cid:18) 1 IZI − 1 v 1 p−1 (cid:19)p−1 . This, together with (16), proves that (w, ζ) ∈ Ap(−1, 1) with a constant depending on C, C1 and the Ap constants of (u, v) and (u1, v1). Now, we use the following estimate for the Bessel functions, which is a consequence of (5), (6) and −1 < α < −1/2: and z1/2Jα(z) ≤ C(1 + zα+1/2), z ≥ 0, z1/2Jα+1(z) ≤ C(1 + zα+1/2) −1 , z ≥ 0. In particular, there exists a constant C such that, for x ∈ (−1, 1) and n ≥ 0, we have M 1/2 n Jα(Mnx) ≤ Cx−1/2(1 + Mnxα+1/2) 11 and M 1/2 n Jα+1(Mnx) ≤ C x−1/2 1 + Mnxα+1/2 . Moreover, the inequality (17) gives 2α+1/2xα+1/2(x + M −1 n )−(α+1/2) ≤ 1 + Mnxα+1/2 ≤ 2xα+1/2(x + M −1 n )−(α+1/2) so that we get and M 1/2 n Jα(Mnx) ≤ Cxα(x + M −1 n )−(α+1/2) M 1/2 n Jα+1(Mnx) ≤ Cx−(α+1)(x + M −1 n )α+1/2. (18) (19) To handle these expressions, the following result will be useful: Lemma 3. Let 1 < p < ∞, a sequence {Mn} of positive numbers that tends to infinity, two nonnegative functions U and V defined on the interval (−1, 1), −1 < α < −1/2 and δ > 1 (δ = 1 if U = V ). If (3) and (4) are satisfied, then (cid:16)U (x)p(x + M −1 n ) (cid:16)U (x)p(x + M −1 n ) p(α+1/2) x(2α+1)(1−p), V (x)p(x + M −1 n ) p(α+1/2) −p(α+1/2) x2α+1, V (x)p(x + M −1 n ) −p(α+1/2) p(−1, 1), x(2α+1)(1−p)(cid:17) ∈ Aδ x2α+1(cid:17) ∈ Aδ p(−1, 1), (20) (21) "uniformly", i.e., with Aδ p constants independent of n. Proof. As a first step, let us observe that (3) and (4) imply (cid:16)U (x)px(2α+1)(1− 1 2 p), V (x)px(2α+1)(1− 1 p(−1, 1). 2 p)(cid:17) ∈ Aδ To prove this, just put U (x)px(2α+1)(1− 1 2 p) =hU (x)px(2α+1)(1−p)i1/2hU (x)px(2α+1)i1/2 (the same with V ) and check the Aδ p condition using the Cauchy-Schwarz inequality and (3), (4). Now, (17) yields hU (x)p(x + M −1 n ) p(α+ 1 2 ) x(2α+1)(1−p)i−δ 2hU (x)px(2α+1)(1− 1 1 ≥ 2 p)i−δ + 1 2hU (x)pM −p(α+ 1 2 ) n x(2α+1)(1−p)i−δ and hV (x)p(x + M −1 n ) p(α+ 1 2 ) ≤ 2−pδ(α+ 1 x(2α+1)(1−p)i−δ 2 )hV (x)px(2α+1)(1− 1 2 p)i−δ + 2−pδ(α+ 1 (cid:16)U (x)pM is the same constant of the pair −p(α+ 1 2 ) n x(2α+1)(1−p), V (x)pM −p(α+ 1 2 ) n Thus, Lemma 2 gives (20) with an Aδ p constant independent of n, since the Aδ p constant of the pair x(2α+1)(1−p)i−δ . −p(α+ 1 2 ) n 2 )hV (x)pM x(2α+1)(1−p)(cid:17) (cid:16)U (x)px(2α+1)(1−p), V (x)px(2α+1)(1−p)(cid:17) 12 i.e., it does not depend on n. The proof of (21) follows the same argument, since hU (x)p(x + M −1 n ) −p(α+ 1 2 ) ≤ 2−pδ(α+ 1 and hV (x)p(x + M −1 n ) −p(α+ 1 2 ) x2α+1iδ 2 )hU (x)px(2α+1)(1− 1 2 p)iδ x2α+1iδ ≥ 1 2 p)iδ 2hV (x)px(2α+1)(1− 1 2 )hU (x)pM 2hV (x)pM 1 + + 2−pδ(α+ 1 p(α+ 1 2 ) n x2α+1iδ x2α+1iδ . p(α+ 1 2 ) n We already have all the ingredients to start with the proof of Theorem 2. Let us take the same decom- position Snf = T1,nf + T2,n + T3,nf as in (13) in the previous section and consider each term separately. 5.1. The first term As in the proof of Theorem 1, by using (19) we have kU T1,nf kp n Jα(Mny)yα+1 U (x)pM p/2 n Jα+1(Mnx)px2α+1−αp dx U (x)p(x + M −1 n )p(α+1/2)x(2α+1)(1−p) dx. V (x)p(x + M −1 n )p(α+1/2)x(2α+1)(1−p) dx, f (x)pV (x)px2α+1 dx = CkV f kp Lp((−1,1),dµα). p dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) p dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Now, by the Ap condition (20), this is bounded by p dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) x − y −1 −1 x − y n Jα(Mny)yα+1 f (y)M 1/2 f (y)M 1/2 −1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Lp((−1,1),dµα) =Z 1 Z 1 −1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CZ 1 Z 1 CZ 1 n Jα(Mnx)xα+1(cid:12)(cid:12)(cid:12) −1(cid:12)(cid:12)(cid:12)f (x)M 1/2 CZ 1 −1 p which, by (18) is in turn bounded by 5.2. The second term The definition of T2,n and (18) yield Now, by the Ap condition (21), this is bounded by −1 −1 y − x y − x n Jα+1(Mny)yα Z 1 Z 1 f (y)yM 1/2 f (y)yM 1/2 −1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Lp((−1,1),dµα) =Z 1 −1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CZ 1 CZ 1 −1(cid:12)(cid:12)(cid:12)f (x)xM 1/2 n Jα+1(Mnx)xα(cid:12)(cid:12)(cid:12) CZ 1 −1 p dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 13 which, by (19) is in turn bounded by kU T2,nf kp n Jα+1(Mny)yα U (x)pM p/2 n Jα(Mnx)px2α+1−αp dx U (x)p(x + M −1 n )−p(α+1/2)x2α+1 dx. p V (x)p(x + M −1 n )−p(α+1/2)x2α+1 dx, f (x)pV (x)px2α+1 dx = CkV f kp Lp((−1,1),dµα). 5.3. The third term Taking limits when n → ∞ in (20) we get (2), so the proof of the boundedness of the third summand in Theorem 1 is still valid for Theorem 2. 6. Proof of Theorem 3 The following lemma is a small variant of a result proved in [8]. We give here a proof for the sake of completeness. Lemma 4. Let ν > −1. Let h be a Lebesgue measurable nonnegative function on [0, 1], {ρn} a positive sequence such that ρn = +∞ and 1 ≤ p < ∞. Then lim n→∞ n→∞Z 1 lim 0 ρ1/2 n Jν(ρnx)ph(x) dx ≥ MZ 1 0 h(x)x−p/2 dx (22) (in particular, that limit exists), where M is a positive constant independent of h and {ρn}. Proof. We can assume that h(x)xνp is integrable on (0, δ) for some δ ∈ (0, 1), since otherwise Z 1 0 ρ1/2 n Jν(ρnx)ph(x) dx = ∞ for each n, as follows from (5), and (22) is trivial. Assume also for the moment that h(x)x−p/2 is integrable on (0, 1). For each x ∈ (0, 1) and n, let us put ϕ(x, n) = (ρnx)1/2Jν (ρnx) −r 2 π cos(cid:16)ρnx − νπ 2 − π 4(cid:17) . The estimate (6) gives for each x ∈ (0, 1). Moreover, in case ρnx ≥ 1 the same estimate gives lim n→∞ ϕ(x, n) = 0 ϕ(x, n) ≤ C ρnx ≤ C with a constant C independent of n and x, while for ρnx ≤ 1 it follows from (5) that (23) (24) Without loss of generality we can assume that ρn ≥ 1. Then, (23) and (24) give ϕ(x, n) ≤ C(xν+1/2 + 1) with a constant C independent of x and n, so that, by the dominate convergence theorem, h(x)x−p/2 dx = 0. (25) lim 0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n→∞Z 1 Therefore, (ρnx)1/2Jν(ρnx) −r 2 n→∞Z 1 lim 0 lim λ→∞ 1 2πZ 2π 0 π π − νπ 2 ϕ(x, n) ≤ C(cid:16)(ρnx)ν+1/2 + 1(cid:17) . 4(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) cos(cid:16)ρnx − 2πZ π cos(cid:16)ρnx − 0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) r 2 n→∞Z 1 g(λt)f (t) dt =bg(0)bf (0) = 14 π 1 0 p νπ 2 p π − 4(cid:17)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2πZ π 1 0 g(t) dt f (t) dt ρ1/2 n Jν (ρnx)ph(x) dx = lim h(x)x−p/2 dx. (26) Now we use Fej´er's lemma: if f ∈ L1(0, 2π), and g is a continuous, 2π-periodic function, then right hand side of (26) gives where bf ,bg denote the Fourier transforms of f , g. After a change of variables, Fej´er's lemma applied to the n→∞Z 1 lim 0 ρ1/2 n Jν(ρnx)ph(x) dx = MZ 1 0 h(x)x−p/2 dx for some constant M , thus proving (22). Finally, in case h(x)x−p/2 is not integrable on (0, 1), let us take the sequence of increasing measurable sets Kj = {x ∈ (0, 1) : h(x)x−p/2 ≤ j}, j ∈ N, and define hj = h on Kj and hj = 0 on (0, 1) \ Kj. Applying (22) to each hj and then the monotone convergence theorem proves that n→∞Z 1 lim 0 ρ1/2 n Jν(ρnx)ph(x) dx = ∞, which is (22). We can now prove Theorem 3. Proof of Theorem 3. The first partial sum of the Fourier expansion is S0f = e0Z 1 −1 f e0 dµα = (α + 1)Z 1 −1 f (x)x2α+1 dx, so that the inequality kS0(f )U kLp((−1,1),dµα) ≤ Ckf V kLp((−1,1),dµα) gives, by duality, U (x)px2α+1 ∈ L1((−1, 1), dx), V (x)−p′ x2α+1 ∈ L1((−1, 1), dx). In fact, this is needed just to ensure that the partial sums of the Fourier expansions of all functions in Lp(V p dµα) are well defined and belong to Lp(U p dµα). These are the last two integrability conditions of Theorem 3. Now, if kSn(f )U kLp((−1,1),dµα) ≤ Ckf V kLp((−1,1),dµα) then the difference Snf − Sn−1f = enZ 1 = enZ 1 −1 −1 f en dµα + e−nZ 1 f en dµα + enZ 1 −1 −1 f e−n dµα f en dµα is bounded in the same way. Taking even and odd functions, and using that Re en is even and Im en is odd, gives kU Re enkLp((−1,1),dµα)kV −1 Re enkLp′ ((−1,1),dµα) ≤ C (27) and the same inequality with Im en. Recall that Re en(x) = 2α/2Γ(α + 1)1/2 snα Jα(sn) Jα(snx) (snx)α . Taking into account that Jν (x) is an even function (recall that Jα(z)/zα is taken as an even function) and Jα(sn) ≤ Cs−1/2 (this follows from (6)), Lemma 4 gives n lim inf −1(cid:12)(cid:12)(cid:12)(cid:12) n→∞ Z 1 1 Jα(sn) p Jν(snx)(cid:12)(cid:12)(cid:12)(cid:12) h(x) dx ≥ CZ 1 −1 15 h(x)x−p/2 dx for every measurable nonnegative function h. Therefore, lim inf n→∞ kU Re enkLp((−1,1),dµα) ≥ C(cid:18)Z 1 −1 U (x)px−pα− p p 2 +2α+1 dx(cid:19) 1 and the corresponding lower bound for lim inf n kV −1 Re enkLp′ ((−1,1),dµα) holds. The same bounds hold for Im en. Thus, (27) implies (cid:18)Z 1 −1 U (x)px−pα− p 2 +2α+1 dx(cid:19) 1 p(cid:18)Z 1 −1 V (x)−p′ x−p′α− p ′ 2 +2α+1 dx(cid:19) 1 p′ ≤ C or, in other words, the first two integrability conditions of Theorem 3. Take now f = U/(1 + V + U V ) and any measurable set E ⊆ (−1, 1). Then f ∈ L2(dµα) by Holder's in- equality, the obvious inequality f ≤ U V −1 and the integrability conditions U ∈ Lp(dµα), V −1 ∈ Lp′ (dµα), already proved. Since {ej}j∈Z is a complete orthonormal system in L2((−1, 1), dµα), we have Sn(f χE) → f χE in the L2(dµα) norm. Therefore, there exists some subsequence Snj (f χE) converging to f χE almost everywhere. Fatou's lemma then gives Z 1 −1 f χEpU p dµα ≤ lim inf j→∞ Z 1 −1 Snj (f χE)pU p dµα. Under the hypothesis of Theorem 3, each of the integrals on the right hand side is bounded by CpZ 1 −1 f χEpV p dµα (observe, by the way, that f V ∈ Lp(dµα), since f V ≤ 1). Thus, Z 1 −1 f χEpU p dµα ≤ CpZ 1 −1 f χEpV p dµα for every measurable set E ⊆ (−1, 1). This gives f U ≤ Cf V almost everywhere, and U ≤ CV . Acknowledgment We thank the referee for his valuable suggestions, which helped us to make the paper more readable. References [1] N. B. Andersen and M. de Jeu, Elementary proofs of Paley-Wiener theorems for the Dunkl transform on the real line, Int. Math. Res. Not. 30 (2005), 1817 -- 1831. [2] A. Benedek and R. Panzone, On mean convergence of Fourier-Bessel series of negative order, Studies in Appl. Math. 50 (1971), 281 -- 292. [3] A. Benedek and R. Panzone, Mean convergence of series of Bessel functions, Rev. Un. Mat. Argentina 26 (1972/73), 42 -- 61. [4] J. J. Betancor, ´O. Ciaurri and J. L. Varona, The multiplier of the interval [−1, 1] for the Dunkl transform on the real line, J. Funct. Anal. 242 (2007), 327 -- 336. [5] ´O. Ciaurri and J. L. Varona, A Whittaker-Shannon-Kotel'nikov sampling theorem related to the Dunkl transform, Proc. Amer. Math. Soc. 135 (2007), 2939 -- 2947. [6] C. F. Dunkl, Integral kernels with reflections group invariance, Canad. J. Math. 43 (1991), 1213 -- 1227. [7] M. F. E. de Jeu, The Dunkl transform, Invent. Math. 113 (1993), 147 -- 162. [8] J. J. Guadalupe, M. P´erez, F. J. Ruiz and J. L. Varona, Two notes on convergence and divergence a.e. of Fourier series with respect to some orthogonal systems, Proc. Amer. Math. Soc. 116 (1992), 457 -- 464. [9] J. J. Guadalupe, M. P´erez, F. J. Ruiz and J. L. Varona, Mean and weak convergence of Fourier-Bessel series, J. Math. Anal. Appl. 173 (1993), 370 -- 389. [10] R. Hunt, B. Muckenhoupt and R. Wheeden, Weighted norm inequalities for the conjugate function and Hilbert transform, Trans. Amer. Math. Soc. 176 (1973), 227 -- 251. 16 [11] H. Mejjaoli and K. Trim`eche, A variant of Cowling-Price's theorem for the Dunkl transform on R, J. Math. Anal. Appl. 345 (2008), 593 -- 606. [12] B. Muckenhoupt, Weighted norm inequalities for the Hardy maximal function, Trans. Amer. Math. Soc. 165 (1972), 207 -- 226. [13] C. J. Neugebauer, Inserting Ap weights, Proc. Amer. Math. Soc. 87 (1983), 644 -- 648. [14] A. Nowak and K. Stempak, Relating transplantation and multipliers for Dunkl and Hankel transforms, Math. Nachr. 281 (2008), 1604 -- 1611. [15] A. Nowak and K. Stempak, Riesz transforms for the Dunkl harmonic oscillator, Math. Z. 262 (2009), 539 -- 556. [16] M. Rosler and M. Voit, An uncertainty principle for Hankel transforms, Proc. Amer. Math. Soc. 127 (1999), 183 -- 194. [17] G. N. Watson, A treatise on the theory of Bessel functions, Cambridge Univ. Press, Cambridge, 1944. [18] G. M. Wing, The mean convergence of orthogonal series, Amer. J. Math. 72 (1950), 792 -- 808. [19] E. T. Whittaker and G. N. Watson, A Course of Modern Analysis, Cambridge Univ. Press, 1952. 17
1602.07077
2
1602
2016-07-08T05:07:49
On the Carleman ultradifferentiable vectors of a scalar type spectral operator
[ "math.FA" ]
A description of the Carleman classes of vectors, in particular the Gevrey classes, of a scalar type spectral operator in a reflexive complex Banach space is shown to remain true without the reflexivity requirement. A similar nature description of the entire vectors of exponential type, known for a normal operator in a complex Hilbert space, is generalized to the case of a scalar type spectral operator in a complex Banach space.
math.FA
math
Methods of Functional Analysis and Topology Vol. 21 (2015), no. 4, pp. 361 -- 369 ON THE CARLEMAN ULTRADIFFERENTIABLE VECTORS OF A SCALAR TYPE SPECTRAL OPERATOR 6 1 0 2 l u J 8 ] MARAT V. MARKIN To Academician Yu. M. Berezansky in honor of his 90th jubilee Abstract. A description of the Carleman classes of vectors, in particular the Gevrey classes, of a scalar type spectral operator in a reflexive complex Banach space is shown to remain true without the reflexivity requirement. A similar nature description of the entire vectors of exponential type, known for a normal operator in a complex Hilbert space, is generalized to the case of a scalar type spectral operator in a complex Banach space. . A F h t a m [ 2 v 7 7 0 7 0 . 2 0 6 1 : v i X r a Never cut what you can untie. Joseph Joubert 1. Introduction The description of the Carleman classes of ultradifferentiable vectors, in particular the Gevrey classes, of a normal operator in a complex Hilbert space in terms of its spectral measure established in [10] (see also [12] and [11]) is generalized in [15, Theorem 3.1] to the case of a scalar type spectral operator in a complex reflexive Banach space. Here, the reflexivity requirement is shown to be superfluous and a similar nature description of the entire vectors of exponential type, known for a normal operator in a complex Hilbert space (see, e.g., [12]), is generalized to the case of a scalar type spectral operator in a complex Banach space. 2. Preliminaries For the reader's convenience, we shall outline in this section certain essential prelimi- naries. 2.1. Scalar type spectral operators. Henceforth, unless specified otherwise, A is sup- posed to be a scalar type spectral operator in a complex Banach space (X, k · k) and EA(·) to be its spectral measure (the resolution of the identity), the operator's spectrum σ(A) being the support for the latter [1, 4]. In a complex Hilbert space, the scalar type spectral operators are precisely those similar to the normal ones [23]. A scalar type spectral operator in complex Banach space has an operational calculus analogous to that of a normal operator in a complex Hilbert space [1, 3, 4]. To any Borel measurable function F : C → C (or F : σ(A) → C, C is the complex plane), there corresponds a scalar type spectral operator F (A) := ZC F (λ) dEA(λ) = Zσ(A) F (λ) dEA(λ) 2010 Mathematics Subject Classification. Primary 47B40; Secondary 47B15. Key words and phrases. Scalar type spectral operator, normal operator, Carleman classes of vectors. 361 362 MARAT V. MARKIN defined as follows: F (A)f := lim n→∞ Fn(A)f, f ∈ D(F (A)), (D(·) is the domain of an operator), where D(F (A)) := nf ∈ X(cid:12)(cid:12) lim n→∞ Fn(A)f existso Fn(·) := F (·)χ{λ∈σ(A) F (λ)≤n}(·), n ∈ N, (χδ(·) is the characteristic function of a set δ ⊆ C, N := {1, 2, 3, . . . } is the set of natural numbers) and Fn(A) := Zσ(A) Fn(λ) dEA(λ), n ∈ N, are bounded scalar type spectral operators on X defined in the same manner as for a normal operator (see, e.g., [3, 21]). In particular, (2.1) An = ZC λn dEA(λ) = Zσ(A) λn dEA(λ), n ∈ Z+, (Z+ := {0, 1, 2, . . . } is the set of nonnegative integers). The properties of the spectral measure EA(·) and the operational calculus, exhaustively delineated in [1, 4], underly the entire subsequent discourse. Here, we shall outline a few facts of particular importance. Due to its strong countable additivity, the spectral measure EA(·) is bounded [2, 4], i.e., there is such an M > 0 that, for any Borel set δ ⊆ C, (2.2) kEA(δ)k ≤ M. The notation k · k has been recycled here to designate the norm in the space L(X) of all bounded linear operators on X. We shall adhere to this rather common economy of symbols in what follows adopting the same notation for the norm in the dual space X ∗ as well. For any f ∈ X and g∗ ∈ X ∗, the total variation v(f, g∗, ·) of the complex-valued Borel measure hEA(·)f, g∗i (h·, ·i is the pairing between the space X and its dual X ∗) is a finite positive Borel measure with (2.3) v(f, g∗, C) = v(f, g∗, σ(A)) ≤ 4M kf kkg∗k (see, e.g., [15]). Also (Ibid.), F : C → C (or F : σ(A) → C) being an arbitrary Borel measurable function, for any f ∈ D(F (A)), g∗ ∈ X ∗, and an arbitrary Borel set σ ⊆ C, (2.4) In particular, Zσ F (λ) dv(f, g∗, λ) ≤ 4M kEA(σ)F (A)f kkg∗k. (2.5) ZC F (λ) dv(f, g∗, λ) = Zσ(A) F (λ) dv(f, g∗, λ) ≤ 4M kF (A)f kkg∗k. The constant M > 0 in (2.3) -- (2.5) is from (2.2). The following statement allowing to characterize the domains of the Borel measurable functions of a scalar type spectral operator in terms of positive Borel measures is also fundamental for our discussion. Proposition 2.1. ([14, Proposition 3.1]) . Let A be a scalar type spectral operator in a complex Banach space (X, k · k) and F : C → C (or F : σ(A) → C) be Borel measurable function. Then f ∈ D(F (A)) iff (i) For any g∗ ∈ X ∗, Zσ(A) F (λ) dv(f, g∗, λ) < ∞. The inclusions (2.6) are obvious. C(mn)(A) ⊆ C{mn}(A) ⊆ C∞(A) ⊆ X ON THE CARLEMAN ULTRADIFFERENTIABLE VECTORS 363 (ii) {g∗∈X ∗ kg∗k=1}Z{λ∈σ(A) F (λ)>n} sup F (λ) dv(f, g∗, λ) → 0 as n → ∞. Subsequently, the frequent terms "spectral measure" and "operational calculus" will be abbreviated to s.m. and o.c., respectively. 2.2. The Carleman classes of vectors. Let A be a densely defined closed linear operator in a complex Banach space (X, k · k) and {mn}∞ n=0 be a sequence of positive numbers and ∞ The subspaces of C∞(A) C∞(A) := D(An). \n=0 C{mn}(A) := (cid:8)f ∈ C∞(A)(cid:12)(cid:12)∃α > 0 ∃c > 0 : kAnf k ≤ cαnmn, n ∈ Z+(cid:9) , C(mn)(A) := (cid:8)f ∈ C∞(A)(cid:12)(cid:12)∀α > 0 ∃c > 0 : kAnf k ≤ cαnmn, n ∈ Z+(cid:9) are called the Carleman classes of ultradifferentiable vectors of the operator A corres- ponding to the sequence {mn}∞ n=0 of Roumieu and Beurling type, respectively. If two sequences of positive numbers (cid:8)mn(cid:9)∞ ∀γ > 0 ∃c = c(γ) > 0 : m′ n=0 and (cid:8)m′ n ≤ cγnmn, n ∈ Z+, n(cid:9)∞ n=0 are related as follows: we also have the inclusion (2.7) C{m′ n}(A) ⊆ C(mn)(A), the sequences being subject to the condition ∃γ1, γ2 > 0, ∃c1, c2 > 0 : c1γn 1 mn ≤ m′ n ≤ c2γn 2 mn, n ∈ Z+, their corresponding Carleman classes coincide (2.8) C{mn}(A) = C{m′ n}(A), C(mn)(A) = C(m′ n)(A). Considering Stirling's formula and the latter, E {β}(A) := C{[n!]β }(A) = C{nβn}(A), E (β)(A) := C([n!]β )(A) = C(nβn)(A) with β ≥ 0 are the well-known Gevrey classes of strongly ultradifferentiable vectors of A of order β of Roumieu and Beurling type, respectively (see, e.g., [10, 11, 12]). In particular, E {1}(A) and E (1)(A) are the well-known classes of analytic and entire vectors of A, respectively [6, 19]; E {0}(A) and E (0)(A) (i.e., the classes C{1}(A) and C(1)(A) corresponding to the sequence mn ≡ 1) are the classes of entire vectors of exponential and minimal exponential type, respectively (see, e.g., [22, 12]). If the sequence of positive numbers {mn}∞ n=0 satisfies the condition (WGR) ∀α > 0 ∃c = c(α) > 0 : cαn ≤ mn, n ∈ Z+, (2.9) the scalar function (2.10) T (λ) := m0 ∞ Xn=0 λn mn , λ ≥ 0, (00 := 1) first introduced by S. Mandelbrojt [13], is well-defined (cf. [12]). The function is contin- uous, strictly increasing, and T (0) = 1. 364 MARAT V. MARKIN As is shown in [10] (see also [12] and [11]), the sequence {mn}∞ n=0 satisfying the condition (WGR), for a normal operator A in a complex Hilbert space X, the equalities (2.11) C{mn}(A) = [t>0 C(mn)(A) = \t>0 D(T (tA)), D(T (tA)) are true, the normal operators T (tA), t > 0, defined in the sense of the operational calculus for a normal operator (see, e.g., [3, 21]) and the function T (·) being replaceable with any nonnegative, continuous, and increasing on [0, ∞) function F (·) satisfying (2.12) c1F (γ1λ) ≤ T (λ) ≤ c2F (γ2λ), λ ≥ R, with some γ1, γ2, c1, c2 > 0 and R ≥ 0, in particular, with S(λ) := m0 sup n≥0 λn mn , λ ≥ 0, or P (λ) := m0(cid:20) ∞ Xn=0 λ2n m2 n(cid:21)1/2 , λ ≥ 0, (cf. [12]). In [15, Theorem 3.1], the above is generalized to the case of a scalar type spectral operator A in a reflexive complex Banach space X. The reflexivity requirement dropped, proved were the inclusions (2.13) C{mn}(A) ⊇ [t>0 C(mn)(A) ⊇ \t>0 D(T (tA)), D(T (tA)) only, which is a deficiency for statements like [16, Theorem 5.1] and [18, Theorem 3.2]. 3. The Carleman classes of a scalar type spectral operator Theorem 3.1. Let (cid:8)mn(cid:9)∞ n=0 be a sequence of positive numbers satisfying the condition (WGR) (see (2.9)). Then, for a scalar type spectral operator A in a complex Banach space (X, k · k), equalities (2.11) are true, the scalar type spectral operators T (tA), t > 0, defined in the sense of the operational calculus for a scalar type spectral operator and the function T (·) being replaceable with any nonnegative, continuous, and increasing on [0, ∞) function F (·) satisfying (2.12). Proof. We are only to prove the inclusions inverse to (2.13), the rest, including the latter, having been proved in [15, Theorem 3.1]. Consider an arbitrary vector f ∈ C{mn}(A) (f ∈ C(mn)(A)). Then necessarily, f ∈ C∞(A) and for a certain α > 0 (an arbitrary α > 0), there is a c > 0 such that (3.14) (see Preliminaries). For any g∗ ∈ X ∗, kAnf k ≤ cαnmn, n ∈ Z+, (3.15) Zσ(A) 2α λ(cid:1) dv(f, g∗, λ) = Zσ(A) T (cid:0) 1 ∞ Xn=0 λn 2nαnmn dv(f, g∗, λ) by the Monotone Convergence Theorem; = ∞ Xn=0Zσ(A) λn 2nαnmn dv(f, g∗, λ) = ∞ Xn=0 1 2nαnmn Zσ(A) λn dv(f, g∗, λ) by (2.5) and (2.1); ON THE CARLEMAN ULTRADIFFERENTIABLE VECTORS 365 ≤ ∞ Xn=0 1 2nαnmn 4M kAnf kkg∗k by (3.14); ≤ 4M c ∞ Xn=0 1 2n kg∗k = 8M ckg∗k < ∞. For an arbitrary ε > 0, one can fix an N ∈ N such that, (3.16) M 2c 2N −2 < ε/2. Due to the strong continuity of the s.m., for any n ∈ N, (cid:13)(cid:13)EA(cid:0)(cid:8)λ ∈ σ(A)(cid:12)(cid:12)T ( 1 2α λ) > k(cid:9)(cid:1) Anf(cid:13)(cid:13) → 0 Hence, there is a K ∈ N such that as k → ∞. (3.17) N Xn=0 1 2nαnmn 4M (cid:13)(cid:13)EA(cid:0)(cid:8)λ ∈ σ(A)(cid:12)(cid:12)T ( 1 2α λ) > k(cid:9)(cid:1) Anf(cid:13)(cid:13) < ε/2 whenever k ≥ K. Similarly to (3.15), for k ≥ K, we have sup {g∗∈X ∗ kg∗k=1}Znλ∈σ(A)(cid:12)(cid:12)T ( 1 ∞ = sup {g∗∈X ∗ kg∗k=1} 1 2α λ)>ko T (cid:0) 1 2nαnmn Znλ∈σ(A)(cid:12)(cid:12)T ( 1 Xn=0 ∞ 2α λ(cid:1) dv(f, g∗, λ) 2α λ)>ko λn dv(f, g∗, λ) by (2.4) and (2.1); 1 ≤ ≤ + sup sup 2nαnmn {g∗∈X ∗ kg∗k=1} Xn=0 {g∗∈X ∗ kg∗k=1}(cid:20) N Xn=0 Xn=N +1 4M (cid:13)(cid:13)EA(cid:0)(cid:8)λ ∈ σ(A)(cid:12)(cid:12)T ( 1 4M (cid:13)(cid:13)EA(cid:0)(cid:8)λ ∈ σ(A)(cid:12)(cid:12)T ( 1 4M (cid:13)(cid:13)EA(cid:0)(cid:8)λ ∈ σ(A)(cid:12)(cid:12)T ( 1 2nαnmn 2nαnmn ∞ 1 1 2α λ) > k(cid:9)(cid:1)(cid:13)(cid:13) kAnf k kg∗k(cid:21) 2α λ) > k(cid:9)(cid:1) Anf(cid:13)(cid:13) kg∗k 2α λ) > k(cid:9)(cid:1) Anf(cid:13)(cid:13) kg∗k ≤ {g∗∈X ∗ kg∗k=1}(cid:20) N Xn=0 sup 1 2nαnmn 4M (cid:13)(cid:13)EA(cid:0)(cid:8)λ ∈ σ(A)(cid:12)(cid:12)T ( 1 2n kg∗k(cid:21) + 4M 2c ∞ 1 Xn=N +1 2α λ) > k(cid:9)(cid:1) Anf(cid:13)(cid:13) + ≤ N Xn=0 1 2nαnmn 4M (cid:13)(cid:13)EA(cid:0)(cid:8)λ ∈ σ(A)(cid:12)(cid:12)T ( 1 by (2.2) and (3.14); 2α λ) > k(cid:9)(cid:1) Anf(cid:13)(cid:13) kg∗k M 2c 2N −2 by (3.16) and (3.17); < ε/2 + ε/2 = ε 366 MARAT V. MARKIN and we conclude that (3.18) sup {g∗∈X ∗ kg∗k=1}Znλ∈σ(A)(cid:12)(cid:12)T ( 1 2α λ)>ko T (cid:0) 1 2α λ(cid:1) dv(f, g∗, λ) → 0 as k → ∞. By Proposition 2.1, (3.15) and (3.18) imply f ∈ D(T ( 1 2α A)). Considering that for f ∈ C{mn}(A), α > 0 is fixed and for f ∈ C(mn)(A), α > 0 is arbitrary, we infer that in the former case and in the latter. f ∈ [t>0 f ∈ \t>0 D(T (tA)) D(T (tA)) Since f ∈ C{mn}(A) (f ∈ C(mn)(A)) is arbitrary, we have proved the inclusions C{mn}(A) ⊆ [t>0 C(mn)(A) ⊆ \t>0 D(T (tA)), D(T (tA)), which along with their inverses (2.13) imply equalities (2.11) to be true. (cid:3) 4. The Gevrey classes The sequence mn := [n!]β (mn := nβn) with β > 0 satisfying the condition (WGR) and the corresponding function T (·) being replaceable with F (λ) = eλ1/β , λ ≥ 0, (see [15] for details, cf. also [12]), in [15, Corollary 4.1] describing the Gevrey classes of vectors of a scalar type spectral operator in a reflexive complex Banach space, the reflexivity requirement can be dropped as well and we have the following Corollary 4.1. Let β > 0. Then, for a A scalar type spectral operator in a complex Banach space (X, k · k), E {β}(A) = [t>0 E (β)(A) = \t>0 D(etA1/β ), D(etA1/β ). Corollary 4.1 generalizes the corresponding result of [10] (see also [12, 11]) for a normal operator A in a complex Hilbert space and, for β = 1, gives a description of the analytic and entire vectors of a scalar type spectral operator A in a complex Banach space. 5. The entire vectors of exponential type Observe that the sequence mn ≡ 1 generating the entire vectors of exponential type does not meet the condition (WGR) (see (2.9)) and thus, this case falls outside the realm of Theorem 3.1. As is known (cf., e.g., [12, 7]), for a normal operator A in a complex Hilbert space X, E {0}(A) = [α>0 EA(∆α)X and E (0)(A) = \α>0 EA(∆α)X = EA({0})X = ker A := (cid:8)f ∈ X(cid:12)(cid:12)Af = 0(cid:9) ON THE CARLEMAN ULTRADIFFERENTIABLE VECTORS 367 (EA(·) is the spectral measure of A) with We are to generalize the above to the case of a scalar type spectral operator A in a ∆α := (cid:8)λ ∈ C(cid:12)(cid:12)λ ≤ α(cid:9) , α > 0. complex Banach space X. Theorem 5.1. For a scalar type spectral operator A in a complex Banach space (X, k·k), (i) E {0}(A) = Sα>0 EA(∆α)X, (ii) E (0)(A) = Tα>0 EA(∆α)X = EA({0})X = ker A := (cid:8)f ∈ X(cid:12)(cid:12)Af = 0(cid:9), where ∆α := (cid:8)λ ∈ C(cid:12)(cid:12)λ ≤ α(cid:9), α > 0. Proof. Let f ∈ Sα>0 EA(∆α)X (f ∈ Tα>0 EA(∆α)X), i.e., f = EA(cid:0)(cid:8)λ ∈ C (cid:12)(cid:12) λ ≤ α(cid:9)(cid:1) f for some (any) α > 0. Then, by the properties of the o.c., f ∈ C∞(A). Furthermore, as follows form the Hahn-Banach Theorem and the properties of the o.c. (in particular, (2.1)), kAnf k = (cid:13)(cid:13)(cid:13)(cid:13) λn dEA(λ)f(cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) ZC Z{λ∈Cλ≤α} g∗∈X ∗, kg∗k=1(cid:12)(cid:12)(cid:12)(cid:12) (cid:28)Z{λ∈Cλ≤α} g∗∈X ∗, kg∗k=1(cid:12)(cid:12)(cid:12)(cid:12) Z{λ∈Cλ≤α} g∗∈X ∗, kg∗k=1Z{λ∈Cλ≤α} λn dEA(λ)f(cid:13)(cid:13)(cid:13)(cid:13) λn dEA(λ)f, g∗(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) λn dhEA(λ)f, g∗i(cid:12)(cid:12)(cid:12)(cid:12) λn dv(f, g∗, λ) sup = = ≤ sup sup ≤ sup αnv(f, g∗, {λ ∈ Cλ ≤ α}) g∗∈X ∗, kg∗k=1 ≤ sup 4M kf kkg∗kαn ≤ 4M [kf k + 1] αn, α > 0, g∗∈X ∗, kg∗k=1 by (2.3); which implies that f ∈ E {0}(A) (f ∈ E (0)(A)). Conversely, for an arbitrary f ∈ E {0}(A) (f ∈ E (0)(A)), kAnf k ≤ cαn, n ∈ Z+, with some (any) α > 0 and some c > 0. Then, for any γ > α and g∗ ∈ X ∗, we have γnv(f, g∗, {λ ∈ Cλ ≥ γ}) ≤ Z{λ∈Cλ≥γ} ≤ ZC λn dv(f, g∗, λ) λn dv(f, g∗, λ) by (2.5); ≤ 4M kAnf kkg∗k ≤ 4M ckg∗kαn, n ∈ Z+. Therefore, v(f, g∗, {λ ∈ Cλ ≥ γ}) ≤ 4M ckg∗k(cid:18) α γ(cid:19)n , n ∈ Z+. Considering that α/γ < 1 and passing to the limit as n → ∞, we conclude that v(f, g∗, {λ ∈ Cλ ≥ γ}) = 0, g∗ ∈ X ∗, 368 MARAT V. MARKIN and the more so hEA({λ ∈ Cλ ≥ γ})f, g∗i = 0, Whence, as follows from the Hahn-Banach Theorem, g∗ ∈ X ∗. EA({λ ∈ Cλ ≥ γ})f = 0, which, considering that γ > α is arbitrary, by the strong continuity of the s.m., implies that Hence, by the additivity of the s.m., EA({λ ∈ Cλ > α})f = 0. f = EA({λ ∈ Cλ ≤ α})f + EA({λ ∈ Cλ > α})f = EA({λ ∈ Cλ ≤ α})f, which implies that f ∈ Sα>0 EA(∆α)X (f ∈ Tα>0 EA(∆α)X). An immediate implication of Theorem 5.1 is the following generalization of the well- known result on the denseness of exponential type vectors of a normal operator in a complex Hilbert space (see, e.g., [12]), which readily follows by the strong continuity of the s.m. and joins a number of similar results of interest for approximation and qualitative theories (see [22, 12, 7, 8, 9]). (cid:3) Corollary 5.1. For a scalar-type spectral operator A in a complex Banach space (X, k·k), (· is the closure of a set in the strong topology of X). E {0}(A) = X Hence, for any positive sequence {mn}∞ n=0 satisfying the condition (WGR), in par- ticular, for mn = [n!]β with β > 0, due to inclusion (2.7) with m′ n ≡ 1, which implies E {0}(A) ⊆ C(mn)(A), C(mn)(A) = X. 6. Final remarks Observe that, for a normal operator in a complex Hilbert, equalities (2.11) have not only the set-theoretic but also a topological meaning [10] (see also [11, 12]). By analogy, this also appears to be true for a scalar type spectral operator in a complex Banach space, although the idea was entertained by the author neither in [15] nor here. For a normal operator in a complex Hilbert space, Theorems 5.1 and 3.1 can be considered as generalizations of Paley-Wiener Theorems relating the smoothness of a square-integrable on the real axis R function f (·) to the decay of its Fourier transform f (·) as x → ±∞ [20], which precisely corresponds to the case of the self-adjoint differential operator A = i (i is the imaginary unit) in the complex Hilbert space L2(R) [12]. Observe that, in Lp(R) with 1 ≤ p < ∞, p 6= 2, the same operator fails to be spectral d dx [5] (the domain of A = i W 1 d dx in X = Lp(R), 1 ≤ p < ∞, is understood to be the subspace p (R) := (cid:8)f ∈ Lp(R)(cid:12)(cid:12)f (·) is absolutely continuous on R and f ′ ∈ Lp(R)(cid:9)). Theorem 3.1 entirely substantiates the proof of the "only if" part of [16, Theorem 5.1], where inclusions (2.13) turn out to be insufficient, and of [18, Theorem 3.2]. It appears to be fundamental for qualitative results of this nature (cf. [17]). Acknowledgments. The author's utmost appreciation is to Drs. Miroslav L. and Valentina I. Gorbachuk, whose work and life have become a perpetual source of in- spiration, ideas, and goodness for him. ON THE CARLEMAN ULTRADIFFERENTIABLE VECTORS 369 References 1. N. Dunford, A survey of the theory of spectral operators, Bull. Amer. Math. Soc. 64 (1958), 217 -- 274. 2. N. Dunford and J. T. Schwartz with the assistance of W. G. Bade and R. G. Bartle, Linear Operators. Part I : General Theory, Interscience Publishers, New York, 1958. 3. 4. , Linear Operators. Part II : Spectral Theory. Self Adjoint Operators in Hilbert Space, Interscience Publishers, New York, 1963. , Linear Operators. Part III : Spectral Operators, Interscience Publishers, New York, 1971. 5. R. Farwig and E. Marschall, On the type of spectral operators and the nonspectrality of several differential operators on Lp, Integral Equations Operator Theory 4 (1981), no. 2, 206 -- 214. 6. R. Goodman, Analytic and entire vectors for representations of Lie groups, Trans. Amer. Math. Soc. 143 (1969), 55 -- 76. 7. M. L. Gorbachuk and V. I. Gorbachuk, On the approximation of smooth vectors of a closed operator by entire vectors of exponential type, Ukrain. Mat. Zh. 47 (1995), no. 5, 616 -- 628. (Ukrainian); English transl. Ukrainian Math. J. 47 (1995), no. 5, 713 -- 726. 8. 9. , On the well-posed solvability in some classes of entire functions of the Cauchy problem for differential equations in a Banach space, Methods Funct. Anal. Topology 11 (2005), no. 2, 113 -- 125. , On completeness of the set of root vectors for unbounded operators, Methods Funct. Anal. Topology 12 (2006), no. 4, 353 -- 362. 10. V. I. Gorbachuk, Spaces of infinitely differentiable vectors of a nonnegative self-adjoint operator, Ukrain. Mat. Zh. 35 (1983), no. 5, 617 -- 621. (Russian); English transl. Ukrainian Math. J. 35 (1983), no. 5, 531 -- 534. 11. V. I. Gorbachuk and M. L. Gorbachuk, Boundary Value Problems for Operator Differential Equations, Kluwer Academic Publishers, Dordrecht -- Boston -- London, 1991. (Russian edition: Naukova Dumka, Kiev, 1984) 12. V. I. Gorbachuk and A. V. Knyazyuk, Boundary values of solutions of operator-differential equations, Russian Math. Surveys 44 (1989), no. 3, 67 -- 111. 13. S. Mandelbrojt, Series de Fourier et Classes Quasi-Analytiques de Fonctions, Gauthier-Villars, Paris, 1935. 14. M. V. Markin, On an abstract evolution equation with a spectral operator of scalar type, Int. J. Math. Math. Sci. 32 (2002), no. 9, 555 -- 563. 15. 16. 17. 18. , On the Carleman classes of vectors of a scalar type spectral operator, Ibid. 2004 (2004), no. 60, 3219 -- 3235. , On scalar type spectral operators and Carleman ultradifferentiable C0-semigroups, Ukrain. Mat. Zh. 60 (2008), no. 9, 1215 -- 1233; English transl. Ukrainian Math. J. 60 (2008), no. 9, 1418 -- 1436. , On the Carleman ultradifferentiability of weak solutions of an abstract evolution equa- tion, Modern Analysis and Applications, Oper. Theory Adv. Appl., vol. 191, pp. 407 -- 443, Birkhauser Verlag, Basel, 2009. , On the generation of Beurling type Carleman ultradifferentiable C0-semigroups by scalar type spectral operators, Methods Funct. Anal. Topology (to appear). 19. E. Nelson, Analytic vectors, Ann. of Math. (2) 70 (1959), no. 3, 572 -- 615. 20. R. E. A. C. Paley and N. Wiener, Fourier Transforms in the Complex Domain, Amer. Math. Soc. Coll. Publ., vol. 19, Amer. Math. Soc., New York, 1934. 21. A. I. Plesner, Spectral Theory of Linear Operators, Nauka, Moscow, 1965. (Russian) 22. Ya. V. Radyno, The space of vectors of exponential type, Dokl. Akad. Nauk BSSR 27 (1983), no. 9, 791 -- 793. (Russian) 23. J. Wermer, Commuting spectral measures on Hilbert space, Pacific J. Math. 4 (1954), no. 3, 355 -- 361. Department of Mathematics, California State University, Fresno 5245 N. Backer Avenue, M/S PB 108 Fresno, CA 93740-8001 E-mail address: [email protected] Received 04/06/2015
1302.1958
2
1302
2015-08-06T09:17:20
Variance of operators and derivations
[ "math.FA", "math.OA" ]
The variance of a bounded linear operator $a$ on a Hilbert space $H$ at a unit vector $h$ is defined by $D_h(a)=\|ah\|^2-|<ah,h>|^2$. We show that two operators $a$ and $b$ have the same variance at all vectors $h\in H$ if and only if there exist scalars $\sigma,\lambda$ with $|\sigma|=1$ such that $b=\sigma a+\lambda1$ or $a$ is normal and $b=\sigma a^*+\lambda1$. Further, if $a$ is normal, then the inequality $D_h(b)\leq\kappa D_h(a)$ holds for some constant $\kappa$ and all unit vectors $h$ if and only if $b=f(a)$ for a Lipschitz function $f$ on the spectrum of $a$. Variants of these results for C$^*$-algebras are also proved. We also study the related, but more restrictive inequalities $\|bx-xb\|\leq \|ax-xa\|$ supposed to hold for all $x\in B(H)$ or for all $x\in B(H^n)$ and all positive integers $n$. We consider the connection between such inequalities and the range inclusion $d_b(B(H))\subseteq d_a(B(H))$, where $d_a$ and $d_b$ are the derivations on $B(H)$ induced by $a$ and $b$. If $a$ is subnormal, we study these conditions in particular in the case when $b$ is of the form $b=f(a)$ for a function $f$.
math.FA
math
VARIANCE OF OPERATORS AND DERIVATIONS BOJAN MAGAJNA Abstract. The variance of a bounded linear operator a on a Hilbert space H at a unit vector ξ is defined by Dξ(a) = kaξk2 − haξ, ξi2. We show that two operators a and b have the same variance at all vectors ξ ∈ H if and only if there exist scalars σ, λ ∈ C with σ = 1 such that b = σa + λ1 or a is normal and b = σa∗ + λ1. Further, if a is normal, then the inequality Dξ(b) ≤ κDξ(a) holds for some constant κ and all unit vectors ξ if and only if b = f (a) for a Lipschitz function f on the spectrum of a. Variants of these results for C∗-algebras are also proved, where vectors are replaced by pure states. We also study the related inequalities kbx − xbk ≤ kax − xak supposed to hold for all x ∈ B(H) or for all x ∈ B(Hn) and all n ∈ N. We consider the connection between such inequalities and the range inclusion db(B(H)) ⊆ da(B(H)), where da and db are the derivations on B(H) induced by a and b. If a is subnormal, we study these conditions in particular in the case when b is of the form b = f (a) for a function f . 1. Introduction and notation The expected value of a quantum mechanical quantity represented by a self- adjoint operator a on a complex Hilbert space H in a state ω is ω(a), while the variance of a is defined by Dω(a) = ω(a∗a) − ω(a)2. If a is the multiplication by a bounded measurable function on L2(µ) for a probability measure µ and ω is the state x 7→ hx1, 1i, where 1 ∈ L2(µ) is the constant function, these notions reduce to the classical notions of probability calculus. We may define the variance by the same formula for all (not necessarily selfadjoint) operators a ∈ B(H). For a general vector state ω(a) := haξ, ξi, coming from a unit vector ξ ∈ H, the variance Dω(a) = kaξk2 − haξ, ξi2 means just the square of the distance of aξ to the set of all scalar multiples of ξ. (Thus Dω(a) = ηa(ξ)2, where η is the function considered by Brown and Pearcy in [7].) We will prove that an operator a is almost determined by its variances: if a, b ∈ B(H) are such that Dω(a) = Dω(b) for all vector states ω then b = αa + β1 or a is normal and b = αa∗ + β1 for some α, β ∈ C with α = 1 (Theorem 2.3). We will also deduce a variant of this statement for C∗-algebras, where vector states are replaced by pure states. 2010 Mathematics Subject Classification. Primary 47B06, 47B47, 46L07; Secondary 47A60, 47B15, 47B20. Key words and phrases. bounded linear operator, variance, state, derivation, completely bounded map, subnormal operator. Acknowledgment. I am grateful to Miran Cerne for discussions concerning complex anaysis topics, to Matej Bresar and Spela Spenko for conversations from which some of the questions studied in this paper have emerged, and to Victor Shulman for the correspondence concerning a question about Besov spaces. I am especially grateful to the anonymous referee for all his comments and corrections of the paper. The author was supported in part by the Ministry of Science and Education of Slovenia. 1 2 BOJAN MAGAJNA Then we will study the inequality (1.1) Dω(b) ≤ κDω(a), where κ is a positive constant (which may be taken to be 1 if we replace b by κ−1/2b). If (1.1) holds for all vector states ω, then we will show that there exists a Lipschitz function f : σap(a) → σap(b), where σap(·) denotes the approximate point spectrum, such that if a is normal then b = f (a) (Theorem 3.5). For a general a, however, f is perhaps not nice enough to allow the definition of f (a). Therefore we will also consider stronger variants of (1.1). kdb(x)k ≤ κkda(x)k (∀x ∈ B(H)), For 2 × 2 matrices (1.1) implies that b = αa + β1 for some scalars α, β ∈ C (Lemma 3.2). But for general operators the condition (1.1) is not very restrictive for it does not even imply that b commutes with a. For example, if a is hyponormal (1.1) holds with b = a∗ and κ = 1. A simple computation (Lemma 4.1) shows, however, that for a vector state ω = ωξ and a normal operator a the quantity Dω(a) is just the square of the norm of the operator da(ξ ⊗ ξ∗), where da is the derivation on B(H), defined by da(x) = ax− xa, and ξ⊗ ξ∗ is the rank one operator on H, defined by (ξ ⊗ ξ∗)η = hη, ξiξ. Thus we will also study the condition (1.2) where a, b ∈ B(H) and κ > 0 are fixed. We will show (Theorem 4.2) that if equality holds in (1.2) and κ = 1 then either b = σa + λ1 for some scalars σ, λ ∈ C with σ = 1 or there exist a unitary u and scalars α, β, λ, µ in C with β = α such that a = αu∗ + λ1 and b = βu + µ1. This will also be generalized to C∗-algebras. For a normal operator a Johnson and Williams [20] proved that the condition (1.2) is equivalent to the range inclusion db(B(H)) ⊆ da(B(H)). Their work was continued by several researchers, including Williams [40], Fong [15], Kissin and Shulman [22], Bresar [8] and in [9] in different contexts, but still restricted to special classes of operators a (such as normal, isometric or algebraic). It is known that the range inclusion db(B(H)) ⊆ da(B(H)) does not imply (1.2) in general since it does not even imply that b is in the bicommutant (a)′′ of a [19]. However the author does not know of any operators a, b satisfying (1.2) for which the range inclusion does not hold. The corresponding purely algebraic problem for operators on an (infinite dimensional) vector space V, where B(H) is replaced by the algebra L(V) of all linear operators on V and the condition (1.2) is replaced by the inclusion of the kernels ker da ⊆ ker db, was studied in [25]. By the Hahn-Banach theorem the inclusion db(B(H)) ⊆ da(B(H)) (the norm closure) is equivalent to the requirement that for each ρ ∈ B(H)♯ (the dual of B(H)) the condition aρ − ρa = 0 implies bρ − ρb = 0, where aρ and bρ are functionals on B(H) defined by (aρ)(x) = ρ(xa) and (ρa)(x) = ρ(ax). The operator spaces B(H) and B(H)♯ are quite different (if H is infinite dimensional), so in general we can not expect a strong connection between (1.2) and a formally similar condition kbρ − ρbk ≤ κkaρ − ρak (∀x ∈ B(H)♯). Question. Does (1.2) imply at least that the centralizer Ca of a in B(H)♯ (that is, the set of all ρ ∈ B(H)♯ satisfying aρ = ρa) is contained in Cb? A stronger condition than (1.2), namely that (1.2) holds for all x ∈ Mn(B(H)) and all n ∈ N (where a and b are replaced by the multiples a(n) and b(n) acting on Hn), implies that (1.2) holds in any representation of the C∗-algebra generated by a, b and 1 (Lemma 6.1) and that b is contained in the C∗-algebra generated 3 In a special situation (when H is a cogenerator for by a and 1 (Corollary 6.2). Hilbert modules over the operator algebra A0 generated by a and 1) it follows that b must be in A0 (Proposition 6.3). If a is, say, subnormal (a restriction of a normal operator to an invariant subspace), this means that b = f (a) for a function f in the uniform closure of polynomials on σ(a). Perhaps for a general subnormal operator a (1.2) does not imply that b = f (a) for a function f , but when it does, it forces on f certain degree of regularity. For example, if a is the operator of multiplication on the Hardy space H 2(G) by the identity function on G, where G is a domain in C bounded by finitely many nonintersecting analytic Jordan curves, (1.2) implies that b is an analytic Toeplitz operator with a symbol f which is continuous also on the boundary of G (Proposition 7.3). Let us call a complex function f on a compact set K ⊆ C a Schur function if the supremum over all (finite) sequences λ = (λ1, λ2, . . .) ⊆ K of norms of matrices Λ(f ; λ) =(cid:20) f (λi) − f (λj) λi − λj (cid:21) , regarded as Schur multipliers, is finite. (Here the quotient is interpreted as 0 if λi = λj.) If a is normal the work of Johnson and Williams [20] tells us that b = f (a) satisfies (1.2) if and only if f is a Schur function on σ(a). In the 'only if' direction we extend this to general subnormal operators (Proposition 7.1), in the other direction only to subnormal operators with nice spectra (Theorem 7.10). In the last section we will investigate the condition (1.2) in the case when a is subnormal and b = f (a) for a function f . If a is normal, a known effective method of studying such commutator estimates is based on double operator inte- grals (see [2] and the references there), which are defined via spectral projection valued measures. But, since invariant subspaces of a normal operator are not necessarily invariant under its spectral projections, a different method is needed for subnormal operators. In Section 7 we will 'construct' for a given subnormal operator a and suitable function f on σ(a) a completely bounded map Ta,f on B(H) such that Ta,f commutes with the left and the right multiplication by a and aTa,f (x) − Ta,f (x)a = f (a)x − xf (a) for all x ∈ B(H). For b = f (a) this implies (1.2) and also the range inclusion db(B(H)) ⊆ da(B(H)). By the above mentioned result from [20] even if a is normal the functions f considered here must be Schur. By [20] every Schur function on σ(a) is complex differentiable relative to σ(a) at each nonisolated point of σ(a) (thus holomorphic on the interior of σ(a)) and f ′ is bounded. The construction of Ta,f applies to the subclass that includes all func- tions for which f ′ is Lipschitz of order α > 0. Only if σ(a) is sufficiently nice are we able to find Ta,f for all Schur functions. We will denote by S the norm closure and by S the weak* closure of a subset S in B(H). 2. Variance of operators Definition 2.1. For a bounded operator a on a Hilbert space H and a nonzero vector ξ ∈ H let Dξ(a) = (kaξk2kξk2 − haξ, ξi2)kξk−2. Thus, if ξ is a unit vector and ω : x 7→ hxξ, ξi is the corresponding vector state on B(H), then Dξ(a) = ω(a∗a) − ω(a)2, 4 BOJAN MAGAJNA and this formula can be used to define the variance Dω(a) of a in any (not just vector) state ω. Remark 2.2. (i) It is clear from the definition that Dξ(a) is just the square of the distance of aξ to the set Cξ of scalar multiples of ξ. Hence, if Dξ(b) ≤ Dξ(a) for all ξ ∈ H, then in particular each eigenvector of a is also an eigenvector for b. Consequently Dξ(b) = 0 for all unit vectors ξ ∈ H if and only if b ∈ C1. (ii) Dξ(αa + β1) = α2Dξ(a) for all a, b ∈ B(H) and α, β ∈ C. (iii) Dξ(a∗) = Dξ(a) for all ξ ∈ H if and only if a is normal. Theorem 2.3. If operators a, b ∈ B(H) satisfy Dξ(b) = Dξ(a) for all ξ ∈ H, then there exist α, β ∈ C with α = 1 such that b = αa + β or a is normal and b = αa∗ + β. Proof. We assume that neither a no b is a scalar multiple of the identity, otherwise the proof is easy. If b is of the form b = αa + β1 (α, β ∈ C), then the hypothesis Dξ(b) = Dξ(a) for all ξ ∈ H, that is, α2Dξ(a) = Dξ(a), clearly implies that α = 1. To deduce a similar conclusion in the case whenf b = αa∗ + β1, replacing b with b − β1, we may assume that b = αa∗. From the hypothesis we have that α2Dξ(a∗) = Dξ(a), which implies that a and a∗ have the same eigenvectors, hence, if dimH < ∞, we can see inductively that a is normal. To prove the same in general, we consider the distance d(a, C1) := inf λ∈C ka − λ1k = d(a∗, C1). We may assume that this distance (and also d(b, C1) = αd(a∗, C1)) is achieved at λ = 0 (otherwise we just consider a − λ1 instead of a). Then by [34, Theorem 2] there exists a sequence of unit vectors ξn ∈ H such that limnhaξn, ξni = 0 and limn kaξnk = kak. From Dξn (a) = α2Dξn(a∗) it now follows kak2 = lim n→∞kaξnk2 = α2 lim n→∞ka∗ξnk2 ≤ α2ka∗k2 = α2kak2. This implies that α ≥ 1 and similarly we prove (by exchanging the roles of a and a∗) that α ≤ 1. Thus α = 1 and from Dξ(a∗) = Dξ(a) (for all ξ ∈ H) we now see that a must be normal. To prove the theorem, we will now assume that neither b nor b∗ is of the form αa + β1 and show that this leads to a contradiction. For any two nonzero vectors ξ, η ∈ H we expand the function f (z) := Dξ+zη(a)kξ + zηk2 = ka(ξ + zη)k2kξ + zηk2 − ha(ξ + zη), ξ + zηi2 of the complex variable z into powers of z and z, f (z) = Dξ(a)kξk2+2Re (D1z)+2Re (D2z2)+D3z2+2Re (D4z2z)+Dη(a)kηk2z4. Among the coefficients Dj we will need to know only D2, which is D2 = haη, aξihη, ξi − haη, ξihη, aξi. Thus, from the equality Dξ+zη(a) = Dξ+zη(b), by considering the coefficients of z2 we obtain hbη, bξihη, ξi − hbη, ξihη, bξi = haη, aξihη, ξi − haη, ξihη, aξi. From this we see that if η is orthogonal to ξ and aξ then η must be orthogonal to bξ or to b∗ξ. In other words, if for a fixed ξ we denote H0(ξ) = {ξ, aξ}⊥, H1(ξ) = {ξ, aξ, bξ}⊥, H2 = {ξ, aξ, b∗ξ}⊥, then H0(ξ) = H1(ξ) ∪ H2(ξ). Since Hj(ξ) are vector spaces, this implies that H1(ξ) = H0(ξ) or else H2(ξ) = H0(ξ). In the first case we have bξ ∈ Cξ + Caξ, 5 while in the second case b∗ξ ∈ Cξ + Caξ. Since this holds for all ξ ∈ H, it follows that H is the union of the two sets F1 = {ξ ∈ H : bξ ∈ Cξ + Caξ} and F2 = {ξ ∈ H : b∗ξ ∈ Cξ + Caξ}. ◦ Fi. We will consider the case when Since F1 and F2 are closed, by Baire's theorem at least one of them has nonempty F16= ∅ and in appropriate places point interior out the differences with the other case, which is similar. Since a /∈ C1, there exists ◦ a vector ξ ∈ F1 such that ξ and aξ are linearly independent. (Namely, if aξ = αξξ F1, where αξ ∈ C, then considering this equality for the vectors ξ, ζ and for all ξ ∈ ◦ (1/2)(ξ + ζ) in F1, where ξ and ζ are linearly independent, it follows easily that αξ must be independent of ξ for ξ in an open subset of H, hence a must be a scalar multiple of 1.) Let ◦ ◦ ◦ U = {ξ ∈ F1: ξ and aξ are linearly independent}. ◦ For any ξ, η ∈ U and z ∈ C let ξ(z) = (1 − z)ξ + zη. If ξ and η are such that the 'segment' ξ(z) (z ≤ 1) is contained in U , then we have (2.1) bξ(z) = α(z)aξ(z) + β(z)ξ(z) for some scalars α(z), β(z) ∈ C. To see that the coefficients α and β are holomorphic (in fact rational) functions of z (for fixed ξ and η), for any fixed z0 with z0 ≤ 1 we take the inner product of both sides of (2.1) with the vectors ξ(z0) and aξ(z0) to obtain two equations from which we compute α(z) and β(z) by Cramer's rule (if z is near z0). (From the condition Dξ(z)(b) = Dξ(z)(a) and (2.1) we also conclude that α(z) = 1, hence α must be constant, which we could use to somewhat simplify the proof in the present case. But this argument is not available in the other case, F26= ∅, since we do not know if Dξ(z)(b∗) = Dξ(z)(a), hence we will not use it.) when Since α and β are rational functions, it follows from (2.1) that bξ(z) is contained in the two-dimensional space S(z) spanned by ξ(z) and aξ(z) for all z ∈ C. (Here we have used that the singular points are isolated and that the set of z for which bξ(z) /∈ S(z) is open.) It is known that this implies, since b is not in L := C1 + Ca, that L contains an operator of rank one (see [26, 2.5] and use that 1 ∈ L). Thus, replacing a by a + λ1 for a suitable λ ∈ C, we may assume that a is of rank 1. Let H0 be a 2-dimensional subspace of H containing the range of a and the orthogonal complement of the kernel of a and set K = H⊥ 0 . Then bξ ∈ Cξ for all ξ ∈ K, which easily implies that bK = λ1K for a scalar λ. Replacing b by b− λ1, we may assume that bK = 0. Since also bξ ∈ Caξ + Cξ ⊆ H0 for all ξ ∈ H0, we see that both operators a and b now live on the two dimensional space H0. (In the case when F26= ∅, the same arguments reduce the proof to the case when a and b∗, hence also b, live on the same two dimensional space.) Thus, it only remains to show that for two 2 × 2 complex matrices a and b the condition Dξ(b) = Dξ(a) implies that b ∈ Ca + C1. Now instead of proving this here directly we just refer to Lemma 3.2(i) below, where a sharper result is proved. Corollary 2.4. If elements a, b in a C∗-algebra A ⊆ B(H) satisfy Dω(b) = Dω(a) for all pure states ω on A, then there is a projection p in the center Z of the weak* closure R of A and central elements u1, z1 ∈ Rp, u2, z2 ∈ Rp⊥, with u1, u2 unitary, such that bp = u1a + z1 and bp⊥ = u2a∗ + z2 and ap⊥ is normal. (cid:3) ◦ 6 BOJAN MAGAJNA Proof. Since the condition Dω(b) = Dω(a) persists for all weak* limits of pure states on A and such states are precisely the restrictions of weak* limits of pure states on R by [16, Theorem 5], the proof immediately reduces to the case A = R. Let Z be the center of R, ∆ the maximal ideal space of Z, for each t ∈ ∆ let Rt be the closed ideal of R generated by t and set R(t) := R/(Rt). For any a ∈ R let a(t) denotes the coset of a in R(t). Since each pure state on R(t) can be lifted to a pure state on R, we have Dω(b(t)) = Dω(a(t)) for each pure state ω on R(t) and each t ∈ ∆. Since R(t) is a primitive C∗-algebra by [18], it follows from Theorem 2.3 that there exist scalars α(t), β(t), with α(t) = 1, such that (2.2) or (2.3) b(t) = α(t)a(t) + β(t)1 b(t) = α(t)a(t)∗ + β(t)1 and a is normal. Let F1 be the set of all t ∈ ∆ for which (2.2) holds, F2 the set of all those t for which (2.3) holds and U the set of all t such that a(t) is not a scalar. Since for each x ∈ R the function t 7→ kx(t)k is continuous on ∆ by [16], it is easy to see that U is open and F1, F2 are closed. To show that the coefficients α and β in (2.2) and (2.3) are continuous functions of t on U , let t ∈ U be fixed, note that the center of R(t) is C1 and that R(t) is generated by projections, so there is a projection pt ∈ R(t) such that (1−pt)a(t)pt 6= 0. We may lift 1 − pt and pt to positive elements x, y in R with xy = 0 [21, 4.6.20]. Then from (2.2) (2.4) x(s)b(s)y(s) = α(s)x(s)a(s)y(s) (∀s ∈ ∆), and k(xay)(s)k 6= 0 for s in a neighborhood of t by continuity. Now let c ∈ Z be the element whose Gelfand transform is the function s 7→ kx(s)a(s)y(s)k and let φ : R → Z be a bounded Z-module map such that φ(xay) = c. (Such a map may be obtained simply as the completely bounded Z-module extension to R of the map Z(xay) → Z, z 7→ z(xay), since Z is injective [6].) Since φ is a Z module map, φ is just a collection of maps φs : R(s) → Z(s) = C, hence from (2.4) we obtain α(s)c(s) = (φ(xby))(s). Since c(t) = kx(t)a(t)y(t)k 6= 0, it follows that α is continuous in a neighborhood of t, hence continuous on U . Then, denoting by q0 the projection corresponding to a clopen neighborhood U0 ⊆ U of t, we have from (2.2) that β(t)q0(t) = e(t), where e = (bq0 − αaq0) ∈ Rq0, hence βU0 represents a central element of Rq0 and is therefore continuous. Since ∆ (hence also U ) is a Stonean space and α (hence also β) are bounded continuous functions, they have continuous extensions to U (see [21, p. 324]). If q ∈ Z is the projection that corresponds to U , then aq⊥ is a scalar in Rq⊥, and it follows easily that bq⊥ must also be a scalar. So we have only to consider the situation in Rq, which means that we may assume that U = ∆, hence that α and ◦ β are defined and continuous throughout ∆. The interior F := F 1 of F1 is a clopen subset such that (2.2) holds for t ∈ F . Since the complement F c (= F c 1 ) is contained in F2, (2.3) holds if t ∈ F c. Finally, to conclude the proof, just let p ∈ Z be the projection that corresponds to F , and let u1, z1 ∈ Zp, u2, z2 ∈ Zp⊥ be elements that corresponds to functions αF , βF , αF c and βF c (respectively). (cid:3) 7 We note that the converse of Corollary 2.4 also holds, the proof follows easily from the well-known fact [21, p. 268] that if ω is a pure state on a C∗-algebra R then ω(xz) = ω(x)ω(a) for all x ∈ R and all z in the center of R. 3. The inequality Dξ(b) ≤ Dξ(a) Lemma 3.1. For any two operators a, b ∈ B(H) and any state ω on B(H) the following estimate holds: Dω(b) − Dω(a) ≤ 2kb − ak(kak + kbk). Proof. Since ω(b∗b− a∗a) ≤ kb∗b− a∗ak = k(b∗− a∗)b + a∗(b− a)k ≤ kb− ak(kbk + kak) and(cid:12)(cid:12)ω(b)2 − ω(a)2(cid:12)(cid:12) = (ω(b) +ω(a))ω(b) − ω(a) ≤ (kak +kbk)kb− ak, we have Dω(b) − Dω(a) = ω(b∗b − a∗a) − (ω(b)2 − ω(a)2) ≤ ω(b∗b − a∗a) +(cid:12)(cid:12)(ω(b)2 − ω(a)2)(cid:12)(cid:12) ≤ 2kb − ak(kak + kbk). (cid:3) Lemma 3.2. Let a, b ∈ M2(C) (2 × 2 complex matrices), 0 < ε < 1/2, and let αi and βi (i = 1, 2) be the eigenvalues of a and b (respectively). (i) If Dξ(b) ≤ Dξ(a) for all unit vectors ξ ∈ C2, then b = θa + τ for some scalars (ii) If Dξ(b) ≤ Dξ(a) + ε8 for all unit vectors ξ ∈ C2, then β2 − β1 ≤ α2 − α1 + 2ε(kak + 2kbk + 1). θ, τ ∈ C with θ ≤ 1. Proof. (i) Since Dξ(a) = Dξ(a − λ1) for all λ ∈ C, we may assume that one of the eigenvalues of a is 0, say aξ2 = 0 for a unit vector ξ2 ∈ C2. Then from 0 ≤ Dξ2(b) ≤ Dξ2 (a) = 0 we see that ξ2 is also an eigenvector for b, hence (replacing b by b − λ1 for a λ ∈ C) we may assume that bξ2 = 0. So, choosing a suitable orthonormal basis {ξ1, ξ2} of C2, we may assume that a and b are of the form a =(cid:20) α1 γ 0 0 (cid:21) , b =(cid:20) β1 δ 0 0 (cid:21) . Now we compute for any unit vector ξ = (λ, µ) ∈ C2 (using λ2 + µ2 = 1) that Dξ(a) = kaξk2 − haξ, ξi2 = (α12 + γ2)λ2 −(cid:12)(cid:12)α1λ2 + γλµ(cid:12)(cid:12) = λ2[α12µ2 + γ2λ2 − 2Re (α1γλµ)]. 2 Using this and a similar expression for Dξ(b), the condition Dξ(b) ≤ Dξ(a) can be written as (3.1) (α12 − β12)µ2 + (γ2 − δ2)λ2 − 2Re ((α1γ − β1δ)λµ) ≥ 0, which means that the matrix M =(cid:20) α12 − β12 β1δ − α1γ γ2 − δ2 (cid:21) β1δ − α1γ is nonnegative. This is equivalent to the conditions β1 ≤ α1, δ ≤ γ and det M ≥ 0. 8 BOJAN MAGAJNA Since det M = −α1δ − β1γ2, the condition det M ≥ 0 means that α1δ = β1γ. If α1 6= 0, it follows that b is of the form b =(cid:20) β1 β1 α1 γ 0 0 (cid:21) = β1 α1 a = θa, where θ := β1 α1 , hence θ ≤ 1. If α1 = 0, then β1 = 0 (since β1 ≤ α1), hence again b = θa, where θ = δ/γ if γ 6= 0. (ii) As above, replacing a and b by a−λ1 and b−µ1, where λ and µ are eigenvalues of a and b, we may assume that a and b are of the form a =(cid:20) α1 γ 0 0 (cid:21) , b =(cid:20) β1 δ1 δ2 0 (cid:21) . The norms of the new a and b are at most two times greater than the norms of original ones, which will be taken into account in the final estimate. If ξ = (0, 1), then Dξ(b) = δ22 and Dξ(a) = 0, hence the condition Dξ(b) ≤ Dξ(a) + ε8 shows that δ2 ≤ ε4. Thus, denoting by b0 the matrix 0 (cid:21) , 0 we have that kb − b0k ≤ ε4, hence by Lemma 3.1 b0 =(cid:20) β1 δ1 Dξ(b) − Dξ(b0) ≤ 2ε4(kb0k + kbk) ≤ 4kbkε4 for all unit vectors ξ ∈ C2. It follows that Dξ(b0) ≤ Dξ(a) + 4kbkε4 + ε8 ≤ Dξ(a) + ε4(4kbk + 1). The same calculation that led to (3.1) shows now that λ2[(α12 − β12)µ2 + (γ2 − δ12)λ2 − 2Re ((α1γ − β1δ1)λµ)] ≥ −ε4(4kbk + 1) for all λ, µ ∈ C with λ2 + µ2 = 1. We may choose the arguments of λ and µ so that (α1γ − β1δ1)λµ is positive, hence the above inequality implies that t[(α12 − β12)(1 − t) + (γ2 − δ12)t − 2α1γ − β1δ1pt(1 − t)] ≥ −ε4(4kbk + 1) for all t ∈ [0, 1]. Setting t = ε2, it follows (since γ ≤ kak and δ1 ≤ kbk) that (α12 − β12)(1 − ε2) + (kak2 + kbk2)ε2 ≥ −ε2(4kbk + 1), hence α12 − β12 ≥ −ε2(kak2 + kbk2 + 4kbk + 1), so β1 ≤ α1 + ε(kak + 2kbk + 1). Taking into account that α2 and β2 were initially reduced to 0 (by which the norms of a and b may have increased at most by a factor 2), this proves (ii). (cid:3) The approximate point spectrum of an operator a will be denoted by σap(a). Definition 3.3. If a, b ∈ B(H) are such that Dξ(b) ≤ Dξ(a) for all ξ ∈ H, then we can define a function f : σap(a) → σap(b) as follows. Given α ∈ σap(a), let (ξn) be a sequence of unit vectors in H such that limk(a − α1)ξnk = 0. Then from the condition Dξn (b) ≤ Dξn (a) we conclude that limk(b− λn1)ξnk = 0, where λn = hbξn, ξni. We will show that the sequence (λn) converges, so we define f (α) = lim λn. Proposition 3.4. The function f is well-defined and Lipschitz: f (β) − f (α) ≤ β − α for all α, β ∈ σap(a). 9 Proof. To slightly simplify the computation, we assume that a and b are contrac- tions; for general a and b the proof is essentially the same. Given ε > 0, choose unit vectors ξ, η ∈ H such that k(a − α1)ξk < ε and k(a − β1)ηk < ε. Let p be the projection onto the span of {ξ, η} and let c be the operator on pH defined by cξ = αξ and cη = βη. Then kapH − ck2 ≤ k(a − c)ξk2 + k(a − c)ηk2 < 2ε2. (3.2) Let λ = hbξ, ξi, µ = hbη, ηi and d the operator on pH defined by dξ = λξ and dη = µη. Then, using the conditions Dξ(b) ≤ Dξ(a) and Dη(b) ≤ Dη(a), we have (3.3) kbpH − dk2 ≤ k(b − λ1)ξk2 + k(b − µ1)ηk2 ≤ k(a − α1)ξk2 + k(a − β1)ηk2 < 2ε2. Now by Lemma 3.1 and Remark 2.2(i) and since kdk ≤ kbk, kck ≤ kak we infer from (3.2) and (3.3) that Dξ(d) ≤ Dξ(b) + 4kd − bpHkkbk < Dξ(b) + 4ε√2 and Dξ(c) > Dξ(a) − 4ε√2, hence (since Dξ(b) ≤ Dξ(a)) Dξ(d) ≤ Dξ(c) + 8ε√2 for all ξ ∈ H with kξk = 1. By Lemma 3.2 (ii) we now conclude that (3.4) where κ is a constant. µ − λ ≤ β − α + κε 1 8 , If (ξn) and (ηn) are two sequences of unit vectors in H such that limk(a − α1)ξnk = 0 and limk(a − β1)ηnk = 0, we infer from (3.4) (since ε can be taken to tend to 0 as n → ∞) that (3.5) Further, if β = α and we put in (3.4) λn = hbξn, ξni instead of λ and λm = hbξm, ξmi instead of µ, we conclude that (λn) is a Cauchy sequence, hence it converges to a point λ ∈ C. From limk(b − λn1)ξnk = 0 it follows now that limk(b − λ1)ξnk = 0, hence λ ∈ σap(b). Similarly the sequence (µn) = (hbηn, ηni) converges to some µ and (3.5) implies that lim supµn − λn ≤ β − α. µ − λ ≤ β − α. (cid:3) This shows that f is a well-defined Lipschitz function. Theorem 3.5. Let a, b ∈ B(H). If a is normal, then there exists a constant κ such that Dξ(b) ≤ κDξ(a) for all ξ ∈ H if and only if b = f (a) for a Lipschitz function f on σ(a). In this case Dω(b) ≤ κDω(a) for all states ω. Proof. Assume that Dξ(b) ≤ Dξ(a) for all ξ ∈ H. We may assume that a is not a scalar (otherwise the proof is trivial). First consider the case when a can be represented by a diagonal matrix diag (αj) in some orthonormal basis (ξj ) of H. If f : σ(a) → σap(b) is defined as in Definition 3.3, then bξj = f (αj)ξj for all j, hence b = f (a). For a general normal a, first suppose that H is separable. Then by Voiculescu's version of the Weyl-von Neumann-Bergh theorem [39], given ε > 0, there exists a diagonal normal operator c = diag (γj) such that ka− ck2 < ε, where k ·k2 denotes the Hilbert-Schmidt norm. Let (ξj ) be an orthonormal basis of H consisting of 10 BOJAN MAGAJNA eigenvectors of c, so that cξj = γjξj . Since Dξj (b) ≤ Dξj (a), by Remark 2.2(i) there exist scalars βj ∈ C such that k(b − βj1)ξjk ≤ k(a − γj1)ξjk = k(a − c)ξjk, hence Xj k(b − βj1)ξjk2 ≤Xj k(a − c)ξjk2 < ε2. In particular kb − dk < ε, where d is the diagonal operator defined by dξj = βjξj. Since d and c commute, it follows that kbc − cbk = k(b − d)c − c(b − d)k < 2εkck ≤ 2ε(kak + ε) ≤ 4εkak (if ε ≤ kak), hence also kba − abk = k(bc − cb) + b(a − c) − (a − c)bk ≤ 4ε(kak + kbk). Since this holds for all ε > 0, it follows that a and b commute. If a has a cyclic vector this already implies that b is in (a)′′ hence a measurable function of a, but in general we need an additional argument to prove this. Let f : σ(a) → σap(b) be defined as in Definition 3.3. (Note that σ(a) = σap(a) since a is normal.) Let e(·) be the projection valued spectral measure of a, ξ ∈ H any separating vector for the von Neumann algebra (a)′′ generated by a and ε > 0. If α is any point in σ(a), U is any Borel subset of σ(a) containing α and ξU := ke(U )ξk−1e(U )ξ, then k(a− α1)ξUk converges to 0 as the diameter of U shrinks to 0. For each U let βU = hbξU , ξUi so that k(b− βU 1)ξUk ≤ k(a− α1)ξUk; then f (α) = limU→{α} βU by the definition of f . Thus, since by Proposition 3.4 f is a Lipschitz function, for each α ∈ σ(a) there is an open neighborhood Uα with the diameter at most ε such that f (α)− βU < ε for all Borel subsets U ⊆ Uα and f (α2)− f (α1) < ε if α1, α2 ∈ Uα. By compactness we can cover σ(a) with finitely many such neighborhoods Uαi and this covering then determines a partition of σ(a) into finitely many disjoint Borel sets ∆j (say j = 1, . . . , n) such that each ∆j is contained in some Uαi(j) . Let ej = e(∆j). Now we can estimate, denoting βj = β∆j , k(b − f (a))ejξk ≤k(b − βj1)ejξk + (βj − f (αi(j))kejξk + k(f (αi(j))1 − f (a))ejξk ≤k(a − αj 1)ejξk + (βj − f (αi(j))kejξk + k(f (αi(j))1 − f (a))ejξk ≤3εkejξk. (Here we have used the spectral theorem to estimate the term k(f (αi(j))1−f (a))ejξk from above by supα∈∆j f (αi(j) − f (α)kejξk ≤ εkejξk.) Since b commutes with a, hence also with all spectral projections of a, it follows that k(b − f (a))ξk2 =k ej(b − f (a))ejξk2 = n Xj=1 kej(b − f (a))ejξk2 n Xj=1 Xj=1 ≤9ε2 n kejξk2 = 9ε2kξk2. Thus k(b−f (a))ξk ≤ 3εkξk and, since this holds for all ε > 0 and separating vectors of (a)′′ are dense in H, we conclude that b = f (a). If H is not necessarily separable, H can be decomposed into an orthogonal sum of separable subspaces Hk that reduce both a and b and are such that σ(aHk) = σ(a). For each k there exists a Lipschitz function fk such that bHk = f (aHk). Since for any two k, j the space Hk ⊕Hj is also separable, there also exists a function f such 11 that b(Hj ⊕ Hk) = f (a(Hj ⊕ Hk)) and it follows easily that fk = f = fj. Thus b = f (a). Conversely, if b = f (a) for a function f such that f (α2) − f (α1) ≤ κα2 − α1 for all α1, α2 ∈ σ(a) and some constant κ, then for a fixed unit vector ξ ∈ H denote by µ the probability measure on Borel subsets of σ(a) defined by µ(·) = he(·)ξ, ξi. Since Dξ(a) is just the square of the distance of aξ to Cξ and similarly for Dξ(b), the estimate k(f (a) − f (α))1ξk2 =Zσ(a) f (λ) − f (α)2 dµ(λ) ≤Zσ(a) κλ − α2 dµ(λ) = κk(a − α1)ξk2 implies that Dξ(b) ≤ κDξ(a). Finally, since any state ω is in the weak*-closure of the set of all convex combi- nations of vector states and each such combination can be represented as a vector state on B(Hn) for some n ∈ N, the argument of the previous paragraph (applied to a(n) and b(n) = f (a(n)) implies that Dω(b) ≤ Dω(a). (cid:3) A variant of the above Theorem 3.5 was proved in [20] and generalized to C∗- algebras in [9], but both under the much stronger hypothesis that k[b, x]k ≤ κk[a, x]k for all elements x, where [a, x] denotes the commutator ax − xa. (See Lemma 4.1 below for the explanation of the connection between the two conditions.) The following Corollary was proved in [9, 5.2] for prime C∗-algebras, but under a much stronger assumption about the connection between a and b instead of the inequality Dω(b) ≤ Dω(a) for pure states ω. Corollary 3.6. Let A be a unital C∗-algebra, a, b ∈ A, a normal. If Dω(b) ≤ Dω(a) for all states ω on A, then b = f (a) for a function f on σ(a) such that f (µ) − f (λ) ≤ µ − λ for all λ, µ ∈ σ(a). If A is prime, it suffices to assume the condition for pure states only. Proof. The first statement follows immediately from Theorem 3.5 since we may assume that A ⊆ B(H) for a Hilbert space H and each vector state on B(H) restricts to a state on A. For the second statement, we note that the C∗-algebra generated by a and b is contained in a separable prime C∗-subalgebra A0 of A by [14, 3.1] (an elementary proof of this is in [24, 3.2]), and A0 is primitive by [29, p. 102], hence we may assume that A0 is an irreducible C∗-subalgebra of B(H). But then each vector state on B(H) restricts to a pure state on A0, and each pure state on A0 extends to a pure state on A. (cid:3) Corollary 3.7. Let a, b ∈ B(H) satisfy Dξ(b) ≤ Dξ(a) for all ξ ∈ H. If a is essentially normal, then this implies that b = f ( a) for a Lipschitz function f on the essential spectrum of a, where a denotes the coset of a in the Calkin algebra. Proof. Any state ω on the Calkin algebra can be regarded as a state on B(H) annihilating the compact operators. By Glimm's theorem (see [21, 10.5.55] or [16]) such a state ω is a weak* limit of vector states, hence Dω(b) ≤ Dω(a). The conclusion follows now from Corollary 3.6. (cid:3) 12 BOJAN MAGAJNA Theorem 3.8. Let A ⊆ B(H) be a C∗-algebra a, b ∈ A and a normal. Denote by R the weak* closure of A and by Z the center of R. Then the inequality Dω(b) ≤ Dω(a) holds for all pure states ω on A if and only if b is in the norm closure of the set S of all elements of the form Pj pjfj(a) (finite sum), where pj are orthogonal projections in Z with the sum Pj pj = 1 and fj are functions on σ(a) such that fj(µ) − fj(λ) ≤ µ − λ for all λ, µ ∈ σ(a). Proof. Note that g(a(t)) = g(a)(t) for each continuous function g on σ(a). We will use the notation from the proof of Corollary 2.4. Similarly as in that proof, the condition that Dω(b) ≤ Dω(a) for all pure states ω on A implies the same condition for all pure states on R(t) for all t ∈ ∆ and it follows then from Corollary 3.6 that for each t there exists a Lipschitz function ft on σ(a(t)) with the Lipschitz constant 1 such that b(t) = ft(a(t)). By Kirzbraun's theorem each ft can be extended to a Lipschitz function on σ(a), denoted again by ft, with the same Lipschitz constant 1. Given ε > 0, since ∆ is extremely disconnected and for each x ∈ R the function t 7→ kx(t)k is continuous on ∆ by [16], each t ∈ ∆ has a clopen neighborhood Ut such that kft(a)(s) − b(s)k ≤ ε for all s ∈ Ut. Let (Uj) be a finite covering of ∆ by such neighborhoods Uj := Utj and for each j let pj be the central projection in R that corresponds to the clopen set Uj, and set fj := ftj . Then kb −Xj pjfj(a)k ≤ ε. Since this can be done for all ε > 0, b is in the closure of the set S as stated in the theorem. Conversely, suppose that for each ε > 0 there exists an element c ∈ R of the form c =Pj pjfj(a), where pj ∈ Z are projections with the sum 1 and fj are Lipschitz functions with the Lipschitz constant 1, such that kb − ck < ε. Then for each pure state ω on R and x ∈ R, z ∈ Z the equality ω(zx) = ω(z)ω(x) holds [21, 4.3.14]). In particular ωZ is multiplicative, hence ω(pj0 ) = 1 for one index j0 and ω(pj) = 0 if j 6= j0. It follows now by a straightforward computation that Dω(c) = Dω(fj0 (a)), which is at most Dω(a) by the same computation as in the last part of the proof of Theorem 3.5. Now, since kb− ck < ε, it follows from Lemma 3.1 (by letting ε → 0) that Dω(b) ≤ Dω(a). (cid:3) 4. Is a derivation determined by the norms of its values? Given an operator a ∈ B(H), we will denote by da the derivation on B(H) defined by da(x) = ax − xa. For any vectors ξ, η ∈ H we denote by ξ ⊗ η∗ the rank one operator on H defined by (ξ ⊗ η∗)(ζ) = hζ, ηiξ. The following lemma enables us to interpret the results of the previous section in terms of derivations. Lemma 4.1. For each unit vector ξ ∈ H and a ∈ B(H) we have the equality kda(ξ ⊗ ξ∗)k2 = max{Dξ(a), Dξ(a∗)}. Thus, if a is normal, then kda(ξ ⊗ ξ∗)k2 = Dξ(a). Proof. Denote x = ξ⊗ ξ∗. The square of the norm of da(x) = aξ⊗ ξ∗− ξ⊗ (a∗ξ)∗ is equal to the spectral radius of the operator T := da(x)∗da(x), which is the largest eigenvalue of the restriction of T to the span H0 of ξ and a∗ξ. If ξ and a∗ξ are linearly independent, then the matrix of TH0 in the basis {ξ, a∗ξ} can easily be computed to be 13 (cid:20) Dξ(a) 0 hξ, aξi(kaξk2 − ka∗ξk2) Dξ(a∗) (cid:21) . Thus kda(x)k2 = max{Dξ(a), Dξ(a∗)}. By continuity (considering perturbations of a) we see that this equality holds even if ξ and a∗ξ are linearly dependent . (cid:3) Theorem 4.2. If a, b ∈ B(H) are such that (4.1) then either b = σa + λ1 for some scalars σ, λ ∈ C with σ = 1 or there exist a unitary u and scalars α, β, λ, µ in C with β = α such that a = αu∗ + λ1 and b = βu + µ1. k[b, x]k = k[a, x]k for all x ∈ B(H), A variant of this theorem was proved in [9, 5.3, 5.4] in general C∗-algebras, but under the additional assumption that a and b are normal. The methods in [9] are different from those we will use below. The author is not able to deduce Theorem 4.2 as a direct consequence of the previous results; for a proof we will need two additional lemmas. We denote by a(n) the direct sum of n copies of an operator a ∈ B(H), thus a(n) acts on Hn. We will also use the usual notation [x, y] := xy−yx, so that da(x) = [a, x]. Remark 4.3. We will need the following, perhaps well-known, general fact: for any bounded linear operators S, T : X → Y between Banach spaces the inequality (4.2) implies kT ♯♯vk ≤ kS♯♯vk (v ∈ X ♯♯), where T ♯♯ denotes the second adjoint of T . This follows from [20, 1.1, 1.3], but here is a slightly more direct proof. The inequality (4.2) simply means that there is a contraction Q from the range of S into the range of T such that T = QS. But then T ♯♯ = Q♯♯S♯♯, which clearly implies the desired conclusion. kT xk ≤ kSxk (x ∈ X) The content of the following lemma was observed already by Kissin and Shulman k[b, x]k ≤ k[a, x]k in the proof of [22, 3.3]. Lemma 4.4. [22] Let a, b ∈ B(H) and suppose that (4.3) for all x ∈ K(H). If a is normal, then k[b(n), x]k ≤ k[a(n), x]k for all x ∈ Mn(B(H)) (n × n matrices with the entries in B(H)) and all n ∈ N. Proof. Since da = (daK(H))♯♯ (the second adjoint in the Banach space sense), it follows from Remark 4.3 that (4.3) holds for all x ∈ B(H). Suppose now that a is normal and note that (a)′ is a C∗-algebra by the Fuglede- Putnam theorem. Since (4.3) holds for all x ∈ B(H), b ∈ (a)′′. Further, by (4.3) the map [a, x] 7→ [b, x] is a contraction from da(B(H)) to db(B(H)). Clearly this map is a homomorphism of (a)′-bimodules, hence by [33, 2.1, 2.2, 2.3] it is a complete contraction, which is equivalent to the conclusion of the lemma. (cid:3) Remark 4.5. We will use below the following well-known fact. Given cj, ej ∈ B(H), an identity of the form Pn j=1 cjxej = 0, if it holds for all x ∈ B(H), implies that all cj must be 0 if the ej are linearly independent. (See e. g. [3, Theorem 5.1.7]). 14 BOJAN MAGAJNA We refer to [6] or [28] for the definition of the injective envelope of an operator space used in the following lemma. Lemma 4.6. Let R = da(B(H)) and let S be the operator system S =(cid:26)(cid:20) λ z∗ µ (cid:21) : λ, µ ∈ C y, z ∈ R(cid:27) . y If a does not satisfy any quadratic equation over C then the C∗-algebra C∗(S) generated by S is irreducible and the injective envelope I(S) of S is M2(B(H)). Proof. Since S contains the diagonal 2×2 matrices with scalar entries, each element of S′ (the commutant of S) is a block diagonal matrix, that is, of the form c ⊕ e, where c, e ∈ B(H). To prove the irreducibility of C∗(S) means to prove that each selfadjoint such element c ⊕ e is a scalar multiple of 1. Since c ⊕ e commutes with elements of S, we have that cy = ye for all y ∈ R. Setting y = ax − xa in the last identity we obtain cax − cxa − axe + xae = 0 for all x ∈ B(H). (4.4) Since in (4.4) the left coefficients ca,−c,−a and 1 are not all 0, it follows that 1, a, e, ae are linearly dependent. Thus, if 1, a and e are linearly independent, then ae = α1 + βa + γe for some scalars α, β, γ ∈ C. Using this, we may rearrange (4.4) into (4.5) (ca + α1)x + (β1 − c)xa + (γ1 − a)xe = 0. If 1, a and e were linearly independent, then (4.5) would imply that a = γ1, but this would be in contradiction with the assumption about a. Hence 1, a and e are linearly dependent, say e = α1 + βa (α, β ∈ C). Then (4.4) can be rewritten as (4.6) (c − α1)ax + (α1 − βa − c)xa + βxa2 = 0. Since 1, a and a2 are linearly independent by assumption, we infer from (4.6) that β = 0 and c = α1. But then e = α1 and c ⊕ e = α(1 ⊕ 1). This proves the irreducibility of C∗(S). Since S contains nonzero compact operators, the identity map on S has a unique completely positive extension to C∗(S) by the Arveson boundary theorem [5], which implies that C∗(S) ⊆ I(S). (Otherwise a projection B(H) → I(S) restricted to C∗(S) would be a completely positive extension of idS, different from idC ∗(S).) But since C∗(S) is irreducible and contains nonzero compact operators, it follows that C∗(S) ⊇ M2(K(H)), hence I(S) must contain the injective envelope I(M2(K(H))), which is known to be M2(B(H)) [6]. Proof of Theorem 4.2. If a (or b) is a scalar multiple of 1 the proof is easy, so we assume from now on that this is not the case. If a satisfies a quadratic equation of the form (cid:3) a2 + βa + γ1 = 0 (β, γ ∈ C), then each element of (a)′′ is a polynomial in a (this holds for any algebraic operator a by [37]), hence in particular b is a linear polynomial in a, say b = σa + λ1. Then the condition (4.1) obviously implies that σ = 1. Hence we may assume that a does not satisfy any quadratic equation over C. By Lemma 4.1 the assumption (4.1) implies that max{Dξ(b), Dξ(b∗)} = max{Dξ(a), Dξ(a∗)} for all unit vectors 15 ξ ∈ H, hence for each non-zero ξ ∈ H at least one of the following four equalities hold: Dξ(b) = Dξ(a), Dξ(b) = Dξ(a∗), Dξ(b∗) = Dξ(a), Dξ(b∗) = Dξ(a∗). (4.7) Since the functions of the form H ∋ ξ 7→ kξk2Dξ(a) are continuous, it follows that H is the union of four closed sets Fi, where F1 = {ξ ∈ H; kξk2Dξ(b) = kξk2Dξ(a)} and so on. By Bair's theorem at least one of the sets Fi has nonempty interior and then, since functions of the form ξ 7→ kξk2Dξ(a) are polynomial (more precisely, for any fixed vectors ξ, η the function z 7→ kξ + zηk2Dξ+zη is a polynomial in z and z), at least one of the equalities (4.7) must hold for all nonzero ξ ∈ H. In each case it follows then by Theorem 2.3 that b must have the form b = σa + λ1 or b = σa∗ + λ1, where σ = 1. Moreover, in the second case, which we assume from now on (otherwise the proof is already completed), we deduce now from (4.1) that k[a∗, x]k = k[a, x]k for all x ∈ B(H), hence (setting x = a) a must be normal. Replacing b by αb+β for suitable α, β ∈ C we may assume without loss of generality that b = a∗. Denote by Ra and Rb the ranges of the derivations da and db and by Sa and Sb the corresponding operator systems (as in Lemma 4.6). Since a is normal, by Lemma 4.4 the map φ : Ra → Rb, φ([a, x]) := [b, x] (x ∈ B(H)) is completely contractive and the same holds for its inverse. Hence φ is completely isometric and consequently the map Φ : Sa → Sb, Φ(cid:18)(cid:20) α y z∗ β (cid:21)(cid:19) :=(cid:20) α φ(z)∗ φ(y) β (cid:21) is completely positive with completely positive inverse, hence also completely iso- metric (see [28]). But then Φ extends to a complete isometry ψ between the injective envelopes I(Sa) and I(Sb) (since both Φ and Φ−1 extend to complete contractions which must be each other's inverse by rigidity). Since a (and b = a∗) does not sat- isfy any quadratic equation over C, these injective envelopes are both M2(B(H)) by Lemma 4.6. Hence ψ is a unital surjective complete isometry of M2(B(H)) = B(H2). Thus by [6, 4.5.13] or [21, Ex. 7.6.18] (and since all automorphisms of B(H2) are inner) ψ is necessarily of the form ψ(y) = w∗yw (y ∈ B(H2)), where w ∈ B(H2) is unitary. Since by definition ψ fixes the projections of H2 on the two summands, w must commute with these two projections (by the multiplicative domain argument, see [28, p. 38]), consequently w is of the form w = u ⊕ v for unitaries u, v ∈ B(H). It follows now from the definition of ψ that φ is of the form φ(y) = uyv (y ∈ Ra), that is φ([a, x]) = u[a, x]v. Hence u[a, x]v = [b, x] for all x ∈ B(H), which can be rewritten as (4.8) uaxv − uxav − bx + xb = 0 (x ∈ B(H)). Thus by Remark 4.5 we see from (4.8) that v, av, 1, and b are linearly dependent. Hence, if 1, v and b are linearly independent, then av = α1 + βb + γv, where α, β γ ∈ C, and (4.8) can be rewritten as (ua − γu)xv − (αu + b)x + (1 − βu)xb = 0. 16 BOJAN MAGAJNA But by Remark 4.5 this implies in particular that ua − γu = 0, hence a = γ1, a possibility which we have excluded in the first paragraph of this proof. So we may assume that 1, v and b are linearly dependent. If v were a scalar, say v = δ, then (4.8) could be rewritten as (δua − b)x − δuxa + xb = 0, which would imply that 1, a and b are linearly dependent, a possibility already taken care of in the beginning of the proof. Thus we may assume that v is not a scalar. Hence b = α1 + βv for suitable α, β ∈ C. Since v is unitary and a = b∗, this concludes the proof. (cid:3) To extend Theorem 4.2 to C∗-algebras we need a lemma. Lemma 4.7. Let A ⊆ B(H) be a C∗-algebra, J a closed ideal in A, and let a, b ∈ A satisfy k[b, x]k ≤ k[a, x]k for all x ∈ A. Then the same inequality holds for all x ∈ A and also for all cosets x ∈ A/J. Proof. The statement about the quotient was observed already in [9, Proof of 5.4] and follows from the existence of a quasicentral approximate unit (ek) in J [4]. Namely, the conditions k[a, ek]k, k[b, ek]k → 0 (from the definition of the quasicen- tral approximate unit) and the well-known property that k yk = limk ky(1 − ek)k (y ∈ A) imply that k[b, x]k = lim k k[a, x(1 − ek)]k = k[ a, x]k. Let A♯♯ be the universal von Neumann envelope of A (= bidual of A) and regard A as a subalgebra in A♯♯ in the usual way. Since d♯♯ a is just the derivation induced by a on A♯♯, it follows from Remark 4.3 that the condition k[b, x]k ≤ k[a, x]k holds for all x ∈ A♯♯. Since A is a quotient of A♯♯, it follows from the previous paragraph (applied to A♯♯ instead of A) that the condition holds also in A. Corollary 4.8. If A is a C∗-algebra and a, b ∈ A are such that k[b, x]k = k[a, x]k for all x ∈ A, then there exist a projection p in the center Z of A and elements s, d ∈ Zp with s unitary, and u, v, c, g, h ∈ Zp⊥ with u, v unitary, such that bp = sa + d and ap⊥ = cu∗ + g, bp⊥ = vcu + h. k k[b, x](1 − ek)k = lim k k[b, x(1 − ek)] ≤ lim (cid:3) Proof. If A is primitive the corollary follows immediately from Theorem 4.2 and Lemma 4.7 since A = B(H) if A is irreducibly represented on H. In general, Lemma 4.7 reduces the proof to von Neumann algebras, where the arguments are similar as in the proof of Corollary 2.4, so we will omit the details. (cid:3) Corollary 4.9. If k[b, x]k ≤ k[a, x]k for all x ∈ A then max{Dω(b), Dω(b∗)} ≤ max{Dω(a), Dω(a∗)} for all pure states ω on A. Proof. If π : A → B(Hπ) is the irreducible representation obtained from ω by the GNS construction, then π(A) = B(Hπ), hence the corollary follows from Lemmas 4.7 and 4.1. (cid:3) 5. An inequality between norms of commutators In this section we study the inequality (5.1) where a, b ∈ B(H) are fixed and κ is a constant. For a normal a it is proved in [20] that (5.1) holds (for some κ) if and only if k[b, x]k ≤ κk[a, x]k (∀x ∈ B(H)), (5.2) db(B(H)) ⊆ da(B(H)). That for normal a (5.1) implies (5.2) can be easily proved as follows. We have seen in the proof of Lemma 4.4 that for normal a the condition (5.1) is equivalent to the fact that the map 17 da(x) 7→ db(x) (x ∈ B(H)) is a completely bounded homomorphism of (a)′-bimodules da(B(H)) → db(B(H)). Then this map can be extended to a completely bounded (a)′-bimodule endomor- phism φ of B(H) by the Wittstock theorem (see [6, 3.6.2]), hence we have db(x) = φ(da(x)) = da(φ(x)) (x ∈ B(H)). When studying the connection between (5.1) and (5.2), it is useful to have in mind a fact (recalled below as Lemma 5.1) concerning operators in B(X, Y ), the space of all bounded linear operators from X into Y , where X and Y are Banch spaces. Denote by X ♯ the dual of X and by T ♯ the adjoint of T ∈ B(X, Y ). The following is well-known (see [20]). Lemma 5.1. Given S, T ∈ B(X, Y ), the inclusion T ♯(Y ♯) ⊆ S♯(Y ♯) holds if and only if there exists a constant κ such that (5.3) for all ξ ∈ X. kT ξk ≤ κkSξk Since da = −(daT(H))♯, where T(H) is the ideal in B(H) of trace class operators, the following is just a special case of Lemma 5.1. Corollary 5.2. Let a, b ∈ B(H). (i) The inclusion db(B(H)) ⊆ da(B(H)) holds if and only if there exists a constant κ such that kdb(t)k1 ≤ κkda(t)k1 for all t ∈ T(H). (ii) The inclusion db(T(H)) ⊆ da(T(H)) is equivalent to the existence of a con- stant κ such that kdb(x)k ≤ κkda(x)k for all x ∈ K(H) or (equivalently, by Lemma 4.7) for all x ∈ B(H). If a is not normal, then the range inclusion (5.2) does not necessarily imply that b ∈ (a)′′ [19], hence it does not imply (5.1). But we will prove that conversely (5.1) implies (5.2), if a satisfies certain conditions which are more general than normality. Proposition 5.3. Denote Ra := da(B(H)). If Ra + (a)′ = B(H), then for each b ∈ B(H) the condition (5.1) implies that Rb ⊆ Ra. Moreover, if Ra = B(H), then there exists a weak* continuous (a)′-bimodule map φ on B(H) such that db = φda = daφ. Proof. By (5.1) the correspondence da(x) 7→ db(x) extends to a bounded map φ0 from Ra into Rb such that φ0da = db. Note that φ0(da(K(H))) ⊆ db(K(H)). Recall that for normed spaces Y ⊆ Z the weak* closure Y of Y in Z ♯♯ can be naturally identified with Y ♯♯, hence in particular da(K(H))♯♯ = da(K(H)) inside K(H)♯♯ = B(H). It follows that φ := (φ0da(K(H)))♯♯ is the weak* continuous extension of φ0 to da(B(H)) = da(K(H)) satisfying φda = db. Since φ0 is an (a)′-bimodule map, so must be φ by continuity, hence in particular db(x) = φ(da(x)) = daφ(x) for all x ∈ Ra = da(B(H)) and consequently db(Ra) ⊆ Ra. Finally, to conclude the proof, note that the assumption Ra + (a)′ = B(H) implies that Rb = db(Ra), since from (5.1) (a)′ ⊆ (b)′ = ker db so that Rb = db(Ra + (a)′) = db(Ra). (cid:3) 18 BOJAN MAGAJNA By duality the condition da(B(H)) = B(H) means that the kernel of daT(H) is 0, that is, (a)′ ∩ T(H) = 0. There are many Hilbert space operators a which do not commute even with any nonzero compact operator. This is so for example, if a is normal and has no eigenvalues. (Namely, (a)′ is a C∗-algebra and contains the spectral projection p corresponding to any nonzero eigenvalue of each h = h∗ ∈ (a)′. If h is compact, then p is of finite rank, hence ap, and therefore also a, has eigenvalues.) For a general normal a ∈ B(H) we can decompose H into the orthogonal sum H = H1⊕H2, where H1 is the closed linear span of all eigenvectors of a and H2 = H⊥ 1 . Then a also decomposes as a1 ⊕ a2, where (a2)′ contains no nonzero compact operators, while a1 is diagonal in an orthonormal basis. (A general subnormal operator, however, can commute with a nonzero trace class operator even if it is pure; an example is in [41, 2.1].) Corollary 5.4. Let a ∈ B(H) and suppose that H decomposes into the orthogonal sum H1 ⊕ H2 of two subspaces which are invariant under a, so that a = a1 ⊕ a2, where ai ∈ B(Hi). If a1 is a diagonalizable normal operator, while (a2)′∩T(H2) = 0 and σp(a2) ∩ σp(a1) = ∅, σp(a∗ 1) = ∅, where σp(c) denotes the set of all eigenvalues of an operator c, then the condition kdb(x)k ≤ kda(x)k (∀x ∈ B(H)) implies that db(B(H)) ⊆ da(B(H)). Proof. Since b ∈ (a)′′, H1 and H2 are invariant subspaces for b, so b also decomposes as b = b1 ⊕ b2, where bi ∈ B(Hi). Relative to the same decomposition of H each x ∈ B(H) can be represented by a 2 × 2 operator matrix x = [xi,j ] and 2) ∩ σp(a∗ db(x) =(cid:20) b1x1,1 − x1,1b1 b2x2,1 − x2,1b1 b2x2,2 − x2,2b2 (cid:21) . b1x1,2 − x1,2b2 Thus it suffices to show that for each pair (i, j) of indexes and for each xi,j ∈ B(Hj,Hi) the element bixi,j − xi,j bj is in the range of the map dai,aj defined on B(Hj,Hi) by dai,aj (y) = aiy − yaj. In the case i = 2 = j this follows from Proposition 5.3 and in the case i = 1 = j this is an elementary special case of a result from [20]. We will now consider the case i = 2 and j = 1, the remaining case i = 1 and j = 2 is treated similarly. From the norm inequality condition we have in particular that kdb2,b1 (x)k ≤ kda2,a1(x)k (∀x ∈ B(H1,H2)). This implies that there exists a bounded (a2)′, (a1)′-bimodule map φ0 : da2,a1(K(H1,H2)) → K(H1,H2) such that φ0da2,a1 = db2,b1. (To prove that φ0 is indeed a bimodule map, we use that (ai)′ ⊂ (bi)′, which follows from (a)′ ⊆ (b)′.) As in the proof of Proposition 5.3 we now extend φ0 weak* continuously to the weak* closure R of the range R of da2,a1 and show that db2,b1 (R) ⊆ R. Finally, let (ξj)j∈J be an orthonormal basis of H1 consisting of eigenvectors of a1 and let αj be the corresponding eigenvalues. If y ∈ 2y∗ξj−αjy∗ξj, ηi = hξj, ya2ηi− ker da1,a2, then for each j ∈ J and η ∈ H2 we have ha∗ hy∗a∗ 1ξj, ηi = −hξj , da1,a2(y)ηi = 0, which means (by the arbitrariness of η) that y∗ξj is an eigenvector for a∗ 2) ∩ σp(a∗ 1) = ∅ and the vectors ξj span H1, we infer that y = 0. Thus ker da1,a2 = 0. Consequently R (which is just the annihilator in B(H1,H2) of ker(da1,a2T(H2,H1)) is equal to B(H1,H2). Therefore db2,b1(B(H1,H2) = db2,b1(R) ⊆ R. 2 with the eigenvalue αj. Since by assumption σp(a∗ (cid:3) 19 Problem. Does Corollary 5.4 still hold if we omit the hypothesis about the disjointness of the point spectra? Perhaps, in general, (5.1) does not even imply that db(B(H)) ⊆ da(B(H)), but no counterexample is known to the author. Note, however, that (5.1) implies that ker daT(H) ⊆ ker dbT(H), hence by duality db(B(H)) ⊆ da(B(H)); in particular db(K(H)) ⊆ da(K(H)) since the weak topology agrees on K(H) with the weak* topology inherited from B(H). More generally, we will see that the question, whether (5.1) implies the inclusion db(B(H)) ⊆ da(B(H)), depends entirely on what happens in the Calkin algebra. For a C∗-algebra A and a ∈ A note that a functional ρ ∈ A♯ annihilates da(A) if and only if [a, ρ] = 0, where [a, ρ] ∈ A♯ is defined by ([a, ρ])(x) = ρ(xa − ax). In other words, the annihilator in A♯ of da(A) is just the centralizer Ca of a in A♯. Proposition 5.5. If a, b ∈ B(H) satisfy k[b, x]k ≤ k[a, x]k for all x ∈ B(H), then k[b, x]k ≤ k[ a, x]k in the Calkin algebra C(H). If this latter inequality implies that C a ⊆ Cb, then Ca ⊆ Cb also holds, hence db(B(H)) ⊆ da(B(H)). Proof. The first statement follows from Lemma 4.7. To prove the rest of the propo- sition, first note that for any a ∈ B(H) and a functional ρ ∈ Ca the normal part (Indeed, from [a, ρ] = 0 we have ρn and the singular part ρs are both in Ca. [a, ρn] = −[a, ρs], where the left side is normal and the right side is singular, hence both are 0.) Further, since ρn is given by a trace class operator t, [a, t] = 0, hence the hypothesis of the proposition implies that [b, t] = 0, so ρn ∈ Cb. Since singular func- tionals annihilate K(H), they can be regarded as functionals on the Calkin algebra C(H). Thus, if the condition k[b, x]k ≤ k[ a, x]k ( x ∈ C(H)) implies that C a ⊆ Cb, then we have ρs ∈ Cb, which means just that ρs ∈ Cb (since ρs annihilates K(H)). Now both ρn and ρs are in Cb, hence so must be their sum ρ. This proves that Ca ⊆ Cb. The Hahn-Banach theorem then implies that db(B(H)) ⊆ da(B(H)). (cid:3) 6. Commutators and the completely bounded norm In this section we will study stronger variants of the condition k[b, x]k ≤ k[a, x]k (x ∈ B(H)) in the context of completely bounded maps. Lemma 6.1. If a, b ∈ B(H) satisfy (6.1) k[b(n), x]k ≤ k[a(n), x]k for all x ∈ Mn(B(H)) and all n ∈ N, then k[π(b), x]k ≤ k[π(a), x]k for all x ∈ B(Hπ) (6.2) for every unital ∗-representation π : A → B(Hπ) of the C∗-algebra A generated by 1, a and b. Proof. First assume that Hπ is separable. Let J = K(H) ∩ A, Hn = [π(J)Hπ], and let πn and πs be the representations of A defined by πn(a) = π(a)Hn and πs(a) = π(a)H⊥ n (a ∈ A), so that π = πn ⊕ πs. By basic theory of representations of C∗-algebras of compact operators πn is a subrepresentation of a multiple id(m) of the identity representation. By Voiculescu's theorem ([38], [4]) the representation π⊕id is approximately unitarily equivalent to πn⊕id, hence π⊕id is approximately unitarily equivalent to a subrepresentation of id(m+1). It follows easily from (6.1) 20 BOJAN MAGAJNA that (6.2) holds for any multiple of the identity representation in place of π, hence it must also hold for any subrepresentation ρ of id(m+1) (to see this, just take in (6.2) for x elements that live on the Hilbert space of ρ). But then it follows from the approximate equivalence that the condition (6.2) holds for π ⊕ id in place of π, hence also for π itself. In general, when Hπ is not necessarily separable, Hπ decomposes into an orthog- onal sum ⊕i∈IHi of separable invariant subspaces for π(A). For a fixed x ∈ B(Hπ) there exists a countable subset J of I such that the norm of the operator [π(b), x] is the same as the norm of its compression to L := ⊕i∈JHi. Since L is separable, it follows from what we have already proved that k[π(b), x]k ≤ k[π(a), x]k. Corollary 6.2. If a, b ∈ B(H) satisfy (6.1) then b is contained in the C∗-algebra B generated by a and 1. Proof. Let π be the universal representation of A = C∗(a, b, 1) and Hπ its Hilbert space. It follows from Lemma 6.1 (that is, from (6.2)) that π(b) ∈ (π(a))′′, hence also π(b) ∈ π(B)′′. But π(B)′′ = π(B), thus π(b) ∈ π(B)∩ π(A) = π(B), where the last equality is by [21, 10.1.4]. (cid:3) (cid:3) A completely contractive Hilbert module H over an operator algebra A (that is, a Hilbert space on which A has a completely contractive representation) is a cogenerator if for each nonzero morphism R : K → L of Hilbert A-modules (that is, a bounded A-module map) there exists a morphism T : L → H such that T R 6= 0 [6, 3.2.7]. Here by an operator algebra we will always mean a norm complete algebra of operators on a Hilbert space. Proposition 6.3. If a, b ∈ B(H) satisfy (6.1), where H is a cogenerator for the operator algebra A0 generated by a and 1, then b ∈ A0. Proof. Let π be the universal representation of the C∗-algebra A generated by 1, a and b. Then Hπ (the Hilbert space of π) is a cogenerator for A0. (Indeed, let R : K → L be a nonzero morphism of Hilbert A0-modules and denote by ρ the completely contractive representation of A0 on L through which the A0-module structure has been introduced on L. There exists a representation σ of A on a Hilbert space L1 ⊇ L such that ρ(a) = σ(a)L for all a ∈ A0 [28], hence L is a Hilbert A0-submodule of L1. Thus R(K) ⊆ L1. Since π is universal (thus L1 is contained in a multiple of Hπ), there exists a morphism T1 : L1 → Hπ of Hilbert A-modules such that T1(R(K)) 6= 0. Then T := T1L : L → Hπ is a morphism of Hilbert A0 modules such that T R 6= 0.) Hence by the Blecher-Solel bicommutation theorem (see [6, 3.2.14]) π(A0) = π(A0)′′. From Lemma 6.1 π(b) ∈ π(A0)′′, hence π(b) ∈ π(A0) ∩ π(A) = π(A0) by [21, 10.1.4]). (cid:3) The author does not know if in Proposition 6.3 the assumption that H is a cogenerator is dispensable. In particular the following problem is open. Problem. If in (6.1) a is subnormal, is then b necessarily of the form b = f (a) for some function f ? Is b necessarily subnormal? 7. Commutators of functions of subnormal operators By Theorem 3.5 and Lemma 4.1 for a normal operator a the condition (5.1) implies that b = f (a) for a Lipschitz function f . However, as observed in [20], (5.1) implies that f must have additional properties. In this section we will study properties of a function f that imply or are implied by an inequality of the form 21 (7.1) k[f (a), x]k ≤ κk[a, x]k ∀x ∈ B(H), where a is a subnormal operator. 7.1. Schur functions. Let us begin with the case when a is a diagonal normal operator. Then there exists an orthonormal basis of H consisting of eigenvectors of a; let λi be the corresponding eigenvalues. If we denote by [xi,j ] the matrix of a general operator x ∈ B(H) with respect to this basis, then the inequality (7.1) assumes the form (7.2) k[(f (λi) − f (λj ))xi,j ]k ≤ κk[(λi − λj )xi,j ]k. In the same way we can express the inequality in Corollary 5.2(i). Since there exist contractive projections from B(H) and from T(H) onto subsets of block diagonal matrices, it follows that the condition (7.1) and its analogue for the trace norm are equivalent to the requirements that the matrix Λ(f ) with the entries (7.3) Λi,j(f ) =( f (λi)−f (λj ) λi−λj 0, , if λi 6= λj if λi = λj is a Schur multiplier on B(H) and T(H) (respectively). In one direction the last statement can be generalized to subnormal operators. Proposition 7.1. Let a ∈ B(H) be a subnormal operator and let f be a Lipschitz function on σ(a). If a is not normal, assume that f is in the uniform closure of the set of rational functions with poles outside σ(a), so that b := f (a) is defined. If k[b, x]k ≤ κk[a, x]k for all x ∈ B(H), then for each sequence (λi) ⊆ σ(a) the matrix Λ(f ; λ) with the entries defined by the right side of (7.3) is a Schur multiplier with the norm at most 2κ. That is, f is a Schur function on σ(a) as defined in the Introduction. Similarly, the condition k[b, x]k1 ≤ κk[a, x]k1 for all x ∈ T(H) implies that Λ(f ; λ) is a Schur multiplier on T(H) with the norm at most 2κ. Proof. First suppose that (λi)m i=1 is a finite subset of the boundary ∂σ(a) of σ(a), where the λi are distinct. Then each λi is an approximate eigenvalue of a [10], hence there exists a sequence of unit vectors ξi,n ∈ H such that limn k(a − λi1)ξi,nk = 0. Since a − λi1 is hyponormal, k(a − λi1)∗ξi,nk ≤ k(a − λi1)ξi,nk and it follows that the sequence (λi − λj )hξi,n, ξj,ni = hλiξi,n, ξj,ni − hξi,n, λjξj,ni converges to limn(haξi,n, ξj,ni − hξi,n, a∗ξj,ni) = 0. Thus limhξi,n, ξj,ni = 0 if i 6= j, so the set {ξ1,n, . . . , ξm,n} is approximately orthonormal if n is large. Therefore for each matrix α = [αi,j] ∈ Mm(C) the norm of the operator x :=Pm i,j=1 αi,j ξi,n⊗ξ∗ j,n is approximately equal to the usual operator norm of α. Further, for large n we have approximate equalities m Xi,j=1 da(x) = and αi,j (aξi,n ⊗ ξ∗ j,n − ξi,n ⊗ (a∗ξj,n)∗) ≈ m Xi,j=1 αi,j (λi − λj )ξi,n ⊗ ξ∗ j,n db(x) ≈ m Xi,j=1 αi,j(f (λi) − f (λj ))ξi,n ⊗ ξ∗ j,n, 22 BOJAN MAGAJNA hence it follows from the assumption kdb(x)k ≤ κkda(x)k that i,j=1k ≤ κk[(λi − λj )αi,j]m (7.4) k[(f (λi) − f (λj ))αi,j ]m i,j=1k. By continuity the estimate (7.4) holds also when λi are not necessarily distinct. This estimate means that for a finite collection λ = (λi)m i=1 of not necessarily distinct elements of ∂σ(a) the matrix Λ(f ; λ) with the entries (7.5) Λi,j(f ; λ) =( f (λi)−f (λj ) λi−λj 0, , if λi 6= λj if λi = λj acts as a Schur multiplier with the norm at most κ on the subspace E ⊆ Mm(C) of matrices of the form [(λi − λj )αi,j]. Note that E (which depends on λ1, . . . , λm) is just the set of all matrices with zero entries on those positions (i, j) for which λi = λj . Let D be the subspace of corresponding block diagonal matrices (that is, matrices in Mm(C) with non-zero entries only on those positions (i, j) for which λi = λj ). Since the natural projection from Mm(C) onto D has Schur norm 1 (and Λ(f, λ)(D) = 0), it follows that the norm of Λ(f ; λ) as a Schur multiplier on Mm(C) is at most 2κ; the same bound 2κ is valid for all m. (Now it already follows from the second half of the proof of [20, 4.1], that for each (non-isolated) point ζ ∈ ∂σ(a) the limit f ′(ζ) := limz∈∂σ(a),z→ζ exists. So we can redefine the matrix Λ(f ; λ) by setting Λi,j = f ′(λi) if λi = λj (with f ′(λi) interpreted as 0 if λi is isolated). Then (7.4) and the continuity imply that the new Λ(f, λ) has Schur norm at most κ. But it is not necessary to use this redefined Λ(f ; λ) in this proof.) If a is normal, then the above argument applies to all points of σ(a) (not just points in ∂σ(a)) since all are approximate eigenvalues, hence we assume from now on that a is not normal. Then by hypothesis f is a uniform limit of rational functions with poles outside σ(a), hence holomorphic on the interior G of σ(a). We can use the first line of (7.5) to define Λi,j(f ; λ) also for all pairwise distinct λ1, . . . , λm from G. When λi = λj ∈ G we do define Λi,j(f ; λ) by setting Λi,j(f ; λ) = f ′(λj ). For fixed elements λ2, . . . , λm of ∂σ(a) consider the function f (z)−f (ζ) z−ζ g(λ1) := Λ(f ; λ1, λ2, . . . , λm) from σ(a) \ {λ2, . . . , λm} into the Banach algebra Sm = Mm(C) equipped with the Schur norm. This function is holomorphic on G and (since f is Lipschitz) bounded (by m2κ). We would like to prove that g is bounded on G by the same bound (2κ) as on ∂σ(a), but we do not know if g can be extended con- tinuously to the closure G of G. (Namely, discontinuities can appear at the pos- sible boundary points λ2, . . . , λm.) We may consider the scalar valued functions gω = ωg for all linear functionals ω on Sm with kωk = 1. If for a fixed ω we denote M = supζ∈∂G\{λ2,...,λm} limz→ζ,z∈G gω(z) = supζ∈∂G\{λ2,...,λm} gω(ζ) and h(z) = gω(z) − M , then h is subharmonic on G and it follows from the extended maximum principle [31, 3.6.9] (and the fact that finite sets are polar [31, p. 56], while ∂G is not polar since G is bounded) that h(ζ) ≤ 0 for all ζ ∈ G. Thus h(ζ) ≤ M for all ζ ∈ G and (since M ≤ supζ∈∂G kg(ζ)k ≤ 2κ) we deduce that supλ1∈G kg(λ1)k ≤ 2κ. Thus the Schur norm of Λ(f ; λ1, λ2, . . . , λm) is at most 2κ for all λ1 ∈ σ(a) and λ2, . . . , λm ∈ ∂σ(a). In the same way, by considering the func- tion λ2 7→ Λ(f ; λ1, λ2, . . . , λm) for fixed λ1 ∈ σ(a) and λ3, . . . λm ∈ ∂σ(a), we can now show that the Schur norm of Λ(f ; λ1, . . . , λm) is at most 2κ for all λ1, λ2 ∈ σ(a) and λ3, . . . , λm ∈ ∂σ(a). Proceeding successively, we see that this must hold for all 23 λi ∈ σ(a) and, since the same bound 2κ is valid for all choices of λ1, . . . , λm and all m, this implies that f is a Schur function on σ(a). This proves the case of B(H) and the proof for T(H) is similar. (cid:3) It is well-known that a matrix is a Schur multiplier on T(H) if and only if its transpose is a Schur multiplier on B(H) and the two multipliers have the same norm. For a rank one operator x the operators da(x) and db(x) have rank at most two and on such operators the trace class norm is equivalent to the usual operator norm. If a is normal, we deduce now from Corollary 5.2, Lemma 4.1 and Theorem 3.5 that each of the two range inclusions db(B(H)) ⊆ da(B(H)) and db(T(H)) ⊆ da(T(H)) implies that b is of the form b = f (a) for a Lipschitz function f on σ(a). Then f is a Schur function by Proposition 7.1. For normal operators the converse of Proposition 7.1 holds. Namely, let a be normal and f a Schur function on σ(a). Given ε > 0, by the Weyl-von Neumann-Bergh theorem [11, Corollary 39.6] there exists a diagonal operator a0 such that σ(a0) ⊆ σ(a), ka − a0k < ε and (approximating f by polynomials) kf (a) − f (a0)k < ε. Then, by what we have already proved for diagonal oper- ators (by the computation preceding Proposition 7.1), for each x ∈ B(H) with kxk = 1 we have k[f (a0), x]k ≤ κk[a0, x]k for a constant κ, hence k[f (a), x]k ≤ k[f (a0), x]k + 2ε ≤ κk[a0, x]k + 2ε ≤ κk[a, x]k + 2κε + 2ε. Since this holds for all ε > 0, we infer that k[f (a), x]k ≤ κk[a, x]k. Thus we may summarize the above discussion in the following theorem proved already by Johnson and Williams in [20] in a somewhat different way. Theorem 7.2. [20] If a ∈ B(H) is normal, then for any b ∈ B(H) the inclu- sion db(B(H)) ⊆ da(B(H)) holds if and only if there exists a constant κ such that kdb(x)k ≤ κkda(x)k for all x ∈ B(H) and this is also equivalent to the condition that b = f (a) for a Schur function f on σ(a). By [22, 6.5], if a is normal, (5.2) implies (5.1) in any C∗-algebra A. The converse is true only under additional assumptions about A (for example, if A is a von Neumann algebra), but since the proof would considerably lengthen the paper, we will not present it here. Following the usual convention, we denote by Rat(K) the algebra of all rational functions with poles outside a compact subset K ⊆ C and, if µ is a positive Borel measure on K, R2(K, µ) is the closure in L2(µ) of Rat(K). As before, for a ∈ B(H) we denote by a the coset in the Calkin algebra C(H). The simplest example of an operator a satisfying the conditions of our next proposition is the unilateral shift. Proposition 7.3. Let K be a compact subset of C, a a subnormal operator with σ(a) ⊆ K such that a is cyclic for the algebra Rat(K) and let c be the minimal normal extension of a. Assume that σ(c) = σ( a), let µ be a scalar spectral measure for c such that a is the multiplication on H := R2(K, µ) by the identity function z. Denote by p the orthogonal projection from K := L2(µ) onto H and assume that the only function h ∈ C(σ(c)) + (L∞(µ) ∩ R2(K, µ)) for which the operator Th defined by Th(ξ) := p(hξ) (ξ ∈ H) is compact is h = 0. Then for each b ∈ B(H) satisfying k[b, x]k ≤ k[a, x]k (x ∈ B(H)) there exists a function f ∈ C(σ(c)) ∩ R2(K, µ) such that b = f (c)H. Moreover, if K is the closure of a domain G bounded by finitely many non- intersecting analytic Jordan curves and a is the multiplication operator by z on the Hardy space H 2(G), f can be extended to a Schur function on K. 24 BOJAN MAGAJNA Proof. It is well-known that a rationally cyclic subnormal operator a can be repre- sented as the multiplication on R2(K, µ) by the independent variable z [12, p. 51] and that (a)′ = R2(K, µ)∩L∞(µ) by Yoshino's theorem [12, p. 52]. Since b ∈ (a)′, it follows that b is the multiplication on R2(K, µ) by a function f ∈ R2(K, µ)∩L∞(µ). Thus b = Tf since H = R2(K, µ) is invariant under multiplications by functions from R2(K, µ) ∩ L∞(µ). It follows from Lemma 4.1 and Bair's theorem (as in the proof of Theorem 4.2) that at least one of the inequalities Dξ(b) ≤ Dξ(a), Dξ(b) ≤ Dξ(a∗) holds for all nonzero ξ ∈ H. Since a is essentially normal by the Berger-Shaw theorem [12, p. 152], by Corollary 3.7 b = g( a) for a continuous function g on σ( a). Further, since c is normal and a is subnormal and essentially normal, an easy computation with 2 × 2 operator matrices (relative to the decomposition K = H ⊕ H⊥) shows that the operator p⊥c∗p is compact, hence (since also p⊥cp = 0) p c = c p. Consequently the map h 7→ Th (= ph(c)H) from C(σ(c)) into the Calkin algebra C(H) is a ∗- homomorphism, thus it must coincide with the ∗-homomorphism h 7→ h( a) since the two coincide on the generator idσ(c). It follows in particular that b = g( a) = Tg, hence the operator Tg−f = Tg − b is compact. But by the hypothesis this is possible only if g − f = 0, hence f = g, therefore continuous. In the case a is the unilateral shift, f is a continuous function on the circle and contained in the closure P 2(µ) of polynomials in L2(µ), where µ is the normalized Lebesgue measure on the circle. It is well known that such a function can be holomorphically extended to the disc D such that the extension (denoted again by f ) is continuous on D. By Proposition 7.1 f is a Schur function on D. Similar arguments apply to multiply connected domains bounded by analytic Jordan curves by [1, 2.11, 1.1], [27, 4.3, 9.4]. (cid:3) 7.2. A sufficient degree of smoothness. By Proposition 7.1 the inequality (7.1) can hold only for Schur functions. But the author does not know if (7.1) holds for all Schur functions and all subnormal operators a, we will prove this for all Schur functions only if σ(a) is nice enough (Theorem 7.10). It follows from the proof in [20, Theorem 4.1] that a Schur function f is complex differentiable in the sense that the limit f ′(ζ0) = limζ→ζ0, ζ∈σ(a)(f (ζ)− f (ζ0))/(ζ − ζ0) exists at each non-isolated point of σ(a). Moreover, from the Lipschitz condition on f we see that f ′ is bounded. However, the boundedness of f ′ is not sufficient for f to be a Schur function. When a is selfadjoint it is proved in [20, 5.1] that (7.1) holds if f (3) is continuous. We will prove (7.1) for subnormal operators a under a much milder condition on f (for example, f ′ Lipschitz suffices), but perhaps our condition on f is still more restrictive than Peller's condition that f is a restriction of a function from the appropriate Besov space (see [30] and [2]), which is sufficient when a is normal. We will start from the special case of the Cauchy-Green formula (7.6) g(λ) = − 1 π ZC ∂g(ζ) ζ − λ dm(ζ), which holds for a compactly supported differentiable function g such that ∂g is bounded. Here m denotes the planar Lebesgue measure and ∂g = (1/2)( ∂g ∂y ). (The proof in [32, 20.3] is valid for functions with the properties just stated.) We note that an operator calculus based on the Cauchy-Green formula was already ∂x + i ∂g developed by Dynkin [13], however we will need rather different results, specific to subnormal operators. 25 Lemma 7.4. If a ∈ B(H) is a subnormal operator and g : C → C is a differentiable function with compact support such that ∂g is bounded and ∂gσ(a) = 0, then (ξ, η ∈ H). (7.7) ∂g(ζ)h(ζ1 − a)−1η, ξi dm(ζ), hg(a)η, ξi = − 1 π ZC\σ(a) Proof. Let c ∈ B(K) be the minimal normal extension of a, e the projection valued spectral measure of c (which is 0 outside σ(c) ⊆ σ(a)), K = σ(a) and H = {ζ ∈ C : ∂g(ζ) 6= 0}. For fixed η ∈ H and ξ ∈ K denote by µ the measure he(·)η, ξi. Then by the spectral theorem g(c) = RK g(λ) de(λ) and (ζ1 − c)−1 = RK(ζ − λ)−1 de(λ) for each ζ ∈ C \ K (in particular for ζ ∈ H since H ∩ K = 0 because of ∂gK = 0), hence by (7.6) 1 1 1 π ZH g(λ) dµ(λ) = − hg(c)η, ξi =ZK π ZKZH π ZH = − 1 π ZC\σ(a) ∂g(ζ)h(ζ1 − c)−1η, ξi dm(ζ) = − ∂g(ζ)(ζ − λ)−1 dm(ζ) dµ(λ) ∂g(ζ)ZK (ζ − λ)−1 dµ(λ) dm(ζ) ∂g(ζ)h(ζ1 − a)−1η, ξi dm(ζ). = − For all ξ ∈ H⊥ the last integrand is 0 since (ζ1 − a)−1η ∈ H, hence g(c)η ∈ H. Thus H is an invariant subspace for g(c) and the usual definition of g(a), namely g(a) := g(c)H (see [12, p. 85]), is compatible with (7.7). To justify the interchange of order of integration in the above computation, let M = supζ∈C ∂g(ζ) and let R be a constant larger than the diameter of the set H − K, so that for each λ ∈ K the disc D(λ, R) with the center λ and radius R contains H. Introduce the polar coordinates by ζ = λ + reiφ. Then by the Fubini-Tonelli theorem ZHZK ∂g(ζ)ζ − λ−1 dµ(λ) dm(ζ) ≤ MZHZK ζ − λ−1 dµ(λ) dm(ζ) = MZKZH ζ − λ−1dm (ζ) dµ(λ) ≤ MZKZD(λ,R) ζ − λ−1 dm(ζ) dµ(λ) dr dφ dµ = 2πM Rµ(K) < ∞. = MZKZ 2π 0 Z R 0 Now, if a and g are as in Lemma 7.4 and if b = g(a), we may compute formally (cid:3) for each x ∈ B(H) [b, x] = [g(a), x] = − ∂g(ζ)[(ζ1 − a)−1, x] dm(ζ) ∂g(ζ)(ζ1 − a)−1[a, x](ζ1 − a)−1 dm(ζ) = [a, Ta,g(x)], 1 π ZC\σ(a) = − 1 π ZC\σ(a) where (7.8) Ta,g(x) := − 1 π ZC\σ(a) ∂g(ζ)(ζ1 − a)−1x(ζ1 − a)−1 dm(ζ). 26 BOJAN MAGAJNA The problem here is, of course, the existence of the integral in (7.8). We have to show that the map (7.9) (η, ξ) 7→ − 1 π ZC\σ(a) ∂g(ζ)hx(ζ1 − a)−1η, (ζ1 − a∗)−1ξi dm(ζ) is a bounded sesquilinear form on H. The following lemma will be helpful. Lemma 7.5. Let a, g and K := σ(a) be as in Lemma 7.4. If (7.10) κ := sup then the sesquilinear form defined by (7.9) is bounded by 2 λ∈KZC\K ∂gζ − λ−2 dm(ζ) < ∞, πkxkκ. 1 2 Proof. For any t > 0, using first the Schwarz inequality and then the inequality αβ ≤ 1 2 (t2α2 + t−2β2) (α, β ≥ 0) to estimate the inner product in the integral in (7.9), we see that the integral in (7.9) is dominated by kxkZK c ∂g(ζ)k(ζ1 − a)−1ηkk(ζ1 − a∗)−1ξk dm(ζ) ≤ [t2ZK c ∂g(ζ)k(ζ1 − a)−1ηk2 dm(ζ) + t−2ZK c ∂g(ζ)k(ζ1 − a)−1ξk2 dm(ζ)]. kxk Using the notation from the proof of Lemma 7.4 (with µ(·) := he(·)ξ, ξi) and (7.10), we have ZC\K ∂g(ζ)k(ζ1 − a)−1ξk2 dm(ζ) =ZH ∂g(ζ)ZK ζ − λ−2 dµ(λ) dm(ζ) =ZKZH ∂g(ζ)ζ − λ−2 dm(ζ) dµ(λ) ≤ κµ(K) = κkξk2. Since a similar estimate holds with η in place of ξ, it follows that ZH ∂g(ζ)k(ζ1 − a)−1ηkk(ζ1 − a∗)−1ξk dm(ζ) ≤ κ(t2kηk2 + t−2kξk2). Taking the infimum over all t > 0 we get 1 π ZH ∂g(ζ)kt(ζ1 − a)−1ηkkt−1(ζ1 − a∗)−1ξk dm(ζ) ≤ 2 π κkηkkξk. (cid:3) Remark 7.6. Lemma 7.5 applies, for example, if ∂g is a Lipschitz function of order α, that is ∂g(ζ)− ∂g(ζ0) ≤ βζ − ζ0α (ζ, ζ0 ∈ C) for some positive constants α and β, with ∂gK = 0. In this case the integral (7.10) may be estimated by noting that the Lipschitz condition (together with ∂gK = 0) implies that ∂g(ζ) ≤ βδ(ζ, K)α, where δ(ζ, K) is the distance from ζ to K. Let R > 0 be so large that for each λ ∈ K the closed dics D(λ, R) with the center λ and radius R contains H, where H is as in the proof of Lemma 7.4. Introducing the polar coordinates by ζ = λ + reiφ, for each λ ∈ K we have ZC\K ∂g(ζ)ζ − λ−2 dm(ζ) ≤ βZH ζ − λ2 dm(ζ) ≤ βZD(λ,R) ζ − λα−2 dm(ζ) δ(ζ, K)α = 2πβα−1Rα. 27 ζ→ζ0, ζ∈σ(a) ζ − ζ0 Definition 7.7. A function f on a compact subset K ⊆ C is in the class L(1+α, K) (where α ∈ (0, 1]) if the limit (7.11) f (ζ) − f (ζ0) f ′(ζ0) = lim exists for each (nonisolated) ζ0 ∈ K and if there exists a constant κ > 0 such that (7.12) f (ζ) − f (ζ0) − f ′(ζ0)(ζ − ζ0) ≤ κζ − ζ01+α and (7.13) for all ζ, ζ0 ∈ K. f ′(ζ) − f ′(ζ0) ≤ κζ − ζ0α We need the following consequence of the Whitney extension theorem. Lemma 7.8. Each f ∈ L(1 + α, K) can be extended to a continuously differentiable function g on C with compact support such that ∂g is a Lipschitz function of order α and ∂g(ζ) = 0 if ζ ∈ K (even though K may have empty interior). Proof. It suffices to extend f to a differentiable function g with ∂g and ∂g Lipschitz of order α and ∂gK = 0, for then we simply replace g by hg, where h is a smooth function (that is, has continuous partial derivatives of all orders) with compact support which is equal to 1 on K. (Namely, gh then has compact support and ∂(hg) and ∂(hg) are easily seen to be Lipschitz of order α with ∂(hg)K = (∂hg + h∂g)K = 0 since hK = 1.) Let ζ = x+ iy, f = f1 + if2 and f ′(ζ) = h1(ζ)+ ih2(ζ), where f1, f2 and h1, h2 are real valued functions on K. It follows from (7.12) and (7.13) that for any ζ, ζ0 ∈ K f1(ζ) = f1(ζ0) + h1(ζ0)(x − x0) − h2(ζ0)(y − y0) + R(ζ, ζ1) and hj(ζ) = hj(ζ0) + Rj(ζ, ζ0) (j = 1, 2), where R and Rj are functions satisfying R(ζ, ζ1) ≤ κζ − ζ01+α and Rj(ζ, ζ0) ≤ κζ − ζ0α. By the Whitney extension theorem [35, p. 177] f1 can be extended to a differentiable function g1 on C such that the partial derivatives of g1 are Lipschitz of order α and (7.14) ∂g1 ∂x = h1, ∂g1 ∂y = −h2 on K. Similarly f2 can be extended to an appropriate function g2 such that (7.15) ∂g2 ∂x = h2, ∂g2 ∂y = h1 on K. Then g := g1 + ig2 is a required extension of f since (7.14) and (7.15) imply that ∂g = 0 on K. (cid:3) In all of the above discussion in this subsection we may replace the operator a by a(∞) acting on H∞, which implies that the map Ta,g defined by (7.8) is completely bounded. Taking in (7.9) ξ and η to be in H∞, we see that (7.16) hTa,g(x), ρi = − ∂g(ζ)hx, (ζ1 − a)−1ρ(ζ1 − a)−1i dm(ζ) = hx, (Ta,g)♯(ρ)i 1 π ZC\σ(a) 28 BOJAN MAGAJNA for each ρ = η ⊗ ξ∗ in the predual of B(H), where (Ta,g)♯(ρ) = − π ZC\σ(a) = − 1 1 π ZC\σ(a) ∂g(ζ)(ζ1 − a)−1ρ(ζ1 − a)−1 dm(ζ) ∂g(ζ)(ζ1 − a)−1η ⊗ ((ζ1 − a∗)−1)ξ)∗ dm(ζ). A similar computation as in the proof of Lemma 7.5 shows that the last integral exists and that k(Ta,g)♯(ρ)k ≤ const.kρk. Therefore we conclude that Ta,g is weak* continuous. Further, if S is any weak* continuous (a)′-bimodule endomorphism of B(H), then S commutes in particular with multiplications by (ζ1− a)−1 and, using (7.16), it follows that S commutes with Ta,g. Collecting all the above results, we have proved the following theorem. Theorem 7.9. For a subnormal operator a ∈ B(H) and a function f ∈ L(1 + α, σ(a)) (α ∈ (0, 1]) let g be the extension of f as in Lemma 7.8. Then the map Ta,g defined by (7.8) is a central element in the algebra of all normal completely bounded (a)′-bimodule endomorphisms of B(H) such that [f (a), x] = [a, Ta,g(x)] = Ta,g([a, x]) for all x ∈ B(H). In particular the range of df (a) is contained in the range of da and (7.1) holds. Now we are going to show that if σ(a) is nice enough, then the Lipschitz type condition on f in Theorem 7.9 can be relaxed: f only needs to be a Schur function. First suppose that σ(a) is the closed unit disc D. For each r ∈ (0, 1) let fr(ζ) = f (rζ). Thus each fr is a holomorphic function on a neighborhood Ωr of D and fr(a) can be expressed as f (a) = 1 fr(ζ)(ζ1 − a)−1 dζ, where Γr is a contour in Ωr surrounding σ(a) once in a positive direction. Then for each x ∈ B(H) we have 2πiRΓr f (rζ)[(ζ−a)−1, x] dζ = 1 2πiZΓr f (rζ)(ζ−a)−1[a, x](ζ−a)−1 dζ, 1 2πiZΓr [fr(a), x] = hence (7.17) where [fr(a), x] = Tr([a, x]) and similarly [fr(a), x] = [a, Tr(x)] Tr(x) = 1 2πiZΓr f (rζ)(ζ − a)−1x(ζ − a)−1 dζ. If the set of completely bounded maps Tr on B(H) (0 < r < 1) is bounded, then it has a limit point, say T , in the weak* topology (which the space of all completely bounded maps on B(H) carries as a dual space, see e.g. [6, 1.5.14 (4)]). T commutes with left and right multiplications by elements of (a)′ (since all Tr do). Since f is continuous, kfr(a) − f (a)k r→1−→ 0 (this holds already if a is replaced by its minimal normal extension). Then from (7.17) we see that [f (a), x] = T ([a, x]), hence [f (a), x] = T ([a, x]) = [a, T x] (x ∈ B(H)). These equalities hold also for a(n) in place of a and for x ∈ Mn(B(H)), and (7.1) is also a consequence of [f (a), x] = T ([a, x]). To estimate the norms of the maps Tr, let c on K ⊇ H be the unitary power dilation of a (so that an = pcnH for all n ∈ N, where p is the orthogonal projec- tion from K onto H, see e.g. [17] or [28]). Let Sr be the map on B(K) defined by Sr(x) = 1 f (rζ)(ζ − c)−1x(ζ − c)−1 dζ. Then (since ζ − a is invertible if 2πiRΓr 29 rλi−rλj ζ ∈ Γr) Tr(x) = pSr(x)H for each x ∈ B(H), where x is regarded as an operator on K by setting xH⊥ = 0. Hence kTrk ≤ kSrk. In the special case when c is diagonal (relative to some orthonormal basis of K) with eigenvalues λi and x = [xi,j], a sim- ple computation shows that Sr(x) is represented by the matrix r[ f (rλi)−f (rλj ) xi,j] (where the quotient is taken to be f ′(rλj ) if λi = λj ). Hence in this case kSrk ≤ κ since f is a Schur function. Since any normal operator c can be approximated uniformly by diagonal operators, it follows from the formula defining Sr that the same estimate must hold for all such c with σ(c) ⊆ D. A similar reasoning applies also to the completely bounded norm, hence it follows that sup0<r<1 kTrkcb < ∞. Let us now consider the case when σ(a) is the closure of its interior U and U is simply connected. Let h be a conformal bijection from D onto U . If the boundary ∂σ(a) of σ(a) is sufficiently nice, say a Jordan curve of class C3, then h can be extended to a bijection, denoted again by h, from D onto U = σ(a), such that h and h−1 are in the class C2 [23, 5.2.4]. Then by Theorem 7.9 and Proposition 7.1 h and h−1 are Schur functions. Let a0 = h−1(a). Note that {a0, 1} generates the same Banach algebra as {a, 1} since h and h−1 can both be uniformly approximated by polynomials (by Mergelyan's theorem). For any Schur function f on σ(a) the composition f0 := f ◦ h is a Schur function on D. (To see this, note that for any λ 6= µ in D we may write f (h(λ))−f (h(µ)) and that the inequality k[xi,j yi,j]kS ≤ k[xi,j ]kSk[yi,j]kS holds for the Schur norm of the Schur product of two matrices.) Note that f (a) = f0(a0) and (a0)′ = (a)′. By the previous paragraph there exists a completely bounded (a0)′-bimodule map T on B(H) such that [f0(a0), x] = [a0, T x] = T ([a0, x]), hence (using a0 = h−1(a) and f0(a0) = f (a)) = f (h(λ))−f (h(µ)) h(λ)−h(µ) λ−µ h(λ)−h(µ) λ−µ (7.18) [h−1(a), T x] = [f (a), x] = T ([h−1(a), x]) for all x ∈ B(H). Now the map T is not a priori normal, but it can be replaced by its normal part Tn in (7.18), hence we may achieve that T is normal. (Namely, let T = Tn + Ts be the decomposition of T into its normal and singular part [36, III.2.15]. This decomposition has similar properties as in the special case of linear functionals [21, 10.1.15]. Then the first equality in (7.18) can be rewritten as [h−1(a), Tnx] − [f (a), x] = −[h−1(a), Tsx]. Since the left side of the last equality is a normal function of x, while the right side is singular, both must be 0. This shows that T can be replaced by Tn in the first equality of (7.18) and a similar argument applies also to the second equality.) By Theorem 7.9 there exists a completely bounded (a)′-bimodule map S on B(H) such that (7.19) and S commutes with all normal (a)′-bimodule maps on B(H) (in particular with T ). From the first equality in (7.18) and in (7.19) (with y = T x) we have now [f (a), x] = [h−1(a), T x] = [a, ST x], while from the remaining two equalities in (7.18) and (7.19) (with y = x) we deduce that [f (a), x] = T ([h−1(a), x]) = T S([a, x]) for all x ∈ B(H). Denoting Ta,f = T S = ST , we have deduced the following theo- rem. [a, Sy] = [h−1(a), y] = S([a, y]) for all y ∈ B(H) Theorem 7.10. For a subnormal a ∈ B(H) suppose that σ(a) is the closure of a simply connected domain bounded by a Jordan curve of class C3. Then for each Schur function f on σ(a) there exists a (normal) completely bounded (a)′-bimodule 30 BOJAN MAGAJNA map Ta,f on B(H) such that [f (a), x] = [a, Ta,f (x)] = Ta,f ([a, x]) for all x ∈ B(H). (In particular the inequality (5.1) holds for b = f (a) with κ = kTa,fk.) In general, the Lipschitz type condition in Theorem 7.9 can be replaced by a similar, but less restrictive condition, which involves a regular modulus of continuity 0 ω(r)/r dr < ∞. (There exists an appropriate version of Whitney's extension theorem [35, p. 194].) But probably even this is too restrictive, for we do not need any requirements about ∂g of the extension g (only requirements about ∂g). ω in the sense of [35, p. 175] (instead of just ω(t) = tα) such thatR 1 References [1] M. B. Abrahamse, Toeplitz Operators in Multiply Connected Regions, Amer. J. Math. Vol. 96 (1974), 261 -- 297. [2] A. B. Aleksandrov, V. V. Peller, D. Potapov and F. Sukochev, Functions of normal operators under perturbations, Adv. Math. 226 (2011), 5216 -- 5251. [3] P. Ara and M. Mathieu, Local multipliers of C∗-algebras, Springer Monographs in Math., Springer-Verlag, Berlin, 2003. [4] W. B. Arveson, Notes on extensions of C∗-algebras, Duke. Math. J. 44 (1977), 329 -- 355. [5] W. B. Arveson, Subalgebras of C∗-algebras II, Acta Mathematica 128 (1972), 271 -- 308. [6] D. P. Blecher and C. Le Merdy, Operator Algebras and their Modules, L.M.S. Monographs, New Series 30, Clarendon Press, Oxford, 2004. [7] A. Brown and C. Pearcy, Structure of Commutators of Operators, Annals of Math. 82 (1965), 112 -- 127. [8] M. Bresar, The range and kernel inclusion of algebraic derivations and commuting maps, Quart. J. Math. 56 (2005), 31 -- 41. [9] M. Bresar, B. Magajna and S. Spenko, Identifying derivations through the spectra of their values, Integral Equations and Operator Theory 73 (2012), 395 -- 411. [10] J. B. Conway, A course in Functional Analysis, GTM 96, Springer, Berlin, 1985. [11] J. B. Conway, A Course in Operator Theory, GSM 21, Amer. Math. Soc., Providence, RI, 2000. [12] J. B. Conway, The Theory of Subnormal Operators, Mathematical Surveys and Monographs 36, Amer. Math. Soc., Providence, R.I., 1991. [13] E. M. Dyn'kin, An operator calculus based on the Cauchy-Green formula, (Russian) Inves- tigations on linear operators and the theory of functions, III. Zap. Naucn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI) 30 (1972), 33 -- 39. [14] G. A. Elliott and L. Zsido, Almost uniformly continuous automorphism groups of operator algebras, J. Operator Theory 8 (1982), 227 -- 277. [15] C. K. Fong, Range inclusion for normal derivations, Glasgow Math. J. 25 (1984), 255 -- 262. [16] J. Glimm, A Stone-Weierstrass theorem for C∗-algebras, Ann. Math. 72 (1960), 216 -- 244. [17] P. Halmos, A Hilbert Space Problem Book, GTM 19, Springer -- Verlag, New York 1982. [18] H. Halpern, Irreducible module homomorphisms of a von Neumann algebra into its center, Trans. Amer. Math. Soc. 140 (1969), 195 -- 221. [19] Y. Ho, Commutants and derivation ranges, Tohoku Math. J. 27 (1975), 509 -- 514. [20] B. E. Johnson and J.P. Williams, The range of a normal derivation, Pacific J. Math. 58 (1975), 105 -- 122. [21] R. V. Kadison and J. R. Ringrose, Fundamentals of the theory of operator algebras, Vols. 1 and 2, Academic Press, London, 1983 and 1986. [22] E. Kissin and V. S. Shulman, On the range inclusion of normal derivations: variations on a theme by Johnson, Williams and Fong, Proc. London Math. Soc. 83 (2001), 176 -- 198. [23] S. G. Krantz, Geometric Function Theory: Explorations in Complex Analysis, Birckhauser, Boston, 2006. [24] B. Magajna, The Haagerup norm on the tensor product of operator modules, J. Funct. Anal. 129 (1995), 325 -- 348. [25] B. Magajna, Bicommutants and ranges of derivations, Lin. and Multilin. Alg. 61 (2013), 1161 -- 1180; a Corrigendum in 62 (2014), 1272-1273. 31 [26] R. Meshulam and P. Semrl, Locally linearly dependent operators and reflexivity of operator spaces, Lin. Alg. Appl. 383 (2004), 143 -- 150. [27] G. J. Murphy, Toeplitz operators on generalized H 2 spaces, Integral Eq. Op. Th. 15 (1992), 825 -- 852. [28] V. I. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge Studies in Ad- vanced Mathematics 78, Cambridge University Press, Cambridge, 2002. [29] G. K. Pedersen, C∗-Algebras and Their Automorphism Groups, Academic Press, London, 1979. [30] V. V. Peller, Hankel operators in the perturbation theory of unitary and self-adjoint operators, Funct. Anal. Appl. 19 (1985) 111 -- 123. [31] T. Ransford, Potential Theory in the Complex Plane, LMS Student Texts 28, Cambridge Univ. Press, Cambridge, 1995. [32] W. Rudin, Real and Complex Analysis, McGraw-Hill Publ. Comp., New York, 1987. [33] R. R. Smith, Completely bounded module maps and the Haagerup tensor product, J. Funct. Anal. 102 (1991), 156 -- 175. [34] J. G. Stampfli, The norm of a derivation, Pacific J. Math. 33 (1970), 737 -- 747. [35] E. M. Stein, Singular Integrals and Differentiability Properties of Functions, Princeton Uni- versity Press, Princeton, 1970. [36] M. Takesaki, Theory of Operator Algebras I, EMS 124, Springer-Verlag, Berlin, 2002. [37] T. R. Turner, Double commutants of algebraic operators, Proc. Amer. Math. Soc. 33 (1972), 415 -- 419. [38] D. V. Voiculescu, A non-commutative Weyl-von Neumann theorem, Rev. Roum. Math. Pures Appl. 21 (1976), 97 -- 113. [39] D. V. Voiculescu, Some results on norm-ideal perturbations of Hilbert space operators, J. Operator Theory 2 (1979), 3 -- 37. [40] J. P. Williams, On the range of a derivation II, Proc. Roy. Irish Acad. Sec. A 74 (1974), 299 -- 310. [41] L. R. Williams, Quasisimilarity and hyponormal operators, J. Operator Theory 5 (1981), 127 -- 139. Department of Mathematics, University of Ljubljana, Jadranska 21, Ljubljana 1000, Slovenia E-mail address: [email protected]
1206.0913
3
1206
2012-08-28T11:27:26
Uniform families of ergodic operator nets
[ "math.FA", "math.DS" ]
We study mean ergodicity in amenable operator semigroups and establish the connection to the convergence of strong and weak ergodic nets. We then use these results in order to show the convergence of uniform families of ergodic nets that appear in topological Wiener-Wintner theorems.
math.FA
math
UNIFORM FAMILIES OF ERGODIC OPERATOR NETS MARCO SCHREIBER Abstract. We study mean ergodicity in amenable operator semigroups and establish the connection to the convergence of strong and weak ergodic nets. We then use these results in order to show the convergence of uniform families of ergodic nets that appear in topological Wiener-Wintner theorems. N PN −1 The classical mean ergodic theorem (see [10, Chapter 2.1]) is concerned with the convergence n=0 Sn for some power bounded operator S on a Banach space X. of the Cesàro means 1 The natural extension of the Cesàro means for representations S of general semigroups is the notion of an ergodic net as introduced by Eberlein [7] and Sato [18]. In the first part of this paper we discuss and slightly modify this concept in order to adapt it better for the study of operator semigroups. Sato showed in [18] that in amenable semigroups there always exist weak ergodic nets. We extend this result and show that even strong ergodic nets exist. Using this fact we then state a mean ergodic theorem connecting the convergence of strong and weak ergodic nets and the existence of a zero element in the closed convex hull of S. In the second part we develop the adequate framework for investigating uniform convergence in so-called topological Wiener-Wintner theorems. In the simplest situation these theorems deal with the convergence of averages n=0 λnSn for some operator S on spaces C(K) and λ in the unit circle T. Assani [2] and Robinson [16] asked when this convergence is uniform in λ ∈ T. Subsequently, their results have been generalised in different ways by Walters [19], Santos and Walkden [17] and Lenz [11, 12]. We propose and study uniform families of ergodic nets as an appropriate concept for unifying and generalizing these and other results. 1 N PN −1 1. Amenable and mean ergodic operator semigroups We start from a semitopological semigroup G and refer to Berglund et al. [3, Chapter 1.3] for an introduction to this theory. Let X be a Banach space and denote by L (X) the set of bounded linear operators on X. We further assume that S = {Sg : g ∈ G} is a bounded representation of G on X, i.e., (i) Sg ∈ L (X) for all g ∈ G and supg∈G kSgk < ∞, (ii) SgSh = Shg for all g, h ∈ G, Date: September 26, 2018. Key words and phrases. amenable semigroups, mean ergodic theorem, topological Wiener-Wintner theo- rems. 1 2 MARCO SCHREIBER (iii) g 7→ Sgx is continuous for all x ∈ X. For a bounded representation S we denote by co S its convex hull and by coS the closure with respect to the strong operator topology. Notice that S as well as co S and coS are topological semigroups with respect to the strong and semitopological semigroups with respect to the weak operator topology. A mean on the space Cb(G) of bounded continuous functions on G is a linear functional m ∈ Cb(G)′ satisfying hm, 1i = kmk = 1. A mean m ∈ Cb(G)′ is called right (left) invariant if hm, Rgf i = hm, f i (hm, Lgf i = hm, f i) ∀g ∈ G, f ∈ Cb(G), where Rgf (h) = f (hg) and Lgf (h) = f (gh) for h ∈ G. A mean m ∈ Cb(G)′ is called invariant if it is both right and left invariant. The semigroup G is called right (left) amenable if there exists a right (left) invariant mean on Cb(G). It is called amenable if there exists an invariant mean on Cb(G) (see Berglund et al. [3, Chapter 2.3] or the survey article of Day [5]). For simplicity, we shall restrict ourselves to right amenable and amenable semigroups, although most of the results also hold for left amenable semigroups. Notice that if S := {Sg : g ∈ G} is a bounded representation of a (right) amenable semigroup G on X, then S endowed with the strong as well as the weak operator topology is also (right) amenable. Indeed, if m ∈ Cb(G)′ is a (right) invariant mean on Cb(G), then m ∈ Cb(S)′ given by hm, f i := h m, f i (f ∈ Cb(S)) defines a (right) invariant mean on Cb(S), where f (g) = f (Sg). In the following, the space L (X) will be endowed with the strong operator topology unless stated otherwise. Definition 1.1. A net (AS net if the following conditions hold. α)α∈A of operators in L (X) is called a strong right (left) S-ergodic (1) AS (2) (AS α ∈ coS for all α ∈ A. α) is strongly right (left) asymptotically invariant, i.e., limα AS αx − AS αSgx = 0 (cid:0)limα AS αx − SgAS αx = 0(cid:1) for all x ∈ X and g ∈ G. α) is called a weak right (left) S-ergodic net if (AS The net (AS α) is weakly right (left) asymp- totically invariant, i.e., if the limit in (2) is taken with respect to the weak topology σ(X, X ′) α) is called a strong (weak) S-ergodic net if it is a strong (weak) right and on X. The net (AS left S-ergodic net. We note that our definition differs slightly from that of Eberlein [7], Sato [18] and Krengel [10, Chapter 2.2]. Instead of condition (1) they require only (1') AS αx ∈ coSx for all α ∈ A and x ∈ X. UNIFORM FAMILIES OF ERGODIC OPERATOR NETS 3 However, the existence of (even strong) ergodic nets in the sense of Definition 1.1 is ensured by Corollary 1.5. Moreover, both definitions lead to the same convergence results (see The- orem 1.7 below). The reason is that if the limit P x := limα AS αx exists for all x ∈ X, then the operator P satisfies P ∈ coS rather than only P x ∈ coSx for all x ∈ X (see Nagel [14, Theorem 1.2]). Here are some typical examples of ergodic nets. Examples 1.2. (a) Let S ∈ L (X) with kSk ≤ 1 and consider the representation S = N )N ∈N given {Sn : n ∈ N} of the semigroup (N, +) on X. Then the Cesàro means (AS by AS N := 1 N N −1Xn=0 Sn form a strong S-ergodic net. (b) In the situation of (a), the Abel means (AS r )0<r<1 given by AS r := (1 − r) ∞Xn=0 rnSn form a strong S-ergodic net. (c) Consider the semigroup (R+, +) being represented on X by a bounded C0-semigroup S = {S(t) : t ∈ R+}. Then (AS s )s∈R+ given by 1 sZ s 0 AS s x := S(t)x dt (x ∈ X) is a strong S-ergodic net. (d) Let S = {Sg : g ∈ G} be a bounded representation on X of an abelian semigroup G. Order the elements of co S by setting U ≤ V if there exists W ∈ co S such that V = W U. Then (AS U )U ∈co S given by AS U := U is a strong S-ergodic net. (e) Let H be a locally compact group with left Haar measure · and let G ⊂ H be a subsemigroup. Suppose that there exists a Følner net (Fα)α∈A in G (see [15, Chapter 4]), i.e., a net of compact sets such that Fα > 0 for all α ∈ A and lim α gFα∆Fα Fα = 0 ∀g ∈ G, where A∆B := (A \ B) ∪ (B \ A) denotes the symmetric difference of two sets A and B. Suppose that S := {Sg : g ∈ G} is a bounded representation of G on X. Then (AS α)α∈A given by AS αx := is a strong right S-ergodic net. 1 FαZFα Sgx dg (x ∈ X) 4 MARCO SCHREIBER If G is a right amenable group in the situation of Example 1.2 (e), then there always exists a Følner net (Fα)α∈A in G (see [15, Theorem 4.10]). Hence, in right amenable groups there always exist strong right S-ergodic nets for each representation S. In [18, Proposition 1], Sato showed the existence of weak right (left) ergodic nets in right (left) amenable operator semigroups. We give a proof for the case of an amenable semigroup. The one-sided case is analogous. Proposition 1.3. Let G be represented on X by a bounded (right) amenable semigroup S = {Sg : g ∈ G}. Then there exists a weak (right) S-ergodic net in L (X). Proof. Let m ∈ Cb(S)′ be an invariant mean. Denote by B the closed unit ball of Cb(S)′ and by ex B the set of extremal points of B. Since m is a mean, we have m ∈ B = co ex B by the Krein-Milman theorem, where the closure is taken with respect to the weak∗-topology. Since )α∈A ⊂ co{δSg : g ∈ G} i=1 λi,αδSgi = m in the weak∗-topology. Since m is invariant, we obtain ex B = {δSg : g ∈ G}, this implies that there exists a net (PNα with limαPNα NαXi=1 (f − RSg f ) = lim α i=1 λi,αδSgi λi,αδSgi λi,αδSgi lim α (f − LSg f ) = 0 ∀g ∈ G, f ∈ Cb(S). Define the net (AS α)α∈A is weakly asymptotically invariant, let x ∈ X and x′ ∈ X ′ and define fx,x′ ∈ Cb(S) by fx,x′(Sg) := hSgx, x′i for g ∈ G. Then for all g ∈ G we have i=1 λi,αSgi ∈ co S for α ∈ A. To see that (AS α)α∈A by AS NαXi=1 α := PNα (cid:10)AS αx − AS αSgx, x′(cid:11) = = = λi,α(cid:0)(cid:10)Sgix, x′(cid:11) −(cid:10)SgiSgx, x′(cid:11)(cid:1) λi,α(fx,x′(Sgi) − RSg fx,x′(Sgi)) λi,αδSgi (fx,x′ − RSg fx,x′) −→ 0 NαXi=1 NαXi=1 NαXi=1 and αx − SgAS (cid:10)AS αx, x′(cid:11) −→ 0 analogously. Hence (AS α)α∈A is a weak S-ergodic net. (cid:3) We now show that the existence of weak ergodic nets actually implies the existence of strong ergodic nets. Theorem 1.4. Let G be represented on X by a bounded semigroup S = {Sg : g ∈ G}. Then the following assertions are equivalent. (1) There exists a weak (right) S-ergodic net. (2) There exists a strong (right) S-ergodic net. Proof. We give a proof for the case of an S-ergodic net. The one-sided case can be shown in a similar way. UNIFORM FAMILIES OF ERGODIC OPERATOR NETS 5 (1)⇒(2): Consider the locally convex space E :=Q(g,x)∈G×X X ×X endowed with the product topology, where X × X carries the product (norm-)topology. Define the linear map Φ : L (X) → E, Φ(T ) = (T Sgx − T x, SgT x − T x)(g,x)∈G×X . By 17.13(iii) in [9] the weak topology σ(E, E′) on the product E coincides with the product α)α∈A is a weak S-ergodic net of the weak topologies of the coordinate spaces. Hence, if (AS σ(E,E ′). on X, then Φ(AS Since the weak and strong closure coincide on the convex set Φ(coS), there exists a net β ) → 0 in the topology of E. By the definition of this topology (BS β xk → 0 for all (g, x) ∈ G × X and hence this means kBS (BS α) → 0 with respect to the weak topology on E and thus 0 ∈ Φ(coS) β )β∈B ⊂ coS with Φ(BS β )β∈B is a strong S-ergodic net. β Sgx − BS β xk → 0 and kSgBS β x − BS (2)⇒(1) is clear. (cid:3) The following corollary is a direct consequence of Proposition 1.3 and Theorem 1.4. Corollary 1.5. Let G be represented on X by a bounded (right) amenable semigroup S = {Sg : g ∈ G}. Then there exists a strong (right) S-ergodic net. If ergodic nets converge we are led to the concept of mean ergodicity. We use the following abstract notion. Definition 1.6. The semigroup S is called mean ergodic if coS contains a zero element P (cf. [3, Chapter 1.1]), called the mean ergodic projection of S. Notice that for P being a zero element of coS it suffices that P Sg = SgP = P for all g ∈ G. Nagel [14] and Sato [18] studied such semigroups and their results are summarized in Krengel [10, Chapter 2]. In the next theorem we collect a series of properties equivalent to mean ergodicity. Most of them can be found in Krengel [10, Chapter 2, Theorem 1.9], but we give a proof for completeness. Denote the fixed spaces of S and S ′ by Fix S = {x ∈ X : Sgx = x ∀g ∈ G} and Fix S ′ = {x′ ∈ gx′ = x′ ∀g ∈ G} respectively and the linear span of the set rg(I − S) = {x − Sgx : x ∈ X ′ : S′ X, g ∈ G} by lin rg(I − S). Theorem 1.7. Let G be represented on X by a bounded right amenable semigroup S = {Sg : g ∈ G}. Then the following assertions are equivalent. (1) S is mean ergodic with mean ergodic projection P . (2) coSx ∩ Fix S 6= ∅ for all x ∈ X. (3) Fix S separates Fix S ′. (4) X = Fix S ⊕ lin rg(I − S). (5) AS αx has a weak cluster point in Fix S for some/every weak right S-ergodic net (AS α) and all x ∈ X. 6 MARCO SCHREIBER (6) AS αx converges weakly to a fixed point of S for some/every weak right S-ergodic net (AS α) and all x ∈ X. (7) AS (AS αx converges weakly to a fixed point of S for some/every strong right S-ergodic net α) and all x ∈ X. (8) AS (AS αx converges strongly to a fixed point of S for some/every strong right S-ergodic net α) and all x ∈ X. The limit P of the nets (AS mean ergodic projection of S mapping X onto Fix S along lin rg(I − S). α) in the weak and strong operator topology, respectively, is the Proof. (1)⇒(2): Since SgP = P for all g ∈ G, we have P x ∈ coSx ∩ Fix S for all x ∈ X. (2)⇒(3): Let 0 6= x′ ∈ Fix S ′. Take x ∈ X such that hx′, xi 6= 0. If y ∈ coSx ∩ Fix S then we have hx′, yi = hx′, xi 6= 0. Hence Fix S separates Fix S ′. (3)⇒(4): Let (AS α) be a weak right S-ergodic net and let x′ ∈ X ′ vanish on Fix S ⊕lin rg(I −S). Then in particular hx′, yi = hx′, Sgyi = (cid:10)S′ gx′, y(cid:11) for all y ∈ X and g ∈ G. Hence x′ ∈ Fix S ′. Since Fix S separates Fix S ′ and x′ vanishes on Fix S, this implies x′ = 0. Hence Fix S ⊕ lin rg(I − S) is dense in X by the Hahn-Banach theorem and it remains to show αx exists} we obtain D = that Fix S ⊕ lin rg(I − S) is closed. For D := {x ∈ X : σ- limα AS Fix S ⊕ lin rg(I − S) and D is closed since (AS α) is uniformly bounded. α) be any weak right S-ergodic net. Then AS αx converges weakly to a fixed (4)⇒(6): Let (AS point of S for all x ∈ X. Indeed, the convergence on Fix S is clear and the weak convergence α). to 0 on lin rg(I − S) follows from weak right asymptotic invariance and linearity of (AS Since the set {x ∈ X : σ- limα AS αx exists} is closed we obtain weak convergence on all of Fix S ⊕ lin rg(I − S). The limit of the net (AS α) is the projection onto Fix S along lin rg(I − S). αx for every (4)⇒(8): An analogous reasoning as in (4)⇒(6) yields the strong convergence of AS strong right S-ergodic net (AS α) and every x ∈ X. (5)⇒(2): Let (AS of a convergent subnet (AS βx α) be a weak right S-ergodic net, take x ∈ X and define P x as the weak limit x) of (AS αx). Then P x ∈ coSx ∩ Fix S for all x ∈ X. α) be a weak right S-ergodic net. Defining P x as the weak limit of AS (6)⇒(1): Let (AS αx for each x ∈ X we obtain P x ∈ coSx for all x ∈ X. Furthermore, for all x ∈ X and g ∈ G we have P x − P Sgx = σ- limα AS αSgx = 0 and P x − SgP x since P x ∈ Fix S. Hence SgP = P Sg = P for all g ∈ G and [14, Theorem 1.2] yield the mean ergodicity of S. The remaining implications are trivial. αx − AS (cid:3) α) is a convergent S-ergodic net, then the limit P x := αx − αx = 0 for each g ∈ G by the asymptotic left invariance. Hence the following corollary is αx is automatically a fixed point of S for each x ∈ X, since SgP x − P x = limα AS If the semigroup S is amenable and (AS limα AS SgAS a direct consequence of Theorem 1.7. Corollary 1.8. Let G be represented on X by a bounded amenable semigroup S = {Sg : g ∈ G}. Then the following assertions are equivalent. (1) S is mean ergodic with mean ergodic projection P . UNIFORM FAMILIES OF ERGODIC OPERATOR NETS 7 (2) coSx ∩ Fix S 6= ∅ for all x ∈ X. (3) Fix S separates Fix S ′. (4) X = Fix S ⊕ lin rg(I − S). (5) AS (6) AS (7) AS (8) AS αx has a weak cluster point for some/every weak S-ergodic net (AS αx converges weakly for some/every weak S-ergodic net (AS αx converges weakly for some/every strong S-ergodic net (AS αx converges strongly for some/every strong S-ergodic net (AS α) and all x ∈ X. α) and all x ∈ X. α) and all x ∈ X. α) and all x ∈ X. The limit P of the nets (AS mean ergodic projection of S mapping X onto Fix S along lin rg(I − S). α) in the weak and strong operator topology, respectively, is the The next result can be found in Nagel [14, Satz 1.8] (see also Ghaffari [8, Theorem 1]). We give a different proof. Corollary 1.9. Let G be represented on X by a bounded amenable semigroup S = {Sg : g ∈ G}. If S is relatively compact with respect to the weak operator topology, then S is mean ergodic. Proof. Since Sx is relatively weakly compact, we obtain that coSx is weakly compact for α) is a weak S-ergodic net, then all x ∈ X by the Krein-Šmulian Theorem. Hence, if (AS αx has a weak cluster point for each x ∈ X. The mean ergodicity of S then follows from AS Corollary 1.8. (cid:3) If the Banach space satisfies additional geometric properties, contractivity of the semigroup implies amenability and mean ergodicity. For uniformly convex spaces with strictly convex dual unit balls this has been shown by Alaoglu and Birkhoff [1, Theorem 6] using the so-called minimal method. In [4, Theorem 6'] Day observed that the same method still works if uniform convexity is replaced by strict convexity. Corollary 1.10. Let X be a reflexive Banach space such that the unit balls of X and X ′ are strictly convex. If the semigroup G is represented on X by a semigroup of contractions S = {Sg : g ∈ G}, then S is mean ergodic. Proof. If S is a contractive semigroup in L (X) and the unit balls of X and of X ′ are strictly convex, then S is amenable by [6, Corollary 4.14]. Since S is bounded on the reflexive space X, it follows that S is relatively compact with respect to the weak operator topology. Hence Corollary 1.9 implies that S is mean ergodic. (cid:3) In some situations (see e.g. Assani [2, Theorem 2.10], Walters [19, Theorem 4], Lenz [12, Theorem 1]) one is interested in convergence of an ergodic net only on some given x ∈ X. Apart from the implication (2)⇒(3) the following result is a direct consequence of Theorem 1.7 and Corollary 1.8 applied to the restriction SYx of S to the closed S-invariant subspace Yx := linSx. Proposition 1.11. Let G be represented on X by a bounded (right) amenable semigroup S = {Sg : g ∈ G} and let x ∈ X. Then the following assertions are equivalent. 8 MARCO SCHREIBER (1) S is mean ergodic on Yx with mean ergodic projection P . (2) coSx ∩ Fix S 6= ∅. (3) Fix SYx separates Fix S′ Yx . (4) x ∈ Fix S ⊕ lin rg(I − S). (5) AS (AS αx has a weak cluster point (in Fix S) for some/every weak (right) S-ergodic net α). (6) AS αx converges weakly (to a fixed point of S) for some/every weak (right) S-ergodic net (AS α). (7) AS αx converges weakly (to a fixed point of S) for some/every strong (right) S-ergodic net (AS α). (8) AS αx converges strongly (to a fixed point of S) for some/every strong (right) S-ergodic net (AS α). The limit P of the nets AS mean ergodic projection of SYx mapping Yx onto Fix SYx along lin rg(I − SYx). α in the weak and strong operator topology on Yx, respectively, is the and take y ∈ Yx such that hx′, yi 6= 0. Since x Proof. (2)⇒(3): Let 0 6= x′ ∈ Fix S′ generates the space Yx this yields hx′, xi 6= 0. If z ∈ coSx ∩ Fix S, then z ∈ Yx and we have hx′, zi = hx′, xi 6= 0. Hence Fix SYx separates Fix S′ Yx The other implications follow directly from Theorem 1.7 and Corollary 1.8 by noticing that the set {y ∈ X : limα AS (cid:3) αy exists} is a closed S-invariant subspace of X. Yx . 2. Uniform families of ergodic operator nets We now use the above results on mean ergodic semigroups in order to study the convergence of uniform families of ergodic nets. Let I be an index set and suppose that the semigroup G is represented on X by bounded semigroups Si = {Si,g : g ∈ G} for each i ∈ I. Moreover, we assume that the Si are uniformly bounded, i.e., supi∈I supg∈G kSi,gk < ∞. Definition 2.1. Let A be a directed set and let (ASi each i ∈ I. Then {(ASi α )α∈A : i ∈ I} is a uniform family of right (left) ergodic nets if α )α∈A ⊂ L (X) be a net of operators for (1) ∀α ∈ A, ∀ε > 0, ∀x1, . . . , xm ∈ X, ∃g1, . . . , gN ∈ G such that for each i ∈ I there exists j=1 ci,jSi,gj ∈ co Si satisfying j=1ci,jSi,gj xkk < ε ∀k ∈ {1, . . . , m}; a convex combination PN α xk −PN α Si,gxk = 0 (cid:18)lim α x − ASi sup i∈I kASi kASi α (2) lim α sup i∈I α )α∈A : i ∈ I} is called a uniform family of ergodic nets if it is a uniform family The set {(ASi of left and right ergodic nets. sup i∈I kASi α x − Si,gASi α xk = 0(cid:19) ∀g ∈ G, x ∈ X. UNIFORM FAMILIES OF ERGODIC OPERATOR NETS 9 α )α∈A : i ∈ I} is a uniform family of (right) ergodic nets, then each (ASi α )α∈A Notice that if {(ASi is a strong (right) Si-ergodic net. Here are some examples of uniform families of ergodic nets. Proposition 2.2. (a) Let S ∈ L (X) with kSk ≤ 1. Consider the semigroup (N, +) being represented on X by the families Sλ = {(λS)n : n ∈ N} for λ ∈ T. Then n(cid:16) 1 N PN −1 n=0 (λS)n(cid:17)N ∈N : λ ∈ To is a uniform family of ergodic nets. (b) In the situation of (a), is a uniform family of ergodic nets. (cid:8)((1 − r)P∞ n=0 rnλnSn)0<r<1 : λ ∈ T(cid:9) (c) Let K be a compact space and ϕ : K → K a continuous transformation. Let H be a Hilbert space and S : f 7→ f ◦ ϕ the Koopman operator corresponding to ϕ on the space C(K, H) of continuous H-valued functions on K. Denote by U (H) the set of unitary operators on H and by Λ the set of continuous maps γ : K → U (H). Consider the semigroup (N, +) and its representations on C(K, H) given by the families Sγ = {(γS)n : n ∈ N} for γ ∈ Λ, where (γS)f (x) = γ(x)Sf (x) for x ∈ K and f ∈ C(K, H). Then n(cid:16) 1 N PN −1 n=0 (γS)n(cid:17)N ∈N : γ ∈ Λo is a uniform family of ergodic nets. (d) Let (S(t))t∈R+ be a bounded C0-semigroup on X. Consider the semigroup (R+, +) being represented on X by the families Sr = {e2πirtS(t) : t ∈ R+} for r ∈ B, where B ⊂ R is bounded. Then is a uniform family of ergodic nets. n(cid:0) 1 0 e2πirtS(t) dt(cid:1)s∈R+ sR s : r ∈ Bo (e) Let S = {Sg : g ∈ G} be a bounded representation on X of an abelian semigroup G. Order the elements of co S by setting U ≤ V if there exists W ∈ co S such that multiplicative maps G → T, and consider the representations Sχ = {χ(g)Sg : g ∈ G} V = W U . Denote by bG the character semigroup of G, i.e., the set of continuous of G on X for χ ∈ bG. Then (cid:26)(cid:16)PN i=1 ciχ(gi)Sgi(cid:17)PN i=1 ciSgi ∈co S : χ ∈ bG(cid:27) is a uniform family of ergodic nets. (f ) Let H be a locally compact group with left Haar measure · and let G ⊂ H be a subsemigroup. Suppose that there exists a Følner net (Fα)α∈A in G. Suppose that S := {Sg : g ∈ G} is a bounded representation of G on X. Consider the representations Sχ = {χ(g)Sg : g ∈ G} of G on X for χ ∈ Λ, where Λ ⊂ bH is uniformly equicontinuous on compact sets. Then n(cid:16) 1 FαRFα χ(g)Sg dg(cid:17)α∈A : χ ∈ Λo 10 MARCO SCHREIBER is a uniform family of right ergodic nets. Proof. (a) Property (1) of Definition 2.1 is clear. To see (2), let x ∈ X and k ∈ N. Then sup λ∈T (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 N N −1Xn=0 (λS)nx − (λS)n+kx(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ sup λ∈T 1 N k−1Xn=0 k(λS)nxk + 1 N N −1+kXn=N k(λS)nxk ≤ 2k N kxk −−−−→ N→∞ 0. (b) (1): Let 0 < r < 1 and ε > 0. Choose N ∈ N such that rN < ε 2. Then for all λ ∈ T we have k(1 − r) rnλnSn − rnλnSnk (1 − r) (1 − rN ) N −1Xn=0 (1 − r) rnλnSn − (1 − r) ∞Xn=0 ∞Xn=0 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (1 − r) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) rn + (1 − r)(cid:12)(cid:12)(cid:12)(cid:12)1 − ≤ (1 − r) ∞Xn=N ≤ rN + rN < ε. rnλnSn − rnλnSn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N −1Xn=0 N −1Xn=0 (1 − rN )(cid:12)(cid:12)(cid:12)(cid:12) N −1Xn=0 1 + rn (1 − r) (1 − rN ) N −1Xn=0 rnλnSn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (2): Let x ∈ X and k ∈ N. Define for each λ ∈ T the sequence x(λ) by x(λ) := n (λS)nx−(λS)n+kx for n ∈ N. Then it follows from (a) that supλ k 1 n k −→ 0. It is well known that Cesàro convergence implies the convergence of the Abel means to the same limit (see [13, Proposition 2.3]). One checks that if the Cesàro convergence is uniform in λ ∈ T, then the convergence of the Abel means is also uniform. Hence N PN −1 n=0 x(λ) we obtain limr↑1 supλ∈T k(1 − r)P∞ n=0 rnx(λ) n k = 0. (c) (1) is clear. To see (2) let f ∈ C(K, H) and k ∈ N. Then k(γS)f k = sup x∈K kγ(x)f (ϕ(x))kH = sup x∈K kf (ϕ(x))kH ≤ kf k, since γ(x) is unitary for all x ∈ K. Hence sup γ∈Λ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 N N −1Xn=0 (γS)nf − (γS)n+kf(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ sup γ∈Λ 1 N k−1Xn=0 k(γS)nf k + 1 N N −1+kXn=N k(γS)nf k ≤ 2k N kf k −−−−→ N→∞ 0. (d) This is a special case of (f) for the Følner net ([0, s])s>0 in R+ and the set Λ := {χr : R+ → T : r ∈ B}, where χr(t) = e2πirt for t ∈ R+. Notice that Λ ⊂ bR is uniformly equicontinuous on compact sets since B is bounded. UNIFORM FAMILIES OF ERGODIC OPERATOR NETS 11 (e) (1) is clear. To see (2) let x ∈ X, g ∈ G and ε > 0. Choose N ∈ N such that g , n=0 Sn N < ε, where M = supg∈G kSgk. Then for all V = W 1 n=0 Sn g ≥ 1 2M 2kxk N PN −1 N PN −1 i=0 ciSgi ∈ co S, we have where W =Pk−1 N −1Xn=0 sup χ∈ bG (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k−1Xi=0 1 N ciχ(gign)Sgignx − ciχ(gi)Sgi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) k−1Xi=0 M 1 N ≤ sup χ∈ bG ≤ sup χ∈ bG k−1Xi=0 ·(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 N 1 N ciχ(gign)Sgignχ(g)Sgx(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N −1Xn=0 (χ(g)Sg)n(x − χ(g)Sgx)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) N −1Xn=0 2M kxk < ε. 1 N kx − (χ(g)Sg)N xk ≤ M (f) (1): Let α ∈ A, ε > 0 and x1, . . . , xm ∈ X. Since Fα is compact and Λ is uniformly equicontinuous on Fα the family {g 7→ χ(g)Sgxk : χ ∈ Λ} is also uniformly equicon- tinuous on Fα for each k ∈ {1, . . . , m}. Hence for each k ∈ {1, . . . , m} we can choose an open neighbourhood Uk of the unity of H satisfying g, h ∈ G, h−1g ∈ Uk ⇒ sup χ∈Λ kχ(g)Sgxk − χ(h)Sgxkk < ε. 1 FαZFα Hence for all χ ∈ Λ and k ∈ {1, . . . , m} we have k=1 Uk is still an open neighbourhood of unity. Since Fα is compact n=1 gnU. Defining V1 := g1U ∩ Fα and n=1 Vn. Then U := Tm there exists g1, . . . , gN ∈ Fα such that Fα ⊂ SN Vn := (gnU ∩ Fα) \ Vn−1 for n = 2, . . . , N we obtain a disjoint union Fα = SN χ(gn)Sgnxk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13) NXn=1ZVn NXn=1 kχ(g)Sgxk − χ(gn)Sgnxkk χ(g)Sgxkdg− NXn=1 Vnε = ε. Vn Fα {z Fα Fα } dg ≤ 1 1 < <ε (2): If x ∈ X and h ∈ G, then we have 1 FαZFα sup χ∈Λ(cid:13)(cid:13)(cid:13)(cid:13) χ(g)Sgx − χ(hg)Shgx dg(cid:13)(cid:13)(cid:13)(cid:13) ≤ sup χ∈Λ Fα△hFα Fα ≤ 1 FαZFα△hFα kχ(g)Sgxk dg kSgkkxk −→ 0. sup g∈G (cid:3) α )α∈A : i ∈ I} be a uniform family of right ergodic nets. If x ∈ X and Si is right Now, let {(ASi amenable and mean ergodic on linSix for each i ∈ I, then it follows from Proposition 1.11 α x → Pix for each i ∈ I, where Pi denotes the mean ergodic projection of SilinSix. The that ASi 12 MARCO SCHREIBER question arises, when this convergence is uniform in i ∈ I. The following elementary example shows that in general this cannot be expected. Example 2.3. Let X = C and let S = IC ∈ L(C) be the identity operator on C. Consider the semigroup (N, +) being represented on C by the families Sλ = {λnIC : n ∈ N} for λ ∈ T. Then for each λ ∈ T the Cesàro means 1 n=0 λn converge, but the convergence is not uniform in λ ∈ T. N PN −1 However, the following theorem gives a sufficient condition for uniform convergence. Theorem 2.4. Let I be a compact set and let G be represented on X by the uniformly bounded right amenable semigroups Si = {Si,g : g ∈ G} for all i ∈ I. Let {(ASi α )α∈A : i ∈ I} be a uniform family of right ergodic nets. Take x ∈ X and assume that (a) Si is mean ergodic on linSix with mean ergodic projection Pi for each i ∈ I, (b) I → R+, i 7→ kASi α x − Pixk is continuous for each α ∈ A. Then lim α sup i∈I kASi α x − Pixk = 0. For the proof of Theorem 2.4 we need a lemma. Lemma 2.5. If {(ASi α )α∈A : i ∈ I} is a uniform family of right ergodic nets, then lim α sup i∈I kASi α x − ASi α ASi β xk = 0 ∀β ∈ A, x ∈ X. Proof. Let β ∈ A, ε > 0 and x ∈ X. Since {(ASi godic nets, there exists g1, . . . , gN ∈ G and for each i ∈ I a convex combinationPN α )α∈A : i ∈ I} is a uniform family of right er- j=1 ci,jSi,gj ∈ j=1 ci,jSi,gj xk < ε/M, where M = supi∈I supg∈G kSi,gk. Now, co Si such that supi∈I kASi choose α0 ∈ A such that for all α > α0 β x−PN kASi α x − ASi α Si,gj xk < ε ∀j = 1, . . . , N. sup i∈I Then for all α > α0 and i ∈ I we obtain kASi α x − ASi α ASi β xk ≤ kASi α PN α x − ASi j=1 ci,jkASi ≤PN < 2ε. j=1 ci,jSi,gj xk + kASi α PN α Si,gj xk + M kPN α x − ASi α ASi j=1 ci,jSi,gj x − ASi j=1 ci,jSi,gj x − ASi β xk β xk Proof of Theorem 2.4. By Proposition 1.11 and the hypotheses the function fα : I → R+ defined by fα(i) = kASi α x − Pixk is continuous for each α ∈ A and limα fα(i) = 0 for all i ∈ I. Compactness and continuity yield a net (iα)α∈A ⊂ I with supi∈I fα(i) = fα(iα) for all α ∈ A. To show that limα supi∈I fα(i) = 0 it thus suffices to show that every subnet of (fα(iα)) has a subnet converging to 0. Let (fαk (iαk )) be a subnet of (fα(iα)) and let ε > 0. Since I is compact, we can choose a subnet of (iαk ), also denoted by (iαk ), such that iαk → i0 for some i0 ∈ I. Since fα converges pointwise to 0, we can take β ∈ A such that fβ(i0) < ε/M, where (cid:3) UNIFORM FAMILIES OF ERGODIC OPERATOR NETS 13 M = supi∈I supg∈G kSi,gk. By continuity of fβ there exists k1 such that for all k > k1 we have fβ(iαk ) − fβ(i0) < ε/M. By Lemma 2.5 there exists k2 > k1 such that for all k > k2 fαk (iαk ) ≤ kA xk + kA Siαk Siαk αk A β Siαk αk x − A Siαk ≤ ε + M kA β ≤ ε + M (fβ(iαk ) − fβ(i0)) + M fβ(i0) ≤ 3ε. x − Piαk xk Siαk Siαk αk A β x − A Siαk αk Piαk xk Hence limα supi∈I fα(i) = 0. (cid:3) We now apply the above theory to operator semigroups on the space C(K) of complex valued continuous functions on a compact metric space K. Corollary 2.6. Let ϕ : K → K be a continuous map, S : f 7→ f ◦ ϕ the corresponding Koop- man operator on C(K), and assume that there exists a unique ϕ-invariant Borel probability measure µ on K. If f ∈ C(K) satisfies Pλf = 0 for all λ ∈ T, where Pλ denotes the mean ergodic projection of {(λS)n : n ∈ N} on L2(K, µ), then (1) lim N→∞ (2) lim r↑1 sup λ∈T sup λ∈T(cid:13)(cid:13)(cid:13) 1 FN Pn∈FN k(1 − r)P∞ λnSnf(cid:13)(cid:13)(cid:13)∞ n=0 rnλnSnf k∞ = 0. = 0 for each Følner sequence (FN )N ∈N in N, Proof. By the hypotheses it follows from Robinson [16, Theorem 1.1] and Proposition 1.11 that the semigroups Sλ = {(λS)n : n ∈ N} are mean ergodic on linSλf ⊂ C(K) for each λ ∈ T. Let now (FN )N ∈N be a Følner sequence in N and consider the uniform family of ergodic nets λnSn(cid:17)N ∈N : λ ∈ To n(cid:16) 1 FN Pn∈FN FN Pn∈FN FN Pn∈FN (cf. Proposition 2.2 (f)). If Pλf = 0 for all λ ∈ T, then also condition (b) in Theorem 2.4 is satisfied since the map λ 7→ 1 λnSnf is continuous for each N ∈ N. Hence limN→∞ supλ∈T k 1 λnSnf k∞ = 0 by Theorem 2.4. The same reasoning applied to the uniform family of ergodic nets yields the second assertion. (cid:8)((1 − r)P∞ n=0 rnλnSn)0<r<1 : λ ∈ T(cid:9) (cid:3) Remark 2.7. In [2, Theorem 2.10] Assani proved the first assertion of Corollary 2.6 for the Følner sequence FN = {0, . . . , N − 1} in N. Generalisations of this result can be found in Walters [19, Theorem 5], Santos and Walkden [17, Prop. 4.3] and Lenz [12, Theorem 2]. We will systematically study and unify these cases in a subsequent paper. Acknowledgement. The author is grateful to Rainer Nagel for his support, valuable discus- sions and comments. 14 MARCO SCHREIBER References 1. L. Alaoglu and G. Birkhoff, General ergodic theorems, Ann. of Math. (2), 41 (1940), pp. 293 -- 309. 2. I. Assani, Wiener Wintner Ergodic Theorems, World Scientific Publishing Co. Inc., River Edge, NJ, 2003. 3. J. F. Berglund, H. D. Junghenn, and P. Milnes, Analysis on Semigroups, Canadian Mathematical Society Series of Monographs and Advanced Texts, John Wiley & Sons Inc., New York, 1989. 4. M. M. Day, Reflexive Banach spaces not isomorphic to uniformly convex spaces, Bull. Amer. Math. Soc., 47 (1941), pp. 313 -- 317. 5. M. M. Day, Semigroups and amenability, in Semigroups (Proc. Sympos., Wayne State Univ., Detroit, Mich., 1968), Academic Press, New York, 1969, pp. 5 -- 53. 6. K. de Leeuw and I. Glicksberg, Applications of almost periodic compactifications, Acta Math. 105 (1961), pp. 63 -- 97. 7. W. F. Eberlein, Abstract ergodic theorems and weak almost periodic functions, Trans. Amer. Math. Soc. 67 (1949), pp. 217 -- 240. 8. A. Ghaffari, Ergodic theory of amenable semigroup actions, Proc. Indian Acad. Sci. Math. Sci. 117 (2007), pp. 177 -- 183. 9. J. L. Kelley and I. Namioka, Linear Topological Spaces, D. Van Nostrand Co., Inc., Princeton, N.J., 1963. 10. U. Krengel, Ergodic Theorems, vol. 6 of de Gruyter Studies in Mathematics, Walter de Gruyter & Co., Berlin, 1985. 11. D. Lenz, Aperiodic order via dynamical systems: diffraction for sets of finite local complexity, in Ergodic Theory, vol. 485 of Contemp. Math., Amer. Math. Soc., Providence, RI, 2009, pp. 91 -- 112. 12. , Continuity of eigenfunctions of uniquely ergodic dynamical systems and intensity of Bragg peaks, Comm. Math. Phys. 287 (2009), pp. 225 -- 258. 13. Y.-C. Li, R. Sato, and S.-Y. Shaw, Convergence theorems and Tauberian theorems for functions and sequences in Banach spaces and Banach lattices, Israel J. Math. 162 (2007), pp. 109 -- 149. 14. R. J. Nagel, Mittelergodische Halbgruppen linearer Operatoren, Ann. Inst. Fourier (Grenoble) 23 (1973), pp. 75 -- 87. 15. A. L. T. Paterson, Amenability, vol. 29 of Mathematical Surveys and Monographs, American Mathe- matical Society, Providence, RI, 1988. 16. E. A. Robinson, Jr., On uniform convergence in the Wiener-Wintner theorem, J. London Math. Soc. 49 (1994), pp. 493 -- 501. 17. S. I. Santos and C. Walkden, Topological Wiener-Wintner ergodic theorems via non-abelian Lie group extensions, Ergodic Theory Dynam. Systems 27 (2007), pp. 1633 -- 1650. 18. R. Sat ¯o, On abstract mean ergodic theorems, Tôhoku Math. J. 30 (1978), pp. 575 -- 581. 19. P. Walters, Topological Wiener-Wintner ergodic theorems and a random L2 ergodic theorem, Ergodic Theory Dynam. Systems 16 (1996), pp. 179 -- 206. Institute of Mathematics, University of Tübingen, Auf der Morgenstelle 10, 72076 Tübingen, Germany E-mail address: [email protected]
1305.6691
1
1305
2013-05-29T04:39:02
Sum theorems for maximally monotone operators of type (FPV)
[ "math.FA" ]
The most important open problem in Monotone Operator Theory concerns the maximal monotonicity of the sum of two maximally monotone operators provided that the classical Rockafellar's constraint qualification holds. In this paper, we establish the maximal monotonicity of $A+B$ provided that $A$ and $B$ are maximally monotone operators such that $\sta(\dom A)\cap\inte\dom B\neq\varnothing$, and $A$ is of type (FPV). We show that when also $\dom A$ is convex, the sum operator: $A+B$ is also of type (FPV). Our result generalizes and unifies several recent sum theorems.
math.FA
math
Sum theorems for maximally monotone operators of type (FPV) 3 1 0 2 y a M 9 2 ] Jonathan M. Borwein∗ and Liangjin Yao† May 29, 2013 Abstract . A F h t a m [ 1 v 1 9 6 6 . 5 0 3 1 : v i X r a The most important open problem in Monotone Operator Theory concerns the maximal mono- tonicity of the sum of two maximally monotone operators provided that the classical Rockafel- lar's constraint qualification holds. In this paper, we establish the maximal monotonicity of A + B provided that A and B are maximally monotone operators such that star(dom A)∩ int dom B 6= ∅, and A is of type (FPV). We show that when also dom A is convex, the sum operator: A + B is also of type (FPV). Our result generalizes and unifies several recent sum theorems. 2010 Mathematics Subject Classification: Primary 47H05; Secondary 49N15, 52A41, 90C25 Keywords: Constraint qualification, convex function, convex set, Fitzpatrick function, linear relation, maximally monotone operator, monotone operator, operator of type (FPV), sum problem. 1 Introduction Throughout this paper, we assume that X is a real Banach space with norm k · k, that X ∗ is the continuous dual of X, and that X and X ∗ are paired by h·,·i. Let A : X ⇒ X ∗ be a set-valued operator (also known as a relation, point-to-set mapping or multifunction) from X to X ∗, i.e., for every x ∈ X, Ax ⊆ X ∗, and let gra A :=(cid:8)(x, x∗) ∈ X × X ∗ x∗ ∈ Ax(cid:9) be the graph of A. Recall that A is monotone if hx − y, x∗ − y∗i ≥ 0, ∀(x, x∗) ∈ gra A ∀(y, y∗) ∈ gra A, ∗CARMA, University [email protected]. Professor at King Abdul-Aziz University, Jeddah. of Newcastle, Newcastle, New South Wales E-mail: Laureate Professor at the University of Newcastle and Distinguished 2308, Australia. †CARMA, University of Newcastle, Newcastle, New South Wales 2308, Australia. E-mail: [email protected]. 1 and maximally monotone if A is monotone and A has no proper monotone extension (in the sense of graph inclusion). Let A : X ⇒ X ∗ be monotone and (x, x∗) ∈ X × X ∗. We say (x, x∗) is monotonically related to gra A if hx − y, x∗ − y∗i ≥ 0, ∀(y, y∗) ∈ gra A. Let A : X ⇒ X ∗ be maximally monotone. We say A is of type (FPV) if for every open convex set U ⊆ X such that U ∩ dom A 6= ∅, the implication x ∈ U and (x, x∗) is monotonically related to gra A ∩ (U × X ∗) ⇒ (x, x∗) ∈ gra A holds. We emphasize that it remains possible that all maximally monotone operators are of type (FPV). Also every (FPV) operator has the closure of its domain convex. See [28, 27, 8, 10] for this and more information on operators of type (FPV). We say A is a linear relation if gra A is a linear subspace. Monotone operators have proven important in modern Optimization and Analysis; see, e.g., the books [1, 8, 14, 15, 20, 27, 28, 25, 40, 41, 42] and the references therein. We adopt standard notation used in these books: thus, dom A :=(cid:8)x ∈ X Ax 6= ∅(cid:9) is the domain of A. Given a subset C of X, int C is the interior of C, bdry C is the boundary of C, aff C is the affine hull of C, C is the norm closure of C, and span C is the span (the set of all finite linear combinations) of C. The intrinsic core or relative algebraic interior of C, iC [40], is defined by iC := {a ∈ C ∀x ∈ aff(C − C),∃δ > 0,∀λ ∈ [0, δ] : a + λx ∈ C}. We then define icC by icC :=(iC, ∅, if aff C is closed; otherwise, The indicator function of C, written as ιC , is defined at x ∈ X by ιC (x) :=(0, ∞, if x ∈ C; otherwise. of C at x is defined by NC(x) :=(cid:8)x∗ ∈ X ∗ supc∈Chc − x, x∗i ≤ 0(cid:9), if x ∈ C; and NC(x) = ∅, if If C, D ⊆ X, we set C − D = {x − y x ∈ C, y ∈ D}. For every x ∈ X, the normal cone operator x /∈ C. We define the support points of C, written as supp C, by supp C := {c ∈ C NC(c) 6= {0}}. For x, y ∈ X, we set [x, y] := {tx + (1 − t)y 0 ≤ t ≤ 1}. We define the centre or star of C by star C := {x ∈ C [x, c] ⊆ C, ∀c ∈ C} [7]. Then C is convex if and only if star C = C. Given f : X → ]−∞, +∞], we set dom f := f −1(R). We say f is proper if dom f 6= ∅. We also set PX : X × X ∗ → X : (x, x∗) 7→ x. Finally, the open unit ball in X is denoted by UX :=(cid:8)x ∈ X kxk < 1(cid:9), the closed unit ball in X is denoted by BX :=(cid:8)x ∈ X kxk ≤ 1(cid:9), and N := {1, 2, 3, . . .}. We denote by −→ and ⇁w* the norm convergence and weak∗ convergence of nets, respectively. A + B : X ⇒ X ∗ : x 7→ Ax + Bx := (cid:8)a∗ + b∗ a∗ ∈ Ax and b∗ ∈ Bx(cid:9) is monotone. Rockafellar Let A and B be maximally monotone operators from X to X ∗. Clearly, the sum operator established the following very important result in 1970. 2 Theorem 1.1 (Rockafellar's sum theorem) (See [24, Theorem 1] or [8].) Suppose that X is reflexive. Let A, B : X ⇒ X ∗ be maximally monotone. Assume that A and B satisfy the classical constraint qualification Then A + B is maximally monotone. dom A ∩ int dom B 6= ∅. Arguably, the most significant open problem in the theory concerns the maximal monotonicity of the sum of two maximally monotone operators in general Banach spaces; this is called the "sum problem". Some recent developments on the sum problem can be found in Simons' monograph [28] and [4, 5, 6, 8, 12, 11, 36, 19, 31, 37, 38, 39]. It is known, among other things, that the sum theorem holds under Rockafellar's constraint qualification when both operators are of dense type or when each operator has nonempty domain interior [8, Ch. 8] and [35]. Here we focus on the case when A is of type (FPV), and B is maximally monotone such that star(dom A) ∩ int dom B 6= ∅. (Implicitly this means that B is also of type (FPV).) In Theorem 3.3 we shall show that A + B is maximally monotone. As noted it seems possible that all maximally monotone operators are of type (FPV). The remainder of this paper is organized as follows. In Section 2, we collect auxiliary results for future reference and for the reader's convenience. In Section 3, our main result (Theorem 3.3) is presented. In Section 4, we then provide various corollaries and examples. We also pose several sig- nificant open questions on the sum problem. We leave the details of proof of Case 2 of Theorem 3.3 to Appendix 5. 2 Auxiliary Results We first introduce one of Rockafellar's results. Fact 2.1 (Rockafellar) (See [23, Theorem 1] or [28, Theorem 27.1 and Theorem 27.3].) Let A : X ⇒ X ∗ be maximally monotone with int dom A 6= ∅. Then int dom A = int dom A and int dom A and dom A are both convex. The Fitzpatrick function defined below has proven to be an important tool in Monotone Operator Theory. Fact 2.2 (Fitzpatrick) (See [17, Corollary 3.9].) Let A : X ⇒ X ∗ be monotone, and set (1) FA : X × X ∗ → ]−∞, +∞] : (x, x∗) 7→ sup (a,a∗)∈gra A(cid:0)hx, a∗i + ha, x∗i − ha, a∗i(cid:1), 3 the Fitzpatrick function associated with A. Suppose also A is maximally monotone. Then for every (x, x∗) ∈ X × X ∗, the inequality hx, x∗i ≤ FA(x, x∗) is true, and the equality holds if and only if (x, x∗) ∈ gra A. The next result is central to our arguments. Fact 2.3 (See [33, Theorem 3.4 and Corollary 5.6], or [28, Theorem 24.1(b)].) Let A, B : X ⇒ X ∗ be maximally monotone operators. Assume Sλ>0 λ [PX (dom FA) − PX (dom FB)] is a closed subspace. If FA+B ≥ h·, ·i on X × X ∗, then A + B is maximally monotone. We next cite several results regarding operators of type (FPV). Fact 2.4 (Simons) (See [28, Theorem 46.1].) Let A : X ⇒ X ∗ be a maximally monotone linear relation. Then A is of type (FPV). The following result presents a sufficient condition for a maximally monotone operator to be of type (FPV). Fact 2.5 (Simons and Verona-Verona) (See [28, Theorem 44.1], [29] or [5].) Let A : X ⇒ X ∗ be maximally monotone. Suppose that for every closed convex subset C of X with dom A∩int C 6= ∅, the operator A + NC is maximally monotone. Then A is of type (FPV). Fact 2.6 (See [2, Lemma 2.5].) Let C be a nonempty closed convex subset of X such that int C 6= ∅. Let c0 ∈ int C and suppose that z ∈ X r C. Then there exists λ ∈ ]0, 1[ such that λc0 + (1 − λ)z ∈ bdry C. Fact 2.7 (Boundedness below) (See [9, Fact 4.1].) Let A : X ⇒ X ∗ be monotone and x ∈ int dom A. Then there exist δ > 0 and M > 0 such that x + δBX ⊆ dom A and supa∈x+δBX kAak ≤ M . Assume that (z, z∗) is monotonically related to gra A. Then hz − x, z∗i ≥ δkz∗k − (kz − xk + δ)M. Fact 2.8 (Voisei and Zalinescu) (See [36, Corollary 4].) Let A, B : X ⇒ X ∗ be maximally monotone. Assume that ic(dom A) 6= ∅,ic (dom B) 6= ∅ and 0 ∈ic [dom A − dom B]. Then A + B is maximally monotone. The proof of the next Lemma 2.9 follows closely the lines of that of [12, Lemma 2.10]. It generalizes both [12, Lemma 2.10] and [3, Lemma 2.10]. Lemma 2.9 Let A : X ⇒ X ∗ be monotone, and let B : X ⇒ X ∗ be a maximally monotone operator. Suppose that star(dom A) ∩ int dom B 6= ∅. Suppose also that (z, z∗) ∈ X × X ∗ with z ∈ dom A is monotonically related to gra(A + B). Then z ∈ dom B. 4 Proof. We can and do suppose that (0, 0) ∈ gra A∩gra B and 0 ∈ star(dom A)∩int dom B. Suppose to the contrary that z /∈ dom B. Then we have z 6= 0. We claim that (2) N[0,z] + B is maximally monotone. Because z Clearly, ic(dom B) 6= 0, we have 1 2 z ∈ ic(dom N[0,z]). 6= ∅ and 0 ∈ ic(cid:2)dom N[0,z] − dom B(cid:3). By Fact 2.8, N[0,z] + B is maximally monotone and hence (2) holds. Since (z, z∗) /∈ gra(N[0,z] + B), there exist λ ∈ [0, 1] and x∗, y∗ ∈ X ∗ such that (λz, x∗) ∈ gra N[0,z], (λz, y∗) ∈ gra B and (3) hz − λz, z∗ − x∗ − y∗i < 0. Now λ < 1, since (λz, y∗) ∈ gra B and z /∈ dom B, by (3), (4) hz,−x∗i + hz, z∗ − y∗i = hz, z∗ − x∗ − y∗i < 0. Since (λz, x∗) ∈ gra N[0,z], we have hz − λz, x∗i ≤ 0. Then hz,−x∗i ≥ 0. Thus (4) implies that (5) hz, z∗ − y∗i < 0. Since 0 ∈ star(dom A) and z ∈ dom A, λz ∈ dom A. By the assumption on (z, z∗), we have hz − λz, z∗ − a∗ − y∗i ≥ 0, ∀a∗ ∈ A(λz). Thence, hz, z∗ − a∗ − y∗i ≥ 0 and hence (6) hz, z∗ − y∗i ≥ hz, a∗i, ∀a∗ ∈ A(λz). Next we show that (7) We consider two cases. hz, a∗i ≥ 0, ∃a∗ ∈ A(λz). Case 1 : λ = 0. Then take a∗ = 0 to see that (7) holds. Case 2 : λ 6= 0. Let a∗ ∈ A(λz). Since (λz, a∗) ∈ gra A, hλz, a∗i = hλz − 0, a∗ − 0i ≥ 0 and hence hz, a∗i ≥ 0. Hence (7) holds. Combining (6) and (7), Hence z ∈ dom B. hz, z∗ − y∗i ≥ 0, which contradicts (5). (cid:4) The proof of Lemma 2.10 is modelled on that of [37, Proposition 3.1]. It is the first in a sequence of lemmas we give that will allow us to apply Fact 2.3. 5 Lemma 2.10 Let A : X ⇒ X ∗ be monotone, and let B : X ⇒ X ∗ be maximally monotone. Let (z, z∗) ∈ X × X ∗. Suppose x0 ∈ dom A ∩ int dom B and that there exists a sequence (an, a∗ n)n∈N in gra A ∩(cid:16) dom B × X ∗(cid:17) such that (an)n∈N converges to a point in [x0, z[, while (8) Then FA+B(z, z∗) = +∞. ni −→ +∞. hz − an, a∗ Proof. Since an ∈ dom B for every n ∈ N, we may pick v∗ n ∈ B(an). We again consider two cases. Case 1 : (v∗ n)n∈N is bounded. Then we have {n∈N} FA+B(z, z∗) ≥ sup ≥ sup = +∞ (by (8) and the boundedness of (v∗ [han, z∗i + hz − an, a∗ [−kank · kz∗k + hz − an, a∗ ni + hz − an, v∗ ni] ni − kz − ank · kv∗ nk] n)n∈N). {n∈N} Hence FA+B(z, z∗) = +∞. Case 2 : (v∗ n)n∈N is unbounded. By assumption, there exists 0 ≤ λ < 1 such that (9) We first show that (10) an −→ x0 + λ(z − x0). n→∞ hz − an, v∗ lim sup ni = +∞. n)n∈N is unbounded and, after passing to a subsequence if necessary, we may assume that nk → +∞. By x0 ∈ int dom B and Fact 2.7, there exist δ0 > 0 and nk 6= 0,∀n ∈ N and that kv∗ Since (v∗ kv∗ K0 > 0 such that (11) Then we have (12) han − x0, v∗ ni ≥ δ0kv∗ nk − (kan − x0k + δ0)K0. han − x0, v∗ n nki ≥ δ0 − kv∗ (kan − x0k + δ0)K0 kv∗ nk , ∀n ∈ N. By the Banach-Alaoglu Theorem (see [26, Theorem 3.15]), there exist a weak* convergent subnet ( nk )n∈N such that v∗ γ kv∗ γ k )γ∈Γ of ( v∗ n kv∗ (13) v∗ γ kv∗ γk ⇁w* v∗ ∞ ∈ X ∗. 6 Using (9) and taking the limit in (12) along the subnet, we obtain (14) Hence λ is strictly positive and hλ(z − x0), v∗ ∞i ≥ δ0. (15) hz − x0, v∗ ∞i ≥ δ0 λ > 0. Now assume contrary to (10) that there exists M > 0 such that Then, for all n sufficiently large, and so (16) n→∞ hz − an, v∗ lim sup ni < M. hz − an, v∗ ni < M + 1, hz − an, v∗ n nki < kv∗ M + 1 kv∗ nk . Then by (9) and (13), taking the limit in (16) along the subnet again, we see that (1 − λ)hz − x0, v∗ ∞i ≤ 0. Since λ < 1, we see hz − x0, v∗ ∞i ≤ 0 contradicting (15), and(10) holds. By (8) and (10), FA+B(z, z∗) ≥ sup n∈N [han, z∗i + hz − an, ani + hz − an, v∗ ni] = +∞. Hence as asserted. FA+B(z, z∗) = +∞, (cid:4) We also need the following two lemmas. Lemma 2.11 Let A : X ⇒ X ∗ be monotone, and let B : X ⇒ X ∗ be maximally monotone. Let (z, z∗) ∈ X × X ∗. Suppose that x0 ∈ dom A ∩ int dom B and that there exists a sequence (an)n∈N in dom A ∩ dom B such that (an)n∈N converges to a point in [x0, z[, and that (17) an ∈ bdry dom B, ∀n ∈ N. Then FA+B(z, z∗) = +∞. 7 (21) Then (20) and (19) imply that Thus, as before, λ > 0 and (22) (cid:10)y∗ n, an − x0(cid:11) ≥ δ. y∗ i ⇁w* y∗ ∞ ∈ X ∗. (cid:10)y∗ ∞, λ(z − x0)(cid:11) ≥ δ. ∞, z − x0(cid:11) ≥ (cid:10)y∗ δ λ > 0. Proof. Suppose to the contrary that (18) (z, z∗) ∈ dom FA+B. By the assumption, there exists 0 ≤ λ < 1 such that (19) an −→ x0 + λ(z − x0). By the Separation Theorem and Fact 2.1, there exists (y∗ n ∈ Ndom B(an). Thus ky∗ y∗ that x0 + δBX ⊆ dom B. Thus (cid:10)y∗ n, an(cid:11) ≥ sup(cid:10)y∗ =(cid:10)y∗ n, x0 + δBX(cid:11) ≥(cid:10)y∗ n, x0(cid:11) + δ. n, x0(cid:11) + sup(cid:10)y∗ Hence (20) nk = 1 and n ∈ Ndom B(an),∀k > 0. Since x0 ∈ int dom B, there exists δ > 0 such n)n∈N in X ∗ such that ky∗ n, δBX(cid:11) =(cid:10)y∗ n, x0(cid:11) + δky∗ nk By the Banach-Alaoglu Theorem (see [26, Theorem 3.15]), there exists a weak∗ convergent and bounded subnet (y∗ i )i∈O such that Since B is maximally monotone, B = B + Ndom B. As an ∈ dom A ∩ dom B, we have FA+B(z, z∗) ≥ sup(cid:2)(cid:10)z − an, A(an)(cid:11) +(cid:10)z − an, B(an) + ky∗ Thus n(cid:11) + hz∗, an(cid:11)(cid:3) , # , nE + hz∗, an(cid:11)k ∀n ∈ N,∀k > 0. ∀n ∈ N,∀k > 0. FA+B(z, z∗) k ≥ sup"Dz − an, A(an) Since (z, z∗) ∈ dom FA+B by (18), on letting k −→ +∞ we obtain k + y∗ B(an) k E +Dz − an, 0 ≥(cid:10)z − an, y∗ n(cid:11), 0 ≥(cid:10)(1 − λ)(z − x0), y∗ ∞(cid:11). ∀n ∈ N. 8 Combining with (21), (19) and taking the limit along the bounded subnet in the above inequality, we have Since λ < 1, which contradicts (22). Hence FA+B(z, z∗) = +∞. (cid:10)z − x0, y∗ ∞(cid:11) ≤ 0, (cid:4) Lemma 2.12 Let A : X ⇒ X ∗ be of type (FPV). Suppose x0 ∈ dom A but that z /∈ dom A. Then there is a sequence (an, a∗ n)n∈N in gra A so that (an)n∈N converges to a point in [x0, z[ and hz − an, a∗ ni −→ +∞. Proof. Since z /∈ dom A, z 6= x0. Thence there exist α > 0 and y∗ Set 0 ∈ X ∗ such that hy∗ 0, z − x0i ≥ α. Un := [x0, z] + 1 n UX , ∀n ∈ N. Since x0 ∈ dom A, Un ∩ dom A 6= ∅. Now (z, ny∗ there exist (an, a∗ n)n∈N in gra A with an ∈ Un such that 0) /∈ gra A and z ∈ Un. As A is of type (FPV), (23) hz − an, a∗ ni > hz − an, ny∗ 0i. As an ∈ Un, (an)n∈N has a subsequence convergent to an element in [x0, z]. We can assume that (24) an −→ x0 + λ(z − x0), where 0 ≤ λ ≤ 1, and since z /∈ dom A, we have λ < 1. Thus, x0 + λ(z − x0) ∈ [x0, z[. Thus by (24) and hz − x0, y∗ (cid:10)z − an, y∗ 0i ≥ α > 0, 0(cid:11) −→ (1 − λ)hz − x0, y∗ 0i ≥ (1 − λ)α > 0. Hence there exists N0 ∈ N such that for every n ≥ N0 (25) Appealing to (23), we have (1 − λ)α 2 > 0. (cid:10)z − an, y∗ 0(cid:11) ≥ (cid:10)z − an, a∗ n(cid:11) > 2 (1 − λ)α and so(cid:10)z − an, a∗ n(cid:11) −→ +∞. This completes the proof. n > 0, ∀n ≥ N0, (cid:4) 9 3 Our main result Before we come to our main result, we need the following two technical results which let us place points in the closures of the domains of A and B. The proof of Proposition 3.1 follows in part that of [37, Theorem 3.4]. Proposition 3.1 Let A : X ⇒ X ∗ be of type (FPV), and let B : X ⇒ X ∗ be maximally monotone. Suppose that dom A ∩ int dom B 6= ∅. Let (z, z∗) ∈ X × X ∗ with z ∈ dom B. Then FA+B(z, z∗) ≥ hz, z∗i. Proof. Clearly, FA+B(z, z∗) ≥ hz, z∗i if (z, z∗) /∈ dom FA+B. Now suppose that (z, z∗) ∈ dom FA+B. We can suppose that 0 ∈ dom A ∩ int dom B and (0, 0) ∈ gra A ∩ gra B. Next, we show that (26) FA+B(tz, tz∗) ≥ t2hz, z∗i and tz ∈ int dom B, ∀t ∈ ]0, 1[ . Fix t ∈ ]0, 1[. As 0 ∈ int dom B, z ∈ dom B, Fact 2.1 and [40, Theorem 1.1.2(ii)] imply (27) tz ∈ int dom B, and Fact 2.1 strengthens this to (28) We again consider two cases. tz ∈ int dom B. Case 1 : tz ∈ dom A. On selecting a∗ ∈ A(tz), b∗ ∈ B(tz), the definition of the Fitzpatrick function (1) shows FA+B(tz, tz∗) ≥ htz∗, tzi + htz, a∗ + b∗i − htz, a∗ + b∗i = htz, tz∗i. Hence (26) holds. Case 2 : tz /∈ dom A. If hz, z∗i ≤ 0, then FA+B(tz, tz∗) ≥ 0 ≥ htz, tz∗i because (0, 0) ∈ gra A ∩ gra B. So we assume that (29) We first show that (30) Set hz, z∗i > 0. tz ∈ dom A. Un := [0, tz] + 1 n UX, ∀n ∈ N. 10 Since 0 ∈ dom A, Un ∩ dom A 6= ∅. Since (tz, nz∗) /∈ gra A and tz ∈ Un, while A is of type (FPV), there is (an, a∗ n)n∈N in gra A with an ∈ Un such that (31) htz, a∗ ni > nhtz − an, z∗i + han, a∗ ni. As an ∈ Un, (an)n∈N has a subsequence convergent to an element in [0, tz]. We can assume that (32) an −→ λz, where 0 ≤ λ ≤ t. As tz ∈ int dom B also λz ∈ int dom B,and so appealing to Fact 2.7, there exist N ∈ N and K > 0 such that (33) We claim that (34) an ∈ int dom B and sup v∗∈B(an) kv∗k ≤ K, ∀n ≥ N. λ = t. Suppose to the contrary that 0 ≤ λ < t. As (an, a∗ n) ∈ gra A and (33) holds, for every n ≥ N sup sup {v∗∈B(an)} FA+B(z, z∗) ≥ ≥ ≥ han, z∗i + hz, a∗ > han, z∗i + 1 ≥ han, z∗i + 1 {v∗∈B(an)} [han, z∗i + hz, a∗ [han, z∗i + hz, a∗ ni − han, a∗ ni − han, a∗ ni − han, a∗ ni − Kkz − ank ni + hz − an, v∗i] ni − Kkz − ank] t nhtz − an, z∗i + 1 t nhtz − an, z∗i − Kkz − ank (since han, a∗ ni − han, a∗ t han, a∗ ni − Kkz − ank (by (31)) ni ≥ 0 by (0, 0) ∈ gra A and t ≤ 1). Divide by n on both sides of the above inequality and take the limit with respect to n. Since (32) and FA+B(z, z∗) < +∞, we obtain (1 − λ t )hz, z∗i = hz − λ t z, z∗i ≤ 0. Since 0 ≤ λ < t, we obtain hz, z∗i ≤ 0, which contradicts (29). Hence λ = t and by (32) tz ∈ dom A so that (30) holds. We next show that (35) Set FA+B(tz, tz∗) ≥ t2hz, z∗i. ∀n ∈ N. Note that Hn ∩ dom A 6= ∅, since tz ∈ dom A\dom A by (30). n UX , Hn := tz + 1 11 Because (tz, tz∗) /∈ gra A and tz ∈ Hn, and A is of type (FPV), there exists (bn, b∗ n)n∈N in gra A such that bn ∈ Hn and (36) htz, b∗ ni + hbn, tz∗i − hbn, b∗ ni > t2hz, z∗i, ∀n ∈ N. As tz ∈ int dom B and bn −→ tz, by Fact 2.7, there exist N1 ∈ N and M > 0 such that (37) sup v∗∈B(bn)kv∗k ≤ M, ∀n ≥ N1. bn ∈ int dom B and We now compute FA+B(tz, tz∗) ≥ ≥ sup sup {c∗∈B(bn)} [hbn, tz∗i + htz, b∗ ni − hbn, b∗ {c∗∈B(bn)}(cid:2)t2hz, z∗i + htz − bn, c∗i(cid:3) , ≥ sup(cid:2)t2hz, z∗i − Mktz − bnk(cid:3) , ni + htz − bn, c∗i] , ∀n ≥ N1 (by (36)) ∀n ≥ N1 ∀n ≥ N1 (by (37)). (38) Thus, because bn −→ tz. Hence FA+B(tz, tz∗) ≥ t2hz, z∗i. Thus (35) holds. FA+B(tz, tz∗) ≥ t2hz, z∗i Combining the above cases, we see that (26) holds. Since (0, 0) ∈ gra(A + B) and A + B is monotone, we have FA+B(0, 0) = h0, 0i = 0. Since FA+B is convex, (26) implies that tFA+B(z, z∗) = tFA+B(z, z∗) + (1 − t)FA+B(0, 0) ≥ FA+B(tz, tz∗) ≥ t2hz, z∗i, ∀t ∈ ]0, 1[ . Letting t −→ 1− in the above inequality, we obtain FA+B(z, z∗) ≥ hz, z∗i. (cid:4) We have one more block to put in place: Proposition 3.2 Let A : X ⇒ X ∗ be of type (FPV), and let B : X ⇒ X ∗ be maximally monotone. Suppose star(dom A) ∩ int dom B 6= ∅, and (z, z∗) ∈ dom FA+B. Then z ∈ dom A. Proof. We can and do suppose that 0 ∈ star(dom A) ∩ int dom B and (0, 0) ∈ gra A ∩ gra B. As before, we suppose to the contrary that (39) z /∈ dom A. Then z 6= 0. By the assumption that z /∈ dom A, Lemma 2.12 implies that there exist (an, a∗ in gra A and 0 ≤ λ < 1 such that (40) hz − an, a∗ ni −→ +∞ and an −→ λz. n)n∈N We yet again consider two cases. Case 1 : There exists a subsequence of (an)n∈N in dom B. 12 We can suppose that an ∈ dom B for every n ∈ N. Thus by (40) and Lemma 2.10, we have FA+B(z, z∗) = +∞, which contradicts our original assumption that (z, z∗) ∈ dom FA+B. Case 2 : There exists N1 ∈ N such that an 6∈ dom B for every n ≥ N1. Now we can suppose that an /∈ dom B for every n ∈ N. Since an /∈ dom B, Fact 2.1 and Fact 2.6 shows that there exists λn ∈ [0, 1] such that (41) λnan ∈ bdry dom B. By (40), we can suppose that (42) λnan −→ λ∞z. Since 0 ∈ star(dom A) and an ∈ dom A, λnan ∈ dom A. Then (40) implies that (43) λ∞ < 1. We further split Case 2 into two subcases. Subcase 2.1 : There exists a subsequence of (λnan)n∈N in dom B. We may again suppose λnan ∈ dom B for every n ∈ N. Since 0 ∈ star(dom A) and an ∈ dom A, λnan ∈ dom A. Then by (41) and (42), (43) and Lemma 2.11, FA+B(z, z∗) = +∞, which contradicts the hypothesis that (z, z∗) ∈ dom FA+B. Subcase 2.2 : There exists N2 ∈ N such that λnan 6∈ dom B for every n ≥ N2. We can now assume that λnan 6∈ dom B for every n ∈ N. Thus an 6= 0 for every n ∈ N. Since 0 ∈ int dom B, (41) and (42) imply that 0 < λ∞ and then by (43) (44) 0 < λ∞ < 1. Since 0 ∈ int dom B, (41) implies that λn > 0 for every n ∈ N. By (40), kan − zk 9 0. Then we can and do suppose that kan − zk 6= 0 for every n ∈ N. Fix n ∈ N. Since 0 ∈ int dom B, there exists 0 < ρ0 ≤ 1 such that ρ0BX ⊆ dom B. As 0 ∈ star(dom A) and an ∈ dom A, λnan ∈ dom A. Set (45) and take bn := λnan b∗ n ∈ A(λnan). τ0 := 1 Next we show that there exists εn ∈ (cid:3)0, 1 λnh2kzk + 2kank + 2 + (kank + 1) 2λnkz−ank Hn ⊆ dom B and inf(cid:13)(cid:13)B(cid:0)Hn(cid:1)(cid:13)(cid:13) ≥ n(1 + τ0kb∗ (46) i, we have n(cid:2) such that with Hn := (1 − εn)bn + εnρ0UX and ρ0 nk), while εn max{kank, 1} < 1 2kz − ankλn. For every s ∈ ]0, 1[, (41) and Fact 2.1 imply that (1− s)bn + sρ0BX ⊆ dom B. By Fact 2.1 again, (1 − s)bn + sρ0UX ⊆ int dom B = int dom B. 13 Now we show the second assertion of (46). Let k ∈ N and (sk)k∈N be a positive sequence such that sk −→ 0 when k −→ ∞. It suffices to show (47) lim k→∞ inf(cid:13)(cid:13)B(cid:0)(1 − sk)bn + skρ0UX(cid:1)(cid:13)(cid:13) = +∞. Suppose to the contrary there exist a sequence (ck, c∗ and L > 0 such that supk∈N kc∗ (again see [26, Theorem 3.15]), there exist a weak* convergent subnet, (c∗ that c∗ assumption that λnan /∈ dom B. Hence (47) holds and so does (46). kk ≤ L. Then ck −→ bn = λnan. By the Banach-Alaoglu Theorem k)k∈N such ∞) ∈ gra B, which contradicts our k)k∈N in gra B ∩(cid:2)(cid:0)(1 − sk)bn + skρ0UX(cid:1) × X ∗(cid:3) [9, Corollary 4.1] shows that (λnan, c∗ ∞ ∈ X ∗. β)β∈J of (c∗ β ⇁w* c∗ εnρ0 2λnkz−ank and thus 0 < tn < 1 4 . Thus Set tn := (48) tnλnz + (1 − tn)(1 − εn)bn ∈ Hn. ∗)n∈N in gra A ∩ (Hn × X ∗) such that Next we show there exists (fan,fan (49) nk. We consider two further subcases. (cid:10)z −fan,fan ∗(cid:11) ≥ −τ0kb∗ Subcase 2.2a: (cid:0)tnλnz + (1 − tn)(1 − εn)bn, (1 + tn)b∗ ∗) :=(cid:0)tnλnz + (1 − n(cid:1) ∈ gra A. Set (fan,fan n(cid:1). Since (0, 0) ∈ gra A, hbn, b∗ tn)(1 − εn)bn, (1 + tn)b∗ nE ∗E =Dtnλnz − tnλnz − (1 − tn)(1 − εn)bn, (1 + tn)b∗ Then we have ni ≥ 0. Dtnλnz −fan,fan =D − (1 − tn)(1 − εn)bn, (1 + tn)b∗ (50) On the other hand, (45) and the monotonicity of A imply that n)(1 − εn)bn, b∗ nE = −D(1 − t2 nE. nE ≥ −Dbn, b∗ nE =Dtnλnz + (1 − tn)(1 − εn)bn − bn, (1 + tn)b∗ Dtnλnz + (1 − tn)(1 − εn)bn − bn, tnb∗ n − b∗ nE ≥ 0. Thus (51) Dtnλnz − [1 − (1 − tn)(1 − εn)] bn, b∗ nE ≥ 0. Since 1 − (1 − tn)(1 − εn) > 0 and hbn, b∗ and thus ni = hbn − 0, b∗ n − 0i ≥ 0, (51) implies that htnλnz, b∗ ni ≥ 0 hz, b∗ ni ≥ 0. Then byfan ∗ = (1 + tn)b∗ Dz −fan,fan n and tnλn ≤ 1, (50) implies that ∗E ≥ −Dbn, b∗ nE ≥ −kbnk · kb∗ 14 nk ≥ −kank · kb∗ nk ≥ −τ0kb∗ nk. Hence (49) holds. ni ≥ 0) nE ∗ − b∗ nE (since hbn, b∗ we have (1 − εn)λnan ∈ dom A, hence dom A ∩ Hn 6= ∅. Since tnλnz + (1 − tn)(1 − εn)bn ∈ Hn n(cid:1) /∈ gra A. By 0 ∈ star(dom A) and an ∈ dom A, n(cid:1) /∈ gra A and A is of type (FPV), there exists nE > 0 Subcase 2.2b: (cid:0)tnλnz+(1−tn)(1−εn)bn, (1+tn)b∗ by (48), (cid:0)tnλnz + (1 − tn)(1 − εn)bn, (1 + tn)b∗ ∗) ∈ gra A such thatfan ∈ Hn and (fan,fan Dtnλnz + (1 − tn)(1 − εn)bn −fan,fan ∗ − (1 + tn)b∗ ⇒Dtnλnz − [1 − (1 − tn)(1 − εn)]fan + (1 − tn)(1 − εn)(bn −fan),fan nE ≥Dtnλnz −fan, tnb∗ >Dtnλnz + (1 − tn)(1 − εn)bn −fan, tnb∗ nE ⇒Dtnλnz − [1 − (1 − tn)(1 − εn)]fan,fan ∗ − b∗ nE ∗E +Dtnλnz −fan, tnb∗ >D(1 − tn)(1 − εn)(bn −fan), b∗ n −fan nE nE >Dtnλnz −fan, tnb∗ ⇒Dtnλnz − [1 − (1 − tn)(1 − εn)]fan,fan ∗ − b∗ nE. nE +Dtnλnz − [tn + εn − tnεn]fan, b∗ ∗E >Dtnλnz −fan, tnb∗ ⇒Dtnλnz − [tn + εn − tnεn]fan,fan ∗E ≥ ∗ − 0i ≥ 0 and tn + εn − tnεn ≥ tn ≥ tnλn,D [tn + εn − tnεn]fan,fan ∗i = hfan − 0,fan Since hfan,fan ∗E. Thus tnλnDfan,fan nE nE +Dtnλnz − [tn + εn − tnεn]fan, b∗ ∗E >Dtnλnz −fan, tnb∗ Dtnλnz − tnλnfan,fan ⇒D tnλnz − tnλnfan nE +D tnλnz − [tn + εn − tnεn]fan nE ∗E >Dtnλnz −fan, ,fan nE +Dz −(cid:20)1 + tn − εn(cid:21) 1 ∗E >Dtnλnz −fan, ⇒Dz −fan,fan nE λnfan, b∗ ∗E > − ⇒Dz −fan,fan nk(cid:16)kzk + (kank + 1)(cid:0)1 + nk(cid:0)kzk + kank + 1(cid:1) − kb∗ 1 λnkb∗ λn(cid:20)2kzk + 2kank + 2 + (kank + 1) ∗E > −kb∗ ⇒Dz −fan,fan 1 nk 1 λn 2λnkz − ank Finally, combining all the subcases, we deduce that (49) holds. (cid:21) = −τ0kb∗ nk. 1 λn b∗ 2λnkz − ank ρ0 (cid:17) 1 λn b∗ λntn , b∗ λntn εn ρ0 Since εn < 1 n andfan ∈ Hn, (42) shows that fan −→ λ∞z. n ∈ B(fan) by (46). Then by (46) again, nk ≥ n(1 + τ0kb∗ kw∗ nk), (52) (53) Take w∗ ∀n ∈ N. 15 Then by (49), we have −τ0kb∗ nk +Dz −fan, w∗ Thus (54) − τ0kb∗ nk kw∗ nk ∗E +Dz −fan, w∗ nE +Dz∗,fanE ≤Dz −fan,fan ≤ FA+B(z, z∗) nkE +D z∗ ,fanE ≤ +Dz −fan, w∗ n kw∗ kw∗ nk kw∗ nk FA+B(z, z∗) . nE +Dz∗,fanE By the Banach-Alaoglu Theorem (see [26, Theorem 3.15]), there exist a weak* convergent subnet, ( w∗ i kw∗ i k )i∈I of w∗ nk such that n kw∗ (55) w∗ i kw∗ i k ⇁w* w∗ ∞ ∈ X ∗. Combine (52), (53) and (55), by FA+B(z, z∗) < +∞, and take the limit along the subnet in (54) to obtain On the other hand, since 0 ∈ int dom B, Fact 2.7 implies that there exists ρ1 > 0 and M > 0 such that Use (52), (53) and (55), and take the limit along the subnet in the above inequality to obtain Then (44) shows that (56) Thus Hence which contradicts (56). ∞E ≤ 0. Dz − λ∞z, w∗ ∞E ≤ 0. Dz, w∗ Dfan, w∗ Dfan, w∗ n kw∗ nE ≥ ρ1kw∗ nkE ≥ ρ1 − Dλ∞z, w∗ ∞E ≥ Dz, w∗ nk − (kfank + ρ1)M. (kfank + ρ1)M ∞E ≥ ρ1. kw∗ nk . ρ1 λ∞ > 0, Combining all the above cases, we have arrived at z ∈ dom A. We are finally ready to prove our main result. The special case in which B is the normal cone (cid:4) operator of a nonempty closed convex set was first established by Voisei in [34]. 16 Theorem 3.3 ((FPV) Sum Theorem) Let A, B : X ⇒ X ∗ be maximally monotone with star(dom A)∩int dom B 6= ∅. Assume that A is of type (FPV). Then A+B is maximally monotone. Proof. After translating the graphs if necessary, we can and do assume that 0 ∈ star(dom A) ∩ int dom B and that (0, 0) ∈ gra A ∩ gra B. By Fact 2.2, dom A ⊆ PX (dom FA) and dom B ⊆ PX (dom FB). Hence, (57) [λ>0 λ(cid:0)PX (dom FA) − PX(dom FB)(cid:1) = X. Thus, by Fact 2.3, it suffices to show that (58) FA+B(z, z∗) ≥ hz, z∗i, ∀(z, z∗) ∈ X × X ∗. Take (z, z∗) ∈ X × X ∗. Then (59) FA+B(z, z∗) = sup {x,x∗,y∗} [hx, z∗i + hz − x, x∗i + hz − x, y∗i − ιgra A(x, x∗) − ιgra B(x, y∗)] . Suppose to the contrary that there exists η > 0 such that (60) so that (61) FA+B(z, z∗) + η < hz, z∗i, (z, z∗) is monotonically related to gra(A + B). Then by Proposition 3.1 and Proposition 3.2 , (62) Now by Lemma 2.9, (63) z ∈ dom A\dom B. z /∈ dom A. Indeed, if z ∈ dom A, Lemma 2.9 and (61) show that z ∈ dom B. Thus, z ∈ dom A ∩ dom B and hence FA+B(z, z∗) ≥ hz, z∗i which contradicts (60). Thence we have established (63). Thus (62) implies that there exists (an, a∗ n)n∈N in gra A such that (64) an −→ z. By (62), an /∈ dom B for all but finitely many terms an. We can suppose that an /∈ dom B for all n ∈ N. Fact 2.1 and Fact 2.6 show that there exists λn ∈ ]0, 1[ such that (65) λnan ∈ bdry dom B. 17 By (64), we can assume that (66) λn −→ λ∞ ∈ [0, 1] and thus λnan −→ λ∞z. Then by (65) and (62) (67) We consider two cases. λ∞ < 1. Case 1 : There exists a subsequence of (λnan)n∈N in dom B. We can suppose that λnan ∈ dom B for every n ∈ N. Since 0 ∈ star(dom A) and an ∈ dom A, λnan ∈ dom A. Then by (65),(66), (67) and Lemma 2.11, FA+B(z, z∗) = +∞, which contradicts (60) that (z, z∗) ∈ dom FA+B. Case 2 : There exists N ∈ N such that λnan 6∈ dom B for every n ≥ N . We can suppose that λnan 6∈ dom B for every n ∈ N. Thus an 6= 0 for every n ∈ N. Following the pattern of Subcase 2.2 in the proof of Proposition 3.2 1, we obtain a contradiction. Combing all the above cases, we have FA+B(z, z∗) ≥ hz, z∗i for all (z, z∗) ∈ X × X ∗. Hence A + B is maximally monotone. Remark 3.4 In Case 2 in the proof of Theorem 3.3 (see Appendix 5 below), we use Lemma 2.9 to deduce that kan − zk 6= 0. Without the help of Lemma 2.9, we may still can obtain (77) as follows. For the case of an = z, consider whether(cid:0)(1 − εn)bn, 0(cid:1) =(cid:0)(1 − εn)λnz, 0(cid:1) ∈ Hn × X ∗ is in gra A or not. We can deduce that there exists (fan,fan (cid:10)z −fan,fan ∗(cid:11) ≥ 0. Hence (77) holds, and the proof of Theorem 3.3 can be achieved without Lemma 2.9. ∗)n∈N in gra A ∩ (Hn × X ∗) such that (cid:4) ♦ 4 Examples and Consequences We start by illustrating that the starshaped hypothesis catches operators whose domain may be non-convex and have no algebraic interior. Example 4.1 (Operators with starshaped domains) We illustrate that there are many choices of maximally monotone operator A of type (FPV) with non-convex domain such that ic dom A = int dom A = ∅ and star(dom A) 6= ∅. Let f : R2 → ]−∞, +∞] be defined by (x, y) 7→(max{1 − √x, y} +∞, if x ≥ 0; otherwise. 1We banish the details to Appendix 5 to spare the readers. 18 Consider an infinite dimensional Banach space X containing a nonempty closed and convex set C such that icC = ∅. It is not known whether all spaces have this property but all separable or reflexive spaces certainly do [8]. Define A : (R2 × X) ⇒ (R2 × X ∗) by (v, w) ⇒(cid:16)∂f (v), ∂ιC (w)(cid:17) = ∂F (v, w), where F := f ⊕ ιC . Define k · k on R2 × X by k(v, w)k := kvk + kwk. Then f is proper convex and lowers semicontinuous and so, therefore, is F . Indeed, [22, Example before Theorem 23.5, page 218] shows that dom ∂f is not convex and consequently dom A is not convex. (Many other candidates for f are given in [8, Chapter 7].) Clearly, A = ∂F is maximally monotone. Let w0 ∈ C and v0 = (2, 0). Consider (v0, w0) ∈ R2 × X. Since v0 = (2, 0) ∈ int dom ∂f , v0 ∈ star(dom ∂f ) since dom f is convex. Thus (v0, w0) ∈ star(dom A). Since icC = ∅ and so int C = ∅, it follows that ic dom A = int dom A = ∅. [28, Theorem 48.4(d)] shows that A = ∂F is of type (FPV). ♦ The next example gives all the details of how to associate the support points of a convex set to a subgradient. In [18], [13] and [8, Exercise 8.4.1, page 401] the construction is used to build empty subgradients in various Fr´echet spaces and incomplete normed spaces. Example 4.2 (Support points) Suppose that X is separable. We can always find a compact convex set C ⊆ X such that span C 6= X and span C = X [8]. Take x0 /∈ span C. Define f : X → ]−∞, +∞] by (68) f (x) := min{t ∈ R x + tx0 ∈ C}, ∀x ∈ X. By direct computation f is proper lower semicontinuous and convex, see [18]. By the definition of f , dom f = C + Rx0. Let t ∈ R and c ∈ C. We shall establish that (69) ∂f (tx0 + c) =(NC(c) ∩ {y∗ ∈ X ∗ hy∗, x0i = −1}, ∅, if c ∈ supp C; otherwise. . Thence, also dom ∂f = Rx0 + supp C. First we show that the implication (70) tx0 + c = sx0 + d, where t, s ∈ R, c, d ∈ C ⇒ t = s and c = d holds. Let t, s ∈ R and c, d ∈ C. We have (t − s)x0 = d − c ∈ span C − span C = span C. Since x0 /∈ span C, t = s and then c = d. Hence we obtain (70). By (70), we have (71) We next show that (69) holds. f (tx0 + c) = −t, ∀t ∈ R,∀c ∈ C. 19 Since dom f = C + Rx0, by (71), we have ∀s ∈ R,∀d ∈ C x∗ ∈ ∂f (tx0 + c) ⇔Dx∗, sx0 + d − (tx0 + c)E ≤ f (sx0 + d) − f (tx0 + c) = −s + t, ⇔Dx∗, (s − t)x0 + (d − c)E ≤ −s + t, ⇔Dx∗, sx0 + (d − c)E ≤ −s, ⇔Dx∗, sx0E ≤ −s and Dx∗, d − cE ≤ 0, ⇔Dx∗, sx0E ≤ −s and x∗ ∈ NC(c), x∗ 6= 0, ⇔Dx∗, x0E = −1, and c ∈ supp C. ∀s ∈ R,∀d ∈ C x∗ ∈ NC(c) ∀s ∈ R,∀d ∈ C ∀s ∈ R,∀d ∈ C ∀s ∈ R Hence (69) holds. As a concrete example of C consider, for 1 ≤ p < ∞, any order interval C := {x ∈ ℓp(N) : α ≤ x ≤ β} where α < β ∈ ℓp(N). The example extends to all weakly compactly generated (WCG) spaces [8] with a weakly compact convex set in the role of C. ♦ We gave the last example in part as it allows one to better understand what the domain of a maximally monotone operator with empty interior can look like. While the star may be empty, it has been recently proven [32], see also [16], that for a closed convex function f the domain of ∂f is always pathwise and locally pathwise connected. An immediate corollary of Theorem 3.3 is the following which generalizes [37, Corollary 3.9]. Corollary 4.3 (Convex domain) Let A, B : X ⇒ X ∗ be maximally monotone with dom A ∩ int dom B 6= ∅. Assume that A is of type (FPV) with convex domain. Then A + B is maximally monotone. An only slightly less immediate corollary is given next. Corollary 4.4 (Nonempty interior) (See [5, Theorem 9(i)] or Fact 2.8.) Let A, B : X ⇒ X ∗ be maximally monotone with int dom A ∩ int dom B 6= ∅. Then A + B is maximally monotone. Proof. By the assumption, there exists x0 ∈ int dom A ∩ int dom B. We first show that A is of type (FPV). Let C be a nonempty closed convex subset of X, and suppose that dom A∩ int C 6= ∅. Let x1 ∈ dom A∩int C. Fact 2.1 and [40, Theorem 1.1.2(ii)] imply that [x0, x1[ ⊆ int dom A = int dom A. Since x1 ∈ int C, there exists 0 < δ < 1 such that x1 + δ(x0 − x1) ∈ int dom A∩ int C. Then NC + A is maximally monotone by Corollary 4.3 and [28, Theorem 48.4(d)]. Hence by Fact 2.5, A is of type (FPV), see also [5]. 20 Since x0 ∈ int dom A, Fact 2.1 and [40, Theorem 1.1.2(ii)] imply that x0 ∈ star(dom A) and hence we have x0 ∈ star(dom A) ∩ int dom B. Then by Theorem 3.3, we deduce that A + B is maximally monotone. (cid:4) Corollary 4.5 (Linear relation) (See [12, Theorem 3.1].) Let A : X ⇒ X ∗ be a maximally monotone linear relation, and let B : X ⇒ X ∗ be maximally monotone. Suppose that dom A ∩ int dom B 6= ∅. Then A + B is maximally monotone. Proof. Apply Fact 2.4 and Corollary 4.3 directly. (cid:4) The proof of our final Corollary 4.6 is adapted from that of [37, Corollary 2.10] and [12, Corol- lary 3.3]. Moreover, it generalizes both [37, Corollary 2.10] and [12, Corollary 3.3]. Corollary 4.6 (FPV property of the sum) Let A, B : X ⇒ X ∗ be maximally monotone with dom A ∩ int dom B 6= ∅. Assume that A is of type (FPV) with convex domain. Then A + B is of type (F P V ). Proof. By Corollary 4.3, A + B is maximally monotone. Let C be a nonempty closed convex subset of X, and suppose that dom(A + B) ∩ int C 6= ∅. Let x1 ∈ dom A ∩ int dom B and x2 ∈ dom(A + B) ∩ int C. Then x1, x2 ∈ dom A, x1 ∈ int dom B and x2 ∈ dom B ∩ int C. Hence λx1 + (1− λ)x2 ∈ int dom B for every λ ∈ ]0, 1] by Fact 2.1 and [40, Theorem 1.1.2(ii)] and so there exists δ ∈ ]0, 1] such that λx1 + (1 − λ)x2 ∈ int C for every λ ∈ [0, δ]. Thus, δx1 + (1 − δ)x2 ∈ dom A ∩ int dom B ∩ int C. By Corollary 4.4, B + NC is maximally monotone. Then, by Corollary 4.3 (applied A and B +NC to A and B), A+B +NC = A+(B +NC ) is maximally monotone. By Fact 2.5, A + B is of type (F P V ). (cid:4) We have been unable to relax the convexity hypothesis in Corollary 4.6. We finish by listing some related interesting, at least to the current authors, questions regarding the sum problem. Open Problem 4.7 Let A : X ⇒ X ∗ be maximally monotone with convex domain. Is A neces- sarily of type (FPV)? Let us recall a problem posed by S. Simons in [27, Problem 41.2] Open Problem 4.8 Let A : X ⇒ X ∗ be of type (FPV), let C be a nonempty closed convex subset of X, and suppose that dom A ∩ int C 6= ∅. Is A + NC necessarily maximally monotone? More generally, can we relax or indeed entirely drop the starshaped hypothesis on dom A in Theorem 3.3? Open Problem 4.9 Let A, B : X ⇒ X ∗ be maximally monotone with dom A ∩ int dom B 6= ∅. Assume that A is of type (FPV). Is A + B necessarily maximally monotone? 21 If all maximally monotone operators are type (FPV) this is no easier than the full sum problem. Can the results of [32] help here? Acknowledgment Jonathan Borwein and Liangjin Yao were partially supported by various Australian Research Coun- cil grants. References [1] H.H. Bauschke and P.L. Combettes, Convex Analysis and Monotone Operator Theory in Hilbert Spaces, Springer-Verlag, 2011. [2] H.H. Bauschke, X. Wang, and L. Yao, "An answer to S. Simons' question on the maximal monotonicity of the sum of a maximal monotone linear operator and a normal cone operator", Set-Valued and Variational Analysis, vol. 17, pp. 195 -- 201, 2009. [3] H.H. Bauschke, X. Wang, and L. Yao, "On the maximal monotonicity of the sum of a maximal monotone linear relation and the subdifferential operator of a sublinear function", Proceedings of the Haifa Workshop on Optimization Theory and Related Topics. Contemp. Math., Amer. Math. Soc., Providence, RI, vol. 568, pp. 19 -- 26, 2012. [4] J.M. Borwein, "Maximal monotonicity via convex analysis", Journal of Convex Analysis, vol. 13, pp. 561 -- 586, 2006. [5] J.M. Borwein, "Maximality of sums of two maximal monotone operators in general Banach space", Proceedings of the American Mathematical Society, vol. 135, pp. 3917 -- 3924, 2007. [6] J.M. Borwein, "Fifty years of maximal monotonicity", Optimization Letters, vol. 4, pp. 473 -- 490, 2010. [7] J.M. Borwein and A.S. Lewis, Convex Analyis andd Nonsmooth Optimization, Second ex- panded edition, Springer, 2005. [8] J.M. Borwein and J. Vanderwerff, Convex Functions: Constructions, Characterizations and Counterexamples, Encyclopedia of Mathematics and its Applications, 109, Cambridge Uni- versity Press, 2010. [9] J.M. Borwein and L. Yao,"Structure theory for maximally monotone operators with points of continuity", Journal of Optimization Theory and Applications, vol 156, pp. 1 -- 24, 2013 (Invited paper). [10] J.M. Borwein and L. Yao,"Some results on the convexity of the closure of the domain of a maximally monotone operator", Optimization Letters, in press; http://dx.doi.org/10.1007/s11590-012-0564-7. 22 [11] J.M. Borwein and L. Yao, "Recent progress on Monotone Operator Theory", Infinite Products of Operators and Their Applications, Contemporary Mathematics, in press; http://arxiv.org/abs/1210.3401v2. [12] J.M. Borwein and L. Yao, "Maximality of the sum of a maximally monotone linear relation and a maximally monotone operator", Set-Valued and Variational Analysis, in press; http://arxiv.org/abs/1212.4266v1. [13] A. Brøndsted and R.T. Rockafellar, "On the subdifferentiability of convex functions", Pro- ceedings of the American Mathematical Society , vol. 16, pp. 605 -- 611, 1995. [14] R.S. Burachik and A.N. Iusem, Set-Valued Mappings and Enlargements of Monotone Opera- tors, Springer-Verlag, 2008. [15] D. Butnariu and A.N. Iusem, Totally Convex Functions for Fixed Points Computation and Infinite Dimensional Optimization, Kluwer Academic Publishers, 2000. [16] C. De Bernardi and L. Vesel´y, "On support points and support functionals of convex sets," Israel J. Math. vol. 171, pp. 15-27, 2009. [17] S. Fitzpatrick, "Representing monotone operators by convex functions", in Work- shop/Miniconference on Functional Analysis and Optimization (Canberra 1988), Proceedings of the Centre for Mathematical Analysis, Australian National University, vol. 20, Canberra, Australia, pp. 59 -- 65, 1988. [18] R. B. Holmes, Geometric Functional Analysis and its Applications, Springer-Verlag, New York, 1975. [19] M. Marques Alves and B.F. Svaiter, "A new qualification condition for the maximality of the sum of maximal monotone operators in general Banach spaces", Journal of Convex Analysis, vol. 19, pp. 575 -- 589, 2012. [20] R.R. Phelps, Convex Functions, Monotone Operators and Differentiability, 2nd Edition, Springer-Verlag, 1993. [21] R.R. Phelps and S. Simons, "Unbounded linear monotone operators on nonreflexive Banach spaces", Journal of Nonlinear and Convex Analysis, vol. 5, pp. 303 -- 328, 1998. [22] R.T. Rockafellar, Convex Analysis, Princeton Univ. Press, Princeton, 1970. [23] R.T. Rockafellar, "Local boundedness of nonlinear, monotone operators", Michigan Mathe- matical Journal, vol. 16, pp. 397 -- 407, 1969. [24] R.T. Rockafellar, "On the maximality of sums of nonlinear monotone operators", Transactions of the American Mathematical Society, vol. 149, pp. 75 -- 88, 1970. [25] R.T. Rockafellar and R.J-B Wets, Variational Analysis, 3nd Printing, Springer-Verlag, 2009. [26] R. Rudin, Functional Analysis, Second Edition, McGraw-Hill, 1991. 23 [27] S. Simons, Minimax and Monotonicity, Springer-Verlag, 1998. [28] S. Simons, From Hahn-Banach to Monotonicity, Springer-Verlag, 2008. [29] A. Verona and M.E. Verona, "Regular maximal monotone operators", Set-Valued Analysis, vol. 6, pp. 303 -- 312, 1998. [30] A. Verona and M.E. Verona, "Regular maximal monotone operators and the sum theorem", Journal of Convex Analysis, vol. 7, pp. 115 -- 128, 2000. [31] A. Verona and M.E. Verona, "On the regularity of maximal monotone operators and related results"; http://arxiv.org/abs/1212.1968v3, December 2012. [32] L. Vesel´y "A parametric smooth variational principle and support properties of convex sets and functions", Journal of Mathematical Analysis and Applications, vol. 350, pp. 550-561, 2009. [33] M.D. Voisei, "The sum and chain rules for maximal monotone operators", Set-Valued and Variational Analysis, vol. 16, pp. 461 -- 476, 2008. [34] M.D. Voisei, "A Sum Theorem for (FPV) operators and normal cones ", Journal of Mathe- matical Analysis and Applications, vol. 371, pp. 661 -- 664, 2010. [35] M.D. Voisei and C. Zalinescu, "Strongly-representable monotone operators", Journal of Con- vex Analysis , vol. 16, pp. 1011 -- 1033, 2009. [36] M.D. Voisei and C. Zalinescu, "Maximal monotonicity criteria for the composition and the sum under weak interiority conditions", Mathematical Programming (Series B), vol. 123, pp. 265 -- 283, 2010. [37] L. Yao, "The sum of a maximal monotone operator of type (FPV) and a maximal monotone operator with full domain is maximally monotone", Nonlinear Anal., vol. 74, pp. 6144 -- 6152, 2011. [38] L. Yao, "The sum of a maximally monotone linear relation and the subdifferential of a proper lower semicontinuous convex function is maximally monotone", Set-Valued and Variational Analysis, vol. 20, pp. 155 -- 167, 2012. [39] L. Yao, On Monotone Linear Relations and the Sum Problem in Banach Spaces, Ph.D. thesis, University of British Columbia (Okanagan), 2011; http://hdl.handle.net/2429/39970. [40] C. Zalinescu, Convex Analysis in General Vector Spaces, World Scientific Publishing, 2002. [41] E. Zeidler, Nonlinear Functional Analysis and its Applications II/A: Linear Monotone Oper- ators, Springer-Verlag, 1990. [42] E. Zeidler, Nonlinear Functional Analysis and its Applications II/B: Nonlinear Monotone Operators, Springer-Verlag, 1990. 24 5 Appendix Proof of Case 2 in the proof of Theorem 3.3. Proof. Case 2 : There exists N ∈ N such that λnan 6∈ dom B for every n ≥ N . We can and do suppose that λnan 6∈ dom B for every n ∈ N. Thus an 6= 0 for every n ∈ N. Since 0 ∈ int dom B, (66) and (65) imply that 0 < λ∞ and hence by (67) (72) 0 < λ∞ < 1. By (63), kan − zk 6= 0 for every n ∈ N. Fix n ∈ N. Since 0 ∈ int dom B, there exists 0 < ρ0 ≤ 1 such that ρ0BX ⊆ dom B. Since 0 ∈ star(dom A) and an ∈ dim A, λnan ∈ dom A. Set (73) and take bn := λnan Next we show that there exists εn ∈(cid:3)0, 1 (74) Hn ⊆ dom B and n(cid:2) such that inf(cid:13)(cid:13)B(cid:0)Hn(cid:1)(cid:13)(cid:13) ≥ n(1 + τ0kb∗ where Hn := (1 − εn)bn + εnρ0UX and τ0 := 1 b∗ n ∈ A(λnan). nk), εn max{kank, 1} < 1 2kz − ankλn. λnh2kzk + 2kank + 2 + (kank + 1) 2λnkz−ank ρ0 i. For every ε ∈ ]0, 1[, by (65) and Fact 2.1, (1 − ε)bn + ερ0BX ⊆ dom B. By Fact 2.1 again, (1 − ε)bn + ερ0UX ⊆ int dom B = int dom B. Now we show the second part of (74). Let k ∈ N and (sk)k∈N be a positive sequence such that sk −→ 0 when k −→ ∞. It suffices to show (75) lim k→∞ inf(cid:13)(cid:13)B(cid:0)(1 − sk)bn + skρ0UX(cid:1)(cid:13)(cid:13) = +∞. Suppose to the contrary there exist a sequence (ck, c∗ and L > 0 such that supk∈N kc∗ orem (see [26, Theorem 3.15]), there exist a weak* convergent subnet, (c∗ that c∗ assumption that λnan /∈ dom B. kk ≤ L. Then ck −→ bn = λnan. By the Banach-Alaoglu The- k)k∈N such ∞) ∈ gra B, which contradicts our k)k∈N in gra B ∩(cid:2)(cid:0)(1 − sk)bn + skρ0UX(cid:1) × X ∗(cid:3) [9, Corollary 4.1] shows that (λnan, c∗ ∞ ∈ X ∗. β)β∈J of (c∗ β ⇁w* c∗ Hence (75) holds and so does (74). εnρ0 2λnkz−ank and thus 0 < tn < 1 4 . Thus Set tn := (76) Next we show there exists (fan,fan (77) tnλnz + (1 − tn)(1 − εn)bn ∈ Hn. ∗)n∈N in gra A ∩ (Hn × X ∗) such that (cid:10)z −fan,fan ∗(cid:11) ≥ −τ0kb∗ nk. 25 We consider two subcases. Then we have Subcase 2.1 : (cid:0)tnλnz + (1 − tn)(1 − εn)bn, (1 + tn)b∗ Then set (fan,fan n(cid:1) ∈ gra A. ∗) :=(cid:0)tnλnz + (1 − tn)(1 − εn)bn, (1 + tn)b∗ nE = −D(1 − t2 n(cid:1). Since (0, 0) ∈ gra A, hbn, b∗ nE nE =Dtnλnz − tnλnz − (1 − tn)(1 − εn)bn, (1 + tn)b∗ nE. nE ≥ −Dbn, b∗ n)(1 − εn)bn, b∗ nE =Dtnλnz + (1 − tn)(1 − εn)bn − bn, (1 + tn)b∗ Dtnλnz −fan,fa∗ =D − (1 − tn)(1 − εn)bn, (1 + tn)b∗ Dtnλnz + (1 − tn)(1 − εn)bn − bn, tnb∗ On the other hand, (73) and the monotonicity of A imply that n − b∗ (78) nE ≥ 0 ni ≥ 0. Thus (79) Dtnλnz − [1 − (1 − tn)(1 − εn)] bn, b∗ nE ≥ 0. Since 1 − (1 − tn)(1 − εn) > 0 and hbn, b∗ and thus ni = hbn − 0, b∗ n − 0i ≥ 0, (79) implies that htnλnz, b∗ ni ≥ 0 Since 0 ∈ star(dom A) and an ∈ dom A, we have (1 − εn)λnan ∈ dom A, hence dom A ∩ Hn 6= ∅. hz, b∗ ni ≥ 0. Then byfan ∗ = (1 + tn)b∗ n and tnλn ≤ 1, (78) implies that Hence (77) holds. nk ≥ −τ0kb∗ nk. nk ≥ −kank · kb∗ n(cid:1) /∈ gra A. nE ≥ −Dbn, b∗ Dz −fan,fa∗ nE ≥ −kbnk · kb∗ Subcase 2.2 : (cid:0)tnλnz + (1 − tn)(1 − εn)bn, (1 + tn)b∗ Since tnλnz + (1 − tn)(1 − εn)bn ∈ Hn by (76),(cid:0)tnλnz + (1 − tn)(1 − εn)bn, (1 + tn)b∗ ∗) ∈ gra A such thatfan ∈ Hn and A is of type (FPV), there exists (fan,fan Dtnλnz + (1 − tn)(1 − εn)bn −fan,fan ⇒Dtnλnz − [1 − (1 − tn)(1 − εn)]fan + (1 − tn)(1 − εn)(bn −fan),fan nE ≥Dtnλnz −fan, tnb∗ >Dtnλnz + (1 − tn)(1 − εn)bn −fan, tnb∗ nE ⇒Dtnλnz − [1 − (1 − tn)(1 − εn)]fan,fan ∗ − b∗ nE ∗E +Dtnλnz −fan, tnb∗ >D(1 − tn)(1 − εn)(bn −fan), b∗ n −fan nE nE >Dtnλnz −fan, tnb∗ ⇒Dtnλnz − [1 − (1 − tn)(1 − εn)]fan,fan ∗ − b∗ nE ∗ − b∗ nE (since hbn, b∗ ∗ − (1 + tn)b∗ nE > 0 26 n(cid:1) /∈ gra A and ni ≥ 0) ∗E >Dtnλnz −fan, tnb∗ nE. nE +Dtnλnz − [tn + εn − tnεn]fan, b∗ ⇒Dtnλnz − [tn + εn − tnεn]fan,fan ∗E ≥ ∗ − 0i ≥ 0 and tn + εn − tnεn ≥ tn ≥ tnλn,D [tn + εn − tnεn]fan,fan ∗i = hfan − 0,fan Since hfan,fan ∗E. Thus tnλnDfan,fan nE nE +Dtnλnz − [tn + εn − tnεn]fan, b∗ ∗E >Dtnλnz −fan, tnb∗ Dtnλnz − tnλnfan,fan ⇒D tnλnz − tnλnfan nE +D tnλnz − [tn + εn − tnεn]fan nE ∗E >Dtnλnz −fan, ,fan nE +Dz −(cid:20)1 + tn − εn(cid:21) 1 ∗E >Dtnλnz −fan, ⇒Dz −fan,fan nE λnfan, b∗ nk(cid:16)kzk + ∗E > − ⇒Dz −fan,fan nk(cid:0)kzk + kank + 1(cid:1) − kb∗ (kank + 1)(cid:0)1 + 1 λnkb∗ λn(cid:20)2kzk + 2kank + 2 + (kank + 1) ∗E > −kb∗ ⇒Dz −fan,fan 1 nk 1 λn 2λnkz − ank Hence combining all the subcases, we have (77) holds. (cid:21) = −τ0kb∗ 2λnkz − ank 1 λn 1 λn nk. (cid:17) λntn λntn , b∗ b∗ b∗ εn ρ0 ρ0 Since εn < 1 (80) (81) Take w∗ n andfan ∈ Hn, (66) shows that fan −→ λ∞z. n ∈ B(fan) by (74). Then by (74) again, nk ≥ n(1 + τ0kb∗ kw∗ nk), Then by (77), we have ∀n ∈ N. −τ0kb∗ nk +Dz −fan, w∗ Thus (82) − τ0kb∗ nk kw∗ nk ∗E +Dz −fan, w∗ nE +Dz∗,fanE ≤Dz −fan,fan ≤ FA+B(z, z∗) nkE +D z∗ ,fanE ≤ +Dz −fan, w∗ n kw∗ kw∗ nk kw∗ nk FA+B(z, z∗) . nE +Dz∗,fanE By the Banach-Alaoglu Theorem (see [26, Theorem 3.15]), there exist a weak* convergent subnet, ( w∗ i kw∗ i k )i∈I of w∗ nk such that n kw∗ (83) w∗ i kw∗ i k ⇁w* w∗ ∞ ∈ X ∗. Combining (80), (81) and (83), by FA+B(z, z∗) < +∞, take the limit along the subnet in (82) to obtain Dz − λ∞z, w∗ ∞E ≤ 0. 27 Then (72) shows that (84) Dz, w∗ ∞E ≤ 0. On the other hand, since 0 ∈ int dom B, Fact 2.7 implies that there exists ρ1 > 0 and M > 0 such that Combining (80), (81) and (83), take the limit along the subnet in the above inequality to obtain Thus Hence which contradicts (84). nk − (kfank + ρ1)M. (kfank + ρ1)M ∞E ≥ ρ1. kw∗ nk . Dfan, w∗ Dfan, w∗ n kw∗ nE ≥ ρ1kw∗ nkE ≥ ρ1 − Dλ∞z, w∗ ∞E ≥ Dz, w∗ ρ1 λ∞ > 0, (cid:4) 28
1005.2936
6
1005
2013-08-21T02:26:56
Maximal and area integral characterizations of Bergman spaces in the unit ball of $\mathbb{C}^n$
[ "math.FA", "math.CV" ]
In this paper, we present maximal and area integral characterizations of Bergman spaces in the unit ball of $\mathbb{C}^n.$ The characterizations are in terms of maximal functions and area integral functions on Bergman balls involving the radial derivative, the complex gradient, and the invariant gradient. As an application, we obtain new maximal and area integral characterizations of Besov spaces. Moreover, we give an atomic decomposition of real-variable type with respect to Carleson tubes for Bergman spaces.
math.FA
math
where wk is the complex conjugate of wk. We also write The open unit ball in Cn is the set Cn = C × ··· × C hz, wi = z1w1 + ··· + znwn, z =pz12 + ··· + zn2. Bn = { z ∈ Cn : z < 1}. Sn = { z ∈ Cn : z = 1}. MAXIMAL AND AREA INTEGRAL CHARACTERIZATIONS OF BERGMAN SPACES IN THE UNIT BALL OF Cn ZEQIAN CHEN AND WEI OUYANG Abstract. In this paper, we present maximal and area integral characterizations of Bergman spaces in the unit ball of Cn. The characterizations are in terms of maximal functions and area integral functions on Bergman balls involving the radial derivative, the complex gradient, and the invariant gradient. As an application, we obtain new maximal and area integral characterizations of Besov spaces. Moreover, we give an atomic decomposition of real-variable type with respect to Carleson tubes for Bergman spaces. 1. Introduction and main results Let C denote the set of complex numbers. Throughout the paper we fix a positive integer n, and let denote the Euclidean space of complex dimension n. Addition, scalar multiplication, and conjugation are defined on Cn componentwise. For z = (z1,··· , zn) and w = (w1,··· , wn) in Cn, we write The boundary of Bn will be denoted by Sn and is called the unit sphere in Cn, i.e., Also, we denote by Bn the closed unit ball, i.e., Bn = {z ∈ Cn : z ≤ 1} = Bn ∪ Sn. The automorphism group of Bn, denoted by Aut(Bn), consists of all bi-holomorphic mappings of Bn. Traditionally, bi-holomorphic mappings are also called automorphisms. For α ∈ R, the weighted Lebesgue measure dvα on Bn is defined by dvα(z) = cα(1 − z2)αdv(z) where cα = 1 for α ≤ −1 and cα = Γ(n + α + 1)/[n!Γ(α + 1)] if α > −1, which is a normalizing constant so that dvα is a probability measure on Bn. In the case of α = −(n + 1) we denote the resulting measure by (1 − z2)n+1 , For α > −1 and p > 0, the (weighted) Bergman space Ap and call it the invariant measure on Bn, since dτ = dτ ◦ ϕ for any automorphism ϕ of Bn. Bn with dτ (z) = α consists of holomorphic functions f in dv kfkp, α =(cid:18)ZBn f (z)pdvα(z)(cid:19)1/p < ∞, 2010 Mathematics Subject Classification: 32A36, 32A50. Key words: Bergman space, Bergman metric, maximal function, area integral function, atomic decomposition. 1 2 Z. Chen and W. Ouyang where the weighted Lebesgue measure dvα on Bn is defined by dvα(z) = cα(1 − z2)αdv(z) and cα = Γ(n + α + 1)/[n!Γ(α + 1)] is a normalizing constant so that dvα is a probability measure on Bn. Thus, Ap α = H(Bn) ∩ Lp(Bn, dvα), where H(Bn) is the space of all holomorphic functions in Bn. When α = 0 we simply write Ap for Ap 0. These are the usual Bergman spaces. Note that for 1 ≤ p < ∞, Ap α is a Banach space under the norm k kp, α. If 0 < p < 1, the space Ap Recall that D(z, γ) denotes the Bergman metric ball at z α is a quasi-Banach space with p-norm kfkp p,α. with γ > 0, where β is the Bergman metric on Bn. It is known that D(z, γ) = {w ∈ Bn : β(z, w) < γ} β(z, w) = 1 2 log 1 + ϕz(w) 1 − ϕz(w) , z, w ∈ Bn, whereafter ϕz is the bijective holomorphic mapping in Bn, which satisfies ϕz(0) = z, ϕz(z) = 0 and ϕz ◦ ϕz = id. As is well known, maximal functions play a crucial role in the real-variable theory of Hardy spaces (cf. [6]). In this paper, we first establish a maximal-function characterization for the Bergman spaces. To this end, we define for each γ > 0 and f ∈ H(Bn) : (1.1) (Mγf )(z) = sup w∈D(z,γ)f (w), ∀z ∈ Bn. Then we have the following result. Theorem 1.1. Suppose γ > 0 and α > −1. Let 0 < p < ∞. Then for any f ∈ H(Bn), f ∈ Ap only if Mγf ∈ Lp(Bn, dvα). Moreover, (1.2) where "≈" depends only on γ, α, p, and n. kfkp,α ≈ kMγfkp,α, α if and The norm appearing on the right-hand side of (1.2) can be viewed an analogue of the so-called nontangential maximal function in Hardy spaces. The proof of Theorem 1.1 is fairly elementary (see §2), using some basic facts and estimates on the Bergman balls. notation. For any f ∈ H(Bn) and z = (z1, . . . , zn) ∈ Bn we define In order to state the area integral characterizations of the Bergman spaces, we require some more and call it the radial derivative of f at z. The complex and invariant gradients of f at z are respectively defined as Now, for fixed 1 < q < ∞ and γ > 0, we define for each f ∈ H(Bn) and z ∈ Bn : (1) The radial area integral function ∇f (z) =(cid:16) ∂f (z) ∂z1 zk , . . . , ∂f (z) ∂f (z) ∂zk Rf (z) = nXk=1 ∂zn (cid:17) and e∇f (z) = ∇(f ◦ ϕz)(0). R (f )(z) = ZD(z,γ) (1 − w2)Rf (w)qdτ (w)! 1 ∇ (f )(z) = ZD(z,γ) (1 − w2)∇f (w)qdτ (w)! 1 . . q q Aγ,q Aγ,q (2) The complex gradient area integral function Bergman spaces 3 (3) The invariant gradient area integral function Aγ,q ∇ (f )(z) = ZD(z,γ) ∇f (w)qdτ (w)! 1 q . We state the second main result of this paper as follows. Theorem 1.2. Suppose 1 < q < ∞, γ > 0, and α > −1. Let 0 < p < ∞. Then, for any f ∈ H(Bn) the following conditions are equivalent: (a) f ∈ Ap α. (b) Aγ,q R (f ) is in Lp(Bn, dvα). (c) Aγ,q ∇ (f ) is in Lp(Bn, dvα). (d) Aγ,q (f ) is in Lp(Bn, dvα). ∇ Moreover, the quantities kAγ,q R (f )kp,α, kAγ,q ∇ (f )kp,α, kAγ,q ∇ (f )kp,α, are all comparable to kf − f (0)kp,α, where the comparable constants depend only on q, γ, α, p, and n. For 0 < p < ∞ and −∞ < α < ∞, we fix a nonnegative integer k with pk + α > −1 and define the generalized Bergman space Ap α as introduced in [8] to be the space of all f ∈ H(Bn) such that (1−z2)kRkf ∈ Lp(Bn, dvα). One then easily observes that Ap α is independent of the choice of k and consistent with the traditional definition when α > −1. Let N be the smallest nonnegative integer such that pN + α > −1. Put kfkp,α = f (0) +(cid:18)ZBn (1 − z2)pNRN f (z)pdvα(z)(cid:19) 1 It is known that the family of the generalized Bergman spaces Ap Equipped with (1.3), Ap holomorphic functions in the unit ball of Cn, such as the classical diagonal Besov space Bs Sobolev space W p (see e.g. [8] for an overview). α becomes a Banach space when p ≥ 1 and a quasi-Banach space for 0 < p < 1. α covers most of the spaces of p and the k,β, which has been extensively studied before in the literature under different names , f ∈ Ap α . (1.3) p There are various characterizations for Bs k,β involving complex-variable quantities in terms of radical derivatives, complex and invariant gradients, and fractional differential operators (for a review and details see [8] and references therein). However, as an application of Theorems 1.1 and 1.2, we obtain new maximal and area integral characterizations of the Besov spaces as follows, which can be considered as a unified characterization for such spaces involving real-variable quantities. p or W p Corollary 1.1. Suppose γ > 0 and α ∈ R. Let 0 < p < ∞ and k be a positive integer such that pk + α > −1. Then for any f ∈ H(Bn), f ∈ Ap (1.4) γ (f ) ∈ Lp(Bn, dvα), where α if and only if M(k) w∈D(z,γ)(1 − w2)kRkf (w), γ (f )(z) = sup z ∈ Bn. M(k) Moreover, kf − f (0)kp,α ≈ kMγ(Rkf )kp,α, (1.5) where "≈" depends only on γ, α, p, k, and n. Corollary 1.2. Suppose 1 < q < ∞, γ > 0 and α ∈ R. Let 0 < p < ∞ and k be a nonnegative integer such that pk + α > −1. Then for any f ∈ H(Bn), f ∈ Ap Rk+1 (f ) is in Lp(Bn, dvα), where α if and only if Aγ,q (1.6) Moreover, Aγ,q Rk+1 (f )(z) = ZD(z,γ)(cid:12)(cid:12)(1 − w2)k+1Rk+1f (w)(cid:12)(cid:12)q q dτ (w)! 1 . (1.7) where "≈" depends only on q, γ, α, p, k, and n. kf − f (0)kp,α ≈ kAγ,q Rk+1(f )kp,α, 4 Z. Chen and W. Ouyang α if and only if Rkf ∈ Lp(Bn, dvα+pk) To prove Corollaries 1.1 and 1.2, one merely notices that f ∈ Ap and applies Theorems 1.1 and 1.2 respectively to Rkf with the help of Lemma 2.1 below. The paper is organized as follows. In Sect. 2 we will prove Theorems 1.1 and 1.2. An atomic decomposition of real-variable type with respect to Carleson tubes for Bergman spaces will be pre- sented in Sect. 3 via duality method. Finally, in Sect. 4, we will prove Theorem 1.2 through using the real-variable atomic decomposition of Bergman spaces established in the preceding section. In what follows, C always denotes a constant depending (possibly) on n, q, p, γ or α but not on f, which may be different in different places. For two nonnegative (possibly infinite) quantities X and Y, by X . Y we mean that there exists a constant C > 0 such that X ≤ CY and by X ≈ Y that X . Y and Y . X. Any notation and terminology not otherwise explained, are as used in [9] for spaces of holomorphic functions in the unit ball of Cn. For the sake of convenience, we collect some elementary facts on the Bergman metric and holo- 2. Proofs of Theorems 1.1 and 1.2 morphic functions in the unit ball of Cn as follows. Lemma 2.1. (cf. [9, Lemma 2.20]) For each γ > 0, for all a and z in Bn with β(a, z) < γ. 1 − a2 ≈ 1 − z2 ≈ 1 − ha, zi Lemma 2.2. (cf. [9, Lemma 2.24]) Suppose γ > 0, p > 0, and α > −1. Then there exists a constant C > 0 such that for any f ∈ H(Bn), C f (z)p ≤ (1 − z2)n+1+αZD(z,γ) f (w)pdvα(w), ∀z ∈ Bn. Lemma 2.3. (cf. [9, Lemma 2.27]) For each γ > 0, for all z in ¯Bn and u, v in Bn with β(u, v) < γ. 1 − hz, ui ≈ 1 − hz, vi 2.1. Proof of Theorem 1.1. We need the following result (cf. [1, Lemma 5]). Lemma 2.4. For fixed γ > 0, there exist a positive integer N and a sequence {ak} in Bn such that (1) Bn = ∪kD(ak, γ), and (2) each z ∈ Bn belongs to at most N of the sets D(ak, 3γ). Proof of Theorem 1.1. Let p > 0. By Lemmas 2.4, 2.2, and 2.1, we have ZBn Mγ(f )(z)pdvα(z) ≤Xk ZD(ak,γ) Mγ(f )(z)pdvα(z) sup 1 sup w∈D(z,γ)f (w)pdvα(z) =Xk ZD(ak,γ) .Xk ZD(ak,γ) .Xk ZD(ak,γ) .Xk ZD(ak,3γ) f (u)pdvα(u) . NZBn f (u)pdvα(u) (1 − w2)n+1+αZD(w,γ) f (u)pdvα(u)dvα(z) (1 − ak2)n+1+αZD(ak,3γ) f (u)pdvα(u)! dvα(z) w∈D(z,γ) 1 where N is the constant in Lemma 2.4 depending only on γ and n. (cid:3) Bergman spaces 5 2.2. Proof of Theorem 1.2. Recall that B(Bn) is defined as the space of all f ∈ H(Bn) so that kfkB = sup z∈Bn ∇f (z) < ∞. B(Bn) with the norm kfk = f (0) + kfkB is a Banach space and called the Bloch space. Then, the following interpolation result holds. Lemma 2.5. (cf. [9, Theorem 3.25]) Let 1 < p < ∞. Suppose α > −1 and 1 p = 1 − θ p′ for 0 < θ < 1 and 1 ≤ p′ < ∞. Then α(Bn) =hAp′ Ap with equivalent norms. α (Bn),B(Bn)iθ Moreover, to prove Theorem 1.2 for the case 0 < p ≤ 1, we will use atom decomposition for Bergman spaces due to Coifman and Rochberg [4] (see also [9, Theorem 2.30]) as follows. Proposition 2.1. Suppose p > 0, α > −1, and b > n max{1, 1/p} + (α + 1)/p. Then there exists a sequence {ak} in Bn such that Ap α consists exactly of functions of the form where {ck} belongs to the sequence space ℓp and the series converges in the norm topology of Ap α. Moreover, f (z) = , z ∈ Bn, ck (1 − hz, aki)b (1 − ak2)(pb−n−1−α)/p ∞Xk=1 ZBn f (z)pdvα(z) ≈ infnXk ckpo, where the infimum runs over all the above decompositions. Also, we need a characterization of Carleson type measures for Bergman spaces as follows, which can be found in [8, Theorem 45]. Proposition 2.2. Suppose n + 1 + α > 0 and µ is a positive Borel measure on Bn. Then, there exists a constant C > 0 such that if and only if for each s > 0 there exists a constant C > 0 such that µ(Qr(ζ)) ≤ Cr2(n+1+α), ∀ζ ∈ Sn and r > 0, ZBn (1 − z2)s 1 − hz, win+1+α+s dµ(w) ≤ C for all z ∈ Bn. We are now ready to prove Theorem 1.2. Note that for any f ∈ H(Bn), (1 − z2)Rf (z) ≤ (1 − z2)∇f (z) ≤ ∇f (z), ∀z ∈ Bn [9, Lemma 2.14]). We have that (d) implies (c), and (c) implies (b) in Theorem 1.2. Then, it (cf. remains to prove that (b) implies (a), and (a) implies (d). Proof of (b) ⇒ (a). Since Rf (z) is holomorphic, by Lemma 2.2 we have C (1 − z2)n+1ZD(z,γ) Rf (w)qdv(w) ≤ CγZD(z,γ) Rf (w)qdτ (w). Rf (z)q ≤ Then, (1 − z2)Rf (z) ≤C(1 − z2) ZD(z,γ) Rf (w)qdτ (w)! 1 ≤Cγ ZD(z,γ) (1 − w2)Rf (w)q dτ (w)! 1 q q = CγAγ,q R (f )(z). 6 Z. Chen and W. Ouyang Hence, for any p > 0, if Aγ,q that f ∈ Ap α (cf. [9, Theorem 2.16]). R (f ) ∈ Lp(Bn, dvα) then (1 − z2)Rf (z) is in Lp(Bn, dvα), which implies (cid:3) atomic decomposition, and then the remaining case via complex interpolation. The proof of (a) ⇒ (d) is divided into two steps. We first prove the case 0 < p ≤ 1 using the Proof of (a) ⇒ (d) for 0 < p ≤ 1. To this end, we write An immediate computation yields that fk(z) = (1 − ak2)(pb−n−1−α)/p . (1 − hz, aki)b and Then we have ∇fk(z) = bak(1 − ak2)(pb−n−1−α)/p (1 − hz, aki)b+1 bhz, aki(1 − ak2)(pb−n−1−α)/p Rfk(z) = (1 − hz, aki)b+1 ∇fk(z)2 =(1 − z2)(∇fk(z)2 − Rfk(z)2) =b2(1 − z2)(1 − ak2)2(pb−n−1−α)/p ak2 − hz, aki2 1 − hz, aki2(b+1) . By Lemmas 2.1 and 2.3 one has Aγ,q ∇ (fk)(z) = ZD(z,γ) ∇fk(w)qdτ (w)! 1 q ≤ b(1 − ak2)(pb−n−1−α)/p(cid:16)ZD(z,γ) Cγb(1 − ak2)(pb−n−1−α)/p , ≤ 1 − hz, akib 1 1 − hw, akiqb dτ (w)(cid:17) 1 q where we have used the fact v(D(z, γ)) ≈ (1 − z2)n+1. Note that vα(Qr) ≈ r2(n+1+α) (cf. Corollary 5.24]), by Proposition 2.2 we have Hence, for 0 < p ≤ 1 we have for f =P∞ ∞Xk=1 (f )(z)pdvα ≤ ∇ ∇ ZBn Aγ,q ZBn Aγ,q ZBn Aγ,q 1 − hz, akipb (fk)(z)pdvα(z) ≤ CbpZBn (1 − ak2)(pb−n−1−α) k=1 ckfk withPk ckp < ∞, ckpZBn Aγ,q (f )(z)pdvα ≤ Cp,α infn ∞Xk=1 This concludes that ∇ ∇ The proof is complete. (fk)(z)pdvα ≤ Cp,α ∞Xk=1 ckp. ckpo ≤ Cp,αZBn f (z)pdvα(z). dvα(z) ≤ Cp,α. Proof of (a) ⇒ (d) for p > 1. Set E = Lq(Bn, χD(0,γ)dτ ; Cn). Consider the operator [9, (cid:3) Note that ϕz(D(0, γ)) = D(z, γ) and the measure dτ is invariant under any automorphism of Bn (cf. [9, Proposition 1.13]), we have T (f )(z, w) =(cid:0) ∇f(cid:1)(ϕz(w)), kT (f )(z)kE =(cid:18)ZBn(cid:12)(cid:12)(cid:0) ∇f(cid:1)(ϕz(w))(cid:12)(cid:12)q f ∈ H(Bn). χD(0,γ)(w)dτ (w)(cid:19) 1 q =(cid:18)ZBn(cid:12)(cid:12) ∇f (w)(cid:12)(cid:12)q χD(z,γ)(w)dτ (w)(cid:19) 1 q = Aγ,q ∇ (f )(z). Bergman spaces 7 On the other hand, Aγ,q ∇ (f )(z) ≤(cid:2)Cγ(1 − z2)−n−1v(D(z, γ))(cid:3) 1 2 kfkB ≤ CkfkB. This follows that T is bounded from B into L∞ from A1 α to L1 α(Bn, E). Thus, by Lemma 2.5 and the well known fact that α (Bn, E). Notice that we have proved that T is bounded Lp(Bn, E) = (L1 α(Bn, E), L∞ α (Bn, E))θ with θ = 1 − α(Bn, E) for any 1 < p < ∞, i.e., , 1 p we conclude that T is bounded from Ap α into Lp ∇ kAγ,q ∀f ∈ Ap α, where C depends only on q, γ, n, p, and α. The proof is complete. (cid:3) Remark 2.1. From the proofs of that (b) =⇒ (a) and that (a) ⇒ (d) for p > 1 we find that Theorem 1.2 still holds true for the Bloch space. That is, for any f ∈ H(Bn), f ∈ B if and only if one (or equivalently, all) of Aγ,q (f )kp,α ≤ Ckfkp,α, R (f ), Aγ,q kfkB ≈ kAγ,q ∇ (f ), and Aγ,q ∇ R (f )kL∞(Bn) ≈ kAγ,q (f ) is (or, are) in L∞(Bn). Moreover, (f )kL∞(Bn), ∇ (f )kL∞(Bn) ≈ kAγ,q ∇ (2.1) where "≈" depends only on q, γ, and n. 3. Atomic decomposition for Bergman spaces We let d(z, w) = 1 − hz, wi 2 , z, w ∈ Bn. 1 It is known that d satisfies the triangle inequality and the restriction of d to Sn is a metric. As usual, d is called the nonisotropic metric. For any ζ ∈ Sn and r > 0, the set Qr(ζ) = {z ∈ Bn : d(z, ζ) < r} is called a Carleson tube with respect to the nonisotropic metric d. We usually write Q = Qr(ζ) in short. As usual, we define the atoms with respect to the Carleson tube as follows: a ∈ Lq(Bn, dvα) is said to be a (1, q)α-atom if there is a Carleson tube Q such that (1) a is supported in Q; (2) kakLq(Bn,dvα) ≤ vα(Q) (3) RBn a(z) dvα(z) = 0. q −1; 1 The constant function 1 is also considered to be a (1, q)α-atom. Note that for any (1, q)α-atom a, for 1 < q < ∞, kak1,α =ZQ advα ≤ vα(Q)1−1/qkakq,α ≤ 1. Recall that Pα is the orthogonal projection from L2(Bn, dvα) onto A2 K α(z, w)f (w)dvα(w), ∀f ∈ L1(Bn, dvα), α > −1, Pαf (z) =ZBn where α, which can be expressed as K α(z, w) = 1 (cid:0)1 − hz, wi(cid:1)n+1+α , z, w ∈ Bn. Pα extends to a bounded projection from Lp(Bn, dvα) onto Ap We have the following useful estimates. α (1 < p < ∞). Lemma 3.1. For α > −1 and 1 < q < ∞ there exists a constant Cq,α,n > 0 such that for any (1, q)α-atom a. kPα(a)k1, α ≤ Cq,α,n To prove Lemma 3.1, we need first to show an inequality for reproducing kernel K α associated with d, which is essentially borrowed from [7, Proposition 2.13]. 8 Z. Chen and W. Ouyang Lemma 3.2. For α > −1 there exists a constant δ > 0 such that for all z, w ∈ Bn, ζ ∈ Sn satisfying d(z, ζ) > δd(w, ζ), we have K α(z, w) − K α(z, ζ) ≤ Cα,n d(w, ζ) d(z, ζ)2(n+1+α)+1 . Proof. Note that We have d dt(cid:18) K α(z, w) − K α(z, ζ) =Z 1 K α(z, w) − K α(z, ζ) ≤Z 1 0 0 (1 − hz, ζi − thz, w − ζi)n+1+α(cid:19) dt. 1 (n + 1 + α)hz, w − ζi 1 − hz, ζi − thz, w − ζin+2+α dt. Write z = z1 + z2 and w = w1 + w2, where z1 and w1 are parallel to ζ, while z2 and w2 are perpendicular to ζ. Then and so Since w1 − ζ = 1 − hw, ζi, hz, wi − hz, ζi = hz2, w2i − hz1, w1 − ζi hz, wi − hz, ζi ≤ z2w2 + w1 − ζ. z22 = z2 − z12 < 1 − z12 < (1 + z1)(1 + z1) ≤ 1 − hz1, ζi = 21 − hz, ζi, and similarly we have w22 ≤ 21 − hw, ζi, hz, wi − hz, ζi ≤ 21 − hz, ζi1/21 − hw, ζi1/2 + 1 − hw, ζi δ(cid:17) 1 = 2d(w, ζ)[d(z, ζ) + d(w, ζ)] ≤ 2(cid:16)1 + 1 δ d2(z, ζ). This concludes that there is δ > 1 such that hz, w − ζi < 1 21 − hz, ζi, ∀z, w ∈ Bn, ζ ∈ Sn, whenever d(z, ζ) > δd(w, ζ). Then, we have 1 − hz, ζi − thz, w − ζi > 1 − hz, ζi − thz, ζ − wi > 1 21 − hz, ζi. Therefore, K α(z, w) − K α(z, ζ) ≤ 2n+3+α(n + 1 + α)(1 + 1/δ)d(w, ζ)d(z, ζ) 1 − hz, ζin+2+α d(w, ζ) d(z, ζ)2(n+1+α)+1 ≤ Cα,n and the lemma is proved. (cid:3) Proof of Lemma 3.1. When a is the constant function 1, the result is clear. Thus we may suppose a is a (1, q)α-atom. Let a be supported in a Carleson tuber Qr(ζ) and δr ≤ √2, where δ is the constant in Lemma 3.2. Since Pα is a bounded operator on Lq(Bn, dvα), we have ZQδr Pα(a)dvα(z) ≤ vα(Qδr)1− 1 q kPα(a)kq, α ≤ kPαkLq vα(Qδr)1− 1 q kakq,α ≤ kPαkLq . Next, if d(z, ζ) > δr then Bergman spaces 9 (cid:12)(cid:12)(cid:12)ZBn 1 a(w) (1 − hz, wi)n+1+α dvα(w)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) a(w)(cid:20) ZQr (ζ) ≤ CZQr(ζ) a(w) ≤ CrZQr (ζ) a(w)dvα(w) (1 − hz, wi)n+1+α − d(z, ζ)2(n+1+α)+1 dvα(w) d(w, ζ) 1 (1 − hz, ζi)n+1+α(cid:21) dvα(w)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 d(z, ζ)2(n+1+α)+1 ≤ Cr d(z, ζ)2(n+1+α)+1 . Then Zd(z, ζ)>δrPα(a)dvα(z) ≤ CrZd(z, ζ)>δr = CrXk≥0Z2kδr<d(z, ζ)≤2k+1δr ≤ CrXk≥0 vα(Q2k+1δr) (2kδr)2(n+1+α)+1 ≤ Cr 1 d(z, ζ)2(n+1+α)+1 dvα(z) 1 d(z, ζ)2(n+1+α)+1 dvα(z) (2k+1δr)2(n+1+α) (2kδr)2(n+1+α)+1 ≤ C, ∞Xk=0 where we have used the fact that vα(Qr) ≈ r2(n+1+α) in the third inequality (cf. [9, Corollary 5.24]). Thus, we get ZBn Pα(a)dvα(z) =ZQδr Pα(a)dvα(z) +Zd(z, ζ)>δr Pα(h)dvα(z) ≤ C, where C depends only on q, n and α. (cid:3) Recall that kak1,α ≤ 1 for any (1, q)α-atom a. Then, we define A1,q admits a decomposition Now we turn to the atomic decomposition of A1 and Xi f =Xi λi ≤ Cqkfk1, α, λiPαai α (α > −1) with respect to the Carleson tubes. α which α as the space of all f ∈ A1 where for each i, ai is an (1, q)α-atom and λi ∈ C so thatPi λi < ∞. We equip this space with the norm kfkA1,q α = infnXi λi : f =Xi λiPαaio where the infimum is taken over all decompositions of f described above. α is a Banach space. By Lemma 3.1 we have the contractive inclusion α. We will prove in what follows that these two spaces coincide. That establishes the α. In fact, we will show the remaining It is easy to see that A1,q A1,q α ⊂ A1 "real-variable" atomic decomposition of the Bergman space A1 inclusion A1 Theorem 3.1. Let 1 < q < ∞ and α > −1. For every f ∈ A1 (1, q)α-atoms and a sequence {λi} of complex numbers such that (3.1) α by duality. α ⊂ A1,q α there exist a sequence {ai} of λiPαai λi ≤ Cqkfk1, α. Moreover, f =Xi and Xi kfk1, α ≈ infXi λi where the infimum is taken over all decompositions of f described above and " ≈ " depends only on α, n, and q. 10 Z. Chen and W. Ouyang α can be identified with B (with equivalent norms) under the integral pairing α is the Bloch space B (we refer to [9] for details). The Banach f (rz)g(z)dvα(z), f ∈ A1 α, g ∈ B. dual of A1 Recall that the dual space of A1 r→1−ZBn hf, giα = lim (cf. [9, Theorem 3.17].) In order to prove Theorem 3.1, we need the following result, which can be found in [1] (see also [9, Theorem 5.25]). Lemma 3.3. Suppose α > −1 and 1 ≤ p < ∞. Then, for any f ∈ H(Bn), f is in B if and only if there exists a constant C > 0 depending only on α and p such that 1 vα(Qr(ζ))ZQr (ζ) f − fα, Qr (ζ)pdvα ≤ C fα, Qr (ζ) = 1 Qr(ζ)ZQr (ζ) f (z)dvα(z). for all r > 0 and all ζ ∈ Sn, where Moreover, kfkB ≈ sup r>0,ζ∈S(cid:16) 1 vα(Qr(ζ))ZQr (ζ) f − fα, Qr (ζ)pdvα(cid:17) 1 p , where "≈" depends only on α, p, and n. As noted above, we will prove Theorem 3.1 via duality. To this end, we first prove the following precisely, duality theorem. Proposition 3.1. For any 1 < q < ∞ and α > −1, we have (A1,q (i) Every g ∈ B defines a continuous linear functional ϕg on A1,q (3.2) f (rz)g(z)dvα(z), ϕg(f ) = lim α by ∀f ∈ A1,q α . r→1−ZBn (ii) Conversely, each ϕ ∈ (A1,q α )∗ is given as (3.2) by some g ∈ B. Moreover, we have α )∗ = B isometrically. More (3.3) Proof. Let p be the conjugate index of q, i.e., 1/p + 1/q = 1. We first show B ⊂ (A1,q For any (1, q)α-atom a, by Lemma 3.3 we have kϕgk ≈ g(0) + kgkB, ∀g ∈ B. α )∗. Let g ∈ B. (cid:12)(cid:12)(cid:12)(cid:12)ZBn Pαa(z)g(z)dvα(z)(cid:12)(cid:12)(cid:12)(cid:12) = hPα(aj), giα =(cid:12)(cid:12)(cid:12)(cid:12)ZBn agdvα(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)ZBn ≤(cid:18)ZQ aqdvα(cid:19)1/q(cid:18)ZQ g − gQpdvα(cid:19)1/p vα(Q)ZQ g − gQpdvα(cid:19)1/p ≤(cid:18) 1 ≤ CkgkB. a(g − gQ)dvα(cid:12)(cid:12)(cid:12)(cid:12) On the other hand, for the constant function 1 we have Pα1 = 1 and so Thus, we deduce that (cid:12)(cid:12)(cid:12)(cid:12)ZBn Pα1(z)g(z)dvα(z)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)ZBn (cid:12)(cid:12)(cid:12)(cid:12)ZBn f ¯gdvα(cid:12)(cid:12)(cid:12)(cid:12) ≤ CkfkA1,q α g(z)dvα(z)(cid:12)(cid:12)(cid:12)(cid:12) = g(0). (g(0) + kgkB) ϕg(f ) ≤ C(g(0) + kgkB)kfkA1, q α for any finite linear combination f of (1, q)α-atoms. Hence, g defines a continuous linear functional ϕg on a dense subspace of A1, q α such that α and ϕg extends to a continuous linear functional on A1, q Bergman spaces 11 for all f ∈ A1, q α . Next let ϕ be a bounded linear functional on A1, q α . Note that Hq(Bn, dvα) = H(Bn) ∩ Lq(Bn, dvα) ⊂ A1, q α . Then, ϕ is a bounded linear functional on Hq(Bn, dvα). By duality there exists g ∈ Hp(Bn, dvα) such that ϕ(f ) =ZBn f ¯gdvα, ∀f ∈ Hq(Bn, dvα). Let Q = Qr(ζ) be a Carleson tube. For any f ∈ Lq(Bn, dvα) supported in Q, it is easy to check that af = (f − fQ)χQ/[kfkLqvα(Q)1/p] is a (1, q)-atom. Then, ϕ(Pαaf ) ≤ kϕk and so Hence, for any f ∈ Lq(Bn, dvα) we have (cid:12)(cid:12)(cid:12)(cid:12)ZQ (cid:12)(cid:12)ϕ(Pα[(f − fQ)χQ])(cid:12)(cid:12) ≤ kϕkkfkLqvα(Q)1/p. f (g − gQ)dvα(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)ZQ (f − fQ)¯gdvα(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)ZBn (f − fQ)χQ¯gdvα(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)ZBn Pα[(f − fQ)χQ]¯gdvα(cid:12)(cid:12)(cid:12)(cid:12) = ϕ(Pα[(f − fQ)χQ]) vα(Q)ZQ g − gQp dvα(cid:19)1/p (cid:18) 1 ≤ 2kϕk. This concludes that ≤ kϕk k(f − fQ)χQkLq(Bn, dvα) vα(Q)1/p ≤ 2kϕk kfkLq(Q, dvα) vα(Q)1/p. By Lemma 3.3 we have that g ∈ B and kgkB ≤ Ckϕk. Therefore, ϕ is given as (3.2) by g with g(0) + kgkB ≤ Ckϕk. (cid:3) Now we are ready to prove Theorem 3.1. Proof of Theorem 3.1. By Lemma 3.1 we know that A1,q α ⊂ A1 α)∗ = (A1,q α )∗. Hence, by duality we have kfk1,q ≈ kfkA1,q 3.1 we have (A1 α α. On the other hand, by Proposition (cid:3) . Remark 3.1. (1) One would like to expect that when 0 < p < 1, Ap α also admits an atomic decomposition in terms of atoms with respect to Carleson tubes. However, the proof of Theorem 3.1 via duality cannot be extended to the case 0 < p < 1. At the time of this writing, this problem is entirely open. (2) The real-variable atomic decomposition of Bergman spaces should be known to specialists in the case p = 1. Indeed, based on their theory of harmonic analysis on homogeneous spaces, Coifman and Weiss [5] claimed that the Bergman space A1 admits an atomic decomposition in terms of atoms with respect to (Bn, , dv), where (z, w) = (cid:12)(cid:12)z − w(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)1 − 1 zwhz, wi(cid:12)(cid:12)(cid:12), otherwise. z + w, This also applies to A1 α because (Bn, , dvα) is a homogeneous space for α > −1 (see e.g. [7]). However, the approach of Coifman and Weiss is again based on duality and therefore not constructive and cannot be applied to the case 0 < p < 1. Recently, the present authors [3] extend this result to the case 0 < p < 1 through using a constructive method. if z, w ∈ Bn\{0}, 12 Z. Chen and W. Ouyang 4. Area integral inequalities: Real-variable methods In this section, we will prove the area integral inequality for the Bergman space A1 α via atomic decomposition established in Section 3. Theorem 4.1. Suppose 1 < q < ∞, γ > 0, and α > −1. Then, (4.1) kAγ,q ∇ (f )k1,α . kfk1,α, ∀f ∈ H(Bn). to involve a real-variable method. This is the assertion (a) ⇒ (d) of Theorem 1.2 in the case p = 1. The novelty of the proof here is The following lemma is elementary. Lemma 4.1. Suppose 1 < q < ∞, γ > 0 and α > −1. If f ∈ Aq α, then ZBn Aγ,q ∇ (f )(z)qdvα ≈ZBn f (z) − f (0)qdvα, where " ≈ " depends only on q, γ, α, and n. Proof. Note that vα(D(z, γ)) ≈ (1 − z2)n+1+α. Then ZBn Aγ,q ∇ (f )(z)qdvα =ZBnZD(z,γ) (1 − w2)−1−n ∇f (w)qdv(w)dvα(z) =ZBn vα(D(w, γ))(1 − w2)−1−n ∇f (w)qdv(w) ≈ZBn ∇f (w)qdvα(w) ≈ZBn f (w) − f (0)qdvα(w). In the last step we have used [9, Theorem 2.16 (b)]. (cid:3) Proof of (4.1). By Theorem 3.1, it suffices to show that for 1 < q < ∞, γ > 0, and α > −1 there exists C > 0 such that for all (1, q)α-atoms a. Given an (1, q)α-atom a supported in Q = Qr(ζ), by Lemma 4.1 we have kAγ,q ∇ (Pαa)k1,α ≤ C Z2Q Aγ,q ∇ (Pαa)dvα ≤ vα(2Q)1− 1 ≤ Cvα(Q)1− 1 ≤ Cvα(Q)1− 1 q ∇ (Pαa)(cid:3)q q(cid:16)Z2Q(cid:2)Aγ,q q(cid:16)ZBn Pαa(z) − Pαa(0)qdvα(cid:17) 1 dvα(cid:17) 1 q q kakq,α ≤ C, where 2Q = Q2r(ζ). On the other hand, dvα(z) q Aγ,q ∇ (Pαa)dvα =Z(2Q)c(cid:16)ZD(z,γ) ∇Pαa(w)qdτ (w)(cid:17) 1 Z(2Q)c =Z(2Q)c(cid:16)ZD(z,γ)(cid:12)(cid:12)(cid:12)ZQ ∇w[K α(w, u) − K α(w, ζ)]a(u)dvα(u)(cid:12)(cid:12)(cid:12) ≤kakq,αZ(2Q)c(cid:16)ZD(z,γ)(cid:16)ZQ ∇w[K α(w, u) − K α(w, ζ)] ≤Z(2Q)c(cid:16)ZD(z,γ) u∈Q ∇w[K α(w, u) − K α(w, ζ)]qdτ (w)(cid:17) 1 sup q q q dτ (w)(cid:17) 1 q−1 dvα(u)(cid:17)q−1 q dvα(z), dvα(z) q dτ (w)(cid:17) 1 dvα(z) where (2Q)c = Bn\2Q. An immediate computation yields that Bergman spaces 13 ∇w[K α(w, u) − K α(w, ζ)] =(n + 1 + α)h ¯u =(n + 1 + α) ¯ζ (1 − hw, ζi)n+2+αi (1 − hw, ui)n+2+α − ¯u(1 − hw, ζi)n+2+α − ¯ζ(1 − hw, ui)n+2+α (1 − hw, ui)n+2+α(1 − hw, ζi)n+2+α and Moreover, and Rw[K α(w, u) − K α(w, ζ)] =(n + 1 + α)h hw, ui =(n + 1 + α)hw, ui(1 − hw, ζi)n+2+α − hw, ζi(1 − hw, ui)n+2+α (1 − hw, ζi)n+2+αi (1 − hw, ui)n+2+α − hw, ζi (1 − hw, ui)n+2+α(1 − hw, ζi)n+2+α . (cid:12)(cid:12)∇w[K α(w, u) − K α(w, ζ)](cid:12)(cid:12)2 =(n + 1 + α)2nu21 − hw, ζi2(n+2+α) + 1 − hw, ui2(n+2+α) 1 − hw, ui2(n+2+α)1 − hw, ζi2(n+2+α) − − (1 − hw, ζi)n+2+α(1 − hu, wi)n+2+αhζ, ui 1 − hw, ui2(n+2+α)1 − hw, ζi2(n+2+α) 1 − hw, ui2(n+2+α)1 − hw, ζi2(n+2+α) o, (1 − hw, ui)n+2+α(1 − hζ, wi)n+2+αhu, ζi (cid:12)(cid:12)Rw[K α(w, u) − K α(w, ζ)](cid:12)(cid:12)2 = (n + 1 + α)2nhw, ui21 − hw, ζi2(n+2+α) + hw, ζi21 − hw, ui2(n+2+α) − hw, uihζ, wi(1 − hw, ζi)n+2+α(1 − hu, wi)n+2+α − hw, ζihu, wi(1 − hw, ui)n+2+α(1 − hζ, wi)n+2+α 1 − hw, ui2(n+2+α)1 − hw, ζi2(n+2+α) 1 − hw, ui2(n+2+α)1 − hw, ζi2(n+2+α) 1 − hw, ui2(n+2+α)1 − hw, ζi2(n+2+α) o. Then, we have (cid:12)(cid:12)∇w[K α(w, u) − K α(w, ζ)](cid:12)(cid:12)2 = −(cid:12)(cid:12)Rw[K α(w, u) − K α(w, ζ)](cid:12)(cid:12)2 (n + 1 + α)2 1 − hw, ui2(n+2+α)1 − hw, ζi2(n+2+α) ×n(u2 − hw, ui2)1 − hw, ζi2(n+2+α) + (1 − hw, ζi2)1 − hw, ui2(n+2+α) + (hw, uihζ, wi − hζ, ui)(1 − hw, ζi)n+2+α(1 − hu, wi)n+2+α + (hw, ζihu, wi − hu, ζi)(1 − hw, ui)n+2+α(1 − hζ, wi)n+2+αo. Note that for any f ∈ H(Bn), ∇f (z)2 = (1 − z2)(∇f (z)2 − Rf (z)2), z ∈ Bn 14 Z. Chen and W. Ouyang (cf. [9, Lemma 2.13]). It is concluded that (cid:12)(cid:12) ∇w[K α(w, u) − K α(w, ζ)](cid:12)(cid:12) (n + 1 + α)(1 − w2) 1 2 ≤ 1 − hw, uin+2+α1 − hw, ζin+2+α ×n(1 − hw, ui2)1 − hw, ζi2(n+2+α) + (1 − hw, ζi2)1 − hw, ui2(n+2+α) +(cid:2)hw, u − ζihζ, wi + (hw, ζi2 − 1) + (1 − hζ, ui)(cid:3) × (1 − hw, ζi)n+2+α(1 − hu, wi)n+2+α +(cid:2)hw, ζ − uihu, wi + (hw, ui2 − 1) + (1 − hu, ζi)(cid:3) × (1 − hw, ui)n+2+α(1 − hζ, wi)n+2+αo 1 (n + 1 + α)(1 − w2) 2 (M1 + M2 + M3 + M4) 1 2 , 2 1 1 − hw, uin+2+α1 − hw, ζin+2+α ≤ where for w ∈ D(z, γ), u ∈ Qr(ζ) and z ∈ Bn, ζ ∈ Sn. M1 =1 − hw, ζin+2+α1 − hu, win+2+α hw, u − ζihζ, wi + (1 − hζ, ui) , M2 =1 − hw, uin+2+α1 − hζ, win+2+α hw, ζ − uihu, wi + (1 − hu, ζi) , M3 =(1 − hw, ui2)1 − hζ, win+2+α(cid:12)(cid:12)(1 − hw, ζi)n+2+α − (1 − hw, ui)n+2+α(cid:12)(cid:12) , M4 =(1 − hw, ζi2)1 − hu, win+2+α(cid:12)(cid:12)(1 − hw, ui)n+2+α − (1 − hw, ζi)n+2+α(cid:12)(cid:12) , Z(2Q)c dτ (w)(cid:17) 1 u∈Q(cid:12)(cid:12) ∇w[K α(w, u) − K α(w, ζ)](cid:12)(cid:12)q (Pαa)dvα ≤Z(2Q)c(cid:16)ZD(z,γ) ≤(n + 1 + α)Z(2Q)c (I1 + I2 + I3 + I4)dvα(z), Aγ,q ∇ sup q Hence, dvα(z) where We first estimate I1. Note that q q q q 2 2 q 2 3 q 2 M q 2 1 q 2 4 q 2 M q 2 M 2 M sup u∈Q sup u∈Q sup u∈Q sup u∈Q , , , . (1 − w2) (1 − w2) (1 − w2) (1 − w2) I1 =(cid:16)ZD(z,γ) I2 =(cid:16)ZD(z,γ) I3 =(cid:16)ZD(z,γ) I4 =(cid:16)ZD(z,γ) 1 − hw, uiq(n+2+α)1 − hw, ζiq(n+2+α) dτ (w)(cid:17) 1 1 − hw, uiq(n+2+α)1 − hw, ζiq(n+2+α) dτ (w)(cid:17) 1 1 − hw, uiq(n+2+α)1 − hw, ζiq(n+2+α) dτ (w)(cid:17) 1 1 − hw, uiq(n+2+α)1 − hw, ζiq(n+2+α) dτ (w)(cid:17) 1 M1 ≤ (hw, u − ζi + 1 − hζ, ui)1 − hw, ζin+2+α1 − hw, uin+2+α ≤(cid:16)21 − hu, ζi × 1 − hw, ζin+2+α1 − hw, uin+2+α ≤(cid:18)21 − hu, ζi 1 21 − hz, ζi × 1 − hw, ζin+2+α1 − hw, uin+2+α ≤(cid:16)Cγr1 − hz, ζi 2(cid:1) + 1 − hζ, ui(cid:17) 2(cid:0)1 − hw, ζi 2(cid:1) + 1 − hζ, ui(cid:19) 2(cid:0)Cγ1 − hz, ζi 2 + r2(cid:17) 1 − hw, ζin+2+α1 − hw, uin+2+α, 2 + 1 − hu, ζi 2 + q q 1 1 1 1 1 1 1 where the second inequality is the consequence of the following fact which has appeared in the proof of Lemma 3.2 Bergman spaces 15 2 (1 − hw, ζi the third inequality is obtained by Lemma 2.3 and the fact hw, u − ζi ≤ 21 − hu, ζi 1 1 2 + 1 − hu, ζi 1 2 ); 1 − hu, ζi 1 2 < r < 1 21 − hz, ζi 1 2 for u ∈ Q and z ∈ (2Q)c. Since 1 1 − hz, ui 1 2 ≥ 1 − hz, ζi ≥ 1 − hz, ζi 2 − 1 − hu, ζi 1 2 − 21 − hz, ζi 1 1 2 by Lemmas 2.1 and 2.3 we have 1 2 ≥ 1 21 − hz, ζi 1 2 , 2 2 1 q q 2 (n+2+α) 2 + r2(cid:3) q 2 + r2(cid:1) q ≤ Cγ(cid:0)r1 − hz, ζi 1 − hz, ζin+ 3 dτ (w)! 1 dτ (w)! 1 2 + r2(cid:1) 1 q 2 (n+2+α) 2 +α q 1 2 sup u∈Q I1 ≤ ZD(z,γ) ≤Cγ ZD(z,γ) ≤Cγ (1 − z2) ≤Cγ(cid:16) I1dvα(z) .Z(2Q)c (1 − w2) 1 − hw, ui q sup u∈Q q (1 − z2) 1 − hz, ui 2(cid:0)r1 − hz, ζi q 1 − hz, ζiq(n+2+α) r 1 2 + 2 1 1 q q q 2 (n+2+α)1 − hw, ζi 2 (n+2+α)1 − hz, ζi 2(cid:2)Cr1 − hz, ζi 2(cid:0)Cr1 − hz, ζi ! 1 2 + r2(cid:1) q d(z, ζ)2(n+1+α)+1(cid:17). dvα(z) +Z(2Q)c r 2 1 2 r d(z, ζ)2(n+1+α)+ 1 d(z, ζ)2(n+1+α)+ 1 2 Hence, Z(2Q)c r d(z, ζ)2(n+1+α)+1 dvα(z). The second term on the right hand side has been estimated in the proof of Lemma 3.1. The first term can be estimated as follows: Zd(z, ζ)>2r r1/2 d(z, ζ)2(n+1+α)+1/2 dvα(z) = r1/2Xk≥0Z2kr<d(z, ζ)≤2k+1r ≤ r1/2Xk≥0 (2kr)2(n+1+α)+1/2 ≤ Cr1/2 ∞Xk=0 vα(Q2k+1r) (2k+1r)2(n+1+α) (2kr)2(n+1+α)+1/2 ≤ C, 1 d(z, ζ)2(n+1+α)+1/2 dvα(z) where we have used the fact that vα(Qr) ≈ r2(n+1+α) in the third inequality (cf. [9, Corollary 5.24]). By the same argument we can estimate I2 and omit the details. Next, we estimate I3. Note that M3 ≤(1 − hw, ui2)1 − hw, ζin+2+α(cid:12)(cid:12)(1 − hw, ζi)n+2+α − (1 − hw, ui)n+2+α(cid:12)(cid:12) (1 − hw, tζ + (1 − t)ui)n+2+αdt(cid:12)(cid:12)(cid:12)(cid:12) ≤21 − hw, ui1 − hw, ζin+2+α(cid:12)(cid:12)(cid:12)(cid:12)Z 1 ×(cid:12)(cid:12)(cid:12)(cid:12)hw, ζ − uiZ 1 (1 − hw, tζ + (1 − t)ui)n+1+αdt(cid:12)(cid:12)(cid:12)(cid:12) ≤Cγ1 − hw, ui1 − hw, ζin+2+αr1 − hz, ζin+3/2+α, =2(n + 2 + α)1 − hw, ui1 − hw, ζin+2+α d dt 0 0 where the last inequality is achieved by the following estimates 1 − hw, tζ + (1 − t)ui ≤ Cγ1 − hz, tζ + (1 − t)ui ≤ Cγ1 − hz, ui + hz, ζ − ui ≤ Cγ1 − hz, ζi as shown above. conclude that Similarly, we can estimate I4 and omit the details. Therefore, combining above estimates we Z(2Q)c Aγ( ∇Pαa)dvα ≤ C, 16 and Z. Chen and W. Ouyang hw, ζ − ui ≤ Cγr1 − hz, ζi 1 2 , q q q 2 r for any w ∈ D(z, γ) and u ∈ Qr(ζ). Thus, by Lemmas 2.1 and 2.3 2 (n+ 3 2 1 − hw, ζi 2 (n+ 3 2 1 − hz, ζi 1 − hw, uiq(n+1+α)+ q 2 1 − hz, ζi 2 1 − hz, ζi (1 − w2) sup u∈Q sup u∈Q I3 ≤Cγ ZD(z,γ) ≤Cγ ZD(z,γ) (1 − z2) ≤Cγ(cid:16) 4 q(cid:17) 1 1 − hz, ζiq(n+1+α)+ 3 Z(2Q)c I3dvα(z) ≤ CγZ(2Q)c (1 − z2) 1 − hz, uiq(n+1+α)+ q 2 r ≤ Cγ Hence, 2 r r q 2 q q q 1 2 q q d(z, ζ)2(n+1+α)+ 1 2 . 1 2 r d(z, ζ)2(n+1+α)+ 1 2 dvα(z) ≤ Cγ, q q 2 +α) 2 (n+2+α) dτ (w)! 1 2 (n+2+α) dτ (w)! 1 2 +α) q q where C depends only on q, γ, n, and α. Remark 4.1. We remark that whenever Ap α have an atomic decomposition in terms of atoms with respect to Carleson tubes for 0 < p < 1, the argument of Theorem 1.2 works as well in this case. However, as noted in Remark 3.1 (1), the problem of the atomic decomposition of Ap α with respect to Carleson tubes in 0 < p < 1 is entirely open. Remark 4.2. The area integral inequality in case 1 < p < ∞ can be also proved through using the method of vector-valued Calder´on-Zygmund operators for Bergman spaces. This has been done in [2]. (cid:3) Acknowledgement. This research was supported in part by the NSFC under Grant No. 11171338. References [1] D. B´ekoll´e, C. Berger, L. Coburn, and K. Zhu, BMO in the Bergman metric on bounded symmetric domains, J. Funct. Anal. 93 (1990), 310-350. [2] Zeqian Chen and W. Ouyang, Real-variable characterizations of Bergman spaces, Acta Analysis Functionalis Applicata 13 (2011), 246-259. [3] Zeqian Chen and Wei Ouyang, Atomic decomposition of real-variable type for Bergman spaces in the unit ball of Cn, arXiv: 1303.2182. [4] R. Coifman and R. Rochberg, Representation theorems for holomorphic and harmonic functions in Lp, Asterisque 77 (1980), 11-66. [5] R. Coifman and G. Weiss, Extension of Hardy spaces and their use in analysis, Bull. Amer. Math. Soc. 83 (1977), 569-643. [6] E. Stein, Harmonic analysis: Real-Variable Methods, Orthogonality, and Oscillatory Integrals, Princeton Univer- sity Press, Princeton, NJ, 1993. [7] E. Tchoundja, Carleson measures for the generalized Bergman spaces via a T (1)-type theorem, Ark. Mat. 46 (2008), 377-406. [8] R. Zhao and K. Zhu, Theory of Bergman spaces in the unit ball of Cn, Memoires de la Soc.Math.France 115 (2008), pp.103. [9] K. Zhu, Spaces of Holomorphic Functions in the Unit Ball, Springer-Verlag, New York, 2005. Wuhan Institute of Physics and Mathematics, Chinese Academy of Sciences, 30 West District, Xiao- Hong-Shan, Wuhan 430071, China Wuhan Institute of Physics and Mathematics, Chinese Academy of Sciences, 30 West District, Xiao- Hong-Shan, Wuhan 430071, China and Graduate University of Chinese Academy of Sciences, Beijing 100049, China
1006.5532
1
1006
2010-06-29T08:09:17
Ultraregular generalized functions
[ "math.FA", "math.AP" ]
Algebras of ultradifferentiable generalized functions are introduced. We give a microlocal analysis within these algebras related to the regularity type and the ultradifferentiable property.
math.FA
math
ULTRAREGULAR GENERALIZED FUNCTIONS KHALED BENMERIEM AND CHIKH BOUZAR Abstract. Algebras of ultradifferentiable generalized functions are introduced. We give a microlocal analysis within these algebras related to the regularity type and the ultradifferentiable property. 1. Introduction The introduction by J. F. Colombeau of the algebra of generalized functions G (Ω) , see [5], containing the space of distributions as a subspace and having the algebra of smooth functions as a subalgebra, has initiated different directions of research in the field of differential algebras of generalized functions, see [1], [7], [13], [14] and [3]. The current research of the regularity problem in algebras of generalized func- tions of Colombeau type is based on the Oberguggenberger subalgebra G∞ (Ω) , see [6], [9], [10] and [12]. This subalgebra plays the same role as C∞ (Ω) in D′ (Ω) , and has indicated the importance of the asymptotic behavior of the representa- tive nets of a Colombeau generalized function in studying regularity problems. However, the G∞−regularity does not exhaust the regularity questions inherent to the Colombeau algebra G (Ω) , see [15]. The purpose of this work is to introduce and to study new algebras of gener- alized functions of Colombeau type, denoted by GM,R (Ω) , measuring regularity both by the asymptotic behavior of the nets of smooth functions representing a Colombeau generalized function and by their ultradifferentiable smoothness of . Elements of GM,R (Ω) are called ultrareg- Denjoy-Carleman type M = (Mp)p∈Z+ ular generalized functions. This paper is the composition of our two papers [4] and [2]. 2000 Mathematics Subject Classification. Primary 46F30, secondary 35A18, 46F10. Key words and phrases. Ultradifferentiable functions, Denjoy-Carleman classes, Product of distributions, Generalized functions, Colombeau algebra, Wave front, Microlocal analysis, Gevrey generalized functions. 1 2 KHALED BENMERIEM AND CHIKH BOUZAR 2. Regular generalized functions Let Ω be a non void open subset of Rn, define X (Ω) as the space of elements +,∃m ∈ Z+, (uε)ε of C ∞ (Ω)]0,1] such that, for every compact set K ⊂ Ω, ∀α ∈ Zn ∃C > 0,∃η ∈ ]0, 1] ,∀ε ∈ ]0, η] , x∈K ∂αuε (x) ≤ Cε−m. sup By N (Ω) we denote the elements (uε)ε ∈ X (Ω) such that for every compact set K ⊂ Ω,∀α ∈ Zn +,∀m ∈ Z+, ∃C > 0,∃η ∈ ]0, 1] ,∀ε ∈ ]0, η] , x∈K ∂αuε (x) ≤ Cεm. sup Definition 1. The Colombeau algebra, denoted by G (Ω), is the quotient algebra G (Ω) = X (Ω) N (Ω) . G (Ω) is a commutative and associative differential algebra containing D′ (Ω) as a subspace and C ∞ (Ω) as a subalgebra. The subalgebra of generalized func- tions with compact support, denoted GC (Ω) , is the space of elements f of G (Ω) satisfying : there exist a representative (fε)ε∈]0,1] of f and a compact subset K of Ω,∀ε ∈ ]0, 1] , suppfε ⊂ K. guggenberger in [14], as the quotient algebra X ∞ (Ω) N (Ω) One defines the subalgebra of regular elements G∞ (Ω), introduced by Ober- , where X ∞ (Ω) is the space of elements (uε)ε of C ∞ (Ω)]0,1] such that, for every compact K ⊂ Ω, ∃m ∈ Z+,∀α ∈ Zn +, ∃C > 0,∃η ∈ ]0, 1] ,∀ε ∈ ]0, η] , x∈K ∂αuε (x) ≤ Cε−m. sup It is proved in [14] the following fundamental result G∞ (Ω) ∩ D′ (Ω) = C ∞ (Ω) . This means that the subalgebra G∞ (Ω) plays in G (Ω) the same role as C ∞ (Ω) in D′ (Ω) , consequently one can introduce a local analysis by defining the gen- eralized singular support of u ∈ G (Ω) . This was the first notion of regularity in Colombeau algebra. Recently, different measures of regularity in algebras of generalized functions have been proposed, see [6], [12] and [15]. For our needs we recall the essential definitions and results on R-regular generalized functions, see [6]. (R3) Nl1 + N ′ l2 ≤ N"l1+l2 , ∀ (l1, l2) ∈ Z2 +. Define the R−regular moderate elements of X (Ω) , by X R (Ω) =(cid:26)(uε)ε ∈ X (Ω) ∀K ⊂⊂ Ω,∃N ∈ R,∀α ∈ Zn +, ∃C > 0,∃η ∈ ]0, 1] ,∀ε ∈ ]0, η] : sup x∈K ∂αuε (x) ≤ Cε−Nα(cid:27) ULTRAREGULAR GENERALIZED FUNCTIONS 3 m)m∈Z+ be two elements of R m. A non void subspace R of R Definition 2. Let (Nm)m∈Z+ , (N ′ (N ′ m)m∈Z+ , if ∀m ∈ Z+, Nm ≤ N ′ ular, if the following (R1)-(R3) are all satisfied : For all (Nm)m∈Z+ ∈ R and (k, k′) ∈ Z2 Nm+k + k′ ≤ N ′ m)m∈Z+ in R , there exists (N"m)m∈Z+ ∈ R such that +, there exists (N ′ m , ∀m ∈ Z+. m)m∈Z+ ∈ R such that For all (Nm)m∈Z+ and (N ′ (R1) Z+ + , we write (Nm)m∈Z+ ≤ Z+ + is said reg- (R2) max (Nm, N ′ m) ≤ N"m , ∀m ∈ Z+. For all (Nm)m∈Z+ and (N ′ m)m∈Z+ in R , there exists (N"m)m∈Z+ ∈ R such that and its ideal N R (Ω) =(cid:26)(uε)ε ∈ X (Ω) ∀K ⊂⊂ Ω,∀N ∈ R,∀α ∈ Zn +, ∃C > 0,∃η ∈ ]0, 1] ,∀ε ∈ ]0, η] : sup x∈K ∂αuε (x) ≤ CεNα(cid:27) . Proposition 1. 1. The space X R (Ω) is a subalgebra of X (Ω), stable by differ- entiation. 2. The set N R (Ω) is an ideal of X R (Ω) . Remark 1. From (R1), one can show that N R (Ω) = N (Ω) . Definition 3. The algebra of R-regular generalized functions, denoted by GR (Ω) , is the quotient algebra GR (Ω) = X R (Ω) N (Ω) . Z+ Example 1. The Colombeau algebra G (Ω) is obtained when R = R G R + (Ω) = G (Ω) . Example 2. When A =n(Nm)m∈Z+ ∈ R : ∃a ≥ 0,∃b ≥ 0, Nm ≤ am + b,∀m ∈ Z+o , we obtain a differential subalgebra denoted by GA (Ω). Z+ + , i.e. 4 KHALED BENMERIEM AND CHIKH BOUZAR Example 3. When R = B the set of all bounded sequences, we obtain the Ober- guggenberger subalgebra, i.e. GB (Ω) = G∞ (Ω). Example 4. If we take R = {0}, the condition (R1) is not hold, however we can define G0 (Ω) = X R (Ω) N (Ω) as the algebra of elements which have all derivatives locally uniformly bounded for small ε, see [15]. We have the following inclusions G0 (Ω) ⊂ GB (Ω) ⊂ GA (Ω) ⊂ G (Ω) . 3. Ultradifferentiable functions We recall some classical results on ultradifferentiable functions spaces. A se- is said to satisfy the following conditions: quence of positive numbers (Mp)p∈Z+ (H1) Logarithmic convexity, if (H2) Stability under ultradifferential operators, if there are constants A > 0 M 2 p ≤ Mp−1Mp+1, ∀p ≥ 1. and H > 0 such that (H3)′ Non-quasi-analyticity, if Mp ≤ AH pMqMp−q,∀p ≥ q. Mp−1 Mp < +∞. ∞Xp=1 Remark 2. Some results remain valid, see [11], when (H2) is replaced by the following weaker condition : (H2)′ Stability under differential operators, if there are constants A > 0 and H > 0 such that Mp+1 ≤ AH pMp,∀p ∈ Z+. The associated function of the sequence (Mp)p∈Z+ is the function defined by ln tp Mp , t ∈ R∗ +. fM (t) = sup p Some needed results of the associated function are given in the following propo- sitions proved in [11]. ULTRAREGULAR GENERALIZED FUNCTIONS 5 Proposition 2. A positive sequence (Mp)p∈Z+ only if satisfies condition (H1) if and and satisfies condition (H2) if and and Mp = M0 sup t>0 p = 0, 1, ... tp exp(cid:16)fM (t)(cid:17), Proposition 3. A positive sequence (Mp)p∈Z+ only if , ∃A > 0,∃H > 0,∀t > 0, (1) Remark 3. We will always suppose that the sequence (Mp)p∈Z+ dition (H1) and M0 = 1. 2fM (t) ≤ fM (Ht) + ln (AM0) . A differential operator of infinite order P (D) = Pγ∈Zn + ferential operator of class M = (Mp)p∈Z+ c > 0 such that ∀γ ∈ Zn +, (2) aγ ≤ c hγ Mγ . satisfies the con- aγDγ is called an ultradif- , if for every h > 0 there exist a constant The class of ultradifferentiable functions of class M, denoted by E M (Ω) , is the space of all f ∈ C ∞ (Ω) satisfying for every compact subset K of Ω, ∃c > 0,∀α ∈ Zn +, (3) x∈K ∂αf (x) ≤ cα+1Mα. sup This space is also called the space of Denjoy-Carleman. = (p!σ)p∈Z+ , σ > 1, we obtain E σ (Ω) the Gevrey space Example 5. If (Mp)p∈Z+ of order σ, and A (Ω) := E 1 (Ω) is the space of real analytic functions defined on the open set Ω. The basic properties of the space E M (Ω) are summarized in the following proposition, for the proof see [11]. Proposition 4. The space E M (Ω) is an algebra. Moreover, if (Mp)p∈Z+ sat- isfies (H2)′ , then E M (Ω) is stable by any differential operator of finite order with coefficients in E M (Ω) and if (Mp)p∈Z+ satisfies (H2) then any ultradiffer- ential operator of class M operates also as a sheaf homomorphism. The space DM (Ω) = E M (Ω) ∩ D (Ω) is well defined and is not trivial if and only if the sequence (Mp)p∈Z+ Remark 4. The strong dual of DM (Ω), denoted D′M (Ω) , is called the space of Roumieu ultraditributions. satisfies (H3)′ . 6 KHALED BENMERIEM AND CHIKH BOUZAR 4. Ultraregular generalized functions In the same way as G∞ (Ω) , GR (Ω) forms a sheaf of differential subalgebras of G (Ω), consequently one defines the generalized R−singular support of u ∈ G (Ω), denoted by singsuppR u, as the complement in Ω of the largest set Ω′ such that u/Ω′ ∈ GR (Ω′) , where u/Ω′ means the restriction of the generalized function u on Ω′. This new notion of regularity is linked with the asymptotic limited growth of generalized functions. Our aim in this section is to introduce a more precise notion of regularity within the Colombeau algebra taking into account both the asymptotic growth and the smoothness property of generalized functions. We introduce general algebras of ultradifferentiable R−regular generalized functions of class M, where the sequence M = (Mp)p∈Z+ satisfies the conditions (H1) with M0 = 1, (H2) and (H3)′ . Definition 4. The space of ultraregular moderate elements of class M, denoted X M,R (Ω) , is the space of (fε)ε ∈ C ∞ (Ω)]0,1] satisfying for every compact K of Ω, ∃N ∈ R,∃C > 0,∃ε0 ∈ ]0, 1] ,∀α ∈ Zn (4) +,∀ε ≤ ε0, x∈K ∂αfε (x) ≤ C α+1Mαε−Nα. sup The space of null elements is defined as N M,R (Ω) := N (Ω) ∩ X M,R (Ω) . The main properties of the spaces X M,R (Ω) and N M,R (Ω) are given in the following proposition. Proposition 5. 1) The space X M,R (Ω) is a subalgebra of X (Ω) stable by action of differential operators 2) The space N M,R (Ω) is an ideal of X M,R (Ω) . Proof. 1) Let (fε)ε , (gε)ε ∈ X M,R (Ω) and K a compact subset of Ω, then ∃N ∈ R,∃C1 > 0,∃ε1 ∈ ]0, 1] , such that ∀β ∈ Zn +,∀x ∈ K,∀ε ≤ ε1, (5) ∃N ′ ∈ R,∃C2 > 0,∃ε2 ∈ ]0, 1] , such that ∀β ∈ Zn 2 Mβε−N ′ +,∀x ∈ K,∀ε ≤ ε2, β. (6) It clear from (R2) that (fε + gε)ε ∈ X M,R (Ω) . Let α ∈ Zn 1 Mβε−Nβ, (cid:12)(cid:12)∂βfε (x)(cid:12)(cid:12) ≤ C β+1 (cid:12)(cid:12)∂βgε (x)(cid:12)(cid:12) ≤ C β+1 β(cid:19)(cid:12)(cid:12)∂α−βfε (x)(cid:12)(cid:12)(cid:12)(cid:12)∂βgε (x)(cid:12)(cid:12) . αXβ=0(cid:18)α +, then ∂α (fεgε) (x) ≤ ULTRAREGULAR GENERALIZED FUNCTIONS 7 From (R3) ∃N" ∈ R such that, ∀β ≤ α, Nα−β + N ′ we have MpMq ≤ Mp+q, then for ε ≤ min{ε1, ε2} and x ∈ K, we have β ≤ N"α, and from (H1), εN "α Mα ∂α (fεgε) (x) ≤ β εN ′ β(cid:19) εNα−β αXβ=0(cid:18)α Mα−β(cid:12)(cid:12)∂α−βfε (x)(cid:12)(cid:12) × Mβ(cid:12)(cid:12)∂βgε (x)(cid:12)(cid:12) β(cid:19)C α−β+1 αXβ=0(cid:18)α C β+1 × 1 2 ≤ ≤ C α+1, where C = max{C1C2, C1 + C2}, then (fεgε)ε ∈ X M,R (Ω). Let now α, β ∈ Zn +, where β = 1, then for ε ≤ ε1 and x ∈ K, we have From (R1), ∃N ′ ∈ R, such that Nα+1 ≤ N ′ such that Mα+1 ≤ AH αMα, we have (cid:12)(cid:12)∂α(cid:0)∂βfε(cid:1) (x)(cid:12)(cid:12) ≤ C α+2 (cid:12)(cid:12)∂α(cid:0)∂βfε(cid:1) (x)(cid:12)(cid:12) ≤ AC 2 1 (C1H)α Mαε−N ′ α ≤ C α+1Mαε−N ′ α, 1 Mα+1ε−Nα+1. α, and from (H2)′ ,∃A > 0, H > 0, which means(cid:0)∂βfε(cid:1)ε ∈ X M,R (Ω) . 2) The facts that N M,R (Ω) = N (Ω) ∩ X M,R (Ω) ⊂ X M,R (Ω) and N (Ω) = N R (Ω) is an ideal of X R (Ω) give that N M,R (Ω) is an ideal of X M,R (Ω) . (cid:3) The following definition introduces the algebra of ultraregular generalized func- tions. Definition 5. The algebra of ultraregular generalized functions of class M = (Mp)p∈Z+ (7) , denoted GM,R (Ω) , is the quotient algebra GM,R (Ω) = X M,R (Ω) N M,R (Ω) . The basic properties of GM,R (Ω) are given in the following assertion. Proposition 6. GM,R (Ω) is a differential subalgebra of G (Ω) . Proof. The algebraic properties hold from proposition 5. Example 6. If we take the set R = B we obtain as a particular case the algebra GM,B (Ω) of [12] denoted there by GL (Ω) . (cid:3) 8 KHALED BENMERIEM AND CHIKH BOUZAR = (p!σ)p∈Z+ we obtain a new subalgebra Gσ,R (Ω) Example 7. If we take (Mp)p∈Z+ of G (Ω) called the algebra of Gevrey regular generalized functions of order σ. = (p!σ)p∈Z+ we obtain Example 8. If we take both the set R = B and (Mp)p∈Z+ a new algebra, denoted Gσ,∞ (Ω) , that we will call the Gevrey-Oberguggenberger algebra of order σ. Remark 5. In [3] is introduced an algebra of generalized Gevrey ultradistributions containing the classical Gevrey space E σ (Ω) as a subalgebra and the space of Gevrey ultradistributions D′ 3σ−1 (Ω) as a subspace. It is not evident how to obtain, without more conditions, that X M,R (Ω) is stable by action of ultradifferential operators of class M, however we have the following result. Proposition 7. Suppose that the regular set R satisfies as well the following condition : For all (Nk)k∈Z+ ∈ R , there exist an (N ∗ k )k∈Z+ ∈ R, and positive numbers h > 0, L > 0,∀m ∈ Z+,∀ε ∈ ]0, 1] , (8) hkε−Nk+m ≤ Lε−N ∗ m. Xk∈Z+ Then the algebra X M,R (Ω) is stable by action of ultradifferential operators of class M. Proof. Let (fε)ε ∈ X M,R (Ω) and P (D) = Pβ∈Zn + erator of class M, then for any compact set K of Ω, ∃ (Nm)m∈Z+ ∈ R,∃C > 0,∃ε0 ∈ ]0, 1] , such that ∀α ∈ Zn +,∀x ∈ K,∀ε ≤ ε1, aβDβ be an ultradifferential op- ∂αfε (x) ≤ C α+1Mαε−Nα. For every h > 0 there exists a c > 0 such that ∀β ∈ Zn +, aβ ≤ c hβ Mβ . Let α ∈ Zn +, then hα Mα ∂α (P (D) fε) (x) ≤ Xβ∈Zn ≤ c Xβ∈Zn + + aβ hα Mα(cid:12)(cid:12)∂α+βfε (x)(cid:12)(cid:12) Mα(cid:12)(cid:12)∂α+βfε (x)(cid:12)(cid:12) . hα hβ Mβ ULTRAREGULAR GENERALIZED FUNCTIONS 9 From (H2) and (fε)ε ∈ X M,R (Ω) , then ∃C > 0,∃A > 0,∃H > 0, hβε−Nα+β, Mα ∂α (P (D) fε) (x) ≤ Aα Xβ∈Zn hα + consequently by condition (8), there exist (N ∗ k )k∈Z+ ∈ R, h > 0, L > 0,∀ε ∈ ]0, 1] , hα Mα ∂α (P (D) fε) (x) ≤ AαLε−N ∗ α which shows that (P (D) fε)ε ∈ X M,R (Ω) . Example 9. The sets {0} and B satisfy the condition (8). The space E M (Ω) is embedded into GM,R (Ω) for all R by the canonical map (cid:3) σ : E M (Ω) → GM,R (Ω) , u → [uε] where uε = u for all ε ∈ ]0, 1] , which is an injective homomorphism of algebras. Proposition 8. The following diagram E M (Ω) → C ∞ (Ω) → D′ (Ω) ↓ GM,B (Ω) → GB (Ω) → G (Ω) ↓ ↓ is commutative. Proof. The embeddings in the diagram are canonical except the embedding D′ (Ω) → G (Ω) , which is now well known in framework of Colombeau generalized functions, see [7] for details. The commutativity of the diagram is then obtained easily from the commutativity of the classical diagram C ∞ (Ω) → D′ (Ω) G (Ω) ց ↓ (cid:3) A fundamental result of regularity in G (Ω) is the following. Theorem 9. We have GM,B (Ω) ∩ D′ (Ω) = E M (Ω) . Proof. Let u = cl (uε)ε ∈ GM,B (Ω)∩ C ∞ (Ω) , i.e. (uε)ε ∈ X M,B (Ω), then we have for every compact set K ⊂ Ω,∃N ∈ Z+,∃c > 0,∃η ∈ ]0, 1] , ∀α ∈ Zn +,∀ε ∈ ]0, η] : sup x∈K ∂αu (x) ≤ cα+1Mαε−N . When choosing ε = η, we obtain ∀α ∈ Zn +, sup x∈K ∂αu (x) ≤ cα+1Mαη−N ≤ cα+1 1 Mα, 10 KHALED BENMERIEM AND CHIKH BOUZAR where c1 depends only on K. Then u is in E M (Ω). This shows that GM,B (Ω) ∩ C ∞ (Ω) ⊂ E M (Ω). As the reverse inclusion is obvious, then we have proved GM,B (Ω) ∩ C ∞ (Ω) = E M (Ω). Consequently GM,B (Ω) ∩ D′ (Ω) = (cid:0)GM,B (Ω) ∩ GB (Ω)(cid:1) ∩ D′ (Ω) = GM,B (Ω) ∩(cid:0)GB (Ω) ∩ D′ (Ω)(cid:1) = GM,B (Ω) ∩ C ∞ (Ω) = E M (Ω) (cid:3) Proposition 10. The algebra GM,R (Ω) is a sheaf of subalgebras of G (Ω) . Proof. The sheaf property of GM,R (Ω) is obtained in the same way as the sheaf properties of GR (Ω) and E M (Ω). (cid:3) We can now give a new tool of GM,R-local regularity analysis. Definition 6. Define the (M,R)-singular support of a generalized function u ∈ G (Ω) , denoted by singsuppM,R (u) , as the complement of the largest open set Ω′ such that u ∈ GM,R (Ω′) . The basic property of singsuppM,R is summarized in the following proposition, which is easy to prove by the facts above. Proposition 11. Let P (x, D) = Pα≤m differential operator with GM,R (Ω) coefficients, then (9) singsuppM,R (P (x, D) u) ⊂ singsuppM,R (u) ,∀u ∈ G (Ω) aα (x) Dα be a generalized linear partial We can now introduce a local generalized analysis in the sense of Colombeau algebra. Indeed, a generalized linear partial differential operator with GM,R (Ω) coefficients P (x, D) is said (M,R)−hypoelliptic in Ω, if (10) singsuppM,R (P (x, D) u) = singsuppM,R (u) ,∀u ∈ G (Ω) Such a problem in this general form is still in the beginning. Of course, a mi- crolocalization of the problem (10) will lead to a more precise information about solutions of generalized linear partial differential equations. A first attempt is done in the following section. 5. Affine ultraregular generalized functions Although we have defined a tool for a local (M,R)−analysis in G (Ω), it is not clear how to microlocalize this concept in general. We can do it in the general situation of affine ultraregularity. This is the aim of this section. ULTRAREGULAR GENERALIZED FUNCTIONS 11 Definition 7. Define the affine regular sequences A =n(Nm)m∈Z+ ∈ R : ∃a ≥ 0,∃b ≥ 0, Nm ≤ am + b,∀m ∈ Z+o . A basic (M,A)−microlocal analysis in G (Ω) can be developed due to the fol- lowing result. Proposition 12. Let f = cl (fε)ε ∈ GC (Ω) , then f is A−ultraregular of class if and only if ∃a ≥ 0,∃b ≥ 0,∃C > 0,∃k > 0, ∃ε0 ∈ ]0, 1] ,∀ε ≤ ε0, M = (Mp)p∈Z+ such that (11) F (fε) (ξ) ≤ Cε−b exp(cid:16)−fM (kεa ξ)(cid:17) ,∀ξ ∈ Rn, where F denotes the Fourier transform. Proof. Suppose that f = cl (fε)ε ∈ GC (Ω) ∩ GM,A (Ω) , then ∃K compact of Ω,∃C > 0,∃N ∈ A,∃ε1 > 0,∀α ∈ Zn +,∀x ∈ K,∀ε ≤ ε0, suppfε ⊂ K, such that (12) so we have, ∀α ∈ Zn +, ∂αfε ≤ C α+1Mαε−Nα, ξαF (fε) (ξ) ≤(cid:12)(cid:12)(cid:12)(cid:12)Z exp (−ixξ) ∂αfε (x) dx(cid:12)(cid:12)(cid:12)(cid:12) F (fε) (ξ) ≤ C inf ≤ Cε−b inf ξα F (fε) (ξ) ≤ C α+1Mαε−Nα. α (C αMα ε−Nα) ξα α ((cid:18)ε−aC ξ (cid:19)α Mα) ≤ Cε−b exp(cid:18)−fM(cid:18) εa ξ√nC(cid:19)(cid:19) . F (fε) (ξ) ≤ Cε−b exp(cid:16)−fM (kεa ξ)(cid:17) , then, ∃C > 0,∀ε ≤ ε0, Therefore Hence ∃C > 0,∃k > 0, i.e. we have (11). 12 KHALED BENMERIEM AND CHIKH BOUZAR Suppose now that (11) is valid, then from inequality (1) of the Proposition 3, ∃C, C ′, C ′′ > 0,∃ε0 ∈ ]0, 1] ,∀ε ≤ ε0, ∂αfε (x) ≤ Cε−b sup ξ α H H k H εa ξ(cid:19)(cid:19) × ξα exp(cid:18)−fM(cid:18) k ×Z exp(cid:18)−fM(cid:18) k εa ξ(cid:19)(cid:19) dξ εaξ(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) exp(cid:18)−fM(cid:18) k ηα exp(cid:16)−fM (η)(cid:17) . k(cid:1), and Nm = am + b, then f ∈ GM,A (Ω) . ≤ C ′ε−aα−b sup ≤ C ′′ε−aα−b sup ∂αfε (x) ≤ C α+1Mαε−Nα, H η ξ εa ξ(cid:19)(cid:19) (cid:3) Due to the Proposition 2, then ∃C > 0,∃N ∈ A, such that where C = max(cid:0)C, 1 1 Corollary 13. Let f = cl (fε)ε ∈ GC (Ω) , then f is a Gevrey affine ultraregular generalized function of order σ, i.e. f ∈ Gσ,A (Ω), if and only if ∃a ≥ 0,∃b ≥ 0,∃C > 0,∃k > 0, ∃ε0 > 0,∀ε ≤ ε0, such that (13) In particular, f is a Gevrey generalized function of order σ, i.e. f ∈ Gσ,∞ (Ω), if and only if ∃b ≥ 0,∃C > 0,∃k > 0, ∃ε0 > 0,∀ε ≤ ε0, such that (14) F (fε) (ξ) ≤ Cε−b exp(cid:16)−kεa ξ F (fε) (ξ) ≤ Cε−b exp(cid:16)−k ξ σ(cid:17) ,∀ξ ∈ Rn. σ(cid:17) ,∀ξ ∈ Rn. Using the above results, we can define the concept of GM,A−wave front of u ∈ G (Ω) and give the basic elements of a (M,A)−generalized microlocal analysis within the Colombeau algebra G (Ω). Definition 8. Define the cone PM A (f ) ⊂ Rn\{0} , f ∈ GC (Ω), as the comple- ment of the set of points having a conic neighborhood Γ such that ∃a ≥ 0,∃b ≥ 0,∃C > 0,∃k > 0, ∃ε0 > 0,∀ε ≤ ε0, such that (15) F (fε) (ξ) ≤ Cε−b exp(cid:16)−fM (kεa ξ)(cid:17) ,∀ξ ∈ Rn. Proposition 14. For every f ∈ GC (Ω) , we have 1 1. The setPM 2. PM A (f ) is a closed subset. A (f ) = ∅ ⇐⇒ f ∈ GM,A (Ω) . Proof. The proof of 1. is clear from the definition, and 2. holds from Proposition 12. (cid:3) ULTRAREGULAR GENERALIZED FUNCTIONS 13 Proposition 15. For every f ∈ GC (Ω) , we have (ψf ) ⊂ MXA Proof. Let ξ0 /∈PM A (f ), i.e. ∃Γ a conic neighborhood of ξ0,∃a ≥ 0,∃b > 0,∃k1 > 0,∃c1 > 0,∃ε1 ∈ ]0, 1] , such that ∀ξ ∈ Γ,∀ε ≤ ε1, (16) Let χ ∈ DM (Ω) , χ = 1 on neighborhood of suppf , then χψ ∈ DM (Ω) ,∀ψ ∈ E M (Ω), hence, see [11], ∃k2 > 0,∃c2 > 0,∀ξ ∈ Rn, (17) (f ) ,∀ψ ∈ E M (Ω) . MXA F (fε) (ξ) ≤ c1ε−b exp(cid:16)−fM (k1εa ξ)(cid:17) , F (χψ) (ξ) ≤ c2 exp(cid:16)−fM (k2 ξ)(cid:17) . F (χψfε) (ξ) = ZRn F (fε) (η)F (χψ) (ξ − η) dη = ZA F (fε) (η)F (χψ) (ξ − η) dη + +ZB F (fε) (η)F (χψ) (ξ − η) dη, Let Λ be a conic neighborhood of ξ0 such that, Λ ⊂ Γ, then we have, for ξ ∈ Λ, ×ZA so ∃c > 0,∃k > 0, (18) where A = {η;ξ − η ≤ δ (ξ + η)} and B = {η;ξ − η > δ (ξ + η)}. Take δ sufficiently small such that η ∈ Γ, ξ 2 < η < 2ξ ,∀ ∈ Λ,∀η ∈ A, then ∃c > 0,∀ε ≤ ε1, (cid:12)(cid:12)(cid:12)(cid:12)ZA F (fε) (η)F (χψ) (ξ − η) dη(cid:12)(cid:12)(cid:12)(cid:12) ≤ cε−b exp(cid:18)−fM(cid:18)k1εaξ 2(cid:19)(cid:19) × exp(cid:16)−fM (k2 ξ − η)(cid:17) dη, (cid:12)(cid:12)(cid:12)(cid:12)ZA F (fε) (η)F (χψ) (ξ − η) dη(cid:12)(cid:12)(cid:12)(cid:12) ≤ cε−b exp(cid:16)−fM (kεa ξ)(cid:17) . (cid:12)(cid:12)(cid:12)(cid:12)ZB F (fε) (η)F (χψ) (ξ − η) dη(cid:12)(cid:12)(cid:12)(cid:12) ≤ cε−qZB ηm exp(cid:16)−fM (k2 ξ − η)(cid:17) dη ≤ cε−qZB ηm exp(cid:16)−fM (k2δ (ξ + η))(cid:17) dη. F (fε) (ξ) ≤ cε−q ξm . Hence for ε ≤ min (ε1, ε2) , ∃c > 0, such that we have As f ∈ GC (Ω) , then ∃q ∈ Z+,∃m > 0,∃c > 0,∃ε2 > 0,∀ε ≤ ε2, 14 KHALED BENMERIEM AND CHIKH BOUZAR We have, from Proposition 3, i.e. inequality (1), ∃H > 0,∃A > 0,∀t1 > 0,∀t2 > 0, so (19) −fM (t1 + t2) ≤ −fM(cid:18) t1 H(cid:19) −fM(cid:18) t2 H(cid:19) + ln A , (cid:12)(cid:12)(cid:12)(cid:12)ZB F (fε) (η)F (χψ) (ξ − η) dη(cid:12)(cid:12)(cid:12)(cid:12) ≤ cAε−q exp(cid:18)−fM(cid:18)k2δ H ξ(cid:19)(cid:19) × ×ZB ηm exp(cid:18)−fM(cid:18) k2δ H η(cid:19)(cid:19) dη. (cid:12)(cid:12)(cid:12)(cid:12)ZB bfε (η)bψ (ξ − η) dη(cid:12)(cid:12)(cid:12)(cid:12) ≤ cε−q exp(cid:16)−fM (kεa ξ)(cid:17) . Hence ∃c > 0,∃k > 0, such that (20) A (ψf ) . A (f ) of a generalized function f at a point x0 and the affine wave Consequently, (18) and (20) give ξ0 /∈PM We definePM of class M = (Mp) of f at x0, denoted byPM XM (f ) =\nXM front set of class M in G (Ω) . Definition 9. Let f ∈ G (Ω) and x0 ∈ Ω, the cone of affine singular directions (21) A,x0 (f ), is (φf ) : φ ∈ DM (Ω) , φ = 1 on a neighborhood of x0o . A,x0 (cid:3) A XM A,x0 (f ) = ∅ ⇐⇒ x0 /∈ singsuppM,A (f ) . Proof. See the proof of the analogical Lemma in [3]. A (f ) ⊂ Ω × Rn\{0} , if there exist Definition 10. A point (x0, ξ0) /∈ W F M φ ∈ DM (Ω) , φ ≡ 1 on a neighborhood of x0, a conic neighborhood Γ of ξ0, and numbers a ≥ 0, b ≥ 0, k > 0, c > 0, ε0 ∈ ]0, 1] , such that ∀ε ≤ ε0,∀ξ ∈ Γ, (cid:3) F (φfε) (ξ) ≤ cε−b exp(cid:16)−fM (kεa ξ)(cid:17) . A (f ) ⊂ Ω × Rn\{0} means ξ0 /∈PM A are given in the following proposition. Remark 6. A point (x0, ξ0) /∈ W F M The basic properties of W F M A,x0 (f ) . The following lemma gives the relation between the local and microlocal (M,A)−analysis in G (Ω). Lemma 16. Let f ∈ G (Ω), then ULTRAREGULAR GENERALIZED FUNCTIONS 15 Proposition 17. Let f ∈ G (Ω), then 1) The projection of W F M 2) If f ∈ GC (Ω) , then the projection of W F M 3) W F M 4) W F M A (f ) ,∀α ∈ Zn +. A (f ) ,∀g ∈ GM,A (Ω) . A (∂αf ) ⊂ W F M A (gf ) ⊂ W F M A (f ) on Ω is the singsuppM,A (f ) . A (f ) on Rn\{0} isPM A (f ) . Proof. 1) and 2) holds from the definition, Proposition 14 and Lemma 16. 3) Let (x0, ξ0) /∈ W F M A (f ). Then ∃φ ∈ DM (Ω) , φ ≡ 1 on U , where U is a neighborhood of x0, there exist a conic neighborhood Γ of ξ0, and ∃a ≥ 0,∃b > 0,∃k2 > 0,∃c1 > 0,∃ε0 ∈ ]0, 1] , such that ∀ξ ∈ Γ,∀ε ≤ ε0, (22) F (φfε) (ξ) ≤ c1ε−b exp(cid:16)−fM (k2εa ξ)(cid:17) . We have, for ψ ∈ DM (U) such that ψ (x0) = 1, F (ψ∂fε) (ξ) = F (∂ (ψfε)) (ξ) − F ((∂ψ) fε) (ξ) ≤ ξF (ψφfε) (ξ) + F ((∂ψ) φfε) (ξ) . As W F M Then A (ψf ) ⊂ W F M A (f ) , so (22) holds for both F (ψφfε) (ξ) and F ((∂ψ) φfε) (ξ) . with some c′ > 0, k3 > 0, (k3 < k2), such that ξF (ψφfε) (ξ) ≤ cε−b ξ exp(cid:16)−fM (k2εa ξ)(cid:17) ≤ c′ε−b−a exp(cid:16)−fM (k3εa ξ)(cid:17) , εa ξ ≤ c′ exp(cid:16)fM (k2εa ξ) −fM (k3εa ξ)(cid:17) A (∂f ) . 4) Let (x0, ξ0) /∈ W F M for ε sufficiently small. Hence (22) holds for F (ψ∂fε) (ξ) , which proves (x0, ξ0) /∈ W F M A (f ). Then there exist φ ∈ DM (Ω) , φ (x) = 1 on a neighborhood U of x0, a conic neighborhood Γ of ξ0, and numbers a1 ≥ 0, b1 > 0, k1 > 0, c1 > 0, ε1 ∈ ]0, 1] , such that ∀ε ≤ ε1,∀ξ ∈ Γ, F (φfε) (ξ) ≤ c1ε−b1 exp(cid:16)−fM (k1εa1 ξ)(cid:17) . Let ψ ∈ DM (Ω) and ψ = 1 on suppφ, then F (φgεfε) = F (ψgε) ∗ F (φfε) . We have ψg ∈ GM,A (Ω) , then ∃c2 > 0, ∃a2 ≥ 0,∃b2 > 0,∃k2 > 0,∃ε2 > 0,∀ξ ∈ Rn,∀ε ≤ ε2, (23) F (ψgε) (ξ) ≤ c2ε−b2 exp(cid:16)−fM (k2εa2 ξ)(cid:17) . 16 KHALED BENMERIEM AND CHIKH BOUZAR We have F (φgεfε) (ξ) = ZA F (φfε) (η)F (ψgε) (ξ − η) dη + +ZB F (φfε) (η)F (ψgε) (ξ − η) dη, where A and B are the same as in the proof of Proposition 15. By the same reasoning we obtain the proof. (cid:3) A micolocalization of Proposition 11 is expressed in the following result. Corollary 18. Let P (x, D) = Pα≤m ferential operator with GM,A (Ω) coefficients, then W F M A (P (x, D) f ) ⊂ W F M A (f ) ,∀f ∈ G (Ω) aα (x) Dα be a generalized linear partial dif- Remark 7. The reverse inclusion will give a generalized microlocal ultraregularity of linear partial differential operator with coefficients in GM,A (Ω). The first case of G∞-microlocal hypoellipticity has been studied in [10]. A general interesting (M,A)-generalized microlocal elliptic ultraregularity is to prove the problem of following inclusion W F M A (P (x, D) f ) ∪ Char(P ),∀f ∈ G (Ω) , A (f ) ⊂ W F M where P (x, D) is a generalized linear partial differential operator with GM,A (Ω) coefficients and Char(P ) is the set of generalized characteristic points of P (x, D). 6. Generalized Hormander's theorem A (f ) + W F M We recall the following fundamental lemma, see [9] for the proof. A (g) =(cid:8)(x, ξ + η) : (x, ξ) ∈ W F M We extend the generalized Hormander's result on the wave front set of the product, the proof follow the same steps as the proof of theorem 26 in [3]. Let f, g ∈ G (Ω) , we define (24) W F M A (g)(cid:9) , Lemma 19. LetP1,P2 be closed cones in Rm\{0} , such that 0 /∈P1 +P2 , i)P1 +P2 ii) For any open conic neighborhood Γ of P1 +P2 in Rm\{0} , one can find open conic neighborhoods of Γ1, Γ2 in Rm\{0} of, respectively, P1,P2 , such = (P1 +P2) ∪P1 ∪P2 A (f ) , (x, η) ∈ W F M Rm/{0} then that Γ1 + Γ2 ⊂ Γ The principal result of this section is the following theorem. ULTRAREGULAR GENERALIZED FUNCTIONS 17 Theorem 20. Let f, g ∈ G (Ω) , such that ∀x ∈ Ω, A (f ) + W F M (25) (x, 0) /∈ W F M A (g) , then A (g) A (f )∪ W F M A (f ) ∪ W F M A (φf ) ∪PM (26) W F M A (f ) + W F M A (f ) + W F M A (φf ) +PM A (f g) ⊆(cid:0)W F M A (g)(cid:1) ∪ W F M Proof. Let (x0, ξ0) /∈(cid:0)W F M A (g)(cid:1)∪ W F M A (φg)(cid:17) ∪PM DM (Ω), φ (x0) = 1, ξ0 /∈(cid:16)PM (25) we have 0 /∈PM A (φf ) +PM (φg)(cid:17)∪XM ξ0 /∈(cid:16)XM (φf )∪XM (φf ) +XM Let Γ0 be an open conic neighborhood ofPM such that XM (φf ) ⊂ Γ1, XM (φg) =XM A (φf ) +PM (φg) ⊂ Γ2 and Γ1 + Γ2 ⊂ Γ0 A A (φg) in Rm\{0} such that ξ0 /∈ Γ0 then, from lemma 19 ii), there exist open cones Γ1 and Γ2 in Rm\{0} (φf ) +XM A A A (g) , then ∃φ ∈ A (φg) . From A (φg) then by lemma 19 i), we have A A A A A Rm\{0} (φg) Define Γ = Rm\Γ0, so (27) Let ξ ∈ Γ and ε ∈ ]0, 1] F (φfεφgε) (ξ) = (F (φfε) ∗ F (φgε)) (ξ) Γ ∩ Γ2 = ∅ and (Γ − Γ2) ∩ Γ1 = ∅ = ZΓ2 F (φfε) (ξ − η)F (φgε) (η) dη +ZΓc 2 = I1 (ξ) + I2 (ξ) F (φfε) (ξ − η)F (φgε) (η) dη From (27), ∃a1 ≥ 0, b1 ≥ 0, k1 > 0, c1 > 0, ε1 ∈ ]0, 1] , such that ∀ε ≤ ε1,∀ξ ∈ Γ2, we can show easily by the fact that (φgε) ∈ GC (Ω) that ∀a2 ≥ 0,∀k2 > 0,∃b2 ≥ 0,∃c2 > 0, , ∃ε2 ∈ ]0, 1] , such that ∀ε ≤ ε2, F (φfε) (ξ − η) ≤ c1ε−b1 exp −fM (k1εa1 ξ) F (φgε) (η) ≤ c2ε−b2 expfM (k2εa2 η) ,∀η ∈ Rn, Let γ > 0 sufficiently small such that ξ − η ≥ γ (ξ + η) ,∀η ∈ Γ2. Hence for ε ≤ min (ε1, ε2) , I1 (ξ) ≤ c1c2ε−b1−b2ZΓ2 exp(cid:16)−fM (k1εa1 ξ − η) +fM (k2εa2 η)(cid:17) dη from proposition 3, ∃H > 0,∃A > 0,∀t1 > 0,∀t2 > 0, (28) −fM (t1 + t2) ≤ −fM(cid:18) t1 H(cid:19) −fM(cid:18) t2 H(cid:19) + ln A , 18 then KHALED BENMERIEM AND CHIKH BOUZAR H H H εa1γ ξ(cid:19)(cid:19) εa1γ η(cid:19) +fM (k2εa2 η)(cid:19) dη εa1γ ξ(cid:19)(cid:19) H 2 εa1γ − k2εa2(cid:19)η(cid:19)(cid:19) dη I1 (ξ) ≤ c1c2ε−b1−b2 exp(cid:18)−fM(cid:18)k1 exp(cid:18)−fM(cid:18)k1 ×ZΓ2 ≤ c1c2ε−b1−b2 exp(cid:18)−fM(cid:18)k1 exp(cid:18)−fM(cid:18)(cid:18) k1 ×ZΓ2 I1 (ξ) ≤ cε−b exp(cid:16)−fM (kεa1 ξ)(cid:17) 2∩{η≤rξ} F (φfε) (ξ − η)F (φgε) (η) dη +ZΓc H and k1 take k = γk1 c1c2 Let r > 0, I2 (ξ) = ZΓc = I21 (ξ) + I22 (ξ) H 2 εa1γ − k2εa2 > 0, then ∃b = b (b1 + b2, a1, a2, k1, k2, H) ,∃c = 2∩{η≥rξ} F (φfε) (ξ − η)F (φgε) (η) dη ≤ c′ε−b′ Choose r sufficiently small such that {η ≤ r ξ} =⇒ ξ − η /∈ Γ1. Then ξ − η ≥ (1 − r)ξ ≥ (1 − 2r)ξ +η , consequently ∃c > 0,∃a1, a2, b1, k1, k2 > 0,∃ε1 > 0 such that ∀ε ≤ ε1, I21 (ξ) ≤ cε−bZΓ2 exp(cid:16)−fM (k1εa1 ξ − η) −fM (k2εa2 η)(cid:17) ≤ cε−b exp(cid:16)−fM (k′ exp(cid:16)−fM (k′ 1εa1 ξ)(cid:17)Z exp(cid:16)−fM (k1εa1 η) −fM (k2εa2 η)(cid:17) dη 1εa1 ξ)(cid:17) If η ≥ r ξ , we have η ≥ η + r ξ 0,∃b2 > 0,∃ε2 > 0 such that ∀ε ≤ ε2 I21 (ξ) ≤ cε−b1−b2ZΓ2 ≤ cε−b1−b2ZΓ2 ≤ cε−b1−b2 exp(cid:18)−fM(cid:18) k1r exp(cid:16)fM (k2εa2 ξ − η) −fM (k1εa1 η)(cid:17) dη exp(cid:18)fM (k2εa2 ξ − η) −fM(cid:18) k1 , and then ∃c > 0,∃a1, b1, k1 > 0,∀a2, k2 > εa1 η + εa1 ξ(cid:19)(cid:19) dη exp(cid:18)fM (k2εa2 ξ − η) −fM(cid:18) k1 εa1 ξ(cid:19)(cid:19)ZΓ2 k1r 2 2H 2 2 εa1 η(cid:19)(cid:19) dη 2H ULTRAREGULAR GENERALIZED FUNCTIONS 19 sufficiently smalls, we obtain ∃a, b, c > 0,∃ε3 > 0, such that if we take k2, 1 a2 ∀ε ≤ ε3 I21 (ξ) ≤ cε−b exp(cid:16)−fM (kεa ξ)(cid:17) , which finishes the proof. (cid:3) References [1] ANTONEVICH A. B., RADYNO Ya. V., On general method of constructing algebras of new generalized functions. Soviet. Math. Dokl., vol. 43:3, p. 680-684, 1991. [2] BENMERIEM K., BOUZAR C., Affine ultraregular Generalized function. Banach Center Publications, vol. 88, p. 39-53, 2010. [3] BENMERIEM K., BOUZAR C., Generalized Gevrey ultradistributions. New York J. Math. 15, p. 37-72. 2009. [4] BENMERIEM K., BOUZAR C., Ultraregular Generalized function of Colombeau type, J. Math. Sci. Univ. Tokyo, Vol. 15, No. 4, p. 427 -- 447, 2008. [5] COLOMBEAU J. F., New generalized functions and multiplication of distributions. North- Holland, 1984. [6] A. DELCROIX. Regular rapidly decreasing nonlinear generalized functions. Application to microlocal regularity, J. Math. Anal. Appl., 327, p. 564-584, 2007. [7] GROSSER M., KUNZINGER M., OBERGUGGENBERGER M., STEINBAUER R., Geo- metric theory of generalized functions. Kluwer, 2001. [8] H ORMANDER L., The analysis of linear partial differential operators, Vol. 1 : Distribution theory and Fourier analysis. Springer, 2nd edition 1990. [9] H ORMANN G., KUNZINGER M., Microlocal properties of basic operations in Colombeau algebras. J. Math. Anal. Appl., 26:1, p. 254-270, 2001. [10] H ORMANN G., OBERGUGGENBERGER M., PILIPOVIC S., Microlocal hypoelliptic- ity of linear differential operators with generalized functions as coefficients. Trans. Amer. Math. Soc., 358, p. 3363-3383, 2006. [11] KOMATSU H., Ultradistributions I, J. Fac. Sci. Univ. Tokyo, Sect. IA 20, p. 25-105, 1973. [12] MARTI J. A., Regularity, local and microlocal analysis in theories of generalized functions. Acta Appl Math, 105, p. 267 -- 302, 2009. [13] NEDELJKOV M., PILIPOVIC S., SCARPALEZOS D.. The linear theory of Colombeau generalized functions, Longman, 1998. [14] OBERGUGGENBERGER M., Multiplication of distributions and applications to partial differential equations, Longman, 1992. [15] OBERGUGGENBERGER M., Regularity theory in Colombeau algebras. Bull. T. CXXXIII. Acad. Serbe Sci. Arts, Cl. Sci. Math. Nat. Sci. Math., 31, p. 2006. University of Mascara, Algeria. E-mail address: [email protected] Oran-Essenia University, Algeria. E-mail address: [email protected]
1504.03157
1
1504
2015-04-13T12:52:50
Properties of differential operators with vanishing coefficients
[ "math.FA", "math.AP" ]
In this paper, we investigate the properties of linear operators defined on $L^p(\Omega)$ that are the composition of differential operators with functions that vanish on the boundary $\partial \Omega$. We focus on bounded domains $\Omega \subset \mathbb{R}^d$ with Lipshitz continuous boundary. In this setting we are able to characterize the spectral and Fredholm properties of a large class of such operators. This includes operators of the form $Lu = \text{div}( \Phi \nabla u)$ where $\Phi$ is a matrix valued function that vanishes on the boundary, as well as operators of the form $Lu = D^{\alpha} (\varphi u)$ or $L = \varphi D^{\alpha} u$ for some function $\varphi \in \mathscr{C}^1(\bar{\Omega})$ that vanishes on $\partial \Omega$.
math.FA
math
PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS DANIEL JORDON Abstract. In this paper, we investigate the properties of linear operators defined on Lp(Ω) that are the composition of differential operators with func- tions that vanish on the boundary ∂Ω. We focus on bounded domains Ω ⊂ Rd with Lipshitz continuous boundary. In this setting we are able to character- ize the spectral and Fredholm properties of a large class of such operators. This includes operators of the form Lu = div(Φ∇u) where Φ is a matrix val- ued function that vanishes on the boundary, as well as operators of the form Lu = Dα(ϕu) or L = ϕDαu for some function ϕ ∈ C 1( ¯Ω) that vanishes on ∂Ω. 1. Introduction In this article we study the properties of linear operators when we allow the leading coefficient functions to vanish on the boundary of the domain. For example, the differential equation: (1) Lu = −div(Φ∇u) = f, where f ∈ Lp(Ω), and Φ(x) ∈ Cd×d has been extensively studied when Φ is uni- formly positive definite on ¯Ω. The operator L is called uniformly elliptic. For more on such operators, see [3, 6, 8, 12, 14] and the references therein. Less is known when the uniform positivity assumption on Φ is relaxed. In [6, §6.6], Trudinger and Gilberg partially relax the condition. In particular, they assume Φ ∈ C 0,γ( ¯Ω) for some γ ∈ (0, 1), and if x0 ∈ ∂Ω then there exists a suitably chosen y ∈ Rd such that Φ(x0) · (x0 − y) 6= 0. With this restriction, they establish existence and unique- ness of solutions to (1). In [15], Murthy and Stampacchia studied the properties of weak solutions to (1) in the case where there exists a positive function m with m−1 ∈ Lp(Ω) such that (2) v · Φ(x)v ≥ m(x)v2, for a.e. x ∈ Ω, and all v ∈ Rd. Studies on the properties of solutions to Lu = f , when L is non-uniformly elliptic, can be found in [18–20], as well as [2], and [5], where the authors assume restrictions on Φ that are similar to (2). The results presented here address the Fredholm properties of L in the case when Φ = 0 on ∂Ω, and/or when v · Φ(x)v ≥ m(x)v for some positive function m ∈ C 1( ¯Ω) with m−1 6∈ Lp(Ω). Other examples of a differential equation with vanishing coefficients arise when studying linear stability of solutions to non-linear PDE. The operator (3) Lu = (ϕu)xxx + (ϕu)x + bux, 1 2 DANIEL JORDON defined on Lp(−1, 1) where ϕ(x) = a cos2( π compactly supported solutions to 2 x) and a, b ∈ R, arose when studying ut = (u2)xxx + (u2)x. Our results can be used to accomplish two goals. The first is assessing the solvability of the boundary value problem Lu = f where u = g on the boundary. To that end, we analyze the Fredholm properties of L. Our second goal is establishing well-posedness (or ill-posedness) of the Cauchy problem ut = Lu, where u = g when t = 0. For this goal, we present results on the spectrum of L. The operators studied here are linear differential operators on Lp(Ω), elliptic or otherwise, that have coefficient functions on the leading order derivative term that vanish on ∂Ω. For the matrix valued function Φ : ¯Ω → Cd×d, we only require that at least one eigenvalue vanishes on ∂Ω. The operators shown in (1) and (3) are examples of operators that can be analyzed using the results presented here. 2. The main results 2.1. Preliminaries. Throughout this article we make the following assumptions: • The domain Ω ⊂ Rd is open and bounded and ∂Ω is Lipshitz continuous. • The function ϕ : ¯Ω → R is such that ϕ ∈ C k( ¯Ω) for some positive integer k, ϕ > 0 on Ω ⊂ Rd, and ker ϕ = ∂Ω. • The ambient function space for the differential operator L is Lp(Ω) where 1 < p < ∞. We say the scalar valued function ϕ is simply vanishing on ∂Ω if for each y ∈ ∂Ω there exists an a 6= 0 such that (4) lim x→y ϕ(x) dist(x, ∂Ω) = a, where the limit is taken in Ω. For the matrix valued function Φ : ¯Ω → Cd×d we make restrictions on its eigenfunctions, defined as the functions ϕi such that Φ(x)v = ϕi(x)v for some v ∈ Cd. We say the matrix valued function Φ : ¯Ω → Cd×d is simply vanishing on ∂Ω if Φ(x) is positive semi-definite in ¯Ω and for each fixed i, the eigenfunction ϕi is either strictly positive on ¯Ω or simply vanishing on ∂Ω, with at least one i such that ϕi is simply vanishing on ∂Ω. 2.2. Results. In this section, we summarize the results proved in this article. In the following theorem, ⌊a⌋ denotes the integer part of a. Theorem 2.1. Let Ω ⊂ Rd be a bounded open set and let ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. Fix m ∈ N and 1 < p < ∞. Assume k ∈ N is such that and that the boundary ∂Ω is C k. If u ∈ Lp(Ω) and ϕmu ∈ W k+m,p(Ω) then k > d p + (m − 1)(cid:4) d p(cid:5), and there exists a c > 0, independent of u, such that u ∈ W κ,p(Ω), where κ := k − m(cid:4) d p(cid:5), kukW κ,p(Ω) ≤ ckϕmukW k+m,p(Ω). The above result is proven in section 5.1 as Theorem 5.7, with the estimate proven in Remark 5.8. The following theorem is proven in section 6.1 as Theorem 6.6. PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 3 Theorem 2.2. Let Ω ⊂ Rd be an open and bounded set with C 0,1 boundary. Let ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. If A is Fredholm on Lp(Ω) with domain W k,p(Ω) for some k > 0, then ϕmA, m ≥ 1, is not closed on its natural domain, D(ϕmA) = {u ∈ Lp : u ∈ D(A), ϕmAu ∈ Lp(Ω)} ≡ D(A), but ϕmA is closable. The above theorem tells us that even simple operators do not have the desirable property of being closed on their 'natural domain'. For example, the operator Lu = sin(πx)uxx is not closed on W 2,p(0, 1) for any p ∈ (1, ∞) by Theorem 2.2. We can use Theorem 2.1 to get an estimate on the properties of the domain, as demonstrated in the following example. Example 2.3. Let Ω = (0, 1) and consider the operator L acting on Lp(Ω) defined by Lu = ϕuxxx where ϕ is simply vanishing. The operator L is of the form ϕA where A is Fredholm. By Theorem 2.2, L is not closed on its natural domain, W 3,p(Ω), but is closable. Let ¯L denote the closure of L, and let ϕ(k) : Lp(Ω) → W 3−k,p(Ω) denote the multiplication operator u 7→ ϕ(k)u where the function ϕ(k) denotes the k-th derivative of the function ϕ, and as an abuse of notation we set ϕ = ϕ(0). After rewriting L as Lu = (ϕu)xxx − 3(ϕ(1)u)xx + 3(ϕ(2)u)x − ϕ(3)u, we see that D( ¯L) = D(A3ϕ) ∩ D(A2ϕ(1)) ∩ D(Aϕ(2)) ∩ D(ϕ(3)), where A is the derivative operator on Lp(Ω). Specifically, if ϕ ∈ C 3( ¯Ω) is simply vanishing, then the fact that D(A3) = W 3,p(Ω) implies D(A3ϕ) ⊂ W 2,p(Ω) by Theorem 2.1. By the same theorem, we have W 2,p(Ω) ⊂ D(Akϕ(3−k)) for k = 0, 1, 2 so the best we can do is D( ¯L) ⊂ W 2,p(Ω). More concretely, if we set ϕ(x) = sin(πx) and fix p = 2, then one can construct functions u ∈ D( ¯L) such that u 6∈ W 3,2(Ω) but u ∈ W 2,2(Ω). See [9, §3.1]. The following theorem speaks about the range of the multiplication operator. It is proved in section 5.2 as Theorem 5.9. Theorem 2.4. Let Ω ⊂ Rd be open and bounded with C 0,1 boundary and assume the function ϕ ∈ C k( ¯Ω) is simply vanishing on ∂Ω. Then the range of the operator u 7→ ϕmu is closed in W k,p(Ω) whenever k ≥ m and is not closed when k < m. If we know the range of the multiplication operator u 7→ ϕmu is closed in W k,p(Ω) then we necessarily have for some constant c > 0. kukLp(Ω) ≤ ckϕmukW k,p(Ω) The matrix valued analogs of Theorems 2.1 and 2.4 are proved in section 5.3. One implication is illustrated in the following example. Example 2.5. Let Ω ⊂ Rd be open and bounded with C 0,1 boundary. Let Φ ∈ C 1( ¯Ω; Rd×d) be simply vanishing on ∂Ω. Then by the matrix analog of Theorem 2.4 (which is Theorem 5.11) we know that the range of Φm, m ∈ N, is closed in W 1,2(Ωd) := W 1,2(Ω) × · · · × W 1,2(Ω), d copies {z } 4 DANIEL JORDON if and only if m = 1. This implies Φm : L2(Ωd) → W 1,2(Ωd) is semi-Fredholm if and only if m = 1. Now, it is well known that the weak gradient ∇ : W 1,2(Ω) ⊂ L2(Ω) → L2(Ωd) and its adjoint, div(·) : L2(Ωd) → L2(Ω), are semi-Fredholm. Thus, the non-uniformly elliptic operator, (5) Lu = div(Φm∇u), is semi-Fredholm on L2(Ω) if and only if m = 1. Theorem 2.6. Let Ω ⊂ Rd be open and bounded with C 0,1 boundary, m, k ∈ N with 0 < m < k, and let A be densely defined on Lp(Ω). Assume the following: • The operator A is closed on Lp(Ω) and D(A) ⊂ W k,p(Ω). • There exists a u ∈ D(A) such that u /∈ W m,p • The resolvent set ρ(A) is non-empty. (Ω) ∩ W k,p(Ω). 0 If ϕ ∈ C k( ¯Ω) is simply vanishing then σess(Aϕm) = σess(ϕmA∗) = C. Moreover, if either Aϕm or ϕmA∗ is Fredholm then σp(ϕmA∗) = C. In the above theorem, σp(L) and σess(L) denote the point spectrum and essential spectrum of L respectively. The definition of the essential spectrum is given in section 6.2 and the result is proven as Theorem 6.14. Its import is demonstrated in the following example. Example 2.7. This example continues from Example 2.3, where Ω = (0, 1), and L = ϕuxxx. For any u ∈ W 3,p(Ω), there exists a, b ∈ R such that u + ax + b ∈ W 1,p (Ω), which implies 0 W 3,p(Ω) = W 3,p(Ω) ∩ W 1,p 0 (Ω) ⊕ span{1, x}. 0 Now consider the multiplication operator ϕ : Lp(Ω) → W 3,p(Ω) given by u 7→ ϕu where ϕ is simply vanishing on ∂Ω. Then we know that the range of ϕ is W 3,p(Ω) ∩ W 1,p (Ω), which has co-dimension 2 in W 3,p(Ω). Thus, if we let A denote three applications of the weak derivative operator on Lp(Ω), then we know that ρ(A) is nonempty, and D(A) = W 3,p(Ω) ⊂⊂ Lp(Ω) by the Rellich-Kondrachov Theorem. Then we see that Aϕ has finite dimensional nullspace and the range has finite co-dimension. This shows Aϕ is Fredholm. Applying Theorem 2.6 yields σp( ¯L) = σp(ϕA) = C. More concretely, we have σp(sin(πx)uxxx) = C. The same holds if we set ϕ(x) = sin2(πx), but not necessarily when we set ϕ(x) = sinm(πx) where m ∈ N and m ≥ 3. 2.3. Outline of the article. The article is structured as follows: • Section 3 introduces the notation and basic definitions that are used in this article. • Section 4 goes over some basic properties of closed and Fredholm operators. • Section 5 covers the properties of the operators u 7→ ϕu and u 7→ Φu. The domain and range of the operator u 7→ ϕu is covered in sections 5.1 and 5.2 respectively. Matrix valued operators are handled in section 5.3 PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 5 • We study various properties of differential operators composed with van- ishing operators in section 6. In particular, we focus on the Fredholm properties and spectra of the operators Aϕ and ϕA where A is a Fredholm differential operator. 3. Notation and definitions We will review some of the basic definitions and introduce the notation used in this article. We will use capital letters, such as W , X, Y , or Z, to denote a Banach space. We use B(X, Y ) to denote the set of bounded linear operators from X to Y , and C(X, Y ) to denote the set of closed and densely defined linear operators from X to Y . The sets B(X) and C(X) denote the sets B(X, X) and C(X, X) respectively. The domain and range of a linear operator A will be denoted by D(A) and R(A) respectively, and we use N(A) to denote the nullspace of A. If A ∈ C(X, Y ), then D(A) equipped with the graph norm, kxkD(A) := kxkX + kAxkY , x ∈ D(A), is a Banach space, and we call k · kD(A) the A-norm. When referring to the com- position of two linear operators A and B, the subspace D(AB) = { x ∈ D(B) : Bx ∈ D(A) }, is called the natural domain of AB. If A is a linear operator from X to Y , any closed operator A1 where D(A) ⊂ D(A1) and A = A1 on D(A) is called a closed extension of A. We call A closable if there exists a closed extension of A. We denote ¯A as the closure of A, and it is the 'smallest' closed extension, in the sense that D( ¯A) ⊂ D(A1) for any operator A1 that is a closed extension of A. An operator is closable if every sequence {xn} ⊂ D(A) where xn → 0 in X and Axn → y in Y implies y = 0. A Banach space Y is said to be continuously embedded in another Banach space X if there exists an operator P ∈ B(Y, X) that is one-to-one. The space Y is said to be compactly embedded in X if P is also compact and we write Y ⊂⊂ X whenever Y is compactly embedded in X. For Sobolev spaces, we take P to be the inclusion operator, which we denote as ι. Most of the analysis takes place in Lp(Ω) and the Sobolev spaces W k,p(Ω), where k ∈ N, Ω ⊂ Rd is an open and bounded set, and, unless stated otherwise, 1 < p < ∞. The closure of a set Ω ⊂ Rd will be denoted by ¯Ω and the boundary of Ω will be denoted by ∂Ω. The space C k(Ω) denotes the space of all functions from Ω to R that are k-times continuous differentiable everywhere in Ω, and C k 0 (Ω) ⊂ C k(Ω) denotes the subspace of those functions with compact support in Ω. The space W k,p 0 (Ω) in W k,p(Ω). If u is weakly differentiable, we let Dαu denote the α-th weak derivative of u, where α = (α1, . . . , αd) ∈ Zd + is a multi-index, and we let α = α1 + · · · + αd denote the order of α. We use ∇(k)u to denote the vector of all weak derivatives of u with order k, and set ∇u = ∇(1)u to be the gradient of u. (Ω) denotes the closure of C ∞ 0 We say ∂Ω is C k,γ if for each point y ∈ ∂Ω, there exists an r > 0 and a C k,γ function γ : Rd−1 → R such that Ω ∩ B(y, r) = {x ∈ B(y, r) : xd > γ(x1, . . . , xd−1)}, where B(x, r) := {y ∈ Rd : x − y < r}, and C k,γ is a Holder space. 6 DANIEL JORDON Definition 3.1. Let Ω ⊂ Rd be a bounded and open set. For any ϕ ∈ C 1( ¯Ω), we will call the multiplication operator u 7→ ϕu vanishing if ker ϕ = ∂Ω. As an abuse of notation, we will use ϕ to refer to the multiplication operator u 7→ ϕu. Definition 3.2. Let Ω ⊂ Rd be a bounded and open set. Take dist(x, ∂Ω) := inf y∈∂Ω x − y to be the distance from x to the boundary of Ω. Let ϕ ∈ C 1( ¯Ω). We say ϕ is simply vanishing on ∂Ω if ker ϕ = ∂Ω, and for each y ∈ ∂Ω there exists an a 6= 0 such that (6) lim x→y ϕ(x) dist(x, ∂Ω) = a, where the limit is taken within Ω. The multiplication operator u 7→ ϕu is called simply vanishing on ∂Ω if the function ϕ is simply vanishing on ∂Ω. Definition 3.3. Let Ω ⊂ Rd be a bounded and open set. We say the function ϕ ∈ C m( ¯Ω) is vanishing of order m on ∂Ω if ker ϕ = ∂Ω, Dαϕ = 0 on ∂Ω when α < m, and ∇(m)ϕ 6= 0 on ∂Ω. The multiplication operator u 7→ ϕu is called vanishing of order m on ∂Ω if the function ϕ is vanishing of order m on ∂Ω. Functions that are vanishing of order 1 are simply vanishing functions. To see why, take Ω ⊂ Rd with C 1 boundary and assume ϕ ∈ C k( ¯Ω) is simply vanishing. Fix any y ∈ ∂Ω and let Ωn = B(y, n−1) ∩ Ω. Since ∂Ω is C 1, there exists a point xn ∈ Ωn such that xn − y = dist(xn, ∂Ω). Given ϕ is simply vanishing, there exists an a 6= 0 such that a = lim n→∞ ϕ(xn) dist(xn, ∂Ω) = lim n→∞ ϕ(xn) − ϕ(y) xn − y = ∇ϕ(y), which shows ∇ϕ(y) 6= 0. Since y ∈ ∂Ω was arbitrary, we see that ∇ϕ 6= 0 on ∂Ω whenever ∂Ω is C 1. We have a similar definition for matrix-valued functions. Let Φ : ¯Ω → Cd×d be Hermitian for each x ∈ ¯Ω. Then there exists a unitary matrix U(x) and a real diagonal matrix D(x) such that Φ = UDU∗, (7) by Schur's decomposition theorem. If Φ ∈ C 1( ¯Ω; Cd×d), we can choose U and D in C 1( ¯Ω; Cd×d) so that (7) holds. Thus, we lose no generality by assuming the operator D has the form D = diag(ϕ1, . . . , ϕd) for some functions ϕi ∈ C 1( ¯Ω) where i = 1, . . . , d. Definition 3.4. Let Ω ⊂ Rd be open and bounded, and let Φ ∈ C m( ¯Ω; Cd×d). We say the function Φ is vanishing of order m if Φ is positive semi-definite on ¯Ω, and the matrix D = diag(ϕ1, . . . , ϕd) in its Schur decomposition has the following property: for each i = 1, . . . , d, either ϕi > 0 on ¯Ω or ϕi is vanishing of order mi on ∂Ω with mi ≤ m, with at least one function ϕi that is vanishing of order m on ∂Ω. The multiplication operator u 7→ Φu is called vanishing of order m if the function Φ is vanishing of order m. 3.1. Fredholm and semi-Fredholm operators. We will utilize Fredholm oper- ator theory when describing the properties of the multiplication operators u 7→ ϕu and u 7→ Φu. In this section, we briefly review the theory of Fredholm operators. Let X and Y be Banach spaces. An operator A : X → Y is called Fredholm if (a) The domain of A is dense in X, PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 7 (b) The operator A is closed on its domain, (c) The nullspace of A is finite dimensional, (d) The range of A is closed in Y , (e) The co-dimension of the range of A is finite dimensional, where the co-dimension of a closed subspace M ⊂ Y , denoted co-dim M , is the dimension of the quotient space Y /M . We use F(X, Y ) to denote the set of Fredholm operators from X to Y and write F(X) in place of F(X, X). Note that property (e) is equivalent to requiring that the nullspace of the adjoint operator A∗ be finite dimensional. The index of a Fredholm operator A is defined as ind(A) := dim N(A) − co-dim R(A). The set of semi-Fredholm operators from X to Y , denoted F+(X, Y ), is the set of operators that satisfy all the properties of Fredholm opertors except possibly property (e). The set of semi-Fredholm operators from X to X will be denoted by F+(X). We note that our definition for F+(X, Y ) is sometimes referred to as the set of upper semi-Fredholm operators. The following characterization of Fredholm operators is useful. Theorem 3.5 ([17], Theorem 7.1). Let X and Y be Banach spaces. Then A ∈ F(X, Y ) if and only if there exists closed subspaces X0 ⊂ X and Y0 ⊂ Y where Y0 is finite dimensional and X0 has finite co-dimension such that X = X0 ⊕ N(A), Y = R(A) ⊕ Y0. Moreover, there exists operators A0 ∈ B(Y, X), K1 ∈ B(X), and K2 ∈ B(Y ) where • The N(A0) = Y0, • The R(A0) = X0 ∩ D(A), • A0A = I − K1 on D(A), • AA0 = I − K2 on Y , • The N(K1) = X0, while K1 = I on N(A), • The N(K2) = R(A), while K2 = I on Y0. Note that K1 and K2 are projection operators and their ranges are finite dimen- sional. The operator A0 from Theorem 3.5 will be referred to as the pseudo-inverse of A, since AA0A = A and A0AA0 = A0. An equivalent characterization of semi-Fredholm operators is as follows: if X and Y are Banach spaces and A ∈ C(X, Y ), then A is not semi-Fredholm if and only if there exists a bounded sequence {xk} ⊂ D(A) having no convergent subsequence such that {Axk} converges. A proof of this equivalence can be found in [16] or [17, p. 177]. Remark 3.6. Suppose Ω ⊂ Rd is open and bounded, ϕ ∈ C m( ¯Ω) is simply vanishing on ∂Ω and ζ ∈ C m( ¯Ω) is vanishing of order m on ∂Ω. Then the multiplication operator u 7→ ϕmu is semi-Fredholm from Lp(Ω) to W k,p(Ω) if and only if the mapping u 7→ ζu is semi-Fredholm. Moreover, D(ϕm) = D(ζ). To see why, use the fact that the multiplication operators u 7→ ζϕ−mu and u 7→ ϕmζ−1u are one-to- one and onto Lp(Ω) and that the composition of a semi-Fredholm operator with a Fredholm operator is semi-Fredholm. 8 DANIEL JORDON 4. Basic properties of closed operators Fredholm operators are closed under composition. That is, if X, Y , and Z are Banach spaces, then A ∈ F(X, Y ) and B ∈ F(Y, Z) implies BA ∈ F(X, Z) with ind(BA) = ind(B) + ind(A). Moreover, if B ∈ C(Y, Z) and BA ∈ F(X, Z) then we necessarily have B ∈ F(Y, Z). These claims are proved, respectively, in [17, p. 157] as Theorem 7.3, and [17, p. 162] as Theorem 7.12. As for semi-Fredholm operators, we have the following. Lemma 4.1 ([16], Lemma 4). Let X, Y , and Z be Banach spaces and let A ∈ F(X, Y ), and B ∈ C(Y, Z). If BA ∈ F+(X, Z), then B ∈ F+(Y, Z). One of the theorems that we use throughout this article is the following conse- quence of the Closed Graph Theorem. A proof of the Closed Graph Theorem can be found in many functional analysis textbooks, such as [17, p. 62] or [11, p. 166]. Lemma 4.2. Let X be a Banach space and Y ⊂ X. If there exists a norm that converts Y into a Banach space then there exists a c > 0 such that kykX ≤ ckykY for all y ∈ Y . If Y = X then their norms are equivalent. Proof. Let ι denote the inclusion map from Y to X. It is a closed operator with domain equal to Y , so by the Closed Graph Theorem it is bounded. If Y = X then apply the above argument to the inclusion map from X to Y . (cid:3) The above lemma is useful for showing that special subsets of Lp(Ω) have certain properties - such as compactness - since they can inherit such properties from other Sobolev spaces. One of the most important consequences of compactness is the following theorem. Theorem 4.3. Let X and Y be Banach spaces. If A ∈ C(X, Y ) with D(A) ⊂⊂ X, then A ∈ F+(X, Y ). Proof. We prove this by contradiction. Suppose A is not semi-Fredholm. Since A ∈ C(X, Y ), this implies there exists a bounded sequence {xn} ⊂ D(A) having no convergent subsequence, such that {Axn} is convergent in Y . But if {xn} is bounded in X and {Axn} is convergent in Y , then {xn} is a bounded sequence in the A-norm. Since D(A) ⊂⊂ X there exists a subsequence of {xn} that is convergent in X, the desired contradiction. (cid:3) Remark 4.4. We know an operator A ∈ C(X, Y ) has closed range if and only if there exists a c > 0 such that inf z∈N(A) kx − zkX ≤ ckAxkY , for all x ∈ D(A). Theorem 4.3 can be used to quickly establish estimates involving differential oper- ators. We can, for example, establish Poincar´e inequalities. It is also well known that W 1,p 0 Ω ⊂ Rd. Since the weak gradient operator ∇ is closed on W 1,p closed by Theorem 4.3. This implies the existence of a c > 0 such that 0 (Ω) ⊂⊂ Lp(Ω) for any bounded and open set (Ω), we get R(∇) is kukLp(Ω) = inf a∈N(∇) ku − akLp(Ω) ≤ ck∇ukLp(Ω), for any u ∈ W 1,p 0 (Ω). PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 9 5. Properties of vanishing operators 5.1. The domain of a vanishing operator. In this section we establish proper- ties of the domain of the vanishing operator ϕ : Lp(Ω) → W k,p(Ω). In particular, we will establish the embedding of the domain of ϕ in various Sobolev spaces. If ϕ ∈ C 1( ¯Ω), then the mapping u 7→ ϕu is bounded from Lp(Ω) to Lp(Ω). This implies that a natural choice for its domain is Lp(Ω). Whenever a vanishing operator ϕ is composed with a differential operator A to form Aϕ - A being an operator that is closed on W k,p(Ω) - it makes sense to think of ϕ as a densely defined operator that maps some subset of Lp(Ω) to the space W k,p(Ω). This and subsequent sections rely heavily on Hardy's inequality, so we include the statement for the reader's convenience. Theorem 5.1 ([21], Hardy's Inequality). Let Ω ⊂ Rd be a bounded open set with C 0,1 boundary, and δ(x) = inf y∈∂Ω x − y. Then, for all u ∈ C ∞ 0 (Ω), where c > 0 depends on Ω, p, d, and m. kδ−mukLp(Ω) ≤ ck∇(m)ukLp(Ω), See [7, 13] for recent developments on the assumptions necessary for Hardy's inequality. The interested reader should consult [14, §2.7] for a treatment of optimal constants for Hardy's inequality. We begin with basic properties of the domain and range of the multiplication operator u 7→ ϕmu. Lemma 5.2. Let Ω ⊂ Rd be open and bounded and ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. For each k, m ∈ Z+, the multiplication operator ϕm : Lp(Ω) → W k,p(Ω) defined as u 7→ ϕmu is closed on D(ϕm) = {u ∈ Lp(Ω) : ϕmu ∈ W k,p(Ω)}. Moreover, if ϕ ∈ C m( ¯Ω) and m ≤ k then R(ϕm) = W k,p(Ω) ∩ W m,p 0 (Ω). Proof. The proof is broken into two claims. Claim 1: The multiplication operator u 7→ ϕmu is closed on D(ϕm). We first show it is closable. Suppose un → 0 in Lp(Ω) and ϕmun → y in W k,p(Ω). Then we know that ϕmun → y in Lp(Ω). But since ϕ is bounded and un → 0 in Lp(Ω) we can conclude ϕmun → 0 = y in Lp(Ω). This shows the multiplication operator ϕm is closable on its domain. But any closed extension cannot be defined on a set larger than D(ϕm), implying the domain of any closed extension must be D(ϕm). This completes the proof of the claim. Claim 2: If ϕ ∈ C m( ¯Ω) and m ≤ k then R(ϕm) = W k,p(Ω) ∩ W m,p Take v ∈ W k,p(Ω)∩W m,p (Ω) we can apply Hardy's inequal- ity (Theorem 5.1) to show that ϕ−mv ∈ Lp(Ω). Since this implies ϕ−mv ∈ D(ϕm) we have W k,p(Ω) ∩ W m,p (Ω). Since v ∈ W m,p (Ω) ⊂ R(ϕm). (Ω). 0 For the other direction, first note that since ϕm ∈ C m( ¯Ω) ∩ W m,p (Ω) there exists a sequence {φn} ⊂ C ∞ 0 (Ω) such that φn → ϕm in W m,p(Ω) and {Dαφn} is uniformly bounded when α ≤ m1. Let u ∈ D(ϕm) be arbitrary, and set vn = φn − 0 0 0 0 1One such example are the functions φn = ϕm1Ωn ∗ η1/3n, where 1Ωn is an indicator function for the set Ωn = {x ∈ Ω : dist(x, ∂Ω) > 1/n} and ηǫ is a mollifier. 10 DANIEL JORDON ϕm. Since Dα(φnu) is bounded by cDα(ϕu) for some constant c and Dα(vnu)p → 0 almost everywhere when α ≤ m we have lim n→∞ kDα(φnu) − Dα(ϕmu)kLp(Ω) = lim n→∞ kDα(vnu)kLp(Ω) = 0, when α ≤ m by dominated convergence. Noting that φnu ∈ W m,p ϕmu ∈ W m,p (Ω) and completes the proof. 0 0 (Ω) shows (cid:3) The following lemma establishes the relative compactness of D(ϕm) in Lp(Ω) when ϕm maps to W k,p(Ω) for k > m. It uses the relative compactness of W m+1,p(Ω) in W m,p(Ω). This is implied by the Rellich-Kondrachov Theorem, which establishes that for 1 ≤ p < ∞, W 1,p(Ω) is compactly embedded in Lp(Ω) whenever Ω is a bounded domain with Lipshitz continuous boundary. See [1, p. 168] Thoerem 6.3 for the full statement and proof of the Rellich-Kondrachov Theorem. Lemma 5.3. Let Ω ⊂ Rd be open and bounded with C 0,1 boundary and let ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. If D(ϕm) = {u ∈ Lp(Ω) : ϕmu ∈ W m+1,p(Ω)} then D(ϕm) ⊂⊂ Lp(Ω). Proof. Since Ω is bounded and ∂Ω is C 0,1, we can use the Rellich-Kondrachov Theorem to establish W m+1,p(Ω) ⊂⊂ W m,p(Ω). Suppose {un} ⊂ D(ϕm) is such that kunkD(ϕm) ≤ 1 for each n. Then kϕmunkW m+1,p(Ω) ≤ 1 for each n, so there exists a subsequence that is convergent in W m,p(Ω). After relabeling the convergent subsequence, we take this to be the entire sequence. Applying Hardy's inequality (Theorem 5.1) yields lim n,k→∞ kun − ukkLp(Ω) ≤ lim n,k→∞ ckϕmun − ϕmukkW m,p(Ω) = 0, completing the proof. (cid:3) Next we establish the embedding of D(ϕ) in various Sobolev spaces. To do so, we will use the fact that when Ω ⊂ Rd is open and bounded with C k boundary, the map from C k( ¯Ω) to C k(∂Ω) can be extended to a continuous surjective linear map from W k,p(Ω) to W k−1/p,p(∂Ω) where 1 < p < ∞ (see [3, p. 158] Theorem 3.79). u 7→ u∂Ω 0 We would like to highlight the fact that the trace map T on W k,p(Ω) is defined on a Banach space and has range that is onto the Banach space W k−1/p,p(∂Ω). As for the nullspace of T , a classical result states that when ∂Ω is C 1, T u = 0 if and only if u ∈ W 1,p (Ω); (see [4, p. 259] Theorem 2, or [3, p. 138] Corollary 3.46). Let Ω be an open and bounded set with C k boundary and let T denote the continuous trace operator from W k,p(Ω) onto W k−1/p,p(∂Ω). We will need a trace- like operator that is one-to-one. To define this new operator, first set W0 := W 1,p (Ω) ∩ W k,p(Ω) = N(T ). Clearly W0 is a closed subspace of W k,p(Ω). Next let W k denote the quotient space W k,p(Ω)/W0 and define the operator T from W k to W k−1/p,p(∂Ω) as 0 T [u] = T u, [u] ∈ W k. PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 11 The operator T is well-defined, linear, and one-to-one. To see that T is closed, take {[un]} ⊂ W k such that [un] → [u] and T [un] → y as n → ∞. Then k[un]−[u]k W k → 0 implies the existence of a sequence {vn} ⊂ W0 such that un − vn → u, in W k,p(Ω). Since T [un] → y, we know that T un = T (un − vn) → y as n → ∞. By the boundedness of T we get T u = y. This implies T [u] = y and proves that T is closed. Applying the Closed Graph Theorem shows that T is bounded. Moreover, the fact that T is surjective implies that T is surjective as well. This tells us that T −1 exists and is a bounded linear operator from W k−1/p,p(∂Ω) onto W k, by the Bounded Inverse Theorem. We also need the following general Sobolev space theorem. We use ⌊a⌋ to denote the integer part of a. Theorem 5.4 ([1] p. 85, Sobolev Imbedding). Let Ω ⊂ Rd be open and bounded with a C 0,1 boundary. Assume u ∈ W k,p(Ω), and that kp > d. Set κ = k −j d pk − 1, any number in (0, 1), otherwise. Then there exists a function u∗ such that u∗ = u a.e. and u∗ ∈ C κ,γ( ¯Ω). γ =( 1 −(cid:16) d p −j d pk(cid:17) , if d p 6∈ N We can now establish the following lemma. Lemma 5.5. Let Ω ⊂ Rd be a bounded open set and let ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. Assume k ∈ N with kp > d and that the boundary ∂Ω is C k. Then for every u ∈ Lp(Ω) where ϕu ∈ W k+1,p(Ω) we have ϕDαu ∈ W 1,p 0 (Ω) ∩ W k,p(Ω), for any α with α = 1. Proof. The proof is divided into two claims. Claim 1: If α = 1 then Dα(ϕu) = uDαϕ on ∂Ω. Since kp > d and ϕu ∈ W k+1,p(Ω), we know there exists a γ ∈ (0, 1) dependent on d and p such that ϕu ∈ C p(cid:5),γ k−(cid:4) d ( ¯Ω), by Theorem 5.4. This shows ϕu ∈ C 1( ¯Ω). Also, since u = ϕ−1ϕu whenever ϕ 6= 0, the fact that ϕ ∈ C 1( ¯Ω) and is nonzero in Ω implies u ∈ C 1(Ω). Now, since ϕ is simply vanishing on ∂Ω, we know that for any y ∈ ∂Ω, lim x→y x − y ϕ(x) = lim x→y x − y ϕ(x) − ϕ(y) = ∇ϕ(y)−1, where the limit is taken in Ω. Note that ∇ϕ 6= 0 on ∂Ω, so that this is well defined. Next, given ϕu ∈ W 1,p(Ω) and u ∈ Lp(Ω), we know ϕu ∈ W 1,p (Ω) by Lemma 5.2. We already know ϕu is continuous on ¯Ω, which implies ϕu = 0 on ∂Ω. Thus, for any y ∈ ∂Ω we can take any sequence in Ω that converges to y and obtain 0 (8) lim x→y u(x) = lim x→y ϕ(x)u(x) − ϕ(y)u(y) x − y x − y ϕ(x) = ∇(ϕu)(y)∇ϕ(y)−1. By Leibniz's rule, Dα(ϕu) = uDαϕ+ϕDαu when α = 1, so the above limit implies Dα(ϕu) = uDαϕ on ∂Ω. 12 DANIEL JORDON Claim 2: The function ϕDαu is in W 1,p Set W0 := W 1,p 0 (Ω) ∩ W k,p(Ω) and W k := W k,p(Ω)/W0. By assumption, ϕu ∈ W k+1,p(Ω) so Dα(ϕu) ∈ W k,p(Ω) whenever α = 1. This implies the coset [Dα(ϕu)] is in W k and that its trace T [Dα(ϕu)] is in W k−1/p,p(∂Ω). By claim 1, 0 (Ω) ∩ W k,p(Ω) whenever α = 1. uDαϕ = Dα(ϕu) on ∂Ω, which shows that uDαϕ∂Ω ∈ W k−1/p,p(∂Ω) and that T −1(uDαϕ∂Ω) ∈ W k. The one-to-one nature of T −1 implies [Dα(ϕu)] = [uDαϕ]. Since these cosets are equal, there exists a function v ∈ W0 such that uDαϕ = Dα(ϕu) − v. But again, Dα(ϕu) = uDαϕ + ϕDαu, so it must be the case that v = ϕDαu. We then conclude that ϕDαu ∈ W0 = W 1,p (Ω) ∩ W k,p(Ω), completing the proof. (cid:3) 0 Theorem 5.6. Let Ω ⊂ Rd be a bounded open set and let ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. Assume k ∈ N with kp > d and that ∂Ω is C k. Then u ∈ Lp(Ω) Proof. Fix the multi-index α with α ≤ κ. Choose a finite sequence of multi-indices p(cid:5). and ϕu ∈ W k+1,p(Ω) implies u ∈ W κ,p(Ω) where κ := k −(cid:4) d {αn}n≤α each with αn = 1 such that P αn = α. By assumption, kp > d so we may apply Lemma 5.5 to u to obtain (9) ϕDα1 u ∈ W 1,p (Ω) ∩ W k+1−α1,p(Ω). 0 Applying Hardy's inequality to ϕDα1 u yields (10) Dα1 u ∈ Lp(Ω). Moreover, given α ≤ κ = k −(cid:4) d p(cid:5), we know that (11) k + 1 − α1 ≥ k + 1 − α ≥ 1 +(cid:4) d p(cid:5) > d p . Since (9), (10), and (11) all hold, we may apply Lemma 5.5 to Dα1 u to obtain ϕDα1+α2u ∈ W 1,p (Ω) ∩ W k−1,p(Ω). Another application of Hardy's inequality shows Dα1+α2 u ∈ Lp(Ω). We continue inductively applying Lemma 5.5 and Hardy's inequality at each step to finally show that 0 ϕDαu ∈ W 1,p 0 (Ω) ∩ W k+1−α,p(Ω), and Dαu ∈ Lp(Ω). Since this applies to any multi-index α with α ≤ κ, we see that u ∈ W κ,p(Ω), completing the proof. (cid:3) Iteratively applying the above theorem yields the following. Theorem 5.7. Let Ω ⊂ Rd be a bounded open set and let ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. Fix m ∈ N and 1 < p < ∞. Assume k ∈ N is such that (12) k > d p + (m − 1)(cid:4) d p(cid:5), and that ∂Ω is C k. If u ∈ Lp(Ω) and ϕmu ∈ W k+m,p(Ω) then (13) p(cid:5). u ∈ W κ,p(Ω), where κ := k − m(cid:4) d PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 13 Proof. For convenience, we define the variables κ1, . . . , κm as follows p(cid:5), κj := k + m − j − j(cid:4) d j = 1, . . . , m. We know ϕm−1u ∈ Lp(Ω), ϕmu ∈ W k+m,p(Ω), and k + m − 1 > p−1d. By Theo- rem 5.6, this implies ϕm−1u ∈ W κ1,p(Ω). If m > 1, we see that κ1 − 1 > p−1d, and we have ϕm−2u ∈ Lp(Ω) and ϕm−1u ∈ W κ1,p(Ω). Thus we get ϕm−2u ∈ W κ2,p(Ω) by the same theorem. Continuing inductively, we apply Theorem 5.6 at each step to get ϕm−ju ∈ W κj ,p(Ω) for m > j. When j = m − 1 we have u ∈ Lp(Ω) and ϕu ∈ W κm−1,p(Ω). Since we may apply Theorem 5.6 one more time to get u ∈ W κm,p(Ω), as desired. κm−1 − 1 = k − (m − 1)(cid:4) d p(cid:5) > d p , (cid:3) Remark 5.8. There is an implicit estimate accompanying Theorem 5.7. Assume ϕ ∈ C 1( ¯Ω) and consider the multiplication operator ϕm : Lp(Ω) → W k+m,p(Ω) where k satisfies (12). Theorem 5.7 tells us that u ∈ D(ϕm) implies u ∈ W κ,p(Ω) where κ is given by (13). Thus, D(ϕm) ⊂ W κ,p(Ω). By the closedness of ϕm, D(ϕm) is a Banach space with the operator norm. We can conclude that, for some constants c0, c1 > 0 and for all u ∈ D(ϕm), (14) (15) kukW κ,p(Ω) ≤ c0kukD(ϕm) = c0kukLp(Ω) + c0kϕmukW k+m,p(Ω) ≤ c1kϕmukW k+m,p(Ω), where (14) follows from Lemma 4.2 applied to the Banach spaces D(ϕm) and W κ,p(Ω) and (15) from Hardy's inequality. 5.2. The range of a vanishing operator. Having a closed range is a very useful property for linear operators. As we will see in section 6.2, it is often necessary for establishing basic properties of the spectrum. Showing the multiplication operator u 7→ ϕu has closed range requires keeping track of the multiplicity of the roots of the function ϕ. This is formally established in the following result. Theorem 5.9. Let Ω ⊂ Rd be open and bounded with C 0,1 boundary and assume the function ϕ ∈ C k( ¯Ω) is simply vanishing on ∂Ω. Then the range of the operator u 7→ ϕmu is closed in W k,p(Ω) whenever k ≥ m and is not closed when k < m. Proof. As usual, we treat ϕ as an operator from some dense subset of Lp(Ω) to W k,p(Ω). We start with the following. Claim 1: If k ≥ m then the range of ϕm is closed in W k,p(Ω). If k = m then R(ϕm) = W m,p (Ω) as discussed in Lemma 5.2, which clearly establishes the closedness of R(ϕm) in W m,p(Ω). If k > m then we may apply Lemma 5.3 to show D(ϕm) ⊂⊂ Lp(Ω). Invoking Theorem 4.3 proves ϕm is semi- Fredholm, which implies R(ϕm) is closed in W k,p(Ω). 0 Claim 2: If k < m then the range of ϕm is not closed in W k,p(Ω). We prove this claim by contradiction. Suppose ϕm has closed range in W k,p(Ω). This implies ϕm is semi-Fredholm from Lp(Ω) to W k,p(Ω). Since ϕk is Fredholm from Lp(Ω) to W k,p (Ω), and since ϕm = ϕm−kϕk is semi-Fredholm from Lp(Ω) to W k,p(Ω), Lemma 4.1 implies ϕm−k is semi-Fredholm from W k,p (Ω) to W k,p(Ω). 0 (Ω), so ϕm−kv ∈ 0 (Ω) we know that v = ϕk−mu ∈ C k Now, for any function u ∈ C ∞ 0 0 14 DANIEL JORDON C ∞ 0 (Ω), implying ϕm−k is onto the subspace C ∞ 0 (Ω). This implies C ∞ 0 (Ω) ⊂ R(ϕm−k). Since R(ϕm−k) is closed, we know that W k,p we also know that ϕk ∈ W k,p that ϕm−kv = ϕk, which implies v = ϕ2k−m. But ϕ2k−m cannot be in W k,p Hardy's inequality would then show (Ω) is a subspace of R(ϕm−k). But (Ω) such (Ω) as (Ω), so there exists a function v ∈ W k,p 0 0 0 0 kϕk−mkLp(Ω) = kϕ−kϕ2k−mkLp(Ω) ≤ ckϕ2k−mkW k,p(Ω) < ∞. This is our desired contradiction. (cid:3) Example 5.10 (The Legendre differential equation). Set Ω = (−1, 1). Let us analyze the operator L given by Lu(x) = d dx(cid:20)(1 − x2) d dx u(x)(cid:21) , acting on Lp(Ω), where as usual we assume 1 < p < ∞. Let A denote the derivative operator on Lp(Ω) and ϕ(x) = (1 − x2). The domain of A is W 1,p(Ω), the nullspace of A is span{1}, and the range of A is equal to Lp(Ω). Since ϕ is simply vanishing on ∂Ω, we know the range of the multiplication operator u 7→ ϕu is equal to W 1,p (Ω) by Lemma 5.2. If u ∈ W 1,p(Ω), then u ∈ C 0,1/p( ¯Ω) by Sobolev Imbedding (Theorem 5.4). Thus, (Ω). we can find a unique line l(x) such that u + l = 0 on ∂Ω, implying u + l ∈ W 1,p Since u ∈ W 1,p(Ω) was arbitrary, this implies 0 0 W 1,p(Ω) = W 1,p 0 (Ω) ⊕ span{1, x}. If we take A to be the restriction of the derivative operator to W 1,p 0 dim N( A) = 0 and co-dim R( A) = 1. Since A, A, and ϕ : Lp(Ω) → W 1,p Fredholm we know that L = AϕA is Fredholm, with 0 (Ω), then (Ω) are all ind(L) = ind( AϕA) = ind( A) + ind(ϕ) + ind(A) = −1 + 0 + 1 = 0, and N(L) = span{1}. In terms of the domain of L, we automatically get D(L) ⊂ D(A) = W 1,p(Ω). The interesting thing to note is that L cannot be semi-Fredholm if D(L) ⊆ W 2,p(Ω). To see this, first note that A maps W 2,p(Ω) onto W 1,p(Ω) and that the range of ϕ : W 1,p(Ω) → W 1,p (Ω) 0 cannot be semi-Fredholm, we know that L = AϕA cannot be semi-Fredholm. (Ω) is not closed. Since this implies that ϕA : W 2,p(Ω) → W 1,p 0 5.3. Matrix-valued functions. One of our goals is to aid in the analysis of (16) Lu = div(Φ∇u), when the matrix-valued function Φ : ¯Ω → Cd×d is positive semi-definite for each x ∈ ¯Ω. With that end in mind, this section focuses on the multiplication operator u 7→ Φu where Φ ∈ C 1( ¯Ω; Cd×d) and u(x) ∈ Cd for almost every x ∈ Ω. As we will see shortly, the properties that were established for the multiplication operator u 7→ ϕu apply for the multiplication operator u 7→ Φu as well. In order for the operator L defined in (16) to be uniformly elliptic, the matrix Φ : ¯Ω → Cd×d must be uniformly positive definite. This section, like the ones before it, focus on the violation of this positivity assumption. Specifically, we assume Φ PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 15 is vanishing of order m (recall Definition 3.4). Another way to express this is as follows: for each fixed V ⊂⊂ Ω we have (17) inf x∈V inf v∈Cd ¯v · Φ(x)v ≥ cV v, where cV > 0 is the smallest eigenvalue of Φ(x) for x ∈ V . Moreover, the speed at which cV goes to zero is proportional to am where a = inf y∈∂V dist(y, ∂Ω). We take Lp(Ωd) to be the space of all measurable functions u = (u1, . . . , ud) such that ui ∈ Lp(Ω) for i = 1, . . . , d. The norm of Lp(Ωd) is taken to be In other words, kukLp(Ωd) := kuikLp(Ω). d Xi=1 Lp(Ωd) = Lp(Ω) × · · · × Lp(Ω). d copies {z } The space W k,p(Ωd) is defined as the subset of u = (u1, . . . , ud) ∈ Lp(Ωd) where ui ∈ W k,p(Ω) for each i = 1, . . . , d. We assume Φ : Lp(Ωd) → W k,p(Ωd) for some k ∈ Z+. Theorem 5.11. Let Ω ⊂ Rd be open and bounded with C 0,1 boundary and assume Φ ∈ C k( ¯Ω; Cd×d) is vanishing of order m. Then the range of the operator u 7→ Φu is closed in W k,p(Ωd) whenever k ≥ m and is not closed when k < m. Proof. We know there exists U ∈ C k( ¯Ω; Cd×d) and D ∈ C k( ¯Ω; Rd×d) such that Φ = UDU∗, where D = diag(ϕ1, . . . , ϕd) and ϕi ∈ C k( ¯Ω) for each i = 1, . . . , d. Since U is one-to-one and onto Lp(Ωd), it suffices to prove the claim for the operator D. Now, by our definition of W k,p(Ωd), it must be the case that R(D) is closed in W k,p(Ωd) if and only if the multiplication operators u 7→ ϕiu have closed range in W k,p(Ω) for each i = 1, . . . , d. With this in mind, we simply apply Theorem 5.9 for each diagonal function ϕi, yielding the desired conclusion. (cid:3) Theorem 5.12. Let Ω ⊂ Rd be a bounded open set and let Φ ∈ C k( ¯Ω; Cd×d) be vanishing of order m. Assume k ∈ N is such that and that the boundary ∂Ω is C k. If u ∈ Lp(Ωd) and Φu ∈ W k+m,p(Ωd) then k > d p + (m − 1)(cid:4) d p(cid:5), (18) and there exists a c > 0, independent of u, such that p(cid:5), u ∈ W κ,p(Ωd), where κ := k − m(cid:4) d kukW κ,p(Ωd) ≤ ckΦukW k+m,p(Ωd). (19) Proof. We know there exists U ∈ C k( ¯Ω; Cd×d) and D ∈ C k( ¯Ω; Rd×d) such that Φ = UDU∗, where D = diag(ϕ1, . . . , ϕd) and ϕi ∈ C k( ¯Ω) for each i = 1, . . . , d. As in the above theorem, it suffices to prove the claim for the operator D. Given u = (u1, . . . , ud) ∈ Lp(Ωd) and Du ∈ W k+m,p(Ωd) then we have ui ∈ If ϕi > 0 on ¯Ω then ui ∈ Lp(Ω) and ϕiui ∈ W k+m,p(Ω) for each i = 1, . . . , d. W k+m,p(Ω), and if ϕi is vanishing of order j ≤ m we apply Theorem 5.7 to get ui ∈ W κj,p(Ω), p(cid:5). where κj := k − j(cid:4) d In either case, ui ∈ W κ,p(Ω) for each i = 1, . . . , d. The proof of inequality (19) mirrors that of Remark 5.8 and is omitted. (cid:3) 16 DANIEL JORDON 6. Differential operators composed with vanishing operators In this section we examine differential operators that are composed with van- ishing operators. By 'differential operator' we mean any operator that is closed on the subspace W k,p(Ω) ⊂ Lp(Ω), k ≥ 1, and maps to either Lp(Ω) or Lp(Ωd). We pay particular attention to linear differential operators that are Fredholm or semi-Fredholm. Many of these results use compactness of nested Sobolev spaces. 6.1. Compactness. One of the salient features of the Sobolev space W k,p(Ω) is its compactness relationship with the ambient space Lp(Ω). In this section, we explore the implications of compactness on the composition of differential operators with vanishing functions. We start with the following general result for Fredholm operators. Theorem 6.1. Let X and Y be Banach spaces. If A ∈ F(X, Y ) then D(A) ⊂⊂ X if and only if its pseudo-inverse is compact from Y to X. Proof. Since A is closed we may equip D(A) with the A-norm and convert it into a Banach space, which we call W . Claim 1: If A ∈ F(X, Y ) with W ⊂⊂ X then the pseudo-inverse of A is compact from Y to X. Given that A is Fredholm from X to Y , we know it is also Fredholm from W to Y . Let A0 denote the pseudo-inverse of A : W → Y , and let ι : W → X denote the inclusion map from W to X. The assumption that W ⊂⊂ X tells us that ι is compact, which implies ι A0 : Y → X is compact as well since A0 ∈ B(Y, W ). If we let A0 denote the pseudo-inverse of A : X → Y then we see that A0 = ι A0, so A0 is compact. Claim 2: If A ∈ F(X, Y ) and the pseudo-inverse of A is compact from Y to X then W ⊂⊂ X. We are told A is Fredholm, so we know N(A) is finite dimensional and that there exists a closed subspace X0 ⊂ X such that X = X0 ⊕ N(A). Suppose {xn} ⊂ D(A) with kxnkD(A) ≤ c. Then for each n we have the decomposition xn = an + bn where an ∈ X0 and bn ∈ N(A). Since kankD(A) ≤ c for all n, {Aan} is a bounded sequence in Y . Given that A0, the pseudo-inverse of A, is compact from Y to X, there exists a subsequence of {an} = {A0Aan} that is convergent in X. Also, since {bn} is bounded and N(A) is finite dimensional, every subsequence of {bn} has a further subsequence that is convergent. Thus, we can find a subsequence of {bn} along the convergent subsequence of {an} that is convergent. With this we can conclude that {xn} = {an + bn} contains a convergent subsequence in X. This proves the claim and completes the proof of the theorem. (cid:3) As a consequence of Theorem 6.1 we have the following. Theorem 6.2. Let X, Y , and Z be Banach spaces. Suppose A ∈ F(X, Y ) where D(A) ⊂⊂ X. If B ∈ C(Y, Z) but is not semi-Fredholm then BA is not closed on its natural domain. Proof. The proof is by contradiction. Assume BA is closed on its natural domain, D(BA) = {x ∈ D(A) : Ax ∈ D(B)}. PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 17 Claim 1: There exists a c > 0 such that kxkD(A) ≤ ckxkD(BA) holds for all x ∈ D(BA). If BA was closed on D(BA) then D(BA) would be a Banach space with the BA- norm. Since A is Fredholm it must be closed on its domain D(A), so D(A) is also a Banach space with the A-norm. We know that D(BA) ⊂ D(A), so by Lemma 4.2 there exists a c > 0 such that kxkD(A) ≤ ckxkD(BA) whenever x ∈ D(BA). Claim 2: There exists a sequence that converges in D(BA) but does not converge in D(A). Since B is not semi-Fredholm, there exists a bounded sequence {xn} ⊂ D(B) such that {Bxn} converges but {xn} has no convergent subsequence. Given that A is Fredholm we know A has a pseudo-inverse, which we denote by A0. We then set yn = A0xn and notice that Ayn = AA0xn = (I − K)xn, where K is a projection into some finite dimensional subspace of Y . Since {xn} has no convergent subsequence and K projects to a finite dimensional subspace, {Kxn} is eventually zero. Thus, {Ayn} has no convergent subsequence and {BAyn} converges. Since D(A) ⊂⊂ X, we know that A0 is compact by Theorem 6.1 so {yn} has a convergent subsequence in X (which, after relabeling, we take to be the entire sequence). Using claim 1, we have kym − ynkD(A) = kym − ynkX + kAym − AynkY ≤ ckym − ynkD(BA) = ckym − ynkX + ckBAym − BAynkZ. We have established that {BAyn} and {yn} converge in Z and X respectively, so {yn} is convergent in D(BA). But we know that {Ayn} does not converge in Y , hence {yn} cannot converge in D(A). This is the desired contradiction. (cid:3) If ϕ ∈ C 1( ¯Ω) is simply vanishing, then by Theorem 5.9 the range of the multipli- cation operator u 7→ ϕu is not closed in Lp(Ω). Thus, ϕ cannot be semi-Fredholm from Lp(Ω) to Lp(Ω). The above theorem then says ϕmA is never closed on its natural domain for any m > 0. However, we can partially make up for this loss by showing that ϕmA is closable. Before we begin we will need a few more tools from classical functional analysis. The adjoint operator of A, denoted A∗, is a map from the dual Y ∗ to X ∗, where (20) A∗y∗(x) = y∗(Ax), for all x ∈ D(A), and some y∗ ∈ Y ∗. The set of appropriate y∗ ∈ Y ∗ for which (20) holds is D(A∗). The following two results are needed. Lemma 6.3 ([11] Lemma 131, p. 137). Let X be a normed vector space. Suppose that a sequence {xn} ⊂ X is bounded, and lim l∗(xn) = l∗(x) for each l∗ ∈ V where V is a dense subset of X ∗. Then the sequence {xn} converges to x weakly. Theorem 6.4 ([17], Theorems 7.35 and 7.36, p. 178). Let X, Y , and Z be Banach spaces, and assume that A ∈ C(X, Y ) where R(A) is closed in Y with finite co- dimension. Let B be a densely defined operator from Y to Z. Then (BA)∗ exists, (BA)∗ = A∗B∗, and both (BA)∗ and BA are densely defined. With the above lemma and theorem, we can conclude the following. 18 DANIEL JORDON Theorem 6.5. Let X, Y , and Z be Banach spaces. Suppose A ∈ F(X, Y ) and B is a densely defined linear operator from Y to Z. Then BA is closable. Proof. Since A is Fredholm, the domains of BA and (BA)∗ are both dense, by Theorem 6.4. Suppose we have a sequence {xn} ⊂ D(BA) where xn → 0 and BAxn → z as n → ∞. Then for each w∗ ∈ D((BA)∗), w∗(z) = lim n→∞ w∗(BAxn) = lim n→∞ (BA)∗w∗(xn) = 0. Since D((BA)∗) is dense in X ∗ and BAxn is bounded, this implies that BAxn → 0 weakly as n → ∞, by Lemma 6.3. We know that weak limits must coincide with strong limits so z = 0. Thus BA is closable. (cid:3) Theorems 6.2 and 6.5 yield the following result. Theorem 6.6. Let Ω ⊂ Rd be an open and bounded set with C 0,1 boundary. Let ϕ ∈ C 1( ¯Ω) be simply vanishing on ∂Ω. If A : Lp(Ω) → Lp(Ω) is Fredholm with D(A) ⊂ W 1,p(Ω), then ϕmA, for m ≥ 1, is not closed on its natural domain but is closable. Proof. Theorem 5.9 tells us that the range of the multiplication operator u 7→ ϕmu is not closed in Lp(Ω), so ϕm is not semi-Fredholm from Lp(Ω) to Lp(Ω). Since Ω is bounded with C 0,1 boundary, this implies D(A) ⊂ W 1,p(Ω) ⊂⊂ Lp(Ω) by the Rellich-Kondrachov Theorem. Theorem 6.2 then tells us that ϕmA is not closed on its natural domain, but by Theorem 6.5 ϕmA is closable. (cid:3) Our next result makes use of the following theorem. Theorem 6.7 ([17] Theorem 7.22, p. 170). Let X and Y be Banach spaces. If A ∈ F(X, Y ) and Y is reflexive, then A∗ ∈ F(Y ∗, X ∗) and ind(A∗) = −ind(A). We now have the following: Theorem 6.8. Let X and Y be Banach spaces where Y is also reflexive. If A ∈ F(X, Y ) with D(A) ⊂⊂ X then D(A∗) ⊂⊂ Y ∗. Proof. By Theorem 6.7 we know that A∗ ∈ F(Y ∗, X ∗). Also, if x∗ ∈ D(A∗) then (21) A∗ 0A∗x∗(x) = A∗x∗(A0x) = x∗(AA0x) = x∗(x − K2x) = (I − K ∗ 2 )x∗(x). Thus, from (21) and Theorems 6.4 and 3.5 we conclude: A∗A∗ 0 = (A0A)∗ = I − K ∗ 1 , which implies A∗ 0 is the pseudo-inverse of A∗. Since A ∈ F(X, Y ), and D(A) ⊂⊂ X, we know that the pseudo-inverse A0 is compact from Y to X by Theorem 6.1. 0 is compact from X ∗ to Y ∗. If we Given A0 is compact from Y to X, we know A∗ then apply Theorem 6.1 to A∗ we get D(A∗) ⊂⊂ Y ∗. (cid:3) 0A∗ = (AA0)∗ = I − K ∗ 2 , A∗ 6.2. The Spectrum. Let X be a Banach space and A be a densely defined linear operator from X to X. The resolvent set of A, denoted ρ(A), is the set of all λ ∈ C such that A − λ has a bounded inverse. The complement of ρ(A) in C is called the spectrum of A, and is denoted as σ(A). We let σp(A) denote the point spectrum of A: σp(A) := {λ ∈ C : Ax = λx for some x ∈ D(A)}. PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 19 We define the essential spectrum as: σess(A) = \K∈K(X) σ(A + K), where K(X) is the set of all compact operators on X. This set is sometimes referred to as Schechter's essential spectrum. Another useful characterization of the essential spectrum is given in the theorem below. Theorem 6.9 ([17] Theorem 7.27, p. 172). Let X be a Banach space and assume A ∈ C(X). Then λ 6∈ σess(A) if and only if A − λ ∈ F(X) and ind(A − λ) = 0. We will also need the notion of relatively compact operators. If A ∈ C(X, Y ), an operator B : X → Z is called compact relative to A if B is compact from the Banach space D(A) to Z. The following theorem is a more robust version of the Fredholm Alternative since it is stated for any Fredholm operator A (not just the identity operator) and the perturbations to A can be any operator that is compact relative to A. A proof of this theorem can be found in [10, p. 281] Theorem 1, or [17, p. 162] Theorem 7.10. Theorem 6.10 ([10] Theorem 1, p. 281). If A ∈ F(X, Y ) and B is compact relative to A then A + B ∈ F(X, Y ) and ind(A + B) = ind(A). Remark 6.11. With the help of Theorems 6.9 and 6.10, we see that the essen- tial spectrum is invariant under relatively compact perturbations. Given that the identity map on X is compact relative to A ∈ C(X) whenever D(A) ⊂⊂ X, we know that either σess(A) = ∅ or σess(A) = C. This fact makes calculating the essential spectrum of differential operators relatively easy whenever we can use the Rellich-Kondrachov Theorem. Theorem 6.12. Let X be a Banach space, A ∈ C(X), and B ∈ B(X) be a one- to-one operator where D(A) 6⊂ R(B). If D(AB) is dense in X with D(AB) ⊂⊂ X and ρ(A) is nonempty then σess(AB) = C. Proof. The proof is by contradiction. Suppose σess(AB) 6= C. As mentioned in Remark 6.11, σess(AB) 6= C implies σess(AB) = ∅ since D(AB) ⊂⊂ X. This implies AB ∈ F(X) and ind(AB) = 0 by Theorem 6.9. Since D(AB) ⊂⊂ X, any bounded operator on X is compact relative to AB. In particular, B is compact relative to AB. By Theorem 6.10, Fredholm operators and their indices are invariant under relatively compact perturbations. Thus, for any η ∈ C we have AB − ηB ∈ F(X) and ind(AB) = ind(AB − ηB) = 0. Now, if η ∈ ρ(A) then N(A − η) = {0} and R(A − η) = X, implying A − η maps D(A) to X. But since B does not map to all of D(A) we get R((A − η)B) 6= X. In other words co-dim R(AB − ηB) > 0. Recall that N(B) = {0} by assumption, and N(A − η) = {0} when η ∈ ρ(A), which gives N(cid:0)(A − η)B(cid:1) = {0}. Thus, ind(AB) = ind(AB − ηB) = dim N(AB − ηB) − co-dim R(AB − ηB) < 0, which contradicts the fact that ind(AB) = 0. (cid:3) 20 DANIEL JORDON The proof of the above theorem yields the following corollary. Corollary 6.13. Let X be a Banach space, A ∈ C(X), and B ∈ B(X) be a one-to- one operator where the range of B is not the entirety of D(A). If AB is Fredholm with D(AB) ⊂⊂ X and ρ(A) is nonempty then ind(AB) < 0. In some cases, we are concerned with the adjoint operator ϕmA∗ instead Aϕm. For example, one might be interested in the spectral properties of the operator L given by Lu = −(1 − x2)uxx on Ω = (−1, 1). This operator is the adjoint of Aϕu = −((1 − x2)u)xx where A is the Laplacian on L2(Ω) and ϕ(x) = (1 − x2). In this case, we have the following theorem. Theorem 6.14. Let Ω ⊂ Rd be open and bounded with C 0,1 boundary, m, k ∈ N with m < k, and let A be densely defined on Lp(Ω). Assume (a) A is closed on Lp(Ω) and D(A) ⊂ W k,p(Ω). (b) There exists a u ∈ D(A) such that u /∈ W m,p (c) The resolvent set ρ(A) is non-empty. 0 (Ω) ∩ W k,p(Ω). If ϕ ∈ C k( ¯Ω) is simply vanishing then σess(Aϕm) = σess(ϕmA∗) = C. Moreover, if either Aϕm or ϕmA∗ is Fredholm then σp(ϕmA∗) = C. Remark 6.15. By Theorem 6.6, ϕmA∗ is never closed on its natural domain when A is Fredholm. Thus, we examine its closure since statements about the essential spectrum are uninformative for operators that are not closed. Proof. As usual, let ϕm denote the multiplication operator u 7→ ϕmu from Lp(Ω) to W k,p(Ω). The proof is broken into 4 claims. Claim 1: Aϕm is closed on its natural domain and D(Aϕm) ⊂⊂ Lp(Ω). Given ∂Ω is C 0,1, we see that D(A) ⊂ W k,p(Ω) ⊂⊂ Lp(Ω) by the Rellich- Kondrachov Theorem. Since A is closed on D(A), A must be semi-Fredholm on Lp(Ω) by Theorem 4.3. With the assumption that m < k, Lemma 5.3 implies D(ϕm) ⊂⊂ Lp(Ω), and applying Theorem 4.3 yields ϕm is semi-Fredholm from Lp(Ω) to W k,p(Ω). Since both A and ϕm are semi-Fredholm, Aϕm is closed on its natural domain. The fact that completes the proof of the claim. D(Aϕm) ⊂ D(ϕm) ⊂⊂ Lp(Ω) Claim 2: σess(Aϕm) = C Applying Lemma 5.2 to ϕm shows that R(ϕm) = W m,p (Ω) ∩ W k,p(Ω). This and assumption (b) implies ϕm does not map to all of D(A). Given claim 1, ρ(A) is non-empty, and the multiplication operator ϕm : Lp(Ω) → Lp(Ω) is bounded, we may apply Theorem 6.12 to prove σess(Aϕm) = C. 0 Claim 3: If Aϕm is not Fredholm then σess(ϕmA∗) = C. Fix λ ∈ C. Claim 1 implies the identity is compact relative to Aϕm. Since Aϕm is not Fredholm, Aϕm − λ cannot be Fredholm by Theorem 6.10. By Theorem 6.7, this implies ϕmA∗ − ¯λ is not Fredholm. Applying Theorem 6.9 to ϕmA∗ − ¯λ shows ¯λ ∈ σess(ϕmA∗). Noting that λ ∈ C was arbitrary completes the proof of the claim. PROPERTIES OF DIFFERENTIAL OPERATORS WITH VANISHING COEFFICIENTS 21 Claim 4: If Aϕm or ϕmA∗ is Fredholm then σess(ϕmA∗) = σp(ϕmA∗) = C. Given Lp(Ω) is reflexive, Aϕm is Fredholm if and only if ϕmA∗ is Fredholm by Theorem 6.7. Since ρ(A) is non-empty and D(Aϕm) ⊂⊂ Lp(Ω), Corollary 6.13 then implies ind(Aϕm) < 0. Thus, ind(ϕmA∗) = −ind(Aϕm) > 0. Moreover, by Theorem 6.8 and claim 1, we have that D(ϕmA∗) ⊂⊂ Lp(Ω), so for any λ ∈ C, we can conclude (22) ind(ϕmA∗ − λ) = ind(ϕmA∗) > 0. From (22) we have dim N(ϕmA∗ − λ) > 0, which implies λ ∈ σp(ϕmA∗). By Theorem 6.9, (22) also implies λ ∈ σess(ϕmA∗). Since λ ∈ C was arbitrary, we are done. (cid:3) 7. Acknowledgements Parts of this paper have grown out of work that I completed in my thesis. I would like to thank my thesis advisor, Jay Douglas Wright, for his teachings and innumerable insightful discussions. References [1] Robert A. Adams and John J. F. Fournier, Sobolev spaces, Second, Pure and Applied Math- ematics (Amsterdam), vol. 140, Elsevier/Academic Press, Amsterdam, 2003. MR2424078 (2009e:46025) [2] C. V. Coffman, M. M. Marcus, and V. J. Mizel, Boundary value problems for nonuniformly elliptic equations with measurable coefficients, Ann. Mat. Pura Appl. (4) 110 (1976), 223– 245. MR0466946 (57 #6819) [3] Fran¸coise Demengel and Gilbert Demengel, Functional spaces for the theory of elliptic partial differential equations, Universitext, Springer, London, 2012. Translated from the 2007 French original by Reinie Ern´e. MR2895178 [4] Lawrence C. Evans, Partial differential equations, Graduate Studies in Mathematics, vol. 19, American Mathematical Society, Providence, RI, 1998. MR1625845 (99e:35001) [5] Bruno Franchi, Raul Serapioni, and Francesco Serra Cassano, Irregular solutions of linear de- generate elliptic equations, Potential Anal. 9 (1998), no. 3, 201–216. MR1666899 (99m:35087) [6] David Gilbarg and Neil S. Trudinger, Elliptic partial differential equations of second order, Springer-Verlag, Berlin-New York, 1977. Grundlehren der Mathematischen Wissenschaften, Vol. 224. MR0473443 (57 #13109) [7] Piotr Haj lasz, Pointwise Hardy inequalities, Proc. Amer. Math. Soc. 127 (1999), no. 2, 417– 423. MR1458875 (99c:46028) [8] Qing Han and Fanghua Lin, Elliptic partial differential equations, Second, Courant Lecture Notes in Mathematics, vol. 1, Courant Institute of Mathematical Sciences, New York; Amer- ican Mathematical Society, Providence, RI, 2011. MR2777537 (2012c:35077) [9] Daniel Jordon, Spectral properties of differential operators with vanishing coefficients, Pro- Quest LLC, Ann Arbor, MI, 2013. Thesis (Ph.D.)–Drexel University. MR3187296 [10] Tosio Kato, Perturbation theory for nullity, deficiency and other quantities of linear opera- tors, J. Analyse Math. 6 (1958), 261–322. MR0107819 (21 #6541) [11] , Perturbation theory for linear operators, Classics in Mathematics, Springer-Verlag, Berlin, 1995. Reprint of the 1980 edition. MR1335452 (96a:47025) [12] N. V. Krylov, Lectures on elliptic and parabolic equations in Holder spaces, Graduate Studies in Mathematics, vol. 12, American Mathematical Society, Providence, RI, 1996. MR1406091 (97i:35001) [13] Juha Lehrback, Weighted Hardy inequalities beyond Lipschitz domains, Proc. Amer. Math. Soc. 142 (2014), no. 5, 1705–1715. MR3168477 22 DANIEL JORDON [14] Vladimir Maz'ya, Sobolev spaces with applications to elliptic partial differential equations, augmented, Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences], vol. 342, Springer, Heidelberg, 2011. MR2777530 (2012a:46056) [15] M. K. V. Murthy and G. Stampacchia, Boundary value problems for some degenerate-elliptic operators, Ann. Mat. Pura Appl. (4) 80 (1968), 1–122. MR0249828 (40 #3069) [16] Martin Schechter, Remarks on a previous paper: "On the essential spectrum of an arbitrary operator. I", (J. Math. Anal. Appl. 13 (1966), 205–215), J. Math. Anal. Appl. 54 (1976), no. 1, 138–141. MR0435900 (55 #8851) [17] , Principles of functional analysis, Second, Graduate Studies in Mathematics, vol. 36, American Mathematical Society, Providence, RI, 2002. MR1861991 (2002j:46001) [18] Neil S. Trudinger, On the regularity of generalized solutions of linear, non-uniformly elliptic equations, Arch. Rational Mech. Anal. 42 (1971), 50–62. MR0344656 (49 #9395) [19] [20] , Maximum principles for linear, non-uniformly elliptic operators with measurable coefficients, Math. Z. 156 (1977), no. 3, 291–301. MR0470460 (57 #10214) , Harnack inequalities for nonuniformly elliptic divergence structure equations, Invent. Math. 64 (1981), no. 3, 517–531. MR632988 (82m:35051) [21] Andreas Wannebo, Hardy inequalities, Proc. Amer. Math. Soc. 109 (1990), no. 1, 85–95. MR1010807 (90h:26025) E-mail address: [email protected] Transportation Research Institute, University of Michigan, Ann Arbor, MI 48109
1301.2539
2
1301
2013-10-30T09:22:27
Sobolev spaces on Riemannian manifolds with bounded geometry: General coordinates and traces
[ "math.FA", "math.DG" ]
We study fractional Sobolev and Besov spaces on noncompact Riemannian manifolds with bounded geometry. Usually, these spaces are defined via geodesic normal coordinates which, depending on the problem at hand, may often not be the best choice. We consider a more general definition subject to different local coordinates and give sufficient conditions on the corresponding coordinates resulting in equivalent norms. Our main application is the computation of traces on submanifolds with the help of Fermi coordinates. Our results also hold for corresponding spaces defined on vector bundles of bounded geometry and, moreover, can be generalized to Triebel-Lizorkin spaces on manifolds, improving a result by Skrzypczak.
math.FA
math
SOBOLEV SPACES ON RIEMANNIAN MANIFOLDS WITH BOUNDED GEOMETRY: GENERAL COORDINATES AND TRACES NADINE GROSSE AND CORNELIA SCHNEIDER Abstract. We study fractional Sobolev and Besov spaces on noncompact Riemannian manifolds with bounded geometry. Usually, these spaces are defined via geodesic normal coordinates which, depending on the problem at hand, may often not be the best choice. We consider a more general definition sub ject to different local coordinates and give sufficient conditions on the corresponding coordinates resulting in equivalent norms. Our main application is the computation of traces on submanifolds with the help of Fermi coordinates. Our results also hold for corresponding spaces defined on vector bundles of bounded geometry and, moreover, can be generalized to Triebel-Lizorkin spaces on manifolds, improving [Skr90]. 1. Introduction The main aim of this paper is to consider fractional Sobolev spaces on noncompact Riemannian manifolds, equivalent characterizations of these spaces and their traces on submanifolds. We address the problem to what extend results from classical analysis on Euclidean space carry over to the setting of Riemannian mani- folds – without making any unnecessary assumptions about the manifold. In particular, we will be interested in noncompact manifolds since the compact case presents no difficulties and is well understood. Let (M , g ) denote an n-dimensional, complete, and noncompact Riemannian manifold with Riemannian metric g . Fractional Sobolev spaces on manifolds H s p (M ), s ∈ R, 1 < p < ∞, can be defined similar to p (Rn ), usually characterized via corresponding Euclidean spaces H s H s p = (Id − ∆)−s/2Lp , by replacing the Euclidean Laplacian ∆ with the Laplace-Beltrami operator on (M , g ) and using an auxil- iary parameter ρ, see Section 3.1. The spaces H s p (M ) were introduced and studied in detail in [Str83] and generalize in a natural way classical Sobolev spaces on manifolds, W k p (M ), which contain all Lp functions on M having bounded covariant derivatives up to order k ∈ N, cf. [Aub76, Aub82]. To avoid any confusion, let us emphasize that in this article we study exactly these fractional Sobolev spaces H s p (M ) defined by means of powers of ∆. But we shall use an alternative characterization of these spaces on manifolds with bounded geometry as definition – having in mind the proof of our main theorem. To be more precise, on manifolds with bounded geometry, see Definition 18, one can alternatively define frac- p (M ) via localization and pull-back onto Rn , by using geodesic normal coordinates tional Sobolev spaces H s and corresponding fractional Sobolev spaces on Rn , cf. [Tri92, Sections 7.2.2, 7.4.5] and also [Skr98, Defi- nition 1]. Unfortunately, for some applications the choice of geodesic normal coordinates is not convenient, which is why we do not wish to restrict ourselves to these coordinates only. The main application we have in mind are traces on submanifolds N of M . But also for manifolds with symmetries, product manifolds or warped products, geodesic normal coordinates may not be the first and natural choice and one is interested in coordinates better suited to the problem at hand. Date : October 31, 2013. 2010 Mathematics Subject Classification. 46E35, 53C20. Key words and phrases. Sobolev spaces, Riemannian manifolds, bounded geometry, Fermi coordinates, traces, vector bun- dles, Besov spaces, Triebel-Lizorkin spaces. 1 kf kH s,T p Therefore, we introduce in Definition 11 Sobolev spaces H s,T (M ) in a more general way, containing all those p complex-valued distributions f on M such that := Xα∈I p (Rn )!1/p k(hαf ) ◦ καkp H s is finite, where T = (Uα , κα , hα )α∈I denotes a trivialization of M consisting of a uniformly locally finite covering Uα , local coordinates κα : Vα ⊂ Rn → Uα ⊂ M (not necessarily geodesic normal coordinates) and a subordinate partition of unity hα . Of course, the case of local coordinates κα being geodesic normal coordinates is covered but we can choose from a larger set of trivializations. Clearly, we are not interested in all T but merely the so called admissible trivializations T , cf. Definition 12, yielding the coincidence (M ) = H s H s,T p (M ), p (1) cf. Theorem 14. As pointed out earlier, our main applications in mind are Trace Theorems. In [Skr90, Theorem 1], traces on manifolds were studied using the Sobolev norm (1) with geodesic normal coordinates. Since these coordinates in general do not take into account the structure of the underlying submanifold where the trace is taken, one is limited to so-called geodesic submanifolds. This is highly restrictive, since geodesic submanifolds are very exceptional. Choosing coordinates that are more adapted to the situation will immediately enable us to compute the trace on a much larger class of submanifolds. In particular, we consider Riemannian manifolds (M , g ) with submanifolds N such that (M , N ) is of bounded geometry, see Definition 18, i.e., (M , g ) is of bounded geometry, the mean curvature of N and its covariant derivatives are uniformly bounded, the injectivity radius of (N , gN ) is positive and there is a uniform collar of N . The coordinates of choice for proving Trace Theorems are Fermi coordinates, introduced in Definition 20. We show in Theorem 26 that for a certain cover with Fermi coordinates there is a subordinated partition of unity such that the resulting trivialization is admissible. The main Trace Theorem itself is stated in Theorem 27, where we prove that if M is a manifold of dimension n ≥ 2, N a submanifold of dimension k < n, and (M , N ) of bounded geometry, we have for s > n−k p , TrN H s p (M ) = B s− n−k p p,p (N ). (2) i.e., there is a linear, bounded and surjective trace operator TrN with a linear and bounded right inverse ExM from the trace space into the original space such that TrN ◦ ExM = Id, where Id denotes the identity on operator N . The spaces on the right hand side of (2) are Besov spaces obtained via real interpolation of the spaces H s p , cf. Remark 17. When just asking for TrN to be linear and bounded, one can reduce the assumptions on (M , N ) further by replacing the existence of a collar of N with a uniform local collar, cf. Remark 33. We believe that the method presented in this article is very well suited to tackle the trace problem on manifolds. One could also think of computing traces using atomic decompositions of the spaces H s p (M ) as established in [Skr98], which is often done when dealing with traces on hyperplanes of Rn or on domains. But on (sub-)manifolds it should be complicated (if not impossible) to obtain a linear and continuous exten- sion operator from the trace space into the source space – which by our method follows immediately from corresponding results on Rn . In Section 5, we establish analogous results for vector bundles of bounded geometry. An application of our trace result for vector bundles, Theorem 47, may be found in [GN12], where the authors classify boundary value problems of the Dirac operator on spinC bundles of bounded geometry, deal with the existence of a solution, and obtain some spectral estimates for the Dirac operator on hypersurfaces of bounded geometry. As another application of our general coordinates spaces with symmetries are considered in Section 6.1. We restrict ourselves to the straight forward case where the symmetry group is discrete and obtain a general- ization of a theorem from [Tri83, Section 9.2.1], where the author characterizes Sobolev spaces on the tori Tn := Rn/Zn via weighted Sobolev spaces on Rn containing Zn periodic distributions only. 2 (M ), s ∈ R, 1 < p < ∞, Finally, in Section 6.2 we deal with the larger scale of Triebel-Lizorkin spaces F s,T p,q (M ), s ∈ R, 0 < p < ∞, 0 < q ≤ ∞ or p = q = ∞, linked with fractional Sobolev spaces via F s,T p,2 (M ) = H s,T p and the general scale of Besov spaces B s,T p,q (M ), s ∈ R, 0 < p, q ≤ ∞ defined via real interpolation of the spaces F s,T p,q (M ), cf. Definition 56. We will show that an admissible trivialization T again guarantees coincidence with the corresponding spaces F s p,q (M ), B s p,q (M ) – obtained from choosing geodesic normal coordinates, cf. [Tri92, Sections 7.2, 7.3] – and that trace results from Euclidean space carry over to our setting of submanifolds N of M , where (M , N ) is of bounded geometry. In particular, if now − 1(cid:19) , > k max (cid:18)0, s− n−k s− n−k Tr F s Tr B s p p p,q (M ) = B (N ) and (N ), p,q (M ) = B p,p p,q cf. Theorem 59. The restriction (3) is natural and best possible also in the Euclidean case. n − k p we have s − 1 p (3) Acknowledgement. We are grateful to Sergei V. Ivanov who kindly answered our question on mathover- flow concerning the equivalence of different characterizations on manifolds of bounded geometry. Moreover, we thank Hans Triebel for helpful discussions on the sub ject. The second author thanks the University of Leipzig for the hospitality and support during a short term visit in Leipzig. 2. Preliminaries and notations General notations. Let N be the collection of all natural numbers, and let N0 = N ∪ {0}. Let Rn be the n-dimensional Euclidean space, n ∈ N, C the complex plane, and let Bn r denote the ball in Rn with center 0 and radius r (sometimes simply denoted by Br if there is no danger of confusion). Moreover, index sets are always assumed to be countable, and we use the Einstein sum convention. Let the standard coordinates on Rn be denoted by x = (x1 , x2 , . . . , xn ). The partial derivative oper- ators in direction of the coordinates are denoted by ∂i = ∂ /∂ xi for 1 ≤ i ≤ n. The set of multi- indices a = (a1 , . . . , an), ai ∈ N0 , i = 1, . . . , n, is denoted by Nn 0 , and we shall use the common notation ∂ a f (∂x1 )a1 ···(∂xn )an , where f is a function on Rn . As usual, let a = a1 + · · · + an be the Daf = ∂ a1 1 ...∂ an n f = order of the derivative Daf . Moreover, we put xa = (x1 )a1 · · · (xn )an . For a real number a, let a+ := max(a, 0), and let [a] denote its integer part. For p ∈ (0, ∞], the number p′ is defined by 1/p′ := (1 − 1/p)+ with the convention that 1/∞ = 0. All unimportant positive constants will be denoted by c, occasionally with subscripts. For two non-negative expressions (i.e., functions or functionals) A, B , the symbol A . B (or A & B ) means that A ≤ c B (or c A ≥ B ) for a suitable constant c. If A . B and A & B , we write A ∼ B and say that A and B are equivalent. Given two (quasi-) Banach spaces X and Y , we write X ֒→ Y if X ⊂ Y and the natural embedding of X into Y is continuous. Function spaces on Rn . Lp (Rn ), with 0 < p ≤ ∞, stands for the usual quasi-Banach space with respect to the Lebesgue measure, quasi-normed by kf kLp(Rn ) := (cid:18)ZRn f (x)p dx(cid:19) 1 p with the usual modification if p = ∞. For p ≥ 1, Lp (Rn ) is even a Banach space. Let D(Rn ) denote the space of smooth functions with compact support, and let D ′ (Rn ) denote the corresponding distribution space. By S (Rn ) we denote the Schwartz space of all complex-valued rapidly decreasing infinitely differentiable functions on Rn and by S ′ (Rn ) the dual space of all tempered distributions on Rn . For a rigorous definition of the Schwartz space and ’rapidly decreasing’ we refer to [Tri83, Section 1.2.1]. For f ∈ S ′ (Rn ) we denote by bf the Fourier transform of f and by f ∨ the inverse Fourier transform of f . p (Rn ) contains all f ∈ S ′ (Rn ) with Let s ∈ R and 1 < p < ∞. Then the (fractional) Sobolev space H s (cid:0)(1 + ξ 2 )s/2 bf (cid:1)∨ ∈ Lp (Rn ), ξ ∈ Rn , 3 [Tri92, Section 1.3.2]. In particular, for k ∈ N0 , these spaces coincide with the classical Sobolev spaces cf. W k p (Rn ), usually normed by p (Rn ) = W k H k p (Rn ), H 0 p (Rn ) = Lp (Rn ), i.e., Lp(Rn ) p (Rn ) =  Xa≤k kDaf kp kf kW k Furthermore, Besov spaces B s p,p (Rn ) can be defined via interpolation of Sobolev spaces. In particular, let (·, ·)Θ,p stand for the real interpolation method, cf. [Tri92, Section 1.6.2]. Then for s0 , s1 ∈ R, 1 < p < ∞, p,p (Rn ) := (cid:0)H s0 p (Rn )(cid:1)Θ,p and 0 < Θ < 1, we put B s p (Rn ), H s1 , where s = Θs0 + (1 − Θ)s1. Note that B s p,p (Rn ) does not depend on the choice of s0 , s1 , Θ. The following lemma about pointwise multipliers and diffeomorphisms may be found in [Tri92, Sections 4.2,4.3], where it was proven in a more general setting. 1/p . Lemma 1. Let s ∈ R and 1 < p < ∞. (i) Let f ∈ H s p (Rn ) and ϕ a smooth function on Rn such that for al l a with a ≤ [s] + 1 we have Daϕ ≤ Ca . Then there is a constant C only depending on s, p, n and Ca such that kϕf kH s p (Rn ) ≤ C kf kH s p (Rn ) . (ii) Let f ∈ H s p (Rn ) with supp f ⊂ U ⊂ Rn for U open and let κ : V ⊂ Rn → U ⊂ Rn be a diffeomorphism such that for al l a with a ≤ [s] + 1 we have Da κ ≤ Ca .Then there is a constant C only depending on s, p, n and Ca such that kf ◦ κkH s p (Rn ) ≤ C kf kH s p (Rn ) . Vector-valued function spaces on Rn . Let D(Rn , Fr ) be the space of compactly supported smooth functions on Rn with values in Fr where F stands for R or C and r ∈ N . Let D ′ (Rn , Fr ) denote the p (Rn ) from above, p (Rn , Fr ) is defined in correspondence with H s corresponding distribution space. Then, H s cf. [Triebel, Fractals and spectra, Section 15]. Moreover, Besov spaces B s p,p (Rn , Fr ) are defined as the spaces p,p (Rn , Fr ) := (cid:0)H s0 p (Rn , Fr )(cid:1)Θ,p B s p,p (Rn ) from above; B s p (Rn , Fr ), H s1 where (·, ·)Θ,p again denotes the real interpolation method with s0 , s1 ∈ R, 1 < p < ∞, and 0 < Θ < 1 with s = Θs0 + (1 − Θ)s1 . p (Rn ,Fr ) and (cid:16)Pr p (Rn ,F)(cid:17) 1 p i=1 kϕikp are equivalent where ϕ = (ϕ1 , . . . , ϕr ) ∈ Lemma 2. The norms kϕkH s H s H s p (Rn , Fr ). The analogous statement is true for Besov spaces. Proof. The equivalence for Sobolev spaces follows immediately from their definition. The corresponding result for Besov spaces can be found in [Gro12, Lemma 26]. (cid:3) Notations concerning manifolds. Before starting we want to make the following warning or excuse: For a differential geometer the notations may seem a little overloaded at first glance. Usually, when interested in equivalent norms, one merely suppresses diffeomorphisms as transition functions. This provides no problem when it is clear that all constants appearing are uniformly bounded – which is obvious for finitely many bounded charts (on closed manifolds) and also known for manifolds of bounded geometry with geodesic nor- mal coordinates. But here we work in a more general context where the aim is to find out which conditions the coordinates have to satisfy in order to ignore those diffeomorphisms in the sequel. This is precisely why we try to be more explicit in our notation. Let (M n , g ) be an n-dimensional complete manifold with Riemannian metric g . We denote the volume element on M with respect to the metric g by dvolg . For 1 < p < ∞ the Lp -norm of a compactly supported smooth function v ∈ D(M ) is given by kvkLp (M ) = (cid:0)RM v p dvolg (cid:1) 1 p . The set Lp (M ) is then the completion of D(M ) with respect to the Lp -norm. The space of distributions on M is denoted by D ′ (M ). A cover (Uα )α∈I of M is a collection of open subsets of Uα ⊂ M where α runs over an index set I . The cover is called locally finite if each Uα is intersected by at most finitely many Uβ . The cover is called uniformly locally finite if there exists a constant L > 0 such that each Uα is intersected by at most L sets Uβ . 4 A chart on Uα is given by local coordinates – a diffeomorphism κα : x = (x1 , . . . , xn ) ∈ Vα ⊂ Rn → κα (x) ∈ Uα . We will always assume our charts to be smooth. A collection A = (Uα , κα )α∈I is called an atlas of M . Moreover, a collection of smooth functions (hα )α∈I on M with and Xα is called a partition of unity subordinated to the cover (Uα )α∈I . The triple T := (Uα , κα , hα )α∈I is called a trivialization of the manifold M . supp hα ⊂ Uα , 0 ≤ hα ≤ 1 hα = 1 on M . (4) Using the standard Euclidean coordinates x = (x1 , . . . , xn ) on Vα ⊂ Rn , we introduce an orthonormal frame (eα i )1≤i≤n on T Uα by eα := (κα )∗ (∂i ). In case we talk about a fixed chart we will often leave out i the superscript α. Then, in those local coordinates the metric g is expressed via the matrix coefficients ij =( αΓk α = g (ei , ej ) and the corresponding Christoffel symbols Γk ij ) : Vα → R defined by gij ◦ κ−1 gij (= gα ij ) : ei ej = (Γk ij ◦ κ−1 α )ek where ∇M denotes the Levi-Civita connection of (M , g ). In Vα → R are defined by ∇M local coordinates, Γk ij = gkl (∂j gil + ∂i gj l − ∂l gij ) 1 2 where g ij is the inverse matrix of gij . If α, β ∈ I with Uα ∩ Uβ 6= ∅, we define the transition function α (Uα ∩ Uβ ) → κ−1 µαβ = κ−1 β ◦ κα : κ−1 β (Uα ∩ Uβ ). Then, αβ (x)gβ gα ij (x) = ∂iµk αβ (x)∂j µl (5) kl (µαβ (x)). Example 3 (Geodesic normal coordinates). Let (M n , g ) be a complete Riemannian manifold. Fix z ∈ M and let r > 0 be smaller than the injectivity radius of M . For v ∈ T ≤r z M := {w ∈ TzM gz (w, w) ≤ r2 }, we denote by cv : [−1, 1] → M the unique geodesic with cv (0) = z and cv (0) = v . Then, the exponential map z M → M is a diffeomorphism defined by expM : T ≤r expM z (v) := cv (1). Let S = {pα}α∈I be a set of points z in M such that (U geo := Br (pα ))α∈I covers M . For each pα we choose an orthonormal frame of Tpα M and α pα ◦λα : V geo := Bn α = expM α , κgeo call the resulting identification λα : Rn → Tpα M . Then, Ageo = (U geo r → α α )−1 at pα .) α equals the tangent map (dκgeo (Note that λ−1 U geo α )α∈I is an atlas of M – called geodesic atlas. Notations concerning vector bundles. Let E be a hermitian or Riemannian vector bundle over a Rie- mannian manifold (M n , g ) of rank r with fiber product h., .iE and connection ∇E : Γ(T M ) ⊗ Γ(E ) → Γ(E ). Here Γ always denotes the space of smooth sections of the corresponding vector bundle. We set F = R if E is a Riemannian vector bundle and F = C if E is hermitian. Let A = (Uα , κα : Vα → Uα )α∈I be an atlas of (M , g ) and let ζα : Uα × Fr → E Uα be local trivializations of E . Note that here ’trivialization’ has the usual meaning in connection with the ordinary definition of a vector bundle. We apologize that in lack of a better notion we also call T a trivialization but hope there will be no danger of confusion. We set ξα := ζα ◦ (κα × Id) : Vα × Fr → E Uα . We call AE = (Uα , κα , ξα )α∈I an atlas of E . In case we already start with a trivialization T = (Uα , κα , hα )α∈I on M , TE = (Uα , κα , ξα , hα )α∈I is called a trivialization of E . ∂ yρ (cid:17)1≤ρ≤r Let y = (y 1 , . . . , y r ) be standard coordinates on Fr and let (cid:16)∂ρ := ∂ be the corresponding local α (p), ∂ρ (cid:1) form a local frame of Ep for p ∈ Uα . As before, we suppress ρ (p)) := ξα (cid:0)κ−1 frame. Then, eρ (p)(= eα α in the notation if we talk about a fixed chart. In those local coordinates, the fiber product is represented by hρσ := heρ , eσ iE ◦ κα : Vα → F. Hence, if ϕ, ψ ∈ Γ(E Uα ) we have for ϕ = ϕρ eρ and ψ = ψσ eσ that α )ϕρ ¯ψσ , hϕ, ψiE = (hρσ ◦ κ−1 where ¯a denotes the complex conjugate of a. Let Christoffel symbols Γσ iρ : Uα → F for E be defined by α (cid:17) eσ , where ei = (κα )∗∂i . If the connection ∇E is metric, i.e., ei heσ , eρ iE = h∇E ei eρ = (cid:16) Γσ ∇E iρ ◦ κ−1 ei eσ , eρ iE + heσ , ∇ei eρ iE , we get ∂ihστ = Γρ iσ hτ ρ + Γρ (6) iτ hρσ . For all α, β ∈ I with Uα ∩ Uβ 6= ∅, transition functions µαβ : κ−1 α (Uα ∩ Uβ ) → GL(r, F) are defined by ξ−1 β ◦ ξα (x, u) = (µαβ (x), µαβ (x) · u). Here, GL(r, F) denotes the general linear group of F-valued r × r matrices. 5 Flows. Let x′ (t) = F (t, x(t)) be a system of ordinary differential equations with t ∈ R, x(t) ∈ Rn and F ∈ C∞ (R × Rn , Rn ). Let the solution of the initial value problem x′ (t) = F (t, x(t)) with x(0) = x0 ∈ Rn be de- noted by xx0 (t) and exist for 0 ≤ t ≤ t0 (x0 ). Then, the flow Φ : dom ⊂ R× Rn → Rn with dom ⊂ {(t, x) 0 ≤ t ≤ t0 (x)} is defined by Φ(t, x0 ) = xx0 (t). Higher order ODE’s x(d) (t) = F (t, x(t), . . . , x(d−1) (t)) can be trans- ferred back to first order systems by introducing auxiliary variables. The corresponding flow then obviously depends not only on x0 = x(0) but the initial values x(0), x′ (0), . . . , x(d−1) (0): Φ(t, x(0), . . . , x(d−1) (0)). Example 4 (Geodesic flow). Let (M n , g ) be a Riemannian manifold. Let z ∈ M , v ∈ TzM . Let κ : V ⊂ Rn → U ⊂ M be a chart around z . The corresponding coordinates on V are denoted by x = (x1 , . . . , xn ). We consider the geodesic equation in coordinates: xk = −Γk ij xi xj with initial values x(0) = κ−1 (z ) ∈ Rn and x′ (0) = κ∗ (v)(= dκ−1 (v)). Here Γk ij are the Christoffel symbols with respect to the coordinates given by κ. Let x(t) be the unique solution and Φ(t, x(0), x′ (0)) denotes the corresponding flow. Then, cv (t) = κ(x(t)) is the geodesic described in Example 3 and expM z (v) = κ ◦ Φ(1, κ−1(z ), κ∗ (v)). Lemma 5. [Sch01, Lemma 3.4 and Corollary 3.5] Let x′ (t) = F (t, x(t)) be a system of ordinary differential equations as above. Suppose that Φ(t, x) is the flow of this equation. Then there is a universal expression Expra only depending on the multi-index a such that xΦ(t, x0 ) ≤ Expra (cid:18) sup ′ ≤ a, t(cid:19) 0≤τ ≤t n(cid:12)(cid:12)(cid:12)Da x F (τ , Φ(τ , x0 ))(cid:12)(cid:12)(cid:12)o (cid:12)(cid:12)(cid:12) a ′ Da for al l t ≥ 0 where Φ(t, x0 ) is defined. Moreover, a corresponding statement holds for ordinary differential equations of order d. 3. Sobolev spaces on manifolds of bounded geometry From now on let M always be an n-dimensional manifold with Riemannian metric g . Definition 6. [Shu, Definition A.1.1] A Riemannian manifold (M n , g ) is of bounded geometry if the following two conditions are satisfied: (i) The injectivity radius rM of (M , g ) is positive. (ii) Every covariant derivative of the Riemann curvature tensor RM of M is bounded, i.e., for all k ∈ N0 there is a constant Ck > 0 such that (∇M )kRM g ≤ Ck . Remark 7. i) Note that Definition 6(i) implies that M is complete, cf. [Eic07, Proposition 1.2a]. ii)[Shu, Definition A.1.1 and below] Property (ii) of Definition 6 can be replaced by the following equivalent property which will be more convenient later on: Consider a geodesic atlas Ageo = (U geo α , κgeo α )α∈I as in α ∩ U geo Example 3. For all k ∈ N there are constants Ck such that for all α, β ∈ I with U geo 6= ∅ we have for β the corresponding transition functions µαβ := (κgeo β )−1 ◦ κgeo α that for all a ∈ Nn Daµαβ ≤ Ck , 0 with a ≤ k and all charts. iii) [Eic91, Theorem A and below] Consider a geodesic atlas Ageo as above. Let gij denote the metric in these coordinates and g ij its inverse. Then, property (ii) of Definition 6 can be replaced by the following equivalent property: For all k ∈ N0 there is a constant Ck such that Da gij ≤ Ck , Da g ij ≤ Ck , for all a ∈ Nn 0 with a ≤ k . (7) Example 8 (Geodesic trivialization). Let (M , g ) be of bounded geometry (this includes the case of closed manifolds). Then, there exists a geodesic atlas, see Example 3, that is uniformly locally finite: Let S be a maximal set of points {pα}α∈I ⊂ M such that the metric balls B r (pα ) are pairwise disjoint. 2 Then, the balls {Br (pα )}α∈I cover M , and we obtain a (uniformly locally finite) geodesic atlas Ageo = (U geo := Br (pα ), κgeo α )α∈I ). For an argument concerning the uniform local finiteness of the cover we refer α to Remark 23.ii. Moreover, there is a partition of unity hgeo subordinated to (U geo α )α∈I such that for all α k ∈ N0 there is a constant Ck > 0 such that Da (hgeo α ◦ κgeo α ) ≤ Ck for all multi-indices a with a ≤ k , [Tri92, Proposition 7.2.1] and the references therein. The resulting trivialization is denoted by T geo = cf. α , hgeo α , κgeo (U geo α )α∈I and referred to as geodesic trivialization. 6 3.1. Sobolev norm on manifolds of bounded geometry using geodesic normal coordinates. On manifolds of bounded geometry it is possible to define spaces H s p (M ) using local descriptions (geodesic nor- mal coordinates) and norms of corresponding spaces H s p (Rn ). Definition 9. Let (M n , g ) be a Riemannian manifold of bounded geometry with geodesic trivialization α )α∈I as above. Furthermore, let s ∈ R and 1 < p < ∞. Then the space H s T geo = (U geo α , hgeo α , κgeo p (M ) contains all distributions f ∈ D ′ (M ) such that Xα∈I p (Rn )! 1 p α kp α f ) ◦ κgeo k(hgeo H s α is only defined on V geo is finite. Note that although κgeo α f ) ◦ κgeo α ⊂ Rn , (hgeo α is viewed as a function on Rn extended by zero, since supp (hgeo α f ) ⊂ U geo α . p (M ) generalize in a natural way the classical Sobolev spaces W k Remark 10. The spaces H s p (M ), k ∈ p (M ) := Pk l=0 k∇lf kLp (M ) , then W k N0 , 1 < p < ∞, on Riemannian manifolds M : Let kf kW k p (M ) is the completion of D(M ) in the W k p (M )-norm, cf. [Aub76], [Aub82]. As in the Euclidean case, on manifolds M of bounded geometry one has the coincidence p (M ) = H k W k 1 < p < ∞, k ∈ N0 , p (M ), (8) (9) cf. [Tri92, Section 7.4.5]. Alternatively, the fractional Sobolev spaces H s p (M ) on manifolds with bounded geometry can be characterized with the help of the Laplace-Beltrami operator, cf. [Tri92, Section 7.2.2 and Theorem 7.4.5]. This approach was originally used by [Str83] and later on slightly modified in [Tri92, Section 7.4.5] in the following way: Let 1 < p < ∞ and ρ > 0. Let s > 0, then H s p (M ) is the collection of all f ∈ Lp (M ) such that f = (ρ Id −∆)−s/2h p (M ) = khkLp(M ) . Let s < 0, then H s for some h ∈ Lp (M ), with the norm kf kH s p (M ) is the collection of all f ∈ D ′ (M ) having the form f = (ρ Id −∆)l h with h ∈ H 2l+s (M ), where l ∈ N such that 2l + s > 0, and p (M ) . Let s = 0, then H 0 kf kH s p (M ) = khkH 2l+s p (M ) = Lp (M ). p In particular, the spaces H s p (M ) with s < 0 are independent of the number l appearing in their definition in the sense of equivalent norms, cf. [Str83, Definition 4.1]. The additional parameter ρ > 0 used by Triebel ensures that (9) also holds in this context as well. In particular, for 2 ≤ p < ∞ one can choose ρ = 1, cf. [Tri92, Rem. 1.4.5/1, p. 301]. Technically, it is possible to extend Definition 9 to the limiting cases when p = 1 and p = ∞. However, p (Rn ) already in the classical situation when M = Rn the outcome is not satisfactory: the resulting spaces H s have not enough Fourier multipliers, cf. [Tri92, p. 6, p. 13], and there is no hope for a coincidence in the sense of (9). Therefore, we restrict ourselves to 1 < p < ∞, but emphasize that the boundary cases are included in the outlook about F - and B -spaces in Section 6.2. 3.2. Sobolev norms on manifolds of bounded geometry using other trivializations. For many applications the norm given in (8) is very useful. In particular, it enables us to transfer many results known on Rn to manifolds M of bounded geometry. The choice of geodesic coordinates, however, often turns out to be far too restrictive if one needs to adapt the underlying coordinates to a certain problem, e.g., to submanifolds N of M in order to study traces. Therefore, in order to replace the geodesic trivializations in (8) we want to look for other ’good’ trivializations which will result in equivalent norms (and hence yield the same spaces). Definition 11. Let (M n , g ) be a Riemannian manifold together with a uniformly locally finite trivialization T = (Uα , κα , hα )α∈I . Furthermore, let s ∈ R and 1 < p < ∞. Then the space H s,T (M ) contains all p distributions f ∈ D ′ (M ) such that := Xα∈I p (Rn )! 1 p k(hαf ) ◦ καkp H s is finite. Here again (hαf ) ◦ κα is viewed as function on Rn , cf. (8) and below. 7 kf kH s,T p In general, the spaces H s,T (M ) do depend on the underlying trivialization T . One of our main aims will be p to investigate under which conditions on T this norm is equivalent to the H s p (M )-norm. For that we will use the following terminology. Definition 12. Let (M n , g ) be a Riemannian manifold of bounded geometry. Moreover, let a uniformly locally finite trivialization T = (Uα , κα , hα )α∈I be given. We say that T is admissible if the following conditions are fulfilled: (B1) A = (Uα , κα )α∈I is compatible with geodesic coordinates, i.e., for Ageo = (U geo β , κgeo β )β∈J being a geodesic atlas of M as in Example 3 there are constants Ck > 0 for k ∈ N0 such that for all α ∈ I and β ∈ J with Uα ∩ U geo 6= ∅ and all a ∈ Nn 0 with a ≤ k β Da (µβα = (κgeo Da (µαβ = (κα )−1 ◦ κgeo β )−1 ◦ κα ) ≤ Ck . β ) ≤ Ck (B2) For all k ∈ N there exist ck > 0 such that for all α ∈ I and all multi-indices a with a ≤ k and Da (hα ◦ κα ) ≤ ck . Remark 13. i) If (B1) is true for some geodesic atlas, it is true for any refined geodesic atlas. This follows immediately from Remark 7.ii. ii) Condition (B1) implies in particular the compatibility of the charts in T among themselves, i.e., for all k ∈ N0 there are constants Ck > 0 such that for all multi-indices a with a ≤ k and all α, β ∈ I with Uα ∩ Uβ 6= ∅ we have Da (κ−1 α ◦ κβ ) ≤ Ck . This is seen immediately when choosing z ∈ Uα ∩ Uβ , considering )−1 ◦κβ )). )◦ ((κgeo α ◦κgeo α ◦κβ ) = Da ((κ−1 around z , applying the chain rule to Da (κ−1 the exponential map κgeo z z z The same works for charts belonging to different admissible trivializations. Theorem 14. Let (M , g ) be a Riemannian manifold of bounded geometry, and let T = (Uα , κα , hα )α∈I be an admissible trivialization of M . Furthermore, let s ∈ R and 1 < p < ∞. Then, (M ) = H s H s,T p (M ), p i.e., for admissible trivializations of M the resulting Sobolev spaces H s,T p (M ) do not depend on T . Proof. The proof is based on pointwise multiplier assertions and diffeomorphism properties of the spaces p (Rn ), see Lemma 1. Let T = (Uα , κα , hα )α∈I be an admissible trivialization. Let a geodesic trivialization H s T geo = (U geo β , κgeo β , hgeo β )β∈J of M , see Example 8, be given. If α ∈ I is given, the index set A(α) collects all β ∈ J for which Uα ∩ U geo 6= ∅. The cardinality of A(α) can be estimated from above by a constant β independent of α since the covers are uniformly locally finite. We assume f ∈ H s p (M ). By Lemma 1 and Definition 12 we have for all α ∈ I β f ) ◦ κα(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H s p (Rn ) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ Xβ∈A(α) (cid:13)(cid:13)(cid:13)(hαhgeo β f ) ◦ κα(cid:13)(cid:13)(cid:13)H s Xβ∈A(α) (hαhgeo k(hαf ) ◦ κα kH s p (Rn ) p (Rn ) β (cid:13)(cid:13)(cid:13)H s . Xβ∈A(α) (cid:13)(cid:13)(cid:13)(hαhgeo = Xβ∈A(α) (cid:13)(cid:13)(cid:13)(hα hgeo β )−1 ) ◦ κα(cid:13)(cid:13)(cid:13)H s β f ) ◦ κgeo β ◦ (κgeo β f ) ◦ (κgeo p (Rn ) p (Rn ) . Xβ∈A(α) (cid:13)(cid:13)(cid:13)(hgeo β (cid:13)(cid:13)(cid:13)H s β f ) ◦ κgeo . p (Rn ) In particular, the involved constant can be chosen independently of α. Then p (Rn ) .  Xα∈I ,β∈A(α) (M ) = Xα∈I p (Rn )!1/p β kp β f ) ◦ κgeo k(hgeo k(hαf ) ◦ καkp kf kH s,T . kf kH s p (M ) H s H s p where the last estimate follows from Pα∈I , β∈A(α) = Pβ∈J, α∈A(β ) and the fact that the covers are uniformly (M ) = H s locally finite. The reverse inequality is obtained analogously. Thus, H s,T p (M ). (cid:3) p In view of Remark 7.iii, we would like to have a similar result for trivializations satisfying condition (B1). 8 1/p Lemma 15. Let (M , g ) be a Riemannian manifold with positive injectivity radius, and let T = (Uα , κα , hα )α∈I be a uniformly local ly finite trivialization. Let gij be the coefficient matrix of g and g ij its inverse with respect to the coordinates κα . Then, (M , g ) is of bounded geometry and T fulfil ls (B1) if, and only if, the fol lowing is fulfil led: For al l k ∈ N0 there is a constant Ck > 0 such that for al l multi-indices a with a ≤ k , Da g ij ≤ Ck Da gij ≤ Ck (10) and holds in al l charts κα . Proof. Let (10) be fulfilled. Then, (M , g ) is of bounded geometry since RM in local coordinates is given by a polynomial in gij , g ij and its derivatives. Moreover, condition (B1) follows from [Sch01, Lemma 3.8] – we shortly sketch the argument here: Let Γk ij denote the Christoffel symbols with respect to coordinates κα for α ∈ I . By (4) and (10), there are constants Ck > 0 for k ∈ N0 such that DaΓk ij ≤ Ck for all α ∈ I and all a ∈ Nn 0 with a ≤ k . Moreover, fix r > 0 smaller than the injectivity radius of M . Let β = Br (pgeo Ageo = (U geo β ), κgeo β )β∈J be a geodesic atlas of M where r > 0 is smaller than the injectivity radius. We get that (κα )−1 ◦ κgeo β (x) = Φ(1, κ−1 α (pβ ), κ∗ α (λβ (x))) where Φ is the geodesic flow. Then, together with Lemma 5 it follows that (κα )−1 ◦ κgeo and all its derivatives are uniformly bounded independent on α and β β . Moreover, note that (κgeo α (Uα ∩ U geo r → (κgeo β )−1 (Uα ∩ U geo β )−1 ◦ κα : κ−1 β ) ⊂ Bn β ) ⊂ Bn r is bounded by r. Hence, together with the chain rule applied to ((κgeo β )−1 ◦ κα ) ◦ ((κα )−1 ◦ κgeo β ) = Id condition (B1) follows for all (α, β ). Conversely, let (M , g ) be of bounded geometry, and let condition (B1) be fulfilled. Then, by Remark 7.iii and the transformation formula (5) for α ∈ I and β ∈ J , condition (10) follows. (cid:3) 3.3. Besov spaces on manifolds. Similar to the situation on Rn we can define Besov spaces on manifolds via real interpolation of fractional Sobolev spaces H s p (M ). , (11) where s = Θs0 + (1 − Θ)s1 . Definition 16. Let (M , g ) be a manifold of bounded geometry. Furthermore, let s0 , s1 ∈ R, 1 < p < ∞ and 0 < Θ < 1. We define p (M )(cid:1)Θ,p p,p (M ) := (cid:0)H s0 p (M ), H s1 B s Remark 17. The fractional Sobolev spaces H si p (M ) appearing in Definition 16 above should be understood in the sense of Definition 11. For the sake of simplicity we restrict ourselves to admissible trivializations T when defining Besov spaces on M . This way, by Theorem 14, we can omit the dependency on the trivializations T from our notations in 11 since resulting norms are equivalent and yield the same spaces. Note that our spaces are well-defined since (11) is actually independent of s0 and s1 . An explanation is given in [Tri92, Theorem 7.3.1]. Furthermore, an equivalent norm for f ∈ B s p,p (M ) is given by p,p(M ) = Xα∈I p,p (Rn )! 1 p k(hαf ) ◦ κα kp kf kB s B s We sketch the proof. By ℓp (H s p ) we denote the sequence space containing all sequences {fα}α∈I such that the norm p ! 1 p ) := Xα∈I p kfαkp kfαkℓp (H s H s is finite, similar for ℓp (B s p,p ) with obvious modifications. Let A(α) = {β ∈ I Uβ ∩ Uα 6= ∅}, and let Λα = (cid:16)Pβ∈A(α) hβ (cid:17) ◦ κα . We define a linear and bounded operator Λ : ℓp (H s p (Rn )) −→ H s p (M ), via Λ {fβ }β∈I = Xβ∈I 9 (Λβ fβ ) ◦ κ−1 β , (12) . where (Λβ fβ ) ◦ κ−1 β is extended outside Uβ by zero. Furthermore, we consider Ψ : H s p (M ) −→ ℓp (H s p (Rn )), given by Λ ◦ Ψ = Id Ψ(f ) = {(hαf ) ◦ κα}α∈I which is also a linear and bounded operator. In particular, we have that (identity in H s p (M )). Having arrived at a standard situation of interpolation theory we use the method of retraction/coretraction, cf. [Tri78, Theorem 1.2.4], reducing (12) to the question whether p (cid:1)Θ,p(cid:17), = ℓp(cid:16)(cid:0)H s0 p )(cid:17)Θ,p (cid:16)ℓp (H s0 p ), ℓp (H s1 p , H s1 for 1 < p < ∞, s0 , s1 ∈ R, 0 < Θ < 1, and s = Θs0 + (1−Θ)s1 , which can be found in [Tri78, Theorem 1.18.1]. Since by definition of Besov spaces the right hand side of (13) coincides with ℓp (B s p,p ), this proves (12). (13) 4. Coordinates on submanifolds and Trace Theorems From now on let N k ⊂ M n be an embedded submanifold, meaning, there is a k -dimensional manifold N ′ and an injective immersion f : N ′ → M with f (N ′ ) = N . The aim of this section is to prove a Trace Theorem for M and N . We restrict ourselves to submanifolds of bounded geometry in the following sense: Definition 18. Let (M n , g ) be a Riemannian manifold with a k -dimensional embedded submanifold (N k , g N ). We say that (M , N ) is of bounded geometry if the following is fulfilled (i) (M , g ) is of bounded geometry. (ii) The injectivity radius rN of (N , g N ) is positive. (iii) There is a collar around N (a tubular neighbourhood of fixed radius), i.e., there is r∂ > 0 such that r∂ (x) and B⊥ for all x, y ∈ N with x 6= y the normal balls B⊥ r∂ (y ) are disjoint where r∂ (x) := {z ∈ M distM (x, z ) ≤ r∂ , ∃ε0∀ε < ε0 : distM (x, z ) = distM (BN B⊥ ε (x), z )} with BN ε (x) = {u ∈ N distN (u, x) ≤ ε} and distM and distN denote the distance func- tions in M and N , respectively. N B⊥ r∂ (y ) y x B⊥ r∂ (x) BN ε (x) (iv) The mean curvature l of N given by X Y − ∇N l(X, Y ) := ∇M for all X, Y ∈ T N , X Y and all its covariant derivatives are bounded. Here, ∇M is the Levi-Civita connection of (M , g ) and ∇N the one of (N , g N ). Remark 19. i) If the normal bundle of N in M is trivial, condition (iii) in Definition 18 simply means that {z ∈ M distM (z , N ) ≤ r∂ } is diffeomorphic to Bn−k r∂ × N . Then z (cid:0)ti νi (cid:1) r∂ × N → M ; (t, z ) 7→ expM F : Bn−k is a diffeomorphism onto its image, where (t1 , ..., tn−k ) are the coordinates for t with respect to a standard orthonormal basis on Rn−k and (ν1 , . . . , νn−k ) is an orthonormal frame for the normal bundle of N in M . If the normal bundle is not trivial (e.g. consider a noncontractible circle N in the infinite Mobius strip M ), F still exists locally, which means that for all x ∈ N and ε smaller than the injectivity z (cid:0)ti νi(cid:1) is a diffeomorphism onto its radius of N , the map F : Bn−k r∂ × BN ε (x) → M ; (t, z ) 7→ expM image. All included quantities are as in the case of a trivial vector bundle, but νi is now just a local orthonormal frame of the normal bundle. By abuse of notation, we suppress here and in the following the dependence of F on ε and x. 10 ii) The illustration below on the left hand side shows a submanifold N of a manifold M that admits a collar. On the right hand side one sees that for M = R2 the submanifold N describing the curve which for large enough x contains the graph of x 7→ x−1 together with the x-axes does not have a collar. This situation is therefore excluded by Definition 18. However, to a certain extend, manifolds as in the picture on the right hand side can still be treated, cf. Example 32 and Remark 33. M N expM z z F (B n−k rδ × N ) x−1 M = R2 N iii) Although our notation (M , N ) hides the underlying metric g , this is obviously part of the definition and fixed when talking about M . iv) If N is the boundary of the manifold M , the counterpart of Definition 18 can be found in [Sch01, Definition 2.2], where also Fermi coordinates are introduced and certain properties discussed. In Section 4.1, we adapt some of the methods from [Sch01] to our situation. Note that the normal bundle of the boundary of a manifold is always trivial, which explains why in [Sch01, Definition 2.2] condition (iii) of Definition 18 reads as in Remark 19.i. 4.1. Fermi coordinates. In this subsection we will introduce Fermi coordinates, which are special coordi- nates adapted to a submanifold N of M where (M , N ) is of bounded geometry. The resulting trivialization is used to prove the Trace Theorem in Section 4.2. Definition 20 (Fermi coordinates). We use the notations from Definition 18. Let (M n , N k ) be of 2 r∂ (cid:9), where rN is the injectivity radius of N and rM the one bounded geometry. Let R = min (cid:8) 1 2 rN , 1 4 rM , 1 of M . Let there be countable index sets IN ⊂ I and sets of points {pN α }α∈IN and {pβ }β∈I \IN in N and M \ UR (N ), respectively, where UR (N ) := ∪x∈N B⊥ R (x). Those sets are chosen such that (i) The collection of the metric balls (BN R (pN α ))α∈IN gives a uniformly locally finite cover of N . Here the balls are meant to be metric with respect to the induced metric g N . (ii) The collection of metric balls (BR (pβ ))β∈I \IN covers M \ UR (N ) and is uniformly locally finite on all of M . := F (Bn−k 2R (pN 2R × BN We consider the covering (Uγ )γ∈I with Uγ = BR (pγ ) for γ ∈ I \ IN and Uγ = UpN γ )) γ with γ ∈ IN . Coordinates on Uγ are chosen to be geodesic normal coordinates around pγ for γ ∈ I \ IN . Otherwise, if γ ∈ IN , coordinates are given by Fermi coordinates κγ : VpN γ := Bn−k 2R × B k 2R → UpN γ γ (x)) (cid:0)ti νi(cid:1) (λN where (t1 , . . . , tn−k ) are the coordinates for t with respect to a standard orthonormal basis on Rn−k , (ν1 , . . . , νn−k ) is an orthonormal frame for the normal bundle of BN 2R (pN γ ) in M , expN is the exponen- tial map on N with respect to the induced metric g N , and λN γ : Rk → TpN N is the choice of an orthonormal γ frame on TpN N . γ , (t, x) 7→ expM expN pN γ (14) 11 κγ Rn−k 2R (x, t) 2R Rk N VpN γ expM expN pN γ γ (x)) (ti νi ) (λN UpN γ pN γ expN pN γ (λN γ (x)) M Before giving a remark on the existence of the points {pγ }γ∈I claimed in the Definition above, we prove two lemmata. Lemma 21. Let (M n , N k ) be of bounded geometry, and let C > 0 be such that the Riemannian curvature tensor fulfil ls RM ≤ C and mean curvature of N l ≤ C . Fix z ∈ N and R as in Definition 20. Let U = F (Bn−k 2R × BN 2R (z )), and let a chart κ for U be defined as above. Then there is a constant C ′ > 0 only depending on C , n and k , such that gij ≤ C ′ and g ij ≤ C ′ where gij denotes the metric g with respect to κ. Proof. For N being the boundary of M this was shown in [Sch01, Lemma 2.6]. We follow the idea given there and use the extension of the Rauch comparison theorem to submanifolds of arbitrary codimension given by Warner in [Wa66, Theorem 4.4]. For the comparison, let MC and M−C be two complete n-dimensional Riemannian manifolds of constant sectional curvature C and −C , respectively. In each of them we choose a k -dimensional submanifold NC and N−C , points p±C ∈ N±C and a chart of M±C around p±C given by Fermi coordinates such that all eigenvalues of the second fundamental form with respect to those coordinates at p±C are given by ±C (this is always possible, cf. [SpIV, Chapter 7]). Let (νi )1≤i≤n−k be an orthonormal frame of the normal bundle of U ∩ N and (ei )1≤i≤k be an orthonormal frame of T U ∩N N obtained via geodesic flow on N . Let the frame (ν1 , . . . , νn−k , e1 , ..., ek ) be transported to all of U via parallel transport along geodesics normal to N – the transported vectors are also denoted by νi and ei , respectively. Then, we are in the situation to apply [Wa66, Theorem 4.4]: Let now p ∈ U and v ∈ TpU with v ⊥ νi for all 1 ≤ i ≤ n − k . Then, the comparison theorem yields constants C1 , C2 > 0, depending only on C, n, and k , such that C1 v 2 E ≤ gp (v , v) ≤ C2 v 2 E , where .E denotes the Euclidean metric with respect to the basis (ei ). Moreover, we have gp (νi , νj ) = δij and gp (νi , el ) = 0 for all 1 ≤ i, j ≤ n − k and 1 ≤ l ≤ k , since this is true for p ∈ U ∩ N , and this property is preserved by parallel transport. Altogether this implies the claim. (cid:3) The previous lemma enables us to show that (N , g N ) is also of bounded geometry. Lemma 22. If (M , N ) is of bounded geometry, then (N , g N ) is of bounded geometry. Proof. Since Definition 18 already includes the positivity of the injectivity radius of N , it is enough to show that (∇N )kRN , where RN is the Riemannian curvature of (N , g N ), is bounded for all k ∈ N0 : Let z ∈ N . We consider geodesic normal coordinates κgeo : B k 2R → U geo = BN 2R (z ) on N around z and Fermi coordinates κ : Bn−k 2R → U = F (Bn−k 2R × B k 2R × BN 2R (z ) on M around z , cf. Definition 20. Let gij be the metric with respect to the coordinates given by κ, and let g ij be its inverse. Since M is of bounded geometry, Lemma 21 yields a constant C independent on z such that we have gij ≤ C and g ij ≤ C . Together with the uniform boundedness of RM , l and their covariant derivatives, we obtain that their representations RM ijkl , lrs and their derivatives in the coordinates given by κ are uniformly bounded for all 1 ≤ i, j, k , l ≤ n, n − k + 1 ≤ r, s ≤ n − k . Then the claim follows by the Gauss’ equation [SpIV, p. 47], g (RN (U, V )W, Z ) = g (RM (U, V )W, Z ) + g (l(U, Z ), l(V , W )) − g (l(U, W ), l(V , Z )) for all U, V , W, Z ∈ T N , 12 and the formulas for covariant derivatives of tensors along N . We refer to [Sch01, Lemma 2.22], where everything is stated for hypersurfaces but the formulas remain true for arbitrary codimension sub ject to obvious modifications. (cid:3) Remark 23. i) By construction the covering (Uγ )γ∈I is uniformly locally finite and (U ′ γ := Uγ ∩ N , κ′ γ = κγ κ−1 γ ) )γ∈IN gives a geodesic atlas on N . Moreover, none of the balls BR (pβ ) with β ∈ I \ IN intersects γ (U ′ N . ii) Existence of points pN α and pβ as claimed in Definition 20 (note that the proofs of Lemmas 21 and 22 only use the definition of Fermi coordinates on a single chart and the existence of the points pN α and pβ ): We α }α∈IN of points in N such that the metric balls BN choose a maximal set {pN (pN α ) are pairwise disjoint. Then, R 2 R (pN the balls BN α ) cover N . Since by Lemma 22 the submanifold (N , g N ) is of bounded geometry, the volume of metric balls in N with fixed radius is uniformly bounded from above and from below away from zero. Let a ball BN 2R (pN α ) be intersected by L balls BN 2R (pN α′ ). Then the union of the L balls BN 2R (pN α′ ) forms a subset 4R (pN 2R (pN α ). Comparison of the volumes gives an upper bound on L. Hence, the balls in (BN of BN α ))α∈IN cover N uniformly locally finite. Moreover, choose a maximal set of points {pβ }β∈I \IN ⊂ M \ UR (N ) such (pβ ) are pairwise disjoint in M . Then the balls BR (pβ ) cover M \ UR (N ). Trivially that the metric balls B R 2 the balls BR (pβ ) for β ∈ I \ IN cover ∪BR (pβ ), and by volume comparison as above this cover is uniformly locally finite. Lemma 24. The atlas (Uγ , κγ )γ∈I introduced in Definition 20 fulfil ls condition (B1). Proof. For all γ ∈ I \ IN the chart κγ is given by geodesic normal coordinates and, thus, condition (B1) follows from Remark 7.ii. Let now γ ∈ IN . Then the claim follows from [Sch01, Lemma 3.9]. We sketch the proof. Consider a chart α ), κN ,geo) in (N , g N ) both given by geodesic normal coordinates α ), κgeo ) in M and a chart (BN 2R (pN (B4R (pN around pN α for α ∈ IN . Note that 4R < rM by Definition 20. Let Φ2 be the geodesic flow in (N , g N ) with respect to the coordinates given by κN ,geo , cf. Example 4. Let Φ1 be the corresponding geodesic flow in (M , g ) given by κgeo . Then, Φ2 (1, 0, x) = (κN ,geo)−1 ◦ expN (λN α (x)) pN α and by (14), κα (t, x) = κgeo ◦ Φ1 (1, Φ2 (1, 0, x), (κgeo)∗ (ti νi )) with t = (t1 , . . . , tn−k ) ∈ Rn−k . Since (M , g ) is of bounded geometry, the coefficient matrix gij of g with respect to κgeo , its inverse and all its derivatives are uniformly bounded by (7). Moreover, by Lemma 22 (N , g N ) is also of bounded geometry and, thus, we get an analogous statement for the coefficient matrix of g N with respect to κN ,geo. Hence, applying Lemma 5 to the differential equation of the geodesic flows, see Example 4 and (4), we obtain that (κgeo )−1 ◦ κα and all its derivatives are bounded independent on α. Conversely, (κα )−1 ◦ κgeo : (κgeo )−1 ◦ κα (Bn−k 2R × B k 2R ) ⊂ Bn 4R → Bn−k 2R ×B k 2R is bounded independent on α. Hence, by using the chain rule on ((κα )−1 ◦ κgeo) ◦ ((κgeo )−1 ◦ κα) = Id one sees that also the derivatives of (κα )−1 ◦ κgeo are uniformly bounded, which gives the claim. (cid:3) Lemma 25. There is a partition of unity subordinated to the Fermi coordinates introduced in Definition 20 fulfil ling condition (B2). Proof. By Lemma 22, (N , g N ) is of bounded geometry. Then, by Example 8, there is a partition of unity α subordinated to a geodesic atlas (U ′ h′ α ), κ′ 2R (pN α := Uα ∩ N = BN α = κα κ−1 α ) )α∈IN of N such that for α (U ′ each a ∈ Nk 0 the derivatives Da (h′ α ◦ κ′ α ) are uniformly bounded independent of α. Since by construction the balls BN R (pN α ) already cover N the functions h′ α can be chosen such that supp h′ α ⊂ BN R (pN α ). Choose a function ψ : Rn−k → [0, 1] that is compactly supported on Bn−k 2 R ⊂ Rn−k and ψ Bn−k = 1. Set 3 R hα = (ψ × (h′ α )) ◦ κ−1 α ◦ κ′ on Uα and zero outside. Then, supp hα ⊂ Uα and all Da (hα ◦ κα ) are uniformly α bounded by a constant depending on a but not on α ∈ IN . Let S ⊂ M be a maximal set of points containing the set {pβ }β∈I \IN of Definition 20 such that the metric (p)}p∈S are pairwise disjoint. Then (BR (p))p∈S forms a uniformly locally finite cover of M. We balls in {B R 2 equip this cover with a geodesic trivialization (BR (p), κgeo p , hgeo p )p∈S , see Example 8. For β ∈ I \ IN we have by construction κβ = κgeo pβ and set hβ =  (1 − Pα∈IN , where Pβ ′∈I \IN else. 0, hgeo pβ Pβ ′ ∈I\IN hgeo p β ′ hgeo p′ β 6= 0 hα ) 13 Next, we will argue that all hβ are smooth: It suffices to prove the smoothness in points x ∈ M on the boundary of {Pβ ′ ∈I \IN hgeo 6= 0}. For all other x smoothness follows by smoothness of the functions hα p′ p . Let now x ∈ M as specified above. Then Pβ ′∈I \IN β hgeo and hgeo pβ ′ (x) = 0 and, thus, x ∈ UR (N ) (cf. Remark 23.ii). Together with ψ Bn−k = 1 this implies that for ε small enough there is a neighbourhood Bε (x) ⊂ UR (N ) such that Pα∈IN R hα (y ) = 1 for all y ∈ Bε (x). Thus, hβ Bε (x) = 0 and hβ is smooth in x for all β ∈ I \ IN . hβ + Pα∈IN Moreover, by construction Pβ∈I \IN hα = 1. Hence, (hγ )γ∈I gives a partition of unity subor- dinated to the Fermi coordinates. The uniform boundedness of all Da (hβ ◦ κβ ) follows from the uniform α ◦ κβ ) and Da ((κβ ′ )−1 ◦ κβ ) together with Remark pβ ′ ◦ κβ ′ ), Da (κ−1 boundedness of all Da (hα ◦ κα ), Da (hgeo 13.ii. (cid:3) Collecting the last two lemmata we obtain immediately: Theorem 26. Let (M , N ) be of bounded geometry. Let T FC be a trivialization of M given by Fermi coor- dinates as in Definition 20 together with the subordinated partition of unity of Lemma 25. Then, T FC is an admissible trivialization. 4.2. Trace Theorem. Let (M n , g ) be a Riemannian manifold together with an embedded submanifold N k where k < n. For f ∈ D(M ) the trace operator is defined by pointwise restriction, TrN f := f (cid:12)(cid:12)N . Let X (M ) and Y (N ) be some function or distribution spaces on M and N , respectively. If TrN extends to a continuous map from X (M ) into Y (N ), we say that the trace exists in Y (N ). If this extension is onto, we write TrN X (M ) = Y (N ). For fractional Sobolev spaces on manifolds we have the following trace result. Theorem 27. Let (M n , g ) be a Riemannian manifold together with an embedded k-dimensional submanifold N . Let (M , N ) be of bounded geometry. If 1 < p < ∞ and s > n−k p , then TrN is a linear and bounded s− n−k p p,p operator from H s p (M ) onto B (N ), i.e., TrN H s p (M ) = B s− n−k p p,p (N ). (15) Remark 28. For (M , N ) = (Rn , Rk ) this is a classical result, cf. [Tri83, p. 138, Remark 1] and the references given therein. Here we think of Rk ∼= {0}n−k × Rk ⊂ Rn . Furthermore, in [Tri83, p. 138, Remark 1] it is also shown that TrRk has a linear bounded right inverse – an extension operator ExRn . p (Rn ) and η ∈ D(Rn ) we have TrRk (ηf ) = Note that TrRk respects products with test functions, i.e., for f ∈ H s η Rk TrRk f . Moreover, if κ is a diffeomorphism on Rn such that κ(Rk ) = Rk , then TrRk (f ◦ κ) = TrRk f ◦ κRk . Proof of Theorem 27. Via localization and pull-back we will reduce (15) to the classical problem of traces on hyperplanes Rk in Rn . The proof is similar to [Skr90, Theorem 1], but the Fermi-coordinates enable us to drop some of the restricting assumptions made there. By Theorem 26 we have an admissible trivialization T = (Uα , κα , hα )α∈I of M by Fermi coordinates and the subordinated partition of unity from Lemma 25. Moreover, by the construction of the Fermi coordinates it ) = {0}n−k × B k is clear that κ−1 α (N ∩ UpN 2R and, thus, their restriction to N gives a geodesic trivialization α α := Uα ∩ N , κ′ T N ,geo = (U ′ α ) , h′ α := κα κ−1 α := hα U ′ α )α∈IN of N . α (U ′ 1. Step: Let f ∈ H s p (M ). We define the trace operator via (TrN f )(x) := Xα∈IN p (Rn ) and supp TrRk ((hα f ) ◦ κα ) ⊂ V ′ Note that TrN is well-defined since (hα f ) ◦ κα ∈ H s α = Vα ∩ Rk . Moreover, for fixed x ∈ N the summation is meant to run only over those α for which x ∈ U ′ α . Hence, the summation only runs over finitely many α due to the uniform locally finite cover. Obviously, TrN is linear s− n−k and TrN D(M ) is given by the pointwise restriction. In order to show that TrN : H s p p (M ) → B p,p 14 TrRk [(hα f ) ◦ κα ] ◦ (κ′ α )−1 (x), x ∈ N . (N ) is (Rk ) (Rk ) (16) k(hαf ) ◦ καkp p (Rn ) , H s bounded, we set A(α) := {β ∈ IN Uα ∩ Uβ 6= ∅}. Since the cover is uniformly locally finite, the number of elements in A(α) is bounded independent of α. Together with Lemma 1 and Remark 28 we obtain = Xβ∈IN kTrN f kp β kp k(h′ β TrN f ) ◦ κ′ s− n−k s− n−k p p (Rk ) B B (N ) p,p p,p . Xβ∈IN ; α∈A(β ) β (cid:3)(cid:1) kp β ) (cid:0)TrRk [(hα f ) ◦ κα ] ◦ (cid:2)(κ′ k(h′ β ◦ κ′ α )−1 ◦ κ′ s− n−k p B p,p = Xβ∈IN ; α∈A(β ) β (cid:3) kp kTrRk [(hαhβ f ) ◦ κα ] ◦ (cid:2)(κ′ α )−1 ◦ κ′ s− n−k p (Rk ) B p,p . Xβ∈IN ; α∈A(β ) kTrRk [(hαhβ f ) ◦ κα ] kp n−k s− p B p,p p (Rn ) . Xα∈IN . Xα∈IN ;β∈A(α) k(hαhβ f ) ◦ καkp H s and hence, kTrN f kp . kf kp p (M ) , where the involved constants do not depend on f . H s s− n−k p B (N ) p,p 2. Step: We will show that TrN is onto by constructing a right inverse – an extension operator ExM . Firstly, 2R , supp ψ2 ∈ Bn−k 2R , ψ1 ≡ 1 on B k let ψ1 ∈ D(Rk ), and ψ2 ∈ D(Rn−k ) such that supp ψ1 ∈ B k R and ψ2 ≡ 1 on Bn−k 2 R . Then, we put ψ := ψ1 × ψ2 ∈ D(Rn ). 3 s− n−k Let f ′ ∈ B p (N ). Then we define the extension operator by p,p (ExM f ′ )(x) := (cid:26)Pα∈IN 2R × Bn−k α ) is compactly supported in Vα = B k α f ′ ) ◦ κ′ Note that the use of ψ is to ensure that ψExRn ((h′ 2R for all α ∈ IN . Hence, one sees immediately that ExM is well-defined and calculates TrN (ExM f ′ ) = f ′ . s− n−k (N ) → H s p p (M ) is bounded, we use Lemma p,p α )] ◦ κ−1 α f ′ ) ◦ κ′ [ψExRn ((h′ α (x), x ∈ U2R (N ) 0, otherwise. Thus, TrN is onto. Moreover, in order to show that ExM : B 1 and Remark 28 again, which give β (cid:17)(cid:17) ◦ κα(cid:13)(cid:13)(cid:13) p (Rn ) . Xα∈I ; β∈A(α) (cid:13)(cid:13)(cid:13)(cid:16)hα (cid:16)(cid:2)ψExRn ((h′ p (M ) = Xα∈I β )(cid:3) ◦ κ−1 kExM f ′kp k(hαExM f ′ ) ◦ καkp β f ′ ) ◦ κ′ H s H s Xβ∈IN ; α∈I ; Uα∩Uβ 6=∅ p (Rn ) . Xβ∈IN β )kp β )kp k(hα ◦ κβ )ψExRn ((h′ β f ′ ) ◦ κ′ kExRn ((h′ β f ′ ) ◦ κ′ . H s H s p (Rn ) . Xβ∈IN β kp = kf ′kp β f ′ ) ◦ κ′ k(h′ . s− s− B B p,p p,p Note that the estimate in the second line uses (hα ◦ κβ )ψ ∈ D(Vβ ). This finishes the proof. n−k p (N ) n−k p (Rk ) p H s p (Rn ) (cid:3) According to the coincidence with the classical Sobolev spaces W k p (M ) in the case of bounded geometry, cf. (9), we obtain the following trace result as a consequence. Corollary 29. Let (M n , g ) be a Riemannian manifold together with an embedded k-dimensional submanifold N . Let (M , N ) be of bounded geometry. If m ∈ N, 1 < p < ∞, and m > n−k p , then TrN is a linear and m− n−k p p,p bounded operator from W m p (M ) onto B (N ), i.e., TrN W m p (M ) = B m− n−k p p,p (N ). 15 Example 30. Our results generalize [Skr90] where traces were re- stricted to submanifolds N which had to be totally geodesic. By using Fermi coordinates we can drop this extremely restrictive as- sumption and cover more (sub-)manifolds. For example, consider the case where M is a surface of revolution of a curve γ and N a circle obtained by the revolution of a fixed point p ∈ M . This resulting circle is a geodesic if and only if the rotated curve has an extremal point at p. But there is always a collar around N , hence, this situation is also covered by our assumptions. z N p γ M y x Remark 31. We proved even more than stated. In Step 2 above it was shown that there exists a linear and bounded extension operator ExM from the trace space into the original space such that where Id stands for the identity in B s− n−k p p,p (N ). TrN ◦ ExM = Id, The first part of the Trace Theorem 27 (i.e., the boundedness of the trace operator) can be extended to an even broader class of submanifolds. We give an example to illustrate the idea. 1 ) and (M n , N k Example 32. Let (M n , N k 2 ) be manifolds of bounded geometry with N1 ∩ N2 = ∅. Set N := N1 ∪ N2 . Clearly, TrN = TrN1 + TrN2 (where TrN1 f and TrN2 f are viewed as functions on N that s− n−k equal zero on N2 and N1 , respectively), and TrN is a linear bounded operator from H s p (N ). p (M ) to B p,p One may think of N = Graph(x 7→ x−1 ) ∪ x − axis ⊂ R2 , where (R2 , N ) does not posses a uniform collar, cf. Definition 18.iii. The boundedness of TrN is no longer expectable for an arbitrary infinite union of Ni , e.g., consider Ni = R × {i−1} ⊂ R2 , i ∈ N and put f = ψ1 × ψ2 with ψ1 , ψ2 ∈ D(R). Then N = ⊔iNi ֒→ R2 is an embedding, when N is equipped with standard topology on each copy of R. But one cannot expect the trace operator to be bounded, since not every function f ∈ C∞ c (R2 ) restricts to a compactly supported function on N (N as a subset of R2 is not intersection compact). This problem can be circumvented when requiring that the embedded submanifold N has to be a closed subset of M . However, even in this situation on can find submanifolds N for which the trace operator is not R × {i + j bounded in the sense of (16), e.g. consider N = ⊔i∈N ⊔i−1 i } ֒→ R2 . j=0 Remark 33. The above considerations give rise to the following generalization of Step 1 of the Trace Theorem 27. Assume that N is a k -dimensional embedded submanifold of (M n , g ) fulfilling (i), (ii) and (iv) of Definition 18 – but not (iii). Lemmas 21 and 22 remain valid, since their proofs do not use (iii). We replace (iii) with the following weaker version: α ))α∈IN be a uniformly locally finite cover of N . Set Uα = F (Bn−k 2R (pN 2R × BN (iii)′ Let (U ′ α = B2R (pN α )) as before. Then, (Uα )α∈IN is a uniformly locally finite cover of ∪α∈IN Uα . Condition (iii)′ excludes the negative examples from above . Furthermore, (iii)′ together with the complete- ness of N implies that N is a closed subset of M . ∩ N 6= With this modification, one can still consider Fermi coordinates as in Definition 20 but in general UpN α BN α ) =: U ′ 2R (pN . Also the partition of unity can be constructed as in Lemma 25 when making the following pN α step in between: Following the proof of Lemma 25 we define the map hα = (ψ × (h′ α ◦ κ′ α )) ◦ κ−1 α for α ∈ IN . Since in general Pα∈IN hα (x) can be bigger than one, those maps cannot be part of the desired partition hα′ )−1 where Pα′ ∈IN of unity. Hence, we put hα = hα (Pα′∈IN hα′ 6= 0 and hα = 0 else. Smoothness and uniform boundedness of the derivatives of hα follow as in Lemma 25. Then one proceeds as before, defining hβ for I \ IN . Now the proof of Step 1 of the Trace Theorem 27 carries over when replacing hα by hα in the definition of . Pα∈IN k(hαf ) ◦ κα kp TrN and in the estimate (16). This leads to kTrN f kp p (Rn ) . Finally, H s s− n−k p B p,p 16 (N ) kTrN f kp s− n−k p B p,p (N ) . Xα∈IN p (Rn ) = Xα∈IN k(hα ( Xβ∈IN k(hαf ) ◦ καkp H s p (Rn ) . Xα∈IN . Xα∈IN ; β∈A(α) k(hα hβ f ) ◦ καkp H s demonstrates the boundedness of the trace operator TrN under this generalized assumptions on the subman- ifold N . k(hαf ) ◦ καkp p (Rn ) ≤ kf kp p (M ) , H s H s hβ )f ) ◦ καkp H s p (Rn ) 5. Vector bundles The results about function spaces on manifolds of bounded geometry obtained so far can be transferred to certain vector bundles. For that we need a concept of bounded geometry for vector bundles. After giving such a definition, we shall proceed along the lines of the previous section – introducing synchronous trivialization along geodesic normal coordinates and Fermi coordinates and stating a corresponding Trace Theorem. 5.1. Vector bundles of bounded geometry. Definition 34. [Shu, Section A1.1] Let E be a vector bundle over a Riemannian manifold (M , g ) of bounded α , κgeo α , ξα )α∈I where (U geo α , κgeo geometry together with an atlas AE = (U geo α )α∈I is a geodesic atlas of M as in Example 3. Let µαβ denote the transition functions belonging to ξα , ξβ . The vector bundle E together with the choice of an atlas AE is said to be of bounded geometry if for all k ∈ N0 there is a constant Ck α ∩ U geo such that Da µαβ ≤ Ck for all α, β ∈ I with U geo 6= ∅ and all multi-indices a with a ≤ k . β We give an example of a special trivialization ξα : Definition 35 (Synchronous trivialization along geodesic normal coordinates). Let (E , ∇E , h., .iE ) be a Riemannian or hermitian vector bundle over a Riemannian manifold (M n , g ). Let M be of bounded geometry, and let the connection ∇E be metric. Let Ageo = (U geo α , κgeo α )α∈I be a geodesic atlas of M as in Example 3, and let pα denote the center of the ball U geo α . The choice of the orthonormal frame on Tpα M – already used in the definition of the geodesic coordinates, cf. Example 3 – is again denoted by λα : Rn → Tpα M . We choose an orthonormal frame (e1 (pα ), . . . , er (pα )) for each Epα (α ∈ I ). Then, E U geo is trivialized by parallel transport along radial geodesics emanating from pα as follows: α cv X v ρ (t) ∈ Ecv (t) be the unique solution of the differential equation ∇E For 1 ≤ ρ ≤ r, let X v ρ = 0 with ρ (0) = eρ (pα ) and cv (t) being the unique geodesic with cv (0) = pα and cv (0) = v ∈ T ≤r X v pα M , where r is smaller than the injectivity radius of M . Then the trivialization by parallel transport is given by α × Cr 7→ uρX λα (x) ξ geo : (x, u) ∈ V geo (1) ∈ E U geo and is called synchronous trivialization (along geodesic ρ α α normal coordinates). Ageo α , ξ geo α , κgeo E = (U geo α )α∈I is called a geodesic atlas of E . Note that by construction, hστ (0) = δστ for all α ∈ I . Since eσ on Uα is obtained by the parallel transport for a metric connection, we get hστ = δστ on each Uα and, hence, Γρ iσ = − Γσ iρ , cp. (6). Remark 36. In [Eic07, Section 1.A.1] one can find another definition of E being of bounded geometry: A hermitian or Riemannian vector bundle (E , ∇E , h., .iE ) over (M , g ) with metric connection is of bounded geometry, if (M , g ) is of bounded geometry and if the curvature tensor of E and all its covariant derivatives are uniformly bounded. In [Eic91, Theorem B] it was shown that bounded geometry of E in the sense of [Eic07, Section 1.A.1] is equivalent to the following condition: For all k ∈ N0 there is a constant Ck such that for all α ∈ I , 1 ≤ i ≤ n, 1 ≤ ρ, σ ≤ r and all multi-indices a with a ≤ k , Da Γσ iρ ≤ Ck , where Γσ iρ denote the Christoffel symbols with respect to ξ geo α . Our next aim is to compare the two definitions of bounded geometry of E given above: (17) Theorem 37. Let (E , ∇E , h., .iE ) be a hermitian or Riemannian vector bund le over a Riemannian manifold (M , g ). Let (M , g ) be of bounded geometry, and let ∇E be a metric connection. Moreover, let Ageo E = α )α∈I be a geodesic atlas of E , see Definition 35. Then, E together with Ageo (U geo α , κgeo α , ξ geo E is of bounded 17 geometry in the sense of Definition 34 if, and only if, it is of bounded geometry in the sense of [Eic91, Section 1.A.1], cf. Remark 36. α ∩ U geo α ∩ U geo β . Then, for all z ∈ U geo Proof. Let α, β ∈ I . For simplicity, we assume that pα , pβ ∈ U geo the β α ∩ U geo geodesics joining z with pα and pβ , respectively, are completely contained in U geo β . For atlases not satisfying this assumption, one can switch to a refined atlas and use the composition of pairs of charts, each pair satisfying the assumption from above. ρ (t) and cv (t) be defined as above. Let Γσ iρ be the Christoffel symbols for ∇E with Let v ∈ Tpα M and let X v respect to ξ geo β . We put Y v (t) = (ξ geo β )−1X v (t). Moreover, let Φ1 (t) := Φ1 (t, 0, (κgeo β )∗ (v)) = (κgeo β )−1 cv (t) be the geodesic flow on (U geo β , κgeo β ). Then the initial value problem ∇E cv X v ρ = 0 with X v ρ (0) = eρ (pα ) reads σ (0) = (ξ geo Γρ σ + Φi β )−1 eσ (pα ). We denote the corresponding ρ = 0 with Y v iσ Y v in local coordinates as ∂tY v 1 β (cid:16)Pr β )−1 eρ (pα )(cid:17). By (17), Γσ α (x, u) = ξ geo i=1 uρΦλα (x)(1, (ξ geo flow by Φv (t, Y v ρ (0)) and have ξ geo iρ and all its derivatives are uniformly bounded. Moreover, the same is true for the geodesic flow Φ1 since (M , g ) is of bounded geometry. Then by Lemma 5, E together with T geo is bounded in the sense of Definition 34. Conversely, let E be a vector bundle of bounded geometry in the sense of Definition 34. Since ξ geo is a α synchronous trivialization, hρσ = δρσ , see Definition 35 and below. Let now p be any point in M and κ geodesic coordinates on a ball around p with radius r. Let V be a unit radial vector field starting at p. Then its derivatives are uniformly bounded at distances between r 10 and r from p, since (M , g ) is of q ( r bounded geometry. For a point q ∈ M , let vi be n unit vectors that span TqM . We set pi = expM 2 vi ). Let (ei σ (pi ))σ be an orthonormal frame of Epi . We consider geodesic normal coordinates and a synchronous trivialization around those pi . Moreover, let vi be the vector vi parallel transported to pi along cvi . Since the σ (expM transition functions of E and all its derivatives are uniformly bounded, ei pi (−tvi ) and its derivatives are uniformly bounded for t ∈ ( r 10 , r). In particular, uniformly bounded means in this context, that the bound may depend on the order of the derivatives but not on i. Moreover, since the synchronous trivialization is vi ei defined by parallel transport along radial geodesics, we have ∇E σ (q) = 0 for all i and σ . For a synchronous trivialization over q , those equations give a linear system on the Christoffel symbols Γρ lτ , whose coefficients are polynomials in the components of ei σ (q), their first derivatives and vi with respect to the geodesic coordinates around q . But those are uniformly bounded as explained above. Moreover, by construction, this system has a unique solution. Hence, the Christoffel symbols and all its derivatives in the synchronous trivialization around q are uniformly bounded. (cid:3) Example 38. (i) Let (M , g ) be a Riemannian manifold of bounded geometry. Then its tangent bundle equipped with its Levi-Civita connection is trivially of bounded geometry. (ii) Let (M , g ) be a Riemannian spin manifold of bounded geometry with chosen spin structure, i.e. we have chosen a double cover PspinM of the oriented orthonormal frame bundle such that it is compatible to the double covering Spin(n) → SO(n), cf. [Fr00, Section 1.5 and 2.5]. We denote 2 ] the associated spinor bundle, where κ : Spin(n) → U (C[ n 2 ] ) is the spin by S = Pspin (M ) ×κ C[ n representation, cf. [Fr00, Section 2.1]. The connection on S is induced by the Levi-Civita connection on M . Hence, the Riemannian curvature of S and all its covariant derivatives are uniformly bounded. In this spirit, any natural vector bundle E over a manifold of bounded geometry equipped with a geodesic trivialization of E is of bounded geometry. (iii) Let (M , g ) be a Riemannian SpinC manifold of bounded geometry. Here the spinor bundle S described above may not exist globally (but it always exists locally). But a SpinC -structure assures the existence of a SpinC -bundle S ′ that is a hermitian vector bundle of rank 2[ n 2 ] , endowed with a natural scalar product and with a connection ∇S ′ that parallelizes the metric. Moreover, the SpinC -bundle is endowed with a Clifford multiplication denoted by ’·’, where · : T M → EndC (S ′ ) is such that at every point x ∈ M ’·’ defines an irreducible representation of the corresponding Clifford algebra. The 2 ]−1 – denoted by L and called the auxiliary determinant line bundle det S ′ has a root of index 2[ n line bundle associated to the SpinC -structures, [Fr00, Section 2.5]. The square root of L always 1 2 is defined even globally, [Fr00, Appendix D]. The connection on S ′ exists locally but S ′ = S ⊗ L is the twisted connection of the one on the spinor bundle coming from the Levi-Civita connection (as described in (ii)) and a connection on L. Hence, for S ′ being of bounded geometry, we not only 18 need that (M , g ) is a bounded geometry but also that the curvature of the auxiliary line bundle and its covariant derivatives are uniformly bounded. 5.2. Sobolev spaces on vector bundles. We start with two definitions of Sobolev spaces on vector bundles E over M . The first one is for vector bundles of bounded geometry only. where r is the rank of E . < ∞, kϕkp p (M ,E ) = W k ϕp dvolg for ϕ ∈ D(M , E ). Definition 39. [Shu, Section A1.1] Let E with trivialization TE = (U geo α , ξα , hgeo α , κgeo α )α∈I be a vector bundle of bounded geometry over a Riemannian manifold (M n , g ). Then, for s ∈ R, 1 < p < ∞, the Sobolev space Hs p (M , E ) contains all distributions ϕ ∈ D ′ (M , E ) with p (M ,E ) := Xα∈I p (Rn ,Cr )! 1 p α ϕ)kp kξ ∗ α (hgeo kϕkHs H s Let E be a hermitian or Riemannian vector bundle over a complete Riemannian manifold (M , g ) of rank r with fiber product h., .iE and connection ∇E : Γ(T M ) ⊗ Γ(E ) → Γ(E ). In general, ξα and ∇E have nothing to do with each other. But one can alternatively use the connection in order to define Sobolev spaces: For k ∈ N0 , 1 < p < ∞, let the W k p (M , E )-norm be defined by kXi=1 ZM ∇E · · · ∇E {z } i times p (E ) is defined to be the completion of D(M , E ) with respect to the H k Then the space W k p (M , E )-norm. Theorem 40. Let (E , ∇E , h., .iE ) be of bounded geometry. In case that ξα is the synchronous trivialization p (M , E ) = Hk along geodesic normal coordinates W k p (M , E ) for al l k ∈ N0 and 1 < p < ∞. Proof. We briefly sketch the proof which is straightforward. Let ϕ ∈ D(Uα , E Uα ). By induction we have ϕ(cid:17) = Xl≤k α (cid:16)∇E ((∇E )k ϕ)i1 ,...,ik := ξ ∗ eik where the coefficients dj1 ,...,jl are itself polynomials in gij , g ij , Γρ iσ and their derivatives (and depend on i1 , . . . , ik ). Moreover, again by induction, one has that the coefficients of the leading terms, i.e., l = k , are given by dj1 ,...,jk = g i1 j1 · · · g ik jk . By Remark 7.iii, all those coefficients are uniformly bounded. Moreover, using the fact that ξα is obtained by synchronous trivialization, we have hρσ = δρσ , see below Definition 35. Hence, there are constants Ck > 0 with E = Xσ (∇E )k ϕ2 g i1 j1 · · · g ik jk ((∇E )k ϕ)σ j1 ,...,jk ◦ κ−1 i1 ,...,ik ((∇E )k ϕ)σ α ≤ Ck Xγl ;γl ≤k;l∈1,2 for all α and all ϕ ∈ D(Uα , E Uα ). Together with a uniform upper bound on det gij which follows again from Remark 7.iii, we obtain k(∇E )k ϕ(ei1 , . . . , eik )kLp (Uα ,E Uα ) ≤ C Pγ ,γ ≤k kDγ (ξ ∗ α (κα ))kLp (Vα ,Fr ) . On the other hand, by the remark on the leading coefficients d from above α ((∇E )k ϕ(ei1 , . . . , eik )) = ∂m1 · · · ∂mk (ξ ∗ gi1m1 · · · gikmk ξ ∗ α (ϕ)) + terms with lower order derivatives. Thus, as above α (ϕ)) = Xl≤k ∂m1 · · · ∂mk (ξ ∗ i1 ,...,il : Vα → R are again polynomials in gij , g ij , Γσ where the functions d′ iρ and their derivatives. In the same way as before we obtain p (Vα ,Fr ) ≤ C Xl≤k kξ ∗ α (ϕ)kH s 19 d′ i1 ,...,il ξ ∗ α ((∇E )lϕ(ei1 , . . . , eil )). α (ϕ))Dγ2 (ξ ∗ Dγ1 (ξ ∗ α (ϕ)) · · · ∇E ei1 dj1 ,...,jl ∂j1 · · · ∂jl (ξ ∗ α (ϕ)), k(∇E )lϕkLp (Uα ,E Uα ) . Let now ϕ ∈ D(M , E ). Then, using Example 8, the uniform local finiteness of the cover and the local inequalities from above we see that for k ∈ N0 , Lp (M ,E ) = Xα∈I (cid:13)(cid:13)ξ ∗ (cid:13)(cid:13)(∇E )k ϕ(cid:13)(cid:13)p α (∇E )k ϕ)(cid:13)(cid:13)p α (hgeo Lp (Vα ,Fr ) α (∇E )k−i ϕ! (cid:13)(cid:13)(cid:13) α (∇E )k (hgeo kXi=1 (cid:18)k i (cid:19)(∇M )ihgeo = Xα∈I (cid:13)(cid:13)(cid:13)ξ ∗ p α ϕ) − Lp (Vα ,Fr ) Lp (Uα ,E Uα )! . Xα∈I k(∇E )k (hgeo kXi=1 α (∇E )k−iϕkp α ϕ)kp k(∇M )ihgeo Lp (Uα ,E Uα ) + kXi=1 . Xα∈I k(∇E )k−i ϕkp α ϕ)kp k(∇E )k (hgeo Lp (Uα ,E Uα ) + Lp (M ,E ) kXi=1 . Xα∈I k(∇E )k−i ϕkp Lp (M ,E ) . Using this estimate inductively, there is a constant C ′ > 0 with p (M ,E ) ≤ C ′(cid:16) Xα∈I p (Rn ,Fr )(cid:17)1/p α ϕ)kp α (hgeo kξ ∗ kϕkW k H k α ϕ)kp kξ ∗ α (hgeo p (Rn ,Fr ) + H k . On the other hand, kXi=0 Lp (Uα ,E Uα ) . Xα∈I Xα∈I p (Rn ,Fr ) . Xα∈I α ϕ)kp α ϕ)kp k(∇E )i (hgeo kξ ∗ α (hgeo H k iXj=0 kXi=0 Lp (Uα ,E Uα ) . Xα∈I . Xα∈I k(∇E )i−j ϕkp The coincidence of the corresponding spaces follows since D(M , E ) is dense in W k p (M , E ) for k ∈ N, cf. [Str83, Theorem 4.3]. (cid:3) iXj=0 kXi=0 kϕkp p (Uα ,E Uα ) . kϕkp p (M ,E ) . W k W k α (∇E )i−j ϕkp k(∇M )j hgeo Lp (Uα ,E Uα ) The above considerations give rise to the following definition. Definition 41. Let (E , ∇E , h., .iE ) be of bounded geometry. In case that ξα is the synchronous trivialization p (M , E ) := Hs p (E ) := W s along geodesic normal coordinates we set H s p (M , E ) for all s ∈ R and 1 < p < ∞. 5.3. Sobolev norms on vector bundles of bounded geometry via trivializations. As for Sobolev spaces on manifolds we look for ’admissible’ trivializations of a vector bundle E such that the resulting Sobolev norms are equivalent to those obtained when using a geodesic trivialization of E . kϕkH s,TE p Definition 42. Let E be a hermitian or Riemannian vector bundle of rank r over (M n , g ) with a uniformly locally finite trivialization TE = (Uα , κα , ξα , hα )α∈I . Furthermore, let s ∈ R and 1 < p < ∞. Then the space (E ) contains all distributions ϕ ∈ D ′ (M , E ) such that H s,TE p (E ) := Xα∈I p (Rn ,Fr )! 1 p k(ξα )∗ (hαϕ)kp H s Definition 43. Let (E , ∇E , h., .iE ) be a hermitian or Riemannian vector bundle of rank r and of bounded geometry over a Riemannian manifold (M n , g ). Let TE = (Uα , κα , ξα , hα )α∈I be a uniformly locally finite trivialization of E . Using the notations from above, we say that TE is an admissible trivialization for E if the following are fulfilled: (C1) T := (Uα , κα , hα )α∈I is an admissible trivialization of M . 20 is finite. (C2) TE is compatible with the synchronous trivialization along geodesic coordinates, i.e., for Ageo E = (U geo β , ξ geo β , κgeo β )β∈J being a geodesic atlas of E , cf. Definition 35, there are constants Ck > 0 for k ∈ N0 such that for all α ∈ I and β ∈ J with Uα ∩ U geo 6= ∅ and all a ∈ Nn 0 with a ≤ k , β Da µαβ ≤ Ck and Da µβα ≤ Ck . For vector bundles of bounded geometry we have corresponding results as on manifolds of bounded geometry. We start with the formulation of the analog of Theorem 14. The proof follows in the same way. Theorem 44. Let E be a hermitian or Riemannian vector bund le over a Riemannian manifold (M , g ). Let TE = (Uα , κα , ξα , hα )α∈I be an admissible trivialization of E . Furthermore, let s ∈ R and 1 < p < ∞. Then, H s,TE p (E ) = H s p (E ). (18) 5.4. Trace Theorem for vector bundles. Definition 45 (Synchronous trivialization along Fermi coordinates). Let (M , N ) be of bounded geometry, and let E be a hermitian or Riemannian vector bundle of bounded geometry over M . Let T F C = (Uγ , κγ , hγ )γ∈I be a trivialization of M using Fermi coordinates (adapted to N ). We refer to Section 4.1 (also concerning the notation). In case that γ ∈ I \ IN , we trivialize E Uγ via synchronous trivialization along the underlying geodesic coordinates as described in Definition 35. In case that γ ∈ IN , we first trivialize E Uγ ∩N along the underlying geodesic coordinates on N . Then, we trivialize by parallel transport along geodesics emanating at N and being normal to N . The resulting trivialization is denoted by T F C E = (Uγ , κγ , ξγ , hγ )γ∈I . Next, we state corresponding results to Lemma 24 and Theorem 27. Lemma 46. The trivialization T FC E introduced in Definition 45 fulfil ls condition (C2). Proof. The proof is the same as in Lemma 24. (cid:3) Theorem 47. Let E be a hermitian or Riemannian vector bund le of bounded geometry over a Riemannian manifold (M , g ) together with an embedded k-dimensional submanifold N . Let (M , N ) be of bounded geom- etry. If 1 < p < ∞ and s > n−k p , then the pointwise restriction TrN : D(M , E ) → D(N , E N ) extends to a s− n−k linear and bounded operator from H s p (E N ), i.e., p (E ) onto B p,p Moreover, TrN has a linear and bounded right inverse, an extension operator ExM : B TrN H s p (E ) = B s− n−k p p,p (E N ). (19) s− n−k p p,p (E N ) → H s p (E ). Proof. We start with the case that E = Rn × Fr is the trivial bundle over Rn . In this case the claim follows immediately from the Trace Theorem on (Rn , Rk ) and Lemma 2. γ = Uγ ∩ N , κ′ The rest of the proof follows along the lines of Theorem 27, using that by construction (U ′ γ = (κ−1 γ = ξγ (Vγ ∩Rk )×Fr , h′ γ )−1 , ξ ′ γ )γ∈IN gives a geodesic trivialization of E N . γ = hγ U ′ γ U ′ (cid:3) 6. Outlooks 6.1. Spaces with symmetries - a first straightforward example. The aim of this subsection is to give an application of admissible trivializations to spaces with symmetries. We consider manifolds M , where a countable discrete group G acts in a convenient way and show that the Sobolev spaces of functions on the resulting orbit space M /G and the weighted Sobolev spaces of G-invariant functions on M coincide. This is in spirit of Theorem 48. [Tri83, Section 9.2.1] Let 1 < p < ∞ and consider the weight ρ(x) = (1 + x)−κ on Euclidean space Rn where κp > n. Let Tn := Rn/Zn denote the torus and π : Rn → Tn the natural projection. Put H s p,π (Rn , ρ) := {f ∈ D ′ (Rn ) ρf ∈ H s p (Rn ) and f is π-periodic}, then p,π (Rn , ρ). p (Tn ) = H s H s 21 This is just a special case of the theorem given in [Tri83, Section 9.2.1], where more generally Besov and Triebel-Lizorkin spaces are treated, cf. Section 6.2. The proof uses Fourier series. With the help of admis- sible trivializations, we want to present a small generalization of this result for manifolds with G-actions. We start by introducing our setup. In order to avoid any confusion with the metric g , elements of the group G are denoted by h. Definition 49 (G-manifold). Let (M , g ) be a Riemannian manifold, and let G be a countable discrete group that acts freely and properly discontinuously on M . If, additionally, g is invariant under the G-action (which means that h : p ∈ M 7→ h · p ∈ M is an isometry for all h ∈ G), we call (M , g ) a G-manifold. By [Lee01, Corollary 12.27] the orbit space fM := M /G of a G-manifold is again a manifold. From now on we restrict ourselves to the case where fM is closed. Let π : M → fM be the corresponding pro jection. If (M , g ) is a G-manifold, then there is a Riemannian metric g on fM such that π∗ g = g . Let now eT = (Uα , κα , hα )α∈I be an admissible trivialization of fM . In particular, this means we assume that ( fM , g ) is of bounded geometry and, hence, so is (M , g ). Then there are Uα,h ⊂ M with π−1 (Uα ) = ⊔h∈GUα,h and Uα,h = h · Uα,e for all α ∈ I . Here e is the identity element of G. Let πα,h := π Uα,h : Uα,h → Uα denote the corresponding diffeomorphism. Setting κα,h := π−1 α,h ◦ κα : Vα → Uα,h and hα,h := (cid:26)hα ◦ πα,h 0 we have hα,h ◦ κα,h = hα ◦ κα for all α ∈ I , h ∈ G. This way we obtain an admissible trivialization T = (Uα,h , κα,h , hα,h)α∈I ,h∈G of M , which we call G-adapted trivialization. on Uα,h , else, Da (ρ ◦ κα,h) ≤ Ck . Definition 50 (G-adapted weight). Let (M , g ) be a G-manifold with a G-adapted trivialization T as above. A weight function ρ : M → (0, ∞) on M is called G-adapted, if there exist a constant Ck > 0 for all k ∈ N0 such that for a ∈ Nn 0 with a ≤ k and all α ∈ I , Xh∈G Remark 51. The notion of a G-adapted weight is independent on the chosen admissible trivialization on M /G. This follows immediately from the compatibility of two admissible trivializations, cf. Remark 13.ii. Example 52. We give an example of a weight adapted to the G-action. Take a geodesic trivialization on fM as in Example 3 and let T be an admissible trivialization of M constructed from eT on fM as above. There is an injection ι : G → N, since G is countable, and we set X(α,h)∈I×G; p∈Uα,h ι(h)−2 hα,h (p). ρ(p) = Since the covering is locally finite, the summation is always finite. Moreover, Definition 12 and the uniform finiteness of the cover yield for fixed α ∈ I and all a ∈ Nn 0 with a ≤ k (k ∈ N0 ), Xh∈G Da (ρ ◦ κα,h ) ≤ Xh∈G X(α′ ,h′ )∈I×G; ι(h′ )−2 Da (hα′ ,h′ ◦ κα,h) Uα,h ∩U α′ ,h′ 6=∅ ι(h′ )−2 (cid:12)(cid:12)(cid:12)Da (hα′ ◦ κα′ )(cid:12)(cid:12)(cid:12) k Xh∈G Xa′ ≤a X(α′ ,h′ )∈I×G; ′ ≤C ′ α′ ,h′ 6=∅ Uα,h ∩U k Xh∈G X(α′ ,h′ )∈I×G; k Xh∈G X(α′ ,h′ )∈I×G, ≤C ′′ ι(h′ )−2 = C ′′ ι(h′ )−2 Uα,h∩U α′ ,h′ 6=∅ Uα,e∩U α′ ,h−1 h′ 6=∅ k Xh∈G X(α′ ,h′ )∈I×G, k L Xh∈G k L Xi∈N =C ′′ ι(hh′ )−2 ≤ C ′′ ι(h)−2 ≤ C ′′ Uα,e∩U α′ ,h′ 6=∅ i−2 < ∞, 22 where L is the multiplicity of the cover and the constants C ′ k , C ′′ k do not depend on h ∈ G and α ∈ I . In particular, together with Remark 51, this example demonstrates that each G-manifold admits a G-adapted weight. We fix some more notation. Let s ∈ R and 1 < p < ∞. Then the space H s p (M , ρ) consists of all distributions f ∈ D ′ (M ) such that k(ρ ◦ κα,h ) ((hα f ′ ) ◦ κα ) kp H s p (Rn ) kf kH s p (M ,ρ) := kρf kH s p (M ) < ∞. Moreover, we call a distribution f ∈ D ′ (M ) G-invariant, if f (ϕ) = f (h∗ϕ) holds for all ϕ ∈ D(M ) and h ∈ G. The space of all G-invariant distributions in H s p (M , ρ) is denoted by H s p (M , ρ)G . Theorem 53. Let (M , g ) be a G-manifold of bounded geometry where fM = M /G is closed, and let ρ be a G-adapted weight on M . Furthermore, let s ∈ R and 1 < p < ∞. Then p ( fM ) = H s H s p (M , ρ)G . Proof. It suffices to show that the norms of the corresponding spaces are equivalent. We work with a geodesic trivialization eT geo of (M /G, g) and a G-adapted trivialization T of M constructed from T geo as described above. Note that the closedness of M /G implies, that ρ∪αUα,e ≥ c > 0 for some constant c > 0 (since then ∪αUα,e is compact). Let f ′ ∈ H s p (M /G) and set f = f ′ ◦ π . Then, p (M ,ρ) = Xα∈I ,h∈G p (Rn ) = Xα∈I ,h∈G kf kp k(hα,hρf ) ◦ κα,h kp H s H s . Xα∈I k(hαf ′ ) ◦ καkp p (Rn ) = kf ′kp p (M/G) . H s H s p (M , ρ)G . Since f is G-invariant, there is a unique f ′ with f = f ′ ◦ π . Then, Let now f ∈ H s p (M/G) = Xα∈I p (Rn ) = Xα∈I 1 ◦ κα,e ) ((hα,e ρf ) ◦ κα,e ) kp kf ′kp k(hαf ′ ) ◦ καkp k( H s p (Rn ) H s H s ρ . Xα∈I p (Rn ) ≤ Xα∈I ,h∈G k ((hα,eρf ) ◦ κα,e ) kp H s Here we used the uniform boundedness of 1 ρ ◦ κα,e and its derivatives, which follows from the corresponding statement for ρ ◦ κα,e and the lower bound ρ ◦ κα,e ≥ c > 0. (cid:3) Remark 54. The restriction to closed manifolds fM (i.e., compact manifolds without boundary) in Theorem 53 should not be necessary. In case that fM is noncompact, one needs to modify the definition of G-adapted weights in a suitable way to assure the weight is bounded away from zero with respect to the ’noncompact directions’ of fM . We conclude our considerations with an example of a G-manifold other than the torus, which is covered by Theorem 53. Example 55. Let ( fM , g) be a closed manifold. Let G be a subgroup of the fundamental group π1 ( fM ) of fM . Note that G is countable since π1 ( fM ) is. Let (M , g ) be the G-cover of ( fM , g ) where g = π∗ g . Then, (M , g ) is a G-manifold with fM = M /G. 6.2. Triebel-Lizorkin spaces on manifolds. In order to keep our considerations as easy as possible, we have been concentrating on (fractional) Sobolev spaces on Riemannian manifolds M so far. This last para- graph is aimed at the reader who is more interested in the general theory of Besov and Triebel-Lizorkin spaces – also referred to as B- and F-spaces in the sequel. We now want to sketch how those previous results generalize to Triebel-Lizorkin spaces on manifolds. k ((hα,hρf ) ◦ κα,h ) kp p (Rn ) = kf kp p (M ,ρ) . H s H s 23 By the Fourier-analytical approach, Triebel-Lizorkin spaces F s p,q (Rn ), s ∈ R, 0 < p < ∞, 0 < q ≤ ∞, consist of all distributions f ∈ S ′ (Rn ) such that p,q (Rn ) = (cid:13)(cid:13)(cid:13)(cid:16) ∞Xj=0 (cid:12)(cid:12)2js (ϕj bf )∨ (·)(cid:12)(cid:12)q (cid:17)1/q (cid:13)(cid:13)(cid:13)Lp (Rn ) (cid:13)(cid:13)f (cid:13)(cid:13)F s (usual modification if q = ∞) is finite. Here {ϕj }∞ j=0 denotes a smooth dyadic resolution of unity, where ϕ0 = ϕ ∈ S (Rn ) with supp ϕ ⊂ {y ∈ Rn : y < 2} x ≤ 1, and for each j ∈ N put and ϕ(x) = 1 if ϕj (x) = ϕ(2−j x) − ϕ(2−j+1 x). The scale F s p,q (Rn ) generalizes fractional Sobolev spaces. In particular, we have the coincidence (20) p,2 (Rn ) = H s F s p (Rn ), 1 < p < ∞, s ∈ R, [Tri83, p. 51]. In general, Besov spaces on Rn are defined in the same way by interchanging the order cf. in which the ℓq - and Lp -norms are taken in (20). Hence, the Besov space B s p,q (Rn ), s ∈ R, 0 < p, q ≤ ∞ consists of all distributions f ∈ S ′ (Rn ) such that Lp (Rn )(cid:17)1/q p,q (Rn ) = (cid:16) ∞Xj=0 2jsq (cid:13)(cid:13)(ϕj bf )∨(cid:13)(cid:13)q (cid:13)(cid:13)f (cid:13)(cid:13)B s (usual modification if p = ∞ and/or q = ∞) is finite. In particular, if p = q , p,p (Rn ) = F s B s p,p (Rn ), 0 < p < ∞, and we extend this to p = ∞ by putting F s p,q (Rn ) and B s ∞,∞ (Rn ). The scales F s ∞,∞ (Rn ) := B s p,q (Rn ) were studied in detail in [Tri83, Tri92], where the reader may also find further references to the literature. (21) On Rn one usually gives priority to Besov spaces, and they are mostly considered to be the simpler ones compared to Triebel-Lizorkin spaces. However, the situation is different on manifolds M , since B-spaces lack the so-called localization principle, cf. [Tri92, Theorem 2.4.7(i)], which is used to define F-spaces on M (as p,q (Rn ) inside p (Rn ) by F s was already done in Definition 11 for fractional Sobolev spaces, now replacing H s of the norm). Then Besov spaces on M are introduced via real interpolation of Triebel-Lizorkin spaces (in order to compute traces we have to generalize the B-spaces on M from Definition 16 and allow 0 < p ≤ 1). Definition 56. Let (M n , g ) be a Riemannian manifold with an admissible trivialization T = (Uα , κα , hα )α∈I and let s ∈ R. (i) Let either 0 < p < ∞, 0 < q ≤ ∞ or p = q = ∞. Then the space F s,T p,q (M ) contains all distributions f ∈ D ′ (M ) such that p,q (M ) := Xα∈I p,q (Rn )! 1 p k(hαf ) ◦ καkp kf kF s,T F s is finite (with the usual modification if p = ∞). (ii) Let 0 < p, q ≤ ∞, and let −∞ < s0 < s < s1 < ∞. Then p,q (M ) = (cid:0)F s0 ,T p,p (M )(cid:1)Θ,q p,p (M ), F s1 ,T B s,T with s = (1 − Θ)s0 + Θs1 . Remark 57. Restricting ourselves to geodesic trivializations T geo , the spaces from Definition 56 coincide with the spaces F s p,q (M ), introduced in [Tri92, Definition 7.2.2, 7.3.1]. The space B s,T p,q (M ) and B s p,q (M ) is independent of the chosen numbers s0 , s1 ∈ R and, furthermore, for s ∈ R and 0 < p ≤ ∞ we have the coincidence (22) p,p (M ) = F s,T B s,T (23) p,p (M ). This follows from [Tri92, Theorem 7.3.1], since the arguments presented there are based on interpolation and completely oblivious of the chosen trivialization T . In particular, (23) yields that for f ∈ B s p,p (M ) a quasi-norm is given by p,p (M ) = Xα∈I p,p (Rn )! 1 p k(hαf ) ◦ καkp kf kB s,T B s (24) . 24 Now we can transfer Theorem 14 to F- and B-spaces. Theorem 58. Let (M n , g ) be a Riemannian manifold with an admissible trivialization T = (Uα , κα , hα )α∈I . Furthermore, let s ∈ R and let 0 < p, q ≤ ∞ (0 < p, q < ∞ or p = q = ∞ for F-spaces). Then p,q (M ) = F s F s,T p,q (M ) and p,q (M ) = B s B s,T p,q (M ). Proof. For F-spaces the proof is the same as the one of Theorem 14. The claim for B-spaces then follows from Definition 56.ii. (cid:3) Trace theorem. The generalization of the Trace Theorem 27 is stated below. improves [Skr90, Theorem 1, Corollary 1]. In particular, this result Theorem 59. Let (M n , g ) be a Riemannian manifold together with an embedded k-dimensional submanifold N and (M n , N k ) be of bounded geometry. Furthermore, let 0 < p, q ≤ ∞ (0 < p, q < ∞ or p = q = ∞ for F-spaces) and let > k (cid:18) 1 − 1(cid:19)+ p Then TrN = Tr is a linear and bounded operator from F s p,q (M ) onto B s− n−k p p,q (N ) and B s p,q (M ) onto n − k p B (N ), respectively, i.e., s− n−k p p,p s − . (25) TrN F s p,q (M ) = B s− n−k p p,p (N ) and TrN B s p,q (M ) = B s− n−k p p,q (N ). (26) Proof. The proof of (26) runs along the same lines as the proof of Theorem 27. Choosing Fermi coordinates, via pull back and localization the problem can be reduced to corresponding trace results in Rn on hyperplanes Rk , cf. [Tri92, Theorem 4.4.2], where the proof for k = n−1 may be found. The result for general hyperplanes – and condition (25) – follows by iteration. The assertion for B-spaces follows then from Definition 56.ii. (cid:3) References [Aub76] T. Aubin. Espaces de Sobolev sur les vari´et´es Riemanniennes. Bul l. Sci. Math. 100: 149–173, 1976. [Aub82] T. Aubin. Nonlinear Analysis on Manifolds. Monge-Amp`ere Equations, New York, Springer, 1982. [Eic07] J. Eichhorn. Global analysis on open manifolds. Nova Science Publishers, Inc., New York x+644, 2007. [Eic91] J. Eichhorn. The boundedness of connection coefficients and their derivatives. Math. Nachr. 152: 144–158, 1991. [Fr00] T. Friedrich. Dirac operators in Riemannian geometry. Graduate Studies in Mathematics, American Mathematical Society, Providence, RI 25: xvi+195, 2000. [Gro12] N. Grosse. Solutions of the equation of a spinorial Yamabe-type problem on manifolds of bounded geometry. Comm. Part. Diff. Eq. 37(1): 58–76 2012. [GN12] N. Grosse, R. Nakad. Boundary value problems for noncompact boundaries of Spinc manifolds and spectral estimates. arXiv:1207.4568 [Lee01] J.M. Lee. Introduction to topological manifolds. Second edition, Graduate Texts in Mathematics (202), Springer, New York, 2011. [Sch01] T. Schick. Manifolds with boundary and of bounded geometry. Math. Nachr., 223:103–120, 2001. [Shu] M.A. Shubin, Spectral theory of elliptic operators on noncompact manifolds. M´ethodes semi-classiques, Vol. 1 (Nantes, 1991), Ast´erisque 207:35–108 1992. [Skr90] L. Skrzypczak. Traces of Function Spaces of F s p,q –Bs p,q Type on Submanifolds. Math. Nachr. 46: 137–147, 1990. [Skr98] L. Skrzypczak. Atomic decompositions on manifolds with bounded geometry. Forum Math. 10: 19–38, 1998. [SpIV] M. Spivak. A comprehensive introduction to differential geometry, Vol.IV. Second edition, Publish or Perish Inc. Wilm- ington, Del. viii+561, 1979. [Str83] R.S. Strichartz. Analysis of the Laplacian on the complete Riemannian manifold. J. Funct. Anal. 52: 48–79, 1983. [Tr86] H. Triebel. Spaces of Besov-Hardy-Sobolev type on complete Riemannian manifolds. Ark. Mat. 24(2): 299–337, 1986. [Tri83] H. Triebel. Theory of function spaces, volume 78 of Monographs in Mathematics. Birkhauser Verlag, Basel, 1983. [Tri92] H. Triebel. Theory of function spaces II, volume 84 of Monographs in Mathematics. Birkhauser Verlag, Basel, 1992. [Tri78] H. Triebel. Interpolation theory, function spaces, differential operators, volume 18. North-Holland Publishing Co., Amsterdam, 1978. [Wa66] F.W. Warner. Extensions of the Rauch comparison theorem to submanifolds. Trans. Amer. Math. Soc. 122: 341–356, 1966. 25 Nadine Grosse Mathematical Institute University of Leipzig Augustusplatz 10 04109 Leipzig Germany Cornelia Schneider Applied Mathematics III University of Erlangen–Nuremberg Cauerstrasse 11 91058 Erlangen Germany [email protected] [email protected] 26
1205.4394
3
1205
2012-07-31T16:17:34
Lax-Halmos Type Theorems in H^p Spaces
[ "math.FA" ]
In this paper we characterize for 0 < p \leq \infty, the closed subspaces of Hp that are invariant under multiplication by all powers of a finite Blaschke factor B, except the first power. Our result clearly generalizes the invariant subspace theorem obtained by Paulsen and Singh [9] which has proved to be the starting point of important work on constrained Nevanlinna-Pick interpolation. Our method of proof can also be readily adapted to the case where the subspace is invariant under all positive powers of B (z). The two results are in the mould of the classical Lax-Halmos Theorem and can be said to be Lax-Halmos type results in the finitre multiplicity case for two commuting shifts and for a single shift respectively.
math.FA
math
Lax-Halmos Type Theorems On H p Spaces Niteesh Sahni and Dinesh Singh Abstract. In this paper we characterize for 0 < p ≤ ∞, the closed subspaces of H p that are invariant under multiplication by all powers of a finite Blaschke factor B, except the first power. Our result clearly generalizes the invariant subspace theorem obtained by Paulsen and Singh [9] which has proved to be the starting point of important work on constrained Nevanlinna-Pick interpolation. Our method of proof can also be readily adapted to the case where the subspace is invariant under all positive powers of B (z) . The two results are in the mould of the classical Lax-Halmos Theorem and can be said to be Lax-Halmos type results in the finitre multiplicity case for two commuting shifts and for a single shift respectively. 1. INTRODUCTION In recent times a great deal of interest has been generated in the Banach algebra H ∞ 1 and subsequently also in a class of related algebras in the context of problems dealing with invariant subspaces and their use in solving Nevanlinna-Pick type interpolation problems for these algebras. We refer to [3], [4], [11], [15], [18], and [21]. Note that H ∞ 1 = {f (z) ∈ H ∞ : f ′(0) = 0} is a closed subalgebra of H ∞, the Banach algebra of bounded analytic functions on the open unit disc. The starting point, in the sequence of papers cited above, is an invariant subspace theorem first proved by Paulsen and Singh in a special case [18, Theorem 4.3] and subsequently in more general forms in [4] and [17]. This invariant subspace result is crucial to the solutions of interpolation problems of the Pick-Nevanlinna type as presented in the papers cited above. This theorem characterizes the closed subspaces of the Hardy spaces that are left invariant under the action of every element of the algebra H ∞ 1 . In this paper we present a far reaching generalization of this invariant subspace theorem by presenting a complete characterization of the invariant subspaces of the Banach algebra H ∞ 1 ; B is a finite Blaschke product}. In the special case where B(z) = z we arrive at the first or original invariant subspace theorem mentioned above for the Banach algebra H ∞ 1 (B) stands for the closed subalgebra of H ∞ generated by B2 and B3. Our paper also deals with a second and related problem of characterizing the invariant subspaces on H p of the algebra H ∞(B) which consists of the Banach algebra generated byB. For the case p = 2, this problem has been tackled in the far more general setting 1 (B) = {f (B(z)) : f ∈ H ∞ 1 . We also note that H ∞ 2000 Mathematics Subject Classification. Primary 05C38, 15A15; Secondary 05A15, 15A18. Key words and phrases. Finite Blaschke factor B, subalgebra of H∞ generated by B2 and B3, invariant subspace, H p. 1 2 NITEESH SAHNI AND DINESH SINGH of de Branges spaces in [19] and in the same year, for the classical H pspaces for all values of p ≥ 1, this problem has been tackled in [13]. However, for this second problem, we claim some novelty and completeness on two counts; for one we have shown that in the case when 0 < p < 1 we have an explicit description of the invariant subspaces, and second, our proof is more elementary and different from that in [13] since we do not use their general inner-outer factorization theorem [13, page 112]. Finally, we wish to observe that the two main results presented in this paper can be interpreted as being in the mould of the classical Lax-Halmos Theorem in the case of finite multiplicity, see [8],[9], and [14], for two commuting shifts as represented by multiplication by B2 and by B3 and for a single shift represented by multiplication by B except that unlike the classical versions of the Lax-Halmos theorem we work entirely in the scalar valued setting of the classical Hardy spaces and our characterisations are also inside this scalar setting. Let D denote the open unit disk, and let its boundary, the unit circle, be denoted by T. The Lebesgue space Lp on the unit circle is the collection of complex valued functions f on the unit circle such that R f pdm is finite, where dm is the normalized Lebesgue measure on T. The Hardy space H p is the closure in Lp of the analytic polynomials. For p ≥ 1, H p can be viewed as the the following closed subspace of Lp: (cid:26)f ∈ Lp :Z f zndm = 0 for all n ≥ 1(cid:27) . For 1 ≤ p < ∞, H p is a Banach space under the norm kf kp =(cid:18)Z f pdm(cid:19) 1 p . H ∞ is a Banach algebra under the essential supremum norm. The Hardy space H 2 turns out to be a Hilbert space under the inner product < f, g >=Z f gdm. For a detailed account of H p spaces the reader can refer to [5], [7], [10], and [12]. By a finite Blaschke factor B(z) we mean B(z) = z − αj 1 − αj z , n Yj=1 where αj ∈ D. Throughout we shall assume α1 = 0 as this does not affect generality. The operator of multiplication by B(z) denoted by TB is an isometric operator on H p. We call a closed subspace M of H p to be B -invariant if TBM ⊂ M . By an invariant subspace M (in H p) of any subalgebra A of H ∞ we mean a closed subspace of H p such that f g ∈ M for all f in A and for all g ∈ M . Let Bj(z) denote the product of the first j factors in B(z): Bj(z) = j Yi=1 z − αi 1 − αiz , αj ∈ D. LAX-HALMOS TYPE THEOREMS ON H p SPACES 3 2. Preliminary Results Note: We shall assume throughout this paper that the Blaschke product B so indicated is fixed and has n zeros that may not necessarily be distinct. Theorem 1. ([19, Theorem 3.3]) The collection {ejm = (1 − αj+12) 2 (1 − αj+1z)−1BjBm : 0 ≤ j ≤ n − 1, m = 0, 1, 2, . . .} is an orthonormal basis for H 2. j=0 ⊕ej0H 2(B) where H 2(B) stands for the closed linear 1 Consequently H 2 = Pn−1 span of {Bm : m = 0, 1, 2, ...} in H 2. Let (ϕ1, . . . , ϕr) be an r tuple of H ∞ functions (r ≤ n). Suppose each ϕj has the representation n−1 The matrix A = (ϕij )n×r is called the B− matrix of (ϕ1, . . . , ϕr). The matrix A is called B-inner if A∗A = I. Suppose an H ∞ function ψ has the representation ϕj = ei0ϕij. Xi=0 ψ = n−1 Xj=0 ej0ψj, for some ψ0, . . . , ψn−1 ∈ H 2(B), then ψ is called B-inner if n−1 ψj 2 = 1 a.e. Xj=0 It has been proved in [19] that ψ is B-inner if and only if {Bmψ : m = 0, 1, ...} is an orthonormal set in H 2. Theorem 2. ([19, Theorem 4.1]) Let M be a B-invariant subspace of H 2. Then there is an r ≤ n such that M = ϕ1H 2(B) ⊕ ϕ2H 2(B) ⊕ · · · ⊕ ϕrH 2(B) for some B-inner functions ϕ1, . . . , ϕr, and the B-matrix of (ϕ1, . . . , ϕr) is B-inner. Further, this representation is unique in the sense that if M = ψ1H 2(B) ⊕ ψ2H 2(B) ⊕ · · · ⊕ ψsH 2(B) r then r = s, and ϕi = αij ψj for scalars αij such that the matrix (αij ) is unitary. Pj=1 there exist constants Ci,p, i = 1, . . . , k such that kfikp ≤ Ci,p kf kp. Lemma 2. For 1 ≤ p < 2, we can write H p = e00H p (B) ⊕ e10H p (B) ⊕ · · · ⊕ en−1,0H p (B) , where each ej0 is as in Theorem 1 for each j. We shall also make use of the following results to establish certain key facts that are central to the proof of the main results. Lemma 1. ([13, Proposition 3]) Let 1 ≤ p ≤ ∞ and let ϕ1, . . . , ϕk (k ≤ n) be B-inner functions. Then for any f ∈ H ∞ such that f (z) = ϕi (z) fi (B (z)), k Pi=1 4 NITEESH SAHNI AND DINESH SINGH Proof. It is trivial to note that e00H p (B) ⊕ e10H p (B) ⊕ · · · ⊕ en−1,0H p (B) ⊂ H p. To establish the opposite inclusion take an arbitrary element f ∈ H p. Since H ∞ is dense in H p, there exists a sequence {fk} of H ∞ functions that converges to f in the norm of H p. In view of Theorem 1 we can write (2.1) fk = e00f (1) k + e10f (2) k + · · · + en−1,0f (n) k , where f (j) the estimate k ∈ H 2 (B) for all j = 1, ..., n. By Lemma 1, we have for all j = 1, . . . n, (2.2) ≤ Dkj kfkkp (cid:13)(cid:13)(cid:13) f (j) k (cid:13)(cid:13)(cid:13)p for some constants Dkj. Equation (2.2) implies thatnf (j) all j = 1, . . . , n, and hence f (j) k → f (j) in H p (B). Therefore fk → e00f (1)+e10f (2)+ · · · + en−1,0f (n) as k → ∞ in H p. Hence f = e00f (1) + e10f (2) + · · · + en−1,0f (n). This completes the proof of the assertion. (cid:3) k o is a Cauchy sequence for Lemma 3. ([6, Lemma 4.1]) Suppose {fn}∞ n=1 is a sequence of H p functions, p > 2, which converges to an H p function f in the H 2 norm. Then there exists a sequence {gn}∞ n=1 of H ∞ functions such that gnfn → f in the H p norm (weak-star convergence when p = ∞). Further the sequence {gn}∞ n=1 is uniformly bounded, and converges to the constant function 1 a.e. Lemma 4. Let p > 2. Suppose an H p function f is of the form f = ϕ1h1 + · · ·+ϕnhn, where h1, . . . , hn ∈ H 2 (B), and ϕ1, . . . , ϕn are B-inner , then h1, . . . , hn belong to H p (B). Proof. Let f = ϕ1h1 + ϕ2h2 + · · · + ϕrhr, where h1, h2, . . . , hr ∈ H 2 (B), and ϕ1, . . . , ϕn are B− inner. Also f can be identi- fied with a bounded linear functional Ff ∈ L∗ q (1 ≤ q < 2) such that Ff (g) =Z f g and Ff (g) ≤ δ kgkq for all g ∈ Lq for some δ > 0. Now for any l ∈ span(cid:8)1, B, B2, ...(cid:9) such that l is a polynomial in B, we have (cid:12)(cid:12)(cid:12)(cid:12) Z f ϕ1l(cid:12)(cid:12)(cid:12)(cid:12) ≤ δ kϕ1lkq ≤ δ1 klkq LAX-HALMOS TYPE THEOREMS ON H p SPACES 5 for some δ1 > 0. Also note that Z f ϕ1l(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) Z (ϕ1h1 + ϕ2h2 + · · · + ϕrhr) ϕ1l(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) Z ϕ1h1ϕ1l +Z ϕ2h2ϕ1l + · · · +Z ϕrhr ϕ1l(cid:12)(cid:12)(cid:12)(cid:12) Z ϕ1h1ϕ1l(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) Z h1l(cid:12)(cid:12)(cid:12)(cid:12) . Here R ϕjhjϕ1l = 0, j = 2, ..., r because ϕ1H 2 (B) ⊥ ϕj H 2 (B), and R ϕ1h1ϕ1l = R h1l because ϕ1 is B− inner. Therefore, Z h1l(cid:12)(cid:12)(cid:12)(cid:12) Now any analytic polynomial k ∈ Lq can be written as ≤ δ1 klkq . (cid:12)(cid:12)(cid:12)(cid:12) k = e00k1 (B) + · · · + er−10kr (B) and (2.3) (cid:12)(cid:12)(cid:12)(cid:12) Z h1k(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) Z h1k1 +Z e10h1k2 + · · · +Z er−10h1kr(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:12)(cid:12)(cid:12)(cid:12) Z h1k1(cid:12)(cid:12)(cid:12)(cid:12) Z er−10h1kr(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) Z h1k1(cid:12)(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12)(cid:12) Z e10h1k2(cid:12)(cid:12)(cid:12)(cid:12) + · · · +(cid:12)(cid:12)(cid:12)(cid:12) . It is easily checked that all integrals in the above equation except the first integral shall be zero. We show this by examining one of the above integrals in question: Let h1 = α0 + α1B + α2B2 + · · · , k2 = β0 + β1B + β2B2 + · · · . Now, Z e10h1k2 = (cid:10)e10h1, k2(cid:11) z = (cid:28) = 0. 1 − α2z (cid:0)α0 + α1B + α2B2 + · · ·(cid:1) , β0 + β1B + β2B 2 + · · ·(cid:29) 6 NITEESH SAHNI AND DINESH SINGH Let k1 = γ0 + γ1B + γ2B2 + · · · so that (cid:12)(cid:12)(cid:12)(cid:12) Z h1k1(cid:12)(cid:12)(cid:12)(cid:12) = α0 γ0 = (cid:12)(cid:12)(cid:10)h1, k1(cid:11)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) Z h1(cid:12)(cid:12)(cid:12)(cid:12) Z k1(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) Z h1(cid:12)(cid:12)(cid:12)(cid:12) Z k(cid:12)(cid:12)(cid:12)(cid:12) ≤ AZ k ≤ A kkkq polynomials, and so it can be extended to a bounded linear functional on Lq. Call q implies that there where A = R h1. Thus h1acts as a bounded linear functional on the space of this extension as F . So F (g) =R h1g for all g ∈ Lq. But F ∈ L∗ exists G ∈ Lp such that F (g) =R Gg for all g ∈ Lq. In particular we have Z (G − h1) zn = 0 for all n ∈ Z. Hence h1 = G ∈ Lp. In a similar fashion we get h2, ..., hr ∈ Lp. (cid:3) 3. The B2 and B3 invariant subspaces of H p Theorem 3. Let M be a closed subspace of H p, 0 < p ≤ ∞, such that M is 1 (B) but not invariant under H ∞ (B). Then there exist B− invariant under H ∞ inner functions J1, . . . , Jr (r ≤ n) such that k M = Xj=1  ⊕ hϕji  ⊕ ⊕B2JlH p (B) r Xl=1 where k ≤ 2r − 1, and for all j = 1, 2, ..., k, ϕj = (α1j + α2jB)J1 + (α3j + α4jB)J2 + ... + (α2r−1,j + α2r,jB)Jr. Remark 1. The proof shall also show that the matrix A = (αij )2r×k satisfies A∗A = I, and αst 6= 0 for some (s, t) ∈ {1, 3, . . . , 2r − 1} × {1, 2, . . . , k}. Also, when 0 < p < 1, the right hand side should be interpreted as being the closure of the sum in the H p metric. Proof. The case p = 2. Using the initial line of argument as in [4] we define It is easily seen that M1 is a B−invariant subspace of H 2. M1 = H ∞ (B) · M . Observe that (3.1) Therefore (3.2) M1 ⊃ M ⊃ H ∞ 1 (B) · M ⊃ B2H ∞ (B) · M = B2M1. B2M1 ⊂ M ⊂ M1 Note that all containments in the above equation are strict. For if M = M1 or M = B2M1 it would then mean that M is invariant under H ∞ (B), which is a LAX-HALMOS TYPE THEOREMS ON H p SPACES 7 contradiction. By Theorem 2, there exist B− inner functions J1, . . . , Jr (r ≤ n) such that M1 = J1H 2 (B) ⊕ · · · ⊕ JrH 2 (B) = (cid:0)hJ1i ⊕ hBJ1i ⊕ B2J1H 2(B)(cid:1) ⊕ · · · ⊕(cid:0)hJri ⊕ hBJri ⊕ B2JrH 2(B)(cid:1) = (hJ1i ⊕ hBJ1i ⊕ · · · ⊕ hJri ⊕ hBJri) ⊕ B2M1. So M1 ⊖ B2M1 has dimension 2r, and hence M ⊖ B2M1 has dimension k, where k ≤ 2r − 1. Let ϕ1, . . . , ϕk be an orthonormal basis for M ⊖ B2M1. Now M = (cid:2)M ⊖ B2M1(cid:3) ⊕ B2M1 k ⊕ hϕj i =   Xj=1  ⊕ B2J1H 2 (B) ⊕ · · · ⊕ B2JrH 2 (B) . Since ϕj ∈ M ⊖ B2M1 ⊂ M1 ⊖ B2M1, we see that each ϕj is of the form α1jJ1 + α2jJ1B + α3jJ2 + α4jJ2B + · · · + α2r−1,jJr + α2r,jJrB. The conditions kϕjk2 2 = 1 and hϕj, ϕii = 0 for j 6= i, imply that α1j2 + α2j2 + · · · + α2r,j2 = 1, and α1jα1i + α2j α2i + · · · + α2r,jα2r,i = 0. In addition the k tuples (αi1, αi2, . . . , αik), i = 1, 3, . . . , 2r − 1 cannot be simultaneously zero, otherwise M would become B− invariant. The case 0 < p < 1. Observe that every H p function f , can be written as f = IO, where I is an inner function and O ∈ H p is an outer function. Choose n such that 2np > 2, so that we can express f as a product of H 2 functions: f = IO 1 2n O 1 2n · · · O 1 2n . We first show that M ∩ H 2 6= [0]. Let 0 6= f ∈ M . Then f can be written as where f1, f2,. . . ,fm ∈ H 2. In view of Theorem 1, we can express each fl as f = f1f2 · · · fm, fl = e00g(l) 1 + · · · + er0g(l) r , for some g(l) 1 , . . . , g(l) for each g(l) j , there exists k(l) j ∈ H 2such that g(l) j = k(l) r ∈ H 2(cid:0)B2(cid:1). It is known that the operator T : H 2 −→ H 2 defined by T h = h(cid:0)B2 (z)(cid:1) is an isometry, and its range is H 2(cid:0)B2(cid:1) (see [2]). So qjl (z) := exp  j (cid:0)B2 (z)(cid:1). Define (z)(cid:12)(cid:12)(cid:12) Here ∼ denotes the harmonic conjugate. Then qjl (z) ≤ 1, and thus hl (z) := q1l (z) q2l (z) · · · qrl (z) ∈ H ∞. ^ k(l) j (z)(cid:12)(cid:12)(cid:12)   −(cid:12)(cid:12)(cid:12) k(l) j 1 2 1 2 − i 2 (cid:12)(cid:12)(cid:12) . Note that hl(cid:0)B2(z)(cid:1) fl (z) = e00hl(cid:0)B2(z)(cid:1) g(l) = e00hl(cid:0)B2(z)(cid:1) k(l) r (cid:0)B2 (z)(cid:1) , which clearly belongs to H ∞. This implies that h1(cid:0)B2 (z)(cid:1) · · · hm(cid:0)B2 (z)(cid:1) f = h1(cid:0)B2 (z)(cid:1) f1 · · · hm(cid:0)B2 (z)(cid:1) fm ∈ H ∞. Since h1(cid:0)B2 (z)(cid:1) · · · hm(cid:0)B2 (z)(cid:1) ∈ H ∞ its Cesaro means (cid:8)pn(cid:0)B2(cid:1)(cid:9), which is a sequence of polynomials, shall converge 1 (z) + · · · + er0hl(cid:0)B2(z)(cid:1) g(l) 1 (cid:0)B2 (z)(cid:1) + · · · + er0hl(cid:0)B2(z)(cid:1) k(l) r (z) 8 NITEESH SAHNI AND DINESH SINGH to h1(cid:0)B2 (z)(cid:1) · · · hm(cid:0)B2 (z)(cid:1) a.e.. Hence, by the Dominated Convergence The- orem, we see that pn(cid:0)B2(cid:1) f → h1(cid:0)B2 (z)(cid:1) · · · hm(cid:0)B2 (z)(cid:1) f in H p. Therefore, h1(cid:0)B2 (z)(cid:1) · · · hm(cid:0)B2 (z)(cid:1) f ∈ M , because M is invariant under B2. This estab- lishes M ∩ H 2 6= {0}. Next we claim that M ∩ H 2 is dense in M . The density will also imply that M ∩ H 2 is not B− invariant, otherwise it would force M to be B− invariant, which is not possible. It is trivial to note that M ∩ H 2 ⊆ M (bar denotes closure in H p). Let f ∈ M . We can express f as where each fl ∈ H 2, and m is chosen so that 2mp > 2. As argued previously we can express each fl as f = f1f2 · · · f2m , for certain k(l) 1 , . . . , k(l) r ∈ H 2. Define fl = e00k(l) 1 (cid:0)B2 (z)(cid:1) + · · · + er0k(l) q(jl) n (z) = exp  −k(l) j (z) 1 2 − i n 1 2 r (cid:0)B2 (z)(cid:1) ,   ^ j (z) k(l) and (cid:12)(cid:12)(cid:12) q(jl) n (cid:12)(cid:12)(cid:12) Since h(1) so that Moreover (∼ denotes the harmonic conjugate which exists for L2 functions). Then q(jl) ≤ 1. For each l = 1, ..., 2m, the function h(l) n (z) = q(1l) n (z) · · · q(rl) n n ∈ H ∞ (z) belongs to H ∞ and h(l) n (cid:0)B2 (z)(cid:1) multiplies fl into H ∞. This implies that (B2 (z))f ∈ H ∞ h(1) n (B2 (z)) · · · h(2m) n (B2 (z)) · · · h(2m) n h(1) n (B2 (z)) · · · h(2m) n n (B2 (z)) → 1 a.e., we have (B2 (z))f → f a.e. h(1) n (B2 (z)) · · · h(2m) n h(1) n (B2 (z)) · · · h(2m) n (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) p (B2 (z))f − f(cid:12)(cid:12)(cid:12) (B2 (z))f − f(cid:12)(cid:12)(cid:12) p → 0 a.e ≤ 2p f p so by the Dominated Convergence Theorem, n (B2 (z)) · · · h(2m) h(1) n (B2 (z))f → f ∈ H p. We claim that h(1) follows. For each n, there exists a sequence of polynomials {p(n) n (B2) · · · h(2m) (B2)f ∈ M . To prove this claim we proceed as n k } such that p(n) k (B2(z)) → h(1) n (B2 (z)) · · · h(2m) n ...h(2m) n n (B2 (z)) k } is the Cesaro means of h(1) The sequence {p(n) and it converges boundedly It is then easy to see by means of the Dominated Convergence and pointwise. k (B2)f converges to h(1) Theorem that p(n) (B2)f in H p. The claim now follows in view of the fact that M is invariant under B2 and the fact that p(n) k (B2)f ∈ M . By the validity of the result for the case p = 2, we have n (B2) · · · h(2m) n M ∩ H 2 =  2r−1 Xj=1 ⊕ hϕji  ⊕ B2J1H 2 (B) ⊕ · · · ⊕ B2JrH 2 (B) , LAX-HALMOS TYPE THEOREMS ON H p SPACES 9 and hence M =   =   2r−1 Xj=1 Xj=1 2r−1 ⊕ hϕji ⊕ hϕji  ⊕ B2J1H 2 (B) ⊕ · · · ⊕ B2JrH 2 (B)  ⊕ B2(cid:16)J1H 2 (B) ⊕ · · · ⊕ JrH 2 (B)(cid:17) (bar denotes closure in H p). We can easily see that B2(J1H p(B)⊕...⊕JrH p(B)) ⊂ taking the closure in H p of all three subspaces we shall get equality throughout B2(cid:16)J1H 2 (B) ⊕ · · · ⊕ JrH 2 (B)(cid:17) ⊂ B2(cid:16)J1H p (B) ⊕ · · · ⊕ JrH p (B)(cid:17) and so upon so that B2(cid:16)J1H 2 (B) ⊕ · · · ⊕ JrH 2 (B)(cid:17) = B2(cid:16)J1H p (B) ⊕ · · · ⊕ JrH p (B)(cid:17) and this gives us the characterisation for the case 0 < p < 1. The case 1 ≤ p < 2. The arguments and conclusions above in the case 0 < p < 1 are also valid for this case and so certainly ⊕ hϕji M =  2r−1 Xj=1  ⊕(cid:16)J1H p (B) ⊕ · · · ⊕ JrH p (B)(cid:17) where the bar denotes closure in H p, 1 ≤ p < 2. Let N = J1H p (B) ⊕ · · · ⊕ JrH p (B) so that M = ( ⊕ hϕji) ⊕ N . Then as a closed subspace of H p, N is invariant 2r−1 Pj=1 under multiplication by B. It can be verified that N ∩ H 2 = J1H 2 (B) ⊕ · · · ⊕ JrH 2 (B) . Any arbitrary g ∈ JiH p (B) can be written as g = Jif , for some f ∈ H p (B). Then the Cesaro means of f denoted by the sequence of polynomials, {pn}, is such that pn(z) → f (z) in H p. Hence pn(B) → f (B) in H p. But Ji ∈ H ∞, so pn(B)Ji → Jif (B) in H p. Because N is B− invariant, we have pn(B)Ji ∈ N , and the fact that N is closed implies that Jif ∈ N . This establishes that J1H p (B) ⊕ · · · ⊕ JrH p (B) ⊂ N . Now we establish the inclusion in the other direction. In a fashion, similar to as shown above, for any f ∈ N , we can construct an outer function K ∈ H ∞ such that Kf ∈ N ∩ H 2. Therefore, (3.3) Kf = J1h1 + J2h2 + · · · + Jrhr, for some uniquely determined h1, h2, . . . , hr ∈ H 2(B) ⊂ H p (B). Since f ∈ H p, by Lemma 2, we can express it uniquely as (3.4) f = e00f1 + e10f2 + · · · + en−1,0fn, for some f1, . . . , fn ∈ H p (B). Therefore, (3.5) Kf = e00Kf1 + e10Kf2 + · · · + en−1,0Kfn. Because J1, J2, . . . , Jr are B− inner, we can write J1 = e00ϕ10 + e10ϕ11 + · · · + en−1,0ϕ1,n−1 J2 = e00ϕ20 + e10ϕ21 + · · · + en−1,0ϕ2,n−1 ... Jr = e00ϕr0 + e10ϕr1 + · · · + en−1,0ϕr,n−1, 10 NITEESH SAHNI AND DINESH SINGH where the B− matrix (ϕij )r×n satisfies (ϕij )r×n (ϕji)n×r = I. Equation 3.3 now becomes Kf = e00 (ϕ10h1 + ϕ20h2 + · · · + ϕr0hr) + (3.6) e10 (ϕ11h1 + ϕ21h2 + · · · + ϕr1hr) + · · · + en−1,0 (ϕ1,n−1h1 + ϕ2,n−1h2 + · · · + ϕr,n−1hr) From equations (3.5) and (3.6) we see that Kf1 = ϕ10h1 + ϕ20h2 + · · · + ϕr0hr ... Kfn = ϕ1,n−1h1 + ϕ2,n−1h2 + · · · + ϕr,n−1hr. This in matrix form can be written as (3.7) (Kfi)1×n = (hi)1×r (ϕij)r×n Taking the conjugate transpose we get (3.8) By multiplying equations (3.7) and (3.8) we get: Thus for j = 1, ..., r, we have 2 (cid:12)(cid:12)(cid:12)(cid:12) h1 K(cid:12)(cid:12)(cid:12)(cid:12) = f12 + · · · + fn2 ≤ (f1 + · · · + fn)2 . 2 hr (cid:0)Kfi(cid:1)n×1 = (ϕji)n×r(cid:0)hi(cid:1)r×1 K(cid:12)(cid:12)(cid:12)(cid:12) + · · · +(cid:12)(cid:12)(cid:12)(cid:12) K(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ≤ f1 + · · · + fn hj hj K ∈ H p. and this clearly implies that Then from (3.3) we conclude that f is in J1H p (B) ⊕ · · · ⊕ JrH p (B) so that N ⊂ J1H p (B) ⊕ · · · ⊕ JrH p (B) and so N = J1H p (B) ⊕ · · · ⊕ JrH p (B) and this then ∈ Lp. Because K is outer, we have hj K implies that M = k Pj=1 ⊕ hϕji! ⊕Pr l=1 ⊕B2JlH p (B). The case 1 < p ≤ ∞. Let M1 = M denote the closure of M in H 2. Suppose M1 is invariant under multiplication by B(z), then by Theorem 2, we can write M1 = ϕ1H 2 (B) ⊕ · · · ⊕ ϕnH 2 (B) , for some B− inner functions ϕ1, . . . , ϕn. It follows that any element f ∈ M can be written as f = ϕ1h1 + · · · + ϕnhn, for some h1, . . . , hn ∈ H 2 (B). By Lemma 2, hj ∈ H p (in fact hj ∈ H p (B)). We ⊂ l o∞ claim that Φk = ϕkhk ∈ M . Since Φk ∈ M , so there exists a sequence nh(k) l −→ Φk in H 2 as l → ∞. Moreover we can write M such that h(k) l=1 and h(k) l = e00h(k,1) l + · · · + en−1,0h(k,n) l Φk = e00Φ(1) k + · · · + en−1,0Φ(n) k . LAX-HALMOS TYPE THEOREMS ON H p SPACES 11 l (B) −→ ej0Φ(j) Therefore, ej0h(k,,j) isometry on H 2 (B), so we have h(k,j) h(k,j) l H ∞ such that g(k,j) (z) −→ Φ(j) (z) h(k,j) k (B) in H 2. But multiplication by ej0 is an k (B) in H 2. This implies that ⊂ (B) −→ Φ(j) l k (z) in H 2. By Lemma 2, there exists a sequence ng(k,j) k (z) in H p as l → ∞. Define (z) −→ Φ(j) l l o∞ l=1 l g(k) l = g(k,1) l (cid:0)B2(cid:1) · · · g(k,n) l (cid:0)B2(cid:1) is uniformly bounded and converges to 1 a.e. Consider l o∞ so that ng(k) l=1 g(k) l h(k) l ej−1,0g(k) l h(k,j) l n n = = Xj=1 Xj=1 (cid:0)B2(cid:1) · · · g(k,n) (cid:0)B2(cid:1) · · · g(k,j−1) l l l ej−1,0g(k,1) l (cid:0)B2(cid:1) · · · g(k,n) l (cid:0)B2(cid:1) h(k,j) l We now show that g(k,1) sequence θl = g(k,1) bounded and converges to 1a.e., and the sequence ψl = g(k,j) It can be shown that θlψl −→ Φ(j) l h(k) by the invariance of M we have Φk ∈ M . This means that M can be written as (cid:0)B2(cid:1) h(k,j) (cid:0)B2(cid:1) g(k,j+1) in H p. l −→ Φk in H p. Now (cid:0)B2(cid:1) is uniformly (cid:0)B2(cid:1) · · · g(k,n) in H p, and hence g(k) in H p. Note that the −→ Φ(j) −→ Φ(j) h(k,j) k k k l l l l l l where M = ϕ1N1 ⊕ ϕ2N2 ⊕ · · · ⊕ ϕnNn, Nj = {h ∈ H p (B) : ϕjh ∈ M } is a closed subspace of H p. It is easy to see that Nj is invariant under B2 and B3, and is also dense in H 2 (B). Note that all Njs cannot be B− invariant simultaneously. For if they are then it would imply that M is also B−invariant which is not possible. Thus, some Nj is not invariant under B. Without loss of generality assume that N1is not invariant. We show that even this is not possible. For any f ∈ H p (B), we can find a sequence {fl} in N1 such that fl → f in H 2 because N1 is dense in H 2 (B). Once again we can write fl = e00f (1) l + · · · + en−1,0f (n) l and f = e00f (1) + · · · + en−1,0f (n) l o∞ l → f (j) in H 2. Again by Lemma 3, there exists a sequence ng(j) so that f (j) H ∞, such that g(j) (B) · · · g(n) that Bmglfl → Bmf in H p for m ≥ 2. Thus Bmf ∈ N1 for m ≥ 2. Let us choose f = 1. So we have B2H p (B) ⊂ N1. This inclusion must be strict as N1 is not invariant under B. So N1 is of the form l → f (j) in H p. Taking gl = g(1) (B), it follows l f (j) l=1 in l l N1 = A ⊕ B2H p (B) , where A is a non zero subspace. We know that A ( N1 H 2 (B) . 12 NITEESH SAHNI AND DINESH SINGH If 1 and B belong to A, then N1 = H p (B) which is not possible. So A = hα + βBi, where α 6= 0. Again the density of N1 implies that there exists a sequence (cid:8)αn (α + βB) + B2fn(cid:9) ⊂ N1 that converges to 1 in H 2. This gives B2fn → 0, ααn → 1, and βαn → 0 Therefore, β = 0, which means that there cannot be a sequence in N1 that converges to B in H 2 norm. This contradicts the fact that N1 is dense in H 2 (B). This contradiction stems from the fact that M1 is assumed to be invariant under B. Thus M1 is invariant under B2 and B3 but not under B. Now by the validity of our result on H 2, there exist B− inner functions J1, . . . , Jr, r ≤ n, such that M1 =  2r−1 Xj=1 ⊕ hϕj i  ⊕ B2J1H 2 (B) ⊕ · · · ⊕ B2JrH 2 (B) 1J1 + αj 2J1B + αj 3J2 + αj where ϕj = αj 2nJnB. From the form of ϕj it is clear that ϕj ∈ H p. Using the arguments already used in the proof it can be shown that ϕj ∈ M . Also essentially repeating the arguments as in the previous case we can easily establish that B2J1, . . . , B2Jr ∈ M . Thus 4J2B + · · · + αj 2n−1Jn + αj ⊕ hϕji 2r−1 Xj=1    ⊕ B2J1H p (B) ⊕ · · · ⊕ B2JrH p (B) ⊂ M. To establish the reverse inclusion consider any f ∈ M . By virtue of the character- ization of M1, f = α1ϕ1 + · · · + α2n−1ϕ2n−1 + B2J1h1 + · · · + B2Jnhn. Note that B2J1h1 + · · · + B2Jrhr = f − α1ϕ1 − · · · − α2r−1ϕ2r−1 ∈ H p. Hence, by Lemma 4, h1, . . . , hn ∈ H p (B). Hence f ∈ 2r−1 Pj=1 so that M ⊂ 2r−1 Pj=1 ⊕ hϕj i! ⊕ B2J1H p (B) ⊕ · · · ⊕ B2JrH p (B) ⊕ hϕj i! ⊕ B2J1H p (B) ⊕ · · · ⊕ B2JrH p (B) . This completes the proof of the theorem. (cid:3) 4. The B− invariant subspaces of H p The ideas from the above proof can be applied to derive a new factorization free proof of the following invariant subspace theorem obtained in [13] for the cases 1 ≤ p ≤ ∞, p 6= 2. In addition we have extended the theorem to the case 0 < p < 1. Theorem 4. Let M be a closed subspace of H p, 0 < p ≤ ∞, p 6= 2, such that M is invariant under H ∞ (B). Then there exist B− inner functions J1, . . . , Jr, r ≤ n, such that M = J1H p (B) ⊕ · · · ⊕ JrH p (B) . When 0 < p < 1, then, as the proof will show, the right hand side is to be read as being dense in M i.e. its closure in the H p metric is all of M . LAX-HALMOS TYPE THEOREMS ON H p SPACES 13 Proof. The idea of the proof is quite similar to the proof of the Theorem 3. We shall only sketch the details. Using the fact that every 0 6= f ∈ H p can be written as a product of an appropriate number of H 2 functions, we can construct an outer function O(z), in a manner identical to the proof of Theorem 3, such that O(B(z))f ∈ M ∩ H 2. Thereby establishing that M ∩ H 2 6= {0}. Now M ∩ H 2 is a closed subspace of H 2 and invariant under H ∞(B), so by Theorem 2 there exist B− inner functions J1, . . . , Jr, with r ≤ n, such that M ∩ H 2 = J1H 2 (B) ⊕ · · · ⊕ JrH 2 (B) . Next we show that M ∩ H 2 is dense in M . It is trivial to note that M ∩ H 2 ⊆ M , where the bar denotes closure in H p. In order to establish the reverse inequality, we follow the same arguments used in the proof of Theorem 3 to construct a sequence l=0, such that for any f ∈ M , Ol(B(z))f ∈ M ∩ H 2, and of outer functions {Ol(z)}∞ Ol(B(z))f → f in H p, as l → ∞. The proof of Theorem 3 also establishes that, for 0 < p < 1, M ∩ H 2 takes the form J1H p (B) ⊕ · · · ⊕ JrH p (B) and for 1 ≤ p < 2, M ∩ H 2 = J1H p (B) ⊕ · · · ⊕ JrH p (B) thereby establishing the characterization for M in these cases. Next we deal with the case when 2 < p ≤ ∞. As in the proof of Theorem 3, we consider M1 = M , the closure of M in H 2. Since M is B− invariant, we have M1 is B− invariant. So by Theorem 2 there exist B− inner functions J1, . . . , Jr, r ≤ n, such that M1 = J1H 2(B) ⊕ · · · ⊕ JrH 2(B). Thus any arbitrary f ∈ M can be written as f = J1h1 + · · · + Jrhr, for some h1, . . . , hr ∈ H 2. By Lemma 4, these h1, . . . , hr ∈ H p, and hence M ⊂ J1H 2(B) ⊕ · · · ⊕ JrH 2(B). To prove the inclusion in the reverse, we need to establish that, for each k = 1, . . . , r, JkH p(B) ⊂ M . For an arbitrary h ∈ H p(B), consider Ψk = Jkh. Now proceeding in the same fashion as in the proof of Theorem 3 (where we show that Φk ∈ M ), it follows that Ψk ∈ M . (cid:3) Acknowledgement 1. The second author thanks Vern Paulsen for useful dis- cussions. The first author thanks the Shiv Nadar University, Dadri, Uttar Pradesh, and the Mathematical Sciences Foundation, New Delhi, for the facilities given to complete this work. References [1] J. Agler and J.E. McCarthy, Pick Interpolation and Hilbert Function Spaces, Graduate Stud- ies in Mathematics, 44, American Mathematical Society, Providence, RI, 2002. [2] C. Cowen and B. Mccluer, Composition Operators on Spaces of Analytic Functions, CRC Press, 1994. [3] K. R. Davidson, and R. Hamilton, Nevanlinna-Pick interpolation and factorization of linear functionals, Integral Equations and Operator theory, 70(2011),125-149. [4] K.R. Davidson, V.I. Paulsen, M. Raghupathi, and D. Singh, A constrained Nevanlinna-Pick interpolation problem, Indiana University Mathematics Journal, 58(2009), 709 -- 732. [5] P.L. Duren, Theory of H p Spaces, Academic Press, London-New York, 1970. [6] T.W. Gamelin, Uniform Algebras, AMS Chelsea 1984. [7] J.B. Garnett, Bounded Analytic Functions, Academic Press, 1981. [8] P.R. Halmos, Shifts on Hilbert spaces, J. Reine Angew. Math. 208(1961), 102-112. [9] H. Helson, Harmonic analysis, Hindustan Book Agency, 1995. [10] K. Hoffman, Banach Spaces of Analytic Functions, Prentice Hall, 1962. [11] M. Jury, G. Knese, and S. McCullough, Nevanlinna-Pick interpolation on distinguished vari- eties in the bidisk, J. Funct. Anal., 262(2012), 3812-3838. [12] P. Koosis, Introduction to H p spaces, Cambridge University Press, Cambridge, 1998. 14 NITEESH SAHNI AND DINESH SINGH [13] T.L. Lance and M.I. Stessin, Multiplication Invariant Subspaces of Hardy Spaces, Can. J. Math, 49(1997) 100-118. [14] P.D. Lax, Translation invariant spaces, Acta Math. 101(1959), 163-178. [15] S. McCullough, and T.T. Trent, Invariant subspaces and Nevanlinna-Pick Kernels, J. Funct. Anal., 178(2000), 226-249. [16] M. Raghupathi, Abrahamse's interpolation theorem and Fuchsian groups, J. Math. Anal. Appl,, J. Math. Anal. Appl., 355 (2009), 258 -- 276. [17] M. Raghupathi, Nevanlinna-Pick interpolation for C + BH∞, Integral Equations and Oper- ator theory, 63 (2009), 103-125. [18] V.I. Paulsen and D. Singh, Modules over subalgebras of the disk algebra, Indiana Univ. Math. Jour. 55 (2006), 1751-1766. [19] D. Singh and V. Thukral, Multiplication by finite Blaschke factors on de Branges spaces, J. Operator Theory 37(1997), 223-245. [20] N. K. Nikolski, Operators, Functions and Systems: an easy reading, Vol. 1, Amer. Math. Soc., 2002. [21] M. Raghupathi, and D. Singh, Function theory in real Hardy spaces, Math. Nachr., 284 (2011), 920-930. University of Delhi, Delhi, 110007, Shiv Nadar University, Dadri, Uttar Pradesh E-mail address: [email protected] University of Delhi, Delhi, 110007 E-mail address: [email protected]
1709.09396
2
1709
2019-04-05T08:54:00
Backward Shift Invariant Subspaces in Reproducing Kernel Hilbert Spaces
[ "math.FA" ]
In this note, we describe the backward shift invariant subspaces for a large class of reproducing kernel Hilbert spaces. This class includes in particular de Branges-Rovnyak spaces (the non-extreme case) and the range space of co-analytic Toeplitz operators.
math.FA
math
BACKWARD SHIFT INVARIANT SUBSPACES IN REPRODUCING KERNEL HILBERT SPACES EMMANUEL FRICAIN, JAVAD MASHREGHI and RISHIKA RUPAM Abstract In this note, we describe the backward shift invariant subspaces for an abstract class of repro- ducing kernel Hilbert spaces. Our main result is inspired by a result of Sarason concerning de Branges -- Rovnyak spaces (the non-extreme case). Furthermore, we give new applications in the context of the range space of co-analytic Toeplitz operators and sub-Bergman spaces. 1. Introduction A celebrated theorem of Beurling describes all (non-trivial) closed invariant sub- spaces of the Hardy space H 2 on the open unit disc D which are invariant with respect to the backward shift operator S∗. They are of the form KΘ = (ΘH 2)⊥, where Θ is an inner function. The result of Beurling was the cornerstone of a whole new direction of research lying at the interaction between operator the- ory and complex analysis. It was generalized in many ways. See for instance [1, 3, 5, 7, 17]. Sarason [22] classified the non-trivial closed backward shift invari- ant subspaces of the de Branges -- Rovnyak spaces H (b), where b is a non-extreme point of the closed unit ball of H ∞: they are of the form KΘ ∩ H (b), where Θ is an inner function. In other words, the closed invariant subspaces for S∗H (b) are the trace on H (b) of the closed invariant subspaces for S∗. This naturally leads to the following question: (Q): let H1 and H2 be two reproducing kernel Hilbert spaces on D such that H1 ⊂ H2; assume that the shift operator S (multiplication by the independent variable) is contractive on H2 and if T = SH2, its adjoint T ∗ maps H1 con- tractively into itself. Then, is it true that every closed invariant subspace E of T ∗H1 has the form E ∩ H1, where E is a closed invariant subspace of T ∗ (as an 2010 Mathematics Subject Classification. 30J05, 30H10, 46E22. Key words and phrases. backward shift operators, toeplitz operators, de Branges -- Rovnyak spaces. The first and third authors were supported by Labex CEMPI (ANR-11-LABX-0007-01) and the grant ANR-17-CE40 -0021 of the French National Research Agency ANR (project Front). The second author was supported by grants from NSERC (Canada). 1 9 1 0 2 r p A 5 ] . A F h t a m [ 2 v 6 9 3 9 0 . 9 0 7 1 : v i X r a 2 FRICAIN, MASHREGHI and RUPAM operator on H2)? In other words, are the closed invariant subspaces for T ∗H1 the trace on H1 of the closed invariant subspaces for T ∗? It should be noted that, of course, the interesting situation is when H1 is not a closed subspace of H2. Sarason's result says that the answer to question (Q) is affirmative in the situation where H2 = H 2 and H1 = H (b), with b a non- extreme point of the closed unit ball of H ∞. However, it should be noted that question (Q) has a negative answer in the case where H1 = D is the Dirichlet space and H2 = H 2 is the Hardy space. Indeed, let (zn)n≥1 be a non-Blaschke sequence of D which is a zero set for A2 and put M := {f ∈ A2 : f (zn) = 0, n ≥ 1}, where A2 is the Bergman space of D. Define N = {F ∈ Hol(D) : F ′ = f, f ∈ A2 ⊖ M }. It is not difficult to see that N is a non-trivial closed subspace of D, which is S∗ invariant. Then, observe that N cannot be of the form E ∩D, where E is a closed subspace of H 2 invariant with respect to S∗. Indeed, assume on the contrary that there exists a closed subspace E of H 2, invariant with respect to S∗, such that N = E ∩ D. Since N is non-trivial, the subspace E is also non-trivial, and by Beurling's theorem, there exists an inner function Θ such that E = KΘ. Thus N = KΘ ∩ D. Observe now that for every n ≥ 1, the Cauchy kernel kλn belongs to N (because its derivative is up to a constant the reproducing kernel of A2 at point λn and thus it is orthogonal to M ). Then kλn ∈ KΘ, n ≥ 1. To get a contradiction, it remains to see that, since (zn)n≥1 is not a Blaschke sequence, then the sequence of Cauchy kernels kλn , n ≥ 1, generates all H 2, and we deduce that H 2 ⊂ KΘ, which is absurd. The aim of this note is to present a general framework where the answer to the question (Q) is affirmative. Note that in [2], Aleman -- Malman present another general situation of reproducing kernel Hilbert spaces where they extend Sarason's result. In Section 2, we first recall some basic facts on reproducing kernel Hilbert spaces and on the Sz.-Nagy -- Foias model for contractions. Then, in Section 3, we study the properties of multiplication operators in our general context and prove that the scalar spectral measures of the minimal unitary dilation of T ∗H1 are absolutely continuous. We also show that when H2 = H 2, then the reproducing kernel Hilbert space H1 satisfies an interesting division property, the so-called F -property. In Section 4, we give an analogue of Beurling's theorem in our general context and give an application to cyclic vectors for the backward shift. In Section 5, we show that our main theorem can be applied to H (b) spaces and range space of co-analytic Toeplitz operators. We also provide a BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 3 new application in the context of sub-Bergman Hilbert space which was recently studied in [26, 27, 28]. 2. Preliminaries We first recall some standard facts on reproducing kernel Hilbert spaces. See [21] for a detailed treatment of RKHS. 2.1. Reproducing kernel Hilbert spaces and multipliers Let H be a Hilbert space of complex valued functions on a set Ω. We say that H is a reproducing kernel Hilbert space (RKHS) on Ω if the following two conditions are satisfied: (P1) for every λ ∈ Ω, the point evaluations f 7−→ f (λ) are bounded on H ; (P2) for every λ ∈ Ω, there exists a function f ∈ H such that f (λ) 6= 0. According to the Riesz representation theorem, for each λ ∈ Ω, there is a function kH λ in H , called the reproducing kernel at point λ, such that f (λ) = hf, kH λ iH , (f ∈ H ). Note that according to (P2), we must have kH λ sequence in H , then 6≡ 0. Moreover if (fn)n is a fn → f weakly in H =⇒ ∀λ ∈ Ω, lim n→∞ fn(λ) = f (λ). (2.1) A multiplier of H is a complex valued function ϕ on Ω such that ϕf ∈ H for all f ∈ H . The set of all multipliers of H is denoted by Mult(H ). Using the closed graph theorem, we see that if ϕ belongs to Mult(H ), then the map Mϕ,H : (cid:12)(cid:12)(cid:12)(cid:12) H −→ H f 7−→ ϕf (2.2) is bounded on H . When there is no ambiguity, we simply write Mϕ for Mϕ,H . It is well-known that if we set kϕkMult(H ) = kMϕkL(H ), ϕ ∈ Mult(H ), then Mult(H ) becomes a Banach algebra. Moreover, using a standard argu- ment, we have M ∗ ϕkH λ = ϕ(λ)kH λ , (λ ∈ Ω), which gives ϕ(λ) ≤ kϕkMult(H ), (λ ∈ Ω). (2.3) (2.4) See for instance [10, Chapter 9] or [21]. Let H1, H2 be two RKHS such that H1 ⊂ H2. If (fn)n is a sequence in H1 which is convergent in the weak topology of H2, we cannot deduce that it also 4 FRICAIN, MASHREGHI and RUPAM converges in the weak topology of H1. However, the following result shows that on the bounded subsets of H1 the above conclusion holds. Lemma 2.1. Let H1, H2 be two RKHS on a set Ω such that H1 ⊂ H2, let (fn)n be a sequence in H1 bounded in H1-norm by a constant C, and let f ∈ H2. Assume that (fn)n converges to f in the weak topology of H2. Then the following holds: (i) f ∈ H1, (ii) fn → f in the weak topology of H1, (iii) kf kH1 ≤ C. Proof. Since (fn)n is uniformly bounded in the norm of H1, it has a weakly convergent subsequence. More explicitly, there is a subsequence (fnk )k that converges to some g ∈ H1 in the weak topology of H1. Using (2.1), we easily see that the two functions f and g coincide on Ω. Therefore f ∈ H1. Second, since each H1−weakly convergent subsequence of (fn)n has to converge weakly to f in H1, we conclude that (fn)n itself also converges to f in the weak topology of H1. Third, the weak convergence in H1 implies completing the proof. kf kH1 ≤ lim inf n→∞ kfnkH1 ≤ C, 2.2. H ∞ functional calculus for contractions Let T be a contraction on a Hilbert space H . We recall that T is said to be completely non-unitary if there is no nonzero reducing subspaces H0 for T such that T H0 is a unitary operator. We recall that for a completely non-unitary contraction T on H , we can define an H ∞-functional calculus with the following properties (see [4, Theorem 2.1, page 117]): (P3) for every f ∈ H ∞, we have kf (T )k ≤ kf k∞. (P4) If (fn)n is a sequence of H ∞ functions which tends boundedly to f on the open unit disc D (which means that supn kfnk∞ < ∞ and fn(z) → f (z), n → +∞ for every z ∈ D), then fn(T ) tends to f (T ) WOT (for the weak operator topology). (P5) If (fn)n is a sequence of H ∞ functions which tends boundedly to f almost everywhere on T = ∂D, then fn(T ) tends to f (T ) SOT (for the strong operator topology). Finally, we recall that every contraction T on a Hilbert space H has a unitary dilation U on K (which means that H ⊂ K and T n = PH U nH , n ≥ 1) which is minimal (in the sense that K = W∞ −∞ U nH ). BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 5 2.3. A general framework In this note, we consider two analytic reproducing kernel Hilbert spaces H1 and H2 on the open unit disc D (which means that their elements are analytic on D) and such that H1 ⊂ H2. A standard application of the closed graph theorem shows that there is a constant C such that kf kH2 ≤ Ckf kH1, (f ∈ H1). (2.5) Denote by χ the function χ(z) = z, z ∈ D. Furthermore, we shall assume the following two properties: χ ∈ Mult(H2) and kχkMult(H2) ≤ 1, and if X := M ∗ χ,H2 (recall notation (2.2)), then XH1 ⊂ H1 and kXkL(H1) ≤ 1. The restriction of X to H1 is denoted by XH1 := XH1. (2.6) (2.7) 2.4. Range Spaces Let X , Y be two Hilbert spaces and T ∈ L(X , Y ). We define M(T ) as the range space equipped with the range norm. More explicitly, M(T ) = R(T ) = T X and kT xkM(T ) = kP(kerT )⊥xkX , x ∈ X , where P(kerT )⊥ denotes the orthogonal projection from X onto (kerT )⊥. It is easy to see that M(T ) is a Hilbert space which is boundedly contained in Y . A result of Douglas [6] says that if A ∈ L(X1, Y ) and B ∈ L(X2, Y ), then M(A) ≖ M(B) ⇐⇒ AA∗ = BB∗. (2.8) Here the notation M(A) ≖ M(B) means that the Hilbert spaces M(A) and M(B) coincide as sets and, moreover, have the same Hilbert space structure. We also recall that if A, B ∈ L(X1, Y ) and C ∈ L(Y ), then C is a contraction from M(A) into M (B) ⇐⇒ CAA∗C∗ ≤ BB∗. (2.9) See also [11, Corollaries 16.8 and 16.10]. 3. Multiplication operators Note that (2.6) implies kχkMult(H2) = 1. Indeed, according to (2.4), we have 1 = sup z∈D χ(z) ≤ kχkMult(H2) ≤ 1. More generally, since Tn≥0 M n non-unitary contraction. Hence, we get the following consequence. χ,H2H2 = {0}, we see that Mχ,H2 is a completely 6 FRICAIN, MASHREGHI and RUPAM Lemma 3.1. Let H2 be a reproducing kernel Hilbert space of analytic functions on D satisfying (2.6). Then Mult(H2) = H ∞ and for every ϕ ∈ H ∞, we have Mϕ,H2 = ϕ(Mχ,H2 ) with kϕkMult(H2) = kϕk∞. (3.1) Proof. Let ϕ ∈ H ∞ and consider the dilates ϕr(z) = ϕ(rz), 0 < r < 1, z ∈ D. If ϕ(z) = P∞ n=0 anzn and f ∈ H2, observe that ϕr(Mχ,H2 )f = ∞ Xn=0 anrnM n χ,H2f = ∞ Xn=0 anrnχnf = ϕrf. Moreover, by (P5), we have ϕr(Mχ,H2 )f → ϕ(Mχ,H2 )f in H2 as r → 1. Then, using (2.1), we get on one hand ϕr(λ)f (λ) = (ϕr(Mχ,H2 )f )(λ) → (ϕ(Mχ,H2 )f )(λ), as r → 1, (λ ∈ D), and on the other hand, ϕr(λ)f (λ) → ϕ(λ)f (λ), r → 1 (λ ∈ D). We thus deduce In particular, ϕ ∈ Mult(H2) and Mϕ,H2 = that ϕf = ϕ(Mχ,H2 )f ∈ H2. ϕ(Mχ,H2 ). Moreover, by (P3), we have kϕkMult(H2) = kϕ(Mχ,H2 )k ≤ kϕk∞. If we combine with (2.4), we get (3.1), as claimed. Lemma 3.2. Let H1 and H2 be two reproducing kernel Hilbert spaces of ana- lytic functions on D such that H1 ⊂ H2. Assume that H1 and H2 satisfy (2.6) and (2.7). Then the minimal unitary dilation of XH1 has an absolutely con- tinuous scalar spectral measure. In particular, for every f, g ∈ H1, there exists uf,g ∈ L1(T) such that hX n H1f, giH1 = ZT znuf,g(z) dm(z). (3.2) Proof. Let f, g ∈ H1 and let µf,g be the scalar spectral measure associated to the minimal unitary dilation of the contraction XH1 . Then, we have hX n H1 f, giH1 = ZT zn dµf,g(z). (3.3) Let us prove that µf,g is absolutely continuous with respect to normalized Lebesgue measure m on T. Let F be a closed Borel subset of T such that m(F ) = 0. Then, we can construct a bounded sequence of polynomials (qn)n such that qn(z) → χF (z), as n → +∞, for every z ∈ D. Indeed, Let f be the Fatou function associated to F , that is a function f in the disc algebra (that is the closure of polynomials for the sup norm) such that f = 1 on F and f < 1 on D \ F (See [20, page 116] or [18]). Now take f n, n ≥ 0. The functions f n are BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 7 still in the disc algebra. Then if we take ε > 0, we can find a polynomial qn such that f n(z) − qn(z) ≤ sup z∈D ε 2 . In particular, we have for every z ∈ F , 1 − qn(z) ≤ ε/2. On the other hand, for z ∈ D \ F , we can find n0 such that for n ≥ n0, f n(z) ≤ ε/2 (because fn(z) < 1 and thus f n(z) → 0, as n → ∞). Therefore, for n ≥ n0, we have qn(z) ≤ qn(z) − f n(z) + f n(z) ≤ ε 2 + ε 2 = ε. Hence qn(z) tends to 1 for z ∈ F and to 0 for z ∈ D \ F . In other words, qn tends to χF pointwise. On the other hand, we have of course qn(z) ≤ 1 + sup z∈D ε 2 , which proves that the sequence (qn)n is also bounded, and we are done. Now, since (qn)n converges boundedly to 0 on D and since X is a completely unitary contraction, we deduce from (P4) that (qn(X))n converges WOT to 0 in L(H2). Hence it implies that (qn(XH1 )f )n converges weakly to 0 in H2. On the other hand, by von Neumann inequality, we have kqn(XH1)f kH1 ≤ kqnk∞kf kH1 ≤ Ckf kH1, where C = supn kqnk∞ < +∞. By Lemma 2.1, we deduce that (qn(XH1 )f )n converges weakly to 0 in H1. But, according to (3.3), we have which gives that hqn(XH1 )f, giH1 = ZT qn(z) dµf,g(z), n→+∞ZT lim qn(z) dµf,g(z) = 0. It remains to apply dominated Lebesgue convergence theorem to get ZT χF (z) dµf,g(z) = 0, which implies that µf,g(F ) = 0. Hence µf,g is absolutely continuous with respect to m, as claimed. Theorem 3.3. Let H1 and H2 be two reproducing kernel Hilbert spaces of analytic functions on D such that H1 ⊂ H2. Assume that H1 and H2 satisfy (2.6) and (2.7). Let ϕ ∈ H ∞. Then M ∗ ϕ,H2 maps H1 into itself, and if f, g ∈ H1, we have hM ∗ ϕ,H2f, giH1 = ZT ϕ∗(z)uf,g(z) dm(z), (3.4) 8 FRICAIN, MASHREGHI and RUPAM where ϕ∗(z) = ϕ(z). Proof. Let us first assume that ϕ is holomorphic on D and let us consider the Taylor series of ϕ, ϕ(z) = P∞ n=0 anzn. Then we have M ∗ ϕ,H2 = ϕ(Mχ,H2 )∗ = ∞ Xn=0 anX n. (3.5) H1 ⊂ H1. Now using Since XH1 ⊂ H1, the last equation implies that M ∗ ϕ,H2 that P∞ n=0 an < ∞ and (3.2), we get hM ∗ ϕ,H2f, giH1 = ∞ ∞ = Xn=0 Xn=0 = ZT = ZT anhX n H1 f, giH1 znuf,g(z) dm(z) anZT ∞ anznuf,g(z) dm(z) Xn=0 ϕ∗(z)uf,g(z) dm(z). This proves (3.4) for ϕ which is holomorphic on D. We also observe that (cid:12)(cid:12)hM ∗ ϕ,H2f, giH1(cid:12)(cid:12) ≤ ZT ϕ∗(z)uf,g(z) dm(z) ≤ kϕk∞ZT uf,g(z) dm(z). But by spectral theorem, we know thatRT uf,g(z) dm(z) = kµf,gk ≤ kf kH1kgkH1, which gives kM ∗ (3.6) Now let ϕ ∈ H ∞ and define the dilates ϕr(z) = ϕ(rz), 0 < r < 1, z ∈ D. Observe that ϕr are holomorphic on D. By the previous argument, we get that M ∗ ϕ,H2f kH1 ≤ kϕk∞kf kH1. ϕr,H2 maps H1 into itself and hM ∗ ϕr,H2 f, giH1 = ZT ϕ∗ ruf,g dm, f, g ∈ H1. (3.7) Since ϕr converges boundedly to ϕ on D as r → 1, and since Mχ,H2 is a com- pletely non unitary contraction on H2, we get that M ∗ ϕr,H2 f converges weakly to M ∗ ϕ,H2f in H2 as r → 1. On the other hand, using (3.6), we have kM ∗ ϕr,H2f kH1 ≤ kϕrk∞kf kH1 ≤ kϕk∞kf kH1. BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 9 Lemma 2.1 now implies that M ∗ ϕr,H2 f converges weakly to M ∗ ϕ,H2f in H1 as r → 1. Letting r → 1 in (3.7) and using dominated convergence, we deduce that formula (3.4) is satisfied by ϕ, completing the proof. ϕ,H2f belongs to H1 and M ∗ Remark 3.4. It follows immediately from (3.4) that for ϕ ∈ H ∞, we have kM ∗ ϕ,H2kL(H1) ≤ kϕk∞. Given a bounded operator T on a Hilbert space H , the family of all closed T -invariant subspaces of H is denoted by Lat(T ). Corollary 3.5. Let H1 and H2 be two reproducing kernel Hilbert spaces of analytic functions on D such that H1 ⊂ H2. Assume that H1 and H2 satisfy (2.6) and (2.7). Then, for every ϕ ∈ H ∞, we have Lat(XH1) ⊂ Lat(M ∗ ϕ,H2H1). Proof. Let ϕ ∈ H ∞, ϕr(z) = ϕ(rz), 0 < r < 1, and let E ∈ Lat(XH1 ). Note that (3.5) implies that M ∗ ϕr,H2E ⊂ E. On the other hand, as we have seen in the proof of Theorem 3.3, M ∗ ϕr,H2 → M ∗ ϕ,H2, as r → 1, in the weak operator topology of L(H1). Since a norm-closed subspace is also weakly closed [19], we conclude that M ∗ ϕ,H2E ⊂ E, as claimed. To conclude this section, we show that Theorem 3.3 has an interesting appli- cation in relation with the F-property. Recall that a linear manifold V of H 1 is said to have the F-property if whenever f ∈ V and θ is an inner function which is lurking in f , i.e., f /θ ∈ H 1 or equivalently θ divides the inner part of f , then we actually have f /θ ∈ V . This concept was first introduced by V. P. Havin [16] and it plays a vital role in the analytic function space theory. Several classical spaces have the F-properties. the list includes Hardy spaces H p, Dirichlet space D, BMOA, VMOA, and the disc algebra A. See [13, 15, 24]. However, for the Bloch spaces B and B0, we know that B ∩ H p and B0 ∩ H p do not have the F-property [14]. Using the tools developed in Section 2.3, we will see that in the situation when H1 ⊂ H 2 satisfies (2.7), then H1 has the F-property. First, let us note that H2 = H 2 satisfies (2.6) and Mχ,H2 = S is the classical forward shift operator. Thus, X = M ∗ χ,H2 = S∗ is the backward shift operator f (z) − f (0) (S∗f )(z) = , f ∈ H 2, z ∈ D. z In this context, if H1 is a reproducing kernel Hilbert space such that H1 ⊂ H 2, the condition (2.7) can be rephrased as (3.8) Recall that for ψ ∈ L∞(T), the Toeplitz operator Tψ is defined on H 2 by Tψ(f ) = P+(ψf ) where P+ is the Riesz projection (the orthogonal projection from L2(T) and kS∗H1k ≤ 1. S∗H1 ⊂ H1 10 FRICAIN, MASHREGHI and RUPAM onto H 2). If ϕ ∈ H ∞ = Mult(H 2), then Mϕ,H 2 = Tϕ and M ∗ situation, we get the following result. ϕ,H 2 = Tϕ. In this Theorem 3.6. Let H1 be a reproducing kernel Hilbert space contained in H 2, and assume that it satisfies (3.8). Then the space H1 has the F-property. Moreover, if f ∈ H1 and θ is an inner function which divides f , then ≤ kf kH1. (cid:13)(cid:13)(cid:13)(cid:13) f θ(cid:13)(cid:13)(cid:13)(cid:13)H1 Proof. Assume that f ∈ H1 and that θ is an inner function so that f /θ ∈ In fact, by Smirnov Theorem [20], we actually have ψ := f /θ ∈ H 2. H 1. Therefore, Tθ(f ) = P+(θf ) = P+(ψ) = ψ. But according to Theorem 3.3 and Remark 3.4, Tθ acts contractively on H1. Hence ψ = Tθ(f ) ∈ H1 and f as claimed. (cid:13)(cid:13)(cid:13)(cid:13) θ(cid:13)(cid:13)(cid:13)(cid:13)H1 = (cid:13)(cid:13)Tθf(cid:13)(cid:13)H1 ≤ kf kH1, 4. Invariant subspaces and cyclicity The following result says that under certain circumstances, the closed invariant subspaces of XH1 = XH1 are exactly the trace on H1 of the closed invariant subspaces of X. Despite the following characterization, the implication (i) =⇒ (ii) is the essential part of the result. Theorem 4.1. Let H1 and H2 be two analytic reproducing kernel Hilbert spaces on D such that H1 ⊂ H2 and satisfying (2.6) and (2.7). Assume that there exists an outer function ϕ ∈ H ∞ such that R(M ∗ ϕ,H2) ⊂ H1. Then, for every E ⊂ H1, the following assertions are equivalent. (i) E is a closed subspace of H1 invariant under XH1; (ii) there is a closed subspace E of H2 invariant under X = M ∗ χ,H2 such that E = E ∩ H1. Moreover, E = H1 if and only if E = H2. The proof will be based on the following lemma, which extends [11, Lemmata 17.21 and 24.30] in our general context. Lemma 4.2. Let H1 and H2 be two analytic reproducing kernel Hilbert spaces on D such that H1 ⊂ H2 and satisfying (2.6) and (2.7). Assume that there exists an outer function ϕ ∈ H ∞ such that R(M ∗ ϕ,H2) ⊂ H1. Then, for every BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 11 E ∈ Lat(XH1 ), the space M ∗ of H1. ϕ,H2E is dense in E with respect to the norm topology Proof. According to Corollary 3.5, we know that M ∗ ϕ,H2E ⊂ E. Now let ϕ,H2E in the H1-topology. In particular, for every n ≥ 0, we have g ∈ E, g ⊥ M ∗ 0 = hM ∗ ϕ,H2X n H1 g, giH1. Observe now that M n Hence, by Theorem 3.3, we get χ,H2 Mϕ,H2 = Mχnϕ,H2, which gives M ∗ ϕ,H2X n = M ∗ χnϕ,H2. 0 = hM ∗ χnϕ,H2g, giH1 = ZT ϕ∗(z)znug,g(z) dm(z), 0 . Since ϕ∗ is outer and 0 . Since ug,g ≥ 0, this for every n ≥ 0. We thus deduce that ϕ∗ug,g ∈ H 1 ug,g ∈ L1(T), Smirnov Theorem [8] implies that ug,g ∈ H 1 gives ug,g = 0, that is g = 0, completing the proof of the Lemma. Proof of Theorem 4.1. (ii) =⇒ (i): Let E be a closed subspace of H2, invariant under X = M ∗ χ,H2 such that E = E ∩ H1. First, let us check that E is a closed subspace of H1. The verification essentially owes to (2.5). To do so, let f ∈ H1 be in the H1-closure of E ∩ H1. Then there is a sequence (fn)n in E ∩ H1 which converges to f in the norm topology of H1. Since H1 is boundedly contained in H2, the sequence (fn)n also converges to f in H2. Since E is closed in H2, the function f must belong to E. Hence, f ∈ E = E ∩ H1, which proves that E is closed in H1. The fact that E is invariant under XH1 = M ∗ χ,H2 H1 is immediate. ϕ,H2 ) ⊂ H1, the mapping M ∗ (i) =⇒ (ii): A standard argument using the closed graph theorem implies that, according to R(M ∗ ϕ,H2 from H2 into H1 is a bounded operator. Now let E be a closed subspace of H1, and assume that E is invariant under XH1 . Denote by E the closure of E in the H2-topology. It is clear that E is a closed subspace of H2 which is invariant under X. Let us prove that E = E ∩ H1. The inclusion E ⊂ E ∩ H1 is trivial. For the reverse inclusion, let us verify that M ∗ ϕ,H2E ⊂ E. (4.1) Let f ∈ E. By definition, there is a sequence (fn)n in E which converges to f in the H2-topology. Then, since M ∗ ϕ,H2 is bounded from H2 into H1, the sequence (M ∗ ϕ,H2f in the H1-topology. Since fn ∈ E, Corollary 3.5 implies that M ∗ ϕ,H2f ∈ E, which proves (4.1). In particular, we have ϕ,H2fn ∈ E and since E is closed in H1, then M ∗ ϕ,H2fn)n tends to M ∗ M ∗ ϕ,H2(E ∩ H1) ⊂ E, 12 FRICAIN, MASHREGHI and RUPAM and since E ∩ H1 is a closed subspace of H1 invariant with respect to XH1, it follows from Lemma 4.2 that M ∗ ϕ,H2(E ∩ H1) is dense in E ∩ H1, which implies Thus we have E = E ∩ H1. E ∩ H1 ⊂ E. It remains to prove that E = H1 if and only if E = H2. Assume first that ϕ,H2) ⊂ E. But note that ker (Mϕ,H2) = {0}, whence ϕ,H2) is dense in H2. Hence we get E = H2. Conversely, assume that E = H1. Then R(M ∗ R(M ∗ E = H2. Then E = E ∩ H1 = H2 ∩ H1 = H1, which concludes the proof. Remark 4.3. As already noted, R(M ∗ ϕ,H2 ) is always dense in H2 and thus under the hypothesis of Theorem 4.1 (that is if there exists an (outer) function ϕ ∈ H ∞ such that R(M ∗ ϕ,H2) ⊂ H1), then automatically H1 is dense in H2. Theorem 4.1 has an immediate application in characterization of cyclic vec- tors. Corollary 4.4. Let H1 and H2 be two analytic reproducing kernel Hilbert spaces on D such that H1 ⊂ H2 and satisfying (2.6) and (2.7). Suppose that there exists an outer function ϕ ∈ H ∞ such that R(M ∗ ϕ,H2) ⊂ H1. Let f ∈ H1. Then the following assertions are equivalent. (i) f is cyclic for X = M ∗ χ,H2. (ii) f is cyclic for XH1 = XH1. Proof. (i) =⇒ (ii): Assume that f is cyclic for X in H2 and denote by E the subspace of H1 defined by E = Span(X n H1 f : n ≥ 0) H1 . It is clear that E is a closed subspace of H1, invariant with respect to XH1. Assume that E 6= H1. Then, according to Theorem 4.1, there exists a closed subspace E of H2, E 6= H2, invariant under X such that E = E ∩ H1. In particular, f ∈ E, and thus it is not cyclic for X, which is contrary to the hypothesis. Thus E = H1 and f is cyclic for XH1. (ii) =⇒ (i): Let g ∈ R(M ∗ ϕ,H2). Then g ∈ H1 and if f is cyclic for XH1, there exists a sequence of polynomials (pn) such that kpn(XH1)f − gkH1 → 0, as n → ∞. Since H1 is contained boundedly in H2, then we have kpn(XH2)f − gkH2 → 0, as n → ∞. BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 13 Thus, R(M ∗ get ϕ,H2) ⊂ Span(X ∗nf : n ≥ 0) . Since R(M ∗ ϕ,H2) is dense in H2, we H2 Span(X ∗nf : n ≥ 0) H2 = H2, completing the proof. We can apply Theorem 4.1 and Corollary 4.4 to some specific reproducing kernel Hilbert spaces contained in the Hardy space H 2 on D. Theorem 4.5. Let H1 be a reproducing kernel Hilbert space contained in H 2 that satisfies (3.8) and assume that there exists an outer function ϕ ∈ H ∞ such that TϕH 2 ⊂ H1. Then, for every E ( H1, the following assertions are equivalent. (i) E is a closed subspace of H1 invariant under XH1; (ii) there is an inner function Θ such that E = KΘ ∩ H1. Moreover, if f ∈ H1, then f is cyclic for S∗H1 if and only if f is cyclic for S∗. Proof. It is sufficient to combine Theorem 4.1 and Corollary 4.4 with Beurl- ing's theorem. Remark 4.6. Note that the hypothesis TϕH 2 ⊂ H1 implies that polynomials belong to H1. Regarding the last part of Theorem 4.5, let us mention that a well-known theorem of Douglas -- Shapiro -- Shields [7] says that a function f in H 2 is cyclic for S∗ if and only if f has no bounded type meromorphic pseudo continuation across T to De = {z : 1 < z ≤ ∞}. 5. Applications In this section, we give some examples of RKHS for which our main Theorem 4.1 can be applied. 5.1. A general RKHS Let H2 be an analytic RKHS on D satisfying (2.6). Let ϕ ∈ H ∞ and H1 := M(M ∗ ϕ,H2). Recall the definition of the range space from Section 2.4. Then H1 is also an analytic RKHS on D which is contained in H2. Observe that H1 satisfies (2.7). Indeed, since Mϕ,H2Mχ,H2 = Mχ,H2 Mϕ,H2, we have XM ∗ ϕ,H2 = M ∗ ϕ,H2g ∈ H1 for some g ∈ (kerM ∗ ϕ,H2X, which implies that XH1 ⊂ H1. Moreover, if f = M ∗ ϕ,H2)⊥, then kXf kH1 = kXM ∗ ϕ,H2gkH1 ϕ,H2XgkH1 = kM ∗ = kXgkH2 ≤ kgkH2 = kf kH1 14 FRICAIN, MASHREGHI and RUPAM Thus H1 satisfies (2.7). In this context, we get immediately from Theorem 4.1 the following. Corollary 5.1. Let H2 be an analytic RKHS satisfying (2.6). Let ϕ be an ϕ,H2). Then, for every E ⊂ H1, the outer function in H ∞ and let H1 := M(M ∗ following assertions are equivalent. (i) E is a closed subspace of H1, invariant under XH1 ; (ii) There is a closed subspace E of H2, invariant under X = M ∗ χ,H2 such that E = E ∩ H1. Moreover E = H1 if and only if E = H2. 5.2. The space M(ϕ) Let H2 = H 2 be the Hardy space on D, ϕ an outer function in H ∞ and H1 = M(Tϕ) which we denote for simplicity M(ϕ). The space H 2 trivially satisfies (2.6) and according to the discussion at the beginning of Subsection 5.1, the space M(ϕ) is an analytic RKHS contained in H 2 and satisfying (2.7) (or equivalently (3.8)). Again, for simplicity, we write Xϕ = XM(ϕ) = S∗M(ϕ). In this context, we can apply Theorem 4.5 which immediately gives the fol- lowing. Corollary 5.2. Let ϕ be an outer function. Then the following assertions are equivalent. (i) E is a closed subspace of M(ϕ), E 6= M(ϕ), and E is invariant under Xϕ. (ii) There is an inner function Θ such that E = KΘ ∩ M(ϕ). 5.3. de Branges -- Rovnyak space H (b) Let b ∈ b(H ∞) -- the closed unit ball of H ∞. The de Branges -- Rovnyak space H (b) is defined as H (b) = M((I − TbTb)1/2). For details on de Branges -- Rovnyak spaces, we refer to [11, 23]. Here we will just recall what will be useful for us. It is well -- known that H (b) is an analytic RKHS contractively contained in H 2 and invariant with respect to S∗. Moreover, the operator Xb = S∗H (b) acts as a contraction on H (b). In particular, the space H (b) satisfies the hypothesis (3.8). Assume now that b is a non-extreme point of b(H ∞), meaning that log(1 − b) ∈ L1(T). Thus, there exists a unique outer function a such that a(0) > 0 and a2 + b2 = 1 a.e. on T. This function a is called the pythagorean mate of b. It is well-known that R(Ta) ⊂ H (b). We can then apply Theorem 4.5 to H1 = H (b) and H2 = H 2 to recover the following result due to Sarason ([22], Theorem 5). BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 15 Corollary 5.3 (Sarason). Let b be a non-extreme point of the closed unit ball of H ∞, and let E be a closed subspace of H (b), E 6= H (b). Then the following are equivalent. (i) E is invariant under Xb. (ii) There exists an inner function Θ such that E = KΘ ∩ H (b). Remark 5.4. As already noted, hypothesis of Theorem 4.5 implies that poly- nomials belongs to H1. In the case when H1 = H (b), we know that it necessarily implies that b is non-extreme. In the extreme case, the backward shift invariant subspaces have been described by Suarez [25], also using some Sz.-Nagy-Foias model theory, but the situation is rather more complicated. 5.4. Sub-Bergman Hilbert space The Bergman space A2 on D is defined as the space of analytic functions f on D satisfying kf k2 A2 := ZD f (z)2dA(z) < ∞, where dA(z) is the normalized area measure on D. In [26, 27, 28], an analogue of de Branges -- Rovnyak spaces was considered in this context. Recall that the Toeplitz operator on A2(D) with symbol ϕ ∈ L∞(D) is defined as Tϕ(f ) = PA2 (ϕf ), where PA2 is the Bergman projection (that is the orthogonal projection from ϕ = Tϕ. Given ϕ ∈ L∞(D), we define the L2(D, dA) onto A2). It is clear that T ∗ sub -- Bergman Hilbert space H (ϕ) as H (ϕ) = M((I − TϕTϕ)1/2). In other words, H (ϕ) = (I − TϕT ∗ product ϕ)1/2A2 and it is equipped with the inner h(I − TϕT ∗ ϕ)1/2f, (I − TϕT ∗ ϕ)1/2giH(ϕ) := hf, giA2, for every f, g ∈ A2⊖ker(I−TϕT ∗ ϕ). We keep the same notation as the de Branges -- Rovnyak spaces, but there will be no ambiguity because in this subsection, the ambient space is A2 (in contrast with the de Branges -- Rovnyak spaces for which the ambient space is H 2). We refer the reader to [27] for details about this space. The shift operator (also denoted S in this context), defined as S = Tz, is clearly a contraction and S∗ = Tz. As we have seen, the de Branges -- Rovnyak spaces are invariant with respect to the backward shift operator which acts as a contraction on them. In the context of sub -- Bergman Hilbert spaces, the ana- logue of this property is also true. The proof is the same but we include it for completeness. 16 FRICAIN, MASHREGHI and RUPAM Lemma 5.5. Let b ∈ b(H ∞). Then S∗ acts as a contraction on H (b). Proof. We first prove that S∗ acts as a contraction on H (b). According to (2.9), we should prove that that is S∗(I − TbTb)S ≤ I − TbTb, Tz(I − TbTb)Tz ≤ I − TbTb. But, if ϕ, ψ ∈ L∞(D, dA) and at least one of them is in H ∞, then TψTϕ = Tψϕ. (5.1) (5.2) See [29, Proposition 7.1]. Then (5.1) is equivalent to Tz2(1−b2) ≤ T1−b2, that is 0 ≤ T(1−z2)(1−b2). Since (1 − z2)(1 − b2) ≥ 0 on D, the latter inequality is satisfied (see also [29, Proposition 7.1]) and thus S∗ acts as a contraction on H (b). To pass to the H (b) case, we use a well -- known relation between H (b) and H (b): let f ∈ A2; then f ∈ H (b) if and only if Tbf ∈ H (b) and kf k2 H (b) = kf k2 A2 + kTbf k2 H (b). So let f ∈ H (b). Since TbS∗f = S∗Tbf and H (b) is invariant with respect to S∗, we get that TbS∗f ∈ H (b), whence S∗f ∈ H (b) and kS∗f k2 H (b) = kS∗f k2 = kS∗f k2 A2 + kTbS∗f k2 A2 + kS∗Tbf k2 H (b) H (b) H (b) = kf k2 Hence S∗ is a contraction on H (b), completing the proof. A2 + kTbf k2 ≤ kf k2 H (b). According to Lemma 5.5, we see that A2 satisfies (2.6) and H (b) satisfies (2.7). We will show that under the additional hypothesis that b is a non-extreme point of the closed unit ball of H ∞, we can apply our Corollary 5.1 to H1 = H (b) and H2 = A2. Corollary 5.6. Let b be a non -- extreme point of the unit ball of H ∞ and a its pythagorean mate. Then the following are equivalent. (i) E is a closed subspace of H (b), invariant under Xb = S∗H (b). (ii) There is a closed subspace E of A2, invariant under S∗, such that E = E ∩ H (b). BACKWARD SHIFT INVARIANT SUBSPACES IN RKHS 17 Proof. It is known that since b is analytic, then H (b) = H (¯b) with equiv- alent norms, see [26, page 641]. Moreover, according to (5.2), we have I − T¯bTb = T1−b2 = Ta2 = T¯aTa. This identity implies by (2.8) that H (¯b) ≖ M(T¯a). Hence H (b) = M(T¯a) = M(T ∗ a ) with equivalent norms. We then apply Corollary 5.1 to H2 = A2 and H1 = H (b) = M(T ∗ a ), which gives the result. Acknowledgements. We would like to warmly thank the anonymous ref- In an eree for his/her remarks leading to a real improvement of the paper. earlier version, we had stated Theorem 4.1 with the additional hypothesis that for every E ∈ Lat(XH1), the space M ∗ ϕ,H2E is dense in E. It was the referee's suggestion to use the Sz.-Nagy -- Foias theory to obtain that particular property as a consequence of the other hypothesis in the theorem (see Lemma 4.2). References 1. Aleksandrov A.B., Invariant subspaces of the backward shift operator in the space H p (p ∈ (0, 1)). Zap. Nauchn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 92:7 -- 29, 318, 1979. Investigations on linear operators and the theory of functions, IX. 2. Aleman A.; Malman B., Hilbert spaces of analytic functions with a contractive backward shift. Preprint on arXiv:1805.11842. 3. Aleman A.; Richter S; Sundberg C., Invariant subspaces for the backward shift on Hilbert spaces of analytic functions with regular norm. In Bergman spaces and related topics in complex analysis, volume 404 of Contemp. Math., pages 1 -- 25. Amer. Math. Soc., Provi- dence, RI, 2006. 4. Bercovici H.; Foias C.; Kerchy L.; Sz.-Nagy B., Harmonic analysis of operators on Hilbert space Universitext, Springer Verlag. Revised and Enlarged Edition, 2010. 5. Bolotnikov, V.; Rodman, L., Finite dimensional backward shift invariant subspaces of Arveson spaces. Linear Algebra Appl., 349:265 -- 282, 2002. 6. Douglas, R.G., On majorization, factorization, and range inclusion of operators on Hilbert space. Proc. Amer. Math. Soc., 17:413 -- 415, 1966. 7. Douglas, R.G.; Shapiro, H.S.; Shields, A.L., Cyclic vectors and invariant subspaces for the backward shift operator. Ann. Inst. Fourier (Grenoble), 20(fasc. 1):37 -- 76, 1970. 8. Duren P., Theory of H p spaces, volume 38 of Pure and Applied Mathematics,. Academic Press, New York-London, 1970. 9. El-Fallah, O.; Kellay, K.; Mashreghi, J.; Ransford, T., A primer on the Dirichlet space, volume 203 of Cambridge Tracts in Mathematics. Cambridge University Press, Cambridge, 2014. 10. Fricain, E.; Mashreghi, J., The theory of H(b) spaces. Vol. 1, volume 20 of New Mathe- matical Monographs. Cambridge University Press, Cambridge, 2016. 11. Fricain, E.; Mashreghi, J., The theory of H(b) spaces. Vol. 2, volume 21 of New Mathe- matical Monographs. Cambridge University Press, Cambridge, 2016. 12. Garnett J., Bounded Analytic Functions. Graduate Texts in Mathematics 236. Revised First Edition, Springer, 2007. 18 FRICAIN, MASHREGHI and RUPAM 13. Girela, D.; Gonz´alez,C., Division by inner functions. In Progress in analysis, Vol. I, II (Berlin, 2001), pages 215 -- 220. World Sci. Publ., River Edge, NJ, 2003. 14. Girela, D.; Gonz´alez,C.; Pel´aez, J. ´A., Multiplication and division by inner functions in the space of Bloch functions. Proc. Amer. Math. Soc., 134(5):1309 -- 1314, 2006. 15. Girela, D.; Gonz´alez,C.; Pel´aez, J. ´A., Toeplitz operators and division by inner functions. In Proceedings of the First Advanced Course in Operator Theory and Complex Analysis, pages 85 -- 103. Univ. Sevilla Secr. Publ., Seville, 2006. 16. Havin, V.P., The factorization of analytic functions that are smooth up to the boundary. Zap. Naucn. Sem. Leningrad. Otdel. Mat. Inst. Steklov. (LOMI), 22:202 -- 205, 1971. 17. Izuchi, K.; Nakazi, T., Backward shift invariant subspaces in the bidisc. Hokkaido Math. J., 33(1):247 -- 254, 2004. 18. Koosis, P., An introduction to H p-spaces, vol. 115 of Cambridge Tracts in Mathematics Cambridge University Press, Cambridge 1998. 19. Lax, P.D., Functional analysis. Pure and Applied Mathematics (New York). Wiley- Interscience [John Wiley & Sons], New York, 2002. 20. Mashreghi, J., Representation theorems in Hardy spaces, volume 74 of London Mathemat- ical Society Student Texts. Cambridge University Press, Cambridge, 2009. 21. Paulsen, V.I.; Raghupathi, M., An introduction to the theory of reproducing kernel Hilbert spaces, volume 152 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 2016. 22. Sarason, D., Doubly shift-invariant spaces in H 2. J. Operator Theory, 16(1):75 -- 97, 1986. 23. Sarason, D., Sub-Hardy Hilbert spaces in the unit disk. Wiley-interscience, 1994. 24. Shirokov N.A., Analytic functions smooth up to the boundary. Lecture Notes in Mathe- matics, 1312, Springer Verlag, Berlin, 1988. 25. Su´arez, D., Backward shift invariant spaces in H 2. Indiana Univ. Math. J., 46(2):593 -- 619, 1997. 26. Sultanic, S., Sub-Bergman Hilbert spaces. Journal of mathematical analysis and applica- tions, 324(1):639 -- 649, 2006. 27. Zhu, K., Sub-Bergman Hilbert spaces on the unit disk. Indiana University Mathematics Journal, 45(1):165 -- 176, 1996. 28. Zhu, K., Sub-Bergman Hilbert spaces in the unit disk. II. J. Funct. Anal., 202(2):327 -- 341, 2003. 29. Zhu, K., Operator theory in function spaces. Number 138. American Mathematical Soc., 2007. Laboratoire Paul Painlev´e, Univer- sit´e Lille 1, 59 655 Villeneuve d'Ascq C´edex E-mail : [email protected] lille1.fr Laboratoire Paul Painlev´e, Univer- sit´e Lille 1, 59 655 Villeneuve d'Ascq C´edex E-mail : [email protected] lille1.fr D´epartement de math´ematiques et de statistique, Universit´e Laval, Qu´ebec, QC, Canada G1K 7P4 E-mail : [email protected]
1807.05417
1
1807
2018-07-14T16:19:49
Differential structure associated to axiomatic Sobolev spaces
[ "math.FA" ]
The aim of this note is to explain in which sense an axiomatic Sobolev space over a general metric measure space (\`a la Gol'dshtein-Troyanov) induces - under suitable locality assumptions - a first-order differential structure.
math.FA
math
Differential structure associated to axiomatic Sobolev spaces Nicola Gigli∗ Enrico Pasqualetto† July 17, 2018 Abstract The aim of this note is to explain in which sense an axiomatic Sobolev space over a general metric measure space (`a la Gol'dshtein-Troyanov) induces – under suitable locality assumptions – a first-order differential structure. MSC2010: primary 46E35, secondary 51Fxx Keywords: axiomatic Sobolev space, locality of differentials, cotangent module Contents Introduction 1 General notation 2 Axiomatic theory of Sobolev spaces 3 Cotangent module associated to a D-structure Introduction 1 2 3 10 An axiomatic approach to the theory of Sobolev spaces over abstract metric measure spaces has been proposed by V. Gol'dshtein and M. Troyanov in [6]. Their construction covers many important notions: the weighted Sobolev space on a Riemannian manifold, the Haj lasz Sobolev space [7] and the Sobolev space based on the concept of upper gradient [2, 3, 8, 9]. A key concept in [6] is the so-called D-structure: given a metric measure space (X, d, m) and an exponent p ∈ (1, ∞), we associate to any function u ∈ Lp loc(X) a family D[u] of non- negative Borel functions called pseudo-gradients, which exert some control from above on the variation of u. The pseudo-gradients are not explicitly specified, but they are rather supposed ∗SISSA, Via Bonomea 265, 34136 Trieste. E-mail address: [email protected] †SISSA, Via Bonomea 265, 34136 Trieste. E-mail address: [email protected] 1 to fulfil a list of axioms. Then the space W 1,p(X, d, m, D) is defined as the set of all functions in Lp(m) admitting a pseudo-gradient in Lp(m). By means of standard functional analytic techniques, it is possible to associate to any Sobolev function u ∈ W 1,p(X, d, m, D) a uniquely determined minimal object Du ∈ D[u] ∩ Lp(m), called minimal pseudo-gradient of u. More recently, the first author of the present paper introduced a differential structure on general metric measure spaces (cf. [4, 5]). The purpose was to develop a second-order differential calculus on spaces satisfying lower Ricci curvature bounds (or briefly, RCD spaces; we refer to [1,12,13] for a presentation of this class of spaces). The fundamental tools for this theory are the Lp-normed L∞-modules, among which a special role is played by the cotangent module, denoted by L2(T ∗X). Its elements can be thought of as 'measurable 1-forms on X'. The main result of this paper – namely Theorem 3.2 – says that any D-structure (satisfying suitable locality properties) gives rise to a natural notion of cotangent module Lp(T ∗X; D), whose properties are analogous to the ones of the cotangent module L2(T ∗X) described in [4]. Roughly speaking, the cotangent module allows us to represent minimal pseudo-gradients as pointwise norms of suitable linear objects. More precisely, this theory provides the existence of an abstract differential d : W 1,p(X, d, m, D) → Lp(T ∗X; D), which is a linear operator such that the pointwise norm du ∈ Lp(m) of du coincides with Du in the m-a.e. sense for any function u ∈ W 1,p(X, d, m, D). 1 General notation For the purpose of the present paper, a metric measure space is a triple (X, d, m), where (X, d) m 6= 0 is a complete and separable metric space, is a non-negative Borel measure on X, finite on balls. (1.1) Fix p ∈ [1, ∞). Several functional spaces over X will be used in the forthcoming discussion: L0(m) : Lp(m) : Lp loc(m) : L∞(m) : L0(m)+ : Lp(m)+ : Lp loc(m)+ : LIP(X) : Sf(X) : the Borel functions u : X → R, considered up to m-a.e. equality. the functions u ∈ L0(m) for which up is integrable. the functions u ∈ L0(m) with uB ∈ Lp(cid:0)mB(cid:1) for any B ⊆ X bounded Borel. the functions u ∈ L0(m) that are essentially bounded. the Borel functions u : X → [0, +∞], considered up to m-a.e. equality. the functions u ∈ L0(m)+ for which up is integrable. the functions u ∈ L0(m)+ with uB ∈ Lp(cid:0)mB(cid:1)+ for any B ⊆ X bounded Borel. the Lipschitz functions u : X → R, with Lipschitz constant denoted by Lip(u). the functions u ∈ L0(m) that are simple, i.e. with a finite essential image. Observe that for any u ∈ Lp that the space Sf(X) is strongly dense in Lp(m) for every p ∈ [1, ∞]. loc(m)+ it holds that u(x) < +∞ for m-a.e. x ∈ X. We also recall 2 Remark 1.1 In [6, Section 1.1] a more general notion of Lp loc(m) is considered, based upon the concept of K-set. We chose the present approach for simplicity, but the following discussion would remain unaltered if we replaced our definition of Lp (cid:4) loc(m) with the one of [6]. 2 Axiomatic theory of Sobolev spaces We begin by briefly recalling the axiomatic notion of Sobolev space that has been introduced by V. Gol'dshtein and M. Troyanov in [6, Section 1.2]: Definition 2.1 (D-structure) Let (X, d, m) be a metric measure space. Let p ∈ [1, ∞) be fixed. Then a D-structure on (X, d, m) is any map D associating to each function u ∈ Lp loc(m) a family D[u] ⊆ L0(m)+ of pseudo-gradients of u, which satisfies the following axioms: A1 (Non triviality) It holds that Lip(u) χ{u>0} ∈ D[u] for every u ∈ Lp A2 (Upper linearity) Let u1, u2 ∈ Lp loc(m) be fixed. Consider g1 ∈ D[u1] and g2 ∈ D[u2]. Suppose that the inequality g ≥ α1 g1 + α2 g2 holds m-a.e. in X for some g ∈ L0(m)+ and α1, α2 ∈ R. Then g ∈ D[α1 u1 + α2 u2]. loc(m)+ ∩ LIP(X). A3 (Leibniz rule) Fix a function u ∈ Lp loc(m) and a pseudo-gradient g ∈ D[u] of u. Then for every ϕ ∈ LIP(X) bounded it holds that g supX ϕ + Lip(ϕ) u ∈ D[ϕ u]. A4 (Lattice property) Fix u1, u2 ∈ Lp loc(m). Given any g1 ∈ D[u1] and g2 ∈ D[u2], one has that max{g1, g2} ∈ D(cid:2) max{u1, u2}(cid:3) ∩ D(cid:2) min{u1, u2}(cid:3). A5 (Completeness) Consider two sequences (un)n ⊆ Lp satisfy gn ∈ D[un] for every n ∈ N. Suppose that there exist u ∈ Lp such that un → u in Lp loc(m) and gn → g in Lp(m). Then g ∈ D[u]. loc(m) and (gn)n ⊆ Lp(m) that loc(m) and g ∈ Lp(m) Remark 2.2 It follows from axioms A1 and A2 that 0 ∈ D[c] for every constant map c ∈ R. Moreover, axiom A2 grants that the set D[u] ∩ Lp(m) is convex and that D[α u] = α D[u] for every u ∈ Lp loc(m) and α ∈ R \ {0}, while axiom A5 implies that each set D[u] ∩ Lp(m) is closed in the space Lp(m). (cid:4) Given any Borel set B ⊆ X, we define the p-Dirichlet energy of a map u ∈ Lp(m) on B as For the sake of brevity, we shall use the notation Ep(u) to indicate Ep(uX). Ep(uB) := inf(cid:26)ZB g ∈ D[u](cid:27) ∈ [0, +∞]. (2.1) gp dm(cid:12)(cid:12)(cid:12)(cid:12) Definition 2.3 (Sobolev space) Let (X, d, m) be a metric measure space. Let p ∈ [1, ∞) be fixed. Given a D-structure on (X, d, m), we define the Sobolev class associated to D as Moreover, the Sobolev space associated to D is defined as Sp(X) = Sp(X, d, m, D) :=(cid:8)u ∈ Lp loc(m) : Ep(u) < +∞(cid:9). W 1,p(X) = W 1,p(X, d, m, D) := Lp(m) ∩ Sp(X, d, m, D). 3 (2.2) (2.3) Theorem 2.4 The space W 1,p(X, d, m, D) is a Banach space if endowed with the norm kukW 1,p(X) :=(cid:16)kukp Lp(m) + Ep(u)(cid:17)1/p for every u ∈ W 1,p(X). (2.4) For a proof of the previous result, we refer to [6, Theorem 1.5]. Proposition 2.5 (Minimal pseudo-gradient) Let (X, d, m) be a metric measure space and let p ∈ (1, ∞). Consider any D-structure on (X, d, m). Let u ∈ Sp(X) be given. Then there exists a unique element Du ∈ D[u], which is called the minimal pseudo-gradient of u, such that Ep(u) = kDukp Lp(m). Both existence and uniqueness of the minimal pseudo-gradient follow from the fact that the set D[u]∩Lp(m) is convex and closed by Remark 2.2 and that the space Lp(m) is uniformly convex; see [6, Proposition 1.22] for the details. In order to associate a differential structure to an axiomatic Sobolev space, we need to be sure that the pseudo-gradients of a function depend only on the local behaviour of the function itself, in a suitable sense. For this reason, we propose various notions of locality: Definition 2.6 (Locality) Let (X, d, m) be a metric measure space. Fix p ∈ (1, ∞). Then we define five notions of locality for D-structures on (X, d, m): L1 If B ⊆ X is Borel and u ∈ Sp(X) is m-a.e. constant in B, then Ep(uB) = 0. L2 If B ⊆ X is Borel and u ∈ Sp(X) is m-a.e. constant in B, then Du = 0 m-a.e. in B. L3 If u ∈ Sp(X) and g ∈ D[u], then χ{u>0} g ∈ D[u+]. L4 If u ∈ Sp(X) and g1, g2 ∈ D[u], then min{g1, g2} ∈ D[u]. L5 If u ∈ Sp(X) then Du ≤ g holds m-a.e. in X for every g ∈ D[u]. Remark 2.7 In the language of [6, Definition 1.11], the properties L1 and L3 correspond to locality and strict locality, respectively. (cid:4) We now discuss the relations among the several notions of locality: Proposition 2.8 Let (X, d, m) be a metric measure space. Let p ∈ (1, ∞). Fix a D-structure on (X, d, m). Then the following implications hold: L3 =⇒ L2 =⇒ L1, L4 ⇐⇒ L5 L1 + L5 =⇒ L2 + L3. (2.5) Proof. L2 =⇒ L1. Simply notice that Ep(uB) ≤RB(Du)p dm = 0. 4 L3 =⇒ L2. Take a constant c ∈ R such that the equality u = c holds m-a.e. in B. Given that Du ∈ D[u − c] ∩ D[c − u] by axiom A2 and Remark 2.2, we deduce from L3 that Given that u − c = (u − c)+ − (c − u)+, by applying again axiom A2 we see that χ{u>c} Du ∈ D(cid:2)(u − c)+(cid:3), χ{u<c} Du ∈ D(cid:2)(c − u)+(cid:3). χ{u6=c} Du = χ{u>c} Du + χ{u<c} Du ∈ D[u − c] = D[u]. Hence the minimality of Du grants that ZX (Du)p dm ≤Z{u6=c} (Du)p dm, which implies that Du = 0 holds m-a.e. in {u = c}, thus also m-a.e. in B. This means that the D-structure satisfies the property L2, as required. L4 =⇒ L5. We argue by contradiction: suppose the existence of u ∈ Sp(X) and g ∈ D[u] such L5 =⇒ L4. Since Du ≤ g1 and Du ≤ g2 hold m-a.e., we see that Du ≤ min{g1, g2} holds m-a.e. as well. Therefore min{g1, g2} ∈ D[u] by A2. that m(cid:0){Du > g}(cid:1) > 0, whence h := min{Du, g} ∈ Lp(m) satisfies R hp dm < R (Du)p dm. Since h ∈ D[u] by L4, we deduce that Ep(u) <R (Du)p dm, getting a contradiction. L1+L5 =⇒ L2+L3. Property L1 grants the existence of (gn)n ⊆ D[u] withRB(gn)p dm → 0. Hence L5 tells us that RB(Du)p dm ≤ limnRB(gn)p dm = 0, which implies that Du = 0 holds because we know that h = max{h, 0} ∈ D(cid:2) max{u, 0}(cid:3) = D[u+] for every h ∈ D[u] by A4 and 0 ∈ D[0], in particular u+ ∈ Sp(X). Given that u+ = 0 m-a.e. in the set {u ≤ 0}, one has that Du+ = 0 holds m-a.e. in {u ≤ 0} by L2. Hence for any g ∈ D[u] we have Du+ ≤ χ{u>0} g by L5, which implies that χ{u>0} g ∈ D[u+] by A2. Therefore L3 is proved. (cid:3) m-a.e. in B, yielding L2. We now prove the validity of L3: it holds that D[u] ⊆ D[u+], Definition 2.9 (Pointwise local) Let (X, d, m) be a metric measure space and p ∈ (1, ∞). Then a D-structure on (X, d, m) is said to be pointwise local provided it satisfies L1 and L5 (thus also L2, L3 and L4 by Proposition 2.8). We now recall other two notions of locality for D-structures that appeared in the literature: Definition 2.10 (Strong locality) Let (X, d, m) be a metric measure space and p ∈ (1, ∞). Consider a D-structure on (X, d, m). Then we give the following definitions: i) We say that D is strongly local in the sense of Timoshin provided χ{u1<u2} g1 + χ{u2<u1} g2 + χ{u1=u2} (g1 ∧ g2) ∈ D[u1 ∧ u2] (2.6) whenever u1, u2 ∈ Sp(X), g1 ∈ D[u1] and g2 ∈ D[u2]. 5 ii) We say that D is strongly local in the sense of Shanmugalingam provided χB g1 + χX\B g2 ∈ D[u2] for every g1 ∈ D[u1] and g2 ∈ D[u2] (2.7) whenever u1, u2 ∈ Sp(X) satisfy u1 = u2 m-a.e. on some Borel set B ⊆ X. The above two notions of strong locality have been proposed in [11] and [10], respectively. We now prove that they are actually both equivalent to our pointwise locality property: Lemma 2.11 Let (X, d, m) be a metric measure space and p ∈ (1, ∞). Fix any D-structure on (X, d, m). Then the following are equivalent: i) D is pointwise local. ii) D is strongly local in the sense of Shanmugalingam. iii) D is strongly local in the sense of Timoshin. Proof. i) =⇒ ii) Fix u1, u2 ∈ Sp(X) such that u1 = u2 m-a.e. on some E ⊆ X Borel. Pick g1 ∈ D[u1] and g2 ∈ D[u2]. Observe that D(u2 − u1) + g1 ∈ D(cid:2)(u2 − u1) + u1(cid:3) = D[u2] by A2, so that we have (cid:0)D(u2 − u1) + g1(cid:1) ∧ g2 ∈ D[u2] by L4. Since D(u2 − u1) = 0 m-a.e. on B by L2, we see that χB g1 + χX\B g2 ≥(cid:0)D(u2 − u1) + g1(cid:1) ∧ g2 holds m-a.e. in X, whence accordingly we conclude that χB g1 + χX\B g2 ∈ D[u2] by A2. This shows the validity of ii). ii) =⇒ i) First of all, let us prove L1. Let u ∈ Sp(X) and c ∈ R satisfy u = c m-a.e. on some Borel set B ⊆ X. Given any g ∈ D[u], we deduce from ii) that χX\B g ∈ D[u], thus accordingly Ep(uB) ≤RB(χX\B g)p dm = 0. This proves the property L1. To show property L4, fix u ∈ Sp(X) and g1, g2 ∈ D[u]. Let us denote B := {g1 ≤ g2}. Therefore ii) grants that g1 ∧ g2 = χB g1 + χX\B g2 ∈ D[u], thus obtaining L4. By recalling Proposition 2.8, we conclude that D is pointwise local. i) + ii) =⇒ iii) Fix u1, u2 ∈ Sp(X), g1 ∈ D[u1] and g2 ∈ D[u2]. Recall that g1 ∨ g2 ∈ D[u1 ∧ u2] by axiom A4. Hence by using property ii) twice we obtain that χ{u1≤u2} g1 + χ{u1>u2} (g1 ∨ g2) ∈ D[u1 ∧ u2], χ{u2≤u1} g2 + χ{u2>u1} (g1 ∨ g2) ∈ D[u1 ∧ u2]. (2.8) The pointwise minimum between the two functions that are written in (2.8) – namely given by χ{u1<u2} g1 + χ{u2<u1} g2 + χ{u1=u2} (g1 ∧ g2) – belongs to the class D[u1 ∧ u2] as well by property L4, thus showing iii). iii) =⇒ i) First of all, let us prove L1. Fix a function u ∈ Sp(X) that is m-a.e. equal to some constant c ∈ R on a Borel set B ⊆ X. By using iii) and the fact that 0 ∈ D[0], we have that χ{u<c} g ∈ D(cid:2)(u − c) ∧ 0(cid:3) = D(cid:2) − (u − c)+(cid:3) = D(cid:2)(u − c)+(cid:3), χ{u>c} g ∈ D(cid:2)(c − u) ∧ 0(cid:3) = D(cid:2) − (c − u)+(cid:3) = D(cid:2)(c − u)+(cid:3). (2.9) 6 Since u − c = (u − c)+ − (c − u)+, we know from A2 and (2.9) that χ{u6=c} g = χ{u<c} g + χ{u>c} g ∈ D[u − c] = D[u], whence Ep(uB) ≤RB(χ{u6=c} g)p dm = 0. This proves the property L1. To show property L4, fix u ∈ Sp(X) and g1, g2 ∈ D[u]. Hence (2.6) with u1 = u2 := u (cid:3) simply reads as g1 ∧ g2 ∈ D[u], which gives L4. This proves that D is pointwise local. Remark 2.12 (L1 does not imply L2) In general, as we are going to show in the following example, it can happen that a D-structure satisfies L1 but not L2. Let G = (V, E) be a locally finite connected graph. The distance d(x, y) between two vertices x, y ∈ V is defined as the minimum length of a path joining x to y, while as a reference measure m on V we choose the counting measure. Notice that any function u : V → R is locally Lipschitz and that any bounded subset of V is finite. We define a D-structure on the metric measure space (V, d, m) in the following way: D[u] :=ng : V → [0, +∞](cid:12)(cid:12)(cid:12)(cid:12)(cid:12)u(x) − u(y)(cid:12)(cid:12) ≤ g(x) + g(y) for any x, y ∈ V with x ∼ yo (2.10) for every u : V → R, where the notation x ∼ y indicates that x and y are adjacent vertices, i.e. that there exists an edge in E joining x to y. We claim that D fulfills L1. To prove it, suppose that some function u : X → R is constant on some set B ⊆ V , say u(x) = c for every x ∈ B. Define the function g : V → [0, +∞) as g(x) :=( 0 c +(cid:12)(cid:12)u(x)(cid:12)(cid:12) if x ∈ B, if x ∈ V \ B. Hence g ∈ D[u] and RB gp dm = 0, so that Ep(uB) = 0. This proves the validity of L1. On the other hand, if V contains more than one vertex, then L2 is not satisfied. Indeed, consider any non-constant function u : V → R. Clearly any pseudo-gradient g ∈ D[u] of u is not identically zero, thus there exists x ∈ V such that Du(x) > 0. Since u is trivially constant on the set {x}, we then conclude that property L2 does not hold. (cid:4) Hereafter, we shall focus our attention on the pointwise local D-structures. Under these locality assumptions, one can show the following calculus rules for minimal pseudo-gradients, whose proof is suitably adapted from analogous results that have been proved in [2]. Proposition 2.13 (Calculus rules for Du) Let (X, d, m) be a metric measure space and let p ∈ (1, ∞). Consider a pointwise local D-structure on (X, d, m). Then the following hold: i) Let u ∈ Sp(X) and let N ⊆ R be a Borel set with L1(N ) = 0. Then the equality Du = 0 holds m-a.e. in u−1(N ). ii) Chain rule. Let u ∈ Sp(X) and ϕ ∈ LIP(R). Then ϕ′ ◦ u Du ∈ D[ϕ ◦ u]. More precisely, ϕ ◦ u ∈ Sp(X) and D(ϕ ◦ u) = ϕ′ ◦ u Du holds m-a.e. in X. 7 iii) Leibniz rule. Let u, v ∈ Sp(X) ∩ L∞(m). Then u Dv + v Du ∈ D[uv]. In other words, uv ∈ Sp(X) ∩ L∞(m) and D(uv) ≤ u Dv + v Du holds m-a.e. in X. Proof. Step 1. First, consider ϕ affine, say ϕ(t) = α t + β. Then ϕ′ ◦ u Du = α Du ∈ D[ϕ ◦ u] by Remark 2.2 and A2. Now suppose that the function ϕ is piecewise affine, i.e. there exists a is sequence (ak)k∈Z ⊆ R, with ak < ak+1 for all k ∈ Z and a0 = 0, such that each ϕ[ak,ak+1] an affine function. Let us denote Ak := u−1(cid:0)[ak, ak+1)(cid:1) and uk := (u ∨ ak) ∧ ak+1 for every index k ∈ Z. By combining L3 with the axioms A2 and A5, we can see that χAk Du ∈ D[uk] for every k ∈ Z. Called ϕk : R → R that affine function coinciding with ϕ on [ak, ak+1), we deduce from the previous case that ϕ′ k ◦ uk Duk ∈ D[ϕk ◦ uk] = D[ϕ ◦ uk], whence we have that ϕ′ ◦ uk χAk Du ∈ D[ϕ ◦ uk] by L5, A2 and L2. Let us define (vn)n ⊆ Sp(X) as vn := ϕ(0) + n −1 Xk=0(cid:0)ϕ ◦ uk − ϕ(ak)(cid:1) + Xk=−n(cid:0)ϕ ◦ uk − ϕ(ak+1)(cid:1) for every n ∈ N. Hence gn :=Pn k=−n ϕ′ ◦ uk χAk Du ∈ D[vn] for all n ∈ N by A2 and Remark 2.2. Given that loc(m) and gn → ϕ′ ◦ u Du in Lp(m) as n → ∞, we finally conclude one has vn → ϕ ◦ u in Lp that ϕ′ ◦ u Du ∈ D[ϕ ◦ u], as required. Step 2. We aim to prove the chain rule for ϕ ∈ C 1(R)∩ LIP(R). For any n ∈ N, let us denote n(t) → ϕ′(t) for all t ∈ R \ D. In particular, the functions gn := ϕ′ by ϕn the piecewise affine function interpolating the points (cid:0)k/2n, ϕ(k/2n)(cid:1) with k ∈ Z. We call D ⊆ R the countable set (cid:8)k/2n : k ∈ Z, n ∈ N(cid:9). Therefore ϕn uniformly converges to ϕ and ϕ′ n ◦ u Du converge m-a.e. to ϕ′ ◦ u Du by L2. Moreover, Lip(ϕn) ≤ Lip(ϕ) for every n ∈ N by construction, so that (gn)n is a bounded sequence in Lp(m). This implies that (up to a not relabeled subsequence) gn ⇀ ϕ′ ◦ u Du weakly in Lp(m). Now apply Mazur lemma: for any n ∈ N, n→ ϕ′ ◦ u Du there exists (αn strongly in Lp(m). Given that gn ∈ D[ϕn ◦ u] for every n ∈ N by Step 1, we deduce from i ϕi. Finally, it clearly i=n ⊆ [0, 1] such that PNn i = 1 and hn := PNn i=n αn i=n αn axiom A2 that hn ∈ D[ψn ◦ u] for every n ∈ N, where ψn := PNn loc(m), whence ϕ′ ◦ u Du ∈ D[ϕ ◦ u] by A5. holds that ψn ◦ u → ϕ ◦ u in Lp Step 3. We claim that i=n αn i )Nn i gi Du = 0 m-a.e. in u−1(K), for every K ⊆ R compact with L1(K) = 0. (2.11) For any n ∈ N \ {0}, define ψn := n d(·, K) ∧ 1 and denote by ϕn the primitive of ψn such that ϕn(0) = 0. Since each ψn is continuous and bounded, any function ϕn is of class C 1 and Lipschitz. By applying the dominated convergence theorem we see that the L1-measure of the ε-neighbourhood of K converges to 0 as ε ց 0, thus accordingly ϕn uniformly converges to idR as n → ∞. This implies that ϕn ◦ u → u in Lp loc(m). Moreover, we know from Step 2 that ψn ◦ u Du ∈ D[ϕn ◦ u], thus also χX\u−1(K) Du ∈ D[ϕn ◦ u]. Hence χX\u−1(K) Du ∈ D[u] by A5, which forces the equality Du = 0 to hold m-a.e. in u−1(K), proving (2.11). Step 4. We are in a position to prove i). Choose any m′ ∈ P(X) such that m ≪ m′ ≪ m and call µ := u∗m′. Then µ is a Radon measure on R, in particular it is inner regular. We can thus 8 find an increasing sequence of compact sets Kn ⊆ N such that µ(cid:0)N \Sn Kn(cid:1) = 0. We already know from Step 3 that Du = 0 holds m-a.e. in Sn u−1(Kn). Since u−1(N ) \Sn u−1(Kn) is m-negligible by definition of µ, we conclude that Du = 0 holds m-a.e. in u−1(N ). This shows the validity of property i). Step 5. We now prove ii). Let us fix ϕ ∈ LIP(R). Choose some convolution kernels (ρn)n and define ϕn := ϕ ∗ ρn for all n ∈ N. Then ϕn → ϕ uniformly and ϕ′ n → ϕ′ pointwise L1-a.e., whence accordingly ϕn ◦ u → ϕ ◦ u in Lp n ◦ u Du → ϕ′ ◦ u Du pointwise m-a.e. in X. Since ϕ′ n ◦ u Du ≤ Lip(ϕ) Du for all n ∈ N, there exists a (not relabeled) subsequence such that ϕ′ n ◦ u Du ∈ D[ϕn ◦ u] for all n ∈ N because the chain rule holds for all ϕn ∈ C 1(R) ∩ LIP(R), hence by combining Mazur lemma and A5 as in Step 2 we obtain that ϕ′◦u Du ∈ D[ϕ◦u], so that ϕ◦u ∈ Sp(X) and the inequality D(ϕ ◦ u) ≤ ϕ′ ◦ u Du holds m-a.e. in X. Step 6. We conclude the proof of ii) by showing that one actually has D(ϕ ◦ u) = ϕ′ ◦ u Du. We can suppose without loss of generality that Lip(ϕ) = 1. Let us define the functions ψ± as n ◦ u Du ⇀ ϕ′ ◦ u Du weakly in Lp(m). We know that ϕ′ loc(m) and ϕ′ ψ±(t) := ±t − ϕ(t) for all t ∈ R. Then it holds m-a.e. in u−1(cid:0){±ϕ′ ≥ 0}(cid:1) that which forces the equality D(ϕ ◦ u) = ±ϕ′ ◦ u Du to hold m-a.e. in the set u−1(cid:0){±ϕ′ ≥ 0}(cid:1). Du = D(±u) ≤ D(ϕ ◦ u) + D(ψ± ◦ u) ≤(cid:0)ϕ′ ◦ u + ψ′ This grants the validity of D(ϕ ◦ u) = ϕ′ ◦ u Du, thus completing the proof of item ii). Step 7. We show iii) for the case in which u, v ≥ c is satisfied m-a.e. in X, for some c > 0. Call ε := min{c, c2} and note that the function log is Lipschitz on the interval [ε, +∞), then choose any Lipschitz function ϕ : R → R that coincides with log on [ε, +∞). Now call C the ± ◦ u(cid:1) Du = Du, constant log(cid:0)kuvkL∞(m)(cid:1) and choose a Lipschitz function ψ : R → R such that ψ = exp on the interval [log ε, C]. By applying twice the chain rule ii), we thus deduce that uv ∈ Sp(X) and the m-a.e. inequalities D(uv) ≤ ψ′ ◦ ϕ ◦ (uv) D(cid:0)ϕ ◦ (uv)(cid:1) ≤ uv(cid:0)D log u + D log v(cid:1) Dv = uv(cid:18) Du u + v(cid:19) = u Dv + v Du. Therefore the Leibniz rule iii) is verified under the additional assumption that u, v ≥ c > 0. Step 8. We conclude by proving item iii) for general u, v ∈ Sp(X) ∩ L∞(m). Given any n ∈ N and k ∈ Z, let us denote In,k :=(cid:2)k/n, (k + 1)/n(cid:1). Call ϕn,k : R → R the continuous function that is the identity on In,k and constant elsewhere. For any n ∈ N, let us define un,k := u − vn,ℓ := v − k − 1 n ℓ − 1 n , un,k := ϕn,k ◦ u − , vn,ℓ := ϕn,ℓ ◦ v − k − 1 n ℓ − 1 n for all k ∈ Z, for all ℓ ∈ Z. Notice that the equalities un,k = un,k and vn,ℓ = vn,ℓ hold m-a.e. in u−1(In,k) and v−1(In,ℓ), respectively. Hence Dun,k = Dun,k = Du and Dvn,ℓ = Dvn,ℓ = Dv hold m-a.e. in u−1(In,k) and v−1(In,ℓ), respectively, but we also have that D(un,k vn,ℓ) = D(un,k vn,ℓ) is verified m-a.e. in u−1(In,k) ∩ v−1(In,ℓ). 9 Moreover, we have the m-a.e. inequalities 1/n ≤ un,k, vn,ℓ ≤ 2/n by construction. Therefore for any k, ℓ ∈ Z it holds m-a.e. in u−1(In,k) ∩ v−1(In,ℓ) that D(uv) ≤ D(un,k vn,ℓ) + k − 1 n Dvn,ℓ + Dun,k ℓ − 1 n k − 1 n Dvn,ℓ + ℓ − 1 n Dun,k ≤ vn,ℓ Dun,k + un,k Dvn,ℓ + ≤(cid:18)v + 4 n(cid:19) Du +(cid:18)u + 4 n(cid:19) Dv, where the second inequality follows from the case u, v ≥ c > 0, treated in Step 7. This implies that the inequality D(uv) ≤ u Dv + v Du + 4 (Du + Dv)/n holds m-a.e. in X. Given that n ∈ N is arbitrary, the Leibniz rule iii) follows. (cid:3) 3 Cotangent module associated to a D-structure It is shown in [4] that any metric measure space possesses a first-order differential structure, whose construction relies upon the notion of Lp(m)-normed L∞(m)-module. For completeness, we briefly recall its definition and we refer to [4,5] for a comprehensive exposition of this topic. Definition 3.1 (Normed module) Let (X, d, m) be a metric measure space and p ∈ [1, ∞). Then an Lp(m)-normed L∞(m)-module is any quadruplet (cid:0)M , k · kM , · , · (cid:1) such that i) (cid:0)M , k · kM(cid:1) is a Banach space, ii) (M , ·) is an algebraic module over the commutative ring L∞(m), iii) the pointwise norm operator · : M → Lp(m)+ satisfies f · v = f v m-a.e. for every f ∈ L∞(m) and v ∈ M , (3.1) kvkM =(cid:13)(cid:13)v(cid:13)(cid:13)Lp(m) for every v ∈ M . A key role in [4] is played by the cotangent module L2(T ∗X), which has a structure of L2(m)-normed L∞(m)-module; see [5, Theorem/Definition 1.8] for its characterisation. The following result shows that a generalised version of such object can be actually associated to any D-structure, provided the latter is assumed to be pointwise local. Theorem 3.2 (Cotangent module associated to a D-structure) Let (X, d, m) be any metric measure space and let p ∈ (1, ∞). Consider a pointwise local D-structure on (X, d, m). Then there exists a unique couple (cid:0)Lp(T ∗X; D), d(cid:1), where Lp(T ∗X; D) is an Lp(m)-normed L∞(m)-module and d : Sp(X) → Lp(T ∗X; D) is a linear map, such that the following hold: i) the equality du = Du is satisfied m-a.e. in X for every u ∈ Sp(X), ii) the vector space V of all elements of the form Pn partition of X and (ui)i ⊆ Sp(X), is dense in the space Lp(T ∗X; D). i=1 χBi dui, where (Bi)i is a Borel 10 Uniqueness has to be intended up to unique isomorphism: given another such couple (M , d′), there is a unique isomorphism Φ : Lp(T ∗X; D) → M such that Φ(du) = d′u for all u ∈ Sp(X). The space Lp(T ∗X; D) is called cotangent module, while the map d is called differential. Proof. (cid:12)(cid:12)Φ(ω)(cid:12)(cid:12) =Xi=1 i=1 χBi dui, with (Bi)i Borel partition of X and u1, . . . , un ∈ Sp(X). Notice that the requirements that Φ is L∞(m)-linear Uniqueness. Consider any element ω ∈ V written as ω = Pn and Φ ◦ d = d′ force the definition Φ(ω) :=Pn Xi=1 χBi d′ui = χBi Dui = Xi=1 i=1 χBi d′ui. The m-a.e. equality χBi dui = ω n n grants that Φ(ω) is well-defined, in the sense that it does not depend on the particular way of representing ω, and that Φ : V → M preserves the pointwise norm. In particular, one has that the map Φ : V → M is (linear and) continuous. Since V is dense in Lp(T ∗X; D), we can uniquely extend Φ to a linear and continuous map Φ : Lp(T ∗X; D) → M , which also preserves the pointwise norm. Moreover, we deduce from the very definition of Φ that the identity Φ(h ω) = h Φ(ω) holds for every ω ∈ V and h ∈ Sf(X), whence the L∞(m)-linearity of Φ follows by an approximation argument. Finally, the image Φ(V) is dense in M , which implies that Φ is surjective. Therefore Φ is the unique isomorphism satisfying Φ ◦ d = d′. Existence. First of all, let us define the pre-cotangent module as Pcm :=((cid:8)(Bi, ui)(cid:9)n i=1 (cid:12)(cid:12)(cid:12)(cid:12) n ∈ N, u1, . . . , un ∈ Sp(X), (Bi)n i=1 Borel partition of X ) . provided D(ui − vj) = 0 holds m-a.e. on Bi ∩ Cj for every i, j. The equivalence class of an We define an equivalence relation on Pcm as follows: we declare that(cid:8)(Bi, ui)(cid:9)i ∼(cid:8)(Cj, vj)(cid:9)j element (cid:8)(Bi, ui)(cid:9)i of Pcm will be denoted by [Bi, ui]i. We can endow the quotient Pcm/ ∼ with a vector space structure: [Bi, ui]i + [Cj, vj]j := [Bi ∩ Cj, ui + vj]i,j, λ [Bi, ui]i := [Bi, λ ui]i, (3.2) for every [Bi, ui]i, [Cj, vj]j ∈ Pcm/ ∼ and λ ∈ R. We only check that the sum operator is well-defined; the proof of the well-posedness of the multiplication by scalars follows along the same lines. Suppose that (cid:8)(Bi, ui)(cid:9)i ∼ (cid:8)(B′ k) = 0 m-a.e. on Bi ∩ B′ words D(ui − u′ whence accordingly k and D(vj − v′ k, u′ k)(cid:9)k and (cid:8)(Cj, vj)(cid:9)j ∼ (cid:8)(C ′ ℓ) = 0 m-a.e. on Cj ∩ C ′ ℓ)(cid:9)ℓ, in other ℓ for every i, j, k, ℓ, ℓ, v′ L5 k + v′ ≤ D(ui − u′ D(cid:0)(ui + vj)− (u′ This shows that (cid:8)(Bi ∩ Cj, ui + vj)(cid:9)i,j ∼ (cid:8)(B′ k)+ D(vj − v′ operator defined in (3.2) is well-posed. Now let us define k ∩ C ′ ℓ)(cid:1) ℓ) = 0 ℓ, u′ k + v′ holds m-a.e. on (Bi ∩ Cj)∩ (B′ k ∩ C ′ ℓ). ℓ)(cid:9)k,ℓ, thus proving that the sum (cid:13)(cid:13)[Bi, ui]i(cid:13)(cid:13)Lp(T ∗X;D) := n Xi=1(cid:18)ZBi (Dui)p dm(cid:19)1/p 11 for every [Bi, ui]i ∈ Pcm/ ∼ . (3.3) Such definition is well-posed: if (cid:8)(Bi, ui)(cid:9)i ∼(cid:8)(Cj, vj)(cid:9)j then for all i, j it holds that L5 Dui − Dvj ≤ D(ui − vj) = 0 m-a.e. on Bi ∩ Cj, i.e. that the equality Dui = Dvj is satisfied m-a.e. on Bi ∩ Cj. Therefore one has that Xi (cid:18)ZBi (Dui)p dm(cid:19)1/p =Xi,j (cid:18)ZBi∩Cj =Xj (cid:18)ZCj (Dui)p dm(cid:19)1/p (Dvj)p dm(cid:19)1/p , =Xi,j (cid:18)ZBi∩Cj (Dvj)p dm(cid:19)1/p which grants that k · kLp(T ∗X;D) in (3.3) is well-defined. The fact that it is a norm on Pcm/ ∼ easily follows from standard verifications. Hence let us define Lp(T ∗X; D) := completion of (cid:0)Pcm/ ∼, k · kLp(T ∗X;D)(cid:1), d : Sp(X) → Lp(T ∗X; D), du := [X, u] for every u ∈ Sp(X). Observe that Lp(T ∗X; D) is a Banach space and that d is a linear operator. Furthermore, given any [Bi, ui]i ∈ Pcm/ ∼ and h = Pj λj χCj ∈ Sf(X), where (λj)j ⊆ R and (Cj)j is a Borel partition of X, we set (cid:12)(cid:12)[Bi, ui]i(cid:12)(cid:12) :=Xi χBi Dui, h [Bi, ui]i := [Bi ∩ Cj, λj ui]i,j. One can readily prove that such operations, which are well-posed again by the pointwise locality of D, can be uniquely extended to a pointwise norm · : Lp(T ∗X; D) → Lp(m)+ and to a multiplication by L∞-functions L∞(m)×Lp(T ∗X; D) → Lp(T ∗X; D), respectively. Therefore the space Lp(T ∗X; D) turns out to be an Lp(m)-normed L∞(m)-module when equipped with the operations described so far. In order to conclude, it suffices to notice that du =(cid:12)(cid:12)[X, u](cid:12)(cid:12) = Du holds m-a.e. for every u ∈ Sp(X) and that [Bi, ui]i =Pi χBi dui for all [Bi, ui]i ∈ Pcm/ ∼, giving i) and ii), respectively. (cid:3) In full analogy with the properties of the cotangent module that is studied in [4], we can show that the differential d introduced in Theorem 3.2 is a closed operator, which satisfies both the chain rule and the Leibniz rule. Theorem 3.3 (Closure of the differential) Let (X, d, m) be a metric measure space and let p ∈ (1, ∞). Consider a pointwise local D-structure on (X, d, m). Then the differential operator d is closed, i.e. if a sequence (un)n ⊆ Sp(X) converges in Lp loc(m) and dun ⇀ ω weakly in Lp(T ∗X; D) for some ω ∈ Lp(T ∗X; D), then u ∈ Sp(X) and du = ω. loc(m) to some u ∈ Lp 12 Proof. Since d is linear, we can assume with no loss of generality that dun → ω in Lp(T ∗X; D) by Mazur lemma, so that d(un − um) → ω − dum in Lp(T ∗X; D) for any m ∈ N. In particular, one has un − um → u − um in Lp as n → ∞ for all m ∈ N, whence u − um ∈ Sp(X) and D(u − um) ≤ ω − dum holds m-a.e. for all m ∈ N by A5 and L5. Therefore u = (u − u0) + u0 ∈ Sp(X) and loc(m) and D(un − um) =(cid:12)(cid:12)d(un − um)(cid:12)(cid:12) → ω − dum in Lp(m) m→∞(cid:13)(cid:13)D(u − um)(cid:13)(cid:13)Lp(m) ≤ lim kdu − dumkLp(T ∗X;D) = lim kdun − dumkLp(T ∗X;D) = 0, lim m→∞ = lim m→∞ lim n→∞ kω − dumkLp(T ∗X;D) m→∞ which grants that dum → du in Lp(T ∗X; D) as m → ∞ and accordingly that du = ω. (cid:3) Proposition 3.4 (Calculus rules for du) Let (X, d, m) be any metric measure space and let p ∈ (1, ∞). Consider a pointwise local D-structure on (X, d, m). Then the following hold: i) Let u ∈ Sp(X) and let N ⊆ R be a Borel set with L1(N ) = 0. Then χu−1(N ) du = 0. ii) Chain rule. Let u ∈ Sp(X) and ϕ ∈ LIP(R) be given. Recall that ϕ ◦ u ∈ Sp(X) by Proposition 2.13. Then d(ϕ ◦ u) = ϕ′ ◦ u du. iii) Leibniz rule. Let u, v ∈ Sp(X) ∩ L∞(m) be given. Recall that uv ∈ Sp(X) ∩ L∞(m) by Proposition 2.13. Then d(uv) = u dv + v du. Proof. i) We have that du = Du = 0 holds m-a.e. on u−1(N ) by item i) of Proposition 2.13, thus accordingly χu−1(N ) du = 0, as required. ii) If ϕ is an affine function, say ϕ(t) = α t + β, then d(ϕ ◦ u) = d(α u + β) = α du = ϕ′ ◦ u du. Now suppose that ϕ is a piecewise affine function. Say that (In)n is a sequence of intervals whose union covers the whole real line R and that (ψn)n is a sequence of affine functions such n coincide L1-a.e. in the interior of In, that ϕIn n ◦ f df = ϕ′ ◦ f df holds m-a.e. on f −1(In) for all n, we have that d(ϕ ◦ f ) = d(ψn ◦ f ) = ψ′ = ψn holds for every n ∈ N. Since ϕ′ and ψ′ To prove the case of a general Lipschitz function ϕ : R → R, we want to approximate ϕ with a sequence of piecewise affine functions: for any n ∈ N, let us denote by ϕn the function so that d(ϕ ◦ u) = ϕ′ ◦ u du is verified m-a.e. on Sn u−1(In) = X. that coincides with ϕ at (cid:8)k/2n : k ∈ Z(cid:9) and that is affine on the interval (cid:2)k/2n, (k + 1)/2n(cid:3) for every k ∈ Z. readily check that, up to a not relabeled subsequence, ϕn → ϕ uniformly on R and ϕ′ pointwise L1-almost everywhere. The former grants that ϕn ◦ u → ϕ ◦ u in Lp that ϕ′ It is clear that Lip(ϕn) ≤ Lip(ϕ) for all n ∈ N. Moreover, one can n → ϕ′ loc(m). Given n −ϕ′p ◦u (Du)p → 0 n − ϕ′p ◦ u (Du)p dm → 0 n ◦ u du → ϕ′ ◦ u du in as n → ∞ by the dominated convergence theorem. In other words, ϕ′ the strong topology of Lp(T ∗X; D). Hence Theorem 3.3 ensures that d(ϕ ◦ u) = ϕ′ ◦ u du, thus proving the chain rule ii) for any ϕ ∈ LIP(R). pointwise m-a.e. by the latter above together with i), we obtain R ϕ′ n −ϕ′p ◦u (Du)p ≤ 2p Lip(ϕ)p (Du)p ∈ L1(m) for all n ∈ N and ϕ′ 13 iii) In the case u, v ≥ 1, we argue as in the proof of Proposition 2.13 to deduce from ii) that d(uv) uv whence we get d(uv) = u dv + v du by multiplying both sides by uv. = d log(uv) = d(cid:0) log(u) + log(v)(cid:1) = d log(u) + d log(v) = du u + dv v , In the general case u, v ∈ L∞(m), choose a constant C > 0 so big that u + C, v + C ≥ 1. By the case treated above, we know that d(cid:0)(u + C)(v + C)(cid:1) = (u + C) d(v + C) + (v + C) d(u + C) = (u + C) dv + (v + C) du = u dv + v du + C d(u + v), while a direct computation yields (3.4) (3.5) d(cid:0)(u + C)(v + C)(cid:1) = d(cid:0)uv + C(u + v) + C 2(cid:1) = d(uv) + C d(u + v). By subtracting (3.5) from (3.4), we finally obtain that d(uv) = u dv + v du, as required. This completes the proof of the Lebniz rule iii). (cid:3) Acknowledgements. This research has been supported by the MIUR SIR-grant 'Nonsmooth Differential Geometry' (RBSI147UG4). References [1] L. Ambrosio, Calculus, heat flow and curvature-dimension bounds in metric measure spaces. Proceedings of the ICM 2018, 2018. [2] L. Ambrosio, N. Gigli, and G. Savar´e, Calculus and heat flow in metric measure spaces and applica- tions to spaces with Ricci bounds from below, Invent. Math., 195 (2014), pp. 289–391. [3] J. Cheeger, Differentiability of Lipschitz functions on metric measure spaces, Geom. Funct. Anal., 9 (1999), pp. 428–517. [4] N. Gigli, Nonsmooth differential geometry - an approach tailored for spaces with Ricci curvature bounded from below. Accepted at Mem. Amer. Math. Soc., arXiv:1407.0809, 2014. [5] , Lecture notes on differential calculus on RCD spaces. Preprint, arXiv:1703.06829, 2017. [6] V. Gol'dshtein and M. Troyanov, Axiomatic theory of Sobolev spaces, Expositiones Mathematicae, 19 (2001), pp. 289–336. [7] P. Haj lasz, Sobolev spaces on an arbitrary metric space, Potential Analysis, 5 (1996), pp. 403–415. [8] J. Heinonen and P. Koskela, Quasiconformal maps in metric spaces with controlled geometry, Acta Math., 181 (1998), pp. 1–61. [9] N. Shanmugalingam, Newtonian spaces: an extension of Sobolev spaces to metric measure spaces, Rev. Mat. Iberoamericana, 16 (2000), pp. 243–279. [10] , A universality property of Sobolev spaces in metric measure spaces, Springer New York, New York, NY, 2009, pp. 345–359. [11] S. Timoshin, Regularity in metric spaces, (2006). PhD thesis, ´Ecole polytechnique f´ed´erale de Lausanne, available at: https://infoscience.epfl.ch/record/85799/files/EPFL TH3571.pdf. [12] C. Villani, In´egalit´es isop´erim´etriques dans les espaces m´etriques mesur´es [d'apr`es F. Cavalletti & A. Mondino]. S´eminaire Bourbaki, available at: http://www.bourbaki.ens.fr/TEXTES/1127.pdf. [13] , Synthetic theory of Ricci curvature bounds, Japanese Journal of Mathematics, 11 (2016), pp. 219– 263. 14
1911.11301
1
1911
2019-11-26T01:19:16
The recovery of complex sparse signals from few phaseless measurements
[ "math.FA", "cs.IT", "cs.IT" ]
We study the stable recovery of complex $k$-sparse signals from as few phaseless measurements as possible. The main result is to show that one can employ $\ell_1$ minimization to stably recover complex $k$-sparse signals from $m\geq O(k\log (n/k))$ complex Gaussian random quadratic measurements with high probability. To do that, we establish that Gaussian random measurements satisfy the restricted isometry property over rank-$2$ and sparse matrices with high probability. This paper presents the first theoretical estimation of the measurement number for stably recovering complex sparse signals from complex Gaussian quadratic measurements.
math.FA
math
THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS YU XIA AND ZHIQIANG XU Abstract. We study the stable recovery of complex k-sparse signals from as few phaseless measurements as possible. The main result is to show that one can employ ℓ1 minimization to stably recover complex k-sparse signals from m ≥ O(k log(n/k)) complex Gaussian random quadratic measurements with high probability. To do that, we establish that Gaussian random measurements satisfy the restricted isometry property over rank-2 and sparse matrices with high probability. This paper presents the first theoretical estimation of the measurement number for stably recovering complex sparse signals from complex Gaussian quadratic measurements. 9 1 0 2 v o N 6 2 ] . A F h t a m [ 1 v 1 0 3 1 1 . 1 1 9 1 : v i X r a 1. Introduction 1.1. Compressive Phase Retrieval. Suppose that x0 ∈ Fn is a k-sparse signal, i.e., kx0k0 ≤ k, where F ∈ {R, C}. We are interested in recovering x0 from yj = haj, x0i2 + wj, j = 1, . . . , m, where aj ∈ Fn is a measurement vector and wj ∈ R is the noise. This problem is called [21]. To state conveniently, we let A : Fn×n → Rm be a linear compressive phase retrieval map which is defined as A(X) = (a∗1Xa1, . . . , a∗mXam), where X ∈ Fn×n, aj ∈ Fn, j = 1, . . . , m. We abuse the notation and set A(x) := A(xx∗) = (ha1, xi2, . . . ,ham, xi2), where x ∈ Fn. We also set x0 := {cx0 : c = 1, c ∈ F}. The aim of compressive phase retrieval is to recover x0 from noisy measurements y = A(x0) + w, with y = (y1, . . . , ym)T ∈ Rm and w = (w1, ..., wm)T ∈ Rm. One question Yu Xia was supported by NSFC grant (11901143), Zhejiang Provincial Natural Science Foundation (LQ19A010008), Education Department of Zhejiang Province Science Foundation (Y201840082). Zhiqiang Xu was supported by NSFC grant (91630203, 11688101), Beijing Natural Science Foundation (Z180002). 1 2 YU XIA AND ZHIQIANG XU in compressive phase retrieval is: how many measurements yj, j = 1, . . . , m, are needed to stably recover x0? For the case F = R, in [8], Eldar and Mendelson established that m = O(k log(n/k)) Gaussian random quadratic measurements are enough to stably recover k- sparse signals x0. For the complex case, Iwen, Viswanathan and Wang suggested a two-stage strategy for compressive phase retrieval and show that m = O(k log(n/k)) measurements can guarantee the stable recovery of x0 [16]. However, the strategy in [16] requires the measurement matrix to be written as a product of two random matrices. Hence, it still remains open whether one can stably recover arbitrary complex k-sparse signal x0 from m = O(k log(n/k)) Gaussian random quadratic measurements. One of the aims of this paper is to confirm that m = O(k log(n/k)) Gaussian random quadratic measurements are enough to guarantee the stable recovery of arbitrary complex k-sparse signals. In fact, we do so by employing ℓ1 minimization. 1.2. ℓ1 minimization. Set A := (a1, . . . , am)T ∈ Fm×n. One classical result in compressed sensing is that one can use ℓ1 minimization to recover k-sparse signals, i.e., argmin x∈Fn {kxk1 : Ax = Ax0} = x0, provided that the measurement matrix A meets the RIP condition [4]. Recall that a matrix A satisfies the k-order RIP condition with RIP constant δk ∈ [0, 1) if (1 − δk)kxk2 2 ≤ kAxk2 2 ≤ (1 + δk)kxk2 2 holds for all k-sparse vectors x ∈ Fn. Using tools from probability theory, one can show that Gaussian random matrix satisfies k-order RIP with high probability provided m = O(k log(n/k)) [1]. Naturally, one is interested in employing ℓ1 minimization for compressive phase retrieval. We consider the following model: (1.1) argmin x∈Fn {kxk1 : Ax = Ax0}. Although the constrained conditions in (1.1) is non-convex, the model (1.1) is more amenable to algorithmic recovery. In fact, one already develops many algorithms for solving (1.1) [38, 39]. For the case F = R, the performance of (1.1) was studied in [30, 12, 35, 17]. Particularly, in [30], it was shown that if A ∈ Rm×n is a random Gaussian matrix with THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 3 m = O(k log(n/k)), then holds with high probability. The methods developed in [30] heavily depend on Ax0 is a argmin x∈Rn {kxk1 : Ax = Ax0} = ±x0 real vector and one still does not know the performance of ℓ1 minimization for recovering complex sparse signals. As mentioned in [30]: "The extension of these results to hold over C cannot follow the same line of reasoning". In this paper, we extend the result in [30] to complex case with employing the new idea about the RIP of quadratic measurements. 1.3. Our contribution. In this paper, we study the performance of ℓ1 minimization for recovering complex sparse signals from phaseless measurements y = A(x0) + w, where kwk2 ≤ ǫ. Particularly, we focus on the model (1.2) min x∈Cn kxk1 s.t. kA(x) − A(x0)k2 ≤ ǫ. Our main idea is to lift (1.2) to recover rank-one and sparse matrices, i.e., min X∈Hn×n kXk1 s.t. kA(X) − yk2 ≤ ǫ, rank(X) = 1. Throughout this paper, we use Hn×n to denote the n × n Hermitian matrices set. Thus we require A satisfies restricted isometry property over low-rank and sparse matrices: Definition 1.1. We say that the map A : Hn×n → Rm satisfies the restricted isometry property of order (r, k) if there exist positive constants c and C such that the inequality (1.3) ckXkF ≤ 1 mkA(X)k1 ≤ CkXkF holds for all X ∈ Hn×n with rank(X) ≤ r and kXk0,2 ≤ k. Throughout this paper, we use kXk0,2 to denote the number of non-zero rows in X. Since X is Hermitian, we have kXk0,2 = kX∗k0,2. We next show that Gaussian random map A satisfies RIP of order (2, k) with high probability provided m & k log(n/k). Theorem 1.2. Assume that the linear measurement A(·) is defined as A(X) = (a∗1Xa1, . . . , a∗mXam), with aj independently taken as complex Gaussian random vectors, i.e., aj ∼ N (0, 1 N (0, 1 2 In×n)i. If 2 In×n) + m & k log(n/k), 4 YU XIA AND ZHIQIANG XU under the probability at least 1 − 2 exp(−c0m), the linear map A satisfies the restricted isometry property of order (2, k), i.e. 0.12kXkF ≤ 1 mkA(X)k1 ≤ 2.45kXkF , for all X ∈ Hn×n with rank(X) ≤ 2 and kXk0,2 ≤ k (also kX∗k0,2 ≤ k). In the next theorem, we present the performance of (1.2) with showing that one can employ ℓ1 minimization to stably recovery complex k-sparse signals from the phaseless measurements provided A satisfies restricted isometry property of order (2, 2ak) with a proper choice of a > 0. Theorem 1.3. Assume that A(·) satisfy the RIP condition of order (2, 2ak) with RIP constant c, C > 0 satisfying (1.4) 4C √a − C a c − > 0. For any k sparse signals x0 ∈ Cn, the solution to (1.2) x# satisfies (1.5) , kx#(x#)∗ − x0x∗0k2 ≤ C1 2ǫ √m where Furthermore, we have C1 = 1 a + 4√a + 1 c − 4C√a − C a . (1.6) min c∈C,c=1kc · x# − x0k2 ≤ min{2√2C1ǫ/(√mkx0k2), 2q2√2C1 · √ǫ · (n/m)1/4}. According to Theorem 1.2, if aj, j = 1, . . . , m are complex Gaussian random vectors, then A satisfies RIP of order (2, 2ak) with constants c = 0.12 and C = 2.45 with high probability provided m & 2ak log(n/2ak). To guarantee (1.4) holds, it is enough to require a > (8C/c)2. Therefore, combining Theorem 1.2 and Theorem 1.3 with ǫ = 0, we can obtain the following corollary: Corollary 1.4. Suppose that x0 ∈ Cn is a k-sparse signal. Assume that A = (a1, . . . , am)T where aj, j = 1, . . . , m is Gaussian random vectors, i.e., aj ∼ N (0, 1 2 In×n)i. If m & k log(n/k), then 2 In×n) + N (0, 1 holds under the probability at least 1 − 2 exp(−c0m). Here c0 > 0 is an absolute constant. argmin x∈Cn {kxk1 : Ax = Ax0} = x0 THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 5 2. Proof of Theorem 1.2 We first introduce Bernstein-type inequality which plays a key role in our proof. sub-exponential random variables and K := Lemma 2.1. [28] Let ξ1, . . . , ξm be i.i.d. maxj kξjkψ1. Then for every ǫ > 0, we have ≥ ǫ  ≤ 2 exp(cid:18)−c0m min(cid:18) ǫ2 P  where c0 > 0 is an absolute constant. Xj=1 Xj=1 ξj − 1 m 1 m E( m m K 2 , ξj)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ǫ K(cid:19)(cid:19) , We next introduce some key lemmas needed to prove Theorem 1.2, and then present the proof of Theorem 1.2. Lemma 2.2. If z1, z2, z3 and z4 are independently drawn from N (0, 1). When t ∈ [−1, 0], we have 1 + z2 2 + tz2 3 + tz2 Ez2 4 = 2(cid:18) 1 + t2 1 − t (cid:19) . 4 = Ez2 Proof. When t = 0, we have Ez2 2 = 2. We consider the case where −1 ≤ t < 0. Taking coordinates transformation as z1 = ρ1 cos θ, z2 = ρ1 sin θ, z3 = ρ2 cos φ, and z4 = ρ2 sin φ, we obtain that 2 + tz2 3 + tz2 1 + z2 1 + z2 1 + z2 2 + tz2 3 + tz2 Ez2 (cid:19) dz1dz2dz3dz4 2 (cid:19) dρ1dρ2 1 + ρ2 ρ2 2 1 + z2 2 + tz2 3 + tz2 0 2π(cid:19)2ZR4 z2 4 =(cid:18) 1 2π(cid:19)2Z 2π =(cid:18) 1 =Z ∞ 0 Z ∞ =Zρ1>√−tρ2 +Zρ1≤√−tρ2 2t2 2 1 − t 1 − t = + 0 0 dθZ 2π ρ1ρ2ρ2 ρ1ρ2(ρ2 1 + tρ2 1 + tρ2 0 dφZ ∞ 0 Z ∞ 2 exp(cid:18)− 2) exp(cid:18)− 2 − ρ2 ρ1ρ2(cid:0)−tρ2 2(1 + t2) 1 − t = . 2 + z2 3 + z2 4 2 exp(cid:18)− 2 1 + z2 z2 1 + tρ2 4 exp(cid:18)− ρ1ρ2ρ2 1 + ρ2 ρ2 2 (cid:19) dρ1dρ2 2 2 (cid:19) dρ1dρ2 1 + ρ2 ρ2 2 1(cid:1) exp(cid:18)− 1 + ρ2 ρ2 2 2 (cid:19) dρ1dρ2 Here, we evaluate the last integrals using the integration by parts. One can also use Maple to check the integrals. (cid:3) 6 YU XIA AND ZHIQIANG XU Lemma 2.3. Set which is equipped with Frobenius norm. The covering number of X at scale ǫ > 0 is less than or equal to (cid:16) 9√2en X := {X ∈ Hn×n kXkF = 1, rank(X) ≤ 2, kXk0,2 ≤ k} ǫk (cid:17)4k+2 . Proof. Note that where and X = {X ∈ Hn×n : X = U ΣU∗, Σ ∈ Λ, U ∈ U}, Λ = {Σ ∈ R2×2 : Σ = diag(λ1, λ2), λ2 1 + λ2 2 = 1} U = {U ∈ Cn×2 : U∗U = I, kUk0,2 ≤ k} = ∪#T =kUT . Here T ⊂ {1, . . . , n}, and UT := {U ∈ Cn×2 : U∗U = I, U = UT,:}. The UT,: ⊂ Cn×2 is a matrix obtained by keeping the rows of U indexed by T and setting the rows off the index set T as 0. Note that kUkF = √2 for all U ∈ UT and that the real dimension of UT is at most 4k for any fixed support T with #T = k. We use QT to denote ǫ/3-net of UT with #QT ≤ (9√2/ǫ)4k. Then Qǫ := ∪#T =kQT is a ǫ/3-net of U with ≤ 9√2en ǫk !4k We use Λǫ to denote the ǫ/3-net of Λ with #Λǫ ≤ (9/ǫ)2. k (cid:17)k 9√2 ǫ !4k #Qǫ ≤(cid:16) en . Set Nǫ := {U ΣU∗ U ∈ Qǫ, and Σ ∈ Λǫ}. Then for any X = U ΣU∗ ∈ X , there exists U0Σ0U∗0 ∈ Nǫ with kU − U0kF ≤ ǫ/3 and kΣ − Σ0kF ≤ ǫ/3. So, we have kU ΣU∗ − U0Σ0U∗0kF ≤ kU ΣU∗ − U0ΣU∗kF + kU0ΣU∗ − U0Σ0U∗kF + kU0Σ0U∗ − U0Σ0U∗0kF ≤ kU − U0kFkΣU∗k + kU0kkΣ − Σ0kFkU∗k + kU0Σ0kkU∗ − U0kF ≤ ǫ. Therefore, Nǫ is an ǫ-net of X with #Nǫ ≤ #Qǫ · #Λǫ ≤ 9√2en ǫk !4k (9/ǫ)2 ≤ 9√2en ǫk !4k+2 THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 7 provided with n ≥ k and ǫ ≤ 1. (cid:3) We now have the necessary ingredients to prove Theorem 1.2. Proof of Theorem 1.2. Without loss of generality, we assume that kXkF = 1. We first consider EkA(X)k1. Noting that rank(X) ≤ 2 and kXkF = 1, we can write X in the form of X = λ1u1u∗1 + λ2u2u∗2, where λ1, λ2 ∈ R satisfying λ2 1,hu1, u2i = 0. Therefore, we obtain that 1 + λ2 2 = 1 and u1, u2 ∈ Cn satisfying ku1k2 = ku2k2 = a∗kXak = λ1u∗1ak2 + λ2u∗2ak2, where u∗1ak and u∗2ak are independently drawn from N (0, 1 Xj=1(cid:12)(cid:12)λ1u∗1aj2 + λ2u∗2aj2(cid:12)(cid:12) = (2.1) 4(cid:12)(cid:12) ξ =(cid:12)(cid:12)λ1z2 1 mkA(X)k1 = 2 + λ2z2 3 + λ2z2 1 + λ1z2 1 m m where the ξj are independent copies of the following random variable 2 ) + N (0, 1 Xj=1 1 m 2 )i. Then m ξj, 2 ) are independent. Without loss of generality, we assume that √2 2 , 1]. Note that ξ can also be rewritten as where z1, z2, z3, z4 ∼ N (0, 1 λ1 ≥ λ2 and hence λ1 ∈ [ (2.2) with t := λ2/λ1 satisfying t ≤ 1. Noting that EkA(X)k1 = E(ξ), we next consider E(ξ). According to (2.2), we have 1 + z2 2 + tz2 3 + tz2 ξ = λ1(cid:12)(cid:12)z2 (2.3) E(ξ) ≤ λ1E(z2 1 + z2 2 + z2 3 + z2 4) ≤ 2, as E(z2 j ) = 1 2 for j = 1, . . . , 4. On the other hand, when t ≥ 0, we obtain that 4(cid:12)(cid:12) √2 2 . (2.4) E(ξ) ≥ λ1E(z2 1 + z2 2) ≥ For t ∈ [−1, 0], Lemma 2.2 shows that (2.5) E(ξ) = λ1(cid:18) 1 + t2 1 − t (cid:19) ≥ 0.57. Combining (2.3), (2.4) and (2.5), we obtain that 0.57 ≤ E(ξ) ≤ 2. 8 YU XIA AND ZHIQIANG XU We next consider the bounds of kA(X)k1. Note that ξ is a sub-exponential variable with kξkψ1 ≤ P4 i=1 kzikψ1 ≤ 4. Here kxkψ1 := supp≥1 p−1(Exp)1/p. To state conveniently, we set X := {X ∈ Hn×n : kXkF = 1, rank(X) ≤ 2, kXk0,2 ≤ k}. We use Nǫ to denote an ǫ-net of the set X respect to Frobenius norm k · kF , i.e. for any X ∈ X , there exists X0 ∈ Nǫ such that kX − X0kF ≤ ǫ. Based on Lemma 2.1 and equality (2.1), we obtain that (2.6) holds with probability at least 1 − 2 · #Nǫ · exp(− c0 0.57 − ǫ0 ≤ 16 mǫ2 0). 1 mkA(X0)k1 ≤ 2 + ǫ0, for all X0 ∈ Nǫ Assume that X ∈ X with X0 ∈ Nǫ such that kX − X0kF ≤ ǫ. Note that A is continuous about X ∈ X and X is a compact set. We can set and UA := max X∈X 1 mkA(X)k1 U0 := max{UA : A satisfies (2.6)}. Note that rank(X − X0) ≤ 4 and kXk2,0 ≤ k,kX0k2,0 ≤ k. We can decompose X − X0 as X − X0 = X1 + X2 where X1, X2 ∈ X and hX1, X2i = 0 which leads to 1 mkA(X − X0)k1 = 1 mkA(X1 + X2)k1 ≤ ≤ U0kX1kF + U0kX2kF ≤ 1 mkA(X1)k1 + 1 mkA(X2)k1 √2U0kX1 + X2kF ≤ √2U0ǫ. We obtain that 1 mkA(X)k1 ≤ (2.7) 1 mkA(X0)k1 + 1 mkA(X − X0)k1 ≤ 2 + ǫ0 + √2U0ǫ. According to the definition of U0, (2.7) implies U0 ≤ 2+ ǫ0 +√2U0ǫ and hence which implies that U0 ≤ 2 + ǫ0 1 − √2ǫ . 1 mkA(X0)k1− We also have 1 1 − √2ǫ mkA(X)k1 ≥ Hence, we obtain that the following holds with probability at least 1− 2·#Nǫ·exp(− c0 16 mǫ2 0) (cid:18)0.57 − ǫ0 − 1 − √2ǫ(cid:19)kXkF , for all X ∈ X . 1 mkA(X−X0)k1 ≥ 0.57−ǫ0− mkA(X)k1 ≤(cid:18) 2 + ǫ0 √2U0ǫ ≥ 0.57−ǫ0− ǫ(cid:19)kXkF ≤ 2 + ǫ0 2 + ǫ0 1 − √2ǫ √2 √2 ǫ. 1 THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 9 Taking ǫ = ǫ0 = 0.1, according to Lemma 2.3, we obtain #Nǫ ≤(cid:16) 90√2en m ≥ O(k log(en/k)), we obtain that for all X ∈ X 1 mkA(X)k1 ≤ 2.45kXkF , 0.12kXkF ≤ k . Thus when (cid:17)4k+2 holds with probability at least 1 − 2 exp(−cm). (cid:3) 3. Proof of Theorem 1.3 We first introduce the convex k-sparse decomposition of signals which was proved inde- pendently in [3] and [37]. Lemma 3.1. [3, 37] Suppose that v ∈ Rp satisfying kvk∞ ≤ θ, kvk1 ≤ sθ where θ > 0 and s ∈ Z+. Then we have v = N Xi=1 λiui, 0 ≤ λi ≤ 1, λi = 1, N Xi=1 where ui is s-sparse with (supp(ui)) ⊂ supp(v), and kuik1 ≤ kvk1, kuik∞ ≤ θ. We also need the following lemma: Lemma 3.2. If x, y ∈ Cd, and hx, yi ≥ 0, then 1 2kxk2 kxx∗ − yy∗k2 F ≥ 2kx − yk2 2. Similarly, we have kxx∗ − yy∗k2 F ≥ 1 2kyk2 2kx − yk2 2. Proof. To state conveniently, we set a := kxk2, b := kyk2 and t := hx,yi kxk2kyk2 calculation shows that . A simple kxx∗ − yy∗k2 F − where 1 2kxk2 2kx − yk2 2 = h(a, b, t) h(a, b, t) := a4 + b4 − 2(ab)2t2 − 1 2 a2(a2 + b2 − 2abt). 10 YU XIA AND ZHIQIANG XU Hence, to this end, it is enough to show that h(a, b, t) ≥ 0 provided a, b ≥ 0 and 0 ≤ t ≤ 1. For any fixed a and b, h(a, b, t) achieves the minimum for either t = 0 or t = 1. For t = 0, we have (3.1) h(a, b, 0) = a4 + b4 − 1 2 a4 − 1 2 a2b2 = 1 2 (a2 − 1 2 b2)2 + 7 8 b4 ≥ 0. When t = 1, we have (3.2) 1 h(a, b, 1) = a4 + b4 − a2(a2 + b2) − 2(ab)2 + a3b 2 1 = (a − b)2( a2 + b2 + 2ab) ≥ 0 2 Combining (3.1) and (3.2), we arrive at the conclusion. (cid:3) Now we have enough ingredients to prove Theorem 1.3. Proof of Theorem 1.3. We assume that x# is a solution to (1.2). Noting exp(iθ)x# is also a solution to (1.2) for any θ ∈ R, without loss of generality, we can assume that hx#, x0i ∈ R and hx#, x0i ≥ 0. We consider the programming (3.3) min X∈Hn×n kXk1 s.t. kA(X) − yk2 ≤ ǫ, rank(X) = 1. Then a simple observation is that X # is the solution to (3.3) if and only if X # = x#(x#)∗. Set X0 := x0x∗0 and H := X # − X0 = x#(x#)∗ − x0x∗0. To this end, it is enough to consider the upper bound of kHkF . Denote T0 = supp(x0). Set T1 as the index set which contains the indices of the ak largest elements of x# in magnitude, and T2 contains the T c indices of the next ak largest elements, and so on. For simplicity, we set T01 := T0 ∪ T1 and ¯H := HT01,T01, where HS,T denotes the sub-matrix of H with the row set S and the column set T . Therefore, it is enough to consider kHkF . We claim that + 1(cid:19) k ¯HkF ≤ a + 4√a + 1 c − 4C√a − C which implies the conclusion (1.5). According to Lemma 3.2, we obtain that 2ǫ √m kHkF ≤ k ¯HkF + kH − ¯HkF ≤(cid:18) 1 a (3.4) + 4 √a 0 1 , a min c∈C,c=1kc · x# − x0k2 ≤ kx# − x0k2 ≤ √2kHkF /kx0k2 ≤ 1 a + 4√a + 1 c − 4C√a − C a 2√2ǫ √mkx0k2 . For the case where kx0k2 + kx0k1 ≤ 2√2C1 min c∈C,c=1kc · x# − x0k2 ≤ min(cid:26)2√2C1 ǫ√mkx0k2 2 + kx0k2kx0k1 1/n + kx0k2 ǫ √m ≥ kx0k2 ≥ kx0k2 2√2C1 ǫ √mkx0k2 ,kx0k2 + kx0k1(cid:27) . , we obtain that 1/√n = kx0k2 1(1/n + 1/√n), THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 11 Furthermore, we also have min c∈C,c=1kc · x# − x0k2 ≤ kx# − x0k2 ≤ kx#k2 + kx0k2 ≤ kx#k1 + kx0k2 ≤ kx0k1 + kx0k2. Here, we use kx#k1 ≤ kx0k1. Combining the above two inequalities, we obtain that which implies kx0k1 ≤p2√2C1·√ǫ·(n/m)1/4. Hence, when kx0k2+kx0k1 ≤ 2√2C1 we have ǫ√mkx0k2 , kx0k2 + kx0k1 ≤ 2kx0k1 ≤ 2q2√2C1 · √ǫ · (n/m)1/4. We arrive at the conclusion (1.6). We next turn to prove (3.4). The first inequality in (3.4) follows from (3.5) kH − ¯HkF ≤ (cid:18) 1 a + 4 √a(cid:19)k ¯HkF and the second inequality follows from (3.6) k ¯HkF ≤ 1 c − 4C√a − C a 2ǫ √m . To this end, it is enough to prove (3.5) and (3.6). Step 1: We first present the proof of (3.5). A simple observation is that (3.7) kH − ¯HkF ≤ Xi≥2,j≥2 = Xi≥2,j≥2 kHTi,TjkF + Xi=0,1Xj≥2 kHTi,TjkF + 2 Xi=0,1Xj≥2 kHTi,TjkF + Xj=0,1Xi≥2 kHTi,TjkF . kHTi,TjkF We first consider the first term on the right side of (3.7). Note that (3.8) Xi≥2,j≥2 kx# kHTi,TjkF = Xi≥2,j≥2 1 akkHT c = 0 ,T c 2 Tjk2 = Xi≥2 Tik2 · kx# 1 akkHT0,T0k1 ≤ 0k1 ≤ Tik2 kx#  1 akHT0,T0kF ≤ 1 akkx# 1 ak ¯HkF . 0 k2 ≤ T c 1 12 YU XIA AND ZHIQIANG XU Ti−1k1/√ak, for i ≥ 2. The second Here, the first inequality follows from kx# inequality is based on kH − HT0,T0k1 ≤ kHT0,T0k1. Indeed, according to kX #k1 ≤ kX0k1, we have Tik2 ≤ kx# we have T0,T0k1 ≤ kX0k1 − kX # kH − HT0,T0k1 = kX # − X # T0,T0k1 ≤ kX0 − X # We next turn toPi=0,1Pj≥2 kHTi,TjkF . When i ∈ {0, 1}, noting that kx# 1 (3.9) Xj≥2 √akx# The last inequality is based on kx#k1 ≤ kx0k1, which leads to kHTi,TjkF = kx# Tik2 ·Xj≥2 √akkx# 0 k1kx# Tjk2 ≤ Tik2 ≤ kx# T c 1 T0,T0k1 = kHT0,T0k1. Tjk2 ≤ kx# Tj−1k1/√ak, Tik2kx# T01 − x0k2. kx# 0 k1 ≤ kx0k1 − kx# T0k1 ≤ kx# T c T0 − x0k1 ≤ √kkx# T0 − x0k2 ≤ √kkx# T01 − x0k2. Substituting (3.8) and (3.9) into (3.7), we obtain that (3.10) kH − ¯HkF ≤ Xi≥2,j≥2 1 ak ¯HkF + kHTi,TjkF + Xi=0,1Xj≥2 T01k2kx# ≤ kHTi,TjkF + Xj=0,1Xi≥2 T01 − x0k2 ≤(cid:18) 1 kHTi,TjkF √a(cid:19)k ¯HkF . + a 4 Here, the first inequality is based on kx# follows from Lemma 3.2. T1k2 ≤ √2kx# T01k2, and the second inequality 2√2 √a kx# T0k2+kx# Step 2: We next prove (3.6). Since kA(H)k2 ≤ kA(X #) − yk2 + kA(X0) − yk2 ≤ 2ǫ, we have (3.11) 2ǫ √m ≥ 1 mkA( ¯H)k1 − 1 √mkA(H)k2 ≥ In order to get the lower bound of 1 lower bound of 1 k ¯Hk0,2 ≤ 2k, based on the RIP of measurement A, we obtain that (3.12) 1 mkA(H)k1 ≥ mkA( ¯H)k1 − 1 mkA( ¯H)k1 and the upper bound of 1 1 mkA( ¯H)k1 ≥ ck ¯HkF . We next consider an upper bound of 1 1 mkA(H − ¯H)k1. mkA(H − ¯H)k1, we need to estimate the mkA(H − ¯H)k1. As rank( ¯H) ≤ 2 and H − ¯H = (HT0,T c 01 + HT c mkA(H − ¯H)k1. Since H − ¯H can be written as 01,T0) + (HT1,T c 01,T1) + HT c 01 + HT c 01,T c 01, THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 13 we have 01+HT c 01,T0)k1+ 1 mkA(HT1,T c 1 mkA(HT0,T c (3.13) 1 mkA(H− ¯H)k1 ≤ We next calculate the upper bound of 1 respectively. According to the RIP condition, for i ∈ {0, 1}, we have (3.14) 1 mkA(HTi,T c mkA(HTi,T c 01 + HT c 01 +HT c 01+HT c 1 01,T1)k1+ 1 mkA(HT c 01,T c 01)k1. 01,Ti)k1, i ∈ {0, 1}, and 1 mkA(HT c 01,T c 01)k1, 01,Ti)k1 ≤Xj≥2 ≤ CXj≥2 2C √akx# ≤ mkA(HTi,Tj + HTj ,Ti)k1 ≤Xj≥2 (kx# )∗kF + kx# (x# Tj (x# Ti Tj Ti CkHTi,Tj + HTj,TikF Tik2kx# kx# )∗kF ) = 2CXj≥2 Tjk2 Tik2kx# T01 − x0k2. Here, the first inequality follows from HTi,T c 01 + HT c 01,Ti =Xj≥2 (HTi,Tj + HTj,Ti) =Xj≥2 (x# Ti (x# Tj )∗ + x# Tj (x# Ti )∗) and the last inequality is obtained by (3.9). Then it remains to estimate the upper bound of 1 mkA(HT c 01,T c 01)k1. Note that HT c 01,T c 01 = x# T c 01 (x# T c 01 )∗ T c 01k∞ ≤ kx# with kx# Φ := Diag(P h(x# T c and then Φ−1x# T c 01 01 T1k1/(ak). Set θ := max{kx# 01k1/(ak)}. We assume that )) is the diagonal matrix with diagonal elements being the phase of x# T c T1k1/(ak),kx# T c 01 is a real vector. According to Lemma 3.1, we have Φ−1x# T c 01 = N Xi=1 λiui, 0 ≤ λi ≤ 1, λi = 1, N Xi=1 where ui is ak-sparse, and which leads to kuik1 ≤ kx# T c 01k1, kuik∞ ≤ θ, kuik2 ≤pkuik1kuik∞ ≤qθkx# T c 01k1. 14 YU XIA AND ZHIQIANG XU If θ = kx# If θ = kx# T c T1k1/(ak), we have kuik2 ≤skx# T c ak T1k1kx# 01k1 ≤rkH − HT0,T0k1 ak 01k1/(ak), we have kuik2 ≤skx# T c T c 01k1kx# 01k1 ≤rkH − HT0,T0k1 ak ak ak =skHT1,T c 01k1 ≤rkHT0,T0k1 ak ≤rkHT0,T0kF a ≤rk ¯HkF a . 01,T c ak =skHT c 01k1 ≤rkHT0,T0kF ≤rkHT0,T0k1 ak a ≤rk ¯HkF a . Thus we can obtain that HT c 01,T c 01 = x# T c , for i = 1, . . . , N. a kuik2 ≤rk ¯HkF )∗ = N Xi=1 01 01 (x# T c λiΦui! N Xi=1 λiλjΦ(uiu∗j + uj u∗i )Φ−1 +Xi λiΦui!∗ i Φuiu∗i Φ−1, λ2 =Xi<j 01)k1 ≤Xi<j ≤Xi<j ≤ C k ¯HkF Cλ2 ikuiu∗ikF Cλiλjk(uiu∗j + uj u∗i )kF +Xi 2Cλiλjkuik2kujk2 +Xi a Xi = C k ¯HkF a λi!2 Cλ2 ikuik2 2 (3.15) Since (3.16) based on the RIP condition, we can obtain that 1 mkA(HT c 01,T c Here, the third line follows from (3.15). Now combing (3.14) and (3.16), we obtain that (3.17) 1 mkA(H − ¯H)k1 ≤ ≤ 1 mkA(HT0,T c 2C √akx# T0k2kx# 2√2C √a kx# ≤ ≤ C(cid:18) 4 √a T01k2kx# a(cid:19)k ¯HkF . 1 + 01 + HT c 01,T0)k1 + 2C √akx# 1 mkA(HT1,T c T1k2kx# 01 + HT c 01,T1)k1 + T01 − x0k2 + C k ¯HkF a T01 − x0k2 + 1 mkA(HT c 01,T c 01)k1 T01 − x0k2 + C k ¯HkF a THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 15 The last inequality uses Lemma 3.2. Based on (3.12), (3.17) and (3.11), we obtain that 1 mkA(H − ¯H)k1 2ǫ √m ≥ 1 mkA( ¯H)k1 − ≥ ck ¯HkF − C(cid:18) 4 √a + 1 a(cid:19)k ¯HkF =(cid:18)c − 4C √a − C a(cid:19)k ¯HkF . According to the condition (1.4), it implies that 1 k ¯HkF ≤ Thus, we arrive at the conclusion (3.6). c − 4C√a − C a 2ǫ √m . (cid:3) References [1] Richard Baraniuk, Mark Davenport, Ronald DeVore and Michael Wakin. A simple proof of the restricted isometry property for random matrices. Construction Approximation, 28(3), 253-263, 2008. [2] T Tony Cai, Xiaodong Li, Zongming Ma. Optimal rates of convergence for noisy sparse phase retrieval via thresholded wirtinger flow. The Annals of Statistics, 44(5):2221 -- 2251, 2016. [3] T Tony Cai and Anru Zhang. Sparse representation of a polytope and recovery in sparse signals and low-rank matrices. IEEE Transactions on Information Theory, 60(1):122-132, 2014. [4] Emmanuel J Cand`es, Justin Romberg, and Terence Tao, Stable signal recovery from incomplete and inaccurate measurements, Communications on Pure and Applied Mathematics, 59(8):1207-1223, 2006. [5] Emmanuel J Cand`es, Xiaodong Li, and Mahdi Soltanolkotabi. Phase retrieval via wirtinger flow: Theory and algorithms. IEEE Transactions on Information Theory, 61(4):1985 -- 2007, 2015. [6] Emmanuel J Cand`es, Thomas Strohmer, and Vladislav Voroninski. Phaselift: Exact and stable signal recovery from magnitude measurements via convex programming. Communications on Pure and Applied Mathematics, 66(8):1241 -- 1274, 2013. [7] Yuxin Chen and Emmanuel Cand`es. Solving random quadratic systems of equations is nearly as easy as solving linear systems. In Advances in Neural Information Processing Systems, pages 739 -- 747, 2015. [8] Yonina C. Eldar, Shahar Mendelson. Phase retrieval: Stability and recovery guarantees. In Applied and Computational Harmonic Analysis, 36(3): 473 -- 494, 2014. [9] James R Fienup and J C Dainty. Phase retrieval and image reconstruction for astronomy. Image Re- covery: Theory and Application, pages 231 -- 275, 1987. [10] James R Fienup. Phase retrieval algorithms: a comparison. Applied optics, 21(15):2758 -- 2769, 1982. [11] Bing Gao and Zhiqiang Xu. Phaseless recovery using the Gauss -- Newton method. IEEE Transactions on Signal Processing, 65(22):5885 -- 5896, 2017. [12] Bing Gao, Yang Wang and Zhiqiang Xu. Stable Signal Recovery from Phaseless Measurements. Journal of Fourier Analysis and Applications, 22(4):787 -- 808, 2016. [13] Ralph W Gerchberg and Saxton W Owen. A practical algorithm for the determination of phase from image and diffraction plane pictures. Optik, 35(2):237-250, 1972. [14] Tom Goldstein and Christoph Studer. Phasemax: Convex phase retrieval via basis pursuit. IEEE Trans- actions on Information Theory, 64(4):2675 -- 2689, 2018. [15] Robert W Harrison. Phase problem in crystallography. Journal of the Optical Society of America A- Optics Image Science and Vision, 10(5):1046 -- 1055, 1993. [16] Mark Iwen, Aditya Viswanathan, and Yang Wang. Robust sparse phase retrieval made easy. Applied and Computational Harmonic Analysis, 42(1):135 -- 142, 2015. [17] Ran Lu, On the strong restricted isometry property of Bernoulli random matrices, Journal of Approxi- mation Theory, 245:1-22, 2019. [18] Shahar Mendelson, Alain Pajor and Nicole Tomczak-Jaegermann. Uniform uncertainty principle for Bernoulli and subgaussian ensembles. Constructive Approximation, 28(3): 277-289, 2008. 16 YU XIA AND ZHIQIANG XU [19] Jianwei Miao, Tetsuya Ishikawa, Qun Shen, and Thomas Earnest. Extending x-ray crystallography to allow the imaging of noncrystalline materials, cells, and single protein complexes. Annual Review of Physical Chemistry, 59(1):387 -- 410, 2008. [20] Rick P Millane. Phase retrieval in crystallography and optics. Journal of the Optical Society of America A-Optics Image Science and Vision, 7(3):394 -- 411, 1990. [21] Matthew L Moravec, Justin K Romberg, and Richard G Baraniuk. Compressive phase retrieval. in Proceedings of SPIE, 6701, 2007. [22] Praneeth Netrapalli, Prateek Jain, and Sujay Sanghavi. Phase retrieval using alternating minimization. IEEE Transactions on Signal Processing, 63(18):4814 -- 4826, 2015. [23] Yaniv Plan and Roman Vershynin. One-bit compressed sensing by linear programming. Communications on Pure and Applied Mathematics, 66(8):1275 -- 1297, 2011. [24] Yaniv Plan and Roman Vershynin. The generalized lasso with non-linear observations. IEEE Transac- tions on information theory, 62(3):1528 -- 1537, 2016. [25] Yoav Shechtman, Yonina C Eldar, Oren Cohen, Henry Nicholas Chapman, Jianwei Miao, and Mordechai Segev. Phase retrieval with application to optical imaging: a contemporary overview. IEEE signal processing magazine, 32(3):87 -- 109, 2015. [26] Ju Sun, Qing Qu, and John Wright. A geometric analysis of phase retrieval. Foundations of Computa- tional Mathematics, 18(5):1131 -- 1198, 2018. [27] Yan Shuo Tan and Roman Vershynin. Phase retrieval via randomized kaczmarz: theoretical guarantees. Information and Inference: A Journal of the IMA, 2017. [28] R. Vershynin. Introduction to the non-asymptotic analysis of random matrices, Cambridge University Press, 2010. [29] Roman Vershynin. High-dimensional probability: An introduction with applications in data science, volume 47. Cambridge University Press, 2018. [30] Vladislav Voroninski and Zhiqiang Xu. A strong restricted isometry property, with an application to phaseless compressed sensing. Applied and Computational Harmonic Analysis, 40(2):386 -- 395, 2016. [31] Ir`ene Waldspurger. Phase retrieval with random Gaussian sensing vectors by alternating projections, IEEE Transactions on Information Theory, 64(5): 3301 -- 3312, 2018. [32] Adriaan Walther. The question of phase retrieval in optics. Journal of Modern Optics, 10(1):41 -- 49, 1963. [33] Gang Wang, Georgios B Giannakis, and Yonina C Eldar. Solving systems of random quadratic equations via truncated amplitude flow. IEEE Transactions on Information Theory, 64(2):773 -- 794, 2018. [34] Gang Wang, Liang Zhang, Georgios B Giannakis, Mehmet Akcakaya, and Jie Chen. Sparse phase retrieval via truncated amplitude flow. IEEE Transactions on Signal Processing, 66(2):479 -- 491, 2018. [35] Yang Wang and Zhiqiang Xu. Phase retrieval for sparse signals. Applied and Computational Harmonic Analysis, 37(3):531 -- 544, 2014. [36] Huishuai Zhang and Yingbin Liang. Reshaped wirtinger flow for solving quadratic system of equations. Advances in Neural Information Processing Systems, pages 2622 -- 2630, 2016. [37] Guangwu Xu and Zhiqiang Xu. On the ℓ1-Norm Invariant Convex k-Sparse Decomposition of Signals. Journal of the Operations Research Society of China, 1:537 -- 541, 2013. [38] P. Schniter and S. Rangan. Compressive Phase Retrieval via Generalized Approximate Message Passing, in Proceedings of Allerton Conference on Communication, Control, and Computing, Monticello, IL, USA, Oct. 2012. [39] Zai Yang, Cishen Zhang, and Lihua Xie. Robust compressive phase retrieval via L1 minimization with application to image reconstruction, arXiv:1302.0081. THE RECOVERY OF COMPLEX SPARSE SIGNALS FROM FEW PHASELESS MEASUREMENTS 17 Department of Mathematics, Hangzhou Normal University, Hangzhou 311121, China E-mail address: [email protected] LSEC, Inst. Comp. Math., Academy of Mathematics and System Science, Chinese Academy of Sciences, Beijing, 100091, China School of Mathematical Sciences, University of Chinese Academy of Sciences, Beijing 100049, China E-mail address: [email protected]
1502.00825
2
1502
2015-07-08T15:18:06
A general fibre theorem for moment problems and some applications
[ "math.FA" ]
The fibre theorem \cite{schm2003} for the moment problem on closed semi-algebraic subsets of $\R^d$ is generalized to finitely generated real unital algebras. As an application two new theorems on the rational multidimensional moment problem are proved. Another application is a characterization of moment functionals on the polynomial algebra $\R[x_1,\dots,x_d]$ in terms of extensions. Finally, the fibre theorem and the extension theorem are used to reprove basic results on the complex moment problem due to Stochel and Szafraniec \cite{stochelsz} and Bisgaard \cite{bisgaard}.
math.FA
math
A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS KONRAD SCHM UDGEN Abstract. The fibre theorem [11] for the moment problem on closed semi- algebraic subsets of Rd is generalized to finitely generated real unital algebras. As an application two new theorems on the rational multidimensional moment problem are proved. Another application is a characterization of moment functionals on the polynomial algebra R[x1, . . . , xd] in terms of extensions. Finally, the fibre theorem and the extension theorem are used to reprove basic results on the complex moment problem due to Stochel and Szafraniec [12] and Bisgaard [2]. AMS Subject Classification (2000). 44A60, 14P10. Key words: moment problem, real algebraic geometry 1. Introduction This paper deals with the classical multidimensional moment problem. A useful result on the existence of solutions is the fibre theorem [11] for closed semi-algebraic subsets of Rd. The crucial assumption for this theorem is the existence of sufficiently many bounded polynomials on the semi-algebraic set. The proof given in [11] was based on the decomposition theory of states on ∗-algebras. An elementary proof has been found by T. Netzer [7], see also M. Marshall [6, Chapter 4]. In the present paper we generalize the fibre theorem from the polynomial algebra R[x1, . . . , xd] to arbitrary finitely generated unital real algebras and we add further statements (Theorem 3). This general result and these additional statements allow us to derive a number of new applications. The first application developed in Section 3 is about the rational moment problem on semi-algebraic subsets of Rd. Pioneering work on this problem was done by J. Cimpric, M. Marshall, and T. Netzer in [5]. The main assumption therein is that the preordering is Archimedean. We essentially use the fibre theorem to go beyond the Archimedean case and derive two basic results on the rational multidimensional moment problem (Theorems 7 and 8). The second application given in Section 4 concerns characterizations of moment functionals in terms of extension to some appropriate larger algebra. We prove an extension theorem (Theorem 21) that provides a necessary and sufficient condition for a linear functional on R[x1, . . . , xd] being a moment functional. In the case d = 2 this result leads to a theorem of J. Stochel and F. H. Szafraniec [12] on the complex moment problem (Theorem 22). A third application of the general fibre theorem is a very short proof of a theorem of T.M. Bisgaard [2] on the two-sided complex moment problem (Theorem 23). Let us fix some definitions and notations which are used throughout this paper. A complex ∗-algebra B is a complex algebra equipped with an involution, that is, an antilinear mapping B ∋ b → b∗ ∈ B satisfying (bc)∗ = c∗b∗ and (b∗)∗ = b for j bj, where bj ∈ B. A linear functional L on B is called positive if L is nonnegative on b, c ∈ B. Let P B2 denote the set of finite sums Pj b∗ P B2, that is, if L(b∗b) ≥ 0 for all b ∈ B. j bj of hermitian squares b∗ 1 2 KONRAD SCHM UDGEN A ∗-semigroup is a semigroup S with a mapping s∗ → s of S into itself, called involution, such that (st)∗ = t∗s∗ and (s∗)∗ = s for s, t ∈ S. The semigroup ∗- algebra C[S] of S is the complex ∗-algebra is the vector space of all finite sums Ps∈S αss, where αs ∈ C, with product and involution (cid:0)Xs βtt(cid:1) :=Xs,t αsβtst, (cid:0)Xs αss(cid:1)(cid:0)Xt αss(cid:1)∗ :=Xs αs s∗. The polynomial algebra R[x1, . . . , xd] is abbreviated by Rd[x]. By a finite real algebra we mean a real algebra which is finite as a real vector space. 2. A generalization of the fibre theorem Throughout this section A is a finitely generated commutative real unital algebra. By a character of A we mean an algebra homomorphism χ : A → R satisfying χ(1) = 1. We equip the set A of characters of A with the weak topology. Let us fix a set {p1, . . . , pd} of generators of the algebra A. Then there is a algebra homomorphism π : R[x] → A such that π(xj ) = pj, j = 1, . . . , d. If J denotes the kernel of π, then A is isomorphic to the quotient algebra Rd[x]/J . Further, each character χ is completely determined by the point xχ := (χ(p1), . . . , χ(pd)) of Rd. For simplicity we will identify χ with xχ and write f (xχ) := χ(f ) for f ∈ A. Under this identification A is a real algebraic subvariety of Rd given by (1) A = Z(J ) := {x ∈ Rd : p(y) = 0 for p ∈ J }. In the special case A = R[x1, . . . , xd] we can take p1 = x1, . . . , pd = xd and get A = Rd. Definition 1. A preordering of A is a subset T of A such that T · T ⊆ T , T + T ⊆ T , 1 ∈ T , a2T ∈ T for all a ∈ A. LetP A2 denote the set of finite sumsPi a2 A is commutative, P A2 is invariant under multiplication and hence the smallest i of squares of elements ai ∈ A. Since preordering of A. For a preordering T of A, we define K(T ) = {x ∈ A : f (x) ≥ 0 for all f ∈ T }. If I is an ideal of A, its zero set is defined by Z(I) = {x ∈ A : f (x) = 0 for f ∈ I}. The main concepts are introduced in the following definition, see [11] or [6]. Definition 2. A preordering T of A has the • moment property (MP) if each T -positive linear functional L on A is a moment functional, that is, there exists a positive Borel measure µ ∈ M( A) such that (2) f (x) dµ(x) for all f ∈ A, L(f ) =Z A • strong moment property (SMP) if each T -positive linear functional L on A is a K(T ) -- moment functional, that is, there is a positive Borel measure µ ∈ M( A) such that supp µ ⊆ K(T ) and (2) holds. To state our main result (Theorem 3) we need some preparations. Suppose that T a finitely generated preordering of A and let f = {f1, . . . , fk} be a sequence of generators of T . We consider a fixed m-tuple h = (h1, . . . , hm) of elements hk ∈ A. Let h(K(T )) denote the subset of Rm defined by (3) For λ = (λ1, . . . , λm) ∈ Rm let K(T )λ be the subset of A given by K(T )λ = {x ∈ K(T ) : h1(x) = λ1, . . . , hm(x) = λm} h(K(T )) = {(h1(x), . . . , hm(x)); x ∈ K(T )}. A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS 3 and Tλ the preordering of A generated by the sequence f(λ) := {f1, . . . , fk, h1 − λ1·1, λ1·1 − h1·1, . . . , hm − λm·1, λm·1 − hm} Then K(T ) is the disjoint union of fibre sets K(T )λ = K(Tλ), where λ ∈ h(K(T )). Let Iλ be the ideal of A generated by h1 − λ1·1, . . . , hm − λm·1. Then we have Tλ := T + Iλ and the preordering Tλ/Iλ of the quotient algebra A/Iλ is generated by πλ(f) := {πλ(f1), . . . , πλ(fk)}, where πλ : A → A/Iλ denotes the canonical map. Further, let Iλ := I(Z(Iλ)) denote the ideal of all elements f ∈ A which vanish on the zero set Z(Iλ) of Iλ. Clearly, Iλ ⊆ Iλ and Z(Iλ) = Z(Iλ). Set Tλ := T + Iλ. Then Tλ/Iλ is a preordering of the quotient algebra A/Iλ. Note that in general we have Iλ 6= Iλ and equality holds if and only if the ideal j ∈ Iλ for finitely many elements aj ∈ A Iλ is real. The latter means that Pj a2 always implies that aj ∈ Iλ for all j. Theorem 3. Let A be a finitely generated commutative real unital algebra and let T be a finitely generated preordering of A. Suppose that h1, . . . , hm are elements of A that are bounded on the set K(T ). Then the following are equivalent: (i) T has property (SMP) (resp. (MP)) in A. (ii) Tλ satisfies (SMP) (resp. (MP)) in A for all λ ∈ h(K(T )). (ii)′ Tλ satisfies (SMP) (resp. (MP)) in A for all λ ∈ h(K(T )). (iii) Tλ/Iλ has (SMP) (resp. (MP)) in A/Iλ for all λ ∈ h(K(T )). (iii)′ Tλ/Iλ has (SMP) (resp. (MP)) in A/Iλ for all λ ∈ h(K(T )). Proposition 4(i) gives the implication (i)→(ii), Proposition 4(ii) yields the equiva- lences (ii)↔(iii) and (ii)′ ↔(iii)′, and Proposition 4(iii) implies equivalence (ii)↔(ii)′. The main assertion of Theorem 3 is the implication (ii)→(i). Its proof is lengthy and technically involved. In the proof given below we reduce the general case to the case Rd[x]. Proposition 4. Let I be an ideal and T a finitely generated preordering of A. Let I be the ideal of all f ∈ A which vanish on the zero set Z(I) of I. (i) If T satisfies (SMP) (resp. (MP)) in A, so does T + I. (ii) T + I satisfies (SMP) (resp. (MP)) in A if and only if (T + I)/I does in A/I. (iii) T + I obeys (SMP) (resp. (MP)) in A if and only if T + I does. Proof. (i): See e.g. [10], Proposition 4.8. (ii) See e.g. [10], Lemma 4.7. (iii): It suffices to show that both preorderings T + I and T + I have the same positive characters and the same positive linear functionals on A. For the sets of characters, using the equality Z(I) = Z(I) we obtain K(T + I) = K(T ) ∩ Z(I) = K(T ) ∩ Z(I) = K(T + I). Since T + I ⊆ T + I, a (T + I)-positive functional is trivially (T + I)-positive. Conversely, let L be a (T + I)-positive linear functional on A. Let f ∈ I and take a polynomial f ∈ Rd[x] such that π( f ) = f . Recall that A ∼= Rd[x]/J and Z(I) ⊆ A = Z(J ) by (1). Then I := π−1(I) is an ideal of Rd[x]. Let x ∈ Z(I)(⊆ Rd). Clearly, J ⊆ I and hence Z(I) ⊆ Z(J ) = A. Let g ∈ I and choose g ∈ I such that g = π(g). Since x annhilates J and x ∈ Z(I), we have g(x) = g(x) = 0. That is, x ∈ Z(I). From the equality Z(I) = Z(I) we obtain g(x) = 0 for g ∈ I. Thus, since f ∈ I, it follows that f (x) = f (x) = 0. That is, the polynomial f vanishes on Z(I). 4 KONRAD SCHM UDGEN Conversely, let L be a (T + I)-positive linear functional on A. Recall that A ∼= Rd[x]/J and A = Z(J ) by (1). Then I := π−1(I) is an ideal of Rd[x]. We verify that Z(I) ⊆ Z(I). Let x ∈ Z(I)(⊆ Rd). Clearly, J ⊆ I and hence Z(I) ⊆ Z(J ) = A. Let g ∈ I and choose g ∈ I such that g = π(g). Since x annhilates J and x ∈ Z(I), we have g(x) = g(x) = 0. That is, x ∈ Z(I) = Z(I). Now let f ∈ I. We choose f ∈ Rd[x] such that π( f ) = f . Then f (x) = f (x) for x ∈ Z(J ) = A. Hence, since f vanishes on Z(I) and Z(I) ⊆ Z(I), the polynomial f vanishes on Z(I). Therefore, by the real Nullstellensatz [6, Theorem 2.2.1, (3)], there are m ∈ N and g ∈P Rd[x]2 such that p := ( f )2m + g ∈ I. Upon multiplying p by some even power of f we can assume that 2m = 2k for some k ∈ N. Then (4) π(p) = f 2k + π(g) ∈ I, where π(g) ∈X A2. Being (T + I)-positive, L annihilates I and is nonnegative on P A2. Hence, by (4), we obtain 0 = L(p) = L(cid:0)f 2k(cid:1) + L(π(g)), L(π(g)) ≥ 0, L(cid:0)f 2k(cid:1) ≥ 0. This implies that L(f 2k inequality holds. By a repeated application of this inequality we derive ) = 0. Since L is nonnegative onP A2, the Cauchy-Schwarz L(f )2k L(1)2k−1 ≤ L(f 4)2k−2 L(1)2k−2+2k−1 ≤ . . . ≤ L(f 2)2k−1 ≤ L(f 2k )L(1)1+···+2k−1 = 0. Therefore, L(f ) = 0. That is, L annihilates I. Hence L is (T + I)-positive. This completes the proof of (iii). (cid:3) Proof of the implication (ii)→(i) of Theorem 3: As noted at the beginning of this section, A is (isomorphic to) the quotient algebra Rd[x]/J and A = Z(J ) by (1). We choose polynomials hj ∈ Rd[x] such that π(hj) = hj. Then T := π−1(T ) is a preordering of Rd[x]. Since J ⊆ T , each T - positive character of Rd[x] annihilates J , so it belongs to A = Z(J ). This implies that K( T ) = K(T ) and hj(x) = hj(x) for x ∈ K(T ), so that h(K( T )) = h(K(T )). Since hj is bounded on K(T ) by (ii), so is hj on K( T ). For λ ∈ h(K(T )), Iλ := π−1(Iλ) is an ideal of Rd[x]. It is easily verified that ( T )λ = T + Iλ. We will apply Proposition 4(ii) twice to A = Rd[x]/J , that is, with A replaced by Rd[x] and I by J in Proposition 4(ii). Because Tλ = (( T )λ + J )/J has (SMP) (resp. (MP)) in A = Rd[x]/J by assumption (ii), so does ( T )λ in Rd[x] by the if direction of Proposition 4(ii). For the algebra Rd[x] the implication (ii)→(i) of Theorem 3 was proved in [11]. Since ( T )λ = T + Iλ and hj is bounded on K( T ), this result applies to T . Hence T obeys (SMP) (resp. (MP)) in Rd[x] and so does T = ( T + J )/J in A = Rd[x]/J by the only if direction of Proposition 4(ii). (cid:3) We state the special case T =P A2 of Proposition 4(ii) separately as Corollary 5. If I is a ideal of A, then I +P A2 obeys (MP) (resp. (SMP)) in A if and only if P(A/I)2 does in A/I. Our applications to the rational moment problem in Section 3 are based on the following corollary. Corollary 6. Let us retain the notation of Theorem 3. Suppose that for each λ ∈ h(K(T )) there exist a finitely generated real unital algebra Bλ and a surjective algebra homomorphism ρλ : Bλ → A/Iλ (or ρλ : Bλ → A/Iλ) such that P B2 obeys (MP) in Bλ. Then T has (MP) in A. λ A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS 5 Proof. Let λ ∈ h(K(T )). We denote by J λ the kernel of the homomorphism ρλ : λ has (MP) Bλ → A/Iλ. Since P B2 in Bλ. Therefore P(Bλ/J λ)2 satisfies (MP) in Bλ/J λ by Corollary 5. Since the Hence P(A/Iλ)2 has (MP) in A/Iλ as well. Consequently, since P(A/Iλ)2 ⊆ λ has (MP) in Bλ, it is obvious that J λ +P B2 algebra homomorphism ρλ is surjective, the algebra Bλ/J λ is isomorphic to A/Iλ. T /Iλ, T /Iλ obeys (MP) in A/Iλ. Then T has (MP) by Theorem 3,(iii)→(i). The proof under the assumption ρλ : Bλ → A/Iλ is almost the same; in this (cid:3) case the implication (iii)′ →(i) of Theorem 3 is used. Theorem 3 is formulated for a commutative unital real algebra A. In Sections 5 and 6 we are concerned with commutative complex semigroup ∗-algebras. This case can be easily reduced to Theorem 3 as we discuss in what follows. If B is a commutative complex ∗-algebra, its hermitian part A = Bh := {b ∈ B : b = b∗} is a commutative real algebra. Conversely, suppose that A is a commutative real algebra. Then its com- plexification B := A + iA is a commutative complex ∗-algebra with involution (a1 + ia2)∗ := a1 − ia2 and scalar multiplication (α + iβ)(a1 + ia2) := αa1 − βa2 + i(αa2 + βa1), α, β ∈ R, a1, a2 ∈ A, (5) 1 + a2 2. b∗b = (a1 − ia2)(a1 + ia2) = a2 and A is the hermitian part Bh of B. Let b ∈ B. Then we have b = a1 + ia2 with a1, a2 ∈ A and since A is commutative, we get 1 + a2 2 + i(a1a2 − a2a1) = a2 Hence, if T is a preordering of A, then b∗b T ⊆ T for b ∈ B. Further, each R-linear functional L on A has a unique extension L to a C-linear P B2 =P A2. functional on B. By (5), L is nonnegative onP A2 if and only of L is nonnegative on P B2, that is, L is a positive linear functional on B. In particular, 3. Rational moment problems in Rd We begin with some notation and some preliminaries to state the main results. For q ∈ Rd[x] and a subset D ⊆ Rd[x] we put Z(q) := {x ∈ Rd : q(x) = 0}, ZD := ∪q∈D Z(q). Let D(Rd[x]) denote the family of all multiplicative subsets D of Rd[x] such that 1 ∈ D and 0 /∈ D. The real polynomials in a single variable y are denoted by R[y]. Let {f1, . . . , fk} be a k-tuple of polynomials of Rd[x] and D ∈ D(Rd[x]). Then A := D−1Rd[x] is a real unital algebra which contains Rd[x] as a subalgebra. Let T be the pre- ordering of A generated by f1, . . . , fk. Further, we fix an m-tuple h = {h1, . . . , hm} of elements hj ∈ A. For λ ∈ Rd let Iλ be the ideal of A generated by hj − λj · 1, j = 1, . . . , m. Recall that h(K(T )) is defined by (3). We consider the following assumptions: (i) The functions h1, . . . , hm ∈ A are bounded on the set h(K(T ). (ii) For all λ ∈ h(K(T )) there are a finitely generated set Eλ ∈ D(R[y]), a finite commutative unital real algebra Cλ, and a surjective homomorphism ρλ : E −1 λ R[y] ⊗ Cλ → D−1R[x]/Iλ ≡ A/Iλ. The following two theorems are the main results of this section. 6 KONRAD SCHM UDGEN Theorem 7. Suppose that the multiplicative set D ∈ D(Rd[x]) is finitely generated and assume (i) and (ii). Then the (finitely generated) real algebra A obeys (MP). For each T -- positive linear functional L on A there is a positive regular Borel mea- sure µ on A ∼= Rd\ZD such that L(f ) =R A f dµ for all f ∈ A. If the set D is not finitely generated, a number of technical difficulties appear: In general the algebra A is no longer finitely generated and the character space A is not locally compact in the corresponding weak topology. Recall that the fibre theorem requires a finitely generated algebra, because it is based on the Krivine- Stengle Positivstellensatz. However, circumventing these technical problems we have the following general result concerning the multidimensional rational moment problem. Theorem 8. Let D0 ∈ D(Rd[x]) and let {f1, . . . , fk} be a k-tuple of polynomials of Rd[x]. Suppose that for each finitely generated subset D ∈ D(Rd[x]) of D0 there exists an m-tuple h = {h1, . . . , hm} of elements hj ∈ A := D−1Rd[x] such that D, A, and the preordering T of A generated by f1, . . . , fk satisfy assumptions (i) and (ii). Let T0 be the preordering of the algebra A0 := D−1 Rd[x] generated by f1, . . . , fk. Then for each T0 -- positive linear functional L0 on A0 there exists a regular positive Borel measure µ on Rd such that f ∈ L1(Rd; µ) and 0 (6) L0(f ) =ZRd f dµ for all f ∈ A0. The proofs of these theorems require a number of preparatory steps. The fol- lowing two results for d = 1 are crucial and they are of interest in itself. Proposition 9. Suppose that E ∈ D(R[y]). Then for each positive linear func- tional L on the real algebra B := E −1R[y] there exists a positive regular Borel measure µ on R such that f ∈ L1(R, µ) and (7) L(f ) =Z f (x) dµ(x) for f ∈ B. If E is finitely generated, so is the algebra B and P B2 satisfies (MP) in B. In this case, ZE is a finite set and µ(ZE ) = 0. The proof of Proposition 9 is given below. First we derive the following corollary. Corollary 10. Suppose that E ∈ D(R[y]) is finitely generated and C is a finite commutative real unital algebra. Then P (E −1R[y]⊗C)2 obeys (MP) in the algebra E −1R[y] ⊗ C. Proof. Throughout this proof we abbreviate A := E −1R[y] ⊗ C and B := E −1R[y]. Suppose that L is a positive linear functional on the algebra E −1R[y] ⊗ C. Let I denote the ideal of elements of C which vanish on all characters of C. Since C is a finite algebra, the character set C is finite and each positive linear functional on C is a linear combination of characters, so it vanishes on I. Let f ∈ B. Clearly, L(f 2 ⊗·) is a positive linear functional on C. Thus C has a positive linear functional, C is not empty, say C = {η1, . . . , ηn} with n ∈ N. Further, L(f 2 ⊗ c) = 0 for c ∈ I. Since the unital algebra C is spanned by its squares, we therefore have L(g ⊗ c) = 0 for all g ∈ B and c ∈ C. We choose elements e1, . . . , en ∈ C such that ηj(ek) = δjk for j, k = 1, . . . , n and define linear functionals Lk on the algebra B by Lk(·) = L(·⊗ek). Since ηj(e2 k − ek ∈ I. Hence k − ek) = 0 for all j, we conclude that e2 Lk(f 2) = L(f 2 ⊗ ek) = L(f 2 ⊗ e2 k) = L((f ⊗ ek)2) ≥ 0, f ∈ B. A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS 7 That is, Lk is a positive linear functional on B. Therefore, by Proposition 9, there exists a positive regular Borel measure µk on R such that Lk(f ) = R f dµk for f ∈ B and µk(ZE ) = 0. A = ∪n From Lemma 12 below we obtain B = {χt; t ∈ R\ZE}. It is easily verified that j=1{χt ⊗ ηj; t ∈ R\ZE}. We define a positive regular Borel measure µ on A by µ(Pj Mj ⊗ ηj) =Pj µj(Mj) for Borel sets Mj ⊆ R\ZE. Let f ∈ B and c ∈ C. Since the element c −Pj ηj(c)ej is annihilated by all ηk, it belongs to I. Therefore, L(f ⊗ c) =Pj ηj(c)L(f ⊗ ej) and hence L(f ⊗ c) = nXj=1 =Xj ηj(c)L(f ⊗ ej) =Xj ηj (c)Z A P A2 obeys (MP) in the algebra A. (f ⊗ ej) dµ =Z A ηj(c)Z B f dµj (f ⊗ c) dµ, that is, L is given by the integral with respect to the measure µ. This shows that (cid:3) In the proof of Proposition 9 we use the following lemmas. We retain the notation established above. Lemma 11. Let B be as in Proposition 9 and let f = p p ∈ R[y]. Then f (y) ≥ 0 for all y ∈ R\Z(q) if and only if f ∈P B2. Proof. Suppose that f ≥ 0 on R\Z(q). Then q2f = pq ≥ 0 on R\Z(q) and hence on the whole real line, since Z(q) is empty or finite. Since each nonnegative polynomial in one (!) variable is a sum of two squares in R[y], we have pq = p2 1 + p2 2 with p1, p2 ∈ R[y]. Therefore, since q ∈ E, we get f = ( p1 converse implication is obvious. q )2 ∈ P B2. The q )2 + ( p1 (cid:3) q ∈ B, where q ∈ E and Lemma 12. For t ∈ Rd\ZD, p ∈ Rd[x], and q ∈ D, we define χt( p A = {χt; t ∈ Rd\ZD}. q ) = p(t) q(t) . Then Proof. First we note that for any t ∈ Rd\ZD, χt is a well-defined character on A, that is, χt ∈ A. Conversely, suppose that χ ∈ A. Put tj = χ(xj ) for j = 1, . . . , d. Then t = (t1, . . . , td) ∈ Rd and χ(p(x)) = p(χ(x1), . . . χ(xd))) = p(t) for p ∈ Rd[x]. For q ∈ D we have 1 = χ(1) = χ(qq−1) = χ(q)χ(q−1) = q(t)χ(q−1). Hence q(t) 6= 0 for all q ∈ D, that is, t ∈ Rd\ZD. Further, χ(q−1) = q(t)−1 and therefore χ( p q ) = χ(p)χ(q−1) = p(t)q(t)−1 = χt( p (cid:3) q ). Thus χ = χt. Set E := A + Cc(Rd). Let f = p q ∈ A, where p ∈ Rd[x], q ∈ D, and ϕ ∈ Cc(Rd). Then f + ϕ ∈ E and we define (8) f + ϕ ≥ 0 if f (x) + ϕ(x) ≥ 0 for x ∈ Rd\Z(q). Since Z(q) is nowhere dense in Rd, (E, ≥) is a real ordered vector space. In the proofs of the following two lemmas we modify some arguments from Choquet's approach to the moment problem based on adapted spaces, see [3] or [1]. Lemma 13. For each positive linear functional L on the ordered vector space (E, ≥) there is a regular positive Borel measure µ on Rd such that (9) L(f ) =Z f dµ f or f ∈ A. 8 KONRAD SCHM UDGEN Proof. Fix p ∈ Rd[x] and q ∈ D. Put g = p2+1 by setting g = +∞ for x ∈ Z(q) and abbreviate kxk2 = x2 arbitrary. Obviously, the set . We define g on the whole space Rd d. Let ε > 0 be 1 + · · · + x2 q2 Kε := {x ∈ Rd : kxk2 + 1 + g ≤ ε−1} is compact and q2(x) ≥ ε for x ∈ Kε. Hence the compact set Kε and the closed set U(q) := {x ∈ Rd : q2(x) ≤ ε/2} are disjoint, so by Urysohn's lemma there exists a function ηε ∈ Cc(Rd) such that ηε = 1 on Kε, ηε = 0 on U(q) and 0 ≤ ηε ≤ 1 on Rd. Then we have gηε ∈ Cc(Rd) and (10) g(x) ≤ g(x)ηε(x) + ε[(kxk2 + 1)g(x) + g2(x)] for x ∈ Rd\Z(q). (Indeed, by the definitions of ηε and Kε we have g(x) = g(x)ηε(x) if x ∈ Kε and g(x) < ε[(kxk2 + 1)g(x) + g2(x)] if x /∈ Kε.) Since the restriction of L to Cc(Rd) is a positive linear functional on Cc(Rd), by Riesz' theorem there exists a positive regular Borel measure µ on Rd such that (11) L(ϕ) =Z ϕ dµ for ϕ ∈ Cc(R). Using that L is a positive functional with respect to the order relation ≥ on E and inequality (10) and applying (11) with ϕ = gηε we derive L(g) ≤ L(gηε) + εL((cid:0)kxk2 + 1)g + g2(cid:1) =Z gηεdµ + εL(cid:0)(kxk2 + 1)g + g2(cid:1) ≤Z gdµ + εL(cid:0)(kxk2 + 1)g + g2(cid:1). Since ε > 0 was arbitrary, we conclude that L(g) ≤R gdµ. To prove the converse inequality, let Uq be the set of η ∈ Cc(Rd) such that 0 ≤ η ≤ 1 on Rd and η vanishes in some neighborhood of Z(q). Since Z(q) is nowhere dense in Rd and using again by the positivity of L and (11) we obtain Z gdµ = sup η∈Uq Z gη dµ = sup η∈Uq L(gη) ≤ L(g). Therefore, L(g) =R gdµ. Thus we have proved the equality in (9) for elements of the form g = p2+1 q2 Setting p = 0, (9) holds for all elements 1 q2 by linearity. Since A is spanned by its squares, the equality (9) holds for all f ∈ A. (cid:3) q2 , where q ∈ D, and hence for p2 . for f ∈ A. Lemma 14. Let (µi)i∈I be a net of positive regular Borel measures on Rd which converges vaguely to a positive regular Borel measure µ on Rd. Let L is a linear functional on A such that L(f ) =R f dµi for i ∈ I and f ∈ A. Then L(f ) =R f dµ proof. For η ∈ Uq we have f η ∈ Cc(Rd) and hence limiR f η dµi =R f η dµ. Then q , where q ∈ D. Let Uq be as in the preceding Proof. Let f ∈ E, f ≥ 0 and f = p Z f dµ = sup η∈UqZ f η dµ = sup η∈Uq i Z f ηdµi ≤ lim i Z f dµi = L(f ), lim (12) that is, f ∈ L1(µ). Since E is spanned by its squares, E ⊆ L1(X; µ). Now we use notation and facts from this proof of Lemma 13 and take an element g = p2+1 q2 as thererin Setting h = (kxk2 + 1)g + g2, inequality (10) means that g ≤ gηε + εh. A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS 9 Using this inequality and (12), first with f = g and then with f = h, we derive (cid:12)(cid:12)(cid:12)(cid:12)L(g) −Z g dµ(cid:12)(cid:12)(cid:12)(cid:12) = L(g) −Z g dµ =Z g dµi −Z g dµ =Z (g − gηε) dµi −Z (g − gηε) dµ +Z gηε dµi −Z gηε dµ ≤ ε(cid:18)Z h dµi +Z hdµ(cid:19) +Z gηε dµi −Z gηε dµ ≤ ε(L(h) + L(h)) +Z gηε dµi −Z gηε dµ. Since gηε ∈ Cc(Rd), limi R gηε dµi =R gηε dµ by the vague convergence. Therefore, taking limi in the preceding inequality yields L(g) −R g dµ ≤ 2 εL(h). Hence L(g) =R g dµ by letting ε → +0. Arguing as in the proof of Lemma 13 this implies that L(f ) =R f dµ for all f ∈ A. and only if f ∈P B2. Hence L(f ) ≥ 0. Since R[y] ⊆ B, A is a cofinal subspace of Proof of Proposition 9. Set E := B + Cc(R). Let (E, ≥) be the real ordered vector space defined above, see (8). Let f ∈ B. By Lemma 11, we have f ≥ 0 if the ordered vector space (E, ≥). Therefore, L can be extended to a positive linear functional L on (E, ≥). Applying Lemma 13, with d = 1 and A replaced by B, to the functional L yields (8). (cid:3) Suppose in addition that E is finitely generated. Let q1, . . . , qr be generators of E. Then the algebra B is generated by y, q−1 r , so B is finitely generated. Further, ZE = ∩jZ(qj), so the set ZE is finite. If q ∈ E, then q−2 ∈ A and hence L(q−2) = R q−2dµ < ∞. Therefore, µ(Z(q)) = 0, so that µ(ZE ) = 0. Hence, by Borel measure on A. Thus, P B2 has property (MP). Lemma 12, the integral in (8) is over the set A ∼= R\ZE and µ is a positive regular (cid:3) Remarks. 1. It is possible that there is no nontrivial positive linear functional 1 , . . . , q−1 on the algebra B in Proposition 9; for instance, this happens if ZE = R. 2. Let B be a real unital algebra of rational functions in one variable and consider the following conditions: (i) All positive linear functionals on B can be represented as integrals with respect 1 y2+1 , 1 to some positive regular Borel measure on R (or on the character set B), (ii) f ∈ B and χ(f ) ≥ 0 for all χ ∈ B implies that f ∈P B2. It seems to be of interest to characterize those algebras B for which (i) or (ii) holds. By Proposition 9 and Lemma 11 the algebra E −1R[y] satisfies (i) and (ii). Since B := R[ y2+2 ] is isomorphic to the polynomial algebra R[x1, x2], (i) does not hold for B. For B := R[x, x2+1 ] condition (i) is true, while (ii) is not fulfilled. 3. Another important open problem is the following: Characterize the finitely generated real unital algebras A for which P A2 has (MP). R[y] ⊗ Cλ. Then, by Corollary 10, P B2 Proof of Theorem 7. First we note that the algebra A is finitely generated, because D is finitely generated by assumption. Let us fix λ ∈ h(K(T )) and abbre- viate Bλ := E −1 λ has (MP) in the algebra Bλ. Therefore, Corollary 6 applies and shows that T satisfies (MP) in A. The description of A was given in Lemma 12. (cid:3) λ 1 Proof of Theorem 8. Let I denote the net of all finitely generated multiplicative subsets D of Rd[x] such that D ⊆ D0. Fix D ∈ I. Since D satisfies the assumptions (i) and (ii), Theorem 7 applies to D and L0⌈AD. Hence there exists a regular positive Borel measure µD on Rd such that L0(f ) = R f dµD for f ∈ AD. Since R 1 dµD = µF (Rd) = L0(1) for all D ∈ I, it follows that the set {µD; D ∈ I} 10 KONRAD SCHM UDGEN is relatively vaguely compact. Hence there is a subnet of the net (µD)D∈I which converges vaguely to some regular positive Borel measure µ on Rd. For notational simplicity let us assume that the net (µD)D∈I itself has this property. Now we fix F ∈ I and apply Lemma 14 to the net (µD)D⊇F , the algebra AF and the functional L = L0⌈AF to conclude that L0(f ) = R f dµ for f ∈ AF . Since each function f ∈ A0 is contained in some algebra AF with F ∈ I, (6) is satisfied. This completes the proof. (cid:3) The general fibre Theorem 3 fits nicely to the multidimensional rational moment problem, because in general the corresponding algebra A contains more bounded functions on the semi-algebraic set K(T ) than in the polynomial case. We illustrate the use of our Theorems 7 and 8 by some examples. Example 15. First let d = 2. Suppose that D contains x1 − α and the semi- algebraic set K(T ) is a subset of {(x1, x2) : x1 − α ≥ c} for some α ∈ R and c > 0. Then h1 := (x1 − α)−1 is in A and bounded on K(T ), so assumption (i) holds. Let λ ∈ h1(K(T )). Then x1 = λ−1 + α in Aλ , so Aλ consists of rational functions in x2 with denominators from some set Eλ ∈ D(R[x2]). Hence assumption (ii) is satisfied with Cλ = C. Thus Theorem 7 applies if D is finitely generated. Replacing D by D0 and T by T0 in the preceding, the assertion of Theorem 8 holds as well. The above setup extends at once to arbitrary d ∈ N, d ≥ 2, if we assume that xj−αj ∈ D and xj−αj ≥ c on K(T ) for some αj ∈ R, c > 0, and j =, . . . , d − 1. Example 16. Again we take d = 2. Let E ∈ D(R[x2]) and let p be a nonconstant polynomial from R[x1]. Suppose that D is generated by E and p2 + α for some α > 0. Obviously, h1 = (p2 + α)−1 ∈ A is bounded on each semi-algebraic set K(T ). Let λ ∈ h1(K(T )). Then we have p(x1)2 = λ−1 + α in the fibre algebra Aλ. Let C be the quotient algebra of R[x1] by the ideal generated by p(x1)2 − λ−1 − α. Then Aλ is generated by two subalgebras which are isomorphic to C and E −1R[x2], respectively. Hence assumptions (i) and (ii) are fulfilled, so P A2 has (MP) by Theorem 7 if D is finitely generated. In the general case Theorem 8 applies. In the following two examples we suppose that the sets D are finitely generated. xj = λd+jλ−1 Eλ = {1}, we have E −1 λ j Example 17. Suppose that D ∈ D(R[x]) contains the polynomials qj = 1 + x2 j = 1, . . . , d. Let T = P(AD)2. Then K(T ) = dAD = Rd by Lemma 12. Setting hj = qj(x)−1, hd+j = xj qj(x)−1 for j = 1, . . . , d, all hl ∈ AD are bounded on K(T ). Let λ ∈ h(K(T )). Then λj = qj(λ)−1 6= 0 and λd+j = λj qj(λ)−1; hence we have for j = 1, . . . , d in the fibre algebra (AD)λ. Thus, (AD)λ = R. Taking j for R[x] = R[x], so (i) and (ii) are obviously satisfied. Therefore, P(AD)2 obeys (MP) by Theorem 7. That is, each positive linear functional on the algebra AD is given by some positive measure on dAD = Rd. The same conclusion and almost the same reasoning are valid if we assume d. In this case we set instead that D contains the polynomial q = 1 + x2 hj = xj q(x)−1 for j = 1, . . . , d and hd+1 = q(x)−1. 1 + · · · + x2 Example 18. Suppose now that D ∈ D(R[x]) contains only the polynomials qj = 1 + x2 j for j = 1, . . . , d − 1. Let T =P(AD)2. Then hj := qj (x)−1 and hd+j−1 := xjqj(x)−1 for j = 1, . . . , d − 1 are in AD and bounded on K(T ) =dAD. Arguing as Then, again by Theorem 7, P(AD)2 satisfies (MP). The same is true if we in Example 17 it follows that xj = λd+j−1λ−1 (AD)λ. Therefore (AD)λ is an algebra E −1 xd for some finitely generated set Eλ ∈ D(R[xd]). for j = 1, . . . , d − 1 in the algebra R[xd] of rational functions in the variable assume instead that the polynomial q = 1 + x2 d−1 is in D ∈ D(R[x]). 1 + · · · + x2 λ j A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS11 Remark. Assumption (ii) is needed to ensure that the fibre preorderings Tλ satisfy (MP). The crucial result for this is Proposition 9 which states that the cone P (E −1R[y])2 obeys (MP) for finitely generated E ∈ D(R[y]). A similar assertion holds for several other algebras of rational functions as well; a sample is 2+1 ]. Replacing E −1R[y] by such an algebra the fibre theorem will lead R[x1, x2, to further results on the multidimensional rational moment problem. 1 x2 4. An extension theorem In this section we derive a theorem which characterizes moment functional on Rd in terms of extensions. Throughout let A denote the real algebra of functions on (Rd)× := Rd\{0} generated by the polynomial algebra Rd[x] and the functions (13) fkl(x) := xkxl(x2 1 + · · · + x2 d)−1, where k, l = 1, . . . , d, x ∈ Rd\{0}. Clearly, these functions satisfy the identity (14) dXk,l=1 fkl(x)2 = 1. That is, the functions fkl, k, l = 1, . . . , d, generate the coordinate algebra C(Sd−1) of the unit sphere Sd−1 in Rd. The next lemma describes the character set A of A. Lemma 19. The set A is parameterized by the disjoint union of Rd\{0} and Sd−1. For x ∈ Rd\{0} the character χx is the evaluation of functions at x and for t ∈ Sd−1 the character χt acts by χt(xj ) = 0 and χt(fkl) = fkl(t), where j, k, l = 1, . . . , d. Proof. It is obvious that for any x ∈ Rd\{0} the point evaluation χx at x is a character on the algebra A satisfying (χ(x1), . . . , χ(xd)) 6= 0. Conversely, let χ be a character of A such that x := (χ(x1), . . . , χ(xd)) 6= 0. Then the identity (x2 1 + · · · + x2 (χ(x1)2 + · · · + χ(xd)2)χ(fkl) = χ(xk)χ(xl) d)fkl = xkxl implies that and therefore χ(fkl) = (χ(x1)2 + · · · + χ(xd)2)−1χ(xk)χ(xl) = fkl(x). Thus χ acts on the generators xj and fkl, hence on the whole algebra A, by point evaluation at x, that is, we have χ = χx. Next let us note that the quotient of A by the ideal generated by Rd[x] is (isomorphic to) the algebra C(Sd−1). Therefore, if χ is a character of A such that (χ(x1), . . . , χ(xd)) = 0, then it gives a character on the algebra C(Sd−1). Clearly, each character of C(Sd−1) comes from a point of Sd−1. Conversely, each point t ∈ Sd−1 defines a unique character of A by χt(fkl) = fkl(t) and χt(xj ) = 0 for all k, l, j. (cid:3) Theorem 20. The preordering P A2 of the algebra A satisfies (MP), that is, for each positive linear functional L on A there exist positive Borel measures ν0 on Sd−1 and ν1 ∈ M(Rd\{0}) such that for all polynomials g we have L(g(x, f11(x), . . . , fdd(x))) (15) =ZSd−1 g(0, f11(t), . . . , fdd(t)) dν0(t) +ZRd\{0} g(x, f11(x), . . . , fdd(x)) dν1(x). 12 KONRAD SCHM UDGEN Proof. It suffices to prove that P A2 has (MP). The assertions follow then from the definition of the property (MP) and the explicit form of the character set given in Lemma 19. From the description of A it is obvious that the functions fkl, k, l = 1, . . . , d, are bounded on A, so we can take them as functions hj in Theorem 3. Let us fix a non-empty fire for λ = (λkl), where λkl ∈ R for all k, l, and consider the quotient algebra A/Iλ of A by the fibre ideal Iλ. In the algebra A/Iλ we have χ(fkl) = λkl for χ ∈ A and all k, l. Let χ = χx, where x ∈ Rd\{0}. Then χx(fkl) = fkl(x) = λkl. Since 1 = Pk fkk(x) =Pk λkk, there is a k such that λkk 6= 0. From the equality λkk = fkk(x) = x2 it follows that xk 6= 0. Thus λkl λkk 1 +· · ·+x2 , so that d)−1 k(x2 (16) = fkl(x) fkk(x) = xl kk xk xk xl = λklλ−1 for l = 1 . . . , d. If χ = χt for t ∈ Sd, then χ(xl) = χ(xk) = 0, so (16) holds trivially. That is, in the quotient algebra A/Iλ we have the relations (16) and fkl = λkl. This implies that A/Iλ is an algebra of polynomials in the single variable xk. Hence it follows from Hamburger's theorem that the preorderingP(A/Iλ)2 satisfies (MP) in A/Iλ. Therefore, by Theorem 3,(iii)→(i), T =P A2 obeys (MP) in A. The main result of this section is the following extension theorem. (cid:3) Theorem 21. A linear functional L on Rd[x] is a moment functional if and only if it has an extension to a positive linear functional L on the larger algebra A. Proof. Assume first that L has an extension to a positive linear functional L on A. By Theorem 20, the functional L on A is of the form described by equation (15). We define a positive Borel measure µ on Rd by µ({0}) = ν0(Sd−1), µ(M \{0}) = ν1(M \{0}). Let p ∈ Rd[x]. Setting g(x, 0, . . . , 0) = p(x) in the equation of Theorem 20, we get L(p) = L(p) = ν0({0})p(0) +ZRd\{0} g(x, 0, . . . , 0) dν1(x) =ZRd p(x) dµ(x). Thus L is moment functional on Rd[x] with representing measure µ. Conversely, suppose that L is a moment functional on Rd[x] and let µ be a representing measure. Since fkl(t, 0, . . . , 0) = δk1δl1 for t ∈ R, t 6= 0, we have limt→0 fkl(t, 0, . . . , 0) = δk1δl1. Hence there is a well-defined character on the alge- bra A given by and χ(p) = p(0) for p ∈ R[x]. Then, for f ∈ Rd[x], we have χ(f ) = lim t→0 f (t, 0, . . . , 0), f ∈ A, (17) L(f ) = µ({0})χ(f ) +ZRd\{0} f (x) dµ(x). For f ∈ A we define L(f ) by the right-hand side of (17). Then L is a positive (cid:3) linear functional on A which extends L. Remarks. 1. The problem of characterizing moment sequences in terms of ex- tensions have been studied in several papers such as [12], [9], and [4]. 2. Another type of extension theorems has been derived in [9]. The main differ- ence to the above theorem is that in [9], see e.g. Theorem 2.5, a function h(x) := (1 + x2 1 + · · · + x2 d + p1(x)2 + · · · + pk(x)2)−1 is added to the algebra, where p1, . . . , pk ∈ Rd[x] are fixed. Then h(x) is bounded on the character set and so are xjh and xj xkh for j, k = 1, . . . , d. The existence A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS13 assertions of the results in [9] follow also from Theorem 3. Note that in this case the representing measure for the extended functional is unique (see [9, Theorem 2.5]). 2. The measure ν1 in Theorem 20 and the representing measure µ for the func- tional L in Theorem 21 are not uniquely determined by L. (A counter-example can be easily constructed by taking an appropriate measure supported by a coordinate axis.) Let µrad denote the measure on [0, +∞) obtained by transporting µ by the mapping x → kxk2. Then, as shown in [8, p. 2964, Nr 2.], if µrad is determinate on [0, +∞), then µ is is uniquely determined by L. Thus, Theorem 21 fits nicely to the determinacy results via disintegration of measures developed in [8, Section 8]. 5. Application to the complex moment problem Given a complex 2-sequence s = (sm,n)(m,n)∈N2 the complex moment problem asks when does there exist a positive Borel measure µ on C such that the function zmzn on C is µ-integrable and 0 (18) zmzn dµ(z) for all (m, n) ∈ N2 0. smn =ZC The semigroup ∗-algebra C[N2 0 with involution (m, n) := 0, is the ∗-algebra C[z, z] with involution given by z∗ = z. If L 0] of the ∗-semigroup N2 (n, m), (m, n) ∈ N2 denotes the linear functional on C[z, z] defined by then (18) means that L(zmzn) = sm,n, (m, n) ∈ N2 0 Ls(p) =ZC p(z, z) dµ(z), p ∈ C[z, z]. Clearly, N2 0 is a subsemigroup of the larger ∗-semigroup N+ = {(m, n) ∈ Z2 : m + n ≥ 0} with involution (m, n)∗ = (n, m). The following fundamental theorem was proved by J. Stochel and F.H. Szafraniec [12]. Theorem 22. A linear functional L on C[z, z] is a moment functional if and only if L has an extension to a positive linear functional L on the ∗-algebra C[N+]. In [12] this theorem was stated in terms of semigroups: A complex sequence s = (sm,n)(m,n)∈N2 0 if and only if there exists a positive semidefinite sequence s = (sm,n)(m,n)∈N+ on the ∗-semigroup N+ such that sm,n = sm,n for all (m, n) ∈ N2 0. is a moment sequence on N2 0 In order to prove Theorem 22 we first describe the semigroup ∗-algebra C[N+]. Clearly, C[N+] is the complex ∗-algebra generated by the functions zmzn on C\{0}, where m, n ∈ Z and m + n ≥ 0. If r(z) denotes the modulus and u(z) the phase of z, then zmzn = r(z)m+nu(z)m−n. Setting k = m + n, it follows that C[N+] = Lin {r(z)ku(z)2m−k; k ∈ N0, m ∈ Z} . The functions r(z) und u(z) itself are not in C[N+], but r(z)u(z) = z and v(z) := u(z)2 = zz−1 are in C[N+] and they generate the ∗-algebra C[N+]. Writing z = x1 + ix2 with x1, x2 ∈ R, we get 1 + v(z) = 1 + x1 + ix2 x1 − ix2 = 2 x2 1 + i x1x2 1 + x2 x2 2 , 1 − v(z) = 2 x2 2 − i x1x2 1 + x2 x2 2 . 14 KONRAD SCHM UDGEN This implies that the complex algebra C[N+] is generated by the five functions (19) x1, x2, x2 1 1 + x2 x2 2 , x2 2 1 + x2 x2 2 , x1x2 1 + x2 x2 2 . Obviously, the hermitian part C[N+]h of the complex ∗-algebra C[N+] is just the real algebra generated by the functions (19). This real algebra is the special case d = 2 of the ∗-algebra A treated in Section 4. Therefore, if we identify C with R2, the assertion of Theorem 22 follows at once from Theorem 21. 6. Application to the two-sided complex moment problem The two-sided complex moment problem is the moment problem for the ∗- semigroup Z2 with involution (m, n) := (n, m). Given a sequence s = (sm,n)(m,n)∈Z2 it asks when does there exist a positive Borel measure µ on C× := C\{0} such that the function zmzn on C× is µ-integrable and smn =ZC× zmzn dµ(z) for all (m, n) ∈ Z2. Note that this requires conditions for the measure µ at infinity and at zero. The following basic result was obtained by T.M. Bisgaard [2]. Theorem 23. A linear functional L on C[Z2] is a moment functional if and only if L is a positive functional, that is, L(f ∗f ) ≥ 0 for all f ∈ C[Z2]. In terms of ∗-semigroups the main assertion of this theorem says that each pos- itive semidefinite sequence on Z2 is a moment sequence on Z2. This result is somewhat surprising, since C× has dimension 2 and no additional condition (such as strong positivity or some appropriate extension) is required. First we reformulate the semigroup ∗-algebra C[Z2]. Clearly, C[Z2] is generated by the functions z, z, z−1, z −1 on the complex plane, that is, C[Z2] is the ∗-algebra C[z, z, z−1, z−1] of complex Laurent polynomials in z and z. A vector space basis of this algebra is {zkzl; k, l ∈ Z}. Writing z = x1 + ix2 with x1, x2 ∈ R we have z−1 = x1 − ix2 1 + x2 x2 2 and z −1 = x1 + ix2 1 + x2 x2 2 . Hence C[Z2] is the complex unital ∗-algebra generated by the hermitian functions (20) x1, x2, y1 := x1 x2 1 + x2 2 , y2 := x2 x2 1 + x2 2 on R2\{0}. All four functions are unbounded on R2\{0} and we have (21) (y1 + iy2)(x1 − ix2) = 1. Proof of Theorem 23: As above we identify C and R2 in the obvious way. As discussed at the end of Section 2, the hermitian part of the complex ∗-algebra C[Z2] is a real algebra A. First we determine the character set A of A. Obviously, the point evaluation at each point x ∈ R2\{0} defines uniquely a character χx of A. From (21) it follows at once that there is no character χ on A for which χ(x1) = χ(x2) = 0. Thus, The three functions A = {χx; x ∈ R2, x 6= 0 }. h1(x) = x1y1 = x2 1 1 + x2 x2 2 , h2(x) = x2y2 = x2 2 1 + x2 x2 2 , h3(x) = 2x1y2 = x1x2 1 + x2 x2 2 are elements of A and they are bounded on A ∼= R2\{0}. The same reasoning as in the proof of Theorem 20 shows that the fibre algebra A/Iλ for each nonempty fibre A GENERAL FIBRE THEOREM FOR MOMENT PROBLEMS AND SOME APPLICATIONS15 is a polynomial algebra in a single variable. Therefore, by Hamburger's theorem, the preorderingP(A/Iλ)2 satisfies (MP) in A/Iλ and so doesP A2 in A by Theorem 3,(iii)→(i). Since A ∼= R2\{0} = C×, this gives the assertion. (cid:3) Remark. The algebra A generated by the four functions x1, x2, y1, y2 on C× is an interesting structure: The generators satisfy the relations 1 + y2 x1y1 + x2y2 = 1 and (x2 1 + x2 2)(y2 2) = 1 and there is a ∗-automorphism Φ of the real algebra A (and hence of the complex ∗-algebra C[Z2]) given by Φ(xj) = yj and Φ(yj) = xj, j = 1, 2. Acknowledgements. The author would like to thank T. Netzer for valuable discussions on the subject of this paper and M. Wojtylak for useful comments. References [1] Berg, C., Christensen, J.P.R. and P. Ressel, Harmonic Analysis on Semigroups, Graduate Texts in Math., Springer-Verlag, Berlin, 1984. [2] Bisgaard, T.M., The two-sided complex moment problem, Ark. Mat. 27(1989), 23 -- 28. [3] Choquet, G., Lectures on Analysis, vol. III, Benjamin, Reading, 1969. [4] Cichon, C., Stochel, J. and Szafraniec, F.H., Extending positive definiteness. Trans. Amer. Math. 363(2010), 545 -- 577. [5] Cimpric, J., Marshall, M. and Netzer,T., On the real multidimensional rational K-moment problem, Trans. Amer. Math. Soc 363(2011), 5773 -- 5788. [6] Marshall, M., Positive polynomials and sums of squares, Math. Surveys and Monographs 146, Amer. Math. Soc., 2008. [7] Netzer, T., An elementary proof of Schmudgen's theorem on the moment problem of closed semi-algebraic sets, Proc. Amer. Math. Soc. 136(2008), 529 -- 537. [8] Putinar, M. and Schmudgen, K., Multivariate determinateness, Indiana Univ. Math. J. 57(2008), 2931 -- 2968. [9] Putinar, M. and Vasilescu, F.-H., Solving the moment problem by dimension extension, Ann. Math. 149(1999), 1087 -- 1069. [10] Scheiderer, C., Non-existence of degree bounds for weighted sums of squares representations, J. Complexity 2(2005), 823 -- 844. [11] Schmudgen, K., On the moment problem of closed semi-algebraic sets, J. Reine Angew. Math. 558(2003), 225 -- 234. [12] Stochel, J. and Szafraniec, F.H., The complex moment problem and subnormality: a polar decomposition approach. J. Funct. Anal. 159(1998), 432 -- 491. Universitat Leipzig, Mathematisches Institut, Augustusplatz 10/11, D-04109 Leipzig, Germany E-mail address: [email protected]