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
1212.1604
1
1212
2012-12-07T13:17:37
On some inequalities for s-logarithmically convex functions in the second sense via fractional integrals
[ "math.FA", "math.CA" ]
In this paper, we establish some new Hadamard type inequalities for s-logarithmically convex functions in the second sense via fractional integrals by using Lemma 1 which has been proved by Sarikaya et al. in the paper [3].
math.FA
math
ON SOME INEQUALITIES FOR s-LOGARITHMICALLY CONVEX FUNCTIONS IN THE SECOND SENSE VIA FRACTIONAL INTEGRALS HAVVA KAVURMACI AND MEVL UT TUNC¸ Abstract. In this paper, we establish some new Hadamard type inequalities for s-logarithmically convex functions in the second sense via fractional inte- grals by using Lemma 1 which has been proved by Sarıkaya et al. in the paper [3]. 1. Introduction The following result is well known in the literature as Hadamard's inequality [1]. Theorem 1. Let f : I ⊂ R → R be a convex function on the interval I of real numbers and a, b ∈ I with a < b. Then (1.1) f(cid:18) a + b 2 (cid:19) ≤ 1 b − aZ b a f (x) dx ≤ f (a) + f (b) 2 The following definitions is well known in the literature: Definition 1. A function f : I → R, ∅ 6= I ⊆ R, where I is a convex set, is said to be convex on I if inequality f (tx + (1 − t) y) ≤ tf (x) + (1 − t) f (y) holds for all x, y ∈ I and t ∈ [0, 1]. In [2], Akdemir and Tun¸c were introduced the class of s-logarithmically convex functions in the first and second sense as the following: Definition 2. A function f : I ⊂ R0 → R+ is said to be s-logarithmically convex in the first sense if (1.2) for some s ∈ (0, 1], where x, y ∈ I and αs + βs = 1. f (αx + βy) ≤ [f (x)]αs [f (y)]βs Definition 3. A function f : I ⊂ R0 → R+ is said to be s-logarithmically convex in the second sense if (1.3) f (tx + (1 − t) y) ≤ [f (x)]ts for some s ∈ (0, 1], where x, y ∈ I and t ∈ [0, 1]. [f (y)](1−t)s 2000 Mathematics Subject Classification. 26D10, 26A15, 26A16, 26A51. Key words and phrases. Hadamard's inequality, s-geometrically convex functions . 1 2 HAVVA KAVURMACI AND MEVL UT TUNC¸ Clearly, when taking s = 1 in Definition 2 or Definition 3, then f becomes the standard logarithmically convex function on I. Definition 4. Let f ∈ L1[a, b]. The Riemann-Liouville integrals J α order α > 0 with a ≥ 0 are defined by a+ f and J α b−f of and J α a+ f (x) = 1 Γ(α) J α b−f (x) = 1 Γ(α) ∞ x Za b Zx (x − t)α−1 f (t)dt, x > a (t − x)α−1 f (t)dt, x < b respectively where Γ(α) = e−uuα−1du. Here is J 0 a+ f (x) = J 0 b−f (x) = f (x). In the case of α = 1, the fractional integral reduces to the classical integral. For some recent results connected with fractional integral inequalities see [3]- [11]. In [3], Sarıkaya et. al. proved the following results for fractional integrals. Lemma 1. Let f : [a, b] → R be a differentiable mapping on (a, b) with a < b. If f ′ ∈ L [a, b] , then the following equality for fractional integrals holds: R0 f (a) + f (b) − Γ (α + 1) 2 (b − a)α [J α b− f (a) + J α a+ f (b)] [(1 − t)α − tα] f ′ (ta + (1 − t) b) dt. = b − a 2 2 Z 1 0 Theorem 2. Let f : [a, b] → R be a differentiable mapping on (a, b) with a < b. If f ′ is convex on [a, b] , then the following inequality for fractional integrals holds: − f (a) + f (b) 2 b − a (cid:12)(cid:12)(cid:12)(cid:12) 2 (α + 1)(cid:18)1 − b− f (a) + J α Γ (α + 1) 2 (b − a)α [J α 2α(cid:19) [f ′ (a) + f ′ (b)] . 1 a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) = In the present paper, we will establish several Hermite-Hadamard type inequali- ties for the class of functions whose derivatives in absolute value are s-logarithmically convex functions in the first and second sense via Riemann-Liouville fractional in- tegral. 2. Hadamard type inequalities for s-logarithmically convex functions Theorem 3. Let I ⊃ [0, ∞) be an open interval and f : I → (0, ∞) is differentiable. If f ′ ∈ L [a, b] and f ′ is s-logarithmically convex functions in the second sense on ON SOME INEQUALITIES 3 [a, b] for some fixed s ∈ (0, 1] and µ, η > 0 with µ + η = 1, then the following inequality for fractional integrals with α > 0 holds: f (a) + f (b) 2 − Γ (α + 1) 2 (b − a)α [J α b− f (a) + J α (2.1) ≤ (cid:12)(cid:12)(cid:12)(cid:12) b − a 2 (Z 1/2 0 a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) 1 µ dt +Z 1 1/2 µ [(1 − t)α − tα] µ [tα − (1 − t)α] 1 µ dt + η × f ′ (b) s η ψ(cid:18) s η , s η(cid:19)(cid:27) where (2.2) Ψ (ψ) =(cid:26) 1, ψ−1 ln ψ , ψ = 1, 0 < ψ < 1 and ψ (u, v) = f ′ (a)u f ′ (b)−v , u, v > 0. Proof. By Lemma 1 and since f ′ is s-logarithmically convex functions in the second sense on [a, b] , we have ≤ ≤ ≤ (2.3) (cid:12)(cid:12)(cid:12)(cid:12) 0 0 2 2 b − a b − a Z 1 Z 1 2 (Z 1/2 +Z 1 b − a 1/2 0 f (a) + f (b) 2 − Γ (α + 1) 2 (b − a)α [J α b− f (a) + J α a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) (1 − t)α − tα f ′ (ta + (1 − t) b) dt (1 − t)α − tα f ′ (a)ts f ′ (b)(1−t)s dt [(1 − t)α − tα] f ′ (a)ts f ′ (b)(1−t)s dt [tα − (1 − t)α] f ′ (a)ts f ′ (b)(1−t)s dt) , for all t ∈ [0, 1] . Using the well known inequality mn ≤ µm side of (2.3), we have 1 µ + ηn 1 η , on the right ≤ = 0 2 b − a f (a) + f (b) (cid:12)(cid:12)(cid:12)(cid:12) 2 (Z 1/2 +Z 1 2 (Z 1/2 + ηZ 1 b − a 1/2 0 0 − Γ (α + 1) 2 (b − a)α [J α b−f (a) + J α µ [(1 − t)α − tα] 1 µ dt +Z 1/2 0 a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) η f ′ (a) ts η f ′ (b) (1−t)s η dt µ [tα − (1 − t)α] 1 µ dt +Z 1 1/2 η f ′ (a) ts η f ′ (b) (1−t)s η dt) µ [(1 − t)α − tα] f ′ (a) ts η f ′ (b) (1−t)s η µ [tα − (1 − t)α] 1 µ dt 1/2 1 µ dt +Z 1 dt(cid:27) If 0 < λ ≤ 1, 0 < u, v ≤ 1, then (2.4) λuv ≤ λuv. 4 HAVVA KAVURMACI AND MEVL UT TUNC¸ When ψ (u, v) ≤ 1, by (2.4), we get that (2.5) f ′ (a) ts η f ′ (b) (1−t)s η dt ≤Z 1 0 f ′ (a) st η f ′ (b) s(1−t) η dt = f ′ (b) s η ψ(cid:18) s η , s η(cid:19) . (cid:3) From (2.3) to (2.5), (2.1) holds. Remark 1. If we take α = 1, in Theorem 3, then the inequality (2.1) become the inequality f (a) + f (b) 2 − 1 b − aZ b a ≤ b − a 2 (cid:20) µ2 µ + 1 + η × f ′ (b) s η ψ(cid:18) s η , s η(cid:19)(cid:21) . The corresponding version for powers of the absolute value of the first derivative f (x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) is incorporated in the following result: Z 1 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) Theorem 4. Let I ⊃ [0, ∞) be an open interval and f : I → (0, ∞) is differentiable. If f ′ ∈ L [a, b] and f ′ is s-logarithmically convex functions in the second sense on [a, b] for some fixed s ∈ (0, 1] and µ, η > 0 with µ + η = 1 and p, q > 1, then the following inequality for fractional integrals with α > 0 holds: (2.6) ≤ f (a) + f (b) 2 b − a 2 (αp + 1) 1 p − Γ (α + 1) 2 (b − a)α [J α f ′ (b)s (ψ (sq, sq)) 1 q b− f (a) + J α a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) where 1/p + 1/q = 1, and ψ (u, v) is defined as in (2.2). Proof. By Lemma 1 and since f ′ is s-logarithmically convex functions in the second sense on [a, b] , we have (2.7) ≤ b − a 2 Z 1 0 f (a) + f (b) 2 − Γ (α + 1) 2 (b − a)α [J α b− f (a) + J α (1 − t)α − tα f ′ (a)ts f ′ (b)(1−t)s a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) dt for all t ∈ [0, 1] . Using the well known Holder inequality, on the right side of (2.7) and making the change of variable we have (2.8) ≤ (cid:12)(cid:12)(cid:12)(cid:12) f (a) + f (b) 2 b − a 2 (cid:18)Z 1 0 − Γ (α + 1) 2 (b − a)α [J α (1 − t)α − tαp dt(cid:19) 1 p (cid:18)Z 1 0 b− f (a) + J α a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) f ′ (a)qts It is know that for α, t1, t2 ∈ [0, 1] , f ′ (b)q(1−t)s 1 q . dt(cid:19) tα 1 − tα 2 ≤ t1 − t2α , therefore (2.9) Z 1 0 (1 − t)α − tαp dt ≤Z 1 0 1 − 2tαp dt = 1 αp + 1 . ON SOME INEQUALITIES 5 Since f ′ is s-logarithmically convex functions on [a, b] and ψ (u, v) ≤ 1, we obtain (2.10) Z 1 0 f ′ (a)qts f ′ (b)q(1−t)s dt ≤ f ′ (b)sq ψ (sq, sq) From (2.8) to (2.10), (2.6) holds. (cid:3) A different approach leads to the following result. Theorem 5. Let I ⊃ [0, ∞) be an open interval and f : I → (0, ∞) is differentiable. If f ′ ∈ L [a, b] and f ′q is s-logarithmically convex functions in the second sense on [a, b] for some fixed s ∈ (0, 1] and µ, η > 0 with µ + η = 1 and q ≥ 1, then the following inequality for fractional integrals with α > 0 holds: (2.11) (cid:12)(cid:12)(cid:12)(cid:12) ≤ f (a) + f (b) 2 b − a q−(1−α)(q−1) q 2 b−f (a) + J α − Γ (α + 1) 2 (b − a)α [J α α + 1 (cid:19)1− 1 (cid:18) 2α − 1 q (cid:18) µ2 α + µ a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) sq + η f ′ (b) η ψ(cid:18) sq η 1 q , sq η (cid:19)(cid:19) where ψ (u, v) is defined as in (2.2). Proof. By Lemma 1 and using the well known power mean inequality, we have (cid:12)(cid:12)(cid:12)(cid:12) ≤ ≤ b − a 0 2 Z 1 2 (cid:18)Z 1 0 b − a f (a) + f (b) 2 − Γ (α + 1) 2 (b − a)α [J α b− f (a) + J α (1 − t)α − tα f ′ (ta + (1 − t) b) dt (1 − t)α − tα dt(cid:19) 1− 1 q (cid:18)Z 1 0 a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) (1 − t)α − tα f ′ (ta + (1 − t) b)q dt(cid:19) 1 q It is easily check that Z 1 0 (1 − t)α − tα dt = 2 α + 1(cid:18)1 − 1 2α(cid:19) . Since f ′q is s-logarithmically convex and using the well known inequality mn ≤ µm η , we obtain µ + ηn 1 1 Z 1 0 (1 − t)α − tα f ′ (ta + (1 − t) b)q dt ≤ Z 1 ≤ Z 1 ≤ µZ 1 0 0 0 It is easily check that (1 − t)α − tα f ′ (a)qts f ′ (b)q(1−t)s dt 1 − 2tα f ′ (a)qts f ′ (b)q(1−t)s dt 1 − 2t α µ dt + ηZ 1 0 f ′ (a) qts η f ′ (b) q(1−t)s η dt. µZ 1 0 1 − 2t α µ dt = µ 1 α µ + 1 = µ2 α + µ . 6 HAVVA KAVURMACI AND MEVL UT TUNC¸ Afterwards, when ψ (u, v) ≤ 1, by (2.4), we get that (2.12) f ′ (a) qts η f ′ (b) q(1−t)s Z 1 0 Therefore f (a) + f (b) 2 − (cid:12)(cid:12)(cid:12)(cid:12) ≤ ≤ b − a 2 (cid:18) 2 2 (cid:18) 2 α + 1(cid:18)1 − α + 1(cid:18)1 − b − a which comletes the proof. η 0 dt ≤Z 1 Γ (α + 1) 2 (b − a)α [J α 2α(cid:19)(cid:19)1− 1 2α(cid:19)(cid:19)1− 1 1 1 f ′ (a) sqt η f ′ (b) sq(1−t) η dt = f ′ (b) sq η ψ(cid:18) sq η , sq η (cid:19) . b− f (a) + J α 1 − 2t α a+ f (b)](cid:12)(cid:12)(cid:12)(cid:12) µ dt + ηZ 1 η ψ(cid:18) sq η sq 0 + η f ′ (b) 0 q (cid:18)µZ 1 q (cid:18) µ2 α + µ 1 q dt(cid:19) f ′ (a) qts η f ′ (b) q(1−t)s η 1 q , sq η (cid:19)(cid:19) (cid:3) References [1] J. Hadamard: ´Etude sur les propri´et´es des fonctions enti`eres et en particulier d'une fonction consider´ee par Riemann, J. Math Pures Appl., 58, (1893) 171 -- 215. [2] A. O. Akdemir and M. Tun¸c: On some integral inequalities for s-logarithmically convex functions, submitted. [3] Sarıkaya, M. Z., Set, E., Yaldiz, H. and Ba¸sak, N.: Hermite-Hadamard's inequalities for frac- tional integrals and related fractional inequalities, Mathematical and Computer Modelling, In Press. [4] Belarbi, S. and Dahmani, Z.: On some new fractional integral inequalities, J. Ineq. Pure and Appl. Math., 10(3), Art. 86 (2009). [5] Dahmani, Z.: New inequalities in fractional integrals, International Journal of Nonlinear Science, 9(4), 493-497 (2010). [6] Dahmani, Z.: On Minkowski and Hermite-Hadamard integral inequalities via fractional inte- gration, Ann. Funct. Anal. 1(1), 51-58 (2010). [7] Dahmani, Z., Tabharit, L. and Taf, S.: Some fractional integral inequalities, Nonl. Sci. Lett. A., 1(2), 155-160 (2010). [8] Dahmani, Z., Tabharit, L. and Taf, S.: New generalizations of Gruss inequality using Riemann-Liouville fractional integrals, Bull. Math. Anal. Appl., 2(3), 93-99 (2010). [9] Ozdemir, M.E., Kavurmacı, H. and Avcı, M.: New inequalities of Ostrowski type for map- pings whose derivatives are (α, m)-convex via fractional integrals, RGMIA Research Report Collection, 15, Article 10, 8 pp (2012). [10] Ozdemir, M.E., Kavurmacı, H. and Yıldız, C¸ .: Fractional integral inequalities via s-convex functions, arXiv:1201.4915v1 [math.CA] 24 Jan 2012. [11] Ozdemir, M.E., Dragomir, S. S. and Yıldız, C¸ .: The Hadamard inequality for convex function via fractional integrals, RGMIA Volume 15, art.14, 2012. Agrı Ibrahim C¸ ec¸en University, Faculty of Science and Arts, Department of Math- ematics, Agrı, Turkey. E-mail address: [email protected] Kilis 7 Aralık University, Faculty of Science and Arts, Department of Mathematics, Kilis, 79000, Turkey. E-mail address: [email protected]
1808.06225
4
1808
2019-11-14T13:20:46
Inversion problem in measure and Fourier-Stieltjes algebras
[ "math.FA" ]
In this paper we study the inversion problem in measure and Fourier-Stieltjes algebras from qualitative and quantitative point of view extending the results obtained by N. Nikolski.
math.FA
math
INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI Abstract. In this paper we study the inversion problem in measure and Fourier -- Stieltjes algebras from qualitative and quantitative point of view extending the results obtained by N. Nikolski in [N]. 1. Introduction We are going to collect first some basic facts from Banach algebra theory and harmonic analysis in order to fix the notation (our main reference for Banach algebra theory is [Ż], for harmonic analysis check [R]). For a commutative unital Banach algebra A, the Banach algebra of all complex-valued Borel regular measures equipped with the convolution Gelfand space of A (the set of all multiplicative-linear functionals endowed with weak∗- topology) will be denoted by △(A) and the Gelfand transform of an element x ∈ A is a surjection bx : △(A) → σ(x) defined by the formula: bx(ϕ) = ϕ(x) for ϕ ∈ △(A), where σ(x) := {λ ∈ C : µ − λδ0 is not invertible} is the spectrum of an element x. Let G be a locally compact Abelian group with its unitary dual bG and let M(G) denote the product and the total variation norm. It is also a ∗-algebra with involution µ → eµ defined for any Borel set E ⊂ G by eµ(E) := µ(−E). A measure µ ∈ M(G) is called hermitian, if µ = eµ or, equivalently, its Fourier-Stieltjes transform is real-valued. The Fourier-Stieltjes transform will be treated as a restriction of the Gelfand transform to bG. Note that we have a direct sum decomposition M(G) = Mc(G) ⊕ Md(G) where Mc(G) is the ideal Date: November 15, 2019. 2010 Mathematics Subject Classification. Primary 43A05; Secondary 43A30. Key words and phrases. Inverse, Measure Algebras, Fourier -- Stieltjes algebras. PO was supported by foundations managed by The Royal Swedish Academy of Sciences. MW was partially supported by National Science Centre (NCN) grant 2016/21/N/ST1/02499, long term structural funding -- Methusalem grant of the Flemish Government -- and by European Research Council Consolidator Grant 614195 RIGIDITY.. 1 2 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI of continuous (non-atomic) measures and Md(G) is the subalgebra of discrete (atomic) measures. For µ ∈ M(G) we will write µ = µc + µd with µc ∈ Mc(G) and µd ∈ Md(G). We recall the problem investigated by N. Nikolski in [N], which will be our main point of interest. Problem 1. Let µ ∈ M(G) satisfy kµk ≤ 1 and suppose that inf γ∈bG bµ(γ) = δ > 0. What is the minimal value of δ0 > 0 such that for any δ > δ0 the measure µ is automatically invertible? What can be said about the norm of the inverse? It is clear that δ0 < 1 2 ν where ν is a probability measure with non-negative Fourier-Stieltjes transform satisfying σ(ν) = D 2 is not enough as the example of a measure µ := 1 2 δ0 + 1 shows (here δ0 is the Dirac delta at the point 0). On the other hand, it was proved in [N] that any δ > 1√2 does the job and moreover the norm of the inverse is bounded by (2δ2 − 1)−1. For the readers convenience we will reprove this result now with a slightly simpler (but less general) approach. Let µ ∈ M(G) satisfy kµk ≤ 1 and inf γ∈bG bµ(γ) = δ > 1√2. Then, by the generalization of Wiener lemma (see 5.6.9 in [R]): (µ ∗ eµ)({0}) = Xx∈Gµ({x})2 = lim α ZΓcfα(γ)bµ(γ)2dγ ≥ δ2 > 1 2 , where {fα} is a system of continuous positive-definite functions with compact supports subordinated to a neighborhood base {Vα} of 0 with fα(0) = 1. The second ingredient is a version of Lemma 1.4.3 from [N] adapted to our situation. Lemma 1.1. Let µ ∈ M(G) and let µ = λδ0 + ν where ν({0}) = 0 and 1 kµk ≤ 1. Then µ is invertible and kµ−1k ≤ 1 By the lemma, k(µ ∗ eµ)−1k ≤ 1 observe that µ−1 = (µ ∗ eµ)−1 ∗ eµ. 2δ−1 . 2δ2−1 and in order to finish the argument one has only to However, the minimal value of δ0 > 0 seems to be unknown and the question on improv- 2 < δ ≤ λ ≤ ing the aforementioned bounds is stated in [N]. The first aim of this paper is to give a proof of the following fact (Theorem 2.4): if 2 then µ is invertible, which is the final µ ∈ M(G) satisfies kµk ≤ 1 and inf γ∈bG bµ(γ) > 1 INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 3 solution of the problem (proving δ0 = 1 sharp. 2) -- in view of the previous discussion the result is In the next part of this article we attack the problem of estimating the norm of the inverse. In addition to the upper bounds mentioned above, the paper [N] contains a result on the lower bound. We will not reproduce the whole discussion here but what is crucial for us is that one cannot hope for a bound better than (2δ − 1)−1 and the norm-controlled inversion cannot hold for any δ ≤ 1 2 for any infinite locally compact Abelian group. The most important result is Theorem 2.8 which states that under additional group-theoretic assumption the norm-controlled inversion holds true for δ > −1+√33 ≃ 0, 593 - this improves the result of N. Nikolski (δ > 1√2 ≃ 0, 707). Later, we show in Theorem 2.14 that for measures with discrete parts supported on independent subsets of locally compact Abelian 8 groups consisting of elements of infinite order we obtain the optimal bound. The section is concluded with a discussion of another special case -- real measures on groups of exponent two. The last, independent part of the paper is devoted to the study of analogous problems for Fourier-Stieltjes algebras. Particularly, we prove in Theorem 3.2 that the qualitative 2. We also prove (Theorem 3.7) the inversion problem has a positive solution for any δ > 1 norm-controlled inversion for δ > 1√2 in this context. 2. Measure algebras 2.1. Qualitative result. is invertible. Lemma 2.1. Let µ ∈ M(G) satisfy kµk ≤ 1 and suppose that inf γ∈Γ cµd(γ) > 1 Proof. Recalling that σ(µd) = σMd(G)(µd) and △(Md(G)) = △(l1(G)) = bbG (the Bohr compactfication of bG) we get 2. Then µ (2.1) ∀ϕ∈△(M (G)) cµd(ϕ) > 1 2 . 4 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI Of course, 1 2 < r(µd) ≤ kµdk so (2.2) kµck = kµk − kµdk < 1 2 . Suppose now, towards the contradiction, that µ is not invertible. Then 0 ∈ σ(µ) and there exists ϕ0 ∈ △(M(G)) for which ϕ0(µ) = 0. This gives ϕ0(µd) = −ϕ0(µc) implying that ϕ0(µd) = ϕ0(µc). But by (2.1) we have ϕ0(µd) > 1 2 and by (2.2) we get ϕ0(µc) ≤ r(µc) ≤ kµck < 1 2 which is a contradiction. (cid:3) Here we need to cite the theorem of I. Glicksberg and I. Wik (see (2) in [GW]). A straightforward corollary of this theorem is the following observation. Theorem 2.2. Let µ ∈ M(G). Then cµd(bG) ⊂ bµ(bG). Corollary 2.3. Let µ ∈ M(G). Then inf γ∈bG cµd(γ) ≥ inf γ∈bG bµ(γ). Theorem 2.4. Let µ ∈ M(G) satisfy kµk ≤ 1 and inf γ∈bG bµ(γ) > 1 Proof. By Corollary 2.3 we have inf γ∈bG cµd(γ) ≥ inf γ∈bG bµ(γ) > 1 We are prepared now to ultimately solve the inversion problem. just need to apply Lemma 2.1. 2. Then µ is invertible. 2. To finish the proof we (cid:3) 2.2. Quantitative results. 2.2.1. The general case. In this section we will attack the problem of improving the con- stant δ0 giving the following assertion: if µ ∈ M(G) satisfy kµk ≤ 1 and inf γ∈bG bµ(γ) ≥ δ > δ0 then kµ−1k can be bounded as a function of δ (norm-controlled inversion). As it was explained in the introduction, if G is non-discrete, then δ0 ≥ 1 2. In addition, a non-trivial argument from [N] shows that the same statement holds true for arbitrary infinite discrete groups. Let us also recall that δ0 ≤ 1√2 We start with an elementary lemma. . Lemma 2.5. Let (xn)∞n=1 be a non-increasing sequence of non-negative numbers satisfying (2.3) ∞Xn=1 xn ≤ 1, INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 5 (2.4) x2 n ≥ δ2, where 1 ≥ δ > 1 2 . ∞Xn=1 Then the following inequalities hold true (2.5) (2.6) x1 ≥ δ2, x1 + x2 ≥ δ. Proof. In order to prove (2.5) we use consecutively (2.4), monotonicity of the sequence (xn)∞n=1 and then (2.3) δ2 ≤ x2 n ≤ x1 xn ≤ x1. ∞Xn=1 ∞Xn=1 For the second inequality we first observe that by (2.4) and the monotonicity: δ2 ≤ ∞Xn=1 x2 n = x2 1 + Now, we apply (2.3) which gives ∞Xn=2 x2 n ≤ x2 1 + x2 ∞Xn=2 xn. (2.7) δ2 ≤ x2 1 + x2(1 − x1). If x1 ≥ δ then inequality (2.6) is obvious, so we are allowed to assume that x1 < δ and then (2.7) can be rewritten as (2.8) Thus, it is enough to justify (2.9) This is equivalent to x2 ≥ 1 δ2 − x2 1 − x1 . x1 + 1 δ2 − x2 1 − x1 ≥ δ for x1 ∈ [δ2, δ]. 2x2 1 − (1 + δ)x1 + δ − δ2 ≤ 0 for x1 ∈ [δ2, δ]. This is a quadratic inequality with the discriminant (3δ−1)2 > 1 lead to x1 ∈ [ 1−δ the last inequality simply follows from the assumption δ > 1 2 , δ] and therefore the assertion will be proved as long as 1−δ 4. Elementary calculations 2 ≤ δ2. However, 2 establishing (2.6) and finishing (cid:3) the proof of the lemma. This lemma can be applied to obtain the following proposition. 6 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI µd = 2 for every γ ∈ bG. If anδτn with a1 ≥ a2 ≥ . . . and for some τn ∈ G, Proposition 2.6. Let G be a locally compact Abelian group and let µ ∈ M(G) satisfy kµk ≤ 1 and bµ(γ) ≥ δ > 1 ∞Xn=1 then a1 ≥ δ2 and a1 + a2 ≥ δ. Proof. Clearly, kµdk ≤ 1 which is equivalent to the assumption (2.3) from Lemma 2.5. We can treat µd as an element of l2(Gd) (Gd is the original group G equipped with the It is well-known that (Gd) is canonically discrete topology). Then cµd ∈ L2(cid:16)(Gd)(cid:17). isomorphic to bbG -- the Bohr compactification of bG. By Theorem 2.2 we have cµd(γ) ≥ δ for every γ ∈ bG and since bG is dense in bbG we obtain by Parseval's identity: ∞Xn=1an2 =ZbbG cµd2(x)dx ≥ δ2 (here dx is the normalized Haar measure on bbG), which proves (2.4) from Lemma 2.5 and enables us to use the assertion of the lemma to finish the argument. (cid:3) In order to proceed further, we need yet another technical lemma. Lemma 2.7. Let G be a locally compact Abelian group and for x ∈ G define Sx = {γ(x) : γ ∈ bG} ⊂ {z ∈ C : z = 1}. If x is of order n < ∞ then Sx = Zn and if x is of infinite order then Sx is a dense subgroup of {z ∈ C : z = 1}. Proof. In case of element x ∈ G of finite order n we simply observe that the closed subgroup generated by x is isomorphic to Zn and as Znb = Zn the result follows by the well-known fact that characters on the closed subgroup can be extended to the whole group. If x is of infinite order we formally verify first that Sx is a subgroup of the circle group and since every infinite subgroup of the circle group is dense it is enough to show #Sx = ∞. Let H be a closed subgroup of G generated by x. By definition, H is a monothetic group1 and by the results from Section 2.3 in [R] we obtain that either H = Z or H is a compact 1A locally compact Abelian group is called monothetic if it contains a dense homomorphic image of Z. INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 7 group whose dual is a subgroup of Td (the circle group with the discrete topology). The first case is elementary and for the second one we continue as follows: since H is infinite, cH is also infinite. Now, if cH contains an element of infinite order then we are done, as every character in cH is uniquely defined by its action on x. Otherwise, cH contains elements of arbitrarily high order and again since x topologically generates H we obtain that for a fixed γ0 ∈ cH the set {γk 0 (x) : k ∈ N} ⊂ Sx has the same number of elements as the order of γ0, which finishes the proof. (cid:3) We are ready now to prove one of the two main theorems of this section. Note that the group-theoretic assumption from the next theorem is automatically satisfied for some classical groups such as R and Z. Theorem 2.8. Let µ ∈ M(G) satisfy kµk ≤ 1 and bµ(γ) > δ > 1 If the order of element τ2 − τ1 is infinite then anδτn, a1 ≥ a2 ≥ . . . . ∞Xn=1 µd = 2 for every γ ∈ bG. Let 1 − δ + √17δ2 + 6δ − 7 a1 ≥ 4 1 3 2 ≥ 1 2 and δ − 2 3 , for δ > kµ−1k ≤ 3δ − 2 −(1 + δ) + √17δ2 + 6δ − 7 2 kµ−1k ≤ for δ > −1 + √33 8 ≃ 0, 593. Proof. Since both multiplication by a constant of modulus one and convolution with a Dirac delta do not violate the assumptions of the theorem and do not change the norm of the inverse, we are allowed to work with a measure ν := cµ ∗ δ−τ1 where ca1 = a1 instead of the original measure µ. Then νd = a1δ0 + ca2δτ2−τ1 + ρ, where ρ := c anδτn−τ1. ∞Xn=3 As the order of τ2 − τ1 is infinite, basing on Lemma 2.7, we are allowed to pick a sequence (γn)∞n=1 ⊂ bG such that (2.10) bδτ2−τ1(γn) · ca2 = γn(τ2 − τ1) · ca2 −−−→n→∞ −a2. 8 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI By our assumptions and Theorem 2.2 we have (2.11) a1 + ca2bδτ2−τ1(γn) + bρ(γn) ≥ δ. Using elementary inequalities, kρk = kµdk − a1 − a2 and passing with the parameter n to infinity we obtain, by (2.10), (2.12) kµdk − 2a2 ≥ δ, which is equivalent to a2 ≤ kµdk − δ 2 ≤ 1 − δ 2 . By Proposition 2.6 we get a1+a2 ≥ δ. Combining this with (2.12) we obtain a1 ≥ 3 and by Lemma 1.1 we get kµ−1k ≤ 1 In order to get a more refined bound we recall that by (2.8) (from the proof of Lemma 2.5) 3δ−2 for δ > 2 3. 2 δ− 1 2 and (2.12): (2.13) or equivalently a12 − It is a quadratic inequality with discriminant D := 17 δ2 − a12 1 − a1 1 − δ 2 ≥ 1 − δ 2 a1 + 2 δ − 7 4 δ2 + 3 be positive as δ > 1 2. With the aid of Viète formulas (note that 1−δ 1 − δ 2 − δ2 ≥ 0. 4, which is easily shown to 2 − δ2 < 0), we finally get a1 ≥ The elementary verification of the inequality 2 iff δ > −1+√33 1 8 finishes the argument by Lemma 1.1. 2 + √D 1−δ 2 2 +√D 1−δ . 2 ≥ 3 2 δ− 1 2 and the condition 1−δ 2 +√D 2 > (cid:3) It is worth to state the full formulation for the most classical case of G = Z. Theorem 2.9. Let f ∈ A(T) satisfy kfk ≤ 1 and f(t) ≥ δ > 1 2 for every t ∈ T. Then 1 for δ > 2 3 , (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 f(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ 3δ − 2 −(1 + δ) + √17δ2 + 6δ − 7 2 for δ > −1 + √33 8 ≃ 0, 593. (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 1 f(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ At last, we deal with the case of elements of finite order. Theorem 2.10. Let µ ∈ M(G) satisfy kµk ≤ 1 and bµ(γ) > δ for every γ ∈ bG. Let µd = anδτn, a1 ≥ a2 ≥ . . . . ∞Xn=1 INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 9 If the order of element τ2 − τ1 is equal to n ∈ N then a1 ≥ δ − kµ−1k ≤ 2f(δ) − 1 1 − δ 2(1 − sin π 2n) 1 for δ > := f(δ) and 2n 2 − sin π 3 − sin π 2n . Proof. We perform the same reductions as in the proof of Theorem 2.8 and arrive to the following point (check (2.9)): a1 + ca2γ(τ2 − τ1) + bρ(γ) ≥ δ for every γ ∈ bG. the set of all vertices of a regular n-gon inscribed in the unit circle with one vertex at the As the order of τ2 − τ1 is equal to n we know, by Lemma 2.7, that {γ(τ2 − τ1) : γ ∈ bG} is point 1. Using elementary geometry2 we find γ0 ∈ bG such that csgn(a2)γ0(τ2 − τ1) + 1 ≤ 2 sin π 2n . This gives a1 + ca2γ0(τ2 − τ1) − a1 − a2 ≤ a1 + ca2γ(τ2 − τ1) − (a1 − a2) ≤ 2a2 sin 1−δ Repeating the estimates from the proof of Theorem 2.8 we easily get a2 ≤ a1 ≥ δ − 1−δ 2(1−sin π 2n ) and finish the argument using Lemma 1.1. 2(1−sin π π 2n 2n ), (cid:3) Remark 2.11. It is possible to obtain a more refined bound in Theorem 2.10 applying the same method as in the proof of Theorem 2.8. 2.2.2. Independent case. We pass now to the analysis of a restricted class of measures (in fact, the discrete part is the most important) in order to get the estimates for the inverse. The first (trivial) lemma will help us to deal with measures with the discrete part consisting of a finite sum of Dirac deltas at points constituting an independent set. Lemma 2.12. Let x1, x2, . . . , xN , N ∈ N be a finite sequence of non-negative numbers satisfying x1 ≥ x2 ≥ . . . ≥ xN and PN 2. Assume also that the n=1 xn ≤ 1 and let δ > 1 2We estimate the distance of any point of the unit circle from the nearest vertex of the regular n-gon inscribed in it by the distance of a midpoint of an arc joining two neighboring vertices. Then (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (−1)n−1xn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≥ δ and (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXn=1 NXn=2 x1 ≥ δ + NXn=2 xn(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≥ δ x1 − xn. 10 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI following inequalities hold true: Proof. As the sequence (xn)N x1 ≥ PN n=1(−1)n−1xn ≥ δ. Since PN n=1 is non-increasing we have PN n=1(−1)n−1xn ≥ 0. Of course, n=1 xN ≤ 1 the second inequality proves the assertion. (cid:3) We also need a generalization of the classical Kronecker's approximation theorem (see 5.1.3 in [R]). By definition, a subset E ⊂ G is independent if for every choice of distinct points x1, . . . , xk of E and integers n1, . . . , nk either or n1x1 = n2x2 = . . . = nkxk = 0 n1x1 + n2x2 + . . . + nkxk 6= 0. Theorem 2.13. Suppose E is a finite independent subset of a locally compact Abelian group G consisting of elements of infinite order and let f be a function on E satisfying f(x) = 1 for every x ∈ E. Then, for any fixed ε > 0, there exists γ ∈ bG such that γ(x) − f(x) < ε for every x ∈ E. Now, we are ready to prove the theorem on the norm of the inverse of a measure with a discrete part supported on an independent set of points of infinite order. Theorem 2.14. Let µ ∈ M(G) satisfy kµk ≤ 1, inf γ∈bG bµ(γ) = δ > 1 of µd be independent and consist of elements of infinite order. Then kµ−1k ≤ 1 Proof. First of all, by Corollary 2.3 we have inf γ∈bG cµd(γ) ≥ δ. We will deal first with a simpler case, so let us assume that the support of µd is finite, i.e. 2 and let the support 2δ−1 . µd = akδyk for some N ∈ N, complex numbers ak and yk ∈ G, k = 1, 2, . . . , N . NXk=1 INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS We can, of course, arrange the numbers (ak)N k=1 so that a1 ≥ a2 ≥ . . . ≥ aN. addition, as convolution of a measure with a Dirac delta and multiplication by a constant of modulus one do not change the norm of a measure, the norm of the inverse and do not violate the estimate bµ ≥ δ, we can assume, without losing the generality, the following form of µd: 11 In µd = a1δ0 + akδyk where a1 > 0. NXk=2 a1 + a1 + (2.14) for k ∈ {2, . . . , N}. 2, any sequence of signs (ξn)N n=2 and let (αk)N γ0(yk) − αk ≤ Now, we perform elementary estimates Let us now fix ε > 0 with δ − ε > 1 k=2 be a sequence of complex numbers of modulus one such that αkak = ξkak for k = 2, . . . , N. By Theorem 2.13 there is γ0 ∈ bG such that ε PN k=2 ak cµd(γ0) −(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXk=2akξk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) akγ0(yk)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) −(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXk=2akξk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ NXk=2 ≤(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXk=2akξk!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ a1 + akγ0(yk)! − a1 + NXk=2 NXk=2akγ0(yk) − akξk = NXk=2akγ0(yk) − αk ≤ ε by (2.14). NXk=2akγ0(yk) − akαk = NXk=2akξk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≥ δ − ε. Recalling that cµd(γ0) ≥ δ we get (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) This estimate holds true for every choice of signs ξk and every sufficiently small ε > 0 so a1 + a1 + = we obtained, in fact, the following conclusion: a1 + (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) NXk=2akξk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≥ δ > 1 2 for every choice of signs ξk Now, we are in position to use Lemma 2.12 which gives The assertion of the theorem follows now from Lemma 1.1. a1 ≥ δ + N−1Xk=2 ak ≥ δ > 1 2 . Let us move on now to the general case, so let µd be given by the formula (we use the same 12 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI reduction as before): µd = a1δ0 + akδyk where ∞Xk=2 a1 > 0, a1 ≥ a2 ≥ a3 ≥ ... and for every N ∈ N the set {yk}N of elements of infinite order and Pn an < ∞. Put NXk=2 νN = a1δ0 + akδτk . k=2 is independent, consists 2. Then there exists N ∈ N, N := N(ε) such that Fix ε > 0 such that δ − ε > 1 inf γ∈bG cνN (γ) ≥ δ − ε. By the initial part of the proof we get A moment of consideration (we can assume that N(ε) tends to ∞ as ε → 0) leads to a1 ≥ δ − ε + NXn=2an. ∞Xn=2an + δ ≥ δ which finishes the proof. a1 ≥ (cid:3) The proof of the last theorem gives us a bit more. Theorem 2.15. Let µ ∈ M(G) be a measure satisfying the assumptions of Theorem 2.14 and let a > 0 be the mass of the greatest atom in µd. Then kµ−1k ≤ 1 2(kµdk + δ − a) − 1 Remark 2.16. The assumptions of Theorem 2.14 might seem to be quite restrictive but for the classical groups of harmonic analysis: Tn and Rn the situation covered by the theorem is a generic one. 2.2.3. Groups of exponent two. In this section we will deal with a special case of real measures on groups of exponent two (i.e. groups such that 2x = 0 for every x), which has connection to the so-called dyadic analysis. As it will be shown in Theorem 2.18 we obtain the norm-controlled inversion for every δ > 1 2. We start with finite groups Zn 2 : for x, y ∈ Zn ordering on Zn 2 . For convenience, we introduce the following (non-standard) 2 = {0, 1}n we define x < y if the last digit of x non-equal to INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 13 the corresponding digit of y is equal to 0. This order is sometimes called the colexicographic order; the only difference is that we read the strings of digits from right to left. For example, for n = 3 we have 000 < 100 < 010 < 110 < 001 < 101 < 011 < 111. The most important property of this order is that the second half of strings end with 0, and the second half ends with 1. What is more, if we have a string ending with 0, it is easy to find the corresponding string that ends with 1 -- you need to shift by 2n−1. the set of all homomorphisms f : Zn We will also need a precise description of the dual group of Zn 2 which is by definition 2 → {−1, 1}. It is clear that every such f is uniquely determined by its values on the standard system of generators of Zn 2 : 1000...0, 01000...0, ..., 000...1 and for every choice of signs from {−1, 1}n we get a homomorphism. Thus we can write that every homomorphism f ∈ cZn 2 is bijectively identified with an element y ∈ Zn 2 by the formula fy(x) = (−1)<x,y>n where < x, y >n is a formal scalar product on Zn 2 (for example: < 110, 011 >3= 1 · 0 + 1 · 1 + 0 · 1 = 1). Thus, the Fourier-Stieltjes transform of δx for x ∈ Zn 2 is of the form: We will use the following two symbolic operations on Zn 2 : cδx(y) = (−1)<x,y>n for y ∈ Zn 2 . (1) x′ for x ∈ Zn (2) x0, x1 for x ∈ Zn−1 2 (respectively) after the last digit of x. 2 is the element of Zn−1 2 obtained from x by erasing the last digit. is the element of Zn 2 obtained from x by inserting 0 or 1 Lemma 2.17. Let n be a positive integer and let µ ∈ M(Zn kµk ≤ 1 and bµ(γ) ≥ δ > 1 support of µ is greater or equal then δ. 2 for every γ ∈ Zn 2 ) be a real measure satisfying 2 . Then the mass of the greatest atom in the Proof. Let µ = 2n Xk=1 axk δxk, where the elements xk ∈ Zn 2 are ordered as explained before. 14 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI Then, using the introduced notation, our assumption can be rewritten as: ∀l∈{1,...,2n}(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2n (−1)<xk,xl>naxk(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≥ δ. Xk=1 The proof will proceed by induction on n. The case of n = 1 is elementary as the inequalities ax1 + ax2,ax1 − ax2 ≥ δ for real numbers ax1, ax2 clearly imply that the modulus of one of these numbers is greater than δ. Suppose that we have already proven the assertion for s ≤ n − 1. By the definition of our order, the last digit of xl is 0 for l ∈ {1, . . . , 2n−1} which gives for l ∈ {1, . . . , 2n−1}: < x′l, x′k >n−1 for k ∈ {1, . . . , 2n−1} < x′l, x′k−2n−1 >n−1 for k ∈ {2n−1 + 1, 2n} Inserting this to (2.15), we obtain < xl, xk >n=< x′l0, x′k0 >n=( (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2n−1Xk=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∀l∈{1,...,2n−1} l∈{2n−1+1,...,2n} 2n−1Xk=1 ∀ Using similar arguments for elements ending with 1, we get (axk + axk+2n−1 )(−1)<x′ k,x′ . l>n−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) l>n−1(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (axk − axk+2n−1 )(−1)<x′ k,x′ . (2.15) (2.16) (2.17) We perform a substitution bxk := axk + axk+2n−1 , ck := axk − axk+2n−1 for k ∈ {1, . . . , 2n−1}. It is straightforward to verify that the original assertion (for n) reduces to two analogous problems of reduced order (n − 1). Hence we are allowed to use the inductive hypothesis, which yields bxk1 ≥ δ and cxk2 ≥ δ for some k1, k2 ∈ {1, . . . , 2n−1}. If k1 6= k2, we get a contradiction with the norm assumption: 1 ≥ bxk1 + cxk2 ≥ 2δ > 1; indeed, note that bxk1 + ck2 6 axk1 + axk1+2n−1 + axk2 + axk2+2n−1. Thus k1 = k2 and it is enough to use the argument for n = 1 to finish the proof. (cid:3) Theorem 2.18. Let G be a locally compact Abelian group satisfying 2x = 0 for every x ∈ G. Let µ ∈ M(G) be a real measure such that kµk ≤ 1, bµ(γ) ≥ δ > 0 for every γ ∈ bG. Then the mass of the greatest atom in the support of µd is greater than or equal to δ. Moreover, µ is invertible and kµ−1k ≤ 1 2δ−1 . INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 15 Proof. By Theorem 2.2 we have cµd(γ) ≥ δ for every γ ∈ bG and obviously kµdk ≤ 1. Let µd have the following form: (2.18) µd = ∞Xk=1 Fix ε > 0 such that δ − ε > 1 2 and find Nε ∈ N for which akδxk for some ak ∈ R and xk ∈ G. ∞Xk=Nε+1ak < ε. Define µε as a sum as in (2.18) from k = 1 to Nε. Then clearly kµεk ≤ 1 and cµε(γ) ≥ 2 for every γ ∈ bG. Now it is enough to observe that µε may be treated as an δ − ε > 1 element of M(ZNε 2 ) and apply Lemma 2.17. To finish the proof we pass with ε to zero and use Lemma 1.1. (cid:3) 3. Fourier -- Stieltjes algebras We switch now to Fourier -- Stieltjes algebras (the standard reference for this part is [Ey], see also a recent monograph [KL]). Let G be a locally compact group and let B(G) be the Fourier -- Stieltjes algebra of the group G i.e. the linear span of continuous positive- definite functions on G equipped with the norm given by the duality B(G) = (C∗(G))∗, where C∗(G) is the full group C∗-algebra of G. It is well-known that B(G) with pointwise product is a commutative semisimple unital Banach algebra with G embedded in △(B(G)) via evaluation functionals. A continuous bounded function on G is called (weakly) almost periodic if the set of all its translates is precompact in the (weak) uniform topology, the set of all such functions will be denoted by AP (G) (W AP (G)). A classical result of Eberlein (see [E] for the original work and [GL] for a different proof) states that B(G) ⊂ W AP (G). Moreover, there exists an invariant mean M ∈ (W AP (G))∗ (the existence of the mean and related matters are discussed in detail in the book [B]). Restricting this mean to B(G) we obtain a decomposition B(G) = Bc(G) ⊕ (B(G) ∩ AP (G)) where Bc(G) = {f ∈ B(G) : M(f) = 0} is a closed ideal in B(G) and B(G)∩AP (G) is a closed subalgebra. Here Bc(G) is the analogue of the ideal of continuous measures and B(G)∩AP (G) is the analogue of the 16 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI subalgebra of discrete measures for commutative groups. The investigations on Fourier -- Stieltjes algebras are natural generalizations of those concerning measure algebras, as in case of an Abelian group G we have a canonical isomorphism of B(G) with M(bG). As we work already 'on the Fourier transform side', the variant of Problem 1 can be rephrased in our context as follows. Problem 2. Let G be a locally compact group and let f ∈ B(G) satisfy kfk ≤ 1 and inf x∈G f(x) = δ > 0. What is the minimal value of δ0 > 0 such that for any δ > δ0 the element f is automatically invertible? What can be said about the norm of the inverse? 3.1. Qualitative result. The inspection of the proof of Theorem 2.4 shows that in order to obtain an analogue for Fourier -- Stieltjes algebras (and answer the first question from Problem 2) we need only to show the version of Theorem 2.2 for Fourier-Stieltjes algebras (indeed, we have only used the basic facts from Gelfand theory, the orthogonal decompo- sition of a measure into continuous and discrete part and the density of bG in △(Md(G)) -- all of these facts have standard counterparts for Fourier -- Stieltjes algebras). However, the original proof cannot be repeated verbatim and we use a different approach in the second part of the argument. Theorem 3.1. Let G be a locally compact group and let f ∈ B(G) have the decomposition f = g + h with g ∈ (AP (G) ∩ B(G)) and h ∈ Bc(G). Then g(G) ⊂ f(G). Proof. Without loss of generality, it is enough to show g(e) ∈ f(G) (the general case being obtained via considering shifted f). First we prove the existence of distinct group elements (xn)∞n=1 satisfying (3.1) h(x1) < ε and h(xnx−1 j ) < ε for j < n and n > 1. Since h ∈ Bc(G) and the mean M satisfies M(f) ≥ 0 if f ≥ 0 we clearly have inf x∈G h(x) = 0 so there is no problem with the choice of x1. Suppose that we have already picked x1, . . . , xn−1 and consider an auxiliary function u ∈ Bc(G) defined as follows: u(x) = n−1Xj=1 h(xx−1 j )2. INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 17 As u ∈ Bc(G) w also have inf x∈G u(x) = 0 and we are able to choose xn different from x1, . . . , xn−1 or h(e) < ε which immediately finishes the proof. Let us consider now the following set of functions in B(G) ∩ AP (G). X = {gxn : n ∈ N} where gx(y) = g(xy) for x, y ∈ G. Since g ∈ AP (G) this set is relatively compact in the uniform topology. Hence there exists a subsequence (gxnk )k∈N which is a Cauchy sequence. It follows that for k > N we have kgxnk+1 − gxnkk∞ < ε. In particular, But now, gxnk+1 nk ) − gxnk (x−1 (x−1 nk ) = g(xnk+1x−1 nk ) − g(e) < ε. f(xnk+1x−1 nk ) − g(e) = h(xnk+1x−1 nk ) + g(xnk+1x−1 nk ) − g(e) ≤ 2ε. (cid:3) For completeness, let us state a variant of Theorem 2.4 for Fourier-Stieltjes algebras. 2. Then f is invertible. Theorem 3.2. Let G be a locally compact group and let f ∈ B(G) satisfy kfk ≤ 1 and inf x∈G f(x) > 1 3.2. Quantitative results. The aim of this short section is to show that Nikolski's esti- mate for δ > 1√2 can be achieved for Fourier-Stieltjes algebras as well. In order to do that, we need to recall some facts about absolute continuity and L-spaces from [OW]. Definiton 3.3 (cf. [Ex, Definition 2.] and [OW, Definition 4.7]). Let G be a locally compact group and let f, g ∈ B(G). Let zs(f) and zs(g) be the central supports of f and g, respectively, where f and g are viewed as functionals on C∗(G). We say that f is absolutely continuous with respect to g (denoted by f ≪ g) if zs(f) ≤ zs(g). We call f and g mutual ly singular if zs(f)zs(g) = 0. Proposition 3.4 (cf. [OW, Proposition 4.18] and [OW, Corollary 4.19]). Let X ⊂ B(G) be a closed subspace that is left and right invariant by the action of G. Then there exists 18 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI a central projection z ∈ W ∗(G) := (C∗(G))∗∗ such that X = z · B(G). Consequently, we have a splitting B(G) = X ⊕ X⊥, where X⊥ := (1 − z) · B(G), which is orthogonal, i.e. all elements from X⊥ are mutually singular with elements from X. We are now able to show the following fact. Proposition 3.5. Let f ∈ B(G). Then there exists an orthogonal decomposition f = c1 + f0, where f0 and 1 are mutually singular. Moreover, c = m(f), where m is the unique invariant mean on B(G). Proof. Since the subspace consisting of constant functions is G-invariant, we get an orthog- onal decomposition by the previous proposition. Denote the constant appearing therein by ϕ(f). By uniqueness of the decomposition, it is easy to see that ϕ : B(G) → C is a linear functional that satisfies ϕ(1) = 1. Since the decomposition is orthogonal, we get kfk = ϕ(f) +kf0k, so ϕ is a continuous functional. Note also that ϕ is G-invariant (from both left and right). Indeed, for any g ∈ G the equality f = ϕ(f)1 + f0 implies that g · f = ϕ(f)1 + g · f0, but also g · f = ϕ(g · f)1 + (g · f)0, hence ϕ(g · f) = ϕ(f) by uniqueness; in the same way we obtain invariance from the right. This is almost exactly the definition of an invariant mean, but we do not know if ϕ is a positive functional. We will not show this directly; in the next lemma we will prove that there exists a unique G-invariant functional that attains the value 1 at 1, hence it has to coincide with the invariant mean. (cid:3) Lemma 3.6 (see also [LL, Proposition 4.10]). Let ϕ : B(G) → C be a linear functional such that ϕ(1) = 1 and ϕ(f) = ϕ(g · f) = ϕ(f · g) for any g ∈ G. Then it is equal to the unique invariant mean on B(G). Proof. It suffices to show that there is at most one functional ϕ satisfying the assumptions of the lemma. Recall that in order to show that two such functional are equal, it suffices to show that their kernels coincide; this will force them to be proportional and the condition ϕ(1) = 1 will finish the proof. Therefore we have to check that the kernel of ϕ is uniquely specified by the conditions spelled out in the lemma. Note that kerϕ is a G-invariant INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 19 subspace, so it is of the form z · B(G) for some central projection in W ∗(G). Let ztriv be the central projection such that ztrivB(G) = C1. We will show that z = 1 − ztriv. Since ϕ(1) = 1, it cannot be the case that ztriv ≤ z. As ztriv is a minimal projection, it follows that z ≤ 1 − ztriv. If the inequality was strict, then the codimension of the kernel would be greater than 1, which cannot happen for a linear funcitonal, hence z = 1 − ztriv. (cid:3) With these tools we are able to conclude. 2δ2−1. f k 6 1 Proposition 3.7. Let f ∈ B(G) satisfy kfk 6 1 and f > δ, where δ > 1√2. Then f is invertible and k 1 Proof. As f > δ, f2 > δ2 > 1 2. By Proposition 3.5 we get f2 = m(f2)1 + f0. Since the mean m(f) is characterised as being the unique constant that can be uniformly approximated with convex combinations of translates of f, we get m(f2) ≥ δ2 > 1 2. As kf2k 6 1, we get that kf0k ≤ 1 − δ2 < 1 2. This means that we can write the inverse of f2 using the Neumann series (just like in Lemma 1.1) and quickly obtain the estimate for the norm of the inverse by (cid:3) 1 2δ2−1; it just remains to note that 1 f = f f2 . The proofs contained in this note clearly suggest that the most important information concerning invertibility is contained in the discrete (or almost periodic) part. In the non- commutative setting it is possible that the almost periodic part is very simple. Namely, there is a class of groups, called minimally almost periodic, for which constants are the only almost periodic functions; an example of such a group is SL2(R). For them we get a better result. 2. Then f is invertible and k 1 Proposition 3.8. Let G be a minimally almost periodic group and let f ∈ B(G) satisfy kfk 6 1 and f ≥ δ > 1 Proof. Let f = fap + fc be a decomposition of f into almost periodic and continuous parts, and recall that m(fc) = 0. It follows that m(f) = m(fap). As the only almost periodic functions are the constants, we get fap = m(f)1, hence m(f) = m(f) ≥ δ. We have an 2, so kfck ≤ 1 − δ < 1 orthogonal decomposition f = m(f)1 + fc, where m(f) ≥ δ > 1 f k ≤ 1 2δ−1 . 2 20 PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI and the usual procedure involving the Neumann series proves invertibility and provides the estimate. (cid:3) There are two more simple cases in which we get the same estimate for the inverse; this time there are very strong assumptions about the function. Proposition 3.9. Let G be a locally compact group and let f ∈ B(G) satisfy kfk ≤ 1 and f > δ > 1 2δ−1 if one of the following conditions is satisfied: 2. Then f is invertible and k 1 f k 6 1 (i) f is real and has constant sign; (ii) the absolute value of f is constant. Proof. (i) We may assume that f is positive. Recall the decomposition from Proposition 3.5: f = m(f)1 + f0. Since f > δ, we get m(f) > δ, so f = m(f)(cid:16)1 + f0 inverted using the Neumann series, which also gives the estimate. c , which gives k 1 (ii) In this case we have f f = c > δ2, hence 1 m(f )(cid:17) can be 2δ−1. Note f k 6 1 δ2 6 1 that here the assumption δ > 1 f = f 2 is not necessary. 4. Concluding remarks (1) The critical constant δ0 = 1 2 is the same as for measure algebras on 'analytic' semigroups as discussed in part 2.2. of [N]. (cid:3) (2) The problem of obtaining an estimate of the inverse for measures µ ∈ M(G) sat- 2 , 1√2(cid:17) seems to be very difficult even for isfying kµk ≤ 1 and inf γ∈bG bµ(γ) ∈ (cid:16) 1 G = T and the considered measures being finite sums of Dirac deltas supported at rational points of the circle. Then the estimation of the inverse is translated to an elementary but surprisingly hard problem on bounding the entries of the inverses of circulant matrices in terms of their eigenvalues. It should be also noted that it cannot be solved by simply proving that the mass of a single atom is greater 2 δ1 ∈ M(Z2) shows (such examples can be also than δ as the example µ = 1 2 δ0 + i constructed for non-torsion groups such as Z). INVERSION PROBLEM IN MEASURE AND FOURIER -- STIELTJES ALGEBRAS 21 (3) Theorem 2.4 is sharp for any non-discrete locally compact Abelian group as it was discussed before but Theorem 3.2 is not necessarily sharp even for non-compact groups as there are examples of non-compact groups for which the Wiener-Pitt phenomenon does not occur (check [C]). However, this Theorem is sharp for discrete groups containing an infinite Abelian subgroup (consult [OW]) 5. Acknowledgements The authors would like to thank Maria Roginskaya and Michał Wojciechowski for many valuable discussions. They would also like to thank Nikolai Nikolski for careful reading of the manuscript and useful remarks. Parts of this work were completed during a visit of Mateusz Wasilewski to Chalmers University of Technology; he thanks the university and the host for a stimulating research environment. References [B] [C] [E] [Ey] [Ex] [GW] [GL] [KL] [LL] R. B. Burckel: Weakly almost periodic functions on semigroups, Gordon and Breach, New York, 1970 M. Cowling: The Fourier-Stieltjes algebra of a semisimple group, Colloq. Math., vol. 41, no. 1, pp. 89 -- 94, 1979. W. F. Eberlein: Abstract ergodic theorems and weak almost periodic functions, Transactions of the Amer. Math. Soc., vol. 67, pp. 217 -- 240, 1949. P. Eymard: L'algebre de Fourier d'un groupe localement compact, Bull. Soc. Math. France, vol. 92, pp. 181 -- 236, 1964. R. Exel: The F. and M. Riesz theorem for C∗-algebras, J. Operator Theory, vol. 23, no.2, pp. 351 -- 368, 1990. I. Glicksberg, I. Wik: The range of Fourier-Stieltjes transforms of parts of a mea- sures, Lecture Notes in Mathematics, vol. 266, pp 73 -- 77, 1971. Colin C. Graham, Anthony T. M. Lau: Relative weak compactness of orbits in Banach spaces associated with locally compact groups, Trans. Amer. Math. Soc., vol.359, no. 3, pp. 1129 -- 1160, 2007. E. Kaniuth, Anthony T. M. Lau: Fourier and Fourier-Stieltjes algebras on Locally Compact Abelian Groups, AMS 2018. Anthony T. M. Lau, J. Ludwig: Fourier-Stieltjes algebra of a topological group, Adv. Math., vol. 229, no. 3, p. 2000 -- 2023, 2012. 22 [N] PRZEMYSŁAW OHRYSKO AND MATEUSZ WASILEWSKI N. Nikolski: In search of the invisible spectrum, Ann. Inst. Fourier (Grenoble), vol. 49, no. 6, pp. 1925 -- 1998, 1999. [OW] P. Ohrysko, M. Wasilewski: Spectral theory of Fourier-Stieltjes algebras, J. Math. Anal. Appl., vol. 473, no. 1, pp. 174 -- 200, 2019. W. Rudin: Fourier analysis on groups, Wiley Classics Library, 1990. W. Żelazko: Banach algebras, Elsevier Science Ltd and PWN, February 1973. [R] [Ż] Chalmers University of Technology and the University of Gothenburg E-mail address: [email protected] KU Leuven E-mail address: [email protected]
1302.0652
1
1302
2013-02-04T11:27:50
Coupling and relaxed commutant lifting
[ "math.FA" ]
A Redheffer type description of the set of all contractive solutions to the relaxed commutant lifting problem is given. The description involves a set of Schur class functions which is obtained by combining the method of isometric coupling with results on isometric realizations. For a number of special cases, including the case of the classical commutant lifting theorem, the description yields a proper parameterization of the set of all contractive solutions, but examples show that, in general, the Schur class function determining the contractive lifting does not have to be unique. Also some sufficient conditions are given guaranteeing that the corresponding relaxed commutant lifting problem has only one solution.
math.FA
math
COUPLING AND RELAXED COMMUTANT LIFTING A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK Abstract. A Redheffer type description of the set of all contractive solutions to the relaxed commutant lifting problem is given. The description involves a set of Schur class functions which is obtained by combining the method of isometric coupling with results on isometric realizations. For a number of special cases, including the case of the classical commutant lifting theorem, the description yields a proper parameterization of the set of all contractive solutions, but examples show that, in general, the Schur class function deter- mining the contractive lifting does not have to be unique. Also some sufficient conditions are given guaranteeing that the corresponding relaxed commutant lifting problem has only one solution. 0. Introduction This paper is devoted to the relaxed commutant lifting theorem in [14]. This theorem is a generalization of the classical commutant lifting theorem [19], and it includes as special cases the Treil-Volberg lifting theorem [20], and its weighted version due to Biswas, Foias and Frazho [12]. To state the relaxed commutant lifting theorem, let us first recall the general setup. The starting point is a lifting data set {A, T ′, U ′, R, Q} consisting of five Hilbert space operators. The operator A is a contraction mapping H into H′, the operator U ′ on K′ is a minimal isometric lifting of T ′ on H′, and R and Q are operators from H0 to H, satisfying the following constraints T ′AR = AQ and R∗R ≤ Q∗Q. Given this data set the relaxed commutant lifting theorem in [14] states that there exists a contraction B from H to K′ such that (0.1) ΠH′ B = A and U ′BR = BQ. Here ΠH′ is the orthogonal projection from K′ onto H′. In fact, [14] provides an explicit construction for a contraction B satisfying (0.1). In the sequel we say that B is a contractive interpolant for {A, T ′, U ′, R, Q} if B is a contraction from H into K satisfying (0.1). In this paper we present a Redheffer type formula to describe the set of all contractive interpolants for {A, T ′, U ′, R, Q}. In order to state our main results we need some auxiliary operators. To this end, let D◦ be the positive square root of 1991 Mathematics Subject Classification. Primary 47A20, 47A57; Secondary 47A48. Key words and phrases. commutant lifting, isometric coupling, isometric realization, parameterization. 1 2 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK Q∗Q − R∗R, and set (0.2) F = DAQH0 and F ′ =   D◦ DT ′ AR DAR   H0. Notice that F is a subspace of DA and F ′ is a subspace of D◦ ⊕ DT ′ ⊕ DA. Here we follow the convention that for a contraction C, the symbol DC denotes the positive square root of I − C∗C and DC stands for the closure of the range of DC. Furthermore, D◦ = D◦H0. Since T ′AR = AQ, we know from formula (4.11) in [14] that there exists a unique unitary operator ω mapping F onto F ′ such that (0.3) ω(DAQh) =   D◦ DT ′ AR DAR   h, h ∈ H0. We also need the projections ΠT ′ and ΠA defined by (0.4) ΠT ′ = (cid:2) 0 I 0 (cid:3) :   D◦ DT ′ DA   → DT ′ , ΠA = (cid:2) 0 0 I (cid:3) :   D◦ DT ′ DA   → DA. Notice that the previous definitions only relied upon the operators A, T ′, R and Q. The minimal isometric lifting U ′ did not play a role. Recall that all minimal isometric liftings of the same contraction are isomorphic. So without loss of gen- erality, in our main theorem, we can assume that U ′ = V is the Sz.-Nagy-Schaffer minimal isometric lifting of T ′ which acts on H′ ⊕ H 2(DT ′ ). The definitions of a minimal isometric lifting and the Sz.-Nagy-Schaffer lifting are presented in the next section. Finally, given Hilbert spaces U and Y, we write S(U , Y) for the set of all operator-valued functions which are analytic on the open unit disk D and whose values are contractions from U to Y. We refer to S(U , Y) as the Schur class associated with U and Y. We are now ready to state our first main result. Theorem 0.1. Let {A, T ′, V, R, Q} be a lifting data set, where V on H′ ⊕ H 2(DT ′ ) is the Sz.-Nagy-Schaffer minimal isometric lifting of T ′. Then all contractive in- terpolants for this date set are given by (0.5) Bh = (cid:20) ΠT ′ F (λ)(IDA − λΠAF (λ))−1DAh (cid:21) , Ah h ∈ H, where F is any function from the Schur class S(DA, D◦ ⊕ DT ′ ⊕ DA) satisfying F (0)F = ω. In general, formula (0.5) does not establish a one to one correspondence between B and the parameter F . It can happen that different F 's yield the same B. For instance, assume H0, H and H′ to be equal to C, let A, R and Q be the zero operator on C, and take for T ′ the identity operator on C. Since T ′ is an isometry, the Sz.- Nagy-Schaffer minimal isometric lifting V of T ′ is equal to T ′. The latter implies that there is only one contractive interpolant B for the data set {A, T ′, V, R, Q}, namely B = A. The fact that R and Q are the zero operators on C implies that F = {0} and F ′ = {0}. It follows that for this data set {A, T ′, V, R, Q} the only contractive interpolant B is given by formula (0.5) where for F we can take any function in the Schur class S(C, C). The previous example can be seen as a special case of our second main theorem. COUPLING AND RELAXED COMMUTANT LIFTING 3 Theorem 0.2. Let B be a contractive interpolant for the data set {A, T ′, V, R, Q} where V is the Sz.-Nagy-Schaffer minimal isometric lifting of T ′. Then there is a one to one mapping from the set of all F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω such that B is given by (0.5) onto the set S(GB , G′ B being given by B), with GB and G′ (0.6) GB = DB ⊖ DBQH0 and G′ B = (D◦ ⊕ DB) ⊖(cid:20) D◦ DBR (cid:21) H0. Our proof of the above theorem also provides a procedure to obtain a mapping of the type referred to in the theorem. It is interesting to specify Theorems 0.1 and 0.2 for the case when in the lifting data set {A, T ′, V, R, Q} the operators A, T ′, R and Q are zero operators. In this case the intertwining condition V BR = BQ, where V is the Sz.-Nagy-Schaffer minimal isometric lifting of T ′ = 0, is trivially fulfilled, and hence B is a contractive interpolant if and only if 0 Bh = (cid:20) where Θ is any function in H 2 ball(L(H, H′)). The latter means that Θ is a L(H, H′)- valued analytic function on D such that for each h ∈ H the function Θ(·)h belongs to the Hardy space H 2(H′), and kΘ(·)hkH 2(Y) ≤ khk. It follows that Theorems 0.1 and 0.2 have the following corollaries. Θ(·)h (cid:21) , h ∈ H, Corollary 0.3. Let F be any function in the Schur class S(H, H′ ⊕ H), and let Π and Π′ be the orthogonal projections of H′ ⊕ H on H and H′, respectively. Then the function Θ defined by (0.7) Θ(λ) = Π′F (λ)(IH − λΠF (λ))−1 belongs to H 2 way ball(L(H, H′)), and any function in H 2 ball(L(H, H′)) is obtained in this Corollary 0.4. Let Θ ∈ H 2 ball(L(H, H′)). Then there is a one to one mapping from the set of all F in S(H, H′ ⊕ H) such that (0.7) holds onto the set S(DΓ, DΓ), where Γ is the contraction from H into H 2(H′) defined by (Γh)(λ) = Θ(λ)h, h ∈ H, λ ∈ D. When H = H′ = C, and hence Θ is a scalar function, Corollary 0.3 can be found in [18], page 490, provided Θ is of unit H 2 norm. For p × q matrix functions Θ, when H = Cq and H′ = Cp, Corollary 0.3 is Theorem 2.2 in [3]. For the general operator valued case Corollary 0.3 seems to be new. Corollary 0.4 seems to be new even in the scalar case. Notice that in the scalar case the space DΓ in Corollary 0.4 consists of the zero element only if Θ is of unit H 2 norm, and DΓ = C otherwise. Another case of special interest is the classical commutant lifting problem. As we know from [14] the commutant lifting theorem can be obtained by applying the relaxed commutant lifting theorem to the data set {A, T ′, U ′, IH, Q} where H0 = H, the operator R is the identity operator on H and Q is an isometry; see [14]. In this case, the space G′ B in Theorem 0.2 consists of the zero element for any choice of the contractive interpolant B. In other words, for the case of the classical commutant lifting formula (0.5) provides a proper parameterization, that is, for every contractive interpolant B for {A, T ′, V, R, Q} there exists a unique F 4 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω such that B is given by (0.5). Finally, it is noted that this formula also yields the Redheffer type parameterization for the commutant lifting theorem presented in Section XIV of [13]. If in Theorem 0.1 we take F (λ) ≡ ωΠF , where ΠF is the orthogonal projection of DA onto F , then the contractive interpolant B in (0.5) is precisely the central solution presented in [14]. From Theorem 0.1 we see that F = DA implies that there is a unique contrac- tive interpolant (which is known from Theorem 3.1 in [14]). Other conditions of uniqueness will be given in the final section of the paper. We shall prove Theorems 0.1 and 0.2 by combining the method of isometric coupling with some aspects of isometric realization theory. The theory of isometric couplings originates from [1], [2], and was used to study the commutant lifting problem for the first time in [5] -- [9]; see also, Section VII.7 in [13]. The paper consist of six sections not counting this introduction. The first two sections have a preliminary character, and review the notions of an isometric lifting (Section 1), and an isometric realization (Section 2). In the third section we develop the notion of an isometric coupling of a pair of contractions which provides the main tool in this paper. In Section 4 we prove Theorem 0.1 for the case when R∗R = Q∗Q, and in Section 5 we prove Theorem 0.1 in its full generality. In the final section we prove Theorem 0.2, and we present a few sufficient conditions for the case when (0.5) provides a proper parameterization, and also conditions for uniqueness of the solution. We conclude this introduction with a few words about notation and terminology. Throughout capital calligraphic letters denote Hilbert spaces. The Hilbert space direct sum of U and Y is denoted by U ⊕ Y or by (cid:20) U Y (cid:21) . The set of all bounded linear operators from H to H′ is denoted by L(H, H′). The identity operator on the space H is denoted by IH or just by I, when the underlying space is clear from the context. By definition, a subspace is a closed linear manifold. If M is a subspace of H, then H ⊖ M stands for the orthogonal complement of M in H. Given a subspace M of H, the symbol ΠM will denote the orthogonal projection of H onto M viewed as an operator from H to M, and PM will denote the orthogonal projection of H onto M viewed as an operator on H. Note that Π∗ M is the canonical embedding from M into H, and hence PM = Π∗ MΠM. Instead of Π∗ M we shall sometimes write EM, where the capital E refers to embedding. A subspace M of H is said to be cyclic for an operator T on H whenever H = ∞ _n=0 T nM = span {T nM n = 0, 1, 2, . . .}. Finally, by definition, a L(H, H′)-valued Schur class function is a function in S(H, H′), i.e., an operator-valued function which is analytic on the open unit disk D and whose values are contractions from H to H′. COUPLING AND RELAXED COMMUTANT LIFTING 5 1. Isometric liftings In this section we review some facts concerning isometric liftings that are used throughout this paper. For a more complete account we refer to the book [16] (see also Chapter VI in [13], and Section 11.3 in [14]). Let T ′ on H′ be a contraction. Recall that an operator U on K is a isometric lifting of T ′ if H′ is a subspace of K and U is an isometry satisfying ΠH′ U = T ′ΠH′. Isometric liftings exist. In fact, the Sz.-Nagy-Schaffer isometric lifting V of T ′ is given by (1.1) V = (cid:20) T ′ EDT ′ S (cid:21) on (cid:20) 0 H′ H 2(DT ′ ) (cid:21) . Here S is the unilateral shift on the Hardy space H 2(DT ′ ) and E is the canonical embedding of DT ′ onto the space of constant functions in H 2(DT ′ ). To see that V in (1.1) is an isometric lifting of T ′ note that any operator U on K = H′ ⊕ M is an isometric lifting of T ′ if and only if U admits an operator matrix representation of the form (1.2) U = (cid:20) is an isometry. T ′ Y1DT ′ Y2 (cid:21) on (cid:20) H′ 0 M (cid:21) where Y = (cid:2) Y1 Y2 (cid:3) : (cid:20) DT ′ M (cid:21) → M An isometric lifting U of T ′ is called minimal when H′ is cyclic for U . The Sz.-Nagy-Schaffer isometric lifting of T ′ is minimal. If the isometric lifting U is given by (1.2), then the lifting is minimal if and only if the space Y1DT ′ is cyclic for Y2. Two isometric liftings U1 on K1 and U2 on K2 of T ′ are said to be isomorphic if there exists a unitary operator Φ from K1 onto K2 such that ΦU1 = U2Φ and Φh = h for all h ∈ H′. Minimality of an isometric lifting is preserved under an isomorphism, and two minimal isometric liftings of T ′ are isomorphic. Finally, when U on K is a isometric lifting of T ′, then the subspace K′, given by K′ = ∞ _n=0 U nH′, is reducing for U , that is, both K′ and its orthogonal complement K = K ⊖ K′ are invariant under U . Furthermore, in that case the operator U ′ = ΠK′ U K′ on K′ is a minimal isometric lifting of T ′, and the operator U admits a operator matrix decomposition of the form (1.3) U = (cid:20) U ′ 0 0 U (cid:21) on (cid:20) K′ K (cid:21) , where U is an isometry on K. We shall call U ′ in (1.3) the minimal isometric lifting of T ′ associated with U . The following proposition summarizes the results referred to above in a form that will be convenient for this paper. For details we refer to Section 11.3 in [14]. Theorem 1.1. Let T ′ be a contraction on H′, let V on H′ ⊕ H 2(DT ′ ) be the Sz.- Nagy-Schaffer (minimal) isometric lifting of T ′, and let U on H⊕M be an arbitrary isometric lifting of T ′ given by (1.2). Then there exists a unique isometry Φ from 6 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK H′ ⊕ H 2(DT ′ ) into H′ ⊕ M such that U Φ = ΦV and ΦH′ = IH′ . In fact, Φ is given by Φ = (cid:20) IH′ 0 0 Λ (cid:21) : (cid:20) where Λ is defined by H′ H 2(DT ′ ) (cid:21) → (cid:20) H′ M (cid:21) , Λh = ∞ Xn=0 Y n 2 Y1hn, h(λ) = ∞ Xn=0 λnhn ∈ H 2(DT ′ ), 2 )−1m for each with Y1 and Y2 as in (1.2). Moreover, (Λ∗m)(λ) = Y ∗ m ∈ M. Finally, Φ is unitary if and only if U is a minimal isometric lifting of T ′, and in that case the isometric liftings V and U of T ′ are isomorphic. 1 (I − λY ∗ The isometry Φ introduced in the above theorem will be referred to as the unique isometry associated with T ′ that intertwines V with U . Since V is uniquely deter- mined by T ′, we shall denote this isometry simply by ΦU, T ′ . When U on K is an isometric lifting of T ′ and U ′ on K′ is the minimal isometric lifting of T ′ as- sociated with U , then the operator ΠK′ΦU, T ′ is the unique isometry associated with T ′ that intertwines V with U ′, that is, ΦU ′, T ′ = ΠK′ΦU, T ′ or, equivalently, Π∗ K′ΦU ′, T ′ = ΦU, T ′ . 2. Isometric realizations In this section we review some of the classical results on controllable isometric realizations, and we prove a few additional results that will be useful in the later sections. We say that {Z, B, C, D; X , U , Y} (or simply {Z, B, C, D}) is a realization of a L(U , Y)-valued function G if (2.1) G(λ) = D + λC(IX − λZ)−1B for all λ in some open neighborhood of the origin in the complex plane. Here Z is an operator on X and B is an operator from U into X while C is an operator mapping X into Y and D is an operator from U into Y (where X , U and Y are all Hilbert spaces). In this case, we refer to the function defined by the right hand side of (2.1) as the associated transfer function. A realization {Z, B, C, D} is called isometric if the operator (2.2) M = (cid:20) D C B Z (cid:21) : (cid:20) U X (cid:21) → (cid:20) Y X (cid:21) is an isometry. The 2 × 2 operator matrix in (2.2) is called the system matrix associated with the realization {Z, B, C, D}. The transfer function of an isometric realization belongs to the Schur class S(U , Y), that is, if {Z, B, C, D} is an isometric realization, then the function G defined by (2.1) is a contractive analytic L(U , Y)- valued function on D. Conversely, if G ∈ S(U , Y), then there is an isometric realization {Z, B, C, D} such that (2.1) holds for all λ in D. The transfer function of a realization can also be expressed in terms of the system matrix M . In fact, if {Z, B, C, D} is a realization and M is the associated system matrix, then in a neighborhood of the origin the transfer function G is also given by (2.3) G(λ) = ΠY M (IU ⊕X − λJX M )−1Π∗ U , COUPLING AND RELAXED COMMUTANT LIFTING 7 where JX is the partial isometry from Y ⊕ X to U ⊕ X given by JX = (cid:20) 0 0 0 IX (cid:21) : (cid:20) Y X (cid:21) → (cid:20) U X (cid:21) . Indeed, for λ sufficiently close to zero we have G(λ) = D + λC(IX − λZ)−1B = (cid:2) D C (cid:3)(cid:20) IU λ(IX − λZ)−1B (cid:21) 0 IU = (cid:2) D C (cid:3)(cid:20) = (cid:2) D C (cid:3)(cid:20) IU = (cid:2) D C (cid:3)(cid:18)IU ⊕X − λ(cid:20) 0 λ(IX − λZ)−1B (IX − λZ)−1 (cid:21)(cid:20) IU 0 (cid:21) −λB IX − λZ (cid:21)−1(cid:20) IU 0 (cid:21) B Z (cid:21)(cid:19)−1(cid:20) IU 0 IX (cid:21)(cid:20) D C 0 (cid:21) 0 0 = ΠY M (IU ⊕X − λJX M )−1Π∗ U . Since the right side of (2.1) is a Schur class function if M in (2.2) is an isometry, the same holds true for the right hand side of (2.3). Notice that the function G defined by (2.3) can also be written in the form: If for a realization {Z, B, C, D; X , U , Y} the space W∞ G(λ) = ΠY (IY⊕X − λM JX )−1M Π∗ U . n=0 Z nBU is equal to X , then the realization or the pair {Z, B} is called controllable. In other words, a realization is controllable if and only if the space BU is cyclic for Z. In terms of the system matrix M in (2.2) the realization {Z, B, C, D} is controllable if and only if (2.4) X = ΠX ∞ _n=0 (JX M )n(cid:20) U {0} (cid:21) . The above condition (2.4) is also equivalent to the requirement that {JX M, Π∗ U } is a controllable pair. In the particular case when U = Y in (2.2), condition (2.4) can be written in an even simpler form. This is the contents of the next lemma. Lemma 2.1. Let M be as in (2.2), and assume U = Y. Then {Z, B} is controllable if and only if U ⊕ {0} is cyclic for M , that is, (2.5) ∞ _n=0 M n(cid:20) U {0} (cid:21) = (cid:20) U X (cid:21) . Proof. Let EU be the canonical embedding of U into U ⊕ X , and define M0 to be the operator M0 = (cid:20) 0 B Z (cid:21) : (cid:20) U X (cid:21) → (cid:20) U X (cid:21) . 0 Then M0 = M − EU (cid:2) C D (cid:3). This feedback relation implies that the pair {M0, EU } is controllable if and only if the pair {M, EU } is controllable. Thus (2.5) holds if and only if {M0, EU } is controllable. Now notice that for all integers n ≥ 1, we have M n 0 EU = (cid:20) 0 Z n−1B Z n (cid:21)(cid:20) IU 0 (cid:21) = (cid:20) 0 0 Z n−1B (cid:21) : U → (cid:20) U X (cid:21) . 8 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK It follows that ∞ _n=0 M n 0 EU U = (cid:20) U {0} (cid:21) ⊕ ∞ _n=1 M n 0 EU = U ⊕ ∞ _n=1 Z n−1BU . We conclude that (2.5) holds if and only if the pair {Z, B} is controllable. (cid:3) A realization {Z, B, C, D} or the pair {C, Z} is called observable if CZ nx = 0 for all integers n ≥ 0 implies that the vector x is equal to zero. Since the orthogonal complement of Ker CZ n is equal to the closure of Im (Z ∗)nC∗, we see that observability of the realization {Z, B, C, D} is equivalent to the controllability of the dual realization {Z ∗, C∗, B∗, D∗}. Two realizations {Z1 on X1, B1, C1, D1} and {Z2 on X2, B2, C2, D2} are said to be unitarily equivalent if D1 = D2 and there exists a unitary operator W mapping X1 onto X2 such that W Z1 = Z2W, W B1 = B2 and C2W = C1. Unitary equivalence does not change the transfer function. More precisely, when two realizations are unitary equivalent, then their transfer functions coincide in a neighborhood of zero. For isometric controllable realizations the converse is also true. In fact we have the following theorem. Theorem 2.2. Let G be a L(U , Y)-valued function. Then G ∈ S(U , Y) if and only if G admits an isometric realization. In this case, G admits a controllable isomet- ric realization and all controllable isometric realizations of G are unitarily equiv- alent. In particular, formula (2.1) provides a one to one correspondence between the L(U , Y)-contractive analytic functions on D and (up to unitary equivalence) the controllable isometric realizations of L(U , Y)-valued functions. The above result appears in a somewhat different form in [19] as a theorem representing a Schur class function as a characteristic operator function. A full proof, with isometric systems replaced by their dual ones, can be found in [4] which also gives additional references. In Section 1.3 of [15] the theorem is proved using the Naimark dilation theory. We conclude this section with a proposition that will be useful in the later sections. The starting point is an isometry Y of the type appearing in (1.2). More precisely, (2.6) Y = (cid:2) Y1 Y2 (cid:3) : (cid:20) D′ M (cid:21) → M. Proposition 2.3. Let Y in (2.6) be an isometry. Assume M = D ⊕ X , and let ΠD and ΠX be the orthogonal projections of M onto D and X , respectively. Put (2.7) where ΠD ′⊕D is the orthogonal projection of D′ ⊕ M onto D′ ⊕ D, and F (λ) = ΠD ′⊕DY ∗(IM − λJ ′ X Y ∗)−1Π∗ D, λ ∈ D, X : D′ ⊕ M → M, J ′ X (d′ ⊕ m) = ΠX m. J ′ Then F belongs to the Schur class S(D, D′ ⊕ D) and (2.8) 1 (IM − λY ∗ Y ∗ 2 )−1Π∗ where Π and Π′ are the orthogonal projections of D′⊕D onto D and D′, respectively. D = Π′F (λ)(cid:16)ID − λΠF (λ)(cid:17)−1 , λ ∈ D, COUPLING AND RELAXED COMMUTANT LIFTING 9 It will be convenient first to prove a lemma. Let Γ be a contraction from M into E1 ⊕ M. Partition Γ as a 2 × 1 operator matrix, as follows (2.9) Γ = (cid:20) Γ1 Γ2 (cid:21) : M → (cid:20) E1 M (cid:21) . Furthermore, let E2 be a subspace of M, and consider the function (2.10) Ξ(λ) = Γ1(IM − λΓ2)−1Π∗ E2 , λ ∈ D, Here ΠE2 is the orthogonal projection of M onto E2. Since Γ is a contraction, the same holds true for Γ2, and hence I − λΓ2 is invertible for each λ ∈ D. Thus Ξ is well-defined on D. Next, let X be the orthogonal complement of E2 in M, and thus M = E2 ⊕ X . Then Γ also admits a 3 × 2 operator matrix representation, namely (2.11) Put (2.12) Γ =   D1 C1 D2 C2 B Z   : (cid:20) E2 X (cid:21) →   E1 E2 X .   F (λ) = (cid:20) D1 D2 (cid:21) + λ(cid:20) C1 C2 (cid:21) (IX − λZ)−1B, λ ∈ D. Again, since Γ is a contraction, the operator Z is a contraction, and hence F is well-defined on D. Lemma 2.4. Let Ξ and F be the functions defined by (2.10) and (2.12), respectively. Then F belongs to the Schur class S(E2, E1 ⊕ E2) and (2.13) Ξ(λ) = Π1F (λ)(I − Π2F (λ))−1 λ ∈ D, where Π1 and Π2 are the orthogonal projections of E1 ⊕ E2 onto E1 and E2, respec- tively. Proof. The function F is the transfer function of the system (cid:8)Z, B,(cid:20) C1 C2 (cid:21) ,(cid:20) D1 D2 (cid:21) ; X , E2, E1 ⊕ E2(cid:9). By (2.11) the system matrix corresponding to this system is equal to Γ, and hence it is a contraction. This implies that F belongs to the Schur class S(E2, E1 ⊕ E2); cf., Theorem 4.1 in [10] where this is proved for time-variant systems. To prove (2.13) fix λ ∈ D. Using the partitioning of Γ in (2.11) we see that for each e ∈ E2 we have Ξ(λ)e = (cid:2) D1 C1 (cid:3)(cid:18)I − λ(cid:20) D2 C2 B Z (cid:21)(cid:19)−1(cid:20) e 0 (cid:21) . To find Ξ(λ)e we have to compute the first column of the inverse of the 2 × 2 operator matrix (2.14) (cid:20) I − λD2 −λC2 I − λZ (cid:21) . −λB Since I − λZ is invertible, the Schur complement ∆(λ) of I − λZ in (2.14) is well- defined and is given by ∆(λ) := I − λD2 − λ2C2(I − λZ)−1B = I − λΠ2F (λ). 10 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK It follows (cf., Remark 1.2 in [11]) that (cid:20) I − λD2 −λC2 I − λZ (cid:21)−1 −λB = (cid:20) (I − λΠ2F (λ))−1 λ(I − λZ)−1B(I − λΠ2F (λ))−1 ∗ ∗ (cid:21) . Thus Ξ(λ)e = (cid:2) D1 C1 (cid:3)(cid:20) (I − λΠ2F (λ))−1e λ(I − λZ)−1B(I − λΠ2F (λ))−1e (cid:21) = (D1 − λC1(I − λZ)−1B)(I − λΠ2F (λ))−1e = Π1F (λ)(I − λΠ2F (λ))−1e. Since e is an arbitrary element of E2, this proves (2.13). Proof of Proposition 2.3. Since Y is assumed to be an isometry, Y ∗ is a con- traction. Now apply Lemma 2.4 with D′ in place of E1, with Y ∗ in place of the contraction Γ in (2.9), and with D in place of E2. With these choices the function Ξ in (2.10) coincides with the function defined by the left hand side of (2.8). Thus in order to finish the proof it remains to show that with Γ = Y ∗, E1 = D′, and E2 = D the function F in (2.12) is also given by (2.7). But this follows by applying to F in place of G that the function G in (2.1) is also given by (2.3). Indeed, since F is the transfer function of the system (cid:3) (cid:8)Z, B,(cid:20) C1 C2 (cid:21) ,(cid:20) D1 D2 (cid:21) ; X , D, D′ ⊕ D(cid:9), and the system matrix of this system is equal to Y ∗, the equivalence between (2.1) and (2.3) yields in a straightforward way that F in (2.12) is also given by (2.7). (cid:3) 3. Isometric couplings Throughout this section {T ′, A} is a pair of contractions, T ′ on a Hilbert space H′ and A from a Hilbert space H to H′. An isometric coupling of {T ′, A} is a pair {U on K, τ} of operators such that U is an isometric lifting of T ′, acting on K (and thus H′ ⊂ K), and τ is an isometry from H to K with ΠH′ τ = A. If the space K is of no interest, then we will just write {U, τ}. An isometric coupling {U on K, τ} of {T ′, A} is called minimal if, in addition, the space H′ ∨ τ H is cyclic for U , that is, K = ∞ _n=0 U n(H′ ∨ τ H). There exist minimal isometric couplings of {T ′, A}. To see this, let U be the operator on H′ ⊕ H 2(DT ′ ) ⊕ H 2(DA) given by the following operator matrix rep- resentation T ′ 0 EDT ′ SDT ′ 0 0 SDA 0 0   . U =   Here EDT ′ is the canonical embedding of DT ′ onto the space of constant functions of H 2(DT ′ ), and SDT ′ and SDA are the unilateral shifts on H 2(DT ′ ) and H 2(DA), respectively. Notice that the operator defined by the 2 × 2 operator matrix in the left upper corner of the matrix for U is the Sz.-Nagy-Schaffer minimal isometric COUPLING AND RELAXED COMMUTANT LIFTING 11 lifting of T ′. Since SDA is an isometry, we conclude that U is also an isometric lifting of T ′. Now let τ be the isometry defined by τ =   A 0 EDA DA   : H →   H′ H 2(DT ′ ) H 2(DA)   where EDA is the canonical embedding of DA onto the space of constant functions of H 2(DA). Then {U, τ} is a minimal isometric coupling of {T ′, A}. Two isometric couplings {U1 on K1, τ1} and {U2 on K2, τ2} of {T ′, A} are said to be isomorphic if there exists a unitary operator Ψ from K1 to K2 such that ΨU1 = U2Ψ, Ψτ1 = τ2 and Ψh = h for all h ∈ H′. In this case (3.1) ΨU1τ1 = U2τ2. Minimality is preserved under isomorphic equivalence. Indeed, when the pairs {U1 on K1, τ1} and {U2 on K2, τ2} are isomorphic isometric couplings of {T ′, A}, and Ψ from K1 to K2 is an isomorphism between the two isometric couplings, then ∞ _n=0 U n 2 (H′ ∨ τ2H) = ∞ ∞ (ΨU1Ψ∗)n(H′ ∨ Ψτ1H) = _n=0 1 ((Ψ∗H′) ∨ Ψ∗Ψτ1H) = Ψ U n _n=0 _n=0 ∞ = Ψ ΨU n 1 Ψ∗(H′ ∨ Ψτ1H) ∞ _n=0 U n 1 (H′ ∨ τ1H). We say that an isometric coupling {U on K, τ} of {T ′, A} is special if K is a Hilbert direct sum of the space H′, the space DA and some Hilbert space X , that is, K = H′ ⊕ DA ⊕ X , and the action of τ is given by τ h = Ah ⊕ DAh ⊕ 0, where 0 is the zero vector in X . In other words, an isometric coupling {U on K, τ} of {T ′, A} is special if, in addition, DA is a subspace of M, where M = K ⊖ H′, and τ admits a matrix representation of the form τ = (cid:20) A DA DA (cid:21) : H → (cid:20) H′ M (cid:21) . Π∗ The importance of special isometric couplings follows from Theorem 3.4 below. To prove this theorem we need a few auxiliary propositions. The first also settles the question of existence of special isometric couplings. Proposition 3.1. Every isometric coupling is isomorphic to a special isometric coupling. Proof. Let {U on K, τ} be an isometric coupling of {T ′, A}, and put M = K ⊖ H′. Since τ is an isometry and ΠH′ τ = A, the operator τ admits a matrix representation of the form: (3.2) τ = (cid:20) A ΓDA (cid:21) : H → (cid:20) H′ M (cid:21) where Γ : DA → M is an isometry; see Section IV.1 of [13] or Section XXVII.5 of [17]. Now let D = Im Γ, and put X = M ⊖ D. Then D is closed, and we can view Γ as a unitary operator from DA 12 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK onto D. Define the unitary operator σ by σ =   IH′ 0 0 0 Γ 0 0 0 IX   :   H′ DA X   →   H′ D X   . Also define U0 = σ∗U σ and τ0 = σ∗τ . Then {U0, τ0} is a special isometric coupling of {T ′, A} which is isomorphic to {U, τ}. (cid:3) Since minimality of isometric couplings is preserved under isomorphisms, and isometric couplings do exist (see the third paragraph of this section), the above proposition shows that any {T ′, A} admits a special minimal isometric coupling. Recall that an isometric lifting U of T ′ can always be represented (see (1.2)) in the following form: (3.3) U = (cid:20) T ′ Y1DT ′ Y2 (cid:21) on (cid:20) H′ 0 M (cid:21) where Y = (cid:2) Y1 Y2 (cid:3) : (cid:20) DT ′ M (cid:21) → M is an isometry. According to (1.3) this U also admits a matrix representation of the form: U = (cid:20) U ′ 0 0 U (cid:21) on (cid:20) K′ K (cid:21) where K′ = ∞ _n=0 U nH′, K = K ⊖ K′. Here U ′ on K′ is the minimal isometric lifting of T ′ associated with U (see Section 1), and U is an isometry on K. We can now state the next proposition. Proposition 3.2. Let {U, τ } be an isometric coupling of {T ′, A}, where U is de- termined by (3.3) and τ by (3.2). Set D = Im Γ, where Γ is given by (3.2), and for Y in (3.3) consider the following operator matrix representation: (3.4) Y = (cid:20) D C B Z (cid:21) : (cid:20) DT ′ ⊕ D X (cid:21) → (cid:20) D X (cid:21) where X = M ⊖ D. Then {U, τ } is a minimal isometric coupling of {T ′, A} if and only if the pair {Z, B} is controllable. Proof. Since τ is given by (3.2), the space H′ ⊕ τ H is equal to H′ ⊕ D. Thus we have to show that H′ ⊕D is cyclic for U if and only if the pair {Z, B} is controllable. To do this we associate with U two auxiliary operators, namely U = (cid:20) 0 Y1DT ′ Y2 (cid:21) on (cid:20) H′ M (cid:21) , M = (cid:20) 0 Y1 Y2 (cid:21) on (cid:20) DT ′ M (cid:21) . 0 0 Notice that the range of U − U belongs to H′. Since H′ ⊂ H′ ⊕ D, this implies that H′ ⊕ D is cyclic for U if and only if H′ ⊕ D is cyclic for U . By induction one proves that for n = 1, 2, 3, . . . we have U n = (cid:20) 0 0 Y n−1 Y1DT ′ Y n 2 2 (cid:21) on (cid:20) H′ M (cid:21) , M = (cid:20) 0 Y n−1 2 0 Y1 Y n 2 (cid:21) on (cid:20) DT ′ M (cid:21) . On the other hand (cid:2) Y n−1 2 Y1DT ′ Y n 2 (cid:3)(cid:20) H′ D (cid:21) = (cid:2) Y n−1 2 Y1 Y n 2 (cid:3)(cid:20) DT ′ D (cid:21), n = 1, 2, 3, . . . , and hence H′ ⊕ D is cyclic for U if and only if DT ′ ⊕ D is cyclic for M . COUPLING AND RELAXED COMMUTANT LIFTING 13 It remains to prove that DT ′ ⊕ D is cyclic for M if and only if the pair {Z, B} is controllable. Using Y = (cid:2) Y1 Y2 (cid:3) and (3.4), we see that M = (cid:20) D C B Z (cid:21) , DD and C = Π∗ where D = Π∗ DC with ΠD equal to the orthogonal projection of DT ′ ⊕ D onto D. By employing Lemma 2.1 with U = Y = DT ′ ⊕ D, we see that DT ′ ⊕D is cyclic for M if and only if the pair {Z, B} is controllable, which completes the proof. (cid:3) Proposition 3.3. Let {U1 on K1, τ1} and {U2 on K2, τ2} be special isometric cou- plings of {T ′, A}. For j = 1, 2 set Xj = Kj ⊖ (H′ ⊕ DA), and let Y (j) be the isometry from DT ′ ⊕ DA ⊕ Xj into DA ⊕ Xj corresponding to Uj via (3.3). Consider the following operator matrix representation Y (j) = (cid:20) Dj Cj Bj Zj (cid:21) : (cid:20) DT ′ ⊕ DA Xj Xj (cid:21) (cid:21) → (cid:20) DA for j = 1, 2. Then {U1, τ1} and {U2, τ2} are isomorphic if and only if {Z1, B1, C1, D1} and {Z2, B2, C2, D2} are unitarily equivalent realizations. Proof. Assume that {Z1, B1, C1, D1} and {Z2, B2, C2, D2} are unitarily equivalent, that is, D1 = D2 and there exists a unitary operator W from X1 onto X2 such that (3.5) W Z1 = Z2W, W B1 = B2 and C1 = C2W. Now let Φ be the unitary operator from K1 onto K2 defined by Then Φh = h for all h in H′. Because {U1, τ1} and {U2, τ2} are special, we see that Φ =   IH′ 0 0 τj =   A DA 0 0 IDA 0 0 0 W   :   : H →   H′ DA Xj     H′ DA X1   →   H′ DA X2   . for j = 1, 2. (3.6) (3.7) (3.8) Hence Φτ1 = τ2. Using the appropriate operator matrix decomposition we arrive at IH′ 0 0 ΦU1 =   =   D1Π∗ W B1Π∗ T ′ 0 DT ′ DT ′ D1Π∗ DT ′ DT ′ B1Π∗ 0 DA C1 DA Z1   0 IDA D1Π∗ B1Π∗ 0 0 0 W   T ′ 0 0 D1Π∗ DT ′ DT ′ C1 DA DT ′ DT ′ W B1Π∗ DA W Z1   . A similar calculation shows that U2Φ =   =   (3.9) T ′ 0 DT ′ DT ′ D2Π∗ DT ′ DT ′ B2Π∗ T ′ 0 DT ′ DT ′ D2Π∗ DT ′ DT ′ B2Π∗ 0 DA C2 DA Z2 0   DA C2W DA Z2W D2Π∗ B2Π∗ D2Π∗ B2Π∗ 0 IDA 0 0 0 W IH′ 0 0     .     14 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK Because D1 = D2 and (3.5) holds, we see that ΦU1 = U2Φ. In other words, {U1, τ1} and {U2, τ2} are isomorphic. Conversely assume that {U1, τ1} and {U2, τ2} are isomorphic. Then there exists a unitary operator Φ from K1 onto K2 such that Φh = h for all h in H′ and Φτ1 = τ2 and ΦU1 = U2Φ. Because τ1 and τ2 admit matrix representations of the form presented in (3.7) and Im DA is dense in DA, we see that Φh = h for all h ∈ DA. So Φ admits a matrix representation as in (3.6) where W is a unitary operator from X1 onto X2. By combining ΦU1 = U2Φ with the matrix representations for ΦU1 in (3.8) and U2Φ in (3.9), we see that D1 = D2, W Z1 = Z2W, W B1 = B2 and C1 = C2W. Hence {Z1, B1, C1, D1} and {Z2, B2, C2, D2} are unitarily equivalent realizations. (cid:3) Theorem 3.4. Let {T ′, A} be a pair of contractions, T ′ acting on H′ and A from H into H′. Then there is a one to one map from the set of minimal isometric couplings of {T ′, A}, with isomorphic ones being identified, onto the Schur class S(DA, DT ′ ⊕DA). This map is defined as follows. Let {U, τ } be a minimal isometric coupling of {T ′, A}, which may be assumed to be special, by Proposition 3.1. Define (3.10) F{U,τ }(λ) = ΠDT ′ ⊕DA Y ∗(IM − λJ ′ X Y ∗)−1Π∗ DA , where Y is the isometry uniquely determined by U via (3.3), X = M ⊖ DA, and J ′ X is the partial isometry from DT ′ ⊕ DA ⊕ X to DA ⊕ X given by J ′ X = (cid:20) 0 0 0 IX (cid:21) : (cid:20) DT ′ ⊕ DA X (cid:21) → (cid:20) DA X (cid:21) . Then {U, τ } 7→ F{U,τ } is the desired map. Proof. We know from Proposition 3.1 that every isometric coupling is isomorphic to a special one. So without loss of generality we can assume the isometric couplings to be special. From Proposition 3.2 and Section 1 it is clear that there is a one to one cor- respondence between the special minimal isometric couplings of {T ′, A} and the isometries Y mapping the space DT ′ ⊕ DA ⊕ X into DA ⊕ X , where X is some Hilbert space and the pair {ΠX Y Π∗ DT ′ ⊕DA} is controllable. In fact, this one to one correspondence is provided by (3.3). Furthermore, formula (3.4) estab- lishes a one to one correspondence between the isometries Y mapping the space DT ′ ⊕ DA ⊕ X into DA ⊕ X and the isometric realizations X , ΠX Y Π∗ {Z, B, C, D; X , DT ′ ⊕ DA, DA}, X and B = ΠX Y Π∗ and in this one to one correspondence Z = ΠX Y Π∗ DT ′ ⊕DA. From Theorem 2.2 we know that there is a one to one correspondence between the controllable isometric realizations, with the unitarily equivalent ones being iden- tified, and the S(DT ′ ⊕ DA, DA) Schur class functions. Next, note that the map G 7→ F , where F (λ) = G(λ)∗, is a one to one map from S(DT ′ ⊕ DA, DA) onto S(DA, DT ′ ⊕ DA). Following up all these one to one correspondences and using the results of Section 2 we see that the map from a special minimal isometric coupling {U, τ } to F is given by F = F{U,τ }. To complete the proof, it remains to apply Proposition 3.3. (cid:3) We conclude this section with a lemma that will be useful in the next section. COUPLING AND RELAXED COMMUTANT LIFTING 15 Lemma 3.5. Let {U1 on K1, τ1} and {U2 on K2, τ2} be isomorphic isometric cou- plings of {T ′, A}, and let V on H′ ⊕ H 2(DT ′ ) be the Sz.-Nagy-Schaffer minimal isometric lifting of T ′. For j = 1, 2 let Φj be the unique isometry associated with T ′ intertwining V and Uj. Then Φ∗ 1τ1 = Φ∗ 2τ2. Proof. Let Ψ from K1 to K2 be an isomorphism from {U1, τ1} to {U2, τ2}. Define Θ from H′ ⊕ H 2(DT ′ ) into K1 by setting Θ = Ψ∗Φ2. Then Θ is an isometry, Θh = Ψ∗Φ2h = Ψ∗h = h for all h ∈ H′, and ΘV = Ψ∗Φ2V = Ψ∗U2Φ2 = U1Ψ∗Φ2 = U1Θ. So, by Theorem 1.1 (see also the last paragraph of Section 1), the operator Θ is the unique isometry associated with T ′ intertwining V and U1, that is, Θ = Φ1. It follows that Φ1 = Ψ∗Φ2, and hence Φ∗ 2τ2, which completes the proof. (cid:3) 2Ψτ1 = Φ∗ 1τ1 = Φ∗ 4. Main theorem for the case when R∗R = Q∗Q In this section {A, T ′, V, R, Q} is a lifting data set, with V on H′ ⊕H 2(DT ′ ) being the Sz.-Nagy-Schaffer minimal isometric lifting of T ′. In particular, T ′AR = AQ and R∗R ≤ Q∗Q. Recall that B from H into H′ ⊕ H 2(DT ′ ) is a contractive interpolant for {A, T ′, V, R, Q} if B is a contraction satisfying ΠH′ B = A and V BR = BQ. Our aim is to prove Theorem 0.1 assuming that R∗R = Q∗Q. First let us reformulate Theorem 0.1 for this case. For this purpose note that for R∗R = Q∗Q the spaces F and F ′ defined by (0.2) are given by F = DAQH0 and F ′ = (cid:20) DT ′ AR DAR (cid:21) H0. Observe that F ⊂ DA and F ′ ⊂ DT ′ ⊕ DA. Furthermore, the unitary operator ω mapping F onto F ′ in (0.3) is now determined by (4.1) ω(DAQh) = (cid:20) DT ′ AR DAR (cid:21) h, h ∈ H0. The following is the main result of this section. Theorem 4.1. Let {A, T ′, V, R, Q} be a lifting data set, where V on H′⊕H 2(DT ′ ) is the Sz.-Nagy-Schaffer minimal isometric lifting of T ′, and assume that R∗R = Q∗Q. Then all contractive interpolants B for {A, T ′, V, R, Q} are given by (4.2) Bh = (cid:20) ΠT ′ F (λ)(I − λΠAF (λ))−1DAh (cid:21) , Ah h ∈ H, where F is any function in S(DA, DT ′ ⊕ DA) satisfying F (0)F = ω. Here ω is the unitary operator defined in (4.1) while ΠT ′ and ΠA are the projections given by ΠT ′ = (cid:2) I 0 (cid:3) : (cid:20) DT ′ DA (cid:21) → DT ′ and ΠA = (cid:2) 0 I (cid:3) : (cid:20) DT ′ DA (cid:21) → DA. The proof of the above theorem will be based on a further refinement (which we present in two propositions) of the theory of isometric couplings presented in the previous section. In fact, to obtain contractive interpolants for {A, T ′, V, R, Q} we 16 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK shall need isometric couplings {U, τ } of {T ′, A} satisfying the additional intertwin- ing relation U τ R = τ Q. This is the contents of the first proposition (Proposition 4.2 below). The existence of such couplings is guaranteed by the second proposi- tion (Proposition 4.3 below), which is based on Theorem 3.4. In the sequel, for simplicity, we shall write V for the space H′ ⊕ H 2(DT ′ ). Proposition 4.2. Let {A, T ′, V, R, Q} be a lifting data set, with V on V being the Sz.-Nagy-Schaffer minimal isometric lifting of T ′, and assume that R∗R = Q∗Q. Let {U on K, τ } be an isometric coupling of {T ′, A} satisfying U τ R = τ Q, and let Φ be the unique isometry from V into K associated with T ′ intertwining V with U . Then (4.3) B = Φ∗τ is a contractive interpolant for {A, T ′, V, R, Q}, and all contractive interpolants for this data set are obtained in this way. More precisely, if B a contractive interpolant for {A, T ′, V, R, Q}, then there exists a minimal special isometric coupling {U, τ } of {T ′, A} such that B = Φ∗τ and U τ R = τ Q. Proof. First let us show that B defined by (4.3) is a contractive interpolant for the data set {A, T ′, V, R, Q}. Obviously, B is a contraction. Put K′ = Im Φ. Recall (see Section 1) that ΦΦ∗ is the orthogonal projection of K onto K′. From Theorem 1.1 we know that K′ = Wn≥0 U nH′ is a reducing subspace for U . It follows that U commutes with ΦΦ∗. Since Φ∗Φ is the identity operator on V = H′ ⊕ H 2(DT ′ ), we obtain V BR = Φ∗U Φ(Φ∗τ )R = Φ∗(U ΦΦ∗)τ R = Φ∗ΦΦ∗U τ R = Φ∗τ Q = BQ. Thus B is a contractive interpolant for {A, T ′, V, R, Q}. To prove the reverse implication, assume that B is a contractive interpolant. We have to construct a minimal special isometric coupling {U, τ } of {T ′, A} satisfying U τ R = τ Q such that B is given by (4.3). Since B is a contraction, we may consider the subspaces F = DBRH0 and F ′ = DBQH0. Using V BR = BQ with R∗R = Q∗Q, and the fact that V is an isometry, we see that for each h ∈ H0 we have kDBQhk2 = kQhk2 − kBQhk2 = kRhk2 − kV BRhk2 = kRhk2 − kBRhk2 = kDBRhk2. Hence there exists a unique unitary operator ω from F onto F ′ such that ωDBR = DBQ. Next, define the subspaces G = DB ⊖ F and G′ = DB ⊖ F ′. Notice that DB = F ⊕ G and DB = F ′ ⊕ G′. Thus ω defines a partial isometry Ω on DB as follows: Ω = (cid:20) ω 0 0 0 (cid:21) : (cid:20) F G (cid:21) → (cid:20) F ′ G′ (cid:21) . Observe that DΩ coincides with G. Define VΩ to be the Sz.-Nagy-Schaffer minimal isometric lifting of Ω on VΩ = DB ⊕ H 2( G). Thus VΩ has the following operator COUPLING AND RELAXED COMMUTANT LIFTING 17 matrix representation VΩ =   0 0 0 ω 0 0 0 E G S G   :   F G H 2( G)   →   F ′ G′ H 2( G) .   Here E G is the canonical embedding of G onto the space of constant functions in H 2( G), and S G is the unilateral shift on the Hardy space H 2( G). Since VΩ is a minimal isometric lifting of Ω, we have VΩ = DB ⊕ H 2( G) = W∞ Ω DB. Now, n=0 V n put (4.4) (4.5) UΩ = (cid:20) V τΩ = (cid:20) VΩ (cid:21) , 0 VΩ (cid:21) on (cid:20) V DB DB (cid:21) : H → (cid:20) V Π∗ VΩ (cid:21) . 0 B Since V on V = H′ ⊕ H 2(DT ′ ) is an isometric lifting of T ′, the operator UΩ is an isometric lifting of T ′, and A = ΠH′ τΩ. It follows that {UΩ, τΩ} is an isometric coupling of {T ′, A}. Notice that V ∨ τΩH = V ⊕ DB. Because H′ is cyclic for V and DB is cyclic for VΩ, the reducing decomposition of UΩ in (4.4) shows that H′ ∨ τΩH is cyclic for UΩ. In other words, the isometric coupling {UΩ, τΩ} is minimal. Since V BR = BQ, the construction of UΩ and τΩ implies that (4.6) UΩτΩR = τΩQ and B = ΠV τΩ. Indeed, for h ∈ H0 we have UΩτΩRh = (cid:20) V 0 VΩ (cid:21)(cid:20) BRh DBRh (cid:21) = (cid:20) V BRh VΩDBRh (cid:21) . 0 However, DBRH0 ⊂ F , and hence VΩDBRh = ωDBRh = DBQh, which follows from the definition of ω. Since, by assumption, V BRh = BQh, we see that UΩτΩRh = (cid:20) BQh DBQh (cid:21) = τΩQh, which proves the first identity in (4.6). The second is clear from the definition of τΩ. From the construction of UΩ it follows that the unique isometry ΦΩ associated with T ′ that intertwines V with UΩ is equal to Π∗ V , where ΠV is the orthogonal projection of V ⊕ VΩ onto V. This together with the second identity in (4.6) yields (4.7) B = ΠV τΩ = Φ∗ ΩτΩ. By Proposition 3.1 and the fact that minimality of isometric couplings is pre- served under isomorphisms, there exists a minimal special isometric coupling {U, τ } of {T ′, A} which is isomorphic to {UΩ, τΩ}. Using Lemma 3.5 and formula (4.7) we obtain B = Φ∗τ , where Φ is the unique isometry associated with T ′ that intertwines V with U . It remains to prove that U τ R = τ Q. Let Ψ be the isomorphism that trans- forms {UΩ, τΩ} into {U, τ }. In particular, ΨτΩ = τ . Moreover formula (3.1) yields ΨUΩτΩ = U τ . Since UΩτΩR = τΩQ, it follows that U τ R = ΨUΩτΩR = ΨτΩQ = τ Q. (cid:3) 18 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK Proposition 4.3. Let {A, T ′, V, R, Q} be a lifting data set. Assume that R∗R = Q∗Q, and let ω be the unitary operator defined in (4.1). Consider a minimal special isometric coupling {U, τ } of {T ′, A}, and let F{U,τ } be the function in the Schur class S(DA, DT ′ ⊕ DA) defined by (3.10). Then U τ R = τ Q if and only F{U,τ }(0)F = ω. In particular, there exists a special isometric coupling {U, τ } of {T ′, A} satisfying U τ R = τ Q. It will be convenient first to prove the following lemma. Lemma 4.4. Let {A, T ′, U ′, R, Q} be a lifting data set satisfying R∗R = Q∗Q. Let {U, τ } be a special isometric coupling of {T ′, A}, and consider its operator matrix representation of the form (4.8) U = (cid:20) T ′ Y1DT ′ Y2 (cid:21) on (cid:20) H′ 0 M (cid:21) where Y = (cid:2) Y1 Y2 (cid:3) : (cid:20) DT ′ M (cid:21) → M is an isometry. Then U τ R = τ Q if and only if Y F ′ = ω∗. Proof. Since the coupling is special, the space DA is a subspace of M and τ = (cid:20) A DA DA (cid:21) : H → (cid:20) H′ M (cid:21) . Π∗ It follows that for h in H0, we have T ′ARh Y1DT ′ ARh + Y2DARh (cid:21) U τ Rh = (cid:20) =   Thus T ′ 0 AQh DARh (cid:21) = (cid:20) Y1DT ′ Y2 (cid:21)(cid:20) ARh DARh (cid:21)   . DAR (cid:21) h = DAQh, Y (cid:20) DT ′ ARh U τ R = τ Q ⇐⇒ Y (cid:20) DT ′ AR h ∈ H0. Since Y is an isometry, we see that U τ R = τ Q if and only if Y F ′ is a unitary operator from F ′ onto F with the same action as ω∗. Because of the uniqueness of ω, this proves the lemma. (cid:3) Proof of Proposition 4.3. Let Y be the isometry determined by the operator matrix representation for U in (4.8), and set F = F{U,τ }. From Lemma 4.4 we know that U τ R = τ Q if and only if Y F ′ = ω∗. Thus we have to show that F (0)F = ω ⇐⇒ Y F ′ = ω∗. (4.9) By consulting (3.10) we see that F (0) = ΠDT ′ ⊕DA Y ∗Π∗ DA. The fact that F ′ ⊂ DT ′ ⊕ DA allows us to view ω as an isometry from F into DT ′ ⊕ DA. Since F ⊂ DA, it follows that the first condition in (4.9) is equivalent to Y ∗F = (cid:20) ω γ (cid:21) : F → (cid:20) DT ′ ⊕ DA X (cid:21) , where γ is some operator from F into X . However, ω is an isometry and Y ∗F is a contraction. This implies that γ = 0. We conclude that the first condition in (4.9) is equivalent to Y ∗F = ω. By taking adjoints, and using that ω is a unitary operator from F onto F ′, we see that the same holds true for the second condition in (4.9). COUPLING AND RELAXED COMMUTANT LIFTING 19 Finally, for each λ ∈ D define the operator F (λ) from DA into DT ′ ⊕DA by setting F (λ)d = ωΠF d for d ∈ DA. Here ΠF is the orthogonal projection of DA onto F . Then F belongs to the Schur class S(DA, DT ′ ⊕ DA). Hence, by Theorem 3.4, there exists a minimal special isometric coupling {U, τ } of {T ′, A} such F = F{U,τ }. Since F (0)F = ω, we conclude that U τ R = τ Q. (cid:3) Proof of Theorem 4.1. We split the proof into two parts. Part 1. Assume that B is a contractive interpolant for the data set {A, T ′, V, R, Q}. Since R∗R = Q∗Q, we know from Proposition 4.2 that there exists a (minimal) special isometric coupling {U on K, τ } of {T ′, A} such that U τ R = τ Q and B = Φ∗τ , where Φ is the unique isometry from V = H′ ⊕ H 2(DT ′ ) into K associated with T ′ intertwining V with U . Now write U = (cid:20) T ′ Y1DT ′ Y2 (cid:21) on (cid:20) H′ M (cid:21) , 0 where M = K ⊖ H′. Since {U, τ } is special, we have DA ⊂ M, and τ = (cid:20) A DA DA (cid:21) : H → (cid:20) H′ M (cid:21) . Π∗ The identity B = Φ∗τ and the formula for Φ in Theorem 1.1 show that B = A ⊕ Λ∗Π∗ DA DA, where (Λ∗m)(λ) = Y ∗ 1 (I − λY ∗ 2 )−1, m ∈ M and λ ∈ D. It follows that Bh = (cid:20) Y ∗ 1 (I − λY ∗ Ah 2 )−1Π∗ DA DAh (cid:21) , h ∈ H. To obtain the expression for B given in (4.2) we apply Proposition 2.3 with D = DA and D′ = DT ′ . It follows that (4.2) holds with F ∈ S(DA, DT ′ ⊕ DA) given by F (λ) = ΠDT ′ ⊕DAY ∗(IM − λJ ′ X Y ∗)−1, λ ∈ D. Here X = M ⊖ DA, the operator ΠDT ′ ⊕DA is the orthogonal projection of DT ′ ⊕ M onto DT ′ ⊕ DA, and J ′ X = (cid:2) 0 PX (cid:3) : (cid:20) DT ′ M (cid:21) → M, Y = (cid:2) Y1 Y2 (cid:3) : (cid:20) DT ′ M (cid:21) → M. In other words, using the terminology introduced in Theorem 3.4, we have F = F{U,τ }. Since U τ R = τ Q, Proposition 4.3 shows that F (0)F = ω, which completes the first part of the proof. Part 2. Let F be any function in S(DA, DT ′ ⊕ DA) satisfying F (0)F = ω. We have to show that B defined by (4.2) is a contractive interpolant for the given data set. According to Theorem 3.4 there is a minimal special isometric coupling {U, τ } of {T ′, A} such that F = F{U,τ }, where F{U,τ } is defined by (3.10). The fact that F (0)F = ω yields U τ R = τ Q, by Proposition 4.3. Since B is given by (4.2), we can use Proposition 2.3 (with D = DA and D′ = DT ′ ) and Theorem 1.1 to show that B = Φ∗τ , where Φ is the unique isom- etry associated with T ′ intertwining V (the Sz.-Nagy-Schaffer minimal isometric lifting of T ′) with U . This allows us to apply Proposition 4.2 to show that B is a contractive interpolant. (cid:3) 20 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK 5. Proof of the first main theorem In this section we shall prove Theorem 0.1. The proof will be based on the analogous result for the case when R∗R = Q∗Q, which was proved in the preceding section, and on Proposition 5.1 below, which allows us to reduce the general case to the case when R∗R = Q∗Q. Throughout this section {A, T ′, V, R, Q} is a lifting data set with V being the Sz.-Nagy-Schaffer minimal isometric lifting of T ′. As before, put D◦ = D◦H0, where D◦ is the positive square root of Q∗Q − R∗R. Introduce the following operators: D◦ (cid:21) , T ′ 0 0 0 A◦ = (cid:20) A R◦ = (cid:20) R ID◦ (cid:21) : (cid:20) H D◦ (cid:21) : H0 → (cid:20) H D◦ (cid:21) → (cid:20) H′ D◦ (cid:21) , Q◦ = (cid:20) Q D◦ (cid:21) , 0 ◦ = (cid:20) T ′ 0 (cid:21) on (cid:20) H′ D◦ (cid:21) , 0 (cid:21) : H0 → (cid:20) H   H 2(DT ′ ) H 2(D◦)   H′ D◦ . on   V = T ′ 0 EDT ′ DT ′ 0 0 0 0 ED◦ 0 0 SDT ′ 0 0 0 0 SD◦   ◦, and that the quintet ◦, V , R◦, Q◦} Here EDT ′ and ED◦ are the canonical embeddings of DT ′ and D◦ onto the spaces of constant functions of H 2(DT ′ ) and H 2(D◦), respectively, and SDT ′ and SDA are the forward shifts on H 2(DT ′ ) and H 2(DA), respectively. Identifying H 2(DT ′ ⊕D◦) with H 2(DT ′ ) ⊕ H 2(D◦) it is straightforward to check that V is the Sz.-Nagy-Schaffer minimal isometric lifting of T ′ (5.1) is a lifting data set satisfying R∗ Proposition 5.1. If B from H⊕D◦ to H′ ⊕D◦⊕H 2(DT ′ )⊕H 2(D◦) is a contractive interpolant for the data set (5.1), then the operator B from H to H′ ⊕ H 2(DT ′ ), defined by {A◦, T ′ ◦R◦ = Q∗ ◦Q◦. (5.2) B = ΠH′⊕H 2(DT ′ ) BΠ∗ H, is a contractive interpolant for the data set {A, T, V, R, Q}, and all contractive interpolants for {A, T, V, R, Q} are obtained in this way. Proof. Let B be a contractive interpolant for the data set (5.1). Then B is of the following form (5.3) where B = A 0 H′ D◦ Γ1DA Γ2DA : (cid:20) H 0 ID◦ 0 0 D◦ (cid:21) →   Γ2 (cid:21) : DA → (cid:20) H 2(DT ′ ) (cid:20) Γ1     H 2(D◦) (cid:21) is a contraction. H 2(DT ′ ) H 2(D◦)   , Moreover, V BR◦ = BQ◦. Now, using this B, let B be the operator defined by (5.2). In other words B = (cid:20) A Γ1DA (cid:21) : H → (cid:20) H′ H 2(DT ′ ) (cid:21) . COUPLING AND RELAXED COMMUTANT LIFTING 21 By virtue of V BR◦ = BQ◦ it follows that (cid:20) T ′ EDT ′ DT ′ SDT ′ (cid:21)(cid:20) AR Γ1DAR (cid:21) = (cid:20) AQ Γ1DAQ (cid:21) . 0 Thus B is a contraction, A = ΠH′ B and V BR = BQ, that is, B is a contractive interpolant for the data set {A, T ′, V, R, Q}. Next, let B from H to V = H′ ⊕ H 2(DT ′ ) be an arbitrary contractive interpolant for the data set {A, T ′, V, R, Q}. We have to show that B is given by (5.2), where B is some contractive interpolant for the data set (5.1). In fact, from (5.3) we see that it suffices to find a contraction Γ from DB into H 2(D◦) such that the operator B, given by (5.4) B =   B 0 ΓDB 0 ID◦ 0   : (cid:20) H D◦ (cid:21) →   V D◦ H 2(D◦) ,   satisfies the intertwining relation W BR◦ = BQ◦, where W is the operator which one obtains by interchanging the second and the third column and the second and third row in the operator matrix for V . Put D◦ (cid:21) → (cid:20) V (5.5) B◦ = (cid:20) B 0 (cid:21) on (cid:20) V ID◦ (cid:21) : (cid:20) H V◦ = (cid:20) V D◦ (cid:21) . D◦ (cid:21) , 0 0 0 0 Since V BR = BQ, we have V◦B◦R◦ = B◦Q◦. Now, notice that B◦ = B ⊕ ID◦ is a contraction. Furthermore, V◦ is a partial isometry, and the Sz.-Nagy-Schaffer minimal isometric lifting of V◦ is equal to W . Thus (cid:8)B◦, V◦, W , R◦, Q◦(cid:9) is a lifting data set. Since R∗ set (5.6) has a contractive interpolant B. By identifying the spaces ◦Q◦, we know from Theorem 4.1 that the data ◦R◦ = Q∗ (5.6) H′ ⊕ D◦ ⊕ H 2(DT ′ ) ⊕ H 2(D◦) and H′ ⊕ H 2(DT ′ ) ⊕ D◦ ⊕ H 2(D◦). one sees that this operator B is also a contractive interpolant for the data set (5.1), and from (5.4) it follows that with this choice of B the identity (5.2) holds. (cid:3) Proof of Theorem 0.1. We split the proof into two parts. Part 1. Let B be a contractive interpolant for the data set {A, T ′, V, R, Q}. Then B is of the form (5.2) for some contractive interpolant B for {A◦, T ′ ◦, V , R◦, Q◦}. ◦Q◦, we can use Theorem 4.1 to find a formula for B. To write Since R∗ this formula, we need the subspaces ◦R◦ = Q∗ F◦ = DA◦ Q◦H0 and F ′ ◦ = (cid:20) DT ′ DA◦ R◦ ◦ A◦R◦ and the unitary operator ω◦ from F◦ onto F ′ ω◦DA◦Q◦ = (cid:20) DT ′ ◦ given by ◦ A◦R◦ DA◦ R◦ (cid:21) . In this setting, (5.7) DA◦ = (cid:20) DA 0 0 (cid:21) 0 and DT ′ ◦ = (cid:20) DT ′ 0 (cid:21) H0, 0 ID◦ (cid:21) . 22 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK A straightforward computation shows that DA◦ Q◦ = (cid:20) DAQ 0 (cid:21) : H0 → (cid:20) H D◦ (cid:21) and ◦A◦R◦ (cid:20) DT ′ DA◦ R◦ (cid:21) =   DT ′ AR D◦ DAR 0   : H0 →   H′ D◦ H D◦ .   By interchanging in the last column the first two coordinate spaces and identifying the vector x ⊕ 0 with the vector x, we see that (5.8) F◦ = F , F ′ ◦ = F ′ and ω◦ = ω, where the subspaces F and F ′ and the unitary operator ω are defined in Section 0. Let us now apply Theorem 4.1 to B. It follows that (5.9) Bx = (cid:20) A◦x DA◦ x (cid:21) , x ∈ (cid:20) H D◦ (cid:21) , ΠT ′ ◦ F◦(λ)(cid:0)IDA◦ − λΠA◦ F◦(λ)(cid:1)−1 where F◦ ∈ S(DA◦ , DT ′ ◦ ⊕ DA◦ ) satisfies F◦(0)F◦ = ω◦. Here ΠT ′ ◦ = (cid:2) I 0 (cid:3) : (cid:20) DT ′ DA◦ (cid:21) → DT ′ ◦ ◦ and ΠA◦ = (cid:2) 0 I (cid:3) : (cid:20) DT ′ DA◦ (cid:21) → DA◦ . ◦ From (5.7) we see that we can identify in a canonical way DA◦ with DA, and DT ′ with D◦ ⊕ DT ′ . This together with (5.8) shows that we can view F◦ as a function F from the Schur class S(DA, D◦ ⊕ DT ′ ⊕ DA) satisfying F (0)F = ω and (5.10) ◦ B(cid:20) h d0 (cid:21) =   Ah 0 ΠT ′ F (λ)(I − λΠAF (λ))−1DAh ΠD◦ F (λ)(I − λΠAF (λ))−1DAh 0 d0 0 0 ,   Here ΠT ′ and ΠA are the projections given by (0.4) and d0 (cid:21) ∈ (cid:20) H (cid:20) h D◦ (cid:21) . ΠD◦ = (cid:2) I 0 0 (cid:3) :   D◦ DT ′ DA   → D◦. Since B is obtained from B via (5.2), we conclude that B has the desired form (0.5). Part 2. The reverse implication is proved in a similar way. Indeed, assume that B is given by (0.5), where F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) satisfies F (0)F = ω. Using the identifications made in the first part of the proof, we can view F as a function ◦ ⊕ DA◦) satisfying F◦(0)F◦ = ω◦. But then we can use Theorem F◦ ∈ S(DA◦ , DT ′ 4.1 to show that B defined by (5.9) is a contractive interpolant for the data set ◦, V , R◦, Q◦}. Since B is also given by (5.10), we conclude that B and B are {A◦, T ′ related as in (5.2). Thus Proposition 5.1 implies that B is a contractive interpolant for {A, T ′, V, R, Q}. (cid:3) COUPLING AND RELAXED COMMUTANT LIFTING 23 6. Parameterization and uniqueness of solutions In this section we prove the second main theorem (Theorem 0.2). As a conse- quence of this theorem we obtain conditions on the lifting data set {A, T ′, V, R, Q} guaranteeing that the parameterization in Theorem 0.1 is proper, that is, con- ditions on {A, T ′, V, R, Q} implying that for every contractive interpolant B for {A, T ′, V, R, Q} there exists a unique F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω such that B = BF . Here BF is the contractive interpolant for {A, T ′, V, R, Q} pro- duced by the Schur class function F from S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω as in Theorem 0.1, that is, (6.1) BF h = (cid:20) ΠT ′ F (λ)(IDA − λΠAF (λ))−1DAh (cid:21) , Ah h ∈ H. We shall also present conditions on {A, T ′, V, R, Q} implying the existence of a unique interpolant for {A, T ′, V, R, Q}. To shorten the notation in this section we define V = H′ ⊕ H 2(DT ′ ) and (6.2) Also, for a given contractive interpolant B for {A, T ′, V, R, Q} we define the spaces FB and F ′ V = H′ ⊕ D◦ ⊕ H 2(DT ′ ) ⊕ H 2(D◦). B by (6.3) FB = DBQH0 and F ′ B = (cid:20) D◦ DBR (cid:21) H0. Notice that GB and G′ B in (0.6) are then given by (6.4) GB = DB ⊖ FB and G′ B = (D◦ ⊕ DB) ⊖ F ′ B. With the above notation and definitions we can reformulate Theorem 0.2 as follows. Theorem 6.1. Let {A, T ′, V, R, Q} be a lifting data set with V the Sz.-Nagy- Schaffer minimal isometric lifting of T ′, and let B be a contractive interpolant for the data set {A, T ′, V, R, Q}. Then there exists a one to one mapping from the set of all F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω such that B = BF onto the Schur class S(GB, G′ B), with GB and G′ B as in (6.4). For the proof of Theorem 6.1 it will be convenient to first prove two lemma's. Let ◦, V , R◦, Q◦} be as defined in Section 5. Given a contractive interpolant B for {A◦, T ′ the data set {A, T ′, V, R, Q}, we let B◦ and V◦ be the operators defined by (5.5). Furthermore, as in the previous section, we define W to be the operator which one obtains by interchanging the second and the third column and the second and third row in the operator matrix for V . Recall that both {A◦, T ′ ◦, V , R◦, Q◦} and {B◦, V◦, W , R◦, Q◦} are lifting data sets. From the construction of W from V we see that both W and V are minimal isometric liftings of both T ′ ◦ and V◦. Lemma 6.2. Let B be a contractive interpolant for {A, T ′, V, R, Q} and let the pair {U on K, τ} be an isometric coupling of {V◦, B◦}. Then ◦, A◦}; (i) the pair {U, τ} is an isometric coupling of {T ′ (ii) the pair {U, τ} is minimal as an isometric coupling of {V◦, B◦} if and only if {U, τ} is minimal as an isometric coupling of {T ′ ◦, A◦}; (iii) the operator U is an isometric lifting of both T ′ and T ′ ◦; moreover, ΦU,T ′ is the canonical embedding of V into K, and ΦU,T ′ ◦ v = v for all v ∈ V; (iv) the contractive interpolant B = ΠV Φ∗ U,T ′ ◦ τ Π∗ H. 24 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK Furthermore, two isometric couplings {U1, τ1} and {U2, τ2} of {V◦, B◦} are iso- morphic as isometric couplings of {V◦, B◦} if and only if they are isomorphic as isometric couplings of {T ′ ◦, A◦}. Proof. First remark that K can be decomposed as V ⊕ D◦ ⊕ M for some Hilbert space M. Relative to this direct sum decomposition the operators U and τ admit operator matrix representations of the form (6.5) U =   V 0 0 0 0 ∗ 0 0 ∗   on   V D◦ M   and τ =   B 0 ∗ 0 ID◦ 0   : (cid:20) H D◦ (cid:21) →   V D◦ M ,   where ∗ represents operators which are not specified any further. (i) Because V is an isometric lifting of T ′ and U is an isometric lifting of V , as we can see from (6.5), we obtain that U is an isometric lifting of T ′. From (6.5) we can immediately see that U also is an isometric lifting of the zero operator on D◦. Hence U is an isometric lifting of T ′ ◦. Since τ is as in (6.5) and ΠH′ B = A, we see that ΠH′⊕D◦ τ = A◦. So {U, τ} is an isometric coupling of {T ′ (ii) Assume that {U, τ} is minimal as an isometric coupling of {V◦, B◦}, and thus that the space (V ⊕ D◦) ∨ τ (H ⊕ D◦) is cyclic for U . ◦, A◦}. Notice that in general we have for every operator W on a Hilbert space L with U and Y subspaces L that ∞ _n=0 ∞ ∞ W n(U ∨ Y) = ( _n=0 W nU) ∨ ( _n=0 W nY). Applying this with U in (6.5), the fact that V on V is a minimal isometric lifting of T ′ and the fact that (V ⊕ D◦) ∨ τ (H ⊕ D◦) is cyclic for U yield ∞ K = U n((V ⊕ D◦) ∨ τ (H ⊕ D◦)) ∞ U nV) ∨ ( _n=0 ∞ U n(D◦ ∨ τ (H ⊕ D◦))) = V ∨ ( U n(D◦ ∨ τ (H ⊕ D◦))) _n=0 V V nH′) ∨ ( Π∗ ∞ = ( U nH′) ∨ ( U n(D◦ ∨ τ (H ⊕ D◦))) _n=0 U n(D◦ ∨ τ (H ⊕ D◦))) ∞ _n=0 = ( = ( _n=0 _n=0 ∞ ∞ ∞ _n=0 _n=0 _n=0 ∞ = U n((H′ ⊕ D◦) ∨ τ (H ⊕ D◦)). Hence (H′ ⊕ D◦) ∨ τ (H ⊕ D◦) is cyclic for U , and thus {U, τ} is minimal as an isometric coupling of {T ′ ◦, A◦}. Conversely, assume that {U, τ} is minimal as an isometric coupling of {T ′ ◦, A◦}. In other words, (H′ ⊕ D◦) ∨ τ (H ⊕ D◦) is cyclic for U . Note that H′ is a subspace of V and hence (H′ ⊕D◦)∨τ (H⊕D◦) is a subspace of (V ⊕D◦)∨τ (H⊕D◦). COUPLING AND RELAXED COMMUTANT LIFTING 25 This implies that (V ⊕D◦)∨τ (H⊕D◦) is cyclic for U as well. Thus {U, τ} is minimal as an isometric coupling of {V◦, B◦}. (iii) We already showed, in (i), that U is an isometric lifting of both T ′ and T ′ ◦. From (6.5) we see that V is the minimal isometric lifting of T ′ associated with U and thus, using the remark in the final paragraph of Section 1, we obtain that ΦU,T ′ = Π∗ V IV = Π∗ V . ◦ is the unique isometry associated with T ′ V ΦV,T ′ = Π∗ Because ΦU,T ′ U , we see that the isometry ΦU,T ′ ◦ Π∗ V satisfies ΦU,T ′ ◦ Π∗ U ΦU,T ′ ◦ Π∗ V = ΦU,T ′ ◦ V Π∗ V = ΦU,T ′ ◦ that intertwines V with V h = h for all h ∈ H′ and ◦ Π∗ V V. Hence ΦU,T ′ U . Thus for all v ∈ V we have ◦ Π∗ V is the unique isometry associated with T ′ that intertwines V and ΦU,T ′ ◦ v = ΦU,T ′ ◦ Π∗ V v = ΦU,T ′ v = v. (iv) In the proof of (iii) we saw that ΦU,T ′ obtain that ◦ Π∗ V = ΦU,T ′ = Π∗ V . Hence from (6.5) we B = ΠV τ Π∗ H = ΠV Φ∗ U,T ′ ◦ τ Π∗ H. It remains to prove the final statement of the lemma. For this purpose, let {U1 on K1, τ1} and {U2 on K2, τ2} be isometric couplings of {V◦, B◦}. First assume that {U1, τ1} and {U2, τ2} are isomorphic as isometric couplings of {V◦, B◦}. We can immediately see from the definition of an isomorphism and the fact that H′ ⊕ D◦ is a subspace of V ⊕ D◦, that every isomorphism from {U1, τ1} to {U2, τ2} as isometric couplings of {V◦, B◦} also is an isomorphism from {U1, τ1} to {U2, τ2} as isometric couplings of {T ′ ◦, A◦}. Hence {U1, τ1} and {U2, τ2} are isomorphic as isometric couplings of {T ′ ◦, A◦}. Conversely, assume that {U1, τ1} and {U2, τ2} are isomorphic as isometric cou- ◦, A◦} and that Ψ is an isomorphism from {U1, τ1} to {U2, τ2}. Then V h = h for each h ∈ H′. Since V is the plings of {T ′ ΨΠ∗ minimal isometric lifting of T ′ associated with U1, we obtain V is an isometry from V to K2 with ΨΠ∗ U2ΨΠ∗ V = ΨU1Π∗ V = ΨΠ∗ V V. Thus with (iii) we see that ΨΠ∗ V = ΦU2,T ′ = Π∗ V . Since Ψ is an isomorphism between isometric couplings of {T ′ ◦, A◦}, the operator Ψ is the identity on H′ ⊕ D◦. In particular, Ψd = d for each d ∈ D◦. Hence Ψ also is an isomorphism from {U1, τ1} to {U2, τ2} as isometric couplings of {V◦, B◦}. (cid:3) Lemma 6.3. Let B be a contractive interpolant for the data set {A, T ′, V, R, Q}, and let {U on K, τ} be an isometric coupling of {T ′ ◦, A◦} such that B = ΠV Φ∗ U,T ′ ◦ τ Π∗ H. Then there exists an isometric coupling { U , τ } of {T ′ such that { U , τ } also is an isometric coupling of {V◦, B◦}. ◦, A◦}, isomorphic to {U, τ}, Proof. From the remark in the last paragraph of Section 1 we can conclude that ΦU,T ′ ◦ associated with U . Since U ′ is minimal and V is the Sz.-Nagy-Schaffer minimal isometric lifting ◦ , where U ′ on K′ is the minimal isometric lifting of T ′ K′ ΦU ′,T ′ ◦ = Π∗ 26 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK ◦, we see that the unique isometry ΦU ′,T ′ of T ′ V with U ′ is unitary. Define Ψ from V ⊕ W to K, where W = K ⊖ K′, by ◦ associated with T ′ ◦ that intertwines Ψ = (cid:20) ΦU ′,T ′ 0 ◦ 0 IW (cid:21) : (cid:20) V W (cid:21) → (cid:20) K′ W (cid:21) . Then Ψ is a unitary operator with Ψx = x for all x ∈ H′ ⊕ D◦. So the operators U = Ψ∗U Ψ and τ = Ψ∗τ form an isometric coupling { U , τ } of {T ′ ◦, A◦} that is isomorphic to {U, τ}. Since ΦU ′,T ′ ◦ intertwines V and U ′, we obtain that V is the minimal isometric ◦ associated with U . Recall that V also is an isometric lifting of V◦. lifting of T ′ Hence U is an isometric lifting of V◦. Φ U ,T ′ ◦ Because V is the minimal isometric lifting of T ′ ◦ associated with U , we have . Since { U , τ } and {U, τ} are isomorphic, Lemma 3.5 implies that = Π∗ V B = ΠV Φ∗ H = ΠV τ Π∗ H. H = ΠV Π V τ Π∗ H = ΠV Φ∗ τ Π∗ τ Π∗ U,T ′ U ,T ′ ◦ ◦ Note that ΠH′⊕D◦ τ = A◦, and thus, since both τ and A◦D◦ are isometries, we get that ΠV⊕D◦ τ = B◦. Hence { U , τ } is an isometric coupling of {V◦, B◦}. (cid:3) Proof of Theorem 6.1. Let SB be the set defined by (6.6) We have to show that there exists a one to one mapping from SB onto S(GB , G′ SB = {F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) F (0)F = ω and B = BF }. B). ◦, A◦} and Proposition 4.3 to the lifting ◦, V , R◦, Q◦}, and using the identities in (5.8), we obtain that the By applying Theorem 3.4 to the pair {T ′ data set {A◦, T ′ mapping (6.7) {U, τ} 7→ F{U,τ} given by Theorem 3.4 is a one to one mapping from the set of (equivalence classes of) minimal isometric couplings {U, τ} of {T ′ ◦, A◦} satisfying U τ R◦ = τ Q◦ onto the set of all functions F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) satisfying F (0)F = ω. Moreover, ◦, V , R◦, Q◦}, we obtain from Proposition 4.2 and Proposition 5.1, applied to {A◦, T ′ that the mapping (6.7) maps the set of (equivalence classes of) minimal isometric couplings {U, τ} of {T ′ H onto the set SB defined by (6.6). ◦, A◦} satisfying U τ R◦ = τ Q◦ and B = ΠV Φ∗ τ Π∗ U,T ′ ◦ Then, using Lemma 6.2 and Lemma 6.3, we obtain that there exists a one to one mapping from SB onto the set of (equivalence classes of) minimal isometric couplings {U, τ} of {V◦, B◦} satisfying U τ R◦ = τ Q◦. Note that, because B is a contractive interpolant for {A, T ′, V, R, Q} and thus V BR = BQ, we have that {B, V, V, R, Q} is a lifting data set, and that the lifting data set {B◦, V◦, W , R◦, Q◦} is constructed from {B, V, V, R, Q} in the same way ◦, V , R◦, Q◦} from {A, T ′, V, R, Q} in Section 5. Moreover, as we constructed {A◦, T ′ since V is an isometry and thus DV = {0}, we get that FB and F ′ B in (6.3) correspond to {B, V, V, R, Q} as F and F ′ correspond to {A, T ′, V, R, Q}. Hence there exists a unique unitary operator ωB from FB to F ′ B defined by ωBDBQ = (cid:20) D◦ DBR (cid:21) . By again applying Theorem 3.4, Proposition 4.3 and the identities in (5.8), but now to the pair {V◦, B◦} and the lifting data set {B◦, V◦, W , R◦, Q◦}, we obtain that COUPLING AND RELAXED COMMUTANT LIFTING 27 there exists a one to one mapping from the set of (equivalence classes of) minimal isometric couplings {U, τ} of {B◦, V◦, W , R◦, Q◦} satisfying U τ R◦ = τ Q◦ onto the set of all functions H ∈ S(DB, D◦⊕DB) satisfying H(0)FB = ωB. Thus there exists a one to one mapping from SB onto the set of all functions H ∈ S(DB , D◦ ⊕ DB) satisfying H(0)FB = ωB. For each H ∈ S(DB, D◦ ⊕ DB) we have that H(0)FB = ωB if and only if there exists a (unique) G in S(GB , G′ B), with GB and G′ B as in (6.4), such that (6.8) H(λ) = (cid:20) ωB 0 G(λ) (cid:21) : (cid:20) FB GB (cid:21) → (cid:20) F ′ B G′ 0 B (cid:21) , λ ∈ D. Hence there exists a one to one mapping from the set SB onto S(GB , G′ B). (cid:3) In fact, in the proof of Theorem 6.1 we do not only show that there exists a one to one mapping from the set of F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω such that B = BF onto S(GB , G′ B), but we actually indicate how such a mapping can be constructed. To be more specific, the construction in the reverse way goes as follows. Assume that G is a Schur class function from S(GB, G′ B), with B some contractive interpolant for {A, T ′, V, R, Q}. Define H ∈ S(DB, D◦ ⊕ DB) by (6.8). Then H satisfies H(0)FB = ωB, and thus from Section 2 we obtain that there exists an isometry M from D◦ ⊕ DB ⊕ Y to DB ⊕ Y, for some Hilbert space Y, such that H(λ) = ΠD◦⊕DB M ∗(IDB ⊕Y − λJ ∗ Y M ∗)−1Π∗ DB , λ ∈ D, where M satisfies the controllability type condition (6.9) Y = ΠY ∞ _n=0 (JY M )n(cid:20) D◦ ⊕ DB {0} (cid:21) and M F ′ B = ω∗ B. Here JY is the partial isometry given by JY = (cid:20) 0 0 IY (cid:21) : (cid:20) DB Y (cid:21) → (cid:20) D◦ ⊕ DB Y 0 (cid:21) . Notice that because V◦ and B◦ are as in (5.5) we obtain that DV◦ = ΠD◦ , DV◦ = D◦, DB◦ = DB and DB◦ = DBΠH. Thus we can define U = (cid:20) V◦ M D◦ΠD◦ M (DB ⊕ Y) (cid:21) on (cid:20) V ⊕ D◦ DB ⊕ Y (cid:21) 0 and τ = (cid:20) B◦ DB DBΠH (cid:21) : H ⊕ D◦ → (cid:20) V ⊕ D◦ DB ⊕ Y (cid:21) . Π∗ Then { U , τ } is a special isometric coupling of {V◦, B◦}. Because M satisfies (6.9), the coupling { U , τ } is minimal and U τ R◦ = τ Q◦. Hence by Lemma 6.2 we obtain that { U , τ } also is a minimal isometric coupling of {T ′ τ Π∗ H. According to Proposition 3.1, the coupling { U , τ } is isomorphic to a special isomet- ric coupling {U on K, τ} of {T ′ ◦, A◦}. This isometric coupling {U, τ} is minimal, satisfies U τ R◦ = τ Q◦ and, by Lemma 3.5, we have that B = ΠV Φ∗ H. The isometry U defines an isometry Y from DT ′ ⊕ D◦ ⊕ DA ⊕ X to DA ⊕ X , with ◦ on H′ ⊕ D◦ instead of T ′ on H′. Then X = K ⊖ (H′ ⊕ D◦ ⊕ DA), by (1.2), with T ′ ◦, A◦} with B = ΠV Φ U ,T ′ Π∗ U,T ′ ◦ ◦ 28 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK the functions F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω satisfying B = BF corresponding to the function G ∈ S(GB, G′ B) is given by F (λ) = ΠDT ′ ⊕D◦⊕DA Y ∗(I − λJ ∗ X Y ∗)−1ΠDA , λ ∈ D, with JX being the partial isometry given by JX = (cid:20) 0 0 0 IX (cid:21) : (cid:20) DA X (cid:21) → (cid:20) DT ′ ⊕ D◦ ⊕ DA X (cid:21) . From Theorem 6.1 we immediately obtain the next corollary. Corollary 6.4. Let B be a contractive interpolant for {A, T ′, V, R, Q}. Then there is a unique F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω such that B = BF if and only if FB = DB or F ′ B = D◦ ⊕ DB. The next lemma gives some sufficient conditions on {A, T ′, V, R, Q} under which the parameterization in Theorem 0.1 is proper. To this end, define the subspace F ′ A of D◦ ⊕ DA by (6.10) F ′ A = (cid:20) D◦ DAR (cid:21) H0. Lemma 6.5. Let {A, T ′, V, R, Q} be a lifting data set. Then (i) F = DA implies that there exists a unique contractive interpolant B and that FB = DB; A = D◦ ⊕ DA implies that F ′ B = D◦ ⊕ DB for every contractive interpolant (ii) F ′ B. If either F = DA or F ′ A = D◦ ⊕ DA holds, then the mapping F → BF given by (6.1) is one to one from the set of all F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) satisfying F (0)F = ω onto the set of all contractive interpolants. Proof. Let B be a contractive interpolant for {A, T ′, V, R, Q}. Then ΠH′ B = A, hence there exists a contraction Γ from DA to H 2(DT ′ ) such that B = (cid:20) A ΓDA (cid:21) : H → (cid:20) H′ H 2(DT ′ ) (cid:21) . From this we obtain that for all h ∈ H kDBhk2 = khk2 − kBhk2 = khk2 − kAhk2 − kΓDAhk2 = kDAhk2 − kΓDAk2 = kDΓDAhk2. Note that DΓ ⊂ DA and thus DΓDAH = DΓ. Hence there exists a unitary operator γ from DΓ onto DB such that DB = γDΓDA. (i) Assume that F = DA. Then there is only one F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω. Hence, by Theorem 0.1, there can be only one contractive interpolant B for {A, T ′, V, R, Q}, and for this contractive interpolant B there can be only one F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω such that B = BF . Moreover, we have FB = DBQH0 = γDΓDAQH0 = γDΓDAQH0 = γDΓF = γDΓDA = DB. COUPLING AND RELAXED COMMUTANT LIFTING 29 (ii) Assume that F ′ A = D◦ ⊕ DA. Then we have that F ′ 0 0 DAR (cid:21) H0 B = (cid:20) D◦ = (cid:20) ID◦ = (cid:20) ID◦ DBR (cid:21) H0 = (cid:20) ID◦ γDΓ (cid:21)(cid:20) D◦ γDΓ (cid:21)(cid:20) D◦ γDΓ (cid:21)(cid:20) D◦ DAR (cid:21) H0 = (cid:20) ID◦ DA (cid:21) = (cid:20) 0 0 0 0 0 γDΓ (cid:21) F ′ DB (cid:21) . γDΓDA (cid:21) = (cid:20) D◦ D◦ A 0 The final statement of the lemma follows immediately from (i), (ii) and Corol- (cid:3) lary 6.4. For the classical commutant lifting theorem, that is, when H = H0, R = IH and Q is an isometry on H, we have already seen in Section 0 that the parameterization in Theorem 0.1 is proper. This result also follows from Lemma 6.5 (ii). Indeed, if H = H0, R = IH and Q is an isometry on H, then F ′ A = {0} ⊕ DAIHH = {0} ⊕ DA = D◦ ⊕ DA. Finally we derive some sufficient conditions on {A, T ′, V, R, Q} guaranteeing that there is only one contractive interpolant. From Lemma 6.5 we already know that the condition F = DA is such a condition. In the same way we can see that the condition F ′ = DT ′ ⊕ D◦ ⊕ DA is sufficient. For the classical commutant lifting theorem the combination of these two con- ditions is also a necessary condition. That is, if H = H0, R = IH and Q is an isometry on H, then there is only one contractive interpolant if and only if F = DA or F ′ = DT ′ ⊕ D◦ ⊕ DA. We can see this as follows. If the parametrization in Theorem 0.1 for the lifting data set {A, T ′, V, R, Q} is proper, then there is only one contractive interpolant for {A, T ′, V, R, Q} if and only if there is only one F in S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω. The latter is equivalent to the condition 'F = DA or F ′ = DT ′ ⊕ D◦ ⊕ DA'. Notice that when T ′ is an isometry, then the Sz.-Nagy-Schaffer minimal isometric lifting of T ′ is T ′ itself. So in that case there also is only one contractive interpolant B for {A, T ′, V, R, Q}, namely B = A. In the next lemma we summarize the above, and improve the condition F ′ = DT ′ ⊕ D◦ ⊕ DA a bit further. Proposition 6.6. Assume that for {A, T ′, V, R, Q} either T ′ is an isometry, F = DA or DT ′ ⊕ DA ⊂ F ′. Then there exists a unique contractive interpolant for {A, T ′, V, R, Q}. Proof. We have already seen above that the requirement T ′ is an isometry and the equality F = DA are both sufficient conditions. So assume that we have DT ′ ⊕ DA ⊂ F ′. Define for all F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) the Schur class functions F◦ = ΠD◦ F , FT ′ = ΠDT ′ F and FA = ΠDA F . Hence for all λ ∈ D F (λ) =   F◦(λ) FT ′ (λ) FA(λ)   : DA →   D◦ DT ′ DA   . Then we have (6.11) ΠT ′ F (λ)(IDA − λΠAF (λ))−1DA = FT ′ (λ)(IDA − λFA(λ))−1DA, λ ∈ D. 30 A.E. FRAZHO, S. TER HORST, AND M.A. KAASHOEK All F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω admit a matrix representation of the form F (λ) = (cid:20) ω 0 G(λ) (cid:21) : (cid:20) F 0 G (cid:21) → (cid:20) F ′ G′ (cid:21) , λ ∈ D, for some G ∈ S(G, G′) where G = DA ⊖ F and G′ = (D◦ ⊕ DT ′ ⊕ DA) ⊖ F ′. Hence, because DT ′ ⊕ DA ⊂ F ′ all F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω have identical FT ′ and FA and thus from (6.11) we see that BF is the same operator for all F ∈ S(DA, D◦ ⊕ DT ′ ⊕ DA) with F (0)F = ω. Hence by Theorem 0.1 there is only one contractive interpolant. (cid:3) Acknowledgement. We thank Ciprian Foias for useful discussions on an earlier version of this paper. Added in proof. At the IWOTA 2004 conference in Newcastle, when a preliminary version of this paper had been completed, the authors learned that W.S. Li and D. Timotin had a preprint ready in which the coupling method was also used to study the relaxed commutant lifting problem and the set of its solutions. Although the same method was used in the same area the two papers turned out to be quite complementary in style and results. We are happy that the editors of Integral Equations and Operator Theory agreed to publish the final versions of both papers, one directly after the other in this volume. References [1] V.M. Adamjan and D.Z. Arov, Scattering operators and contraction semigroups in Hilbert space, Doklady 165 (1965), 1377 -- 1380. [2] V.M. Adamjan and D.Z. Arov, On the unitary couplings of isometric operators, Mat. Issled. Kisinev 1 (1966), 3- 66 (Russian). [3] D. Alpay, V. Bolotnikov, Y. Peretz, On the tangential interpolation problem for H2 functions. Trans. Amer. Math. Soc. 347 (1995), 675 -- 686. [4] T. Ando, De Branges spaces and analytic functions, Lecture notes of the division of Applied Mathematics Research Institute of Applied Electricity, Hokkaido University, Sapporo, Japan, 1990. [5] R. Arocena, Generalized Toeplitz kernels and dilations of intertwining operators, Integral Equations and Operator Theory, 6 (1983), 759 -- 778. [6] R. Arocena, On the parameterization of Adamjan, Arov and Krein, Publ. Math. Orsay 83 (1983), 7 -- 23. [7] R. Arocena, On generalized Toeplitz kernels and their relation with a paper of Adamjan, Arov and Krein, in: Functional Analysis Homomorphy and Approximation Theory Math. Studies 86, North-Holland Amsterdam, 1984, pp. 1 -- 22. [8] R. Arocena, A theorem of Naimark, linear systems and scattering operators, J. Funct. Anal. 69 (1986), 281 -- 288. [9] R. Arocena, Unitary extensions of isometries and contractive intertwining dilations, in: The Gohberg Anniversary Collection II, OT 41, Birkauser Verlag Basel, 1989, pp. 13 -- 23. [10] D.Z. Arov, M.A. Kaashoek and D.R. Pik, Optimal time-variant systems and factorization of operators, I: minimal and optimal systems, Integral Equations and Operator Theory, 31 (1998), 389 -- 420. [11] H. Bart, I. Gohberg and M.A. Kaashoek, Minimal factorization of matrix and operator functions, OT 1, Birkhauser Verlag, Basel, 1979. [12] A. Biswas, C. Foias and A. E. Frazho, Weighted Commutant Lifting, Acta Sci. Math. (Szeged), 65 (1999), 657-686. [13] C. Foias and A. E. Frazho, The Commutant Lifting Approach to Interpolation Problems, OT 44, Birkhauser Verlag, Basel, 1990. [14] C. Foias, A.E. Frazho, and M.A. Kaashoek, Relaxation of metric constrained interpolation and a new lifting theorem, Integral Equations and Operator Theory, 42 (2002), 253 -- 310. COUPLING AND RELAXED COMMUTANT LIFTING 31 [15] A.E. Frazho and M.A. Kaashoek, A Naimark dilation perspective of Nevanlinna-Pick inter- polation, Integral Equations and Operator theory, 42 (2002), 253 -- 310. [16] B. Sz.-Nagy and C. Foias, Harmonic Analysis of Operators on Hilbert Space, North Holland Publishing Co., Amsterdam-Budapest, 1970. [17] I. Gohberg, S. Goldberg and M. Kaashoek, Classes of Linear Operators Vol.II, OT 63, Birkhauser Verlag, Basel, 1993. [18] D. Sarason, Exposed points in H 1. I, in: The Gohberg anniversary collection, Vol. II, OT 41, Birkhauser Verlag, Basel, 1989, pp. 485 -- 496. [19] B. Sz.-Nagy and C. Foias, Dilation des commutants d'op´erateurs, C. R. Acad. Sci. Paris, s´erie A, 266 (1968), 493-495. [20] S. Treil and A. Volberg, A fixed point approach to Nehari's problem and its applications, in: The Harold Widom Anniversary Volume, OT 71, Birkhauser Verlag, Basel, 1994, pp.165-186. Department of Aeronautics and Astronautics, Purdue University, West Lafayette, IN 47907, USA E-mail address: [email protected] Afdeling Wiskunde,, Faculteit der Exacte Wetenschappen, Vrije Universiteit, De Boelelaan 1081a, 1081 HV Amsterdam, The Netherlands E-mail address: [email protected] Afdeling Wiskunde, Faculteit der Exacte Wetenschappen, Vrije Universiteit, De Boelelaan 1081a, 1081 HV Amsterdam, The Netherlands E-mail address: [email protected]
1206.1244
1
1206
2012-06-06T14:33:37
New examples of K-monotone weighted Banach couples
[ "math.FA" ]
Some new examples of K-monotone couples of the type (X, X(w)), where X is a symmetric space on [0, 1] and w is a weight on [0, 1], are presented. Based on the property of the w-decomposability of a symmetric space we show that, if a weight w changes sufficiently fast, all symmetric spaces X with non-trivial Boyd indices such that the Banach couple (X, X(w)) is K-monotone belong to the class of ultrasymmetric Orlicz spaces. If, in addition, the fundamental function of X is t^{1/p} for some p \in [1, \infty], then X = L_p. At the same time a Banach couple (X, X(w)) may be K-monotone for some non-trivial w in the case when X is not ultrasymmetric. In each of the cases where X is a Lorentz, Marcinkiewicz or Orlicz space we have found conditions which guarantee that (X, X(w)) is K-monotone.
math.FA
math
New examples of K-monotone weighted Banach couples Sergey V. Astashkin∗, Lech Maligranda† and Konstantin E. Tikhomirov∗ Abstract Some new examples of K-monotone couples of the type (X, X(w)), where X is a symmetric space on [0, 1] and w is a weight on [0, 1], are presented. Based on the property of the w- decomposability of a symmetric space we show that, if a weight w changes sufficiently fast, all symmetric spaces X with non-trivial Boyd indices such that the Banach couple (X, X(w)) is K-monotone belong to the class of ultrasymmetric Orlicz spaces. If, in addition, the fundamental function of X is t1/p for some p ∈ [1,∞], then X = Lp. At the same time a Banach couple (X, X(w)) may be K-monotone for some non-trivial w in the case when X is not ultrasymmetric. In each of the cases where X is a Lorentz, Marcinkiewicz or Orlicz space we have found conditions which guarantee that (X, X(w)) is K-monotone. 2 1 0 2 n u J 6 ] . A F h t a m [ 1 v 4 4 2 1 . 6 0 2 1 : v i X r a 1 Introduction One of the fundamental problems in interpolation theory is to find a description of all interpolation spaces between two fixed Banach spaces X0 and X1, which form a Banach couple ¯X = (X0, X1), i.e., the description of all intermediate Banach spaces X with respect to ¯X such that every linear operator T : ¯X → ¯X maps X into X boundedly. between fixed Banach spaces X0 and X1, which are defined as follows: X0 + X1, and the inequality An important role in the interpolation theory is played by the K-monotone spaces if x ∈ X, y ∈ K(t, y; X0, X1) ≤ K(t, x; X0, X1) holds for all t > 0, then y ∈ X and kykX ≤ CkxkX for some constant C ≥ 1 independent of x and y. Here K(t, x; X0, X1) = inf{kx0kX0 + tkx1kX1 : x = x0 + x1, x0 ∈ X0, x1 ∈ X1} is the classical K-functional of Peetre. ∗Research partially supported by RFBR grant 10-01-00077, G. S. Magnusons found of the Royal Swedish Academy of Sciences-project number FOAMagn09-028 and Lulea University of Technology. †Research partially supported by the Swedish Research Council (VR) grant 621-2008-5058. 2010 Mathematics Subject Classification: 46E30, 46B20, 46B42 Key words and phrases: K-monotone couples, w-decomposable Banach lattices, symmetric spaces, ultrasymmetric spaces, weighted symmetric spaces, Lorentz spaces, Marcinkiewicz spaces, Orlicz spaces, regularly varying functions 1 A couple ¯X = (X0, X1) is called K-monotone (or Calder´on-Mityagin couple) if all interpolation spaces between X0 and X1 are K-monotone. By a theorem due to Brudnyı and Krugljak [9, Theorem 4.4.5] all interpolation spaces with respect to a K-monotone Banach couple (X0, X1) can be represented in the form X = (X0, X1)K Φ , where Φ is a Banach lattice of measurable functions on (0,∞) and kxk(X0,X1)K Φ = kK(·, x; X0, X1)kΦ. Moreover, even if (X0, X1) is not K-monotone, every interpolation space X with respect to (X0, X1) which happens to be a K-monotone space satisfies X = (X0, X1)K Φ for some suitable Φ, and of course this is only up to equivalence of norms ( Brudnyı and Krugljak [9, Theorem 3.3.20]). Therefore, the problem of finding new examples of K-monotone couples or K-monotone spaces becomes very important. Calder´on [10] and independently Mitjagin [26] proved that the couple (L1, L∞) is K- monotone. Several years later Sedaev and Semenov [33] proved that a weighted couple (L1(w0), L1(w1)) is K-monotone (cf. also Cwikel-Kozlov [13] for another proof) and then Sedaev [32] generalized this result to the couples of the form (Lp(w0), Lp(w1)) (1 ≤ p ≤ ∞). Finally, Sparr [35], [36] showed that (Lp(w0), Lq(w1)) is a K-monotone couple for 0 < p, q ≤ ∞. There are other proofs of Sparr's result, for example, in papers of Dmitriev [17], Cwikel [11] and of Arazy-Cwikel [2]. In [15], Cwikel and Nilsson considered the problem of K-monotonicity from a some- what different point of view. Namely, they studied the problem when a weighted Banach couple (X(w0), Y (w1)), with X, Y being separable Banach lattices with the Fatou prop- erty on a measure space (Ω, Σ, µ), is K-monotone for all weights w0, w1 on Ω. They proved that this can happen if and only if X = Lp(v0) and Y = Lq(v1) for some weights v0, v1 and some numbers 1 ≤ p, q < ∞. In their proof the concept of a decomposable Banach lattice on a measure space is essentially used. A Banach lattice X is called decomposable if for any convergent series P∞n=1 fn in X with pairwise disjoint fn (n = 1, 2, . . . ) and any (formal) series P∞n=1 gn, gn ∈ X, kgnkX ≤ kfnkX (n = 1, 2, . . . ), such that all gn are pair- wise disjoint, we have P∞n=1 gn ∈ X and kP∞n=1 gnkX ≤ C kP∞n=1 fnkX with a constant C independent of fn, gn. This notion or some variants of it were introduced earlier by Cwikel [12] and Cwikel-Nilsson [14]. Note that the problem of K-monotonicity of weighted couples (X(w0), Y (w1)) can be reduced to considering couples of the form (X, Y (w)). Therefore, in what follows, we will examine couples with one weight only. We will say that a weight w is non-trivial if either w or 1/w is unbounded. In [37], the concept of w-decomposability of a Banach lattice, which generalizes in a sense the previous one due to Cwikel, was introduced. A theorem proved in [37] states that, whenever X is a Banach lattice with the Fatou property, the couple (X, X(w)) is K-monotone if and only if X is w-decomposable (see Theorem 3.1 below in Section 3). Earlier Kalton [18] showed that in the case of symmetric sequence spaces with the Fatou property the K-monotonicity of a couple (X, Y (w)) for some non-trivial weight w implies that X = lp and Y = lq for some 1 ≤ p, q ≤ ∞ (note, however, that there exist examples of shift-invariant sequence spaces X with the Fatou property, such that (X, X(2−k)) is K-monotone but X is not isomorphic to lp for any 1 ≤ p ≤ ∞ [5], [6]). 2 Tikhomirov's theorem from [37] allows us to examine whether the result of Kalton extends to symmetric function spaces. We will see that this is not the case and the situation here will be essentially different. The paper is organized as follows. After the introduction, in Section 2, some necessary definitions and notations are collected. In the first part, we recall necessary information about symmetric spaces on [0, 1] and then, in the second part, regularly varying convex Orlicz functions on [0,∞) and regularly varying quasi-concave functions on [0, 1] are discussed. In Section 3 we consider the notion of a w-decomposable Banach lattice, which plays a central role in these investigations. Using the Krivine theorem we show that it can be essentially simplified in the case of symmetric function spaces. Namely, we prove condition (9) which means that for any w-decomposable symmetric space X there exists p ∈ [1,∞] (depending on X) such that X has, roughly speaking, both "restricted lower and upper p-estimates". In particular, its fundamental function ϕ satisfies condition (13) for some p, which means that the function ϕp is "almost additive" near zero. Section 4 contains results on the w-decomposability of Lorentz and Marcinkiewicz spaces on [0, 1]. If ϕ is a concave increasing function on [0, 1] with γϕ > 0 and 1 ≤ p < ∞, then the couple (X, X(w)) with X = Λp,ϕ([0, 1]) and a given non-trivial weight w is K- monotone if and only if condition (10) holds. This couple is K-monotone for some weight w if and only if ϕ is equivalent to a regularly varying function at 0 of order p. Moreover, for any weight w on [0, 1] we can construct a concave function ϕ on [0, 1] such that the couple (X, X(w)) with X = Λ1,ϕ([0, 1]) is K-monotone and Λ1,ϕ([0, 1]) 6= L1[0, 1]. We obtain analogous results for Marcinkiewicz spaces, as a consequence of a new duality theorem which is of independent interest. It states that under suitable mild conditions on a Banach lattice X, the weighted couple (X, X(w)) is K-monotone if and only if the couple (X′, X′(w)) is K-monotone, where X′ means the Kothe dual to X. Section 5 deals with conditions of w-decomposability of Orlicz spaces LF [0, 1]. It is shown, in Theorem 6, that if an Orlicz function F satisfies the ∆2-condition for large arguments, then LF [0, 1] is w-decomposable if and only if it satisfies some restricted p- upper and p-lower estimates (see condition (32)). Moreover, it is proved, in Theorem 7, that if an Orlicz function F is equivalent to an Orlicz function which is regularly varying at ∞ of order p ∈ [1,∞), then the Orlicz space LF = LF [0, 1] is w-decomposable for some weight w on [0, 1] and therefore the couple (LF , LF (w)) is K-monotone. Finally, in Section 6, we prove that if a symmetric space X on [0, 1] with non-trivial Boyd indices is w-decomposable with respect to a weight changing sufficiently fast, then X is an ultrasymmetric Orlicz space. The result implies that, for such a weight w, every K- monotone couple (X, X(w)) with X having the Fatou property must be an ultrasymmetric Orlicz space. Moreover, if its fundamental function is of the form ϕX(t) = t1/p for some 1 ≤ p ≤ ∞, then X = Lp. 2 Preliminaries Let us collect necessary information and results, in two parts, on symmetric (rear- rangement invariant) spaces and regularly varying functions. 3 2a. Symmetric spaces. Let (Ω, Σ, µ) be a complete σ-finite measure space and L0 = L0(Ω) be the space of all classes of µ-measurable real-valued functions defined on Ω. A Banach space X = (X,k · kX ) is said to be a Banach lattice on Ω if X is a linear subspace of L0(Ω) and satisfies the so-called ideal property, which means that if y ∈ X, x ∈ L0 and x(t) ≤ y(t) for µ-almost all t ∈ Ω, then x ∈ X and kxkX ≤ kykX. We also assume that the support of the space X is Ω (supp X = Ω), that is, there is an element x0 ∈ X such that x0(t) > 0 µ-a.e. on Ω. We will say that X has the Fatou property if 0 ≤ xn ↑ x ∈ L0 with xn ∈ X and supn∈N kxnkX < ∞ imply that x ∈ X and kxnkX ↑ kxkX . A Banach lattice X is said to be p-convex (1 ≤ p < ∞), respectively q-concave (1 ≤ q < ∞), if there is a constant C > 0 such that Xk=1 xkp)1/pkX ≤ C( Xk=1 n n k( respectively, n n kxkkp X )1/p, ( Xk=1 kxkkq X )1/q ≤ Ck( xkq)1/qkX, Xk=1 for any choice of vectors x1, x2, . . . , xn in X and any n ∈ N. If in the above definitions vec- tors x1, x2, . . . , xn ∈ X are pairwise disjoint, then X is said to satisfy an upper p-estimate and lower q-estimate, respectively. Of course, p-convexity implies upper p-estimate and q-concavity implies lower q-estimate of a Banach lattice X. More properties can be found in the book [22]. Let w be a weight on (Ω, Σ, µ), i.e., positive finite a.e. function, and let X be a Banach lattice on (Ω, Σ, µ). Then the weighted space X(w) on (Ω, Σ, µ) is defined by X(w) = {x ∈ Ω : xw ∈ X} with the norm kxkX(w) = kxwkX. In what follows, we will always suppose that the weight w is non-trivial, that is, w or 1/w is an unbounded function on (Ω, Σ, µ). C For two Banach spaces E and F the symbol E ֒→ F means that the embedding E ⊂ F is continuous with the norm which is not bigger than C, i.e., kxkF ≤ CkxkE for all x ∈ E. By a symmetric space (symmetric Banach function space), we mean a Banach lattice X = (X,k · kX) on I = [0, 1] with the Lebesgue measure m satisfying the following for any two equimeasurable functions x, y ∈ L0(I) (that is, they additional property: have the same distribution functions dx(λ) = dy(λ), where dx(λ) = m({t ∈ I : x(t) > λ}), λ ≥ 0) the condition x ∈ X implies that y ∈ X and kxkX = kykX. In particular, kxkX = kx∗kX, where x∗(t) = inf{λ > 0 : dx(λ) ≤ t}, t ≥ 0. Recall that a non-negative function ϕ : [0, 1] → [0,∞) is called quasi-concave if it is non-decreasing on [0, 1] with ϕ(0) = 0 and if ϕ(t) is non-increasing on (0, 1]. The t fundamental function ϕX of a symmetric space X on I is defined by the formula ϕX(t) = kχ[0, t]kX, t ∈ I. It is well known that every fundamental function is quasi-concave on I. Taking ϕX(t) := inf s∈(0,1)(1 + t s )ϕX(s) we obtain a concave function ϕX satisfying ϕX (t) ≤ ϕX(t) ≤ 2ϕX(t) for all t ∈ I. For any quasi-concave function ϕ on I the 4 Marcinkiewicz space Mϕ is defined by the norm kxkMϕ = sup t∈I,t>0 ϕ(t)x∗∗(t), x∗∗(t) = 1 t Z t 0 x∗(s)ds. ֒→ This is a symmetric space on I with the fundamental function ϕMϕ(t) = ϕ(t) and X MϕX . The fundamental function of a symmetric space X = (X,k · kX) is not necessarily concave but we can introduce an equivalent norm on X in such a way that the fundamental function will be concave (take kxk1 For any symmetric function space X with a concave fundamental function ϕ = ϕX there is also the smallest symmetric space with the same fundamental function. This space is the Lorentz space given by the norm X = max(kxkX ,kxkM ϕX ), x ∈ X). 1 kxkΛϕ =ZI x∗(t)dϕ(t) := ϕ(0+)kxkL∞(I) +ZI x∗(t)ϕ′(t)dt. 1 1 ֒→ X ֒→ MϕX . A non-trivial symmetric function space X C1֒→ C2֒→ L1(I), where C1 = ϕX (1), C2 = 1/ϕX(1) (see [7], Corollary 6.7 on page 78 or We have then embeddings ΛϕX on I = [0, 1] is an intermediate space between the spaces L1(I) and L∞(I) and L∞(I) X Theorem 4.1 on page 91 of [21] for a similar result when the underlying measure space is (0,∞).) The lower and upper Boyd indices αX resp. βX and the dilation indices γX resp. δX of a symmetric space X on I = [0, 1] with the fundamental function ϕX = ϕ are defined as follows: αX := lim t→0+ lnkσtkX→X ln t , βX := lim t→∞ lnkσtkX→X ln t , σtx(s) = x(s/t)χI(s/t) and γX := γϕ = lim t→0+ ln ¯ϕ(t) ln t , δX := δϕ = lim t→∞ ln ¯ϕ(t) ln t , ¯ϕ(t) = sup s,st∈I ϕ(st) ϕ(s) . We have the relations 0 ≤ αX ≤ γX ≤ δX ≤ βX ≤ 1 (see [21], pp. 101-102 and [24], p. 28). A function F : [0,∞) → [0,∞) is called an Orlicz function if it is convex and increasing with F (0) = 0. For a given Orlicz function F the Orlicz space LF = LF (I) on I = [0, 1] is defined as LF (I) = {x ∈ L0(I) : IF (cx) < ∞ for some c = c(x) > 0}, where IF (x) := RI F (x(t))dt. The Orlicz space LF is a symmetric space on I with the so-called Luxemburg-Nakano norm defined by kxkLF = inf {λ > 0 : IF (x/λ) ≤ 1} . An Orlicz function F satisfies the ∆2-condition for large u if there exist constants C ≥ 1, u0 ≥ 0 such that F (2u) ≤ CF (u) for all u ≥ u0. 5 The following notation will be used throughout the text: f ≈ g means that the functions f and g are equivalent with the constant C > 0, that is, C−1f (t) ≤ g(t) ≤ Cf (t) for all points t of the whole set on which these functions are defined, or at all points of some explicitly designated subset of that set. In the case when the constant of equivalence is not important we will write just f ≈ g. By [r] we will denote the integer part of a real number r. More information about Banach lattices and symmetric spaces can be found, for ex- C ample, in [7], [21] and [22]; about Orlicz spaces one can read e.g. in [20] and [25]. 2b. Regularly varying convex and concave functions. An Orlicz function F on [0,∞) is called regularly varying at ∞ of order p (1 ≤ p < ∞) if F (tu) F (t) lim t→∞ = up for all u > 0. (1) The following result is due to Kalton [18, Lemma 6.1]. Lemma 2.1. Let F be an Orlicz function. The following three conditions are equivalent: (a) F is equivalent to a regularly varying Orlicz function at ∞ of order p ∈ [1,∞). (b) There exists a constant C > 0 such that for any u ∈ (0, 1] we can find t0 = t0(u) with F (tu) F (t) C ≈ up for all t ≥ t0. Although we do not need it here, there is an analogous definition to the one above for Orlicz functions which are regularly varying of order p at 0 instead of at ∞ (see e.g. [18]). However, we do need to consider quasi-concave functions which are regularly varying of order p at 0. Before recalling the definition of these we should point out that it is not quite analogous to the definitions for regularly varying Orlicz functions, because the power p which appeared in (1) and in the corresponding definition in [18] will be replaced in (2) by the power 1/p. A function ϕ : [0, 1] → [0,∞) which is quasi-concave and satisfies ϕ(0) = 0 is said to be regularly varying at zero of order p (1 ≤ p ≤ ∞) if ϕ(tu) ϕ(t) lim t→0+ = u1/p for all u > 0. (2) Abakumov and Mekler [1, Theorem 5] proved that a quasi-concave function ϕ is equiv- alent to a quasi-concave regularly varying function at zero of order p ∈ [1,∞] if and only if ϕ(tu) ϕ(t) ≈ u1/p for all u > 0. lim sup t→0+ The following lemma is an immediate consequence of this result (see also the proof of Theorem 5 in [1]). 6 Lemma 2.2. A quasi-concave function ϕ on [0, 1] is equivalent to a quasi-concave function which is regularly varying at zero of order p ∈ [1,∞] if and only if for some C > 0 and any N ∈ N there exists τ (N) ∈ (0, 1] such that for all 0 < t ≤ τ (N), 0 < tN ≤ 1 we have ϕ(Nt) ϕ(t) C ≈ N 1/p. (3) Recall that the fundamental function of an Orlicz space LF on [0, 1] with the Luxemburg- F −1(1/t) for 0 < t ≤ 1 and ϕLF (0) = 0, where F −1 is the inverse Nakano norm is ϕLF (t) = of F (see formula (9.23) in [20] on page 79 of the English version or Corollary 5 in [25] on page 58). The function ϕLF is quasi-concave but not necessarily concave on [0, 1] (see [20] or [25]). 1 The notions of regularly varying Orlicz and quasi-concave functions are closely inter- related. Using Lemmas 2.1 and 2.2 and routine arguments we establish the following quantitative result showing a connection between an regularly varying Orlicz function F and the fundamental function of the corresponding Orlicz space LF . Proposition 2.3. Suppose that p ∈ [1,∞) and let F be an Orlicz function such that both F and its complementary function F ∗ satisfy the ∆2-condition for large u. Then the following conditions are equivalent: (a) There exists a constant C′ > 0 such that for any N ∈ N there exists τ (N) ∈ (0, 1] with F (u) F (uN−1/p) C ′ ≈ N for all u ≥ F −1(1/τ (N)). (4) (b) There exists a constant C > 0 such that for any N ∈ N the fundamental function ϕLF satisfies condition (3) with the same τ (N). 3 w–decomposable Banach lattices Later on C will denote a constant whose value may be different in its different appear- ances. The following notion was introduced in paper [37] and it will be very important for us. Let X be a Banach lattice on (Ω, Σ, µ) and w be a weight on Ω. We say that X is w-decomposable if there exists C > 0 such that for any n ∈ N and for all x1, . . . , xn, y1, . . . , yn in X satisfying the conditions: kxikX = kyikX , i = 1, 2, . . . , n, (5) and inf w(supp xi ∪ supp yi) ≥ 2 sup w(supp xi+1 ∪ supp yi+1), i = 1, 2, . . . , n − 1, (6) we have that k n Xi=1 xikX C ≈ k n Xi=1 yikX. 7 (7) To clarify the meaning of condition (6), consider the following example: let X be a Banach lattice of Lebesgue measurable functions on [0, 1] and w(t) = 1/t (0 < t ≤ 1). Then (6) is equivalent to the following inequality 2 sup(supp xi ∪ supp yi) ≤ inf(supp xi+1 ∪ supp yi+1), i = 1, 2, . . . , n − 1. In other words, there are some intervals [ai, bi] ⊂ [0, 1] (depending on xi, yi) such that 2bi ≤ ai+1 (i = 1, 2, . . . , n − 1), supp xi ⊂ [ai, bi] and supp yi ⊂ [ai, bi] (i = 1, 2, . . . , n). It is not hard to see that 1/t-decomposability is equivalent to 1/tq-decomposability and, more generally, w-decomposability and wq-decomposability are equivalent for any weight w and any q > 0 (see [38], Corollary 2.2 on page 61). It turns out that the w-decomposability of a Banach lattice X guarantees the K– monotonicity of the weighted couple (X, X(w)). More precisely, Tikhomirov in [37] ob- tained the following generalization of Kalton's results from [18]. Theorem 3.1. Suppose X is a Banach lattice on a σ–finite measure space (Ω, Σ, µ) with supp X = Ω which has the Fatou property and w is a (non-trivial) weight on Ω. Then the Banach couple (X, X(w)) is K–monotone if and only if X is w-decomposable. In the case of symmetric spaces on [0, 1] the notion of w-decomposability can be clarified by using the well–known Krivine theorem. Proposition 3.2. Let w be a weight on [0, 1]. A symmetric space X on [0, 1] is w- decomposable if and only if there exist C > 0 and 1 ≤ p ≤ ∞ such that for any n ∈ N and for all x1, x2, . . . , xn ∈ X satisfying the conditions inf w(supp xi) ≥ 2 sup w(supp xi+1), 1 ≤ i ≤ n − 1, we have that (8) (9) C ≈ n Xi=1 X!1/p kxikp , n (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 xi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X where, as usual, in the case p = ∞ the right hand side should be replaced by max1≤i≤n kxikX. Proof. By Krivine's theorem (see [22, Theorem 2.b.6] or [31]), there exists p ∈ [1/βX, 1/αX] such that for every m ∈ N there are pairwise disjoint equimeasurable functions y1, y2, . . . , ym ∈ X,kykkX = 1 (k = 1, 2, . . . , m), such that for any αk ∈ R (k = 1, 2, . . . , m) we have 1 2k(αk)kp ≤(cid:13)(cid:13)(cid:13) m Xk=1 αkyk(cid:13)(cid:13)(cid:13)X ≤ 2k(αk)kp. Obviously, the support of each function yk has measure not greater than 1/m. (10) Suppose that a symmetric space X is w-decomposable and that, for some n ∈ N, functions x1, . . . , xn in X satisfy condition (8). Without loss of generality we may assume that xi 6= 0 for each i = 1, 2, . . . , n. We choose m ∈ N sufficiently large so that the support of each xi has measure greater than 1/m (and so of course we also have m ≥ n). For this choice of m we consider the disjoint measurable functions y1, y2, . . . , ym,kykkX = 1 (k = 1, 2, . . . , m), obtained as it is described in the previous paragraph. In fact, we will only 8 need the first n of these functions, and we will only need special case of (10) for sequences (αk) which satisfy αk = 0 for k > n. We may assume without loss of generality, that the support of yi is contained in the support of xi for each i = 1, 2, . . . , n. (If not, since X is symmetric, we can simply replace each yi by an equimeasurable function which has this property and the above mentioned special case of (10) will remain valid.) Thus condition (8) implies that condition (6) is satisfied and therefore, applying w-decomposability (see (7)) and then the special case of (10), we obtain that αi n Xi=1 (cid:13)(cid:13)(cid:13) xi kxikX(cid:13)(cid:13)(cid:13)X ≈(cid:13)(cid:13)(cid:13) n Xi=1 αiyi(cid:13)(cid:13)(cid:13)X ≈ k(αk)n i=1kp for all choices of real numbers αi. In particular, when αi = kxikX we obtain (9). Since the reverse implication is obvious, the proof is complete. For a given weight w consider the sets Mk := {t ∈ [0, 1] : w(t) ∈ [2k, 2k+1)}, k ∈ Z. Let (wr)∞r=1 be the non-increasing rearrangement of the sequence (m(Mk))+∞k=−∞ weight w is non-trivial it follows that wr > 0 for all r = 1, 2, . . . . . Since the For some fixed n ∈ N, let x1, x2, . . . , xn be functions in X. Suppose first that these n functions satisfy condition (8). Then it is easy to see that card{i : Mk ∩ supp xi 6= ∅} ≤ 1 for each k ∈ Z. Alternatively, more or less conversely, suppose that the functions xi satisfy card{k : Mk ∩ supp xi 6= ∅} ≤ 1 for each i ∈ {1, 2, . . . , n}, i.e., for each i, there exists a unique ki ∈ Z for which supp xi ⊂ Mki. Furthermore, suppose k1 < k2 < . . . < kn. While this is not sufficient to imply that the collection of functions x1, x2, . . . , xn satisfies condition (8), it does imply that (after relabelling) the collection of functions x1, x3, x5, . . . satisfies (8) and so does the collection x2, x4, . . . . By {M r}∞r=1 we will denote any rearrangement of the sets Mk (k = 0,±1,±2, . . .) such that m(M r) = wr, r = 1, 2, . . .. Thus, by Proposition 3.2, we obtain the following result. Theorem 3.3. Suppose w is a non-trivial weight on [0, 1]. A symmetric space X on [0, 1] is w-decomposable if and only if there exist C > 0 and 1 ≤ p ≤ ∞ such that for any n ∈ N and for all x1, x2, . . . , xn ∈ X satisfying the condition supp xi ⊂ M i, 1 ≤ i ≤ n, (11) we have (9). Next, we will need some corollaries of Theorem 3.3. Firstly, using the symmetry of the norm in X, we get 9 Corollary 3.4. Let w be a non-trivial weight on [0, 1]. A symmetric space X on [0, 1] is w-decomposable if and only if there exist C > 0 and 1 ≤ p ≤ ∞ such that for any n ∈ N and for all pairwise disjoint x1, x2, . . . , xn ∈ X satisfying the condition m(supp xi) ≤ wi, 1 ≤ i ≤ n, (12) we have (9). Corollary 3.5. A symmetric space X on [0, 1] is w-decomposable for some non-trivial weight w on [0, 1] if and only if there exist C > 0, 1 ≤ p ≤ ∞, and a sequence of disjoint intervals {∆k}∞k=1 from [0, 1] such that for any n ∈ N and for all x1, x2, . . . , xn ∈ X satisfying the condition supp xi ⊂ ∆i (1 ≤ i ≤ n) we have (9). Corollary 3.6. Let w be a non-trivial weight on [0, 1] and let the sequence (wr)∞r=1 be as above. Suppose that X is a w-decomposable symmetric space X on [0, 1] with the fundamental function ϕ. Then there exist some C > 0 and p ∈ [1,∞] such that, for every sequence of reals (τr)∞r=1 satisfying 0 < τr ≤ wr (r ∈ N), we have ∞ C ≈ ∞ Xr=1 ϕp(τr)!1/p Xr=1 ϕ(cid:0) τr(cid:1) with the natural modification for p = ∞. Corollary 3.7. Let w be a non-trivial weight on [0, 1] such that a symmetric space X on [0, 1] is w-decomposable. Then there exist C > 0 and 1 ≤ p ≤ ∞ such that condition (3) is fulfilled with τ (N) = wN (N ∈ N). In particular, the fundamental function ϕ of X is equivalent to a regularly varying function at zero of order p and αX = γϕ = δϕ = βX = 1/p. (13) Proof. First we note that condition (3) is an immediate consequence of (13). Moreover, it is well known that the assertion of Krivine's theorem holds for both p = 1/αX and p = 1/βX (see [22, p. 141], [31] and [3]). Therefore, coincidence of the Boyd indices and dilation indices follows from an inspection of the proof of Proposition 3.2 and the inequalities αX ≤ γϕ ≤ δϕ ≤ βX (cf. [21, p. 102] and [24, p. 28]). Let us show that, conversely, (13) can be derived from (3) with τ (N) = wN for a large class of weights w. Theorem 3.8. Let w be a weight on [0, 1] such that qwr+1 ≤ wr (r = 1, 2, . . . ) for some q > 1 and let ϕ be a quasi–concave function on [0, 1]. Suppose there exist C > 0 and 1 ≤ p ≤ ∞ such that ϕ satisfies (3) with τ (N) = wN (N = 1, 2, . . . ). Then, for any sequence of reals (τr)∞r=1 such that 0 < τr ≤ wr (r = 1, 2, . . . ), estimate (13) holds. Proof. We present the proof for 1 ≤ p < ∞ since the case p = ∞ needs only minor changes. Firstly, it is easy to see that condition (3) can be extended as follows: we can find a (possibly different) constant C > 0 such that for every real z ≥ 1 and τ (z) := τ ([z]) we have ϕ(zt) ϕ(t) C ≈ z1/p if 0 < t ≤ τ (z). 10 (14) Let us show that for every m ∈ N there is a constant C(m) > 0 such that for all even N ∈ N satisfying the inequality N m ≤ qN/2 and all z ∈ [1, N] we have ϕ(zmt) ϕ(t) C(m) ≈ zm/p if 0 < t ≤ τ (N). (15) In fact, by the assumption, τ (N/2) ≥ qN/2τ (N), whence zkt ≤ zmt ≤ N mτ (N) ≤ qN/2τ (N) ≤ τ (N/2) ≤ 1 (k = 0, 1, . . . , m) provided that t ≤ τ (N). Therefore, using the quasi–concavity of ϕ and equivalence (14) for max(1, z/2) we obtain that ϕ(zkt) ϕ(zk−1t) ≈ ϕ(max(1, z/2)zk−1t) ϕ(zk−1t) ≈ z1/p if 0 < t ≤ τ (N), with a constant of equivalence depending on p. Multiplying these relations for all k = 1, 2, . . . , m, we come to (15). Next, let ϕ0(s) = lim sup t→0+ ϕ(ts) ϕ(t) for s > 0. Clearly, condition (14) implies ϕ0(s) ≈ s1/p(s > 0). On the other hand, in view of Boyd's result [8] (see also [24, Theorem 2.2]) ϕ0(s) ≥ sγϕ if 0 < s ≤ 1 and ϕ0(s) ≥ sδϕ if s > 1. Since γϕ ≤ δϕ it follows that γϕ = δϕ = 1 p > 0. Therefore, there exist A > 0 and κ > 0 such that (16) ϕ(st) ϕ(s) ≤ Atκ for all 0 ≤ t ≤ 1. sup 0<s≤1 Let us prove that (13) is a consequence of (15) and (16). Take a natural number m0 ≥ 2 such that κ m0 > 1 and consider an arbitrary sequence (τr)∞r=1 satisfying τr ≤ wr, r = 1, 2, . . . . Since the non-increasing rearrangement (τ∗r )∞r=1 of this sequence also satisfies τ∗r ≤ wr for r = 1, 2, . . . we can assume without loss of generality that the sequence (τr)∞r=1 is itself non-increasing. Further, set I = {r ∈ N : τr rm0 ≥ τ1}, J = N\ I. Clearly, 1 ∈ I. By (16) and the choice of m0, rm0(cid:17) ≤ A(cid:16) ϕ(τ1) ≤ C1ϕ(τ1). τr(cid:17) ≤ ϕ(cid:16) r−m0(cid:17)κ ϕ(cid:16)Xr∈J ∞ Xr=2 τ1 ∞ Xr=2 Analogously, ϕp(τr) ≤ Xr∈J ∞ Xr=2 ϕp(τ1/rm0) ≤ Ap ∞ Xr=2 r−p κ m0ϕp(τ1) ≤ C2 ϕp(τ1). Thus, it is sufficient to prove equivalence (13) for (τr)r∈I. If card I < ∞ then there is nothing to prove. So, assume that card I = ∞. Choose a positive integer i0 ∈ I, i0 ≥ 2 such that for N = 2 [i0/2] we have N m0 ≤ qN/2. Denote δr = (τr/τi0)1/m0 for r ∈ I ∩{1, 2, . . . , i0}. Then, by the definition of I, δr ≤ (τ1/τi0)1/m0 ≤ 11 i0 ≤ 2N. Applying (15) in the case m = m0, z = max(1, δr/2) for all r ∈ I, r ≤ i0, we get with a constant of equivalence depending on m0 and p. The last formula implies that ϕ(τr) ϕ(τi0) = ϕ(δm0 r τi0) ϕ(τi0) ≈ δm0/p r Xr∈I∩{1,2,...,i0} ϕp(τr) ≈ ϕp(τi0) τi0 , τr. τi0(cid:19)1/p =(cid:18) τr Xr∈I∩{1,2,...,i0} τr/τi0(cid:17)1/(m0+1) ≤ i0. we get On the other hand, setting δ :=(cid:16)Pr∈I∩{1,2,...,i0} δ ≤(cid:16) Xr∈I∩{1,2,...,i0} τ1/τi0(cid:17)1/(m0+1) Therefore, again by (15), we obtain ϕp(τi0) τi0 Xr∈I∩{1,2,...,i0} τr = δm0+1ϕp(τi0) ≈ ϕp(δm0+1τi0) = ϕp(cid:16) Xr∈I∩{1,2,...,i0} τr(cid:17), with a constant depending on m0 and p. Combining the above formulas and noting that i0 can be arbitrarily large, we conclude that equivalence (13) holds and the proof is complete. Theorem 3.8 allows us to construct non-trivial quasi–concave functions satisfying con- dition (13), for a large class of weights. For example, let w(t) = 1/t (0 < t ≤ 1). In this case wr = 2−r, r = 1, 2, . . . Define ϕ(t) = t log e t (0 < t ≤ 1). Obviously, ϕ is quasi–concave. Elementary calculations show that (3) is fulfilled for ϕ with p = 1 and τ (N) = wN = 2−N (N = 1, 2, . . . ). Thus, by Theorem 3.8, ϕ satisfies (13). 4 w–decomposable Lorentz and Marcinkiewicz spaces For 1 ≤ p < ∞ and any increasing concave function ϕ, ϕ(0) = 0, the Lorentz space Λp,ϕ consists of all classes of measurable functions x on [0, 1] such that kxkΛp,ϕ =(cid:18)Z 1 0 [x∗(t)ϕ(t)]p dt t (cid:19)1/p < ∞. The space Λp,ϕ was investigated by Sharpley [34] and Raynaud [30], who proved that if 0 < γϕ ≤ δϕ < 1, then Λp,ϕ is a symmetric space on [0, 1] with an equivalent norm Λp,ϕ =(cid:18)Z 1 kxk⋆ 0 x∗(s) ds (cf. 0 [x∗∗(t)ϕ(t)]p dt t (cid:19)1/p , where x∗∗(t) = 1 [34], Lemma 3.1). Moreover, if γϕ > 0, then applying Corollary 3 on page 57 of [21] to the function ψ = ϕp (1 ≤ p < ∞) (see also [24, Theorem 6.4(a)]), we obtain that there exists a constant K = K(p) ≥ 1 such that t R t K−1ϕp(t) ≤Z t 0 ϕp(s) s ds ≤ Kϕp(t) (0 < t ≤ 1). (17) 12 Therefore, the fundamental function ϕΛp,ϕ(t) is equivalent to ϕ(t). Inequalities (17) imply also that, if γϕ > 0, then the space Λ1,ϕ coincides with the Lorentz space Λϕ with the norm kxkΛϕ :=Z 1 0 x∗(t)dϕ(t). Recall also that the Kothe dual of the Lorentz space Λϕ is isometric to the Marcinkiewicz space M ϕ with ϕ(t) = t ϕ(t) and its norm is kxkM ϕ = sup 0<t≤1 ϕ(t)x∗∗(t) = sup 0<t≤1 (cf. [21], Theorem 5.2 on page 112). 1 ϕ(t)Z t 0 x∗(s)ds We will prove that condition (13) is necessary and sufficient for Lorentz and Marcin- kiewicz spaces to be w-decomposable. We start by proving a specific geometric property of Lorentz spaces. Proposition 4.1. Let ϕ be an increasing non-negative concave function on [0, 1] such that γϕ > 0, and let 1 ≤ p < ∞. Then for arbitrary b > 1 there exists a constant C = C(b, ϕ, p) > 0 with the following property: for any two-sided non-decreasing sequence (aj)+∞j=−∞ b−jχ(aj−1,aj ] belongs to Λp,ϕ, we have of reals from [0, 1] such that the function x = P+∞j=−∞ kxkp Λp,ϕ C ≈ +∞ Xj=−∞ b−pjϕp(aj − aj−1). (18) Proof. Since γϕ > 0, there exist κ > 0 and A > 0 such that inequality (16) holds. Choose a constant C1 = C1(ϕ) > 1 satisfying the inequality (C1 + 1)κ A ≥ 2 K 2, (19) where K is the constant from (17), and denote by I the set of all indices j ∈ Z such that aj − aj−1 ≥ C1 aj−1. We prove the following equivalences: and, if b > C1 + 1, Z aj aj−1 ϕp(t) t dt ≈ ϕp(aj − aj−1), j ∈ I ϕp(t) t dt, kxkp Λp,ϕ ≈Xj∈I b−pjZ aj b−pjϕp(aj − aj−1) ≈ Xj∈I aj−1 +∞ Xj=−∞ b−pjϕp(aj − aj−1), (20) (21) (22) with constants which depend only on b, ϕ and p. At first, if j ∈ I then, by (16) and (19), ϕ(aj) ≥ ϕ((C1 + 1)aj−1) ≥ (C1 + 1)κ A 13 ϕ(aj−1) ≥ 2K 2 ϕ(aj−1). Combining this with (17) and the inequality ϕ(aj) ≤ ϕ(aj − aj−1) + ϕ(aj−1) ≤ 2 ϕ(aj − aj−1), (23) we obtain 1 2K ϕp(aj − aj−1) ≤ 1 2K ϕp(aj) ≤ 1 2K [2 ϕp(aj) − (2K 2)p ϕp(aj−1)] ϕp(aj) K − K ϕp(aj−1) 1 2K ≤ ≤ Z aj ≤ Z aj 0 0 [2 ϕp(aj) − 2K 2 ϕp(aj−1)] = ϕp(t) ϕp(t) ϕp(t) dt −Z aj−1 dt ≤ Kϕp(aj) ≤ 2pKϕp(aj − aj−1), dt =Z aj aj−1 t t 0 dt t ϕp(t) t which implies (20). Now, assuming b > C1 + 1, we show that the set I is unbounded from below. In fact, otherwise there is j0 ∈ Z such that aj − aj−1 < C1aj−1 for all j ≤ j0. Then, we have aj0 ≤ (C1 + 1)j0−j aj (j ≤ j0) and by (17) and the concavity of ϕ, kxkp Λp,ϕ ≥ sup j≤j0 1 K ≥ sup j≤j0 ϕp(t) b−pjZ aj 0 dt ≥ (C1 + 1)p(j−j0) t 1 K sup j≤j0 b−pjϕp(aj) bpj ϕp(aj0) = ∞. Therefore, for a given i /∈ I we can find k = max{j < i : j ∈ I}. Further, from the definition of I it follows that ai < (C1 + 1)i−k ak. Since ϕ is concave and 2ak−1 ≤ ak, we get Z ai ai−1 ϕp(t) t dt ≤ Z (C1+1)i−kak ak ϕp(t) t dt and so b−piZ ai ai−1 ϕ(t) t dt )Z (C1+1)i−kak ak ϕ(t) t dt ak ) ak ak 2 ≤ ϕp−1((C1 + 1)i−kak)Z (C1+1)i−kak ≤ 2p−1(C1 + 1)(p−1)(i−k) ϕp−1( ak ≤ 2p−1(C1 + 1)p(i−k) ϕp−1( 2 ak ≤ 2p(C1 + 1)p(i−k)ϕp−1( 2 ≤ 2p(C1 + 1)p(i−k) Z ak ϕp(t) ≤ 2p(C1 + 1)p(i−k) Z ak dt ≤ 2p(cid:16) C1 + 1 (cid:17)p(i−k) ) Z ak ϕp(t) ϕp(t) ak−1 ak/2 ak/2 dt dt b t t t b−pk Z ak ak−1 ϕ(ak) ak ϕ(t) t 14 dt ϕp(t) t dt. Since b > C1 + 1 we obtain (21). In a similar way, applying (23) for j = k, we get b−piϕp(ai − ai−1) ≤ b−pi(C1 + 1)p(i−k)ϕp(ak) ≤ 2p(cid:18)C1 + 1 b (cid:19)p(i−k) b−pkϕp(ak − ak−1), which implies (22). Relations (20)–(22) imply (18), so we proved the statement for b > C1 + 1. To extend this result to all b > 1 it suffices to prove the following: whenever (18) holds for some b > 1 and arbitrary non-decreasing sequence (aj)∞j=−∞ with a constant C, it is automatically fulfilled for b1/2 with a constant not exceeding 2pbpC. Indeed, if y = +∞ Xj=−∞ b−j/2χ(aj−1,aj ] and z = b−jχ(a2j−2,a2j ], +∞ Xj=−∞ then kykp Λp,ϕ = = bp/2 ≈ On the other hand, +∞ +∞ Xj=−∞ Xj=−∞ Xj=−∞ +∞ dt t aj−1 ϕp(t) b−pj/2 Z aj b−p(2j−1)/2 Z a2j−1 b−pj Z a2j ϕp(t) a2j−2 a2j−2 ϕp(t) dt t + +∞ Xj=−∞ b−pj Z a2j a2j−1 ϕp(t) dt t dt t = kzkp Λp,ϕ. b−pjϕp(a2j − a2j−2) 2pbp/2 ≈ b−pjϕp(a2j − a2j−1) + b−p(2j−1)/2ϕp(a2j−1 − a2j−2), so we get an analog of (18) for y and b1/2 and the proof is complete. Remark 4.2. For the space Λ1,ϕ = Λϕ the result can be proved also by using the following well-known formula (cf. formula 5.1 in [21] on page 108) kxkΛϕ = +∞ Xj=−∞ (b−j − b−j−1)ϕ(aj). Let, as above, for a given weight w, Mk = {t ∈ [0, 1] : w(t) ∈ [2k, 2k+1)} (k ∈ Z) and (wr)∞r=1 be the non-increasing rearrangement of the sequence (m(Mk))+∞k=−∞ Theorem 4.3. Let ϕ be an increasing concave function on [0, 1] such that γϕ > 0, 1 ≤ p < ∞ and let w be a weight on [0, 1]. Then the Lorentz space X := Λp,ϕ is w-decomposable if and only if ϕ satisfies condition (13). . 15 Proof. If X = Λp,ϕ is w-decomposable then, by Corollary 3.6, the relation (13) holds for the fundamental function ϕX. Since, as it was mentioned above, ϕ ≈ ϕX, then (13) is fulfilled for ϕ as well. Conversely, suppose that ϕ satisfies (13). Let n ∈ N and x1, x2, . . . , xn be non-negative functions from X satisfying (12). Evidently, there exist x′1, x′2, . . . , x′n ∈ X taking their values from the set {2−k}∞k=−∞ ∪ {0} and such that xi(t) ≈ x′i(t) (0 < t ≤ 1). Clearly, m(supp x′i) = m(supp xi) ≤ wi (1 ≤ i ≤ n) and x′i(t) = 2−ko = Xi=1 m{t : x′i(t) = 2−k} mnt : Xi=1 n n 2 for all integer k. Therefore, applying (13), we get that n Xi=1 ϕp(m{t : x′i(t) = 2−k}) ≈ ϕp(m{t : n Xi=1 x′i(t) = 2−k}) (k ∈ Z). (24) (25) (26) On the other hand, Proposition 4.1 yields kx′ikp X ≈ +∞ Xk=−∞ n and 2−pkϕp(m{t : x′i(t) = 2−k}) (1 ≤ i ≤ n) x′ikp X ≈ k Xi=1 +∞ Xk=−∞ 2−pkϕp(m{t : x′i(t) = 2−k}) n Xi=1 with a constant which depends only on ϕ and p. Combining relations (25) and (26) with (24), we obtain (9) for x′i and so for xi. The proof is complete. In particular, from the above theorem and a remark after Theorem 3.8 it follows that t is 1/t-decomposable and the Lorentz space Λϕ generated by the function ϕ(t) = t log e therefore the Banach couple (Λ(ϕ), Λ(ϕ)( 1 t )) is K-monotone. Theorem 4.4. Suppose that ϕ is an increasing concave function on [0, 1] such that γϕ > 0 and 1 ≤ p < ∞. The following conditions are equivalent: (a) there exists a weight w on [0, 1] such that the Lorentz space Λp,ϕ is w-decomposable; (b) ϕ is equivalent to a regularly varying function at zero of order p. Proof. First, if X := Λp,ϕ is w-decomposable for some weight w on [0, 1], then, by Corol- lary 3.7, as in the proof of the previous theorem, we conclude that ϕ is equivalent to a regularly varying function at zero of order p. Conversely, suppose that ϕ is equivalent to a function that varies regularly at zero with order p, that is, ϕ satisfies (3) for some τ (N) (N = 1, 2, . . . ). Consider a family (MN )∞N =1 of pairwise disjoint measurable subsets of [0, 1] such that m(M2) = min(τ (2), 1/4), m(MN ) = min(τ (N), m(MN−1) 2 ), N > 2, and let M1 := [0, 1] \S∞N =2 MN . Set w(t) := 2N for all t ∈ MN and N ∈ N. Clearly, m(MN +1) ≤ m(MN )/2 (N ∈ N). Therefore, by Theorem 3.8, ϕ satisfies (13) for any sequence (τN )∞N =1 majorized by the sequence (m(MN ))∞N =1. To complete the proof it remains to apply Theorem 4.3. 16 It is obvious that Lp-spaces (1 ≤ p ≤ ∞) are w-decomposable for every weight w. On the other hand, we show that for an arbitrary weight w there exist w-decomposable Lorentz spaces Λϕ different from L1. Theorem 4.5. Let w be an arbitrary weight on [0, 1]. Then there exists an increasing concave function ϕ such that the space Λϕ is w-decomposable and Λϕ 6= L1. Proof. As above, Mk = {t ∈ [0, 1] : w(t) ∈ [2k, 2k+1)} for k ∈ Z and (wr)∞r=1 is the non-increasing rearrangement of the sequence (m(Mk))+∞k=−∞ . Define G(α) := min{α, wr}, α ≥ 0. ∞ Xr=1 it follows that Then G(1) = 1, G(0) = 0 and G is increasing and continuous at zero. Let (tk)∞k=0 be a sequence from (0, 1] such that t0 = 1, 0 < tk < tk−1/3 for k ≥ 1 and (27) G(tk+1) ≤ 2−k tk, k = 0, 1, . . . . Then we set ϕ′k(t) = maxi=0,1,...,k{2iχ[0,ti](t)}, k = 0, 1, . . . and ϕ′(t) = limk→∞ ϕ′k(t) (0 < t ≤ 1). It is easy to see that ϕ′k and ϕ′ are non-increasing functions on(0, 1]. Moreover, since tkϕ′(tk) = tk 2k ≤ Z 1 Xk=0 ∞ 2 3 tk−1 2k−1 = 2 3 tk−1ϕ′(tk−1) 3(cid:19)k Xk=0(cid:18)2 < ∞. ∞ 0 ϕ′(tk)tk ≤ ϕ′(t)dt ≤ Therefore, the function ϕ(t) :=R t In view of Theorem 4.3, it suffices to show that for some constant C ≥ 1 and for any sequence of reals (dr)∞r=1 such that 0 < dr ≤ wr (r = 1, 2, . . . ) we have 0 ϕ′(s) ds is well-defined, increasing and concave on (0, 1]. We shall prove that the Lorentz space Λϕ is w-decomposable. ∞ Xr=1 dr! . dr! ≤ ϕ(dr) ≤ Cϕ ∞ Xr=1 Note that the left hand side of this inequality is an immediate consequence of the concavity ϕ ∞ Xr=1 of ϕ. Further, since ϕk(t) := R t 0 ϕ′k(s) ds ↑ ϕ(t), then limk→∞P∞r=1 ϕk(dr) = P∞r=1 ϕ(dr). P∞r=1 ϕk(dr) ϕk (P∞r=1 dr) ≤ 3, k ≥ 0. Therefore, it is enough to prove that (28) Noting that P∞r=1 dr ≤ t0 = 1, we set k0 := maxnk = 0, 1, 2,· · · : ∞ Xr=1 dr ≤ tko. 17 From the definition of ϕk it follows that ϕk ∞ Xr=1 dr! = 2k ∞ Xr=1 dr = ∞ Xr=1 ϕk(dr) if 0 ≤ k ≤ k0. Since tk0+1 < ∞ Pr=1 Let k > k0 be arbitrary. The inequality dr > tk implies that dr ≤ tk0, then, again by the definition of ϕk, dr ≤ 2ϕk0+1 ∞ Xr=1 ϕk0+1(dr) ≤ 2k0+1 Xr=1 ∞ ∞ dr! . ∞ Xr=1 Pr=1 dr! > ϕk(tk) = 2ktk. ϕk ∞ Xr=1 Moreover, since we obtain ϕk+1(dr) =(2k+1dr = 2 ϕk(dr), 2ktk+1 + ϕk(dr), if dr ≤ tk+1, if dr > tk+1, (29) (30) (31) Hence, for any k > k0, by (31) and (27), we obtain ∞ ∞ ∞ ϕk(dr) = Xr=1 ϕk+1(dr) − Xr=1 Xr=1 P∞r=1 ϕk+1(dr) ϕk+1 (P∞r=1 dr) ≤ P∞r=1 ϕk(dr) ϕk (P∞r=1 dr) ≤ P∞r=1 ϕk(dr) ϕk (P∞r=1 dr) min(2ktk+1, 2kdr) ≤ 2kG(tk+1). + P∞r=1 ϕk+1(dr) −P∞r=1 ϕk(dr) G(tk+1) + tk ϕk (P∞r=1 dr) ≤ P∞r=1 ϕk(dr) ϕk (P∞r=1 dr) + 2−k. Applying the last estimate together with (29) and (30), we obtain (28). It is easy to see that ϕ(t) is not equivalent to t, and therefore Λϕ 6= L1. The proof is complete. Remark 4.6. Theorem 4.5 can be easily extended to the spaces Λp,ψ with p ∈ (1,∞). Indeed, let w be an arbitrary weight on [0, 1] and ϕ be the function from the proof of Theorem 4.5. Set ψ := ϕ1/p. Clearly, ψ is an increasing concave function not equivalent to the function t1/p. Therefore, Λp,ψ 6= Lp. Since relation (13) is fulfilled for ψ as well, then, by Theorem 4.3, the space Λp,ψ is w-decomposable. Our next goal is to prove analogous results for Marcinkiewicz spaces Mϕ. To make use of the duality of Lorentz and Marcinkiewicz spaces we will need the following statement which is of interest in its own right. 18 Theorem 4.7. Let X be a Banach lattice on a σ-finite measure space (Ω, Σ, µ) with suppX = Ω which has the Fatou property and w be a non-trivial weight on Ω. Then the couple (X, X(w)) is K-monotone if and only if (X′, X′(w)) is K-monotone, where X′ is the Kothe dual of X. The proof follows from Theorem 3.1 proved in [37] and the following result. Theorem 4.8. Let X be a Banach lattice on a σ-finite measure space (Ω, Σ, µ) with suppX = Ω which has the Fatou property and w be a non-trivial weight on Ω. Then X is w-decomposable if and only if its Kothe dual X′ is w-decomposable. Proof. Suppose that X is w-decomposable. Let n ∈ N and the functions x′1, x′2, . . . , x′n, y′1, y′2, . . . , y′n ∈ X′ satisfy (5) (with the norm from X′) and (6). Take a function x ∈ X, kxkX = 1, such that supp x ⊂ supp x′i and n n Si=1 x′ikX ′ ≤ 2ZΩ Xi=1 k n Xi=1 x′i(t)x(t) dµ. Now, consider yi ∈ X such that supp yi ⊂ supp y′i, kyikX = kxχsupp x′ kyikX ZΩ y′i(t)yi(t) dµ, 1 ≤ i ≤ n. ky′ikX ′ ≤ 2 ikX and Then, according to the hypothesis, k n Xi=1 yikX ≤ Ck n Xi=1 xχsupp x′ ikX = C, and, therefore, k n Xi=1 y′ikX ′ ≥ ≥ ≥ y′i(t)yi(t) dµ y′j(t) dµ = n n 1 n yi(t) 1 2C C ZΩ Xi=1 Xj=1 Xi=1 ky′ikX ′kyikX = Xi=1 ZΩ x′i(t)x(t)χsupp x′ 1 2C 1 2C n i n n 1 C Xi=1 ZΩ kx′ikX ′kxχsupp x′ Xi=1 1 4Ck Xi=1 (t) dµ ≥ n ikX x′ikX ′. Certainly, the same argument can be applied to get the opposite estimate. The proof is complete. Since M′ϕ = Λ ϕ (cf. [21], p. 117) and δϕ + γ ϕ = 1 for any increasing concave function ϕ on [0, 1] (cf. [21], Theorem 4.12 on page 107 or [24], p. 28), then by Theorems 4.3, 4.4 and 4.8 we immediately obtain the following statements. 19 Corollary 4.9. Let ϕ be an increasing concave function on [0, 1] such that δϕ < 1 and let w be a weight on [0, 1]. Then the Marcinkiewicz space Mϕ is w-decomposable if and only if ϕ(t) = t/ϕ(t) satisfies (13) with p = 1. Corollary 4.10. If ϕ is an increasing concave function on [0, 1] such that δϕ < 1, then the space Mϕ is w-decomposable for some weight w on [0, 1] if and only if ϕ is equivalent to a regularly varying function at zero of order ∞. In the paper [18], Kalton proved that if X and Y are symmetric sequence spaces with the Fatou property such that the couple (X, Y (w)) is K-monotone for some non-trivial weight w, then X = lp and Y = lq with 1 ≤ p, q ≤ ∞. The results in this section and Theorem 3.1 show that in the case of symmetric function spaces on [0, 1] the situation is completely different. The following theorems present new examples of K-monotone Banach couples of weighted Lorentz and Marcinkiewicz function spaces. The first theorem follows from Theorem 3.1, Theorem 4.4, Theorem 4.5 and Remark 2 and the second one from Theorem 3.1, Theorem 4.8 on the duality and Corollary 4.10. Theorem 4.11. If ϕ is an increasing concave function on [0, 1] such that γϕ > 0 and 1 ≤ p < ∞, then the weighted couple (Λp,ϕ, Λp,ϕ(w)) is K-monotone for some (non-trivial) weight w on [0, 1] if and only if ϕ is equivalent to a regularly varying function at zero of order p. On the other hand, for arbitrary weight w on [0, 1] and 1 ≤ p < ∞ there exists an increasing concave function ϕ on [0, 1] such that the couple (Λp,ϕ, Λp,ϕ(w)) is K-monotone and Λp,ϕ 6= Lp. Theorem 4.12. If ϕ is an increasing concave function on [0, 1] such that δϕ < 1, then the weighted couple (Mϕ, Mϕ(w)) is K-monotone for some (non-trivial) weight w on [0, 1] if and only if ϕ is equivalent to a regularly varying function at zero of order ∞. 5 w–decomposable Orlicz spaces As we have seen in the previous section, in order to check the property of w-decomposa- bility for Lorentz spaces, it is enough to consider only characteristic functions (Theorem 4.3). In this section we will prove that in the case of Orlicz spaces it is sufficient to examine scalar multiples of characteristic functions. As above, for a weight w on [0, 1] let Mk := {t ∈ [0, 1] : w(t) ∈ [2k, 2k+1)} (k ∈ and { ¯Mr}∞r=1 Z), (wr)∞r=1 be the non-increasing rearrangement of the sequence (m(Mk))+∞k=−∞ denote any rearrangement of the sets Mk such that m( ¯Mr) = wr, r = 1, 2, . . . Theorem 5.1. Let an Orlicz function F satisfy the ∆2-condition for large u and let w be a weight on [0, 1]. Then, the Orlicz space LF = LF [0, 1] is w-decomposable if and only if there exists p ∈ [1,∞) such that for any n ∈ N, all measurable sets Ak ⊂ M k and reals ck (1 ≤ k ≤ n) we have n with a constant independent of ck, Ak (1 ≤ k ≤ n) and n ∈ N. If, in addition, the com- plementary function F ∗ satisfies the ∆2-condition for large u, then the w-decomposability of LF implies that F is equivalent to a regularly varying Orlicz function at ∞ of order p. ckχAkkp LF ≈ kckχAkkp LF (32) k Xk=1 n Xk=1 20 Proof. Suppose, first, that LF is w-decomposable. By Proposition 3.2, there is p ∈ [1,∞] such that (9) holds for X = LF , which implies (32). Since F satisfies the ∆2-condition for large u > 0, then αX > 0. Therefore, by Corollary 3.7, p < ∞. Conversely, let n ∈ N and yk ∈ LF , supp yk ⊂ ¯Mk, 1 ≤ k ≤ n. We may (and will) assume that yk are positive bounded functions and Taking into account Theorem 3.3, we need to show that n Xk=1 kykkp LF = 1. k n Xk=1 ykkp LF ≈ 1, with a constant independent from n and yk. For each 1 ≤ k ≤ n we set (33) (34) and ck = kykkLF 2ϕLF (m(supp yk)) yk(t) :=(yk(t), 0, if yk(t) ≥ ck, if yk(t) < ck. Applying (32) to the functions ckχsuppyk and taking into account the definition of ck and (33) we get k n Xk=1 ckχsuppykkp LF ≤ C1 = C1 n n Xk=1 Xk=1 cp kϕLF (m(suppyk))p 2−p kykkp LF = 2−pC1. Up to equivalence of norms the Orlicz space LF = LF [0, 1] depends only on the behaviour of F for large enough u > 0. Therefore, we may assume that F (2u) ≤ C2F (u) for all u > 0. Then, from the last inequality it follows that n Xk=1 m(suppyk)F (ck) ≤ C3, (35) where C3 is a constant independent of n and yk. Moreover, from the definition of ck and yk we have kykkLF ≤ kykkLF and kykkLF ≥ kykkLF − kckχsupp ykkLF = 1 2kykkLF . Next, let us show that there is rk ∈ [ck, supt yk(t)] such that F (rk) = F (cid:18) rk kykkLF(cid:19)Z 1 0 F (yk(t))dt. 21 (36) (37) In fact, consider the function Hk(t) := From the equality R 1 0 F ( yk(t) kyk(t)kLF F (yk(t)) F (cid:16) yk(t) kyk(t)kLF(cid:17) , t ∈ supp yk. )dt = 1 it follows that Hk(t) ≤Z 1 0 inf F [yk(t)]dt ≤ sup t∈supp yk Hk(t). t∈supp yk yk(t) ≥ ck, by the continuity of F , equality (37) holds for some rk Thus, since inf t∈supp yk from the interval [ck, supt yk(t)]. Next, define dk ∈ [0, 1] (k = 1, 2, . . . , n) as follows: dk =(ϕ−1 LF (cid:16)kykkLF rk (cid:17) , m(supp yk), if kykkLF ≤ rkϕLF (m(supp yk)), if kykkLF > rkϕLF (m(supp yk)). Clearly, by the definition of dk, rkϕLF (dk) ≤ kykkLF . On the other hand, since rk ≥ ck, we obtain rkϕLF (dk) ≥ 1 2kykkLF , (38) (39) LF whence dk ≥ ϕ−1 (kykkLF /(2rk)). Hence, taking into account that F satisfies the ∆2- condition with constant C2 for all u > 0, the formula ϕLF (t) = 1/F −1(1/t) (see formula (9.23) in [20] on page 79 of the English version or Corollary 5 in [25] on page 58) and (37), we have dkF (rk) ≥ F (rk) kykkLF(cid:17) ≥ F (cid:16) 2rk 1 C2 F (rk) M(cid:16) rk kykkLF(cid:17) = 1 C2 1 Z0 F [yk(t)] dt. (40) Conversely, from the equality 1/dk = F (1/ϕLF (dk)), (38) and (37) it follows that dkF (rk) = F (rk) ϕLF (dk)) ≤ 1 F ( F (rk) F (cid:16) rk kykkLF(cid:17) = 1 Z0 F [yk(t)] dt. (41) Now, by the definition of dk, we have dk ≤ m(supp yk). Therefore, we can define the scalar multiples of characteristic functions fk(t) := rkχBk(t), where Bk ⊂ supp yk and m(Bk) = dk. According to (38), (39) and (36), we have 1 4kykkLF ≤ kfkkLF ≤ kykkLF , k = 1, 2, . . . , n. 22 Therefore, in view of (32) and (33), we obtain k n Xk=1 fkkp LF ≈ n Xk=1 kfkkp LF ≈ n Xk=1 kykkp LF = 1, with constants which depend only on p. Hence, taking into account that F satisfies the ∆2-condition, we conclude that (34) will be proved once we show that k n Xk=1 ykkLF ≈ k n Xk=1 fkkLF with constants independent of n and yk. Since the functions fk (respectively, yk) are pairwise disjoint, in view of estimate (41), we find that Z 1 0 n F [ Xk=1 n fk(t)] dt = Xk=1 ≤ Z 1 0 Conversely, by (40) and (35), we get n Xk=1Z 1 0 F (yk(t)) dt n dkF (rk) ≤ Xk=1 F [ yk(t)] dt. Z 1 0 F [ n Xk=1 yk(t)] dt ≤ F [yk(t)] dt + m(supp yk)F (ck) n Xk=1 n 0 Xk=1Z 1 ≤ C2Z 1 0 n Xk=1 F [ fk(t)] dt + C3, and we come to the desired result. In order to obtain the second assertion of the theorem it is sufficient to apply Corollary 3.6, Lemmas 2.1 and 2.2, Proposition 2.3 and the elementary observation that condition (a) in that proposition implies the equivalence of F to an Orlicz function which is regularly varying at ∞ of order p. Remark 5.2. Arguing in the same way as in the proof of Theorem 5.1 we may obtain the following result: Let an Orlicz function F satisfy the ∆2-condition for large u and 1 < p, q < ∞. The Orlicz space LF [0, 1] satisfies the upper p-estimate, respectively the lower q-estimate, if and only if there exsists a constant C > 0 such that for any n ∈ N, all pairwise disjoint measurable sets Ak and reals ck we have respectively, k n Xk=1 ckχAkkLF ≤ C( n Xk=1 kckχAkkp LF )1/p, n ( Xk=1 kckχAkkq LF )1/q ≤ Ck n Xk=1 ckχAkkLF . 23 However, an inspection of the proof of results from [19] (pages 120-121 and 124) shows that the first of these inequalities is equivalent to either of the following conditions: the Orlicz space LF [0, 1] is p-convex or LF [0, 1] satisfies the upper p-estimate or there exists an Orlicz function F1 equivalent to F for large arguments such that F1(u1/p) is a convex function on [0,∞). At the same time, the second of them is equivalent to either of the following conditions: the Orlicz space LF [0, 1] is q-concave or LF [0, 1] satisfies the lower q-estimate or there exists an Orlicz function F1 equivalent to F for large arguments such that F1(u1/q) is a concave function on [0,∞). The following result is analogous to Theorem 4.4 for Lorentz spaces. Theorem 5.3. Let F be an Orlicz function equivalent to an Orlicz function which is regularly varying at ∞ of order p ∈ [1,∞). Then there is a weight w on [0, 1] such that the Orlicz space LF is w-decomposable and, consequently, the couple (LF , LF (w)) is K-monotone. Proof. By Corollary 3.5, it is sufficient to find a sequence of pairwise disjoint intervals {∆k}∞k=1 from [0, 1] such that for any n ∈ N and x1, x2, . . . , xn ∈ X satisfying the condition supp xi ⊂ ∆i (1 ≤ i ≤ n), relation (9) holds. First, since F is equivalent to a regularly varying Orlicz function at ∞ of order p, then Lemma 1 and a simple compactness argument (see also [18, Lemma 6.1]) show that there exists a constant C1 > 1 such that for every k ∈ N there is vk > 0 such that for all v ≥ vk and u ∈ [k−2/8, 1] we have that F (uv) C1≈ upF (v). (42) Let v > 0, ε > 0 be arbitrary and ∆ be an interval from [0, 1] such that m(∆) ≤ ε/F (v). Moreover, suppose that z ∈ LF , z ≥ 0 and supp z ⊂ ∆. Then Z{t∈∆: z(t)≤v} F [z(t)] dt ≤ F (v) m(∆) ≤ ε. Let {∆k}∞k=1 be a sequence of disjoint intervals from [0, 1] such that m(∆k) ≤ 2−k−1(F (vk))−1 (k = 1, 2, . . . ). Then, as it was noted above, for every z ∈ LF such that z ≥ 0 and supp z ⊂ ∆k, we have (43) F [z(t)] dt ≤ 2−k−1 (k = 1, 2, . . . ). Z{t∈∆k: z(t)≤vk} Suppose that {xk}∞k=1 is an arbitrary sequence from LF such that xk ≥ 0 and supp xk ⊂ ∆k (k = 1, 2, . . . ). To prove (9) we assume that or, equivalently, = 1, xi(cid:13)(cid:13)(cid:13)LF n (cid:13)(cid:13)(cid:13) Xi=1 Xi=1 Z∆i n F [xi(t)] dt = 1. (44) 24 If λi := kxikLF (i = 1, 2, . . . ), then 0 ≤ λi ≤ 1 and Z∆i Fhxi(t) λi i dt = 1 (i = 1, 2, . . . ). Denote I1 := {i = 1, 2, . . . , n : λi ≤ i−2/8}, I2 := {1, 2, . . . , n} \ I1. Then λp i ≤ Xi∈I1 1 8Xi∈I1 i−2p ≤ 1 4 . Now, let i ∈ I2, i.e., λi ≥ i−2/8. Then, if xi(t) ≥ λivi, from (42) it follows that C−1 1 λp i Fhxi(t) λi i ≤ F [xi(t)] ≤ C1λp i Fh xi(t) λi i. Moreover, by (43) and (45), we have (45) (46) (47) Z{t∈∆i: xi(t)>λivi} Fh xi(t) λi i dt = 1 −Z{t∈∆i: xi(t)≤λivi} Fhxi(t) λi i dt ≥ 1 − 2−i−1 ≥ 3 4 , whence, taking into account the left hand side of (47), we obtain Z∆i F [xi(t)] dt ≥ C−1 1 λp i Z{t∈∆i: xi(t)>λivi} Fhxi(t) λi i dt ≥ 3 4 C−1 1 λp i , i ∈ I2. Combining this with (44) and (46), we get n Xi=1 λp i =Xi∈I1 λp i +Xi∈I2 λp i ≤ 1 4 + 4 3 C1 n Xi=1 Z∆i F [xi(t)] dt ≤ 2C1, and the first inequality in (9) is proved. On the other hand, using the right hand side of (47) and (45), we infer that Xi∈I2Z{t∈∆i: xi(t)>λivi} F [xi(t)] dt ≤ C1Xi∈I2 Xi=1 ≤ C1 n λp i Z{t∈∆i: xi(t)>λivi} Fh xi(t) λi i dt (48) λp i . At the same time, by (43) and the convexity of F , we obtain Xi∈I2Z{t∈∆i: xi(t)≤λivi} λiZ{t∈∆i: xi(t)≤λivi} Fhxi(t) λi i dt 2−i−1 = 1 2 F [xi(t)] dt ≤ Xi∈I2 Xi=1 ≤ ∞ 25 and, by (45) and the definition of I1, Hence, taking into account (44), we get Xi∈I1Z∆i Xi∈I2Z{t∈∆i: xi(t)>λivi} 1 4 . λiZ∆i Fh xi(t) λi i dt ≤ F [xi(t)] dt ≤Xi∈I1 F [xi(t)] dt = 1 −Xi∈I2Z{t∈∆i: xi(t)≤λivi} F [xi(t)] dt ≥ F [xi(t)] dt − Xi∈I1Z∆i i ≥ 1/(4C1), and so the proof of (9) is complete. 1 4 . From this and (48) it follows thatPn i=1 λp 6 Ultrasymmetric Orlicz spaces and w–decomposa- bility In the previous sections we have examined the problem of the K-monotonicity of weighted couples generated by Lorentz, Marcinkiewicz and Orlicz spaces. We have seen that the central role in the question is played by the notion of w-decomposibility. It turns out that studying the last property in a natural way leads to the so-called ultrasymmetric Orlicz spaces. Recall that a symmetric space X on [0, 1] is ultrasymmetric if X is an interpolation space between the Lorentz space ΛϕX and the Marcinkiewicz space MϕX . These spaces were studied by Pustylnik [27], who proved that they embrace all possible generalizations of Lorentz-Zygmund spaces and have a simple analytical description. Moreover, one could substitute ultrasymmetric spaces into almost all results concerning classical spaces such as Lorentz-Zygmund spaces, and so they are very useful in many applications (see, for example, Pustylnik [28] and [29]). Pustylnik asked about a description of ultrasymmetric Orlicz spaces (see [27], p. 172). In the case of reflexive Orlicz spaces this problem was solved in [4]: such a space is ultrasymmetric if and only if it coincides (up to equivalence of norms) with a Lorentz space Λp,ϕ for some 1 < p < ∞ and some increasing concave function ϕ on [0, 1]. As it was said above, the class of w–decomposable symmetric spaces is closely related to the class of ultrasymmetric Orlicz spaces. Our next theorem shows that in the case when a weight w changes sufficiently fast any w–decomposable symmetric space with non-trivial Boyd indices is an ultrasymmetric Orlicz space. Again, as above, for a weight w defined on [0, 1], let Mk := {t ∈ [0, 1] : w(t) ∈ [2k, 2k+1)} (k ∈ Z) and (wk)∞k=1 be the non-increasing rearrangement of the sequence (m(Mk))+∞k=−∞ Theorem 6.1. Let X be a symmetric space on [0, 1] with non-trivial Boyd indices and w be a weight on [0, 1] satisfying the condition: there are k0 ∈ N and c0 > 0 such that wk2k ≥ c0 for k ≥ k0. (49) 26 (a) If X is w–decomposable, then X is an ultrasymmetric Orlicz space. (b) If X has the Fatou property and (X, X(w)) is a K-monotone couple, then X is an ultrasymmetric Orlicz space. Proof. (a) Firstly, taking into account the boundedness of the dilation operator and The- orem 3.3, a symmetric space X is w–decomposable if and only if it is v–decomposable, where v(u) = w(cu) for some c > 0. Therefore, we may assume that c0 = 1. Denote Ik := [2−k, 2−k+1), ¯χIk := χIk/ϕ(2−k) (k = 1, 2, . . . ), where ϕ is the fundamental function of X. From (49) it follows that m(supp ¯χIk) ≤ wk for all k ≥ k0. Applying Corollary 3.4 to scalar multiples of ¯χIk (k ≥ k0), we get that ( ¯χIk)∞k=k0 spans lp for some p ∈ [1,∞) (p 6= ∞ because the Boyd indices of X are non-trivial). Obviously, replacing ( ¯χIk)∞k=k0 with ( ¯χIk)∞k=1 does not change this property, so for all ak ∈ R (k = 1, 2, . . . ) Then, taking into account [4, Proposition 2], we get ∞ Xk=1 (cid:13)(cid:13)(cid:13) ak ¯χIk(cid:13)(cid:13)(cid:13)X ≈ k(ak)klp. X = (L1, L∞)K lp((ϕ(2−k )2−k)∞ k=1). By Corollary 3.7, δϕ = βX < 1. Therefore, limt→∞ kσtkX→X/t = 0, and we can apply [21, Theorem II.6.6, p. 137] in the case when A is the identity operator, to obtain kxkX ≈ k(ϕ(2−k) x∗∗(2−k))∞k=1klp ≈ k(ϕ(2−k) x∗(2−k))∞k=1klp t (cid:19)1/p ≈ (cid:18)Z 1 [x∗(t)ϕ(t)]p dt 0 , and we conclude that Next, denote F (u) =Z u 0 F (t) t X = Λp,ϕ. (50) dt, where F (t) =( t 1 ϕ−1(1) ϕ−1( 1 t ) if 0 ≤ t ≤ 1, if t ≥ 1. Since F (t)/t is increasing on (0,∞), then F (u) is a convex function and for u > 0 we have that F (u/2) ≤Z u u/2 F (t) t dt ≤ F (u) ≤ F (u). Moreover, by Corollary 3.7, we have that γϕ = αX > 0, which implies that F satisfies the ∆2–condition for all u > 0. Therefore, for all u > 0 F (u) ≥ F (u/2) ≥ c F (u), that is, the functions F and F are equivalent on (0,∞). 27 Now, we recall the following definition due to Kalton [18] (see also [4], where the notion is used): For an Orlicz function F and 1 ≤ p < ∞, define the function Ψ∞F,p(u, C) for 0 < u ≤ 1, C > 1 to be the supremum (possibly ∞) of all N such that there exist ak−1 ≥ 2 for k = 2, . . . , N such that for all k either Fak (u) ≥ Cup 1 ≤ a1 < a2 < . . . < aN , or up ≥ CFak (u), where Fa(u) := F (au) have that To complete the proof it suffices to verify that for some C0 > 0, C1 > 0 and r > 0 we for a, u > 0. F (a) ak Ψ∞F,p(u, C0) ≤ C1u−r for all u ∈ (0, 1]. Indeed, once it is done, we can apply Theorem 1 from [4] to conclude that the Orlicz space LF is ultrasymmetric and that it coincides with a Lorentz space Λp,ψ generated by some increasing concave function ψ. Since the fundamental function of LF is equivalent to ϕ, then LF = Λp,ϕ, and, in view of (50), the proof is complete. Since the functions F and F are equivalent, then, by [4, Lemma 1], it is sufficient to prove the inequality for F , i.e., to prove that for some C0 > 0, C1 > 0 and r > 0 we have Ψ∞F ,p(u, C0) ≤ C1u−r for all u ∈ (0, 1]. (51) In view of w-decomposability, Corollary 3.7, Lemma 2.2 and the inequality wk ≥ 2−k, there is a constant C > 0 such that for any l = 1, 2, . . . ϕ(lt) ϕ(t) C ≈ l1/p if 0 < t ≤ 2−l. Since 0 < αX ≤ βX < 1 it follows that 0 < γϕ ≤ δϕ < 1. Therefore, from the definition of F it follows that both F and its complementary function satisfy the ∆2-condition. Hence, by Proposition 2.3 and by the definition of F once more, we obtain that there exists a constant C1 > 0 such that, for any l ∈ N and for all x ≥ F −1(2l), we have 1 C1l ≤ F (xl−1/p) F (x) ≤ C1 l . By standard arguments, there are constants C2 > 0 and C3 > 0 such that 2 up ≤ C−1 F (ua) F (a) ≤ C2up, (52) for all 0 < u ≤ 1 and any a satisfying F (a) ≥ C32u−p. Suppose that 1 ≤ a1 < a2 < . . . < aN , ak ak−1 ≥ 2 for k = 2, . . . , N such that for all k either F (uak) F (ak) ≥ 2C2up or F (uak) F (ak) ≤ 1 2C2 up. Then, by (52), we have that F (aN ) ≤ C32u−p, which implies F (a12N−1) ≤ C32u−p. Hence, N ≤ C4u−p, that is, Ψ∞F ,p(u, 2C2) ≤ C4u−p (0 < u ≤ 1), and (51) is proved. (b) This part follows immediately from (a) and Theorem 3.1. 28 Using equality (50) from the proof of Theorem 6.1, we obtain the following corollary. Corollary 6.2. Let X be a symmetric space on [0, 1] and w be a weight on [0, 1] satisfying the condition (49). Assume that either X is w–decomposable or X has the Fatou property and (X, X(w)) is K-monotone couple. If ϕX (t) = t1/p for some 1 < p < ∞, then X = Lp. Remark 6.3. Using Krivine's theorem and the arguments from the beginning of the proof of Theorem 6.1, the last assertion can be proved for p = 1 and p = ∞ as well. Remark 6.4. It is well known that there is a regularly varying at ∞ Orlicz function F such that the corresponding Orlicz space LF is not ultrasymmetric (see [18]). Thus, Theorems 4.4 and 5.3 show that condition (49) on the weight w from Theorem 6.1 and Corollary 6.2 is essential. Remark 6.5. Conversely, if LF is an ultrasymmetric reflexive Orlicz space on [0, 1], then there is a weight w on [0, 1] such that LF is w-decomposable and, equivalently, the Banach couple (LF , LF (w)) is K-monotone. In fact, in that case F is regularly varying at ∞ of order p ∈ (1,∞) (cf. [4]) and we can apply Theorem 5.3. Examples. Theorem 5.3 guarantees that a weighted couple of Orlicz spaces (LF , LF (w)) on [0, 1] is K-monotone for some weight w on [0, 1] if F is equivalent to an Orlicz function which is regularly varying at ∞ of order p ∈ [1,∞). We present some examples of such Orlicz functions below. 1. The function F (u) = up(1 + ln u) for p ≥ (3 + √5)/2 is an Orlicz function on (0,∞) which is regularly varying at ∞ of order p (cf. [24, Example 4]). 2. The function F (u) = up[1 + c sin(p ln u)] for 0 < c < 1/√2 and p ≥ (1 − √2c√1−2c2 )−1 is an Orlicz function on (0,∞) which is not regularly varying but it is equivalent to up and 1 4up ≤ F (u) ≤ 2up for all u > 0 (cf. [24, Example 10] and [25], Example 5 on p. 93 with c = 1/√5 and p ≥ 6). 3. Let an Orlicz function F be equivalent for large u to the function F (u) = up(ln u)q1(ln ln u)q2 . . . (ln . . . ln u)qn, where p ∈ (1,∞) and q1, . . . , qn are arbitrary real numbers. It is easy to see that F is equivalent to a regularly varying function at ∞ of order p (in fact, the corresponding Orlicz space LF is even ultrasymmetric [4]). 4. Some more examples of Orlicz functions that are equivalent to some regularly varying functions at ∞ of order p are given by Kalton [18]. References [1] E. V. Abakumov and A. A. Mekler, A concave regularly varying leader for equi–concave functions, J. Math. Anal. Appl. 187 (1994), no. 3, 943–951. [2] J. Arazy and M. Cwikel, A new characterization of the interpolation spaces between Lp and Lq, Math. Scand. 55 (1984), no. 2, 253–270. [3] S. V. Astashkin, On finite representability of lp–spaces in symmetric spaces, Algebra i Analiz 23 (2011), no. 2, 77–101 (Russian); English transl. in St. Petersburg Math. J. 23 (2012), no. 2, 257–273. 29 [4] S. V. Astashkin and L. Maligranda, Ultrasymmetric Orlicz spaces, J. Math. Anal. Appl. 347 (2008), no. 2, 273–285. [5] S. V. Astashkin and K. E. Tikhomirov, On stability of K-monotonicity of Banach couples, Rev. Mat. Complut. 23 (2010), no. 1, 113–137. [6] S. V. Astashkin and K. E. Tikhomirov, On stably K-monotone Banach couples, Funktsional. Anal. i Prilozhen. 44 (2010), no. 3, 65–69 (Russian); English transl. in Funct. Anal. Appl. 44 (2010), no. 3, 212–215. [7] C. Bennett and R. Sharpley, Interpolation of Operators, Academic Press, Boston 1988. [8] D. W. Boyd, Indices for the Orlicz spaces, Pacific J. Math. 65 (1971), 325–335. [9] Ju. A. Brudnyı and N. Ya. Krugljak, Interpolation Functors and Interpolation Spaces. I, North-Holland, Amsterdam 1991. [10] A. P. Calder´on, Spaces between L1 and L∞ and the theorem of Marcinkiewicz, Studia Math. 26 (1966), no. 3, 273–299. [11] M. Cwikel, Monotonicity properties of interpolation spaces, Ark. Mat. 14 (1976), no. 2, 213–236. [12] M. Cwikel, K-divisibility of the K-functional and Calder´on couples, Ark. Mat. 22 (1984), no. 1, 39–62. [13] M. Cwikel and I. Kozlov, Interpolation of weighted L1 spaces-a new proof of the Sedaev- Semenov theorem, Illinois J. Math. 46 (2002), no. 2, 405–419. [14] M. Cwikel and P. G. Nilsson, The coincidence of real and complex interpolation methods for couples of weighted Banach lattices, in: Interpolation Spaces and Allied Topics in Analysis (Lund, 1983), Lecture Notes in Math. 1070, Springer, Berlin 1984, 54–65. [15] M. Cwikel and P. G. Nilsson, Interpolation of weighted Banach lattices, Mem. Amer. Math. Soc. 165 (2003), 1–105. [16] M. Cwikel, P. G. Nilsson and G. Schechtman, A characterization of relatively decomposable Banach lattices, Mem. Amer. Math. Soc. 165 (2003), 106–127. [17] V. I. Dmitriev, On the interpolation of operators in Lp spaces, Dokl. Akad. Nauk SSSR 260 (1981), no. 5, 1051–1054; English transl. in Soviet Math. Dokl. 24 (1981), no. 2, 373–376 (1982). [18] N. J. Kalton, Calder´on couples of rearrangement invariant spaces, Studia Math. 106 (1993), no. 3, 233–277. [19] A. Kami´nska, L. Maligranda and L. E. Persson, Type, cotype and convexity properties of Orlicz spaces, Publicaciones del Departamento de Analysis Matematico, Facultad de Matematicas, Universidad Complutense de Madrid, No. 42, 1996-97, 113-126. [20] M. A. Krasnoselskiı and Ja. B. Rutickiı, Convex Functions and Orlicz Spaces, Fizmatgiz, Moscow 1958 (Russian); English transl. Noordhoff, Groningen 1961. [21] S. G. Krein, Yu. I. Petunin and E. M. Semenov, Interpolation of Linear Operators, Amer. Math. Soc., Providence 1982. [22] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces. II. Function Spaces, Springer- Verlag, Berlin-New York 1979. [23] G. G. Lorentz, Relations between function spaces, Proc. Amer. Math. Soc. 12 (1961), 127– 132. 30 [24] L. Maligranda, Indices and interpolation, Dissertationes Math. (Rozprawy Mat.) 234 (1985), 1–49. [25] L. Maligranda, Orlicz Spaces and Interpolation, Seminars in Mathematics 5, University of Campinas, Campinas SP, Brazil 1989. [26] B. S. Mitjagin, An interpolation theorem for modular spaces, Mat. Sb. (N.S.) 66 (1965), 473–482 (Russian); English transl. in Lecture Notes in Math. 1070, Springer, Berlin 1984, 10–23. [27] E. Pustylnik, Ultrasymmetric spaces, J. London Math. Soc. (2) 68 (2003), no. 1, 165–182. [28] E. Pustylnik, Sobolev type inequalities in ultrasymmetric spaces with applications to Orlicz- Sobolev embeddings, J. Funct. Spaces Appl. 3 (2005), no. 2, 183–208. [29] E. Pustylnik, Ultrasymmetric sequence spaces in approximation theory, Collect. Math. 57 (2006), no. 3, 257–277. [30] Y. Raynaud, On Lorentz-Sharpley spaces, Israel Math. Conf. Proc., Vol. 5, Bar-Ilan Univ., Ramat Gan 1992, 207–228. [31] H. P. Rosenthal, On a theorem of J. L. Krivine concerning block finite representability of lp in general Banach spaces, J. Funct. Anal. 28 (1978), no. 2, 197–225. [32] A. A. Sedaev, A description of the interpolation spaces of the couple (Lp a1), and certain related questions, Dokl. Akad. Nauk SSSR 209 (1973), 798–800; English transl. in Soviet Math. Dokl. 14 (1973), 538–541. a0 , Lp [33] A. A. Sedaev and E. M. Semenov, The possibility of describing interpolation spaces in terms of Peetre's K-method, Optimizacija 4(21) (1971), 98–114 (Russian). [34] R. Sharpley, Spaces Λα(X) and interpolation, J. Funct. Anal. 11 (1972), 479–513. [35] G. Sparr, Interpolation des espaces Lp w, C. R. Acad. Sci. Paris S´er. A 278 (1974), 491–492. [36] G. Sparr, Interpolation of weighted Lp-spaces, Studia Math. 62 (1978), no. 3, 229–271. [37] K. E. Tikhomirov, K–monotone weighted couples of Banach lattices, Sibirsk. Math. Zh. 52 (2011), no. 1, 187–200 (Russian); English transl. in Siberian Math. J. 52 (2011), no. 1, 147–158. [38] K. E. Tikhomirov, K–monotone weighted couples of Banach lattices, Candidate of Sciences Dissertation, Samara State University, Samara 2011, 115 pp. (Russian). Sergey V. Astashkin and Konstantin E. Tikhomirov Department of Mathematics and Mechanics, Samara State University Acad. Pavlova 1, 443011 Samara, Russia E-mail addresses: [email protected], [email protected] Lech Maligranda Department of Engineering Sciences and Mathematics Lulea University of Technology SE-971 87 Lulea, Sweden E-mail address: [email protected] 31
1807.03061
1
1807
2018-07-09T12:02:27
On the norm-continuity for evolution family arising from non-autonomous forms
[ "math.FA" ]
We consider evolution equations of the form \begin{equation*}\label{Abstract equation} \dot u(t)+ A(t)u(t)=0,\ \ t\in[0,T],\ \ u(0)=u_0, \end{equation*} where $A(t),\ t\in [0,T],$ are associated with a non-autonomous sesquilinear form $\mathfrak a(t,\cdot,\cdot)$ on a Hilbert space $H$ with constant domain $V\subset H.$ In this note we continue the study of fundamental operator theoretical properties of the solutions. We give a sufficient condition for norm-continuity of evolution families on each spaces $V, H$ and on the dual space $V'$ of $V.$ The abstract results are applied to a class of equations governed by time dependent Robin boundary conditions on exterior domains and by Schr\"odinger operator with time dependent potentials.
math.FA
math
ON THE NORM-CONTINUITY FOR EVOLUTION FAMILY ARISING FROM NON-AUTONOMOUS FORMS∗ OMAR EL-MENNAOUI AND HAFIDA LAASRI Abstract. We consider evolution equations of the form u(t) + A(t)u(t) = 0, t ∈ [0, T ], u(0) = u0, where A(t), t ∈ [0, T ], are associated with a non-autonomous sesquilinear form a(t, ·, ·) on a Hilbert space H with constant domain V ⊂ H. In this note we continue the study of fundamental operator theoretical properties of the solutions. We give a sufficient condition for norm-continuity of evolution families on each spaces V, H and on the dual space V ′ of V. The abstract results are applied to a class of equations governed by time dependent Robin boundary conditions on exterior domains and by Schrödinger operator with time dependent potentials. Introduction Throughout this paper H, V are two separable Hilbert spaces over C such that V is densely and continu- ously embedded into H (we write V ֒→ H). We denote by (· ·)V the scalar product and k · kV the norm d on V and by (· ·)H , k · kH the corresponding quantities in H. Let V ′ be the antidual of V and denote by h., .i the duality between V ′ and V. As usual, by identifying H with H ′ we have V ֒→ V ′ see d H ∼= H ′ ֒→ d e.g., [5]. Let a : [0, T ] × V × V → C be a non-autonomous sesquilinear form, i.e., a(t; ·, ·) is for each t ∈ [0, T ] a sesquilinear form, (1) such that a(·; u, v) is measurable for all u, v ∈ V, (2) a(t, u, v) ≤ M kukV kvkV and Re a(t, u, u) ≥ αkuk2 V (t, s ∈ [0, T ], u, v ∈ V ), for some constants M, α > 0 that are independent of t, u, v. Under these assumptions there exists for each t ∈ [0, T ] an isomorphism A(t) : V → V ′ such that hA(t)u, vi = a(t, u, v) for all u, v ∈ V. It is well known that −A(t), seen as unbounded operator with domain V, generates an analytic C0-semigroup on V ′. The operator A(t) is usually called the operator associated with a(t, ·, ·) on V ′. Moreover, we associate an operator A(t) with a(t; ·, ·) on H as follows D(A(t)) = {u ∈ V ∃f ∈ H such that a(t; u, v) = (f v)H for all v ∈ V } A(t)u = f. It is not difficult to see that A(t) is the part of A(t) in H. In fact, we have D(A(t)) = {u ∈ V : A(t)u ∈ H} and A(t)u = A(t)u. Furthermore, −A(t) with domain D(A(t)) generates a holomorphic C0-semigroup on H which is the restriction to H of that generated by −A(t). For all this results see e.g. [22, Chapter 2] or [3, Lecture 7]. We now assume that there exist 0 < γ < 1 and a continuous function ω : [0, T ] −→ [0, +∞) such that (3) a(t, u, v) − a(s, u, v) ≤ ω(t − s)kukVγ kvkVγ (t, s ∈ [0, T ], u, v ∈ V ), ∗THIS WORK IS SUPPORTED BY DEUTSCHE FORSCHUNGSGEMEINSCHAFT DFG (GRANT LA 4197/8-1) 1 2 with (4) OMAR EL-MENNAOUI AND HAFIDA LAASRI sup t∈[0,T ] ω(t) tγ/2 < ∞ and Z T 0 ω(t) t1+γ/2 < ∞ where Vγ := [H, V ] is the complex interpolation space. square root property for some t0 ∈ [0, T ] (and then for all t ∈ [0, T ] by [4, Proposition 2.5] ), i.e., In addition we assume that a(t0; ·, ·) has the (5) D(A 1 2 (t0)) = V. Recall that for symmetric forms, i.e., if a(t; u, v) = a(t; v, u) for all t, u, v, then the square root property is satisfied. Under the assumptions (1)-(5) it is known that for each x0 ∈ V the non-autonomous homogeneous Cauchy problem (6) ( u(t) + A(t)u(t) = 0 u(0) = x0, a.e. on [0, T ], has a unique solution u ∈ MR (V, H) := L2(0, T ; V ) ∩ H 1(0, T ; H) such that u ∈ C([0, T ]; V ). This result has been proved by Arendt and Monniaux [4] (see also [10]) when the form a satisfies the weaker condition (7) a(t, u, v) − a(s, u, v) ≤ ω(t − s)kukV kvkVγ (t, s ∈ [0, T ], u, v ∈ V ). In this paper we continue to investigate further regularity of the solution of (6). For this it is necessary to associate to the Cauchy problem (6) an evolution family which means that: U :=nU (t, s) : 0 ≤ s ≤ t ≤ To ⊂ L(H) (i) U (t, t) = I and U (t, s) = U (t, r)U (r, s) for every 0 ≤ r ≤ s ≤ t ≤ T, (ii) for every x ∈ X the function (t, s) 7→ U (t, s)x is continuous into H for 0 ≤ s ≤ t ≤ T (iii) for each x0 ∈ H, U (·, s)x0 is the unique solution of (6). Definition 0.1. Let Y ⊆ H be a subspace. An evolution family U ⊂ L(H) is said to be norm continuous in Y if U ⊂ L(Y ) and the map (t, s) 7→ U (t, s) is norm continuous with value in L(Y ) for 0 ≤ s < t ≤ T. If the non-autonomous form a satisfies the weaker condition (7) then it is known that (6) is governed by an evolution family which is norm continuous in V [15, Theorem 2.6], and norm continuous in H if in addition V ֒→ H is compact [15, Theorem 3.4]. However, for many boundary value problem the compactness embedding fails. In this paper we prove that the compactness assumption can be omitted provided a satisfies (3) instead of (7). This will allow us to consider a large class of examples of applications. One of the main ingredient used here is the non-autonomous returned adjoint form a∗ r : [0, T ] × V × V −→ C defined by (8) a ∗ r(t, u, v) := a(T − t, v, u) (t, s ∈ [0, T ], u, v ∈ V ). The concept of returned adjoint forms appeared in the work of D. Daners [7] but for different interest. Furthermore, [15, Theorem 2.6] cited above will be also needed to prove our main result. We note that the study of regularity properties of the evolution family with respect to (t, s) in general Banach spaces has been investigated in the case of constant domains by Komatsu [14] and Lunardi [17], and by Acquistapace [1] for time-dependent domains. We illustrate our abstract results by two relevant examples. The first one concerns the Laplacian with non-autonomous Robin boundary conditions on a unbounded Lipschitz domain. The second one traits a class of Schrödinger operators with time dependent potential. ON THE NORM-CONTINUITY FOR EVOLUTION FAMILY ARISING FROM NON-AUTONOMOUS FORMS∗1 3 1. Preliminary results Let a : [0, T ] × V × V → C a non-autonomous sesquilinear form satisfying (1) and (2). Then the following well known result regarding L2-maximal regularity in V ′ is due to J. L. Lions Theorem 1.1 (Lions, 1961). For each given s ∈ [0, T ) and x0 ∈ H the homogenuous Cauchy problems (9) ( u(t) + A(t)u(t) = 0 u(s) = x, a.e. on [s, T ], has a unique solution u ∈ MR (V, V ′) := MR (s, T ; V, V ′) := L2(s, T ; V ) ∩ H 1(s, T ; V ′). Recall that the maximal regularity space MR (V, V ′) is continuously embedded into C([s, T ], H) [21, page 106]. A proof of Theorem 1.1 using a representation theorem of linear functionals, known in the literature as Lions's representation Theorem can be found in [21, page 112] or [8, XVIII, Chpater 3, page 513]. Furthermore, we consider the non-autonomous adjoint form a∗ : [0, T ] × V × V −→ C of a defined by ∗(t; u, v) := a(t; v, u) a for all t ∈ [0, t] and u, v ∈ V. Finally, we will need to consider the returned adjoin form a∗ V −→ C given by r : [0, T ] × V × a ∗ r(t, u, v) := a ∗(T − t, u, v). Clearly, the adjoint form is a non-autonomous sesquilinear form and satisfies (1) and (2) with the same constant M, α. Moreover, the adjoint operators A∗(t), t ∈ [0, T ] of A(t), t ∈ [0, T ] coincide with the operators associated with a∗ on H. Thus applying Theorem 1.1 to the returned adjoint form we obtain that the Cauchy problem associated with A∗ r(t) := A∗(T − t) (10) ( v(t) + A∗ v(s) = x, r(t)v(t) = 0 a.e. on [s, T ], has for each x ∈ H a unique solution v ∈ M R(V, V ′). Accordingly, for every (t, s) ∈ ∆ := {(t, s) ∈ [0, T ]2 : t ≤ s} and every x ∈ H we can define the following family of linear operators (11) U (t, s)x := u(t) and U ∗ r (t, s)x := v(t), (t, s) ∈ ∆} and {U ∗ where u and v are the unique solutions in M R(V, V ′) respectively of (9) and (10). Thus each family (t, s) ∈ ∆} yields a contractive, strongly continuous evolution {U(t, s) : family on H [15, Proposition]. In the autonomous case, i.e., if a(t, ·, ·) = a(·, ·) for all t ∈ [0, T ], then one knows that −A0, the operator associated with a0 in H, generates a C0-semigroup (T (t))t≥0 in H. In this case U (t, s) := T (t − s) yields a strongly continuous evolution family on H. Moreover, we have r (t, s) : (12) U (t, s)′ = T (t − s)′ = T ∗(t − s) = U ∗(t, s) = U ∗ r (t, s). Here, T (·)′ denote the adjoint of T (·) which coincides with the C0-semigroup (T ∗(t))t≥0 associated with the adjoint form a∗. In the non-autonomous setting however, (12) fails in general even in the finite dimensional case, see [7, Remark 2.7]. Nevertheless, Proposition 1.2 below shows that the evolution families U and U ∗ r can be related in a similar way. This formula appeared in [7, Theorem 2.6]. Proposition 1.2. Let U and U ∗ r be given by (11). Then we have (13) for all x ∈ H and (t, s) ∈ ∆. x = U (T − s, T − t)x r (t, s)(cid:3)′ (cid:2)U ∗ The equality (13) will play a crucial role in the proof of our main result. We include here a new proof for the sake of completeness. 4 OMAR EL-MENNAOUI AND HAFIDA LAASRI Proof. (of Proposition 1.2) Let Λ = (0 = λ0 < λ1 < ... < λn+1 = T ) be a subdivision of [0, T ]. Let ak : V × V → C for k = 0, 1, ..., n be given by ak(u, v) := ak,Λ(u, v) := a(r; u, v)dr for u, v ∈ V. 1 λk+1 − λk Z λk+1 λk All these forms satisfy (2) with the same constants α, M. The associated operators in V ′ are denoted by Ak ∈ L(V, V ′) and are given for all u ∈ V and k = 0, 1, ..., n by (14) Aku := Ak,Λ := 1 λk+1 − λk Z λk+1 λk A(r)udr. Consider the non-autonomous form aΛ : [0, T ] × V × V → C defined by (15) aΛ(t; ·, ·) :=(ak(·, ·) an(·, ·) if t ∈ [λk, λk+1) if t = T . Its associated time dependent operator AΛ(·) : [0, T ] ⊂ L(V, V ′) is given by (16) AΛ(t) :=(Ak An if t ∈ [λk, λk+1) if t = T . Next denote by Tk the C0−semigroup associated with ak in H for all k = 0, 1...n. Then applying Theorem 1.1) to the form aΛ we obtain that in this case the associated evolution family UΛ(t, s) is given explicitly for λm−1 ≤ s < λm < ... < λl−1 ≤ t < λl by (17) UΛ(t, s) := Tl−1(t − λl−1)Tl−2(λl−1 − λl−2)...Tm(λm+1 − λm)Tm−1(λm − s), and for λl−1 ≤ a ≤ b < λl by (18) UΛ(t, s) := Tl−1(t − s). By [20, Theorem 3.2] we know that (UΛ)Λ converges weakly in M R(V, V ′) as Λ → 0 and lim Λ→0 kUΛ − U kMR(V,V ′) = 0 The continuous embedding of M R(V, V ′) into C([0, T ]; H) implies that lim Λ→0 UΛ = U in the weak operator topology of L(H). Now, let (t, s) ∈ ∆ with λm−1 ≤ s < λm < ... < λl−1 ≤ t < λl be fixed. Applying the above approximation argument to a∗ r one obtains that (19) (20) U ∗ Λ,r(t, s) = T ∗ = T ′ l−1,r(t − λl−1)T ∗ l−1,r(t − λl−1)T ′ l−2,r(λl−1 − λl−2)...T ∗ l−2,r(λl−1 − λl−2)...T ′ m,r(λm+1 − λm)T ∗ m,r(λm+1 − λm)T ′ m−1,r(λm − s) m−1,r(λm − s) where Tk,r and T ∗ k,r are the C0-semigroups associated with (21) ak,r(u, v) := 1 λk+1 − λk Z λk+1 λk a(T − r; u, v)dr = 1 λk+1 − λk Z T −λk T −λk+1 a(r; u, v)dr k,r, respectively. Recall that T ∗ and its adjoint a∗ On the other hand, the last equality in (21) implies that Tk,r coincides with the semigroup associated with ak,ΛT where ΛT is the subdivision ΛT := (0 = T − λn+1 < T − λn < ... < T − λ1 < T − λ0 = T ). It follows from (17)-(18) and (19)-(20) that k,r = T ′ k,r. Λ,r(t, s)i′ hU ∗ = Tm−1,r(λm − s)Tm,r(λm+1 − λm)...Tl−2,r(λl−1 − λl−2)Tl−1,r(t − λl−1) = Tm−1(cid:16)(T − s) − (T − λm)(cid:17)Tm(cid:16)(T − λm) − (T − λm+1)(cid:17)...Tl−1(cid:16)(T − λl−1) − (T − t)(cid:17) = UΛT (T − s, T − t) Finally, the desired equality (13) follows by passing to the limit as Λ = ΛT → 0. (cid:3) ON THE NORM-CONTINUITY FOR EVOLUTION FAMILY ARISING FROM NON-AUTONOMOUS FORMS∗2 5 Remark 1.3. The coerciveness assumption in (2) may be replaced with (22) Re a(t, u, u) + ωkuk2 H ≥ αkuk2 V (t ∈ [0, T ], u ∈ V ) for some ω ∈ R. In fact, a satisfies (22) if and only if the form aω given by aω(t; ·, ·) := a(t; ·, ·) + ω(· ·) satisfies the second inequality in (2). Moreover, if u ∈ M R(V, V ′) and v := e−w.u, then v ∈ M R(V, V ′) and u satisfies (9) if and only if v satisfies v(t) + (ω + A(t))v(t) = 0 t−a.e. on [s, T ], v(s) = x. 2. Norm continuous evolution family In this section we assume that the non-autonomous form a satisfies (2)-(5). Thus as mentioned in the introduction, under theses assumptions the Cauchy problem (9) has L2-maximal regularity in H. Thus for each x ∈ V, U (·, s)x ∈ MR (V, H) := MR (s, T ; V, H) := L2(s, T ; V ) ∩ H 1(s, T ; H). Moreover, U (·, s)x ∈ C[s, T ]; V ) by [4, Theorem 4.2]. From [15, Theorem 2.7] we known that the re- striction of U to V defines an evolution family which norm continuous. The same is also true for the Cauchy problem (10) and the assocaited evolution family U ∗ r inherits all properties of a. In the following we establish that U can be extended to a strongly continuous evolution family on V ′. r since the returned adjoint form a∗ Proposition 2.1. Let a be a non-autonomous sesquilinear form satisfying (2)-(5). Then U can be extended to a strongly continuous evolution family on V ′, which we still denote U. Proof. Let x ∈ H and (t, s) ∈ ∆. Then using Proposition 1.2 and the fact that U and U ∗ strongly continuous evolution families on V and H we obtain that r define both kU (t, s)xkV ′ = sup < U (t, s)x, v > kvkV =1 v∈V = sup (U (t, s)xv)H = sup (xU (t, s)′v)H kvkV =1 v∈V = sup kvkV =1 v∈V = sup kvkV =1 v∈V kvkV =1 v∈V (xU ∗ r (T − s, T − t)v)H < x, U ∗ r (T − s, T − t)v > ≤ kxkV ′kU ∗ ≤ ckxkV ′ r (T − s, T − t)kL(V ) where c > 0 is such that sup t,s∈∆ kU ∗ r (t, s)kL(V ) ≤ c. Thus, the claim follows since H is dense in V ′. (cid:3) Let ∆ := {(t, s) ∈ ∆ t ≥ s}. The following theorem is the main result of this paper Theorem 2.2. Let a be a non-autonomous sesquilinear form satisfying (2)-(5). Let {U (t, s) : (t, s) ∈ ∆} given by (11). Then the function (t, s) 7→ U (t, s) is norm continuous on ∆ into L(X) for X = V, H and V ′. Proof. The norm continuity for U in the case where X = V follows from [15, Theorem 2.7]. On the other hand, applying [15, Theorem 2.7] to a∗ r is also norm continuous on ∆ with values in L(V ). Using Proposition 1.2, we obtain by similar arguments as in the proof of Lemma 2.1 r we obtain that U ∗ (23) kU (t, s) − U (t′, s′)kL(V ′) ≤ kU ∗ r (T − s, T − t)x − U ∗ r (T − s′, T − t′)xkL(V ) for all x ∈ V ′ and (t, s), (t′, s′) ∈ ∆. This implies that U is norm continuous on ∆ with values in L(V ′). Finally, the norm continuity in H follows then by interpolation. (cid:3) 6 OMAR EL-MENNAOUI AND HAFIDA LAASRI 3. Examples This section is devoted to some relevant examples illustrating the theory developed in the previous sections. We refer to [4] and [19] and the references therein for further examples. (i) Laplacian with time dependent Robin boundary conditions on exterior domain Let Ω be a bounded domain of Rd with Lipschitz boundary Γ. Denote by σ the (d − 1)-dimensional Hausdorff measure on Γ. Let Ωext denote the exterior domain of Ω, i.e., Ωext := Rd \ Ω. Let T > 0 and α > 1/4. Let β : [0, T ] × Γ −→ R be a bounded measurable function such that β(t, ξ) − β(t, ξ) ≤ ct − sα for some constant c > 0 and every t, s ∈ [0, T ], ξ ∈ Γ. We consider the from a : [0, T ] × H 1(Ωext) × H 1(Ωext) −→ C defined by a(t; u, v) :=ZΩext ∇u · ∇vdξ +ZΩext β(t, ·)uΓ¯vΓdσ where u → uΓ : H 1(Ωext) −→ L2(Γ, σ) is the trace operator which is bounded [2, Theorem 5.36]. The operator A(t) associated with a(t; ·, ·) on H := L2(Ωext) is minus the Laplacian with time dependent Robin boundary conditions Here ∂ν is the weak normal derivative. Thus the domain of A(t) is the set ∂ν u(t) + β(t, ·)u = 0 on Γ. D(A(t)) =nu ∈ H 1(Ωext) △u ∈ L2(Ωext), ∂νu(t) + β(t, ·)uΓ = 0o and for u ∈ D(A(t)), A(t)u := − △u. Thus similarly as in [4, Section 5] one obtains that a satisfies (2)-(5) with γ := r0 + 1/2 and ω(t) = tα where r0 ∈ (0, 1/2) such that r0 + 1/2 < 2α. We note that [4, Section 5] the authors considered the Robin Laplacian on the bounded Lipschitz domain Ω. The main ingredient used there is that the trace operators are bounded from H s(Ω) with value in H s−1/2(Γ, σ) for all 1/2 < s < 3/4. This boundary trace embedding theorem holds also for unbounded Lipschitz domain, and thus for Ωext, see [18, Theorem 3.38] or [6, Lemma 3.6]. Thus applying [4, Theorem 4.1] and Theorem 2.2 we obtain that the non-autonomous Cauchy problem (24) ( u(t) − △u(t) = 0, u(0) = x ∈ H 1(Ωext) ∂νu(t) + β(t, ·)u = 0 on Γ has L2-maximal regularity in L2(Ωext) and its solution is governed by an evolution family U (·, ·) that is norm continuous on each space V, L2(Ωext) and V ′. 3.1. Non-autonomous Schrödinger operators. Let m0, m1 ∈ L1 be a measurable function such that there exist positive constants α1, α2 and κ such that Loc(Rd) and m : [0, T ] × Rd −→ R α1m0(ξ) ≤ m(t, ξ) ≤ α2m0(ξ), m(t, ξ) − m(s, ξ) ≤ κt − sm1(ξ) for almost every ξ ∈ Rd and every t, s ∈ [0, T ]. Assume moreover that there exist a constant c > 0 and s ∈ [0, 1] such that for u ∈ C∞ c (Rd) (25) m1(ξ)u(ξ)2dξ ≤ ckukHs(Rd). ZRd Consider the non-autonomous Cauchy problem (26) ( u(t)− △u(t) + m(t, ·)u(t) = 0, u(0) = x ∈ V. Here A(t) = − △ + m(t, ·) is associated with the non-autonomous form a : [0, T ] × V × V −→ C given by V :=(cid:26)u ∈ H 1(Rd) :ZRd m0(ξ)u(ξ)2dξ < ∞(cid:27) ON THE NORM-CONTINUITY FOR EVOLUTION FAMILY ARISING FROM NON-AUTONOMOUS FORMS∗3 7 and a(t; u, v) =ZRd ∇u · ∇vdξ +ZRd m(t, ξ)u(ξ)2dξ. The form a satisfies also (2)-(5) with γ := s and ω(t) = tα for α > s 2 and s ∈ [0, 1]. This example is taken from [19, Example 3.1]. Using our Theorem 2.2 we have that the solution of Cauchy problem (26) is governed by a norm continuous evolution family on L2(Rd), V and V ′. References [1] P. Acquistapace. Evolution operators and strong solutions of abstract linear parabolic equations. Differential Integral Equations 1 (1988), no. 4, 433-457. [2] R. A. Adams, J. J. F. Fournier. Sobolev spaces. Second edition. Pure and Applied Mathematics (Amsterdam), 140. Elsevier/Academic Press, Amsterdam, 2003. [3] W. Arendt. Heat kernels. 9th Internet Seminar (ISEM) 2005/2006. Available at. https://www.uni- ulm.de/mawi/iaa/members/professors/arendt.html [4] W. Arendt, S. Monniaux. Maximal regularity for non-autonomous Robin boundary conditions. Math. Nachr. 1- 16(2016) /DOI: 10.1002/mana.201400319 [5] H. Brézis. Functional Analysis, Sobolev Spaces and Partial Differential Equations. Springer, Berlin 2011. [6] M. Costabel. Boundary integral operators on Lipschitz domains: elementary results. SIAM J. Math. Anal. 19 (1988), no. 3, 613-626. [7] D. Daners. Heat kernel estimates for operators with boundary conditions. Math. Nachr. 217 (2000), 13-41. [8] R. Dautray and J.L. Lions. Analyse Mathématique et Calcul Numérique pour les Sciences et les Techniques. Vol. 8, Masson, Paris, 1988. [9] O. El-Mennaoui, H. Laasri. Stability for non-autonomous linear evolution equations with Lp− maximal regularity. Czechoslovak Mathematical Journal. 63 (138) 2013. [10] O. El-Mennaoui, H. Laasri. On evolution equations governed by non-autonomous forms. Archiv der Mathematik (2016), 1-15, DOI 10.1007/s00013-016-0903-5 [11] K. J. Engel, R. Nagel. One-Parameter Semigroups for Linear Evolution Equations, Springer-Verlag, 2000. [12] S. Fackler J.-L. Lions' problem concerning maximal regularity of equations governed by non-autonomous forms. Ann. Inst. H. Poincaré Anal. Non Linéaire 34 (2017) [13] T. Kato. Perturbation theory for linear operators. Springer-Verlag, Berlin 1992. [14] H. Komatsu, Abstract analyticity in time and unique continuation property of solutions of a parabolic equation, J. Fac. Sci. Univ. Tokyo, Sect. 1 9 (1961), 1-11. [15] H. Laasri. Regularity properties for evolution family governed by non-autonomous forms. Archiv der Mathematik (2018), https://doi.org/10.1007/s00013-018-1175-z. [16] J.L. Lions. Equations Différentielles Opérationnelles et Problèmes aux Limites. Springer-Verlag, Berlin, Göttingen, Heidelberg, 1961. [17] A. Lunardi. Differentiability with respect to (t, s) of the parabolic evolution operator. Israel J. Math. 68 (1989), no. 2, 161-184. [18] W. Mclean. Strongly elliptic systems and boundary integral equations. Cambridge University Press, Cambridge, 2000. [19] E. M. Ouhabaz. Maximal regularity for non-autonomous evolution equations governed by forms having less regularity. Arch. Math. 105 (2015), 79-91. Princeton Univ. Press 2005. [20] A. Sani, H. Laasri, Evolution Equations governed by Lipschitz Continuous Non-autonomous Forms. Czechoslovak Mathematical Journal. 65 (140) (2015), 475-491. [21] R. E. Showalter. Monotone Operators in Banach Space and Nonlinear Partial Differential Equations. Mathematical Surveys and Monographs. American Mathematical Society, Providence, RI, 1997. [22] H. Tanabe. Equations of Evolution. Pitman 1979. University of Wuppertal School of Mathematics and Natural Sciences Arbeitsgruppe Funktionalanalysis, Gaussstrasse 20 D-42119 Wuppertal, Germany E-mail address: [email protected] Ibn Zohr University, Faculty of Sciences Departement of mathematics, Agadir, Morocco E-mail address: [email protected]
1211.3511
2
1211
2013-04-20T13:53:54
On Kadison-Schwarz type quantum quadratic operators on $\bm_2(\mathbb{C})$
[ "math.FA", "math.DS" ]
In the present paper we study description of Kadison-Schwarz type quantum quadratic operators acting from $\bm_2(\mathbb{C})$ into $\bm_2(\mathbb{C})\o\bm_2(\mathbb{C})$. Note that such kind of operator is a generalization of quantum convolution. By means of such a description we provide an example of q.q.o. which is not a Kadision-Schwartz operator. Moreover, we study dynamics of an associated nonlinear (i.e. quadratic) operators acting on the state space of $\bm_2(\mathbb{C})$.
math.FA
math
ON KADISON-SCHWARZ TYPE QUANTUM QUADRATIC OPERATORS ON M2(C) FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV Abstract. In the present paper we study description of Kadison-Schwarz type quantum qua- dratic operators acting from M2(C) into M2(C) ⊗ M2(C). Note that such kind of operator is a generalization of quantum convolution. By means of such a description we provide an example of q.q.o. which is not a Kadision-Schwartz operator. Moreover, we study dynamics of an associated nonlinear (i.e. quadratic) operators acting on the state space of M2(C). Mathematics Subject Classification: 46L35, 46L55, 46A37. 60J99. Key words: quantum quadratic operators; Kadison-Schwarz operator. 3 1 0 2 r p A 0 2 ] . A F h t a m [ 2 v 1 1 5 3 . 1 1 2 1 : v i X r a 1. Introduction It is known that one of the main problems of quantum information is characterization of positive and completely positive maps on C∗-algebras. There are many papers devoted to this problem (see for example [5, 14, 24, 25]). In the literature the completely positive maps have proved to be of great importance in the structure theory of C∗-algebras. However, general positive (order-preserving) linear maps are very intractable[14, 13]. It is therefore of interest to study conditions stronger than positivity, but weaker than complete positivity. Such a condition is called Kadison-Schwarz property, i.e a map φ satisfies the Kadison-Schwarz property if φ(a)∗φ(a) ≤ φ(a∗a) holds for every a. Note that every unital completely positive map satisfies this inequality, and a famous result of Kadison states that any positive unital map satisfies the inequality for self-adjoint elements a. In [23] relations between n-positivity of a map φ and the Kadison-Schwarz property of certain map is established. Certain relations between complete positivity, positivite and the Kadison-Schwarz property have been considered in [1],[3],[2]. Some spectral and ergodic properties of Kadison-Schwarz maps were investigated in [9, 10, 22]. In [20] we have studied quantum quadratic operators (q.q.o.), i.e. maps from M2(C) into M2(C)⊗ M2(C), with the Kadison-Schwarz property. It is found some necessary conditions for the trace-preserving quadratic operators to be the Kadison-Schwarz ones. Since trace-preserving maps arise naturally in quantum information theory (see e.g. [21]) and other situations in which one wishes to restrict attention to a quantum system that should properly be considered a subsystem of a larger system with which it interacts. Note that in [6, 7] quantum quadratic op- erators acting on a von Neumann algebra were defined and studied. Certain ergodic properties of such operators were studied in [17, 18] (see for review [8]). In the present paper we continue our investigation, i.e. we are going to study further properties of q.q.o. with Kadison-Schwarz property. We will provide an example of q.q.o. which is not a Kadision-Schwarz operator, and study its dynamics. We should stress that q.q.o. is a generalization of quantum convolution (see [26]). Some dynamical properties of quantum convolutions were investigated in [11]. Note that a description of bistochastic Kadison-Schwarz mappings from M2(C) into M2(C) has been provided in [19]. 1 2 FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV 2. Preliminaries In what follows, by M2(C) we denote an algebra of 2 × 2 matrices over complex filed C. By M2(C)⊗ M2(C) we mean tensor product of M2(C) into itself. We note that such a product can be considered as an algebra of 4 × 4 matrices M4(C) over C. In the sequel 1I means an identity matrix, i.e. 1I = (cid:18) 1 0 0 1 (cid:19). By S(M2(C)) we denote the set of all states (i.e. linear positive functionals which take value 1 at 1I) defined on M2(C). Definition 2.1. A linear operator ∆ : M2(C) → M2(C) ⊗ M2(C) is said to be (a) -- a quantum quadratic operator (q.q.o.) if it satisfies the following conditions: (i) unital, i.e. ∆1I = 1I ⊗ 1I; (ii) ∆ is positive, i.e. ∆x ≥ 0 whenever x ≥ 0; (b) -- a Kadison-Schwarz operator (KS) if it satisfies (2.1) ∆(x∗x) ≥ ∆(x∗)∆(x) for all x ∈ M2(C). One can see that if ∆ is unital and KS operator, then it is a q.q.o. A state h ∈ S(M2(C)) is called a Haar state for a q.q.o. ∆ if for every x ∈ M2(C) one has (2.2) Remark 2.2. Note that if a quantum convolution ∆ on M2(C) becomes a ∗-homomorphic map with a condition (h ⊗ id) ◦ ∆(x) = (id ⊗ h) ◦ ∆(x) = h(x)1I. Lin((1I ⊗ M2(C))∆(M2(C))) = Lin((M2(C) ⊗ 1I)∆(M2(C))) = M2(C) ⊗ M2(C) then a pair (M2(C), ∆) is called a compact quantum group [26]. It is known [26] that for given any compact quantum group there exists a unique Haar state w.r.t. ∆. Remark 2.3. Let U : M2(C) ⊗ M2(C) → M2(C) ⊗ M2(C) be a linear operator such that U(x ⊗ y) = y ⊗ x for all x, y ∈ M2(C). If a q.q.o. ∆ satisfies U∆ = ∆, then ∆ is called a quantum quadratic stochastic operator. Such a kind of operators were studied and investigated in [17]. Each q.q.o. ∆ defines a conjugate operator ∆∗ : (M2(C) ⊗ M2(C))∗ → M2(C)∗ by ∆∗(f )(x) = f (∆x), f ∈ (M2(C) ⊗ M2(C))∗, x ∈ M2(C). (2.3) (2.4) One can define an operator V∆ by which is called a quadratic operator (q.c.). Thanks to the conditions (a) (i),(ii) of Def. 2.1 the operator V∆ maps S(M2(C)) to S(M2(C)). V∆(ϕ) = ∆∗(ϕ ⊗ ϕ), ϕ ∈ S(M2(C)), 3. Quantum quadratic operators with Kadison-Schwarz property on M2(C) In this section we are going to describe quantum quadratic operators on M2(C) as well as find necessary conditions for such operators to satisfy the Kadison-Schwarz property. Recall [4] that the identity and Pauli matrices {1I, σ1, σ2, σ3} form a basis for M2(C), where σ1 =(cid:18) 0 1 1 0 (cid:19) σ2 =(cid:18) 0 −i i 0 (cid:19) σ3 =(cid:18) 1 0 −1 (cid:19) . 0 In this basis every matrix x ∈ M2(C) can be written as x = w01I + wσ with w0 ∈ C, w = (w1, w2, w3) ∈ C3, here wσ = w1σ1 + w2σ2 + w3σ3. ON DYNAMICS OF QUADRATIC OPERATORS 3 Lemma 3.1. [24] The following assertions hold true: (a) x is self-adjoint iff w0, w are reals; (b) Tr(x) = 1 iff w0 = 0.5, here Tr is the trace of a matrix x; (c) x > 0 iff kwk ≤ w0, where kwk =pw12 + w22 + w32. Note that any state ϕ ∈ S(M2(C)) can be represented by ϕ(w01I + wσ) = w0 + hw, fi, (3.1) where f = (f1, f2, f3) ∈ R3 with kfk ≤ 1. Here as before h·,·i stands for the scalar product in C3. Therefore, in the sequel we will identify a state ϕ with a vector f ∈ R3. In what follows by τ we denote a normalized trace, i.e. τ (x) = 1 2 Tr(x), x ∈ M2(C), Let ∆ : M2(C) → M2(C) ⊗ M2(C) be a q.q.o. with a Haar state τ . Then one has τ ⊗ τ (∆x) = τ (τ ⊗ id)(∆(x)) = τ (x)τ (1I) = τ (x), x ∈ M2(C), which means that τ is an invariant state for ∆. Let us write the operator ∆ in terms of a basis in M2(C) ⊗ M2(C) formed by the Pauli matrices. Namely, ∆1I = 1I ⊗ 1I; ∆(σi) = bi(1I ⊗ 1I) + Xj=1 b(1) ji (1I ⊗ σj) + Xj=1 ij , bijk ∈ C, (i, j, k ∈ {1, 2, 3}). where bi, b(1) ij , b(2) One can prove the following 3 3 3 b(2) ji (σj ⊗ 1I) + bml,i(σm ⊗ σl), Xm,l=1 i = 1, 2, 3, Theorem 3.2. [20, Proposition 3.2] Let ∆ : M2(C) → M2(C)⊗ M2(C) be a q.q.o. with a Haar state τ , then it has the following form: (3.2) ∆(x) = w01I ⊗ 1I + 3 hbml, wiσm ⊗ σl, Xm,l=1 where x = w0 + wσ, bml = (bml,1, bml,2, bml,3) ∈ R3, m, n, k ∈ {1, 2, 3}. Let us turn to the positivity of ∆. Given vector f = (f1, f2, f3) ∈ R3 put (3.3) β(f)ij = bki,jfk. 3 Xk=1 Define a matrix B(f) = (β(f)ij)3 ij=1. By kB(f)k we denote a norm of the matrix B(f) associated with Euclidean norm in C3. Put S = {p = (p1, p2, p3) ∈ R3 : p2 1 + p2 2 + p2 3 ≤ 1} and denote kBk = sup f∈S kB(f)k. Proposition 3.3. [20, Proposition 3.3] Let ∆ be a q.q.o. with a Haar state τ , then kBk ≤ 1. 4 FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV Let ∆ : M2(C) → M2(C) ⊗ M2(C) be a liner operator with a Haar state τ . Then due to Theorem 3.2 ∆ has a form (3.2). Take arbitrary states ϕ, ψ ∈ S(M2(C)) and f, p ∈ S be the corresponding vectors (see (3.1)). Then one finds that 3 (3.4) ∆∗(ϕ ⊗ ψ)(σk) = bij,kfipj, k = 1, 2, 3. Xi,j=1 Thanks to Lemma 3.1 the functional ∆∗(ϕ ⊗ ψ) is a state if and only if the vector f∆∗(ϕ,ψ) =(cid:18) 3 Xi,j=1 bij,1fipj, 3 3 bij,2fipj, Xi,j=1 Xi,j=1 bij,3fipj(cid:19). satisfies kf∆∗(ϕ,ψ)k ≤ 1. So, we have the following Proposition 3.4. [20, Proposition 4.1] Let ∆ : M2(C) → M2(C) ⊗ M2(C) be a liner operator with a Haar state τ . Then ∆∗(ϕ ⊗ ψ) ∈ S(M2(C)) for any ϕ, ψ ∈ S(M2(C)) if and only if one holds (3.5) ≤ 1 for all f, p ∈ S. 2 3 3 Xk=1(cid:12)(cid:12)(cid:12)(cid:12) bij,kfipj(cid:12)(cid:12)(cid:12)(cid:12) Xi,j=1 From the proof of Proposition 3.3 and the last proposition we conclude that kBk ≤ 1 holds if and only if (3.5) is satisfied. Remark 3.5. Note that characterizations of positive maps defined on M2(C) were considered in [15] (see also [12]). Characterization of completely positive mappings from M2(C) into itself with invariant state τ was established in [24] (see also [16]). Next we would like to recall (see [20]) some conditions for q.q.o. to be the Kadison-Schwarz ones. xm =(cid:0)hbm1, wi,hbm2, wi,hbm3, wi(cid:1), fm = ϕ(σm), αml = hxm, xli − hxl, xmi, Let ∆ : M2(C) → M2(C) ⊗ M2(C) be a linear operator with a Haar state τ , then it has a form (3.2). Now we are going to find some conditions to the coefficients {bml,k} when ∆ is a Kadison-Schwarz operator. Given x = w0 + wσ and state ϕ ∈ S(M2(C)) let us denote (3.6) (3.7) where m, l = 1, 2, 3. Here and what follows [·,·] stands for the usual cross product in C3. Note that here the numbers αml are skew-symmetric, i.e. αml = −αml. By π we shall denote mapping {1, 2, 3, 4} to {1, 2, 3} defined by π(1) = 2, π(2) = 3, π(3) = 1, π(4) = π(1). q(f, w) =(cid:0)hβ(f)1, [w, w]i,hβ(f)2, [w, w]i,hβ(f)3, [w, w]i(cid:1), where β(f)m =(cid:0)β(f)m1, β(f)m2, β(f)m3(cid:1) (see (3.3)) Theorem 3.6. [20, Theorem 3.6] Let ∆ : M2(C) → M2(C) ⊗ M2(C) be a Kadison-Schwarz operator with a Haar state τ , then it has the form (3.2) and the coefficients {bml,k} satisfy the following conditions γml = [xm, xl] + [xm, xl], Denote (3.8) (3.9) 3 kwk2 ≥ i Xm=1 fmαπ(m),π(m+1) + 3 kxmk2 Xm=1 q(f, w) − i (cid:13)(cid:13)(cid:13)(cid:13) 3 Xm=1 (3.10) ON DYNAMICS OF QUADRATIC OPERATORS 5 fmγπ(m),π(m+1) − [xm, xm](cid:13)(cid:13)(cid:13)(cid:13) 3 fkαπ(k),π(k+1) Xk=1 ≤ kwk2 − i Xm=1 kxmk2. − 3 for all f ∈ S, w ∈ C3. Here as before xm =(cid:0)hbm1, wi,hbm2, wi,hbm3, wi(cid:1), bml = (bml,1, bml,2, bml,3) and q(f, w), αml and γml are defined in (3.8),(3.6),(3.7), respectively. Remark 3.7. The provided characterization with [14, 24] allows us to construct examples of positive or Kadison-Schwarz operators which are not completely positive (see next section). Now we are going to give a general characterization of KS-operators. Let us first give some notations. For a given mapping ∆ : M2(C) → M2(C) ⊗ M2(C), by ∆(σ) we denote the vector (∆(σ1), ∆(σ2), ∆(σ3)), and by w∆(σ) we mean the following (3.11) w∆(σ) = w1∆(σ1) + w2∆(σ2) + w3∆(σ3), where w ∈ C3. Note that the last equality (3.11), due to the linearity of ∆, also can be written as w∆(σ) = ∆(wσ). Theorem 3.8. Let ∆ : M2(C) → M2(C) ⊗ M2(C) be a unital ∗-preserving linear mapping. Then ∆ is a KS-operator if and only if one has (3.12) i[w, w]∆(σ) + (w∆(σ))(w∆(σ)) ≤ 1I ⊗ 1I, for all w ∈ C3 with kwk = 1. Proof. Let x ∈ M2(C) be an arbitrary element, i.e. x = w01I + wσ. Then x∗ = w01I + wσ. Therefore Consequently, we have (3.13) (3.14) (3.15) x∗x =(cid:0)w02 + kwk2(cid:1)1I +(cid:0)w0w + w0w − i(cid:2)w, w(cid:3)(cid:1)σ ∆(x) = w01I ⊗ 1I + w∆(σ), ∆(x∗) = w01I ⊗ 1I + w∆(σ) ∆(x∗x) =(cid:0)w02 + kwk2(cid:1)1I ⊗ 1I +(cid:0)w0w + w0w − i(cid:2)w, w(cid:3)(cid:1)∆(σ) ∆(x)∗∆(x) = w021I ⊗ 1I +(cid:0)w0w + w0w)∆(σ) + (w∆(σ))(w∆(σ)) ∆(x∗x) − ∆(x)∗∆(x) = kwk21I ⊗ 1I − i(cid:2)w, w(cid:3)∆(σ) − (w∆(σ))(w∆(σ)). From (3.14)-(3.15) one gets So, the positivity of the last equality implies that kwk21I ⊗ 1I − i(cid:2)w, w(cid:3)∆(σ) − (w∆(σ))(w∆(σ)) ≥ 0. Now dividing both sides by kwk2 we get the required inequality. Hence, this completes the proof. (cid:3) 6 FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV 4. An Example of q.q.o. which is not Kadision-Schwarz one In this section we are going to study dynamics of (5.2) for a special class of quadratic operators. Such a class operators associated with the following matrix {bij,k} given by: b11,1 = ε; b12,1 = 0; b13,1 = 0; b22,1 = 0; b23,1 = ε; b33,1 = 0; b11,2 = 0; b12,2 = 0; b13,2 = ε; b22,2 = ε; b23,2 = 0; b33,2 = 0; b11,3 = 0; b12,3 = ε; b13,3 = 0; b22,3 = 0; b23,3 = 0; b33,3 = ε; and bij,k = bji,k. Via (3.2) we define a liner operator ∆ε, for which τ is a Haar state. In the sequel we would like to find some conditions to ε which ensures positivity of ∆ε. It is easy that for given {bijk} one can find a form of ∆ε as follows (4.1) ∆ε(x) = w01I ⊗ 1I + εω1σ1 ⊗ σ1 + εω3σ1 ⊗ σ2 + εω2σ1 ⊗ σ3 +εω3σ2 ⊗ σ1 + εω2σ2 ⊗ σ2 + εω1σ2 ⊗ σ3 +εω2σ3 ⊗ σ1 + εω1σ3 ⊗ σ2 + εω3σ3 ⊗ σ3, where as before x = w01I + wσ. Theorem 4.1. A linear operator ∆ε given by (4.1) is a q.q.o. if and only if ε ≤ 1 3 . Proof. Let x = w01I + wσ be a positive element from M2(C). Let us show positivity of the matrix ∆ε(x). To do it, we rewrite (4.1) as follows ∆ε(x) = w01I + εB, here ω1 − 2iω3 − ω2 ω3 B =  ω2 + iω1 ω2 + iω1 ω2 − iω1 −ω3 ω1 + ω2 ω2 − iω1 ω1 + ω2 −ω3 ω1 + 2iω3 − ω2 −ω2 − iω1 −ω2 − iω1 −ω2 + iω1 −ω2 + iω1 ω3 ,   where positivity of x yields that w0, ω1, ω2, ω3 are real numbers. In what follows, without loss of generality, we may assume that w0 = 1, and therefore kwk ≤ 1. It is known that positivity of ∆ε(x) is equivalent to positivity of the eigenvalues of ∆ε(x). Let us first examine eigenvalues of B. Simple algebra shows us that all eigenvalues of B can be written as follows λ1(w) = ω1 + ω2 + ω3 + 2qω2 λ2(w) = ω1 + ω2 + ω3 − 2qω2 λ3(w) = λ4(w) = −ω1 − ω2 − ω3 1 + ω2 2 + ω2 3 − ω1ω2 − ω1ω3 − ω2ω3 1 + ω2 2 + ω2 3 − ω1ω2 − ω1ω3 − ω2ω3 Now examine maximum and minimum values of the functions λ1(w), λ2(w), λ3(w), λ4(w) on the ball kwk ≤ 1. One can see that (4.2) λ3(w) = λ4(w) ≤ 3 Xk=1 ωk ≤ √3 3 Xk=1 √3 ωk2 ≤ ON DYNAMICS OF QUADRATIC OPERATORS 7 Note that the functions λ3,λ4 can reach values ±√3 at ±(1/√3, 1/√3, 1/√3). Now let us rewrite λ1(w) and λ2(w) as follows (4.3) (4.4) λ1(w) = ω1 + ω2 + ω3 + λ2(w) = ω1 + ω2 + ω3 − One can see that 2 √2q3(ω2 √2q3(ω2 2 1 + ω2 2 + ω2 3) − (ω1 + ω2 + ω3)2 1 + ω2 2 + ω2 3) − (ω1 + ω2 + ω3)2 (4.5) (4.6) λk(hω1, hω2, hω3) = hλk(ω1, ω2, w3), λ1(hω1, hω2, hω3) = hλ2(ω1, ω2, w3), if h ≥ 0, if h ≤ 0. where k = 1, 2. Therefore, the functions λk(w), k = 1, 2 reach their maximum and minumum on the sphere ω2 3 = 1 (i.e. kwk = 1). Hence, denoting t = ω1 + ω2 + ω3 from (4.5) and (4.4) we introduce the following functions 1 + ω2 2 + ω2 g1(t) = t + 2 √2 √3 − t2, g2(t) = t − 2 √2 √3 − t2 where t ≤ √3. One can find that the critical values of g1 are t = ±1, and the critical value of g2 is t = −1. Consequently, extremal values of g1 and g2 on t ≤ √3 are the following: g1(t) = − √3, max t≤√3 g2(t) = −3, max t≤√3 min t≤√3 min t≤√3 g1(t) = 3, g2(t) = √3. Therefore, from (4.5),(4.6) we conclude that (4.7) − 3 ≤ λk(w) ≤ 3, for any kwk ≤ 1, k = 1, 2. It is known that for the spectrum of 1I + εB one has Therefore, So, if Sp(1I + εB) = 1 + εSp(B) Sp(1I + εB) = {1 + ελk(w) : k = 1, 4} ε ≤ 1 max kwk≤1λk(w) , k = 1, 4 then one can see 1 + ελk(w) ≥ 0 for all kwk ≤ 1 k = 1, 4. This implies that the matrix 1I + εB is positive for all w with kwk ≤ 1. Now assume that ∆ε is positive. Then ∆ε(x) is positive whenever x is positive. This means that 1 + ελk(w) ≥ 0 for all kwk ≤ 1 (k = 1, 4). From (4.2),(4.7) we conclude that ε ≤ 1/3. This completes the proof. (cid:3) Theorem 4.2. Let ε = 1 3 then the corresponding q.q.o. ∆ε is not KS-operator. 8 FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV Proof. It is enough to show the dissatisfaction of (3.10) at some values of w (kwk ≤ 1) and f = (f1, f1, f2). Assume that f = (1, 0, 0), then a little algebra shows that (3.10) reduces to the following one (4.8) where √A + B + C ≤ D A = ε(ω2ω3 − ω3ω2) − iε2(2ω2ω3 − 2ω12 − ω2ω1 + ω1ω2 − ω1ω3 + ω3ω1)2 B = ε(ω1ω2 − ω2ω1) − iε2(2ω1ω2 − 2ω32 − ω1ω3 + ω3ω1 − ω3ω2 + ω2ω3)2 C = ε(ω3ω1 − ω1ω3) − iε2(2ω3ω1 − 2ω22 − ω3ω2 + ω2ω3 − ω2ω1 + ω1ω2)2 D = (1 − 3ε2)(ω12 + ω22 + ω32) −iε2(ω3w2 − ω2ω3 + ω2ω1 − ω1ω2 + ω1ω3 − ω3ω1) Now choose w as follows: Then calculations show that ω1 = − 1 9 ; ω2 = 5 36 ; ω3 = 5i 27 A = C = Hence, we find 9594 19131876 ; B = 1625 3779136 ; D = 19625 ; 86093442 589 17496 . r 9594 19131876 + 19625 86093442 + 1625 3779136 > 589 17496 which means that (4.8) is not satisfied. Hence, ∆ε is not a KS-operator at ε = 1/3. (cid:3) Recall that a linear operator T : Mk(C) → Mm(C) is completely positive if for any positive matrix (cid:0)aij)(cid:1)n i,j=1 is positive for all n ∈ N. Now we are interested when the operator ∆ε is completely positive. It is known [5] that the complete positivity of ∆ε is equivalent to the positivity of the following matrix i,j=1 ∈ Mk(Mn(C)) the matrix (cid:0)T (aij)(cid:1)n ∆ε =(cid:18) ∆ε(e11) ∆ε(e12) ∆ε(e21) ∆ε(e22) (cid:19) here eij (i, j = 1, 2) are the standard matrix units in M2(C). From (4.1) one can calculate that ∆ε(e11) = 1 2 1I ⊗ 1I + εB11, ∆ε(e22) = 1 2 1I ⊗ 1I − εB11 ∆ε(e12) = εB12, ∆ε(e21) = εB∗12 where B11 =  1 2 0 0 −i 0 − 1 0 0 0 − 1 0 0 2 1 0 0 i 2 2   , B12 =  0 1+i 0 0 1+i 0 i 2 0 i 1−i 2 −i −i 2 1−i 2 0 0 0   ON DYNAMICS OF QUADRATIC OPERATORS 9 Hence, we find where 1I8 is the unit matrix in M8(C) and 2 ∆ε = 1I8 + εB B = 1 0 0 2i 0 0 0 1 + i 0 0 −1 0 0 −1 0 0 −2i −2i 0 1 − i 1 − i 0 0 0   1 + i 0 0 0 1 + i −2i 0 0 2i 0 2i 0 1 1 − i −2i −2i 0 1 + i −1 0 0 2i 0 2i 1 0 −2i 0 0 1 0 0 1 − i 0 0 0 2i 0 0 −1   So, the matrix ∆ε is positive if and only if ε ≤ 1 λmax(B) , where λmax(B) = max λ∈Sp(B)λ. One can easily calculate that λmax(B) = 3√3. Therefore, we have the following Theorem 4.3. Let ∆ε : M2(C) → M2(C) ⊗ M2(C) be given by (4.1). Then ∆ε is completely positive if and only if ε ≤ 1 3√3 . 5. Dynamics of ∆ε Let ∆ be a q.q.o. on M2(C). Let us consider the corresponding quadratic operator defined by V∆(ϕ) = ∆∗(ϕ ⊗ ϕ), ϕ ∈ S(M2(C)). From Theorem 3.2 one can see that the defined operator V∆ maps S(M2(C)) into itself if and only if kBk ≤ 1 or equivalently (3.5) holds. From (3.4) we find that 3 (5.1) Xi,j=1 Here as before S = {f = (f1, f2f p3) ∈ R3 : f 2 V∆(ϕ)(σk) = bij,kfifj, f ∈ S. 1 + f 2 2 + f 2 3 ≤ 1}. So, (5.1) suggests us to consider of the following nonlinear operator V : S → S defined by 3 (5.2) V (f)k = bij,kfifj, Xi,j=1 k = 1, 2, 3. where f = (f1, f2, f3) ∈ S. by (5.2) was investigated in [20, Theorem 4.4]. It is worth to mention that uniqueness of the fixed point (i.e. (0, 0, 0)) of the operator given In this section, we are going to study dynamics of the quadratic operator Vε corresponding to ∆ε (see (4.1)), which has the following form (5.3) Let us first find some condition on ε which ensures (3.5). Vε(f )1 = ε(f 2 Vε(f )2 = ε(f 2 Vε(f )3 = ε(f 2 1 + 2f2f3) 2 + 2f1f3) 3 + 2f1f2)   10 FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV Lemma 5.1. Let Vε be given by (5.3). Then Vε maps S into itself if and only if ε ≤ 1√3 satisfied. is Proof. "If" part. Assume that Vε maps S into itself. Then (3.5) is satisfied. Take f = (1/√3, 1/√3, 1/√3), p = f. Then from (3.5) one finds 3 Xi,j=1 3 Xk=1(cid:12)(cid:12)(cid:12)(cid:12) 2 bij,kfipj(cid:12)(cid:12)(cid:12)(cid:12) = 3ε2 ≤ 1 one finds which yields ε ≤ 1/√3. "only if" part. Assume that ε ≤ 1/√3. Take any f = (f1, f2, f3), p = (p1, p2, p3) ∈ S. Then Xk=1(cid:12)(cid:12)(cid:12)(cid:12) bij,kfipj(cid:12)(cid:12)(cid:12)(cid:12) = ε2(f1p1 + f3p2 + f2p32 + f3p1 + f2p2 + f1p32 Xi,j=1 3 3 2 ≤ ε2((f 2 +(p2 +f2p1 + f1p2 + f3p32) 1 + p2 2 + f 2 ≤ ε2(1 + 1 + 1) = 3ε2 ≤ 1. 2 + f 2 2 + p2 1 + f 2 1 + p2 3 )(p2 3)(f 2 3) + (f 2 2 + p2 1 + f 2 3 )) 3 + f 2 2 + f 2 1 )(p2 1 + p2 2 + p2 3) This completes the proof. (cid:3) Remark 5.2. We stress that condition (3.5) is necessary for ∆ to be a positive operator. Namely, from Theorem 4.1 and Lemma 5.1 we conclude that if ε ∈ ( 1 ] then the operator ∆ε is not positive, while (3.5) is satisfied. 3, 1√3 (5.4) In what follows, to study dynamics of Vε we assume ε ≤ 1√3 . Recall that a vector f ∈ S is a fixed point of Vε if Vε(f) = f. Clearly (0, 0, 0) is a fixed point of Vε. Let us find others. To do it, we need to solve the following equation ε(f 2 ε(f 2 ε(f 2   then Vε has a unique fixed point (0, 0, 0) in S. If ε = 1√3 Proposition 5.3. If ε < 1√3 Vε has the following fixed points (0, 0, 0) and (± 1√3 Proof. It is clear that (0, 0, 0) is a fixed point of Vε. If fk = 0, for some k ∈ {1, 2, 3} then due to ε ≤ 1√3 , one can see that only solution of (5.4) belonging to S is f1 = f2 = f3 = 0. Therefore, we assume that fk 6= 0 (k = 1, 2, 3). So, from (5.4) one finds 1 + 2f2f3) = f1 2 + 2f1f3) = f2 3 + 2f1f2) = f3 We have the following ,± 1√3 ,± 1√3 ) in S. then (5.5) Denoting (5.6) f1 f2 f1 f3 f2 f3 2+2f2f3 2+2f1f3 2+2f2f3 2+2f1f2 2+2f1f3 2+2f1f2 = f1 f2 = f1 f3 = f2 f3   x = f1 f2 , y = f1 f3 , z = f2 f3 ON DYNAMICS OF QUADRATIC OPERATORS 11 From (5.5) it follows that (5.7) According to our assumption x, y, z are nonzero, so from (5.7) one gets   1+ 2x x(cid:18) x(cid:0)1+ 2 xy(cid:1) z − 1(cid:19) = 0 xy(cid:1) y(cid:18) y(cid:0)1+ 2 1+2yz − 1(cid:19) = 0 z(cid:18) z(cid:0)1+ 2x z(cid:1) 1+2yz − 1(cid:19) = 0 x(cid:0)1+ 2 xy(cid:1)  y(cid:0)1+ 2 xy(cid:1)  z(cid:0)1+ 2x z(cid:1) y(cid:0)1 + 2x z (cid:1) 1+2yz = 1 1+2yz = 1 x(1 + 2yz) = 1 = 1 1+ 2x z (5.8) (5.9) where 2x 6= −z and 2yz 6= −1. Dividing the second equality of (5.8) to the first one of (5.8) we find which with xz = y yields Simplifying the last equality one gets y + 2x2 = x + 2y2. This means that either y = x or x + y = 1 2 . (y − x)(1 − 2(y + x)) = 0. Assume that x = y. Then from xz = y, one finds z = 1. Moreover, from the second equality y = 1 + 2y. So, y2 + y − 2 = 0 therefore, the solutions of the last one are 2 − y. We note that y 6= 1/2, since x 6= 0. So, from of (5.8) we have y + 2 y1 = 1, y2 = −2. Hence, x1 = 1, x2 = −2. the second equality of (5.8) we find Now suppose that x + y = 1 2 , then x = 1 y + = 1 + 4 1 − 2y 4y2 1 − 2y . So, 2y2 − y − 1 = 0 which yields the solutions y3 = − 1 z3 = − 1 Consequently, solutions of (5.8) are the following ones 2 and x4 = − 1 2 , y4 = 1. Therefore, we obtain x3 = 1, Now owing to (5.6) and we need to solve the following equations 2 , z4 = −2. (1, 1, 1), (cid:0)1,− ( f1 f2 f2 f3 1 2 ,− 1 2 1 2(cid:1), (cid:0) − , 1,−2(cid:1), (−2,−2, 1). = xk, = zk, k = 1, 4. According to our assumption fk 6= 0, therefore we consider cases when xkzk 6= 0. Now let us start to consider several cases: case 1. Let x2 = 1, z2 = 1. Then from (5.9) one gets f1 = f2 = f3. So, from (5.4) we 3ε2 ≤ 1. 3ε . Now taking into account f1 2 ≤ 1 one gets 2 = f1, i.e. f1 = 1 find 3εf1 2 + f2 2 + f3 1 12 FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV . Due to Lemma 5.1 the operator Vε is well defined From the last inequality we have ε ≥ 1√3 iff ε ≤ 1√3 the last ones to (5.4) we get f1 + 3f1 , therefore, one gets ε = 1√3 ;± 1√3(cid:1). case 2. Let x2 = 1, z2 = −1/2. Then from (5.9) one finds f1 = f2, 2f2 = −f3. Substituting 3ε . Taking into account f1 3, due to Lemma 5.1 in this case the operator Vε is not well defined, therefore, we conclude that there is not a fixed point of Vε belonging to S. . Hence, in this case a solution is (cid:0) ± 1√3 3ε , f2 = − 1 9ε2 ≤ 1. This means ε ≥ q 2 2ε = 0. Then, we have f1 = − 1 ;± 1√3 3ε , f3 = 2 2 ≤ 1 we find 1 9ε2 + 4 9ε2 + 1 2 + f3 2 + f2 Using the same argument for the rest cases we conclude the absence of solutions. This shows that if ε < 1/√3 the operator Vε has unique fixed point in S. If ε = 1/√3, then Vε has three fixed points belonging to S. This completes the proof. (cid:3) Now we are going to study dynamics of operator Vε. Theorem 5.4. Let Vε be given by (5.3). Then the following assertions hold true: (i) if ε < 1/√3, then for any f ∈ S one has V n (ii) if ε = 1/√3, then for any f ∈ S with f /∈(cid:8)(± 1√3 as n → ∞. ε (f) → (0, 0, 0) as n → ∞. )(cid:9) one has V n ,± 1√3 ,± 1√3 ε (f) → (0, 0, 0) Proof. Let us consider the following function ρ(f) = f 2 1 + f 2 2 + f 2 3 . Then we have ρ(Vε(f)) = ε2(cid:0)(f 2 ≤ ε2(cid:0)f 2 ≤ ε2(cid:0)f 2 = 3ε2(f 2 1 + 2f2f3)2 + (f 2 1 + 2f2f3 + f 2 3 + f 2 1 + f 2 1 + f 2 3 ) = 3ε2ρ(f) 2 + f 2 2 + f 2 2 + 2f1f3)2 + (f 2 2 + 2f1f3 + f 2 2 + f 2 3 + f 2 1 + f 2 3 + 2f1f2)2(cid:1) 3 + 2f1f2(cid:1) 1 + f 2 2 ) 3 + f 2 This means (5.10) ρ(Vε(f)) ≤ 3ε2ρ(f). Due to ε2 ≤ 1 3 from (5.10) one finds ρ(V n+1 ε (f)) ≤ ρ(V n ε (f)), which yields that the sequence {ρ(V n of {ρ(V n (i). First we assume that ε < 1√3 ε (f))}. ε (f))} is convergent. Next we would like to find the limit , then from (5.10) we obtain ρ(V n ε (f)) ≤ 3ε2ρ(V n−1 ε (f)) ≤ · · · ≤ (3ε2)nρ(f). This yields that ρ(V n (ii). Now let ε = 1√3 1 + f 2 Case (a). Let f 2 ε (f)) → 0 as n → ∞, for all f ∈ S. . Then consider two distinct subcases. 1 + f 2 2 + f 2 3 < 1 and denote d = f 2 1 + 2f2f3)2 + (f 2 3 )2 + (f 2 1 + f 2 2 + f 2 ρ(Vε(f)) ≤ ε2(cid:0)(f 2 ≤ ε2(cid:0)(f 2 = 3ε2d2 = dd = dρ(f). Hance, we have ρ(Vε(f)) ≤ dρ(f). This means ρ(V n n → ∞. 2 + f 2 3 . Then one gets 1 + f 2 2 + 2f1f3)2 + (f 2 3 )2 + (f 2 2 + f 2 3 + 2f1f2)2(cid:1) 2 )2(cid:1) ε (f)) ≤ dnρ(f) → 0. Hence, V n 1 + f 2 3 + f 2 ε (f) → 0 as ON DYNAMICS OF QUADRATIC OPERATORS 13 Case (b). Now take f 2 3 = 1 and assume that f is not a fixed point. Therefore, we may assume that fi 6= fj for some i 6= j, otherwise from Lemma 5.3 one concludes that f is a fixed point. Hence, from (5.3) one finds 2 + f 2 1 + f 2 Vε(f)1 = ε(f 2 1 + 2f2f3) = ε(1 − f 2 2 − f 2 3 + 2f2f3) = ε(1 − (f2 − f3)2). Similarly, one gets Vε(f)2 = ε(1 − (f1 − f3)2), Vε(f)3 = ε(1 − (f1 − f2)2). It is clear that Vε(f)k ≤ ε (k = 1, 2, 3). According to our assumption fi 6= fj (i 6= j) we conclude that one of Vε(f)k is strictly less than 1√3 3 < 1. Therefore, from the case (a), one gets that V n , this means Vε(f)2 2 + Vε(f)2 1 + Vε(f)2 ε (f) → 0 as n → ∞. (cid:3) Acknowledgement The authors acknowledges the MOHE Grant FRGS11-022-0170. The first named author thanks the Junior Associate scheme of the Abdus Salam International Centre for Theoretical Physics, Trieste, Italy. The authors would like to thank to an anonymous referee whose useful suggestions and comments improve the content of the paper. References [1] S.J. Bhatt, Stinespring representability and Kadison's Schwarz inequality in non-unital Banach star algebras and applications, Proc. Indian Acad. Sci. (Math. Sci.), 108 (1998), 283 -- 303. [2] R. Bhatia, C. Davies, More Operator Versions of the Schwarz Inequality, Commun. Math. Phys. 215 (2000), 239 -- 244. [3] R. Bhatia, R. Sharma, Some inequalities for positive linear maps, Linear Alg. Appl. 436 (2012), 1562 -- 1571. [4] O. Bratteli and D. W. Robertson, Operator algebras and quantum statistical mechanics. I, Springer, New York-Heidelberg-Berlin 1979. [5] M.-D. Choi, Completely Positive Maps on Complex Matrices, Linear Algebra Appl. 10 (1975), 285-290. [6] N. N. Ganikhodzhaev, F. M. Mukhamedov, On quantum quadratic stochastic processes, and some ergodic theorems for such processes, Uzb. Matem. Zh. 1997, no. 3, 820. (Russian) [7] N.N. Ganikhodzhaev, F. M. Mukhamedov, Ergodic properties of quantum quadratic stochastic processes, Izv. Math. 65 (2000), 873890. [8] R. Ganikhodzhaev, F. Mukhamedov, U. Rozikov, Quadratic stochastic operators and processes: results and open problems, Inf. Dim. Anal. Quantum Probab. and Related Topics 14(2011), 279-335. [9] U. Groh, The peripheral point spectrum of Schwarz operator on C ∗-algebras, Math. Z. 176 (1981), 311 -- 318. [10] U. Groh, Uniform ergodic theorems for identity preserving Schwarz maps on W ∗-algebras, J. Operator Theory 11 (1984), 395 -- 404. [11] U. Franz, A. Skalski, On ergodic properties of convolution operators associated with compact quantum groups, Colloq. Math. 113 (2008), 13 -- 23. [12] A. Kossakowski, A Class of Linear Positive Maps in Matrix Algebras, Open Sys. & Information Dyn. 10(2003) 213 -- 220. [13] W.A. Majewski, On non-completely positive quantum dynamical maps on spin chains, J. Phys. A: Math. Gen. 40 (2007) 11539-11545. [14] W.A. Majewski, M. Marciniak, On a characterization of positive maps, J. Phys. A: Math. Gen. 34 (2001) 5863-5874. [15] W.A. Majewski, M. Marciniak, On the structure of positive maps between matrix algebras, In book: Noncommutative harmonic analysis with applications to probability, 249 -- 263, Banach Center Publ., 78, Polish Acad. Sci. Inst. Math., Warsaw, 2007. [16] W.A. Majewski, M. Marciniak, Decomposability of extremal positive unital maps on M2(C). In book: M. Bozejko, (ed.) et al., Quantum probability, Banach Center Publications 73(2006), 347 -- 356. 14 FARRUKH MUKHAMEDOV AND ABDUAZIZ ABDUGANIEV [17] F.M. Mukhamedov, On ergodic properties of discrete quadratic dynamical system on C ∗-algebras. Method of Funct. Anal. and Topology, 7(2001), No.1, 63 -- 75. [18] F.M. Mukhamedov, On decomposition of quantum quadratic stochastic processes into layer-Markov pro- cesses defned on von Neumann algebras, Izvestiya Math. 68(2004), No.5. 1009 -- 1024. [19] F. Mukhamedov, A. Abduganiev, On description of bistochastic Kadison-Schwarz operators on M2(C), Open Sys. & Inform. Dynam. 17(2010), 245-253. [20] F. Mukhamedov, H. Akin, S. Temir, A. Abduganiev, On quantum quadratic operators on M2(C) and their dynamics, Jour. Math. Anal. Appl. 376(2011), 641 -- 655. [21] M.A. Nielsen, I.L. Chuang, Quantum Computation and Quantum Information, Cambridge University Press, Cambridge, 2000. [22] A. G. Robertson, A Korovkin theorem for Schwarz maps on C ∗-algebras, Math. Z. 156(1977), 205 -- 206. [23] A. G. Robertson, Schwarz inequalities and the decomposition of positive maps on C ∗-algebras Math. Proc. Camb. Philos. Soc. 94(1983), 291 -- 296. [24] M.B. Ruskai, S. Szarek, E. Werner, An analysis of completely positive trace-preserving maps on M2, Lin. Alg. Appl. 347 (2002) 159-187. [25] E. Stormer, Positive linear maps of operator algebras, Acta Math. 110(1963), 233 -- 278. [26] S.L. Woronowicz, Compact matrix pseudogroups. Comm. Math. Phys. 111 (1987), 613-665. Farrukh Mukhamedov, Department of Computational & Theoretical Sciences, Faculty of Science, International Islamic University Malaysia, P.O. Box, 141, 25710, Kuantan, Pahang, Malaysia E-mail address: [email protected], farrukh [email protected] Abduaziz Abduganiev, Department of Computational & Theoretical Sciences, Faculty of Science, International Islamic University Malaysia, P.O. Box, 141, 25710, Kuantan, Pahang, Malaysia E-mail address: [email protected]
1703.09677
2
1703
2018-06-11T16:15:22
A multiplier algebra functional calculus
[ "math.FA" ]
This paper generalizes the classical Sz.-Nagy--Foias $H^{\infty}(\mathbb{D})$ functional calculus for Hilbert space contractions. In particular, we replace the single contraction $T$ with a tuple $T=(T_1, \dots, T_d)$ of commuting bounded operators on a Hilbert space and replace $H^{\infty}(\mathbb{D})$ with a large class of multiplier algebras of Hilbert function spaces on the unit ball in $\mathbb C^d$.
math.FA
math
A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS KELLY BICKEL, MICHAEL HARTZ, AND JOHN E. MCCARTHY Abstract. This paper generalizes the classical Sz.-Nagy–Foias H ∞(D) functional calculus for Hilbert space contractions. In particular, we replace the single con- traction T with a tuple T = (T1, . . . , Td) of commuting bounded operators on a Hilbert space and replace H ∞(D) with a large class of multiplier algebras of Hilbert function spaces on the unit ball in Cd. 1. Introduction 1.1. Overview. One seminal result connecting operator theory and complex function theory is the classical Sz.-Nagy–Foias H ∞(D) functional calculus [29], which says that for a completely non-unitary contraction T on a Hilbert space K, the polynomial functional calculus (1) ΦT : C[z] → B(K), T 7→ p(T ), extends to a weak-∗ continuous, contractive algebra homomorphism on H ∞(D) (with related statements for general contractions). This allows one to make sense of f (T ) for fairly general functions f and operators T and has both illuminated the structure of C0 contractions and related operators, see [8, 10, 28, 29], and led to breakthroughs related to the invariant subspace problem, see [11, 12]. In this paper, we generalize the Sz.-Nagy–Foias H ∞(D) functional calculus in two nontrivial ways: first, we replace the single contraction T ∈ B(K) with a tuple T = (T1, . . . , Td) of commuting (but not necessarily contractive) bounded operators on K and secondly, we replace H ∞(D) with a multiplier algebra associated to any of a large class of reproducing kernel Hilbert spaces H on the open unit ball Bd ⊆ Cd. Results of this type were proved by J. Eschmeier, who established an H ∞(Bd)- functional calculus for completely non-unitary tuples of commuting operators sat- isfying von Neumann's inequality over the unit ball [19] and by R. Clouâtre and K. Davidson, who studied the Drury-Arveson space H 2 d on Bd and its multiplier algebra Mult(H 2 d ). They proved that completely non-unitary commuting row con- tractions admit a Mult(H 2 d )-functional calculus [13]. This current paper generalizes these earlier investigations and provides new arguments that avoid technical analyses of specific multiplier algebras. 2010 Mathematics Subject Classification. Primary 47A60; Secondary 47A13, 46E22. Key words and phrases. Functional calculus, multiplier algebra, unit ball. K.B. was partially supported by National Science Foundation Grant DMS 1448846. M.H. was partially supported by a Feodor Lynen Fellowship. J.M. was partially supported by National Science Foundation Grant DMS 1565243. 1 2 K. BICKEL, M. HARTZ, AND J. MCCARTHY 1.2. Main Result. Here is our setting: Let T = (T1, . . . , Td) be a tuple of commuting operators on a Hilbert space K and let U = (U1, . . . , Ud) be a spherical unitary on K, i = I. We say T is completely non-unitary if it has i.e. each Ui is normal andP UiU ∗ no non-zero reducing subspace M with T M a spherical unitary. Let H be a reproducing kernel Hilbert space on Bd whose multiplier algebra, Mult(H), contains the polynomials C[z1, . . . , zd]. Thus, the polynomial functional calculus for T , (2) ΦT : C[z1, . . . , zd] → B(K), T 7→ p(T ), is defined on a subalgebra of Mult(H). We are interested in well-behaved extensions of the polynomial functional calculus ΦT , namely ways to make sense of f (T ) for a larger class of functions. Let A(H) denote the norm closure of the polynomials in Mult(H). Then, we say T admits an A(H)-functional calculus if ΦT extends to a completely contractive algebra homomor- phism A(H) → B(K); this loosely says T admits an A(H) version of von Neumann's inequality. Moreover, we say T is Mult(H)-absolutely continuous if in addition ΦT extends to a weak-∗ continuous algebra homomorphism Mult(H) → B(K). Thus, the following two questions naturally arise: (Q1) When does T admit an A(H)-functional calculus? (Q2) If T admits an A(H)-functional calculus, is T Mult(H)-absolutely continuous? If H is a complete Nevanlinna-Pick space, then (Q1) has a concrete answer in terms of the reproducing kernel of H (see Section 5). However, even in the case when H is the Hardy space on the unit ball, so that A(H) is the ball algebra, we are not aware of an explicit answer to (Q1). Nevertheless, Arveson's dilation theorem shows that one can rephrase (Q1) in terms of dilations, and the resulting dilation has a very explicit form if H satisfies some additional assumptions, which we now describe. We say that H is a unitarily invariant space on Bd if the reproducing kernel K of H is of the form ∞ K(z, w) = anhz, win with a0 = 1 and an > 0 for all n ∈ N. Xn=0 If in addition limn→∞ an/an+1 = 1, we say that H is a regular unitarily invariant space. A discussion of this condition can be found in Subsection 2.1. For now, we merely mention that this class of spaces includes many well-known spaces on Bd such as the Bergman space, the Hardy space, the Dirichlet space and the Drury-Arveson space. Then our main result is the following answer to (Q2) for completely non-unitary tuples T : Theorem 1.1. Let H be a regular unitarily invariant space on Bd. Let T = (T1, . . . , Td) be a tuple of commuting operators on a Hilbert space that admits an A(H)-functional calculus. If T is completely non-unitary, then T is Mult(H)-absolutely continuous. This loosely says that if T satisfies an A(H) version of von Neumann's inequality and is completely non-unitary, then we can make sense of f (T ) for all f ∈ Mult(H). A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 3 This theorem partially extends Corollary 1.7 in [19], which establishes an H ∞(Bd)- functional calculus for completely non-unitary tuples which satisfy von Neumann's inequality over the unit ball. Moreover, if H = H 2 d , the Drury-Arveson space, then a theorem of Müller–Vasilescu [24] and Arveson [6] shows that a tuple of commuting operators T admits an A(H)-functional calculus if and only if T is a row contraction, i ≤ I. Thus, in this setting, we recover Theorem 4.3 in [13]. We remark that in this last case, our proof does not require the detailed description of the dual of A(H 2 d ) from [14], which was used in [13]. i=1 TiT ∗ i.e. Pd Even in the one variable setting, Theorem 1.1 appears to be new. For example, consider the one-variable Dirichlet space D. Because the Dirichlet space has an ir- reducible complete Nevanlinna-Pick kernel, there is an explicit characterization (see Theorem 5.1) for when T admits an A(D)-functional calculus, and this characteriza- tion does not require T to be a contraction. Then for such T which are completely non-unitary, Theorem 1.1 says that T is Mult(D)-absolutely continuous. 1.3. Outline of Paper. Recall that the Sz.-Nagy–Foias H ∞(D) functional calculus requires two structural results about contractions: first, each contraction T decom- poses as T = Tcnu ⊕ U, with Tcnu completely non-unitary and U unitary; and second, Sz.-Nagy's dilation theorem, namely each contraction admits a unitary dilation. If T = U is unitary, the spectral theorem implies that U has a weak-∗ continuous H ∞(D) functional calculus if and only if its scalar spectral measure is absolutely continuous with respect to Lebesgue measure. If T = Tcnu is completely non-unitary, then the spectral measure of its minimal unitary dilation is absolutely continuous with respect to Lebesgue measure, from which the general calculus follows [29]. This paper follows the classical outline. In Section 2, we recall some necessary background material on reproducing kernel Hilbert spaces and multivariable operator theory. In particular, Proposition 2.2 notes that every tuple of commuting operators T splits as T = Tcnu ⊕ U, for Tcnu completely non-unitary and U a spherical unitary. Using standard dilation theoretic arguments, we show in Theorem 2.4 that (Q1) can be rephrased in terms of concrete dilations of T for regular unitarily invariant spaces H. Section 3 develops some properties of Mult(H)-Henkin measures, the analogue of measures which are absolutely continuous with respect to Lebesgue measure in our setting. This section closely follows part of the work of Clouâtre and Davidson on Ad- Henkin measures [14]. In particular, adapting their arguments, we extend in Lemma 3.3 a theorem of Henkin by showing that Mult(H)-Henkin measures form a band. Moreover, Lemma 3.4 gives the following answer to (Q2) for spherical unitaries: a spherical unitary U is Mult(H)-absolutely continuous if and only if its scalar spectral measure is Mult(H)-Henkin. In Proposition 3.5, we also provide an answer to (Q2) for operator tuples T in terms of a minimal dilation of T . Section 4 proves our main result, Theorem 1.1. The proof requires the dilation the- orem (Theorem 2.4) as well as Lemma 4.1, which says analytic polynomials are weak-∗ dense in L∞(µ), for µ a totally singular measure. Then Theorem 4.3 also answers (Q2) for commuting operator tuples which may have spherical unitary summands. 4 K. BICKEL, M. HARTZ, AND J. MCCARTHY In Section 5, we consider the case when H is a complete Nevanlinna-Pick space. In this setting, Corollary 5.2 characterizes those operator tuples for which (Q1) and (Q2) have affirmative answers. 2. Preliminaries 2.1. Unitarily invariant spaces. We begin by discussing unitarily invariant repro- ducing kernel Hilbert spaces on the unit ball Bd; in what follows, we only consider the situation d < ∞. Background material on reproducing kernel Hilbert spaces can be found in [3] and [26]. Recall from the introduction that a unitarily invariant space on Bd is a reproducing kernel Hilbert space on Bd whose reproducing kernel K is of the form ∞ K(z, w) = anhz, win with a0 = 1 and an > 0 for all n ∈ N. of convergence of the power series P∞ Moreover, we say that H is a regular unitarily invariant space if in addition, we have limn→∞ an/an+1 = 1. Since H is a space on Bd, it is natural to assume that the radius n=0 antn is 1, so that the limit limn→∞ an/an+1, if it exists, is necessarily equal to 1. Thus, we regard this condition as a regularity condition on the sequence (an). If H is a regular unitarily invariant space, then the polynomials are automatically multipliers of H (see, for example [20, Corollary 4.4 (1)]). Regularity is very useful to us since it guarantees that the tuple (Mz1, . . . , Mzd) is a spherical unitary tuple modulo the compacts (see the proof of Theorem 2.4 below). Examples of regular unitarily invariant spaces are the spaces with reproducing kernels K(z, w) = 1 (1 − hz, wi)α , where α ∈ (0, ∞), which includes in particular the Drury-Arveson space (α = 1), the Hardy space (α = d) and the Bergman space (α = d + 1). A related scale of spaces is given by the kernels ∞ Xn=0 K(z, w) = (n + 1)shz, win, Xn=0 where s ∈ R. Here, the Drury-Arveson space corresponds to the choice s = 0. If s = −1, then we obtain the Dirichlet space on Bd. 2.2. The weak-∗ topology on the multiplier algebra. Let H be a unitarily invariant space on Bd. Since 1 ∈ H, every multiplier ϕ ∈ Mult(H) is uniquely determined by its associated multiplication operator Mϕ. Thus, the assignment ϕ 7→ Mϕ allows us to regard Mult(H) as a unital subalgebra of B(H). It is not hard to see that Mult(H) is closed in the weak operator topology and a fortiori in the weak-∗ topology. Therefore, Mult(H) is a dual space in its own right, namely the dual of B(H)∗/ Mult(H)⊥. We endow Mult(H) with the resulting weak-∗ topology. Since unitarily invariant spaces are separable, so is B(H)∗/ Mult(H)⊥, hence the weak-∗ topology on bounded subsets of Mult(H) is metrizable. Moreover, one easily A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 5 verifies that on bounded subsets of Mult(H), the weak-∗ topology coincides with the topology of pointwise convergence on Bd. 2.3. Decomposition of operator tuples. The proof of [13, Theorem 4.1] implicitly contains the fact that every commuting operator tuple T = (T1, . . . , Td) is the direct sum of a completely non-unitary tuple and a spherical unitary. The authors of [13] attribute the argument to Jörg Eschmeier. For completeness, we repeat the relevant part of the argument, beginning with this lemma: Lemma 2.1. Let U = (U1, . . . , Ud) be a tuple of commuting operators on a Hilbert space. Then U is a spherical unitary if and only if (3) UkU ∗ k = d Xk=1 d Xk=1 U ∗ k Uk = I. Proof. It is obvious that every spherical unitary satisfies (3). Conversely, suppose that U satisfies (3). Then U and U ∗ are spherical isometries. By a theorem of Athavale [7], U therefore extends to a spherical unitary V = (V1, . . . , Vd), say Vk =(cid:20)Uk Ak 0 Bk(cid:21) . k=1 VkV ∗ From Pd k=1 UkU ∗ k = I and Pd 0 I(cid:21) =(cid:20)I +Pd (cid:20)I 0 k = I, we deduce k=1 AkA∗ ∗ k ∗ ∗(cid:21) , so that Ak = 0 for k = 1, . . . , d. Hence, U is the restriction of the spherical unitary V to a reducing subspace, so U itself is a spherical unitary. (cid:3) The desired decomposition of operator tuples follows easily. Note that there are no contractivity assumptions on the operator tuple. Proposition 2.2. Let T be a tuple of commuting operators on a Hilbert space K. Then there exist unique complementary reducing subspaces Kcnu and Ku for T such is a spherical unitary. is completely non-unitary and U := T(cid:12)(cid:12)Ku is a spherical unitary and let Ku be the closed linear span of all elements of U. Then is completely that Tcnu := T(cid:12)(cid:12)Kcnu Proof. Let U be the set of all reducing subspaces M of T with the property that T(cid:12)(cid:12)M Ku is reducing for T , and if we define Kcnu = K ⊖ Ku, then T(cid:12)(cid:12)Kcnu Let U = T(cid:12)(cid:12)Ku . Observe that each M ∈ U is contained in the space non-unitary by definition. d ker I − UkU ∗ Xk=1 U ∗ k Uk! d k! ∩ ker I − Xk=1 k =Pd k=1 UkU ∗ and hence, so is Ku. This shows that Pd spherical unitary by Lemma 2.1. k=1 U ∗ k Uk = IKu, so that U is a 6 K. BICKEL, M. HARTZ, AND J. MCCARTHY u ⊕ K′ Finally, if K = K′ cnu is another decomposition as in the statement, then by definition, K′ cnu ∩ Ku, the tuple T is both spherical unitary and completely non-unitary. Hence, this subspace is trivial, so K′ (cid:3) u ⊂ Ku. On the reducing subspace Ku ⊖ K′ u = Ku and thus, K′ cnu = Kcnu. u = K′ Remark 2.3. According to our definition, an operator tuple T is completely non- unitary if and only if T has no non-zero reducing subspace M with T M a spherical unitary. If T is a row contraction, then it is not hard to check that if T is completely non-unitary in our sense, then it does not even have a non-zero invariant subspace M with T M a spherical unitary. This is the definition used in [13]. 2.4. Dilations. If T = (T1, . . . , Td) is a tuple of commuting operators on a Hilbert space K, and if S = (S1, . . . , Sd) is a tuple of commuting operators on a larger Hilbert space L ⊃ K, then we say that T dilates to S if L decomposes as L = L− ⊕ K ⊕ L+ such that with respect to this decomposition, Sk =  0 ∗ 0 ∗ Tk 0 ∗ ∗ ∗  , for k = 1, . . . , d. Recall that the polynomials are multipliers of every regular unitarily invariant space H. We let Mz denote the d-shift on H, namely the tuple of multiplication operators (Mz1, . . . , Mzd). Abstract dilation theory in the form of Arveson's dilation theorem yields the following result (cf. [6, Section 8]). Theorem 2.4. Let H be a regular unitarily invariant space on Bd and let T be a tuple of commuting operators on a Hilbert space K. Then the following are equivalent: (i) T admits an A(H)-functional calculus. (ii) T dilates to M κ z ⊕ U for some cardinal κ and spherical unitary U. Moreover, if K is separable, then the dilation M κ space. z ⊕ U can be realized on a separable Proof. The implication (ii) ⇒ (i) is clear, since M κ z tautologically admits an A(H)- functional calculus, and U admits an A(H)-functional calculus since the multiplier norm dominates the supremum norm. Conversely, suppose that (i) holds. By Arveson's dilation theorem (see, for example, [25, Corollary 7.7]), the A(H)-functional calculus for T dilates to a ∗-representation π : C ∗(A(H)) → B(L), for some Hilbert space L. A result of Sarason (see, for example, [3, Lemma 10.2]) implies that L decomposes as L = L− ⊕ K ⊕ L+ such that π(Mzk) =  0 ∗ 0 ∗ Tk 0 ∗ ∗ ∗  for k = 1, . . . , d. It remains to show that π(Mz) is unitarily equivalent to M κ z ⊕ U. Since limn→∞ an/an+1 = 1, an application of [20, Theorem 4.6] shows that there exists a short exact sequence of C ∗-algebras 0 −→ K(H) −→ C ∗(A(H)) −→ C(∂Bd) −→ 0, A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 7 where K(H) is the ideal of all compact operators on H, the first map is the inclusion map and the second map sends Mzk to zk for k = 1, . . . , d. In this setting, general re- sults about representations of C ∗-algebras (see, for example, the discussion preceding Theorem 1.3.4 and Corollary 2 in Section 1.4 in [5]) show that π splits as an orthog- onal direct sum π = π1 ⊕ π2, where π1 is unitarily equivalent to a multiple of the identity representation and π2 annihilates the compact operators and hence can be regarded as a representation of C(∂Bd). Consequently, π1(Mz) is unitarily equivalent to M κ z for some cardinal κ, and π2(Mz) is a spherical unitary tuple. The additional claim follows because if H is separable, then the dilation π, which is obtained from Stinespring's dilation theorem, can be realized on a separable space (see the discussion following [25, Theorem 4.1]). (cid:3) In the setting of Theorem 2.4, we say that M κ z ⊕ U ∈ B(L)d is a minimal dilation of T if L is the smallest reducing subspace for M κ z ⊕ U that contains K. If T acts on a separable Hilbert space, then so does every minimal dilation of T . Note that minimial dilations of this type are generally not unique up to unitary equivalence. Indeed, if H = H 2, the Hardy space on the unit disc, and if T = Mz ∈ B(H 2), then Mz itself and the bilateral shift are both minimal dilations of Mz of the form Mz ⊕ U, but they are not unitarily equivalent. 2.5. Normal tuples. We briefly recall a few notions from the spectral theory for commuting normal operators. Let N = (N1, . . . , Nd) be a commuting tuple of normal operators on a Hilbert space K. By the Putnam-Fuglede theorem, the unital C ∗- algebra C ∗(N) = C ∗(N1, . . . , Nd) is commutative. Moreover, the maximal ideal space X of C ∗(N) can be identified with a compact subset of Cd via X → Cd, ρ 7→ (ρ(N1), . . . , ρ(Nd)). The range of this map is called the joint spectrum of N, denoted by σ(N). By the Gelfand-Naimark theorem, C ∗(N) is isomorphic to C(σ(N)) via a ∗-homomorphism that sends Ni to zi for each i = 1, . . . , d. A well-known result about representations of algebras of continuous functions (see [16, Theorem IX.1.14]) shows that there exists a projection-valued spectral measure E on σ(N) such that p(N, N ∗) =Zσ(N ) p(z, z) dE for every p ∈ C[z1, . . . , zd, z1, . . . , zd]. If K is separable, then W ∗(N), the von Neu- mann algebra generated by N, admits a separating vector x ∈ K (see [16, Corollary IX.7.9]), so the scalar-valued measure µ = hE(·)x, xi has the same null sets as E. Any scalar-valued measure with this property is called a (scalar-valued) spectral measure for N. It is unique up to mutual absolute continuity. Moreover, if µ is a scalar- valued spectral measure for N, then the isomorphism between C ∗(N) and C(σ(N)) extends to a ∗-isomorphism between W ∗(N) and L∞(µ) which is a weak-∗-weak-∗ homeomorphism (see [18, Theorem II.2.5]). 8 K. BICKEL, M. HARTZ, AND J. MCCARTHY 3. Henkin measures We begin by recalling a few definitions of measures on the unit sphere which are related to the ball algebra. Background material on this topic can be found in [27, Chapter 9]. Let M(∂Bd) denote the space of complex regular Borel measures on ∂Bd, which can be identified with the dual space of C(∂Bd), the set of continuous functions on ∂Bd. A probability measure µ ∈ M(∂Bd) is called a representing measure for the origin if p(0) =Z∂Bd p dµ holds for every polynomial p (or, equivalently, for every function in the ball algebra). A Borel set E is said to be totally null if it is null for every representing measure of the origin. A measure µ ∈ M(∂Bd) is said to be totally singular if it is singular with respect to every representing measure of the origin. Let H be a unitarily invariant space on Bd such that the polynomials are multipliers of H. Let A(H) denote the norm closure of the polynomials. Since the supremum norm is dominated by the multiplier norm, A(H) is contained in the ball algebra and so, every function in A(H) extends uniquely to a continuous function on Bd. Then, we say that a functional ρ ∈ A(H)∗ is Mult(H)-Henkin if it extends to a weak-∗ continuous functional on Mult(H). We say that a measure µ ∈ M(∂Bd) is a Mult(H)-Henkin measure if the functional ρµ ∈ A(H)∗ defined by ρµ(f ) =Z∂Bd f dµ, f ∈ A(H) is Mult(H)-Henkin. The classical Henkin measures are precisely the H ∞(Bd)-Henkin measures in our terminology. Moreover, Mult(H 2 d )-Henkin measures are called Ad-Henkin in [13]. The following lemma is a straightforward generalization of [19, Lemma 1.1]. In particular, specializing to the case K = C, we see that the notion of Mult(H)-Henkin measures (or functionals) only depends on Mult(H) and not on the particular choice of Hilbert space H. Lemma 3.1. Let H be a unitarily invariant space on Bd such that the polynomials are multipliers of H, let K be a Hilbert space and let Φ : A(H) → B(K) be a bounded linear map. Then the following are equivalent: (i) Φ extends to a weak-∗ continuous linear map from Mult(H) to B(K). (ii) Whenever (pn) is a sequence of polynomials such that pnMult(H) ≤ 1 for all n ∈ N and limn→∞ pn(z) = 0 for all z ∈ Bd, the sequence (Φ(pn)) converges to 0 in the weak-∗ topology of B(K). If Φ is multiplicative, then so is its weak-∗ continuous extension to Mult(H). More- over, the set of Mult(H)-Henkin functionals forms a norm closed subspace of A(H)∗. Proof. The implication (i) ⇒ (ii) is trivial. For the proof of the reverse implication, we first observe that unitary invariance of H implies that there exists a strongly con- tinuous one-parameter unitary group (Ut)t∈R on H defined by (Utf )(z) = f (eitz) for A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 9 f ∈ H and z ∈ Bd. If ϕ ∈ Mult(H), then UtMϕU ∗ t = Mϕ(t), where ϕ(t)(z) = ϕ(eitz). Consequently, if ϕ ∈ Mult(H), then ϕ(t) belongs to Mult(H) with ϕ(t)Mult(H) = ϕMult(H), and the map t 7→ ϕ(t) is continuous in the strong operator topology. Us- ing these two facts, a routine argument involving the Fejér kernel (cf. [23, Lemma I n=0 ϕn is the homogeneous expansion of ϕ as an analytic function on Bd, then the sequence of polynomials (ψn) defined by 2.5]) shows that if ϕ ∈ Mult(H) and if ϕ = P∞ Xj=0 Xk=0 n + 1 ψn = n k 1 ϕj satisfies ψnMult(H) ≤ ϕMult(H) and converges to ϕ in the strong operator topology. With this observation in hand, the proof of [19, Lemma 1.1] carries over to the present setting. Moreover, the first additional claim follows from the fact that multiplication is separately continuous in the weak-∗ topology. To show that the set of Mult(H)-Henkin functionals is norm closed inside of A(H)∗, observe that the argument in the first part of the proof shows that the unit ball of A(H) is weak-∗ dense in the unit ball of Mult(H). Thus, the contraction Mult(H)∗ → A(H)∗, ρ 7→ ρ(cid:12)(cid:12)A(H), restricts to an isometry between the predual of Mult(H) and the set of Mult(H)- Henkin functionals. Since the predual of Mult(H) is complete, it follows that the set of Mult(H)-Henkin functionals is closed in A(H)∗. (Alternatively, this can also be seen by using the characterization (ii) of Mult(H)-Henkin functionals.) (cid:3) Remark 3.2. Since the multiplier norm dominates the supremum norm over Bd, it also follows from Lemma 3.1 that every classical Henkin measure is a Mult(H)-Henkin measure. The converse is false in general. Indeed, there are Mult(H 2 2 )-Henkin mea- sures that are not classical Henkin measures [21]. According to Henkin's theorem [22], see also [27, Theorem 9.3.1], classical Henkin measures form a band. Clouâtre and Davidson showed that Mult(H 2 d )-Henkin mea- sures form a band as well [14, Theorem 5.4]. Their proof generalizes to Mult(H)- Henkin measures. If ν, µ ∈ M(∂Bd) are complex measures, then we write ν ≪ µ if ν ≪ µ. Lemma 3.3. Let H be a unitarily invariant space on Bd such that the polynomials are multipliers of H. Then the Mult(H)-Henkin measures form a band. That is, if µ, ν ∈ M(∂Bd) such that µ is Mult(H)-Henkin and ν ≪ µ, then ν is Mult(H)-Henkin. Proof. While the relevant part of the proof of [14, Theorem 5.4] does generalize to this setting, other parts do not easily generalize. Thus, for clarity, we include the proof here. Given a subset E of ∂Bd, let kf kE denote the value supz∈E f (z). By the Glicksberg-König-Seever decomposition theorem and Henkin's theorem (see Theorem 9.4.4, Theorem 9.3.1 and Section 9.8 in [27]), we can decompose µ = µa +µs, where µa is H ∞(Bd)-Henkin and µs is totally singular. Since every H ∞(Bd)-Henkin measure is Mult(H)-Henkin, the assumption implies that µs = µ − µa is Mult(H)- Henkin. Since ν ≪ µ, by the Radon-Nikodym theorem, there is a function h ∈ L1(µ) 10 K. BICKEL, M. HARTZ, AND J. MCCARTHY such that ν = hµ = hµa + hµs. By Henkin's theorem [27, Theorem 9.3.1], hµa is H ∞(Bd)-Henkin and hence, Mult(H)-Henkin. It therefore suffices to show that hµs is Mult(H)-Henkin. Let ε > 0 and let g be a continuous function on ∂Bd such that g − hL1(µs) < ε. Since µs is totally singular, Rainwater's lemma (see [27, Lemma 9.4.3]) shows that µs is concentrated on an Fσ set which is totally null. Thus, there exists a totally null compact set K ⊂ ∂Bd such that µs − µs(cid:12)(cid:12)KM (∂Bd) < ε/(g∂Bd + 1). By Bishop's theorem (see [27, Theorem 10.1.2]), there exists a function in the ball algebra which agrees with g on K and is bounded by g∂Bd. Thus, there is a polynomial p such that p − gK < ε and p∂Bd ≤ g∂Bd. Clearly, pµs is Mult(H)-Henkin since µs is Mult(H)-Henkin. Moreover, pµs − hµsA(H)∗ ≤ pµs − hµsM (∂Bd) ≤ (p − g)µsKM (∂Bd) + (p − g)(µs − µsK)M (∂Bd) + gµs − hµsM (∂Bd) ≤ µs(K) p − gK + p − g∂Bdµs − µsKM (∂Bd) + g − hL1(µs) ≤ ε(3 + µsM (∂Bd)). Since ε > 0 was arbitrary and since the set of Mult(H)-Henkin functionals is norm closed in A(H)∗ by Lemma 3.1, it follows that integration against hµs is Mult(H)- Henkin, as desired. (cid:3) The spectral measure of a commuting tuple of normal operators (see Subsection 2.5) is only unique up to mutual absolute continuity. Nevertheless, Lemma 3.3 shows that it is meaningful to say that the spectral measure of a commuting spherical unitary is Mult(H)-Henkin. Using Lemma 3.1 and Lemma 3.3, the following result can now be proved just like [13, Lemma 3.1]. Lemma 3.4. Let H be a unitarily invariant space on Bd such that the polynomials are multipliers of H, and let U be a spherical unitary on a separable Hilbert space. Then U is Mult(H)-absolutely continuous if and only if the spectral measure of U is Mult(H)-Henkin. (cid:3) We also obtain an answer to (Q2) in terms of minimal dilations of an operator tuple. Recall from the discussion at the end of Subsection 2.4 that minimal dilations are not necessarily unique. The following result generalizes [13, Theorem 3.2], at least in the separable case. Since we are dealing with dilations as opposed to co-extensions, the proof is somewhat more complicated. Proposition 3.5. Let H be a regular unitarily invariant space on Bd. Let T = (T1, . . . , Td) be a tuple of commuting operators on a separable Hilbert space K and z ⊕ U ∈ B(Hκ ⊕ L) be a minimal dilation of T , where U is spherical unitary. let M κ Then T is Mult(H)-absolutely continuous if and only if the spectral measure of U is Mult(H)-Henkin. A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 11 Proof. If the spectral measure of U is Mult(H)-Henkin, then U is Mult(H)-absolutely continuous by Lemma 3.4. Since M κ is Mult(H)-absolutely continuous and since z p(T ) = PKp(M κ continuous as well. z ⊕U)(cid:12)(cid:12)K for every polynomial p, it follows that T is Mult(H)-absolutely Conversely, suppose that T is Mult(H)-absolutely continuous. Let E denote the projection-valued spectral measure of U and let M = {(x, y) ∈ L ⊕ L : hE(·)x, yi is Mult(H)-Henkin}. We have to show that M = L ⊕ L. For y ∈ L, let Since the map My = {x ∈ L : (x, y) ∈ M}. L → A(H)∗, x 7→ hE(·)x, yi, is linear and continuous, and since the set of all Mult(H)-Henkin functionals forms a norm closed subspace of A(H)∗ by Lemma 3.1, we see that each My is a closed subspace of L. We claim that each My is reducing for U. To this end, note that Z∂Bd p dhE(·)Uix, yi = hp(U)Uix, yi =Z∂Bd pzidhE(·)x, yi and Z∂Bd p dhE(·)U ∗ i x, yi = hp(U)U ∗ i x, yi =Z∂Bd pzidhE(·)x, yi holds for every polynomial p. Clearly, the measures zihE(·)x, yi and zihE(·)x, yi are absolutely continuous with respect to hE(·)x, yi, hence Lemma 3.3 shows that if x ∈ My, then also Uix ∈ My and U ∗ i x ∈ My. Thus, each My is reducing for U. Next, we show that if x, y ∈ K, then (PLx, PLy) ∈ M. Let p be a polynomial. Then Z∂Bd p dhE(·)PLx, PLyi = hp(U)PLx, PLyi = hp(M κ z ⊕ U)x, yi − hp(M κ z )PHκx, PHκyi = hp(T )x, yi − hp(M κ z )PHκx, PHκyi. Since T and M κ is Mult(H)-Henkin, so that (PLx, PLy) ∈ M. z are both Mult(H)-absolutely continuous, we see that hE(·)PLx, PLyi Finally, since M κ z ⊕ U is a minimal dilation of T , it follows in particular that L is the smallest reducing subspace for U that contains PLK. From the preceding two paragraphs, we therefore deduce that if y ∈ PLK, then My = L. Another application of Lemma 3.3 shows that if (x, y) ∈ M, then also (y, x) ∈ M, since hE(·)x, yi = hE(·)y, xi. Thus, for every y ∈ L, the space My contains PLK. Using minimiality of the dilation again, we conclude that My = L for every y ∈ L, so that M = L ⊕ L, as desired. (cid:3) 12 K. BICKEL, M. HARTZ, AND J. MCCARTHY 4. Mult(H)-absolute continuity The goal of this section is to prove our main result, Theorem 1.1. Before proceeding with the proof, we require a lemma about totally singular measures on ∂Bd. It is a well-known consequence of the classical F. and M. Riesz theorem that if µ is a measure on the unit circle which is singular with respect to Lebesgue measure, then the analytic polynomials are weak-∗ dense in L∞(µ) (see, for example, [17, Exercise VI.7.10]). The following lemma is a generalization of this fact to higher dimensions. Lemma 4.1. Let µ be a positive totally singular measure on ∂Bd. Then the analytic polynomials are weak-∗ dense in L∞(µ). Proof. Let f ∈ L1(µ) such that pf dµ = 0 Z∂Bd for all analytic polynomials p. By the Hahn-Banach theorem, we need to show that f = 0 ∈ L1(µ). To this end, observe that the measure f µ annihilates the ball algebra, hence the Cole-Range theorem (see [27, Theorem 9.6.1]) shows that there exists a representing measure ρ for the origin such that f µ ≪ ρ. On the other hand, since µ is totally singular, µ and hence f µ is singular with respect to ρ. Thus, f µ = 0, as desired. (cid:3) Let H be a regular unitarily invariant space on Bd. To prove Theorem 1.1, we will use the dilation of Theorem 2.4 and so, require information about operator tuples of z ⊕ U, where Mz = (Mz1, . . . , Mzd) ∈ B(H)d and U ∈ B(L)d is a spherical the form M κ unitary. Let us assume that L is separable, and let µ ∈ M(∂Bd) be a scalar valued spectral measure of U. By the Glicksberg-König-Seever decomposition theorem [27, Theorem 9.4.4] and Henkin's theorem [27, Theorem 9.3.1] (see also the discussion in [27, Section 9.8]), we can decompose µ = µa + µs, where µa is H ∞(Bd)-Henkin and µs is totally singular. Correspondingly, U = Ua ⊕ Us, where µa is a spectral measure of Ua and µs is a spectral measure of Us. Let La, Ls denote the spaces on which Ua, Us act, respectively. The following lemma plays a key role in the proof of Theorem 1.1. Lemma 4.2. In the setting of the preceding paragraph, let M be the unital weak-∗ closed subalgebra of B(Hκ ⊕ La ⊕ Ls) generated by M κ z ⊕ Ua ⊕ Us, and let Ms = {N ∈ W ∗(Us) : 0 ⊕ 0 ⊕ N ∈ M}. Then Ms is a weak-∗ closed ideal of the von Neumann algebra W ∗(Us). Proof. It is clear that Ms is a weak-∗ closed subspace of W ∗(Us). To show that it is an ideal, let N ∈ Ms. For p ∈ C[z1, . . . , zd], 0 ⊕ 0 ⊕ p(Us)N = p(M κ z ⊕ Ua ⊕ Us)(0 ⊕ 0 ⊕ N) belongs to M and hence, p(Us)N ∈ Ms. The spectral measure µs of Us is totally singular and so, Lemma 4.1 implies that the analytic polynomials are weak-∗ dense in L∞(µs). Recall that W ∗(Us) and L∞(µs) are isomorphic as von Neumann algebras A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 13 via an isomorphism that sends the tuple Us to (z1, . . . , zd). Thus, operators of the form p(Us), where p is an analytic polynomial, are weak-∗ dense in W ∗(Us). Since Ms is weak-∗ closed, we conclude that RN ∈ Ms for all R ∈ W ∗(Us), and by the commutativity of W ∗(Us), NR is also in Ms. Thus, Ms is an ideal. (cid:3) We are now in a position to prove Theorem 1.1. Proof of Theorem 1.1. Let T = (T1, . . . , Td) ∈ B(K)d be a commuting operator tuple on a Hilbert space K that admits an A(H)-functional calculus. Suppose that T is not Mult(H)-absolutely continuous. We will show that T has a spherical unitary summand. In a first step, we observe that it suffices to consider the case where K is separable. Indeed, since T is not Mult(H)-absolutely continuous and since the weak-∗ and weak operator topologies coincide on bounded subsets of B(K), there exist by Lemma 3.1 a sequence of polynomials (pn) in the unit ball of Mult(H) that converges to 0 in the weak-∗ topology of Mult(H) and x, y ∈ K such that the sequence (hpn(T )x, yi) does not converge to 0. Let K0 be the smallest reducing subspace for T that contains x and is not Mult(H)-absolutely continuous. Moreover, has a spherical unitary summand, then so does T . Thus, we may assume that y. Then K0 is separable, and T(cid:12)(cid:12)K0 if T(cid:12)(cid:12)K0 K is separable. By Theorem 2.4, the tuple T dilates to M κ z ⊕U, where κ is a countable cardinal and U is a spherical unitary on a separable Hilbert space. As in the discussion preceding Lemma 4.2, we can decompose U = Ua ⊕ Us. Since T is not Mult(H)-absolutely continuous, Lemma 3.1 guarantees the existence of a sequence of polynomials (pn) in the unit ball of Mult(H) such that (pn) converges to 0 in the weak-∗ topology of Mult(H) but the bounded sequence (pn(T )) does not converge to 0 in the weak-∗ topology of B(K). Since K is separable, the weak-∗ topology on the unit ball of B(K) is metrizable. Then, by passing to a subsequence, we can assume that (pn(T )) actually converges to a non-zero operator A ∈ B(K). Observe that (pn(M κ z )) converges to 0 weak-∗. Moreover, (pn) is bounded in the supremum norm on Bd and converges to 0 pointwise in Bd. Hence, as the spectral measure of Ua is H ∞(Bd)-Henkin, the sequence (pn(Ua)) converges to 0 weak-∗ as well. From pn(T ) = PKpn(M κ z ⊕ Ua ⊕ Us)(cid:12)(cid:12)K and the fact that (pn(T )) converges to the non-zero operator A, we deduce that the bounded sequence (pn(Us)) does not converge to 0 weak-∗. By passing to another subsequence, we can assume that (pn(Us)) converges to a non-zero contraction R ∈ W ∗(Us), so that A = PK(0 ⊕ 0 ⊕ R)(cid:12)(cid:12)K. Let M and Ms be defined as in Lemma 4.2. By our previous arguments, R belongs to Ms. By Lemma 4.2, Ms is a weak-∗ closed ideal of the von Neumann algebra W ∗(Us). Thus, there exists an orthogonal projection P ∈ Ms such that P S = S for all S ∈ Ms (see, for example, [9, Section III.1.1.13]). Since M κ z ⊕ Ua ⊕ Us dilates T , the space K is semi-invariant for M κ z ⊕ Ua ⊕ Us and hence for every operator in M. 14 K. BICKEL, M. HARTZ, AND J. MCCARTHY Therefore, the map M → B(K), S 7→ PKS(cid:12)(cid:12)K, is multiplicative, which we will use repeatedly. Since K is semi-invariant and thus reducing for the self-adjoint operator 0⊕0⊕P ∈ M, we see that Q = PK0⊕0⊕P(cid:12)(cid:12)K is an orthogonal projection. Recall that M is commutative and so, Q and Tk commute for 1 ≤ k ≤ d. Moreover, from P R = R, we deduce that QA = A, which implies Q 6= 0. We finish the proof by showing that the restriction of T to the range of Q is a spherical unitary. Applying Lemma 4.2 again, we see that P (Us)k and P (Us)∗ k belong to Ms for 1 ≤ k ≤ d, hence K is reducing for each operator 0 ⊕ 0 ⊕ P (Us)k. Moreover, for 1 ≤ k ≤ d. In particular, each QTk is the compression of a normal operator to a reducing subspace and thus, is normal. Moreover, QTk = PK(0 ⊕ 0 ⊕ P (Us)k)(cid:12)(cid:12)K d which completes the proof. Xk=1 QTk(QTk)∗ = PK(0 ⊕ 0 ⊕ P )(cid:12)(cid:12)K = Q, (cid:3) Moreover, in the presence of a spherical unitary summand, we obtain the following characterization of Mult(H)-absolute continuity. Theorem 4.3. Let H be a regular unitarily invariant space on Bd, and let T = (T1, . . . , Td) be a commuting tuple of operators on a separable Hilbert space that admits an A(H)-functional calculus. Let T = Tcnu ⊕ U be the decomposition of T from Proposition 2.2. Then T is Mult(H)-absolutely continuous if and only if the spectral measure of U is Mult(H)-Henkin. Proof. This follows immediately from Theorem 1.1 and Lemma 3.4. (cid:3) 5. Complete Nevanlinna-Pick Kernels In this section, we obtain a refined version of Theorem 1.1 when H is a complete Nevanlinna-Pick space. Specifically, we use the fact that in this case, there exists a more checkable condition for T to admit an A(H)-functional calculus. Background material on Nevanlinna-Pick spaces can be found in the book [3]. Let H be a regular unitarily invariant space on Bd with reproducing kernel K(z, w) = n=0 anhz, win. A straightforward generalization of [3, Theorem 7.33] shows that H is an irreducible complete Nevanlinna-Pick space if and only if the sequence (bn) given by P∞ (4) satisfies bn ≥ 0 for all n ≥ 1. bntn = 1 − ∞ Xn=1 1 n=0 antn P∞ A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 15 For example, the class of regular unitarily invariant complete Nevanlinna-Pick spaces on Bd includes the spaces with reproducing kernels K(z, w) = 1 (1 − hz, wi)α for α ∈ (0, 1] and the spaces with kernels ∞ K(z, w) = (n + 1)shz, win Xn=0 for s ≤ 0, and hence in particular the Drury-Arveson space and the Dirichlet space (see Subsection 2.1). Let H be a regular unitarily invariant complete Nevanlinna-Pick space on Bd with kernel K and let (bn) be the sequence of Equation (4). We write 1/K(T, T ∗) ≥ 0 if N Xn=1 bn Xα=n(cid:18)n α(cid:19)T α(T ∗)α ≤ I for all N ∈ N. This definition goes back to work of Agler [1]. For more discussion, see [15, Section 5]. The following result is [15, Theorem 5.4]. Theorem 5.1. Let H be a regular unitarily invariant complete Nevanlinna-Pick space on Bd with kernel K. Let T = (T1, . . . , Td) be a tuple of commuting operators on a Hilbert space. Then the following are equivalent: (i) The tuple T satisfies 1/K(T, T ∗) ≥ 0. (ii) The tuple T admits an A(H)-functional calculus. We remark that if H is not a complete Nevanlinna-Pick space, then one can often still make sense of the condition 1/K(T, T ∗) ≥ 0, see [2, 4]. In general, however, not every tuple which admits an A(H)-functional calculus satisfies 1/K(T, T ∗) ≥ 0, as the example of the Bergman space and the unilateral shift shows. Rather, the condition 1/K(T, T ∗) ≥ 0 is related to the existence of co-extensions, whereas admitting an A(H)-functional calculus is equivalent to the existence of dilations (see Theorem 2.4). In the complete Nevanlinna-Pick setting, there is no difference, see [15] for more discussion. The following refinement of Theorem 4.3 in the complete Nevanlinna-Pick setting is almost immediate. Corollary 5.2. Let H be a regular unitarily invariant complete Nevanlinna-Pick space on Bd with kernel K. Let T = (T1, . . . , Td) be a commuting tuple of operators on a separable Hilbert space and let T = Tcnu ⊕ U be the decomposition of Proposition 2.2. Then the following are equivalent: (i) The tuple T admits an A(H)-functional calculus and is Mult(H)-absolutely continuous. (ii) The tuple T satisfies 1/K(T, T ∗) ≥ 0 and the spectral measure of U is Mult(H)- Henkin. Proof. This follows from Theorem 5.1 and Theorem 4.3. (cid:3) 16 K. BICKEL, M. HARTZ, AND J. MCCARTHY References 1. Jim Agler, The Arveson extension theorem and coanalytic models, Integral Equations Operator Theory 5 (1982), no. 5, 608–631. 2. Jim Agler and John E. McCarthy, Nevanlinna-Pick kernels and localization, Operator theoretical methods (Timişoara, 1998), Theta Found., Bucharest, 2000, pp. 1–20. 3. , Pick interpolation and Hilbert function spaces, Graduate Studies in Mathematics, vol. 44, American Mathematical Society, Providence, RI, 2002. 4. C.-G. Ambrozie, M. Engliš, and V. Müller, Operator tuples and analytic models over general domains in Cn, J. Operator Theory 47 (2002), no. 2, 287–302. 5. William Arveson, An invitation to C ∗-algebras, Springer-Verlag, New York-Heidelberg, 1976, Graduate Texts in Mathematics, No. 39. 6. , Subalgebras of C ∗-algebras. III. Multivariable operator theory, Acta Math. 181 (1998), no. 2, 159–228. 7. Ameer Athavale, On the intertwining of joint isometries, J. Operator Theory 23 (1990), no. 2, 339–350. 8. Hari Bercovici, Operator theory and arithmetic in H ∞, Mathematical Surveys and Monographs, vol. 26, American Mathematical Society, Providence, RI, 1988. 9. Bruce Blackadar, Operator algebras, Encyclopaedia of Mathematical Sciences, vol. 122, Springer- Verlag, Berlin, 2006, Theory of C ∗-algebras and von Neumann algebras, Operator Algebras and Non-commutative Geometry, III. 10. Scott W. Brown and Bernard Chevreau, Toute contraction à calcul fonctionnel isométrique est réflexive, C. R. Acad. Sci. Paris Sér. I Math. 307 (1988), no. 5, 185–188. 11. Scott W. Brown, Bernard Chevreau, and Carl M. Pearcy, Contractions with rich spectrum have invariant subspaces, J. Operator Theory 1 (1979), no. 1, 123–136. 12. 13. Raphaël Clouâtre and Kenneth R. Davidson, Absolute continuity for commuting row contrac- , On the structure of contraction operators II, J. Funct. Anal. 76 (1988), no. 1, 30–55. tions, J. Funct. Anal. 271 (2016), no. 3, 620–641. 14. , Duality, convexity and peak interpolation in the Drury-Arveson space, Adv. Math. 295 (2016), 90–149. 15. Raphaël Clouâtre and Michael Hartz, Multiplier algebras of complete Nevanlinna-Pick spaces: Dilations, boundary representations and hyperrigidity, arXiv:1612.03936 (2016). 16. John B. Conway, A course in functional analysis, second ed., Graduate Texts in Mathematics, vol. 96, Springer-Verlag, New York, 1990. 17. , The theory of subnormal operators, Mathematical Surveys and Monographs, vol. 36, American Mathematical Society, Providence, RI, 1991. 18. Kenneth R. Davidson, C ∗-algebras by example, Fields Institute Monographs, vol. 6, American Mathematical Society, Providence, RI, 1996. 19. Jörg Eschmeier, Invariant subspaces for spherical contractions, Proc. London Math. Soc. (3) 75 (1997), no. 1, 157–176. 20. Kunyu Guo, Junyun Hu, and Xianmin Xu, Toeplitz algebras, subnormal tuples and rigidity on reproducing C[z1, . . . , zd]-modules, J. Funct. Anal. 210 (2004), no. 1, 214–247. 21. Michael Hartz, Henkin measures for the Drury-Arveson space, Michigan Math. J. (to appear), arXiv:1701.07777. 22. G. M. Henkin, The Banach spaces of analytic functions in a ball and in a bicylinder are non- isomorphic, Funkcional. Anal. i Priložen. 2 (1968), no. 4, 82–91. 23. Yitzhak Katznelson, An introduction to harmonic analysis, third ed., Cambridge Mathematical Library, Cambridge University Press, Cambridge, 2004. 24. V. Müller and F.-H. Vasilescu, Standard models for some commuting multioperators, Proc. Amer. Math. Soc. 117 (1993), no. 4, 979–989. A MULTIPLIER ALGEBRA FUNCTIONAL CALCULUS 17 25. Vern Paulsen, Completely bounded maps and operator algebras, Cambridge Studies in Advanced Mathematics, vol. 78, Cambridge University Press, Cambridge, 2002. 26. Vern I. Paulsen and Mrinal Raghupathi, An introduction to the theory of reproducing kernel Hilbert spaces, Cambridge Studies in Advanced Mathematics, Cambridge University Press, 2016. 27. Walter Rudin, Function theory in the unit ball of Cn, Classics in Mathematics, Springer-Verlag, Berlin, 2008, Reprint of the 1980 edition. 28. Béla Sz.-Nagy and Ciprian Foias, Sur les contractions de l'espace de Hilbert. vii. triangulations canoniques. fonction minimum, Acta Sci. Math. 25 (1964), 12–37. 29. Béla Sz.-Nagy, Ciprian Foias, Hari Bercovici, and László Kérchy, Harmonic analysis of operators on Hilbert space, second ed., Cambridge Mathematical Library, Universitext Springer, New York, 2010. Department of Mathematics, Bucknell University, 701 Moore Ave, Lewisburg, PA 17837, USA E-mail address: [email protected] Department of Mathematics, Washington University in St. Louis, One Brookings Drive, St. Louis, MO 63130, USA E-mail address: [email protected] Department of Mathematics, Washington University in St. Louis, One Brookings Drive, St. Louis, MO 63130, USA E-mail address: [email protected]
1112.2329
1
1112
2011-12-11T07:38:42
To the Many-Hilbert-Space Theory of Quantum Measurements
[ "math.FA" ]
In this work, a connection between some spectral properties of direct integral of operators in the direct integral of Hilbert spaces and their coordinate operators has been investigated.
math.FA
math
To the Many-Hilbert-Space Theory of Quantum Measurements by Z.I.ISMAILOV and E.OTKUN ÇEVIK Karadeniz Technical University, Faculty of Sciences, Department of Mathematics 61080 Trabzon, TURKEY e-mail address : [email protected] Abstract: In this work, a connection between some spectral properties of direct integral of operators in the direct integral of Hilbert spaces and their coordinate operators has been investigated. Keywords: Direct integral of Hilbert spaces and operators; spectrum and resolvent sets; compact operators;Schatten-von Neumann operator classes; power and polynomially bounded operators. 2000 AMS Subject Classification: 47A10;81Q35. 1.Introduction It is known that the general theory of linear closed operators in Hilbert spaces and its applications to physical problems have been investigated by many researchers(for example,see ).But many physical problems require studying the theory of linear operators in direct sums or in general direct integrals of Hilbert spaces. The concepts of direct integral of Hilbert spaces and direct integral of operators as a generalization of the concept of direct sum of Hilbert spaces and direct sum of operators were introduced to mathematics and developed in 1949 by John von Neumann in his work .These subjects were incorporated in several works(see .A spectral theory of some operators on a finite sum of Hilbert spaces was investigated by N.Dunford .Note that,in terms of application ,there are some results in papers in the finite sum cases .Also in the infinite direct sum cases there is a work ,in which some spectral and compactness properties are surveyed. Furthermore,some spectral investigations of the direct integral of operators in the direct integral of Hilbert spaces have been provided by T.R.Chow ,T.R.Chow,F.Gilfeather , E.A.Azoff L.A.Fialkow . It must be noted that the theory of direct integral of Hilbert spaces and operators on the these spaces has important role in the representation theory of locally compact groups,in the theory of decomposition rings of operators to factors,invariant measures,reduction theory,von Neumann algebras and ets.On the other hand, many physical problems of today 1 arising in the modelling of processes of multiparticle quantum mechanics , quantum field theory and in the physics of rigid bodies support to study a theory of direct integral of operators in the direct integral of Hilbert spaces (see [19] and references in it). Numereous scientific investigations have been done to explain quantum measurements .Dealing with these subjects, S.Machida and M.Namiki (see, also have offered many-Hilbert- space theory lately.Also,they find that decoherence of the wave function is only necessity to formulate quantum mechanical measurements. A direct integral space of continiously many Hilbert spaces and a continious superselection rule is starring in their theory.The direct integral space structure is assigned to the measurement devise which reflects its macroscopic features.On the other hand,observed system is left to be defined by a single Hilbert space.Furthermore,they have examined double slit experiments and negative result experiments in the framework of the many-Hilbert-space theory.Note that a direct integral space of continiously many Hilbert spaces often arises in the quantum version of Lax-Phillips theory .In this investigation the direct integral space is introduced in order to the allow the generator of motion to have a spectrum in real axis ,which is a necessary condition for the application of the Lax-Phillips theory. In second section of this paper a connection between parts of spectrum, resolvent set of direct integral of operators defined in the direct integral of Hilbert spaces and parts of the spectrum of ‘’coordinate operators ‘’has been established . Note that the another approach to analogous problem has been used in work . In this present paper sharp formulas for the connection are given. In third section the compactness properties of these operators have been examined.Finally,in fourth section in special case the analogous questions for the power and polynomially bounded operators have been provided. In the special case of direct sum of Hilbert spaces, these questions have been investigated in Along this paper the triplet be any measure space and the Hilbert spaces are looked at will become infinite dimensional. In addition ,the space of compact operators and Schatten-von Neumann classes in any Hilbert space will be denoted by and respectively. 2. On the spectrum of direct integral of operators In this section, the relationship between the spectrum and resolvent sets of the direct integral of operators and its coordinate operators will be investigated. Before of all prove the following result. Theorem 2.1.For the operator are true , in the Hilbert space = 2 Proof. For any there exist element such that and .Then almost everywhere with respect to measure µ it is true Since ,then there exist which satisfy the above equality and . This means that Hence and from this it is obtained that The proof of the second proposition is clear. Actually ,in one special case the following stronger assertions are true. Theorem 2.3. Assume that every one-point set is measurable and its measure is positive. For the parts of spectrum and resolvent sets of the operator in Hilbert space = ; the following claims are true ; ; ; Proof.Here only the first and second relations of theorem will be prove. The validity of other propositions can be proved by the similar ideas. such that .So Assumed that . Then there exist .Hence ) for some for every and .Then On the contrary,assumed that and from this is true at least one index is hold ,i.e. for some .In this case for the element and we have .Consequently, . Now we prove the second relation on the continuous spectrum.Let . In this case by the is a one-to-one operator, definition of continuous spectrum and is dense in . From this and definition of direct integral it implies that for every and an operator is a one-to-one operator in , is dense in ,then . Since or . This means that 3 It is easy to prove the inverse implication. On the other hand the simple calculations show that the following relations are true. Corollary 2.4. Under the assumptions of last theorem we have , . Corollary 2.5.Let Λ be any countable set , = and be any measure with property for every point .In this case the formulas , , , . are true. Note that when is counting measure the analogous results have been established in work 3. SOME COMPACTNESS PROPERTIES of DIRECT INTEGRAL of OPERATORS In this section the compactness and spectral properties between direct integral of operators and their ‘’coordinate operators’’have been established.In general ,there is not any relation between mentioned operators in compactness means. Example3.1. -counting measure, . In this case for every ) ,but Example 3.2.In some cases from the relations no implies ) for every . Indeed, from the definition of direct integral of operators on the set having null µ-measure the coordinate operators may be defined by arbitrary way. But in certain situations there are concrete results. 4 Theorem 3.3. Let Λ be any countable set , = and be any measure with property for every point .Then ,then for every , . (1) If (2)Let Λ infinite countable set and for every .In this case if and only if This theorem is proved by analogous scheme of the proof in theorem 4.6 in . Now give one characterizating theorem on the point spectrum of compact direct integral of operators. Theorem 3.4.Assumed that in the Hilbert space = , and .In this case there exist countable subset of Λ such that the set is minimal and . From the definition of singular number s(.) (or characteristic numbers ) of any compact operator in any Hilbert space and Theorems 2.1 and 3.4 it is easy to prove the validity of the following result. Theorem 3.5. Assumed that in the Hilbert space = , and .In this case there exist countable subset of Λ such that ; (2) İf then for every , ; (3) Let .Then if and only if the series =1 =1 ( ) converges. (4) If and the series is convergent,then (5) If and sup ,then for every ( ). (6) If , sup and for some , ,then for every . Proof.The validity of the claims (1) and (2) are clear.Prove third assertion of theorem.If the operator ,then the series is convergent.In this case by the first proposition of this theorem and important theorem on the convergence of the rearrangement series it is obtained that the series is convergent. 5 On the contrary, if the series is convergent ,then the series being a rearrangement of the above series,is also convergent.So . Now prove (4).If for every then from the inequality and first claim the validity of this assertion is clear.Now consider the general case.In this case the operator can be written in form . Then On the other hand ,since and , then from the (3) of this theorem it implies that with . Therefore,by the important theorem of the operator theory . Furthermore ,by using proposition (2) of this theorem it is easy to prove the claim(5).On the other hand, the claim(6) is one of the corollary of (5). Remark 3.6.Note that for the some in representation Theorem 3.4. may be hold in In these situations corresponding conditions for such index in the Theorem 3.5(3-6) may be omitted ,for example,as in the following assertion. Theorem 3.7. Assumed that in the Hilbert space = , and .In this case there exist countable subset of Λ such that If , , and ,then . 6 4.Power and Polynomially Boundednessity of the Direct Sum Operators In this section let us and is the counting measure .Here a connection of power (and polynomially ) boundedness property of the direct sum operators in the direct sum Hilbert spaces and its coordinate operators were established.In advance,give some necessary definitions for the later. Definition 4.1 .Let ℋ be any Hilbert space. (1).An operator ℋ is called power bounded ( ℋ if there exist a constant such that for any is satisfied (3.1) (2)Operator ℋ polynomially bounded ℋ ,if there exist a constant such that for any polynomial is satisfied , (3.2) where =sup . (3) The smallest number satisfying (3.1) (resp.(3.2)) is called the power bound (resp.polynomial bound) of the operator and will be denoted by (resp. . Before of all note that the following theorem is true. , then for every , . Theorem 4.2. If This result is a one of the corollary of following equality . In general,the inverse of last assertion may be not true. Example 4.3. Let , , , , . Consequently, It is easy to see that and = . Hence,for any ,but Example 4.4. Let , , and , ,i.e.for every , ,but In this case and = Therefore, . Actually,it is true the following result . Theorem 4.5. if and only if and . Proof.If ,then from the following relation 7 for each .From this it is determined that for any it is implies that . On the other hand,it is clear that for each Therefore, . and On the contrary,if for any , ,then from the equality it is obtained that . Now polynomially boundness property of the direct sum operators will be investigated.In advance,note that the following proposition is true. , and ,then for every , Theorem 4.6. If . Unfortunately,the inverse of last theorem may be not true in general. Example 4.7. Let , , It is known that and is a nilpotent operator with power of nilpotency 2 ,for we have any .In this case for any polynomial function . In the other words ,for every , .Unfortunately,for the polynomial we have i.e. . But in general case the following result is true. , and .In order to the Theorem 4.8. Let necessary and sufficient conditions are and . Proof. Assumed that for every polynomial p(.) .and . In this case since for every polynomial function ),then .From last relation it is obtained that .Hence . Now let us ,i.e.for any and polynomial p(.) it is valid 8 . Then it is clear that , .In the other hand , from last equality it implies that for every is hold .Hence, .This completes the proof of the theorem. Acknowledgment The authors are grateful to G.İsmailov(student of Trabzon Kanuni Anadolu High School) for his helping suggestion to english version and other technical discussion. REFERENCES some sum of Hilbert spaces and 1. Dunford, N., Schwartz, J. T., Linear Operators, I , II,Interscience, New York, 1958,1963. 2. Rofe-Beketov F.S.,Kholkin A.M.,Spectral Analysis of Differential Equations,First ed.,World Scientific Monograph Series in Mathematics,New Jersey,Vol.7,2005. 3. von Neumann,J.,On rings of operators.Reduction theory,Ann.Math.,50,1949,pp.401-485. 4. Dixmier,J.,Les algebres d’operateurs dans l’espace Hilbertien,Geauther-Villars,Paris,1957. 5. Schwartz,J.T., algebras,Gordon and Breach,New York ,1967. 6. Mackey,G.W.,The theory of unitary group representations,University of Chicago Press,1976. 7. Naimark,M.A.,Fomin,S.V.,Continuous direct sum of Hilbert applications,Uspehi Mat.Nauk,10,2(64),1955,111-142(in Russian). 8. Dunford,N.,Spectral operators in a direct spaces,Mathematics,v.50,1963,p.1041-1043. 9. Dunford,N.,A spectral theory for certain operators on direct sum spaces,Mat.Ann,162,1966,pp.294-330. 10. Sokolov,M.S.,An abstract approach to some spectral problems of direct sum differential operators,Electr.J.Diff.Equat.,v.2003(2003),no.75,pp.1-6. 11. Zettl, A., Sturm-Liouville Theory, Amer. Math. Soc., Math. Survey and Monographs vol. 121, USA, 2005. 12. Ismailov, Z.I., Multipoint Normal Differential Operators for First Order, Opusc. Math., 29, 4, (2009), 399-414. 13. Ismailov Z.I.,Otkun Çevik E.,Unluyol E.,Compact inverses of multipoint normal differential operators for first order,Electr.J.Diff.Equat.,v.2011(2011),no.89,pp.1-11. 14. Chow ,T.R.,A spectral theory for the direct sum operators,Math.Ann.188(1970),285-303. 15. Chow of ,T.R.,Gilfeather,Functions operators,Proc.Amer.Math.Soc.,v.29,2,1971 ,pp.325-330. 16. Azoff,E.A.,Spectrum and direct integral,Trans.Amer.Math.Soc.,197(1974),211-223. 17. Azoff,E.A.,Clancey,K.F.,Spectral multiplicity of integral direct for operators,J.Operator theory,3,1980,pp.213-235. of Hilbert integral of integral direct of normal 9 of in Quantum quasinilpotent sum direct of note 18. Fialkow,L.A.,A operators,Proc.Amer.Math.Soc.,v.48.1,1975,pp.125-131. 19. Albeverio,S.,Gesztesy,F.,Hoegh-Krohn,R.and Holden,H.,Solvable Models Mechanics,New-York,Springer,1988. 20. Machida ,S.,Namiki,M.,Theory of measurements in quantum-mechnics-mechanism of reduction of the wave packet,Prog.Theor.Phys.63(5),1980,pp.1457-1473. 21. Namiki,M.,Many-Hilbert-Space Theory of quantum measurements,Foundations of physics,v.18,1,1988,pp,29-55. 22. Namiki.M.,Pascazio,S.,Superselection rules and fluctuations in the Many-Hilbert-Space approach to quantum measurements ,Found.Phys.Lett.,4(3),1991,pp.203-???. 23. Wheeler J.A.,Zurek,W.H.,Quantum theory and measurements,Princeton University Press,1993. 24. Lax,P.D.,Phillips,R.S.,Scattering theory,Academic Press,New York,1967. 25. Halmos,P.R..,A Hilbert space problem book,Springer-Verlag,New York,1982. 26. Ionascu,E.J.,On power bounded operators,Proc.Amer.Math.Soc.,125,5,1997,pp.1435-1441. 10
1705.09861
2
1705
2018-06-25T21:59:21
Binary frames with prescribed dot products and frame operator
[ "math.FA" ]
This paper extends three results from classical finite frame theory over real or complex numbers to binary frames for the vector space ${\mathbb Z}_2^d$. Without the notion of inner products or order, we provide an analog of the "fundamental inequality" of tight frames. In addition, we prove the binary analog of the characterization of dual frames with given inner products and of general frames with prescribed norms and frame operator.
math.FA
math
BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR VERONIKA FURST AND ERIC P. SMITH Abstract. This paper extends three results from classical finite frame theory over real or complex numbers to binary frames for the vector space Zd 2. Without the no- tion of inner products or order, we provide an analog of the "fundamental inequality" of tight frames. In addition, we prove the binary analog of the characterization of dual frames with given inner products and of general frames with prescribed norms and frame operator. 1. Introduction Due to applications in signal and image processing, data compression, sampling theory, and other problems in engineering and computer science, frames in finite- dimensional spaces have received much attention from pure and applied mathemati- cians alike, over the past thirty years (see, for example, Chapter 1 of [6]). The redundant representation of vectors inherent to frame theory is central to the idea of efficient data storage and transmission that is robust to noise and erasures. Frames for Cd and Rd have been extensively studied (see [9] for a standard intro- duction to frame theory, [16] for applications, and [13] for an exposition at the un- dergraduate level). Noting the similarity between frames and error-correcting codes, Bodmann, Le, Reza, Tobin, and Tomforde ([3]) introduced the concept of binary frames, that is, finite frames for the vector space Zd 2. Binary Parseval frames robust to erasures were characterized in [2], and their Gramian matrices were studied in [1]. A more generalized approach to binary frames was taken in [15]. We begin with a brief introduction to classical frame theory terminology. Let Hd denote the Hilbert space Cd or Rd. Definition 1.1. A (finite) frame for a Hilbert space Hd is a collection {xj}K j=1 of vectors in Hd for which there exist finite constants A, B > 0 such that for every y ∈ Hd, Akyk2 ≤ KXj=1 hy, xji2 ≤ Bkyk2. The constants A and B are known as frame bounds. An A-tight frame is one for which A = B, and a Parseval frame is one for which A = B = 1. The vectors xj in the above definition need not be orthogonal or even linearly independent. An orthonormal basis is most closely resembled by a Parseval frame, for which we have the (not necessarily unique) reconstruction formula: 1 2 V. FURST AND E. P. SMITH Proposition 1.2. A collection of vectors {xj}K Hd is a Parseval frame for Hd if and only if j=1 in a finite-dimensional Hilbert space y = KXj=1 hy, xji xj for each y ∈ Hd. Definition 1.3. Let {xj}K The corresponding frame operator S : Hd → Hd is defined by j=1 be a frame for the finite-dimensional Hilbert space Hd. S(x) = KXj=1 hx, xji xj. It can be seen as the composition S = ΘΘ∗ of the synthesis operator Θ : CK → Hd and its adjoint, the analysis operator Θ∗ : Hd → CK, given by the formulas Θ     c1 c2 ... cK     = KXj=1 respectively. cjxj and Θ∗(x) =   hx, x1i hx, x2i ... hx, xKi   , The frame operator is a bounded, invertible, self-adjoint operator satisfying AId ≤ S ≤ BId. Here and in what follows, we use Id to denote the d × d identity matrix and 0d to denote the d × d zero matrix. A frame is Parseval if and only if its frame operator is the identity operator. From both a pure and an applied point-of-view, construction of frames with desired properties has been a central question ([4]). In particular, much attention has been paid to tight frames with prescribed norms and general frames with both prescribed norms and frame operator. In the case of tight frames, the answer, the so-called "fundamental frame inequality," was provided by Casazza, Fickus, Kovacevi´c, Leon, and Tremain: Theorem 1.4 ([7], Corollary 4.11). Given real numbers a1 ≥ a2 ≥ . . . ≥ aK > 0, K ≥ d, there exists a λ-tight frame {xj}K j=1 for a d-dimensional Hilbert space Hd with prescribed norms kxjk2 = aj for 1 ≤ j ≤ K if and only if λ = 1 d KXj=1 aj ≥ a1. Casazza and Leon generalized this result to frames with prescribed frame operators (the classical case is when S = λId): Theorem 1.5 ([8], Theorem 2.1). Let S be a positive self-adjoint operator on a d- dimensional Hilbert space Hd with eigenvalues λ1 ≥ λ2 ≥ . . . ≥ λd > 0. Given real BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 3 numbers a1 ≥ a2 ≥ . . . ≥ aK > 0, K ≥ d, there is a frame {xj}K operator S and kxjk2 = aj for all 1 ≤ j ≤ K if and only if j=1 for Hd with frame KXj=1 aj = dXj=1 λj and kXj=1 aj ≤ kXj=1 λj for every 1 ≤ k ≤ d. This can be seen as a consequence of the classical Schur-Horn theorem ([4]). In [5], Cahill, Fickus, Mixon, Poteet, and Strawn introduce a so-called eigenstep method for constructing all frames with a given frame operator and set of norms (see also [12], and [4] for a survey of the topic). A different approach was taken by Christensen, Powell, and Xiao in [10], extending j=1, a sequence Theorem 1.4 to the setting of dual frame pairs. Given a frame {xj}K {yj}K j=1 is called a dual frame if for every y ∈ Hd, y = KXj=1 hy, xji yj = KXj=1 hy, yji xj. Theorem 1.6 ([10], Theorem 3.1). Given a sequence of numbers {aj}K K > d, the following are equivalent: (1) There exist dual frames {xj}K j=1 and {yj}K j=1 for Hd such that hxj, yji = aj for j=1 ⊂ H with all 1 ≤ j ≤ K. (2) There exists a tight frame {xj}K j=1 and dual frame {yj}K j=1 for Hd such that hxj, yji = aj for all 1 ≤ j ≤ K. (3) PK j=1 aj = d. The goal of this paper is to extend the theory of frames with prescribed norms (or inner products) from the classical Hilbert spaces of Cd and Rd to the binary space Zd 2. We provide analogs of Theorems 1.4, 1.5, and 1.6 for binary frames. The challenge, of course, is the lack of an inner product, positive elements, and guaranteed eigenvalues. Section 2 contains background material on binary frames. In Section 3, we explore dual binary frames and prove the binary version of Theorems 1.6 and 1.4. In Section 4, we construct binary frames with prescribed "norms" and frame operator, as an analog to Theorem 1.5. We conclude in Section 5 with examples and a catalog. 2. Binary Frames In [3], Bodmann, Le, Reza, Tobin, and Tomforde introduce a theory of frames over the d-dimensional binary space Zd 2 is the direct product Z2 ⊕ · · · ⊕ Z2 having d ≥ 1 copies of Z2. The main trouble in defining frames in a binary space stems from the lack of an ordering on Z2. Without an order, there can be no inner product defined for binary space. In spite of this, [3] establishes the dot product as the analog of the inner product on Rd and Cd. 2 where Zd 4 V. FURST AND E. P. SMITH Definition 2.1 ([3]). The dot product on Zd given by 2 is defined as the map (·, ·) : Zd 2×Zd 2 → Z2 (a, b) = dXn=1 a[n]b[n]. Due to the degenerate nature of the dot product (note that (a, a) = 0 need not imply a = 0), it fails to help define a frame in the manner of Definition 1.1. However, when working over finite-dimensional spaces in the classical case, a frame is merely a spanning sequence of vectors. This motivates the definition of a frame in binary space. Definition 2.2 ([3]). A frame is a sequence of vectors F = {fj}K span(F ) = Zd 2. j=1 in Zd 2 such that The synthesis, analysis, and frame operators of F are defined similarly as in Defi- nition 1.3 and are denoted ΘF , Θ∗ F , and SF , respectively. Definition 2.3 ([3]). The synthesis operator of a frame F = {fj}K matrix whose ith column is the ith vector in F . The analysis operator Θ∗ transpose of the synthesis operator. Explicitly, j=1 is the d × K F is the ΘF =  f1 · · · fK   and Θ∗ F =  - f ∗ 1 - ... - f ∗ K -   . The frame operator is SF = ΘF Θ∗ F . It is demonstrated in [3] that the spanning property of F is necessary and sufficient for F to have a reconstruction identity with a dual family G. This fact is summed up in the following theorem and is shown by choosing a basis consisting of d vectors in F (without loss of generality, assumed to be f1, . . . , fd) and applying the Riesz Representation Theorem to the linear functionals γi defined by γi(fj) = δij. Theorem 2.4 ([3], Theorem 2.4). The family F = {fj}K only if there exist vectors G = {gj}K j=1 such that for all y ∈ Zd 2, j=1 in Zd 2 is a frame if and (1) y = KXj=1 (y, gj)fj. In the proof, gi is defined as the unique vector satisfying γi(y) = (y, gi) for every y for 1 ≤ i ≤ d, and gi = 0 for d < i ≤ K. Equation (1) can be rewritten as ΘF Θ∗ G = Id, which is equivalent to ΘGΘ∗ F = Id. Consequently, G is a dual frame to F . We will refer to the dual frame G as a natural dual to F . Note that this definition is unrelated to the usual definition of the canonical dual in Cd or Rd as {S −1(fj)}, where S = ΘΘ∗ is the BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 5 frame operator from Definition 1.3. Although SF is no longer necessarily invertible, we still have dXj=1 dXj=1 (gi, fj)fj = δijfj = fi for i ≤ d. Propositions 2.5 and 2.6 make clear that the natural dual frame is unique, up to permutation, if and only if K = d. Proposition 2.5. Let F = {fj}K Then H is a dual frame of F if and only if Θ∗ satisfying ΘF C = 0d. j=1 be a frame for Zd H = Θ∗ 2 with a natural dual frame G. G + C for some K × d matrix C Proof. Given the existence of a matrix C with Θ∗ immediate that ΘF Θ∗ letting C = Θ∗ H − Θ∗ H = ΘF Θ∗ G gives ΘF C = ΘF Θ∗ G + C and ΘF C = 0d, it is G = Id. Conversely, if H is a dual frame of F , then (cid:3) G = Id − Id = 0d. H − ΘF Θ∗ H = Θ∗ As in Rd and Cd (see [13], Proposition 6.3), the following result still holds in Zd 2; however, since the proof in [13] uses the invertibility of the frame operator, we provide a modified proof here. Proposition 2.6. Let F = {fj}K frame if and only if F is a basis. j=1 be a frame for Zd 2. Then F has a unique dual Proof. Since a frame is a spanning set, F is a basis if and only if the vectors fj are linearly independent and K = d. This is equivalent to the only K × d matrix C satisfying ΘF C = 0d being the zero matrix. By Proposition 2.5, this happens if and only if the (unique choice of) natural dual G is the only dual frame of F . (cid:3) The diagonal of the Gramian matrix Θ∗ when F and H are a dual frame pair, the diagonal of the cross-Gramian matrix Θ∗ is the vector whose ith element is (hi, fi). Definition 2.7 ([3]). A Parseval frame for Zd Zd 2 is a sequence of vectors F = {fj}K j=1 ⊂ F ΘF is the vector whose ith entry is (fi, fi); F ΘH 2 such that y = KXj=1 (y, fj)fj for all y ∈ Zd 2. Note that a binary Parseval frame must be a binary frame. In matrix notation, F is a Parseval frame for Zd If a collection of vectors {xj} ⊂ Zd 2 satisfies (xi, xj) = 0 for all i 6= j and (xi, xi) = 1 for all i, we say, through a slight abuse of terminology, that {xj} is an orthonormal set. An easy, matrix theoretical consequence of the definitions of frame and Parseval frame is the following proposition: 2 if and only if ΘF Θ∗ F = Id. Proposition 2.8. Let F = {fj}K j=1 be a sequence of vectors in Zd 2. (1) The rows of ΘF are linearly independent if and only if F is a frame. (2) The rows of ΘF are orthonormal if and only if F is a Parseval frame. 6 V. FURST AND E. P. SMITH In the remainder of this paper, unless otherwise noted, all vectors are elements of the binary vector spaces Zd 2 . All operations are performed modulo 2; for exam- ple, Tr(A) represents the usual trace of a matrix A, but Tr(AB) ≡ Tr(BA) (mod 2) for two binary matrices A and B. All frames refer to binary frames. Throughout, we denote the standard orthonormal basis in Zd 2 by {ε1, ε2, . . . , εd}. 2 or ZK 3. Dual and Parseval Binary Frames If F = {fj}K j=1 is a frame for Zd 2 and K = d, then the vectors {fj} must be linearly independent and hence a basis with a unique (natural) dual G. In this case, (fj, gj) = γj(fj) = 1 for all 1 ≤ j ≤ K. In this section, we are largely concerned with the question of which sequences α ∈ ZK 2 satisfy α[j] = (fj, hj) for a dual frame pair (F , H); so we assume K > d. We use kαk0 to denote the number of nonzero entries in a vector α, the parity of which will be fundamental in this paper. j=1 be a frame for Zd Lemma 3.1. Let F = {fj}K 2 and let π be a permutation of the set {1, 2, . . . , K}. Denote by Fπ the frame {fπ(j)}K j=1. Then a frame H is a dual frame of F if and only if Hπ is a dual frame of Fπ. Furthermore, if α is a sequence in ZK 2 , then the dual frame pair (F , H) satisfies (fj, hj) = α[j] for every j if and only if the dual frame pair (Fπ, Hπ) satisfies (fπ(j), hπ(j)) = α[π(j)] for every j. Proof. Suppose ΘF Θ∗ H = Id and diag(Θ∗ ΘFπΘ∗ fπ(1) · · · fπ(1) · · · fπ(K)   HΘF ) = α. Then - h∗ π(1) - ... π(K) - - h∗ π(1) - ... π(K) -      P ∗P   - h∗ - h∗ fπ(K)   · · · fK     - h∗ 1 - ... - h∗ K -   Hπ =   =   =   f1 = ΘF Θ∗ = Id H where P =   - ε∗ π−1(1) - ... - ε∗ π−1(K) -   , and here we use εi to indicate the ith standard basis vector in ZK 2 . Thus Fπ and Hπ are dual frames and (fj, hj) = α[j] for each j implies (fπ(j), hπ(j)) = α[π(j)]. For the converse statements, let σ = π−1. (cid:3) The next theorem and corollary are the analog of Theorem 1.6 in binary space. BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 7 Theorem 3.2. Given α ∈ ZK (fi, hi) = α[i] for every i if and only if kαk0 ≡ d (mod 2). 2 , there exists a dual frame pair (F , H) for Zd 2 with Proof. Suppose (F , H) is a dual frame pair for Zd i. Then 2 such that (fi, hi) = α[i] for every kαk0 ≡ Tr(Θ∗ HΘF ) ≡ Tr(ΘF Θ∗ H) ≡ Tr(Id) ≡ d (mod 2). Conversely, suppose kαk0 ≡ d (mod 2). We consider three cases. Case 1: kαk0 = d. Let fj = εj for 1 ≤ j ≤ d and let fj be arbitrary for d < j ≤ K. A natural dual frame is then given by gj = εj for 1 ≤ j ≤ d and gj = 0 for d < j ≤ K. Define β ∈ ZK 2 by β[j] = 1 for 1 ≤ j ≤ d and β[j] = 0 for d < j ≤ K, and let π be a permutation of {1, 2, . . . , K} such that β[j] = α[π(j)]. It follows that diag(Θ∗ GΘF ) = β. By Lemma 3.1, (Fπ−1, Gπ−1) is the desired dual frame pair with (fπ−1(j), gπ−1(j)) = β[π−1(j)] = α[j] for each j. Case 2: kαk0 = d + 2t for some positive integer t ≤ K −d 2 . Consider the frame F defined in Case 1 above, but set fj = ε1 for all d + 1 ≤ j ≤ d + 2t. A natural dual frame G of F is the same as that defined in Case 1 above. Let C be the K × d matrix whose top d × d block is 0d and rows d + 1 through d + 2t are equal to ε∗ 1. The remaining rows of C are zeros. Due to the introduction of an even number of ε∗ 1's, we see that ΘF C = 0d, and hence H given by the rows of Θ∗ G + C is a dual frame of F , by Proposition 2.5. Since diag(Θ∗ HΘF ) is the vector composed of d + 2t ones followed by K − (d + 2t) zeros, Lemma 3.1 again implies the existence of the desired dual frame pair. H = Θ∗ Case 3: kαk0 = d − 2t for some positive integer t ≤ d 2. Consider again the frame F defined in Case 1, except set fd+1 = εd + εd−1 + · · · + εd−2t+2 + εd−2t+1. We again take the same natural dual frame G as in Case 1 above. Let C be the K × d matrix whose top d − 2t rows are zeros, row d − 2t + 1 through row d + 1 are f ∗ d+1, and the remaining rows are zeros. Then ΘF C = 0d, and hence H defined by Θ∗ H = Θ∗ G + C is a dual frame of F . Due to the presence of an even number of ones in fd+1, (fd+1, fd+1) = 0 while (fj, hj) = 0 for d − 2t + 1 ≤ j ≤ d. Consequently, diag(Θ∗ HΘF ) consist of ones in its first d − 2t entries followed by K − (d − 2t) zeros. Lemma 3.1 again completes the proof. (cid:3) Corollary 3.3. Given α ∈ ZK dual frame H for Zd 2 , there exists a Parseval frame F and a corresponding 2 with (fi, hi) = α[i] for every i if and only if kαk0 ≡ d (mod 2). Proof. The necessity of the condition on kαk0 follows immediately from Theorem 3.2. The sufficiency depends on slight modifications of the frame F constructed in the proof of Theorem 3.2. In Case 1, instead of letting fj for d < j ≤ K be arbitrary, set each of those vectors to be the zero vector, ~0, in Zd 2. Proposition 2.8 implies F is a Parseval frame. Similarly, the frame built in Case 2 is a Parseval frame if we set fj = ~0 for 2t+1 ≤ j ≤ K. The frame built in Case 3 is not a Parseval frame; however, consider instead the frame F ′ defined as f ′ d+2 = fd+1, and j = ~0 for d + 3 ≤ j ≤ K. By Proposition 2.8, F ′ is a Parseval frame. Note that each f ′ column of the matrix C constructed in Case 3 is still a (possibly trivial) dependence j = fj for 1 ≤ j ≤ d + 1, f ′ 8 V. FURST AND E. P. SMITH relation among the columns of ΘF ′, which implies ΘF ′C = 0d. Since the natural dual G of F constructed in Case 3 is still a natural dual of F ′, the frame H with analysis operator Θ∗ j, hj) = (fj, hj) for all j since hd+2 = 0. (cid:3) G + C is a dual frame of F ′; moreover, (f ′ H = Θ∗ Remark 3.4. The Parseval frames built in Cases 1 and 2 of the above corollary, in fact, satisfy (fj, fj) = α[j] for each j after a suitable permutation, as allowed by Lemma 3.1. However, this is not true in Case 3. By constructing a Parseval frame that satisfies (fj, fj) = α[j] for each j in the case when kαk0 = d−2t for some positive integer t ≤ d 2 , we will prove the binary analog of Theorem 1.4. Theorem 3.5. Given nonzero α ∈ ZK (fi, fi) = α[i] for every i if and only if kαk0 ≡ d (mod 2). 2 , there exists a Parseval frame F for Zd 2 with Proof. Since a Parseval frame is self-dual, the necessity of the condition on kαk0 follows immediately from Theorem 3.2. For sufficiency, Remark 3.4 implies that we need only construct Parseval frames satisfying (fi, fi) = α[i] for every i for kαk0 < d. Since kαk0 ≡ d (mod 2) and kαk0 6= 0, we must have d ≥ 3. For d = 3 and kαk0 = 1, we build the Parseval frame 2 that, after suitable permutation and inclusion of copies of ~0, satisfies any for Z4 kαk0 = 2. 2 Given any odd dimension d, suppose we have constructed, without zero vectors, corresponding to kαk0 = 1, 3, . . . , d − the Parseval frames F 1, F 3, . . . , F d−4 for Zd−2 Zd with two zero entries and unioning the augmented vectors with εd−1, εd ∈ Zd 2. Then 2 corresponding to kαk0 = 3, 5, . . . , d − 2. 4. For each odd n, create the collection eF n+2 by augmenting each vector of F n eF 3, eF 5, . . . , eF d−2 are Parseval frames for Zd Let eF 1 = {~1 + ε1,~1 + ε2, . . . ,~1 + εd,~1} where ~1 represents the vector with d ones in 2. Then eF1 is a Parseval frame for Zd If the dimension d is even, we create the Parseval frames eF n+1 from eF n for each eF n with a last entry of 0 and insert the vector εd ∈ Zd In both cases, permutation of the vectors and possible inclusion of copies of ~0 (cid:3) n = 1, 3, . . . , d − 3, corresponding to kαk0 = 2, 4, . . . , d − 2 : augment each vector of 2 corresponding to kαk0 = 1. 2. finishes the proof. and note that we may permute the vectors as needed. Moreover, we may insert any number of copies of ~0 to satisfy any K > 4. By augmenting each vector with a last entry of 0 and inserting the vector ε4 ∈ Z4 2, we construct the Parseval frame     0 1 1   ,  1 0 1   ,  1 1 0   ,  1 1 1   ,     0 1 1 0   ,   1 0 1 0   ,   1 1 0 0   ,     1 1 1 0   ,   0 0 0 1     BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 9 4. Binary Frames with Prescribed Frame Operator In the previous section, we gave a necessary and sufficient condition on α ∈ ZK 2 for the existence of a Parseval frame F for Zd 2 with (fj, fj) = α[j] for every j. In classical frame theory over R or C, the characterization has been broadened to frames with a given frame operator and specified values for kfjk (the case of a Parseval frame is when S = I), as in Theorem 1.5. In the classical case, the frame operator is a symmetric, invertible, positive definite matrix. For a binary frame F , SF = ΘF Θ∗ F is not necessarily invertible; for example, the zero matrix is the frame operator of any frame in which every vector occurs twice. Consequently, we must first characterize those binary symmetric matrices that are frame operators of binary frames. Throughout this section, we rely heavily on the idea of vector parity in Zd 2. Definition 4.1. Describe a vector v ∈ Zd is even if kvk0 ≡ 0 (mod 2). If a vector is not even, then it is odd. 2 as even if (v, v) = 0. Equivalently, a vector Lemma 4.2. (1) The sum of two even vectors is an even vector. (2) The sum of two odd vectors is an even vector. (3) The sum of an odd vector and an even vector is an odd vector. Proof. This follows from the above definition and the observation that if u, v ∈ Zd 2, then (u + v, u + v) = (u, u) + (u, v) + (v, u) + (v, v) = (u, u) + (v, v). (cid:3) As a consequence of this lemma, we note that a collection of only even vectors cannot span Zd 2. Given a d × d symmetric matrix S, we call A a factor of S if S = AA∗. We say A is a minimal factor if it has the minimum number of columns over all factors of S. Minimal factorization of symmetric binary matrices also arises in the computation of the covering radius of Reed-Muller codes ([11]). Theorem 4.3 ([17], Theorem 1). Every binary symmetric matrix S can be factorized as S = AA∗ for some binary matrix A. The number of columns of a minimal factor of S is rank(S) if diag(S) 6= ~0 and rank(S) + 1 if diag(S) = ~0. Proposition 4.4. If S = AA∗ for some d × m matrix A where m = rank(S) or m = rank(S) + 1, then rank(A) = rank(S). Proof. Frobenius's rank inequality and the fact that, for properly sized matrices C and D, rank(CD) ≤ min{rank(C), rank(D)} (see, for example, [14]) imply that rank(A) + rank(A∗) ≤ m + rank(S) ≤ m + rank(A). If m = rank(S), the above inequalities simplify to rank(A) ≤ rank(S) ≤ rank(A), and hence rank(A) = rank(S). If m = rank(S) + 1, we instead have 2 rank(A) ≤ 2 rank(S) + 1 ≤ 2 rank(A) + 1. Since 2 rank(A) = 2 rank(S) + 1 is impossible, we must have rank(A) = rank(S). (cid:3) 10 V. FURST AND E. P. SMITH We can use factors of a matrix to construct frames with a given frame operator. Minimal factors correspond to minimal frames, that is, frames with the fewest number of elements. In the previous section, we disregarded frames {fj}K j=1 that were bases since they corresponded to unique duals {gj}K j=1 with predetermined values for the dot products (fj, gj). In this section, however, we are concerned with (fj, fj), so we do not rule the case K = d out of consideration. Theorem 4.5. Let S be a d × d symmetric matrix with rank(S) = d. There exists a (minimal) d-element frame F (i.e., basis) such that S = ΘF Θ∗ F if and only if diag(S) 6= ~0. If diag(S) = ~0, then there exists a (minimal) (d + 1)-element frame F such that S = ΘF Θ∗ F . F . If diag(S) = ~0, Proof. Suppose there exists a d-element frame F such that S = ΘF Θ∗ then ΘF is a d×d square matrix with all even rows, which cannot span Zd 2. Hence, the columns of ΘF cannot span Zd 2, contradicting F being a frame. Conversely, suppose diag(S) 6= ~0. By Theorem 4.3, there exists a d × k matrix A such that S = AA∗ and k = rank(S). By Proposition 4.4, rank(A) = rank(S) = k = d. Thus, A is a d × d matrix whose columns span Zd 2. Defining F by ΘF = A constructs the d-element frame. Now suppose diag(S) = ~0. By Theorem 4.3, there exists a minimal d × k factor A such that S = AA∗ and k = rank(S)+1. By Proposition 4.4, rank(A) = rank(S) = d. Therefore, the columns of A span Zd 2, and we can construct a (d + 1)-element frame by taking ΘF = A. (cid:3) Remark 4.6. The columns of ΘF can be augmented by copies of the zero vector without affecting S, so non-minimal frames can always be constructed from minimal frames by including any number of copies of the zero vector. Next we construct minimal frames whose prescribed frame operators are not of full rank. Let r(A) and c(A) denote the number of rows and columns of a matrix A, respectively. Lemma 4.7 ([17], Section 4). Let S be a d × d symmetric matrix of rank(S) < d. There exists a permutation matrix P and a nonsingular matrix T such that S = P ∗(cid:20) L M M ∗ K (cid:21) P = P ∗T (cid:20) L 0 0 0 (cid:21) T ∗P where L is a symmetric matrix with r(L) = rank(L) = rank(S). Corollary 4.8. A d × d symmetric matrix S with diag(S) = ~0 must have even rank. Proof. It is known that if rank(S) = d, then d must be even (see, for example, [11] Section 9.3). Suppose rank(S) < d. The rank(S) × rank(S) symmetric matrix L constructed in Lemma 4.7 has rank(L) = rank(S). Since the diagonal elements of S are the diagonal elements of L and K, diag(L) = ~0. So rank(S) must be even. (cid:3) Theorem 4.9. Let S be a d × d symmetric matrix with rank(S) < d. There exists a k-element frame F such that S = ΘF Θ∗ F if and only if k ≥ 2d − rank(S). BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 11 Proof. Necessity follows from Frobenius's rank inequality rank(ΘF ) + rank(Θ∗ F ) ≤ k + rank(S) since frames are spanning sets. Conversely, let k be an integer such that k ≥ 2d − rank(S). Let L, P, T be the matrices guaranteed by Lemma 4.7, and let V = P ∗T . Suppose diag(L) 6= ~0. By Theorem 4.3 there exists a factor H of L such that and r(H) = c(H) = rank(L). Consider the augmented matrix (cid:20) L 0 0 0 (cid:21) =(cid:20) H 0 (cid:21)(cid:2) H ∗ 0 (cid:3) A =(cid:20) H 0 0 B (cid:21) where the columns of B are the standard basis vectors of Zd−rank(S) twice. Then r(A) = d, c(A) = 2d − rank(S), and AA∗ =(cid:20) L 0 0 0 (cid:21). By construction, , each repeated 2 rank(A) = d. Since V is nonsingular, V A is a d × (2d − rank(S)) matrix of rank d such that S = V AA∗V ∗. If k = 2d − rank(S), let a minimal frame F be the columns of V A; if k > 2d − rank(S), augment the columns of V A with the necessary number of zero vectors. Now suppose diag(L) = ~0. By Corollary 4.8, rank(L) = rank(S) must be even. As above (by Theorem 4.3) we can factorize L with a matrix H, but now r(H) = rank(L) and c(H) = rank(L) + 1. In this case, we build the following augmented matrix:   eA = 0 r1 H (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12) r2 (cid:12)(cid:12)(cid:12)(cid:12) eB 0   where r1 = [1 1 1 · · · 1] is a vector of length c(H), r2 = [1 0 0 0 · · · 0] has length 2 basis vectors of Zd−(rank(S)+1) , each repeated twice. Since the (i, j) entry of the product 2(d − rank(S)) − 1, and the columns of eB are the zero vector followed by the standard eAeA∗ can be viewed as the dot product of the ith and jth rows of eA, it is easy to see that eAeA∗ = (cid:20) L 0 0 0 (cid:21). vector. Since c(H) is odd, the vector [r1 r2] is even, as is each row of eB. By construction, rank(eA) = d, r(eA) = r(H) + 1 + d − rank(S) − 1 = d, and c(eA) = of V eA if k = 2d − rank(S) or these columns together with k − 2d + rank(S) copies of c(H) + 1 + 2(d − rank(S) − 1) = 2d − rank(S). As above let F consist of the columns Indeed, since diag(L) = ~0, each row of H is an even the zero vector if k > 2d − rank(S). (cid:3) Theorems 4.5 and 4.9 provide minimal (and non-minimal) frames with frame oper- ator S, subject only to restrictions based on rank(S). In what follows, sometimes we 12 V. FURST AND E. P. SMITH will make additional assumptions on S, which allow the construction of frames with frame operator S in different, and sometimes more intuitive, ways. Definition 4.10. A d × d symmetric matrix S is said to be parity indicative if, for every 1 ≤ i ≤ d, the diagonal entry Sii = 1 if and only if the ith row of S is odd. Lemma 4.11. Let S be a d × d symmetric matrix, and suppose S = AA∗ for some d × m matrix A. If every column of A is odd, then S is parity indicative. Conversely, if S is parity indicative and the columns of A are linearly independent, then every column of A must be odd. Proof. Assume S = AA∗ =Pm i where each column ai of A is odd. The ith row i=1 aia∗ j satisfying aj[i] = 1. Suppose the ith row of S is odd. of S equals the sum of those a∗ Then this sum must be composed of an odd number of nonzero terms by Lemma 4.2. That is, there is an odd number of indices j having aj[i] = 1. Consequently, the ith j=1 aj[i]aj[i] = 1. On the other hand, if the ith row of S is even, then Lemma 4.2 implies that an even number of aj have aj[i] = 1, resulting in Sii = 0. row of A is an odd vector, and Sii =Pm To show the converse, assume that the columns of A are linearly independent. Suppose some of the columns of A are even; denote the odd columns by {oj} and the even columns by {ej}. If for every index l, ej[l] = 1 for an even number of the vectors {ej}, then P ej = ~0, contradicting the linear independence of the columns of A. So, assume that there exists an index i such that the number of the vectors {ej} that satisfy ej[i] = 1 is odd. If an odd number of the vectors {oj} are such that oj[i] = 1, then the ith row of A is even, so Sii = 0; on the other hand, the ith row of S equals the sum of an odd number of rows e∗ j , which is odd, by Lemma 4.2. If there are an even number of the vectors {oj} with oj[i] = 1, then the ith row of A is odd, so Sii = 1; on the other hand, the ith row of S equals the sum of an odd number of e∗ j , which is even, by Lemma 4.2. Therefore, S is not parity indicative. (cid:3) j plus an odd number of rows o∗ j plus an even number of o∗ Lemma 4.12. Let S be a d × d symmetric matrix, and suppose S = AA∗ for some d × m matrix A. If S is parity indicative, diag(S) = ~0, and c(A) = rank(A) + 1, then either every column of A is even or every column of A is odd. Proof. Denote the odd columns and even columns of A by {oj} and {ej}, respectively, and assume both sets are nonempty. Since each row of S is even, for every i, an even number of the vectors {oj} must have oj[i] = 1, by Lemma 4.2. It follows that P oj = ~0; that is, Ax = ~0 where x[j] = 1 if j is the index of an odd column and x[j] = 0 if j is the index of an even column. But since every row of A is even, A~1 = ~0. Hence Ay = ~0 for y = ~1 + x. Since the nonzero linearly independent vectors x and y are both contained in the null space of A, the rank-nullity theorem implies rank(A) ≤ c(A) − 2 = rank(A) + 1 − 2 = rank(A) − 1, a contradiction. Therefore, either {oj} or {ej} must be empty. (cid:3) One additional useful fact is required before we state our main result. BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 13 Lemma 4.13. (1) Suppose e, o1, o2, o3 ∈ Zd 2 are four vectors such that e is even and o1, o2, o3 are odd. Then there exists three even vectors f1, f2, f3 and an odd vector p such that ee∗ + o1o∗ 1 + o2o∗ 2 + o3o∗ 3 = f1f ∗ 1 + f2f ∗ 2 + f3f ∗ 3 + pp∗, and span{e, o1, o2, o3} = span{f1, f2, f3, p}. (2) Suppose e1, e2, e3, o ∈ Zd 2 are four vectors such that e1, e2, e3 are even and o is odd. Then there exists an even vector f and odd vectors p1, p2, p3 such that e1e∗ 1 + e2e∗ 2 + e3e∗ 3 + oo∗ = f f ∗ + p1p∗ 1 + p2p∗ 2 + p3p∗ 3, and span{e1, e2, e3, o} = span{f, p1, p2, p3}. Proof. For part (1), let f1 = e + o1 + o2, f2 = e + o1 + o3, f3 = e + o2 + o3, and p = o1 + o2 + o3. By Lemma 4.2, f1, f2 and f3 are even and p is odd. For part (2), let f = e1 + e2 + e3, and p1 = e1 + e2 + o, p2 = e1 + e3 + o, p3 = e2 + e3 + o. Lemma 4.2 implies f is even and p1, p2, p3 are odd. Easy computations show that the given equalities are satisfied. (cid:3) We are now ready for the binary analog of Theorem 1.5: necessary and sufficient conditions on pairs of matrices S and vectors α such that S is the frame operator of a frame with vector "norms" determined by α. The necessary condition is easy. Theorem 4.14. Let F = {fi}K vector in ZK 2 defined by α[i] = (fi, fi) for each i. Then kαk0 ≡ Tr(S) (mod 2). i=1 be a frame such that S = ΘF Θ∗ F . Let α be the Proof. kαk0 ≡ Tr(Θ∗ F ΘF ) ≡ Tr(ΘF Θ∗ F ) ≡ Tr(S) (mod 2). (cid:3) Sufficiency breaks down into three possible scenarios. If S is parity indicative, then a minimal frame F with frame operator S must consist of only odd vectors or can attain any nonzero vector α with kαk0 ≡ Tr(S) (mod 2) in the sense that (fi, fi) = α[i] for each i; if S is not parity indicative, a minimal frame must contain at least one even vector. This is shown in Theorems 4.15 and 4.17. Nonminimal frames can be constructed to correspond to any nonzero α with kαk0 ≡ Tr(S) (mod 2) if S is parity indicative or to any such α with at least one zero entry if S is not parity indicative (Corollary 4.16 and Theorem 4.17). The frame elements can be permuted in any way without affecting the frame op- F = S and Θ eF = ΘF P ∗ for some permutation matrix P , Indeed, if ΘF Θ∗ erator. then Θ eF Θ∗ eF = ΘF P ∗P Θ∗ F = ΘF Θ∗ F = S. Therefore, in what follows, we need only construct frames with the correct number of odd elements, corresponding to kαk0, in order to attain the dot products prescribed by α. Theorem 4.15. Let S be a d × d parity indicative symmetric matrix. Let K = 2d − rank(S), and let α ∈ ZK (1) Suppose diag(S) 6= ~0. 2 be a nonzero vector with kαk0 ≡ Tr(S) (mod 2). 14 V. FURST AND E. P. SMITH (a) If rank(S) = d, there exists a (minimal) K-element frame F such that S = ΘF Θ∗ F and (fi, fi) = α[i] for every i only if kαk0 = K. (b) If rank(S) < d, there exists a (minimal) K-element frame F such that S = ΘF Θ∗ F and (fi, fi) = α[i] for every i. (2) Suppose diag(S) = ~0. Then rank(S) < d. (a) If rank(S) = d − 1, there exists a (minimal) K-element frame F such that S = ΘF Θ∗ F and (fi, fi) = α[i] for every i only if kαk0 = K. (b) If rank(S) < d − 1, there exists a (minimal) K-element frame F such that S = ΘF Θ∗ F and (fi, fi) = α[i] for every i. Proof. In case (1a), Theorem 4.5 implies the existence of a K = d - element frame F whose frame operator is S. The d columns of ΘF must be linearly independent, so every fi must be odd, by Lemma 4.11. (As a corollary of Theorem 4.14, we note that d ≡ Tr(S) (mod 2) for any d × d, full rank, parity indicative symmetric matrix S with diag(S) 6= ~0.) For case (1b), instead of using the result of Theorem 4.9, we rely directly on The- orem 4.3 to construct a d × rank(S) matrix A with rank(A) = rank(S) such that AA∗ = S. Since the columns of A are linearly independent, Lemma 4.11 implies that they are all odd. As in the proof of Theorem 4.9, we consider an augmented matrix ΘF =(cid:2) A B (cid:3) but with more care taken in the choice of B. By letting the columns of B be d−rank(S) of the standard basis vectors not in the column space of A, each repeated twice, we construct a frame F for kαk0 = 2d−rank(S). Replacing any identical pair of columns of B, say {εl, εl}, with {εl +εn, εl +εn} for any other basis vector εn 6= εl, the columns of ΘF still span, ΘF Θ∗ F is still equal to S, but now F contains two fewer odd vectors. In this way, we are able to construct a frame F with dot products (fi, fi) satisfying any kαk0 = rank(S) + 2m for 0 ≤ m ≤ d − rank(S). (Note that, by the proof of Theorem 4.14, rank(S) ≡ Tr(S) (mod 2).) By Lemma 4.13, any four vectors consisting of three odds and one even can be substituted by four vectors consisting of three evens and one odd, having the same span and no effect on ΘF Θ∗ F . Each substitution allows us to increase the number of even vectors by two, until only two odd vectors remain in F if rank(S) is even or one odd vector remains if rank(S) is odd. Therefore, we can build a frame with 2d − rank(S) elements, corresponding to any nonzero α with kαk0 ≡ rank(S) ≡ Tr(S) (mod 2). Now let S be parity indicative with diag(S) = ~0. Since every row of S is even, rank(S) < d. In case (2a), Theorem 4.9 implies the existence of a (d + 1)-element frame F whose frame operator is S. Since F is a spanning set, it must contain an odd vector. By Lemma 4.12, every vector in F must be odd. (Note that, by Corollary 4.8, case (2a) can only occur if rank(S) is even, hence d is odd.) Lastly, we assume in case (2b) that rank(S) ≤ d − 2. By Theorem 4.3 and Proposi- tion 4.4, there exists a d × (rank(S) + 1) matrix A with rank(A) = rank(S) such that AA∗ = S. By Lemma 4.12, either every column of A is even or every column is odd. Since Sii = 0 for every i, every row of A is even, hence the sum of the columns of A is ~0. Moreover, since diag(S) = ~0, we know that rank(S) must be even, by Corollary 4.8. If every column of A were odd, then the sum of all rank(S) + 1 columns would BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 15 have to be odd, by Lemma 4.2, yielding a contradiction. So every column of A must be even. n ∈ span{b∗ j : b∗ 2 n of B such that b∗ Augment A with a column of zeros and call the resulting matrix B. Then each column of B is even, each row of B is even, and B has an even number of columns. Consider a row b∗ j is a row of B and j 6= n}. Replace n by its complement (that is, add ~1∗ for ~1 ∈ Zrank(S)+2 b∗ n), and call the resulting matrix C. Then CC ∗ = S, rank(C) = rank(S) + 1, and C is composed of rank(S) + 2 odd columns. As in case (1b), we now augment C with d−(rank(S)+1) of the standard basis vectors not in the column space of C, each repeated twice, to construct ΘF . In doing so, we construct a frame F consisting of rank(S) + 2 + 2(d − (rank(S) + 1)) = 2d − rank(S) vectors, with frame operator S, such that every element of F is odd. By Theorem 4.14, kαk0 must be even. As in case (1b), we can replace pairs of odd elements of F by even vectors until only two odd vectors remain. (cid:3) to b∗ Corollary 4.16. Let S be a d × d parity indicative symmetric matrix. Let K > 2d − rank(S). Let α ∈ ZK 2 be a nonzero vector such that kαk0 ≡ Tr(S) (mod 2). Then there exists a K-element frame F such that S = ΘF Θ∗ F and (fi, fi) = α[i] for every i. Proof. Since S is parity indicative, a K-element frame with K > 2d − rank(S) is necessarily non-minimal and can be constructed by augmenting the minimal frames of the previous theorem. Consider first the minimal frame F guaranteed by case (1a) of Theorem 4.15. Adding the zero vector to F allows us to apply Lemma 4.13 and create frames satisfying (fi, fi) = α[i] for any kαk0 ≡ Tr(S) (mod 2) with 0 < kαk0 < d. Similarly, for case (2a), including the zero vector allows the construction of a frame corresponding to any kαk0 = 2, 4, 6, . . . d + 1. In either case, the addition of two identical copies of odd vectors or two identical copies of even vectors provides frames for any kαk0 ≡ Tr(S) (mod 2) when kαk0 ≥ d + 2 (case (1a)) or kαk0 ≥ d + 3 (case (2a)). Similarly in cases (1b) and (2b), the addition of two identical copies of an odd vector or two identical copies of an even vector yield frames for 2d − rank(S) < kαk0. (cid:3) Theorem 4.17. Let S be a d × d symmetric matrix that is not parity indicative. Let K ≥ 2d − rank(S) or if rank(S) = d and diag(S) = ~0, let K ≥ d + 1 . Let α ∈ ZK 2 be a nonzero vector such that kαk0 ≡ Tr(S) (mod 2). Then there exists a K-element frame F such that S = ΘF Θ∗ F and (fi, fi) = α[i] for every i only if kαk0 6= K. Proof. We use Theorem 4.5 and Remark 4.6 or Theorem 4.9 to construct a K-element frame F such that ΘF Θ∗ F = S. By Lemma 4.11, F must contain an even vector. Of course, F must also contain an odd vector, in order to span. Let m ≡ Tr(S) (mod 2) represent the number of odd elements of F and K −m be the number of even elements. By Lemma 4.13, F may be replaced by a frame with two more or two fewer odd vectors. Through repeated applications, we can construct a frame F corresponding to any kαk0 = 1, 3, 5, . . . , K −1 if m is odd and K is even, any kαk0 = 1, 3, 5, . . . , K −2 if m is odd and K is odd, any kαk0 = 2, 4, 6, . . . , K − 1 if m (≥ 2) is even and K (> m) is odd, or any kαk0 = 2, 4, 6, . . . , K − 2 if m (≥ 2) is even and K (> m) is even. (cid:3) 16 V. FURST AND E. P. SMITH 5. Examples and Data 5.1. Examples. In this subsection we consider two symmetric matrices S and build frames with various α's to illustrate the main result of Section 4. The algorithm for factoring a matrix as S = AA∗ and for reducing A into a minimal factor can be found in [17]. Example 5.1. Consider the identity matrix S = I4 =   1 0 0 0 0 1 0 0 0 0 1 0 0 0 0 1   and note that it is a symmetric, parity indicative, full rank matrix. Any frame with frame operator S is a Parseval frame. By Theorem 4.15, a minimal 4-element such Parseval frame must satisfy kαk0 = 4 where α[i] = (fi, fi) for each i; clearly, this follows from the Parseval frame necessarily being an orthonormal basis. Corollary 4.16 guarantees Parseval frames in Z4 2 of length K = 5 with either 2 or 4 odd vectors corresponding to any α ∈ Z5 2 with kαk0 = 2, 4. To begin the construction, factor S as I4I ∗ 4 . Appending the zero-column to the left factor yields the matrix ΘF =   1 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 1 0   , the columns of which constitute a frame with frame operator S and α = (1, 1, 1, 1, 0). To obtain any other α ∈ Z5 2 with kαk0 = 4, simply permute the columns. We utilize Lemma 4.13 to reduce the number of odd vectors by 2. Let e be the zero-column and o1, o2, o3 be the first, second, and third columns of ΘF , respectively. Replacing e, o1, o2, o3 with their counterparts constructed in Lemma 4.13 results in ΘF ′ =   1 1 0 0 1 1 0 1 0 1 0 1 1 0 1 0 0 0 1 0   . Taking the columns of ΘF ′ as frame vectors builds the frame F ′ satisfying α = (0, 0, 0, 1, 1). Again, the columns of F ′ can be permuted to acquire any α ∈ Z5 2 with kαk0 = 2. Notice that a permutation of F ′ appears in the proof of Theorem 3.5. Example 5.2. Suppose we wish to find a frame for Z3 2 of length 7 with frame operator S =  0 1 0 1 1 1 0 1 0   . BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 17 Since this rank 2, symmetric matrix is not parity indicative, we apply Theorem 4.17. In doing so, we follow the proof of Theorem 4.9 and factor S as S = P ∗T(cid:20) L 0 0 0 (cid:21) T ∗P where P is the 3 × 3 identity matrix, L =(cid:20) 0 1 1 1 (cid:21) , and T =  1 0 0 0 1 0 1 0 1   . Then A =(cid:20) H 0 0 B (cid:21) =  1 1 0 0 1 0 0 0 0 0 1 1   , and we append three zero-columns to T A to build the frame F : ΘF =hT A ~0 ~0 ~0i =  1 1 0 0 0 0 0 1 0 0 0 0 0 0 1 1 1 1 0 0 0   . Letting (fi, fi) = α[i] for each i, we see that F satisfies α = (1, 0, 1, 1, 0, 0, 0). We in- crease or decrease the number of odd vectors as desired, by applying Lemma 4.13 first to {f1, f3, f4, f5} and then to {f4, f5, f6, f7}. We obtain frames F1 and F2 satisfying ΘF1 =   ΘF2 =   1 1 1 0 1 0 0 1 0 1 0 1 0 0 0 1 0 0 1 0 0 1 1 0 0 0 0 0 1 0 0 0 0 0 0 1 1 1 0 1 1 1   , α1 = (0, 0, 0, 0, 1, 0, 0);   , α2 = (1, 0, 1, 0, 1, 1, 1). By Theorem 4.17, kαk0 = 7 is unattainable. j=1 in Zd 5.2. Data. An exhaustive search for frame operators ΘF Θ∗ F and kαk0 associated with F = {fj}K 2 was performed, using Python 3.6, for various dimensions and frame lengths (i.e. various d's and K's). The tables contained in this subsection hold information about the number of symmetric matrices that are frame operators and the set of kαk0 that accompany them. We include summaries for dimensions d = 2, . . . , 5. Because every frame in Zd 2 must have at least d vectors, and because 2d is the minimum number of vectors needed to ensure every symmetric matrix is a frame operator (Theorems 4.5, 4.9), the computations were performed for K = d, . . . , 2d. For d = 2, . . . , 5, in the table containing information about the d-dimensional binary space, the entry in the row labeled {αmin, αmin + 2, . . . , αmin + 2t} and column labeled K = k0 shows the number of symmetric matrices S in d-dimensional space such that for each α ∈ {αmin, αmin + 2, . . . , αmin + 2t} there exists a frame F = {fj}k0 j=1 such that S = ΘF Θ∗ F and α[i] = (fi, fi) for 1 ≤ i ≤ k0. 18 V. FURST AND E. P. SMITH In each table, the sum of the entries of the last column represents all possible symmetric d × d binary matrices. There are 2d(d+1)/2 such matrices, which becomes prohibitively large as the dimension d increases. {kαk0} {1} {2} {1, 3} {2, 4} K 2 3 4 2 2 0 1 3 2 0 2 4 0 0 2 Table 1. Number of attainable frame operators of frames for Z2 2 K {kαk0} 5 3 4 12 0 0 {1} 12 21 0 {2} 4 0 {3} 0 0 {4} 0 {1, 3} 0 {2, 4} {1, 3, 5} 0 {2, 4, 6} 0 6 0 0 0 0 1 0 28 24 0 6 31 24 32 8 0 0 0 8 Table 2. Number of attainable frame operators of frames for Z3 2 {kαk0} K 6 5 4 0 168 0 {2} 28 0 0 {4} 224 392 0 {1, 3} 0 {2, 4} 0 {1, 3, 5} {2, 4, 6} 0 {1, 3, 5, 7} 0 {2, 4, 6, 8} 0 7 0 0 0 420 441 0 56 0 0 0 8 0 0 0 0 504 448 0 63 0 0 511 448 64 512 64 0 Table 3. Number of attainable frame operators of frames for Z4 2 {kαk0} 7 6 5 0 0 448 {5} 0 0 28 {6} 6720 0 0 {1, 3} 6720 13020 0 {2, 4} 0 {1, 3, 5} 0 {2, 4, 6} 0 {1, 3, 5, 7} 0 {2, 4, 6, 8} {1, 3, 5, 7, 9} 0 {2, 4, 6, 8, 10} 0 K 8 0 0 0 0 13888 15120 0 840 0 0 0 0 9 0 0 0 0 0 15988 15345 0 1008 0 0 0 10 0 0 0 0 0 0 16368 15360 0 1023 0 0 16383 15360 16384 1024 0 1024 Table 4. Number of attainable frame operators of frames for Z5 2 BINARY FRAMES WITH PRESCRIBED DOT PRODUCTS AND FRAME OPERATOR 19 We thank Erich McAlister for his very valuable feedback. Acknowledgement References [1] Z. Baker, B. Bodmann, S. Branum, M. Bullock, and J. McLaney, What is odd about binary Parseval frames?, Involve (2017), in press. [2] B. Bodmann, B. Camp, and D. Mahoney, Binary frames, graphs, and erasures, Involve 7:2 (2014), 151–169. [3] B. Bodmann, M. Le, L. Reza, M. Tobin, and M. Tomforde, Frame theory for binary vector spaces, Involve 2:5 (2009), 589–602. [4] M. Bownik and J. Jasper, Existence of frames with prescribed norms and frame operator, pp. 103–117 in Excursions in harmonic analysis, Vol 4, Appl. Numer. Harmon. Anal. New York: Birkhauser/Springer, 2015. [5] J. Cahill, M. Fickus, D. Mixon, M. Poteet, and N. Strawn, Constructing finite frames of a given spectrum and set of lengths, Appl. Comput. Harmon. Anal. 35:1 (2013), 52–73. [6] P. Casazza and G. Kutyniok, ed. Finite frames: theory and applications, Appl. Numer. Harmon. Anal. New York: Birkhauser/Springer, 2013. [7] P. Casazza, M. Fickus, J. Kovacevi´c, M. Leon, and J. Tremain, A physical interpretation of tight frames, pp. 51–76 in Harmonic analysis and applications, Appl. Numer. Harmon. Anal. Boston: Birkhauser, 2006. [8] P. Casazza and M. Leon, Existence and construction of finite frames with a given frame operator, Int. J. Pure Appl. Math. 63:2 (2010), 149–157. [9] O. Christensen, An introduction to frames and Riesz bases, Appl. Numer. Harmon. Anal. Boston: Birkhauser, 2003. [10] O. Christensen, A. Powell, and X. Xiao, A note on finite dual frame pairs, Proc. Amer. Math. Soc. 140:11 (2012), 3921–3930. [11] G. Cohen, I. Honkala, S. Litsyn, and A. Lobstein, Covering Codes, North-Holland Math. Li- brary. New York: Elsevier, 1997. [12] M. Fickus, D. Mixon, M. Poteet, and N. Strawn, Constructing all self-adjoint matrices with prescribed spectrum and diagonal, Adv. Comp. Math. 39:3-4 (2013), 585–609. [13] D. Han, K. Kornelson, D. Larson, and E. Weber, Frames for undergraduates, Student Math. Library. Providence: American Mathematical Society, 2007. [14] R. Horn and C. Johnson, Matrix Analysis. New York: Cambridge University Press, 1985. [15] R. Hotovy, D. Larson, and S. Scholze, Binary frames, Houston J. Math. 41:3 (2015), 875–899. [16] J. Kovacevi´c and A. Chebira, Life Beyond Bases: The Advent of Frames (Part II). IEEE SP Mag. 24:5 (2007), 115–125. [17] A. Lempel, Matrix factorization over GF (2) and trace-orthogonal bases of GF (2n)∗, SIAM J. Comput. 4:2 (1975), 175–186. Veronika Furst, Department of Mathematics, Fort Lewis College, Durango, Col- orado 81301, USA E-mail address: furst [email protected] Eric P. Smith, Department of Mathematics, Fort Lewis College, Durango, Col- orado 81301, USA E-mail address: [email protected]
1705.02625
1
1705
2017-05-07T14:01:50
Strongly extreme points and approximation properties
[ "math.FA" ]
We show that if $x$ is a strongly extreme point of a bounded closed convex subset of a Banach space and the identity has a geometrically and topologically good enough local approximation at $x$, then $x$ is already a denting point. It turns out that such an approximation of the identity exists at any strongly extreme point of the unit ball of a Banach space with the unconditional compact approximation property. We also prove that every Banach space with a Schauder basis can be equivalently renormed to satisfy the sufficient conditions mentioned. In contrast to the above results we also construct a non-symmetric norm on $c_0$ for which all points on the unit sphere are strongly extreme, but none of these points are denting.
math.FA
math
STRONGLY EXTREME POINTS AND APPROXIMATION PROPERTIES TROND A. ABRAHAMSEN, PETR H ´AJEK, OLAV NYGAARD, AND STANIMIR TROYANSKI Abstract. We show that if x is a strongly extreme point of a bounded closed convex subset of a Banach space and the identity has a geometrically and topologically good enough local approxi- mation at x, then x is already a denting point. It turns out that such an approximation of the identity exists at any strongly ex- treme point of the unit ball of a Banach space with the uncondi- tional compact approximation property. We also prove that every Banach space with a Schauder basis can be equivalently renormed to satisfy the sufficient conditions mentioned. In contrast to the above results we also construct a non-symmetric norm on c0 for which all points on the unit sphere are strongly extreme, but none of these points are denting. 1. Introduction Let X be a (real) Banach space and denote by BX its unit ball, SX its unit sphere, and X ∗ its topological dual. Let A be a non-empty set in X. By a slice of A we mean a subset of A of the form S(A, x∗, ε) := {x ∈ A : x∗(x) > M − ε} where ε > 0, x∗ ∈ X ∗ with x∗ 6= 0, and M = supx∈A x∗(x). We will simply write S(x∗, ε) for a slice of a set when it is clear from the setting what set we are considering slices of. Definition 1.1. Let B be a non-empty bounded closed convex set in a Banach space X and let x ∈ B. Then x a) is an extreme point of B if for any y, z in B we have x = y + z 2 ⇒ y = z = x. 7 1 0 2 y a M 7 ] . A F h t a m [ 1 v 5 2 6 2 0 . 5 0 7 1 : v i X r a b) is a strongly extreme point of B if for any sequences (yn)∞ n=1, (zn)∞ n=1 in B we have n kx − lim yn + zn 2 k = 0 ⇒ lim n kyn − znk = 0. 2010 Mathematics Subject Classification. Primary: 46B20; Secondary: 46B04. Key words and phrases. denting point, strongly extreme point. The second author was financially supported by GACR 16-073785 and RVO: 67985840. The fourth author was partially supported by MTM2014-54182-P and the Bulgarian National Scientific Fund under Grant DFNI-I02/10. 1 2 T. A. ABRAHAMSEN, P. H ´AJEK, O. NYGAARD, AND S. TROYANSKI When B is the unit ball, the above condition can be replaced by n kx ± xnk = 1 ⇒ lim lim n kxnk = 0. In this case we say that the norm is midpoint locally uniformly rotund (MLUR) at x. c) is a point of continuity for the map Φ : B → X if Φ is weak to norm continuous at x. When Φ is the identity mapping we just say that x is a point of continuity (PC). d) is a denting point of B if for every ε > 0 and δ > 0 there exists a slice S(x∗, δ) of B with diameter less than ε. e) is a locally uniformly rotund (LUR) point of BX if for any sequence (xn)∞ n kx + xnk = 2 lim lim n=1 we have n kxnk = 2kxk = 2 ⇒ lim n kx − xnk = 0. It is well known that LUR points are denting points and that denting points are strongly extreme points [KR82]. Trivially strongly extreme points are extreme points. The importance of denting points became clear in the sixties when the Radon-Nikod´ym Property (RNP) got its geometric description. In particular, it became clear that extreme points in many cases are al- ready denting as every bounded closed convex set in a space with the RNP has at least one denting point. The "extra" an extreme point needs to become a denting points is precisely described in the following Theorem 1.2. [LLT88] Let x be an extreme point of continuity of a bounded closed convex set C in X. Then x is a denting point of C. It is well known that all points of the unit sphere of ℓ1 are points of continuity for the unit ball Bℓ1. So from Theorem 1.2 we get that every extreme point of the unit ball of any subspace of ℓ1 automatically gets the "extra" to become denting. However, despite the theoretical elegance of Theorem 1.2, it is not always easy to to check whether the identity mapping is weak to norm continuous at a certain point of a bounded closed convex set. For this reason it is natural to look for geometrical conditions which ensure weak to norm continuity of the identity operator at x when we approximate it strongly by maps that are weak to norm continuous at x. One such idea could be to assume that x is strongly extreme (not just extreme as in Theorem 1.2) and that the identity map is approximated strongly by finite rank operators. But this is not enough to give the extreme point the "extra" needed to be denting: Consider x = x(t) ≡ 1 ∈ BC[0,1]. Then x is strongly extreme in BC[0,1], but the identity map I : BC[0,1] → BC[0,1] is not weak to norm continuous there (see the next paragraph), and limn kPnx − xk = 0, where (Pn) are the projections corresponding to the Schauder basis in C[0, 1]. Clearly Pn is weak to norm continuous at any point of BC[0,1] (as any compact operator is). STRONGLY EXTREME POINTS AND APPROXIMATION PROPERTIES 3 Actually C[0, 1] belongs to the class LD2 of Banach spaces where all slices of the unit ball have diameter 2. Naturally, in such spaces no point of the unit sphere can be a PC point. See e.g. the references in [AHN+16] for more information about the class LD2. Assuming x is strongly extreme, we need to make stronger assump- tions of the approximating sequence of the identity. One such condition which we impose is related to the behaviour of the approximating map- pings close to the point x. In particular we obtain as a corollary that in Banach spaces with the unconditional compact approximation prop- erty (UKAP) (see Definition 2.8), every strongly extreme point in the unit ball is PC and therefore denting. In particular, we have that this conclusion holds for Banach spaces with an unconditional basis with unconditionally basis constant 1. Further we show that every Banach space with a Schauder basis can be renormed to satisfy the conditions of Theorem 2.1. Nevertheless we construct, in Section 3, a closed convex body in c0 where all boundary points are strongly extreme, but none of them is denting. This body is not symmetric; we refer to its gauge as a non- symmetric equivalent norm on c0. Thus, we construct a non-symmetric equivalent norm on c0 for which all points on the unit sphere are strongly extreme points, but none of these points is denting. In fact every slice of the unit ball of c0 with this non-symmetric norm has diameter at least 1/√2. In Section 4 we investigate when equivalent norms conserve strongly extreme and denting points of the corresponding unit balls. The notation and conventions we use are standard and follow [JL01]. When considered necessary, notation and concepts are explained as the text proceeds. 2. Weak to norm continuity of the identity map Our most general result on how to force a strongly extreme point x to be denting in terms of approximating the identity map I : X → X at x is the following Theorem 2.1. Let x be a strongly extreme point of a non-empty bounded closed convex set C in a Banach space X. Let x be a point of continuity for a sequence Φn : C → X, n = 1, 2, . . . of maps such that (1) n kΦnx − xk = 0 lim and (2) where lim n lim ε→0+ fn(ε) = 0, fn(ε) = sup{dist((1 + λ)Φny − λy, C) : y ∈ C,kΦnx − Φnyk ≤ ε}, for some λ ∈ (0, 1]. Then x is a denting point of C. 4 T. A. ABRAHAMSEN, P. H ´AJEK, O. NYGAARD, AND S. TROYANSKI The proof follows from Theorem 1.2 and the next proposition which is an interplay between weak and norm topology. With B(x, ρ) we denote the ball with center at x and radius ρ. Proposition 2.2. Let x be a strongly extreme point of a convex set C of a normed space X and let 0 < λ ≤ 1. Assume that for every η > 0 there exist a weak neighbourhood W of x and a map Φ : W ∩ C → X such that (3) and (4) Then x is PC. Φ(W ∩ C) ⊂ B(x, η) sup w∈W ∩C dist((1 + λ)Φw − λw, C) < η. Proof. Since x is strongly extreme point, for every ε > 0 we can find δ > 0 such that kx − u, v ∈ C ⇒ ku − vk < λε. (5) u + v 2 k < δ, Set η = min{δ, λε}/2. There is a weak neighbourhood W of x and a map Φ satisfying (3) and (4). Set Ψ = I − Φ and pick an arbitrary w ∈ W ∩ C. Put y+ = (1 − λ)x + λw. Since x, w ∈ C we get y+ ∈ C by convexity. Since Φw + λΨw − y+ = (1 − λ)(Φw − x), we have from (3) kΦw + λΨw − y+k ≤ (1 − λ)η < η. (6) Having in mind (4) we can find y− ∈ C such that k(Φw − λΨw) − y−k < η. (7) This and (6) imply kx − y+ + y− 2 1 2k(Φw + λΨw − y+) + (Φw − λΨw − y−)k k ≤ kx − Φwk + < kx − Φwk + η ≤ 2η. From (5) we get On the other hand, using again (6) and (7), we get ky+ − y−k < λε. ky+ − y−k = ky+ − (Φw + λΨw) − y− + (Φw − λΨw) + 2λΨwk > 2λkΨwk − 2η. Hence 2λkΨwk < ky+ − y−k + 2η < λε + 2λε = 3λε. STRONGLY EXTREME POINTS AND APPROXIMATION PROPERTIES 5 This and (3) imply kw − xk ≤ kΦw − xk + kΨwk < 2ε. Since w is an arbitrary element of W ∩C we get that W ∩C ⊂ B(x, 2ε). (cid:3) Remark 2.3. If x is PC for C we get that x satisfies the hypotheses of Proposition 2.2 just taking Φ = I, λ ∈ (0, 1]. Proof of Theorem 2.1. Let {εn} be a sequence of positive numbers tend- ing to 0. Since Φn : C → X, n = 1, 2, . . . is weak to norm continuous at x there is weak neighbourhood Vn of x such that Φn(Vn ∩ C) ⊂ B(x, εn), n = 1, 2, . . . Thus the conditions of Theorem 2.1 imply that for every η > 0 we can find n = n(η) such that (3) and (4) hold for W = Vn and Φ = Φn. (cid:3) Recall that every linear compact operator is weak to norm continuous on bounded sets. This together with Theorem 2.1 gives Corollary 2.4. Let (X,k · k) be a Banach space and let · be an equivalent (not necessarily symmetric) norm on X with corresponding unit ball C. Let x be a strongly extreme point of C. Let Tn : X → X be linear compact operators such that (8) (9) n kTnx − xk = 0, lim lim n lim ε→0+ sup{(1 + λ)Tny − λy : y ≤ 1,kTn(x − y)k ≤ ε} = 1, for some λ ∈ (0, 1]. In particular the above is satisfied if (10) n (1 + λ)Tn − λI = 1. lim Then x is a denting point of C. Proof. It is enough to prove that (9) implies (2). Indeed, since there exists k > 0 such that k · k ≤ k·, then for every u ∈ X \ C we have dist(u, C) = inf{ku − vk : v ∈ C} ≤ k inf{u − v : v ∈ C} u − (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) u (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) u(cid:12)(cid:12)(cid:12)(cid:12) = k(u − 1). (cid:3) Remark 2.5. The functions fn defined in Theorem 2.1 can be discon- tinuous at 0. Indeed, X be a Banach space, e ∈ BX , and e∗ ∈ SX ∗ be such that e∗(e) = kek = ke∗k = 1. Define a (norm one) projection P on X by P x = e∗(x)e and put f (ε) = sup{kP y − Ryk : kyk ≤ 1,kP (e − y)k ≤ ε}, 6 T. A. ABRAHAMSEN, P. H ´AJEK, O. NYGAARD, AND S. TROYANSKI where R = I − P . Now, if the norm k · k on X is either strictly convex or Gateaux differentiable at e∗, then f is discontinuous at 0. Indeed, let ε = 0,kyk ≤ 1, and P (e − y) = 0. We get e∗(e − y)e = 0. Hence e∗(y) = e∗(e) = 1. By the strict convexity of the norm or the Gateaux differentiability of the norm at e∗, we have y = e. This implies Ry = Re = 0, so f (0) = 1. In order to prove that f is discontinuous at 0 we simply apply Corollary 2.4 with Tn = P . Since e is strongly extreme, but not denting, we get limε→0+ f (ε) > 1. Remark 2.5 shows that one cannot replace the limit condition (9) of Corollary 2.4, by lim n sup{(1 + λ)Tny − λy : y ≤ 1, Tnx = Tny} = 1. The conditions in Corollary 2.4 are essential. Let us illustrate this by examples. Example 2.6. Consider the space c of convergent sequences endowed with its natural norm. Let e = (1, 1, . . .) ∈ Sc and Pn the projection on c which projects vectors onto their n first coordinates. Clearly e is a strongly extreme point of Bc which is not denting. Moreover, it is evident that the condition limn kPne − ek = 0 fails and that the condition (9) (moreover (8)) holds for λ = 1 (and thus for all λ ∈ (0, 1]). It follows that the approximation condition in Corollary 2.4 is essential. Example 2.7. Consider again c endowed with its natural norm. Let e ∈ c be is as in Example 2.6. Define a projection P on c by P x = limn x(n)e and put Pn = P for all n. By construction Pne = e. For z = (0, 1, 1, . . .) we have Pnz = P z = e. Now, for any λ ∈ (0, 1] we have k(1 + λ)Pnz − λzk = k(1 + λ)e − λzk = 1 + λ. Thus lim n lim ε→0+ sup{k(1 + λ)Pny − λyk : y ∈ BX,kPn(e − y)k < εk} ≥ 1 + λ. It follows that the condition (2) in Theorem 2.1 is essential. We now present our results in terms of an approximation property introduced and studied by Godefroy, Kalton and Saphar. Definition 2.8. [GKS93] A Banach space X is said to have the un- conditional compact approximation property (UKAP) if there exists a sequence (Tn) of linear compact operators on X such that limn kTnx − xk = 0 for every x ∈ X and limn kI − 2Tnk = 1. Clearly Banach spaces X with the UKAP satisfy condition (10) for λ = 1. Clearly also Banach spaces with an unconditional basis with basis constant 1 have the UKAP (simply put Tn = Pn the projection onto the n first vectors of the basis). Thus we immediately have the following corollary. STRONGLY EXTREME POINTS AND APPROXIMATION PROPERTIES 7 Corollary 2.9. If X has the UKAP, in particular if X has an un- conditional basis with unconditional basis constant 1, then all strongly extreme points in BX are denting points. Let us mention that the global condition (10) is much stronger than the local condition (9), even in the case when it holds for all x in SX . This will be clear from the discussion below and in particular from Example 2.12 which shows that the condition (10) is strictly stronger than (9). For that example we will use the following result. Proposition 2.10. Let X be a Banach space and x a locally uniformly rotund (LUR) point in SX. Let (Tn) be a sequence of linear bounded operators on X, with limn kTnk = 1, and which satisfies condition (8) in Corollary 2.4 for x ∈ SX . Then condition (9) holds for λ = 1 (and thus for all λ ∈ (0, 1]). Proof. Pick an arbitrary sequence (εn) with εn > 0 and limn εn = 0. First we show that (11) lim n diamDn = 0, where Dn = {y ∈ BX : kTn(x − y)k < εn}. To this end note that it suffices to show that (12) yn ∈ BX, n kTn(x − yn)k = 0 ⇒ lim lim n kx − ynk = 0. Indeed, kTnkkx + ynk ≥ kTn(x + yn)k = k2Tnx + Tn(yn − x)k ≥ 2kTnxk − kTn(yn − x)k. Hence lim inf n kx + ynk ≥ 2. Since kynk ≤ 1 we get limn kx + ynk = 2. Since x is a LUR point, we get that (12) holds, and thus (11) holds. In order to prove (9) for λ = 1, it is enough to show that lim n dn = 1, where dn = sup{kTnx − Rnyk : y ∈ Dn}, Rn = I − Tn. Since x ∈ Dn, we have dn ≥ kTnx − Rnxk. So we get from (8) that lim inf dn ≥ 1. Now, pick arbitrary y ∈ Dn. Then we have kTnx − Rnyk ≤ kTnxk + kRnyk ≤ kTnxk + kRnxk + kRn(y − x)k ≤ kTnxk + kRnxk + kRnkk(y − x)k ≤ kTnxk + kRnxk + (kTnk + 1)diamDn. Hence lim sup dn ≤ 1. (cid:3) 8 T. A. ABRAHAMSEN, P. H ´AJEK, O. NYGAARD, AND S. TROYANSKI Proposition 2.11. Let (Tn) be a bounded sequence of linear compact operators on X, Rn = I − Tn, and (fn) ⊂ SX ∗ a separating family for X. Then the norm u = (cid:18) ∞ Xn=1 2−n(kRnuk2 + f 2 n(u))(cid:19) 1 2 is LUR at x ∈ X provided limn kRnxk = 0. Moreover, if the operators (Tn) commute and limn kTnk = 1, then limn Tn = 1. Proof. Pick a sequence (xk) ⊂ X with limk xk + x = x = xk. By convex arguments ([DGZ93, Fact 2.3 p. 45]) we have (13) (14) k kRnxkk = kRnxk, n = 1, 2, . . . lim fn(xk) = fn(x), n = 1, 2, . . . . lim k First we show that (xk) is norm compact. Given ε > 0 we can find n with kRnxk < ε. Using (13) we can find kε such that kRnxkk < ε for k < kε. The set K = {x1, x2, . . . , xkε} ∪ kxkTn(BX) is norm compact. We show that K is an ε-net for (xk). Indeed, pick xk, k > kε. Then kxk − Tnxkk = kRnxkk < ε and Tnxk ∈ kxkTn(BX). So (xk) is norm compact. Since (fn) is total, we get from (14) that limk kxk − xk = 0. Thus the norm · is LUR at the point x ∈ X. Now, let us prove the moreover part. As (Tn) commute, we have ∞ Tmu2 = 2−n(kRnTmuk2 + f 2 n(Tmu)) ∞ = Xn=1 Xn=1 Xn=1 ≤ = kTmk2u2. ∞ 2−n(kTmRnuk2 + (T ∗ mfn(u))2) 2−n(kTmk2kRnuk2 + kT ∗ mk2f 2 n(u)) Hence Tm2 ≤ kTmk2 for all m = 1, 2, . . . , so lim supm Tm ≤ lim supm kTmk. Since limm kTmx − xk = 0, we get lim inf m Tm ≥ 1, and so limm Tm = 1 provided limm kTmk = 1. (cid:3) It is now easy to give the announced example which shows that condition (10) can fail as condition (9) holds for every x in SX. Example 2.12. Consider c0 endowed with the norm k · k defined by kxk = supi,j≥1(x(i) − x(j)) where x = (x(k)) ∈ c0. Clearly k · k is equivalent to the canonical norm on c0. Let Pn be the projection onto the n first vectors in the canonical basis (ek) of c0 and let · be the norm on c0 given in Proposition 2.11 where fn = 0 for every n. Then (c0,·) fulfils the conditions of Proposition 2.10 and thus satisfies STRONGLY EXTREME POINTS AND APPROXIMATION PROPERTIES 9 condition (9) for every x in Sc0. Nevertheless we have Pk − λRk > 1 for any λ ∈ (0, 1], so condition (10) fails. For the latter, just consider (Pk − λRk)(Pk+1 From the two preceding propositions we also get i=1 ei). Corollary 2.13. Let X be a Banach space with a Schauder basis. Then there exists an equivalent norm k · k on X for which the sequence of projections Pn onto the first n vectors of the basis, satisfy (9) for λ = 1. On the other hand we have Proposition 2.14. There exists an equivalent norm k · k on C[0, 1] such that (9) does not hold for any λ > 0 and any sequence (Tn) of compact linear operators on Xwhen x ∈ C[0, 1] with kxk = 1 and limn kTnx − xk∞ = 0. Proof. The norm on C[0, 1] constructed in [AHN+16, Theorem 2.4] is midpoint locally uniformly rotund and has the diameter two property, i.e. all non-empty relatively weakly open subsets of the unit ball have diameter 2. In particular, in this norm all points on the unit sphere are strongly extreme, but none are denting. Thus the conclusion follows from Theorem 2.1. (cid:3) 3. An MLUR non-symmetric norm on c0 without denting points on the unit sphere We will now construct a non-symmetric equivalent norm on c0 for which all points on the unit sphere are strongly extreme, but none of these points are denting. Proposition 3.1. There exists in c0 an equivalent (non-symmetric) MLUR norm · with no denting points. Moreover, every open slice of the unit ball corresponding to · has diameter ≥ 1/√2. Proof. With J we denote the set of all finite strictly increasing se- quences (jk)k≥0 of natural numbers. For x = (x(k))k≥1 ∈ ℓ∞ and j ∈ J we put Q(x, j) = x(j0) +Xk≥1 2−kx+(jk), x+ = max{x, 0} 2−kx2(k))1/2, kxk = sup{Q(x, j) : j ∈ J}, q(x) = (Xk≥1 x = (kxk2 + q2(x))1/2. Clearly (15) kxk∞ ≤ kxk ≤ x ≤ (kxk2 + kxk2 ∞)1/2 ≤ √2kxk ≤ 3kxk∞. 10 T. A. ABRAHAMSEN, P. H ´AJEK, O. NYGAARD, AND S. TROYANSKI Claim 3.2. For every x ∈ c0 and every ε > 0, there exists m = m(x, δ) ∈ N and δ = δ(x, ε) > 0 such that (16) max{kx + yk,kx − yk} ≥ kxk + δ whenever (17) y ∈ ℓ∞,kRmyk∞ > ε, where Pm is the projection onto the first m vectors of the canonical basis of c0 and Rm = I − Pm. Proof. Pick m such that (18) and put δ = ε/2m+3. Using the definition of k·k we can find j = (ji)p such that kRmxk∞ < ε/8, k=0 kxk − δ/2 < Q(x, j). (19) We may assume that p ≥ m. Choose i ≥ 1 such that (20) and put j1 = (jk)i−1 (21) Pick y satisfying (17). There is r > m with y(r) > ε. We have (22) k=0. Using (20) we get Q(x, j) ≤ Q(x, j1) + 2−i+1kRmxk∞. ji−1 ≤ m < ji (x(r) + y(r))+ + (x(r) − y(r))+ ≥ y(r) − x(r) > ε − kRmxk∞. Put j2 = (j0, j1, . . . , ji−1, r). Since r > m ≥ ji−1 we have j2 ∈ J. So kx ± yk ≥ Q(x ± y, j2) = Q(x±, j1) + 2−i(x(r) ± y(r))+. Since i ≤ ji−1 we get from (20), (21), and (22) (kx + yk + kx − yk)/2 ≥ Q(x, j1) + 2−i−1(ε − kRmk∞) ≥ Q(x, j) + 2−i−1(ε − 5kRmxk∞) ≥ Q(x, j) + 2−m−1(ε − 5kRmxk∞). This and (18), (19) imply (16). (cid:3) Let x ∈ c0, xn ∈ ℓ∞, and limn x ± xn = x. By convex arguments as in [DGZ93, p. 45] we have (23) and (24) n kx ± xnk = kxk, lim lim n xn(k) = 0, k = 0, 1, . . . . We have to show that limn kxnk∞ = 0. Assume the contrary. Then we may assume that kxnk∞ ≥ 2ε > 0 for all n = 1, 2, . . . . From (24) it follows that there exists m such that kPmxnk∞ < ε for all n which are STRONGLY EXTREME POINTS AND APPROXIMATION PROPERTIES 11 sufficiently big. But, this contradicts (23) and Claim 3.2. Hence · is MLUR. Now, put C = {u ∈ c0 : u ≤ 1}. We will show that every non-void slice of C has k · k∞ diameter ≥ 1/√2. To this end, let f ∈ ℓ1, a ∈ R, H = {u ∈ c0 : f (u) > a}, and S = H ∩ C 6= ∅. Claim 3.3. For every x ∈ H with x < 1, there is y ∈ S with x − y ≥ kxk. Proof. Choose δ > 0 such that (kxk + 2δ)2 + q2(x) + δ < 1, f (x) − δ > a. There exists a natural number m such that kRmxk∞ < δ, (26) 0 < kxk − x(m) < √2mδ, (25) (27) and f (em) < δ (28) where em is the standard basis vector number m in c0. Put y = x − kxkem. Pick arbitrary j ∈ J. We will show that (29) Q(y, j) ≤ kxk + 2δ. If m < j0, then y(jk) = x(jk) for k = 0, 1, 2, . . . . So Q(y, j) = Q(x, j) ≤ kxk. If m > j0, we get Q(y, j) ≤ Q(x, j) ≤ kxk since y+(m) = 0. If m = j0 Q(y, j) ≤ kxk − x(m) + kRmxk∞. This and (26) imply (29). From (27) we get q2(y) < q2(x) + δ. From this, (25), and (29) we have y < 1. Since kxk < 1 we get from (28) that y ∈ H. Thus y ∈ S, and kx − yk∞ = kxk. Using (15) we get x − y ≥ kx − yk∞ = kxk, and the claim follows. Since x ∈ S can be chosen with ·-norm arbitrarily close to 1 we get from (15) that the diameter of S ≥ 1/√2. (cid:3) (cid:3) 4. Weak continuous perturbation of the norm It is natural to expect that weak continuous perturbations of the norm would preserve points of continuity. We will show that simi- lar perturbations of the norm preserve strongly extreme points of the corresponding ball. Theorem 4.1. Let X be a Banach space and let B be the unit ball corresponding to an equivalent norm · on X. Assume that the re- striction ·SX of · to SX is continuous at x ∈ SX 12 T. A. ABRAHAMSEN, P. H ´AJEK, O. NYGAARD, AND S. TROYANSKI a) with respect to the weak topology. Then x is a point of continuity for BX provided x is a point of continuity for B. b) with respect to a topology σ = σ(X, Y ), Y ⊂ X ∗ with the prop- erty that for every bounded sequence in X there exists a σ- Cauchy subsequence, and let x be an extreme point of BX. Then x is a strongly extreme point of BX provided it is a strongly ex- treme point for B. Proof. From the assumption x = 1. a). Pick ε > 0. Since x is a point of continuity for B, there is a weakly open set W containing x such that diam(W ∩ B) < ε. Since · is uniformly norm continuous, we can find δ > 0 such that diam(W ∩ (1 + 2δ)B) < 2ε. Since ·SX is weakly continuous at x, we can find a weakly open set V containing x such that V ∩ SX ⊂ (1 + δ)B. Since k · k is uniformly norm continuous, we can find η ∈ (0, 1) such that V ∩ (BX \ ηBX ) ⊂ (1 + 2δ)B. Finally there is a weakly open set U containing x such that U ∩ ηBX = ∅. Hence x ∈ U ∩ V ∩ W and diam(U ∩ V ∩ W ∩ BX ) < 2ε. b). Pick an arbitrary sequence (un) ⊂ X with limn kx ± unk = 1. In order to prove that limkunk = 0, it is enough to show that there exists a subsequence of (un) converging to zero. To this end we assume without loss of generality that σ − lim (um − un) = 0. m,n Taking into account that kx ± um − un 2 k ≤ kx ± umk + kx − unk 2 , we get m,n kx ± lim um − un 2 k = 1. Put um,n = (um−un)/2, α± and σ−limm,n(x±um,n)/(1+α± Since ·SX is σ-continuous at x, we get m,n = kx±um,nk−1. Clearly limm,n α± m,n) = x, and (x±um,n)/(1+α± m,n = 0 m,n) ∈ SX . lim (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) m,n(cid:12)(cid:12)(cid:12)(cid:12) = x = 1. x ± um,n 1 + α± (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) m,n(cid:12)(cid:12)(cid:12)(cid:12) This implies limm,n x ± um,n = 1. Having in mind that x is a strongly extreme point of B, we get limm,n um,n = 0. Hence (un) is a norm Cauchy sequence. It follows that there is u ∈ X with limn kun−uk = 0. Thus kx ± uk = 1, which implies that u = 0 as x is an extreme point. Now limn kunk = 0 which finishes the proof. Remark 4.2. The requirement that x is an extreme point of BX is essential for part b). Indeed, in finite dimensional Banach spaces, all (cid:3) STRONGLY EXTREME POINTS AND APPROXIMATION PROPERTIES 13 norms are weak(= norm) continuous. Clearly a point x can be extreme for the ball B and not extreme for BX. We end with a result that follows directly by combining Theorems 1.2 and 4.1. Corollary 4.3. Let k·kW be a weakly continuous semi norm on bounded sets in a Banach space (X,k · k) and let · be a lattice norm on R2. Let x = (kxk,kxkW ). Let B be the unit ball of (X,·). Then the strongly extreme points of B and BX coincide, and the denting points of B and BX coincide. References [AHN+16] T. A. Abrahamsen, P. H´ajek, O. Nygaard, J. Talponen, and S. Troy- anski, Diameter 2 properties and convexity, Studia Math. 232 (2016), no. 3, 227–242. MR 3499106 [DGZ93] R. Deville, G. Godefroy, and V. Zizler, Smoothness and renormings in Banach spaces, Pitman Monographs and Surveys in Pure and Applied Mathematics, vol. 64, Longman Scientific & Technical, Harlow; copub- lished in the United States with John Wiley & Sons, Inc., New York, 1993. MR 1211634 (94d:46012) [JL01] [KR82] [GKS93] G. Godefroy, N. J. Kalton, and P. D. Saphar, Unconditional ideals in Ba- nach spaces, Studia Math. 104 (1993), 13–59. MR 1208038 (94k:46024) W. B. Johnson and J. Lindenstrauss (eds.), Handbook of the geometry of Banach spaces. Vol. I, North-Holland Publishing Co., Amsterdam, 2001. MR 1863688 K. Kunen and H. Rosenthal, Martingale proofs of some geometrical re- sults in Banach space theory, Pacific J. Math. 100 (1982), no. 1, 153– 175. MR 661446 B. L. Lin, P. K. Lin, and S. L. Troyanski, Characterizations of denting points, Proc. Amer. Math. Soc. 102 (1988), 526–528. MR 0928972 (89e:46016) [LLT88] 14 T. A. ABRAHAMSEN, P. H ´AJEK, O. NYGAARD, AND S. TROYANSKI Department of Mathematics, University of Agder, Postboks 422, 4604 Kristiansand, Norway. E-mail address: [email protected] URL: http://home.uia.no/trondaa/index.php3 Mathematical Institute, Czech Academy of Science, Zitn´a 25, 115 67 Praha 1, Czech Republic and Department of Mathematics, Faculty of Electrical Engineering, Czech Technical University in Prague, Zikova 4, 160 00, Prague, Czech Republic E-mail address: [email protected] Department of Mathematics, University of Agder, Postboks 422, 4604 Kristiansand, Norway E-mail address: [email protected] URL: http://home.hia.no/olavn/ Institute of Mathematics and Informatics, Bulgarian Academy of Science, bl.8, acad. G. Bonchev str. 1113 Sofia, Bulgaria and De- partamento de Matem´aticas, Universidad de Murcia, Campus de Es- pinardo, 30100 Espinardo (Murcia), Spain E-mail address: [email protected]
0909.5672
5
0909
2010-06-02T12:55:33
The Cauchy problem for Schr\"odinger-type partial differential operators with generalized functions in the principal part and as data
[ "math.FA", "math.AP" ]
We set-up and solve the Cauchy problem for Schr\"odinger-type differential operators with generalized functions as coefficients, in particular, allowing for distributional coefficients in the principal part. Equations involving such kind of operators appeared in models of deep earth seismology. We prove existence and uniqueness of Colombeau generalized solutions and analyze the relations with classical and distributional solutions. Furthermore, we provide a construction of generalized initial values that may serve as square roots of arbitrary probability measures.
math.FA
math
THE CAUCHY PROBLEM FOR SCHR ODINGER-TYPE PARTIAL DIFFERENTIAL OPERATORS WITH GENERALIZED FUNCTIONS IN THE PRINCIPAL PART AND AS DATA G UNTHER H ORMANN Dedicated to Peter Michor on the occasion of his 60th birthday celebrated at the Central European Seminar in Mikulov, Czech Republic, May 2009 Abstract. We set-up and solve the Cauchy problem for Schrodinger-type differential operators with generalized functions as coefficients, in particular, allowing for distributional coefficients in the principal part. Equations involving such kind of operators appeared in models of deep earth seismology. We prove existence and uniqueness of Colombeau generalized solutions and analyze the relations with classical and distributional solutions. Furthermore, we provide a construction of generalized initial values that may serve as square roots of arbitrary probability measures. 1. Introduction Partial (and pseudo-) differential operators of Schrodinger-type with variable coefficients in the principal part arise in the form of so-called paraxial equations in models of wave propagation. These are based on narrow-angle symbol approximations of wave operators and have been applied in integrated optics, underwater acoustic tomography, reflection seismic imaging, time-reversal mirror experiments (cf. [2, 5, 41]), and recently in [12] to seismic wave propagation near the core- mantle boundary inside the Earth (at approximately 2800 km depth). Since paraxial equations are used to split the wave fields according to a prescribed principal direction of propagation they are also called one-way wave equations. The leading-order approximation leads to model equations of Schrodinger-type, where the material properties are still reflected by the regularity structure of the coefficients in the principal part. Under strong smoothness conditions on the wave speed function (i.e., the coefficients in the original wave operators) well-posedness of the one-way wave Cauchy problems has been discussed in [20, 39]. In [12] this smoothness assumption has been considerably relaxed by allowing the coefficients to be of Holder- or Sobolev-type regularity below log-Lipschitz continuity. Recall that in general existence of distributional solutions to the second- order wave equation may fail below log-Lipschitz regularity of the coefficients (cf. [10]). Beyond such coefficient regularity barrier unique solvability of the Cauchy problem holds in the sense of Colombeau generalized functions (cf. [17]), even in a covariant setting. The relevance of coefficients with low Holder or Sobolev regularity has been shown in a variety of geophysical applications, e.g. in the study of phase transitions in Earth's lowermost mantle ([40]) or in exploration geophysics ([13, 21, 22, 24, 32, 36, 42]). The model analyzed in [12] describes the paraxial approximation to seismic wave propagation near the core-mantle boundary by the following Schrodinger-type equation with depth z as evolution variable, the 2-dimensional lateral x-variable, and a pseudodifferential time-frequency dependence (1) ∂zu − i(cid:0)∂x1 (c1∂x1u) + ∂x2(c2∂x2 u)(cid:1) = 0, Date: November 19, 2018. 2000 Mathematics Subject Classification. Primary: 46F30; Secondary: 35D99. Research supported by the FWF START grant Y237. 1 where c1(z, x, τ ) and c2(z, x, τ ) are strictly positive symbols of order −1 in τ , continuously differen- tiable with respect to z, but a non-Lipschitz dependence regarding the x-regularity, e.g. of Sobolev type H r+1 or Holder continuous with exponent r, where r < 1. Roughly speaking, the main result of [12] is the unique solvability of the corresponding Cauchy problem for the wave component u in C([0,∞[, S ′(R; H 2(R2)) (i.e., continuous dependence on z, arbitrary S ′ quality with respect to time or frequency, but lateral x-regularity of class H 2), given initial data in L1(R, H 2(R2)) (at z = 0) and a right-hand side in C1([0,∞[, L1(R, L2(R2))). The emphasis in this result is on the (Holder or) Sobolev regularity of the solution (with respect to x) in relation to the initial data regularity under lowest possible regularity assumptions on the coefficient (cf. the brief discussion on inverse analysis of medium regularity at the end of [12]). In the current paper we strive for a widening of the possible range of applications and thereby also simply for a more unified general mathematical set-up. Thus, we allow for discontinuous or distributional coefficients, initial data, and right-hand sides. In fact, a natural method is then to place the whole Cauchy problem into the context of nonlinear theories of generalized functions, in particular the differential algebras of Colombeau generalized functions (cf. [7, 8, 9, 19, 29]). In view of models from geophysics this amounts to including discontinuous material properties (e.g. at fault zones and geological boundaries or cracks) and Dirac-type data such as strong impulsive sources (e.g. explosions or earth quakes). A quantum field theoretic application of Colombeau generalized functions in terms of regularizations of Wightman distributions has been given in [18]. In the simpler context of quantum mechanics one will of course be interested in allowing for a singular zero order term in the operator to represents a potential and non-smooth data corresponding to an initial probability distribution. The square of the modulus of a solution to the standard Schrodinger equation is usually interpreted as (time evolution of a) probability density. Why should one not turn this around in the sense of allowing now for generalized initial data which represent a "square root of a given arbitrary probability measure"?1 We will show below how to construct a Colombeau generalized function whose square is associated with a given probability measure in the sense of distributional shadows. The issue of squares of distribution theoretic objects or generalized functions which contain or model measures (in particular, Dirac- type terms) also arises in general relativity theory, e.g. in the pseudo-Riemannian metric associated with impulsive pp-waves (cf. [19, Section 5.3] and [23, 34, 35]) and in the stress energy tensor corresponding to ultrarelativistic Reissner-Nordstrøm fields (cf. [33]). A different regularization approach for powers of Dirac measures as initial value appears in [27] in the context of (semilinear) heat equations with singular potentials. In our analysis of the Cauchy problem we will return to denoting the evolution variable by t and furthermore incorporate also first-order terms in the differential operator. Note that unlike in the specific geophysical model situation of [12] we neglect additional pseudodifferential aspects in the symbol, since these played only the role of an external parameter upon a partial Fourier transform. In summary of the basic structure, let T > 0 be arbitrary. We consider the Cauchy problem for a generalized function u on Rn × [0, T ] in the form (2) ∂tu − i nXk=1 ∂xk (ck∂xk u) − iV u = f u t=0 = g, (3) where ck (k = 1, . . . , n), V , and f are generalized functions on Rn × [0, T ] and g is a general- ized function on Rn. We note that basics of a theory for abstract variational problems in the context of Colombeau-spaces have been presented in [16]. As indicated in [16, Section 8] the Lax-Milgram-type theorem established there provides solutions to Dirichlet-problems involving differential operators similar to the spatial derivatives appearing in the left-hand side of our above differential equation. However, this does not directly extend to our kind of evolution problem and 1My thanks go to Gebhard Grubl and Michael Oberguggenberger who brought up this viewpoint in joint discussions at the University of Innsbruck in April 2002. 2 furthermore involves less strict solution spaces than we are willing to accept here, namely GH 1 as compared to GH∞ . Colombeau-generalized solutions to linear and nonlinear Schrodinger equations with constant coefficient principal part have been constructed previously in [4, 37, 38]. We may point out that our current paper has both focus and aim different from the enormous literature on the prominent questions concerning properties of Lp-solutions and spectral theory for Schrodinger equations with standard principal part (i.e., with constant or smoooth coefficients) and singular potentials. Instead we investigate here the feasibility of extending the solution concept and the basic analysis to the case of generalized functions and, in particular, non-smooth or highly singular principal parts. The plan of the paper is as follows. In Section 2 we briefly review the regularization approach to generalized functions in the sense of Colombeau and show how square roots of probability measures can be implemented in this framework. Section 3 presents the main result of unique existence of generalized solutions to our Schrodinger-type Cauchy problem (2- 3). Furthermore, we discuss the relation of Colombeau generalized solutions with classical and distributional solution concepts. 2. Regularization of coefficients and data The basic idea of regularization methods is to replace non-smooth data by approximating nets of smooth functions, e.g. instead of g consider C∞ ∋ gε → g (ε → 0). More generally, we may replace g by a net (gε)0<ε≤1 of C∞ functions, convergent or not, but with moderate asymptotics with respect to ε and identify regularizing nets whose differences compared to the moderateness scale are negligible. For a modern introduction to Colombeau algebras we refer to [19]. Here we will also make use of constructions and notations from [15]. Construction of generalized functions based on a locally convex topological vector space E: Let E be a locally convex topological vector space whose topology is given by the family of seminorms {pj}j∈J . The elements of ME := {(uε)ε ∈ E(0,1] : ∀j ∈ J ∃N ∈ N pj(uε) = O(ε−N ) as ε → 0} and NE := {(uε)ε ∈ E(0,1] : ∀j ∈ J ∀q ∈ N pj(uε) = O(εq) as ε → 0}, are called E-moderate and E-negligible, respectively. With operations defined componentwise, e.g. (uε) + (vε) := (uε + vε) etc., NE becomes a vector subspace of ME. We define the generalized functions based on E as the factor space GE := ME/NE. If E is a differential algebra then NE is an ideal in ME and GE becomes a differential algebra too. By particular choices of E we reproduce several standard Colombeau algebras of generalized functions. For example, E = C with the absolute value as norm yields the generalized complex numbers GE = eC and for any Ω ⊆ Rd open, E = C∞(Ω) with the topology of compact uniform convergence of all derivatives provides the so-called special Colombeau algebra GE = G(Ω). Recall that Ω 7→ G(Ω) is a fine sheaf, thus, in particular, the restriction uB of u ∈ G(Ω) to an arbitrary open subset B ⊆ Ω is well-defined and yields uB ∈ G(B). Moreover, we may embed D′(Ω) into G(Ω) by appropriate localization and convolution regularization. In case E ⊆ D′(Ω) certain generalized functions can be projected into the space of distributions by taking weak limits: we say that u ∈ GE is associated with w ∈ D′(Ω), if uε → w in D′(Ω) as ε → 0 holds for any (hence every) representative (uε) of u. This fact is also denoted by u ≈ w. In the current paper we will consider open strips of the form ΩT = Rn× ]0, T [⊆ Rn+1 (with T > 0 arbitrary) and employ the spaces E = H ∞(ΩT ) = {h ∈ C∞(ΩT ) : ∂αh ∈ L2(ΩT ) ∀α ∈ Nn+1} 3 with the family of (semi-)norms khkHk =(cid:16) Xα≤k k∂αhk2 L2(cid:17)1/2 (k ∈ N), as well as E = W ∞,∞(ΩT ) = {h ∈ C∞(ΩT ) : ∂αh ∈ L∞(ΩT ) ∀α ∈ Nn+1} with the family of (semi-)norms khkW k,∞ = max α≤k k∂αhkL∞ (k ∈ N). (Note that ΩT clearly satisfies the strong local Lipschitz property [1, Chapter IV, 4.6, p. 66] and therefore the Sobolev embedding theorem [1, Chapter V, Theorem 5.4, Part II, p. 98] implies that every element of H ∞(ΩT ) and W ∞,∞(ΩT ) belongs to C∞(ΩT ).) To avoid overloaded subscripts we shall make use of the following (abuses of) notation in the sequel GL2 (Rn × [0, T ]) := GH∞(ΩT ) and GL∞ (Rn × [0, T ]) := GW ∞,∞(ΩT ). For example and in explicit terms we will represent a generalized initial value g ∈ GL2(Rn × [0, T ]) by a net (gε) with the moderateness property and similarly for the right-hand side and the coefficients. Asymptotically negligible errors in the ∀k ∃m : kgεkHk = O(ε−m) (ε → 0) initial data, e.g. by using a representative (egε) instead, are expressed by estimates of the form ∀k ∀p : kgε − egεkHk = O(εp) (ε → 0). Similar constructions and notations will be used in case of E = H ∞(Rn) and E = W ∞,∞(Rn). Note that by Young's inequality ([14, Proposition 8.9.(a)]) any standard convolution regularization with a scaled mollifier of Schwartz class provides embeddings L2 ֒→ GL2 and Lp ֒→ GL∞ (1 ≤ p ≤ ∞). As an example of a detailed regularization model we construct Colombeau generalized positive square roots of arbitrary probability measures, which can serve as initial values in the Cauchy problem analyzed below. Proposition 2.1. Let µ be a (Borel) probability measure on Rn. Choose ρ ∈ L1(Rn)∩W ∞,∞(Rn) to be positive with R ρ = 1 and satisfying ρ(x) ≥ x−m0 when x ≥ 1 with some m0 > n. Set ρε(x) = 1 (1) hε is positive and setting φε := √hε the net (φε)ε∈ ]0,1] represents an element φ ∈ G(Rn) such that φ2 ≈ µ; (2) there exists a generalized function g ∈ GL2(Rn) such that g2 ≈ µ and the class of (gεΩ)ε∈ ]0,1] is equal to φΩ in G(Ω), or by slight abuse of notation gΩ = φΩ, for every bounded open subset Ω ⊆ Rn. ε ) and hε := µ ∗ ρε, then we have εn ρ( x Proof. (1) The Borel measure µ is regular since it is finite and Rn is locally compact and second countable ([6, Proposition 7.2.3] or also [14, Theorem 7.8]). Thus µ(Rn) = 1 implies that we can find a compact subset A ⊆ Rn such that µ(A) ≥ 1/2. A variant of Young's inequality for measures (cf. [14, Proposition 8.49]) applied to ∂αhε = µ ∗ ∂αρε directly implies that the net (hε) is C∞(Rn)-moderate, even with global L∞-norms, since kµ ∗ ∂αρεkL∞ ≤ kµkvar k∂αρεkL∞ = k∂αρkL∞ε−n−α. (Here k kvar denotes the total variation norm on the space of finite Radon measures on Rn, which gives kµkvar = µ(Rn) = 1 in case of the positive probability measure µ). Furthermore, for any x ∈ Rn and ε > 0 we have ρε(x − y)dµ(y) ≥Z hε(x) = µ ∗ ρε(x) =Z ρε(x − y)dµ(y) ≥ µ(A) · min ρε(z) > 0 z∈{x}−A Rn A 4 and hence that φε = √hε ∈ C∞(Rn) (0 < ε ≤ 1). Moreover, if K ⊆ Rn is compact then we have2 with r(K) := max{x : x ∈ K − A} > 0 the estimate (∗) inf x∈K hε(x) = inf x∈K µ ∗ ρε(x) ≥ µ(A) · inf x∈K min z∈{x}−A 1 2εn min z≤r(K) ≥ ρ(z/ε) ≥ 1 2εn r(K)m0 ρε(z) ≥ εm0 = 1 2 · min z∈K−A εm0−n 2 r(K)m0 ρ(z/ε) εn (0 < ε < 1/r(K)). If α ∈ Nn is arbitrary then ∂αφε is a linear combination of terms of the form ∂β1hε ··· ∂βk hε/hl/2 with appropriate β1, . . . , βk ∈ Nn and l ∈ N. Hence C∞(Rn)-moderateness of (hε) together with the lower bounds for inf x∈K hε(x) obtained above prove C∞(Rn)-moderateness of φε. Finally, we have that by construction φ2 is represented by hε = µ ∗ ρε and thus clearly converges to µ as ε → 0 in the sense of distributions. (2) For j ∈ N let Kj be the closed ball of radius 2j around 0 in Rn. Let χ0 ∈ D(Rn), 0 ≤ χ0 ≤ 1 with χ0(x) = 1 when x ≤ 1 and χ0(x) = 0 when x > 2 and put χj(x) := χ0(2−jx) (j ≥ 1). Thus we have χj(x) = 1 when x ≤ 2j, χj(x) = 0 when x > 2j+1, and for every γ ∈ Nn ε k∂γχjkL2 = 2(−γ+n/2)jk∂γχ0kL2. We define the net (gε)ε∈ ]0,1] of smooth functions on Rn by gε := χj φε, if 2−j−1 < ε ≤ 2−j (j ∈ N). For every ε ∈ ]0, 1] the function gε is smooth and has compact support, hence belongs to H ∞(Rn). We have to show that (gε) is an H ∞-moderate net. Let α ∈ Nn arbitrary and 2−j−1 < ε ≤ 2−j. By the Leibniz rule we have that ∂αgε is a linear combination of terms of the form ∂βφε · ∂α−βχj (β ≤ α). Therefore we have by the triangle inequality that k∂αgεkL2 is bounded above by a linear combination of terms of the form k∂βφε · ∂α−βχjkL2 ≤ k∂α−βχjkL2k∂βφεkL∞(Kj+1) ≤ 2nj/2 kχ0kHαk∂βφεkL∞(Kj+1) ≤ ε−n/2 kχ0kHαk∂βφεkL∞(Kj+1) and it remains to prove ε-moderate bounds for k∂βφεkL∞(Kj+1) when β ≤ α (note that the coupling of j with ε has to be taken into account here). As already noted in (1) the latter is in turn bounded above by a linear combination of terms of the form sup x∈Kj+1 ∂β1hε(x)··· ∂βk hε(x) hε(x)l/2 with suitable β1, . . . , βk ∈ Nn and l ∈ N. Thanks to the estimate (∗) in the proof of (1) we have 1/hε(x)l/2 ≤ 2 r(Kj+1)m0 ε−(m0−n) uniformly with respect to x ∈ Kj+1, where r(Kj+1) = O(2j+1) = O(1/ε). Thus it remains to consider the factors ∂βq hε(x) in the above supremum. As has already been noted in the first paragraph of the proof of (1) we have global L∞-estimates with moderate ε-dependence. Thus in summary, we have shown the H ∞-moderateness of (gε). Finally, let Ω ⊆ Rn be bounded and open. Choose j sufficiently large so that Ω ⊆ Kj. Then we have for 0 < ε < 2−j by construction of gε that gεΩ = φεΩ and hence equality of the corresponding classes in G(Ω). Moreover, the latter also implies that g2 ≈ µ, since the support of any test function in D(Rn) is contained in some open bounded subset. (cid:3) Remark 2.2. (i) Property (∗) established in course of the proof shows in fact that the class of (hε) in G(Rn) as well as φ are strictly positive generalized functions (cf. [26, Theorem 3.4]). In this sense, φ provides a strictly positive square root of the positive measure µ. 2We may exclude the case A = K = {x0} without loss of generality 5 (ii) For specific choices of ρ in L1(Rn) ∩ H ∞(Rn) such that √ρ ∈ H ∞(Rn) we could obtain that (φε) is also H ∞-moderate and directly defines a square root in GL2(Rn) without having to undergo the cut-off procedure in part (2) of Proposition 2.1 (which, on the other hand, cannot be avoided for general ρ ∈ H ∞). For example, putting ρ(x) = c(1+x2)−(n+1)/2 with a suitable normalization constant c > 0 provides such a mollifier. However, the above Proposition leaves considerably more flexibility in adapting the regularization to particular applications. 3. Generalized function solutions and coherence properties We come now to the main existence and uniqueness result for generalized solutions to the Cauchy problem (2-3). We recall that a regularization of an arbitrary finite-order distribution which meets the log-type conditions on the coefficients ck and V in the following statement is easily achieved by employing a re-scaled mollification process as described in [28]. Theorem 3.1. Let ck (k = 1, . . . , n) as well as V be real3 generalized functions in GL∞(Rn×[0, T ]), f in GL2(Rn × [0, T ]), and g be in GL2(Rn). Assume that the following log-type conditions hold for some (hence all) representatives (ckε)ε∈]0,1] of ck (k = 1 . . . , n) and (Vε) of V : k∂tckεkL∞ = O(log( 1 ε )) and k∂tVεkL∞ = O(log( 1 ε )) (ε → 0). In addition let the positivity conditions ckε(x, t) ≥ c0 for all (x, t) ∈ Rn × [0, T ], ε ∈ ]0, 1], k = 1, . . . , n with some constant c0 > 0 be met.4 Then the Cauchy problem (4) ∂tu − i nXk=1 ∂xk (ck∂xk u) − iV u = f, u t=0 = g, (5) has a unique solution u ∈ GL2(Rn × [0, T ]). Proof. In terms of representatives (4-5) means that we have a family of Cauchy problems parametrized by ε ∈ ]0, 1]: (6) ∂tuε − i nXk=1 ∂xk (ckε∂xk uε) − iVεuε = fε uε t=0 = gε on ΩT = Rn× ]0, T [, on Rn, (7) where ckε (k = 1, . . . , n) and Vε (are real-valued and) belong to W ∞,∞(ΩT ), ckε satisfies the log-type and the positivity condition as stated above, gε ∈ H ∞(Rn), and fε ∈ H ∞(ΩT ). Our strategy is to solve the corresponding problem at fixed, but arbitrary, parameter value ε and thereby produce a solution candidate (uε). The substance of the proof lies in the efforts to show that each uε belongs to H ∞(Rn×]0, T [) and in addition satisfies moderateness estimates in every derivative with respect to the L2-norm. We will obtain basic energy estimates by standard variational methods as discussed by Dautray-Lions in [11, Chapter XVIII, §7, Section 1], but we will need to perform an additional analysis of the dependence of the constants in these estimates on the various norms of the coefficient functions. Step 1 (basic estimates and regularity): We apply the set-up and constructions in [11, Chapter XVIII, §7, Section 1] to the Hilbert space triple H 1(Rn) ⊂ L2(Rn) ⊂ H −1(Rn) and the sesquilinear form aε(t; ϕ, ψ) := nXk=1 hckε(t) ∂xk ϕ∂xk ψiL2(Rn) + hVε(t)ϕψi (ϕ, ψ ∈ H 1(Rn)), 3In the sense that they possess representating nets of real-valued functions. 4For any other representative the conditions thus hold for sufficiently small ε with c0/2 instead. 6 where we have used the short-hand notation v(t) for a function x 7→ v(x, t). The basic conditions [11, (7.1) and (7.2), p. 621] on the quadratic form aε are easily seen to be met: continuous differentiability of the map t 7→ aε(t; ϕ, ψ) follows from smoothness of ckε (k = 1, . . . , n) and Vε, the form aε(t; ., .) is hermitian since each ckε and Vε is real-valued, and the positivity conditions on the coefficients ckε (k = 1, . . . , n) in the statement of Theorem 3.1 yield the following coercivity estimate aε(t; ϕ, ϕ) + λεkϕk2 L2 ≥ c0kϕk2 H 1 ∀ϕ ∈ H 1(Rn), where λε = c0+kVεkL∞ (thus we have λ = λε and α = c0 in [11, (7.2), ii) on p. 621] ). Furthermore, the hypotheses [11, (7.5) and (7.6), pp. 621-622] on the initial values hold by our assumptions on gε and fε. We claim that [11, Theorem 1, pp. 621-622] in combination with the remark on additional regularity [11, Remark 3, p. 625] implies that we obtain a unique solution uε ∈ C([0, T ], H 1(Rn)) ∩ C1([0, T ], H −1(Rn)) (S1) to the Cauchy problem (6-7) for every ε ∈ ]0, 1]. In fact, the basic theorem gives existence and uniqueness with uε ∈ L2([0, T ], H 1(Rn)) and u′ ε ∈ L2([0, T ], H −1(Rn)) only. However, as men- tioned in [11, Remark 3, p. 625] the existence proof given in Dautray-Lions directly shows that the solution as well as its t-derivative are L∞-functions of t with values in H 1 or H −1, respectively. The same remark states that even continuity with respect to t holds (an earlier reference for this fact is [25, Remark 10.2, p. 302], which also indicates a proof based on convolution regulariza- tion). Finally, we note that thanks to C([0, T ], H 1) ⊆ C([0, T ], H −1) an application of [11, Chapter XVIII, §1, Section 2, Proposition 7, p. 477] shows that the solution belongs to C1([0, T ], H −1). Moreover, a careful inspection of all constants appearing in the derivation of the a priori estimate [11, (7.16) and (7.17) on p. 624] shows that we may deduce the following basic energy estimate uniformly for all t ∈ [0, T ] (8) H 1 ≤ C2ε eC1ε kuε(t)k2 H 1 + kgεk2 TZ 0 (cid:16)kfε(τ )k2 L2 + k∂tfε(τ )k2 H−1(cid:17) dτ  , where the specifics on the constants C1ε and C2ε according to the derivation in [11, pp. 623-624] are as follows: C2ε = O(T (c0 + kVεkL∞)) = O(ε−p) (ε → 0) for some p ∈ N, and 1≤k≤n k∂tckεkL∞ + k∂tVεkL∞ )(cid:17) (ε → 0), C1ε = O(cid:16) T c0 · ( max which by the log-type conditions on (ckε) and (Vε) imply moderateness estimates for the first-order spatial derivatives in the form ∃N1 ∈ N : k∂xluεkL2(Rn×[0,T ]) ≤ √T sup t∈[0,T ]kuε(t)kH 1 = O(ε−N1 ) (ε → 0, l = 1, . . . , n). As a preparation for the procedure in Step 2 we claim that additional regularity properties for uε hold in the form (∗) (at fixed ε without precise asymptotic estimates of the norms). These can be obtained from the concept of mild solutions and evolution systems on L2(Rn) (cf. [30, Chapter 5]) for the self-adjoint k=1 ∂xk (ckε(x, t)∂xk v)(x) (v ∈ H 2(Rn)) familiy of operators (Aε(t))t∈[0,T ], where Aε(t)v(x) := Pn with common (i.e., t-independent) domain H 2(Rn): In fact, Equation (4) has the form uε ∈ L∞([0, T ], H 2(Rn)) ∩ W 1,∞([0, T ], H 1(Rn)) ∂tuε − iAε(t)uε = fε + iVεuε =: Fε, where we already know from (S1) that Fε ∈ C([0, T ], L2(Rn)) ∩ C1([0, T ], H −2(Rn)). Let the strongly continuous evolution system corresponding to (Aε(t))t∈[0,T ] be denoted by (Uε(t, s))0≤s≤t≤T , then we necessarily have by Duhamel's formula uε(t) = Uε(t, 0)gε + tZ 0 Uε(t, τ )Fε(τ ) dτ ∈ C1 w([0, T ], L2(Rn)), 7 where C1 Rn with values in L2(Rn) for every t. Writing the differential equation now in the form w([0, T ], L2(Rn)) means weakly continuously differentiable as t-dependent distribution on −iAε(t)uε = −∂tuε + Fε ∈ Cw([0, T ], L2(Rn)) we may employ elliptic regularity (spatial, with respect to x) for the second-order operator Aε(t) and deduce uε ∈ Cw([0, T ], H 2(Rn)) ⊆ L∞([0, T ], H 2(Rn)). Moreover, we may now state in addition that Fε ∈ Cw([0, T ], H 1(Rn)) and use Duhamel's formula again to show that also uε ∈ w([0, T ], H 1(Rn)) ⊆ W 1,∞([0, T ], H 1(Rn)) as claimed. C1 Step 2 (higher x-derivatives): We take the partial xj-derivative on both sides in Equations (6) and (7) to obtain a similar differential equation for ∂xj uε in the form nXk=1 nXk=1 ∂xk (ckε∂xk ∂xj uε) − iVε∂xj uε = i ∂t∂xj uε − i and the initial condition ∂xj uε(0) = ∂xj gε. Thanks to (∗) we have f1ε ∈ L2([0, T ], L2(Rn)) and ∂tf1ε ∈ L2([0, T ], H −1(Rn)), which corresponds to conditions (7.5-6) of [11, Chapter XVIII, §7, Section 1, Theorem 1]. Thus we may apply (S1) and (8) with ∂xj uε, ∂xj gε, f1ε replacing uε, gε, fε, respectively, but with the same constants C1 and C2. Collecting the results for j = 1, . . . , n we arrive at ∂xk (∂xj ckε∂xk uε) + i∂xj Vεuε + ∂xj fε =: f1ε (S2) uε ∈ C([0, T ], H 2(Rn)) ∩ C1([0, T ], L2(Rn)) k∂xl∂xj uεkL2(Rn×[0,T ]) ≤ n√T sup and ∃N2 ∈ N : (ε → 0, j, l = 1, . . . , n). Similarly, taking higher partial x-derivatives in Equations (6) and (7) produces the same kind of differential equations and initial conditions for these higher derivatives, where we may always con- sider the already estimated lower order derivatives of uε as part of the right-hand side. Therefore we obtain successively for arbitrary m ∈ N (Sm) t∈[0,T ] kuε(t)kH 2 = O(ε−N2 ) uε ∈ C([0, T ], H m(Rn)) ∩ C1([0, T ], H m−1(Rn)) x uεkL2(Rn×[0,T ]) = O(ε−Nm ) 0 ,α ≤ m : k∂α ∃Nm ∈ N∀α ∈ Nn (ε → 0). and Step 3 (t-derivatives): We first note that due to (Sm) we have ∂tuε(0) ∈ H ∞(Rn) and Equation (6) evaluated at t = 0 provides moderate ε-asymptotics of its Sobolev norms. Taking the partial t-derivative in Equation (6) now yields again the same kind of differential equation for ∂tuε as above when interpreted in the following way: ∂t∂tuε − i nXk=1 ∂xk (ckε∂xk ∂tuε) − iVε∂tuε = i nXk=1 ∂xk (∂tckε∂xk uε) + i∂tVεuε + ∂tfε =: ef1ε. 0 , α ≤ m, By property (Sm) (now replacing (∗) in a similar argument in Step 2) we deduce the required regularity conditions on ef1ε and ∂tef1ε ∈ L2([0, T ], H −1(Rn)) to apply (S1) and (8) now with ∂tuε, ∂tuε(0), ef1ε replacing uε, gε, fε, respectively. We obtain for α ∈ Nn 0 , α ≤ 1, ∃M1 ∈ N : k∂t∂α uε ∈ C1([0, T ], H 1(Rn)) ∩ C2([0, T ], H −1(Rn)), x uεkL2(Rn×[0,T ]) = O(ε−M1 ). Furthermore, inserting here the same procedure as in Step 1 now yields for any m ∈ N and for all α ∈ Nn x uεkL2(Rn×[0,T ]) = O(ε−Mm ). uε ∈ C1([0, T ], H m(Rn)) ∩ C2([0, T ], H m−1(Rn)), Thus in turn we may now deduce from the above differential equation for ∂tuε that ∂2 t uε(0) ∈ H ∞(Rm) holds with moderate spatial Sobolev norms. Hence we may now play the same game again with ∂2 t uε etc. and finally arrive at the statements that for arbitrary d, m ∈ N and for all α ∈ Nn uε ∈ Cd([0, T ], H m(Rn))∩Cd+2([0, T ], H m−1(Rn)), ∃Mmd ∈ N : k∂d x uεkL2(Rn×[0,T ]) = O(ε−Mmd ). 0 , α ≤ m, we have ∃Mm ∈ N : k∂t∂α t ∂α 8 Therefore we may conclude that (uε) is moderate net and its class GL2 (Rn × [0, T ]) defines a solution u to the Cauchy problem (4-5). The proof of uniqueness requires to show negligibility of any solution (uε) assuming that the right-hand side (fε) and the initial data (gε) are negligible. But this follows successively from the corresponding variant of the energy estimate (8) applied at each step of the sequence of differentiated differential equations used in course of the existence proof. (cid:3) We note that in case of smooth coefficients an easy integration by parts argument shows that any solution to the Cauchy problem obtained from the variational method as in [11, Chapter XVIII, §7, Section 1]) also provides a solution in the sense of distributions. The following statement ensures in addition the coherence with the Colombeau generalized solution. Corollary 3.2. Let V and ck (k = 1, . . . , n) belong to C∞(ΩT ) ∩ L∞(ΩT ) with bounded time derivatives of first-order, g0 ∈ H 1(Rn), and f0 ∈ C1([0, T ], L2(Rn)). Let u denote the unique Colombeau generalized solution to the Cauchy problem (4-5), where g, f denote standard embed- dings of g0, f0, respectively. Then u ≈ w, where w ∈ C([0, T ], H 1(Rn)) is the unique distributional solution obtained from the variational method. Proof. Let (uε) ∈ u be as in the proof of Theorem 3.1, then vε := uε − w satisfies the following Cauchy problem ∂tvε − i nXk=1 ∂xk (ck∂xk vε) − iV vε = fε − f0, vε t=0= gε − g0. Moreover, we have the energy estimate (8), where now both C1ε =: C1 and C2ε =: C2 both are independent of ε, in the form sup t∈[0,T ]kvε(t)k2 H 1 ≤ C2eC1(cid:16)kgε − g0k2 H 1 + TZ 0 (cid:16)kfε(τ ) − f0(τ )k2 L2 + k∂tfε(τ ) − ∂tf0(τ )k2 H−1(cid:17) dτ(cid:17). Since the smoothing process via mollification ensures convergence to 0 of the norms on the right- hand side of the above estimate we obtain even norm convergence of uε to w. (cid:3) Remark 3.3. To compare our main result with the evolution systems solution constructed in [12] for the case of Sobolev coefficients and spatial dimension n = 2 we drop the additional pseudodifferential aspects in that model. Then we obtain also coherence of the Colombeau concept with the functional analytic solution as we sketch in the following discussion: Let 0 < r < 1 and V = 0, c10, c20 ∈ C1([0, T ], H r+1(R2)), g0 ∈ H 2(R2), f0 ∈ C1([0, T ], L2(R2)). Let v denote the unique H 2-valued solution to the Cauchy problem corresponding to (4-5) with initial value g0 and right-hand side f0 (and coefficients cj0 replacing cj) according to [12, Theorem 4.2]. Let g, f , c1, c2 denote the embeddings of g0, f0, c10, c20 into corresponding Colombeau spaces, respectively, and let u ∈ GL2(R2 × [0, T ]) be the unique Colombeau solution to (4-5). Subtracting the equations satisfied by a representative (uε) of u and v we deduce the following equations for the difference hε := uε − v which imply an energy estimate of the form (8) for hε and with suitably adapted initial value and right-hand side efε instead. As noted in the proof of Corollary 3.2 we have appropriate norm convergence of the standard regularizations gε, fε of the data g0,f0. In addition, we now have to call on uniform boundedness of k∂tcεkL∞(R2×[0,T ]) and on the C1([0, T ], H r+1(R2))-norm convergence ckε − ck0 → 0 (ε → 0). Using the continuity of H r+1(R2) · H 2(R2) → H r+1(R2) we obtain efε → 0 in C1([0, T ], H r+1(R2)). Thus we finally obtain the convergence uε → v in C([0, T ], H 1(R2)), and in particular we have u ≈ v. 9 ∂thε − i ∂xk (ckε∂xk hε) = fε − f0 + i 2Xk=1 2Xk=1 ∂xk ((ckε − ck0)∂xk v) =: efε, hε t=0= gε − g0, Example 3.4. Consider the strictly positive square root of δ represented by (√ρε)ε∈ ]0,1], where ρ is a mollifier as in Proposition 2.1. (Note that we obtain hε = δ ∗ ρε = ρε in this case.) To simplify technical matters, assume in addition that ρ satisfies the conditions discussed in Remark 2.2(ii) and moreover √ρ ∈ L1(Rn). (For example, any suitably normalized function of the form x 7→ (1 + x2)−m/2 with m > 2n would do.) Then we may consider the class g ∈ GL2(Rn), given directly by (√ρε)ε∈ ]0,1], as a square root of δ. We have g2 ≈ δ, but g ≈ 0, which is easily seen by action on a test function ϕ upon substituting y = x/ε inRpρε(x) ϕ(x)dx = εn/2Rpρ(y) ϕ(εy)dy and applying dominated convergence (thereby using that √ρ ∈ L1). In essence, this effect has already been observed earlier in the generalized function model of ultrarelativistic Reissner-Nordstrøm fields in [33, Equations (15) and (17)]. The Cauchy problem (4-5) with generalized initial value g, right-hand side f = 0, constant coeffi- cients ck = 1 (k = 1, . . . , n) and V = 0, written out for representatives then reads ∂tuε = i∆uε, uεt=0 = √ρε. µt uε(t, x)2 dx =Z (Ut√ρε)(x) · (Ut√ρε)(x) dx =Z ε denote the positive measure on Rn with density function uε(t, .)2 for the pρε(x)2 dx =Z ε : t ∈ R, ε ∈ ]0, 1]} is a family probability measures on Rn and kµt The solution is given by the action of the strongly continuous unitary group Ut := exp(it∆) (t ∈ R) of operators on L2(Rn) in the form uε(t, x) = (Ut√ρε)(x). Let t ∈ R and µt Lebesgue measure. By unitarity of Ut we obtain ε(Rn) =Z hence {µt ε ∈ ]0, 1]) holds in the Banach space of finite measures. We claim that for any t 6= 0 the net (µt on finite measures (cf. [3, §30]): Since √ρε ∈ L1(Rn) we obtain from the L1-L∞-estimate for the Schrodinger propagator ([31, §4.4, Theorem 1]) that kuε(t, .)kL∞ ≤ k√ρεkL1/(4πt)n/2 and therefore for any ψ ∈ Cc(Rn) ε)ε∈ ]0,1] converges to 0 with respect to the vague topology εkvar = 1 (t ∈ R, ρε(x) dx = 1, Rn Rn Rn Rn ε, ψi ≤Z hµt Rn uε(t, x)2ψ(x) dx ≤ (4πt)−n kψkL1k√ρεkL1 = (4πt)−n kψkL1k√ρkL1 εn/2 → 0 (ε → 0). ε → 0 in S ′(Rn) (when t 6= 0), whereas ε)ε∈ ]0,1] can certainly not be weakly convergent as net of finite measures (i.e., in the weak-∗ ε(Rn) = The same obviously holds with ψ ∈ S (Rn), hence also µt (µt topology in the dual of the space of bounded continuous functions on Rn) since hµt 1 6→ 0 as ε → 0 (cf. also [3, Theorem 30.8]). ε, 1i = µt Acknowledgement: The author wishes to express his sincere thanks to two anonymous referees. In particular, we thank the one for hints to additional references of related work, and we are deeply grateful to the second referee for patient comments and clarifications that lead to crucial corrections in Theorem 3.1 and its proof. References [1] R. Adams. Sobolev Spaces. Academic Press, New York, 1975. [2] G. Bal, G. Papanicolaou, and L. Ryzhik. Self-averaging in time reversal for the parabolic wave equation. Stochastics and Dynamics, 2:507 -- 531, 2002. [3] H. Bauer. Measure and integration theory, volume 26 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, 2001. [4] C. Bu. Generalized solutions to the cubic Schrodinger equation. Nonlinear Anal., 27(7):769 -- 774, 1996. [5] J. F. Claerbout. Imaging the Earth's interior. Blackwell Scientific Publications, Palo Alto, 1985. [6] D. L. Cohn. Measure theory. Birkhauser, Boston, 1980. 10 [7] J. F. Colombeau. Une multiplication g´en´erale des distributions. C. R. Acad. Sci. Paris, S´er.I 296:357 -- 360, 1983. [8] J. F. Colombeau. New generalized functions and multiplication of distributions. North-Holland, Amsterdam, 1984. [9] J. F. Colombeau. Multiplication of distributions. A tool in mathematics, numerical engineering and theoretical physics. Number 1532 in Lecture Notes in Math. Springer-Verlag, New York, 1992. [10] F. Colombini and N. Lerner. Hyperbolic operators with non-Lipschitz coefficients. Duke Math. J., 77(3):657 -- 698, 1995. [11] R. Dautray and J.-L. Lions. Mathematical analysis and numerical methods for science and technology. Vol. 5. Springer-Verlag, Berlin, 1992. [12] M. V. de Hoop, G. Hormann, and M. Oberguggenberger. Evolution systems for paraxial wave equations of Schrodinger-type with non-smooth coefficients. J. Differential Equations, 245(6):1413 -- 1432, 2008. [13] S. Dolan, C. Bean, and B. Riollet. The broad-band fractal nature of heterogeneity in the upper crust from petrophysical logs. Geophys. Jour. Int., 132:489 -- 507, 1998. [14] G. B. Folland. Real Analysis. John Wiley and Sons, New York, 1999. [15] C. Garetto. Topological structures in Colombeau algebras: topological C-modules and duality theory. Acta Appl. Math., 88(1):81 -- 123, 2005. [16] C. Garetto and H. Vernaeve. Hilberte-modules: Structural properties and applications to variational problems. Trans. Amer. Math. Soc., 2009. to appear. [17] J. D. E. Grant, E. Mayerhofer, and R. Steinbauer. The wave equation on singular space-times. Comm. Math. Phys., 285(2):399 -- 420, 2009. [18] H. Grosse, M. Oberguggenberger, and I. T. Todorov. Generalized functions for quantum fields obeying quadratic exchange relations. Tr. Mat. Inst. Steklova, 228(Probl. Sovrem. Mat. Fiz.):90 -- 100, 2000. [19] M. Grosser, M. Kunzinger, M. Oberguggenberger, and R. Steinbauer. Geometric theory of generalized func- tions. Kluwer, Dordrecht, 2001. [20] L. Halpern and L. N. Trefethen. Wide-angle one-way wave equations. J. Acoust. Soc. Amer., 84:1397 -- 1404, 1988. [21] F. J. Herrmann. A scaling medium representation. A discussion on well-logs, fractals and waves. PhD thesis, Technische Universiteit Delft, 1997. [22] G. Hormann and M. V. de Hoop. Geophysical modeling with Colombeau functions: Microlocal properties and Zygmund regularity. In A. Delcroix, M. Hasler, J.-A. Marti, and V. Valmorin, editors, Nonlinear Algebraic Analysis. Cambridge Scientific Publishers, 2004. [23] M. Kunzinger and R. Steinbauer. A note on the Penrose junction conditions. Classical Quantum Gravity, 16(4):1255 -- 1264, 1999. [24] X. Li and J. Haury. Characterization of heterogeneities from sonic velocity measurements using the wavelet transform. Expanded Abstracts Soc. Expl. Geophys., 1995:488 -- 491, 1995. [25] J. L. Lions and E. Magenes. Non-homogeneous boundary value problems and applications, volume 1. Springer- Verlag, Berlin, New York, 1972. [26] E. Mayerhofer. On Lorentz geometry in algebras of generalized functions. Proc. Roy. Soc. Edinburgh Sect. A, 138(4):843 -- 871, 2008. [27] M. Nedeljkov, S. Pilipovi´c, and D. Rajter- ´Ciri´c. Heat equation with singular potential and singular data. Proc. Roy. Soc. Edinburgh Sect. A, 135(4):863 -- 886, 2005. [28] M. Oberguggenberger. Hyperbolic systems with discontinuous coefficients: generalized solutions and a trans- mission problem in acoustics. J. Math. Anal. Appl., 142:452 -- 467, 1989. [29] M. Oberguggenberger. Multiplication of distributions and applications to partial differential equations. Long- man Scientific & Technical, 1992. [30] A. Pazy. Semigroups of linear operators and applications to partial differential equations. Springer-Verlag, New York, 1983. [31] J. Rauch. Partial differential equations, volume 128 of Graduate Texts in Mathematics. Springer-Verlag, New York, 1991. [32] A. Saucier and J. Muller. Use of multifractal analysis in the characterization of geological information. Fractals, 1(3):617 -- 628, 1993. [33] R. Steinbauer. The ultrarelativistic Reissner-Nordstrøm field in the Colombeau algebra. J. Math. Phys., 38(3):1614 -- 1622, 1997. [34] R. Steinbauer. Geodesics and geodesic deviation for impulsive gravitational waves. J. Math. Phys., 39(4):2201 -- 2212, 1998. [35] R. Steinbauer and J. A. Vickers. The use of generalized functions and distributions in general relativity. Classical Quantum Gravity, 23(10):R91 -- R114, 2006. [36] B. Z. Steinberg and J. J. McCoy. Toward local effective parameter theories using multiresolution decomposition. Jour. Acoustic Soc. Am., 96(2):1130 -- 1143, 1994. [37] M. Stojanovi´c. Nonlinear Schrodinger equation with singular potential and initial data. Nonlinear Anal., 64(7):1460 -- 1474, 2006. [38] M. Stojanovi´c. Perturbed Schrodinger equation with singular potential and initial data. Commun. Contemp. Math., 8(4):433 -- 452, 2006. 11 [39] L. N. Trefethen and L. Halpern. Well-posedness of one-way wave equations and absorbing boundary conditions. Math. Comp., 47:421 -- 435, 1986. [40] R. van der Hilst, M. V. de Hoop, P. Wang, S.-H. Shim, P. Ma, and L. Tenorio. Seismo-stratigraphy and thermal structure of earth's core-mantle boundary region. Science, 315:1813 -- 1817, 2007. [41] S. C. Wales and J. J. McCoy. A comparison of parabolic wave theories for linearly elastic solids. Wave Motion, 5:99 -- 113, 1983. [42] K. Wapenaar. Seismic reflection and transmission coefficients of a self-similar interface. Geophys. J. Int., 135:585 -- 594, 1998. Fakultat fur Mathematik, Universitat Wien E-mail address: [email protected] 12
1801.08885
6
1801
2018-07-25T07:18:12
Fractional powers and singular perturbations of quantum differential Hamiltonians
[ "math.FA", "math-ph", "math-ph" ]
We consider the fractional powers of singular (point-like) perturbations of the Laplacian, and the singular perturbations of fractional powers of the Laplacian, and we compare such two constructions focusing on their perturbative structure for resolvents and on the local singularity structure of their domains. In application to the linear and non-linear Schr\"{o}dinger equations for the corresponding operators we outline a programme of relevant questions that deserve being investigated.
math.FA
math
FRACTIONAL POWERS AND SINGULAR PERTURBATIONS OF QUANTUM DIFFERENTIAL HAMILTONIANS ALESSANDRO MICHELANGELI, ANDREA OTTOLINI, AND RAFFAELE SCANDONE Dedicated to Gianfausto Dell'Antonio on the occasion of his 85th birthday Abstract. We consider the fractional powers of singular (point-like) pertur- bations of the Laplacian, and the singular perturbations of fractional powers of the Laplacian, and we compare such two constructions focusing on their per- turbative structure for resolvents and on the local singularity structure of their domains. In application to the linear and non-linear Schrodinger equations for the corresponding operators we outline a programme of relevant questions that deserve being investigated. 1. Background: at the edge of fractional quantum mechanics and zero-range interactions At the edge of the theory of quantum Hamiltonians with zero-range interactions, the theory of partial differential operators, and the recent mainstream of fractional quantum mechanics, there are two constructions, each of which is classical in nature, the combination of which has been receiving an increasing attention in the recent times. The first is the construction of fractional powers of a differential operator with non-negative symbol and more generally the fractional power of a non-negative self-adjoint operator on L2(Rd). For concreteness, let us focus on the negative Laplacian −∆ = − d(cid:88) j=1 ∂2 ∂x2 j , x ≡ (x1, . . . , xd) ∈ Rd , that we shall simply call the Laplacian. In this case the definition of (−∆)s/2 for s ∈ R is obvious in terms of the corresponding power of the Fourier multiplier for −∆: ((−∆)s/2f )(x) = (ps(cid:98)f )∨(x) , 2(cid:82) where (Ff )(p) ≡ (cid:98)f (p) = (2π)− d Rd e−ixpf (x)dx is the present convention for In fact, F(−∆)F∗ gives an explicit multiplication form the Fourier transform. and hence an explicit spectral decomposition for −∆ as a non-negative self-adjoint operator on L2(Rd), thus (−∆)s/2 given by the identity above coincides with the construction with functional calculus. The second construction is that kind of perturbation of a given pseudo-differential operator that heuristically amounts to add to it a potential supported at one point only, whence the jargon of singular perturbation [2, 3]. This is a typical restriction- extension construction, where first one restricts the initial operator to sufficiently smooth functions vanishing in neighbourhoods of a given point x0 ∈ Rd, and then one builds an operator extension of such restriction that is self-adjoint on L2(Rd). Date: July 26, 2018. Key words and phrases. Fractional Quantum Mechanics, fractional Laplacian, singular per- turbations of differential operators, Kreın-Visik-Birman self-adjoint extension theory. 1 2 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE This procedure, when the initial operator is −∆, is known to be equivalent to the somewhat more concrete scheme of taking the limit ε ↓ 0 in Schrodinger operators of the form −∆ + Vε, where Vε is a regular potential essentially supported at a scale ε−1 around x0 and magnitude diverging with ε, which shrink to a delta-like profile centred at x0 [2, 3]. Both constructions are well known and relevant in several contexts: Sobolev spaces, fractional Sobolev norms, high and low regularity theory for non-linear PDE, etc., concerning the former; solvable yet realistic models with computable spectral features (eigenvalues, scattering matrix,...) in quantum mechanics, chem- istry, biology, acoustics, concerning the latter. Recently, especially for the solution theory of non-linear Schrodinger equations whose linear part is governed by singular Hamiltonians of point interactions[8, 12], as well as for linear Schrodinger-like equations with singular perturbations of frac- tional powers of the Laplacian [14, 16, 4, 11, 18, 20, 10, 15, 17], the interest has increased around various ways of combining the two constructions above. The goal of this work is to set up the problem in a systematic way for two operations that in a sense do not commute, to make the rigorous constructions of the operators of interest, and to make qualitative and quantitative comparisons. We shall consider, in the appropriate sense that we are going to specify, the class of singular perturbations of the d-dimensional Laplacian, supported at a point x0 ∈ Rd, in the homogeneous and inhomogeneous case, namely (informally speaking), (1.1) hτ = −∆ + singular perturbation at x0 hτ + 1 = −∆ + 1 + singular perturbation at x0 , and then we shall combine this construction with that of fractional powers, thus considering on the one hand the operators (1.2) τ = (−∆ + singular perturbation at x0)s/2 hs/2 (hτ + 1)s/2 = (−∆ + 1 + singular perturbation at x0)s/2 and on the other hand the operators (1.3) k(s/2) τ d(s/2) τ = (−∆)s/2 + singular perturbation at x0 = (−∆ + 1)s/2 + singular perturbation at x0 . All operators in (1.1) -- (1.3) above are meant to be self-adjoint. Let us remark that hs/2 τ and (hτ + 1)s/2 are going to be genuine fractional powers of a non-negative self-adjoint operator on L2(R3) (to this aim one has to consider non-restrictively only those singular perturbations that produce non-negative oper- ators), whereas the different notation for the superscript s/2 in k(s/2) and in d(s/2) is to indicate that the latter operators are instead singular perturbations of s/2-th powers (and not fractional powers of singular perturbations). , for some J ∈ N, is going to be a parameter that qualifies one element in the infinite family of self-adjoint realisations of the considered singular perturbation, with the customary convention that 'τ = ∞' denotes the absence of perturbation. In all cases τ ∈ (R ∪ {∞})J 2 τ τ As for the choice of the point x0, there is no loss of generality in choosing x0 = 0, which we shall do henceforth. The knowledge of singular quantum Hamiltonians of the type (1.1) is well es- tablished in the literature and we review them in Section 2. The study of their fractional powers, hence of the operators of the type (1.2), has only started re- cently and in the second part of Section 2 we give an account of the main known facts about them. FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 3 In comparison to (1.1) and (1.2), we shall then discuss the rigorous construction of operators of the type (1.3) (Sections 3 -- 7). Our presentation will have two main focuses, which reflect into the formulation of our results: 1. to qualify the nature of the perturbation in the resolvent sense (finite rank vs infinite-rank perturbations); 2. to qualify the natural decomposition of the domain of the considered oper- ators into a regular component and a singular component, and to determine the boundary condition constraining such two components. The first issue is central for deducing an amount of properties from the unperturbed to the perturbed operators. The second issue also arises naturally, as one can see heuristically that the considered operators must act in an ordinary way on those functions supported away from the perturbation centre, and therefore their domains must contain a subspace of H s-regular functions, where s is the considered power, next to a more singular component that is the signature of the perturbation. In this respect we are going to highlight profound differences between two con- structions, say, (−∆ + 1 + singular perturbation at x0)s/2 (−∆ + 1)s/2 + singular perturbation at x0 , that are therefore 'non-commutative'. We organized the material as follows: nians is presented in Section 2; • the analysis of operators of fractional power of point interaction Hamilto- • on the other hand, the construction of singular perturbations of fractional Laplacians is presented in Sections 3 through 7, with the general set-up in Section 3, the detailed discussion of the paradigmatic scenario of rank-one perturbations in Sections 4 and 5 (homogeneous case) and in Section 6 (in- homogeneous case), and an outlook on high-rank perturbations in Section 7; • last, in Section 8 we outline an amount of relevant questions that deserve being investigated in application to the linear and non-linear Schrodinger equations governed by the operators constructed in this work. 2. Singular perturbations of the Laplacian and their fractional powers In this Section we qualify the operators hτ , hs/2 , and (hτ + 1)s/2 of (1.1)-(1.2). Whereas the former is well known in all dimensions in which it is not trivial, namely d = 1, 2, 3, the latter two operators have been studied mainly in three dimensions, thus in this Section we choose d = 3 for our presentation. τ (cid:16) (cid:17)∨ λx e−√ 4πx = 1 1 (2π) 3 2 p2 + λ (x) , x, p ∈ R3 , For chosen λ > 0 we set (2.1) Gλ(x) := whence (−∆ + λ)Gλ = δ(x) (2.2) as a distributional identity on R3. We also set h := −∆ (cid:22) C∞ (2.3) as an operator closure with respect to the Hilbert space L2(R3). 0 (R3\{0}) 4 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE As well known[13], h is densely defined, closed, and symmetric on L2(R3), with (2.4) D(h) = f ∈ H 2(R3) hf = −∆f . Its Friedrichs extension given by the self-adjoint Laplacian on L2(R3) with domain H 2(R3), and its adjoint is the operator η (p2 + λ)2 + η, ξ ∈ C ξ p2 + λ f λ ∈ D(h) ,  (cid:110) , R3 (cid:12)(cid:12)(cid:12)(cid:90) (cid:111) (cid:98)f (p) dp = 0 g ∈ L2(R3) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:98)g(p) =(cid:99)f λ(p) + (cid:17) (cid:16)(cid:99)f λ(p) + η . = H 2(R3) (cid:117) span{Gλ} (p2 + λ)2 D(h∗) = (2.5) (cid:92) (h∗ + λ)g) (p) = (p2 + λ) ( The structure in Eq. (2.5) is typical of a well-known decomposition (see, e.g., Eq. (2.5) and (2.10) in [7]). We observe that D(h∗) does not depend on λ, only the splitting of g ∈ D(h∗) into a H 2-function plus a less regular component does. The last identity in (2.5) may be re-interpreted distributionally as h∗g = −∆g + (2π) 3 2 ξ δ(x)g , where neither −∆g nor (2π) 3 2 ξ δ(x)g belongs to L2(R3), but their sum does. The fact that ker(h∗ + λ1) = span{Gλ} indicates that h has deficiency index 1 and hence admits a one-(real-)parameter family of self-adjoint extensions. They can be classified in terms of the standard parametrisation of the Kreın-Visik-Birman self-adjoint extension theory (see, e.g.,[7, Sec. 3]), identifying each of them as a restriction of h∗ by imposing in (2.5) Birman's self-adjointness condition η = τ ξ for some τ ∈ R ∪ {∞} (see, e.g., [13, Theorem 1 and Corollary 1]), whence the following Theorem. Theorem 2.1. Let λ > 0. (i) The self-adjoint extensions in L2(R3) of the operator h form the family (hτ )τ∈R∪{∞}, where h∞ = hF , the Friedrichs extension, and all other (proper) extensions are given by hτ = h∗ (cid:22) D(hτ ), where g ∈ L2(R3) (cid:110) g = F λ + 8π (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:98)g(p) =(cid:99)f λ(p) + ξ ∈ C , √ τ F λ(0) Gλ λ ξ p2 + λ τ ξ (p2 + λ)2 + f λ ∈ D(h) (cid:12)(cid:12)(cid:12) F λ ∈ H 2(R3) (cid:111) .  (2.6) D(hτ ) := = (ii) Each extension is semi-bounded from below and (2.7) inf σ(hτ + λ1) (cid:62) 0 ⇔ τ (cid:62) 0 inf σ(hτ + λ1) > 0 ⇔ τ > 0 (hτ + λ1) is invertible ⇔ τ (cid:54)= 0 . (iii) For each τ ∈ R the quadratic form of the extension hτ is given by (2.8) (2.9) D[hτ ] = H 1(R3) (cid:117) span{Gλ} hτ [F λ + κλGλ] = (cid:107)∇F λ(cid:107)2 √ τ 8 L2(R3) − λ(cid:107)F λ + κλGλ(cid:107)2 κλ2 for any F λ ∈ H 1(R3) and κλ ∈ C. + λ L2(R3) + λ(cid:107)F λ(cid:107)2 L2(R3) FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 5 Remark 2.2. The τ -parametrisation depends on the initial choice of the λ-shift and thus the same extension is described by infinitely many different pairs (λ, τ ) -- of course with a unique τ once λ is chosen. This is clear by inspecting the boundary condition at x = 0 between regular and singular component of a generic g ∈ D(hτ ): for any two pairs (λ, τ ) and (λ(cid:48), τ(cid:48)) identifying the same extension hτ , owing to (2.6) one has (p2+λ)(p2+λ(cid:48)) , ξ √ ξ(cid:48) ξ(cid:48) p2 + λ := 8π λ τ(cid:48) (2π)3 p2 + λ(cid:48) (cid:98)g = (cid:99)F λ + (cid:82) R3 (cid:99)F λdp = τ (2π)3 (cid:82) R3(cid:100)F λ(cid:48) (cid:0)1 + 2 τ − 2λ√ = (cid:100)F λ(cid:48) (cid:0)(cid:82) R3 (cid:99)F λdp(cid:1) , (cid:100)F λ(cid:48) := (cid:99)F λ + ξ(cid:48)(λ(cid:48)−λ) (cid:0)1 + 2 (cid:1) , (cid:1), or equivalently τ(cid:48) − 2λ(cid:48)√ λ(cid:48) √ dp = τ(cid:48) (2π)3 λ(cid:48) ξ(cid:48) · τ λ(cid:48) √ τ(cid:48)√ λ λ(cid:48)−√ √ √ λ(cid:48) τ(cid:48)√ λ λ(cid:48)−√ 8π λ √ √ λ( √ λ( τ = √ ξ λ) τ 8π + λ) λ . with and also whence necessarily 1 = τ (2.10) Thus, when referring to the extension hτ , we shall always implicitly declare the choice of λ, and any other λ(cid:48) and τ(cid:48) satisfying (2.10) actually identify the same extension. In the literature it is customary to find the second expression of (2.6) with the alternative extension parameter α defined by τ − 2λ √ λ 8π α := (2.11) thus re-writing (hτ )τ∈R∪{∞} as (hα)α∈R∪{∞} with hα := h∗ (cid:22) D(hα) (see [13, Re- mark 3] and [2, Eq. (I.1.1.26)]). Of course, owing to Remark 2.2, in particular to formula (2.10), the parameter α identifies unambiguously an extension irrespec- tively of the shift λ chosen for the explicit domain decomposition. From this and a bit of further spectral analysis [2, Sec. I.1.1] one then deduces the following. , Theorem 2.3. (i) The self-adjoint extensions in L2(R3) of the operator h form the family (hα)α∈R∪{∞}, where h∞ = hF , the Friedrichs extension, and all other (proper) extensions are given by hα := h∗ (cid:22) D(hα), where, for arbitrary λ > 0 D(hα) = g = F λ + (α + √ 4π )−1F λ(0) Gλ λ (cid:12)(cid:12)(cid:12) F λ ∈ H 2(R3) (cid:111) (2.12) (hα + λ) g = (−∆ + λ) F λ . (ii) The spectrum of hα is given by σess(hα) = σac(hα) = [0, +∞) , σsc(hα) = ∅ , (2.13) σp(hα) = ∅ {−(4πα)2} if α ∈ [0, +∞] if α ∈ (−∞, 0) . (cid:110) (cid:40) The negative eigenvalue −(4πα)2, when it exists, is non-degenerate and the corresponding eigenfunction is x−1e−4πα x. Thus, α (cid:62) 0 corresponds to a non-confining, 'repulsive' contact interaction. 6 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE (iii) For each α ∈ R the quadratic form of the extension hα is given by D[hα] = H 1(R3) (cid:117) span{Gλ} (2.14) hα[F λ + κλ Gλ] = −λ(cid:107)F λ + κλ Gλ(cid:107)2 L2(R3) + (cid:107)∇F λ(cid:107)2 L2(R3) + λ(cid:107)F λ(cid:107)2 L2(R3) +(cid:0)α + (cid:1)κλ2 √ λ 4π for arbitrary λ > 0. (iv) The resolvent of hα is given by (2.15) (hα + λ1)−1 = (−∆ + +λ1)−1 + (α + √ 4π )−1Gλ(cid:105)(cid:104)Gλ λ for arbitrary λ > 0. For the operators hα + λ1 with λ sufficiently large, and the operators hα with α (cid:62) 0, the self-adjoint functional calculus defines the powers (hα + λ1)s/2 or hs/2 for s ∈ R. A surely relevant regime is surely s ∈ [0, 2]: the integer powers s = 2, 1, 0 correspond, respectively, to the considered operator, its square root whose domain is then the form domain of the considered operator, and the identity; the fractional powers in between are of interest when one needs to discuss the corresponding linear or non-linear dynamics in spaces of convenient fractional regularity. α From the thorough analysis of such fractional powers made in [8], one has the following. Theorem 2.4. Let α (cid:62) 0, λ > 0, and s ∈ [0, 2], and set (i) One has α ) = D(hα + λ1)s/2 (cid:101)H s α(R3) := D(hs/2 (cid:107)g(cid:107)(cid:101)H s α(R3) := (cid:107)(hα + λ1)s/2g(cid:107)L2(R3) . (cid:101)H s α(R3) = H s(R3) (cid:107)ψ(cid:107)(cid:101)H s α(R3) ≈ (cid:107)ψ(cid:107)H s(R3) (cid:101)H s α(R3) = H s(R3) (cid:117) span{Gλ} (cid:110) (cid:107)F λ + κλ Gλ(cid:107)(cid:101)H s g = F λ +(cid:0)α + (cid:1)−1 α(R3) ≈ (cid:107)F λ(cid:107)H s(R3) + (1 + α)κλ (cid:101)H s α(R3) = (cid:107)F λ + κλ Gλ(cid:107)(cid:101)H s α(R3) ≈ (cid:107)F λ(cid:107)H s(R3) where here '≈' denotes the equivalence of norms. F λ(0) Gλ λ 4π √ (2.16) (2.17) (2.18) (2.19) if s ∈ [0, 1 2 ) , (cid:12)(cid:12)(cid:12) F λ ∈ H s(R3) (cid:111) if s ∈ ( 1 2 , 3 2 ) , if s ∈ ( 3 2 , 2) , (ii) For g ∈ D(hs/2 α ), (hα + λ1)s/2g = (−∆ + λ1)s/2g (2.20) − 4 sin sπ 2 (iii) One has the resolvent identity (2.21) (hα + λ1)−s/2 = (−∆ + λ1)−s/2 − (cid:90) +∞ 0 dt (cid:90) +∞ 0 ts/2 (cid:104)Gλ+t, g(cid:105)L2(R3) √ 4πα + λ + t Gλ+t . − s 4t 4πα+ 2 sin √ sπ 2 λ+t Gλ+t(cid:105)(cid:104)Gλ+t . dt 2 and s = 3 The transition cases s = 1 2 too can be qualified, however with less explicit formulas -- see [8, Prop. 8.1 and 8.2]. α(R3) decomposes into a regular H s-component and a singular component with local singularity x−1, precisely as the domain of hα itself, and for s > 3 2 regular and singular parts are constrained by Thus, as described in Theorem 2.4, for s > 1 2 , (cid:101)H s FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 7 a local boundary condition among them of the same type as in (2.12); low powers s < 1 2 , instead, only produce domains where no leading singularity can be singled out. It is remarkable that one has such an explicit and clean knowledge of the frac- tional powers of the singular perturbations of the Laplacian: the singular pertur- bation yields an operator hα that is not a (pseudo-)differential operator any longer and its powers are not simply recovered as Fourier multipliers, nor is it a priori ob- vious how the fractional power affects the local boundary condition between regular and singular components of elements in D(hα). α , and (hα + λ1)s/2 for s ∈ It is also worth remarking that whereas hα, hs/2 2 , 2) share the same singular behaviour x−1 of the functions in their domain, one ( 1 noticeable difference is given by equations (2.15) and (2.21): indeed, (hα + λ1)−1 is a rank-one perturbation of (−∆ + λ1)−1, while (hα + λ1)−s/2 is not a finite rank perturbation of (−∆ + λ1)−s/2. 3. Singular perturbations of the fractional Laplacian: general setting for the homogeneous case τ In comparison to hs/2 , the fractional power of a singular perturbation of the Laplacian, we start discussing in this Section the rigorous construction of operators of the type k(s/2) , as introduced informally in (1.3), namely the self-adjoint singular perturbations of the fractional power of the Laplacian. Then, in Section 4 we shall consider the concrete cases of dimension d = 3, and in Section 5 the case d = 1. τ For chosen d ∈ N, λ > 0, and s ∈ R we set (3.1) Gs,λ(x) := 1 (2π) d 2 1 ps + λ (x) , x, p ∈ Rd , (cid:16) (cid:17)∨ whence (3.2) ((−∆)s/2 + λ) Gs,λ = δ(x) as a distributional identity on Rd. We also set (3.3) k(s/2) := (−∆)s/2 (cid:22) C∞ 0 (Rd\{0}) as an operator closure with respect to the Hilbert space L2(Rd). Thus, in compar- ison to Section 2, when d = 3 one has G2,λ = Gλ and k(1) = h. The domain of k(s/2), as a consequence of the operator closure in (3.3), is a space of functions with H s-regularity and vanishing conditions at x = 0 for each function and its partial derivatives. The amount of vanishing conditions depends on d and s, to classify which we introduce the intervals (cid:40) (3.4) I (d) n := (0, d 2 ) 2 + n − 1, d ( d n = 0 2 + n) n = 1, 2, . . . For our purposes it is convenient to use momentum coordinates to express the vanishing conditions that qualify the domain of k(s/2): thus, with the notation p ≡ (p1, . . . , pd) ∈ Rd, by means of an approximation argument (see Appendix A) 8 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE we find (3.5) D(k(s/2)) = H s   = 0 (Rd\{0}) = C∞ H s(Rd) 0 (Rd\{0}) (cid:107) (cid:107)Hs f ∈ H s(Rd) such that Rd pγ1 (cid:82) d (cid:98)f (p) dp = 0 γ1, . . . , γd ∈ N0 , (cid:80)d d (cid:98)f (p) dp = 0 is the same as(cid:0) ∂γ1 j=1 γj (cid:54) n − 1 1 ··· pγd 0 if s ∈ I (d)  if s ∈ I (d) ··· ∂γd ∂x γd d ∂xγ1 1 Clearly,(cid:82) Rd pγ1 1 ··· pγd notation x ≡ (x1, . . . , xd) ∈ Rd. n , n = 1, 2, . . . f(cid:1)(0) = 0, with the The expression of D(k(s/2)) for the endpoint values s = d 2 + n require an amount of extra analysis besides the arguments proof of (3.5): we do not discuss it here, an omission that does not affect the conceptual structure of our presentation. Being densely defined, closed, and positive, either the symmetric operator k(s/2) is already self-adjoint on L2(Rd), or it admits infinitely many self-adjoint extensions. As well known, the infinite multiplicity of such extensions is quantified by the deficiency index of k(s/2), which is the quantity J (s, d) := dim ker(cid:0)(k(s/2))∗ + λ1(cid:1) (3.6) for one, and hence for all λ > 0. The self-adjointness of k(s/2) is equivalent to J (s, d) = 0. It is not difficult to compute J (s, d) for generic values of d and s and to identify a natural basis of the J (s, d)-dimensional space ker(cid:0)(k(s/2))∗ + λ1(cid:1). Lemma 3.1. For given d ∈ N and s > 0, (3.7) s ∈ I (d) n ⇒ J (s, d) = In particular, when s ∈ I (d) (cid:110) ker(cid:0)(k(s/2))∗ + λ1(cid:1) = span (3.8) n for some n ∈ N, then uλ γ1,...,γd (cid:18)d + n − 1 (cid:19) (cid:12)(cid:12)(cid:12) γ1, . . . , γd ∈ N0 , d . (cid:111) , γj (cid:54) n − 1 d(cid:88) j=1 where (3.9) 1 ··· pγd pγ1 ps + λ It is worth noticing, comparing (3.1) and (3.9), that γ1,...,γd (p) := d . (cid:99)uλ s ∈ I (d) to uλ 0,...,0 = (2π) (3.10) Proof of Lemma 3.1. When s ∈ I (d) then ker(cid:0)(k(s/2))∗ + λ1(cid:1) is trivial and J (s, d) = 0, consistently with (3.7). When n , n = 1, 2, . . . , then u ∈ ker(cid:0)(k(s/2))∗ +λ1(cid:1) = ran(cid:0)k(s/2) +λ1(cid:1)⊥ 0 , we see from (3.5) that k(s/2) is self-adjoint: is equivalent 2 Gs,λ . d and one argues from (3.5) that ker(cid:0)(k(s/2))∗ + λ1(cid:1) is spanned by linearly indepen- ∀f ∈ D(k(s/2)) 0 = dent functions of the form uλ j=1 γj (cid:54) n − 1. Such functions are as (cid:90) R3(cid:98)u(p)(ps + λ)(cid:98)f (p)dp , with(cid:80)d γ1,...,γd FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 9 (cid:18)d + n − 1 (cid:19) d many as the linearly independent monomials in d variables with degree at most equal to n − 1, and therefore their number equals (cid:3) . The knowledge of ker(cid:0)(k(s/2))∗+λ1(cid:1) and of the inverse of the Friedrichs extension of k(s/2) are the two inputs for the Kreın-Visik-Birman extension theory (see, e.g., [7, Sec. 3]), by means of which we can produce the whole family of self-adjoint extensions of k(s/2). Such a construction is particularly clean in the case, relevant in applications, of deficiency index one: the comprehension of this case is instructive to understand the case of higher deficiency index. Moreover, as we shall see, in this case the self- adjoint extensions of k(s/2) turns out to be rank-one perturbations, in the resolvent sense: we will use the jargon J = 1 or 'rank one' interchangeably. the J (s, d) = 1 scenario when s ∈ I (d) high-J scenario. This corresponds to analysing the regimes s ∈ ( 1 s ∈ (1, 2) when d = 2, s ∈ ( 3 Therefore, in this work we make the presentational choice to discuss in detail 1 , deferring to Section 7 an outlook on the 2 ) when d = 1, 2 , 3 2 ) when d = 3, etc. Indeed, when s ∈ I (d) (3.10) that ker(cid:0)(k(s/2))∗ +λ1(cid:1) = span{Gs,λ}, and the function Gs,λ may or may not The construction of the self-adjoint extensions of k(s/2) in any such regimes is technically the very same, irrespectively of d, except for a noticeable peculiarity when d = 1, as opposite to d = 2, 3, . . . and hence J (s, d) = 1, we know from Lemma 3.1 and have a local singularity as x → 0. As follows from the d-dimensional distributional identity 2 , 5 1 2 s 2 Γ( s 2 ) 1 ps = 2 d−s 2 Γ( d−s 2 ) , s ∈ (0, d) , (cid:17) (cid:92)(cid:16) 1 xd−s Gs,0 has a singularity ∼ x−(d−s) when s < d, it has a logarithmic singularity when s = d, and it is continuous at x = 0 when s > d, with asymptotics (3.11) xd−s Gs,λ(x) Gs,λ(x) := x→0−−−−→ Λ(d) x→0−−−−→ Gs,λ(0) = (cid:0)2d−1π (2π) d s Γ( d−s 2 ) 2 2s− d , 2 Γ( d 2 ) 2 −1Γ( d 2 ) λ d (cid:1)−1 s ∈ (0, d) , s > d . s−d s s sin πd s Now, all the considered regimes s ∈ (1, 2) when d = 2, s ∈ ( 3 2 ) when d = 3, etc. lie below the transition value s = d between the local singular and the local regular behaviour of Gs,λ, whereas the regime s ∈ ( 1 2 ) when d = 1 lies across the transition value s = 1. of ker(cid:0)(k(s/2))∗ + λ1(cid:1) for higher deficiency index J (s, d). The same type of distinction clearly occurs for the spanning functions (3.8)-(3.9) 2 , 3 2 , 5 In the present context, the peculiarity described above when d = 1 results in certain different steps of the construction of the self-adjoint extensions of k(s/2) and ultimately in the type of parametrisation of such extensions, as we shall see. Therefore, we articulate our discussion on the extensions of k(s/2) when the deficiency index is one discussing first the three-dimensional case (Section 4) and then the one-dimensional case (Section 5). As commented already, for generic d (cid:62) 2 the discussion and the final results are completely analogous to d = 3. 10 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE 4. Rank-one singular perturbations of the fractional Laplacian: homogeneous case in dimension three In terms of the general discussion of Sec. 3, we consider here the operator k(s/2) 2 ). k(s/2) acts as the fractional Laplacian (−∆)s/2 on the 2 , 5 on L2(R3) when s ∈ ( 3 domain (cid:110) (cid:12)(cid:12)(cid:12)(cid:90) R3 (cid:111) (cid:98)f (p) dp = 0 (4.1) D(k(s/2)) = f ∈ H s(R3) and its deficiency index is 1. One has the following construction. Theorem 4.1. Let s ∈ ( 3 2 , 5 2 ) and λ > 0. (i) The self-adjoint extensions in L2(R3) of the operator k(s/2) form the family (k(s/2) )τ∈R∪{∞}, where k(s/2)∞ is its Friedrichs extension, namely the self- adjoint fractional Laplacian (−∆)s/2, and all other extensions are given by τ D(cid:0)k(s/2) (cid:1) := τ (4.2) (4.3) and = (cid:0)k(s/2) τ τ ξ g ∈ L2(R3) (cid:110) + λ1(cid:1)g := F−1(cid:16) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:98)g(p) =(cid:99)f λ(p) + ξ ∈ C , f λ ∈ H s(R3) , (cid:82) (cid:16)(cid:99)f λ + g = F λ + 2πs2 sin( 3π τ (s−3) (ps + λ) F λ(0) Gs,λ s )λ2− 3 τ ξ s (ps + λ)2 + ξ ps + λ R3(cid:99)f λ(p) dp = 0 (cid:12)(cid:12)(cid:12) F λ ∈ H s(R3) (cid:111) (cid:17)(cid:17)  , (ps + λ)2 . (ii) Each extension is semi-bounded from below and τ inf σ(k(s/2) inf σ(k(s/2) + λ1) (cid:62) 0 ⇔ τ (cid:62) 0 + λ1) > 0 ⇔ τ > 0 + λ1) is invertible ⇔ τ (cid:54)= 0 . τ (k(s/2) τ (4.4) (iii) For each τ ∈ R the quadratic form of the extension k(s/2) τ is given by (4.5) (4.6) k(s/2) τ D[k(s/2) τ ] = H s 2 (R3) (cid:117) span{Gs,λ} [F λ + κλGs,λ] = (cid:107)∇ s L2(R3) − λ(cid:107)F λ + κλGs,λ(cid:107)2 2 F λ(cid:107)2 +λ(cid:107)F λ(cid:107)2 τ (s − 3) L2(R3) κλ2 L2(R3) + 2πs2λ2− 3 s sin( 3π s ) for any F λ ∈ H s/2(R3) and κλ ∈ C. (iv) For τ (cid:54)= 0, one has the resolvent identity + λ1)−1 = ((−∆)s/2 + λ1)−1 + (k(s/2) (4.7) s )λ2− 3 s 2πs2 sin( 3π τ (s − 3) Gs,λ(cid:105)(cid:104)Gs,λ . τ Proof. The whole construction is based upon the Kreın-Visik-Birman self-adjoint extension of k(s/2) + λ1 is the Fourier multiplier (ps + λ), one has the following formula for the adjoint (see, e.g., [7, Theorem 2.2]): extension scheme. Since ker(cid:0)(k(s/2))∗ + λ1(cid:1) = span{Gs,λ} and the Friedrichs  g ∈ L2(R3) D(cid:0)(k(s/2))∗(cid:1) = (cid:0)(k(s/2))∗ + λ1(cid:1)g = F−1(cid:16) (cid:98)g(p) =(cid:99)f λ(p) + η, ξ ∈ C , f λ ∈ H s(R3) , (cid:82) (cid:16)(cid:99)f λ + (ps + λ)2 + ps + λ R3(cid:99)f λ(p) dp = 0 (ps + λ) (cid:17)(cid:17) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) η η ξ . (ps + λ)2 FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 11 transform reads Each element of the one-parameter family of self-adjoint extensions of k(s/2) is identified (see, e.g., [7, Theorem 3.4]) by the Birman self-adjointness condition η = τ ξ for some τ ∈ R ∪ {∞}. This establishes the first line of (4.2). Setting (cid:99)F λ := (cid:99)f λ + (ps + λ)−2τ ξ, the boundary condition between F λ and ξ in Fourier (cid:99)F λ(p) dp = ξ 2(cid:82) R3 (cid:99)F λdp, and using (3.1) with d = 3, the second line 4π2τ (s−3) s sin( 3π s ) is a restriction of (k(s/2))∗, from the above action of R3 Then, from F λ(0) = (2π)− 3 of (4.2) follows. Since k(s/2) the adjoint one deduces (4.3). This completes the proof of part (i). s2λ2− 3 (cid:90) (*) . τ Part (ii) lists standard facts of the Kreın-Visik-Birman theory -- see [7, Theorems The quadratic form is characterised in the extension theory [7, Theorem 3.6] by 3.5 and 5.1]. the formulas D[k(s/2) whence (4.5), and (k(s/2) whence (4.6). The proof of part (iii) is completed. ] (cid:117) ker(cid:0)(k(s/2))∗ + λ1(cid:1) ('F ' stands for Friedrichs), ] = D[k(s/2) τ +λ1)[F λ+κλGs,λ] = ((−∆)s/2+λ1)[F λ]+τκλ2(cid:107)Gs,λ(cid:107)2 F τ L2(R3), Kreın's resolvent formula for deficiency index 1 [7, Theorem 6.6] prescribes f (k(s/2) τ + λ1)−1 = ((−∆)s/2 + λ1)−1 + βλ,τ Gs,λ(cid:105)(cid:104)Gs,λ for some scalar βλ,τ to be determined, whenever (k(s/2) τ (cid:54)= 0. Thus, for a generic h ∈ L2(R3), the element g := (k(s/2) reads, in view of (4.2) and of the resolvent formula above, τ +λ1) is invertible, hence for τ +λ1)−1h ∈ D(k(s/2) ) τ (cid:98)g(p) = (cid:99)F λ(p) + (cid:98)h(p) (cid:99)F λ(p) := ξλ ps + λ (cid:90) (cid:98)h(q) βλ,τ (2π)3 , R3 dq . ξλ := ps + λ τ (s−3) 2πs2 sin( 3π qs + λ The boundary condition (*) for F λ and ξλ then implies 1 = βλ,τ , s )λ2− 3 which determines βλ,τ and proves (4.7), thus completing also the proof of (iv). (cid:3) In analogy to what argued in Remark 2.2, the τ -parametrisation of the family (k(s/2) )τ∈R∪{∞} depends on the initially chosen shift λ > 0, meaning that with a different choice λ(cid:48) > 0 the same self-adjoint realisation previously identified by τ with shift λ is now selected by a different extension parameter τ(cid:48). In certain contexts it is more convenient to switch onto a natural parametrisation that identifies one extension irrespectively of the infinitely many different pairs (λ, τ ) attached to it by the parametrisation of Theorem 4.1. We shall do it in the next Theorem: observe that indeed, as compared to Theorem 4.1, here below λ > 0 is arbitrary. τ s Theorem 4.2. Let s ∈ ( 3 2 , 5 2 ). (i) The self-adjoint extensions in L2(R3) of the operator k(s/2) form the family (k(s/2) )α∈R∪{∞}, where k(s/2)∞ is its Friedrichs extension, namely the self- adjoint fractional Laplacian (−∆)s/2, and all other extensions are given, for arbitrary λ > 0, by α  g = F λ + D(k(s/2) α ) = (4.8) F λ(0) α − λ 3 s −1 2πs sin( 3π s ) Gs,λ (k(s/2) α + λ) g = ((−∆)s/2 + λ) F λ . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) F λ ∈ H s(R3)  12 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE (ii) For each α ∈ R the quadratic form of the extension k(s/2) α is given by (4.9) (4.10) k(s/2) α D[k(s/2) α ] = H s 2 (R3) (cid:117) span{Gs,λ} [F λ + κλGs,λ] = (cid:107)∇ s L2(R3) − λ(cid:107)F λ + κλGs,λ(cid:107)2 L2(R3) 2 F λ(cid:107)2 L2(R3) +(cid:0)α − λ −1 3 s 2πs sin( 3π s ) (cid:1)κλ2 + λ(cid:107)F λ(cid:107)2 for arbitrary λ > 0. (iii) The resolvent of k(s/2) α is given by (4.11) (k(s/2) α + λ1)−1 = ((−∆)s/2 + λ1)−1 +(cid:0)α − λ −1 3 s 2πs sin( 3π s ) (cid:1)−1 Gs,λ(cid:105)(cid:104)Gs,λ for arbitrary λ > 0. (iv) Each extension is semi-bounded from below, and (4.12) (4.13) σsc(k(s/2) α ) = ∅ , σess(k(s/2) α σdisc(k(s/2) α ) = ) = σac(k(s/2) α ) = [0, +∞) , (cid:40) ∅ α = −(cid:0)2πα s sin(− 3π s )(cid:1) s if α (cid:62) 0 if α < 0 , {E(s) α } where the eigenvalue E(s) α is non-degenerate and is given by 3−s , the (non-normalised) eigenfunction being Gs,λ=E(s) α . E(s) Proof. Reasoning as in Remark 2.2, we seek for the relation τ = τ (λ) that ensures that all the pairs (λ, τ (λ)), with λ > 0, preserve the decomposition (4.2)-(4.3) and thus label the same element of the family of extensions. For chosen λ and τ , a function g ∈ D(k(s/2) τ ) decomposes uniquely as ξ ps + λ , F λ ∈ H s(R3) ξ ∈ C (cid:99)F λdp = ξ 4π2τ (s−3) s sin( 3π s ) s2λ2− 3 . Let now λ(cid:48) > 0 and τ(cid:48) ∈ R be such that for the same function g in the domain of the same self-adjoint realisation k(s/2) one also has τ (cid:98)g = (cid:99)F λ + (cid:98)g = (cid:100)F λ(cid:48) + (cid:90) (cid:90) R3 R3 (cid:100)F λ(cid:48) , , ξ(cid:48) F λ(cid:48) ∈ H s(R3) ξ(cid:48) ∈ C The new splitting of g is equivalent to ps + λ(cid:48) , dp = ξ(cid:48) 4π2τ(cid:48)(s−3) (cid:48)2− 3 s sin( 3π s ) s2λ . ξ(cid:48) = ξ , F λ(cid:48) and the boundary condition for F λ(cid:48) ξ(cid:48) (cid:16) (cid:90) ξ 4π2τ (s−3) s sin( 3π s ) s2λ2− 3 + ps + λ R3 ξ − ps + λ = F λ + and ξ(cid:48) is equivalent to dp = ξ(cid:48) (cid:17) − ξ(cid:48) ps + λ(cid:48) (*) ξ(cid:48) ps + λ(cid:48) , 4π2τ(cid:48)(s−3) (cid:48)2− 3 s sin( 3π s ) s2λ . Let us analyze the integral in (*). Both summands in the integrand diverge, with two identical divergences that cancel out. Thus, by means of the identity r2(rs + λ)−1 = r2−s − λr2−s(rs + λ)−1, one has (cid:90) (cid:16) (cid:17) 1 ps + λ − 1 ps + λ(cid:48) R3 dp = 4π lim R→+∞ (cid:16)(cid:90) R (cid:16)(cid:90) R 0 (cid:90) R (cid:90) R 0 − r2 dr − rs + λ λ(cid:48) dr (cid:17) r2 rs + λ(cid:48) dr λ dr rs−2(rs + λ) (cid:17) = 4π lim 0 R→+∞ 4π2 s s sin( 3π s ) λ 1− 3 rs−2(rs + λ(cid:48)) − λ(cid:48) 1− 3 4π2 s s sin( 3π s ) 0 . = FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 13 Plugging the result of the above computation into (*) yields τ (3 − s) − sλ τ(cid:48)(3 − s) − sλ(cid:48) λ2− 3 s = λ(cid:48)2− 3 s , which shows, in complete analogy to (2.10) when s = 2, that all pairs (λ, τ ) such that (**) − τ (3 − s) − sλ s )λ2− 3 2πs2 sin( 3π s =: α indeed label the same extension (the pre-factor −2πs2 sin( 3π s ) having being added for convenience). Thus, α ∈ R∪{∞} defined in (**) is the natural parametrisation we were aiming for (and the Friedrichs case τ → +∞ corresponds to α → +∞). = (cid:0)α − (cid:1)−1 Upon replacing 2πs2 sin( 3π τ (s−3) s )λ2− 3 s 1 2πs sin( 3π s )λ1− 3 s in the formulas of Theorem 4.1 we deduce at once all formulas of parts (i), (ii), and (iii), together of course with the certainty, proved above, that the decompositions are now λ- independent. Since the deficiency index is 1, and hence all extensions are a rank-one per- turbation, in the resolvent sense, of the self-adjoint fractional Laplacian, then all extensions have the same essential spectrum [0, +∞) of the latter, and additionally may have at most one negative non-degenerate eigenvalue, in any case all exten- sions are semi-bounded from below -- all these being general facts of the extension theory, see, e.g., [7, Theorem 5.9 and Corollary 5.10]. This proves, in particular, the first line in (4.12). The occurrence of a negative eigenvalue Eα = −λ of an extension k(s/2) , for some λ > 0, can be read out from the resolvent formula (4.11) as the pole of (k(s/2) α α + λ1)−1, that is, imposing α − i.e., = 0 , 1 2πs sin( 3π s 3−s α = −λ s )λ1− 3 s (cid:0)2πs sin(− 3π s )(cid:1)−1 λ = (cid:0)2πα s sin(− 3π s )(cid:1) s 3−s . . The identity above can be only satisfied by some λ > 0 when α < 0, because sin(− 3π s ) > 0, in which case Alternatively, one can argue from (4.2)-(4.3) that the eigenvalue −λ must corre- 1ps+λ )∨, that is, an element of the domain with only spond to the eigenfunction ( singular component, and to the parameter τ = 0, hence with f λ ≡ 0 in the notation therein. Then, setting τ = 0 in (**) yields the same condition above on α and λ. This proves (4.13) and the second line in (4.12) when α < 0, and it also qualifies the eigenfunction. When such a bound state is absent, and therefore when α (cid:62) 0, for what argued ) = [0, +∞). This proves the second line in (cid:3) before one has σ(k(s/2) (4.12) when α (cid:62) 0, and completes the proof of part (iv). ) = σess(k(s/2) α α Mirroring the observations made in the conclusions of Section 2, we see that the elements of the domains of both the operators hs/2 (the fractional power of the singular perturbation of (−∆)) and k(s/2) (the singular perturbation of the fractional power (−∆)s/2) split into a regular H s-component plus a singular com- ponent; however, in the former case the local singularity is x−1 for all considered 2 , 2], whereas in the latter it is the singularity of Gs,λ, namely x−(3−s) powers s ∈ ( 1 for all powers s ∈ ( 3 α α 2 , 5 2 ). 14 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE In either case, a local boundary condition constrains regular and singular com- ponents: working out the asymptotics as x → 0 in (2.12) and (4.8) by means of (3.11) we find (4.14) g(x) ∼ (α + (4πx)−1) g(x) ∼ (α + Λsx−(3−s)) s ∈ ( 3 2 , 2] ) , s ∈ ( 3 2 , 5 where Λs is defined in (3.11). (Observe that Λs=2 = (4π)−1, consistently.) g ∈ D(hs/2 α ) , g ∈ D(k(s/2) as x → 0 , as x → 0 , Furthermore, whereas (hα + λ1)−s/2 is not a finite rank perturbation of (−∆ + + λ1)−1 is indeed a rank-one perturbation of ((−∆)s/2 + λ1)−1. λ1)−s/2, (k(s/2) 2 ) , α α 5. Rank-one singular perturbations of the fractional Laplacian: homogeneous case in dimension one In terms of the general discussion of Sec. 3, we consider here now the operator k(s/2) on L2(R) when s ∈ ( 1 2 )\{1}. 2 , 3 We start with identifying the Friedrichs extension k(s/2) of k(s/2). Unlike the depends on whether s < 1 or s > 1. F three-dimensional case, the structure of k(s/2) Proposition 5.1. Let s ∈ ( 1 F 2 , 1) ∪ (1, 3 2 ). (cid:40) (i) The quadratic form of the Friedrichs extension k(s/2) if s ∈ ( 1 2 , 1) if s ∈ (1, 3 2 ) D[k(s/2) H s/2(R) H s/2 [f, g] = (cid:104)∇ s (R\{0}) 2 g(cid:105) . 2 f,∇ s ] = F F 0 k(s/2) F (5.1) of k(s/2) is (ii) When s ∈ ( 1 2 , 1), one has F D(k(s/2) k(s/2) F ) = H s(R) f = (−∆) s 2 f . (5.2) (5.3) (iii) When s ∈ (1, 3 (cid:26) 2 ), for every λ > 0 one has f = φ − φ(0) Gs,λ(0) D(k(s/2) + λ1)f = ((−∆) s 2 + λ)φ . ) = F (k(s/2) F (cid:27) (cid:12)(cid:12)(cid:12) φ ∈ H s(R) Gs,λ In particular, D(k(s/2) λ1) has an everywhere defined and bounded inverse on L2(R) with ) ⊂ H s(R)(cid:117)span{Gs,λ}. In this regime of s, (k(s/2) F + F (5.4) (k(s/2) F + λ1)−1 = ((−∆) s 2 + λ1)−1 − 1 Gs,λ(0) Gs,λ(cid:105)(cid:104)Gs,λ . Proof. Following the standard form construction of the Friedrichs extension (see, e.g., [7, Theorem A.2]), the Friedrichs form domain is the completion of D(k(s/2)) with respect to the H s D[k(s/2) F ] = H s 2 -norm, and therefore 0 (R\{0}) Hs/2(R) = (cid:107) (cid:107) (cid:40) H s/2(R) H s/2 (R\{0}) 0 if s ∈ ( 1 2 , 1) if s ∈ (1, 3 2 ) , last identity being proved precisely as (3.5). Moreover, k(s/2) for all the sequences (fn)n, (gn)n of H s whence k(s/2) [f, g] = (cid:104)∇ s 2 f,∇ s 2 g(cid:105). This completes the proof of part (i). [f, g] = lim 2 -approximants of f and g respectively, F F n→+∞(cid:104)∇ s 2 fn,∇ s 2 gn(cid:105) FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 15 The self-adjoint operator associated with the form (5.1) is qualified by the for- mulas D(k(s/2) F + λ1) = (k(s/2) F + λ1)f = uf f ∈ D[k(s/2) F ] (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:104)∇ s ∃ uf ∈ L2(R) such that 2 f,∇ s 2 g(cid:105) + λ(cid:104)f, g(cid:105) = (cid:104)uf , g(cid:105) ∀ g ∈ D[k(s/2) ] F  valid for any λ > 0. unique.) In particular, D(k(s/2) is. When s ∈ ( 1 f ∈ H s(R) and uf = (−∆) s When instead s ∈ (1, 3 (By density of H s/2 (R\{0}) in L2(R), the above uf is ) is independent of λ, only its internal decomposition 2 , 1) the condition identifying f and uf is clearly equivalent to F 0 2 f , which yields part (ii) of the thesis. (cid:90) 2(cid:98)f p s p s R ∀g such that 2 ) the above condition reads 2(cid:98)g dp + λ (cid:90) (cid:90) R (p s R (cid:90) (cid:90) (cid:98)f(cid:98)g dp = R(cid:99)uf(cid:98)g dp 2 + 1)(cid:98)g(p)2dp < +∞ and (cid:1)(cid:98)g dp = 0 (cid:0)(ps + λ)(cid:98)f −(cid:99)uf R (*) that is, (cid:90) R(cid:98)g dp = 0 , for all the g's indicated in (*) and for some uf ∈ L2(R). It is easy to see that this is the same as (cid:98)f (p) = (cid:99)uf (p) + c ps + λ , (cid:99)uf (p) = (ps + λ)(cid:98)f (p) − c for some c ∈ C. Now, the condition for f to vanish at x = 0 and belong to H s/2(R) is equivalent to (cid:90) R (cid:12)(cid:12)(cid:12)(p s 2 + 1)(cid:99)uf (p) + c ps + λ (cid:12)(cid:12)(cid:12)2 dp . dp and + ∞ > The finiteness of the second integral above is guaranteed by s > 1 and uf ∈ L2(R), whereas the vanishing of the first integral is the same as (cid:90) R 0 = ps + λ (cid:99)uf (p) + c (cid:90) (cid:99)uf (p) ps + λ R (cid:90) R (cid:90) (cid:82) dp = −c dp ps + λ = −2πc Gs,λ(0) , i.e., √ 2π c = − 1 ((−∆) s 2 + λ)−1uf dx . R Gs,λ(0) Therefore, as a distributional identity, uf = ((−∆) In turn, the condition uf ∈ L2(R) is equivalent to φ := ((−∆) s in terms of such φ the previous identity reads 2π c δ = ((−∆) 2 + λ)f − s 2 + λ) (cid:16) (cid:82) f + √ s (cid:17) R((−∆) s 2 + λ)−1uf Gs,λ . Gs,λ(0) 2 + λ)−1uf ∈ H s(R): e the condition f (0) = 0 reads(cid:82) f = φ − R φ dx Gs,λ(0) Gs,λ , R φ dx = φ(0). We have thus found f = φ − φ(0) Gs,λ(0) Gs,λ , uf = ((−∆) s 2 + λ)φ , 16 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE which proves (5.3). The following inversion formula is then a straightforward con- sequence: (k(s/2) F + λ1)−1u = φu − φu(0) Gs,λ(0) 2 + λ)−1u ∈ H s(R) . φu := ((−∆) Gs,λ s On the other hand, φu(0) = 1√ 2π (cid:90) R (cid:98)u(p) ps + λ dp = (cid:104)Gs,λ, u(cid:105) , whence (5.4) and the conclusion of the proof of part (iii). (cid:3) As a special case of (5.4) above, (5.5) hs,λ := = + λ1)−1 Gs,λ (cid:17) 2π (k(s/2) √ F  F−1(cid:16) F−1(cid:16) 1 1 (ps + λ)2 (ps + λ)2 − s − 1 2 ) . and (as follows from (3.11)) Gs,λ(0) = (λ1− 1 if s ∈ ( 1 2 , 1) if s ∈ (1, 3 ps + λ λ s 1 (cid:17) s s sin π s )−1, Indeed, (cid:104)Gs,λ, Gs,λ(cid:105) = whence (cid:107)Gs,λ(cid:107)2 s−1 s s2 sin π s λ2− 1 L2(R)/Gs,λ(0) = s−1 λ s . We can now establish the following construction. Theorem 5.2. Let s ∈ ( 1 2 ) and λ > 0. Set (5.6) ω(s) := , θs := 0 1 if s < 1 if s > 1 . 2 , 1) ∪ (1, 3 s2 sin ( π s ) s − 1 (cid:40) (cid:17) (cid:1) D(cid:0)k(s/2) + λ1(cid:1)g (5.7) τ (5.8) (cid:0)k(s/2) τ (5.9) (i) The self-adjoint extensions in L2(R) of the operator k(s/2) form the family )τ∈R∪{∞}, where k(s/2)∞ is its Friedrichs extension, already qualified (k(s/2) in Proposition 5.1, and all other extensions are given by τ (cid:41) (cid:12)(cid:12)(cid:12) F λ ∈ H s(R) − θs Gs,λ(0) F λ(0) Gs,λ (cid:40) (cid:16) ω(s)λ2− 1 s g = F λ + := := ((−∆) s 2 + λ)F λ . τ (ii) Each extension is semi-bounded from below and τ inf σ(k(s/2) inf σ(k(s/2) + λ1) (cid:62) 0 ⇔ τ (cid:62) 0 + λ1) > 0 ⇔ τ > 0 + λ1) is invertible ⇔ τ (cid:54)= 0 . τ (k(s/2) τ (cid:40) (iii) For each τ ∈ R the quadratic form of the extension k(s/2) τ is given by (5.10) k(s/2) τ (5.11) D[k(s/2) H s/2(R) (cid:117) span{Gs,λ} H s/2 [F λ + κλGs,λ] = (cid:107)∇ s (R\{0}) (cid:117) span{Gs,λ} ] = 0 τ if s ∈ ( 1 2 , 1) if s ∈ (1, 3 2 ) L2(R) 2 F λ(cid:107)2 +λ(cid:107)F λ(cid:107)2 L2(R) − λ(cid:107)F λ + κλGs,λ(cid:107)2 κλ2 L2(R) + τ ω(s)λ2− 1 s for any F λ ∈ D[k(s/2) F ] and κλ ∈ C. FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 17 (iv) For τ (cid:54)= 0, one has the resolvent identity (5.12) (k(s/2) τ + λ1)−1 = ((−∆)s/2 + λ1)−1 + (cid:16) ω(s)λ2− 1 s τ (cid:17)Gs,λ(cid:105)(cid:104)Gs,λ . − θs Gs,λ(0) Proof. We proceed along the line of the proof of Theorem 4.1, based upon the Kreın-Visik-Birman self-adjoint extension scheme and Proposition 5.1. (I) The adjoint of k(s/2) is qualified by D(cid:0)(k(s/2))∗(cid:1) = g ∈ L2(R) (cid:0)(k(s/2))∗ + λ1(cid:1)g = (k(s/2) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) F + λ1)(f λ + η hs,λ) , (cid:98)g(p) =(cid:99)f λ(p) + η(cid:100)hs,λ + η, ξ ∈ C , f λ ∈ H s(R) , (cid:82) ξ ps + λ R(cid:99)f λ(p) dp = 0  where hs,λ is the function (5.5), and the self-adjoint restrictions of (k(s/2))∗ are qualified by the self-adjointness condition η = τ ξ for some τ ∈ R∪{∞}. For clarity of presentation, let us split the discussion into the two regimes s < 1 and s > 1. First case: s ∈ ( 1 Let F λ := f λ + τ ξ hs,λ. When f λ and ξ run over their possible domains, F λ 2 , 1). spans the whole Friedrichs domain H s(R). Moreover, (cid:90) R (cid:99)F λ(p) dp = ξ 2πτ λ2− 1 s ω(s) . (II) Thus, the first line in (I) and (II) yield (5.7). Owing to (5.2) and to the fact that is a restriction of (k(s/2))∗, one deduces (5.8) from (I). Thus, part (i) is proved. k(s/2) τ ] = + λ1)[F λ + ] (cid:117) ker(cid:0)(k(s/2))∗ + λ1(cid:1) and (5.1) imply (5.10), whereas (k(s/2) Parts (ii) and (iii) are proved as in Theorem 4.1: in particular, D[k(s/2) D[k(s/2) κλGs,λ] = ((−∆)s/2 + λ1)[F λ] + τκλ2(cid:107)Gs,λ(cid:107)2 L2(R) and (5.1) imply (5.11). F τ τ Kreın's resolvent formula for deficiency index 1 and (5.2) prescribe + λ1)−1 = ((−∆)s/2 + λ1)−1 + βλ,τ Gs,λ(cid:105)(cid:104)Gs,λ (k(s/2) τ for some scalar βλ,τ to be determined, whenever (k(s/2) τ (cid:54)= 0. Thus, for a generic h ∈ L2(R), the element g := (k(s/2) τ reads, in view of (4.2) and of the resolvent formula above, τ +λ1) is invertible, hence for + λ1)−1h ∈ D(k(s/2) ) τ (cid:98)g(p) = (cid:99)F λ(p) + ξλ ps + λ , (cid:99)F λ(p) := (cid:98)h(p) ps + λ , ξλ := βλ,τ 2π qs + λ R dq . (cid:90) (cid:98)h(q) The boundary condition (I) for F λ and ξλ then implies 1 = βλ,τ ω(s)λ2− 1 determines βλ,τ and proves (5.12), thus completing also the proof of (iv). s τ , which Second case: s ∈ (1, 3 2 ). Let F λ := f λ +τ ξ hs,λ. (This is for consistency with the first case, but such F λ is not going to be the same as the F λ of the thesis: functions will be renamed later.) When f λ and ξ run over their possible domains, F λ spans the whole Friedrichs domain (5.3). In particular F λ(0) = 0, which shows that the boundary condition in g between F λ and g − F λ cannot have the form (II) any longer. Owing to (5.3), we can re-write (III) D(k(s/2) F ) (cid:51) F λ = f λ + τ ξ hs,λ = φλ τ,ξ − φλ τ,ξ(0) Gs,λ(0) Gs,λ , as follows from (3.11) and (5.6), then (V) becomes λ s s − 1 Gs,λ(0) (cid:16) ω(s) λ2− 1 s τ 1 = ω(s) λ2− 1 s , (cid:17) φλ τ,ξ(0) Gs,λ . − 1 Gs,λ(0) g = φλ τ,ξ + (VI) Moreover, (VII) 18 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE τ,ξ running over the whole H s(R) when F λ runs over the whole D(k(s/2) φλ (5.5) this is the same as F ). Using s − 1 λ s 1 ps + λ = (cid:100)φλ τ,ξ − φλ √ τ,ξ(0) 2π Gs,λ(0) 1 ps + λ , (cid:99)f λ + whence the identification τ ξ (ps + λ)2 − τ ξ (cid:100)φλ τ,ξ = (cid:99)f λ + τ ξ (ps + λ)2 , √ φλ τ,ξ(0) 2π Gs,λ(0) = τ ξ s − 1 λ s . (In fact, a straightforward computations confirms that assuming the first of (IV), the second follows.) From (I) (with η = τ ξ) we then see that a generic g ∈ D(k(s/2) ) has the form (cid:98)g = (cid:99)F λ + = (cid:100)φλ τ,ξ + √ 2π ξ(cid:98)Gs,λ = (cid:100)φλ (cid:17) φλ (cid:16) 1 λ s s − 1 τ − 1 τ,ξ + τ,ξ(0) Gs,λ(0) (cid:17)(cid:98)Gs,λ (cid:16)√ 2π ξ − φλ (cid:98)Gs,λ , τ τ,ξ(0) Gs,λ(0) where we used (III) in the second step and the second identity of (IV) in the third step. Since (IV) (V) (k(s/2) τ + λ1)g = (k(s/2) F + λ1)F λ = ((−∆) s 2 + λ)φλ τ,ξ , τ the whole D(k(s/2) Part (i) is proved. Renaming φλ having used the second line of (I) in the first identity and (5.3) in the second identity. τ,ξ into F λ, now F λ runs over the whole H s(R) when g runs over ): this fact and (VI) then yield (5.7), whereas (VII) yields (5.8). The proof of parts (ii) and (iii) follows the same scheme as in the case s ∈ ( 1 ] (cid:117) ker(cid:0)(k(s/2))∗ + λ1(cid:1) and (5.1) imply (5.10), whereas ] = D[k(s/2) +λ1)[F λ +κλGs,λ] = ((−∆)s/2 +λ1)[F λ]+τκλ2(cid:107)Gs,λ(cid:107)2 L2(R) and (5.1) imply Concerning part (iv), Kreın's resolvent formula and (5.4) prescribe for all τ (cid:54)= 0 thus, D[k(s/2) (k(s/2) (5.11). 2 , 1): F τ τ (VIII) (k(s/2) τ + λ1)−1 = (k(s/2) = ((−∆) F + λ1)−1 + βλ,τ Gs,λ(cid:105)(cid:104)Gs,λ 2 + λ1)−1 + 1 βλ,τ − s for some scalar βλ,τ to be determined. Owing to (VIII) and to (5.7), in order for g := (k(s/2) τ + λ1)−1h = ((−∆) s 2 + λ1)−1h + βλ,τ − 1 τ to belong to D(k(s/2) ((−∆) s λ1)F λ(cid:105) = F λ(0), one must necessarily have βλ,τ = ω(s)λ2− 1 (5.12). ) for a generic h ∈ L2(R), keeping into account that F λ := 2 +λ1)−1h is a generic function in H s(R) and that (cid:104)Gs,λ, h(cid:105) = (cid:104)Gs,λ, ((−∆) s 2 + s /τ . Then (VIII) yields (cid:3) Analogously to the change of parametrisation from Theorem 4.1 to Theorem 4.2, we deduce from Theorem 5.2 the following version. (cid:16) (cid:16) (cid:17)Gs,λ(cid:105)(cid:104)Gs,λ (cid:17)(cid:104)Gs,λ, h(cid:105) Gs,λ Gs,λ(0) Gs,λ(0) FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 19 Theorem 5.3. Let s ∈ ( 1 2 , 1) ∪ (1, 3 Θ(s, λ) := (cid:0)λ1− 1 , λ > 0 . (i) The self-adjoint extensions in L2(R) of the operator k(s/2) form the family (5.13) (k(s/2) α (5.14) 2 ) and s s sin ( π s )(cid:1)−1  g = F λ + )α∈R∪{∞}, where for arbitrary λ > 0 D(k(s/2) ) = α F λ ∈ H s(R) + λ) g = ((−∆)s/2 + λ) F λ . (k(s/2) α F λ(0) α − Θ(s, λ) Gs,λ  The Friedrichs extension k(s/2) , already qualified in Proposition 5.1, cor- responds to α = ∞ when s ∈ ( 1 2 ). For generic s, the extension with α = ∞ is the ordinary self-adjoint fractional Laplacian (−∆)s/2 on L2(R). 2 , 1) and to α = 0 when s ∈ (1, 3 F (ii) For each α ∈ R the quadratic form of the extension k(s/2) if s ∈ ( 1 2 , 1) if s ∈ (1, 3 2 ) and α (cid:54)= 0 L2(R) H s/2(R) (cid:117) span{Gs,λ} H s/2 [F λ + κλGs,λ] = (cid:107)∇ s L2(R) − λ(cid:107)F λ + κλGs,λ(cid:107)2 (R\{0}) (cid:117) span{Gs,λ} D[k(s/2) is given by ] = α α 0 (cid:40) 2 F λ(cid:107)2 + λ(cid:107)F λ(cid:107)2 (cid:16) L2(R) + θs Θ(s, λ) + 1 α − Θ(s, λ) (cid:17)−1κλ2 (5.15) k(s/2) α (5.16) for arbitrary λ > 0. (iii) The resolvent of k(s/2) α is given by (5.17) (k(s/2) α + λ1)−1 = ((−∆)s/2 + λ1)−1 +(cid:0)α − Θ(s, λ)(cid:1)−1 Gs,λ(cid:105)(cid:104)Gs,λ for arbitrary λ > 0. (iv) For each α ∈ R the extension k(s/2) α is semi-bounded from below, and (5.18) σess(k(s/2) α ) = σac(k(s/2) σsc(k(s/2) α ) = ∅ , (5.19) σdisc(k(s/2) α ) = where the eigenvalue −E(s) α is non-degenerate and is given by α ) = [0, +∞) , (cid:40) ∅ α = (cid:0)αs sin( π s )(cid:1) s α } {−E(s) 1−s E(s) if (s − 1) α (cid:54) 0 if (s − 1) α > 0 (5.20) the (non-normalised) eigenfunction being Gs,λ=E(s) α . Proof. For any two pairs (λ, τ ) and (λ(cid:48), τ(cid:48)) identifying the same self-adjoint reali- sation k(s/2) ) decomposes as τ τ , a function g ∈ D(k(s/2) g = F λ + A(λ, τ )F λ(0)Gs,λ = F λ(cid:48) − θs ω(s)λ2− 1 A(λ, τ ) := s τ , Gs,λ(0) + A(λ(cid:48), τ(cid:48))F λ(cid:48) (0)Gs,λ(cid:48) and the uniqueness of the decomposition implies (I) F λ(cid:48) = F λ + A(λ, τ )F λ(0)Gs,λ − A(λ(cid:48), τ(cid:48))F λ(cid:48) (0)Gs,λ(cid:48) . 20 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE In order for F λ(cid:48) Gs,λ(cid:48) must cancel out, that is, to belong to H s(R), the non-H s singularities at x = 0 of Gs,λ and (II) A(λ, τ )F λ(0) = A(λ(cid:48), τ(cid:48))F λ(cid:48) (0) . Plugging (II) into (I) and evaluating of (I) at x = 0 yields (III) A(λ, τ )F λ(0) (0) = F λ(0) + A(λ, τ )F λ(0) = F λ(cid:48) F λ(cid:48) (0) (cid:90) 1 2π (cid:16) (cid:17) dp . 1 ps + λ − 1 ps + λ(cid:48) A straightforward analysis of the integral above (exploiting the compensation of singularities when s ∈ ( 1 2 , 1)) shows that 1 (cid:16) (cid:17) 1−s 1 1 s (cid:1) s − λ(cid:48) 1−s − ps + λ R ps + λ(cid:48) dp = (cid:90) 1 2π (IV) R (cid:0)λ s sin π s = Θ(s, λ) − Θ(s, λ(cid:48)) . Combining (III) and (IV) together implies that (λ, τ ) and (λ(cid:48), τ(cid:48)) are linked by the relation i.e., (V) α − Θ(s, λ) = 1 A(λ, τ ) + Θ(s, λ) = + Θ(s, λ(cid:48)) =: α , 1 A(λ(cid:48), τ(cid:48)) (cid:16) ω(s)λ2− 1 s − θs Θ(s, λ) τ (cid:17)−1 , which gives the natural extension parameter α. It is immediate from (V) that the Friedrichs case τ → +∞ corresponds to α → +∞ when s ∈ ( 1 2 , 1) and to α = 0 when s ∈ (1, 3 2 ). Upon replacing (V) in the formulas of Theorem 4.1 we deduce parts (i), (ii), and (iii). Moreover, arguing as in the analogous point of the proof of Theorem 4.2, formulas (5.18) follow, and one also concludes that each k(s/2) may have at most one negative non-degenerate eigenvalue Eα = −λ, λ > 0. α The occurrence of Eα is read out from the resolvent formula (5.17) as the pole + λ1)−1, that is, imposing α − Θ(s, λ) = 0 and hence of (k(s/2) α (VI) When s < 1, (VI) can be only satisfied by some λ > 0 when α < 0, in which case s )(cid:1)−1 . s−1 s s sin( π α =(cid:0)λ λ =(cid:0)αs sin( π s )(cid:1) s λ =(cid:0)αs sin( π s )(cid:1) s 1−s 1−s (s < 1 , α < 0) . (s > 1 , α > 0) . When instead s > 1, a solution λ > 0 to (VI) exists only when α > 0, and is given by (cid:3) Hence we proved also (5.19), which completes the proof of part (iv). In the regime s ∈ (1, 3 α 2 ) Theorem 5.3(ii) can be re-phrased in the following even can be equivalently qualified as more natural formulation, which shows that k(s/2) a form perturbation of (−∆)s/2. Proposition 5.4. Let s ∈ (1, 3 2 ). The self-adjoint extensions in L2(R) of k(s/2) form the family (k(s/2) )α∈R∪{∞}, where α = 0 labels the Friedrichs extension given by (5.1), α = ∞ labels the ordinary self-adjoint fractional Laplacian (−∆)s/2, and for α ∈ R \ {0} one has D[k(s/2) α ] = H s/2 [g] = (cid:107)∇ s 0 α k(s/2) α (R\{0}) (cid:117) span{Gs,λ} = H s/2(R) 2 g (cid:107)2 g(0)2 L2(R) − 1 α (5.21) FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 21 for every λ > 0. Proof. In view of Theorem 5.3(ii), we only need to prove the second line of (5.21) for Θ(s,λ) − a generic g ∈ D[k(s/2) α−Θ(s,λ) )−1 for short and decompose ]. We set Σ := ( g = F λ + κλGs,λ for some F λ ∈ H s/2 (R\{0}) and κλ = g(0)/Gs,λ(0). Applying (5.16), we find α 1 1 0 k(s/2) α L2(R) + (cid:107)∇ s [g] = −λ(cid:107)g(cid:107)2 = (cid:107)∇ s 2 g(cid:107)2 − 2 Re κλ L2(R) + κλ2(cid:0)(cid:107)∇ s (cid:0)(cid:104)∇ s 2 (g − κλGs,λ)(cid:107)2 2 Gs,λ(cid:107)2 2 Gs,λ(cid:105) + λ(cid:104)g, Gs,λ(cid:105)(cid:1) . L2(R) + λ(cid:107)g − κλGs,λ(cid:107)2 L2(R) + λ(cid:107)Gs,λ(cid:107)2 L2(R) + Σ(cid:1) L2(R) + Σκλ2 2 g,∇ s L2(R) + λ(cid:107)Gs,λ(cid:107)2 2 g,∇ s 2 Gs,λ(cid:107)2 Since (cid:107)∇ s Θ(s, λ) and analogously (cid:104)∇ s 2 g(cid:107)2 [g] = (cid:107)∇ s k(s/2) α L2(R) + 2 + λ)Gs,λ(cid:105) = Gs,λ(0) = L2(R) = (cid:104)Gs,λ, ((−∆) s 2 Gs,λ(cid:105) + λ(cid:104)g, Gs,λ(cid:105) = g(0), then g(0)2 Θ(s, λ)2 (Θ(s, λ) + Σ) − 2 g(0)2 Θ(s, λ) (cid:16) The coefficient of the g(0)2-term above amounts to 1 Θ(s, λ)2 whence indeed k(s/2) α Θ(s, λ) + [g] = (cid:107)∇ s 1 1 α−Θ(s,λ) Θ(s,λ) − 2 g (cid:107)2 L2(R) − α−1g(0)2. 1 (cid:17) − 2 Θ(s, λ) = − 1 α , . (cid:3) 6. Rank-one singular perturbations of the fractional Laplacian: inhomogeneous case For completeness of presentation, in this Section we work out the inhomogeneous version of the operator k(s/2) of Section 4, for concreteness in dimension d = 3. That is, instead of constructing a singular perturbation of (−∆)s/2, we consider the singular perturbation of the fractional power (−∆ + 1)s/2. This is going to be the operator d(s/2) introduced informally in (1.3). τ τ The conceptual scheme is the very same as in Sections 3 and 4, and only certain explicit computations are modified in an easy way. Thus, we content ourselves to state the main results without proofs. For chosen λ > 0 and s ∈ R we set x, p ∈ R3 , Gs,λ(x) := 1 1 2 (x) , (2π) 3 (p2 + λ)s/2 (6.1) whence (−∆ + λ1)s/2 Gs,λ = δ(x) distributionally. We also set 0 (R3\{0}) (6.2) as an operator closure with respect to the Hilbert space L2(R3). Thus, in compar- ison to Section 2, G2,λ = Gλ and d(1) x→0−−−−→ Λs = := (−∆ + λ1)s/2 (cid:22) C∞ λ = h + λ1. Moreover, x3−s Gs,λ(x) s ∈ (0, 3) d(s/2) λ , (cid:17)∨ (cid:16) (6.3) Gs,λ(x) x→0−−−−→ Gs,λ(0) = , s > 3 . Γ( 3−s 2 ) 2 2s− 3 (2π) 3 2 Γ( s 2 ) Γ( s−3 2 ) s−3 2 Γ( s 2 ) 2 λ 8π 3 Reasoning as in (3.5) and in Lemma 3.1, we see that when s ∈ ( 3 index of d(s/2) equals 1. λ 2 , 5 2 ) the deficiency One has the following construction. Theorem 6.1. Let s ∈ ( 3 2 , 5 2 ) and λ > 0. 22 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE (i) The self-adjoint extensions in L2(R3) of the operator d(s/2) form the family (d(s/2) λ,∞ is its Friedrichs extension, namely the self- adjoint fractional shifted Laplacian (−∆ + λ1)s/2, and all other extensions are given by )τ∈R∪{∞}, where d(s/2) λ,τ λ λ,τ (cid:1) := D(cid:0)d(s/2) g ∈ L2(R3) (cid:110) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:98)g(p) =(cid:99)f λ(p) + ξ ∈ C , f λ ∈ H s(R3) , (cid:82) where (cid:99)F λ =(cid:99)f λ(p) + (p2 + λ)−sτ ξ, and 2 ) F λ(0)Gs,λ g = F λ + 8π 2 λs− 3 τ Γ(s− 3 2 Γ(s) τ ξ = 3 (p2 + λ)s + ξ (p2 + λ)s/2 R3(cid:99)f λ(p) dp = 0 (cid:12)(cid:12)(cid:12) F λ ∈ H s(R3) (cid:111) ,  (6.4) (6.5) λ,τ g := (−∆ + λ1)s/2F λ . d(s/2) (ii) Each extension is semi-bounded from below and (6.6) λ,τ inf σ(d(s/2) inf σ(d(s/2) d(s/2) λ,τ ) (cid:62) 0 ⇔ τ (cid:62) 0 ) > 0 ⇔ τ > 0 is invertible ⇔ τ (cid:54)= 0 . λ,τ (iii) The eigenvalue zero of the extension d(s/2) normalised) eigenfunction is Gs,λ. When τ < 0 the operator d(s/2) one non-degenerate negative eigenvalue Eτ < τ . λ,τ =0 is non-degenerate and the (non- admits (iv) For each τ ∈ R the quadratic form of the extension d(s/2) is given by λ,τ λ,τ (6.7) (6.8) D[d(s/2) 2 (R3) (cid:117) span{Gs,λ} [F λ + κλGs,λ] = (cid:107)(−∆ + λ1)s/4F λ(cid:107)2 ] = H λ,τ s d(s/2) λ,τ for any F λ ∈ H s 2 (R3) and κλ ∈ C. (v) For τ (cid:54)= 0, one has the resolvent identity )−1 = (−∆ + λ)−s/2 + 8π (d(s/2) 3 λ,τ (6.9) 2 λs− 3 2 Γ(s) τ Γ(s− 3 2 ) Gs,λ(cid:105)(cid:104)Gs,λ . L2(R3) + τ Γ(s− 3 2 λs− 3 8π 2 ) 2 Γ(s) 3 κλ2 or k(s/2) It is clear that, as opposite to hs/2 τ : in fact, the domain of d(s/2) , the shift λ > 0 is inherent the very is independent of construction of the operator d(s/2) λ > 0, but its action is not (the difference d(s/2) trivial operator), thus also the adjoint (d(s/2) λ-dependent (the adjoints (hs/2)∗ and (k(s/2))∗ are λ-independent, instead). λ is a bounded, yet non- )∗ and its self-adjoint restrictions are − d(s/2) λ(cid:48) λ,τ λ λ τ Let us also elaborate further on part (iii) of the Theorem. As in the previous Sections, both statements are classical facts in the Kreın-Visik-Birman extension scheme. Chosen λ > 0 and s ∈ ( 3 2 ), the negative eigenvalue Eτ of the extension d(s/2) for τ < 0 is obtained as follows. Let g be the corresponding eigenfunction λ,τ and decompose(cid:98)g =(cid:99)f λ + (p2 + λ)−sτ + (p2 + λ)−s/2 according to (6.4) (without loss 2 , 5 of generality we re-absorbed ξ in f λ). Then the condition d(s/2) the property (6.5), reads 2 τ = Eτ(cid:99)f λ + (p2 + λ)−sτ Eτ + (p2 + λ)− s 2 Eτ , λ,τ g = Eτ g, owing to (p2 + λ) whence s 2(cid:99)f λ + (p2 + λ)− s (cid:98)f (p) = (cid:17) . (cid:16) 1 (p2 + λ) s 2 − Eτ τ Eτ (p2 + λ)s − τ − Eτ (p2 + λ) s 2 FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 23 Figure 1. Plot of the negative eigenvalue Eτ of the extension d(s/2) vs τ for λ = 1 and s = 1.8 (blue curve). The reference λ,τ orange curve gives the corresponding value of τ . Indeed, Eτ < τ and Eτ→0 = 0. The fact that f λ ∈ H s(R3) is then obvious, whereas the condition (cid:82) R3(cid:99)f λdp = 0 selects the value of Eτ in terms of τ (and of λ) -- a numerical example is provided in Figure 1. Let us conclude by remarking that two relevant features of the homogeneous case are present in the inhomogeneous case too. First, and most importantly, the operator d(s/2) the resolvent sense, of (−∆ + λ)s/2, precisely as k(s/2) of (−∆)s/2. τ λ,τ is a rank-one perturbation, in is a rank-one perturbation Second, the elements of the domain of d(s/2) decompose into a regular H s-part and a singular part, constrained to the former by a local boundary condition, where the local singularity when s ∈ ( 3 2 ) is of the form x3−s as x → 0. 2 , 5 λ,τ 7. High deficiency index (high fractional power) scenario Let us outline in this Section how the previous constructions of the self-adjoint extensions of the operators k(s/2) or d(s/2) We recall from Lemma 3.1 that when s ∈ I (d) λ get modified when s > d 2 + 1. (cid:110) ker(cid:0)(k(s/2))∗ + λ1(cid:1) = span (cid:110) )∗ + λ1(cid:1) = span ker(cid:0)(d(s/2) λ uλ γ1,...,γd vλ γ1,...,γd n ( d 2 + n − 1, d (cid:12)(cid:12)(cid:12) γ1, . . . , γd ∈ N0 , (cid:12)(cid:12)(cid:12) γ1, . . . , γd ∈ N0 , 2 + n), n ∈ N, one has (cid:111) d(cid:88) γj (cid:54) n − 1 , j=1 d(cid:88) j=1 (cid:111) γj (cid:54) n − 1 , (7.1) (7.2) and analogously having defined (7.3) (cid:99)vλ γ1,...,γd (p) := 1 ··· pγd pγ1 d (p2 + λ) s 2 . operator T in some subspace D(T ) of ker(cid:0)(k(s/2))∗ + λ1(cid:1) ∼= CJ (s,d), hence labelled The same extension scheme applied in Section 4 provides an analogous classifica- tion of all the self-adjoint extensions in the case of generic deficiency index J (s, d), where now each extension of k(s/2) is an operator k(s/2) labelled by a self-adjoint by some N × N hermitian matrix, 1 (cid:54) N (cid:54) J (s, d). T -10-9-8-7-6-5-4-3-2τ-5-10-15-20Eτ 24 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE Explicitly (see, e.g., [7, Theorem 3.4]), g ∈ L2(R3) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) g = f + (k(s/2) w ∈ ker(cid:0)(k(s/2))∗ + λ1(cid:1) ∩ D(T )⊥ + λ1)−1(T u + w) + u 0 (R3\{0}) , u ∈ D(T ) , where f ∈ H s F  D(k(s/2) T ) = (7.4) (k(s/2) T + λ1)g = (k(s/2) + λ1)Fλ F Fλ := f + (k(s/2) F + λ1)−1(T u + w) ∈ H s(R3) , with k(s/2) F denoting the Friedrichs extension. Analogously, the self-adjoint extensions of d(s/2) form a family of operators d(s/2) λ,T λ with (7.5) D(d(s/2) λ,T ) = g ∈ L2(R3) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) where f ∈ H s g = f + (d(s/2) w ∈ ker(cid:0)d(s/2) λ,F )−1(T v + w) + v 0 (R3\{0}) , v ∈ D(T ) , )∗ ∩ D(T )⊥ λ  d(s/2) λ,T g = d(s/2) λ,F Fλ Fλ := f + (d(s/2) λ,F )−1(T v + w) ∈ H s(R3) . The theory provides also a counterpart classification of the the quadratic forms of the extensions (see, e.g., [7, Theorem 3.6]). The above formulas show that for high powers s the operators k(s/2) and d(s/2) have a richer variety (a J (s, d)2-parameter family) of self-adjoint extensions. The parametrising matrix T determines a more complicated set of 'boundary condition' between the 'Friedrichs' part of a generic element of the extension domain, and the remaining part: the resulting constraint involves the evaluation at x = 0 of some number of partial derivatives of the former component. λ This construction produces finite-rank perturbations in the resolvent sense, hence extensions that are all semi-bounded from below and may admit a (finite) number of negative eigenvalues, up to J (s, d), counting the multiplicity. Unlike the case of deficiency index 1, depending on the extension parameter T the large-p vanishing behaviour in momentum space of the singular component may be milder than that of the Green function, and therefore the local singularity of g in position space may be more severe than the behaviour of the Green function as x → 0. function g ∈ D(k(s/2) identical. Let us comment on how the worst leading singularity at x = 0 of a generic λ,T ) is ) depends on s and d -- the discussion for g ∈ D(d(s/2) T As expressed by (7.4), such a singularity is due to those functions of type uλ γ1,...,γd that span D(T ). When s ∈ I (d) the worst local singularity occurs when such functions decrease at infinity in momentum coordinates with the slowest possible vanishing rate compatible with s and d, that is, when γ1 + ··· + γd = n − 1. p−(s+1−n) as p → +∞. Then u(x) ≈ x−(d−1+n−s) as x → 0. Since the map Let u be any such most singular function, which then behaves as (cid:98)u(p) ≈ n n (cid:51) s (cid:55)→ d − 1 + n − s I (d) 2 − 1, d is monotone decreasing and takes values in ( d such that D(T ) (cid:51) u, then the functions in D(k(s/2) ranges from x− d 2 to x− d−1 s increases in I (3) 1 . T 2 as long as s increases in I (d) 2 ), if the extension k(s/2) is ) display a local singularity that n , precisely as (4.14) when T FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 25 Noticeably, at the transition values s ∈ N + 1 2 the above picture undergoes a discontinuity in s, due to the further control of one more derivative in D(k(s/2)), as a consequence of Sobolev's Lemma, and consequently to emergence in ker(cid:0)(k(s/2))∗ + λ1(cid:1) of elements that in momentum coordinates vanish more slowly at infinity. 8. Applications and perspectives Besides the operator-theoretic and functional-analytic interest per se of the con- structions of the operators (1.3), our discussion is profoundly inspired to an amount of natural applications. Singular perturbations model point-like impurities and more generally point- like interactions. In the realm of the evolutive equations of relevance for quantum mechanics, they naturally govern the evolution of systems subject to such 'singular potentials'. This includes the linear Schrodinger evolution i∂tu = hαu as well as the class of semi-linear Schrodinger equations i∂tu = hαu + N (u) of the free Laplacian plus a point-like perturbation, with physically relevant non- linearities such as the power-law local non-linearity N (u) = uγ−1u or the Hartree type non-local non-linearity N (u) = (w ∗ u2)u. The reconstruction of the linear propagator from the resolvent of hα is already known in the literature [19, 1], as well as the dispersive properties and space-time estimates of such a propagator [5, 9], and the existence, completeness, and Lp-boundedness of the wave operators for the pairs (hα,−∆) [6]. In addition, for the study of the non-linear problem in a suitable space (the energy space in the first place, as well as other spaces of lower or higher regularity), the knowledge is needed of the corresponding singular norms, namely the norms α(R3) = (cid:107)(hα + 1)s/2u(cid:107)L2(R3) considered in Theorem 2.4 above and [8]. In this (cid:107)u(cid:107)(cid:101)H s respect, and with such tools, the study of certain non-linear Schrodinger equations with singular potentials has already started [12]. An analogous systematic knowledge for k(s/2) This is even more needed due to the relevance of various models of singular per- turbations of fractional differential operators. A relevant example are the powers of the quantum-mechanical semi-relativistic energy operator perturbation of which yields precisely operators of the type d(s/2) Section 6 or, in the case of zero rest energy, of the type k(s/2) √−∆ + m2, the singular m2,τ considered in as in Section 4. is by know lacking. τ τ and d(s/2) λ,τ What one finds in the literature is an increasing amount of recent studies [14, 16, 4, 11, 18, 20, 10, 15] where the singular perturbation of the fractional Laplacian is approached through Green's function methods (together with Wick-like rotations to obtain the propagator from the resolvent) that have the virtue of highlighting the singular structure carried over by what we denoted with Gs,λ and Gs,λ, however with no specific concern to the multiplicity of self-adjoint realisations and the associated local boundary conditions, or to the increase of the deficiency index with the power s. As above, for each extension one would like to qualify the linear propagator, its space-time estimates, the fractional norms, and to use these tools for the associated non-linear problems. We trust that the research programme emerging from the above considerations may be successfully addressed over the next future! 26 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE Appendix A. Characterisation of H s 0 (Rd\{0}) We show in this Appendix how to prove the characterisation (3.5) of the space 0 (Rd \ {0}). H s It is not restrictive to fix d = 3 and to discuss and compare the first two regimes s ∈ I (3) n , n = 2, 3, . . . , is completely analogous. 1 . The argument for s ∈ I (3) and s ∈ I (3) 0 Thus, let us prove the following property. Lemma A.1. 0 (R3\{0}) = H s (cid:40) H s(R3) (cid:8)f ∈ H s(R3)(cid:12)(cid:12)(cid:82) R3 (cid:98)f (p) dp = 0(cid:9) if s ∈ ( 3 if s ∈ [0, 3 2 ) 2 , 5 2 ) . Proof. We consider first the case s ∈ [0, 3 2 ). The inclusion 0 (R3\{0}) ⊂ H s(R3) H s is obvious. For the other inclusion, for any f ∈ H s(R3) and for arbitrary ε > 0 we want to find fε ∈ C∞ 0 (R3\{0}) such that (cid:107)fε − f(cid:107)H s (cid:54) ε, and by means of a standard density argument, it is not restrictive to assume further that f ∈ S(R3) and (cid:98)f is compactly supported. Let φ ∈ C∞(R3) be such that φ(x) = 0 for x (cid:54) 1 , φ(x) = 1 for x (cid:62) 2 , and let ψ := φ − 1, φn(x) := φ(nx), and ψn(x) := ψ(nx), for n ∈ N. Thus, 0 (R3) and φnf ∈ C∞(R3), with φnf vanishing in a neighbourhood of x = 0. ψ ∈ C∞ 2(cid:107)(cid:104)p(cid:105)s((cid:98)ψn ∗ (cid:98)f )(cid:107)L2 , Moreover, where Ds := (1 − ∆)s/2 and (cid:104)p(cid:105) =(cid:112)1 + p2. Clearly, (cid:98)ψn(p) = 1 (cid:107)φnf − f(cid:107)H s = (cid:107)Ds(φnf − f )(cid:107)L2 = (cid:107)Ds(ψnf )(cid:107)L2 = (2π) n3(cid:98)ψ( p (cid:12)(cid:12)(cid:12)2 dq (cid:98)f (q)(cid:98)ψn(p − q) (cid:1)(cid:12)(cid:12)2 dq (cid:104)p(cid:105)2s(cid:12)(cid:12)(cid:98)ψ(cid:0) p−q (cid:90) (cid:54) (2π)3(cid:107)(cid:98)f(cid:107)2 dp(cid:104)p(cid:105)2s(cid:12)(cid:12)(cid:12)(cid:90) supp (cid:98)f (cid:90) (cid:90) (cid:90) supp (cid:98)f dp(cid:104)p(cid:105)2s(cid:98)ψ(p − q)2 (cid:107)φnf − f(cid:107)2 n ). Therefore, H s = (2π)3 (cid:46) n2s (cid:90) dp R3 L2 n 3 The last step above follows from the continuity of q (cid:55)→(cid:82) (cid:46) n2s−3 R3 dp(cid:104)p(cid:105)2s(cid:98)ψ(p − q)2. In particular, we can choose n := n(ε) sufficiently large such that R3 n6 supp (cid:98)f R3 n→+∞ −−−−−−→ 0 . dq n (cid:107)φnf − f(cid:107)H s (cid:54) ε 2 . For such n, we can find a smooth function x (cid:55)→ χ(x) that produces a slow cut-off at infinity so that (cid:107)χφnf − φnf(cid:107)H s (cid:54) ε . 2 0 (R\{0}) satisfying We have thus identified a function fε := χφnf ∈ C∞ (cid:107)fε − f(cid:107)H s (cid:54) (cid:107)χφnf − φnf(cid:107)H s + (cid:107)φnf − f(cid:107)H s (cid:54) ε . Let us discuss now the case s ∈ ( 3 2 , 5 continuous embedding H s(R3) (cid:44)→ C 0,s− 3 2 ). Owing to Sobolev's Lemma, one has the 2 (R3) and hence any limit in the H s-norm FRACTIONAL POWERS, SINGULAR PERTURBATIONS, QUANTUM HAMILTONIANS 27 of elements in C∞ prove the inclusion 0 (R3\{0}) must vanish at the origin. Therefore, we only need to , (cid:12)(cid:12)(cid:12)(cid:90) (cid:98)f (p) dp = 0 (cid:111) (cid:17) 1{p(cid:54)1} (cid:12)(cid:12){p (cid:54) 1}(cid:12)(cid:12) , H s 0 (R3\{0}) ⊃ (cid:110) (cid:99)fR := (cid:98)f · 1{p(cid:54)R} −(cid:16)(cid:90) f ∈ H s(R3) p(cid:54)R R > 0 , H s = (2π)3 part of the proof, (cid:107)φnf − f(cid:107)2 R → +∞, it is not restrictive to assume from the beginning that f ∈ S(R3) with that is, for any f ∈ H s(R3) with f (0) = 0 and for arbitrary ε > 0 we want to find fε ∈ C∞ 0 (R3\{0}) such that (cid:107)fε − f(cid:107)H s (cid:54) ε. Since the function fR defined by (cid:98)f (p)dp has the obvious properties fR ∈ S(R3),(cid:82) R3(cid:99)fR(p) dp = 0, and (cid:107)f − fR(cid:107)H s → 0 as f (0) = 0 and with compactly supported (cid:98)f . With the same notation as in the first dp(cid:104)p(cid:105)2s(cid:12)(cid:12)(cid:12)(cid:90) (cid:12)(cid:12)(cid:12)2 dq (cid:98)f (q)(cid:98)ψn(p − q) dp(cid:104)p(cid:105)2s(cid:12)(cid:12)(cid:12)(cid:90) dq (cid:98)f (q)(cid:0)(cid:98)ψn(p − q) − (cid:98)ψn(p)(cid:1)(cid:12)(cid:12)(cid:12)2 supp (cid:98)f supp (cid:98)f (cid:90) (cid:90) (cid:1)(cid:12)(cid:12)2 dq (cid:104)p(cid:105)2s(cid:12)(cid:12)(cid:98)ψ(cid:0) p−q (cid:1) − (cid:98)ψ(cid:0) p (cid:90) dq (cid:104)p(cid:105)2s(cid:12)(cid:12)(cid:12)(cid:90) 1 supp (cid:98)f (cid:12)(cid:12)(cid:12)2 dt (∇(cid:98)ψ)(p − tq) · q (cid:90) 1 dt(∇(cid:98)ψ)(p − tq)2 (cid:90) (cid:90) (cid:54) (2π)3(cid:107)(cid:98)f(cid:107)2 (cid:90) where we used the condition (cid:82) n−1 for q ∈ supp (cid:98)f q (cid:55)→(cid:82) R3 (cid:98)f (q) dq = 0 in the second step, the bound q (cid:46) 0 dt(cid:104)p(cid:105)2s(∇(cid:98))ψ(p − tq)2 in the last step. In particular, we can choose in the penultimate step, and the continuity of the function supp (cid:98)f R3 n→+∞ −−−−−−→ 0 , R3 (cid:46) n2s−2 R3 dp(cid:82) 1 (cid:46) n2s−5 dp(cid:104)p(cid:105)2s (cid:46) n2s supp (cid:98)f n = (2π)3 R3 R3 dp R3 n n n6 dp 0 0 dq n L2(R3) (cid:90) (cid:90) n n := n(ε) sufficiently large such that (cid:107)φnf − f(cid:107)H s (cid:54) ε 2 . For such n, we can find a smooth function x (cid:55)→ χ(x) that produces a slow cut-off at infinity so that (cid:107)χφnf − φnf(cid:107)H s (cid:54) ε . 2 0 (R\{0}) satisfying We have thus identified a function fε := χφnf ∈ C∞ (cid:107)fε − f(cid:107)H s (cid:54) (cid:107)χφnf − φnf(cid:107)H s + (cid:107)φnf − f(cid:107)H s (cid:54) ε, (cid:3) which concludes the proof. 2 , n + 3 When n = 2, 3, . . . and s ∈ In = (n + 1 2 ), Sobolev's Lemma guarantees 0 (R3\{0}) comes with the vanishing at x = 0 that the closure in the H s-norm of C∞ of the the function and its first partial derivatives up to order n. Then one can 0 (R3 \{0}) by repeating an analogous argument complete the characterisation of H s as above, now replacing f with its partial derivatives. This yields the formula 3 (cid:98)f (p) dp = 0 f ∈ H s(R3) such that R3 pγ1 0 (R3\{0}) = H s s ∈ (n + 1 2 , n + 3 1 pγ2 2 pγ3 (cid:82) 2 ) . γ1, γ2, γ3 ∈ N0 , γ1 + γ2 + γ3 (cid:54) n − 1  ,  28 A. MICHELANGELI, A. OTTOLINI, AND R. SCANDONE References [1] S. Albeverio, Z. Brze´zniak, and L. Dabrowski, Fundamental solution of the heat and Schrodinger equations with point interaction, J. Funct. Anal., 130 (1995), pp. 220 -- 254. [2] S. Albeverio, F. Gesztesy, R. Høegh-Krohn, and H. Holden, Solvable Models in Quan- tum Mechanics, Texts and Monographs in Physics, Springer-Verlag, New York, 1988. [3] S. Albeverio and P. Kurasov, Singular perturbations of differential operators, vol. 271 of London Mathematical Society Lecture Note Series, Cambridge University Press, Cambridge, 2000. Solvable Schrodinger type operators. [4] E. Capelas de Oliveira and J. J. Vaz, Tunneling in fractional quantum mechanics, J. Phys. A, 44 (2011), pp. 185303, 17. [5] P. D'Ancona, V. Pierfelice, and A. Teta, Dispersive estimate for the Schrodinger equa- tion with point interactions, Math. Methods Appl. Sci., 29 (2006), pp. 309 -- 323. [6] G. Dell'Antonio, A. Michelangeli, R. Scandone, and K. Yajima, Lp-Boundedness of Wave Operators for the Three-Dimensional Multi-Centre Point Interaction, Ann. Henri Poincar´e, 19 (2018), pp. 283 -- 322. [7] M. Gallone, A. Michelangeli, and A. Ottolini, Kreın-Visik-Birman self-adjoint exten- sion theory revisited, SISSA preprint 25/2017/MATE (2017). [8] V. Georgiev, A. Michelangeli, and R. Scandone, On fractional powers of singular per- turbations of the Laplacian, Journal of Functional Analysis, 275 (2018) 1551-1602. [9] F. Iandoli and R. Scandone, Dispersive estimates for Schrodinger operators with point in- teractions in R3, in Advances in Quantum Mechanics: Contemporary Trends and Open Prob- lems, A. Michelangeli and G. Dell'Antonio, eds., Springer INdAM Series, vol. 18, Springer International Publishing, pp. 187 -- 199. [10] S. Jarosz and J. J. Vaz, Fractional Schrodinger equation with Riesz-Feller derivative for delta potentials, J. Math. Phys., 57 (2016), pp. 123506, 16. [11] E. K. Lenzi, H. V. Ribeiro, M. A. F. dos Santos, R. Rossato, and R. S. Mendes, Time dependent solutions for a fractional Schrodinger equation with delta potentials, J. Math. Phys., 54 (2013), pp. 082107, 8. [12] A. Michelangeli, A. Olgiati, and R. Scandone, The singular Hartree equation in frac- tional perturbed Sobolev spaces, Journal Nonlin. Math. Phys. 25, (2018) pp. 1-32, 4 [13] A. Michelangeli and A. Ottolini, On point interactions realised as Ter-Martirosyan- Skornyakov Hamiltonians, Rep. Math. Phys., 79 (2017), pp. 215 -- 260. [14] S. I. Muslih, Solutions of a particle with fractional δ-potential in a fractional dimensional space, Internat. J. Theoret. Phys., 49 (2010), pp. 2095 -- 2104. [15] M. M. Nayga and J. P. Esguerra, Green's functions and energy eigenvalues for delta- perturbed space-fractional quantum systems, J. Math. Phys., 57 (2016), pp. 022103, 7. [16] E. C. d. Oliveira, F. S. Costa, and J. J. Vaz, The fractional Schrodinger equation for delta potentials, J. Math. Phys., 51 (2010), pp. 123517, 16. [17] A. Sacchetti, Stationary solutions of a fractional Laplacian with singular perturbation, arXiv:1801.01694 (2018). [18] T. Sandev, I. Petreska, and E. K. Lenzi, Time-dependent Schrodinger-like equation with nonlocal term, J. Math. Phys., 55 (2014), pp. 092105, 10. [19] S. Scarlatti and A. Teta, Derivation of the time-dependent propagator for the three- interaction, J. Phys. A, 23 (1990), dimensional Schrodinger equation with one-point pp. L1033 -- L1035. [20] J. D. Tare and J. P. H. Esguerra, Bound states for multiple Dirac-δ wells in space- fractional quantum mechanics, J. Math. Phys., 55 (2014), pp. 012106, 10. (A. Michelangeli) International School for Advanced Studies -- SISSA, via Bonomea 265, 34136 Trieste (Italy). E-mail address: [email protected] (A. Ottolini) Department of Mathematics, Stanford University, 450 Serra Mall, Stan- ford CA 94305 (USA). E-mail address: [email protected] (R. Scandone) International School for Advanced Studies -- SISSA, via Bonomea 265, 34136 Trieste (Italy). E-mail address: [email protected]
1904.03592
2
1904
2019-04-09T08:53:07
Tracial moment problems on hypercubes
[ "math.FA", "math.OA" ]
In this paper we introduce the "tracial $K$-moment problem" and the "sequential matrix-valued $K$-moment problem" and show the equivalence of the solvability of these problems. Using a Haviland's theorem for matrix polynomials, we solve these $K$-moment problems for the case where $K$ is the hypercube $[-1,1]^n$.
math.FA
math
TRACIAL MOMENT PROBLEMS ON HYPERCUBES CONG TRINH LE Abstract. In this paper we introduce the tracial K-moment prob- lem and the sequential matrix-valued K-moment problem and show the equivalence of the solvability of these problems. Using a Haviland's the- orem for matrix polynomials, we solve these K-moment problems for the case where K is the hypercube [−1, 1]n. 1. Introduction Let R[X] := R[X1, . . . , Xn] denote the algebra of polynomials in n vari- ables X1, . . . , Xn with real coefficients. For a real-valued linear functional L on R[X] and a closed subset K ⊆ Rn, the K-moment problem asks when does there exist a positive Borel measure σ on Rn supported on K such that L(f ) = Z f (x)dσ(x), ∀f ∈ R[X]? Each functional of this form is called a K-moment functional. Another kind of K-moment problems for sequences of numbers is stated as follows. For a multi-index sequence s = (sα)α∈Nn 0 , the K-moment problem for sequences asks when does there exist a positive Borel measure σ on Rn supported on K such that sα = Z xαdσ(x), ∀α ∈ Nn 0 , where xα = xα1 sequence. 1 . . . xαn n . Each sequence of this form is called a K-moment For each sequence s = (sα)α∈Nn 0 , the Riesz functional Ls on R[X] associ- ated to s is defined by Ls(xα) = sα, ∀α ∈ Nn 0 . It is easy to check that s is a K-moment sequence if and only if Ls is a K-moment functional. In the case K = R (resp. [0, +∞), [0, 1]), we go back to the Hamburger (resp. Stieltjes, Hausdorff ) moment problem which was solved completely by many authors (see, for example, a very excellent book of Schmudgen [10] Date: April 10, 2019. 2010 Mathematics Subject Classification. 44A60; 30E05; 13J30; 47A57; 11E25. Key words and phrases. Moment problem; matrix polynomial; moment functional; mo- ment sequence; hypercube. 1 2 CONG TRINH LE on the moment problem). If K is a compact basic closed semi-algebraic set in Rn, the K-moment problem was solved by Schmudgen (1991, [9]). One of the power tools to solve K-moment problems is the Haviland theorem, which was proved by E.K. Haviland (1935) and stated as follows. Given a linear functional L on R[X] and a closed subset K in Rn. Then L is a K-moment functional if and only if L(f ) ≥ 0 for all f ≥ 0 on K. Here the notation f ≥ 0 on K means that f (x) ≥ 0, ∀x ∈ K. A similar meaning is defined for the notation f > 0 on K. For a fix integer t > 0, we denote by Mt(R[X]) the algebra of t×t matrices with entries in R[X], and by St(R[X]) the subalgebra of symmetric matrices. Each element A ∈ Mt(R[X]) is a matrix whose entries are polynomials in R[X], which is called a polynomial matrix. A is also called a matrix polynomial, because it can be viewed as a polynomial in X1, . . . , Xn whose coefficients come from Mt(R). Namely, we can write A as A(X) = d Xα=0 AαX α, where α = (α1, · · · , αn) ∈ Nn . . . X αn n , Aα ∈ Mt(R). To unify notation, throughout the paper each element of Mt(R[X]) is called a matrix polynomial. 0 , α := α1 + . . . + αn, X α := X α1 1 Let A(X) = d Xα=0 AαX α be a matrix polynomial. Denote Ad(X) := Xα=d AαX α. If Ad 6≡ 0, the matrix polynomial A is called of degree d. In particular, when d n = 1 and A(x) = Aixi is a univariate matrix polynomial of degree d, Xi=0 the non-zero matrix Ad is called the leading coefficient of A. Recently, there are some researchs on (truncated) noncommutative K- moment problems, e.g. Dette and Studden (2002, [4]), Burgdorf and Klep (2011, [1]), Burgdorf, Klep and Povh (2013, [2]), Cimpric and Zalar (2013, [3]), Kimsey and Woerdeman (2013, [6]), Kimsey (2011, [5]) and the refer- ences therein. In particular, in [3], Cimpric and Zalar proved a version of the Haviland theorem for matrix polynomials, and they applied this devel- opment to solve the matrix-valued moment problem on compact basic closed semi-algbraic sets. In the case where K ⊆ Rn is a basic closed semi-algebraic set defined by a subset G = {g1, . . . , gm} of R[X], almost all solutions of the matrix-valued K-moment problems are obtained using the positivity of the given linear functional on the quadratic module M (G)t in Mt(R[X]), where M (G) is the TRACIAL MOMENT PROBLEMS ON HYPERCUBES 3 quadratic module in R[X] generated by G and r M (G)t = { Xi=1 miAT i Air ∈ N, mi ∈ M (G), Ai ∈ Mt(R[X])}. The main aim of this paper is to introduce some kinds of K-moment problems for matrix polynomials and solve these matrix-valued K-moment problems in the case where K is the hypercube [−1, 1]n. The most advantage to work on the hypercube [−1, 1]n (and in general, on convex, compact polyhedra with non-empty interior) is that any matrix polynomial which is positive definite on [−1, 1]n can be represented in terms of positive definite scalar matrices (see Theorem 3.1 for the case n = 1 and Corollary 4.2 for n ≥ 1). Then the solution of the matrix-valued [−1, 1]n-moment problem is obtained basing on the positivity of the given linear functional on the cone of positive definite scalar matrices, which is well described. The paper is organized as follows. In Section 2 we introduce the defi- nition of the tracial K-moment problem and the sequential matrix-valued K-moment problem and establish an equivalence of the solvability of these problems (Proposition 2.8). Moreover, we also prove in this section a version of Haviland's theorem for matrix polynomials which was proved mainly by Cimpric and Zalar in [3]. Next, in Section 3 we solve the tracial (resp. sequen- tial matrix-valued) [−1, 1]-moment problem, which is established based on a representation of univariate matrix polynomials positive definite on [−1, 1]. A solution of the tracial (resp. sequential matrix-valued) [0, 1]-moment prob- lem is also given in this section. Finally, in Section 4 we solve the tracial (resp. sequential matrix-valued) moment problem on the general hypercube [−1, 1]n. Notation. For any matrix A ∈ Mt(R[X]), the notation A ≥ 0 means A is positive semidefinite, i.e. for each x ∈ Rn, vT A(x)v ≥ 0 for all v ∈ Rt; A > 0 means A is positive definite, i.e. for each x ∈ Rn, vT A(x)v > 0 for all v ∈ Rt \ {0}. 2. Haviland's theorem for matrix polynomials 2.1. Matrix-valued measures and integrals. In this section we recall some basic notions of the matrix-valued measures and integrals on a closed subset of Rn, which can be seen from the thesis of Kimsey [5]. Throughout this subsection, let X ⊆ Rn be a non-empty closed set. Definition 2.1. Denote by B(X) the smallest σ-algebra generated from the open (or equivalently closed) subsets of X. A measure σ defined on B(X) is called a Borel measure. A Borel measure σ on B(X) is called finite if σ(X) < +∞. Denote by m(X) the set of all finite measures on X. A measure σ ∈ m(X) is called positive if σ(E) ≥ 0 for all E ∈ B(X). The set of all finite positive Borel measures on X is denoted by m+(X). For each σ ∈ m(X), the support of σ is defined by supp(σ) := {E ∈ B(X) : σ(E) > 0}, 4 CONG TRINH LE where σ(E) := σ(E) for all E ∈ B(X). Definition 2.2. Let t be a positive integer. Let σij ∈ m(X), i, j = 1, . . . , t. Define the matrix-valued function E : B(X) −→ Mt(R) by E(A) := (σij(A))i,j=1,...,t ∈ Mt(R), ∀A ∈ B(X). The matrix-valued function E defined by this way is called a matrix-valued measure on X. The set of all matrix-valued measure on X is denoted by M (X). The set supp(E) := t [i,j=1 supp(σij) is called the support of the matrix-valued measure E. If σij = σji for all i, j = 1, . . . , t, we say that E is a symmetric measure. In addition, if for all A ∈ B(X) and for all v ∈ Rt we have vT E(A)v ≥ 0, then E is called a positive semidefinite matrix-valued measure. The set of all positive semidefinite matrix-valued measures on X is denoted by M +(X). Definition 2.3. Let E = (σij) ∈ M (X). A function f : X → R is called E-measurable if f is σij-measurable for every i, j = 1, . . . , t. Let E = (σij) ∈ M (X) and f : X → R be E-measurable. The matrix- valued integral of f with respect to the matrix measure E is defined by ZX f (x)dE(x) := (cid:16)ZX f (x)dσij(x)(cid:17)i,j=1,...,t ∈ Mt(R). 2.2. Tracial K-moment problems. Let L be a real-valued linear func- tional on Mt(R[X]), where R[X] := R[X1, . . . , Xn] and K ⊆ Rn a closed subset. Definition 2.4 (cf. [3]). L is called a tracial K-moment functional exists a positive semidefinite matrix-valued E ∈ M +(Rn) such that if there supp(E) ⊆ K and L (F) = Z tr(F(x)dE(x)), ∀F ∈ Mt(R[X]). (2.1) The matrix-valued measure E ∈ M +(Rn) satisfying (2.1) is called a repre- senting measure of the tracial K-moment functional L . Problem 2.5. Let L be a linear functional on Mt(R[X]) and K ⊆ Rn a closed set. The tracial K-moment problem asks when does there exist a matrix-valued measure E ∈ M +(Rn) satisfying the conditions (2.1)? The following definition of matrix-valued K-moment sequences is learned from Kimsey in [5]. Definition 2.6. Let S = (Sα)α∈Nn 0 be a multi-indexed sequence of symmet- ric matrices in Mt(R) and K ⊆ Rn a closed subset. The sequence S is called a matrix-valued K-moment sequence if there exists a positive semidefinite matrix-valued E ∈ M +(Rn) such that supp(E) ⊆ K and Sα = Z xαdE(x), ∀α ∈ Nn 0 , (2.2) TRACIAL MOMENT PROBLEMS ON HYPERCUBES 5 where xα = xα1 1 . . . xαn n with α = (α1, . . . , αn) and x = (x1, . . . , xn) ∈ Rn. The matrix-valued measure E ∈ M +(Rn) satisfying (2.2) is called a rep- resenting measure of the matrix-valued K-moment sequence S. Problem 2.7. Let S = (Sα)α∈Nn 0 be a multi-indexed sequence of symmetric matrices in Mt(R) and K ⊆ Rn a closed set. The sequential matrix- valued K-moment problem asks when does there exist a matrix-valued measure E ∈ M +(Rn) satisfying the conditions (2.2)? For each matrix-valued sequence S = (Sα)α∈Nn 0 , let us consider the real- valued function LS on Mt(R[X]) defined as follows. For each F(X) = d Xα=0 FαX α ∈ Mt(R[X]), define LS(F) := d Xα=0 tr(FαSα). It is easy to check that LS is linear, which is called the Riesz functional associated to the sequence S. Proposition 2.8. Let S = (Sα)α∈Nn 0 be a sequence of symmetric matrices in Mt(R) and K ⊆ Rn a closed set. Then S is a matrix-valued K-moment sequence if and only if LS is a tracial K-moment functional. Proof. Assume that S is a matrix-valued K-moment sequence. Then there exists a matrix-valued measure E in M +(Rn) supported in K and Sα = Z xαdE(x), ∀α ∈ Nn 0 . Then, for each F(X) = have d Xα=0 FαX α ∈ Mt(R[X]), by definition of LS, we LS(F) = Xα tr(FαSα) = Xα tr(cid:16)Z xαFαdE(x)(cid:17) = Z tr(cid:16)Xα FαxαdE(x)(cid:17) = Z tr(F(x)dE(x)). This implies that LS is a tracial K-moment functional. Conversely, assume that E is a matrix-valued measure in M +(Rn) satis- fying supp(E) ⊆ K and LS(F) = Z tr(F(x)dE(x)), ∀F ∈ Mt(R[X]). 6 CONG TRINH LE In particular, for each coordinate matrix Wk,l of the algebra Mt(R) and each α ∈ Nn 0 , we have LS(X αWk,l) = Z xαtr(Wk,ldE(x)) = Z xαdEl,k. Here we use the fact that for each k, l = 1, . . . , t and for each matrix A = (Ai,j)i,j ∈ Mt(R), we have tr(Wk,lA) = Al,k. On the other hand, by definition of the associated Riesz functional, we have LS(X αWk,l) = tr(Wk,lSα) = (Sα)l,k. It follows that ST α = (cid:0)(Sα)l,k(cid:1)k,l = (cid:16)Z xαdEl,k(x)(cid:17)k,l = (cid:16)Z xαdE(x)(cid:17)T . Equivalently, Sα = Z xαdE(x), i.e. S is a matrix-valued K-moment se- (cid:3) quence. 2.3. Haviland's theorem for matrix polynomials. Theorem 2.9. Let L be a real-valued linear functional on Mt(R[X]) and K ⊆ Rn a closed set. Then the following are equivalent. (1) L is a tracial K-moment functional, i.e. there exists a positive semi- definite matrix-valued measure E on Rn whose support is contained in K such that L (F) = Z tr(FdE). (2) L (F) ≥ 0 for all F ≥ 0 on K. (3) L (F + ǫI) ≥ 0 for all F ≥ 0 on K and for all ǫ > 0. (4) L (F) ≥ 0 for all F > 0 on K. Proof. The equivalence of (1) and (2) was proved by Cimpric and Zalar [3, Theorem 3 and Remark 5]. (2) → (3) is obvious, since L (I) ≥ 0 by hypothesis. Now we check (3) → (2). for n ∈ N. Then by the assumption of Take F ≥ 0 on K. Consider ǫ = (3), 1 n 0 ≤ L (F + 1 n I) = L (F) + 1 n L (I), where L (I) ≥ 0 by hypothesis. Letting n → +∞, we get L (F) ≥ 0, i.e. we have (2). It is obvious that (2) → (4). Now we verify (4) → (3). Take F ≥ 0 on K and ǫ > 0. Then F + ǫI > 0 on K. Hence by the assumption we have L (F + ǫI) ≥ 0, i.e. we have (3). The proof is complete. (cid:3) TRACIAL MOMENT PROBLEMS ON HYPERCUBES 7 3. Tracial moment problems on the intervals [−1, 1] and [0, 1] Firstly we propose a version of the Bernstein theorem (cf. [10, Prop. 3.4]) for matrix polynomials, representing a matrix polynomial positive definite on the interval [−1, 1]. This is a special case of the Handelman's Positivstel- lensatz for matrix polynomials established in [7], but the proof given here is special for the case of one-variable matrix polynomials. Theorem 3.1. Let F ∈ Mt(R[x]) be a matrix polynomial of degree d > 0. Assume F(x) > 0 for all x ∈ [−1, 1]. Then there exists a positive number N ∈ N and positive definite matrices Gi, i = 1, . . . , N + d such that F(x) = N +d Xi=0 Gi(1 + x)i(1 − x)N +d−i. The main idea of the proof is learned from the proof [10, Prop. 3.4] for polynomials, using the Goursat transform and the Pólya theorem for homo- geneous matrix polynomials established in [8]. Let F be as in Theorem 3.1. The Goursat transform of F is defined by F(x) := (1 + x)dF( 1 − x 1 + x ). Lemma 3.2. deg( F) = d, Fd > 0, and F(t) > 0 for all t ∈ [0, +∞), where Fd denotes the leading coefficient of the matrix polynomial F. Proof. Let F(x) = Pd Fixi, with Fd 6≡ 0. Then we can write F as i=0 d Xi=0 The leading coefficient of F is F(x) = Fi(1 − x)i(1 + x)d−i. Fd := d Xi=0 (−1)iFi = F(−1) which is positive definite by the assumption of F. For every t ∈ [0, +∞), x := ∈ (−1, 1]. Observe that t = 1 − x 1 + x , and hence 1 − t 1 + t F(t) = (1 + t)dF(x). It follows from the assumption of F that F(t) > 0. (cid:3) Lemma 3.3. Let G ∈ Mt(R[x]) be a matrix polynomial of degree d with a positive definite leading coefficient Gd. If G(x) > 0 for all x ∈ [0, +∞) then there exists a positive integer N and positive definite matrices Hi, i = 0, . . . , N + d such that (1 + x)N G(x) = N +d Xi=0 Hixi. 8 CONG TRINH LE Proof. Assume G(x) = Pd R2 Gixi. Denote by ∆ the standard simplex in i=0 ∆ = {(x, y) ∈ R2x ≥ 0, y ≥ 0, x + y + 1}. Let Gh(x, y) ∈ Mt(R[x, y]) be the homogenization of G with respect to the new variable y, defined by Gh(x, y) := d Xi=0 Gixiyd−i. For any (x, y) ∈ ∆, if y = 0, we have x = 1, then If y > 0, we have Gh(1, 0) = Gd > 0. Gh(x, y) = ydG( x y ) > 0, since x y ∈ [0, +∞). In summary, we have Gh(x, y) > 0 for all (x, y) ∈ ∆. It follows from the Pólya theorem for homogeneous matrix polynomials proved by Scherer and Hol [8, Theorem 3] that there exist a positive integer N and positive definite matrices Gk,l, with k, l ∈ N0, k + l = 0, . . . , N + d such that (x + y)N Gh(x, y) = Xk+l=N +d Gk,lxkyl. (3.1) Substituting y = 1 in both sides of the expression (3.1), observing that Gh(x, 1) = G(x), we get (1 + x)N G(x) = Xk+l=N +d Gk,lxk = N +d Xk=0 Gk,N +d−kxk. For each k = 0, . . . , N + d, the matrices Hk := Gk,N +d−k satisfy conclusion of the lemma. (cid:3) Now we are ready to prove Theorem 3.1. Proof of Theorem 3.1. Let F satisfy the hypothesis of the theorem. It follows from Lemma 3.2 that the Goursat transform F satisfy the following properties • deg( F) = d; • Fd > 0; • F(t) > 0 for all t ∈ [0, +∞). Ir follows from Lemma 3.3 that there exist a positive integer N and positive definite matrices Hi, i = 0, . . . , N + d such that (1 + t)N F(t) = N +d Xi=0 Hiti, ∀t ∈ [0, +∞). (3.2) TRACIAL MOMENT PROBLEMS ON HYPERCUBES 9 For any x ∈ (−1, 1], there exists a unique t ∈ [0, +∞) such that x = Then t = 1 − x 1 + x , (1 + t)−1 = 1 + x 2 , and F(t) = (1 + t)dF(x). Substituting into (3.2), we get 1 − t 1 + t . (1 + t)N (1 + t)dF(x) = N +d Xi=0 Hi(cid:0) 1 − x 1 + x(cid:1)i . It follows that N +d F(x) = (1 + t)−N +d Hi(cid:0) Xi=0 N +d 1 − x 1 + x(cid:1)i Hi(cid:0) 1 − x 1 + x(cid:1)i Xi=0 1 2N +d (1 + x)N +d Xi=0 (cid:16) 1 2N +d N +d = = Hi(cid:17)(1 − x)i(1 + x)N +d−i. Hence the matrices Gi := of the theorem. 1 2N +d Hi, i = 0, . . . , N + d, satisfy the conclusion (cid:3) It follows from the Haviland theorem for matrix polynomials (Theorem 2.9) and Theorem 3.1 the following solution of the tracial [−1, 1]-moment problem. Corollary 3.4. Let L be a real-valued linear functional on Mt(R[x]). Then the following are equivalent. (1) L is a tracial [−1, 1]-moment functional. (2) L ((1+x)k(1−x)lG) ≥ 0 for all positive definite matrices G ∈ Mt(R) and k, l ∈ N0. Proof. If (1) holds, for each G positive definite and k, l ∈ N0, since (1 + x)k(1 − x)lG ≥ 0 on [−1, 1], it follows from Theorem 2.9, (1) ⇒ (2), that L ((1 + x)k(1 − x)lG) ≥ 0. Hence we have (1) → (2). Conversely, assume (2) holds. Then for each F ∈ St(R[x]) of degree d, F > 0 on [−1, 1], it follows from Theorem 3.1 that F(x) = N +d Xi=0 Gi(1 + x)i(1 − x)N +d−i for some N > 0 and Gi > 0 for all i = 0, . . . , N + d. Then L (F) = N +d Xi=0 L (Gi(1 + x)i(1 − x)N +d−i) ≥ 0. 10 CONG TRINH LE It follows from Theorem 2.9, (4) ⇒ (1), that L is a tracial [−1, 1]-moment functional, i.e. (2) → (1). The proof is complete. (cid:3) As a consequence of this result, we obtain the following solution of the tracial [0, 1]-moment problem, which was also solved by Cimpric and Zalar [3, Coro. 1], using sums of Hermitian squares. Corollary 3.5. Let L be a real-valued linear functional on Mt(R[x]). Then the following are equivalent. (1) L is a tracial [0, 1]-moment functional. (2) L (xk(1 − x)lG) ≥ 0 for all positive definite matrices G ∈ Mt(R) and k, l ∈ N0. Proof. Observe that there is a bijection from the interval [−1, 1] onto the interval [0, 1] given by x 7→ x + 1 2 , ∀x ∈ [−1, 1]. Using this bijection, it follows from Theorem 3.1 that for F ∈ St(R[x]) of degree d, if F > 0 on [0, 1] then F(x) = N +d Xi=0 Gixi(1 − x)N +d−i for some N > 0 and Gi > 0 for all i = 0, . . . , N + d. Then the result follows from this representation and Theorem 2.9. (cid:3) For the sequential matrix-valued [−1, 1]-problem, we have the following solution. Corollary 3.6. Let S = (Si)i∈N0 be a sequence of symmetric matrices in Mt(R). Then S is a matrix-valued [−1, 1]-moment sequence if and only if k Xi=0 l Xj=0 (−1)j(cid:18)k i(cid:19)(cid:18)l j(cid:19)tr(GSi+j) ≥ 0 for all positive definite matrices G ∈ Mt(R) and k, l ∈ N0. Proof. Let LS be the Riesz functional on Mt(R[x]) associated to the se- quence S. For each G ∈ St(R[x]), G > 0, and k, l ∈ N0, we have k l (−1)j(cid:18)k LS(cid:0)(1 + x)k(1 − x)lG(cid:1) = LS(cid:0) Xi=0 Xj=0 Xj=0 Xi=0 (−1)j(cid:18)k = k i(cid:19)(cid:18)l i(cid:19)(cid:18)l j(cid:19)xi+j G(cid:1) j(cid:19)LS(xi+jG) l l = k Xi=0 Xj=0 (−1)j(cid:18)k i(cid:19)(cid:18)l j(cid:19)tr(GSi+j). TRACIAL MOMENT PROBLEMS ON HYPERCUBES 11 By Proposition 2.8, S is a matrix-valued [−1, 1]-moment sequence if and only if LS is a tracial [−1, 1]-moment functional. It follows from Corollary 3.4 that this is equivalent to LS(cid:0)(1 + x)k(1 − x)lG(cid:1) ≥ 0 for all positive definite matrices G ∈ Mt(R) and k, l ∈ N0. This implies the result. (cid:3) Similarly, using Corollary 3.5, we have the following solution of the se- quential matrix-valued [0, 1]-moment problem. Corollary 3.7. Let S = (Si)i∈N0 be a sequence of symmetric matrices in Mt(R). Then S is a matrix-valued [0, 1]-moment sequence if and only if k Xi=0 (−1)i(cid:18)k i(cid:19)tr(GSi+l) ≥ 0 for all positive definite matrices G ∈ Mt(R) and k, l ∈ N0. 4. Tracial moment problems on the hypercube [−1, 1]n In this section we consider a convex, compact polyhedron K ⊆ Rn with non-empty interior. Assume that K is the basic closed semi-algebraic set defined by linear polynomials L1, . . . , Lm ∈ R[X] := R[X1, . . . , Xn], i.e. K = {x ∈ RnL1(x) ≥ 0, . . . , Lm(x) ≥ 0}. The following version of Handelman's Positivstellensatz for matrix polyno- mials was proved by the the authors in [7]. Theorem 4.1. Let F ∈ St(R[X]) be a matrix polynomial of degree d > 0. If F(x) > 0 for all x ∈ K, then there exist a positive integer N and positive definite matrices Gα ∈ Mt(R), α ∈ Nm 0 , α = 0, . . . , N + d, such that F(X) = N +d Xα=0 GαLα1 1 . . . Lαm m . Applying Theorem 4.1 for the linear polynomials l1 = 1 + X1, l2 = 1 − X1, . . . , l2n−1 = 1 + Xn, l2n = 1 − Xn we obtain the following representation for matrix polynomials positive defi- nite on the hypercube [−1, 1]n. Corollary 4.2. Let F ∈ St(R[X]) be a matrix polynomial of degree d > 0. If F(x) > 0 for all x ∈ [−1, 1]n, then there exist a positive integer N and positive definite matrices Gα ∈ Mt(R), α ∈ N2n 0 , α = 0, . . . , N + d, such that N +d F(X) = Xα=0 Gαlα1 1 lα2 2 . . . lα2n−1 2n−1 lα2n 2n . Now, applying the Haviland theorem for matrix polynomials (Theorem 2.9) and Corollary 4.2 we obtain the following solution of the tracial [−1, 1]n- moment problem. 12 CONG TRINH LE Corollary 4.3. Let L be a real-valued linear functional on Mt(R[X]). Then the following are equivalent. (1) L is a tracial [−1, 1]n-moment functional. (2) L (lα1 2 . . . lα2n−1 2n−1 lα2n 2n 1 lα2 Mt(R) and α = (α1, . . . , α2n) ∈ N2n 0 . G) ≥ 0 for all positive definite matrices G ∈ Observe that the solution of the tracial K-moment problem, where K is a convex, compact polyhedron in Rn with non-empty interior, can be solved by the same way, using the linear polynomial Li instead of lj in Corollary 4.3. Finally, applying Proposition 2.8 and Corollary 4.3 we obtain the following solution of the sequential matrix-valued [−1, 1]n-moment problem. Corollary 4.4. Let S = (Sα)α∈Nn 0 be a sequence of symmetric matrices in Mt(R). Then S is a matrix-valued [−1, 1]n-moment sequence if and only if β(cid:19)tr(GSβ) ≥ 0 for all positive definite matrices G ∈ Mt(R) and α ∈ N2n 0 . (−1)e(α)(cid:18)α Xβ∈N2n 0 ,β≤α Here, for α = (α1, . . . , α2n) and β = (β1, . . . , β2n), β ≤ α means βi ≤ αi n for all i = 1, . . . , 2n; (cid:0)α β(cid:1) := (cid:0)α1 β1(cid:1) . . .(cid:0)α2n β2n(cid:1); and e(α) := α2i. Xi=1 Proof. The proof is similar to that of Corollary 3.6, using Proposition 2.8, Corollary 4.3 and the identity (1+X1)α1(1−X1)α2 . . . (1+Xn)α2n−1 (1−Xn)α2n = Xβ∈N2n 0 ,β≤α (−1)e(α)(cid:18)α β(cid:19)X β, where α = (α1, . . . , α2n). (cid:3) Acknowledgements. The author would like to express his sincere grati- tude to Prof. Konrad Schmudgen for his series of lectures on moment prob- lems given at the Vietnam Institute for Advanced Study in Mathematics (VIASM) which motivates the author to study the matrix-valued moment problems. This paper was finished during the visit of the author at VIASM. He thanks VIASM for financial support and hospitality. References [1] S. Burgdorf and I. Klep The truncated tracial moment problem, J. Oper. Theory 68(1)(2012), 141 -- 163. [2] S. Burgdorf, I. Klep and J. Povh, The tracial moment problem and trace- optimization of polynomials, Math. Prog. 137, Issue 1 -- 2 (2013), 557 -- 578. [3] J.Cimpri c, A. Zalar, Moment problems for operator polynomials, J. Math. Anal. Appl. 401(1) (2013), 307-316. [4] H. Dette and W.J. Studden, Matrix measures, moment spaces and Favard's the- orem for the interval [0, 1] and [0, +∞), Linear Algebra Appl. 345 (2002), 169 -- 193. TRACIAL MOMENT PROBLEMS ON HYPERCUBES 13 [5] D. P. Kimsey, Matrix-valued moment problems, Ph.D. dissertation, Drexel Univer- sity, Philadelphia, PA, 2011. [6] D. P. Kimsey and H. J. Woerdeman, Truncated matrix-valued moment problems on Rd, Cd, and Td, Trans. Amer. Math. Soc. 365(10)(2013), 5393-5430. [7] C.-T. Lê, T.H.B. Du, Handelman's Positivstellensatz for polynomial matrices posi- tive definite on polyhedra, Positivity 22 (3) (2018), 449-460. [8] C.W. Scherer, C.W.J. Hol, Matrix sum-of-squares relaxations for robust semi- definite programs, Math. Program. 107 no. 1-2, Ser. B (2006), 189 -- 211. [9] K. Schm udgen, The K-moment problem for compact semi-algebraic sets, Math. Ann. 289(1) (1991), 203-206. [10] K. Schm udgen, The moment problem, Springer, 2017. Department of Mathematics, Quy Nhon University, 170 An Duong Vuong, Quy Nhon, Binh Dinh, Vietnam E-mail address: [email protected]
1507.06860
1
1507
2015-07-24T14:35:15
What is odd about binary Parseval frames?
[ "math.FA" ]
This paper examines the construction and properties of binary Parseval frames. We address two questions: When does a binary Parseval frame have a complementary Parseval frame? Which binary symmetric idempotent matrices are Gram matrices of binary Parseval frames? In contrast to the case of real or complex Parseval frames, the answer to these questions is not always affirmative. The key to our understanding comes from an algorithm that constructs binary orthonormal sequences that span a given subspace, whenever possible. Special regard is given to binary frames whose Gram matrices are circulants.
math.FA
math
WHAT IS ODD ABOUT BINARY PARSEVAL FRAMES? ZACHERY J. BAKER, BERNHARD G. BODMANN, MICAH G. BULLOCK, SAMANTHA N. BRANUM, AND JACOB E. MCLANEY Abstract. This paper examines the construction and properties of bi- nary Parseval frames. We address two questions: When does a binary Parseval frame have a complementary Parseval frame? Which binary symmetric idempotent matrices are Gram matrices of binary Parseval frames? In contrast to the case of real or complex Parseval frames, the answer to these questions is not always affirmative. The key to our un- derstanding comes from an algorithm that constructs binary orthonor- mal sequences that span a given subspace, whenever possible. Special regard is given to binary frames whose Gram matrices are circulants. 1. Introduction Much of the literature on frames, from its beginnings in non-harmonic Fourier analysis [7] to comprehensive overviews of theory and applications [6, 13, 14] assumes an underlying structure of a real or complex Hilbert space to study approximate expansions of vectors. Indeed, the correspondence be- tween vectors in Hilbert spaces and linear functionals given by the Riesz representation theorem provides a convenient way to characterize Parseval frames, sequences of vectors that behave in a way that is similar to orthonor- mal bases without requiring the vectors to be linearly independent [6]. In- corporating linear dependence relations is useful to permit more flexibility for expansions and to suppress errors that may model faulty signal trans- missions in applications [20, 21, 17, 18, 12, 16, 3]. The concept of frames has also been established even in vector spaces without (definite) inner product [4, 11]. In fact, the well known theory of binary codes could be seen as a form of frame theory, in which linear dependence relations among binary vectors are examined [19, 10, 2]. Here, binary vector spaces are defined over the finite field with two elements; a frame for a finite dimensional binary vector space is simply a spanning sequence [4]. In a preceding paper [5], the study of binary codes from a frame theoretic perspective has lead to additional combinatorial insights in the design of error-correcting codes. Date: January 30, 2018. 2010 Mathematics Subject Classification. 42C15. Key words and phrases. frames, Parseval frames, finite-dimensional vector spaces, bi- nary numbers, orthogonal extension principle, switching equivalence, Gram matrices. This research was supported by NSF grant DMS-1109545 and DMS-1412524. 1 2 Z. J. BAKER, B. G. BODMANN, M. G. BULLOCK, S. N. BRANUM, AND J. MCLANEY The present paper is concerned with binary Parseval frames. These binary frames provide explicit expansions of binary vectors using a bilinear form that resembles the dot product in Euclidean spaces. In contrast to the inner product on real or complex Hilbert spaces, there are many nonzero vectors whose dot product with themselves vanishes. Such vectors have special significance in our results. Counting the number of non-vanishing entries motivates calling them even vectors, whereas the others are called odd. As a consequence of the degeneracy of the bilinear form, there are some striking differences with frame theory over real or complex Hilbert spaces. In this paper, we explore the construction and properties of binary Parseval frames, and compare them with real and complex ones. Our main results are as follows: In the real or complex case, it is known that each Parseval frame has a Naimark complement [6]. The complementarity is most easily formulated by stating that the Gram matrices of two complementary Parseval frames sum to the identity. We show that in the binary case, not every Parseval frame has a Naimark complement. In addition, we show that a necessary and sufficient condition for its existence is that the Parseval frame contains at least one even vector. Moreover, we study the structure of Gram matrices. The Gram matrices of real or complex Parseval frames are characterized as symmetric or hermit- ian idempotent matrices. The binary case requires the additional condition that at least one column vector of the matrix is odd. The general results we obtain are illustrated with examples. Special re- gard is given to cyclic binary Parseval frames, whose Gram matrices are circulants. 2. Preliminaries We define the notions of a binary frame and a binary Parseval frame as in a previous paper [4]. The vector space that these sequences of vectors span is the direct sum Zn 2 = Z2 ⊕ · · · ⊕ Z2 of n copies of Z2 for some n ∈ N. Here, Z2 is the field of binary numbers with the two elements 0 and 1, the neutral element with respect to addition and the multiplicative identity. Definition 2.1. A binary frame is a sequence F = {f1, . . . , fk} in a binary vector space Zn 2 such that span F = Zn 2 . A simple example of a frame is the canonical basis {e1, e2, . . . , en} for 2 . The ith vector has components (ei)j = δi,j, thus (ei)i = 1 is the only i=1 is expanded Zn non-vanishing entry for ei. Consequently, a vector x = (xi)n in terms of the canonical basis as x =Pn Frames provide similar expansions of vectors in linear combinations of the frame vectors. Parseval frames are especially convenient for this purpose, because the linear combination can be determined with little effort. In the real or complex case, this only requires computing values of inner products between the vector to be expanded and the frame vectors. Although we i=1 xiei. With the help of this dot product, we define a Parseval frame for Zn 2 . Definition 2.3. A binary Parseval frame is a sequence of vectors F = {f1, . . ., fk} in Zn 2 , the sequence satisfies the reconstruction identity 2 such that for all x ∈ Zn k (2.1) x = (x, fj)fj . Xj=1 WHAT IS ODD ABOUT BINARY PARSEVAL FRAMES? 3 cannot introduce a non-degenerate inner product in the binary case, we define Parseval frames using a bilinear form that resembles the dot product on Rn. Other choices of bilinear forms and a more general theory of binary frames have been investigated elsewhere, see [9]. Definition 2.2. The dot product on Zn Z2 given by 2 is the bilinear map (·, ·) : Zn 2 × Zn 2 → xn(cid:19) , y1 yn!! := (cid:18) x1 ... ... n Xi=1 xiyi. To keep track of the specifics of such a Parseval frame, we then also say that F is a binary (k, n)-frame. In the following, we use matrix algebra whenever it is convenient for establishing properties of frames. We write A ∈ Mm,n(Z2) when A an m × n 2 to Zm matrix with entries in Z2 and identify A with the linear map from Zn 2 induced by left multiplication of any (column) vector x ∈ Zn 2 with A. We let A∗ denote the adjoint of A ∈ Mm,n(Z2); that is, (Ax, y) = (x, A∗y) for all x ∈ Zn 2 and consequently, A∗ is the transpose of A. 2 , y ∈ Zm Definition 2.4. Each frame F = {f1, . . . , fk} is associated with its anal- ysis matrix ΘF , whose ith row is given by the ith frame vector for i ∈ {1, 2, . . . , k}. Its transpose Θ∗ F is called the synthesis matrix. With the help of matrix multiplication, the reconstruction formula (2.1) of a binary (k, n)-frame F with analysis matrix ΘF is simply expressed as (2.2) where In is the n × n identity matrix. We also note that for any x, y ∈ Zn 2 , then F ΘF = In , Θ∗ hΘF x, ΘF yi = hx, yi which motivates speaking of ΘF as an isometry, as in the case of real or complex inner product spaces. Another way to interpret identity (2.2) is in terms of the column vectors of ΘF . Again borrowing a concept from Euclidean spaces, we introduce orthonormality. Definition 2.5. We say that a sequence of vectors {v1, v2, . . . , vr} in Zn 2 is orthonormal if (vi, vj) = δi,j for i, j ∈ {1, 2, . . . , r}, that is, the dot product of the pair vi and vj vanishes unless i = j, in which case it is equal to one. 4 Z. J. BAKER, B. G. BODMANN, M. G. BULLOCK, S. N. BRANUM, AND J. MCLANEY Inspecting the matrix identity (2.2), we see that a binary k × n matrix Θ is the analysis matrix of a binary Parseval frame if and only if the columns of Θ form an orthonormal sequence in Zk 2. The orthogonality relations between the frame vectors are recorded in the Gram matrix, whose entries consists of the dot products of all pairs of vectors. Definition 2.6. The Gram matrix of a binary frame F = {f1, f2, . . . , fk} for Zn 2 is the k × k matrix G with entries Gi,j = (fj, fi). It is straightforward to verify that the Gram matrix of F is expressed as the composition of the analysis and synthesis matrices, G = ΘF Θ∗ F . The identity (2.2) implies that the Gram matrix of a Parseval frame satisfies the equations G = G∗ = G2 . For frames over the real or complex numbers, these equations characterize the set of all Gram matrices of Parseval frames as orthogonal projection matrices. However, in the binary case, this is only a necessary condition, as shown in the following proposition and the subsequent example. Proposition 2.7. If M is binary matrix that satisfies M = M 2 = M ∗ and it has only even column vectors, then M is not the Gram matrix of a binary Parseval frame. Proof. If G is the Gram matrix of a Parseval frame with analysis operator Θ, then GΘ = ΘΘ∗Θ = Θ, and thus for each column ω of Θ, we obtain the eigenvector equation Gω = ω. By the orthonormality of the columns of Θ, each ω is odd. On the other hand, if M has only even columns, then any eigenvector corresponding to eigenvalue one is even, because it is a linear combination of the column vectors of M . This means M cannot be the Gram matrix of a binary Parseval frame. (cid:3) The following example shows that idempotent symmetric matrices that are not Gram matrices of binary Parseval frames exist for any odd dimension k ≥ 3. Example 2.8. Let k ≥ 3 be odd and let M be the k × k matrix whose entries are all equal to one except for vanishing entries on the diagonal, Mi,j = 1 − δi,j , i, j ∈ {1, 2, . . . , k}. This matrix satisfies M = M 2 = M ∗, but only has even columns and by the preceding proposition, it is not the Gram matrix of a binary Parseval frame. As shown in Section 4, having only even column vectors is the only way a binary symmetric idempotent matrix can fail to be the Gram matrix of a Parseval frame. The construction of Example 2.8 is intriguing, because the alternative choice where k is odd and all entries of M are equal to one is the WHAT IS ODD ABOUT BINARY PARSEVAL FRAMES? 5 Gram matrix of a binary Parseval frame. The relation between these two alternatives can be interpreted as complementarity, which will be explored in more detail in the next section. 3. Complementarity for binary Parseval frames Over the real or complex numbers, each Parseval frame has a so-called Naimark complement [6]; if G is the Gram matrix of a real or complex Parseval frame, then it is an orthogonal projection matrix, and so is I − G, which makes it the Gram matrix of a complementary Parseval frame. We adopt the same definition for the binary case. Definition 3.1. Two binary Parseval frames F and G having analysis oper- ators ΘF ∈ Mk,n(Z2) and ΘG ∈ Mk,k−n(Z2) are complementary if ΘF Θ∗ F + ΘGΘ∗ G = Ik . We also say that F and G are Naimark complements of each other. There is an equivalent statement of complementarity in terms of the block matrix U = (ΘF ΘG) formed by adjoining ΘF and ΘG being orthogonal, meaning U U ∗ = U ∗U = I, just as in the real case (or as U being unitary in the complex case). Proposition 3.2. Two binary Parseval frames F and G having analysis operators ΘF ∈ Mk,n(Z2) and ΘG ∈ Mk,k−n(Z2) are complementary if and only if the block matrix (ΘF ΘG) is an orthogonal k × k matrix. Proof. In terms of the block matrix (ΘF ΘG), the complementarity is ex- pressed as (ΘF ΘG)(ΘF ΘG)∗ = Ik . Since U = (ΘF ΘG) is a square matrix, U U ∗ = I is equivalent to U ∗ also being a left inverse of U , meaning U U ∗ = U ∗U = Ik, U is orthogonal. (cid:3) In the binary case, not every Parseval frame has a Naimark complement. For example, if k ≥ 3 is odd and n = 1, the frame consisting of k vectors {1, 1, . . . , 1} in Z2 is Parseval, and the Gram matrix G is the k × k matrix whose entries are all equal to one. However, I − G ≡ I + G is the matrix M appearing in Example 2.8, which is not the Gram matrix of a binary Parseval frame. This motivates the search for a condition that characterizes the existence of complementary Parseval frames. 3.1. A simple condition for the existence of complementary Parse- val frames. We observe that if F is a Parseval frame with analysis operator ΘF that extends to an orthogonal matrix, then the column vectors of ΘF are a subset of a set of n orthonormal vectors. This is true in the binary as well as the real or complex case. Thus, one could try to relate the construction of a complementary Parseval frame to a Gram-Schmidt orthogonalization strategy. Indeed, this idea allows us to formulate a concrete condition that 6 Z. J. BAKER, B. G. BODMANN, M. G. BULLOCK, S. N. BRANUM, AND J. MCLANEY characterizes when F has a complementary Parseval frame. We prepare this result with a lemma about extending orthonormal sequences. Lemma 3.3. A binary orthonormal sequence Y = {v1, v2, . . . , vr} in Zk 2 with r ≤ k − 1 extends to an orthonormal sequence {v1, v2, . . . , vk} if and 2 whose entries are all equal i=1 vi 6= ιk, where ιk is the vector in Zk only if Pr to one. Proof. If the sequence extends, then {v1, v2, . . . , vk} forms a Parseval frame for Zk i=1(ιk, vi)vi = ιk. On the other hand, the orthonormality forces the set {v1, v2, . . . , vk} to be linearly independent, so ιk cannot be expressed as a linear combination of a proper subset. 2, and by the orthonormality, Pk i=1 vi = Pk To show the converse, we use an inductive proof. Let V be the analysis operator associated with an orthonormal sequence {v1, v2, . . . , vs}, r ≤ s ≤ i=1 vi 6= ιk. To extend the sequence by one vector, we need to find vs+1 with (vs+1, vs+1) = (vs+1, ιk) = 1 and with (vj , vs+1) = 0 for all 1 ≤ j ≤ s. Using block matrices this is summarized in the equation k − 1 satisfying Ps (3.1) (cid:18) V ι∗ k (cid:19) vs+1 =(cid:18) 0s 1 (cid:19) , where 0s is the zero vector in Zs 2. the orthonormal sequence. This is all that is needed if s = k − 1. In order to verify that this equation is consistent, we note that by the orthonormality of the sequence {v1, v2, . . . , vs}, the vector ιk is a linear com- i=1 vi = ιk. Thus, there exists vs+1 which extends bination if and only if Ps chosen so that Ps+1 there is one choice of vs+1 such that Ps+1 Next, we need to show that if s ≤ k − 2, then a solution vs+1 can be i=1 vi 6= ιk, so that the iterative extension procedure can be continued. The solution set of equation (3.1) forms an affine subspace of 2 having dimension k − (s + 1), thus contains 2k−s−1 elements. If s ≤ k − 2, Zk then there are at least two elements in this affine subspace. Consequently, (cid:3) We are ready to characterize the complementarity property for binary Parseval frames. The condition that determines the existence of a Naimark complement is whether at least one frame vector is even, that is, its entries sum to zero. i=1 vi 6= ιk. Theorem 3.4. A binary (k, n)-frame F with n < k has a complementary Parseval frame if and only if at least one frame vector is even. Proof. We first rewrite the condition on the frame vectors in the equivalent form ΘF ιn 6= ιk. Re-expressed in terms of the column vectors {ω1, ω2, . . . , ωn} of ΘF , we claim that a complementary Parseval frame exists if and only if Pn i=1 ωn 6= ιk. On the other hand, the existence of a complementary Parseval frame is equivalent to the sequence of column vectors having an extension to an orthonormal sequence of k elements. WHAT IS ODD ABOUT BINARY PARSEVAL FRAMES? 7 The preceding lemma provides the existence of such an extension if and (cid:3) i=1 ωn 6= ιk, which finishes the proof. only if Pn 3.2. A catalog of binary Parseval frames with the complementarity property. A previous work contained a catalogue of binary Parseval frames for Zn 2 when n was small [4]. Here, we wish to compile a list of the binary Parseval frames that have a complementary Parseval frame. For notational convenience, we consider ΘF instead of the sequence of frame vectors. By Proposition 3.2, every such ΘF is obtained by a selection of columns from a binary orthogonal matrix, so we could simply list the set of all orthogonal matrices for small k. However, such a list quickly becomes extensive as k increases. To reduce the number of orthogonal matrices, we note that although the frame depends on the order in which the columns are selected to form ΘF , the Gram matrix does not. Identifying frames whose Gram matrices coincide has already been used to avoid repeating information when examining real or complex frames [1] and binary frames [4]. We consider an even coarser underlying equivalence relation [8, 12, 3] that has also appeared in the context of binary frames [4]. Definition 3.5. Two families F = {f1, f2, . . . , fk} and G = {g1, g2, . . . gk} in Zn 2 are called switching equivalent if there is an orthogonal n × n matrix U and a permutation π of the set {1, 2, . . . k} such that fj = U gπ(j) for all j ∈ {1, 2, . . . } . Representing the permutation π by the associated permutation matrix P with entries Pi,j = δi,π(j) gives that if F and G are switching equivalent, then ΘF = P ΘGU with an orthogonal n × n matrix U and a k × k permutation matrix P . Alternatively, switching equivalence is stated in the form of an identity for the corresponding Gram matrices. Theorem 3.6 ([4]). Two binary (k, n)-frames F and G are switching equiv- alent if and only if their Gram matrices are related by conjugation with a k × k permutation matrix P , GF = P GG P ∗ . We deduce a consequence for switching equivalence and Naimark comple- ments, which is inferred from the role of the Gram matrices in the definition of complementarity. Corollary 3.7. If F and G are switching equivalent binary (k, n)-frames, then F has a Naimark complement if and only if G does. Thus, to provide an exhaustive list, we only need to ensure that at least one representative of each switching equivalence class appears as a selection of columns in the orthogonal matrices we include. To reduce the number of representatives, we identify matrices up to row and column permutations. 8 Z. J. BAKER, B. G. BODMANN, M. G. BULLOCK, S. N. BRANUM, AND J. MCLANEY Definition 3.8. Two matrices A, B ∈ Mk,k(Z2) are called permutation equiv- alent if there are two permutation matrices P1, P2 ∈ Mk,k(Z2) such that A = P1BP ∗ 2 . Proposition 3.9. If U1 and U2 are permutation equivalent binary orthogo- nal matrices, then each (k, n)-frame F formed by a sequence of n columns of U1 is switching equivalent to a (k, n)-frame G formed with columns of U2. Proof. Without loss of generality, we can assume that the analysis matrix ΘF is formed by the first n columns of U1. By the equivalence of U1 and U2, U1P2 = P1U2 with permutation matrices P1 and P2. The right multiplication of U1 with P2 gives a column permutation, which identifies a sequence of columns in P1U2 that is identical to the first n columns of U1. If G is obtained with the corresponding columns in U2, then the Gram matrices of F and G are related by GF = P1GGP ∗ 1 , which proves the switching equivalence. (cid:3) A list of permutation-inequivalent orthogonal k × k matrices allows us to obtain the Gram matrix of each binary (k, n)-frame with a Naimark com- plement by selecting an appropriate choice of n columns from an orthogonal k × k matrix to form Θ and then by applying a permutation matrix P to obtain GF = P ΘΘ∗P ∗. Accordingly, each representative of an equivalence class of orthogonal ma- trices can be chosen so that the columns are in lexicographical order. Table 1 contains a complete list of representatives of binary orthogonal matrices for k ∈ {3, 4, 5, 6} from each permutation-equivalence class. Each column vec- tor in our list is recorded by the integer obtained from the binary expansion with the entries of the vector. For example, if a frame vector in Z4 2 is f1 = (1, 0, 1, 1), then it is represented by the integer 20 + 22 + 23 = 13. Ac- cordingly, in Z4 2, the standard basis is recorded as the sequence of numbers 1, 2, 4, 8. 4. Gram matrices of binary Parseval frames The preceding section on complementarity hinged on the problem that if G is the Gram matrix of a binary Parseval frame, then I − G may not, although it is symmetric and idempotent. Again, there is a simple condi- tion that needs to be added; Gram matrices of binary Parseval frames are symmetric and idempotent and have at least one odd column, that is, a column whose entries sum to one. Because of the identity G2 = G, having an odd column is equivalent to having a non-zero diagonal entry. Indeed, it has been shown that for any binary symmetric matrix G without vanishing diagonal, there is a factor Θ such that G = ΘΘ∗ and the rank of Θ is equal to that of G [15]. The assumptions needed for our proof are stronger, but our algorithm for producing Θ appears to be more straightforward than the factorization procedure for general symmetric binary matrices. Theorem 4.1. A binary symmetric idempotent matrix M is the Gram ma- trix of a Parseval frame if and only if it has at least one odd column. WHAT IS ODD ABOUT BINARY PARSEVAL FRAMES? 9 8 4 4 4 Orthogonal k × k Matrices, 3 ≤ k ≤ 6 k Integer corresponding to binary column vectors 2 3 1 2 8 4 1 11 13 14 7 2 16 8 5 1 11 19 25 26 4 8 19 21 22 7 11 13 14 16 7 4 16 32 2 6 1 8 19 35 49 50 4 11 16 35 41 42 4 11 19 25 26 32 4 16 35 37 38 8 7 8 19 21 22 32 7 7 11 13 14 16 32 13 14 28 44 55 59 21 22 28 47 52 59 25 26 28 47 55 56 31 37 38 44 52 59 31 41 42 44 55 56 31 47 49 50 52 56 31 47 55 59 61 62 of permutation equivalence Table 1. Representatives classes of orthogonal matrices. Up to switching equivalence, the Gram matrix of each binary (k, n)-frame with a Naimark complement is obtained by selecting appropriate columns in one of the listed k × k orthogonal matrices. Proof. First, we re-express the condition on the columns of a symmetric k × k matrix M in the equivalent form of the matrix Ik + M having at least one even column or row. This, in turn, is equivalent to the inequality (Ik + M )ιk 6= ιk. Next, we recall that both M and Ik + M are assumed to be idempotent. 2 is in the range of an idempotent P if We observe that any vector y ∈ Zk and only if P y = y if and only if y is in the kernel of Ik + P . Assuming M is the Gram matrix of a Parseval frame, then M = ΘΘ∗ where Θ has orthonormal columns and and (M + Ik)Θ = 0. Combining the two properties gives (cid:18) M + Ik ι∗ k (cid:19) Θ =(cid:18) 0k,n ιn (cid:19) . This is inconsistent if and only if ιk is in the span of the columns of the idempotent M + Ik, which is equivalent to (M + Ik)ιk = ιk. 10Z. J. BAKER, B. G. BODMANN, M. G. BULLOCK, S. N. BRANUM, AND J. MCLANEY Conversely, assuming that M is symmetric and idempotent and that the range of Ik + M does not contain ιk, we construct a matrix Θ with orthonor- mal columns such that M = ΘΘ∗. The assumption on M is equivalent to M ιk 6= 0, so at least one row or column of M is odd. Let this column be ω1, then the fact that M is idempotent gives M ω1 = ω1. j ΘΘ∗ωi for each i, j ∈ {1, 2, . . . , n} imply M = ΘΘ∗. Next, we follow an inductive strategy similar to an earlier proof. What we need is an orthonormal sequence {ω1, . . . , ωn} such that n is the rank of M and M ωi = ωi for all i ∈ {1, 2, . . . , n}. In that case, the range of M is the span of the sequence, and so is the range of M ∗. Thus, the identities j M ωi = δi,j = ω∗ ω∗ Proceeding inductively, we need to extend a given orthonormal sequence {ω1, . . . , ωs} in the kernel of Ik+M by one vector if s ≤ n−1 and if ιk is not in the span of the columns of Ik + M combined with the orthonormal sequence. Let V be a matrix formed by a maximal set of linearly independent rows in Ik + M , then if M has rank n, rank nullity gives that V has k − n rows. Letting Y be the analysis matrix of the orthonormal sequence {ω1, . . . , ωs}, then extending it by one vector requires solving the equation (4.1) V Y ι∗ k    ωs+1 =   0k−n 0s 1   . step, we need to show that if s ≤ n − 2, then ι∗ rows of the matrix formed by V , Y and ω∗ by the fact that V Y ∗ = 0, so if ιk =Ps+1 Moreover, in order to guarantee the induction assumption at the next k is not in the span of the s+1. As before, this is obtained i=1 ciωi + v with v being in the span of the columns of V ∗, then Y v = 0 and orthonormality forces ci = 1 for all i ∈ {1, 2, . . . , s + 1}. The solutions of the equation (4.1) form an affine subspace of dimension k − (k − n) − s − 1 = n − s − 1, so if s ≤ n − 2, then there are at least two solutions, one of which does not satisfy the sum (cid:3) identity ιk =Ps+1 i=1 ciωi + v. 5. Binary cyclic frames and circulant Gram matrices Next, we examine a special type of frame whose Gram matrices are cir- culants. We recall that a cyclic subspace V of Zk 2 has the property that it is closed under cyclic shifts, that is, the cyclic shift S, which is characterized by Sej = ej+1 (mod) k, leaves V invariant. Definition 5.1. A frame F = {f1, f2, . . . , fk} for Zn 2 is called a binary cyclic frame if the range of the analysis operator is invariant under the cyclic shift S. If F is also Parseval, then we say that is a binary cyclic (k, n)-frame. Since the range of the Gram matrix G belonging to a Parseval frame is identical to the set of eigenvectors corresponding to eigenvalue one, we have a simple characterization of Gram matrices of binary cyclic Parseval frames. Theorem 5.2. A binary frame F = {f1, f2, . . . , fk} for Zn 2 is a cyclic Par- seval frame if and only if its Gram matrix GF is a symmetric idempotent WHAT IS ODD ABOUT BINARY PARSEVAL FRAMES? 11 circulant matrix, that is, GF = G∗ column vectors. F = G2 F and SGF S∗ = GF , with only odd F = G2 Proof. If GF is the Gram matrix of a binary cyclic Parseval frame, then from the Parseval property, we know that GF = G∗ F . Moreover, by the cyclicity of the frame, the eigenspace corresponding to eigenvalue one of GF is invariant under S, and thus if x = GF x, then Sx = SGF x = GF Sx. Using this identity repeatedly and writing y = Sk−1x = S∗x gives y = SGF S∗y for all y in the range of GF . By the symmetry of GF , the range of GF is identical to that of G∗ F , so hGF x, yi = hSGF S∗x, yi for all x, y in the range of GF establishes the circulant property GF = SGF S∗. If GF is a circulant, then each column vector generates all the others by applying powers of the cyclic shift to it. Thus, if one column vector is odd, so are all the other column vectors. Applying Theorem 4.1 then yields that the Gram matrices of binary cyclic Parseval frames are symmetric idempotent circulant matrices with only odd column vectors. Conversely, if G is a symmetric idempotent circulant and each column vector is odd, then Theorem 4.1 again yields that it is the Gram matrix of a binary Parseval frame F with G = ΘF Θ∗ F . Moreover, the range of G is invariant under the cyclic shift, because one column vector generates all the others by applying powers of the cyclic shift to it. Since the range of G is identical to that of ΘF , F is a cyclic binary Parseval frame. (cid:3) Since adding the identity matrix changes odd columns of G to even columns, we conclude that complementary Parseval frames do not exist for binary cyclic Parseval frames. Corollary 5.3. If F is a binary cyclic Parseval frame, then it has no com- plementary Parseval frame. In Table 2, we provide an exhaustive list of the Gram matrices of cyclic binary Parseval frames with 3 ≤ k ≤ 20. Factoring these into the cor- responding analysis and synthesis matrices shows that many of these ex- amples contain repeated frame vectors. In an earlier paper, such repeated vectors have been associated with a trivial form of redundancy incorporated in the analysis matrix ΘF . Table 3 lists the circulant Gram matrices of rank n < k ≤ 20 paired with k × n analysis matrices, for which no repetition of frame vectors occurs. References [1] Balan, R., Equivalence relations and distances between Hilbert frames. Proceedings AMS 127 (8) (1999), 2353-2366. [2] Betten, A., Braun, M., Fripertinger, H., Kerber, A., Kohnert, A. and Wassermann, A., Error-correcting linear codes, in: Algorithms and Computation in Mathematics 18, Springer-Verlag, Berlin, 2006, xl+798 pp. [3] Bodmann, B. G. and Paulsen, V. I., Frames, graphs and erasures, Linear Algebra Appl. 404 (2005), 118–146. 12Z. J. BAKER, B. G. BODMANN, M. G. BULLOCK, S. N. BRANUM, AND J. MCLANEY [4] Bodmann, B. G. Le, M., Reza, L., Tobin, M. and Tomforde, M., Frame theory for binary vector spaces, Involve, 2 (2009), no. 5, 589-602. [5] Bodmann, B. G., Camp, B. and Mahoney, D., Binary frames, graphs and erasures, Involve 7, 151-169 (2014). [6] Christensen, O., An Introduction to Frames and Riesz Bases, Birkhauser, Boston, (2003). [7] Duffin, R. J. and Schaeffer, A. C., A class of nonharmonic Fourier series, Trans. Amer. Soc. 72 (1952), 341-366. [8] Goyal, V. K., Kovacevi´c, J., and Kelner, J. A., Quantized frame expansions with erasures, Appl. Comp. Harm. Anal., vol. 10, pp. 203–233, 2001. [9] Hotovy, R., Larson, D. and Scholze, S., Binary frames, Houston Journal of Mathe- matics, to appear. [10] Haemers, W. H., Peeters, R., and van Rijckevorsel, J. M., Binary codes of strongly regular graphs, Des. Codes Cryptogr. 17, 187–209 (1999). [11] Han, D., Kornelson, K., Larson, D. and Weber, E., Frames for Undergraduates, Stu- dent Mathematical Library, 40, American Mathematical Society, Providence, RI, 2007. xiv+295 pp. [12] Holmes, R. B., Paulsen, V.I., Optimal frames for erasures. Linear Algebra Appl. 377 (2004), 31–51. DOI 10.1016/j.laa.2003.07.012. [13] Kovacevi´c, J. and Chebira, A., Life beyond bases: The advent of frames (Part I), IEEE Signal Processing Magazine 24 (2007), no. 4, 86-104. [14] Kovacevi´c, J. and Chebira, A., Life beyond bases: The advent of frames (Part II), IEEE Signal Processing Magazine 24 (2007), no. 5, 15-125. [15] Lempel, A., Matrix factorization over GF (2) and trace-orthogonal bases of GF (2n), SIAM J. Comput. 4 (1975), 175–186. [16] Puschel, M., Kovacevi´c, J., Real, tight frames with maximal robustness to erasures. In: Data Compression Conference, 2005. Proceedings. DCC 2005, pp. 63–72 (2005). [17] Rath, G. and Guillemot, C., Performance analysis and recursive syndrome decoding of DFT codes for bursty erasure recovery, IEEE Trans. on Signal Processing, vol. 51 (2003), no. 5, 1335-1350. [18] Rath, G. and Guillemot, C., Frame-theoretic analysis of DFT codes with erasures, IEEE Transactions on Signal Processing, vol. 52 (2004), no. 2, 447-460. [19] MacWilliams, F. J. and Sloane, N. J., The Theory of Error-Correcting Codes, North Holland, Amsterdam, 1977. [20] Marshall, T., J., Coding of real-number sequences for error correction: A digital signal processing problem. IEEE Journal on Selected Areas in Communications 2 (1984), no. 2, 381–392. [21] Marshall, T., Fourier transform convolutional error-correcting codes. In: Signals, Systems and Computers, 1989. Twenty-Third Asilomar Conference on, vol. 2, pp. 658–662 (1989). Department of Mathematics, University of Houston, Houston, TX 77204- 3008, USA E-mail address: [email protected] E-mail address: [email protected] E-mail address: [email protected] E-mail address: [email protected] E-mail address: [email protected] WHAT IS ODD ABOUT BINARY PARSEVAL FRAMES? 13 k 3 4 5 6 7 8 9 First row of circulant k × k Gram matrix, 3 ≤ k ≤ 20 1 0 0 1 1 1 1 0 0 0 1 0 0 0 0 1 1 1 1 1 1 0 0 0 0 0 1 0 1 0 1 0 1 0 0 0 0 0 0 1 1 1 1 1 1 1 1 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0 1 0 0 1 0 0 1 1 1 0 1 1 0 1 1 1 1 1 1 1 1 1 1 1 10 1 0 0 0 0 0 0 0 0 0 1 0 1 0 1 0 1 0 1 0 11 1 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 1 1 12 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 1 0 0 0 1 0 0 0 13 1 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 1 1 1 1 14 1 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 15 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 1 0 0 0 0 1 0 0 0 0 1 0 0 1 0 0 1 0 0 1 0 0 1 0 0 1 0 0 1 0 1 1 0 0 1 1 0 1 0 0 1 1 1 0 1 0 0 1 1 0 0 1 0 1 1 1 1 1 0 1 1 0 1 1 0 1 1 0 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 16 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 17 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 1 0 1 1 1 0 0 1 1 1 0 1 0 0 1 1 1 0 1 0 0 0 1 1 0 0 0 1 0 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 18 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 1 0 1 0 1 0 0 0 1 0 1 0 0 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 1 0 19 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 20 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 1 0 0 0 1 0 0 0 1 0 0 0 1 0 0 0 Table 2. Circulant Gram matrices of binary cyclic (k, n)- frames, first row shown. 14Z. J. BAKER, B. G. BODMANN, M. G. BULLOCK, S. N. BRANUM, AND J. MCLANEY k = 9 Circulant Gram matrix 1 1 1 0 1 1 0 1 1 1 1 1 1 0 1 1 0 1 1 1 1 1 1 0 1 1 0 0 1 1 1 1 1 0 1 1 1 0 1 1 1 1 1 0 1 1 1 0 1 1 1 1 1 0 0 1 1 0 1 1 1 1 1 1 0 1 1 0 1 1 1 1 1 1 0 1 1 0 1 1 1 1 1 1 0 1 0 0 1 1 0 0 1 0 1 1 1 1 1 1 0 1 0 0 1 1 0 0 1 0 1 1 1 1 1 1 0 1 0 0 1 1 0 0 1 0 0 1 1 1 1 1 0 1 0 0 1 1 0 0 1 1 0 1 1 1 1 1 0 1 0 0 1 1 0 0 0 1 0 1 1 1 1 1 0 1 0 0 1 1 0 0 0 1 0 1 1 1 1 1 0 1 0 0 1 1 1 0 0 1 0 1 1 1 1 1 0 1 0 0 1 1 1 0 0 1 0 1 1 1 1 1 0 1 0 0 0 1 1 0 0 1 0 1 1 1 1 1 0 1 0 0 0 1 1 0 0 1 0 1 1 1 1 1 0 1 1 0 0 1 1 0 0 1 0 1 1 1 1 1 0 0 1 0 0 1 1 0 0 1 0 1 1 1 1 1 1 0 1 0 0 1 1 0 0 1 0 1 1 1 1 1 1 0 1 0 0 1 1 0 0 1 0 1 1 1 1 0 0 1 0 1 1 0 0 1 1 0 1 0 0 0 1 0 0 1 0 1 1 0 0 1 1 0 1 0 0 0 1 0 0 1 0 1 1 0 0 1 1 0 1 1 0 0 1 0 0 1 0 1 1 0 0 1 1 0 0 1 0 0 1 0 0 1 0 1 1 0 0 1 1 1 0 1 0 0 1 0 0 1 0 1 1 0 0 1 1 1 0 1 0 0 1 0 0 1 0 1 1 0 0 0 1 1 0 1 0 0 1 0 0 1 0 1 1 0 0 0 1 1 0 1 0 0 1 0 0 1 0 1 1 1 0 0 1 1 0 1 0 0 1 0 0 1 0 1 1 1 0 0 1 1 0 1 0 0 1 0 0 1 0 0 1 1 0 0 1 1 0 1 0 0 1 0 0 1 1 0 1 1 0 0 1 1 0 1 0 0 1 0 0 0 1 0 1 1 0 0 1 1 0 1 0 0 1 0 0 0 1 0 1 1 0 0 1 1 0 1 0 0 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 0 1 1 1 1 1 1 1 1 1 0 1 1 1 1 1 0 1 1 1 1 1 1 1 1 1 0 1 1 1 1 1 0 1 1 1 1 1 1 1 1 1 0 1 1 1 1 1 0 1 1 1 1 1 1 1 1 1 0 1 1 1 1 1 0 1 1 1 1 1 1 1 1 1 0 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 1 1 0 1 1 1 1 0 1 1 1 1 1 1 1 1 0 1 1 0 1 1 0 1 1 0 1 1 1 1 1 1 0 1 1 0 1 1 0 1 1 0 1 1 1 1 1 1 0 1 1 0 1 1 0 1 1 0 0 1 1 1 1 1 0 1 1 0 1 1 0 1 1 1 0 1 1 1 1 1 0 1 1 0 1 1 0 1 1 1 0 1 1 1 1 1 0 1 1 0 1 1 0 0 1 1 0 1 1 1 1 1 0 1 1 0 1 1 1 0 1 1 0 1 1 1 1 1 0 1 1 0 1 1 1 0 1 1 0 1 1 1 1 1 0 1 1 0 0 1 1 0 1 1 0 1 1 1 1 1 0 1 1 1 0 1 1 0 1 1 0 1 1 1 1 1 0 1 1 1 0 1 1 0 1 1 0 1 1 1 1 1 0 0 1 1 0 1 1 0 1 1 0 1 1 1 1 1 1 0 1 1 0 1 1 0 1 1 0 1 1 1 1 1 1 0 1 1 0 1 1 0 1 1 0 1 1 1 Corresponding analysis matrix 1 0 0 1 1 1 1 1 1 1 0 0 1 1 1 1 1 1 1 0 0 0 1 0 1 1 1 1 0 0 0 1 0 1 1 0 0 0 0 0 1 0 0 0 1 1 1 1 1 0 0 0 0 1 1 1 0 0 0 0 0 0 1 1 0 1 0 1 1 1 1 1 0 1 1 0 1 0 1 0 1 1 0 0 1 0 1 1 0 1 1 1 1 0 1 1 1 0 1 1 0 0 1 0 0 0 1 1 0 0 0 1 1 0 0 0 0 1 1 1 1 0 1 0 0 0 0 1 1 0 0 1 0 0 0 0 1 0 1 1 0 0 0 0 0 1 0 0 0 1 1 1 1 1 0 0 0 0 1 1 1 0 0 0 0 0 0 1 n = 7 n = 9 1 0 1 0 0 0 1 0 0 0 1 0 0 1 0 0 1 0 1 0 0 0 0 0 1 0 1 1 1 1 0 0 0 1 1 0 1 1 1 1 1 0 0 1 1 0 1 1 0 1 1 0 0 1 1 0 1 1 1 0 1 0 0 0 1 1 1 1 1 1 1 0 0 0 0 1 1 1 0 1 1 0 0 0 0 0 1 1 0 1 0 0 1 1 0 0 0 1 0 0 0 0 0 1 1 0 0 1 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 1 1 0 0 1 1 1 1 1 1 1 1 1 1 1 0 0 1 1 1 1 1 1 1 1 1 1 1 0 0 1 1 1 1 1 1 1 1 1 1 1 0 0 1 1 1 1 1 1 1 1 1 1 1 0 0 0 1 0 1 1 1 1 1 1 1 1 0 0 0 1 0 1 1 1 1 1 1 0 0 0 0 0 1 0 1 1 1 1 0 0 0 0 0 0 0 1 0 1 1 0 0 0 0 0 0 0 0 0 1 0 0 0 1 1 1 1 1 1 1 1 1 0 0 0 0 1 1 1 1 1 1 1 0 0 0 0 0 0 1 1 1 1 1 0 0 0 0 0 0 0 0 1 1 1 0 0 0 0 0 0 0 0 0 0 1 1 0 0 1 1 1 1 0 0 1 1 1 1 1 1 1 0 0 1 1 1 1 0 0 1 1 1 1 1 1 1 0 0 1 1 1 1 0 0 0 1 0 1 1 1 1 0 0 1 1 1 1 0 0 0 1 0 1 1 1 1 0 0 1 1 0 0 0 0 0 1 0 1 1 1 1 0 0 0 0 1 1 1 1 1 0 0 1 1 1 1 0 0 0 0 1 1 1 1 1 0 0 1 1 0 0 0 0 0 0 1 1 1 1 1 0 0 0 0 0 0 0 0 0 1 0 1 1 1 1 0 0 0 0 0 0 0 0 0 1 0 1 1 0 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 0 1 1 1 1 1 0 0 0 0 0 0 0 0 0 0 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 1 n = 11 n = 13 Table 3. Gram and analysis matrices of binary cyclic (k, n)- frames, k > n, whose vectors do not repeat. k = 15
1906.03021
1
1906
2019-06-07T11:33:15
Convergence in variation for the multidimensional generalized sampling series and applications to smoothing for digital image processing
[ "math.FA" ]
In this paper we study the problem of the convergence in variation for the generalized sampling series based upon averaged-type kernels in the multidimensional setting. As a crucial tool, we introduce a family of operators of sampling-Kantorovich type for which we prove convergence in L^p on a subspace of L^p(R^N): therefore we obtain the convergence in variation for the multidimensional generalized sampling series by means of a relation between the partial derivatives of such operators acting on an absolutely continuous function f and the sampling-Kantorovich type operators acting on the partial derivatives of f. Applications to digital image processing are also furnished.
math.FA
math
Convergence in variation for the multidimensional generalized sampling series and applications to smoothing for digital image processing Laura Angeloni, Danilo Costarelli, Gianluca Vinti Department of Mathematics and Computer Science University of Perugia, Via Vanvitelli 1, 06123, Perugia, Italy [email protected] - [email protected] [email protected] Abstract In this paper we study the problem of the convergence in variation for the generalized sampling series based upon averaged-type kernels in the multidimensional setting. As a crucial tool, we introduce a family of operators of sampling-Kantorovich type for which we prove convergence in Lp on a subspace of Lp(RN ): therefore we obtain the convergence in variation for the multidimensional generalized sampling series by means of a relation between the partial derivatives of such operators acting on an absolutely continuous function f and the sampling-Kantorovich type operators acting on the partial derivatives of f. Applications to digital image processing are also furnished. AMS 2010 Mathematics Subject Classification: 41A30, 41A05 Key words and phrases: convergence in variation; multidimensional general- ized sampling series; sampling-Kantorovich operators; variation diminishing type property; smoothing in digital image processing 1 1 Introduction In this paper we present approximation results in BV-spaces for the generalized sampling series in the multidimensional frame. The above sampling series, defined as (I) (Swf )(t) := f χ(wt − k), t ∈ R, w > 0, (cid:18) k (cid:19) w (cid:88) k∈Z has been introduced by P.L. Butzer (see, e.g., [24]) and revealed to be very interesting for both the theoretical and applicative aspects. Indeed, several approximation results have been obtained in the last forty years with respect to different notions of convergence, such as uniform, Lp, and modular conver- gence ([18, 37, 16]). Moreover, they have many important applications to Signal Theory since they provide an approximate sampling formula, which allows to reconstruct not necessarily band-limited signals. By its multivariate generaliza- tion, that is, (II) (Swf )(t) := f χ(wt − k), t ∈ RN , w > 0, (cid:18) k (cid:19) w (cid:88) k∈ZN several problems of image processing can be treated (see, e.g., [17]). A natural setting to study some issues involving digital images is furnished by the spaces of functions of bounded variation ([36, 44, 47]). In this direction, it is interesting to have results about estimates and convergence in variation for the above discrete operators. In the case of functions of one-variable, the problem has been faced in [6] for the operators (I) based upon averaged type kernels. Here, also with the aim to consider applicative aspects, we deal with the multivariate case, namely we study results about approximation in variation by means of the operators (II) based upon multidimensional product kernels of averaged type, namely where ¯χi,m(t) := 1 m ¯χm(t) := ¯χi,m(ti) χi(t + v) dv, for some m ∈ N, and χi : R −→ R are suitable one-dimensional kernels. We will use the multidimensional generalization of the classical Jordan variation N(cid:89) (cid:90) m i=1 2 − m 2 2 introduced by Tonelli (see, e.g., [45]) and later generalized by T. Radó ([40]) and C. Vinti ([48]) to the case of functions of N-variables. In particular, we establish a variation diminishing-type property (Proposi- tion 1) and a convergence theorem (Theorem 1). Due to form of the above discrete sampling type operators, to obtain results about approximation in variation is a very delicate problem. The strategy we propose is to introduce a new family of Kantorovich type operators defined as w (cid:88) k∈ZN (cid:90) kj +1 w kj w (cid:18) k1 w f (cid:19)  χ(wt − k), (III) (Kw,jf )(t) := , . . . , u . . . , kN w du t ∈ RN, w > 0, j = 1, . . . , N, for which we establish Lp-convergence in a subspace of Lp(RN ) (Theorem 4), therefore obtaining a convergence result for the new class of operators (III). We notice that the above result is obtained through the use of the τ−modulus of continuity (see [34, 33, 42]) which seems to be the most suitable approach in this setting. The crucial point in order to reach the convergence in variation is to find a link between the two classes of operators: to this aim we prove (Proposition 2) a relation between the gradient of the multivariate generalized sampling series of a function f and the family of Kantorovich-type discrete operators acting on the partial derivatives of f. Some of the results of the present paper may have an applicative interpre- tation: in particular, the variation diminishing type property (Proposition 1), ensuring that the generalized sampling series (II) have smaller variation than the function on which they act, can be viewed as a smoothing procedure. Indeed, if the function f is a gray scale digital image, the action of the operators and the consequent variation diminishing corresponds to reducing the jumps of gray levels with respect to the original image, hence producing a smoothing effect. In Section 6 we discuss such digital image processing applications in details, furnishing some numerical examples. 2 Notations and preliminaries Our results will be set in the frame of the space of multivariate functions of bounded variation: in particular we will consider the concept of variation intro- duced by Tonelli ([45]) for two variables, extended to the general case of RN by 3 Radó and C. Vinti ([40, 48]). Here we will recall it. For a function f : RN → R and x = (x1, . . . , xN ) ∈ RN, we will use the notation j = (x1, . . . , xj−1, xj+1, . . . , xN ) ∈ RN−1, x = (x (cid:48) (cid:48) (cid:48) j, xj), j, xj), f (x) = f (x x an N−dimensional interval I =(cid:81)N if we are interested in the j−th coordinate of x, j = 1, . . . , N. Moreover, given j] we will denote the (N − 1)−dimensional interval obtained deleting by I the j−th coordinate, i.e., i=1[ai, bi], by I(cid:48) j = [a(cid:48) j, b(cid:48) I = [a (cid:48) j, b j] × [aj, bj], (cid:48) j = 1, . . . , N. α =(cid:0) x1 Given a vector x ∈ RN and α ∈ R, we will use the usual notation for products and quotients, i.e., αx = (αx1, . . . , αxN ) and, for α (cid:54)= 0, x By Rloc(RN ) we will denote the space of locally Riemann integrable functions on RN, while M (RN ) will denote the space of all the measurable and bounded functions f : RN → R. Definition 1 A function f ∈ M (RN ) is said to be of bounded variation (f ∈ j,·)] (the usual Jordan one-dimensional variation of the j−th BV (RN )) if VR[f (x(cid:48) section of f) is finite for a.e. x(cid:48) α , . . . , xN α (cid:1). (cid:90) j ∈ RN−1 and j < +∞, j,·)] dx (cid:48) (cid:48) VR[f (x for every j = 1, . . . , N. RN−1 For more details about BV -spaces, see, e.g., [9, 5, 10, 11, 4, 13]. The first step is to consider, for I = (cid:81)N Let us now recall how to compute the multidimensional Tonelli variation. i=1[ai, bi] and j = 1, . . . , N the (N − 1)−dimensional integrals (cid:90) Φj(f, I) := [a(cid:48) j ,b(cid:48) j ] V[aj ,bj ][f (x j,·)]dx (cid:48) (cid:48) j, j,·)] is the usual one-dimensional (Jordan) variation of the where V[aj ,bj ][f (x(cid:48) j − th section of f. Let now Φ(f, I) be the Euclidean norm of the vector (Φ1(f, I), . . . , ΦN (f, I)), namely Φ(f, I) := Φ2 j (f, I)  1 2 ,  N(cid:88) j=1 4 where Φ(f, I) = +∞ if Φj(f, I) = +∞ for some j = 1, . . . , N. Then the variation of f on I ⊂ RN is defined as m(cid:88) VI [f ] := sup Φ(f, Jk), k=1 where the supremum is taken over all the finite families of N−dimensional in- tervals {J1, . . . , Jm} which form partitions of I. Passing to the supremum over all the intervals I ⊂ RN, we obtain the variation of f over the whole RN, i.e., V [f ] := sup I⊂RN VI [f ]. It is well known that, for every f ∈ BV (RN ), ∇f exists a.e. in RN and ∂f L1(RN ), for every j = 1, . . . , N (see e.g. [40, 48]). ∂xj ∈ We now recall the notion of absolute continuity in the sense of Tonelli. sense of Tonelli (f ∈ ACloc(RN )) if, for every interval I =(cid:81)N Definition 2 A function f : RN → R is locally absolutely continuous in the i=1[ai, bi] and for j,·) : [aj, bj] → R is absolutely every j = 1, 2, . . . , N, the j−th section of f, f (x(cid:48) continuous for almost every x(cid:48) j ∈ [a(cid:48) j, b(cid:48) j]. It is a well known result that, if f ∈ BV (RN ) ∩ ACloc(RN ), then (cid:90) V [f ] = RN ∇f (x) dx (see [40, 48, 35, 12]), that is, an integral representation for the variation of f holds. We will therefore denote by AC(RN ) the space of all the functions f ∈ BV (RN ) ∩ ACloc(RN ) (absolutely continuous functions). We will study approximation results for the multivariate generalized sampling series, namely a family of discrete operators defined as (Swf )(t) := f χ(wt − k), t ∈ RN , w > 0 : (cid:18) k (cid:19) w (cid:88) k∈ZN such operators are the multidimensional version of the generalized sampling series (see, e.g., [21, 22, 17]). Here χ is a kernel, that is, a function χ : RN → R that satisfies the following assumptions: 5 (χ1) χ ∈ L1(RN ) is such that(cid:80) k∈ZN χ(u − k) = 1, for every u ∈ RN; (cid:80) k∈ZN χ(u − k) < +∞, where the convergence of the (χ2) Aχ := supu∈RN series is uniform on the compact sets of RN. The above assumptions are quite standard working with discrete families of operators: see e.g., [21]. We point out that the operators (Swf )w>0 are well- defined, for example, for every f ∈ BV (RN ): indeed in this case f is bounded and since f (t) ≤ M, for some M > 0 and for every t ∈ RN, we have (Swf )(t) ≤ M by (χ2). χ(wt − k) ≤ M Aχ, In particular, in the present paper we will study the convergence in variation for the multivariate generalized sampling series with product kernels of averaged type, that is kernels of the form where ¯χi,m(t) := 1 m ¯χm(t) := ¯χi,m(ti) (1) χi(t + v) dv, for some m ∈ N, and χi : R −→ R is a (one-dimensional) kernel for every i = 1, . . . , N (i.e., satisfying (χ1) and (χ2) with N = 1). Notice that it is easy to see that ¯χm is a kernel itself and moreover, by the Fubini-Tonelli theorem, (cid:88) k∈ZN N(cid:89) (cid:90) m i=1 2 − m 2 (cid:107) ¯χi,m(cid:107)L1(R) = χi(t + v) dv χi(t + v) dt dv = (cid:107)χi(cid:107)L1(R) (2) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt (cid:90) R ≤ 1 m 2 2 m − m (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 (cid:90) m (cid:90) (cid:90) m (cid:107) ¯χm(cid:107)1 ≤ N(cid:89) − m R 2 2 (cid:107)χi(cid:107)L1(R). and, by this, We point out that is it easy to give examples of product kernels of averaged type (see Section 6) to which our results can be applied. i=1 6 The corresponding multivariate generalized sampling series (associated to the averaged product kernel ¯χm) will be denoted as ( ¯Sm w f )(t) := ¯χm(wt − k), t ∈ RN , w > 0. It is easy to see that ¯χm is differentiable and, taking into account that, (cid:18) k (cid:19) w f (cid:88) k∈ZN (cid:104) (cid:16) obviously, ∂ ¯χm ∂tj (t) = 1 m (cid:17)(cid:105) (cid:16) (cid:17) − χj (cid:16) (cid:104) ¯χi,m(ti) χj tj + m 2 tj − m 2 , t ∈ RN , it is possible to write, for every j = 1, . . . , N, t ∈ RN, w > 0, ∂ ¯Sm wtj − kj + w f ∂tj ¯χi,m(wti−ki) (cid:88) (t) = w m χj w f (cid:19)(cid:89) m 2 k∈ZN i(cid:54)=j (cid:12)(cid:12)(cid:12)(cid:12) ∂ ¯Sm w f ∂tj (cid:12)(cid:12)(cid:12)(cid:12) ≤ w m (t) M Notice that, since f is bounded, ∂ ¯Sm w f ∂tj (t) exists for every t ∈ RN since (cid:89) ¯χi,m(wti − ki)(cid:104)(cid:12)(cid:12)(cid:12)χj (cid:16) wtj − kj + m 2 (cid:89) (cid:18) k i(cid:54)=j (cid:88) N(cid:89) i=1 k∈ZN i(cid:54)=j ≤ 2w m M Aχi, (cid:17) − χj (cid:12)(cid:12)(cid:12)χj (cid:17)(cid:12)(cid:12)(cid:12) + (cid:16) (cid:17)(cid:105) . wtj − kj − m 2 (3) (cid:16) wtj − kj − m 2 (cid:17)(cid:12)(cid:12)(cid:12)(cid:105) (4) again by (χ2) for each one-dimensional kernel χi. One of the main ideas of this paper is to establish a relation between the gradient of the multivariate generalized sampling series of a function f and a family of Kantorovich-type discrete operators, that we now introduce, acting on the partial derivatives of f. This generalizes the analogous result, in the one-dimensional case, obtained in [6], similarly to what happens between the Bernstein polynomials and their Kantorovich version (see, e.g., [1]). We therefore introduce the following family of multidimensional discrete op- erators of sampling-Kantorovich type: w (cid:88) k∈ZN (cid:90) kj +1 w kj w (cid:18) k1 w f (cid:19)  χ(wt−k), t ∈ RN , w > 0, , . . . , u . . . , kN w du (Kw,jf )(t) := j = 1, . . . , N. Notice that, similarly to the case of (Swf )w>0, it is easy to see that the operators (Kw,jf )w>0, j = 1, . . . , N, are well-defined if, for example f ∈ BV (RN ), since f is in particular bounded. 7 3 Results for the multidimensional generalized sampling series w f )w>0 map BV (RN ) into AC(RN ). We will first prove that the operators ( ¯Sm Such result is a kind of "asymptotic" variation diminishing property: indeed, choosing m ∈ N sufficiently large, we obtain the classical variation diminishing property for the multidimensional generalized sampling series. Proposition 1 Let f ∈ BV (RN ). Then ¯Sm m ∈ N, and w f ∈ AC(RN ), for every w > 0, V [ ¯Sm w f ] ≤ 1 m (cid:107)χi(cid:107)1V [f ]. (5) N(cid:89) i=1 Proof. Let us fix w > 0 and m ∈ N. By (3) and (4), the partial derivatives of w f exist for every t ∈ RN, j = 1 . . . , N, and are bounded: this implies that ¯Sm the sections of the function f are locally absolutely continuous, namely f ∈ [ai, bi] ⊂ RN and let {J1, . . . , Jp} be a partition ACloc(RN ). Let us fix I = N(cid:89) of I, with Jk = i=1 [(k)aj,(k) bj], k = 1, . . . p. For every fixed j = 1, . . . N, N(cid:89) j=1 k = 1, . . . p, similarly to the proof of Proposition 1 of [6], it is possible to show that, since ( ¯Sm j,·) ∈ ACloc(R), a.e. t(cid:48) w f )(t(cid:48) j ∈ [(k)a(cid:48) j,(k) b(cid:48) j], j,·)] (cid:48) V[(k)aj ,(k)bj ][f (t (cid:89) i(cid:54)=j (cid:107) ¯χi,m(cid:107)1(cid:107)χj(cid:107)1 V[(k)aj ,(k)bj ][( ¯Sm w f )(t j,·)] ≤ 1 (cid:48) m and therefore, by (2), Φj( ¯Sm w f, Jk) = (cid:90) V[(k)aj ,(k)bj ][( ¯Sm w f )(t [(k)a(cid:48) j ,(k)b(cid:48) j ] (cid:90) (cid:48) j j,·)] dt (cid:48) N(cid:89) j,·)] dt (cid:48) (cid:48) j (cid:107)χi(cid:107)1 ≤ 1 m = 1 m V[(k)aj ,(k)bj ][f (t [(k)a(cid:48) j ,(k)b(cid:48) j ] i=1 (cid:107)χi(cid:107)1. Φj(f, Jk) N(cid:89) i=1 8 This implies that, for every k = 1, . . . , p,  N(cid:88)  N(cid:88) 1 m j=1 j=1 Φ( ¯Sm w f, Jk) = = [Φj( ¯Sm w f, Jk)]2 1 m2 [Φj(f, Jk)]2  N(cid:88) j=1 2 ≤  1  1 2 N(cid:89) i=1 N(cid:89) i=1 [Φj(f, Jk)]2 (cid:107)χi(cid:107)1 = 1 m (cid:107)χi(cid:107)1Φ(f, Jk). N(cid:89) i=1 (cid:107)χi(cid:107)2 1  1 2 Summing now over k = 1, . . . , p and passing to the supremum over all the possible partitions {J1, . . . , Jp} of I, we conclude that and hence, by the arbitrariness of I ⊂ RN, N(cid:89) i=1 N(cid:89) i=1 (cid:107)χi(cid:107)1VI [f ] (cid:107)χi(cid:107)1V [f ]. VI [ ¯Sm w f ] ≤ 1 m V [ ¯Sm w f ] ≤ 1 m This implies that ¯Sm w f ∈ BV (RN ) and hence ¯Sm w f ∈ AC(RN ). We will now give the result that establishes a relation between the partial derivatives of the multidimensional sampling series and the multidimensional sampling-Kantorovich type operators acting on the partial derivatives of the function. Proposition 2 If f ∈ AC(RN ), then for every t ∈ RN, j = 1, . . . , N, (cid:18) m(cid:88) i=1 (cid:19)(cid:18) j, tj − m − 2(i − 1) (cid:48) t 2w (cid:19) , ∂ ¯Sm w f ∂tj w > 0, m ∈ N. (t) = 1 m Kw,j ∂f ∂tj 9 Proof. By (3), since f is locally absolutely continuous, we have that ∂ ¯Sm w f ∂tj (t) = w m ¯χi,m(wti − ki) wtj − kj + m 2 (cid:104) (cid:16) (cid:18) k(cid:48) χj j w (cid:19) (cid:17) − χj (cid:16) (cid:17)(cid:105) wtj − kj − m 2 (cid:104) (cid:16) , u du + f , 0 ¯χi,m(wti − ki) χj wtj − kj + m 2 i(cid:54)=j (cid:19)(cid:89) (cid:19)(cid:89) (cid:19)(cid:89) i(cid:54)=j = w m − χj w ∂f ∂tj k∈ZN wtj − kj − m 2 0 k∈ZN (cid:88) (cid:88) (cid:16) (cid:88) (cid:88) j w i(cid:54)=j (cid:19)(cid:89) (cid:18) k(cid:48) (cid:17)(cid:105) (cid:18) k(cid:48) (cid:18) k(cid:48) ∂f ∂tj f w (cid:18) k (cid:90) kj (cid:90) kj (cid:90) kj 0 w w (cid:17) (cid:17) (cid:17) . wtj − kj + m 2 wtj − kj − m 2 (cid:17) (cid:16) wtj −(cid:101)kj − m (cid:17) 2 wtj − kj − m 2 (cid:17) = , 0 , u j w j w w m k∈ZN du + f (cid:19) (cid:19) (cid:18) k(cid:48) (cid:18) k(cid:48) − w m ¯χi,m(wti − ki)χj (cid:16) (cid:16) Now putting in the first series(cid:101)kj = kj − m and(cid:101)ki = ki for every i (cid:54)= j, there ¯χi,m(wti −(cid:101)ki)χj (cid:16) ¯χi,m(wti − ki)χj ∂ ¯Sm w f ∂tj ∂f ∂tj (cid:33) holds du + f (t) = k∈ZN w m i(cid:54)=j i(cid:54)=j j w j w j w , u , u , 0 , 0 0 w w ¯χi,m(wti − ki)χj 0 0 k∈ZN , 0 , u j w j w j w du + f du + f ∂f ∂tj ∂f ∂tj (cid:88) (cid:101)k∈ZN (cid:88) (cid:88) (cid:88) (cid:26)(cid:18) m(cid:88) (cid:90) (cid:101)kj +m (cid:90) kj (cid:90) kj +m  m(cid:88) (cid:18) (cid:33)(cid:89) (cid:32)(cid:101)k(cid:48) (cid:19)(cid:89) (cid:18) k(cid:48) (cid:89) (cid:19) (cid:32)(cid:101)k(cid:48) (cid:19) (cid:18) k(cid:48) (cid:19) (cid:18) k(cid:48)  ∂f (cid:18) k(cid:48) (cid:90) kj +i (cid:89) (cid:19)(cid:16) (cid:18) (cid:17) ∂tj j, tj − m (cid:19)(cid:18) (cid:48) 2w j, tj − m − 2(i − 1) ¯χi,m(wti − ki)χj + . . . + ∂f ∂tj ∂f ∂tj ∂f ∂tj (cid:19) kj +i−1 k∈ZN k∈ZN Kw,j Kw,j Kw,j i(cid:54)=j i(cid:54)=j i(cid:54)=j j w j w (cid:48) t du du , u , u i=1 kj w t . w w w 2w ∂f ∂tj i=1 − w m = = = = w m w m 1 m 1 m (cid:16) wtj − kj − m 2 (cid:16) ¯χi,m(wti − ki)χj (cid:19)(cid:18) wtj − kj − m 2 j, tj − m − 2(m − 1) (cid:48) t 2w (cid:17) (cid:19)(cid:27) 10 4 Convergence in Lp for the multidimensional sampling- Kantorovich type operators We now study the problem of the convergence in Lp for the multidimensional sampling-Kantorovich type operators that we introduced. Such result will be also fundamental in order to prove the convergence in variation for ( ¯Swf )w>0. We recall that convergence in Lp holds for the multidimensional general- ized sampling series (Swf )w>0 (see [17]) assuming that f belongs to a suit- able subspace of Lp(RN ), namely Λp. Since the definition of our operators (Kw,jf )w>0 is very close to that one of (Swf )w>0, it is natural to expect that the Lp−convergence holds within the same subspace, that is what we will now prove. Let us recall the definition of Λp (see [17]). Let us consider an admissible partition over the i−th axis, i.e., Σi := (xi,ji)ji∈Z such that 0 < ∆ := min i=1,...,N ji∈Z(xi,ji − xi,ji−1) ≤ max inf i=1,...,N (xi,ji − xi,ji−1) =: ∆ < +∞. sup ji∈Z We say that Σ = (xj)j∈ZN ⊂ RN, xj = (x1,j1, . . . , xN,jN ), j = (j1, . . . , jN ) ∈ ZN, is an admissible sequence if it is the cartesian product of admissible parti- tions Σi = (xi,ji)ji∈Z. For a fixed admissible sequence Σ, the lp(Σ)− norm of a function f : RN → R is defined as where Qj =(cid:81)N (cid:107)f(cid:107)lp(Σ) := (cid:88)  1 i=1[xi,ji−1 − xi,ji[ and ∆j :=(cid:81)N f (x)p∆j p sup x∈Qj j∈ZN , 1 ≤ p < +∞ i=1(xi,ji − xi,ji−1) is the volume of Qj. With such notations the subspace Λp is defined as Λp := {f ∈ M (RN ) : (cid:107)f(cid:107)lp(Σ) < +∞, for every admissible sequence Σ}. In [17] it is proved that Λp is a proper linear subspace of Lp(RN ), together with other properties concerning such space. In particular, we recall the following important result of convergence in Lp for the τ−modulus of smoothness (see [34, 33, 42]) τr(f ; δ, M (RN ))p := (cid:107)ωr(f ;·, δ, M (RN ))(cid:107)p, 11 where ωr(f ; x, δ, M (RN )) := sup ∆r hf (t) :=(cid:80)r (cid:40) hf (t) : t, t + hr ∈ N(cid:89) (cid:32) (cid:33) i=1 (cid:20) xi − δr 2 , xi + δr 2 (cid:21)(cid:41) , j=0(−1)r+j r δ > 0, and ∆r order r at t ∈ RN with increment h ∈ RN. Proposition 3 (Prop. 7 of [17]) If f ∈ Λp ∩ Rloc(RN ), 1 ≤ p < +∞, r ∈ N, then f (t + jh) denotes the differences of j τr(f ; δ, M (RN ))p = 0. lim δ→0+ We point out that, of course, Λp is a non trivial subspace since it contains, for example, all the functions in M (RN ) with compact support. We are now ready to state the result of convergence in Lp for the multidi- mensional sampling-Kantorovich type operators. Let us now assume that χ is a kernel with compact support, i.e., χ(t) = 0 if t (cid:54)∈ [−T, T ]N, T > 0, not necessarily of product type, namely χ satisfies the assumption: (χ) there exists T > 0 such that χ(t) = 0 for t (cid:54)∈ [−T, T ]N and(cid:80) k∈ZN χ(u − k) = 1, for every u ∈ RN. We point out that, since χ has compact support, obviously χ ∈ L1(RN ) (and therefore (χ1) holds) and satisfies the condition (χ2) since χ is bounded and the series reduces to a finite sum. Proposition 4 Let f ∈ Λp ∩ Rloc(RN ), 1 ≤ p < +∞. Then, for every j = 1, . . . , N, Proof. By assumption (χ) there holds, for w > 0, j = 1, . . . , N and t ∈ RN, (Kw,jf )(t) − f (t) = du χ(wt − k) − f (t) w w f lim k∈ZN w→+∞(cid:107)Kw,jf − f(cid:107)p = 0. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:90) kj +1 (cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) w (cid:90) kj +1 (cid:88) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)w (cid:90) kj +1 ≤ (cid:88) (cid:18) k1 (cid:18) k1 (cid:18) k1 k∈ZN k∈ZN kj w kj w kj w w w f = w , . . . , u . . . , w f , . . . , u . . . , , . . . , u . . . , w (cid:19) kN w kN w (cid:19) kN w (cid:19) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)  χ(wt − 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)χ(wt − k). du − f (t) du − f (t) 12 du − f (t) f ; t, 2T w , M (RN ) , (cid:19) Since χ has support contained in [−T, T ]N, χ(wt − k) = 0 if wt − k > T , and therefore the series reduces to a finite sum over the indexes k ∈ ZN such that w , for every i = 1, . . . , N: for such k (cid:12)(cid:12) ≤ T w w kj w w f , . . . , u . . . , we have that (cid:18) k1 (cid:90) kj +1 (wt − k) ∈ [−T, T ]N, namely(cid:12)(cid:12)ti − ki (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)w (cid:19) (Kw,jf )(t) − f (t) ≤ (cid:88) (cid:18) (cid:18) (wt−k)∈[−T,T ]N ≤ ω1 kN w 2T w f ; t, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)w (cid:18) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ ω1 (cid:18) k1 (cid:90) kj +1 (cid:19) (cid:88) (cid:19) kj w w f w (wt−k)∈[−T,T ]N by the definition of the modulus of smoothness. Therefore (cid:19) kN w (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)χ(wt − k) du − f (t) , . . . , u, . . . , , M (RN ) χ(wt − k) ≤ Aχω1 f ; t, 2T w , M (RN ) , by (χ2). Passing to the Lp−norm we obtain (cid:107)Kw,jf − f(cid:107)p ≤ Aχτ1 and the thesis follows by Proposition 3. (cid:18) f ; (cid:19) , p , M (RN ) 2T w 5 Convergence in variation for the multidimen- sional generalized sampling series We are now ready to prove the main result of the paper, that is, the convergence in variation for the multidimensional generalized sampling series with product kernels of averaged type. Since we will use results of the previous section, we assume here that χ is a kernel with compact support, that is, χ : RN → R satisfies assumption (χ). Theorem 1 Let f ∈ AC(RN ) be such that ∂f j = 1, . . . , N. Then, for every m ∈ N, ∈ Λ1 ∩ Rloc(RN ), for every ∂xj w→+∞ V [ ¯Sm lim w f − f ] = 0. 13 Proof. Since f ∈ AC(RN ), by Proposition 1, ¯Sm and m ∈ N: then w f ∈ AC(RN ), for every w > 0 (cid:90) ∇ ¯Sm w f (t) − ∇f (t) dt. (cid:12)(cid:12)(cid:12)(cid:12) dt (t) (cid:48) j, tj − m − 2(i − 1) j, tj − m − 2(i − 1) 2w (cid:48) t t 2w (cid:19)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt − ∂f ∂tj (t) (cid:19) (cid:19) Now, by Proposition 2, V [ ¯Sm w f − f ] ≤ = RN j=1 i=1 RN i=1 1 m RN j=1 Kj,w RN m Kj,w V [ ¯Sm ∂f ∂tj w f − f ] = N(cid:88) (cid:90) ( ¯Sm ∂tj (cid:90) N(cid:88) ≤ N(cid:88) − m(cid:88) (cid:90) N(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)Kw,j ≤ N(cid:88) N(cid:88) m(cid:88) N(cid:88) (cid:12)(cid:12)(cid:12)(cid:12) ∂ w f )(t) − ∂f (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 (cid:19)(cid:18) (cid:18) m(cid:88) ∂tj (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) m(cid:88) (cid:18) (cid:19)(cid:18) (cid:90) (cid:18) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) m(cid:88) (cid:18) (cid:13)(cid:13)(cid:13)(cid:13)1 (cid:12)(cid:12)(cid:12)(cid:12) ∂f (cid:18) (cid:90) m(cid:88) N(cid:88) j, tj − m − 2(i − 1) − ∂f ∂tj ∂f ∂tj ∂f ∂tj ∂f ∂tj ∂f ∂tj RN ∂tj 1 m + 1 m RN i=1 j=1 i=1 (cid:48) t 2w + j=1 j=1 i=1 j=1 + := Lj + 1 m j=1 j=1 i=1 I i j. j, tj − m − 2(i − 1) (cid:48) t 2w j, tj − m − 2(i − 1) (cid:48) t 2w (cid:19) (cid:19) − ∂f ∂tj − ∂f ∂tj (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dt (cid:12)(cid:12)(cid:12)(cid:12) dt (t) (t) Now, Lj → 0 as w → +∞ by Proposition 4, while I i j → 0 as w → +∞, for every i = 1, . . . , m, j = 1, . . . , N, by the continuity in L1 of the translation operator. Therefore the theorem is proved. 6 Examples and applications to smoothing for digital image processing In this section, we provide some basic examples of kernels for which the above results hold, and we discuss some applications of the variation diminishing-type property (5) to smoothing in digital image processing. 14 Figure 1: On the left: the Fejér kernel F . On the right: the averaged Fejér kernel with m = 1. As stated in Section 2, in this paper we consider product kernels ¯χm of averaged type. As a first example, we can consider the multivariate product kernel of the averaged type generated by the well-known Fejér kernel (see, e.g., Fig. 1 left and [27]), defined by: F (x) := sinc2(x/2), x ∈ R, where the sinc-function (see, e.g., [26, 7]) is of the form: x (cid:54)= 0, x = 0. sinc(x) := sin(πx)/πx, 1, 1 2 (cid:40) Now, the one-dimensional averaged Fejér kernel (see, e.g., Fig. 1, right) is the following: (cid:90) m/2 (cid:18) t + v (cid:19) 2 ¯Fm(t) := 1 sinc2 2 m −m/2 dv, t ∈ R, m ∈ N, and the corresponding multivariate version (see, e.g., Fig. 2 for the case N = 2) is: N(cid:89) Fm(t) := ¯Fm(ti), t ∈ RN . i=1 It is well-known that the Fejér kernel has unbounded support and satisfies as- sumptions (χ1) and (χ2) (see e.g., [31, 8]), then for the multivariate generalized sampling series based upon Fm holds the variation diminishing type property established in Proposition 1, when multivariate signals of bounded variation are considered. 15 Figure 2: The bivariate averaged Fejér kernel with m = 1. Moreover, also the relation established in Proposition 2 holds, where the j-th first order partial derivatives of the generalized sampling series of a given absolutely continuous function f is related with Kw,j , i.e., the corresponding sampling series of the Kantorovich type of the j-th first order partial derivatives of f. ∂f ∂tj However, the convergence results proved in Section 4 can not be applied to the generalized sampling series based upon Fm since the kernels Fm do not have compact support, hence assumption (χ) is not satisfied. Other examples of one-dimensional kernels with unbounded support that can be used to define product averaged type kernels can be found, e.g., in [38, 28, 29, 30]. In order to recall examples of kernels of one variable such that the corre- sponding multivariate averaged type versions also satisfy assumption (χ), we recall the definition of the well-known central B-spline of order n ∈ N (see, e.g., Fig. 3 and [46, 25, 2, 3, 20]), defined by: n(cid:88) i=0 (cid:18)n (cid:19)(cid:16) n i 2 (cid:17)n−1 + Mn(x) := 1 (n − 1)! (−1)i + x − i x ∈ R, , (6) where (x)+ := max{x, 0} denotes "the positive part" of x ∈ R (see e.g., [41, 16 Figure 3: On the left: the central B-spline M2. On the right: the central B-spline M3 which corresponds to ¯M2,1. 32]). The functions Mn(x) are non-negative, continuous with compact support contained in [−n/2, n/2], and satisfy conditions (χ1) and (χ2). Now, let us denote by ¯Mn,m(t) := m−1 (cid:90) m/2 −m/2 Mn(t + v) dv, t ∈ R, the averaged B-spline kernel of order n ∈ N. Recalling the following well-known property: M(cid:48) n(t) = Mn−1(t + 1/2) − Mn−1(t − 1/2), t ∈ R, (n ≥ 2) for m = 1, we have: ¯M(cid:48) n,1(t) = Mn(t + 1/2) − Mn(t − 1/2) = M(cid:48) n+1(t), t ∈ R, (n ≥ 1), i.e., ¯Mn,1(t) = Mn+1(t) + k, k ∈ R. Now, since ¯Mn,1 belongs to L1(R) (see [6]), we must have k = 0 and therefore we conclude that ¯Mn,1(t) = Mn+1(t), t ∈ R, for every n ∈ N, namely, the averaged kernel with m = 1 generated by a central B-spline of order n is a B-spline itself of order n + 1. In view of the above remark, we can explicitly state that the multivariate averaged type kernel with m = 1 and generated by Mn (see, e.g., Fig. 4 for the case n = 2 in two space dimension) is the following: Mn 1 (t) := ¯Mn,1(ti) = Mn+1(ti), t ∈ RN . N(cid:89) N(cid:89) i=1 i=1 17 Figure 4: The bivariate averaged B-spline kernel of order 2 with m = 1 which coincides with M2 i=1 M3(ti). 1(t) =(cid:81)2 In practice, in the latter case the multivariate generalized sampling operators based upon averaged B-spline Mn with m = 1 coincide with the usual general- ized sampling series based upon the multivariate central B-spline of order n + 1. Generally speaking, we can also define the multivariate average central B- spline kernels as follows: N(cid:89) i=1 (cid:90) m/2 −m/2 N(cid:89) i=1 Mn m(t) := ¯Mn,m(ti) = m−1 Mn(ti + v) dv, t ∈ RN . Since Mn m has compact support it satisfies also assumption (χ); thus the above sampling series fulfills the results of both Section 3 and Section 5. Further, for the latter examples of kernels also Proposition 4 and Theorem 1 can be applied. Now, at the end of this section we consider some applications to smoothing in digital processing. For basic facts concerning this numerical tool for imaging, see e.g., [19, 39, 43]. It is well-known that any static gray scale image is a bivariate signal with 18 compact support; then it can be (naturally) modeled as follows: m(cid:88) m(cid:88) IA(x, y) := aij · 1ij(x, y) ((x, y) ∈ R2), i=1 j=1 for every image (matrix) A = (aij)ij, i, j = 1, 2, ..., m, where 1ij(x, y), i, j = 1, 2, ..., m, is the characteristic function of the sets (i − 1, i] × (j − 1, j] (i.e. 1ij(x, y) = 1, for (x, y) ∈ (i − 1, i] × (j − 1, j] and 1ij(x, y) = 0 otherwise). Note that the above function IA(x, y) is defined in such a way that to every pixel (i, j) it is associated the corresponding gray level aij. Moreover, by the above representation IA of the image A it turns out that IA ∈ BV (R2), hence one can consider approximations of A by means of the bivariate generalized sampling series based upon the averaged type kernels. The main advantage that can be achieved by the above procedure is expressed by w IA have no-bigger variation than IA, for sufficiently large Proposition 1: ¯Sm m ∈ N, i.e., the operators provide an approximated version of the original image A, but regularized. In this sense, the generalized sampling series can be used for smoothing of images. Clearly, in order to visualize an approximation (new image) of the original w IA, we need to sample the operators, for w > 0, image A by means of ¯Sm with a fixed sampling rate concordant with the dimension of A. Obviously, the sampling rate is chosen arbitrarily hence one can also consider different (high) sampling rates. The effect of the proposed procedure can be strongly noticed at the edges of the figures, where the jumps of gray levels are reduced with respect to the corresponding ones in the original image. Now we can give the following practical examples of image reconstruction in order to show the smoothing capabilities of the above operators. An optimized version of the above described algorithm for image reconstruction and smooth- ing can be implemented by means of the MATLAB programming language, following the indications outlined in [14, 15] in case of the so-called sampling Kantorovich algorithm for image enhancement. For the numerical experiments, we consider the well-known images of Lena and Baboon with 150 × 150 pixel resolution (see Fig. 5). First of all, we reconstruct the original images in Fig. 5 by using the bivariate averaged Fejér kernel with m = 4. According to Proposition 1, and observing 19 Figure 5: On the left: Lena. On the right: Baboon. Figure 6: On the left: the reconstruction of Lena (of 150 × 150 pixel) by means of the operator ¯Sm w , w = 4, based upon the bivariate averaged Fejér kernel with m = 4. On the right: the reconstruction of Baboon made as for Lena. 20 Figure 7: On the left: the reconstruction of Lena (of 150 × 150 pixel) by means of the operator ¯Sm w , w = 10, based upon the bivariate averaged Fejér kernel with m = 4. On the right: the reconstruction of Baboon made as for Lena. that (cid:107)F(cid:107)1 = 1, we have: V [ ¯S4 wf ] ≤ 1 4 V [f ], (7) i.e., the total variation of any reconstructed images is, at least, 4 times smaller than that of the original ones, producing the smoothing effect. In Fig. 6 we have the reconstruction of Lena and Baboon (of 150 × 150 pixel resolution) by w , w = 4, based upon the bivariate averaged Fejér means of the operator ¯Sm kernel with m = 4. By detailed analysis of the edges (especially in case of Lena, at the contours of the hat) it is possible to observe the smoothing of the analyzed images. In Fig. 7 we have the reconstruction of Lena and Baboon (of 150 × 150 pixel resolution) by means of the operator ¯Sm w , w = 10, based upon the bivariate averaged Fejér kernel with m = 4. The main differences that can be observed between the images in Fig. 6 and Fig. 7 are that, for big values of w the images are closer to the original and consequently the edges tend to be more clear. Finally, in Fig. 8 we have the reconstruction of Lena and Baboon (of 150 × 150 pixel resolution) by means of the operator ¯Sm w , w = 6, based upon the bivariate central B-spline of order 3 with m = 4. Note that, also for the latter case one can state a inequality as that given in (7) since also (cid:107)M3(cid:107)1 = 1. Acknowledgments The authors are members of the Gruppo Nazionale per l'Analisi Matematica, la Probabilitá e le loro Applicazioni (GNAMPA) of the Istituto Nazionale di Alta 21 Figure 8: On the left: the reconstruction of Lena (of 150 × 150 pixel) by means of the operator ¯Sm w , w = 6, based upon the bivariate averaged central B-spline of order 3 with m = 4. On the right: the reconstruction of Baboon made as for Lena. Matematica (INdAM). The authors are partially supported by the "Department of Mathematics and Computer Science" of the University of Perugia (Italy). The first and the third author have been partially supported within the project "Metodi di teoria degli operatori e di Analisi Reale per problemi di approssimazione ed applicazioni", funded by the 2017 basic research fund of the University of Perugia. Finally, the second author of the paper has been partially supported within a 2018 GNAMPA-INdAM Project: "Dinamiche non autonome, analisi reale e appli- cazioni". References [1] O. Agratini, An approximation process of Kantorovich type, Math. Notes, Miskolc, 2 (1) (2001), 3-10. [2] G. Allasia, R. Cavoretto, A. De Rossi, A class of spline functions for landmark-based image registration, Math. Methods Appl. Sci., 35 (2012), 923 -- 934. [3] G. Allasia, R. Cavoretto, A. De Rossi, Lobachevsky spline functions and interpolation to scattered data, Comput. Appl. Math., 32 (2013), 71 -- 87. 22 [4] L. Ambrosio, S. Di Marino, Equivalent definitions of BV space and of total variation on metric measure spaces, Journal of Functional Analysis, 266 (7) (2014), 4150 -- 4188. [5] L. Angeloni, Approximation results with respect to multidimensional ϕ- variation for nonlinear integral operators, Z. Anal. Anwendungen, 32 (1) (2013), 103 -- 128. [6] L. Angeloni, D. Costarelli, G. Vinti, A characterization of the convergence in variation for the generalized sampling series, Ann. Acad. Sci. Fenn. Math., 43 (2018), 755-767. [7] L. Angeloni, D. Costarelli, G. Vinti, A characterization of the absolute con- tinuity in terms of convergence in variation for the sampling Kantorovich operators, Mediterr. J. Math., 16 (2019), 44. DOI:10.1007/s00009-019- 1315-0 [8] L. Angeloni, D. Costarelli, G. Vinti, Quantitative estimates for sampling type operators with respect to the Jordan variation, submitted (2019). [9] L. Angeloni, G. Vinti, Convergence and rate of approximation for linear integral operators in BV ϕ−spaces in multidimensional setting, J. Math. Anal. Appl., 349 (2009), 317 -- 334. [10] L. Angeloni, G. Vinti, Approximation in variation by homothetic operators in multidimensional setting, Differential Integral Equation, 26 (5-6) (2013), 655 -- 674. [11] L. Angeloni, G. Vinti, Convergence and rate of approximation in BV ϕ(RN + ) for a class of Mellin integral operators, Atti della Accademia Nazionale dei Lincei, Classe di Scienze Fisiche, Matematiche e Naturali, Rendiconti Lincei Matematica e Applicazioni, 25 (3) (2014), 217 -- 232. [12] L. Angeloni, G. Vinti, Convergence in variation and a characterization of the absolute continuity, Integral Transforms Spec. Funct., 26 (10) (2015), 829 -- 844. [13] J. Appell, J. Banaś, N. Merentes, Bounded variation and around, 17 De Gruyter Series in Nonlinear Analysis and Applications, De Gruyter, Berlin, Germany, (2014). 23 [14] F. Asdrubali, G. Baldinelli, F. Bianchi, D. Costarelli, A. Rotili, M. Seracini, G. Vinti, Detection of thermal bridges from thermographic images by means of image processing approximation algorithms, Applied Mathematics and Computation, 317 (2018) 160 -- 171. [15] F. Asdrubali, G. Baldinelli, F. Bianchi, D. Costarelli, L. Evangelisti, A. Rotili, M. Seracini, G. Vinti, A model for the improvement of thermal bridges quantitative assessment by infrared thermography, Applied Energy, 211 (2018) 854 -- 864. [16] C. Bardaro, P.L. Butzer, R.L. Stens, G. Vinti, Prediction by samples from the past with error estimates covering discontinuous signals, IEEE Trans. Inform. Theory, 56 (1) (2010), 614 -- 633. [17] C. Bardaro, I. Mantellini, R. Stens, J. Vautz, G. Vinti, Generalized sam- pling approximation for multivariate discontinuous signals and application to image processing, New Perspectives on Approximation and Sampling Theory-Festschrift in honor of Paul Butzer's 85th birthday, Birkhauser (2014), 87-114. [18] C. Bardaro, G. Vinti, A general approach to the convergence theorems of generalized sampling series, Appl. Anal., 64 (1997), 203 -- 217. [19] S. Bi, X. Han, Y. Yu, An L1 image transform for edge-preserving smoothing and scene-level intrinsic decomposition, ACM Transactions on Graphics, 34 (4) (2015) No. 78. [20] M. Bozzini, L. Lenarduzzi, M. Rossini, R. Schaback, Interpolation with variably scaled kernels, IMA Journal of Numerical Analysis, 35 (1) (2015) 199 -- 219. [21] P.L. Butzer, A. Fisher, R.L. Stens, Generalized sampling approximation of multivariate signals: theory and applications, Note di Matematica, 10 (1) (1990), 173 -- 191. [22] P.L. Butzer, A. Fisher, R.L. Stens, Generalized sampling approximation of multivariate signals: general theory, Atti Sem. Mat. Fis. Univ. Modena, 41 (1) (1993), 17 -- 37. 24 [23] P.L. Butzer, R.J. Nessel, Fourier Analysis and Approximation I, Academic Press, New York-London, 1971. [24] P.L. Butzer, W. Splettstösser, R.L. Stens, The sampling theorem and linear prediction, Jahresberichte Dt. Math.-Verein, 90 (1988), 1 -- 70. [25] M. Charina, C. Conti, K. Jetter, G. Zimmermann, Scalar multivariate sub- division schemes and box splines, Computer Aided Geometric Design, 28 (5) (2011), 285-306. [26] L. Coroianu, S.G. Gal, Lp- approximation by truncated max-product sam- pling operators of Kantorovich-type based on Fejer kernel, Journal of Inte- gral Equations and Applications, 29 (2) (2017) 349 -- 364. [27] D. Costarelli, A.M. Minotti, G. Vinti, Approximation of discontinuous sig- nals by sampling Kantorovich series, Journal of Mathematical Analysis and Applications, 450 (2) (2017), 1083 -- 1103. [28] D. Costarelli, G. Vinti, Approximation by max-product neural network oper- ators of Kantorovich type, Results in Mathematics, 69 (3) (2016), 505 -- 519. [29] D. Costarelli, G. Vinti, Max-product neural network and quasi-interpolation operators activated by sigmoidal functions, J. Approx. Theory, 209 (2016), 1 -- 22. [30] D. Costarelli, G. Vinti, Pointwise and uniform approximation by multivari- ate neural network operators of the max-product type, Neural Networks, 81 (2016), 81 -- 90. [31] D. Costarelli, G. Vinti, An inverse result of approximation by sampling Kantorovich series, Proceedings of the Edinburgh Mathematical Society, 62 (1) (2019), 265-280. [32] L. D'Amore, R. Campagna, A. Galletti, L. Marcellino, A. Murli, A smooth- ing spline that approximates Laplace transform functions only known on measurements on the real axis, Inverse Problems 28 (2) (2012), 025007. [33] E.P. Dolženko, E.A. Sevast'janov, Approximations of functions in the Haus- dorff metric by means of piecewise monotone (in particular, rational) func- tions (Russian), Mat. Sb. (N.S.), 101 (143) no. 4 (1976), 508 -- 541. 25 [34] P.P. Korovkin, Attempt at an axiomatic construction in certain problems in the theory of approximation of functions of one variable (Russian), Kalinin. Gos. Ped. Inst. Učen. Zap., 69 (1969), 91 -- 109. [35] E. Giusti, Minimal Surfaces and Functions of Bounded Variation, Mono- graphs in Mathematics, vol. 80, Birkhäuser Verlag, Basel, 1984. [36] Y. Gousseau, J.M. Morel, Are Natural Images of Bounded Variation?, SIAM J. Math. Anal., 33 (3) (2001), 634 -- 648. [37] I. Mantellini, G. Vinti, Approximation results for nonlinear integral opera- tors in modular spaces and applications, Ann. Polon. Math., 81 (1) (2003), 55 -- 71. [38] O. Orlova, G. Tamberg, On approximation properties of generalized Kantorovich-type sampling operators, Journal of Approximation Theory, 201 (2016), 73 -- 86. [39] J. Pan, X. Yang, H. Cai, B. Mu, Image noise smoothing using a modified Kalman filter, Neurocomputing, 173 (3) (2016), 1625 -- 1629 [40] T. Radó, Length and Area, American Mathematical Society Colloquium Publications, vol. 30, American Mathematical Society, New York, 1948. [41] L. Romani, M.A. Sabin, The conversion matrix between uniform B-spline and Bézier representations, Computer aided geometric design, 21 (6) (2004), 549 -- 560. [42] B. Sendov, V.A. Popov, The Averaged Moduli of Smoothness, Wiley, Chich- ester, U.K., 1988. [43] M. Sonka, V. Hlavac, R. Boyle, Image processing, analysis, and machine vision, Cengage Learning, Australia - Brazil - Japan, (2014). [44] D. Strong, T. Chan, Edge-preserving and scale-dependent properties of total variation regularization, Inverse Problems, 19 (6) (2003). [45] L. Tonelli, Su alcuni concetti dell'analisi moderna, Ann. Scuola Norm. Su- per. Pisa. (2), 11 (1942), 107 -- 118. [46] M. Unser, Ten good reasons for using spline wavelets, Wavelets Applications in Signal and Image Processing, 3169 (5) (1997), 422 -- 431. 26 [47] T. Valkonen, K. Bredies, F. Knoll, Total Generalized Variation in Diffusion Tensor Imaging, SIAM J. Imaging Sci., 6 (1) (2013), 487 -- 525. [48] C. Vinti, Perimetro -- variazione, Ann. Scuola Norm. Sup. Pisa (3), 18 (1964), 201 -- 231. 27
1807.05599
3
1807
2018-12-07T20:37:11
Inequalities for $L^p$-norms that sharpen the triangle inequality and complement Hanner's Inequality
[ "math.FA", "math.CA" ]
In 2006 Carbery raised a question about an improvement on the na\"ive norm inequality $\|f+g\|_p^p \leq 2^{p-1}(\|f\|_p^p + \|g\|_p^p)$ for two functions in $L^p$ of any measure space. When $f=g$ this is an equality, but when the supports of $f$ and $g$ are disjoint the factor $2^{p-1}$ is not needed. Carbery's question concerns a proposed interpolation between the two situations for $p>2$. The interpolation parameter measuring the overlap is $\|fg\|_{p/2}$. We prove an inequality of this type that is stronger than the one Carbery proposed. Moreover, our stronger inequalities are valid for all $p$.
math.FA
math
INEQUALITIES FOR Lp-NORMS THAT SHARPEN THE TRIANGLE INEQUALITY AND COMPLEMENT HANNER'S INEQUALITY ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB p + (cid:107)g(cid:107)p p ≤ 2p−1((cid:107)f(cid:107)p Abstract. In 2006 Carbery raised a question about an improvement on the naıve norm inequality (cid:107)f + g(cid:107)p p) for two functions in Lp of any measure space. When f = g this is an equality, but when the supports of f and g are disjoint the factor 2p−1 is not needed. Carbery's question concerns a proposed interpolation between the two situations for p > 2. The interpolation parameter measuring the overlap is (cid:107)f g(cid:107)p/2. We prove an inequality of this type that is stronger than the one Carbery proposed. Moreover, our stronger inequalities are valid for all p. December 6, 2018 8 1 0 2 c e D 7 ] . A F h t a m [ 3 v 9 9 5 5 0 . 7 0 8 1 : v i X r a 1. Introduction and main theorem Since zp is a convex function of z for p ≥ 1, for any measure space, the Lp unit ball, Bp := {f : (cid:82) fp ≤ 1}, is convex. One way to express this is with Minkowski's triangle inequality (cid:107)f + g(cid:107)p ≤ (cid:107)f(cid:107)p + (cid:107)g(cid:107)p. Another is the inequality p + (cid:107)g(cid:107)p p ≤ 2p−1(cid:0)(cid:107)f(cid:107)p (cid:107)f + g(cid:107)p (cid:1) , p (1.1) valid for any functions f and g on any measure space. There is equality if and only if f = g and, in Theorem 1.1, we improve (1.1) substantially when f and g are far from equal. In 2006 Carbery proposed [3] several plausible refinements of (1.1) for p ≥ 2, of which the strongest was (cid:90) (cid:18) f + gp ≤ 1 + (cid:107)f g(cid:107)p/2 (cid:107)f(cid:107)p(cid:107)g(cid:107)p (cid:19)p−1(cid:90) (fp + gp) . (1.2) There is equality in (1.2) both when f = g and when f g = 0. Thus, (1.2), if true, can be viewed as a refinement of (1.1) in which there is equality not only when f = g but also when f g = 0. The ratio Γ = (cid:107)f g(cid:107)p/2 (cid:107)f(cid:107)p(cid:107)g(cid:107)p varies between 1 and 2p−1, interpolating between the two cases of equality in (1.2). varies between 0 and 1 and, therefore, the factor of (1 + Γ)p−1 c(cid:13) 2018 by the authors. This paper may be reproduced, in its entirety, for non-commercial purposes. Work partially supported by NSF grants DMS -- 1501007 (E.A.C.), DMS -- 1363432 (R.L.F.), PHY -- 1265118 (E.H.L.). 1 2 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB We propose and prove a strengthening of (1.2) in which Γ is replaced by the quantity (cid:18)(cid:107)f(cid:107)p (cid:101)Γ := (cid:107)f g(cid:107)p/2 (cid:19)−2/p p + (cid:107)g(cid:107)p 2 p , (1.3) which is smaller by virtue of the arithmetic-geometric mean inequality. Our improved inequalities are not restricted to p > 2, but are valid for all p ∈ R, as stated in Theorem 1.1. There we write (cid:107)f(cid:107)p := (cid:18)(cid:90) (cid:19)1/p fp for all p (cid:54)= 0. p and (cid:107)f(cid:107)p(cid:107)g(cid:107)p), while our inequality involves only two ((cid:107)f g(cid:107)p/2 and (cid:107)f(cid:107)p We note that inequality (1.2) involves three kinds of quantities on the right side ((cid:107)f g(cid:107)p/2, (cid:107)f(cid:107)p p+ (cid:107)g(cid:107)p p), a simplification that is essential for our proof. Theorem 1.1 (Main Theorem). For all p ∈ (0, 1] ∪ [2,∞) and functions f and g on any p + (cid:107)g(cid:107)p measure space, (cid:90) f + gp ≤ (cid:32) (cid:33)p−1(cid:90) 1 + 22/p(cid:107)f g(cid:107)p/2 p + (cid:107)g(cid:107)p ((cid:107)f(cid:107)p p)2/p (fp + gp ) . (1.4) The inequality reverses if p ∈ (−∞, 0)∪ (1, 2), where, for p ∈ (1, 2), it is assumed that f and g are positive almost everywhere. For p > 2, (resp. for p ∈ (0, 1) ) the inequality is false if (cid:101)Γ is raised to any power q > 1, For p < 0, (resp. for p ∈ [1, 2] ) the reversed inequality is false if (cid:101)Γ is raised to any power (resp. q < 1). q > 1, (resp. for q < 1). For p > 0, p (cid:54)= 1, 2, (cid:107)f(cid:107)p,(cid:107)g(cid:107)p < ∞, there is equality in (1.4) if and only if f and g have disjoint supports, up to a null set, or are equal almost everywhere. For p < 0, (cid:107)f(cid:107)p,(cid:107)g(cid:107)p < ∞, there is equality in (1.4) if and only if f and g are equal almost everywhere. We note that in proving the theorem, we may always assume that f and g are non- negative. In fact, the right side of (1.4) only depends on f and g and the left side does not decrease for p ≥ 0 and does not increase for p < 0 if f and g are replaced by f and g. The latter follows since f + g ≤ f + g implies f + gp ≤ (f + g)p for p > 0 and f + gp ≥ (f + g)p for p < 0. Carbery proved that his proposed inequality is valid when f and g are characteristic functions. Our theorem can also be easily proved in this special case. Theorem 1.1 may be viewed as a refinement of Minkowski's inequality. Since (1.1), like Minkowski's inequality, is a direct expression of the convexity of Bp, it is equivalent to Minkowski's inequality. We recall the simple argument: For any unit vectors u, v ∈ Lp, (1.1) says that (cid:107)(u + v)/2(cid:107)p ≤ 1, and then by continuity, (cid:107)λu + (1 − λ)v(cid:107)p ≤ 1 for all HANNER COMPLEMENT -- December 6, 2018 3 λ ∈ (0, 1). Suppose 0 < (cid:107)f(cid:107)p,(cid:107)g(cid:107)p < ∞, and define λ = (cid:107)f(cid:107)p/((cid:107)fp(cid:107) + (cid:107)g(cid:107)p), u = (cid:107)f(cid:107)−1 and v = (cid:107)g(cid:107)−1 p f , p g. Then (cid:107)f + g(cid:107)p p = ((cid:107)f(cid:107)p + (cid:107)g(cid:107)p)p(cid:107)λu + (1 − λ)v(cid:107)p p ≤ ((cid:107)f(cid:107)p + (cid:107)g(cid:107)p)p , which is Minkowski's inequality. When p = 1 and f, g ≥ 0, (1.1) is an identity; otherwise when p > 1, there is equality in (1.1) if and only if f = g. When the supports of f and g are disjoint, however, (1.1) is far from an equality and the factor 2p−1 is not needed. There is equality in Minkowski's inequality whenever f is a multiple of g or vice-versa. Hence although (1.1) is equivalent to Minkowski's inequality, it becomes an equality in fewer circumstances. There is another well-known refinement of Minkowski's inequality for 1 < p < ∞, namely Hanner's inequality, [4, 2, 6] which gives the exact modulus of convexity of Bp, the unit ball in Lp. For p ≥ 2, and unit vectors u and v, Hanner's inequality says that (cid:13)(cid:13)(cid:13)(cid:13) u + v 2 (cid:13)(cid:13)(cid:13)(cid:13)p p (cid:13)(cid:13)(cid:13)(cid:13) u − v 2 (cid:13)(cid:13)(cid:13)(cid:13)p p + ≤ 1, (1.5) which is also a consequence of one of Clarkson's inequalities [1]. When u and v have disjoint supports, (cid:107)u+v(cid:107)p p = 2, and then the left hand side is 22−p, so that for unit vectors u and v, the condition uv = 0, which yields equality in the inequality of Theorem 1.1, does p = (cid:107)u−v(cid:107)p not yield equality in Hanner's inequality. On the other hand, while one can derive a bound on the modulus of convexity in Lp from (1.4), one does not obtain the sharp exact result provided by Hanner's inequality. Both inequalities express a quantitative strict convexity property of Bp, but neither implies the other; they provide complimentary information, with the information provided by Theorem 1.1 being especially strong when f and g have small overlap as measured by (cid:107)f g(cid:107)p/2. Our proof of Theorem 1.1 consists of three parts: Part A: We show how to reduce the inequality to a simpler one involving only one function, namely α := f /(f + g) for f, g > 0, which takes values in [0, 1], and a reference measure that is a probability measure. This exploits the fact that the only important quantity is the ratio of f to g. This part is very easy. Part B: In the second part, which is more difficult than Part A, we show that Theo- rem 1.1 is true if it is true when the function α is constant. (This is the same as saying f and g are proportional to each other.) When α := f /(f + g) is constant and the reference measure is a probability measure, (1.4) yields the inequality for numbers α ∈ [0, 1] and p ∈ [0, 1] ∪ [2,∞], (cid:32) 1 + 1 ≤ (cid:33)2/pp−1 2αp/2(1 − α)p/2 αp + (1 − α)p (αp + (1 − α)p) . (1.6) 4 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB with the reverse inequality for p /∈ [0, 1] ∪ [2,∞]. 2αp/2(1 − α)p/2 αp + (1 − α)p Remark 1.2. The quantity R := decreases as q increases. Thus for p ≥ 2, the inequality (cid:33)q(cid:33)p−1 (cid:32) (cid:32) 1 ≤ 1 + 2αp/2(1 − α)p/2 αp + (1 − α)p lies in [0, 1] for all α and p. Therefore, Rq (αp + (1 − α)p) , (1.7) strengthens as q increases, and for p ∈ [0, 1], it strengthens as q decreases. Likewise, for p ∈ [1, 2] the reverse of (1.7) is stronger for smaller q, and for p < 0, it is stronger for larger q. Part C: With Parts A and B complete, the proof reduces to a seemingly elementary inequality, parametrized by p, for a number α ∈ [0, 1]. The proof of this is Part C. While the validity of (1.6) appears to be a consequence of Theorem 1.1, one can also view Theorem 1.1 as a consequence of (1.6). Theorem 1.3. For all numbers α ∈ [0, 1], inequality (1.6) is valid for all p ∈ (0, 1]∪ [2,∞), and the reverse inequality is valid for all p ∈ (−∞, 0) ∪ (1, 2). For p > 2, (resp. for p ∈ (0, 1) ) inequality (1.7) is false if q > 2/p, (resp. for q < 2/p). for For p < 0, for p ∈ [1, 2] ) the reverse inequality is false if q > 2/p, (resp. (resp. q < 2/p). For p > 0, p (cid:54)= 1, 2, there is equality if and only if α ∈ {0, 1/2, 1}. For p < 0, there is equality if and only if α = 1/2. 1.1. Restatement of Theorem 1.3 in terms of means. Inequality (1.6) can be restated in terms of qth power means [5]: For x, y > 0, define if q ∈ R \ {0} Mq(x.y) = ((xq + yq)/2)1/q and M0(x, y) = √ xy . Note that M0(x, y) is the geometric mean of x and y and M−1(x, y) is their harmonic mean. Corollary 1.4. For all x, y > 0, and all p ∈ [0, 1] ∪ [2,∞] (cid:18) Mp(x, y) + M−p(x, y) (cid:19)p−1 2 1 (x, y) ≤ M p Mp(x, y) , (1.8) while the reverse inequality is valid for all p ∈ (−∞, 0) ∪ (1, 2). M−p(x, y) Proof. A simple calculation shows that for all p > 0, Mp(x, y) x = α and y = 1 − α, the inequality (1.6) can be written as = 22/pxy (xp + yp)2/p . Thus, taking (cid:18) (cid:19)p−1 ≤ 1 2 M−p(α, 1 − α) Mp(α, 1 − α) 1 + p (α, 1 − α) , M p Then by homogeneity and the fact that M1(α, 1 − α) = 1/2, (1.6) is equivalent to (1.8) (cid:3) HANNER COMPLEMENT -- December 6, 2018 5 The following way to write our inequality sharpens and complements the arithmetic- geometric mean inequality for any two numbers x, y > 0, provided one has information on Mp(x, y). Corollary 1.5 (Improved and complemented AGM inequality). For all x, y > 0, and all p > 2, 1 − (cid:18) A (cid:19)p(cid:48) Mp (cid:32) ≥ 1 2 1 − (cid:19)2(cid:33) (cid:18) G Mp (cid:32) 1 − (cid:19)2(cid:33) (cid:18) G Mp(cid:48) (cid:18) A (cid:19)p Mp(cid:48) ≥ 1 − (1.9) ≥ 1 2 √ (cid:80)n where p(cid:48) = p/(p − 1), A = (x + y)/2 and G = Remark 1.6. Since p, p(cid:48) ≥ 1, all of the quantities being compared in these inequalities are non-negative. xy. Despite the classical appearance of (1.8), we have not been able to find it in the literature, most of which concerns inequalities for means Mq(x1, . . . , xn) = ( 1 j )1/p of an n-tuple n of non-negative numbers, often with more general weights. The obvious generalization of j=1 xp (1.8) from two to three non-negative numbers x, y, and z is false as one sees by taking z = 0: Then there is no help from M−p(x, y, z) on the right. A valid generalization to more variables probably involves means over M−p(xj, xk) for the various pairs. In any case, as far as we know, (1.8) is new. A truly remarkable feature of the inequality (1.8) is that it is surprisingly close to equality uniformly in the arguments. To see this, let f (α, p) denote the right hand side of (1.6). Contour plots of this function for various ranges of p are shown in Figs. 1, 2 and 3 below. Fig. 1 Fig. 2 Fig. 3 Fig. 1 is a contour plot of this function in [1/2, 1] × [2, 4]. The contours shown in Fig. 1 range from 1.00001 to 1.018. Note that the function f is identically 1 along three sides of plot: α = 1/2, 1, and p = 2. The maximum value for 2 ≤ p ≤ 4, near 1.018, occurs towards the middle of the segment at p = 4. Fig. 2 is a contour plot of f on [1/2, 1]× [1, 2]. The contours range from 0.9961 (the small closed contour) to 0.99999999 (close to the boundary). Amazingly, the function in (1.6) is quite close -- within two percent -- to the constant 1 over the range p ≥ 1 and α ∈ [0, 1]. 6 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB Moreover, the "landscape" is quite flat: The gradient has a small norm over the whole domain. Fig. 3 is a contour plot of f in the domain [0, 1/2] × [0, 1]. The contours in Fig. 3 range from 1.0000001 to 1.06. Higher values are to the right. For p in this range, the maximum is not so large -- about 1.06 -- but the landscape gets very "steep" near α = 1 and p = 0. The proof of the inequality is especially delicate in this case. For p < 0, there is equality only at α = 1/2, and the inequality is not so uniformly close to an identity. The contour plot is less informative, and hence is not recorded here. This is the case in which the inequality is easiest to prove. It is possible to give a simple direct proof of the inequality for certain integer values of p, as we discuss in Section 5. We also give a simple proof that for p > 2 and for p < 0, validity of the inequality at p implies validity of the inequality at 2p, and we briefly discuss an application of this to the problem in which functions are replaced by operators and integrals are replaced by traces. Remark 1.7. We close the introduction by briefly discussing one other way to write the inequality (1.6). Introduce a new variable s ∈ (0, 1) through √ s α = 1 + 2 (cid:32) (cid:33) 1 − s 2 p (s) η (1.10) (1.11) (1.12) Rewriting (1.6), and taking the 1 p−1 root of both sides, we may rearrange terms to obtain. 2 ≤ η 1 p−1 (s) 1 + 1 p−1 (s) + (1 − s)η = η 2−p p(p−1) (s) for 0 ≤ s ≤ 1, where √ s)p + (1 − √ s)p (1 + 2 . η(s) := Taking the 1 p−1 eliminate the change of direction in the inequality at p = 1, and it now take on a non-trivial form at p = 1: Define p−1 (s) + (1 − s)η for p (cid:54)= 1, and one easily computes the limit at p = 1: fp(s) := η 1 2−p p(p−1) (s) − 2 , f1(s) := (2 − s)(1 − √ 1−√ s 2 s) (1 + √ s) 1+ √ 2 − 2 . s Theorem 1.3 is equivalent to the assertion that for all s ∈ (0, 1) fp(s) ≥ 0 for p ∈ (−∞, 0) ∪ (2,∞) and fp(s) ≤ 0 for p ∈ (0, 2) . (1.13) In this form, the inequality is easy to check for some values of p. For example, for p = −1, 1−s and f−1(s) = (1 − s)1/2 + (1 − s)−1/2 − 2. which is clearly positive. One can η(s) = 1 give simple proofs of (1.13) for other integer values of p, e.g., p = 3 and p = 4 along these HANNER COMPLEMENT -- December 6, 2018 7 line, but this change of variables is not what we use to prove the general inequality. It is, however, convenient for checking optimality of of the power 2/p in (1.6). 2. Part A. Reduction from two functions to one While Theorem 1.1 involves two functions f and g one can use the arbitrariness of the (cid:82) 1 = 1). We have already observed that it suffices to prove the inequality in the case where measure to reduce the question to a single function defined on a probability space (that is, f and g are both non-negative. For non-negative functions f and g, set α = f /(f + g) , 1 − α = g/(f + g) . Replacing the underlying measure dx by the new measure (f + g)p dx/(cid:107)f + g(cid:107)p p we see that it suffices to prove the following inequality for p ∈ [0, 1] ∪ [2,∞], and also to prove the reverse inequalities for p /∈ [0, 1] ∪ [2,∞] : (cid:32) (cid:33)p−1(cid:0)(cid:107)α(cid:107)p for a single function 0 ≤ α ≤ 1 on a probability space, i.e.,(cid:82) 1 = 1. 22/p (cid:107)α(1 − α)(cid:107)p/2 ((cid:107)α(cid:107)p p )2/p p + (cid:107)1 − α(cid:107)p p + (cid:107)1 − α(cid:107)p p 1 ≤ 1 + (2.1) (cid:1) 3. Part B. Reduction to a constant function In this section we prove the following. Proposition 3.1. If p ∈ [0, 1] ∪ [2,∞), then inequality (2.1) is true for all functions α (which is equivalent to (1.2) for all f, g) if and only if it is true for all constant functions, that is, for all numbers α ∈ [0, 1], (cid:32) 1 + 1 ≤ (cid:33)2/pp−1 2αp/2(1 − α)p/2 αp + (1 − α)p (αp + (1 − α)p) . (3.1) If p /∈ [0, 1] ∪ [2,∞), then the reverse of inequality (2.1) is true for all functions α (which is equivalent to the reverse of (1.2) for all f, g) if and only if it is true for all constant functions, that is, for all numbers α ∈ [0, 1], the reverse of (3.1) holds. Moreover, for p (cid:54)= 0, 1, 2, there is equality in (2.1) if and only if max{α(x), 1 − α(x)} is constant almost everywhere. To prove this Proposition we need a definition and a lemma. Definition 3.2. Fix p ∈ R and for 0 ≤ a ≤ 1, let h(a) := ap/2(1 − a)p/2 and let b(a) := ap+(1−a)p. Clearly, b determines the unordered pair a and 1−a and, therefore, b determines h. Thus, we can consider the function b (cid:55)→ H(b) := h(a−1(b)) (in which the dependence on p is suppressed in the notation). 8 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB Lemma 3.3 (convex/concave H). The function b (cid:55)→ H(b) is strictly convex when p ∈ (2,∞) and strictly concave when p ∈ (−∞, 2), p (cid:54)= 0, 1. Proof. To prove this lemma we use the chain rule to compute the second derivative of H. As a first step we define a useful reparametrization as follows: e2x := a/(1 − a). A quick computation shows that h = (2 cosh x)−p and b = 2 cosh(px)(2 cosh x)−p. Thus, h = b/(2 cosh(px)). By symmetry, we can restrict our attention to the half-line x ≥ 0. We now compute the first two derivatives: db/dx = 21−pp sinh((p − 1)x) (cosh x)p+1 dh/dx = −p tanh x ( 2 cosh x )p (dH/db)(x) = dh/dx db/dx (d/dx)(dH/db)(x) = cosh(x) sinh x 2 sinh((p − 1)x) = − (p − 1) tanh x − tanh((p − 1)x) 2 sinh((p − 1)x) tanh((p − 1)x) (d2H/db2)(x) = (d/dx)(dH/db)(x) db/dx (3.2) (3.3) (3.4) (3.5) (3.6) Our goal is to show that (3.6) has the correct sign (depending on p) for all x ≥ 0. Clearly, the quantity (3.2) is nonpositive for p ∈ (0, 1] and nonnegative elsewhere. We claim that the quantity (3.5) is nonpositive for p ∈ (−∞, 0]∪[1, 2] and nonnegative elsewhere. In fact, the denominator is always positive. For the numerator we write t = p − 1 and use the fact that for all x > 0, t tanh(x) − tanh(tx) > 0 for t > 1 and for −1 < t < 0, while the inequality reverses, and is strict for other values of t except t ∈ {−1, 0, 1}. To see this, fix x > 0, and define f (t) := t tanh(x) − tanh(tx). Evidently f (t) = 0 for t = −1, 0, 1. Then, since f(cid:48)(cid:48)(t) = 2x2 sinh(tx)/ cosh3(tx), f(cid:48)(cid:48)(t) > 0 for t > 0, and f(cid:48)(cid:48)(t) < 0 for t < 0. It follows that f (t) > 0 for −1 < t < 0 and t > 1, while f (t) < 0 for 0 < t < 1 and t < −1. According to (3.6) the products of the signs of (3.2) and (3.5) yield the strict convex- (cid:3) ity/concavity properties of H(b) shown in rows 2 to 4 of the table below. HANNER COMPLEMENT -- December 6, 2018 9 p < 0 0 < p < 1 1 < p < 2 p > 2 db dx (cid:0) dH db (cid:1) d dx ≥ 0 ≤ 0 ≤ 0 ≥ 0 ≥ 0 ≤ 0 ≥ 0 ≥ 0 H(b) concave concave concave convex p(p − 1) Direction ≥ 0 ≥ ≤ 0 ≤ ≥ 0 ≥ ≥ 0 ≤ Fig. 6: Table of signs determining the direction of the main inequality (1.4). Lemma 3.4 (constant b(α(x))). For f, g ≥ 0, define α(x) = f (x)/(f (x) + g(x)). Then with b(α(x)) is almost everywhere constant if max{α(x), 1− α(x)} is constant almost everywhere, which is true if and only if only if either f and g have essentially disjoint support, or else f (x) := max{f (x), g(x)} and g(x) := min{f (x), g(x)} are proportional. Proof. Let F = {x : f (x) > 0} and G = {x : g(x) > 0}. Then α(x) = 1 on F\G, and 1 − α(x) = 1 on G\F . If the measure of F ∩ G is zero, b(α(x)) = 1 almost everywhere with respect to (f (x) + g(x))dx. Conversely, since b(a) = 1 if and only if a ∈ {0, 1}, if b(α(x)) = 1 almost everywhere, then almost everywhere α(x) ∈ {0, 1}, which means that f and g have essentially disjoint supports. For b ∈ [21−p, 1), there is a unique a ∈ [1/2, 1) such that b(a) = b. Therefore, if b(α(x)) = b ∈ [21−p, 1), there is a unique a ∈ [1/2, 1) such that α(x) ∈ {a, 1 − a} almost everywhere, and this is the case if and only if max{α(x), 1 − α(x)} = f (x)/( f (x) + g(x)) = a almost (cid:3) everywhere. Proof of Proposition 3.1. Consider the ratio in (2.1). The numerator is the integral(cid:82) H(b(α(x))). By Jensen's inequality (recalling that(cid:82) 1 = 1) and the convexity/concavity of H in Lemma 3.3, this integral is bounded from below by H(B) in the convex case and from above in the B := (αp + (1 − α)p) . (3.7) concave case, where That is, (cid:90) (cid:82) H(b(y)) (3.8) for p ≥ 2, while the reverse is true for p ≤ 2. Moreover, by the strict convexity/concavity of H(b), the inequality in (3.8) is strict unless b(α(x)) is constant when p /∈ {0, 1, 2}. By the first part of Lemma 3.4, b(α(x)) is constant if and only if max{α(x), 1− α(x)} is a constant, B ≥ H(B) B 10 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB necessarily belonging to [1/2, 1]. Then, taking into account the signs of 2/p and p − 1 in the various ranges, (cid:32) (cid:18) 2(cid:82) H(b(y))dy (cid:19)2/p(cid:33)p−1 (cid:32) (cid:18) 2H(B) (cid:19)2/p(cid:33)p−1 1 + B ≥ 1 + B for p ∈ (0, 1]∪ [2,∞), with the reverse in equality for p ∈ (−∞, 0)∪ [1, 2]. The last two rows in Fig. 6 summarize the interaction of the convexity/concavity properties of H(b) and the signs of the exponents p/2 and p−1 in the direction of the inequality in (3.9) for the different ranges of p, and taking into account the cases of equality discussed above, this yields the result as stated. Thus, it suffices for us to prove 1 ≤ (cid:32) 1 + (cid:18) 2H(B) (cid:19)2/p(cid:33)p−1 B B (3.9) for p ∈ [0, 1] ∪ [2,∞] and the reverse inequality for p /∈ [0, 1] ∪ [2,∞]. We do not know what the number B is, but that does not matter. In each case the range of b is an interval and, therefore, the average value B lies in this same interval. Consequently, whatever B might be, there is a number α such that B = αp + (1 − α)p. (Note that it is not claimed that this (cid:3) number α is related in any particular way to the function α(x).) Proof of Theorem 1.1. This is immediate from Proposition 3.1 and Theorem 1.3. (cid:3) 4. Part C. Proof of Theorem 1.3 4.1. Proof of the inequality. First we prove the inequality (cid:32) 1 + (αp + (1 − α)p) 2αp/2(1 − α)p/2 αp + (1 − α)p (cid:33)2/pp−1 ≥ 1 for all α ∈ (0, 1) (4.1) if p ∈ [0, 1] ∪ [2,∞), and the reverse inequality if p ∈ (−∞, 0] ∪ [1, 2]. For p > 0, there is evidently equality for α ∈ {0, 1 2 , 1}, and for p < 0, there is equality for α = 1/2. Thus for the proof of (4.1) it suffices to consider α ∈ (1/2, 1) for p > 0, and α ∈ (0, 1/2) if p < 0, and it is convenient to change variables (cid:18) 1 − α (cid:19)p ∈ (0, 1] α t := Moreover, for fixed c we introduce the function f (t) := − 1 c ln(1 + tc) + ln(1 + t) + and c := 1/p . (cid:18) ln 1 + (cid:18) 4t (t + 1)2 (cid:19)c(cid:19) . 1 − c c HANNER COMPLEMENT -- December 6, 2018 11 By taking logarithms we see that the claimed inequality (4.1) is equivalent to f (t) ≥ 0 for t ∈ (0, 1) if p ∈ [0, 1] ∪ [2,∞) (that is, c ∈ (0, 1/2] ∪ [1,∞)), and the reverse inequality in (4.1) is equivalent to the reverse inequality if p ∈ (−∞, 0]∪ [1, 2] (that is, c ∈ (−∞, 0]∪ [1/2, 1]). We shall show that for c > 0 the derivative f(cid:48) has a unique sign change in (0, 1) and it changes sign from + to − if c ∈ (0, 1/2)∪ (1,∞) and from − to + if c ∈ (1/2, 1). Moreover, for c < 0 we shall show that the derivative f(cid:48) is positive on (0, 1). Since f (0) = f (1) = 0 for c > 0, this proves that f ≥ 0 if c ∈ (0, 1/2) ∪ (1,∞) and that f ≤ 0 if c ∈ (1/2, 1). Moreover, since f (1) = 0 for c < 0, this proves that f ≤ 0 if c < 0. Thus, we have reduced the proof of Theorem 1.3 to proving the above sign change properties of f(cid:48). In order to discuss the sign changes of f(cid:48) we compute  . (4.2) f(cid:48)(t) = (1 − c)(1 − t) t(1 + t) tc − t (1 − c)(tc + 1)(1 − t) − + 1  (cid:16) (1+t)2 1 (cid:17)c 4t Clearly, it suffices to consider the sign changes of the second factor and therefore to consider the sign changes of g(t) := (cid:18) (1 + t)2 (cid:19)c − 4t (cid:18) (1 − c)(tc + 1)(1 − t) (cid:19) − 1 tc − t . (4.3) We shall show that for c > 0, g has a unique sign change in (0, 1) and it changes sign from − to + if c ∈ (0, 1/2) and from + to − if c ∈ (1/2,∞). Moreover, for c < 0 we shall show that g is negative on (0, 1). Clearly, these properties of g imply the claimed properties of f(cid:48) and therefore will conclude the proof. We next observe that the second term in (4.3) is positive. Lemma 4.1. For any c ∈ R \ {1} and t ∈ (0, 1), (1 − c)(tc + 1)(1 − t) tc − t > 1 . Proof. First, consider the case c ∈ [0, 1). Then concavity of the map t (cid:55)→ tc implies 1 − c + ct − tc ≥ 0, therefore (1−c)(1−t) tc−t ≥ 1, and the claim follows from tc + 1 > 1. Next, for c > 1 the argument is similar using convexity of the map t (cid:55)→ tc. Finally, for c < 0 convexity of t (cid:55)→ t1−c implies that (1 − c)(1 + t−c)(1 − t) (1 − c)(tc + 1)(1 − t) tc − t − 1 = 1 − t1−c − 1 > (1 − c)(1 − t) 1 − t1−c − 1 ≥ 0 . This concludes the proof of the lemma. (cid:3) 12 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB Because of Lemma 4.1, we can define (cid:18) (1 + t)2 (cid:19) 4t − ln (cid:18) (1 − c)(tc + 1)(1 − t) (cid:19) − 1 tc − t h(t) := c ln . (4.4) We shall show that for c > 0, h has a unique sign change in (0, 1) and it changes sign from − to + if c ∈ (0, 1/2) and from + to − if c ∈ (1/2,∞). Moreover, for c < 0 we shall show that h is negative on (0, 1). Clearly, these properties of h imply the claimed properties of g and therefore will conclude the proof. We will prove this by investigating sign changes of h(cid:48). Namely, we shall show that for c > 0, h(cid:48) has a unique sign change in (0, 1) and it changes sign from + to − if c ∈ (0, 1/2) and from − to + if c ∈ (1/2,∞). Moreover, for c < 0 we shall show that h(cid:48) is positive on (0, 1). Let us show that this implies the claimed properties of h. Indeed, an elementary limiting argument shows that h(0) = and  −∞ −2c ln 2 − ln(1 − c) +∞ if c < 0 , if c ∈ (0, 1) , if c > 1 . The function −2c ln 2− ln(1− c) is convex on (0, 1) and vanishes at c = 0 and c = 1/2. From this we conclude that h(1) = 0 for all c . h(0) < 0 if c < 1/2 , h(0) = 0 if c = 1/2 , h(0) > 0 if c > 1/2 . Because of this behavior of h(0) and h(1), the claimed properties of h(cid:48) imply the claimed properties of h. Therefore in order to complete the proof of Theorem 1.3 we need to discuss the sign changes of h(cid:48). We compute h(cid:48)(t) = (1 + t)(tc − t)2(cid:16) (1−c)(tc+1)(1−t) v(t) tc−t (cid:17) − 1 with v(t) := t(2c2 − 1) − t2c2 + 2c(1 − 2c)(tc − tc+1) + t2c(1 − 2c2) + (t1+2c − 1)(1 − c)2 + t2c−1c2 . We shall show that for c > 0, v has a unique sign change in (0, 1) and it changes sign from + to − if c ∈ (0, 1/2) and from − to + if c ∈ (1/2,∞). Moreover, for c < 0 we shall show that v is positive on (0, 1). Since, by Lemma 4.1 the denominator in the above expression for h(cid:48) is positive, these properties of v clearly imply those of h(cid:48) and therefore complete the proof of the theorem. HANNER COMPLEMENT -- December 6, 2018 13 In order to prove the claimed properties of v we shall study the sign changes of v(cid:48)(cid:48). We shall show that for c > 0, v(cid:48)(cid:48) has a unique sign change in (0, 1) and it changes sign from + to − if c ∈ (0, 1/2) and from − to + if c ∈ (1/2,∞). Moreover, for c < 0 we shall show that v(cid:48)(cid:48) is positive. Let us now argue that these properties of v(cid:48)(cid:48) indeed imply the claimed properties of v. We compute v(cid:48)(t) = 2c2 − 1 − 2c2t + 2c(1 − 2c)(ctc−1 − (c + 1)tc) + 2c(1 − 2c2)t2c−1 + (1 − c)2(1 + 2c)t2c + c2(2c − 1)t2c−2, v(cid:48)(cid:48)(t) = 2c · [−c + c(1 − 2c)(c − 1)tc−2 − c(1 − 2c)(c + 1)tc−1 + (2c − 1)(1 − 2c2)t2c−2 + (1 − c)2(1 + 2c)t2c−1 + c(2c − 1)(c − 1)t2c−3], v(cid:48)(cid:48)(cid:48)(t) = 2c(1 − 2c)(c − 1)tc−3w(t) (4.5) (4.6) and finally with w(t) := c(c − 2) − (c + 1)ct − 2tc(1 − 2c2) + (1 − c)(1 + 2c)tc+1 − c(2c − 3)tc−1 . From these formulas we easily infer that v(1) = v(cid:48)(1) = v(cid:48)(cid:48)(1) = 0 , v(cid:48)(cid:48)(cid:48)(1) = 2c(1 − 2c)(c − 1)2 . In particular, v(cid:48)(cid:48)(cid:48)(1) > 0 if c ∈ (0, 1/2) and v(cid:48)(cid:48)(cid:48)(1) < 0 if c ∈ (−∞, 0) ∪ (1/2, 1) ∪ (1,∞). This means that v is convex near t = 1 if c ∈ (−∞, 0) ∪ (1/2,∞) and concave near t = 1 if c ∈ (0, 1/2). Let us discuss the behavior near t = 0. v(0) = +∞, and v(cid:48)(cid:48)(0) > 0. If c > 1/2, then v(0) = −(1 − c)2 and v(cid:48)(cid:48)(0) < 0. If c < 1/2, then v(t) behaves like t2c−1c2, so This behavior of v near 0 and 1, together with the claimed sign change properties of v(cid:48)(cid:48), imply the claimed sign change properties of v and will therefore complete the proof of Theorem 1.3. This is because, for example, if v is convex near t = 1 with v(1) = v(cid:48)(1) = 0, and v has a single inflection point t0 ∈ (0, 1), then v is positive on [t0, 0), and v is concave on (0, t0). Thus, we are left with studying the sign changes of v(cid:48)(cid:48). In order to do so, we need to distinguish several cases. For c < 1 we will argue via the sign changes of v(cid:48)(cid:48)(cid:48), while for c > 1 we will argue directly. Case c ∈ (0, 1). We want to show that v(cid:48)(cid:48) changes sign from + to − if c ∈ (0, 1/2) and from − to + if c ∈ (1/2, 1). Since v(cid:48)(cid:48)(0) > 0 if c ∈ (0, 1/2), v(cid:48)(cid:48)(0) < 0 if c ∈ (1/2, 1), v(cid:48)(cid:48)(1) = 0, and v(cid:48)(cid:48)(cid:48)(1) > 0, it suffices to show that v(cid:48)(cid:48)(cid:48) changes sign only once on (0, 1). Because of (4.6) this is the 14 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB same as showing that w changes sign only once on (0, 1). Notice that w(0) = +∞, and w(1) = c − 1 < 0. Moreover, w(cid:48)(cid:48)(t) = c(1 − c)tc−3p(t) with p(t) := t2(c + 1)(1 + 2c) + 2t(1 − 2c2) + 2c2 − 7c + 6 . Indeed, when c ∈ (0, 1/2) this follows from the The quadratic polynomial p is positive. fact that all its coefficients are positive. When c ∈ (1/2, 1) we observe that the parabola p is minimized on R at t = 2c2−1 , which is positive for c ∈ (1/2, 1). (c+1)(1+2c) , and its minimal value is (5−3c2)+c(11−8c2) The fact that p is positive means that w is convex. Since w(0) = +∞ and w(1) < 0, we (c+1)(1+2c) conclude that w has only one root. Case c ∈ (−∞, 0). We want to show that v(cid:48)(cid:48) is positive. Since v(cid:48)(cid:48)(1) = 0, it suffices to show that v(cid:48)(cid:48)(cid:48) is negative which, by (4.6), is the same as showing that w is negative. Clearly, w(0) = −∞, w(cid:48)(cid:48)(0) < 0, and w(1) = c − 1 < 0, and w(cid:48)(1) = (3c− 1)(c− 1) > 0, so it suffices to show that w(cid:48)(cid:48) < 0 on (0, 1). For this it suffices to show that p > 0 on (0, 1). We have p(0) > 0, and p(1) = 9−4c > 0. Thus if (1+c)(1+2c) ≤ 0 we have proved the claim. Consider the case when (1 + c)(1 + 2c) > 0. The vertex of the parabola is t0 = 2c2−1 clearly (c+1)(1+2c) > 1. If c ∈ (−1/2, 0), then (c+1)(1+2c) . If c < −1 then clearly 2c2−1 2c2−1 (c+1)(1+2c) < 0. Case c ∈ (1,∞). We want to show that v(cid:48)(cid:48) changes sign from − to +. We begin with the case c ∈ (1, 2). We write (4.5) as v(cid:48)(cid:48)(t) = 2cq(t) with q(t) := −c + c(1 − 2c)(c − 1)tc−2 − c(1 − 2c)(c + 1)tc−1 + (2c − 1)(1 − 2c2)t2c−2 +(1 − c)2(1 + 2c)t2c−1 + c(2c − 1)(c − 1)t2c−3. Clearly q(0) = −∞ and q(1) = 0. It is enough to show that q(cid:48) changes sign from + to −. We have q(cid:48)(t) = t2c−4(2c − 1)(c − 1)m(t) with m(t) := c(2 − c)t1−c + c(c + 1)t2−c + 2(1 − 2c2)t + (c − 1)(1 + 2c)t2 + c(2c − 3) . We shall show that m(t) changes sign only once from + to −. Clearly m(0) = +∞ and m(cid:48)(cid:48)(0) > 0. Next, m(1) = 1 − c < 0, and m(cid:48)(cid:48)(1) = (c − 1)(c2 + 2c + 2) > 0. Thus it suffices to show m(cid:48)(cid:48) > 0 on (0, 1). Since m(cid:48)(cid:48)(0) > 0, m(cid:48)(cid:48)(1) > 0, then m(cid:48)(cid:48) > 0 will follow from m(cid:48)(cid:48)(cid:48) having the constant sign. We have m(cid:48)(cid:48)(cid:48)(t) = t−c−2c2(c − 1)(c − 2)(c + 1)(1 − t) < 0. HANNER COMPLEMENT -- December 6, 2018 15 This finishes the case c ∈ (1, 2). If c = 2, then q(t) = (t − 1)(5t2 − 16t + 8), and we see that it changes sign only once. In what follows we assume c > 2. Let us rewrite (4.5) as v(cid:48)(cid:48)(t) = 2ct2c−3u(t) with u(t) := −ct3−2c + c(1 − 2c)(c − 1)t1−c − c(1 − 2c)(c + 1)t2−c + (2c − 1)(1 − 2c2)t + (1 − c)2(1 + 2c)t2 + c(2c − 1)(c − 1) . We need to show that u changes sign only once. We have u(0) = −∞, and u(cid:48)(cid:48)(0) < 0. At the point t = 1, we have u(1) = 0, u(cid:48)(1) = −(2c− 1)(c− 1)2 < 0, u(cid:48)(cid:48)(1) = −2(2c− 1)(c− 1)2 < 0. It suffices to show that u(cid:48)(cid:48) < 0 on (0, 1). Since u(cid:48)(cid:48)(0) < 0, u(cid:48)(cid:48)(1) < 0, the latter claim will follow from showing that u(cid:48)(cid:48)(cid:48) has a constant sign. We have u(cid:48)(cid:48)(cid:48)(t) = t−2−cc(2c − 1)(c − 1)b(t) with b(t) := c3 − c − tc(c + 1)(c − 2) + 2t2−c(2c − 3) . The factor b has the property that b(0) = +∞, b(1) = (c + 6)(c− 1) > 0. On the other hand, b(cid:48)(t) = −c(c + 1)(c − 2)t1−c is negative, so b is positive. (cid:18) tc−1 + 2(2c − 3) c(c + 1) (cid:19) This concludes the proof of the inequality of Theorem 1.3. 4.2. Sharpness of the exponent 2/p. The sharpnes of the exponent 2/p is easily checked using the variables introduced in Remark 1.7. If one rpelaces the power of 2/p in (1.6) and kames the transforations described there, one is led to the function gr,p(s) := η 1 p−1 (s) 1 + − 2 . (4.7) (cid:32) (cid:32) (cid:33)r(cid:33) 1 − s 2 p (s) η instead of fp(s). A motivation for this reparametrization is that for fixed p, the function on the right hand side of (1.6) is equal to 1 up to order O((α − 1/2)4) at α = 1/2. In the variable s, the leading term in Taylor expansion in s will be second order, and we proves the sharpness by an expansion at this point. Proof of the second paragraph of Theorem 1.3. For fixed r > 0, define the function gr,p(s) by (4.7). By the arithmetic-geometric mean inequality, (1− s)p/2 ≤ η(s) for all p, and hence (1 − s)/η2/p(s) < 1 for p > 0, while (1 − s)/η2/p(s) > 1 for p < 0. Therefore, for fixed s and p, gr,p(s) decreases as r increases for p > 0, and does the opposite for p < 0. A Taylor expansion shows that gr,p(s) = p(1 − r)s + o(s) . 16 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB It follows that gr,p(s) ≥ 0 on [0, 1] is false (near s = 0) for p > 2 and r > 1, and for p < 0 and r < 1. Likewise, it follows that gr,p(s) ≤ 0 on [0, 1] is false for p ∈ (0, 2) and r < 1. Since the exponent q in (1.7) corresponds to r(2/p), this together with the reamrks leading to (1.13) justifies the statements referring to q in Theorem 1.3. Consideration of the argument shows that for p > 0, p (cid:54)= 1, 2, there is equality if and only (cid:3) if α ∈ {0, 1/2, 1} and for p < 0, if and only if α = 1/2. The proof of Theorem 1.3 is now complete. By what has been explained above, the inequality of Theorem 1.1 is proved. Concerning the cases of equality, we have seen in Section 3, that for all values of p under consideration, if there is equality then max{α(a), 1 − α(x)} is constant, and then by what has just been proved here, and in Lemma 3.4 for p > 0, this constant is either 1, in which case f and g have essentially disjoint support, or 1/2 in which case = g. For p < 0, there is equality only in case f = g. Finally, it is evident that there is equality in these cases. 5. Doubling arguments and a generalization to Schatten norms. 5.1. Doubling arguments. We begin this section with a simple proof showing that if the inequality (1.4) is valid for some p ≥ 2 or some p < 0, then it is also valid for 2. Since the inequality (1.4) holds as an identity for p = 2, and is simple to prove for p = −1 (see Remark 1.7), this yields a simple proof of infinitely many cases of the inequality (1.4). The proof is not only simple and elegant; it applies to certain non-commutative generalizations of (1.4) for which the reductions in parts A and B of the proof we have just presented are not applicable, as we discuss. To introduce the doubling argument we present a direct proof of Theorem 1.1 for p = 4. Direct proof of Theorem 1.1 for p = 4. Suppose f, g ≥ 0, By homogeneity, we may suppose that (cid:107)f(cid:107)4 4 = 2. Define 4 + (cid:107)g(cid:107)4 X := f g , Y := f 2 + g2 , α := (cid:107)X(cid:107)2 and β := (cid:107)Y (cid:107)2 . (5.1) (cid:90) (cid:90) By the arithmetic-geometric mean inequality, X ≤ 1 2 Y , and hence 1 2 This yields α ≤ 1 and β ≤ 2. Then (f + g)2 = Y + 2X and hence (f 4 + g4 + 2f 2g2)dµ = X 2dµ ≤ 1 4 Y 2dµ = + (cid:90) 1 4 (cid:90) 1 2 X 2dµ . It suffices to prove that β + 2α ≤ 21/2(1 + α)3/2. Note that β2 =(cid:82) (f 2 + g2)2dµ = 2 + 2α2, 4 = (cid:107)Y + 2X(cid:107)2 ≤ (cid:107)Y (cid:107)2 + 2(cid:107)X(cid:107)2 = β + 2α . (cid:107)f + g(cid:107)2 (5.2) and then since α ∈ [0, 1]. Thus it suffices to show that (1 + α2)1/2 ≤ (1 + α)3/2 − 21/2α for all 0 ≤ α ≤ 1 . (5.3) HANNER COMPLEMENT -- December 6, 2018 17 Squaring both sides, this is equivalent to 1+α2 ≤ (1+α)3+2α2−23/2α(1+α)3/2. This reduces to 23/2(1 + α)3/2 ≤ 3 + 4α + α2. Squaring both sides again, this reduces to (α2 − 1)2 ≥ 0, (cid:3) completing the proof. What made this proof work is the fact that the inequality holds for p = 2 -- as an identity, but that is unimportant. Then, using Minkowski's inequality, as in (5.2), together with the numerical inequality (5.3) we arrive at the inequality for p = 4. This is a first instance of the general doubling proposition, to be proved next. The inequality (5.3) is s special case of the general inequality (5.4) proved below. This strategy can be adapted to give direct proof of the inequality for other integer values expansion of (f + g)p = f p + gp + mixed terms. Under the assumption that(cid:82) (f p + gp) = 2, of p; e.g., p = 3. When p is an integer, and f and g are non-negative, one has the binomial one is left with estimating the mixed terms, and one can use Holder for this. When p is not an integer, there is no useful expression for (f + g)p − f p − gp. Proposition 5.1 (A "doubling" argument). Suppose that for some p ≥ 2, (1.4) is valid for all f, g ≥ 0. Then (1.4) is valid with p replaced by 2p for all f, g ≥ 0. Likewise, if for some p < 0 the reverse of (1.4) is valid for all f, g > 0, then the reverse of (1.4) is valid with p replaced by 2p for all f, g > 0. The proof of Proposition 5.1 relies on the following lemma. Lemma 5.2. For t ∈ R, define ψt on [0,∞) by ψt(α) = (1 + α)1+t − (1 + α2)t − 2tα . (5.4) Then for t ∈ [0, 1], ψt(α) ≥ 0 on [0,∞), while for t > 1, ψt(α) ≤ 0 on [0,∞). Proof. We write ψt(α) = (1 + α)t − (1 + α2)t − (2t − (1 + α)t)α. Therefore, ψt(α) α(1 − α) = (1 + α)t − (1 + α2)t α(1 − α) − 2t − (1 + α)t 1 − α . Defining a := 1 + α2, b := 1 + α and c := 2, and defining ϕ(α) := xt, the right hand side is the same as ϕ(b) − ϕ(a) b − a − ϕ(c) − ϕ(b) c − b . For α ∈ [0, 1) we have a < b < c and therefore this quantity is positive when ϕ is concave, and negative when ϕ is convex. For α ∈ (1,∞) we have a > b > c and therefore this quantity (cid:3) is negative when ϕ is concave, and positive when ϕ is convex. Proof of Proposition 5.1. Let f, g ∈ L2p with (cid:107)f(cid:107)2p 2p + (cid:107)g(cid:107)2p Y := f 2 + g2, and γ := (cid:107)X(cid:107)p and β := (cid:107)Y (cid:107)p. By the triangle inequality we have 2p = 2. Define X := f g and (cid:107)f + g(cid:107)2 2p = (cid:107)Y + 2X(cid:107)p ≤ (cid:107)Y (cid:107)p + 2(cid:107)X(cid:107)p = β + 2γ ≥ (cid:107)Y (cid:107)p + 2(cid:107)X(cid:107)p = β + 2γ if p ≥ 2 , if p < 0 . 18 ERIC A. CARLEN, RUPERT L. FRANK, PAATA IVANISVILI, AND ELLIOTT H. LIEB (Note that the triangle inequality reverses for p < 0.) We now use the assumption that the inequality (1.4) is valid for p. Applying the inequality with exponent p to the functions f 2 and g2, which satisfy (cid:107)f 2(cid:107)p p = (cid:107)f(cid:107)2p p + (cid:107)g2(cid:107)p βp = (cid:107)f 2 + g2(cid:107)p and similarly βp ≥ 2(cid:0)1 + γ2(cid:1)p−1 for p < 0. To summarize, we have shown that 2p + (cid:107)g(cid:107)2p = 2(cid:0)1 + γ2(cid:1)p−1 p ≤ 2(cid:0)1 + (cid:107)f 2g2(cid:107)p/2 (cid:1)p−1 2p = 2, we obtain for p ≥ 2, ≤ 21/p(1 + γ2)1−1/p + 2γ ≥ 21/p(1 + γ2)1−1/p + 2γ if p ≥ 2 , if p < 0 . (cid:107)f + g(cid:107)2 2p According to Lemma 5.2 (with t = 1− 1/p and α = γ) this is bounded from above for p ≥ 2 and from below for p < 0 by 21/p(1 + γ)2−1/p = 21/p(1 + (cid:107)f g(cid:107)p)2−1/p, which is the claimed (cid:3) inequality. 5.2. A generalization to Schatten norms. For p ∈ [1,∞), an operator A on some Hilbert space belongs the Schatten p-class Sp in case (A∗A)p/2 is trace class, and the Schatten p norm on Sp is defined by (cid:107)A(cid:107)p = (Tr[(A∗A)p/2])1/p. One possible non-commutative analog of (part of) Theorem 1.1 would assert that for positive A, B ∈ Sp, p > 2. (cid:32) 1 + (cid:33)2/pp−1 Tr(A + B)p ≤ Tr[Bp/4Ap/2Bp/4] 2(cid:107)A(cid:107)p 2(cid:107)B(cid:107)p 1 p + 1 p Tr ( Ap + Bp ) . (5.5) Note that for p = 2, (5.5) holds as an identity. In this setting, it is not clear how to implement analogs of Parts A and B of our proof for functions. However, the direct proofs sketched at the beginning of this section do allow us to prove the valididty of (5.5) for all p = 2k, k ∈ N. Theorem 5.3. If (5.5) is valid for some p ≥ 2 and all positive A, B ∈ Sp, then it is valid for 2p and all A, B ∈ S2p. In particular, since (5.5) holds as an identity for p = 2, it is valid for p = 2k for all k ∈ N. Proof. Let A and B be positive operators in S2p, and assume that (cid:107)A(cid:107)2p by homogeneity, entails no loss of generality. Define 2p +(cid:107)B(cid:107)2p 2p = 2, which, Note that X := 1 2 (AB + BA) and Y = A2 + B2 . (cid:107)X(cid:107)p ≤ 1 2 ((cid:107)AB(cid:107)p + (cid:107)BA(cid:107)p) . By definition, the Lieb -- Thirring inequality [7], and cyclicity of the trace, (cid:107)AB(cid:107)p p = Tr[(BA2B)p/2] ≤ Tr[Bp/2ApBp/2] = Tr[Ap/2BpAp/2] . Define β := (cid:107)Y (cid:107)p and γ := (Tr[Bp/2ApBp/2])1/p . HANNER COMPLEMENT -- December 6, 2018 19 2p = (cid:107)Y + 2X(cid:107)p ≤ (cid:107)Y (cid:107)p + 2(cid:107)X(cid:107)p ≤ β + 2γ. Since (cid:107)A2(cid:107)p p + (cid:107)B2(cid:107)p p = 2, Therefore, (cid:107)A + B(cid:107)2 we can apply (5.5) to deduce that βp = (cid:107)A2 + B2(cid:107)p p ≤ 2 (cid:16) 1 + (Tr[B2p/4ApB2p/4])2/p(cid:17)p−1 = 2(cid:0)1 + γ2(cid:1)p−1 . Altogether (cid:107)A + B(cid:107)2 2p ≤ 21/p(1 + γ2)1−1/p + 2γ and, by Lemma 5.2, the right side is bounded above by 21/p(1 + γ)2−1/p, which proves the (cid:3) inequality, Acknowledgement We thank Anthony Carbery for useful correspondence. References [1] J. A. Clarkson, Uniformly convex spaces. Trans. Amer. Math. Soc., 40 (1936), 396 -- 414. [2] K. Ball, E. A. Carlen, E. H. Lieb, Sharp uniform convexity and smoothness inequalities for trace norms. Invent. Math. 115 (1994), no. 3, 463 -- 482. [3] A. Carbery, Almost orthogonality in the Schatten -- von Neumann classes. J. Operator Theory 62 (2009), no. 1, 151 -- 158. [4] O. Hanner, On the uniform convexity of Lp and (cid:96)p. Ark. Math. 3 (1956), 239 -- 244. [5] G. Hardy, J. E. Littlewood and G. Polya, Inequalities Cambridge Univ. Press, Cambridge, 1934. [6] E. H. Lieb and M. Loss, Analysis, second edition, American Mathematical Society (2014). [7] E. H. Lieb and W. Thirring, Inequalities for the moments of the eigenvalues of the Schrodinger hamilton- ian and their relation to Sobolev inequalities, in Studies in Mathematical Physics, E.H. Lieb, B. Simon, A. Wightman eds., Princeton University Press, 269-303 (1976). (Eric A. Carlen) Department of Mathematics, Hill Center, Rutgers University, 110 Frel- inghuysen Road, Piscataway NJ 08854-8019, USA E-mail address: [email protected] (Rupert L. Frank) Mathematisches Institut, Ludwig-Maximilans Universitat Munchen, There- sienstr. 39, 80333 Munchen, Germany, and Mathematics 253-37, Caltech, Pasadena, CA 91125, USA E-mail address: [email protected] (Paata Ivanisvili) Department of Mathematics, University of California, Irvine, CA 92617, USA E-mail address: [email protected] (Elliott H. Lieb) Departments of Mathematics and Physics, Princeton University, Princeton, NJ 08544, USA E-mail address: [email protected]
1705.04342
2
1705
2017-09-21T19:02:06
An invertibility criterion in a C*-algebra acting on the Hardy space with applications to composition operators
[ "math.FA", "math.CV", "math.OA" ]
In this paper we prove an invertibility criterion for certain operators which is given as a linear algebraic combination of Toeplitz operators and Fourier multipliers acting on the Hardy space of the unit disc. Very similar to the case of Toeplitz operators we prove that such operators are invertible if and only if they are Fredholm and their Fredholm index is zero. As an application we prove that for "quasi-parabolic" composition operators the spectra and the essential spectra are equal.
math.FA
math
AN INVERTIBILITY CRITERION IN A C*-ALGEBRA ACTING ON THE HARDY SPACE WITH APPLICATIONS TO COMPOSITION OPERATORS U GUR G UL AND BEYAZ BAS¸AK KOCA Abstract. In this paper we prove an invertibility criterion for certain opera- tors which is given as a linear algebraic combination of Toeplitz operators and Fourier multipliers acting on the Hardy space of the unit disc. Very similar to the case of Toeplitz operators we prove that such operators are invertible if and only if they are Fredholm and their Fredholm index is zero. As an application we prove that for "quasi-parabolic" composition operators the spectra and the essential spectra are equal. 1. introduction In this paper we investigate the invertibility of elements in the C*-algebra Ψ gen- erated by Toeplitz operators and Fourier multipliers. The C*-algebra Ψ is defined to be Ψ = Ψ(QC, C([0,∞])) = C ∗({Tϕ : ϕ ∈ QC} ∪ {Dϑ : ϑ ∈ C([0,∞])}) the C*-algebra generated by Toeplitz operators with QC symbols and Fourier multi- pliers with continuous symbols. This C*-algebra was introduced by the first author in [3] in order to study the spectral properties of a class of composition operators. In [3], the first author showed that Ψ/K(H 2) is a commutative C*-algebra with identity and determined its maximal ideal space. The maximal ideal space M of Ψ/K(H 2) is found to be homeomorphic to a certain subset of M (QC) × [0,∞] which can be described as M ∼= (M∞(QC(R)) × [0,∞]) ∪ (M (QC(R)) × {∞}) where M∞(QC) is the fiber of M (QC) at infinity. In this paper we show that if T = P Tϕj Dϑj +P Dνj Tψj ∈ Ψ is written as a finite sum or as an infinite sum converging in the operator norm where ψj, ϕj ∈ QC and νj, ϑj ∈ C([0,∞]) then T is invertible if and only if T is Fredholm and has Fredholm index zero. We do this through constructing a homotopy H : [0, 1] → Ψ which is defined as H(w) :=X Tϕj Dϑw j +X Dνw j Tψj where νw j (t) := νj(t − ln w), ϑw j (t) := ϑj(t − ln w) and H(0) := Tϕ where ϕ := P λjϕj +P µjψj with λj := limt→∞ ϑj(t) and µj = limt→∞ νj(t). We observe that this homotopy acts continuously on finite sums hence keeps on to act continuously on infinite sums which converge in operator norm. We apply this result to show 2000 Mathematics Subject Classification. 47B33. Key words and phrases. Composition Operators, Hardy Spaces, Essential Spectra. Submitted May 2, 2017. 1 2 U. G UL, B. KOCA that the class of composition operators that the first author studied in [3] have spectra equal to their essential spectra. The class of composition operators that was studied in [3] is the class of composition operators with symbols ϕ which have upper half-plane re-incarnation C−1 ◦ ϕ ◦ C(z) = z + ψ(z) for a bounded analytic function ψ satisfying ℑ(ψ(z)) > ǫ > 0 for all z ∈ H. We call this class of composition operators "quasi-parabolic". 2. preliminaries In this section we fix the notation that we will use throughout and recall some preliminary facts that will be used in the sequel. Let S be a compact Hausdorff topological space. The space of all complex valued continuous functions on S will be denoted by C(S). For any f ∈ C(S), k f k∞ will denote the sup-norm of f , i.e. k f k∞= sup{ f (s) : s ∈ S}. For a Banach space X, K(X) will denote the space of all compact operators on X and B(X) will denote the space of all bounded linear operators on X. The open unit disc will be denoted by D, the open upper half-plane will be denoted by H, the real line will be denoted by R and the complex plane will be denoted by C. The one point compactification of R will be denoted by R which is homeomorphic to T. For any z ∈ C, ℜ(z) will denote the real part, and ℑ(z) will denote the imaginary part of z, respectively. For any subset S ⊂ B(H), where H is a Hilbert space, the C*-algebra generated by S will be denoted by C ∗(S). The Cayley transform C will be defined by C(z) = z − i z + i . For any a ∈ L∞(R) (or a ∈ L∞(T)), Ma will be the multiplication operator on L2(R) (or L2(T)) defined as Ma(f )(x) = a(x)f (x). For convenience, we remind the reader of the rudiments of the theory of Toeplitz operators and commutative C*-algebras. Let A be a commutative Banach algebra. Then its maximal ideal space M (A) is defined as M (A) = {x ∈ A∗ : x(ab) = x(a)x(b) ∀a, b ∈ A} where A∗ is the dual space of A. If A has identity then M (A) is a compact Hausdorff topological space with the weak* topology. The Gelfand transform Γ : A → C(M (A)) is defined as Γ(a)(x) = x(a). If A is a commutative C*-algebra with identity, then Γ is an isometric *-isomorphism between A and C(M (A)). If A is a C*-algebra and I is a two-sided closed ideal of A, then the quotient algebra A/I is also a C*-algebra (see [6] and [5]). For a Banach algebra A, we denote by com(A) the closed ideal in A generated by the commutators {a1a2−a2a1 : a1, a2 ∈ A}. It is an algebraic fact that the quotient 3 algebra A/com(A) is a commutative Banach algebra. For a ∈ A the spectrum σA(a) of a on A is defined as σA(a) = {λ ∈ C : λe − a is not invertible in A}, where e is the identity of A. In particular the spectrum σ(T ) of a linear bounded operator T : X → X where X is a Banach space is defined as σ(T ) := σB(X)(T ). Recall that a bounded linear operator T on a Hilbert space H is called Fredholm if the range of T is closed, dim ker(T ) and dim ker(T ∗) are finite. The Fredholm index ind is defined as ind(T ) = dim(ker(T )) − dim(ker(T ∗)) It is a very well known fact that ([5]) when the set of Fredholm operators F ⊂ B(H) is equipped with operator norm topology and Z is equipped with discrete topology, the index function ind : F → Z is continuous. The essential spectrum σe(T ) of an operator T acting on a Banach space X is the spectrum of the coset of T in the Calkin algebra B(X)/K(X), the algebra of bounded linear operators modulo compact operators. The following Atkinson's characterization for Fredholm operators is also well known: Theorem 1. [5, p.28, Theorem 1.4.16] A bounded linear operator T on a Hilbert space H is Fredholm if and only if T + K(H) is invertible in the quotient algebra B(H)/K(H), where K(H) is the algebra of all compact operators on H. the Hardy space of the upper half-plane will be denoted by H p(H). For 1 ≤ p < ∞ the Hardy space of the unit disc will be denoted by H p(D) and The two Hardy spaces H 2(D) and H 2(H) are isometrically isomorphic. An iso- metric isomorphism Φ : H 2(D) −→ H 2(H) is given by z + i(cid:19) √π(z + i)(cid:19)g(cid:18) z − i Φ(g)(z) =(cid:18) 1 (1) The mapping Φ has an inverse Φ−1 : H 2(H) −→ H 2(D) given by Φ−1(f )(z) = e 1 2 iπ 2 (4π) (1 − z) 1 − z (cid:19) f(cid:18) i(1 + z) Using the isometric isomorphism Φ, one may transfer Fatou's theorem in the unit disc case to upper half-plane and may embed H 2(H) in L2(R) via f −→ f ∗ where f ∗(x) = limy→0 f (x + iy). This embedding is an isometry. Throughout the paper, using Φ, we will go back and forth between H 2(D) and H 2(H). We use the property that Φ preserves spectra, compactness and essential spectra i.e. if T ∈ B(H 2(D)) then σB(H2(D))(T ) = σB(H2(H))(Φ ◦ T ◦ Φ−1), K ∈ K(H 2(D)) if and only if Φ ◦ K ◦ Φ−1 ∈ K(H 2(H)) and hence we have (2) We also note that T ∈ B(H 2(D)) is essentially normal if and only if Φ ◦ T ◦ Φ−1 ∈ B(H 2(H)) is essentially normal. σe(T ) = σe(Φ ◦ T ◦ Φ−1). The Toeplitz operator with symbol a is defined as Ta = P MaH2 , 4 U. G UL, B. KOCA where P denotes the orthogonal projection of L2 onto H 2. A good reference about Toeplitz operators on H 2 is Douglas' treatise ([2]). Although the Toeplitz operators treated in [2] act on the Hardy space of the unit disc, the results can be transfered to the upper half-plane case using the isometric isomorphism Φ introduced by equation (1). In the sequel the following identity will be used: Φ−1 ◦ Ta ◦ Φ = Ta◦C−1, (3) where a ∈ L∞(R). We also employ the fact (4) for any a ∈ L∞(R), which is a consequence of Theorem 7.11 of [2] (pp. 160–161) and equation (3). For any subalgebra A ⊆ L∞(R) the Toeplitz C*-algebra generated by symbols in A is defined to be k Ta ke=k Ta k=k a k∞ It is a well-known result of Sarason (see [7]) that the set of functions T (A) = C ∗({Ta : a ∈ A}). H ∞ + C = {f1 + f2 : f1 ∈ H ∞(D), f2 ∈ C(T)} is a closed subalgebra of L∞(T). The following theorem of Douglas [2] will be used in the sequel. Theorem 2 (Douglas' Theorem). Let a,b ∈ H ∞+C then the semi-commutators Tab − TaTb ∈ K(H 2(D)), Tab − TbTa ∈ K(H 2(D)), and hence the commutator is compact. [Ta, Tb] = TaTb − TbTa ∈ K(H 2(D)) Let QC be the C*-algebra of functions in H ∞ + C whose complex conjugates also belong to H ∞ + C. Let us also define the upper half-plane version of QC as the following: QC(R) = {ϕ ∈ L∞(R) : ϕ ◦ C−1 ∈ QC}. Going back and forth with Cayley transform one can deduce that QC(R) is a closed subalgebra of L∞(R). By Douglas' theorem and equation (3), if a, b ∈ QC(R), then TaTb − Tab ∈ K(H 2(H)). Let scom(QC(R)) be the closed ideal in T (QC(R)) generated by the semi-commutators {TaTb − Tab : a, b ∈ QC(R)}. Then we have com(T (QC(R))) ⊆ scom(QC(R)) ⊆ K(H 2(H)). By Proposition 7.12 of [2] and equation (3) we have com(T (QC(R))) = scom(QC(R)) = K(H 2(H)). Now consider the symbol map (5) defined as Σ(a) = Ta. This map is linear but not necessarily multiplicative; however if we let q be the quotient map Σ : QC(R) → T (QC(R)) q : T (QC(R)) → T (QC(R))/scom(QC(R)), 5 then q ◦ Σ is multiplicative; moreover by equations (4) and (5), we conclude that q ◦ Σ is an isometric *-isomorphism from QC(R) onto T (QC(R))/K(H 2(H)). The maximal ideal space M (QC(R)) is fibered over R in the following way: For any x ∈ M (QC(R)) consider x = xC( R) then x ∈ M (C( R)) = R. Hence M (QC(R)) is fibered over R, i.e. where M (QC(R)) = [t∈ R Mt(QC), Mt(QC) = {x ∈ M (QC(R)) : x = xC( R) = δt, δt(f ) = f (t)}. We also remind the reader about the very important fact in the theory of Toeplitz operators that any Toeplitz operator Tϕ with symbol ϕ ∈ L∞ is invertible if and only if Tϕ is Fredholm and the Fredholm index ind(Tϕ) = 0 is zero. The proof of this fact can also be found in [2]. This fact will be used in the proof of our main result in this paper. Let ϕ : D −→ D or ϕ : H −→ H be a holomorphic self-map of the unit disc or the upper half-plane. The composition operator Cϕ on H p(D) or H p(H) with symbol ϕ is defined by Cϕ(g)(z) = g(ϕ(z)), z ∈ D or z ∈ H. Composition operators of the unit disc are always bounded [1] whereas composition operators of the upper half-plane are not always bounded. For the boundedness problem of composition operators of the upper half-plane see [4]. The composition operator Cϕ on H 2(D) is carried over to ( ϕ(z)+i z+i )C ϕ on H 2(H) through Φ, where ϕ = C ◦ ϕ ◦ C−1, i.e. we have ΦCϕΦ−1 = T( ϕ(z)+i z+i )C ϕ. (6) However this gives us the boundedness of Cϕ : H 2(H) → H 2(H) for ϕ(z) = pz + ψ(z), where p > 0, ψ ∈ H ∞ and ℑ(ψ(z)) > ǫ > 0 for all z ∈ H: have Let ϕ : D → D be an analytic self-map of D such that ϕ = C−1 ◦ ϕ ◦ C, then we where If ΦC ϕΦ−1 = Tτ Cϕ τ (z) = ϕ(z) + i z + i . ϕ(z) = pz + ψ(z) with p > 0, ψ ∈ H ∞ and ℑ(ψ(z)) > ǫ > 0, then T 1 ΦC ϕΦ−1 is always bounded we conclude that Cϕ is bounded on H 2(H). is a bounded operator. Since τ The Fourier transform F f of f ∈ S(R) (the Schwartz space, for a definition see [6]) is defined by The Fourier transform extends to an invertible isometry from L2(R) onto itself with inverse (F f )(t) = 1 √2π Z +∞ −∞ e−itxf (x)dx. (F −1f )(t) = 1 √2π Z +∞ −∞ eitxf (x)dx. 6 U. G UL, B. KOCA The following is a consequence of a theorem due to Paley and Wiener [6]. Let 1 < p < ∞. For f ∈ Lp(R), the following assertions are equivalent: (i) f ∈ H p, (ii) supp( f ) ⊆ [0,∞) A reformulation of the Paley-Wiener theorem says that the image of H 2(H) under the Fourier transform is L2((0,∞)). By the Paley-Wiener theorem we observe that the operator Dϑ = F −1MϑF for ϑ ∈ C([0,∞]) maps H 2(H) into itself, where C([0,∞]) denotes the set of con- tinuous functions on [0,∞) which have limits at infinity. Since F is unitary we also observe that Let F be defined as k Dϑ k=k Mϑ k=k ϑ k∞ F = {Dϑ ∈ B(H 2(H)) : ϑ ∈ C([0,∞])}. We observe that F is a commutative C*-algebra with identity and the map D : C([0,∞]) → F given by D(ϑ) = Dϑ for a.e. is an isometric *-isomorphism by equation above. Hence F is isometrically *- isomorphic to C([0,∞]). The operator Dϑ is usually called a "Fourier Multiplier." We will also need the fact that, under the Fourier transform the Beurling type invariant subspace eiηxH 2(H), for η > 0, is mapped onto L2 η((0,∞)) := {f ∈ L2(0,∞) : f (t) = 0 t ∈ (0, η)} for all η > 0 and F TeiηxF −1 = Sη where Sη : L2((0,∞)) → L2((0,∞)) is defined as Sηf (t) := f (t−η) if t ≥ η and Sηf (t) = 0 if 0 ≤ t < η. Similarly we have F Te−iηxF −1 = S ∗ η f (t) = f (t + η). Here Teiηx and Te−iηx are Toeplitz operators with symbols eiηz and e−iηz respectively. η((0,∞)) is dense in L2((0,∞)), ∪η>0eiηxH 2(H) is also dense in H 2(H). Since ∪η>0L2 In [3] the first author studied the C*-algebra Ψ generated by Toeplitz operators with QC symbols and Fourier multipliers with continuous symbols. He proved that the commutator [Tϕ, Dϑ] = TϕDϑ − DϑTϕ ∈ K(H 2) of any Toeplitz operator with QC symbol and a Fourier multiplier with continuous symbol is compact which im- plies that the Calkin algebra Ψ/K(H 2) is a commutative C*-algebra with identity. The maximal ideal space M of Ψ/K(H 2) is also studied in [3] and is found to be η where S ∗ M ∼= (M∞(QC(R)) × [0,∞]) ∪ (M (QC(R)) × {∞}) ⊂ M (QC) × [0,∞] where M∞(QC) := {x ∈ M (QC) : xC( R) = δ∞, δ∞(f ) = limt→∞ f (t)} is the fiber of M (QC) at infinity. The Gelfand transform Γ of Ψ/K(H 2) looks like Γ(cid:16)hX Tϕj Dϑji(cid:17) (x, t) =(P ϕj (x) ϑj (t) P ϕj (x) ϑj (∞) 3. the main result if x ∈ M∞(QC(R)) if t = ∞ In this section we prove the main result of this paper which asserts that any sum T = P Tϕj Dϑj +P Dνj Tψj ∈ Ψ convergent in the operator norm is invertible if and only if T is Fredholm and has Fredholm index zero. This may be regarded as a generalization of the fact that any Toeplitz operator Tϕ with a bounded symbol ϕ ∈ L∞ is invertible if and only if Tϕ is Fredholm and has Fredholm index zero. In 7 proving this fact our main technical tool will be a homotopy H : [0, 1] → Ψ which carries T to a Toeplitz operator. The homotopy H for T =P Tϕj Dϑj +P Dνj Tψj is defined as follows: H(w) :=X Tϕj Dϑw j (t) = ϑj(t − ln w), νw j +X Dνw where w ∈ (0, 1], ϑw j (t) = νj(t − ln w) and for w = 0, H(0) := Tϕ, ϕ = P λjϕj +P µjψj and λj = limt→∞ ϑj(t), µj = limt→∞ νj(t). It is easily seen that when T = P Tϕj Dϑj +P Dνj Tψj consists of finite sums, H is continuous. And thus H keeps on being continuous when T consists of infinite sums which are both convergent in operator norm. In the proof of this fact we will always work with finite sums since it is enough to prove it for finite sums. In this section, unless otherwise stated, H 2 will always be understood as H 2(H) and QC will always be understood as QC(R). j Tψj Here is our main theorem: Theorem 3. Let T =P Tϕj Dϑj +P Dνj Tψj ∈ Ψ be such that νj, ϑj ∈ C([0,∞]), ψj, ϕj ∈ QC. Then T ∈ Ψ is invertible ⇔ T is Fredholm and ind(T ) = 0 Proof. (⇒): trivial. (⇐): Let T be Fredholm and ind(T ) = 0. Let H : [0, 1] → Ψ be the homotopy con- structed above. Since σe(T ) = {P ϕj (x)ϑj (t) +P ψj(x)νj (t) : x ∈ M∞(QC), t ∈ [0,∞]}∪{P λj ϕj(x)+P µj ψj(x) : x ∈ M (QC)} and σe(H(w)) = {P ϕj(x)ϑj (t)+ P ψj(x)νj (t) : x ∈ M∞(QC), t ∈ [− ln w,∞]} ∪ {P λj ϕj(x) +P µj ψj(x) : x ∈ M (QC)}, we have σe(H(w)) ⊆ σe(T ) for all w ∈ [0, 1]. Hence if T is Fredholm then 0 6∈ σe(T ) ⇒ 0 6∈ σe(H(w)) for all w ∈ [0, 1] which implies that H(w) is Fredholm for all w ∈ [0, 1]. Since H is continuous and ind is continuous on the set of Fredholm operators, we have ind(H(0)) = ind(H(w)) = ind(H(1)) = ind(T ) = 0 ∀w ∈ [0, 1]. Hence ind(H(0) = ind(Tϕ) = 0 which implies that Tϕ is invertible since any Toeplitz operator with L∞ symbol is invertible iff it is Fredholm with Fredholm index zero. Since invertible elements in Ψ form an open subset and H is continu- ous, there is a w0 ∈ (0, 1] such that H(w) is invertible for all 0 ≤ w < w0. Now suppose that H(w0) is not invertible. Sη where Mg : L2((0,∞)) → L2((0,∞)), Mgf (t) := g(t)f (t) is the multiplication operator. Hence we have For η = ln(w0) − ln(w), where w0 > w, we have Mϑw j = S ∗ η Mϑ w0 j Dϑw j = (F −1S ∗ ηF )Dϑ w0 j (F −1SηF ) = Te−iηx Dϑ w0 j Teiηx . Since ψj, ϕj ∈ QC we have Tϕj Te−iηx − Te−iηxTϕj ∈ K(H 2) and Tψj Teiηx − TeiηxTψj ∈ K(H 2) for all η > 0(see [2]). Hence we have j Tψj = j +X Dνw Teiηx +X Te−iηx Dν +X Dν = Te−iηxH(w0)Teiηx + K(w, w0). where K(w, w0) ∈ K(H 2) is a compact operator. H(w) =X Tϕj Dϑw X Tϕj Te−iηx Dϑ = Te−iηx(X Tϕj Dϑ w0 j w0 j w0 j Teiηx Tψj w0 j Tψj )Teiηx + K(w, w0) 8 U. G UL, B. KOCA Hence we have H(w) = Te−iηxH(w0)Teiηx + K(w, w0) for some compact operator K(w, w0) ∈ K(H 2). Since Fredholm index is stable under compact perturbations this implies that the operator Te−iηx H(w0)Teiηx is Fredholm with index 0 for all η > 0. Hence if H(w0) is non-invertible then ker(Te−iηxH(w0)Teiηx ) 6= {0} for some η > 0 since Te−iηx H(w0)Teiηx is Fredholm with index 0. If this is not the case i.e. if we have ker(Te−iηx H(w0)Teiηx ) = {0} for all η > 0, since Te−iηx H(w0)Teiηx is Fredholm with index 0, Te−iηx H(w0)Teiηx is invertible for all η > 0. And this implies that H(w0) maps each eiηxH 2 onto itself in a one to one manner. Hence H(w0) maps ∪η>0eiηxH 2 onto itself in a one to one manner. Since H(w0) is Fredholm, ran(H(w0)) ⊆ H 2 is closed in H 2 and we have ∪η>0eiηxH 2 ⊆ ran(H(w0)). Since ∪η>0eiηxH 2 is dense in H 2 and ind(H(w0)) = 0, this implies that ran(H(w0)) = H 2 which in turn implies that ker(H(w0)) = {0} and hence H(w0) is invertible which is a contradiction. Hence we should have ker(Te−iηxH(w0)Teiηx ) 6= {0} for some η > 0. This implies that either ker(H(w0))∩ (eiηxH 2) 6= {0} or H(w0)(eiηxH 2) ∩ ker(Te−iηx ) 6= {0} should hold for some η > 0. Now suppose that ψj, ϕj ∈ H ∞ for all j ∈ N. Since ψj, ϕj ∈ H ∞ for all j, we have H(w0)(eiηxH 2) ⊆ (eiηxH 2) for all η > 0 and ker(Te−iηx ) = (eiηxH 2)⊥, we have H(w0)(eiηxH 2) ∩ ker(Te−iηx ) = {0} for all η > 0. So if H(w0) is non-invertible we should have ker(H(w0))∩ (eiηxH 2) 6= {0}. Since H(w0) is Fredholm, ker(H(w0)) is finite dimensional an η0 := max{η > 0 : ker(H(w0)) ∩ (eiηxH 2) 6= {0}} exists. For such η0 > 0, let H0 := Te−iη0xH(w0)Teiη0 x. Then since ker(H(w0)) ∩ (eiη0xH 2) 6= {0}, H0 is non-invertible. But since ker(H(w0)) ∩ (ei(η0+δ)xH 2) = {0} for all δ > 0 and H(w0)(ei(η0+δ)xH 2) ⊆ ei(η0+δ)xH 2 we have ker(Te−i(η0 +δ)x H(w0)Tei(η0 +δ)x ) = ker(Te−iδx H0Teiδx ) = {0} and this implies that Te−iδx H0Teiδx is invertible for all δ > 0. This again implies that H0 maps eiδxH 2 onto itself in a one to one man- ner which implies that H0 should be invertible. This contradicts our assumption that H(w0) is non-invertible. Hence H(w0) should be invertible. Therefore, in this case(where ψj, ϕj ∈ H ∞ for all j), if H(w) is invertible for all 0 ≤ w < w0 then H(w0) is also invertible. So by transfinite induction H(w) is invertible for all w ∈ [0, 1] and in particular H(1) = T is invertible. Now suppose that ϕj and ψj are continuous for all j. Since trigonometric poly- nomials are dense in continuous functions, it is enough to prove the claim when k=m bkzk are a trigonometric polynomials ϕj(z) := Pm k=−m akzk and ψj(z) = Pm for all j. One can write ϕj and ψj in the form ϕj(z) = z−m m′ Xk=0 ajzk = z−mqj(z), ψj(z) = z−mpj(z) where z = ( x−i x+i ) and qj and pj are analytic polynomials. In this case we have H(w) = Te−iηx H(w0)Teiηx + K(w, w0) w0 j = Te−iηx (X Tϕj Dϑ = Te−iηx (X Tz−mTqj Dϑ = Tz−mTe−iηx (X Tqj Dϑ w0 j w0 j +X Dν w0 j +X Dν +X Dν w0 j w0 j Tψj )Teiηx + K(w, w0) Tz−mTpj)Teiηx + K(w, w0) Tpj )Teiηx + K0 + K(w, w0) 9 w0 j w0 j have +P Dν H(w) = Te−iηx Tz−m H(w0)Teiηx + K(w, w0) for some K0 ∈ K(H 2) since Dνj Tz−m − Tz−mDνj ∈ K(H 2) ∀j, Tz−mTe−iηx = Te−iηxTz−m and the sum is finite. Let H(w0) =P Tqj Dϑ Tpj . Then we where K(w, w0) = K0 + K(w, w0) ∈ K(H 2). Since Tz−mTe−iηx = Te−iηx Tz−m we have ker( H(w0)) ∩ eiηxH 2 6= {0} for some η > 0. If this is not the case i.e. if ker( H(w0))∩eiηxH 2 = {0} ∀η > 0 then H(w)− K(w, w0) = Tz−mTe−iηx H(w0)Teiηx is invertible for all η > 0. And this implies that H(w0)(eiηxH 2) = zm(eiηxH 2) ∀η > 0. Since ran( H(w0) ⊆ H 2 is closed, this implies that ran( H(w0)) = zmH 2 and this in turn implies that H(w0) = Tz−m H(w0) is invertible which contradicts our assumption. Since ker( H(w0) is finite dimensional, an η0 := max{η > 0 : ker( H(w0)) ∩ (eiηxH 2) 6= {0}} exists. Now let H0 := Te−iη0 x H(w0)Teiη0x, then ker( H0) 6= {0}. Since ker( H(w0)) ∩ (ei(η0+δ)xH 2) = {0} ∀δ > 0 we have Tz−mTe−iδx H0Teiδx = Te−iδx Tz−m H0Teiδx is invertible for all δ > 0. This implies that H0(eiδxH 2) = zm(eiδxH 2) for all δ > 0. Since ran( H0) ⊆ H 2 is closed, this implies that ran(H0) = zmH 2. Since ind(Tz−m H0) = 0 this implies that ind( H0) = −m which in turn implies that ker( H0) = {0}. This contradiction implies that H(w0) is invertible. Hence if ψj , ϕj ∈ H ∞ or ψj, ϕj are continuous for all j then T = P Tϕj Dϑj + P Dνj Tψj is Fredholm with index zero implies that T is invertible. So if ψj, ϕj ∈ (H ∞ + C) ∩ QC = QC for all j then T =P Tϕj Dϑj +P Dνj Tψj is Fredholm with index zero implies that T is invertible. (cid:3) A reinterpretation of this theorem would be relating the invertibility of a generic element in Ψ to the invertibility of a related Toeplitz operator which is the corollary below: Corollary 4. For any T = P Tϕj Dϑj +P Dνj Tψj ∈ Ψ such that T is Fredholm ψj ϕj ∈ QC and νj, ϑj ∈ C([0,∞]) ∀j ∈ N, T is invertible if and only if Tϕ ∈ Ψ is invertible where ϕ =P λjϕj +P µjψj and µj = limt→∞ νj(t), λj = limt→∞ ϑj(t). Proof. If T is invertible then T is Fredholm with ind(T ) = 0. Using the homotopy H : [0, 1] → Ψ constructed in the beginning of this section, since ind is continuous we have H(w) is Fredholm ∀w ∈ [0, 1] and ind(H(w)) = ind(T ) = 0 ∀w ∈ [0, 1]. In particular H(0) = Tϕ is Fredholm and ind(Tϕ) = 0, since any Fredholm Toeplitz operator Tϕ with ind(Tϕ) = 0 is invertible, Tϕ is invertible. On the other hand if Tϕ is invertible, since T is Fredholm, by the proof of Theorem 3, H(w) is invertible ∀w ∈ [0, 1], in particular H(1) = T is invertible. (cid:3) 4. Applications of the Main Results An immediate application of Corollary 4 shows that the essential spectrum and the spectrum of a quasi-parabolic composition operator coincide: Theorem 5. Let ϕ : D → D and ϕ : H → H be such that ϕ(z) = 2iz+η(z)(1−z) and ϕ(w) = w + ψ(w) where η ∈ QC(T) ∩ H ∞, ℑ(η(z)) > δ > 0 for all z ∈ D, ψ ∈ QC(R) ∩ H ∞, ℑ(η(w)) > δ > 0 for all w ∈ H. Then C ϕ : H 2(H) → H 2(H) is bounded and σe(C ϕ) = σ(C ϕ). We also have σe(Cϕ) = σ(Cϕ). 2i+η(z)(1−z) 10 U. G UL, B. KOCA Proof. The boundedness of C ϕ was shown in [3]. In particular we have ∞ C ϕ = Tτ nDϑn Xn=0 where τ (x) = iα − ψ(x) and ϑn(t) = (−it)ne−αt λ 6∈ σe(C ϕ) we have n! ∞ for some α > 0. So for any λ − C ϕ = λ − Tτ nDϑn Xn=0 where the series on the Right Hand Side converges in the operator norm. Since λn := limt→∞ ϑn(t) = 0, by Corollary 1 we have λ − C ϕ is invertible if and only if λ is invertible which is certainly the case since λ 6= 0. Hence λ 6∈ σ(C ϕ) ⇒ σ(C ϕ) ⊆ σe(C ϕ) ⇒ σ(C ϕ) = σe(C ϕ). The same argument applies to Cϕ since ∞ Φ ◦ Cϕ ◦ Φ−1 = T z+i+η◦C(z) Xn=0 where Φ : H 2(D) → H 2(H) is the isometric isomorphism √π(z + i)(cid:19)f(cid:18) z − i z + i(cid:19), Φ(f )(z) =(cid:18) z+i 1 Tτ nDϑn C is the Cayley transform and τ (x) = iα − η ◦ C(x). Hence we also have σ(Cϕ) = σe(Cϕ). (cid:3) References [1] C.C. Cowen, B.D. MacCluer Composition Operators on Spaces of Analytic Functions, CRC Press, 1995. [2] R.G. Douglas, Banach algebra techniques in Operator Theory, GTM Springer, 1998. [3] U. Gul, Essential spectra of quasi-parabolic composition operators on Hardy spaces of ana- lytic functions, J. Math. Anal. Appl. 377, (2011), 2, 771–791 [4] V. Matache, Composition operators on Hardy spaces of a half-plane, Proc. Amer. Math. Soc. 127, (1999), no. 5, 1483–1491. [5] G.J. Murphy, C ∗-algebras and Operator Theory, Academic Press, New York-London, 1990. [6] W. Rudin, Functional Analysis, McGraw Hill Inc., 1973. [7] D. Sarason, Function Theory on the Unit Circle, Lecture Notes for a Conference at Virginia Polytechnic and State University, Blacksburg, Virginia, 1978. ugur gul, Hacettepe University, Department of Mathematics, 06800, Beytepe, Ankara, TURKEY E-mail address: [email protected] beyaz bas¸ak koca, Istanbul University, Department of Mathematics, 34134, Vezneciler, Istanbul, TURKEY E-mail address: [email protected]
1201.3531
2
1201
2012-11-06T08:06:17
On the peripheral point spectrum and the asymptotic behavior of irreducible semigroups of Harris operators
[ "math.FA" ]
Given a positive, irreducible and bounded C_0-semigroup on a Banach lattice with order continuous norm, we prove that the peripheral point spectrum of its generator is trivial whenever one of its operators dominates a non-trivial compact or kernel operator. For a discrete semigroup, i.e. for powers of a single operator T, we show that the point spectrum of some power T^k intersects the unit circle at most in 1. As a consequence, we obtain a sufficient condition for strong convergence of the C_0-semigroup and for a subsequence of the powers of T, respectively.
math.FA
math
ON THE PERIPHERAL POINT SPECTRUM AND THE ASYMPTOTIC BEHAVIOR OF IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS MORITZ GERLACH Abstract. Given a positive, irreducible and bounded C0-semigroup on a Ba- nach lattice with order continuous norm, we prove that the peripheral point spectrum of its generator is trivial whenever one of its operators dominates a non-trivial compact or kernel operator. For a discrete semigroup, i.e. for powers of a single operator T , we show that the point spectrum of some power T k intersects the unit circle at most in 1. As a consequence, we obtain a sufficient condition for strong convergence of the C0-semigroup and for a subsequence of the powers of T , respectively. 1. Introduction E. B. Davies proved in [9] that σp(A)∩iR ⊆ {0} if A is the generator of a positive and contractive semigroup on ℓp for some 1 ≤ p < ∞. This result was generalized by V. Keicher in [14] to positive, bounded and strongly continuous semigroups on atomic Banach lattices with order continuous norm. A further generalization to (w)-solvable semigroups on super-atomic Banach lattices was found by M. Wolff [19]. Independently, G. Greiner observed in [13, Thm. 3.2 and Kor. 3.11] that σp(A)∩ iR ⊆ {0} for the generator A of a positive, contractive and strongly continuous semigroup on Lp(Ω, µ), 1 ≤ p < ∞, if the semigroup contains a kernel operator. This covers in particular the atomic case because every regular operator on an order complete atomic Banach lattice is a kernel operator. Greiner's result is also in the center of a survey article by W. Arendt [4] that includes a generalization to Abel bounded semigroups. Under the assumption that the semigroup is irreducible, we generalize Greiner's result in the present article in two respects. First, we consider general Banach lattices with order continuous norm instead of Lp and secondly, we merely assume that one operator of the semigroup dominates a non-trivial compact operator. Un- der suitable assumptions, the latter holds in particular if the semigroups contains a Harris operator, i.e. one operator dominates a non-trivial kernel operator. More precisely, we prove the following theorem in Section 3. Theorem. Let E be a Banach lattice with order continuous norm and let T = (T (t))t∈[0,∞) be a positive, irreducible, bounded and strongly continuous semigroup on E with generator A. Assume that there exists t0 > 0 such that T (t0) ≥ K > 0 where K is a compact or a kernel operator. Then σp(A) ∩ iR ⊆ {0}. This result remains true for semigroups of Harris operators which are merely Abel bounded, see Proposition 3.2. 2010 Mathematics Subject Classification. Primary 47A11; Secondary: 47B65, 47G10, 47D06. Key words and phrases. peripheral point spectrum, irreducible, Harris operator, compact op- erator, asymptotic behavior. 1 2 MORITZ GERLACH We also consider discrete semigroups by which we mean a family T = (T n)n∈N0 for some bounded operator T . If T is positive, irreducible and power bounded such that T m dominates a non-trivial compact or kernel operator for some m ∈ N, then σp(T n) ∩ Γ ⊆ {1} for some n ∈ N where Γ denotes the unit circle. This result, presented as The- orem 3.4, is optimal in the sense that, under these assumptions, one can neither expect σp(T ) ∩ Γ ⊆ {1} for the point spectrum of T nor σ(T n) ∩ Γ ⊆ {1} for the spectrum of some operator T n, see Examples 3.3 and 3.6. One important application of the above-mentioned results lies in the analysis of the asymptotic behavior of a semigroup. Greiner deduced from his result that a positive and contractive strongly continuous semigroup on Lp converges strongly to an equilibrium if it admits a strictly positive fixed point and contains a kernel operator [13, Kor. 3.11]. See also [4, Thm. 4.2] for a proof in the irreducible case. In Section 4, we generalize this to bounded and irreducible semigroups of Harris operators on Banach lattices with order continuous norm. In addition, we obtain an analogous result for discrete semigroups in Theorem 4.3. If T is a positive, power bounded and irreducible Harris operator with non-trivial fixed space, then there exists some k ∈ N such that the subsequence T nk converges strongly as n tends to infinity. If T is even strongly positive, i.e. T x is a quasi- interior point of E+ for all x > 0, we show that the whole sequence T n converges strongly. The main tools for our study of the asymptotic behavior are a so-called zero-two law by G. Greiner [13, Thm. 3.7] and a result by D. Axmann [7, Satz 3.5] on dis- jointness of powers of an irreducible operator. In order to improve the accessibility of both results we present their proofs in the appendices. 2. Notation and Tools Let us fix some notation and recall some well-known facts from the theory of Banach lattices and positive operators. The reader is referred to [16] for a more detailed introduction. Throughout, E denotes a (real) Banach lattice with order continuous norm. We write E+ for the positive cone of E and x > 0 for an element x ∈ E+ different from zero. For y ∈ E+ we denote by Ey := {x ∈ E : x ≤ cy for some c > 0} = [c≥0 [−cy, cy] the principal ideal generated by y. If Ey is dense in E, then y is called a quasi- interior point of E+. We say that a positive functional x′ ∈ E′ is strictly positive if hx′,xi > 0 for all x 6= 0. By Fix(T ) := ker(I − T ) we denote the fixed space of a linear operator T : E → E. 2.1. Positive Discrete and Strongly Continuous Semigroups. For a conve- nient simultaneous dealing with discrete and strongly continuous semigroups we use the following definition. Definition 2.1. Let R = N0 or R = [0,∞). A family T = (T (t))t∈R of bounded linear operators on E is called a semigroup on E if T (t + s) = T (t)T (s) for all t, s ∈ R. A semigroup T = (T (t))t∈[0,∞) is called strongly continuous if the mappings t 7→ T (t)x are continuous for all x ∈ E. A semigroup T = (T (t))t∈N0 is said to be discrete. A semigroup T is called positive if T (t) is positive for all t ∈ R and it is called bounded if supt∈R kT (t)k < ∞. IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 3 Given a semigroup T = (T (t))t∈R, a subset A ⊆ E is said to be T -invariant if T (t)A ⊆ A for all t ∈ R. The semigroup T is called irreducible if there are no closed T -invariant ideals in E beside {0} and E. By Fix(T ) := \t∈R ker(I − T (t)) we denote the fixed space of a semigroup T . It will be convenient to call a single operator T irreducible (power bounded) if the discrete semigroup (T n)n∈N0 is irreducible (bounded). By EC we denote the complexification of E, which is called a complex Banach lattice (see [16, Sec. 2.2]), and by TC : EC → EC the complexification of a linear operator T . If T = (T (t))t∈[0,∞) is a strongly continuous semigroup on E with generator A, then so is TC = (T (t)C)t∈[0,∞) with generator AC. We recall that a strongly continuous semigroup with generator A is called Abel bounded if Reλ ≤ 0 for all λ ∈ σ(AC) and supλ>0 kλ(λ − AC)−1k < ∞. A discrete semigroup T = (T n)n∈N0 is called Abel bounded if λ ≤ 1 for all λ ∈ σ(TC) and supλ>1 k(λ − 1)(λ − TC)−1k < ∞. Lemma 2.2. Let T = (T (t))t∈R be a positive and Abel bounded semigroup on E which is strongly continuous or discrete. Assume that T (t)z ≥ z for some z ∈ E, z > 0, and all t ∈ R. Then there exists 0 < x′ ∈ Fix(T ′). If in addition T is irreducible, x′ is strictly positive and z is a quasi-interior point of E+ with Fix(T ) = span{z}. Proof. The existence of 0 < x′ ∈ Fix(T ′) follows from [17, Lem. V 4.8] in the dis- crete case and from the proof of [6, Prop. 4.3.6] for a strongly continuous semigroup, respectively. Now, assume that T is irreducible. Since the absolute null ideal N (x′) := {x ∈ E : hx′,xi = 0} 6= E is closed and T -invariant, it follows that N (x′) = {0}, i.e. x′ is strictly positive. Thus, hx′, T (t)z − zi = 0 for all t ∈ R implies that z ∈ Fix(T ). By the same argument, we observe that T (t)x = x for all t ∈ R and x ∈ Fix(T ), i.e. Fix(T ) is a sublattice of E. Let x ∈ Fix(T ). Then x+ := x ∨ 0 ∈ Fix(T ) and x− := x − x+ ∈ Fix(T ) and the principal ideals generated by x+ and x− are T -invariant and disjoint. The irreducibility of T implies that either x+ is a quasi-interior point of E+ and x− = 0 or vice versa. Consequently, Fix(T ) is totally ordered and hence one-dimensional by [17, Prop. II 3.4], i.e. Fix(T ) = span{z}. (cid:3) Next, we note a consequence of the well-known mean ergodic theorem. Lemma 2.3. Let T : E → E be a positive and Abel bounded linear operator. If there exists a quasi-interior point e ∈ Fix(T ) of E+, then T is mean ergodic, i.e. E = Fix(T ) ⊕ (I − T )E. Proof. For n ∈ N and x ∈ E let C(n)x := 1 averages of T . M := supn∈NkC(n)k < ∞, see [10, 1.5]. Hence, k=0 T kx denote the Cesàro It is well-known that the Abel boundedness of T implies that n Pn−1 (cid:13)(cid:13)(cid:13)(cid:13) T n n (cid:13)(cid:13)(cid:13)(cid:13) ≤ 2(cid:13)(cid:13)(cid:13)(cid:13) T n n + 1(cid:13)(cid:13)(cid:13)(cid:13) for all n ∈ N. Now, it follows from lim 1 for all x ∈ E. ≤ 2kC(n + 1)k ≤ 2M n T nx = 0 for all x ∈ Ee that lim 1 n T nx = 0 4 MORITZ GERLACH Due to the order continuity of the norm on E, for every c > 0 the C(n)-invariant order interval [−ce, ce] is σ(E, E′)-compact, see [16, Thm. 2.4.2]. Thus, the se- quence (C(n)x) has a weak cluster point for every x ∈ Ee and the mean ergodic theorem [15, §2 Thm. 1.1] yields that lim C(n)x exists for all x ∈ Ee. Consequently, by the uniform boundedness of the operators C(n), lim C(n)x exists for all x ∈ E. The desired ergodic decomposition E = Fix(T ) ⊕ (I − T )E now follows from [15, §2 Thm. 1.3]. (cid:3) We will use the following version of the famous splitting theorem by Jacobs, de Leeuw and Glicksberg. Let Lσ(E) denote the space of all bounded linear operator on E endowed with the weak operator topology. Theorem 2.4 (Jacobs -- de Leeuw -- Glicksberg). Let T = (T (t))t∈R be a positive and bounded semigroup on E and denote by S the closure of {T (t) : t ∈ R} in Lσ(E). If there exists a quasi-interior point e ∈ Fix(T ) of E+, then S contains a positive projection Q commuting with every operator in S such that the closed subspace F := QE includes Fix(T ) and QS := {QT : T ∈ S } ⊆ S is a norm bounded group of positive operators with neutral element Q. If in addition T is irreducible, Q is strictly positive in the sense that Qy > 0 for all y > 0. Proof. It follows from the uniform boundedness principle that the closure of a norm bounded set in Lσ(E) is bounded in norm, again. Moreover, the operators of S are positive and they commute since T is Abelian. Now, let x ∈ Ee, i.e. x ∈ [−ce, ce] for some c > 0. Then {T (t)x : t ∈ R} is a subset of [−ce, ce] and hence, by the order continuity of the norm [16, Thm. 2.4.2], relatively weakly compact. Since Ee is dense in E, {T (t) : t ∈ R} is relatively compact in Lσ(E) [11, Cor. A 5]. Now, it follows from [15, §2 Thm 4.1] that there exists a (positive) projection Q ∈ S that commutes with every operator in S and QS is a (bounded) group with neutral element Q. Since Q is in the Lσ(E)-closure of T , one has that Fix(T ) ⊆ Fix(Q) ⊆ F . If in addition T is irreducible, then Q is strictly positive because the closed ideal N (Q) := {y ∈ E : Qy = 0} is T -invariant and distinct from E. 2.2. Atoms. We recall that an element a ∈ E+ is said to be an atom (of E) if the generated ideal (cid:3) Ea := {x ∈ E : x ≤ ca for some c > 0} is one-dimensional. If E contains no atoms, E is said to be diffuse. Remark 2.5. Let a ∈ E+ be an atom and T : E → E be positive. Then clearly T Ea ⊆ ET a. If in addition T is invertible with a positive inverse, i.e. T is a lattice isomorphism, then ET a = T T −1ET a ⊆ T ET −1T a = T Ea = span{T a}. Hence, every lattice isomorphism on E maps atoms to atoms. The lemma below is a more general formulation of [14, Prop. 3.5], followed by an analogous assertion in the discrete case. Lemma 2.6. Let T = (T (t))t∈R be a positive and bounded strongly continuous group on E, i.e. T is a uniformly bounded family of positive linear operators on E such that T (0) = I, T (t)T (s) = T (t + s) for all t, s ∈ R, and the mappings t 7→ T (t)x are continuous from R to E for every x ∈ E. Then Fix(T ) contains every atom of E. IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 5 Proof. Since T is a group, every operator T (t) is a lattice isomorphism and hence maps atoms to atoms by Remark 2.5. Fix an atom a ∈ E+. Then for every disjoint atom b ∈ E+ one has that a− b = a + b and hence ka− bk ≥ kak. Thus, it follows from limt→0 T (t)a = a that T (t)a ∈ Ea for all t ∈ R. Therefore, T (t)a = exp(λt)a for a λ ∈ R and all t ∈ R by the semigroup law. Since the group T is bounded, we conclude that λ = 0 and hence a ∈ Fix(T ). Lemma 2.7. Suppose that E has atoms and let T be an irreducible lattice isomor- phism on E such that (cid:3) (2.1) Then there exists n ∈ N such that T n = I. Proof. Let a ∈ E+ be an atom. The ideal sup{kT kk : k ∈ Z} < ∞. J := {x ∈ E : x ≤ c(a + T a + ··· + T ma) for some c ≥ 0, m ∈ N} is T -invariant and hence J = E. Since the norm on E is order continuous, J equals J⊥⊥, the band generated by J. Thus, for every x > 0 we find some n ∈ N0 such In particular, T −1a ∧ T n−1a > 0 for some n ∈ N. As T −1 that x ∧ T na > 0. maps atoms to atoms by Remark 2.5, this implies that T n−1a = cT −1a for some c > 0 and T na = ca. Now, it follows from assumption (2.1) that c = 1. Therefore, T ka ∈ Fix(T n) for all k ∈ N0. Since every T ka is an atom, we conclude that E = J = span{a, T a, . . . , T n−1a} ⊆ Fix(T n), which completes the proof. (cid:3) 2.3. Kernel and Harris Operators. First, we recall that a linear operator T : E → E is called regular if it is the difference of two positive operators. Every regular operator is bounded [16, Prop. 1.3.5] and admits a modulus. Moreover, since E is order complete, the regular operators on E form a Banach lattice with respect to the regular norm kTkr := kTkk [16, Prop. 1.3.6]. We denote by E′ ⊗ E the space of all finite rank operators on E. The elements of (E′ ⊗ E)⊥⊥, the band generated by E′ ⊗ E in the space of all regular operators, are called (regular) kernel operators. A regular operator T on E is called a Harris operator if T n 6∈ (E′ ⊗ E)⊥ for some n ∈ N. Note that a positive operator is a Harris operator if and only if some power In the case where E = Lp, the band dominates a non-trivial kernel operator. (E′ ⊗ E)⊥⊥ consists precisely of those operators given by a measurable kernel (see [17, Prop. IV 9.8] or [16, Thm. 3.3.7]). The following well-known lemma states that the kernel operators form an algebra ideal in the regular operators. Lemma 2.8. Let K be a kernel operator. Then KT ∈ (E′ ⊗ E)⊥⊥ and T K ∈ (E′ ⊗ E)⊥⊥ for every regular operator T . Proof. We may assume that K, T ≥ 0. Denote by J the (lattice) ideal of regular operators generated by E′ ⊗ E and let A := [0, K] ∩ J. Then K = sup A by [16, Prop. 1.2.6] and Kx = sup{Ax : A ∈ A } (x ∈ E+), since A is upwards directed. Now, consider the directed set A T := {AT : A ∈ A }. As every operator of A is dominated by a finite rank operator, so is every AT , i.e. A T ⊆ J. Now, it follows from KT x = sup{AT x : A ∈ A } (x ∈ E+) 6 MORITZ GERLACH that KT = sup A T and hence KT ∈ (E′ ⊗ E)⊥⊥. By a similar argument, using in addition the fact that T is order continuous because of the order continuity of the norm, one observes that T K ∈ (E′ ⊗ E)⊥⊥. (cid:3) The following lemma shows that under certain conditions some power of a Harris operator always dominates a non-trivial compact operator. Lemma 2.9. Let T : E → E be a positive and power bounded Harris operator such that Fix(T ) contains a quasi-interior point of E+ and that there exists a strictly positive element in Fix(T ′). Then T n ≥ C > 0 for some n ∈ N and some compact operator C. Proof. After replacing T with a suitable power T m, we may assume that T ≥ K > 0 for some kernel operator K : E → E. Let S := T − K. Since the kernel operators form an algebra ideal, see Proposition 2.8, Kn := T n − Sn ∈ (E′ ⊗ E)⊥⊥ for all n ∈ N. Let e ∈ Fix(T ) be a quasi-interior point of E+ and x′ ∈ Fix(T ′) be strictly positive. It follows from e = T e = Ke + Se > Se that Sn+1e ≤ Sne for all n ∈ N. By the order continuity of the norm, u := limn→∞ Sne ∈ Fix(S) exists and satisfies 0 ≤ u < e. Since x′ is strictly positive, we infer from T u = Ku + Su ≥ u and hx′, T u − ui = 0 that u ∈ Fix(T ). Hence, v := e − u > 0 is a fixed point of T and lim Snv = 0. As the operators Kn are uniformly bounded, it follows from Knv = T nv − Snv → v (n → ∞) that lim K 2 nv = v. In particular, K 2 m 6= 0 for a suitable m ∈ N large enough. that lim C2 k ≤ K 2 kC2 k v = K 2 mvk ≤ kCk(Ckv − Kmv)k + k(Ck − Km)Kmvk It follows from [16, Prop. 1.2.6] that Km = sup([0, Km]∩ J) where J is the ideal generated by E′ ⊗ E. Thus, there exists a (bounded) sequence (Ck) ⊆ [0, Km] ∩ J such that lim Ckx = Kmx for x ∈ {v, Kmv}. Since every Ck is dominated by a finite rank operator, C2 k is compact for every k ∈ N by [16, Cor. 3.7.15]. We conclude from kv − K 2 mv 6= 0. Therefore, 0 < C2 m ≤ T 2m for a suitable k ∈ N. (cid:3) Arendt proved in [2, Kor. 1.26] that compact operators are disjoint from lattice isomorphisms on a diffuse and order complete Banach lattice. This is an essential tool for our spectral analysis in Section 3. In order to be more self-contained, we give a proof of this result for a Banach lattice with order continuous norm. Theorem 2.10. Let V, K : E → E be positive operators on a diffuse Banach lattice E with order continuous norm. If V is a lattice isomorphism and K is compact, then V ∧ K = 0. Proof. First, we show that S := I ∧ K = 0. Aiming for a contradiction, we assume that Sx > 0 for some x ∈ E+. Since E is Archimedean, we find some c > 0 such that w := cSx − x is not negative, i.e. it has a non-trivial positive part. Denote by P the band projection onto {w+}⊥⊥, the band generated by w+. Then 0 < w+ = P w = P (cSx − x) = cP Sx − P x. It follows easily from S ≤ I that S is an orthomorphism, i.e. it commutes with every band projection. Thus, SP x > 0 and hence P x > 0. Since E is diffuse, we are able to construct a sequence (xn) ⊆ E of pairwise disjoint elements, each satisfying 0 < xn < P x, by applying [16, Lem. 2.7.12] inductively. Let Pn be the band projection onto {xn}⊥⊥ and define un := Pnx/kPnxk. As the orthomorphism S is dominated by the compact operator K, S itself is compact by [1, Thm. 16.21]. Hence, after passing to a subsequence, (Sun) converges to some y ∈ E. By the order IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 7 continuity of the norm, the disjoint and order bounded sequence (Pny) converges to zero [16, Thm. 2.4.2]. Now, it follows from kPnxkun = Pnx = PnP x ≤ cPnSP x = cPnSx = ckPnxkSun for all n ∈ N that and hence Pny = lim m→∞ PnSum ≤ lim m→∞ PnScSum = cSPny kunk ≤ ckPnSunk ≤ ckPnSun − Pnyk + ckPnyk ≤ ckSun − yk + c2kSPnyk. for all n ∈ N. The right-hand side tends to zero, which is a contradiction to kunk = 1. Hence, S = I ∧ K = 0. In order to complete the proof, we recall that the mapping T 7→ T V is a lattice homomorphism on the lattice of all regular operators on E, see [1, Thm. 7.4]. Therefore, V ∧ K = (I ∧ KV −1)V and hence vanishes by the first part of the proof since KV −1 is compact. (cid:3) 3. Triviality of the Peripheral Point Spectrum In this section we give the proofs of our main results, which are inspired by [4, Thm. 3.1]. There, the semigroup is assumed to contain a kernel operator, which is a less general hypothesis than ours. Again, let E be a Banach lattice with order continuous norm. Theorem 3.1. Let T = (T (t))t∈[0,∞) be a positive, irreducible and bounded strongly continuous semigroup with generator A. If there exists t0 > 0 such that T (t0) ≥ K > 0 where K is a compact or a kernel operator, then σp(AC) ∩ iR ⊆ {0}. Proof. Let α ∈ R such that iα ∈ σp(AC) ∩ iR. Then T (t)Cz = exp(iαt)z for some z ∈ EC\{0}. Since 0 < z = T (t)Cz ≤ T (t)z for all t ≥ 0, by Lemma 2.2 there exists a strictly positive ϕ ∈ Fix(T ′) and Fix(T ) = span{e}, where e := z is a quasi-interior point of E+. In view of Lemma 2.9, we may now assume that T (t0) ≥ K > 0 for a compact operator K. Denote by S the closure of {T (t) : t ≥ 0} in Lσ(E), the bounded linear operators on E endowed with the weak operator topology. Then, by Theorem 2.4, S contains a strictly positive projection Q commuting with every operator in S such that Fix(T ) ⊆ F := QE and that QS := {QS : S ∈ S } ⊆ S is norm bounded group of positive operators with neutral element Q. Hence, the restriction of the elements of S to F defines a group of lattice isomorphisms. If x ∈ F , then x = Qx ≤ Qx and since ϕ is also a strictly positive fixed point of Q′, it follows that hϕ, Qx − xi = 0 and hence Qx = x. Thus, F is a closed sublattice of E and hence a Banach lattice. Moreover, the norm on F is order continuous because every monotone order bounded sequence in F converges (see [16, Thm. 2.4.2]). Since Q is strictly positive and the quasi-interior point e belongs to F , QKF e > 0. Thus, T (t0)F = QT (t0)F dominates the non-trivial compact operator QKF . If F were diffuse, compact operators would be disjoint from lattice isomorphisms by Theorem 2.10. Thus, there exists at least one atom a ∈ F . Since a ∈ Fix(T ) = span{e} by Lemma 2.6, we obtain that e is an atom of F and consequently, F = Fe = span{e} = Fix(T ). It remains to check that z ∈ FC = QCEC. Then z ∈ Fix(TC) and therefore α = 0. Let w ∈ (EC)′ such that hT (t)Cz, wi = hz, wi = kzk > 0. Since QC is in the closure of TC in Lσ(EC), there exists t ≥ 0 such that hT (t)Cz − QCz, wi < kzk. 8 MORITZ GERLACH This implies that hQCz, wi ≥ hT (t)Cz, wi − hT (t)Cz − QCz, wi > 0 and hence QCz 6= 0. Now, choose w ∈ (EC)′ vanishing on z and let ε > 0. Again, there exists t ≥ 0 such that which implies that hT (t)Cz − QCz, wi < ε hQCz, wi ≤ hT (t)Cz − QCz, wi + hT (t)Cz, wi < ε. Since ε > 0 was arbitrary, hQCz, wi = 0. Thus, QCz ∈ span{z} and consequently QCz = z ∈ FC because QC is a projection. (cid:3) As we will see next, this result remains true for Abel bounded semigroups of Harris operators. Proposition 3.2. Let T = (T (t))t∈[0,∞) be a positive, irreducible and Abel bounded strongly continuous semigroup with generator A. If T (t0) is a Harris operator for some t0 > 0, then σp(AC) ∩ iR ⊆ {0}. Proof. Assume that T (t)Cz = exp(iαt)z for some z ∈ EC\{0}, α ∈ R and all t ≥ 0. Since z ≤ T (t)z for all t ≥ 0, it follows from Lemma 2.2 that there exists a strictly positive linear form ϕ ∈ Fix(T ′). Now, we endow E with the order continuous lattice norm kxkϕ := hϕ,xi (x ∈ E). Let (F,k·kF ) be the completion of (E,k·kϕ), i.e. the closure of E in the bidual (E,k·kϕ)′′. Then T uniquely extends to a positive strongly continuous contraction semigroup T = ( T (t))t∈[0,∞) on F . We show that T is irreducible. Let J be a closed and T -invariant ideal in (F,k·kF ). Then J := J ∩ E is a closed and T -invariant ideal in (E,k·k) and hence J = {0} or J = E. Since ϕ is order continuous, E is an ideal in F by [17, Lem. IV 9.3]. This implies that J is the closure of J in F and hence either J = {0} or J = F . Thus, T is irreducible. Next, we verify that T (t0) is a Harris operator. To simplify notation, we assume that T := T (t0) 6∈ (E′ ⊗ E)⊥; otherwise we replace t0 with nt0 for a suitable n ∈ N. Hence, there exist w′ ∈ E′, w ∈ E and z ∈ E+ such that (T ∧ w′ ⊗ w)z > 0. In view of Nakano's carrier theorem [16, Thm. 1.4.11], the strictly positive ϕ is a weak order unit of E′. Since w′ ⊗ w = sup m∈N ((mϕ ∧ w′) ⊗ w) = sup m∈N ((mϕ ⊗ w) ∧ (w′ ⊗ w)), one has that 0 < T ∧ w′ ⊗ w = T ∧ (w′ ⊗ w) = T ∧ sup{(mϕ ⊗ w) ∧ (w′ ⊗ w) : m ∈ N} = sup{T ∧ (mϕ ⊗ w) ∧ (w′ ⊗ w) : m ∈ N}. Hence, m(T ∧(ϕ⊗w)) ≥ T ∧(mϕ⊗w) > 0 for some m ∈ N. Since the extensions ϕ ⊗ w and T leave the ideal E invariant, we conclude that ( T ∧ ( ϕ ⊗ w))z = inf = inf F { T (z − y) + ϕ(y)w : y ∈ F, 0 ≤ y ≤ z} E {T (z − y) + ϕ(y)w : y ∈ E, 0 ≤ y ≤ z} This shows that T = T (t0) 6∈ (F ′ ⊗ F )⊥. = (T ∧ (ϕ ⊗ w))z > 0. IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 9 Finally, we conclude from Theorem 3.1 that σp(AC) ∩ iR ⊆ σp( AC) ∩ iR ⊆ {0} where A is the generator of T . (cid:3) We continue with a consideration of discrete semigroups, i.e. powers of a single operator. Let Γ := {z ∈ C : z = 1} denote the unit circle. The following simple example shows that one cannot expect that σp(TC) ∩ Γ ⊆ {1} in analogy to Theorem 3.1. Example 3.3. On E = R2, consider the matrix 1 (3.1) T := (cid:18)0 1 0(cid:19) Then T is a positive, irreducible, contractive and compact operator but −1 ∈ σp(TC). However, the following theorem states that the point spectrum of some power of TC is trivial. Theorem 3.4. Let T : E → E be a positive, irreducible and power bounded op- erator. If there exists m ∈ N such that T m ≥ K > 0 for a compact or a kernel operator K, then σp(T n Proof. If σp(TC)∩Γ = ∅, then the assertion follows for n = 1. Thus, we may assume that there exists α ∈ R and z ∈ EC such that z 6= 0 and TCz = exp(iα)z. Since C ) ∩ Γ ⊆ {1} for some n ∈ N. 0 < z = exp(inα)z = T n C z ≤ T nz (n ∈ N), Lemma 2.2 applied to the discrete semigroup T := (T n)n∈N0 yields that there exists a strictly positive ϕ ∈ Fix(T ′) and Fix(T ) = span{e}, where e := z is In view of Lemma 2.9, we may now assume that a quasi-interior point of E+. T m ≥ K > 0 for a compact operator K. As in the proof of Theorem 3.1, we conclude from Theorem 2.4 that there ex- ists a strictly positive projection Q : E → E commuting with T and having the following properties: Its range F := QE is a closed sublattice of E with order con- tinuous norm and includes Fix(T ). Moreover, the restriction of T to F is a lattice isomorphism with sup{kT k F dominates the non-trivial compact operator QKF and it follows from Theorem 2.10 that F is not diffuse. Next, we show that TF is irreducible. Let J ⊆ F be a closed and T -invariant ideal with corresponding band projection P : F → J and let e′ := P e. Since T e′ ∈ J, we observe that Fk : k ∈ Z} < ∞. On the other hand, T m T e′ = T P e = P T P e ≤ P T e = P e = e′. Now, it follows from hϕ, e′ − T e′i = 0 that e′ ∈ Fix(T ) = span{e} because ϕ is strictly positive. Hence, e′ = 0 or e′ = e which implies that either J = {0} or J = F . Now, we conclude from Lemma 2.7, applied to the restriction of T to F , that there exists some n ∈ N such that F ⊆ Fix(T n). Let exp(iβ) ∈ σp(T n C ) and denote by ξ1, . . . , ξn the nth roots of exp(iβ), i.e. (ξk)n = exp(iβ) for all 1 ≤ k ≤ n. Then we infer from exp(iβ) − T n C = (ξ1 − TC)(ξ2 − TC) . . . (ξn − TC) that ξk ∈ σp(TC) for at least one 1 ≤ k ≤ n. Pick y ∈ EC\{0} such that TCy = ξky. Now, we observe that y ∈ FC = QCEC by the same arguments as in the proof of Theorem 3.1. Indeed, since QC is in the closure of TC = (T k C)k∈N0 in Lσ(EC), for w ∈ (EC)′ satisfying hw, yi = kyk we find some j ∈ N0 such that hT j Cy − QCy, wi < kyk. 10 MORITZ GERLACH Cy, wi = ξj As hT j khy, wi = kyk, it follows that hQCy, wi > 0 and hence QCy 6= 0. On the other hand, for every w ∈ (EC)′ vanishing on y and for every ε > 0 there exists j ∈ N0 such that hT j Cy − QCy, wi < ε and hence hQCy, wi < ε. Thus, QCy ∈ span{y} which implies that QCy = y ∈ FC since QC is a projection. Altogether, we proved that y = T n C y = ξn k y = exp(iβ)y and hence exp(iβ) = 1. (cid:3) Remark 3.5. (a) The assertion of Theorem 3.4 remains true for every positive, ir- reducible and Abel bounded Harris operator T : E → E. This can be proven analogously to Proposition 3.2. (b) It is well known that σp(TC) ∩ Γ ⊆ {1} for every power bounded operator T : E → E which is strongly positive, i.e. T x is a quasi-interior point of E+ for all x > 0, c.f. [17, Prop. V 5.6]. This can also be obtained from the proof of Theorem 3.4 by observing that such an operator T is irreducible and its restriction to F is again strongly positive and therefore no lattice isomorphism unless F contains atoms. (c) The assumption T being irreducible cannot be dropped in Theorem 3.1 (and Theorem 3.4). Indeed, let (T1(t))t∈[0,∞) be a positive and bounded strongly continuous semigroup on E such that T1(t0) ≥ K > 0 for some compact op- erator K and some t0 > 0. Now, let (T2(t))t∈[0,∞) be another positive and bounded strongly continuous semigroup on E such that iα is an eigenvalue of its generator for some α 6= 0. Then T (t) := T1(t) ⊕ T2(t) defines a positive and bounded strongly continuous semigroup on E ⊕ E where T (t0) dominates the compact operator K ⊕ 0 but iα is an eigenvalue of its generator. (d) If A is the generator of a bounded and positive strongly continuous semigroup T = (T (t))t∈[0,∞) such that T (t0) is compact for some t0 > 0, it is easy to see that σ(AC) ∩ iR ⊆ {0}. Indeed, in this case T is eventually norm continuous [11, Lem. II 4.22] and hence σ(AC)∩iR is bounded [11, Thm. II 4.18]. Since the boundary spectrum is cyclic by [12, Thm. 2.4], it follows that σ(AC)∩iR ⊆ {0}. It is natural to ask if even σ(T n C ) ∩ Γ is trivial for some n ∈ N under the assumptions of Theorem 3.4. The following example shows that this is not the case. Example 3.6. Fix 1 < p < ∞ and let E := ℓp. We construct a positive and irreducible contraction T on E such that σp(TC) = ∅ and σ(T n C ) ∩ Γ = Γ for all n ∈ N. Since E is atomic, T is automatically a kernel operator and hence satisfies the assumptions of Theorem 3.4. Let (bn) ⊆ [0, 1] be a decreasing sequence with Q∞n=1(1 − bn) = 1 2 and let 0 ≤ an ≤ 1 − (1 − bn)p be small enough such that P∞n=1 an ≤ 1. For x = (xn) ∈ ℓp we define T x :=(cid:18) ∞ Xn=1 anxn, (1 − b1)x1, (1 − b2)x2, (1 − b3)x3, . . .(cid:19). If x ≥ 0, then it follows from Jensen's inequality that kT xkp =(cid:18) ∞ Xn=1 anxp (1 − bn)pxp anxn(cid:19)p Xn=1 n ≤ ∞ Xn=1 n + ∞ + ∞ Xn=1 (1 − bn)pxp n ≤ kxkp which shows that T is a contraction. It is well-known that the complexification of E equals ℓp(C), the p-summable sequences in C. Now, assume that there is λ ∈ σp(TC) and denote by z = (zn) ∈ EC, IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 11 z 6= 0, a corresponding eigenvector. Since TC is injective and contractive, 0 < λ ≤ 1 and it follows by induction that zn+1 = 1 λn (1 − b1)(1 − b2) . . . (1 − bn)z1 for all n ∈ N. Hence, z is not a null sequence. This is impossible and therefore σp(TC) = ∅. In order to calculate the peripheral spectrum of TC, we first point out that the discrete semigroup (T n C )n∈N0 is not strongly stable. Indeed, for e1 = (1, 0, 0, . . . ) we have that n kT ne1k ≥ Yk=1 (1 − bk) ≥ 1 2 for all n ∈ N. Thus, it follows from the characterization of strong stability by Arendt, Batty, Lyubich and Vu [5, Thm. 5.1] that σ(TC) ∩ Γ is uncountable. Since σ(TC) ∩ Γ is cyclic [17, Thm. V 4.9], it is dense in Γ and hence σ(TC) ∩ Γ = Γ. By factorizing λ−T n C )∩Γ = Γ for all n ∈ N. C as in the proof of Theorem 3.4 one observes that also σ(T n 4. Strong Convergence of the Semigroup The asymptotic behavior of a bounded strongly continuous semigroup T = (T (t))t∈[0,∞) on a complex Banach space X is highly related to the peripheral (point) spectrum of its generator A. If, for instance, T is known to be asymptotically almost periodic, i.e. X = X0 ⊕ XAP where X0 := {x ∈ X : lim t→∞kT (t)xk = 0} and XAP := span{x ∈ D(A) : Ax = iαx for some α ∈ R}, then σp(A) ∩ iR ⊆ {0} already implies that limt→∞ T (t)x exists for all x ∈ X. This is the case if the generator A has compact resolvent [6, Prop. 5.4.7] or if the peripheral spectrum σ(A)∩ iR is countable and T is totally ergodic [6, Thm. 5.5.5]. In the following, we prove a convergence result for semigroups of Harris operators with a non-trivial fixed space by adjusting techniques developed by Greiner to arbitrary operator semigroups that might by discrete or strongly continuous. Our main tool is Greiner's zero-two law, see Theorem A.1 in the appendix. Again, let E be a Banach lattice with order continuous norm. We start with the following proposition, a generalized version of [13, Kor. 3.9], that yields sufficient conditions for strong convergence of a semigroup. Proposition 4.1. Let T = (T (t))t∈R be a positive, bounded and irreducible semi- group on E which is strongly continuous or discrete such that Fix(T (t)) = Fix(T ) 6= {0} for all t ∈ R and assume that there are r, s ∈ R, r > s, such that T (r)∧T (s) > 0. Then lim t→∞ T (t)x = hx′, xie (x ∈ E) for some strictly positive x′ ∈ Fix(T ′) and a quasi-interior point e ∈ Fix(T ) of E+. Proof. By Lemma 2.2, there exists a strictly positive x′ ∈ Fix(T ′) and a quasi- interior point e of E+ with Fix(T ) = span{e}. We may assume that hx′, ei = 1. By assumption, E2 := {y ∈ E : (T (t) ∧ T (t + τ ))y = 0 for all t ≥ 0} 6= E 12 MORITZ GERLACH for τ := r − s. Since T is irreducible, it follows from Greiner's zero-two law, Theorem A.1, that E2 = {0} and lim t→∞T (t) − T (t + τ )e = 0. Hence, T (t)(I − T (τ ))y ≤ T (t) − T (t + τ )e → 0 (t → ∞) for all y ∈ [−e, e], i.e. lim T (t)z = 0 for all z ∈ D := (I − T (τ ))[−e, e]. As D is total in (I − T (τ ))E and T is bounded, limt→∞ T (t)x = 0 for all x ∈ (I − T (τ ))E. By Lemma 2.3, T (τ ) is mean ergodic, i.e. E = Fix(T (τ )) ⊕ (I − T (τ ))E. Since Fix(T (τ )) = span{e}, the corresponding mean ergodic projection is given by x′ ⊗ e which completes the proof. (cid:3) It seems to be rather technical to assume that two operators T (r) and T (s) are disjoint. However, by Theorem B.2 due to Axmann, this holds if the semigroup contains an irreducible Harris operator. This leads us to the following theorems, our main results in this section. Theorem 4.2. Let T = (T (t))t∈[0,∞) be a positive, bounded, irreducible and strongly continuous semigroup on E with generator A such that Fix(T ) 6= {0}. If T (t0) is a Harris operator for some t0 > 0, then lim t→∞ T (t)x = hx′, xie (x ∈ E) for some strictly positive x′ ∈ Fix(T ′) and a quasi-interior point e ∈ Fix(T ) of E+. Proof. By Lemma 2.2, Fix(T ) = span{e} for a quasi-interior point e of E+ and there exists a strictly positive element x′ ∈ Fix(T ′). Hence, Theorem 3.1 implies that σp(AC) ∩ iR = {0} and therefore (4.1) (see [11, Cor. IV 3.8]). Next, we prove that T := T (t0) is irreducible. Let J ⊆ E be a T -invariant closed ideal and P : E → J the corresponding band projection. Then T P e ∈ J and hence T P e = P T P e ≤ P T e = P e. Since hx′, P e − T P ei = 0 and x′ is strictly positive, it follows that P e ∈ Fix(T ) = span{e} which implies that J = {0} or J = E. Therefore, it follows from Theorem B.2 that there exist some natural numbers n < m such that T n ∧ T m > 0. Now, the assertion follows from Proposition 4.1. Fix(T (t)) = Fix(T ) = span{e} (t > 0) (cid:3) In the discrete case, we obtain strong convergence of a subsequence (T nk)k∈N for a fixed n ∈ N which is optimal in view of Example 3.3. If the operator is not only irreducible but even strongly positive, the sequence (T k)k∈N itself converges strongly to a projection of rank one. Theorem 4.3. Let T : E → E be a positive, power bounded and irreducible Harris operator with Fix(T ) 6= {0}. Then there exists n ∈ N such that T nk converges strongly as k tends to infinity. If T is even strongly positive, i.e. T x is a quasi-interior point of E+ for all x > 0, then limk→∞ T kx = hx′, xie for some strictly positive x′ ∈ Fix(T ′) and a quasi-interior point e ∈ Fix(T ) of E+. Proof. By Theorem B.2, there are natural numbers a < b such that T a ∧ T b > 0. Let n := b − a. Now, we conclude as in the proof of Proposition 4.1 that E = Fix(T n) ⊕ (I − T n)E IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 13 and limk→∞ T kx = 0 for all x ∈ (I − T n)E. If T is strongly positive, then σp(TC) ∩ Γ ⊆ {1} by Remark 3.5 (b). There- fore, Fix(T k) = Fix(T ) for all k ∈ N and the assertion follows immediately from Proposition 4.1. (cid:3) Remark 4.4. It is natural to ask whether Theorem B.2 holds true for positive and irreducible operators that merely dominate non-trivial compact operators, which would allows us to generalize Theorem 4.2 and 4.3 accordingly. This is not the case since there exists a positive, compact and irreducible operator T on L2(Γ), the square-integrable functions on the unit circle endowed with the Lebesgue measure, such that T n ∧ T m = 0 for all n 6= m. In fact, based on a work of Varopoulos [18], Arendt constructed a self-adjoint and compact Markov operator T on L2(Γ) such that T n ∧ T m = 0 whenever n 6= m [3, Ex. 3.7]. By Theorem 4.5 below, there exists a T -invariant band B in L2(Γ) such that the restriction of T to B is irreducible and still compact. Since B is of the form {f ∈ L2(Γ) : f = 0 on A} for some measurable A ⊆ Γ by [17, III §1 Ex.2] and Γ\A is not a nullset, B is in turn isomorphic to L2(Γ) by [8, Cor. 6.6.7 and Thm. 9.2.2]. Theorem 4.5. Let T : E → E be a positive and compact operator. If there exist a quasi-interior point e ∈ Fix(T ) of E+ and a strictly positive ϕ ∈ Fix(T ′), then there are finitely many disjoint T -invariant bands B1, . . . , BN ⊆ E distinct from {0} such that E = B1 ⊕ ··· ⊕ BN and the restriction of T to Bk is irreducible for all k = 1, . . . , N . Proof. We assume that there is no T -invariant band A ⊆ E except {0} such that the restriction of T to A is irreducible. Then, in particular, T is not irreducible on E and hence there is a T -invariant closed ideal A1 distinct from {0} and E. As the norm on E is order continuous, every closed ideal is a projection band. Denote by P1 : E → A1 the corresponding band projection. Since T P1e = P1T P1e ≤ P1T e = P1e and hP1e − T P1e, ϕi = 0, the strict positivity of ϕ implies that T P1e = P1e. By linearity, T (I−P1)e = (I−P1)e and thus (I−P1)E = A⊥1 is a non-trivial T -invariant band, too. We may assume that hϕ, ei = 1 and hP1e, ϕi ≤ 1 2he, ϕi; otherwise we replace A1 with A⊥1 and P1 with I − P1. A2 ⊆ A1 distinct from {0} and A1 with band projection P2 : E → A2 such that By our assumption, TA1 is not irreducible and hence we find a T -invariant band hP2e, ϕi ≤ 1 2hP1e, ϕi. Inductively, we obtain a decreasing sequence An+1 ⊆ An of T -invariant bands with projections Pn : E → An satisfying hPne, ϕi ≤ 2−nhe, ϕi (n ∈ N). Since ϕ is strictly positive and the norm on E is order continuous, this implies that lim Pne = inf Pne = 0. Now, consider the sequence xn := Pne/kPnek. The compactness of T yields that a subsequence of (T xn) = (xn) converges to some x with kxk = 1. On the other hand, x ∈ An for every n ∈ N because xk ∈ An whenever k ≥ n. Thus, e = lim(I − Pn)e ∈ {x}⊥ which implies that x = 0, a contradiction. We proved the existence of a T -invariant band {0} 6= B1 ⊆ E such that TB1 is irreducible. If B1 6= E we may apply the same argument to the restriction of T to B⊥1 to obtain a T -invariant band {0} 6= B2 ⊆ B⊥1 such that TB2 is irreducible. 14 MORITZ GERLACH Continue this construction inductively as long as B1⊕···⊕Bn 6= E. Suppose that this process does not terminate after finitely many steps, i.e. we obtain an infinite sequence Bn of disjoint non-trivial bands such that TBn is irreducible. Denote by Qn : E → Bn the corresponding band projections and let yn := Qne/kQnek. Then a subsequence of T yn = yn converges to some y ∈ E with kyk = 1 since T is compact. On the other hand, Qky = limn→∞ Qkyn = 0 for every k ∈ N implies that y ∈ B⊥k for all k ∈ N. This shows that every yk is contained in {y}⊥ and so is y. Hence, y = 0 contradicting kyk = 1. We conclude that the process of constructing B1, B2, . . . ends after finitely many steps, which completes the proof. (cid:3) Appendix A. Greiner's zero-two law We present the proof of Greiner's zero-two law from [13] in a reformulation for Banach lattices with order continuous norm and without any continuity condition on the semigroup. Throughout, let T = (T (t))t∈R be a positive and bounded semigroup on E, a Banach lattice with order continuous norm, and fix τ > 0. Theorem A.1 (Greiner's zero-two law). Assume that Fix(T ) contains a quasi- interior point e of E+ and that there exists a strictly positive element in Fix(T ′). Then E2 := {y ∈ E : (T (t) ∧ T (t + τ ))y = 0 for all t ∈ R} and E0 := E⊥2 are T -invariant bands. Moreover, if P denotes the band projection onto E2, then and T (t) − T (t + τ )P e = 2P e for all t ∈ R lim t→∞T (t) − T (t + τ )(I − P )e = 0. To simplify notation, for t ∈ R we define the positive operators S(t) := T (t) ∧ T (t + τ ) and D(t) := T (t) − T (t + τ ) on E. It follows immediately that (A.1) S(t)x + 1 2 D(t)x = x for all x ∈ Fix(T ) and t ∈ R. Further properties of S(t) and D(t) are provided by the following lemma. Lemma A.2. Let x′ ∈ Fix(T ′) be strictly positive and τ ∈ R. Then the following assertions hold. (a) D(t)T (s) ≥ D(t + s) and T (s)D(t) ≥ D(t + s) for all t, s ∈ R. Moreover, if (b) S(t)T (s) ≤ S(t + s) and T (s)S(t) ≤ S(t + s) for all t, s ∈ R. Moreover, if (c) If limt→∞ S(t)x > 0 for all 0 < x ∈ Fix(T ), then limt→∞ S(t)mx > 0 for all x ∈ Fix(T ), then limt→∞ D(t)x ∈ Fix(T ). x ∈ Fix(T ), then limt→∞ S(t)x ∈ Fix(T ). m ∈ N and x ∈ Fix(T ), x > 0. Proof. (a) For all t, s ∈ R one observes that D(t)T (s) = T (t) − T (t + τ ) · T (s) ≥ (T (t) − T (t + τ ))T (s) = D(t + s) and similarly that T (s)D(t) ≥ D(t + s). Let x ∈ Fix(T ), x ≥ 0. Then D(t)x = D(t)T (s)x ≥ D(t + s)x IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 15 for all t, s ∈ R. Hence, by the order continuity of the norm, y := lim D(t)x exists in E and T (s)y = lim t→∞ T (s)D(t)x ≥ lim t→∞ D(t + s)x = y ≥ 0. Finally, we conclude from hx′, T (s)y− yi = 0 that y ∈ Fix(T ) because x′ is strictly positive. (b) For all t, s ∈ R one observes that S(t)T (s) = ≤ 1 2 1 2 (T (t + s) + T (t + τ + s) − D(t)T (s)) (T (t + s) + T (t + τ + s) − D(t + s)) = S(t + s) and similarly that T (s)S(t) ≤ S(t + s). Hence, S(t)x is increasing for all positive x ∈ Fix(T ). Since 0 ≤ S(t) ≤ 1 2 (T (t) + T (t + τ )), we conclude as in the proof of part (a) that lim S(t)x exists in Fix(T ) for all x ∈ Fix(T ). (c) Let x ∈ Fix(T ), x > 0, and define recursively xk := limt→∞ S(t)xk−1 for all k ∈ N where x0 := x. Then xk ∈ Fix(T ) by part (b) and xk > 0 by assumption. It follows by induction that S(t)mx − xm = m Xj=1 S(t)m−j(S(t)xj−1 − xj ) for all m ∈ N and t ∈ R. As kS(t)k ≤ supt∈R kT (t)k, we conclude that S(t)mx = xm > 0 lim t→∞ for all m ∈ N. (cid:3) The key for the proof of the zero-two law is the following combinatorial lemma. 2 √m . m Xj=1 + 2(cid:19) ≤ (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) j − 1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)m j (cid:19) −(cid:18) m Lemma A.3. For every m ∈ N 2−m(cid:18) m j − 1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)m j (cid:19) −(cid:18) m Xj=1 Proof. For k ∈ N and m = 2k − 1 we have k−1 (cid:18)(cid:18)m j (cid:19) −(cid:18) m j − 1(cid:19)(cid:19) + 2 Xj=1 = 2(cid:18) m k − 1(cid:19) = (cid:18)2k k (cid:19) (cid:18)(cid:18)m + 1 j (cid:19) −(cid:18)m + 1 j − 1(cid:19)(cid:19) + 1 Xj=1 (cid:12)(cid:12)(cid:12)(cid:12) j − 1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)m + 1 j (cid:19) −(cid:18)m + 1 + 2 = 2 1 2 m+1 Xj=1 + 1. k = = It follows from Stirling's formula that the central binomial coefficient can be esti- mated by (cid:18)2k k (cid:19) ≤ 22k √2k (k ∈ N). 16 MORITZ GERLACH Thus, we obtain that 2−m(cid:18) m Xj=1 (cid:12)(cid:12)(cid:12)(cid:12) j − 1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)m j (cid:19) −(cid:18) m which completes the proof. + 2(cid:19) = 2−(m+1)(cid:18) m+1 (cid:12)(cid:12)(cid:12)(cid:12) Xj=1 k (cid:19) ≤ = 2−m(cid:18)2k (cid:18)m + 1 j − 1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) j (cid:19) −(cid:18)m + 1 2 √m + 1 ≤ 2 √m , + 2(cid:19) (cid:3) Proof of Theorem A.1. By Lemma A.2 (b), we have S(t)T (s)y ≤ S(t)T (s)y ≤ S(t + s)y = 0 for all y ∈ E2 and t, s ∈ R, which shows that E2 is T -invariant. Let x′ ∈ Fix(T ′) be strictly positive. Since 0 ≤ T (t)P e = P T (t)P e ≤ P T (t)e = P e and hx′, P e − T (t)P ei = 0 for all t ∈ R, it follows that P e ∈ Fix(T ). Define e0 := (I − P )e ∈ E0 and e2 := P e ∈ E2. As e0 = e− e2 ∈ Fix(T ), we conclude that E0, which equals the closure of the principle ideal generated by e0, is T -invariant. It follows immediately from (A.1) that D(t)e2 = 2e2 for all t ∈ R. Hence, it remains to show that lim D(t)e0 = 0. For simplicity, we omit the index 0 and write E = E0, e = e0, T (t) = T (t)E0 and so on. Now, lim S(t)y > 0 for all y ∈ Fix(T ), y > 0, by Lemma A.2 (b) and the definition of E0. Assume that h := lim D(t)e > 0. As h ≤ 2e, there exists m ∈ N such that k :=(cid:18)h − 2 √m e(cid:19)+ > 0. Moreover, k ∈ Fix(T ) since the fixed space is a sublattice. Indeed, if y ∈ Fix(T ), it follows from T (t)y ≥ T (t)y = y and hx′, T (t)y − yi = 0 for all t ∈ R that y ∈ Fix(T ) because x′ is strictly positive. Now, Lemma A.2 (c) yields that S(t0)mk > 0 for some t0 ∈ R. Let t1 := m(t0 + τ ) and define the operator (A.2) U := T (t1) − 2−mS(t0)m(cid:18)I + T (τ )(cid:19)m . It follows from S(t0)(I + T (τ )) ≤ T (t0 + τ ) + T (t0)T (τ ) = 2T (t0 + τ ) that U is positive. Moreover, (A.3) T (nt1) = U n + Rn2−m(cid:18)I + T (τ )(cid:19)m for all n ∈ N where R1 := S(t0)m and Rn+1 := U nR1 + RnT (t1). We infer from e = U ne + Rne that 0 ≤ Rne ≤ e and 0 ≤ U ne ≤ e for all n ∈ N. Now, by Lemma A.2 (a) and Lemma A.3, we obtain that h ≤ D(nt1)e = T (nt1)(I − T (τ ))e Xj=0 e m m j (cid:19)T j(τ )(I − T (τ ))(cid:12)(cid:12)(cid:12)(cid:12) ≤ 2U ne + Rn2−m(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)m j − 1(cid:19)T j(τ )(cid:12)(cid:12)(cid:12)(cid:12) = 2U ne + Rn2−m(cid:12)(cid:12)(cid:12)(cid:12) (cid:18) m (cid:18)m j (cid:19)T j(τ ) − Xj=1 Xj=0 (cid:12)(cid:12)(cid:12)(cid:12) j − 1(cid:19)(cid:12)(cid:12)(cid:12)(cid:12) (cid:18)m j (cid:19) −(cid:18) m e(cid:19) = 2U ne + Rn2−m(cid:18)2e + m+1 m e ≤ 2U ne + Rn 2 √m Xj=1 e ≤ 2U ne + 2 √m e IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 17 for every n ∈ N. Let y := limn→∞ U ne ≥ 0. Then h ≤ 2y + 2√m e and hence 0 < k =(cid:18)h − 2 √m e(cid:19)+ ≤ 2y. Since y is a fixed point of U , equation (A.3) yields that T (nt1)y ≥ y ≥ 0 and we conclude from hx′, T (nt1)y− yi = 0 that T (nt1)y = y for every n ∈ N. By equation (A.2) we have 0 = S(t0)m(cid:18)I + T (τ )(cid:19)m y ≥ S(t0)my ≥ 0. Therefore, 0 < S(t0)mk ≤ 2S(t0)my = 0 which contradicts the preceded observation that S(t0)mk > 0. Hence, h = lim D(t)e = 0. (cid:3) Appendix B. Axmann's theorem We give a proof Axmann's theorem from [7, Satz 3.5] stating that not all powers of an irreducible Harris operator can be disjoint. A proof of this for E = Lp can be found in [4, Sec. 6]. We start with a version for L-spaces and reduce the general case to it in what follows. Let us recall that a Banach lattice E is said to be a L-space if kx + yk= kxk + kyk holds for all x, y ∈ E+. Proposition B.1. Let T be a positive and irreducible operator on a L-space E such that T 6∈ (E′ ⊗ E)⊥. Then there is n ∈ N, n ≥ 2, such that T ∧ T n > 0. Proof. Aiming for a contradiction, we assume that T ∧ T n = 0 for all n ≥ 2. Since E is a L-space, we may identify E′ with C(K) for some compact space K by Kakutani's theorem [16, Thm. 2.1.3]. For n, m ∈ N, n ≥ 2, define An := {T ′h + T ′n(1 − h) : h ∈ C(K), 0 ≤ h ≤ 1} ⊆ C(K) and It follows from our assumption and Synnatschke's theorem [16, Prop. 1.4.17] that On,m := {t ∈ K : h(t) < 1/m for some h ∈ An}. inf An = (T ′ ∧ T ′n)1 = (T ∧ T n)′1 = 0 for all n ≥ 2. Now, we show that each of the open sets On,m is dense in K. Assume the opposite. Then there exists a non-empty open set U ⊆ K\On,m for some n, m ∈ N. By Urysohn's theorem, we can construct a continuous function g : K → [0, 1 m for some t0 ∈ U . Hence, g > 0 is a lower bound of An, which is impossible. Therefore, every On,m and, by Baire's theorem, also G := ∩n,mOn,m is dense in K. Note that for all t ∈ G and n ≥ 2 one has that m ] vanishing on On,m such that g(t0) = 1 hT ′′δt ∧ T ′′nδt, 1i = inf{h(t) : h ∈ An} = 0 and consequently T ′′δt ∧ T ′′nδt = 0. By assumption, there are x ∈ E+ and y′ ∈ E′+ such that T is not disjoint from R = y′ ⊗ x. Then R′ = x ⊗ y′ corresponds to a rank-one operator µ ⊗ g on C(K) for some µ ∈ C(K)′+ and g ∈ C(K)+. Since R′ ∧ T ′ ≥ (R ∧ T )′ > 0 there exists some e ∈ C(K)+, such that := (R′ ∧ T ′)e > 0. For all 0 ≤ h ≤ e and t ∈ K it follows from that = (R′ ∧ T ′)h + (R′ ∧ T ′)(e − h) ≤ R′h + T ′(e − h) (t) ≤ hR′h, δti + he − h, T ′′δti = g(t)hµ, hi + he − h, T ′′δti. (cid:3) Finally, we consider w := T v. If w = 0, then the closed ideal Ev is T -invariant for all n ∈ N because t ∈ G. and non-trivial since T 6= 0. If w > 0, then the closure of the T -invariant ideal J := {z ∈ E : z ≤ c(w + T w + ··· + T kw) for some c > 0 and m ∈ N} 6= 0. Then z′ is non-trivial because w ∈ J and v ∈ J⊥. In both cases, T cannot be irreducible. Thus, we conclude that T ∧ T n > 0 for some n ≥ 2. Theorem B.2. Let T be a positive and irreducible operator on E, a Banach lattice with order continuous norm, such that T 6∈ (E′ ⊗ E)⊥. Then there is n ∈ N, n ≥ 2, such that T ∧ T n > 0. Proof. Fix λ > kTk and y′ ∈ E′+, y′ := (λ − T ′)−1y′ satisfies λz′ − T ′z′ = y′ and hence T ′z′ ≤ λz′. This implies that the closed ideal {x ∈ E : hx, z′i = 0} is T -invariant and thus equal to {0}. Therefore, kxkz′ := hz′,xi defines an order continuous lattice norm on E. Let (F,k·kF ) be the completion of (E,k·kz′), i.e. the closure of E in (E,k·kz′)′′. Then F is an L-space and, since kT xkF ≤ λkTkF for all x ∈ E, T uniquely extends to a positive operator T on F . Now, it follows as in the proof of Proposition 3.2 that T is irreducible and T 6∈ (F ′ ⊗ F )⊥. by [17, Lem. IV 9.3]. Thus, for x ∈ E+ and n ∈ N we observe that F { T (x − y) + T ny : y ∈ F, 0 ≤ y ≤ x} E {T (x − y) + T ny : y ∈ E, 0 ≤ y ≤ x} Since the norm on E is order continuous, so is z′ and hence E is an ideal in F ( T ∧ T n)x = inf = inf = (T ∧ T n)x. 18 MORITZ GERLACH Taking the infimum over all 0 ≤ h ≤ e yields that (t) ≤ (g(t)µ ∧ T ′′δt)e for all t ∈ K. Now, fix t ∈ {s ∈ K : (s) > 0} ∩ G which exists since G is dense in K. Then ν := g(t)µ ∧ T ′′δt > 0 because (t) > 0. As ν is dominated by g(t)µ and µ corresponds to x ∈ E, ν itself corresponds to a vector v ∈ E, v > 0, since E is an ideal in E′′. This vector v satisfies v ∧ T nv ≤ T ′′δt ∧ T ′′n+1δt = 0 As E+ is dense in F+, this shows that T ∧ T n = 0 if and only if T ∧ T n = 0. Thus, the assertion follows from Proposition B.1. (cid:3) Acknowledgements The author was supported by the graduate school Mathematical Analysis of Evolution, Information and Complexity during the work on this article and he would like to thank Wolfgang Arendt for many helpful discussions and the anonymous referee for his/her constructive comments. References [1] C. D. Aliprantis and O. Burkinshaw. Positive operators, volume 119 of Pure and Applied Mathematics. Academic Press Inc., Orlando, FL, 1985. [2] W. Arendt. Über das Spektrum regulärer Operatoren. PhD thesis, Eberhard-Karls-Universität zu Tübingen, 1979. [3] W. Arendt. On the o-spectrum of regular operators and the spectrum of measures. Math. Z., 178(2):271 -- 287, 1981. [4] W. Arendt. Positive semigroups of kernel operators. Positivity, 12(1):25 -- 44, 2008. [5] W. Arendt and C. J. K. Batty. Tauberian theorems and stability of one-parameter semigroups. Trans. Amer. Math. Soc., 306(2):837 -- 852, 1988. [6] W. Arendt, C. J. K. Batty, M. Hieber, and F. Neubrander. Vector-valued Laplace transforms and Cauchy problems, volume 96 of Monographs in Mathematics. Birkhäuser Verlag, Basel, 2001. IRREDUCIBLE SEMIGROUPS OF HARRIS OPERATORS 19 [7] D. Axmann. Struktur-und Ergodentheorie irreduzibler Operatoren auf Banachverbänden. PhD thesis, Eberhard-Karls-Universität zu Tübingen, 1980. [8] V. I. Bogachev. Measure theory. Vol. I, II. Springer-Verlag, Berlin, 2007. [9] E. B. Davies. Triviality of the peripheral point spectrum. J. Evol. Equ., 5(3):407 -- 415, 2005. [10] R. Émilion. Mean-bounded operators and mean ergodic theorems. J. Funct. Anal., 61(1):1 -- 14, 1985. [11] K. Engel and R. Nagel. One-parameter semigroups for linear evolution equations, volume 194 of Graduate Texts in Mathematics. Springer-Verlag, New York, 2000. [12] G. Greiner. Zur Perron-Frobenius-Theorie stark stetiger Halbgruppen. Math. Z., 177(3):401 -- 423, 1981. [13] G. Greiner. Spektrum und Asymptotik stark stetiger Halbgruppen positiver Operatoren. Sitzungsber. Heidelb. Akad. Wiss. Math.-Natur. Kl., pages 55 -- 80, 1982. [14] V. Keicher. On the peripheral spectrum of bounded positive semigroups on atomic Banach lattices. Arch. Math. (Basel), 87(4):359 -- 367, 2006. [15] U. Krengel. Ergodic theorems, volume 6 of de Gruyter Studies in Mathematics. Walter de Gruyter & Co., Berlin, 1985. [16] P. Meyer-Nieberg. Banach lattices. Universitext. Springer-Verlag, Berlin, 1991. [17] H. H. Schaefer. Banach lattices and positive operators. Springer-Verlag, New York, 1974. Die Grundlehren der mathematischen Wissenschaften, Band 215. [18] N. T. Varopoulos. Sets of multiplicity in locally compact abelian groups. Ann. Inst. Fourier (Grenoble), 16(fasc. 2):123 -- 158, 1966. [19] M. P. H. Wolff. Triviality of the peripheral point spectrum of positive semigroups on atomic Banach lattices. Positivity, 12(1):185 -- 192, 2008. University of Ulm, Institute of Applied Analysis, 89069 Ulm, Germany E-mail address: [email protected]
1212.2817
1
1212
2012-12-12T13:54:17
Covering $L^p$ spaces by balls
[ "math.FA", "math.GN" ]
We prove that, given any covering of any separable infinite-dimensional uniformly rotund and uniformly smooth Banach space $X$ by closed balls each of positive radius, some point exists in $X$ which belongs to infinitely many balls.
math.FA
math
COVERING Lp SPACES BY BALLS Vladimir P. Fonf ∗, Michael Levin †and Clemente Zanco ‡ 2 1 0 2 c e D 2 1 ] . A F h t a m [ 1 v 7 1 8 2 . 2 1 2 1 : v i X r a Abstract We prove that, given any covering of any separable infinite-dimensional uni- formly rotund and uniformly smooth Banach space X by closed balls each of positive radius, some point exists in X which belongs to infinitely many balls. 2000 Mathematics Subject Classification: Primary 46B20; Secondary 54D20. Key words and phrases: point finite coverings, slices, uniformly rotund spaces, uni- formly smooth spaces ∗Research of the first author was supported in part by Israel Science Foundation, Grant # 209/09 and by the Istituto Nazionale di Alta Matematica of Italy. †Research of the second author was supported in part by Israel Science Foundation, Grant # 836/08 ‡Research of the third author was supported in part by the Ministero dell'Universit`a e della Ricerca Scientifica e Tecnologica of Italy and by the Center for Advanced Studies in Mathematics at the Ben- Gurion University of the Negev, Beer-Sheva, Israel. 1 1 Introduction and statement of the main result In the present paper X always denotes a real separable infinite-dimensional Banach space; by ball in X we mean a closed ball. Let A be a collection of subsets of X. We say that A is point-finite if every point of X belongs to at most finitely many elements of A. A point x of X is said to be a regular point for A if there is a neighborhood of x that meets at most finitely many elements of A; x is said to be a singular point for A otherwise. If every point of X is a regular point for A, we say that A is locally-finite (clearly, that is equivalent to the requirement that no compact set in X meets infinitely many members of A). A is said a covering of X if each point of X belongs to some member of A. The aim of the present paper is to take a step forward to answering the following Question 1.1 Which infinite-dimensional Banach spaces admit point-finite coverings by balls (each of positive radius)? In order to explain how such a question arises, we recall the following two results (the first one is well known as "the Corson's Theorem"). Theorem 1.2 ([Co]) No (infinite-dimensional) reflexive Banach space admits locally fi- nite coverings by bounded closed convex sets. Theorem 1.3 ([MZ]) Any real Banach space X can be covered by bounded closed convex sets, each with nonempty interior, in such a way that no point of X belongs to more than two of them. The family of sets exhibited in the general construction used to prove Theorem 1.3 is very far from being a family of balls in the original norm of X. Moreover, some classical Banach spaces admit point-finite coverings by balls. For instance, it is easy to check that the covering of c0 that can be obtained by translating the unit ball without overlapping interiors is even locally finite. V. Klee proved in [Kl] that the space l1(Γ) for suitable (uncountable) Γ can be covered by translates of its unit ball without overlapping them at all. So Question 1.1 seems to be very natural, though providing it with a complete answer seems not to be an easy matter. A first step in that direction have been recently made with the following Theorem, that excludes Hilbert spaces from the class of spaces Question 1.1 asks for. Even if, looking for spaces outside that class, Hilbert spaces appear as the simplest ones to be considered, up to now no elementary argument for getting such exclusion seems to be available. Theorem 1.4 ([FZ], Theorem 3.2) No covering by balls, each of positive radius, of the infinite-dimensional separable Hilbert space can be point-finite. We refer to [FZ] also for more details and references concerning the subjects involved in the present Introduction. The goal of this paper is to extend Theorem 1.4 to a considerably wider class of spaces. In fact we prove the following 2 Theorem 1.5 (Main result) No covering by balls, each of positive radius, of any infinite- dimensional uniformly rotund and uniformly smooth separable Banach space can be point- finite. It is very well known that those Banach spaces that are uniformly rotund or uniformly smooth are reflexive (in fact, super-reflexive). Moreover, if a Banach space X is uniformly rotund or uniformly smooth, then an equivalent norm can be put on X under which X is both uniformly rotund and uniformly smooth. Among those spaces that are both uniformly rotund and uniformly smooth there are Lp(µ) spaces for any measure µ and p ∈ (1, +∞), so in particular we claim Corollary 1.6 No covering by balls, each of positive radius, of a separable infinite- dimensional Lp(µ) space, 1 < p < ∞, µ any measure, can be point-finite. The proof of Theorem 1.5 we provide here is based on a key result of [FZ]; however, after that, it follows a completely different way than what was used in [FZ] to get Theorem 1.4. In fact the argument there runs as follows. Separable polyhedral Banach spaces are first characterized as those whose unit sphere under some equivalent norm admits a point-finite covering by slices of the unit ball that do not contain the origin. (Recall that a Banach space is called "polyhedral" if the unit ball of any its finite-dimensional subspace is a polytope.) As a consequence, if the unit sphere of some separable Banach space X admits such a covering, then X must be isomorphically polyhedral. It is well known that no (infinite-dimensional) Hilbert space is isomorphically polyhedral. Next, point-finite coverings of the Hilbert space by balls (if any) are easily reduced to point- finite coverings of the unit sphere by balls that do not contain the origin. Finally, to get a contradiction, these coverings are reduced to point-finite coverings of the unit sphere by slices of the unit ball via the following observation: whenever two spheres in an inner product space do not coincide and have nonempty intersection, such an intersection lies in some hyperplane; this hyperplane splits each of the two balls determined by those spheres in two complementary slices. Unfortunately such a situation characterizes inner product spaces (see [Am] (15.17)), so the argument cannot be applied outside that class of spaces. Our argument here in proving Theorem 1.5 has an essential topological component: Corson's Theorem 1.2, which is based on Brouwer's fixed point Theorem, is now our basic tool. We use it in connection with suitable considerations of geometrical nature. Throughout the paper we use standard Geometry of Banach Spaces notation as in In particular, for x ∈ X and r > 0, B(x, r) and S(x, r) respectively denote the [JL]. closed ball and the sphere with center at x and radius r; moreover, B(0, 1) and S(0, 1) are denoted in short respectively by BX and SX. 2 Proof of the main result In order to prove Theorem 1.5, first of all we notice that, X being separable, we can confine ourselves to prove our theorem for countable coverings. (In fact, let {xn} be any 3 sequence dense in X: since each ball has nonempty interior, for some n0 it must happen that xn0 belongs to uncountably many balls.) We start by borrowing from [FZ] the following Proposition. It describes a quite general situation in which a sequence of slices of the unit ball of any separable Banach space cannot be point-finite. Proposition 2.1 ([FZ], Proposition 2.1) Let {fi}∞ functionals on X and {αi}∞ sequence {Si}∞ i=1 of slices of BX defined by i=1 be a sequence of norm-one linear i=1 a sequence in the interval (0, 1) converging to 0. Then the Si = {x ∈ BX : fi(x) ≥ αi}, i = 1, 2, ... is not point-finite. Next we point out a very simple (probably well known) fact. Roughly speaking, it simply states that any point of a sphere of any uniformly smooth Banach space admits "almost flat" neighborhoods (relative to the sphere) of "big" dia- mater provided that the radius of the sphere is "big". We make this sentence precise in the following way. Fact 2.2 Let X be uniformly smooth. For any ε > 0 there exists b > 0 such that, for any R > b and x ∈ RSX, if Γx is the hyperplane supporting RBX at x, then dist(y, Γx) ≤ ε ∀y ∈ RSX ∩ B(x, 2). (1) Proof. For t ∈ SX, denote by ft the (only) norm-one linear functional such that ft(t) = 1. Fix ε > 0. By definition of uniform smoothness, for any ε > 0 there exists δ = δ(ε) > 0 such that for any w, z ∈ SX with w − z ≤ δ, the following estimate holds 1 − fz(w) = z + (w − z) − 1 − fz(w − z) ≤ εw − z/2. So for x = Rz, y = Rw ∈ RSX with y − x ≤ min{Rδ, 2} it is true that dist(y, Γx) = R − fz(y) ≤ εy − x/2 ≤ ε. By assuming b = 2/δ we are done. (cid:4) The previous fact allows us to say that, in a uniformly smooth space, a "big" sphere intersecting a "small" ball splits it in two parts that are not "too far" from being slices of the small ball. So Proposition 2.1 leads us to the following Proposition, which will be crucial for our purposes and may have interest by itself. Proposition 2.3 Let B = {B(xn, Rn)}∞ n=1 be a countable collection of balls in a uniformly smooth Banach space X such that Rn goes to infinity with n. If B is not locally finite, then it is not point-finite. 4 Proof. Since B is not locally finite, there exist a point x ∈ X and a sequence {nk} of integers such that dist(x, B(xnk , Rnk) → 0 as k → ∞. (2) Without loss of generality we may assume x = 0 and {nk} = {k}. Moreover, we may assume that 0 /∈ B(xn, Rn) for every n and that Rn > 2 and xn < Rn + 1. For any n, let zn be the point at which the segment [0, xn] meets S(xn, Rn) and let Γn be the hyperplane supporting B(xn, Rn) at zn. Let 0 < βn < 1 and fn ∈ SX ∗ be such that Γn = {t ∈ X : fn(t) = βn}. Because of (2), clearly zn → 0 and βn → 0 as n → ∞. By Fact 2.2, for any i ∈ N big enough there is ni such that dist(y, Γni) < 1/i for every y ∈ S(xni, Rni) ∩ B(zni, 2): hence the set Ti = {t ∈ B(zni, 2) : fni(t) ≥ βni + 1/i} is a slice of B(zni, 2) contained in B(xni, Rni). We can choose ni+1 > ni for every i. Under all our assumptions, for every i big enough (since BX ⊂ B(zni, 2)) the set Si = {t ∈ BX : fni(t) ≥ βni + 1/i} is a slice of BX that is contained in Ti. The sequence {Si} of such slices of BX satisfies the assumptions of Proposition 2.1, so it is not point-finite. Since we have we are done. (cid:4) Si ⊂ Ti ⊂ B(xni, Rni) The following Fact sounds, in some sense, as the converse of Fact 2.2. Fact 2.4 Let X be uniformly rotund and b > 0. Let {Bn = B(xn, Rn)}∞ of balls in X such that Rn > b and x0 /∈ intBn for every n. Put n=1 be a sequence Fn = conv(Bn \ intB(x0, b)) and let Then Rn → ∞ with n. dist(x0, Fn) → 0 as n goes to ∞. (3) Proof. Let us recall that the following statement is equivalent to X being uniformly rotund: (UR) for every β > 0 there exists α = α(β) > 0 such that, for any x ∈ SX and f ∈ SX ∗ with f (x) = 1, the slice {t ∈ BX : f (t) ≥ 1 − α} of BX is contained in B(x, β). Of course we may assume that function α(β) is increasing; it goes to 0 as β does. Without loss of generality may assume x0 = 0. 5 Assume to the contrary that some subsequence {Rnk} of {Rn} is bounded: without loss of generality we may assume that {Rn} itself is bounded. For any n, put εn = xn − Rn ≥ 0 and Bn = B(xn, Rn + εn); moreover, let fn ∈ SX ∗ be such that fn( Bn) ⊂ (−∞, 0]. Let αn = α( α = α(β) is the number defined in condition (UR). We have b Rn + εn ), where {t ∈ Bn : fn(t) ≥ −αn} ⊂ bBX hence (Bn \ intbBX ) ⊂ ( Bn \ intbBX ) ⊂ {t ∈ Bn) : fn(t) ≤ −αn}. This last set is closed and convex, so Fn = conv(Bn \ intbBX ) ⊂ {t ∈ Bn : fn(t) ≤ −αn}. (4) Since fn ∈ SX ∗, (4) implies dist(0, Fn) ≥ αn for every n: that contradicts (3) because {Rn + εn} is bounded, so {αn} = {α( )} is far away from 0. (cid:4) b Rn + εn The following three Lemmas are of a technical nature. The first one is something like a local version of Corson's Theorem 1.2. Lemma 2.5 Let X be reflexive. Let x0 ∈ X, a > b > c > 0 and F a collection of closed convex subsets of X contained in B(x0, a) \ intB(x0, c) such that F covers B(x0, a) \ intB(x0, b). Then F is not locally finite in X. Proof. Without loss of generality we may assume x0 = 0. Suppose by contradiction that F is locally finite. Set Fn = {(a/b)nF : F ∈ F }, n ∈ Z where αF = {αy : y ∈ F }. Denote F ′ = ∪n∈ZFn. Then F ′ covers X \ {0} and 0 is the only singular point of F ′. Split X into two closed half spaces X = X + ∪ X − such that X + ∩ X − is a hyperplane of X, 0 /∈ X + ∩ X − and 0 ∈ X −. Let F + = {X + ∩ F : F ∈ F ′} and let F − be the covering of X − which is a symmetric reflection of F + with respect to the hyperplane X + ∩ X − (made via any line not contained in X + ∩ X −). Then F + ∪ F − is a locally finite covering of X by bounded closed convex sets and we get a contradiction with Corson's Theorem 1.2. (cid:4) Lemma 2.6 Let X be both uniformly rotund and uniformly smooth. Consider a closed hyperplane X ′ ⊂ X and a ball B(x0, a), x0 ∈ X ′, a > 0. Assume that B = {Bn}∞ n=1 is a countable point-finite collection of balls and F = {Fn}∞ n=1 is a countable collection of closed convex sets such that F covers B(x0, a) ∩ X ′, Fn ⊂ Bn ∩ B(x0, a) and x0 /∈ intBn for every n. Then there is a point y ∈ B(x0, a) ∩ X ′, y 6= x0, that is a singular point for F . 6 Proof. Assume that every y ∈ B(x0, a) ∩ X ′, y 6= x0, is a regular point of F . Take any n ∩ Fn ∩ X ′ and b such that a > b > 0 and define B ′ F ′ = {F ′ n ∩ B(x0, c) = ∅ for every n and then our Lemma will follow from Lemma 2.5 applied to the collection F ′ with X being replaced by X ′. n = B ′ n=1. We will show that there is c > 0 such that b > c and F ′ n = conv(Bn \ intB(x0, b)), F ′ n}∞ Assume that such c does not exist. Then any arbitrarily small neighborhood of x0 n for some n so, since x0 /∈ intBn, from Fact 2.4 we deduce that there is a intersect B ′ subsequence of balls in B whose radii go to infinity for which x0 is a singular point. By Proposition 2.3, this contradicts the pointwise finiteness of B. The Lemma is proved. (cid:4) Lemma 2.7 Let X be both uniformly rotund and uniformly smooth. Let B = {Bn}∞ {Bn = B(xn, Rn)}∞ and B# moreover, for every n we have that B# B#. n=1 = 1 = B1 n=1 is a covering of X; n ⊂ Bn and any x0 ∈ intBn is a regular point for n=1 be a countable point-finite covering of X by balls. Put B# n+1 = conv((Bn+1 \ (B1 ∪ . . . ∪ Bn)). Then B# = {B# n }∞ Proof. Assume that, for some n, x0 ∈ IntBn is a singular point for B#. Then for some subsequence {B# i=1 with ni > n it happens that x0 /∈ Bni for every i and B(x0, 1/i) ni}∞ intersects the set B# nj for every j ≥ i. Note that B# ni ⊂ conv(Bni \ Bn) ⊂ Bni. Then B(x0, 1/i) intersects conv(Bnj \ Bn) for j ≥ i. Let b > 0 such that B(x0, b) ⊂ Bn (we assumed x0 ∈ intBn): of course conv(Bnj \ B(x0, b) ⊃ conv(Bnj \ Bn) hence, since and x0 /∈ Bni for every i, from Fact 2.4 we get that Rni −→ ∞ as i −→ ∞. By Proposition 2.3 this contradicts the assumption that B is point-finite. (cid:4) We are now ready to prove our main result. Proof of Theorem 1.5. Assume that B = {Bn}∞ n=1 is a point-finite covering by balls of a uniformly rotund and uniformly smooth Banach space X. Consider the covering B# from Lemma 2.7 and let S be the set of the points that are singular for B#. By Theorem 1.2 we have S 6= ∅ and by Lemma 2.7 we have S ⊂ ∪n∂Bn. Clearly S is closed in X, hence by the Baire cathegory theorem there are x0 ∈ S, a > 0 and Bm such that S∩B(x0, a) ⊂ ∂Bm. Take a closed hyperplane X ′ in X passing through x0 and intersecting Bm only at x0. Then, by applying Lemma 2.6 to the collection F = {B# n ∈ B#} of closed convex sets in X with respect to the hyperplane X ′, we get a contradiction. (cid:4) n ∩ B(x0, a) : B# 7 References [Am] D. Amir, Characterizations of inner product spaces, Operator Theory: Advances and Applications, 20, Birkhauser Verlag, Basel, 1986. [Co] H. H. Corson, Collections of convex sets which cover a Banach space, Fund. Math. 49 (1961), 143-145. [FZ] V. P. Fonf and C. Zanco, Covering the unit sphere of certain Banach spaces by sequences of slices and balls, to appear in Canad. Math. Bull., doi:10.4153/CMB- 2012-027-7. [JL] W. B. Johnson and J. Lindenstrauss, Basic Concepts in the Geometry of Banach Spaces, Handbook of the Geometry of Banach Spaces vol. 1, edited by W.B. Johnson and J. Lindenstrauss, Elsevier Science B.V. (2001), 1-84. [Kl] V. Klee, Dispersed Chebyshev sets and covering by balls, Math. Ann. 257 (1981), 251 -- 260. MR 84e:41036 [MZ] A. Marchese and C. Zanco, On a question by Corson about point-finite coverings, Israel J. Math. 189 (2012), 55 -- 63. Vladimir P. Fonf Department of Mathematics Ben-Gurion University of the Negev 84105 Beer-Sheva, Israel E-mail address: [email protected] Michael Levin Department of Mathematics Ben-Gurion University of the Negev 84105 Beer-Sheva, Israel E-mail address: [email protected] Clemente Zanco Dipartimento di Matematica Universit`a degli Studi Via C. Saldini, 50 20133 Milano MI, Italy E-mail address: [email protected] ph. ++39 02 503 16164 fax ++39 02 503 16090 8
1610.01535
1
1610
2016-10-05T17:21:51
A retract theorem for nilpotent Lie groups
[ "math.FA", "math.OA" ]
Let $G= \exp(\g)$ be a connected, simply connected nilpotent Lie group. We show that for every $G$-invariant smooth sub-manifold $M$ of $g^*$, there exists an open relatively compact subset $\mathcal{M}$ of $M$ such that for any smooth adapted field of operators $(F(l))_{l\in M}$ supported in $G\cdot \mathcal{M}$ there exists a Schwartz function $f$ on $G$ such that $\pi_l(f)= \op_{F(l)}$ for all $l\in M$. This retract theorem can then be used to show that for every Lie group $\G$ of automorphisms of $G$ containing the inner automorphisms of $G$ with locally closed $\G$-orbits in $\g^*$, the proper $\G$-prime two-sided closed ideals of $L^1(G)$ are the kernels of $\G$-orbits in $\hat{G}$.
math.FA
math
A retract theorem for nilpotent Lie groups Ying-Fen Lin, Jean Ludwig and Carine Molitor-Braun Abstract 1 Let G = exp(g) be a connected, simply connected nilpotent Lie group. We show that for every G-invariant smooth sub-manifold M of g∗, there exists an open relatively compact subset M of M such that for any smooth adapted field of operators (F (l))l∈M supported in G · M there exists a Schwartz function f on G such that πl(f ) = opF (l) for all l ∈ M . This retract theorem can then be used to show that for every Lie group G of automorphisms of G containing the inner automorphisms of G with locally closed G-orbits in g∗, the proper G-prime two-sided closed ideals of L1(G) are the kernels of G-orbits in bG. 1 Introduction For a connected, simply connected, nilpotent Lie group G, the description of its spectrum and of the Fourier inversion theorem is due to Kirillov [3], who showed that the dual space bG of G is in one-to-one correspondence with the space g∗/G of co-adjoint orbits of G. R. Howe proved in [2] that for every irreducible unitary representation (π, Hπ) of G and every smooth linear operator a on Hπ there exists a Schwartz function fa on G such that π(fa) = a. He also showed that the mapping a 7→ fa is linear and continuous with respect to the Fr´echet topology of the space B∞(Hπ) of smooth linear operators on Hπ and the Fr´echet space S(G) of Schwartz functions on G. In this paper, we study a version of the Fourier inversion theorem for nilpotent Lie groups. More precisely, we generalise the result of R. Howe's mentioned above by con- structing a continuous retract from the space of adapted smooth kernel functions defined on a smooth G-invariant sub-manifold M of g∗ and supported in a subset G · M of M, where M is a relatively compact open subset of M, into the space S(G). We will prove this result, which we call the retract theorem, by proceeding an induction on the length I of the largest index set I for which (B × g∗)I ∩ M 6= ∅. In order to do so we will apply the variable group techniques developed in [11], which have already been used in [10]. Once we have the retract theorem, we can apply it to study the G-prime ideals of the Banach algebra L1(G). Here G denotes a Lie subgroup of the automorphism group of G with the property that the G-orbits in g∗ are all locally closed. The retract theorem 1keywords: nilpotent Lie group, irreducible representation, co-adjoint orbit, Fourier inversion, retract, compact group action -- 2010 Mathematics Subject Classification: 22E30, 22E27, 43A20 1 implies that the Schwartz functions contained in the kernel of a G-orbit Ω in bG are dense in the L1(G)-kernel of Ω. Using the methods in [7], it follows that every G-prime ideal in L1(G) is the kernel of such an G-orbit Ω. One can for instance use this result for the study of bounded irreducible representations (π, X) of a Lie group G on a Banach space X. Restricting the representation π to the nilradical G, one obtains the kernel ker(πG) of πG in the algebra L1(G). This ideal ker(πG) is then G-prime. If ker(πG) is given as the kernel in L1(G) of a G-orbit G · π0 ⊂ bG for some π0 ∈ bG, then one use π0 to make an analysis of π as Mackey did in the case of unitary representations. For connected, simply connected, nilpotent Lie groups, J. Ludwig showed in [6] that the closed prime ideals of L1(G) coincide with the kernels of the irreducible unitary rep- resentations. In 1984, D. Poguntke studied the action of an abelian compact group K on a nilpotent Lie group [12] and characterised the K-prime ideals as kernels of K-orbits. In [4], R. Lahiani and C. Molitor-Braun identified the K-prime ideals with hull contained in the generic part of the dual space of G for a general compact Lie subgroup K of the automorphism group of G. In [7] and [8], it was shown that for an exponential Lie group G, the G-prime ideals are also kernels of G-orbits. In this way the bounded irreducible Banach space representations of an exponential Lie group could be determined. The paper is organised in the following way: in Section 2 we recall the definition of induced representations and of kernel functions, we explain the notion of variable nilpotent Lie groups and their Lie algebras, of index sets for co-adjoint orbits and of adapted kernel functions on a G-invariant sub-manifold of g∗. In Section 3, we introduce the main theorem of the paper, the Retract Theorem, and in Section 4 we present the proof of the theorem, dividing it into several steps. As an application of the Retract Theorem, in the last section (Section 5) we show that every G-prime ideal in L1(G) is the kernel of a G-orbit. 2 Notations and generalities 2.1 Representations and kernel functions Let G = exp(g) be a connected, simply connected, nilpotent Lie group and g be its Lie alge- bra. All the irreducible unitary representations of G (and hence of L1(G)) are obtained (up to equivalence) in the following way: Let l ∈ g∗ and p = p(l) be an arbitrary polarisation of l in g (a maximal isotropic subalgebra of g for the bilinear form (X, Y ) 7→ hl, [X, Y ]i). Let P (l) = exp(p(l)). The induced representation denoted by πl := indG P (l)χl on the Hilbert space Hl, Hl = L2(G/P (l), χl) := {ξ : G → C; ξ measurable , ξ(gp) = χl(p−1)ξ(g), g ∈ G, p ∈ P (l), kξk2 2 =ZG/P (l) ξ(g)2d g < ∞}, where d g is the invariant measure on G/P (l), is unitary and irreducible. Here χl is the character defined on P (l) by χl(g) = e−ihl,log gi for all g ∈ P (l). Two different polarisations 2 for the same l give equivalent representations. The same is true for the case of two linear forms l and l′ belonging to the same co-adjoint orbit. of Zk, . . . , Zn and lk = lgk for all l ∈ g∗. The polarisation p(l)Z = p(l) := Pn One particular way to obtain a polarisation is the following: Let {Z1, . . . , Zn} denote a Jordan-Holder basis of g, for 1 ≤ k ≤ n, let gk := span{Zk, . . . , Zn} be the linear span k=1 gk(lk) of l in g, with gk(lk) := {U ∈ gk; hl, [U, gk]i ≡ 0}, is called the Vergne polarisation at l with respect to the basis Z1, . . . , Zn. We refer to [1] for more details on the theory of irreducible representations of nilpotent Lie groups. Let πl = indG P (l)χl. The corresponding representation of L1(G), also denoted by πl, is obtained via the formula πl(f )ξ := RG f (x)(cid:0)πl(x)ξ(cid:1)dx, for all ξ ∈ Hl. If f ∈ L1(G), then πl(f ) is a kernel operator, i.e. it is of the form (cid:0)πl(f )ξ(cid:1)(g) =ZG/P (l) F (l, g, u)ξ(u)du, where F is the operator kernel given by F (l, g, u) =ZP (l) f (ghu−1)χl(h)dh for g, u ∈ G. If f is a Schwartz function, then the kernel function F belongs to C∞ and satisfies the covariance relation F (l, gh, g′h′) = χl(h)χl(h′)F (l, g, g′) for h, h′ ∈ P (l) and g, g′ ∈ G, and is a Schwartz function on G/P (l) × G/P (l). 2.2 Group actions Let G = exp(g) be a connected, simply connected, nilpotent Lie group and A be a Lie subgroup of the automorphism group Aut(G) of G acting smoothly on G. This action will be denoted by A × G 7→ G (a, x) 7→ a · x. The action of A on G induces naturally actions of A on g, g∗, bG, L1(G), and on S(G). These group actions will lead to examples for our retract theory and provide an important application of retracts. 2.3 Variable Lie algebras and groups We will prove our main theorem by induction; in our proofs, new parameters and new variations will appear. This may be handled most easily by the concept of variable Lie structures. Such structures were already considered in [5], [11], [10] and [9], among others. 3 Definition 2.3.1. 1. Let g be a real vector space of finite dimension n and B be an arbitrary nonempty set. We say that (B, g) is a variable (nilpotent) Lie algebra if (a) For every β ∈ B, there exists a Lie bracket [·, ·]β defined on g such that gβ := (g, [·, ·]β) is a nilpotent Lie algebra. (b) There exists a fixed basis Z = Z 0 = {Z1 = Z 0 1 , . . . , Zn = Z 0 n} of g such that the structure constants ak ij(β) defined by [Zi, Zj]β := nXk=1 ak ij(β)Zk satisfy the following property: For all β ∈ B and k ≤ max{i, j}, ak This means that {Z1, . . . , Zn} is a Jordan-Holder basis for gβ = (g, [·, ·]β). ij(β) = 0. 2. Assume that B is a smooth manifold. If the structure constants ak ij(β) vary smoothly on B, we say that (B, g) is a smooth variable (nilpotent) Lie algebra. We will denote (B, g) = (g, [·, ·]β)β∈B for the variable Lie algebra. For the rest of the paper we will assume that all variable Lie algebras are smooth. If B is reduced to a singleton, we have in fact no dependency on β in B but a fixed Lie algebra. To each variable Lie algebra, we associate a variable Lie group Gβ. The vari- able Lie group G := (Gβ)β may be identified with the collection of Lie algebras (g, [·, ·]β)β equipped with the corresponding Campbell-Baker-Hausdorff multiplications. If G = (Gβ)β is a (smooth) variable Lie group endowed with a fixed Jordan-Holder basis, then the cor- responding Vergne polarisations, induced representations and operator kernels all depend on β ∈ B and l ∈ g∗. 2.4 Ludwig-Zahir indices Let (B, g) be a smooth variable Lie algebra. We assume that g is equipped with a fixed basis Z = Z 0 = {Z1 = Z 0 n}, which is a Jordan-Holder basis for every (g, [·, ·]β). 1 , . . . , Zn = Z 0 Let (β, l) ∈ B × g∗. The Ludwig-Zahir indices I(β, l) defined in [11] can be obtained in the following way: Let gβ(l) := {U ∈ g; hl, [U, g]βi ≡ 0} be the stabiliser of l in gβ = (g, [·, ·]β) and let aβ(l) be the maximal ideal contained in gβ(l). If aβ(l) = gβ(l) = g, then χ(β,l)(x) := e−ihl,logβ xi is a character on Gβ and nothing has to be done. In this case, there are no Ludwig-Zahir indices, i.e. I(β, l) = ∅. Otherwise, let j1(β, l) = max{j ∈ {1, . . . , n}; Z 0 k1(β, l) = max{k ∈ {1, . . . , n}; hl, [Z 0 j 6∈ aβ(l)}, and j1(β,l), Z 0 k]βi 6= 0}. We let X1(β, l) Y1(β, l) Z1(β, l) c(β, l) : = Z 0 k1(β,l), : = Z 0 j1(β,l), : = [Z 0 k1(β,l), Z 0 : = hl, Z1(β, l)i. j1(β,l)]β, and 4 We then consider g1(β, l) := {U ∈ g; hl, [U, Y1(β, l)]βi = 0} (2.1) which is an ideal of co-dimension one in gβ. A Jordan-Holder basis of (g1(β, l), [·, ·]β) is given by Z 1(β, l) = {Z 1 i (β, l); i 6= k1(β, l)} with Z 1 i (β, l) := Z 0 i − hl, [Z 0 i , Y1(β, l)]βi c(β, l) X1(β, l), i 6= k1(β, l). (2.2) i (β, l) = Z 0 One sees that Z 1 i , if i > k1(β, l). As previously we may now compute the indices j2(β, l), k2(β, l) of l1 := lg1(β,l) with respect to this new basis and construct the corresponding subalgebra g2(β, l) with its associated basis {Z 2 i (β, l) ; i 6= k1(β, l), k2(β, l)}. This procedure stops after a finite number r of steps. Let IZ (β, l) = I(β, l) =(cid:0)(j1(β, l), k1(β, l)), . . . , (jr(β, l), kr(β, l))(cid:1), which is called the Ludwig-Zahir index of l in gβ with respect to the basis {Z1, . . . , Zn}. The construction in [11] shows that the final subalgebra gr(β, l) obtained by this construction coincides with the Vergne polarisation of l in gβ with respect to the basis Z 0 (see also [10], [9]). Note that the length I = 2r of the index set I = I(β, l) gives us the dimension of the co-adjoint orbit Ad∗(Gβ)ℓ. The vectors Y1(β, l), · · · , Yr(β, l) together with the stabiliser gβ(l) of l in gβ span the polarisation pβ(l) = gr(β, l) and g = ⊕r i=1 RXi(β, l) ⊕r i=1 RYi(β, l) ⊕ gβ(l). Let us introduce the following notations: For any index set I ∈ (N2)r ≡ N2r with r = 0, · · · , dim(g)/2, we let (B × g∗)I := {(β, l) ∈ B × g∗; I(β, l) = I} and (B × g∗)I ∩ ΣI := {(β, l) ∈ (B × g∗)I; l(Zji) = l(Zki) = 0 for 1 ≤ i ≤ r}. This last line corresponds to the Pukanszky section associated to the index I. In fact, in [9] it was proved that the indices js(β, l), ks(β, l) coincide with the Pukanszky indices of the given layer (if one does not make any distinction between the j's and the k's). For many I's, the subset (B × g∗)I is empty. Hence it is reasonable to define I :=nI ∈ dim(g)/2[j=0 (N2)j; (B × g∗)I 6= ∅o, and B × g∗ = [I∈I (B × g∗)I. This gives a partition of B × g∗ into the different layers (B × g∗)I. The set I may be r′)} ∈ I, ordered lexicographically: we say that I < I ′ if either 2r = I < I ′ = 2r′ or there exists a ∈ {1, . . . , r} such that if I = {(j1, k1), · · · , (jr, kr)}, I ′ = {(j′ 1), · · · , (j′ r′, k′ 1, k′ (js, ks) = (j′ a), s, k′ s) if s < a and (ja, ka) < (j′ a, k′ 5 which means that either ja < j′ a). a or (ja = j′ a and ka < k′ This allows us to define (B × g∗)≤I := {(β, l) ∈ (B × g∗)J ; J ≤ I} = [J≤I (B × g∗)J . By induction on the length of the index sets, it is easy to see that for every I ∈ I there exists a smooth function PI on B × g∗, which is polynomial in l for fixed β ∈ B such that (B × g∗)I = {(β, l); PI ′(β, l) = 0 for I ′ > I and PI(β, l) 6= 0}. (2.3) 2.5 Co-adjoint orbits For any index set I, we consider the subspace sI of g∗ which is given by For each β ∈ B, let sI = span{Z ∗ j ; j ∈ I}. Σβ,I := {(β, l) ∈ ({β} × g∗)I; l ∈ sI }. Then Σβ,I is locally closed in sI, since we have the smooth functions PI ′, I ′ ∈ I, defined on B × g∗ as in (2.3). Let d := I. For l ∈ g∗, let Ωβ,l = {Ad∗ β(g)l; g ∈ G} be the Gβ-orbit of l. Then dim(Ωβ,l) = d for l ∈ (B × g∗)I. There exist functions pj : (B × g∗)I × Rd → R, j = 1, · · · , n, which are rational in l ∈ g∗ and polynomial in z ∈ Rd for fixed β ∈ B such that for every (β, l) ∈ (B × g∗)I, Ωβ,l =n nXi=1 pi(β, l, z)Z ∗ i ; z ∈ Rdo. Furthermore if we write I = {i1 < · · · < id}, then pij (β, l, z) = zj for j = 1, · · · , d, and for i 6∈ I, we have pi(β, l, z) = hl, Zii + p′ i(β, l, z1, · · · , zj), ij < i < ij+1. Definition 2.5.1. A subset M of B × g∗ is called G-invariant if for every (β, l) ∈ M the element g · (β, l) := (β, Ad∗ β(g)l) is also contained in M. 6 2.6 Schwartz functions Let r ∈ N \ {0}, we define the space of (generalised) Schwartz functions S(Rr, B, G) ≡ S(Rr, B, g) ≡ S(Rr, B, Rn) to be the set of all functions f from Rr × B × G to C such that the function f defined by f (α, β, (x1, . . . , xn)) := f (α, β, expβ(x1Z1 + · · · + xnZn)) for α ∈ Rr, β ∈ B is smooth on Rr × B × Rn and that k fkK,T1,...,Ts,A1,A2,B1,B2 = β∈K;α∈Rr;x∈Rnh sup sup αr1xs1 ri≤Ai;sj≤Bj;i,j∈{1,2} T1T2 · · · Ts ∂r2 ∂αr2 ∂s2 ∂xs2 < +∞, f (α, β, (x1, . . . , xn))i for any compact subset K of B, any finite collection T1, . . . Ts of smooth vector fields defined on the manifold B, and any A1, A2, B1, B2 ∈ N. The function space S(Rr, B, G) is equipped with the topology defined by the collection of all these semi-norms. One may of course also use coordinates of the second kind to define the semi-norms on S(Rr, B, G). Note that the space S(Rr, B, G) does not depend on the choice of the Jordan-Holder basis. 2.7 Kernel functions Let S be a subset of B ×g∗ and L be a smooth manifold. We say that a mapping F : S → L is smooth, if the restriction of F to any smooth manifold N contained in S is smooth. Let B × g∗ be a smooth variable nilpotent Lie group with Jordan-Holder basis Z. For any (β, l) ∈ B × g∗ denote the Vergne polarisation at (β, l) associated to Z. We put π(β, l) := indG P (β,l)χl, with P (β, l) := expβp(β, l), for the corresponding family of induced unitary representations. Then the mapping (β, l) 7→ p(β, l) is smooth on each subset (B × g∗)I. For each index set I with length dI and (β, l) ∈ B × g∗, choose a Malvev basis R(β, l) = {R1(β, l), · · · , RdI (β, l)} of g relative to p(β, l), such that the mappings (β, l) 7→ R(β, l) are smooth on the different layers (B × g∗)I. Definition 2.7.1. Let M be any smooth G-invariant manifold of B × g∗ and let r ∈ N. We denote by Dc M,r the space of all functions F : Rr × M × G × G → C satisfying the following conditions. 1. F satisfies the covariance condition for every (β, l) ∈ M with respect to p(β, l), i.e. F (α, (β, l), x ·β p, y ·β q) = χl(p)χl(q)F (α, (β, l), x, y)) for all α ∈ Rr, p, q ∈ P (β, l) and x, y ∈ G. 7 2. The function F satisfies the following compatibility condition F (α, (β, Ad∗ β(g)l), x, y) = F (α, (β, l), x ·β g, y ·β g), for α ∈ Rr, (β, l) ∈ M and x, y, g ∈ G. This compatibility condition reflects the unitary equivalence of the representations π(β,l) and π(β,Ad∗ β (g)l). 3. The support of F in (β, l) is compact modulo G, i.e. there exists a compact subset C of M such that F (·, (β, l), ·, ·) is 0 outside the subset of G · C. 4. The function F has the Schwartz space property, i.e. for any I ∈ I the function FRr×M ∩(B×sI )×G×G is smooth and that kF kD,A1,A2,B1,B2,C1,C2 := (β,l)∈M,α∈Rr,x,x′∈Rrh sup D(β,l) ∂r2 ∂αr2 ∂s2 ∂xs2 ∂t2 ∂(x′)t2 where sup αr1xs1(x′)t1 ri≤Ai,sj≤Bj,tk≤Ck;i,j,k∈{1,2} F (α, (β, l), x, x′)i < ∞, (2.4) F (α, (β, l), x, x′) := F (α, (β, l), expβ(x1R1) · · · expβ(xrRr), expβ(x′ 1R1) · · · expβ(x′ rRr)), for any smooth differential operator D = D(β,l) on the manifold M, and any A1, A2, B1, B2, C1, C2 ∈ N. The space Dc M,r will be equipped with the topology defined by the collection of all these semi-norms. This does of course not depend on the choice of the smooth Malcev basis of g with respect to the smooth family of Vergne polarisations. Definition 2.7.2. Let M ⊂ B × g∗. A field F = (F (β, l))(β,l)∈M of kernel functions is called adapted if it satisfies the conditions in Definition 2.7.1. For an adapted field of kernel functions F on M, denote by opF the field of smooth operators defined through their kernel functions. For (β, l) ∈ M, the operator opF (β,l) acts on the space L2(G/P (β, l), χ(β,l)) in the following way: opF (β,l)ξ(g) =ZG/P (β,l) F (β, l)(g, x)ξ(x)d x. Remarks 2.7.3. a) If we impose the condition that the support of (β, l) be contained in the set G · C0 for a fixed subset C0 of M, we will denote the space of kernel functions by DC0 M . b) One has a similar definition of the kernel functions if one takes another smooth family of polarisations together with a smooth family of Malcev bases. 8 3 The retract theorem In this section, we state the main theorem of the paper which will be proved in the next section. Theorem 3.1. Let B × G be a smooth variable nilpotent Lie group, I = {(j1, k1) < · · · < (jr, kr)} be an index set and let M be a smooth G-invariant sub-manifold of B×g∗ contained in (B × g∗)≤I such that MI := M ∩ (B × g∗)I 6= ∅. Let π(β, l) be defined as previously from the smooth family of Vergne polarisations for (β, l) ∈ M. Then there exists an open nonempty relatively compact subset M ⊂ MI with closure M contained in MI such that the following holds: For any adapted kernel function F ∈ DM M , there is a function f in the Schwartz space S(Rr, B, G) such that π(β,l)(f (α, β, ·)) has F (α, (β, l), ·, ·) as an operator kernel for all (α, (β, l)) ∈ Rr ×M. Moreover the mapping F 7→ f is continuous with respect to the corresponding function space topologies. If the variation is trivial, we get the following theorem. Theorem 3.2. Let g be a nilpotent Lie algebra with Jordan-Holder basis Z. Let M be a smooth G-invariant sub-manifold of g∗. Let I := max{J ∈ IZ : M ∩ g∗ J 6= ∅}. Let πl = π(l) be defined as previously from the smooth family of Vergne polarisations for l ∈ M. Then there exists an open, relatively compact nonempty subset M ⊂ g∗ I of M such that M ⊂ M ⊂ MI , M is compact and that the following holds: For any kernel function F ∈ DM M , there is a function f in the Schwartz space S(G) such that πl(f ) has F (l, ·, ·) as an operator kernel for all l ∈ M. Moreover, the Schwartz function f may be constructed such that the mapping F 7→ f is continuous with respect to the corresponding function space topologies. Remark 3.2.1. If M is contained in g∗ we have the following (well known) result. Imax, where Imax is the maximal index set in I, then Theorem 3.3. Let B × G be a simply connected, connected smooth variable nilpotent Lie group and M = (B × g∗)gen := (B × g∗)Imax be the space of generic co-adjoint orbits. Let M be an open relatively compact subset of M such that M ⊂ M. For every adapted field of kernel functions F ∈ DM M , there exists a unique Schwartz function f = R(F ) : G → C such that π(β,l)(f ) = opF (β,l) for any (β, l) ∈ B × g∗, and the mapping F 7→ R(F ) is continuous. Proof. It suffices to apply the Fourier inversion formula. For each F ∈ DM M , let f (β, g) = R(F )(β, g) :=ZΣβ,Imax tr(π(β,l)(g) ◦ opF (β,l))Pa(β, l)dl, g ∈ G, where Pa(β, l) is the Pfaffian of the polynomial Q(l) = det (hl, [Zi, Zj]βii,j∈Imax). It follows from [11] that the function f is Schwartz and the Fourier inversion theorem tells us that π(β,l)(f ) = opF (β,l) for any (β, l) ∈ B × g∗. 9 4 Proof of the retract theorem The proof of Theorem 3.1 proceeds by induction on the length I of the largest index set I for which (B × g∗)I ∩ M 6= ∅ and it will be done in several steps. 4.0.1 The case I = ∅ Suppose that all the elements (β, l) ∈ M are characters of gβ, which means that their index sets are empty. Let us replace the variable group (B, G) by the group (C, G), where C = B as a manifold, and the multiplications coming from C are abelian, i.e. [U, V ]γ = 0 for every U, V ∈ g, γ ∈ C. We identify now the group G with its Lie algebra and then U ·γ V = U + V for every U, V ∈ g and γ ∈ C. This also means that χl is a character on Gγ = expγg, for all (γ, l) ∈ C × g∗. Now take M = M. Let F ∈ S(Rr × M) be a kernel function with compact support in the variables (γ, l). As Rr × M is a sub-manifold of Rr × C × g∗, the function F may be extended to a Schwartz function eF (in the sense of Section 2.6 and 2.7) on 3 eF , where F −1 Rr × C × g∗ with compact support in the variables (γ, l). Let f := (2π)nF −1 denotes the partial inverse Fourier transform in the variable l which is the third variable in Rr × C × g∗. Then f ∈ S(Rr × C × g∗). For all (α, (γ, l)) ∈ Rr × M, we have 3 π(γ,l)(cid:0)f (α, γ, ·)(cid:1) = bf 3(α, γ, l) = (2π)nF3F −1 = F (α, (γ, l)). 3 F (α, (γ, l)) In particular, π(γ,l)(f (α, γ, ·)) = 0 if (α, (γ, l)) ∈ Rr × (M \ C). The continuity of the map F 7→ f is obvious. This proves the first step in the induction procedure. 4.0.2 Reducing B There are two cases where we can reduce the manifold B. 1. Suppose that there exists a smooth function ϕ : B → R+ which is not constant on the subset BM := pB(M), where pB : B × g∗ → B is the canonical projection. Let β0 ∈ B such that ϕ(β0) ∈]a, b[ for some b > a > 0 and let B0 := {β ∈ B; a 2 < ϕ(β) < 2b} and M0 := {(β, l) ∈ M; β ∈ B0}. Suppose that the theorem holds for the pair (B0, M0). Let us show that the result remains true for the pair (B, M). Let M0 be an open relatively compact subset as in the theorem for (B0, M0). We let M := {(β, l) ∈ M; a < ϕ(β) < b} ∩ M0. We will show that M works for (B, M). Note that since M0 is open in M0, we have that M is open in M. Let F be a kernel function defined on Rr × M × G × G such that its support in (β, l) is contained in G · M ⊂ M0. By assumption, there exists f ∈ S(Rr × B0 × G) such that π(β,l)(f (·, β, ·)) admits F (·, (β, l), ·, ·) as an operator kernel if (β, l) ∈ M0. In 10 c (R) with 0 ≤ ϑ ≤ 1, ϑ ≡ 1 on [a, b] and ϑ ≡ 0 on [0, a particular, π(β,l)(f (·, β, ·)) = 0 if (β, l) ∈ M0 \ G · M0. As B0 is a sub-manifold of B, we may extend f to a function in S(Rr × B × G) which we denote also by f . Choose ϑ ∈ C∞ 2 ] ∪ [2b, +∞[. We define φ ∈ C∞(M) by φ(β, l) := ϑ(ϕ(β)). Then φ ≡ 1 on G · M0 and φ ≡ 0 on M \ G · M0. By taking g := φ · f , we have that π(β,l)(g(·, β, ·)) = φ(β, l) · π(β,l)(f (·, β, ·)). If (β, l) ∈ M ⊂ M0, then π(β,l)(g(·, β, ·)) = π(β,l)(f (·, β, ·)) and it admits F (·, (β, l), ·, ·) If (β, l) ∈ M0 \ G · M0, then π(β,l)(f (·, β, ·)) = 0 and as an operator kernel. π(β,l)(g(·, β, ·)) = 0. If (β, l) ∈ M \M0, then ϕ(β) ∈ [0, a 2 ]∪[2b, +∞[, hence φ(β, l) = 0 and so π(β,l)(g(·, β, ·)) = 0. Hence the result is true for the function g. 2. If there exists a smooth sub-manifold B0 of B such that pB(M) ⊂ B0, then we can apply our theorem to the pair (B0, M). Since every smooth function f0 on B0 × G can be extended to a smooth function f on B × G, the retract theorem also holds for (B, M). Remark 4.0.3. Let B and M be given as in the statement of the theorem. Let pB : M → B; pB(β, l) = β, If we denote by M max the subset of M consisting of all be the canonical projection. (β, l) ∈ M for which the rank of dpB(β, l) is maximal, then M max is open in M and the subset pB(M max) of B is a smooth sub-manifold of B. If pB(M max) contains at least 2 elements, by the reasoning in Subsection 4.0.2, using a non-constant smooth function ϕ0 on pB(M max), which can be extended to a smooth function ϕ of B, we reduce B to Bmax and we can always assume in this way that pB(M) is a smooth sub-manifold of B. If pB(M max) is a singleton {β0}, then M = M max and pB(M) is obviously a smooth sub-manifold of B. 4.1 Reducing to smoothly varying subspaces depending on B Let M ⊂ B × g∗ be a smooth G-invariant sub-manifold of B × g∗. Let us fix the largest index IM = I =(cid:0)(j1, k1), · · · , (jr, kr)(cid:1) = (j1, k1) × I1, where I1 = ((j2, k2), · · · , (jr, kr)) is the index set of (β, lg1(β,ℓ)), such that the open subset MI := (B × g∗)I ∩ M of M is nonempty. Let pB : M → B; (β, l) 7→ β, be the projection onto the first variable and set BM := pB(M), which is a smooth sub-manifold of B by Remark 4.0.3. Let c1 := gj1+1 = span{Zj1+1, . . . , Zn} ⊂ g and let n1 β := [g, c1]β + [Zj1, gk1+1]β ⊂ c1 for β ∈ B. (4.5) 11 Then, by the definition of the indices (j1, k1), we have n1 β ⊂ ker(l) ∩ c1 ⊂ aβ(l) if (β, l) ∈ (B × g∗)≤I. (4.6) It is easy to see that n1 β is an ideal in g. Let Zβ := [Zk1, Zj1]β for β ∈ B. We fix a scalar product h·, ·i on g such that {Z1, . . . , Zn} is an orthonormal basis and r=j1+1 arZr of 1) with respect to the given we identify c∗ c1. Denote by k · k2 the Euclidean norm on c1 (and hence on c∗ scalar product. We also identify 1 with c1 by identifying Pn 1 with the element Pn r=j1+1 arZ ∗ r ∈ c∗ (n1 β)⊥ := {q ∈ c∗ 1; hq, n1 βi = {0}} with a subspace of c1. For all β ∈ B, we write c1 = n1 orthogonal projection of c1 onto (n1 by β)⊥. For each β ∈ B, a generating subset of n1 β ⊕ (n1 β)⊥ and define pβ to be the β is given V (β) = {v1(β), · · · , vs(β)} := {[Za, Za′]β; a = 1, · · · , n, a′ = j1 + 1, · · · , n} ∪ {[Zb, Zj1]β; b = k1 + 1, · · · , n}. Let aj,j′(β) := hvj(β), vj′(β)i for 1 ≤ j, j′ ≤ s. Fix 0 ≤ k ≤ s, let Ik = {J ⊂ {1, · · · , s}; J = k} and for β ∈ B, let hk(β) := XJ∈Ik det(cid:0)(cid:0)aj,j′(β)(cid:1)j,j′∈J(cid:1)2 . It is easy to check that hk(β) 6= 0 ⇔ v1(β), . . . , vs(β) have at least rank k, hk(β) = 0 ⇔ v1(β), . . . , vs(β) have rank r < k. Let n1 ∈ N and put f0 := hn1+1 and f1 := hn1. Let B≤n1 = {β ∈ B; f0(β) = 0}, B≥n1 = {β ∈ B; f1(β) 6= 0}, Bn1 = {β ∈ B; f0(β) = 0 and f1(β) 6= 0}. One sees that B≥n1 is open in B, and hence is a sub-manifold of B. Again, according to the reduction argument in Subsection 4.0.2 we can assume that f1(β) 6= 0 for all β ∈ B. On the other hand, let n1 := maxβ∈BM dim(n1 β), then we have B≤n1 Bn1 B≥n1 := {β ∈ B; dim(n1 := {β ∈ B; dim(n1 := {β ∈ B; dim(n1 β) ≤ n1}, β) = n1}, β) ≥ n1}. 12 Note that if we want n1 β to be of fixed dimension and to have n1 β)⊥ and pβ to vary smoothly with respect to β, we must restrict to Bn1. But in general Bn1 is not a sub-manifold of B. Therefore we must find a smooth sub-manifold inside Bn1 containing an open subset of the smooth manifold BM = pB(M). We have to distinguish the following two cases: β, (n1 Case 1: If the differential df0 is not identically zero on BM , we may define Bmax M := {β ∈ BM ; df0(β) 6= 0} and Bmax := {β ∈ B; df0(β) 6= 0}. By assumption, Bmax is a nonempty open subset of B. Case 2: Assume that df0 is identically zero on BM . If dim(BM ) < dim(B), we may build a function γ ∈ C∞(B) such that γ ≡ 0 on BM and dγ is not identically zero on BM . We put ef0 = f0 + γ. Then ef0 ≡ 0 on BM and def0 is not identically zero on BM . We then define Bmax M := {β ∈ BM ; def0(β) 6= 0}, Bmax := {β ∈ B; ef0(β) = 0, def0(β) 6= 0}. By the construction of ef0, we have again that Bmax is an open subset of B and Bmax M ⊂ Bmax. If dim(BM ) = dim(B), then BM is open in B and we take a smooth function f 6= 0 in B supported on BM . Let Bmax := {β ∈ B; ef (β) 6= 0} and Bmax M := {β ∈ BM ; ef (β) 6= 0}. β vary smoothly on the smooth sub-manifold Bmax of B, since In the two cases, the ideals n1 dim(n1 β) = n1 on Bmax M . The projection pβ also varies smoothly on Bmax. Remark 4.1.1. According to Remark 4.0.3, we can now assume that B = Bmax. 2 is now smooth on B, we can take β0 ∈ B Furthermore, since the function β 7→ kpβ(Zβ)k2 and 0 < δ < R < ∞ such that δ < kpβ0(Zβ0)k2 < R and by using the reduction argument, we can then assume that the number kpβ(Zβ)k2 is contained in the interval [δ, R] for any β ∈ B. 4.1.2 On the manifold M Let us focus on the manifold M again. Let (β0, l0) ∈ M be fixed, but arbitrary. There exist 0 < δ < R < ∞ such that 0 < δ < min{hl0, Zβ0i, kpβ0(Zβ0)k2} < max{hl0, Zβ0i, kpβ0(Zβ0)k2} < R. This is due to the fact that M ⊂ (B × g∗)I. According to Remark 4.1.1 we can now assume that 0 < δ < min{kpβ(Zβ)k2} < max{kpβ(Zβ)k2} < R 13 for all β ∈ B. We define M δ,R = M red := {(β, l) ∈ M; 0 < δ < min{hl, Zβi} < max{hl, Zβi} < R}. Obviously, M red is open in M and thus is a smooth sub-manifold of M. On the other hand, we define (B × g∗)≤I,R,δ = n(β, l) ∈ (B × g∗)≤I; 1 2 δ < min{hl, Zβi} < max{hl, Zβi} < Ro. 3 2 4.1.3 Reducing M Now we show that if the retract theorem holds for (B, M red), then it remains true for (B, M). Assume that the result is true for (B, M red). Let M be the open subset in M red given by the assumption. We will show that one may take the same manifold M for (B, M) such that the theorem remains true for (B, M). As M red is open in M, the set M also has a nonempty interior in M. Moreover, pB(M) ⊂ pB(M red) ⊂ Bn1 ⊂ B≥n1. Let ∅ 6= C ⊂ M be compact and let F be a kernel function defined on Rr × M × G × G whose support in (β, l) is contained in G · C. The restriction of F to Rr × M red × G × G is a kernel function for (B, M red). By assumption, there exists f ∈ S(Rr × B × G) such that π(β,l)(f (·, β, ·)) admits In particular, π(β,l)(f (·, β, ·)) = 0 F (·, (β, l), ·, ·) as an operator kernel if (β, l) ∈ M red. if (β, l) ∈ M red \ C. As ∅ 6= C ⊂ M is compact, there exist δ1, R1 ∈ R+ such that 0 < δ < δ1 ≤ min{hl, Zβi} ≤ max{hl, Zβi} ≤ R1 < R for all (β, l) ∈ C, as C ⊂ M ⊂ M ⊂ (B × g∗)I. Let u ∈ C∞ c (R) be odd such that u ≡ 1 on [δ1, R1] and u ≡ 0 on [0, δ] ∪ [R, +∞[. There exists χ ∈ S(R) such that For (β, l) ∈ (B × g∗)J with J ≤ I, we have Zβ ∈ aβ(l) ⊂ gβ(l) and π(β,l)RZβ (ψ(β, ·)) = bχ = u. Let us define a function ψ on Sβ{β} × exp(RZβ) by ψ(β, exp(sZβ)) := χ(s). bχ(hl, Zβi)IH(β,l) = u(hl, Zβi)IH(β,l). We define a function g on Rr × B × G by g(·, β, ·) := f (·, β, ·) ∗ ψ(β, ·). This implies that π(β,l)(g(·, β, ·)) = u(hl, Zβi)π(β,l)(f (·, β, ·)). If (β, l) ∈ C, then u(hl, Zβi) = 1 and π(β,l)(g(·, β, ·)) = π(β,l)(f (·, β, ·)) admits F (·, (β, l), ·, ·) as an operator kernel. If (β, l) ∈ M red \C, then π(β,l)(f (·, β, ·)) = 0, hence π(β,l)(g(·, β, ·)) = 14 0 and F (·, (β, l), ·, ·) = 0. If (β, l) ∈ M \ M red, then hl, Zβi /∈ [δ, R], i.e. u(hl, Zβi) = 0, which implies that π(β,l)(g(·, β, ·)) = 0. Hence, the mapping F 7→ g satisfies the property of the retract for (B, M). 4.2 Construction of a new variable group 4.2.1 The mapping α(β, l) β)⊥ and For (β, l) ∈ (B × g∗)≤I, we have seen in (4.6) that n1 hl, Zβi = hl, pβ(Zβ)i = hq, pβ(Zβ)i. For (β, l) ∈ (B×g∗)≤I,R,δ, we have that hl, [Zk1, Zj1]βi = hl, Zβi > δ 2 > 0 implies hl, [Zk1, Zj1]βi 6= 0 and j1(β, l) = j1, k1(β, l) = k1. β ⊂ ker(l). Let q := lc1 ∈ (n1 Take an odd function ϕ ∈ C ∞(R) with the properties that ϕ(s) = 0 for 0 ≤ s < δ/4 and s > 2R, 1 > ϕ(s) > 0 for s ∈]δ/4, δ/2[∪]3R/2, 2R[ and ϕ(s) = 1 for 3R/2 ≥ s ≥ δ/2. For every (β, q) ∈ B × c∗ 1, we construct the vector α(β, q) ∈ c1 ≃ (c1)∗ by α(β, q) := ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i)pβ(q) +(cid:0)1 − ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i)(cid:1)pβ(Zβ). (4.7) Then by the construction, α(β, q) ∈ (n1 other hand, for (β, l) ∈ (B × g∗)≤I,R,δ and q = lc1, we have that β)⊥ ⊂ c∗ 1 ≡ c1 for every (β, q) ∈ B × c∗ 1. On the α(β, q) = ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i)pβ(q) +(1 − ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i))pβ(Zβ) = ±pβ(q) + (1 − 1)pβ(Zβ) = ±q. (4.8) This is due to the fact that pβ(q) = q as n1 β ⊂ ker(q) for q = lc1, if (β, l) ∈ (B × g∗)≤I. We will show that hα(β, q), Zβi = ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i)hpβ(q), Zβi +(1 − ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i))kpβ(Zβ)k2 2 > 0 on B × c∗ and ϕ ≥ 0 on R+, we have 1. In fact, let us first notice that hpβ(q), Zβi = hq, pβ(Zβ)i. As ϕ is an odd function A := ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i)hpβ(q), Zβi ≥ 0. Since 0 ≤ ϕ ≤ 1 on R+, B :=(cid:0)1 − ϕ(kpβ(Zβ)k2)ϕ(hq, pβ(Zβ)i)(cid:1)kpβ(Zβ)k2 2 ≥ 0. If none of the ϕ(·)'s is equal to zero and if hpβ(q), Zβi 6= 0, then A > 0. If hpβ(q), Zβi = 0, then ϕ(hpβ(q), Zβi) = 0 and so B > 0, as now by Remark 4.1.1 kpβ(Zβ)k2 > 0. If one of the ϕ(·)'s is equal to zero, then again B > 0. 15 For (β, q) ∈ B × c∗ 1, let We have that and δ(β, q) := ad∗ β(Zj1)α(β, q) ∈ g∗. hδ(β, q), Zk1i = hα(β, q)), Zβi > 0 hδ(β, q), [g, g]βi = hα(β, q), [[g, g]β, Zj1]βi ⊂ hα(β, q), [g, [g, Zj1]β]βi βi = {0}, ⊂ hα(β, q), [g, c1]βi ⊂ hα(β, q), n1 by the definition of α(β, q) in (4.7). This means that δ(β, q) is an algebra homomorphism of gβ = (g, [·, ·]β) which does not vanish at the vector Zk1. Hence the subspace g1(β, q) := ker(δ(β, q)) is an ideal of gβ of co-dimension 1 and g = RZk1 ⊕ g1(β, q). (4.9) Furthermore g1(β, q) contains c1 for any (β, q) ∈ B × c∗ 1. In fact, hδ(β, q), c1i = hα(β, q), [c1, Zj1]βi = 0 as α(β, q) ∈ (n1 β)⊥ and [c1, Zj1]β ⊂ n1 β. 4.2.2 The new variable group (B1, G1) In order to construct a new variation in the induction procedure, we put For (β, y, q) ∈ B1, we define a Jordan-Holder basis B1 := B × R × c∗ 1. Z 1(β, y, q) = { Z 1 1 (β, q), · · · , Z 1 n−1(β, q)} of g1(β, y, q) = ker(δ(β, q)) by αβ,y,q k = αβ,q k := hα(β, q), [Zk, Zj1]βi hα(β, q), Zβi and eZ 1(β, y, q) = eZ 1(β, q) := {Z1 − αβ,q = { Z 1 1 (β, q), · · · , Z 1 n−1(β, q)}. 1 Zk1, , · · · , Zk1−1 − αβ,q k1−1Zk1, Zk1+1, · · · , Zn} 16 In particular, for (β, l) ∈ (B × g∗)≤I,R,δ we have by (2.1) that g1(β, y, lc1) = g1(β, l) = g1(β, lc1). In fact, in this case j1(β, l) = j1, k1(β, l) = k1 and g1(β, y, lc1) = {U ∈ g hδ(β, lc1), Ui = 0} = {U ∈ g hl, [U, Zj1]βi = 0} = g1(β, l), as α(β, lc1) = ε · lc1 with ε = ±1 if (β, l) ∈ (B × g∗)≤I,R,δ. For each k, we also have that αβ,q k = hl, [Zk, Zj1]βi hl, Zβi . (4.10) This new basis eZ 1(β, y, q) coincides then, up to normalisation, with the basis obtained in Section 2.4, both procedures and bases generate the same indices. Furthermore by (4.10), for (β, l) ∈ (B × g∗)≤I,R,δ, we have Z 1(β, y, l1) = Z 1(β, l), (4.11) where Z 1(β, l) is defined in Section 2.4 and l1 = lc1. For any (β, y, q) ∈ B1, let us write [ Z 1 u(β, q), Z 1 v (β, q)] = n−1Xw=1 γ(β, q)u,v w Z 1 w(β, q) for u < v in {1, · · · , n − 1}. We obtain in this way a new variable Lie algebra (B1, g1), where g1 = Rn−1, B1 = B × R × c∗ 1 and the canonical basis Z 1 = {Z 1 1 , · · · , Z 1 n−1} of g1 satisfies, by definition, [Z 1 u, Z 1 v ](β,q) = n−1Xw=1 γ(β, q)u,v w Z 1 w, for u < v in {1, · · · , n − 1}. This means that the new variable Lie algebra (B1, g1) with g1 ≡ g1(β, q) is defined such that (g1, [·, ·](β,q)) ≡ (g1(β, q), [·, ·](β,q)). Given (β, l) ∈ B × g∗, let us define l1 ∈ g∗ i ) := l( Z 1 i ) = l(Zi+1) if i ≥ k1. We also define a map 1 by l1(Z 1 1}. One has l1(Z 1 i (β, q)) for all i ∈ {1, . . . , n − ι1 : B × g∗ → B1 × g∗ 1 (β, l) 7→ ((β, hl, Zk1i, α(β, lc1)), l1), (4.12) where l1 ≡ lg1(β,lc1 ). We see that ι1 is obviously smooth, injective and even a diffeomor- phism onto its image. 17 Using (4.9) we can identify every l ∈ g∗ with the pair (v, l1) where v := hℓ, Zk1i and l1 := lg1 ≡ lg1(β,lc1 ). We can then transfer the natural action of G on B × g∗ to B1 × g∗ using the mapping ι1. This gives us 1 g · ((β, v, q), l1) = ((β, v + hAd∗ β(g)l1). β(g)l1, Zk1i, Ad∗ β(g)q), Ad∗ Then we have automatically the relation ι1(g · (β, l)) = g · (ι1(β, l)) for any g ∈ G and (β, l) ∈ B × g∗. Consider now the smooth manifold (B × g∗)0 ≤I,R,δ := {(β, l) ∈ (B × g∗)≤I,R,δ; hl, Zk1i = hl, Zj1i = 0}. Then obviously the smooth manifold (B × g∗)≤I,R,δ is diffeomorphic with the manifold R2 × (B × g∗)0 ≤I,R,δ. The mapping Φ : R2 × (B × g∗)0 ≤I,R,δ → (B × g∗)≤I,R,δ given by Φ(s, t, (β, l)) :=(cid:16)β, Ad∗(exp( s hl, Zβi Zj1)exp( t hl, Zβi Zk1))l(cid:17) is such a diffeomorphism. Hence every smooth G-invariant sub-manifold M of (B×g∗)≤I,R,δ can be decomposed into a direct product of R2 with the smooth manifold M 0, where M 0 := {(β, l) ∈ M; hl, Zk1i = hl, Zj1i = 0}. For (β, l) ∈ B × g∗, one has l1(Z 1 i ) = l(Zi+1) if i ≥ k1. We remark that for (β, l) and (β, l′) in M with ι1(β, l) = ι1(β, l′) we have that l and l′ have the same restriction to g1(β, l) = g1(β, l′), so they are on the same co-adjoint orbit and l′ = Ad∗(y)l for some y ∈ P (β, l) and hence i ) = l(Zi) if i < k1 and l1(Z 1 F (β, l) = F (β, l′) by the conditions on the operator fields defined over M given in Definition 2.7.1. We denote the new variable Lie group by G1 = (B1, G1) where G1 = (expβ1g1)β1∈B1 and expβ1g1 is the connected, simply connected, nilpotent Lie group associated to the Lie algebra (g1, [·, ·]β1). 4.3 Induction step To simplify notations, from now on we will omit the subscript β in the notations of the mul- tiplication and the exponential map, unless the subscript is crucial for the understanding. There are two preliminary steps to check. 18 4.3.1 Induction hypothesis In this subsection, we will prove the result for (B, M) using induction. Let M1 = ι1(M) be as constructed in (4.12). Let us recall that l1 = l1(β, lc1) ≡ lg1(β,lc1 ) for (β, l) ∈ (B × g∗)≤I. The Vergne polarisation p(β, l) for l in (g, [·, ·]β), obtained by the procedure of Ludwig- Zahir (see [11], [9]), is also the Vergne polarisation for l1 in (g1, [·, ·]β). Let us denote by P (β, l) = expβp(β, l)) the corresponding subgroup. The associated induced representations ∼= will be denoted by π(β,l) := indG indG G1 π((β,lc1 ),l1), as usual. Since M is G-invariant, the manifold M1 = ι1(M) is also G-invariant in B1 × g∗ P (β,l)χl, respectively, π((β,lc1 ),l1) := indG1 P (β,l)χl1. Then π(β,l) 1. Hence we can write M1 as a direct product manifold R2 × M 0 1 , where M 0 1 := {((β, 0, q), l1); hl1, Zj1i = 0} is G1 invariant. Note that M 0 we have that ∞ > R > hl1, Zβi > δ > 0. The induction hypothesis in B1 × g∗ M 0 1 of M 0 the required properties of the theorem. 1)I1 and for every ((β, v, lc1), l1) ∈ M1 1 and 1 with 1)I1 gives us an open relatively compact non-empty subset M0 1 is contained in (B1 × g∗ 1 ⊂ (B1 × g∗ We choose now a relatively compact open subset M1 of M1 such that M1 ⊂ M1 and M1 is contained in G · M0 1. Let M := ι−1 1 (M1) and M0 := ι−1 1 (M0 1). Then M is non-empty open with its closure M contained in M and M is contained in G · M0. We take a kernel function F ∈ Dc M such that its support is contained in Rr × G · M × G × G. Given the kernel function F , we will now define a kernel function for the variable group (B1, G1). For simplicity, we will omit the subscripts β or (β, v, lc1) in the nota- tions of the multiplication and the exponential map, and we will identify g1, g′ 1 ∈ G1 = ≡ g1 with the corresponding elements in G1. In the following computa- tions, the parameters β and (β, v, lc1) will indicate how to multiply group elements or how to decompose smoothly the group elements, even if it is not marked explicitly. For ι1(β, l) = ((β, hl, Zk1i, lc1), l1) ∈ M1, we put (cid:0)(G1)(β1)(cid:1)β1∈B1 F1(α, u, t, ((β, hl, Zk1i, lc1), l1), g1, g′ 1) := (2π)n−j1+1 · c(β, l) · F (α, (β, l), exp((u + t)X) · g1, exp(tX) · g′ 1), for α ∈ Rr, u, t ∈ R and g1, g′ function F1 has its support S1 := ι1(S) contained in G · M1, and belongs to Dc operator field F1 is smooth on M1, since the mappings F and c are both smooth. 1 ∈ G1, where c(β, l) := hl, [Zk1, Zj1]βi 6= 0 and X = Zk1. This M1. The By the induction hypothesis, there exists h ∈ S(Rr+2, B1, G1) such that π((β1,lc1 ),l1)(cid:0)h(α, u, t, β1, ·)(cid:1) admits F1(α, u, t, (β1, l1), ·, ·) as an operator kernel for all (β1, l1) ∈ M 0 of the retract function f will now be done in several steps. 1 . The construction 19 4.3.2 Definition of the retract function on the original group For (β, v, q) ∈ B1, let us first define eh by eh(α, u, t, (β, v, q), g1) := ZRZc1 h(α, u, t, (β, v, q), g1 · exp(yY ) · exp(Z))e−iq(Z)dZdy, where Y = Zj1 and Z = Zβ = [X, Y ]β with X = Zk1. The integral converges, as h is Schwartz in g1 (for fixed β1), and it is of rapidly decreased in q ∈ (c1)∗, because it is a Fourier transform in Z. For all (β, v, q) ∈ B × R × c∗ 1, we then define ef (α, (β, v, q), exp(uX) · g1 · exp(yY ) · exp(Z)) = eiq(Z)ZReh(α, u, t, (β, v, q), g−t 1 )e−ityq([X,Y ]β)dt with g = exp(uX) · g1 and g−t 1 Rr × (B × c∗ := exp(−tX) · g1 · exp(tX). The function ef is smooth on 1 by construction, we may define f by 1) × G. As ef is of rapid decrease in q ∈ c∗ f (α, β, g) :=Z(c1)∗ ef (α, (β, 0, q), g)dq, α ∈ Rr, β ∈ B, g ∈ G. One can see that f ∈ S(Rr, B, G) (in the sense of Section 2.6). 4.3.3 The retract property Let us now compute π(β,l)(cid:0)f (α, β, ·)(cid:1) for (β, l) ∈ M. Since the manifold M is contained in (B × g∗)≤I,R,δ we have that c1 ⊂ aβ(l) ⊂ gβ(l). If we identify exp(c1) and c1, as well as exp(Z) and Z, for any function ξ(β) ∈ H(β,l) (the representation space of π(β,l)) and any g ∈ G, we have that (cid:16)π(β,l)(cid:0)f (α, β, ·)(cid:1)ξ(β)(cid:17)(g) = ZGβ /c1Zc1 = ZGβ /c1Zc1 = ZGβ /c1Zc1Z(c1)∗ = ZGβ /c1Zc1Z(c1)∗ 2π(cid:17)n−j1ZGβ/c1 = (cid:16) 1 f (α, β, g · Z)(cid:16)π(β,l)(g)π(β,l)(Z)ξ(β)(cid:17)(g)dZd g f (α, β, g · Z)e−il(Z)(cid:16)π(β,l)(g)ξ(β)(cid:17)(g)dZd g f (α, (β, 0, q), g · Z)e−il(Z)(cid:16)π(β,l)(g)ξ(β)(cid:17)(g)dqdZd g f (α, (β, 0, q), g)eiq(Z)e−il(Z)(cid:16)π(β,l)(g)ξ(β)(cid:17)(g)dqdZd g f (α, (β, 0, lc1), g)(cid:16)π(β,l)(g)ξ(β)(cid:17)(g)d g. We use the following smooth decomposition: X = Zk1, g1 = g1(β, l) which gives us g = exp(sX) · g1 with s = s(g, β, lc1), g1 = g1(g, β, lc1). 20 We then obtain (using the fact that c1 ⊂ (aβ(l)) for all our (β, l)'s) that: · exp((s − u)X) · g1)d g1du · g1)d g1dr f (α, (β, 0, lc1), exp(uX) · g1) · with gr 1 = exp(rX) · g1 · exp(−rX) f (α, (β, 0, lc1), exp((s − r)X) · g1) · f (α, (β, 0, lc1), exp((s − r)X) · gr 1) · ξ(β, 0, lc1)(r)(g−1 1 f (α, (β, 0, lc1), exp(uX) · g1) · ξ(β)(g−1 1 with ξ(β, 0, lc1)(v)(g1) := ξ(β)(exp(vX) · g1) f (α, (β, 0, lc1), g)(cid:16)π(β,l)(g)ξ(β)(cid:17)(exp(sX) · g1)d g (2π)n−j1 ·(cid:16)π(β,l)(cid:0)f (α, β, ·)(cid:1)ξ(β)(cid:17)(exp(sX) · g1) = ZGβ /c1 = ZRZG1/c1 = ZRZG1/c1 ξ(β, 0, lz)(s − u)(cid:16)(cid:0)exp(−(s − u)X) · g1 · exp((s − u)X)(cid:1)−1 · g1(cid:17)d g1du = ZRZG1/c1 ξ(β, 0, lc1)(r)(cid:16)(cid:0)exp(−rX) · g1 · exp(rX)(cid:1)−1 · g1(cid:17)d g1dr with s − u = r = ZRZG1/c1 = ZRZG1/c1 = ZRZG1/c1ZR = ZRZG1/exp(RY )·c1ZRZR (cid:16)π((β,0,lc1 ),l1)(w1) ξ(β, 0, lc1)(r)(cid:17)(g1)dtdyd w1dr (as l(Y ) = 0, for l1 ≡ lg1(β,0,lc1 )) = ZRZG1/exp(RY )·c1ZRZR (cid:16)π((β,0,lc1 ),l1)(w1) ξ(β, 0, lc1)(r)(cid:17)(g1)dtdyd w1dr (as (cid:0)exp(yY )(cid:1)r−t c(β, l)ZRZG1/exp(RY )·c1ZRZR (cid:16)π((β,0,lc1 ),l1)(w1) ξ(β, 0, lc1)(r)(cid:17)(g1)dtdyd w1dr = exp(yY + y(r − t)[Zk1, Zj1]β), by covariance of h, with 1)(cid:16)π((β,0,lc1 ),l1)(g1) ξ(β, 0, lc1)(r)(cid:17)(g1)d g1dr )(cid:16)π((β,0,lc1 ),l1)(g1) ξ(β, 0, lc1)(r)(cid:17)(g1)dtd g1dr h(α, s − r, t, (β, 0, lc1), wr−t · (exp(yY )r−t) h(α, s − r, t, (β, 0, lc1), wr−t 1 · exp(yY )) · ei(r−t)c(β,l)y f (α, (β, 0, lc1), exp((s − r)X) · gr c(β, l) = hl, [Zk1, Zj1]βi 6= 0 as before) 1 = (with l1 ≡ lg1) h(α, s − r, t, (β, 0, lc1), gr−t 1 h(α, s − r, t, (β, 0, lc1), wr−t 1 · exp(c(β, l)−1yY ))eir ye−ity (with y = c(β, l)y) 1 21 = = = = 1 h(α, s − r, t, (β, 0, lc1), wr−t ·exp(c(β, l)−1 yY ) · exp(y′Y ) · Z) · e−il(Z)eir ye−ity c(β, l)ZRZG1/exp(RY )·c1ZRZRZRZc1 (cid:16)π((β,0,lc1 ),l1)(w1) ξ(β, 0, lc1)(r)(cid:17)(g1)dZdy′dtdyd w1dr c(β, l)ZRZG1/exp(RY )·c1ZRZRZRZc1 ·e−il(Z)eir ye−ity(cid:16)π(β1,l1)(w1) ξ(β, 0, lc1)(r)(cid:17)(g1)dZdy′′dtdyd w1dr (for y′′ = y′ + c(β, l)−1y) h(α, s − r, t, (β, 0, lc1), wr−t · exp(y′′Y ) · Z) 1 1 1 h(α, s − t, t, (β, 0, lc1), w1 · exp(y′′Y ) · Z) 1 2π · 1 c(β, l)ZRZG1/exp(RY )·c1ZRZc1 ·e−il(Z)(cid:16)π((β,0,lc1 ),l1)(w1) ξ(β, 0, lc1)(t)(cid:17)(g1)dZdy′′dtd w1 1 c(β, l)ZR 1 2π · π((β,0,lc1 ),l1)(cid:0)h(α, s − t, t, (β, 0, lc1), ·) ξ(β, 0, lc1)(t)(cid:1)(g1)dt as e−il(Z)π((β,0,lc1 ),l1)(w1) = π((β,0,lc1 ),l1)(w1 · exp(y′′Y ) · Z). Let us finish the computation for (β, l) ∈ M. It suffices to take (β, l) ∈ M 0. Then ((β, 0, lc1), l1) ∈ M 0 1 and by the induction hypothesis, · 1 (cid:16)π(β,l)(cid:0)f (α, β, ·)(cid:1)ξ(β)(cid:17)(g) = (cid:16)π(β,l)(cid:0)f (α, β, ·)(cid:1)ξ(β)(cid:17)(exp(sX) · g1) c(β, l)ZRZG1/P (β,l) = (cid:16) 1 2π(cid:17)n−j1+1 = ZRZG1/P (β,l) = ZG/P (β,l) ξ(β, 0, lc1)(t)(g1)d g1dt F (α, (β, l), g, g)ξ(β)(g)dg. F1(α, s − t, t, ((β, 0, lc1), l1), g1, g1) F (α, (β, l), exp(sX) · g1, exp(tX) · g1)ξ(β)(exp(tX) · g1)dg1dt Hence for every (β, l) ∈ M, we have the required result. The algorithm used to build the retract function f respects the semi-norms defining the topology of our function spaces. So the retract map F 7→ f is continuous. 5 G-prime ideals in L1(G) In this section, we will study the structure of the A-prime ideals in L1(G) by using the retract theorem. 22 5.1 A retract defined on closed orbits Let G be a Lie group of automorphisms of a connected, simply connected, nilpotent Lie group G = exp(g) containing the inner automorphisms of G. For instance take any simply connected Lie group G and let G be the nilradical of G. Let l0 ∈ g∗ be fixed, we consider the orbit Ω = Ωl0 := G · l0 in g∗, let O = Ol0 be the G-orbit of l0. We assume that Ω is locally closed in g∗. In particular we can write Ω = Ω ∩ U, where Ω denotes the closure of Ω in g∗ and U is G-invariant open subset of g∗. It is then a G-invariant smooth sub-manifold of g∗ diffeomorphic to the manifold G/Gl0, where Gl0 denotes the stabiliser Gl0 := {α ∈ G; α · l0 = l0}. The G-orbit G · (G · l0) in the orbit space g∗/G is then locally closed and homeomorphic to the quotient G/GO, where GO is the stabiliser of the set O in G. In fact, we have that GO = G · Gl0. For a Jordan-Holder basis Z = {Z1, · · · , Zn} of g and g ∈ G, let g · Z := {Ad(g)Z1, · · · , Ad(g)Zn}, which is again a Jordan-Holder basis of g. For every index set I, we have the following relation (see [9]): Ad∗(g)g∗ I,g·Z = g∗ I,Z, g ∈ G. (5.13) For an index set I and a Jordan-Holder basis Z of g, recall that sI :=Xi∈I RZ ∗ i , ΣI,Z := sI ∩ g∗ I,Z, and the mapping EI : Rd × ΣI,Z → g∗ I,Z is given by EI(s1, t1, · · · , sr, tr; l) := Ad∗(exp(s1Zj1)exp(t1Zk1) · · · exp(srZjr)exp(trZkr ))l. We have that EI is a bijection and EI(Rd × {l}) is the G-orbit of l. Let Υ : g∗ I,Z → ΣI,Z; Υ(l) := Ad∗(G)l ∩ ΣI,Z = pΣI,Z (E−1 I (l)), where pΣI,Z is the projection of Rd × ΣI,Z onto ΣI,Z. For the orbit Ω, we need to construct a finite partition of unity (ψi)i∈Γ consisting of smooth G-invariant functions ψi : Ω → R+ such that for every i ∈ Γ the support of each function ψi is contained in an open subset of g∗ In order to do that let ϕ : R → R+ be a smooth function with compact support which vanishes in a neighbourhood of 0. We define a function ψ : g∗ I,gi·Z for some gi ∈ G. ≤I → R+ by ψ(l) := ϕ(PI(Υ(l)) if l ∈ g∗ I and ψ(l) := 0 if l ∈ g∗ <I, 23 where PI is a smooth function on B × g∗ defined in Section 2.4. We see that ψ is smooth (since ϕ vanishes in a neighbourhood of 0) and is G-invariant by the construction. Let UI,Z := {l ∈ Ω; ψ(l) 6= 0}. Now assume that g∗ We have that Ω ∩ g∗ non-empty open subset of Ω contained in g∗ I = g∗ I,g·Z 6= ∅ and Ω ∩ g∗ I and I,Z be the maximal layer with respect to Z such that Ω ∩ g∗ I,Z 6= ∅. I ′,g·Z = ∅ for g ∈ G and I ′ > I. Moreover, UI,Z is a Ω ⊂ [g∈G Ad∗(g)UI,Z. Let C be a compact subset of g∗ contained in Ω, then there exists a finite subset Γ ⊂ G such that C ⊂ [g∈Γ Ad∗(g)UI,Z. Hence there is a finite partition of unity (ψi)i∈Γ consisting of smooth G-invariant functions ψi : Ω → R+ such that the support of each function ψi is contained in Ad∗(gi)UI,Z ⊂ g∗ for every gi ∈ Γ. I,gi·Z Suppose we have a smooth adapted operator field F on Ω supported on G · C, we can write F =Xi∈Γ ψiF. According to the retract theorem, for each i ∈ Γ there is a (retract) Schwartz function fi on G such that πl(fi) = opψiF (l) for every l ∈ Ω. For f :=Pi∈Γ fi, we have that πl(f ) = Xi∈Γ = Xi∈Γ πl(fi) ψiopF (l) = opF (l). Hence for every smooth adapted kernel function supported on G·C, we have build a retract function. 24 5.2 G-prime ideals Let us first recall the definition of G-prime ideals. Definition 5.2.1. A two-sided closed ideal I in L1(G) is called G-prime, if I is G-invariant and if, for all G-invariant two-sided ideals I1 and I2 of L1(G), the following implication holds I1 ∗ I2 ⊂ I ⇒ I1 ⊂ I or I2 ⊂ I. Denote by P rim∗(G) the collection of all the kernels of irreducible unitary representa- tions of L1(G). For a closed subset C of P rim∗(G), let ker(C) := \P ∈C P. For a subset I of L1(G), denote by h(I) the subset h(I) := {P ∈ P rim∗(G); I ⊂ P }. The set h(I) is then closed in P rim∗(G) with respect to the Fell topology. We have the following result for G-prime ideals of L1(G) which can be viewed as an application of the retract theorem. Theorem 5.3. Let G be a simply connected, connected nilpotent Lie group and let G be a Lie group of automorphisms of G containing the inner automorphisms, which acts smoothly on the group G, such that every G-orbit in g∗ is locally closed. If I is a proper G-prime ideal of L1(G), then there exists an G-orbit Ωl0 in g∗ such that Moreover, all the kernels of G-orbits are G-prime ideals. I = ker(Ωl0). Proof. For any G-orbit Ω in g∗, the retract theorem tells us that the Schwartz functions contained in ker(Ω) are dense in ker(Ω) (see [7, proof of Proposition 4.1] and [4]). From the proof of [8, Theorem 1.2.12], it follows that the hull of a prime ideal I is the closure of an G-orbit in P rim∗(G) ≃ bG. On the other hand, the density of S(G) ∩ ker(Ω) implies that ker(Ω)N is contained in the minimal ideal J(Ω) with hull Ω for some N ∈ N. This tells us that ker(Ω)N ⊂ J(Ω) ⊂ I, since the minimal ideal with hull Ω is contained in every ideal with hull Ω. Since I is G-prime, we have that I = ker(Ω). Obviously the ideal ker(Ω) is G-prime for any G-orbit Ω in g∗. To see this, let I1 and I2 be two G-invariant ideals of L1(G) such that I1 ∗ I2 ⊂ ker(Ω). This means that I1 ∗ I2 ⊂ ker(Ω) ⊂ ker(πl) for some l ∈ Ω. We have then either I1 or I2 is contained in ker(πl), since πl is irreducible. But if I1 is contained in ker(πl), it is also contained in ker(πk·l) since I1 is G-invariant. Hence I1 ⊂ ker(Ω) and the proof is complete. 25 References [1] L. Corwin and P. Greenleaf, Representations of nilpotent Lie groups and their applications. Part 1: Basic theory and examples, Cambridge Studies in Ad- vanced Mathematics, 18 (1990). [2] R. Howe, On a connection between nilpotent groups and oscillatory integrals associated to singularities, Pacific J. Math. 73 (1977), no. 2, 329-363. [3] A. A. Kirillov, Unitary representations of nilpotent Lie groups, Uspekhi Mat. Nauk. 17 (1962), 53-104. [4] R. Lahiani and C. Molitor-Braun, Compact actions, retract theory and prime ideals, Illinois J. Math. 55, no. 3 (2011), 1235-1266. [5] H. Leptin, J. Ludwig, Unitary representation theory of exponential Lie groups, De Gruyter Expositions in Mathematics 18, 1994. [6] J. Ludwig, On Primary Ideals in the Group Algebra of a Nilpotent Lie Group, Math. Ann. 262 (1983), 287-304. [7] J. Ludwig and C. Molitor-Braun, Exponential actions, Orbits and their Ker- nels, Bull. Austral. Math. Soc. 57 (1998), 497-513. [8] J. Ludwig, C. Molitor-Braun, Repr´esentations irr´eductibles born´ees des groupes de Lie exponentiels, Canad. J. Math. 53, no. 5 (2001), 944-978. [9] J. Ludwig, C. Molitor-Braun and L. Scuto, On Fourier's inversion theorem in the context of nilpotent Lie groups, Acta Sci. Math. (Szeged) 73 (2007), 547-591. [10] J. Ludwig and D. Muller, Sub-Laplacians of Holomorphic Lp-type on Rank One AN-Groups and Related Solvable Groups, J. Func. Anal. 170 (2000), 366-427. [11] J. Ludwig and H. Zahir, On the nilpotent ∗-Fourier transform, Lett. Math. Phys. 30 (1994), 23-34. [12] D. Poguntke, Uber das Synthese-Problem fur nilpotente Liesche Gruppen, Math. Ann. 269 (1984), 431-467. Ying-Fen Lin, Pure Mathematics Research Centre, Queen's University Belfast, Belfast, BT7 1NN, U.K. E-mail: [email protected] Jean Ludwig, Universit´e de Lorraine, Institut Elie Cartan de Lorraine, UMR 7502, Metz, F-57045, France. E-mail: [email protected] Carine Molitor-Braun, Unit´e de Recherche en Math´ematiques, Universit´e du Luxembourg, 6, rue Coudenhove-Kalergi, L-1359 Luxembourg, Luxembourg. E-mail: [email protected] 26
1309.2393
2
1309
2013-11-22T18:20:02
A Survey on Reproducing Kernel Krein Spaces
[ "math.FA" ]
This is a survey on reproducing kernel Krein spaces and their interplay with operator valued Hermitian kernels. Existence and uniqueness properties are carefully reviewed. The approach we follow in this survey uses a more abstract but very useful concept of linearization or Kolmogorov decomposition, as well as the underlying concept of Krein space induced by a selfadjoint operator and that of Krein space continuously embedded. The operator range feature of reproducing kernel spaces is emphasized. We include a careful presentation of Hermitian kernels on complex regions that points out a universality property of the Szego kernels with respect to reproducing kernel Krein spaces of holomorphic functions.
math.FA
math
A SURVEY ON REPRODUCING KERNEL KREIN SPACES AURELIAN GHEONDEA Abstract. This is a survey on reproducing kernel Kreın spaces and their interplay with operator valued Hermitian kernels. Existence and uniqueness properties are carefully re- viewed. The approach used in this survey involves the more abstract, but very useful, concept of linearisation or Kolmogorov decomposition, as well as the underlying concepts of Kreın space induced by a selfadjoint operator and that of Kreın space continuously embedded. The operator range feature of reproducing kernel spaces is emphasised. A careful presentation of Hermitian kernels on complex regions that points out a univer- sality property of the Szego kernels with respect to reproducing kernel Kreın spaces of holomorphic functions is included. Acknowledgement. Work supported by a grant of the Romanian National Authority for Scientific Research, CNCS UEFISCDI, project number PN-II-ID-PCE-2011-3-0119. Key words and phrases. Kreın space, reproducing kernel, Hermitian kernel, linearisation, Kolmogorov decomposition, induced space, holomorphy. 2010 Mathematics Subject Classification. 46C07, 46C20, 47B32, 47B50, 32A25. 3 1 0 2 v o N 2 2 ] . A F h t a m [ 2 v 3 9 3 2 . 9 0 3 1 : v i X r a 1. Introduction A kernel K on a set X and with operator entries is the analog of an abstract matrix double indexed on X and for which each entry is a bounded linear operator acting between appropriate spaces. The adjoint K ♯ of the kernel K can be defined by analogy with the adjoint of matrices. A Hermitian kernel K is defined by the property K = K ♯. Roughly speaking, a reproducing kernel Kreın space on a set X is a Kreın space R of functions on X for which there exists a Hermitian kernel with the property that the evaluations of the functions in R can be calculated in terms of the kernel K. One of the main problems is that of associating a reproducing kernel Kreın space R to a given Hermitian kernel K. The classical theory says that for any reproducing kernel Hilbert space R its reproducing kernel K is positive semidefinite and uniquely determined by R. Conversely, to any positive semidefinite kernel K there is associated a unique reproducing kernel Hilbert space such that K is its reproducing kernel. These facts can be proven also for a slightly more general situation relating in a similar fashion reproducing kernel Pontryagin spaces to Hermitian kernels with finite negative (or positive) signatures. For genuine indefinite Hermitian kernels the correspondence with reproducing kernel Kreın spaces is much more complicated. There is a variety of characterisations of those Hermitian kernels that produce reproducing kernel Kreın spaces in terms of boundedness with respect to positive semidefinite kernels and in terms of decomposability as a difference of two positive semidefinite kernels. Uniqueness of the reproducing kernel Kreın space, provided that it exists, is also problematic. Characterisations of uniqueness are available in terms of lateral spectral gaps as well as in terms of maximal uniformly positive subspaces. 1 2 A. GHEONDEA When Hermitian kernels on regions in either a single complex variable or several complex variables have certain holomorphy properties, existence of reproducing kernel Kreın spaces is guaranteed. This fact is related to a certain universality property of the Szego kernel that allows the construction of the corresponding reproducing kernel Kreın space inside the Hardy space or, respectively, the Drury-Arveson space. The approach used in this survey involves a more abstract, but very useful, concept of linearisation or Kolmogorov decomposition, the underlying concept of Kreın spaces induced by selfadjoint operators, as well as the concept of Kreın spaces continuously embedded, pointing out the operator range feature of reproducing kernel spaces. 2. Kreın Spaces and their Linear Operators A Kreın space K is a complex linear space on which it is defined an indefinite scalar product [·, ·] such that K is decomposed in a direct sum (2.1) K = K+[ +]K− in such a way that K± are Hilbert spaces with scalar products ±[·, ·], and the direct sum in (2.1) is orthogonal with respect to the indefinite scalar product [·, ·], i.e. K+ ∩K− = {0} and [x+, x−] = 0 for all x± ∈ K±. The decomposition (2.1) gives rise to a positive definite scalar product h·, ·i by setting hx, yi := hx+, y+i−hx−, y−i, where x = x++x−, y = y++y−, and x±, y± ∈ K±. The scalar product h·, ·i defines on K a structure of Hilbert space. Subspaces K± are orthogonal with respect to the scalar product h·, ·i, too. One denotes by P± the corresponding orthogonal projections onto K±, and let J = P+ − P−. The operator J is a symmetry, i.e. a selfadjoint and unitary operator, J ∗J = JJ ∗ = J 2 = I. Unless otherwise specified, any Kreın space is considered as a Banach space with the norm given by an arbitrary norm induced by the positive definite inner product associated to a fundamental decomposition. For two Kreın spaces K1 and K2, we denote by L(K1, K2) the Banach space of all bounded linear operators T : K1 → K2. Given a Kreın space (K, [·, ·]), the cardinal numbers (2.2) κ+(K) = dim(K+), κ−(K) = dim(K−), do not depend on the fundamental decomposition and they are called, respectively, the geometric ranks of positivity/negativity of K. The operator J is called a fundamental symmetry of the Kreın space K. Note that [x, y] = hJx, yi, (x, y ∈ K). If T is a densely defined operator from a Kreın space K1 to another Kreın space K2, the adjoint of T can be defined as an operator T ♯ defined on the set of all y ∈ K2 for which there exists hy ∈ K1 such that [T x, y] = [x, hy], and T ♯y = hy. In addition, T ♯ = J1T ∗J2, where T ∗ denotes the adjoint operator of T with respect to the Hilbert spaces (K1, h·, ·iJ1) and (K2, h·, ·iJ2). The symbol ♯ denotes the adjoint when at least of the spaces K1 or K2 is indefinite. In the case of an operator T defined on the Kreın space K, T is called symmetric if T ⊆ T ♯, i.e. if the relation [T x, y] = [x, T y] holds for each x, y ∈ Dom(T ), and T is called selfadjoint if T = T ♯. In this survey, a bit of geometry of Kreın spaces is used. Thus, a (closed) subspace L of a Kreın space K is called regular if K = L + L⊥, where L⊥ = {x ∈ K [x, y] = 0 for all y ∈ L}. Regular spaces of Kreın spaces are important since they are exactly the analogue of Kreın subspaces, that is, if we want L be a Kreın space with the restricted indefinite inner product and the same strong topology, then it should be regular. REPRODUCING KERNEL KREIN SPACES 3 In addition, let us recall that, given a subspace L of a Kreın space, one calls L non- negative (positive) if the inequality [x, x] ≥ 0 holds for x ∈ L (respectively, [x, x] > 0 for all x ∈ L \ {0}). Similarly one defines non-positive and negative subspaces. A subspace L is called degenerate if L ∩ L⊥ 6= {0}. Regular subspaces are non-degenerate. As a consequence of the Schwarz inequality, if a subspace L is either positive or negative it is nondegenerate. A remarkable class of subspaces are those regular spaces that are either positive or negative, for which the terms uniformly positive, respectively, uniformly negative are used. These notions can be defined for linear manifolds also, that is, without assuming closedness. A linear operator V defined from a subspace of a Kreın space K1 and valued into another Kreın space K2 is called isometric if [V x, V y] = [x, y] for all x, y in the domain of V . Note that isometric operators between genuine Kreın spaces may be unbounded and different criteria of boundedness are available. However, in this presentation, a unitary operator between Kreın spaces means that it is a bounded isometric operator that have bounded inverse. 3. Hermitian Kernels Let X be a nonempty set and let H = {Hx}x∈X be a family of Kreın spaces with inner products denoted by [·, ·]Hx. A mapping K defined on X × X such that K(x, y) ∈ L(Hy, Hx) for all x, y ∈ X is called an H-kernel on X. In case Hx = H for all x ∈ X, where H is some fixed Kreın space, one talks about an H-kernel on X, while, even more particularly, if H = C, one talks about a scalar kernel on X, or simply a kernel on X. To any H-kernel K one associates its adjoint K ♯ defined by K ♯(x, y) = K(y, x)♯, for all x, y ∈ X. The H-kernel K is called Hermitian if K = K ♯, that is, (3.1) K(x, y) = K(y, x)♯, x, y ∈ X. Denote by F (H) the set of all H-vector fields f on X, that is, f = {fx}x∈X such that fx ∈ Hx, for all x ∈ X, and let F0(H) denote the set of all f ∈ F (H) of finite support, that is, the set supp(f ) = {x ∈ X fx 6= 0} is finite. Alternatively, one can view any f ∈ F (H) as a function f : X →Sx∈X Hx such that f (x) ∈ Hx for all x ∈ X. If K is a Hermitian H-kernel then one can introduce on F0(H) an inner product [·, ·]K defined by (3.2) [f, g]K = Xx,y∈X [K(x, y)f (y), g(x)]Hx, f, g ∈ F0(H). The H-kernel K is called positive semidefinite if (3.3) Xx,y∈X [K(x, y)h(y), h(x)]Hx ≥ 0, h ∈ F0(H). Every positive semidefinite H-kernel is Hermitian. Also, a Hermitian H-kernel is positive semidefinite if and only if the corresponding inner product in (3.2) is nonnegative. Let us denote by Kh(H) the class of all Hermitian H-kernels and by K+(H) the subclass of all positive semidefinite H-kernels. On Kh(H) one defines addition, subtraction and multiplication with real numbers in a natural way. Moreover, on Kh(H) one has a natural 4 A. GHEONDEA partial order defined as follows: if H, K ∈ Kh(H) then H ≤ K means [f, f ]H ≤ [f, f ]K, for all f ∈ F0(H). With this definition one gets (3.4) K+(H) = {H ∈ Kh(H) H ≥ 0}, and K+(H) is a strict cone of Kh(H), that is, it is closed under addition and multiplication with nonnegative numbers, and K+(H)T −K+(H) = {0}, where 0 denotes the null kernel. More generally, one can define the signatures of K, denoted by κ±(K), as the posi- tive/negative signatures of the inner product [·, ·]K. Then, if κ(K) = min{κ−(K), κ+(K)} denotes the definiteness signature of K, the Hermtian H-kernel K is called quasi semi- definite if κ(K) < ∞, that is, either κ−(K) or κ+(K) is finite. A pairing (·, ·)F can be defined for arbitrary f, g ∈ F (H), provided at least one of f and g has finite support, by (3.5) (f, g)F = Xx∈X [f (x), g(x)]Hx. When restricted to F0(H) this pairing becomes a nondegenerate inner product. To each H-kernel K one associates the convolution operator, denoted also by K, and defined by (3.6) K : F0(H) → F (H), K(x, y)f (y), f ∈ F0(H). (Kf )(x) =Xy∈X Then (3.7) [f, g]K = (Kf, g)F , f, g ∈ F0(H). Consequently, the kernel K is positive semidefinite (Hermitian) if and only if the corre- sponding convolution operator K is positive semidefinite (Hermitian), that is, (Kf, g)F ≥ 0 ((Kf, g)F = (f, Kg)F), for all f, g ∈ F0(H). Similar assertions can be made about the signatures κ±(K), with appropriate definitions of signatures of Hermitian operators on inner product spaces. At this point, it is worth noting that there is no restriction of generality if one assumes that all spaces Hx are Hilbert. To see this, fixing a fundamental symmetry Jx on each Kreın space H, one can refer to the Hilbert spaces (Hx; h·, ·iJx) and, one considers the kernel H on X defined by H(x, y) = JxK(x, y)Jy, for all x, y ∈ X. Taking into account that A♯ = JxA∗Jy for any bounded linear operator A : Hx → Hy, it follows that all notions like Hermitian, positive semidefinite, signatures, etc. defined for the kernel K, have word- for-word transcriptions for the kernel H. However, since this survey is focusing on the indefinite case, allowing right from the beginning Kreın spaces Hx brings more symmetry and simpler formulas. A reproducing kernel inner product space, with respect to the set X, the collection of Kreın spaces H = {Hx}x∈X, and H-kernel K, is an inner product space (R, [·, ·]R) subject to the following conditions: (rk1) R ⊆ F (H). (rk2) For all y ∈ X and all h ∈ Hy the map X ∋ x 7→ K(x, y)h ∈ Hx belongs to R. (rk3) [f (x), h]Hx = [f, K(·, x)h]R, for all f ∈ R, x ∈ X, and h ∈ Hx. The axiom (rk3) is usually called the reproducing property while the H-kernel K is called the reproducing kernel of R. Note that K is necessarily a Hermitian H-kernel. REPRODUCING KERNEL KREIN SPACES 5 According to the axiom (rk2), it is useful to consider the notation Ky = K(·, y) : X → L(Hy, R), for y ∈ X. An immediate consequence of the axioms (rk1)–(rk3) is that the inner product [·, ·]R is nondegenerate. Also, recalling that on any nondegenerate inner product space the weak topology is separated, the following minimality property holds (rk4) The span of {KxHx x ∈ X} is weakly dense in R. Also, the reproducing kernel is uniquely determined by the reproducing kernel inner prod- uct space. Given a Hermitian H-kernel K on X, consider the subspace R0(H) of F (H) spanned by Kxh, for all x ∈ X and all h ∈ Hx, on which one can define the inner product (3.8) [ mXi=1 Kxihi, nXj=1 Kyj kj]R0 = mXi=1 nXj=1 [K(yj, xi)hi, kj]Hyj , This definition can be proven to be correct: vectors in R0(H) may have different rep- resentations as linear combinations of Kyk but the definition in (3.8) is independent of these. The subspace R0(H) of F (H) is the range of the convolution operator K defined at (3.6) and the inner product [·, ·]R0 is nondegenerate. In addition, (R0(H); [·, ·]R0) is a reproducing kernel inner product space with reproducing kernel K. In case the reproducing kernel inner product space (R; [·, ·]R) is a Kreın space, one talks about a reproducing kernel Kreın space. The uniqueness of the reproducing kernel of a Kreın space has a stronger characterisation. Theorem 3.1. Let K be a Kreın space of H-valued vector fields on X, that is, K ⊆ F (H). For each x ∈ X consider the linear operator E(x) : K → Hx of evaluation at x, that is, E(x)f = f (x) for all f ∈ K. Then, K has a reproducing kernel if and only if E(x) is bounded for all x ∈ X. In this case, the reproducing kernel of K is (3.9) K(x, y) = E(x)E(y)♯, x, y ∈ X, hence uniquely determined by K. With the notation as in Theorem 3.1, it is useful to note that, if K is a reproducing kernel Kreın space with reproducing kernel K, then Kx can be viewed as a linear operator Hy → K and Kx = E(x)♯ for all x ∈ X. In particular, (3.10) K(x, y) = K ♯ xKy, x, y ∈ X. Conversely, given a Hermitian H-kernel K on X, the questions on existence and unique- ness of a reproducing kernel Kreın space R ⊆ F (H) such that K is its reproducing kernel are much more difficult. If (R; [·, ·]R) is a reproducing kernel Kreın space with reproducing kernel K, in view of the axioms (rk1)–(rk3) and the minimality property (rk4), the inner products [·, ·]R and [·, ·]R0 coincide on R0(H) and R0(H) is dense in R. Thus, existence of a reproducing kernel Kreın space associated to a given H-kernel K on X depends heavily on the possibility of "completing" the inner product space (R0(H); [·, ·]R0) to a Kreın space R inside F (H), which is a core problem in the theory of indefinite inner product spaces. In order to tackle these questions, let us note that if R is a reproducing kernel Hilbert space with reproducing kernel K, then K should be positive semidefinite. Conversely, if K is a positive semidefinite H-kernel on X, then the inner product space (R0, [·, ·]R0) 6 A. GHEONDEA defined at (3.8) is a pre-Hilbert space and the existence of a reproducing kernel Hilbert space R with reproducing kernel K depends on whether (R0, [·, ·]R0) has a completion inside of F (H). Actually, this is always the case and, moreover, uniqueness holds as well, but these two facts are slightly more general, namely, they are true if K is quasi semidefinite, in which case it is obtained a unique reproducing kernel Pontryagin space, cf. Section 5. In this survey, existence and uniqueness of reproducing kernel Kreın spaces associated to Hermitian kernels are approached through the more abstract, but very useful, concept of linearisation or Kolmogorov decomposition. By definition, a linearisation, sometimes called Kolmogorov decomposition, of the H-kernel K is a pair (K; V ), subject to the following conditions: (kd1) K is a Kreın space and V (x) ∈ L(Hx, K) for all x ∈ X. (kd2) K(x, y) = V (x)♯V (y) for all x, y ∈ X. The linearisation (K; V ) is called minimal if the following condition holds as well: (kd3) K = Wx∈X V (x)Hx. Two linearisations (Kj; Vj), j = 1, 2, of the same H-kernel K are called unitary equiv- alent if there exists a Kreın space bounded unitary operator U : K1 → K2 such that V2(x) = UV1(x) for all x ∈ X. There is a close connection between the notion of reproducing kernel Kreın space with reproducing H-kernel K and that of minimal linearisation of K. Proposition 3.2. (1) Let R be a reproducing kernel Kreın space with reproducing H- kernel K. Then, letting (3.11) V (x) = Kx = K(·, x) ∈ L(Hx, R), the pair (R; V ) is a minimal linearisation of the H-kernel K. Conversely, letting (K; V ) be a minimal linearisation of the H-kernel K, then, (3.12) R = {V (·)♯ k ∈ K}, is a vector subspace of F (H) which, with respect to the the inner product defined by (3.13) [V (·)♯h, V (·)♯k]R = [h, k]K, h, k ∈ K, is a Kreın space with reproducing kernel K. (2) In the correspondence defined at (3.12) and (3.13), two unitary equivalent minimal realisations of the same H-kernel K produce the same reproducing kernel Krein space and hence, the correspondence between reproducing kernel Krein spaces and minimal linearisa- tions is one-to-one, provided that unitary equivalent minimal linearisations are identified. 4. Some Examples Example 4.1. Matrices. Let X = {1, 2, . . . , n} and Hx = C for each x ∈ X. A kernel on X is simply a map K : X × X → C and hence, it can be identified with the n × n complex matrix [K(i, j)]n i,j=1 is Hermitian. The vector space F = F0 = Cn hence, in view of the definition (3.6), the convolution operator K : F → F is simply the linear operator K : Cn → Cn associated to the matrix K. i,j=1. The kernel K is Hermitian if and only the matrix [K(i, j)]n REPRODUCING KERNEL KREIN SPACES 7 If K is a Hermitian kernel on X then the inner product [·, ·]K is the familiar inner i,j=1. The signatures κ±(K) coincide, i,j=1, i,j=1 product associated to the Hermitian matrix [K(i, j)]n respectively, with the number of positive/negative eigenvalues of the matrix [K(i, j)]n counted with multiplicities. K is positive semidefinite if and only if the matrix [K(i, j)]n is positive semidefinite. Assuming that the kernel K on X is Hermitian, for each j = 1, . . . , n, the "map" i=1 in Cn = F . The vector space R0, the Kj : X → C is simply the column vector [K(i, j)]n range of the convolution operator K, is the vector subspace of Cn generated by all the column vectors [K(i, j)]n i=1. In this particular case, R0 is the reproducing kernel Kreın space with reproducing kernel K. The inner product [·, ·]R0 is defined at (3.8). Example 4.2. Operator Block Matrices. Let X = {1, 2, . . . , n} but, this time, for each x ∈ X one denotes by Hx a Hilbert space and let H = {Hx}x∈X. An H-kernel on X, originally defined as a map K on X × X such that K(x, y) ∈ L(Hy, Hx), is naturally identified with the operator block matrix [K(i, j)]n i,j=1. The vector space F (H) = F0(H) together with its natural inner product (·, ·)F defined at (3.5) is naturally identified with i=1 Hi. Then, the convolution operator K associated to the H- kernel K is identified with the bounded linear operator K : H → H naturally associated to the operator block matrix [K(i, j)]n i,j=1. For each j = 1, . . . , n, the map Kj = K(·, j) is the operator block column matrix [K(i, j)]n the Hilbert space H =Ln i=1 : Hj → H. The H-kernel K on X is Hermitian if and only if the corresponding operator block matrix [K(i, j)]n i,j=1 is Hermitian, if and only if the convolution operator K ∈ L(H) is selfadjoint (Hermitian). In this case, the vector space R0(H) is identified with the range of the convolution operator K as a subspace of H and is spanned by the ranges of the operator block column matrices Kj = [K(i, j)]n i=1 : Hj → H, for j = 1, . . . , n. The inner product [·, ·]R0 is defined as in (3.8). Example 4.3. The Hardy space H 2(D). Let D denote the open unit ball in the complex field and consider the Szego kernel (4.1) S(z, w) = 1 1 − zw = ∞Xn=0 wnzn, z, w ∈ D, the series converging absolutely and uniformly on any compact subset of D × D. Clearly, S is a Hermitian scalar kernel on D and for every w ∈ D Sw(z) = S(z, w) = ∞Xn=0 wnzn, z, w ∈ D, the series converging absolutely and uniformly on any compact subset of D, Sw is a holomorphic function on D, and its Taylor coefficients are 1, w, w2, . . .. In order to describe the reproducing kernel inner product space associated to the Szego kernel S, in view of (3.8), one has to consider the vector space generated by the functions Sw, w ∈ D, and complete it with respect to the inner product (4.2) hSu, Swi = Su(w) = S(w, u) = ∞Xn=0 unwn, u, v ∈ D. 8 A. GHEONDEA This completion is called the Hardy space, denoted by H 2(D), and in view of (4.2), is (4.3) H 2(D) = {f f (z) = ∞Xn=0 anzn, 2Xn=0 an2 < ∞}. Since H 2(D) is a Hilbert space, it follows that the Szego kernel is positive semidefinite. The Hardy space H 2(D) has some other special properties, among which, a distinguished role is played by the boundary values of its functions. More precisely, for each f ∈ H 2(D), (4.4) there exists ef a function defined on the unit circle T = ∂D, such that The function ef is uniquely determined by f , a.e. on T, and ef ∈ L2(T). Usually, there is no distinction between a function f in H 2(D) and ef , that is, the tilde sign is not used at all. In this way, an isometric embedding of H 2(D) into L2(T) is defined, in particular, ef (eit) = lim a.e. t ∈ [0, 2π). f (reit), r→1− (4.5) hf, giH 2(D) = hf, giL2(T) = f (eit)g(eit)dt, f, g ∈ H 2(D). 1 2πZ 2π 0 With respect to this embedding, H 2(D) is identified with the subspace of L2(T) of all functions f whose Fourier coefficients of negative index vanish. Example 4.4. The Drury-Arveson Space. Let Br(ξ) be the open ball of radius r and n=1 ηnξn for center ξ in the Hilbert space G = CN , with inner product hξ, ηiG = η · ξ =PN n=1 in CN . We write Br instead of Br(0). The Szego kernel is n=1 and η = (ηn)N ξ = (ξn)N (4.6) S(ξ, η) = 1 1 − hξ, ηiG , ξ, η ∈ B1, and note that S is a scalar Hermitian kernel on B1. We now describe a minimal lineari- G⊗n be the Fock space associated to G = CN , where G⊗0 = C sation of S. Let F (G) = and G⊗n is the n-fold Hilbert space tensor product of G with itself, hence F (G) is a Hilbert space. Let ∞Ln=0 (4.7) Pn = 1 n! Xπ∈Sn π be the orthogonal projection of G⊗n onto its symmetric part, where π(ξ1 ⊗ . . . ⊗ ξn) = ξπ−1(1) ⊗ . . . ⊗ ξπ−1(n) for any element π of the permutation group Sn on n symbols. Recall Pn)F (G). For ξ ∈ B1 set ξ⊗0 = 1 and let that the symmetric Fock space is F s(G) = ( ∞Ln=0 G⊗n =Xn≥0 ξ⊗n denote the n-fold tensor product ξ ⊗ . . . ⊗ ξ, n ≥ 1. Note that 1 kMn≥0 ξ⊗nk2 F (G) =Xn≥0 kξ⊗nk2 kξk2n G = . 1 − kξk2 G Hence Ln≥0 (4.8) ξ⊗n ∈ F s(G) and one can define the mapping VS from B1 into F s(G), VS(ξ) =Mn≥0 ξ⊗n, ξ ∈ G. REPRODUCING KERNEL KREIN SPACES 9 The pair (F s(G); VS) is a minimal linearisation of the kernel S. In order to see this, VS(ξ) is also viewed as a bounded linear operator from C into F s(G) by VS(ξ)λ = λVS(ξ), λ ∈ C, so that, for ξ, η ∈ B1, VS(η)∗VS(ξ) = hVS(ξ), VS(η)iF s(G) =Xn≥0 hξ⊗n, η⊗niG⊗n =Xn≥0 (η · ξ)n = 1 1 − η · ξ = S(ξ, η). In particular, this shows that the Szego kernel is positive semidefinite. The set {VS(ξ) ξ ∈ B1} is total in F s(G) since for n ≥ 1 and ξ ∈ G one has dn dtn V (tξ)t=0 = n!ξ⊗n. The reproducing kernel Hilbert space associated to the Szego kernel S, called the Drury- Arveson space and denoted by H 2(B1), is given by the completion of the linear space generated by the functions Sη = S(·, η), η ∈ B1, with respect to the inner product defined by hSη, SξiH 2(B1) = S(ξ, η), see (3.8). We use the multiindex notation: for n = (n1, . . . , nN ) ∈ NN N . Then, anξn, a function f holomorphic in B1, with Taylor series representation f (z) = Pn∈NN 0 , let n = n1 + · · · + nN , n! = n1! · · · nN !, and ξn = ξn1 belongs to H 2(B1) if and only if 1 · · · ξnN 0 (4.9) kf k2 H 2(B1) = Xn∈NN 0 n! n! an2 < ∞. Note that there exists a unitary operator Φ from the Drury-Arveson space H 2(B1) onto F s(G) such that ΦSξ = VS(ξ), ξ ∈ B1. If N = 1, the Drury-Arveson space coincides with the Hardy space H 2(D) described at the previous example. Example 4.5. Holomorphic Kernels. Given two Kreın spaces G and H let Ω be a sub- region of the open unit disc D in the complex plane and Θ : Ω → L(G, H). One considers the following kernels (4.10) (4.11) (4.12) KΘ(z, w) = K eΘ(z, w) = I − Θ(z)Θ(w)♯ 1 − zw I − eΘ(z)eΘ(w)♯ 1 − zw , , DΘ(z, w) =  KΘ(z, w) eΘ(z) − eΘ(w) z − w Θ(z) − Θ(w) z − w K eΘ(z, w)   , where eΘ(z) = Θ(z)♯. These are operator valued Hermitian holomorphic kernels that are of interest in connection to Schur classes and their generalisations. The classical Schur class corresponds to the case when G and H are Hilbert spaces and the Schur kernel KΘ is positive semidefinite. In case G and H are Kreın or even genuine Pontryagin spaces, in order to define the Schur class, positive semidefiniteness has to be imposed on all KΘ, K eΘ and DΘ. Given κ ∈ N, the generalized Schur class corresponds to the requirement that the negative signatures of each kernel KΘ, K eΘ and DΘ are κ. 10 A. GHEONDEA When G = H and for appropriate regions in the complex field, the Carath´eodory kernel CΘ (4.13) CΘ(z, ζ) = 1 2 Θ(z) − Θ(ζ)♯ 1 − ζz , as well as the Nevanlinna kernel NΘ (4.14) NΘ(z, ζ) = Θ(z) − Θ(ζ)♯ z − ζ , are of special interest. As in the case of the Schur classes, positive semidefiniteness of the corresponding kernels define the Carath´eodory class and, respectively, the Nevanlinna class of holomorphic functions. Generalized classes correspond to the case when the appropriate kernels have fixed negative signatures. Example 4.6. Toplitz Kernels. Let H be a Kreın space and X = Z, the set of integer numbers. An H-kernel on Z, K : Z × Z → L(H), is called a Toplitz Kernel if K(i, j) = T (i − j) for some function T : Z → L(H), called the symbol of K. One considers the set T(H) of all Toplitz H-kernels. In the following it is considered the subclass of Toplitz Hermitian H-kernels Th(H) and the subclass of Toplitz positive semidefinite H-kernels T+(H). One notes that Th(H) is closed under addition, subtraction and (left and right) multiplication with bounded operators on H. Also, T+(H) is a strict cone of Th(H). Considering the complex vector space F0(H) of all functions h : Z → H with finite support and for an arbitrary Hermitian kernel H ∈ Kh(H) one associates the inner product space (F0, [·, ·]H) as in the previous sections. On the vector space F0(H) one considers two operators, the forward shift S+ defined by (S+h)(n) = h(n − 1), for all h ∈ F0(H) and n ∈ Z, and the backward shift S− defined by (S−h)(n) = h(n + 1), for all h ∈ F0(H) and all n ∈ Z. If H ∈ Kh(H) then it is a Toplitz kernel if and only if [S+h, g]H = [h, S−g]H, f, g ∈ F0(H). If H is a Hermitian Toplitz H-kernel then both S+ and S− are isometric with respect to the inner product [·, ·]H, that is, for all h, g ∈ F0(H) the equalities [S+h, S+g]H = [h, g]H and [S−h, S−g]H = [h, g]H hold. The converse is also true, if either S+ or S− is isometric with respect to the Hermitian H-kernel H then H is Toplitz. Let H be a Toplitz Hermitian H-kernel. A Naımark dilation of H is, by definition, a triple (U, Q; K) with the following properties: (nd1) K is Kreın space, U ∈ L(K) is a unitary operator, and Q ∈ L(H, K). (nd2) K = Wn∈Z U nQH. (nd3) H(i, j) = Q♯U i−jQ, i, j ∈ Z. If the Toplitz Hermitian H-kernel H has a Naımark dilation (U, Q; K), letting V (n) = U nQ ∈ L(H, K) it is readily verified that the pair (V ; K) is a linearisation of H. REPRODUCING KERNEL KREIN SPACES 11 5. Quasi Semidefinite Kernels In the following, for simplicity and because most of the applications presented in this survey do not require the full generality as considered in Section 3, there is consider only the case of H-kernels, that is, Hx = H for some fixed Kreın space H. A Hermitian kernel H : X × X → L(H) is associated to an indefinite inner product space [·, ·]H on F (H), the vector space of all complex valued functions f : X → H, as in (3.2). Recall that κ±(H), the signatures of H, are defined as the positive/negative signatures of the inner product [·, ·]H, while its rank of indefiniteness is defined by κ(H) = min{κ−(H), κ+(H)}. A Hermitian H-kernel H is quasi semidefinite if κ(H) < ∞, that is, either κ−(H) or κ+(H) is finite. Theorem 5.1. Let H be quasi semidefinite H-kernel on X. Then: (a) H admits a minimal linearisation to a Pontryagin space (K; V ) with κ(H) = κ(K), unique up to a unitary equivalence. (b) There exists a unique reproducing kernel Pontryagin space R on X and with re- producing kernel H. On the vector space F (H) of H-fields on X, the inner product [·, ·]H, see (3.2) is defined. Since κ(H) is finite, the inner product space (F (H); [·, ·]H) is decomposable. To make a choice, assume that κ−(H) < ∞, hence (5.1) F (H) = F−[+]F0[+]F+, where F0 is the isotropic subspace, F± are positive/negative subspaces, and dim(F−) = κ−(H) < ∞. Factoring out F0 one can assume, without loss of generality, that F0 = 0. Then, let K+ denote the completion of (F+; [·, ·]H) to a Hilbert space and then K = F−[+]K+ is a Pontryagin space with κ−(K) = κ−(H). For arbitrary x ∈ X, the linear operator V (x) : H → K is defined by assigning to each vector h ∈ H the H-vector field f : X → H given by f (x) = h and f (y) = 0 for all y ∈ X, y 6= x. Then (K; V ) is a minimal lineariasation of H. The uniqueness of minimal linearisations of H, modulo unitary equivalence, follows from the fact that any dense linear subspace of a Pontryagin space with finite negative/positive signature contains a maximal negative/positive subspace and the continuity of isometric densely defined operators between Pontryagin spaces with the same negative/positive signature. From Proposition 3.2 (1), once one gets a minimal linearisation (K; V ) of H one imme- diately obtains a reproducing kernel Pontryagin space R with reproducing kernel H as defined in (3.12) and (3.13). The uniqueness of the reproducing kernel Pontryagin space R follows from Proposition 3.2 (2) and the uniqueness, modulo unitary equivalence, of the minimal linearisation of H. There is a more direct but longer way of constructing the reproducing kernel Kreın space associated to a quasi positive semidefinite H-kernel H, by considering the vector space R0(H) linearly generated by Hxh, for x ∈ X and h ∈ H on which the inner product [·, ·]R0 is defined as in (3.8) in such a way that (R0(H); [·, ·]R0) is a reproducing kernel inner product space with H its reproducing kernel. The inner product space (R0(H); [·, ·]R0) is nondegenerate and its negative signature is the same with κ−(H), hence finite. Then R0(H) is decomposable, hence R0(H) = R−[+]R0,+, where dim(R−) = κ−(H) < ∞ 12 A. GHEONDEA is negative definite, while R0,+ is positive definite. It can be proven that R0,+ has a completion to a Hilbert space R+ inside F (H) and then R = R−[+]R+ is the reproducing kernel Kreın space of H. The next theorem points out a property of propagation to arbitrarily large domains of holomorphy for quasi semidefinite holomorphic kernels. Theorem 5.2. Let K be a Hermitian holomorphic H-kernel on some region Ω and let Ω0 be a subregion of Ω. If κ−(K0) < ∞, where K0 is the restriction of K to Ω0 × Ω0, then κ−(K) = κ−(K0). The idea of proof is to use Cauchy's Theorem in order to prove the propagation property for positive semidefinite kernels, first for disks around the origin of the complex plane, then for union of regions for which the kernel is positive definite, and finally to use the decomposition of K = K+ − K−, where K± are positive semidefinite kernels and K− has a reproducing Hilbert space of dimension κ−(K), that can be obtained from (5.1). 6. Induced Kreın Spaces Let (H, h·, ·iH) be a Hilbert space and consider a bounded selfadjoint operator A on H. A new inner product on H is defined by (6.1) [h, k]A = hAh, kiH, h, k ∈ H. In this section the properties of some Kreın spaces associated with this inner product are described. By definition, a Kreın space induced by A is a pair (K, Π), where (ik1) K is a Kreın space and Π ∈ L(H, K) has dense range. (ik2) [Πx, Πy]K = hAx, yiH, for all x, y ∈ H. There exist at least two main constructions of Kreın spaces induced by selfadjoint operators, and they are related by certain unitary equivalences. Two Kreın spaces (Ki, Πi), i = 1, 2, induced by the same A, are unitary equivalent if there exists a unitary operator U ∈ L(K1, K2) such that UΠ1 = Π2. Example 6.1. The Induced Kreın Space (KA; ΠA). Let A denote the selfadjoint operator with respect to the Hilbert space (H, h·, ·iH). Let H− and H+ be the spectral subspaces corresponding to the semiaxis (−∞, 0) and, respectively, (0, +∞) and the operator A. Then one gets the decomposition H = H− ⊕ Ker A ⊕ H+. Note that (H−, −[·, ·]A) and (H+, [·, ·]A) are positive definite inner product spaces and hence they can be completed to Hilbert spaces K− and, respectively, K+. We can build the Kreın space (KA, [·, ·]A) by letting (6.2) KA = K−[+]K+. The operator ΠA ∈ L(H, KA) is, by definition, the composition of the orthogonal pro- jection of H onto H ⊖ Ker A with the embedding of H ⊖ Ker A into KA. With these definitions, it is readily verified that (KA, ΠA) is a Kreın space induced by A. REPRODUCING KERNEL KREIN SPACES 13 In order to take a closer look at the strong topology of the Kreın space KA, consider the seminorm H ∋ x 7→ kA1/2xk = hA1/2x, A1/2xi1/2. The kernel of this seminorm is ex- actly Ker A and the completion of H⊖Ker A with respect to this norm is exactly the space KA. Moreover, the strong topology of KA is induced by the extension of this seminorm. The positive definite inner product on KA associated with the norm kA1/2 · k is hA·, ·i. Hence, if A = SAA is the polar decomposition of A and SA denotes the corresponding selfadjoint partial isometry, then it follows that SA can be extended by continuity to KA and this extension is exactly the fundamental symmetry of KA corresponding to hA·, ·i. Example 6.2. The Induced Kreın Space (BA; ΠBA). With notation as in Example 6.1, consider the polar decomposition of A = SAA as in Example 6.1. Define the space BA = Ran(A1/2) endowed with the positive definite inner product h·, ·iBA given by (6.3) hA1/2x, A1/2yiBA = hPH⊖Ker Ax, yiH, x, y ∈ H. This positive definite inner product is correctly defined and (BA, h·, ·iBA) is a Hilbert space. To see this, just note that the operator A1/2 : H ⊖ Ker A → BA is a Hilbert space unitary operator. On BA one defines the inner product [·, ·]BA by (6.4) Since the operator A1/2 and SA commute it follows that [A1/2x, A1/2y]BA = hSAx, yiH, x, y ∈ H. [a, b]BA = hSAa, biBA, a, b ∈ BA. The operator SABA is a symmetry. We define now a linear operator ΠBA : H → BA by (6.5) ΠBAh = Ah, h ∈ H. It follows that (BA, [·, ·]BA) is a Kreın space induced by A. The Kreın space induced by a selfadjoint operator as in Example 6.2 is a genuine operator range subspace. More precisely, to any Kreın space K continuously embedded in the Hilbert space H one associates a selfadjoint operator A ∈ L(H) in the following way: let ι : K → H be the inclusion operator which is supposed bounded and take A = ιι♯ ∈ L(H). Clearly A is selfadjoint and (K, ι♯) is a Kreın space induced by A. Conversely, from Example 6.2 it is easy to see that BA is a Kreın space continuously embedded in H. The connection between the induced Kreın spaces (KA, ΠA) and (BA, ΠBA) is explained by the following Proposition 6.3. The induced Kreın spaces (KA, ΠA) and (BA, ΠBA) are unitary equiv- alent, more precisely, the mapping extends uniquely to a unitary operator V ∈ L(KA, BA) such that V ΠA = ΠBA. KA ∋ x 7→ Ax ∈ BA, 7. Uniqueness of Induced Kreın Spaces The two examples of induced Kreın spaces described in Example 6.1 and Example 6.2 turned out to be unitary equivalent. However, in general, not all possible Kreın spaces induced by a fixed selfadjoint operator are unitary equivalent. We denote by ρ(T ) the resolvent set of the operator T . 14 A. GHEONDEA Theorem 7.1. Let A be a bounded selfadjoint operator in the Hilbert space H. The following statements are equivalent: (i) The Kreın space induced by A is unique, modulo unitary equivalence. (ii) There exists ǫ > 0 such that either (0, ǫ) ⊂ ρ(A) or (−ǫ, 0) ⊂ ρ(A). (iii) For some (equivalently, for any) Kreın space {K, Π} induced by A, the range of Π contains a maximal uniformly definite subspace of K. The equivalence of (ii) and (iii) is a simple matter of spectral theory for bounded selfadjoint operators. The implication (iii)⇒(i) comes from the fact that any densely defined isometric operator whose domain contains a maximal uniformly definite subspace is bounded. So only the idea of the implication (i)⇒(ii) is clarified. Assuming that the statement (ii) does not hold, there exists a decreasing sequence of numbers (µn)n≥1 with µn ∈ σ(A) and 0 < µn < 1 for all n ≥ 1, such that µn → 0 (n → ∞), and there exists a decreasing sequence of numbers (νn)n≥1 with −νn ∈ σ(−A) and 0 < νn < 1 for all n ≥ 1, such that νn → 0 (n → ∞). Then, letting µ0 = ν0 = 1, there exist sequences of orthonormal vectors {en}n≥1 and {fn}n≥1 such that (7.1) en ∈ E((µn, µn−1])H, fn ∈ E([−νn−1, −νn))H, n ≥ 1. As a consequence, one also gets (7.2) [Aei, fj] = 0, i, j ≥ 1. Define the sequence (λn)n≥1 by λn = max{p1 − µ2 n,p1 − ν2 n}. Then 0 < λn ≤ 1, λn ↑ 1 (n → ∞). Consider the subspace Sn of the Kreın space KA, defined by Sn = Cen +Cfn, n ≥ 1, and then define the operators Un ∈ L(Sn), Un = 1 p1 − λ2 n (cid:20) 1 −λn λn −1 (cid:21) , n ≥ 1. The operators Un are isometric in Sn. Further, one defines the linear manifold D0 in KA by and note that the closure of D0 = Wk≥1 By construction, the linear manifold D0 = [k≥1 Sk {ek, fk} is a regular subspace in KA. D = D+ +D− = Ran(Π) is dense in KA, where A = A+ −A− is the Jordan decomposition of A and D± = Ran(A±). Also, D0 ⊆ D and the following decomposition holds D = D0 +(D ∩ D⊥ 0 ). Then define a linear operator U in KA, with domain D0 and the same range, by USn = Un, n ≥ 1, and U(D ∩ D⊥ 0 ). The operator U is isometric, it has dense range 0 ) = I(D ∩ D⊥ REPRODUCING KERNEL KREIN SPACES 15 as well as dense domain, and it is unbounded since it maps uniformly definite subspaces into subspaces that are not uniformly definite. (KA, Π) is a Kreın space induced by A. Indeed, ΠH = UΠAH ⊇ D, and the latter is dense in KA. Further, [Πx, Πy] = [UΠAx, UΠAy] = [ΠAx, ΠAy] = [Ax, y], x, y ∈ H. Since ΠA is bounded it follows that Π is closed. By the Closed Graph Principle, it follows that Π is bounded. Since U is unbounded it follows that (KA, ΠA) is not unitary equivalent with (KA, Π). Thus, a contradiction with the assertion (i) is obtained. Let Kk be two Kreın spaces continuously embedded into the Kreın space H and denote by ιk : Kk → H the corresponding embedding operators, that is ιkh = h, h ∈ Kk, k = 1, 2. We say that the Kreın spaces K1 and K2 correspond to the same selfadjoint operator A in H if ι1ι♯ 2 = A. If A is nonnegative, equivalently, the Kreın spaces Kk are actually Hilbert spaces, this implies that K1 = K2. As a consequence of Theorem 7.1 the following corollary is obtained. 1 = ι2ι♯ Corollary 7.2. Given a selfadjoint operator A ∈ L(H), the following statements are mutually equivalent: (a) There is a unique Kreın space K continuously embedded in H and associated to A. (b) There exists ǫ > 0 such that either (−ǫ, 0) ⊂ ρ(A) or (0, ǫ) ⊂ ρ(A). (c) There exists a Kreın space K continuously embedded in H, ι : K → H such that ιι♯ = A and Ran(ι♯) contains a maximal uniformly definite subspace of K. Indeed, let Ki be two Kreın spaces continuously embedded in H and let ιi : K → H be the embedding operators, i = 1, 2. Assume that the induced Kreın spaces (Ki, ι♯ i) are unitary equivalent, that is, there exists a unitary operator U ∈ L(K1, K2) such that Uι2 = ι1. Then ι1 = ι2U ♯ and taking into account that ιi are embeddings, it follows that U ♯x = ι2U ♯x = ι1x = x, x ∈ K2, hence K1 = K2 and ι1 = ι2. This shows that the two Kreın spaces coincide. We can now apply Theorem 7.1 and get the equivalence of the statements (a), (b), and (c). 8. Existence of Reproducing Kernel Kreın Spaces The next theorem clarifies the problem of existence of reproducing kernel Kreın spaces associated to Hermitian H-kernels. Notation is as in Section 3. In addition, two positive semidefinite H-kernels are called independent if for any P ∈ K+(H) such that P ≤ H, K, it follows P = 0. Theorem 8.1. Let H ∈ Kh(H). The following assertions are equivalent: (1) There exists L ∈ K+(H) such that −L ≤ H ≤ L. (1)′ There exists L ∈ K+(H) such that [f, g]H ≤ [f, f ]1/2 L [g, g]1/2 L , f, g ∈ F0(H). (2) H = H1 − H2 with H1, H2 ∈ K+(H). (2)′ H = H+ − H− with H± ∈ K+(H) independent. (3) There exists a Kolmogorov decomposition (V ; K) of H. (4) There exists a Kreın space with reproducing kernel H. 16 A. GHEONDEA Note that letting f = g in (1)′ one obtains (1). Conversely, let L ∈ K+(H) be such that −L ≤ H ≤ L, that is [f, f ]H ≤ [f, f ]L, for all f ∈ F0(H). Let f, g ∈ F0(H). Since H is Hermitian, one gets 4 Re[f, g]H = [f + g, f + g]H − [f − g, f − g]H and hence 4 Re[f, g]H ≤ [f + g, f + g]L + [f − g, f − g]L = 2[f, f ]L + 2[g, g]L. Let λ ∈ C be chosen such that λ = 1 and Re[f, λg] = [f, λg]. Then (8.1) [f, λg]H ≤ 1 2 [f, f ]L + 1 2 [g, g]L. We distinguish two possible cases. First, assume that either [f, f ]L = 0 or [g, g]L = 0. To make a choice assume [f, f ]L = 0. Consider the inequality (8.1) with g replaced by tg for t > 0. Then [f, g]H ≤ t 2 [g, g]H. Letting t → 0 one gets [f, g]H = 0. Second case, assuming that both [f, f ]L and [g, g]L are nontrivial, in (8.1) replace f by L . Thus, the assertions (1) g to get [f, g]H ≤ [f, f ]1/2 L [g, g]1/2 L [f, f ]−1/2 and (1)′ are equivalent. f and g by [g, g]−1/2 L In order to describe the idea of the proof of the implication (1)′ ⇒(2)′, let KL be the quotient-completion of (F0(H), [·, ·]L) to a Hilbert space. More precisely, letting NL = {f ∈ F0 [f, f ]L = 0} denote the isotropic subspace of the positive semidefinite inner product space (F0(H), [·, ·]L), one considers the quotient F0(H)/NL and complete it to a Hilbert space KL. The inequality (1)′ implies that the isotropic subspace NL is contained into the isotropic subspace NH of the inner product (F0, [·, ·]H). Therefore, [·, ·]H uniquely induces an inner product on KL, also denoted by [·, ·]L such that the inequality in (1)′ still holds for all f, g ∈ KL. By the Riesz Representation Theorem one gets a selfadjoint and contractive operator A ∈ L(KL), such that (8.2) [f, g]H = [Af, g]L, f, g ∈ KL. Let A = A+ −A− be the Jordan decomposition of A in KL. Then A± are also contractions and hence (8.3) [A±f, f ]L ≤ [f, f ]L, f ∈ KL. It can be proven that the nonnegative inner products [A±·, ·] uniquely induce kernels H± ∈ K+(H) such that [f, f ]H± ≤ [f, f ]L, f ∈ F0(H), and H = H+ − H−. Indeed, the inner product [A+·, ·] restricted to F0(H)/NH can be extended to an inner product [·, ·]+ on F0(H) by letting it be null onto NL and hence (8.4) [f, f ]+ ≤ [f, f ]L, f ∈ F0(H). Let x, y ∈ X be arbitrary and x 6= y. Clearly, one can identify the Kreın space Hx[+]Hy with the subspace of all H-fields f ∈ F0(H) such that supp f ⊆ {i, j}. With this identification, one considers the restrictions of the inner products [·, ·]+ and [·, ·]L to Hx[+]Hy. The inner product [·, ·]L is jointly continuous with respect to the strong topology of Hx[+]Hy. By (8.4) and the equivalence of (1) and (1)′ one concludes that the inner product [·, ·]+ is also jointly continuous with respect to the strong topology of Hx[+]Hy and hence, by the Riesz Representation Theorem, there exists a selfadjoint operator S ∈ L(Hx[+]Hy) such that [f, g]+ = [Sf, g]Hx[+]Hy, f, g ∈ Hx[+]Hy. Define H+(x, y) = PHxSHy, H+(x, x) = PHxSHx, H+(j, j) = PHySHy and H+(j, i) = PHy SHx = H+(x, y)♯. REPRODUCING KERNEL KREIN SPACES 17 In this way one obtains a kernel H+ ∈ Kh(H) such that H+ ≤ L and [f, g]H+ = [f, g]+, f, g ∈ F0(H). Since the inner product [·, ·]+ is nonnegative it follows that H+ ∈ K+(H). Similarly one constructs the kernel H− ∈ K+(H) such that H− ≤ L and [f, g]H− = [f, g]−, f, g ∈ F0(H), where the inner product [f, g]− is the extension of the restriction of the inner product [A−f, g] to F0(H)/NL, by letting it be null onto NH. From A = A+−A−, (8.2) and the constructions of the kernels H+ and H− one concludes that H = H+ − H−. Let P ∈ K+(H) be such that P ≤ H±. Then (8.5) [f, f ]P ≤ [f, f ]L, f ∈ F0(H). As before, [·, ·]P induces a nonnegative inner product [·, ·]P on KL such that (8.3) holds for all f ∈ KL. From P ≤ H± one concludes that [f, f ]P ≤ [A±f, f ]L, f ∈ KL, and, since A+A− = 0 this implies [f, f ]P = 0 for all f ∈ KL. Since by (8.5) one gets NL ⊆ NP this implies that the inner product [·, ·]P is null onto the whole F0(H) and hence P = 0. The implications (2)′ ⇒(2) and (2)⇒(1) are clear. The most interesting implication is (1)′ ⇒(3). In a fashion similar to (1)′ ⇒(2)′, one considers the quotient-completion Hilbert space KL, the representation (8.2) and the Jordan decomposition A = A+ − A−. The latter yields in a canonical way a Kreın space (K, [·, ·]H). We again consider NL and NH, the isotropic spaces of the inner product spaces (F0(H), [·, ·]L) and, respectively, (F0(H), [·, ·]H). From the inequality (1)′ one gets NL ⊆ NH. For every x ∈ X and every vector h ∈ Hx one considers the function h ∈ F0(H) defined by (8.6) h(y) =(cid:26) h, 0, y = x, y 6= x. This identification of vectors with functions in F0(H) yields a natural embedding Hx ֒→ F0(H). With this embedding one defines linear operators V (x) : Hx → K by V (x)h = h + NH ∈ F0(H)/NH ⊆ K, h ∈ Hx. It follows that the linear operators V (x) are bounded, for all x ∈ X, and that (K; V ) is a minimal linearisation of the H-kernel H. The idea of the proof of (3)⇒(1) deserves an explanation as well. Let (K, [·, ·]) be a Kreın space and {V (x)}x∈X be a family of bounded linear operators V (x) ∈ L(Hx, K), x ∈ X, such that H(x, y) = V (y)♯V (x), x, y ∈ X. Fix on K a fundamental symmetry J and for each x ∈ X fix a fundamental symmetry Jx on Hx. Then, defining the kernel L by L(x, y) = JyV (y)∗V (x), x, y ∈ X, 18 A. GHEONDEA it can be proven that L ∈ K+(H) and that −L ≤ H ≤ L. Example 8.2. (1) Let E be a reflexive real Banach space which is not a Hilbert space. For example, one can take 1 < p < 2 and E = ℓp R. Let E ′ denote its topological dual space. On X = E × E ′ consider the Hermitian form (8.7) H((e, ϕ), (f, ψ)) = ϕ(f ) + ψ(e), e, f ∈ E, ϕ, ψ ∈ E ′. The Hermitian form H can be viewed as a Hermitian scalar kernel on X × X and it can be proven that it cannot be written as a difference of two positive semidefinite scalar kernels. Briefly, the idea is that X is a selfdual Banach space when given the norm k(e, ψ)k2 E ′, and H is jointly continuous with respect to this norm, but the topological inner product space (X; H; k · kX) is not decomposable. X = kek2 E + kψk2 9. Uniqueness of Reproducing Kernel Kreın Spaces Quasi semidefinite Hermitian kernels are associated with reproducing kernel Kreın spaces which are unique, equivalently, they have linearisations having the uniqueness property modulo unitary equivalence. In the general Hermitian case, the existence of a reproducing kernel Kreın space does not imply that it is unique, equivalently, the exis- tence of a linearisation does not imply its uniqueness, modulo unitary equivalence. Recall that, two linearisations (K; V ) and (H, U) of the same H-kernel H are unitary equiva- lent if there exists a unitary operator Φ ∈ L(K, H) such that for all x ∈ X one gets U(x) = ΦV (x). Let H be an H-kernel. If L ∈ K+(H) is such that −L ≤ H ≤ L then one denotes by KL the quotient completion of (F0(H), [·, ·]L) to a Hilbert space and by A = AL ∈ L(KL) the Gram operator of the inner product [·, ·]H with respect to the positive semidefinite inner product [·, ·]L, that is, [h, k]H = [ALh, k]L for all h, k ∈ KL. Theorem 9.1. Let H be an H-kernel on a set X which has a minimal linearisation, equivalently, it is associated to a reproducing kernel of a reproducing kernel Kreın space on X. The following assertions are equivalent: (i) The H-kernel H has unique minimal linearisation, modulo unitary equivalence. (ii) For any (equivalently, there exists a) positive semidefinite H-kernel L such that −L ≤ H ≤ L there exists ǫ > 0 such that either (0, ǫ) ⊂ ρ(AL) or (−ǫ, 0) ⊂ ρ(AL). (iii) H has a minimal linearisation (equivalently, any minimal linearisation) (K; V ) that has fundamental decomposition K = K+[+]K− such that either K+ or K− is contained in the linear manifold generated by V (x)Hx, x ∈ X. (iv) The reproducing kernel Kreın space with reproducing kernel H is unique. In order to explain the implication (i)⇒(ii), assume that there exists a positive semi- definite H-kernel L such that −L ≤ H ≤ L and for any ǫ > 0 one gets (0, ǫ) ∩ σ(AL) 6= ∅ and (−ǫ, 0) ∩ σ(AL) 6= ∅. From Theorem 7.1 it follows that there exists two Kreın spaces (K, Π) and (H, Φ) induced by the same selfadjoint operator AL, which are not unitary equivalent. It is easy to see that the operator Ψ : Ran(Π) → Ran(Φ) defined by ΨΠf = Φf, f ∈ KL, REPRODUCING KERNEL KREIN SPACES 19 is isometric, densely defined with dense range, and it is unbounded due to the non- unitary equivalence of the two induced Kreın spaces. As a consequence, Ψ is closable and its closure, denoted also by Ψ, shares the same properties. Let (V ; K) be the minimal linearisation of H defined as in the proof of Theorem 8.1, (1)′ ⇒(3). Define a new minimal linearisation (H; U) and prove that it is not uni- tary equivalent with (V ; H). More precisely, let U(x) = ΨV (x) for all x ∈ X. Since Ran(V (x)) ⊆ D(Ψ) and Ψ is closed it follows, via the Closed Graph Principle, that U(x) ∈ L(Hx, K) for all x ∈ X. Let x, y ∈ X be arbitrary and fix vectors h ∈ Hx and k ∈ Hy. Then [U(y)k, U(x)h] = [ΨV (y)k, ΨV (x)h] = [V (y)k, V (x)h] = [H(x, y)k, h], and hence U ♯ xUy = H(x, y). Also, _j∈X U(y)Hy = _j∈X ΨV (y)Hy = Clos Ran(Φ) = H. Thus, (H; U) is a minimal linearisation of the H-kernel H. On the other hand, since the operator Ψ is unbounded it follows that the two minimal linearisations (K; V ) and (H; U) are not unitary equivalent. For the implication (ii)⇒(i), let (K; V ) and (H; U) be two minimal linearisations of H. Let J and Jx be fundamental symmetries on K and, respectively, Hx, x ∈ X. We consider the positive definite H-kernel LV defined by LV (x, y) = JyV (y)∗V (x), x, y ∈ X, and as in the proof of Theorem 8.1 it follows that −LV ≤ H ≤ LV . We define a linear operator ΠV : F0(H) → K by (9.1) V (x)hx, h = (hx)x∈X ∈ F0(H). ΠV (h) =Xx∈X Taking into account of the axiom (kd2) in the definition of a minimal linearisation one obtains (9.2) [ΠV h, ΠV k]H = [h, k]K, h, k ∈ F0(H), that is, the operator ΠV is isometric from (F0(H), [·, ·]H) into (K, [·, ·]K). In addition, ΠV is also isometric when considered as a linear operator from (F0, [·, ·]LV ). Similarly, considering the positive semidefinite H-kernel LU defined by LU (x, y) = JyU(y)∗U(x), x, y ∈ X, one gets −LU ≤ H ≤ LU and, defining the linear operator ΠU : F0(H) → H by (9.3) one obtains (9.4) ΠU (h) =Xx∈X U(x)hx, h = (hx)x∈X ∈ F0(H), [ΠU h, ΠU k]H = [h, k]K, h, k ∈ F0(H), that is, the operator ΠU is isometric from (F0, [·, ·]H) into (K, [·, ·]H) and ΠU is also isometric when considered as a linear operator from (F0, [·, ·]LU ) into (K, h·, ·iJ). 20 A. GHEONDEA Let L = LV + LU and clearly −L ≤ H ≤ L. Since LV ≤ L it follows that KL is contractively embedded into KLV and hence ΠV induces a bounded operator ΠV : AL → K. From (9.2) it follows (9.5) Π∗ V JΠV = AL. By assumption, there exists ǫ > 0 such that either (−ǫ, 0) ⊂ ρ(AL) or (0, ǫ) ⊂ ρ(AL) and hence from (9.5) and taking into account that by the minimality axiom (b) of the minimal linearisation the operator ΠV has dense range, it follows that there exists a uniquely determined unitary operator ΦV : KAL → K such that (9.6) ΦV h = ΠV h, h ∈ AL, where KAL is the Kreın space induced by the operator AL. Similarly, performing the same operations with respect to the linearisation (H; V ) one gets a uniquely determined unitary operator ΦU : KAL → H such that (9.7) ΦU h = ΠU h, h ∈ AL. Define the unitary operator Φ : K → H by Φ = ΦU Φ−1 V . Taking into account of (9.6) and (9.7), the definition of the operator ΠV as in (9.1), and the definition of ΠU as in (9.3), it follows that Φ(Xx∈X V (x)hx) =Xx∈X U(x)hx, (hx)x∈X ∈ F0(H). This implies readily that for all x ∈ X one has ΦV (x) = U(x) and hence the two Kol- mogorov decompositions (K; V ) and (H; U) are unitary equivalent. As a consequence of Theorem 9.1 one can obtain a rather general sufficient condition of nonuniqueness. Let K and H be two positive semidefinite H-kernel. Then one consid- ers the Hilbert spaces KK and KH, obtained by quotient completion of (F0, [·, ·]K) and, respectively, of (F0, [·, ·]H). If H ≥ K then KH is contractively embedded into KK. The kernel H is K-compact if the embedding of KH into KK is a compact operator. Corollary 9.2. Let H+, H− ∈ K+(H) be two independent kernels, both of them of infinite rank. If there exists a kernel K ∈ K+(H) such that H+ and H− are K-compact, then the minimal linearisations of the kernel H+ − H− are not unique, modulo unitary equivalence. Let H = H1 − H2. Clearly −K ≤ H ≤ K. Let A± ∈ L(KK) denote the Gram operator of the kernel H±. Since H+ and H− are independent it follows that A = A+ − A− is the Gram operator of H. Since H± are of infinite rank and K-compact it follows that A± are compact operators of infinite rank in L(KK) and hence the spectra σ(A±) are accumulating to 0. Then the spectrum σ(A) is accumulating to 0 from both sides. This clearly contradicts the condition (ii) in Theorem 9.1 and hence the kernel H has non- unique minimal linearisations. (cid:3) REPRODUCING KERNEL KREIN SPACES 21 10. Holomorphic Kernels: Single Variable Domains Let Ω be a domain, a nonempty open subset, in the complex field C, and let H be a Hilbert space. A kernel K : Ω × Ω → L(H) is called holomorphic if it is holomorphic in the first variable, that is, for each w ∈ Ω, the map Ω ∋ z 7→ K(z, w) ∈ L(H) and conjugate holomorphic in the second variable, that is, for each z ∈ Ω, the map Ω ∋ w 7→ K(z, w) ∈ L(H) is holomorphic. Recall that, for Banach space valued functions of complex variable, strong holomorphy is the same with weak holomorphy. If K is Hermitian, then K is a holomorphic kernel if and only if the map Ω ∋ z 7→ K(z, w) ∈ L(H) is holomorphic for all w ∈ Ω. Theorem 10.1. Let H be a Hilbert space and, for some r > 0, let K be a holomorphic H-kernel on Dr = {z ∈ C z < r}. Then, there exists 0 < r′ ≤ r and a reproducing kernel Kreın space on Dr′ with reproducing kernel KDr′ × Dr′. The first step in the proof of this theorem is to observe that, without loss of generality, one can assume r > 1. Indeed, if r ≤ 1 then, for some 0 < ρ < r small enough, the H- kernel Kρ(z, w) = K(ρz, ρw) is holomorphic on Dr/ρ. If Kρ is the reproducing kernel Kreın space with reproducing kernel Kρ restricted to Dr′′, for some 0 < r′′ ≤ r/ρ, let r′ = r′′ρ and let K denote the vector space of functions f : Dr′ → H such that f (z) = F (z/ρ) for some F ∈ Kρ and all z ∈ Dr′. On K there is defined the inner product [·, ·]K f (z) = F (z/ρ), g(z) = G(z/ρ), F, G ∈ Kρ. [f, g]K = [F, G]Kρ, Then (K; [·, ·]K) is a Kreın space. For each w ∈ Dr′ the map z 7→ K(z, ρw) belongs to K and, for each F ∈ Kρ, f (z) = F (z/ρ), and h ∈ H, [f, K(·, ρw)h]K = [F, K(ρ·, ρw)h]Kρ = [F (ρw), h]H = [f (w), h]H, hence K is a reproducing kernel Kreın space with reproducing kernel KDr′ × Dr′. There are two main ideas of the proof. First, the Szego kernel S, see Example 4.3, plays a distinguished role in holomorphy, and allows us to construct the convolution kernel of K in the Hardy space H 2(D). The second idea is that, once the convolution operator represented as a selfadjoint bounded operator on a Hilbert space of functions is defined, the construction of the induced Kreın space as in Example 6.2 will provide the reproducing kernel Kreın space with reproducing kernel K. Here are a few details. Letting r > 1, there is considered the Hardy space H 2(D) ⊗ H, identified with a space n=0 anzn, where (an)n≥0 is a sequence of vectors in H such H < ∞. Also, the inner product on the Hilbert space H 2(D) ⊗ H Thus, on H 2(D) ⊗ H one can define the analog of the convolution operator K hf (eit), g(eit)iHdt, f, g ∈ H 2(D) ⊗ H. is n≥0 kank2 of H-valued functions f (z) =P∞ that kf k2 =P∞ 2π Z 2π 2π Z 2π (Kf )(z) = hf, gi = (10.1) 0 1 1 0 K(z, eit)f (eit)dt, f ∈ H 2(D) ⊗ H. Letting M = supz,w≤1 kK(z, w)k < ∞, it follows that kKf k ≤ Mkf kH 2(D)⊗H, hence the convolution operator K is a bounded linear operator in H 2(D) ⊗ H. On the other hand, hKf, giH 2(D)⊗H = hK(eit, eis)f (eis), g(eit)iHdtds, 1 4π2 Z 2π 0 Z 2π 0 22 A. GHEONDEA hence K is selfadjoint. Since, for any w ∈ D1 and any h ∈ H, the function fw(z) = h/(1 − wz) belongs to H 2(D) ⊗ H and by the Cauchy formula, (Kfw)(z) = K(z, w)h, the range of the convolution operator K contains all the functions Kw(·)h, for w ∈ D and h ∈ H. Then, one can use the construction of the induced Kreın space (BK; ΠBK ) inside of H 2(D) ⊗ H, as in Example 6.2 but applied to the bounded selfadjoint operator K in the Hilbert space H 2(D) ⊗ H, in order to get the reproducing kernel Kreın space K with reproducing kernel K. 11. Holomorphic Kernels: Several Variables Domains In this section it is considered the analog of Theorem 10.1 in case the Hermitian kernels are defined on domains in CN for N ≥ 2. We use the notation as in Example 4.4 where the Drury-Arveson space was constructed as the reproducing kernel Hilbert space associated to the Szego kernel. For simplicity, it is considered only scalar valued kernels. Recall that a scalar-valued Hermitian kernel K, defined on a nonempty open subset O of G = CN , is holomorphic on O if K(·, η) is holomorphic on O for each fixed η ∈ O. Since K is Hermitian, it follows that K is conjugate holomorphic in the second variable. Theorem 11.1. Let r > 0 and let K be a Hermitian holomorphic kernel on the open ball Br in CN . Then, there exist 0 < r′ ≤ r and a reproducing kernel Kreın space K on Br′ with reproducing kernel KBr′ × Br′. To a certain extent, the proof of this theorem follows a pattern similar to that of the proof of Theorem 10.1, namely, first a scaling argument can be used in order to reduce the proof to the case r > 1, then the convolution kernel of K can defined on the Drury-Arveson space H 2(B1) associated to the Szego kernel S, see Example 4.4, and it can be proven that this convolution operator is bounded and selfadjoint. Finally, the construction of type (BA; ΠBA), see Example 6.2, can be used in order to produce a reproducing kernel space K with reproducing kernel K. This proof shows, once again, a certain universality property of the Szego kernel with respect to holomorphic Hermitian kernels. Let K be a scalar Hermitian holomorphic kernel on Br, r > 0. Since K is Hermitian, it follows that ζ 7→ K(ξ, ζ) is holomorphic on Br for each ξ ∈ Br, that is, letting {ej}N denote the canonical orthonormal basis of G = CN , the conjugation can be defined by j=1 ξ = NXj=1 hξ, ejiej → NXj=1 hξ, ejiej = ξ, so that the function f (ξ, η) = K(ξ, η) is separately holomorphic on Br × Br. By Hartogs' Theorem f is holomorphic on Br×Br, hence f is locally bounded. Similar to the argument provided for Theorem 10.1, without loss of generality one can suppose that r > 1. Hence there exist 1 < ρ < r and C > 0 such that: (11.1) and (11.2) K(ξ, η) ≤ C for all ξ, η ∈ Bρ K(ξ, η) = Xm≥0 pm(ξ, η) REPRODUCING KERNEL KREIN SPACES 23 uniformly on Bρ, where each pm, m ≥ 0, is an m-homogeneous complex polynomial on 2N variables. There exists a continuous linear functional Am on Pm(G × G)⊗m, see (4.7), such that (11.3) pm(ξ, η) = Am((ξ, η)⊗m), for all ξ, η ∈ CN . Using Cauchy Inequalities, for Bρ, one gets (11.4) hence (11.5) kAmk ≤ C 1 ρm , Xm≥0 kAmk2 ≤ CXm≥0 1 ρ2m = C 1 1 − 1/ρ2 = C ′ < ∞. By the Riesz Representation Theorem, there exist am ∈ Pm(G × G)⊗m, m ≥ 0, such that (11.6) and (11.7) Am((ξ, η)⊗m) = h(ξ, η)⊗m, ami(G×G)⊗m, kamk = kAmk, (with a0 = A0 ∈ C). Since Pm(G × G)⊗m is isometrically isomorphic to (PmG⊗m)⊕(m+1), it is deduced that there are ak m ∈ PmG⊗m, k = 0, . . . , m, such that (11.8) and (11.9) h(ξ, η)⊗m, ami(G×G)⊗m = miG⊗m, mXk=0 hbk m(ξ, η), ak mXk=0 kak mk2 = kamk2, where b0 (11.6), and (11.8), 0 = 1 and bk m(ξ, η) = ξ⊗(m−k) ⊗ η⊗k, m ≥ 1, k = 0, . . . , m. By (11.2), (11.3), K(ξ, η) = Xm≥0 mXk=0 hbk m(ξ, η), ak mi =Xk≥0Xm≥k hbk m(ξ, η), ak mi, where the series converge absolutely on η by (11.4). Using all these, it can be shown that Kη ∈ H 2(B1) for all η ∈ B1, where Kη(ξ) = K(ξ, η). Then letting Kaη = Kη, η ∈ B1, one gets a bounded linear operator in H 2(B1) such that K(ξ, η) = hKaξ, aηiH 2(B1). This operator K is selfadjoint and it is the analog of the convolution operator for which, applying the construction of type (BA; ΠBA) as in Example 6.2, one gets a reproducing kernel Kreın space K with reproducing kernel KB1 × B1. 24 A. GHEONDEA 12. Comments The theory of reproducing kernel Hilbert spaces and their positive semidefinite kernels originates with the works of S. Zaremba [44], G. Szego [42], S. Bergman [7], and S. Bochner [8]. E.H. Moore [33] also contributed significantly to this theory, but the first systemati- sation and abstract presentation belongs to N. Aronszajn [4]. A different but equivalent theory belongs to L. Schwartz [40], whose work remained almost unnoticed for a long time, although it was the first to consider reproducing kernel Kreın spaces (Hermitian spaces, as called there). So far, monographs on this subject have been written by T. Ando [3], S. Saitoh [38], and the forthcoming title of S. Saitoh and Y. Sawano [39], which are good sources for the large area of applications of the technique of reproducing kernel spaces in complex functions theory, ordinary and partial differential equations, integral equations, and approximation and numerical analysis. So far, there are two monographs devoted to indefinite inner product spaces and their linear operators, J. Bognar [9] and T.Ya. Azizov and I.S. Iokhvidov [6], where proofs, examples, and counter-examples of the facts recalled in Sections:ks can be found. The introductory material on Hermitian kernels in Section 3 follows T. Constantinescu and A. Gheondea [14]. Theorem 3.1 is classical. The concept of linearisation originates with J. Mercer [32], for the scalar case, and A.N. Kolmogorov [25], for the operator valued case. The Hardy Space H 2(D) originates with the G. Szego Kernel [42]. For the theory of Hardy spaces there are monographs of P.L. Duren [19] and P. Koosis [26]. Bergman kernel is also another important example of a positive semidefinite kernel but it falls out of this chapter concern, see P.L. Duren and B. Schuster [20]. The Drury-Arveson space originates with S.W. Drury [21] and W.B. Arveson [5]. The holomorphic kernels considered at Example 4.5 make the main object of investigation of the monograph of D. Alpay et al. [2]. The investigations of A.V. Potapov [36], L. de Branges [10]–[12], M.G. Kreın and H. Langer [27]–[31], and H. Dym [22] highly motivate the interest for Hermitian kernels with or without finite negative signatures. The study of Toplitz type kernels is related to the investigations on operator dilations of M.A. Naımark [34] and B. Sz.-Nagy [43]. Our short presentation as in Example 4.6 follows [14]. We only mention that there is a more general and powerful theory of kernels invariant under actions of ∗-semigroups presented in [15], motivated by problems in mathematical physics as in D.E. Evans and J.T. Lewis [23] and K.R. Parthasaraty, K. Schmidt [35], and many others. Theorem 5.1 essentially belongs to P. Sorjonen [41] but this result follows from the more general theory of L. Schwartz [40] that has been obtained about ten years before. Theorem 5.2 can be found in D. Alpay et al. [2]. Examples 6.1, 6.2, and Theorem 7.1 on induced Kreın spaces can be found in [14], while Corollary 7.2 belongs to T. Hara [24]. Similar and, in a certain way, equivalent uniqueness conditions, can be found in [13] and M.A. Dritschel [18]. B. ´Curgus and H. Langer [17] proves that once non-equivalent induced Kreın spaces exist, there are infinitely many. The characterisations of existence of reproducing Kreın spaces associated to given Her- mitian kernels as in Theorem 8.1 belong essentially to L. Schwartz [40], but our presen- tation follows [14]. Example 8.2 is from [40] as well, cf. [2]. The uniqueness Theorem 9.1 is from [14] while Corollary 9.2 is from [40]. REPRODUCING KERNEL KREIN SPACES 25 The result in Theorem 10.1 on single variable holomorphic kernels belongs to D. Alpay [1], while its several variables generalisation in Theorem 11.1 is from [16]: for the basics of several complex variables holomorphic functions facts used during the explanation of the ideas of the proof, see R.M. Range [37]. References [1] D. Alpay, Some remarks on reproducing kernel Kreın spaces, Rocky Mountain J., 21(1991), 1189– 1205. [2] D. Alpay, A. Dijksma, J. Rovnyak, H. de Snoo, Schur Functions, Operator Colligations, and Reproducing Kernel Pontryagin Spaces, Birkhauser Verlag, Basel 1997. [3] T. Ando, Reproducing Kernel Spaces and Quadratic Inequalities, Lecture Notes, Hokkaido Univer- sity, Research Institute of Applied Electricity, Division of Applied Mathematics, Sapporo, Japan, 1987. [4] N. Aronszajn, Theory of reproducing kernels, Trans. Amer. Math. Soc. 68(2950), 337-404. [5] W.B. Arveson, Subalgebras of C ∗-algebras. III: Multivariable operator theory, Acta Math. 181(1998), 159–228. [6] T.Ya. Azizov, I.S. Iokhvidov, Linear Operators in Spaces with an Indefinite Metric, Pure and Applied Mathematics, John Wiley & Sons, Ltd., Chichester 1989. [7] S. Bergman, Uber die Entwicklung der harmonischen Funktionen der Ebene und des Raumes nach Orthogonalfunktionen, Math. Ann. 86(1922), 238–271. [8] S. Bochner, Uber orthogonale Systeme analytischer Funktionen, Math. Zeitschr. 14(1922), 180– 207. [9] J. Bognar: Indefinite Inner Product Spaces, Springer-Verlag, Berlin 1976. [10] L. de Branges: Complementation theory in Kreın spaces, Trans. Amer. Math. Soc., 305(1988), 277–291. [11] L. de Branges, Krein spaces of analytic functions, J. Functional Analysis, 81(1988), 219–259. [12] L. de Branges, A construction of Krein spaces of analytic functions, J. Functional Analysis, 98(1991), 1–41. [13] T. Constantinescu, A. Gheondea, Elementary rotations of linear operators in Kreın spaces, J. Operator Theory, 29(1993), 167–203. [14] T. Constantinescu, A. Gheondea, Representations of hermitian kernels by means of Kreın spaces, Publ. RIMS. Kyoto Univ., 33(1997), 917–951. [15] T. Constantinescu, A. Gheondea, Representations of Hermitian kernels by means of Kreın spaces II. Invariant kernels, Commun. Math. Phys. 216(2001), 409–430. [16] T. Constantinescu, A. Gheondea, On L. Schwartz's boundedness condition for kernels, Posi- tivity 10(2006), 65–86. [17] B. ´Curgus, H. Langer, Continuous embeddings, completions and complementation in Krein spaces, Rad. Mat. 12(2003), 37–79. [18] M.A. Dritschel, The essential uniqueness property of linear operators in Kreın spaces, J. Funct. Anal. 118(1993), 198–248. [19] P.L. Duren, Theory of H [20] P.L. Duren, A. Schuster, Bergman Spaces, Amer. Math. Soc., Providence R.I. 2004. [21] S.W. Drury, A generalization of von Neumanns inequality to the complex ball, Proc. Amer. Math. p Spaces, Academic Press, INC., New York, 1970. Soc. 68(1978), 300–304. [22] H. Dym, J-Contractive Matrix Functions, Reproducing Kernel Hilbert Spaces and Interpolation, CBMS Regional Conference Series in Mathematics, Amer. Math. Soc., Providence R.I. 1989. [23] D.E. Evans, J.T. Lewis: Dilations of Irreducible Evolutions in Algebraic Quantum Theory, Comm. Dublin Inst. Adv. Studies Ser. A No. 24, Dublin Institute for Advanced Studies, Dublin, 1977. [24] T. Hara, Operator inequalities and construction of Kreın spaces, Integral Equations and Operator Theory, 15(1992), 551–567. [25] A.N. Kolmogorov: Stationary sequences in Hilbert space, Bull. Math. Univ. Moscow, 2(1941), 1–40. [26] P. Koosis, Introduction to H p Spaces, Cambridge Mathematical Press, Cambridge, 1970. 26 A. GHEONDEA [27] M.G. Krein, H. Langer, Uber einige Fortsetzungsprobleme, die eng mit der Theorie hermitescher Operatoren im Raume Πκ zusammenhangen, I. Einige Funktionenklassen und ihre Darstellungen, Math. Nachr. 77(1977), 187–236. [28] M.G. Krein, H. Langer, Uber einige Fortsetzungsprobleme, die eng mit der Theorie hermitescher Operatoren im Raume Πκ zusammenhangen, II. Verallgemeinerte Resolventen, u-Resolventen und ganze Operatoren, J. Funct. Anal. 30(1978), 390–447. [29] M.G. Krein, H. Langer, On some extension problems which are closely connected with the theory of hermitian operators in a space Πκ. III: Indefinite analogues of the Hamburger and Stieltjes moment problem, Part I, Beitr. Ana. 14(1979), 25-40; Part II, Beitr. Anal. 15(1981), 27–45. [30] M.G. Krein, H. Langer, Some propositions on analytic matrix functions related to the theory of operators in the space Πκ, Acta. Sci. Math. (Szeged) 1981, 181–205. [31] M.G. Krein, H. Langer, On some extension problems which are closely connected with the theory of hermitian operators in a space Πκ, IV. Continuous analogues of orthogonal polynomials on the unit circle with respect to an indefinite weight and related continuation problems for some classes of functions, J. Operator Theory 13(1985), 299-417. [32] J. Mercer, Functions of positive and negative type and their connections with the theory of integral equations, Philos. Trans. Roy. Soc. London Ser. A 209(1909), 415–446. [33] E.H. Moore, General Analysis, Mem. Amer. Philos. Soc. I, 1935; II, 1939. [34] M.A. Naımark: Positive-definite operator functions on a commutative group [Russian], Izvestya Akad. Nauk SSSR, 1(1943), 234–244. [35] K.R. Parthasaraty, K. Schmidt: Positive-Definite Kernels, Continous Tensor Products and Central Limit Theorems of Probability Theory, Lecture Notes in Mathematics, Vol. 272, Springer- Verlag, Berlin 1972. [36] V.P. Potapov, The multiplicative structure of J-contractive matrix functions, Trudy Moskov. Mat. Obshch. 4(1955), 125–236. [37] R.M. Range, Holomorphic Functions and Integral Representations in Several Complex Variables, Graduate Texts in Mathematics Vol. 108, Springer-Verlag, Berlin 1986. [38] S. Saitoh, Theory of Reproducing Kernels and its Applications, Pitman Research Notes in Mathe- matics Series, vol. 189, Longman Scientific and Technical, Harlow 1988 [39] S. Saitoh, Y. Sawano, The Theory of Reproducing Kernels – 60 Years Since N. Aronszajn, (to appear) [40] L. Schwartz, Sous espace Hilbertiens d'espaces vectoriel topologiques et noyaux associ´es (noyaux reproduisants), J. Analyse Math., 13(1964), 115–256. [41] P. Sorjonen, Pontrjagin Raume mit einem reproduzierenden Kern, Ann. Acad. Sci. Fenn. Ser. A I Math. Dissertationes, 594(1973), 1–30. [42] G. Szego, Uber die Randwerte einer analytischen Funktion, Math. Ann. 84(1921), 232–244. [43] B. Sz.-Nagy, Prolongement des transformations de l'espace de Hilbert qui sortent de cet espace, in Appendice au livre "Le¸cons d'analyse fonctionnelle" par F. Riesz et B. Sz.-Nagy, pp. 439-573 Akademiai Kiado, Budapest, 1955. [44] S. Zaremba, L'´equation biharmonique et une classe remarquable de fonctions fondamentales har- monique, Krak. Anz. (1907), 147–196. Department of Mathematics, Bilkent University, 06800 Bilkent, Ankara, Turkey, and Institutul de Matematica al Academiei Romane, C.P. 1-764, 014700 Bucures¸ti, Romania E-mail address: [email protected] and [email protected]
1005.1585
2
1005
2010-12-29T12:56:59
A really simple elementary proof of the uniform boundedness theorem
[ "math.FA" ]
I give a proof of the uniform boundedness theorem that is elementary (i.e. does not use any version of the Baire category theorem) and also extremely simple.
math.FA
math
A really simple elementary proof of the uniform boundedness theorem Alan D. Sokal∗ Department of Physics New York University 4 Washington Place New York, NY 10003 USA [email protected] May 7, 2010 revised October 20, 2010 to appear in the American Mathematical Monthly Abstract I give a proof of the uniform boundedness theorem that is elementary (i.e., does not use any version of the Baire category theorem) and also extremely simple. Key Words: Uniform boundedness; gliding hump; sliding hump; Baire category. Mathematics Sub ject Classification (MSC 2000) codes: 46B99 (Primary); 46B20, 46B28 (Secondary). 0 1 0 2 c e D 9 2 ] A F . h t a m [ 2 v 5 8 5 1 . 5 0 0 1 : v i X r a ∗Also at Department of Mathematics, University College London, London WC1E 6BT, England. One of the pillars of functional analysis is the uniform boundedness theorem: Uniform boundedness theorem. Let F be a family of bounded linear op- erators from a Banach space X to a normed linear space Y . If F is pointwise bounded (i.e., supT ∈F kT xk < ∞ for all x ∈ X ), then F is norm-bounded (i.e., supT ∈F kT k < ∞). The standard textbook proof (e.g., [17, p. 81]), which goes back to Stefan Banach, Hugo Steinhaus, and Stanis law Saks in 1927 [3], employs the Baire category theorem or some variant thereof.1 This proof is very simple, but its reliance on the Baire category theorem makes it not completely elementary. By contrast, the original proofs given by Hans Hahn [7] and Stefan Banach [2] in 1922 were quite different: they began from the assumption that supT ∈F kT k = ∞ and used a “gliding hump” (also called “sliding hump”) technique to construct a sequence (Tn ) in F and a point x ∈ X such that limn→∞ kTnxk = ∞.2 Variants of this proof were later given by T. H. Hildebrandt [11] and Felix Hausdorff [9, 10].3 These proofs are elementary, but the details are a bit fiddly. Here is a real ly simple proof along similar lines: Lemma. Let T be a bounded linear operator from a normed linear space X to a normed linear space Y . Then for any x ∈ X and r > 0, we have sup x′∈B (x,r) kT x′k ≥ kT kr , (1) where B (x, r) = {x′ ∈ X : kx′ − xk < r}. Proof. For ξ ∈ X we have max(cid:8)kT (x + ξ )k, kT (x − ξ )k(cid:9) ≥ 1 2 (cid:2)kT (x + ξ )k + kT (x − ξ )k(cid:3) ≥ kT ξ k , (2) where the second ≥ uses the triangle inequality in the form kα − β k ≤ kαk + kβ k. Now take the supremum over ξ ∈ B (0, r). (cid:3) Proof of the uniform boundedness theorem. Suppose that supT ∈F kT k = ∞, and choose (Tn )∞ n=1 in F such that kTnk ≥ 4n . Then set x0 = 0, and for n ≥ 1 use the lemma to choose inductively xn ∈ X such that kxn − xn−1k ≤ 3−n and kTnxnk ≥ 2 3 3−nkTnk. The sequence (xn ) is Cauchy, hence convergent to some x ∈ X ; and it is easy to see that kx − xnk ≤ 1 2 3−n and hence kTnxk ≥ 1 6 3−nkTnk ≥ 1 6 (4/3)n → ∞. (cid:3) 1 See [4, p. 319, note 67] concerning credit to Saks. 2 Hahn’s proof is discussed in at least two modern textbooks: see [14, Exercise 1.76, p. 49] and [13, Exercise 3.15, pp. 71–72]. 3 See also [18, pp. 63–64] and [21, pp. 74–75] for an elementary proof that is closely related to the standard “nested ball” proof of the Baire category theorem; and see [8, Problem 27, pp. 14–15 and 184] and [12] for elementary proofs in the special case of linear functionals on a Hilbert space. 2 Remarks. 1. As just seen, this proof is most conveniently expressed in terms of a sequence (xn ) that converges to x. This contrasts with the earlier “gliding hump” proofs, which used a series that sums to x. Of course, sequences and series are equivalent, so each proof can be expressed in either language; it is a question of taste which formulation one finds simpler. 2. This proof is extremely wasteful from a quantitative point of view. A quantita- tively sharp version of the uniform boundedness theorem follows from Ball’s “plank theorem” [1]: namely, if P∞ n=1 kTnk−1 < ∞, then there exists x ∈ X such that limn→∞ kTnxk = ∞ (see also [15]). 3. A similar (but slightly more complicated) elementary proof of the uniform boundedness theorem can be found in [6, p. 83]. 4. “Gliding hump” proofs continue to be useful in functional analysis: see [20] for a detailed survey. 5. The standard Baire category method yields a slightly stronger version of the uniform boundedness theorem than the one stated here, namely: if supT ∈F kT xk < ∞ for a nonmeager (i.e., second category) set of x ∈ X , then F is norm-bounded. 6. The uniform boundedness theorem has generalizations to suitable classes of non-normable and even non-metrizable topological vector spaces (see, e.g., [19, pp. 82– 87]). I leave it to others to determine whether any ideas from this proof can be carried over to these more general settings. 7. More information on the history of the uniform boundedness theorem can be found in [4, pp. 302, 319n67], [5, pp. 138–142], and [16, pp. 21–22, 40–43]. Acknowledgments. I wish to thank Keith Ball for reminding me of Hahn’s “gliding hump” proof; it was my attempt to fill in the details of Keith’s sketch that led to the proof reported here. Keith informs me that versions of this proof have been independently devised by at least four or five people, though to his knowledge none of them have bothered to publish it. I also wish to thank David Edmunds and Bob Megginson for helpful correspondence, and three anonymous referees for valuable suggestions concerning the exposition. Finally, I wish to thank Jurgen Voigt and Markus Haase for drawing my attention to the elementary proofs in [6, 18, 21]. This research was supported in part by U.S. National Science Foundation grant PHY–0424082. References [1] K. Ball, The plank problem for symmetric bodies, Invent. Math. 104 (1991) 535–543. [2] S. Banach, Sur les op´erations dans les ensembles abstraits et leur application aux ´equations int´egrales, Fund. Math. 3 (1922) 133–181; also in OEuvres avec des Commentaires , vol. 2, ´Editions scientifiques de Pologne, Warsaw, 1979. 3 [3] S. Banach and H. Steinhaus, Sur le principe de la condensation des singularit´es, Fund. Math. 9 (1927) 50–61. [4] G. Birkhoff and E. Kreyszig, The establishment of functional analysis, Historia Math. 11 (1984) 258–321. [5] J. Dieudonn´e, History of Functional Analysis , North-Holland, Amsterdam, 1981. [6] I. Gohberg, S. Goldberg, and M. A. Kaashoek, Basic Classes of Linear Opera- tors , Birkhauser, Basel, 2003. [7] H. Hahn, Uber Folgen linearer Operationen, Monatsh. Math. Phys. 32 (1922) 3–88. [8] P. R. Halmos, A Hilbert Space Problem Book , 2nd ed., Springer-Verlag, New York, 1982. [9] F. Hausdorff, Zur Theorie der linearen metrischen Raume, J. Reine Angew. Math. 167 (1932) 294–311. [10] J. Hennefeld, A nontopological proof of the uniform boundedness theorem, Amer. Math. Monthly 87 (1980) 217. [11] T. H. Hildebrandt, On uniform limitedness of sets of functional operations, Bul l. Amer. Math. Soc. 29 (1923) 309–315. [12] S. S. Holland Jr., A Hilbert space proof of the Banach-Steinhaus theorem, Amer. Math. Monthly 76 (1969) 40–41. [13] B. D. MacCluer, Elementary Functional Analysis , Springer-Verlag, New York, 2009. [14] R. E. Megginson, An Introduction to Banach Space Theory , Springer-Verlag, New York, 1998. [15] V. Muller and J. Vrsovsk´y, Orbits of linear operators tending to infinity, Rocky Mountain J. Math. 39 (2009) 219–230. [16] A. Pietsch, History of Banach Spaces and Linear Operators , Birkhauser, Boston, 2007. [17] M. Reed and B. Simon, Methods of Modern Mathematical Physics , vol. I, Func- tional Analysis , Academic Press, New York, 1972. [18] F. Riesz and B. Sz.-Nagy, Functional Analysis , Ungar, New York, 1955. [19] H. H. Schaefer and M. P. Wolff, Topological Vector Spaces , 2nd ed., Springer- Verlag, New York, 1999. 4 [20] C. Swartz, Infinite Matrices and the Gliding Hump , World Scientific, River Edge, NJ, 1996. [21] J. Weidmann, Linear Operators in Hilbert Spaces , Springer-Verlag, New York, 1980. 5
1010.0513
1
1010
2010-10-04T09:21:14
The Range of Localization Operators and Lifting Theorems for Modulation and Bargmann-Fock Spaces
[ "math.FA" ]
We study the range of time-frequency localization operators acting on modulation spaces and prove a lifting theorem. As an application we also characterize the range of Gabor multipliers, and, in the realm of complex analysis, we characterize the range of certain Toeplitz operators on weighted Bargmann-Fock spaces. The main tools are the construction of canonical isomorphisms between modulation spaces of Hilbert-type and a refined version of the spectral invariance of pseudodifferential operators. On the technical level we prove a new class of inequalities for weighted gamma functions.
math.FA
math
THE RANGE OF LOCALIZATION OPERATORS AND LIFTING THEOREMS FOR MODULATION AND BARGMANN-FOCK SPACES KARLHEINZ GR OCHENIG AND JOACHIM TOFT Abstract. We study the range of time-frequency localization operators acting on modulation spaces and prove a lifting theorem. As an application we also characterize the range of Gabor multipliers, and, in the realm of complex analysis, we characterize the range of certain Toeplitz operators on weighted Bargmann- Fock spaces. The main tools are the construction of canonical isomorphisms between modulation spaces of Hilbert-type and a refined version of the spectral invariance of pseudodifferential operators. On the technical level we prove a new class of inequalities for weighted gamma functions. 1. Introduction 0 1 0 2 t c O 4 ] . A F h t a m [ 1 v 3 1 5 0 . 0 1 0 1 : v i X r a The precise description of the range of a linear operator is usually difficult, if not impossible, because this amounts to a characterization of which operator equations are solvable. In this paper we study the range of an important class of pseudodiffer- ential operators, so-called time-frequency localization operators, and we prove an isomorphism theorem between modulation spaces with respect to different weights. The guiding example to develop an intuition for our results is the class of multi- plication operators. Let m ≥ 0 be a weight function on Rd and define the weighted m =(cid:16)RRd f (x)pm(x)p dx(cid:17)1/p space Lp = kf mkLp. Let Ma be the multiplication operator defined by Maf = af . Then Lp m is precisely the range of the multiplication operator M1/m. We will prove an similar result for time-frequency localization operators between weighted modulation spaces. To set up terminology, let π(z)g(t) = e2πiξ·tg(t − x) denote the time-frequency shift by z = (x, ξ) ∈ R2d acting on a function g on Rd. The corresponding transform is the short-time Fourier transform of a function defined by m(Rd) by the norm kfkLp Vgf (z) =ZRd f (t)g(t − x) e2πiξ·t dt = hf, π(z)gi . The standard function spaces of time-frequency analysis are the modulation spaces. The modulation space norms measure smoothness in the time-frequency space (phase space in the language of physics) by imposing a norm on the short-time Key words and phrases. Localization operator, Toeplitz operator, Bargmann-Fock space, mod- ulation space, Sjostrand class, spectral invariance, Hermite function. K. G. was supported in part by the project P2276-N13 of the Austrian Science Foundation (FWF). 1 2 KARLHEINZ GR OCHENIG AND JOACHIM TOFT Fourier transform of a function f . As a special case we mention the modulation spaces M p m(Rd) for 1 ≤ p ≤ ∞ and a non-negative weight function m. Let h(t) = 2d/4e−πt2 = 2d/4e−πt·t, t ∈ Rd denote the (normalized) Gaussian. Then the modulation space M p by the norm m(Rd) is defined The localization operator Ag function, and the symbol or multiplier m is defined formally by the integral m with respect to the "window" g, usually some test kfkM p m = kVhfkLp m . mf =ZR2d Ag m(z)Vgf (z)π(z)g dz . Localization operators constitute an important class of pseudodifferential operators and occur under different names such as Toeplitz operators or anti-Wick operators. They were introduced by Berezin as a form of quantization [2], and are nowadays applied in mathematical signal processing for time-frequency masking of signals and for phase-space localization [11]. An equivalent form occurs in complex analysis as Toeplitz operators on Bargmann-Fock space [3, 4, 7]. In hard analysis they are used to approximate pseudodifferential operators and in some proofs of the sharp Garding inequality and the Fefferman-Phong inequality [22, 23, 27]. For the analysis of localization operators with time-frequency methods we refer to [9] and the references given there, for a more analytic point of view we recommend [29,30]. A special case of our main result can be formulated as follows. By an isomorphism between two Banach spaces X and Y we understand a bounded and invertible operator from X onto Y . Theorem 1.1. Let m be a non-negative continuous symbol on R2d satisfying m(w+ z) ≤ eawm(z) for w, z ∈ R2d and assume that m is radial in each time-frequency coordinate. Then for suitable test functions g the localization operator Ag m is an isomorphism from the modulation space M p µ/m(Rd) for all 1 ≤ p ≤ ∞ and moderate weights µ. µ(Rd) onto M p We see that the range of a localization operator exhibits the same behavior as the multiplication operators. For the precise formulation with all assumptions stated we refer to Section 4. The above isomorphism theorem can also be intepreted as a lifting theorem, quite in analogy with the lifting property of Besov spaces [32]. However, whereas Besov spaces Bp,q s with different smoothness s are isomorphic via Fourier multipliers, the case of modulation spaces is more subtle because in this case the time-frequency smoothness is parametrized by a weight function on R2d rather than by a single numbers ∈ R. The lifting operators between modulation spaces are precisely the localization operators with the weight m. The isomorphism theorem for localization operators stated above is preceded by many contributions, which were already listed in [19]. In particular, in [10] it was shown that for moderate symbol functions with subexponential growth the localization operator is a Fredholm operator, i.e., it differs from an invertible op- erator only by a finite-rank operator. Here we show that such operators are even THE RANGE OF LOCALIZATION OPERATORS 3 isomorphisms between the corresponding modulation spaces, even under weaker conditions than needed for the Fredholm property. In the companion paper [19] we proved the isomorphism property for weight functions of polynomial type, i.e., c(1 + z)−N ≤ m(z) ≤ C((1 + z)N . (1) The contribution of Theorem 1.1 is the extension of the class of possible weights. In particular, the isomorphism property holds also for subexponential weight func- tions of the form m(z) = eazb or m(z) = eaz/ log(e+z) for a > 0, 0 < b < 1, which are often considered in time-frequency analysis. We remark that this generality comes at the price of imposing the radial symmetry on the weight m. Therefore, our results are not applicable to all situations covered by [19], where no radial symmetry is required. Although the extension to weights of ultra-rapid growth looks like a routine generalization, it is not. The proof of Theorem 1.1 for weights of polynomial type is based on a deep theorem of Bony and Chemin [6]. They construct a one-parameter mt(Rd) for t ∈ R. group of isomorphisms from L2(Rd) onto the modulation spaces M 2 Unfortunately, the pseudodifferential calculus developed in [6] requires polynomial growth conditions, and, to our knowledge, an extention to symbols of faster growth is not available. On a technical level, our main contribution is the construction of canonical iso- m(Rd) of Hilbert type. In morphisms between L2(Rd) and the modulation spaces M 2 fact, such an isomorphism is given by a time-frequency localization operator with Gaussian window h(t) = 2d/4e−πt2. Theorem 1.2. Assume that m is a continuous moderate weight function of at most exponential growth and radial in each time-frequency coordinate. Then the localization operator Ah m is an isomorphism from L2(Rd) onto M 2 1/m(Rd). C−1 ≤Z ∞ 0 θ(px/π) xn n! e−xdx Z ∞ 0 This theorem replaces the result of Bony and Chemin. The main point is that Theorem 1.2 also covers symbols of ultra-rapid growth. Its proof requires most of our efforts. We take a time-frequency approach rather than using classical methods from pseudodifferential calculus. In the course of its proof we will establish new inequalities for weighted gamma functions of the form 1 xn n! θ(px/π) e−xdx ≤ C for all n ∈ N ∪ {0} . The proof method for Theorem 1.1 may be of interest in itself. Once the canonical isomorphisms are in place (Theorem 1.2), the proof of Theorem 1.1 proceeds as fol- lows. It is easy to establish that Am is an isomorphism from M 2√m to M 2 1/√m, so that the composition Ag m is an isomorphism on M 2√m. Using the canonical isomor- phisms of Theorem 1.2, one shows next that the operator V = Ah√mAg 1/√m is an isomorphism on L2 and that the (Weyl) symbol of this operator belongs to a generalized Sjostrand class. After these technicalities we apply the machinery of spectral invariance of pseudodifferential operators from [16] to conclude that V is invertible on all modulation spaces M p µ (with µ compatible with the conditions on 1/mAg 1/mAg mAh 4 KARLHEINZ GR OCHENIG AND JOACHIM TOFT m and the window g). Since V is a composition of three isomorphisms, we then deduce that Ag m is an isomorphism from M p µ onto M p µ/m. As an application we prove (i) a new isomorphism theorem for so-called Ga- bor multipliers, which are a discrete version of time-frequency localization oper- ators, and (ii) an isomorphism theorem for Toeplitz operators between weighted Bargmann-Fock spaces of entire functions. To formulate this result more explicitly, for a non-negative weight function µ and 1 ≤ p ≤ ∞, let F p µ(Cd) be the space of entire functions of d complex variables defined by the norm =ZCd F (z)pµ(z)pe−pπz2/2dz < ∞ , kFkp F p µ µ/m(Cd) for every 1 ≤ p, q ≤ ∞ and every moderate weight µ. and let P be the usual projection from L1 loc(Cd) to entire functions on Cd. The Toeplitz operator with symbol m acting on a function F is defined to be TmF = P (mF ). Then we show that the Toeplitz operator Tm is an isomorphism from µ (Cd) onto F p,q F p,q The paper is organized as follows: In Section 2 we provide the precise definition of modulation spaces and localization operators, and we collect their basic properties. In particular, we investigate the Weyl symbol of the composition of two localization operators. In Section 3, which contains our main contribution, we construct the canonical isomorphisms and prove Theorem 1.2. In Section 4 we derive a refinement of the spectral invariance of pseudodifferential operators and prove the general isomorphism theorem, of which Theorem 1.1 is a special case. Finally in Section 5 we give applications to Gabor multipliers and Toeplitz operators. 2. Time-Frequency Analysis and Localization Operators We first set up the vocabulary of time-frequency analysis. For the notation we follow the book [15]. For a point z = (x, ξ) ∈ R2d in phase space the time-frequency shift of a function f is π(z)f (t) = e2πiξ·tf (t − x), t ∈ Rd. loc(Rd) The short-time Fourier transform: Fix a non-zero function g ∈ L1 which is usually taken in a suitable space of Schwartz functions. Then the short- time Fourier transform of a function or distribution f on Rd is defined to be Vgf (z) = hf, π(z)gi (2) provided the scalar product is well-defined for every z ∈ R2d. Here g is called a "window function". If f ∈ Lp(Rd) and g ∈ Lp′(Rd) for the conjugate parameter p′ = p/(p − 1), then the short-time Fourier transform can be written in integral form as z ∈ R2d , Vgf (z) =ZRd f (t)g(t − x) e−2πiξ·t dt. In general, the bracket h·,·i extends the inner product on L2(Rd) to any dual pairing between a distribution space and its space of test functions, for instance g ∈ S(Rd) and f ∈ S′(Rd), but time-frequency analysis often needs larger distribution spaces. THE RANGE OF LOCALIZATION OPERATORS 5 Weight functions: We call a locally bounded, strictly positive weight function m on R2d moderate, if z∈R2d(cid:18)m(z + y) m(z) sup m(z) (cid:19) := v(y) < ∞ m(z − y) , for all y ∈ R2d . The resulting function v is a submultiplicative weight function, i.e., v is even and satisfies v(z1 + z2) ≤ v(z1)v(z2) for all z1, z2 ∈ R2d, and then m satisfies m(z1 + z2) ≤ v(z1)m(z2) for all z1, z2 ∈ R2d . (3) Given a submultiplicative weight function v on R2d, any weight satisfying the condition (3) is called v-moderate. For a fixed submultiplicative function v the set Mv := {m ∈ L∞loc(R2d) : 0 < m(z1 + z2) ≤ v(z1)m(z2) ∀z1, z2 ∈ R2d} contains all v-moderate weights. We will use several times that every v-moderate weight m ∈ Mv satisfies the following bounds: 1 v(z1 − z2) ≤ m(z1) m(z2) ≤ v(z1 − z2) for all z1, z2 ∈ R2d . (4) This follows from (3) by replacing z1 with z1 − z2. Modulation spaces M p,q m for arbitrary weights: For the general definition of modulation spaces we choose the Gaussian function h(t) = 2d/4e−πt·t as the canoni- cal window function. Then the short-time Fourier transform is defined for arbitrary elements in the Gelfand-Shilov space (S1/2 1/2 )′(Rd) of generalized functions. The mod- ulation space M p,q 1/2 )′(Rd) such that the norm m (Rd), 1 ≤ p, q < ∞ consists of all elements f ∈ (S1/2 m =(cid:16)ZRd(cid:16)ZRd Vhf (x, ξ)pm(x, ξ)p dx(cid:17)q/p If p = ∞ or q = ∞, we make the usual modification and replace the If m = 1, then we usually write M p,q is finite. integral by the supremum norm k · kL∞. instead of M p,q m = M p,p m and M p = M p,p. m . We also set M p dξ(cid:17)1/q = kVhfkLp,q m (5) kfkM p,q The reader who does not like general distribution spaces, may interprete M p,q m (Rd) as the completion of the finite linear combinations of time-frequency shifts H0 = span{π(z)h : z ∈ R2d} with respect to the M p,q m -norm for 1 ≤ p, q < ∞ and as a weak∗-closure, when p = ∞ or q = ∞. These issues arise only for extremely rapidly decaying weight functions. If m ≥ 1 and 1 ≤ p, q ≤ 2, then M p,q m (Rd) is in fact a subspace of L2(Rd). If m is of polynomial type (cf. (1)), then M p,q m (Rd) is a subspace of tempered distributions. This is the case that is usually considered, although the theory of modulation spaces was developed from the beginning to include arbitrary moderate weight functions [12, 15, 17]. Norm equivalence: Definition (5) uses the Gauss function as the canonical window. The definition of modulation spaces, however, does not depend on the 6 KARLHEINZ GR OCHENIG AND JOACHIM TOFT particular choice of the window. More precisely, if g ∈ M 1 then there exist constants A, B > 0 such that v , g 6= 0, and m ∈ Mv, AkfkM p,q m ≤ kVgfkLp,q m ≤ BkfkM p,q m = BkVhfkLp,q m . We will usually write (6) for the equivalent norms. kVgfkLp,q m ≍ kfkM p,q m Localization operators: Given a non-zero window function g ∈ M 1 symbol or multiplier m on R2d, the localization operator Ag by v (Rd) and a m is defined informally m(z)Vgf (z)π(z)g dz , (7) mf =ZR2d Ag provided the integral exists. A useful alternative definition of Ag definition m is the weak (8) While in general the symbol m may be a distribution in a modulation space of the form M∞1/v(R2d) [9,30], we will investigate only localization operators whose symbol is a moderate weight function. mf, kiL2(Rd) = hmVgf, VgkiL2(R2d) . hAg Taking the short-time Fourier transform of (7), we find that Vg(Ag mf )(w) =ZR2d m(z)Vgf (z)hπ(z)g, π(w)gi dz =(cid:0)(mVgf ) ♮ Vgg(cid:1)(w) , with the usual twisted convolution ♮ defined by Since (F ♮ G)(w) =ZR2d F 7→ZR2d F (z)G(w − z)e2πiz1·(z2−w2) dz. F (z)hπ(z)g, π(·)gi dz is the projection from arbitrary tempered distributions on R2d onto functions of the form Vgf for some distribution f , the localization operator can been seen as the composition of a multiplication operator and the projection onto the space of short-time Fourier transforms. In this light, localization operators resemble the classical Toeplitz operators, which are multiplication operators following by a projection onto analytic functions. Therefore they are sometimes called Toeplitz operators [19,28,30]. If the window g is chosen to be the Gaussian, then this formal similarity can be made more precise. See Proposition 5.5. 2.1. Mapping Properties of Localization Operators. The mapping proper- ties of localization operators on modulation spaces resemble closely the mapping properties of multiplication operators between weighted Lp-spaces. The bound- edness of localization operators has been investigated on many levels of general- ity [9, 29, 33]. We will use the following boundedness result from [10, 30]. Lemma 2.1. Let m ∈ Mv and µ ∈ Mw. Fix g ∈ M 1 operator Ag µm(Rd). vw(Rd). Then the localization 1/m is bounded from M p,q µ (Rd) to M p,q THE RANGE OF LOCALIZATION OPERATORS 7 REMARK: The condition on the window g is required to make sense of Vgf for f µ (Rd) and of Vgk for k in the dual M p′,q′ µm)′ 1/(µm)(Rd) = (M p,q in the domain space M p,q of the target space M p,q µm for the full range of parameters p, q ∈ [1,∞]. For fixed p, q ∈ [1,∞] weaker conditions may suffice, because the norm equivalence (6) still v for r ≤ min(p, p′, q, q′) [31]. holds after relaxing the condition g ∈ M 1 v into g ∈ M r On a special pair of modulation spaces, Ag 1/m is even an isomorphism [19]. Likewise A1/m is an isomorphism from M 2 Lemma 2.2. Let g ∈ M 1 from M 2 θ (Rd) onto M 2 the composition Ag morphism on M 2 1/mAg 1/θ(Rd). 1/θ(Rd). v , m ∈ Mv, and set θ = m1/2. Then Ag 1/θ(Rd) onto M 2 m is an isomorphism on M 2 θ (Rd), and Ag m is an isomorphism θ (Rd). Consequently 1/m is an iso- mAg Lemma 2.2 is based on the equivalence hAg mf, fi = hm,Vgf2i = kVgf · θk2 2 ≍ kfk2 M 2 θ , (9) 1/mAg and is proved in detail in [19, Lemma 3.4]. 2.2. The Symbol of Ag m. The composition of localization operators is no longer a localization operator, but the product of two localization operators still has a well behaved Weyl symbol. In the following we use the time-frequency calculus of pseudodifferential operators as developed in [16, 18]. Compared to the standard pseudodifferential operator calculus it is more restrictive because it is related to the constant Euclidean geometry on phase space, on the other hand, it is more general because it works for arbitrary moderate weight functions (excluding exponential growth). Given a symbol σ(x, ξ) on Rd × Rd ≃ R2d, the corresponding pseudodifferential operator in the Weyl calculus Op(σ) is defined formally as Op(σ)f (x) =ZZR2d σ(cid:16)x + y 2 , ξ(cid:17)e2πi(x−y)·ξf (y) dydξ with a suitable interpretation of the integral. If C is a class of symbols, we write Op(C) = {Op(σ) : σ ∈ C} for the class of all pseudodifferential operators with symbols in C. For the control of the symbol of composite operators we will use the following characterization of the generalized Sjostrand class from [16]. For the formulation associate to a submultiplicative weight v(x, ξ) on R2d the rotated weight on R4d defined by v(x, ξ, η, y) = v(−y, η). (10) For radial weights, to which we will restrict later, the distinction between v and v is unnecessary. Theorem 2.3. Fix a non-zero g ∈ M 1 in M∞,1 H ∈ L1 v . An operator T possesses a Weyl symbol ), if and only if there exists a (semi-continuous) function , T ∈ Op(M∞,1 v v(R2d) such that v hT π(z)g, π(y)gi ≤ H(y − z) for all y, z ∈ R2d . 8 KARLHEINZ GR OCHENIG AND JOACHIM TOFT v REMARK: This theorem says the symbol class M∞,1 is characterized by the off- diagonal decay of its kernel with respect to time-frequency shifts. This kernel is in fact dominated by a convolution kernel. The composition of operators can then studied with the help of convolution relations. Clearly this is significantly easier than the standard approaches that work with the Weyl symbol directly and the twisted product between Weyl symbols. See [20] for results in this direction. Theorem 2.4. Assume that g ∈ M 1 θ ∈ Mv1/2. Then Ag 1/θ ∈ Op(M∞,1 Proof. We distinguish the window g of the localization operator Ag m from the win- dow h used in the expression of the kernel hT π(z)h, π(y)hi. Choose h to be the Gaussian, then h ∈ M 1 v for every submultiplicative weight v. Let us first write the kernel hT π(z)h, π(y)hi informally and justify the convergence of the integrals later. Recall that vs(Rd), T ∈ Op(M∞,1 ) for s ≥ 1/2, and vs−1/2). θT Ag vs T (Ag 1/θf ) = T(cid:16)ZR2d θ(u)−1hf, π(u)giπ(u)g du(cid:17) θT Ag Then hAg =ZZR4d Now set 1/θπ(z)h, π(y)hi = hT Ag 1/θπ(z)h, Ag θπ(y)hi 1 θ(u)(cid:10)π(z)h, π(u)g(cid:11)(cid:10)T π(u)g, π(u′)g(cid:11) θ(u′)(cid:10)π(y)h, π(u′)g(cid:11) dudu′ . G(z) = hg, π(z)hi = Vhg(z) and G∗(z) = G(−z) (11) and let H be a dominating function in L1 u). Since time-frequency shifts commute up to a phase factor, we have v(R2d), so that hT π(u)g, π(u′)gi ≤ H(u′− Before substituting all estimates into (11), we recall that θ is √v-moderate by assumption and so (4) says that hπ(z)h, π(u)gi = G(z − u) . θ(u′) θ(u) ≤ v(u′ − u)1/2 for all u, u′ ∈ R2d . Now by (11) we get θT Ag G(z − u)H(u′ − u)G(y − u′) dudu′ θ(u′) θ(u) G(z − u)v(u′ − u)1/2H(u′ − u)G(y − u′) dudu′ 1/θπ(z)h, π(y)hi hAg ≤ ZR2dZR2d ≤ ZR2dZR2d = (cid:16)G ∗ (v1/2H) ∗ G∗(cid:17)(z − y) . vs(Rd) and thus G ∈ L1 θT Ag Thus the kernel of Ag assumption g ∈ M 1 1/θ is dominated by the function G ∗ (v1/2H) ∗ G∗. By vs ) and thus vs(R2d), and T ∈ Op(M∞,1 THE RANGE OF LOCALIZATION OPERATORS 9 H ∈ L1 vs(R2d). Then v1/2H ∈ L1 vs−1/2(R2d). Consequently vs ⊆ L1 The characterization of Theorem 2.3 now implies that Ag G ∗ (v1/2H) ∗ G∗ ∈ L1 vs−1/2 ∗ L1 vs ∗ L1 vs−1/2 . θT Ag (12) 1/θ ∈ Op(M∞,1 vs−1/2). Corollary 2.5. Assume that g ∈ M 1 submultiplicative weight. Then Ag 1/mAg v2w(Rd) and m ∈ Mv and w is an arbitrary m ∈ Op(M∞,1 v w )). Proof. In this case T is the identity operator and Id ∈ Op(M∞,1 v0 ) for every sub- multiplicative weight v0(x, ξ, η, y) = v0(η, y). In particular Id ∈ Op(M∞,1 ev2 ). Now replace the weight θ in Theorem 2.4 by m and the condition θ ∈ Mv1/2 by m ∈ Mv and modify the convolution inequality (12) in the proof of Theorem 2.4. 3. Canonical Isomorphisms between Modulation Spaces of Hilbert-Type In [19] we have used a deep result of Bony and Chemin [6] about the existence of isomorphisms between modulation spaces of Hilbert type and then extended those isomorphisms to arbitrary modulation spaces. Unfortunately the result of Bony and Chemin is restricted to weights of polynomial type and does not cover weights moderated by superfast growing functions, such as v(z) = eazb for 0 < b < 1. In this section we construct explicit isomorphisms between L2(Rd) and the modu- lation spaces M 2 θ (Rd) for a general class of weights. We will assume that the weights are radial in each time-frequency variable. Precisely, consider time-frequency vari- ables which we identify by (x, ξ) ≃ z = x + iξ ∈ Cd ≃ R2d, (x1, ξ1; x2, ξ2; . . . ; xd, ξd) = (z1, z2, . . . , zd) ∈ Cd ≃ R2d. Then the weight function m should satisfy (13) (14) m(z) = m0(z1, . . . ,zd) for z ∈ R2d d + = [0,∞)d. Without loss of generality, we may also for some function m0 on R assume that m is continuous on R2d. (Recall that only weights of polynomial type occur in the lifting results in [19]. On the other hand, no radial symmetry is needed in [19].) For each multi-index α = (α1, . . . , αd) ∈ Nd variate Hermite function by 0 we denote the corresponding multi- hα(t) = dYj=1 hαj (tj), where hn(x) = 21/4πn/2 n!1/2 eπx2 dn dxn (e−2πx2 ) is the n-th Hermite function in one variable with the normalization khnk2 = 1. 10 KARLHEINZ GR OCHENIG AND JOACHIM TOFT Then the collection of all Hermite functions hα, α ≥ 0, is an orthonormal basis of L2(Rd). By identifying R2d with Cd via (13), the short-time Fourier transform of hα with respect to h(t) = 2d/4e−πt2 is simply α! (cid:17)1/2 zα e−πz2/2 = e−πix·ξ eα(z) e−πz2 Vhhα(z) = e−πix·ξ(cid:16)πα for z ∈ Cd . (15) REMARK: We mention that a formal Hermite expansion f = Pα cαhα defines a distribution in the Gelfand-Shilov space (S1/2 1/2)′, if and only if the coefficients satisfy cα = O(eǫα) for every ǫ > 0. The Hermite expansion then converges 1/2)′ × S1/2 in the weak∗ topology. Here we have used the fact that the duality (S1/2 1/2 extends the L2-form h·,·i on S1/2 1/2 , and likewise the duality of the modulation spaces 1/θ(Rd). Consequently, the coefficient cα of a Hermite expansion is θ (Rd) × M 2 M 2 uniquely determined by cα = hf, hαi for f ∈ (S1/2 In the following we take the existence and convergence of Hermite expansions for functions and distributions in arbitrary modulation spaces for granted. By dµ(z) = e−πz2 dz we denote the Gaussian measure on Cd. Lemma 3.1. Assume that θ(z) = O(eaz) and that θ is radial in each coordinate. θ (Rd). (a) Then the monomials zα, α ≥ 0, are orthogonal in L2 (b) The finite linear combinations of the Hermite functions are dense in M 2 By using polar coordinates zj = rjeiϕj , where rj ≥ 0 and ϕj ∈ [0, 2π), we get 1/2 )′. See [21] for details. θ(Cd, µ). zα = rαeiα·ϕ and dz = r1 · · · rd dϕdr (16) and the condition on θ in Lemma 3.1 can be recast as (17) for some appropriate function θ0 on [0,∞)d, and r = (r1, . . . , rd) and ϕ = (ϕ1, . . . , ϕd) as usual. θ(z) = θ0(r), Proof. (a) This is well-known and is proved in [11,14]. In order to be self-contained, we recall the arguments. By writing the integral over R2d in polar coordinates in each time-frequency pair, (16) and (17) give ZR2d zαzβθ(z)2 e−πz2 dz =ZZRd +×[0,2π)d ei(α−β)·ϕrα+βe−πr2 θ0(r)2r1 · · · rd dϕdr . The integral over the angles ϕj is zero, unless α = β, whence the orthogonality of the monomials. (b) Density: Assume on the contrary that the closed subspace in M 2 by the Hermite functions is a proper subspace of M 2 zero f ∈ (M 2 hα ∈ M 2 0 in (S1/2 θ (Rd) spanned θ (Rd). Then there exists a non- 1/θ(Rd), such that hf, hαi = 0 for all Hermite functions 0. Consequently the Hermite expansion of f =Pαhf, hαihα = θ (Rd), α ∈ Nd 1/2 )′, which contradicts the assumption that f 6= 0. θ (Rd))′ = M 2 THE RANGE OF LOCALIZATION OPERATORS 11 Definition 1. The canonical localization operator Jm is the localization operator Ah m associated to the weight m and to the Gaussian window h = h0. Specifically, Jmf =ZR2d m(z)hf, π(z)hiπ(z)h dz . (18) (19) For m = θ2 we obtain hJmf, fi = hmVhf, VhfiR2d = kVhf θk2 2 = kfk2 M 2 θ whenever f is in a suitable space of test functions. Our main insight is that localization operators with respect to Gaussian windows In view of the connection to and radial symbols have rather special properties. the localization operators on the Bargmann-Fock space (see below) this is to be expected. Theorem 3.2. If θ is a continuous, moderate function and radial in each time- frequency coordinate, then each of the mappings Jθ : M 2 J1/θ : M 2 Jθ : L2(Rd) → M 2 θ (Rd) → L2(Rd), 1/θ(Rd) → L2(Rd), J1/θ : L2(Rd) → M 2 1/θ(Rd) θ (Rd) is an isomorphism. The proof is non-trivial and requires a number of preliminary results. In these investigations we will play with different coefficients of the form or, more generally, τα(θ) := hJθhα, hαi, τα,s(θ) := τα(θs) = hJθshα, hαi =ZR2d θ(z)s πα α! zα2e−πz2 dz , (20) when θ is a weight function and s ∈ R. We note that the τα,s(θ) are strictly positive, since θ is positive. If θ ≡ 1, then τα,s(θ) = 1, so we may consider the coefficients τα,s(θ) as weighted gamma functions. Proposition 3.3 (Characterization of M 2 moderate and radial function. Then θ with Hermite functions). Let θ be a kfk2 M 2 θ =Xα≥0 hf, hαi2τα(θ2) . (21) Proof. Let f = Pα≥0 cαhα be a finite linear combination of Hermite functions. Since the short-time Fourier transform of f with respect to the Gaussian h is given by Vhf (z) =X cαVhhα(z) = e−πix·ξXα≥0 cαeα(z)e−πz2/2 12 KARLHEINZ GR OCHENIG AND JOACHIM TOFT in view of (15), definition (18) gives kfk2 M 2 θ = ZR2d Vhf (z)2θ(z)2 dz = Xα,β≥0 cαcβZR2d eα(z)eβ(z)θ(z)2 e−πz2 dz =Xα≥0 cα2τα(θ2) . In the latter equalities it is essential that the weight θ is radial in each time- frequency coordinate so that the monomials eα are orthogonal in L2 θ(Cd, µ). In the next proposition, which is due to Daubechies [11], we represent the canon- ical localization operator by a Hermite expansion. Proposition 3.4. Let θ be a moderate, continuous weight function on R2d that is radial in each time-frequency coordinate. Then the Hermite function hα is an eigenfunction of the localization operator Jθ with eigenvalue τα(θ) for α ∈ Nd 0, and Jθ possesses the eigenfunction expansion Proof. By Lemma 3.1(a) we find that, for α 6= β, τα(θ)hf, hαihα for all f ∈ L2(Rd) . (22) Jθf =Xα≥0 hJθhβ, hαi = ZCd = ZCd θ(z)Vhhβ(z) Vhhα(z) dz θ(z)eβ(z)eα(z)e−πz2 dz = 0 . This implies that Jθhα = chα and therefore c = chhα, hαi = hJθhα, hαi = τα(θ). For a (finite) linear combination f =Pβ≥0 cβhβ, we obtain Jθf =Xα≥0 hJθf, hαihα =Xα≥0Xβ≥0 cβhJθhβ, hαihα =Xα≥0Xβ≥0 τβ(θ)δα,βcβhα =Xα≥0 τα(θ)cαhα . 1/θ(Rd) and because 1/θ(Rd) is one-to-one and possesses dense range in M 2 The proposition follows because the Hermite functions span M 2 the coefficients of a Hermite expansion are unique and given by cα = hf, hα). Corollary 3.5. If θ is moderate and radial in each coordinate, then Jθ : L2(Rd) → M 2 Proof. The coefficients in Jθf =Pα≥0 τα(θ)hf, hαihα are unique. If Jθf = 0, then τα(θ)hf, hαi = 0, and since τα(θ) > 0 we obtain hf, hαi = 0 and thus f = 0. Clearly the range of Jθ in M 2 1/θ(Rd) contains the finite linear combinations of Hermite functions, and these are dense in M 2 1/θ(Rd) by Proposition 3.1. 1/θ(Rd). THE RANGE OF LOCALIZATION OPERATORS 13 To show that Jθ maps L2(Rd) onto M 2 1/θ(Rd) is much more subtle. For this we need a new type of inequalities valid for the weighted gamma functions in (20). By Proposition 3.4 the number τα,s(θ) is exactly the eigenvalue of the localization operator Jθs corresponding to the eigenfunction hα. Proposition 3.6. If θ ∈ Mw is continuous and radial in each time-frequency coordinate, then the mapping s 7→ τα,s(θ) is "almost multiplicative". This means that for every s, t ∈ R there exists a constant C = C(s, t) such that C−1 ≤ τα,s(θ)τα,t(θ)τα,−s−t(θ) ≤ C for all multi-indices α . (23) Proof. The upper bound is easy. By Lemma 3.1 the Hermite function hα is a common eigenfunction of Jθs, Jθt, and Jθ−s−t. Since the operator JθsJθtJθ−s−t is bounded on L2(Rd) by repeated application of Lemma 2.1, we obtain that τα,s(θ)τα,t(θ)τα,−s−t(θ) = kJθsJθtJθ−s−thαkL2 ≤ CkhαkL2 = C (24) for all α ≥ 0. The constant C is operator norm of JθsJθtJ−s−t on L2(Rd). For the lower bound we rewrite the definition of τα,s(θ) and make it more explicit by using polar coordinates zj = rjeiϕj , rj ≥ 0, ϕj ∈ [0, 2π), in each variable. Then by assumption θ(z) = θ0(r) for some continuous moderate function θ0 on Rd +, and we obtain τα,s(θ) =ZR2d θ(z)s πα dz + α! zα2e−πz2 = (2π)dZRd . . .Z ∞ . . .Z ∞ 0 0 0 = (2π)dZ ∞ =Z ∞ 0 θ0(r) πα α! r2αe−πr2 r1 · · · rd dr θ0(r1, . . . , rd) dYj=1 1 αj! (πr2 j )αj e−πr2 j r1 · · · rd dr1 · · · drd θ0(cid:0)pu1/π, . . . ,pud/π(cid:1) dYj=1 uαj j αj! e−uj du1 · · · dud . (25) We focus on a single factor in the integral first. The function fn(x) = xne−x/n! takes its maximum at x = n and fn(n) = 1 n! nne−n = (2πn)−1/2(cid:0)1 + O(n−1)(cid:1) by Stirling's formula. Furthermore, fn is almost constant on the interval [n − √n/2, n +√n/2] of length √n. On this interval the minimum of fn is taken at one of the endpoints n ± √n/2, where the value is fn(n ± √n/2) = 1 n! (n ± √n/2)ne−(n±√n/2) = 1 n! nn en(cid:0)1 ± 1 2√n(cid:1)ne∓√n/2 . 14 Since KARLHEINZ GR OCHENIG AND JOACHIM TOFT (2πn)1/2(cid:16) n e(cid:17)n n! lim n→∞ = 1 and lim n→∞(cid:0)1 ± 1 2√n(cid:1)n e∓√n/2 = e−1/8 by Stirling's formula and straight-forward applications of Taylor's formula, we find that fn(x) ≥ c √n for x ∈ [n − √n/2, n + √n/2] and all n ≥ 1 . (26) Now consider the products of the fn's occuring in the integral above. For α = For n = 0 we use the inequality f0(x) ≥ e−1/2 for x ∈ [0, 1/2]. (α1, . . . , αd) ∈ Nd 0 define the boxes Cα = dYj=1(cid:2)αj − 2−1√αj , αj + 2−1 max(√αj, 1)(cid:3) ⊆ Rd , with volume vol (Cα) = Qd have j=1pmax(2−1, αj). Consequently, on the box Cα we dYj=1 uαj j αj! e−uj ≥ C0 dYj=1 1 pmax(2−1, αj) = C0(cid:0)vol Cα(cid:1)−1 (27) for some constant C0 > 0 which is independent of α ∈ Nd 0. Next, to take into account the coordinate change in (25), we define the box (αj + 2−1 max(√αj, 1))1/2 √π i ⊆ Rd , Furthermore the length of each edge of Dα is Dα = , √π dYj=1h (αj − 2−1√αj)1/2 π−1/2(cid:16)(cid:0)αj + √αj/2(cid:1)1/2 if z1, z2 ∈ Dα, when αj ≥ 1 and likewise for αj = 0. Consequently, −(cid:0)αj − √αj/2(cid:1)1/2(cid:17) ≤ π−1/2 then z1 − z2 ⊆ [−π−1/2, π−1/2]d . (28) After these preparations we start the lower estimate of τα,s(θ). Using (27) we obtain τα,s(θ) = Z ∞ 0 0 . . .Z ∞ vol (Cα)ZCα dYj=1 θ0(cid:0)pu1/π, . . . ,pud/π(cid:1)s θ0(cid:0)pu1/π, . . . ,pud/π(cid:1)s 1 uαj j αj! e−uj du1 · · · dud du1 · · · dud . ≥ C0 Since θ is continuous, the mean value theorem asserts that there is a point z = z(α, s) = (z1, z2, . . . , zd) ∈ Cα, such that τα,s(θ) ≥ C0 θ0(cid:0)pz1/π, . . . ,pzd/π(cid:1)s . THE RANGE OF LOCALIZATION OPERATORS 15 Note that the point with coordinates ζ = ζ(α, s) =(cid:0)pz1/π, . . . ,pzd/π(cid:1) is in Dα, consequently Finally τα,s(θ) ≥ C0 θ0(ζ(α, s)) for ζ(α, s) ∈ Dα . τα,s(θ) τα,t(θ) τα,−s−t(θ) ≥ C 3 0 θ0(ζ)s θ0(ζ′)t θ0(ζ′′)−s−t for points ζ, ζ′, ζ′′ ∈ Dα. Since the weight θ is a w-moderate, θ0 satisfies θ0(z1) θ0(z2) ≥ 1 w(z1 − z2) z1, z2 ∈ Rd . (29) Since ζ, ζ′, ζ′′ ∈ Dα, the differences ζ−ζ′′ and ζ′−ζ′′ are in the cube [−π−1/2, π−1/2]d as observed in (28). We conclude the non-trivial part of this estimate by w(z)(cid:17)−s−t 0(cid:16) w(ζ′ − ζ′′)t ≥ C 3 w(ζ − ζ′′)s τα,s(θ)τα,t(θ)τα,−s−t(θ) ≥ C 3 The proof is complete. 0 max z∈[−π−1/2,π−1/2]d 1 1 = C . The next result provides a sort of symbolic calculus for the canonical localization operators Jθs. Although the mapping s → Jθs is not homomorphism from R to operators, it is multiplicative modulo bounded operators. Theorem 3.7. Let θ and µ be two moderate, continuous weight functions on R2d that are radial in each time-frequency variable. For every r, s ∈ R there exists an operator Vs,t that is invertible on every M 2 µ(Rd) such that JθsJθtJθ−s−t = Vs,t . Proof. For s, t ∈ R fixed, set γ(α) = τα,s(θ)τα,t(θ)τα,−s−t(θ) and Vs,tf =Xα≥0 γ(α)hf, hαihα . Clearly, Vs,t = JθsJθtJθ−s−t. Since C−1 ≤ γ(α) ≤ C for all α ≥ 0 by Proposition 3.6, Proposition 3.3 implies that Vs,t is bounded on every modulation space M 2 µ. Like- s,t f = Pα≥0 γ(α)−1hf, hαihα is bounded on wise the formal inverse operator V −1 M 2 µ(Rd), consequently Vs,t is invertible on M 2 µ(Rd). We can now finish the proof of Theorem 3.2. Proof of Theorem 3.2. Choose s = 1 and t = −1, then JθJ1/θ = V1,−1 is invertible on L2. Similarly, the choice s = −1, t = 1 yields that J1/θJθ = V−1,1 is invertible θ . The factorization JθJ1/θ = V1,−1 implies that J1/θ is one-to-one from L2 to on M 2 M 2 θ and that Jθ maps M 2 θ onto L2. The factorization J1/θJθ = V−1,1 implies that θ to L2 and that J1/θ maps L2 onto M 2 Jθ is one-to-one from M 2 θ . θ to L2 and that J1/θ is an We have proved that Jθ is an isomorphism from M 2 isomorphism from L2 to M 2 θ . The other isomorphisms are proved similarly. 16 KARLHEINZ GR OCHENIG AND JOACHIM TOFT REMARK: In dimension d = 1 the invertibility of JθJθ−1 follows from the equiv- alence τn,1(θ)τn,−1(θ) ≍ 1, which can be expressed as the following inequality for weighted gamma functions: C−1 ≤Z ∞ 0 θ0(px/π) xn n! e−xdx Z ∞ 0 xn n! e−xdx ≤ C (30) 1 θ0(px/π) for all n ≥ 0. Here θ0 is the same as before. It is a curious and fascinating fact that this inequality implies that the localization operator J1/θ is an isomorphism between L2(R) and M 2 θ (R). 4. The General Isomorphism Theorems In Theorem 4.3 we will state the general isomorphism theorems. The strategy of the proof is similar to that of Theorem 3.2 in [19]. The main tools are the theorems about the spectral invariance of the generalized Sjostrand classes [16] and the existence of a canonical isomorphism between L2(Rd) and M 2 θ (Rd) established in Theorem 3.2. 4.1. Variations on Spectral Invariance. We first introduce the tools concerning the spectral invariance of pseudodifferential operators. Recall the following results from [16]. Theorem 4.1. Let v be a submultiplicative weight on R2d such that lim n→∞ v(nz)1/n = 1 for all z ∈ R2d , (31) ) and T is invertible on L2(Rd), and let v be the same as in (10). If T ∈ Op(M∞,1 then T −1 ∈ Op(M∞,1 for 1 ≤ p, q ≤ ∞ and all µ ∈ Mv. ). v v Consequently, T is invertible simultaneously on all modulation spaces M p,q µ (Rd) Condition (31) is usually called the Gelfand-Raikov-Shilov (GRS) condition. We prove a more general form of spectral invariance. Since we have formulated all results about the canonical localization operators Jθ for radial weights only, we will assume from now on that all weights are radial in each coordinate. In this case v(x, ξ, η, y) = v(−η, y) = v(y, η), v ) and T is invertible on M 2 If T ∈ Op(M∞,1 As a consequence T is invertible on every modulation space M p,q and we do not need the somewhat ugly distinction between v and v. Theorem 4.2. Assume that v satisfies the GRS-condition, θ2 ∈ Mv, and that both v and θ are radial in each time-frequency coordinate. θ (Rd), then T is invertible on L2(Rd). µ (Rd) for 1 ≤ p, q ≤ ∞ and µ ∈ Mv. Proof. Set eT = JθT J1/θ. By Theorem 3.2, J1/θ is an isomorphism from L2(Rd) θ (Rd) onto L2(Rd), therefore eT is onto M 2 an isomorphism on L2(Rd). θ (Rd) and Jθ is an isomorphism from M 2 THE RANGE OF LOCALIZATION OPERATORS 17 (32) Theorem 4.1 on the spectral invariance of the symbol class M∞,1 θ M 2 θ ↑ J1/θ L2(Rd) T−→ M 2 ↓ Jθ eT−→ L2(Rd) By Theorem 2.4 the operator eT is in Op(M∞,1 inverse operator eT also possesses a symbol in M∞,1 Now, since eT −1 = J−1 T −1 = J1/θeT −1Jθ . 1/θT −1J−1 , we find that θ v1/2 ). Since eT is invertible on L2(Rd), v1/2 , i.e., eT −1 ∈ Op(M∞,1 v1/2 implies that the v1/2 ). Applying Theorem 2.4 once again, the symbol of T −1 must be in M∞,1. As a consequence, T −1 is bounded on L2(Rd). ) and T is invertible on L2(Rd), it follows that T is also Since T ∈ Op(M∞,1 invertible on M p,q v µ (Rd) for every weight µ ∈ Mv and 1 ≤ p, q ≤ ∞. 4.2. An isomorphism theorem for localization operators. We now combine all steps and formulate and prove our main result, the isomorphism theorem for time-frequency localization operators with symbols of superfast growth. v2w(Rd), µ ∈ Mw and m ∈ Mv be such that m is Theorem 4.3. Let g ∈ M 1 radial in each time-frequency coordinate and v satisfies (31). Then the localization µ/m(Rd) for 1 ≤ p, q ≤ ∞. operator Ag m is an isomorphism from M p,q µ (Rd) onto M p,q Proof. Set T = Ag (1) T = Ag (2) T is invertible on M 2 1/mAg 1/mAg m possesses a symbol in M∞,1 θ (Rd) by Lemma 2.2. m. We have already established that v w by Corollary 2.5. These are the assumptions of Theorem 4.2, and therefore T is invertible on M p,q for every µ ∈ Mw ⊆ Mvw and 1 ≤ p, q ≤ ∞. The factorization T = Ag implies that Ag M p,q µ/m(Rd) onto M p,q Now we change the order of the factors and consider the operator T ′ = Ag m is one-to-one from M p,q µ/m(Rd) and that Ag µ (Rd) to M p,q µ (Rd). µ (Rd) 1/mAg m 1/m maps µ/m(Rd) to M p,q µ/m(Rd) and factoring through M p,q v w and is invertible on M 2 from M p,q a symbol in M∞,1 that T ′ is invertible on all modulation spaces M p,q T ′ = Ag that Ag 1/m now yields that Ag µ (Rd) onto M p,q mAg m maps M p,q 1/m is one-to-one from M p,q µ/m(Rd). 1/m µ (Rd). Again T ′ possesses 1/θ(Rd). With Theorem 4.2 we conclude µ/m(Rd). The factorization of µ (Rd) and µ/m(Rd) to M p,q mAg As a consequence Ag m is bijective from M p,q µ (Rd) onto M p,q µ/m(Rd), and Ag 1/m is bijective from M p,q µ/m(Rd) onto M p,q µ (Rd). 18 KARLHEINZ GR OCHENIG AND JOACHIM TOFT 5. Consequences for Gabor Multipliers and Toeplitz Operators on Bargmann Fock Space 5.1. Gabor Multipliers. Gabor multipliers are time-frequency localization oper- ators whose symbols are discrete measures. Their basic properties are the same, but the discrete definition makes them more accessible for numerical computations. Let Λ = AZ2d for some A ∈ GL(2d, R) be a lattice in R2d, g, γ suitable window functions and m a weight sequence defined on Λ. Then the Gabor multiplier Gg,γ,Λ is defined to be m m(λ)hf, π(λ)giπ(λ)γ . (33) Gg,γ,Λ m f =Xλ∈Λ m The boundedness of Gabor multipliers between modulation spaces is formulated and proved exactly as for time-frequency localization operators. See [13] for a detailed exposition of Gabor multipliers and [9] for general boundedness results that include distributional symbols. Proposition 5.1. Let g, γ ∈ M 1 m ∈ Mv. Then Gg,γ,Λ vw(Rd), m be a continuous moderate weight function is bounded from M p,q µ (R2d) to M p,q µ/m(Rd). Gabor frame for L2(Rd). This means that the frame operator Sg,Λf = M g,g,Λ In the following we will assume that the set G(g, Λ) = {π(λ)g : λ ∈ Λ} is a = Pλ∈Λhf, π(λ)giπ(λ)g is invertible on L2(Rd). From the rich theory of Gabor frames we quote only the following result: If g ∈ M 1 v and G(g, Λ) is a frame for L2(Rd), 1 ≤ p ≤ ∞, and m ∈ Mv, then hf, π(λ)gipm(λ)p(cid:17)1/p for all f ∈ M p kfkM p m(Rd). (34) 1 m ≍(cid:16)Xλ∈Λ We refer to [15] for a detailed discussion, the proof, and for further references. Our main result on Gabor multipliers is the isomorphism theorem. Theorem 5.2. Assume that g ∈ M 1 If m ∈ Mv is radial in each time-frequency coordinate, then Gg,g,Λ phism from M p,q v2w(Rd) and that G(g, Λ) is a frame for L2(Rd). is an isomor- µ/m(Rd) for every µ ∈ Mw and 1 ≤ p, q ≤ ∞. µ (R2d) onto M p,q The proof is similar to the proof of Theorem 4.3, and we only sketch the necessary modifications. In the following we fix the lattice Λ and choose γ = g and drop the reference to these additional parameter by writing Gg,γ,Λ m in analogy to the localization operator Ag Proposition 5.3. If g ∈ M 1 isomorphism from M 2 v (Rd) and m ∈ Mv and θ = m1/2, then Gg m is an as Gg θ (Rd) onto M 2 1/θ(Rd). m. m m Proof. By Proposition 5.1 Gg mfkM 2 Next we use the characterization of M 2 multipliers: θ to M 2 θ/m = M 2 1/θ and thus m is bounded from M 2 kGg 1/θ ≤ CkfkM 2 θ by Gabor frames and relate it to Gabor mf, fi =Xλ∈Λ hGg m(λ)hf, π(λ)gi2 ≍ kfk2 (35) M 2 θ . . θ THE RANGE OF LOCALIZATION OPERATORS 19 This identity implies that Gg equivalent norm on M 2 whence Gg m is onto as well. m is one-to-one on M 2 θ and that kGg θ . Since Gg m is self-adjoint, it has dense range in M 2 mfkM1/θ is an 1/θ, Using the weight 1/m instead of m and 1/θ instead of θ, we see that Gg 1/m is an isomorphism from M 2 1/θ(Rd) onto M 2 θ (Rd). Proof of Theorem 5.2. We proceed as in the proof of Theorem 4.3. Define the operators T = Gg θ (Rd) to M 2 θ (Rd), and since T is a composition of two isomorphisms, T is an isomorphism on M 2 1/m. By Proposition 5.1 T maps M 2 θ (Rd). Likewise T ′ is an isomorphism on M 2 m and T ′ = Gg 1/mGg 1/θ(Rd). mGg As in Corollary 2.5 we verify that the symbol of T is in M∞,1 v w (R2d). Theorem 4.2 then asserts that T is invertible on L2(Rd) and the general spectral invariance implies that T is an isomorphism on M p,q µ (Rd). Likewise T ′ is an isomorphism on M p,q µ/m(Rd). This means that each of the factors of T must be an isomorphism on the correct space. 5.2. Toeplitz Operators. The Bargmann-Fock space F = F 2(Cd) is the Hilbert space of all entire functions of d variables such that =Z d C F (z)2e−πz2 kFk2 F dz < ∞ . (36) Here dz is the Lebesgue measure on Cd = R2d. Related to the Bargmann-Fock space F 2 are spaces of entire functions satisfying weighted integrability conditions. Let m be a moderate weight on Cd and 1 ≤ m (Cd) is the Banach space of all entire functions of d p, q ≤ ∞. Then the space F p,q complex variables, such that =ZRd(cid:16)ZRd F (x + iy)pm(x + iy)pe−pπx+iy2/2dx(cid:17)q/p kFkq Let P be the orthogonal projection from L2(Cd, µ) with Gaussian measure dµ(z) = dy < ∞ . F p,q (37) m e−πz2 dz onto F 2(Cd). Then P is given by the formula F (z)eπzwe−πz2 dz . P F (w) =ZCd (38) We remark that a function on R2d with a certain growth is entire if and only if it satisfies F = P F . The classical Toeplitz operator on Bargmann-Fock space with a symbol m is defined by TmF = P (mF ) for F ∈ F 2(Cd). Explicitly Tm is given by the formula (39) m(z)F (z)eπzw e−πz2 dz . TmF (w) =ZCd See [1–4, 11, 14] for a sample of references. 20 KARLHEINZ GR OCHENIG AND JOACHIM TOFT In the following we assume that the symbol of a Toeplitz operator is continu- ous and radial in each coordinate, i.e., m(z1, . . . , zd) = m0(z1, . . .zd) for some continuous function m0 from Rd Theorem 5.4. Let µ ∈ Mw, and let m ∈ Mv be a continuous moderate weight function such that one of the following conditions is fulfilled: + to R+. (1) Either m is radial in each coordinate, (2) or m is of polynomial type. Then the Toeplitz operator Tm is an isomorphism from F p,q 1 ≤ p, q ≤ ∞. µ (Cd) onto F p,q µ/m(Cd) for The formulation of Theorem 5.4 looks similar to the main theorem about time- frequency localization operators. In fact, after a suitable translation of concepts, it is a special case of Theorem 4.3. To explain the connection, we recall the Bargmann transform that maps distri- butions on Rd to entire functions on Cd. Bf (z) = F (z) = 2d/4e−πz2/2ZRd f (t)e−πt2 e2πt·zdt z ∈ Cd . (40) If f is a distribution, then we interpret the integral as the action of f on the function e−πt2e2πt·z. The connection to time-frequency analysis comes from the fact that the Bargmann transform is just a short-time Fourier transform in disguise [15, Ch. 3]. As before, we use the normalized Gaussian h(t) = 2d/4e−πt2 as a window. Identifying the pair (x, ξ) ∈ Rd × Rd with the complex vector z = x + iξ ∈ Cd, the short-time Fourier transform of a function f on Rd with respect to h is Vhf (z) = eπix·ξBf (z)e−πz2/2 . (41) In particular, the Bargmann transform of the time-frequency shift π(w)h is given by B(π(w)h)(z) = e−πiu·ηeπwze−πw2/2 w, z ∈ Cd, w = u + iη . (42) It is a basic fact that the Bargmann transform is a unitary mapping from L2(R) onto the Bargmann-Fock space F 2(Cd). Furthermore, the Hermite functions hα, α ≥ 0 are mapped to the normalized monomials eα(z) = πα/2(α!)−1/2zα. Let m′(z) = m(z). Then (41) implies that the Bargmann transform B maps m (Rd) isometrically to F p,q M p,q m′ (Cd). By straight-forward arguments it follows that m (Rd) onto the Fock space the Bargmann transform maps the modulation space M p,q F p,q m′ (Cd) [15, 25]. The connection between time-frequency localization operators and Toeplitz op- erators on the Bargmann-Fock space is given by the following statement. Proposition 5.5. Let m be a moderate weight function on R2d ≃ Cd and set m′(z) = m(z). The Bargmann transform intertwines Jm and Tm′, i.e., for all f ∈ L2(Rd) (or f ∈ (S1/2 1/2)∗) we have B(Jmf ) = Tm′(Bf ) . (43) THE RANGE OF LOCALIZATION OPERATORS 21 Proof. This fact is well-known [7,11]. For completeness and consistency of notation, we provide the formal calculation. We take the short-time Fourier transform of Jm = Ah m with respect to h. On the one hand we obtain that hJmf, π(w)hi = eπiu·ηB(Jmf )(w) e−πw2/2 , (44) and on the other hand, after substituting (41) and (42), we obtain that hJmf, π(w)hi = ZR2d = ZCd = eπiu·ηZCd m(z)Vhf (z)Vh(π(w)h)(z) dz m(z)Bf (z) B(π(w)h)(z)e−πz2 dz (45) m(z) Bf (z)eπzw e−πz2 dz e−πw2/2 . Comparing (44) and (45) we obtain that B(Jmf )(w) =ZCd m(z)Bf (z)eπzw e−πz2 dz = Tm′Bf (w) . Proof of Theorem 5.4. First assume that (1) holds, i.e., m is radial in each vari- able. The Bargmann transform is an isomorphism between the modulation space m (Rd) and the Bargmann-Fock space F p,q M p,q m′ (Cd) for arbitrary moderate weight function m and 1 ≤ p, q ≤ ∞. By Theorem 4.3 the canonical localization operator µ (Rd) onto M p,q Jm is an isomorphism from M p,q µ/m(Rd). Since Tm′ is a composition of three isomorphism (Proposition 5.5 and (46)), the Toeplitz operator Tm′ is an isomorphism from F p,q µ′/m′(Cd). µ′ (Cd) onto F p,q F p,q ↑ B M p,q µ µ′ µ′/m′ Tm′ −→ F p,q ↑ B Jm−→ M p,q µ/m . (46) Finally replace m′ and µ′ by m and µ. By using Theorem 3.2 in [19] instead of Theorem 4.3, the same arguments show that the result follows when (2) is fulfilled. References [1] V. Bargmann. On a Hilbert space of analytic functions and an associated integral transform. Comm. Pure Appl. Math., 14:187–214, 1961. [2] F. A. Berezin. Wick and anti-Wick symbols of operators. Mat. Sb. (N.S.), 86(128):578–610, 1971. [3] C. A. Berger and L. A. Coburn. Toeplitz operators on the Segal-Bargmann space. Trans. Amer. Math. Soc., 301(2):813–829, 1987. [4] C. A. Berger and L. A. Coburn. Heat flow and Berezin-Toeplitz estimates. Amer. J. Math., 116(3):563–590, 1994. 22 KARLHEINZ GR OCHENIG AND JOACHIM TOFT [5] P. Boggiatto, E. Cordero, and K. Grochenig. Generalized anti-Wick operators with symbols in distributional Sobolev spaces. Integral Equations Operator Theory, 48(4):427–442, 2004. [6] J.-M. Bony and J.-Y. Chemin. Espaces fonctionnels associ´es au calcul de Weyl-Hormander. Bull. Soc. Math. France, 122(1):77–118, 1994. [7] L. A. Coburn. The Bargmann isometry and Gabor-Daubechies wavelet localization operators. In Systems, approximation, singular integral operators, and related topics (Bordeaux, 2000), volume 129 of Oper. Theory Adv. Appl., pages 169–178. Birkhauser, Basel, 2001. [8] J. B. Conway. A course in functional analysis. Springer-Verlag, New York, second edition, 1990. [9] E. Cordero and K. Grochenig. Time-frequency analysis of localization operators. J. Funct. Anal., 205(1):107–131, 2003. [10] E. Cordero and K. Grochenig. Symbolic calculus and fredholm property for localization operators. J. Fourier Anal. Appl., 12(4):345–370, 2006. [11] I. Daubechies. Time-frequency localization operators: a geometric phase space approach. IEEE Trans. Inform. Theory, 34(4):605–612, 1988. [12] H. G. Feichtinger and K. Grochenig. Banach spaces related to integrable group representa- tions and their atomic decompositions. I. J. Functional Anal., 86(2):307–340, 1989. [13] H. G. Feichtinger and K. Nowak. A first survey of Gabor multipliers. In Advances in Gabor analysis, Appl. Numer. Harmon. Anal., pages 99–128. Birkhauser Boston, Boston, MA, 2003. [14] G. B. Folland. Harmonic Analysis in Phase Space. Princeton Univ. Press, Princeton, NJ, 1989. [15] K. Grochenig. Foundations of time-frequency analysis. Birkhauser Boston Inc., Boston, MA, 2001. [16] K. Grochenig. Time-frequency analysis of Sjostrand's class. Revista Mat. Iberoam., 22(2):703– 724, 2006. [17] K. Grochenig. Weight functions in time-frequency analysis. In e. a. L. Rodino, M.-W. Wong, editor, Pseudodifferential Operators: Partial Differential Equations and Time-Frequency Analysis, volume 52, pages 343 – 366. Fields Institute Comm., 2007. [18] K. Grochenig and Z. Rzeszotnik. Banach algebras of pseudodifferential operators and their almost diagonalization. Ann. Inst. Fourier (Grenoble), 58(7):2279–2314, 2008. [19] K. Grochenig and J. Toft. Isomorphism properties of Toeplitz operators and pseudo- differential operators between modulation spaces. J. Anal. Math., To appear. [20] A. Holst, J. Toft and P. Wahlberg. Weyl product algebras and modulation spaces. J. Funct. Anal., 251:463-491, 2007. [21] A. J. E. M. Janssen. Bargmann transform, Zak transform, and coherent states. J. Math. Phys., 23(5):720–731, 1982. [22] N. Lerner. The Wick calculus of pseudo-differential operators and some of its applications. Cubo Mat. Educ., 5(1):213–236, 2003. [23] N. Lerner and Y. Morimoto. A Wiener algebra for the Fefferman-Phong inequality. In Sem- inaire: Equations aux D´eriv´ees Partielles. 2005–2006, S´emin. ´Equ. D´eriv. Partielles, pages Exp. No. XVII, 12. ´Ecole Polytech., Palaiseau, 2006. [24] M. A. Shubin. Pseudodifferential Operators and Spectral Theory. Springer-Verlag, Berlin, second edition, 2001. Translated from the 1978 Russian original by Stig I. Andersson. [25] M. Signahl, J. Toft Mapping properties for the Bargmann transform on modulation spaces, Preprint, arXiv:1002.3061. [26] S. Thangavelu. Lectures on Hermite and Laguerre expansions, volume 42 of Mathemati- cal Notes. Princeton University Press, Princeton, NJ, 1993. With a preface by Robert S. Strichartz. [27] J. Toft. Regularizations, Decompositions and Lower Bound Problems in the Weyl Calculus. Comm. Partial Differential Equations, 25 (7& 8): 1201–1234, 2000. [28] J. Toft. Subalgebras to a Wiener type algebra of pseudo-differential operators. Ann. Inst. Fourier (Grenoble), 51(5):1347–1383, 2001. THE RANGE OF LOCALIZATION OPERATORS 23 [29] J. Toft. Continuity properties for modulation spaces, with applications to pseudo-differential calculus. I. J. Funct. Anal., 207(2):399–429, 2004. [30] J. Toft. Continuity and Schatten-von Neumann Properties for Toeplitz Operators on Modu- lation Spaces in: J. Toft, M. W. Wong, H. Zhu (Eds), Modern Trends in Pseudo-Differential Operators, Operator Theory Advances and Applications Vol 172, Birkhauser Verlag, Basel: pp. 313–328, 2007. [31] J. Toft. Multiplication properties in pseudo-differential calculus with small regularity on the symbols. J. Pseudo-Differ. Oper. Appl. 1(1): 101–138, 2010. [32] H. Triebel. Theory of function spaces. Birkhauser Verlag, Basel, 1983. [33] M. W. Wong. Wavelets Transforms and Localization Operators, volume 136 of Operator Theory Advances and Applications. Birkhauser, 2002. Faculty of Mathematics, University of Vienna, Nordbergstrasse 15, A-1090 Vi- enna, Austria E-mail address: [email protected] Department of Computer science, Physics and Mathematics, Linnaeus University, Vaxjo, Sweden E-mail address: [email protected]
1707.01282
5
1707
2017-10-31T01:09:29
Addition Formulas of Leaf Functions According to Integral Root of Polynomial Based on Analogies of Inverse Trigonometric Functions and Inverse Lemniscate Functions
[ "math.FA" ]
The second derivative of a function r(t) with respect to a variable t is equal to -n times the function raised to the 2n-1 power of r(t); using this definition, an ordinary differential equation is constructed. Graphs with the horizontal axis as the variable t and the vertical axis as the variable r(t) are created by numerically solving the ordinary differential equation. These graphs show several regular waves with a specific periodicity and waveform depending on the natural number n. In this study, the functions that satisfy the ordinary differential equation are presented as the leaf functions. For n = 1, the leaf function become a trigonometric function. For n = 2, the leaf function becomes a lemniscatic elliptic function. These functions involve an additional theorem. In this paper, based on the additional theorem for n = 1 and n = 2, the double angle and additional theorem for n = 3 are presented.
math.FA
math
Addition Formulas of Leaf Functions According to Integral Root of Polynomial Based on Analogies of Inverse Trigonometric Functions and Inverse Lemniscate Functions Kazunori Shinohara Department of Mechanical System Engineering, Daido University 10-3 Takiharu-cho, Minami-ku, Nagoya 457-8530, Japan Abstract Inverse trigonometric functions arcsin(r) and arccos(r) are obtained by integrating the reciprocals of the root of the polynomial 1 − t2. Based on this analogy, the inverse lemniscate functions arcsl(r) and arccl(r) are created by integrating the reciprocal of the root of the polynomial 1 − t4. The lemniscate functions are then extended to the Jacobi elliptic functions, which in turn are improved using the theta functions. However, to the best of our knowledge, the integration of the reciprocals of the root of the polynomial 1 − t6 has never been published. The inverse functions based on 1 − t6 are defined as arcsleaf3(r) and arccleaf3(r), and they were determined to produce waves different from those of the trigonometric and lemniscate functions. This paper presents the addition formulas of the artificially created functions sleaf3(l) and cleaf3(l). These formulas were numerically verified through examples. Keywords: Leaf functions, Trigonometric functions, Lemniscatic elliptic functions, Inverse functions, Addition formulas 1 Introduction The second derivative of a function r(l) with respect to a variable l is equal to -n times the function raised to the 2n-1 power of r(l) using this definition, an ordinary differential equation is constructed. Graphs with the horizontal axis as the variable l and the vertical axis as the variable r(l) are created by numerically solving the ordinary differential equation. These graphs show several regular waves with a specific periodicity and waveform depending on the natural number n [1] [2].   lrd 2 dl 2 The definition of the arc leaf function arcsleafn(r) is obtained by the above equation. arcsleaf (1.1) (1.2)   lrn   1  l n 2 r   r n  0 dt t  1 2 n In case of n=3, the above ordinary differential equations are as follows:   lrd 2 dl 2 3   5 lr (1.3) These equations are related in the nonlinear equations (Quintic Duffing equation) [3]. To obtain the solution accuracy, or to grasp behavior of the solution, the addition formula often be demanded in the solution process [4] [5] [6]. Historically, the trigonomic function (in n=1 of the Eq. (1.2)) and the lemniscatic elliptic function (in n=2 of the Eq. (1.2)) have been discussed [7] [8] [9] [10] [11]. However, in case of n=3, the integration of the reciprocals of the root of the polynomial 1− t6 has never been published. The variable n represents a natural number. For n = 1, the above equation is written as follows: arcsleaf (1.4) arcsin   l   r   r 1   r 10 dt  2 t 1  2   l   2 l  2 cos   l   l  or sin2 sleaf sleaf The function arcsin(r) represents inverse trigonometric functions. The double angle and the addition formulas of the function sin(l) are well known, as follows: sin sin(  or For n = 2, Eq.(1.2) is rewritten as follows:   l cos 1  l    l 2 sleaf   l 2 cleaf sin   l  1 sin    l 1 )  2 sleaf (1.5) (1.6)   l 1   cos  l  cleaf   1 l sleaf cleaf   l   l  l l 1 1 2 1 2 l 1 1 2 1 1 1 dt 1  t arcsleaf   r 2 r  0  arcsl   r  l 4 (1.7) Function arcsl(r) represents the inverse lemniscatic elliptic function [12] [13] [14]. The double angle and the addition formulas are as follows: sl  l 2   2   lsl 1  4     lsl 1      4 lsl  lsl 1  l 2     lsl 1 1  or sleaf  l 1 2  l 2     or sleaf 2  l 2   2   sleaf l 1  2  sleaf 1       2 sleaf    4 l 4    l 2     (1.8)     lsl 1 4 4      lsl lsl  2 2       2 lsl lsl 1  1   sleaf l 1 1 2    2  1   2  sleaf   1  2  l 2 sleaf 4       2 l 1 2 sleaf 2  sleaf  l 2  l 2  2 1   2   sleaf 2 4    l 1     (1.9) In this study, the double angles with respect to the leaf functions sleaf3(l) and cleaf3(l) were derived by using the analogous of the trigonometric functions and lemniscatic elliptic function. Next, the addition formulas of these leaf functions were derived. 2 Double angle 2.1 Leaf function: sleaf3(l) For n = 3, the double angle is derived. Variables l1 and l2 are defined as follows: l 1 l 2 r   1 0 r   2 0 dt 1  6 t dt 1  t 6 3  1 l sleaf Variable l1 is related to variable r1 as follows: r  1 Variable l2 is related to variable r2 as follows: (2.1) (2.2) (2.3)  l 2 3 2 2l sleaf (2.4) r  2 We try to find the relationship between r1 and r2 when the following relationship exists. l  1 The following equation is obtained by using Eqs. (2.1), (2.2), and (2.5).  The relationship between r1 and r2 is obtained as the double angle formula of the leaf function: (2.5) (2.6) dt 1  dt 1    2 r 2 t t r 1 6 0 6 0 (2.7) 2 r 1  r r 1 6  2 2 r 81 6  2  By differentiating it with respect to variable r2, the following equation is obtained. dr 1 dr 2 The following equation is transformed by using the variable r2. r 2 40 6  2   r 81 6  2 r 12 2 r 6 2 16 1 (2.8)   3 2 1  1 r 6 1   r 81 6  2 r 20 6   2 3  2 r 8 12 2 1 The following equation is obtained by using Eqs. (2.8) and (2.9). 1  1 r 6 1 dr 1 dr 2  3   r 81 6  2 2 r r 20 8 6 12   2 2 1 r 2 40 6  2   r 81 6  2 3 2  r 16 12 2 r 1 6 2   2  1 r 6 2 The following equation is derived from the above equation. dr 1 1   2 r 6 1 dr 2 1  r 6 2 (2.9) (2.10) (2.11) Eq. (2.6) satisfies the above equation. The double angle of the leaf function is obtained as follows: (1) In case of an inequality  l 2 m  1 2    3     1 2    3  m  (See [1] [2] for the constant π3.) 2    sleaf 3  l 2   2 sleaf    l 1  3  sleaf 81  sleaf   6 l 3 3 6    l (2) In case of an inequality 2 m     1 2    3   l 2 m     3 2    3  sleaf  l 2  3  2 sleaf    l sleaf 1  3    6 sleaf l 81  3 3 6    l (2.12) (2.13) The differentiation of the leaf function sleaf3(l) is defined by the domain of the variable l. The sign is defined as follows [1] [2]. (1) In case of the inequality  l 2 m  2 m     1 2    3     1 2    3  (2.14) (2.15) 3  1     l sleaf sleaf d dl (2) In case of the inequality 3   6 l sleaf   l 3  1   sleaf   6 l 3 d dl 2 m     1 2    3   l 2 m     3 2    3     6 l 3 1  sleaf depends on the domain of variable l. The sign of the term Because of change in sign, it is necessary to classify Eqs. (2.12) and (2.13). 2.2 Leaf function: cleaf3(l) The leaf function sleaf3(l) is related to the leaf function cleaf3(l) as follows[1] [2]: 2    l   cleaf 3 2    l 2   sleaf 3 2    l  cleaf 3 2    l   1 (2.16)  sleaf 3 (2.17) Variable l in the above equation is replaced by variable 2l.   cleaf 3  cleaf 3  sleaf 3 sleaf 3 2   l 2  l 2  l 2  l 2 1            2 2 2 2 By substituting Eq. (2.12) or Eq. (2.13) into Eq. (2.17), we get the following equation. cleaf 3  l 2       l cleaf 2     2 cleaf l 81   3 2  cleaf 2      cleaf l 8 3 4    l 1   cleaf 8  6 3 3   8 l 3 (2.18) When the function cleaf3(2l) is derived from Eq. (2.17), both plus and minus signs occur in the function cleaf3(2l). We have to choose the plus sign so that the initial condition, namely, cleaf3(0) =1 [1] [2] is satisfied. 3 Additional formulas 3.1 Leaf function: sleaf3(l1+l2) With respect to arbitrary variables l1 and l2, the following equation is satisfied. 4    l 1 (3.1) (3.2) (3.3) (3.4) (3.5) (3.6)  l  3 3   sleaf sleaf    l  1 l   2      sleaf sleaf l 41  2 3          3 sleaf sleaf l l  3 2 3 1         4 2 sleaf sleaf l l 4  1 3 4 3 3 3  l 3 2 2 2 sleaf      sleaf l 4  1  2   sleaf l 1 3    2 sleaf 2 3   4 l 1 2   3 sleaf l  1   2   l 1     sleaf 2  l 3 3  sleaf 3 l   l 1  sleaf  sleaf 41   2   2  l 2  l 2 3 3 2  The above equation is set as follows:   llg , 2 1  lg 1,    llp , 2 12 2  llp , 13  llp , 2 11  llp , 2 13 sleaf  l 13    2 l   l 2       2 2  llp , 2 11   sleaf   l 13  l 2  sleaf  3 l  2  sleaf 3  l 2    l 13 sleaf  l  1  lp 12 , l 2    sleaf    l 13 3 sleaf 3  l 2   sleaf    l 13 sleaf 3  l 2  3  llp , 2 13   41  sleaf 4    l 13  sleaf 3  l 2 2    4  sleaf 2    l 13  sleaf 3  l 2  4   llp ,  2 12 l  1 sleaf 3   3     3  l 2  lp  13 l  1 ,  l 2  For Eqs. (3.4)–(3.6), the following equations are obtained by differentiating with respect to variable l1.   l  1   l    l , 2 3 3 2  3 sleaf    l sleaf 2 3   5 l 1 3 (3.7)  lp 1 1 l  1  sleaf l  1 sleaf l  2 sleaf 2    l 13 sleaf 3  l 2  sleaf 2    l 13   sleaf 3    l 13 sleaf  l  1  sleaf  l  1   2  l 2  sleaf 3   l 13  l 2 3     l 13 sleaf  l  1 (3.8)    sleaf l 3 13   l 13 sleaf 8  sleaf  l  1 2  sleaf 3  l 2 2     l 13 sleaf  l  1   sleaf 16 3    l 13   2   2 2  sleaf 8    l sleaf 3 13  l sleaf    l 13 2   sleaf 3  l 2   The following equation is substituted into Eq. (3.7).  2   l  3  sleaf   5 l 3 sleaf l 2  3  l 2 4     l 13 sleaf  l  1 (3.9) (3.10) For Eqs. (3.4)–(3.6), the following equations are obtained by differentiating with respect to variable l2.  lp  11 l  sleaf  3 l  (3.11) sleaf 3 sleaf 3    l 13   l 13 sleaf  l  1  5  l  l     l , 2 2 2 2 2  lp  1 2 l  2 , l 2        l 1 sleaf 3 3    l 1  l 2   sleaf  3 l  2  sleaf 3 sleaf 3 2    l 1  3  sleaf 3  l 2   2    l 2 2 sleaf 3   l  sleaf 3 3  sleaf  3 l  2 2  lp  1 3 l  2 , l 2   16  sleaf  l 2 3 3  sleaf 2    l 1 3  l 2  3  sleaf  l  2 2    l 1 sleaf 2    l 1 3  2  sleaf  l 2 3 2        l 2    l 1 sleaf  3 l  2 (3.12) 2    l 2 4    l 1 3 sleaf   l 3 sleaf 8  sleaf  l  3 2  l 2  3 sleaf  l  2 (3.13)  8 sleaf  l 2 3   sleaf 3 For Eq. (3.2), the following equation is obtained by differentiating with respect to variable l1.  2    l 2   lp 11 , l 2  ,  lg  1 l  1    lp 11 , l 2 2      2  lg 1 , l 2  lp 1 2   l ,  lp 3 1 2 2 l , 2 l l ,  lp 1 3  lp 1 2    2 2 , l    lp  11 l  1  lg , 2 1   lp  1 3 l  1  2   l , 2 2 , l 2   2  lp  1 2 l  1 , l 2 2    lp 1 3 , l 2     (3.14) Using Eqs. (3.11)–(3.13), the numerator of Eq. (3.16) is expanded as follows: 2 2 2 2 2 2  2      2 , l , l , l , l , l , l       lp 11  lp 1 3  lp  1 2 l  1   lp  11 l  1  Using Eqs. (3.7)–(3.9), the numerator of Eq. (3.14) is expanded as follows (see Appendix B):   lp 2  11     2 sleaf 8 3  2  8      l sleaf l sleaf 3 1 3     sleaf l    l 1 3 l  1     sleaf sleaf l 3 1 3    sleaf l    l 3 l    lp 1 2   lp  1 3 l  1     sleaf l 3 1  16     l sleaf 3 1   16  sleaf 8 3    sleaf l 3  8 sleaf 3 sleaf 3  l  24   sleaf 3  24   sleaf 3   l     l 1    l    l 1    l    l 1   l 1    l 1 (3.15) sleaf 3 sleaf 3 sleaf 3 sleaf 3 sleaf 3  lp 1 2  l        , l , l        5 8  2 5 2 5 2 2 6 2 6 3 2 5 2 8 2 3 4 2 2 4 2 2 2 2 2 On the other hand, for Eq. (3.2), the following equation is obtained by differentiating with respect to variable l2. 2  2   , l 2    lp 11 , l 2  , l 2   lp  11 l   lg  1 l    2  lp 11 , l 2 2      2  lg 1 , l 2  lp 1 2   l ,  lp 13 2 2   , l 2  2  lp , 1 2    lp 13  l , 2 l 2 2  lg , 2 1   lp  13 l   2 2 , l 2  l  2  lp  1 2 l  2 , l 2 2    lp 13 , l 2     (3.16) 2 ,   lp  11 l  2    l 8  1    l sleaf 1 3 l  2  2  lp 1 2 , l 2  3 sleaf    l 2 3 5     l sleaf 1  16 sleaf  6 , l 2  2  lp  1 2 l     l    l 1 3 2 3 2 5   2   l     l 1 6  8 sleaf sleaf  l 3 3 2 3 3 2     l sleaf 1   16 sleaf  3 3 5 8 2    l    l 1   l sleaf ,   lp 2  11   2  sleaf 8  2  8    sleaf sleaf 3 4   2  l sleaf   l 1 3 3 2 3 3 , , l l    lp 1 1  lp 1 3     24 sleaf       l sleaf 1   sleaf    l l  1      sleaf sleaf l 24 1    sleaf l    sleaf l l  3 5 2 3 4 3 8 2 3 2 3  l  2 Using the above equation, the following equation is derived. 2      lp 1 2 , l 2 2   , l 2    lp  1 3 l  2 2 3   2 (3.17) (3.18)  l 2  ,  lg  1 l  1 , l 2   lg  1 l  2 The following equation that satisfies Eq. (3.18) (see Appendix A) is derived. 2 2 , l l    0,  lg 1  lg 1 Using the initial condition sleaf3(0)=0 and ∂sleaf3(0)/∂l=1, the function g(l1+l2, 0) is written as follows: (3.19)   0 3 2 3 l   l 1 sleaf      sleaf l 41  3 1   3 l sleaf    sleaf sleaf sleaf    l  2          sleaf l 4 0   3 2      l sleaf sleaf l 0   3 2 1          2 sleaf l l 4 0   1 4 2 3 3 4 2 2 3 2 3 2   0 3  sleaf l  1 3 l  1 l    2  2  2 l     sleaf 2  sleaf l 3 1  23    0  sleaf 3 4    0   lg 1   2 2   0, l    l sleaf  3 1   l sleaf l 41   3 1  2   l l sleaf  1 3 2   Using Eqs. (3.19) and (3.20), the following equation is obtained. 4    0 3 (3.20)    lg 1 , l 2  2     l 2   sleaf l 13  sleaf    sleaf 41    sleaf  sleaf 41      l  13    l 13    l 13    l 13 4  lg  1 sleaf  3 l  2  sleaf 3 sleaf 3  sleaf 3 3 4 l 0, 2   l 2 2     l 2   l  2    2 l 2   l 2 sleaf 3  sleaf 4     sleaf l 13  sleaf 4     l 13 2 sleaf    l   1     2 sleaf l 3 2  23    l 2    4 sleaf l 2 3    l 13 sleaf 3    2 l 13 4 (3.21) The differentiation ∂sleaf3(l1)/dl1 and ∂sleaf3(l2)/dl2 depends on the domains of variables l1 and l2. It is necessary to classify domains. The additional formulas can be summarized as follows. (i) In case both (4m-1)π3/2≦l1≦(4m+1)π3/2 and (4k-1)π3/2≦l2≦(4k+1)π3/2 or both (4m+1)π3/2≦l1≦(4m+3)π3/2 and (4k+1)π3/2≦l2≦(4k+3)π3/2 (Variables m and k represent integers; for the constant π3, see [1][2].)  sleaf 3 2 l   l 1   sleaf  sleaf 41   2    3 3    l 1    l 1 4 3 2 6   l sleaf 3 1  sleaf 41  3   sleaf l  2 3     2 sleaf l       sleaf l l sleaf 1   3 2 3         4 2 sleaf l l sleaf 4  3 2 1  23       sleaf sleaf l l 2 1 3 3       2 sleaf l sleaf l 4  1  4 3 2 3 2 3 3  sleaf 1  3     sleaf l 1 2 2 6     l 1    l 4 2 3 (3.22) (ii) In case both (4m+1)π3/2≦l1≦(4m+3)π3/2 and (4k-1)π3/2≦l2≦(4k+1)π3/2 or both (4m-1)π3/2≦l1≦(4m+1)π3/2 and (4k+1)π3/2≦l2≦(4k+3)π3/2 (Variables m and k represent integers)    sleaf 1  3     sleaf l 1 2 2 6     l 1    l 4 2 3 (3.23)  sleaf 3 2 l   l 1   sleaf  sleaf 41   2    3 3    l 1    l 1 4 3 2 6   l sleaf 3 1  sleaf 41  3   sleaf l  2 3     2 sleaf l       sleaf l l sleaf 1   3 2 3         4 2 sleaf l l sleaf 4  3 2 1  23       sleaf sleaf l l 1 2 3 3       2 sleaf sleaf l l 4  1  4 3 2 3 2 3 3 3.2 Leaf function: cleaf3(l1+l2) With respect to arbitrary variables l1 and l2, the following equation is satisfied.  cleaf 3 2 l   l 1   sleaf 3  sleaf 41  3  2       l 1    l 2 4  l  2 sleaf     cleaf l 3  3 1 l   2         cleaf sleaf l l 41  1 2 3 3      3 cleaf sleaf cleaf l l  1 3 2 3 3         2 2 sleaf l l cleaf 4  3 1 3 2 4  sleaf 3  l 2  2  sleaf 4  3  23    l 2  cleaf 3   4 l 1 2   l 1 cleaf  3 l  1   2     cleaf 3  l 2 4    l 1 (3.24) The additional formulas of the leaf function cleaf3(l1 + l2) can be similarly proved as described in section 3.1. The differentials ∂cleaf3(l1)/dl1 and ∂cleaf3(l2)/dl2 depend on the domains of variables l1 and l2. It is necessary to classify domains. The additional formulas can be summarized as follows. (i) In case both 2kπ3≦l1≦(2k+1)π3 and (4m-1)π3/2≦l2≦(4m+1)π3/2 or (2k+1)π3≦l1≦(2k+2)π3 and (4m+1)π3/2≦l2≦(4m+3)π3/2 (Variables m and k represent integers)  cleaf 3 2 l   l 1   sleaf 3  sleaf 41  3  2    3    l 1    l 2 4    cleaf l 3 1  sleaf 41  3   cleaf l  3 2     2 cleaf l 3 1 6       l sleaf sleaf l 1   3 2 2 3         4 2 l cleaf l sleaf 4  3 1 2 3  23       sleaf cleaf l l 1 2 3 3      2 l sleaf cleaf 4  3 3   4 l 1 2 1  l 2  cleaf  3    2 cleaf 3 2 6     l 1    l 1 4 (3.25) (ii) In case both (2k+1)π3≦l1≦(2k+2)π3 and (4m-1)π3/2≦l2≦(4m+1)π3/2 or 2kπ3≦l1≦(2k+1)π3 and (4m+1)π3/2≦l2≦(4m+3)π3/2 (Variables m and k represent integers)  cleaf 3 2 l   l 1   sleaf  sleaf 41   3 2    3 3    l 1    l 2 4    cleaf l 3 1  sleaf 41  3   cleaf l  3 2     2 cleaf l 3 1 3 6       l sleaf sleaf l 1   3 2 2         4 2 l cleaf l sleaf 4  3 1 2  23       cleaf sleaf l l 1 2 3 3      2 sleaf l cleaf 4  3   4 l 1 3 2 3 1  l 2  cleaf  3    2 cleaf 3 2 6     l 1    l 1 4 (3.26) 4 Numerical example The numerical data of the leaf functions sleaf3(l) and cleaf3(l) are presented in Table 1. Using Eqs. (3.22), (3.23), (3.25), and (3.26), the values of sleaf3(l1+ l2) and cleaf3(l1+l2) can be calculated by using the values of sleaf3(l1), sleaf3(l2), cleaf3(l1), and cleaf3(l2). As an example, consider l1 = 0.2 and l2 = 0.3. 6 3 4 3 2 6 3 4 2     sleaf Then, sleaf3(0.2+0.3) can be obtained. We have to choose Eq. (3.22) for the inequalities -π3/2≦0.2≦π3/2 and -π3/2≦0.3≦π3/2. By substituting l1= 0.2 and l2 = 0.3 into Eq. (3.22), we get the following equation:              sleaf sleaf sleaf sleaf 2.0 13.0 3.0 12.0    3 3 3 3                 2 2 4 sleaf sleaf sleaf sleaf 3.0 2.0 41 4 2.0 3.0   3 3  23            sleaf sleaf sleaf 2.0 3.0 3.0 2.0  3 3 3                2 2 sleaf sleaf sleaf 3.0 4 3.0 2.0 2.0        199999 .0 299984 .0 299984 .0 1            4 2 2 299984 199999 .0 .04 .0 299984 199999      23       3 199999 299984 .0 299984 .0 .0            2 2 299984 .0 299984 .04 .0 199999        3.02.0   sleaf  sleaf 41  3   199999 .0  .041    199999 .0  199999 .041  (4.1) 2494431         .0  Therefore, the value of sleaf3(0.5) is obtained as follows: sleaf 3   3.02.0  49944 2494431 sleaf 3   5.0 .0 .0     (4.2) 3 4  4     2 6 4 3 1 3 3 6 3 4 Table 1 Numerical data of leaf functions sleaf3(l) and cleaf3(l) l sleaf3(l) 0.000000 0.100000 0.199999 0.299984 0.399883 0.499443 0.598009 0.694183 0.785387 0.867486 0.934768 0.980708 0.999692 0.989090 0.950393 0.888560 0.810064 0.720972 0.625896 0.527828 0.428461 cleaf3(l) 1.000000 0.985184 0.942810 0.878184 0.797825 0.707632 0.611979 0.513647 0.414176 0.314304 0.214324 0.114325 0.014325 -0.085670 -0.185670 -0.285660 -0.385580 -0.485220 -0.583990 -0.680640 -0.772770 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7 1.8 1.9 2.0 l sleaf3(l) 0.328621 0.228649 0.128651 0.028651 -0.071350 -0.171350 -0.271340 -0.371280 -0.470980 -0.569930 -0.667000 -0.759970 -0.845200 -0.917390 -0.970090 -0.997240 -0.995140 -0.964110 -0.908270 -0.833880 -0.747280 cleaf3(l) -0.856490 -0.926290 -0.975670 -0.998770 -0.992410 -0.957500 -0.898590 -0.822090 -0.734190 -0.639750 -0.541980 -0.442740 -0.342940 -0.242970 -0.142980 -0.042980 0.057024 0.157024 0.257019 0.356971 0.456727 2.1 2.2 2.3 2.4 2.5 2.6 2.7 2.8 2.9 3.0 3.1 3.2 3.3 3.4 3.5 3.6 3.7 3.8 3.9 4.0 4.1 5 Conclusions By using the analogy of the inverse trigonometric and lemniscate functions, higher order of functions were artificially created as leaf functions. The following conclusions can be drawn from this study: ・The artificially created leaf function r = sleaf3(l) corresponds to r = sin(l) and r = sl(l), whereas the artificially created leaf function r = cleaf3(l) corresponds to r = cos(l) and r = cl (l). The waves obtained from the trigonometric functions sin(l) and cos(l) have a periodicity of 6.28… and an amplitude of 1, while the waves obtained from the lemniscate functions have a periodicity of 5.24… and an amplitude of 1. The artificially created functions sleaf3(l) and cleaf3(l) also produce regular waves with 4.85 … periodicity and amplitude of 1. These functions produce waves different from those of both the trigonometric and lemniscate functions. The prefix "s" and "c" of these functions are classified by the following initial conditions: r(0) = 0, dr(0)/dl = 1 and r(0) = 1, dr(0)/dl = 0. ・Additional formulas of leaf functions sleaf3(l1 +l2) and cleaf3(l1 +l2) were also derived. ・The values of the leaf functions sleaf3(l1 +l2) can be numerically obtained using the values of sleaf3(l1), sleaf3(l2), while those of cleaf3(l1 +l2) can be obtained using values of sleaf3(l1), sleaf3(l2), cleaf3(l1), and cleaf3(l2) through the additional formulas. This study will be the first step in clarifying solution behaviors of nonlinear equations consisting of both the second-order differentials and the higher order polynomials. References [1] K. Shinohara, Special function: leaf function r=sleafn(l) (First report), Bulletin of Daido University, 51 (2015), 23–38. http://id.nii.ac.jp/1277/00000049/ [2] K. Shinohara, Special function: leaf function r=cleafn(l) (Second report), Bulletin of Daido University, 51 (2015), 39–68. http://id.nii.ac.jp/1277/00000050/ [3] Ivana Kovacic, Michael J. Brennan, The Duffing Equation: Nonlinear Oscillators and their Behaviour, John Wiley & Sons, 2011. https://doi.org/10.1002/9780470977859 [4] A. Beléndez, G. Bernabeu, J. Francés, D.I. Méndez, S. Marini, An accurate closed-form approximate solution for the quintic Duffing oscillator equation, Mathematical and Computer Modelling, 52 (2010), 637-641. https://doi.org/10.1016/j.mcm.2010.04.010 [5] A. Kargar, M. Akbarzade, Analytical solution of nonlinear cubic-quintic Duffing oscillator using global error minimization method, Advanced Studies in Theoretical Physics, 6 (2012), 467–471. [6] A. E. Zúñiga, Exact solution of the cubic-quintic Duffing oscillator, Applied Mathematical Modelling, 37 (2013), 2574-2579. https://doi.org/10.1016/j.apm.2012.04.005 [7] N. I. Akhiezer, Elements of the Theory of Elliptic Functions, (translations of mathematical monographs), Vol. 79, American Mathematical Society, 1990. [8] D. F. Lawden, Elliptic Functions and Applications, Applied Mathematical Sciences, Vol. 80, Springer, 1989. https://doi.org/10.1007/978-1-4757-3980-0 [9] C. H. Liang, Notes on the ellipsoidal function, Springer, 2016. [10] H. McKean and V. Moll, Elliptic Curves: Function Theory, Geometry, Arithmetic, Cambridge University Press, 1999. [11] R. Ranjan, Elliptic and Modular Functions from Gauss to Dedekind to Hecke, Cambridge University Press, 2017. https://doi.org/10.1017/9781316671504 [12] G.C. Fagnano, Produzioni Matematiche Del Marchese Giulio Carlo De' Toschi Di Fagnano, Pesaro Gavelli, 1750. [13] L. Euler, Leonhardi Euleri Opera Omnia Series (cid:31)-(cid:31), Birkhaeuser, 1911. [14] C. F. Gauss, Disquisitiones Arithmeticae, Yale University Press, 1965. Appendix A Let g(l1, l2) be a differentiable function. The necessary and sufficient condition for satisfying g(l1, l2) = g(l1+l2,0) is that holds. Function h(x,y) is     l l 2 2 ,  lg 1 l  1 defined as follows:  yxh ,  xg xy ,       lg 1 l  , 2 y (A.1) x y  x y 2 (A.2) l 1 l By differentiating the above equation with respect to y, we obtain the following equation.  yxh ,  y   xg  xy ,  l  1  (A.3) xy ,  l  (A.4)  xg   xg   xg  l  2 y        y y y y     2 l  1 y  xy ,  l  2 xy ,  l  1  Therefore, if the equation ∂g/∂l1=∂g/∂l2 holds, the following equation holds.  (A.5)   yxh  , y  0  , l  b  l 1  bxh  (A.6) From the above equation, we find that h(x,y) is a function of x, not a function of y. Therefore, the following equation holds for any constants a and b.  axh , Here, we set the following equation: x a From the above two equations, the following equation is obtained.  axh , (A.8) (A.9) (A.7)  2  2 l 1 l 1 l 1 l 1 l 1      l l l l l l l l , , 2 2 2 2 2  lg 1 , 2 2 2 2 1 lh     2  2    1 lg     2  2  2  2     bxh ,   l 2 l 1 , 1 lh     2 l 2  2    1 lg     2 l 2 l 1  l 2 l 1 , l 2 l 1   2  2 l 2  2     lg 1  l 2 0, (A.10) The following equation is obtained by using Eqs. (A.6), (A.9), and (A.10). , l 2    lg 1  l 2 0, (A.11)    xy ,  xg Conversely, if the above equation holds, the following relational expression can be obtained by using Eqs. (A.1) and (A.11).  yxh , We differentiate the above equation with y. Because g(2x, 0) is a function of x and not a function of y, the following equation is obtained. (A.12)  0,2 xg   y   yxh ,    xg 0,2  0   y   lg 1  y   yxh ,  y   xg    xg     y xy ,  l  1  y xy ,  l  1    xg    l  1 y  xy ,  l  2  xg   y  xy ,  l  2 l  2 y   y   0 Therefore, the following equation is obtained.   xg  xg    y y   xy ,  l  1    xy , l 2 2 l  l 1  2 Here, we use the following equation: x y Using Eq. (A.15), we obtain the following equation:  lg  1 l  1  lg  1 l   2 l 1     l l l , , 2 2 2 2 (A.13) (A.14) (A.15) (A.16) (A.17) (A.18) Appendix B The first term of the Eq. (3.15) is transformed as follows:  lp 1 1  lp 2 1    l l , , 2 2 sleaf 3  sleaf 4   l 2        l 1 3   l 1 sleaf  l  1 3 3  sleaf  l 2 3 2 4       l 2    l 1 sleaf  l  2 sleaf  l  1 3 3  sleaf   l 2 6   3   sleaf  l 2 3   1  sleaf  3    l 1  3  sleaf 6    l 1 3 sleaf  l 2 3  l  2   sleaf 4  l 2 4    sleaf 3    l 1 3 3  l 2  3 6 sleaf  l    l 1  sleaf  l  1 2 3 2 , l   lp  1 2 l  1   l      l 2 2 3 2 2   l 1 2 , l   lp  1 1 l  1 sleaf  l  1   l 1  sleaf 3 3  sleaf  l  2 sleaf  l  sleaf  l  1 3   l 1 3     l 2 6 2    l   l 1 3  1  3  sleaf 6    l 1 3 sleaf 3  sleaf    l 1 3 sleaf 3  sleaf   l 1 3 sleaf  l  1 3 2  sleaf  l  sleaf  l  1  sleaf 3 3   l 1    l 1 sleaf  l 2 6   3    3 sleaf  sleaf  sleaf   l 1  3  sleaf 4 3    l 1  l 3 2  l 2    4 sleaf 6    l 1 3 sleaf  l 2 3 3 4  sleaf    l  1   sleaf l   l  3 2 3 2 The following equation is applied [1] [2].   l 1  (B.1) sleaf 3   l  1   sleaf 3   6 l (B.2) d dl
1910.13285
1
1910
2019-10-29T14:24:31
Boundedness of operators on certain power-weighted Morrey spaces beyond the Muckenhoupt weights
[ "math.FA" ]
We prove that for operators satistying weighted inequalities with $A_p$ weights the boundedness on a certain class of Morrey spaces holds with weights of the form $|x|^\alpha w(x)$ for $w\in A_p$. In the case of power weights the shift with respect to the range of Muckenhoupt weights was observed by N.~Samko for the Hilbert transform, by H.~Tanaka for the Hardy-Littlewood maximal operator, and by S.~Nakamura and Y.~Sawano for Calder\'on-Zygmund operators and others. We extend the class of weights and establish the results in a very general setting, with applications to many operators. For weak type Morrey spaces, we obtain new estimates even for the Hardy-Littlewood maximal operator. Moreover, we prove the necessity of certain $A_q$ condition.
math.FA
math
BOUNDEDNESS OF OPERATORS ON CERTAIN POWER-WEIGHTED MORREY SPACES BEYOND THE MUCKENHOUPT WEIGHTS JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Abstract. We prove that for operators satistying weighted in- p weights the boundedness on a certain class of equalities with A Morrey spaces holds with weights of the form xαw(x) for w ∈ A p. In the case of power weights the shift with respect to the range of Muckenhoupt weights was observed by N. Samko for the Hilbert transform, by H. Tanaka for the Hardy-Littlewood maximal oper- ator, and by S. Nakamura and Y. Sawano for Calder´on-Zygmund operators and others. We extend the class of weights and establish the results in a very general setting, with applications to many operators. For weak type Morrey spaces, we obtain new estimates even for the Hardy-Littlewood maximal operator. Moreover, we prove the necessity of certain A q condition. 1. Introduction For 1 ≤ p < ∞ and 0 ≤ λ < n, let the Morrey space Lp,λ(w) be the collection of all measurable functions f such that (1.1) kf kLp,λ(w) := sup x∈Rn,r>0(cid:18) 1 rλZB(x,r) f pw(cid:19)1/p < ∞. We also consider the weak Morrey space W Lp,λ(w), for which kf kW Lp,λ(w) := x∈Rn,r>0,t>0(cid:18)tp w({y ∈ B(x, r) : f (y) > t}) sup rλ (cid:19)1/p < ∞. (Here and in what follows w(A) stands for the integral of w over A.) Clearly, Lp,λ(w) ⊂ W Lp,λ(w). N. Samko proved in [7] that the Hilbert transform is a bounded operator on Lp,λ(xα) for 0 < λ < 1 and λ − 1 < α < λ + p − 1. This 2010 Mathematics Subject Classification. Primary . Key words and phrases. Morrey spaces, Muckenhoupt weights, Hardy- Littlewood maximal operator, Calder´on-Zygmund operators. The first author is supported by the grants MTM2014-53850-P of the Minis- terio de Econom´ıa y Competitividad (Spain) and grant IT-641-13 of the Basque Gouvernment. 1 2 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL range of values of α shows a shift with respect to the corresponding range in the Ap class, which is −1 < α < p − 1. In [10], H. Tanaka explored the boundedness on Lp,λ(w) of the Hardy-Littlewood maximal operator and was able to describe necessary conditions and sufficient conditions, but not a characterization. Nevertheless, for power weights w(x) = xα he obtained the sharp range λ − n ≤ α < λ + n(p − 1), which in the one-dimensional case coincides with the range obtained by Samko for the Hilbert transform except at the endpoint α = λ − n. Later on, S. Nakamura and Y. Sawano in [6] studied the boundedness of the Riesz transforms and other singular integrals and obtained similar shifted ranges for the case of Lp,λ(xα) (with open left endpoint). In [3] the authors of this paper proved a general result involving Muckenhoupt weights, under the assumptions of the extrapolation the- orem for Ap weights. When particularized for the Hardy-Littlewood maximal operator or for Calder´on-Zygmund operators, the bounded- ness on Lp,λ(w) was obtained for w ∈ Ap ∩ RHσ in the range 0 ≤ λ ≤ n/σ′, which for w(x) = xα gives the range λ − n ≤ α < n(p − 1). In this paper the results in [3] for Lp,λ(w) spaces are extended to the weights xαw(x) for w ∈ Ap ∩ RHσ and 0 < α < λ. (We recall the definitions of the weight classes Ap and the reverse Holder classes RHσ in Section 2.) In particular, we prove the following theorem. Theorem 1.1. 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 (1.2) kgkLp0 (w) ≤ Ckf kLp0 (w), where C does not depend on the pair (f, g) and it depends on w only (defined at the beginning of Section 2). Then for in terms of [w]Ap0 1 < p < ∞ and w ∈ Ap ∩ RHσ it holds (1.3) kgkLp,λ(xαw) ≤ Ckf kLp,λ(xαw), for 0 ≤ λ < n/σ′ and 0 ≤ α < λ. In particular, for power weights of the form xβ = xαw(x), the estimate (1.3) holds for λ − n < β < λ + n(p − 1). If the hypothesis holds for p0 = 1, then (1.3) also holds for p = 1, and in the case of power weights for λ − n < β < λ. When we say that (1.2) holds for every w ∈ Ap0 we mean that if the right-hand side is finite for a fixed w, then also the left-hand side is finite for the same w and the inequality holds. The conclusion of the if f is in Lp,λ(xαw), theorem is to be understood in the same way: then g is in the same space and the inequality holds. BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 3 To make things clear let us say that the weights appearing in the the- orem are always in some Muckenhoupt class. Indeed, xαw ∈ Ap+λ/n for 0 ≤ α < λ and w ∈ Ap. Being in Ap+λ/n is not a particular restric- tion for the weights in our theorem, because we show in Proposition 5.2 the necessity of u ∈ Ap+λ/n for the boundedness of the Hardy- Littlewood maximal operator in the Morrey space Lp,λ(u) (and even for the weak-type boundedness). We are thus forced to deal with Muck- enhoupt weights. When we say that the results go beyond the Muck- enhoupt range, we mean that for a fixed value of p the boundedness of the involved operators holds for weights which are not necessarily in Ap. This theorem has a number of applications because we know that for many operators T , the pairs (f , T f ) satisfy its assumptions. In particular, we recover the results for the Hardy-Littlewood maximal operator (except the left endpoint, that is, β = λ − n), the Hilbert transform and the Calder´on-Zygmund operators mentioned above. But it extends also to Littlewood-Paley operators, rough singular integrals and others. Moreover, in those examples the case p = 1 of the theorem provides a weak-type result, from L1,λ(xβ) to W L1,λ(xβ) for λ − n < β < λ, which for β > 0 is new even for the Hardy-Littewood maximal operator. We present some preliminary results in Section 2. The proof of The- orem 1.1 is in Section 3, where we also prove another theorem suited to operators satisfying the assumptions of the so-called limited range extrapolation. In Section 4 we establish embeddings which allow to de- fine the operators in the Morrey spaces by restriction. In Section 5 we prove the necessity of the Ap+λ/n condition for the Hardy-Littlewood maximal operator and of Ap+λ for the Hilbert transform in the case of weak-type estimates. This implies the necessity of the range of power weights for positive exponents. We also give an easy proof of the ne- cessity for negative exponents. In the case of the strong estimates and power weights this was proved by Tanaka in [10] checking his more general necessary condition. We extend the necessity to the weak-type estimates. In the same section we prove the estimate for the left end- point for power weights (that is, for the weight xλ−n). 2. Preliminary results Let w ∈ Lloc 1 (Rn) with w > 0 almost everywhere. We say that w is a Muckenhoupt weight belonging to Ap for 1 < p < ∞ if [w]Ap ≡ sup B w(B) B (cid:18)w1−p′(B) B (cid:19)p−1 < ∞, 4 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL where the supremum is taken over all Euclidean balls B in Rn. The quantity [w]Ap is the Ap constant of w. We say that w belongs to A1 if, for any Euclidean ball B, w(B) B ≤ cw(x) for almost all x ∈ B. The A1 constant of w, denoted by [w]A1, is the smallest constant c for which the inequality holds. We say that a nonnegative locally integrable function w on Rn be- longs to the reverse Holder class RHσ for 1 < σ < ∞ if it satisfies the reverse Holder inequality with exponent σ, i.e., (cid:18) 1 BZB σ w(x)σdx(cid:19) 1 ≤ c BZB w(x)dx, where the constant c is universal for all Euclidean balls B ⊂ Rn. Remark 2.1. Some results for weights are the following. We use repeat- edly the first two throughout the paper. (i) If Mh < ∞ a.e., then (Mh)1/s ∈ A1 and its A1 constant de- pends on s, but not on h. Moreover, (Mh)1/s ∈ A1 ∩ RHσ if s > σ. (ii) Let w ∈ RHσ. For any ball B and any measurable E ⊂ B it holds that (2.1) w(E) w(B) ≤ c(cid:18)E B(cid:19)1/σ′ . Since w ∈ Ap implies that w ∈ RHσ for some σ, the inequality holds for each Ap weight for the appropriate σ. (iii) Weights simultaneously in Ap and RHσ can be described ([4]) as (2.2) Ap ∩ RHσ = {w : wσ ∈ Aσ(p−1)+1}. Remark 2.2. When dealing with the definition of the norm (1.1), we only need to take into account two types of balls: balls centered at the origin and balls of the form B(x, r) with r < x/4. Indeed, if we have a ball B(x, r) with r ≥ x/4, it holds that B(x, r) ⊂ B(0, 5r), and since the radii are comparable we can replace the smaller ball by the larger one. The following lemma provides an estimate which is used in the proofs of the theorems. BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 5 Lemma 2.3. Let 1 ≤ p < ∞ and 0 ≤ λ < n. Let f ≥ 0 in Lp,λ(xαw) and r > 0. If α < λ, then (2.3) (cid:18)ZB(0,r) p f pw(cid:19) 1 ≤ Cr λ−α p kf kLp,λ(xαw). The constant depends only on α, λ and p. Proof. Let Aj = B(0, 2−j+1r) \ B(0, 2−jr), j ∈ N. Then ZB(0,r) f pw ≤ C ≤ C ∞Xj=1ZAj ∞Xj=1 f (y)pw(y)(cid:18) y 2−jr(cid:19)α dy (2−jr)λ−αkf kp Lp,λ(xαw). If α < λ the series is convergent and (2.3) follows. (cid:3) Remark 2.4. We deal with integrals of the typeRB f pw where w is a certain Ap weight and B is a ball. This can be written asR f pwχB. But wχB cannot be an Ap weight for any p because Ap-weights cannot vanish in a set of positive measure. Hence, the second proof of Theorem 6.1 of [1] is not correct, because it claims that rλ−nχB(x,r) is an A1 weight. 3. Main theorems In this section we first prove Theorem 1.1, starting with the assump- tion for p0 = 1. This case is important because the proof of the theorem is simpler and the general case p0 ∈ (1, ∞) can be reduced to this one by a scaling argument. In the applications it is also significant because for a number of operators weighted weak-type (1, 1) estimates are known and our result provides weak-type Morrey estimates for them. Proof of Theorem 1.1. Case p0 = 1. We assume first that (1.2) holds with p0 = 1. Let Br := B(x, r) be one of the balls considered in Remark 2.2. Let w ∈ Ap. We have (3.1) ZBr g(y)pyαw(y)dy ≤ gpw, rαZBr CxαZBr if x = 0; gpw, if 0 < r ≤ x/4. 6 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Proof for 1 < p < ∞. In both cases we are left with the integral of gpw on the ball Br, which we handle as in [3]. Using duality we have p gpw(cid:19) 1 (cid:18)ZBr ghw ≤ZRn ZBr = sup Lp′ (w,Br ) h : khk ghw. =1ZBr s ≤ cZRn 1 Fix such a function h and we have for s > 1 that (3.2) gM(hswsχBr ) f M(hswsχBr ) 1 s , 1 provided that M(hswsχBr ) s ∈ A1. According to Remark 2.1 (i) we need M(hswsχBr )(x) < ∞ a.e. We check that hswsχBr ∈ L1 for ap- propriate s > 1, and get a bound for its integral for future use. To this end, we choose s > 1 such that w1−p′ ∈ Ap′/s, which is possible because w1−p′ ∈ Ap′. We have (cid:18)ZBr s hsws−1w(cid:19) 1 ≤(cid:18)ZBr (3.3) hp′ w(cid:19) 1 p′(cid:18)ZBr (Br)− 1 w(s−1) p′ p′−s s − 1 p′ +1(cid:19) 1 ≤ c Br 1 s w1−p′ p′ ≤ c w(Br) 1 p r− n 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 cnrn = Br ≤ w(Br) 1 p w1−p′ (Br) 1 p′ . We split the last integral of (3.2) into the integral over B2r and over its complement. On the one side we have ZB2r f M(hswsχBr ) 1 s ≤(cid:18)ZB2r f pw(cid:19) 1 p(cid:18)ZB2r M(hswsχBr ) p′ p′ s w1−p′(cid:19) 1 . The last term is bounded by a constant because M is bounded on Lp′/s(w1−p′) and we get a constant times the norm of h in Lp′(w, Br), which is 1. Now we have f pw ≤ (3.4) ZB2r C(2r)λ−αkf kp Lp,λ(xαw), Cx−α(2r)λkf kp Lp,λ(xαw), if x = 0; if 0 < r ≤ x/4; where in the first case we use (2.3), and in the second case we use x ∼ y for y ∈ B2r. To deal with the integral on Rn \ B2r we decompose it into annuli and use that on B2j+1r \ B2j r the maximal operator is comparable to BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 7 (2jr)−nRBr hsws. Using (3.3) we get ∞Xj=1ZB2j+1 r\B2j r 1 p w(Br) 2jn/srn f (3.5) ≤ ≤ f pw! 1 ∞Xj=1 ZB2j+1 r\B2j r f pw! 1 ∞Xj=1 ZB2j+1 r p jn s′ 2− jn pσ′ , 2 p w1−p′ (B2j+1r) 2jn/srn 1 p′ w(Br) 1 p where in the last step we use w (B2j+1r) 1 p w1−p′ (B2j+1r) 1 p′ ≤ CB2j+1r = C ′2jnrn, and (2.1) for Br and B2j+1r. If Br is centered at the origin, we use (2.3) in the last term of (3.5) to get (3.6) C ∞Xj=1 2j[(λ−α− n σ′ ) 1 p + n s′ ]r(λ−α) 1 p kf kLp,λ(xαw). The series is convergent if we choose s close enough to 1, because we assume λ < n/σ′. If the ball is centered at x and r ≤ x/4, we can assume x = 2N r for some N ≥ 2, increasing slightly r if necessary. For the integral in (3.5) we distinguish the cases j ≤ N −2 and j ≥ N −1. In the first case, if y ∈ B2j+1r, then y ∼ x. In the second case, B2j+1r ⊂ B(0, 2j+2r). As a consequence, for j ≤ N − 2, ZB2j+1 r f pw ≤ Cx−αZB2j+1 r f (y)pw(y)yαdy ≤ Cx−α(2jr)λkf kp Lp,λ(xαw). For j ≥ N − 1 we use (2.3) and we obtain f pw ≤ C(2jr)λ−αkf kp Lp,λ(xαw). f pw ≤ZB(0,2j+2r) ZB2j+1 r kf kLp,λ(xαw) N −2Xj=1 Inserting this into (3.5) we obtain a constant times x− α p r λ p 2j[(λ− n σ′ ) 1 p + n s′ ] + r(λ−α) 1 p 2j[(λ−α− n σ′ ) 1 p + n s′ ]! . ∞XN −1 8 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL The first sum is bounded independently of N if we take s close enough to 1, because λ < n/σ′. For this same reason the series is convergent and the value of its sum is s′ ]r(λ−α) 1 s′ ] ≤ C ′x− α p r p = Cx− α p r C2N [(λ−α− n σ′ ) 1 p + n λ p 2N [(λ− n σ′ ) 1 p + n λ p , where C ′ is independent of N because the exponent of 2N in the middle term is negative. Taking into account (3.4), (3.6) and the recent bound, the right-hand side of (3.1) is bounded as desired. Proof for p = 1. The proof is similar but easier because we do not need to use a duality argument and there is no h as in the previous situation. We have w ∈ A1 ∩ RHσ (which is the same as wσ ∈ A1). In the construction of the A1 weight M(wsχBr )1/s we choose 1 < s < σ. Then M(wsχBr )1/s ≤ M(ws)1/s ≤ w a.e. and when we integrate on B2r we obtain (3.4) with p = 1. When we are in B2j+1r \ B2j r we have M(wsχBr )1/s ≤ C(cid:18)ws(Br) (2jr)n(cid:19)1/s . Using that w ∈ RHs (because s < σ), ws(Br)1/s ≤ CBr−1/s′w(Br). Instead of (3.5) we have now ∞Xj=1ZB2j+1 r\B2j r f w 2jn/s′ w(Br) w(B2j+1r) ≤ ∞Xj=1ZB2j+1 r\B2j r The proof continues as before. f w 2j( n s′ − n σ′ ). Case p0 > 1. By the usual extrapolation theorem the assumption (1.2) is valid for any p0 ∈ (1, ∞). Given p > 1 and w ∈ Ap ∩ RHσ, we choose p0 > 1 for which w ∈ Ap/p0. The assumption holds in the form kgp0kL1(v) ≤ Ckf p0kL1(v), for v ∈ A1 ⊂ Ap0 and we can apply the previous part of the proof to the pair (f p0, gp0) to get the Morrey estimate with exponent p/p0. (cid:3) We can generalize Theorem 1.1 to a setting in which weighted in- equalities is a restricted range are assumed. Theorem 3.1. Let 1 < b < ∞ and 1 ≤ p0 < b. Let F be a collection of nonnegative measurable pairs of functions. Assume that for every (f, g) ∈ F and every w ∈ Ap0 ∩ RH(b/p0)′ we have (3.7) kgkLp0 (w) ≤ Ckf kLp0 (w), BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 9 where C does not depend on the pair (f, g) and it depends on w only in terms of the Ap0 and RH(b/p0)′ constants of w. Then if 1 < p < b and w ∈ Ap ∩ RHσ with σ > (b/p)′, it holds that kgkLp,λ(xαw) ≤ Ckf kLp,λ(xαw), for 0 ≤ λ < n( 1 of the form xβ = xαw(x) and for σ′ − p b ) and 0 ≤ α < λ. In particular, for power weights it holds that λ − n(cid:16)1 − p b(cid:17) < β < λ + n(p − 1), kgkLp,λ(xβ) ≤ Ckf kLp,λ(xβ). Moreover, if the hypothesis holds for p0 = 1, then the results are valid for p = 1. Proof. The proof is similar to that of Theorem 1.1. Starting with p0 = 1, to apply the hypothesis in (3.2) we need to assume s > b′ because the weight has to be in A1 ∩RHb′. The estimate (3.3) and the boundedness of M in Lp′/s(w1−p′) need w1−p′ ∈ Ap′/s. We are assuming that w ∈ Ap∩RHσ with σ > (b/p)′, in particular, w ∈ Ap∩RH(b/p)′. According to (2.2) this is the same as saying w(b/p)′ ∈ A(b/p)′(p−1)+1, which by duality yields w(b/p)′(1−q′) ∈ Aq′ with q = (b/p)′(p−1)+1, that is, w1−p′ ∈ Ap′/b′. Then there exists s > b′ for which w1−p′ ∈ Ap′/s as needed. The proof continues as before, and we only need to add the condition that makes the series convergent. This condition is λ < n( 1 If we assume p0 > 1 in (3.7), by the usual extrapolation theorem we can consider any p0 ∈ (1, b). Given p and w we proceed again as before by choosing p0 close enough to 1 such that w ∈ Ap/p0 and working with the pairs (f p0, gp0). (cid:3) σ′ − p b ). The formulation of the extrapolation theorem in terms of pairs of functions provides several extensions as corollaries (see [2, p. 21 -- 22]). In a similar way, we can get similar extensions in the Morrey set- ting. We state the scaling and weak-type extensions in the following two corollaries, and leave to the interested reader the extension to the vector-valued setting. Corollary 3.2. Let 0 < p− ≤ p0 < p+ ≤ ∞. Let F be a collection of nonnegative measurable pairs of functions. Assume that for every (f, g) ∈ F and every w ∈ A p0 p− ∩ RH(cid:16) p+ p0 (cid:17)′ we have (3.8) kgkLp0 (w) ≤ Ckf kLp0 (w), 10 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL where C does not depend on the pair (f, g) and it depends on w only in terms of the A p0 p0 (cid:17)′ constants of w. Then if p− < p < p+ p− and RH(cid:16) p+ and w ∈ A p p− ∩ RHσ with σ >(cid:16) p+ p(cid:17)′ , it holds that kgkLp,λ(xαw) ≤ Ckf kLp,λ(xαw), for 0 ≤ λ < n( 1 of the form xβ = xαw(x) and for σ′ − p p+ ) and 0 ≤ α < λ. In particular, for power weights λ − n(cid:18)1 − p p+(cid:19) < β < λ + n( p p− − 1), kgkLp,λ(xβ) ≤ Ckf kLp,λ(xβ). it holds that Moreover, if the hypothesis holds for p0 = p−, then the conclusion is valid for p = p−. If p+ = ∞ this is a corollary to Theorem 1.1 and if p+ < ∞ to Theorem 3.1. The proof is immediate from the theorems if we read (3.8) as kgp−kLfp0 (w) ≤ Ckf p−kLfp0 (w), and b = p+ p− . for every w ∈ Aep0 ∩ RH(b/ ep0)′ with ep0 = p0 Corollary 3.3. If in Theorem 1.1, Theorem 3.1 or Corollary 3.2 the assumptions hold as weak-type inequalities, that is, with kgkLp0,∞(w) in- stead kgkLp0(w), then the conclusions also hold in the weak sense, that is, with kgkW Lp,λ(xαw) instead of kgkLp,λ(xαw). of p− To prove this case, the weak-type hypothesis can be read as a strong type inequality for the pair (f, tχ{g>t}) with constants uniform in t. 4. Embeddings and applications The proof of Theorem 1.1 shows that ZRn f M(hswsχBr ) 1 s ≤ Ckf kLp,λ(xαw), for 0 ≤ λ < n, w ∈ Ap and appropriate s > 1. This implies the continuous embedding Lp,λ(xαw) ֒→ L1(M(hswsχBr ) 1 s ). BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 11 In particular, choosing the ball B(0, 1) and h = cw−1, we have Lp,λ(xαw) ֒→ L1((1 + x)−β) for some β < n. Since M(hswsχBr ) s ∈ A1, we have 1 Lp,λ(xαw) ⊂ [u∈A1 L1(u). By the scaling argument at the end of the proof of the same theorem, if p > 1, (4.1) Lq(u), q > 1. Lp,λ(xαw) ⊂ [u∈Aq (The right-hand side is independent of q; see [5] and [3].) Suitable em- beddings can be written for the weighted spaces appearing in Theorem 3.1, hence in Corollary 3.2. The applications to boundedness of operators in Morrey spaces are corollaries of the general theorems and of the embeddings just men- tioned. For instance, the basic one is the following. Corollary 4.1. Assume that for some p0 ∈ [1, ∞), T is an operator Lp0(w) into the space of measurable functions and acting fromSw∈Ap0 satisfying kT f kLp0 (w) ≤ Ckf kLp0 (w) (4.2) for all f ∈ Lp0(w) and w ∈ Ap0, with a constant depending on [w]Ap0 . Then for every 1 < p < ∞ (and also p = 1 if p0 = 1), w ∈ Ap ∩ RHσ, 0 ≤ λ < n/σ′ and 0 < α < λ, we have that T is well defined on Lp,λ(xαw) by restriction and, moreover, (4.3) kT f kLp,λ(xαw) ≤ Ckf kLp,λ(xαw). For power weights xβ, T is well defined and bounded on Lp,λ(xβ) if λ − n < β < λ + n(p − 1). If (4.2) is replaced by the weak estimate from L1(w) to L1,∞(w) for w ∈ A1, then (4.3) holds from L1,λ(xαw) to W L1,λ(xαw). The definition by embedding is guaranteed by (4.1) and the size estimate by Theorem 1.1. There are many operators satisfying the assumptions of the theo- rem: the Hardy-Littlewood maximal operator, Calder´on-Zygmund op- erators, rough operators with kernel x−nΩ(x/x) with Ω ∈ L∞(Sn−1) and integral zero, commutators (in this case the weighted weak-type (1, 1) does not hold), square functions (including some Littlewood- Paley type operators, Lusin area integral, gλ functions, Marcinkiewicz integral), etc. 12 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL Similar corollaries can be written for our other general results. All the applications mentioned in [3] for the spaces Lp,λ(w) are now ex- tended to Lp,λ(xαw) with 0 ≤ α < λ by the theorems in this paper. Note that in [3] the space Lp,ν(w) was denoted as Lr p(λ, w, Rn) with n + rp = ν. 5. Necessary conditions for M and H, and the endpoint for M Tanaka proved in [10] that M is bounded on Lp(xβ) if and only if λ − n ≤ β < λ + n(p − 1). The necessity of the upper bound means that 0 ≤ α < λ in Theorem 1.1 is optimal. Tanaka's proof uses a general necessary condition involving duality in Morrey spaces. Avoiding duality, we prove first a necessary condition in terms of the Aq scale from which the necessity of the upper bound follows, and next the necessity of the lower bound is proved in a direct way. In all cases our necessary conditions are valid for the weak-type (p, p) estimates. The sufficiency for λ − n < β < λ + n(p − 1) comes from Theorem 1.1. We give in Proposition 5.7 below a direct proof of the boundedness at the endpoint Lp,λ(xλ−n) for 1 < p < ∞ and the corresponding weak estimate for p = 1. The weak estimates are not in [10]. pn Lemma 5.1. Let 1 ≤ p < ∞ and 0 ≤ λ < n. The embedding n−λ ) ֒→ Lp,λ(w) holds with constant depending only on n, λ L and p, not on w. n−λ (w n Proof. Let B be a ball of radius r. Then 1 rλZB f pw ≤ f pn n−λ w 1 rλ(cid:18)ZB n n n−λ(cid:19)1− λ B λ n ≤ cn,λkf kp L pn n−λ (w n n−λ ) . (cid:3) Proposition 5.2. Let 1 ≤ p < ∞ and 0 ≤ λ < n. If M is bounded from Lp,λ(w) to W Lp,λ(w), then w ∈ Ap+λ/n. Proof. Let B be a ball of radius r. Define f = σχB with σ nonnegative to be chosen later. For x ∈ B, we have Mf (x) ≥ σ(B)/B. If t < σ(B)/B, then B = {x ∈ B : Mf (x) > t}. Assuming that M is bounded from Lp,λ(w) to W Lp,λ(w) we have tw(B)1/p rλ/p ≤ CkσχBkLp,λ(w) ≤ C ′kσχBk pn n−λ (w) L = C ′(cid:18)ZB pn n−λ w σ n pn n−λ(cid:19) n−λ , where we used Lemma 5.1 in the second inequality. Let t tend to σ(B)/B and choose σ such that σ = σ n = w. n−λ , that is, σ1−p− λ n−λ w pn n ZB(0,δ) xβdx ∼ δβ+n, BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 13 We get w(B)σ(B)p+ λ n −1 Bp+ λ n ≤ C, with a constant independent of B. Therefore, w ∈ Ap+λ/n. To be precise, we do not know a priori that σ(B) is finite for the n = w. As usual, to overcome this problem, we define σǫ (cid:3) = w + ǫ for ǫ > 0 and let ǫ tend to 0. choice σ1−p− λ 1−p− λ by σ n ǫ Proposition 5.3. Let 1 ≤ p < ∞ and 0 ≤ λ < n. If M is bounded from Lp,λ(xβ) to W Lp,λ(xβ), then λ − n ≤ β < λ + n(p − 1). Proof. According to the previous proposition, xβ ∈ Ap+λ/n and this requires β < λ + n(p − 1). For the lower bound, first we observe that if the characteristic func- tion of a ball centered at the origin is in W Lp,λ(xβ), then λ − n ≤ β. Indeed, let δ be small and t < 1. Then for every x ∈ B(0, δ), the function is bigger than t at x. Since we want δβ+n ≤ Kδλ for small δ. Therefore, β + n ≥ λ is necessary. Let f be the characteristic function of the ball centered at 1 with radius 1/2. This function is clearly in Lp,λ(xβ) for any β and for 0 ≤ λ < n. The maximal operator acting on f satisfies Mf (x) ≥ c for some c > 0 and x ∈ B(0, 1). Consequently, Mf /∈ W Lp,λ(xβ) for β < λ − n. (cid:3) A direct proof of the necessity of β < λ + n(p − 1) is obtained as follows. For β ≥ λ + n(p − 1), the function x−nχB(0,1) is in Lp,λ(xβ) and is not locally integrable. Therefore, β < λ + n(p − 1) is necessary. Proposition 5.4. Let 1 ≤ p < ∞ and 0 ≤ λ < 1. transform H is bounded from Lp,λ(w) to W Lp,λ(w), then w ∈ Ap+λ. If the Hilbert Proof. First we observe that for characteristic functions of sets the norms in Lp,λ(w) and W Lp,λ(w) are the same. Given an interval I, let I ′ be the adjacent interval of the same length, placed at the right of I. Note that for x ∈ I ′, H(χI)(x) > 1/(2π). Assuming the weak boundedness of H we have kχI ′kW Lp,λ(w) ≤ k2πH(χI)χI ′kW Lp,λ(w) ≤ 2πkH(χI)kW Lp,λ(w) ≤ CkχIkLp,λ(w). Interchanging the role of I and I ′ we deduce that kχIkW Lp,λ(w) and kχI ′kW Lp,λ(w) are comparable. ≤ Ck2πH(σχI)χI ′kW Lp,λ(w) ≤ 2πCkH(σχI)kW Lp,λ(w) ≤ C ′kσχIkLp,λ(w). σ(I) I w(I)1/p Iλ/p ≤ C(cid:13)(cid:13)(cid:13)(cid:13) Using Lemma 5.1 we get σ(I) I χI ′(cid:13)(cid:13)(cid:13)(cid:13)W Lp,λ(w) Iλ/p ≤ C ′′(cid:18)ZI w(I)1/p σ(I) I 14 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL For σ to be chosen later, we observe that H(σχI)(x) > σ(I)(2πI)−1 for x ∈ I ′. Also w(I) Iλ ≤ kχIkp W Lp,λ(w) ≤ CkχI ′kp W Lp,λ(w), where the first inequality holds by the definition of the norm. Then p 1−λ w σ p 1 1−λ(cid:19) 1−λ p , 1 1−λ . (cid:3) from which w ∈ Ap+λ follows if we choose σ = σ 1−λ w Remark 5.5. In the case of the Hilbert transform one could prefer to assume that it is defined a priori only for Schwartz functions through the principal value formula. The proof given here can be adapted to such assumption by approximating the involved functions by smooth ones. Without affecting the proof one can take the intervals I and I ′ separated by a distance equal to their length, instead of taking them adjacent, so that there is some room for the approximation. The proof of Proposition 5.4 can be adapted to higher dimensions to obtain the necessity of the condition Ap+λ/n for the Riesz transforms and other singular integral operators satisfying a nondegeneracy con- dition (see [9, Chapter V, §4.6] for the similar result in the Lebesgue setting). Proposition 5.6. Let 1 ≤ p < ∞ and 0 ≤ λ < 1. If H is bounded from Lp,λ(xβ) to W Lp,λ(xβ), then λ − 1 < β < λ + p − 1. Proof. The condition β < λ + p − 1 is a consequence of xβ ∈ Ap+λ as required by the previous proposition. As in the proof for the maximal operator in Proposition 5.3, the estimate H(χ1,2)(x) ≥ 1/(2π) for x ∈ (0, 1) is enough to get β ≥ λ − 1. To rule out the Morrey estimate for β ≥ λ − 1 we consider χ(0,1), which is in Lp,λ(xλ−1). A direct computation shows that Hχ(0,1)(x) = π−1 log(x/x − 1). Then Hχ(0,1)(x) ≥ c(− log x)χ(0,1/4)(x), and the last function is not in W Lp,λ(xλ−1). (cid:3) The result in this proposition was proved for the strong estimates (hence, 1 < p < ∞) by N. Samko in [7, Theorem 4.7], and in [8] she BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 15 discussed the necessity of a more general condition. Our result gives also the necessity for the weak estimates (1 ≤ p < ∞). In the next proposition we give a direct proof of the boundedness of M for the Morrey spaces with weight xλ−n (the endpoint of the allowed range). In [10] this is a consequence of a certain sufficient condition involving duality. Proposition 5.7. Let 0 ≤ λ < n. M is bounded on Lp,λ(xλ−n) for 1 < p < ∞ and from L1,λ(xλ−n) to W L1,λ(xλ−n). Proof. Let f ∈ Lp,λ(xλ−n) with λ > 0. Assume that it is nonnegative. Consider the ball Br := B(x, r). Decompose f as f1 + f2, where f1 = f χB2r . Using the subadditivity of M we have Mf (y) ≤ Mf1(y) + Mf2(y). Using the boundedness of M on Lp(xλ−n) we have ZBr (Mf1)(y)pyλ−ndy ≤ C1ZB2r f (y)pyλ−ndy ≤ C2rλkf kp Lp,λ(xλ−n). On the other hand, Mf2(y) is almost constant on Br in the sense that Mf2(y1) ∼ Mf2(y2) ∼ Mf2(x) for y1, y2 ∈ Br. Then (5.1) Moreover, ZBr (Mf2)(y)pyλ−ndy ≤ C1(Mf2)(x)pZBr yλ−ndy. Mf2(x) ∼ sup R>2r 1 RnZBR\B2r f. We distinguish two types of balls as in Remark 2.2. In the case of a ball centered at 0, we have 1 RnZBR\B2r f ≤ 1 Rn(cid:18)ZBR f p(cid:19)1/p BR1/p′ ≤ Ckf kLp,λ(xλ−n), using (2.3) with w ≡ 1 and α = λ − n. Since the last integral in (5.1) is Crλ we get the desired estimate. For a ball centered at x 6= 0 with radius r ≤ x/4 we consider first R ≥ x/2. In such case, 1 RnZBR\B2r f ≤ 1 RnZB(0,3R) f ≤ Ckf kLp,λ(xλ−n). 16 JAVIER DUOANDIKOETXEA AND MARCEL ROSENTHAL The last integral in (5.1) is Cxλ−nrn and this is bounded by Crλ because r < x and λ − n is negative. Let now 2r < R < x/2. Then 1 RnZBR\B2r f ≤ 1 Rn(cid:18)ZBR f p(cid:19)1/p Rn/p′ ≤ Cx n−λ p R λ−n p kf kLp,λ(xλ−n). Replacing the last integral in (5.1) by Cxλ−nrn the needed estimate holds because Rλ−nrn ≤ rλ due to R > r. The proof of the weak type for p = 1 is similar. (cid:3) The range λ − n < β < λ + n(p − 1) for weights of type xβ corre- sponds to all the power weights in Ap+λ/n ∩ RHn/(n−λ). On the other hand, the result in Theorem 1.1 is valid for λ < n/σ′, that is, we need w ∈ RHσ for some σ > n/(n − λ) to get the estimate. Such a σ exists for any weight in RHn/(n−λ) by the self-improvement property of the reverse Holder inequalities (Gehring's lemma). The endpoint weight xλ−n, for which the estimates for M hold, is not in RHn/(n−λ), but it is in RHσ for every σ < n/(n − λ). One could guess the necessity of a reverse Holder condition of this type in Proposition 5.2, but we have not been able to get it. The weights in Ap+λ/n ∩ RHn/(n−λ) are characterized by the factor- ization u−λ/nw with u ∈ A1 and w ∈ Ap ∩ RHn/(n−λ). The result in Theorem 1.1 covers all the weights of this type for which u is a power weight in A1. By translation invariance power weights can be taken to be centered at a point different from the origin. The sufficiency of Ap+λ/n ∩ RHn/(n−λ) in Theorem 1.1 remains an open question for us. References [1] Adams, D. R.: Morrey spaces, Lecture Notes in Applied and Numerical Har- monic Analysis. Birkhauser/Springer, Cham (2015). [2] 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, Birkhauser, Basel (2011). [3] Duoandikoetxea, J., Rosenthal, M.: Extension and boundedness of operators on Morrey spaces from extrapolation techniques and embeddings. J. Geom. Anal. 28, 3081 -- 3108 (2018). [4] Johnson, R., Neugebauer, C. J.: Change of variable results for A p and reverse Holder RH r-classes. Trans. Amer. Math. Soc. 328, 639 -- 666 (1991). [5] Knese, G., McCarthy, J. E., Moen, K.: Unions of Lebesgue spaces and A1 majorants. Pacific J. Math. 280, 411 -- 432 (2016). [6] Nakamura, S., Sawano, Y.: The singular integral operator and its commutator on weighted Morrey spaces. Collect. Math. 68, 145 -- 174 (2017). [7] Samko, N.: Weighted Hardy and singular operators in Morrey spaces. J. Math. Anal. Appl. 350, 56 -- 72 (2009). BOUNDEDNESS OF OPERATORS ON WEIGHTED MORREY SPACES 17 [8] Samko, N.: On a Muckenhoupt-type condition for Morrey spaces. Mediterr. J. Math. 10, 941 -- 951 (2013). [9] Stein, E. M.: Harmonic analysis: real-variable methods, orthogonality, and oscillatory integrals. Princeton University Press, Princeton, NJ (1993). [10] Tanaka, H.: Two-weight norm inequalities on Morrey spaces. Ann. Acad. Sci. Fenn. Math. 40, 773 -- 791 (2015). (J. D.) 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]
1202.6390
1
1202
2012-02-28T21:52:17
Higher Order Spreading Models
[ "math.FA" ]
We introduce the higher order spreading models associated to a Banach space $X$. Their definition is based on $\ff$-sequences $(x_s)_{s\in\ff}$ with $\ff$ a regular thin family and the plegma families. We show that the higher order spreading models of a Banach space $X$ form an increasing transfinite hierarchy $(\mathcal{SM}_\xi(X))_{\xi<\omega_1}$. Each $\mathcal{SM}_\xi (X)$ contains all spreading models generated by $\ff$-sequences $(x_s)_{s\in\ff}$ with order of $\ff$ equal to $\xi$. We also provide a study of the fundamental properties of the hierarchy.
math.FA
math
HIGHER ORDER SPREADING MODELS S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS Abstract. We introduce the higher order spreading models associated to a Banach space X. Their definition is based on F -sequences (xs)s∈F with F a regular thin family and the plegma families. We show that the higher order spreading models of a Banach space X form an increasing transfinite hierarchy (SMξ(X))ξ<ω1 . Each SMξ(X) contains all spreading models generated by F -sequences (xs)s∈F with order of F equal to ξ. We also provide a study of the fundamental properties of the hierarchy. 1. Introduction Spreading models have been invented by A. Brunel and L. Sucheston [7] in the middle of 70's and since then have a constant presence in the evolution of the Banach space theory. Recall that a sequence (en)n in a seminormed space (E, k·k∗) is called a spreading model of the space X if there exists a sequence (xn)n in X which is Schreier almost isometric to (en)n, that is for some null sequence (δn)n of positive reals we have (1) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) kXi=1 aixni(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXi=1 < δk, aiei(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) i=1 ∈ [−1, 1]k. We also say that the sequence for every k ≤ n1 < . . . < nk and (ai)k (xn)n which satisfies (1) generates (en)n as a spreading model. By an iterated use of Ramsey's theorem [17], Brunel and Sucheston proved that every bounded sequence in a Banach space X has a subsequence generating a spreading model. It is easy to see that any sequence (en)n satisfying (1) is spreading1. The impor- tance of the spreading models arise from the fact that they connect in an asymptotic manner the structure of an arbitrary Banach space X to the corresponding one of spaces generated by spreading sequences. The definition of the spreading model resembles the finite representability2 of the space generated by the sequence (en)n into the space (X, k · k). However there exists a significant difference between the two concepts. Indeed in the frame of the finite representability there are two classi- cal achievements: Dvoretsky's theorem [8] asserting that ℓ2 is finitely representable in every Banach space X and also Krivine's theorem [12] asserting that for every linearly independent sequence (xn)n in X there exists a 1 ≤ p ≤ ∞ such that ℓp is block finitely representable in the subspace generated by (xn)n. On the other hand 12010 Mathematics Subject Classification: Primary 46B03, 46B06, 46B25, 46B45, Secondary 05D10 2Keywords: Spreading models, Ramsey theory, thin families, plegma families 1A sequence (en)n in a seminormed space (E, k · k∗) is called spreading if for every n ∈ N, k1 < . . . < kn in N and a1, . . . , an ∈ R we have that k Pn j=1 aj ejk∗ = k Pn j=1 aj ekj k∗ 2A Banach space Y is finitely representable in X if for every finite dimensional subspace F of Y and every ε > 0 there exists T : F → Y bounded linear injection such that kT k · kT −1k < 1 + ε. 1 2 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS E. Odell and Th. Schlumprecht [15] have shown that there exists a reflexive space X admitting no ℓp as a spreading model. Thus the spreading models of a space X lie strictly between the finitely representable spaces in X and the spaces that are isomorphic to a subspace of X. The spreading models associated to a Banach space X can be considered as a cloud of Banach spaces, including many members with regular structure, surround- ing the space X and offering information concerning the local structure of X in an asymptotic manner. Our aim is to enlarge that cloud and to fill in the gap between spreading models and the spaces which are finitely representable in X. More pre- cisely we extend the Brunel -- Sucheston concept of a spreading model and we show that under the new definition the spreading models associated to a Banach space X form a whole hierarchy of classes of spaces indexed by the countable ordinals. The first class of this hierarchy is the classical spreading models. The initial step of this extension has already been done in [4] where the class of k-spreading models was defined for every positive integer k. The transfinite extension introduced in the present paper requires analogue ingredients that we are about to describe. The first one is the F -sequences; that is sequences of the form (xs)s∈F where the index set F is a regular thin family of finite subsets of N (see Definition 2.7). Typical examples of such families are the k-element subsets of N and also the max- imal elements of the ξth-Schreier family Sξ (see [2]). A subsequence of (xs)s∈F is a restriction of the F -sequence on an infinite subset of N, i.e. it is of the form (xs)s∈F ↾M where F ↾ M = F ∩ [M ]<∞. Among others we study the convergence of F -sequences in a topological space (X, T ). In this setting we show that when the closure of (xs)s∈F in (X, T ) is a compact metrizable space then we can always restrict to an infinite subset M of N where the subsequence (xs)s∈F ↾M is subordi- s ∈ F ↾ M (see Definition 5.8 and Theorem 5.10). nated ; that is if bF = {t ∈ [N]<∞ : ∃s ∈ F such that t is an initial segment of s} then there exists a continuous map ϕ : bF ↾ M → X with ϕ(s) = xs for every The second ingredient is the plegma families which extend the corresponding one in [4]. Roughly speaking a plegma family is a sequence (s1, . . . , sl) of non empty finite subsets of N where the first elements of s1, ..., sl are in increasing order and they lie before their second elements which are also in increasing order and so on (see Definition 3.1). Here the plegma families do not necessarily include sets of equal size. The F -sequences and the plegma families are the key components for the def- inition of the higher order spreading models which goes as follows. Given an F - sequence (xs)s∈F in a Banach space X and a sequence (en) in a seminormed space (E, k · k∗) we will say that (xs)s∈F generates (en)n as an F -spreading model if for some null sequence (δn)n of positive reals we have (2) kXi=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) aixsi(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXi=1 < δk, aiei(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) i=1 ∈ [−1, 1]k and every plegma family (si)k for every (ai)k i=1 in F with k ≤ min s1. Note that the F -sequences (xs)s∈F generate a higher order spreading model just as the ordinary sequences (xn)n do in the classical definition. Moreover since the family of all k-element subsets of N is a regular thin family, the above definition extends the classical definition of the spreading model as well as the one of k- spreading models given in [4]. HIGHER ORDER SPREADING MODELS 3 Brunel -- Sucheston's theorem [6] is extended in the frame of the bounded F - sequences (xs)s∈F in a Banach space X. Namely, every bounded F -sequence in X contains a subsequence (xs)s∈F ↾M generating an F -spreading model. The proof is based on the fact that plegma families with elements in a regular thin family satisfy strong Ramsey properties. It is notable that the concept of the F -spreading model is independent of the particular family F and actually depends only on the order3 of the family F . Namely, if (en)n is an F -spreading model then it is also a G-spreading model for every regular thin family G with o(G) ≥ o(F ). This fact allow us to classify all the F -spreading models of a Banach space X as an increasing transfinite hierarchy of the form (SMξ(X))ξ<ω1 . Let's point out that the ξ-spreading models of X have a weaker asymptotic relation to the space X as ξ increases to ω1. The infinite graphs with vertices from a regular thin family and edges the plegma pairs is the key for the proof of the above results. Specifically, it is shown that if G and F are two regular thin families with o(G) ≥ o(F ) then there exist an infinite subset M of N and a plegma preserving map ϕ : G ↾ M → F (that is (ϕ(s1), ϕ(s2)) is a plegma pair in F whenever (s1, s2) is plegma pair in G ↾ M ). Moreover, it is also shown that such an embedding is forbidden if we wish to go from families of lower to families of higher order. More precisely, if o(F ) < o(G) then for every M ∈ [N]∞ and ϕ : F ↾ M → G there exists L ∈ [M ]∞ such that for every plegma pair (s1, s2) in F ↾ L neither (φ(s1), φ(s2)) nor (φ(s2), φ(s1)) is a plegma pair in G (see Theorems 3.16 and 3.18). The paper is organized as follows. In Section 2 we review some basic facts concerning families of finite subsets of N and we define the regular thin families. In Section 3 we study the plegma families and their properties. In Section 4 we introduce the definition of the higher order spreading models. In Section 5 we deal with F -sequences (xs)s∈F in a general topological space. Finally, in Section 6 we study the F -sequences which generate several classes of spreading sequences as spreading models. In this last section we show that several well known results concerning the classical spreading models remain valid in the higher order setting. For instance, we show that a subordinated, seminormalized and weakly null F - sequence generates an unconditional spreading model. The present paper is an updated version of the first part of [3]. The second part which deals with certain examples will be presented elsewhere. 1.1. Preliminary notation and definitions. By N = {1, 2, . . .} we denote the set of all positive integers. Throughout the paper we shall identify strictly increasing sequences in N with their corresponding range i.e. we view every strictly increasing sequence in N as a subset of N and conversely every subset of N as the sequence resulting from the increasing ordering of its elements. We will use capital letters as L, M, N, ... to denote infinite subsets and lower case letters as s, t, u, ... to denote finite subsets of N. For every infinite subset L of N, [L]<∞ (resp. [L]∞) stands for the set of all finite (resp. infinite) subsets of L. For an L = {l1 < l2 < ...} ∈ [N]∞ and a positive integer k ∈ N, we set L(k) = lk. Similarly, for a finite subset s = {n1 < .. < nm} of N and for 1 ≤ k ≤ m we set s(k) = nk. For an L = {l1 < l2 < ...} ∈ [N ]∞ and a finite subset s = {n1 < .. < nm} (resp. for an infinite subset 3The order of F , denoted by o(F ) is a countable ordinal which measures the complexity F (see Section 2 for the precise definition). For example the family of k-element subsets of N has order k, while the ξth-Schreier has order o(Sξ) = ωξ 4 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS N = {n1 < n2 < ...} of N), we set L(s) = {ln1, ..., lnm} = {L(s(1)), ..., L(s(m))} (resp. L(N ) = {ln1, ln2, ...} = {L(N (1)), L(N (2)), ...}). For s ∈ [N]<∞ by s we denote the cardinality of s. For L ∈ [N]∞ and m ∈ N, we denote by [L]m the set of all s ∈ [L]<∞ with s = m. Also for every nonempty s ∈ [N]<∞ and 1 ≤ k ≤ s we set sk = {s(1), . . . , s(k)} and s0 = ∅. Moreover, for s, t ∈ [N]<∞, we write t ⊑ s (resp. t ⊏ s) to denote that t is an initial (resp. proper initial) segment of s. Also, for every s, t ∈ [N]<∞ we write t < s if either at least one of them is the empty set, or max t < min s. Concerning Banach space theory, although the notation that we follow is the standard one, we present for the sake of completeness some basic concepts that we will need. Let X be a Banach space. We say that a sequence (xn)n in X is bounded (resp. seminormalized) if there exists M > 0 (resp. M1, M2 > 0) such that kxnk ≤ M (resp. M1 ≤ kxnk ≤ M2) for all n ∈ N. The sequence (xn)n is called Schauder basic if there exists a constant C ≥ 1 such that (3) anxn(cid:13)(cid:13)(cid:13), kXn=1 (cid:13)(cid:13)(cid:13) anxn(cid:13)(cid:13)(cid:13) ≤ C(cid:13)(cid:13)(cid:13) mXn=1 for every k ≤ m in N and a1, ..., am ∈ R. Finally, we say that (xn)n is C- unconditional, if for every m ∈ N, F ⊆ {1, ..., m} and a1, ..., am ∈ R, it holds that (4) (cid:13)(cid:13)(cid:13)Xn∈F anxn(cid:13)(cid:13)(cid:13) ≤ C(cid:13)(cid:13)(cid:13) mXn=1 anxn(cid:13)(cid:13)(cid:13). 2. Regular thin families In this section we define the regular thin families of finite subsets of N and we study their basic properties. The definition is based on two well known concepts, namely that of regular families traced back to [2] and that of thin families defined in [14] and extensively studied in [16] and [13]. 2.1. On families of finite subsets of N. We start with a review of the basic concepts concerning families of finite subsets of N. For a more detailed exposition the reader can refer to [5]. 2.1.1. Ramsey properties of families of finite subsets of N. For a family F ⊆ [N]<∞ and L ∈ [N]∞, we set (5) F ↾ L = {s ∈ F : s ⊆ L} = F ∩ [L]<∞. Recall the following terminology from [11]. Let F ⊆ [N]<∞ and M ∈ [N]∞. We say that F is large in M if for every L ∈ [M ]∞, F ↾ L is nonempty. We say that F is very large in M if for every L ∈ [M ]∞ there exists s ∈ F such that s ⊑ L. The following is a restatement ( see [11]) of a well known theorem of C. St. J. A. Nash-Williams [14] and F. Galvin and K. Prikry [9]. Theorem 2.1. Let F ⊆ [N]<∞ and M ∈ [N]∞. If F is large in M then there exists L ∈ [M ]∞ such that F is very large in L. HIGHER ORDER SPREADING MODELS 5 2.1.2. The order of a family of finite subsets of N. Let F ⊆ [N]<∞ be a nonempty family of finite subsets of N. The order of F ⊆ [N]<∞ is defined as follows (see also [16]). First, we assign to F its (⊑ −)closure, i.e. the set (6) bF = {t ∈ [N]<∞ : ∃s ∈ F with t ⊑ s}, which is a tree under the initial segment ordering. If bF is ill-founded (i.e. there exists an infinite sequence (sn)n in bF such that sn ⊏ sn+1) then we set o(F ) = ω1. Otherwise for every maximal element s of bF we set o bF (s) = 0 and recursively for every s in bF we define (7) o bF (s) = sup{o bF (t) + 1 : t ∈ bF and s ⊏ t}. The order of F denoted by o(F ) is defined to be the ordinal o bF (∅). For instance o({∅}) = 0 and o([N]k) = k, for every k ∈ N. For every n ∈ N, we define (8) F(n) = {s ∈ [N]<∞ : n < s and {n} ∪ s ∈ F }, where n < s means that either s = ∅ or n < min s. It is easy to see that for every nonempty family F ⊆ [N]<∞ we have that (9) o(F ) = sup{o(F(n)) + 1 : n ∈ N}. 2.1.3. Regular families. A family R ⊆ [N]<∞ is said to be hereditary if for every s ∈ F and t ⊆ s we have that t ∈ F and spreading if for every n1 < ... < nk and m1 < ... < mk with n1 ≤ m1, ..., nk ≤ mk, we have that {m1, ..., mk} ∈ R whenever {n1, ..., nk} ∈ R. Also R is called compact if the set {χs ∈ {0, 1}N : s ∈ R} of characteristic functions of the members of R, is a closed subset of {0, 1}N under the product topology. A family R of finite subsets of N will be called regular if it is compact, hereditary and spreading. Notice that for every regular family R, bR = R and R(n) is also regular for every n ∈ N. Moreover, using equation (9) and by induction on the order of R we easily get the following. Proposition 2.2. Let R be a regular family. Then o(R ↾ L) = o(R), for every L ∈ [N]∞. Exploiting the method of [16] we have the next result. Proposition 2.3. For every ξ < ω1 there exists a regular family Rξ with o(Rξ) = ξ. Proof. For ξ = 0 we set R0 = {∅}. We proceed by induction on ξ < ω1. Assume that for some ξ < ω1 and for each ζ < ξ we have defined a regular family Rζ with o(Rζ ) = ζ. If ξ is a successor ordinal, i.e. ξ = ζ + 1, then we set If ξ is a limit ordinal, then we choose a strictly increasing sequence (ζn)n such that ζn → ξ and we set Rξ =n{n} ∪ s : n ∈ N, s ∈ Rζ and n < so. Rξ =[n ns ∈ Rζn : min s ≥ no =[n Rζn ↾ [n, +∞). It is easy to check that Rξ is a regular family with o(Rξ) = ξ for all ξ < ω1. (cid:3) 6 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS We will need some combinatorial properties of regular families. To this end we give the following definition. For every R ⊆ [N]<∞ and L ∈ [N]∞, let (10) L(R) = {L(s) : s ∈ R}. Notice that o(R) = o(L(R)) and if R is compact (resp. hereditary) then L(R) is also compact (resp. hereditary). It is also easily verified that if R is spreading then L1(R) ⊆ L2(R), for every L1 ⊆ L2 in [N]∞ and more generally, (11) for every k ∈ N and L1, L2 ∈ [N]∞ satisfying {L1(j) : j > k} ⊆ {L2(j) : j > k} (where L(R(k)) = {L(s) : s ∈ R(k)}). L1(R(k)) ⊆ L2(R(k)), Proposition 2.4. Let R, S be regular families of finite subsets of N with o(R) ≤ o(S). Then for every M ∈ [N]∞ there exists L ∈ [M ]∞ such that L(R) ⊆ S. Proof. If o(R) = 0, i.e. R = {∅}, then the conclusion trivially holds. Suppose that for some ξ < ω1 the proposition is true for every regular families R′, S′ ⊆ [N]<∞ such that o(R′) < ξ and o(R′) ≤ o(S′). Let R, S be regular with o(R) = ξ and let M ∈ [N]∞. By equation (9) we have that o(R(1)) < o(R). Hence o(R(1)) < o(S) and so there is some l1 ∈ N such that o(R(1)) ≤ o(S(l1)). Since S is spreading we have that o(S(l1)) ≤ o(S(n)) for all n ≥ l1 and therefore we may suppose that l1 ∈ M . Since R(1) and S(l1) are regular families, by our inductive hypothesis there is L1 ∈ [M ]∞ such that L1(R(1)) ⊆ S(l1). Proceeding in the same way we construct a strictly increasing sequence (lj)j in M and a decreasing sequence M = L0 ⊃ L1 ⊃ ... of infinite subsets of M such that (i) lj+1 ∈ Lj, (ii) lj+1 > Lj(j) and (iii) Lj(R(j)) ⊆ S(lj ), for all j ≥ 1. We set L = {lj}j and we claim that L(R) ⊆ S. Indeed, by the above construction we have that for every k ∈ N, {L(j)}j>k ⊆ {Lk(j)}j>k. Therefore by (11) and (iii) above, we get that (12) L(R(k)) ⊆ Lk(R(k)) ⊆ S(lk). It is easy to see that L(R(k)) = L(R)(lk) and so by (12) we have that L(R)(lk) ⊆ S(lk). Since this holds for every k ∈ N, we conclude that L(R) ⊆ S. (cid:3) The next corollary is an immediate consequence. Corollary 2.5. Let R, S be regular families of finite subsets of N with o(R) = o(S). Then for every M ∈ [N]∞ there exists L ∈ [M ]∞ such that L(R) ⊆ S and L(S) ⊆ R. 2.1.4. Thin families. A family F of finite subsets of N is called thin if there do not exist s, t in F such that s is a proper initial segment of t. The following result is contained in [14] and [16]. Since it plays a crucial role in the sequel for the sake of completeness we present its proof. Proposition 2.6. Let F ⊆ [N]<∞ be a thin family. Then for every finite partition i=1Fi, (k ≥ 2) of F and every M ∈ [N]∞ there exist L ∈ [M ]∞ and F = ∪k 1 ≤ i0 ≤ k such that F ↾ L ⊆ Fi0 . Proof. It suffices to show the result only for k = 2 since the general case follows easily by induction. So let F = F1 ∪ F2 and M ∈ [N]∞. Then either there is L ∈ [M ]∞ such that F1 ↾ L = ∅ or F1 is large in M . In the first case it is clear HIGHER ORDER SPREADING MODELS 7 In the second case by Theorem 2.1 there is L ∈ [M ]∞ such that F ↾ L ⊆ F2. that F1 is very large in L. We claim that F ↾ L ⊆ F1. Indeed, let s ∈ F ↾ L. We choose N ∈ [L]∞ such that s ⊑ N and let t ⊑ N such that t ∈ F1. Then s, t are ⊑-comparable members of F and since F is thin s = t ∈ F1. Therefore F ↾ L ⊆ F1. (cid:3) 2.2. Regular thin families. We are now ready to introduce the main concept of this section. Definition 2.7. A family F of finite subsets of N will be called regular thin if (a) F is thin and (b) the ⊑-closure bF of F is a regular family. We will use the following notation. For a family R ⊆ [N]<∞ we set The next lemma allow us to construct regular thin families from regular ones. (13) M(R) = {s ∈ R : s is ⊑ -maximal in R}. Notice that a family F ⊆ [N]<∞ is thin if and only if F = M(bF ). Lemma 2.8. Let R be a regular family. Then the family M(R) is thin and satisfies \M(R) = R. Therefore M(R) is regular thin with o(M(R)) = o(R). Proof. Since M(R) ⊆ R and R is hereditary, we have that \M(R) ⊆ R. To show that R ⊆ \M(R) notice that for every s ∈ R there exists a t ∈ M(R) such that s ⊑ t, otherwise R would not be compact. Hence \M(R) = R and clearly M(R) is thin. Thus M(R) is regular thin. Finally, by the definition of the order, we have o(M(R)) = o( \M(R)) and hence o(M(R)) = o(R). (cid:3) Corollary 2.9. For every ξ < ω1 there is a regular thin family Fξ with o(Fξ) = ξ. Proof. Let ξ < ω1 and Rξ be a regular family with o(Rξ) = ξ. Then Fξ = M(Rξ) is as desired. (cid:3) all regular thin families and the set of all regular ones. Moreover, the inverse map sends each regular family R to M(R). Corollary 2.10. The map which sends F to bF is a bijection between the set of Proof. By the definition of regular thin families, the map F → bF sends each regular thin family to a regular one. By Lemma 2.8 we get that the map is 1-1, onto and the inverse map sends each regular family R to M(R). (cid:3) Remark 1. If F is a regular thin family with o(F ) = k < ω, then it is easy to see that there exists n0 such that F ↾ [n0, ∞) = {s ∈ [N]k : min s ≥ n0}. Therefore, for each k < ω, the family [N]k is essentially the unique regular thin family of order k. However this does not remain valid for regular thin families of order ξ ≥ ω. For instance for every unbounded increasing map f : N → N the family F = {s ∈ [N]<∞ : s = f (min s)}, is a regular thin family of order ω. Lemma 2.11. Let R be a regular family and L ∈ [N]∞. Then M(R) ↾ L = M(R ↾ L) and setting M = M(R), \M ↾ L = R ↾ L and o(M ↾ L) = o(R). Proof. It is easy to see that M(R) ↾ L ⊆ M(R ↾ L). To show the converse inclusion let s ∈ M(R ↾ L) and assume that s /∈ M(R). Since R ↾ L ⊆ R, s ∈ R and therefore there exists some t ∈ M(R) with s ⊏ t. Since R is spreading this 8 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS yields that exists t′ ∈ R ↾ L with s ⊏ t′. Thus s /∈ M(R ↾ L), a contradiction. Therefore s ∈ M(R). Since s ∈ [L]<∞, we have that s ∈ M(R) ↾ L. Therefore M(R ↾ L) = M(R) ↾ L. Since M ↾ L = M(R) ↾ L ⊆ R ↾ L and R ↾ L is hereditary, we have that \M ↾ L ⊆ R ↾ L. Conversely, let s ∈ R ↾ L. Since R ↾ L is compact there is t ∈ M(R ↾ L) = M(R) ↾ L with s ⊑ t. Hence s ∈ \M ↾ L and \M ↾ L = R ↾ L. Finally, o(M ↾ L) = o( \M ↾ L) = o(R ↾ L) = o(R), where the last equality (cid:3) follows by Proposition 2.2. Corollary 2.12. Let F be a regular thin family and L ∈ [N]∞. Then F ↾ L = M(bF ↾ L), \F ↾ L = bF ↾ L and o(F ↾ L) = o(F ). Proof. Since F is thin we have that F = M(bF ). Setting R = bF in Lemma 2.11 Corollary 2.13. Let F be a regular thin family. Then for every M ∈ [N]∞ there exists L ∈ [M ]∞ such that F ↾ L is very large in L. the result follows. (cid:3) Proof. If F is regular thin then since bF is spreading, bF ↾ N is nonempty for every N ∈ [N]∞. Since bF ↾ N = \F ↾ N , we get that F ↾ N is nonempty too, i.e. F is large in N. Therefore, by Theorem 2.1, for every M ∈ [N]∞ there exists L ∈ [M ]∞ such that F ↾ L is very large in L. (cid:3) Definition 2.14. For two families F , G of finite subsets of N, we write F ⊑ G (resp. F ⊏ G) if every element in F has an extension (resp. proper extension) in G and every element in G has an initial (resp. proper initial) segment in F . The following proposition is a consequence of a more general result from [10]. Proposition 2.15. Let F , G ⊆ [N]<∞ be regular thin families with o(F ) < o(G). Then for every M ∈ [N]∞ there exists L ∈ [M ]∞ such that F ↾ L ⊏ G ↾ L. Proof. By Corollary 2.13 we have that there exists L1 ∈ [M ]∞ such that both F , G are very large in L1. So for every L ∈ [L1]∞ and every t ∈ G ↾ L there exists s ∈ F ↾ L such that s, t are comparable. Let G1 be the set of all elements of G which have a proper initial segment in F and G2 = G \ G1. By Proposition 2.6 there exist i0 ∈ {1, 2} and L ∈ [L1]∞ such that G ↾ L ⊆ Gi0 . It suffices to show that i0 = 1. Indeed, if i0 = 2 then for every t ∈ G ↾ L there is s ∈ F such that t ⊑ s. This in conjunction with Corollary 2.12 yields that o(G) = o(G ↾ L) ≤ o(F ) which is a contradiction. (cid:3) A similar but weaker result holds when o(F ) = o(G). Proposition 2.16. Let F , G ⊆ [N]<∞ be regular thin families with o(F ) = o(G). Then there exists L0 ∈ [N]∞ such that for every M ∈ [N]∞ there exists L ∈ [L0(M )]∞ such that L0(F ) ↾ L ⊑ G ↾ L. M ∈ [N]∞. Notice that L0(F ) and G are large in L0(M ). Hence by Theorem 2.1 there exists L ∈ [L0(M )]∞ such that L0(F ) and G are very large in N . Since (cid:3) Proof. By Proposition 2.4 there exists L0 ∈ [N]∞ such that L0(bF ) ⊆ bG. Let L0(bF ) ⊆ bG, we conclude that L0(F ) ↾ L ⊑ G ↾ L. Technically the above two propositions are incorporated in one as follows. HIGHER ORDER SPREADING MODELS 9 Corollary 2.17. Let F , G ⊆ [N]<∞ be regular thin families with o(F ) ≤ o(G). Then there exists L0 ∈ [N]∞ such that for every M ∈ [N]∞ there exists L ∈ [L0(M )]∞ such that L0(F ) ↾ L ⊑ G ↾ L. Proof. If o(F ) < o(G), we set L0 = N. Then L0(F ) = F and L0(M ) = M and the conclusion follows by Proposition 2.15. If o(F ) = 0(G) the result is immediate by Proposition 2.16. (cid:3) 3. Plegma families In this section we introduce the notion of plegma families initially appeared in [4] for k-subsets of N. Here we do not assume that the members of a plegma family are necessarily of the same cardinality. 3.1. Definition and basic properties. We begin by stating the definition of a plegma family. Definition 3.1. Let l ∈ N and s1, ..., sl be nonempty finite subsets of N. The l−tuple (sj)l j=1 will be called a plegma family if the following are satisfied. (i) For every i, j ∈ {1, ..., l} and k ∈ N with i < j and k ≤ min(si, sj), we have that si(k) < sj(k). (ii) For every i, j ∈ {1, ..., l} and k ∈ N with k ≤ min(si, sj − 1), we have that si(k) < sj(k + 1). For instance a pair ({n1}, {n2}) of singletons is plegma iff n1 < n2 and a pair of doubletons ({n1, m1}, {n2, m2}) is plegma iff n1 < n2 < m1 < m2. More generally for two non empty s, t ∈ [N]<∞ with s ≤ t the pair (s, t) is a plegma pair iff s(1) < t(1) < s(2) < t(2) < ... < s(s) < t(s). Of course the situation is more involved when the size of a plegma family is large. Below we gather together some stability properties of plegma families. We omit the proof as it is a direct application of the definition. Proposition 3.2. Let (sj)l are satisfied. j=1 be a family of finite subsets of N. Then the following (i) If (sj)l j=1 is a plegma family then (sjm )k every 1 ≤ k ≤ l and 1 ≤ j1 < . . . < jk ≤ l. m=1 is also a plegma family, for (ii) The family (sj)l j=1 is a plegma family iff (sj1 , sj2 ) is a plegma pair, for every 1 ≤ j1 < j2 ≤ l. (iii) If (sj)l j=1 is a plegma family then (tj)l j=1 is also a plegma family, whenever ∅ 6= tj ⊑ sj for 1 ≤ j ≤ l. (iv) If (sj)l j=1 is a plegma family then (L(sj))l j=1 is also a plegma family, for every L ∈ [N]∞. such that s1, ..., sl ∈ F and (sj)l For every family F ⊆ [N]<∞ and l ∈ N we denote by Plml(F ) the set of all (sj)l j=1 j=1 is a plegma family. We also set Plm(F ) = l=1 Plml(F ). Our mai aim is to show that for every l ∈ N, Plml(F ) is a Ramsey family. To this end we need some preparatory lemmas. S∞ Lemma 3.3. Let F be a regular thin family and l ∈ N. Then for every (sj)l Plml(F ) we have that s1 ≤ . . . ≤ sl. j=1 ∈ 10 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS Proof. By (ii) of Proposition 3.2 it suffices to show the conclusion for l = 2. Assume on the contrary that there exists a plegma pair (s1, s2) in F with s1 > s2. We pick s ∈ [N]<∞ such that s = s1, s2 ⊏ s and s(s2 + 1) > max s1. By the definition of the plegma family, we have that for every 1 ≤ k ≤ s2, s1(k) < s2(k) = s(k). Hence, for every 1 ≤ k ≤ s1, we have that s1(k) ≤ s(k). By the spreading property s2 6∈ F , which is a contradiction. of bF we get that s ∈ bF . But since s2 is a proper initial segment of s we get that Lemma 3.4. Let F be a thin family of finite subsets of N and l ∈ N. Let (sj)l j=1sj ⊑ ∪l j=1 in Plml(F ) with s1 ≤ . . . ≤ sl, t1 ≤ . . . ≤ tl and ∪l j=1tj. Then (sj)l j=1 = (tj)l j=1, (tj)l (cid:3) j=1 and consequently ∪l j=1sj = ∪l j=1tj. j=msj and t = ∪l Proof. Suppose that for some 1 ≤ m ≤ l we have that (si)i<m = (ti)i<m. We will show that sm = tm. Let s = ∪l j=mtj. Then by our assumptions s ⊑ t. Moreover since sm ≤ . . . ≤ sl and tm ≤ . . . ≤ tl, we easily conclude that sm(j) = s(cid:0)(j − 1)(l − m + 1) + 1(cid:1), for all 1 ≤ j ≤ sm and similarly tm(j) = t(cid:0)(j − 1)(l − m + 1) + 1(cid:1), for all 1 ≤ j ≤ tm. Hence, as s ⊑ t, we get that for all 1 ≤ j ≤ min{tm, sm}, sm(j) = tm(j). Therefore sm and tm are ⊑-comparable. Since F is thin we have sm = tm. By induction on 1 ≤ m ≤ l, we obtain that sj = tj for every 1 ≤ j ≤ l. (cid:3) Theorem 3.5. Let M be an infinite subset of N, l ∈ N and F be a regular thin family. Then for every finite partition Plml(F ↾ M ) = ∪p i=1Pi, there exist L ∈ [M ]∞ and 1 ≤ i0 ≤ p such that Plml(F ↾ L) ⊆ Pi0. Proof. Let U = {∪l j=1 ∈ Plml(F ↾ M )}. By Lemmas 3.3 and 3.4 we get that U is thin and the map Φ : Plml(F ↾ M ) → U sending each plegma l-tuple (sj)l j=1sj, is an onto bijection. We set Ui = Φ(Pi), for 1 ≤ j ≤ p. Then U = ∪p i=1Ui and by Proposition 2.6 there exist j0 and L ∈ [M ]∞ such that U ↾ L ⊆i0 Ui0 or equivalently Plml(F ↾ L) ⊆ Pj0 . (cid:3) j=1 with si ∈ F ↾ M for 1 ≤ i ≤ l, to its union ∪l j=1sj : (sj)l 3.2. Plegma paths. In this subsection we introduce the definition of the plegma paths in finite subsets of N and we present some of their properties. Such paths will be used in the next subsection for the study of plegma preserving maps. Definition 3.6. Let k ∈ N and s0, ..., sk be nonempty finite subsets of N. We will say that (sj)k j=0 is a plegma path of length k from s0 to sk, if for every 0 ≤ j ≤ k−1, the pair (sj, sj+1) is plegma. Similarly, a sequence (sj)j of nonempty finite subsets of N will be called an infinite plegma path if for every j ∈ N the pair (sj, sj+1) is plegma. The next simple lemma is very useful for the following. Lemma 3.7. Let (s0, . . . , sk−1, s) be a plegma path of length k from s0 to s such that s0 < s. Then k ≥ min{si : 0 ≤ i ≤ k − 1}. Proof. Suppose that k < min{si : 0 ≤ i ≤ k − 1}. Then s(1) < sk−1(2) < sk−2(3) < . . . < s1(k) < s0(k + 1), which contradicts that s0 < s. (cid:3) For a family F ⊆ [N]<∞ a plegma path in F is a (finite or infinite) plegma path which consists of elements of F . It is easy to verify the existence of infinite plegma paths in F whenever F is very large in an infinite subset L of N. In particular, let HIGHER ORDER SPREADING MODELS 11 s ∈ F ↾ L satisfying the next property: for every j = 1, ..., s − 1 there exists l ∈ L such that s(j) < l < s(j + 1). Then it is straightforward that there exists s′ ∈ F ↾ L such that the pair (s, s′) is plegma and moreover s′ shares the same property with s. Based on this one can built an infinite plegma path in F of elements having the above property. The above remarks motivate the following definition. For every F ⊆ [N]<∞ and L ∈ [N]∞, we set The proof of the next lemma follows the same lines with the one of Lemma 2.11. (14) F ↾↾ L =ns ∈ F ↾ L : ∀1 ≤ j ≤ s − 1 ∃ l ∈ L with s(j) < l < s(j + 1)o. Lemma 3.8. Let F be a regular thin family and L ∈ [N]∞. Then \F ↾↾ L = bF ↾↾ L, i.e. s ∈ F ↾↾ L iff s is ⊑-maximal in bF ↾↾ L. We are now ready to present the main result of this subsection. In terms of graph theory it states that in the (directed) graph with vertices the elements of F ↾↾ L and edges the plegma pairs (s, t) in F ↾↾ L, the distance between two vertices s0 and s with s0 < s is equal to the cardinality of s0. Theorem 3.9. Let F be a regular thin family and L ∈ [N]∞. Assume that F is very large in L. Then for every s0, s ∈ F ↾↾ L with s0 < s there exists a plegma path (s0, . . . , sk−1, s) in F ↾↾ L of length k = s0 from s0 to s. Moreover k = s0 is the minimal length of a plegma path in F ↾↾ L from s0 to s. Proof. By Lemmas 3.7 and 3.3 every plegma path in F from s0 to s is of length at least s0. Therefore for s0 < s a plegma path of the form (s0, . . . , sk−1, s) with s0, ..., sk−1, s ∈ F and k = s0 certainly is of minimal length. We will actually prove a slightly more general result. Namely we will show that t from t to s such that all its elements except perhaps t belong to F ↾↾ L. for every t in bF ↾↾ L and s ∈ F ↾↾ L with t < s there exists a plegma path of length For the proof we will use induction on the length of t. The case t = 1 is trivial, since for every s ∈ [N]<∞ with t < s the pair (t, s) is already a plegma path of length 1 from t to s. Suppose that for some k ∈ N the above holds for all t in exist n1 < n2 < . . . < nk+1 in N such that nj − nj−1 > 1, for 2 ≤ j ≤ k and t = {L(nj) : 1 ≤ j ≤ k + 1}. We set t0 = {L(nj − 1) : 2 ≤ j ≤ k + 1}. Since nj − 1 > nj−1 we have that t0 is of equal cardinality and pointwise strictly bF ↾↾ L with t = k. Let t ∈ bF ↾↾ L with t = k + 1 and s ∈ F ↾↾ L with t < s. Then there greater than t \ {max t}. Hence, since bF is spreading, we have t0 ∈ bF ↾↾ L and moreover t0 cannot be a ⊑-maximal element of bF ↾↾ L. By Lemma 3.8 we have that F ↾↾ L is the set of all ⊑-maximal elements of bF ↾↾ L. Therefore, we conclude that t0 ∈ bF \ F . Thus, since t0 = k, by the inductive hypothesis, there exists a plegma Let l = s1. Since (t0, s1) is a plegma pair with t0 ∈ bF \ F and s1 ∈ F , arguing as in Lemma 3.3, we see that l ≥ k + 1. Moreover since s1 ∈ F ↾↾ L, there exist m1 < . . . < ml in N such that mj − mj−1 > 1 and s1 = {L(mj) : 1 ≤ j ≤ l}. Notice that n2 ≤ m1 < n3 ≤ m2 < ... < nk ≤ mk−1 < mk+1 − 1. path (t0, s1, . . . , sk−1, s) of length k = t0 from t0 to s with all s1, . . . , sk−1, s in F ↾↾ L. We set w = t0 ∪ {L(mj − 1) : k + 1 ≤ j ≤ l} and let L′ ∈ [L]∞ such that w is an initial segment of L′. Notice that w = l. Since F is very large in L there exists 12 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS that t0 < s0 ≤ l and therefore t0 ⊏ s0 ⊑ w. s0 ∈ F with s0 initial segment of L′. Using again that bF is spreading it is shown It is easy to check that (t, s0) and (s0, s1) are plegma pairs. Hence the sequence (t, s0, . . . , sk−1, s) is a plegma path of length k + 1 from t to s with s0, . . . , sk−1, s ∈ F ↾↾ L. The proof of the inductive step as well as of the theorem is complete. (cid:3) We close this section by presenting an application of the above theorem. We start with the following definition. Let X be a set, M ∈ [N]∞, F ⊆ [N]<∞ and ϕ : F → X. We will say that ϕ is hereditarily nonconstant in M if for every L ∈ [M ]∞ the restriction of ϕ on F ↾ L is nonconstant. In particular if M = N then we will simply say that ϕ is hereditarily nonconstant. Lemma 3.10. Let F be a regular thin family, X be a set and ϕ : F → X be hereditarily nonconstant. Then for every N ∈ [N]∞ there exists L ∈ [N ]∞ such that for every plegma pair (s1, s2) in F ↾ L, ϕ(s1) 6= ϕ(s2). Proof. By Theorem 3.5 there exists an L ∈ [N ]∞ such that either ϕ(s1) 6= ϕ(s2), for all plegma pairs (s1, s2) in F ↾ L, or ϕ(s1) = ϕ(s2) for all plegma pairs (s1, s2) in F ↾ L. The second alternative is excluded. Indeed, suppose that ϕ(s1) = ϕ(s2), for every plegma pair (s1, s2) in F ↾ L. By Corollary 2.13 we may also assume that F ↾ L is very large in L. Let s0 be the unique initial segment of L0 =(cid:8)L(2ρ) : ρ ∈ N(cid:9) in 0 =(cid:8)L(2ρ) : ρ ∈ N and ρ > k(cid:9). By Theorem 3.9 for F ↾ L and let k = s0. We set L′ every s ∈ F ↾ L′ Therefore for every s ∈ F ↾ L′ which contradicts that ϕ is hereditarily nonconstant. 0 there exist a plegma path (s0, s1, . . . , sk−1, s) of length k in F ↾ L. 0 we have that ϕ(s) = ϕ(sk−1) = . . . = ϕ(s1) = ϕ(s0), (cid:3) Proposition 3.11. Let F be a regular thin family, M ∈ [N]∞ and ϕ : F → N be hereditarily non constant in M . Let also g : N → N. Then there exists N ∈ [M ]∞ such that for every plegma pair (s1, s2) in F ↾ N , ϕ(s2) − ϕ(s1) > g(n), where min s2 = N (n). Proof. By Theorem 3.5 there exists L ∈ [M ]∞ such that one of following holds. (i) For every plegma pair (s1, s2) in F ↾ L, we have ϕ(s1) = ϕ(s2). (ii) For every plegma pair (s1, s2) in F ↾ L, we have ϕ(s1) > ϕ(s2). (iii) For every plegma pair (s1, s2) in F ↾ L, we have ϕ(s1) < ϕ(s2). Since ϕ is hereditarily non constant in M , by Lemma 3.10 case (i) is excluded. Similarly case (ii) cannot occur since otherwise, (ϕ(sn))n would form a strictly decreasing sequence in N whenever (sn)n is an infinite plegma path in F ↾ L. Therefore, case (iii) holds. We choose N ∈ [L]∞ such that for every n ≥ 2, we have that Let (s1, s2) be a plegma pair in F ↾ N and let n ∈ N such that min s2 = N (n). Notice for every 1 ≤ k ≤ s1, we have g(j). j≤n (cid:12)(cid:12)(cid:12)nl ∈ L : N (n − 1) < l < N (n)o(cid:12)(cid:12)(cid:12) ≥ max (cid:12)(cid:12)(cid:12)nl ∈ L : s1(k) < l < s2(k)o(cid:12)(cid:12)(cid:12) ≥ g(n). (cid:12)(cid:12)(cid:12)nl ∈ L : s2(k − 1) < l < s2(k)o(cid:12)(cid:12)(cid:12) ≥ g(n). Similarly for every s1 < k ≤ s2, we have HIGHER ORDER SPREADING MODELS 13 The above yield that there exist t1, . . . , tg(n) ∈ F ↾ L such that the (g(n) + 2)-tuple (s1, t1, . . . , tg(n), s2) is plegma. Hence ϕ(s2) − ϕ(s1) > g(n). (cid:3) Corollary 3.12. Let F be a regular thin family, M ∈ [N]∞ and ϕ : F → N be hereditarily non constant in M . Then there exists N ∈ [M ]∞ such that for every plegma pair (s1, s2) in F ↾ N we have ϕ(s2) − ϕ(s1) > 1. 3.3. Plegma preserving maps. Let F ⊆ [N]<∞ and ϕ : F → [N]<∞. We will say that the map ϕ is plegma preserving if (ϕ(s1), ϕ(s2)) is a plegma pair whenever (s1, s2) is a plegma pair in F . Lemma 3.13. Let F ⊆ [N]<∞ and ϕ : F → [N]<∞. If ϕ is plegma preserving then for every l ∈ N and (sj)l j=1 is a plegma l-tuple. j=1 ∈ Plm(F ) we have that (ϕ(sj ))l Proof. Let l ∈ N and (sj)l j=1 be a plegma l-tuple in F . Then for every 1 ≤ j1 < j2 ≤ l we have that (sj1 , sj2 ) is plegma and thus (ϕ(sj1 ), ϕ(sj2 )) is plegma. Hence, by (ii) of Proposition 3.2, (ϕ(sj))l (cid:3) j=1 is a plegma l-tuple. Proposition 3.14. Let F be a regular thin family and ϕ : F → [N]∞. Then for every M ∈ [N]∞ there is L ∈ [M ]∞ such that exactly one of the following holds. (i) The restriction of ϕ on F ↾ L is plegma preserving. (ii) For every (s1, s2) ∈ Plm2(F ↾ L) neither (ϕ(s1), ϕ(s2)) nor (ϕ(s2), ϕ(s1)) is a plegma pair. Proof. Assume that there is M ∈ [N]∞ such that for every L ∈ [M ]∞ neither (i) nor (ii) holds true. Then by Theorem 3.5 there exists L ∈ [M ]∞ such that for every (s1, s2) ∈ Plm2(F ↾ N ) we have that (ϕ(s2), ϕ(s1)) is plegma. But this is impossible. Indeed, otherwise for an infinite plegma path (sn)n in F ↾ N the sequence (min sn)n would form a strictly decreasing infinite sequence in N. (cid:3) For a family F ⊆ [N]<∞ and a plegma preserving map ϕ : F → [N]<∞ we will say that ϕ is normal provided that ϕ(s1) ≤ ϕ(s2) for every plegma pair (s1, s2) in F and ϕ(s) ≤ s for every s ∈ F . Theorem 3.15. Let F be a regular thin family, M ∈ [N]∞ and ϕ : F ↾ M → [N]<∞ be a plegma preserving map. Then there exists L ∈ [M ]∞ such that the restriction of ϕ on F ↾ L is a normal plegma preserving map. Proof. By Theorem 3.5 there exists N ∈ [M ]∞ such that either (a) ϕ(s1) ≤ ϕ(s2), for every plegma pair (s1, s2) in F ↾ N , or (b) ϕ(s1) > ϕ(s2), for every plegma pair (s1, s2) in F ↾ N . Alternative (b) cannot occur since otherwise for an infinite plegma path (sn)n in F ↾ N the sequence (ϕ(sn))n would form a strictly decreasing sequence in N. By Theorem 2.6 there exists L ∈ [N ]∞ such that either (c) ϕ(s) ≤ s, for every s ∈ F ↾ L, or (d) ϕ(s) > s, for every s ∈ F ↾ L. We claim that (d) cannot hold true. Indeed, since ϕ on F ↾ L is plegma preserving, using a plegma path of enough large length, we may choose s0, s in F ↾↾ L such that min s0 < min s and min ϕ(s0) < min ϕ(s). Let k0 = s0. Then by Proposition 3.9 there exists a plegma path (si)k0 i=0 in F ↾↾ L from s0 to s = sk0 of length k0. By Lemma 3.13 (ϕ(si))k0 i=0 is also a plegma path of length k0 from ϕ(s0) to ϕ(sk0 ) and by Lemma 3.7 we have that (15) min{ϕ(si) : 0 ≤ i ≤ k0 − 1} ≤ k0. 14 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS Moreover by Lemma 3.4, we have s0 ≤ s1 ≤ ... ≤ sk0 . Hence if (d) holds true then (16) min{ϕ(si) : 0 ≤ i ≤ k0 − 1} > min{si : 0 ≤ i ≤ k0 − 1} ≥ k0, which contradicts (15). Therefore we conclude that (a) and (c) hold true i.e. the restriction of ϕ on (cid:3) F ↾ L is a normal plegma preserving map. 3.4. Plegma preserving maps between thin families. In this subsection we are concerned with the question of the existence of a plegma preserving map ϕ : G → F , where G and F are regular thin families. We shall show that such maps exist only when o(G) ≥ o(F ). We start with the positive result. Theorem 3.16. Let F , G be regular thin families with o(F ) ≤ o(G). Then for every M ∈ [N]∞ there is N ∈ [N]∞ and a plegma preserving map ϕ : G ↾ N → F ↾ M . Moreover, for every l ∈ N and t ∈ G ↾ N , if min t ≥ N (l) then min ϕ(t) ≥ M (l). Proof. Let M ∈ [N]∞. By Corollary 2.17 there exists L0 ∈ [N]∞ and N ∈ [L0(M )]∞ such that L0(F ) ↾ N ⊑ G ↾ N . Thus for every t ∈ G ↾ N there exists a unique st ∈ F such that L0(st) ⊑ t. Moreover, L0(st) ⊑ t ⊆ N ⊆ L0(M ) and therefore st ⊆ M . We define ϕ : G ↾ N → F ↾ M , by setting ϕ(t) = st. To see that ϕ is plegma preserving, let (t1, t2) be a plegma pair in G ↾ N . Then L0(ϕ(ti)) ⊑ ti, for i ∈ {1, 2} and therefore by (iv) of Proposition 3.2 (L0(ϕ(t1), (L0(ϕ(t2)) and (ϕ(t1), ϕ(t2)) are also plegma pairs. Hence ϕ is plegma preserving. Finally, let l ∈ N and t ∈ G ↾ N with min t ≥ N (l). Since L0(ϕ(t)) ⊑ t, we have that min L0(ϕ(t)) = min t and therefore L0(min ϕ(t)) = min L0(ϕ(t)) = min t ≥ N (l) ≥ L0(M )(l) = L0(M (l)). Hence min ϕ(t) ≥ M (l). (cid:3) For the following we shall need the next definition. Let F ⊆ [N]<∞ and L ∈ [N]∞. We define (17) L−1(F ) =nt ∈ [N]<∞ : L(t) ∈ Fo. It is easy to see that for every family F ⊆ [N]<∞ and L ∈ [N]∞ the following are satisfied. (a) If F is very large in L then the family L−1(F ) is very large in N. (b) If F is regular thin then so does the family L−1(F ). (c) o(L−1(F )) = o(F ↾ L). In particular if F is regular thin then o(L−1(F )) = o(F ). Lemma 3.17. Let F be a regular thin family, L ∈ [N]∞ such that F is very large in L. Let ϕ : F ↾ L → [N]<∞ be a normal plegma preserving map. Let ψ : L−1(F ) → [N]<∞ defined by ψ(u) = ϕ(L(u)) for every u ∈ L−1(F ). Then ψ is a normal plegma preserving map which in addition satisfies the following property. If u ∈ L−1(F ) ↾↾ N and w = ψ(u) then u(i) ≤ w(i) for every 1 ≤ i ≤ w. Proof. It is easy to check that ψ is a normal plegma preserving map. Therefore we pass to the proof of the property of ψ. First, by induction on k = u(1), we shall show that u(1) ≤ ψ(u)(1), for all u ∈ L−1(F ) ↾↾ N. Indeed, if u(1) = 1 then obviously ψ(u)(1) ≥ 1 = u(1). Suppose that for some k ∈ N and every u ∈ L−1(F ) ↾↾ N with u(1) = k we have that ψ1(u)(1) ≥ k. Let u ∈ L−1(F ) ↾↾ N with u(1) = k + 1. Since L−1(F ) is regular thin and very large in N, we easily see that there exists a unique u′ ∈ L−1(F ) with u′ ⊑ {u(ρ) − 1 : 1 ≤ ρ ≤ u}. Notice that (u′, u) is a plegma pair HIGHER ORDER SPREADING MODELS 15 in L−1(F ) and u′(1) = k. Since ψ is a normal plegma preserving map we have that (ψ(u′), ψ(u)) is also a plegma pair. Hence ψ(u)(1) > ψ(u′)(1) ≥ u′(1) = k, that is ψ(u)(1) ≥ k + 1 = u(1). Suppose now that for some i ∈ N and every u ∈ L−1(F ) ↾↾ N with i ≤ ψ(u), u(i) ≤ ψ(u)(i). Let u ∈ L−1(F ) ↾↾ N with i+1 ≤ ψ(u). Since L−1(F ) is very large in N, there exists u′ ∈ L−1(F ) ↾↾ N such that {u(ρ)−1 : 2 ≤ ρ ≤ u} ⊑ u′. Observe that (u, u′) is plegma pair in L−1(F ), u ≤ u′ and u(i + 1) = u′(i) + 1. Since ψ is normal plegma preserving, we have that (ψ(u), ψ(u′)) is also a plegma pair and in addition i+1 ≤ ψ(u) ≤ ψ(u′). Hence, ψ(u)(i+1) > ψ(u′)(i) ≥ u′(i) = u(i+1)−1, that is ψ(u)(i + 1) ≥ u(i + 1). By induction on i ∈ N the proof is complete. (cid:3) Theorem 3.18. Let F , G be regular thin families with o(F ) < o(G) and M ∈ [N]∞. Then there is no plegma preserving map from F ↾ M to G. More precisely for every M ∈ [N]∞ and ϕ : F ↾ M → G and there exists L ∈ [M ]∞ such that for every plegma pair (s1, s2) in F ↾ L neither (φ(s1), φ(s2)) nor (φ(s2), φ(s1)) is a plegma pair. Proof. Assume that there exist M ∈ [N]∞ and ϕ : F ↾ M → G such that ϕ is plegma preserving. By Theorem 3.15 there exists L ∈ [M ]∞ such that the restriction of ϕ on F ↾ L is a normal plegma preserving map. By Corollary 2.13 we may also assume that F is very large in L. Since o(L−1(F )) = o(F ) < o(G) by Proposition 2.15 we have that there exists N ∈ [N]∞ such that L−1(F ) ↾ N ⊏ G ↾ N . We may assume that N (i + 1) − N (i) > 1 and therefore L−1(F ) ↾ N ⊆ L−1(F ) ↾↾ N. Let ψ : L−1(F ) → [N]<∞ defined by ψ(u) = ϕ(L(u)) for every u ∈ L−1(F ). Pick u0 ∈ L−1(F ) ↾ N and set w0 = ψ(u0). Since L−1(F ) ↾ N ⊏ G ↾ N , we have that u0 ∈ bG \ G and since ϕ takes values in G, we have that w0 ∈ G. We are now ready to derive a contradiction. Indeed, by Lemma 3.17 we have that ψ is a normal plegma preserving map which implies that w0 ≤ u0. Moreover, since L−1(F ) ↾ N ⊆ L−1(F ) ↾↾ N, we have that u0 ∈ L−1(F ) ↾↾ N. Hence, again by Lemma 3.17, we get that u0(i) ≤ w0(i), for every 1 ≤ i ≤ w0. Summarizing we have that u0 ∈ bG \ G, w0 ≤ u0 and u0(i) ≤ w0(i), for every 1 ≤ i ≤ w0. Since bG is spreding we conclude that w0 ∈ bG \ G, which is impossible. Therefore there is no M ∈ [N]∞ and ϕ : F ↾ M → G such that ϕ is plegma preserving. By Proposition 3.14, we get that for every M ∈ [N]∞ and ϕ : F ↾ M → G there is L ∈ [M ]∞ such that for every plegma pair (s1, s2) in F ↾ L neither (φ(s1), φ(s2)) nor (φ(s2), φ(s1)) is plegma. (cid:3) 4. The hierarchy of spreading models In this section we define the class of the ξ-spreading models of a Banach space X for every countable ordinal 1 ≤ ξ < ω1. The definition is a transfinite extension of the corresponding one of the finite order spreading models given in [4]. The basic ingredients of this extension are the concepts of the F -sequences in X, i.e. sequences of the form (xs)s∈F with xs ∈ X for every s ∈ F and the plegma families with members in F where F is a regular thin family. 4.1. The F -spreading models of a Banach space X. Let X be a Banach space and F ⊆ [N]<∞ be a regular thin family. By the term F -sequence in X we will mean a map ϕ : F → X. An F -sequence in X will be usually denoted by (xs)s∈F , where xs = ϕ(s) for all s ∈ F . Also, for every M ∈ [N]∞, the map ϕ : F ↾ M → X 16 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS will be called an F -subsequence of (xs)s∈F and will be denoted by (xs)s∈F ↾M . An F -sequence (xs)s∈F in X will be called bounded (resp. seminormalized) if there exists C > 0 (resp. 0 < c < C) such that kxsk ≤ C (resp. c ≤ kxsk ≤ C) for every s ∈ F . Lemma 4.1. Let (xs)s∈F be a bounded F -sequence in X. Let k ∈ N, N ∈ [N]∞ and δ > 0. Then there exists M ∈ [N ]∞ such that kXj=1 ajxtj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXj=1 ≤ δ ajxsj(cid:13)(cid:13)(cid:13)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) kXj=1 an j xtj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXj=1 ≤ δ 3 , an j xsj(cid:13)(cid:13)(cid:13)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) for every (tj)k j=1, (sj)k j=1 ∈ Plmk(F ↾ M ) and a1, ..., ak ∈ [−1, 1]. Proof. Let (an)n0 (an Nn0 satisfying 1 , ..., an n=1 be a δ 3l −net of the unit ball of (Rk, k · k∞). Setting an = k ) for every 1 ≤ n ≤ n0, we inductively construct N = N0 ⊇ N1 ⊇ . . . ⊇ for every 1 ≤ n ≤ n0 and every (sj)k j=1, (tj)k j=1 ∈ Plmk(F ↾ Nn). have been constructed. Define gn : Plmk(F ↾ Nn−1) → [0, lC] by gn((sj)k j xsj k. By dividing the interval [0, lC] into disjoint intervals of length δ The inductive step of the construction has as follows. Suppose that N0, . . . , Nn−1 j=1) = 3 and j=1) ≤ δ 3 , applying Theorem 3.5, there is Nn ⊆ Nn−1 such that gn((tj )k for every (tj)k kPk j=1)−gn((sj )k j=1 ∈ Plmk(F ↾ Nn). j=1, (sj)k j=1 an We set M = M (Nn0). By (19) and taking into account that (an)n0 n=1 is a δ of the unit ball of (Rk, k · k∞) it is easy to see that L is as desired. Lemma 4.2. Let (xs)s∈F be a bounded F -sequence in X. Let l ∈ N, N ∈ [N]∞ and δ > 0. Then there exists M ∈ [N ]∞ such that 3 −net (cid:3) for every 1 ≤ k ≤ l, (tj)k j=1, (sj)k j=1 ∈ Plmk(F ↾ M ) and a1, ..., ak ∈ [−1, 1]. Proof. It follows easily by an iterated use of Lemma 4.1. (cid:3) Lemma 4.3. Let (xs)s∈F be a bounded F -sequence in X. Then for every sequence (δn)n of positive real numbers and N ∈ [N]∞ there exists M ∈ [N ]∞ satisfying kXj=1 ajxtj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXj=1 ≤ δ ajxsj(cid:13)(cid:13)(cid:13)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) kXj=1 ajxtj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXj=1 ajxsj(cid:13)(cid:13)(cid:13)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ δl, for every 1 ≤ k ≤ l, a1, ..., ak ∈ [−1, 1] and (tj)k that s1(1), t1(1) ≥ M (l). j=1, (sj )k j=1 ∈ Plmk(F ↾ M ) such Proof. It is straightforward by Lemma 4.2 and a standard diagonalization. (cid:3) Hence, assuming in the above lemma that (δn)n is a null sequence, we get that for every l ∈ N and every sequence(cid:0)(sn 1 (1) → ∞ the sequence (kPl j=1 ajxn sn j )l j=1(cid:1)n of plegma l−tuples in F ↾ M with sj k)n is convergent and its limit is independent (18) (19) (20) (21) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) HIGHER ORDER SPREADING MODELS 17 from the choice of ((sn under which the natural Hamel basis (en)n satisfies j )l j=1)n. Actually, we may define a seminorm k · k∗ on c00(N) (22) ≤ δl, (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) kXj=1 ajxsj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXj=1 ajej(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) for all 1 ≤ k ≤ l, (ai)k i=1 in [−1, 1] and (sj)k j=1 ∈ Plmk(F ↾ M ) with s1(1) ≥ M (l). Let's notice here that there exist bounded F -sequences in Banach spaces such that no seminorm resulting from Lemma 4.3 is a norm. For example this happens in the case where (xs)s∈F is constant. Moreover, even if the k · k∗ is a norm on c00(N), the sequence (en)n is not necessarily Schauder basic. We are now ready to give the definition of the F -spreading models of a Banach space X. Definition 4.4. Let X be a Banach space, F be a regular thin family, (xs)s∈F be an F -sequence in X. Let (E, k · k∗) be an infinite dimensional seminormed linear space with Hamel basis (en)n. Also let and M ∈ [N]∞ and (δn)n be a null sequence of positive real numbers. We will say that the F -subsequence (xs)s∈F ↾M generates (en)n as an F -spreading i=1 in [−1, 1] and model (cid:0)with respect to (δn)n(cid:1) if for every l ∈ N, 1 ≤ k ≤ l, (ai)k j=1 ∈ Plmk(F ↾ M ) with s1(1) ≥ M (l), we have (sj)k (23) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) kXj=1 ajxsj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXj=1 ajej(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ δl. We will also say that (xs)s∈F admits (en)n as an F -spreading model if there exists M ∈ [N]∞ such that (xs)s∈F ↾M generates (en)n as an F -spreading model. Finally, for a subset A of X, we will say that (en)n is an F -spreading model of A if there exists an F -sequence (xs)s∈F in A which admits (en)n as an F -spreading model. The next remark is straightforward. Remark 2. Let M ∈ [N]∞ such that (xs)s∈F ↾M generates (en)n as an F -spreading model. Then the following are satisfied. (i) The sequence (en)n is spreading, i.e. for every n ∈ N, k1 < . . . < kn in N and a1, . . . , an ∈ R we have that kPn j=1 ajejk∗ = kPn (ii) For every M ′ ∈ [M ]∞ we have that (xs)s∈F ↾M ′ generates (en)n as an F - j=1 ajekj k∗. spreading model. (iii) For every null sequence (δ′ n)n of positive reals there exists M ′ ∈ [M ]∞ such that (xs)s∈F ↾M ′ generates (en)n as an F -spreading model with respect to (δ′ n)n. By Lemma 4.3 we get the following. Theorem 4.5. Let F be a regular thin family and X a Banach space. Then every bounded F -sequence in X admits an F -spreading model. In particular for every bounded F -sequence (xs)s∈F in X and every N ∈ [N]∞ there exists M ∈ [N ]∞ such that (xs)s∈F ↾M generates F -spreading model. 18 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS 4.2. Spreading models of order ξ. In this subsection we show that Definition 4.4 is independent of the particular regular thin family F and actually depends on the order of F . More precisely we have the following. Lemma 4.6. Let F , G be regular thin families with o(F ) ≤ o(G). Let X be a Banach space and (xs)s∈F be an F -sequence in X which admits an F -spreading model (en)n. Then there exists a G-sequence (wt)t∈G with {wt : t ∈ G} ⊆ {xs : s ∈ F } and which admits (en)n as a G-spreading model. Proof. Let M ∈ [N]∞ and (δn) ց 0 be such that (xs)s∈F ↾M generates (en)n as an F -spreading model with respect to (δn)n. By Theorem 3.16 there exist N ∈ [N]∞ and a plegma preserving map ϕ : G ↾ N → F ↾ M such that min ϕ(t) ≥ M (l), for every l ∈ N and t ∈ G ↾ N with min t ≥ N (l). For every t ∈ G ↾ N let wt = xϕ(t) and for every t ∈ G \ (G ↾ N ) let wt = xs0 where s0 is an arbitrary element of F . We claim that (wt)t∈G↾N generates (en)n as a G-spreading model with respect to (δn)n. Indeed, fix l ∈ N, 1 ≤ k ≤ l, (aj )k j=1 in [0, 1] and (tj )k j=1 ∈ G ↾ N with t1(1) ≥ N (l). Let sj = ϕ(tj ), for all 1 ≤ j ≤ k. Then (sj)k j=1 ∈ Plml(F ↾ M ) and s1(1) ≥ M (l). Therefore, since (xs)s∈F ↾M generates (en)n as an F -spreading model with respect to (δn)n, we have kXj=1 ajwtj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) and the proof is complete. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) kXj=1 ajej(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) kXj=1 ajxsj(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) kXj=1 ajej(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ δl (cid:3) Corollary 4.7. Let X be a Banach space, A ⊆ X and F , G be regular thin families with o(F ) = o(G). Then (en)n is an F -spreading model of A iff (en)n is a G- spreading model of A. The above permits us to give the following definition. Definition 4.8. Let A be a subset of a Banach space X and ξ ≥ 1 be a countable ordinal. We will say that (en)n is a ξ-spreading model of A if there exists a regular thin family F with o(F ) = ξ such that (en)n is an F -spreading model of A. The set of all ξ-spreading models of A will be denoted by SMξ(A). Notice that by Lemma 4.6 we have (24) SMζ(A) ⊆ SMξ(A), for every 1 ≤ ζ < ξ < ω1. The following is an extension of Example 1 in [4]. It shows that for a given ξ < ω1 and a regular thin family G there exists a norm on c00(G) such that setting A = {es : s ∈ G} (where (es)s∈F is the natural Hamel basis of c00(G)), we have SMζ(A) ( SMξ(A), for every ζ < ξ. Example 1. Let 1 ≤ ξ < ω1, G be a regular thin family of order ξ and (es)s∈G be the natural Hamel basis of c00(G). Let (E, k·k) be a Banach space with a normalized spreading and 1-unconditional basis (en)n which in addition is not equivalent to the usual basis of c0. Let XG be the completion of c00(G) under the norm k · kF defined HIGHER ORDER SPREADING MODELS 19 by (25) kxkG = supn(cid:13)(cid:13)(cid:13) lXi=1 ati ei(cid:13)(cid:13)(cid:13) : l ∈ N, (ti)l for every x =Pt∈G atet ∈ c00(G). i=1 ∈ Plml(G) and l ≤ t1(1)o, Let A = {et : t ∈ G}. It is easy to see that (et)t∈G generates (en)n as an G- spreading model, i.e. (en)n belongs to SMξ(A). Let ζ < ξ. We claim that for every (e′ n)n is generated by a constant F -sequence with o(F ) = ζ or it is isometric to the usual basis of c0. Thus (en)n /∈ SMζ(A). n)n ∈ SMζ(A) either (e′ Indeed, let (e′ n)n ∈ SMζ (A). Then there exists a regular thin family F of order ζ, an F -sequence (xs)s∈F in A and M ∈ [N]∞ such that (xs)s∈F ↾M generates (e′ n)n as an F -spreading model. Since {xs}s∈F ⊆ A, we may define ϕ : F ↾ M → G by choosing for each s ∈ F ↾ M an element ϕ(s) ∈ G satisfying eϕ(s) = xs. Assume for the following that (e′ n)n ∈ SMζ(A) is not generated by a constant F -sequence with o(F ) = ζ. By part (ii) of Remark 2 we have that ϕ is hereditarily nonconstant and by Lemma 3.10 there exists N ∈ [M ]∞ such that for every plegma pair (s1, s2) in F ↾ N , ϕ(s1) 6= ϕ(s2). Moreover since o(F ) < o(G) by Theorem 3.18 we have that there exists L ∈ [N ]∞ such that for every plegma pair (s1, s2) in G ↾ L neither (ϕ(s1), ϕ(s2)), nor (ϕ(s2), ϕ(s1)) is a plegma pair. Therefore, by part (i) of Proposition 3.2, we conclude that for every 1 ≤ k ≤ l, (sj)k j=1 ∈ Plmk(F ↾ L) and (ti)l i=1 ∈ Plml(G), we must have Hence, for every k ∈ N, a1, . . . , ak ∈ R and (sj)k j=1 ∈ Plmk(F ↾ L), we have (26) (27) (cid:12)(cid:12)(cid:8)j ∈ {1, . . . , k} : ϕ(sj) ∈ {ti : 1 ≤ i ≤ l}(cid:9)(cid:12)(cid:12) ≤ 1. ajxsj(cid:13)(cid:13)(cid:13)G kXj=1 ajeϕ(sj)(cid:13)(cid:13)(cid:13)G =(cid:13)(cid:13)(cid:13) kXj=1 max 1≤j≤k (25),(26) = (cid:13)(cid:13)(cid:13) aj, i.e. (e′ n)n is isometric to the usual basis of c0. A natural question arising from the above is the following. Question. Let X be a separable Banach space. countable ordinal ξ such that SMζ(X) = SMξ(X) for every ζ > ξ? Is it true that there exists a The above question can be also stated in an isomorphic version, i.e. whether every sequence in SMζ(X) is equivalent to some sequence in SMξ(X) and vice versa. Remark 3. In a forthcoming paper we will provide examples establishing the hierarchy of the higher order spreading models and also illustrating the boundaries of the theory. Specifically we will show the following. (1) For every countable limit ordinal ξ there exist a Banach space X such that SMξ(X) properly includes up to equivalence ∪ζ<ξSMζ(X). (2) There exist a Banach space X such that for every ξ < ω1 and every (en)n ∈ SMξ(X), the space E generated by (en)n does not contain any isomorphic copy of c0 or ℓp for all 1 ≤ p < ∞. The above results require a deeper study of the structure of F -sequences generating ℓ1-spreading models (see also [4] for the finite order case). 20 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS 5. F -sequences in topological spaces Let (X, T ) be a topological space and F be a regular thin family. As we have already defined in the previous section, an F -sequence in X is any map of the form ϕ : F → X and generally an F -subsequence in X is any map of the form ϕ : F ↾ M → X. In this section we will study the topological properties of F - sequences. The particular case where F = [N]k, k ∈ N had been studied in [4]. 5.1. Convergence of F -sequences. We introduce the following natural definition of convergence of F -sequences. Definition 5.1. Let (X, T ) be a topological space, F a regular thin family, M ∈ [N]∞, x0 ∈ X and (xs)s∈F an F -sequence in X. We will say that the F -subsequence (xs)s∈F ↾M converges to x0 if for every U ∈ T with x0 ∈ U there exists m ∈ N such that for every s ∈ F ↾ M with min s ≥ M (m) we have that xs ∈ U . It is immediate that if an F -subsequence (xs)s∈F ↾M in a topological space X is convergent to some x0, then every further F -subsequence is also convergent to x0. Also notice that if o(F ) ≥ 2 then the convergence of (xs)s∈F ↾M does not in general imply that {xs : s ∈ F ↾ M } is a relatively compact subset of X. For instance, let i=s(1) ei, where (ei)i is the usual basis of c0. According to Definition 5.1 the [N]2-sequence (xs)s∈[N]2 weakly = {xs : s ∈ [N]2} ∪ {0} which is not a weakly converges to zero but {xs : s ∈ [N]2} compact subset of c0. (xs)s∈[N]2 be the [N]2-sequence in c0 defined by xs =Ps(2) w Proposition 5.2. Let (X, TX ), (Y, TY ) be two topological spaces and f : Y → X be a continuous map. Let F be a regular thin family, M ∈ [N]∞ and (ys)s∈F an F -sequence in Y . Suppose that the F -subsequence (ys)s∈F ↾M is convergent to some y ∈ Y . Then the F -subsequence (f (ys))s∈F ↾M is convergent to f (y). Proof. Let UX ∈ TX , with f (y) ∈ UX . By the continuity of f there exists UY ∈ TY such that y ∈ UY and f [UY ] ⊆ UX . Since (ys)s∈F ↾M is convergent to y, there exists m ∈ N such that for every s ∈ F ↾ M with min s ≥ M (m) we have that ys ∈ UY and therefore f (ys) ∈ f [UY ] ⊆ UX . (cid:3) For the rest of this section we shall restrict to F -sequences in metric spaces. Definition 5.3. Let (X, ρ) be a metric space, F a regular thin family, M ∈ [N]∞ and (xs)s∈F an F -sequence in (X, ρ). We will say that the F -subsequence (xs)s∈F ↾M is Cauchy if for every ε > 0 there exists m ∈ N such that for every s1, s2 ∈ F ↾ M with min s1, min s2 ≥ M (m), we have that ρ(xs1 , xs2 ) < ε. Proposition 5.4. Let M ∈ [N]∞, F be a regular thin family and (xs)s∈F be an F -sequence in a complete metric space (X, ρ). Then the F -subsequence (xs)s∈F ↾M is Cauchy if and only if (xs)s∈F ↾M is convergent. Proof. If the F -subsequence (xs)s∈F ↾M is convergent, then it is straightforward that (xs)s∈F ↾M is Cauchy. Concerning the converse we have the following. Suppose that the F -subsequence (xs)s∈F ↾M is Cauchy. Let (sn)n be a sequence in F ↾ M such that min sn → ∞. It is immediate that (xsn )n forms a Cauchy sequence in X. Since (X, ρ) is complete, there exists x ∈ X such that the sequence (xsn )n converges to x. We will show that the F -subsequence (xs)s∈F ↾M converges to x. Indeed, let ε > 0. Since (xs)s∈F ↾M is Cauchy, there exists k0 ∈ N such that for HIGHER ORDER SPREADING MODELS 21 every t1, t2 ∈ F ↾ M with min t1, min t2 ≥ M (k0) we have that ρ(xt1 , xt2 ) < ε 2 . Since the sequence (xsn )n converges to x and min sn → ∞, there exists n0 ∈ N ) < ε/2. Hence for every s ∈ F ↾ M such such that min sn0 ≥ M (k0) and ρ(x, xsn0 , xs) < ε and the that min s ≥ M (k0), we have that ρ(x, xs) ≤ ρ(x, xsn0 proof is completed. (cid:3) ) + ρ(xsn0 Lemma 5.5. Let M ∈ [N]∞, F be a regular thin family and (xs)s∈F an F -sequence in a metric space (X, ρ). Suppose that for every ε > 0 and L ∈ [M ]∞ there exists a plegma pair (s1, s2) in F ↾ L such that ρ(xs1 , xs2 ) < ε. Then the F -subsequence (xs)s∈F ↾M has a further Cauchy subsequence. Proof. Let (εn)n be a sequence of positive reals such that P∞ n=1 εn < ∞. Using Theorem 3.5, we inductively construct a decreasing sequence (Ln)n in [M ]∞, such that for every n ∈ N and for every plegma pair (s1, s2) in F ↾ Ln we have that ρ(xs1 , xs2 ) < εn. Let L′ be a diagonalization of (Ln)n, i.e. L′(n) ∈ Ln for all n ∈ N, and L = {L′(2n) : n ∈ N}. exists n0 ∈ N such that P∞ We claim that the F -subsequence (xs)s∈F ↾L is Cauchy. Indeed let ε > 0. There 2 . Let s0 be the unique initial segment of {L(n) : n ≥ n0} in F . If max s0 = L(k) then we set k0 = k + 1. Then for every s1, s2 ∈ F ↾ L with min s1, min s2 ≥ L(k0), by Theorem 3.9 there exist plegma j=1 in F ↾ L′ from s0 to s1, s2 respectively. Then for i = 1, 2 we paths (s1 have that j=1, (s2 εn < ε n=n0 j )s0 j )s0 ρ(xs0 , xsi ) ≤ ρ(xsi j , xsi j+1 ) < s0−1Xj=0 s0−1Xj=0 εn0+j < ε 2 which implies that ρ(xs1 , xs2 ) < ε. Definition 5.6. Let ε > 0, L ∈ [N]∞, F be a regular thin family and (xs)s∈F an F -sequence in a metric space X. We will say that the subsequence (xs)s∈F ↾L is plegma ε-separated if for every plegma pair (s1, s2) in F ↾ L, ρ(xs1 , xs2 ) > ε. (cid:3) The following proposition is actually a restatement of Lemma 5.5. Proposition 5.7. Let M ∈ [N]∞, F be a regular thin family and (xs)s∈F an F -sequence in a metric space X. Then the following are equivalent. (i) The F -subsequence (xs)s∈F ↾M has no further Cauchy subsequence. (ii) For every N ∈ [M ]∞ there exist ε > 0 and L ∈ [N ]∞ such that the subse- quence (xs)s∈F ↾L is plegma ε-separated. Proof. (i)⇒(ii): Assume that (ii) is not true. Then there is N ∈ [M ]∞ such that for every ε > 0 and L ∈ [N ]∞ there exists a plegma pair (s1, s2) in F ↾ L such that ρ(xs1 , xs2 ) < ε. By Lemma 5.5 the F -subsequence (xs)s∈F ↾N has a further Cauchy subsequence. Since N ⊆ M this means that (xs)s∈F ↾M has a further Cauchy subsequence which is a contradiction. (ii)⇒(i): Suppose that (i) does not hold. Then there exists N ∈ [M ]∞ such that (xs)s∈F ↾N is Cauchy. Let ε > 0 and L ∈ [N ]∞. Then (xs)s∈F ↾L is also Cauchy and therefore (xs)s∈F ↾L is not plegma ε-separated, a contradiction. (cid:3) 5.2. Subordinated F -sequences. By identifying every subset of N with its char- acteristic function, a thin family F becomes a discrete subspace of {0, 1}N (under the usual product topology) with bF being its closure. This in particular yields that every φ : F → (X, T ) is automatically continuous. In this subsection we show 22 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS that for every regular thin family F , M ∈ [N]∞ and ϕ : F ↾ M → (X, T ) such that the closure of ϕ(F ↾ M ) is a compact metrizable subspace of X there exist following definition. Definition 5.8. Let (X, T ) be a topological space, F be a regular thin family, M ∈ [N]∞ and (xs)s∈F be an F -sequence in X. We say that (xs)s∈F ↾M is subordinated L ∈ [M ]∞ and a continuous extension bϕ : bF ↾ L → (X, T ). We start with the (with respect to (X, T )) if there exists a continuous map bϕ : bF ↾ M → (X, T ) with bϕ(s) = xs for every s ∈ F ↾ M . Assume that (xs)s∈F ↾M is subordinated. Then since F is dense in bF , there exists a unique continuous map bϕ : bF ↾ M → (X, T ) witnessing this. Moreover, for the same reason we have {xs : s ∈ F ↾ M } = bϕ(cid:0)bF ↾ M(cid:1), where {xs : s ∈ F ↾ M } is the T -closure of {xs : s ∈ F ↾ M } in X. Therefore {xs : s ∈ F ↾ M } is a countable compact metrizable subspace of (X, T ) with Cantor-Bendixson index at most o(F ) + 1. Another property of subordinated F -sequences is stated in the next proposition. Proposition 5.9. Let (X, T ) be a topological space, F be a regular thin family and (xs)s∈F be an F -sequence in X. Let M ∈ [N]∞ such that (xs)s∈F ↾M is subordinated. (X, T ) is the continuous map witnessing the fact that (xs)s∈F ↾M is subordinated Proof. Via the identity map we may consider the family F as an F -sequence in let (ys)s∈F be the F -sequence in Y , with ys = s for every s ∈ F . As we have already noticed (ys)s∈F converges to the empty set. Then (xs)s∈F ↾M is a convergent F -subsequence in X. In particular, if bϕ : bF ↾ M → then (xs)s∈F ↾M is convergent to bϕ(∅). the metric space Y = bF i.e. Hence, since bϕ : bF ↾ M → (X, T ) is continuous, by Proposition 5.2 we get that (cid:0)bϕ(ys)(cid:1)s∈F ↾M converges to bϕ(∅). Since bϕ(ys) = bϕ(s) = xs for every s ∈ F ↾ M , this means that (xs)s∈F ↾M is convergent to bϕ(∅). Theorem 5.10. Let F be a regular thin family and (xs)s∈F be an F -sequence in a topological space (X, T ). Then for every M ∈ [N]∞ such that {xs : s ∈ F ↾ M } is a compact metrizable subspace of (X, T ) there exists L ∈ [M ]∞ such that (xs)s∈F ↾L is subordinated. (cid:3) Proof. We will use induction on the order of the regular thin family F . If o(F ) = 0 (i.e. the family F is the singleton F = {∅}) the result trivially holds. Let ξ < ω1 and assume that the theorem is true when o(F ) < ξ. We fix a regular thin family F with o(F ) = ξ, an F -sequence (xs)s∈F in a topological space (X, T ) and M ∈ [N]∞ such that {xs : s ∈ F ↾ M } is a compact metrizable subspace of (X, T ). By passing to an infinite subset of M if it is necessary we may also suppose that F is very large in M . Let ρ be a compatible metric for the subspace X0 = {xs : s ∈ F ↾ M }. We shall construct (a) a strictly increasing sequence (mn)n in M , (b) a decreasing sequence M = M0 ⊇ M1 ⊇ ... of infinite subsets of M (c) a sequence bϕn of maps with bϕn : \F(mn) ↾ Mn → X and (d) a decreasing sequence of closed balls (Bn)n in X0 such that for every n ∈ N the following are satisfied: (i) mn = min Mn−1 and Mn ⊆ Mn−1 \ {mn}, (ii) diam Bn < 1/n, HIGHER ORDER SPREADING MODELS 23 (iii) the map bϕn is continuous, (iv) bϕn(u) = x{mn}∪u, for every u ∈ F(mn) ↾ Mn and (v) {bϕn(u) : u ∈ \F(mn) ↾ Mn} ⊆ Bn. We shall present the general inductive step of the above construction so let us assume that the construction has been carried out up to some n ∈ N. We set mn+1 = min Mn. Since F is very large in M we have that G = F(mn+1) = {u ∈ [N]<∞ : mn+1 < u and {mn+1} ∪u ∈ F } is a regular thin family. For each u ∈ G we set yu = x{mn+1}∪u and we form the G-sequence (yu)u∈G. Let M ′ n = Mn \ {mn+1}. Since Y = {yu : u ∈ G ↾ M ′ n} ⊆ {xs : s ∈ F ↾ M }, we have that the closure of Y in (X, T ) is also a compact metrizable subspace of (X, T ). Thus Y is a totally bounded metric space and therefore by Theorem 2.1 and passing to an infinite subset of M ′ n if it is necessary, we may also suppose that there exists a ball Bn+1 of X0 with diam Bn+1 < (n + 1)−1 and such that (28) {yu : u ∈ G ↾ M ′ n} ⊆ Bn+1. n = Mn \ {mn+1} such that the sis, there exists an infinite subset Mn+1 of M ′ Moreover, o(G) = o(cid:0)F(mn+1)(cid:1) < o(F ) = ξ. Hence, by our inductive hypothe- G-subsequence (yu)u∈G↾Mn+1 is subordinated. Let bϕn+1 : bG ↾ Mn+1 → X be the continuous map witnessing this fact. Then bϕn+1(u) = yu = x{mn+1}∪u, for every u ∈ F(mn+1) ↾ Mn+1 and by the continuity of bϕn+1 we have (29) (cid:8)bϕn+1(u) : u ∈ \F(mn+1) ↾ Mn+1(cid:9) ⊆ {yu : u ∈ F(mn+1) ↾ Mn+1} which completes the proof of the inductive step. We set M ′ = {mn : n ∈ N}. Since lim diam Bn = 0 and X0 is a compact metric (28) ⊆ Bn+1, space there exists a strictly increasing sequence (kn)n and x0 ∈ X0 such that (30) lim dist(cid:0)x0, Bkn(cid:1) = 0. We set L = {mkn : n ∈ N} and we define bϕ : bF ↾ L → X as follows. For s = ∅, we set bϕ(∅) = x0. Otherwise, if n is the unique positive integer such that mkn = min s we set bϕ(s) = bϕkn (s \ {min s}) = xt. It is easy to check that bϕ is well defined. To see that bϕ is continuous let (sn)n be a sequence in bF ↾ L and s ∈ bF ↾ L such that (30) we obtain that bϕ(sn) → x0 = bϕ(∅). Otherwise, let mkn0 for all but finitely many n and since bϕkn0 sn → s. If s = ∅ we have that min sn → +∞, thus using condition (v) and equation = min s. Then (sn), (cid:3) , for all but finitely many n. Therefore, bϕ(sn) = bϕkn0 is continuous, bϕ(sn) → bϕ(s). min sn = min s = mkn0 6. F -sequences generating spreading models Let (xn)n be a sequence in a Banach space X generating a spreading model (en)n. It is well known (see [6], [7]) that if (xn)n is norm convergent then the seminorm in the space E generated by (en)n is not a norm. Furthermore if (xn)n is weakly null and seminormalized then (en)n is 1-unconditional. In this section we show that analogues of these results remain true in the higher order setting of ξ-spreading models. We begin with a short review of the basic properties of spreading sequences. (31) Hence, (32) (cid:13)(cid:13)(cid:13) k−1Xj=1 (cid:13)(cid:13)(cid:13)ek − ek+1(cid:13)(cid:13)(cid:13)∗ ajej + akek(cid:13)(cid:13)(cid:13)∗ ak (cid:13)(cid:13)(cid:13) k−1Xj=1 =(cid:13)(cid:13)(cid:13) k−1Xj=1 ajej + akek(cid:13)(cid:13)(cid:13)∗ ajej + akek+1(cid:13)(cid:13)(cid:13)∗ +(cid:13)(cid:13)(cid:13) k−1Xj=1 1 ≤ = 0. ajej + anek+1(cid:13)(cid:13)(cid:13)∗! = 0. 24 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS quence (en)n in E is called spreading if kPn 6.1. Spreading sequences. Let (E, k · k∗) be a seminormed linear space. A se- j=1 ajekj k∗, for every n ∈ N, a1, . . . , an ∈ R and k1 < . . . < kn in N. As we have already mentioned every spreading model of any order of a Banach space is a spreading sequence. In this subsection we shall briefly recall some well known results on spreading sequences that we shall later use (for a more detailed exposition see [1], [5], [6], [7]). We start with the following elementary lemma. j=1 ajejk∗ = kPn Lemma 6.1. Let (E, k · k∗) be a seminormed linear space and (en)n be a spreading sequence in E. Then the following are equivalent. (i) There exist k ∈ N and a1, . . . , ak ∈ R not all zero with kPk (ii) For every n, m ∈ N, ken − emk∗ = 0. j=1 aieik∗ = 0. Proof. The implication (ii)⇒(i) is straightforward. To show the converse let k ∈ N j=1 ajejk∗ = 0. Since (en)n is and a1, . . . , an ∈ R not all zero, such that kPk spreading we may suppose that aj 6= 0 for all 1 ≤ j ≤ n. Moreover notice that Since (en)n is spreading we get that ken − emk∗ = 0 for every n, m ∈ N. (cid:3) Spreading sequences in seminormed linear spaces satisfying (i) or (ii) of the above lemma will be called trivial. As a consequence we have that if (en)n is non trivial then the restriction of the seminorm k · k∗ to the linear subspace generated by (en)n is actually norm. As in [4], we classify the non trivial spreading sequences into the following three categories: (1) the singular, i.e. the non trivial spreading sequences which are not Schauder basic, (2) the unconditional and (3) the conditional Schauder basic spreading sequences, i.e. the spreading sequences which are Schauder basic but not unconditional. Proposition 6.2. Let (en)n be a non trivial spreading sequence. (i) If (en)n is weakly null then it is 1-unconditional. (ii) If (en)n is unconditional then either it is equivalent to the usual basis of ℓ1 Proof. (i) See [1]. (ii) Since (en)n is unconditional there exist C > 0 such that i=1 ei(cid:13)(cid:13)(cid:13) = 0). or it is norm Ces`aro summable to 0 (i.e. limn→∞(cid:13)(cid:13)(cid:13) 1 nPn (cid:13)(cid:13)(cid:13)Pn i=1 εiaiei(cid:13)(cid:13)(cid:13) ≤ C(cid:13)(cid:13)(cid:13)Pn i=1 aiei(cid:13)(cid:13)(cid:13), for every n ∈ N, a1, . . . , an ∈ R and ε1, . . . , εn ∈ {−1, 1}. Also since it is spreading and nontrivial there exists M > 0 such that kenk = M for all n ∈ N. Suppose that (en)n is not Ces`aro summable to zero. Then there exist θ > 0 and a strictly increasing sequence of natural numbers (pn)n such that k 1 i=1 eik > θ, for all i ∈ {1, . . . , pn}. Hence for every n ∈ N there nk = 1 such that x∗ exist x∗ i=1 ei) > θ. For every n ∈ N, we set n with kx∗ n( 1 pnPpn pnPpn HIGHER ORDER SPREADING MODELS 25 In = {1, . . . , pn} and let An = {i ∈ In : x∗ n(ei) > θ θ < x∗ n(cid:16) 1 pn Xi∈In ei(cid:17) = 1 pn x∗ n(cid:16) Xi∈An ei(cid:17) + 1 pn 2 }. Then we have 1 pn n(cid:16) Xi∈In\An ei(cid:17) ≤ x∗ AnC + θ 2 Hence An ≥ θ to show that (en)n is equivalent to the usual basis of ℓ1. a1, . . . , an ∈ R and choose n0 ∈ N such that An0 ≥ n. Then 2C pn which gives that limn→∞ An = +∞. We are now ready Indeed, let n ∈ N, M nXi=1 ai ≥(cid:13)(cid:13)(cid:13) nXi=1 aiei(cid:13)(cid:13)(cid:13) ≥ ≥ 1 C(cid:13)(cid:13)(cid:13) 1 C 1 aiei(cid:13)(cid:13)(cid:13) = C(cid:13)(cid:13)(cid:13) nXi=1 nXi=1 aieAn0 (i)(cid:17) ≥ n(cid:16) nXi=1 aieAn0 (i)(cid:13)(cid:13)(cid:13) nXi=1 θ 2C · x∗ ai. (cid:3) Proposition 6.3. Let (en)n be a singular sequence and let E be the space gener- ated by (en)n. Then there is e ∈ E \ {0} such that (en)n is weakly convergent to e. Moreover if e′ n)n is spreading, 1-unconditional and Ces`aro summable to zero. n = en − e then (e′ Proof. Since (en)n is equivalent to all its subsequences and it is not Schauder basic, every subsequence of (en)n is not Schauder basic. In particular, every subsequence of (en)n cannot be non trivial weak-Cauchy or weakly null. Hence, by Rosenthal's ℓ1-theorem [18] (en)n is weakly convergent to a non zero element e ∈ E. Let e′ n = en − e. To show that (e′ n)n is spreading, let n ∈ N, λ1, ..., λn ∈ R and block subsequence (wm)m of (e′ such that kn < supp(wm) for all m ≥ m0. Then by (33) we have n)n is weakly null we may choose a convex n)n which norm converges to zero. Let m0 ∈ N be i=1 λi = λ. Since (e′ (34) λie′ (cid:13)(cid:13)(cid:13) nXi=1 ki − λwm(cid:13)(cid:13)(cid:13), for all m ≥ m0. Hence, by taking limits, we get that(cid:13)(cid:13)(cid:13)Pn i − λwm(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) n)n is non trivial. Finally, since (e′ that is the sequence (e′ mk, by Lemma 6.1 we have that (e′ n)n is weakly null, by Proposition 6.2 it is also 1-unconditional and norm Ces´aro summable to zero. (cid:3) n)n is spreading. Moreover, since ken − emk = ke′ i(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13)Pn i=1 λie′ n − e′ ki(cid:13)(cid:13)(cid:13), nXi=1 i=1 λie′ λie′ In the following the above decomposition en = e′ n + e of a singular spreading sequence (en)n will be called the natural decomposition of (en)n. 6.2. F -sequences generating non singular spreading models. We start with a characterization of the F -sequences in a Banach space X which generate a trivial spreading model. (33) k1 < ... < kn in N. IfPn i(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13) nXi=1 Generally let Pn λie′ i=1 λi = 0, then notice that nXi=1 λiei(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) nXi=1 λieki(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) nXi=1 λie′ ki(cid:13)(cid:13)(cid:13) 26 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS Theorem 6.4. Let X be a Banach space, F be a regular thin family and (xs)s∈F be an F -sequence in X and M ∈ [N]∞. Let (E, k · k∗) be an infinite dimensional seminormed linear space with Hamel basis (en)n such that (xs)s∈F ↾M generates (en)n as an F -spreading model. Then the following are equivalent: (i) The sequence (en)n is trivial. (ii) For every ε > 0 and every L ∈ [M ]∞, there exists a plegma pair (s1, s2) in F ↾ L such that kxs1 − xs2 k < ε. (iii) The F -subsequence (xs)s∈F ↾M contains a further norm Cauchy subsequence. (iv) There exists x ∈ X such that every subsequence of (xs)s∈F ↾M contains a further subsequence convergent to x. Proof. (i)⇒(ii): Let ε > 0 and L ∈ [M ]∞. Since the F -subsequence (xs)s∈F ↾L also generates (en)n as an F -spreading model (see Remark 2), there exists n0 ∈ N such that for every plegma pair (s1, s2) in F ↾ L with min s1 ≥ L(n0), we have Let (s1, s2) be such a plegma pair. Since (en)n is trivial we have ke1 − e2k∗ = 0 (35) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)xs1 − xs2(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13)e1 − e2(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) and therefore by (35) we obtain that(cid:13)(cid:13)(cid:13)xs1 − xs2(cid:13)(cid:13)(cid:13) < ε. < ε. (ii)⇒(iii): This follows by Lemma 5.5. (iii)⇒(i): Using that (xs)s∈F ↾M contains a further norm Cauchy subsequence, 2 ))n of plegma pairs in F ↾ M such that 1 , sn we easily construct a sequence ((sn sn 1 (1) → ∞ and kxsn 1 − xsn 2 k < 1/n. Then we have (cid:13)(cid:13)e1 − e2k∗ = lim n→∞ (36) kxsn 1 − xsn 2 k = 0. Thus, by Lemma 6.1, we conclude that (en)n is trivial. (iv)⇒(iii): It is straightforward. (i)⇒(iv) Since every subsequence of (xs)s∈F ↾M generates (en)n as an F -spreading model we have that for every L ∈ [M ]∞ (xs)s∈F ↾M generates a trivial spreading model. By the implication (i)⇒(iii) and Proposition 5.4, we have that every sub- sequence of (xs)s∈F ↾M contains a further convergent subsequence. It remains to show that all the convergent subsequences of (xs)s∈F ↾M have a common limit. To this end, let L1, L2 ∈ [M ]∞, x1, x2 ∈ X such that (xs)s∈F ↾Li converges to xi for i ∈ {1, 2} and let ε > 0. Hence there exists n0 ∈ N such that for every s ∈ F ↾ L1 and t ∈ F ↾ L2 with min s ≥ L1(n0) and min t ≥ L2(n0) we have (37) kx1 − xsk, kx2 − xtk < ε 3 Since (xs)s∈F ↾M generates the trivial sequence (en)n as an F -spreading model, we may also assume that (38) (39) for every plegma pair (s1, s2) in F ↾ M with min s1 ≥ M (n0). It is easy to see that we can choose s1 ∈ F ↾ L1 with min s1 ≥ L1(n0) and s2 ∈ F ↾ L2 with min s2 ≥ L2(n0) such that (s1, s2) is a plegma pair. Then by (37) and (38) we have , < ε 3 (cid:13)(cid:13)xs1 − xs2(cid:13)(cid:13) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)xs1 − xs2(cid:13)(cid:13) −(cid:13)(cid:13)e1 − e2(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:13)(cid:13)(cid:13)x1 − x2(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)x1 − xs1(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13)xs1 − xs2(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13)x2 − xs2(cid:13)(cid:13)(cid:13) < ε. HIGHER ORDER SPREADING MODELS Since (39) holds for every ε > 0 we get that x1 = x2. The proof is complete. 27 (cid:3) We proceed to present a sufficient condition for an F -sequence to generate a Schauder basic spreading model. We need the next definition. Definition 6.5. Let A be a countable seminormalized subset of a Banach space X. We say that A admits a Skipped Schauder Decomposition (SSD) if there exist C ≥ 1 and a pairwise disjoint sequence (Ak)k of finite subsets of A such that the following are satisfied. k=1Ak = A. (i) ∪∞ (ii) For every N ∈ [N]∞ not containing two successive integers and every se- quence (xk)k∈N with xk ∈ Ak for all n ∈ N , (xk)k∈N is a Schauder basic sequence of constant C. The following proposition is well known but for the sake of completeness we outline its proof. Proposition 6.6. Let (xn)n be a seminormalized weakly null sequence in a Banach space X. Then for every ε > 0 the set A = {xn : n ∈ N} admits a SSD with constant C = 1 + ε. Proof. We may assume that X has a Schauder basis (en)n with basis constant K=1 (for example we may assume that X = C[0, 1]). By induction and using the sliding hump argument, we define (1) a partition (Fn)n of N into finite pairwise disjoint sets and (2) a sequence (yn)n of finitely supported vectors in X such that the following are fulfilled: (i) For every k ∈ N and n ∈ Fk, we have kxn − ynk < ε/2k and (ii) for every k2 > k1 with k2 − k1 > 1, n1 ∈ Fk1 and n2 ∈ Fk2 , we have max supp(yn1) < min supp(yn2). Setting Ak = {xn : n ∈ Fk}, k ∈ N, it is easy to check that (Ak)k satisfies conditions (i) and (ii) of Definition 6.5. (cid:3) Theorem 6.7. Let A be a subset of a Banach space X. If A admits a SSD with constant C, then every non trivial spreading model of any order of A is Schauder basic of constant C. Proof. Let 1 ≤ ξ be a countable ordinal and (en)n be a non trivial spreading model of A of order ξ. Let F be a regular thin family with o(F ) = ξ, (xs)s∈F be an F -sequence in A and M ∈ [N]∞ such that (xs)s∈F ↾M generates (en)n as an F - spreading model. Let (Ak)k be a partition of A satisfying condition (ii) of Def. 6.5. Finally, let ϕ : F ↾ M → N, defined by ϕ(s) = k if xs ∈ Fk. Observe that ϕ is hereditarily nonconstant in M . Indeed, otherwise there exists L ∈ [M ]∞ and k0 ∈ N such that xs ∈ Fk0 for every s ∈ F ↾ L and s ∈ F ↾ L. Since Fk0 is finite, by Proposition 2.6 there exists N ∈ [L]∞ such that (xs)s∈F ↾N is constant. By part (ii) of Remark 2, the F -sequence (xs)s∈F ↾N also generates (en)n as an F -spreading model. But then, since (xs)s∈F ↾N is constant, the sequence (en)n should be trivial which is a contradiction. Hence ϕ is hereditarily nonconstant in M and therefore by Corollary 3.12 there exists N ∈ [M ]∞ such that ϕ(s2) − ϕ(s1) > 1 for every plegma pair (s1, s2) in F ↾ N . By the SSD property of A we have that 28 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS for every 1 ≤ m < l ∈ N and for every (40) k mXj=1 ajxsj k ≤ Ck ajxsj k, lXj=1 for every plegma l-tuple (sj)l (en)n is a Schauder basic sequence with constant C. j=1 in F ↾ N and a1, ..., al ∈ R. This easily yields that (cid:3) 6.3. Spreading models generated by subordinated F -sequences. Let (xn)n be a weakly convergent sequence in a Banach space X which is not norm Cauchy and assume that it generates a spreading model (en)n. It is well known (see [6], [7]) that (en)n is either an unconditional or a singular spreading sequence. In [4] we extended this fact for subordinated k-sequences. Here we will show that similar results also hold true for F -sequences where F is a regular thin family. 6.3.1. Unconditional spreading models. Let n ∈ N and for every 1 ≤ i ≤ n let Fi ⊆ [N]<∞. We will say that (Fi)n i=1 is completely plegma connected if for every choice of si ∈ Fi, the n-tuple (si)n i=1 is a plegma family. Also for a subset A of a Banach space X, convA denotes the convex hull of A. Lemma 6.8. Let X be a Banach space, n ∈ N, F1, . . . , Fn be regular thin families, family (Fi)n every i = 1, ..., n. and L ∈ [N]∞. Assume that for every i = 1, ..., n, there exists a continuous map bϕi : cFi ↾ L → (X, w). Then for every ε > 0 there exists a completely plegma connected i=1 such that Fi ⊆ [Fi ↾ L]<∞ and dist(cid:16)bϕi(∅), conv bϕi(Fi)(cid:17) < ε, for i=1(cid:1) := max{o(Fi) : 1 ≤ i ≤ n}. Proof. We will use induction on o(cid:0)(Fi)l i=1(cid:1) = 0, i.e. Fi = {∅} for all 1 ≤ i ≤ n the result follows trivially. Let o(cid:0)(Fi)l i=1(cid:1) < ξ. Let n ∈ N, 1 ≤ ξ < ω1 and suppose that the lemma holds true if o(cid:0)(Fi)n i=1(cid:1) = ξ and assume L ∈ [N]∞ and F1, . . . , Fn be regular thin families with o(cid:0)(Fi)l that for every 1 ≤ i ≤ n, there exists a continuous map bϕi :cFi ↾ L → (X, w). every singleton {l} with l ∈ L belongs to cFi. By the continuity of bϕi, we get that w- liml∈Lbϕi({l}) = bϕi(∅). By Mazur's theorem, we may choose a finite subset Fix i ∈ {1, ..., n}. We may suppose that Fi is very large in L and therefore mi} of L such that 1 < ... < li Λi = {li If (41) dist(cid:16)bϕi(∅), conv bϕi(Λi)(cid:17) < ε/2, for every 1 ≤ i ≤ n. We may also assume that (42) Λ1 < ... < Λn. j < s and {li Let Λ = ∪n We set Gi i=1Λi and let M = {l ∈ L : l > max Λ}. Fix 1 ≤ i ≤ n and 1 ≤ j ≤ mi. j = (Fi)(li j ↾ j is a continuous j ) = {s ∈ [N]<∞ : li map and since o(Gi our inductive assumption we may choose a completely plegma connected family M → (X, w), defined by bϕi j} ∪ s). Notice that bϕi j (s) = bϕi({li i=1(cid:17) < o(cid:0)(Fi)n j ) < o(Fi), o(cid:16)(cid:0)(cid:0)Gi j=1(cid:1)n j(cid:1)mi (cid:0)(cid:0)Gi j(cid:1)mi j=1(cid:1)n j )(cid:17) < ε/2, dist(cid:16)bϕi j(∅), conv bϕi j} ∪ s ∈ Fi} and let bϕi j : cGi i=1(cid:1) = ξ. Therefore using i=1 such that Gi j ↾ M ]<∞ and j ⊆ [Gi j(Gi (43) HIGHER ORDER SPREADING MODELS 29 (44) for every 1 ≤ i ≤ n and 1 ≤ j ≤ mi. j}∪s : s ∈ Gi Hence, equation (43) translates to for every 1 ≤ i ≤ n and 1 ≤ j ≤ mi. i=1 is completely j (Gi j ). For every 1 ≤ i ≤ n and 1 ≤ j ≤ mi we set F i j(cid:9). By equation j) = bϕi(F i j =(cid:8){li j=1(cid:1)n j(cid:1)mi i=1, we easily see that(cid:0)(cid:0)F i j(∅) = bϕi({li j}) and bϕi j )(cid:17) < ε/2, j=1(cid:1)n j(cid:1)mi (42) and the choice of(cid:0)(cid:0)Gi plegma connected. Moreover, observe that bϕi dist(cid:16)bϕi({li j}), conv bϕi(F i For every 1 ≤ i ≤ n, let Fi =Smi By (44) we have dist(cid:0)x, conv bϕi(Fi)(cid:1) < ε/2, for every x ∈ conv bϕi(Λi) and therefore by (41) we conclude that dist(cid:16)bϕi(∅), conv bϕi(Fi)(cid:17) < ε. i=1 is a completely plegma connected family with Fi ⊆ [Fi ↾ L]<∞, for every i = 1, ..., n. Finally, fix 1 ≤ i ≤ n. Theorem 6.9. Let X be a Banach space, F be a regular thin family and L ∈ [N]∞. Let (xs)s∈F ↾L be an F -subsequence in X generating an F -spreading model (en)n. Also assume that (xs)s∈F ↾L is seminormalized, subordinated (with respect to the weak topology of X) and weakly null. Then (en)n is an 1-unconditional spreading sequence. j . Clearly (Fi)n j=1 F i (cid:3) Proof. We first show that (en)n is non trivial. Indeed, otherwise by Theorem 6.4 there exists M ∈ [L]∞ and x0 ∈ X such that the F -subsequence (xs)s∈F ↾M is norm convergent to x0. Since M ⊆ L, (xs)s∈F ↾M is also weakly null and therefore x0 = 0. But this is a contradiction since (xs)s∈F ↾L is seminormalized. We proceed to show that (en)n is 1-unconditional. Fix n ∈ N, 1 ≤ p ≤ n and a1, . . . , an ∈ [−1, 1]. It suffices to show that for every ε > 0 we have Indeed, fix ε > 0. Since (xs)s∈F ↾L generates (en)n as an F -spreading model, by passing to a final segment of L if it is necessary we may assume that + ε. i6=p nXi=1 aiei(cid:13)(cid:13)(cid:13)∗ aiei(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < nXi=1 aiei(cid:13)(cid:13)(cid:13)∗ <(cid:13)(cid:13)(cid:13) and (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) nXi=1 ε 3 aixsi(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) nXi=1 aiei(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) < ε 3 , (45) (cid:13)(cid:13)(cid:13) (46) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) nXi=1 i6=p aixsi(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) nXi=1 i6=p for every plegma n-tuple (si)n i=1 in F ↾ L. Since (xs)s∈F ↾L is subordinated with for all i = 1, ..., n), there exist a completely plegma connected family (Fi)n sequence (xi)n respect to the weak topology, there exists a continuous map bϕ : F ↾ L → (X, w) such that bϕ(s) = xs for every s ∈ F ↾ L. Since (xs)s∈F ↾L is weakly convergent to bϕ(∅) we have that bϕ(∅) = 0. Therefore by Lemma 6.8 (for Fi = F and bϕi = bϕ, i=1 in X such that Fi ⊆ [F ↾ L]<∞, xi ∈ conv bϕ(Fi) and kxik < ε/3, for every 1 ≤ i ≤ n. Let (µs)s∈Fp be a sequence in [0, 1] such thatPs∈Fp and xp =Ps∈Fp µsbϕ(t) and for each i 6= p choose si ∈ Fi. By the above we have and kxpk = kPs∈Fp that for every s ∈ Fp the n-tuple (s1, . . . , sp−1, s, sp+1, . . . , sn) is a plegma family µsxsk < ε 3 . i=1 and a µs = 1 30 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS Therefore by (46) we have nXi=1 i6=p (cid:13)(cid:13)(cid:13) aiei(cid:13)(cid:13)(cid:13)∗ i6=p aixsi(cid:13)(cid:13)(cid:13) + ≤(cid:13)(cid:13)(cid:13) nXi=1 µs(cid:13)(cid:13)(cid:13) nXi=1 ≤ Xs∈Fp aiei(cid:13)(cid:13)(cid:13)∗ =(cid:13)(cid:13)(cid:13) nXi=1 i6=p + ε. The proof is complete. i6=p ε 3 ≤(cid:13)(cid:13)(cid:13) nXi=1 aixsi + apxs(cid:13)(cid:13)(cid:13) + aixsi + ap Xs∈Fp ≤ Xs∈Fp 2ε 3 ε 3 µsxs(cid:13)(cid:13)(cid:13) + ap aiei(cid:13)(cid:13)(cid:13)∗ µs(cid:16)(cid:13)(cid:13)(cid:13) nXi=1 + ε 3 + ε 3(cid:17) + 2ε 3 (cid:3) 6.3.2. Singular or isomorphic to ℓ1 spreading models. We proceed to show an ana- logue of Theorem 6.9 for subordinated F -sequences which are not weakly null. We will need the following lemma. are both Ces´aro summable to zero. Suppose that for every n ∈ N and λ1, . . . , λn Lemma 6.10. Let (en)n and (een)n be two non trivial spreading sequences which withPn i=1 λi = 0, we have that (47) to zero, we have limm→∞ Then the map en →een extends to a linear isometry from < (en)n > onto < (een)n >. Proof. Let n ∈ N and λ1, . . . , λn ∈ R. Since (en)n (resp. (een)n) is Ces´aro summable j=1een+j = 0). Let λ =Pn j=1 en+j = 0 (resp. limm→∞ λ m = 0 and therefore, mPm 1 1 nXi=1 =(cid:13)(cid:13)(cid:13) . λieei(cid:13)(cid:13)(cid:13)∗∗ nXi=1 λiei(cid:13)(cid:13)(cid:13)∗ (cid:13)(cid:13)(cid:13) mPm i=1 λi −Pm m→∞(cid:13)(cid:13)(cid:13) nXi=1 m→∞(cid:13)(cid:13)(cid:13) nXi=1 j=1 (47) = lim = lim i=1 λi. ThenPn λiei(cid:13)(cid:13)(cid:13)∗ (cid:13)(cid:13)(cid:13) nXi=1 λiei − λ m λieei − en+j(cid:13)(cid:13)(cid:13)∗ mXj=1 mXj=1een+j(cid:13)(cid:13)(cid:13)∗∗ λ m nXi=1 =(cid:13)(cid:13)(cid:13) λieei(cid:13)(cid:13)(cid:13)∗∗ (cid:3) The next lemma is from [4]. We reproduce it for the sake of completeness. i=1 ai = 0. i=1 aieik = kPn Lemma 6.11. Let X be a Banach space, F be a regular thin family and (xs)s∈F ↾L be an F -subsequence in X. Let x0 ∈ X and set x′ s = xs − x0, for all s ∈ F ↾ L. Assume that (xs)s∈F ↾L and (x′ respectively. Then the following hold. s)s∈F ↾L generate F -spreading models (en)n and (een)n (i) kPn i=1 aieeik, for every n ∈ N and a1, . . . , an ∈ R with Pn (ii) The sequence (en)n is trivial if and only if (een)n is trivial. (iii) The sequence (en)n is equivalent to the usual basis of ℓ1 if and only if (een)n i=1 ai = 0, we have Pn with Pn si . Since (en)n and (een)n are Proof. (i) Notice that for every n ∈ N, s1, ..., sn in F ↾ L and a1, . . . , an ∈ R generated by (xs)s∈F ↾L and (xs)s∈F ↾L the result follows. i=1 aixsi = Pn is equivalent to the usual basis of ℓ1. i=1 aix′ HIGHER ORDER SPREADING MODELS 31 (ii) It follows by part (i) and Lemma 6.1. i=1 a′ i=1 a′ 1, . . . , a′ i/2 and an+i = −a′ to the usual basis of ℓ1. n ∈ N and a′ ai = a′ (iii) We fix ε > 0. If (een)n is not equivalent to the usual basis of ℓ1 then there exist ieeik < ε. Setting i=1 aieik = kP2n kP2n i = 1 and kPn i/2, for all 1 ≤ i ≤ n, we haveP2n i=1 ai = 0 and therefore, i=1 ai = 1, (en)n is also not equivalent (cid:3) n ∈ R such that Pn i=1 aieeik < ε. SinceP2n Theorem 6.12. Let X be a Banach space, F be a regular thin family and L ∈ [N]∞. Let (xs)s∈F ↾L be an F -subsequence in X generating a non trivial F -spreading model (en)n. Also assume that (xs)s∈F ↾L is subordinated and let x0 be the weak-limit of (xs)s∈F ↾L. Finally, let x′ s = xs − x0, for every s ∈ F ↾ L. If x0 6= 0 then exactly one of the following holds. (i) The sequence (en)n as well as every spreading model of (x′ s)s∈F ↾L is equiv- alent to the usual basis of ℓ1. (ii) The sequence (en)n is singular and if en = e′ n+e is its natural decomposition n)n is the unique (up to isometry) F -spreading model then kek = kx0k and (e′ of (x′ s)s∈F ↾L. s)s∈F ↾L. By passing to an infinite If (en)n is equivalent to the usual basis of ℓ1 then by Lemma 6.11, we have that subset of L if it is necessary we may assume that (x′ F -spreading model. Proof. Let (een)n be an F -spreading model of (x′ s)s∈F ↾L generates (een)n as an the same holds for (een)n and hence (i) is satisfied. Otherwise, again by Lemma 6.11, (een)n is also non trivial and not equivalent to the ℓ1-basis. Let us denote by k · k∗ (resp. k · k∗∗) the norm of the space generated by (en)n (resp. (een)n). Since (een)n is non trivial, we have that keenk∗∗ > 0 and therefore (by passing to a final Theorem 6.9, (een)n is 1-unconditional. Moreover, since (een)n is not equivalent to the usual basis of ℓ1, by Proposition 6.2 (ii), we conclude that (een)n is norm Ces`aro segment of L if it is necessary) we may assume that (x′ It is also easy to see that (x′ s)s∈F ↾L is seminormalized. s)s∈F ↾M is subordinated and weakly null. Therefore by summable to zero. Hence, by part (i) of Lemma 6.11, we have For every n ∈ N choose a sequence (sn Since xs − x0 = x′ s, for every s ∈ F ↾ L, we have j=1 ∈ Plmn(F ↾ L) such that min sn 1 ≥ L(n). n + e be the natural decomposition of (en)n. By (50) and the fact that (e′ it is singular. Let By (48) and (50), we get that (en)n is not Schauder basic, i.e. en = e′ n)n is Ces`aro summable to zero, we have that kek = kx0k. To complete the proof it n)n are isometrically equivalent. By Lemma 6.10 remains to show that (een)n and (e′ (48) (49) (50) Therefore ej − 1 n lim 1 n nXj=1 n→∞(cid:13)(cid:13)(cid:13) n→∞(cid:13)(cid:13)(cid:13)x0 − n→∞(cid:13)(cid:13)(cid:13) 1 n lim lim j )n = lim 2nXj=n+1 ej(cid:13)(cid:13)(cid:13)∗ j(cid:13)(cid:13)(cid:13) = lim n→∞(cid:13)(cid:13)(cid:13) n→∞(cid:13)(cid:13)(cid:13) ej(cid:13)(cid:13)(cid:13)∗ n→∞(cid:13)(cid:13)(cid:13) nXj=1 nXj=1 = lim 1 n 1 n 1 n 1 n 2nXj=n+1eej(cid:13)(cid:13)(cid:13)∗∗ nXj=1een(cid:13)(cid:13)(cid:13)∗∗ nXj=1eej − j(cid:13)(cid:13)(cid:13) = lim n→∞(cid:13)(cid:13)(cid:13) j(cid:13)(cid:13)(cid:13) = kx0k > 0. 1 n xsn x′ sn xsn nXj=1 nXj=1 1 n = 0. = 0. 32 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS it suffices to show that , λieei(cid:13)(cid:13)(cid:13)∗∗ i=1 λ1 = 0. (51) λie′ (cid:13)(cid:13)(cid:13) nXi=1 i(cid:13)(cid:13)(cid:13)∗ =(cid:13)(cid:13)(cid:13) nXi=1 for every n ∈ N and λ1, . . . λn ∈ R with Pn λ1, . . . λn ∈ R with Pn λiei(cid:13)(cid:13)(cid:13)∗ (cid:13)(cid:13)(cid:13) nXi=1 nXi=1 k→∞(cid:13)(cid:13)(cid:13) such that limk→∞ min sk and the proof is complete. =(cid:13)(cid:13)(cid:13) i(cid:13)(cid:13)(cid:13)∗ nXi=1 = lim λixsk λie′ i=1 λi = 0. For each k ∈ N choose (sk 1 = +∞. Then Indeed, fix n ∈ N and j=1 ∈ Plmn(F ↾ L) j )n i(cid:13)(cid:13)(cid:13) = lim k→∞(cid:13)(cid:13)(cid:13) nXi=1 λix′ sk i(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) nXi=1 λieei(cid:13)(cid:13)(cid:13)∗∗ (cid:3) 6.3.3. Weakly relatively compact F -sequences. Let X be a Banach space and ξ < ω1. By SMwrc (X) we will denote the set of all spreading sequences (en)n such that there exists a weakly relatively compact subset W of X which admits (en)n as a ξ-spreading model. We also set ξ SMwrc(X) = [ξ<ω1 SMwrc ξ (X). w Hence (en)n ∈ SMwrc (X) if and only if there exists an F -sequence (xs)s∈F such ξ is a weakly compact subset of X and for some L ∈ [N]∞, that {xs : s ∈ F } (xs)s∈F ↾L generates (en)n as an F -spreading model. The F -sequences with weakly relatively compact range will be called weakly relatively compact ("wrc" in short). The following proposition says that every wrc F -sequence always contains a subor- dinated subsequence. Proposition 6.13. Let X be a Banach space, F be a regular thin family and (xs)s∈F be a weakly relatively compact F -sequence in X. Then for every M ∈ [N]∞ there exists L ∈ [M ]∞ such that the F -subsequence (xs)s∈F ↾L is subordinated with respect to the weak topology. ξ ζ w (X) ⊆ SMwrc Proof. Let M ∈ [N]∞. Since the weak topology on every separable weakly compact subset of a Banach space is metrizable, we have that {xs : s ∈ F } is compact metrizable. By Theorem 5.10 the result follows. (cid:3) Proposition 6.14. Let X be a Banach space, ξ < ω1 and (en)n ∈ SMwrc (X). Then for every regular thin family G with o(G) ≥ ξ there exist a weakly relatively compact G-sequence (wt)t∈G in X and L ∈ [N]∞ such that (wt)t∈G↾L is subordi- nated with respect to the weak topology and generates (en)n as a G-spreading model. Consequently SMwrc Proof. Since (en)n ∈ SMwrc (X) there exists a weakly relatively compact subset A of X such that A admits (en)n as a ξ-spreading model. Hence there exists a regular thin family F of order ξ, an F -sequence (xs)s∈F in A and M ∈ [N]∞ such that (xs)s∈F ↾M generates (en)n as an F -spreading model. By Lemma 4.6 there exist a G-sequence (wt)t∈G and N ∈ [N]∞ such that (wt)t∈G↾N generates (en)n as a G-spreading model and moreover {wt : t ∈ G} ⊆ {xs : s ∈ F } ⊆ A. Hence (wt)t∈G is a weakly relatively compact G-sequence. By Proposition 6.13 there exists L ∈ [N ]∞ such that (wt)t∈G↾L is subordinated with respect to the weak topology. Clearly (wt)t∈G↾L also generates (en)n as a G-spreading model and the proof is complete. (cid:3) (X), for every 1 ≤ ζ < ξ < ω1. ξ ξ HIGHER ORDER SPREADING MODELS 33 Proposition 6.14 implies that every (en)n in SMwrc(X) is generated by a subor- dinated F -subsequence. Hence, by Theorems 6.9 and 6.12 we obtain the following. Corollary 6.15. Let X be a Banach space, F be a regular thin family and (xs)s∈F be a weakly relatively compact F -sequence. Let (en)n be a spreading sequence and assume that (xs)s∈F admits (en)n as an F -spreading model. Then exactly one of the following holds: (i) The sequence (en)n is trivial. (ii) The sequence (en)n is singular. x0 ∈ X such that if en = e′ the F -subsequence (x′ generates the sequence (e′ In this case there exist L ∈ [N]∞ and n + e is the natural decomposition of (en)n then s = xs − x0 for all s ∈ F ↾ L, s)s∈F ↾L, defined by x′ n)n as an F -spreading model and kx0k = kek. (iii) The sequence (en)n is Schauder basic. In this case (en)n is unconditional. 6.4. F -sequences generating singular spreading models. Let X be a Banach space and (xn)n be a sequence in X which generates a singular spreading model (en)n and let en = e′ n + e be the natural decomposition of (en)n. It can be shown that there exists x ∈ X \ {0} such that kxk = kek and setting x′ n)n is the unique spreading model of (x′ n)n. In the following we will present an extension of this fact for F -sequences in a Banach space X. We start with the next lemma. n = xn − x, (e′ Lemma 6.16. Let F be a regular thin family, M ∈ [N]∞ and (xs)s∈F be an F - sequence in a Banach space X such that (xs)s∈F ↾M generates a singular F -spreading model (en)n. Then there exists L ∈ [M ]∞ satisfying the next property. For every ε > 0 there exists m0 ∈ N such that (52) 1 n nXj=1 xsj − 1 m mXj=1 (cid:13)(cid:13)(cid:13) xtj(cid:13)(cid:13)(cid:13) < ε, for every n, m ≥ m0 and (sj)n t1(1) ≥ L(m). j=1, (tj)m j=1 ∈ Plm(F ↾ L) with s1(1) ≥ L(n) and Proof. Initially we notice that a weaker version of the lemma holds true, that is for every ε > 0 there exists k0 ∈ N such that for every n, m ≥ k0 and every (sj)n+m j=1 ∈ Plm(F ↾ M ) with s1(1) ≥ M (n + m), we have Indeed, let ε > 0. Since (en)n is singular, it is weakly convergent to some e and moreover setting e′ n)n is Ces´aro summable to zero. Hence < ε/4, for all n ≥ n0. Hence for n = en −e, the sequence (e′ i=1 e′ (53) 1 n (cid:13)(cid:13)(cid:13) 1 m xsj − nXj=1 we may choose n0 ∈ N such that (cid:13)(cid:13)(cid:13) 1 nPn =(cid:13)(cid:13)(cid:13) en+i(cid:13)(cid:13)(cid:13)∗ every n, m ≥ n0, we have mXi=1 nXi=1 (cid:13)(cid:13)(cid:13) 1 m ei − (54) 1 n xsn+j(cid:13)(cid:13)(cid:13) < ε. mXj=1 i(cid:13)(cid:13)(cid:13)∗ nXi=1 e′ i − 1 m mXi=1 1 n Since (xs)s∈F ↾M generates (en)n as an F -spreading model we can find k0 ≥ n0 such that for every n, m ≥ k0 and every (sj )n+m j=1 ∈ Plm(F ↾ M ) with s1(1) ≥ M (n + m) equation (53) is satisfied. < ε/2 e′ n+i(cid:13)(cid:13)(cid:13)∗ 34 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS Let (εk)k be a sequence of positive real numbers such thatPk εk < +∞. By the above we can choose an increasing sequence (nk)k in N such that for every k ∈ N, n, m ≥ nk and (sj)n+m j=1 ∈ Plm(F ↾ M ) with s1(1) ≥ M (n + m), it holds that (55) xsj − 1 m 1 n nXj=1 (cid:13)(cid:13)(cid:13) mXj=1 xsn+j(cid:13)(cid:13)(cid:13) < εk. We may also assume that F is very large in M and 2nk < nk+1, for every k ∈ N. We set L = {M (2nk + nk+1) : k ∈ N} and we shall show that L satisfies the conclusion of the lemma. To this end we shall use an appropriate map sending each s ∈ F ↾ L to a plegma family in F ↾ M . First, for every s ∈ F ↾ L and p = 1, ..., s, let k(s(p)) be the unique positive integer k satisfying s(p) = L(k) = M (2nk +nk+1). We define Φ : F ↾ L → Plm(F ↾ M ) as follows. For every s ∈ F ↾ L, we assign j is the unique element of F ↾ M , where vs the nk(s(1))-tuple Φ(s) = (cid:0)vs satisfying j(cid:1)nk(s(1)) j=1 vs j ⊑(cid:8)M(cid:0)2nk(s(p)) + nk(s(p))+1 − nk(s(1)) + j(cid:1) : p = 1, . . . , s(cid:9). j , j = 1, ..., nk(s(1)), follows easily from the fact that F is regular (56) The existence of vs thin and very large in M . Below we state some useful properties of Φ. Their verification is straightforward. (P1) For every s ∈ F ↾ L, Φ(s) ∈ Plmnk (F ↾ M ), vs 1(1) > M (nk + nk+1) and vs nk = s, where k = k(s(1)), (P2) for every (s1, s2) ∈ Plm2(F ↾ L), the concatenation Φ(s1)⌢Φ(s2) belongs to Plm(F ↾ M ). We are now ready to prove that L is actually the desired set. Fix a positive integer k and let us denote by s the unique element of F ↾ L such that s ⊑ {L(i) : i ≥ k}. Notice that s(1) = L(k) = M (2nk + nk+1) and therefore k(s(1)) = k. Also let mk = max{nk, k + s + 1}. We claim that (57) 1 nk nkXj=1 xvs j − 1 m mXj=1 (cid:13)(cid:13)(cid:13) xtj(cid:13)(cid:13)(cid:13) < k+sXl=k εl, for every m ≥ mk and (tj )m j=1 ∈ F ↾ L with t1(1) ≥ L(m). 1 , . . . , vwl+1 1 , . . . , vwl Indeed, let m ≥ mk and (tj )m j=1 ∈ F ↾ L with t1(1) ≥ L(m). Notice that max s = L(k + s − 1) < L(m) ≤ t1(1) = min t1. Hence, by Theorem 3.9 there exists a plegma path (wl)l0 l=0 in F ↾ L from w0 = s to wl0 = t1 of length l0 = s. Notice that k(wl(1)) ≥ k + l which implies that nk(wl(1)) ≥ nk+l and therefore length nk+l + nk+l+1 with vwl Hence by (55) we get 1 , . . . , vwl (cid:0)vwl we have that (cid:0)vwl (cid:13)(cid:13)(cid:13) nk+l , vwl+1 1 (1) > M (nk(wl(1)) + nk(wl(1)+1)) > M (nk+l + nk+l+1). nk+l(cid:1) is a subfamily of Φ(wl). Thus, by properties (P1) and (P2) above, nk+l+1(cid:1) is a plegma family in F ↾ L of nk+l+1Xj=1 nk+sXj=1 j (cid:13)(cid:13)(cid:13) < (cid:13)(cid:13)(cid:13) < εk+l, k+s−1Xl=k nk+lXj=1 nkXj=1 for every l = 0, ..., l0 − 1. Thus, nk+l+1 nk+s xvs j − xv wl+1 j (cid:13)(cid:13)(cid:13) 1 nk xv wl j − xv t1 1 nk+l (59) (58) εl. 1 1 HIGHER ORDER SPREADING MODELS 35 1 − (60) xv t1 j 1 m Now (57) follows by (59) and (60). Similarly, since m > k + s = k + l0 we have that nm > nk+s. Also since t1(1) ≥ L(m) = M (2nm + nm+1) we have that k(t1(1)) ≥ m. Hence nk(t1(1)) ≥ Finally, by (57) and a triangle inequality we obtain that M (2nm + nm+1) ≥ M (nk+s + m) and so again by (55), we have 1 , ..., vt1 nm > nk+s which implies that (cid:0)vt1 Therefore,(cid:0)vt1 nk+sXj=1 nk+s 1 , ..., vt1 nk+s(cid:1) is a proper subfamily of Φ(t1). nk+s, t1, ..., tm(cid:1) is a plegma family in F ↾ L. Moreover t1(1) ≥ xtj(cid:13)(cid:13)(cid:13) < εk+s. (cid:13)(cid:13)(cid:13) mXj=1 xtj(cid:13)(cid:13)(cid:13) < 2 k+sXl=k mXj=1 nXj=1 (cid:13)(cid:13)(cid:13) j=1, (tj )m for every k ∈ N, n, m ≥ mk and (sj)n j=1 ∈ Plm(F ↾ L) with s1(1) ≥ L(n) (cid:3) and t1(1) ≥ L(m). SincePk εk < +∞ the proof is complete. Theorem 6.17. Let F be a regular thin family, M ∈ [N]∞ and (xs)s∈F be an F - sequence in a Banach space X such that (xs)s∈F ↾M generates a singular F -spreading model (en)n. Let en = e′ n + e be the natural decomposition of (en)n. Then there exist x ∈ X with kxk = kek∗ and L ∈ [M ]∞ such that setting x′ s = xs − x the F -subsequence (x′ n)n as a unique (up to isometry) F -spreading model. s)s∈F ↾L admits (e′ xsj − 1 m (61) 1 n εl, Proof. We start by determining the element x ∈ X. Let L ∈ [M ]∞ satisfying Lemma 6.16. For every k ∈ N we set Ak =n 1 n nXi=1 xsi : (si)n i=1 ∈ Plm(F ↾ L) and s1(1) ≥ n ≥ ko. Clearly the sequence (Ak)k is decreasing and by Lemma 6.16, diam(Ak) → 0. Therefore there exists a unique x ∈ X such that ∩∞ k=1Ak = {x}. We continue to show that kek = kxk. Notice that by the choice of x, we have that for every ε > 0 there exists n0 ∈ N such that for all n ≥ n0 For each n ∈ N we pick (sn have i )n i=1 ∈ Plm(F ↾ L), with sn i (1) ≥ L(n). By (62), we Also, since (xs)s∈F ↾L generates (en)n as an F -spreading model, we get Moreover, since (e′ n)n is Ces`aro summable to zero, we have = 0. (62) (63) (64) (65) lim i 1 n 1 n xsn nXj=1 xsj − x(cid:13)(cid:13)(cid:13) < ε. (cid:13)(cid:13)(cid:13) n (cid:13)(cid:13)(cid:13) − x(cid:13)(cid:13)(cid:13) = 0. nXi=1 ei(cid:13)(cid:13)(cid:13)∗(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) i(cid:13)(cid:13)(cid:13) −(cid:13)(cid:13)(cid:13) nXi=1 nXi=1 n (cid:13)(cid:13)(cid:13)Pn − e(cid:13)(cid:13)(cid:13)∗ i=1 ei n = 0. xsn 1 n lim 1 n lim n (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13) 36 S. A. ARGYROS, V. KANELLOPOULOS, AND K. TYROS Hence, kek∗ (65) = lim 1 n nXi=1 n (cid:13)(cid:13)(cid:13) ei(cid:13)(cid:13)(cid:13)∗ (64) = lim xsn (63) = kxk. 1 n nXi=1 n (cid:13)(cid:13)(cid:13) i(cid:13)(cid:13)(cid:13) We proceed now to show that (e′ n)n is the unique F -spreading model of (x′ where x′ s = xs − x, s ∈ F ↾ L. Let N ∈ [L]∞ such that (x′ s)s∈F ↾L, s)s∈F ↾N generates an (66) F -spreading model (een)n. We will show that (een)n is isometric to (e′ for every n ∈ N, by (62) we conclude that (een)n is Ces´aro summable to zero. Hence by Lemma 6.10 it suffices to show that n)n. Since, nXj=1 nXj=1 xsj − x, x′ sj = 1 n 1 n , λieei(cid:13)(cid:13)(cid:13)∗∗ (67) λie′ (cid:13)(cid:13)(cid:13) nXi=1 i(cid:13)(cid:13)(cid:13)∗ =(cid:13)(cid:13)(cid:13) nXi=1 for every n ∈ N and λ1, . . . λn ∈ R with Pn λ1, . . . λn ∈ R withPn λiei(cid:13)(cid:13)(cid:13)∗ (cid:13)(cid:13)(cid:13) nXi=1 nXi=1 i=1 λi = 0. Also let ((sk k→∞(cid:13)(cid:13)(cid:13) and the proof is complete. such that limk→∞ sk 1(1) = +∞. Then =(cid:13)(cid:13)(cid:13) i(cid:13)(cid:13)(cid:13)∗ nXi=1 = lim λixsk λie′ Indeed, let n ∈ N and i=1 λ1 = 0. j )n j=1)k be a sequence in Plmn(F ↾ L) i(cid:13)(cid:13)(cid:13) = lim k→∞(cid:13)(cid:13)(cid:13) nXi=1 λix′ sk i(cid:13)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:13) nXi=1 λieei(cid:13)(cid:13)(cid:13)∗∗ (cid:3) We close by a strengthening of Theorem 6.17 for Banach spaces with separable dual. We will need the following lemma. Lemma 6.18. Let F be a regular thin family and (ys)s∈F be an F -sequence in a Banach space X. Let L ∈ [N]∞ and suppose that for every ε > 0 and N ∈ [L]∞ there exist k ∈ N, λ1, . . . , λk > 0 and (sj)k j=1 λj = 1 j=1 λj ysj k < ε. Then for every x∗ ∈ X ∗, ε > 0 and N ∈ [L]∞ there exists j=1 ∈ Plm(F ↾ N ) such thatPk M ∈ [N ]∞ such that x∗(ys) < ε for every s ∈ F ↾ M . and kPk Proof. Let x∗ ∈ X ∗, ε > 0 and N ∈ [L]∞. By Proposition 2.6 there exists M ∈ [N ]∞ such that exactly one of the following holds. (a) x∗(ys) < ε, for every s ∈ F ↾ M , or (b) x∗(ys) ≥ ε, for every s ∈ F ↾ M , or (c) x∗(ys) ≤ −ε, for every s ∈ F ↾ M . It suffices to show that cases (b) and (c) cannot occur. Indeed, suppose (b) holds true (the proof for case (c) is similar). By our assumption there exist k ∈ N, λ1, . . . , λk > 0 and (sj)k j=1 λj = 1 and j=1 ∈ Plm(F ↾ M ) such thatPk j=1 λj ysj k < ε. But then ε >(cid:13)(cid:13)(cid:13) kXj=1 λj ysj(cid:13)(cid:13)(cid:13) ≥ x∗(cid:16) kXj=1 λjysj(cid:17) = kXj=1 λj x∗(ysj ) ≥ ε, (cid:3) which is a contradiction. kPk (68) Corollary 6.19. Let X be a Banach space with separable dual. Let F be a regular thin family, (xs)s∈F be an F -sequence in X and M ∈ [N]∞ such that (xs)s∈F ↾M generates a singular F -spreading model (en)n. Let en = e′ n + e be the natural decomposition of (en)n. Then there exist x ∈ X with kxk = kek∗ and N ∈ [M ]∞ HIGHER ORDER SPREADING MODELS 37 such that setting x′ (up to isometry) F -spreading model. s = xs −x, (x′ s)s∈F ↾N is weakly null and admits (e′ n)n as a unique s)s∈F ↾L admits (e′ Proof. By Theorem 6.17, there exist x ∈ X with kxk = kek∗ and L ∈ [M ]∞ such that (x′ n)n as a unique F -spreading model. By applying Lemma 6.18 (for xs in place of ys) and a standard diagonalization for a countable dense subset of X ∗ we may choose N ∈ [L]∞ such that the F -subsequence (x′ s)s∈F ↾N is in addition weakly null. (cid:3) References [1] F. Albiac, N. J. Kalton, Topics in Banach Space Theory, Springer, 2006. [2] D. E. Alspach, S. A. Argyros, Complexity of weakly null sequences, Dissertationes Math., (1992). [3] S.A. Argyros, V. Kanellopoulos and K. Tyros, Higher order Spreading Models in Banach space theory, preprint, available at http://arxiv.org/abs/1006.0957. [4] S. A. Argyros, V. Kanellopoulos and K. Tyros, Finite order Spreading Models, preprint, avail- able at http://arxiv.org/abs/1105.2732. [5] S.A. Argyros, S. Todorcevic, Ramsey methods in analysis, Birkhauser Verlag, Basel, 2005. [6] B. Beauzamy, J.-T. Laprest´e, Mod`eles ´etal´es des espaces de Banach, Travaux en Cours, Her- mann, Paris, 1984, iv+210 pp. [7] A. Brunel, L. Sucheston On B-convex Banach spaces, Math. Systems Theory 7 (1974), no. 4, 294 -- 299. [8] A. Dvoretzky, Some results on convex bodies and Banach spaces, 1961 Proc. Internat. Sympos. Linear Spaces (Jerusalem, 1960) pp. 123 -- 160 Jerusalem Academic Press, Jerusalem; [9] F. Galvin, K. Prikry, Borel sets and Ramsey's theorem, J. Symbolic Logic, 38, (1973), 193 -- 198. [10] I. Gasparis, A dichotomy theorem for subsets of the power set of the natural numbers, Proc. Amer. Math. Soc., 129, (2001), no. 3, 759 -- 764. [11] W.T Gowers, Ramsey Methods in Banach Spaces, in Handbook of the geometry of Banach Spaces, vol. 2, (2003), 1072-1097. [12] J. L. Krivine, Sous-espaces de dimension finie des espaces de Banach r´eticul´es, Ann. of Math. (2) 104 (1976), no. 1, 1 -- 29. [13] J. Lopez-Abad, S. Todorcevic, Pre-compact families of finite sets of integers and weakly null sequences in Banach spaces, Topology Appl. 156 (2009), no. 7, 1396 -- 1411. [14] C.St.J.A. Nash-Williams, On well quasi-ordering transfinite sequences, Proc. Cambr. Phil. Soc., 61, (1965), 33-39. [15] E. Odell, Th. Schlumprecht, On the richness of the set of p's in Krivine's theorem, Geometric aspects of functional analysis (Israel, 1992 -- 1994), 177 -- 198, Oper. Theory Adv. Appl., 77, Birkhauser, Basel, 1995. [16] P. Pudlak, V. Rodl, Partition theorems for systems of finite subsets of integers, Discrete Math. 39, (1982), no. 1, 67 -- 73. [17] F. P. Ramsey, On a problem of formal logic, Proc. London Math. Soc. 30 (2), (1929), 264286. (2001) [18] H. Rosenthal, A characterization of Banach spaces containing l1, Proc. Nat. Acad. Sci. U.S.A. 71 (1974), 2411 -- 2413. National Technical University of Athens, Faculty of Applied Sciences, Department of Mathematics, Zografou Campus, 157 80, Athens, Greece E-mail address: [email protected] National Technical University of Athens, Faculty of Applied Sciences, Department of Mathematics, Zografou Campus, 157 80, Athens, Greece E-mail address: [email protected] Department of Mathematics, University of Toronto, Toronto, Canada, M5S 2E4 E-mail address: [email protected]
1310.2187
3
1310
2015-03-11T17:23:43
Schur--Agler and Herglotz--Agler classes of functions: positive-kernel decompositions and transfer-function realizations
[ "math.FA" ]
We discuss transfer-function realization for multivariable holomorphic functions mapping the unit polydisk or the right polyhalfplane into the operator analogue of either the unit disk or the right halfplane (Schur/Herglotz functions over either the unit polydisk or the right polyhalfplane) which satisfy the appropriate stronger contractive/positive real part condition for the values of these functions on commutative tuples of strict contractions/strictly accretive operators (Schur--Agler/Herglotz--Agler functions over either the unit polydisk or the right polyhalfplane). As originally shown by Agler, the first case (polydisk to disk) can be solved via unitary extensions of a partially defined isometry constructed in a canonical way from a kernel decomposition for the function (the {\em lurking-isometry method}). We show how a geometric reformulation of the lurking-isometry method (embedding of a given isotropic subspace of a Kre\u{\i}n space into a Lagrangian subspace---the {\em lurking-isotropic-subspace method}) can be used to handle the second two cases (polydisk to halfplane and polyhalfplane to disk), as well as the last case (polyhalfplane to halfplane) if an additional growth condition at $\infty$ is imposed. For the general fourth case, we show how a linear-fractional-transformation change of variable can be used to arrive at the appropriate symmetrized nonhomogeneous Bessmertny\u{\i} long-resolvent realization. We also indicate how this last result recovers the classical integral representation formula for scalar-valued holomorphic functions mapping the right halfplane into itself.
math.FA
math
SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES OF FUNCTIONS: POSITIVE-KERNEL DECOMPOSITIONS AND TRANSFER-FUNCTION REALIZATIONS JOSEPH A. BALL AND DMITRY S. KALIUZHNYI-VERBOVETSKYI Abstract. We discuss transfer-function realization for multivariable holomor- phic functions mapping the unit polydisk or the right polyhalfplane into the operator analogue of either the unit disk or the right halfplane (Schur/Herglotz functions over either the unit polydisk or the right polyhalfplane) which sat- isfy the appropriate stronger contractive/positive real part condition for the values of these functions on commutative tuples of strict contractions/strictly accretive operators (Schur -- Agler/Herglotz -- Agler functions over either the unit polydisk or the right polyhalfplane). As originally shown by Agler, the first case (polydisk to disk) can be solved via unitary extensions of a partially de- fined isometry constructed in a canonical way from a kernel decomposition for the function (the lurking-isometry method). We show how a geometric reformulation of the lurking-isometry method (embedding of a given isotropic subspace of a Kreın space into a Lagrangian subspace -- the lurking-isotropic- subspace method) can be used to handle the second two cases (polydisk to halfplane and polyhalfplane to disk), as well as the last case (polyhalfplane to halfplane) if an additional growth condition at ∞ is imposed. For the general fourth case, we show how a linear-fractional-transformation change of variable can be used to arrive at the appropriate symmetrized nonhomoge- neous Bessmertnyı long-resolvent realization. We also indicate how this last result recovers the classical integral representation formula for scalar-valued holomorphic functions mapping the right halfplane into itself. 1. Introduction For U, Y coefficient separable Hilbert spaces, we define the operator-valued Schur class S(U ,Y) (over the unit disk D) to consist of all holomorphic functions S on the unit disk D with values in the closed unit ball of the space L(U ,Y) of bounded linear operators from U to Y, i.e., subject to kS(ζ)k ≤ 1 for all ζ ∈ D. The fol- lowing result linking the theories of holomorphic functions, linear operators, and input/state/output linear systems is now well known (see e.g. [10] for a full discus- sion where multivariable extensions are also treated). Theorem 1.1. Given a function S : D → L(U ,Y), the following are equivalent. (1) S ∈ S(U ,Y). 2010 Mathematics Subject Classification. 32A10; 47A48; 47A56. Key words and phrases. Schur -- Agler class; Herglotz -- Agler class; Bessmertnyı long resol- vent representation; positive-kernel decomposition; transfer-function realization; lurking-isometry method; lurking-isotropic-subspace method. The authors were partially supported by US -- Israel BSF grant 2010432. The second author was also partially supported by NSF grant DMS-0901628, and wishes to thank the Department of Mathematics at Virginia Tech for hospitality during his sabbatical visit in January -- May 2013, when a significant part of work on the paper was done. 1 2 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI (2) The de Branges -- Rovnyak kernel KS(ω, ζ) = I − S(ω)∗S(ζ) 1 − ωζ is a positive kernel on D, i.e., there is an auxiliary Hilbert space X and a holomorphic function H : D → L(U ,X ) which gives rise to a Kolmogorov decomposition for KS: (3) S has a unitary transfer-function realization, i.e., there is an auxiliary KS(ω, ζ) = H(ω)∗H(ζ). Hilbert state space X and a unitary colligation matrix U =(cid:20)A B C D(cid:21) : (cid:20)X U(cid:21) →(cid:20)X Y(cid:21) so that (3′) Condition (3) above holds where the colligation matrix U is taken to be any S(ζ) = D + ζC(I − ζA)−1B for ζ ∈ D. of (i) coisometric, (ii) isometric, or (iii) contractive. It is natural to seek extensions of the Schur class to the multivariable setting where the disk D is replaced by the polydisk Dd = {ζ = (ζ1, . . . , ζd) ∈ Cd : ζk < 1 for k = 1, . . . , d}. We therefore define the d-variable Schur class Sd(U ,Y) to consist of holomorphic functions S : Dd → L(U ,Y) subject to kS(ζ)k ≤ 1 for all ζ ∈ Dd. It was the pro- found observation of Agler [1] that, unless d ≤ 2, a characterization of Sd(U ,Y) of the same form as Theorem 1.1 is not possible. Instead, we define what is now called the Schur -- Agler class, denoted as SAd(U ,Y), to consist of holomorphic functions S : Dd → L(U ,Y) such that kS(T1, . . . , Td)k ≤ 1 whenever T = (T1, . . . , Td) is a commutative d-tuple of strict contraction operators on a fixed separable infinite- dimensional Hilbert space K. Here the functional calculus defining S(T1, . . . , Td) can be given by S(T1, . . . , Td) = Xn∈Zd + Sn ⊗ T n Snζn is the multivariable Taylor expansion for S centered at the origin 0 ∈ Dd and where we use standard multivariable notation: (convergence in the strong operator topology) where S(ζ) = Pn∈Zd if n = (n1, . . . , nd) ∈ Zd +. 1 ··· ζnd d , T n = T n1 1 ··· T nd ζn = ζn1 d + The following result due to Agler [1] (see also [2, 13]) has had a profound impact on the subject over the years. Theorem 1.2. Given a function S : Dd → L(U ,Y), the following are equivalent. (1) S ∈ SAd(U ,Y). (2) S has an Agler decomposition in the sense that there exists d L(Y)-valued positive kernels K1, . . . , Kd on Dd such that I − S(ω)∗S(ζ) = dXk=1 (1 − ωkζk)Kk(ω, ζ). (1.1) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 3 (3) S has a unitary Givone -- Roesser d-dimensional transfer-function realiza- tion, i.e., there is an auxiliary Hilbert state space X with a d-fold orthog- onal direct-sum decomposition X = X1 ⊕ ··· ⊕ Xd together with a unitary colligation matrix U =(cid:20)A B C D(cid:21) : (cid:20)X U(cid:21) →(cid:20)X Y(cid:21) so that S(ζ) = D + C(I − P (ζ)A)−1P (ζ)B for ζ ∈ Dd. (1.2) where we have set P (ζ) = ζ1P1 + ··· + ζdPd where Pk is the orthogonal projection of X onto Xk for each k = 1, . . . , d. (3′) Condition (3) above holds where the colligation matrix U is taken to be any of (i) coisometric, (ii) isometric, or (iii) contractive. The goal of this paper is to study parallel results for an assortment of linear- fractional transformed versions of the Schur -- Agler class. Specifically, we seek anal- ogous characterizations of the following classes of holomorphic functions: (1) The Herglotz -- Agler class over the unit polydisk Dd, denoted HA(Dd,L(U)) defined as the class of all holomorphic functions F : Dd → L(U) such that F (T ) has positive real part F (T ) + F (T )∗ ≥ 0 for all commutative d-tuples T = (T1, . . . , Td) of strict contractions on the Hilbert space K. (2) The Schur -- Agler class over the right polyhalfplane Πd = {z = (z1, . . . , zd) ∈ Cd : zk + zk > 0 for k = 1, . . . , d}, denoted by SA(Πd,L(U ,Y)), consisting of all holomorphic L(U ,Y)-valued functions s on Πd such that ks(A)k ≤ 1 for all strictly accretive com- mutative d-tuples A = (A1, . . . , Ad) of operators on K (i.e., such that Ak + A∗k ≥ cI for some constant c > 0, k = 1, . . . , d). (3) The Herglotz -- Agler class over Πd, denoted by HA(Πd,L(U)), consisting of all holomorphic functions f on Πd such that f (A) + f (A)∗ ≥ 0 for all strictly accretive commutative d-tuples A of operators on K. (4) The Nevanlinna -- Agler class over the upper polyhalfplane zk − zk 2i > 0 for k = 1, . . . , d}, denoted by NA((iΠ)d,L(U)), consisting of all holomorphic L(U)-valued (iΠ)d = {z = (z1, . . . , zd) ∈ Cd : 2i (ef (eA) − ef (eA)∗) ≥ 0 whenever eA = functions ef on (iΠ)d such that 1 (eA1, . . . , eAd) is a commutative d-tuple of operators on K, each with strictly positive-definite imaginary part (i.e., such that 1 constant c > 0, k = 1, . . . , d). 2i (Ak − A∗k) ≥ cI for some To be consistent with the more detailed notation used for these variants of the Schur -- Agler class, we will also use the notation SA(Dd,L(U ,Y)) for the Schur -- Agler class SAd(U ,Y) over the polydisk Dd discussed above. 4 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI We note that our convention is to use the term Herglotz for functions with values having positive real part, and Nevanlinna for functions with values having positive imaginary part; we recognize that these conventions are by no means universal (see e.g. [6]). For the single-variable case such realization results have been explored in a sys- tematic way in [35] and [12]. For the multivariable setting, apart from the now clas- sical Schur -- Agler class over the polydisk SAd(U ,Y), the only results along these lines which we are aware of are those in the recent paper of Agler -- McCarthy -- Young [3] and of Agler -- Tully-Doyle -- Young [4, 5]. The approach in [35] (in the single-variable setting) is to use a linear-fractional- transformation (LFT) change of variables (on the domain and/or the range side) to reduce the desired result to the corresponding result for the Schur class over the unit disk. This is also the main tool in [3, 4, 5]: use an LFT Cayley-transform change of variables to reduce results for the Nevanlinna -- Agler class to the corre- sponding known results for the Schur -- Agler class. However the procedure is rather intricate due to the added subtleties involved in handling points at infinity in the multivariable case. In contrast, the approach in [12] is to apply a projective version of the lurk- ing isometry argument (roughly, a lurking-isotropic-subspace argument in a Kreın- space setting) to arrive at the desired realization result via a direct but unified Kreın-space geometric argument. One of the main contributions of the present work is to extend this approach to the multivariable setting. The main difficulty is to guarantee that a naturally defined isotropic subspace is actually a graph space with respect to a system of coordinates not coming from a fundamental decompo- sition of the ambient Kreın space. We show how this difficulty can be overcome for the case of the Dd-Herglotz -- Agler class and Πd-Schur -- Agler class. For the Πd- Herglotz -- Agler class, we are able to overcome the difficulty only in a special case (associated with the imposition of a growth condition at infinity), thereby recov- ering parallel results from [3]. For the most general Πd-Herglotz -- Agler function f , we follow the LFT change-of-variable approach of [4] combined with the more general realization formalism (Schur complement of an operator pencil) suggested by the work of Bessmertnyı (see [18, 19, 20, 21, 28]) to arrive at a realization for- mula for the most general Πd-Herglotz -- Agler function. We note that the original Bessmertnyı class involved additional symmetries leading to strong rigidity results. It was conjectured in [9] that an appropriate weakening of the metric conditions for the Bessmertnyı operator pencil should lead to a representation for the most general Πd-Herglotz -- Agler function. Here we show that this conjecture is correct once one identifies the appropriate modification: one must allow the nonhomoge- neous skew-adjoint term in the nonhomogeneous Bessmertnyı operator pencil to be unbounded (more precisely, a certain flip Π-impedance-conservative system node in the sense of [35]). There has been a lot of work on transfer-function realization for the single- variable Schur and Herglotz classes over the right half plane. The most influential for our point of view toward multivariable generalizations is the work of Arov- Nudelman [7] and of Staffans and collaborators (see [34, 35, 12, 29, 37] as well as the treatise [36] and the references there). There is also a complementary approach to such realization theory (upper halfplane rather than right halfplane version) with emphasis on the theory of selfadjoint extensions of densely defined symmetric SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 5 operators on a Hilbert space (see [14, 15, 16] as well as the recent book [6] and the references there). The paper is organized as follows. Section 2 highlights the main ideas from Kreın-space geometry and from infinite-dimensional systems theory, in particular, the idea of a system node, which will be used in the later sections. Section 3 presents our results for the Herglotz -- Agler class over the polydisk while Section 4 does the same for the Schur -- Agler class over the right polyhalfplane. Section 5 presents our results for the restricted Herglotz -- Agler class over the right polyhalfplane where a growth condition at infinity is imposed on the functions to be realized. With this added restriction, the lurking-isotropic-subspace method from [12] adapts well to lead to a classical type realization (but with an in general unbounded Π-impedance- conservative system node) for the Πd-Herglotz -- Agler function. Section 6 identifies the nonhomogeneous unbounded Bessmertnyı operator pencils which then lead to a realization for the most general Herglotz -- Agler function over Πd. We also mention that the results parallel to the results of this paper for the four classes under discussion (Dd-Schur -- Agler, Dd-Herglotz -- Agler, Πd-Schur -- Agler and Πd-Herglotz -- Agler) for the rational matrix-valued (Cayley) inner case, where the emphasis is on obtaining realizations with finite-dimensional state space, are obtained in our companion paper [11]. We shall have occasion to need a Cayley transform (with both scalar and operator argument) acting between the right halfplane and the unit disk. Following the conventions in [28] and [11], we shall make use of the following version: ζ ∈ D 7→ w = w ∈ Π 7→ ζ = 1 + ζ 1 − ζ ∈ Π, with inverse given by w − 1 w + 1 ∈ D. (1.3) For ζ = (ζ1, . . . , ζd) a point in the unit polydisk Dd, we continue to use the notation for the corresponding point in the right polyhalfplane Πd. Similarly, given a point w = (w1, . . . , wd) in the right polyhalfplane Πd, we use the notation 1 + ζ 1 − ζ :=(cid:18) 1 + ζ1 1 − ζ1 , . . . , 1 + ζd 1 − ζd(cid:19) w − 1 w + 1 :=(cid:18) w1 − 1 w1 + 1 , . . . , wd − 1 wd + 1(cid:19) (1.4) (1.5) for the associated point in the polydisk. Acknowledgement. The authors are thankful to the anonymous referee for his/her careful reading and constructive remarks. 2. Preliminaries 2.1. Decompositions of the identity. Given a Hilbert space X , we shall say that a collection of d operators (Y1, . . . , Yd) on X forms a d-fold positive decom- position of the identity IX if each Yk is a selfadjoint contraction (0 ≤ Yk ≤ IX k=1 Yk = IX ). In case (Y1, . . . , Yd) = (P1, . . . , Pd) consists of orthogonal projection operators (necessar- ily with pairwise orthogonal ranges) we shall say that (P1, . . . , Pd) forms a d-fold spectral decomposition of IX . Note that d-fold spectral decompositions (P1, . . . , Pd) for 1 ≤ k ≤ d) which together sum up to the identity (Pd 6 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI arise in the realization formula for the Schur -- Agler class in Theorem 1.2. We shall see that the more general positive decompositions are needed for the realization formulas for functions in the Πd-Schur -- Agler class and in the Πd-Herglotz -- Agler class, as already discovered in [3, 4, 5]. From the definitions we see that any spectral decomposition (P1, . . . , Pd) is also a positive decomposition. There is also a result in the converse direction: if (Y1, . . . , Yd) is a positive decomposition of IX , then there exist a Hilbert space eX , , and an isometric embedding ι : X → eX a spectral decomposition (P1, . . . , Pd) of I eX such that Yk = ι∗Pkι for k = 1, . . . , d. This can be seen as a consequence of the Naimark dilation theorem (apply [30, Theorem 4.6] with the measurable space X taken to be the finite set {k ∈ N : 1 ≤ k ≤ d}). To prove the result for this simple case of the Naimark dilation theorem, i=1 X by simply define an isometric embedding of X intoLd  ι = onto the k-th block in the direct-sum spaceLd Q1 ... Qd i=1 X . where Qk provides a factorization Yk = Q∗kQk and take Pk equal to the projection 2.2. Basics on the geometry of Kreın spaces. In this Section we review some basics about the geometry of Kreın spaces and Kreın-space operator theory which we shall need in the sequel. Other resources on this topic is a similar survey section in the paper [12] as well as the more complete treatises [22, 8]. A Kreın space by definition is a linear space K endowed with an indefinite inner product [·,·] which is complete in the following sense: there are two subspaces K+ and K− of K such that the restriction of [·,·] to K+ × K+ makes K+ a Hilbert space while the restriction of −[·,·] to K− × K− makes K− a Hilbert space, and K = K+[ +]K− is a [·,·]-orthogonal direct sum decomposition of K. In this case the decomposition K = K+[ +]K− is said to form a fundamental decomposition for K. Fundamental decompositions are never unique except in the trivial case where one of K+ or K− is equal to the zero space. Unlike the case of Hilbert spaces where closed subspaces all look the same, there is a rich geometry for subspaces of a Kreın space. A subspace M of a Kreın space K is said to be positive, isotropic, or negative depending on whether [u, u] ≥ 0 for all u ∈ M, [u, u] = 0 for all u ∈ M (in which case it follows that [u, v] = 0 for all u, v ∈ M as a consequence of the Cauchy-Schwarz inequality), or [u, u] ≤ 0 for all u ∈ M. Given any subspace M, we define the Kreın-space orthogonal complement M[⊥] to consist of all v ∈ K such that [u, v] = 0 for all u ∈ K. Note that the statement that M is isotropic is just the statement that M ⊂ M[⊥]. If it happens that M = M[⊥], we say that M is a Lagrangian subspace of K. Examples of such subspaces arise from placing appropriate Kreın-space inner products on the direct sum H1 ⊕ H2 of two Hilbert spaces and looking at graphs of operators of an appropriate class. SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 7 Example 2.1. Suppose that H′ and H are two Hilbert spaces and we take K to be the external direct sum H′ ⊕ H with inner product (cid:20)(cid:20)x y(cid:21) ,(cid:20)x′ y′(cid:21)(cid:21) =(cid:28)(cid:20)IH′ 0 −IH(cid:21)(cid:20)x 0 y(cid:21) ,(cid:20)x′ y′(cid:21)(cid:29)H′⊕H where h·,·iH′⊕H is the standard Hilbert-space inner product on the direct-sum Hilbert space H′ ⊕ H. In this case it is easy to find a fundamental decomposition: take K+ =(cid:2) H {0}(cid:3) and K− =h {0} H to H′ and let M be the graph of T : H′i. Now let T be a bounded linear operator from M = GT =(cid:26)(cid:20)T x x(cid:21) : x ∈ H(cid:27) ⊂ K. Then a good exercise is to work out the following facts: • GT is negative if and only if kTk ≤ 1. • GT is isotropic if and only if T is isometric (T ∗T = IH). • GT is Lagrangian if and only if T is unitary: T ∗T = IH and T T ∗ = IH′ . Example 2.2. Let H be a Hilbert space and set K equal to the direct-sum space K = H ⊕ H with indefinite inner product given by In this case a choice of fundamental decomposition is not so obvious; one such choice is IH IH 0(cid:21)(cid:20)x (cid:20)(cid:20)x y(cid:21) ,(cid:20)x′ K+ =(cid:26)(cid:20)x y′(cid:21)(cid:21) =(cid:28)(cid:20) 0 y(cid:21) ,(cid:20)x′ x(cid:21) : x ∈ H(cid:27) , K− =(cid:26)(cid:20)−y y′(cid:21)(cid:29)H⊕H y (cid:21) : y ∈ H(cid:27) . . Then the exercise parallel to that suggested in Example 2.1 is to work out the following: given a closed operator T ∈ L(H) with dense domain D(T ) ⊂ H, let GT =(cid:26)(cid:20)T x x(cid:21) : x ∈ D(T )(cid:27) ⊂ K be its graph space. Then: • GT is negative if and only if T is dissipative: hT x, xi + hx, T xi ≤ 0 for all x ∈ D(T ). • GT is maximal negative, i.e., GT is negative and is not contained in any properly larger negative subspace, if and only if T is maximal dissipative, i.e., T is dissipative and has no proper dissipative extension. An equivalent condition is T is dissipative and the operator I + T is onto (or equivalently wI + T is onto for all w in the right halfplane Π+) (see [31]). • GT is isotropic if and only if T is skew-symmetric or T ⊂ −T ∗, i.e. hT x, xiH+ hx, T xiH = 0 for all x ∈ D(T ). • GT is Lagrangian if and only if T is skew-adjoint or T = −T ∗ (i.e., T ⊂ −T ∗ and T and T ∗ have the same domain: y, z ∈ H such that hT x, yiH = hx, ziH implies that y ∈ D(T ) and z = −T y. A closely related result is proved in Corollary 2.7 below. We shall have use for the following connection between Examples 2.1 and 2.2. Consider Example 2.1 for the case where H′ = H and call this Kreın space K1. Let 8 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI K2 be the Kreın space H ⊕ H with Kreın-space inner product the negative of the inner product as in Example 2.2: (cid:20)(cid:20)x y(cid:21) ,(cid:20)x′ y′(cid:21)(cid:21)K2 =(cid:28)(cid:20) 0 −IH −IH 0 (cid:21)(cid:20)x y(cid:21) ,(cid:20)x′ y′(cid:21)(cid:29)H⊕H . Note that this modification of the inner product just interchanges positive and negative subspaces and preserves isotropic and Lagrangian subspaces. Then the operator Γ defined by Γ := 1 √2(cid:20) IH −IH IH(cid:21) : (cid:20)H H(cid:21) →(cid:20)H H(cid:21) IH defines a Kreın-space isomorphism between the Kreın space K1 as in Example 2.2 and the Kreın space K2 as in Example 2.1; this is a consequence of the bijectivity of Γ along with the identity Γ∗(cid:20) 0 −I 0 (cid:21) Γ =(cid:20)I −I 0 −I(cid:21) . 0 It follows that Γ maps K1-Lagrangian subspaces to K2-Lagrangian subspaces. In particular, if U is unitary and does not have 1 as an eigenvalue, we define the Cayley transform Y of T according to the formula D(Y ) = Ran(I − U ) and Y = (I + U )(I − U )−1. (2.1) Note that ΓGU = I 1 −I x (cid:21) : x ∈ H(cid:27) √2(cid:20) I I(cid:21)(cid:26)(cid:20)U x =(cid:26)(cid:20)(I + U )x (I − U )x(cid:21) : x ∈ H(cid:27) =(cid:26)(cid:20)(I + U )(I − U )−1 =(cid:26)(cid:20) Y IH(cid:21) y : y ∈ D(Y )(cid:27) = GY . I (cid:21) (I − U )x : x ∈ H(cid:27) There results the following fact concerning the Cayley transform map of this type. Proposition 2.3. Suppose that U is a linear operator on a Hilbert space H which does not have 1 as an eigenvalue. Define the Cayley transform Y of U as in (2.1). Then U is unitary if and only if Y is skew-adjoint. 2.3. Well-posed linear systems and system nodes. A continuous-time in- put/state/output (i/s/o) linear system is a system of equations of the form Σ : (cid:26) x(t) = Ax(t) + Bu(t) y(t) = Cx(t) + Du(t) (2.2) where x(t) takes values in the state space X , u(t) takes values in the input space U, and y(t) takes values in the output space Y. Under the assumption that the system matrix (sometimes also called the colligation matrix) Σ :=(cid:20)A B C D(cid:21) : (cid:20)X U(cid:21) →(cid:20)X Y(cid:21) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 9 consists of bounded operators, imposition of the initial condition x(0) = 0 and application of the Laplace transform bx(w) =Z ∞ 0 e−wtx(t) dt, leads to the input-output relation in the frequency domain by(w) =Z ∞ 0 e−wty(t) dt e−wtu(t) dt, 0 bu(w) =Z ∞ by(w) = TΣ(w)bu(w) where is the transfer function of the linear system Σ (2.2). The converse question of when an L(U ,Y)-valued function f on the right halfplane Π can be realized as f = TΣ TΣ(w) = D + C(wI − A)−1B C D(cid:21) has generated much interest over the years. The for some system Σ = (cid:20)A B obvious necessary condition is that w 7→ f (w) be analytic on some right halfplane but this is not sufficient if we limit our attention to systems Σ consisting only of bounded operators A, B, C, D. While it is clear that the right generalization of the state operator A is that it should be the generator of a C0-semigroup, exactly how to handle the remaining operators B, C, D so as to get a meaningful theory containing compelling examples of interest was not so clear, but some progress was made already in the 1970s (see [26, 27]). It is now understood that a useful notion of generalized system matrix Σ is that associated with so-called well-posed system. Roughly, a well-posed linear system is an i/s/o linear system for which the integral form of the system operators A, B, C, D satisfy natural compatibility conditions and the integral form of the system matrix Ct 0 0 Dt 0(cid:21) : (cid:20) x(0) (cid:20)At Bt of the system matrix Σ =(cid:20)A B C D(cid:21) appearing in (2.2)) is the notion of system node makes sense as a bounded operator from X ⊕ L2 ([0, t)) for each U t > 0 (see [36] for complete details). The "right" infinitesimal object (the analogue u[0,t)(cid:21) 7→(cid:20) x(t) y[0,t)(cid:21) defined as follows; this notion is well laid out in the work of Staffans [36, 35] where it is acknowledged that much of the idea was already anticipated in the earlier work of Salamon [32] and Smuljan [33]. ([0, t)) to X ⊕ L2 Y (2.3) We first make some preliminary observations. A system node Σ is still an oper- ator from X ⊕ U to X ⊕ Y but now allowed to be unbounded with some domain D(Σ) ⊂ X ⊕ U. We may then split Σ in the form Σ2(cid:21) Σ =(cid:20)Σ1 where Σ1 : D(Σ) → X and Σ2 : D(Σ) → Y. However we allow the possibility that D(Σ) does not split D(Σ) =(cid:20)D(Σ)1 D(Σ)2(cid:21) as the direct sum of a linear manifold D(Σ)1 in X with a linear manifold D(Σ)2 in U. To keep the parallel with the classical case, we therefore write Σ =(cid:20)A&B C&D(cid:21) 10 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI splitting (cid:2)AX B(cid:3) where AX and B are operators mapping the spaces X and U with the notation A&B (and similarly C&D) suggesting that the common domain of A&B and C&D in X ⊕ U may not have a splitting. However A&B will have a into a larger "rigged" space X−1 which is part of a so-called Gelfand triple defined as follows. We assume that A is an in general unbounded closed operator with dense domain D(A) := X1 in X and with nonempty resolvent set. Then X1 is a Hilbert space in its own right with respect to the X1-norm given by kxk1 := k(αI − A)xkX where α is any fixed number in the resolvent set of A. While the norm depends on the choice of α, any other choice α′ of α leads to the same space but with an equivalent norm, as long as α and α′ are in the same connected component of the resolvent set of A. If A is the generator of a C0-semigroup (the most interesting case for us), one can take the connected component of the resolvent set to be a right half plane. Note that (αI − A)X1 can be viewed as an isometry from X1 onto X . We next construct another Hilbert space X−1 as the completion of X in the norm kxk−1 := k(αI − A)−1xkX . If {xn}n∈Z+ is a sequence in X1 converging to x ∈ X in X -norm, then the sequence {(αI − A)xn}n∈Z+ is Cauchy in X−1-norm and hence converges to an element x−1 ∈ X−1. One can check that this element x−1 ∈ X−1 is independent of the choice of sequence {xn} ⊂ X1 converging in X -norm to x; we denote this element y ∈ X−1 by y = (αI − A)x and then define an extension AX : X → X−1 by AX : x 7→ αx − (αI − A)x ∈ X−1 if x ∈ X . We then have that the extended operator (αI − A)X is an isometry from X onto X−1 and we have the nested inclusions X1 ⊂ X ⊂ X−1 with continuous and dense injections. We will on occasion simplify the notation AX to simply A when the meaning is clear; thus for x ∈ X and x not necessarily in X1 = D(A), the element Ax is still defined but as an element of X−1. It is also useful to note the role of these spaces in duality pairings. First, we note that the constructions in the previous paragraph can be carried out using the 1 = (αI − A∗)−1X operator A∗ in place of A. When this is done we get spaces X ⋆ (with the norm kxk1,⋆ := k(αI − A∗)xkX and X ⋆ −1 equal to the completion of X in −1-norm kxk−1,⋆ = k(αI − A∗)−1xkX with the properties that (αI − A∗)−1 the X ⋆ is an isometry from X onto X ⋆ 1 and (αI − A∗) extends to an isometry from X onto X ⋆ −1 with the nesting X ⋆ −1. Given any x ∈ X , we can view x as a linear functional on X ⋆ 1 using the X -pairing: 1 ⊂ X ⊂ X ⋆ ℓx(x⋆ 1) := hx⋆ 1, xiX for x⋆ 1 ∈ X ⋆ 1 . (2.4) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 11 If we write x⋆ 1 = (αI − A∗)−1y with y ∈ X , then ℓx(x⋆ 1) = hx⋆ 1, xiX = h(αI − A∗)−1y, xiX = hy, (αI − A)−1xiX ≤ kykXk(αI − A)−1xkX = kx⋆ 1k1,⋆kxk−1 with equality if we take x⋆ conclude that the linear-functional norm of ℓx is equal to the X−1-norm of X : 1 = (αI − A∗)−1(αI − A)−1x (or y = (αI − A)−1x). We kℓxk(X ⋆ 1 )∗ = kxk−1 1 )∗. It is not difficult to see that this and X can be identified with a subspace of (X ⋆ subspace is dense and hence, after taking completions, we have that X−1 is naturally isomorphic to the dual space of X ⋆ 1 via the X -pairing (2.4). For our application to infinite-dimensional linear systems, in practice the unbounded closed operator A in this construction will also be taken to be the generator of a C0-semigroup, so as to make sense of a differential equation of the form dx dt (t) = Ax(t). We are now ready to introduce the notion of system node. Definition 2.4. By a system node Σ on the collection of three Hilbert spaces (U ,X ,Y), we mean a linear operator such that: Σ :=(cid:20)A&B C&D(cid:21) : (cid:20)X U(cid:21) ⊃ D(Σ) →(cid:20)X Y(cid:21) (1) Σ is a closed operator with domain D(Σ) dense in(cid:20)X U(cid:21), (2) If we define an operator A with domain D(A) = {x ∈ X : [ x Ax = Σ [ x 0 ] for x ∈ D(A), 0 ] ∈ D(Σ)} by (2.5) then A is the generator of a C0-semigroup on X. (3) Let AX : X → X−1 be the extension of the operator A : X1 = D(A) → X (as in (2.5)) as described in the previous paragraph. Then there is a bounded linear operator B : U → X−1 so that we recover the operator ] → X−1 to D(Σ): A&B as the restriction of the operator(cid:2)AX B(cid:3) : [ X U Σ =(cid:2)AX B(cid:3) D(Σ). (4) C&D is a bounded operator from D(Σ) to Y C&D ∈ L(D(Σ),Y) where D(Σ) carries the graph norm. (5) The domain D(Σ) of Σ is characterized as D(Σ) =(cid:26)(cid:20)x U(cid:21) : AX x + Bu ∈ X(cid:27) . Given u ∈ U , there exists xu ∈ D(Σ) so that [ xu A consequence of this definition of system node is the following fact: u ] ∈ D(Σ). u(cid:21) ∈(cid:20)X (2.6) (2.7) 12 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI in part (5) of Definition 2.4 to check that this [ xu Indeed, it suffices to take xu =(cid:0)(αI − A)X(cid:1)−1 u ] = AX(cid:0)(αI − A)X(cid:1)−1 (cid:2)AX B(cid:3) [ xu = (A − αI)X(cid:0)(αI − A)X(cid:1)−1 = α ·(cid:0)(αI − A)X(cid:1)−1 Bu ∈ X . Bu + Bu Bu. To see this, we use the criterion u ] is in D(Σ): Bu + α ·(cid:0)(αI − A)X(cid:1)−1 Bu + Bu In the definition of system node, we took pains to write the top component in ] does not split; yet in the end the form A&B to indicate that its domain in [ X U we found a splitting by extending to a larger space(cid:2) X−1 restriction of an extended operator (cid:2)AX B(cid:3) : (cid:2) X−1 split. Similarly, there is at least a partial splitting for the operator C&D : D(Σ) → Y. Indeed, we have seen that X1 = D(A) = {x ∈ X : [ x 0 ] ∈ D(Σ} is a dense subset of X . We may therefore define an operator C : X1 → Y by U (cid:3) and writing A&B as the U (cid:3) → Y whose domain does Cx = C&D [ x 0 ] for x ∈ X1. (2.8) Since C&D is a bounded operator from D(Σ) (graph norm) to Y, it follows that C so defined is a bounded operator from X1 (graph norm induced by the operator A) to Y. In practice we assign no independent meaning to the D in C&D except under some additional hypotheses (e.g., for the case of a regular system -- see the paper of Weiss [39] or [36, Section 5.6]). To this point we have at least versions of all the usual constituents for a linear input/state/output linear system: A : X1 → X main operator or state dynamics, B : U → X−1 input or control operator, C : X1 → Y output or observation operator, C&D : D(Σ) → Y combined observation/feedthrough operator, (2.9) and we lack in general an independent well-defined feedthrough operator D. The formula for xu in (2.7) can use any point w in the connected component (e.g., an appropriate right half plane) of the resolvent set of A containing α. For clarity, Bu to indicate the dependence of xu on the follows that the expression let us write xu(w) =(cid:0)(wI − A)X(cid:1)−1 point w in the right half plane. From the fact that(cid:2) xu(w) TΣ(w) : u 7→ C&D(cid:20)xu(w) u (cid:21) = C&D(cid:20)(cid:0)(wI − A)X(cid:1)−1 (cid:3) ∈ D(Σ) = D(C&D) it (cid:21) Bu u u is well defined and defines the transfer function of the system node. Let α be any fixed point in the resolvent set of A. Then we can recover the value of the transfer function at any point w in the same connected component of the resolvent set of A from its value at the fixed point α according to the recipe TΣ(w)u = (TΣ(w) − TΣ(α)) u + TΣ(α)u = C(xu(w) − xu(α)) + TΣ(α)u =(cid:16)(α − w)C(wI − A)−1(cid:0)(αI − A)X(cid:1)−1 B + TΣ(α)(cid:17) u Conversely, start with any semigroup generator A on X with domain D(A) = X1 with induced Gelfand rigging X1 ⊂ X ⊂ X−1, any input operator B : U → X−1, (2.10) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 13 and an output operator C : X1 → Y along with a value TΣ(α) ∈ L(U ,Y) for the transfer function at the point α. Define D(Σ) =(cid:26)(cid:20)x u(cid:21) : AX x + Bu ∈ X(cid:27) with A&B =(cid:2)AX B(cid:3) D(Σ), C&D [ x u ] = C(x − xu(α)) + TΣ(α)u. (2.11) Then Σ so defined is a system node with value of its transfer function at α equal to the prescribed value TΣ(α). Note here that x− xu(α) is in X1 since [ x are in D(Σ) and hence so also is u ] and(cid:2) xu(α) (cid:3) u (cid:20)x − xu(α) 0 (cid:21) =(cid:20)x u(cid:21) −(cid:20)xu(α) u (cid:21) (2.12) resulting in x − xu(α) ∈ X1. Moreover, we recover the transfer function TΣ at a general point from A, B, C and TΣ(α) via the formula (2.10). (see [34, Lemma 2.2] for more complete details). Given a system node Σ, it is possible to make sense of the associated system of differential equations (cid:20) x(t) y(t)(cid:21) = Σ(cid:20)x(t) u(t)(cid:21) , as long as u ∈ C2([0,∞);U) and(cid:20)x(0) u(0)(cid:21) ∈ D(Σ) (see [29, Proposition 2.6]). Appli- x(0) = x0 (2.13) cation of the Laplace transform 0 e−wtx(t) dt x(t) 7→bx(w) :=Z ∞ by(w) = C(wI − A)−1x(0) + TΣ(w)bu(w) to the system equations (2.13) leads us to the input-output property of the transfer function (2.10): for w with sufficiently large real part. We mention that any well-posed linear system (2.3) is the integral form of the dynamical system associated with a system node (see e.g. [36]); however there are system nodes for which the associated dynamical system (2.13) is not well-posed (i.e., one or more of the block operators Bt 0 appearing in (2.3) fail to exist as bounded operators between the appropriate spaces), despite the fact that the infinitesimal form of the system equations (2.13) does make sense. 0, Dt 0, Ct The following examples of system nodes will be useful in the sequel. In this discussion we make use of Kreın-space geometry notions discussion in Section 2.2. Example 2.5. Suppose that Σ = (cid:20)A&B operator such that its graph space Y(cid:21) is a closed C&D(cid:21) : (cid:20)X U(cid:21) ⊃ D(Σ) → (cid:20)X  D(Σ) ⊂ X Y X U A&B C&D 0 I 0 I G(Σ) := 14 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI is a Lagrangian subspace of K := X ⊕ Y ⊕ X ⊕ U, where K is given a Kreın space structure using the signature operator J = 0 0 IX 0 0 IY 0 0 0 IX 0 0 0 0 0 −IU  . Ct 0 0 Dt Then Σ is a system node (see [12, Proposition 4.9]). In fact, Schur-class functions over the right halfplane Π are characterized as those functions s having a realization s(w) = TΣ(w) as in (2.10) with a system node Σ of this form (see [12, Theorem 4.10]). These systems are also characterized by the energy-balance property that the block matrixh At Bt 0i associated with the integral form (2.3) of the system equa- tions is unitary for each t (see [35]). For this reason any such system node is said to be a Π-scattering conservative system node. In particular, any Π-scattering conser- vative system is well-posed. For more recent information concerning Π-scattering conservative system nodes and the related notion of Π-scattering dissipative system we refer to the recent work of Malinen, Staffans, and Weiss ([29], [37], [38]). In par- nodes (where the block matrixh At Bt ticular, the result of [37] is that a linear operator S : (cid:20)X 0i in (2.3) is assumed only to be contractive), Y(cid:21) is a Π-scattering passive system node if and only if it is closed with its graph equal to a maximal J -negative subspace of X ⊕ Y ⊕ X ⊕ U. Much of this work also has a focus of fitting physical examples into this framework; a recent accomplishment was to fit Maxwell's equations into this framework (see [40]). U(cid:21) →(cid:20)X 0 Dt Ct 0 Example 2.6. Suppose that Σ = (cid:20)A&B C&D(cid:21) : (cid:20)X U(cid:21) ⊃ D(Σ) → (cid:20)X U(cid:21) is a closed operator with output space Y taken to be the same as the input space U such that (1) its graph G(Σ) is a Lagrangian subspace of K, but where now K is given the Kreın-space inner product induced by the signature operator J ′ given by  , u(cid:21) is in D(Σ). (2) for each u ∈ U there is an xu ∈ X so that(cid:20)xu J ′ = 0 0 0 0 0 IX 0 −IU 0 IX 0 −IU 0 0 0 0 and (2.14) Then Σ is a system node (see [12, Proposition 4.11]). In fact, Herglotz functions over the right halfplane which also satisfy the growth condition at infinity lim t→+∞ t−1f (t)u = 0 for each u ∈ U are characterized as those functions f having a realization f (w) = TΣ(w) as in (2.10) with a system node Σ of this form (see [12, Theorem 4.12]). The trajectories (u(t), x(t), y(t)) satisfy the alternative energy-conservation law kx(t)k2 X − kx0k2 X = 2Z t 0 Rehy(t), u(t)iU dt SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 15 and Σ is called a Π-impedance-conservative system node (see [35]). As the transfer function for a Π-impedance-conservative system node need not be bounded in the right halfplane, it follows that Π-impedance-conservative system nodes need not be well-posed in general (see [34]). In connection with Example 2.6 we shall have use for the following additional fact. 0 operator such that system node as in Example 2.6. (1) Y is skew-adjoint: Y = −Y∗, and U(cid:21) is a closed densely defined u(cid:21) ∈ D(Y). Corollary 2.7. Suppose that Y : D(Y) ⊂(cid:20)X U(cid:21) →(cid:20)X (2) for each u ∈ U there is an xu ∈ X such that(cid:20)xu Set J ′0 =(cid:20)IX then Y := −ΣJ′0 : D(Y) ⊂(cid:20)X D(Y) =n(cid:20)x 0 −IU(cid:21) and set Σ := −YJ ′0. Then Σ is a Π-impedance-conservative U(cid:21) is a closed operator with the dense domain U(cid:21) : J′0(cid:20)x Conversely, if Σ is a Π-impedance-conservative system node as in Example 2.6, −u(cid:21) ∈ D(Σ)o Proof. Given any closed, densely defined operator Σ : D(Σ) ⊂ (cid:20)X u(cid:21) =(cid:20) x satisfying conditions (1) and (2). U(cid:21) →(cid:20)X u(cid:21) ∈(cid:20)X U(cid:21) → (cid:20)X U(cid:21), the following translation of the definition of adjoint operator to graph spaces is well known: where here the orthogonal complement is with respect to the standard Hilbert space inner product. More generally, compute the J ′-orthogonal complement (where (cid:18)(cid:20)Σ I(cid:21)D(Σ)(cid:19)⊥ =(cid:20) I −Σ∗(cid:21)D(Σ∗) 0(cid:21) by the definition (2.14) of J ′) as follows: J ′0y1(cid:21) ∈(cid:18)(cid:20)Σ I(cid:21)D(Σ)(cid:19)⊥J ′ J ′0 J ′ =(cid:20) 0 J ′0 y2(cid:21) ∈(cid:18)(cid:20)Σ (cid:20)y1 I(cid:21)D(Σ)(cid:19)⊥ =(cid:20) I −Σ∗(cid:21)D(Σ∗) I y2(cid:21) =(cid:20)J ′0y2 ⇔ J ′(cid:20)y1 y2(cid:21) ∈(cid:20)−J ′0Σ∗J ′0 ⇔(cid:20)y1 (cid:20)Σ I(cid:21)D(Σ) =(cid:20)−J ′0Σ∗J ′0 Σ = −J ′0Σ∗J ′0. I (cid:21)D(Σ∗J ′0). (cid:21)D(Σ∗J ′0) Thus condition (1) in Example 2.6 for Σ to be a Π-impedance-conservative system node translates to or simply An equivalent condition is: the operator Y := ΣJ ′0 (or equivalently −Y = −ΣJ ′0) is skew-adjoint: Y = −Y∗. 16 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI Conversely, if Y is any skew-adjoint operator with dense domain in (cid:20)X U(cid:21), then Σ = −YJ ′0 satisfies condition (1) in Example 2.6. Since(cid:20)x u(cid:21) ∈ D(Σ) if and only if (cid:20) x −u(cid:21) ∈ D(Y), condition (2) in Example 2.6 is equivalent to condition (2) in the statement of the corollary. (cid:3) The next result gives a model for Π-impedance-conservative system nodes as described in Corollary 2.7. Proposition 2.8. Let (T, V0, R) be a triple of operators such that: (1) T is a densely defined skew-adjoint operator on the Hilbert space X , (2) V0 ∈ L(U ,X ) is a bounded linear operator from the input-output space U (3) R is a bounded skew-adjoint operator on U. into X , and U(cid:21) →(cid:20)X u(cid:21) ∈(cid:20)X Define an operator Σ : D(Σ) ⊂(cid:20)X D(Σ) =(cid:26)(cid:20)x V ∗0 x + V ∗0 T (x − V0u) + Ru(cid:21) ∈(cid:20)X U(cid:21) as follows. Set U(cid:21) : x − V0u ∈ D(T )(cid:27) U(cid:21) for (cid:20)x Σ(cid:20)x u(cid:21) =(cid:20) T (x − V0u) + V0u and then define u(cid:21) ∈ D(Σ). (2.15) (2.16) Equivalently, Σ can be defined as the system node constructed from the data A = T ∈ L(X1,X ), B = (I − T )V0 ∈ L(U ,X−1), C = V ∗0 (I − T )∗ ∈ L(X1,U), α = 1 and TΣ(1) = V ∗0 V0 + iR ∈ L(U) (2.17) according to the recipe (2.11). Then Σ is a Π-impedance-conservative system node. Conversely, any Π-impedance-conservative system node arises in this way from a triple of operators T, V0, R satisfying conditions (1), (2), and (3) above. Remark 2.9. We note that, in case T is bounded, the operator Σ given by (2.15) and (2.16) is simply I − T 0 0 (2.18) 0 T Σ =(cid:20)I 0 V ∗0(cid:21)(cid:20) (I − T )∗ −T (cid:21)(cid:20)I One way to make sense of the first term in this formula for the general case where T 0 V0(cid:21) +(cid:20)0 0 R(cid:21) . is allowed to be unbounded is as follows. We may view eΣ =(cid:20) (I − T )∗ −T (cid:21) = X−1(cid:21), where X−1 is the rigged level- (cid:20) T x2(cid:21) : T (x1 − x2) ∈ X(cid:27) . −T (cid:21) as an operator from(cid:20)X D =(cid:26)(cid:20)x1 I + T (−1) space associated with the skew-adjoint operator T as explained in discussion at the beginning of this section. It is natural to introduce a domain X(cid:21) to(cid:20)X−1 X(cid:21)(cid:27) =(cid:26)(cid:20)x1 x2(cid:21) ∈(cid:20)X I − T I − T (2.19) T x2(cid:21) : eΣ(cid:20)x1 SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 17 the operator Σ via the formula (2.18) with domain given by and define an operator eΣ0 : D ⊂(cid:20)X X(cid:21) →(cid:20)X X(cid:21) by eΣ0 = eΣD. Then we may define 0 V0(cid:21)(cid:20)x u(cid:21) ∈ D(eΣ0)(cid:27) . D(Σ) =(cid:26)(cid:20)x u(cid:21) : (cid:20)I Proof of Proposition 2.8. We first show that Σ defined as in (2.15), (2.16) satisfies conditions (1) and (2) in Corollary 2.7. As for condition (2), note that if we set xu = V0u for each u ∈ U, then xu − V0u = 0 ∈ D(T ), so condition (2) is satisfied. It remains to verify condition (1). Toward this end, by the representation (2.18) for Σ as explained in Remark 2.9, 0 it suffices to show that the operator Σ′ :=(cid:20)−I I(cid:21)(cid:20) T I + T −T + R(cid:21) =(cid:20) −T with domain D as in (2.19) is skew-adjoint. Note that(cid:20)y1 T + I −T + R(cid:21) y2(cid:21) ∈ D(cid:18)(cid:20) −T T − I I − T means that the sesquilinear form 0 0 T + I −T + R(cid:21)∗(cid:19) T − I y2(cid:21)(cid:29) x2(cid:21) ,(cid:20)y1 T − I (cid:28)(cid:20) −T T + I −T + R(cid:21)(cid:20)x1 x2 ] ∈ D(cid:0)(cid:2) −T T−I T − I = hT (x1 − x2),−y1 + y2i − hx2, y1i + hx1, y2i + hRx2, y2i T +I −T +R(cid:3)(cid:1) is continuous in the argument [ x1 defined for [ x1 x2 ] in the X ⊕X norm. As the second, third, and fourth terms are automatically bounded in the [ x1 x2 ]-argument and the element x1 − x2 is an arbitrary element of D(T ) (e.g., take x2 = 0 and x1 ∈ D(T )), it follows that necessarily y2 − y1 ∈ D(T ∗) = D(T ) (i.e., [ y1 y2 ] ∈ D) and the calculation above continues as y2(cid:21)(cid:29) x2(cid:21) ,(cid:20)y1 (cid:28)(cid:20) −T T + I −T + R(cid:21)(cid:20)x1 = hx1 − x2, T ∗(y2 − y1)i − hx2, y1i + hx1, y2i + hRx2, y2i x2(cid:21) , −(cid:20) −T x2(cid:21) ,(cid:20) −T y2 + T y1 + y2 =(cid:28)(cid:20)x1 T y2 − T y1 − y1 + R∗y2(cid:21)(cid:29) =(cid:28)(cid:20)x1 y2(cid:21)(cid:29) T + I −T + R(cid:21)(cid:20)y1 where we use R is bounded with R = −R∗ in the last two steps. The skew- adjointness of the operator Σ′ now follows as wanted. As a consequence of Corollary 2.7 it follows in particular that Σ is a system node. The equivalence of the original definition (2.16) of Σ with the alternative formulation based on the data set (2.17) is a simple consequence of the identities (2.11) along with computation of the value of the transfer function at the point 1 ∈ Π: TΣ(1) = V ∗0 V0 + iR; this in turn is a routine verification which we leave to the reader. For the converse statement, we let Σ′ =(cid:20)A′&B′ C′&D′(cid:21) be any Π-impedance-conser- vative system node. As (cid:20)−(A′&B′) C′&D′ (cid:21) is skew-adjoint, from the duality theory of system nodes (see e.g. [36]) one can see that necessarily A′ = −A′∗ and C′ = B′∗; here B′ ∈ L(U ,X ′ −1), C′ ∈ L(X ′1,U) and computation of the adjoint B′∗ is with respect to the duality pairing between X ′ −1 and X ′1 via the X ′ pairing (2.4) (note that X ′1 = X ′⋆ 1 since A′ = −A′∗). Set T = A′. As T is skew-adjoint, both 1 and −1 are in the resolvent set of T and we may define a bounded operator V0 ∈ L(U ,X ′) T − I 18 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI by V0 = (I − T )−1B′. We then have B′ = (I − T )V0 and C′ = V ∗0 (I + T ). We now have all the ingredients to define another Π-impedance-conservative system node Σ0 =(cid:20)I 0 V ∗0(cid:21)(cid:20) T 0 I + T zero). When we write Σ0 in the form I − T −T (cid:21)(cid:20)I 0 0 V0(cid:21) as in (2.15), (2.16) (with R taken equal to C0&D0(cid:21) , Σ0 =(cid:20)A0&B0 we see that the construction gives that D(Σ′) = D(Σ0), i.e., D(A0&B0) = D(A′&B′), with A0&B0 = A′&B′, so A0 := A0&B0D(Σ0)∩[ X0 ] = A′&B′D(Σ0)∩[ X0 ] =: A′, B0 = B′ ∈ L(U ,X−1), C0 := C0&D0D(Σ0)∩[ X0 ] = C′&D′D(Σ)∩[ X0 ] =: C′. As observed in (2.12), for any fixed choice of α ∈ Π (e.g., α = 1) an element [ x u ] ∈ D(Σ′) can be decomposed as (cid:20)x u(cid:21) =(cid:20)x − (αI − A)−1Bu 0 (cid:21) +(cid:20)(αI − A)−1Bu (cid:21) u where each summand is again in D(Σ′) = D(Σ0). Then making use of the formulas (2.11) we compute and similarly 0 Ax + Bu C(x − (αI − A)−Bu)(cid:21) +(cid:20) C(x − (αI − A)−1Bu)(cid:21) +(cid:20) TΣ(α)u(cid:21) TΣ0(α)u(cid:21) (TΣ(α) − TΣ0 (α)) u(cid:21) . Ax + Bu 0 Σ(cid:20)x u(cid:21) =(cid:20) Σ0(cid:20)x u(cid:21) =(cid:20) u(cid:21) = Σ0(cid:20)x Σ(cid:20)x 0 −I(cid:3) and Σ0(cid:2) I 0 0 0 R ] with R := TΣ(α) − TΣ0(α) equal to a bounded operator on u(cid:21) +(cid:20) 0 −I(cid:3) are skew-adjoint, it is necessarily the case that (cid:3) from which we read off Thus Σ = Σ0 + [ 0 0 U. As both Σ(cid:2) I 0 R = −R∗ as well. We need a generalization of the formula for the transfer function (2.10) to the setting where the resolvent (wI − A)−1 is replaced by the structured resolvent (Y (w) − A)−1. For the statement of this result it is convenient to assume that A is maximal dissipative. Recall from the discussion in Example 2.2 above that a densely defined closed Hilbert-space operator A is maximal dissipative if it is dissipative (RehAx, xi ≤ 0 for all x ∈ D(A)) and I + A is onto; it then follows that wI + A is onto for all w ∈ Π and that A is the generator of a contractive semigroup (see [31]). Proposition 2.10. Let A be a maximal dissipative operator on the Hilbert space X (and hence A is the generator of a contractive C0-semigroup with resolvent set containing the right halfplane Π), and suppose that (Y1, . . . , Yd) is a d-fold positive decomposition of IX . For w = (w1, . . . , wd) ∈ Cd, we set Y (w) = w1Y1 +···+wdYd. Then the following observations hold true: SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 19 (1) For w ∈ Πd, Y (w) − A is invertible with k(Y (w) − A)−1k ≤ minj Re wj 1 . (2.20) (2) For w ∈ Πd, the operator (Y (w) − A)−1 is a bicontinuous bijection from X (3) For w ∈ Πd, the operator Y (w) − A : X1 → X has an extension onto X1. (Y (w) − A)X : X → X−1 which is a bicontinuous bijection from X onto X−1. (4) For any system node Σ =(cid:20)A&B C&D(cid:21) containing A as its state operator/semi- group generator, it holds that (cid:20)((Y (w) − A)X )−1Bu u (cid:21) ∈ D(Σ). Proof. (1) For w = (w1, . . . , wd) ∈ Πd and x ∈ D(A), we have k(Y (w) − A)xk · kxk ≥ h(Y (w) − A)x, xi ≥ Reh(Y (w) − A)x, xi ≥ RehY (w)x, xi = dXk=1 h(Re wk)Ykx, xi dXk=1 ≥ minj(Re wj) = minj(Re wj)kxk2. hYkx, xi (2.21) Thus Y (w) − A is bounded below and has a left inverse. A similar argument with Y (w) − A replaced by (Y (w) − A)∗ shows that Y (w) − A also has a right inverse. The computation (2.21) gives the estimate (2.20). (2) Next note that for w ∈ Πd and x ∈ X , A(Y (w) − A)−1x = −x + Y (w)(Y (w) − A)−1x ∈ X from which we conclude that (Y (w) − A)−1 maps X into X1. Conversely if x ∈ X1, then y = (Y (w) − A)x ∈ X and we recover x as x = (Y (w) − A)−1y. We conclude that (Y (w) − A)−1 maps X bijectively to X1. The fact that (Y (w) − A)−1 is bicontinuous is then a consequence of the open mapping theorem. (3) We use the dual version of a result from part (2): for w ∈ Πd, Y (w)∗ − A∗ : X ⋆ 1 → X is a bicontinuous bijection from X ⋆ adjoints with respect to the X -pairing, we get an operator 1 (with the (1, ⋆)-norm) onto X . If we then take (Y (w)∗ − A∗)∗ : X → X−1 which must also be a bicontinuous bijection, but now from X to X−1. Since the duality is with respect to the X -pairing (2.4), it is clear that this map provides an extension of the operator Y (w) − A : X1 → X : (Y (w) − A)X := (Y (w)∗ − A∗)∗ : X → X−1. 20 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI (4) To show that(cid:20)((Y (w) − A)X )−1Bu D(Σ) in Definition 2.4 we need only show that u (cid:21) is in D(Σ), by the characterization of (cid:2)AX B(cid:3)(cid:20)((Y (w) − A)X )−1Bu u (cid:21) ∈ X . (2.22) As Bu ∈ X−1, (Y (w) − A)−1 maps X−1 to X by part (3) and Y (w) is a bounded operator on X , we conclude that Y (w)((Y (w) − A)X )−1Bu ∈ X . We now rewrite the expression in (2.22) as (cid:2)AX B(cid:3)(cid:20)((Y (w) − A)X )−1Bu u (cid:21) = [(A − Y (w))X + Y (w)]((Y (w) − A)X )−1Bu + Bu = Y (w)((Y (w) − A)X )−1Bu to conclude that (2.22) holds as desired. (cid:3) Remark 2.11. We note that as a consequence of property (4) in Proposition 2.10 it is possible to define the transfer function associated with the structured resolvent (Y (w) − A)−1 via TΣ,{Yk}(w) = [C&D] (cid:20)(cid:0)(Y (w) − A)X(cid:1)−1 u Bu (cid:21) . It is tempting to view this as the transfer function for a multidimensional linear system Pd k=1 Yk ∂x ∂tk (t) = Ax(t) + Bu(t) y(t) = Cx(t) + Du(t). (2.23) where t = (t1, . . . , td) (a continuous-time version of a multidimensional linear sys- tem of Fornasini -- Marchesini type -- see [25]). However, in general it is not clear how to extend the operators Y1, . . . , Yd ∈ L(X ) to operators YkX−1 in a sensible way. Special cases where this is possible are: (1) the case where B and C are bounded, i.e., B ∈ L(U ,X ) and C ∈ L(X ,Y): in this situation the system (2.23) makes sense with state space taken simply to be X , and (2) in case the operators Yk commute with A: in this case one can extend Yk to X−1 via the formula YkX−1 = (αI − A)Yk(αI − A)−1 : X−1 → X−1. We note that continuous-time counterparts of Fornasini -- Marchesini models of a somewhat different form have been considered in the literature (see [23] and the references there). 3. The Herglotz -- Agler class over the polydisk In this section we present our results for the Herglotz -- Agler class over the poly- disk Dd. Theorem 3.1. Given a function F : Dd → L(U), the following are equivalent. (1) F is in the Herglotz -- Agler class HA(Dd,L(U)). SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 21 (2) F has a Dd-Herglotz -- Agler decomposition, i.e., there exist L(U)-valued pos- itive kernels K1, . . . , Kd on Dd such that F (ω)∗ + F (ζ) = (1 − ωkζk)Kk(ω, ζ). (3.1) dXk=1 (3) There exist a Hilbert space X , a d-fold spectral decomposition (P1, . . . , Pd) of IX with associated operator pencil P (ζ) = ζ1P1 + ··· + ζdPd (see Section 2.1), and a bounded colligation matrix U =(cid:20)A B C D(cid:21) : (cid:20)X U(cid:21) →(cid:20)X U(cid:21) with block matrix entries A, B, C, D satisfying the relations A∗A = AA∗ = IX , B = AC∗, D + D∗ = CC∗ (= B∗B) (3.2) such that F (ζ) = D + C(I − P (ζ)A)−1P (ζ)B. (3.3) Proof. (1)⇒(2): Use that F ∈ HA(Dd,L(U)) if and only if S(ζ) := [F (ζ)−I][F (ζ)+ I]−1 is in the Schur -- Agler class SA(Dd,L(U)). Then S has a Schur -- Agler decom- position I − S(ω)∗S(ζ) = dXk=1 (1 − ωkζk)eKk(ω, ζ) for L(U)-valued positive kernels eK1, . . . , eKd on Dd. A routine computation gives I − S(ω)∗S(ζ) = 2[F (ω)∗ + I]−1 (F (ω)∗ + F (ζ)) [F (ζ) + I]−1. This leads us to the Dd-Herglotz -- Agler decomposition (3.1) with Kk(ω, ζ) = 1 2 [F (ω)∗ + I] eKk(ω, ζ) [F (ζ) + I]. Notice that Agler in [1] proved the implications (1)⇒(2) in Theorems 1.2 and 3.1 simultaneously, while we show here that they are, in fact, equivalent. ζ1I fX1 decomposition (3.1) in the form (2)⇒(3): Since each kernel Kk is positive, each Kk has a Kolmogorov decom- H(ζ) = Colk=1,...,d[Hk(ζ)] be the associated block column matrix function defining position, i.e., there is a function Hk : Dd → L(U , eXk) (for some auxiliary Hilbert space eXk) so that we have the factorization Kk(ω, ζ) = Hk(ω)∗Hk(ζ). We let a function on Dd with values in L(U , eX ), where we set eX = eX1 ⊕ ··· ⊕ eXd (written . We now may rewrite the Agler as columns). We set P (ζ) =  eG = span  u : u ∈ U , ζ ∈ Dd F (ω)∗ + F (ζ) = H(ω)∗(I eX − P (ω)∗P (ζ))H(ζ). eX eX We consider the subspace H(ζ) F (ζ) ⊂ U U   ... ζdI fXd P (ζ)H(ζ) IU (3.4) 22 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI where the ambient space eX ⊕ U ⊕ eX ⊕ U is given the Kreın-space inner product induced by the signature matrix eJ given by −I eX 0 0 0 0 0 0 IU 0 0 I eX 0 0 IU 0 0  . eJ =   Then one can check IU IU 0 IU 0 0 P (ζ)H(ζ) P (ω)H(ω) 0 0 0 IU H(ζ) F (ζ) 0 0 I eX 0 H(ω) F (ω) We claim that −I eX 0 0 0  u′+  u, where the last equality follows as a consequence of the Agler decomposition (3.4). = h[−H(ω)∗H(ζ) + F (ω)∗ + H(ω)∗P (ω)∗P (ζ)H(ζ) + F (ζ)] u, u′i = 0 * We conclude that eG is a eJ -isotropic subspace. We then extend eG to a J -Lagrangian subspace G, where the ambient space eX ⊕ U ⊕ eX ⊕ U extended to a space of the form X ⊕ U ⊕ X ⊕ U with X ⊃ eX and the Kreın-space Gramian matrix J of the same block form as eJ above. Indeed, suppose that(cid:20) x and ζ ∈ Dd, 0 =*J(cid:20) x x + u, u′iU . Hence u = H(ζ)∗P eX x for all ζ ∈ Dd; in particular, u = H(0)∗P eX element of G has the form(cid:20) 0 =(cid:28)J(cid:20) 0(cid:21) ∈ G. As G is isotropic, we then must have, for all u′ ∈ U 0(cid:21) ," H(ζ) # u′+ = h−H(ζ)∗P eX (cid:21). We must also have (cid:21)(cid:29) = −kxk2 (cid:21) ,(cid:20) {0}# = {0}. G ∩" X U {0} x. Thus our F (ζ) P (ζ)H(ζ) H(0)∗P fX x H(0)∗P fX x IU x 0 0 x 0 0 H(0)∗P fX x u 0 u 0 x 0 0 which enables us to conclude that x = 0 and hence also(cid:20) claim follows. X x 0 0 H(0)∗P fX x (cid:21) = 0, and the We are now able to conclude that G is a graph space: A&B C&D 0 IX 0 IU G =  for some closed linear operator (cid:2) A&B construction the vector(cid:2) P (ζ)H(ζ)u D = (cid:2) D1 Lemma 3.4 from [12] to conclude that in fact (cid:2) A&B (cid:20)x u(cid:21) : (cid:20)x C&D(cid:3) with domain D ⊂ [ X (cid:3) is in D for all ζ ∈ Dd and u ∈ U; in particular, U (cid:3) for some linear manifold D1 ⊂ X . We are now in position to apply C&D(cid:3) = [ A B u ] ∈ D for all u ∈ U and hence D splits: C D ] is bounded with by setting ζ = 0 ∈ Dd we see that [ 0 u(cid:21) ∈ D ]. Furthermore, by U u SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 23 domain equal to all of [ X U produced a colligation matrix U = [ A B ], and moreover the identities (3.2) hold. We have now In particular, it follows that, for any u ∈ U, there is a corresponding(cid:2) x′ u′(cid:3) in [ X that U ] so C D ] so that A B C D 0 IX 0 IU G =   u = (cid:20)A B C D(cid:21)(cid:20)P (ζ)H(ζ)u (cid:21) =(cid:20)H(ζ)u (cid:20)X U(cid:21) . (cid:20)x′ u′(cid:21) . F (ζ)u(cid:21) for all u ∈ U A B C D 0 IX 0 IU H(ζ) F (ζ) P (ζ)H(ζ) IU u From the bottom two components of (3.5) we read off u′ = u. x′ = P (ζ)H(ζ)u, Then the top two components of (3.5) give (3.5) (3.6) which we rewrite as a linear system of operator equations Solving the first equation in (3.6) for H(ζ) gives CP (ζ)H(ζ) + D = F (ζ). (cid:26) AP (ζ)H(ζ) + B = H(ζ) H(ζ) = (IX − AP (ζ))−1B. (We note that the inverse on the right hand side exists since A is unitary and kP (ω)k < 1 for ω ∈ Dd.) Plugging this last expression into the second of equations (3.6) then yields F (ζ) = D + CP (ζ)(I − AP (ζ))−1B = D + C(I − P (ζ)A)−1P (ζ)B and condition (3) in the statement of Theorem 3.1 follows. (3)⇒(2): We assume that F has the representation (3.3) where the coefficient matrices A, B, C, D satisfy the relations (3.2). Then we compute F (ζ)∗ + F (ω) = D∗ + B∗(I − P (ζ)∗A∗)−1P (ζ)∗C∗ + D + CP (ω)(I − AP (ω))−1B = [D + D∗] + B∗(I − P (ζ)∗A∗)−1P (ζ)∗A∗B + B∗AP (ω)(I − AP (ω))−1B = B∗(I − P (ζ)∗A∗)−1X(I − AP (ω))−1B where we have set X equal to X = (I − P (ζ)∗A∗)(I − AP (ω)) + P (ζ)∗A∗(I − AP (ω)) + (I − P (ζ)∗A∗)AP (ω) = I − P (ζ)∗A∗ − AP (ω) + P (ζ)∗P (ω) + P (ζ)∗A∗ − P (ζ)∗P (ω) + AP (ω) − P (ζ)∗P (ω) = I − P (ζ)∗P (ω). We conclude that the Agler decomposition (3.1) holds with Kk(ζ, ω) = B∗(I − P (ζ)∗A∗)−1Pk(I − AP (ω))−1B, 24 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI where Pk = P eXk Theorem 3.1 follows. is the orthogonal projection of eX onto eXk, and condition (2) in (2)⇒(1): Given an Agler decomposition (3.1), we may rewrite it in the form (3.4). From the proof of (2)⇒(3)⇒(2) we see that we may assume that H(ζ) is holomorphic in ζ ∈ Dd. If T = (T1, . . . , Td) is a commutative d-tuple of strict contractions on K, it is straightforward to verify that the formula (3.4) leads to F (T )∗ + F (T ) = H(T )∗ (IX⊗K − P (T )∗P (T )) H(T ). k=1 ζkPk, we see that P (T ) =Pd k=1 Pk⊗Tk has kP (T )k < 1. Hence F (T )∗ + F (T ) ≥ 0 and we conclude that F ∈ HA(Dd,L(U)). From the diagonal form of P (ζ) =Pd (cid:3) Remark 3.2. Given a representation for F ∈ HA(Dd,L(U)) as in (3.3) and (3.2), let us separate out the selfadjoint and skew-adjoint parts of F (0) = D to rewrite the formula (3.3) as where Re F (0) = 1 (3.2) we see that F (ζ) = Re F (0) + C(I − P (ζ)A)−1P (ζ)B + R 2 (D + D∗) and we set R = 1 2 (D− D∗) = −R∗. From the relations 1 2 1 2 B∗(cid:0)I + 2A(I − P (ζ)A)−1P (ζ)(cid:1) B F (ζ) − R = B∗(cid:0)I + 2AP (ζ)(I − AP (ζ))−1(cid:1) B = = V ∗(I − AP (ζ))−1(I + AP (ζ))V, (3.7) where V := 1√2 B is such that V ∗V = Re F (0). We note that Agler [1] obtained the representation (3.7) for a function in the Herglotz -- Agler class HA(Dd,L(U)) starting with the realization S(ζ) = D + C(I − P (ζ)A)−1P (ζ)B (with U = [ A B C D ] unitary as in part (3) of Theorem 1.2) for the associated function S(ζ) = (F (ζ) − I)(F (ζ) + I)−1 (3.8) in the Schur -- Agler class SA(Dd,L(U)) as follows. The fact that S arises as the Cayley transform of a function in the Herglotz -- Agler class implies that I − S(ζ) = 2(F (ζ) + I)−1 is injective: in particular, D = S(0) has the property that I − D is invertible; moreover, the fact that U is unitary implies that the U0 := A + B(I − D)−1C is also unitary. This follows from the two identities, I (cid:20)A B C D(cid:21)(cid:20) (cid:20)A∗ C∗ B∗ D∗(cid:21)(cid:20) (I − D)−1C(cid:21) =(cid:20) (I − D∗)−1B∗(cid:21) =(cid:20) I U0 (I − D)−1C(cid:21) , (I − D∗)−1B∗(cid:21) , U∗0 which, when combined with that fact that [ A B C D ] is unitary, imply that both U0 and U∗0 are isometries. In terms of our notation, the result of Agler [1, Proof of theorem 1.8] is that then F has the representation (3.7) with A = U0, V = 1 √2 B. (3.9) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 25 The formula (3.7) can be further adjusted as follows: F (ζ) − R = V ∗(I − AP (ζ))−1(I + AP (ζ))V = V ∗(U − P (ζ))−1(U + P (ζ))V (3.10) where we set U = A∗. Here we still have V ∗V = Re F (0) and U = A∗ is unitary. We shall use this representation for a Dd-Herglotz -- Agler function in Section 6. For an alternate direct derivation of the realization (3.10) for a Dd-Herglotz -- Agler function, see [11], where the result is given for the rational matrix-valued case with the additional constraint that F have zero real part on the unit circle; in this case one can arrange that the state space X is finite-dimensional. 4. The Schur -- Agler class over the right polyhalfplane In this section we present our realization results for the Schur -- Agler class over the right polyhalfplane Πd. Theorem 4.1. Given a function s : Πd → L(U ,Y), the following are equivalent. (1) s is in the right polyhalfplane Schur -- Agler class SA(Πd,L(U ,Y)). (2) s has a Πd-Schur -- Agler decomposition, i.e., there exist positive kernels K1, . . . , Kd on Πd such that I − s(z)∗s(w) = dXk=1 (zk + wk)Kk(z, w). (4.1) for all z, w ∈ Πd. (3) There exists a state space X and a d-fold positive decomposition of IX (Y1, . . . , Yd) (see Section 2.1) together with a Π-scattering-conservative sys- tem node (see Example 2.5) so that C&D(cid:21) : D ⊂(cid:20)X U =(cid:20)A&B U(cid:21) →(cid:20)X Y(cid:21) s(w) = C&D(cid:20)(cid:0)(Y (w) − A)X(cid:1)−1 IU B (cid:21) , (4.2) where we have set Y (w) = w1Y1 + ··· + wdYd. Proof. (1)⇒(2): The proof uses the component-wise multivariable Cayley trans- form from Πd to Dd; for this purpose it is convenient to use the condensed notation (1.4) for the point 1+ζ point ζ = (ζ1, . . . , ζd) ∈ Πd. function S ∈ SA(Dd,L(U ,Y)) in the Schur -- Agler class over Dd via 1−ζd(cid:17) in the polydisk Dd associated with the For s ∈ SA(Πd,L(U ,Y)) in the Schur -- Agler class over Πd we associate the , . . . , 1+ζd 1−ζ =(cid:16) 1+ζ1 1−ζ1 S(ζ) = s(cid:18) 1 + ζ 1 − ζ(cid:19) Then by Theorem 1.2 S has a Dd-Schur -- Agler decomposition IU − S(ω)∗S(ζ) = dXk=1 (1 − ωkζk)eKk(ω, ζ). 26 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI Using the relation (where we use the convention (1.5)), we next get that w − 1(cid:19) s(w) = S(cid:18) w + 1 dXk=1(cid:18)1 − zk − 1 zk + 1 · IU − s(z)∗s(w) = wk − 1 wk + 1(cid:19) bKk(z, w) w + 1(cid:19) . w − 1 where we set This in turn leads to , z + 1 bKk(z, w) = eKk(cid:18) z − 1 dXk=1 IU − s(z)∗s(w) = (zk + wk)Kk(z, w) with Kk given by and (2) follows. Kk(z, w) = 2 1 zk + 1bKk(z, w) 1 wk + 1 Schur -- Agler decomposition (4.1) in the form (2)⇒(3): We use the Kolmogorov decompositions Kk(z, w) = Hk(z)∗Hk(w) I − s(z)∗s(w) = H(z)∗(P (z)∗ + P (w))H(w) (where Hk : Πd → L(U , eXk)) of the positive kernels Kk(z, w) to rewrite the Πd- . We intro- where we set H(w) = Col1≤k≤d Hd(w) with P (w) = eX0 := span(cid:8)Ran H(w) : w ∈ Πd(cid:9) ⊂ dMk=1 eXk = eX duce the subspace wdI fXd w1I fX1 (4.4) (4.3) ... and we introduce operators eY1, . . . ,eYd on eX0 by Pk eX0 eYk = P eX0 where P eX0 Pk = P eXk is the orthogonal projection of eX onto its subspace eX0 (4.4) and where is the orthogonal projection of eX onto eXk. It is easily verified that We next view (4.3) as the statement that the subspace eY1, . . . ,eYd form a positive decomposition of the identity on eX0. ⊂ eX0 eX0  u : w ∈ Πd, u ∈ U eG := span  eY (w)H(w) s(w) H(w) IU Y U  =: eK SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 27 0 IY 0 0 (4.5) 0 0 I eX0 0 0 I eX0 0 IU product induced by the indefinite Gramian matrix To this end we first check the necessary condition that  . We next check that eG can be expressed as a graph space 0 I eX0 0 0 0 0 0 −IU is an isotropic subspace of eK, where eK is considered as a Kreın space with inner eJ := eD0 eG = eA&eB eC&eD eC& eDi with dense domain eD0 ⊂ eX0 ⊕ U associated with a closed operator eU = h eA& eB mapping into eX0 ⊕ Y.  = {0} eG ∩ eX 0(cid:21)(cid:29) = ky′k2 0 =(cid:28)eJ(cid:20) x′ 0(cid:21) ,(cid:20) x′ 0 =*eJ" eY (w)H(w) as follows. We suppose that x′ ⊕ y′ ⊕ 0 ⊕ 0 ∈ eG. As eG is isotropic, it follows that from which we see that y′ = 0 and x′ ⊕ 0 ⊕ 0 ⊕ 0 ∈ eG. As eG is isotropic, we must 0(cid:21)+ = hH(w)u, x′i eX0 for all u ∈ U and w ∈ Πd. We now use the condition (4.4) to conclude that necessarily x′ = 0 as well. # u,(cid:20) x′ 0 0 then also have Y {0} {0} y′ 0 y′ 0 s(w) H(w) IU We next observe that (4.3) for z = w = te, where e = (1, . . . , 1), becomes I − s(te)∗s(te) = 2tH(te)∗H(te), so for any u ∈ U we have kH(te)uk2 = 1 2t (kuk2 − ks(te)uk2) → 0 as t → ∞. Therefore spanu,wn(cid:20)H(w) and we now see thatnh H(w) IU (cid:21) uo ⊃ spanu,w,tn(cid:20)H(w) − H(te) (cid:21) uo ⊃ spanu,wn(cid:20)H(w) 0 (cid:21) uo {0}(cid:21) , =(cid:20)fX0 IU i uo has dense span in eX0 ⊕ U. We conclude that eG is indeed a graph space as claimed. 0 28 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI K := X ⊕ Y ⊕ X ⊕ U where we may arrange that X is a Hilbert space containing By Proposition 2.5 in [12], we may embed eG into a J -Lagrangian subspace G of eX and where we set (4.6) Furthermore, it is argued there that one can arrange that this (possibly) enlarged Lagrangian subspace G is also a graph space: 0 IY 0 0 0 0 IX 0  . X Y {0} {0} 0 IX 0 0 0 0 0 −IU J = G ∩  = {0}. U =(cid:20)A&B C&D(cid:21) : D(U) ⊂(cid:20)X U(cid:21) →(cid:20)X Y(cid:21) D(U). I(cid:21)D(U) = G =(cid:20)U A&B C&D 0 IX 0 IU Hence there is a closed operator so that As discussed in Example 2.5, U is a Π-scattering-conservative system node. It is shown in the proof of Proposition 4.9 from [12] that then the main operator A of U (given by (2.5) with U in place of Σ) is maximal dissipative (as defined in Example 2.2 above). We let Y1, . . . , Yd be a positive decomposition of the identity on X which extends eY1, . . . ,eYd; e.g., one way to do this is YkX⊖ eX0 X⊖ eX0 = I , = Yk for k = 1, . . . , d Y1X⊖ eX0 Yk eX0 = 0 for k = 2, . . . , d, and extend by linearity. We then set Y (w) = w1Y1 + ··· + wdYd. As G contains eG, we conclude that Y (w)H(w)u for each w ∈ Πd and u ∈ U. Thus, for each such w and u there is ah xw,u w,ui ∈ D(U) so that u′ (4.7) (4.8)   s(w)u H(w)u u Y (w)H(w)u s(w)u H(w)u u  ∈  = A&B C&D 0 IU 0 IU A&B C&D 0 IX 0 IU D(U) (cid:20)xw,u u′w,u(cid:21) . From the bottom two rows of (4.8) we read off H(w)u = xw,u, u = u′w,u. SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 29 Plugging these back into the first two rows of (4.8) gives AX H(w)u + Bu = Y (w)H(w)u C&D(cid:20)H(w)u u (cid:21) = s(w)u. (4.9) Since U is a Π-scattering-conservative system node, a consequence of Proposition 2.10 is that Y (w) − A is invertible for each w ∈ Πd with an extension (Y (w) − A)X : X → X−1 having the property that ((Y (w) − A)X )−1 : X−1 → X . We may therefore solve the first of equations (4.9) for H(w)u to get and furthermore H(w)u =(cid:0)(Y (w) − A)X(cid:1)−1 Bu. (4.10) (cid:20)(cid:0)(Y (w) − A)X(cid:1)−1 u Bu (cid:21) ∈ D(C&D) for each w ∈ Πd. We can now substitute (4.10) into the second of equations (4.9) to arrive at the desired realization formula (4.2) for s, and (3) follows. (3)⇒(2): Assume that s has a realization as in (4.2). For w ∈ Πd, set I B B. H(w) =(cid:0)(Y (w) − A)X(cid:1)−1 (cid:21) u AX H(w)u + Bu = A&B(cid:20)(cid:0)(Y (w) − A)X(cid:1)−1 = AX(cid:0)(Y (w) − A)X(cid:1)−1 = −Bu + P (w)(cid:0)(Y (w) − A)X(cid:1)−1 = Y (w)(cid:0)(Y (w) − A)X(cid:1)−1 (cid:21) =(cid:20)A&B C&D(cid:21)(cid:20)H(w)u U(cid:20)H(w)u (cid:21) =(cid:20)Y (w)H(w)u = Y (w)H(w)u. Bu + Bu s(w)u Bu u u Bu + Bu (cid:21) . (4.11) Observe that Combining this with (4.3) gives By Proposition 4.9 from [12], the fact that U is a Π-scattering-conservative system node tells us that the graph of U is a Lagrangian subspace of X ⊕ Y ⊕ X ⊕ U in the Kreın-space inner product induced by J as in (4.6). In particular it holds that 0 =*J =*J A&B C&D 0 IX 0 IU Y (w)H(w)u s(w)u H(w)u u (cid:21)+ u (cid:20)H(w)u  , (cid:21) , A&B C&D 0 IX 0 IU Y (z)H(z)u′ s(z)u′ H(z)u′ u′ u′ (cid:20)H(z)u′  + = h[H(z)∗(Y (z)∗ + Y (w))H(w) + (s(z)∗s(w) − IU )] u, u′i. (4.12) 30 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI By the arbitrariness of u, u′ ∈ U, we conclude that H(z)∗ dXk=1 (zk + wk)Yk! H(w) = IU − s(z)∗s(w). (4.13) Then (4.13) leads to the Agler decomposition (4.1) with Kk(z, w) = H(z)∗YkH(w) = [Y 1/2 k H(z)]∗[Y 1/2 k H(w)]. (2)⇒(1): We write the Agler decomposition (4.1) in the form (4.3). Then, if A = (A1, . . . , Ad) is a commutative d-tuple of strictly accretive operators, the functional calculus gives I − s(A)∗s(A) = H(A)∗(P (A)∗ + P (A))H(A). From the diagonal form P (w) = w1P1 + ··· + wdPd of P (w), we see that P (A)∗ + P (A) ≥ 0, and hence I − s(A)∗s(A) ≥ 0, or ks(A)k ≤ 1. We conclude that s ∈ SA(Πd,L(U ,Y)) and (1) follows. (cid:3) 5. The Herglotz -- Agler class over the polyhalfplane In this section we present our realization results for a restricted class of Herglotz -- Agler functions over the right polyhalfplane Πd, where a growth condition (5.1) is imposed at infinity. Theorem 5.1. Given a function f : Πd → L(U), the following are equivalent: (1) f ∈ HA(Πd,L(U)) and also f satisfies the growth condition at +∞: lim t→+∞ t−1f (te)u = 0 for each u ∈ U . (5.1) where e = (1, . . . , 1) ∈ Πd. itive kernels K1, . . . , Kd on Πd such that (2) f has a Πd-Herglotz -- Agler decomposition, i.e., there exist L(U)-valued pos- f (z)∗ + f (w) = dXk=1 (zk + wk)Kk(z, w) (5.2) and in addition f satisfies the growth condition (5.1). (3) There exists a Hilbert state space X and a positive decomposition of the identity (Y1, . . . , Yd) on X along with an Π-impedance-conservative system node (see Example 2.6) C&D(cid:21) : D(Y) ⊂(cid:20)X Y =(cid:20)A&B U(cid:21) →(cid:20)X U(cid:21) f (w) = C&D(cid:20)(cid:0)(Y (w) − A)X(cid:1)−1 (cid:21) . IU B such that Proof. We show (5.3) (1)⇒(2)⇒(3)⇒(2)⇒(1). (1)⇒(2): We note that f is in the Herglotz -- Agler class HA(Πd,L(U)) if and only if s = (f − I)(f + I)−1 is in the Schur -- Agler class SA(Πd,L(U)). Thus by SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 31 Theorem 4.1 this s has a Πd-Schur -- Agler decomposition I − s(z)∗s(w) = dXk=1 (zk + wk)eKk(z, w) for some L(U)-valued positive kernels eKk on Πd. A standard computation gives I − s(z)∗s(w) = 2(f (z)∗ + I)−1[f (z)∗ + f (w)](f (w) + I)−1 from which we see that (5.2) holds with Kk(z, w) = 1 2 (f (z)∗ + I)eKk(z, w)(f (w) + I). (2)⇒(3): We use the Kolmogorov decompositions Kk(z, w) = Hk(z)∗Hk(w) of the positive kernels Kk (with Hk : Πd → L(U , eXk) say) to rewrite the Agler decomposition (5.2) in the form f (z)∗ + f (w) = H(z)∗(P (z)∗ + P (w))H(w) (5.4) where we have set  . wdI eXd Just as in the proof of Theorem 4.1, we introduce the subspace . . . ... H1(w) Hd(w) w1I eX1 H(w) =  , P (w) = eX0 := span{Ran H(w) : w ∈ Πd} ⊂ and introduce operators eY1, . . . ,eYd on eX0 by eYk = P eX0 We then view (5.2) as the statement that the subspace Pk eX0 . dMk=1 eXk =: eX (5.5) (5.6) (5.7) is an isotropic subspace of eK when eK is given the Kreın-space inner product induced by the Gramian matrix ⊂ eX0 eX0 U U  =: eK eG = span  eY (w)H(w) f (w) H(w) IU  u : u ∈ U , w ∈ Πd  . 0 I eX0 0 −IU 0 0 0 0 0 0 0 0 0 I eX0 0 −IU eJ = We show next that eG is a graph space, i.e.,  = {0}. eG ∩ U {0} {0} eX0 32 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI Toward this end, suppose that x ⊕ u ⊕ 0 ⊕ 0 ∈ eG. Our goal is to show that then necessarily x = 0 and u = 0. As eG is eJ -isotropic, we necessarily have, for all u′ ∈ U, + f (w)u′ H(w)u′  , 0 =*eJ eY (w)H(w)u′ =(cid:28)(cid:20)H(w)u′ −u′ (cid:21) ,(cid:20)x u(cid:21)(cid:29) = hu′, H(w)∗x − ui u′ x u 0 0  (5.8) (5.9) from which we conclude that We note that, as a consequence of (5.4), u = H(w)∗x for all w ∈ Πd. H(te)∗H(te) = f (te)∗ + f (te) 2t . The growth assumption (5.1) then implies that H(te) → 0 strongly as t → +∞. To conclude the proof of (5.7), we now need only specialize (5.9) to the case w = te and take a weak limit as t → +∞ to show that u = 0. It then follows from (5.8) that x is orthogonal to span{H(w)u : w ∈ Πd, u ∈ U}. As a consequence of (5.5), this in turn forces x = 0 and (5.7) follows. We next embed eG into a J -Lagrangian subspace G of X ⊕ U ⊕ X ⊕ U, where X is a Hilbert space containing eX as a subspace and the indefinite Gramian matrix has the same form as eJ in (5.6) above: J = 0 0 0 0 0 IX 0 −IU 0 IX 0 −IU 0 0 0 0 in such a way that G is still a graph subspace. That this is possible follows via a minor adjustment of Proposition 2.5 in [12] as indicated in the proof of Theorem 4.12 there. We also note that condition (2) in Example 2.6 is automatic since the  (5.10) subspace eG satisfies this condition by construction. It then follows that G has the form (4.7) for a closed operator U =(cid:20)A&B C&D(cid:21) : D(U) ⊂(cid:20)X U(cid:21) →(cid:20)X Y(cid:21) . That U is a Π-impedance-conservative system node follows from the fact that G is J -Lagrangian (with J given by (5.10)). We also extend the positive decomposition on X just as in the proof of (2)⇒(3) in Theorem 4.1 above. That we recover f (w) as the transfer-function for the system node U, i.e., the formula (5.3) holds, now follows exactly as in the proof of Theorem 4.1. of the identityeY1, . . . ,eYd on eX0 to a positive decomposition of the identity Y1, . . . , Yd (3)⇒(2): We follow the proof of (3)⇒(2) in Theorem 4.1. If we define H(w) = i (well-defined by Proposition 2.10), we arrive at (4.11) and A&Bh (Y (w)−AX )−1 B I SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 33 (4.12), but with J given by (5.10) rather than by (4.5), leading to the adjusted final conclusion 0 = h[H(z)∗(Y (z)∗ + Y (w))H(w) − (f (z)∗ + f (w))] u, u′i. This leads to the Agler decomposition (5.2) with Kk(z, w) = H(z)∗YkH(w) = [Y 1/2 k H(z)]∗[Y 1/2 k H(w)] as in the proof of (3)⇒(2) in Theorem 4.1. It remains to show that the growth condition (5.1) necessarily holds if f has a realization (5.3) from a Π-impedance-conservative system node. To see this, we note that then the single-variable Herglotz function f (se) (s ∈ Π) has an impedance- conservative system-node realization. That the growth condition (5.1) holds now follows from the result for the single-variable case (see Theorem 7.4 in [35]). (2)⇒(1): The proof is parallel to the proofs of (2)⇒(1) in Theorems 3.1 and 4.1. Write the Agler decomposition (5.2) in the form (5.4) and observe that the functional calculus gives f (A)∗ + f (A) = H(A)∗(P (A)∗ + P (A))H(A). If A is a strictly accretive commutative d-tuple, from the diagonal form of P (w) we see that P (A)∗ + P (A) ≥ 0. We conclude that f ∈ HA(Πd,L(U)), and (1) follows. (cid:3) In [3] a criterion was given for when a Πd-Herglotz -- Agler function has a real- ization involving the structured resolvent (P (w) − A)−1 coming from a spectral decomposition (P1, . . . , Pd) (so P (w) = w1P1 + ··· + wdPd) rather than just a pos- itive decomposition (Y1, . . . , Yd) of the identity). We give our version of a result of this type, with realization in terms of a Π-impedance-conservative system node rather than in the form presented in [3]. Theorem 5.2. Suppose that f : Πd → L(U) is a Πd-Herglotz -- Agler function satis- fying the growth condition (5.1). Then the following are equivalent: (1) There exists a Hilbert space X and a d-fold spectral decomposition (P1, . . . , Pd) of IX along with a Π-impedance-conservative system node such that C&D(cid:21) : D(U) ⊂(cid:20)X Y =(cid:20)A&B U(cid:21) →(cid:20)X U(cid:21) f (w) = C&D(cid:20)(cid:0)(P (w) − A)X(cid:1)−1 (cid:21) . 1−ζ(cid:17) − Iihf(cid:16) 1+ζ 1−ζ(cid:17) + Ii−1 1−ζd(cid:17) if ζ = (ζ1, . . . , ζd) ∈ Dd), then S is in the (where by convention , . . . , 1+ζd (1.4) 1+ζ Schur -- Agler class SA(Dd,L(U)) and S has a realization as in (1.2) where U = [ A B C D ] is unitary with the additional property that 1 is not in the point spectrum of U. (2) If S(ζ) = C(f )(ζ) :=hf(cid:16) 1+ζ 1−ζ = (cid:16) 1+ζ1 1−ζ1 B IU Proof. (1)⇒(2): We suppose that we are given a Π-impedance-conservative system node Y as in condition (1). We set H(w) =(cid:0)(P (w) − A)X(cid:1)−1 B : U → X 34 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI (well-defined by Proposition 2.10) and verify that (cid:20)A&B C&D(cid:21)(cid:20)H(w)u u (cid:21) =(cid:20)P (w)H(w)u f (w)u (cid:21) (5.11) for all u ∈ U. Furthermore, working as in the proof of (3) ⇒ (2) in Theorem 5.1, we see that H(w) so defined provides an Agler decomposition for f : f (z)∗ + f (w) = (zk + wk)H(z)∗PkH(w). dXk=1 If we set eY = h −(A&B) C&D i, then the fact that Y = (cid:2) A&B C&D(cid:3) is an impedance- conservative system node means that eY is skew-adjoint (see Corollary 2.7): eY∗ = −eY. Moreover we can rewrite the identity (5.11) in the form eY(cid:20)H(w)u As eY is skew-adjoint, easily verified properties of the Cayley transform imply that U := (eY − I)(eY + I)−1 is unitary and the point 1 is not in the point spectrum of U (this is another version of Proposition 2.3). U =: [ A B (cid:21) =(cid:20)−P (w)H(w)u C D ] provides a Dd-scattering-conservative realization of S := C(f ). It remains only to check that f (w)u (5.12) (cid:21) u We rewrite (5.12) in terms of U as or equivalently We reorganize this using linearity to get u f (w)u (I + U)(I − U)−1(cid:20)H(w)u (I + U)(cid:20)H(w)u U(cid:20)(P (w) − I)H(w)u (cid:21) =(cid:20)−P (w)H(w)u (cid:21) (cid:21) = (I − U)(cid:20)−P (w)H(w)u (cid:21) −(f (w) + I)u (cid:21) =(cid:20)(P (w) + I)H(w)u −(f (w) − I)u (cid:21) . f (w)u u 1 w + 1(cid:19) (f (w) + I). (wk − 1)PkH(w) wk + 1eHk(cid:18) w − 1 dXk=1 dXk=1 wk + 1eHk(ζ)(f (w) + I) dXk=1 ζkeHk(ζ)(f (w) + I) = P (ζ)eH(ζ)(f (w) + I) wk − 1 = = We introduce L(U ,Xk)-valued functions eHk on the polydisk Dd according to the relation (where we again use the convention (1.5)) Hk(w) := PkH(w) = (5.13) Then we note that (P (w) − I)H(w) = SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 35 where we set and ζ = w − 1 w + 1 (as in (1.5)) for w ∈ Πd Similarly one can verify that (P (w) + I)H(w) = dXk=1 and we have arrived at the pair of identities dXk=1 eHk(ζ). eH(ζ) = wk + 1eHk(ζ)(f (w) + I) = eH(ζ)(f (w) + I), wk + 1 Fix a vector v ∈ U. Define uw = (I + f (w))−1v so v = (I + f (w))uw. Then (5.13) with u = uw can be rewritten in the form (5.14) (P (w) − I)H(w) = P (ζ)eH(ζ)(I + f (w)) (P (w) + I)H(w) = eH(ζ)(I + f (w)). (cid:21) =(cid:20) eH(ζ)v −S(ζ)v(cid:21) . U(cid:20)P (ζ)eH(ζ)v −v Writing out U = [ A B C D ], one can now solve (5.14) in the standard way to arrive at S(ζ) = D + CP (ζ)(I − AP (ζ))−1B, i.e., the unitary colligation matrix U with the additional property that U does not have 1 as an eigenvalue provides a Dd-scattering-conservative realization for S = C(f ). (2)⇒(1): We suppose that S = C(f ) has a Dd-scattering-conservative realization C D ] does not have 1 (1.2) where the associated unitary colligation matrix U = [ A B as an eigenvalue. Then we know that U also has the defining property U(cid:20)P (ζ)eH(ζ) (cid:21) u =(cid:20)eH(ζ) S(ζ)(cid:21) u for all u ∈ U and ζ ∈ Dd where eH(ζ) =   provides a Dd-Schur -- Agler decomposition (1.1). Since 1 is not an eigenvalue of U by assumption we may form the Cayley transform eH1(ζ) ... (5.15) eHd(ζ) I eY := (I + U)(I − U)−1 with D(eY) = Ran(I − U). By construction we have By Proposition 2.3, eY is skew-adjoint: eY : (I − U)(cid:20)x eY = −eY∗. u(cid:21) 7→ (I + U)(cid:20)x u(cid:21) . (5.16) (5.17) 36 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI From (5.15) we note that Notice that (I − S(ζ)) = 2(f (w) + I)−1 is invertible. Hence (5.17) leads to (I − U)(cid:20)P (ζ)eH(ζ) (I + U)(cid:20)P (ζ)eH(ζ) eY(cid:20)(P (ζ) − I)eH(ζ)(I − S(ζ))−1 I I I (I − S(ζ))u (cid:21) u =(cid:20)(P (ζ) − I)eH(ζ)u (cid:21) , (cid:21) u =(cid:20)(P (ζ) + I)eH(ζ) (I + S(ζ))u (cid:21) . (cid:21) u =(cid:20)(P (ζ) + I)eH(ζ)(I − S(ζ))−1u wk + 1eHk(cid:18) w − 1 w + 1(cid:19) (I + f (w)) (I + S(ζ))(I − S(ζ))−1u 1 Let us set where we use the convention (1.5) as usual, and note that Then we compute I + f (w) = 2(I − S(ζ))−1. Hk(w) = (cid:21) . From these identities we see that the identity (5.18) is equivalent to eY(cid:20)−H(w) I u ] ∈ D(eY). In particular, for u ∈ U then there is an xu = −H(w)u ∈ X so that [ xu As we have already observed that eY is skew-adjoint (see (5.16)), we now have all the hypotheses needed in order to apply Corollary 2.7 to conclude that Y = C&D(cid:3) is a Π-impedance-conservative system node. Moreover the eY(cid:2) −I 0 0 I(cid:3) =: (cid:2) A&B relation (5.19) leads to the relation Y(cid:20)H(w) I (cid:21) u =(cid:20)P (w)H(w) f (w) (cid:21) u. (5.18) (5.19) (P (ζ) − I)eH(ζ)(I − S(ζ))−1 = and similarly (P (ζ) + I)eH(ζ)(I − S(ζ))−1 = 2 = − = − wk + 1 − 1(cid:19) eHk(ζ)(I − S(ζ))−1 wk + 1eHk(ζ)(I − S(ζ))−1 Hk(w) = −H(w) dXk=1(cid:18) wk − 1 dXk=1 dXk=1 dXk=1(cid:18) wk − 1 dXk=1 wk + 1eHk(ζ)(I − S(ζ)−1 dXk=1 (cid:21) u =(cid:20)P (w)H(w) wkHk(w) = P (w)H(w). + 1(cid:19) eHk(ζ)(I − S(ζ))−1 (cid:21) u. wk + 1 f (w) 2wk = = SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 37 We now recover f (w) as the transfer-function for the Π-impedance-conservative system node Y exactly as in the proof of (2)⇒(3) in Theorem 5.1. (cid:3) Remark 5.3. The implication (2)⇒(1) in Theorem 5.2 is due essentially to Agler -- McCarthy -- Young [3] (without explicit reference to Π-impedance-conservative sys- tem nodes), where the result is worked out for the scalar-valued case in the Nevan- linna -- Agler (rather than Herglotz -- Agler) setting. 6. Bessmertnyı long resolvent representations for Herglotz -- Agler functions Bessmertnyı long-resolvent representations were introduced by Bessmertnyı in connection with the study of general rational matrix functions of several variables, with a special symmetrized form of such a representation handling functions f in the Πd-Herglotz -- Agler class having an extension to Ωd :=Sλ∈T(λΠ)d ⊂ Cd satisfying additional symmetry conditions (see [17, 18, 19, 20, 21, 28] and our companion paper [11] for more detail). In [9], a relaxation of the symmetrized Bessmertnyı long-resolvent representation was proposed which handles more general Herglotz -- Agler-class functions (i.e., the homogeneity property is discarded). Given a 2 × 2-block operator pencil (cid:20)V11(w) V12(w) V21(w) V22(w)(cid:21) =: V(w) = V0 + w1V1 + ··· + wdVd ∈ L(U ⊕ X ) (6.1) we define the transfer function fV(w) associated with the operator pencil V by fV(w) := V11(w) − V12(w)V22(w)−1V21(w). (6.2) wherever the formula makes sense. Let us say that a representation f (w) = fV(w) of the form (6.2) for a given f is a B-realization if the operator pencil V(w) satisfies (1) the B-symmetry condition V(w) = −V(−w)∗, namely V0 + V∗0 = 0, Vk = V∗k for k = 1, . . . , d, (6.3) and (2) the B-positivity condition Vk = V∗k ≥ 0 for k = 1, . . . , d with dXk=1 V22,k strictly positive definite. (6.4) It was shown in [9] that any function f of the form (6.2) with pencil V(w) satisfying conditions (6.3) and (6.4) is in the Herglotz -- Agler class HA(Πd,L(U)). a B-realization with operator pencil V(w) = V0 + w1V1 + ··· + wdVd given by We note that the realization (5.3) in Theorem 5.1 can formally be considered as V0 =(cid:20) D&C −(B&A)(cid:21) = −V∗0, Vk =(cid:20)0 0 0 Yk(cid:21) = V∗k ≥ 0 for k = 1, . . . , d meeting all the constraints (6.3) -- (6.4) for a B-realization of f with the exception that V0 is unbounded. We now make precise how a general Herglotz -- Agler-class function (i.e., with the growth condition at infinity (5.1) removed) can be com- pletely characterized in terms of possession of a nonhomogeneous Bessmertnyı-type representation of the form (6.2) subject to (6.3) and (6.4) but with possibly un- bounded skew-adjoint operator V0. 38 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI We shall consider unbounded pencils of the form (6.1) but with the constant term a possibly unbounded operator on(cid:20)U Agler system node properties: V0,21 V0,22(cid:21) V0 =(cid:20)V0,11 V0,12 X(cid:21) satisfying what we shall call the Herglotz -- The gist of conditions (HA1) and (HA2) is that the reorganized colligation matrix ]. (HA1) V0 is skew-adjoint on [ U X (HA2) For each u ∈ U there is an xu ∈ X so that [ u 0(cid:21) (cid:20)A&B C&D(cid:21) =(cid:20) 0 −IX 0 (cid:21) V0(cid:20) 0 IX IU IU xu ] ∈ D(V0). is an impedance-conservative system node as discussed in Example 2.6. To reduce the number of subscripts and to suggest the connection with system nodes in the work of Staffans et al. [29, 35, 36, 12], we shall use the notation encoded in the following formal definition. Definition 6.1. We shall say that the pencil (6.1) is a Herglotz -- Agler operator pencil if the following conditions hold: (1) V0 has the form I −(B&A)(cid:21) :=(cid:20) 0 V0 =(cid:20) D&C C&D(cid:21) is an impedance-conservative system node as in Example 2.6. where(cid:20)A&B (2) The homogeneous part of the pencil, VH(w) :=Pd 0(cid:21)(cid:20)A&B C&D(cid:21)(cid:20)0 k=1 wkVk has each 0(cid:21) −I I I Vk,21 Vk,22(cid:21) Vk =(cid:20)Vk,11 Vk,12 dXk=1(cid:20)Vk,11 Vk,12 Vk,21 Vk,22(cid:21) =(cid:20)VH,11(e) 0 0 I(cid:21) a bounded positive semidefinite operator on U ⊕ X with the sum having the form VH(e) = (6.5) where we use the notation e = (1, . . . , 1) as in (5.1). Let us suppose that V(w) = V0 + VH(w) =(cid:20) D&C is a Herglotz -- Agler operator pencil. Thus in particular (cid:20)0 −I −(B&A)(cid:21) +(cid:20)VH,11(w) VH,12(w) VH,21(w) VH,22(w)(cid:21) 0 (cid:21) V0(cid:20)0 I I all the properties delineated in Definition 2.4, with the additional property that V0 = −V∗0 (see Example 2.6). In particular we may write I 0(cid:21) has Here AX : X → X−1 is the extension of the skew-adjoint operator A defined by . B&A = (cid:2)B AX(cid:3)(cid:12)(cid:12)D(V0) x(cid:21) ∈ D(V0)(cid:27) with Ax = −V0(cid:20)0 x(cid:21) D(A) =(cid:26)x : (cid:20)0 SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 39 and with B : U → X−1 constructed via Bu = B&A(cid:20) u xu(cid:21) − AX x where xu is any choice of vector in X such that [ u xu ] ∈ D(V0); it can be shown that B is well defined, i.e., the formula for Bu is independent of the choice of xu ∈ X . Furthermore, the domain of V0 is the same as the domain of B&A and has the precise characterization D(V0) = D(B&A) =(cid:26)(cid:20)u x(cid:21) ∈(cid:20)U X(cid:21) : Bu + AX x ∈ X(cid:27) . The following proposition gives some additional key properties for Herglotz -- Agler operator pencils. Proposition 6.2. Suppose that V(w) = V0 + VH(w) :=(cid:20) D&C −(B&A)(cid:21) +(cid:20)VH,11(w) VH,12(w) VH,21(w) VH,22(w)(cid:21) is a Herglotz -- Agler operator pencil. Then, for a given w ∈ Πd, the formal Bess- mertnyı transfer function (6.2), fV(w) = V11(w) − V12(w)V22(w)−1V21(w) = (D + VH,11(w)) − (C + VH,12(w))(VH,22(w) − A)−1(−B + VH,21(w)), can be interpreted as a bounded operator on U equal to the sum of five well-defined bounded terms where fV(w) = fV,1(w) + fV,2(w) + fV,3(w) + fV,4(w) + fV,5(w) (6.6) I fV,1(w) = D&C(cid:20) (VH,22(w) − A)−1B(cid:21) , fV,2(w) = VH,12(w)(cid:0)(VH,22(w) − A)X(cid:1)−1 fV,3(w) = −C(VH,22(w) − A)−1VH,21(w), fV,4(w) = −VH,12(w)(VH,22(w) − A)−1VH,21(w), fV,5(w) = VH,11(w). VH,21(w), (6.7) Directly in terms of the Bessmertnyı pencil V(w), we have, for each w ∈ Πd, (cid:20) −V22(w)−1V21(w)(cid:21) ∈ L(U ,D(V0)), I V11(w)&V12(w) ∈ L(D(V0),U), (6.8) (6.9) (6.10) and we recover fV(w) as the composition of bounded operators fV(w) = V11(w)&V12(w) ·(cid:20) −V22(w)−1V21(w)(cid:21) ∈ L(U). I Proof. Analysis of fV,1: It follows from part (4) of Proposition 2.10 that the oper- atorh ((VH,22(w)−A)X )−1 that h ((VH,22(w)−A)X )−1 IU IU i maps D(V0) into D(V0). Furthermore, one can check i is bounded as an operator from U to D(V0) (with D(V0) equipped with the graph norm of VH,22(w)). As D&C has domain equal to B B 40 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI B IU D(V0), we see that D&C maps D(V0) into U. Moreover, part (4) of Definition 2.4 assures us that D&C is bounded as an operator from D(V0) into U. We conclude that fV,1(w) = D&C ·h ((VH,22(w)−A)X )−1 Analysis of fV,2: By part (2) of Proposition 2.10,(cid:0)(VH,22(w) − A)X(cid:1)−1 it follows that fU,2(w) = VH,12(w) ·(cid:0)(VH,22(w) − A)X(cid:1)−1 B maps U boundedly into X . As VH,12(w) is bounded as an operator from X into U, B is bounded as an i ∈ L(U). operator on U. Analysis of fV,3: It is a consequence of part(2) of Proposition 2.10 that the operator (VH,22(w)− A)−1 is bounded from X into X1. From the definition (2.8) of C, we see that C maps X1 boundedly into U. Since also VH,21(w) is bounded as an operator from U to X , it follows that fV,3(w) = −C · (VH,22(w) − A)−1 · VH,21(w) defines a bounded operator on U. Analysis of fV,4: Note that VH,21(w) ∈ L(U ,X ), by part (3) of Proposition 2.10 (VH,22(w) − A)−1 is bounded as an operator from X into X1 and hence as an operator from X into itself, and VH,12(w) is bounded as an operator from X into U. It follows that fV,4(w), as a composition of bounded operators, is bounded as an operator on U. Analysis of fV,5: This is the easiest term: VH,11(w) is a bounded operator on U from the definition of Herglotz-Agler pencil (Definition 6.1). Verification of formula (6.10): We first write out V(w) in terms of constant term and homogeneous part: Thus V(w) = V0 + VH,k(w) =(cid:20) D&C −(B&A)(cid:21) +(cid:20)VH,11(w) VH,12(w) VH,21(w) VH,22(w)(cid:21) . V11(w)&V12(w) = D&C +(cid:2)VH,11(w) VH,12(w)(cid:3) . The first term maps D(U0) boundedly into U while the second term is bounded from the larger space [ U ] into U. It follows that the sum indeed is bounded from X D(U0) into U, verifying property (6.9). Similarly, (cid:20) I −V22(w)−1V21(w)(cid:21) =(cid:20) =(cid:20) I (VH,22(w) − A)−1(B − VH,21(w)(cid:21) (VH,22(w) − A)−1B(cid:21) −(cid:20) I Proposition 2.10 while the second term maps U boundedly into (cid:2) 0 The first term maps U boundedly into D(U0) as a consequence of part (4) of (and hence also boundedly into D(U0)) by part (2) of Proposition 2.10; this verifies property (6.8). Thus the composition in the formula (6.10) defines a bounded operator on U. Working out the various pieces in detail, we see that the result agrees with the formula for fV(w) in (6.6). (cid:3) 0 (VH,22(w) − A)−1VH,21(w)(cid:21) . X1(cid:3) ⊂ D(U0) The following is the main result of this section. Theorem 6.3. Given a function f : Πd → L(U), the following are equivalent: (1) f is in the Herglotz -- Agler class HA(Πd,L(U)). SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 41 (2) f has a Πd-Herglotz -- Agler decomposition, i.e., there exist L(U)-valued pos- itive kernels K1, . . . , Kd on Πd such that f (z)∗ + f (w) = (3) There exists a Hilbert space X and a Herglotz -- Agler pencil dXk=1 V(w) = V0 + VH(w) =(cid:20) D&C −(B&A)(cid:21) +(cid:20)VH,11(w) VH,12(w) VH,21(w) VH,22(w)(cid:21) (zk + wk)Kk(z, w). such that f (w) = fV(w) (with fV(w) as in (6.6) -- (6.7) or (6.10)). Proof. (1) ⇔ (2): This is already done in the proof of Theorem 5.1. F over the polydisk Dd: (1) or (2) ⇒ (3): We start with the representation for a Herglotz -- Agler function F (ζ) = R + V ∗(U − P (ζ))−1(U + P (ζ))V (6.11) where U is unitary, P (ζ) = ζ1P1 + ··· + ζdPd is a spectral decomposition of the identity on the state space X , R = F (0)−F (0)∗ as in formula (3.10). We use the tuple version of the mutually inverse Cayley changes of variable (1.3): and V ∗V = F (0)+F (0)∗ 2 2 1 + ζ 1 − ζ ∈ Πd, w ∈ Πd 7→ w − 1 w + 1 ∈ Dd, ζ ∈ Dd 7→ in the Herglotz -- Agler class over the polydisk Dd and hence we can represent F (ζ) with the conventions (1.4) and (1.5) in force. If we set F (ζ) = f(cid:16) 1+ζ as in (6.11). Moreover, we recover f (w) from F (ζ) via f (w) = F(cid:16) w−1 1−ζ(cid:17), then F is w+1(cid:17). This leads to the formula f (w) = R + V ∗M (w)V (6.12) where we set M (w) =(cid:18)U − P(cid:18) w − 1 w + 1(cid:19)(cid:19)−1(cid:18)U + P(cid:18) w − 1 w + 1(cid:19)(cid:19) . We compute further M (w) =(cid:0)U − (P (w) − I)(P (w) + I)−1(cid:1)−1(cid:0)U + (P (w) − I)(P (w) + I)−1(cid:1) = (P (w)U + U − P (w) + I)−1 (P (w)U + U + P (w) − I) = (P (w)(U − I) + (U + I))−1 (P (w)(U + I) + (U − I)) (6.13) Let us split out the eigenspace of U for eigenvalue z = 1 (if any) by writing U in the form U =(cid:20)I 0 U0(cid:21) 0 with respect to the decomposition X = X (1) ⊕X (0) (X (1) equal to the 1-eigenspace for U and X (0) equal to the orthogonal complement of X (1) in X ). Then U0 is unitary but does not have 1 as an eigenvalue. We have U − I =(cid:20)0 0 U0 − I(cid:21) , U + I =(cid:20)2I 0 U0 + I(cid:21) . 0 0 42 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI and hence M (w) =(cid:18)P (w)(cid:20)0 0 U0 − I(cid:21) +(cid:20)2I 0 0 0 U0 + I(cid:21)(cid:19)−1 ·(cid:18)P (w)(cid:20)2I T = (I + U0)(I − U0)−1. Let us set 0 0 U0 + I(cid:21) +(cid:20)0 0 U0 − I(cid:21)(cid:19) 0 Then T is a possibly unbounded skew-adjoint operator on X0 (by Proposition 2.3) and we have the following relations between T and U0: U0 = (T − I)(T + I)−1 = I − 2(T + I)−1 We may then continue the computation (6.13) to get M (w) =(cid:20) 1 0 0 0 0 2 I 0 (T + I). ⇒ U0 − I = −2(T + I)−1, (U0 − I)−1(cid:21)(cid:18)P (w)(cid:20)0 (U0 − I)−1 = − I(cid:21) +(cid:20)I (U0 + I)(U0 − I)−1(cid:21) +(cid:20)0 0 I(cid:21) +(cid:20)I 0 I(cid:21)(cid:19)(cid:20)I 0 −(T + I)∗(cid:21) (U0 + I)(U0 − I)−1(cid:21)(cid:19)−1 I(cid:21)(cid:19)(cid:20)2I 0 −T(cid:21)(cid:19)−1 0 −(T + I)−1(cid:21) ·(cid:18)P (w)(cid:20)I =(cid:20)I 0 −(T + I)(cid:21)(cid:18)P (w)(cid:20)0 ·(cid:18)P (w)(cid:20)I 0 −T(cid:21) +(cid:20)0 =(cid:20)I 0 −(T + I)(cid:21) N (w)(cid:20)I 0 U0 − I(cid:21) 0 0 0 0 0 0 0 0 0 0 0 0 0 · 0 1 2 · (6.14) 0 0 0 0 0 If we write out the block matrix decomposition where, due to the identity −(T + I)∗ = T − I arising from T = −T ∗, N (w) is given by N (w) =(cid:18)P (w)(cid:20)0 I(cid:21) +(cid:20)I 0 −T(cid:21)(cid:19)−1 ·(cid:18)P (w)(cid:20)I 0 −T(cid:21) +(cid:20)0 0 I(cid:21)(cid:19) ·(cid:20)I Pk,01 Pk,00(cid:21) wk(cid:20)Pk,11 Pk,10 P (w) =(cid:20)P11(w) P10(w) P01(w) P00(w)(cid:21) = dXk=1 X (0)(cid:21) of X , we can write out of P (w) with respect to the decomposition X = (cid:20)X (1) (cid:18)P (w)(cid:20)0 =(cid:20)I 0 P00(w) − T(cid:21)−1 (I − T 2)−1(cid:21) . 0 −T(cid:21)(cid:19)−1 I(cid:21) +(cid:20)I more explicitly P10(w) (6.15) 0 0 0 0 0 =(cid:20)I −P10(w)(P00(w) − T )−1 (P00(w) − T )−1 0 (cid:21) , SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 43 P (w)(cid:20)I 0 −T(cid:21) +(cid:20)0 0 and from (6.15) we see that N (w) is given by N (w) =(cid:20)I −P10(w)(P00(w) − T )−1 (P00(w) − T )−1 0 0 P01(w) 0 I(cid:21) =(cid:20)P11(w) −P10(w)T I − P00(w)T(cid:21) (I − P00(w)T )(I − T 2)−1(cid:21) . −P10(w)T (I − T 2)−1 (cid:21)(cid:20)P11(w) P01(w) (6.16) At this stage it is convenient to introduce the Gelfand triple (or rigging) of X (0) associated with the (possibly unbounded) skew-adjoint operator T : 1 X (0) := Dom T = Ran(I − T )−1, X (0) −1 := completion of X (0) in X (0) Then we see that (I − T )−1 is well defined as an element of L(X (0),X (0) L(X (0) −1 ,X (0)). A careful inspection of the formula (6.16) for N (w) shows that −1 -norm: kxk−1 = k(I − T )−1xkX (0) for x ∈ X (0). ) and of 1 (6.17) (6.18) 1 (cid:21) −1(cid:21) →(cid:20)X (1) N (w) : (cid:20)X (1) X (0) X (0) J := −T (I − T 2)−1, from which it follows that the formula (6.14) gives sense for M (w) as an element of L(X (1) ⊕ X (0)). However the formula (6.16) (and (6.15)) lacks symmetry. To fix this we introduce the operator J by Thus J = −J∗ and we can consider J as an element of L(X (0),X (0) as L(X (0) L(X (0) ) as well −1 ,X (0)). While the operator I − P00(w)T makes sense as an element of ,X (0)) (as well as L(X (0),X (0) −1 )), the individual terms in the additive de- 1 1 composition I − P00(w)T = (I − T 2) + (T 2 − P22(w)T ) make sense only as elements in L(X (0) (6.19) −1 ). Nevertheless, we proceed to get a more symmetric formula for N (w) as follows. Note first that the decomposition (6.19) leads to ,X (0) 1 (I − P00(w)T )(I − T 2)−1 = I − (P00(w) − T )T (I − T 2)−1 0 P10(w)J (P00(w) − T )−1 From (6.16) and the definition (6.18) of J, we then have N (w) =(cid:20)I −P10(w)(P00(w) − T )−1 = I + (P00(w) − T )J : X (0) → X (0). (cid:21)(cid:20)P11(w) I + (P00(w) − T )J(cid:21) =(cid:20)P11(w) − P10(w)(P00(w) − T )−1P01(w) −P10(w)(P00(w) − T )−1 (P00(w) − T )−1 + J (cid:21) −1(cid:21) to(cid:20)X (1) X (0)(cid:21) rather than which a priori makes sense only as an operator from(cid:20)X (1) X (0) 1 (cid:21) as in (6.17), due to the decoupling of an ∞ − ∞ cancellation occurring to(cid:20)X (1) X (0) (P00(w) − T )−1P01(w) in the application of the decomposition (6.19). This in turn leads to difficulties P01(w) (6.20) 44 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI in understanding M (w) as a bounded operator on X (1) ⊕ X (0) from the formula (6.14). Continuation of the analysis with an extra assumption: Assuming for the moment that T is bounded (as is the case for the situation where the state space X is finite-dimensional as in the setting discussed in [11]), this difficulty does not occur and we may continue as follows. From (6.20) we see that N (w) =(cid:20)P11(w) 0 0 J(cid:21) −(cid:20)P10(w) −I (cid:21) (P00(w) − T )−1(cid:2)P01(w) If we block-decompose the operator V : U → X = (cid:20)X (1) combine (6.12) with (6.14) and (6.21) while noting the simplification (6.21) I(cid:3) . V0(cid:21) and then X (0)(cid:21) as V =(cid:20)V1 (I + T )J(I + T )∗ = (I + T )[−(I + T )−1T (I − T )−1](I + T )∗ (by (6.18)) = −T (since T ∗ = −T ). we arrive at f (w) = R + V ∗1 P11(w)V1 − V ∗0 T V0 − [V ∗1 P10(w) + V ∗0 (I + T )](P00(w) − T )−1[P01(w)V1 − (I + T )∗V0]. We have arrived at a Bessmertnyı long-resolvent representation for f (6.22) where the operator pencil V(w) =(cid:20)V11(w) V12(w) V21(w) V22(w)(cid:21) is given by f (w) = fV(w) V11(w) = R + V ∗0 (I + T )J(I + T )∗V0 + V ∗1 P11(w)V1, V12(w) = V ∗0 (I + T ) + V ∗1 P10(w), V21(w) = −(I + T )∗V0 + P01(w)V1, V22(w) = −T + P00(w). Thus the associated linear pencil V21(w) V22(w)(cid:21) = V0 + w1V1 + ··· + wdVd (6.23) (6.24) (6.25) has coefficients −(I + T )∗V0 V(w) =(cid:20)V11(w) V12(w) V0 =(cid:20) R − V ∗0 T V0 =(cid:20)R 0 0(cid:21) +(cid:20)V ∗0 Vk =(cid:20)V ∗1 Pk,11V1 V ∗1 Pk,10 k=1 Pk,00 = IX (0) (sincePd Pk,01V1 0 0 0 −T V ∗0 (I + T ) (cid:21) I(cid:21)(cid:20) −T −(I + T )∗ −T (cid:21)(cid:20)V0 Pk,00 (cid:21) for k = 1, . . . , d. I + T 0 0 I(cid:21) , Moreover, it is easily checked that V0 is skew-adjoint and that Pk ≥ 0 for each k with Pd k=1 Pk = IX ), and hence V(w) is a Herglotz -- Agler pencil and Theorem 6.3 is completely proved in case T = (I + U0)(I − U0)−1 is bounded on X (0). Back to the general case: For the general case (where T is allowed to be unbounded), the formula (6.25) for Vk still makes good sense and the Vk's meet SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 45 property (2) in Definition 6.1. The next step is to make sense of the formula (6.24) for V0. The first term in the formula (6.24) for V0 can always be added in later so we focus on the second term V′0: V′0 =(cid:20)V ∗0 0 0 I(cid:21)(cid:20) −T −(I + T )∗ −T (cid:21)(cid:20)V0 I + T 0 0 I(cid:21) . We view V′0 as a possibly unbounded operator with dense domain in(cid:20)U X(cid:21) given by Then the flip of V′0, namely D(V′0) =(cid:26)(cid:20)u x(cid:21) ∈(cid:20)U 0 (cid:21) V′0(cid:20)0 0(cid:21) =(cid:20)I X(cid:21) : x − V0u ∈ D(T )(cid:27) . 0 V ∗0(cid:21)(cid:20) T −T (cid:21)(cid:20)I I − T I + T 0 I I (cid:20)0 −I I 0 0 V0(cid:21) has exactly the form of the model Π-impedance-conservative system node given in Proposition 2.8. We can now conclude that V(w) given by (6.23), (6.24), (6.25) is indeed a Herglotz -- Agler pencil. It remains only to verify that f (w) = fV(w). From the formula (6.12) for f combined with the formula (6.14) for M (w) and the formula (6.20) for N (w), we know that f (w) has the representation f (w) = R + V ∗1 P11(w)V1 − V ∗1 P10(w)(P00(w) − T )−1P01(w)V1 − V ∗0 (I + T )(P00(w) − T )−1P01(w)V1 + V ∗1 P10(w)[T (I + T )−1 + (P00(w) − T )−1(I − P00(w)T )(I + T )−1]V0 + V ∗0 (I + T )(P00(w) − T )−1(I − P00(w)T )(I + T )−1V0. (6.26) On the other hand the Bessmertnyı transfer function associated with the pencil V (6.23) can be written as fV(w) = [V11(w)&V12(w)](cid:20) −V22(w)−1V21(w)(cid:21) I = R + V ∗0 (cid:2)−T + (I + T )(P00(w) − T )−1(I − T )(cid:3) V0 + V ∗1 P11(w)V1 − V ∗1 P10(w)(P00(w) − T )−1P01(w)V1 − V ∗0 (I + T )(P00(w) − T )−1P01(w)V1 + V ∗1 P10(w)(P00(w) − T )−1(I − T )V0. (6.27) −1 and V ∗0 ∈ L(X (0),U) has no extension to X (0) Note that care must be taken in writing the first term of the expression after the R term: the individual expressions −V ∗0 T V0 and V ∗0 (I + T )(P00(w) − T )−1(I − T )V0 make no sense since the operators −T and (I + T )(P00(w)− T )−1(I − T ) map X (0) into X (0) −1 ; as we shall see in detail below, the combination −T + (I + T )(P00(w)− T )−1(I − T ) fortuitously maps X (0) back into itself so that the combined term V ∗0 (−T +(I +T )(P00(w)−T )−1(I−T ))V0 makes good sense as a bounded operator on U for each w ∈ Πd; roughly speaking, this is where we couple back together the ∞ − ∞ cancellation introduced earlier to make our formulas once again make sense. Note also that the formula (6.27) agrees with the formula (6.22) once one takes care to rearrange the terms in (6.22) so that the result makes sense as a well-defined bounded operator on U defining the operator f (w). 46 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI By the analysis done above with the extra assumption imposed, we see that the two expressions (6.26) and (6.27) agree in the special case where the skew-adjoint operator T is bounded. Once the operator −T +(I +T )(P00(w)−T )−1P01(w)(I−T ) is exhibited more explicitly as a bounded operator on X (0) (even in the case where T itself is unbounded), it is possible to verify the equality of the two expressions (6.26) and (6.27) by approximating the unbounded case by the bounded case and then taking limits. As this is really about algebra, however, perhaps more satisfying is to verify the equality between (6.26) and (6.27) directly by brute-force algebra. Toward this goal, we note that each term in (6.26) can be paired with an identical term in (6.27) once we establish the validity of the two identities: (I + T )(P00(w) − T )−1(I − P00(w)T )(I + T )−1 = −T + (I + T )(P00(w) − T )−1(I − T ) P10(w) [T (I + T )−1 + (P00(w) − T )−1(I − P00(w)T )(I + T )−1] = P10(w)(P00(w) − T )−1(I − T ). (6.28) In particular, (6.28) demonstrates how the expression −T + (I + T )(P00(w) − T )−1(I − T ) actually defines a bounded operator on X . These two identities can be verified directly by brute-force algebra; we leave the details to the reader (or as an exercise for MATHEMATICA). This concludes the proof of (1) or (2) ⇒ (3) in Theorem 6.3. (3) ⇒ (2): We assume that f (w) = fV(w) for a Herglotz -- Agler pencil V(w) =(cid:20)V11(w)&V12(w) V21(w)&V22(w)(cid:21) =(cid:20) D&C −(B&A)(cid:21) +(cid:20)VH,11(w) VH,12(w) VH,21(w) VH,22(w)(cid:21) . Thus, as explained in Proposition 6.2, if we set xw = −V22(w)−1V21(w) for w ∈ Πd, then for each u ∈ U we have and We may also compute (cid:20) u xwu(cid:21) ∈ D(cid:18)(cid:20)V11(w)&V12(w) V21(w)&V22(w)(cid:21)(cid:19) f (w)u = (V11(w)&V12(w))(cid:20) u xwu(cid:21) −V22(w)−1V21(w)u(cid:21) −V22(w)−1V21(w)u(cid:21) (as a vector in X (0) u u −1 ) (6.29) (V21(w)&V22(w))(cid:20) =(cid:2)V21(w) V22(w)(cid:3)(cid:20) = V21(w)u − V21(w)u = 0. Thus (6.29) can be expanded to the identity (cid:20)V11(w)&V12(w) V21(w)&V22(w)(cid:21)(cid:20) u xwu(cid:21) =(cid:20)f (w) 0 (cid:21) u. (6.30) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 47 In addition to u ∈ U and w ∈ Πd, choose another pair u′ ∈ U and z ∈ Πd and consider the sesquilinear form Q(z, w)[u, u′] := As a consequence of (6.30) we see that (cid:28)(cid:20)V11(w)&V12(w) V21(w)&V22(w)(cid:21)(cid:20) u xwu(cid:21) ,(cid:20) u′ Q(z, w)[u, u′] =(cid:28)(cid:20)f (w)u xzu′(cid:21)(cid:29) +(cid:28)(cid:20) u (cid:21) ,(cid:20) u′ xzu′(cid:21)(cid:29) +(cid:28)(cid:20) u xwu(cid:21) ,(cid:20)V11(z)&V12(z) V21(z)&V22(z)(cid:21)(cid:20) u′ xwu(cid:21) ,(cid:20)f (z)u′ (cid:21)(cid:29) 0 0 xzu′(cid:21)(cid:29) . = h(f (z)∗ + f (w)) u, u′i. (6.31) On the other hand, from the decomposition of V(w) as V(w) = V0 + VH(w) with V0 = −V∗0, we have Q(z, w)[u, u′] =DV(w)(cid:20) u =D (VH(z)∗ + VH(w))(cid:20) xwu(cid:21) ,(cid:20) u′ xzu′(cid:21)E +D(cid:20) u −V22(w)−1V21(w)(cid:21) u,(cid:20) I xwu(cid:21) , V(z)(cid:20) u′ xzu′(cid:21)E I −V22(z)−1V21(z)(cid:21) u′E (zj + wj)hH(z)∗VjH(w)u, u′i (6.32) where we set H(w) =(cid:20) (6.32) gives us −V22(w)−1V21(w)(cid:21) ∈ L(U ,U ⊕ X ). Combining (6.31) and I f (z)∗ + f (w) = (zj + wj )H(z)∗VjH(w) = dXj=1 dXj=1 where Vj ≥ 0 on U ⊕ X by assumption, and (2) follows. (cid:3) We next illustrate Theorem 6.3 by looking at some special cases. Special case 1: V1 = 0. We note that the case V1 = 0 in the proof of Theorem 6.3 is exactly the case where the representation (6.27) for f collapses to f (w) = R + V ∗0 [−T + (I + T )(P00(w) − T )−1(I − T )]V0. (6.33) where P00(w) = w1P1,00 + ··· + wdPd,00 is a positive decomposition of IX (0) , i.e., f has a representation exactly as in part (3) of Theorem 5.1. In general, from the property (6.5) for a Herglotz -- Agler pencil, we have VH(te) = tVH(e) = t(cid:20)VH,11(e) 0 0 I(cid:21) . Thus, in the general representation (6.27) for f , f (te) takes on the simplified form f (te) = R + V ∗0 [−T + (I + T )(tI − T )−1(I − T )]V0 + tV ∗1 V1. We have already seen that 1 t lim t→+∞ [R + V ∗0 (−T + (I + T )(tI − T )−1(I − T ))V0] = 0. We conclude that in general 1 t lim t→+∞ f (te) = V ∗1 V1. 48 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI Thus the growth condition at ∞ (5.1) is equivalent to the condition that V1 = 0. In this way we arrive at Theorem 5.1 as a corollary of Theorem 6.3. Special case 2: X (1) = {0}. This corresponds to the case where Pk = Pk,00 and P00(w) = w1P1,00 + ··· + wdPd,00 is a spectral decomposition (not just a positive decomposition) of IX (0) . Then the representation (6.27) collapses again to (6.33), but this time with the stronger property that w1P1,00 + ··· + Pd,00wd is a spectral decomposition of IX (0) , i.e., exactly the conclusion of part (1) of Theorem 5.2. On the one hand, the condition X (1) = {0} means that the unitary operator U in the Herglotz representation (3.10) for the function F ∈ HA(Dd,U) given by 1 − ζ(cid:19) F (ζ) = f(cid:18) 1 + ζ 6.3 is that the colligation matrix U = (cid:20)A B C D(cid:21) in the unitary Givone -- Roesser does not have 1 as an eigenvalue. On the other hand, condition (1) in Theorem representation S(ζ) = D + C(I − P (ζ)A)−1P (ζ)B (6.34) for the SA(Dd,U)-class function S(ζ) =(cid:18)f(cid:18) 1 + ζ 1 − ζ(cid:19) − I(cid:19)(cid:18)f(cid:18) 1 + ζ 1 − ζ(cid:19) + I(cid:19)−1 = (F (ζ) − I)(F (ζ) + I)−1 (6.35) does not have 1 as an eigenvalue. To see that these conditions match up, we recall from the discussion in Remark 2.11 (note formulas (3.7), (3.9) and (3.10)) that we have the following connection between the unitary operator U in the Herglotz representation (3.10) for F (ζ) and the unitary colligation matrix U = (cid:20)A B C D(cid:21) generating the Givone -- Roesser representation (6.34) for S(ζ) given by (6.35): U∗ = U0 := A − B(I − D)−1C. As was observed in Remark 3.2, I − D is injective since I − D = I − S(0) = 2(F (0) + I)−1. By the theory of Schur complements, given that I − D is injective, then I − U is injective if and only if its Schur complement U0 = I − B(I − D)−1C is injective. As U0 is unitary, this in turn is equivalent to U = (U0)∗ not having 1 as an eigenvalue. In this way we recover Theorem 5.2 as a corollary of Theorem 6.3. Special case 3: X (0) = {0}. In this case, M (w) = P (w), and the representation (6.27) collapses to f (w) = V(w) = R + V ∗P (w)V. Special case 4: R = 0, U = U∗, V0 = 0. In this case, T = T ∗ = −T ∗, hence T = 0, and the linear pencil A(w) is homogeneous: Moreover, V(e) =(cid:20)V ∗1 V1 0 0 0 0 0 0 I(cid:21) P (w)(cid:20)V1 V(w) = VH(w) =(cid:20)V ∗1 I(cid:21) . I(cid:21), as in (6.5). The representation f (w) = V ∗1 (P11(w) − P10(w)P00(w)−1P01(w))V1 = VH,11(w) − VH,12(w)VH,22(w)−1VH,21 (6.36) (6.37) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 49 is then Bessmertnyı's long-resolvent representation in the infinite-dimensional set- ting, as in [28], i.e., f belongs to the Bessmertnyı class Bd(U). Remark 6.4. Notice that for a function f ∈ Bd(U) one can always find a ho- mogeneous Bessmertnyı pencil VH(w) satisfying the condition (6.5). First of all, Indeed, if ker VH,22(e) 6= {0}, then the maximum VH,22(e) must be invertible. principle and positivity of the coefficients Vk, k = 1, . . . , d, force VH(w) to have the form VH(w) = in contradiction with condition (6.5) for a Herglotz-Agler pencil. Therefore, one can replace the pencil VH(w) by the pencil U Ran VH,22(e) Ker VH,22(e)  , 0 0 U 0 0 0 VH,11(w) Ran VH,22(e) Ker VH,22(e)  :   → 0 VH,22(e)−1(cid:21) , eVH,12(w) eVH,22(w) 0 VH,22(e)−1(cid:21) VH(w)(cid:20)I eVH,21(w) bVH(w) =(cid:20)I f (w) = bVH,11(w) − bVH,12(w)bVH,22(w)−1bVH,21. 0 0 which satisfies the condition (6.5) and provides another Bessmertnyı transfer-function realization of f : In fact, the special case described in Remark 6.4 covers the whole class Bd(U). In the following theorem, we collect the characterizations of this class from [28], together with the two additional characterizations: via its image in the class HA(Dd,L(U)) under the Cayley transform over the variables, and as this special case of Theorem 6.3. Theorem 6.5. Given a function f : Πd → L(U), the following are equivalent: (1) f is in the Herglotz -- Agler class HA(Πd,L(U)) and can be extended to a (λΠ)d which satisfies the holomorphic L(U)-valued function on Ωd := Sλ∈T following conditions: (a) Homogeneity: f (λw1, . . . , λw2) = λf (w1, . . . , wd) for every λ ∈ C\{0} (b) Real Symmetry: f (w1, . . . , wd) = f (w1, . . . , wd)∗ for every w ∈ Ωd. kernels K1, . . . , Kd on Πd such that the identity (2) f has a Πd-Bessmertnyı decomposition, i.e., there exist L(U)-valued positive and w = (w1, . . . , wd) ∈ Ωd. f (w) = wkKk(z, w), z, w ∈ Πd, holds, or equivalently, the following two identities hold: f (z)∗ ± f (w) = (zk ± wk)Kk(z, w), z, w ∈ Πd. dXk=1 dXk=1 (3) f belongs to the Bessmertnyı class Bd(U), i.e., there exist a Hilbert space X and a Bessmertnyı pencil V(w) = VH(w) =(cid:20)VH,11(w) VH,12(w) VH,21(w) VH,22(w)(cid:21) = wkVk, dXk=1 50 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI where Vk ∈ L(U ⊕ X ) are positive semidefinite operators, so that (6.5) holds and f (w) = fV(w) (with fV(w) as in (6.37)). (4) The double Cayley transform of f , S(ζ) = C(f ) :=hf(cid:16) 1 + ζ 1 − ζ(cid:17) − Iihf(cid:16) 1 + ζ 1 − ζ(cid:17) + Ii−1 , belongs to the Schur -- Agler class SA(Dd,L(U)) and has a unitary Givone -- Roesser realization (1.2) with a unitary and self-adjoint colligation matrix U = U∗ = U−1. (5) The Cayley transform of f over the variables, F (ζ) = f(cid:16) 1 + ζ 1 − ζ(cid:17), belongs to the Herglotz -- Agler class HA(Dd,L(U)) and has a representation (6.11) with R = 0, U = U−1 = U∗, and Ran V contained in the 1-eigenspace of U . (6) There exist a Hilbert space X , its subspaces X (0), X (1) with X = X (1)⊕X (0), a decomposition of IX , P (w) = w1P1 + ··· + wdPd = (cid:20)P11(w) P10(w) P01(w) P00(w)(cid:21) (with respect to the two-fold decomposition of X ), and V1 ∈ L(U ,X (1)), such that f has the form (6.36). Proof. The equivalence of statements (1), (2), (3), and (4) has been proved in [28]. (5)⇒(6) has been shown above, in the first paragraph of Special case 4. This is an application of the construction in the proof of Theorem 6.3 to this special case. (6)⇒(3) is obvious. (4)⇒(5). The function F is related to S = C(f ) as in (3.8). Let U =(cid:20)A B C D(cid:21) be a unitary colligation matrix providing the transfer-function realization for S as in (4), i.e., U∗ = U, together with the spectral decomposition of IX , P (ζ) = ζ1P1 + ··· + ζdPd. As it was shown in Remark 3.2, F has a representation (3.10) with skew-adjoint R, unitary U = U∗0 = A∗ − C∗(I − D∗)−1B∗ and V = 1√2 B. Now we also have that R = R∗ = −R∗, which means that R = 0, and that U = U∗ = A − B(I − D)−1B∗. It follows that the space X has an orthogonal decomposition X = X (1) ⊕ X (0), where X (1) is the 1-eigenspace of U and X (0) is the (-1)-eigenspace of U . With respect to this decomposition, let us write U0 =(cid:20) R − V ∗0 T V −(I + T )∗V0 (cid:21) , V ∗0 (I + T ) −T 0 U = U0 =(cid:20)I 0 −I(cid:21) , A =(cid:20)A11 A10 Since A00 is a self-adjoint and contractive, the identity −I = A00 + B0(I − D)−1B∗0 A01 A00(cid:21) , B =(cid:20)B1 B0(cid:21) . is possible only if B0 = 0. Thus im V ⊆ X (1), which completes the proof. Special case 5: the single-variable case. When we specialize Theorem 6.3 to the single-variable case, some simplifications occur. In particular the Herglotz pencil (6.23) involves only two operators, namely the flip of a Π-impedance-conservative system node which can be assumed to be the canonical form (6.24) (cid:3) SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 51 along with a single positive operator necessarily of the diagonal form U1 =(cid:20)V ∗1 V1 0 Then the pencil U(w) has the form U(w) =(cid:20) R − V ∗0 T V0 −(I + T )∗V0 V ∗0 (I + T ) −T 0 I(cid:21) . (cid:21) + w(cid:20)V ∗1 V1 0 0 I(cid:21) and the Bessmertnyı transfer-function realization for f ∈ H(Π,L(U)) becomes f (w) = R + wV ∗1 V1 + V ∗0 (−T + (I + T )(wI − T )−1(I + T )∗)V0. (6.38) For simplicity, let us now assume that U = C (so f is scalar-valued). Then R ∈ C is just a purely imaginary number and V ∗0 V0 is an operator on C and so can be identified with a nonnegative real number α (the image of the operator V ∗0 V0 acting on 1 ∈ C). From the representation (6.38), we see that there is no harm in cutting the state space X (0) (on which T is acting) down to the smallest subspace reducing for T which contains the range of the rank-1 operator V0, i.e., we may assume that V0 · 1 is a cyclic vector for T . Then the spectral theorem tells us that there is a measure ν on the imaginary line iR so that T is unitarily equivalent to T = M−ζ : f (ζ) 7→ −ζf (ζ) 1 1−ζ . As this function must be in L2(ν), we conclude that 1+ζ2 acting on L2(ν). Without loss of generality we take the cyclic vector V0 · 1 to be the function 1 1−ζ ∈ L2(ν), dν(ζ) is a finite measure. When this is done then we see that the i.e., that adjoint operator V ∗0 : L2(ν) 7→ C is given by 1 1 − ζ V ∗0 : f (ζ) 7→ZiR f (ζ) dν(ζ). 1 Then the Bessmertnyı realization (6.38) for f (w) collapses to the integral represen- tation formula f (w) = αw + R +ZiR(cid:20) ζ 1 + ζ2 + 1 ζ + w(cid:21) dν(ζ). This agrees with the classical Nevanlinna integral representation for holomorphic functions taking the right halfplane into itself. Actually the formula is usually stated for holomorphic functions taking the upper halfplane into itself (see [24, Theorem 1 page 20]); however the correspondence ef (ω) 7→ f (w) := −ief(iw) be- tween ef in the Nevanlinna class and f in the Herglotz class enables one to easily convert one integral representation to the other. We also point out that our proof (starting with the Herglotz representation (3.10) for the Herglotz function F (ζ) on the disk and then separating out the 1-eigenspace of the unitary operator U in that representation) is just an operator-theoretic analogue of the proof of the integral representation formula in [24], where one starts with the integral Herglotz representation F (ζ) = R +ZT t + ζ t − ζ dµ(t) and then separates out any point mass of µ at the point 1 on the circle. 52 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI Remark 6.6. In recent work [4], Agler -- Tully-Doyle -- Young obtain a realization formula for the most general scalar-valued Nevanlinna -- Agler function on the upper polyhalfplane. It is a straightforward matter to adjust the formulas to the right polyhalfplane setting which we have here and to extend the results to the operator- valued case. The result amounts to combining our formulas (6.12), (6.14) and (6.15), i.e., f (w) =R + V ∗(cid:20)I 0 0 −(I + T )(cid:21)(cid:18)P (w)(cid:20)0 0 I(cid:21)(cid:19)(cid:20)I 0 −T(cid:21) +(cid:20)0 0 I(cid:21) +(cid:20)I 0 −T(cid:21)(cid:19)−1 0 −(I − T )−1(cid:21) V. ·(cid:18)−P (w)(cid:20)I 0 0 0 0 · (6.39) 0 R However the associated Bessmertnyı pencil V(w) =" (−P (w)(cid:2) I 0 0 −T(cid:3) + [ 0 0 0 I ])h I 0 −T(cid:3)# V ∗(cid:2) I 0 −(I+T )(cid:3) 0 −(I−T )−1i V P (w) [ 0 0 0 I ] +(cid:2) I 0 0 0 lacks the symmetry properties of what we are calling a Herglotz -- Agler operator pencil (see Definition 6.1). Hence there is no easy analogue of the proof of (3) ⇒ (2) in Theorem 6.3 and it is not all transparent from the presentation (6.39) why the resulting function (6.39) has a Πd-Herglotz -- Agler decomposition (5.2); indeed, it takes several pages of calculations in [4] (see Propositions 3.4 and 3.5 there) to arrive at this result. There are other results in [4] and [5] using realization theory to characterize various types of boundary behavior of the function f at infinity; we do not go into this topic here. References [1] J. Agler, On the representation of certain holomorphic functions defined on a polydisc, In Topics in operator theory: Ernst D. Hellinger Memorial Volume, Oper. Theory Adv. Appl., Vol. 48, pp. 47 -- 66, Birkhauser, Basel, 1990. [2] J. Agler and J.E. McCarthy, Nevanlinna -- Pick interpolation on the bidisk, J. Reine Angew. Math. 506 (1999), 191 -- 204. [3] J. Agler, J.E. McCarthy, and N.J. Young, Operator monotone functions and Lowner func- tions of several variables, Annals of Mathematics 176 (2012), 1783 -- 1826. [4] J. Agler, R. Tully-Doyle, and N.J. Young, Nevanlinna representations in several variables, arXiv:1203.2261v2. [5] J. Agler, R. Tully-Doyle, and N.J. Young, Boundary behavior of analytic functions of two variables via generalized models, Indag. Math. (N.S.) 23 (2012) no. 4, 995 -- 1027. [6] Y. Arlinskii, S. Belyi, and E. Tsekanovskii, Conservative Realizations of Herglotz -- Nevanlinna Functions, Operator Theory: Advances and Applications 217, Birkhauser, 2011. [7] D.Z. Arov and M.A. Nudelman, Passive linear stationary dynamical scattering systems with continuous time, Integral Equations and Operator Theory 24 (1996), 1 -- 45. [8] T.A. Azizov and I..S. Iokhvidov, Linear Operators in Spaces with an Indefinite Metric, Wiley, 1980. [9] J. A. Ball. Multidimensional circuit synthesis and multivariable dilation theory, Multidi- mens. Syst. Signal Process 22 (2011) no. 1-3, 27 -- 44. [10] J.A. Ball and V. Bolotnikov, Canonical de Branges -- Rovnyak model transfer-function re- alization for multivariable Schur-class functions, in: Hilbert Spaces of Analytic Functions (Ed. J. Mashreghi, T. Ransford, and K. Seip), CRM Proceedings & Lecture Notes 51, Amer. Math. Soc., Providence, 2010. [11] J.A. Ball and D.S. Kaliuzhnyi-Verbovetskyi, Rational Cayley inner Herglotz -- Agler functions: positive-kernel decompositions and transfer-function realizations. Linear Algebra Appl. 456 (2014), 138 -- 156. SCHUR -- AGLER AND HERGLOTZ -- AGLER CLASSES 53 [12] J.A. Ball and O.J. Staffans, Conservative state-space realizations of dissipative system be- haviors, Integral Equations and Operator Theory 54 (2006), 151 -- 213. [13] J. A. Ball and T. Trent, Unitary colligations, reproducing kernel Hilbert spaces, and Nevanlinna-Pick interpolation in several variables, J. Funct. Anal. 157 (1998), 1 -- 61. [14] J. Behrndt, S. Hassi, and H. de Snoo, Functional models for Nevanlinna families, Opuscula Mathematica 28 (2008) no. 3, 233 -- 245. [15] J. Behrndt, S. Hassi, and H. de Snoo, Boundary relations, unitary colligations, and functional models, Complex Analysis and Operator Theory 3 (2009), 57 -- 98. [16] S. Belyi, S. Hassi, H. de Snoo, and E. Tsekanovskii, A general realization theorem for matrix- valued Herglotz -- Nevanlinna functions, Linear Algebra and its Applications 419 (2006) no. 2 -- 3, 331 -- 358. [17] M. F. Bessmertnyı. Functions of Several Variables in the Theory of Finite Linear Structures. Ph. D. Thesis, Kharkov University, Kharkov, 1982. (Russian). [18] M. F. Bessmertnyı. On realizations of rational matrix functions of several complex variables. In: Interpolation Theory, Systems Theory and Related Topics: The Harry Dym Anniversary Volume (D. Alpay, I. Gohberg, and V. Vinnikov, eds.), Oper. Theory Adv. Appl., Vol. 134, pp. 157 -- 185, Birkhauser-Verlag, Basel, 2002. [19] M. F. Bessmertnyı. On realizations of rational matrix functions of several complex vari- ables. II. Translated from the Russian by V. Katsnelson. In: Reproducing Kernel Spaces and Applications, Oper. Theory Adv. Appl., Vol. 143, pp. 135 -- 146, Birkhauser, Basel, 2003. [20] M. F. Bessmertnyı. On realizations of rational matrix functions of several variables. III. In: Current Trends in Operator Theory and Its Applications, Oper. Theory Adv. Appl., Vol. 149, pp. 133 -- 138, Birkhauser, Basel, 2004. [21] M. F. Bessmertnyı. Functions of several variables in the theory of finite linear structures. I. Analysis. In: Operator Theory, Systems Theory and Scattering Theory: Multidimensional Generalizations, Oper. Theory Adv. Appl. 157, pp. 91 -- 106, Birkhauser, Basel, 2005. [22] J. Bogn´ar, Indefinite Inner Product Spaces, Springer-Verlag, New York-Heidelberg-Berlin, 1974. [23] D. Bors and M. Majewski, On the existence of an optimal solution of the Mayer prob- lem governed by 2D continuous counterpart of the Fornasini-Marchesini model, Multidi- mens. Syst. Signal Process 24 (2013) no. 4, 657 -- 665. [24] W.F. Donoghue, Monotone Matrix Functions and Analytic Continuation, Die Grundlehren der mathematischen Wissenshaften in Einzeldarstellungen Band 207, Springer, New York, 1974. [25] E. Fornasini and G. Marchesini, Doubly-indexed dynamical systems: state-space models and structural properties, Math. Systems Theory 12 (1978(, 59 -- 72. [26] P.A. Fuhrmann, Linear Operators and Systems in Hilbert Space, McGraw-Hill, New York, New York, 1981. [27] J.W. Helton, Systems with infinite-dimensional state space: the Hilbert space approach, Re- cent Trends in System Theory, Proc. IEEE 64 (1976) no. 1, 145 -- 160. [28] D. S. Kalyuzhnyı-Verbovetzkiı. On the Bessmertnyı class of homogeneous positive holomor- phic functions of several variables. In: Current Trends in Operator Theory and Its Applica- tions, Oper. Theory Adv. Appl. 149, pp. 255-289, Birkhauser, Basel, 2004. [29] J. Malinen, O.J. Staffans and G. Weiss, When is a linear system conservative? Quart. Appl. Math. 64 (2006) no. 1, 61 -- 91. [30] V. Paulsen, Completely Bounded Maps and Operator Algebras, Cambridge Studies in Ad- vanced Mathematics 78, Cambridge University Press, 2002. [31] R.S. Phillips, Dissipative operators and hyperbolic systems of partial differential equations, Trans. Amer. Math. Soc. 90 (1959), 193 -- 254. [32] D. Salamon, Infinite dimensional linear systems with unbounded control and observation: a functional analytic approach, Trans. Amer. Math. Soc. 300 (1987), pp. 383 -- 431. [33] Y.L. Smuljan, Invariant subspaces of semigroups and the Lax-Phillips scheme, deposited in VINITI, N 8009-1386, Odessa, 49 pp., 1986. [34] O.J. Staffans, Passive and conservative continuous-time impedance and scatterings systems. Part I: well-posed systems, Math. Control Signals Systems 15 (2002), 291 -- 315. 54 J.A. BALL AND D.S. KALIUZHNYI-VERBOVETSKYI [35] O.J. Staffans, Passive and conservative infinite-dimensional impedance and scattering sys- tems (from a personal point of view), in Mathematical Systems in Biology, Communica- tion, Computation and Finance (MTNS2002 Notre Dame, Indiana) (ed. J. Rosenthal and D.S. Gilliam) pp. 373 -- 414, IMA Volume 314 Springer, 2003. [36] O.J. Staffans, Well-posed Linear Systems, Encyclopedia of Mathematics and Its Applications 103, Cambridge University Press, 2005. [37] O.J. Staffans, On scattering passive system nodes and maximal scattering dissipative opera- tors, Proc. Amer. Math. Soc. 141 no. 4 (2013), 1377 -- 1383. [38] O.J. Staffans and G. Weiss, A physically motivated class of scattering passive linear systems, SIAM J. Control Optim. 50 No. 5 (2012), 3083 -- 3112. [39] G. Weiss, Transfer functions of regular linear systems. Part I: characterizations of regularity, Trans. Amer. Math. Soc. 3342 (1994), pp. 827 -- 854. [40] G. Weiss and O.J. Staffans, Maxwell's equations as a scattering passive linear system, SIAM J. Control Optim. 51 (2013), 3722 -- 3756. Department of Mathematics, Virginia Tech, Blacksburg, VA, 24061 E-mail address: [email protected] Department of Mathematics, Drexel University, 3141 Chestnut St., Philadelphia, PA, 19104 E-mail address: [email protected]
1909.06634
1
1909
2019-09-14T17:22:42
Decay estimates connected with Toeplitzness of composition operators
[ "math.FA" ]
In a higher dimensional version of an earlier conjecture of Nazarov and Shapiro, the truth of which would imply that any composition operator on the second Hardy space is weakly asymptotically Toeplitz, Shayya proved that the arithmetic means of the main diagonal entries of the matrix of coefficients connected with the conjecture, decay like 1/log N. But the proof does not extend to other diagonal entries. We prove that the arithmetic means of the entries along each diagonal decay like 1/(log N loglog N) uniformly.
math.FA
math
DECAY ESTIMATES CONNECTED WITH TOEPLITZNESS OF COMPOSITION OPERATORS Faruk F. Abi-Khuzam Abstract. Let f ∈ L∞(T d) with kf kL∞ (T d) ≤ 1, ν ∈ Zd, n, k ∈ Z and put bn,n−k = RE f (x)ne−2πi(n−k)ν ·xdx, E = {x ∈ T d : f (x) = 1}. Shayya conjectured that, if f (ξ) = 0 for all ξ in a half-space S of lattice points, and ν ∈ −S, and f (0) 6= 0, then limn→∞ bn,n−k = 0, k ∈ Z. This is a higher dimensional version of an earlier conjecture of Nazarov and Shapiro, the truth of which would imply that any composition operator is weakly asymptotically Toeplitz on the Hardy space H 2. Shayya proved that the arithmetic means of the main diagonal entries, {bn,n}, decay like {log N }−1, but the proof does not extend to other diagonal entries. Using an elementary method, we extend and improve all known results. In particular, we show that the arithmetic means of the entries of each diagonal, {bn,n−k }, decay like {log N log log N }−1 uniformly in k ∈ Z. 1. INTRODUCTION Let φ be a function holomorphic in the open unit disk U of the complex plane and satisfying the two conditions φ(0) 6= 0, φ(z) < 1, z ∈ U. The composition operator Cφ : H(U ) → H(U ), defined via φ by Cφf = f ◦ φ, restricts to a bounded linear operator on the Hardy space H 2(U ), and so has an infinite matrix (cα.β) associated to it and defined by φ(z)αz−β dz 2πiz . cα,β =Z∂U If S is the forward shift operator on H 2(U ), and S∗ is its adjoint, then Cφ is said to be weakly asymptotically Toeplitz , or WAT, if the sequence {S∗nCφSn} of bounded linear operators on H 2(U ) converges in the weak operator topology. Equivalently, Cφ is WAT, if If E = {z ∈ ∂U : φ(z) = 1}, and we put lim n→∞ cn,n−k = 0, k ∈ Z. bn,n−k =ZE φ(z)nzk−n dz 2πiz , 2000 Mathematics Subject Classification. Primary 42B05; Secondary 47B33. Key words and phrases. Toeplitz operator, weakly asymptotically Toeplitz. 1 2 FARUK F. ABI-KHUZAM then, clearly, cn,n−k → 0, if and only if bn,n−k → 0, as n → ∞.The WAT conjecture on the circle, formulated by Nazarov and Shapiro [4], is the statement that (1.1) lim n→∞ bn,n−k = 0, k ∈ Z. At present, it is not known if (1.1) is true, even when k = 0. But the arithmetic means are more tame. In [5], it was shown that (1.2) lim N→∞ 1 N bn,n−k2 = 0, k ∈ Z. N Xn=1 An infinite series carrying the full sequence {cn,n−k} was considered in [1], where it was shown, using complex analytic methods, that cn,n−k2 n < ∞, k ≥ 0, ∞ Xn=1 from which it follows, through a simple summation by parts argument, that the arithmetic means of the sequence {cn,n−k2} tend to 0, and also that ∞ Xn=1 1 n(n + 1) n Xj=1 cj,j−k2 < ∞, k ≥ 0. j=1 cj,j−k2 = 0, The convergence of this later series implies that lim inf n→∞ and this anticipates later work on the arithmetic means of the sequence {bn,n−k2}. was established as part of a study of these prob- lems undertaken recently by Shayya [6] in the setting of functions defined on the d-dimensional torus T d. He proved the existence of a constant C, essentially de- pendent on φ(0), such that The convergence ofP∞ n Pn bn,n−k2 log n n=1 n (1.3) 1 N bn,n2 ≤ C log N π , M ≥ 1, N ≥ 2, M+N Xn=M thereby obtaining a quantitative strengthening of (1.2) , in the case k = 0. Shayya's method uses a positive Borel measure µ such that bn,n2 =Z 1/2 −1/2 e−2πinθdµ(θ), bn,n2 n n=1 . This measure, obtained as Shayya's result (1.3) tells us that the lim supn→∞ an application of Herglotz's theorem, is only available when k = 0. along with the convergence of the series P∞ n Pn j=1 bn,n2 < ∞, while, j=1 bn,n2 = 0. So if the limit exists, by the argument above, lim inf n→∞ it would have to be 0, and a natural question would then be to find the rate of decay to 0. In the present work we consider the more general problem of the arithmetic means of bn,n−k2, k ∈ Z. We obtain, uniformly in k ∈ Z, a quantitative strengthening of (1.2). In addition, we show that these means actually decay much faster than suggested by (1.3) and we obtain, among other results, the inequality n Pn log n log n (1.4) 1 N + 3 bn,n−k2 ≤ C/(1 − 1/ log 3) log(N + 3) log[ log(N +4) log 4 ] , k ∈ Z, for all M ≥ 1, and N ≥ 1, where the constant C depends only on φ(0). M+N Xn=M TOEPLITZNESS OF COMPOSITION OPERATORS 3 2. PRELIMINARIES AND STATEMENT OF RESULTS We shall consider the problem of the decay of the arithmetic means in the setting of functions defined on T d, as formulated in [6]. Since we shall be working on the coefficients bn,n−k2 with k ∈ Z, we will not have available a measure as in [6], and so we shall proceed along different, rather elementary, lines. However, there will be two important common ingredients between the present paper and [6], namely the use of Helson's and Lowdenslager's generalization of Szego's theorem [3], and the restriction to the set where the boundary values of the given function have unit modulus. We shall obtain a bound on the sum of the series ∞ Xm=−∞ m6=N bm,m−k2 m − N , where k, N ∈ Z. The usefulness of this stems from the fact that the bound will be independent of k and N , thereby making it possible to obtain uniform bounds on the arithmetic means of {bm,m−k2}. We now give a summary, taken mostly from [6], of material needed in connection with the Helson-Lowdenslager result. Definition 1. A set S ⊂ Zd is said to be a half-space of lattice points in Rd if (i) 0 6∈ S (ii) ξ 6= 0 implies ξ ∈ S or −ξ ∈ S (iii) ξ, ξ′ ∈ S implies ξ + ξ′ ∈ S. Clearly, if S is a half-space of lattice points in Rd, then so is its reflection in the origin. i.e. −S = {−ξ : ξ ∈ S}. Also [6], if S1 and S2 are half-spaces of lattice points in Rd1 and Rd2 respectively, then S1 × Rd2 ∪ Rd1 × S2 contains a half-space of lattice points in Rd1 × Rd2 (e.g., S1 × Rd2 ∪ {0} × S2). We shall need the following generalization of Szego's theorem [3], stated as corollary A in [6]. Lemma 1. Suppose f ∈ L2(T d), S is a half-space of lattice points in Rd, and f (ξ) = 0 for all ξ ∈ S. then log f (x)dx ≥ log f (0). ZT d As a simple application of this lemma, suppose S is a half-space of lattice points in Rd, f ∈ L∞(T d), kf kL∞(T d) ≤ 1, and f (ξ) = 0 for all ξ ∈ S. For 0 < r < 1, and ν ∈ −S, define F : T d × T d → C by F (x, y) = e2πi(ν,−ν)·(x,y) − rf (x)f (y). Then, [6], F (ξ, η) = 0 for all (ξ, η) ∈ S × Rd ∪ Rd × (−S), and so for all (ξ, η) in a half-space of lattice points in Rd × Rd. Lemma 1 applies and we conclude that ZT dZT d log F (x, y)dxdy ≥ log F (0, 0) = log r f (0)2. 4 FARUK F. ABI-KHUZAM Also, using the equality log u = 2 log+ u − log u, we have, for any measur- able subset of T d, (2.1) log F (x, y)dxdy ZEZE = 2ZT dZT d log F (x, y)dxdy ≤ZT dZT d log+ F (x, y)dxdy −ZT dZT d ≤ 2 log(1 + r) − log r f (0)2 ≤ log log F (x, y)dxdy 4 r f (0)2 . We shall also need two sequences of functions logj x and Lq(x). The first is the iterated logarithmic function, and the second is a product of different iterates: If j ≥ 1 is an integer, we define the function logj as follows: log1 x = log x, logj x = log ◦ logj−1 x. We will need to use the iterated logarithm on the domain where it is well- defined and positive, and this, of course, is trivial to determine.In terms of the iterated logarithm we define the function Lq by q Lq(x) = logj x, q ≥ 1. Yj=1 Also, for a positive differentiable function g, we define [2] the function a(.; g) by a(x; g) = xg′(x)/g(x), and note that a(x; L1) = log′ q+1(x) = , a(x; Lq) = 1 log x 1 . xLq(x) 1 log x 1 + q Xj=2 1 logj x  , q ≥ 2. Clearly, a(x; Lq) decreases monotonically to 0 as x → ∞, and so there exist positive constants αq and γq , depending only on the integer q, such that 0 < αq < 1, and logq+1 x > 0, 0 < a(x; Lq) < αq, x ≥ γq. Theorem 1. Suppose f ∈ L∞(T d) with kf kL∞(T d) ≤ 1, and ν ∈ Zd. For n, k ∈ Z, define bn,n−k =ZE f (x)ne−2πi(n−k)ν·xdx, where E = {x ∈ T d : f (x) = 1}.When f (0) 6= 0, put C = log 16 f (0)4 . (i) If f (ξ) = 0 for all ξ in a half-space S of lattice points, ν ∈ −S, f (0) 6= 0, and N ∈ Z, then (2.2) ∞ Xm=−∞ m6=N bm,m−k2 m − N ≤ log 16 f (0)4 , k ∈ Z. (ii) If f (ξ) = 0 for all ξ in a half-space S of lattice points, ν ∈ −S, f (0) 6= 0, M ≥ 1, and p ≥ 1, then (2.3) 1 p + 1 M+p Xm=M bm,m−k2 ≤ C log(p + 1) , k ∈ Z. TOEPLITZNESS OF COMPOSITION OPERATORS 5 (iii) Suppose g is positive, and 0 < a(x; g) < α, for all x ≥ γ ≥ 1. If f (ξ) = 0 for all ξ in a half-space S of lattice points, ν ∈ −S, f (0) 6= 0, M ≥ 1, and p ≥ 1, then (2.4) (cid:18)Z p+1 1 dt tg(t + γ)(cid:19) M+p Xm=M bm,m−k2 ≤ C 1 − α · p + γ g(p + γ) , k ∈ Z. (iv) Suppose Lq, αq, and γq are as defined above. If f (ξ) = 0 for all ξ in a half-space S of lattice points, ν ∈ −S, f (0) 6= 0, M ≥ 1, and p ≥ 1, then (2.5) bm,m−k2 < C/(1 − αq) Lq(p + γq)[logq+1(p + 1 + γq) − logq+1(1 + γq)] , k ∈ Z. 1 p + γq M+p Xm=M Also, for a fixed q ≥ 1, there is a constant C′ depending only on q and C, such that, for all large p, 1 p M+p Xm=M uniformly in k ∈ Z. bm,m−k2 < C′ Lq+1(p) = C′ log p · log log p · · · logq+1 p , The inequality in part (ii) represents a sharpening, and an extension of (1.3). The inequality in part (iv) shows that, uniformly in k ∈ Z, the decay of the arith- metic means of the sequence {bm,m−k2} is faster than that of L−1 q+1 , for any q ≥ 1. 3. PROOF OF THEOREM 1 The definition of bn,n−k gives us that bn.n−k2 =ZEZE(cid:16)f (x)f (y)e−2πiν·(x−y)(cid:17)n and, for N, k ∈ Z, we have, since f (x) = 1 for x ∈ E, .e2πikν·(x−y)dxdy, bn+N,n+N −k2 + b−n+N,−n+N −k2 = 2ZEZE(cid:16)f (x)f (y)e−2πiν·(x−y)(cid:17)N .e2πikν·(x−y) Re(cid:16)f (x)f (y)e−2πiν·(x−y)(cid:17)n Multiplying by rn, 0 < r < 1, dividing by n and summing over n ≥ 1, we obtain dxdy. bn+N,n+N −k2 + b−n+N,−n+N −k2 n rn ∞ Xn=1 = 2ZEZE(cid:16)f (x)f (y)e−2πiν·(x−y)(cid:17)N where F (x, y) = e2πiν·(x−y) − rf (x)f (y) . integral of log F (x, y) we get that .e2πikν·(x−y). log 1 F (x, y) dxdy, In view of the estimate (2.1) on the bn+N,n+N −k2 + b−n+N,−n+N −k2 n rn ≤ log 16 r2 f (0)4 , ∞ Xn=1 and letting r → 1−, we obtain bn+N,n+N −k2 + b−n+N,−n+N −k2 n ≤ C. ∞ Xn=1 6 FARUK F. ABI-KHUZAM Reindexing we arrive at ∞ Xm6=N m=−∞ bm,m−k2 m − N ≤ C, k ∈ Z, and this completes the proof of part (i). If M ≥ 1, p ≥ 1, then, for a given N , the partial sums in (2.2) for M ≤ m ≤ M + p, m 6= N , are also bounded by the constant C uniformly in k, N ∈ Z, and we may sum them over N from M to M + p to get, after an interchange in the order of summation (3.1) M+p M+p Xm=M XN =M N 6=m bm,m−k2 m − N ≤ M+p XN =M C = C(p + 1), k ∈ Z. N =M N 6=m On the left-hand side of (3.1) we use the inequality PM+p m − N  bm,m−k 2!   j  ≥ log(p + 1) M+p Xm=M ≥ M+p Xm=M XN =M Xj=1 N 6=m M+p so that 1 1 p M+p Xm=M bm,m−k 2!  Xm=M log(p + 1) M+p and the inequality in (2.3) follows. 1 m−N ≥Pp j=1 1 j : bm,m−k 2! , bm,m−k 2 ≤ C(p + 1), k ∈ Z, For proof of part (iii), first divide both sides of (2.2) by g(N − M + γ), N ≥ M ≥ 1,and proceed exactly as in the proof of part (ii). In the resulting inequality the left-hand side is M+p ( Xm=M bm,m−k 2) M+p XN =M N 6=m 1 m − N g(N − M + γ) M+p ≥ ( Xm=M bm,m−k 2) 1 jg(j + γ) p Xj=1 ≥ M+p Xm=M while the right-hand side is bm,m−k 2Z p+1 1 dt tg(t + γ) , C M+p XN =M 1 g(N − M + γ) < C{ 1 g(γ) +Z p 0 dt g(t + γ) }. The conditions on a(x; g), along with the positivity of g, imply that both functions g and x/g(x) are increasing on [γ, ∞). An application of the Cauchy mean-value theorem gives a ξ ∈ (γ, x) such that dt g(t) γ R x x g(x) − γ g(γ) = 1 1 − a(ξ; g) < 1 1 − α . TOEPLITZNESS OF COMPOSITION OPERATORS 7 Since γ ≥ 1, 0 < α < 1,we have −(cid:18) 1 1 − α(cid:19) γ g(γ) + 1 g(γ) < 0, and it follows that (3.2) 1 g(γ) +Z p 0 dt g(t + γ) < 1 1 − α · p + γ g(p + γ) , which gives the required bound and the inequality in (2.4) follows. If q ≥ 1, and Lq is the function defined in section 2, and αq and γq are the positive constants associated with it, we may apply the result in part (iii) with g(x) = Lq(x). This gives us (cid:18)Z p+1 1 dt tLq(t + γq)(cid:19) M+p Xm=M bm,m−k2 ≤ C 1 − αq · p + γq Lq(p + γq) , k ∈ Z. For the integral on the left-hand side we have Z p+1 1 dt tLq(t + γq) ≥Z p+1 1 dt (t + γq)Lq(t + γq) = logq+1(p + 1 + γq) − logq+1(1 + γq), and it follows that 1 p + γq M+p Xm=M for all k ∈ Z. bm,m−k2 ≤ C/(1 − αq) Lq(p + γq)[logq+1(p + 1 + γq) − logq+1(1 + γq)] , If q ≥ 1 is fixed, since Lq(p + γq)[logq+1(p + 1 + γq) − logq+1(1 + γq)] ∼ Lq+1(p) as p → ∞, there is a constant C′depending only on q and C, such that, for all large p, 1 p M+p Xm=M bm,m−k2 < C′ Lq+1(p) , uniformly in k ∈ Z. In the special case where q = 1, Lq(x) = log x and we may take γ = 3, and log 3 . Then Theorem 1 gives us that α = 1 1 p + 3 M+p Xm=M bm,m−k2 ≤ C/(1 − 1/ log 3) log(p + 3)[log log(p + 4) − log log 4] , k ∈ Z. There are some interesting questions that present themselves in connection with the study of the problems considered in this note. One question concerns the rate . Even in the case k = 0, this of decay of the tail end of the series P∞ is not known. Another question concerns the series P∞ n known if this series is convergent, even in the case q = 1, and k = 0. n=1 Lq(n) bn,n−k2 . It is not bn,n−k2 n=1 n 8 FARUK F. ABI-KHUZAM 4. REFERENCES [1] F. ABI KHUZAM AND B. SHAYYA, A remark on the WAT conjecture, preprint, 2009. [2]W. HAYMAN, A generalisation of Stirling's formula, J. Reine. Angew. Math. 196 (1956), 67-95. [3]H. HELSON AND D. LOWDENSLAGER, Prediction theory and Fourier series in several variables, Acta Math. 99 (1958), 165-202. [4] F. NAZAROV AND J. SHAPIRO, On the Toeplitzness of composition op- erators, Complex Var. Elliptic Wqu. 52 (2007), 193-210. [5] J. SHAPIRO, Every composition operator is ( mean) asymptotically Toeplitz, J. Math. Anal. Appl. 333 (2007), 523-529. [6] B.SHAYYA, The WAT conjecture on the Torus, Proc. Amer. Math. Soc. 139 ( 2011), 3633-3643. Department of Mathematics, American University of Beirut, Beirut Lebanon E-mail address: [email protected]
1711.00091
3
1711
2019-10-30T08:01:27
The invertibility of U-fusion cross Gram matrices of operators
[ "math.FA" ]
For applications like the numerical solution of physical equations a discretization scheme for operators is necessary. Recently frames have been used for such an operator representation. In this paper, we apply fusion frames for this task. We interpret the operator representation using fusion frames as a generalization of fusion Gram matrices. We present the basic definition of $U$-fusion cross Gram matrices of operators for a bounded operator $U$. We give sufficient conditions for their (pseudo-)invertibility and present explicit formulas for the inverse. In particular, we characterize fusion Riesz bases and fusion orthonormal bases by such matrices. Finally, we look at which perturbations of fusion Bessel sequences preserve the invertibility of the fusion Gram matrix of operators.
math.FA
math
The invertibility of U -fusion cross Gram matrices of operators Mitra Shamsabadia, Ali Akbar Arefijamaala, Peter Balazsb,∗ aDepartment of Mathematics and Computer Sciences, Hakim Sabzevari University, Sabzevar, Iran bAcoustics Research Institute, Austrian Academy of Sciences, Vienna, Austria Abstract For applications like the numerical solution of physical equations a discretization scheme for operators is necessary. Recently frames have been used for such an operator representation. In this paper, we apply fusion frames for this task. We interpret the operator representation using fusion frames as a generalization of fusion Gram matrices. We present the basic definition of U -fusion cross Gram matrices of operators for a bounded operator U . We give sufficient conditions for their (pseudo-)invertibility and present explicit formulas for the inverse. In particular, we characterize fusion Riesz bases and fusion orthonormal bases by such matrices. Finally, we look at which perturbations of fusion Bessel sequences preserve the invertibility of the fusion Gram matrix of operators. 1. Introduction and motivation For the representation (and modification) of functions a standard approach is using orthonormal bases (ONBs). It can be hard to find a 'good' orthonormal basis, in the sense that it sometimes cannot fulfill given properties, as formally expressed e.g. in the Balian-Low theorem [34]. For solving this problem, frames were introduced by Duffin and Schaeffer [28] and widely developed by many authors [17, 22, 26, 29]. In recent years, frames have been the focus of active research, both in theory [2, 18, 30] and applications [11, 19, 24]. Also, several generalizations have been investigated, e.g. [1, 3, 4, 44, 47], among them fusion frames [20, 21, 32], which are the topic of this paper. For a numerical treatment of operator equations, used for example for solving integral equations in acoustics [38], the involved operators have to be discretized to be handled numerically. The (Petrov- )Galerkin approach [31] is a particular and well-known way for this discretization. For an operator O, the matrix M defined by Mk,l = hOψl, φki is called the matrix corresponding to the operator O, or the system matrix. The standard way for the discretization of operators is using bases [33], but recently the general theory for frames has been developed [8, 10]. Frames were also used in numerics [35], in particular in an adaptive approach [25, 45]. In [12, 13] sufficient and necessary conditions of the invertibility of such matrices is investigated. Note that the system matrix of the identity is the cross Gram matrix of the ∗Corresponding author Email addresses: [email protected] (Mitra Shamsabadi), [email protected] (Peter Balazs) Preprint submitted to Elsevier October 31, 2019 two sequences {ψk}k∈I and {φk}k∈I . Therefore, in [15] the concept of matrix representation of operators using frames is reinterpreted as a generalization of the Gram matrix to investigate the inverses. As the concept of domain decomposition is a particularly relevant topic in this field, the extension of the approach to operator representations to fusion frame is very useful [40]. In this paper, we therefore look at those U -fusion cross Gram matrices. In particular, we investi- gate the (pseudo-)invertibility of U -fusion Gram matrices of operators. In Section 2, we review basic notations and preliminaries. In Section 3, we give necessary as well as sufficient conditions for the (pseudo-)invertibility of the U -cross Gram matrices, characterize fusion orthonormal bases and fusion Riesz bases by those properties and give formulas for the (pseudo-)inverses. Finally, in Section 4, some stability results are discussed. 2. Preliminaries and Notations Throughout this paper, H is a separable Hilbert space, I a countable index set and IH the identity operator on H and {ei}i∈I an orthonormal basis for H. The orthogonal projection on a subspace V ⊆ H is denoted by πV . We will denote the set of all linear and bounded operators between Hilbert spaces H1 and H2 by B(H1,H2) and for H1 = H2 = H, it is represented by B(H). We denote the range and the null space of an operator U by ran (U ) and ker (U ), respectively. For a closed range operator U ∈ B(H1,H2), the pseudo-inverse of U is defined the unique operator U† ∈ B(H2,H1) satisfying ker(cid:0)U†(cid:1) = ran (U )⊥ , ran(cid:0)U†(cid:1) = ker (U )⊥ , and U U†U = U. The operator U has closed range if and only if U∗ has closed range and (U∗)† = (cid:0)U†(cid:1)∗, see e.g. Lemma 2.4.1, Lemma 2.5.2]. [22, For 0 < p < ∞, the space of all compact operators (which are the closure of finite-rank operators and denoted by K(H)) T on H such that its singular values {λn}n∈I belonging to ℓp is called the Schatten p-class of H. It is denoted by Sp(H) which is a Banach space with the norm kTkp = Xn p λnp! 1 . (2.1) An operator T ∈ B(H) is called trace class if trace(T ) := Pi∈I hT ei, eii < ∞, for every orthonormal 2 =Pi∈I kT eik2 < ∞. It basis {ei}i∈I for H. It is shown that T is trace class if and only if T ∈ S1(H). Also, the class of Hilbert- Schmidt operators of H is denoted by S2(H), and T ∈ S2(H) if and only if kTk2 is well-known that K(H) and Sp(H) are two sided ∗-ideal of B(H), that is, a Banach algebra under the norm (2.1) and the finite rank operators are dense in (Sp(H),k.kp). Moreover, for T ∈ Sp(H), one has kTkp = kT ∗kp and kTk ≤ kTkp. If S1 ∈ B(H,H1) and S2 ∈ B(H2,H), then kS1Tkp ≤ kS1kkTkp and kT S2kp ≤ kS2kkTkp. For more information about these operators, see [33, 41, 42, 48]. 2 2.1. Fusion frames We now review some definitions and primary results of fusion frames. For more information see [20, 21, 32]. For each sequence {Wi}i∈I of closed subspaces in H, the space Xi∈IM Wi!ℓ2 =({fi}i∈I : fi ∈ Wi,Xi∈I (cid:28){fi}i∈I ,{gi}i∈I(cid:29) =Xi∈I hfi, gii, kfik2 < ∞) , with the inner product is a Hilbert space. We now give the central definition of fusion frames: Definition 2.1. Let {Wi}i∈I be a family of closed subspaces of H and {ωi}i∈I be a family of weights, i.e. ωi > 0, i ∈ I. The sequence {(Wi, ωi)}i∈I is called a fusion frame for H if there exist constants 0 < AW ≤ BW < ∞ such that AWkfk2 ≤Xi∈I i kπWi fk2 ≤ BWkfk2, ω2 (f ∈ H). The constants AW and BW are called fusion frame bounds. If we have the upper bound, we call {(Wi, ωi)}i∈I a Bessel fusion sequence. A fusion frame is called tight, if AW and BW can be chosen to be equal, and Parseval if AW = BW = 1. If ωi = ω for all i ∈ I, the collection {(Wi, ωi)}i∈I is called it is called a Riesz decomposition of H if for every f ∈ H there is a unique choice of fi ∈ Wi such that ω-uniform. A fusion frame {(Wi, ωi)}i∈I is said to be a fusion orthonormal basis if H =Li∈I Wi and f =Pi∈I fi. A family of subspaces is called complete if span (Wi) = H. It is clear that every fusion orthonormal basis is a Riesz decomposition of H. Moreover, a family {Wi}i∈I of closed subspaces of H is a fusion orthonormal basis if and only if it is a 1-uniform Parseval fusion frame [20]. Furthermore, the synthesis operator TW : (Pi∈IL Wi)ℓ2 → H for a Bessel fusion sequence W = {(Wi, ωi)}i∈I is defined by TW ({fi}i∈I ) =Xi∈I ωifi. The adjoint operator T ∗W : H → (Pi∈IL Wi)ℓ2 which is called the analysis operator is given by T ∗W f = {ωiπWi f}i∈I , (f ∈ H). Both are bounded by √BW . If W = {(Wi, ωi)}i∈I is a fusion frame, the fusion frame operator SW : H → H, which is defined by i πWi f , is bounded (with bound BW ), invertible and positive [20, 32]. SW f = TW T ∗W f =Pi∈I ω2 3 Every Bessel fusion sequence V = {(Vi, υi)}i∈I is called a Gavrut¸a-dual of W = {(Wi, ωi)}i∈I , if f =Xi∈I ωiυiπVi S−1 W πWi f, (f ∈ H), TV φV W T ∗W = IH, for more details see [32]. From here on, for simplicity we say dual instead of Gavrut¸a-dual. The sequence W Wi, ωi(cid:1)(cid:9)i∈I is also a fusion frame for H and a dual of W, called the canonical dual of W [20, 32]. Rephrasing that, a Bessel fusion sequence V = {(Vi, υi)}i∈I is a dual of a fusion frame W = {(Wi, ωi)}i∈I if and only if of subspaces fW :=(cid:8)(cid:0)S−1 where the bounded operator φV W : (Pi∈IL Wi)ℓ2 → (Pi∈IL Vi)ℓ2 is given by and kφV Wk ≤(cid:13)(cid:13)S−1 W (cid:13)(cid:13). Also, V is called a pseudo-dual of W if TV φV W T ∗W is an invertble operator, see [23] for discrete case. Another approach to duality [36, 37] uses any bounded operator O : (Pi∈IL Wi)ℓ2 → (Pi∈IL Vi)ℓ2 . Starting with two fusion frames the duality is defined analogously to (2.2), i.e. TV O T ∗W = IH. We stick to the Gavrut¸a duals, but all results herein can be adapted to this other definition of duality. Let {Wi}i∈I be a family of closed subspaces of H and {ωi}i∈I a family of weights. We say that {(Wi, ωi)}i∈I is a fusion Riesz basis for H if span (Wi) = H and there exist constants 0 < C ≤ D < ∞ such that for each finite subset J ⊆ I and all (fj ∈ Wj, j ∈ J) we have φV W ({fi}i∈I ) = {πVi S−1 W fi}i∈I (2.2) (2.3) Remark 2.2. Note that weights are not included in the definition of the Riesz decomposition. Obviously, we have that f =Pi∈I fi with unique fis, if and only if f =Pi∈I wifi with the same uniqueness. The next theorem explores fusion Riesz bases with respect to local frames and their operators. Theorem 2.3. [20] Let {Wi}i∈I be a fusion frame for H and {eij}j∈Ji be an orthonormal basis for Wi, for each i ∈ I. Then the following conditions are equivalent: (1) {Wi}i∈I is a Riesz decomposition of H. (2) The synthesis operator TW is one to one. (3) The analysis operator T ∗W is onto. (4) {Wi}i∈I is a fusion Riesz basis for H. (5) {eij}i∈I,j∈Ji is a Riesz basis for H. For some results we need a version of Theorem 2.3 formulated for general family of subspaces: Proposition 2.4. Let W = {(Wi, ωi)}i∈I be a family of closed subspaces and {fij}j∈Ji be a Riesz basis for Wi, for each i ∈ I with bounds Ai and Bi, respectively, such that 0 < inf i∈I Then the following conditions are equivalent: Ai ≤ sup i∈I Bi < ∞. 4 CXj∈J 2 kfjk2 ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xj∈J ωjfj(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≤ DXj∈J kfjk2. (2.4) (1) W is a Riesz decomposition of H. (2) The synthesis operator TW is bounded and bijective. (3) The analysis operator T ∗W is bounded and bijective. (4) W is a fusion Riesz basis for H. (5) {ωifij}i∈I,j∈Ji is a Riesz basis for H. Proof. (1) ⇔ (2) ⇔ (3) by Theorem 2.3 for any family of closed subspaces (as those conditions imply a fusion frame property). (4) ⇔ (5) by relating the inequality (2.4) for fusion Riesz basis W to the one for Riesz bases {wifij}i∈I,j∈Ji see e.g. [22, Theorem 3.3.7]. (4) ⇔ (2) The equation (2.4) is equivalent to TW being injective, having closed range and being bounded. By the completeness we have (4) ⇒ (2). By the surjectivity of TW we have completeness and so (2) ⇒ (4). The following characterizations of fusion Riesz bases will be used frequently in this note, which is a generalization of [43, Theorem 2.2] to non-uniform fusion frames. Proposition 2.5. Let W = {(Wi, ωi)}i∈I be a fusion frame in H. Then the following are equivalent: (1) W is a fusion Riesz basis. (2) S−1 (3) ω2 W Wi ⊥ Wj for all i, j ∈ I, i 6= j. i πWi S−1 W πWj = δijπWj for all i, j ∈ I. Proof. (1) ⇒ (2) was proved in [20, Proposition 4.3]. (2) ⇒ (1) Suppose that {eij}j∈Ji is an orthonormal basis for Wi, for all i ∈ I. Then for any f ∈ H we have Xi∈I Xj∈Ji hf, ωieiji2 =Xi∈I i kπWi fk2 . ω2 It easily follows that {eij}i∈I,j∈Ji is a weighted frame [9] (with weights ωi > 0) for H with the frame operator SW . Moreover, by Proposition 2.4 it is enough to show that {ωieij}i∈I,j∈Ji is a Riesz basis for H or equivalently, that the sequences {ωieij}i∈I,j∈Ji and {S−1 W ωieij}i∈I,j∈Ji are biorthogonal. This immediately follows from the reconstruction formula f =Xi∈I Xj∈Ji hf, S−1 W ωieijiωieij, (f ∈ H). (2) ⇒ (3) Suppose that {(Wi, ωi)}i∈I is a fusion frame in H and f, g ∈ H. By (2) we obtain hπWi S−1 W πWj f, gi = hS−1 W πWj f, πWi gi = 0 (i 6= j). In particular, suppose {eij}j∈Ji is an orthonormal basis for Wi, then we know by the argument above that the sequence {ωieij}i∈I,j∈Ji is a Riesz basis for H with the frame operator SW . Hence, {S−1/2 W ωieij}i∈I,j∈Ji is an orthonormal basis for H. Let f, g ∈ H and take πWi f = Xj∈Ji cij eij, dijeij , πWi g = Xj∈Ji 5 for all i ∈ I, for some {cij}j∈Ji and {dij}j∈Ji in ℓ2. Then, for all i ∈ I we have i πWi S−1 (cid:10)ω2 W πWi f, g(cid:11) = DS−1/2 = hXj∈Ji = Xj,k∈Ji = *Xj∈Ji W ωiπWi f, S−1/2 W ωiπWi gE cij S−1/2 W ωieiki dikS−1/2 W ωieij,Xk∈Ji cijdikDS−1/2 cijeij,Xk∈Ji W ωieij, S−1/2 W ωieikE dikeik+ = hπWi f, gi . So, ω2 W πWi = πWi . i πWi S−1 (3) ⇒ (2) Let f ∈ Wi and g ∈ Wj , where i 6= j. Then hS−1 W f, gi = hS−1 W πWi f, πWj gi = hπWj S−1 W πWi f, gi = 0. From the Proposition 2.5, it easily follows that for a fusion Riesz basis W 1 W πWi fi}i∈I = { ω2 i i fi is its bounded inverse. φW W (fi) = {πWi S−1 W fi}i∈I = {πWi S−1 Therefore, it is bounded and the operator fi 7→ w2 fi}i∈I . (2.5) All items in Proposition 2.5 include a dependency on weights, for (2) the weight is included in the definition of SW . So, the question arises how dependent the Riesz property is on the considered weights. It is worthwhile to mention that, by the fusion frame definition, the family of weights belongs to ℓ∞+ assuming that the subspaces are non-empty, see [37, Remark 2.4]. Moreover, if W = {(Wi, wi)}i∈I is a fusion Riesz basis, then (2.4) shows that (i ∈ I). Using Proposition 2.4 the following lemma immediately follows. Lemma 2.6. Let W = {(Wi, ωi)}i∈I and V = {(Wi, vi)}i∈I be a family of subspaces in H with different weights. Then W is a fusion Riesz basis if and only if V is a fusion Riesz basis. √C ≤ wi ≤ √D, This could also be seen as direct consequence of Proposition 2.4, as the Riesz decomposition property, e.g. item (1) is independent of the weight. In the sequel, for a given fusion Riesz basis W = {(Wi, wi)}i∈I , we denote by W ′ the 1-uniform family of subspaces {(Wi, 1)}i∈I. We will use the following criterion for the invertibility of operators. Proposition 2.7. [33] Let F : H → H be invertible on H. Suppose that G : H → H is a bounded operator and kGh − F hk ≤ υkhk for all h ∈ H, where υ ∈ [0, ). Then G is invertible on H and kF −1k 1 G−1 =P∞k=0[F −1(F − G)]kF −1. 6 3. U -fusion cross Gram matrix of operators In this section, we extend the notion of cross Gram matrices [15] to fusion frames and discuss on their invertibility. We interpret the representation of operators using fusion frames [14] as a generalization of the Gram matrix of operators: Definition 3.1. Let W = {(Wi, ωi)}i∈I be a Bessel fusion sequence for H and V = {(Vi, υi)}i∈I a fusion frame for H. For U ∈ B(H), the matrix operator GU,W,V :(cid:0)Pi∈IL Wi(cid:1)ℓ2 →(cid:0)Pi∈IL Wi(cid:1)ℓ2 given by GU,W,V = φW V T ∗V U TW , is called the U -fusion cross Gram matrix. If U = IH, it is called fusion cross Gram matrix and denoted by GW,V . We use GW for GW,W ; the so called fusion Gram matrix. Note that [GU,W,V (fi)]j ="φW V (vkπVk UXi wifi)k∈I# j∈I =Xi wivjπWj S−1 V πVj U Bj,i {z fi =Xi Bj,ifi, } where Bj,i : Wi → Wj . Therefore GU,W,V is a block-matrix of operators [6]1, which motivates the name (cross-)Gram matrix. Clearly, using (2.3), U -fusion cross Gram matrices are well-defined and kGU,W,V k = kφW V T ∗V U TWk ≤ kφW V kkT ∗V k kUk kTWk ≤ (cid:13)(cid:13)S−1 V (cid:13)(cid:13)kUkpBW BV √BW BV kUk. ≤ AV We have chosen to use GU,W,V = φW V T ∗V U TW instead of the 'naive' GU,W,V = T ∗V U TW . The reason for that is that, by this definition, GU,W,V maps Pi∈IL Wi into itself and is a projection as in the Hilbert space case (albeit an oblique one, see Remark 3.2). We can represent an operator U ∈ B(H) from its U -fusion cross Gram matrix. Suppose W is a dual fusion frame of V , then TWGU,W,V T ∗W S−1 W = TW φW V T ∗V U TW T ∗W S−1 W = U. (3.1) From the ideal property of Sp(H) in B(H) it follows that if U is compact, trace class and Hilbert Schmidt, so is GU,W,V . By using (3.1) we can deduce the converse with some assumptions. Moreover, φV W G∗U,W,V = GU ∗,V,W φ∗W V . 1This could be called a generalized subband matrix, motivated by system identification applications [39]. 7 Remark 3.2. Note that as in the discrete Hilbert space frame case GV,W = G2 V,W and so this an oblique projection whenever V is a dual fusion frame of W . Also TV GV,W = TV , and GV,W (φV W T ∗W ) = φV W T ∗W , but in general the operator is not self-adjoint. Using the above identities and the definition of GV,W we achieve ker (GV,W ) = ker (TV ) and ran (GV,W ) =(cid:8){πVi S−1 W fi}i∈I : {fi}i∈I ∈ ran (T ∗W )(cid:9) . In the next result, we are going to characterize Gram matrices of fusion orthonormal bases. Proposition 3.3. Let W = {(Wi, ωi)}i∈I be a fusion Riesz basis. The following are equivalent. (1) W is a fusion orthonormal basis. (2) GW = IPi∈I L Wi . (3) GW ′ = IPi∈I L W ′ i . Proof. (1) ⇒ (2) Assume that W is a fusion orthonormal basis. Applying (2.5) for all f = {fi}i∈I ∈ Pi∈IL Wi we have (2) ⇒ (1) By Proposition 2.5 we obtain SW ′ = SW ′ TW T −1 W GW f = φW W T ∗W TW f 1 w2 i = {πWi fi}i∈I = f . wjfji∈I wiπWiXj∈I wj fji∈I wiπWiXj∈I = φW W =  = Xi∈I = Xi∈I = TW φW W T ∗W TW T −1 W = TWGW T −1 πWi TW T −1 W W = IH. w2 i πWi S−1 W πWi TW T −1 W W ′ is a fusion orthonormal basis. (1) ⇒ (3) It follows from (1) ⇒ (2) and the fact that W is a fusion orthonormal basis if and only if (3) ⇒ (1) It is enough to show that W ′ is a Parseval fusion frame by Proposition 3.23 of [20]. S2 W ′ = TW ′ T ∗W ′ TW ′T ∗W ′ Now, the invertibility of SW ′ implies SW ′ = IH. = TW ′GW ′ T ∗W ′ = SW ′ . 8 3.1. Invertibility of U -cross Gram matrices We now discuss the relationship between the invertibility of Gram matrices and their associated operators. Proposition 3.4. Let W = {(Wi, ωi)}i∈I be a fusion frame in H and U ∈ B(H). The following are equivalent: (1) W is a fusion Riesz basis and U is invertible. (2) GU,W,W is invertible. (3) GU,W,W is onto. (4) GU,W,W is one to one. (5) GU,fW ,W is invertible. (6) GU,fW ,W is onto. (7) GU,fW ,W is one to one. Proof. (1) ⇒ (2) is trivial by the invertibility of φW W , see (2.5). is a fusion Riesz basis by Theorem 2.3. So, φW W is invertible and then (2) ⇒ (1) By using the invertibility GU,W,W = φW W T ∗W U TW , the operator TW is injective and so W U = (T ∗W )−1 φ−1 W WGU,W,W T −1 W is invertible. (2) ⇒ (3) and (2) ⇒ (4) are trivial. (3) ⇒ (1) If GU,W,W = φW W T ∗W U TW is onto, then φ∗W W = φW W is invertible. Hence, T ∗W U TW = φ−1 W WGU,W,W is onto. Thus, W is a fusion Riesz basis and U is invertible. (1) ⇒ (5) Using Theorem 2.9 of [5],fW is also a fusion Riesz basis and so by (2.5), TfW is invertible by Theorem 2.3. Hence, GU,fW ,W = φfW W T ∗W U TfW is invertible by the invertibility of U and φfW W{fi}i∈I = W fi(cid:9)i∈I , for all {fi}i∈I ∈Pi∈IL Wi. (cid:8)S−1 GU,fW ,W = φfW W T ∗W U TfW is invertible and therefore fW is a fusion Riesz basis by Theorem 2.3. By Theorem 2.9 of [5], W is also a fusion Riesz basis. Applying Theorem 2.3, T ∗W and TfW are invertible and it immediately follows the desired result. (5) ⇒ (1) Note that φfW W is invertible by the definition of φfW W . Also, TfW is one to one since (1) ⇔ (4) ⇔ (6) ⇔ (7) can be proved in an analogue way. Similar to the question, when the inverse of a frame multiplier is a multiplier again [7, 16], we can show that for fusion Riesz bases the operator keeps its structure. To state our results in a accessible way, we need a new definition: Definition 3.5. Let W = {(Wi, wi)}i∈I be a fusion frame and V = {(Vi, vi)}i∈I a Bessel fusion sequence. The alternate cross-fusion frame operator LV W : H → H is defined by We denote LW W as LW . LV W = TV φV W T ∗W . 9 It follows that kLV Wk ≤ √BW BV AW and (LV W )∗ =Xi∈I πWi S−1 W πVi . Obviously, LV W is an invertible operator if and only if V is a pseudo-dual of W . In particular, LV W = IH if and only if V is a dual of W . In the following we summarize the basic properties of LW . Proposition 3.6. Let W = {(Wi, wi)}i∈I be a fusion frame. Then LW is positive, self-adjoint and invertible operator. Proof. It is easy to see that φW W is self-adjoint and so, LW is self-adjoint. Moreover, for all f ∈ H we have w2 i πWi S−1 W πWi f, f+ W πWi f, wi · S−1/2 hLW f, fi = *Xi∈I = Xi∈IDwi · S−1/2 = Xi∈I 2! W πWi f(cid:13)(cid:13)(cid:13) i(cid:13)(cid:13)(cid:13)S−1/2 W (cid:13)(cid:13)(cid:13) ≥ (cid:13)(cid:13)(cid:13)S1/2 (cid:13)(cid:13)(cid:13)S1/2 W (cid:13)(cid:13)(cid:13) kT ∗W fk2 ≥ AW w2 −2 W πWi fE 2kfk2, where AW is a lower bound of the fusion frame W . This shows that LW is positive and an invertible operator in B(H). As an easy consequence of Proposition 2.5, we state Corollary 3.7. Let W = {(Wi, ωi)}i∈I be a fusion Riesz basis. Then LW = SW ′ . Theorem 3.8. Let W = {(Wi, ωi)}i∈I and V = {(Vi, υi)}i∈I be fusion frames in H and U an invertible operator in B(H). Then the following assertions hold. (1) If GU,W,W is invertible, then W is a fusion Riesz basis and W ′ U −1S−1 W ′ ,W,W . G−1 U,W,W = GS−1 (2) If GU,V,W is invertible and V is a dual of W , then V is a fusion Riesz basis and G−1 (3) If GU,W,V is invertible, then W is a fusion Riesz basis and U,V,W = GU −1,V,W . G−1 U,W,V = G(LW V U SW ′ )−1,W,W , Proof. (1) By Proposition 3.4 W is a fusion Riesz basis. Using Corollary 3.7 follows that GU,W,WGS−1 W ′ U −1S−1 W ′ ,W,W = φW W T ∗W U TW φW W T ∗W S−1 W ′ U−1S−1 W ′TW = φW W T ∗W S−1 W ′ TW = I(Pi∈I L fWi)ℓ2 . With the same we have GS−1 W ′ U −1S−1 W ′ ,W,WGU,W,W = I(Pi∈I L fWi)ℓ2 . 10 (2) According to (2.2) we have TV φV W T ∗W = IH. In addition, the invertibility of GU,V,W implies that φV W T ∗W has a right inverse. Using Proposition 2.3 V is a fusion Riesz basis, and therefore, Hence, φV W T ∗W TV = I(Pi∈I L Vi)ℓ2 . GU,V,WGU −1,V,W = φV W T ∗W U TV φV W T ∗W U−1TV = φV W T ∗W TV = I(Pi∈I L Vi)ℓ2 . Similarly, GU −1,V,WG−1 U,V,W = I(Pi∈I L fWi)ℓ2 . (3) It follows from Corollary 3.7. Repeating the previous argument and using (2.2) leads to a characterization for fusion Riesz bases due to U -fusion cross Gram matrices. Theorem 3.9. Let V = {(Vi, υi)}i∈I be a dual fusion frame of W = {(Wi, ωi)}i∈I in H. The following are equivalent: (1) V is a fusion Riesz basis. (2) GV,W = I(Pi∈I L Vi)ℓ2 . (3) GV,W has a left inverse. Consider W = {(Wi, ωi)}i∈I , V = {(Vi, υi)}i∈I and Z = {(Zi, zi)}i∈I as fusion frames in H and U1, U2 ∈ B(H), it is obvious to see that GU1,W,V GU2,W,Z = GU1TW φW Z T ∗ Z U2,W,V . In particular, if V = {(Vi, υi)}i∈I is a dual of Z = {(Zi, zi)}i∈I , then GU1,V,WGU2,V,Z = GU1U2,V,W . As a special case, if U is invertible and V a fusion Riesz basis such that V is a dual of W , then GU,V,W has an inverse in the form of Gram matrices The above identity is also proved in Theorem 3.8. (GU,V,W )−1 = GU −1,V,W . 11 3.2. Pseudo-inverses In the following we discuss the pseudo-inverse of U -fusion cross Gram matrices with closed range, and under some conditions we represent their pseudo-inverse as a U -fusion cross Gram matrix again motivated by the discrete frame case [15]. In the following we state a sufficient condition for a U -fusion cross Gram matrix having closed range. Lemma 3.10. Let V = {(Vi, υi)}i∈I be a dual fusion frame of W = {(Wi, ωi)}i∈I in H and U an operator in B(H) such that U V = {(U Vi, υi)}i∈I is also a fusion frame in H. Then GU,V,W has closed range and Proof. The dual condition (2.2) follows that φV W T ∗W has closed range. Thus, ran (GU,V,W ) = ran (φV W T ∗W ) . ran (GU,V,W ) = ran (φV W T ∗W U TV ) = ran (φV W T ∗W TUV ) = ran (φV W T ∗W ) . The assumptions in the above lemma are fulfilled, in particular, for U being invertible [32]. Theorem 3.11. Let V = {(Vi, υi)}i∈I be a dual fusion frame of W = {(Wi, ωi)}i∈I in H, also let U ∈ B(H) and GU,V,W have closed range. The following are equivalent: (1) G†U,V,W = GU †,V,W . (2) ran (φV W T ∗W U∗) = ran (T ∗V U∗) and ran (φV W T ∗W U ) = ran (T ∗V U ). (3) φV W T ∗W U∗ = T ∗V S−1 Proof. (2) ⇒ (1) Applying (2.2) follows that V U∗ and φV W T ∗W U = T ∗V S−1 V U . GU,V,WGU †,V,WGU,V,W = φV W T ∗W U TV φV W T ∗W U†TV φV W T ∗W U TV = φV W T ∗W U U†U TV = GU,V,W . In addition, (2.2) is also follows that TW φ∗V W is surjective. Hence, we have ran(cid:0)GU †,V,W(cid:1) = ran(cid:0)φV W T ∗W U†TV(cid:1) = ran (φV W T ∗W U∗) (by (2)) = ran (T ∗V U∗) = ran (T ∗V U∗TW φ∗V W ) = ran(cid:0)G∗U,V,W(cid:1) . 12 Moreover, ker(cid:0)GU †,V,W(cid:1) = ker(cid:0)φV W T ∗W U†TV(cid:1) = ker(cid:0)U†TV(cid:1) = ran(cid:0)T ∗V (U†)∗(cid:1)⊥ = ran (T ∗V U )⊥ (by (2)) = ran (φV W T ∗W U )⊥ = ker (U∗TW φ∗V W ) = ker(cid:0)G∗U,V,W(cid:1) . Thus, (1) is obtained. The converse follows immediately from the above identities. To show (2) ⇒ (3) using Douglas's theorem [27], there is an operator C ∈ B(H) such that So, φV W T ∗W U = T ∗V U C. φV W T ∗W U = T ∗V U C = T ∗V S−1 = T ∗V S−1 V TV T ∗V U C V TV φV W T ∗W U = T ∗V S−1 V U. The other identity is obtained similarly. The converse is clear. Note that fusion Riesz bases satisfy in the assumptions of Theorem 3.11. These assumptions might In particular, the failure to fulfill this equality makes the concept of seem obvious, but are not [32]. duality of fusion frames interesting, so fusion frames are not 'just another' generalization of frames. Remark 3.12. In the proof of Theorem 3.11(1) we can replace the fusion frame appeared in the right side by any fusion frame Z such that V is its dual. More precisely, let W and Z be fusion frames and V is a dual of both W and Z. Then if and only if φV Z T ∗Z U∗ = T ∗V S−1 G†U,V,W = GU †,V,Z, V U∗ and φV Z T ∗Z U = T ∗V S−1 V U . Using a similar argument as Theorem 3.11 we obtain the following. Theorem 3.13. Let W = {(Wi, ωi)}i∈I be a fusion frame in H, also let U ∈ B(H) and GU,V,W have closed range. The following are equivalent: (1) (GU,W,W )† = GL−1 W ,W,W . (2) The operators φW W T ∗W L−1 W U∗ and T ∗W U∗ have the same range as φW W T ∗W U and T ∗W (L−1 W )∗U , W U †L−1 respectively, for all i ∈ I W U∗ = T ∗W S−1 (3) φW W T ∗W L−1 W U∗ and φW W T ∗W U = T ∗W S−1 W U . 13 4. Stability of U -cross Gram Matrices of Operators In this section, we state a general stability for the invertibility of U -fusion cross Gram matrices, compare to the results on the invertibility of multipliers [46]. Theorem 4.1. Let W = {(Wi, υi)}i∈I be a Bessel fusion sequence and U1, U2 ∈ B(H) with kU1 − U2k < µ. Also, let V = {(Vi, υi)}i∈I and Z = {(Zi, υi)}i∈I be fusion frames on H such that GU1,W,V is invertible and 2 + ǫkfk, (4.1) , (4.2) 2 2 υ2 υ2 Xi∈I for all f ∈ H, in addition i kπZi fk2! 1 i kπZi f − πVi fk2! 1 + λ2 Xi∈I i kπVi fk2! 1 kU2k + kU2k < (cid:13)(cid:13)(cid:13)G−1 U1,W,V(cid:13)(cid:13)(cid:13) B(cid:13)(cid:13)S−1 V (cid:13)(cid:13) where ǫ > 0, B = max{BW , BV , BZ} andPi∈I υi2 < ∞. Then GU2,W,Z is also invertible. ≤ λ1 Xi∈I √B(cid:19)√B(cid:13)(cid:13)S−1 Z (cid:13)(cid:13) Xi∈I µ +(cid:18)λ1 + λ2 + υi2! 1 Proof. First note that −1 υ2 ǫ 2 υ2 k(SV − SZ ) fk = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xi∈I ≤ Xi∈I i (πVi f − πZi f )(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 Xi∈I υi2! 1 υ2 i k(πVi f − πZi f )k2! 1 2 , for all f ∈ H. Therefore, 2 i (πVi − πZi ) S−1 (cid:13)(cid:13)(cid:0)S−1 Z (cid:1) f(cid:13)(cid:13) V − S−1 Z f(cid:13)(cid:13) = (cid:13)(cid:13)S−1 V (SV − SZ) S−1 Z f(cid:13)(cid:13) V (cid:13)(cid:13)(cid:13)(cid:13)(SV − SZ ) S−1 ≤ (cid:13)(cid:13)S−1 V (cid:13)(cid:13) Xi∈I υi2! 1 2 Xi∈I(cid:13)(cid:13)υ2 ≤ (cid:13)(cid:13)S−1 V (cid:13)(cid:13) Xi∈I υi2! 1 ≤ (cid:13)(cid:13)S−1 λ1 Xi∈I Z f(cid:13)(cid:13)2! 1 i(cid:13)(cid:13)πVi S−1 i(cid:13)(cid:13)πZi S−1 V (cid:13)(cid:13) Xi∈I υi2! 1 2(cid:16)λ1pBZ + λ2pBV + ǫ(cid:17)(cid:13)(cid:13)S−1 ≤ (cid:13)(cid:13)S−1 Z f(cid:13)(cid:13)2! 1 Z f(cid:13)(cid:13)2! 1 Z (cid:13)(cid:13)kfk. + λ2 Xi∈I υ2 υ2 2 2 2 14 Z f(cid:13)(cid:13) + ǫ(cid:13)(cid:13)S−1 Using the assumption (4.1) and the above computations imply that 2 υ2 υ2 υ2 υ2 V − S−1 V πVi − S−1 V πVi − S−1 Xi∈I Z πZi )f(cid:13)(cid:13)2! 1 i(cid:13)(cid:13)(S−1 ≤ Xi∈I V πZi )f(cid:13)(cid:13) +(cid:13)(cid:13)(S−1 i(cid:0)(cid:13)(cid:13)(S−1 ≤ Xi∈I V (cid:13)(cid:13)kπVi f − πZi fk +(cid:13)(cid:13)S−1 i(cid:0)(cid:13)(cid:13)S−1 ≤  Xi∈I kπVi f − πZi fk2! 1 +(cid:13)(cid:13)S−1 i(cid:13)(cid:13)S−1 V (cid:13)(cid:13)2 V (cid:13)(cid:13)(cid:16)λ1pBZ + λ2pBV + ǫ(cid:17)kfk+ ≤ h(cid:13)(cid:13)S−1 (cid:13)(cid:13)S−1 V (cid:13)(cid:13)(cid:13)(cid:13)S−1 V (cid:13)(cid:13)(cid:16)λ1pBZ + λ2pBV + ǫ(cid:17)1 +(cid:13)(cid:13)S−1 ≤ (cid:13)(cid:13)S−1 Z (cid:13)(cid:13)pBZ Xi∈I υi2! 1 2 V πZi − S−1 2 2 Z πZi )f(cid:13)(cid:13)(cid:1)2! 1 Z (cid:13)(cid:13)kπZi fk(cid:1)2! 1 Z (cid:13)(cid:13) Xi∈I V − S−1 υ2 2 i kπZi fk2! 1 2(cid:16)λ1pBZ + λ2pBV + ǫ(cid:17)kfk υi2! 1 Z (cid:13)(cid:13)pBZ Xi∈I 2kfk. Finally, applying (4.2) we obtain kGU1,W,V − GU2,W,Zk = kφW V T ∗V U1TW − φW Z T ∗Z U2TWk ≤ kφW V T ∗V U1TW − φW V T ∗V U2TWk + kφW V T ∗V U2TW − φW Z T ∗Z U2TWk V πVi − S−1 υ2 V πVi − S−1 i(cid:13)(cid:13)(S−1 V (cid:13)(cid:13)pBW BV kU1 − U2k + Xi∈I(cid:13)(cid:13)υiπWi (S−1 ≤ (cid:13)(cid:13)S−1 V (cid:13)(cid:13)pBW BV µ +"Xi∈I ≤ (cid:13)(cid:13)S−1 V (cid:13)(cid:13)pBW BV µ + ≤ (cid:13)(cid:13)S−1 V (cid:13)(cid:13)(cid:16)λ1pBZ + λ2pBV + ǫ(cid:17)1 +(cid:13)(cid:13)S−1 (cid:13)(cid:13)S−1 < (cid:13)(cid:13)(cid:13)G−1 U1,W,V(cid:13)(cid:13)(cid:13) −1 . 2 Z πZi )(cid:13)(cid:13)2# 1 Z (cid:13)(cid:13)pBZ Xi∈I 2 Z πZi )U2TW(cid:13)(cid:13)2! 1 kU2kpBW υi2! 1 2 kU2kpBW Therefore, GU2,W,Z is also invertible by Proposition 2.7. 15 Remark 4.2. It is worthwhile to mention that if we consider in Theorem 4.1 1. the perturbation condition kπZi f − πVi fk ≤ λ1υi kπZi fk + λ2υi kπVi fk + ǫkfk we can replace the assumption Pi∈I υi2 < ∞ by bounded weights. Hence, uniform fusion frames satisfy a slightly different version of this theorem. 2. the condition Xi∈I kπZi f − πVi fk2! 1 2 ≤ λ1 Xi∈I i kπZi fk2! 1 2 υ2 + λ2 Xi∈I i kπVi fk2! 1 2 υ2 + ǫkfk, instead of (3.4) we get the same result substitutingPi∈I υi4 forPi∈I υi2. As a consequence we obtain the following result. Corollary 4.3. Let W = {(Wi, υi)}i∈I be a fusion Riesz basis with bounds AW and BW and Z = {(Zi, υi)}i∈I a fusion frame in H such that (4.1) holds. Also, U ∈ B(H) with kU − IHk < µ. If µ +(cid:18)λ1 + λ2 + ǫ √B(cid:19)√B(cid:13)(cid:13)S−1 Z (cid:13)(cid:13) Xi∈I 2 υi2! 1 B(cid:13)(cid:13)S−1 W (cid:13)(cid:13) where ǫ > 0 and B = max{BW , BZ} andPi∈I υi2 < ∞. Then GU,W,Z is also invertible. , kUk + kUk < AW 5. Acknowledgement The first and the second authors were supported in part by the Iranian National Science Foundation (INSF) under Grant 97018155. The last author was partly supported by the START project FLAME Y551-N13 of the Austrian Science Fund (FWF) and the DACH project BIOTOP I-1018-N25 of Austrian Science Fund (FWF). He also thanks Nora Simovich for help with typing. References [1] A. Aldroubi, A. Baskakov, I. Krishtal. Slanted matrices, banach frames, and sampling. J. Funct. Anal., 255(7):1667 -- 1691, 2008. [2] A. Aldroubi, C. Cabrelli, A. C¸ akmak, U. Molter, A. Petrosyan. Iterative actions of normal operators. J. Funct. Anal., 272(3):1121 -- 1146, 2017. [3] S. T. A. Ali, J.-P. Antoine, and J.-P. Gazeau. Continuous frames in Hilbert space. Ann. Physics, 222(1):1 -- 37, 1993. [4] J.-P. Antoine and P. Balazs. Frames and semi-frames. J. Phys. A , 44:205201, 2011. 16 [5] A. A. Arefijamaal and F. A. Neyshaburi. Some properties of alternate duals and approximate alternate duals of fusion frames. Turkish J. Math., 41(5):1191 -- 1203, 2018. [6] F. V. Atkinson, H. Langer, R. Mennicken, and A. A. Shkalikov. The essential spectrum of some matrix operators. Math. Nachr., 167(1):5 -- 20, 1994. [7] P. Balazs. Basic definition and properties of Bessel multipliers. J. Math. Anal. Appl., 325(1):571 -- 585, 2007. [8] P. Balazs. Matrix-representation of operators using frames. Sampl. Theory Signal Image Process., 7(1):39 -- 54, 2008. [9] P. Balazs, J.-P. Antoine, and A. Grybos. Weighted and controlled frames: Mutual relationship and first numerical properties. Int. J. Wavelets Multiresolut. Inf. Process., 8(1):109 -- 132, 2010. [10] P. Balazs and K. Grochenig. A guide to localized frames and applications to Galerkin-like representations of operators. In I. Pesenson, H. Mhaskar, A. Mayeli, Q. T. L. Gia, and D.-X. Zhou, editors, Frames and Other Bases in Abstract and Function Spaces, Applied and Numerical Harmonic Analysis series (ANHA). Birkhauser/Springer, 2017. [11] P. Balazs, N. Holighaus, T. Necciari, and D. Stoeva. Frame theory for signal processing in psychoacoustics. In R. Balan, J. J. Benedetto, W. Czaja, and K. Okoudjou, editors, Excursions in Harmonic Analysis Vol. 5, Springer, 2017. [12] P. Balazs and G. Rieckh. Oversampling operators: Frame representation of operators. Analele Universitatii "Eftimie Murgu", 18(2):107 -- 114, 2011. [13] P. Balazs and G. Rieckh. Redundant representation of operators. arXiv:1612.06130. [14] P. Balazs, M. Shamsabadi, A. A. Arefijamaal and C. Gardon. Representation of operators using fusion frames. preprint. [15] P. Balazs, M. Shamsabadi, A. A. Arefijamaal, and A. Rahimi. U -cross Gram matrices and their invertibility. J. Math. Anal. Appl., 476(2):367-390, 2019. [16] P. Balazs and D. Stoeva. Representation of the inverse of a multiplier. J. Math. Anal. Appl., 422:981 -- 994, 2015. [17] J. Benedetto and P. Ferreira. Modern sampling theory. Mathematics and applications. Birkhauser, 2001. [18] B. G. Bodmann, P. G. Casazza. The road to equal-norm Parseval frames. J. Funct. Anal., 258(2): 397 -- 420, 2010. [19] H. Bolcskei, F. Hlawatsch, and H. G. Feichtinger. Frame-theoretic analysis of oversampled filter banks. IEEE Trans. Signal Processing, 46(12):3256 -- 3268, 1998. [20] P. G. Casazza and G. Kutyniok. Frames of subspaces. Cont. Math., 2004. [21] P. G. Casazza, G. Kutyniok, and S. Li. Fusion frames and distributed processing. Appl. Comput. Harmon. Anal., 254(1):114 -- 132, 2008. [22] O. Christensen. Frames and Bases. An Introductory Course. Applied and Numerical Harmonic Analysis. Basel Birkhauser, 2008. [23] O. Christensen and R. S. Laugesen. Approximately dual frames in Hilbert spaces and applications to Gabor frames. Sampl. Theory Signal Image Process, 9(3):77 -- 89, 2010. 17 [24] N. Cotfas and J. P. Gazeau. Finite tight frames and some applications. J. Phys. A, 43(19):193001, 2010. [25] S. Dahlke, T. Raasch, M. Werner, M. Fornasier, and R. Stevenson. Adaptive frame methods for elliptic operator equations: The steepest descent approach. IMA J. Numer. Anal., 27(4):717 -- 740, 2007. [26] I. Daubechies, A. Grossmann, and Y. Meyer. Painless non-orthogonal expansions. J. Math. Phys., 27:1271 -- 1283, 1986. [27] R. G. Douglas, On majorization, factorization and range inclusion of operators onHilbert space. Proc. Amer. Math. Soc., 17(2):413 -- 415, 1996. [28] R. J. Duffin and A. C. Schaeffer. A class of nonharmonic Fourier series. Trans. Amer. Math. Soc., 72:341 -- 366, 1952. [29] H. G. Feichtinger and T. Strohmer. Gabor Analysis and Algorithms - Theory and Applications. Birkhauser Boston, 1998. [30] H. Fuhr, J. Lemvig. System bandwidth and the existence of generalized shift-invariant frames. J. Funct. Anal., 276(2):563 -- 601, 2019. [31] L. Gaul, M. Kogler, and M. Wagner. Boundary Element Methods for Engineers and Scientists. Springer, 2003. [32] P. Gavrut¸a. On the duality of fusion frames. J. Math. Anal. Appl., 333(2):871 -- 879, 2007. [33] I. Gohberg, S. Goldberg, and M. A. Kaashoek. Basic Classes of Linear Operators. Birkhauser, 2003. [34] K. Grochenig. Foundations of Time-Frequency Analysis. Birkhauser Boston, 2001. [35] H. Harbrecht, R. Schneider, and C. Schwab. Multilevel frames for sparse tensor product spaces. Numer. Math., 110(2):199 -- 220, 2008. [36] S. B. Heineken and P. Morillas. Properties of finite dual fusion frames. Linear Algebra and its Applications, 453:1 -- 27, 2014. [37] S. B. Heineken, P. Morillas, A. Benavente, and M. Zakowicz. Dual fusion frames. Arch. Math. (Basel), 103(4):355 -- 365, 2014. [38] W. Kreuzer, P. Majdak, and Z. Chen. Fast multipole boundary element method to calculate head-related transfer functions for a wide frequency range. J. Acoust .Soc. Am., 126(3):1280 -- 1290, 2009. [39] D. Marelli and M. Fu. Performance analysis for subband identification. IEEE Trans. Signal Process., 51(12):3128 -- 3142, 2003. [40] P. Oswald. Stable space splittings and fusion frames. In Wavelets XIII (V. Goyal, M. Papadakis, D. Van de Ville, eds.), Proceedings of SPIE San Diego, volume 7446, 2009. [41] A. Pietsch. Operator Ideals. North-Holland Publishing Company, 1980. [42] R. Schatten. Norm Ideals of Completely Continuous Operators. Springer Berlin, 1960. [43] M. Shamsabadi and A. A. Arefijamaal. The invertibility of fusion frame multipliers. Linear Multilinear Algebra, 65(5):1062 -- 1072, 2016. [44] M. Speckbacher and P. Balazs. Reproducing pairs and the continuous nonstationary Gabor transform on lca groups. J. Phys. A, 48:395201, 2015. 18 [45] R. Stevenson. Adaptive solution of operator equations using wavelet frames. SIAM J. Numer. Anal., 41(3):1074 -- 1100, 2003. [46] D. T. Stoeva and P. Balazs. Invertibility of multipliers. Appl. Comput. Harmon. Anal., 33(2):292 -- 299, 2012. [47] W. Sun. G-frames and g-Riesz bases. J. Math. Anal. Appl., 322(1):437 -- 452, 2006. [48] J. Weidmann. Linear Operators in Hilbert Spaces. Springer New York, 1980. 19
1309.4280
1
1309
2013-09-17T12:22:22
From local to global ideal-triangularizability
[ "math.FA" ]
Let L be a Banach lattice with order continuous norm, and let S be a multiplicative semigroup of ideal-triangularizable positive compact operators on L such that, for every pair A, B in S, the atomic diagonal of the commutator AB-BA is equal to zero. We prove that the semigroup S is ideal-triangularizable.
math.FA
math
FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY ROMAN DRNOVSEK, MARKO KANDI ´C Abstract. Let L be a Banach lattice with order continuous norm, and let S be a multiplicative semigroup of ideal-triangularizable positive compact operators on L such that, for every pair {S, T } ⊆ S, the atomic diagonal of the commutator ST − T S is equal to zero. We prove that the semigroup S is ideal-triangularizable. Math. Subj. Classification (2000): 47A15, 47B65. Key words: Invariant subspaces, Banach lattices, closed ideals, positive operators, idem- potents, the atomic diagonal, semigroups of operators. 1. Introduction Let L be a normed Riesz space, and let L+ denote the positive cone of L. Elements of L+ are called positive vectors. A linear subspace J of L is an ideal in L whenever x ≤ y and y ∈ J imply x ∈ J. A band in L is an order closed ideal. A band B in L is a projection band whenever L = B ⊕ Bd. In this case, there exists a positive projection onto B with a kernel Bd, and the direct sum is also an order direct sum. A Banach lattice L is said to have an order continuous norm whenever every decreasing net {xα}α with infimum 0 of positive vectors in L converges in norm to 0. It is well known that a Banach lattice with order continuous norm is Dedekind complete, and that its closed ideals are bands [2]. Moreover, every band in a Dedekind complete Banach lattice is a projection band. A non-zero vector a ∈ L+ is an atom in a normed Riesz space L if 0 ≤ x, y ≤ a and x ∧ y = 0 imply either x = 0 and y = 0, or equivalently, if 0 ≤ x ≤ a implies x = λa for some λ ≥ 0, i.e., the principal ideal Ba generated by a is one dimensional. It turns out that Ba is a projection band [13]. The decomposition L = Ba ⊕ Bd a implies that for an arbitrary (positive) vector x ∈ L there exist a (positive) scalar λx and a (positive) vector Date: June 4, 2018. THE PAPER WILL APPEAR IN LINEAR AND MULTILINEAR ALGEBRA. 1 2 ROMAN DRNOVSEK, MARKO KANDI ´C yx ∈ Bd a such that x = λxa + yx. The linear functional ϕa : L → R associated to the atom a is defined by ϕa(x) = λx. A positive vector x ∈ L is a quasi-interior point in L if the principal ideal in L generated by x is dense in L. A (linear) operator T between Riesz spaces L1 and L2 is said to be positive if T maps the positive cone L+ 1 into the positive cone L+ 2 . The absolute kernel N (T ) of T is defined as N (T ) = {x ∈ L1 : T x = 0}. If it is a zero ideal, then T is said to be strictly positive. The range ideal R(T ) of T is the ideal generated by the range of T . Recall that every positive operator T on a Banach lattice is continuous, and its spectral radius r(T ) belongs to its spectrum σ(T ). Throughout the text we assume that functionals and operators acting on normed Riesz spaces are continuous. A family F of operators on a normed Riesz space L is said to be ideal-reducible when- ever there exists a nontrivial closed ideal in L that is invariant under all operators from the family F . Otherwise, we say that F is ideal-irreducible. The following very useful proposition is proved in [8]. Proposition 1.1. Let L be a normed Riesz space, and let S be a semigroup of positive operators on L. The following statements are equivalent: (a) S is ideal-reducible; (b) there exist a nonzero positive functional ϕ ∈ L∗ and a nonzero positive vector f ∈ L+ such that ϕ(Sf ) = {0}; (c) there exist nonzero positive operators A and B on L such that ASB = {0}; (d) some nonzero semigroup ideal of S is ideal-reducible. If there is a chain C that is maximal as a chain of closed ideals of L and that has the property that every ideal in C is invariant under all the operators in a family F , then F is said to be ideal-triangularizable, and C is an ideal-triangularizing chain for F . Every maximal chain of closed ideals is also maximal as a chain of closed subspaces of L [6]. Let I and J be closed ideals in L that are invariant under every operator from a family F . If I ⊆ J, then F induces a family bF of operators on the quotient normed Riesz space J/I as follows. For each T ∈ F the operator bT is defined on J/I by bT (x + I) = T x + I. Any such family bF is called a family of ideal-quotients of the family F . A set P of properties is said to be inherited by ideal-quotients if every family of ideal-quotients of a family of operators satisfying P also satisfies the same properties. Ideal-triangularizability of operators on FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY1 3 Banach lattices is in practice often reduced to ideal-reducibility of operators. The details are contained in the following lemma that was proved in [5]. Lemma 1.2 (The Ideal-triangularization Lemma). Let P be the set of properties inher- ited by ideal-quotients. If every family of operators on a Banach lattice of dimension greater than one which satisfies P is ideal-reducible, then every such family is ideal- triangularizable. Let T be an operator on a Banach lattice L with an order continuous norm. If B1 and B2 are bands of L such that B2 ⊆ B1, then the quotient Banach lattice B1/B2 is isometrically lattice isomorphic to the band B := B1 ∩ Bd 2, and so the norm of B is also order continuous. If T leaves the bands B1 and B2 invariant, then the induced operator T on B1/B2 can be identified with the compression of T on B, that is, with the operator PBT B, where PB denotes the band projection on B. For the terminology not explained in the text about normed Riesz spaces or Banach lattices and operators acting on them we refer the reader to classical textbooks [1], [2], [16] and [13]. Finally, we explain the organization of the paper. Section 2 deals with the structure of positive idempotents of finite ranks. In Section 3 the atomic diagonal operator D introduced recently in [12] is connected with the known band projection onto the center of a Dedekind complete Banach lattice. The main result of the paper is Theorem 4.5 (and its finite-dimensional special case Corollary 4.6) in Section 4. It reveals the relation between the atomic diagonal operator and the ideal-triangularizability of semigroups of positive operators. We also prove Theorem 4.2 that completes [12, Theorem 3] and extends [3, Theorem 3.8] and [14, Corollary 27]. 2. The structure of positive idempotents of finite rank The structure result for positive idempotents on Lp-spaces (1 ≤ p < ∞) was given by Zhong in [19]; see also the book [15, Section 8.7] for a complete treatment. In this section we extend this result to Banach lattices with order continuous norm. Let L be a normed Riesz space, and let P be a positive idempotent operator of rank one on L. Then there exist a positive vector x ∈ L and a positive linear functional ϕ on L such that ϕ(x) = 1 and P = x ⊗ ϕ, i.e., P y = ϕ(y)x for y ∈ L. The ideal-irreducibility 4 ROMAN DRNOVSEK, MARKO KANDI ´C of such an idempotent is characterized by the following lemma that is a generalization of [15, Lemma 8.7.11]. Lemma 2.1. Let L be a normed Riesz space, and let P = x ⊗ ϕ be a positive idempotent of rank one on L, where x ∈ L is a positive vector and ϕ a positive linear functional on L. The following statements are equivalent: (a) P is ideal-irreducible; (b) The vector x is a quasi-interior point of L, and ϕ is a strictly positive functional on L; (c) P and P ∗ are strictly positive operators on L and L∗, respectively. Proof. To see that (a) implies (b), assume that P is ideal-irreducible. If y is a positive vector in the absolute kernel N (ϕ) of ϕ, then P y = ϕ(y)x = 0 implies that N (ϕ) is invariant under P . Since P is ideal-irreducible, we have N (ϕ) = {0}, so that ϕ is strictly positive on L. If 0 ≤ y ≤ λx with y ∈ L and λ ≥ 0, then P y ≤ P y ≤ λP x = λϕ(x)x implies that the closure of the principal ideal generated by the positive vector x is invariant under P . Hence, x needs to be a quasi-interior point in L. Assume that (b) holds. To prove (a), assume that J is a proper closed ideal that is invariant under the operator P . By [8, Lemma 1.1], there exists a nonzero positive functional ψ on L which is zero on J. Then for y ∈ J we have 0 = ψ(P y) = ψ(x)ϕ(y). Since x is a quasi-interior point, ψ(x) > 0 by [1, Lemma 4.15], and so we conclude that ϕ(y) = 0. Since ϕ is a strictly positive functional, this implies that y = 0, so that J = {0}. Therefore, the operator P does not have nontrivial closed invariant ideals. Assume again that (b) holds. To show (c), assume that P y = 0 and P ∗ψ = 0 for some positive vectors y and ψ in L and L∗, respectively. Since ϕ is a strictly positive functional on L, the equality 0 = P y = ϕ(y)x implies that y = 0. Since x is a quasi-interior point in L, the equality 0 = P ∗(ψ) = ψ(x)ϕ implies that ψ = 0 by [1, Lemma 4.15]. To see that (c) implies (b), suppose that P and P ∗ are strictly positive operators on L and L∗, respectively. Let ψ be an arbitrary positive functional on L with ψ(x) = 0. Then P ∗ψ = (ϕ ⊗ x)ψ = ψ(x)ϕ = 0 implies ψ = 0, since P ∗ is strictly positive. It follows from [1, Lemma 4.15] that x is a quasi-interior point in L. Suppose now that ϕ(y) = 0 for some positive vector y ∈ L. Then P y = (x ⊗ ϕ)(y) = ϕ(y)x = 0. Since P is strictly positive, we have y = 0, so that ϕ is strictly positive on L. (cid:3) FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY2 5 Note that in the case when any of the equivalent statements of Lemma 2.1 holds for a positive ideal-triangularizable idempotent P of rank one, then L is lattice isomorphic to R, and P is just the identity on R. Positive idempotents of finite rank are treated in the following propositions that ex- tend [15, Proposition 8.7.12]. It should be noted that these propositions contain addi- tional statements regarding the structure of finite rank idempotents that are also ideal- triangularizable. Proposition 2.2. Let L be a Banach lattice with order continuous norm and P a positive idempotent of finite rank r such that P and P ∗ are strictly positive. There exist pairwise disjoint bands L1, . . . , Lr in L such that L = L1 ⊕ · · · ⊕ Lr and P is of the form P = P1 ⊕ · · · ⊕ Pr, where Pj is an ideal-irreducible positive idempotent of rank 1 on Lj. If, in addition, P is ideal-triangularizable, then r = dim L and P is the identity operator on L. Proof. If r = 1, then the assertion follows from Lemma 2.1 and the remark following it. Assume r > 1. Let x and y be linearly independent positive vectors in the range of the operator P. Replacing y by x + y (if necessary) we may assume 0 ≤ x ≤ y and x 6= y. Let t0 be the supremum of the nonempty set {t ∈ R : t ≥ 0, ty ≤ x} that is bounded from above. Since t0y ≤ x ≤ y and x 6= y, we have t0 < 1. Pick s ∈ (t0, 1) and let z = x − sy. The vectors z+ and z− are both nonzero, since z is neither positive nor negative. From z = P z = P z+ − P z− ≤ P z+ it easily follows z+ ≤ P z+. Since P is strictly positive, P (P z+ − z+) = 0 implies P z+ = z+. Let J be the closed ideal generated by the vector z+. Note that J is invariant under P and it is nontrivial, since z− is nonzero. Therefore, the operator P can be decomposed as with respect to the decomposition L = J ⊕ J d. (cid:20) P1 P2 0 P4 (cid:21) We claim that P2 = 0. Since P 2 = P , we have P 2 1 = P1, P 2 4 = P4 and P1P2 +P2P4 = P2. We conclude that P1P2P4 = 0. Since the operator P is strictly positive, the operator P1 is also strictly positive, and so P2P4 = 0. Taking the adjoints we have P ∗ the adjoint P ∗ is strictly positive, so is the operator P ∗ 2 = 0. Since 4 on (J d)∗, and hence P2 = 0 as 4 P ∗ claimed. We finish the proof by induction on r. (cid:3) (1) P = 0 XQ XQY QY 0 Q 0 0 0  = X I 0  Q(cid:2) 0 I Y (cid:3) , where Q is an idempotent of rank r such that Q and Q∗ are strictly positive operators on B2 and B∗ 2, respectively. Moreover, there exist pairwise disjoint bands L1, . . . , Lr in B2 with B2 = L1 ⊕ · · · ⊕ Lr such that Q = Q1 ⊕ · · · ⊕ Qr, where Qj is a strictly positive idempotent of rank one on Lj. If, in addition, P is ideal-triangularizable, then the dimension of B2 is finite and Q is the identity operator on B2. Proof. Let B1 be the absolute kernel of the operator P , and let B be the band generated by B1 and the range of the operator P . Let B2 = B ∩ Bd 1 and B3 = Bd. Let 6 ROMAN DRNOVSEK, MARKO KANDI ´C Proposition 2.3. Let L be a Banach lattice with order continuous norm and P a positive idempotent of finite rank r on L. There exist pairwise disjoint bands B1, B2 and B3 with L = B1 ⊕ B2 ⊕ B3 such that P can be decomposed as  0 X Z 0 Q Y 0 0 0  be the operator matrix corresponding to the operator P with respect to the decomposition L = B1 ⊕ B2 ⊕ B2. Since P is an idempotent, we have Q2 = Q, X = XQ, Y = QY and Z = XY , so that the decomposition (1) is proved. It also shows that the rank of Q must be r. We claim that Q and Q∗ are strictly positive operators on B2 and B∗ 2, respectively. Suppose that for some positive vector x ∈ B2 we have Qx = 0. Then Xx = XQx = 0, so that P x = 0 2 satisfies Q∗ϕ = 0. Then ϕ is implies x = 0. Suppose that a positive functional ϕ in B∗ zero on the range of the operator P which implies ϕ = 0. The last two assertions hold by Proposition 2.2. (cid:3) 3. The atomic diagonal operator Throughout this section, let L be a Dedekind complete Banach lattice. Denote by Lr(L) the Dedekind complete Riesz space of all regular operators on L, i.e., those operators that are linear combinations of positive operators. It is well known that Lr(L) becomes a Banach lattice algebra with respect to the regular norm defined by kT kr := kT k. The center Z(L) is the ideal in Lr(L) generated by the identity operator I, i.e., Z(L) = {T ∈ Lr(L) : T ≤ λI for some λ ≥ 0}. FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY3 7 If T ∈ Z(L), then the operator norm and the regular norm of T coincide. Since Z(L) is also a band in Lr(L), we have a band decomposition Lr(L) = Z(L) ⊕ Z(L)d. Let P be the band projection onto Z(L). By a result of Voigt [18], P is a contraction with respect to the operator norm. Schep [17] proved that the component P(T ) of a positive operator T in Z(L) is (2) P(T ) = inf( nXi=1 PiT Pi : 0 ≤ Pi ≤ I, P 2 i = Pi, Pi = I) . nXi=1 Let A be the band generated by all atoms in L, and let A ⊆ A be the maximal set of pairwise disjoint atoms of norm one. Given a ∈ A, we denote by Pa the band projection onto the band Ba. Let T be a positive operator on L. In [12] it is proved that the operator D(T ) = sup(Xa∈F PaT Pa : F is a finite subset of A) exists, since it is a supremum of an increasing net that is bounded from above, and it also satisfies 0 ≤ D(T ) ≤ T. If L is atomic (i.e., A = L), then D(T ) = P(T ). For a general L we have Proposition 3.1. Let T be a positive operator on L. If PA denotes the band projection onto the band A, then D(T ) = PAP(T ). Proof. Since PA ∈ Z(L), it commutes with P(T ), and so PAP(T ) = PAP(T )PA. Applying nXi=1 Pi = I) PA. Pi = I) = nXi=1 Qi = PA) . nXi=1 (PAPi)T (PiPA) : 0 ≤ Pi ≤ I, P 2 i = Pi, QiT Qi : 0 ≤ Qi ≤ PA, Q2 i = Qi, PiT Pi : 0 ≤ Pi ≤ I, P 2 i = Pi, Since 0 ≤ PA ≤ I, PA is order continuous, so that Schep's formula (2) we obtain PAP(T ) = PA inf( nXi=1 PAP(T ) = inf( nXi=1 = inf( nXi=1 Since A is an atomic Dedekind complete Banach lattice, PAP(T ) = D(T ). (cid:3) We use the equality from Proposition 3.1 to extend the operator D to the operator on the whole space Lr(L), so we define D(T ) := PAP(T ) for T ∈ Lr(L). This extension is called the atomic diagonal operator and the operator D(T ) on L is said to be the atomic diagonal of an operator T ∈ Lr(L). T =(cid:20) T1 T2 T3 T4 (cid:21) , T =(cid:20) T1 T3 T2 T4 (cid:21) and PA =(cid:20) I 0 0 0 (cid:21) . 8 ROMAN DRNOVSEK, MARKO KANDI ´C Proposition 3.2. The following assertions hold for the atomic diagonal operator D: (a) D is a band projection onto the band {T ∈ Lr(L) : T ≤ λPA for some λ ≥ 0} in Lr(L) that can be identified with the center Z(A); (b) kD(T )k ≤ kP(T )k ≤ kT k for all T ∈ Lr(L); (c) P(T ) = D(T ) for every positive compact operator T on L. Proof. For the proof of (a) just note that, with respect to the band decomposition L = A ⊕ Ad, an operator T ∈ Lr(L), its modulus T and the band projection PA have the block forms To show (b), use Voigt's theorem [18, Theorem 1.4] to obtain that kD(T )k = kPAP(T )k ≤ kP(T )k ≤ kT k. If PAd is the band projection on Ad, then the operator P(T ) − D(T ) = PAdP(T ) = PAdP(T )PAd ∈ Z(L) is dominated by a positive compact operator PAdT PAd, and so it is equal to 0 by [17, Corollary 1.7]. This proves (c). (cid:3) 4. The ideal-triangularizability We first extend [12, Theorem 3] by showing the implication that remained unproven in [12]. In the proof we will make use of the following extension of Ringrose's Theorem; see [4, Theorem 2.1]. Recall that if T is a power-compact operator (i.e., some power of T is a compact operator) on a Banach space X, then the algebraic multiplicity m(T, λ) of a non-zero complex number λ is the dimension of the subspace ker ((λ−T )k), where k is the smallest natural number such that ker ((λ − T )k) = ker ((λ − T )k+1). On the other hand, the geometric multiplicity of λ is the dimension of the subspace ker (λ − T ). We say that a chain C of closed subspaces of X is a complete chain if it contains arbitrary intersections and closed linear spans of its members. If a closed subspace M is in a complete chain C, then the predecessor M− of M in C is defined as the closed linear span of all proper subspaces of M belonging to C. Theorem 4.1. Let T be a power-compact operator on a Banach space X, and let C be a complete chain of closed subspaces invariant under T . Let C ′ be a subchain of C of all FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY4 9 subspaces M ∈ C such that M− 6= M. For each M ∈ C ′, define TM to be the quotient operator on M/M− induced by T . Then Moreover, for each non-zero complex number λ we have σ(T ) \ {0} = [M∈C ′ m(T, λ) = XM∈C ′ σ(TM) \ {0}. m(TM, λ). The following theorem completes [12, Theorem 3]. It extends [3, Theorem 3.8], where the implication (c) ⇒ (a) has been proved for a positive compact operator on a Banach lattice lp (1 < p < ∞). It is also a generalization of [14, Corollary 27], where positive trace-class operators acting on Lp(µ) (1 ≤ p < ∞) are considered. A diagonal entry of an operator T on a Banach lattice L is the number ϕa(T a), where ϕa is the linear functional associated to an atom a ∈ L+. Theorem 4.2. Let T be a positive power-compact operator on a Banach lattice L with order continuous norm. The following conditions are mutually equivalent: (a) T is ideal-triangularizable; (b) T − D(T ) is quasinilpotent; (c) The diagonal entries of T consists precisely of eigenvalues (except maybe zero) of the operator T repeated according to their algebraic multiplicities. Proof. It suffices to prove that (c) implies (a), since other implications were proved in [12]. Let C be a maximal chain of closed ideals invariant under the operator T . We will prove that, in fact, C is an ideal-triangularizing chain for T . Maximality of C implies that {0} and L are elements of C, and that C is a complete chain. Therefore, we only need to prove that for every J ∈ C the dimension of the quotient space J/J− is less than or equal to one. Maximality of C also implies that the induced operator TJ is ideal-irreducible on J/J−, and so it is not quasinilpotent by [9, Theorem 1.3] provided dim(J/J−) ≥ 2. Note that TJ can be identified with the compression of T to the band J ∩ J d −. Let B be the set of all ideals J in C with dim(J/J−) ≥ 2. Assume that B is not empty. Let λ be the largest number among the diagonal entries of the compressions of the operator T onto J ∩ J d −, where J runs over B. Pick any ideal J ∈ B such that λ is the diagonal entry of the compression of the operator T onto J ∩ J d −. We claim that λ = r(TJ ). Otherwise we would have λ < r(TJ), and so the number of appearances of 10 ROMAN DRNOVSEK, MARKO KANDI ´C r(TJ ) as a diagonal entry of the compression of T onto some one-dimensional quotient space (induced by an ideal in C) would be precisely the algebraic multiplicity of r(TJ ). In view of Theorem 4.1 this would imply that r(TJ) is not in the spectrum of TJ which is absurd. So, λ = r(TJ ) as claimed. Therefore, the spectral radius r(TJ ) appears on the diagonal of the compression of T onto J ∩ J d −. By [12, Corollary 1], this compression is ideal-reducible, contradicting the maximality of the chain C. This implies that B is an empty set, and so C is an ideal-triangularizing chain for the operator T . (cid:3) Now we turn our attention to multiplicative semigroups of ideal-triangularizable positive operators. We first recall [5, Theorem 4.5]. Theorem 4.3. Every semigroup of quasinilpotent positive compact operators on a Banach lattice is ideal-triangularizable. As a consequence we obtain the following result. Theorem 4.4. Let L be an atomless Banach lattice with order continuous norm. If S is a semigroup of ideal-triangularizable positive compact operators on L, then it is ideal- triangularizable. Proof. By [8, Proposition 4.5], every operator in S is quasinilpotent, and so Theorem 4.3 can be applied. (cid:3) Theorems 4.3 and 4.4 do not hold without the assumption that the operators are compact. Namely, in [10] there was constructed an irreducible semigroup of square- zero positive operators on Lp[0, 1) (1 ≤ p < ∞) in which any finite number of elements generate an ideal-triangularizable semigroup. So, one may seek for extensions of Theorem 4.4 by relaxing the assumption that L is atomless. It will turn out that in this case the atomic diagonal operator plays important role. Let L be a Banach lattice with order continuous norm, and let S be a semigroup of positive operators on L. If S is ideal-triangularizable, then D(ST ) = D(S)D(T ) for every pair {S, T } ⊆ S, as was shown in the proof of [12, Proposition 3]. It follows that D(ST ) = D(T S) or equivalently D(ST − T S) = 0 for every pair {S, T } ⊆ S. The following main theorem of the paper treats the question when the converse implication holds. Note also that it is a generalization of Theorem 4.4. FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY5 11 Theorem 4.5. Let S be a semigroup of ideal-triangularizable positive compact operators on a Banach lattice L with order continuous norm such that D(ST ) = D(T S) for every pair {S, T } ⊆ S. Then the semigroup S is ideal-triangularizable. Proof. With no loss of generality we may assume that S is closed under positive scalar multiplication. Let us prove that the assumptions are satisfied for the closure of S, so that we may also assume that S is a closed set. To this end, let a sequence {Sk}k ⊆ S converge to a positive compact operator S. By [17, Theorem 1.6], D(Sk) is a positive compact operator for each k. Since the diagonal operator D is continuous by Theorem 3.2, the sequence {D(Sk)}k converges to the operator D(S), so that the sequence {Sk − D(Sk)}k converges to the operator S − D(S). Since the operators Sk − D(Sk) are all quasinilpotent by Theorem 4.2, the compact operator S −D(S) is quasinilpotent as well by [15, Corollary 7.2.11.]. Thus, Theorem 4.2 implies that S is ideal-triangularizable. Since D(ST ) = limk→∞ D(SkT ) = limk→∞ D(T Sk) = D(T S) for every T ∈ S, we conclude (because of symmetry) that D(ST ) = D(T S) for every pair {S, T } in the closure of S. Therefore, we have shown that there is no loss of generality in assuming that S is a closed set. We first show that S is ideal-reducible. Assume otherwise. It follows from Theorem 4.3 that any semigroup consisting of quasinilpotent positive compact operators and multiples of the identity operator is ideal-triangularizable. Therefore, there exists an operator A ∈ S such that r(A) = 1 and A is not the identity operator I. Since A is ideal-triangularizable, [12, Proposition 4] implies that its spectrum σ(A) is contained in [0, 1]. We distinguish two cases: Case 1: The geometric multiplicity of the eigenvalue 1 ∈ σ(A) is equal to its algebraic multiplicity. Then the sequence {Ak}k converges to a nonnegative idempotent E ∈ S of finite rank satisfying E 6= I. Since E is ideal-triangularizable, it follows from Proposition 2.3 that there exist bands B1, B2 and B3 such that, with respect to the decomposition L = B1 ⊕ B2 ⊕ B3, E has the block-triangular form E = 0  , 0 X XY 0 Y 0 I 0 where only one of the bands B1 and B3 may be equal to zero, and the dimension n = dim B2 is finite, so that B2 is atomic and isomorphic to Rn by [16, Corollary 1, p.70]. Assume first that B1 = {0} and Y = 0, so that E has the form E =(cid:20) I 0 0 0 (cid:21) . 12 Let ROMAN DRNOVSEK, MARKO KANDI ´C S =(cid:20) S1 S2 S3 S4 (cid:21) ∈ S and T =(cid:20) T1 T2 T3 T4 (cid:21) ∈ S be arbitrary operators in S. The operators ES · T E and T E · ES are both in S, and they have the same diagonal. From ES · T E =(cid:20) S1T1 + S2T3 0 0 (cid:21) 0 and T E · ES =(cid:20) T1S1 T1S2 T3S1 T3S2 (cid:21) we conclude that Since D(S1T1) + D(S2T3) = D(T1S1). tr(D(S1T1)) = tr(S1T1) = tr(T1S1) = tr(D(T1S1)), we obtain that tr(D(S2T3)) = 0, so that D(S2T3) = 0. Since S is ideal-irreducible, we can choose an atom a ∈ B2 and an operator T ∈ S such that f = T3a is a nonzero positive vector in B3. If ϕa is the linear functional associated to the atom a, then ϕa(Sf ) = ϕa(S2T3a) + ϕa(S4T3a) = 0, as D(S2T3) = 0 and S4T3a ⊥ a. This is a contradiction with Proposition 1.1. The case when B3 = {0} and X = 0 can be handled similarly. Hence, it remains to consider the case when X and Y are not both zero. We consider only the case X 6= 0, since the case Y 6= 0 is similar. Let be an arbitrary operator. Then S1 S2 S3 S4 S5 S6 S7 S8 S9  ∈ S S = ES = SE = XS4 + XY S7 ∗ 0 ∗ 0 0 S4X + S5 0 ∗ 0 ∗ ∗ S5 + Y S8 ∗ 0  S7XY + S8Y  . ∗ ∗ and Since the operator ES is ideal-triangularizable, its (1, 1) block XS4 + XY S7 is ideal- triangularizable as well, and so is the operator XS4. The equality D(SE) = D(ES) implies that D(XS4) = 0, and so the operator XS4 is quasinilpotent by Theorem 4.2. Since σ(XS4)\{0} = σ(S4X)\{0}, S4X is a nilpotent operator on a finite-dimensional Banach lattice B2. Since the operator SE is ideal-triangularizable, its (2, 2) block S4X +S5 is ideal-triangularizable as well, and so is S4X. It follows that D(S4X) = 0. Since X 6= 0, FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY6 13 there is an atom a ∈ B2 such that f = Xa is a nonzero vector in B1. The vectors S1f and S7f are disjoint with the atom a, and so ϕa(Sf ) = ϕa(S4f ) = ϕa(S4Xa) = 0, as D(S4X) = 0. This is again a contradiction with Proposition 1.1. Case 2: The geometric multiplicity of the eigenvalue 1 ∈ σ(A) is smaller than its algebraic multiplicity. Then, by the Riesz decomposition theorem, the Banach space L can be decomposed into a direct sum of a finite-dimensional subspace L1 and a (possibly zero) closed subspace L2 such that A has the block-diagonal form A =(cid:20) I + N 0 C (cid:21) , 0 where N is a nilpotent operator on L1 of the nilpotency index k ≥ 2, and r(C) < 1 provided L2 is nonzero. Now lim m→∞ Am k−1(cid:1) =(cid:20) N k−1 0 (cid:0) m 0 0 (cid:21) ∈ S. This proves that the semigroup S contains a square-zero operator M 6= 0. Let N (M) and R(M) denote the absolute kernel and the range ideal of the operator M, respectively. We claim that R(M) ⊆ N (M). If y is a vector in R(M), then there exist positive vectors x1, . . . , xn ∈ E and positive scalars λ1, . . . , λn such that y ≤ Then we have λjMxj. 0 ≤ My ≤ λjM 2xj = 0, nXj=1 nXj=1 which proves the claim. Now, if z ∈ N (M)d then Mz ∈ R(M) ⊆ N (M). This shows that, with respect to the band decomposition L = N (M) ⊕ N (M)d, we have Let S be an arbitrary operator in S, and let 0 M =(cid:20) 0 X 0 (cid:21) . S =(cid:20) S1 S2 S3 S4 (cid:21) be its block operator matrix with respect to the band decomposition L = N (M)⊕N (M)d. Then MS =(cid:20) XS3 XS4 0 0 (cid:21) and SM =(cid:20) 0 S1X 0 S3X (cid:21) . 14 ROMAN DRNOVSEK, MARKO KANDI ´C Since D(MS) = D(SM), we have D(XS3) = 0 and D(S3X) = 0. Since MS and SM are ideal-triangularizable operators, the operators XS3 and S3X are also ideal- triangularizable. Theorem 4.2 implies that XS3 and S3X are both quasinilpotent, so that MS and SM are quasinilpotent as well. Let J be the semigroup ideal in S generated by the operator M. Pick any S ∈ J . Then S is a product of operators from S, and the operator M appears in this product at least once. If M appears either at the beginning or at the end of this product, then S is quasinilpotent by the observation above. Otherwise there exist S1 and S2 in S such that S = S1MS2. By the well known equality for the spectral radius, we have r(S) = r(S1MS2) = r(MS2S1) = 0. This implies that J consists of quasinilpotent compact operators, so that it is ideal-triangularizable by Theorem 4.3. This is again a contradiction with Proposition 1.1. We have shown that the semigroup S is ideal-reducible. We will finish the proof by applying the Ideal-triangularization lemma. The property D(ST ) = D(T S) is inherited by ideal-quotients. Indeed, let J be a closed ideal in L that is invariant under the semigroup S. Let SJ denote the restriction of the operator S to J, and let PJ be the band projection onto J. Then we have Pa(SJ)(T J )Pa = Xa∈A∩J Pa(ST )JPa = D(SJT J) = Xa∈A∩J = Xa∈A∩J PaPJST PJ Pa = PJ Xa∈A PaST Pa! PJ = PJ D(ST )PJ. This implies that D(SJT J) = PJ D(ST )PJ = PJ D(T S)PJ = D(T JSJ). This proves that the property of zero diagonals of commutators of operators from S is inherited to restrictions on invariant closed ideals. Since the induced operator TJ of the operator T on the quotient Banach lattice E/J can be identified with the compression of the operator T to J d, we can (similarly as above) prove that the property that the diagonal of every commutator of operators from S is zero is inherited by induced operators on the quotient Banach lattice E/J. Since ideal-triangularizability is inherited by ideal-quotients by [11, Proposition 2.3], we finish the proof by applying the Ideal-triangularization lemma. (cid:3) In the finite-dimensional case Theorem 4.5 can be stated as follows. FROM LOCAL TO GLOBAL IDEAL-TRIANGULARIZABILITY7 15 Corollary 4.6. Let S be a semigroup of nonnegative n × n matrices. Suppose that for every matrix S ∈ S there exists a permutation matrix PS (depending on S) such that the matrix PSSP −1 there exists a permutation matrix P such that every matrix in the semigroup P SP −1 is is upper triangular. If D(ST ) = D(T S) for every pair S, T ∈ S, then S upper triangular. The following example which was already introduced in [7] shows that Theorem 4.5 and Corollary 4.6 do not hold for collections of ideal-triangularizable nonnegative matrices with zero diagonals of commutators. Example 4.7. Let e1, e2, . . ., en be the standard basis vectors of IRn, where n ≥ 3. Define ideal-triangularizable nilpotent matrices by Ai = eieT i+1 for i = 1, 2, . . . , n − 1, and An = eneT 1 . Then the collection {A1, A2, . . . , An} has the property that D(AiAj) = 0 for all 1 ≤ i, j ≤ n. We claim that the collection is not ideal-triangularizable. Assume the contrary. Then the sum S = A1 + A2 + . . . + An is ideal-triangularizable. Since all the diagonal entries of S are zero, S must be nilpotent which contradicts the fact that Sn = I. The following example shows that Theorem 4.5 and Corollary 4.6 do not hold without the assumption that operators are positive. Example 4.8. Let A = 0 0 0 0 −1 0 0 0 0 0 0 1 −1 1 0 0  and B = 0 0 0 0 0 0 0 0 1 1 0 0 1 1 0 0  . Then A2 = B2 = AB = BA = 0, so that S = {0, A, B} is a semigroup of ideal- triangularizable matrices such that D(S) = 0 for all S ∈ S. However, S is not ideal- triangularizable, as the diagonal of the matrix A + B = is zero, but the matrix is not nilpotent. 0 0 1 1 0 0 1 1 1 1 0 0 1 1 0 0  Acknowledgments. This work was supported in part by the Slovenian Research Agency. 16 ROMAN DRNOVSEK, MARKO KANDI ´C References [1] Y. A. Abramovich, C. D. Aliprantis, An Invitation to Operator Theory. American Mathematical Society, Providence (2002). [2] C. D. Aliprantis, O. Burkinshaw, Positive operators, Reprint of the 1985 original, Springer, Dor- drecht (2006). [3] J. Bernik, L. W. Marcoux, H. Radjavi, Spectral conditions and band reducibility of operators, J. Lond. Math. Soc. (2) 86 (2012), no. 1, 214 -- 234. [4] R. Drnovsek, Once more on positive commutators, Studia Math. 211 (2012), no. 3, 241 -- 245. [5] R. Drnovsek, Common invariant subspaces for collections of operators, Integral Equat. Oper. Th. 39 (2001), 253 -- 266. [6] R. Drnovsek, Triangularizing semigroups of positive operators on an atomic normed Riesz spaces, Proc. Edin. Math. Soc. 43 (2000), 43 -- 55. [7] R. Drnovsek, M. Jesenko, M. Kandi´c, Positive commutators and collections of operators, Oper. Matrices 6 (3) (2012), 535 -- 542. [8] R. Drnovsek, M. Kandi´c, Ideal-triangularizability of semigroups of positive operators, Integral Equat. Oper. Th. 64 (4) (2009), 539 -- 552. [9] R. Drnovsek, M. Kandi´c, More on positive commutators, J. Math. Anal. Appl. 373 (2011), 580 -- 584. [10] R. Drnovsek, D. Kokol-Bukovsek, L. Livshits, G. MacDonald, M. Omladic, H. Radjavi, An irre- ducible semigroup of non-negative square-zero operators, Integral Equat. Oper. Th. 42 (2002), no. 4, 449 -- 460. [11] M. Kandi´c, Ideal-triangularizability of upward directed sets of positive operators, Ann. Funct. Anal. 2 (1) (2011), 206 -- 219. [12] M. Kandi´c, Multiplicative coordinate functionals and ideal-triangularizability, Positivity, doi:10.1007/s11117-013-0222-z. [13] W. A. J. Luxemburg, A. C. Zaanen, Riesz spaces I, North-Holland, Amsterdam (1971). [14] G. MacDonald, H. Radjavi, Standard triangularization of semigroups of non-negative operators, J. Funct. Anal. 219 (2005), 161 -- 176. [15] H. Radjavi, P. Rosenthal, Simultaneous Triangularization, Springer-Verlag, New York, 2000. [16] H. H. Schaefer, Banach lattices, Springer-Verlag, Berlin-Heidelberg-New York, 1974. [17] A. R. Schep, Positive diagonal and triangular operators, J. Oper. Theory 3 (1980), 165 -- 178. [18] J. Voigt, The projection onto the center of operators in a Banach lattice, Math. Z. 199 (1988), no. 1, 115 -- 117. [19] Y. Zhong, Functional positivity and invariant subspaces of semigroups of operators, Houston J. Math. 19 (1993), 239 -- 262. Roman Drnovsek, Marko Kandi´c : Faculty of Mathematics and Physics University of Ljubljana Jadranska 19 1000 Ljubljana Slovenia E-mails : [email protected], [email protected]
1109.4793
1
1109
2011-09-22T12:59:46
Structure constants of the Weyl calculus
[ "math.FA" ]
We find some explicit bounds on the ${\mathcal L}(L^2)$-norm of pseudo-differential operators with symbols defined by a metric on the phase space. In particular, we prove that this norm depends only on the "structure constants" of the metric and a fixed semi-norm of the symbol. Analogous statements are made for the Fefferman-Phong inequality.
math.FA
math
STRUCTURE CONSTANTS OF THE WEYL CALCULUS WEN DENG Abstract. We find some explicit bounds on the L(L2)-norm of pseudo-differential operators with symbols defined by a metric on the phase space. In particular, we prove that this norm depends only on the "structure constants" of the metric and a fixed semi-norm of the symbol. Analogous statements are made for the Fefferman-Phong inequality. 1. Introduction The class of symbols Sm 1,0 consists of smooth functions a defined on the phase space Rn × Rn such that for all multi-indices α, β, (1.1) (∂α ξ ∂β x a)(x, ξ) ≤ Cα,β(1 + ξ)m−α. 1,0. We have Property A. If a is in S0 The best constants Cα,β in (1.1) are called the semi-norms of the symbol a in the Fr´echet space Sm 1,0, then a(x, D) defines a bounded operator on L2(Rn). One might ask some very natural questions: the operator norm ka(x, D)kL(L2(Rn)) is bounded by which constant? Is it a semi-norm of the symbol a? If yes, then which semi-norm? Questions of the same type might be asked for the constant C in the following inequality: Property B (Fefferman-Phong inequality). If a is a non-negative symbol belonging to S2 1,0, then there exists C > 0 such that, for all u ∈ S(Rn), (1.2) Reha(x, D)u, uiL2(Rn) + Ckuk2 L2(Rn) ≥ 0. We can pose similar questions in many other examples of classes of symbols, such as the semi-classical symbols, Shubin's class, etc. As a particular example, the class Σm, defined as the set of smooth functions a on Rn × Rn × R+ such that for all multi-indices α, β, (1.3) ∀x, ξ ∈ Rn, τ ∈ R+, (∂α ξ ∂β x a)(x, ξ, τ ) ≤ Cα,β(1 + ξ + τ )m−α, is useful for Carleman estimates. One would like to check the Property A and Property B independent of the parameter τ . Several authors like Bony [1], Boulkhemair [3], Lerner-Morimoto [7], have al- ready considered these questions and they were able to identify the constants. The constants in Properties A, B are always a constant Cn times a semi-norm of the symbol, whose order depends only on the dimension n. Although the problem is well-understood for a single class of pseudo-differential calculus, including the class S(m, g) developed by Hormander, we want to address a more general and useful Date: July 9, 2018. 2000 Mathematics Subject Classification. Key words and phrases. Weyl calculus, phase space, Fefferman-Phong inequality. 1 2 WEN DENG question, having in mind the class Σm depending on the non-compact parameter τ ≥ 0 which is defined in (1.3) and is useful for Carleman estimates. In this paper, we consider the Weyl quantization for pseudodifferential operators and we choose the framework with a metric g on the phase space. The metric g is assumed to be admissible, that is slowly varying, satisfying the uncertainty principle and is temperate (see Definition 2.1, 2.6 below). The so-called structure constants of g are closely related to these properties. We can define very general classes of symbols S(m, g) attached to the metric g and a g-admissible weight m (see Definition 2.3) and we have an effective symbolic calculus. The following results are classical: (see [5, chapter 18], [6, chapter 2]) (1.4) (1.5) L2-boundedness: Fefferman-Phong: a ∈ S(1, g) =⇒ kawkL(L2(Rn)) ≤ C, a ∈ S(λ2 g, g), a ≥ 0 =⇒ aw + C ≥ 0. The question that we would like to address is the following: what happens if we change the metric g but keep the same structure constants? We intend to show that the constants involved in (1.4), (1.5) depend only on the structure constants of the metric g and a fixed semi-norm of a. Since it may happen that the metric g depends on a non-compact parameter with uniform structure the class Σm), this fact is useful explicitly or implicitly in many constants (e.g. examples where these metrics are used and it seems useful to rely on a more stable argument than referring to "inspection of the proofs". Remark. An abstract functional analysis argument does not seem to work. Our method is to follow the proofs, by carefully computing all the constants. 2. Metric on the phase space In this section, we introduce the definitions of the admissible metric and exhibit its properties. We use the Weyl quantization which associates to a symbol a the operator aw defined by (2.1) (awu)(x) = ZZ e2iπ(x−y)·ξa( x + y 2 , ξ)u(y)dydξ. Consider the symplectic space R2n equipped with the symplectic form σ = Pn dxj. Given a positive-definite quadratic form Γ on R2n, we define j=1 dξj∧ (2.2) Γσ(T ) = sup Γ(Y )=1 σ(T, Y )2, which is also a positive-definite quadratic form. Let g be a measurable map from R2n into the cone of positive-definite quadratic forms on R2n, i.e. for each X ∈ R2n, gX is a positive definite quadratic form on R2n. Definition 2.1 (Slowly varying metric). We say that g is a slowly varying metric on R2n, if there exists C0 ≥ 1 such that for all X, Y, T ∈ R2n, (2.3) gX(X − Y ) ≤ C −1 0 =⇒ C −1 0 ≤ gX(T ) gY (T ) ≤ C0. Definition 2.2 (Slowly varying weight). Let g be a slowly varying metric on R2n. A function m : R2n → (0, +∞) is called a g-slowly varying weight if there exists STRUCTURE CONSTANTS OF THE WEYL CALCULUS 3 µm ≥ 1 such that for all X, Y ∈ R2n, (2.4) gX(Y − X) ≤ µ−1 m =⇒ µ−1 m ≤ m(X) m(Y ) ≤ µm. Definition 2.3 (Class of symbols). Let g be a slowly varying metric on R2n and m be a g-slowly varying weight. The class of symbols S(m, g) is defined as the subset of functions a ∈ C ∞(R2n) satisfying that for all k ∈ N, there exists Ck > 0 such that for all X, T1, · · · , Tk ∈ R2n, a(k)(X)(T1, · · · , Tk) ≤ Ckm(X) Y1≤j≤k gX(Tj)1/2. For a ∈ S(m, g), l ∈ N, we denote (2.5) kak(l) S(m,g) = max 0≤k≤l sup a(k)(X)(T1, · · · , Tk)m(X)−1. X,Tj∈R2n gX (Tj )=1 The space S(m, g) equipped with the countable family of semi-norms (k · k(l) is a Fr´echet space. S(m,g))l∈N For a slowly varying metric g on the phase space R2n, we can introduce some partition of unity related to g. Define the g-ball near X ∈ R2n (2.6) UX,r = {Y, gX(X − Y ) ≤ r2}, we have the following theorem, which is Theorem 2.2.7 in [6]. Theorem 2.4 (Partition of unity). Let g be a slowly varying metric on R2n and C0 > 0 given in (2.3). Then for all r ∈ (0, C −1/2 ], there exists a family (ϕY )Y ∈R2n 0 of smooth functions supported in UY,r such that (2.7) (2.8) ∀k ∈ N, sup Y ∈R2n kϕY k(k) S(1,g) ≤ C(k, r, n, C0), ∀X ∈ R2n, ZR2n ϕY (X)gY 1/2dY = 1, where C(k, r, n, C0) is a positive constant depending only on k, r, n, C0 and gY is the determinant of gY with respect to the standard Euclidean norm. Proof. As in the proof of Theorem 2.2.7 in [6], let χ0 ∈ C ∞ such that χ0(t) = 1 on t ≤ 1/2, χ0(t) = 0 on t ≥ 1. Define for r ∈ (0, C −1/2 0 (R+; [0, 1]) non-increasing ], 0 Since ωY (X) is supported in UY,r and χ0 is non-increasing, by (2.3) we have gY 1/2dY. ω(X, r) = ZR2n χ0(cid:0)r−2gY (X − Y )(cid:1) } {z gX1/2dY = ZR2n χ0(cid:0)r−2C0gX(X − Y )(cid:1)C −n =ωY (X) 0 ω(X, r) ≥ ZR2n χ0(Z2)dZC −2n 0 r2n, and an estimate from above of the same type, i.e. there exists a positive constant C1 = C1(r, n, C0) such that C −1 1 ≤ ω(X, r) ≤ C1. 4 WEN DENG Now let us check the derivatives of ωY (X). Using the notation h, iY the inner-product associated to gY , we have ω′ Y (X)T = χ′ 0(cid:0)r−2gY (X − Y )(cid:1)r−2hX − Y, T iY , and by induction, for k ≥ 1, T ∈ R2n, ω(k) Y (X)T k is a finite sum of terms of type (2.9) cp,kχ(p) 0 (cid:0)r−2gY (X − Y )(cid:1)r−2phX − Y, T i2p−k Y gY (T )k−p, where cp,k is a constant depending only on p, k and p ∈ [k/2, k] ∩ N. Since the support of χ(p) 0 , the term (2.9) can be bounded 0 from above by is included in [0, 1] and r2 ≤ C −1 cp,kkχ(p) 0 kL∞r−2p(r2)(2p−k)/2C k/2 0 gX(T )k/2, so that for all k ≥ 1, ω(k) S(1, g) and moreover, Y (X)T k ≤ C(k, r, C0)gX(T )k/2. This implies that ωY is in (2.10) ∀k ∈ N, sup Y ∈R2n kωY k(k) S(1,g) ≤ C(k, r, C0). Now we choose a non-negative function χ1 ∈ C ∞ t ≤ 1, then 0 (R+; [0, 1]) such that χ1(t) = 1 on ω(k)(X, r)T k = (cid:12)(cid:12)ZR2n ≤ sup Y ∈R2n ω(k) Y (X)T kχ1(cid:0)r−2gY (X − Y )(cid:1)gY 1/2dY(cid:12)(cid:12) S(1,g)gX(T )k/2ZR2n kωY k(k) χ1(cid:0)r−2gY (X − Y )(cid:1)gY 1/2dY ≤ C(k, r, n, C0)gX(T )k/2, which implies that ω(·, r) is a symbol in S(1, g) with kω(·, r)k(k) Since ω is bounded from below by C −1 S(1,g) ≤ C ′(k, r, n, C0). 1 , the function ω(·, r)−1 is also in S(1, g) and (2.11) We define kω(·, r)−1k(k) S(1,g) ≤ C ′′(k, r, n, C0). ϕY (X) = ωY (X)ω(X, r)−1, then the estimate (2.7) follows from (2.10), (2.11) and moreover, the family (ϕY )Y ∈R2n satisfies the requirements of Theorem 2.4. (cid:3) A direct consequence of Theorem 2.4 is the following. Proposition 2.5. Let g be a slowly varying metric on R2n and m be a g-slowly varying weight. Let C0, µm be given in (2.3), (2.4) respectively. Let a be a symbol in S(m, g). Then for all 0 < r ≤ min(C −1/2 , µ−1/2 m ), 0 a(X) = ZR2n aY (X)gY 1/2dY, where aY has support included in UY,r and (2.12) ∀k ∈ N, sup Y ∈R2n kaY k(k) S(m(Y ),gY ) ≤ C(k, r, C0, n, µm)kak(k) S(m,g). STRUCTURE CONSTANTS OF THE WEYL CALCULUS 5 Proof. Define aY (X) = a(X)ϕY (X). Since ϕY is supported in UY,r, we have, for k ≥ 0, X ∈ UY,r, T ∈ R2n, a(k) Y (X)T k = (cid:12)(cid:12)(cid:12) X0≤l≤k ≤ X0≤l≤k l(cid:19)a(l)(X)T l · ϕ(k−l) (cid:18)k Y (X)T k−l(cid:12)(cid:12)(cid:12) ck,lkak(l) S(m,g)m(X)gX(T )l/2kϕY k(k−l) S(1,g)gX(T )(k−l)/2 ≤ C(k)kak(k) ≤ C(k)µmC k/2 S(m,g)kϕY k(k) 0 kak(k) S(1,g)m(X)gX(T )k/2 S(m,g)kϕY k(k) S(1,g)m(Y )gY (T )k/2, which completes the proof. (cid:3) For two positive-definite quadratic forms Γ1, Γ2 on R2n, the harmonic mean Γ1∧Γ2 is defined by (2.13) 2 )−1, which is also a positive-definite quadratic form on R2n. Γ1 ∧ Γ2 = 2(Γ−1 1 + Γ−1 Definition 2.6 (Admissible metric). We say that g is an admissible metric on R2n if g is slowly varying (see Definition 2.1) and there exist C ′ 0 > 0, N0 ∈ N such that for all X, Y, T ∈ R2n, (2.14) uncertainty principle (2.15) temperance X(T ), gX(T ) ≤ gσ gX(T ) ≤ C ′ where gσ is given by (2.2) and ∧ given by (2.13). 0gY (T )(cid:0)1 + (gσ X ∧ gσ Y )(X − Y )(cid:1)N0, We may suppose C ′ 0 = C0 in the sequel, where C0 is given in (2.3). Then the constants (C0, N0) appearing in (2.3), (2.15) are called the structure constants of the metric g. Definition 2.7 (Admissible weight). Suppose that g is an admissible metric on R2n. A function m : R2n → (0, +∞) is called a g-admissible weight if m is a g- slowly varying weight (see Definition 2.2) and there exist µm > 0, νm ∈ N such that for all X, Y ∈ R2n, (2.16) m(X) ≤ µmm(Y )(cid:0)1 + (gσ X ∧ gσ Y )(X − Y )(cid:1)νm. The constants (µm, νm) appearing in (2.4), (2.16) are called the structure constants of the g-admissible weight m. Let g be an admissible metric on R2n. We define for X ∈ R2n, (2.17) λg(X) = inf T 6=0(cid:16)gσ X (T ) gX (T )(cid:17)1/2 . Then the uncertainty principle (2.14) can be expressed by gX ≤ λg(X)−2gσ X, λg(X) ≥ 1. Lemma 2.8 ([6, Remark 2.2.17]). For any s ∈ R, λs with structure constants (µλs constants of the metric g (C0, N0). g is an admissible weight, g) in (2.4), (2.16) depending only on the structure g , νλs 6 WEN DENG Proof. We first verify that λs T ∈ R2n, we have g is a g-slowly varying weight. For gX(X − Y ) ≤ C −1 0 , C −1 0 gX(T ) ≤ gY (T ) ≤ C0gX(T ), C −1 0 gσ X(T ) ≤ gσ Y (T ) ≤ C0gσ X(T ), which implies gσ Y (T ) gY (T ) Taking the infimum with respect to T , we get gσ X(T ) gX(T ) C −2 ≤ 0 ≤ C 2 0 gσ X(T ) gX(T ) . C −2 0 λg(X)2 ≤ λg(Y )2 ≤ C 2 0 λg(X)2, so that λg is g-slowly varying with µλg = C0 and so is λs check that λs g is temperate. We have for all X, Y, T ∈ R2n, g with µλs g = C s 0 . Next we which gives gX(T ) ≥ C −1 X(T ) ≤ C0gσ gσ X ∧ gσ X ∧ gσ 0 gY (T )(cid:0)1 + (gσ Y (T )(cid:0)1 + (gσ 0 λg(Y )2(cid:0)1 + (gσ Y )(X − Y )(cid:1)−N0, Y )(X − Y )(cid:1)N0, Y )(X − Y )(cid:1)2N0. g with νλs λg(X)2 ≤ C 2 X ∧ gσ Thus λg is temperate with νλg = N0 and so is λs the proof of Lemma 2.8. g = sN0. This completes (cid:3) The composition a♯b of two symbols is defined by awbw = (a♯b)w and we have, with the notations [X, Y ] = σ(X, Y ), D = (2iπ)−1∂, (a♯b)(X) = 22nZZR2n×R2n a(Y )b(Z)e−4iπ[X−Y,X−Z]dY dZ, For a ∈ S(m1, g), b ∈ S(m2, g), we have the asymptotic expansion (a♯b)(X) = exp(cid:0)iπ[DY , DZ](cid:1)(cid:0)a(Y )b(Z)(cid:1)Y =Z=X. a♯b(x, ξ) = X0≤k<p wk(a, b)(x, ξ) + rp(a, b)(x, ξ), (2.18) (2.19) (2.20) (2.21) (2.22) (2.23) with wk(a, b) = 2−k Xα+β=k (−1)β α!β! Dα ξ ∂β x a Dβ ξ ∂α x b ∈ S(m1m2λ−k g , g), rp(a, b)(X) = Rp(cid:0)a(X) ⊗ b(Y )(cid:1)Y =X ∈ S(m1m2λ−p g , g), Rp = Z 1 0 (1 − θ)p−1 (p − 1)! exp θ 4iπ [∂X , ∂Y ]dθ(cid:16) 1 4iπ [∂X , ∂Y ](cid:17)p Notice w1(a, b) = 1 asymptotic (2.20) at p = 2 is 4iπ {a, b}, where {, } denotes the Poisson bracket, so that the (2.24) a♯b = ab + 1 4iπ {a, b} + r2(a, b). STRUCTURE CONSTANTS OF THE WEYL CALCULUS 7 Definition 2.9 (The main distance function). Let g be an admissible metric on R2n. Define the main distance function, for r > 0, X, Y ∈ R2n, (2.25) δr(X, Y ) = 1 + (gσ X ∧ gσ Y )(UX,r − UY,r), where UX,r is given in (2.6) and g(U − V ) = inf X∈U,Y ∈V g(X − Y ). Lemma 2.10 ([6, Lemma 2.2.24], Integrability of δr). Let g be an admissible met- ric with structure constants (C0, N0). Then there exist positive constants N1 = N1(n, C0, N0), C = C(n, C0, N0) such that for all r ∈ (0, C −1/2 ], 0 (2.26) X∈R2nZR2n sup δr(X, Y )−N1gY 1/2dY ≤ C < +∞, Proof. Suppose r ≤ C −1/2 Y ′ ∈ UY,r, T ∈ R2n, we have 0 . Using the slowness and temperance of g, for X ′ ∈ UX,r, (gσ X ∧ gσ Y )(T ) ≥ C −1 0 (gσ X ′ ∧ gσ Y ′)(T ) ≥ C −2 ≥ C −3 0 gσ 0 gσ ≥ C −3−N0 0 X ′(T )(cid:0)1 + (gσ X(T )(cid:0)1 + C0(gσ X (T )(cid:0)1 + (gσ X ′ ∧ gσ X ∧ gσ X ∧ gσ Y ′)(X ′ − Y ′)(cid:1)−N0 Y )(X ′ − Y ′)(cid:1)−N0 Y )(X ′ − Y ′)(cid:1)−N0. gσ Taking the infimum in X ′ ∈ UX,r, Y ′ ∈ UY,r, we get (2.27) We have also X(T ) ≤ C 3+N0 gσ 0 δr(X, Y )N0(gσ X ∧ gσ Y )(T ). gX(T ) gY (T ) ≤ C 2 0 gX ′(T ) gY ′(T ) ≤ C 3 ≤ C 3 X ′ ∧ gσ 0(cid:0)1 + (gσ 0(cid:0)1 + C0(gσ (cid:0)1 + (gσ 0 Y ′)(X ′ − Y ′)(cid:1)N0 Y )(X ′ − Y ′)(cid:1)N0 Y )(X ′ − Y ′)(cid:1)N0. X ∧ gσ X ∧ gσ ≤ C 3+N0 By taking the infimum in X ′, Y ′, we get the following inequality (2.28) Then gX(T ) gY (T ) ≤ C 3+N0 0 δr(X, Y )N0. 1+gX(X − Y ) ≤ 1 + 3gX(X − X ′) + 3gX(X ′ − Y ′) + 3gX(Y ′ − Y ) 0 ≤ 3C 3+N0 ≤ 3C 3+N0 ≤ 9C 6+2N0 0 0 δr(X, Y )N0(cid:0)1 + gX(X − X ′) + gX(X ′ − Y ′) + gY (Y ′ − Y )(cid:1) by (2.28) δr(X, Y )N0(cid:0)1 + 2r2 + gσ δr(X, Y )2N0(cid:0)1 + (gσ X(X ′ − Y ′)(cid:1) Y )(X ′ − Y ′)(cid:1) by (2.27), X ∧ gσ so that 1 + gX(X − Y ) ≤ 9C 6+2N0 0 δr(X, Y )2N0+1. In the other hand, we have gY 1/2 gX1/2 ≤ C n(3+N0) 0 δr(X, Y )nN0, 8 WEN DENG so that for N1 = nN0 + (n + 1)(2N0 + 1) > 0, ZR2n δr(X, Y )−N1gY 1/2dY ≤ C(n, C0, N0)ZR2n δr(X, Y )−N1+nN0gX1/2dY ≤ C ′(n, C0, N0)ZR2n(cid:0)1 + gX (X − Y )(cid:1)−(n+1) = C ′(n, C0, N0)ZR2n (1 + Z2)−(n+1)dZ < +∞. gX1/2dY The proof of the lemma is complete. (cid:3) 3. L2-boundedness In this section, we prove the L2-boundedness of pseudo-differential operators with symbol in S(1, g) and make precise the operator norms. 3.1. The constant metric case. Proposition 3.1. Suppose that g is a positive-definite quadratic form (constant metric) on R2n with g ≤ gσ. Then there exists a constant C(n) > 0 depending only on the dimension n such that for all a ∈ S(1, g), kawkL(L2(Rn)) ≤ C(n)kak(2n+1) S(1,g) . Proof. Since g is a constant metric, according to Lemma 4.4.25 in [6], there exist symplectic coordinates (x, ξ) such that g = X1≤j≤n λ−1 j (dxj2 + dξj2), gσ = X1≤j≤n λj(dxj2 + dξj2), with λj > 0. g ≤ gσ is expressed as min 1≤j≤n λj ≥ 1. As a result, we have g ≤ dx2 + dξ2 := Γ0, which implies S(1, g) ⊂ S(1, Γ0) and for all a ∈ S(1, g), (3.1) ∀l ∈ N, kak(l) S(1,Γ0) ≤ kak(l) S(1,g). By Theorem 1.1.4 in [6] and aw = (J 1/2a)(x, D), where J t is introduced in Lemma 4.1.2 in [6], we obtain that kawkL(L2(Rn)) ≤ C(n)kak(2n+1) S(1,Γ0), where C(n) depends only on n. Together with (3.1), we complete the proof of the proposition. (cid:3) 3.2. The general case. Theorem 3.2. Let g be an admissible metric on R2n with structure constants (C0, N0) (see Definition 2.6). Then there exist C = C(n, C0, N0) > 0 and l = l(n, C0, N0) ∈ N such that for all a ∈ S(1, g) (see Definition 2.3), kawkL(L2(Rn)) ≤ Ckak(l) S(1,g). STRUCTURE CONSTANTS OF THE WEYL CALCULUS 9 Proof. Using the partition in Proposition 2.5, we write aw = ZR2n aw Y gY 1/2dY, where aY is supported in UY,r and satisfies (2.12). By Proposition 3.1, we have supY kaw S(1,g) < +∞. The following lemma is useful. Y kL(L2(Rn)) ≤ C(r, n, C0, N0)kak(2n+1) Lemma 3.3 (Cotlar). Let H be a Hilbert space and (Ω, A, ν) a measured space such that ν is a σ-finite positive measure. Let (Ay)y∈Ω be a measurable family of bounded operators on H such that y∈ΩZΩ sup kA∗ yAzk1/2 L(H)dν(z) ≤ M, y∈ΩZΩ sup kAyA∗ zk1/2 L(H)dν(z) ≤ M. Then for all u ∈ H, we have ZZΩ×Ω hAyu, AzuiHdν(y)dν(z) ≤ M 2kuk2 H, which implies the strong convergence of A = RΩ Aydν(y) and kAkL(H) ≤ M. In order to apply Cotlar's lemma, we should estimate k¯aw Y aw ZkL(L2(Rn)), i.e. a semi-norm of ¯aY ♯aZ in S(1, gY + gZ). Indeed, the following estimate holds. Lemma 3.4. Let g, aY be as above. For any k, N ∈ N, there exist C = C(k, N, n) > 0, l = l(k, N, n) ∈ N such that (3.2) k¯aY ♯aZk(k) S(1,gY +gZ ) ≤ Ck¯aY k(l) S(1,gY )kaZk(l) S(1,gZ )δr(Y, Z)−N . We use some biconfinement estimates, which can be found in [6, section 2.3], to prove Lemma 3.4. Definition 3.5 (Confined symbols). Let g be a positive-definite quadratic form on R2n such that g ≤ gσ. Let a be a smooth function on R2n and U ⊂ R2n. We say that a is g-confined in U, if for all k, N ∈ N, there exits Ck,N > 0 such that for all X, T ∈ R2n, We denote (3.3) (3.4) a(k)(X)T k ≤ Ck,N g(T )k/2(cid:0)1 + gσ(X − U)(cid:1)−N/2 . kak(k,N ) g,U = sup X,T ∈R2n,g(T )=1 a(k)(X)T k(cid:0)1 + gσ(X − U)(cid:1)N/2 , and a(l) g,U = max k≤l kak(k,l) g,U . Theorem 3.6 ([6, Theorem 2.3.2], biconfinement estimate). Let g1, g2 be two positive- definite quadratic forms on R2n such that gj ≤ gσ j . Let aj, j = 1, 2 be gj-confined in Uj, a gj-ball of radius ≤ 1. Then for all k, N ∈ N, for all X, T ∈ R2n, (3.5) (a1♯a2)(k)(X)T k ≤ Ak,N (g1+g2)(T )k/2(cid:16)1+(gσ 1 ∧gσ 2 )(X−U1)+(gσ 1 ∧gσ 2 )(X−U2)(cid:17)−N/2 , with Ak,N = γ(k, N, n)a1(l) g1,U1a2(l) g2,U2, l = 2n + 1 + k + N. Now we begin the proof of Lemma 3.4. 10 WEN DENG Proof of Lemma 3.4. The symbol aY is gY -confined in UY,r, since aY is supported in the gY -ball UY,r. Moreover, we have ∀k, N ∈ N, kaY k(k,N ) gY ,UY,r = sup a(k)(X)T k, X∈UY,r,T ∈R2n gY (T )=1 ∀l ∈ N, aY (l) gY ,UY,r = max k≤l kaY k(k,l) gY ,UY,r = kaY k(l) S(1,gY ). Applying (3.5) to ¯aY ♯aZ and using the triangular inequality (gσ Y ∧ gσ Z)(X − UY,r) + (gσ Y ∧ gσ Z)(X − UZ,r) ≥ 1 2 we get (gσ Y ∧ gσ Z)(UY,r − UZ,r), (¯aY ♯aZ)(k)(X)T k ≤ γ(k, N, n)k¯aY k(l) ×(cid:0)1 + 1 2 (gσ Y ∧ gσ S(1,gZ )(gY + gZ)(T )k/2 S(1,gY )kaZk(l) Z)(UY,r − UZ,r)(cid:1)−N/2 . Using the definition of the distance δr, we complete the proof of Lemma 3.4. (cid:3) End of the proof of Theorem 3.2. Now by Proposition 3.1, Lemma 3.4 and the estimate (2.12), we obtain that for any N > 0, there exists l = l(N, n) ∈ N such that k¯aw Y aw ZkL(L2(Rn)) ≤ C(n)k¯aw Y aw Zk(2n+1) S(1,gY +gZ ) ≤ C(N, n)k¯aY k(l) S(1,gY )kaZk(l) S(1,gZ )δr(Y, Z)−N ≤ C(N, n, C0)(cid:0)kak(l) S(1,g)(cid:1)2δr(Y, Z)−N The same inequality holds for aY ♯¯aZ. Choose N = 2N1, where N1 is given in (2.26), so that max{k¯aw Y aw Zk1/2 L(L2(Rn)), kaw Y ¯aw Zk1/2 L(L2(Rn))} ≤ Ckak(l) S(1,g)δr(Y, Z)−N1, where C = C(n, C0, N1) > 0, l = l(n, N1) ∈ N. Then together with Lemma 2.10, the assumptions of Cotlar's lemma are fulfilled with M = Ckak(l) S(1,g), and this completes the proof of Theorem 3.2. (cid:3) 4. Fefferman-Phong inequality In this section, we prove that the constant in the Fefferman-Phong inequality depends only on the structure constants of the metric and a fixed semi-norm of the symbol. Theorem 4.1 (Fefferman-Phong inequality). Let g be an admissible metric on R2n with structure constants (C0, N0) (see Definition 2.6). Let a be a non-negative symbol in S(λ2 g, g) (see Definition 2.3 and (2.17)). Then the operator aw on L2(Rn) is semi-bounded from below. More precisely, there exist l = l(n, C0, N0) ∈ N, C = C(n, C0, N0) > 0 such that (4.1) aw + Ckak(l) g ,g) ≥ 0. S(λ2 STRUCTURE CONSTANTS OF THE WEYL CALCULUS 11 4.1. The constant metric case. For the constant metric case, we use the results of Sjostrand and refer the readers to [6, page 116] for the detailed proof. Let 1 = Pj∈Z2n χ0(X −j) be a partition of unity, χ0 ∈ C ∞ χ0(X − j). c (R2n). Denote χj(X) = Proposition 4.2 ([6, Proposition 2.5.6]). Suppose a ∈ S(R2n). We say that a belongs to the class A if ωa ∈ L1(R2n), with ωa(Ξ) = supj∈Z2n F (χja)(Ξ), where F is the Fourier transform. We have S0 0,0 ⊂ S0,0;2n+1 ⊂ A ⊂ C 0(R2n) ∩ L∞(R2n), 0,0 = C ∞ where S0 b (R2n) is the space of C ∞ functions on R2n which are bounded as well as all their derivatives, S0,0;2n+1 is the set of functions defined on R2n such that (∂α x a)(x, ξ) ≤ Cαβ for α + β ≤ 2n + 1. A is a Banach algebra for the multiplication with the norm kakA = kωakL1(R2n). ξ ∂β Theorem 4.3 ([6, Theorem 2.5.10]). For all non-negative function a defined on R2n satisfying a(4) ∈ A, then the operator aw is semi-bounded from below. More precisely, where Cn depends only on the dimension n. aw + Cnka(4)kA ≥ 0, 4.2. Proof of Theorem 4.1. We shall use the partition of unity (ϕY )Y ∈R2n given in Theorem 2.4. Let (ψY )Y ∈R2n be a family of real-valued functions supported in UY,2r, equal to 1 on UY,r and (4.2) kψY k(k) S(1,g) ≤ C(k, r, C0). sup Y ∈R2n Indeed, with the same notations as in the proof of Theorem 2.4, the function ψY (X) = χ0(cid:0) 1 we write 2r−2gY (X − Y )(cid:1) satisfies the requirements. Then with aY = ϕY a, (4.3) ψY ♯aY ♯ψY = aY + rY . Lemma 4.4 (Estimate for rY ). For all k, N ∈ N, there exist C = C(k, N, C0) > 0, l = l(k, N, C0) ∈ N such that for all X ∈ R2n, T ∈ R2n with gY (T ) ≤ 1, r(k) Y (X)T k ≤ CkaY k(l) S(λg (Y )2,gY )(cid:0)1 + gσ Y (X − UY,2r)(cid:1)−N . Moreover, there exist C1 = C1(n, C0, N0) > 0, l1 = l1(n, C0, N0) ∈ N such that (4.4) (4.5) rw (cid:13)(cid:13)ZR2n Y gY 1/2dY(cid:13)(cid:13)L(L2(Rn)) ≤ C1kak(l1) S(λ2 g,g), To prove Lemma 4.4, we use the biconfinement estimate for the remainders, the proof of which can be found in [6, section 2.3]. Theorem 4.5 ([6, Theorem 2.3.4], biconfinement estimate). Let g1, g2 be two positive- definite quadratic forms on R2n with gj ≤ gσ j . Let aj, j = 1, 2 be gj-confined in Uj, a gj-ball of radius ≤ 1. Recall (2.20) rp(a1, a2)(X) = (a1♯a2)(X) − X0≤k<p 1 j!(cid:0)iπ[DX1, DX2](cid:1)j(cid:0)a1(X1)a2(X2)(cid:1)X1=X2=X. 12 WEN DENG Then for all k, l, p ∈ N, for all X, T ∈ R2n, we have (cid:0)rp(a1, a2)(cid:1)(k) ×(cid:16)1 + (gσ (4.6) (X)T k ≤ Ak,N,p(g1 + g2)(T )k/2Λ−p 1,2 1 ∧ gσ 2 )(X − U1) + (gσ 1 ∧ gσ 2 )(X − U2)(cid:17)−N/2 with Ak,N,p = C(k, N, p, n)a1(l) g2,U2, l = 2n + 1 + k + p + N and g1,U1a2(l) g2(T )(cid:17)1/2 1 (T ) T ∈R2n,T 6=0(cid:16)gσ inf = T ∈R2n,T 6=0(cid:16)gσ 2 (T ) g1(T )(cid:17)1/2 inf . (4.7) Λ1,2 = Now we use Theorem 4.5 to prove Lemma 4.4. Proof of Lemma 4.4. By the asymptotic formula (2.24), we have ψY ♯aY = aY + 1 4iπ {ψY , aY } +r2(ψY , aY ), =0 {z } since ψY = 1 on the support of aY . The symbol ψY is gY -confined in UY,2r and aY is gY -confined in UY,r, and moreover, we have ∀l ∈ N, ψY (l) gY ,UY,2r = kψY k(l) S(1,gY ), aY (l) gY ,UY,r = λg(Y )2kaY k(l) S(λg(Y )2,gY ). Applying (4.6) to r2(ψY , aY ), we have for all k, N ∈ N, there exist C(k, N, n) > 0, l(k, N, n) ∈ N such that for all X, T ∈ R2n, (cid:0)r2(ψY , aY )(cid:1)(k)(X)T k ≤ C(k, N, n)ψY (l) ≤ C(k, N, n)kψY k(l) (4.8) gY ,UY,2r aY (l) S(1,gY )kaY k(l) gY ,UY,r gY (T )k/2Λ−2 1,2(cid:0)1 + gσ S(λg(Y )2,gY )gY (T )k/2(cid:0)1 + gσ Y (X − UY,2r)(cid:1)−N Y (X − UY,2r)(cid:1)−N , noticing here Λ1,2 defined in (4.7) is equal to λg(Y ). An analogous estimate as (4.8) holds for r2(aY , ψY ). In our case, we write rY , which is defined in (4.3), rY = (ψY ♯aY − aY )♯ψY + (aY ♯ψY − aY ) = r2(ψY , aY )♯ψY + r2(aY , ψY ). Then the estimate (4.4) follows from (4.8) and (3.5). Furthermore, for any k, N ∈ N, there exist C = C(k, N, n, C0) > 0, l = l(k, N, n, C0) ∈ N such that k¯rY ♯rZk(k) S(1,gY +gZ ) ≤ CkaY k(l) S(λg(Y )2,gY )kaZk(l) S(λg(Z)2,gZ )δ2r(Y, Z)−N . Thus we can apply Cotlar's lemma and get the estimate (4.5). (cid:3) Lemma 4.6 (Estimate for ψY ). For all k, N ∈ N, there exist C = C(k, N, C0) > 0, l = l(k, N, C0) ∈ N such that for all X ∈ R2n, T ∈ R2n with gY (T ) ≤ 1, (4.9) (ψY ♯ψY )(k)(X)T k ≤ C(cid:0)kψY k(l) S(1,gY )(cid:1)2(cid:0)1 + gσ Y (X − UY,2r)(cid:1)−N . Moreover, there exists C2 = C2(n, C0, N0) > 0 such that (4.10) kZ ψw Y ψw Y gY 1/2dY kL(L2(Rn)) ≤ C2. STRUCTURE CONSTANTS OF THE WEYL CALCULUS 13 Proof. The inequality (4.9) follows immediately from (3.5). And it follows from (3.5), (4.2) and (4.9) that for all k, N ∈ N, k(ψY ♯ψY )♯(ψZ♯ψZ)k(k) S(1,gY +gZ) ≤ Cδ2r(Y, Z)−N , for some C = C(k, N, n, C0) > 0. Then by choosing N = 2N1 and using Cotlar's lemma, we get the estimate (4.10). (cid:3) End of the proof of Theorem 4.1. The symbol aY is non-negative and uniformly in S(λg(Y )2, gY ), so that we can apply the Fefferman-Phong inequality (Theorem 4.3) for the constant metric gY to get Y + C(n)kaY k(l(n)) aw By Proposition 2.5 and Lemma 2.8, we have S(λg (Y )2,gY ) ≥ 0. kaY k(l(n)) S(λg(Y )2,gY ) ≤ C(n, C0, N0)kak(l(n)) S(λ2 g ,g), so that (4.11) Y + C3kak(l(n)) aw S(λ2 g,g) ≥ 0. where C3 = C3(n, C0, N0) > 0, l(n) ∈ N are constants. Combining (4.3), (4.5), (4.10) and (4.11), we obtain aw = ZR2n = ZR2n aw Y gY 1/2dY ψw Y aw Y ψw Y gY 1/2dY −ZR2n g ,g)ZR2n ψw Y ψw rw Y gY 1/2dY ≥ −C3kak(l(n)) S(λ2 Y gY 1/2dY − C1kak(l1) S(λ2 g,g) ≥ −Ckak(l) g,g), S(λ2 for some C = C(n, C0, N0) > 0 and l = l(n, C0, N0) ∈ N. The proof of Theorem 4.1 is complete. (cid:3) References [1] Jean-Michel Bony, Sur l'in´egalit´e de Fefferman-Phong, Seminaire: ´Equations aux D´eriv´ees Partielles, 1998 -- 1999, S´emin. ´Equ. D´eriv. Partielles, ´Ecole Polytech., Palaiseau, 1999, pp. Exp. No. III, 16. MR 1721321 (2000i:35232) [2] Jean-Michel Bony and Jean-Yves Chemin, Espaces fonctionnels associ´es au calcul de Weyl- Hormander, Bull. Soc. Math. France 122 (1994), no. 1, 77 -- 118. MR 1259109 (95a:35152) [3] Abdesslam Boulkhemair, On the Fefferman-Phong inequality, Ann. Inst. Fourier (Grenoble) 58 (2008), no. 4, 1093 -- 1115. MR 2427955 (2009h:35468) [4] Nils Dencker, The resolution of the Nirenberg-Treves conjecture, Ann. of Math. (2) 163 (2006), no. 2, 405 -- 444. MR 2199222 (2006i:35386) [5] Lars Hormander, The analysis of linear partial differential operators. III, Classics in Math- ematics, Springer, Berlin, 2007, Pseudo-differential operators, Reprint of the 1994 edition. MR 2304165 (2007k:35006) [6] Nicolas Lerner, Metrics on the phase space and non-selfadjoint pseudo-differential operators, Pseudo-Differential Operators. Theory and Applications, vol. 3, Birkhauser Verlag, Basel, 2010. MR 2599384 [7] Nicolas Lerner and Yoshinori Morimoto, On the Fefferman-Phong inequality and a Wiener-type algebra of pseudodifferential operators, Publ. Res. Inst. Math. Sci. 43 (2007), no. 2, 329 -- 371. MR 2341014 (2009b:47083) 14 WEN DENG Wen Deng, Institut de Math´ematiques de Jussieu, Universit´e Pierre-et-Marie- Curie (Paris 6), 4 place Jussieu, 75005 Paris, France. E-mail address: [email protected] URL: http://www.math.jussieu.fr/~wendeng/
1302.5790
1
1302
2013-02-23T12:09:47
Crofton formulae and geodesic distance in hyperbolic spaces
[ "math.FA" ]
The geodesic distance between points in real hyperbolic space is a hypermetric, and hence is a kernel negative type. The proof given here uses an integral formula for geodesic distance, in terms of a measure on the space of hyperplanes. An analogous integral formula, involving the space of horospheres, is given for complex hyperbolic space.By contrast geodesic distance in a projective space is not of negative type.
math.FA
math
CROFTON FORMULAE AND GEODESIC DISTANCE IN HYPERBOLIC SPACES GUYAN ROBERTSON Abstract. The geodesic distance between points in real hyper- bolic space is a hypermetric, and hence is a kernel negative type. The proof given here uses an integral formula for geodesic distance, in terms of a measure on the space of hyperplanes. An analogous integral formula, involving the space of horospheres, is given for complex hyperbolic space. By contrast geodesic distance in a pro- jective space is not of negative type. Introduction Motivated by problems in harmonic analysis on homogeneous spaces, J. Faraut and K. Harzallah showed in [FH] that the geodesic distance d on a real or complex hyperbolic space of dimension n ≥ 2 is a kernel of negative type. Equivalently, √d is a Hilbert space distance [HV: Chapter 5]. On the other hand, in the real hyperbolic plane, there is an explicit Crofton integral formula for the length L of a rectifiable curve C in terms of a measure µ on the space of geodesics [S1, Section 3], [GS, Section 4.4]. The formula has the form R n(γ)dµ(γ) = 2L, where n(γ) is the number of times the geodesic γ meets C and the integral is over the space of all geodesics. (M.W. Crofton proved the corresponding formula for the euclidean plane in 1868.) This paper proves a Crofton formula for geodesic distances in real hyperbolic space of dimension n, from which one obtains an explicit geometric proof of the result of Faraut and Harzallah, in the real case. In this case the measure µ is defined on the space of totally geodesic submanifolds of codimension one. There is also a Crofton formula in the complex case. However, the measure is defined on the space of horospheres, and it is not clear how to deduce that d is of negative type. Date: July 28, 1997. 1991 Mathematics Subject Classification. Primary 22D10. This research was supported by the Australian Research Council. 1 Typeset by AMS-LATEX. 2 G. ROBERTSON Hyperbolic spaces H n F can be defined over R, C, H and (in the case n = 2) the octonions O. In each case the space is two point homoge- neous: if d(x, y) = d(x′, y′) then there is an isometry g of H n F such that gx = x′, gy = y′ [He1, Chapter IX, Proposition 5.1, p.355]. It turns out that this property, together with the existence of a certain invari- ant measure is at the heart of the Crofton formulae. We note that the noncompact riemannian manifolds which are two point homogeneous are exactly the euclidean spaces and hyperbolic spaces [He1]. The Crofton formula that we prove for H n R has the consequence that the geodesic distance d is a hypermetric in the sense of [A]. It follows R, endowed with the metric √d, embeds that any finite subset of H n isometrically in a euclidean sphere (Corollary 3.2). Finally, we complete the investigation of Faraut and Harzallah [FH] for symmetric spaces of rank one, by showing explicitly that the geo- desic distance on a projective space is not of negative type, although it can be expressed by means of a Crofton formula. 1. motivation and preliminaries In the course of the proof of [RS, Proposition 1.4], a geometric ar- gument was given to show that the euclidean distance d(x, y) between points x, y ∈ Rn is a kernel of negative type. In retrospect, the basis of that argument is seen to be the classical Crofton formula asserting that d(x, y) equals the measure of the set of euclidean hyperplanes which meet the line segment [xy]. The relevant measure is the appropriately normalized measure on the space of hyperplanes which is invariant un- der isometries of Rn. This measure lifts to a measure µ on the space of half spaces of Rn. Then d(x, y) = µ(Sx△Sy), where Sx denotes the set of half spaces which contain x. It is natural to try to extend this ar- gument to the rank one symmetric spaces considered by J. Faraut and K. Harzallah [FH]. It turns out that the only such spaces for which the geodesic distance is a kernel of negative type are real or complex hyperbolic spaces, and euclidean spheres. Let F be one of the (skew-) fields R, C or H. Regard Fn+1 as a right vector space over F and define a hermitian form on Fn+1 by means of the formula hz, wi = −z0w0 + z1w1 + · · · + znwn. F is the image in the projective space P n F of The hyperbolic space H n the set of negative vectors {z ∈ Fn+1 : hz, zi < 0}. 3 It is convenient to use the same notation for a negative vector x ∈ F . The geodesic distance between Fn+1 and its equivalence class in P n points x, y in X is given by cosh d(x, y) = hx,yi (hx,xihy,yi) . 1 2 The octionic hyperbolic plane H 2 O requires a more involved definition [M]. 2. real hyperbolic space Denote by X the hyperbolic space H n R [CG]. Then X is the image in the projective space RPn of the set of negative vectors (cid:8)x ∈ Rn+1 : hx, xi = −(x0)2 + (x1)2 + . . . (xn)2 < 0(cid:9) . The orthogonal group G = O(1, n) is the subgroup of GL(n + 1, R) which preserves the form h·,·i, and G acts isometrically and transi- tively on X. The stabilizer in G of the point x0 = (1, 0, . . . , 0) ∈ X is the compact group K = O(1) × O(n). Thus X is isomorphic to the topological homogeneous space G/K. The hyperbolic space H n−1 em- beds naturally into X = H n R as the subspace S0 consisting of all points with last coordinate equal to zero. R Every totally geodesic submanifold of codimension one in X is a G -- translate of S0 [CG, Proposition 2.5.1]. It is convenient simply to refer to such a submanifold as a hyperplane. The space S of all hyper- planes may therefore be identified with the topological homogeneous space G/G(S0), where G(S0) is the subgroup of G consisting of ele- ments which leave S0 globally invariant. By [CG, Lemma 4.2.1], we have G(S0) ∼= O(1, n − 1) × O(1). The groups G and G(S0) are both unimodular [He1, Chapter X, Proposition 1.4]. (A direct proof is given in [BP, Proposition C.4.11].) It follows [N, Chapter 3, p.140, Corollary 4] that there is a nonzero positive G-invariant measure µS on the space S of hyperplanes. We also consider the space H of half spaces in X. These are the G -- translates of the half -- space H0 consisting of points with last coordinate positive. The group G(H0) ∼= O(1, n − 1) is unimodular and there is a corresponding invariant measure µH on H = G/G(H0). There is a natural double covering π : H → S and so by uniqueness of the measures (up to a positive multiple) [N, Chapter 2, p.95, Corollary] we may assume that µS = π∗ ◦ µH. first prove the following Crofton formula. Let [xy] denote the (unique) geodesic between points x, y ∈ X. We 4 G. ROBERTSON Proposition 2.1. There is a constant k > 0 such that, if x, y ∈ X then µS {S ∈ S : S ∩ [xy] 6= ∅} = kd(x, y). For the purposes of the proof, we introduce the notation m(x, y) = µS {S ∈ S : S ∩ [xy] 6= ∅} . Lemma 2.2. If x, y ∈ X then {S ∈ S : S ∩ [xy] 6= ∅} is compact; (a) (b) m(x, y) < ∞. Proof: (a) Let g, h ∈ G. We claim that the point gx0 lies on the hyperplane hS0 if and only if gK ∩ hG(S0) 6= ∅. The crucial fact used in the proof is that G(S0) acts transitively on S0. Therefore gx0 ∈ hS0 ⇐⇒ h−1gx0 ∈ S0 ⇐⇒ h−1gx0 = g0x0, for some g0 ∈ G(S0) ⇐⇒ g−1 0 h−1g ∈ K, for some g0 ∈ G(S0) ⇐⇒ h−1g = g0k0, for some g0 ∈ G(S0), k0 ∈ K ⇐⇒ gK ∩ hG(S0) 6= ∅. This proves our assertion. Furthermore, since [xy] is compact, there exists a compact subset J ⊂ G such that Jx0 = [xy], by [N, p.137, Lemma 1]. A hyperplane hS0 meets [xy] if and only if JK∩hG(S0) 6= ∅, that is h ∈ JKG(S0). The set of such hyperplanes hS0 is therefore compact, being the image of the compact set JK under the quotient map G → G/G(S0). (b) This follows immediately from (a). (cid:3) Proof of Proposition 2.1. We must prove that m(x, y) = kd(x, y) where k > 0 is constant. This is based on the following facts: (a) X is two-point homogeneous with respect to the action of G; (b) It follows that m(x, y) = m(x′, y′) the measure µS is G-invariant. whenever d(x, y) = d(x′, y′). Moreover, by considering a large number of pairwise disjoint geodesic segments of equal length in some fixed geodesic segment [ab], we see that m(x, y) → 0 as d(x, y) → 0. In particular m(x, x) = 0. If we divide [xy] into s segments [xixi+1] of equal length (1 ≤ i ≤ s), then m(x, y) = sm(x1, x2). In this way, if d(x′, y′) = qd(x, y) where q is rational, then m(x′, y′) = qm(x, y). Now m(x, y) < ∞, by Lemma 2, and so by continuity there is a constant k ≥ 0 such that m(x, y) = kd(x, y) for all x, y ∈ X. We must check that k > 0. For this, it is enough to show that m(x, y) > 0 for some 5 x, y. Take x = (1, 0, 0. . . . , ǫ), y = (1, 0, 0. . . . ,−ǫ). Then the hyper- plane S0 meets [xy] transversally at the interior point x0 = (1, 0, . . . , 0) of the geodesic segment [xy]. There is an open neighbourhood eV of the identity in G such that gS0 meets [xy] at an interior point for all g ∈ eV . The image V of eV in G/G(S0) is an open set which is contained in the set of all hyperplanes meeting [xy]. Since V has positive measure, this implies that m(x, y) > 0, as required. (cid:3) Remark 2.3. Unlike the case n = 2 in [S1, Section 3], [GS, Section 4.4], the above proof of the Crofton formula in hyperbolic space is not constructive. However it gives a geometric explanation of why the result is true. We now re-interpret Proposition 1, replacing µS by µH, the invariant measure on the space of all half-spaces. This will allow us to use the method of [RS] to show that the metric on X is of negative type. Given x ∈ X, let Σx denote the set of half-spaces containing x. Lemma 2.4. If x, y ∈ X then µH(Σx △ Σy) = µS {S ∈ S : S ∩ [xy] 6= ∅} . Proof: The double covering π : H → S is given by π(H) = ∂H, the boundary of H. Since a hyperplane S is totally geodesic we have either # ([xy] ∩ S) ≤ 1 or [xy] ⊆ S. Therefore Σx △ Σy = π−1 {S ∈ S : S ∩ [xy] 6= ∅ and [xy] 6⊆ S} . Now µS {S ∈ S : [xy] ⊆ S} = 0. The result follows, since µS = π∗ ◦ µH. Corollary 2.5. If x, y ∈ X then d(x, y) = kµH(Σx △ Σy) where k > 0 is constant. Hence d is a kernel of negative type. (cid:3) Proof: The first assertion is immediate from Lemma 2.4. The second assertion follows by embedding X into L2(H, µH) via x 7→ vx where vx = χx − χx0, and χx is the characteristic function of Σx. Then 2 and so √d is a Hilbert space distance. (c.f. µH(Σx △ Σy) = kvx − vyk2 [RS, Proposition 1.1].) (cid:3) Remark 2.6. Faraut and Harzallah show [FH] that the distance func- tion on real or complex hyperbolic space is a kernel of negative type. See [HV, Chapitre 6, Th´eor`eme 21]. The corresponding result for quater- nionic hyperbolic space is false, because the group of isometries Sp(1, n) has Kazhdan's property (T ) [FH, Th´eor`eme 6.4]. It would be interesting to have a direct proof of this fact, avoiding the use of property (T ). 6 G. ROBERTSON Remark 2.7. Corollary 2.5 is stronger than the result of Faraut and Harzallah. It asserts that the distance d is a measure definite kernel, in the sense of [RS], a concept which is in general strictly stronger than that of negative type. See the next section. Remark 2.8. Since G(H0) is a closed noncompact subgroup of G, it follows from Moore's ergodicity Theorem [Z, Theorem 2.2.6] that any lattice subgroup Γ of G acts ergodically on H = G/G(H0). Note that the group Γ does not have property (T ). Fix an element x ∈ X then µH(Σx△gΣx) = µH(Σx△Σgx) is a positive constant multiple of d(x, gx), and hence unbounded. The action of Γ in [RS, Theorem 2.1] may therefore be chosen to be ergodic. Whether one can find an ergodic action with similar properties for all groups without property (T ) is an open question. Remark 2.9. Consider any Coxeter system (W, S), where W is a Cox- eter group and S is the canonical generating set. Let ∆ denote the associated Coxeter complex. By a result of J. Tits [B, Chap. IV, p.41], the natural distance d(c, c′) between chambers c and c′ is equal to the number of walls of ∆ separating c and c′. This is precisely the Crofton formula for distance, relative to counting measure on the space of walls. Let H denote the space of half spaces ("roots") of ∆ with counting measure. If c is a chamber of ∆, let Hc be the set of roots containing c. Then 2d(c, c′) = #(Hc △ Hc′). (See the proof of [HV, Proposition 6.14].) The argument proceeds as before, showing that d is a kernel of negative type. The same type of argument is applied in [BS, Proof of Theorem 3] to a polygonal complex X, which is locally finite, simply connected and of type (4, 4) or (6, 3). Let B denote the set consisting of all barycentres of faces of X or of edges of X which are not adjacent to faces. An unbounded negative definite kernel N of the above form is defined on B. The kernel N is equivalent to the geodesic distance between elements of B and is invariant under automorphisms of X. This is used in [BS] to show that a properly discontinuous group of automorphisms of X does not have property (T). 3. the hypermetric property for H n R A semimetric space (X, d) is said to be hypermetric if it satisfies the following property: for each finite subset {x1, x2 . . . , xm} of X and inte- gers {t1, t2 . . . , tm} such thatPm i,j=1 titjd(xi, xj) ≤ 0. The corresponding statement, with t1, t2 . . . , tm real numbers satis- fying Pm i=1 ti = 0, says that d is a kernel negative type. It is easy to i=1 ti = 1, we havePm 7 see that if d has the hypermetric property then d is a kernel negative type [A, 1.2]. Suppose that (W, µ) is a measure space and that a semimetric d on a set X is defined by the formula d(x, y) = µ(Sx △ Sy), where for each x ∈ X, Sx is a measurable set. It was proved in [K, Theorem 3.1] that d is then a hypermetric. The next result is therefore an immediate consequence of Corollary 2.5. Corollary 3.1. The geodesic distance d on H n R is a hypermetric. Finite hypermetric spaces have been characterized up to isometry in [A], and there is a detailed exposition of their properties in [DGL]. In view of the fact that the space (H n R, d) has constant negative curvature, the next result may seem slightly surprising. Corollary 3.2. Let {x1, x2 . . . , xm} be any finite subset of H n with the metric √d. Then ({x1, x2 . . . , xm},√d) embeds isometrically in a euclidean sphere in Rp, where p ≥ log2 m. from [HV, Proposition 5.14] that ({x1, x2 . . . , xm},√d) embeds isomet- Proof: Since d is of negative type (being hypermetric) it follows rically in Rp, for some p. The fact that the image is contained in a euclidean sphere is then a consequence of [A, Lemme 1.12]. The esti- mate p ≥ log2 m is provided by [A, Proposition 1.18]. (cid:3) 4. complex hyperbolic space R, endowed C of CPn. Complex hyperbolic space H n C is constructed in a manner similar to that of H n R [CG]. The bihermitian form hz, wi = −z0w0 + z1w1 +· · · + znwn on Cn defines a set of negative vectors in Cn+1, defined by the condition hz, zi < 0, which projects to the subspace H n One might try to mimic the arguments of the preceding section in the complex case. The space H n C is two point homogeneous with respect to the isometry group U(1, n). Natural analogues of hyperplanes are the equidistant hypersurfaces [GP, Section 4], also known as spinal surfaces in [M]. These are, by definition, subspaces of the form S = {z ∈ H n C. There is a natural invariant measure on the space of equidistant hypersurfaces, but we cannot prove an analogue of Lemma 2.2, because the group G(S0) of elements of U(1, n) which leave a given equidistant hypersurface S0 globally invariant does not act transitively on S0 [M, Section 3.2]. An additional problem is that equidistant hypersurfaces are not totally geodesic and a geodesic segment can meet an equidistant hypersurface more than once, without being contained in the hypersurface. However, C : d(z, z1) = d(z, z2)}, where z1, z2 ∈ H n 8 G. ROBERTSON even in the complex case, it is possible to prove a Crofton formula if we consider horospheres instead of equidistant hypersurfaces. Horospheres in complex hyperbolic space are described geometri- cally in [GP, Section 1]. From a group-theoretic point of view they may be described as follows [He2, Chapter II.1]. Let G = U(1, n), so that H n C = G/K where K = U(1) × U(n). Let G = KAN be the Iwasawa decomposition and M the centralizer of A in K. Thus MAN is a minimal parabolic subgroup. Let x0 be the origin K in G/K and ξ0 = Nx0 Then ξ0 is a horosphere and the subgroup of G which maps ξ0 to itself equals MN. The homogeneous space G/MN is the space of horocycles. Note that N is isomorphic to the complex Heisenberg group [He2, Chapter 2, p.215]. The groups G and MN are therefore unimodular [He1, Chapter X, Proposition 1.4], and there is a G-invariant measure on G/MN. Since MN acts transitively on ξ0, the set of horospheres which meet a geodesic segment [xy] is compact, as in Lemma 2.2. Consider the expression m(x, y) =RG/M N n(h)dµ(h), where n(h) is the number of times a horocycle h ∈ G/MN meets the geodesic segment [xy]. Since H n C is two-point homogeneous, and µ is G-invariant, m(x, y) depends only on d(x, y). The same argument as in the proof of Proposition 2.1 establishes the following Crofton formula. Proposition 4.1. There is a constant k > 0 such that, for all x, y ∈ H n C, ZG/M N n[x, y](h)dµ(h) = kd(x, y) where n[x, y](h) is the number of times that h meets [xy]. Remark 4.2. Because it is possible to have 1 < n[x, y](h) < ∞, it is not clear how to use this result to prove that d(x, y) is of negative type, as we did for the real case in Corollary 2.5. In fact, the example in the next section indicates that a Crofton formula alone is not enough. Note that a Crofton formula involving horospheres is clearly also valid in real hyperbolic space. 5. projective space The purpose of this section is twofold. Firstly we give an example to show that the existence of a Crofton formula for a distance func- tion does not in general imply that the distance is of negative type. Secondly, this example completes the classification of riemannian sym- metric spaces of rank one for which the geodesic distance is of negative type. The projective spaces P n F of dimension n ≥ 2 over R, C, H and (in the case n = 2) the octonions O, are compact two-point homogeneous spaces. Proposition 5.1. The geodesic distance d on P n type. F is not of negative 9 The projective spaces all contain P 2 it is enough to consider the case of P 2 used for a vector x ∈ R3 and its equivalence class in P 2 distance between points x, y in P 2 R as a geodesic subspace, and so R. As usual, the same notation is R. The geodesic , R is given by cos d(x, y) = (x,y) ((x,x)(y,y)) 1 2 where (x, y) is the usual inner product on R3. Proposition 5.2. The geodesic distance d on P 2 type. R is not of negative Proof: We show that the metric d is not of negative type by ex- hibiting points x1, x2 . . . , x6 ∈ P 2 R and real numbers t1, t2 . . . , t6 sat- i=1 ti = 0 such that P6 isfying P6 In fact we choose (t1, t2 . . . , t6) = (1, 1, 1,−1,−1,−1) and for notational conve- nience put (x1, x2, x3, x4, x5, x6) = (p1, p2, p3, q1, q2, q3). It is enough to choose these points so that i,j=1 titjd(xi, xj) > 0. X d(pi, pj) +X d(qi, qj) >X d(pi, qj). (5.1) We make the following choices. Let p1 = (1, 0, 1), p2 = (1, 0,−1), p3 = (0, 1, 0), and q1 = (0, 1, 1), q2 = (0, 1,−1), q3 = (1, 0, 0). Then d(pi, pj) = π/2, d(qi, qj) = π/2, for i, j ∈ {1, 2, 3}. Moreover d(pi, qj) = π/3 if both i, j ∈ {1, 2}, π/4 if exactly one of i, j equals 3,  π/2 if i = 3 and j = 3. We therefore obtainP d(pi, pj)+P d(qi, qj) = 3π/2+3π/2 = 3π and P d(pi, qj) = 4π/3 + 4π/4 + π/2 = 17π/6, which proves the inequality (5.1). (cid:3) Example. There is a Crofton formula for geodesic distance d on the space P n R , but as we have just seen, this distance is not of negative type. The proof of a Crofton formula for d(x, y) in terms of totally geodesic hypersurfaces which meet the segment [xy] is exactly the same as for real hyperbolic space in Section 2, except that by compactness there is no need for an analogue of Lemma 2.2. The reason that the analogue of Corollary 2.5 fails is that a totally geodesic hypersurface does not separate P n R into two parts. 10 G. ROBERTSON Remark 5.3. Two point homogeneous riemannian manifolds have been classified completely [He1, Chapter IX §5]. They are the euclidean spaces, the circle, and the symmetric spaces of rank one. The sym- metric spaces of rank one for which the geodesic distance is of negative type are the spheres and the real or complex hyperbolic spaces. This follows from the results of [FH], together with our result for projective spaces. Note that it is easy to give a new proof that the geodesic dis- tance on a sphere is of negative type along the lines of Corollary 2.5, using a measure on the space of half-spaces. Remark 5.4. There is a naturally occurring invariant metric of neg- ative type on the projective space P n F [Hi]. For there is an embedding of P n F as the set of primitive idempotents in the formally real Jordan algebra A of hermitian n× n matrices over F. The trace form defines a euclidean distance dE on A which induces a metric of negative type on F . It follows from [Hi] that P n P n F is also two point homogeneous relative to this metric. ADDENDUM In Remark 2.6 we asked for a direct proof of the fact that the distance function on quaternionic hyperbolic space is a kernel of negative type. This is in fact possible, by considering the following 24 points. For σ ∈ {+1,−1} and ǫ ∈ {±i,±j,±k}, let xσ ǫ = (3, 0, 2σ + 2ǫ). Direct calculation shows that δ ) −X d(xσ X d(xσ ǫ , yρ ǫ and −1 to the yσ ǫ . δ ) = 417.03 − 415.77 > 0. This shows that the condition for negative type fails with the number +1 being associated to the xσ ǫ = (3, 2σ + 2ǫ, 0) and yσ ǫ , xρ δ) +X d(yσ ǫ , yρ References [A] P. Assouad, Sur les in´egalit´es valides dans L1, Europ. J. Combinatorics 5 (1984), 99 -- 112. [BS] W. Ballmann and J. ´Swia tkowski, On L2- cohomology and property (T) for ' automorphism groups of polyhedral cell complexes, preprint, Universitat Bonn, 1996. N. Bourbaki. Groupes et Alg`ebres de Lie, Chapitres IV, V et VI., Masson, Paris 1981. [B] [BP] R. Benedetti and C. Petronio, Lectures on Hyperbolic Geometry, Springer- [CG] Verlag, Berlin-Heidelberg-New York, 1992. S.S. Chen and L. Greenberg, Hyperbolic spaces, Bers Symposium "Contri- butions to Analysis", L.V. Ahlfors (ed.), (Academic Press, 1974), 49-87. [DGL] M. Deza, V.P. Grishukhin and M. Laurent, Hypermetrics in the geometry of numbers, DIMACS Series in Discrete Mathematics and Computer Science 20, 1995. 11 [FH] J. Faraut and K. Harzallah, Distances hilbertiennes invariantes sur un es- pace homog`ene, Ann. Inst. Fourier 24 (1974), 171-217. [GP] W.M. Goldman and J.R. Parker, Dirichlet polyhedra for dihedral groups acting on complex hyperbolic space, J. Geometric Analysis 2 (1992), 517 -- 554. I.M. Gelfand and M.M. Smirnov, Lagrangians satisfying Crofton formulas, Radon transforms, and nonlocal differentials, Advances in Math. 109 (1994), 188 -- 227. [GS] [HV] P. de la Harpe and A. Valette, La Propri´et´e (T) de Kazhdan pour les groupes localement compacts, Ast´erisque 179, Soc. Math. France, 1989. [He1] S. Helgason, Differential Geometry and Symmetric Spaces, Academic Press, New York, 1962. [Hi] [K] [M] [He2] S. Helgason, Geometric Analysis on Symmetric Spaces, Mathematical Sur- veys and Monographs, vol.39, American Mathematical Society, Providence, R.I., 1994. U. Hirzebruch, Uber Jordan-Algebren und kompakte Riemannsche sym- metrische Raume vom Rang 1, Math. Zeitschr. 90 (1965), 339 -- 354. J.B. Kelly, Metric inequalities and symmetric differences, Inequalities II, O. Shisha (ed.), Academic Press (1970), 193 -- 212. G. Mostow, On a remarkable class of polyhedra in complex hyperbolic space, Pac. J. Math. 86 (1980), 171-276. [N] L. Nachbin, The Haar Integral, van Nostrand, Princeton, New Jersey, 1965. [RS] G. Robertson and T. Steger, Negative definite kernels and a dynamical characterization of property (T ) for countable groups, Ergodic Theor. and Dynam. Systems, to appear. L.A. Santal´o, Integral geometry on surfaces of constant negative curvature, Duke Math. J. 10 (1943), 687-704 L.A. Santal´o, Integral Geometry and Geometric Probability, Addison- Wesley, Reading, Mass., 1976. R. Zimmer, Ergodic Theory and Semisimple Groups, Monographs in Math., vol. 81, Birkhauser, Boston, 1984. [S2] [S1] [Z] Department of Mathematics, University of Newcastle, NSW 2308, AUSTRALIA E-mail address: [email protected]
1705.04926
1
1705
2017-05-14T08:14:32
Frames in Quaternionic Hilbert Spaces
[ "math.FA" ]
In this paper, we introduce and study the frames in separable quaternionic Hilbert spaces. Results on the existence of frames in quaternionic Hilbert spaces have been given. Also, a characterization of frame in quaternionic Hilbert spaces in terms of frame operator is given. Finally, a Paley-Wiener type perturbation result for frames in quaternionic Hilbert space has been obtained.
math.FA
math
Akxk2≤ hx, xni2≤ Bkxk2, for all x ∈ H." (1.1) ∞ Xn=1 FRAMES IN QUATERNIONIC HILBERT SPACES S.K. Sharma and Shashank Goel Abstract. In this paper, we introduce and study the frames in separable quaternionic Hilbert spaces. Results on the existence of frames in quaternionic Hilbert spaces have been given. Also, a characterization of frame in quaternionic Hilbert spaces in terms of frame operator is given. Finally, a Paley-Wiener type perturbation result for frames in quaternionic Hilbert space has been obtained. Duffin and Schaeffer [11] indroduced frames for Hilbert spaces while working on some deep problems in non-harmonic Fourier series. They gave the following definition: 1. Introduction "Let H be a Hilbert space. Then {xn}n∈N ⊂ H is said to be a frame for H if there exist finite constants A and B with 0 < A ≤ B such that The positive constants A and B , respectively, are called lower and upper frame bounds for the frame {xn}n∈N . The inequality (1.1) is called the frame inequality for the frame {xn}n∈N . A frame {xn}n∈N in H is said to be • tight if it is possible to choose A = B . • Parseval if it is a tight frame with A = B = 1. Frame theory began to spread among researchers when in 1986, Daubechies, Grossmann and Meyer published a fundamental paper [10] and observed that frames can be also used to provide the series expansions of functions in L2(R). The main property of frames which makes them so useful is their redundancy, due to which, representation of signals using frames is advantageous over basis expansions in a variety of practical applications. Frame works as an important tool in the study of signal and image processing [2], filter bank theory [4], wireless communications [13] and sigma-delta quantization [3]. For more literature on frame theory, one may refer to [5, 8, 6]. Recently, Khokulan, Thirulogasanthar and Srisatkunarajah [14] introduced and studied frames for finite dimensional quaternionic Hilbert spaces. Sharma and Virender [15] study some different types of dual frames of a given frame in a quaternionic Hilbert space and gave various types of reconstructions with the help of dual frame. In this paper, we will introduce and study the frames in separable quaternionic Hilbert spaces . Results on the existence of frames in quaternionic Hilbert spaces have been given. Also, a characterization of frame in quaternionic Hilbert spaces in terms of frame operator is given. Finally, a Paley-Wiener type perturbation result for frames in quaternionic Hilbert space has been obtained. 2. Quaternionic Hilbert space As the quaternions are non-commutative in nature therefore there are two different types of quaternionic Hilbert spaces exist, the left quaternionic Hilbert space and the 2010 Mathematics Subject Classification. 42C15, 42A38. Key words and phrases. Frame, Quaternionic Hilbert spaces. 1 2 Sharma and Goel right quaternionic Hilbert space depending on positions of quaternions. In this section, we will study some basic notations about the algebra of quaternions, right quaternionic Hilbert space and operators on right quaternionic Hilbert spaces. Throughout this paper, we will denote Q to be a non-commutative field of quaternions, I be a non empty set of indicies, VR(Q) be a separable right quaternionic Hilbert space, by the term "right linear operator", we mean a "right Q-linear operator" and B(VR(Q)) denotes the set of all bounded (right Q-linear) operators of VR(Q): B(VR(Q)) := {T : VR(Q) → VR(Q) : kTk< ∞}. The non-commutative field of quaternions Q is a four dimensional real algebra with In Q, 0 denotes the null element and 1 denotes the identity with respect to unity. multiplication. It also includes three so-called imaginary units, denoted by i, j, k . i.e., Q = {x0 + x1i + x2j + x3k : x0, x1, x2, x3 ∈ R} where i2 = j 2 = k2 = −1; ij = −ji = k; jk = −kj = i and ki = −ik = j . For each quaternion q = x0 + x1i + x2j + x3k ∈ Q, define conjugate of q denoted by q as If q = x0 + x1i + x2j + x3k is a quaternion, then x0 is called the real part of q and x1i + x2j + x3k is called the imaginary part of q . The modulus of a quaternion q = x0 + x1i + x2j + x3k is defined as q = x0 − x1i − x2j − x3k ∈ Q. q= (qq)1/2 = (qq)1/2 =qx2 0 + x2 1 + x2 2 + x2 3. For every non-zero quaternion q = x0 + x1i + x2j + x3k ∈ Q, there exists a unique inverse q−1 in Q as q−1 = q q2 = x0 − x1i − x2j − x3k px2 2 + x2 0 + x2 1 + x2 3 . Definition 2.1. A right quaternionic vector space VR(Q) is a linear vector space under right scalar multiplication over the field of quaternionic Q, i.e., VR(Q) × Q → VR(Q) (2.1) and for each u, v ∈ VR(Q) and p, q ∈ Q, the right scalar multiplication (2.1) satisfying the following properties: (u, q) → uq (u + v)q = uq + vq u(p + q) = up + uq v(pq) = (vp)q. Definition 2.2. A right quaternionic pre-Hilbert space or right quaternionic inner product space VR(Q) is a right quaternionic vector space together with the binary mapping h..i : VR(Q) × VR(Q) → Q (called the Hermitian quaternionic inner product) which satisfies following properties: (a) hv1v2i = hv2v1i for all v1, v2 ∈ VR(Q). (b) hvvi > 0 if v 6= 0. (c) hvv1 + v2i = hvv1i + hvv2i for all v, v1, v2 ∈ VR(Q) (d) hvuqi = hvuiq for all v, u ∈ VR(Q) and q ∈ Q. In view of Definition 2.2, a right quaternionic inner product space VR(Q) also have the following properties: (i) hqvui = qhvui for all v, u ∈ VR(Q) and q ∈ Q (ii) pv1 + qv2 ∈ VR(Q), for all v1, v2 ∈ VR(Q) and p, q ∈ Q Frames in quaternionic Hilbert space 3 kuk=phuui, u ∈ VR(Q). Let VR(Q) be right quaternionic inner product space with the Hermitian inner product h..i. Define the quaternionic norm k.k: VR(Q) → R+ on VR(Q) by (2.2) Definition 2.3. The right quaternionic pre-Hilbert space is called a right quaternionic Hilbert space, if it is complete with respect to the norm (2.2) and is denoted by VR(Q). Theorem 2.4 (The Cauchy-Schwarz Inequality). [12] If VR(Q) is a right quaternionic Hilbert space then huvi2≤ huuihvvi, for all u, v ∈ VR(Q). Moreover, a norm defined in (2.2) satisfy the following properties: For the non-commutative field of quaternions Q, define the space ℓ2(Q) by (a) kuqk= kukq, for all u ∈ VR(Q) and q ∈ Q. (b) ku + vk≤ kuk+kvk, for all u, v ∈ VR(Q). (c) kuk= 0 for some u ∈ VR(Q), then u = 0. kqik2< +∞(cid:27) ℓ2(Q) =(cid:26){qi}i∈I ⊂ Q : Xi∈I under right multiplication by quaternionic scalars together with the quaternionic inner product on ℓ2(Q) defined as hpqi =Xi∈I piqi, p = {pi}i∈I and q = {qi} ∈ ℓ2(Q). (2.3) It is easy to observe that ℓ2(Q) is a right quaternionic Hilbert space with respect to quaternionic inner product (2.3). Definition 2.5 ([12]). Let VR(Q) be a right quaternionic Hilbert Space and S be a subset of VR(Q). Then, define the set: • S⊥ = {v ∈ VR(Q) : hvui = 0 ∀ u ∈ S}. • hSi be the right Q-linear subspace of VR(Q) consisting of all finite right Q-linear combinations of elements of S . ai as the following element Let i → ai ∈ R+, i ∈ I be a function on I . Then, define Pi∈I of R+ ∪ {+∞}: Xi∈I ai : J is a non-empty finite subset of I(cid:27). ai = sup(cid:26)Xi∈J ai < +∞, then the set of all i ∈ I such that ai 6= 0 is at most countable. In view of this, given a quaternionic Hilbert space VR(Q) and a map ui is said to be converges absolutely if Pi∈Ikuik< +∞. It is clear that, if Pi∈I i → ui ∈ VR(Q), i ∈ I , the series Pi∈I Pi∈I Theorem 2.6 ([12]). Let VR(Q) be a quaternionic Hilbert space and let N be a subset of VR(Q) such that, for z, z′ ∈ N such that hzz′i = 0 if z 6= z′ and hzzi = 1. Then the following conditions are equivalent: If this happens, then only a finite or countable number of ui is non-zero and the series ui converges to a unique element of VR(Q), independently from the ordering of u′is. u =Xz∈N zhzui where the series Pz∈N zhzui converges absolutely in VR(Q). 4 Sharma and Goel huzihzvi. huvi =Xz∈N (a) For every u, v ∈ VR(Q), the series Pz∈Nhuzihzvi converges absolutely and (b) For every u ∈ VR(Q), kuk2=Pz∈Nhzui2 . (c) N⊥ = 0. (d) hNi is dense in H . Definition 2.7 ([12]). Every quaternionic Hilbert space VR(Q) admits a subset N , called Hilbert basis or orthonormal basis of VR(Q), such that, for z, z′ ∈ N , hzz′i = 0 if z 6= z′ and hzzi = 1 and satisfies all the conditions of Theorem 2.6. Further, if there are two such sets, then they have the same cardinality. Furthermore, if N is a Hilbert basis of VR(Q), then for every u ∈ VR(Q) can be uniquely expressed as Definition 2.8 ([1]). Let VR(Q) be a right quaternionic Hilbert space and T be an operator on VR(Q). Then T is said to be • right Q-linear if T (v1α+v2β) = T (v1)α+T (v2)β, for all v1, v2 ∈ VR(Q) and α, β ∈ Q. • bounded if there exist K ≥ 0 such that kT (v)k≤ Kkvk, for all v ∈ VR(Q). Definition 2.9 ([1]). Let VR(Q) be a right quaternionic Hilbert space and T be an operator on VR(Q). Then the adjoint operator T ∗ of T is defined by hvT ui = hT ∗vui, for all u, v ∈ VR(Q). Further, T is said to be self-adjoint if T = T ∗ . Theorem 2.10 ([1]). Let VR(Q) be a right quaternionic Hilbert space and S & T be two bounded right linear operators on VR(Q). Then (a) T + S and T S ∈ B(VR(Q)). Moreover: kT + Sk≤ kTk+kSk and kT Sk≤ kTkkSk (b) hT vui = hvT ∗ui. (c) (T + S)∗ = T ∗ + S∗ . (d) (T S)∗ = S∗T ∗ . (e) (T ∗)∗ = T. (f ) I∗ = I , where I is an identity operator on VR(Q). (g) If T is an invertible operator then (T −1)∗ = (T ∗)−1 . Theorem 2.11 ([12]). Let VR(Q) be a right quaternionic Hilbert space and let T ∈ B(VR(Q)) be an operator. If T ≥ 0, then there exists a unique operator in B(VR(Q)), indicated by √T , such that √T ≥ 0 and √T√T = T . Furthermore, it turns out that √T commutes with every operator which commutes with T and if T is invertible and self-adjoint, then √T is also invertible and self-adjoint. Theorem 2.12 ([9]). Let VR(Q) be a right quaternionic Hilbert space and T ∈ B(VR(Q)) and let s ∈ Q be such that kTk< s. Then the operator (s−1T )ns−1I Xn≥0 Frames in quaternionic Hilbert space 5 is the right and left algebraic inverse of (sI − T ) and the series converges in the operator norm. 3. Frames in Quaternionic Hilbert space We begin this section with the following definition of frames in right quaternionic Hilbert spaces VR(Q). Definition 3.1. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a sequence in VR(Q). Then {ui}i∈I is said to be a frame for VR(Q), if there exist two finite constants with 0 < A ≤ B such that (3.1) Akuk2≤Xi∈I huiui2≤ Bkuk2, for all u ∈ VR(Q). The positive constants A and B , respectively, are called lower frame and upper frame bounds for the frame {ui}i∈I . The inequality (3.1) is called frame inequality for the frame {ui}i∈I . A sequence {ui}i∈I is called Bessel sequence for right quaternionic Hilbert space VR(Q) with bound B , if {ui}i∈I satisfies the right hand side of the inequality (3.1). A frame {ui}i∈I for right quaternionic Hilbert space VR(Q) is said to be • tight, if it is possible to choose A and B satisfying inequality (3.1) with A = B . • Parseval frame, if it is tight with A = B = 1. • exact, if it ceases to be a frame whenever anyone of its element is removed. Regarding the existence of frames in right quaternionic Hilbert space VR(Q), we have the following examples: Example 3.2. Let N be a Hilbert basis for right quaternionic Hilbert space VR(Q) such that, for each zi, zk ∈ N , i, k ∈ I , we have hzizki =(0, for i 6= k 1, for i = k. (1) Tight and Non-Exact. Let {ui}i∈I be a sequence in VR(Q) defined as ui = ui+1 = zi, i ∈ I. Then {ui}i∈I is tight and non-exact frame for VR(Q) with bound A = 2. Indeed, we have Xi∈I huiui2 = 2Xi∈I hziui2 = 2kuk2, for all u ∈ VR(Q). (2) Non-Tight and Non-exact. Let {ui}i∈I be a sequence in VR(Q) defined as Then {ui}i∈I is non-tight and non-exact frame for VR(Q). Indeed, we have huiui2≤ 2kuk2, for all u ∈ VR(Q). (3) Parseval. Let {ui}i∈I be a sequence in VR(Q) defined as u1 = z1 uik = uik+1 = uik+2 = . . . = uik+1−1 = zi√i , (u1 = z1 ui = zi−1, i ≥ 2, i ∈ I. kuk2≤Xi∈I   6 Sharma and Goel where ik = ik−1 + (k − 1), k ∈ N, i0 = 1. Then {ui}i∈I is a Parseval frame for VR(Q). Indeed, we have Xi∈I = kuk2, for all u ∈ VR(Q). huiui2=Xi∈I 2 i(cid:12)(cid:12)(cid:12)(cid:12) (cid:28) zi√i(cid:12)(cid:12)(cid:12)(cid:12) u(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) for VR(Q). (4) Exact. Let {zi}i∈I be a Hilbert basis of VR(Q). Then {zi}i∈I is an exact frame i ukqk − j Xk=1 i Xk=1 (cid:13)(cid:13)(cid:13)(cid:13) ukqk(cid:13)(cid:13)(cid:13)(cid:13) uiqi converges unconditionally. Next, we show that for a sequence {ui}i∈I in right quaternionic Hilbert space VR(Q) being a Bessel sequence is a sufficient condition for the series Pi∈I uiqi , {qi}i∈I ⊂ ℓ2(Q) to converge unconditionally. Theorem 3.3. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a Bessel sequence for VR(Q) with Bessel bound B . Then, for every sequence {qi}i∈I ∈ ℓ2(Q), the series Pi∈I Proof. For i, j ∈ I, i > j . Then, we have ukqk(cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) Xk=j+1 (cid:28) i kvk=1(cid:12)(cid:12)(cid:12)(cid:12) ukqkv(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) Xk=j+1 Xk=j+1(cid:12)(cid:12)(cid:12)(cid:12) qkhukvi(cid:12)(cid:12)(cid:12)(cid:12) ≤ (cid:18) i kgk=1(cid:18) i qk2(cid:19)1/2 Xk=j+1 Xk=j+1 √B(cid:18) i qk2(cid:19)1/2 Xk=j+1 ≤ Pk=1qk2(cid:27)i∈I Since {qi}i∈I ∈ ℓ2(Q) and (cid:26) i (cid:26) i Pk=1 If view of Theorem 3.3, if {ui}i∈I is a Bessel sequence for VR(Q). Then, the (right) synthesis operator for {ui}i∈I is a right linear operator T : ℓ2(Q) → VR(Q) defined by uiqi(cid:27) unconditionally Hence (cid:26)Pi∈I is a Cauchy sequence in R. Therefore hukvi2(cid:19)1/2 is Cauchy sequence in VR(Q). ukqk(cid:27)i∈I convergent in VR(Q). sup . = sup i = sup kvk=1 (cid:3) T ({qi}i∈I) =Xi∈I uiqi, {qi} ∈ ℓ2(Q). The adjoint operator T ∗ of right synthesis operator T is called the (right) analysis operator. Further, the analysis operator T ∗ : VR(Q) → ℓ2(Q) is given by T ∗(u) = {huiui}i∈I, u ∈ VR(Q). Frames in quaternionic Hilbert space 7 Infact, for u ∈ VR(Q) and {qi}i∈I ∈ ℓ2(Q), we have hT ∗(u){qi}i∈Ii = huT ({qi}i∈I)i uiqi(cid:29) huuiiqi = (cid:28)u(cid:12)(cid:12)(cid:12)(cid:12)Xi∈I = Xi∈I = (cid:28){huiui}i∈I,{qi}i∈I(cid:29). Thus T ∗(u) = {huiui}i∈I, u ∈ VR(Q). Next, we give a characterization for a Bessel sequence in a right quaternionic Hilbert space. Theorem 3.4. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a sequence in VR(Q). Then, {ui}i∈I is a Bessel sequence for VR(Q) with bound B if and only if the right linear operator T : ℓ2(Q) → VR(Q) defined by T ({qi}i∈I) =Xi∈I uiqi, {qi}i∈I ∈ ℓ2(Q) is a well defined and bounded operator with kTk≤ √B . Proof. Let {ui}i∈I be a Bessel sequence for right quaternionic Hilbert space VR(Q). Then, by Theorem 3.3, the T is a well defined and bounded operator with kTk≤ √B . Conversely, let T be a well defined and bounded right linear operator with kTk≤ √B . Then, the adjoint of a bounded right linear operator T is itself bounded and kTk= kT ∗k. Since, for u ∈ VR(Q), we have Xi∈I huiui2 = kT ∗(u)k2 ≤ kT ∗k2kuk2 = kTk2kuk2, it follows that {ui}i∈I is a Bessel sequence for right quaternionic Hilbert space VR(Q) with bound B . (cid:3) Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q). Then, the (right) frame operator S : VR(Q) → VR(Q) for frame {ui}i∈I is a right linear operator given by S(u) = T T ∗(u) = T ({huiui}i∈I) = Xi∈I uihuiui, u ∈ VR(Q). In next result, we discuss some properties of the frame operator for a frame in right quaternionic Hilbert space. 8 Sharma and Goel Theorem 3.5. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with lower and upper frame bounds A and B , respectively and frame operator S . Then S is positive, bounded, invertible and self adjoint right linear operator on VR(Q). Proof. For any u ∈ VR(Q), we have hSuui = (cid:28)Xi∈I = Xi∈I = Xi∈I uihuiui, u(cid:29) huiuihuiui huiui2. This gives Akuk2≤ hSuui ≤ Bkuk2, u ∈ VR(Q). Thus AI ≤ S ≤ BI. (3.2) Hence S is positive and bounded right linear operator on VR(Q). Also, 0 ≤ I − B−1S ≤ B − A I and consequently B kI − B−1Sk = sup kvk=1(cid:12)(cid:12)(cid:12)(cid:12) B − A (cid:28)(I − B−1S)v(cid:12)(cid:12)(cid:12)(cid:12) v(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) ≤ < 1. B Then, S is invertible. Further for any u, v ∈ VR(Q), we have hSuvi = (cid:28)Xi∈I = Xi∈I = Xi∈I = (cid:28)u(cid:12)(cid:12)(cid:12)(cid:12)Xi∈I = huSvi. uihuiui(cid:12)(cid:12)(cid:12)(cid:12) v(cid:29) huiuihuivi huuiihuivi uihuivi(cid:29) Thus S is also a self adjoint right linear operator on VR(Q). (cid:3) Corollary 3.6 (The Reconstruction Formula). Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with frame operator S . Then, for every u ∈ VR(Q) can be expressed as S−1uihuiui. u =Xi∈I Frames in quaternionic Hilbert space Proof. As S is invertible. Therefore, for u ∈ VR(Q), we have u = S−1S(u) = Xi∈I S−1uihuiui. 9 (cid:3) In the next result, we construct a frame with the help a give frame in a right quaternionic Hilbert space. Theorem 3.7. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with lower and upper frame bounds A and B , respectively and frame operator is also a frame for VR(Q) with bounds B−1 and A−1 and right S . Then {S−1ui}i∈I frame operator S−1 . Proof. For u ∈ VR(Q), we have hS−1uiui2 = Xi∈I huiS−1ui2 Xi∈I ≤ BkS−1uk2 ≤ BkS−1k2kuk2. Thus {S−1ui}i∈I is a Bessel sequence for VR(Q). It follows that the right frame operator for {S−1ui}i∈I is well defined. Therefore, we have Xi∈I S−1uihS−1uiui = S−1(cid:18)Xi∈I uihS−1uiui(cid:19) = S−1S(S−1u) = S−1u, u ∈ VR(Q). (3.3) Thus, the right frame operator for {S−1ui}i∈I is S−1 . The operator S−1 commutes with both S and I : VR(Q) → VR(Q) (an identity operator on VR(Q)). Therefore, multiplying the inequality (3.2) with S−1 , we have i.e. B−1I ≤ S−1 ≤ A−1I B−1kuk2≤ hS−1uui ≤ A−1kuk2, u ∈ VR(Q). By using (3.3), we get B−1kuk2≤Xi∈I hS−1uiui2≤ A−1kuk2, u ∈ VR(Q). Therefore {S−1ui}i∈I is frame for VR(Q) with bounds S−1 . 1 B and 1 A and right frame operator (cid:3) In view of Theorem 3.7, we have a following definition: Definition 3.8. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with frame operator S . Then, the frame {S−1ui}i∈I is called the canonical dual of the frame {ui}i∈I . Next, we construct a Parseval frame with the help of a give frame in a right quaternionic Hilbert space. 10 Sharma and Goel Theorem 3.9. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with frame operator S . Then {S−1/2ui}i∈I is a Parseval frame for VR(Q). Proof. By Theorem 2.11, for any u ∈ VR(Q), we have u = S−1/2SS−1/2u uihuiS−1/2ui S−1/2uihuiS−1/2ui. Therefore, for any u ∈ VR(Q) = S−1/2Xi∈I = Xi∈I kuk2 = huui = (cid:28)Xi∈I = Xi∈I = Xi∈I S−1/2uihuiS−1/2ui(cid:12)(cid:12)(cid:12)(cid:12) u(cid:29) huiS−1/2uihS−1/2uiui hS−1/2uiui2. (3.4) (by using (3.4)) . Thus {S−1/2ui}i∈I is a Parseval frame for VR(Q). Next, we give a characterization of Parseval frames {ui}i∈I Hilbert space VR(Q). Theorem 3.10. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with frame operator S . Then {ui}i∈I is a Parseval frame for VR(Q) if and only if S is an identity operator on VR(Q). Proof. Let {ui}i∈I be a Parseval frame for VR(Q). Then, for all u ∈ VR(Q), we have (cid:3) for right quaternionic huiui2 = kuk2 Xi∈I which implies hSuui = huui. Thus, S is an identity operator on VR(Q). Conversely, let S is an identity operator on VR(Q). Thus, for u ∈ VR(Q), we have u = S(u) So, we have uihuiui. = Xi∈I kuk2 = huui = (cid:28)Xi∈I = Xi∈I uihuiuiu(cid:29) huiui2. (3.5) Frames in quaternionic Hilbert space Hence {ui}i∈I is a Parseval frame for right quaternionic Hilbert space VR(Q). 11 (cid:3) In the next theorem, we give a necessary condition for a Parseval frame {ui}i∈I for right quaternionic Hilbert space VR(Q). Theorem 3.11. Let VR(Q) be a right quaternionic Hilbert space and {qi}i∈I be a semi- normalized sequence of quaternions in Q with bounds a and b. If {uiqi}i∈I is a Parseval frame for VR(Q), then {ui}i∈I is a frame for VR(Q) with bounds b−2 and a−2 . Proof. Let {uiqi}i∈I be a Parseval frame for VR(Q). Then, for any u ∈ VR(Q), we have huiqiui2= kuk2. Xi∈I This gives qi2huiui2= kuk2. Xi∈I (3.6) Since {qi}i∈I is a normalized sequence with bounds a and b, we get huiui2, u ∈ VR(Q). qi2huiui2≤ b2Xi∈I a2Xi∈I huiui2≤Xi∈I Therefore, by using (3.5), we have 1 b2kuk2≤Xi∈I huiui2≤ 1 a2 , u ∈ VR(Q). Hence {ui}i∈I is a frame for VR(Q) with bounds b−2 and a−2 . (cid:3) Finally, in this section we give a characterization of frame for right quaternionic Hilbert space in terms of operators. Theorem 3.12. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a sequence in VR(Q). Then {ui}i∈I is a frame for VR(Q) if and only if the right linear operator T : ℓ2(Q) → VR(Q) T ({qi}i∈I) =Xi∈I uiqi, {qi}i∈I ∈ VR(Q) is a well-defined and bounded mapping from ℓ2(Q) onto VR(Q). Proof. Let {ui}i∈I be a frame for VR(Q). Then the frame operator S = T T ∗ for {ui}i∈I is invertible on VR(Q). So T is an onto mapping. Also, by Theorem 3.4, T is well defined and bounded mapping from ℓ2(Q) to VR(Q). Conversely, let T be a well defined, bounded right linear operator from ℓ2(Q) onto VR(Q). Then, by Theorem 3.4, {ui}i∈I is a Bessel sequence for VR(Q). Since T is an onto, there exists an right linear operator T † : VR(Q) → ℓ2(Q) such that u = T T †u = Xi∈I ui(T †u)i, u ∈ VR(Q), 12 Sharma and Goel where (T †u)i denotes the i−th coordinate of T †u. Therefore, for u ∈ VR(Q), we have 2 kuk4 = huui2 ui(T †u)i(cid:12)(cid:12)(cid:12)(cid:12) u(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12) (cid:28)Xi∈I (T †u)i2 Xi∈I ≤ Xi∈I ≤ kT †k2kuk2Xi∈I huiui2. huiui2 Hence {ui}i∈I is a frame for VR(Q) with bounds kT †k−2 and kTk2 . (cid:3) 4. Stability of Frames in Right Quaternionic Hilbert Space Christensen [7] gave a version of Paley-Wiener Theorem for frames in Hilbert spaces. In this section, we give a similar version for frames in right quaternionic Hilbert spaces. Theorem 4.1. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with lower and upper frame bounds A and B , respectively. Let {vi}i∈I be a sequence in VR(Q) and assume that there exist λ, µ ≥ 0 such that (cid:18)λ + √A(cid:19) < 1 and µ (4.1) µ . qi2(cid:19)1/2 (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J viqi(cid:13)(cid:13)(cid:13)(cid:13) µ √B(cid:19)(cid:19)2 (ui − vi)qi(cid:13)(cid:13)(cid:13)(cid:13) ≤ λ(cid:13)(cid:13)(cid:13)(cid:13)Xi∈J uiqi(cid:13)(cid:13)(cid:13)(cid:13) + µ(cid:18)Xi∈J for all finite quaternions qi ∈ Q, i ∈ J ⊆ I with J< +∞. Then {vi}i∈I is a frame for √A(cid:19)(cid:19)2 and B(cid:18)1 +(cid:18)λ + VR(Q) with bounds A(cid:18)1 −(cid:18)λ + Proof. Let J ⊆ I with J< +∞. Then uiqi(cid:13)(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13)(cid:13)Xi∈J (ui − vi)qi(cid:13)(cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J qi2(cid:19)1/2 uiqi(cid:13)(cid:13)(cid:13)(cid:13) ≤ (1 + λ)(cid:13)(cid:13)(cid:13)(cid:13)Xi∈J + µ(cid:18)Xi∈J uiqi(cid:12)(cid:12)(cid:12)(cid:12) kvk=1(cid:12)(cid:12)(cid:12)(cid:12) v(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) (cid:28)Xi∈J qi2(cid:19)1/2 ≤ (cid:18)Xi∈J kvk=1(cid:18)Xi∈J qi2(cid:19)1/2 √B(cid:18)Xi∈J huivi2(cid:19)1/2 (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J uiqi(cid:13)(cid:13)(cid:13)(cid:13) sup . = sup ≤ Since . This gives Frames in quaternionic Hilbert space 13 ≤(cid:18)(1 + λ)√B + µ(cid:19)(cid:18)Xi∈J qi2(cid:19)1/2 . (4.2) viqi(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J T ({qi}i∈I) =Xi∈J Let T : ℓ2(Q) → VR(Q) be a right linear operator defined by viqi, {qi}i∈I ∈ ℓ2(Q). Then, by (4.2), T is well defined and bounded with kTk≤ (cid:18)(1 + λ)√B + µ(cid:19). Therefore, by Theorem 3.4, {vi}i∈I is a Bessel sequence for VR(Q) with the upper bound √B(cid:19)(cid:19)2 B(cid:18)1 +(cid:18)λ + Let U be the right synthesis operator for the frame {ui}i∈I . Define an operator W : VR(Q) → ℓ2(Q) by µ . W (u) = U∗(U U∗)−1u, u ∈ VR(Q). Since {(U U∗)−1ui}i∈I is also a frame for VR(Q) with bounds 1 B and 1 A , then kW (u)k2 = Xi∈I hS−1uiui2 1 Akuk2, u ∈ VR(Q). ≤ Using (4.1) with {qi}i∈I = W u, u ∈ VR(Q), we get ku − T W uk ≤ λkuk+µkW uk ≤ (cid:18)λ + µ √A(cid:19)kuk, u ∈ VR(Q). . Since (cid:18)λ + µ This gives kT Wk≤ 1 + λ + invertible and k(T W )−1k≤ µ √A 1 −(cid:18)λ + Now, for any u ∈ VR(Q), we have 1 . µ √A(cid:19) u = (T W )(T W )−1u = Xi∈I vih(U U∗)−1ui(T W )−1ui. √A(cid:19) < 1, then the operator T W is 14 Sharma and Goel Therefore, for each u ∈ VR(Q), we have 2 vih(U U∗)−1ui(T W )−1ui(cid:12)(cid:12)(cid:12)(cid:12) u(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) kuk4 = (cid:12)(cid:12)(cid:12)(cid:12) (cid:28)Xi∈I 2(cid:18)Xi∈I h(U U∗)−1ui(T W )−1ui(cid:12)(cid:12)(cid:12)(cid:12) ≤ Xi∈I (cid:12)(cid:12)(cid:12)(cid:12) hviui2(cid:19) A(cid:20) hviui2(cid:19). kuk2(cid:18)Xi∈I 1 − (λ + µ√A )(cid:21)2 ≤ 1 1 This gives hviui2≥ A(cid:18)1 −(cid:18)λ + µ √A(cid:19)(cid:19)2 Xi∈I kuk2, u ∈ VR(Q). Hence {vi}i∈I is a frame for VR(Q) with desire bounds. Corollary 4.2. Let VR(Q) be a right quaternionic Hilbert space and {ui}i∈I be a frame for VR(Q) with lower and upper frame bounds A and B , respectively. Let {vi}i∈I be a sequence in VR(Q) and assume that there exist 0 < R < A such that (cid:3) (ui − vi)qik≤ kXi∈I √R(cid:18)Xi∈I qi2(cid:19)1/2 , for all {qi}i∈I ∈ ℓ2(Q). (4.3) Then {vi}i∈I is also a frame for VR(Q) with bounds (√A − √R)2 and (√B + √R)2 . Proof. Take λ = 0 and µ = √R in Theorem 4.1. √A(cid:19) < 1 can not be dropped in the Theorem 4.1. In Remark 4.3. A condition (cid:18)λ + Example 4.4. Let VR(Q) be a right quaternionic Hilbert space, {zi}i∈N be a Hilbert basis of VR(Q), {pi}i∈N be a sequence in Q and {vi}i∈N be a sequence in VR(Q) such that this, regard, we give the following example: µ (cid:3) vi = zi + zi+1pi, i ∈ N. Then, for J ⊆ N with J< ∞ and {qi}i∈N ∈ ℓ2(Q), we have (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J (vi − zi)qi(cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J ≤ sup i zi+1piqi(cid:13)(cid:13)(cid:13)(cid:13) pi(cid:18)Xi∈I qi2(cid:19)1/2 . Thus, if p = sup (1 − p)2 and (1 + p)2 . On taking pi = 1, for all i ∈ N, we get pi< 1, then by Theorem 4.1, {ui}i∈N is a frame for VR(Q) with bounds i vi = zi + zi+1, i ∈ N. (4.4) Frames in quaternionic Hilbert space 15 Then, for any sequence {qi}i∈N ∈ ℓ2(Q) and for J ⊆ N with J< ∞ (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J (vi − zi)qi(cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13)Xi∈J ≤ (cid:18)Xi∈I zi+1qi(cid:13)(cid:13)(cid:13)(cid:13) qi2(cid:19)1/2 . Thus, the condition (4.1) is satisfied with either (λ, µ) = (1, 0) or (0, 1). So, (cid:18)λ+ 1 and {vi}i∈N is not a frame for VR(Q). µ √A(cid:19) ≮ References [1] S.L. Adler, Quaternionic Quantum Mechanics and Quantum Fields, Oxford University Press, New York, 1995. [2] R. Balan, P.G. Casazza and D. Edidin, On signal reconstruction without phase, Appl. Comp. Harm. Anal., 20 (2006), 345-356. [3] J. Benedetto, A. Powell and O. Yilmaz, Sigma-Delta quantization and finite frames, IEEE Trans. Inform. Theory, 52 (2006), 1990-2005. [4] H. Bolcskel, F. Hlawatsch and H.G. Feichtinger, Frame - theoretic analysis of over sampled filter banks, IEEE Trans.Signal Process. 46 (1998), 3256-3268. [5] P.G. Casazza, The art of frame theory, Taiwanese J. of Math., 4 (2)(2000), 129-201. [6] P.G. Casazza and G. Kutyniok, Frames of subspaces, in Wavelets, Frames and Operator Theory (College Park, MD, 2003), Contemp. Math., 345, Amer. Math. Soc., Providence, RI, 2004, 87- 113. [7] O. Christensen, A Paley-Wiener theorem for frames, Proc. Amer. Math. Soc., 123 (1995), 2199-2202. [8] O. Christensen, An introduction to Frames and Riesz Bases, Birkhauser, 2003. [9] F. Colombo, I. Sabadini and D.C. Struppa: Noncommutative Functional Calculus, Theory and applications of slice hyperholomorphic functions, Progress in Mathematics, vol. 289, Birkhauser/Springer Basel AG, Basel, 2011. [10] I. Daubechies, A. Grossmann and Y. Meyer, Painless non-orthogonal expansions, J. Math. Physics, 27 (1986), 1271-1283. [11] R.J. Duffin and A.C. Schaeffer, A class of non-harmonic Fourier series, Trans. Amer. Math. Soc., 72 (1952), 341-366. [12] R. Ghiloni, , V. Moretti and A. Perotti, Continuous slice functional calculus in quaternionic Hilbert spaces, Rev. Math. Phys. 25 (2013), 1350006. [13] R.W. Heath and A.J. Paulraj, Linear dispersion codes for MIMO systems based on frame theory, IEEE Trans.Signal Process. 50 (2002), 2429- 2441. [14] M. Khokulan, K. Thirulogasanthar and S. Srisatkunarajah, Discrete frames on finite dimensional quaternion Hilbert spaces, arXiv: 1302.2836v1 [15] S.K. Sharma and Virender, Dual frames on finite dimensioal quaternionic Hilbert space, Poincare J. Anal. Appl., 2016 (2), 79-88. (S.K. Sharma) Department of Mathematics, Kirori Mal College, University of Delhi, Delhi 110 007, INDIA. E-mail address: [email protected] (Shashank Goel) Department of Mathematics, Amity Institute of Applied Sciences, Amity University, Noida, U.P-201301, INDIA. E-mail address: [email protected]
1306.0724
1
1306
2013-06-04T10:27:24
Wandering subspaces of the Bergman space and the Dirichlet space over polydisc
[ "math.FA" ]
Doubly commutativity of invariant subspaces of the Bergman space and the Dirichlet space over the unit polydisc $\mathbb{D}^n$ (with $ n \geq 2$) is investigated. We show that for any non-empty subset $\alpha=\{\alpha_1,\dots,\alpha_k\}$ of $\{1,\dots,n\}$ and doubly commuting invariant subspace $\s$ of the Bergman space or the Dirichlet space over $\D^n$, the tuple consists of restrictions of co-ordinate multiplication operators $M_{\alpha}|_\s:=(M_{z_{\alpha_1}}|_\s,\dots, M_{z_{\alpha_k}}|_\s)$ always possesses wandering subspace of the form \[\bigcap_{i=1}^k(\s\ominus z_{\alpha_i}\s). \]
math.FA
math
WANDERING SUBSPACES OF THE BERGMAN SPACE AND THE DIRICHLET SPACE OVER POLYDISC A. CHATTOPADHYAY, B. KRISHNA DAS, JAYDEB SARKAR, AND S. SARKAR Abstract. Doubly commutativity of invariant subspaces of the Bergman space and the Dirichlet space over the unit polydisc Dn (with n ≥ 2) is investigated. We show that for any non-empty subset α = {α1, . . . , αk} of {1, . . . , n} and doubly commuting invariant subspace S of the Bergman space or the Dirichlet space over Dn, the tuple consists of restrictions of co- S) always possesses wandering ordinate multiplication operators MαS := (Mzα1 S , . . . , Mzα subspace of the form k (S ⊖ zαiS). k \i=1 1. Introduction A closed subspace W of a Hilbert space H is said to be wandering subspace (following Halmos [Hal]) for an n-tuple of commuting bounded linear operators T = (T1, . . . , Tn) (n ≥ 1) if for all (l1, . . . , ln) ∈ Nn \ {(0, . . . , 0)} and W ⊥ T l1 1 T l2 2 · · · T ln n W H = span{T l1 1 T l2 2 · · · T ln n h : h ∈ W, l1, . . . , ln ∈ N}. In this case, the tuple T is said to have the wandering subspace property. The main purpose of this paper is to investigate the following question: Question: Let (T1, . . . , Tn) be a commuting n-tuple of bounded linear operators on a Hilbert space H. Does there exists a wandering subspace W for (T1, . . . , Tn)? This question has an affirmative answer for the restriction of multiplication operator by the coordinate function Mz, to an invariant subspace of the Hardy space H 2(D) (Beurlings theorem [Beu]) or the Bergman space A2(D) (Aleman, Richter and Sundberg [ARS]) or the Dirichlet space D(D) (Richter [Ric] ) over the unit disc D of the complex plane C. For n ≥ 2, existence of wandering subspaces for general invariant subspaces of the Hardy space H 2(Dn) over the unit polydisc Dn rather fails spectacularly (cf. [Rud]). 2010 Mathematics Subject Classification. 47A13, 47A15, 47A20, 47L99. Key words and phrases. Invariant subspace, Beurling's theorem, Bergman space, Dirichlet space, Hardy space, Doubly commutativity. 1 2 CHATTOPADHYAY, DAS, SARKAR, AND SARKAR Recall that the Hardy space over the unit polydisc Dn = {z = (z1, . . . , zn) ∈ Cn : zi < 1, i = 1, . . . , n} is denoted by H 2(Dn) and defined by H 2(Dn) = {f ∈ O(Dn) : sup 0≤r<1ZTn f (rz)2dθ < ∞}, where dθ is the normalized Lebesgue measure on the torus Tn, the distinguished boundary of Dn, rz := (rz1, . . . , rzn) and O(Dn) denotes the set of all holomorphic functions on Dn (cf. [Rud]). In [Beu], A. Beurling characterize all closed Mz-invariant subspaces of H 2(D) in the follow- ing sense: Let S 6= {0} be a closed subspace of H 2(D). Then S is Mz-invariant if and only if S = θH 2(D) for some inner function θ (that is, θ ∈ H ∞(D) and θ = 1 a.e. on the unit circle T). In particular, Beurlings theorem yields the wandering subspace property for Mz-invariant subspaces of H 2(D) as follows: if S = θH 2(D) is an Mz-invariant subspace of H 2(D) then (1) ⊕zmW, S =Xm≥0 where W is the wandering subspace for MzS given by W = S ⊖ zS = θH 2(D) ⊖ zθH 2(D) = θC. In [Rud], W. Rudin showed that there are invariant subspaces M of H 2(D2) which does not contain any bounded analytic function. In particular, the Beurling like characterization of (Mz1, . . . , Mzn)-invariant subspaces of H 2(Dn), in terms of inner functions on Dn is not possible. On the other hand, the wandering subspace (and the Beurlings theorem) for the shift invariant subspaces of the Hardy space H 2(D) follows directly from the classical Wold decom- position [Wol] theorem for isometries. Indeed, for a closed Mz-invariant subspace S(6= {0}) of H 2(D), Consequently, by Wold decomposition theorem for the isometry MzS on S we have ∞ \m=0 S =Xm≥0 (MzS)mS = M m z S ⊆ M m z H 2(D) = {0}. ∞ \m=0 ∞ \m=0 ⊕zmW ⊕ ( ∞ \m=0 (MzS)mS) =Xm≥0 ⊕zmW, where W = S ⊖ zS and hence, (1) follows. Therefore, the notion of wandering subspaces is stronger (as well as of independent interest) than the Beurlings characterization of Mz- invariant subspaces of H 2(D). To proceed, we now recall the following definition: a commuting n-tuple (n ≥ 2) of bounded j Ti for all linear operators (T1, . . . , Tn) on H is said to be doubly commuting if TiT ∗ 1 ≤ i < j ≤ n. j = T ∗ Natural examples of doubly commuting tuple of operators are the multiplication operators by the coordinate functions acting on the Hardy space or the Bergman space or the Dirichlet space over the unit polydisc Dn (n ≥ 2). WANDERING SUBSPACES OF THE BERGMAN SPACE AND THE DIRICHLET SPACE OVER Dn 3 Let H ⊆ O(Dn) be a reproducing kernel Hilbert space over Dn (see [Aro]) and multi- plication operators {Mz1, . . . , Mzn} by coordinate functions are bounded. Then a closed (Mz1, . . . , Mzn)-invariant subspace S of H is said to be doubly commuting if the n-tuple (Mz1S, . . . , MznS) is doubly commuting, that is, RiR∗ j Ri for all 1 ≤ i < j ≤ n, where Ri = MziS. j = R∗ In [SSW], third author and Sasane and Wick proved that any doubly commuting invariant subspace of H 2(Dn) (where n ≥ 2) has the wandering subspace property (see [Man] for n = 2). Also in [ReT], Redett and Tung obtained the same conclusion for doubly commuting invariant subspaces of the Bergman space A2(D2) over the bidisc D2. In this paper we prove that doubly commuting invariant subspaces of the Bergman space A2(Dn) and the Dirichlet space D(Dn) has the wandering subspace property. Our result on the Bergman space over polydisc is a generalization of the base case n = 2 in [ReT]. Our analysis is based on the Wold-type decomposition result of S. Shimorin for operators closed to isometries [Shi]. The paper is organized as follows: in Section 2, we obtain some general results concerning the wandering subspaces of tuples of doubly commuting operators. We obtain wandering subspaces for doubly commuting shift invariant subspaces of the Bergman and Dirichlet spaces over Dn in Section 3. 2. Wandering subspace for tuple of doubly commuting operators In this section we prove the multivariate version of the S. Shimorin's result for tuple of doubly commuting operators on a general Hilbert space. We show that for a tuple of doubly commuting operators T = (T1, . . . , Tn) on a Hilbert space H, if for any reducing subspace Si of Ti the subspace Si ⊖ TiSi is a wandering subspace for TiSi, i = 1, . . . , n, then (H ⊖ TiH) n Ti=1 is a wandering subspace for T . We fix for the rest of the paper a natural number n ≥ 2 and set Λn := {1, . . . , n}. For a closed subset K of a Hilbert space H, an n-tuple of commuting operators T = (T1, . . . , Tn) on H and a non-empty subset α = {α1, . . . , αk} ⊆ Λn, we write [K]Tα to denote the smallest closed joint Tα := (Tα1, . . . , Tαk)-invariant subspace of H containing K. In other words (2) [K]Tα = ∞ _l1,l2,...,lk=0 T l1 α1T l2 α2 · · · T lk αk (K). If α is a singleton set {i} then we simply write [K]Ti. For a non-empty subset α = {α1, . . . , αk} ⊆ Λn, we also denote by Wα the following subspace of H, (3) Wα = (H ⊖ TαiH). k \i=1 4 CHATTOPADHYAY, DAS, SARKAR, AND SARKAR Again if α is a singleton set {i} then we simply write Wi. Thus with the above notation Wα = Tαi∈α Wαi. A bounded linear operator T on a Hilbert space H is analytic if concave if it satisfies the following inequality kT 2xk2 + kxk2 ≤ 2kT xk2 (x ∈ H). T nH = {0} and it is ∞ Tn=0 Multiplication by coordinate functions on the Dirichlet space over the unit polydisc are con- cave operators as we show in the next section. For a single bounded operator T on a Hilbert space H, the following result ensure the existence of the wandering subspace under certain conditions (see [Ric], [Shi]). Theorem 2.1. (Richter, Shimorin). Let T be an analytic operator on a Hilbert space H satisfies one of the following properties: (i) k T x + y k2≤ 2(k x k2 + k T y k2) (x, y ∈ H), (ii) T is concave. Then H ⊖ T H is a wandering subspace for T , that is, H = [H ⊖ T H]T . The following proposition is essential in order to generalize the above result for certain tuples of commuting operators. Proposition 2.2. Let T = (T1, . . . , Tn) be a doubly commuting tuple of operators on H. Then Wα is Tj-reducing subspace for all non-empty subset α ⊆ Λn and j /∈ α, where Wα is as in (3). Proof. Let α = {α1, . . . , αk} be a non-empty subset of Λn and j /∈ α. First note that Wl = Ker T ∗ l for all 1 ≤ l ≤ n and therefore Wα = y ∈ H. By doubly commutativity of T we have, k Ti=1 Ker T ∗ αi. Let x ∈ Wα, αi ∈ α and hTjx, Tαiyi = hT ∗ αiTjx, yi = hTjT ∗ αix, yi = 0. Therefore, TjWα ⊥ TαiH and hence TjWα ⊆ Wαi for all αi ∈ α. Thus Wα is an invariant subspace for Tj. Also, by commutativity of T we have for all αi ∈ α and y ∈ H. This implies T ∗ This completes the proof. hT ∗ j x, Tαiyi = hT ∗ j x, yi = hT ∗ αiT ∗ j Wα ⊥ TαiH and then T ∗ αix, yi = 0, j T ∗ j Wα ⊆ Wαi for all αi ∈ α . (cid:3) Now we prove the main theorem in the general Hilbert space operator setting. Below for a set α, we denote by #α the cardinality of α. WANDERING SUBSPACES OF THE BERGMAN SPACE AND THE DIRICHLET SPACE OVER Dn 5 Theorem 2.3. Let T = (T1, . . . , Tn) be a doubly commuting tuple of operators on H such that for any reducing subspace Si of Ti, the subspace Si ⊖ TiSi is a wandering subspace for TiSi, i = 1, . . . , n. Then for each non-empty subset α = {α1, . . . , αk} ⊆ Λn, the tuple Tα = (Tα1, . . . , Tαk ) has the wandering subspace property. More- over, the corresponding wandering subspace is given by Wα = (H ⊖ TαiH). k \i=1 Proof. First note that we only need to show H = [Wα]Tα for any non-empty subset α of Λn as the orthogonal property for wandering subspace is immediate. Now for a singleton set α the result follows form the assumption that Wi is a wandering subspace for Ti, i = 1, . . . , n. Now = Wα\{αi} for any αi ∈ α. Because for αi, αj ∈ α, for #α ≥ 2, it suffices to show that [Wα]Tαi one can repeat the procedure to get [Wα]T{αi,αj } = [[Wα]Tαi ]Tαj = Wα\{αi,αj}, and continue this process until the set α \ {αi, αj} becomes a singleton set and finally apply the assumption for singleton set. To this end, let α ⊆ Λn, #α ≥ 2 and αi ∈ α. Consider the set F = Wα\{αi} ⊖ Tαi(Wα\{αi}). Now by Proposition 2.2, Wα\{αi} is a reducing subspace for Tαi and therefore F = {x ∈ Wα\{αi} : x ∈ Ker T ∗ αi} = Wα\{αi} ∩ Wαi = Wα. On the other hand, since Wα\{αi} is a reducing subspace for Tαi then by assumption F = Wα is a wandering subspace for Tαi. Thus [Wα]Tαi = Wα\{αi}. This completes the proof. (cid:3) Combining Theorem 2.1 with the above theorem we have the following result, which is the case of our main interest. Corollary 2.4. Let T = (T1, . . . , Tn) be a commuting tuple of analytic operators on H such that T is doubly commuting and satisfies one of the following properties: (a) Ti is concave for each i = 1, . . . , n (b) kTix + yk2 ≤ 2(kxk2 + kTiyk2) (x, y ∈ H, i = 1, 2, . . . , n). Then for any non-empty subset α = {α1, . . . , αk} ⊆ Λn, Wα is a wandering subspace for Tα = (Tα1, . . . , Tαk), where Wα is as in (3). Proof. Note that if T satisfies condition (a) or (b) then for any 1 ≤ i ≤ n and reducing subspace Si of Ti, TiSi also satisfies condition (a) or (b) respectively. Thus by Theorem 2.1, T satisfies the hypothesis of the above theorem and the proof follows. (cid:3) 6 CHATTOPADHYAY, DAS, SARKAR, AND SARKAR We end this section by proving kind of converse of the above result and part of which is a generalization of [ReT], Theorem 3. Theorem 2.5. Let T = (T1, . . . , Tn) be an n-tuple of commuting operators on H with the property kTix + yk2 ≤ 2(kxk2 + kTiyk2) (x, y ∈ H, i = 1, 2, . . . , n) or Ti is concave for all i = 1, . . . , n. Then, (i) T is doubly commuting on H, and (ii) Ti is analytic for all i = 1, 2, . . . , n if and only if (a) for any non-empty subset α = {α1, . . . , αk} ⊆ Λn, Wα is a wandering subspace for Tα = (Tα1, . . . , Tαk) and for k ≥ 2, [Wα]Tαi (b) Ti commutes with T ∗ = Wα\{αi} for all αi ∈ α, j Tj for all 1 ≤ i < j ≤ n. Proof. The forward direction follows from Theorem 2.3. For the converse, suppose that (a) and (b) hold. To show (i), let 1 ≤ i ≤ n be fixed. By assumption (a), Wi = [W{i,j}]Tj for all 1 ≤ i 6= j ≤ n. This shows that Wi is Tj invariant subspace for all 1 ≤ j 6= i ≤ n. Let x ∈ H. Since [Wi]Ti = H then there exists a sequence xk converging to x such that xk = T m i xm,k, NkXm=0 where each Nk is a nonnegative integer and xm,k is a member of Wi. Now for any 1 ≤ j 6= i ≤ n we have, T ∗ i Tjxk = NkXm=0 T ∗ i TjT m i xm,k = T ∗ i TjT m i xm,k. NkXm=1 On the other hand, TjT ∗ i xk = NkXm=0 TjT ∗ i T m i xm,k = TjT ∗ i T m i xm,k = NkXm=1 T ∗ i TjT m i xm,k, NkXm=1 where the last equality follows from (b). So T ∗ TjT ∗ Finally, by Theorem 3.6 of [Shi] we have that i x. Thus we have (i). i Tjxk = TjT ∗ i xk and by taking limit T ∗ i Tjx = H = [Wi]Ti ⊕ T m i H, ∞ \m=1 for all 1 ≤ i ≤ n. By part (a), [Wi]Ti = H for all 1 ≤ i ≤ n. Thus 1 ≤ i ≤ n and this completes the proof. ∞ Tm=1 T m i H = {0} for all (cid:3) WANDERING SUBSPACES OF THE BERGMAN SPACE AND THE DIRICHLET SPACE OVER Dn 7 3. Wandering subspaces for invariant subspaces of A2(Dn) and D(Dn) In this section we apply the general theorem proved in the previous section to obtain wandering subspaces for invariant subspaces of the Bergman space and the Dirichlet space over polydisc. The Bergman space over Dn is denoted by A2(Dn) and defined by A2(Dn) = {f ∈ O(Dn) :ZDn f (z)2dA(z) < ∞}, where dA is the product area measure on Dn. Below for any invariant subspace S of A2(Dn) and non-empty set α = {α1, . . . , αk} ⊆ Λn, we denote by W S α the following set: W S α := (S ⊖ zαiS). k \i=1 We denote by M = (Mz1, . . . , Mzn) the n-tuple of co-ordinate multiplication operators on A2(Dn). Theorem 3.1. Suppose S is a closed joint M-invariant subspace of A2(Dn). If S is dou- bly commuting then for any non-empty subset α = {α1, . . . , αk} of Λn, W S α is a wandering subspace for MαS = (Mzα1 S, . . . , Mzαk S). The proof of the above theorem follows if we show the tuple of operators (Mz1S . . . , MznS) on S satisfies all the hypothesis of Corollary 2.4. First note that by analyticity of A2(Dn), all the co-ordinate multiplication operators Mzi, i = 1, . . . , n on A2(Dn) are analytic. Then MziS is also analytic for all i = 1, . . . , n. Thus the only hypothesis remains to verify is either condition (a) or (b) of Corollary 2.4. In the next lemma we show that the tuple (Mz1, . . . , Mzn) satisfies condition (b) and therefore so does (Mz1S . . . , MznS). Lemma 3.2. For all 1 ≤ i ≤ n the operators Mzi : A2(Dn) → A2(Dn), f 7→ zif, satisfies for all f, g ∈ A2(Dn). kMzif + gk2 A2(Dn) + kMzigk2 A2(Dn) ≤ 2(cid:16)kf k2 A2(Dn)(cid:17) , Before we prove this lemma, we recall a well known fact regarding the norm of a function in A2(Dn) and prove an inequality. If f is in A2(Dn) with the following power series expansion corresponding to i-th variable: f (z1, · · · , zn) = akzk i , ∞ Xk=0 8 CHATTOPADHYAY, DAS, SARKAR, AND SARKAR where ak ∈ A2(Dn−1) for all k ∈ N, then kf k2 A2(Dn) = kakk2 A2(Dn−1) (k + 1) . ∞ Xk=0 Next we prove the following inequality (can be found in [DuS], page 277, we include the proof for completeness) for any z, w ∈ C and k ∈ N \ {0}, (4) z + w2 k + 1 ≤ 2(cid:18)z2 k + w2 k + 2(cid:19) . To this end, note that 2k(k + 2) Re(z ¯w) ≤ (k + 2)2z2 + k2w2, which follows from the inequality (k + 2)z − kw2 ≥ 0. Now z2 + w2 z2(k + 2)/k + w2k/(k + 2) z2 + w2 + 2 Re(z ¯w) k + 1 z + w2 k + 1 = ≤ + k + 1 = 2(cid:18)z2 k + w2 k + 2(cid:19) . k + 1 ∞ Pk=0 Now we prove the lemma. Proof. Let 1 ≤ i ≤ n be fixed. Let f (z1, · · · , zn) = be the power series expansions of two functions f, g ∈ A2(Dn) with respect to zi-th variable, where ak, bk ∈ A2(Dn−1) for all k ∈ N. Then akzk i and g(z1, · · · , zn) = bkzk i ∞ Pk=0 (Mzif + g) = akzk+1 i + ∞ Xk=0 ∞ Xk=0 bkzk i = b0 + Now (ak−1 + bk)zk i . ∞ Xk=1 kMzif + gk2 A2(Dn) = kb0k2 A2(Dn−1) + A2(Dn−1) ∞ ∞ kak−1 + bkk2 (k + 1) Xk=1 Xk=1 kak−1k2 Xk=1 A2(Dn)(cid:17) . + ∞ kakk2 A2(Dn−1) kbk−1k2 A2(Dn−1) k + 1 k + 1 A2(Dn) + kMzigk2 ≤ kb0k2 A2(Dn−1) + 2 = 2 ∞ Xk=0 = 2(cid:16)kf k2 A2(Dn−1) k + kbkk2 A2(Dn−1) k + 2 ! ( by (4)) ! (cid:3) Thus the result. WANDERING SUBSPACES OF THE BERGMAN SPACE AND THE DIRICHLET SPACE OVER Dn 9 Now we turn our discussion to the Dirichlet space over polydisc. The Dirichlet space over D is denoted by D(D) and defined by D(D) := {f ∈ O(D) : f ′ ∈ A2(D)}. (k + 1)ak2. The Dirichlet space over Dn is denoted For any f = by D(Dn) and defined by ∞ ∞ D = Pk=0 akzk ∈ D(D), kf k2 Pk=0 D(Dn) := {f ∈ O(cid:0)D; D(Dn−1)(cid:1) : f = Another way of expressing D(Dn) is the following D(Dn) := D(D) ⊗ · · · ⊗ D(D) . n-times {z } ∞ Xk=0 akzk, ∞ Xk=0 (k + 1)kakk2 D(Dn−1) < ∞}. In the above, ak ∈ D(Dn−1) for all k ∈ N. The co-ordinate multiplication operators on D(Dn) are also denoted by Mzi, i = 1, . . . , n. Since D(Dn) contains holomorphic functions on Dn then Mzi is analytic for all i = 1, . . . , n. Let 1 ≤ i ≤ n be fixed. Now for f ∈ D(Dn), let i be the Taylor expansion of f corresponding to zi-th variable, where ak ∈ D(Dn−1) f = akzk ∞ Pk=0 for all k ∈ N. Then kf k2 = (k + 1)kakk2 D(Dn−1) and kM 2 zif k2 + kf k2 = = (k + 1)kak−2k2 D(Dn−1) + (k + 1)kakk2 D(Dn−1) (k − 1)kakk2 D(Dn−1) + (k + 1)kakk2 D(Dn−1) ∞ ∞ Xk=0 Xk=0 ∞ Pk=0 ∞ ∞ Xk=2 Xk=0 Xk=1 ∞ = 2 kkakk2 D(Dn−1) = 2kMzif k2. Therefore Mzi is concave for all i = 1, . . . , n. Thus again by Corollary 2.4, we have proved the following result of wandering subspaces for invariant subspaces of Dirichlet space over polydisc. Theorem 3.3. Suppose S is a closed joint (Mz1, . . . , Mzn)-invariant subspace of D(Dn). If S is doubly commuting then for any non-empty subset α = {α1, . . . , αk} of Λn, W S α is a wandering subspace for MαS = (Mzα1 S, . . . , Mzαk S), where W S α = (S ⊖ zαiS). k \i=1 10 CHATTOPADHYAY, DAS, SARKAR, AND SARKAR We conclude the paper with the remark that since the hypothesis of Corollary 2.4 and The- orem 2.5 are same then the same conclusion as in Theorem 2.5 holds for invariant subspaces of Bergman space or Dirichlet space over polydisc. Acknowledgment: First two authors are grateful to Indian Statistical Institute, Bangalore Centre for warm hospitality. The fourth author was supported by UGC Centre for Advanced Study. References [Aro] N. Aronszajn, Theory of reproducing kernels, Trans. of American Mathematical Soc. 68 (1950), 337- 404. [ARS] A. Aleman, S. Richter and C. Sundberg, Beurling's theorem for the Bergman space, Acta. Math. 177 (1996), 275 -- 310. [Beu] A. Beurling, On two problems concerning linear transformations in Hilbert space, Acta. Math. 81 (1949), 239 -- 255 [DuS] P. Duren and A. Schuster, Bergman Spaces, Math. Surveys and Monographs, 100, Amer. Math. Soc. Providence. [Hal] P. R. Halmos, Shifts on Hilbert spaces, J. Reine Angew. Math. 208 (1961) 102 -- 112. [Man] V. Mandrekar, The validity of Beurling theorems in polydiscs, Proc. Amer. Math. Soc. 103 (1988), 145 -- 148. [ReT] D. Redett and J. Tung, Invariant subspaces in Bergman space over the bidisc Proc. Amer Math Soc, 138 (2010) , 2425 -- 2430. S. Richter, Invariant subspaces of the Dirichlet shift, J. Reine Angew. Math. 386 (1988), 205 -- 220. [Ric] [Rud] W. Rudin, Function theory in polydiscs, Benjamin, New York, 1969. [Shi] S. Shimorin, Wold-type decompositions and wandering subspaces for operators close to isometries, J. Reine Angew. Math. 531 (2001), 147 -- 189. [SSW] J. Sarkar, A. Sasane and B. Wick, Doubly commuting submodules of the Hardy module over polydiscs, preprint, arXiv:1302.5312. [Wol] H. Wold, A study in the analysis of stationary time series, Almquist and Wiksell, Uppsala, 1938. (A. Chattopadhyay) Indian Statistical Institute, Statistics and Mathematics Unit, 8th Mile, Mysore Road, Bangalore, 560059, India E-mail address: [email protected], [email protected] (B. K. Das) Indian Statistical Institute, Statistics and Mathematics Unit, 8th Mile, Mysore Road, Bangalore, 560059, India E-mail address: [email protected], [email protected] (J. Sarkar) Indian Statistical Institute, Statistics and Mathematics Unit, 8th Mile, Mysore Road, Bangalore, 560059, India E-mail address: [email protected], [email protected] (S. Sarkar) Department of Mathematics, Indian Institute of Science, Bangalore, 560 012, India E-mail address: [email protected]
1608.03080
1
1608
2016-08-10T08:32:26
The classical theory of calculus of variations for generalized functions
[ "math.FA", "math.DG" ]
We present an extension of the classical theory of calculus of variations to generalized functions. The framework is the category of generalized smooth functions, which includes Schwartz distributions while sharing many nonlinear properties with ordinary smooth functions. We prove full connections between extremals and Euler-Lagrange equations, classical necessary and sufficient conditions to have a minimizer, the necessary Legendre condition, Jacobi's theorem on conjugate points and Noether's theorem. We close with an application to low regularity Riemannian geometry.
math.FA
math
THE CLASSICAL THEORY OF CALCULUS OF VARIATIONS FOR GENERALIZED FUNCTIONS ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO Abstract. We present an extension of the classical theory of calculus of vari- ations to generalized functions. The framework is the category of generalized smooth functions, which includes Schwartz distributions while sharing many nonlinear properties with ordinary smooth functions. We prove full connec- tions between extremals and Euler-Lagrange equations, classical necessary and sufficient conditions to have a minimizer, the necessary Legendre condition, Ja- cobi's theorem on conjugate points and Noether's theorem. We close with an application to low regularity Riemannian geometry. 1. Introduction and motivations Singular problems in the calculus of variations have longly been studied both in mathematics and in relevant applications (see, e.g., [15, 5, 39, 23] and references therein). In this paper, we introduce an approach to variational problems involving singularities that allows the extension of the classical theory with very natural statements and proofs. We are interested in extremizing functionals which are either distributional themselves or whose set of extremals includes generalized functions. Clearly, distribution theory, being a linear theory, has certain difficulties when nonlinear problems are in play. To overcome this type of problems, we are going to use the category of generalized smooth functions, see [9, 10, 12, 13]. This theory seems to be a good candidate, since it is an extension of classical distribution theory which allows to model nonlinear singular problems, while at the same time sharing many nonlinear properties with ordinary smooth functions, like the closure with respect to composition and several non trivial classical theorems of the calculus. One could describe generalized smooth functions as a methodological restoration of Cauchy-Dirac's original conception of generalized function, see [6, 26, 22]. In essence, the idea of Cauchy and Dirac (but also of Poisson, Kirchhoff, Helmholtz, Kelvin and Heaviside) was to view generalized functions as suitable types of smooth set-theoretical maps obtained from ordinary smooth maps depending on suitable infinitesimal or infinite parameters. For example, the density of a Cauchy-Lorentz distribution with an infinitesimal scale parameter was used by Cauchy to obtain classical properties which nowadays are attributed to the Dirac delta, cf. [22]. In the present work, the foundation of the calculus of variations is set for func- tionals defined by arbitrary generalized functions. This in particular applies to any Schwartz distribution and any Colombeau generalized function, and hence justifies the title of the present paper. 2000 Mathematics Subject Classification. 49-XX, 46F-XX, 46F30, 53B20. Key words and phrases. Calculus of variations, Schwartz distributions, generalized functions for nonlinear analysis, low regular Riemannian geometry. A. Lecke has been supported by the uni:doc fellowship programme of the University of Vienna. L. Luperi Baglini has been supported by grant M1876-N35 of the Austrian Science Fund FWF. P. Giordano has been supported by grants P25311-N25 and P25116-N25 of the Austrian Science Fund FWF. 1 2 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO For example, during the last years, the study of low regularly Riemannian and Lorentzian geometry was intensified and made a huge amount of progress (cf. [25, 24, 37, 31, 27, 35]). It was shown that the exponential map is a bi-Lipschitz homeomorphism when metrics g ∈ C1,1 are considered, [32, 25], or that Hawking's singularity theorem still holds when g ∈ C1,1, see [24]. However, calculus of vari- ations in the classical sense may cease to hold when metrics with C1,1 regularity, or below, are considered [17, 28]. This motivates the search for an alternative. In fact, if p, q ∈ Rd and Ω(p, q) denotes the set of all Lipschitz continuous curves connecting p and q, the natural question about what curves γ ∈ Ω(p, q) realize the minimal g-length leads to the corresponding geodesic equation, but the Jacobi equation is not rigorously defined. To be more precise: The Riemannian curvature tensor exists only as an L∞ loc function on Rd and is evaluated along γ. However, the image Im(γ) of γ has Lebesgue-measure zero, if d > 1. Thus we cannot state the Jacobi equations properly. In order to present a possible way out of the aforementioned problems, the singular metric g is embedded as a generalized smooth function. In this way, the embedding ι(g) has derivatives of all orders, valued in a suitable non-Archimedean ring1 ρ(cid:101)R ⊇ R, and behaves very closely to a standard smooth function. We apply our extended calculus of variations to the generalized Riemannian space (ρ(cid:101)Rd, ι(g)), and sketch a way to translate the given problem into the language of generalized smooth functions, solve it there, and translate it back to the standard Riemannian space (Rd, g). Clearly, the process of embedding the singular metric g using ι(g) introduces infinitesimal differences. This is typical in a non-Archimedean setting, but the notion of standard part comes to help: standard real number s, i.e. x− s ≤ r for all r ∈ R>0, then the standard part of x is exactly s. We then show that (assuming that (Rd, g) is geodesically complete) the standard part of the minimal length in the sense of generalized smooth functions is the minimal length in the classical sense, and give a simple way to check if a given (classical) geodesic is a minimizer of the length functional or not. In this way, the framework of generalized smooth functions is presented as a method to solve standard problems rather than a proposal to switch into a new setting. if x ∈ ρ(cid:101)R is infinitely close to a The structure of the present paper is as follows. We start with an introduction into the setting of generalized smooth functions and give basic notions concerning generalized smooth functions and their calculus that are needed for the calculus of variations (Section 2). The paper is self-contained in the sense that it contains all the statements required for the proofs of calculus of variations we are going to present. If proofs of preliminaries are omitted, we clearly give references to where they can be found. Therefore, to understand this paper, only a basic knowledge of distribution theory is needed. In Section 3, we obtain some preliminary lemmas regarding the calculus of variations with generalized smooth functions. The first variation and the notion of critical point will be defined and studied in section 4. We prove the fundamental lemma of calculus of variations and the full connection between critical points of a given functional and solutions of the corresponding Euler-Lagrange equation. In section 5, we study the second variation and define the notion of local minimizer. We also extend to generalized functions classical necessary and sufficient conditions to have a minimizer, and we give a proof of the Legendre condition. In Section 6, we introduce the notion of Jacobi field and extend to generalized functions the definition of conjugate points, so as to prove the corresponding Jacobi theorem. In Section 7, we extend the classical Noether's theorem. We close with an application to C1,1 Riemannian geometry in Section 8. 1I.e. a ring that contains infinitesimal and infinite numbers. CALCULUS OF VARIATIONS FOR GF 3 Note that the work [23] already established the calculus of variations in the setting of Colombeau generalized functions by using a comparable methodological approach. Indeed, generalized smooth functions are related to Colombeau gener- alized functions, and one could say that the former is a minimal extension of the latter so as to get more general domains for generalized functions and hence the closure with respect to composition and a better behaviour on unbounded sets. However, there are some conceptual advantages in our approach. (i) Whereas generalized smooth functions are closed with respect to composi- tion, Colombeau generalized functions are not. This forces [23] to consider only functionals defined using compactly supported Colombeau generalized functions, i.e. functions assuming only finite values, or tempered generalized function. (ii) The authors of [23] are forced to consider the so called compactly supported points c(Ω) (i.e. finite points in Ω ⊆ Rn), where the setting of generalized smooth functions gives the possibility to consider more natural domains like the interval [a, b] ⊆ ρ(cid:101)R. This leads us to extend in a natural way the state- still hold when we take as a, b ∈ ρ(cid:101)R two infinite numbers such that a < b, or as boundary points two unbounded points p, q ∈ ρ(cid:101)Rd. ments of classical results of calculus of variations. Moreover, all our results (iii) Furthermore, the theory of generalized smooth functions was developed to be very user friendly, in the sense that one can avoid cumbersome "ε-wise" proofs quite often, whereas the proofs in [23] frequently use this technique. Thus, one could say that some of the proofs based on generalized smooth functions are more "intrinsic" and close to the classical proofs in a standard smooth setting. This allows a smoother approach to this new framework. (iv) The setting of generalized smooth functions depends on a fixed infinitesimal net (ρε)ε∈(0,1] ↓ 0, whereas the Colombeau setting considers only ρε = ε. This added degree of freedom allows to solve singular differential equations that are unsolvable in the classical Colombeau setting and to prove a more general Jacobi theorem on conjugate points. In [23] only the notion of global minimizer is defined, whereas we define the notion of local minimizer as in [8] using a natural topology in space of generalized smooth curves. (v) (vi) We obtain more classical results like the Legendre condition, and the classical (vii) results about Jacobi fields and conjugate points. In addition, note that the Colombeau generalized functions can be embedded into generalized smooth functions. Thus our approach is a natural extension of [23]. 2. Basic notions The new ring of scalars. In this work, I denotes the interval (0, 1] ⊆ R and we will always use the variable ε for elements of I; we also denote ε-dependent nets x ∈ RI simply by (xε). By N we denote the set of natural numbers, including zero. We start by defining the new simple non-Archimedean ring of scalars that extends the real field R. The entire theory is constructive to a high degree, e.g. no ultrafilter or non-standard method is used. For all the proofs of results in this section, see [13, 9, 12]. Definition 1. Let ρ = (ρε) ∈ RI be a net such that limε→0 ρε = 0+, then (i) ε ) a ∈ R>0} is called the asymptotic gauge generated by ρ. The I(ρ) := {(ρ−a net ρ is called a gauge. 4 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO If P(ε) is a property of ε ∈ I, we use the notation ∀0ε : P(ε) to denote (ii) ∃ε0 ∈ I ∀ε ∈ (0, ε0] : P(ε). We can read ∀0ε as for ε small. (iii) We say that a net (xε) ∈ RI is ρ-moderate, and we write (xε) ∈ Rρ if ∃(Jε) ∈ I(ρ) : xε = O(Jε) as ε → 0+. (iv) Let (xε), (yε) ∈ RI , then we say that (xε) ∼ρ (yε) if ∀(Jε) ∈ I(ρ) : xε = ) as ε → 0+. This is a congruence relation on the ring Rρ of yε + O(J−1 moderate nets with respect to pointwise operations, and we can hence define ε ρ(cid:101)R := Rρ/ ∼ρ, which we call Robinson-Colombeau ring of generalized numbers, [34, 4]. We denote the equivalence class x ∈ ρ(cid:101)R simply by x =: [xε] := [(xε)]∼ ∈ ρ(cid:101)R. In the following, ρ will always denote a net as in Def. 1. The infinitesimal ρ can be chosen depending on the class of differential equations we need to solve for the generalized functions we are going to introduce, see [14]. For motivations exists (zε) ∈ RI such that (zε) ∼ρ 0 (we then say that (zε) is ρ-negligible) and xε ≤ yε + zε for ε small. Equivalently, we have that x ≤ y if and only if there exist representatives [xε] = x and [yε] = y such that xε ≤ yε for all ε. Clearly, concerning the naturality of ρ(cid:101)R, see [12]. We can also define an order relation on ρ(cid:101)R by saying that [xε] ≤ [yε] if there ρ(cid:101)R is a partially ordered ring. The usual real numbers r ∈ R are embedded in ρ(cid:101)R considering constant nets [r] ∈ ρ(cid:101)R. max ([xε], [yε]) := [max(xε, yε)] and the absolute value [xε] := [xε] ∈ ρ(cid:101)R. Note, tomary notation ρ(cid:101)R∗ for the set of invertible generalized numbers. Our notations for intervals are: [a, b] := {x ∈ ρ(cid:101)R a ≤ x ≤ b}, [a, b]R := [a, b]∩ R, and analogously for segments [x, y] := {x + r · (y − x) r ∈ [0, 1]} ⊆ ρ(cid:101)Rn and [x, y]Rn = [x, y] ∩ Rn. Fi- Even if the order ≤ is not total, we still have the possibility to define the infimum min ([xε], [yε]) := [min(xε, yε)], and analogously the supremum function e.g., that x ≤ z and −x ≤ z imply x ≤ z. In the following, we will also use the cus- nally, we write 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 limε→0+ xε − yε = 0 for all representatives x = [xε] and y = [yε]. Topologies on ρ(cid:101)Rn. On the ρ(cid:101)R-module ρ(cid:101)Rn, we can consider the natural extension of the Euclidean norm, i.e. [xε] := [xε] ∈ ρ(cid:101)R, where [xε] ∈ ρ(cid:101)Rn. Even if this generalized norm takes values in ρ(cid:101)R, it shares several properties with usual norms, to consider on ρ(cid:101)Rn topologies generated by balls defined by this generalized norm Definition 2. Let R ∈(cid:110)ρ(cid:101)R∗ like the triangular inequality or the property y · x = y·x. It is therefore natural (cid:111) , c ∈ ρ(cid:101)Rn and x, y ∈ ρ(cid:101)R, then: (cid:111) (cid:110) x ∈ ρ(cid:101)Rn x − c <R r and a set of radii R: ≥0, R>0 for each r ∈ R. (i) We write x <R y if ∃r ∈ R : r ≤ y − x. (ii) BR (iii) B E r (c) := r (c) := {x ∈ Rn x − c < r}, for each r ∈ R>0, denotes an ordinary Euclidean ball in Rn. (cid:111) is a base for a topology on ρ(cid:101)Rn. The der relation a ≤ b and a (cid:54)= b (that we will never use) because for R ∈(cid:110)ρ(cid:101)R∗ r (c) r ∈ R, c ∈ ρ(cid:101)Rn(cid:111) topology generated in the case R = ρ(cid:101)R∗ The relation <R has better topological properties as compared to the usual strict or- ≥0, R>0 ≥0 is called sharp topology, whereas the one with the set of radii R = R>0 is called Fermat topology. We will call sharply open the set of balls (cid:110) BR CALCULUS OF VARIATIONS FOR GF 5 ≥0 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(cid:48)(x0) is infinite, we have f (x) ≈ f (x0) only for x ≈ x0 , see [11, 10]. With an innocuous abuse of language, we write x < y instead of x <ρ(cid:101)R∗ For example, ρ(cid:101)R∗ y and x <R y instead of x <R>0 y. intervals are defined using the relation <, i.e. (a, b) := {x ∈ ρ(cid:101)R a < x < b}. Lemma 3. Let x ∈ ρ(cid:101)R. Then the following are equivalent: ≥0 = ρ(cid:101)R>0. We will simply write Br(c) to denote an open ball in The following result is useful to deal with positive and invertible generalized r (c) for an open ball in the Fermat topology. Also open the sharp topology and B F numbers (cf. [16]). x is invertible and x ≥ 0, i.e. x > 0. (i) (ii) For each representative (xε) ∈ Rρ of x we have ∀0ε : xε > 0. (iii) For each representative (xε) ∈ Rρ of x we have ∃m ∈ N∀0ε : xε > ρm topology is dense in [a, b]. We will also need the following result. Lemma 4. Let a, b ∈ ρ(cid:101)R such that a < b, then the interior int ([a, b]) in the sharp and bounded sets in ρ(cid:101)Rn is by using a net (Aε) of subsets Aε ⊆ Rn. We have two ways of extending the membership relation xε ∈ Aε to generalized points [xε] ∈ ρ(cid:101)R: Internal and strongly internal sets. A natural way to obtain sharply open, closed ε Definition 5. Let (Aε) be a net of subsets of Rn, then (i) (cid:110) [xε] ∈ ρ(cid:101)Rn ∀0ε : xε ∈ Aε (cid:111) (ii) is called the internal set generated by [Aε] := the net (Aε). See [33] for the introduction and an in-depth study of this notion in the case ρε = ε. Let (xε) be a net of points of Rn, then we say that xε ∈ε Aε, and we read it (cid:110) (cid:111) [xε] ∈ ρ(cid:101)Rn xε ∈ε Aε ε) ∼ρ (xε), then as (xε) strongly belongs to (Aε), if ∀0ε : xε ∈ Aε and if (x(cid:48) also x(cid:48) , exists r ∈ ρ(cid:101)R>0 such that K ⊆ Br(0). Analogously, a net (Aε) is sharply and we call it the strongly internal set generated by the net (Aε). ε ∈ Aε for ε small. Moreover, we set (cid:104)Aε(cid:105) := (iii) Finally, we say that the internal set K = [Aε] is sharply bounded if there bounded the internal set [Aε] is sharply bounded. Therefore, x ∈ [Aε] if there exists a representative [xε] = x such that xε ∈ Aε for ε 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 Aε = A ⊆ Rn is simply denoted by [A]. The following theorem shows that internal and strongly internal sets have dual topological properties: Theorem 6. For ε ∈ I, let Aε ⊆ Rn and let xε ∈ Rn. Then we have (i) if and only if [d(xε, Aε)] = 0 ∈ ρ(cid:101)R. [xε] ∈ [Aε] if and only if ∀q ∈ R>0 ∀0ε : d(xε, Aε) ≤ ρq ε. Therefore [xε] ∈ [Aε] [xε] ∈ (cid:104)Aε(cid:105) if and only if ∃q ∈ R>0 ∀0ε : d(xε, Ac ε := Rn\Aε. Therefore, if (d(xε, Ac ε)] > 0. [Aε] is sharply closed and (cid:104)Aε(cid:105) is sharply open. [Aε] = [cl (Aε)], where cl (S) is the closure of S ⊆ Rn. On the other hand (cid:104)Aε(cid:105) = (cid:104)int(Aε)(cid:105), where int (S) is the interior of S ⊆ Rn. ε)) ∈ Rρ, then [xε] ∈ (cid:104)Aε(cid:105) if and only if [d(xε, Ac (ii) (iii) (iv) ε, where Ac ε) > ρq 6 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO Generalized smooth functions and their calculus. Using the ring ρ(cid:101)R, it is easy to probability density by f (x, σ), and if we set σ = [σε] ∈ ρ(cid:101)R>0, where σ ≈ 0, we Definition 7. Let X ⊆ ρ(cid:101)Rn and Y ⊆ ρ(cid:101)Rd be arbitrary subsets of generalized consider a Gaussian with an infinitesimal standard deviation. If we denote this obtain the net of smooth functions (f (−, σε))ε∈I . This is the basic idea we are going to develop in the following points. Then we say that f : X −→ Y is a generalized smooth function if there exists a net fε ∈ C∞(Ωε, Rd) defining f in the sense that X ⊆ (cid:104)Ωε(cid:105), f ([xε]) = [fε(xε)] ∈ Y and (∂αfε(xε)) ∈ Rd ρ for all x = [xε] ∈ X and all α ∈ Nn. The space of generalized smooth functions (GSF) from X to Y is denoted by ρGC∞(X, Y ). Let us note explicitly that this definition states minimal logical conditions to ob- tain a set-theoretical map from X into Y and defined by a net of smooth functions. In particular, the following Thm. 8 states that the equality f ([xε]) = [fε(xε)] is meaningful, i.e. that we have independence from the representatives for all deriva- tives [xε] ∈ X (cid:55)→ [∂αfε(xε)] ∈ ρ(cid:101)Rd, α ∈ Nn. Theorem 8. Let X ⊆ ρ(cid:101)Rn and Y ⊆ ρ(cid:101)Rd be arbitrary subsets of generalized points. Let fε ∈ C∞(Ωε, Rd) be a net of smooth functions that defines a generalized smooth map of the type X −→ Y , then ε) ∈ Rn ρ : (i) (ii) (iii) For all α ∈ Nn, the GSF g : [xε] ∈ X (cid:55)→ [∂αfε(xε)] ∈ (cid:101)Rd is locally Lipschitz ∀α ∈ Nn ∀(xε), (x(cid:48) ∀[xε] ∈ X ∀α ∈ Nn ∃q ∈ R>0 ∀0ε : supy∈BE εq (xε) ∂αuε(y) ≤ ε−q. that g(x) − g(y) ≤ Lx − y for all x, y ∈ U and some L ∈ ρ(cid:101)R. in the sharp topology, i.e. each x ∈ X possesses a sharp neighborhood U such (iv) Each f ∈ ρGC∞(X, Y ) is continuous with respect to the sharp topologies in- ε] ∈ X ⇒ (∂αuε(xε)) ∼ρ (∂αuε(x(cid:48) [xε] = [x(cid:48) ε)). duced on X, Y . (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 vε ∈ C∞(Rn, Rd) defining a generalized smooth map of type X −→ Y such that f = [vε(−)]X . (vii) Subsets S ⊆ ρ(cid:101)Rs with the trace of the sharp topology, and generalized smooth (vi) maps as arrows form a subcategory of the category of topological spaces. We will call this category ρGC∞, the category of GSF. The differential calculus for GSF can be introduced showing existence and unique- ness of another GSF serving as incremental ratio. Theorem 9 (Fermat-Reyes theorem for GSF). Let U ⊆ ρ(cid:101)Rn be a sharply open set, let v = [vε] ∈ ρ(cid:101)Rn, and let f ∈ ρGC∞(U, ρ(cid:101)R) be a generalized smooth map generated map r ∈ ρGC∞(T, ρ(cid:101)R), called the generalized incremental ratio of f along v, by the net of smooth functions fε ∈ C∞(Ωε, R). Then (i) There exists a sharp neighborhood T of U × {0} and a generalized smooth ∀(x, h) ∈ T : f (x + hv) = f (x) + h · r(x, h). such that U × {0}. (cid:104) ∂fε ∂vε (cid:105) (ii) Any two generalized incremental ratios coincide on a sharp neighborhood of (iii) We have r(x, 0) = (xε) v := ∂f ∂v (x) := r(x, 0), so that ∂f for every x ∈ U and we can thus define Df (x)· ∂v ∈ ρGC∞(U, ρ(cid:101)R). CALCULUS OF VARIATIONS FOR GF 7 j If U is a large open set, then an analogous statement holds replacing sharp neigh- borhoods by large neighborhoods. Note that this result permits to consider the partial derivative of f with respect to an arbitrary generalized vector v ∈ ρ(cid:101)Rn which can be, e.g., infinitesimal or (cid:16)ρ(cid:101)Rn(cid:17)j −→ ρ(cid:101)Rd over the ring ρ(cid:101)R will be denoted by Lj(ρ(cid:101)Rn, ρ(cid:101)Rd). For A = [Aε(−)] ∈ Lj(ρ(cid:101)Rn, ρ(cid:101)Rd), we set A := [Aε], the generalized infinite. Using this result, we can also define subsequent differentials Djf (x) as j−multilinear maps, and we set Djf (x) · hj := Djf (x)(h, . . . . . . , h). The set of all the j−multilinear maps number defined by the operator norms of the multilinear maps Aε ∈ Lj(Rn, Rd). The following result follows from the analogous properties for the nets of smooth functions defining f and g. Theorem 10. Let U ⊆ ρ(cid:101)Rn be an open subset in the sharp topology, let v ∈ ρ(cid:101)Rn and f , g : U −→ ρ(cid:101)R be generalized smooth maps. Then ∀r ∈ ρ(cid:101)R ∂v + ∂g ∂v · g + f · ∂g (i) (ii) (iii) (iv) For each x ∈ U , the map df (x).v := ∂f (v) ∂(f +g) ∂v = ∂f ∂(r·f ) ∂v = r · ∂f ∂v (x) ∈ ρ(cid:101)R is ρ(cid:101)R-linear in v ∈ ρ(cid:101)Rn ∂(f·g) ∂v = ∂f Let U ⊆ ρ(cid:101)Rn and V ⊆ ρ(cid:101)Rd be open subsets in the sharp topology and g ∈ ρGC∞(V, U ), f ∈ ρGC∞(U, ρ(cid:101)R) be generalized smooth maps. Then for all x ∈ V and all v ∈ ρ(cid:101)Rd, we have ∂(f◦g) Theorem 11. Let f ∈ ρGC∞(U, ρ(cid:101)R) be a generalized smooth function defined in the sharply open set U ⊆ ρ(cid:101)Rn. Let a, x ∈ ρ(cid:101)Rn such that the line segment [a, x] ⊆ U . We also have a generalization of the Taylor formula: (x) = df (g(x)) . ∂g ∂v (x). ∂v ∂v ∂v ∂v Then, for all n ∈ N we have ∃ξ ∈ [a, x] : f (x) = Djf (a) · (x − a)j + Dn+1f (ξ) (n + 1)! · (x − a)n+1. (2.1) If we further assume that all the n components (x − a)k ∈ ρ(cid:101)R of x − a ∈ ρ(cid:101)Rn are invertible, then there exists ρ ∈ ρ(cid:101)R>0, ρ ≤ x − a, such that j=0 j! n(cid:88) n(cid:88) ∀k ∈ Bρ(0)∃ξ ∈ [a− k, a + k] : f (a + k) = Djf (a) · kj + Dn+1f (ξ) (n + 1)! · kn+1 (2.2) j! j=0 Dn+1f (ξ) (n + 1)! · kn+1 ≈ 0. (2.3) Formula (2.1) corresponds to a direct generalization of Taylor formulas for ordi- nary smooth functions with Lagrange remainder. On the other hand, in (2.2) and (2.3), the possibility that the differential Dn+1f may be infinite at some point is considered, and the Taylor formulas are stated so as to have infinitesimal remainder. The following local inverse function theorem will be used in the proof of Jacobi's theorem (see [9] for a proof). Theorem 12. Let X ⊆ ρ(cid:101)Rn, let f ∈ ρGC∞(X, ρ(cid:101)Rn) and suppose that for some x0 in the sharp interior of X, Df (x0) is invertible in L(ρ(cid:101)Rn, ρ(cid:101)Rn). Then there exists a sharp neighborhood U ⊆ X of x0 and a sharp neighborhood V of f (x0) such that f : U → V is invertible and f−1 ∈ ρGC∞(V, U ). 8 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO +(a) := limt→a a<t GSF is based on the following We can define right and left derivatives as e.g. f(cid:48)(a) := f(cid:48) which always exist if f ∈ ρGC∞([a, b], ρ(cid:101)Rd). One dimensional integral calculus of Theorem 13. Let f ∈ ρGC∞([a, b], ρ(cid:101)R) be a generalized smooth function defined in the interval [a, b] ⊆ ρ(cid:101)R, where a < b. Let c ∈ [a, b]. Then, there exists one and only one generalized smooth function F ∈ ρGC∞([a, b], ρ(cid:101)R) such that F (c) = 0 and F (cid:48)(x) = f (x) for all x ∈ [a, b]. Moreover, if f is defined by the net fε ∈ C∞(R, R) and c = [cε], then F (x) = for all x = [xε] ∈ [a, b]. f(cid:48)(t), fε(s) ds (cid:104)´ (cid:105) xε cε ´ We can thus define ´ (cid:17)(cid:48) (−) c (i) Definition 14. Under the assumptions of Theorem 13, we denote by f (s) ds ∈ ρGC∞([a, b], ρ(cid:101)R) the unique generalized smooth function such that: ´ (cid:16)´ c c f = 0 f u f (s) ds = f (x) for all x ∈ [a, b]. All the classical rules of integral calculus hold in this setting: Theorem 15. Let f ∈ ρGC∞(U, ρ(cid:101)R) and g ∈ ρGC∞(V, ρ(cid:101)R) be generalized smooth functions defined on sharply open domains in ρ(cid:101)R. Let a, b ∈ ρ(cid:101)R with a < b and c, (x) = d dx (−) c (−) u f := (ii) ´ x d ∈ [a, b] ⊆ U ∩ V , then (i) ´ ´ ´ ´ ´ ´ ´ ´ ∀λ ∈ ρ(cid:101)R ´ d d d c (f + g) = c f + c g ´ ´ d d c λf = λ c f e f for all e ∈ [a, b] ´ d e c f = c f + c f = − d c d f c f(cid:48) = f (d) − f (c) c f(cid:48) · g = [f · g]d c − ´ c f · g(cid:48) d d d d (ii) (iii) (iv) (v) (vi) Theorem 16. Let f ∈ ρGC∞(U, ρ(cid:101)R) and ϕ ∈ ρGC∞(V, U ) be generalized smooth functions defined on sharply open domains in ρ(cid:101)R. Let a, b ∈ ρ(cid:101)R, with a < b, such that [a, b] ⊆ V , ϕ(a) < ϕ(b), [ϕ(a), ϕ(b)] ⊆ U . Finally, assume that ϕ([a, b]) ⊆ [ϕ(a), ϕ(b)]. Then ϕ(b) f (t) dt = ϕ(a) a b f [ϕ(s)] · ϕ(cid:48)(s) ds. 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 ρ 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(cid:48) = y/ dε, which cannot be solved for ρ = (ε), but it has a solution for ρ = (e−1/ε). 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. See also [14, 30] for further details. If ϕ ∈ D(Rn), r ∈ R>0 and x ∈ Rn, we use the notations r (cid:12) ϕ for the function x ∈ Rn (cid:55)→ 1 These notations permit to highlight that (cid:12) 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 (cid:12) (x ⊕ ϕ) = rx ⊕ r (cid:12) ϕ. (cid:1) ∈ R and x ⊕ ϕ for the function y ∈ Rn (cid:55)→ ϕ(y − x) ∈ R. Lemma 17. Let b ∈ ρ(cid:101)R be a net such that limε→0+ bε = +∞. Let d ∈ (0, 1). There exists a net (ψε)ε∈I of D(Rn) with the properties: rn · ϕ(cid:0) x r CALCULUS OF VARIATIONS FOR GF 9 ´ supp(ψε) ⊆ B1(0) for all ε ∈ I. ψε = 1 for all ε ∈ I. ´ ∀j ∈ N∀0ε : 1 ≤ α ≤ j ⇒ ψε ≤ 1 + η. ∀η ∈ R>0 ∀0ε : If n = 1, then the net (ψε)ε∈I can be chosen so that (i) (ii) (iii) ∀α ∈ Nn ∃p ∈ N : supx∈Rn ∂αψε(x) = O(bp (iv) (v) (vi) In particular ψb ε (cid:12) ψε satisfies (ii) - (v). xα · ψε(x) dx = 0. ε := b−1 ´ ε) as ε → 0+. ´ 0 −∞ ψε = d. It is worth noting that the condition (iv) of null moments is well known in the study of convergence of numerical solutions of singular differential equations, see e.g. [38, 7, 19] and references therein. Concerning embeddings of Schwartz distributions, we have the following result, where c(Ω) := {[xε] ∈ [Ω] ∃K (cid:98) Ω∀0ε : xε ∈ K} is called the set of compactly supported points in Ω ⊆ Rn. Theorem 18. Under the assumptions of Lemma 17, let Ω ⊆ Rn be an open set and let (ψb ε) be the net defined in 17. Then the mapping Ω : T ∈ E(cid:48)(Ω) (cid:55)→(cid:2)(cid:0)T ∗ ψb ε (cid:1) (−)(cid:3) ∈ ρGC∞(c(Ω), ρ(cid:101)R) ιb uniquely extends to a sheaf morphism of real vector spaces ιb : D(cid:48) −→ ρGC∞(c((−)), ρ(cid:101)R), (−) : C∞(−) −→ ρGC∞(c((−)), ρ(cid:101)R) and satisfies the following properties: (i) ´ If b ≥ dρ−a for some a ∈ R>0, then ιbC∞ is a sheaf morphism of algebras; If T ∈ E(cid:48)(Ω) then supp(T ) = supp(ιb ιb commutes with partial derivatives, i.e. ∂α(cid:0)ιb limε→0+ T ∈ D(cid:48)(Ω) and α ∈ N. Ω(T )); Ω ιb (ii) (iii) (iv) Ω(T )(cid:1) = ιb Ω(T )ε · ϕ = (cid:104)T, ϕ(cid:105) for all ϕ ∈ D(Ω) and all T ∈ D(cid:48)(Ω); Ω (∂αT ) for each Concerning the embedding of Colombeau generalized functions, we recall that the special Colombeau algebra on Ω is defined as the quotient Gs(Ω) := EM (Ω)/N s(Ω) of moderate nets over negligible nets, where the former is EM (Ω) := {(uε) ∈ C∞(Ω)I ∀K (cid:98) Ω∀α ∈ Nn ∃N ∈ N : sup x∈K ∂αuε(x) = O(ε−N )} and the latter is N s(Ω) := {(uε) ∈ C∞(Ω)I ∀K (cid:98) Ω∀α ∈ Nn ∀m ∈ N : sup x∈K ∂αuε(x) = O(εm)}. Using ρ = (ε), we have the following compatibility result: Theorem 19. A Colombeau generalized function u = (uε) + N s(Ω)d ∈ Gs(Ω)d defines a generalized smooth map u : [xε] ∈ c(Ω) −→ [uε(xε)] ∈ (cid:101)Rd which is locally assignment provides a bijection of Gs(Ω)d onto ρGC∞(c(Ω), ρ(cid:101)Rd) for every open set Lipschitz on the same neighborhood of the Fermat topology for all derivatives. This Ω ⊆ Rn. 2.1. Extreme value theorem and functionally compact sets. For GSF, suit- able generalizations of many classical theorems of differential and integral calculus hold: intermediate value theorem, mean value theorems, a sheaf property for the Fermat topology, local and global inverse function theorems, Banach fixed point theorem and a corresponding Picard-Lindelof theorem, see [13, 12, 29, 9]. Even though the intervals [a, b] ⊆(cid:101)R, a, b ∈ R, are neither compact in the sharp nor in the Fermat topology (see [13, Thm. 25]), analogously to the case of smooth functions, a GSF satisfies an extreme value theorem on such sets. In fact, we have: 10 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO Theorem 20. Let f ∈ GC∞(X,(cid:101)R) be a generalized smooth function defined on the subset X of (cid:101)Rn. Let ∅ (cid:54)= K = [Kε] ⊆ X be an internal set generated by a sharply bounded net (Kε) of compact sets Kε (cid:98) Rn , then ∃m, M ∈ K ∀x ∈ K : f (m) ≤ f (x) ≤ f (M ). (2.4) We shall use the assumptions on K and (Kε) given in this theorem to introduce a notion of "compact subset" which behaves better than the usual classical notion of compactness in the sharp topology. Definition 21. A subset K of (cid:101)Rn is called functionally compact, denoted by K (cid:98)f (cid:101)Rn, if there exists a net (Kε) such that (i) K = [Kε] ⊆(cid:101)Rn If, in addition, K ⊆ U ⊆(cid:101)Rn then we write K (cid:98)f U . Finally, we write [Kε] (cid:98)f U if (ii) (iii) ∀ε ∈ I : Kε (cid:98) Rn (Kε) is sharply bounded (ii), (iii) and [Kε] ⊆ U hold. We motivate the name functionally compact subset by noting that on this type of subsets, GSF have properties very similar to those that ordinary smooth functions have on standard compact sets. Remark 22. By [33, Prop. 2.3], any internal set K = [Kε] is closed in the sharp topology. In particular, the open interval (0, 1) ⊆ (cid:101)R is not functionally compact since (i) (ii) (iv) the GSF f (x) = x does not satisfy the conclusion (2.4) of Prop. 20. In the present paper, we need the following properties of functionally compact sets. it is not closed. If H (cid:98) Rn is a non-empty ordinary compact set, then [H] is functionally compact. In particular, [0, 1] = [[0, 1]R] is functionally compact. (cid:101)Rn is not functionally compact since it is not sharply bounded. (iii) The empty set ∅ =(cid:101)∅ (cid:98)f (cid:101)R. (v) The set of compactly supported points(cid:101)Rc is not functionally compact because Theorem 23. Let K ⊆ X ⊆ (cid:101)Rn, f ∈ GC∞(X,(cid:101)Rd). Then K (cid:98)f (cid:101)Rn implies f (K) (cid:98)f (cid:101)Rd. Corollary 24. If a, b ∈(cid:101)R and a ≤ b, then [a, b] (cid:98)f (cid:101)R. Let us note that a, b ∈ (cid:101)R can also be infinite, e.g. a = [−ε−N ], b = [ε−M ] or Theorem 25. Let K (cid:98)f (cid:101)Rn and H (cid:98)f (cid:101)Rd, then K × H (cid:98)f (cid:101)Rn+d. In particular, if ai ≤ bi for i = 1, . . . , n, then(cid:81)n a = [ε−N ], b = [ε−M ] with M > N . Finally, in the following result we consider the product of functionally compact sets: As a corollary of this theorem and Rem. (22).(ii) we get i=1[ai, bi] (cid:98)f (cid:101)Rn. A theory of compactly supported GSF has been developed in [9], and it closely resembles the classical theory of LF-spaces of compactly supported smooth func- tions. It establishes that for suitable functionally compact subsets, the correspond- ing space of compactly supported GSF contains extensions of all Colombeau gen- eralized functions, and hence also of all Schwartz distributions. CALCULUS OF VARIATIONS FOR GF 11 3. Preliminary results for calculus of variations with GSF In this section, we study extremal values of generalized functions at sharply interior points of intervals [a, b] ⊆ ρ(cid:101)R. As in the classical calculus of variations, this variational problems. Since the new ring of scalars ρ(cid:101)R has zero divisors and is not will provide the basis for proving necessary and sufficient conditions for general totally ordered, the following extension requires a more refined analysis than in the classical case. b a d ds b a ∂ ∂s f (τ, s) dτ = The following lemma shows that we can interchange integration and differentia- tion while working with generalized functions. Lemma 26. Let a, b, c, d ∈ ρ(cid:101)R, with a < b and c < d. Let f ∈ ρGC∞(X, Y ) and assume that [a, b] × [c, d] ⊆ X ⊆ ρ(cid:101)R2 and Y ⊆ ρ(cid:101)Rd. Then for all s ∈ [c, d],we have (3.1) Proof. We first note that f (·, s) ∈ ρGC∞([a, b], Y ) by the closure of GSF with respect to composition. Therefore, ∂ (3.1) is well defined as an integral of a GSF. In order to show that also the left hand side of (3.1) is well-defined, we need to prove that also σ ∈ [c, d] (cid:55)→ ∂s f (·, s) ∈ ρGC∞([a, b], ρ(cid:101)Rd), and the right hand side of ´ ρ(cid:101)Rd is a GSF. Let f be defined by the net fε ∈ C∞(cid:0)Ωε, Rd(cid:1), with X ⊆ (cid:104)Ωε(cid:105). Let a f (τ, σ) dτ ∈ ρ(cid:101)R2 and the extreme value theorem 20 applied (cid:12)(cid:12)(cid:12)(cid:12) ∂n [σε] ∈ [c, d], then [a, b] × {[σε]} (cid:98)f to ∂nf ∂σn yields the existence of N ∈ R>0 such that (cid:12)(cid:12)(cid:12)(cid:12) dτ ≤ ρ−N ∂σn fε(τ, σε) · (bε − aε). fε(τ, σε) dτ f (τ, s) dτ. (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dn dσn (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ bε aε bε aε b This proves that also the left hand side of (3.1) is well-defined as a derivative of a GSF. From the classical derivation under the integral sign, the Fermat-Reyes theorem 9, and Thm. 13 about definite integrals of GSF, we obtain d ds b f (τ, s) dτ = d ds a ε (cid:35) (cid:35) b a (cid:34) (cid:20) ∂ ∂s aε bε aε [fε(τ, s)] dτ bε fε(τ, s) dτ fε(τ, s) dτ (cid:21) fε(τ, s) dτ = = = = d ds d ds (cid:34) b b a a ∂ ∂s f (τ, s) dτ. (cid:3) The next result will frequently be used in the following Lemma 27. Let (D,≥) be a directed set and let f : D −→ ρ(cid:101)R be a set-theoretical map such that f (d) ≥ 0 for all d ∈ D, and ∃ limd∈D f (d) ∈ ρ(cid:101)R in the sharp topology. Then limd∈D f (d) ≥ 0. Proof. Note that the internal set [0, +∞) = [[0, +∞)R] is sharply closed by Thm. 6.(iii). (cid:3) 12 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO (i) Remark 28. If x ∈ ρ(cid:101)R, then x ≥ 0 if and only if ∃A ∈ R>0 ∀a ∈ R>A : x ≥ −dρa. Indeed, it suffices to let a → +∞ in f (a) = x + dρa. (ii) Assume that x, y ∈ ρ(cid:101)Rn and Then taking s → 0 in f (s) = sy − x we get x = 0. ∃s0 ∈ ρ(cid:101)R>0 ∀s ∈ ρ(cid:101)R>0 : s ≤ s0 ⇒ x ≤ sy. Definition 29. We call x = (x1, . . . , xd) ∈ ρ(cid:101)Rd componentwise invertible if and only if for all k ∈ {1, . . . , d} we have that xk ∈ ρ(cid:101)R is invertible. Lemma 30. Let f ∈ ρGC∞(U, Y ) where Y ⊆ ρ(cid:101)R and U ⊆ ρ(cid:101)Rd is a sharply open Proof. By Lem. 3, it follows that for V ⊆ ρ(cid:101)R, the set of invertible points in V, i.e. V ∩ ρ(cid:101)R∗ ⊆ V is dense in V (with respect to the sharp topology). This implies that U ∩ (ρ(cid:101)Rd)∗ ⊆ U is dense. By Thm. 8.(iv), f is sharply continuous, so Lem. 27 subset. Then f ≥ 0 if and only if f (x) ≥ 0 for all componentwise invertible x ∈ U . yields that f (x) ≥ 0. The other direction is obvious. (cid:3) Analogously to the classical case, we say that x0 ∈ X is a local minimum of f ∈ ρGC∞(X) if there exists a sharply open neighbourhood (in the trace topology) Y ⊆ X of x0 such that f (x0) ≤ f (y) for all y ∈ Y . A local maximum is defined accordingly. We will write f (x0) = min!, which is a short hand notation to denote that x0 is a (local) minimum of f . Lemma 31. Let X ⊆ ρ(cid:101)R and let f ∈ ρGC∞(X, ρ(cid:101)R). If x0 ∈ X is a sharply interior GSF with respect to composition. Let r ∈ ρ(cid:101)R>0 be such that B2r(0) =: U ⊆ X and f (0) = min! over U . Take any x ∈ ρ(cid:101)R such that 0 < x < r, so that [−x,x] ⊆ U . Proof. Without loss of generality, we can assume x0 = 0, because of the closure of local minimum of f then f(cid:48)(x0) = 0. Thus, if x > 0, by Taylor's theorem 11 there exists ξ ∈ [0, x] such that Set K := [Brε(0)] (cid:98)f B2r(0) ⊆ U and M := maxx∈K f(cid:48)(cid:48)(x) ∈ ρ(cid:101)R≥0. Due to the 2 f (x) = f (0) + f(cid:48)(0) · x + f(cid:48)(cid:48)(ξ) · x2. fact that f (0) is minimal, we have f(cid:48)(cid:48)(ξ) 2 (3.2) 2 x = M As a corollary of Lem. 4 and Thm. 8.(iv), we have 2 x. Therefore, f(cid:48)(0) ≤ M f(cid:48)(0) · x + 2 x2 and −f(cid:48)(0) ≤ M Thus −f(cid:48)(0) · x ≤ M x < 0, we get f(cid:48)(0) ≤ − M follows by Rem. 28.(ii). · x2 = f (x) − f (0) ≥ 0. 2 x since x > 0. Analogously, if we take 2 x and the conclusion (cid:3) Lemma 32. Let a, b ∈ ρ(cid:101)R with a < b and let f ∈ ρGC∞([a, b], ρ(cid:101)Rd) such that Lemma 33. Let a, b ∈ ρ(cid:101)R with a < b and let f ∈ ρGC∞([a, b], ρ(cid:101)R) such that f (x) = 0 for all sharply interior points x ∈ [a, b]. Then f = 0 on [a, b]. Now, we are able to prove the "second - derivative - test" for GSF. f (x0) = min! for some sharply interior x0 ∈ [a, b]. Then f(cid:48)(cid:48)(x0) ≥ 0. Vice versa, if f(cid:48)(x0) = 0 and f(cid:48)(cid:48)(x0) > 0, then f (x0) = min!. CALCULUS OF VARIATIONS FOR GF Proof. As above, we can assume x0 = 0. Let r ∈ ρ(cid:101)R>0 be such that B2r(0) =: U ⊆ X and f (0) = min! over U . Take any x ∈ ρ(cid:101)R such that 0 < x < r, so that [0, x] ⊆ U , and set K := [Brε(0)] (cid:98)f B2r(0) ⊆ U and M := maxx∈K f(cid:48)(cid:48)(cid:48)(x) ∈ ρ(cid:101)R≥0. By 13 Taylor's theorem 11, we obtain that for some ξ ∈ [0, x] f(cid:48)(cid:48)(0)x2 + 1 6 1 2 By assumption, we have that for all a ∈ R>0 f (x) = f (0) + f(cid:48)(0)x + f(cid:48)(cid:48)(cid:48)(ξ)x3. By Lemma 31, we know that f(cid:48)(0) = 0. Thus, we obtain for all a ∈ R>0 0 ≤ f (x) − f (0) + dρa. 1 2 1 6 f (x) − f (0) = f(cid:48)(cid:48)(0)x2 + f(cid:48)(cid:48)(cid:48)(ξ)x3 ≥ −dρa. 2 f(cid:48)(cid:48)(0)x2 + 1 assuming that a > A and dρA < r. We get f(cid:48)(cid:48)(0) ≥ −(cid:0)2 + M Therefore, also 1 conclusion follows from Lem. 27 as a → +∞. by Lem. 3. Since f(cid:48)(0) = 0, for all x ∈ Br(0), Taylor's formula gives Now assume that f(cid:48)(0) = 0 and f(cid:48)(cid:48)(0) > 0, so that f(cid:48)(cid:48)(0) > dρa for some a ∈ R>0 6 M x3 ≥ −dρa. In this inequality we can set x = dρa/3, (cid:1) dρa/3, and the 3 f(cid:48)(cid:48)(cid:48)(ξx)x3, 1 6 2 dρa + 1 6 f(cid:48)(cid:48)(cid:48)(ξx)x(cid:1). Now Mx → 0 as x → 0. 1 2 (cid:12)(cid:12)(cid:12)(cid:12) 1 f(cid:48)(cid:48)(0)x2 + f (x) − f (0) = where ξx ∈ [0, x]. Therefore, f (x) − f (0) > x2(cid:0) 1 (cid:12)(cid:12)(cid:12)(cid:12) ≤ 1 We can hence write f (x)−f (0) > x2(cid:0) 1 ∃s ∈ ρ(cid:101)R>0 : s < r, ∀x ∈ Bs(0) : − 1 2 dρa − 1 f(cid:48)(cid:48)(cid:48)(ξx)x Thus 6 6 which proves that x = 0 is a local minimum. 1 6 dρa < 4 dρa(cid:1) = x2 1 4 1 4 dρa. f(cid:48)(cid:48)(cid:48)(ξx)x < 4 dρa ≥ 0 for all x ∈ Bs(0), (cid:3) For the generalization of Lem. 31 and Lem. 33 to the multivariate case, one can proceed as above, using the ideas of [23]. Note, however, that we do not need this generalization in the present work. 4. First variation and critical points In this section, we define the first variation of a functional and prove that some classical results have their counterparts in this generalized setting, for example the fundamental lemma (Lem. 37) or the connection between critical points and the Euler-Lagrange equations (Thm. 38). Definition 34. If a, b ∈ ρ(cid:101)R and a < b, we define ρGC∞ 0 (a, b) := (cid:111) (cid:110) η ∈ ρGC∞(ρ(cid:101)R, ρ(cid:101)Rd) : η(a) = 0 = η(b) 0 (a, b) is an ρ(cid:101)R-module. . 0 . We also note here that ρGC∞ When the use of the points a, b is clear from the context, we adopt the simplified notation ρGC∞ One of the positive features of the use of GSF for the calculus of variations is their closure with respect to composition. For this reason, the next definition of functional is formally equal to the classical one, though it can be applied to arbitrary generalized functions F and u. 14 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO Theorem 35. Let a, b ∈ ρ(cid:101)R with a < b. Let u ∈ ρGC∞([a, b], ρ(cid:101)Rd) and let F ∈ ρGC∞([a, b] × ρ(cid:101)Rd × ρ(cid:101)Rd, ρ(cid:101)R) and define b I(u) := F (t, u, u) dt. (4.1) Let η ∈ ρGC∞ 0 , then δI(u; η) := d ds I(u + sη) a = a (cid:18) b η (cid:12)(cid:12)(cid:12)(cid:12)s=0 (cid:19) F u(t, u, u) dt. Fu(t, u, u) − d dt Proof. We have (we use Thm. 10, Thm. 15 and Lemma 26) (cid:12)(cid:12)(cid:12)(cid:12)s=0 d ds I(u + sη) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)s=0 (cid:12)(cid:12)(cid:12)(cid:12)s=0 dt b = = = d ds b b a a F (t, u + sη, u + s η) dt a ∂ ∂s F (t, u + sη, u + s η) ηFu(t, u, u) + ηF u(t, u, u) dt (cid:18) b η a (cid:18) = [ηF u(t, u, u)]b a + b = η a Fu(t, u, u) − d dt Fu(t, u, u) − d dt (cid:19) F u(t, u, u) dt. (cid:19) F u(t, u, u) dt We call δI(u; η) the first variation of I. In addition we call u ∈ ρGC∞([a, b], ρ(cid:101)Rd) a critical point of I if δI(u; η) = 0 for all η ∈ ρGC∞ 0 . (cid:3) To prove the fundamental lemma of the calculus of variations, Lem. 37, we first show that every GSF can be approximated using generalized strict delta nets. Lemma 36. Let a, b ∈ ρ(cid:101)R be such that a < b and let f ∈ ρGC∞([a, b], ρ(cid:101)R). Let x ∈ [a, b] and R ∈ ρ(cid:101)R>0 be such that BR(x) ⊆ [a, b]. Assume that Gt ∈ ρGC∞(ρ(cid:101)R, ρ(cid:101)R) satisfy ´ −R Gt = 1 for t ∈ ρ(cid:101)R>0 small. R (i) (ii) For t small, (Gt)t∈ρ(cid:101)R>0 ∀δ ∈ ρ(cid:101)R>0 ∃ρ ∈ ρ(cid:101)R>0 ∀t ∈ Bρ(0) ∩ ρ(cid:101)R>0 ∀y ∈ [−R,−δ] ∪ [δ, R] : Gt(y) = 0. (4.2) (iii) ∃M ∈ ρ(cid:101)R>0 ∃ρ ∈ ρ(cid:101)R∀t ∈ Bρ(0) : ´ −R Gt(y) dy ≤ M . is zero outside every ball Bδ(0), 0 < δ < R, i.e. R R Then Moreover lim t→0+ ´ −R −R f (x − y)Gt(y) dy = R f (x − y)Gt(y) dy = f (x). ´ x−R f (y)Gt(x − y) dy. x+R Proof. We only have to generalize the classical proof concerning limits of convolu- tions with strict delta nets. We first note that R −R f (x − y)Gt(y) dy = f (y)Gt(x − y) dy x+R x−R CALCULUS OF VARIATIONS FOR GF 15 so that these integrals exist because (x − R, x + R) = BR(x) ⊆ [a, b]. Using (i), for t small, let's say for 0 < t < S ∈ ρ(cid:101)R>0, we get (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = f (x − y)Gt(y) dy − f (x) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) −R R R −R R [f (x − y) − f (x)] Gt(y) dy (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ f (x − y) − f (x) · Gt(y) dy. −R we have (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) For each r ∈ ρ(cid:101)R>0, sharp continuity of f at x yields f (x − y) − f (x) < r for all y such that y < δ ∈ ρ(cid:101)R>0, and we can take δ < R. By (ii), for 0 < t < min(ρ, S), (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ r The right hand side of (4.3) can be taken arbitrarily small in ρ(cid:101)R>0 because [−δ, δ] (cid:98)f ρ(cid:101)R, (iii) and because of the extreme value theorem 20 applied to the GSF Gt. (cid:3) Lemma 37 (Fundamental Lemma of the Calculus of Variations). Let a, b ∈ ρ(cid:101)R such that a < b, and let f ∈ ρGC∞([a, b], ρ(cid:101)R). If f (x − y)Gt(y) dy − f (x) Gt(y) dy. (4.3) −R −δ +δ R b f (t)η(t) dt = 0 for all η ∈ ρGC∞ 0 , (4.4) a then f = 0. Proof. Let x ∈ [a, b]. Because of Thm. 8.(iv) and Lem. 4, without loss of generality we can assume that x is a sharply interior point, so that BR(x) ⊆ [a, b] for some , where R ∈ ρ(cid:101)R>0. Let ϕ ∈ D[−1,1](R) be such that x ∈ R and t ∈ ρ(cid:101)R>0, and Gt(x) := [Gt,ε(xε)] for all x ∈ ρ(cid:101)R. Then, for t sufficiently small, we have Gt(x − .) ∈ ρGC∞ a f (y)Gt(x − y) dy = 0. For t small, we both have that Gt(x− .) = 0 on [a, x− R]∪ [x + R, b] and the assumptions of Lem. 36 hold. Therefore ϕ = 1. Set Gt,ε(x) := 1 tε 0 and (4.4) yields (cid:16) x (cid:17) ´ ´ ϕ tε b a R = −R 0 = b f (y)Gt(x − y) dy = f (x − y)Gt(y) dy, x+R x−R f (y)Gt(x − y) dy = (cid:3) and Lem. 36 hence yields f (x) = 0. Thus we obtain the following Theorem 38. Let a, b ∈ ρ(cid:101)R such that a < b, and let u ∈ ρGC∞([a, b], ρ(cid:101)Rd). Then u solves the Euler-Lagrange equations for I given by (4.1), if and only if δI(u; η) = 0 for all η ∈ ρGC∞ u is a critical point of I. 0 , i.e. if and only if Fu − d dt F u = 0 (4.5) 5. second variation and minimizers As in the classical case (see e.g. [8]), thanks to the extreme value theorem 20 and the property of the interval [a, b] of being functionally compact, we can naturally define a topology on the space ρGC∞([a, b], ρ(cid:101)Rd): 16 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO Definition 39. Let a, b ∈ ρ(cid:101)R, with a < b. Let m ∈ N0 and v ∈ ρGC∞([a, b], ρ(cid:101)Rd). (cid:18)(cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) , (cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12)(cid:19) ∈ ρ(cid:101)R, (cid:107)v(cid:107)m := max n≤m 1≤i≤d max dtn vi(Mni) dtn vi(mni) Then where Mni, mni ∈ [a, b] satisfy dn ∀t ∈ [a, b] : dtn vi(mni) ≤ dn dtn vi(t) ≤ dn dtn vi(Mni). The following result permits to calculate the (generalized) norm (cid:107)v(cid:107)m using any net (vε) that defines v. (i) (cid:20) Lemma 40. Under the assumptions of Def. 39, let a = [aε] and b = [bε] be such that aε < bε for all ε. Then we have: If the net (vε) defines v, then (cid:107)v(cid:107)m = (cid:107)v(cid:107)m ≥ 0; ε(t)(cid:12)(cid:12)(cid:21) ∀c ∈ ρ(cid:101)R : (cid:107)c · v(cid:107)m = c · (cid:107)v(cid:107)m; For all u ∈ ρGC∞([a, b], ρ(cid:101)Rd), we have (cid:107)u+v(cid:107)m ≤ (cid:107)u(cid:107)m+(cid:107)v(cid:107)m and (cid:107)u·v(cid:107)m ≤ cm · (cid:107)u(cid:107)m · (cid:107)v(cid:107)m for some cm ∈ ρ(cid:101)R>0. (ii) (iii) (cid:107)v(cid:107)m = 0 if and only if v = 0; (iv) (v) max n≤m 1≤i≤d (cid:12)(cid:12) dn maxt∈[aε,bε] dtn vi ; Proof. By the standard extreme value theorem applied ε-wise, we get the existence of ¯mniε, ¯Mniε ∈ [aε, bε] such that dn dtn vi ε( ¯mniε) ≤ dn ε(t) ≤ dn ε( ¯Mniε). dtn vi dtn vi Hence(cid:12)(cid:12) dn dtn vi max n≤m 1≤i≤d  max n≤m 1≤i≤d max t∈[aε,bε] ∀t ∈ [aε, bε] : ε(t)(cid:12)(cid:12) ≤ max(cid:0)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) dn dtn vi dtn vi ε(t) ε(t) (cid:12)(cid:12)(cid:12)(cid:12) max t∈[aε,bε] dtn vi max dtn vi ε( ¯mniε) dtn vi n≤m 1≤i≤d ε( ¯Mniε)(cid:12)(cid:12)(cid:1). Thus ε( ¯mniε)(cid:12)(cid:12) ,(cid:12)(cid:12) dn (cid:18)(cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) ≤ max (cid:12)(cid:12)(cid:12)(cid:12) ,  max  = (cid:18)(cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) , (cid:18)(cid:12)(cid:12)(cid:12)(cid:12) dn (cid:12)(cid:12)(cid:12)(cid:12) dn dtn vi( ¯mni) n≤m 1≤i≤d dtn vi (cid:12)(cid:12)(cid:12)(cid:12) , ε( ¯mniε) max max = max n≤m 1≤i≤d But ¯mniε, ¯Mniε ∈ [aε, bε], so dtn vi ε( ¯Mniε) dtn vi ε( ¯Mniε) (cid:12)(cid:12)(cid:12)(cid:12)(cid:19) . dtn vi( ¯Mni) . (cid:12)(cid:12)(cid:12)(cid:12)(cid:19) (cid:12)(cid:12)(cid:12)(cid:12)(cid:19) = This proves both that (cid:107)v(cid:107)m is well-defined, i.e. it does not depend on the particular choice of points mni, Mni as in Def. 39, and the claim (i). The remaining properties (ii) - (v) follows directly from (i) and the usual properties of standard Cm-norms. (cid:3) Using these ρ(cid:101)R-valued norms, we can naturally define a topology on the space ρGC∞([a, b], ρ(cid:101)Rd). Definition 41. Let a, b ∈ ρ(cid:101)R, with a < b. Let m ∈ N, u ∈ ρGC∞([a, b], ρ(cid:101)Rd), r ∈ ρ(cid:101)R>0, then (cid:110) v ∈ ρGC∞([a, b], ρ(cid:101)Rd) (cid:107)v − u(cid:107)m < r ∀u ∈ U ∃m ∈ N∃r ∈ ρ(cid:101)R>0 : Bm If U ⊆ ρGC∞([a, b], ρ(cid:101)Rd), then we say that U is a sharply open set if r (u) ⊆ U. r (u) := (cid:111) Bm (ii) (i) CALCULUS OF VARIATIONS FOR GF 17 As in [13, Thm. 2], one can easily prove that sharply open sets form a topology on ρGC∞([a, b], ρ(cid:101)Rd). Using this topology, we can define when a curve is a minimizer numbers a, b ∈ ρ(cid:101)R, a < b. E.g. they can also both be infinite. Definition 42. Let a, b ∈ ρ(cid:101)R, with a < b and let u ∈ ρGC∞([a, b], ρ(cid:101)Rd), then of the functional I. Note explicitly that there are no restrictions on the generalized For all p, q ∈ ρ(cid:101)Rd, we set (cid:110) v ∈ ρGC∞([a, b], ρ(cid:101)Rd) v(a) = p, v(b) = q ρGC∞ bd(p, q) := (cid:111) (i) . (ii) We say that u is a local minimizer of I in ρGC∞ Note that ρGC∞ values". bd(0, 0) = ρGC∞ ∃r ∈ ρ(cid:101)R>0 ∃m ∈ N∀v ∈ Bm 0 . The subscript "bd" stands here for "boundary bd(p, q) if u ∈ ρGC∞ bd(p, q) : I(v) ≥ I(u) r (u) ∩ ρGC∞ bd(p, q) and (5.1) (iii) We define the second variation of I in direction η ∈ ρGC∞ 0 as δ2I(u; η) := Note also explicitly that the points p, q ∈ ρ(cid:101)Rd can have infinite norm, e.g. pε → +∞ as ε → 0. We calculate, by using the standard Einstein's summation conven- tions I(u + sη). (cid:12)(cid:12)(cid:12)(cid:12)0 d2 ds2 (cid:12)(cid:12)(cid:12)(cid:12)0 ∂2 ∂s2 a b (cid:12)(cid:12)(cid:12)(cid:12)0 δ2I(u; η) = = = d2 ds2 b a b a F (t, u + sη, u + s η) dt F (t, u + sη, u + s η) dt Fuiuj (t, u, u)ηiηj + 2Fui uj (t, u, u)ηi ηj + F ui uj (t, u, u) ηi ηj dt, which we abbreviate as b δ2I(u; η) = a Fuu(t, u, u)ηη + 2Fu u(t, u, u)η η + F u u(t, u, u) η η dt. (i) (ii) bd(p, q). Then The following results establish classical necessary and sufficient conditions to decide if a function u is a minimizer for the given functional (4.1). δI(u; η) = 0 for all η ∈ ρGC∞ 0 ; δ2I(u; η) ≥ 0 for all η ∈ ρGC∞ 0 . Theorem 43. Let a, b ∈ ρ(cid:101)R with a < b, let F ∈ ρGC∞([a, b] × ρ(cid:101)Rd × ρ(cid:101)Rd, ρ(cid:101)R), let p, q ∈ ρ(cid:101)Rd and let u be a local minimizer of I in ρGC∞ 0 , the map s ∈ ρ(cid:101)R (cid:55)→ Proof. Let r ∈ ρ(cid:101)R>0 be such that (5.1) holds. Since η ∈ ρGC∞ the sharp topology in its codomain. Therefore, we can find ¯r ∈ ρ(cid:101)R>0 such that bd(p, q) is well defined and continuous with respect to the trace of This shows that the GSF s ∈ B¯r(0) (cid:55)→ I(u + sη) ∈ ρ(cid:101)R has a local minimum at bd(p, q) for all s ∈ B¯r(0). We hence have I(u + sη) ≥ I(u). s = 0. Now, we employ Lem. 31 and Lem. 33 and thus the claims are proven. (cid:3) Theorem 44. Let a, b ∈ ρ(cid:101)R with a < b and p, q ∈ ρ(cid:101)Rd. Let u ∈ ρGC∞ u + sη ∈ ρGC∞ u + sη ∈ Bm r (u) ∩ ρGC∞ bd(p, q) be such that (i) (ii) δI(u; η) = 0 for all η ∈ ρGC∞ 0 . δ2I(v; η) ≥ 0 for all η ∈ ρGC∞ r ∈ ρ(cid:101)R>0 and m ∈ N. 0 and for all v ∈ Bm r (u) ∩ GC∞ bd (p, q), where 18 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO 2r(u) ∩ GC∞ Moreover, if δ2I(v; η) > 0 for all η ∈ ρGC∞ Then u is a local minimizer of the functional I in ρGC∞ bd (p, q), then I(v) > I(u) for all v ∈ Bm v ∈ Bm (cid:107)v − u(cid:107)m > 0. bd (p, q), we set ψ(s) := I(u + s(v − u)) ∈ ρ(cid:101)R for r (u) ∩ GC∞ Proof. For any v ∈ Bm all s ∈ B1(0) so that u + s(v − u) ∈ Bm r (u). Since (v − u)(a) = 0 = (v − u)(b), we have v − u ∈ ρGC∞ 0 , and properties (i), (ii) yield ψ(cid:48)(0) = δI(u; v − u) = 0 and ψ(cid:48)(cid:48)(s) = δ2I(u + s(v − u); v − u) ≥ 0 for all s ∈ B1(0). We claim that s = 0 is a minimum of ψ. In fact, for all s ∈ B1(0) by Taylor's theorem 11 such that (cid:107)η(cid:107)m > 0 and for all bd (p, q) such that bd(p, q). r (u) ∩ GC∞ 0 ψ(s) = ψ(0) + sψ(cid:48)(0) + ψ(cid:48)(cid:48)(ξ) s2 2 2 ψ(cid:48)(cid:48)(ξ) ≥ 0. Finally, for some ξ ∈ [0, s]. But ψ(cid:48)(0) = 0 and hence ψ(s) − ψ(0) = s2 Lem. 27 yields lim 2r(u) ∩ GC∞ r (u) ∩ GC∞ bd (p, q), and take v ∈ Bm s→1− ψ(s) = I(v) ≥ ψ(0) = I(u), bd (p, q), then ψ(cid:48)(cid:48)(ξ) = 0 and hence I(v) = I(u). Now, assume that δ2I(v; η) > 0 for all η ∈ ρGC∞ r (u) ∩ GC∞ which is our conclusion. Note explicitly that if δ2I(v; η) = 0 for all η ∈ ρGC∞ 0 and for all v ∈ Bm 0 such that (cid:107)η(cid:107)m > 0 and for all As above set ψ(s) := I(u + s(v − u)) ∈ ρ(cid:101)R for all s ∈ B3/2(0) so that u + s(v − u) ∈ bd (p, q) such that (cid:107)v − u(cid:107)m > 0. v ∈ Bm 2r(u). We have ψ(cid:48)(0) = 0 and ψ(cid:48)(cid:48)(s) = δ2I(u + s(v − u); v − u) > 0 for all s ∈ Bm B3/2(0) because (cid:107)v−u(cid:107)m > 0. Using Taylor's theorem, we get ψ(1) = ψ(0)+ 1 2 ψ(cid:48)(cid:48)(ξ) for some ξ ∈ [0, 1]. Therefore ψ(1) − ψ(0) = I(v) − I(u) = 1 (cid:3) Lemma 45. Let (ak)k∈N, (bk)k∈N and (ck)k∈N be sequences in ρ(cid:101)R>0. Assume that f ∈ ρGC∞([a1, b1], ρ(cid:101)R). Finally, let ak < t < bk for all k ∈ N, then it holds that → 1 in the sharp topology as k → +∞. Let both (ak)k, (bk)k → 0 and 2 ψ(cid:48)(cid:48)(ξ) > 0. ck ak+bk Proof. We can apply the integral mean value theorem for each ε and each defining net (fε) of f to get the existence of τk ∈ [t − ak, t + bk] such that Now, we take the limit for k → ∞, and the claim follows by assumption and by (cid:3) Thm. 8.(iv), i.e. by sharp continuity of f . We now derive the so-called necessary Legendre condition: Theorem 46. Let a, b ∈ ρ(cid:101)R with a < b and let u ∈ ρGC∞([a, b], ρ(cid:101)Rd) be a minimizer of the functional I. Then is positive semi definite for all t ∈ [a, b], i.e. F u u(t, u(t), u(t)) ∀λ = (λ1, . . . , λd) ∈ ρ(cid:101)Rd. Proof. Let λ = [λε] ∈ ρ(cid:101)Rd and k, h ∈ N be arbitrary. Let t = [tε] ∈ [a, b]. We F ui uj (t, u(t), u(t))λiλj ≥ 0, (5.2) can assume that t is a sharply interior point, because otherwise we can use sharp continuity of the left hand side of (5.2) and Lem. 27. We can also assume that λ is componentwise invertible because of Lem. 30. We want to mimic the classical f (t) = lim k→∞ 1 ck t+bk f (s) ds. t−ak f (τk) = = 1 bk + ak ck bk + ak t+bk t−ak 1 ck t+bk t−ak f (s) ds f (s) ds. CALCULUS OF VARIATIONS FOR GF 19 proof of [21, Thm. 1.3.2], but considering a "regularized" version of the triangular function used there (see Fig. 5.1). In particular: (1) the smoothed triangle must have an infinitesimal height which is proportional to λ, and we will take ρk ε as this infinitesimal; (2) in the proof we need that the derivative at t is equal to λ, and this justifies the drawing of the peak in Fig. 5.1; (3) to regularize the singular points of the triangular function, we need a smaller infinitesimal, and we can take e.g. ρ2k ε . So, consider a net of smooth functions ϑε on [aε, bε], such that the following properties hold: ϑε(x) = 0, for x ≤ tε − ρk (i) ϑε(x) = 0, for x ≥ tε + ρk (ii) (iii) ϑε(x) = λ(x − tε) + ρk (iv) ϑε(x) = −λ(x − tε) + ρk ϑε(x) ≤ ρk ε · λ + 2ρ2k (v) ϑε(x) ≤ 2λ for all x (vi) ε − ρ2k ε . ε + ρ2k ε . ε λ for x ∈ [tε − ρk ε λ for x ∈ [tε + ρ2k ε λ. ε , tε]. ε , tε + ρk ε − ρ2k ε ]. ε + ρ2k 2ρ2k ε λερk ε 2ρ2k ε 2ρ2k ε a tε − ρk ε tε tε + ρk ε b Figure 5.1. This figure illustrates the function ϑε we are consid- ering (blue). The dotted green triangle symbolizes the function which is used in the classical proofs of the Legendre necessary con- dition (cf. [21, Thm. 1.3.2]). The net (ϑε) defines a GSF ϑ := [ϑε(−)] ∈ ρGC∞ point. Setting for simplicity ak := dρk + dρ2k, by assumption we have 0 ≤ δ2I(u, ϑ) = (5.3) Now, setting M := max[a,b] Fuu(t, u, u) and N := max[a,b] Fu u(t, u, u), by (v) we Fuu(t, u, u)ϑϑ + 2F uu(t, u, u) ϑϑ + F u u(t, u, u) ϑ ϑ dt. 0 because t is a sharply interior t−ak t+ak t+ak Fuu(t, u, u)ϑϑ dt where we used the evident notation Gk = O(cid:0)dρk(cid:1) to denote that there exists some A ∈ ρ(cid:101)R>0 such that Gk ≤ A · dρk for all k ∈ N. Using (v) and (vi), we analogously t−ak (cid:12)(cid:12)(cid:12)(cid:12) ≤ M · ϑ(t)2 · 2ak = O(cid:0)dρ3k(cid:1) , (cid:12)(cid:12)(cid:12)(cid:12) ≤ 4N · ϑ(t) · ak · λ = O(cid:0)dρ2k(cid:1) . (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) have t+ak t−ak F uu(t, u, u) ϑϑ dt 20 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO Note that there always exists C ∈ ρ(cid:101)R such that λ ≤ Cdρk. Therefore lim k→+∞ 1 t+ak 2dρk t−ak Fuu(t, u, u)ϑϑ + 2F uu(t, u, u) ϑϑ dt = 0. (5.4) Using Lemma 45, (5.4), (5.3) and Lem. 27, we obtain that F u u(t, u(t) u(t)) ϑ(t) ϑ(t) = lim k→+∞ F u u(t, u, u) ϑ ϑ dt ≥ 0. But (iii) yields ϑ(t) = λ, and this concludes the proof. (cid:3) 1 t+ak 2dρk t−ak As in the classical case, Thm. 43.(ii) motivates to define the accessory integral 6. Jacobi fields b ψ(t, η, η) dt ∀η ∈ ρGC∞ 0 , Q(η) := a where (6.1) (6.2) for all t ∈ [a, b] and (l, v) ∈ ρ(cid:101)Rd × ρ(cid:101)Rd. Note that if u minimizes I, then ψ(t, l, v) := Fuu(t, u, u)ll + 2Fu u(t, u, u)lv + F u u(t, u, u)vv Q(η) ≥ 0 ∀η ∈ ρGC∞ 0 . As usual, we note that η = 0 is a minimizer of the functional Q and we are interested whether there are others. In order to solve this problem, we consider the Euler-Lagrange equations for Q, which are given by d dt ψ η(t, η, η) = ψη(t, η, η), (6.3) in other words d dt {F u u(t, u, u) η + Fu u(t, u, u)η} = Fu u(t, u, u) η + Fuu(t, u, u)η. Since u is given, (6.4) is an ρ(cid:101)R-linear system of second order equations in the unknown GSF η and with time dependent coefficients in ρ(cid:101)R. We call (6.4) the (6.4) Jacobi equations of I with respect to u. As in the classical setting, we define Definition 47. A solution η ∈ ρGC∞ field along u0 = u. 0 of the Jacobi equations (6.4) is called a Jacobi The following result confirms that the intuitive interpretation of a Jacobi field as the tangent space of a smooth family of solutions of the Euler-Lagrange equation still holds in this generalized setting. Lemma 48. Let u ∈ ρGC∞([−δ, δ] × [a, b], ρ(cid:101)Rd), where δ ∈ ρ(cid:101)R>0. We write us := u(s,−) for all s ∈ [−δ, δ]. Assume that each us satisfies the Euler-Lagrange equations (4.5): d dt Then F u(t, us, us) = Fu(t, us, us) ∀s ∈ [−δ, δ]. η(t) := d ds is a Jacobi field along u. us(t) ∀t ∈ [a, b] (cid:12)(cid:12)(cid:12)(cid:12)0 CALCULUS OF VARIATIONS FOR GF Proof. A straight forward calculation gives: F u(t, us, us) − Fu(t, us, us) (cid:19) 21 (cid:3) (cid:12)(cid:12)(cid:12)(cid:12)0 (cid:18) d dt 0 = = d ds d dt (F u u(t, u, u) η + Fu u(t, u, u)η) − Fu u(t, u, u) η − Fuu(t, u, u)η. 6.1. Conjugate points and Jacobi's theorem. The classical key result con- cerning Jacobi fields relates conjugate points and minimizers. The main aim of the present section is to derive this theorem in our generalized framework by extending the ideas of the proof of [21, Thm. 1.3.4]. A crucial notion is hence that of piecewise GSF: Definition 49. We call piecewise GSF an n-tuple (f1, . . . , fn) such that (i) For all i = 1, . . . , n there exist ai, ai+1 ∈ ρ(cid:101)R such that ai < ai+1 and fi ∈ ρGC∞([ai, ai+1], ρ(cid:101)Rd). Note that [a, b] = [a(cid:48), b(cid:48)] implies a = a(cid:48) and b = b(cid:48) because the relation ≤ is antisymmetric. Therefore, the points ai, ai+1 are uniquely determined by the set-theoretical function fi. For all i = 1, . . . , n, we have fi(ai+1) = fi+1(ai+1). (ii) Every pointwise GSF (f1, . . . , fn) defines a set-theoretical function: (iii) For all t ∈(cid:83)n We also use the arrow notation (f1, . . . , fn) : (cid:83)n i=1[ai, ai+1], we set (f1, . . . , fn)(t) := fi(t) if t ∈ [ai, ai+1]. i=1[ai, ai+1] −→ ρ(cid:101)Rd to say that both (i) and (ii) hold. Remark 50. (i) (v) (ii) (iii) Clearly, t ∈ [ai, ai+1] ∩ [ai+1, ai+2] implies t = ai+1, so that condition (ii) yields that the evaluation (iii) is well defined. Since the order relation ≤ is not a total one, we do not have that [ai, ai+1] ∪ [ai+1, ai+2] = [ai, ai+2]. piecewise GSF (f1, f2), then neither the GSF fi nor the points ai are uniquely determined by ν. For this reason, we prefer to stress our notations with piecewise GSF. (g1, . . . , gn) + (f1, . . . , fn) := (g1 + f1, . . . , gn + fn) and r · (f1, . . . , fn) := If ν : [a1, a2]∪ [a2, a3] −→ ρ(cid:101)Rd is a set-theoretical function originating from a symbols like (f1, f2)(t) ∈ ρ(cid:101)Rd. (iv) Every GSF f ∈ ρGC∞([a1, a2], ρ(cid:101)Rd) can be seen as a particular case of a If (g1, . . . , gn), (f1, . . . , fn) : (cid:83)n i=1[ai, ai+1] −→ ρ(cid:101)Rd and r ∈ ρ(cid:101)R, then also (r · f1, . . . , r · fn) are piecewise GSF, and we hence have a structure of ρ(cid:101)R- If (f1, . . . , fn) :(cid:83)n i=1[ai, ai+1] −→ ρ(cid:101)Rd and F ∈ ρGC∞(ρ(cid:101)Rd, ρ(cid:101)Rn), then we can define the composition F◦(f1, . . . , fn) := (F◦f1, . . . , F◦fn) :(cid:83)n ρ(cid:101)Rn. i=1[ai, ai+1] −→ Definition 51. Let x = [xε] ∈ ρ(cid:101)R, then ν(x) := sup(cid:8)b ∈ R xε = O(ρb (iii) dρ(x) := dρ− log xe ∈ ρ(cid:101)R>0. It is worth noting that −e : ρ(cid:101)R −→ R≥0 induces an ultrametric on ρ(cid:101)R that gener- Piecewise GSF inherit from their defining components a well-behaved differential and integral calculus. The former is even more general and taken from [1]. ε)(cid:9) ∈ R ∪ {+∞}. xe := e−ν(x) ∈ R≥0. module. (i) (ii) (vi) ates exactly the sharp topology, see e.g. [2, 11] and references therein. However, we 22 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO will not use this ultrametric structure in the present paper, and we only introduced it to get an invertible infinitesimal dρ(x) that goes to zero with x: it is in fact easy to show that x lim x→0 dρ(x) = 1 in the sharp topology. Definition 52. Let T ⊆ ρ(cid:101)R and let f : T −→ ρ(cid:101)Rd be an arbitrary set-theoretical function. Let t0 ∈ T be a sharply interior point of T . Then we say that f is differentiable at t0 if2 ∃m ∈ ρ(cid:101)Rd : f (t + h) − f (t0) − m · h lim h→0 dρ(h) = 0. (6.5) In this case, using Landau little-oh notation, we can hence write as h → 0. As in the classical case, (6.6) implies the uniqueness of m ∈ ρ(cid:101)Rd, so that we can f (t + h) = f (t0) + m · h + o(dρ(h)) define f(cid:48)(t0) := f (t0) := m, and the usual elementary rules of differential calculus. By the Fermat-Reyes theorem, this definition of derivative generalizes that given for GSF. In particular, this notion of derivative applies to the set-theoretical function induced by a piecewise GSF (f1, . . . , fn). We therefore have that (f1, . . . , fn)(−) is differentiable at each ai < t < ai+1, and (f1, . . . , fn)(cid:48)(t) = f(cid:48) i (t), but clearly there is no guarantee that (f1, . . . , fn)(−) is also differentiable at each point ai. The notion of definite integral is naturally introduced in the following (6.6) i=1[ai, ai+1] −→ ρ(cid:101)Rd be a piecewise GSF, then Definition 53. Let (f1, . . . , fn) :(cid:83)n an+1 (f1, . . . , fn)(t) dt := ai+1 fi(t) dt. n(cid:88) a1 i=1 ai Since our main aim in using piecewise GSF is to prove Jacobi's theorem, we do not need to prove that the usual elementary rules of integration hold, since we will always reduce to integrals of GSF. Having a notion of derivative and of definite integral, also for piecewise GSF we can consider functionals ν := (f1, . . . , fn), a1 = a, an = b =⇒ I(ν) := F (t, ν(t), ν(t))dt ∈ ρ(cid:101)R, b (6.7) and the concept of piecewise GSF (global) minimizer : I(ν) ≤ I(ν) for all ν ∈ ρGC∞ 0 . For the proof of Jacobi's theorem, we will only need this particular notion of global minimizer. Note explicitly that in (6.7) we only need existence of right and left derivatives of GSF, because of Def. 53 and of Def. 14 of definite integral of GSF. a Classically, several proofs of Jacobi's theorem use both some form of implicit function theorem and of uniqueness of solution for linear ODE. Theorem 54 (Implicit function theorem). Let U ⊆ ρ(cid:101)Rn, V ⊆ ρ(cid:101)Rd be sharply open sets. Let F ∈ ρGC∞(U × V, ρ(cid:101)Rd) and (x0, y0) ∈ U × V . If ∂2F (x0, y0) is invertible in L(ρ(cid:101)Rd, ρ(cid:101)Rd), then there exists a sharply open neighbourhood U1 × V1 ⊆ U × V of (x0, y0) such that ∀x ∈ U1 ∃!yx ∈ V1 : F (x, yx) = F (x0, y0). (6.8) 2This definition is based on [1, Def. 2.2]. CALCULUS OF VARIATIONS FOR GF 23 Moreover, the function f (x) := yx for all x ∈ U1 is a GSF f ∈ ρGC∞(U1, V1) and satisfies Df (x) = − (∂2F (x, f (x))) −1 ◦ ∂1F (x, f (x)). (6.9) Proof. The usual deduction of the implicit function theorem from the inverse func- tion theorem in Banach spaces can be easily adapted using Thm. 12 and noting (cid:3) that det [∂2F (−,−)] is a GSF such that det [∂2F (x0, y0)] ∈ ρ(cid:101)R>0. In the next theorem, the dependence of the entire theory on the initial infinites- imal net ρ = (ρε) ↓ 0 plays an essential role. Indirectly, the same important role will reverberate in the final Jacobi's theorem. Theorem 55 (Solution of first order linear ODE). Let A ∈ ρGC∞([a, b], ρ(cid:101)Rd×d), where a, b ∈ ρ(cid:101)R, a < b, and t0 ∈ [a, b], y0 ∈ ρ(cid:101)Rd. Assume that (cid:12)(cid:12)(cid:12)(cid:12) ≤ −C · log dρ ∀t ∈ [a, b], where C ∈ R>0. Then there exists one and only one y ∈ ρGC∞([a, b], ρ(cid:101)Rd) such that (cid:12)(cid:12)(cid:12)(cid:12) A(s) ds (6.10) t0 t (cid:40) Moreover, this y is given by y(t) = exp A(s) ds y(cid:48)(t) = A(t) · y(t) y(t0) = y0 if t ∈ [a, b] (cid:16)´ (cid:17) t t0 (cid:104) (cid:17) · y0 for all t ∈ [a, b]. (cid:16)´ (cid:17)(cid:105) (cid:16)´ t t0 (6.11) exp tε t0ε t t0 (cid:19) A(s) ds = exp A(s) ds Aε(s) ds (cid:16)´ (cid:18) , where t = [tε], Proof. We first note that exp ≤ e−C log dρ ≤ dρ−C. because for all t ∈ [a, b], we have t0 = [t0ε] and A(s) = [Aε(sε)] ∈ ρ(cid:101)Rd×d. This exponential matrix in ρ(cid:101)Rd×d is a GSF (cid:17) · y0 are ρ-moderate. Analogously, (cid:16)− Therefore, all values of y(t) = exp one can prove that also y(k)(t) are moderate for all k ∈ N and t ∈ [a, b]. Considering that derivatives can be calculated ε-wise, we have that this GSF y satisfies (6.11), and this proves the existence part. To show uniqueness, we can proceed as in the smooth case. Assume that z ∈ for all t ∈ [a, b]. Since h(cid:48) = −A · h, we have (hz)(cid:48) = h(cid:48)z + hz(cid:48) = −Ahz + hAz = −Ahz + Ahz = 0. From uniqueness of primitives of GSF, Thm. 13, we have that h· z = h(t0)· z(t0) = y0. Therefore z = h−1 · y0. (cid:3) ρGC∞([a, b], ρ(cid:101)Rd) satisfies (6.11), and set h(t) := exp If α, β ∈ ρ(cid:101)R, we write α = OR(β) to denote that there exists C ∈ R>0 such A(s) ds A(s) ds (cid:17) ´ t t0 t t0 that α ≤ C · β. Therefore, assumption (6.10) can be written as OR(log dρ). Note that this assumption is weaker, in general, than A(s) ds = ´ t t0 (b − a) · max t∈[a,b] A(t) = OR(log dρ). theorem. The following result is the key regularity property we need to prove Jacobi's Lemma 56. Let a, a(cid:48), b ∈ ρ(cid:101)R, with a < a(cid:48) < b, and let K ∈ ρGC∞([a, b] × ρ(cid:101)Rd × ρ(cid:101)Rd, ρ(cid:101)R). Let ν = (η, β) : [a, a(cid:48)] ∪ [a(cid:48), b] −→ ρ(cid:101)Rd be a piecewise GSF which satisfies the Euler-Lagrange equation Ku(t, ν(t), ν(t)) − d dt K u(t, ν(t), ν(t)) = 0 ∀t ∈ [a, a(cid:48)) ∪ (a(cid:48), b]. (6.12) ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO 24 Finally, assume that det(cid:0)K ui uj (a(cid:48), η(a(cid:48)), η(a(cid:48)))i,j=i,...,d (cid:1) ∈ ρ(cid:101)R is invertible, then ν(t) = η(a(cid:48)). (6.13) ν(t) = lim t→a(cid:48) a(cid:48)<t In particular, if β ≡ 0[a(cid:48),b], then η(a(cid:48)) = 0. lim t→a(cid:48) t<a(cid:48) Proof. Set Φ(t, l, v, q) := K u(t, l, v) − q for all t ∈ [a, b] and all l, v, q ∈ ρ(cid:101)Rd. For simplicity, set (t0, l0, v0, q0) := (a(cid:48), η(a(cid:48)), η(a(cid:48)), K u(a(cid:48), η(a(cid:48)), η(a(cid:48))). Our assumption on the invertibility of K u u(a(cid:48), η(a(cid:48)), η(a(cid:48))) = ∂vΦ(t0, l0, v0, q0) makes it possible to apply the implicit function Thm. 54 to conclude that there exists a neighbourhood T × L × V × Q of (t0, l0, v0, q0) such that ∀(t, l, q) ∈ T × L × Q∃!v ∈ V : Φ(t, l, v, q) = Φ(t0, l0, v0, q0). (6.14) But Φ(t0, l0, v0, q0) = K u(a(cid:48), η(a(cid:48)), η(a(cid:48))) − q0 = 0. Moreover, the unique function ϕ defined by Φ(t, l, ϕ(t, l, q), q) = 0 for all (t, l, q) ∈ T × L × Q is a GSF ϕ ∈ ρGC∞(T × L × Q, V ). Now, for all t ∈ [a, a(cid:48)) ∪ (a(cid:48), b], we have Φ(t, ν(t), ν(t), K u(t, ν(t), ν(t))) = K u(t, ν(t), ν(t)) − K u(t, ν(t), ν(t)) = 0. Therefore, uniqueness in (6.14) yields ν(t) = ϕ (t, ν(t), K u(t, ν(t), ν(t))) ∀t ∈ [a, a(cid:48)) ∪ (a(cid:48), b]. (6.15) We now integrate the Euler-Lagrange equation (6.12) on [a, t], obtaining K u(t, ν(t), ν(t)) = Ku(s, ν(s), ν(s)) ds + K u(a, η(a), η(a)) ∀t ∈ [a, a(cid:48)) ∪ (a(cid:48), b]. t a t (cid:18) This entails that we can write ν(t) = ϕ t, ν(t), a Ku(s, ν(s), ν(s)) ds + K u(a, η(a), η(a)) But the function t ∈ [a, a(cid:48)) ∪ (a(cid:48), b] (cid:55)→ on the left and on the right of a(cid:48) because on [a, a(cid:48)) and on (a(cid:48), b] it is a GSF; in fact for t < a(cid:48) we have a Ku(s, ν(s), ν(s)) ds ∈ ρ(cid:101)Rd has equal limits ´ t ∀t ∈ [a, a(cid:48)) ∪ (a(cid:48), b]. (6.16) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) a t Ku(s, ν(s), ν(s)) ds − a(cid:48) a (cid:19) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ Ku(s, ν(s), ν(s)) ds ≤ max t∈[a,a(cid:48)] Ku(s, η(s), η(s)) · t − a(cid:48), and this goes to 0 as t → a(cid:48), t < a(cid:48). Analogously we can proceed for t > a(cid:48) using β. Therefore t lim t→a(cid:48) t<a(cid:48) Ku(s, ν(s), ν(s)) ds = lim t→a(cid:48) t>a(cid:48) a a t Ku(s, ν(s), ν(s)) ds. ν(t) = η(a(cid:48)) = limt→a(cid:48) a(cid:48)<t Applying this equality in (6.16), we get limt→a(cid:48) t<a(cid:48) claimed. Finally, if β ≡ 0[a(cid:48),b], then limt→a(cid:48) a(cid:48)<t In the following definition and below, we use the complete notation ρGC∞ (see Def. 34). Definition 57. Let a, a(cid:48), b ∈ ρ(cid:101)R, where a < a(cid:48) < b. We call a(cid:48) conjugate to a w. r. t. the variational problem (4.1) if there exists a non identically vanishing Jacobi field η ∈ ρGC∞ 0 (a, a(cid:48)) along u[a,a(cid:48)] such that η(a) = 0 = η(a(cid:48)), where ψ is given by (6.2). ν(t) as (cid:3) 0 (a, a(cid:48)) ν(t) = 0. CALCULUS OF VARIATIONS FOR GF 25 In order to prove the important Jacobi's theorem in the present generalized context, which shows that we cannot have minimizers if there are interior points which are conjugate to a, we finally need the following Lemma 58. Let u ∈ ρGC∞([a, b], ρ(cid:101)Rd) and let a(cid:48) ∈ (a, b). Let η ∈ ρGC∞ 0 (a, a(cid:48)) be a Jacobi field along u[a,a(cid:48)], with η(a) = 0 = η(a(cid:48)). Then a(cid:48) ψ(t, η, η) dt = 0. Proof. Since ψ is ρ(cid:101)R-homogeneous of second order in (η, η),we have 2ψ(t, η, η) = ψη(t, η, η)η + ψ η(t, η, η) η. Thus we calculate: a a(cid:48) 2 ψ(t, η, η) dt = a ηψη(t, η, η) + ηψ η(t, η, η) dt (cid:19) a(cid:48) (cid:18) a a(cid:48) η = a ψη(t, η, η) − d dt ψ η(t, η, η) dt by integration by parts = 0 since η is a Jacobi field. (cid:3) After these preparations we can finally prove Theorem 59 (Jacobi). Let a, b ∈ ρ(cid:101)R be such that a < b. Suppose that F ∈ ρGC∞([a, b] × ρ(cid:101)Rd × ρ(cid:101)Rd, ρ(cid:101)R) and u ∈ ρGC∞([a, b], ρ(cid:101)R) are such that a(cid:48) ∈ (a, b) is conjugate to a det F u u(t, u(t), u(t)) ∈ ρ(cid:101)R is invertible for all t ∈ [a, b]. (i) (ii) (iii) For all t ∈ [a, a(cid:48)] Fu u(s, u(s), u(s)) − Fuu(s, u(s), u(s)) (cid:20) d ds t a(cid:48) u u (s, u(s), u(s)) · −F −1 t (cid:21) ds = = OR(log dρ) b a a(cid:48) a b a(cid:48) a(cid:48) −F −1 u u (s, u(s), u(s)) · d ds F u u(s, u(s), u(s)) ds = OR(log dρ). Then u cannot be a local minimizer of I. Therefore, for any r ∈ ρ(cid:101)R>0 there exists bd(u(a), u(b)) and m ∈ N such that (cid:107)v − u(cid:107)m < r but I(u) (cid:54)≤ I(v). v ∈ ρGC∞ Proof. By contradiction, assume that u is a local minimizer, and let η ∈ ρGC∞ 0 (a, a(cid:48)) be a Jacobi field along u[a,a(cid:48)] such that the conditions from Def. 57 hold for η. We want to prove that η ≡ 0. Define ν := (η, 0[a(cid:48),b]), which is a piecewise GSF since η(a(cid:48)) = 0. Since also η(a) = 0, Lem. 58 and homogeneity of ψ yield Q(ν) = ψ(t, ν(t), ν(t)) dt = ψ(t, η(t), η(t)) dt + ψ(t, 0, 0) dt = 0. Therefore, Thm. 43 (necessary condition for u being a minimizer) gives Q(ν) ≥ 0 = Q(ν) for all ν ∈ ρGC∞ 0 (a, b). Thus, ν is a minimizer of the functional Q. Since ν is only a piecewise GSF, we cannot directly apply Thm. 38 (Euler-Lagrange equations). But for all ϕ ∈ ρGC∞ 0 (a, b) and all s ∈ ρ(cid:101)R, we have Q(ν + sϕ) = b ψ(t, ν + sϕ, ν + s ϕ) dt a(cid:48) a = ψ(t, η + sϕ, η + s ϕ) dt + a b a(cid:48) ψ(t, sϕ, s ϕ) dt. (6.17) (cid:12)(cid:12)(cid:12)(cid:12)0 26 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO This shows that s ∈ ρ(cid:101)R (cid:55)→ Q(ν + sϕ) ∈ ρ(cid:101)R is a GSF, and hence s = 0 is a minimum for this function. By Lem. 31 and (6.17), we get δQ(ν, ϕ) = 0 = a(cid:48) a(cid:48) a a = = (cid:18) (cid:18) Q(ν + sϕ) d ds ψη(t, η, η) − d dt ψη(t, η, η) − d dt (cid:19) (cid:19) b a(cid:48) ϕ dt + ψ η(t, η, η) ψ η(t, η, η) ϕ dt. (ϕψη(t, 0, 0) + ϕψ η(t, 0, 0)) dt By the fundamental Lem. 37, this implies that η satisfies the Euler-Lagrange equa- tions for ψ in the interval [a, a(cid:48)). Therefore, ν satisfies the same equations in [a, a(cid:48)) ∪ (a(cid:48), b]. Moreover, ψ η η(a(cid:48), η(a(cid:48)), η(a(cid:48))) = F u u(a(cid:48), u(a(cid:48)), u(a(cid:48))) is invertible by assumption (ii). Thus, all the hypotheses of the regularity Lem. 56 hold, and we derive that η(a(cid:48)) = 0. (cid:21) Fu u(t, u, u) − Fuu(t, u, u) , and (cid:20) d For all t ∈ [a, b], we define ξ(t) := −F −1 u u · dt u u · d ϑ(t) := −F −1 dt (cid:32) (cid:33) ´ order ODE  η z y(a(cid:48)) = = η(a(cid:48)) η(a(cid:48)) (cid:33) (cid:32) (cid:32) y := F u u(t, u, u), (cid:33) (cid:32) (cid:33) 0 1 ξ ϑ · η z = 0 so that we can re-write the Jacobi equations (6.4) for η on [a, a(cid:48)] as a system of first =: A · y ∀t ∈ [a, a(cid:48)] By assumptions (iii), we obtain hence apply Thm. 55 obtaining y ≡ 0 and thus η ≡ 0. a(cid:48) A(t) = OR(log dρ) for all t ∈ [a, a(cid:48)], and we can (cid:3) t Note that if one of the quantities in ((iii)) depends even only polynomially on ε, then we are forced to take e.g. ρε = ε1/ε to fulfill this assumption. This underlines the importance of the parameter ρ making the entire theory dependent on the pa- rameter ρ., in order to avoid unnecessary constraints on the scope of the functionals we look upon. 7. Noether's theorem We first note that any X ∈ ρGC∞ (J × X, Y ), where J ⊆ ρ(cid:101)R, can also be considered In this section, we state and prove Noether's theorem following the lines of [3]. as a family in GSF which smoothly depends on the parameter s ∈ J. In this case, we hence say that (Xs)s∈J is a generalized smooth family in ρGC∞(X, Y ). In particular, we can reformulate in the language of GSF the classical definition of one-parameter group of generalized diffeomorphisms of X as follows: (i) (ii) (iii) X0(x) = x for all x ∈ X, (Xs)s∈ρ(cid:101)R is a generalized smooth family in ρGC∞(X, X), For all s ∈ ρ(cid:101)R, the map Xs : X −→ X is invertible, and X−1 s ∈ ρGC∞(X, X), (iv) Xs ◦ Xt = Xs+t for all s, t ∈ ρ(cid:101)R. In our proofs, we will in fact only use properties (i) and (iii). The proof of Noether's theorem is classically anticipated by the following time- independent version, which the general case is subsequently reduced to. CALCULUS OF VARIATIONS FOR GF Theorem 60. Let K ∈ ρGC∞(cid:16) L × V, ρ(cid:101)R(cid:17) , where L, V ⊆ ρ(cid:101)Rn are sharply open 27 sets. Let w ∈ ρGC∞((a, b), L) be a solution of the Euler-Lagrange equation corre- sponding to K, i.e. for all t ∈ (a, b) w(t) ∈ V , Ku(w(t), w(t)) = Suppose that 0 is a sharply interior point of J ⊆ ρ(cid:101)R and (Xs)s∈J is a generalized K u(w(t), w(t)). d dt (7.1) smooth family in ρGC∞(L, L), such that for all t ∈ (a, b) ∂t Xs(w(t)) ∈ V , (i) (ii) X0(w(t)) = w(t), (iii) K is invariant under (Xs)s∈J along w, i.e. ∂ (cid:18) K(w(t), w(t)) = K Then, the quantity Xs(w(t)), ∂ ∂t Xs(w(t)) (cid:12)(cid:12)(cid:12)(cid:12)s=0 ∂ ∂s s (w(t)) ∈ ρ(cid:101)R X j K uj (w(t), w(t)) (cid:19) ∀s ∈ J. (7.2) (cid:12)(cid:12)(cid:12)(cid:12)s=0 (cid:12)(cid:12)(cid:12)(cid:12)s=0 ∂ ∂s (cid:12)(cid:12)(cid:12)(cid:12)s=0 (cid:12)(cid:12)(cid:12)(cid:12)s=0 (cid:12)(cid:12)(cid:12)(cid:12)s=0 ∂ ∂s ∂ ∂s is constant in t ∈ (a, b). be arbitrary but fixed. Since s = 0 ∈ J is a sharply interior point, we can consider Proof. We first note that both sides of (7.2) are in ρGC∞((a, b), ρ(cid:101)R). Let τ ∈ (a, b) (cid:12)(cid:12)s=0. We obtain (cid:12)(cid:12)(cid:12)(cid:12)s=0 Xs(w), Xs(w) (cid:18) (cid:19) (7.2) = ∂ ∂t d ds K 0 ∂ ∂s τ (ii) = Ku(w, w) a ∂ ∂s Xs(w) + K u(w, w) Xs(w) dt. ∂ ∂t ∂ ∂s Since the Euler-Lagrange equations (7.1) for K are given by Ku(w, w) = d we have dt K u(w, w), 0 = = d dt d dt (K u(w, w)) (cid:18) K u(w, w) Xs(w) + K u(w, w) (cid:19) ∂ ∂t Xs(w) . Xs(w) (cid:3) Which is our conclusion by the uniqueness - part of Thm. 13. Theorem 61 (E. Noether). We are now able to prove Noether's theorem. For the convenience of the reader, in its statement and proof we use the variables t, T , l, L, v, V so as to recall tempus, locus, velocitas resp. Let a, b ∈ ρ(cid:101)Rd, with a < b, and F ∈ ρGC∞(cid:16) [a, b] × ρ(cid:101)Rd × ρ(cid:101)Rd, ρ(cid:101)R(cid:17) ρGC∞([a, b], ρ(cid:101)Rd) be a solution of the Euler-Lagrange equation (4.5) corresponding to F . Suppose that 0 is a sharply interior point of J ⊆ ρ(cid:101)R and (Xs)s∈J is a generalized smooth family in ρGC∞((a, b)×ρ(cid:101)Rd, (a, b)×ρ(cid:101)Rd). We denote by Ts(t, l) := X 1 s (t, l) ∈ ρ(cid:101)Rd for all (t, l) ∈ (a, b) × ρ(cid:101)Rd, the two projections s (t, l) ∈ of Xs on (a, b) and ρ(cid:101)Rd resp. We assume that for all t ∈ (a, b) (cid:104) ∂t Ts(t, u(t)) ∈ ρ(cid:101)R is invertible, T0(t, u(t)) = t and L0(t, u(t)) = u(t), Ts(t, u), Ls(t, u), ∂t Ts(t, u) for all s ∈ J. (a, b) and Ls(t, l) := X 2 (iii) F (t, u(t), u(t)) = F (cid:105) · ∂ . Let u ∈ (i) (ii) ∂ ∂ ∂t Ls(t,u) ∂ ∂t Ts(t,u) 28 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO Then, the quantity F uj (t, u(t), u(t)) (cid:12)(cid:12)(cid:12)(cid:12)s=0 ∂ ∂s Lj s(t, u(t))+ +(cid:2)F (t, u(t), u(t)) − F uk (t, u(t), u(t)) uk(t)(cid:3) ∂ (cid:12)(cid:12)(cid:12)(cid:12)s=0 ∂s Ts(t, u(t)) (7.3) is constant in t ∈ [a, b]. Proof. Since (7.3) is a GSF in t ∈ [a, b], by sharp continuity it suffices to prove the claim for all t ∈ (a, b). Set L := (a, b) × ρ(cid:101)Rd, V := ρ(cid:101)R∗ × ρ(cid:101)Rd (we recall that ρ(cid:101)R∗ denotes the set of all invertible generalized numbers in ρ(cid:101)R). Define K ∈ ρGC∞(L × V, ρ(cid:101)R) by · p ∀(t, l) ∈ L∀(p, v) ∈ V, K(t, l; p, v) := F (7.4) and w ∈ ρGC∞((a, b), L) by w(t) := (t, u(t)) for all t ∈ (a, b). We note that L, V ⊆ ρ(cid:101)Rd+1 are sharply open subsets and that w(t) = (1, u(t)) ∈ V . The notations t, l, for partial derivatives used in the present work result from the symbolic writing K(u1, . . . , ud+1; u1, . . . , ud+1), so that the variables used in (7.4) yield (cid:18) (cid:19) v p (cid:16) Kt(t, l; p, v) = Ft Kp(t, l; p, v) = F (cid:16) Kvj (t, l; p, v) = F uj−1 Klj (t, l; p, v) = Fuj−1 t, l, v p t, l, v p (cid:17) · p (cid:17) · p (cid:16) (cid:16) (cid:17) − F uk (cid:16) (cid:17) (cid:20) d t, l, v p t, l, v p (cid:17) vk p t, l, v p if j = 1 if j = 2, . . . , d + 1, (7.5) if j = 1 if j = 2, . . . , d + 1. (7.6) (cid:21) · uk Kuj (t, l; p, v) = and K uj (t, l; p, v) = From these, for all t ∈ (a, b) and all j = 2, . . . , d + 1, it follows that F uk (t, u, u) − Fuk (t, u, u) Ku1 (w, w) − d dt Kuj (w, w) − d dt K u1(w, w) = K uj (w, w) = Fuj−1(t, u, u) − d dt dt F uj−1 (t, u, u). Therefore, since u satisfies the Euler-Lagrange equations for F , this entails that w is a solution of the analogous equations for K in (a, b). Now, (i) gives (cid:18) ∂ (cid:19) ∂ ∂t Xs(w(t)) = Ts(t, u(t)), ∂ ∂t Ls(t, u(t)) ∂t (cid:18) K(w, w) = F (t, u, u) (cid:34) K Xs(w), Xs(w) = F Ts(t, u), Ls(t, u), (cid:19) ∈ ρ(cid:101)R∗ × ρ(cid:101)Rd = V. (cid:35) ∂ ∂t Ls(t, u) ∂ ∂t Ts(t, u) · ∂ ∂t Ts(t, u). Moreover, (ii) gives X0(w(t)) = (T0(t, u(t)), L0(t, u(t))) = w(t). Finally We can hence apply Thm. 60, and from (7.5), (7.6) we get that K uj (w, w) ∂ ∂s X j s (w) = F uj (t, u(t), u(t)) +(cid:2)F (t, u(t), u(t)) − F uk (t, u(t), u(t)) uk(t)(cid:3) ∂ s(t, u(t))+ Lj (cid:12)(cid:12)(cid:12)(cid:12)s=0 ∂ ∂s is constant in t ∈ (a, b). (cid:12)(cid:12)(cid:12)(cid:12)s=0 Ts(t, u(t)) (cid:3) ∂s ∂ ∂t (cid:12)(cid:12)(cid:12)(cid:12)0 CALCULUS OF VARIATIONS FOR GF 29 8. Application to C1,1 Riemannian metric In this section, we fix any embedding (ιb In the following, we apply what we did so far to the problem of length-minimizers in (Rd, g), where g ∈ C1,1 is a Riemannian metric. Furthermore, we assume that (Rd, g) is geodesically complete. Note that the seeming restriction of considering only Rd as our manifold weighs not so heavy. Indeed, the question of length min- imizers can be considered to be a local one, since it is not guaranteed that global minimizers exist at all, whereas local minimizers always exist. Additionally, note that it was shown that it suffices to consider smooth manifolds (cf. [18, Thm. 2.9]) instead of Ck manifolds with 1 ≤ k < +∞. Therefore, there is no need to consider non-smooth charts. some a ∈ R>0, and where Ω ⊆ Rd is an arbitrary open set, see Thm. 18. Actually, the embedding also depends on the dimension d ∈ N>0, but to avoid cumbersome notations, we denote embeddings always with the symbol ι. Ω)Ω, where b ∈ ρ(cid:101)R satisfies b ≥ dρ−a for ij be the Christoffel symbols of gε, and set Γij := [Γε ij(−)] ∈ ρGC∞(ρ(cid:101)Rd × ρ(cid:101)Rd, ρ(cid:101)R), then for all ε, By [24, Rem. 2.6.2], it follows that we can always find a net of smooth functions (gε ij) such that setting g := ι(g) =: [gε ij(−)] ∈ ρGC∞(ρ(cid:101)Rd, ρ(cid:101)Rd). ij → gij in C0 norm. ij is a Riemannian metric. By Thm. 18.(iii) it follows that gε gε A curve γ ∈ ρGC∞(J, ρ(cid:101)Rd), J being a sharply open subset of ρ(cid:101)R, is said to be a Let Γε geodesic of (ρ(cid:101)Rd, g) if Definition 62. We say that (ρ(cid:101)Rd, g) is geodesically complete if every solution of the geodesic equation belongs to ρGC∞(ρ(cid:101)R, ρ(cid:101)Rd), i.e. if for all p ∈ ρ(cid:101)Rd and all v ∈ ρ(cid:101)Rd there exists a geodesic γ ∈ ρGC∞(ρ(cid:101)R, ρ(cid:101)Rd) of (ρ(cid:101)Rd, g) such that γ(0) = p any geodecis γ ∈ ρGC∞(ρ(cid:101)R, ρ(cid:101)Rd) induces a Colombeau generalized function γc(R) ∈ This definition includes also the possibility that the point p or the vector v could be infinite. By Thm. 19, it follows that if we consider only finite p and v, then Gs(R)d. Therefore, the space (c(Rd), gc(Rd)×c(Rd)) is geodesically complete in the sense of [36]. We recall that c(Ω) is the set of compactly supported (i.e. finite) generalized points in Ω (see Thm. 18). γ(t) + Γij(γ(t)) γi(t) γj(t) = 0 ∀t ∈ J. and γ(0) = v. (8.1) The definition of length of a (non singular) curve needs the following √ Lem. 3 readily implies that xε] ∈ ρ(cid:101)R>0. Remark 63. We set√− = (−)1/2 : x = [xε] ∈ ρ(cid:101)R>0 (cid:55)→ [ √− ∈ ρGC∞(ρ(cid:101)R>0, ρ(cid:101)R>0). Therefore, the square root is defined on every strictly positive infinitesimal, but it cannot be extended to ρ(cid:101)R≥0. (cid:111) (cid:110) λ ∈ ρGC∞([0, 1], ρ(cid:101)Rd) λ(0) = p, λ(1) = q, λ(t) > 0 ∀t ∈ [0, 1] Let p, q ∈ ρ(cid:101)Rd, then (i) ρGC∞ Definition 64. . >0(p, q) := Moreover, for λ ∈ ρGC∞ >0(p, q), we set (cid:0)gij(α(t)) αi(t) αj(t)(cid:1)1/2 1 dt ∈ ρ(cid:101)R. Lg(λ) := Let x = [xε] ∈ ρ(cid:101)Rn, then we set st(x) := limε→0 xε ∈ Rd, if this limit exists. 0 (ii) Note that x ≈ st(x) in this case. 30 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO We are interested only in global minimizers of the functional Lg, i.e. curves Note that (8.1) are the usual geodesic equations for the generalized metric g, whose derivation is completely analogous to that in the smooth case. Thus they are the Euler-Lagrange equations of Lg. λ0 ∈ X(p, q) such that Lg(λ0) ≤ Lg(λ) for all λ ∈ ρGC∞ Lemma 65. Let p, q ∈ Rd and p, q ∈ ρ(cid:101)Rd such that st(p) = p and st(q) = q. Let >0(p, q) :=(cid:8)w ∈ C1([0, 1]R, Rd) w(0) = p, w(1) = q, w(t) > 0 ∀t ∈ [0, 1]R(cid:9) λ = [λε(−)] ∈ ρGC∞ ¯λ ∈ C1 such that λε → ¯λ in C1 as ε → 0, then >0(p, q) be such that there exists >0(p, q). Proof. We calculate: (cid:17)1/2 −(cid:16) ij(λε) λi gε ε λj ε (cid:16) (cid:12)(cid:12)(cid:12)(cid:12) 1 0 By assumption, (gε such that ε ij(λε) λi (cid:17)1/2 −(cid:16) λj ε (cid:16) (cid:12)(cid:12)(cid:12)(cid:12) 1 0 ij(λε) λi gε ε (cid:12)(cid:12)(cid:12)(gε 1 0 ≤ C st(Lg(λ)) = Lg(¯λ). (cid:12)(cid:12)(cid:12)(cid:12) = dt gij(¯λ) ¯λi ¯λj(cid:17)1/2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) gij(¯λ) ¯λi ¯λj(cid:17)1/2 = 1 0 (cid:12)(cid:12)(cid:12)(cid:12) dt ε − gij(¯λ) ¯λi ¯λj λj ij(λε) λi gε ε)1/2 + (gij(¯λ) ¯λi ¯λj)1/2 λj ε (gε ij(λε) λi ε ε)1/2 → (gij(¯λ) ¯λi ¯λj)1/2, so that there exists C ∈ R>0 λj (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) . dt ij(λε) − gij(λε) + gij(λε) − gij(¯λ)) λi ε λj ε + gij(¯λ)( λi ε ε − ¯λi ¯λj) λj (cid:12)(cid:12)(cid:12) dt. We hence obtain the claim by the triangle inequality and by convergence of λε, λε (cid:3) and gε ij to ¯λ, ¯λ and gij respectively. Now, we consider p, q ∈ Rd with p (cid:54)= q. Let u ∈(cid:8)u ∈ C2,1([0, 1], Rd) u(0) = p, u(1) = q(cid:9) Let c0 := u(0). Obviously, u is also the unique solution of p = u(0) q = u(1). be a solution of the geodesic equationu = −Γij(u) ui uj u = −Γij(u) ui uj y = −Γε p = u(0) c0 = u(0). p = y(0) c0 = y(0). ij(y) yi yj Using these initial conditions, for each fixed ε we can solve the following problem for a unique yε ∈ C∞([−dε, dε]R, Rd) and some dε ∈ R>0. Lemma 66. Let u and yε be as above. Then (i) For ε sufficiently small, the solution yε can be extended to a solution yε ∈ C∞([0, 1]R, Rd) of (8.4) such that yε(1) = q. (8.2) (8.3) (8.4) CALCULUS OF VARIATIONS FOR GF 31 yε → u in C2. (ii) (iii) The net (yε) defines a GSF, i.e. y := [yε(−)] ∈ ρGC∞ Proof. ij → Γij locally uniformly. Thus, we Claim (i), (ii): For all i, j , we have that Γε obtain these claims by (8.2) and by continuous dependence on parameters in ODE, see e.g. [25, Lemma 2.3]. >0(p, q). We have to show that for all ε all derivatives of yε are moderate. This is obviously true for yε, yε and yε. The claim follows now from the fact that Claim (iii) I: y := [yε(−)] ∈ ρGC∞([0, 1], ρ(cid:101)Rd) (cid:0)Γεk dn+2 (cid:0)Γεk dn dtn ij (y) yi ε yj ε dtn+2 yε = − dn (cid:18) (cid:1) = P yε, . . . , dtn yε, d dt so that there exists a polynomial P such that (cid:1) ij (y) yi ε yj ε dn+1 dtn+1 yε, Γεk ij , DΓεk ij , . . . , DnΓεk ij (cid:19) . Claim (iii) II: y(t) > 0 for all t ∈ [0, 1] By (ii), we have that yε → u in C2. Furthermore, gε → g in C1 by assumption, and we know that g( u, u) = c > 0 for some c ∈ R>0 since u is a g-geodesic (cf. [20, Lemma 1.4.5]). Therefore, we obtain that gε( yε, yε) > c/2 > 0 for ε > 0 small (cid:3) enough. Finally, the standard part of the generalized length of y is the length of u: Theorem 67. Let u and yε be as above. We conclude (using Lemma 65) that st(Lg(y)) = Lg(u). Proposition 68. Let y = [yε(−)] be as above. In addition, assume that each yε is Lgε-minimizing. Then Lg(y) is minimal. Proof. Let λ = [λε(−)] ∈ ρGC∞ Lg(y) = [Lgε(yε)]. By assumption, for all ε we have Lgε(λε) ≥ Lgε(yε). >0(p, q). We have that Lg(λ) = [Lgε(λε)] and that Therefore, Lg(λ) ≥ Lg(y), as claimed. Corollary 69. Let λ ∈ ρGC∞ small, yε is Lgε-minimizing. Then Lg(y) = Lg(λ). >0(p, q) be a minimizer of Lg and assume that for ε (cid:3) This Corollary 69 gives us a way to answer the question if a certain classical geodesic between two given classical points p and q is a length-minimizer. Furthermore, we are able to prove the following theorem, relating GSF-minimizers to classical minimizers. Theorem 70. Let p, q ∈ Rd and let γ ∈ ρGC∞ Assume that st(Lg(γ)) exists and that there exists w ∈ C1 st(Lg(γ)). >0(p, q) such that Lg(γ) is minimal. >0(p, q) such that Lg(w) = Then w is g - minimizing and a g - geodesic. Proof. Assume to the contrary that there exists a curve σ ∈ C 2 connecting p and q (w. l. o. g. σ is a g - geodesic) such that Lg(σ) < Lg(w). Now we construct (as done above) gε, σε and set σ := [σε]. Then: st (Lg(σ)) = Lg(σ) < Lg(w) = st (Lg(γ)) . 32 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO But, by assumption we have that Lg(γ) ≤ Lg(σ), which implies st (Lg(γ)) ≤ st (Lg(σ)) < st (Lg(γ)) . This is a contradiction. (cid:3) 9. Conclusions We can summarize the present work as follows (i) The setting of GSF allows to treat Schwartz distributions more closely to classical smooth functions. The framework is so flexible and the extensions of classical results are so natural in many ways one may treat it like smooth functions. (ii) One key step of the theory is the change of the ring of scalars into a non- Archimedean one and the use of the strict order relation < to deal with topological properties. So, the use of < and of ρ(cid:101)R-valued norms allows a nat- ural approach to topology, even of infinite dimensional spaces (cf. Def. 41). On the other hand, the use of a ring with zero divisors and a non-total order relation requires a more refined and careful analysis. However, as proved in the present work, very frequently classical proofs can be formally repeated in this context, but paying particular attention to using the relation <, us- ing invertibility instead of being non zero in R and avoiding the total order property. (iii) Others crucial properties are the closure of GSF with respect to composition and the use of the gauge ρ, because they do not force to narrow the theory into particular cases. (iv) The present extension of the classical theory of calculus of variations shows that GSF are a powerful analytical technique. The final application shows how to use them as a method to address problems in an Archimedean setting based on the real field R. Concerning possible future developments, we can note that: (v) A generalization of the whole construction to piecewise GSF seems possible. (vi) A more elegant approach to integration of piecewise GSF could use the exis- tence of right and left limits of (f1, . . . , fn)(−) and hyperfinite Riemann-like sums, i.e. sums ∈ ρ(cid:101)Rd , where int(−) is the integer N(cid:88) f (x(cid:48) (cid:34) Nε(cid:88) i)(xi − xi−1) := (cid:110) [int(xε)] [xε] ∈ ρ(cid:101)R(cid:111) (cid:35) i,ε)(xi,ε − xi−1,ε) fε(x(cid:48) i=1 extended to N ∈ (cid:101)N := i=1 part function. The present work could lay the foundations for further works concerning the possi- bility to extend other results of the calculus of variations in this generalized setting. References [1] Aragona, J., Fernandez, R., Juriaans, S.O., A discontinuous Colombeau differential calculus, Monatsh. Math. 144, 13–29 (2005). [2] Aragona, J., Juriaans, S.O., Some structural properties of the topological ring of Colombeau's generalized numbers, Comm. Algebra 29 (2001). [3] Avez, A., Differential Calculus, John Wiley & Sons Inc., 1986. [4] Colombeau, J.F., Multiplication of distributions - A tool in mathematics, numerical engi- neering and theoretical Physics. Springer-Verlag, Berlin Heidelberg (1992). [5] Davie, A. M., Singular minimisers in the calculus of variations in one dimension. Arch. Ra- tional Mech. Anal. 101(2), 161–177, 1988. [6] Dirac, P.A.M., The physical interpretation of the quantum dynamics, Proc. R. Soc. Lond. A, 113 , 1926–27, 621–641. CALCULUS OF VARIATIONS FOR GF 33 [7] Engquist, B., Tornberg, A.K., Tsai, R., Discretization of Dirac delta functions in level set methods. Journal of Computational Physics, 207:28–51, 2005. [8] Gelfand, I.M., Fomin, S.V., Calculus of variations, Dover Publications, 2000. [9] Giordano, P., Kunzinger, M., Inverse Function Theorems for Generalized Smooth Functions. Invited paper for the Special issue ISAAC - Dedicated to Prof. Stevan Pilipovic for his 65 birthday. Eds. M. Oberguggenberger, J. Toft, J. Vindas and P. Wahlberg, Springer series "Operator Theory: Advances and Applications", Birkhaeuser Basel, 2016. [10] Giordano, P., Kunzinger, M., A convenient notion of compact sets for generalized func- tions. Accepted in Proceedings of the Edinburgh Mathematical Society, 2016. See arXiv 1411.7292v1. [11] Giordano, P., Kunzinger, M., 'New topologies on Colombeau generalized numbers and the Fermat-Reyes theorem'. Journal of Mathematical Analysis and Applications 399 (2013) 229–238. [12] Giordano P., Kunzinger M., Steinbauer R., A new approach to generalized functions for mathematical physics. See http://www.mat.univie.ac.at/giordap7/GenFunMaps.pdf. [13] Giordano, P., Kunzinger, M., Vernaeve, H., Strongly internal sets and generalized smooth functions. Journal of Mathematical Analysis and Applications, volume 422, issue 1, 2015, pp. 56–71. [14] Giordano, P., Luperi Baglini, L., Asymptotic gauges: Generalization of Colombeau type algebras. Math. Nachr. 289, 2-3, 1–28, (2015). [15] Graves, L.M., Discontinuous solutions in the calculus of variations. Bull. Amer. Math. Soc. 36, 831–846, 1930. [16] Grosser, M., Kunzinger, M., Oberguggenberger, M., Steinbauer, R., Geometric theory of generalized functions, Kluwer, Dordrecht (2001). [17] Hawking, S.W., Ellis, G., The large scale structure of space-time, Cambridge University Press (1976) [18] Hirsch, M.W., Differential Topology, Springer (1976). [19] Hosseini, B. , Nigam, N., Stockie, J.M., On regularizations of the Dirac delta distribution, Journal of Computational Physics, Volume 305, 2016, Pages 423–447. [20] Jost, J., Riemannian Geometry and Geometric Analysis, Springer, 2011. [21] Jost, J., Li-Jost, X., Calculus of variations, Cambridge Studies in Advanced Mathematics 64, 1998. [22] 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. [23] 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) [24] Kunzinger, M., Steinbauer, R., Stojkovi´c, M., Vickers, J.A., A regularisation approach to causality theory for C1,1-Lorentzian metrics, Gen. Relativ. Gravit. 46 (2014). [25] Kunzinger, M., Steinbauer, R., Stojkovi´c, M., The exponential map of a C1,1-metric, Diff. Geom. Appl. 34, 14 –24 (2014). [26] Laugwitz, D., Definite values of infinite sums: aspects of the foundations of infinitesimal analysis around 1820. Arch. Hist. Exact Sci. 39 (1989), no. 3, 195–245. [27] Lecke, A., Steinbauer, R., Svarc, R., The regularity of geodesics in impulsive pp-waves, Gen. Relativ. Gravit. 46 (2014). [28] Lecke, A., Non-smooth Lorentzian Geometry and Causality Theory, PhD Thesis, Universitat Wien (2016). [29] Luperi Baglini, L., Giordano, P., Fixed point iteration methods for arbitrary generalized ODE, preprint. [30] Luperi Baglini, L., Giordan, P., The category of Colombeau algebras. In revision for Monat- shefte fur Mathematik. See arXiv 1507.02413. [31] Lytchak, A., Yaman, A., On Hoelder continuous Riemannian and Finsler metrics, Trans. Amer. Math. Soc. 358 (2006). [32] Minguzzi, E., Convex neighborhoods for Lipschitz connections and sprays, Monatshefte fur Mathemati, Volume 177, Issue 4, pp 569-625 (2015) [33] Oberguggenberger, M., Vernaeve, H., Internal sets and internal functions in Colombeau the- ory, J. Math. Anal. Appl. 341 (2008) 649–659. [34] Robinson, A., Function theory on some nonarchimedean fields, Amer. Math. Monthly 80 (6) 87–109; Part II: Papers in the Foundations of Mathematics (1973). [35] Samann, C., Steinbauer, R., Lecke, A., Podolsk´y, J.,Geodesics in nonexpanding impulsive gravitational waves with Λ, part I, Classical and Quantum Gravity 33 (2016). [36] Samann,C. , Steinbauer, R., Geodesic Completeness of Generalized Space-times. In Eds S. Pilipovi´c, J., Pseudo-Differential Operators and Generalized Functions, Volume 245 of the series Operator Theory: Advances and Applications pp 243-253, 2015. 34 ALEXANDER LECKE, LORENZO LUPERI BAGLINI, AND PAOLO GIORDANO [37] Stojkovi´c, M., Causality theory for C1,1 - metrics, PhD Thesis, Universitat Wien (2015). [38] Tornberg, A.K., Engquist, B., Numerical approximations of singular source terms in differ- ential equations, Journal of Computational Physics 200 (2004) 462–488. [39] Tuckey, C., Nonstandard methods in the calculus of variations, Pitman Research Notes in Mathematics Series 297. Longman Scientific & Technical, Harlow, 1993. University of Vienna, Austria E-mail address: [email protected] University of Vienna, Austria E-mail address: [email protected] E-mail address: [email protected]
1806.10752
4
1806
2019-01-28T10:47:49
Paley-Wiener properties for spaces of entire functions
[ "math.FA" ]
We deduce Paley-Wiener results in the Bargmann setting, which give characterisations of Pilipovi{\'c} spaces of low orders, extending the characterisation of a Gr{\"o}chenig test function space, deduced earlier by the third author.
math.FA
math
PALEY-WIENER PROPERTIES FOR SPACES OF ENTIRE FUNCTIONS ELMIRA NABIZADEH, CHRISTINE PFEUFFER, AND JOACHIM TOFT Abstract. We deduce Paley-Wiener results in the Bargmann set- ting. At the same time we deduce characterisations of Pilipović spaces of low orders. In particular we improve the characterisation of the Gröchenig test function space H♭1 = SC , deduced in [12]. 0. Introduction Paley-Wiener theorems characterize functions and distributions with certain restricted supports in terms of estimates of their Fourier-Laplace transforms. For example, let f be a distribution on Rd and let Br0(0) ⊆ Rd be the ball with center at origin and radius r0. Then f is supported in Br0(0) if and only if bf (ζ) . hζiN er0 Im(ζ), bf (ζ) . hζi−N er0 Im(ζ), ζ ∈ Cd, ζ ∈ Cd, for some N ≥ 0. Furthermore, f is supported in Br0(0) and smooth, if and only if for every N ≥ 0 (see e. g. [6, Section 7]). A similar approach for ultra-regular functions of Gevrey types and corresponding ultra-distribution spaces can be done. In fact, let s > 1, D′ s(Rd) be the set of all Gevrey distributions of order s and let Es(Rd) be the set of all smooth functions with Gevrey regularity s. (See [6] s(Rd) and Section 1 for notations.) Then it can be proved that f ∈ D′ is supported in Br0(0), if and only if bf (ζ) . er0 Im(ζ)+rζ bf (ζ) . er0 Im(ζ)−rζ 1 s , ζ ∈ Cd, 1 s , ζ ∈ Cd, for every r > 0. Furthermore f ∈ Es(Rd) is supported in Br0(0), if and only if for some r > 0. We observe that s in the latter result can not be pushed to be smaller, because if s ≤ 1, it does not make any sense to discuss compact support properties of D′ s(Rd) and Es(Rd). 2010 Mathematics Subject Classification. primary 46F05; 32A25; 32A36; sec- ondary 35Q40; 30Gxx. Key words and phrases. Bargmann transform. 1 In the paper we consider analogous Paley-Wiener properties when the Fourier-Laplace transform above is replaced by the reproducing ker- nel ΠA of the Bargmann transform, and the image spaces are replaced by suitable subspaces of entire functions on Cd. These subspaces were considered in [4, 12] and are given by A♭σ (Cd) =[r>0 As(Cd) =[r>0 Ar,♭σ(Cd), Ar,s(Cd), A0,♭σ (Cd) =\r>0 A0,s(Cd) =\r>0 Ar,♭σ(Cd), Ar,s(Cd), (0.1) when σ > 0 and 0 < s < 1 of all entire F on Cd such that 2, where Ar,♭σ (Cd) and Ar,s(Cd) are the sets F (z) . erz 2σ σ+1 respective F (z) . er(loghzi) 1 1−2s . The spaces in (0.1) appear naturally when considering the Bargmann transform images of an extended class of Fourier invariant Gelfand- Shilov spaces, called Pilipović spaces (see [4, 12]). If (z, w) is the scalar product of z, w ∈ Cd, then the reproducing kernel of the Bargmann transform is given by (ΠAF )(z) = π−dhF, e(z, · )− · 2 i, when F is a suitable function or (ultra-)distribution. If z 7→ F (z)eRz−z2 ∈ L1(Cd), R > 0 (0.2) holds and dλ(w) is the Lebesgue measure on Cd, then (ΠAF )(z) = π−dZCd F (w)e(z,w)−w2 dλ(w). A recent Paley-Wiener result with respect to the transform ΠA and to the image spaces (0.1) is given in [12], where it is proved that if c (Cd) = E ′(Cd) ∩ L∞(Cd), then L∞ ΠA(E ′(Cd)) = ΠA(L∞ (0.3) Evidently, L∞ c ⊆ E ′, and the gap between these spaces is rather large. It might therefore be somehow surprising that the first equality holds in (0.3). c (Cd)) = A♭1(Cd). In Section 2 we improve (0.3) in different ways. Firstly we show that A,c(Cd) given c (Cd) in (0.3) with the smaller space L∞ we may replace L∞ by where L∞ A,c(K) is the set of all F · χK, where F is analytic in a neigh- bourhood of K and χK is the characteristic function of K. Secondly, L∞ A,c(Cd) =[ L∞ A,c(K), 2 s(Cd) of all compactly supported we may replace E ′(Cd) with the set E ′ Gevrey distributions of order s > 1. Summing up we improve (0.3) into (0.3)′ s(Cd)) = ΠA(L∞ s > 1. A,c(Cd)) = A♭1(Cd), ΠA(E ′ In Section 2 we also deduce various kind of related mapping prop- erties when A♭1(Cd) in (0.3) is replaced by any of the spaces in (0.1). More precisely, let χ ∈ L∞ c (Cd) be non-negative, radial symmetric in each complex variable zj and bounded from below by a positive con- stant near the origin. Then we prove σ σ 0,♭σ0 2σ − 1 ΠA(A′ 2 , 1), σ0 = ΠA(A♭σ0 ΠA(As(Cd) · χ) = As(Cd), (Cd) · χ) = A♭σ (Cd), σ ∈ ( 1 (Cd) · χ) = A♭σ (Cd), σ ∈ (0, 1 s ∈ [0, 1 2). Some related properties are deduced for σ = 1 2, as well as when A♭σ and As are replaced by A0,♭σ and A0,s, respectively. (Cf. Theorems 2.2 -- 2.8 and Propositions 2.9 -- 2.11.) Finally, in Section 3 we use the results in Section 2 to deduce char- 1 − 2σ 2), σ0 = , , acterizations of Pilipović spaces of small orders. 1. Preliminaries In this section we recall some basic facts. We start by discussing Pilipović spaces and some of their properties. Then we recall some facts on modulation spaces. Finally we discuss the Bargmann transform and some of its mapping properties, and introduce suitable classes of entire functions on Cd. 1.1. The Pilipović spaces. The definition of Pilipović spaces is based on the Hermite functions, which are given by hα(x) = π− d 4 (−1)α(2αα!)− 1 2 e x2 2 (∂αe−x2 ), α ∈ Nd. The Hermite functions are eigenfunctions for the Fourier transform, and for the Harmonic oscillator Hd ≡ x2 − ∆ which acts on functions and (ultra-)distributions defined on Rd. More precisely, we have Hdhα = (2α + d)hα. It is well-known that the set of Hermite functions is a basis of S (Rd) and an orthonormal basis of L2(Rd) (cf. [10]). In particular, if f ∈ L2(Rd), then where L2(Rd) = Xα∈Nd ch(f, α)2, kfk2 f (x) = Xα∈Nd ch(f, α)hα 3 (1.1) + and s ∈ R♭ we set e−( 1 d ), rαα!− 1 2σ , 1 1 +···+ 1 2s rd 1 2s ·α ·α r1 rα, ·α 1 1 +···+ 1 2s rd ·α 1 2s d ), r1 e( 1 rαα! 1 2σ , s ∈ R+ \ { 1 2}, s = ♭σ, s = 1 2, α ∈ Nd s ∈ R+ \ { 1 2}, s = ♭σ, (1.3) (1.4) and ϑr,s(α) ≡ ϑ′ r,s(α) ≡   is the Hermite seriers expansion of f , and ch(f, α) = (f, hα)L2(Rd) is the Hermite coefficient of f of order α ∈ Rd. should belong to the extended set In order to define the full scale of Pilipović spaces, their order s (1.2) of positive real numbers, with extended inequality relations as R♭ = R+ ∪ { ♭σ ; σ ∈ R+ }, s1 < ♭σ < s2 and ♭σ1 < ♭σ2 when s1 < 1 2 ≤ s2 and σ1 < σ2. (Cf. [12].) For such r ∈ Rd rα, s = 1 2 , α ∈ Nd. Definition 1.1. Let s ∈ R♭ = R♭ ∪ {0}, and let ϑr,s and ϑ′ (1.3) and (1.4). (1) H0(Rd) consists of all Hermite polynomials, and H′ sists of all formal Hermite series expansions in (1.1); (2) if s ∈ R♭, then Hs(Rd) (H0,s(Rd)) consists of all f ∈ L2(Rd) 0(Rd) con- r,s be as in such that ch(f, hα) . ϑr,s(α) holds true for some r ∈ Rd s(Rd) (H′ (3) if s ∈ R♭, then H′ series expansions in (1.1) such that + (for every r ∈ Rd +); 0,s(Rd)) consists of all formal Hermite ch(f, hα) . ϑ′ r,s(α) holds true for every r ∈ Rd + (for some r ∈ Rd +). The spaces Hs(Rd) and H0,s(Rd) are called Pilipović spaces of Roumieu respectively Beurling types of order s, and H′ 0,s(Rd) are called Pilipović distribution spaces of Roumieu respectively Beurling types of order s. s(Rd) and H′ 4 Remark 1.2. Let Ss(Rd) and Σs(Rd) be the Fourier invariant Gelfand- Shilov spaces of order s ∈ R+ of Roumieu respective Beurling types (see [12] for notations). Then it is proved in [8, 9] that H0,s(Rd) = Σs(Rd) 6= {0}, H0,s(Rd) 6= Σs(Rd) = {0}, Hs(Rd) = Ss(Rd) 6= {0}, s > s ≤ s ≥ Hs(Rd) 6= Ss(Rd) = {0}, s < , , 1 2 1 2 1 2 1 2 . and In Proposition 1.3 below we give further characterisations of Pilipović spaces. Next we recall the topologies for Pilipović spaces. Let s ∈ R♭, r > 0, and let kfkHs;r and kfkH′ s ∈ R♭ s ∈ R♭, (1.5) (1.6) s;r be given by α∈Nd ch(f, α)ϑ′ r,s(α), kfkHs;r ≡ sup and kfkH′ s;r ≡ sup α∈Nd ch(f, α)ϑr,s(α), when f is a formal expansion in (1.1). Then Hs;r(Rd) consists of all expansions (1.1) such that kfkHs;r is finite, and H′ s;r(Rd) consists of all expansions (1.1) such that kfkH′ s;r is finite. It follows that both s;r(Rd) are Banach spaces under the norms f 7→ kfkHs;r Hs;r(Rd) and H′ and f 7→ kfkH′ s;r, respectively. We let the topologies of Hs(Rd) and H0,s(Rd) be the inductive re- spectively projective limit topology of Hs;r(Rd) with respect to r > 0. 0,s(Rd) are the pro- In the same way, the topologies of H′ s(Rd) and H′ jective respectively inductive limit topology of H′ s;r(Rd) with respect to r > 0. It follows that all the spaces in Definition 1.1 are complete, s(Rd) are Fréchet spaces with semi-norms and that H0,s(Rd) and H′ f 7→ kfkHs;r and f 7→ kfkH′ s;r, respectively. [12]. The proof is therefore omitted. 0(Rd). Then Proposition 1.3. Let s ∈ R+ ∪ {0} and let f ∈ H′ f ∈ H0,s(Rd) (f ∈ Hs(Rd)), if and only if f ∈ C ∞(Rd) and satis- fies kH N d fkL∞ . hN N!2s for every h > 0 (for some h > 0). The following characterisations of Pilipović spaces can be found in From now on we let φ(x) = π− d 4 e− x2 2 . 5 (1.7) 1.2. Spaces of entire functions and the Bargmann transform. Let Ω ⊆ Cd be open. Then A(Ω) is the set of all analytic functions in Ω. If instead Ω ⊆ Cd is closed, then A(Ω) is the set of all functions which are analytic in an open neighbourhood of Ω. In particular, if z0 ∈ Cd is fixed, then A({z0}) is the set of all complex-valued functions which are defined and analytic near z0. We shall now consider the Bargmann transform which is defined by the formula (Vdf )(z) = π− d when f ∈ L2(Rd) (cf. [1]). Here exp(cid:16) − 1 2 (hz, zi + y2) + 2 1 2hz, yi(cid:17)f (y) dy, hz, wi = zjwj, when z = (z1, . . . , zd) ∈ Cd 4ZRd dXj=1 and w = (w1, . . . , wd) ∈ Cd, and otherwise h · , · i denotes the duality between test function spaces and their corresponding duals. It is evident that Vdf is the entire func- tion on Cd, given by (Vdf )(z) =ZRd Ad(z, y)f (y) dy, or where the Bargmann kernel Ad is given by (Vdf )(z) = hf, Ad(z, · )i, (1.8) Ad(z, y) = π− d 4 exp(cid:16) − 1 2 (hz, zi + y2) + 2 1 2hz, yi(cid:17). (Rd) We note that the right-hand side in (1.8) makes sense when f ∈ S ′ and defines an element in A(Cd), since y 7→ Ad(z, y) can be interpreted (Rd) with values in A(Cd). Here and in what follows, as an element in S 1 A(Ω) denotes the set of analytic functions on the open set Ω ⊆ Cd. It was proved in [1] that f 7→ Vdf is a bijective and isometric map from L2(Rd) to the Hilbert space A2(Cd) ≡ B2(Cd) ∩ A(Cd), where B2(Cd) consists of all measurable functions F on Cd such that 1 2 2 kFkB2 ≡(cid:16)ZCd F (z)2dµ(z)(cid:17) 1 2 < ∞. (1.9) Here dµ(z) = π−de−z2 dλ(z), where dλ(z) is the Lebesgue measure on Cd. We recall that A2(Cd) and B2(Cd) are Hilbert spaces, where the scalar product is given by (F, G)B2 ≡ZCd F (z)G(z) dµ(z), F, G ∈ B2(Cd). (1.10) If F, G ∈ A2(Cd), then we set kFkA2 = kFkB2 and (F, G)A2 = (F, G)B2. 6 Furthermore, Bargmann showed that there is a convenient reproduc- ing formula on A2(Cd). More precisely, let F (w)e(z,w) dµ(w), (1.11) (ΠAF )(z) ≡ZCd dXj=1 when z 7→ F (z)eRz−z2 belongs to L1(Cd) for every R ≥ 0. Here (z, w) = zjwj, when z = (z1, . . . , zd) ∈ Cd and w = (w1, . . . , wd) ∈ Cd, is the scalar product of z ∈ Cd and w ∈ Cd. Then it is proved in [1, 2] that ΠA is the orthogonal projection of B2(Cd) onto A2(Cd). In particular, ΠAF = F when F ∈ A2(Cd). In [1] it is also proved that Vdhα = eα, where eα(z) ≡ zα √α! , z ∈ Cd. (1.12) In particular, the Bargmann transform maps the orthonormal basis {hα}α∈Nd of L2(Rd) bijectively into the orthonormal basis {eα}α∈Nd of monomials of A2(Cd). Hence, there is a natural way to identify formal Hermite series expansion by formal power series expansions c(F, α)eα(z), (1.13) F (z) = Xα∈Nd Xα∈Nd ch(f, α)eα(z). by letting the series (1.1) be mapped into (1.14) (1.15) It follows that if f, g ∈ L2(Rd) and F, G ∈ A2(Cd), then (f, g)L2(Rd) = Xα∈Nd (F, G)A2(Cd) = Xα∈Nd ch(f, α)ch(g, α), c(F, α)c(G, α). Here and in what follows, ( · , · )L2(Rd) and ( · , · )A2(Cd) denote the scalar products in L2(Rd) and A2(Cd), respectively. Furthermore, ch(f, α) = c(F, α) when F = Vdf. (1.16) We now recall the following spaces of power series expansions given in [12]. Definition 1.4. Let s ∈ R♭ = R♭ ∪ {0}, and let ϑr,s and ϑ′ (1.3) and (1.4). r,s be as in 7 (1) A0(Cd) consists of all analytic polynomials on Cd, and A′ 0(Cd) consists of all formal power series expansions on Cd in (1.13); (2) if s ∈ R♭, then As(Cd) (A0,s(Cd)) consists of all F ∈ L2(Cd) such that c(F, hα) . ϑr,s(α) holds true for some r > 0 (for every r > 0); 0,s(Cd)) consists of all formal power (3) if s ∈ R♭, then A′ s(Cd) (A′ series expansions in (1.13) such that c(F, hα) . ϑ′ r,s(α) holds true for every r > 0 (for some r > 0). Let f ∈ H′ 0(Rd) with formal Hermite series expansion (1.1). Then the Bargmann transform Vdf of f is defined to be the formal power series expansion (1.14). It follows that Vd agrees with the earlier definition when acting on L2(Rd), that Vd is linear and bijective from H′ 0(Rd) to A′ 0(Cd), and restricts to bijections from the spaces H0,s(Rd), Hs(Rd), H′ s(Rd) and H′ 0,s(Rd) to A0,s(Cd), As(Cd), A′ s(Cd) and A′ 0,s(Cd) (1.17) (1.18) If s ∈ R♭, f ∈ Hs(Rd), g ∈ H′ respectively, when s ∈ R♭. We also let the topologies of the spaces in (1.18) be inherited from the spaces in (1.17). s(Cd), then (f, g)L2(Rd) and (F, G)A2(Cd) are defined by the formula (1.15). It follows that (1.16) holds for such choices of f and g. Furthermore, the duals of Hs(Rd) and As(Cd) can be identified with H′ s(Rd) and s(Cd), respectively, through the forms in (1.15). The same holds true A′ with s(Rd), F ∈ As(Cd) and G ∈ A′ H0,s, H′ 0,s, A0,s, and A′ 0,s in place of Hs, H′ s, As, and A′ s, respectively, at each occurrence. 8 In order to identify the spaces of power series expansions above with spaces of analytic functions, we let 1 1 2σ z2 M 0 r1(loghz1i) r1z1 2 − (r1z1 M1,r,s(z) = 1,r,s(z) =(M1,r,s(z), M2,r,s(z) = 2,r,s(z) =(M2,r,s(z), M 0 2σ r1z12 + · · · + rdzd2, r1z1 σ−1 + · · · + rdzd 2 + (r1z1 z2 1 2σ σ−1 , 1 s ), s + · · · + rdzd s 6= 1 2, s = 1 2, r1z12 + · · · + rdzd2, 1 1−2s , s < 1 2, s = ♭σ, σ > 0, s ≥ 1 2, 1−2s + · · · + rd(loghzdi) σ+1 + · · · + rdzd 2σ σ+1 , 1 s ), s + · · · + rdzd s 6= 1 2, s = 1 2, s = ♭σ, σ > 1, s ≥ 1 2, (1.19) when r ∈ Rd + and z ∈ Cd. For conveniency we set Mr = M1,r,♭1. By [12] we have the following. The proof is therefore omitted. +. Then 2,r,s be as in (1.19) 1,r,s, M2,r,s and M 0 Proposition 1.5. Let M1,r,s, M 0 when s ∈ R♭ and r ∈ Rd A0,s(Cd) = { F ∈ A(Cd) ; F e−M 0 1,r,s ∈ L∞(Cd) for every r ∈ Rd As(Cd) = { F ∈ A(Cd) ; F e−M1,r,s ∈ L∞(Cd) for some r ∈ Rd A′ s(Cd) = { F ∈ A(Cd) ; F e−M2,r,s ∈ L∞(Cd) for every r ∈ Rd 0,s(Cd) = { F ∈ A(Cd) ; F e−M 0 A′ 2,r,s ∈ L∞(Cd) for some r ∈ Rd A′ ♭1(Cd) = A(Cd) Next we recall the link between the Bargmann transform and certain types of short-time Fourier transforms. The short-time Fourier trans- 1/2(Rd) with respect to the window φ ∈ S1/2(Rd) \ 0 is form of f ∈ S ′ defined by 0,♭1(Cd) = A({0}). + }, + }, + }, + }, and A′ s > 0, s > 0, s > ♭1, s > ♭1, Vφf (x, ξ) ≡ hf, φ( · − x)e−ih · ,ξii. We assume from now on that φ is given by (1.7), Let S be the dilation operator, defined by (SF )(x, ξ) = F (2− 1 2 x,−2− 1 2 ξ), 9 (1.20) when F ∈ L1 that loc(R2d). Then it follows by straight-forward computations (Vdf )(z) = (Vdf )(x + iξ) = (2π) d 2 e 1 2 (x2+ξ2)e−ihx,ξiVφf (2 1 2 x,−2 1 2 ξ) = (2π) d 2 e 1 2 (x2+ξ2)e−ihx,ξi(S−1(Vφf ))(x, ξ), (1.21) or equivalently, Vφf (x, ξ) = (2π)− d 2 e− 1 4 (x2+ξ2)e−ihx,ξi/2(Vdf )(2− 1 2 x,−2− 1 2 ξ) = (2π)− d 2 e−ihx,ξi/2S(e− · 2 2 (Vdf ))(x, ξ). (1.22) We observe that (1.21) and (1.22) can be formulated as V ◦ Vd = Vφ, Vd = UV ◦ Vφ, and U −1 where UV is the linear, continuous and bijective operator on D ′(R2d) ≃ D ′(Cd), given by (UVF )(x, ξ) = (2π) d 2 e 1 2 (x2+ξ2)e−ihx,ξiF (2 Let Dd,r(z0) be the polydisc 1 2 x,−2 1 2 ξ). (1.23) { z = (z1, . . . , zd) ∈ Cd ; zj − z0,j < rj, j = 1, . . . , d } with center and radii given by z0 = (z0,1, . . . , z0,d) ∈ Cd Then and r = (r1, . . . , rd) ∈ Rd +. A(Cd) = ∩r∈Rd + A(Dd,r(z)), A({0}) = ∪r∈Rd + A(Dd,r(z0)). 1.3. Hilbert spaces of power series expansions and analytic functions. The spaces in Definition 1.4 can also be dscribed by related unions and intersections of Hilbert spaces of analytic functions and power series expansions as follows. (See also [12].) A weight ω on a Borel set Ω is a positive function in L∞ loc(Ω) such loc(Ω). Let ϑ be a weight on Nd, ω be a weight that 1/ω belongs to L∞ on Cd, and let when F ∈ A′ 0(Cd) is given by (1.13), and 2 , [ϑ](Cd) ≡ Xα∈Nd c(F, α)ϑ(α)2! 1 2 z)2 dµ(z)(cid:19) 1 (ω)(Cd) ≡(cid:18)ZCd F (z)ω(2 1 2 (1.24) , (1.25) kFkA2 kFkA2 when F ∈ A(Cd). We let A2 is finite, and A2 that kFkA2 0(Cd) such [ϑ](Cd) be the set of all F ∈ A′ (ω)(Cd) be the set of all F ∈ A(Cd) such [ϑ] 10 that kFkA2 under these norms. (ω) is finite. It follows that these spaces are Hilbert spaces If ϑ and ω are related to each others as ϑ(α) = 1 α!ZRd + ω0(r)2rα dr! 1 2 , α ∈ Nd (1.26) and ω(z) = e z2 2 ω0(z12, . . . ,zd2), z ∈ Cd, (1.27) for some suitable weight ω0 on Rd +, then the following multi-dimensional version of [7, Theorem (4.1)] shows that A2 (ω)(Cd) with equal norms. Here we identify entire functions with their power series expansions at origin. Theorem 1.6. Let eα be as in (1.12), α ∈ Nd, and let ω0 be a positive measurable function on Rd +. Also let ϑ and ω be weights on Nd and Cd, respectively, related to each others by (1.26) and (1.27), and such that [ϑ](Cd) = A2 rα (α!) 1 2 . ϑ(α), α ∈ Nd, [ϑ](Cd) = A2 (1.28) holds for every r > 0. Then A2 norms. (ω)(Cd) with equality in We omit the proof of Theorem 1.6, since the result is an immediate consequence of [12, Theorem 3.5]. In our situation, the involved weights should satisfy a split condition. In one dimension, (1.26), (1.27) and (1.28) take the forms ϑj(αj) =(cid:18) 1 ω0,j(r)2rαj dr(cid:19) 1 2 αj!ZR+ 2 ω0,j(zj2), zj 2 ωj(zj) = e and rαj (αj!) 1 2 . ϑj(αj), , αj ∈ N, zj ∈ C (1.26)′ (1.27)′ r > 0, αj ∈ N. (1.28)′ Lemma 1.7. For j = 1, . . . , d, let ω0,j be weights on R+, ωj be weights on C and ϑj be weights on N such that (1.26)′ -- (1.28)′ hold, j = j=1 ω0,j(zj), z ≡ (z1, . . . , zd) ∈ Cd. If ϑ 1, . . . , d, and set ω0(z) ≡ Qd and ω are given by (1.26), and (1.27), then ϑ(α) = dYj=1 ϑj(αj), α = (α1, . . . , αd) ∈ Nd (1.29) 11 and ω(z) = dYj=1 ωj(zj), z = (z1, . . . , zd) ∈ Cd. (1.30) Proof. By [12, Theorem 3.5] and its proof, it follows that ω0,j · rαj ∈ L1(R+) for all j ∈ {1, . . . , d} and αj ∈ N. Hence, Fubini's theorem gives ϑ(α) = 1 α!ZRd + ω0(r)2rα dr! 1 2 = 1 α! dYj=1Z ∞ 0 ω0,j(rj)2rαj 2 j drj! 1 dYj=1 = ϑj(αj), and (1.29) follows. The assertion (1.30) follows from the definitions. (cid:3) 1.4. A test function space introduced by Gröchenig. In this sec- tion we recall some comparison results deduced in [12], between a test function space, SC(Rd), introduced by Gröchenig in [5] to handle mod- ulation spaces with elements in spaces of ultra-distributions. The definition of SC(Rd) is given as follows. Here we notice that the continuity of the map Vφ from S1/2(Rd) to S1/2(R2d) implies that its dual, V ∗ 1/2(Rd) is continuous. φ from S ′ 1/2(R2d) to S ′ Definition 1.8. Let φ(x) = π− d sist of all f ∈ S ′(Rd) such that f = V ∗ E ′(Rd) and F ∈ E ′(Rd), respectively. 4 e− x2 2 . Then SC(Rd) and SG(Rd) con- φ F , for some F ∈ L∞(Rd) ∩ It follows that f ∈ SC(Rd), if and only if 2 x−y2 f (x) = (2π)− d F (y, η)e− 1 2ZZR2d eihx,ηi dydη, (1.31) for some F ∈ L∞(Rd) ∩ E ′(Rd). Remark 1.9. By the identity (Vφh, F ) = (h, V ∗ φ F ) and the fact that the map (f, φ) 7→ Vφf is continuous from S (Rd) × S (Rd) to S (R2d), it follows that f = V ∗ φ F is uniquely defined as an element in S ′(Rd) when F ∈ S ′(R2d) (cf. [3]). In particular, the space SG(Rd) in Definition 1.8 is well-defined, and it is evident that SC(Rd) ⊆ SG(Rd). The following is a restatement of [12, Lemma 4.9]. The result is essential when deducing the characterizations of Pilipović spaces in Section 3. 12 Lemma 1.10. Let F ∈ L∞(Cd) ∪ E ′(Cd). Then the Bargmann trans- form of f = V ∗ φ F is given by ΠAF0, where √2ξ)e 1 2 (x2+ξ2)e−ihx,ξi, x, ξ ∈ Rd. F0(x + iξ) = (2π3) d 4 F (√2x,− (1.32) Moreover, the images of SC(Rd) and SG(Rd) under the Bargmann transform are given by { ΠAF ; F ∈ L∞(Cd) ∩ E ′(Cd) } and { ΠAF ; F ∈ E ′(Cd) }, (1.33) respectively. The next result is a straight-forward consequence of [12, Theorem 4.10]. The proof is therefore omitted. Proposition 1.11. It holds SC(Rd) = SG(Rd) = H♭1(Rd). Due to the image properties of the Bargmann transform, the next result is equivalent with the previous one. Proposition 1.12. The sets in (1.33) are equal to A♭1(Cd). In the next section we extend Propositions 1.11 and 1.12 by proving that the conclusions in Proposition 1.12 hold for suitable smaller and larger sets than those in (1.33). We also deduce similar identifications for other Pilipović spaces and their Bargmann images. 2. Paley-Wiener properties for Bargmann-Pilipović spaces In this section we consider spaces of compactly supported functions s(Cd). We show that the images of with interiors in As(Cd) or in A′ such functions under the reproducing kernel ΠA are equal to As(Cd), for some other choice of s ≤ ♭1. In the first part we state the main results given in Theorems 2.2 -- 2.8. They are straight-forward conse- quences of Propositions 2.9 -- 2.12, which in some sense contain more information. Thereafter we deduce results which are needed for their proofs. Depending of the choice of s, there are several different situa- tions for characterizing As(Cd). This gives rise to a quite large flora of main results, where each one takes care of one situation. In order to present the main results, we need the following definition. Definition 2.1. Let t1, t2 ∈ Rd + be such that t1 ≤ t2. Then the function χ ∈ L∞(Cd) is called positive, bounded and radial symmetric with respect to t1 and t2, if the following conditions are fulfilled: • χ ∈ L∞(Cd) ∩ E ′(Dt2(0)) is non-negative; • χ(z1, . . . , zd) = χ0(z1, . . . ,zd) for some function χ0; • χ ≥ c on Dt1(0) for some constant c > 0. 13 The set of positive, bounded and radial symmetric functions with re- t1,t2(Cd), and R∞(Cd) is defined spect to t1 and t2 is denoted by R∞ by R∞(Cd) ≡ [t1≤t2∈Rd + t1,t2(Cd). R∞ 2.1. Main results. We begin with characterizing the largest spaces in our investigations, which appears when s = ♭1, and then proceed with spaces of decreasing order. First we recall that elements in As(Cd) and A0,s(Cd) fulfill conditions of the form F (z) . er1z1 when s = ♭σ and of the form 2σ σ+1 +···+rdzd 2σ σ+1 , (2.1) (2.2) F (z) . er1(loghz1i) 1 1−2s +···+rd(loghzdi) 1 1−2s , when s ∈ [0, 1 2). Theorem 2.2. Let F ∈ A(Cd), σ = 1 and s > 1. Then the following conditions are equivalent: (1) F ∈ A♭1(Cd); (2) The estimate (2.1) holds for some r ∈ Rd +; (3) For some r0 ∈ Rd + and every r ∈ Rd + with r0 ≤ r and every χ ∈ r0,r(Cd), there exists F0 ∈ A(Dr(0)) such that F = ΠA(F0· χ); R∞ + with r0 < r and some χ ∈ R∞ r0,r(Cd), there exists F0 ∈ A(Dr(0)) such that F = ΠA(F0· χ); + and every r ∈ Rd (4) For some r0 ∈ Rd (5) There exists F0 ∈ E ′(Cd) ∩ L∞(Cd) such that F = ΠAF0; (6) There exists F0 ∈ E ′ s(Cd) such that F = ΠAF0. Remark 2.3. Since E ′(Cd) ∩ L∞(Cd) ⊆ E ′(Cd) ⊆ E ′ s(Cd), Theorem 2.2 still holds true after E ′ Theorem 2.4. Let F ∈ A(Cd), σ = 1 and χ ∈ R∞(Cd). Then the following is true: s has been replaces by E ′ in (6). (i) The following conditions are equivalent: (1) F ∈ A0,♭1(Cd); (2) The estimate (2.1) holds for every r ∈ Rd +; (3) There exists F0 ∈ A(Cd) such that F = ΠA(F0 · χ); morphism. (ii) The map F 7→ ΠA(F · χ) from A(Cd) to A0,♭1(Cd) is a homeo- The next result deals with the case when s = ♭σ with σ ∈ ( 1 2, 1). 14 Theorem 2.5. Let F ∈ A(Cd), χ ∈ R∞(Cd), σ ∈ ( 1 2, 1) and let Then the following is true: σ0 = σ . 2σ − 1 (i) The following conditions are equivalent: (1) F ∈ A♭σ(Cd) (F ∈ A0,♭σ(Cd)); (2) The estimate (2.1) holds for some (for every) r ∈ Rd +; (3) There exists F0 ∈ A′ (Cd) (F0 ∈ A′ 0,♭σ0 ♭σ0 (Cd)) such that F = ΠA(F0 · χ); (ii) The mappings F 7→ ΠA(F · χ) from A′ 0,♭σ0 (Cd) to A♭σ (Cd) and from A′ ♭σ0 (Cd) to A0,♭σ(Cd) are homeomorphisms. The next result deals with the case when s = ♭σ with σ = 1 2 . Theorem 2.6. Let F ∈ A(Cd), σ = 1 following is true: 2 and χ ∈ R∞(Cd). Then the (i) The following conditions are equivalent: (1) F ∈ A♭σ(Cd) (F ∈ A0,♭σ(Cd)); (2) The estimate (2.1) holds for some (for every) r ∈ Rd +; (3) There exists F0 ∈ A′ 0,1/2(Cd) (F0 ∈ A0,1/2(Cd)) such that 0,1/2(Cd) to A♭1/2(Cd) and (ii) The mappings F 7→ ΠA(F · χ) from A′ from A0,1/2(Cd) to A0,♭1/2(Cd) are homeomorphisms. F = ΠA(F0 · χ); The next result deals with the case when s = ♭σ with σ ∈ (0, 1 2). 2) and let Theorem 2.7. Let F ∈ A(Cd), χ ∈ R∞(Cd), σ ∈ (0, 1 Then the following is true: σ0 = σ . 1 − 2σ (i) The following conditions are equivalent: (1) F ∈ A♭σ(Cd) (F ∈ A0,♭σ(Cd)); (2) The estimate (2.1) holds for some (for every) r ∈ Rd +; (3) There exists F0 ∈ A♭σ0 (Cd) (F0 ∈ A0,♭σ0 (Cd)) such that F = ΠA(F0 · χ); (ii) The mappings F 7→ ΠA(F · χ) from A♭σ0 from A0,♭σ0 (Cd) to A0,♭σ (Cd) are homeomorphisms. (Cd) to A♭σ (Cd) and In the next result we consider the case when s ∈ [0, 1 2) is real. Theorem 2.8. Let F ∈ A(Cd), χ ∈ R∞(Cd), s ∈ [0, 1 following is true: 2). Then the (i) The following conditions are equivalent: 15 (1) F ∈ As(Cd) (F ∈ A0,s(Cd)); (2) The estimate (2.2) holds for some (for every) r ∈ Rd +; (3) There exists F0 ∈ As(Cd) (F0 ∈ A0,s(Cd)) such that F = (ii) The mappings F 7→ ΠA(F · χ) from As(Cd) to As(Cd) and from ΠA(F0 · χ); A0,s(Cd) to A0,s(Cd) are homeomorphisms. The previous theorems are essentially consequences of Propositions 2.9 -- 2.11, where more detailed information about involved constants are given. Proposition 2.9. Let F ∈ A(Cd), σ = 1, s > 1 and r ∈ Rd the following conditions are equivalent: +. Then + such that r0 < r, (2.1) holds with r0 in place of r; (1) For some r0 ∈ Rd (2) For some t1 ∈ Rd t2 < r and every χ ∈ R∞ such that F = ΠA(F0 · χ); (3) For some t1 ∈ Rd t1 ≤ t2 < r, there exist χ ∈ R∞ that F = ΠA(F0 · χDr(0)); L∞(Cd) such that F = ΠAF0; (4) For some r0 ∈ Rd (5) For some r0 ∈ Rd such that F = ΠAF0. + such that t1 < r, every t2 ∈ Rd + such that t1 < r and every t2 ∈ Rd + with t1 ≤ t1,t2(Cd), there exists F0 ∈ A(Dr(0)) + with t1,t2(Cd) and F0 ∈ A(Dr(0)) such + such that r0 < r, there exists F0 ∈ E ′(Dr(0))∩ + such that r0 < r, there exists F0 ∈ E ′ s(Dr0(0)) Theorems 2.5 -- 2.7 essentially follows from the following proposition. Proposition 2.10. Let τ > 1 let χ ∈ R∞ 2 , r, t1, t2 ∈ Rd t1,t2(Cd). Then the following is true: + be such that t1 ≤ t2, and (1) Let F ∈ A(Cd) be such that F (z) . er0,1z1 2τ +1 +···+r0,dzd 2 2τ +1 (2.3) holds for some r0 ∈ Rd + such that r0 < r. Then for some r0 ∈ + such that r0 < r, there exists F0 ∈ A(Cd) such that F = Rd ΠA(F0 · χ) and 2 2 16 where 2 2τ −1 , 2τ −1 +···+R0,dzd F0(z) . eR0,1z1 2 (cid:18) 2r0 2τ − 1 R0 = 2τ −1 2τ + 1(cid:19) 2τ +1 − 4 t 1 2τ −1 ; (2.4) (2.5) (2) Let r0, R0 ∈ Rd + be such that r0 < r and R0 = 2τ − 1 2 (cid:18) 2r0 2τ + 1(cid:19) 2τ +1 2τ −1 − 4 t 2 2τ −1 , (2.6) F0 ∈ A(Cd) be such that (2.4) holds and let F = ΠA(F0 · χ). Then F ∈ A(Cd) and satisfies (2.4) for some r0 ∈ Rd + such that r0 < r. Theorem 2.8 follows from the following two propositions, where the first one concerns the case when s > 0 and the second one make a more detailed explanation of the case s = 0, i. e. the case of analytic polynomials. Proposition 2.11. Let s ∈ (0, 1 the following is true: +, and let χ ∈ R∞(Cd). Then 2), r ∈ Rd (1) Suppose F ∈ A(Cd) satisfies F (z) . er0,1(loghz1i) 1 for some r0 ∈ Rd such that F = ΠA(F0 · χ) and 1−2s +···+r0,d(loghzdi) (2.7) + such that r0 < r. Then there is an F0 ∈ A(Cd) 1−2s 1 F0(z) . er0,1(loghz1i) 1 1−2s +···+r0,d(loghzdi) 1 1−2s (2.8) + such that r0 < r; (2) Suppose F0 ∈ A(Cd) satisfies (2.8) for some r0 ∈ Rd for some r0 ∈ Rd + such that r0 < r, and let F = ΠA(F0 · χ). Then F ∈ A(Cd) and satisfies (2.7) for some r0 ∈ Rd + such that r0 < r. Proposition 2.12. Let χ ∈ R∞(Cd) and let N ≥ 0 be an integer. Then the following is true: (1) Suppose F ∈ A(Cd) is given by c(F, α)zα, (2.9) F (z) = Xα≤N F0(z) = Xα≤N where {c(F, α)}α≤N ⊆ C. Then there is an F0 ∈ A(Cd) such that F = ΠA(F0 · χ) and c(F0, α)zα, (2.10) where {c(F0, α)}α≤N ⊆ C and satisfies c(F0, α) = 0 when c(F, α) = 0; (2) Suppose F0 ∈ A(Cd) satisfies (2.10) for some {c(F0, α)}α≤N ⊆ C, and let F = ΠA(F0 · χ). Then F ∈ A(Cd) and satisfies (2.9) for some {c(F, α)}α≤N ⊆ C such that c(F, α) = 0 when c(F0, α) = 0. 17 2.2. Preparing results and their proofs. For the proofs of Proposi- tions 2.9 -- 2.11 and thereby of Theorems 2.2 -- 2.8 we need some prepara- tory results. Because the proof of Proposition 2.9 needs some room, we put parts of the statement in the following separate proposition. At the same time we slightly refine some parts concerning the image of compactly supported elements in L∞ under the map ΠA. Proposition 2.13. Let s > 1, r0, r ∈ Rd suppose that either + be such that r0 < r and s(Dr0(0)), F0 ∈ E ′ F0 ∈ E ′ Then F = ΠAF0 ∈ A(Cd) and satisfies 0,s(Dr0(0)), F0 ∈ E ′(Dr0(0)) or F0 ∈ L∞(Dr(0)). F (z) . er1z1+···+rdzd. Proof. By the inclusions E ֒→ E ′ 0,s 0,s+ε ֒→ E ′ s ֒→ E ′ Let r2 = r. First suppose that F0 ∈ E ′ when ε > 0, it suffices to consider the case when F0 ∈ E ′ F0 ∈ L∞(Dr(0)) hold. s(Dr0(0)) holds, choose r1 ∈ Rd + such that r0 < r1 < r2, Ψ(y, η) = e−(y2+η2) and let Φz(y, η) = e(z,y+iη). By identifying Cd with R2d and using the fact that F0 ∈ E ′ s(Dr0(0)) we obtain s(Dr0(0)) or ΠAF0(z) = π−dhF0, ΦzΨi . sup α∈N2d kDα(ΦzΨ)kL∞(Dr1 (0)) hα 1 α!s ! (2.11) for every h1 > 0. We also have Ψ ∈ E1/2(R2d) ֒→ E0,s(R2d), which implies kDαΨkL∞(Dr1 (0)) . hα 2 α!s for every h2 > 0. Furthermore, DαΦz(y, η) = mα(z)e(z,y+iη) ≤ mα(z)er1,1z1+···+r1,dzd, where y+iη ∈ Dr1(0), mα(z) = dYj=1 zαj +αd+j j , z ∈ Cd, α ∈ N2d. 18 By choosing h1 = 4 and h2 = 1 above, and letting y + iη ∈ Dr1(0), Leibnitz rule gives 4−αα!−sDα(ΦzΨ)(y, η) ≤ 4−αα!−sXγ≤α(cid:18)α . 4−αα!−s Xγ≤α(cid:18)α . 4−α Xγ≤α(cid:18)α γ≤α dYj=1 γ(cid:19)DγΦz(y, η) Dα−γΨ(y, η) γ(cid:19)mγ(z)(α − γ)!s! er1,1z1+···+r1,dzd γ(cid:19)mγ(z)γ!−s! er1,1z1+···+r1,dzd γd+j!s ! er1,1z1+···+r1,dzd. · zjγd+j k(cid:1) ≤ 2k when n, k are non-negative sums are bounded by 2α, and that(cid:0)n γj! !s γj!s = (zj α∈N2d kDα(ΦzΨ)kL∞(Dr1 (0)) ! . e2s(z1 In the last inequality we have used that the number of terms in the integers such that k ≤ n. By combining (2.12) with the estimate s )er1,1z1+···+r1,dzd zjγj γj!s 1 s , ≤ eszj ≤ sup we get sup 1 s +···+zd 1 zjγj 1 s )γj (2.12) hα 1 α!s In the last inequality we have used the fact that r1 < r2 and s > 1. From the latter estimate and (2.11) we obtain . er2,1z1+···+r2,dzd. ΠAF0(z) . er2,1z1+···+r2,dzd, s(Dr0(0)). Suppose instead that F0 ∈ L∞(Q) holds, where Q = Dr2(0) ⊆ Cd, and the result follows when F0 ∈ E ′ and let Qj = Dr2,j (0) ⊆ C. Then ΠAF0(z) .ZQ F0(w)e(z,w) dλ(w) dYj=1 ZQj ≤ kF0kL∞ ezj wj dλ(wj)! . er2,1z1+···+r2,dzd, and the result follows in this case as well. (cid:3) 19 In the next lemma we give options on compactly supported functions which are mapped on the basic monomials, eα by the operator ΠA. Lemma 2.14. Let t1, t2 ∈ Rd and let χ0 be such that χ0(z1, . . . ,zd) = χ(z1, . . . , zd). If + be such that t1 ≤ t2, χ ∈ R∞ t1,t2(Cd), Fα,χ(z) = ςαzαχ(z), with ςα = 2−dα! 1 2 Z∆t2 χ0(u)e−u2 u2αu1 · · · ud du!−1 then the following is true: , ∆t = { u ∈ Rd + ; u ≤ t}, (2.13) (1) ΠAFα,χ = eα; (2) for some constant C > 0 which only depends on kχkL∞, c in Definition 2.1 and the dimension d, it holds C −1 dYj=1 2,j (αj + 1)! t−2α t−2 2 α! 1 2 ≤ ςα ≤ Cet12 dYj=1 1,j (αj + 1)! t−2α t−2 1 α! 1 2 (2.14) Proof. By using polar coordinates in each complex variable when inte- grating we get wαχ(w)e(z,w)−w2 dλ(w) Iα(u, z)e−u2 uαχ0(u)u1 · · · ud du, (2.15) ezjuj e−iθj! dθ = dYj=1 Iαj (uj, zj) (2.16) (ΠAFα,χ)(z) = π−dςαZCd = π−dςαZ∆t2 eihα,θi dYj=1 where with Iα(u, z) =Z[0,2π)d Iαj (uj, zj) =Z 2π Iαj (uj, zj) =Z 2π 0 0 By Taylor expansions we get eiαj θj ezj uje−iθj dθj. j e−ikθj k! ! dθj zk j uk eiαj θj ∞Xk=0 ∞Xk=0 (cid:18)Z 2π = 0 20 ei(αj −k)θj dθj(cid:19) zk j uk j k! ! = j uαj 2πzαj αj! j By inserting this into (2.15) and (2.16) we get (ΠAFα,χ)(z) = π−dςαZ∆t2 2Z∆t2 = 2dςαα!− 1 and (1) follows. (2π)d uαzα α! e−u2 uαχ0(u)u1 · · · ud du e−u2 u2αχ0(u)u1 · · · ud du! eα(z) = eα(z) Since χ is non-negative and fullfils χ ≥ c on Dt1(0) we get ςα . et12 α! = et12 α! 1 2 1 u2αu1 · · · u2 du!−1 2 Z∆t1 dYj=1 t2αj +2 2αj + 2!−1 ≍ et12 1,j α! 1 2 dYj=1 1,j (αj + 1)),! t−2α 1 (t−2 which gives the right inequality in (2.14). By the support properties of χ we also have ςα & α! 1 2 Z∆t2 u2αu1 · · · u2 du!−1 , and the left inequality in (2.14) follows by similar arguments, and (2) follows. (cid:3) The next lemma shows that we may estimate entire functions by different Lebesgue norms. We omit the proof, since the result follows from [11, Theorem 3.2]. Lemma 2.15. Suppose s, τ ∈ R and r, r0 ∈ Rd + are such that 1 τ > − 2 Let p, q ∈ [1,∞], F ∈ A(Cd) and set s < 1 2 , and r0 < r. M1,r(z) = r1z1 2τ +1 + · · · + rdzd 2 2 2τ +1 and Then M2,r(z) = r1(loghz1i) 1 1−2s + · · · + rd(loghzdi) 1 1−2s . kF · e−Mj,rkLp(Cd) . kF · e−Mj,r0kLq(Cd). The next lemma relates Lebesgue estimates of entire functions with estimates on corresponding Taylor coefficients. Here we let the Gamma 21 function on Rd + be defined by Γd(x1, . . . , xd) = Γ(xj), dYj=1 1 where Γ is the Gamma function on R+. Lemma 2.16. Let τ > − 1 2, M1,r be the same as in Lemma 2.15, ω(z) = 2 z2−M1,r(z), j = 1, 2, z ∈ Cd, and let α0 = (1, . . . , 1) ∈ Nd. Also let e (cid:19)(cid:19) 1 ϑr(α) =(cid:18)(2τ + 1)(2r)−(2τ +1)(α+α0)(cid:18) Γ ((2τ + 1)(α + α0)) . 2 Xα∈Nd c(F, α)ϑr(α)2! 1 If F ∈ A(Cd) is given by (1.13), then kF · e−M1,rkL2(Cd) = π α! , d 2 2 [ϑr](Cd) = A2 (ω)(Cd) with equality in norms. and A2 Proof. Since e−M1,r(z) = e−rjzj 2 2τ +1 , dYj=1 Lemma 1.7 shows that we may assume that d = 1, giving that r = r1 and α0 = 1. In view of Theorem 1.6 we have A2 in norms, when [ϑ](Cd) = A2 (ω1)(Cd) with equality ϑ(α) =(cid:18) 1 α!Z ∞ 0 e−2rt 2 1 2τ +1 tα dt(cid:19) 1 , provided it is verified that (1.28) holds for every r > 0. By taking u = 2rt 2τ +1 as new variables of integration we obtain 1 ϑ(α) =(cid:18)(2τ + 1)(2r)−(2τ +1)(α+1) 1 α!Z ∞ 0 =(cid:18)(2τ + 1)(2r)−(2τ +1)(α+1)(cid:18)Γ ((2τ + 1)(α + 1)) α! 2 e−uuα(2τ +1)+2τ du(cid:19) 1 (cid:19)(cid:19) 1 2 . Since Γ((2τ + 1)(α + 1)) & rα for every r > 0, by Stirling's formula, (1.28) holds for every r > 0. This implies that ϑ = ϑ1,r, and the result follows. (cid:3) We also need the following version of Stirling's formula. Lemma 2.17. Let α ≥ 0 be an integer and let τ > − 1 2 . Then Γ ((2τ + 1)(α + 1)) α! ≍ (2τ + 1)(2τ +1)·α (α + 1)τ · α!2τ . 22 (2.17) Lemma 2.17 follows by repeated applications of Stirling's formula and the standard limit lim t→∞(cid:16)1 + x t(cid:17)t = ex for every x ∈ R. In order to be self-contained we present the arguments. Proof. The result is obviously true for α = 0. For α ≥ 1 we have ατ ≍ (α + 1)τ . A combination of the latter relations and Stirling's formula gives 2 α2τ (α+1) Γ ((2τ + 1)(α + 1)) α! ≍ ((2τ + 1)α + 2τ )(2τ +1)·α+2τ + 1 2 eα+2τ (α+1) eα αα+ 1 2 · = 2τ e2τ (α+1) (2τ + 1)(2τ +1)·α+2τ + 1 2(cid:18)1 + ≍ (2τ + 1)(2τ +1)·α α2τ (α+1) (2τ + 1)α(cid:19)(2τ +1)·α+2τ + 1 (2τ + 1)α(cid:19)(2τ +1)α ·(cid:18)1 + ≍ (2τ + 1)(2τ +1)·α α2τ (α+1) e2τ α ≍ (2τ + 1)(2τ +1)·α (α + 1)τ α!2τ . α(cid:17)x2α x1, x2 ∈ R+, (cid:16)1 + e2τ α x1 2τ , In the forth relation we have used the fact that (2.18) increases with α and has the limit ex1x2 as α tends to infinity. This gives the result. (cid:3) Proposition 2.9 essentially follows from the following lemma. Lemma 2.18. Let τ > − 1 A(Cd). Then the following is true: 2 , r0, r ∈ Rd + be such that r0 < r and F ∈ (1) If then (2) If then 2 2τ + 1(cid:19) 2τ +1 c(F, α) .(cid:18) 2r0 F (z) . er1z1 2 2τ +1 +···+rdzd ·α α!−τ , 2 2τ +1 ; 2 2τ +1 +···+r0,dzd 2 2τ +1 F (z) . er0,1z1 c(F, α) .(cid:18) 2r 2 2τ + 1(cid:19) 2τ +1 23 ·α α!−τ (2.19) (2.20) (2.21) (2.22) Proof. Let ϑr be the same as in Lemma 2.16. First we prove (1). Sup- pose that (2.19) holds, let r1 ∈ Rd + be such that r0 < r1 < r and let α0 = (1, . . . , 1) ∈ Nd. Also let M1,r be the same as in Lemma 2.15. Then Lemmata 2.15 and 2.16 give kF · e−M1,rkL∞(Cd) . kF · e−M1,r1kL2(Cd) 2 2 1 2 ≍ Xα∈Nd c(F, α)ϑr1(α)2! 1 =Xα∈Nd(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 2 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) r1(cid:19)(α+α0) 2τ +1 c(F, α)ϑr0(α)(cid:18)r0 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) α∈Nd(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) dYj=1 α∈Nd c(F, α)(cid:18)2τ + 1 2r0 (cid:19) 2τ +1 2!(cid:18) r0 Xα∈Nd dYj=1 r1(cid:19)(2τ +1)·(α+α0)! c(F, α)ϑr0(α) (αj + 1)− τ ≍ sup (αj + 1) . sup 2 τ ·α α!τ! . Here the second inequality follows from the fact that is convergent since r0 < r1, and the fifth relation follows from Lemma 2.17. This implies that (2.20) holds and (1) follows. Next assume that (2.21) holds. By Lemmata 2.15 and 2.16 we get kF · e−M1,r0kL∞(Cd) & kF · e−M1,rkL2(Cd) 2 α∈Nd (c(F, α) ϑr(α)) ≥ sup = Xα∈Nd c(F, α)ϑr(α)2! 1 α∈Nd c(F, α)(cid:18)2τ + 1 2r (cid:19) 2τ +1 α∈Nd c(F, α)(cid:18)2τ + 1 2 α dYj=1 ≥ sup 2 α!τ! 2r (cid:19) 2τ +1 (αj + 1) & sup τ 2 α α!τ! , where the third inequality follows from Lemma 2.17. This gives (2). (cid:3) Proposition 2.11 mainly follows from the following result. + be such that r0 < r, s ∈ (0, 1 Lemma 2.19. Let r, r0 ∈ Rd F ∈ A(Cd). Then the following is true: 24 2) and let (1) if (2.7) holds, then 1 2s +···+Rdαd c(F, α) . e−(R1α1 (2) if R0 ∈ Rd c(F, α) . e−(R0,1α1 + is given by 1 2s +···+R0,dαd 1 2s ; r (cid:19) 1−2s 2s ), R = s(cid:18)1 − 2s 2s ), R0 = s(cid:18)1 − 2s r0 (cid:19) 1−2s 2s 1 (2.23) , (2.24) then (2.7) holds with r in place of r0. Proof. Let r1, r2 ∈ Rd + be such that r1 < r < r2. By Lemma 1.7 we may assume that d = 1, and by Lemma 2.15 the result follows if we prove sup α∈N(cid:16)c(F, α)eR2·α 1 2s 2(cid:17) .ZC F (z)e−r(loghzi) 1 1−2s 2 dλ(z) α∈N(cid:16)c(F, α)eR1·α . sup 1 2s 2(cid:17) , (2.25) where R1, R2 ∈ R+ satisfy By Theorem 1.6 we have Rj =(cid:18)1 − 2s 1 , 2s j = 1, 2. rj (cid:19) 1−2s 1−2s 2 dλ(z) ≍Xα∈Nc(F, α)ϑr(α)2, 2α!Z ∞ 1−2s tα dt(cid:19) 1 e−r(loghti) . 0 1 2 ZC F (z)e−r(loghzi) ϑr(α) =(cid:18) π where Let θ = 1 1 − 2s and gr,α(t) = e−r(log t)θ tα. In order to prove (2.25), we need to prove eR2α 1 2s . ϑr(α) . eR1α 1 2s , (2.26) which shall be reached by modifying the proof of (15) in [4]. We have ϑr(α)2 .Z ∞ e e(r1−r)(log t)θ gr1,α(t) dt . sup t≥e (gr1,α(t))Z ∞ e 25 e(r1−r)(log t)θ dt ≍ sup t≥e (gr1,α(t)). By straight-forward computations it follows that gr,α(t) attains its global maximum for tr,α = exp(cid:18)(cid:16) α 2s (cid:19) , θr(cid:17) 1−2s (2.27) and that gr1,α(tr1,α) = exp(cid:16)r 1− 1 2s 1 θ− 1 2s (θ − 1)α 1 2s(cid:17) = e2R1·α 1 2s , and the second inequality in (2.26) follows. In order to prove the first inequality in (2.26), we claim that for some c which is independent of α we have (cid:18)1 − c tr2,α(cid:19)α & e(r−r2)(log tr2,α)θ . (2.28) In fact, by (2.27) and Taylor's formula, (2.28) follows if we prove 1 − αc tr2,α & e(r−r2)((1−2s)α/r2) 1 2s (2.29) for some c > 0 which is independent of α. In view of (2.27) it follows that tr2,α tends to infinity as α turns to infinity, which implies that the left-hand side of (2.29) has the limit one and the right-hand side has the limit zero as α tends to infinity. Hence, (2.28) holds. By (2.28) and the fact that 1 2s > 1 we get ϑr(α)2 & 1 α!Z tr2,α tr2,α−c e−r(log t)θ tα dt & 1 α! e−r(log tr2,α)θ (tr2,α − c)α & 1 α! e−r2(log tr2,α)θ tα r2,α = e2R2·α 1 2s & e2R3·α 1 2s . 1 α! This gives the result. (cid:3) 2.3. Proofs of main results. Next we prove Proposition 2.9 and thereby Theorem 2.2. Proof of Proposition 2.9. It is clear that (2) ⇒ (3) ⇒ (4) ⇒ (5). By Proposition 2.13 it follows that (5) implies (1). We need to prove that (1) implies (2). Suppose (1) holds and let r4 = r. Choose r1, r2, r3 ∈ Rd + such that r2,r3(Cd), and let Fα,χ be r0 < r1 < r2 ≤ r3 < r4 and r1r4 < r2 as in Lemma 2.14 with r2 and r3 in place of r1 and r2. If τ = 1, then Lemma 2.18 (2) gives 2, χ ∈ R∞ Let 1 α!− 1 2 . c(F, α) . rα F0(z) = Xα∈Nd 26 c(F, α)ςαzα. (2.30) (2.31) We claim that the series in (2.31) is uniformly convergent with respect to z in DR(0). In fact, if zj ≤ r4,j, then (2.14) gives rα c(F, α)ςαzα . hαidr−2α 1 rα 2 4 = hαidρα, + satisfies ρj = r1,j r4,j where ρ ∈ Rd gent, Weierstrass theorem shows that (2.31) is uniformly convergent and defines an analytic function in Dr4,j (0). Hence, F0 ∈ A(Dr4,j (0)). Furthermore, by the support of χ we get < 1. SincePα∈Ndhαidρα is conver- r2 2,j ΠA(F0 · χ) = ΠA Xα∈Nd = Xα∈Nd c(F, α)ςαzα · χ! c(F, α)ςαΠA(zα · χ) = Xα∈Nd Hence (2) holds, and the proof is complete. c(F, α)eα = F. (2.32) (cid:3) For future references we observe that if ςα and χ are the same as in Lemma 2.14, then (2.31) shows that the relationship between c(F, α) and c(F0, α) is given by c(F0, α) = c(F, α)ςαα! 1 2 . (2.33) Proof of Theorem 2.4. The equivalence between (1) and (2) is clear. It is also obvious that (3) implies (2) in view of Proposition 2.9, and that (3) implies (4). We shall prove that (1) implies (3) and (4) implies (1). +. Let r, R ∈ r,R(Cd), Fα,χ be as in Lemma 2.14 and let F0 be given by Suppose (1) holds. Then (2.30) holds for every r1 ∈ Rd Rd +, χ ∈ R∞ (2.31). By (2.30) we have for every r1 ∈ Rd +, giving that c(F, α)ςα . hαidr−2αrα 1 c(F, α)ςα . rα 0 for every r0 ∈ Rd +. This implies that the series in (2.31) is locally uniformly convergent with respect to z and defines an entire function on Cd. Hence F0 ∈ A(Cd). Moreover, by (2.32) it follows that ΠA(F0χ) = F , and we have proved that (1) implies (3). Next suppose that (4) holds. Then c(F0, α) . rαα! 1 2 for every r ∈ Rd +. By (2.14) and (2.33) we get 2 . rας −1 α α!− 1 c(F, α) = c(F0, α)ς −1 α . rαR2αα!− 1 2 . Since r ∈ Rd + can be chosen arbitrarily small we get c(F, α) . rα 27 0 α!− 1 2 for every r0 ∈ Rd proved that (4) implies (1), and the result follows. +. This implies that F ∈ A0,♭1(Cd). That is, we have (cid:3) Next we prove Proposition 2.10. Proof of Proposition 2.10. Suppose (2.3) holds, and let r1, r2 ∈ Rd such that r0 < r1 < r2 < r. Then Lemma 2.18 gives + be 2 2τ + 1(cid:19) 2τ +1 ·α α!−τ . By Lemma 2.14 and (2.33) we get c(F, α) .(cid:18) 2r1 2τ + 1(cid:19) 2τ +1 2τ + 1(cid:19) 2τ +1 ·α ·α 2 2 c(F0, α) .(cid:18) 2r1 .(cid:18) 2r2 when α!−τhαidα!t−2α 1 α!−(τ −1)t−2α 1 =(cid:18) 2R0 2τ − 1(cid:19) 2τ −1 2 ·α α!−(τ −1), R0 = This proves (1). 2τ − 1 2 (cid:18) 2R2 2τ + 1(cid:19) 2τ +1 2τ −1 − 4 t 1 2τ −1 . Suppose instead that r0, R0, r1, r2 ∈ Rd + are such that (2.6) hold and r0 < r1 < r2 < r, and that F0 ∈ A(Cd) satisfies (2.4). Also let F = ΠA(F0 · χ). Then By combining the latter estimate with Lemma 2.18 we get 2 2τ + 1(cid:19) 2τ +1 t−2 2 . 2 (cid:18) 2R0 2τ − 1(cid:19) 2τ −1 c(F0, α) .(cid:18) 2r1 c(F, α) .(cid:18) 2r1 <(cid:18) 2r1 2τ + 1(cid:19) 2τ +1 2τ + 1(cid:19) 2τ +1 ·α 2 2 t−2α 2 α!1−τ . ·α α!−τ Hence Lemma 2.14 and (2.33) give and Lemma 2.18 again implies that (2.3) holds with r2 in place of r0. This gives the result. (cid:3) Proof of Theorems 2.5 -- 2.7. First suppose σ ∈ ( 1 and τ = 1 2, 1), and let σ0 = σ 2σ−1 2σ . Then 2σ = 2 2τ + 1 and σ + 1 2σ0 σ0 − 1 = 2 . 2τ − 1 Theorem 2.5 now follows from these observations and Proposition 2.10 in the case τ ∈ ( 1 2, 1). 28 In the same way, Theorem 2.6 follows by choosing τ = 1 in Proposi- tion 2.10. Finally, suppose σ ∈ (0, 1 2 ), and let σ0 = σ 2 2σ 2σ0 = σ + 1 2τ + 1 and = σ0 + 1 1−2σ and τ = 1 2σ . Then 2 , 2τ − 1 and Theorem 2.5 follows from these observations and Proposition 2.10 in the case τ > 1. (cid:3) Next we prove Propositions 2.11 and 2.12 and thereby Theorem 2.8. Proof of Propositions 2.11 and 2.12. Let r, rj, Rj ∈ Rd be such that +, j = 0, 1, 2, 3, r0 < r1 < r2 < r3 < r and Rj = s(cid:18)1 − 2s rj (cid:19) 1−2s 2s . First suppose that F ∈ A(Cd) satisfies (2.7) and let F0 be the formal power series expansion with coefficients given by (2.33). Then F = ΠA(F0 · χ). By Lemma 2.19 we get c(F, α) . e−(R1,1α1 Hence Lemma 2.14 and (2.33) give 1 2s +···+R1,dαd 1 2s ). c(F0, α) . α!t−2α 1 e−(R1,1α1 1 2s +···+R1,dαd 1 2s ) . e−(R2,1α1 1 2s +···+R2,dαd 1 2s ). In the last inequality we have used the fact that s < 1 that R2 < R1 and 2 , which implies t−2α 1 α! . e(R1,1−R2,1)α1 1 2s +···+(R1,d−R2,d)αd 1 2s . By applying Lemma 2.19 again it follows that F0 satisfies (2.8) with r3 in place of r0. This gives (1). Suppose instead that F0 ∈ A(Cd) satisfies (2.8) and let F = ΠA(F0 · χ). Then Lemma 2.19 gives c(F0, α) . e−(R1,1α1 1 2s +···+R1,dαd 1 2s ), and it follows from Lemma 2.14 and (2.33) that c(F, α) . α!−1t2α 2 e−(R1,1α1 1 2s +···+R1,dαd 1 2s ) . e−(R1,1α1 1 2s +···+R1,dαd 1 2s ). By applying Lemma 2.19 again we deduce (2.7) with r2 in place of r0, and Proposition 2.11 follows. Proposition 2.12 is a straight-forward consequence of (2.33). The (cid:3) details are left for the reader. 29 3. Characterizations of Pilipović spaces In this section we combine Lemma 1.10 with Theorems 2.2 -- 2.8 to get characterizations of Pilipović spaces. We begin with the following characterization of H♭1. The result is a straight-forward combination of Lemma 1.10 and Theorem 2.2. The details are left for the reader. Proposition 3.1. Let φ be as in (1.7), r ∈ Rd +, χr be the characteristic function for Dr(0) and let s > 0. Then the following conditions are equivalent: (1) f ∈ H♭1(Rd); (2) f = V ∗ (3) f = V ∗ (4) f = V ∗ s(R2d); φ F for some F ∈ E ′ φ F for some F ∈ E ′(R2d) ∩ L∞(R2d); φ F for some F ∈ E ′(R2d) ∩ L∞(R2d) which satisfies F (x, ξ) = F0(x − iξ)e−ihx,ξiχ(x, ξ) +, χ = χr and F0 ∈ A(Dr(0)). for some r ∈ Rd (3.1) Remark 3.2. It is clear that χ in Proposition 3.1 can be chosen as any χ ∈ R∞(Cd) with suitable support properties. 2 (x2+ξ2) which is absent Remark 3.3. In (1.32) there is a factor e in (3.1). We notice that this factor is not needed in (3.1) because t1,t2(Cd) is invariant under multiplications of such functions. R∞ The next result follows from Lemma 1.10 and Theorems 2.4 -- 2.8. The 1 details are left for the reader. Proposition 3.4. Let φ be as in (1.7) and χ ∈ R∞(Cd). Then the following conditions are equivalent: (1) f ∈ H0,♭1(Rd); (2) f = V ∗ for some F0 ∈ A(Cd). φ F for some F ∈ E ′(R2d)∩ L∞(R2d) which satisfies (3.1) Proposition 3.5. Let φ be as in (1.7), χ ∈ R∞(Cd), σ ∈ ( 1 let 2, 1) and σ0 = σ . 2σ − 1 Then the following conditions are equivalent: (1) f ∈ H♭σ (Rd) (f ∈ H0,♭σ (Rd)); (2) f = V ∗ φ F for some F ∈ E ′(R2d)∩ L∞(R2d) which satisfies (3.1) for some F0 ∈ A′ 0,♭σ0 (Cd) (F0 ∈ A′ ♭σ0 (Cd)). Proposition 3.6. Let φ be as in (1.7), χ ∈ R∞(Cd) and let σ = 1 2. Then the following conditions are equivalent: (1) f ∈ H♭σ (Rd) (f ∈ H0,♭σ (Rd)); 30 (2) f = V ∗ φ F for some F ∈ E ′(R2d)∩ L∞(R2d) which satisfies (3.1) for some F0 ∈ A′ 0,1/2(Cd) (F0 ∈ A0,1/2(Cd)). Proposition 3.7. Let φ be as in (1.7), χ ∈ R∞(Cd), σ ∈ (0, 1 let 2) and Then the following conditions are equivalent: σ σ0 = . 1 − 2σ (1) f ∈ H♭σ (Rd) (f ∈ H0,♭σ (Rd)); (2) f = V ∗ φ F for some F ∈ E ′(R2d)∩ L∞(R2d) which satisfies (3.1) for some F0 ∈ A♭σ0 (Cd) (F0 ∈ A0,♭σ0 (Cd)). Proposition 3.8. Let φ be as in (1.7), χ ∈ R∞(Cd) and let s ∈ (0, 1 2). Then the following conditions are equivalent: (1) f ∈ Hs(Rd) (f ∈ Hs(Rd)); (2) f = V ∗ φ F for some F ∈ E ′(R2d)∩ L∞(R2d) which satisfies (3.1) for some F0 ∈ As(Cd) (F0 ∈ A0,s(Cd)). References [1] V. Bargmann On a Hilbert space of analytic functions and an associated inte- gral transform, Comm. Pure Appl. Math., 14 (1961), 187 -- 214. [2] V. Bargmann On a Hilbert space of analytic functions and an associated in- tegral transform. Part II. A family of related function spaces. Application to distribution theory., Comm. Pure Appl. Math., 20 (1967), 1 -- 101. [3] E. Cordero, S. Pilipović, L. Rodino, N. Teofanov Quasianalytic Gelfand-Shilov spaces with applications to localization operators, Rocky Mt. J. Math. 40 (2010), 1123-1147. [4] C. Fernandez, A. Galbis, J. Toft The Bargmann transform and powers of har- monic oscillator on Gelfand-Shilov subspaces, RACSAM 111 (2017), 1 -- 13. [5] K. H. Gröchenig Foundations of Time-Frequency Analysis, Birkhäuser, Boston, 2001. [6] L. Hörmander The Analysis of Linear Partial Differential Operators, vol I -- III, Springer-Verlag, Berlin Heidelberg NewYork Tokyo, 1983, 1985. [7] A.M.E.M. Janssen, S.J.L. Eijndhoven Spaces of type W, growth of Hermite coefficients, Wigner distribution, and Bargmann transform, J. Math. Anal. Appl. 152 (1990), 368 -- 390. [8] S. Pilipović Generalization of Zemanian spaces of generalized functions which have orthonormal series expansions, SIAM J. Math. Anal. 17 (1986), 477-484. [9] S. Pilipović Tempered ultradistributions, Boll. U.M.I. 7 (1988), 235 -- 251. [10] M. Reed, B. Simon Methods of modern mathematical physics, Academic Press, London New York, 1979. [11] J. Toft The Bargmann transform on modulation and Gelfand-Shilov spaces, with applications to Toeplitz and pseudo-differential operators, J. Pseudo- Differ. Oper. Appl. 3 (2012), 145 -- 227. [12] J. Toft Images of function and distribution spaces under the Bargmann trans- form, J. Pseudo-Differ. Oper. Appl. 8 (2017), 83 -- 139. 31 Department of Mathematics, Linnaeus University, Växjö, Sweden E-mail address: [email protected] Department of Mathematics, University of Regensburg, Regensburg, Germany E-mail address: [email protected] Department of Mathematics, Linnaeus University, Växjö, Sweden E-mail address: [email protected] 32
1704.06950
1
1704
2017-04-23T17:09:39
Self-Adjoint Operators in Extended Hilbert Spaces $H\oplus W$: An Application of the General GKN-EM Theorem
[ "math.FA" ]
We construct self-adjoint operators in the direct sum of a complex Hilbert space $H$ and a finite dimensional complex inner product space $W$. The operator theory developed in this paper for the Hilbert space $H\oplus W$ is originally motivated by some fourth-order differential operators, studied by Everitt and others, having orthogonal polynomial eigenfunctions. Generated by a closed symmetric operator $T_{0}$ in $H$ with equal and finite deficiency indices and its adjoint $T_{1}$, we define \textit{families} of minimal operators $\{\widehat{T}_{0}\}$ and maximal operators $\{\widehat{T}_{1}\}$ in the extended space $H\oplus W$ and establish, using a recent theory of complex symplectic geometry, developed by Everitt and Markus, a characterization of self-adjoint extensions of $\{\widehat{T}_{0}\}$ when the dimension of the extension space $W$ is not greater than the deficiency index of $T_{0}$. A generalization of the classical Glazman-Krein-Naimark (GKN) Theorem - called the GKN-EM Theorem to acknowledge the work of Everitt and Markus - is key to finding these self-adjoint extensions in $H\oplus W.$ We consider several examples to illustrate our results.
math.FA
math
SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES H ⊕ W : AN APPLICATION OF THE GENERAL GKN-EM THEOREM LANCE L. LITTLEJOHN AND RICHARD WELLMAN Abstract. We construct self-adjoint operators in the direct sum of a complex Hilbert space H and a finite dimensional complex inner product space W . The operator theory developed in this paper for the Hilbert space H ⊕ W is originally motivated by some fourth-order differential operators, studied by Everitt and others, having orthogonal polynomial eigenfunctions. Generated by a closed symmetric operator T0 in H with equal and finite deficiency indices and its adjoint T1, we define families of minimal operators { bT0} and maximal operators { bT1} in the extended space H ⊕ W and establish, using a recent theory of complex symplectic geometry, developed by Everitt and Markus, a characterization of self-adjoint extensions of { bT0} when the dimension of the extension space W is not greater than the deficiency index of T0. A generalization of the classical Glazman-Krein- Naimark (GKN) Theorem - called the GKN-EM Theorem to acknowledge the work of Everitt and Markus - is key to finding these self-adjoint extensions in H ⊕ W. We consider several examples to illustrate our results. 7 1 0 2 r p A 3 2 ] . A F h t a m [ 1 v 0 5 9 6 0 . 4 0 7 1 : v i X r a Contents Introduction 1. 2. The von-Neumann Formulas and the GKN Theorem 3. Complex Symplectic Geometry and a Generalization of the GKN Theorem 4. Maximal and Minimal Operators in H ⊕ W 5. Examples 5.1. Example 1: The Legendre Type Self-Adjoint Operator 5.2. Example 2: A Simple First-Order Differential Operator 5.3. Example 3: Variations on the Fourier Self-Adjoint Operator References 1 4 8 13 21 21 25 27 31 1. Introduction In [6, p. 105], the authors list ten open problems related to orthogonal polynomial eigenfunctions of differential equations. The one pertinent to this present paper is the following (paraphrased to simplify the original notation): The GKN theory provides a recipe, in theory, for determining all self-adjoint exten- sions in the Hilbert space L2(I; w) of formally symmetric differential expressions of Date: April 23, 2017 [wellmanlittlejohnEMfinalversion.tex]. 1991 Mathematics Subject Classification. Primary 05C38, 15A15; Secondary 05A15, 15A18. Key words and phrases. symmetric operator, self-adjoint operator, differential operator, maximal operator, mini- mal operator, Glazman-Krein-Naimark theory, symplectic GKN theorem, orthogonal polynomialsct . 1 2 LANCE L. LITTLEJOHN AND RICHARD WELLMAN the form (1.1) ℓ2r[y](u) = 1 w(u) rXj=0 (−1)j (qj(u)y(j)(u))(j) (u ∈ I) on some open interval I = (a, b); we assume here that w > 0 and each coefficient qj is sufficiently differentiable on I. This theory works well in developing the spectral theory for the second-order classical differential equations of Jacobi, Laguerre, and Hermite.1 However, for nonclassical symmetric differential equations (1.1) with orthogonal polynomial solutions, the appropriate right-definite setting is a Hilbert- Sobolev space S with orthogonalizing Sobolev inner product (1.2) hf, gi =Z b a f (u)g(u)w(u)du + pXj=0(cid:16)αjf (j)(a)g(j)(a) + βjf (j)(b)g(j)(b)(cid:17) . The Sobolev space S has the form L2(I; w)⊕ Ck for some k ≤ 2p. Develop a general GKN-type theory for this setting; in particular, provide a 'recipe' for determining the self-adjoint operator having the orthogonal polynomials as eigenfunctions. In this paper, we answer this question. In fact, we will see that we can provide a recipe for all self-adjoint operators, generated by ℓ2r[·], in this Sobolev setting. Our result is a generalization of the Glazman-Krein-Naimark (GKN) theory of self-adjoint extensions of Lagrangian symmetric ordinary differential expressions in a weighted Hilbert space L2(I; w), where I is an interval of the real line R. This work is originally motivated by fourth-order differential equations having non-classical or- thogonal polynomials as eigenfunctions. In each of these fourth-order examples, the orthogonalizing inner product has the form hf, gi = Af (a)g(a) +Z b a f (u)g(u)w(u)du + Bf (b)g(b), where A, B ≥ 0. Indeed, H. L. Krall [13, 14] classified, up to a complex linear change of variable, these orthogonal polynomials which were subsequently named the Legendre type, Laguerre type and Jacobi type polynomials and studied extensively by A. M. Krall in [12]. Following the work of the two Kralls, other contributions connecting orthogonal polynomial eigenfunctions to higher-order differential equations have emerged; all known examples have polynomial eigenfunctions orthogonal with respect to an inner product of the form (1.2). These various contributions are far too numerous to list in this manuscript but we refer to the Erice and Patras reports [5, 6] for further details and the references therein contained. In [3, 4], the authors construct the Legendre type self-adjoint operator, generated by the fourth- order Legendre type differential expression (1.3) L2 in the Hilbert space L2 ℓLT [y](u) := (u2 − 1)2y(4)(u) + 8u(u2 − 1)y(3)(u) + (4A + 12)(u2 − 1)y′′(u) + 8Auy′(u), µ[−1, 1] = {f : [−1, 1] → C f is Lebesgue measurable and hf, fiµ =Z[−1,1] f2 dµ < ∞}, µ[−1, 1]. Here 1The GKN theory is applicable as well to developing the self-adjoint operator theory associated with exceptional orthogonal polynomials. SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 3 where hf, giµ = Af (−1)g(−1) +Z 1 −1 f (u)g(u)du + Af (1)g(1). µ[−1, 1] is isometrically isomorphic to L2(−1, 1)⊕ C2. The classical Glazman-Krein- We note that L2 Naimark (GKN) theory of self-adjoint extensions of Lagrangian symmetric differential expressions is not (immediately) applicable in this situation. To develop the appropriate operator theory in L2 µ[−1, 1], Everitt and Littlejohn studied properties of functions in the maximal domain ∆ of ℓLT [·] in the base space L2(−1, 1). They prove the surprising smoothness condition f ∈ ∆ ⇒ f ′′ ∈ L2(−1, 1) from which it follows that f, f ′ ∈ AC[−1, 1]. Using standard operator-theoretic methods, they then prove that the operator T : D(T ) ⊂ L2 µ[−1, 1] defined by µ[−1, 1] → L2 (T f )(u) =  D(T ) = ∆ −8Af ′(−1) u = −1 ℓLT [f ](u) −1 < u < 1 8Af ′(1) u = 1 is self-adjoint. It is remarkable that ∆ is the domain of the self-adjoint operator T in L2 µ[−1, 1]. Indeed, the expression ℓLT [·] is in the limit-3 case at both singular endpoints x = ±1 in L2(−1, 1) so every self-adjoint operator in L2(−1, 1), generated by ℓLT [·], is necessarily determined by two appropriate boundary restrictions on the space ∆. We will re-examine this Legendre type example in Section 5.1 as an application of the results developed in this paper. To this end, let (W,h·,·iW ) be a finite-dimensional complex inner product space and assume (H,h·,·iH ) is a complex Hilbert space. Then H ⊕ W , the direct sum of H and W, is the Hilbert space defined by (1.4) with inner product (1.5) and associated norm H ⊕ W = {(x, a) x ∈ H, a ∈ W} h(x, a), (y, b)iH⊕W := hx, yiH + ha, biW k(x, a)k2 H⊕W = kxk2 H + kak2 W . Throughout this paper, we refer to H ⊕ W as an extended Hilbert space and call H the base space and W the extension space. Our starting point in this paper - assumptions we keep throughout this article - is a closed, symmetric operator T0 in H having equal and finite deficiency indices, denoted by their common value def(T0), and adjoint operator T1 satisfying the inclusions We call T0 the minimal operator and T1 the maximal operator in H. Then, under the essential assumption that T ∗ 1 = T0 ⊆ T ∗ 0 = T1. def(T0) ≤ dim W, we construct one-parameter families {bT0} of minimal operators and associated maximal operators {bT1} in H ⊕ W, generated by T0 and T1 in H, satisfying the properties def (bT0) = def(T0) (bT0 ∈ {bT0}) 4 and LANCE L. LITTLEJOHN AND RICHARD WELLMAN (cid:16)bT1(cid:17)∗ = bT0 ⊆(cid:16)bT0(cid:17)∗ = bT1 (bT0 ∈ {bT0}, bT1 ∈ {bT1}). W. Both families {bT0} and {bT1} are parametrized by an arbitrary, fixed self-adjoint operator B : W → With the constructions of {bT0} and {bT1} in place, we then appeal to a general theory of complex symplectic algebra, with important applications and implications to boundary value problems in ordinary and partial differential equations, which was developed by Everitt and Markus in a series of remarkable papers [7, 8, 9, 10]. An important consequence of their theory is a generalized GKN theory - which we call GKN-EM theory after the contributions of Everitt and Markus - that we apply to characterize all self-adjoint extensions (respectively, restrictions) of bT0 ∈ {bT0} (respectively, of bT1 ∈ {bT1}). The contents of this paper are as follows. In Section 2, we briefly discuss the Stone-von Neumann theory of self-adjoint extensions of symmetric operators in a Hilbert space as well as the now classic GKN theory, including a statement of the GKN Theorem (Theorem 2.3). Section 3 deals with key complex symplectic geometric results developed by Everitt and Markus and culminates in the GKN- EM Theorem (Theorem 3.2). The families {bT0} and {bT1} of minimal and maximal operators in H⊕W, generated by T0 and T1 in the base space H, are developed in Section 4. Another key notion, the symplectic form [·,·]H⊕W in the extended space H⊕W , essential to our application of the GKN- EM theory, is defined in Section 4. Also, in this section, we apply Theorem 3.2 to characterize all self-adjoint extensions {bT} of bT0 ∈ {bT0}; see the summary theorem given in Theorem 4.4. Lastly, Section 5 deals with several examples to illustrate our results. These examples include another look at the Legendre type example where further light is shed on this particular example. Indeed, we show that, remarkably, continuity is a GKN-EM boundary condition. Notation: R, C and N will denote, respectively, the sets of real numbers, the complex numbers and the positive integers. All inner products in this paper will be denoted by h·,·i , properly subscripted indicating the particular underlying vector space. Ordered pairs in H ⊕ W will be written as (·,·) ; if dim W > 1, then an ordered pair in H ⊕ W will have the form (·, (·,·)). Our base space will be a complex Hilbert space (H,h·,·iH), our extension space will be finite-dimensional complex Hilbert space (W,h·,·iW ) and the extended space will be the direct sum space (H ⊕ W,h·,·iH⊕W ). Linear operators in the base space H will be denoted by T0, T1, T, etc. while operators in the extended space H ⊕ W will be hatted: bT0, bT1, bT , etc. The notation x has property P (x ∈ A) means that property P holds for all x in the set A. Lastly, the cardinality of a set A is denoted by card(A) whereas the dimension of subspace W of some vector space will be written dim W. 2. The von-Neumann Formulas and the GKN Theorem Standard references for topics discussed in this section are [1, 2, 15, 16, 17, 18, 19, 21]. Throughout this paper, the linear operator T0 : D(T0) ⊆ H → H will be an arbitrary closed, symmetric operator in H while T1 : D(T1) ⊆ H → H is a linear operator satisfying the operator inclusions (2.1) T ∗ 1 = T0 ⊆ T ∗ 0 := T1; in particular, we see that T0 and T1 are adjoints of each other. Because of the inclusions in (2.1), we call T0 the minimal operator and T1 the maximal operator. Specific reasons for this notation SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 5 will be discussed below in this section (see also Remark 2.1). Notice that if T0 has a self-adjoint extension T in H, then so T necessarily has the same form as T1; that is, T0 ⊆ T = T ∗ ⊆ T ∗ 0 = T1, T x = T1x (x ∈ D(T )). The general theory of self-adjoint extensions of the minimal operator T0 (equivalently, self-adjoint restrictions of the maximal operator T1) in a Hilbert space - called the Stone-von Neumann theory - is discussed in depth in [2, Chapter XII, Section 4]. Of central importance in this theory are two particular subspaces X± of D(T1), defined by X± := {x ∈ D(T1) T1x = ±ix} , where i = √−1. These spaces are called the positive and negative deficiency spaces of T0. The first von Neumann formula decomposes the maximal domain D(T1) into linearly independent subman- ifolds: Theorem 2.1 (The First von Neumann Formula). D(T1) = D(T0) + X+ + X−. In fact, the sum in this formula is actually an orthogonal direct sum. Indeed, under the graph inner product (2.2) and associated norm (2.3) hx, yi∗ H := hx, yiH + hT1x, T1yiH (x, y ∈ D(T1)) (kxk∗ H )2 = kxk2 H + kT1xk2 H ≥ kxk2 H , D(T1) is a Hilbert space and, with this inner product, D(T0), X+ and X− are closed, orthogonal subspaces of D(T1); see [2, Chapter XII]. Notice that if x ∈ D(T1) and x = x0 + x+ + x−, where x0 ∈ D(T0) and x± ∈ X±, then (2.4) (kxk∗ H)2 =(cid:16)(cid:13)(cid:13)x0(cid:13)(cid:13)∗ H(cid:17)2 +(cid:0)(cid:13)(cid:13)x+(cid:13)(cid:13)∗ H(cid:1)2 +(cid:0)(cid:13)(cid:13)x−(cid:13)(cid:13)∗ H(cid:1)2 . The dimensions of X±, denoted by dim(X±), are called the positive and negative deficiency indices of T0. A key result in the Stone-von Neumann theory is that the equality of these deficiency indices is equivalent to the existence of self-adjoint extensions T of T0 in H. Moreover, if dim(X+) = dim(X−) = 0, T0 = T1 is self-adjoint and is, in fact, the only self-adjoint extension of T0 in H. In the case that dim(X+) = dim(X−), we refer to this common value as the deficiency index and denote it by def(T0). In addition to requiring the equality of these deficiency indices for the entirety of this paper, we assume the deficiency indices are also finite. Thus, another key assumption in this paper is: Condition 2.1. 1 ≤ def(T0):= dim(X+) = dim(X−) < ∞. The second von Neumann formula gives a description of the domain of any self-adjoint extension T of T0 in H: 6 LANCE L. LITTLEJOHN AND RICHARD WELLMAN Theorem 2.2 (The Second von Neumann Formula). Let T : D(T ) ⊆ H → H be a self-adjoint extension of T0. Then there exists an isometric isomorphism V : X+ → X− from the positive deficiency space X+ onto the negative deficiency space X− such that (2.5) T x = T1x (2.6) Conversely, if T and its domain D(T ) are defined through (2.5) and (2.6) for some isometric isomorphism V : X+ → X− , then T is a self-adjoint extension of T0. D(T ) = {x + x+ + V x+ x ∈ D(T0), x+ ∈ X+}. The Glazman-Krein-Naimark (GKN) theory is both a refinement and an application of the Stone-von Neumann theory to self-adjoint operator extensions of ordinary differential expressions. Excellent expositions of this theory can be found in Akhiezer and Glazman [1, Volume II, Chapter 8] and Naimark [15, Part II, Chapter V]. To describe this theory we assume, for the sake of simplicity, that ℓ[·] is a real, 2n-th order Lagrangian symmetrizable differential expression of the form (2.7) ℓ[y](u) = 1 (−1)j(cid:16)qj(u)y(j)(u)(cid:17)(j) nXj=0 w(u) (u ∈ I), where each coefficient qj : I → R in (2.7) is j-times continuously differentiable on I (noting, however, that general 'quasi-differentiable' conditions can be placed on these coefficients; see also [20]). The setting for the study of ℓ[·] is the Hilbert space L2(I; w) = {f : I → C f is Lebesgue measurable and ZI f2 wdu < ∞} endowed with the standard inner product hf, giw =Z b a f (u)g(u)w(u)du (f, g ∈ L2(I; w)). Here I ⊆ R is an open interval and w is a positive (a.e.) Lebesgue measurable function on I. The maximal operator L1 : D(L1) ⊆ L2(I; w) → L2(I; w), generated by ℓ[·], is defined to be L1f = ℓ[f ] f ∈ D(L1) = {f : I → C f (j) ∈ ACloc(I) (j = 0, 1, . . . , 2n − 1); f, ℓ[f ] ∈ L2(I; w)}. In this setting, the term 'maximal' is appropriate; indeed, D(L1) - which is called the maximal domain - is the largest subspace of L2(I; w) for which the expression ℓ[·] acts on and maps into L2(I; w). It is clear that L1 is a densely defined operator. We denote the adjoint of L1 by L0; it is natural then to call L0 the minimal operator generated by ℓ[·]. The GKN theory shows that, in fact, L1 and L0 are adjoint to each other and L0 is a closed symmetric operator in L2(I; w). More explicitly, L∗ 1 = L0 and L0 = L0 ⊆ L∗ 0 = L1. Remark 2.1. The operators T0 and T1, defined earlier, are analogous to the minimal operator L0 and maximal operator L1, respectively. Because of this, we call T0 and T1, respectively, the minimal and maximal operators even though, in the general situation, the terms maximal and minimal may not seem as appropriate as they do in the GKN theory. Likewise, we shall call their respective domains the minimal domain D(T0) and the maximal domain D(T1). The domain D(L0) of the minimal operator is given explicitly by (2.8) D(L0) = {f ∈ D(L1) [f, g]wb a = 0 for all g ∈ D(L1)}, SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 7 where [·,·]wb (2.9) a is the skew-symmetric bilinear form obtained from the classic Green's formula hL1f, giw − hf, L1giw = [f, g]wb a (f, g ∈ D(L1)). Moreover, we note that Condition (2.1) is automatically satisfied in this setting. deficiency indices of L0 are equal since ℓ has real coefficients and thus Indeed, the Moreover, in this case, ℓ[f ] = if ⇐⇒ ℓ[f ] = ℓ[f ] = −if . 0 ≤ def(L0) ≤ 2n. We are now in position to state the GKN Theorem. Notice that this theorem provides a 'recipe' for constructing all self-adjoint extensions of the minimal operator L0 in L2(I; w) by specifying certain restrictions (boundary conditions), using the bilinear form [·,·]w, on the maximal domain D(L1). We emphasize, however, that the original GKN Theorem is valid only for the minimal operator L0 associated with a real Lagrangian symmetrizable differential expressions of even order in the specific Hilbert space L2(I; w). Compare the statement of the GKN Theorem below with that of the GKN-EM Theorem (Theorem 3.2) at the end of the next section. Theorem 2.3 (The GKN Theorem). Suppose L0 and L1 are, respectively, the minimal and maximal operators in L2(I; w), generated by the differential expression ℓ[·], given in (2.7). In addition, let m = def(L0) so 0 ≤ m ≤ 2n. (i) Suppose the set {gj j = 1, . . . , m} ⊆ D(L1) satisfies the two conditions (a) mXj=1 αjgj ∈ D(L0) =⇒ αj = 0 (j = 1, . . . , m) and (b) [gi, gj]wb a = 0 (i, j = 1, . . . , def(L0)). Define the operator S : D(S) ⊆ L2(I; w) → L2(I; w) by Sf = L1f (2.10) (2.11) (2.12) (2.13) f ∈ D(S) = {f ∈ D(L1) [f, gj]wb (j = 1, . . . , def(L0))}. Then S is a self-adjoint extension of the minimal operator L0 in L2(I; w). a = 0 (ii) Conversely, if S : D(S) ⊆ L2(I; w) → L2(I; w) is a self-adjoint extension of the minimal operator L0 in L2(I; w), then there exists a set {gj j = 1, . . . , m} ⊆ D(L1) satisfying the conditions (2.10) and (2.11) such that S is given explicitly by (2.12) and (2.13). Remark 2.2. A collection of vectors {gj j = 1, . . . , m} ⊆ D(L1) satisfying condition (2.10) are said to be linearly independent modulo D(T0) while those that satisfy (2.11) are said to satisfy Glazman symmetry conditions. Further light, as well as a generalization, into these concepts be made in the next section. Remark 2.3. Each of the conditions given in (2.13), is called a 'boundary condition'. In the case that def(L0) = 0, then L1 (= L0) is the only self-adjoint extension of L0 and, in this case, there are no boundary conditions. [f, gj]wb a = 0 (j = 1, . . . , def(L0)), 8 LANCE L. LITTLEJOHN AND RICHARD WELLMAN The GKN-EM Theorem, which we discuss in the next section in Theorem 3.2, is a generalization of the GKN Theorem but, remarkably, is valid in an arbitrary Hilbert space for an arbitrary closed symmetric operator with equal and finite deficiency indices. This theorem is a highlight application of the general complex symplectic theory developed by Everitt and Markus. 3. Complex Symplectic Geometry and a Generalization of the GKN Theorem In a series of papers [7, 8, 9, 10], Everitt and Markus developed an extensive theory of complex symplectic geometry with applications to linear ordinary and partial differential equations. Their work was motivated by their interest in boundary value problems. In this section, we report on their results that pertain to this manuscript. A highlight application of their investigations is an important, and remarkable, generalization of Theorem 2.3; see Theorems 3.1 and 3.2 below. This generalization is key to the results we establish in the next section. Definition 3.1. A complex symplectic space S is a complex vector space together with a conjugate bilinear (sesquilinear) complex-valued function [· : ·] : S×S→ C satisfying the properties (i) [c1x1 + c2x2 : y] = c1[x1 : y] + c2[x2 : y], (ii) [x : y] = −[y : x], (iii) [x : S] = 0 =⇒ x = 0 (non-degenerate condition). We call [· : ·] a (non-degenerate) symplectic form. Complex symplectic spaces are non-trivial generalizations (not merely complexifications) of clas- sical real symplectic spaces of Lagrangian and Hamiltonian mechanics (see [11]). Indeed, complex symplectic spaces have a much wider scope and admit new applications. For example, whereas real symplectic spaces cannot be odd dimensional, it is the case that, for every n ∈ N, there exists complex symplectic spaces of dimension n. Along with their real symplectic counterparts, complex symplectic spaces support the notion of Lagrangian subspaces (see [10] equation (1.10)). Definition 3.2. A subspace L of a complex symplectic space S is called Lagrangian if [L : L] = 0; that is to say, when A Lagrangian L ⊆ S is called a complete Lagrangian when x ∈ S and [x : L] = 0 ⇒ x ∈ L. [x : y] = 0 (x, y ∈ L). We can characterize complete Lagrangian subspaces as follows. This characterization is key for later results. Lemma 3.1. A Lagrangian subspace L ⊆ S is a complete Lagrangian if and only if (3.1) L = {x ∈ S [x : y] = 0 (y ∈ L)}. Proof. Suppose L is a complete Lagrangian subspace of S. By definition of complete, it is clear that {x ∈ S [x : y] = 0 (y ∈ L)} ⊆ L. On the other hand, if x ∈ L then [x : y] = 0 for all y ∈ L since L is Lagrangian. Hence L ⊆ {x ∈ S [x : y] = 0 (y ∈ L)}. Conversely, if L is Lagrangian and given by (3.1), then it is clear that L is a complete. (cid:3) An essential step in the work of Everitt and Markus is a natural generalization of the skew- symmetric bilinear form [·,·]wb Definition 3.3. [x, y]H := hT1x, yiH − hx, T1yiH for x, y ∈ D(T1) . a given by Green's formula (2.9). SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 9 Following the work of Everitt and Markus, we will see below that [·,·]H can be identified with a degenerate symplectic form. We also note that [·,·]H coincides with [·,·]wb a in the case T1 is the maximal differential operator, generated by ℓ[·] (see (2.7)), in the weighted Hilbert space L2(I; w). As shown in [10], the quotient space S′ := D(T1)/D(T0), Notice that, from Theorem 2.1 and Condition 2.1, S (3.2) with zero element 0 = D(T0), is a complex symplectic space when endowed with the form [·,·]H ; we outline the specific details below. has dimension 2def(T0). Indeed one may view S as an isomorphic copy of the orthogonal sum of the deficiency spaces X± of T0. Everitt and Markus call the space S are, of course, cosets x = {x + D(T0)} (x ∈ D(T1)). In this case, we call the vector x a representative vector of the coset {x + D(T0)}. the boundary space of T0. The elements of S ′ ′ ′ ′ We now consider the natural projection φ : D(T1) → S ′ defined by φ(x) = {x + D(T0)} (x ∈ D(T1)) . The following proposition makes clear the connection between a basis of a subspace of S and the notion of linear independence modulo D(T0) which we first encountered in Theorem 2.3 and Remark 2.2. Lemma 3.2. A collection of cosets {φtj}d dimension d of the boundary space S j=1 ⊆ D(T1), is a basis for a subspace of j=1, where {tj}d if and only if the representative vectors {tj}d j=1 satisfy ′ ′ dXj=1 αjtj ∈ D(T0) =⇒ αj = 0 (j = 1, 2, . . . , d); that is to say, {tj}d Proof. The equation Pd j=1 is linearly independent modulo D(T0). j=1 αjφtj = 0 is equivalent to Pd j=1 αjtj ∈ D(T0). The following lemma generalizes the characterization of the domain of the minimal operator; see (cid:3) (2.8). Lemma 3.3. D(T0) = {x ∈ D(T1) [x, y]H = 0 (y ∈ D(T1))} . Proof. Fix x ∈ D(T1) and suppose [x, y]H = 0 (y ∈ D(T1)). Since [x, y]H = −[y, x]H, we see that hT1y, xiH = hy, T1xiH so x ∈ D(T ∗ x ∈ D(T0). Since T ∗ 0 = T1 and T0x = T1x, we see that 1 ) = D(T0). Conversely, let hT1x, yiH = hT0x, yiH = hx, T1yiH (y ∈ D(T1)); that is, for each y ∈ D(T1), [x, y]H = hT1x, yiH − hx, T1yiH = 0. (cid:3) This result allows the boundary space S ′ to be equipped with a complex symplectic form. Definition 3.4 (Boundary Space Symplectic Form). (3.3) [φx : φy]S′ := [x, y]H (x, y ∈ D(T1)). 10 LANCE L. LITTLEJOHN AND RICHARD WELLMAN Lemma 3.3 assures Definition 3.4 above is independent of the choice of representative vectors. Moreover, Lemma 3.3 establishes the non-degeneracy property of Definition 3.1. From the definition of a Lagrangian subspace, the following extension of Lemma 3.2 is clear. Proposition 3.1. A collection of cosets {φtj}d subspace of the boundary space S′ if and only if the representative vectors {tj}d j=1 form a basis for a d-dimensional Lagrangian j=1 satisfy (a) (3.4) (b) (3.5) and dXj=1 αjtj ∈ D(T0) =⇒ αj = 0 (j = 1, . . . , d); [ti, tj]H = 0 (i, j = 1, . . . , d). Notice that the properties (3.4) and (3.5) are identical to those conditions discussed in Theorem 2.3. Because of their importance in the special case when d = def (T0) , we incorporate these two properties into the following definition. Definition 3.5. A collection of vectors {tj j = 1, . . . , def (T0)} ⊆ D(T1) is called a GKN set for T0 if (i) the set {tj j = 1, . . . , def (T0)} is linearly independent modulo the minimal domain D(T0); that is to say (3.6) and if def(T0)Xj=1 αjtj ∈ D(T0) then αj = 0 for j = 1, . . . , def (T0) ; (ii) the set {tj j = 1, . . . , def (T0)} satisfies the symmetry conditions [ti, tj]H = 0 (3.7) Remark 3.1. Observe that if G ⊆ D(T1) is a GKN set for T0, then any non-empty, proper subset P ⊆ G is linearly independent modulo D(T0) and satisfies the symmetry conditions in (3.7). We refer to P as a partial GKN set. However, we note that the only partial GKN sets P that we use in this manuscript are those which satisfy card(P ) = dim(W ) ≤ def(T0), where W is a complex finite-dimensional extension space; see Condition 4.1 in Section 4. (i, j = 1, . . . , def (T0)). We now turn our attention to characterizing complete Lagrangians. A key result of Everitt and Markus in this setting is that not only do complete Lagrangians L exist (see [10, Equations (1.54) and (1.61)]) but their dimensions are precisely that of the deficiency index; that is, (3.8) dim L = def(T0) (see [10, Equation (3.9)]). Moreover, Lemma 3.4. With def(T0) < ∞, a Lagrangian subspace L ⊆ S′ is complete if and only if each of the two conditions hold: (i) dim L = def(T0); (ii) L = {φx [φx : φtj]S′ = 0 (j = 1, 2, . . . ,def(T0))} for some GKN set {tj j = 1, 2, . . . ,def(T0)}. Moreover, in this case, (3.9) φ−1L = {x ∈ D(T1) [x, tj]H = 0 (j = 1, 2, . . . , def(T0))}. SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 11 Proof. Suppose L ⊆ S 3.1, ′ (3.10) is complete. Then, by (3.8), dim L = def(T0) establishing (i). By Lemma L = {φx [φx : φy]S′ = 0 (φy ∈ L)}. Let {φtj j = 1, 2, . . . ,def(T0)} be a basis for L. Then, by Proposition 3.1, {tj j = 1, 2, . . . ,def(T0)} is a GKN set for T0. It follows from (3.10), that (3.11) L = {φx [φx : φtj]S′ = 0 (j = 1, 2, . . . , def(T0))}, proving (ii). Lastly, using the identification in (3.3) along with the identity in (3.11), (3.9) is clear. Conversely, suppose (i) and (ii) hold. It is straightforward to show that L is a subspace of S . Clearly (3.9) follows from (ii). Moreover, since {tj j = 1, 2, . . . ,def(T0)} is a GKN set for T0, we see that ′ [φti : φtj]S′ = [ti, tj]H = 0. It follows by taking linear combinations that L is Lagrangian. Finally, from (3.8), we see that L is complete. (cid:3) The authors in [10, Theorem 1.14 and Remark 1.15] establish the following characterization of self-adjoint extensions of T0 in terms of complete Lagrangian subspaces L of S ′ . Theorem 3.1 (The Finite-Dimensional GKN-EM Theorem). Let T0 and T1 be, respectively, the minimal and maximal operators as defined in Section 2 and let S be given by (3.2). There exists a one-to-one correspondence between the set {T} of all self-adjoint extensions of T0 and the set {L} of all complete Lagrangians L ⊆ S . More specifically, ′ ′ (a) if T is a self-adjoint operator with T0 ⊆ T ⊆ T1, then x ∈ D(T )} L := {φx ∈ S ′ (3.12) is a complete Lagrangian subspace of S ′ (b) If L is a complete Lagrangian subspace of S of dimension def(T0). Moreover, φ−1L = D(T ). , then L has dimension def(T0). Define ′ D(T ) = {x ∈ D(T1) φx ∈ L}. Then T : D(T ) ⊂ H → H given by T x = T1x x ∈ D(T ) is a self-adjoint operator satisfying T0 ⊆ T ⊆ T1. Moreover, φ−1 L = D(T ). Combining Theorem 3.1 with Lemmas 3.1 and 3.4, we are now in position to state and prove an important consequence of Theorem 3.1 which, for our purposes, is key to the results developed in the next section and in the examples of Section 5. We note that the next theorem is an exact generalization of the GKN theorem stated in Theorem 2.3. Theorem 3.2 (The Finite-Dimensional Symplectic GKN-EM Theorem). Suppose T0 and T1 are linear operators satisfying the conditions set forth in Section 2 and [·,·]H is the symplectic form defined in Definition 3.3. In particular, we assume T0 has equal and finite deficiency indices denoted by def (T0). 12 LANCE L. LITTLEJOHN AND RICHARD WELLMAN (i) If the operator T : D(T ) ⊆ H → H is self-adjoint and satisfies T0 ⊆ T ⊆ T1 then there exists a GKN set {tj j = 1, . . . , def (T0)} ⊆ D(T1) of T0 such that (3.13) D(T ) = {x ∈ D(T1) [x, tj]H = 0 (j = 1, . . . , def (T0))}. (ii) If {tj j = 1, . . . , def (T0)} ⊆ D(T1) is a GKN set for T0 then the operator T : D(T ) ⊆ H → H given by (3.14) (3.15) T x = T1x x ∈ D(T ) = {x ∈ D(T1) [x, tj]H = 0 (j = 1, . . . , def(T0))} is self-adjoint and satisfies T0 ⊆ T ⊆ T1. Proof. (i) Suppose T : D(T ) ⊆ H → H is self-adjoint and satisfies T0 ⊆ T ⊆ T1. By Theorem 3.1, (3.16) ′ L = {φx ∈ S ′ x ∈ D(T )} is a complete Lagrangian subspace of S (3.17) Moreover, by Lemma 3.4, there exists a GKN set {tj j = 1, 2, . . . ,def(T0)} for T0 such that of dimension def (T0) from which it follows that φ−1L = D(T ). L = {φx [φx : φtj]S′ = 0 (j = 1, 2, . . . , def(T0))}. and (3.18) φ−1L = {x ∈ D(T1) [x, tj]H = 0 (j = 1, 2, . . . , def(T0))}. Comparing (3.17) with (3.18), we obtain (3.13). (ii) Suppose {tj j = 1, 2, . . . ,def(T0)} is a GKN set for T0. Let (3.19) L = {φx [φx : φtj]S′ = 0 (j = 1, 2, . . . , def(T0))}. By Lemma 3.4, L is a complete Lagrangian subspace of S (3.14) and (3.15). Then, from (3.15) and (3.19), we see that L = {φx x ∈ D(T )} so that ′ of dimension def(T0). Define T as in D(T ) = φ−1L = {x ∈ D(T1) φx ∈ L}. By Theorem 3.1, T is self-adjoint and T0 ⊆ T ⊆ T1. Remark 3.2. In the case that H = L2(I; w) and T0 and T1 are, respectively, the minimal and maximal operators L0 and L1, generated by the ordinary differential expression (2.7), Theorem 3.2 is identical to the classical GKN theorem given in Theorem 2.3. Again, it is remarkable that the GKN theorem extends verbatim to a general Hilbert space with an arbitrary closed symmetric operator having equal deficiency indices. As in the classical GKN setting, we also call the conditions (cid:3) [x, tj]H = 0 for j = 1, . . . , def (T0) 'boundary conditions'. Lastly, we note that, as in Remark 2.3, if def(T0) = 0, there are no such boundary conditions and, in this case, the only self-adjoint extension of T0 is the maximal operator T1 (= T0). SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 13 Remark 3.3. Everitt and Markus discuss other important applications of their results to ordinary and partial differential operators. We refer the reader to Sections 2.1, 2.2 and 4.2 in [10]. They outline the argument given above in Theorem 3.2 for Sturm-Liouville problems (see [10, Section 2, equations (2.23), (2.24), and (2.25)]) as well as for general Shin-Zettl quasi-differential operators (see [10, Section 4; in particular equations (4.57)–(4.61)]). The authors are certain that the most general situation (when T0 has finite and equal deficiency indices), which we prove in Theorem 3.2, was known to Everitt and Markus but we cannot find an exact reference in their joint work. 4. Maximal and Minimal Operators in H ⊕ W deficiency indices def(T0) and adjoint operator T1 satisfying T ∗ by, respectively, the minimal operator T0 and the maximal operator T1 in the base space H. We We remind the reader that T0 : D(T0) ⊆ H → H is a closed, symmetric operator with equal, finite 0 = T1. In this section, we identify a family of minimal operators bT0 : D(bT0) ⊆ H ⊕ W → H ⊕ W and an associated family of maximal operators bT1 : D(bT1) ⊆ H ⊕ W → H ⊕ W in the extended space H ⊕ W generated show that each bT0 is a closed, symmetric operator in H ⊕ W with equal deficiency indices and def(bT0) = def(T0). Moreover, the operators bT0 and bT1 are adjoints of each other just as in the A fundamental assumption in our development of the maximal and minimal operators in H ⊕ W classical case with T0 and T1. 1 = T0 ⊆ T ∗ is the following dimensionality requirement for the extension space: Condition 4.1. dim(W ) 6 def (T0) . Fix a partial GKN set (4.1) {tj j = 1, . . . , dim(W )} ⊆ D(T1); recall, from Remark 3.1 and Condition 4.1, that this set exists and satisfies the two conditions (4.2) and (4.3) dim(W )Xj=1 αjtj ∈ D(T0) =⇒ αj = 0 (j = 1, . . . , dim(W )) [ti, tj]H = 0 (i, j = 1, . . . , dim(W )). It is clear that the maximal operator T1 in the base space is symmetric on ∆0 := D(T0) + span{tj j = 1, . . . , dim(W ) ⊆ D(T1). (4.4) Now let be an orthonormal basis of W and define Ψ : ∆0 → W by {ξj j = 1, . . . , dim(W )} ⊆ W Ψ (tj) = ξj Ψ (s) = 0 (4.5) and extend Ψ to ∆0; that is to say Ψ s + dim(W )Xj=1 (j = 1, . . . , dim(W )) (s ∈ D(T0)). αjtj  = dim(W )Xj=1 αjξj. Note the key fact that Ψ maps the partial GKN set {tj j = 1, . . . , dim(W )} onto W. 14 LANCE L. LITTLEJOHN AND RICHARD WELLMAN generated by T0. Lastly, fix an arbitrary self-adjoint operator B : W → W in the extension space W . With these definitions and conditions in place, we are now in position to define a minimal operator cT0 in H⊕W Definition 4.1. The minimal operator bT0 : D(cid:16)bT0(cid:17) ⊆ H ⊕ W → H ⊕ W is defined to be bT0(x, a) = (T1x, Ba) (x, a) ∈ D(bT0) := {(x, Ψx) x ∈ ∆0} . in fact, a densely defined operator which is both closed and symmetric. Moreover, in Theorem 4.2, At this point, it is unclear why we call bT0 the minimal operator generated by T0; we will justify this terminology in Remark 4.1. In Theorem 4.1 below we show that the minimal operator bT0 is, where it is shown that (bT0)∗ = bT1, we introduce the important linear transformation Ω : D(T1) → W defined by (4.8) Ωx := dim(W )Xj=1 [x, tj]Hξj (x ∈ D(T1)). (4.9) Observe, by definition of the partial GKN set {tj j = 1, . . . , dim W} and Lemma 3.3, that With this transformation Ω, we are now ready to introduce the maximal operator bT1. Definition 4.2. The maximal operator bT1 : D(cid:16)bT1(cid:17) ⊆ H ⊕ W → H ⊕ W is defined by Ωx = 0 (x ∈ ∆0). (4.6) (4.7) (4.10) (4.11) that is (4.12) bT1(x, a) = (T1x, Ba − Ωx) (x, a) ∈ D(bT1) := {(x, a) x ∈ D(T1), a ∈ W} . Note that if (x, Ψx) ∈ D(bT0), then (x, Ψx) ∈ D(bT1). Moreover, in this case, Ωx = 0 by (4.9) so bT1(x, Ψx) = (T1x, BΨx − Ωx) = (T1x, BΨx) = bT0(x, Ψx); Remark 4.1. The term 'maximal' is appropriate; indeed, observe that D(bT1) is the largest linear manifold in H ⊕ W on which an operator representation of T1 makes sense. Moreover, once we establish the fact that (bT0)∗ = bT1, we see that the term 'minimal' is appropriate for the operator bT0. Proposition 4.1. The extension J : D(J) ⊆ H → H of the minimal operator T0, defined by bT0 ⊆ bT1. Jx := T1x x ∈ D(J) := ∆0, is a closed symmetric operator. Proof. Since T0 is densely defined and D(T0) ⊆ ∆0, it is clear that D(J) is dense in H. Now, from Lemma 3.3 and (3.7), we see that Hence, from Definition 3.3, [x, y]H = 0 (x, y ∈ ∆0 ⊆ D(T1)). 0 = [x, y]H = hT1x, yiH − hx, T1yiH = hJx, yiH − hx, JyiH (x, y ∈ D(J)), SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 15 establishing that J is symmetric in H. To show that J is closed, suppose first that dim(∆0) (mod(D (T0))) = 1; that is (4.13) D(J) = D(T0) + span{t1}, where t1 ∈ D(T1)(cid:31)D(T0). Consider a sequence {xj} ⊆ ∆0 and vectors x, y ∈ H such that xj → x and T1xj = Jxj → y where the convergence of both sequences is in H. Of course, we need to show (4.14) x ∈ D(J) and (4.15) Jx = y. Since D(J) ⊆ D(T1) and T1 is closed, we know that x ∈ D(T1) and T1x = y. Hence (4.15) will be established once we show (4.14). Now, by (4.13) and Theorem 2.1, we can write (4.16) xj = vj + αjt1 where t1 = t0 x± ∈ X±. Since xj → x and T1xj → T1x, we see that 1 and vj, t0 1 + t− 1 + t+ 1 + vj) + αjt+ 1 + αjt− 1 (j ∈ N) = (αjt0 1 ∈ D(T0), t± 1 ∈ X± and x = x0 + x+ + x−, where x0 ∈ D(T0) and kxj − xk∗ H → 0 as j → ∞, where k·k∗ H is the graph norm given in (2.3). Since (cid:0)kxj − xk∗ H(cid:1)2 =(cid:16)(cid:13)(cid:13)αjt0 H(cid:17)2 1 + vj − x0(cid:13)(cid:13)∗ we see, from (2.3) and (2.4), that (4.17) (4.18) (4.19) 1 − x+(cid:13)(cid:13)∗ H(cid:1)2 +(cid:0)(cid:13)(cid:13)αjt− 1 − x−(cid:13)(cid:13)∗ H(cid:1)2 , +(cid:0)(cid:13)(cid:13)αjt+ 1 + vj → x0 in H 1 → x+ in H αjt+ αjt− 1 → x− in H. αjt0 1 and t− Since t+ 1 ∈ D(T0), contradicting our choice of t1), we see from either (4.18) or (4.19) that there exists α ∈ C with αj → α. It follows that αjt1 → αt1 in H. Then, from (4.16) and (4.17), we see that 1 both cannot be zero (otherwise, t1 = t0 xj = vj + αjt1 Hence we see that x = (x0 − αt0 follows by induction on dim(W ). 1 1 + vj + αjt1 − αjt0 = αjt0 → (x0 − αt0 1) + αt1 ∈ D(J), as required. The general proof of this proposition 1) + αt1 ∈ D(T0) + span{t1}. (cid:3) Remark 4.2. Proposition 4.1 shows that, on ∆0, the maximal operator T1 is a closed, symmetric operator. Of course, T1 is not, in general, symmetric on D(T1). Theorem 4.1. The operator bT0 is a closed, densely defined symmetric operator in H ⊕ W . 16 LANCE L. LITTLEJOHN AND RICHARD WELLMAN Proof. (i) bT0 is Hermitian: Let (x, Ψx), (y, Ψy) ∈ D(bT0). Then, by Proposition 4.1 and the fact that B is symmetric in W , we see that DbT0(x, Ψx), (y, Ψy)EH⊕W = h(T1x, BΨx), (y, Ψy)iH⊕W = hT1x, yiH + hBΨx, ΨyiW = hx, T1yiH + hΨx, BΨyiW =D(x, Ψx), bT0(y, Ψy)EH⊕W . Hence bT0 is Hermitian. (ii) D(bT0) is dense in H ⊕ W : Since D(T0) is dense in H and Ψ is surjective, it is clear that D(bT0) is dense in H ⊕ W. (iii) bT0 is symmetric in H ⊕ W : (iv) bT0 is closed in H ⊕ W : Suppose that {(xn, Ψxn)} ⊆ D(bT0) is such that This follows immediately from (i) and (ii). (4.20) These conditions in (4.20) and (4.21) are equivalent to We need to show that (x, a) ∈ D(bT0) and bT0(x, a) = (y, b); that is to say, we need to prove: (4.24) (xn, Ψxn) → (x, a) in H ⊕ W bT0(xn, Ψxn) → (y, b) in H ⊕ W. xn → x and T1xn → y in H Ψxn → a and BΨxn → b in W. x ∈ ∆0 T1x = y Ψx = a and (4.21) (4.22) and (4.23) (4.25) (4.26) and (4.27) Since {xn} ⊆ ∆0, we see that T1xn = Jxn so, by Proposition 4.1, Ba = b. x ∈ ∆0 and Jx = T1x = y, establishing (4.24) and (4.25). For the remainder of this proof, write (4.28) and (4.29) x = x0 + t xn = xn,0 + tn, SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 17 where x0, xn,0 ∈ D(T0), and t = dim WXj=1 αjtj, tn = dim WXj=1 αn,jtj. From (4.29) and the definition of Ψ, we see that Ψxn = Ψ(xn,0 + tn) = Ψtn = dim WXj=1 αn,jξj so that, from (4.23), It follows that (4.30) so that T1tn → T1bt. Notice that (4.31) αn,j = (Ψxn, ξj)W → ha, ξjiW . dim WXj=1 tn →bt := ha, ξjiW tj Ψtn → Ψbt = dim WXj=1 ha, ξjiW ξj = a. From (4.22), (4.29) and (4.30), we deduce that xn,0 = xn − tn → x −bt = x0 + t −bt in H T0xn,0 = T1xn − T1tn → y − T1bt in H. Since T0 is closed, we see that x0 + t−bt ∈ D(T0) and, in particular, that t−bt ∈ D(T0). By definition of the partial GKN set {tj 1 ≤ j ≤ dim W}, we must have (4.32) Combining (4.31) and (4.32), we obtain establishing (4.26). Finally, Ψxn = Ψtn → Ψbt = Ψt = Ψ(x − x0) = Ψx = a; t =bt. BΨxn → BΨx = Ba This brings us to the proof of the fundamental relation between the maximal and minimal so, by (4.23), Ba = b which proves (4.27). This completes the proof that bT0 is closed. operators bT1 and bT0. Theorem 4.2. (bT0)∗ = bT1. (cid:3) 18 LANCE L. LITTLEJOHN AND RICHARD WELLMAN Proof. For (x, a) , (y, b) ∈ D(bT1), a calculation shows that DbT1(x, a), (y, b)EH⊕W −D(x, a), bT1(y, b)EH⊕W (4.33) (see Definition 4.50). Notice that when y = tk and b = ξk, we obtain = [x, y]H − hΩx, biW + ha, ΩyiW . (4.34) hΩx, ξjiW = dim WXk=1 h[x, tk]Hξk, ξjiW = [x, tj]H since {ξj j = 1, . . . , dim W} is an orthonormal basis of W. Suppose now that (y, b) ∈ D(bT0) so y ∈ ∆0 and b = Ψy. Then where y0 ∈ D(T0), y = y0 +et, dim WXj=1 et := αjtj (4.35) (4.37) (4.38) (4.39) (4.40) (4.41) and dim WXj=1 By (4.12), bT0(y, Ψy) = bT1(y, Ψy) so, from (4.33), we obtain b = Ψy = Ψet = αjξj. DbT1(x, a), (y, Ψy)EH⊕W −D(x, a), bT0(y, Ψy)EH⊕W = [x, y0]H + [x,et]H − hΩx, Ψy0iW −(cid:10)Ωx, Ψet(cid:11)W + ha, ΩyiW . = [x, y]H − hΩx, biW + ha, ΩyiW We now deal with each of the five terms in (4.35). First, from Lemma 3.3, (4.36) From (4.34), we see that [x, y0]H = 0. [x,et]H −(cid:10)Ωx, Ψet(cid:11)W = dim WXj=1 αj{[x, tj]H − hΩx, ξjiW} = 0. From (4.5), Ψy0 = 0 so Likewise, from (4.9), we see that Ωy = 0 so hΩx, Ψy0iW = 0. ha, ΩyiW . =D(x, a), bT0(y, Ψy)EH⊕W bT1 ⊆ (bT0)∗. Together, (4.36), (4.37), (4.38) and (4.39) show that DbT1(x, a), (y, Ψy)EH⊕W and, hence, we obtain ((x, a) ∈ D(bT1), (y, Ψy) ∈ D(bT0)) To show (bT0)∗ ⊆ bT1, let (x, a) ∈ D((bT0)∗) and set (x∗, a∗) = (bT0)∗(x, a). Then for (y, Ψy) ∈ D(bT0), h(x∗, a∗), (y, Ψy)iH⊕W =D(bT0)∗(x, a), (y, Ψy)EH⊕W =D(x, a), bT0(y, Ψy)EH⊕W (4.42) SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 19 since bT0 is closed. Written out, the identity in (4.42) gives (4.43) In particular, if y ∈ D(T0), then (4.43) reduces to hx∗, yiH + ha∗, ΨyiW = hx, T0yiH + ha, BΨyiW . Thus x ∈ D(T ∗ (4.44) 0 ) = D(T1) and def hx∗, yiH = hx, T0yiH . T1x = x∗. Substituting (4.44) into (4.43) and recalling that B is symmetric in W yields ha∗, ΨyiW = hx, T0yiH + ha, BΨyiW − hT1x, yiH (4.45) In particular, let y = tk so Ψy = ξk. From (4.34), we see that [x, tk]H = hΩx, ξkiW . Hence, we find that 4.45) becomes = −[x, y]H + hBa, ΨyiW . ha∗, ξkiW = −hΩx, ξkiW + hBa, ξkiW (k = 1, 2, . . . , dim W ). (4.46) Since {ξk k = 1, . . . , dim W} is a basis for W, we can conclude from (4.46) that (4.47) a∗ = Ba − Ωx. Consequently, from (4.44) and (4.47), we see that so (x∗, a∗) = (T1x, Ba − Ωx) = bT1(x, a) (cid:3) (4.48) Together Theorem 4.1 and Theorem 4.2 establish the following fundamental operator relationship (bT0)∗ ⊆ bT1. Combining (4.41) and (4.48), we obtain (bT0)∗ = bT1. between bT0 and bT1. Theorem 4.3. bT0 = bT0 ⊆ (bT0)∗ = bT1. Consequently we may apply the Stone-von Neumann theory to the minimal operator bT0. Accord- ingly we define the positive and negative deficiency spaces associated with bT0 in H ⊕ W Definition 4.3 (Deficiency Spaces in the Extended Space H ⊕ W ). Y± := {(x, a) ∈ D(bT1) bT1(x, a) = ±i(x, a)}. Remarkably, as we shall see in the next result, the deficiency spaces Y± ⊆ H ⊕ W and X± ⊆ H are isomorphic. We note that, since B : W → W is self-adjoint, then B ± iI is invertible. Lemma 4.1. (x, a) ∈ Y± if and only if x ∈ X± and a = (B ∓ iI)−1 Ωx. Moreover, the deficiency indices of bT0 are equal and finite and satisfy def(bT0) = def(T0). Proof. Let (x, a) ∈ Y±. Then T1x = ±ix and Ba − Ωx = ±ia. Therefore x ∈ X± and a = (B ∓ iI)−1 Ωx. Conversely if x ∈ X± and a = (B ∓ iI)−1 Ωx then Ba − Ωx = ±ia so bT1(x, a) = ±i(x, a). We see that the mappings X± → Y± given by x → (x, (B ∓ iI)−1 Ωx) are vector space isomorphisms. In particular, dim (X±) = dim (Y±) . This shows that the deficiency indices of the minimal operator bT0 are finite and equal with (4.49) dim (Y+) = dim (Y−) = def(bT0) < ∞. 20 LANCE L. LITTLEJOHN AND RICHARD WELLMAN In particular, equation (4.49) guarantees the GKN-EM theorem applies to bT0. We now define the (degenerate) symplectic form in H ⊕ W associated with the operators bT0 and bT1. We remark that, in equation (4.33), we actually already computed this symplectic form. Definition 4.4 (General Symplectic Form). (cid:3) ((x, a), (y, b) ∈ D(bT1)), [(x, a), (y, b)]H⊕W := [x, y]H − hΩx, biW + ha, ΩyiW (4.50) where [·,·]H is the symplectic form defined in (3.3) and where the mapping Ω is defined in (4.8). We are now in position to apply the GKN-EM Theorem (Theorem 3.2) to the minimal operator bT0 in H ⊕ W and, as a result, characterize all self-adjoint extensions (respectively, restrictions) of bT0 (respectively, the maximal operator bT1). Theorem 4.4 (GKN-EM Theorem in H ⊕ W ). We have the following assumptions/definitions: (i) T0 and T1 are, respectively, the minimal and maximal operators in (H,h·,·iH ) , called the base (complex) Hilbert space, with domains D(T0) and D(T1); T0 is a closed, symmetric operator satisfying T0 ⊆ T1 with T ∗ (ii) The deficiency indices of T0 are assumed to be equal and finite and denoted by def(T0); (iii) [·,·]H is the symplectic form given by 0 = T1 and T ∗ 1 = T0; [x, y]H = hT1x, yiH − hx, T1yiH (x, y ∈ D(T1)), (see Definition 3.3); def(T0) (Condition 4.1) and orthonormal basis {ξj j = 1, . . . , dim W}; (iv) (W,h·,·iW ), the extension space, is a finite dimensional complex Hilbert space with dimW ≤ (v) B : W → W is a self-adjoint operator; (vi) H ⊕ W , the extended space, is the Hilbert space defined in (1.4) with inner product (1.5); (vii) P = {tj j = 1, . . . , dim W} is a partial GKN set (see (4.2) and (4.3)); (viii) ∆0 = D(T0) + span{tj j = 1, . . . , dim W} (see (4.4)); (ix) Ψ : ∆0 → W is defined to be Ψ dim WXj=1 x0 + (x) Ω : D(T1) → W is given by αjtj  = (x0 ∈ D(T0)) dim WXj=1 (see (4.5)); αjξj Ωx = dim WXj=1 [x, tj]H ξj (see (4.8)); (xi) bT0 : D(bT0) ⊆ H ⊕ W → H ⊕ W is the minimal operator in H ⊕ W defined by bT0(x, a) = (T1x, Ba) (x, a) ∈ D(bT0) = {(x, Ψx) x ∈ ∆0} (see Definition 4.1); SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 21 (xii) bT1 : D(bT1) ⊆ H ⊕ W → H ⊕ W is the maximal operator in H ⊕ W defined by bT1(x, a) = (T1x, Ba − Ωx) D(cT1) = {(x, a) x ∈ D(T1); a ∈ W} (see 4.2); (xiii) [·,·]H⊕W is the symplectic form given by [(x, a), (y, b)]H⊕W := [x, y]H − hΩx, biW + ha, ΩyiW (see 4.50). ((x, a), (y, b) ∈ D(bT1), Under these definitions and assumptions, we obtain the following results: Theorems 4.1 and 4.2); (a) bT0 is a closed, symmetric operator satisfying bT0 ⊆ bT1 with (bT0)∗ = bT1 and (bT1)∗ = bT0 (see (b) The deficiency indices of bT0 are equal and finite and def(bT0) = def(T0) (see Lemma 4.1); (c) Suppose bT is a self-adjoint extension of bT0 (equivalently, bT is a self-adjoint restriction of bT1) satisfying bT0 ⊆ bT ⊆ bT1. Then there exists a GKN set {(xj, aj) j = 1, . . . , def(T0)} ⊆ D(bT1) (α) {(xj, aj) j = 1, . . . , def(T0)} is linearly independent modulo D(bT0), (β) [(xj, aj), (xk, ak)]H⊕W = 0 for j, k = 1, . . . , def(T0) such that (see Remark 3.1) satisfying the two conditions (4.51) (4.52) bT (x, a) = (T1x, Ba − Ωx) D(bT ) = {(x, a) ∈ D(bT1) [(x, a), (xj , aj)]H⊕W = 0 (j = 1, . . . , def(T0))}. (d) If bT is defined by (4.51) and (4.52) where {(xj , aj) j = 1, . . . , def(T0)} ⊆ D(bT1) is a GKN set satisfying conditions (α) and (β), then bT is a self-adjoint extension of bT0 (equivalently, bT is a self-adjoint restriction of bT1) in H ⊕ W. 5. Examples 5.1. Example 1: The Legendre Type Self-Adjoint Operator. Throughout this example, we let H = L2(−1, 1) (with its usual inner product) and W = C2, endowed with the weighted Euclidean inner product (5.1) h(a1, b1), (a2, b2)iW = a1a2 + b1b2 A ((a1, b1), (a2, b2) ∈ W ); here A is a fixed, positive constant. Let {ξ1, ξ2} be the orthonormal basis in W given by (5.2) and, for this example, suppose B : W → W is the zero self-adjoint operator. defined earlier in (1.3); that is, ξ1 = (√A, 0) and ξ2 = (0,√A) In [3, 4, 12], the authors discuss the spectral analysis of the Legendre type differential expression (5.3) ℓLT [y](u) :=(cid:0)(1 − u2)2y′′(u)(cid:1)′′ −(cid:0)(8 + 4A(1 − u2))y′(u)(cid:1)′ where A is the same constant appearing in (5.1). This differential expression was first discovered by H. L. Krall [13, 14]. When (5.4) λn = n(n + 1)(n2 + n + 4A − 2) (n ∈ N0), 22 LANCE L. LITTLEJOHN AND RICHARD WELLMAN the equation ℓLT [y] = λny has a polynomial solution y = Pn,A(u) of degree n; that is, (5.5) The sequence {Pn,A}∞ sequence in the Hilbert space L2 µ[−1, 1], where ℓLT [Pn,A] = λnPn,A (n ∈ N0). n=0 is called the Legendre type polynomials; they form a complete orthogonal µ[−1, 1] = {f : [−1, 1] → C f is Lebesgue measurable with Z[−1,1] f2 dµ < ∞}, L2 with inner product hf, giµ = f (−1)g(−1) A +Z 1 −1 f (u)g(u)du + f (1)g(1) A , and where dµ is the Lebesgue-Stieltjes measure given by dµ = dx + 1 A δ(x + 1) + 1 A δ(x − 1). When y = Pn,A, we see from (5.3) and (5.5) that (5.6) ∓ 8AP ′ n,A(±1) = λnPn,A(±1). Various properties of the Legendre type polynomials can be found in [12]. Because the measure µ has jumps at u = ±1, the classic GKN theory is not immediately appli- µ[−1, 1]. In order to construct cable in finding a self-adjoint operator representation T of ℓLT [·] in L2 T, Everitt and Littlejohn [3, 4] first studied properties of functions in the maximal domain D(T1) = {x : (−1, 1) → C x, x′, x′′, x′′′ ∈ ACloc(−1, 1); x, ℓLT [x] ∈ L2(−1, 1)}, where T1 is the maximal operator, generated by ℓLT [·], in the Hilbert space L2(−1, 1). They establish the remarkable smoothness property (5.7) x ∈ D(T1) =⇒ x′′ ∈ L2(−1, 1) and hence, upon making the natural identifications (5.8) we can say that x(±1) = lim u→±1 x(u) x′(±1) = lim u→±1 x′(u), Moreover, they prove that the associated sesquilinear form has the simple formulation x ∈ D(T1) =⇒ x, x′ ∈ AC[−1, 1]. (5.9) [x, y]H = 8(x(1)y′(1) − x′(1)y(1) + x′(−1)y(−1) − x(−1)y′(−1)) (x, y ∈ D(T1)). Considering this last formula and Lemma 3.3, it is apparent that the minimal domain associated with ℓLT [·] is explicitly given by (5.10) The deficiency index of the minimal operator T0, generated by ℓLT [·], in L2(−1, 1) is def(T0) = 2. This follows since each endpoint u = ±1 is in the limit-3 case which can be shown by a Frobenius analysis. We emphasize that we are not seeking to find self-adjoint extensions of T0 in L2(−1, 1) but instead we want to find a self-adjoint representation of ℓLT [·] in L2 µ[−1, 1] which produces the D(T0) =(cid:8)x ∈ D(T1) x(±1) = x′(±1) = 0(cid:9) . (5.11) T x(u) =  x ∈ D(T ) := D(T1) −8Ax′(−1) u = −1 ℓLT [x](u) −1 < u < 1 8Ax′(1) u = 1 SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 23 Legendre type polynomials {Pn,A}∞ and Littlejohn show that the operator T : D(T ) ⊆ L2 n=0 as eigenfunctions. By analyzing functions in D(T1), Everitt µ[−1, 1] → L2 µ[−1, 1] defined by is self-adjoint, has the Legendre type polynomials {Pn,A}∞ spectrum n=0 as eigenfunctions, and has discrete σ(T ) = σp(T ) = {λn n ∈ N0}, where each λn is given in (5.4). It is surprising that the maximal domain D(T1) is the domain of a self-adjoint operator in L2 µ[−1, 1]. By the GKN Theorem, D(T1) cannot be the domain of a self-adjoint extension of T0 in L2(−1, 1). We now show, using the results developed in this paper, how to construct the self-adjoint operator T given in (5.11) in the direct sum space H ⊕ W. Indeed, below, we construct a self-adjoint operator bT that is, essentially, the operator T defined in (5.11). With this alternative approach, we will see how continuity is a GKN-EM boundary condition that produces the Legendre type self-adjoint operator bT . µ[−1, 1] is isometrically isomorphic The first step in our analysis is to observe that the space L2 to the direct sum H ⊕ W = {(x, (a, b)) x ∈ H; (a, b) ∈ W} . Next define tj ∈ D(T1) (j = 1, 2) by t1(u) =(cid:26) √A u near − 1 u near 1 0 t2(u) =(cid:26) 0 u near − 1 √A u near 1; we remark that such functions in D(T1) exist by Naimark's Patching Lemma [15, Lemma 2, Sec- tion 17.3]. It is straightforward to see, using (5.9) and (5.10), that {t1, t2} is a GKN set for T0. Consequently, we see that (5.12) where ∆0 is defined in (4.4). Moreover, ∆0 = {x0 + c1t1 + c2t2 x0 ∈ D(T0); c1, c2 ∈ C}, Ψ(x0 + c1t1 + c2t2) = c1ξ1 + c2ξ2 = hc1√A, c2√Ai, (5.13) where {ξ1, ξ2} is defined in (5.2) and where Ψ : ∆0 → W is the map defined in (4.5). Using (5.9), calculations show that [x, t1]H = 8√Ax′(−1), [x, t2]H = −8√Ax′(1) (x ∈ D(T1)). It follows that Ωx = 8√Ax′(−1)ξ1 − 8√Ax′(1)ξ2 = (8Ax′(−1),−8Ax′(1)) (5.14) (x ∈ D(T1)), where Ω : D(T1) → W is the mapping defined in (4.8). The minimal operator bT0 : D(bT0) ⊆ H ⊕ W → H ⊕ W , in this example, is given by (5.15) bT0(x, Ψx) = (T1x, BΨx) = (ℓLT [x], (0, 0)) D(bT0) = {(x, Ψx) x ∈ ∆0}. (5.16) From the theory we established in Section 4, bT0 is a closed, symmetric operator in H ⊕ W with defbT0 = 2. 24 LANCE L. LITTLEJOHN AND RICHARD WELLMAN (5.17) Using (5.14), we see that the associated maximal operator bT1 : D(bT1) ⊆ H ⊕ W → H ⊕ W is given explicitly by (5.18) From (5.1), (5.9) and (5.14), a calculation shows that the symplectic form [·,·]H⊕W , defined in (4.50), is given by bT1(x, (a, b)) = (ℓLT [x], (−8Ax′(−1), 8Ax′(1))) D(bT1) = {(x, (a, b)) x ∈ D(T1); a, b ∈ C}. (5.19) [(x, (a1, b1)), (y, (a2, b2))]H⊕W = [x, y]H − hΩx, (a2, b2)iW + h(a1, b1), ΩyiW = 8(x(1)y′(1) − x′(1)y(1) + x′(−1)y(−1) − x(−1)y′(−1)) − 8x′(−1)a2 + 8x′(1)b2 + 8Ay′(−1)a1 − 8Ay′(1)b1 for (x, (a1, b1)), (y, (a2, b2)) ∈ D(bT1). Define xj ∈ D(T1), for j = 1, 2, by , x2(u) =(cid:26) √A(u + 1) u near − 1 u near 1. 0 x1(u) =(cid:26) 0 u near − 1 √A(u − 1) u near 1 From (5.9) and (5.10), we see that {x1, x2} is a GKN set for T0. Calculations also show that (5.20) [x, x1]H = 8√Ax(1) and [x, x2]H = −8√Ax(−1) (x ∈ D(T1)). In addition, we see that (5.21) We now claim that Ωx1 = (0,−8A√A) and Ωx2 = (8A√A, 0). Proposition 5.1. {(x1, (0, 0)), (x2, (0, 0))} is a GKN set for bT0. Proof. Suppose that c1(x1, (0, 0)) + c2(x2, (0, 0)) = (c1x1 + c2x2, (0, 0)) ∈ D(bT0). By definition of D(bT0), we see that Ψ(c1x1 + c2x2) = (0, 0). This implies that c1x1 + c2x2 ∈ D(T0). However since {x1, x2} is a GKN set for T0, we must have c1 = c2 = 0. Hence {(x1, (0, 0)),hx2, (0, 0))} is linearly independent modulo D(bT0). Next, using (5.19), we see that [(x1, (0, 0)), (x2, (0, 0))]H⊕W = [x1, x2]H = 0. Calculations also show that This completes the proof of the Proposition. (cid:3) [(xj, (0, 0)), (xj , (0, 0))]H⊕W = 0 (j = 1, 2). We now find the appropriate self-adjoint operator bT in H ⊕ W having the Legendre type poly- nomial vectors {(Pn,A, (Pn,A(−1), Pn,A(1)))}∞ (d), the operator bT : D(bT ) ⊆ H ⊕ W → H ⊕ W , defined by bT (x, (a, b)) = (ℓLT [x], (−8Ax′(−1), 8Ax′(1))) n=0 as eigenfunctions. Indeed, using Theorem 4.4 part D(bT ) = {(x, (a, b)) ∈ D(bT1) [(x, (a, b)), (xj , (0, 0))]H⊕W = 0 (j = 1, 2)} is self-adjoint. (5.22) (5.23) SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 25 We now investigate each of the two boundary conditions in (5.23). From (5.19), (5.20) and (5.21), a calculation shows that 0 = [(x, (a, b)), (x1, (0, 0))]H⊕W = [x, x1]H − hΩx, (0, 0)iW + h(a, b), Ωx1iW = [x, x1]H + h(a, b), (0,−8A√A)iW = 8√Ax(1) − = 8√Ax(1) − 8√Ab, 8A√Ab A implying (5.24) b = x(1). A similar calculation, using (5.20) and (5.21), yields 0 = [(x, (a, b)), (x2, (0, 0))]H⊕W = −8√Ax(−1) + 8√Aa which establishes (5.25) a = x(−1). Hence the domain of bT , given in (5.23), simplifies to (5.26) Notice that this domain (5.26) extends the continuity of each x ∈ D(T1) from (−1, 1) to the closure [−1, 1]. It is remarkable that, in this sense, continuity is a GKN-EM boundary condition. Furthermore, from (5.5), (5.6) and (5.22), notice that D(bT ) = {(x, (x(−1), x(1))) x ∈ D(T1)}. bT (Pn,A, (Pn,A(−1), Pn,A(1))) = (ℓLT [Pn,A], (−8AP ′ = (λnPn,A, (λnPn,A(−1), λnPn,A(1))) = λn(Pn,A, (Pn,A(−1), Pn,A(1))). n,A(−1), 8AP ′ n,A(1))) Moreover, bT is the same operator as T defined in (5.11). Remark 5.1. If B : W → W is an arbitrary self-adjoint operator in W, the operator bS : D(bS) ⊆ H ⊕ W → H ⊕ W, defined by bS(x, (a, b)) = (ℓLT [x], B(a, b) + (−8Ax′(−1), 8Ax′(1))) is self-adjoint in H ⊕ W . However, it is the case that the Legendre type polynomial vectors {(Pn,A, (Pn,A(−1), Pn,A(1)))}∞ 5.2. Example 2: A Simple First-Order Differential Operator. Let H = L2[0, 1] be endowed with the standard L2 inner product D(bS) = {(x, (x(−1), x(1))) x ∈ D(T1)}, n=0 are eigenfunctions of bS if and only if B = 0. hx, yiH =Z 1 −1 x(u)y(u)du and let W = C have the usual Euclidean inner product ha, biW := ab (a, b ∈ W ). 26 LANCE L. LITTLEJOHN AND RICHARD WELLMAN In this example, we show how to construct a self-adjoint operator in H ⊕ W generated by the first-order Lagrangian symmetric differential expression ℓ[x](u) = ix′(u). Our construction can be modified to find numerous other self-adjoint operators in H ⊕ W generated by ℓ[·]. The maximal and minimal domains in H associated with ℓ[·] are respectively given by T1x = ix′ and (5.27) D(T1) = {x : [0, 1] → C x ∈ AC[0, 1]; x′ ∈ H}, T0x = ix′ D(T0) = {x ∈ D(T1) x(0) = x(1) = 0}; see [17, Chapter 13, Example 13.4]. The symplectic form [·,·]H associated with T1is given by [x, y]H = ix(1)y(1) − ix(0)y(0). An elementary calculation shows that the deficiency indices of T0 are equal with def(T0) = 1. Choose {ξ1 = 1} as the orthonormal basis for W . All self-adjoint operators B : W → W have the form Ba = αa for some real number α; we fix one such an operator. Define t1 ∈ D(T1) by and note that (5.28) t1(u) ≡ 1 (0 ≤ u ≤ 1) [x, t1]H = i(x(1) − x(0)). It is clear, from (5.27), that t1 is not the minimal domain; moreover, from (5.28), we see that [t1, t1]H = 0; this shows that {t1} is a GKN set for T0 in H. Moreover, with this GKN set, the reader can readily verify, using Theorem 2.3, that the operator T : D(T ) ⊂ H → H defined by T x = ix′ D(T ) = {x ∈ D(T1) x(0) = x(1)} is self-adjoint in H; see [17, Chapter 13, Example 13.4] for an interesting direct proof of the self- adjointness of T. The operators Ω : D(T1) → W and Ψ : ∆0 → W from Section 4 are given by Ωx = [x, t1]H ξ1 = i(x(1) − x(0)) and The maximal operator bT1 : D(bT1) ⊂ H ⊕ W → H ⊕ W and the minimal operator bT0 : D(bT0) ⊂ H ⊕ W → H ⊕ W can now both be defined. Indeed, Ψ(t0 + at1) = a (t0 ∈ D(T0)). bT1(x, a) = (ix′, Ba − Ωx) = (ix′, αa − i(x(1) − x(0))) D(bT1) = {(x, a) x ∈ D(T1), a ∈ W} SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 27 and bT0(x, Ψx) = (ix′, BΨx) = (ix′, αΨx) D(bT0) = {(x, Ψx) x ∈ ∆0}. The symplectic form [·,·]H⊕W associated with bT1 is given by [(x, a), (y, b)]H⊕W = [x, y]H − hΩx, biW + ha, ΩyiW (5.29) = i(x(1)y(1) − x(0)y(0)) − i(x(1) − x(0))b − ai(y(1) − y(0)) for (x, a), (y, b) ∈ D(bT1). Define x1 ∈ D(T1) by from (5.29), we see that x1(u) =(cid:26) 1 u near 1 0 u near 0; (5.30) (5.31) [(x1, 1/2), (x1, 1/2)]H⊕W = 0; [(x, a), (x1, 1/2)]H⊕W = ix(1) − i(x(1) − x(0))/2 − ai. We now show that {(x1, 1/2)} is a GKN set for cT0. From (5.30), note that We claim that (x1, 1/2) /∈ D(bT0); indeed, otherwise, we have for some t0 ∈ D(T0). However, choosing u = 0 or u = 1 shows that (5.32) is not possible. It now follows that {(x1, 1/2)} is a GKN set for bT0. From (5.30), we see that (u ∈ [0, 1]) x1(u) = t0(u) + (5.32) 1 2 t1(u) (5.33) [(x, a), (x1, 1/2)]H⊕W = 0 if and only if a = x(0) + x(1) 2 . It now follows, from Theorem 3.2 and (5.33), that the operator bT : D(bT ) ⊂ H ⊕ W → H ⊕ W defined by (5.34) is self-adjoint. bT (x, (x(0) + x(1))/2) = (ix′, α(x(0) + x(1))/2 − i(x(1) − x(0))) D(bT ) = {(x, (x(0) + x(1))/2) x ∈ D(T1)} 5.3. Example 3: Variations on the Fourier Self-Adjoint Operator. For this example, we consider the well known Fourier differential expression (5.35) ℓF [y](u) = −y′′(u) (u ∈ [a, b]) where [a, b] is a compact interval. Here, the Hilbert space is H = L2[a, b] and, in the sub-examples below, we will consider W to be either C or C2 with a weighted Euclidean inner product. The maximal operator T1 : D(T1) ⊂ H → H is defined by T1x = ℓF [x] D(T1) = {x : [a, b] → C x, x′ ∈ AC[a, b]; x′′ ∈ L2[a, b]} 28 LANCE L. LITTLEJOHN AND RICHARD WELLMAN while the minimal operator T0 : D(T0) ⊂ H → H is given by T0x = ℓF [x] D(T0) = {x ∈ D(T1)x(a) = x′(a) = x(b) = x′(b) = 0}. [x, y]H = x(b)y′(b) − x′(b)y(b) + x′(a)y(a) − x(a)y′(a) The symplectic form [·,·]H associated with T1 is given by (5.36) Because ℓF [·] is regular, the deficiency index of T0 is def(T0) = 2. Consequently, by the GKN Theorem, every self-adjoint extension of T0 in H will be a certain restriction of the maximal operator defined by two appropriate boundary conditions. One such self-adjoint operator is the classical Fourier trigonometric self-adjoint operator in H, generated by ℓF [f ], with domain (x, y ∈ D(T1)). (cid:8)x ∈ D(T1) x′(a) = x′(b), x(a) = x(b)(cid:9) . We list several examples of self-adjoint operators, generated by ℓF [·] in H ⊕ W . 5.3.1. One Dimensional Extension Spaces. Consider the one dimensional extension space W, with basis {ξ1 = 1}, given by Every self-adjoint operator B : W → W has the form Bz = αz for some α ∈ R; for this example, we fix such a B. Observe the Hilbert space H ⊕ W is suggested in a natural way by the inner product W = C hz1, z2iW = z1z2 (z1, z2 ∈ W ). x(u)y(u)dx + x(b)y(b). Z b a With this particular inner product in mind, H ⊕ W is isomorphic to the Lebesgue-Stieltjes inte- gration space generated by the discontinuous Lebesgue-Stieltjes measure Example 3.1 With the partial GKN set {t1} for T0 in H given by dµ = du + δ(u − b). t1(u) =(cid:26) 1 u near b 0 u near a, a calculation shows that {(x1, 0), (x2, 0)} is a GKN set for cT0 in H ⊕ W where u near b u − a u near a. x1(u) =(cid:26) u − b u near b and x2(u) =(cid:26) 0 u near a 0 We leave it to the reader to check that From these equations, we see that the operator (i) [(x, a), (x1, 0)]H⊕W = x(b) − a (ii) [(x, a), (x2, 0)]H⊕W = −x(a). bT (x, x(b)) = (−x′′, αx(b) + x′(b)) D(bT ) = {(x, x(b)) x ∈ D(T1); x(a) = 0} is self-adjoint in H ⊕ W. SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES 29 Example 3.2 By picking the partial GKN set {t1} for T0 in H, where and the GKN set {(x1, 0), (x2, 0)} for cT0 in H ⊕ W, where x1(u) =(cid:26) 0 u near b 1 u near a t1(u) =(cid:26) 0 u near b u − a u near a, and x2(u) =(cid:26) u − b u near b u near a, 0 the reader can check that the operator bT (x, x′(a)) = (−x′′, αx′(a) + x(a)) D(bT ) =(cid:8)(x, x′(b)) x ∈ D(T1); x(b) = 0(cid:9) is self-adjoint in H ⊕ W. 5.3.2. Two Dimensional Extension Spaces. For the last three examples, let W = C2 have the weighted inner product (cid:10)(z1, z2), (z′ 1, z′ 2)(cid:11)W = z1z′ 1 M + z2z′ 2 N , where M, N > 0. Let {ξ1 = (√M , 0), ξ2 = (0,√N)} be a basis for W. The reader can check that the most general form of a self-adjoint operator B : W → W , using this inner product, has the matrix representation and and (5.39) yielding With z = (z1, z2) ∈ W, calculations show that (5.38) Ωx1 = (M 3/2, 0) and Ωx2 = (0,−N 3/2). 0 = [(x, z), (x1, 0)]H⊕W = −x(a)√M + z1√M 0 = [(x, z), (x2, 0)]H⊕W = x(b)√N − z2√N z1 = x(a) and z2 = x(b). B =(cid:18) α βN/M γ (cid:19) , β (5.37) where α, β ∈ R and β ∈ C. Example 3.3 Define ti, xi ∈ D(T1) (i = 1, 2) by , t1(u) =(cid:26) √M u near a x1(u) =(cid:26) √M (u − a) u near a u near b u near b 0 0 , t2(u) =(cid:26) 0 x2(u) =(cid:26) 0 , u near a √N u near b √N (u − b) u near b. u near a It is the case that {t1, t2} is a GKN set for T0 in H and {(x1, 0), (x2, 0)} is a GKN set for bT0 in H ⊕ W. Moreover, using (5.36), we see that [x, t1]H = x′(a)√M ; [x, t2]H = −x′(b)√N ; [x, x1]H = −x(a)√M ; [x, x2]H = x(b)√N In particular, Ωx = [x, t1]ξ1 + [x, t2]ξ2 = (x′(a)M,−x′(b)N ) (x ∈ D(T1)). 30 LANCE L. LITTLEJOHN AND RICHARD WELLMAN That is, the boundary conditions expressed in (5.38) and (5.39) yield continuity of functions in the domain of the self-adjoint operator bT : D(bT ) ⊂ H ⊕ W → H ⊕ W defined by bT (x, (x(a), x(b))) = (−x′′, B(x(a), x(b)) − (x′(a)M,−x′(b)N )) where B is given in (5.37) and D(bT ) = {(x, (x(a), x(b))) x ∈ D(T1)}, B(x(a), x(b)) :=(cid:18) βN/M γ (cid:19)(cid:18) x(a) x(b) (cid:19) . α β In this case, the setting H ⊕ W can be identified with the Hilbert function space L2 σ[a, b], given by L2 σ[a, b] = {f : [a, b] → C f is Lebesgue measurable on [a, b] and kfkH⊕W < ∞}, where k·kH⊕W is the norm generated by the inner product hx, yiH⊕W =Z b a x(u)y(u)du + x(a)y(a) M + x(b)y(b) N and dσ is the Lebesgue-Stieltjes measure generated by the distribution function σ(u) = 1 M − u 1 N + b   + a u ≤ a a < u < b u ≥ b. Example 3.4 For this example, we switch the roles of {t1, t2} and {x1, x2} which are defined in Example 5.3.2. Note that, in this case, {(t1, 0), (t2, 0)} is a GKN set for bT0 and {x1, x2} is a GKN set for T0. The calculations given in the previous example hold with the exception Again, with z = (z1, z2), we find that and so that Ωx = (−x(a)M, x(b)M ) (x ∈ D(T1)). 0 = [(x, z), (t1, 0)]H⊕W = x′(a)√M − z1√M 0 = [(x, z), (t2, 0)]H⊕W = −x′(b)√N + z2√N z1 = x′(a) and z2 = x′(b). In this case, the operator bT : D(bT ) ⊂ H ⊕ W → H ⊕ W, defined by D(bT ) =(cid:8)(x, (x′(a), x′(b))) x ∈ D(T1)(cid:9) bT (x, (x′(a), x′(b))) = (−x′′, B(x′(a), x′(b)) − (−x(a)M, x(b)M )), with domain is self-adjoint. In this case, for x, y ∈ D(bT ), the inner product on H ⊕ W simplifies to the discrete Sobolev inner product (5.40) hx, yiH⊕W =Z b a x(u)y(u)du + x′(a)y′(a) x′(b)y′(b) M + . N Notice that, because of the derivatives in the discrete part of (5.40), the closure of D(bT ) is not a function space and there is no positive Borel measure generating this inner product. SELF-ADJOINT OPERATORS IN EXTENDED HILBERT SPACES Example 3.5 For our last example, we consider a variation of the last two examples. define ti, xi ∈ D(T1) (i = 1, 2) by t1(u) =(cid:26) √M u near a x1(u) =(cid:26) 0 u near a √N u near b u near b 0 , , u near a t2(u) =(cid:26) 0 √N (u − b) u near b x2(u) =(cid:26) √M (u − a) u near a u near b 0 , 31 Indeed, In this case {t1, t2} is a GKN set for T1 and {(x1, 0), (x2, 0)} is a GKN set for bT0. Moreover, a calculation shows Ωx = (x′(a)M, x(b)N ). With z = (z1, z2), the two boundary conditions yield 0 = [(x, z), (x1, 0)]H⊕W = −x′(b)√N + z2√N 0 = [(x, z), (x2, 0)]H⊕W = −x(a)√M + z1√M z1 = x(a) and z2 = x′(b). These calculation show that the operator bT , given by with domain bT (x, (x(a), x′(b))) = (−x′′, B(x(a), x′(b)) − (x′(a)M, x(b)N )) D(bT ) =(cid:8)(x, (x(a), x′(b))) x ∈ D(T1)(cid:9) , is self-adjoint in H ⊕ W. For each x, y ∈ D(bT ), the 'mixed' inner product in H ⊕ W reduces to x′(b)y′(b) x(a)y(a) x(u)y(u)du + + M . N hx, yiH⊕W =Z b a As in the last example, no positive Borel measure generates this inner product and the closure of D(bT ) in the topology from h·,·iH⊕W is not a function space. References [1] N. I. Akhiezer and I. M. Glazman, Theory of Linear Operators in Hilbert space, Parts I and II, Scottish Academic Press, Pitman Advanced Publishing Program, London, U.K. 1981. [2] N. Dunford and J. T. Schwartz, Linear Operators, Part II, John Wiley Publishers, New York, 1963. [3] W. N. Everitt, A. M. Krall, and L. L. Littlejohn, On some properties of the Legendre type differential expression, Quaestiones Math., 13(1), 1990, 83-116. [4] W. N. Everitt and L. L. Littlejohn, Differential operators and the Legendre type polynomials, Differential and Integral Equations, 1(1), 1988, 97-116. [5] W. N. Everitt and L. L. Littlejohn, Orthogonal polynomials and spectral theory: a survey, Proceedings of the III International Symposium on Orthogonal Polynomials and Applications, Erice, Italy, 1990. IMACS Annals on Computing and Applied Mathematics 9(1991), 21-55; J. C. Baltzer AG, Basel, Switzerland, 1991, 21-55. [6] W. N. Everitt, L. L. Littlejohn and R. Wellman, Orthogonal polynomial solutions of linear ordinary differential equations, Proceedings of the Fifth International Symposium on Orthogonal Polynomials, Special Functions and their Applications (Patras, 1999), J. Comput. Appl. Math. 133 (2001), no. 1-2, 85–109. [7] W. N. Everitt and L. Markus, The Glazman-Krein-Naimark Theorem for Ordinary Differential Operators, Op- erator Theory: Advances and Applications 98(1997), 118-130. [8] W. N. Everitt and L. Markus, Complex Symplectic Geometry with Applications to Ordinary Differential Opera- tors, Trans. Amer. Math. Soc., 351(12), 1999, 4905-4945. [9] W. N. Everitt and L. Markus, Boundary Value Problems and Symplectic Algebra for Ordinary Differential and Quasi-Differential Operators, Math. Surveys Monogr., Volume 61, American Mathematical Society, Providence, Rhode Island, 1999. 32 LANCE L. LITTLEJOHN AND RICHARD WELLMAN [10] W. N. Everitt and L. Markus, Complex Symplectic Spaces and Boundary Value Problems, Bull. Amer. Math. Soc. (N.S.) 42 (2005), no. 4, 461–500. [11] V. Guillemin and S. Sternberg, Symplectic techniques in physics, Cambridge University Press, 1990. [12] A. M. Krall, On Orthogonal Polynomials Satisfying Fourth Order Differential Equations, Proc. Roy. Soc. Edinb. (A), 1981, 271-288. [13] H. L. Krall, Certain Differential Equations for Tchebycheff polynomials, Duke Math J., 4. 1938, 705-718 [14] H. L. Krall, Orthogonal polynomials satisfying a Certain fourth order differential equation, The Pennsylvania State College Studies, No. 6, The Pennsylvania State College, State College, PA. 1940. [15] M. A. Naimark, Linear Differential Operators, Part II, Ungar Publishing Co., New York, 1968. [16] M. Reed and B. Simon, Methods of Modern Mathematical Physics I. Functional Analysis, Academic Press, New York, 1972. [17] W. Rudin, Functional Analysis (2nd edition), McGraw-Hill Publishers, New York (1991). [18] M. Stone, Linear transformations in Hilbert space, American Mathematical Society Colloquium Publications 15, Providence, RI, 1932. [19] J. Weidmann, Linear operators in Hilbert space, Springer-Verlag, Heidelberg, 1980. [20] A. Zettl, Formally self-adjoint quasi-differential operators, Rocky Mountain J. Math., 5(1975), 453-474. [21] A. Zettl, Sturm-Liouville theory, Mathematical Surveys and Monographs, 121. American Mathematical Society, Providence, RI, 2005. Department of Mathematics, Baylor University, Waco, TX 76798-7328, Department of Mathematics and Computer Science, Westminster College, Salt Lake City, UT 84105 E-mail address: [email protected]
1707.05669
3
1707
2018-01-17T17:16:52
Quantum stochastic Lie-Trotter product formula II
[ "math.FA", "math-ph", "math-ph", "math.PR" ]
A natural counterpart to the Lie-Trotter product formula for norm-continuous one-parameter semigroups is proved, for the class of quasicontractive quantum stochastic operator cocycles whose expectation semigroup is norm continuous. Compared to previous such results, the assumption of a strong form of independence of the constituent cocycles is overcome. The analysis is facilitated by the development of some quantum Ito algebra. It is also shown how the maximal Gaussian component of a quantum stochastic generator may be extracted - leading to a canonical decomposition of such genarators, and the connection to perturbation theory is described. Finally, the quantum Ito algebra is extended to quadratic form generators, and a conjecture is formulated for the extension of the product formula to holomorphic quantum stochastic cocycles.
math.FA
math
QUANTUM STOCHASTIC LIE–TROTTER PRODUCT FORMULA II J. MARTIN LINDSAY Abstract. A natural counterpart to the Lie–Trotter product formula for norm-contin- uous one-parameter semigroups is proved, for the class of quasicontractive quantum stochastic operator cocycles whose expectation semigroup is norm continuous. Com- pared to previous such results, the assumption of a strong form of independence of the constituent cocycles is overcome. The analysis is facilitated by the development of some quantum Ito algebra. It is also shown how the maximal Gaussian component of a quantum stochastic generator may be extracted - leading to a canonical decomposi- tion of such generators, and the connection to perturbation theory is described. Finally, the quantum Ito algebra is extended to quadratic form generators, and a conjecture is formulated for the extension of the product formula to holomorphic quantum stochastic cocycles. Contents Introduction 1. Quantum Ito algebra 2. Quantum stochastics 3. Product Formula 4. Maximal Gaussian component of a QS generator 5. Perturbation of QS cocycles 6. Holomorphic QS Cocycles References 1 2 8 11 15 18 19 22 Introduction The Lie product formula in a unital Banach algebra states that (ea/neb/n)n → ea+b as n → ∞. Trotter extended this to C0-semigroups on a Banach space where it holds under com- patibility assumptions on the generators, convergence being in the strong operator sense ([Tro], see e.g. [Dav]). It has been further refined, notably by Chernoff ([Che]) and Kato ([Kat]). These product formulae are widely used in mathematical physics and probability theory – for example in establishing positivity preservation of semigroups, and they have an intimate connection to Feynman–Kac formulae (see e.g. [ReS]). Given that quantum stochastic cocycles may be analysed from their associated semigroups ([LW3]), it is nat- ural to seek product formulae in this context. Further motivation comes from the fact that such cocycles are quantum counterparts to stochastic semigroups in the sense of Skoro- hod ([Sko]). Product formulae have been obtained in a variety of quantum stochastic settings ([PaS], [LS1,3], [DLT], [DGS]). The earliest of these dates from before the advent of quantum stochastic calculus ([HuP]). In all of these works the constituent cocycles 2000 Mathematics Subject Classification. 46L53 (primary); 81S25, 47D06, 46N50 (secondary). Key words and phrases. Lie–Trotter product formula, quantum stochastic cocycle, one-parameter semigroup, series product, concatenation product, quantum Ito algebra, quantum stochastic analysis. 1 2 MARTIN LINDSAY enjoy a strong independence property, namely their respective noise dimension spaces are mutually orthogonal. In this paper a Lie–Trotter product formula is established for quasicontractive element- ary (i.e. Markov-regular) quantum stochastic operator cocycles, with no independence assumption on the driving quantum noise. It is a direct generalisation of the product for- mula proved in [LS1], and is proved by quite different means. Properties of the composition law on the set of quantum stochastic generators that is realised by the stochastic product formula established here are also elucidated. Known in the setting of quantum control theory as the series product ([GoJ]), it is more commonly associated with the perturbation of quantum stochastic cocycles ([EvH]). The composition of stochastic generators also corresponds to the operator product (i.e. standard composition) of the generators of the quantum random walks whose scaled embeddings approximate the constituent cocycles ([BGL]). Analysis of this composition leads to left and right series decompositions of a quantum stochastic generator. A decomposition for such generators in terms of the so- called concatenation product is also given; this yields the generator's maximal Gaussian part. It is conjectured here that, as in the case of orthogonal noises ([LS3]), the more gen- eral Lie–Trotter product formula given in this paper has an extension to the class of holomorphic quantum stochastic cocycles ([LS2]). By contrast, without orthogonality of noise dimension spaces there seems to be no sensible formulation of a Lie–Trotter product formula for quantum stochastic mapping cocycles. The plan of the paper is as follows. In Section 1 some quantum Ito algebra is developed, for studying the series product on the class of stochastic generators considered here. In Section 2 the relevant quantum stochastic analysis is recalled. The quantum stochastic Lie–Trotter product formula is proved in Section 3. In Section 4 the maximal Gaussian component of a quantum stochastic generator is extracted by means of the concatenation product. In the short Section 5, the connection to perturbation theory is described, and in Section 6 the quantum Ito algebra is extended to quadratic form generators and a conjecture for quasicontractive holomorphic quantum stochastic cocycles is formulated. Notation. For a vector-valued function f : R+ → V and subinterval J of R+, fJ denotes the function R+ → V which agrees with f on J and vanishes elsewhere. For Hilbert spaces h and h′, B(h; h′) denotes the space of bounded operators from h to h′ and B(h; h′)1 denotes its closed unit ball, abbreviated to B(h) and B(h)1 respectively when h′ = h. For an operator T ∈ B(h), its real and imaginary parts are denoted Re T and Im T respectively, thus T is dissipative if and only if Re T 6 0. The selfadjoint part of a subset A of an involutive space is denoted Asa. The predual of B(h), that is the space of ultraweakly continuous linear functionals on B(h), is denoted B(h)∗. Algebraic and ultraweak tensor products are denoted by ⊗ and ⊗ respectively and, for vectors ζ, η ∈ h, ωζ,η ∈ B(h)∗ denotes the functional given by T 7→ hζ, T ηi. The symbol ⊂⊂ is used to denote finite subset. 1. Quantum Ito algebra For this section take Hilbert spaces h and H. The block matrix decomposition enjoyed by operators in B(h⊕H) is frequently appealed to below. With respect to the distinguished orthogonal projection ∆ := P{0h}⊕H =(cid:20) 0h 0h;H 0H;h IH (cid:21) ∈ B(h ⊕ H), the composition law on B(h ⊕ H) given by F1 ✁ F2 := F1 + F1∆F2 + F2, QS LIE–TROTTER PRODUCT FORMULA 3 with useful alternative asymmetric expressions F1(I + ∆F2) + F2 and F1 + (I + F1∆)F2, has both and F1 + F2 + F3 +(cid:0)F1∆F2 + F1∆F3 + F2∆F3(cid:1) + F1∆F2∆F3, F1 ✁ F3 + (I + F1∆) F2 (∆F3 + I), (1.1a) (1.1b) as common expression for F1 ✁ (F2 ✁ F3) and (F1 ✁ F2) ✁ F3. Moreover the composition ✁ has 0h⊕H as identity element, and the operator adjoint as involution since (F1 ✁ F2)∗ = F ∗ 2 ✁ F ∗ 1 . The notation ✁ is taken from the quantum control theory literature, where the composi- tion is called the series product. Let b(h ⊕ H) := (B(h ⊕ H), ✁) denote the resulting *-monoid (i.e. involutive semigroup- with-identity), let β ∈ R, and consider the following subsets of b(h ⊕ H): qcβ(h ⊕ H) := c(h ⊕ H) + β∆⊥, qc(h ⊕ H) := c(h ⊕ H) + R+∆⊥, z(h ⊕ H) :=(cid:8)K ⊕ 0H : K ∈ B(h)(cid:9) = ∆⊥ b(h ⊕ H) ∆⊥, c(h ⊕ H) :=(cid:8)F ∈ b(h ⊕ H) : F ∗ ✁ F 6 0}, i(h ⊕ H) :=(cid:8)F ∈ b(h ⊕ H) : F ∗ ✁ F = 0}, i(h ⊕ H)∗ :=(cid:8)F ∗ : F ∈ i(h ⊕ H)}, u(h ⊕ H) := i(h ⊕ H) ∩ i(h ⊕ H)∗, and for F ∈ qc(h ⊕ H) set β0(F ) := inf{β ∈ R : F − β∆⊥ ∈ c(h ⊕ H)}. These classes are relevant to the characterisation of the stochastic generators of quantum stochastic cocycles which are respectively contractive, quasicontractive (with exponential growth bound β0(F )), isometric, coisometric and unitary (see Theorem 2.1 below). Note that, for F ∈ qc(h ⊕ H), F ∈ qcβ(h ⊕ H) if and only if β > β0(F ). Remarks. Further classes are relevant to the characterisation of quantum stochastic cocy- cles which are nonnegative, selfadjoint, partially isometric or projection-valued ([Wil]). The characterisation of the generators of 'pure-noise' (or 'local') nonnegative contraction cocycles (for which h = C) plays an important role in the identification of the minimal dilation of a quantum dynamical semigroup ([Bha]). Let F1, F2 and F in b(h ⊕ H) have respective block matrix formsh K1 M1 and(cid:2) K M L Q−I(cid:3). Then L2 Q2−Ii L1 Q1−Ii,h K2 M2 F1 ✁ F2 =(cid:20)K1 + K2 + M1L2 M1Q2 + M2 Q1Q2 − I (cid:21) , and F ∗ ✁ F =(cid:20)K ∗ + K + L∗L L∗Q + M Q∗Q − I(cid:21) , M ∗ + Q∗L L1 + Q1L2 (1.2) 4 MARTIN LINDSAY moreover, for Z ∈ z(h ⊕ H), β ∈ R and X ∈ b(h ⊕ H), (F + Z)∗ ✁ (F + Z) = Z ∗ + F ∗ ✁ F + Z (F − β∆⊥)∗ ✁ (F − β∆⊥) = F ∗ ✁ F − 2β∆⊥, and F ∗ ✁ X ✁ F = F ∗ ✁ F + (I + ∆F )∗X(I + ∆F ). These identities imply the following relations: (G ✁ F )∗ ✁ (G ✁ F ) = F ∗ ✁ F + G∗ + G if G ∈ z(h ⊕ H), 6 F ∗ ✁ F + 2β∆⊥ = F ∗ ✁ F if G ∈ qcβ(h ⊕ H), if G ∈ i(h ⊕ H). The basic algebraic properties of these subsets of b(h ⊕ H) are collected in the following proposition; using the above observations, their proof is straightforward. Recall that an operator T ∈ B(h) is dissipative if it satisfies Re T 6 0, that is Rehu, T ui 6 0 (u ∈ h). Proposition 1.1. In the *-monoid b(h ⊕ H) the following hold. (a) Its group of invertible elements is given by b(h ⊕ H)× =(cid:8)(cid:2) ∗ ∗ the identity element being 0 =(cid:2) 0 0 (cid:20)M Q−1L − K −M Q−1 ∗ Q(cid:3) − ∆ ∈ b(h ⊕ H) : Q ∈ B(H)×(cid:9); 0 I(cid:3) − ∆, and the inverse of(cid:2) K M L Q(cid:3) − ∆ being I (cid:21) Q−1(cid:2)−L I(cid:3) −(cid:20)K 0 I(cid:21) . 0 −Q−1L Q−1 (cid:21) − ∆ =(cid:20)−M (b) Its centre is z(h ⊕ H). (c) Denoting the class of dissipative operators on h by D(h), qc(h ⊕ H) = c(h ⊕ H) + z(h ⊕ H), and c(h ⊕ H) ∩ z(h ⊕ H) = {T ⊕ 0H : T ∈ D(h)}. (d) qc(h⊕ H), c(h⊕ H), i(h⊕ H) and u(h⊕ H) are submonoids; their groups of invertible elements are given by i(h ⊕ H)× = c(h ⊕ H)× = u(h ⊕ H), and qc(h ⊕ H)× = u(h ⊕ H) + z(h ⊕ H). Moreover, i(h ⊕ H) ∩ z(h ⊕ H) = u(h ⊕ H) ∩ z(h ⊕ H) = i z(h ⊕ H)sa. Remark. Clearly u(h ⊕ H) is also closed under taking adjoints, and so is a sub-*-monoid of b(h ⊕ H). In fact qc(h ⊕ H) and c(h ⊕ H) are sub-*-monoids too; this is not immediately obvious. It follows from Part (e) of Theorem 1.3 below, but there are now also direct proofs. The one given below is based on an elegant argument of Wills, arising as a biproduct of his analysis of partially isometric quantum stochastic cocycles ([Wil]). Proposition 1.2. Let F ∈ b(h ⊕ H) and T ∈ z(h ⊕ H)sa. Then (∆F + I)∗(F ✁ F ∗)(∆F + I) = F ∗ ✁ F + (F ∗ ✁ F ) ∆ (F ∗ ✁ F ), (1.3) and F ∗ ✁ F 6 T if and only if F ✁ F ∗ 6 T. QS LIE–TROTTER PRODUCT FORMULA 5 Proof. By the associativity of ✁, setting F1 = F ∗, F2 = F ✁ F ∗ and F3 = F in the expression (1.1b) for F1 ✁ F2 ✁ F3 we see that F ∗ ✁ F + LHS(1.3) = F ∗ ✁ (F ✁ F ∗) ✁ F = (F ∗ ✁ F ) ✁ (F ∗ ✁ F ) = RHS(1.3) + F ∗ ✁ F, so identity (1.3) holds. Suppose now that F ✁ F ∗ 6 T . Then, since ∆T = 0 = T ∆, LHS(1.3) 6 T and so, since ∆ > 0 and F ∗ ✁ F is selfadjoint, F ∗ ✁ F = (1.3) − (F ∗ ✁ F )∆(F ∗ ✁ F ) 6 (1.3) 6 T. The converse implication follows by exchanging F and F ∗. Consider the following possible block matrix forms for F ∈ b(h ⊕ H): 2 (L∗L + A2) L (cid:20)βI + iH − 1 (cid:20) βI + iH − 1 2 (M M ∗ + B2) −CM ∗ − (I − CC ∗)1/2E∗B (cid:21) , and −L∗C − AD(I − C ∗C)1/2 C − I M C − I(cid:21) , (cid:3) (1.4a) (1.4b) in which β ∈ R, H ∈ B(h)sa, A, B ∈ B(h)+, C ∈ B(H)1 and D, E ∈ B(H; h)1, so that H = Im K. Theorem 1.3 (Cf. [LW1], [GL+]). Let β ∈ R and F ∈ b(h ⊕ H). (a) The following are equivalent: (i) F ∈ qcβ(h ⊕ H). (ii) F has block matrix form (1.4a). (iii) F has block matrix form (1.4b). (iv) F ∗ ∈ qcβ(h ⊕ H). (b) F ∈ i(h ⊕ H) if and only if F has block matrix form (1.4a), with β = 0, A = 0 and C isometric. (c) F ∈ i(h ⊕ H)∗ if and only if F has block matrix form (1.4b), with β = 0, B = 0 and C coisometric. (d) Let H1 ⊕ H2 be an orthogonal decomposition of H. Then, with respect to the inclusion J1 : h ⊕ H1 → h ⊕ H, J ∗ 1 qcβ(h ⊕ H)J1 = qcβ(h ⊕ H1). (1.5) (e) Set H′ := H⊕(H⊕h). Then, in terms of the inclusion J : h⊕H → (h⊕H)⊕(H⊕h) = h ⊕ H′, c(h ⊕ H) = J ∗u(h ⊕ H′)J. Proof. The proof exploits the fact that, for an operator T ∈ B(h ⊕ H), T > 0 if and only 0H;H⊕h(cid:21) ∈ B(h ⊕ H; h ⊕ H′). i where (X, Z, V ) ∈ B(h)+ × B(H)+ × X Z 1/2V ∗X 1/2 X 1/2V Z 1/2 Z 0H1;H2(cid:21) ∈ B(h ⊕ H1; h ⊕ H), and J = Ih ⊕(cid:20) IH Remark. In block matrix form J1 = Ih ⊕(cid:20) IH1 if T has block matrix formh B(H; h)1. β = 0 and, setting S := (I − C ∗C)1/2, −G∗ ✁ G =(cid:20) A2 SD∗A ADS S2 (cid:21) > 0, (a) Set G = F − β∆⊥. Suppose first that (ii) holds. Then G is given by (1.4a) with 6 MARTIN LINDSAY so G ∈ c(h ⊕ H). Thus (ii) implies (i). Conversely, suppose that (i) holds and let(cid:2) K M L C−I(cid:3) be the block matrix decomposition of G. Then G ∈ c(h ⊕ H) so (cid:20)K ∗ + K + L∗L M + L∗C C ∗C − I(cid:21) = G∗ ✁ G 6 0. M ∗ + C ∗L Thus (K ∗ + K + L∗L) 6 0, kCk 6 1 and, for some contraction operator D, M = −L∗C − [−(K ∗ + K + L∗L)]1/2D(I − C ∗C)1/2 so G is given by (1.4a) with β = 0 and H = Im K. Thus (i) implies (ii). Therefore (i) and (ii) are equivalent; taking adjoints we see that (iv) and (iii) are equivalent too. The equivalence of (i) and (iv) follows from Proposition 1.2. (b) This is an immediate consequence of (1.2). (c) This follows from Part (b), by taking adjoints. (d) For F1 ∈ qcβ(h ⊕ H1), F1 = J ∗ 1 F J1, where F =(cid:20)F1 0 0(cid:21) ∈ qcβ(h ⊕ H), 0 so RHS ⊂ LHS in (1.5). For the reverse inclusion, setting ∆1 := 0h ⊕IH1 and ∆ := 0h ⊕IH, ∆ − J1∆1J ∗ 1 = 0h ⊕ IH1 ⊕ IH2 − 0h ⊕ IH1 ⊕ 0H2 = 0h ⊕ 0H1 ⊕ IH2 > 0. Thus, for F ∈ c(h ⊕ H), the operator F1 := J ∗ 1 F J1 is in c(h ⊕ H1) since 1 F ∗J1 + J ∗ 1 ✁ F1 = J ∗ F ∗ 1 c(h ⊕ H)J1 ⊂ c(h ⊕ H1) and so, since J ∗ 1 F J1 + J ∗ 1 F ∗J1∆1J ∗ Therefore J ∗ 1 (β ∈ R), LHS ⊂ RHS. (e) In view of (d), it suffices to show that J ∗u(h ⊕ H′)J ⊃ c(h ⊕ H). Accordingly, let F ∈ c(h ⊕ H). Then, by what we have proved already, F has block matrix form (1.4a) with β = 0. Setting 1 β∆⊥J1 = β∆⊥ 1 F J1 6 J ∗ 1 (F ∗ ✁ F )J1 6 0. F ′ := iH − 1 2 (L∗L + A2) −(L∗C + ADS) L∗T − ADC ∗ −AR L C − I −T D∗A RA S 0 C ∗ − I 0  , 0 0 0 where R = (I − DD∗)1/2, S = (I − C ∗C)1/2, and T = (I − CC ∗)1/2, it is now easily verified that F ′ ∈ u(h ⊕ H′). Since J ∗F ′J = F , (e) follows and the proof is complete. (cid:3) Remarks. The dilation property (e) is effectively proved in [GL+], under the assumption that h and H are separable. If H = h ⊗ k for a Hilbert space k (as it is in the application to QS analysis, where k is the noise dimension space) then, in (e), H′ = h ⊗ k′ where k′ = k ⊕ k ⊕ C. In terms of its block matrix form (cid:2) K M L C−I(cid:3), and bound β0 := β0(F ), the equivalent conditions for F ∈ b(h ⊕ H) to be in qc(h ⊕ H) read respectively as follows: M ∗ + C ∗L (cid:20)2(Re K − β0Ih) + L∗L M + L∗C (cid:20)2(Re K − β0Ih) + M M ∗ L∗ + M C ∗ C ∗C − IH(cid:21) 6 0, and CC ∗ − IH(cid:21) 6 0. L + CM ∗ and set β0 = β0(F ). Then, setting Proposition 1.4 (Left and right series decomposition). Let F =(cid:2) K M L C−I(cid:3) ∈ qc(h ⊕ H) 0(cid:21) , F1 =(cid:20)β0Ih + i Im K 0 3 :=(cid:20)Re K + 1 0 (cid:21) , F r 2 =(cid:20)− 1 C − I (cid:21) , and F r 2 L∗L −L∗ L 2 L∗L − β0Ih M + L∗C 0 3 =(cid:20)− 1 2 :=(cid:20)Re K + 1 2 M M ∗ M −M ∗ 0(cid:21) , F ℓ C − I(cid:21) , 2 M M ∗ − β0Ih L + CM ∗ F ℓ 0 0 QS LIE–TROTTER PRODUCT FORMULA 7 the following hold: F1 ∈ z(h ⊕ H), F ℓ 2 , F r 3 ∈ u(h ⊕ H), F r 2 , F ℓ 3 , ∈ c(h ⊕ H), and F1 ✁ F ℓ 2 ✁ F ℓ 3 = F = F1 ✁ F r 2 ✁ F r 3 , Proof. It follows from Theorem 1.3, and the above block matrix inequalities characterising membership of qc(h ⊕ H), that F ℓ 2 , F r 3 ∈ u(h ⊕ H) and F ℓ 2 ∈ c(h ⊕ H). 3 , F r The two series decompositions follow from the identities F1∆F ℓ 2 = 0, F ℓ 2 ∆F ℓ The rest is clear. 3 =(cid:20)0 L∗ − L∗C 0 0 (cid:21) , F1∆F r 2 = 0 and F r 2 ∆F r 3 =(cid:20) 0 M ∗ − CM ∗ 0(cid:21) . 0 (cid:3) Remarks. (i) The left and right series decompositions are related via the adjoint operation as follows: (F ∗)1 = (F1)∗, (F ∗)ℓ 2 = (F r 3 )∗ and (F ∗)ℓ 3 = (F r 2 )∗. (ii) Let F =(cid:2) K M L C−I(cid:3) ∈ qc(h ⊕ H) and set H := Im K. Then F ∈ i(h ⊕ H) if and only if F has left series decomposition with C isometric, whereas F ∈ i(h ⊕ H)∗ if and only if F has right series decomposition with C coisometric. Now suppose that H has an orthogonal decomposition H1 ⊕ H2. Define injections 0 i(cid:20)H 0 i(cid:20)H 0 0(cid:21) ✁(cid:20)− 1 0(cid:21) ✁(cid:20)0 0 0 0 2 L∗L −L∗ L 2 M M ∗ M −M ∗ 0 (cid:21) ✁(cid:20)0 0 C − I(cid:21) , 0(cid:21) , 0 C − I(cid:21) ✁(cid:20)− 1 L N(cid:21) 7→ F ⊕ 0H2 = L N(cid:21) 7→ Σ(cid:0)F ⊕ 0H1(cid:1) = ι : b(h ⊕ H1) → b(h ⊕ H), F =(cid:20)K M ι′ : b(h ⊕ H2) → b(h ⊕ H), F =(cid:20)K M K M 0 L N 0 0 0 0 , and L 0 N , K 0 M 0 0 0 (1.7) Σ being the sum-flip map B(h ⊕ H2 ⊕ H1) → B(h ⊕ H), and define the composition b(h ⊕ H1) × b(h ⊕ H2) → b(h ⊕ H), (F1, F2) 7→ F1 ⊞ F2 := ι(F1) + ι′(F2), known as the concatenation product in quantum control theory ([GoJ]). Thus L2 N2(cid:21) = (cid:20)K1 M1 L1 N1(cid:21) ⊞(cid:20)K2 M2 K1 + K2 M1 M2 0 N1 0 N2 L1 L2  . Note that In view of the identity ι(F1) = F1 ⊞ 0H2 and ι′(F2) = 0H1 ⊞ F2. ι(F1) ∆ ι′(F2) = 0h⊕H = ι′(F2) ∆ ι(F1), for F1 ∈ b(h ⊕ H1) and F2 ∈ b(h ⊕ H2), the concatenation product is effectively a special case of the series product: F1 ⊞ F2 = ι(F1) ✁ ι′(F2). (1.8) 8 MARTIN LINDSAY We end this section with a significant representation of the quantum Ito algebra. It is relevant to the realisation of QS cocycles as time-ordered exponentials ([Hol]), and also to the convergence of a class of scaled quantum random walks to QS cocycles ([BGL]). Set S := Ih⊕H⊕h + A, where A is the following subalgebra of B(h ⊕ H ⊕ h): thus S consists of the elements of B(h ⊕ H ⊕ h) having block matrix form(cid:20) Ih ∗ ∗ respect to (operator composition and) the involution given by 0 0 Ih(cid:21). With 0 ∗ ∗ (cid:8)T ∈ B(h ⊕ H ⊕ h) : Ph⊕H⊕{0}T = T = T P{0}⊕H⊕h(cid:9); T 7→ T ⋆ := ΞT ∗Ξ, where Ξ := 0 ,  = Ih⊕H⊕h + L Q − I(cid:21) 7→ (cid:20)K M Ih M K 0 Q L 0 Ih 0 M K 0 Q − I L 0 0 IH 0 0 0 Ih Ih 0 0 0 0 S is a sub-*-monoid of B(h ⊕ H ⊕ h), and the following is readily verified. Proposition 1.5 ([Hol], [Bel]). The prescription defines an isomorphism of *-monoids φ : b(h ⊕ H) → S. 2. Quantum stochastics L2(R+; k). The quantum Ito algebra developed in the previous section is applied below For the rest of the paper fix Hilbert spaces h and k, and set bk := C ⊕ k and K := with h = h and H = k ⊗ h, so that h ⊕ H =bk ⊗ h. In this context, the operator ∆ = ∆k := P{0C}⊕k =(cid:20)0 0 0 Ik(cid:21) ∈ B(bk) is ubiquitous; below it is freely ampliated. In this section we collect the quantum stochastic (QS) facts needed below. For more detail, see [L]; for further background, see [Par] and [Mey]. Let F denote the symmetric Fock space over K. We use normalised exponential vectors (g) := e−kgk2/2ε(g) where ε(g) :=(cid:0)(n!)−1/2g⊗n(cid:1)n>0 in terms of which the Fock–Weyl operators are the unitary operators on F determined by the identity W (f )(g) = e−i Imhf,gi(f + g) (f, g ∈ K), and the second quantisation operators are the contractions from F to F ′ determined by the identity Γ(C)ε(g) := ε(Cg) (2.1) where F ′ is the symmetric Fock space over a Hilbert space K′. Note that Γ(C) is isometric if C is, and Γ(C ∗) = Γ(C)∗. There is a useful family of slice maps: Ω(g′, g) := idB(h;h′) ⊗ ω(g′),(g) : B(h; h′)⊗ B(F) → B(h; h′) (g′, g ∈ K), (C ∈ B(K; K′)1, g ∈ K), (g ∈ K), amongst which E := Ω(0, 0) is referred to as the vacuum expectation. (The Hilbert spaces h and h′ are determined by context.) Moreover, the action of the slice maps extends to the class of unbounded operators from h⊗ F to h′ ⊗ F whose domains include h⊗ E, where E := Lin{(g) : g ∈ K}. The following obvious identity is exploited below: We also need two families of endomorphisms of (the von Neumann algebra) B(F): Ω(g′, g)(T ) = E(cid:2)(Ih′ ⊗ W (g′))∗ T (Ih ⊗ W (g))(cid:3). (2.2) ρt : T 7→ RtT Rt and σt : T 7→ IF[0,t[ ⊗ StT S∗ t (t ∈ R+), QS LIE–TROTTER PRODUCT FORMULA 9 where Rt := Γ(rt) and St := Γ(st) for the unitary operators rt and st defined as follows rt : K → K, (rtf )(s) =(f (t − s) f (s) if s ∈ [0, t[ if s ∈ [t, ∞[ , st : K → K [t,∞[, (stf )(s) = f (s − t), where K[t,∞[ := L2([t, ∞[; k). The time-reversal maps (cid:0)ρt(cid:1)t>0 are obviously involutive: ρ2 t = idB(F ), and the time-shift maps (σt)t∈R+ form a semigroup, known as the CCR flow of index k. Both are freely ampliated to act on B(h; h′)⊗ B(F), for Hilbert spaces h and h′. For use below, note the identity r(g′[r,∞[), s∗ Ω(cid:0)g′ [r,∞[, g[r,∞[(cid:1) ◦ σr = Ω(cid:0)s∗ Let X i =(cid:0)X i 0 + Λt(H i)(cid:1)t>0 for a QS integrand (hi, hi−1)-process H i and bounded op- erator X i 0 ∈ B(hi; hi−1) ⊗ IF , where i = 1, 2. Suppose that all of the processes H 1, H 2, X 1 and X 2 are bounded, that is they consist of bounded operators, and H 1 and H 2 are strongly continuous. Then the quantum Ito product formula ([HuP]) reads (g, g′ ∈ K, r ∈ R+). (2.3) r(g[r,∞[)(cid:1) X 1 0 X 2 t X 2 t = X 1 with ∆ abbreviating ∆ ⊗ Ih1⊗F . 0 + Λt(H) where H =(cid:0)H 1 s(cid:1)s>0, (2.4) Let X =(cid:0)Λt(H)(cid:1)t>0 for a bounded QS integrand (h1, h2)-process H with block matrix s ) + (Ibk ⊗ X 1 s (Ibk ⊗ X 2 ∗ ∗ ], then form [ K ∗ s + H 1 s ∆H 2 s )H 2 ds E[Ks]v (v ∈ h1, t ∈ R+), (2.5) E[Xt] v =Z t 0 and, if the process X is bounded then, for bounded operators R ∈ B(h2; h3) ⊗ IF and S ∈ B(h0; h1) ⊗ IF , RΛt(H)S = Λt(cid:0)(Ibk ⊗ R)H·(Ibk ⊗ S)(cid:1) (t ∈ R+). (2.6) A bounded QS (left) cocycle on h, with noise dimension space k, is a bounded process V on h which satisfies V0 = Ih⊗F and Vs+t = Vsσs(Vt) (s, t ∈ R+). It is a QS right cocycle if instead it satisfies Vs+t = σs(Vt)Vs for s, t ∈ R+. If V is a QS left cocycle then, V ∗ := (V ∗ t )t>0 and V r :=(cid:0)ρt(Vt)(cid:1)t>0 define QS right cocycles, and (2.7) V ♯ :=(cid:0)ρt(V ∗ t ) = ρ(Vt)∗(cid:1)t>0 defines a QS left cocycle, called the dual cocycle of V ([Jou]). This said, we work exclusively with QS left cocycles in this paper. Following standard terminology of semigroup theory ([HiP]), β0(V ) := inf{β ∈ R : supt>0 ke−βtVtk < ∞} If V is strongly continuous then is referred to as the exponential growth bound of V . β0(V ) < ∞; V is called quasicontractive if, for some β ∈ R, the QS cocycle (e−βtVt)t>0 is contractive. When V is locally uniformly bounded, it is called elementary (or Markov regular ) if its expectation semigroup(cid:0)E[Vt](cid:1)t>0 is norm continuous. Note that, given a QS cocycle V , the two-parameter family Vr,t := σr(Vt−r) satisfies V0,t = Vt (t ∈ R+), the evolution equation Vr,r = Ih⊗F and Vr,t = Vr,sVs,t (t > s > r), and the biadaptedness property Vr,t ∈(cid:0)B(h) ⊗ I[0,r[(cid:1)⊗(cid:0)B(F[r,t[) ⊗ I[t,∞[(cid:1) (t > r > 0). 10 MARTIN LINDSAY For use below, note that in the notation (2.2), these QS evolutions satisfy the time- covariance identity Ω(c′ [r,t[, c[r,t[)(Vr,t) = Ω(c′ [0,t−r[, c[0,t−r[)(Vt−r) by (2.3), and further evolution identity for c′, c ∈ k, g′, g ∈ L2 Ω(g′ [r,t[, g[r,t[)(Vr,t) = Ω(g′ loc(R+; k) and t > s > r > 0. [r,s[, g[r,s[)(Vr,s)Ω(g′ [s,t[, g[s,t[)(Vs,t) (2.8a) (2.8b) Remark. Note that the relations (2.8a) and (2.8b) also hold when V is a product of such QS evolutions. equation dXt = Xt·F dΛt, X0 = I has a unique weakly regular weak solution, denoted X F , moreover X F is a quasicontractive, elementary QS cocycle, with exponential growth bound at most β, strongly satisfying its QS differential equation. The resulting map F 7→ X F is Theorem 2.1 ([Fag], [LW1,2). ] For F ∈ qcβ(bk⊗ h) where β ∈ R, the (left) QS differential bijective from qc(bk ⊗ h) to the class of quasicontractive elementary QS cocycles; it restricts to bijections from c(bk ⊗ h), i(bk ⊗ h) and i(bk ⊗ h)∗ to the respective subclasses of contractive, isometric and coisometric, elementary QS cocycles, moreover it satisfies X F ∗ The unique operator F associated with a quasicontractive elementary QS cocycle V in = (X F )♯. this way is referred to as the stochastic generator of V . There is a basic class of QS cocycles which plays an important role. Example 2.2. For each c ∈ k, the (Fock) Weyl cocycle W c is given by Proposition 2.3. Let F ∈ qc(bk ⊗ h) and G ∈ qc(bk) ⊗ Ih ⊂ qc(bk ⊗ h). Then Proof. First note that, since X G is of the form (cid:0)Ih ⊗ Vt(cid:1)t>0 for a quasicontractive QS X F X G = X F ✁G and X G X F = X G✁F . cocycle V on C, s ) = F ⊗ Vs = (Ibk ⊗ X G Therefore, by the quantum Ito product formula (2.4), (F ⊗ IF )(Ibk ⊗ X G s )(F ⊗ IF ) (s ∈ R+). d(X F t X G t ) =(cid:2)(Ibk ⊗ X F t )(F ⊗ IF )(Ibk ⊗ X G t ) + (Ibk ⊗ X F t X G t )(F ⊗ IF )(Ibk ⊗ X G + (Ibk ⊗ X F = X F t X G t · (F ✁ G) dΛt. t )(G ⊗ IF ) t )(∆G ⊗ IF )(cid:3) dΛt 0 X G Since X F solutions of the QS differential equation 0 = I and F ✁ G ∈ qc(bk ⊗ h), uniqueness for strongly continuous bounded (see Theorem 2.1) implies that X F X G = X F ✁G. Since F ∗ ∈ qc(bk⊗h) and G∗ ∈ qc(bk)⊗Ih, too. The second identity now follows by duality: dXt = Xt · (F ✁ G) dΛt, X0 = Ih⊗F it follows that X F ∗✁G∗ = X F ∗ X G∗ =(cid:0)X F ∗✁G∗(cid:1)♯ =(cid:0)X F ∗ X G∗(cid:1)♯ =(cid:0)X G∗(cid:1)♯ (cid:0)X F ∗(cid:1)♯ = X G X F . X G✁F = X (F ∗✁G∗)∗ These are QS unitary cocycles (both left and right), and X = W c satisfies moreover X is elementary with stochastic generator W c :=(cid:0)Ih ⊗ W (c[0,t[)(cid:1)t>0 (c ∈ k). X r = X and X ♯ = W −c = X ∗, Fc :=(cid:20)− 1 2 kck2 −hc ci 0 (cid:21) ⊗ Ih ∈ B(bk ⊗ h). QS LIE–TROTTER PRODUCT FORMULA 11 (cid:3) Remarks. (i) More generally, if two quasicontractive QS cocycles commute on their initial space then their (pointwise) product is also a cocycle. (ii) In the light of the identities Fc ✁ Fd = Fc+d − i Imhc, di∆⊥ and F−c = F ∗ c (c, d ∈ k), Proposition 2.3 contains Weyl commutation relations (see [BrR], or [L]) as a special case. The connection to quantum random walks mentioned in the introduction is as follows (for details, in particular the precise meaning of the terminology used, see [BGL]). Theorem 2.4 ([BGL]). For i = 1, 2, let Fi ∈ qc(bk ⊗ h) and let (Gi(h))h>0 be a family in B(bk ⊗ h) satisfying sup(cid:8)kGi(h)kn : n ∈ N, h > 0, nh 6 T(cid:9) < ∞ (cid:0)h−1/2∆⊥ + ∆(cid:1)(cid:0)Gi(h) − ∆⊥(cid:1)(cid:0)h−1/2∆⊥ + ∆(cid:1) → Fi + ∆ as h → 0. Then, for all ϕ ∈ B(F)∗ and T ∈ R+, (T ∈ R+), and sup 06t6T(cid:13)(cid:13)(cid:0) idB(h) ⊗ ϕ(cid:1)(cid:0)X hhi t − X F1✁F2 t (cid:1)(cid:13)(cid:13) → 0 as h → 0, )t>0 denotes the h-scale embedded left quantum random walk where, for h > 0, (X hhi generated by G1(h)G2(h). t 3. Product Formula For proving the quantum stochastic Lie–Trotter product formula it is convenient to define a constant associated with a pair of quantum stochastic generators. Thus, for F1, F2 ∈ b(bk ⊗ h) with respective block matrix formsh K1 M1 C(F1, F2) : = k∆⊥F1∆⊥k + k∆⊥F2∆⊥k + k∆⊥F1∆k · k∆F2∆⊥k L1 N1i andh K2 M2 L2 N2i, set = kK1k + kK2k + kM1kkL2k. Note that this depends only on the first row of F1 and first column of F2: C(F1, F2) = C(∆⊥F1, F2∆⊥). Proposition 3.1. Let Fi ∈ qcβi(bk ⊗ h) with block matrix formh Ki Mi t − X F1✁F2 t X F2 t Li Nii, for i = 1, 2. Then (t ∈ R+). (cid:13)(cid:13)E(cid:2)X F1 (cid:3)(cid:13)(cid:13) 6 t2 et(β1+β2)C(F1, F2)2 Proof. Set V (1) = X F1, V (2) = X F2 and V = X F1✁F2. Since e−t(β1+β2)(V (1) t V (2) t − Vt) = X G1 t X G2 t − X G1✁G2 t (t ∈ R+), where Gi = Fi − βi∆⊥ ∈ c(bk ⊗ h) (i = 1, 2), it suffices to assume that β1 = β2 = 0, so that F1, F2, F1 ✁ F2 ∈ c(bk ⊗ h), and the QS cocycles V (1), V (2) and V are all contractive. Fix t ∈ R+. By the quantum Ito product formula (2.4), V (1)V (2) − V = Λ·(H) where H = H (1) + H (2) + H (3) for the processes given by := (Ibk ⊗ V (1) s )(F1 ⊗ IF )(Ibk ⊗ V (2) s ) − (Ibk ⊗ Vs)(F1 ⊗ IF ) s H (1) H (2) H (3) s s :=(cid:0)Ibk ⊗ (V (1) := (Ibk ⊗ V (1) s − Vs)(cid:1)(F2 ⊗ IF ) )(F1∆ ⊗ IF )(Ibk ⊗ V (2) s V (2) s s )(∆F2 ⊗ IF ) − (Ibk ⊗ Vs)(F1∆F2 ⊗ IF ). 12 MARTIN LINDSAY Therefore, by (2.5), E[V (1) processes given by t V (2) t − Vt] =R t 0 ds E[Ys] where Y = Y (1) + Y (2) + Y (3) for the (K1 ⊗ IF )V (2) s − Vs(K1 ⊗ IF ) Y (1) s Y (2) s Y (3) s s := V (1) := (V (1) := V (1) s s V (2) s − Vs)(K2 ⊗ IF ) (M1 ⊗ IF )(Ik ⊗ V (2) s )(L2 ⊗ IF ) − Vs(M1L2 ⊗ IF ). Note that, since V (1), V (2) and V are contractive, kE[Yr]k 6 2(cid:0)kK1k + kK2k + kM1kkL2k(cid:1) = 2C(F1, F2) In turn, the quantum Ito formula, together with the identities (2.5) and (2.6), imply that, for i = 1, 2, 3 and s ∈ R+, E[Y (i) r ] for processes given by (r ∈ R+). (3.1) ] =R s s 0 dr E[Z (i) r − Vr((K1)2 ⊗ IF )+ (K2 ⊗ IF ) − Vr(K1K2 ⊗ IF )+ ((K1)2 ⊗ IF )V (2) (K1 ⊗ IF )V (2) (M1 ⊗ IF )(Ik ⊗ K1 ⊗ IF )(Ik ⊗ V (2) r r Z (1) r := V (1) r V (1) r V (1) r r Z (2) Z (3) r := Yr(K2 ⊗ IF ), and := V (1) r V (1) r V (1) r (K1M1 ⊗ IF )(Ik ⊗ V (2) (M1 ⊗ IF )(Ik ⊗ V (2) (M1 ⊗ IF )(Ik ⊗ M1 ⊗ IF )(Ik⊗k ⊗ V (2) )(L2 ⊗ IF )V (2) r r )(L2 ⊗ IF ) − Vr(M1L2K1 ⊗ IF ), r )(Ik ⊗ K2 ⊗ IF )(L2 ⊗ IF ) − Vr(K2M1L2 ⊗ IF )+ r − Vr(K2M1L2 ⊗ IF )+ )(Ik ⊗ L2 ⊗ IF )(Ik⊗k ⊗ V (2) )(L2 ⊗ IF ) − Vr((M1L2)2 ⊗ IF ). r Now, by further use of the contractivity of the processes V (1), V (2) and V , together with the estimate (3.1), r kE[Z (1) kE[Z (2) kE[Z (3) r r and so ]k 62C(F1, F2)kK2k, and ]k 62(cid:0)kK1k2 + kK1kkK2k + kM1kkK1kkL2k(cid:1) = 2kK1kC(F1, F2), ]k 62(cid:0)kK1kkM1kkL2k + kM1kkK2kkL2k + kM1k2kL2k2(cid:1) = 2kM1kkL2kC(F1, F2), kE[V (1) t V (2) dr E[Z (1) r + Z (2) r + Z (3) r dsZ s t − Vt]k =(cid:13)(cid:13)(cid:13)Z t dsZ s 6Z t 0 0 0 0 ](cid:13)(cid:13)(cid:13) dr 2 C(F1, F2)2 = t2 C(F1, F2)2, as required. (cid:3) r,t Then r,t X F2 r,t − X F1✁F2 Lemma 3.2. Let F1 ∈ qcβ1(bk ⊗ h), F2 ∈ qcβ2(bk ⊗ h), c′, c ∈ k and r, t ∈ R+ with t > r. (cid:13)(cid:13)Ω(cid:0)c′ (cid:1)(cid:13)(cid:13) 6 (t − r)2 e(t−r)(β1+β2) C(cid:0)F ∗ c′ ✁ F1, F2 ✁ Fc(cid:1)2 [r,t[, c[r,t[(cid:1)(cid:0)X F1 Proof. By definition of the evolutions (cid:0)X G r,t(cid:1)06r6t, for G = F1, F2 and F1 ✁ F2, and (cid:1)(cid:13)(cid:13) t−r − X F1✁F2 In this case, by the and so we may suppose without loss of generality that r = 0. identity (2.2), Example 2.2 and the associativity of ✁, the estimate follows immediately from Propositions 2.3 and 3.1. (cid:3) [0,t−r[, c[0,t−r[(cid:1)(cid:0)X F1 LHS =(cid:13)(cid:13)Ω(cid:0)c′ identity (2.8a) along with the remark following it, t−rX F2 t−r . QS LIE–TROTTER PRODUCT FORMULA 13 Now let S denote the subspace of L2(R+; k) consisting of step functions, whose right- Then C(g′, F1, F2, g) := maxnC(cid:0)F ∗ continuous versions we use when evaluating. For F1, F2 ∈ b(bk ⊗ h) and g′, g ∈ S, set c′ ✁ F1, F2 ✁ Fc(cid:1) : c′ ∈ Ran g′, c ∈ Ran go. Lemma 3.3. Let F1 ∈ qcβ1(bk ⊗ h), F2 ∈ qcβ2(bk ⊗ h), g′, g ∈ S and r, t ∈ R+ with t > r. (cid:13)(cid:13)Ω(cid:0)g′ (cid:1)(cid:13)(cid:13) 6 (t − r)2 e(t−r)(β1+β2) C(g′ Proof. Set β := β1 + β2 and let {r = t0 < t1 < · · · < tn < tn+1 = t} be such that g and g′ are constant, say ci and di respectively, on the interval [ti, ti+1[, and, for i = 0, · · · , n, set Ωi := Ω(g′ [r,t[, g[r,t[(cid:1)(cid:0)X F1 [r,t[, F1, F2, g[r,t[)2. r,t − X F1✁F2 r,t X F2 r,t [ti,ti+1[, g[ti,ti+1[) X 1 i := X F1 ti,ti+1 , X 2 i := X F2 ti,ti+1 and Xi := X F ti,ti+1 , noting that Ωi := Ω(ci sequent remark, we see that [ti,ti+1[, di [ti,ti+1[). Then, again using identity (2.8b) and the sub- −→Y06i6n Ωi(X 1 i X 2 i ) − −→Y06i6n Ωi(Xi). Therefore, using Lemma 3.2, r,t r,t X F2 (cid:1) = r,t − X F1✁F2 Ω(cid:0)g′ [r,t[, g[r,t[(cid:1)(cid:0)X F1 LHS =(cid:13)(cid:13)(cid:13) nXi=0(cid:16) −→Y06p<i nXi=0 eβ(ti−r)(cid:13)(cid:13)Ωi(X 1 nXi=1 (ti+1 − ti)2C(cid:0)F ∗ 6eβ(t−r) i X 2 Ωp(Xp) Ωi(X 1 6 i X 2 i − Xi) −→Yi<p6n i − Xi)(cid:13)(cid:13)eβ(t−ti+1) ci ✁ F1, F2 ✁ Fdi(cid:1)2 Ωp(X 1 p X 2 p )(cid:17)(cid:13)(cid:13)(cid:13) 6 RHS . (cid:3) By a partition P of R+ we mean a sequence (sn) in R+ which is strictly increasing and tends to infinity. Where convenient we identify a partition with its set of terms. For S ⊂⊂ R+, let S denote the mesh of {0} ∪ S, that is, max{ti − ti−1 : 1 6 i 6 k}, where S = {t1 < · · · < tn} and t0 := 0. Let P be a partition of R+, let {t1 < · · · < tN −1} = P ∩ ]r, t[ and set t0 := r and tN := t. Then Lemma 3.4. Let F1 ∈ qcβ1(bk ⊗ h), F2 ∈ qcβ2(bk ⊗ h), g′, g ∈ S and r, t ∈ R+ with t > r. (cid:13)(cid:13)(cid:13)Ω(cid:0)g′ [r,t[, g[r,t[(cid:1)(cid:16) −→Y16j6N Proof. Set β := β1 + β2 and, for j = 1, · · · , N , set Ωj := Ω(cid:0)g′ 6(cid:12)(cid:12)P ∩ [r, t](cid:12)(cid:12) (t − r) e(t−r)(β1+β2) C(g′ [tj−1,tj[, g[tj−1,tj[(cid:1), [r,t[, F1, F2, g[r,t[)2. tj−1,tj − X F1✁F2 (cid:17)(cid:13)(cid:13)(cid:13) X F1 X F2 tj−1,tj r,t X 1 j := X F1 tj−1,tj , X 2 j := X F1 tj−1,tj and Xj := X F1✁F2 tj−1,tj . 14 MARTIN LINDSAY Then, using Lemma 3.3 (and arguing as in its proof), we see that Ωn(X 1 nX 2 n)(cid:17)(cid:13)(cid:13)(cid:13) LHS =(cid:13)(cid:13)(cid:13) NXj=1(cid:16) −→Y16n<j NXj=1 eβ(tj−1−r)(cid:13)(cid:13)Ωj(X 1 NXj=1 6eβ(t−r) 6 Ωn(Xn) Ωj(X 1 j X 2 j − Xj) −→Yj<n6N j − Xj)(cid:13)(cid:13)eβ(t−tj+1) j X 2 (tj − tj−1)2C(g′ [tj−1,tj[, F1, F2, g[tj−1,tj [)2 6eβ(t−r)(cid:12)(cid:12)P ∩ [r, t](cid:12)(cid:12) (t − r) C(g′ [r,t[, F1, F2, g[r,t[)2, as required. We may now prove the main result. Theorem 3.5. Let V 1 and V 2 be quasicontractive elementary QS cocycles on h with noise dimension space k and respective stochastic generators F1 and F2, and let T ∈ R+ and ϕ ∈ B(F)∗. Then sup 06r6t6T(cid:13)(cid:13)(cid:0) idB(h) ⊗ ϕ(cid:1)(cid:0)V P,1✁2 r,t − X F1✁F2 r,t (cid:1)(cid:13)(cid:13) → 0 as (cid:12)(cid:12)P ∩ [0, T ](cid:12)(cid:12) → 0. The notation here is as follows. For a partition P of R+ and 0 6 r 6 t, V P,1✁2 r,t := −→Y16j6N V 1 tj−1,tj V 2 tj−1,tj in which {t1 < · · · < tN −1} = P ∩ ]r, t[, t0 := r and tN := t. Proof. Set β := β0(V 1) + β0(V 2) and V := X F1✁F2. We may suppose without loss of generality that ϕ is of the form ω(h′),(h) for some h′, h ∈ S, since such functionals are total in the Banach space B(F)∗ and (V P,1✁2 − Vr,t) is uniformly bounded by 2eT max{0,β} for [r, t] ⊂ [0, T ]. r,t For any partition P of R+, Lemma 3.4 implies that there is a constant C = C(ϕ, F1, F2) such that, for all subintervals [r, t] of [0, T ], (cid:3) (cid:3) Proposition 3.1 and Lemmas 3.2, 3.3 and 3.4 extend to the case of QS generators The result follows. Fi =h Ki Mi where r,t (cid:13)(cid:13)(cid:0) idB(h) ⊗ ϕ(cid:1)(cid:0)V P,1✁2 − Vr,t(cid:1)(cid:13)(cid:13) 6(cid:12)(cid:12)P ∩ [0, T ](cid:12)(cid:12) T eT β C 2. Li Nii (i = 1, · · · , n), the constant C(F1, F2) being replaced by 0 0(cid:21) ∈ B(bC). fi =(cid:20)kKik kMik ∆⊥f1 ✁(cid:0)f2 ✁ · · · ✁ fn−1(cid:1) ✁ fn ∆⊥ kNik(cid:21) ∈ B(bC) and ∆⊥ = ∆⊥ − I =Z t X F1✁···✁Fn ·(cid:0)F1 ✁ · · · ✁ Fn(cid:1) dΛs. C =(cid:20)1 0 X F1✁···✁Fn kLik 0 s t In this case the quantum Ito product formula allows X F1 t − I to be expressed as a sum of 2n − 1 QS integrals, each of which may be paired with one of the 2n − 1 terms arising from the expansion of F1 ✁ · · · ✁ Fn in the identity · · · X Fn t QS LIE–TROTTER PRODUCT FORMULA 15 As a consequence, Theorem 3.5 extends to quasicontractive elementary QS cocycles V 1, · · · , V n, with respective stochastic generators F1, · · · , Fn, as follows: sup 06r6t6T(cid:13)(cid:13)(cid:0) idB(h) ⊗ ϕ(cid:1)(cid:0)V P,1✁···✁n r,t in the notation − X F1✁···✁Fn r,t (cid:1)(cid:13)(cid:13) → 0 as (cid:12)(cid:12)P ∩ [0, T ](cid:12)(cid:12) → 0, V P,1✁···✁n r,t := V 1 tj−1,tj · · · V n tj−1,tj . −→Y16j6N Remarks. (i) Note that the QS cocycle V = X F1✁···✁Fn is contractive (respectively, iso- metric, or coisometric) provided that all of V 1, · · · , V n are. (ii) Whilst the convergence of Trotter products holds in the above hybrid norm-ultraweak topology, if V is isometric and V 1, · · · , V n are contractive (in particular, if V 1, · · · , V n are isometric) then convergence also holds in the strong operator topology: sup 06r6t6T(cid:13)(cid:13)(cid:0)V P,1✁···✁n r,t − X F1✁···✁Fn r,t (cid:1)ξ(cid:13)(cid:13) → 0 as (cid:12)(cid:12)P ∩ [0, T ](cid:12)(cid:12) → 0 This follows from the uniform continuity of the function (T ∈ R+, ξ ∈ h ⊗ F). [0, T ]2 6 → h ⊗ F , (r, t) 7→ Vr,tξ (ξ ∈ h ⊗ F , T ∈ R+). We now revisit the case of QS cocycles with independent driving noise and show how to view it as a special case of the above theorem. Suppose therefore that the noise dimension space has orthogonal decomposition k = k1 ⊕ · · · ⊕ kn. Denoting the symmetric Fock space over L2(R+; ki) by F i, second quantisation of the natural isometry from L2(R+; ki) to L2(R+; k) (see (2.1)), followed by ampliation, gives an isometry h ⊗ F i → h ⊗ F which in turn induces a normal *-algebra monomorphism Υi : B(h ⊗ F i) → B(h ⊗ F) (i = 1, · · · , n). Corollary 3.6 ([LS1]). For i = 1, · · · , n, let V i be a quasicontractive elementary QS cocycle on h with noise dimension space ki and stochastic generator Fi. Then, for all T ∈ R+ and ϕ ∈ B(F)∗, The notation here is as follows. For a partition P of R+ and 0 6 r 6 t, setting {t1 < · · · < tN −1} = P ∩ ]r, t[, t0 := r and tN := t, − X F1⊞ ··· ⊞Fn r,t (cid:1)(cid:13)(cid:13) → 0 as (cid:12)(cid:12)P ∩ [0, T ](cid:12)(cid:12) → 0. r,t sup 06r6t6T(cid:13)(cid:13)(cid:0) idB(h) ⊗ ϕ(cid:1)(cid:0)V P,1⊞ ··· ⊞n −→Y16j6N V P,1⊞ ··· ⊞n r,t := Υ1(V 1 tj−1,tj ) · · · Υn(V n tj−1,tj ), s,s′) = X ι′(F2) Proof. Suppose first that n = 2. Note that, in the notation (1.7), Υ1(V 1 and (s, s′ ∈ R+, s < s′). Therefore, in view of identity (1.8), for this case Υ2(V 2 the corollary follows immediately from Theorem 3.5. The general case follows similarly via obvious extension of the notation (1.7) and corresponding identity (1.8). (cid:3) s,s′) = X ι(F1) s,s′ s,s′ 4. Maximal Gaussian component of a QS generator In this section it is shown that every QS generator F ∈ qc(bk ⊗ h) enjoys a unique decomposition F1 ⊞ F2 in which F2 is 'pure Gaussian' and F1 is 'wholly non-Gaussian', in senses to be defined below. This amounts to extracting a maximal Gaussian component of the generator and demonstrating its uniqueness. It chimes well with the way that Hunt's formula decomposes the generator of a L´evy process on a compact Lie group into a maximal Gaussian component and a jump part (albeit not exactly uniquely). 16 MARTIN LINDSAY To each contraction C ∈ B(k ⊗ h) we associate the following closed subspaces of k: and kC g :=(cid:8)c ∈ k : (C − I)Ec = 0(cid:9), where Ec := ci ⊗ Ih, p := Lin(cid:8)(Ik ⊗ hu)ξ : ξ ∈ Ran(C − I)∗, u ∈ h(cid:9). kC The former captures the subspace of all 'directions' in which C acts as the identity op- erator; the latter is complementary (see below). The subscripts denote 'Gaussian' and 'preservation' parts. Letting J C p denote the inclusion kC p → k, set Cp := (J C p ⊗ Ih)∗C(J C p ⊗ Ih), the compression of C to kC p ⊗ h. Lemma 4.1. Let C ∈ B(k ⊗ h) be a contraction. Then the following hold: (a) k = kC p ⊕ kC g . (b) C has block matrix formh Cp 0 g i, where I C 0 I C p = kC p . g := IkC g ⊗h. (c) (kC p )Cp Proof. Set I C p := IkC p ⊗h. (a) Let c ∈ k. Then, for all u ∈ h, hc, (Ik ⊗ hu)(C − I)∗ξi = 0 for all ξ ∈ k ⊗ h ⇐⇒ (C − I)(c ⊗ u) = 0. It follows that c ∈ (kC p )⊥ if and only if c ∈ kC g , so (a) follows. (b) Contraction operators T on a Hilbert space H enjoy the following elementary prop- erty. In terms of the orthogonal decomposition H = K⊥ ⊕ K where K := Ker(T − I), T has g ⊗ h ⊂ Ker(C − I) it follows that, in terms p ⊗ h) ⊕ (kC g ⊗ h), C has the claimed block diagonal block matrix form(cid:2) ∗ 0 0 IK(cid:3). Since kC of the orthogonal decomposition k ⊗ h = (kC matrix form. (c) By (a) it suffices to show that (kC p )Ee = 0 and so, by (b), (C − I)EJ C p ∩ kC g = {0}, by (a). Since J C p )Cp g = {0}. Accordingly let e ∈ (kC p e = 0, in other words J C g . Then g . Thus p is an inclusion this implies that e = 0, as (cid:3) p )Cp p e ∈ kC (Cp − I C p e ∈ kC J C required. ∗ [ ∗ ∗ ∗ 0 ]; ∗ C−I ]. We say that (a) F is Gaussian if kC (b) F is wholly non-Gaussian if kC g = k, equivalently C = I so that F has block matrix form p = k, equivalently Cp = C; 1 2 L∗L −L∗ L Definition 4.2. Let F ∈ qc(bk ⊗ h) with block matrix form [ ∗ 0 i, for some L ∈ B(h; k⊗ (c) F is pure Gaussian if it has block matrix formh − 0 i, for some M ∈ B(k ⊗ h; h); h), equivalentlyh − 0 C−I(cid:3); (d) F is pure preservation if it has block matrix form(cid:2) 0 g(bk ⊗ h), wn-g(bk ⊗ h), pg(bk ⊗ h) and pp(bk ⊗ h), for the respective classes of Gaussian, wholly non-Gaussian, pure Gaussian and pure preservation generator. 0 0 ], in other words F ∈ z(bk ⊗ h). (e) F is pure drift if it has block matrix form [ ∗ 0 1 2 M M ∗ M −M ∗ Write 0 Remarks. (i) The terminology has its origins in the work of Schurmann on L´evy–Khintchin- type decompositions for L´evy processes on bialgebras such as algebraic quantum groups QS LIE–TROTTER PRODUCT FORMULA 17 ([Sch]). In the presence of a minimality condition, the Gaussian property for the gener- ator of an elementary unitary QS cocycle U may alternatively be expressed in terms of the cocycle itself as follows: t−1E(cid:2)(U 1 t − I)u1 v1 (U 2 t − I)u2 v2 (U 3 t − I)u3 v3(cid:3) → 0 as t → 0 for some K ∈ B(h) and L ∈ B(h; k ⊗ h). for all u1, · · · , v3 ∈ h and all choices of U 1, U 2 and U 3 from {U, U ∗} where, for X ∈ B(h ⊗ F) and u, v ∈ h, X u v := (ωu,v⊗ idB(F ))(X) = (hu ⊗ IF )X(vi ⊗ IF ) (see [SSS] and [Sch]). determines a decomposition F = F1 ⊞ F ′ (ii) By Theorem 1.3 (a), F ∈ g(bk ⊗ h) if and only if it has block matrix form(cid:2) K −L∗ L 0 (cid:3), (ii) Given a Gaussian QS generator(cid:2) K −L∗ L 0 (cid:3), any orthogonal decomposition k = k1 ⊕ k2 decomposition L =h L1 L2i, F1 is pure Gaussian with F1 =(cid:20)− 1 0 (cid:21) and F ′ 1 in which, corresponding to the block matrix 1 =(cid:20)K + 1 0 (cid:21) . 1L1 −L∗ 1 1L1 −L∗ 2 2 L∗ L2 2 L∗ L1 Since L∗L = L∗ 1L1 + L∗ 2L2, F ′ 1 is pure Gaussian too if and only F is. (iii) In the case of no noise (k = {0}), pp(bk ⊗ h) = ∅ and g(bk ⊗ h), wn-g(bk ⊗ h), pg(bk ⊗ h) and z(bk ⊗ h) all equal b(bk ⊗ h). Otherwise, when k 6= {0}, all of g(bk ⊗ h), wn-g(bk ⊗ h), pg(bk ⊗ h), pp(bk ⊗ h) and z(bk ⊗ h) are selfadjoint and the following relations are easily verified: Also, in view of the series decompositions (Proposition 1.4), g(bk ⊗ h) ∩ wn-g(bk ⊗ h) = ∅; pg(bk ⊗ h) ∩ z(bk ⊗ h) = {0} = g(bk ⊗ h) ∩ pp(bk ⊗ h); pg(bk ⊗ h) ⊂ u(bk ⊗ h) and pp(bk ⊗ h) ⊂ c(bk ⊗ h); g(bk ⊗ h) ✁ g(bk ⊗ h) = g(bk ⊗ h) and pp(bk ⊗ h) ✁ pp(bk ⊗ h) = pp(bk ⊗ h); z(bk ⊗ h) ✁ pg(bk ⊗ h) = g(bk ⊗ h) = pg(bk ⊗ h) ✁ z(bk ⊗ h). z(bk ⊗ h) ✁ pg(bk ⊗ h) ✁ pp(bk ⊗ h) ⊃ i(bk ⊗ h), and pp(bk ⊗ h) ✁ pg(bk ⊗ h) ✁ z(bk ⊗ h) ⊃ i(bk ⊗ h)∗. qc(bk ⊗ h) = [k=k1⊕k2 wn-g(bk1 ⊗ h) ⊞ pg(bk2 ⊗ h). The next result implies that Theorem 4.3. Let F ∈ qc(bk ⊗ h). Then F enjoys a unique decomposition Fwn-g ⊞ Fmg where, for some orthogonal decomposition k1 ⊕ k2 of k, LC g := (J C g ⊗ Ih)∗L, LC Proof. Let(cid:2) K M Fwn-g ∈ wn-g(bk1 ⊗ h) and Fmg ∈ pg(bk2 ⊗ h). L C−I(cid:3) be the block matrix decomposition of F , and set Fwn-g =(cid:20)K + 1 Cp − I(cid:21) and Fmg =(cid:20)− 1 Then F = Fwn-g ⊞ Fmg where 2 (LC LC p p ⊗ Ih)∗L, M C g := M (J C p := (J C g M C p g )∗LC g )∗LC 2(LC LC g g ⊗ Ih) and M C g M C g 0 (cid:21) . p := M (J C p ⊗ Ih). 18 MARTIN LINDSAY By Remark (ii) above, M C that Fwn-g is wholly non-Gaussian. This proves existence. g = −(LC g )∗ so Fmg is pure Gaussian. It follows from Lemma 4.1 For uniqueness, suppose that F = F1 ⊞ F2 where F1 ∈ wn-g(bk1 ⊗ h) and F2 ∈ pg(bk2 ⊗ h) for an orthogonal decomposition k = k1 ⊕ k2. Then F1 and F2 have block matrix decompositions moreover 2L2 −L∗ 2 2 L∗ L2 0 (cid:21) , F1 =(cid:20)K1 M1 L1 C1 − I(cid:21) and F2 =(cid:20)− 1 0(cid:1) ∈ kC p = k1 and kC kC1 g ⊃ k2. g . Then (C − I)E(e1 Suppose that e1 ∈ k1 and(cid:0)e1 F1 is wholly non-Gaussian this implies that e1 = 0. Thus kC that C1 = Cp and L2 = LC 0 ) = 0, so (C1 − I)Ee1 = 0. Since p = k1. It follows g . This implies that F1 = Fwn-g and F2 = Fmg, as required. (cid:3) g = k2 and kC Remarks. (i) Clearly Fmg is the maximal pure Gaussian component of F . (ii) Let F ∈ i(bk ⊗ h). Then the above decomposition of F takes the form 0 (cid:21) ∈ i(bk1 ⊗ h) ⊞ u(bk2 ⊗ h) L1 W − I(cid:21) ⊞(cid:20)− 1 F =(cid:20)K1 −L∗ 1L1, W isometric and kW 2L2 −L∗ 2 2 L∗ L2 with Re K1 = − 1 1W 2 L∗ p = k1; W being unitary if and only if F ∈ u(bk ⊗ h). 5. Perturbation of QS cocycles In case the second of two elementary QS cocycles V 1, V 2 is isometric, there is another, more standard, way of realising the QS cocycle X F1✁F2 where F1 and F2 are the stochastic generators of V 1 and V 2. Theorem 5.1 (Cf. [EvH], [BLS]). Let F1 ∈ qc(bk ⊗ h) and F2 ∈ i(bk ⊗ h) Then X F1✁F2 = X j2,F1X F2 where j2 is the normal *-monomorphic QS mapping cocycle on (the von Neumann algebra) B(h) given by t (x) = X F2 j2 t (x ⊗ IF )(X F2 t )∗ (x ∈ B(h), t ∈ R+), and X j2,F1 is the unique strong solution of the QS differential equation dXt = Xt · Gt dΛt, X0 = I, for the integrand process G :=(cid:0)(idB(bk) ⊗ j2 t )(F1)(cid:1)t>0. Proof. Given the existence of X j2,F1 and its quasicontractivity ([BLS]), the result follows easily from the quantum Ito product formula and uniqueness for weak solutions, which are bounded with locally uniform bounds, of the QS differential equation dXt = Xt · (F1 ✁ F2) dΛt, X0 = Ih⊗F (see Theorem 2.1). (cid:3) Remark. In [BLS] we worked in the equivalent category of QS right cocycles. Applying this result to the Gaussian/non-Gaussian decomposition of QS generators (Theorem 4.3) yields the following result. Recall the injections (1.7) associated with realising the concatenation product in terms of the series product (1.8). Corollary 5.2. Let F ∈ qc(bk ⊗ h) with block matrix form [ ∗ X F = X j2,F1X F2 , where, for the orthogonal decomposition k ⊗ h = (kC p ⊗ h) ⊕ (kC g ⊗ h), ∗ ∗ C−I ]. Then F1 = ι(Fwn-g) and F2 = ι′(Fmg). QS LIE–TROTTER PRODUCT FORMULA Proof. The inclusions ensure that X F2 is unitary and so Theorem 5.1 applies. ι′(cid:0)pg(ckC g ⊗ h)(cid:1) ⊂ pg(bk ⊗ h) ⊂ u(bk ⊗ h) 19 (cid:3) 6. Holomorphic QS Cocycles In this section the setting is extended to holomorphic QS cocycles ([LS2]). Before formulating the conjecture, the corresponding Ito algebra is investigated, mirroring Sec- tion 1. As is customary, we identify each bounded Hilbert space operator T with its associated quadratic form qT , given by qT [ξ] := hξ, T ξi. We also use the notation q(·, ·) for the sesquilinear form associated with a quadratic form q[ · ] by polarisation. Fix Hilbert spaces h and H. Let Q(h ⊕ H) denote the class of quadratic forms Γ on h ⊕ H having the following structure: from h to H, and (cid:26) Dom Γ = D ⊕ H Γ[ξ] = γ[u] −(cid:2)hζ, Lui + heLu, ζi + hζ, (C − I)ζi(cid:3), Dom γ = Dom L = DomeL = D. where D is a subspace of h, C ∈ B(H), γ is a quadratic form on h, L and eL are operators Write Γ ∼ (γ, L,eL, C), and refer to (γ, L,eL, C) as the components of Γ. Also define an For reasons which will become apparent, D is not assumed to be dense in h. ζ(cid:1) ∈ Dom Γ, for ξ =(cid:0)u associated operator on h ⊕ H by F ∆ Γ :=(cid:20)0 L C − I(cid:21) . 0 Thus Dom F ∆ Γ = Dom Γ and Ran F ∆ Γ ⊂ {0} ⊕ H = Ran ∆ where, as usual, ∆ := 0h ⊕ IH. The inclusion obviously implies that Γ ⊂ D′ ⊕ H for any subspace D′ of h. Ran F ∆ (6.1) Note that if Γ ∈ Q(h ⊕ H) with components (γ, L,eL, C), then the adjoint form Γ∗ belongs to Q(h ⊕ H) too, with hξ, F ∆ Γ ξi = −Γ(∆ξ, ξ) and hF ∆ Γ∗ ξ, ξi = −Γ(ξ, ∆ξ) (ξ ∈ Dom Γ). (6.2) Definition 6.1. For Γi ∈ Q(h ⊕ H), with components (γi, Li,eLi, Ci) (i = 1, 2), define Γ1 ✁ Γ2, Γ1∆Γ2 ∈ Q(h ⊕ H) by Γ1 ✁ Γ2 ∼ (γ, L,eL, C), where γ[u] = γ1[u] + γ2[u] − heL1u, L2ui, L = L1 + C1L2, eL = C ∗ Γ1∆Γ2 ∼ (γ, L,eL, C), where γ[u] = −heL1u, L2ui, L = (C1 − I)L2, eL = (C ∗ 2eL1 +eL2 and C = C1C2; 2 − I)eL1 and C = (C1 − I)(C2 − I) + I. Thus, in terms of the associated sesquilinear form, Γ∗ ∼ (γ∗,eL, L, C ∗) and F ∆ Γ∗ =(cid:20)0 0 eL C ∗ − I(cid:21) . 20 Thus MARTIN LINDSAY Γ1 ✁ Γ2 = Γ1 + Γ2 + Γ1∆Γ2, Dom(Γ1 ✁ Γ2) = Dom(Γ1∆Γ2) = Dom Γ1 ∩ Dom Γ2, (Γ1∆Γ2)[ξ] = −hF ∆ Γ∗ 1 F ∆ ξ, F ∆ Γ1∆Γ2 =(cid:0){0} ⊕ (C1 − I)(cid:1)F ∆ Γ2 ξi Γ2 . (ξ ∈ Dom Γ1 ∩ Dom Γ2), and (6.3a) (6.3b) (6.3c) Lemma 6.2. The prescription (Γ1, Γ2) 7→ Γ1∆Γ2 defines an associative and bilinear composition on the vector space Q(h ⊕ H) which is also involutive: (Γ1∆Γ2)∗ = Γ∗ 2∆Γ∗ 1. (6.4) Proof. Let Γi ∈ Q(h ⊕ H) with domain Di ⊕ H (i = 1, 2, 3). Bilinearity follows from the evident linearity of the map Γ 7→ F ∆ Γ , and (6.4) holds since, for ξ ∈ (D1 ∩ D2) ⊕ H, (Γ1∆Γ2)∗[ξ] = (Γ1∆Γ2)[ξ] = −hF ∆ Γ2 ξ, F ∆ Γ∗ 1 ξi = (Γ∗ 2∆Γ∗ 1)[ξ]. Clearly Dom(cid:0)(Γ1∆Γ2)∆Γ3(cid:1) =(cid:0)D1 ∩ D2 ∩ D3(cid:1) ⊕ H = Dom(cid:0)Γ1∆(Γ2∆Γ3)(cid:1) and, for ξ ∈ (D1 ∩ D2 ∩ D3) ⊕ H, (6.3c) implies that Γ2∆Γ3 ξi = hF ∆ Γ∗ 1 hF ∆ Γ∗ 1 ξ, F ∆ ξ,(cid:0){0} ⊕ (C2 − I)(cid:1)F ∆ 2 − I)(cid:1)F ∆ Γ3 ξi = hF ∆ = h(cid:0){0} ⊕ (C ∗ = hF ∆ Γ∗ ξ, F ∆ Γ∗ 1 2∆Γ∗ 1 ξ, F ∆ (Γ1∆Γ2)∗ ξ, F ∆ Γ3 ξi Γ3 ξi Γ3 ξi. Thus ∆ is associative, by (6.3b). (cid:3) Proposition 6.3. The composition (Γ1, Γ2) 7→ Γ1 ✁Γ2 endows Q(h⊕H) with the structure of a *-monoid whose identity element is Γ0 ∼ (0, 0, 0, I), in particular, (Γ1 ✁ Γ2)∗ = Γ∗ 2 ✁ Γ∗ 1. Proof. Let Γ1, Γ2, Γ3 ∈ Q(h ⊕ H). In view of Lemma 6.2, (Γ1 + Γ2 + Γ3) + (Γ1∆Γ2 + Γ2∆Γ3 + Γ1∆Γ3) + Γ1∆Γ2∆Γ3 is a common expression for Γ1 ✁ (Γ2 ✁ Γ3) and (Γ1 ✁ Γ2) ✁ Γ3, which have common domain (D1 ∩ D2 ∩ D3) ⊕ H. It is easily seen that the element Γ0 ∼ (0, 0, 0, I) satisfies Γ0 ✁ Γ = Γ = Γ ✁ Γ0 for all Γ ∈ Q(h ⊕ H). The fact that the adjoint operation defines an involution on the resulting monoid follows from its additivity on Q(h ⊕ H) and the identity (6.4). (cid:3) For the following lemma, recall the range observation (6.1). Lemma 6.4. Let Γi ∈ Q(h ⊕ H) with domain Di ⊕ H (i = 1, 2, 3). Then (Γ1 ✁ Γ2 ✁ Γ3)[ξ] = (Γ1 ✁ Γ3)[ξ] + Γ2(cid:0)(I + F ∆ Proof. Let ξ ∈ D1 ∩ D2 ∩ D3. Then, since Γ∗ 1 Γ3)ξ(cid:1) )ξ, (I + F ∆ (ξ ∈ D1 ∩ D2 ∩ D3). Γ1 ✁ Γ2 ✁ Γ3 − Γ1 ✁ Γ3 = Γ2 + Γ2∆Γ3 + Γ1∆Γ2 + Γ1∆Γ2∆Γ3, the lemma follows by several applications of the identities (6.2) and (6.3c): Γ∗ 1 )ξ, (I + F ∆ Γ3 )ξ(cid:1) Γ3 ξ(cid:1) + Γ2(cid:0)F ∆ Γ2(cid:0)(I + F ∆ Γ3 ξ(cid:1) = Γ2[ξ] + Γ2(cid:0)ξ, F ∆ = Γ2[ξ] +(cid:0)Γ2∆Γ3(cid:1)[ξ] +(cid:0)Γ1∆Γ2(cid:1)[ξ] +(cid:0)Γ1∆Γ2∆Γ3(cid:1)[ξ]. ξ, ξ(cid:1) + Γ2(cid:0)F ∆ Γ∗ 1 Γ2 ξi − hF ∆ Γ∗ 1 = Γ2[ξ] − hF ∆ Γ∗ 2 Γ3 ξi − hF ∆ Γ∗ 1 Γ∗ 1 ξ, F ∆ ξ, F ∆ ξ, F ∆ ξ,(cid:0){0} ⊕ (I − C2)(cid:1)F ∆ Γ3 ξi (cid:3) QS LIE–TROTTER PRODUCT FORMULA 21 Proposition 6.5. Let Γ1, Γ2 ∈ Q(h ⊕ H) and β1, β2 ∈ R, and set Γ = Γ1 ✁ Γ2 and β = β1 + β2. (a) Suppose that, for i = 1, 2, Γ∗ i ✁ Γi > 2βi∆⊥ on Dom Γi. Then Γ∗ ✁ Γ > 2β∆⊥ on Dom Γ. (b) Suppose that, for i = 1, 2, Γ∗ Then Γ∗ ✁ Γ = 0 on Dom Γ. i ✁ Γi = 0 on Dom Γi. Proof. By associativity and Lemma 6.4, for all ξ ∈ Dom Γ1∩Dom Γ2. The result therefore follows since ∆⊥F ∆ Γ2 = 0 on Dom Γ2. (cid:3) Remark. Thus (Γ∗ ✁ Γ)[ξ] = (Γ∗ 2 ✁ Γ2)[ξ] + (Γ∗ 1 ✁ Γ1)(cid:2)(I + F ∆ Γ2)ξ(cid:3) (cid:8)Γ ∈ Q(h ⊕ H) : Γ∗ ✁ Γ = 0 = Γ ✁ Γ∗(cid:9) forms a subgroup of the group of invertible elements of (Q(h ⊕ H), ✁). To complete the discussion of the algebra of the series product on quadratic forms, here is the form generalisation of Proposition 1.2. Proposition 6.6. Let Γ ∈ Q(h ⊕ H). Then (Γ∗ ✁ Γ)[ξ] = (Γ ✁ Γ∗)(cid:2)(I + F ∆ Γ )ξ(cid:3) +(cid:13)(cid:13)F ∆ Γ∗✁Γξ(cid:13)(cid:13)2 Let V ∈ Q(h ⊕ H) be of the form ν ⊕ 0H where ν ∈ Q(h)sa. Then Γ∗ ✁ Γ > V if and only if Γ ✁ Γ∗ > V. (ξ ∈ Dom Γ). (6.5) Proof. Let ξ =(cid:0)u share the same domain. On the one hand, setting Γ1 = Γ2 = Γ∗ ✁ Γ in (6.3a) and (6.3b) yields ζ(cid:1) ∈ Dom Γ. Note that Γ, Γ∗, Γ∗ ✁ Γ, Γ ✁ Γ∗ and Γ∗ ✁ Γ ✁ Γ∗ ✁ Γ all Γ∗✁Γξ(cid:13)(cid:13)2 (cid:0)Γ∗ ✁ Γ ✁ Γ∗ ✁ Γ(cid:1)[ξ] −(cid:0)Γ∗ ✁ Γ(cid:1)[ξ] =(cid:0)Γ∗ ✁ Γ(cid:1)[ξ] −(cid:13)(cid:13)F ∆ (cid:0)Γ∗ ✁ Γ ✁ Γ∗ ✁ Γ(cid:1)[ξ] −(cid:0)Γ∗ ✁ Γ(cid:1)[ξ] =(cid:0)Γ ✁ Γ∗(cid:1)(cid:2)(I + F ∆ Γ )ξ(cid:3). On the other hand, setting Γ1 = Γ∗, Γ2 = Γ ✁ Γ∗ and Γ3 = Γ in Lemma 6.4 yields Thus (6.5) holds. . Now suppose that Γ ✁ Γ∗ > V. Then, since V(cid:2)(I + F ∆ Γ )ξ(cid:3) = ν[u] = V[ξ], (6.5) implies that (Γ∗ ✁ Γ)[ξ] > V[ξ]. Thus Γ∗ ✁ Γ > V. The converse implication follows by exchanging Γ and Γ∗. (cid:3) Now we return to the Hilbert spaces h and k. Let Xhol(h) denote the class of quadratic forms γ on h which are closed, densely defined and satisfy the accretive and semisectorial conditions Re γ + β > 0 and (cid:12)(cid:12)Im γ[u](cid:12)(cid:12) 6 α(cid:0) Re γ[u] + kuk2(cid:1) for some β ∈ R and α ∈ R+, and let Shol(h) denote the class of holomorphic semigroups they generate (see e.g. [Ouh]). (u ∈ Dom γ) In [LS2], a quasicontractive QS cocycle V on h is called holomorphic if its expectation semigroup belongs to Shol(h). Denoting this class of QS cocycle by QSChol(h, k), it is shown there that the correspondence Xhol(h) → Shol(h) extends to a bijection X4 hol(h, k) → QSChol(h, k), Γ 7→ X Γ, 22 MARTIN LINDSAY in which X4 hol(h, k) denotes the subclass of Q(bk ⊗ h) = Q(h ⊕ (k ⊗ h)) consisting of forms Γ ∼ (γ, L,eL, C) such that γ ∈ Xhol(h) and Γ∗✁Γ+2β∆⊥ > 0, for some β ∈ R. We speak of the stochastic form generator of the holomorphic cocycle. If Γ ∼ (γ, L,eL, C) ∈ X4 then it follows from Proposition 6.6 that Γ∗ ∈ X4 C ∈ B(k ⊗ h) is a contraction, and X Γ∗ hol(h, k) hol(h, k) and in [LS2] it is also shown that = (X Γ)♯, the dual QS cocycle defined in (2.7). The bijection extends the above form-semigroup correspondence as follows: if Γ ∼ hol(h, k) and X Γ = (Pt ⊗ IF )t>0 where P is the (γ, 0, 0, I) where γ ∈ Xhol(h) then Γ ∈ X4 holomorphic semigroup with form generator γ. It also extends that of Theorem 2.1, in the sense that if F ∈ qc(bk ⊗ h) with block matrix form(cid:2) K M hol(h, k) given by Γ ∼ (q−K , L, M ∗, C). form in X4 L C−I(cid:3) then X F = X Γ for the If Γ1, Γ2 ∈ X4 hol(h, k) then Γ1 ✁ Γ2 ∈ X4 hol(h, k) provided only that Dom Γ1 ∩ Dom Γ2 is dense. This neatly extends the fact that if γ1, γ2 ∈ Xhol(h) then γ1 +γ2 ∈ Xhol(h) provided only that Dom γ1 ∩ Dom γ2 is dense in h. Conjecture 6.7. Let V 1 and V 2 be quasicontractive holomorphic QS cocycles on h with noise dimension space k and respective stochastic form generators Γ1 and Γ2, and suppose that Dom Γ1 ∩ Dom Γ2 is dense inbk ⊗ h. Then (in the notation of Theorem 3.5), for all T ∈ R+, ϕ ∈ B(F)∗ and u ∈ h, sup 06r6t6T(cid:13)(cid:13)(cid:0) idB(h) ⊗ ϕ(cid:1)(cid:0)V P,1✁2 r,t − X Γ1✁Γ2 r,t (cid:1) u(cid:13)(cid:13) → 0 as (cid:12)(cid:12)P ∩ [0, T ](cid:12)(cid:12) → 0. Moreover, if the QS cocycle X Γ1✁Γ2 is isometric and V 1 and V 2 are contractive then, for all T ∈ R+ and ξ ∈ h ⊗ F, sup 06r6t6T(cid:13)(cid:13)(cid:0)V P,1✁2 r,t − X Γ1✁Γ2 r,t (cid:1) ξ(cid:13)(cid:13) → 0 as (cid:12)(cid:12)P ∩ [0, T ](cid:12)(cid:12) → 0. Remarks. The conjecture has three special cases where it is proven. Theorem 3.5 covers the case where Γ1 and Γ2 are bounded. In the semigroup case, where Γi ∼ (γi, 0, 0, I) for γi ∈ Xhol(h) (i = 1, 2), it reduces to a version of a celebrated result of Kato – as extended by Simon ([Kat]). For the case of independent driving noises a version of the holomorphic counterpart to Corollary 3.6, which includes the Kato–Simon theorem, is proved in [LS3]. Acknowledgements. I am grateful to Mateusz Jurczy´nski and Micha l Gnacik for useful comments on an earlier draft of the paper. Support from the UK-India Education and Re- search Initiative (UKIERI), under the research collaboration grant Quantum Probability, Noncommutative Geometry & Quantum Information, is also gratefully acknowledged. References [Bel] V.P. Belavkin, A new form and a *-algebraic structure of quantum stochastic integrals in Fock space, Rend. Sem. Mat. Fis. Milano 58 (1988), 177–193. [BGL] A.C.R. Belton, M. Gnacik and J.M. Lindsay, Strong convergence of quantum random walks via semigroup decomposition, arXiv :1712.02848 [math.PR]. [BLS] A.C.R. Belton, J.M. Lindsay and A.G. Skalski, Quantum Feynman-Kac perturbations, J. London Math. Soc. (2 ) 89 (2014) no. 1, 275–300. [Bha] B.V.R. Bhat, Cocycles of CCR flows, Mem. Amer. Math. Soc. 149 (2001), no. 709. [BrR] O. Bratteli and D.W. Robinson, "Operator Algebras and Quantum Statistical Mechanics II: Equi- librium states. Models in quantum statistical mechanics," 2nd Edition, Springer-Verlag, Berlin, 1997. [Che] P. R. Chernoff, Product formulas, non-linear semigroups and addition of unbounded operators, Mem. Amer. Math. Soc. 140, 1974. [DGS] B. Das, D. Goswami and K.B. Sinha, A homomorphism theorem and a Trotter product formula for quantum stochastic flows with unbounded coefficients, Comm. Math. Phys. 330 (2014) no. 2, 435–467. QS LIE–TROTTER PRODUCT FORMULA 23 [DLT] B.K. Das, J.M. Lindsay and O. Tripak, Sesquilinear quantum stochastic analysis in Banach space, J. Math. Anal. Applic. 409 (2014) no. 2, 1032–1051. [Dav] E.B. Davies, "One-Parameter Semigroups," Academic Press, London, 1980. [EvH] M.P. Evans and R.L. Hudson, Perturbations of quantum diffusions, J. London Math. Soc. (2 ) (1990), no. 2, 373–384. [Fag] F. Fagnola, Characterization of isometric and unitary weakly differentiable cocycles in Fock space, in, "Quantum Probability & Related Topics," QP-PQ VIII, (Ed. L. Accardi), World Scientific, Singapore, 1993, pp. 143–164. [GL+] D. Goswami, J.M. Lindsay, K.B. Sinha and S.J. Wills, Dilation of Markovian cocycles on a von Neumann algebra, Pacific J. Math. 211 (2003) no. 2, 221–247. [GoJ] J. Gough and M. James, The series product and its application to quantum feedforward and feedback networks, IEEE Trans. Automat. Control 54 (2009) no. 11, 2530–2544. [HiP] E. Hille and R.S. Philips, "Functional Analysis and Semigroups," Coll. Publ. 31, American Math- ematical Society, Providence, R.I., 1957. [Hol] A.S. Holevo, Stochastic representation of quantum dynamical semigroups (Russian), [Translated in Proc. Steklov Math. Inst. (1992) no. 2, 145–154.] Trudy Mat. Inst. Steklov 191 (1989), 130–139. [HuP] R.L. Hudson and K.R. Parthasarathy, Quantum Ito's formula and stochastic evolution, Comm. Math. Phys. 93 (1984) no. 3, 301–323. [Jou] J.-L. Journ´e, Structure des cocycles markoviens sur l'espace de Fock, Probab. Theory Related Fields 75 (1987) no. 2, 291–316. [L] [Kat] T. Kato, Trotter's product formula for an arbitrary pair of self-adjoint contraction semigroups, in, "Topics in Functional Analysis (essays dedicated to M.G. Krein on the occasion of his 70th birthday)," Adv. in Math. Suppl. Stud. 3, Academic Press, London 1978, pp. 185–195. J.M. Lindsay, Quantum stochastic analysis - an introduction, in, "Quantum Independent Incre- ment Process, I: From Classical Probability to Quantum Stochastic Calculus" (Eds. U. Franz & M. Schurmann), Lecture Notes in Mathematics 1865, Springer-Verlag, Heidelberg 2005, pp. 181–271. [LS1] J.M. Lindsay and K. B. Sinha, A quantum stochastic Lie–Trotter product formula, Indian J. Pure Appl. Math. 41 (2010) no. 1, 313–325. [LS2] - - , Holomorphic quantum stochastic contraction cocycles, Preprint. [LS3] - - , Trotter–Kato product formulae for quantum stochastic cocycles, in preparation. [LW1] J.M. Lindsay and S.J. Wills, Existence, positivity, and contractivity for quantum stochastic flows with infinite dimensional noise, Probab. Theory Rel. Fields 116 (2000) no. 4, 505–543. [LW2] - - , Markovian cocycles on operator algebras, adapted to a Fock filtration, J. Funct. Anal. 178 (2000), no. 2, 269–305. [LW3] - -, Quantum stochastic cocycles and completely bounded semigroups on operators spaces, Int. Math. Res. Notices (2014) no. 11, 3096–3139. [Mey] P.-A. Meyer, "Quantum Probability for Probabilists" (2nd Edn.), Lecture Notes in Math. 1538, Springer-Verlag, Berlin 1995. [Ouh] E.M. Ouhabaz, "Analysis of Heat Equations on Domains," London Mathematical Society Mono- graphs 31, Princeton University Press, Princeton, 2005. [Par] K.R. Parthasarathy, "An Introduction to Quantum Stochastic Calculus," Monographs in Mathem- atics 85, Birkhauser, Basel 1992. [PaS] K.R. Parthasarathy and K.B. Sinha, A random Trotter–Kato product formula, "Statistics and Prob- ability: Essays in Honor of C.R. Rao," eds. G. Kallianpur, Paruchuri R. Krishnaiah & J.K. Ghosh, North-Holland, Amsterdam, 1982, pp. 553-566. [ReS] M. Reed and B. Simon, "Methods of Modern Mathematical Physics, I: Functional Analysis (2nd Edn.), II: Fourier Analysis, Self-Adjointness," Academic Press, New York, 1980, 1975. [SSS] L. Sahu, M. Schurmann and K.B. Sinha, Unitary processes with independent increments and rep- resentations of Hilbert tensor algebras, Publ. Res. Inst. Math. Sci. 45 (2009) no. 3, 745–785. [Sch] M. Schurmann, "White noise on bialgebras," Lecture Notes in Math. 1544, Springer-Verlag, Berlin 1993. [Sko] A.V. Skorohod, "Asymptotic Methods in the Theory of Stochastic Differential Equations," Trans- lations of Mathematical Monographs 78, American Mathematical Society, Providence, 1989. [Tro] H.F. Trotter, On the product of semi-groups of operators, Proc. Amer. Math. Soc. 10 (1959), 545–551. [Wil] S.J. Wills, On the generators of quantum stochastic operator cocycles, Markov Process. Related Fields 13 (2007) no. 1, 191–211. Department of Mathematics & Statistics, Lancaster University, Lancaster LA1 4YF, UK E-mail address: [email protected]
1210.4715
1
1210
2012-10-17T12:32:09
Hadamard differentiability via G\^ ateaux differentiability
[ "math.FA" ]
Let $X$ be a separable Banach space, $Y$ a Banach space and $f: X \to Y$ a mapping. We prove that there exists a $\sigma$-directionally porous set $A\subset X$ such that if $x\in X \setminus A$, $f$ is Lipschitz at $x$, and $f$ is G\^ateaux differentiable at $x$, then $f$ is Hadamard differentiable at $x$. If $f$ is Borel measurable (or has the Baire property) and is G\^ ateaux differentiable at all points, then $f$ is Hadamard differentiable at all points except a set which is $\sigma$-directionally porous set (and so is Aronszajn null, Haar null and $\Gamma$-null). Consequently, an everywhere G\^ ateaux differentiable $f: \R^n \to Y$ is Fr\' echet differentiable except a nowhere dense $\sigma$-porous set.
math.FA
math
HADAMARD DIFFERENTIABILITY VIA G ATEAUX DIFFERENTIABILITY LUDEK ZAJ´I CEK Abstract. Let X be a separable Banach space, Y a Banach space and f : X → Y a mapping. We prove that there exists a σ-directionally porous set A ⊂ X such that if x ∈ X \ A, f is Lipschitz at x, and f is Gateaux differentiable at x, then f is Hadamard differentiable at x. If f is Borel measurable (or has the Baire property) and is Gateaux differentiable at all points, then f is Hadamard differentiable at all points except a set which is σ-directionally porous set (and so is Aronszajn null, Haar null and Γ-null). Consequently, an everywhere Gateaux differentiable f : Rn → Y is Fr´echet differentiable except a nowhere dense σ-porous set. 2 1 0 2 t c O 7 1 ] . A F h t a m [ 1 v 5 1 7 4 . 0 1 2 1 : v i X r a 1. Introduction Hadamard derivative of a mapping f : X → Y between Banach spaces, which is stronger than Gateaux derivative but weaker that Fr´echet derivative, was applied many times in the literature. For three formally different but equivalent definitions of Hadamard derivative see Lemma 2.1 below. If X is finite-dimensional, then Hadamard derivative coincides with Fr´echet derivative, but for infinite-dimensional spaces Fr´echet derivative is “much stronger”, even for Lipschitz f . On the other hand, if f is Lipschitz, then Hadamard derivative coincides with Gateaux derivative; in this sense these two types of derivatives are rather close. Using this fact, we obtain the well-known result that if f is everywhere Gateaux differentiable and has the Baire property, than f is Hadamard differentiable at all points except a nowhere dense set. Indeed, [12, Corollary 3.9] easily implies that, under above assumptions, f is locally Lipschitz on a dense open set U. However, an easy example (see Remark 3.10) shows that in general (even if X = Y = R) we cannot find U with “measure null” complement. We prove (Theorem 3.9) that if X is separable and f has the Baire property and is everywhere Gateaux differentiable, then f is Hadamard differentiable at all points except a σ-directionally porous set. This is an interesting additional information, since each σ-directionally porous subset of a separable Banach spaces X is “measure null”: it is Aronszajn (=Gauss) null, and so also Haar null, (see [2, p. 164 and Chap. 6]) and also Γ-null (see [8, Remark 5.2.4]). As an easy consequence of Theorem 3.9, we obtain (see Theorem 3.12) that an ev- erywhere Gateaux differentiable f : Rn → Y is Fr´echet differentiable except a σ-porous set. The main ingredient of the proof of Theorem 3.9 is Theorem 3.4. It asserts that if X is separable and f : X → Y , then there exists a σ-directionally porous set A ⊂ X such that if x ∈ X \ A, f is Lipschitz at x, and f is Gateaux differentiable at x, then f is Hadamard differentiable at x. 2010 Mathematics Subject Classification. Primary: 46G05; Secondary: 26B05, 49J50. Key words and phrases. Hadamard differentiability, Gateaux differentiability, Fr´echet differentiability, σ-porous set, σ-directionally porous set, Stepanoff theorem, Aronszajn null set. The research was supported by the grant GA CR P201/12/0436. 1 2 LUD EK ZAJ´I CEK Lemma 3.6 implies that if X is separable and f has the Baire property and is everywhere Gateaux differentiable, then f is Lipschitz at all points except a σ-directionally porous set. So, Theorem 3.4 together with Lemma 3.6 imply Theorem 3.9. Further, Theorem 3.4 shows that the infinite-dimensional version of Stepanoff theo- rem on Gateaux differentiability from [4] holds also for Hadamard differentiability (see Corollary 3.14). 2. Preliminaries In the following, by a Banach space we mean a real Banach space. If X is a Banach space, we set SX := {x ∈ X : kxk = 1}. The symbol B(x, r) will denote the open ball with center x and radius r. In a metric space (X, ρ), the system of all sets with the Baire property is the smallest σ-algebra containing all open sets and all first category sets. We will say that a mapping f : (X, ρ1) → (Y, ρ2) has the Baire property if f is measurable with respect to the σ- algebra of all sets with the Baire property. In other words, f has the Baire property, if and only if f −1(B) has the Baire property for all Borel sets B ⊂ Y (see [6, Section 32]). Let X, Y be Banach spaces, G ⊂ X an open set, and f : G → Y a mapping. We say that f is Lipschitz at x ∈ G if lim supy→x pointwise Lipschitz if f is Lipschitz at all points of G. ky−xk < ∞. We say that f is kf (y)−f (x)k The directional and one-sided directional derivatives of f at x ∈ G in the direction v ∈ X are defined respectively by f ′(x, v) := lim t→0 f (x + tv) − f (x) t and f ′ +(x, v) := lim t→0+ f (x + tv) − f (x) t . The Hadamard directional and one-sided directional derivatives of f at x ∈ G in the direction v ∈ X are defined respectively by f ′ H(x, v) := lim z→v,t→0 t f (x + tv) − f (x) It is easy to see that f ′(x, v) (resp. f ′ (resp. f ′ H+(x, −v) = −f ′ H+(x, v)). and f ′ H+(x, v) := lim z→v,t→0+ H (x, v)) exists if and only if f ′ f (x + tv) − f (x) . t +(x, −v) = −f ′ +(x, v) It is well-known and easy to prove that, if f is locally Lipschitz on G, then f ′(x, v) = f ′ H(x, v) (resp. f ′ The usual modern definition of the Hadamard derivative is the following: A continuous linear operator L : X → Y is said to be a Hadamard derivative of f at a H+(x, v)) whenever one of these two derivatives exists. +(x, v) = f ′ point x ∈ X if lim t→0 f (x + tv) − f (x) t = L(v) for each v ∈ X and the limit is uniform with respect to v ∈ C, whenever C ⊂ X is a compact set. In this case we set f ′ H (x) := L. The following fact is well-known (see [11]): Lemma 2.1. Let X, Y be Banach spaces, ∅ 6= G ⊂ X an open set, x ∈ G, f : G → Y a mapping and L : X → Y a continuous linear operator. Then the following conditions are equivalent: (i) f ′ (ii) f ′ (iii) if ϕ : [0, 1] → X is such that ϕ(0) = x and ϕ′ H(x) := L, H(x, v) = L(v) for each v ∈ X, +(0) exists, then (f ◦ ϕ)′ +(0) = L(ϕ′ +(0)). HADAMARD DIFFERENTIABILITY VIA G ATEAUX DIFFERENTIABILITY 3 Recall (see, e.g. [11]), that if f is Hadamard differentiable at a ∈ G, then f is Gateaux differentiable at a, and (2.1) if f is locally Lipschitz on G, then also the opposite implication holds. Further, (2.2) if X = Rn, then Hadamard differentiability is equivalent to Fr´echet differentiability. Definition 2.2. Let X be a Banach space. We say that A ⊂ X is directionally porous at a point x ∈ X, if there exist 0 6= v ∈ X, p > 0 and a sequence tn → 0 of positive real numbers such that B(x + tnv, ptn) ∩ A = ∅. (In this case we say that A is porous at x in direction v.) We say that A ⊂ X is directionally porous if A is directionally porous at each point x ∈ A. We say that A ⊂ X is σ-directionally porous if it is a countable union of directionally porous sets. Recall that directional porosity is stronger than (upper) porosity, but (2.3) if X is finite-dimensional, then these two notions are equivalent. Let X be a Banach space, x ∈ X, v ∈ SX and δ > 0. Then we define the open cone C(x, v, δ) as the set of all y 6= x for which kv − y−x ky−xk k < δ. The following easy inequality is well known (see e.g. [9, Lemma 5.1]): (2.4) if u, v ∈ X \ {0}, then (cid:13)(cid:13)(cid:13)(cid:13) u kuk − v kvk 2 kuk ku − vk. (cid:13)(cid:13)(cid:13)(cid:13) ≤ Because of the lack of a reference we supply the proof of the following easy fact. Lemma 2.3. Let X be a Banach space, Y a Banach space, G ⊂ X an open set, a ∈ G, and f : G → Y a mapping. Then the following are equivalent. H+(a, 0) exists, H(a, 0) exists, H(a, 0) = 0, (i) f ′ (ii) f ′ (iii) f ′ (iv) f is Lipschitz at a. Proof. The implications (ii) ⇒ (i) and (iii) ⇒ (ii) are trivial. To prove (iv) ⇒ (iii), suppose that K > 0, δ > 0 and kf (x) − f (a)k ≤ Kkx − ak for each x ∈ B(a, δ). To prove (iii), let ε ∈ (0, K) be given. Then, for z ∈ X with kzk < ε/K and 0 < t < δ, we obtain kf (a + tz) − f (a)k ≤ Kt ε K , and (iii) follows. (cid:13)(cid:13)(cid:13)(cid:13) f (a + tz) − f (a) t (cid:13)(cid:13)(cid:13)(cid:13) ≤ ε, To prove (i) ⇒ (iv), suppose that (iv) does not hold. Then, for each n ∈ N, choose hn ∈ X such that 0 < khnk < n−2 and kf (a + hn) − f (a)k ≥ n2khnk. Set tn := nkhnk and zn := n−1khnk−1hn. Then zn → 0, tn → 0, tn > 0, and (cid:13)(cid:13)(cid:13)(cid:13) f (a + tnzn) − f (a) tn kf (a + hn) − f (a)k nkhnk ≥ n, (cid:13)(cid:13)(cid:13)(cid:13) = which clearly implies that f ′ H+(a, 0) does not exist. (cid:3) 4 LUD EK ZAJ´I CEK We will need also the following special case of [12, Lemma 2.4]. It can be proved by the Kuratowski-Ulam theorem (as is noted in [12]), but the proof given in [12] is more direct. Lemma 2.4. Let U be an open subset of a Banach space X. Let M ⊂ U be a set residual in U and z ∈ U . Then there exists a line L ⊂ X such that z is a point of accumulation of M ∩ L. 3. Results Proposition 3.1. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set, and f : G → Y a mapping. Let A be the set of all points x ∈ G for which there exists ky−xk < ∞ and f is not Lipschitz at a cone C = C(x, v, δ) such that lim supy→x,y∈C x. Then A is a σ-directionally porous set. kf (y)−f (x)k Proof. Let {vn : n ∈ N} be a dense subset of SX. For natural numbers k, p, n, m denote by Ak,p,n,m the set of all points x ∈ A such that (3.1) kf (y) − f (x)k ky − xk < k whenever 0 < ky − xk < 1 p and y − x ky − xk − vn (cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13) < 1 m . Since clearly A is the union of all sets Ak,p,n,m, it is sufficient to prove that each set Ak,p,n,m is directionally porous. So, suppose that natural numbers k, p, n, m and x ∈ Ak,p,n,m are given. It is sufficient to prove that Ak,p,n,m is directionally porous at x. To this end, find a sequence yi → x such that (3.2) kf (yi) − f (x)k kyi − xk > k(12m + 4) for each i ∈ N. Set ri := kyi − xk and xi := x − 6mrivn. It is sufficient to prove that there exists i0 ∈ N such that (3.3) B(xi, ri) ∩ Ak,p,n,m = ∅ for each i ≥ i0. To this end, consider i ∈ N and zi ∈ B(xi, ri) ∩ Ak,p,n,m. Observe that 6mri − ri ≤ kx − xik − kxi − zik ≤ kx − zik ≤ kx − xik + kxi − zik ≤ 6mri + ri, 6mri − 2ri ≤ kx − zik − ri ≤ kyi − zik ≤ kx − zik + ri ≤ 6mri + 2ri. These inequalities imply that there exists i0 ∈ N (independent on i) such that (3.4) 0 < kx − zik < 1 p and 0 < kyi − zik < 1 p , if i ≥ i0. Applying (2.4), we obtain (3.5) x − zi kx − zik − vn (cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) x − zi kx − zik − x − xi kx − xik (cid:13)(cid:13)(cid:13)(cid:13) ≤ 2 kxi − zik 6mri − ri < 2ri 6mri − ri ≤ 1 m and (3.6)(cid:13)(cid:13)(cid:13)(cid:13) yi − zi kyi − zik − vn (cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13) yi − zi kyi − zik − x − xi kx − xik (cid:13)(cid:13)(cid:13)(cid:13) ≤ 2 kyi − xk + kxi − zik kyi − zik < 2 2ri 6mri − 2ri ≤ 1 m . Since zi ∈ Ak,p,n,m, conditions (3.1), (3.4), (3.5) and (3.6) imply that, if i ≥ i0, then kf (x) − f (zi)k kx − zik < k and kf (yi) − f (zi)k kyi − zik < k. HADAMARD DIFFERENTIABILITY VIA G ATEAUX DIFFERENTIABILITY 5 So, using also (3.2), we obtain that, if i ≥ i0, then rik(12m + 4) ≤ kf (x) − f (yi)k ≤ kf (x) − f (zi)k + kf (yi) − f (zi)k < kkx − zik + kkyi − zik ≤ k(6mri + ri + 6mri + 2ri), which is impossible. So we have proved (3.3). (cid:3) Proposition 3.2. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set, and f : G → Y a mapping. Let M be the set of all x ∈ G at which f is Lipschitz and there exists v ∈ X such that f ′ H+(x, v) does not exist. Then M is σ-directionally porous. +(x, v) exists but f ′ Proof. For each k ∈ N, set Mk := {x ∈ M : kf (y) − f (x)k ≤ kky − xk whenever ky − xk < 1/k}. It is clearly sufficient to prove that each Mk is directionally porous. So suppose that k ∈ N and x ∈ Mk. Choose v ∈ X such that f ′ H+(x, v) does not exist. Lemma 2.3 gives that v 6= 0. We will prove that Mk is porous at x in direction v. Since f ′ H+(x, v) does not exist, we can choose ε > 0 such that for each δ > 0 there exists w ∈ B(v, δ) and t ∈ (0, δ) with kt−1(f (x + tw) − f (x)) − yk > ε. +(x, v) =: y exists but f ′ For each n ∈ N, we choose wn ∈ B(v, 1/n) and tn ∈ (0, 1/n) such that (3.7) Set xn := x + tnv and p := ε(3k)−1. It is sufficient to prove that there exists n0 ∈ N such that k(tn)−1(f (x + tnwn) − f (x)) − yk > ε. (3.8) B(xn, ptn) ∩ Mk = ∅ for each n ≥ n0. To this end consider n ∈ N and zn ∈ B(xn, ptn) ∩ Mk. Denote x∗ n := x + tnwn. We have kzn − xnk < ptn, (3.9) Thus there exists en0 ∈ N (independent on n) such that the inequality n ≥ en0 implies kzn − xnk < 1/k, kzn − x∗ nk < 1/k, and thus, since zn ∈ Mk, also nk + kzn − xnk ≤ tn/n + ptn. nk ≤ kxn − x∗ kzn − x∗ kf (xn) − f (zn)k ≤ kkxn − znk and kf (x∗ (3.10) If n ≥ en0, then (3.7), (3.9) and (3.10) imply n) − f (zn)k ≤ kkx∗ n − znk. ε < ≤ ≤ = − y − y f (x∗ kf (x∗ n) − f (x) f (xn) − f (x) f (xn) − f (x) kf (xn) − f (zn)k (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:13)(cid:13)(cid:13)(cid:13) − y(cid:13)(cid:13)(cid:13) → 0, we obtain a contradiction, if n > n0, where n0 ≥ en0 is (cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13) + (cid:13)(cid:13)(cid:13)(cid:13) + kp + k(1/n + p) (cid:13)(cid:13)(cid:13)(cid:13) + (2/3)ε + k/n. f (xn) − f (x) n) − f (zn)k − y tn tn tn tn − y tn + tn Since (cid:13)(cid:13)(cid:13) f (xn)−f (x) tn sufficiently large (independent on n). So (3.8) is proved. (cid:3) Remark 3.3. The corresponding “two-sided” result which works with f ′(x, v) and f ′ clearly follows from Proposition 3.2. H (x, v) 6 LUD EK ZAJ´I CEK Using Proposition 3.2 (together with Remark 3.3) and Lemma 2.1, we immediately obtain: Theorem 3.4. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set and f : G → Y a mapping. Then the set of all points at which f is Lipschitz and Gateaux differentiable but is not Hadamard differentiable is σ-directionally porous. Theorem 3.4 together with [12, Corollary 3.9] have the following consequence. Corollary 3.5. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set, and f : G → Y a pointwise Lipschitz mapping. Then the set of all points from G at which f is Gateaux differentiable but is not Hadamard differentiable is nowhere dense and σ-directionally porous. Lemma 3.6. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set, and f : G → Y a mapping. Suppose that f has the Baire property and, for each line L ⊂ X with G ∩ L 6= ∅, the restriction of f to G ∩ L is continuous. Denote by A the set of all points x ∈ G for which there exists a set Bx ⊂ SX of the second category in SX such that (3.11) lim sup t→0+ kf (x + tv) − f (x)k t < ∞ for each v ∈ Bx and f is not Lipschitz at x. Then A is a σ-directionally porous set. Proof. By Proposition 3.1 it is sufficient to prove that for each x ∈ A there exists v ∈ SX and δ > 0 such that (3.12) lim sup y→x,y∈C(x,v,δ) kf (y) − f (x)k ky − xk < ∞. So fix a point x ∈ A. Find p0 ∈ N with B(x, 1/p0) ⊂ G and, for natural numbers p ≥ p0, k, denote Sp,k := {v ∈ SX : kf (x + tv) − f (x)k t ≤ k for each 0 < t < 1/p.} Since Bx is clearly covered by all sets Sp,k, we can find k and p ≥ p0 such that Sp,k is a second category set (in SX). Without any loss of generality, we can clearly suppose that x = 0 and f (x) = f (0) = 0. We will show that S := Sp,k has the Baire property (in SX). Set E := B(0, 1/p) \ {0} and E∗ := SX × (0, 1/p) (equipped with the product topology). For each (v, t) ∈ E∗ set ϕ((v, t)) := tv. Then ϕ : E∗ → E is clearly a homeomorphism (with ϕ−1(z) = (z/kzk, kzk) for z ∈ E). Set M := {z ∈ E : kf (z)k kzk ≤ k} = {z ∈ E : kf (z)k − kkzk ≤ 0}. Since f has the Baire property, using continuity of norms we easily obtain that the real function z 7→ kf (z)k−kkzk has the Baire property on E, and so M has the Baire property in E. Consequently M ∗ := ϕ−1(M) and C := E∗ \ M ∗ have the Baire property in E∗. For any set A ⊂ E∗ and v ∈ SX define the section Av := {t ∈ (0, 1/p) : (v, t) ∈ A} and the projection π(A) := {v ∈ SX : Av 6= ∅}. It is easy to see that (3.13) S = SX \ π(C). Using the continuity of f on the set {tv : t ∈ (0, 1/p)}, we obtain that (3.14) Cv is open for each v ∈ SX. HADAMARD DIFFERENTIABILITY VIA G ATEAUX DIFFERENTIABILITY 7 Since C has the Baire property, we can write C = H ∪ T , where H is a Gδ set (in E∗) and T is a first category set (in E∗). We have (3.15) π(C) = π(H) ∪ π(T ) = π(H) ∪ (π(T ) \ π(H)). The set π(H) is analytic (see, e.g [5, Exercise 14.3]), and so (see [5, Theorem 21.6]) it has the Baire property. By Kuratowski-Ulam theorem (see, e.g. [5, Theorem 8.41]) there exists a first category set Z ⊂ SX such that Tv is a first category subset of (0, 1/p) for each v ∈ SX \ Z. The inclusion π(T ) \ π(H) ⊂ Z holds. Indeed, suppose on the contrary that there exists v ∈ (π(T ) \ π(H)) \ Z. Then Tv is a nonempty first category set and Hv = ∅. Thus Cv = Hv ∪ Tv = Tv. Using (3.14), we obtain that Cv is a nonempty open first category set, which is a contradiction. Consequently π(T ) \ π(H) is a first category set and thus S has the Baire property by (3.13) and (3.15). Since S is also a second category set, we can choose v ∈ SX and δ > 0 such that S = Sp,k is residual in SX ∩ B(v, δ). Set U ∗ := (SX ∩ B(v, δ)) × (0, 1/p), ψ := ϕ ↾U ∗ and U := C(0, v, δ)∩B(0, 1/p). Then ψ : U ∗ → U is clearly a homeomorphism and ψ(S × (0, 1/p)) ⊂ M. Since S × (0, 1/p) is residual in U ∗ (see [5, Lemma 8.43]), we obtain that M is residual in U. Now consider an arbitrary z ∈ U. By Lemma 2.4 there exists a line L ⊂ X and points zn ∈ M ∩ L ∩ U with zn → z. Since the restriction of f to L ∩ U is continuous, we obtain kf (zn)k kzk , and consequently z ∈ M. So U ⊂ M, which implies (3.12). (cid:3) kznk → kf (z)k In fact, we will use only the following immediate consequence of Lemma 3.6. Lemma 3.7. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set, and f : G → Y a mapping. Suppose that f has the Baire property and, for each line L ⊂ X with G ∩ L 6= ∅, the restriction of f to G ∩ L is continuous. Denote by B the set of all points x ∈ G at which f is Gateaux differentiable and f is not Lipschitz at x. Then B is a σ-directionally porous set. As an immediate application of Lemma 3.7 and Theorem 3.4 we obtain Proposition 3.8. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set, and f : G → Y a mapping. Suppose that f has the Baire property and, for each line L ⊂ X with G ∩ L 6= ∅, the restriction of f to G ∩ L is continuous. Then the set of all points x ∈ G at which f is Gateaux differentiable but is not Hadamard differentiable is σ-directionally porous. As already mentioned in Introduction, [12, Corollary 3.9] easily implies that if f is everywhere Gateaux differentiable and has the Baire property, then f is locally Lipschitz on a dense open set U, and so (see (2.1)) f is Hadamard differentiable at all points except a nowhere dense set (even if X is nonseparable). This result together with Proposition 3.8 give the following theorem. Theorem 3.9. Let X be a separable Banach space, Y a Banach space, G ⊂ X an open set. Suppose that f : G → Y has the Baire property and is everywhere Gateaux differentiable. Then f is Hadamard differentiable at all points of G except a nowhere dense σ-directionally porous set. Remark 3.10. The assumptions of Theorem 3.9 do not imply that f is locally Lipschitz on the complement of a closed σ-directionally porous set even in the case X = Y = R. 8 LUD EK ZAJ´I CEK Indeed, let (an, bn), n ∈ N be the system of bounded intervals such that G := Sn∈N(an, bn) is dense in R and F := R \ G has positive Lebesgue measure. We can clearly define a function f on R such that (i) f (x) = 0 for x ∈ F , (ii) f has continuous derivative on each (an, bn), (iii) f (x) ≤ (x − an)2(x − bn)2 for x ∈ (an, bn), (iv) for each n ∈ N and δ > 0, the derivative f ′ is unbouded both on (an, an + δ) and on (bn − δ, bn). Then f is everywhere differentiable, G is the largest open set on which f is locally Lipschitz and F is not σ-directionally porous. So Theorem 3.9 cannot be proved simply using only the fact that Gateaux and Hadamard differentiability coincide for locally Lipschitz functions. Remark 3.11. In Theorem 3.9, we cannot omit the assumption that f has the Baire property. This follows from Lemma 2.3 and Shkarin’s [12] example which shows that on each separable infinite-dimensional Banach space there exists an everywhere Gateaux differentiable real function which is discontinuous at all points. However, if X is finite-dimensional, the assumption that f has the Baire property can be omitted. Indeed (as is also noted without a reference in [12, Remark 3.5]), if f is everywhere Gateaux differentiable on Rn, than f has the Baire property. This follows, e.g., from the old well-known fact (see [7]) that a partially continuous real function on Rn is in the (n − 1)st Baire class (whose proof works also for a Banach space valued function). So, using also (2.2) and (2.3), we obtain the following result. Theorem 3.12. Let G be an open subset of Rn, Y a Banach space, and f : G → Y an everywhere Gateaux differentiable mapping. Then f is Fr´echet differentiable at all points of G except a nowhere dense σ-porous set. Using Theorem 3.4 and the main result of [4] (which generalizes the result of [10] and improves the result of [3]), we easily obtain Theorem 3.13 below. In its formulation, the system A of subsets of a separable Banach space is used. This system which was defined in [10] is strictly smaller (see [10, Proposition 13]) than the well-known system of Aronszajn null sets (and also than the system of Γ-null sets, see [13]). We will not recall the (slightly technical) definition of the system A. Note only, that all members of A are Borel by definition and that it is easy to see that A is stable with respect to countable unions. Theorem 3.13. Let X be a separable Banach space, Y a Banach space with the Radon- Nikod´ym property, G ⊂ X an open set, and f : G → Y a mapping. Then there exists a set A ∈ A such that if x ∈ G \ A and f is Lipschitz at x, then f is Hadamard differentiable at x. Proof. The main result of [4] says that there exists a set A1 ∈ A such that if x ∈ G \ A1 and f is Lipschitz at x, then f is Gateaux differentiable at x. Let A∗ 2 be the set of all x ∈ G at which f is Lipschitz, Gateaux differentiable, but not Hadamard differentiable. Then A∗ 2 is σ-directionally porous by Theorem 3.4. Using [10, Proposition 14] and [10, Theorem 12] we easily obtain that there exists A2 ∈ A which contains A∗ 2. Now it is clearly sufficient to set A := A1 ∪ A2. (cid:3) As a corollary, we obtain an infinite dimensional analogue of Stepanoff theorem on Hadamard differentiability. HADAMARD DIFFERENTIABILITY VIA G ATEAUX DIFFERENTIABILITY 9 Corollary 3.14. Let X be a separable Banach space, Y a Banach space with the Radon- Nikod´ym property, G ⊂ X an open set, and f : G → Y a pointwise Lipschitz mapping. Then f is Hadamard differentiable at all points of G except a set belonging to A. Remark 3.15. Corollary 3.14 and (2.2) imply the following result. Let Y a Banach space with the Radon-Nikod´ym property, G ⊂ Rn an open set, and f : G → Y a pointwise Lipschitz mapping. Then f is Fr´echet differentiable at all points of G except a set belonging to A. However, it is not probably an improvement of the classical Stepanoff theorem, since from some unpublished results follows that each Borel Lebesgue null subset of Rn belongs to A. For n = 2 this follows from [1, Theorem 7.5] (via [10, Theorem 12]), which is stated without a proof, and for n ≥ 3 from the corresponding theorem recently announced by M. Csornyei and P. Jones. References [1] G. Alberti, M. Csornyei, D. Preiss, Structure of null sets in the plane and applications, European Congress of Mathematics, 3-22, Eur. Math. Soc., Zurich, 2005. [2] Y. Benyamini, J. Lindenstrauss, Geometric Nonlinear Functional Analysis, Vol. 1, Colloquium Publications 48, American Mathematical Society, Providence, 2000. [3] D. Bongiorno, Stepanoff ’s theorem in separable Banach spaces, Comment. Math. Univ. Carolin. 39 (1998), 323-335. [4] J. Duda, On Gateaux differentiability of pointwise Lipschitz mappings, Canad. Math. Bull. 51 (2008), 205-216. [5] A.S. Kechris, Classical Descriptive Set Theory, Springer, New York, 1994. [6] K. Kuratowski, Topology, vol. I, Academic Press, New York, 1966, [7] H. Lebesgue, Sur les fonctions repr´esentable analytiquement, J. Math. Pure Appl. (6), vol. 1 (1905), 139–212. [8] J. Lindenstrauss, D. Preiss, J. Tiser, Fr´echet differentiability of Lipschitz maps and porous sets in Banach spaces, Princeton University Press, Princeton 2012. [9] J. L. Massera, J. J. Schaffer, Linear differential equations and functional analysis, I, Ann. of Math. (2) 67 (1958), 517–573. [10] D. Preiss, L. Zaj´ıcek, Directional derivatives of Lipschitz functions, Israel J. Math. 125 (2001), 1-27. [11] A. Shapiro, On concepts of directional differentiability, J. Optim. Theory Appl. 66 (1990), 477- 487. [12] S.A. Shkarin, Points of discontinuity of Gateaux-differentiable mappings, (Russian), Sibirsk. Mat. Zh. 33 (1992), 176–185. [13] L. Zaj´ıcek, On sets of non-differentiability of Lipschitz and convex functions, Math. Bohem. 132 (2007), 75-85. E-mail address: [email protected] Charles University, Faculty of Mathematics and Physics, Sokolovsk´a 83, 186 75 Praha 8-Karl´ın, Czech Republic
1811.10987
1
1811
2018-11-27T14:00:13
Frechet algebras in abstract Harmonic analysis
[ "math.FA" ]
We provide a survey of the similarities and differences between Banach and Frechet algebras including some known results and examples. We also collect some important generalizations in abstract harmonic analysis; for example, the generalization of the concepts of vector-valued Lipschitz algebras, abstract Segal algebras, Arens regularity, amenability, weak amenability, ideal amenability, etc.
math.FA
math
FR´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS Z. ALIMOHAMMADI AND A. REJALI Abstract. We provide a survey of the similarities and differences between Banach and Fr´echet algebras including some known results and examples. We also collect some important generalizations in abstract harmonic anal- ysis; for example, the generalization of the concepts of vector-valued Lips- chitz algebras, abstract Segal algebras, Arens regularity, amenability, weak amenability, ideal amenability, etc. 0. Introduction The class of Fr´echet algebras which is an important class of locally convex algebras has been widely studied by many authors. For a full understanding of Fr´echet algebras, one may refer to [27, 36]. In the class of Banach algebras, there are many concepts which were generalized to the Fr´echet case. Examples of these concepts are as follows. Biprojective Banach algebras were studied by several authors, notably A. Ya. Helemskii [35] and Yu. V. Selivanov [62]. A number of papers appeared in the literature concerned with extending the notion of biprojectivity and its results to the Fr´echet algebras; see for example [48, 63]. Flat cyclic Fr´echet modules were investigated by A. Yu. Pirkovskii [49]. Let A be a Fr´echet algebra and I be a closed left ideal in A. Pirkovskii showed that there exists a necessary and sufficient condition for a cyclic Fr´echet A-module A♯/I to be strictly flat. He also generalized a number of characterizations of amenable Ba- nach algebras to the Fr´echet algebras (ibid.). In 2008, P. Lawson and C. J. Read [43] introduced and studied the notions of approximate amenability and approxi- mate contractibility of Fr´echet algebras. Recently, the concepts of ϕ-amenability and character amenability, ϕ-contractibility and character contractibility, weak amenability and Ideal amenability for Fr´echet algebras have investigated in sev- eral papers; see [7, 8, 11, 55]. T. G. Honary [37] gave an extension of Johnson's uniqueness of norm theorem. He showed that if A and B are Fr´echet algebras, B is semisimple and T : A → B is a surjective homomorphism with a certain condition, then T is continuous. In special case, if A is a Banach algebra, then every epimorphism T : A → B is automatically continuous. Hence, every semisimple Banach algebra has a unique topology as a Fr´echet algebra. Honary then presented a partial answer to the well-known Michael's problem posed in [46]. Let (X, d) be a metric space and E be a Banach algebra over C. Take α ∈ R with α > 0. E. Biyabani and A. Rejali [18] studied the structure and proper- ties of vector-valued Lipschitz algebras Lipα(X, E) and lipα(X, E) of order α. Later, these concepts were generalized to the class of Fr´echet algebras by Re- jali joint with Ranjbari [54]. They also studied the concept of ideal amenability for vector-valued Fr´echet Lipschitz algebras (ibid.). Furthermore, Rejali et al. 2010 Mathematics Subject Classification. 46H05, 46H25, 46J05,43A20. Key words and phrases. locally convex space, Fr´echet algebra, Banach algebra. 1 2 Z. ALIMOHAMMADI AND A. REJALI [9, 10] introduced and studied the notions of Segal and semisimple Segal Fr´echet algebras. They showed that every continuous linear left multiplier of a Fr´echet algebra (B, qn)n∈N is also a continuous linear left multiplier of any Segal Fr´echet algebra (A, pℓ)ℓ∈N in (B, qn)n∈N. Moreover, they proved that if B is a commutative Fr´echet Q-algebra with an approximate identity, then the spaces of all modular maximal closed ideals of B and any Segal Fr´echet algebra A in B are homeomor- phic. Particularly, B is semisimple if and only if A is semisimple [10]. In [2], we joint with Abtahi introduced and studied the strict, uniform, and compact-open locally convex topologies on an arbitrary algebra B, by the fundamental system of seminorms of a locally convex subalgebra (A, pα). In particular, we studied the strict topology on the multiplier algebra M (A) of a Fr´echet algebra A, and we showed that A is a M (A)-Segal algebra. In the present paper we give a list of the significant results which have been obtained by various authors. Some of the results in this paper were included in the first author's Ph.D. thesis, which was supported by the center of excellence for mathematics at the University of Isfahan. 1. Preliminaries In this section we give some basic definitions and frameworks related to locally convex spaces which will be required throughout the paper; see [27, 36, 45] for more information. In this paper, all vector spaces and algebras are assumed to be over the field C of complex numbers. A locally convex space E is a topological vector space in which the origin has a local base of absolutely convex absorbent sets. Throughout the paper, all locally convex spaces are assumed to be Hausdorff. Every locally convex Hausdorff space E has a fundamental system of seminorms (pα)α∈Λ, or equivalently a family of continuous seminorms satisfying the following properties: (i) For every x ∈ E with x 6= 0, there exists an α ∈ Λ with pα(x) > 0; (ii) For all α, β ∈ Λ, there exist γ ∈ Λ and C > 0 such that max(pα(x), pβ (x)) ≤ Cpγ(x) (x ∈ E). We denote by (E, pα)α∈Λ, the locally convex space with the fundamental system of seminorms (pα)α∈Λ. Suppose that (E, pα)α∈Λ and (F, qβ )β∈B are locally convex spaces. For every linear mapping T : E → F , the following assertions are equivalent: (i) T ∈ L(E, F ), i.e. T is continuous; (ii) T is continuous at 0; (iii) For each β ∈ B there exist an α ∈ A and C > 0, such that qβ(T (x)) ≤ Cpα(x) (x ∈ E). Following [45], a complete metrizable locally convex space is called a Fr´echet space. Indeed, a Fr´echet space is a locally convex space which has a countable fundamental system of seminorms (pℓ)ℓ∈N. It is worth noting to remark that due to Conway [23], Fr´echet spaces are not assumed to be locally convex. These two different definitions make some basic differences. Consider Lp(0, 1) (0 < p < 1), the space of all equivalence classes of measurable functions f : (0, 1) → R such 0 f (x)pdx < ∞. Under the translation metric thatR 1 d(f, g) =Z 1 0 (f − g)xpdx < ∞ (f, g ∈ Lp(0, 1)), Lp(0, 1) is a complete metric space. Therefore, according to the definition pre- sented in [23], Lp(0, 1) is a Fr´echet space. However, Lp(0, 1) has only one nonempty the seriesP∞ k=1 λkxk converges in A. FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 3 open convex set, namely itself. It follows that due to the definition given in [45], Lp(0, 1) is not a Fr´echet space. In this paper, our definition of a Fr´echet space is based on that given in [45]. A Fr´echet space A is called quasinormable if for each zero neighborhood U there is a zero neighborhood V so that for every ε > 0 there exists a bounded set B in A such that V ⊆ B + εU . It should be noted that, by [45, Proposition 25.7], the strong dual space of every Fr´echet space A is a (DF)-space which is a type of locally convex topological vector spaces; see [45, page 297] for details. We denote the strong dual of A by A∗. Following [45, Corollary 24.29], every Fr´echet space A has a web {Cn1,··· ,nk }, that is a family Cn1,··· ,nk , n1, · · · , nk ∈ N, k ∈ N, of absolutely convex subsets of A with the following properties: n=1Cn = A; n=1Cn1,··· ,nk,n = Cn1,··· ,nk for all n1, · · · , nk ∈ N and all k ∈ N; (i) ∪∞ (ii) ∪∞ (iii) for each sequence (nk)k∈N in N there exists a sequence (λk)k∈N in (0, ∞), so that for every sequence (xk)k∈N in A with xk ∈ Cn1,··· ,nk,n for all k ∈ N A barrel set in a locally convex space E is a subset which is convex, balanced, absorbing and closed. Moreover, E is called barrelled if each barrel in E is a zero neighborhood. By [45, Proposition 23.23], every Fr´echet space is barrelled. Let Λ be an index set directed under a (reflexive, transitive, anti-symmetric) relation ≺, let (Eα)α∈Λ be a family of locally convex spaces, and denote, for α ≺ β, by τ β α a continuous linear map of Eβ into Eα. Consider E =(cid:8)(xα)α∈Λ ∈ Yα∈Λ Eα : τ β α (xβ ) = xα whenever α ≺ β(cid:9). E is called the projective limit of the family (Eα)α∈Λ with respect to the mappings τ β α ). The topology of E is the α (α, β ∈ Λ, α ≺ β), and denoted by lim ←− projective topology with respect to the family {(Eα, tα, τα)}α∈Λ, where for each α ∈ Λ, tα is the topology of Eα and τα : E → Eα is the canonical map. The projective limit E is called reduced if for each α ∈ Λ, τα has dense range. By applying [61, II.5.4], every Fr´echet space is isomorphic to a projective limit of a sequence of Banach spaces. (Eα, τ β By a similar way, one can introduce the inductive limit of the family (Eα)α∈Λ. See also [61, II.6.3] and [45, section 24] for more information about the notion of inductive topology. A locally convex space E is called bornological (resp. ultra-bornological) if its topology is the inductive topology corresponding to a system (Eα → E)α∈Λ of normed (resp. Banach) spaces. Clearly, every ultra-bornological space is bornolog- ical. Also, every Fr´echet space is ultra-bornological [45, Remark 24.13]. A topological algebra A is an algebra, which is a topological vector space and the multiplication A × A → A ((a, b) 7→ ab) is separately continuous. Moreover, A is called a locally convex algebra if A in addition is a locally convex space. Let (A, pα)α∈Λ be a locally convex algebra. A left approximate identity (left a.i. for short) in A is a net (ei)i∈I in A, such that pα(eia − a) −→i 0 (a ∈ A) for each α ∈ Λ. The left a.i. is bounded (left b.a.i. for short) if the set {ei : i ∈ I} is bounded. It is called uniformly bounded (left u.b.a.i. for short) if sup{pα(ei) : i ∈ I, α ∈ Λ} < ∞. 4 Z. ALIMOHAMMADI AND A. REJALI "Right a.i.", "right b.a.i.", and "right u.b.a.i." are defined similarly. Furthermore, for two-sided approximate identities, we use of the symbols "a.i.", "b.a.i.", and "u.b.a.i.". Following [27], a complete topological algebra is a Fr´echet algebra if its topology is produced by a countable family of increasing submultiplicative seminorms. Let A be a Fr´echet algebra. A left Fr´echet A-module is a Fr´echet space E together with the structure of an A-module such that the bilinear mapping A × E → E, (a, x) 7→ ax is continuous. Similarly, right Fr´echet A-modules and Fr´echet A-bimodules are defined. Suppose that E and F are left Fr´echet A-modules (resp. right Fr´echet A-modules, Fr´echet A-bimodules). In this paper, the space of all continuous A-module morphisms from E to F is denoted by AL(E, F ) (resp. LA(E, F ), ALA(E, F )). 2. Comparison between Banach and Fr´echet algebras In this section, we compare two Banach and Fr´echet algebras and provide a list of important similarities and differences between them. We use the symbol "(cid:3)" after each result to separate them from one another in the list. 2.1. Similarities. Following results are marked "S". S 2.1. Due to [45, Proposition 22.12], there exists a generalization of the Hahn- Banach theorem and its consequences for locally convex spaces. Let A be a locally convex space, B be a subspace of A, and p be a continuous seminorm on A. Then, the following statements hold. (i) For each f ∈ B∗ there exists an F ∈ A∗ such that F B = f . (ii) For each a ∈ A there exists an f ∈ A∗ such that f (a) = p(a) and f ≤ p. (iii) For each a ∈ A with a 6= 0 there exists an f ∈ A∗ such that f (a) 6= 0. Example. Consider the space CN with the fundamental system of seminorms pn(x) = max 1≤m≤n xm (x = (xn)n∈N ∈ CN). Let the subspace B consist of all convergent complex-valued sequences. By argu- ments above, the map T : B → C, defined by T (x) = lim n xn (x = (xn)n∈N ∈ B), has a continuous linear extension to CN. This extension is denoted by LIM; see [33, Theorem 1], [64, III.9] and also [54, section 4] for details. (cid:3) Before the next result, let us recall some notation. Let A be a locally convex space. The polars, of nonempty sets M ⊆ A and N ⊆ A∗, are defined by M ◦ = (cid:8)f ∈ A∗ : f (a) ≤ 1 for all a ∈ M(cid:9) ⊆ A∗, and ◦N = (cid:8)a ∈ A : f (a) ≤ 1 for all f ∈ N(cid:9) ⊆ A. S 2.2. Following [45, Theorem 22.13], the bipolar theorem is valid in the class of locally convex spaces. Let A be a locally convex space and B be an absolutely convex subset of A. Then, ◦(B◦) = B. (cid:3) S 2.3. We know that if A is a normed space, then by applying Banach-Alaoglu theorem, the closed unit ball of A∗ is weak∗-compact. The Alaoglu-Bourbaki the- orem is a generalization by Bourbaki to dual topologies on locally convex spaces. Let A be a locally convex space. Then, the polar U ◦ of any neighborhood U in A is weak∗-compact; see [45, Theorem 23.5]. (cid:3) FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 5 S 2.4. Let A and B be Banach spaces. Due to the open mapping theorem, if the map T : A → B is continuous, linear and surjective, then T is open. A similar result holds for Fr´echet spaces instead of Banach spaces. Following [45, Theorem 24.30], the open mapping theorem was generalized to the locally con- vex spaces. Let A and B be locally convex spaces. If A has a web and B is ultra- bornological, then every continuous, linear, surjective map T : A → B is open. (cid:3) S 2.5. Consider the continuous map f : A → B between two metric spaces A and B. Following [45], G(f ) =(cid:8)(a, f (a)) : a ∈ A(cid:9) ⊆ A × B, denote the graph of f , is closed. If A and B are also complete, then the converse is true. In general, the closed graph theorem holds in the class of locally convex spaces. Suppose that A and B are locally convex spaces. Exactly like S 2.4, if A has a web and B is ultra-bornological, then every linear map T : A → B with a closed graph is continuous [45, Theorem 24.31]. In particular, if A and B are Fr´echet spaces, then a linear mapping of A into B is continuous if and only if its graph is closed in A × B. (cid:3) S 2.6. In addition to the open mapping theorem and closed graph theorem, there is an extension of the Banach-Steinhaus theorem, known as uniform boundedness principle. Let A and B be locally convex spaces and A be barrelled. Further, let H ⊆ L(A, B) be pointwise bounded, i.e., for each a ∈ A, {T a : T ∈ H} is bounded in B. Then, H is equicontinuous. In other words, for each zero neighborhood V in B there exists a zero neighborhood U in A with T (U ) ⊆ V for all T ∈ H; see [19, Theorem III.§4.2.1] and [45, Proposition 23.27]. Corollary. Let A and B be locally convex spaces, A be barrelled and (Tn)n∈N be a sequence in L(A, B). Also, let (Tna)n∈N be convergent in B for each a ∈ A. Then, the mapping T : A → B, defined by is continuous [45, page 274]. (cid:3) T a = lim n→∞ Tna (a ∈ A), S 2.7. Let A be a Fr´echet algebra, E a right Fr´echet A-module, and F a left Fr´echet A-module. By applying [65, Theorem 2], the completed projective tensor product of E and F , denoted by E ⊗F , is again a Fr´echet algebra. Now, denote by E ⊗AF the completion of the quotient (E ⊗F )/N , where N ⊆ E ⊗F is the closed linear span of all elements of the form x · a ⊗ y − x ⊗ a · y (x ∈ E, y ∈ F, a ∈ A). Following [49, Proposition 3.2], there exists a vector space isomorphism from (E ⊗AF )∗ to LA(E, F ∗). (cid:3) Now, let A be a commutative Fr´echet algebra. For each a ∈ A, let a be the Gelfand transform of a defined by a(ϕ) := ϕ(a), where ϕ is an arbitrary nonzero complex-valued algebra homomorphism on A. Note that every complex- valued algebra homomorphism on a Banach algebra is automatically continuous. However, the question of whether each homomorphism on a Fr´echet algebra is necessarily continuous was posed by E. A. Michael [46]. This question has not been solved. But, only partial answers have been obtained so far; for example see [37] as mentioned in the introduction. See also [27, chapter 10] for details. In this paper, the set of all nonzero complex-valued and continuous algebra homomorphisms on A is denoted by ∆(A). Also, the topology on ∆(A) is the Gelfand topology, that is the coarsest topology such that all Gelfand transforms are continuous functions on ∆(A). 6 Z. ALIMOHAMMADI AND A. REJALI S 2.8. Let A be a commutative unital Fr´echet algebra. Following [27, 3.2], ∆(A) is never empty. Therefore, the Gelfand-Mazur theorem holds for Fr´echet algebras. In other words, if every nonzero element in A is invertible, then A = C. (cid:3) Let B be a Banach algebra with a left b.a.i. (not necessarily u.b.a.i.) and let E be a left Banach B-module. Due to the Cohen factorization theorem, for given ε > 0 and x ∈ B · E, there are a ∈ B and y ∈ E such that x = a·y and kx−yk < ε. Moreover, kak < M , for some M ∈ R+; see [25, page 14]. We mention next an important generalization by Voigt of the Cohen factorization theorem; see [67] for details. S 2.9. Let (A, pℓ) be a Fr´echet algebra and let (ei)i∈I be a left u.b.a.i. for A. Suppose that E is a left Fr´echet A-module and B ⊆ E is bounded. Let us recall aco(B) denote the absolutely convex hull of B, that is aco(B) =(cid:8) nXi=1 λixi : xi ∈ B, λi ∈ C, nXi=1 λi ≤ 1(cid:9). Set B := aco(B) and EB := (lin(B), pB), where pB is the gauge (or Minkowski functional) of B, i.e., the non-negative real function x 7→ pB(x) = inf{t > 0 : x ∈ tB} (x ∈ B). By [67, Theorem 3], if eix −→i x uniformly for x ∈ B, then (A, E) has the {B}-strong left factorization property (hereafter abbreviated {B}-SLFP). In other words, for every zero neighborhood U in E, there exist a ∈ A and a continuous linear mapping T : EB → E such that (i) a(T x) = x, for all x ∈ B; (ii) T x belongs to the closed A-submodule of E generated by x, for all x ∈ B; (iii) T x − x ∈ U , for all x ∈ B; (iv) T B : B → E is continuous. Furthermore, a ∈ A can be chosen to satisfy (v) if ℓ ∈ N is such that sup{pℓ(ei) : i ∈ I} ≤ 1, then pℓ(a) ≤ 1. Remark. Let B be a collection of bounded (resp. compact) subsets of E. Following [67], we say that (A, E) has the B-SLFP if for each B ∈ B, (A, E) has the {B}- SLFP. In particular, if B is the collection of bounded (resp. compact) subsets of A, then this property is called the bounded (resp. compact) SLFP. Note that by [67, Corollary 5], if A has a left u.b.a.i., then A has the compact SLFP. However, factorization properties are proved for certain Fr´echet algebras of differentiable functions which do not possess a u.b.a.i.; see [68]. (cid:3) 2.2. Differences. Following results are marked "D". D 2.10. Every Banach algebra (space) is a Fr´echet algebra (space), but in general the converse is not true. Let (An, k · kn)n∈N be a sequence of normed spaces and A =Qn∈N An. Define the metric d on A by kan − bnkn d(a, b) := ∞Xn=1 2n(cid:0)1 + kan − bnkn(cid:1) (cid:0)a = (an)n∈N, b = (bn)n∈N ∈ A(cid:1). Following [45, Lemma 5.17], (A, d) is a locally convex metric linear space. Also, if each (An, k · kn) is complete, then (A, d) is a Fr´echet space. However, (A, d) is not a Banach space if An 6= {0} for infinitely many n ∈ N. For instance, the space CN of all sequences in C with the metric d(x, y) = ∞Xn=1 xn − ynn 2n(cid:0)1 + xn − ynn(cid:1) (cid:0)x = (xn)n∈N, y = (yn)n∈N ∈ CN(cid:1), FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 7 is a Fr´echet space which is not a Banach space; see [45, Examples 5.18]. (cid:3) D 2.11. The dual space of a Banach space is also a Banach space. Moreover, if A is a Banach algebra, then the module actions of A are jointly continuous on A∗. However, this result is not valid for a Fr´echet case. Let C∞(R), denote the space of all infinitely differentiable functions on R. Following [66], C∞(R) is a Fr´echet algebra with the pointwise product and the family of countable seminorms k · kn, defined by kf kn := 1 k! kf (k)k∞,n (f ∈ C∞(R)). nXk=0 The module action of C∞(R) is not jointly continuous on C∞(R)∗. Therefore, C∞(R)∗ is not a Fr´echet space. By arguments above, we immediately get the following. Lemma. Let A be a metrizable locally convex space. The following statements are equivalent: (i) A is a Banach space; (ii) A is a Fr´echet and (DF)-space. Corollary. Let A be a Fr´echet space. Then, A∗ is a Fr´echet space if and only if A is a Banach space. See also [27] and [45]. (cid:3) D 2.12. For Banach spaces A and B, L(A, B) is also a Banach space; see [45, Proposition 5.6]. However, by using D 2.11 and [45, page 253], L(A, B) is not necessarily a Fr´echet space when A and B are Fr´echet spaces. (cid:3) Before proceeding to the next result, recall that a matrix A = (aj,k)j,k∈N of non- negative numbers is called a Kothe matrix if it satisfies the following conditions. (i) For each j ∈ N there exists a k ∈ N with aj,k > 0. (ii) aj,k ≤ aj,k+1 for all j, k ∈ N. Now, for 1 ≤ p < ∞ define Also, for p = ∞ and p = 0 define λp(A) :=(cid:8)x ∈ CN : kxkk :=(cid:0) ∞Xj=1 := (cid:8)x ∈ CN : kxkk := sup := (cid:8)x ∈ λ∞(A) : < ∞ for all k ∈ N(cid:9). xj aj,kp(cid:1)1/p xjaj,k < ∞ for all k ∈ N(cid:9), and xj aj,k = 0 for all k ∈ N(cid:9), lim j→∞ j∈N λ∞(A) c0(A) respectively. Using the definition above, we can present in D 2.13 an example of non-distinguished Fr´echet space due to Meise and Vogt [45]. Note that a metrizable locally convex space A is distinguished if its strong dual is a bar- relled or bornological locally convex space. Consider the Fr´echet space A and n )n∈N. By applying [45, Remark 25.13], if A is distinguished, let A = lim ←− then A∗ ∼= lim n, (τ m n )∗)n∈N as two locally convex spaces. In other words, the −→ inductive topology coincides with the strong topology on A∗ if and only if A is distinguished. This is an important advantage of distinguished spaces. (An, τ m (A∗ D 2.13. Every Banach space is distinguished [61, Corollary II.7.1]. In general, however, a Fr´echet space fails to be distinguished. For example, consider the Kothe matrix A = (ai,j;k)(i,j)∈N2,k∈N such that (i) for each i, j, k ∈ N with k ≤ i, ai,j;k = ai,j;1 > 0; 8 Z. ALIMOHAMMADI AND A. REJALI (ii) for each m ∈ N, limj→∞ am,j;ma−1 m,j;m+1 = 0. By using [45, Corollary 27.18], λ1(A) is a Fr´echet space which is not distinguished. This example is a generalization of a classical counter-example of Grothendieck and Kothe; see [42, §31.7]. (cid:3) D 2.14. Following [45, Proposition 7.5], if A is a reflexive Banach space and B is a closed subspace of A, then A/B is reflexive. Also, every closed subspace of a reflexive Fr´echet space is reflexive [45, Proposition 23.26]. But, there are reflexive Fr´echet spaces with non-reflexive quotients. Consider the Kothe matrix A = (ai,j;k)(i,j)∈N2,k∈N, where ai,j;k :=((ki)k kj for j < k, i ∈ N for j ≥ k, i ∈ N. By [45, Example 27.21], for each infinite subset I of N2 and each n ∈ N there exists k ∈ N such that inf (i,j)∈I ai,j;na−1 i,j;k = 0. Thus, due to Dieudonn´e-Gomes theorem [45, Theorem 27.9], λ1(A) is reflexive. Since ai,j;1 = 1 and ai,k;k = a1,k;k, λ1(A) has a quotient which is isomorphic to ℓ1; see [45, Proposition 27.22]. However, by applying [45, Corollary 7.10], ℓ1 is non-reflexive. (cid:3) For the next result, we need to remind the reader that a Montel space A is a quasi-barrelled space in which each bounded set is relatively compact. Let A be the Kothe matrix defined as in D 2.14. By [45, Theorem 27.9], λp(A) is a Montel space for each p ∈ [1, ∞]. For this reason, λp(A)∗ is also a Montel space and so is re- flexive; see [45, Proposition 24.25]. Following the proof of [45, Proposition 27.22], the continuous linear map Q : λp(A) → ℓp for p ∈ [1, ∞) defined by Qx = ( 2−jxi,j )i∈N ∞Xj=1 (cid:0)x = (xi,j )i,j ∈ λp(A)(cid:1) is onto. Now, consider Q∗ : (ℓp)∗ = ℓq → λp(A)∗, where q = p p−1 . By us- ing [45, Corollary 27.24], rang(Q∗) is a closed subspace of λp(A)∗ which is non- reflexive. This example shows that not every closed subspace of a reflexive locally convex space is reflexive. D 2.15. Let A be a Banach space and B be a closed subspace of A. Following [45, page 293], A is finite dimensional if and only if it is a Montel space. There- fore, B and A/B are also Montel spaces. Clearly, for Fr´echet spaces, every closed subspace of a Montel space is a Montel space. But, Montel property are in gen- eral not inherited by quotients. For instance, consider the montel space λ1(A) in D 2.14 and the space ℓ1 which is not Montel. (cid:3) Before continuing, we must recall that for a continuous linear map T : A → B between locally convex spaces A and B, there is a unique map ¯T in L(A/ker(T ), B) such that T = ¯T ◦ q, where q : A → A/ker(T ) is the quotient map [45, Proposition 22.11]. Now, consider rang(T ) with the relative topology induced by B. The map T is said to be a topological homomorphism, if ¯T is an isomorphism between A/ker(T ) and rang(T ). D 2.16. Consider the short sequence of Fr´echet spaces with continuous linear maps (2.1) {0} → A T→ B S→ C → {0}. By applying [45, Theorem 26.3], if the short sequence (2.1) is exact, then it is already topologically exact, that is T and S are topological homomorphisms. Also, FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 9 by [45, Proposition 26.4], the short sequence (2.1) is exact if and only if (2.2) {0} → C ∗ S ∗ → B∗ T ∗ → A∗ → {0} is exact. However, in general the short sequence (2.2) is not topologically ex- act. Consider the Kothe matrix A and Q ∈ L(λp(A), ℓp) defined in D 2.14. By [45, Remark 27.23], the dual sequence of the short exact sequence {0} → ker(Q) ֒→ λp(A) Q → ℓp → {0} of Fr´echet spaces is not topologically exact. Remark. Let A be a Fr´echet space and B be a closed subspace of A. Consider the short exact sequence {0} → B j → A q → A/B → {0}, where j : B → A is the inclusion and q : A → A/B is the quotient map. If the dual sequence {0} → (A/B)∗ q∗ → A∗ j∗ → B∗ → {0}, is topologically exact, then we obtain the canonical isomorphisms B∗ ∼= A∗/B◦ and (A/B)∗ ∼= B◦; see [45, Remark 26.5]. (cid:3) In S 2.8, we mentioned an important result concerning the spectrum of Banach and Fr´echet algebras. In the following, some other results in this field have been presented. D 2.17. Let B be a commutative unital Banach algebra. Then, ∆(B) is a nonempty compact space. Following [27, 3.2], in general the spectrum of a commutative Fr´echet algebra A is not a compact space. But, it is a nonempty hemicompact space. In fact, there exists a countable compact exhaustion (Kn)n of A such that for every compact subset K ⊆ A there is n ∈ N so that K ⊆ Kn. (cid:3) Before the next result, we recall that an ideal I of a Fr´echet algebra A is called maximal if I 6= A and I is contained in no other proper ideal of A. D 2.18. Every maximal ideal in a commutative Banach algebra B is closed. Hence, there exists a one-to-one correspondence between the maximal ideals of B and the elements of ∆(B). Opposite to this result maximal ideals of commu- tative Fr´echet algebras are in general not closed. For example, C(R) has some maximal ideal which is not closed; see [27, Example 3.2.11]. (cid:3) D 2.19. Recall that a Hausdorff space is called a k-space if every subset inter- secting each compact subset in a closed set is itself closed. Following [27, chapter 3], examples of k-spaces are locally compact and first countable spaces. Also, the spectrum of a commutative Banach algebra is a k-space. But, there are Fr´echet algebras whose spectrum is not a k-space; see [27, Example 7.1.2]. (cid:3) Suppose that A is a commutative Fr´echet algebra. Let us recall that a point ϕ ∈ ∆(A) is said to be a local peak point for A if there exist a neighborhood U of ϕ in ∆(A) and an element a ∈ A such that a(ϕ) = 1 and a(ψ) < 1 (ψ ∈ U \ {ϕ}). Moreover, a is called a peak point for A if U can be taken to be ∆(A). D 2.20. In the class of Banach algebras, by applying the Rossi's maximum princi- ple, one can see that every locall peak point is a peak point; see [27, Theorem 1.4.9]. However, this result is not valid for Fr´echet algebras; see also [27, Example 9.2.2]. Therefore, the Rossi's maximum modulus theorem fails to be true for Fr´echet al- gebras. (cid:3) 10 Z. ALIMOHAMMADI AND A. REJALI D 2.21. Let B be a commutative unital Banach algebra and let B−1 denote the set of all invertible elements of B. Due to the Arens-Royden theorem, for every f ∈ C(∆(B))−1 there exist b ∈ B−1 and g ∈ C(∆(B)) such that f = b exp(g), It is an open question, whether the Arens-Royden theorem can be extended to Fr´echet algebras. However, following [27], this theorem was proved for a special class of Fr´echet algebras. (cid:3) where exp(g) = P∞ n=0 gn/n!; see [27, Theorem 1.4.8]. 3. Some important generalizations in harmonic analysis In this section, we collect some important generalizations to the Fr´echet al- gebras which have been obtained in recent years. We mark these results by the letter "G", and we use the symbol "(cid:3)" after each one of them. Let B be a Banach algebra. We recall that a left Banach B-module F is called flat if for every exact admissible sequence of right Banach B-modules, the sequence {0} → E1 → E2 → E3 → {0} {0} → E1 ⊗BF → E2 ⊗BF → E3 ⊗BF → {0} is exact. Also, B is amenable if for each Banach B-bimodule E every continuous derivation from B to the dual bimodule E∗ is inner. This definition was introduced by Johnson in [39]. Later, Helemskii and Sheinberg [34] showed that a Banach algebra B is amenable if and only if B♯, the unitization of B, is a flat Banach B-bimodule. In this case, B has a b.a.i. [59, Proposition 2.2.1]. In recent years, a large number of papers concerning amenability of Banach algebras have appeared; see for instance [5, 12, 18, 28, 29, 32, 44]. In [49], the arguments above have generalized as follows: G 3.1. Due to Pirkovskii [49], a Fr´echet algebra A is amenable (resp. biflat) if A♯ (resp. A) is a flat Fr´echet A-bimodule. Also, we say that A has a locally b.a.i. if for each zero neighborhood U ⊆ A there exists C > 0 such that for each finite subset F ⊆ A and each ε > 0 there exists b ∈ CU with a − ab ∈ εU and a − ba ∈ εU for all a ∈ F . Note that if A is normable, then the notions of "b.a.i." and "locally b.a.i." are equivalent [49, Remark 6.4]. By applying [49, Lemma 9.4], a Fr´echet algebra A with a locally b.a.i. is amenable if and only if A is biflat. Furthermore, A is amenable if and only if it is isomorphic to a reduced projective limit of a sequence of amenable Banach algebras [49, Theorem 9.5]. Let us note the following theorem proved by A. Pirkovskii [49, Theorem 9.6]. Theorem. Let A be a Fr´echet algebra. The following conditions are equivalent: (i) A is amenable; (ii) for each Banach A-bimodule E, every continuous derivation from A to E∗ is inner; (iii) for each Fr´echet A-bimodule E, every continuous derivation from A to E∗ is inner. Pirkovskii also introduced the concepts of "virtual diagonal", "bounded approxi- mate diagonal", and "locally bounded approximate diagonal" in the class of Fr´echet algebras. He then extended Johnson's theorem (see [38, Lemma 1.2] and [38, The- orem 1.3]; see also [59, Theorem 2.2.4]). Indeed, for a Fr´echet algebra A, the following statements are equivalent: (i) A is amenable; (ii) A has a locally bounded approximate diagonal. If, in addition, A is quasinormable, then (i) and (ii) are equivalent to the following: (iii) A has a bounded approximate diagonal; FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 11 (iv) A has a virtual diagonal. See [49, Theorem 9.7]. (cid:3) G 3.2. Ghahramani and Loy [26] introduced and studied approximate amenabil- ity and contractibility of Banach algebras. These notions were investigated in the class of Fr´echet algebras by Lawson and Read [43]. A Fr´echet algebra A is called approximately contractible if given any A-bimodule E, and any continuous derivation D : A → E, there is a net (xα) ⊆ E such that D(a) = lim α (a · xα − xα · a) (a ∈ A). Every continuous derivation D with this property is said to be approximately inner. Also, we recall that a Fr´echet algebra A is approximately amenable if given any A-bimodule E, every continuous derivation D : A → E∗ is approximately inner. The authors in [43] generalized the important results of Banach algebras in this field. Following [26], all known approximately amenable Banach algebras have b.a.i. It is nonetheless hoped that it would yield Banach algebras without b.a.i. which had a form of amenability. Note that there are examples of Fr´echet algebras which are approximately contractible, but which do not have a b.a.i. For a good many Fr´echet algebras without b.a.i., it was obtained either that the algebra is approximately amenable, or it is "obviously" not approximately amenable because it has continuous point derivations; see [43, section 3]. Therefore, the situation for Fr´echet algebras is quite close to what was hoped for Banach algebras. (cid:3) Weakly amenable Banach algebras were first introduced by Bade, Curtis and Dales in [16]. The general definition is due to Johnson [39]. Later, Dales, Ghahra- mani and Grønbaek [24] introduced and studied the notion of n-weak amenabil- ity of Banach algebras. Indeed, a Banach algebra B is n-weakly amenable if H 1(B, B(n)) = {0}, where B(n) is the n-th dual space of B. In particular, B is weakly amenable if it is 1-weakly amenable. Recently, several papers dealing with this subject have appeared; see for example [28, 53, 60]. G 3.3. Let A be a Fr´echet algebra. The notion of weak amenability of A was stud- ied by Rejali et al. [11]. In fact, the authors showed that some results in the field of weak amenability of Banach algebras can be generalized for the Fr´echet alge- bras. For example, if A is weakly amenable, then A is essential [11, Theorem 2.3]. Moreover, if I is a quasinormable closed ideal in the Fr´echet algebra A such that A/I and I are weakly amenable, then A is weakly amenable, as well; see [11, Theorem 2.6]. (cid:3) G 3.4. Let B be a Banach algebra and I be a closed ideal of B. The notions of I- weak amenability and ideal amenability of B was introduced and studied by Gorgi and Yazdanpanah in [30]. Indeed, B is I-weakly amenable if every continuous derivation D : B → I ∗ is inner. Moreover, B is called ideally amenable if B is I-weakly amenable for every closed ideal I of B. Clearly, every amenable Banach algebra is ideally amenable, and every ideally amenable Banach algebra is weakly amenable. However, the concepts of I-weak amenability and ideal amenability are different from amenability and weak amenability of Banach algebras (ibid.). In [55], Ranjbari and Rejali generalized these concepts and their results to the Fr´echet algebras. Later, the concept of vector-valued Lipschitz algebras in the class of Fr´echet algebras was studied by the same authors in [54]. Especially, they Proved that Lemma 2.1 of [18] holds for Fr´echet algebras under a certain condition. Then, they studied ideal amenability of vector-valued Fr´echet Lipschitz algebras; see [54, section 4] for details. (cid:3) 12 Z. ALIMOHAMMADI AND A. REJALI Let A be a Banach algebra with a b.a.i. which is contained as closed ideal in a Banach algebra B. Consider the pseudo-unital Banach A-bimodule E, that is E = A · E · A = {a · x · b : a, b ∈ A, x ∈ E}. By using [59, Proposition 2.1.6], E is a Banach B-bimodule in a canonical fachion. Moreover, for each derivation D : A → E∗, there is a unique extension derivation eD : B → E∗ of D such that eD is continuous with respect to the strict topology on B and the weak∗-topology on E∗. In [2], we studied the strict topology on a Fr´echet algebra and then we obtained a similar result as above. G 3.5. Let A be a closed ideal in a Fr´echet algebra (B, pℓ)ℓ∈N, satisfying (cid:8)b ∈ B : b · A = A · b = {0}(cid:9) = {0}, and suppose that B has a b.a.i. with elements in A. Let (E, qn)n∈N be an arbi- trary pseudo-unital Fr´echet A-bimodule. Then, for every continuous derivation D : A → E∗, with respect to the topology induced by (pℓ)ℓ∈N on A and the strong topology on E∗, there exists a unique derivation eD : B → E∗ such that eD is con- tinuous with respect to the strict topology on B and the weak∗-topology on E∗; see [2, Theorem 3.2]. (cid:3) G 3.6. Following [9] and also [10], a Fr´echet algebra (A, pℓ)ℓ∈N is called a Segal Fr´echet algebra in the Fr´echet algebra (B, qn)n∈N if the following conditions hold: (1) A is a dense ideal in (B, qn)n; (2) the identity map (A, pℓ)ℓ → (B, qn)n is continuous; (3) the map (B, qn)n × (A, pℓ)ℓ → (A, pℓ)ℓ with (a, b) 7→ ab, for all a, b ∈ A, is jointly continuous. This definition is in fact a generalization of the definition abstract Segal algebras, in the Banach case; see [20]. Also in [21, Definition 14], this definition has been introduced in the field of topological algebras, as follows: A topological algebra (A, τA) is called a (B, τB)-Segal algebra in the topological algebra (B, τB) if the following assertions hold: (1)′ A is a dense left ideal in (B, τB); (2)′ the identity map from (A, τA) into (A, τB) is continuous; (3)′ the multiplication (B, τB) × (A, τA) → (A, τA), ((a, b) 7→ ab), is continuous. In [9], Rejali et al. investigated Theorem 1.2 of [20] for Segal Fr´echet algebras and obtained an analogous result. Indeed, if (A, pℓ)ℓ∈N is a proper Segal Fr´echet algebra in (B, qn)n∈N which contains an a.i. (ei)i∈I , then (ei)i∈I cannot be bounded in (A, pℓ)ℓ∈N [9, Proposition 3.5]. Furthermore, the authors characterized closed ideals of Segal Fr´echet algebras, and showed that the ideal theorem is also valid for Fr´echet algebras. In other words, every closed ideal of any Segal Fr´echet al- gebra (A, pℓ)ℓ∈N in (B, qn)n∈N is the intersection of a closed ideal of B with A [9, Theorem 3.8]. (cid:3) Suppose that A is a commutative Fr´echet algebra and M (A) is the multiplier algebra with respect to the strict topology. As mentioned in section 0, A is a M (A)-Segal algebra; see [2, section 4] for details. In [14], we studied the notion of multiplier algebra of Fr´echet algebras. Let B be a Banach algebra and ϕ ∈ ∆(B). Kaniuth, Lau and Pym [40] in- troduced and studied the concept of (right) ϕ-amenability for Banach algebras as a generalization of left amenability of Lau algebras. We recall that, B is called FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 13 (right) ϕ-amenable if there exists a continuous linear functional m on B∗ such that m(ϕ) = 1 and m(f ·a) = ϕ(a)m(f ) for all a ∈ B and f ∈ B∗. Furthermore, the no- tion of right character amenability of Banach algebras was introduced and studied by Monfared [47]. Indeed, a Banach algebra B is called right character amenable if for each ϕ ∈ ∆(B) ∪ {0} and all Banach B-bimodules E with the left module action a · x = ϕ(a)x, each continuous derivation D : B → E∗ is inner. Alagh- mandan, Nasr-Isfahani, and Nemati [12] characterized the character amenability of abstract Segal algebras. Specifically, they showed that ϕ-amenability of B is equivalent to ϕA-amenability of A, where A is an abstract Segal algebra in B. Right ϕ-amenability and right character amenability for Fr´echet algebras were introduced and studied in [7]. In the following, some related results about Segal Fr´echet algebras are provided. G 3.7. Let (A, pℓ)ℓ∈N be a Segal Fr´echet algebra in a Fr´echet algebra (B, qn)n∈N. Precisely, similar to Lemma 2.2 of [12], ∆(A) = {ϕA : ϕ ∈ ∆(B)}; see [7, Lemma 6.7]. Also, if ϕ ∈ ∆(B), then B is right ϕ-amenable if and only if A is right ϕA-amenable [7, Theorem 6.8]. The following is now immediate from results above. Corollary. Let (B, qn)n∈N be a Fr´echet algebra and let (A, pℓ)ℓ∈N be a Segal Fr´echet algebra in B. Suppose that B is right character amenable. Then, A is right ϕ-amenable for each ϕ ∈ ∆(A); see [7, Corollary 6.9]. (cid:3) In addition to the above, we joint with Abtahi [1] generalized the notion of Connes amenability of Banach algebras. We then proved some of the previous available results about this notion for Fr´echet algebras. Rahnama and Rejali [51, 52] also introduced and studied the concepts of amenability modulo an ideal of Fr´echet algebras and locally bounded approximate diagonal modulo an ideal of Fr´echet algebras. Finally, the notion of BSE algebras in the class of Fr´echet algebras was investigated by Rejali et al. in [3, 4, 15]. Before continuing, let us note the following arguments. Let B be a Banach algebra. We know that on the bidual space B∗∗ of B, there are two multiplications, called the first and second Arens products, which make B∗∗ into a Banach algebra. Moreover, B is called Arens regular if these two products coincide on B∗∗. There are many papers concerning Arens regularity. Also, this notion was investigated by Rejali et al. in several papers; see for example [6, 17, 41, 56, 57, 58]. Note that if B∗∗ is Arens regular, then so is B. However, Pym [50] gives an example of an Arens regular Banach algebra B such that B∗∗ is not Arens regular. Also, look at the following example. Example. Consider the Banach algebras An = ℓ1(Z, w 1 (1 + t)α (t ∈ Z) for each α > 0. Let ) (n ∈ N), where wα(t) = n A = c0-Yn∈N B = ℓ∞-Yn∈N An =(cid:8)(an) ∈ Yn∈N An =(cid:8)(an) ∈ Yn∈N An : kankn → 0(cid:9) and kankn < ∞(cid:9). An : sup n∈N The following statements hold. (i) by [25, page 35], A is Arens regular. (ii) by [25, Example 9.2], B is not Arens regular. (iii) B is a closed sualgebra of A∗∗ and consequently A∗∗ is not Arens regular. Let G be a locally compact (topological) group with a weight function w on it. Rejali and Vishki [58] showed that the following statements are equivalent: 14 Z. ALIMOHAMMADI AND A. REJALI (i) L1(G, w)∗∗ is Arens regular; (ii) L1(G, w) is Arens regular; (iii) L1(G) is Arens regular or Ω : G × G → (0, 1], defined by Ω(x, y) = w(xy) w(x)w(y) (x, y ∈ G), is 0-cluster, that is limm limn Ω(xm, yn) = 0 = limn limm Ω(xm, yn), for sequences (xm) and (yn) in G with distinct elements. Moreover, Baker and Rejali proved that for a semigroup S with a weight function w on it, ℓ1(S, w) is Arens regular, whenever ℓ1(S) is [17]. Now, the following question is natural. Question. Let S be a semigroup and w be a weight function on S. Are the following statements equivalent? (i) ℓ1(S, w)∗∗ is Arens regular. (ii) ℓ1(S, w) is Arens regular. (iii) ℓ1(S) is Arens regular or Ω is 0-cluster. Later, Rejali et al. showed that if the set of all idempotents of an inverse semigroup S is finite, then ℓ1(S) is Arens regular if and only if S is finite; see [6, Theorem 3.8]. Also in this case, if S admits a bounded below weight for which Ω is 0-cluster, then S is countable; see [41, Theorem 3.4]. By arguments above, another question arises. Question. Let S be an inverse semigroup with finitely many idempotents and w be a weight function on S. Are the following statements equivalent? (i) ℓ1(S, w)∗∗ is Arens regular. (ii) ℓ1(S, w) is Arens regular. (iii) S is finite or Ω is 0-cluster. Remark. In [6, Example 3.9], the authors showed that there exists an inverse semigroup S with infinitely many idempotents such that ℓ1(S) is Arens regular. Arens regularity of Fr´echet algebras was defined by Zivari-Kazempour [69]. Indeed, a Fr´echet algebra A is called Arens regular if the first and second Arens products coincide on A∗∗. Later, we [13] obtained some results in this field. G 3.8. In 1966, Gulick [31] studied Arens products on the bidual space of locally multiplicatively-convex topological (lmc for short) algebras with a hypo-continuous multiplication. We recall that the product A × A → A, on a lmc algebra A, is left (resp. right) hypo-continuous if for each zero neighborhood U and for each bounded set B of A there exists a zero neighborhood V such that V B ⊆ U (resp. BV ⊆ U ). The multiplication in A is called hypo-continuous if it is both left and right hypo-continuous. Now, let A be a Fr´echet algebra. Clearly, A belongs to the class of lmc algebras. Moreover, by applying [61, III.5.1], the multiplication in A is hypo-continuous. Therefore, due to Gulick [31] and [45, Proposition 25.9], A∗∗ is always a Fr´echet algebra under the strong topology and Arens products on A∗∗. Furthermore, Gulick generalized a theorem of Civin and Yood [22] to the lmc algebras, especially Fr´echet algebras. He introduced a bicommutative algebra, that is an lmc algebra whose bidual is commutative. Then, he showed that subalgebras of bicommutative algebras are also bicommutative [31, Theorem 5.3]. (cid:3) G 3.9. Following [22], a Banach algebra B has a weak right identity if there exists a net (eα)α∈Λ in B and an M > 0 such that for every α ∈ Λ, keαk < M and lim α f (aeα − a) = 0 (a ∈ B, f ∈ B∗). FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 15 Civin and Yood proved that B has a weak right identity if and only if B∗∗ has a right unit with respect to the first Arens product [22, Lemma 3.8]. There is a generalization of this theorem in the class of lmc algebras especially Fr´echet algebras. Zivari-Kazempour proved that an lmc algebra A has a right b.a.i. (resp. left b.a.i) if and only if A∗∗ has a right (resp. left) unit with respect to the first (resp. second) Arens product [69, Theorem 3.2]. (cid:3) Acknowledgment. The authors would like to thank the Banach algebra cen- ter of Excellence for Mathematics, University of Isfahan. References [1] F. Abtahi, Z. Alimohammadi, and A. Rejali, Connes amenability of Fr´echet algebras, sub- mitted. [2] F. Abtahi, Z. Alimohammadi, and A. Rejali, On the strict, uniform and compact-open topolo- gies on an algebra, Period. Math. Hungar. 77 (2018), no. 2, 209 -- 223. [3] F. Abtahi, M. Amiri, and A. Rejali, The Bochner-Schoenberg-Eberline property for commu- tative Fr´echet algebras, submitted. [4] F. Abtahi, M. Amiri, and A. Rejali, The Bochner-Schoenberg-Eberline property for Fr´echet C ∗-algebras and uniform Fr´echet algebras, submitted. [5] F. Abtahi, M. Azizi, and A. Rejali, Character amenability of the intersection of Lipschitz algebras, Canad. Math. Bull. 60 (2017), no. 4, 673 -- 689. [6] F. Abtahi, B. Khodsiani, and A. Rejali, Arens regularity of inverse semigroup algebras, Bull. Iranian Math. Soc., 40 (2014), no. 6, 1527 -- 1538. [7] F. Abtahi, S. Rahnama, and A. Rejali, ϕ-amenability and character amenability of Fr´echet algebras, Forum. Math. (2018), DOI 10.1515/forum-2016-0116. [8] F. Abtahi and S. Rahnama, ϕ-contractibility and character contractibility of Fr´echet algebras, Ann. Funct. Anal. 8 (2017), no. 1, 75 -- 89. [9] F. Abtahi, S. Rahnama, and A. Rejali, Segal Fr´echet algebras, Anal. Math. (2017), DOI 10.1007/s10476-017-0601-y. [10] F. Abtahi, S. Rahnama, and A. Rejali, Semisimple Segal Fr´echet algebras, Period. Math. Hungar. 71 (2015), no. 2, 146 -- 154. [11] F. Abtahi, S. Rahnama, and A. Rejali, Weak amenability of Fr´echet algebras, U.P.B. Sci. Bull., Series A, 77 (2015), no. 4, 93 -- 104. [12] M. Alaghmandan, R. Nasr-Isfahani, and M. Nemati, Character amenability and contractibil- ity of abstract Segal algebras, Bull. Aust. Math. Soc. 82 (2010), no. 2, 274 -- 281. [13] Z. Alimohammadi, and A. Rejali, On the Arens regularity of Fr´echet algebras and their biduals, submitted. [14] Z. Alimohammadi, and A. Rejali, On the multiplier algebra of Fr´echet algebras, submitted. [15] M. Amiri, A. Rejali, The multiplier algebra and B.S.E. property of the direct sum of Fr´echet algebras, submitted. [16] W. G. Bade, P. C. Curtis, and H. G. Dales, Amenability and weak amenability for Bearling and Lipschitz algebras, Proc. London Math. Soc. 55 (1987), no. 3, 359-377. [17] J. W. Baker and A. Rejali, On the Arens regularity of weighted convolution algebras, J. London Math. Soc. (2), 40 (1989), no. 3, 535 -- 546. [18] E. Biyabani and A. Rejali, Approximate and Character amenability of vector-valued Lips- chitz algebras, Bull. Korean Math. Soc. 55 (2018), no. 4, 1109 -- 1124. [19] N. Bourbaki, Topological vector spaces, Elements of mathematics, Springer (1987). [20] J. T. Burnham, Closed ideals in subalgebras of Banach algebras, Proc. Amer. Math. Soc. 32, (1972), no.2, 551 -- 555. [21] J. T. Burnham, Segal algebras and dense ideals in Banach algebras, Functional analysis and its applications, (Internat. Conf., Eleventh Anniversary of Matscience, Madras; 1973 dedicated to Alladi Ramakrishnan), Springer, Berlin, 1974, 33 -- 58 Lecture Notes in Math., 399. [22] P. Civin and B. Yood, the second conjugate space of a Banach algebra as an algebra, Pacific J. Math., 11 (1961), 820 -- 847. [23] J. B. Conway, A course in functional analysis, Graduate Texts in Math., Vol. 96, Springer- Verlag, 1985. [24] H. G. Dales, F. Ghahramani, and N. Grønbaek, Derivations into iterated duals of Banach algebras, Studia Math. 128 (1998), no. 1, 19 -- 54. [25] H. G. Dales and A. T.-M. Lau, The second duals of Beurling algebras, Mem. Amer. Math. Soc., 177 (2005). 16 Z. ALIMOHAMMADI AND A. REJALI [26] F. Ghahramani and R. J. Loy, Generalized notions of amenability, J. Funct. Anal. 208 (2004), 229 -- 260. [27] H. Goldmann, Uniform Fr´echet algebras, North-Holland Mathematics Studies, 162. North- Holand, Amesterdam-New York, 1990. [28] M. E. Gordji, F. Habibian, and A. Rejali, Module extension of dual Banach algebras, Bull. Korean Math. Soc. 47 (2010), no. 4, 663 -- 673. [29] M. E. Gordji, A. Jabbari, and G. Hui Kim, Some characterizations of character amenable Banach algebras, Bull. Korean Math. Soc. 52 (2015), no. 3, 761 -- 769. [30] M. E. Gordji and T. Yazdanpanah, Derivations into duals of ideals of Banach algebras, Proc. Indian Acad. Sci. 114 (2004), no. 4, 399 -- 408. [31] S. L. Gulick, The bidual of a locally multiplicatively-convex algebra, Pacific J. Math., 17/1 (1966), 71 -- 96. [32] F. Habibian, H. Samea, and A. Rejali, Ideal amenability and approximate ideal amenability of matrix Banach algebras, Proc. Ro Acad. Series A, 3 (2011), 173 -- 178. [33] D. Hajdukovi´c, The functionals of the kind of Banach limits, publications de L'Institut Mathematique, T. 19(33), 1975. [34] A. Ya. Helemskii and M. V. Sheinberg, Amenable Banach algebras (Russian), Funktsional. Anal. i Prilozhen. 13 (1979), no. 1, 42 -- 48; English transl.: Functional Anal. Appl. 13 (1979), no. 1, 32 -- 37. [35] A. Ya. Helemskii, On a method for calculating and estimating the global homological di- mension of Banach algebras, Mat. Sb., 87 (129) (1972), 122 -- 135. [36] A. Ya. Helemskii, The homology of Banach and topological algebras (Moscow University Press, English transl: Kluwer Academic Publishers, Dordrecht 1989). [37] T. G. Honary, Automatic continuity of homomorphisms between Banach algebras and Fr´echet algebras, Bull. Iranian Math. Soc. 32 (2006), no. 2, 1 -- 11. [38] B. E. Johnson, Approximate diagonals and cohomology of certain annihilator Banach al- gebras, Amer. J. Math. 94 (1972), 685 -- 698. [39] B. E. Johnson, Cohomology in Banach algebras, Mem. Amer. Math. Soc. 127 (1972). [40] E. Kaniuth, A. T. Lau, and J. Pym, on ϕ-amenability of Banach algebras, Math. Proc. Combridge Philos. Soc. 144 (2008), no.1, 85 -- 96. [41] B. Khodsiani, A. Rejali, and H. R. E. Vishki, Arens regularity of certain weighted semigroup algebras and countability, A. Semigroup Forum 92 (2016), 304 -- 310. [42] G. Kothe, topological vector space. (I), Spronger-Verlag, New York, 1969. [43] P. Lawson and C. J. Read, Approximate amenability of Fr´echet algebras, Math. Proc. Cambridge Philos. Soc., 145 (2008), 403 -- 418. [44] A. R. Medghalchi and T. Yazdanpanah, n-weak amenability and strong double limit prop- erty, Bull. Korean Math. Soc. 42 (2005), no. 2, 359 -- 367. [45] R. Meise and D. Vogt, Introduction to functional analysis, (Oxford Science Publications), 1997. [46] E. A. Michael, Locally multiplicatively convex topological algebras, Memoirs Amer. Math. Soc. 11 (1952). [47] M. S. Monfared, Character amenability of Banach algebras, Math. Proc. Combridge Philos. Soc. 144 (2008), no. 3, 697 -- 706. [48] A. Yu. Pirkovskii, Biprojective topological algebras of homological bidimension 1, J. Math. Sci. (New York) 111 (2002), no. 2, 3476 -- 3495. [49] A. Yu. Pirkovskii, Flat cyclic Fr´echet modules, Amenable Fr´echet algebras, and approximate identities, Homology, Homotopy Appl., 11/1 (2009), 81 -- 114. [50] J. S. Pym, Remarks on the second duals of Banach algebras, J. Nigerian Math. Soc., 2 (1983), 31 -- 33. [51] S. Rahnama and A. Rejali, Amenability modulo an ideal of Fr´echet algebras, submitted. [52] S. Rahnama and A. Rejali, Locally bounded approximate diagonal modulo an ideal of Fr´echet algebras, submitted. [53] M. Ramezanpour, Weak amenability of the Lau product of Banach algebras defined by a Banach algebra morphism, Bull. Korean Math. Soc. 54 (2017), no. 6, 1991 -- 1999. [54] A. Ranjbari and A. Rejali, Fr´echet α-Lipschitz vector-valued operator algebra, U.P.B. Sci. Bull., Series A, to appear. [55] A. Ranjbari and A. Rejali, Ideal amenability of Fr´echet algebras, U.P.B. Sci. Bull., Series A, 79 (2017), no. 4, 51 -- 60. [56] A. Rejali, The analogue of weighted qroup algebra for semitopological semigroups, J. Sci, I.R. Iran, 6, no. 2 (1995), 113 -- 120. [57] A. Rejali, The Arens regularity of weighted semigroup algebras, Sci. Math. Japon, 60, no. 1 (2004), 129 -- 137. FR ´ECHET ALGEBRAS IN ABSTRACT HARMONIC ANALYSIS 17 [58] A. Rejali, H. R. E. Vishki, Regularity and amenability of the second dual of weighted group algebras, Proyecciones 26 (2007), 259 -- 267. [59] V. Runde, Lectures on Amenability, volume 1774 of Lecture Notes in Mathematics. Springer, Berlin, 2002. [60] H. Samea, Weak amenability of convolution Banach algebras on compact hypergroups, Bull. Korean Math. Soc. 47 (2010), no. 2, 307 -- 317. [61] H. H. Schaefer, Topological vector spaces, Third printing corrected Graduate text in math- ematics, Vol.3, Springer-Verlag, New York, 1971. [62] Yu. V. Selivanov, Biprojective Banach algebras, Izv. Akad. Nauk SSSR, Ser. Mat., 43 (1979), 1159 -- 1174. [63] Yu. V. Selivanov, Biprojective topological algebras (talk in Sheffield, 12 November 1996), Unpublished manuscript, 1996. [64] D. R. Sherbert, The structure of ideals and point derivations in Banach algebras of Lipschitz functions, Trans. Amer. Math. Soc., 111 (1964), 240 -- 272. [65] H. A. Smith, Tensor products of locally convex algebras, Proc. Am. Math. Soc. 17 (1966), 124 -- 132. [66] J. L. Taylor, Homology and cohomology for topological algebras, Adv. Math., 9 (1972), 137 -- 182. [67] J. Voigt, Factorization in Fr´echet algebras, J. London Math. Soc., 29/2 (1984), 147 -- 152. [68] J. Voigt, Factorization in some Fr´echet algebras of differentiable functions, Studia Math. 77 (1984), 333 -- 348. [69] A. Zivari-Kazempour, On the Arens product and approximate identity in locally convex algebras, Filomat, 30 (2016), no. 6, 1493 -- 1496. Z. Alimohammadi Department of Mathematics, University of Isfahan, Isfahan, Iran [email protected]; [email protected] A. Rejali Department of Mathematics, University of Isfahan, Isfahan, Iran [email protected]
1608.02047
4
1608
2016-12-21T11:17:00
Infinitesimal generators of invertible evolution families
[ "math.FA" ]
A logarithm representation of operators is introduced as well as a concept of pre-infinitesimal generator. Generators of invertible evolution families are represented by the logarithm representation, and a set of operators represented by the logarithm is shown to be associated with analytic semigroups. Consequently generally-unbounded infinitesimal generators of invertible evolution families are characterized by a convergent power series representation.
math.FA
math
INFINITESIMAL GENERATORS OF INVERTIBLE EVOLUTION FAMILIES YORITAKA IWATA Abstract. A logarithm representation of operators is introduced as well as a concept of pre-infinitesimal generator. Generators of invertible evolution fami- lies are represented by the logarithm representation, and a set of operators rep- resented by the logarithm is shown to be associated with analytic semigroups. Consequently generally-unbounded infinitesimal generators of invertible evo- lution families are characterized by a convergent power series representation. . A F h t a m [ 4 v 7 4 0 2 0 . 8 0 6 1 : v i X r a 1. Introduction The logarithm of an injective sectorial operator was introduced by Nollau [12] in 1969. After a long time, the logarithm of sectorial operators were studied again from 1990's [2, 13, 14], and its utility was established with respect to the definition of the logarithms of operators [4, 11] (for a review of sectorial operators, see Hasse [5]). While the sectorial operator has been a generic framework to define the logarithm of operators, the sectorial property is not generally satisfied by the evolution family of operators, where the evolution family correspond to the exponentials of operators in abstract Banach space framework (for the definition of evolution family in this paper, see Sec. 2.1). In this paper we characterize infinitesimal generators of invertible evolution fam- ilies that has not been settled so far. First of all, by introducing a kind of similarity transform, a logarithm representation is obtained for such generators. The loga- rithm representation is utilized to show a convergent power series representation of invertible evolution families generated by certain unbounded operators, although the validity of such a representation is not established for any evolution families generated by unbounded operators. In this context, the concept of pre-infinitesimal generator is introduced. 2. Mathematical settings 2.1. Evolution family on Banach spaces. Let X and B(X) be a Banach space with a norm k · k and a space of bounded linear operators on X. The same notation is used for the norm equipped with B(X), if there is no ambiguity. For a positive and finite T , let elements of evolution family {U (t, s)}−T ≤t,s≤T be mappings: (t, s) → U (t, s) satisfying the strong continuity for −T ≤ t, s ≤ T (for reviews or textbooks, see [1, 6, 10, 15, 17]). The semigroup properties: (1) U (t, r) U (r, s) = U (t, s), 2000 Mathematics Subject Classification. 47D03; 35A20; 34K30. Key words and phrases. invertible evolution family, operator theory, maximal regularity. The author is grateful to Prof. Emeritus Hiroki Tanabe for valuable comments. 1 2 and (2) YORITAKA IWATA U (s, s) = I, are assumed to be satisfied, where I denotes the identity operator of X. Both U (t, s) and U (s, t) are assumed to be well-defined to satisfy (3) U (s, t) U (t, s) = U (s, s) = I, where U (s, t) corresponds to the inverse operator of U (t, s). Since U (t, s) U (s, t) = U (t, t) = I is also true, the commutation between U (t, s) and U (s, t) follows. Oper- ator U (t, s) is a generalization of exponential function; indeed the properties shown in Eqs. (1)-(3) are satisfied by taking U (t, s) as et−s. Evolution family is an ab- stract concept of exponential function valid for both finite and infinite dimensional Banach spaces. Let Y be a dense Banach subspace of X, and the topology of Y be stronger than that of X. The space Y is assumed to be U (t, s)-invariant. Following the definition of C0-(semi)group (cf. the assumption H2 in Sec. 5.3 of Pazy [15] or corresponding discussion in Kato [7, 8]), U (t, s) is assumed to satisfy the boundedness [6, 15]; there exist real numbers M and β such that (4) kU (t, s)kB(X) ≤ M eβt, kU (t, s)kB(Y ) ≤ M eβt. Inequalities (4) are practically reduced to kU (t, s)kB(X) ≤ M eβT , kU (t, s)kB(Y ) ≤ M eβT , when t-interval is restricted to be finite [−T, T ]. 2.2. Pre-infinitesimal generator. The counterpart of the logarithm in the ab- stract framework is introduced. There are two concepts associated with the log- arithm of operators; one is the pre-infinitesimal generator and the other is t- differential of U (t, s). For −T ≤ t, s ≤ T , the weak limit (5) wlim h→0 h−1(U (t + h, s) − U (t, s)) u = wlim h→0 h−1(U (t + h, t) − I) U (t, s) u, is assumed to exist for u, which is an element of a dense subspace Y of X. A linear operator A(t) : Y → X is defined by (6) A(t)u := wlim h→0 h−1(U (t + h, t) − I)u for u ∈ Y and −T ≤ t, s ≤ T , and then let t-differential of U (t, s) in a weak sense be (7) ∂tU (t, s) u = A(t)U (t, s) u. Equation (7) is regarded as a differential equation satisfied by U (t, s)u that implies a relation between A(t) and the logarithm: A(t) = ∂tU (t, s) U (s, t). Let us call A(t) defined by Eq. (6) for a whole family {U (t, s)}−T ≤t,s≤T the pre- infinitesimal generator. Note that pre-infinitesimal generators are not necessarily infinitesimal generators; e.g., in t-independent cases, A defined by Eq. (6) is not necessarily a densely-defined and closed linear operator, while A must be a densely- defined and closed linear operator with its resolvent set included in {λ ∈ C : Reλ > β}. On the other hand, infinitesimal generators are necessarily pre-infinitesimal generators. In the following a set of pre-infinitesimal generators is denoted by G(X). INFINITESIMAL GENERATORS ... 3 2.3. A principal branch of logarithm. The logarithm is defined by the Dunford integral in this paper. Two difficulties of dealing with logarithm are its singularity at the origin and its multi-valued property. By introducing a constant κ ∈ C, the singularity can be handled. Let arg be a function of complex number, which gives the angle between the positive real axis to the line including the point to the origin. For the multi-valued property, a principle branch (denoted by "Log") of the logarithm (denoted by "log") is chosen for any complex number z ∈ C, a branch of logarithm is defined by Logz = log z + i arg Z, where Z is a complex number chosen to satisfy Z = z, −π < arg Z ≤ π, and arg Z = arg z + 2nπ for a certain integer n. Lemma 2.1. Let t and s satisfy 0 ≤ t, s ≤ T . For a given U (t, s) defined in Sec. 2, its logarithm is well defined; there exists a certain complex number κ satisfying (8) Log(U (t, s) + κI) = 1 where an integral path Γ, which excludes the origin, is a circle in the resolvent set of U (t, s) + κI. Here Γ is independent of t and s. Log(U (t, s) + κI) is bounded on X. 2πiRΓ Logλ (λ − U (t, s) − κ)−1dλ, Proof. The logarithm Log holds the singularity at the origin, so that it is necessary to show a possibility of taking a simple closed curve (integral path) excluding the origin in order to define the logarithm by means of the Dunford-Riesz integral. It is not generally possible to take such a path in case of κ = 0. First, U (t, s) is assumed to be bounded for 0 ≤ t, s ≤ T (Eq. (4)), and the spectral set of U (t, s) is a bounded set in C. Second, for κ satisfying (9) κ > M eβT , the spectral set of U (t, s) + κI is separated with the origin. Consequently it is possible to take an integral path Γ including the spectral set of U (t, s) + κI and excluding the origin. Equation (8) follows from the Dunford-Riesz integral [3]. Furthermore, by adjusting the amplitude of κ, an appropriate integral path always exists independent of t and s. Log(U (t, s) + κI) is bounded on X, since Γ is included in the resolvent set of (cid:3) (U (t, s) + κI). According to this lemma, by introducing nonzero κ, the logarithm of U (t, s) + κI is well-defined without assuming the sectorial property to U (t, s). On the other hand Eq. (8) is valid with κ = 0 only for limited cases. 3. Main results 3.1. Logarithm representation of pre-infinitesimal generator. Theorem 3.1. Let t and s satisfy −T ≤ t, s ≤ T , and Y be a dense subspace of X. For U (t, s) defined in Sec. 2, let A(t) ∈ G(X) and ∂tU (t, s) be determined by Eqs. (6) and (7) respectively. If A(t) and U (t, s) commute, an evolution family {A(t)}−T ≤t≤T is represented by means of the logarithm function; there exists a certain complex number κ 6= 0 such that (10) A(t) u = (I + κU (s, t)) ∂tLog (U (t, s) + κI) u, 4 YORITAKA IWATA where u is an element in Y . Note that U (t, s) defined in Sec. 2 is assumed to be invertible. Proof. For U (t, s) defined in Sec. 2, operators Log (U (t, s) + κI) and Log (U (t + h, s) + κI) are well defined for a certain κ (Lemma 2.1). The t-differential in a weak sense is formally written by 1 h {Log (U (t + h, s) + κI) − Log (U (t, s) + κI)} (11) {(λ − U (t + h, s) − κ)−1 − (λ − U (t, s) − κ)−1}dλ wlim h→0 = wlim h→0 = wlim h→0 1 1 h 2πiRΓ Logλ 2πiRΓ Logλ 1 {(λ − U (t + h, s) − κ)−1 U(t+h,s)−U(t,s) (λ − U (t, s) − κ)−1}dλ h where Γ, which is possible to be taken independent of t, s and h for a sufficiently large certain κ, denotes a circle in the resolvent set of both U (t, s) + κI and U (t + h, s) + κI. A part of the integrand of Eq. (11) is estimated as k{(λ − U (t + h, s) − κ)−1 U(t+h,s)−U(t,s) h (λ − U (t, s) − κ)−1}vkX (12) ≤ k{(λ − U (t + h, s) − κ)−1kB(X) k U(t+h,s)−U(t,s) (λ − U (t, s) − κ)−1}vkX , h for v ∈ X. There are two steps to prove the validity of Eq. (11). In the first step, the former part of the right hand side of Eq. (12) satisfies k{(λ − U (t + h, s) − κ)−1kB(X) < ∞, since λ is taken from the resolvent set of U (t + h, s) − κI. In the same way the operator (λ − U (t, s) − κ)−1 is bounded on X and Y . Then the continuity of the mapping t → (λ − U (t, s) − κ)−1 as for the strong topology follows: k(λ − U (t + h, s) − κ)−1 − (λ − U (t, s) − κI)−1kB(X) ≤ k(λ − U (t + h, s) − κ)−1kB(X)k(U (t + h, s) − U (t, s))(λ − U (t, s) − κ)−1kB(X). In the second step, the latter part of the right hand side of Eq. (12) is estimated as (13) 1 h U(t+h,s)−U(t,s) (λ − U (t, s) − κ)−1u(cid:13)(cid:13)(cid:13)X (cid:13)(cid:13)(cid:13) t A(τ )U (τ, s)(λ − U (t, s) − κ)−1u dτ(cid:13)(cid:13)(cid:13)X =(cid:13)(cid:13)(cid:13) hR t+h hR t+h ≤ 1 t kA(τ )U (τ, s)kB(Y,X)k(λ − U (t, s) − κ)−1kB(Y )kukY dτ for u ∈ Y . Because kA(τ )U (τ, s)kB(Y,X) < ∞ is true by assumption, the right hand side of Eq. (13) is finite. Equation (13) shows the uniform boundedness with respect to h, then the uniform convergence (h → 0) of Eq. (11) follows. Consequently the weak limit process h → 0 for the integrand of Eq. (11) is justified, as well as the commutation between the limit and the integral. According to Eq. (11), interchange of the limit with the integral leads to ∂tLog(U (t, s) + κI) u = 1 2πiRΓ dλ (cid:20)(Logλ)(λ − U (t, s) − κ)−1 wlim h→0 (cid:16) U(t+h,s)−U(t,s) h (cid:17) (λ − U (t, s) − κ)−1(cid:21) u INFINITESIMAL GENERATORS ... 5 for u ∈ Y . Because we are also allowed to interchange A(t) with U (t, s), ∂tLog(U (t, s) + κI) u = 1 = 1 (14) 2πiRΓ(Logλ)(λ − U (t, s) − κ)−1A(t) U (t, s) (λ − U (t, s) − κ)−1dλ u 2πiRΓ(Logλ) (λ − U (t, s) − κ)−2 U (t, s) dλ A(t) u for u ∈ Y . A part of the right hand side is calculated as 1 κ λ (λ − U (t, s) − κ)−1 dλ − 1 λ (λ − U (t, s) − κ)−1 dλ 2πiRΓ κ 1 λ dλ 1 λ (λ − U (t, s) − κ)−1 U (t, s) dλ 1 λ (λ − U (t, s) − κ)−1 {λ − κ − (λ − U (t, s) − κ)} dλ = 1 = 1 = 1 = 1 = 1 2πiRΓ 2πiRΓ 2πiRΓ (Logλ) (λ − U (t, s) − κ)−2U (t, s) dλ 2πiRΓ 2πiRΓ 2πiRΓ(λ − U (t, s) − κ)−1 dλ − 1 2πiRΓ(λ − U (t, s) − κ)−1 dλ − 1 2πiRΓ(λ − U (t, s) − κ)−1 dλ −κ(U (t, s) + κI)−1(cid:8) 1 2πiRΓ 2πiRΓ(λ − U (t, s) − κ)−1 dλ 2πiRΓ(λ − U (t, s) − κ)−1dλ − 1 −κ(U (t, s) + κI)−1(cid:8) 1 2πiRΓ(λ − U (t, s) − κ)−1dλ 2πiRν=rP∞ = (I − κ(U (t, s) + κI)−1) = (I − κ(U (t, s) + κI)−1) = I − κ(U (t, s) + κI)−1, λ (U (t, s) + κI)(λ − U (t, s) − κ)−1dλ(cid:9) λ dλ(cid:9) 2πiRΓ U(t,s)n ν n+1 dν = 1 n=1 1 1 1 1 due to the integration by parts, where λ − κ = ν = r is a properly chosen circle large enough to include Γ. (2πi)−1RΓ λ−1dλ = 0 is seen by applying dLogλ/dλ = n=1 U (t, s)nν−n−1 dν = I follows from the singularity of 1/λ. ν−n−1. (2πi)−1Rν=rP∞ Consequently we have A(t) u = {I − κ(U (t, s) + κI)−1}−1 ∂tLog (U (t, s) + κI) u = (U (t, s) + κI)U (t, s)−1 ∂tLog (U (t, s) + κI) u = (I + κU (s, t)) ∂tLog (U (t, s) + κI) u for u ∈ Y . (cid:3) What is introduced by Eq. (10) is a kind of resolvent approximation of A(t) ∂tLog (U (t, s) + κI) = (I + κU (s, t))−1A(t), in which A(t) is approximated by the resolvent of U (s, t). As seen in the following it is notable that there is no need to take κ → 0. This point is different from the usual treatment of resolvent approximations. On the other hand, it is also seen by Eq. (10) that ∂tLog (U (t, s) + κI) = (U (t, s) + κI)κ=0A(t)(U (t, s) + κI)−1 shows a structure of similarity transform, where (U (t, s) + κ)κ=0 means U (t, s) + κ satisfying a condition κ = 0. Under the validity of Theorem 3.1, for −T ≤ t, s ≤ T , let a(t, s) be defined by a(t, s) := Log(U (t, s) + κI), 6 YORITAKA IWATA then Eq. (10) is written as A(t) = (I + κU (s, t)) ∂ta(t, s). Since κ is chosen to separate the spectral set of U (t, s) + κI from the origin, the inverse operator of (I + κU (s, t)) = U (s, t)(U (t, s) + κI) is well defined as (I + κU (s, t))−1 = (U (t, s) + κI)−1U (t, s). It also ensures that ∂ta(t, s) is well defined. Corollary 3.2. Let t and s satisfy 0 ≤ t, s ≤ T . For U (t, s) and A(t) satisfying the assumption of Theorem 3.1, the exponential of a(t, s) is represented by a convergent power series: (15) ea(t,s) =P∞ n=0 a(t,s)n n! , with a relation ea(t,s) = exp(Log(U (t, s) + κI)) = U (t, s) + κI. different t and s are further assumed to commute, If a(t, s) with (16) ∂tea(t,s)us = ∂ta(t, s) ea(t,s)us is satisfied for us ∈ Y , where ∂t denotes a t−differential in a weak sense. Proof. Since a(t, s) is a bounded operator on X (Lemma 2.1), the exponential of a(t, s) is represented by a convergent power series [9]: (17) where ea(t,s) =P∞ n=0 a(t,s)n n! , ea(t,s) = exp(Log(U (t, s) + κI)) = 1 2πiRΓ eLogλ (λ − U (t, s) − κ)−1dλ = U (t, s) + κI is satisfied. Since a(t, s) with different t and s commute, ∂t{a(t, s)}n = n{a(t, s)}n−1 ∂ta(t, s) leads to (18) ∂tea(t,s)us = ea(t,s) (∂ta(t, s))us for us ∈ Y . This is a linear evolution equation satisfied by ea(t,s)us. (cid:3) Further calculi on Eq. (18) lead to ∂tea(t,s)us = ∂t(U (t, s) + κI)us = ∂tU (t, s)us, and ea(t,s) (∂ta(t, s))us = (U (t, s) + κI)∂t(Log(U (t, s) + κI))us = (U (t, s) + κI)(U (t, s) + κI)−1U (t, s)A(t)us = U (t, s)A(t)us = A(t)U (t, s)us, where Theorem 3.1 is applied. As a result ∂tU (t, s)us = A(t)U (t, s)us is obtained. Note that ea(t,s) does not satisfy the semigroup property, while U (t, s) satisfies it. INFINITESIMAL GENERATORS ... 7 4. Abstract Cauchy problem 4.1. Autonomous case. Logarithmic representation is utilized to solve autonomous Cauchy problem (19) ( ∂tu(t) = A(t)u(t) u(s) = us, in X, where A(t) ∈ G(X) : Y → X is assumed to be an infinitesimal generator of U (t, s), −T ≤ t, s ≤ T is satisfied, Y is a dense subspace of X permitting the representation shown in Eq. (10), and us is an element of X. As seen in Eq. (16), under the assumption of commutation, a related Cauchy problem is obtained as (20) ( ∂tv(t, s) = (∂ta(t, s)) v(t, s) v(s, s) = ea(s,s)us, in X, where ∂ta(t, s) = ∂tLog(U (t, s) + κI) is well-defined. It is possible to solve re-written Cauchy problem, and the solution is represented by for us ∈ X (cf. Eq. (17)). v(t, s) = ea(t,s)us =P∞ n=0 a(t,s)n n! us Theorem 4.1. Operator ea(t,s) is holomorphic. Proof. According to the boundedness of a(t, s) on X (Lemma 2.1), ∂n possible to be represented as t ea(t,s) [18] is (21) for a certain κ, where λneλ does not hold any singularity for any finite λ. Following the standard theory of evolution equation, t ea(t,s) = 1 ∂n 2πiRΓ λneλ(λ − a(t, s))−1 dλ, k∂n t ea(t,s)k ≤ Cθ,n π(t sin θ)n is true for a certain constant Cθ,n (n = 0, 1, 2, · · · ), where θ ∈ (0π/2) and arg t < π/2 are satisfied (for the detail, e.g., see [17]). It follows that (22) lim t→+0 sup tnk∂n t ea(t,s)k ≤ lim t→+0 sup tn Cθ,n π(t sin θ)n < ∞. Consequently, for z − t < t sin θ, the power series expansion is uniformly convergent in a wider sense. Therefore ea(t,s) is holomorphic. (cid:3) (z−t)n n! ∂n t ea(t,s) n=0 P∞ Theorem 4.2. For us ∈ X there exists a unique solution u(t) ∈ C([−T, T ]; X) of (19) with a convergent power series representation: (23) u(t) = U (t, s)us = (ea(t,s) − κI)us =(cid:16)P∞ where κ is a certain complex number. n=0 a(t,s)n n! − κI(cid:17) us, Proof. The unique existence follows from the assumption for A(t). ea(t,s) is holo- morphic function (Theorem 4.1) with the convergent power series representation (Eq. (17)). The solution of the original Cauchy problem is obtained as a(t,s)n u(t) = (I + κU (s, t))−1v(t, s) = (I + κU (s, t))−1P∞ n=0 n! us 8 YORITAKA IWATA for the initial value us ∈ X. Note that A(t) is not assumed to be a generator of analytic evolution family, but only a generator of invertible evolution family. (cid:3) For Iλ denoting the resolvent operator of A(t), the evolution operator defined by the Hille-Yosida approximation is written by u(t) = lim λ→0 s IλA(τ ) dτ )us, exp(R t so that more informative representation is provided by Theorem 4.2 compared to the standard theory based on the Hille-Yosida theorem. 4.2. Non-autonomous case. Series representation in autonomous part leads to the enhancement of the solvability. Let Y be a dense subspace of X permitting the representation shown in Eq. (10), and us is an element of X. Let us consider non-autonomous Cauchy problem (24) ( ∂tu(t) = A(t)u(t) + f (t) u(s) = us in X, where A(t) ∈ G(X) : Y → X is assumed to be the infinitesimal generator U (t, s), f ∈ L1(−T, T ; X) is locally Holder continuous on [−T, T ] kf (t) − f (s)k ≤ CH t − sγ for a certain positive constant CH , γ ≤ 1 and −T ≤ t, s ≤ T . The solution of non-autonomous problem does not necessarily exist in such a setting (in general, f ∈ C([−T, T ]; X) is necessary). Theorem 4.3. Let f ∈ L1(−T, T ; X) be locally Holder continuous on [−T, T ]. For us ∈ X there exists a unique solution u(t) ∈ C([−T, T ]; X) for (24) such that u(t) = [P∞ n=0 a(t,s)n n! − κI]us +R t s [P∞ n=0 using a certain complex number κ. a(t,τ )n n! − κI]f (τ )dτ Proof. Let us start with cases with f ∈ C([−T, T ]; X). The unique existence follows from the standard theory of evolution equation. The representation follows from that of U (t, s) and the Duhamel's principle (25) u(t, s) = U (t, s)us +R t = (ea(t,s) − κI)us +R t s U (t, τ )f (τ )dτ s [ea(t,τ ) − κI]f (τ )dτ, where the convergent power series representation of ea(t,s) is valid (cf. Eq. (17)). Next let us consider cases with the locally Holder continuous f (t). According to the linearity of Eq. (24), it is sufficient to consider the inhomogeneous term. For ǫ satisfying 0 < ǫ << T , s R t+ǫ [ea(t,τ ) − κI]f (τ )dτ → R t s [ea(t,τ ) − κI]f (τ )dτ INFINITESIMAL GENERATORS ... 9 is true by taking ǫ → 0. On the other hand, (26) s A(t)U (t, τ )f (τ )dτ s A(t)U (t, τ )f (t)dτ ∂τ U (t, τ )f (t)dτ s s [ea(t,τ ) − κI]f (τ )dτ =R t+ǫ A(t)R t+ǫ s A(t)U (t, τ )(f (τ ) − f (t))dτ +R t+ǫ =R t+ǫ =R t+ǫ s A(t)U (t, τ )(f (τ ) − f (t))dτ −R t+ǫ =R t+ǫ =R t+ǫ −U (t, t + ǫ)f (t) + U (t, s)f (t), s s A(t)U (t, τ )(f (τ ) − f (t))dτ − U (t, t + ǫ)f (t) + U (t, s)f (t) (1 + κU (s, t))∂ta(t, s)[ea(t,τ ) − κI](f (τ ) − f (t))dτ where ∂τ U (t, τ ) = −A(τ )U (t, τ ) is utilized. The Holder continuity and Eq. (22) lead to the strong convergence of the right hand of Eq. (26): [ea(t,τ ) − κI]f (τ )dτ s (1 + κU (s, t))(∂ta(t, s))[ea(t,τ ) − κI](f (τ ) − f (t))dτ + (U (t, s) − I)f (t) (due to ǫ → 0) for f ∈ L1(0, T ; X). A(t) is assumed to be an infinitesimal generator, so that A(t) is a closed operator from Y to X. It follows that s A(t)R t+ǫ → R t A(t)R t =R t and s [ea(t,τ ) − κI]f (τ )dτ ∈ Y R t s [ea(t,τ ) − κI]f (τ )dτ s (1 + κU (s, t))(∂ta(t, s))[ea(t,τ ) − κI](f (τ ) − f (t))dτ + (U (t, s) − I)f (t) ∈ X. The right hand side of this equation is strongly continuous on [−T, T ]. As a result, (∂ta(t, τ ))ea(t,τ )f (τ )dτ s (1 + κU (τ, t))−1A(t)(U (t, τ ) + κI)f (τ )dτ s s [ea(t,τ ) − κI]f (τ )dτ ∂tR t+ǫ = [ea(t,t+ǫ) − κI]f (t + ǫ) +R t+ǫ → f (t) +R t = f (t) +R t = f (t) + A(t)R t s A(t)U (t, τ )f (τ )dτ s [ea(t,τ ) − κI]f (τ )dτ. s [ea(t,τ ) − κI]f (τ )dτ satisfies Eq. (24), and that it is sufficient to assume f ∈ L1(0, T ; X) as Holder continuous. (cid:3) We see that R t This result should be compared to the standard theory of evolution equations in which the inhomogeneous term f is assumed to be continuous on [−T, T ]. If the inhomogeneous term f ∈ Lp([0, T ]; X) is further assumed to be satisfied for 1 < p < ∞, and Y = D(A(t)) = D(A(0)) and A(·) ∈ C([0, T ], L(Y, X)), Eq. (20) with such an inhomogeneous term corresponds to the equation exhibiting the maximal regularity of type Lp [16]. 5. Concluding remark As for the applicability of the theory, the conditions to obtain the logarithmic representation (conditions shown in Sec. 2) are not so restrictive; indeed, they can be satisfied by C0-groups generated by t-independent infinitesimal generators. The most restrictive condition to obtain Eq. (10) is the commutation between A(t) and U (t, s). Such a commutation is trivially satisfied by t-independent A(t) = A, and 10 YORITAKA IWATA also satisfied when the variable t is separable (i.e., for an integrable function g(t), A(t) = g(t)A). In this sense the operators specified in Theorem 3.1 correspond to a moderate generalization of t-independent infinitesimal generators. References [1] W. Arendt, Semigroups and evolution equations: functional calculus, regularity and kernel estimates. Handbook of Differential Equations 1 (2002) 185. [2] K. N. Boyadzhiev, Logarithms and imaginary powers of operators on Hilbert spaces. Collect. Math. 45 3 (1994) 287-300. [3] N. Dunford, Spectral theory I, convergence to projections. Trans. Amer. Math. Soc. 54 1943 185-217. [4] M. Hasse, Spectral properties of operator logarithms. Math. Z. 245 4 (2003) 761-779. [5] M. Hasse, The functional calculus for sectorial operators. Birkhauser, 2006. [6] T. Kato, Perturbation Theory for Linear Operators. Springer-Verlag, 1966. [7] T. Kato, Linear evolution equation of "hyperbolic" type. J. Fac. Sci. Univ. Tokyo 17 (1970) 241-258. [8] T. Kato, Linear evolution equation of "hyperbolic" type II. J. Math. Soc. Japan 25 4 (1973) 648-666. [9] T. Kato, A short introduction to perturbation theory for linear operators. Springer-Verlag; 1982. [10] S. G. Krein, Linear differential equations in Banach space (translated from Russian). Transl. Math. Monogr. 29, Amer. Math. Soc., 1971. [11] C. Martinez and M. Sanz, The theory of fractional powers of operators. North-Holland, 2001. [12] V. Nollau, Uber den Logarithmus abgeschlossener Operatoren in Banachschen Raumen, Acta Sci. Math. 300 (1969) 161-174. [13] N. Okazawa, Logarithms and imaginary powers of closed linear operators. Integral Equations and Operator Theory 38 4 (2000) 458-500. [14] N. Okazawa, Logarithmic characterization of bounded imaginary powers. Progress in Nonlin- ear Differential Equations and Their Applications 42 (2000) 229-237. [15] A. Pazy, Semigroups of linear operators and application to partial differential equations. Springer-Verlag, 1983. [16] J. Pruss and R. J. Schnaubelt, Math. Anal. Appl. 256 (2001) 405-430. [17] H. Tanabe, Equations of evolution. Pitman, 1979. [18] A. E. Taylor, Spectral theory of closed distributive operators. Acta Math. 84 (1951) 189-224. (Yoritaka Iwata) Institute of Innovative Research, Tokyo Institute of Technology; Department of Mathematics, Shibaura Institute of Technology. E-mail address: iwata [email protected]
1603.04897
1
1603
2016-03-15T21:55:51
Locally piecewise affine functions and their order structure
[ "math.FA" ]
Piecewise affine functions on subsets of $\mathbb R^m$ were studied in \cite{Ovchinnikov:02,Aliprantis:06a,Aliprantis:07a,Aliprantis:07}. In this paper we study a more general concept of a locally piecewise affine function. We characterize locally piecewise affine functions in terms of components and regions. We prove that a positive function is locally piecewise affine iff it is the supremum of a locally finite sequence of piecewise affine functions. We prove that locally piecewise affine functions are uniformly dense in $C(\mathbb R^m)$, while piecewise affine functions are sequentially order dense in $C(\mathbb R^m)$. This paper is partially based on \cite{Adeeb:14}.
math.FA
math
LOCALLY PIECEWISE AFFINE FUNCTIONS AND THEIR ORDER STRUCTURE S. ADEEB AND V. G. TROITSKY Abstract. Piecewise affine functions on subsets of Rm were stud- ied in [11, 7, 8, 9]. In this paper we study a more general concept of a locally piecewise affine function. We characterize locally piece- wise affine functions in terms of components and regions. We prove that a positive function is locally piecewise affine iff it is the supre- mum of a locally finite sequence of piecewise affine functions. We prove that locally piecewise affine functions are uniformly dense in C(Rm), while piecewise affine functions are sequentially order dense in C(Rm). This paper is partially based on [3]. 1. Piecewise affine functions We start with a brief overview of affine and piecewise affine functions; we refer the reader to [11, 7] and [9, Chapter 7] for details. Fix m ∈ N. By an affine function we mean a function f : Rm → R of form f (x) = hv, xi+b for some v ∈ Rm and b ∈ R. Its kernel {f = 0} is an affine hyperplane in Rm. It is easy to see that any two affine functions which agree on a non-empty open set are equal. We write A for the vector space of all affine functions on Rm. Clearly, A is a subspace of C(Rm), the space of all real-valued continuous functions on Rm. Let Ω be a convex closed subset of Rm with non-empty interior. Fol- lowing [9], we call such a set a solid domain. A continuous function f : Ω → R is called piecewise affine if it agrees with finitely many Date: December 14, 2017. 2010 Mathematics Subject Classification. Primary: 46A40. Secondary: 46E05. Key words and phrases. affine function, piecewise affine function, locally piece- wise affine function, vector lattice, sublattice. The authors were supported by NSERC grants. 1 2 S. ADEEB AND V. G. TROITSKY affine functions. That is, there exist f1 . . . , fn ∈ A such that for ev- ery x ∈ Ω there exists i ≤ n with f (x) = fi(x). Equivalently (see Definition 7.10 and Theorem 7.13 in [9]), a function f : Ω → R is piece- wise affine iff there exist distinct affine functions f1, . . . , fn and subsets S1, . . . , Sn of Ω such that (1) on Si, f agrees with fi; (2) Si = Int Si 6= ∅ for every i; (3) Int Si ∩ Int Sj = ∅ whenever i 6= j; (4) Sn i=1 Si = Ω. In particular, it follows from (2) that the set {f = fi} has non-empty in- i=1, and (cid:8)(Si, fi)(cid:9)n terior for every i = 1, . . . , n. The sets {fi}n i=1 are called the components, regions, and characteristic pairs of f , respectively. The characteristic pairs of f are unique up to re-ordering. We write PA(Ω) for the space of all piecewise affine functions on Ω. We put PA = PA(Rm). i=1, {Si}n Suppose f ∈ PA(Ω) with characteristic pairs (cid:8)(Si, fi)(cid:9)n i=1. For each pair (i, j) with 1 ≤ i, j ≤ n and i 6= j, the set (cid:8)x ∈ Rm : fi(x) = fj(x)(cid:9), or {fi = fj} is a hyperplane in Rm. Let C be the union of all such hyperplanes. Then C is a closed set with empty interior. It is easy to see that the set (Int Ω)\C is open, dense in Ω, and naturally partitioned into a finite number of convex open subsets. We call these subsets the cells of f . Let K be a cell of f . Since K is disjoint from C, no two components agree at any point of K. It follows that there exists a component fi such that f agrees with fi on K, so that K ⊆ Si. It also follows that for every j 6= i we either have fj(x) < f (x) for all x ∈ K or fj(x) > f (x) for all x ∈ K. Example 1.1. Let m = 1 and f (t) = min(cid:8)t, 1(cid:9). Then f is piecewise affine on R, with components f1(t) = t, f2(t) = −t, and f3(t) = 1, and with regions S1 = [0, 1], S2 = [−1, 1], and S3 = (−∞, −1] ∪ [1, +∞). Note that S3 is not connected. The cells of f are K1 = (−∞, −1), K2 = (−1, 0), K3 = (0, 1), and K4 = (1, +∞). Recall that under pointwise order, the space C(Ω) is a vector lattice. It is easy to see that PA(Ω) is a (vector) sublattice of C(Ω). LOCALLY PIECEWISE AFFINE FUNCTIONS 3 Let A be a subset of a vector lattice V . Recall that A∨ = (cid:8)x1 ∨ · · · ∨ xn : n ∈ N, x1, . . . , xn ∈ A(cid:9) and A∧ = (cid:8)x1 ∧ · · · ∧ xn : n ∈ N, x1, . . . , xn ∈ A(cid:9). If A is a linear subspace of V then A∨∧ = A∧∨ and this set is exactly the sublattice generated by A in V ; see, e.g., [9, Lemma 1.21]. Since A ⊆ PA and PA is a sublattice of C(Rm), it follows that A∧∨ ⊆ PA. In fact, [9, Theorem 7.23] asserts that PA = A∧∨, hence PA is exactly the sublattice of C(Rm) generated by A. Moreover, if Ω is a solid domain in Rm and f ∈ PA(Ω) with component set F = {fi}n i=1 then f ∈ F ∧∨ or, more precisely, f agrees with the restriction to Ω of a function in F ∧∨. Since every function in F ∧∨ is defined on all of Rm, it follows that every function in PA(Ω) extends to a function in PA. We write Bm ∞ for the closed unit ball of ℓm ∞: xi ≤ 1(cid:9). ∞ = (cid:8)x ∈ Rm : max Bm 1≤i≤m For n ∈ N, we write Ωn for the ball of radius n centred at zero in ℓm ∞; i.e., Ωn = nBm ∞. 2. Locally piecewise affine functions A function f : Rm → R is said to be locally piecewise affine if its restriction to any bounded solid domain Ω is in PA(Ω). The following lemma is straightforward. Lemma 2.1. Let f : Rm → R. Then f is locally piecewise affine iff the restriction of f to Ωn is in PA(Ωn) for every n ∈ N. We write LPA for the collection of all locally piecewise affine func- tions on Rm. Clearly, LPA is a sublattice of C(Rm) and A ( PA ( LPA. Let F ⊂ C(Rm). We say that F is locally finite if for every compact set C all but finitely many functions in F vanish on C. That is, the set (cid:8)f ∈ F : ∃x ∈ C f (x) 6= 0(cid:9) is finite. Just as in Lemma 2.1, it is easy to see that F is locally finite iff for every n all but finitely many functions in F vanish on Ωn. Clearly, any subset of a locally finite set 4 S. ADEEB AND V. G. TROITSKY of functions is again locally finite. Additionally, if F ⊂ C(Rm) is a locally finite set of functions, then F ∨, F ∧ and consequently F ∨∧ are all locally finite. If F is a sequence, then F is termed a locally finite sequence of functions. It is a standard fact that C(Rm) is not order complete. Yet we have the following. Lemma 2.2. Every locally finite subset of PA has a supremum and an infimum in C(Rm), which belong to LPA and agree with the pointwise maximum and minimum, respectively. Proof. Let F ⊆ PA be a locally finite set. For every point x ∈ Rm, all but finitely many members of F vanish at x. Hence, h(x) = max(cid:8)f (x) : f ∈ F(cid:9) is defined. Similarly, on every open ball B in Rm, h is the point- wise maximum of a finite subcollection of F , so that h is continuous on B; it follows that h ∈ C(Rm). It follows from the definition that h = sup F in C(Rm). It also follows that on every ball, h agrees with a finite subset of F , hence h is locally piecewise affine. The proof for the infimum is similar. (cid:3) We write 1 for the constant one function on Rm. Lemma 2.3. For every ε > 0 there exists f ∈ PA such that 0 ≤ f ≤ 1, f equals 1 on Bm ∞, and vanishes outside of (1 + ε)Bm ∞. Proof. Put g(x) = kxk∞ = maxi≤mxi for x ∈ Rm. Put h : R+ → R+ so that h is continuous, equals 1 on [0, 1], vanishes on [1 + ε, +∞) and is affine on [1, 1 + ε]. Let f = h ◦ g. Being the composition of two piecewise affine functions, f is itself piecewise affine. It is easy to verify that f satisfies the rest of the requirements. (cid:3) Theorem 2.4. For every 0 ≤ f ∈ LPA there exists a locally finite sequence (fi) in PA+ such that f = sup fi. Proof. Let (cn) be an enumeration of all integer points in Rm. Consider the "boxes" Kn = cn + Bm ∞ and eKn = cn + 2Bm ∞. Fix n. Since f is continuous, it is bounded on Kn. Let M > 0 be such that f (x) < M for all x ∈ Kn. As in Lemma 2.3, we can find a piecewise affine function gn such that 0 ≤ gn ≤ M 1, gn equals M on LOCALLY PIECEWISE AFFINE FUNCTIONS 5 Kn, and gn vanishes outside eKn. Put fn = f ∧ gn. We will show that the sequence (fn) satisfies all the requirements. By construction, 0 ≤ fn ≤ f , fn agrees with f on Kn and vanishes outside eKn. Observe that fn is piecewise affine because on eKn it agrees with f ∧ gn, which is piecewise affine on eKn and fn vanishes outside of eKn. The sequence (fn) is locally finite because for every n the only members of the sequence that do not vanish on Kn are fn itself and the members corresponding to the (finitely many) boxes that meet eKn. Finally, f = sup fn because 0 ≤ fn ≤ f and fn agrees with f on Kn for every n. (cid:3) 3. Characteristic pairs of locally piecewise functions Recall that every piecewise affine function is determined by its com- ponents and regions. The following theorem provides a similar charac- terization for locally piecewise affine functions. Theorem 3.1. A function f : Rm → R is locally piecewise affine iff there exist a sequence of distinct affine functions (fi) and a sequence (Si) of subsets of Rm such that (1) on Si, f agrees with fi; (2) Si = Int Si 6= ∅ for every i; (3) Int Si ∩ Int Sj = ∅ whenever i 6= j; (4) S∞ (5) every bounded set meets only finitely many Si's. i=1 Si = Rm; Proof. Suppose that f satisfies the five conditions; show that it is lo- cally piecewise affine. First, observe that f is continuous. Indeed, sup- pose that xn → x in Rm but f (xn) 6→ f (x). Passing to a subsequence, we can find an ε > 0 such that (cid:12)(cid:12)f (xn) − f (x)(cid:12)(cid:12) > ε and xn ∈ B(x, 1) for every n. By (5), the ball B(x, 1) meets only finitely many Si's. Hence, passing to a further subsequence, we may assume that the sequence (xn) is contained in Si for some i. Since Si is closed, we have x ∈ Si. It follows that f (xn) = fi(xn) → fi(x) = f (x); a contradiction. Thus, f is continuous. By (5) again, every bounded solid domain Ω is contained in the union of finitely many Si's; it follows that the restriction of f to Ω is in PA(Ω), so that f is locally piecewise affine. 6 S. ADEEB AND V. G. TROITSKY Suppose now that f is locally piecewise affine. Fix n ∈ N. The restriction of f to Ωn is in PA(Ωn); consider the affine components of this restriction. By collecting the affine functions associated with Ωn for all n ∈ N, we produce a sequence (fi) in A such that for every n there exists kn such that f agrees with f1, . . . , fkn on Ωn. By dropping repeated terms, we may assume without loss of generality that all the members in this sequence are distinct. For every i, put Ui = Int{f = fi} and Si = Ui. Since fi is a component of the restriction of f to Ωn for some n, it follows that Ui 6= ∅. It is easy to see that (1) and (2) are satisfied. To show (3), suppose that Int Si meets Int Sj for some i and j. This follows that fi and fj agree on an open set, hence fi = fj and, therefore, i = j. Fix x ∈ Rm and take n ∈ N such that x ∈ Int(Ωn). Let K1, . . . , Kl be the cells generated in Ωn by f1, . . . , fkn. Since Sl j=1 Kp is dense in Ωn, we have x ∈ Kp for some p ≤ l. Note that Kp is open and f agrees with some fi on Kp. It follows that Kp ⊆ Ui and, therefore, x ∈ Si. This yields (4). Finally, we will show that Int(Ωn) only meets S1, . . . , Skn; this will immediately yield (5). Suppose that Int(Ωn) meets Si for some i ∈ N. It follows that Int(Ωn) meets Ui. This implies that Ui meets Kr for some r ≤ l. Note that f agrees with fj on Kr for some j ≤ kn. It follows that fj and fi agree on the non-empty open set Kr ∩ Ui, hence j = i and, therefore, i ≤ kn. (cid:3) As before, the sets {fi}∞ i=1, and (cid:8)(Si, fi)(cid:9) are called the components, regions, and characteristic pairs of f , respectively. i=1, {Si}∞ 4. LPA is uniformly dense in C(Rm) It is of interest whether every continuous function on Rm can be approximated in some sense by piecewise affine functions. We will consider two kinds of denseness here: uniform denseness and order denseness. We start with uniform denseness. It is easy to see that PA is not uniformly dense in C(Rm) even for m = 1. LOCALLY PIECEWISE AFFINE FUNCTIONS 7 Theorem 4.1. LPA is uniformly dense in C(Rm). That is, for every f ∈ C(Rm) and every ε > 0 there exists h ∈ LPA such that f − h ≤ ε1. Proof. The proof is analogous to that of Theorem 2.4. Let f ∈ C(Rm) and ε > 0. First, we consider the case when f ≥ 0. Let (cn) be an enumeration ∞ and of all integer points in Rm. Consider the "boxes" Kn = cn + Bm eKn = cn + 2Bm ∞. j (x) 6= e∗ j (y), where e∗ Fix n. Observe that PA( eKn) is a sublattice of C( eKn). It clearly contains constant functions. It is easy to see that it separates points. Indeed, let x, y ∈ eKn with x 6= y. Then xj 6= yj for some j ≤ m. That is, e∗ j is the j-th coordinate functional. Clearly, j is an affine function. Since eKn is compact, it now follows from the e∗ lattice version of Stone-Weierstrass Theorem that PA( eKn) is uniformly dense in C( eKn), see, e.g., Theorem 11.3 on page 88 of [4]. It follows that there is hn ∈ PA( eKn) such that (cid:12)(cid:12)hn(x) − f (x)(cid:12)(cid:12) < ε for all x ∈ eKn. Extend hn to a piecewise affine function hn in PA. Since hn is continuous on eKn, it is bounded on eKn. Let Mn > 0 be such that hn(x) < Mn for all x ∈ eKn. As in Lemma 2.3, we can find gn ∈ PA such that 0 ≤ gn ≤ Mn1, gn equals Mn on Kn, and gn vanishes outside eKn. Consider the sequence of functions (cid:0)hn ∧ gn(cid:1). Since gn vanishes outside of eKn, all but finitely many terms of this sequence vanish on Kn; it follows that the sequence (cid:0)hn ∧ gn(cid:1) is locally finite. Let h = hn ∧ gn; by Lemma 2.2 this supremum exists, agrees with the supn pointwise maximum, and belongs to LPA. For every n and every x ∈ Kn, we have h(x) ≥ (hn ∧ gn)(x) = hn(x) ∧ Mn = hn(x) ≥ f (x) − ε. It follows that h ≥ f − ε1, so that h − f ≥ −ε1. On the other hand, for every n and every x ∈ eKn we have (hn ∧ gn)(x) ≤ hn(x) ≤ f (x) + ε. 8 S. ADEEB AND V. G. TROITSKY Since hn ∧ gn ≤ gn, it vanishes outside of eKn. It follows that hn ∧ gn ≤ f + ε1. Taking sup over n we get h ≤ f + ε1, so that h − f ≤ ε1. Therefore, h − f ≤ ε1. Suppose now that f is an arbitrary function in C(Rm). Then f = f + − f −, with 0 ≤ f +, f − ∈ C(Rm). By the first part of the proof, we find h1 and h2 in LPA such that f + − h1 ≤ ε 1. Put h = h1 − h2, then 1 and f − − h2 ≤ ε 2 2 f − h ≤ f + − h1 + f − − h2 ≤ ε1. (cid:3) 5. PA is σ-order dense in C(Rm) Recall that a sublattice F of an Archimedean vector lattice E is said to be order dense if for every 0 < x ∈ E there exists y ∈ F with 0 < y ≤ x. Equivalently, for every 0 < x ∈ E we have x = sup A where A = {y ∈ F : 0 ≤ y ≤ x}; see, e.g., [6, Theorem 1.34]. Note that the set A may be viewed as an increasing net. Therefore, F is order dense in E iff every element of E+ is the supremum of an increasing net in F+. Let f ∈ C(Rm). Suppose that f > 0; that is, f (x) ≥ 0 for every x ∈ Rm and f (c) > 0 for some c ∈ Rm. It follows easily from Lemma 2.3 that there exists a locally piecewise function h such that 0 < h ≤ f . This means that PA is an order dense sublattice in C(Rm). It follows that if 0 ≤ f ∈ C(Rm) then f is the supremum of an increasing net (fα) in PA, that is, fα ↑ f . This leads to two natural questions. Question 5.1. Can we replace a net with a sequence? Question 5.2. How do we deal with non-positive functions? The answer to the first question is related to the countable sup prop- erty. Recall that a vector lattice E has the countable sup property if every increasing net (xα) in E with x = sup xα has an increasing sub- sequence with the same supremum. Thus, to answer the first question, it suffices to prove that C(Rm) has the countable sup property. Recall that for a sequence (xα) in a vector lattice E, the notation xα ↑ means (xα) is increasing; xα ↑≤ x means xα ↑ and xα ≤ x for LOCALLY PIECEWISE AFFINE FUNCTIONS 9 every α, and xα ↑ x means xα ↑ and x = supα xα. Note that fα ↑ f in C(Ω) does not imply pointwise convergence. For f, g ∈ C(Rm), the notation f < g means f ≤ g and f 6= g. For f ∈ C(Rm) and a solid domain Ω in Rm, we write fΩ for the restriction of f to Ω; we view it as an element of C(Ω). Lemma 5.3. Let (fα) be a net in C(Rm) and f ∈ C(Rm). Then fα ↑ f in C(Rm) iff fαΩn ↑ fΩn in C(Ωn) for every n. Proof. Suppose that fα ↑ f in C(Rm). Fix n. It is clear that (fαΩn) is an increasing net in C(Ωn) and fΩn is an upper bound of this net. Suppose it is not the least upper bound. Then there exists g ∈ C(Ωn) such that fαΩn ≤ g for every α but g(t0) < f (t0) for some t0 ∈ Ωn. Since Ωn is a solid domain, we may assume that t0 ∈ Int Ωn. It follows easily that there exists a neighborhood V of t0 and ε > 0 such that V ⊆ Int Ωn and g(t) < f (t) − ε for all t ∈ V . Find 0 < h ∈ C(Rm) such that h(t0) > 0, h(t) < ε for all t ∈ V , and h vanishes outside V . Then f − h < f , yet fα ≤ f − h for every α; this contradicts f = sup fα. Conversely, suppose that fαΩn ↑ fΩn in C(Ωn) for every n. First, we claim that fα ↑≤ f in C(Rm). Indeed, fix any t0 ∈ Rm. Find n so that t0 ∈ Ωn. It follows from fαΩn ↑ fΩn that fα(t0) ↑≤ f (t0). In particular, f is an upper bound of (fα) in C(Rm). On the other hand, suppose that g is an upper bound of (fα) in C(Rm). Fix any n. Then gΩn ≥ fαΩn ↑ fΩn yields gΩn ≥ fΩn. This is true for all n; therefore, g ≥ f . It follows that f = supα fα. (cid:3) Theorem 5.4. C(Rm) has the countable sup property. Proof. Let (fα)α∈Λ be a net in C(Rm) indexed by some directed set Λ such that fα ↑ f for some f ∈ C(Rm). Fix n ∈ N. By Lemma 5.3, we have fαΩn ↑ fΩn in C(Ωn). Recall that every vector lattice which ad- mits a strictly positive functional has the countable sup property; see, e.g., [5, Theorem 2.6]. Since the Riemann integral is a strictly positive functional on Ωn, it follows that C(Ωn) has the countable sup prop- erty. Therefore, the net (fαΩn) has a subsequence whose supremum in C(Ωn) is fΩn. 10 S. ADEEB AND V. G. TROITSKY We will now inductively construct a double sequence (αn,k) in Λ indexed by pairs (n, k) ∈ N2 satisfying the following two conditions (1) (αn,k) is increasing: if n1 ≤ n2 and k1 ≤ k2 then αn1,k1 ≤ αn2,k2; (2) for every n we have fαn,k Ωn ↑ fΩn as k → ∞. For n = 1, let (α1,k)∞ fα1,k Ω1 k=1 be an increasing sequence in Λ such that ↑ fΩ1 as k → ∞; it exists by the preceding paragraph. Suppose that we have constructed αi,k for all i = 1, . . . , n − 1 and all k. We will now construct αn,k for all k. Find an increasing sequence ↑ fΩn in C(Ωn) as k → ∞. Now choose (βk) in Λ such that fβk Ωn αn,1 to be any index greater than β1 and αn−1,1. Choose αn,2 to be any index greater than β2, αn−1,2, and αn,1. Proceed inductively: once we have defined αn,1, . . . , αn,k−1, choose αn,k to be any index greater than βk, αn−1,k, and αn,k−1. This yields fΩn ≥ fαn,k Ωn ≥ fβk Ωn ↑ fΩn, ↑ fΩn in C(Ωn) as k → ∞. Also, it is easy to see that so that fαn,k Ωn (αn,k) is increasing. Consider (αk,k)∞ k=1. This is clearly an increasing sequence in Λ. If ↑ n ≤ k then αn,k ≤ αk,k, so that fαn,k ≤ fαk,k ≤ f . This yields fαk,k Ωn fΩn in C(Ωn) as k → ∞. It now follows from Lemma 5.3 that fαk,k ↑ f in C(Rm). (cid:3) Corollary 5.5. For every 0 ≤ f ∈ C(Rm) there exists a sequence (hn) of positive piecewise affine functions such that hn ↑ f in C(Rm). Thus, Question 5.1 has the affirmative answer. We now pass to Question 5.2. We answer it through order convergence. In literature, there are two slightly different definitions of an order convergent net in a vector lattice, which are, generally, not equivalent. We refer the reader to [10, 1, 2]. In our case, it will be irrelevant which of the two definitions is used for the following two reasons: first, both definitions yield the properties that we will be using and, second, by [10], the two definitions agree in vector lattices which have the countable sup property; hence, they agree in C(Rm). Proposition 5.6. Every function in C(Rm) is the order limit of an order convergent sequence of piecewise affine functions. LOCALLY PIECEWISE AFFINE FUNCTIONS 11 Proof. Let f ∈ C(Rm). Decompose it as f = f +−f −. By Corollary 5.5, there exist positive sequences (gn) and (hh) in PA such that gn ↑ f + and hn ↑ f − in C(Rm). It follows that the sequence (gn − hn) converges in order to f in C(Rm). (cid:3) Acknowledgement. We would like to thank Foivos Xanthos for valuable discussions. References [1] Y. Abramovich and C.D. Aliprantis, An invitation to Operator theory, Vol. 50. Providence, RI: American Mathematical Society, 2002. [2] Y. Abramovich and G. Sirotkin, On order convergence of nets, Positivity, 9, 2005, 287–292. [3] S. Adeeb, Locally piece-wise affine functions, M.Sc. Thesis, University of Al- berta, Department of Mathematical and Statistical Sciences, 2014. [4] C.D. Aliprantis and O. Burkinshaw, Principles of Real Analysis, Academic Press, New York and London, 3rd Edition, 1998. [5] C.D. Aliprantis and O. Burkinshaw, Locally solid Riesz spaces with applica- tions to economics, 2nd ed., AMS, Providence, RI, 2003. [6] C.D. Aliprantis and O. Burkinshaw, Positive Operators, 2nd ed., Springer 2006. [7] C.D. Aliprantis, D. Harris, and R. Tourky, Continuous piecewise linear func- tions, Macroeconomic Dynamics, 10(1) (2006), 77–99. [8] C.D. Aliprantis, D. Harris, and R. Tourky. Riesz estimators, J. Econometrics, 136(2) (2007), 431–456. [9] C.D. Aliprantis and R. Tourky, Cones and Duality, AMS, Providence, RI, 2007. [10] R.F. Anderson and J.C Mathews, A Comparison of Two Modes of Order Convergence, Proc. Amer. Math. Soc., 18 (1967), 100–104. [11] S. Ovchinnikov, Max-min representation of piecewise linear functions, Beit- rage Algebra Geom. 43 (2002), 297-302. Department of Civil and Environmental Engineering, University of Alberta, Edmonton, AB, T6G 1H9, Canada. E-mail address: [email protected] Department of Mathematical and Statistical Sciences, University of Alberta, Edmonton, AB, T6G 2G1, Canada. E-mail address: [email protected]
1403.0720
2
1403
2015-05-12T03:36:04
Invariant subspaces with no generator and a problem of H. Helson
[ "math.FA" ]
In the almost periodic context, any $H_0^2-$space cannot be generated by one of its elements. Together with cocycle argument, this derives that there exist all kinds of invariant subspaces without single generator, from which we can answer some questions on invariant subspace theory.
math.FA
math
INVARIANT SUBSPACES WITH NO GENERATOR AND A PROBLEM OF H. HELSON JUN-ICHI TANAKA Dedicated to the memory of Henry Helson Abstract. In the almost-periodic context, the H 2 0 −space cannot be generated by one of its elements. Together with a cocycle argument, this implies that there exist all kinds of invariant subspaces without single generator, from which we answer some questions on invariant subspace theory. 1. Introduction The theory of invariant subspaces has been developed in the context of compact abelian groups with ordered duals, which is a natural generalization of such a theory on the unit circle T. Many classical results extend to these cases, nevertheless, one also meets new difficulties. The purpose of this paper is to resolve a longstanding problem formulated by H. Helson in the 1950s. Let Γ be a countable dense subgroup of the real line R, endowed with the discrete topology. Then the dual group K of Γ is a compact abelian group that is metrizable. For λ in Γ, it is customary to denote by χλ the character on K defined by χλ(x) = x(λ). Let σ be the normalized Haar measure on K. A function φ in L1(σ) is analytic if its Fourier coefficients (1) aλ(φ) = ZK φ χλ dσ vanish for all negative λ in Γ. The Hardy space H p(σ), 1 ≤ p ≤ ∞, is defined to be the space of all analytic functions in Lp(σ). For technical reasons, it is useful to define H p 0 (σ) as the subspace of all φ in H p(σ) with a0(φ) = 0. A (weak*-, if p = ∞) closed subspace M of Lp(σ) is invariant if M contains χλM for all positive λ in Γ. When the inclusion is strict, M is said to be simply invariant. Of course, both H p(σ) and H p 0 (σ) are simply invariant subspaces of Lp(σ). If φ is in Lp(σ), and let M[φ] denotes the smallest invariant subspace of Lp(σ) containing φ, then φ is called a single generator of M[φ]. Recall that a function of modulus one is said to be unitary and an analytic Date: Received by editors November 21, 2009, and in revised form, January 17, 2013. 1991 Mathematics Subject Classification. Primary 43A17; Secondary 46J10, 46J15, 28D10. Key words and phrases. Compact groups with ordered duals, Invariant subspaces, Cocycles, Single generators. Partially supported by NSF grant no. 0649765. 1 unitary function is called an inner function. We say a function φ in H p(σ) is outer if it satisfies that log a0(φ) = ZK log φ dσ > −∞ . Let 1 ≤ q ≤ p ≤ ∞, and let M be a simply invariant subspace of Lp(σ). It follows from the properties of outer functions that [M ∩ L∞(σ)]q ∩ Lp(σ) = M, where [M ∩ L∞(σ)]q is the closure of M ∩ L∞(σ) in Lq(σ) (see [4, Chapter V, Section 6] for details). This fact assures that there is a one-to-one correspondence between the invariant subspaces in Lp(σ) and those in Lq(σ). Therefore, in dealing with invariant subspaces, we may restrict our attention to the case of p = 2, in which Hilbert space theory works well. It follows from Szego's theorem that φ is a single generator of H 2(σ) if and only if φ is outer in H 2(σ). However, it has been unknown for a long time whether every simply invariant subspace is singly generated or not. In the literature this has come to be known as the single generator problem (refer to [6, §5.4], [3, Remark, p. 158] and [4, p. 138 and p. 177]). The difficulty seems to center on the case of invariant subspace H 2 0 (σ). In [9, p. 183], it is raised in an equivalent form in connection with stochastic processes. Our objective in this note is to show a negative answer to this problem in the almost periodic settings: Theorem. The invariant subspace H 2 0 (σ) cannot be generated by one of its elements. To the best of author's knowledge, H 2 0 (σ) is the first known example of invariant subspace which cannot be singly generated. On the other hand, by [6, §5.3, Theorem 33], it was shown that every invariant subspace is generated by two of its elements. In more general setting, we can artificially make H 2 0 -spaces to have a single generator. For each t in R, let us denote by et the element of K defined by et(λ) = eiλt for λ in Γ. The map sending t to et embeds R continuously onto a dense subgroup of K. Define a one-parameter group {Tt}t∈R of homeomorphisms on K by (2) Tt x = x + et , x ∈ K. Then the pair (K, {Tt}t∈R) is a strictly ergodic flow, for which σ is the unique invariant probability measure. The flow (K, {Tt}t∈R) is called an almost periodic flow, because if φ is continuous on K, then t → φ(x + et) is a uniformly almost periodic function with exponents in Γ. Let H ∞(dt/π(1+t2)) be the space of all boundary functions of bounded analytic functions in the upper half-plane H, and let H p(dt/π(1 + t2)), 1 ≤ p < ∞, be the closure of H ∞(dt/π(1+t2)) in Lp(dt/π(1+t2)). For a function u(x, t) on K × R, the assertion " t → u(x, t) for σ − a.e. x in K " is sometimes abbreviated to "almost every t → u(x, t) ". Then φ in Lp(σ) lies in H p(σ) if and only if almost every t → φ(x + et) lies in H p(dt/π(1 + t2)). This fact enables us to define Hardy spaces on every ergodic flow (see the end of the next section). 2 Let M be a simply invariant subspace of L2(σ). Set Mλ = χλM for each λ in Γ. Define M+ = ^λ<0 Mλ and M− = _λ>0 Mλ. Since these spaces are at most one dimension apart, M coincides with either or both its versions M+ and M−. When M = M+, M is said to be normalized. For φ in L2(σ), the subspace M[φ] is simply invariant if and only if (3) Z ∞ −∞ log φ(x + et) dt 1 + t2 > −∞ , σ − a.e. x ∈ K, It is well-known that there is a function φ in L2(σ) (see [6, §3.3, Theorem 22 ]). satisfying the inequality (3), while log φ does not belong to L1(σ). Our Theorem asserts that any such function φ must satisfy M[φ]+ = M[φ]−. A unitary Borel function A(x, t) on K × R is said to be a cocycle on K if A(x, t) satisfies the cocycle identity A(x, t + s) = A(x, t) · A(x + et, s) , (x, s, t) ∈ K × R × R . We identify two cocycles which differ only on a set of dσ ×dt−measure zero in K ×R. A one-to-one correspondence is established between normalized invariant subspaces and cocycles (as discussed in [6, §2.3]). More precisely, let M be a simply invariant subspace of L2(σ) with cocycle A(x, t). Then a function φ in L2(σ) lies in M+ if and only if almost every t → A(x, t)φ(x + et) lies in H 2(dt/π(1 + t2)) (see [6, §3.2]). It is easy to see that M+ 6= M− if and only if M+ = qH 2(σ) for some unitary function q on K. Then the cocycle of M has the form q(x) · q(x + et), which is called a coboundary. If a cocycle is a coboundary multiplied by exp(iαt) for some α in R, then such a cocycle is said to be trivial. A trivial cocycle exp(iαt) is not a coboundary only if α lies in R \ Γ. We already know from [8] and [12] that some singly generated subspaces have non- trivial cocycles, but we can strengthen this fact by noting the following: Corollary 1. Let M be a simply invariant subspace of L2(σ). If the cocycle of M is trivial, then M− has no single generator. In other words, if M− is singly generated, then the cocycle of M is always nontrivial, so that M+ = M−. A cocycle with values in {−1, 1} is called a real cocycle. It follows from [10] that there exist real cocycles which are nontrivial. Corollary 2. Let M be a simply invariant subspace of L2(σ) with real cocycle. Then M− has no single generator. A cocycle A(x, t) is said to be analytic if almost every t → A(x, t) lies in H ∞(dt/π(1+ t2)). Then a normalized invariant subspace with analytic cocycle contains always H 2(σ). We say that an analytic cocycle A(x, t) is a Blaschke or a singular cocycle, if almost every t → A(x, t) is an inner function of that type in H ∞(dt/π(1 + t2)). Two cocycles are called cohomologous if one is a coboundary times the other. It is known that every cocycle is cohomologous to a Blaschke cocycle in some restricted class (see 3 [6, §4.6, Theorem 26] and [17]). This fact makes Blaschke cocycles so important for the subject. Using our Theorem, we may answer some questions on analytic cocycles: Corollary 3. In the class of analytic cocycles, the following properties hold: (a) There is a Blaschke cocycle not being cohomologous to any singular cocycle. (b) There is a Blaschke cocycle not having exactly the same zeros as any function in H 2(σ). It would be helpful to understand the basic idea behind the proof of our Theorem. On the one hand, we claim that if φ is a single generator of H 2 0 (σ), then φ must have a very special form. Assume that Γ is the smallest group determined by the nonzero Fourier coefficients of φ (see below for details). Similarly, let Λ be the smallest group determined by the nonzero coefficients of φ. Since Λ is a subgroup of Γ, the dual group of Λ is represented as K/H, where H is the annihilator of Λ in K. Let τ be the normalized Haar measure on K/H, and fix an element α in Γ with aα(φ) 6= 0. Then it can be shown that χαφ lies in L2(τ ) and generates the simply invariant subspace of L2(τ ) with trivial cocycle exp(iαt). We also see that α is independent of Λ, meaning that nα lies in Λ only for n = 0 in the integer group Z. This implies that K and dσ are respectively identified with K/H × T and dτ × dθ/2π, since H is regarded as T. Thus, for each single generator φ of H 2 0 (σ), we derive that Γ 6= Λ. On the other hand, if H 2 0 (σ) with the property that Γ = Λ, which contradicts the existence of single generator of H 2 0 (σ) is singly generated, we may construct a generator φ of H 2 0 (σ). In the next section, we establish some notation and elementary facts about invariant subspaces in the almost periodic setting. Using group characters, we develop certain properties of single generators of H 2 0 -spaces in Section 3. In Section 4, the proof of our Theorem is provided and then Corollaries are proved by using a lemma on cocycles. We conclude the paper with some remarks in Section 5. We refer the reader to [1], [4, Chapter VII], [6] and [16, Chapter VIII] for further details on analyticity on compact abelian groups. Basic results concerning the Hardy space theory based on uniform algebras can be found in [4, Chapter IV] and [13]. Part of this work was done while the author was visiting the University of North Carolina at Chapel Hill, and he would like to acknowledge the hospitality of the De- partment of Mathematics. Especially, he would like to express his sincere gratitude to Professors Joe Cima and Karl Petersen for helpful discussions. Thanks are due to the referee as well for his valuable suggestions, which improved the first version of this paper so much. 4 2. Extension of almost periodic functions It is easy to show that a function φ in H 2(σ) is outer if and only if a0(φ) 6= 0 and almost every t → φ(x + et) is outer in H 2(dt/π(1 + t2)). A weak version of this fact stated below is often used in what follows: Lemma 1. Let M be a simply invariant subspace of L2(σ) with cocycle A(x, t). A function φ in L2(σ) generates M− if and only if log φ does not lie in L1(σ) and almost every t → A(x, t)φ(x + et) is outer in H 2(dt/π(1 + t2)). In particular, H 2 0 (σ) is singly generated by φ if and only if a0(φ) = 0 and almost every t → φ(x + et) is outer in H 2(dt/π(1 + t2)). Proof. Suppose that M[φ] = M− for φ in L2(σ). If log φ lies in L1(σ), then there is a unitary function q on K such that M[φ] = qH 2(σ) by Szego's theorem. This implies that M[φ] 6= M−, so log φ cannot lie in L1(σ). Let B(x, t) be the analytic cocycle defined by the inner part of t → A(x, t)φ(x + et). Let N be the invariant subspace with cocycle AB(x, t). By [6, §3.2, Theorem 21], we see that N− is contained in M−. On the other hand, since almost every t → AB(x, t)ψ(x + et) lies in H 2(dt/π(1 + t2)) for each ψ in M[φ], N+ includes M[φ]. This shows that N+ = M+, so B(x, t) ≡ 1. Then almost every t → A(x, t)φ(x + et) is outer in H 2(dt/π(1 + t2)). Conversely, suppose that M[φ] is contained strictly in M−. Then there is a nonzero function q in M− such that ZK ψφq dσ = 0 , ψ ∈ H ∞(σ) . This shows that φq lies in H 1(σ), so almost every t → φq(x+et) lies in H 1(dt/π(1+t2)). Notice that t → A(x, t)q(x + et) is in H 2(dt/π(1 + t2)). Since φ(x + et)q(x + et) = A(x, t)φ(x + et)A(x, t)q(x + et), and since t → A(x, t)φ(x + et) is outer in H 2(dt/π(1 + t2)), we see that almost every t → A(x, t)q(x + et) is also in H 2(dt/π(1 + t2)). This shows that t → A(x, t)q(x + et) is constant for σ−a.e. x in K, and so is t → q(x+et). It follows from the ergodic theorem that q(x) is constant. We then assume q is a unitary function on K. Therefore, A(x, t) is the coboundary q(x)q(x + et) and M− = qH 2 0 (σ). Thus q does not lie in M−, which is a contradiction. The last part of assertion follows from the fact that the cocycle of H 2(σ) equals 1. Under the assumption that almost every t → φ(x + et) is outer in H 2(dt/π(1 + t2)), we see easily a0(φ) = 0 if and only if log φ does not lie in L1(σ). Then M[φ] = H 2 0 (σ), so the proof is complete. (cid:3) Let L1(dt) be the usual Lebesgue space on R. Using {Tt}t∈R, one may convolve a function φ in Lp(σ), 1 ≤ p < ∞, with a function f in L1(dt) by setting φ(x − et)f (t) dt , (φ ∗ f )(x) = Z ∞ −∞ φ(x + et)f (−t) dt = Z ∞ −∞ 5 where the integral is a Bochner integral. When p = ∞, the convolution φ ∗ f is defined in the same way as the weak*-convergent integral. Under the operation of convolution, Lp(σ) becomes an L1(dt)-module such that kφ ∗ f kp ≤ kφkpkf k1 , φ ∈ Lp(σ) , for f in L1(dt). The Fourier transform f of f is defined by the formula (4) f (λ) = Z ∞ −∞ f (t)e−iλt dt, λ ∈ R , as usual. We see easily aλ(φ∗f ) = aλ(φ) f (λ), if λ is in Γ. The Poisson kernel Pir(t) for H is given by Pir(t) = r/π(t2 + r2), for an r > 0. If φ is in L1(σ), then the convolution φ ∗ Pir is considered as the Poisson integral of t → φ(x + et), that is, (φ ∗ Pir)(x + es) = Z ∞ −∞ φ(x + et)Pir(s − t) dt . Lemma 2. Suppose that H 2 properties: 0 (σ) is singly generated. Then we obtain the following (a) There is a single generator of H 2 (b) If φ is a bounded generator of H 2 with r > 0 and φn for n = 1, 2, · · · . 0 (σ) that is bounded. 0 (σ), then so is each of the functions φ ∗ Pir 0 (σ). Then there is an outer function h in H 2(σ) Proof. Let ψ be a single generator of H 2 such that h = min(1, ψ−1). It follows from Lemma 1 that the bounded function ψh generates H 2 0 (σ), thus we obtain (a). To show (b), we observe that t → (φ ∗ Pir)(x + et) as well as t → φn(x + et) is outer in H 2(dt/π(1 + t2)) for σ − a.e. x in K. Since a0(φ ∗ Pir) = a0(φn) = 0, (b) follows from Lemma 1 immediately. (cid:3) We next introduce a local product decomposition of K, which is useful for studying analytic functions on K. Fix a positive γ in Γ, and let Kγ be the closed subgroup of all x in K such that χγ(x) = 1. Then Kγ × [0, 2π/γ) is identified with K via the map (y, s) → y + es. Let σ1 be the normalized Haar measure on Kγ. Then the probability measure (γ/2π)dσ1 × dt on Kγ × [0, 2π/γ) is carried by the map to dσ on K. The one-parameter group {Tt}t∈R given by (2) is represented as Tt(y, s) = (y + [(t + s)γ/2π]e2π/γ, t + s − [(t + s)γ/2π]2π/γ) on Kγ × [0, 2π/γ), where [t] is the largest integer not exceeding t. Define the home- omorphism T on Kγ by T y = y + e2π/γ. We denote by O(ω, T ) the orbit of a point ω in (Kγ, T ), that is, the set of all T nω for n in Z. Since O(ω, T ) is dense in Kγ, the discrete flow (Kγ, T ) is also a strictly ergodic flow, on which σ1 is the unique invariant probability measure. Since Γ is countable, Kγ is metrizable (see [16, 2.2.6]). 6 A function φ on K has the automorphic extension φ♯ to Kγ × R defined by φ♯(y, t) = φ(y + [tγ/2π]e2π/γ , t − [tγ/2π]2π/γ). Since a function f in H 1(dt/π(1 + t2)) extends analytically to H by f (s + ir) = (f ∗ Pir)(s), we write φ♯(y, z) = (φ♯ ∗ Pir)(y, s), z = s + ir ∈ H , for each φ in H 1(σ). It is clear that (φ♯ ∗ Pir)(y, s) = (φ ∗ Pir)♯(y, s) on Kγ × R. The following is due to a property of Lebesgue sets. Lemma 3. If E1 is a compact subset of Kγ with σ1(E1) > 0, then there is a closed subset E of E1 with σ1(E1) = σ1(E) such that O(ω, T )∩E is dense in E, for σ1−a.e. ω in Kγ. Proof. Recall that the metric density of E1 is 1 at σ1 − a.e. ω in E1, meaning that lim δ→0 σ1(E1 ∩ B(ω, δ)) σ1(B(ω, δ)) = 1 , where B(ω, δ) is the open ball with center ω and radius δ > 0. Define E to be the closure of the set of points of E1 at which the metric density of E1 is 1. Clearly, we have σ1(E1) = σ1(E), since E1 is closed. If σ1(E) = 1, then E = Kγ. Since (Kγ, T ) is strictly ergodic every orbit O(ω, T ) is dense in E. Assume that 0 < σ1(E) < 1. It follows from the ergodic theorem that there is a σ1−null set N in Kγ outside which lim n→∞ 1 n n−1Xj=0 IE(T jω) = σ1(E) , where IE denotes the characteristic function of E. Let Hω be the closure of O(ω, T ) ∩E in Kγ. We claim that if E 6= Hω, then ω lies in N. Indeed, we see that σ1(E \Hω) > 0, since the metric density of E does not vanish identically on E\Hω. Let p be a continuous p dσ < σ1(E). Since function on Kγ such that 0 ≤ p ≤ 1, p ≡ 1 on Hω, and RKγ p(T jω) = ZKγ IE(T jω) ≤ lim n→∞ IE(T jω) = IHω (T jω) for j in Z and since (Kγ, T ) is strictly ergodic, we have p dσ1 < σ1(E) 1 n n−1Xj=0 lim sup n→∞ 1 n n−1Xj=0 by [15, §4.2, Proposition 2.8]. Thus ω has to lie in the null set N. (cid:3) For each φ in H ∞(σ), there is a σ1-null set of Kγ outside which z → φ♯(y, z) is analytic and uniformly bounded on the upper half plane H. Recall that if a family of analytic functions is uniformly bounded, then it forms a normal family. The next proposition may be regarded as a strengthened form of Lusin's theorem for analytic functions on K, so that it has some interest of its own. Here we denote by cl(H) the closure of H in R2. 7 Proposition 1. Let φ be a function in H ∞(σ), and let ǫ > 0. Then there is a closed subset E of Kγ with σ1(E) > 1 − ǫ having the following properties: (a) The convolution (φ♯ ∗ Pir)(y, t) is continuous on E × R, for a given r > 0. (b) For σ1 − a.e. ω in Kγ, the function (φ♯ ∗ Pir)(T jω, z) on (O(ω, T ) ∩ E) × cl(H) extends to (φ♯ ∗ Pir)(y, z) on E × cl(H). Proof. Since φ ∗ Pir lies in H ∞(σ), Lusin's theorem asserts that there is a compact subset F of K with σ(F ) > 1 − ǫ2 on which φ ∗ Pir is continuous. Regarding F as a subset of Kγ × [0, 2π/γ), we choose a compact subset E of Kγ with σ1(E) > 1 − ǫ such that E satisfies the property of Lemma 3 and (5) γ 2πZ 2π/γ 0 IF (y, s) ds > 1 − ǫ , y ∈ E . In addition, we assume that z → (φ♯ ∗ Pir/2)(y, z) is analytic on H and (φ♯ ∗ Pir/2)(y, z) ≤ kφk∞ , y ∈ E . Then the family F = (cid:8) (φ♯ ∗ Pir/2)(y, z) ; y ∈ E(cid:9) forms a normal family on H. Let {yn} be a sequence in E tending to y. Since F is normal, there is a subsequence {yj} of {yn} such that (φ♯ ∗ Pir/2)(yj, z) converges uniformly on compact subsets of H to a bounded analytic function f (z) on H. Let us show that f (z) = (φ♯ ∗Pir/2)(y, z). Indeed, we observe by (5) that F ∩ ({y}×[0, 2π/γ)) contains an infinite compact set of the form {y} × J. Since (φ♯ ∗ Pir)(y, t) = (φ♯ ∗ Pir/2)(y, t + ir/2) = f (t + ir/2), t ∈ J , it follows from the uniqueness principle that f (z) = (φ♯ ∗ Pir/2)(y, z). This shows that if (yn, tn) tends to (y, t), then (φ♯ ∗ Pir)(yn, tn) tends to (φ♯ ∗ Pir)(y, t). Thus (a) holds. We notice that (φ♯ ∗ Pir/2)(y, z) is also continuous on E × H. On the other hand, by Lemma 3, O(ω, T ) ∩ E is dense in E for σ1 − a.e. ω in Kγ. Since (O(ω, T ) ∩ E) × cl(H) is dense in E × cl(H) and since (φ♯ ∗ Pir)(y, z) is continuous on E × cl(H), the function (φ♯ ∗ Pir)(T jω, t) on (O(ω, T ) ∩ E) × cl(H) extends to (φ♯ ∗ Pir)(y, t) on E × H. Thus (b) follows immediately. (cid:3) We make some remarks on Proposition 1. Since t → φ♯(y, t) lies in H ∞(dt/π(1 + t2)) for each y in E, we see that (φ♯ ∗ Pir)(y, t + 2π/γ) = (φ♯ ∗ Pir)(T y, t). Then E ∪ T E ∪ · · ·∪T nE also satisfies the properties (a) and (b) and σ1(E ∪T E ∪· · · ∪T nE) converges to 1, as n → ∞, by the recurrence theorem (see [15, §2.3, Theorem 3.2]). However, to obtain φ itself, we need a version of Fatou's theorem as discussed in [14, Theorem II]. Denote by O(x, {Tt}t∈R) the orbit of x in (K, {Tt}t∈R). With the notation above, when x = (y, s) in Kγ × [0, 2π/γ), we see that O(x, {Tt}t∈R) = O(y, T ) × [0, 2π/γ). For x in K, we say that t → (φ ∗ Pir)(x + et) extends to φ ∗ Pir if, for each ǫ > 0, there is a compact subset F = F (ǫ, φ) of K with σ(F ) > 1 − ǫ such that φ ∗ Pir is continuous 8 on F and O(x, {Tt}t∈R) ∩ F is dense in F . The above proof may be modified so as to apply to functions in H 1(σ) as well. The next lemma is an immediate consequence of Proposition 1. Lemma 4. Let φ be a function in H ∞(σ), and let r > 0. Then there is an invariant σ−null set N = N(φ) in K outside which t → (φ ∗ Pir)(x + et) extends to φ ∗ Pir. Proof. For a given ǫ > 0, let E be a closed subset of Kγ with σ1(E) > 1 − ǫ which has the property (a) and (b) of Proposition 1. Putting F = E × [0, 2π/γ], we regard F as a compact subset of K. By (b) of Proposition 1, we choose an invariant null set N ′ = N ′(φ) in (Kγ, T ) outside which O(ω, T ) ∩ E is dense in E. If we set N = N ′ × [0, 2π/γ), then the σ−null set N satisfies the desired property. (cid:3) Let Ω be a compact metric space on which R acts as a Borel transformation group. This means that there is a one-parameter group {Ut}t∈R of Borel isomorphisms on Ω such that the map (ω, t) → Utω of Ω × R to Ω is a Borel map. The pair (Ω, {Ut}t∈R) is referred to a Borel flow. Especially, (Ω, {Ut}t∈R) is called a continuous flow, if Ut is a homeomorphism on Ω and the map (ω, t) → Utω is continuous on Ω × R. We often write ω + t for the translate Utω of ω by t. Let µ be an invariant probability measure on (Ω, {Ut}t∈R) which is ergodic, meaning that µ(E) = 1 or 0 for each invariant subset E of Ω. A function φ in L1(µ) is analytic if t → φ(ω + t) lies in H 1(dt/π(1 + t2)) for µ − a.e. ω in Ω. Then the ergodic Hardy space H p(µ), 1 ≤ p ≤ ∞, is defined to be the space of all analytic functions in Lp(µ). It follows from [13, Theorem I] that µ is a representing measure for H ∞(µ), for which H ∞(µ) is a weak*-Dirichlet algebra in L∞(µ). This fundamental result enables us to apply the Hardy space theory based on uniform algebras, and most of the machinery of invariant subspaces on an almost periodic flow (K, {Tt}t∈R) can be reconstructed (see [2], [13] and [14] for related topics). As we mentioned earlier, the H 2 0 −spaces may be singly generated in the situation of ergodic flows other than almost periodic flows (see [18] and §5 (b)). Let A(x, t) be a cocycle on an almost periodic flow (K, {Tt}t∈R) and define the Borel flow (K × T, {St}t∈R) by (6) St(x, eiθ) = (Ttx, A(x, t)eiθ), (x, eiθ) ∈ K × T, which is called the skew product of K and T induced by A(x, t). Then dσ × dθ/2π is an invariant probability measure on K ×T. Observe that each function f in L2(dσ×dθ/2π) is represented as f (x, eiθ) = ∞Xn=−∞ φn(x)einθ , where the coefficients φn are in L2(σ). From this fact, it follows easily that dσ × dθ/2π is ergodic on (K × T, {St}t∈R) if and only if A(x, t)n is a coboundary only for n = 0 (see [6, §6.2] for details). 9 3. Approximation to generators We now turn to the structure of compact group K, under the assumption that H 2 0 (σ) is singly generated by φ in H 2 0 (σ). By multiplying by a suitable outer function, if necessary, we can assume that φ is a function in L∞(σ) with 1 ≤ kφk∞ < +∞. Furthermore, we also assume that Γ is the smallest group containing all λ such that aλ(φ) 6= 0, that is, the smallest group over which Fourier series, φ(x) ∼ XΓ∋λ>0 aλ(φ)χλ(x) , holds. Similarly, denote by Λ the smallest group containing all λ such that aλ(φ) 6= 0. We observe that the Fourier series of (cid:0)φ2 + ǫ(cid:1)1/2 = exp(cid:26) 1 2 log (φφ + ǫ)(cid:27) , ǫ > 0 , is represented on Γ, by considering the Taylor series of z → log z at a large positive. This shows that Λ is a subgroup of Γ, since aλ(φ) = lim ǫ→0 aλ(cid:0)(φ2 + ǫ)1/2(cid:1) by (1). Since log φ does not lie in L1(σ), the generator φ cannot be periodic in (K, {Tt}t∈R). Then Γ as well as Λ is a countable dense subgroup of R, endowed with discrete topology. Let H be the annihilator of Λ, meaning that H is the closed subgroup of all x in K such that χλ(x) = 1 for all λ in Λ. Then the dual group of Λ is identified with the quotient group K/H (see [16, 2.1]). We denote by τ the normalized Haar measure on K/H. Let π be the canonical homomorphism of K onto K/H. For each x in K, we write ¯x for π(x) = x+H. When a function ψ on K is represented as ψ = ψ◦π for a function ψ on K/H, we usually identify ψ with ψ, so that ψ(x) = ψ(¯x). Then we say descriptively that ψ is generated by a function on K/H. If 1 ≤ p ≤ ∞, then Lp(τ ) and H p(τ ) are subspaces of Lp(σ) and H p(σ), respectively. Since almost every t → φ(x + et) is outer in H ∞(dt/π(1 + t2)) by Lemma 1, we see that −∞ < log (φ ∗ Pir)(x) = (log φ ∗ Pir)(x) for a given r > 0. Since log φ is not in L1(σ) and log φ ≤ kφk∞, Fubini's theorem shows that ZK log φ ∗ Pir dσ = ZK (log φ ∗ Pir) dσ = ZK log φ dσ = −∞ . Let g = φ ∗ Pir. Then Lemma 1 shows that g is also a bounded generator of H 2 0 (σ). Since Pir(λ) = e−rλ by (4), we obtain aλ(g) = aλ(φ ∗ Pir) = aλ(φ)e−rλ, hence aλ(φ) 6= 0 if and only if aλ(g) 6= 0. Thus the generator g plays the same role as φ. For n = 1, 2, . . . , we then denote by φn the outer function in H ∞(σ) with φn = max(1/n, φ). Since − log n ≤ log φn ≤ kφk∞, each φ−1 is also an outer n function in H ∞(τ ). Putting gn = φn ∗Pir, we obtain a sequence {gn} of outer functions 10 in H ∞(τ ) with kgnk∞ ≤ kφk∞. Notice that t → g(x + et) and t → gn(x + et) extend analytically up to {Re z > −r}. Let us look into the relation between g and gn. Since gn(x) = exp{(log φn ∗ Pir)(x)} , we obtain g1(x) ≥ g2(x) ≥ · · · ≥ gn(x) −→ g(x) , (7) for σ −a.e. x in K. Although g may not be in L∞(τ ), we observe that gn(x) = gn(¯x) and g(x) = g(¯x). By (7), it is easy to see that almost every t → (g/gn)(x + et) n(t) = gn(x+et) and Gx(t) = g(x+et). Let N0 be converges pointwise to 1 on R. Put Gx an invariant null set in K outside which the property of Lemma 4 holds simultaneously for φ and all φn. Moreover, for x in K \ N0, we may assume Gx n(t) and Gx(t) are outer functions in H ∞(dt/π(1 + t2)). Then the family of all analytic extensions Gx n(z) of Gx n → ∞ , n(t) to {Re z > −r} forms a normal family, since Gx The following lemma is crucial in our proof of the Theorem. n(z) ≤ kφk∞. Lemma 5. For a bounded generator φ of H 2 0 (σ), let Λ, H and τ be as above. Choose an α in Γ with aα(φ) 6= 0. Then χαφ is generated by a function on K/H, so lies in L∞(τ ). Consequently, Γ is generated by Λ and α. Proof. Let {δk} be a decreasing sequence tending to 0. Then there is a sequence {fk} in L1(dt) such that fk(α) = 1, kfkk1 = 1 and fk = 0 outside (α − δk, α + δk), by modifying the function t → (1/π) sin2 t/t2 in L1(dt). Since aλ(g) = aλ(φ)e−rλ, we see that χαφ lies in L2(τ ) if and only if so does χαg. Thus we may replace φ with g in our argument. Since aλ(g ∗ fk) = aλ(g) fk(λ), we observe that kg ∗ fk − aα(g)χαk2 aα+λ(g) fk(α + λ)2 → 0 , k → ∞ , 2 = X0<λ<δk by the Parseval theorem and that k(g ∗ fk)g − aα(g)(χαg)k2 ≤ kg ∗ fk − aα(g)χαk2 kgk∞ . From these facts, we conclude that if each (g ∗ fk)g lies in L∞(τ ), then so does χαg. Since the outer function φn lies in L∞(τ ), so do gn and gn ∗ fk. Then each (gn ∗ fk)gn lies in L∞(τ ). Let us show that the sequence {(gn ∗ fk)gn} converges to {(g ∗ fk)g} in L2(σ), from which we obtain that (g ∗ fk)g lies in L∞(τ ). Indeed, in the notation above, if we fix an x in K \ N0, there is a subsequence {gm} of {gn} such that {Gx m(t)} converges pointwise to eiγGx(t) in H ∞(dt/π(1 + t2)) with 0 ≤ γ < 2π, where γ depends on x and {gm}. This implies that (gm ∗ fk)(x + et) → e−iγ(g ∗ fk)(x + et), m → ∞, pointwise in L∞(dt/π(1 + t2)). Note that every subsequence of {gn} contains such a subsequence {gm}. Since e−iγeiγ = 1, the sequence {gn} itself satisfies (gn ∗ fk) gn(x + et) → (g ∗ fk) g(x + et) , n → ∞, 11 pointwise in L∞(dt/π(1 + t2)). Since k(gn ∗ fk) gnk∞ ≤ kgnk2 ∞kfkk1 ≤ kφk2 ∞kfkk1, it follows from the bounded convergence theorem that k(gn ∗ fk) gn − (g ∗ fk) gk2 → 0 , n → ∞, so that (g ∗ fk) g lies in L∞(τ ). Therefore, χα g as well as χα φ is generated by a function on K/H. On the other hand, by the property of Γ, each element in Γ has the form λ + nα for λ in Λ and n in Z, thus the proof is complete. (cid:3) Recall that K/H coincides with the dual group of Λ. Let α be as in Lemma 5 and let C(¯x, t) be the trivial cocycle on K/H defined by C(¯x, t) = exp(iαt). Since α is positive, C(¯x, t) is an analytic cocycle. We denote by (K/H × T, {St}t∈R) the skew product of K/H and T induced by C(¯x, t), which is the continuous flow obtained by St(¯x, eiθ) = (Tt ¯x, C(¯x, t)eiθ), (¯x, eiθ) ∈ K/H × T . Then dτ × dθ/2π is the invariant probability measure on K/H × T (see the end of the preceding section). Let us represent the generator g and all the limits of subsequences of {gn} on K/H × T, which is the smallest product group with such property. Each function ψ on K/H extends naturally to the one on K/H × T by setting ψ(¯x, eiθ) = ψ(¯x). Since g and gn are functions on K/H, they belong to L∞(dτ × dθ/2π). With the above notation, we fix a w in K \ N0. Since Gw n (t) and Gw(t) are outer functions in H 2(dt/π(1 + t2)) which extend analytically to {Re z > −r}, we may n (t) converges pointwise to Gw(t) on R, by multiplying each gn by a assume that Gw suitable constant of modulus one. By regarding Lemma 4, the functions Gw n (t) and Gw(t) extend to gn and g, respectively. However, we obtain the following: Lemma 6. For σ − a.e. x in K, Gx we find two subsequences {gm} and {gk} of {gn} such that Gx to eiβGx(t) and eiγGx(t) with 0 ≤ β < γ < 2π, respectively. n(t) never converges pointwise on R. Consequently, k(t) converge m(t) and Gx Proof. Since 1/n ≤ gn(x) ≤ kφk∞, each g−1 n is also an outer function in H ∞(σ). This implies that almost every t → (g/gn)(x + et) is an outer function in H ∞(dt/π(1 + t2)). Furthermore, since a0(g/gn) = ZK g−1 n dσ = 0 , g/gn dσ = ZK g dσZK Lemma 1 assures that each g/gn is also a single generator of H 2 Denote by F the invariant set of all x in K for which {Gx 0 (σ). n(t)} itself converges. Suppose that F has positive measure. By (7) and the ergodic theorem, (g/gn)(x) converges to an invariant function on F , so to a constant of modulus one on K. Then the bounded convergence theorem shows that a0(g/gn) 6= 0 for large n. Such g/gn cannot be a single generator of H 2 (cid:3) 0 (σ), which contradicts the above observation. 12 Let us mention a few remarks derived from Lemma 6. When 0 ≤ β < 2π, Z(β) denotes the subgroup of T generated by eiβ, that is, Z(β) = (cid:8)eijβ ; j ∈ Z(cid:9) . If β/2π is rational, then the order of Z(β) is finite. Fix two points w and x in K \ N0. We assume by Lemma 6 that a subsequence {gk} of {gn} satisfies that Gw k(t) converge respectively to eijβGw(t) and ei(j+1)βGx(t) for j in Z, by multiplying each gk by a suitable constant of modulus one. Denote by O( ¯w) the orbit O( ¯w, {Tt}t∈R) of ¯w in (K/H, {Tt}t∈R). Then g is determined naturally on O( ¯w) × Z(β) and O(¯x) × Z(β) to represent the limits of the subsequence {gk} of {gn} on them. For each m in Z, we see also that every limit of {gm k } is represented on these product subsets. k (t) and Gx k(t)} of {Gx j (t)ℓ} of {Gx If ℓ is a positive integer, then gℓ as well as φℓ is also a bounded generator of H 2 0 (σ) by Lemma 2. We choose an invariant null set N(ℓ) including N0 outside which a n(t)ℓ} converges to eiγGx(t)ℓ with 0 < γ < 2π. Define the subsequence {Gx invariant null set N1 by N1 = ∪∞ ℓ=1 N(ℓ) . When ℓ = m !, we take again a subsequence j (t)} converging to eiβ(m)Gx(t) with eiβ(m)ℓ = eiγ. Then the order of {Gx Z(β(m)) is larger than m, so ∪∞ m=1 Z(β(m)) is dense in T. Therefore, to represent g and all the limits of subsequences of {gn} on each orbit, the product group K/H × T is the smallest one. Let us explain the meaning more precisely. Under the assumption of Lemma 5, we put hα = χαg. Then hα lies in L2(τ ). Define the group character Pα of K/H × T by the projection Pα(¯x, eiθ) = eiθ. Since (hαPα)(St(¯x, eiθ)) = hα(¯x + et)C(¯x, t)eiθ = hα(¯x + et)eiαteiθ, the function t → (hαPα)(St(¯x, eiθ)) is an outer function in H ∞(dt/π(1 + t2)) for dτ × dθ/2π − a.e. (¯x, eiθ) in K/H × T. Then the outer function Gx(t) equals t → (hαPα)(St(¯x, eiθ)) for some θ with 0 ≤ θ < 2π. In order to represent consistently all kinds of limits of subsequences {Gx k(t)}, we require the family of all outer functions t → (hαPα)(St(¯x, eiθ)) with 0 ≤ θ < 2π. Lemma 7. Let Γ and Λ be as above. Then Λ cannot be equal to Γ. Proof. Let α be as in Lemma 5. Then α lies in Λ if and only if Λ = Γ. We suppose, on the contrary, that α lies in Λ. Since K/H = K, let us consider the skew product (K × T, {St}t∈R) of K and T induced by the cocycle C(x, t) = eiαt. We use freely the notation above. Since F (x, eiθ) = (χα Pα)(x, eiθ) , (x, eiθ) ∈ K × T , is an invariant function that is not constant, dσ × dθ/2π is not an ergodic measure on (K × T, {St}t∈R). Now K is represented as the local product decomposition Kα × [ 0, 2π/α), in which Kα is the closed subgroup of all x in K such that χα(x) = 1. If we put G(x, eiθ) = hα(x) Pα(x, eiθ) , (x, eiθ) ∈ K × T, 13 then, for each x = (y, s) in Kα × [0, 2π/α), the equation (8) G(St(x, eiθ)) = ei(θ+αt)hα(x + et) = ei(θ−αs)g(x + et) holds, since ei(θ+αt)χα(y + es + et) = ei(θ−αs) and hα = χαg. By regarding T as the interval [ 0, 2π/α), K × T is identified with Kα × [ 0, 2π/α) × [ 0, 2π/α). Let E be the subset of K × T defined by E = Kα × {(s, s) ; 0 ≤ s < 2π/α}. Then E is a closed invariant set in (K ×T, {St}t∈R), for which (K, {Tt}t∈R) is isomorphic to (E, {St}t∈R) via the map (y, s) → (y, s, s). We see also that the ergodic measure dσ is carried to (α/2π)dσ1 × ds on E by this map, where σ1 is the normalized Haar measure on Kα. We regard gn, g and hα as the functions on (K × T, {St}t∈R). Recall that almost every Gx n(t) and Gx(t) are outer functions in H ∞(dt/π(1 + t2)). Let x be in K \ N1 and let {gk} be a subsequence of {gn} such that Gx k(t) converges pointwise to t → eiαβeiαthα(x + et) with 0 ≤ β < 2π/α. Notice that t → eiαβeiαthα(x + et) is an outer function in H ∞(dt/π(1 + t2)) and that hα(x + et) = g(x + et). Let x = (y, s) in Kα × [ 0, 2π/α) as above. Since x may be replaced by any point in the orbit O(x) of x, we consider x as a function of s on [ 0, 2π/α). It follows from (8) that eiαβeiαthα(y + es + et) = eiα(β−s)G(St(y + es, eiαs)) , (s, t) ∈ [ 0, 2π/α) × R. Putting t = 0 and replacing y with y + e[sα/2π], if necessary, we observe that eiα(β−s)G(y + es, eiαs) = eiαβe−iαsGy(s) , s ∈ R. This shows that Gy k(s) converges pointwise to s → eiαβ(χα g)(y + es), which cannot be an outer function in H ∞(dt/π(1 + t2)). Hence any subsequence of {Gx n(t)} cannot converge to an outer function in H ∞(dt/π(1 + t2)) for σ − a.e. x in K. Thus we have a contradiction. (cid:3) In view of Lemma 7, we know that there are two possibilities in relation to α and Λ. Either nα lies in Λ only for n = 0 or ℓα lies in Λ for an integer ℓ ≥ 2. We claim that the latter case cannot occur, meaning that α is independent to Λ. Lemma 8. Let Λ, H and α be as above. Then nα lies in Λ if and only if n = 0 in Z. Consequently, H is isomorphic to T, so that K and dσ are identified with K/H × T and dτ × dθ/2π, respectively. Proof. Suppose that ℓα lies in Λ for some ℓ ≥ 2. By Lemma 2, φℓ is also a bounded generator of H 2 0 (σ). It follows from Lemma 6 that χℓα and (χαφ)ℓ lie in L2(τ ), so does φℓ itself. Let Γℓ and Λℓ be the smallest groups determined by the nonzero Fourier coefficients of φℓ and φℓ as above. Then they both are subgroups of Λ. On the other hand, since aλ(φ) = lim ǫ→+0 aλ(cid:0)(φℓ + ǫ)1/ℓ(cid:1) , 14 each λ in Λ with aλ(φ) 6= 0 lies in Λℓ. This implies that Λ = Λℓ = Γℓ. By replacing φ with φℓ in Lemma 7, this gives a contradiction. Thus nα lies in Λ if and only if n = 0. Since C(¯x, t)n is a coboundary only for n = 0, the measure dτ × dθ/2π is ergodic on (K/H × T, {St}t∈R). Define the isomorphism of Λ × Z onto Γ by (λ, n) = λ + nα , (λ, n) ∈ Λ × Z. Then the conjugate map ∗ of is given by ∗(x) = (¯x, eiθ) on K, where χα(x) = eiθ. Indeed, we observe that χλ(¯x)einθ = h(λ, n), (¯x, eiθ)i = χλ+nα(x) = χλ(¯x)χα(x)n , for each (λ, n) in Λ × Z. Via the map ∗ , K is identified with K/H × T, and dτ ×dθ/2π is carried by the map to dσ on K. (cid:3) We notice that the annihilator H of Λ is isomorphic to T, and g(x) as well as φ(x) is constant on almost every coset ¯x = x + H in K/H. 4. Contradiction to existence We may now offer our proof of the main result stated in Section 1. Proof of the Theorem. Suppose, on the contrary, that a bounded function φ generates H 2 0 (σ). Let Γ and Λ be the dense subgroups of R defined as in Section 3 with respect to φ and φ, respectively. Choose an α in Γ with aα(φ) 6= 0. It follows from Lemma 8 that α is independent of Λ and Γ is generated by α and Λ. Let 0 < β < 1. Since the function (1 + βχα)−1 = (−β)kχkα ∞Xk=0 lies in H ∞(σ), (1 + βχα)2 is an outer function in H ∞(σ). Define φ1 = (1 + βχα)2 φ . In view of Lemma 1, φ1 is also a bounded generator of H 2 0 (σ). As above, let Γ1 and Λ1 be the smallest groups determined by the nonzero Fourier coefficients of φ1 and φ1, respectively. Notice that Γ1 is a subgroup of Γ. We claim that the generator φ1 cannot satisfy the property of Lemma 7. Indeed, since φ1 = (1 + β2 + βχα + βχα) φ , we obtain by (1) that aλ(φ1) = (1 + β2)aλ(φ) + βaλ+α(φ) + βaλ−α(φ). Since α does not lie in Λ, if λ is in Λ, then aλ+α(φ) = aλ−α(φ) = 0. Then we have aλ(φ1) = (1 + β2)aλ(φ) and aλ+α(φ1) = βaλ(φ), for each λ in Λ. These facts imply that Λ1 contains Λ and α, so that Γ = Λ1 = Γ1, which contradicts Lemma 7. (cid:3) The next proof is of independent interest, because it suggests that our Theorem is regarded essentially as the converse to Corollary 1. 15 Proof of Corollary 1. We consider the case where the cocycle C(x, t) of M has the form C(x, t) = eiαt. Then M− is the space of all ψ in L2(σ) satisfying that ψ(x) ∼ XΓ∋λ > −α aλ(ψ)χλ(x) . Suppose that M− has a generator φ. Then log φ does not lie in L1(σ) and we may assume that φ is bounded. If ℓα is in Γ for a positive integer ℓ, then the bounded func- tion (χαφ)ℓ is a single generator of H 2 0 (σ) by Lemma 1, which is contrary to Theorem. We next consider the case that α ∈ R \ ∞[n=1 (1/n)Γ . Since C(x, t)n is a coboundary only for n = 0, the measure dσ × dθ/2π is ergodic on the skew product (K × T, {St}t∈R) induced by C(y, t), that is, St(x, eiθ) = (x + et , eiαteiθ), (x, eiθ) ∈ K × T. Let Γ1 be the discrete group generated by Γ and α, and let K1 be the dual group of Γ1. Since (λ, n) = λ + αn is an isomorphism of Γ × Z onto Γ1, the almost periodic flow on K1 is identified with (K × T, {St}t∈R). Then, via the dual map ∗ of , the normalized Haar measure dµ on K1 is identified with dσ × dθ/2π. Define the function φ1 in L2(µ) by φ1(x, eiθ) = φ(x)eiθ. Since log φ does not lie in L1(σ), neither does log φ1 in L1(µ). Since t → φ1 ◦ St(x, eiθ) is outer in H 2(dt/π(1 + t2)) for µ −a.e. (x, eiθ) in K × T, Lemma 1 implies that φ1 is a single generator of H 2 0 (µ), which contradicts our Theorem. (cid:3) Proof of Corollary 2. Denote by C(x, t) the real cocycle of M. Suppose that M− has a generator φ, for which log φ does not lie in L1(σ). It follows from Lemma 1 that almost every t → C(x, t)φ(x + et) is outer in H 2(dt/π(1 + t2)). We may assume that φ is bounded. Since C(x, t)2 ≡ 1, φ2 is a single generator of H 2 0 (σ) by Lemma 1, which contradicts our Theorem. (cid:3) By the same way as above, we may show that if C(x, t) takes only finite values, then M− cannot be singly generated. Indeed, by the cocycle identity, the set of values of C(x, t) forms a group of order k, Then if φ generates M−, then φk is a generator of H 2 0 (σ). Z(2π/k) = (cid:8)ei2πj/k ; j = 0, . . . , k − 1(cid:9) . Let M be the normalized simply invariant subspace of L2(σ) with cocycle A(x, t). Recall that ψ lies in M if and only if almost every t → A(x, t)ψ(x + et) lies in in [6, §3.2]). To prove Corollary 3. we need the following: H 2(dt/π(1+t2)). Denote byfM the invariant subspace with cocycle A(x, t) (as discussed Lemma 9. Let M and fM be as above. If M is singly generated, then (fM)− cannot be singly generated. 16 Proof. Since A(x, t) · A(x, t) ≡ 1, H 2 0 (σ) is the smallest subspace of L2(σ) containing all ψ1ψ2 with ψ1 in M∩L∞(σ) and ψ2 in (fM)−∩L∞(σ) (see [6, §3.2, Theorem 20]). Suppose that (fM)− is singly generated. Then Lemma 1 shows that there are bounded single generators φ1 and φ2 of M and (fM)−, respectively. Thus φ1φ2 is a single generator of 0 (σ), which contradicts our Theorem. H 2 (cid:3) Proof of Corollary 3. (a) Let M be a simply invariant subspace with nontrivial cocycle A(x, t). It follows from [11] that M is singly generated if and only if A(x, t) is coho- mologous to a singular cocycle. On the other hand, by [6, §4.6, Theorem 26], every cocycle is cohomologous to a Blaschke cocycle. By virtue of Lemma 9, we obtain easily a desired Blaschke cocycle. (b) From Lemma 9, we choose a Blaschke cocycle B(x, t) such that the invariant subspace N having the cocycle B(x, t) is not singly generated. We claim that B(x, t) satisfies the desired property. Suppose, on the contrary, that some function ψ in H 2(σ) has exactly the same zeros as B(x, t). By multiplying by a suitable outer function, we assume that ψ is bounded. Then ψ generates the invariant subspace with cocycle B(x, t)S(x, t), where S(x, t) is the singular cocycle determined by the inner part of t → B(x, t)ψ(x + et) in H 2(dt/π(1 + t2)). On the other hand, it follows from [11] and Lemma 1 that there is a function h in L2(σ) such that almost every t → S(x, t)h(x+et) is outer in H 2(dt/π(1 + t2)). Observe that (hψ)(x + et) = B(x, t) · S(x, t)h(x + et) · B(x, t)S(x, t)ψ(x + et) . Since the inner part of t → (hψ)(x + et) is t → B(x, t), the subspace N is singly generated by hψ, thus we have a contradiction. (cid:3) In the proof of (b) above, if the singular cocycle S(x, t) is a coboundary, then h is taken as a unitary function, otherwise log h does not lie in L1(σ). (a) It is sometimes useful to study the spectral measures associated with invariant subspaces. Let M be a simply invariant subspace of L2(σ) and put 5. Remarks Mλ = ^λ≥ν χν M. for each λ in R. Denote by Pλ the orthogonal projection of L2(σ) onto Mλ. By the property that ^−∞<λ<∞ Mλ = {0} and 17 _−∞<λ<∞ Mλ = L2(σ) , we obtain the contuinity of the spectral resolution of identity {I − Pλ}λ∈R on L2(σ), where I is the identity map on L2(σ). Let A(x, t) be the cocycle of M. By Stone's theorem, a unitary group {Vt}t∈R on L2(σ) is defined as Vt φ(x) = A(x, t)Tt φ(x) = −Z ∞ −∞ eiλt dPλφ(x) , φ ∈ L2(σ) , where Tt φ(x) = φ(x + et). For a nonzero function φ in L2(σ), −d(Pλφ, φ) is a finite positive measure on R. On almost periodic flows, by comparing with Lebesgue measure dλ, the type of such measures is uniquely determined. We then say that each of M, A(x, t) and {Vt}t∈R is of absolutely continuous, or singular continuous, or discrete type (as discussed in [6, §2.4]). This fact plays an important role to classify invariant subspaces in this special context. It is easy to observe that A(x, t) and A(x, t) have the same spectral type, so the following is an immediate consequence of Lemma 9. Proposition 2. There is a simply invariant subspace of L2(σ) of either absolutely continuous or singular continuous type which has no single generator. Let w be a nonnegative function in L2(σ) satisfying (3), while log w does not lie in L1(σ). We know that a cocycle is trivial if and only if it is of discrete type (see [6, §2.4, Theorem 15 ]). It follows from Corollary 1 that the type of M[w] has to be continuous. However, we have no idea to decide what kind of continuous spectrum M[w] may have. (b) Using a suitable cocycle, we may construct a skew product on which the H 2 0 −space is singly generated. Indeed, let w be a bounded function as above and let A(x, t) be the cocycle of M[w]. By Lemma 1 we see that almost every t → A(x, t)w(x + et) is outer in H 2(dt/π(1 + t2)). Denote by (K × T, {St}t∈R) the skew If A(x, t)n, n ≥ 1, is a coboundary q(x)q(x + et) with product induced by A(x, t). unitary function q on K, then qwn is a single generator of H 2 0 (σ). It then follows from Theorem that A(x, t)n is a coboundary only for n = 0. Hence dµ = dσ × dθ/2π is an ergodic measure on (K × T, {St}t∈R). If we set φ(x, eiθ) = w(x)eiθ , (x, eiθ) ∈ K × T , then φ is a single generator of H 2 0 (µ), since log φ does not lie in L1(µ) and almost every t → φ(St(x, eiθ)) is outer in H 2(dt/π(1 + t2)) (see [18] for another construction). (c) We have a bit of information on the distribution of zeros of functions in H 2(σ) which are connected with Dirichlet series (refer to [19] for related topics). Let {λn} be a sequence in Γ such that 0 ≤ λ1 < λ2 < · · · < λn −→ λ , n → ∞ , for some λ in Γ. Define a function ψ in H 2(σ) by with P∞ function. n=1 an2 < ∞. Observe that almost every t → ψ(x + et) extends to an entire ψ = ∞Xn=1 anχλn 18 Proposition 3. Let ψ be as above and let δ > 0. Then there is a decreasing sequence {mn} with mn → −∞ such that the number of zeros of z → ψ(x + ez) in the strip Sn = { z = t + iu ; mn > u > mn − δ} is infinite, for σ − a.e. x in K. Proof. Putting νn = λ − λn, we let φ = P∞ n=1 an χνn. Since z → eiλz has no zero, z → ψ(x + ez) has zero at z if and only if so does z → φ(x + ez) at ¯z. For each r > 0, t → φ ∗ Pir(x + et) cannot be an outer function in H 2(dt/π(1 + t2)), even if log φ does not lie in L1(σ). Since φ has no weight at infinity, the inner part of t → φ ∗ Pir(x + et) derives a Blaschke cocycle being not constant. From this fact, we may choose easily a desired decreasing sequence {mn}. (cid:3) [Note on September 9, 2014: Made some changes in wording] [Note on April 13, 2015: Accepted in ANNALES DE L'INSTITUTE FOURIER.] References [1] K. de Leeuw and I. Glicksberg, Quasi-invariance and analyticity of measures on compact groups, Acta Math. 109 (1963), 179 -- 205. [2] F. Forelli, Analytic and quasi-invariant measures, Acta Math. 118 (1967), 33 -- 59. [3] T. Gamelin, H p spaces and extremal functions in H 1, Trans. Amer. Math. Soc. 124 (1966), 158 -- 167. [4] T. Gamelin, Uniform algebras, Prentice-Hall, Englewood Cliffs, N.J., 1969. [5] H. Helson, Cocycles in harmonic analysis, Actes Cong`es Intern. Math. 2 (1970), 477 -- 481. [6] H. Helson, Analyticity on compact abelian groups, in Algebras in Analysis, Academic Press, Lon- don, 1975, 1 -- 62. [7] H. Helson, Compact groups with ordered duals VI, Bull. London Math. Soc. 8 (1976), 278 -- 281. [8] H. Helson, Compact groups with ordered duals VII, J. London Math. Soc. (2) 20 (1979), 509 -- 515. [9] H. Helson and D. Lowdenslager, Prediction theory and Fourier series of several complex variables II, Acta Math. 106 (1961), 175 -- 213. [10] H. Helson and W. Parry, Cocycles and spectra, Arkiv for Mat. 16 (1978), 195 -- 206. [11] H. Helson and J. Tanaka, Singular cocycles and the generator problem, Proc. Internat. 17th Conference on Operator Theory, Theta, 2000, 173 -- 186. [12] C. Moore and K. Schmidt, Coboundaries and homomorphisms for non singular actions and a problem of H. Helson, Proc. London Math. Soc. (3) 40 (1980), 443-475. [13] P. Muhly, Function algebras and flows, Acta Sci. Math. (Szeged) 35 (1973), 111 -- 121. [14] P. Muhly, Function algebras and flows III, Math. Z. 136 (1974), 253 -- 260. [15] K. Petersen, Ergodic Theory, Cambridge University Press, Cambridge, 1983. [16] W. Rudin, Fourier Analysis on Groups, Interscience, New York, 1962. [17] J. Tanaka, Blaschke cocycles and generators, Pacific J. Math. 142 (1990), 357 -- 378. [18] J. Tanaka, Single generator problem, Trans. Amer. Math. Soc. 348 (1996), 4113 -- 4129. [19] J. Tanaka, Dirichlet series induced by the Riemann zeta-function, Studia Math. 187 (2008), 157 -- 184 Department of Mathematics, School of Education, Waseda University, Shinjuku, Tokyo 169-8050, Japan E-mail address: [email protected] 19
1502.01078
2
1502
2016-02-03T08:07:32
Some applications of almost analytic extensions to operator bounds in trace ideals
[ "math.FA", "math-ph", "math-ph", "math.SP" ]
Using the Davies-Helffer-Sj\"ostrand functional calculus based on almost analytic extensions, we address the following problem: Given self-adjoint operators $S_j$, $j=1,2$, in $\mathcal{H}$, and functions $f$ in an appropriate class, for instance, $f \in C_0^{\infty}(\mathbb{R})$, how to control the norm $\|f(S_2) - f(S_1)\|_{\mathcal{B}(\mathcal{H})}$ in terms of the norm of the difference of resolvents, $\|(S_2 - z_0 I_{\mathcal{H}})^{-1} - (S_2 - z_0 I_{\mathcal{H}})^{-1}\|_{\mathcal{B}(\mathcal{H})}$, for some $z_0 \in \mathbb{C}\backslash\mathbb{R}$. We are particularly interested in the case where $\mathcal{B}(\mathcal{H})$ is replaced by a trace ideal, $\mathcal{B}_p(\mathcal{H})$, $p \in [1,\infty)$.
math.FA
math
SOME APPLICATIONS OF ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS IN TRACE IDEALS FRITZ GESZTESY AND ROGER NICHOLS Dedicated with deep admiration to Yurij Makarovich Berezansky on the occasion of his 90th birthday. Abstract. Using the Davies -- Helffer -- Sjostrand functional calculus based on almost analytic extensions, we address the following problem: Given self- adjoint operators Sj, j = 1, 2, in H, and functions f in an appropriate class, for instance, f ∈ C∞ 0 (R), how to control the norm kf (S2)−f (S1)kB(H) in terms of the norm of the difference of resolvents, (cid:13) (cid:13)B(H), for some z0 ∈ C\R. We are particularly interested in the case where B(H) is replaced by a trace ideal, Bp(H), p ∈ [1, ∞). (cid:13)(S2 −z0IH)−1 −(S1 −z0IH)−1 (cid:13) 1. Introduction Yurij M. Berezansky's contributions to analysis in general, and areas such as functional analysis, operator theory, spectral and inverse spectral theory, harmonic analysis, analysis in spaces of functions of an infinite number of variables, stochas- tic calculus, mathematical physics, quantum field theory, integration of nonlinear evolution equations, in particular, are legendary and of a lasting nature. The list of fields his ground breaking work changed in dramatic fashion can easily be continued in many directions as is demonstrated by the extraordinary breadth revealed in his highly influential monographs [2] -- [8]. Since operator theoretic methods frequently play a role in his research interests, we hope our modest contribution to operator bounds in trace ideals will create some joy for him. This paper has its origins in the following question: Given self-adjoint operators Sj, j = 1, 2, in H, and functions f in an appropriate class, for instance, f ∈ C∞ 0 (R), how to control the norm kf (S2) − f (S1)kB(H) in terms of the norm of the difference of resolvents,(cid:13)(cid:13)(S2 − z0IH)−1 − S1 − z0IH)−1(cid:13)(cid:13)B(H), for some z0 ∈ C\R? In particular, the question is just as natural with B(H) replaced by a trace ideal, Bp(H), p ∈ [1, ∞). In fact, our interest in these questions stems from computations of the Witten index (a suitable extension of the Fredholm index) for a class of non-Fredholm model operators D A = d dt + A, dom(D A) = W 1,2(R; H) ∩ dom(A), (1.1) Date: March 30, 2019. 2010 Mathematics Subject Classification. Primary: 47A30, 47A53, 47A60, 47B10; Secondary: 47B25, 30D99, 35Q40. Key words and phrases. Almost analytic extensions, trace ideals, operator bounds. R.N. gratefully acknowledges support from an AMS -- Simons Travel Grant. Appeared in Methods Funct. Anal. Topology, 21, 151 -- 169 (2015). 1 2 F. GESZTESY AND R. NICHOLS in the Hilbert space L2(R; H), where (Af )(t) = A(t)f (t) for a.e. t ∈ R, f ∈ dom(A) =(cid:26)g ∈ L2(R; H)(cid:12)(cid:12)(cid:12)(cid:12) g(t) ∈ dom(A(t)) for a.e. t ∈ R; H < ∞(cid:27), t 7→ A(t)g(t) is (weakly) measurable; R dt kA(t)g(t)k2 (1.2) with A(t), t ∈ R, a family of self-adjoint operators in H with asymptotes A± (in norm resolvent sense). Interesting concrete examles for A± are given by mass- less Dirac-type operators in H = L2(Rd), d ∈ N (the latter are known to be non-Fredholm), see, [10] -- [15]. More precisely, given the sequence of self-adjoint operators H j,n, H j in L2(R; dt; H), j = 1, 2, n ∈ N, and self-adjoint operators A+,n, A+, A− in H, n ∈ N, and a Pushnitski-type relation between the spectral shift functions ξ( · ; H 2,n, H 1,n) and ξ( · ; A+,n, A−) for the pairs, (H 2,n, H 1,n) and (A+,n, A−) of the form ξ(λ; H 2,n, H 1,n) =( 1 π ´ λ1/2 −λ1/2 0, λ < 0, ξ(ν;A+,n,A−) dν (λ−ν 2)1/2 for a.e. λ > 0, n ∈ N, (1.3) we were interested in performing the limit n → ∞ in (1.3) to obtain the analogous relation for the limiting spectral shift functions ξ( · ; H 2, H 1) and ξ( · ; A+, A−) cor- responding to the limiting pairs (H 2, H 1) and (A+, A−), respectively. The latter is instrumental in computing the Witten index for DA. The task of performing the limit n → ∞ in (1.3) is considerably complicated since due to the nature of the ap- proximations involved, no suitable bounds on ξ( · ; H 2,n, H 1,n) and ξ( · ; A+,n, A−) (independent of n ∈ N) are readily available. To circumvent this difficulty one can resort to a distributional approach considering R dλ ξ(λ; H 2,n, H 1,n)f ′(λ) = 1 π R dν ξ(ν; A+,n, A−)F ′(ν), n ∈ N, (1.4) where f ∈ C∞ 0 (R) is arbitrary, and F ′ ∈ C∞ 0 (R) is given by F ′(ν) = ∞ ν 2 dλ f ′(λ)(λ − ν2)−1/2, ν ∈ R. (1.5) Focusing now on the left-hand side of (1.4), one recalls Krein's trace formula, dλ ξ(λ; H 2,n, H 1,n)f ′(λ), f ∈ C∞ 0 (R). lim n→∞ k[f (H2,n) − f (H 1,n)] − [f (H 2) − f (H 1)]kB1(L2(R;H)) = 0, and hence obtain lim n→∞[0,∞) dλ ξ(λ; H 2,n, H 1,n)f ′(λ) = lim n→∞ trB1(L2(R;H))(f (H 2,n) − f (H 1,n)) Thus, given control of resolvents in the form trB1(L2(R;H))(f (H 2,n) − f (H 1,n)) = [0,∞) n→∞(cid:13)(cid:13)(cid:2)(H 2,n − z I)−m2 − (H 1,n − z I)−m2(cid:3) lim − [(H 2 − z I)−m2 − (H 1 − z I)−m2(cid:3)(cid:13)(cid:13)B1(L2(R;H)) = 0, one can hope to control z ∈ C\R. (1.6) (1.7) (1.8) ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 3 = trB1(L2(R;H))(f (H 2) − f (H 1)) = [0,∞) dλ ξ(λ; H 2, H 1)f ′(λ), f ∈ C∞ 0 (R). (1.9) Together with controlling the limit n → ∞ on the right-hand side of (1.4), this leads to dλ ξ(λ; H 2, H 1)f ′(λ) = dν ξ(ν; A+, A−)F ′(ν). (1.10) R 1 π R Without going into further details we note that (1.10) in turn can be used to prove the limiting relation in (1.3) and the latter leads to a compution of the semigroup regularized Witten index, Ws(DA), of DA: Assuming that 0 is a left and a right Lebesgue point of ξ( · ; A+, A−), denoting the corresponding values by ξ(0±; A+, A−), the semigroup regularized Witten index is found to be Ws(DA) := lim t↑∞ trL2(R;H)(cid:0)e−tH 1 − e−tH 2(cid:1) = [ξ(0−; A+, A−) + ξ(0+; A+, A−)]/2, (1.11) see, for instance, [10] -- [15]. (We here use the semigroup regularized Witten index rather than the resolvent regularised one as the former is applicable in the case of d- dimensional Dirac-type operators A±, d ∈ N.) We trust this sufficiently illustrates our interest in using control of resolvents of self-adjoint operators to gain control over their C∞ 0 -functions. We also note a further complication lies in the fact that when studying multi- dimensional Dirac-type operators A±, resolvents alone are not sufficient in the trace class context and hence sufficiently high powers (depending on the space dimension involved) of resolvents have to be employed. Our principal tool to gain control over C∞ 0 -functions of S in terms of (powers of) resolvents of S is furnished by a suitable application of almost analytic extensions efℓ,σ of f in the form of a Davies -- Helffer -- Sjostrand functional calculus [18], [20, Ch. 2], [40, Proposition 7.2], of the form (z)(S − zIH)−1, (1.12) f (S) = π−1C dxdy ∂efℓ,σ ∂z and a refinement due to Khochman [43] of the type f (S) = 1 π C dxdy ∂efℓ,σ ∂z (z)(z − z0)m(S − z0IH)−m(S − zIH)−1, m ∈ N ∪ {0}. (1.13) to be discussed in some detail in Section 2. Section 3 contains our principal results and some applications. Finally, Appendix A recalls various useful facts concerning (powers of) resolvents. We conclude with some comments on the notation employed in this paper: Let H be a separable complex Hilbert space, (·, ·)H the scalar product in H (linear in the second argument), and IH the identity operator in H. Next, if T is a linear operator mapping (a subspace of) a Hilbert space into another, then dom(T ) and ker(T ) denote the domain and kernel (i.e., null space) of T . The spectrum and resolvent set of a closed linear operator in a Hilbert space will be denoted by σ(·) and ρ(·), respectively. t The convergence of bounded operators in the strong operator topology (i.e., pointwise limits) will be denoted by s-lim. 4 F. GESZTESY AND R. NICHOLS The Banach spaces of bounded and compact linear operators on a separable complex Hilbert space H are denoted by B(H) and B∞(H), respectively; the cor- responding ℓp-based trace ideals will be denoted by Bp(H), their norms are abbre- viated by k · kBp(H), p > 1. Moreover, trH(A) denotes the corresponding trace of a trace class operator A ∈ B1(H). The symbol C∞ 0 (R) represents C∞-functions of compact support on R; contin- uous functions on R vanishing at infinity are denoted by C∞(R). 2. Basic Facts on Almost Analytic Extensions And The Functional Calculus For Self-Adjoint Operators In this preparatory section we briefly recall the basics of almost analytic ex- tensions and the ensuing functional calculus for self-adjoint operators, following Davies' detailed treatment in [18], [20, Ch. 2]. One introduces the class Sβ(R), β ∈ R, consisting of all functions f ∈ C∞(R) such that f (m)(x) = x→∞ O((cid:10)xiβ−m(cid:1), m ∈ N0, where hzi = (cid:0)z2 + 1(cid:1)1/2 pointwise multiplication, Sβ(R) · Sγ(R) ⊆ Sβ+γ(R), β, γ ∈ R, and the space , z ∈ C. Then in obvious notation, with "·" denoting is an algebra under pointwise multiplication with A(R) = [β<0 Sβ(R) C∞ 0 (R) ⊂ A(R). (2.1) (2.2) (2.3) In particular, f ∈ A(R) implies f ∈ C∞(R) (the continuous functions vanishing at ±∞) and f (m) ∈ L1(R), m ∈ N. Given f ∈ A(R), one defines an almost analytic extension efℓ,σ, of f to C by f (k)(x)(iy)k z = x + iy ∈ C, ℓ ∈ N, , (2.4) efℓ,σ(z) = σ(x, y) ℓXk=0 k! where σ(x, y) = τ (y/hxi), x, y ∈ R, τ ∈ C∞ 0 (R), τ (s) =(1, 0, s 6 1, s > 2. (2.5) possible (cf., [18]). We note the formula The precise structure of efℓ,σ will not be important and other expresssions for it are ∂efℓ,σ 2(cid:18) ∂efℓ,σ (z) + i ∂efℓ,σ ∂y (z) = ∂x ∂z 1 = [σx(x, y) + σy(x, y)] 1 2 f (k)(x)(iy)k f ℓ+1(x)(iy)ℓ + σ(x, y) 1 2 ℓ! , (2.6) implying the crucial fact, ∂efℓ,σ ∂z (cid:12)(cid:12)(cid:12)(cid:12) z ∈ C, (2.7) O(cid:0)yℓ(cid:1), k! (z)(cid:19) ℓXk=0 (x + iy)(cid:12)(cid:12)(cid:12)(cid:12) = y↓0 ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 5 in particular, ∂efℓ,σ ∂z (x) = 0, x ∈ R. (2.8) Following Helffer and Sjostrand [40, Proposition 7.2], particularly, in the form presented by Davies [18], [20, Ch. 2], one then establishes a functional calculus for self-adjoint operators S in a complex, separable Hilbert space H via the formula f (S) = π−1C dxdy ∂efℓ,σ ∂z (z)(S − zIH)−1. (2.9) Since the integrand is norm continuous, the integral in (2.9) is norm convergent, in particular, one notes that (2.7) and (2.8), together with the standard estimate 6 Im(z)−1, overcome the apparent singularity of the integrand (cid:13)(cid:13)(S −zIH)−1(cid:13)(cid:13)B(H) in (2.9) for z ∈ σ(S) ⊆ R (cf. also (2.8)). The justification for calling this a functional calculus follows upon proving the following facts: • The left-hand side of (2.8) is independent of the choice of ℓ ∈ N and the precise form of σ in (2.5). 0 (R) with supp(f ) ∩ σ(S) = ∅, then f (S) = 0. • If f ∈ C∞ • If f, g ∈ A(R), then (f g)(S) = f (S)g(S), f (S)∗ = f (S), kf (S)kB(H) = kf kL∞(R). • Let z ∈ C\R and fz(x) = (x − z)−1, then fz ∈ A(R) and fz(S) = (S − zIH)−1. In addition, we note that Khochman [43] proved the following extension of (2.9): Lemma 2.1 ([43]). Let m ∈ N, f ∈ C∞ H. Then, 0 (R), and suppose that S is self-adjoint in f (S) = 1 π C dxdy ∂efℓ,σ ∂z (z)(z − z0)m(S − z0IH)−m(S − zIH)−1. (2.10) We will employ (in fact, rederive) (2.10) in the proof of Theorem 3.8. Next, we discuss another extension focusing on semigroups rather than powers of resolvents. Lemma 2.2. Let t > 0, f ∈ C∞ from below in H. Then, 0 (R), and suppose that S is self-adjoint and bounded f (S) = 1 π C dxdy ∂efℓ,σ ∂z (z)etze−tS(S − zIH)−1. (2.11) Proof. We start by noting that if f, g ∈ C∞ 0 (R), it is proved in [20, p. 28] that C dxdy ∂g(f g)ℓ′′,σ′′ ∂z (z)(S − zIH)−1 = C dxdy ∂z 0 (R) and let Et ∈ C∞ ∂( fℓ′,σ′ gℓ,σ) Next, suppose that f ∈ C∞ coincides with etx on an open interval I with supp(f ) ⊂ I. Then (z)(S − zIH)−1. (2.12) 0 (R) denote a function which eEt,ℓ,σ(z) = σ(x, y)etx (ity)k k! ℓXk=0 , z = x + iy, x ∈ I. (2.13) almost analytic extension of g, in light of the identity f (S) = g(S)e−tS, one infers Let fℓ′,σ′ denote an almost analytic extension of f . Setting g = f Et, withegℓ′′,σ′′ an 6 F. GESZTESY AND R. NICHOLS from the Davies -- Helffer -- Sjostrand functional calculus (2.9) applied to g, f (S) = = = = 1 1 1 1 ∂z ∂z ∂z ∂z dxdy dxdy dxdy (z)e−tS(S − zIH)−1 (z)e−tS(S − zIH)−1 (z)e−tS(S − zIH)−1 π C π C π C π C π C ∂egℓ′′,σ′′ ∂(efℓ′,σ′eEt,ℓ,σ) ∂efℓ′,σ′ (z)eEt,ℓ,σ(z)e−tS(S − zIH)−1 ∂eEt,ℓ,σ dxdy efℓ′,σ′ (z) (z)σ(x, y)etx(cid:18) ℓXk=0 ∂efℓ′′,σ′′ dxdy efℓ′,σ′ (z)( 1 2(cid:20)σx(x, y)etx(cid:18) ℓXk=0 k! (cid:19) + σ(x, y)etx(cid:18) ℓXk=0 k! (cid:19)(cid:21) k! (cid:19) + σy(x, y)etx(cid:18) ℓXk=0 2(cid:20)σ(x, y)etxit(cid:18) ℓ−1Xk=0 π C (ity)k (ity)k (ity)k (ity)k dxdy ∂z 1 i + 1 + + k! (cid:19)e−tS(S − zIH)−1 (ity)k k! (cid:19)(cid:21)) × e−tS(S − zIH)−1. (2.14) Exploiting the fact that f (S) is independent of ℓ, we now take the limit ℓ → ∞ in (2.14). Since efℓ′,σ′ has compact support and takes care of the singularity of the resolvent, e−tS is bounded and z-independent, and the exponential series converges uniformly on compact sets, one may pass the limit under the integral to obtain dxdy f (S) = = = + 1 1 π C π C π C π C 1 1 dxdy dxdy ∂ ∂ ∂z ∂z e−tS(S − zIH)−1 ∂(σ(x, y)etz) σ(x, y)etze−tS(S − zIH)−1 ∂efℓ′,σ′ dxdy efℓ′,σ′ (z) ∂z(cid:0)efℓ′,σ′ (z)σ(x, y)etz(cid:1)e−tS(S − zIH)−1 ∂z(cid:0)efℓ′,bσ(z)etz(cid:1)e−tS(S − zIH)−1, ∂z(cid:0)efℓ′,bσ(z)etz(cid:1) = ∂efℓ′,bσ π C etze−tS(S − zIH)−1 ∂efℓ,bσ dxdy etz, ∂z ∂z ∂ 1 which then shows f (S) = and hence (2.11) (renaming ℓ′ andbσ). verify that where bσ = σ′σ (which corresponds to choosing bτ = τ ′τ ). It is a simple matter to (2.15) (2.16) (2.17) (cid:3) ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 7 Historically, the idea of almost analytic (resp., pseudo-analytic) extensions ap- peared in Hormander [39] and Dynkin [28], [29], Melin and Sjostrand [46] (see also [26, Ch. 8], [45, Sect. III.6] for expositions and [42] for an alternative approach). The functional calculus was used by Helffer and Sjostrand in their seminal 1989 paper on Schrodinger operators with magnetic fields [40], which in turn was the basis for the systematic treatment by Davies [18], [20, Ch. 2]. Since these early developments, there has been a large body of literature in connection with spectral theory for Schrodinger and Dirac-type operators applying this functional calculus. While a complete list of references in this context is clearly beyond the scope of this paper, we want to illustrate the great variety of applications that rely on this functional calculus a bit and hence refer to [16], [21], [22], [23], [24], [25], [27], [30], [31], [34], [35], [37], [41], [43], [47], [48], [52], [53], and the references cited therein. While we here exclusively focus on linear operators in a Hilbert space, this func- tional calculus applies to operators in Banach spaces with real spectrum, see, for instance, [1], [17], [18], [19], [32], [33]. Extensions to the case where the spectrum is contained in the unit circle or contained in finitely-many smooth arcs were also treated in [28]. 3. Some Applications In this section we apply the almost analytic extension method and its ensuing functional calculus for self-adjoint operators to derive various norm bounds and convergence properties of operators in trace ideals. We start with the following estimates established in the proof of [20, Theo- rem 2.6.2] (more precisely, (3.1) is proved in [20], but then the rest of Lemma 3.1 is obvious): Lemma 3.1. Let z0 ∈ C\R, f ∈ A(R) and suppose that Sj, j = 1, 2, are self- adjoint in H. Then kf (S2) − f (S1)kB(H) 6 8 ∂ef2,σ (cid:2)z02 + z2(cid:3) dxdy(cid:12)(cid:12)(cid:12)(cid:12) π C ×(cid:13)(cid:13)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:13)(cid:13)B(H). (z)(cid:12)(cid:12)(cid:12)(cid:12) Im(z)2 ∂z In addition, if for some p ∈ [1, ∞), (cid:2)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:3) ∈ Bp(H) for some (and hence for all ) z0 ∈ C\R, then [f (S2) − f (S1)] ∈ Bp(H) (3.1) (3.2) (3.3) 8 ∂z ∂ef2,σ (cid:2)z02 + z2(cid:3) and kf (S2) − f (S1)kBp(H) 6 If(cid:2)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:3) ∈ B∞(H), the inclusion (3.2) extends to p = ∞. f (S2) − f (S1) = π−1 C dxdy(cid:12)(cid:12)(cid:12)(cid:12) π C ×(cid:13)(cid:13)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:13)(cid:13)Bp(H). ∂ef2,σ (z)(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3) Proof. Combining (2.9), (A.1), and (A.2) one obtains, (z)(cid:12)(cid:12)(cid:12)(cid:12) Im(z)2 dxdy (3.4) ∂z and hence kf (S2) − f (S1)kB(H) 8 F. GESZTESY AND R. NICHOLS ∂z ∂z ∂ef2,σ ∂ef2,σ ∂ef2,σ ∂ef2,σ dxdy(cid:12)(cid:12)(cid:12)(cid:12) dxdy(cid:12)(cid:12)(cid:12)(cid:12) dxdy(cid:12)(cid:12)(cid:12)(cid:12) dxdy(cid:12)(cid:12)(cid:12)(cid:12) 6 π−1C 6 π−1C ×(cid:2)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:3)(S1 − z0IH)(S1 − zIH)−1(cid:13)(cid:13)B(H) 6 π−1C ×(cid:13)(cid:13)(S1 − z0IH)(S1 − zIH)−1(cid:13)(cid:13)B(H)(cid:13)(cid:13)(cid:2)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:3)(cid:13)(cid:13)B(H) 6(cid:18) 8 π C ×(cid:13)(cid:13)(cid:2)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:3)(cid:13)(cid:13)B(H). (z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(S2 − zIH)−1 − (S1 − zIH)−1(cid:13)(cid:13)B(H) (z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(S2 − z0IH)(S2 − zIH)−1 (z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(S2 − z0IH)(S2 − zIH)−1(cid:13)(cid:13)B(H) (cid:2)z02 + z2(cid:3) (z)(cid:12)(cid:12)(cid:12)(cid:12) Im(z)2 (cid:19) Precisely the same chain of estimates applies to B(H) replaced by Bp(H), relying on the ideal properties of Bp(H), p ∈ [1, ∞). The case p = ∞ in (3.2) is a consequence of the norm convergent integral on the right-hand side of (3.4). (cid:3) (3.5) ∂z ∂z Combined with a Stone -- Weierstrass approximation argument and the fact that kf (S)kB(H) = kf kL∞(R), f ∈ A(R), Lemma 3.1 yields the following well-known fact, recorded, for instance, in [20, Theorem 2.62], [50, Theorem V.III.20(a)]: Lemma 3.2. Let Sn, n ∈ N, and S be self-adjoint in H, and suppose that Sn converges to S in norm resolvent sense as n → ∞. Then, lim n→∞ kf (Sn) − f (S)kB(H) = 0 (3.6) for all f ∈ C∞(R). Remark 3.3. We note that the functional calculus based on almost analytic ex- tensions is not the only possible approach to address estimates such as (3.1) and (3.3). As a powerful alternative we mention the theory of double operator integrals (DOI), which can prove stronger inequalities of the following type (cf. [9], [12], [56]): Given m ∈ N odd and p ∈ [1, ∞), there exist constants a1, a2 ∈ R\{0} and C = C(f, m, a1, a2) ∈ (0, ∞) such that (cid:13)(cid:13)f (A) − f (B)kBp(H) 6 C(cid:0)(cid:13)(cid:13)(A − a1iIH)−m − (B − a1iIH)−m(cid:13)(cid:13)Bp(H) +(cid:13)(cid:13)(A − a2iIH)−m − (B − a2iIH)−m(cid:13)(cid:13)Bp(H)(cid:1), which permits the use of differences of higher powers m ∈ N of resolvents to control the k · kBp(H)-norm of the left-hand side [f (A) − f (B)] for f ∈ C∞ 0 (R). In fact, this extends to a much larger class of functions f , see [13] for details. f ∈ C∞ 0 (R), (3.7) Moreover, repeatedly differentiating (z)(λ − z)−1, λ ∈ R, (3.8) with respect to λ yields dxdy f (λ) = π−1C ∂efℓ,σ f (m−1)(S) = π−1(−1)m−1(m − 1)!C ∂z This leads to estimates of the type (3.7) but with f replaced by f (m−1). ⋄ dxdy (z)(S − z)−m, λ ∈ R. (3.9) ∂efℓ,σ ∂z ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 9 We recall a useful result: Lemma 3.4. Let p ∈ [1, ∞) and assume that R, Rn, T, Tn ∈ B(H), n ∈ N, satisfy s-limn→∞ Rn = R and s-limn→∞ Tn = T and that S, Sn ∈ Bp(H), n ∈ N, satisfy limn→∞ kSn − SkBp(H) = 0. Then limn→∞ kRnSnT ∗ n − RST ∗kBp(H) = 0. This follows, for instance, from [38, Theorem 1], [51, p. 28 -- 29], or [55, Lemma 6.1.3] with a minor additional effort (taking adjoints, etc.). Next, we describe a typical convergence result: Theorem 3.5. Let Sj,n, n ∈ N, and Sj, j = 1, 2, be self-adjoint in H, and assume that Sj,n converges in strong resolvent sense as n → ∞ to Sj, j = 1, 2, respectively. (i) Suppose that for some (and hence for all ) z0 ∈ C\R, lim n→∞(cid:13)(cid:13)(cid:2)(S2,n − z0IH)−1 − (S2 − z0IH)−1(cid:3) −(cid:2)(S1,n − z0IH)−1 − (S1 − z0IH)−1(cid:3)(cid:13)(cid:13)B(H) = 0. k[f (S2,n) − f (S1,n)] − [f (S2) − f (S1)]kB(H) = 0, f ∈ C∞ 0 (R). (3.10) (3.11) Then, lim n→∞ (ii) Let p ∈ [1, ∞) and suppose that for some (and hence for all ) z0 ∈ C\R, (cid:2)(S2,n − z0IH)−1 − (S1,n − z0IH)−1(cid:3),(cid:2)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:3) ∈ Bp(H), n ∈ N, (3.12) and Then, lim n→∞ lim n→∞(cid:13)(cid:13)(cid:2)(S2,n − z0IH)−1 − (S2 − z0IH)−1(cid:3) −(cid:2)(S1,n − z0IH)−1 − (S1 − z0IH)−1(cid:3)(cid:13)(cid:13)Bp(H) = 0. k[f (S2,n) − f (S1,n)] − [f (S2) − f (S1)]kBp(H) = 0, f ∈ C∞ 0 (R). (3.13) (3.14) Proof. As usual, a combination of identity (A.1), Lemma 3.4, and the assumed strong resolvent convergence of Sj,n to Sj as n → ∞, j = 1, 2, proves sufficiency of the conditions (3.10) and (3.12) for just one z0 ∈ C\R. Thus, assumption (3.10) actually implies lim n→∞(cid:13)(cid:13)(cid:2)(S2,n − zIH)−1 − (S2 − zIH)−1(cid:3) −(cid:2)(S1,n − zIH)−1 − (S1 − zIH)−1(cid:3)(cid:13)(cid:13)B(H) = 0, Next, mimicking (3.4), one obtains, [f (S2,n) − f (S1,n)] − [f (S2) − f (S1)] z ∈ C\R. (3.15) = π−1C dxdy ∂ef2,σ ∂z and hence, 6 π−1 lim n→∞C (z)h(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1(cid:3) −(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3)i, (z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:13)(cid:13)(cid:13)(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1(cid:3) ∂ef2,σ ∂z dxdy(cid:12)(cid:12)(cid:12)(cid:12) (3.16) f ∈ C∞ 0 (R), (3.17) lim n→∞ k[f (S2,n) − f (S1,n)] − [f (S2) − f (S1)]kB(H) 10 F. GESZTESY AND R. NICHOLS −(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3)(cid:13)(cid:13)(cid:13)B(H) 0 (R) implies ef2,σ ∈ C∞ , In this context one observes that f ∈ C∞ 0 (C). Since the exceptional set R in (3.15) has dxdy-measure zero (in addition to the fact that by (2.8), ∂ef2,σ/∂z vanishes on R), an application of the Lebesgue dominated conver- gence theorem to interchange the limit n → ∞ with the integral on the right-hand side of (3.17) requires establishing an n-independent integrable majorant of the integrand in (3.17). Employing identity (A.1) and estimate (A.2), this majorant can be obtained as follows: f ∈ C∞ 0 (R). (3.18) (3.19) (cid:13)(cid:13)(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1(cid:3) −(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3)(cid:13)(cid:13)B(H) 6(cid:13)(cid:13)(S2,n − zIH)−1 − (S1,n − zIH)−1(cid:13)(cid:13)B(H) +(cid:13)(cid:13)(S2 − zIH)−1 − (S1 − zIH)−1(cid:13)(cid:13)B(H) 6(cid:13)(cid:13)(S2,n − z0IH)(S2,n − zIH)−1(cid:13)(cid:13)B(H) ×(cid:13)(cid:13)(S2,n − z0IH)−1 − (S1,n − z0IH)−1(cid:13)(cid:13)B(H) ×(cid:13)(cid:13)(S1,n − z0IH)(S1,n − zIH)−1(cid:13)(cid:13)B(H) +(cid:13)(cid:13)(S2 − z0IH)(S2 − zIH)−1(cid:13)(cid:13)B(H) ×(cid:13)(cid:13)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:13)(cid:13)B(H) ×(cid:13)(cid:13)(S1 − z0IH)(S1 − zIH)−1(cid:13)(cid:13)B(H) 6 8(cid:2)z02 + z2(cid:3)Im(z)−2h(cid:13)(cid:13)(S2,n − z0IH)−1 − (S1,n − z0IH)−1(cid:13)(cid:13)B(H) +(cid:13)(cid:13)(S2 − z0IH)−1 − (S1 − z0IH)−1(cid:13)(cid:13)B(H)i 6 8(cid:2)z02 + z2(cid:3)Im(z)−2C(z0), for some 0 < C(z0) < ∞, independent of n ∈ N, since by assumption (3.10), (cid:13)(cid:13)(S2,n − z0IH)−1 − (S1,n − z0IH)−1(cid:13)(cid:13)B(H) 6 eC(z0) for some 0 < eC(z0) < ∞, independent of n ∈ N. Together with the properties (2.7), (2.8) of ∂ef2,σ/∂z, this establishes the sought integrable majorant, independent of n ∈ N, and thus permits the interchange of the limit n → ∞ with the integral on the right-hand side of (3.17). This completes the proof of (3.11). The proof of (3.14) proceeds exactly along the same lines employing once more (cid:3) the ideal properties of Bp(H), p ∈ [1, ∞). We remark in passing that double operator integral techniques permit one to enlarge the class of functions f to which Theorem 3.5 applies (cf. also Remark 3.9). Remark 3.6. The proof to Theorem 3.5 uses dominated convergence and given (3.18), the crucial observation is that which is obvious if f ∈ C∞ it is possible to once again prove (3.20) if f ∈ Sβ(R) for some β < −1. Indeed, z02 + z2 Im(z)2 < ∞, dxdy(cid:12)(cid:12)(cid:12)(cid:12) C 0 (R), since then efℓ,σ is compactly supported. However, (z)(cid:12)(cid:12)(cid:12)(cid:12) ∂efℓ,σ ∂z (3.20) ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 11 following [20, p. 25], and setting U = {(x, y) hxi < y < 2hxi}, V = {(x, y) 0 < y < 2hxi}, (3.21) one infers σx + iσy 6 chxi−1χU (x, y), for some constant c > 0. Then for f ∈ Sβ(R), z = x + iy ∈ C, (3.22) hxiβ−k−1ykχU (x, y)) + Chxiβ−ℓ−1yℓχV (x, y), where C > 0 is an appropriate constant. Therefore, z = x + iy ∈ C, z02 + x2 + y2 (3.23) ) ∂z (z)(cid:12)(cid:12)(cid:12)(cid:12) 6 C( ℓXk=0 ∂efℓ,σ (z)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) ∂efℓ,σ z02 + z2 Im(z)2 ∂z y2 y2 z02 + 5hxi2 hxiβ−1χU (x, y) z02 + x2 + y2 hxiβ−k−1ykχU (x, y) + Chxiβ−ℓ−1yℓχV (x, y) 6 C( ℓXk=0 ) 6 C2ℓ( ℓXk=0 + Chxiβ−ℓ−1yℓ−2χV (x, y)(cid:8)z02 + 5hxi2(cid:9) 6 C2ℓ( ℓXk=0 hxiβ−3χU (x, y)(cid:8)z02 + 5hxi2(cid:9)) + C2ℓ−2hxiβ−3χV (x, y)(cid:8)z02 + 5hxi2(cid:9) 6 bChxiβ−1(cid:8)χU (x, y) + χV (x, y)(cid:9), hxi2 z = x + iy ∈ C\R, (3.24) where bC = bC(z0) > 0 is a constant. Here, we used z02 + 5hxi2 < eChxi2, for an appropriate constant eC = eC(z0) > 0. Given (3.24), (3.20) holds if β < −1 since C −∞ dxdy hxiβ−1(cid:8)χU (x, y) + χV (x, y)(cid:9) = ∞ = 2 ∞ = 6 ∞ dx hxiβ−1 ∞ dx hxiβ−1(cid:26) 2hxi dx hxiβ < ∞. −∞ −∞ 0 −∞ dy(cid:8)χU (x, y) + χV (x, y)(cid:9) dy + 2hxi hxi dy(cid:27) (3.25) Thus, the majorant (3.24) is integrable as long as f ∈ Sβ(R) for some β < −1 and hence Theorem 3.5 extends from C∞ 0 (R) to Sβ(R), β < −1. Next, we briefly apply Theorem 3.5 to a concrete (1 + 1)-dimensional example treated in great detail in [10] and [11] by alternative methods. Example 3.7. Assuming the real-valued functions φ, θ satisfy φ ∈ ACloc(R) ∩ L∞(R) ∩ L1(R), φ′ ∈ L∞(R), (3.26) 12 F. GESZTESY AND R. NICHOLS θ ∈ ACloc(R) ∩ L∞(R), θ′ ∈ L∞(R) ∩ L1(R), lim t→∞ θ(t) = 1, θ(t) = 0, t→−∞ lim (3.27) we introduce the family of self-adjoint operators A(t), t ∈ R, in L2(R), A(t) = −i d dx + θ(t)φ, dom(A(t)) = W 1,2(R), t ∈ R, (3.28) and its self-adjoint asymptotes as t → ±∞, A+ = −i + φ, A− = −i , dom(A±) = W 1,2(R). (3.29) d dx d dx In addition, we introducing the operator d/dt in L2(cid:0)R; dt; L2(R; dx)(cid:1) by (cid:18) d f(cid:19)(t) = f ′(t) for a.e. t ∈ R, f ∈ dom(d/dt) =(cid:8)g ∈ L2(cid:0)R; dt; L2(R)(cid:1)(cid:12)(cid:12) g ∈ ACloc(cid:0)R; L2(R)(cid:1), g′ ∈ L2(cid:0)R; dt; L2(R)(cid:1)(cid:9) dt = W 1,2(cid:0)R; dt; L2(R; dx)(cid:1). Next, we agree to identify L2(cid:0)R; dt; L2(R; dx)(cid:1) with L2(R2; dtdx) (denoting the lat- ter by L2(R2) for brevity ) and introduce DA in L2(R2) by D A = d dt + A, dom(D A) = W 1,2(R2), (3.32) with A defined as in (1.2) and A(t), t ∈ R, given by (3.28). Moreover, we introduce the nonnegative, self-adjoint operators H j, j = 1, 2, in L2(R2) by H 1 = D∗ ADA, H 2 = DAD∗ A. (3.33) (3.30) (3.31) As shown in [10], the assumptions on φ and θ guarantee that (cid:2)(A+ − zI)−1 − (A− − zI)−1(cid:3) ∈ B1(cid:0)L2(R)(cid:1), (for simplicity, we adopt the abbreviation I = IL2(R) throughout this example ), and thus, the spectral shift function ξ( · ; A+, A−) for the pair (A+, A−) exists and is well-defined up to an arbitrary additive real constant, satisfying z ∈ C\R (3.34) ξ( · ; A+, A−) ∈ L1(cid:0)R; (ν2 + 1)−1dν(cid:1). − +n2I)−1/2 and A+,n = A− +χn(A−)φχn(A−), n ∈ N, (3.35) Introducing χn(A−) = n(A2 the fact A+,n − A− = χn(A−)φχn(A−) ∈ B1(cid:0)L2(R)(cid:1), n ∈ N, implies that also the spectral shift functions ξ( · ; A+,n, A−), n ∈ N, exist and are uniquely determined by (3.36) ξ( · ; A+,n, A−) ∈ L1(R; dν), n ∈ N. (3.37) In fact, as has been shown in [10], the open constant in ξ( · ; A+, A−) can naturally be determined via the limiting procedure lim n→∞ ξ(ν; A+,n, A−) = 1 2π R dx φ(x) = ξ(ν; A+, A−), ν ∈ R. (3.38) In particular, ξ( · ; A+, A−) turns out to be constant in this example. ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 13 Replacing A(t) by An(t) = A− + χn(A−)θ(t)φχn(A−), n ∈ N, t ∈ R, and hence, A by An, DA by DA , H j by H j,n, j = 1, 2, n ∈ N, one verifies the facts, n z ∈ C\[0, ∞), (cid:2)(H 2 − z I)−1 − (H 1 − z I)−1(cid:3) ∈ B1(cid:0)L2(R2)(cid:1), (cid:2)(H 2,n − z I)−1 − (H 1,n − z I)−1(cid:3) ∈ B1(cid:0)L2(R2)(cid:1), n ∈ N, z ∈ C\[0, ∞), (3.40) showing that the spectral shift functions ξ( · ; H 2, H 1) and ξ( · ; H 2,n, H 1,n) for the pairs (H 2, H 1) and (H 2, H 1), n ∈ N, respectively, are well-defined and satisfy ξ( · ; H 2, H 1), ξ( · ; H 2,n, H 1,n) ∈ L1(cid:0)R; (λ2 + 1)−1dλ(cid:1), n ∈ N. Since H j > 0, H j,n > 0, n ∈ N, j = 1, 2, one uniquely introduces ξ( · ; H 2, H 1) and ξ( · ; H 2,n, H 1,n), n ∈ N, by requiring that (3.39) (3.41) ξ(λ; H 2, H 1) = 0, ξ( · ; H 2,n, H 1,n) = 0, λ < 0, n ∈ N. (3.42) As shown in [10], one can now prove the following intimate connection between ξ( · ; A+,n, A−) and ξ( · ; H 2,n, H 1,n), n ∈ N, the Pushnitski-type formula [36], [49], ξ(λ; H 2,n, H 1,n) = 1 π λ1/2 −λ1/2 ξ(ν; A+,n, A−)dν (λ − ν2)1/2 for a.e. λ > 0, n ∈ N. (3.43) Moreover, as shown in [10] and [11], one indeed has the convergence property lim n→∞(cid:13)(cid:13)(cid:2)(H 2,n − z I)−1 − (H 1,n − z I)−1(cid:3) − [(H 2 − z I)−1 − (H 1 − z I)−1(cid:3)(cid:13)(cid:13)B1(L2(R2)) = 0, Thus, Theorem 3.5 applies and hence yields, lim n→∞ k[f (H 2,n) − f (H 1,n)] − [f (H 2) − f (H 1)]kB1(L2(R2)) = 0, z ∈ C\R. (3.44) f ∈ C∞ 0 (R). (3.45) This, in turn permits one to take the limit n → ∞ in (3.43), implying ξ(λ; H 2, H 1) = 1 π λ1/2 −λ1/2 ξ(ν; A+, A−)dν (λ − ν2)1/2 for a.e. λ > 0, n ∈ N. (3.46) Equation (3.46) combined with (3.38) yields ξ(λ; H 2, H 1) = ξ(ν; A+, A−) = 1 2π R dx φ(x) (3.47) for a.e. λ > 0 and a.e. ν ∈ R. As a consequence of (3.47), the semigroup regularized Witten index Ws(DA) of the non-Fredholm operator DA exists and equals Ws(D A) = ξ(0+; H 2, H 1) = ξ(0; A+, A−) = 1 2π R dx φ(x). (3.48) This yields an alternative proof of the principal Witten index results in [10] and [15]. The following result provides an extension of Theorem 3.5 to higher powers of resolvents (necessitated by applications to d-dimensional Dirac-type operators as hinted at in the introduction). Theorem 3.8. Let Sj,n, n ∈ N, and Sj, j = 1, 2, be self-adjoint in H, and assume that Sj,n converges in strong resolvent sense as n → ∞ to Sj, j = 1, 2, respectively. Suppose that for some m ∈ N and some p ∈ [1, ∞), (cid:2)(S2,n − zIH)−m − (S1,n − zIH)−m(cid:3),(cid:2)(S2 − zIH)−m − (S1 − zIH)−m(cid:3) ∈Bp(H), 14 If F. GESZTESY AND R. NICHOLS z ∈ C\R, n ∈ N. (3.49) and for some z0 ∈ C\R, lim z ∈ C\R, n→∞(cid:13)(cid:13)(cid:2)(S2,n − zIH)−m − (S1,n − zIH)−m(cid:3) −(cid:2)(S2 − zIH)−m − (S1 − zIH)−m(cid:3)(cid:13)(cid:13)Bp(H) = 0, (S1,n − z0IH)−m(cid:2)(S2,n − z0IH)−1 − (S1,n − z0IH)−1(cid:3), (S1 − z0IH)−m(cid:2)(S2 − z0IH)−1 − (S1 − z0)−1(cid:3) ∈ Bp(H), n→∞(cid:13)(cid:13)(S1,n − z0IH)−m(cid:2)(S2,n − z0IH)−1 − (S1,n − z0IH)−1(cid:3) − (S1 − z0IH)−m(cid:2)(S2 − z0IH)−1 − (S1 − z0)−1(cid:3)(cid:13)(cid:13)Bp(H) = 0, lim k[f (S2,n) − f (S1,n)] − [f (S2) − f (S1)]kBp(H) = 0, f ∈ C∞ 0 (R). with then lim n→∞ (3.50) (3.52) (3.53) n ∈ N, (3.51) In addition, these results hold upon systematically replacing Bp(H) by B(H) in (3.49) -- (3.53). Proof. Let f ∈ C∞ analytic extension. Following a device due to Khochman [43], one introduces 0 (R2) a compactly supported almost 0 (R) be fixed and efℓ,σ ∈ C∞ g(x) = f (x)(x − z0)m, (3.54) concluding g ∈ C∞ applied to g, one obtains for any self-adjoint operator S in H, 0 (R). By the Davies -- Helffer -- Sjostrand functional calculus (2.9) g(S) = 1 π C dxdy ∂efℓ,σ ∂z and hence, f (S) = (S − z0IH)−mg(S) (z)(z − z0)m(S − zIH)−1, (3.55) ∂z ∂efℓ,σ (z)(z − z0)m(cid:12)(cid:12)(cid:12)(cid:12) = 1 π C dxdy (z)(z − z0)m(S − z0IH)−m(S − zIH)−1. (3.56) Applying (3.56) multiple times choosing H ∈ {S2, S1, S2,n, S1,n}, one infers (cid:13)(cid:13)(cid:2)f (S2,n) − f (S1,n)(cid:3) −(cid:2)f (S2) − f (S1)(cid:3)(cid:13)(cid:13)Bp(H) 1 6 ∂z ∂efℓ,σ dxdy(cid:12)(cid:12)(cid:12)(cid:12) π C ×(cid:13)(cid:13)(cid:2)(S2,n − z0IH)−m(S2,n − zIH)−1 − (S1,n − z0IH)−m(S1,n − zIH)−1(cid:3) −(cid:2)(S2 − z0IH)−m(S2 − zIH)−1 − (S1 − z0IH)−m(S1 − zIH)−1(cid:3)(cid:13)(cid:13)Bp(H), n ∈ N. (3.57) In order to prove the convergence claim in (3.53), the idea is to take the limit n → ∞ and apply dominated convergence in (3.57). However, doing so requires one to obtain an n-independent integrable majorant for the expression under the integral in (3.57) and then to show that the integrand converges to zero pointwise + (S1,n − z0IH)−m(S2,n − zIH)−1 − (S1,n − z0IH)−m(S1,n − zIH)−1 = (S2,n − z0IH)−m(S2,n − zIH)−1 − (S1,n − z0IH)−m(S2,n − zIH)−1 (cid:2)(S2,n − z0IH)−m(S2,n − zIH)−1 − (S1,n − z0IH)−m(S1,n − zIH)−1(cid:3) −(cid:2)(S2 − z0IH)−m(S2 − zIH)−1 − (S1 − z0IH)−m(S1 − zIH)−1(cid:3) −(cid:2)(S2 − z0IH)−m(S2 − zIH)−1 − (S1 − z0IH)−m(S2 − zIH)−1 + (S1 − z0IH)−m(S2 − zIH)−1 − (S1 − z0IH)−m(S1 − zIH)−1(cid:3) =(cid:2)(S2,n − z0IH)−m − (S1,n − z0IH)−m(cid:3)(S2,n − zIH)−1 + (S1,n − z0IH)−m(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1(cid:3) −(cid:8)(cid:2)(S2 − z0IH)−m − (S1 − z0IH)−m(cid:3)(S2 − zIH)−1 + (S1 − z0IH)−m(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3)(cid:9) =(cid:8)(cid:2)(S2,n − z0IH)−m − (S1,n − z0IH)−m(cid:3)(S2,n − zIH)−1 −(cid:2)(S2 − z0IH)−m − (S1 − z0IH)−m(cid:3)(S2 − zIH)−1(cid:9) +(cid:8)(S1,n − z0IH)−m(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1 − (S1 − z0IH)−m(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3)(cid:9), (cid:13)(cid:13)(cid:2)(S2,n − z0IH)−m − (S1,n − z0IH)−m(cid:3)(S2,n − zIH)−1 −(cid:2)(S2 − z0IH)−m − (S1 − z0IH)−m(cid:3)(S2 − zIH)−1(cid:13)(cid:13)Bp(H) 6 eC(z0)Im(z)−1, z ∈ C\R, n ∈ N, For the first term in braces after the final equality in (3.58), one has a bound of the type (3.58) z ∈ C\R, n ∈ N. ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 15 with respect to z as n → ∞. In order to carry this out, one expresses the difference in the k · kBp(H)-norm in (3.57) as follows: (3.59) for a constant eC(z0) > 0 which does not depend on n ∈ N or z ∈ C\R. The estimate in (3.59) follows at once from the triangle inequality, basic properties of the Schatten -- von Neumann trace ideals, the standard resolvent estimate k(S − zIH)−1kB(H) 6 Im(z)−1, z ∈ C\R, (3.60) for an arbitrary self-adjoint operator S in H, and assumption (3.50). Moreover, by Lemma 3.2, one also infers lim n→∞(cid:13)(cid:13)(cid:2)(S2,n − z0IH)−m − (S1,n − z0IH)−m(cid:3)(S2,n − zIH)−1 −(cid:2)(S2 − z0IH)−m − (S1 − z0IH)−m(cid:3)(S2 − zIH)−1(cid:13)(cid:13)Bp(H) = 0, For the second term in braces after the final equality in (3.58), z ∈ C\R. (3.61) (S1,n − z0IH)−m(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1 − (S1 − z0IH)−m(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3) = (S1 − z0IH)−m(cid:2)(S1 − zIH)−1 − (S2 − zIH)−1(cid:3) − (S1,n − z0IH)−m(cid:2)(S1,n − zIH)−1 − (S2,n − zIH)−1(cid:3) = {IH + (z − z0)(S1 − zIH)−1}(S1 − z0IH)−m(cid:2)(S1 − z0IH)−1 − (S2 − z0IH)−1(cid:3) 16 F. GESZTESY AND R. NICHOLS × {IH + (z − z0)(S2 − zIH)−1} − {IH + (z − z0)(S1,n − zIH)−1} × (S1,n − z0IH)−m(cid:2)(S1,n − z0IH)−1 − (S2,n − z0IH)−1(cid:3) × {IH + (z − z0)(S2,n − zIH)−1}, z ∈ C\R, n ∈ N. (3.62) Using (3.60), one finds for any self-adjoint operator H, kIH + (z0 − z)(H − zIH)−1kB(H) 6 1 + (z0 + z)Im(z)−1 6 2(z0 + z)Im(z)−1, z ∈ C\R. (3.63) As a result of (3.52), (3.62) and (3.63), (cid:13)(cid:13)(S1,n − z0IH)−m(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1 − (S1 − z0IH)−m(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3)(cid:13)(cid:13)Bp(H) 6 bC(z0) z02 + z2 Im(z)2 , z ∈ C\R, n ∈ N, (3.52) and another application of Lemma 3.4 immediately imply for a constant bC(z0) > 0 which does not depend on n ∈ N or z ∈ C\R. Moreover, n→∞(cid:13)(cid:13)(S1,n − z0IH)−m(cid:2)(S2,n − zIH)−1 − (S1,n − zIH)−1 z ∈ C\R. lim − (S1 − z0IH)−m(cid:2)(S2 − zIH)−1 − (S1 − zIH)−1(cid:3)(cid:13)(cid:13)Bp(H) = 0, (cid:12)(cid:12)z − z0(cid:12)(cid:12)m(cid:12)(cid:12)(cid:12)(cid:12) (z)(cid:12)(cid:12)(cid:12)(cid:12)(cid:18) eC(z0) Im(z)2 (cid:19), + bC(z0) ∂efℓ,σ [z02 + z2] Im(z) ∂z The estimates in (3.60) and (3.64) show that away from R, which has dxdy-measure equal to zero, the integrand in (3.57) is bounded above by (3.64) (3.65) z ∈ C\R, (3.66) which is integrable with ℓ = 2 in light of (2.7) and the fact that efℓ,σ is compactly supported. Therefore, taking the limit n → ∞ on both sides of (3.57) and then applying dominated convergence in combination with (3.59), (3.61), and (3.65) yields (3.53). Clearly, the proof remains valid with Bp(H) replaced by B(H). (cid:3) Remark 3.9. Although applicable to the Witten index computation described in the introduction, Theorem 3.8 is far from optimal. Indeed, upon communicat- ing Theorem 3.8 to G. Levitina, D. Potapov, and F. Sukochev, they subsequently pointed out to us [44] that an application of the double operator integral method permits one to extend the classes of functions f to the one employed in [56], and more importantly, the DOI approach permits one to dispense with the conditions (3.51) and (3.52) altogether. (Conditions (3.51) and (3.52) are clearly an artifact of the resolvent term (S − zIH)−1 in formula (3.56)). This will be further pursued elsewhere [13]. Remark 3.10. While we exclusively focused on applications to self-adjoint operators S, as long as σ(T ) ⊂ R and the singularity of the resolvent (T − zIH)−1 of T as z approaches the spectrum is uniformly bounded by Im(z)−N for some fixed N ∈ N, choosing ℓ ∈ N sufficiently large in efℓ,σ, one can handle such classes of non-self- adjoint operators T , particularly, operators in Banach spaces with real spectrum. ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 17 In fact, a functional calculus for the case of a non-self-adjoint operator T with σ(T ) ⊂ R and a resolvent that satisfies an estimate of the type k(T − zIH)−1kB(H) 6 cIm(z)−1(cid:18) hzi Im(z)(cid:19)α , z ∈ C\R, (3.67) for some c > 0 and α > 0, was discussed in [18], [19], and in subsequent develop- ments in [1], [17], [32], [33]. Moreover, the case where the spectrum is contained in ⋄ the unit circle or contained in finitely-many smooth arcs was discussed in [28]. Appendix A. Some Useful Resolvent Identities In this appendix we recall some well-known, yet useful relations for (powers of) resolvents. We start by recalling the well-known identity (see, e.g., [54, p. 178]), (T2 − zIH)−1 − (T1 − zIH)−1 = (T2 − z0IH)(T2 − zIH)−1 ×(cid:2)(T2 − z0IH)−1 − (T1 − z0IH)−1(cid:3)(T1 − z0IH)(T1 − zIH)−1, z, z0 ∈ ρ(T1) ∩ ρ(T2), (A.1) where Tj, j = 1, 2, are linear operators in H with ρ(T1) ∩ ρ(T2) 6= ∅. In addition, if S is self-adjoint in H, we recall the elementary estimate, Im(z)−1, z, z0 ∈ C\R. In addition, for m ∈ N, we note (cf. [55, p. 315]), (T2 − zIH)−(m+1) − (T1 − zIH)−(m+1) (cid:13)(cid:13)(S − z0IH)(S − zIH)−1(cid:13)(cid:13)B(H) =(cid:13)(cid:13)IH + (z − z0)(S − zIH)−1(cid:13)(cid:13)B(H) 6 81/2(cid:2)z02 + z2(cid:3)1/2 =(cid:2)(T2 − zIH)−m − (T1 − zIH)−m(cid:3)(T1 − zIH)−1 + (T2 − zIH)−m(cid:2)(T2 − zIH)−1 − (T1 − zIH)−1(cid:3), = (T2 − zIH)−1(cid:2)(T2 − zIH)−m − (T1 − zIH)−m(cid:3) +(cid:2)(T2 − zIH)−1 − (T1 − zIH)−1(cid:3)(T2 − zIH)−m −(cid:2)(T2 − zIH)−1 − (T1 − zIH)−1(cid:3)(cid:2)(T2 − zIH)−m − (T1 − zIH)−m(cid:3), (T2 − zIH)−(m+1) − (T1 − zIH)−(m+1) z ∈ ρ(T1) ∩ ρ(T2). z ∈ ρ(T1) ∩ ρ(T2), (A.2) (A.3) (A.4) and Next, by applying Cauchy's integral formula, f (k)(z) = − k! 2πi ‰Γ dζ f (ζ) (ζ − z)k+1 , z ∈ Ω, (A.5) where f is an analytic function in the open set Ω ⊂ C, and Γ is a counterclockwise- oriented contour encompassing the point z ∈ Ω, to a densely defined, closed linear operator T in H with nonempty resolvent set, one obtains a formula for higher powers of the resolvent of T in terms of a fixed lower power as follows, (T − zIH)−k = − (k − m)!(m − 1)! 2πi[(k − 1)!] ‰Γz dζ (ζ − z)m−k−1(T − ζIH)−m, (A.6) z ∈ ρ(H0), 18 F. GESZTESY AND R. NICHOLS where for each z ∈ ρ(T ), Γz is any counterclockwise-oriented circular contour cen- tered at z which does not intersect or encompass points of σ(T ). The following lemma (cf. [55, p. 210]) states an elementary, yet useful, fact: z ∈ C\R, (A.7) z ∈ C\R, n > m. (A.8) Lemma A.1. Let Sj, j ∈ {1, 2} be self-adjoint operators in H. If for some p ∈ [1, ∞) ∪ {∞} and some m ∈ N, then (cid:2)(S2 − zIH)−m − (S1 − zIH)−m(cid:3) ∈ Bp(H), (cid:2)(S2 − zIH)−n − (S1 − zIH)−n(cid:3) ∈ Bp(H), Proof. It suffices to apply the Cauchy-type formula (A.6) and note that (S2 − zIH)−n − (S1 − zIH)−n = − (n − m)!(m − 1)! 2πi[(n − 1)!] ‰Γz dζ (ζ − z)m−n−1(cid:2)(S2 − ζIH)−m − (S1 − zIH)−m(cid:3), z ∈ C\R, n > m, (A.9) where Γz is a counterclockwise-oriented circular contour centered at z that does not intersect R. (cid:3) Acknowledgments. We are indebted to Alan Carey, Galina Levitina, Denis Potapov, Fedor Sukochev, Yuri Tomilov, and Dmitriy Zanin for helpful discus- sions, and particularly to Yuri Tomilov for pointing out to us a number of key references in connection with almost analytic extensions. References [1] A. B´atkai, and E. Facanga, The spectral mapping theorem for Davies' functional calculus. Rev. Roumaine Math. Pures Appl. 48, 365 -- 372 (2003). [2] Ju. M. Berezanskii, Expansions in Eigenfunctions of Selfadjoint Operators, Translations of Mathematical Monographs, Vol. 17, Amer. Math. Soc., Providence, R.I. 1968. [3] Yu. M. Berezanskii, Selfadjoint Operators in Spaces of Functions of Infinitely Many Vari- ables, Translations of Mathematical Monographs, Vol. 63, Amer. Math. Soc., Providence, R.I. 1986. [4] Yu. M. Berezansky and A. A. Kalyuzhnyi, Harmonic Analysis in Hypercomplex Systems, Mathematics and its Applications, 434, Kluwer, Dordrecht, 1998. [5] Y. M. Berezansky and Y. G. Kondtratiev, Spectral methods in infinite-dimensional analysis. Vol. 1, Mathematical Physics and Applied Mathematics, 12/1, Kluwer Academic, Dordrecht, 1995. [6] Y. M. Berezansky and Y. G. Kondtratiev, Spectral methods in infinite-dimensional analysis. Vol. 2, Mathematical Physics and Applied Mathematics, 12/2, Kluwer Academic, Dordrecht, 1995. [7] Y. M. Berezansky, Z. G. Sheftel, and G. F. Us, Functional Analysis. Vol. I, Operator Theory: Advances and Applications, Vol. 85, Birkhauser, Basel, 1996. [8] Y. M. Berezansky, Z. G. Sheftel, and G. F. Us, Functional Analysis. Vol. II, Operator Theory: Advances and Applications, Vol. 86, Birkhauser, Basel, 1996. [9] M. Sh. Birman and M. Solomyak, Double operator integrals in a Hilbert space, Integral Eq. Oper. Theory 47, 131 -- 168 (2003). [10] A. Carey, F. Gesztesy, G. Levitina, D. Potapov, F. Sukochev, and D. Zanin, On index theory for non-Fredholm operators: a (1 + 1)-dimensional example, arXiv:arXiv:1509.01356, Math. Nachrichten, to appear. [11] A. Carey, F. Gesztesy, G. Levitina, D. Potapov, F. Sukochev, and D. Zanin, Trace formulas for a (1 + 1)-dimensional model operator, preprint, 2014. [12] A. Carey, F. Gesztesy, G. Levitina, D. Potapov, F. Sukochev, and D. Zanin, On index theory for non-Fredholm operators: a (2 + 1)-dimensional example, in preparation. ALMOST ANALYTIC EXTENSIONS TO OPERATOR BOUNDS 19 [13] A. Carey, F. Gesztesy, G. Levitina, R. Nichols, D. Potapov, and F. Sukochev, Double operator integral methods applied to continuity of spectral shift functions, arXiv:1511.07998. [14] A. Carey, F. Gesztesy, G. Levitina, and F. Sukochev, On the index of a non-Fredholm model operator, arXiv:1509.01580, Operators and Matrices, to appear. [15] A. Carey, F. Gesztesy, D. Potapov, F. Sukochev, and Y. Tomilov, On the Witten index in terms of spectral shift functions, arXiv:1404.0740, J. Analyse Math., to appear. [16] G. Carron, T. Coulhon, and E. M. Ouhabaz, Gaussian estimates and Lp-boundedness of Riesz means. J. Evol. Equ. 2, 299 -- 317 (2002). [17] N. S. Claire, Spectral mapping theorem for the Davies -- Helffer -- Sjostrand functional calculus, J. Math. Phys. Anal. Geom. 8, 221 -- 239 (2012). [18] E. B. Davies, The functional calculus, J. London Math. Soc. (2) 52, 166 -- 176 (1995). [19] E. B. Davies, Lp spectral independence and L1 analyticity, J. London Math. Soc. (2) 52, 177 -- 184 (1995). [20] E. B. Davies, Spectral Theory and Differential Operators, Cambridge University Press, Cam- bridge, 1995. [21] M. Dimassi, Spectral shift function in the large coupling constant limit, Ann. H. Poincar´e 7, 513 -- 525 (2006). [22] M. Dimassi and A. T. Duong, Trace asymptotics formula for the Schrodinger operators with constant magnetic fields, J. Math. Anal. Appl. 416, 427 -- 448, (2014). [23] M. Dimassi and V. Petkov, Spectral shift function and resonances for non-semi-bounded and Stark Hamiltonians, J. Math. Pures Appl. 82, 1303 -- 1342 (2003). [24] M. Dimassi and V. Petkov, Spectral shift function for operators with crossed magnetic and electric fields, Rev. Math. Phys. 22, 355 -- 380 (2010). [25] M. Dimassi and J. Sjostrand, Trace asymptotics via almost analytic extensions, in Partial Differential Equations and Mathematical Physics, The Danish-Swedish Analysis Seminar, 1995, L. Hormander and A. Melin (eds.), Birkhauser, Basel, 1996, pp. 126 -- 142. [26] M. Dimassi and J. Sjostrand, Spectral Asymptotics in the Semi-Classical Limit, London Math. Soc. Lecture Note Series, Vol. 268, Cambridge University Press, Cambridge, 1999. [27] M. Dimassi and M. Zerzeri, A tme-dependent approach for the study of spectral shift function, C. R. Acad. Sci. Paris Ser. I 350, 375 -- 378 (2012). [28] E. M. Dy'nkin, An operator calculus based upon the Cauchy -- Green formula, J. Sov. Math. 4, 329 -- 334 (1975). [29] E. M. Dy'nkin, The pseudoanalytic extension, J. Analyse Math. 60, 45 -- 70 (1993). [30] R. Froese, D. Hasler, and W. Spitzer, On the AC spectrum of one-dimensional random Schrodinger operators with matrix-valued potentials, Math. Phys. Anal. Geom. 13, 219 -- 233 (2010). [31] J. Frohlich, M. Griesemer, and I. M. Sigal, Spectral theory for the standard model of non- relativistic QED, Comm. Math. Phys. 283, 613 -- 646 (2008). [32] J. E. Gal´e, P. J. Miana, and T. Pytlik, Spectral properties and norm estimates associated to (k) the C c -functional calculus, J. Operator Theory 48, 385 -- 418 (2002). [33] J. E. Gal´e and T. Pytlik, Functional calculus for infinitesimal generators of holomorphic semigroups, J. Funct. Anal. 150, 307 -- 355 (1997). [34] C. Gerard, Sharp propagation estimates for N -particle systems, Duke Math. J. 67, 483 -- 515 (1992). [35] C. G´erard, A proof of the abstract limiting absorption principle by energy estimates, J. Funct. Anal. 254, 2707 -- 2724 (2008). [36] F. Gesztesy, Y. Latushkin, K. A. Makarov, F. Sukochev, and Y. Tomilov, The index formula and the spectral shift function for relatively trace class perturbations, Adv. Math. 227, 319 -- 420 (2011). [37] M. Griesemer, Exponential decay and ionization thresholds in non-relativistic quantum elec- trodynamics, J. Funct. Anal. 210, 321 -- 340 (2004). [38] H. R. Grumm, Two theorems about Cp, Rep. Math. Phys. 4, 211 -- 215 (1973). [39] L. Hormander, Fourier Integral Operators: Lectures at the Nordic Summer School of Math- ematics, 1969. [40] B. Helffer and J. Sjostrand, Equation de Schrodinger avec champ magn´etique et ´equaton de Harper, in Schrodinger Operators, H. Holden and A. Jensen (eds.), Lecture Notes in Physics, Vol. 345, Springer, Berlin, 1989, pp. 138 -- 197. 20 F. GESZTESY AND R. NICHOLS [41] B. Helffer and J. Sjostrand, On diamagnetism and de Haas -- van Alphen effect, Ann. H. Poincar´e A52, 303 -- 375 (1990). [42] A. Jensen and S. Nakamura, Mapping properties of functions of Schrodinger operators be- tween Lp-spaces and Besov spaces, in Spectral and Scattering Theory and Applications, K. Yajima (ed.), Adv. Studies in Pure Math., Vol. 23, Math. Soc. Japan, Kinokuniya Company, Tokyo, Japan, 1994, pp. 187 -- 209. [43] A. Khochman, Resonances and spectral shift function for the semi-classical Dirac operator, Rev. Math. Phys. 19, 1071 -- 1115 (2007). [44] G. Levitina, D. Potapov, and F. Sukochev, private communication, January 2015. [45] M. Martin and M. Putinar, Lectures on Hyponormal Operators, Operator Theory: Advances and Applications, Vol. 39, Birkhauser, Basel, 1989. [46] A. Melin and J. Sjostrand, Fourier integral operators with complex-valued phase functions, in Fourier Integral Operators and Partial Differential Equations, J. Chazarain (ed.), Lecture Notes in Math., Vol. 459, Springer, Berlin, 1975, pp. 120 -- 223. [47] S. O'Rourke, D. Renfrew, and A. Soshnikov, On fluctuations of matrix entries of regular functions of Wigner matrices with non-identically distributed entries, J. Theoret. Probab. 26, 750 -- 780 (2013). [48] A. Pizzo, D. Renfrew, and A. Soshnikov, On finite rank deformations of Wigner matrices. Ann. Inst. Henri Poincar´e Probab. Stat. 49, 64 -- 94 (2013). [49] A. Pushnitski, The spectral flow, the Fredholm index, and the spectral shift function, in Spec- tral Theory of Differential Operators: M. Sh. Birman 80th Anniversary Collection, T. Suslina and D. Yafaev (eds.), AMS Translations, Ser. 2, Advances in the Mathematical Sciences, Vol. 225, Amer. Math. Soc., Providence, RI, 2008, pp. 141 -- 155. [50] M. Reed and B. Simon, Methods of Modern Mathematical Physics. I: Functional Analysis, revised and enlarged edition, Academic Press, New York, 1980. [51] B. Simon, Trace Ideals and Their Applications, Mathematical Surveys and Monographs, Vol. 120, 2nd ed., Amer. Math. Soc., Providence, RI, 2005. [52] J. Sjostrand and M. Zworski, Complex scaling and the distribution of scattering poles, J. Amer. Math. Soc. 4, 729 -- 769 (1991). [53] E. Skibsted, Smoothness of N -body scattering amplitudes, Rev. Math. Phys. 4, 619 -- 658 (1992). [54] J. Weidmann, Linear Operators in Hilbert Spaces, Graduate Texts in Mathematics, Vol. 68, Springer, New York, 1980. [55] D. R. Yafaev, Mathematical Scattering Theory. General Theory, Amer. Math. Soc., Provi- dence, RI, 1992. [56] D. R. Yafaev, A trace formula for the Dirac operator, Bull. London Math. Soc. 37, 908 -- 918 (2005). Department of Mathematics, University of Missouri, Columbia, MO 65211, USA E-mail address: [email protected] URL: http://www.math.missouri.edu/personnel/faculty/gesztesyf.html Mathematics Department, The University of Tennessee at Chattanooga, 415 EMCS Building, Dept. 6956, 615 McCallie Ave, Chattanooga, TN 37403, USA E-mail address: [email protected] URL: http://www.utc.edu/faculty/roger-nichols/index.php
1103.2090
1
1103
2011-03-10T17:53:30
Random Series of Trace Class Operators
[ "math.FA", "math.OA", "math.PR" ]
In this lecture, we present some results on Gaussian (or Rademacher) random series of trace class operators, mainly due jointly with F. Lust-Piquard. We will emphasize the probabilistic reformulation of these results, as well as the open problems suggested by them. We start by a brief survey of what is known about the problem of characterizing a.s. convergent (Gaussian or Rademacher) series of random vectors in a Banach space. The main result presented here is that for certain pairs of Banach spaces $E,F$ that include Hilbert spaces (and type 2 spaces with the analytic UMD property), we have $$ R(E\hat\otimes F) =R(E)\hat\otimes F + E\hat\otimes R(F) $$ where $R(E)$ denotes the space of convergent Rademacher series with coefficients in $E$ and $E\hat\otimes F$ denotes the projective tensor product.
math.FA
math
Random Series of Trace Class Operators∗ by Gilles Pisier Texas A&M University College Station, TX 77843, U. S. A. and Universit´e Paris VI Equipe d'Analyse, Case 186, 75252 Paris Cedex 05, France 1 1 0 2 r a M 0 1 ] . A F h t a m [ 1 v 0 9 0 2 . 3 0 1 1 : v i X r a (1) (2) In this lecture, we would like to present some recent results on Gaussian (or Rademacher) random series of trace class operators, (cf. mainly [17]). We will emphasize the probabilistic refor- mulation of the results of [17], as well as the open problems suggested by these results. We will start by a survey of what is known about the following problems. Let B be a Banach space. Let (xn) be a sequence of elements of B. Let (gn) be a sequence of i.i.d. Gaussian random variables. Problem 1. Find a necessary and sufficient condition on the sequence (xn) for the a.s. convergence in norm of the series gnxn. ∞Xn=1 A similar question can be raised for series of the form P ±xn, where the signs are chosen at random. More precisely, let (εn) be a sequence of i.i.d. ±1-value random variables such that P (εn = +1) = P (εn = −1) = 1/2. We can ask Problem 2. Find a necessary and sufficient on condition (xn) for the a.s. convergence of the series εnxn. ∞Xn=1 εkxk. Let 0 ≤ p ≤ ∞. We will say that the sequence of partial Let Sn = Pk≤n gkxk and Rn = Pk≤n sums {Sn} converges in Lp(B) if there is a Bochner-measurable B-valued random variable S∞ such that kSn − S∞k → 0 in Lp. For p = 0, this corresponds to the convergence in probability of the series (1). (Recall that the definition of a Bochner-measurable random variable is equivalent to ∗In Proceedings IV CLAPEM (Mexico City, September 1990). 1 requiring that the variable essentially takes its values in a separable subspace of B. Equivalently, the image probability measure on B is a Radon measure.) The following result is well known. In the Gaussian case, it goes back to [5] and [13], while the Rademacher case goes back to [10] and [11] (cf. also [9] for the case p = 0). Theorem 1. The following assertions are equivalent: (i) The sequence of partial sums {Sn} converges a.s. in the norm of B. (ii) The sequence {Sn} converges in Lp(B) for some 0 ≤ p < ∞. (iii) The sequence {Sn} converges in Lp(B) for all 0 ≤ p < ∞. The same equivalence holds with the partial sums {Rn} of the series (2) instead of the series (1). Moreover, when the series (1) or (2) is convergent a.s. to a limit denoted by S, necessarily there is a δ > 0 such that Z exp(δkSk2) < ∞. Corollary. For all 0 < p < q < ∞, the norms induced by Lp(B) and Lq(B) on the set of all a.s. convergent series of the form (1) (resp. (2)) are equivalent. (In particular, they are all equivalent to the norm induced by L2(B).) the convergence in L2(B) Remark. Of course, the preceding result gives an answer to Problem 1: for instance is necessary and sufficient for the a.s. convergence. However, this does not help us (except in the few simple cases below) since we usually cannot compute the norm in L2(B) for a general Banach space B. We are looking for a condition as simple as possible characterizing the a.s. convergence of (1). Remark. Recently, Talagrand [30] gave a remarkable necessary and sufficient condition for the a.s. continuity of Gaussian processes. This can be rephrased as a solution of Problem 1 in full generality, as follows. Let K be the closed unit ball of the dual B∗ equipped with the topology σ(B∗, B) for which it is compact. The series (1) converges a.s. in B iff the "majorizing measure condition" holds. The latter means that there is a positive finite measure m on K such that if we set ∀t, s ∈ K and ∀ε > 0 B(t, ε) = {s ∈ K d(t, s) < ε} then the condition can be stated as d(t, s) =(cid:16)X hxn, t − si2(cid:17)1/2 m(B(t, ε))(cid:19)1/2 t∈KZ δ 0 (cid:18)Log 1 sup lim δ→0 dε = 0. This completely solves the problem when we do not have any extra information on the Banach space B. However, as the discussion below will show, the spaces B which arise in analysis are often given with an additional structure (for instance a function space, or an operator space . . .) which allows to find a simple very explicit necessary and sufficient condition. For instance, it is not easy at all (although Talagrand did do it) to deduce the condition in Example 1 below (the Hilbert space 2 case) from the majorizing measure condition. In this lecture we want to concentrate on Banach spaces B which are as "concrete" as possible. This is somehow the other end of the spectrum from Talagrand's result which completely answers (and quite remarkably so) the "abstract" case. One interesting feature of Theorem 1 is that it allows us to define the Banach space of all a.s. convergent series of the form (1) or (2). Indeed, given a fixed sequence (gn) of i.i.d. Gaussian normal variables on a probability space (Ω, A, P), we define the space G(B) as the subspace of L2(Ω, A, P; B) formed by all the convergent series of the form (1), and we equip it with the norm induced by L2(B). Similarly, we define the space R(B) as the subspace of all convergent series of the form (2) in L2(B). It is not hard to show that G(B) (resp. R(B)) is a Banach space which coincides with the closure in L2(B) of all the finite sums of the form (1) (resp. (2)). By Theorem 1, our Problems 1 and 2 are the same as finding a "simple" description of the Banach spaces G(B) and R(B). Let us first review several known cases to illustrate what we mean by a "simple" characterization. Example 1. If B is a Hilbert space, it is well known that (1) [resp. (2)] converges a.s. iff Moreover, G(B) (resp. R(B)) can be identified with ℓ2(B) with equivalent norms. X kxnk2 < ∞. Example 2. If B = Lp(S, Σ, µ) for some measure space (S, Σ, µ) and if 1 ≤ p < ∞ (we insist that p < ∞) then the series (1) [resp. (2)] is a.s. convergent iff Moreover, there is a constant C (depending only on p) such that xn(s)2!p/2 dµ(s) < ∞. Z Xn dµ(s)(cid:19)1/p 1 C(cid:18)Z (cid:16)X xn(s)2(cid:17)p/2 and similarly with the series (2) instead of (1). ≤(cid:13)(cid:13)(cid:13)X gnxn(cid:13)(cid:13)(cid:13)L2(B) ≤(cid:18)Z (cid:16)X xn(s)2(cid:17)p/2 dµ(s)(cid:19)1/p , In particular, if B = Lp(µ), the spaces G(B) and R(B) can be identified with the space Lp(µ; ℓ2), with equivalent norms. This has been known for a long time, as a consequence of classical inequal- ities. The "modern" way to prove this is to use Fubini's theorem and the Corollary of Theorem 1 and to work with the Lp(B) norm on G(B) or R(B) instead of the L2(B)-norm. The basic examples above have been one of the motivation for the theory of type and cotype of Banach spaces, (cf. [19]). A Banach space B is called of type p (1 ≤ p ≤ 2) if the condition P kxnkp < ∞ is sufficient for the a.s. convergence of the series (2). Similarly, the space B is called of cotype q (2 ≤ q < ∞) if the conditionP kxnkq < ∞ is necessary for the a.s. convergence of the series (2). 3 With this terminology, we find that (by Example 1) if B is isomorphic to a Hilbert space then B is of type 2 and cotype 2. By a well known result of Kwapie´n [12] the converse is also true, so that only in spaces isomorphic to Hilbert do we have as simple a characterization as in Example 1. From Example 2, it is rather easy to deduce (by Holder -- Minkowski) that Lp is of cotype max(p, 2) and, if p 6= ∞, of type min(p, 2). In our discussion, it is natural to ask for which spaces B the series of the form (1) and (2) are equivalent in the sense that (1) converges a.s. iff (2) also does. This was answered in [19] as follows. Theorem 2. The following properties of a Banach space B are equivalent: (i) The space B is of cotype q for some q < ∞. (ii) The a.s. convergence of a series of the form (1) is equivalent to the a.s. convergence of the corresponding series of the form (2). (iii) The spaces G(B) and R(B) can be naturally identified. This is also equivalent to the nonexistence in B of a constant λ for which there exists a sequence of finite dimensional subspaces Bn ⊂ B with Bn λ-isomorphic to ℓn (Note that the natural inclusion G(B) ⊂ R(B) holds for an arbitrary B, it is the converse inclusion which only holds if B has a finite cotype.) ∞. Since all the spaces that we will consider below fall into that category, we will sometimes only state our results for R(B), which is more natural in view of the methods of proof, but the reader should recall that the same results holds for G(B) as well. Example 3. Let B be a Banach lattice of cotype q for some q < ∞. Without loss of generality, we can assume that B is a Banach lattice of functions on some measure space (S, Σ, m). Consider again xn ∈ B. Then the series (2) converges a.s. if (3) (4) Moreover, there is a constant C such that ∈ B. xn(·)2!1/2 ∞X1 C(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)X xn2(cid:17)12/2(cid:13)(cid:13)(cid:13)(cid:13)B ≤(cid:13)(cid:13)(cid:13)X εnxn(cid:13)(cid:13)(cid:13)L2(B) 1 ≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)X xn2(cid:17)1/2(cid:13)(cid:13)(cid:13)(cid:13)B . This useful result can be found (in a different but equivalent or more general formulation) in the work of Maurey [18]. It gives us an identification of R(B) and G(B) with the space B(ℓ2) of all sequences (xn) such that (3) holds equipped with the norm appearing on the left side of (4). For a more thorough discussion of the many inequalities satisfied by a Banach lattice of finite cotype, (cf. e.g. the first chapter in the book [14]). In another direction, we refer the reader to [25, 26] for a class of function spaces which, although they are not Banach lattices, satisfy an inequality analogous to (4). 4 Recently, F. Lust-Piquard has obtained a striking noncommutative version of Example 2 (and of Example 3 but we will not discuss this here, see [16]). Although her results are valid in a more general framework, we will state them only for the simplest examples of noncommutative Lp-spaces, namely the Schatten p-classes Cp. For 1 ≤ p < ∞, we denote by Cp the space of all compact operators x : ℓ2 → ℓ2 such that trxp < ∞, where x = (x∗x)1/2. We equip this space with the norm kxkCp = (trxp)1/p. It is well known that trxp = tr(x∗x)p/2 = tr(xx∗)p/2 for each x in Cp. We will now consider our main Problems 1 and 2 with B = Cp. The first part (Theorem 3) seems perhaps less surprising at first glance than the second one (Theorem 4), because it appears more similar to the commutative case, but see below for a clarifi- cation of the duality between the two cases. Theorem 3 ([15]). Assume 2 ≤ q < ∞. Let B = Cq and let xn ∈ B. Then the series (2) (or (1)) converges a.s. iff the series xnx∗ n are both convergent in the strong operator topology Moreover, the norm in the space R(B) (or G(B)) is equivalent to the expression x∗ nxn and ∞P1 ∞P1 n(cid:17)q/2 nxn(cid:17)q/2 < ∞ and tr(cid:16)X xnx∗ tr(cid:16)X x∗ max n!1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Cq nxn!1/2(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Cq (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ∞X1 ∞X1 ,(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) xnx∗ x∗ < ∞. .  and satisfy (5) (6) ∞P1 (7) (8) Remark. Here the two conditions appearing in (5) are not equivalent and the theorem does not hold if we drop one condition in (5). The second part comes from [15] for 1 < p < 2 and from [17] for p = 1. Surprisingly, the condition is different for the interval [1,2[ than for the interval ]2, ∞[. Theorem 4 ([15, 17]). Assume 1 ≤ p ≤ 2. Let B = Cp and let xn ∈ B. Then the series (2) (or (1)) converges a.s. iff there is a decomposition xn = yn + zn with yn, zn ∈ Cp such that y∗ nyn and ∞P1 znz∗ n converge in the strong operator topology and satisfy Moreover, if we define < ∞. nyn(cid:17)p/2 tr(cid:16)X y∗ (xn)p = inf((cid:13)(cid:13)(cid:13)(cid:13)(cid:16)X y∗ n(cid:17)p/2 < ∞ and tr(cid:16)X znz∗ n(cid:17)1/2(cid:13)(cid:13)(cid:13)(cid:13)Cp) . nyn(cid:17)1/2(cid:13)(cid:13)(cid:13)(cid:13)Cp +(cid:13)(cid:13)(cid:13)(cid:13)(cid:16)X znz∗ 5 where the infimum runs over all possible decompositions xn = yn + zn, then there is a constant C (depending only on p) such that 1 C (xn)p ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ∞X1 εnxn(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)R(Cp) ≤ C(xn)p. Remark. To illustrate the preceding result, let us consider the following illuminating special case given in [15]. Let (εij ) be a collection of i.i.d. symmetric ±1 valued random variables as before, but this time indexed by N × N instead of N. Let A = (aij) be an infinite matrix with complex entries. Then, if 2 ≤ q < ∞, the random matrix (εij, aij) is a.s. in Cq iff we have both q/2 < ∞ and Xj Xi aji2!q/2 < ∞. Xi Xj Xj aij2 bij2 Moreover, if 1 ≤ p < 2, then (εij aij) is a.s. in Cp iff aij can be decomposed as aij = bij + cij with Xi p/2 < ∞ and Xj Xi cij 2!p/2 < ∞. (Note: the case 0 < p < 1 does not seem to be known.) p + 1 Remark. The norm appearing in (8) is equivalent to the dual norm to the norm appearing in (6) when 1 q = 1. This is a special case of the general duality between the intersection of two spaces and the sum of their duals. However, the reader should be warned that (6) and (8) are not equivalent norms when p = q unless p = q = 2! Remark. In the preceding examples, when B = Lp or Cp, we can observe that if 1 < p < ∞ the dual of G(B) (resp. R(B)) can be identified with G(B∗) (resp. R(B∗)). This property has been extensively studied under the name of K-convexity. It means equivalently that there is a natural bounded linear projection from L2(B) onto G(B) or R(B). We refer the reader to [20, 24] for more details on that property. In particular, we proved (cf. [24]) that B is K-convex iff B is of type p for some p > 1 or iff B does not contain a sequence of finite dimensional subspaces uniformly isomorphic to ℓn 1 . In the case p = 1, the space C1 is the space of all trace class operators which can also be viewed as a tensor product. Let us recall the definition of the projective tensor product of two Banach spaces. Let E, F be Banach spaces and let u = xi ⊗ yi ∈ E ⊗ F with xi ∈ E, yi ∈ F . We define nPi=1 kukΛ = inf( nX1 kxikkyik) where the infimum runs over all possible representations of u as a finite sum as above. The projective tensor product Eb⊗F is defined as the completion of E ⊗ F for this norm. It is well known that 6 the spaces of all bounded operators either B(E, F ∗) (from E into F ∗) or B(F, E∗) (from F into E∗). We will see below that for a certain class of Banach spaces E and F , there is a rather simple its dual (Eb⊗F )∗ can be identified with the space of all bounded bilinear forms on E × F , or with characterization of the sequences xn ∈ Eb⊗F such that the series (1) or (2) converges a.s. in Eb⊗F . Quite surprisingly the only known proof (at the moment) of these results uses the factorization of operator valued analytic functions. In the case of C1 this factorization goes back to Sarason [29], following classical work in the matrix case by Wiener -- Masani and Helson -- Lowdenslager. Recently, these classical results have been generalized in [22] to the setting of type 2 Banach spaces (cf. also [8, 21, 23]). To explain more clearly the connection with series of independent random variables as in (1) and (2), we present our results on the infinite dimensional torus ∆ = TTT N, equipped with the probability measure m which is the infinite product of copies of the normalized Haar measure on the one dimensional torus TTT . Let B be a Banach space. Consider first a function F : TTT → B which is in L1(TTT , dt; B). We will say that F is analytic if its Fourier transform vanishes on the negative integers. Let 1 ≤ p ≤ ∞. Now consider a function f in Lp(∆, m; B). Let us denote by En the conditional expectation operator on Lp(∆, m; B) with respect to the σ-algebra generated by the first (n + 1) coordinates (t0, t1, . . . , tn) on the infinite dimensional torus ∆ = TTT N. We will say that f is a Hardy function if for each integer n ≥ 0 the function Enf is analytic as a function of the last variable tn, more precisely, the function t → E0f (t) is analytic and for all n > 0, for any fixed (t0, . . . , tn−1) ∈ TTT n the function t → Enf (t0, . . . , tn−1, t) is analytic in the above sense. We will denote by H p(∆, B) the closed subspace of Lp(∆, m; B) formed by all the Hardy func- tions f . This is the same as the set of all Hardy martingales considered in [4] and [6]. These Hardy martingales are a convenient discretization of a certain kind of stochastic integrals which includes analytic functions of the complex Brownian motion, (cf. e.g. [3, 32, 33]). We will need the following. Definition 1. Let E be a Banach space. We will say that E has the analytic UMD property if there is a constant K such that for all g in H 2(∆, E) and for all choices of signs εn = ±1 the series εn(Eng − En−1g) converges in H 2(∆, E) and satisfies (cid:18)Z (cid:13)(cid:13)(cid:13)X εn(Eng − En−1g)(cid:13)(cid:13)(cid:13) 2 dm(cid:19)1/2 ≤ K(cid:18)Z kgk2dm(cid:19)1/2 . We will denote by K(E) the smallest constant K such that this holds. We observe for further use that if we denote by µ the normalized uniform probability measure on all the choices of signs ε = (εn)n≥0 ∈ {−1, 1}N, then by averaging (9) over all choices of signs we find (denoting dng = Eng − En−1g) ∞Pn=1 (9) (10) hence by Fubini and a simple invariance argument, we also have for all points t ∈ TTT N 2 ZZ (cid:13)(cid:13)(cid:13)X εndng(t)(cid:13)(cid:13)(cid:13) ZZ (cid:13)(cid:13)(cid:13)X εne−itn dng(t)(cid:13)(cid:13)(cid:13) dm(t)dµ(ε) ≤ K 2Z kgk2dm, dm(t)dµ(ε) ≤ 4K 2Z kgk2dm. 2 7 This property is slightly weaker than the so-called UMD property for which we refer to Burkholder's paper [2]. Here we only require that "analytic" martingales (meaning those associated to a Hardy function g) are unconditional in L2(∆, m; B). The UMD property corresponds to the same as (9) but for arbitrary E-valued martingales. We refer to [6] for a discussion of the analytic UMD prop- erty. Note that the spaces Lp or Cp are UMD for 1 < p < ∞, and that L1 has the analytic UMD property (see [6]), while C1 fails it (cf. [8]). Remark. To illustrate the possible applications of Theorems 3 and 4, let us consider their significance for martaingales with values in Cq. Let (Mn) be a martingale with values in a UMD Banach space B. Let us denote as usual dMn = Mn − Mn−1. We can define the vector valued version of the maximal function and of the square function, as follows M ∗ = sup n kMnk and S(M ) =sup n Z (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) nX1 2 ǫkdMk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) dµ(ǫ) 1/2 . Then, it is known (cf. [2]) that for any 1 ≤ r ≤ ∞, there is a constant Kr such that for all B-valued martingales (Mn) in Lr(B), we have (Kr)−1kS(M )kr ≤ kM ∗kr ≤ KrkS(M )kr. It was observed in [1] (cf. also [27] for more information) that the spaces Cq are all UMD spaces when 1 < q < ∞. Thus, we can reformulate Theorems 3 and 4 in a martingale setting as follows. Given a martingale (Mn) with values in Cq, for any ω in our probability space, let xn = dMn(ω) and let us denote if 2 ≤ q < ∞ (resp. 1 ≤ p ≤ 2) by Sq(M )(ω) (resp. Sp(M )(ω)) the expression (6) (resp. (8)). Then, we can state for all 1 ≤ r, q < ∞, there is a constant Kr,q such that for all Cq-valued martingales (Mn), we have (Kr,q)−1kSq(M )kr ≤ kM ∗kr ≤ Kr,qkSq(M )kr. (11) xn = n ∈ E, zm ∈ F , ym ∈ E, zm n ∈ F such that case p = 1 of Theorem 4 to more general projective tensor products. Indeed, we have Theorem 5. Assume that E, F are both Banach spaces of type 2 with the analytic UMD property. Since it is well known that C1 can be identified with ℓ2b⊗ℓ2 it is natural to try to extend the Let B = Eb⊗F and let xn ∈ Eb⊗F . Then the series (1) or (2) converges a.s. iff there are positive scalars λm withP λm < ∞ and elements ym ∞Xm=1 n (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xn R(Eb⊗F ) ≈ R(E)b⊗F + Eb⊗R(F ), n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L2(F ) n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)L2(E) sup m (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)Xn Equivalently, we have the isomorphic identification λm[ym n ⊗ zm + ym ⊗ zm n ] sup m kymkE ≤ 1, sup m kzmkF ≤ 1. ≤ 1, sup such that ǫnym ǫnzm ≤ 1, (12) 8 and the norm of R(Eb⊗F ) is equivalent to the natural norm in the space R(E)b⊗F + Eb⊗R(F ) which is defined as (xn) = infnk(yn)kR(E) b⊗F + k(zn)kE b⊗R(F )o where the infimum runs over all decompositions xn = yn + zn. (Another trivially equivalent norm can be defined as the infimum ofP λm over all possible representations as in (11) above.) Remark. The sum in (12) is not a direct sum. The meaning of (12) is that the three spaces appearing in (9) are all naturally continuously injected into the set of all sequences of elements of Eb⊗F and a sequence (xn) comes from R(Eb⊗F ) iff it can be decomposed as the sum of a sequence in R(E)b⊗F and a sequence in Eb⊗R(F ). For example, if we take E = F = ℓ2, then R(E) ≈ ℓ2(E) and R(F ) ≈ ℓ2(F ), and one can check rather easily that k(yn)kℓ2(E) b⊗F (resp. k(zn)kE b⊗ℓ2(F )) can be identified with (cid:13)(cid:13)(cid:13)(cid:13)(cid:16)X y∗ nyn(cid:17)1/2(cid:13)(cid:13)(cid:13)(cid:13)E b⊗F(cid:18)resp. (cid:13)(cid:13)(cid:13)(cid:13)(cid:16)X znz∗ n(cid:17)1/2(cid:13)(cid:13)(cid:13)(cid:13)E b⊗F(cid:19) . Thus, Theorem 5 extends the case p = 1 of Theorem 4. The proof of Theorem 5 is based on a factorization property of functions of H 1(∆, Eb⊗F ) as a convex hull of tensor products of bounded sets of functions in H 2(E) and H 2(F ). From the viewpoint of Harmonic Analysis the space H p(∆, B) is nothing but the B-valued case of the H p- space associated to the compact group ∆ with its dual group ZN ordered by the lexicographical order, as is explained for instance in the chapter devoted to compact groups with ordered duals in [28]. It is easy to see that if g, h are complex valued functions in H 2(∆), then the pointwise product gh is in H 1(∆) with kghk1 ≤ kgk2khk2. Since the methods described in Rudin's book, (which are due to Helson -- Lowdenslager) extend to the matrix valued case (cf. [7]) they allow us to trivially modify the proofs of the Appendix B in the paper [22] (cf. also [23]) to obtain the following statement (which could alternatively be phrased as a factorization of H 1-functions with values in the space of all nuclear operators from E∗ into F ). Theorem 6. Let E, F be Banach spaces for type 2. Then the natural product mapping defines canonically a surjective norm one mapping H 2(∆) × H 2(∆) → H 1(∆) QE,F : H 2(∆, E)b⊗H 2(∆, F ) → H 1(∆, Eb⊗F ). More explicitly, there is a constant C such that for any f in H 1(∆, Eb⊗F ) there are functions gm ∈ H 2(∆, E) and hm ∈ H 2(∆, F ) such that ∀z ∈ ∆ and (13) f (z) = gm(z) ⊗ hm(z) ∞X1 X kgmkH 2(∆,E)khmkH 2(∆,F ) ≤ Ckf kH 1(∆,E b⊗F ). 9 Using this result, it is not hard to complete the Proof of Theorem 5. We first note that there is obviously a norm one inclusion R(E)b⊗F + Eb⊗R(F ) → R(Eb⊗F ). converse assuming (as we clearly may) that B is a complex Banach space. We will work with the complex version of the series (2). (The corresponding series are sometimes called Steinhaus series.) Equivalently, for every sequence (xn) of the form (11), the series P εnxn is in R(Eb⊗F ) with norm kP εnxnkR(E b⊗F ) ≤ 2P λn. This is immediate from the definitions. We will prove the Consider xn ∈ Eb⊗F such that the series S = P εnxn converges a.s. in Eb⊗F and consider the function f : ∆ → B defined by eitnxn. f (t0, t1, . . .) =Xn≥0 Z∆ kf kdm ≤ 2kSkL1(E b⊗F ). Then, it is easy to check (working with finite sums) that f converges a.s. and that (14) Moreover (since f is a polynomial of degree one in each complex variable) f is clearly in H 1(∆, B). By Theorem 6, we can find gm, hm such that (13) holds. Now let us denote A simple calculation shows that dnf = Enf − En−1f. (15) (16) Let us set xn =Z e−itnf (t0, t1, . . .)dm(t) =Xm Z e−itn gm ⊗ hmdm(t) =Xm Z e−itn [dngm ⊗ hm + gm ⊗ dnhm]dm(t). ym n (t) = e−itn dngm(t), zm(t) = hm(t), ym(t) = gm(t), zm n (t) = e−itndnhm(t). n (t) ∈ L2(∆, dm; R(E)) with ≤ 2K(E)kgmkH 2(∆,E), (17) andP εnzm (18) Then by (10) (the analytic UMD property) we haveP εnym 2 n (t) ∈ L2(∆, dm; R(F )) with R(E) dm(t)(cid:19)1/2 (cid:18)Z (cid:13)(cid:13)(cid:13)X εnym n (t)(cid:13)(cid:13)(cid:13) dm(t)(cid:19)1/2 (cid:18)Z (cid:13)(cid:13)(cid:13)X εnzm n (t)(cid:13)(cid:13)(cid:13) h (t)! ⊗ zm(t) + ym(t) ⊗ Xn εnxn =Xn Z " Xn εnym 2 R(F ) Xn therefore it follows from (16) that we have 10 ≤ 2K(F )khmkH 2(∆,F ), n (t)!# dm(t) εnzm so that recalling (13), (14), (17) and (18) we find (cid:13)(cid:13)(cid:13)X εnxn(cid:13)(cid:13)(cid:13)R(E) b⊗F +E b⊗R(F ) ≤ (2CK(E) + 2CK(F ))kf kH 1(∆,E b⊗F ) ≤ 4(CK(E) + CK(F ))kSkR(E b⊗F ) so that we obtain the announced converse inclusion R(Eb⊗F ) ⊂ R(E)b⊗F + Eb⊗R(F ), and this concludes the proof. Final Remarks. (i) Note that Theorem 5 applies in the case E = Lp, F = Lq and 2 ≤ p, q < ∞. We do not know what happens for 1 < p, q < 2 or even in the case E = Lp, F = L2 for 1 < p < 2. (ii) Similarly it is easy to deduce from Theorem 5 that the space Lpb⊗Lq is of cotype max(p, q) if 2 ≤ p, q < ∞. However it remains an open problem whether Lpb⊗Lq is of finite cotype (cotype 2?) for 1 < p < 2 and 1 < q ≤ 2. For the case p = q = 2, the result goes back to [31]. (iii) The preceding Theorem 5 raises many other natural questions. One fascinating point is that (12) appears like a derivation formula. It seems that R operates on the tensor product exactly as a derivation does on a product of noncommuting objects. Therefore, it is natural to ask the following two questions: Is there a similar result for R(Eb⊗Fb⊗G)? The guess is that we should find the sum The most interesting open case is the case E = F = G = ℓ2. It is a long standing R(E)b⊗Fb⊗G + Eb⊗R(F )b⊗G + Eb⊗Fb⊗R(G). We can also ask about "second derivatives": open question whether the space Eb⊗Fb⊗G has finite cotype in this case. Is there an analogous result for R[R(Eb⊗F )]? Here the guess is that we should have a formula analogous to the second derivative. We must distinguish the first R (associated to a first sequence (ǫ1 (associated to a second sequence (ǫ2 R1 and R2 respectively. Then the guess is that we should have n)) and the second one n), independent of the first one) so let us denote them by R1(R2(Eb⊗F )) ≈ R1(R2(E))b⊗F + R2(E)b⊗R1(F ) + R1(E)b⊗R2(F ) + Eb⊗R1(R2(F )). Even in the case E = F = ℓ2, we could not check this. 11 References [1] J. Bourgain, Vector valued singular integrals and the H 1 -- BMO duality, Probability Theory and Harmonic Analysis, Chao and Woyczynski (eds.), Marcel Dekker, (1986), pp. 1 -- 19. [2] D. Burkholder, A geometric characterization of Banach spaces in which martingale difference sequences are unconditional, Ann. Probab. 9 (1981), 997 -- 1011. [3] K. Carne, The algebra of bounded holomorphic margingales, J. Funct. Anal. 45 (1982), 95 -- 108. [4] G. Edgar, Analytic martingale convergence, J. Funct. Anal. 69 (1986), 268 -- 280. [5] X. Fernique, Int´egrabilit´e des vecteurs gaussiens, C.R. Acad. Sci. Paris A 270 (1970), 1698 -- 1699. [6] D.J.H. Garling, On martingales with values in a complex Banach space, Proc. Cambridge Phil. Soc. 104 (1988), 399 -- 406. [7] H. Helson and D. Lowdenslager, Prediction theory and Fourier series in several variables, Acta Math. 99 (1958), 165 -- 202. [8] U. Haagerup and G. Pisier, Factorization of analytic functions with values in noncommutative L1-spaces, Canadian J. Math. 41 (1989), 882 -- 906. [9] K. Ito and M. Nisio, On the convergence of sums of independent Banach space valued random variables, Osaka J. Math. 5 (1968), 35 -- 48. [10] J.P. Kahane, Some random series of functions, (1968), Heath Mathematical Monographs, Sec- ond Edition, Cambridge University Press, (1985). [11] S. Kwapie´n, A theorem on Rademacher series with vector valued coefficients, in Probability in Banach Spaces, Springer Lecture Notes, no. 526, (1976), 157 -- 158. [12] S. Kwapie´n, Isomorphic characterizations of inner product spaces by orthogonal series with vector coefficients, Studia Math. 44 (1972), 583 -- 595. [13] H. Landau and L. Shepp, On the supremum of a Gaussian process, Sankhya A 32 (1970), 369 -- 378. [14] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces II, Springer Verlag, 1979. [15] F. Lust-Piquard, In´egalit´es de Khintchine dans Cp, (1 < p < ∞), C.R. Acad. Sci. Paris 303 (1986), 289 -- 292. [16] F. Lust-Piquard, A Grothendieck factorization theorem on 22-convex Schatten spaces. Israel J. Math. 79 (1992), 331365. 12 [17] F, Lust-Piquard and G. Pisier, Noncommutative Khintchine and Paley inequalities, Ark. Mat. 29 (1991), no. 2, 241 -- 260. [18] B. Maurey, Type et cotype dans les espaces munis de structure locale inconditionnelle, S´eminaire Maurey -- Schwartz 73-75, Expos´e no. 24 -- 25, Ecole Polytechnique, Paris. [19] B. Maurey and G. Pisier, S´eries de variables al´eatoires vectorielles ind´ependantes et propri´et´es g´eom´etriques des espaces de Banach, Studia Math. 58 (1976), 45 -- 90. [20] G. Pisier, Probabilistic methods in the geometry of Banach spaces, CIME Summer School, June 1985, Springer Lecture Notes, no. 1206, (1986), 167 -- 241. [21] G. Pisier, Factorization de fonctions analytiques `a valeurs op´erateurs, C.R. Acad. Sci. Paris 307 (1988), 955 -- 960. [22] G. Pisier, Factorization of operator valued analytic functions, Adv. Math. 93 (1992), 61 -- 125. [23] G. Pisier, Factorization of operator valued analytic functions and complex interpolation, Festschrift in honor of I. Piatetski-Shapiro, Part II S. Gilbert, R. Howe and P. Sarnak (eds.), Weizmann Science Press of Israel (1990), pp. 197 -- 220. [24] G. Pisier, Holomorphic semi-groups and the geometry of Banach spaces, Ann. Math. 115 (1982), 375 -- 392 [25] G. Pisier, The dual J ∗ of the James space has cotype 2 and the Gordon -- Lewis property, Proc. Cambridge Phil. Soc. 103 (1988), 323 -- 331. [26] G. Pisier and Q. Xu, Random series in the real interpolation spaces between the spaces Vp, GAFA, Springer Lecture Notes, no. 1267, (1987), 185 -- 209. [27] J.L. Rubio de Francia, Martingale and integral transforms of Banach space valued functions, Springer Lecture Notes, no. 1221, (1985), 195 -- 222. [28] W. Rudin, Fourier Analysis on Group, Interscience, New York, 1962. [29] D. Sarason, Generalized interpolation in H ∞, Trans. Amer. Math. Soc. 127 (1967), 179 -- 203. [30] M. Talagrand, Regularity of Gaussian processes, Acta Math. 159 (1987), 99 -- 149. [31] N. Tomczak-Jaegermann, On the moduli of convexity and smoothness and the Rademacher averages of the trace classes Sp (1 ≤ p < ∞), Studia Math. 50 (1974), 163 -- 182. [32] N. Varopoulos, The Helson-Szego theorem and Ap-functions for Brownian motion and several variables, J. Funct. Anal. 39 (1980), 85 -- 121. [33] N. Varopoulos, Probabilistic approach to some problems in complex analysis, Bull. Sci. Math. 105 (1981), 181 -- 224. 13
1609.09648
1
1609
2016-09-30T09:35:04
Separably injective $C_\sigma$-spaces
[ "math.FA" ]
We show that a (complex) $C_\sigma$-space is separably injective if and only if it is linearly isometric to the Banach space $C_0(\Omega)$ of complex continuous functions vanishing at infinity on a substonean locally compact Hausdorff space $\Omega$.
math.FA
math
SEPARABLY INJECTIVE Cσ -SPACES CHO-HO CHU AND LEI LI Abstract. We show that a (complex) Cσ-space is separably injective if and only if it is linearly isometric to the Banach space C0(Ω) of complex continuous functions vanishing at infinity on a substonean locally compact Hausdorff space Ω. 1. Introduction Recently, separably injective Banach spaces have been studied in depth by Avil´es, Cabello S´anchez, Castillo, Gonz´alez and Moreno in [2, 3, 4], where one can find a number of interesting examples of these spaces despite the scarcity of examples of injective Banach spaces. In contrast to the fact that 1-injective Banach spaces are isometric to the Banach space C(Ω) of continuous functions on a compact Hausdorff space Ω [12, 14], 1-separably injective Banach spaces need not be complemented in C(Ω). In view of this, a natural question arises: when is a 1-complemented subspace of C(Ω) separably injective? We address this question in this paper. A Banach space V is 1-separably injective if every continuous linear map T : Y −→ V on a closed subspace Y of a separable Banach space Z admits a norm preserving extension to Z. It is known that C(Ω) itself is separably injective if and only if Ω is an F -space [2]. Abelian C*-algebras with identity are of the form C(Ω) for some Ω. The ones without identity can be represented as the algebra C0(S) of complex continuous functions vanishing at infinity on a locally compact Hausdorff space S and it has been shown lately that C0(S) is separably injective if and only if S is substonean [7]. Following [11], we call S substonean if any two disjoint open σ-compact subsets of S have disjoint compact closures. The compact substonean spaces are exactly the F-spaces defined in [8, 18]. However, infinite discrete spaces are F-spaces without being substonean. We refer to [13, Example 5] for an example of a substonean space which is not an F-space. Noting that the class of 1-complemented subspaces of C(Ω) is identical to that of Cσ-spaces [15, Theorem 3] (see also Remark 2.1), our question amounts to asking for a characterisation of separably injective Cσ-spaces. We give a complete answer by showing that a Cσ-space is separably injective if and only if it is linearly isometric to the function space C0(S) on a substonean locally compact Hausdorff space S. In what follows, all Banach spaces are over the complex field and we will denote by C0(K) the C*-algebra of complex continuous functions vanishing at infinity on a locally compact Hausdorff space K. If K is compact, we omit the subscript 0. Given a function g ∈ C0(K), we denote by g the complex conjugate of g. 1991 Mathematics Subject Classification. Primary 46B20, 17C65, 54G05. Key words and phrases. Separably injective Banach space. Cσ-space. Substonean space. The second author is supported by NSF (China) research grants No.11301285. 1 2 CHO-HO CHU AND LEI LI Let T = {α ∈ C : α = 1} be the circle group. By a T-space, we mean a locally compact Hausdorff space K equipped with a continuous group action σ : (α, ω) ∈ T × K 7→ α · ω ∈ K. A complex Banach space is called a complex Cσ-space if it is linearly isometric to a function space of the form Cσ(K) for some T-space K, defined by Cσ(K) = {f ∈ C0(K) : f (σ(α, ω)) = αf (ω), ∀ω ∈ K}. We note that the definition of a complex Cσ-space in [16] requires K to be compact. To achieve our result, we make substantial use of the Jordan algebraic structure of Cσ(K). Indeed, although Cσ(K) lacks a C*-algebraic structure, it is equipped with a triple product {f, g, h} = f gh (f, g, h ∈ Cσ(K)) which turns it into a JB*-triple with many useful Jordan properties. First, let us give a brief introduction to JB*-triples which generalise C*-algebras. For further references and the geometric origin of JB*-triples, we refer to [6, 17, 19]. A complex Banach space V is a JB*-triple if it admits a continuous triple product {·, ·, ·} : V 3 −→ V which is symmetric and linear in the outer variables, but conjugate linear in the middle variable, and satisfies (i) {x, y, {a, b, c}} = {{x, y, a}, b, c} − {a, {y, x, b}, c} + {a, b, {x, y, c}}; (ii) the operator a a : x ∈ V 7→ {a, a, x} ∈ V has real numerical range and non-negative spectrum; (iv) ka ak = kak2 for a, b, c, x, y ∈ V . We always have k{a, b, c}k ≤ kakkbkkck and (i) is called the Jordan triple identity. A C*-algebra A is a JB*-triple with the triple product {a, b, c} = 1 2 (ab∗c + cb∗a) (a, b, c ∈ A). More generally, the range of a contractive projection on a C*-algebra is a JB*-triple (cf. [6, Theorem 3.3.1]), but not always a C*-algebra. An element e in a JB*-triple is called a tripotent if e = {e, e, e}. Tripotents in C*-algebras are exactly the partial isometries. A subspace W of a JB*-triple V is called a subtriple if a, b, c ∈ W implies {a, b, c} ∈ W . Closed subtriples of a JB*-triple are JB*-triples in the inherited norm and triple product. A triple ideal of V is a subspace J ⊂ V such that {a, b, c} ∈ J whenever one of a, b and c belongs to J. Given a closed triple ideal J ⊂ V , the quotient space V /J is a JB*-triple in the triple product {a + J, b + J, c + J} := {a, b, c} + J. Two elements a, b ∈ V are said to be orthogonal to each other if a b = 0, where a b is the continuous linear map x ∈ V 7→ {a, b, x} ∈ V . Two subspaces I, J ⊂ V are orthogonal if I J := {a b : a ∈ I, b ∈ J} = {0}. The bidual V ∗∗ of a JB*-triple V carries a natural structure of a JB*-triple, with a unique predual, in which the triple product is separately weak* continuous and the natural embedding of V into V ∗∗ identifies V as a subtriple of V ∗∗. Given a closed triple ideal I ⊂ V , the bidual I ∗∗ embeds as a weak* closed triple ideal in SEPARABLY INJECTIVE Cσ-SPACES 3 V ∗∗, which can be decomposed into an ℓ∞-sum V ∗∗ = I ∗∗ ⊕∞ J for some weak* closed triple ideal J ⊂ V ∗∗, orthogonal to I ∗∗ [6, lemma 3.3.16]. A linear map ϕ : V −→ W between two JB*-triples is called a triple homo- morphism if {ϕ(a), ϕ(b), ϕ(c)} = ϕ{a, b, c} for a, b, c ∈ V . The triple isomorphisms between V and W are exactly the surjective linear isometries (cf. [6, Theorem 3.1.7, Theorem 3.1.20]). A JB*-triple V is called abelian if its triple product satisfies {{x, y, z}, u, v} = {x, {y, z, u}, v} = {x, y, {z, u, v}} for all x, y, z, u, v ∈ V . An abelian C*-algebra is an abelian JB*-triple and so is Cσ(K) in the triple product defined above. In fact, Cσ(K) is a closed subtriple of C0(K). By [5, Lemma 2.2, Theorem 3.7], an abelian closed subtriple V of a C*-algebra admits a composition series (Jλ)0≤λ≤µ of closed triple ideals, indexed by ordinals λ, such that the quotient Jλ+1/Jλ is linearly isometric to an abelian C*-algebra, for λ < µ. We recall that (Jλ)0≤λ≤µ is called a composition series if J0 = {0}, Jµ = V and for a limit ordinal λ ≤ µ, the ideal Jλ is the closure of Sλ′<λ Jλ′ . 2. Jordan structure in Cσ-spaces We will make use of the abelian JB*-triple structure of Cσ-spaces to derive our result. To pave the way, we first present some detailed analysis of this structure. Let V be an abelian closed subtriple of a C*-algebra (e.g. a Cσ-space) in this section. One can consider it as as a subtriple of its bidual V ∗∗ via the natural embedding v ∈ V 7→ bv ∈ V ∗∗, where bv(ψ) = ψ(v) for ψ ∈ V ∗. By [5, 9], V ∗∗ is (isometric to and identified as) an abelian von Neumann algebra with identity denoted by 1 and involution z ∈ V ∗∗ 7→ z∗ ∈ V ∗∗. The triple product in V ∗∗ is given by {a, b, c} = ab∗c. Each ψ ∈ V ∗ can be viewed naturally as a functional of V ∗∗. If ψ is a positive functional of V ∗∗, then ψ(z∗) = ψ(z) for each z ∈ V ∗∗. A positive functional ψ ∈ V ∗ is called a normal state of V ∗∗ if ψ(1) = 1. It is called pure if it is an extreme point of the norm closed convex set of normal states in V ∗. Let S be the set of all pure normal states of V ∗∗, which are exactly the multiplica- tive normal states of V ∗∗. Given a projection p ∈ V ∗∗ and s ∈ S, we have s(p) = 0 or 1 since s(p) = s(p2) = s(p)2. If u ∈ V ∗∗ is unitary, then 1 = s(1) = s(u∗u) = s(u)2 for all s ∈ S. We equip S with the weak* topology of V ∗ and call it the pure normal state space of V ∗∗. The nonzero triple homomorphisms from V to C are exactly the set K = ext V ∗ 1 1 , where K ∪ {0} is weak*-compact [9, of extreme points of the dual unit ball V ∗ Proposition 2.3, Corollary 2.4] and S ⊂ K. For each ω ∈ K and tripotent c ∈ V ∗∗, we have ω(c) = ω(cc∗c) = ω(c)ω(c)2 which implies ω(c) = 0 or ω(c) ∈ T. We note that K is a T-space with the natural T-action σ : (α, ω) ∈ T × K 7→ αω ∈ K and we have K = {αs : α ∈ T, s ∈ S}. In fact, each ω ∈ K has a unique representation ω = αs for some α ∈ T and s ∈ S, where s = ω(1)ω. By [9, Theorem 1], the map (2.1) v ∈ V 7→ bvK ∈ Cσ(K) 4 CHO-HO CHU AND LEI LI is a surjective linear isometry, which enables us to identify V with the Cσ-space Cσ(K). Remark 2.1. Let π : C(Ω) −→ C(Ω) be a contractive projection. Then its image π(C(Ω)) is an abelian closed subtriple in some C*-algebra [10] and hence the previous discussion implies that it is a Cσ-space. For each a ∈ V \{0}, let V (a) be the JB*-subtriple generated by a in V . Then there is a surjective linear isometry and triple isomorphism (2.2) φ : C0(Sa) → V (a) ⊂ V which identifies V (a) with the abelian JB*-triple C0(Sa) of continuous functions vanishing at infinity on the triple spectrum Sa ⊂ (0, kak], where Sa ∪{0} is compact [6, Theorem 3.1.12]. Let φ∗∗ : C0(Sa)∗∗ → V (a)∗∗ be the bidual map and let i be the identity in the von Neumann algebra C0(Sa)∗∗. Then e = φ∗∗(i) is the identity in the von Neumann algebra V (a)∗∗ with product and involution given by x · y = {x, e, y}, x ∈ V (a)∗∗ 7→ {e, x, e} ∈ V (a)∗∗. While this abelian von Neumann algebraic structure of V (a)∗∗ will be assumed throughout, it should be noted that V (a)∗∗ need not be a subalgebra of V ∗∗ in its natural embedding. Nevertheless, V (a)∗∗ can always be considered as a subtriple of V ∗∗ and the identity e ∈ V (a)∗∗ is a tripotent in V ∗∗ satisfying {e, a, e} = {φ∗∗(i), φ∗∗(ιa), φ∗∗(i)} = φ∗∗{i, ιa, i} = a. (2.3) For each ρ ∈ K, viewed as a complex-valued triple homomorphism on V (a)∗∗, we have ρ(e) = 0 if and only if ρ(a) = ρ{e, e, a} = 0. The norm closed triple ideal Ja generated by a in V contains V (a) and is the norm closure of {a, V, a}. It has been shown in [5, Lemma 2.2] that Ja is linearly isometric to the abelian C*-algebra C0(Xa) of continuous functions vanishing at infinity on a locally compact Hausdorff space Xa. We need some detail here for later application. In fact, Ja is an abelian C*-algebra with the same product and involution of V (a)∗∗, and e ∈ V (a)∗∗ ⊂ J ∗∗ a , which can be a seen from the following computation using (2.3). is the identity of J ∗∗ {{a, x, a}, e, {a, y, a}} = {a, {x, {a, e, a}, y}, a} ∈ {a, V, a}; {e, {a, x, a}, e} = ea∗xa∗e = ee∗ae∗xe∗ae∗e = ae∗xe∗a = {x, a, a} ∈ Ja. Given ρ ∈ K = ext V ∗ 1 with ρ(e) = 1, its restriction ρJa ∈ J ∗ a . Conversely, each pure normal state ϕ of J ∗∗ a a and can be extended to an extreme point eϕ ∈ ext V ∗ 1 a is a pure normal is an extreme point of state of J ∗∗ the closed unit ball of J ∗ satisfying eϕ(e) = 1. Let denote the pure normal state space of J ∗∗ topology w(J ∗ a , Ja) of J ∗ a . Xa = {ρJa : ρ ∈ K, ρ(e) = 1} a which is locally compact in the weak* We note that for each ρ ∈ K, we have ρ(a) = 0 if and only if ρ(V (a)) = {0}, which in turn is equivalent to ρ(Ja) = ρ({a, V, a}) = {0}. Lemma 2.2. In the above notation, the set K(e) = {ρ ∈ K : ρ(e) = 1} with the relative weak* topology of V ∗ is homeomorphic to Xa = {ρJa : ρ ∈ K, ρ(e) = 1} in the topology w(J ∗ a , Ja). In particular, K(e) is weak* locally compact in K. SEPARABLY INJECTIVE Cσ-SPACES 5 Proof. We show that the restriction map ρ ∈ K(e) 7→ ρJa ∈ Xa is a homeomor- phism in these topologies. It is clearly continuous and surjective. Given ρ, ρ′ ∈ K such that ρJa = ρ′Ja , then ρ(a) = ρ′(a) 6= 0 since ρ(e) = ρ′(e) = 1. For any v ∈ V , we have {a, a, v} ∈ Ja and hence ρ(a)2ρ(v) = ρ{a, a, v} = ρ′{a, a, v} = ρ′(a)2ρ′(v), giving ρ(v) = ρ′(v). This proves injectivity of the map. Finally, to show that the inverse of the map is continuous, let (ργJa) be a net converging to ρJa ∈ Xa. Then again, for each v ∈ V , we have ργ(a) → ρ(a) 6= 0 and ργ(a)2ργ(v) = ργ{a, a, v} → ρ{a, a, v} = ρ(a)2ρ(v), which implies ργ(v) → ρ(v), proving continuity. (cid:3) Remark 2.3. The above lemma enables us to identify the pure normal state space Xa with K(e) and write Xa = {ρ ∈ K : ρ(e) = 1}. We retain the above notation in the sequel. 3. Separably injective Cσ-spaces We characterize separably injective Cσ-spaces in this section. Throughout, let V be a Cσ-space. We will identify V , as in the previous section, with the Cσ-space Cσ(K), where K = ext V ∗ 1 is the set of nonzero triple homomorphisms from V to C. Lemma 3.1. Let V be separably injective. Given a ∈ V of unit norm and the identity e ∈ V (a)∗∗, let K(e) = {ρ ∈ K : ρ(e) = 1}. Then there exists an element va ∈ V such that K(e) ⊂ Ka where and Ka is weak* compact in V ∗. Ka = {ρ ∈ K : ρ(va) = 1} Proof. Since Ka is weak* closed in K ∪ {0}, it is weak* compact. Let φ : C0(Sa) → V (a) ⊂ V be the embedding in (2.2), where the triple spectrum Sa ⊂ (0, 1] can be identified, via the evaluation map as usual, with the pure normal state space of C0(Sa)∗∗. Let χa be the constant function on Sa with value 1 and consider the separable subspace C0(Sa) + Cχa of ℓ∞(Sa). By separable injectivity of V , the embedding φ : C0(Sa) → V (a) ⊂ V admits a norm preserving extension Φ : C0(Sa)+Cχa → V . Let va = Φ(χa) ∈ V. To complete the proof, we show that ρ(va) = 1 for each ρ ∈ K(e). We first observe that the sequence (rn) of odd roots of the identity function ιa in C0(Sa) converges pointwise to the function χa. Let un = φ(rn) ∈ V (a). Let ρ ∈ K(e). Then the map ρ ◦ φ : C0(Sa) → C is a pure normal state of C0(Sa)∗∗. Hence we have ρ(un) = ρ ◦ φ(rn) ∈ [0, 1] and limn ρ(un) = 1. The norm preserving extension Φ satisfies kΦ(χa)k ≤ 1 and kΦ(χa) − 2unk = kΦ(χa) − 2φ(rn)k = kΦ(χa) − 2Φ(rn)k ≤ kχa − 2rnk ≤ 1. It follows that ρ(Φ(χa)) − 2ρ(un) ≤ 1 for all n, which implies Now ρ(Φ(χa)) ≤ 1 gives ρ(va) = ρ(Φ(χa)) = 1. (cid:3) ρ(Φ(χa)) − 2 ≤ 1. 6 CHO-HO CHU AND LEI LI Our next task is to show that a separably injective Cσ-space V is actually linearly isometric to an abelian C*-algebra. We adopt the following strategy. Since V is abelian, it has been noted in Section 1 that there is a composition series (Jλ)0≤λ≤µ of closed triple ideals in V such that for each ordinal λ < µ, the quotient Jλ+1/Jλ is linearly isometric to the C*-algebra C0(Xλ) of continuous functions vanishing at infinity on a locally compact Hausdorff space Xλ, and V ∗∗ is linearly isometric to λ<µ(Jλ+1/Jλ)∗∗ = Lℓ∞ the ℓ∞-sum Lℓ∞ λ<µ C0(Xλ)∗∗. By the uniqueness of predual, V ∗ is linearly isometric to the ℓ1-sum ℓ1 M λ<µ (Jλ+1/Jλ)∗ = ℓ1 M λ<µ C0(Xλ)∗. Given that V is separably injective, we will refine this construction to show that V is isometric to the abelian C*-algebra C0(Xλ). c0M λ<µ Let V be separably injective and let a ∈ V be of unit norm. Consider the closed subtriple V (a) generated by a in V as well as the closed triple ideal Ja, the latter is linearly isometric to the C*-algebra C0(Xa) as shown before, where Xa = {ρ ∈ K : ρ(ea) = 1} is weak* locally compact by Lemma 2.2 and Remark 2.3, and ea is the identity of the von Neumann algebra J ∗∗ a . By separable injectivity and Lemma 3.1, there exists va ∈ V such that ρ(va) = 1 for each ρ ∈ Xa. If Ja = V , then we are done. Otherwise we have the ℓ∞-sum V ∗∗ = J ∗∗ a ⊕∞ (J ∗∗ a ) where (J ∗∗ a ) is, J ∗∗ (J ∗∗ a identifies (J ∗∗ is a nonzero weak* closed triple ideal in V ∗∗, orthogonal to J ∗∗ a , that a ) = {0} (cf. [6, Lemma 3.3.16]). The quotient map V ∗∗ → V ∗∗/J ∗∗ a a ) with the quotient V ∗∗/J ∗∗ a and we can write (3.1) V ∗∗ = J ∗∗ a ⊕∞ (V ∗∗/J ∗∗ a ) = C0(Xa)∗∗ ⊕∞ (V /Ja)∗∗. We have the ℓ1-sum V ∗ = C0(Xa)∗ ⊕1 (V /Ja)∗ where Ja is an M-ideal in V . Consider the quotient map x ∈ V 7→ [x] := x + Ja ∈ [V ] := V /Ja which maps the closed unit ball of V onto the closed unit ball of V /Ja (cf. [1, Corollary 5.6]). Pick [b] = b + Ja ∈ [V ] with unit norm. Let V ([b]) and J[b] be respectively the closed subtriple and triple ideal generated by [b] in [V ]. We can repeat the previous arguments in the setting V ([b]) ⊂ J[b] ⊂ [V ] to deduce that J[b] is linearly isometric to some abelian C*-algebra C0(X[b]) with X[b] = {θ ∈ ext [V ]∗ 1 : θ(e[b]) = 1} [b] ⊂ [V ]∗∗ = V ∗∗/J ∗∗ a ) with e[b] = eeb + J ∗∗ where X[b] is locally compact in the weak* topology of [V ]∗ by Lemma 2.2 and e[b] is the identity of J ∗∗ a , which identifies with a tripotent eeb ∈ (J ∗∗ a . Moreover, the quotient JB*-triple [V ] = V /Ja is abelian and seprably injective by [2, Proposition 4.6], and hence Lemma 3.1 implies that there exists v[b] = evb + Ja ∈ [V ] = V /Ja such that θ(v[b]) = 1 for each θ ∈ X[b]. 1 consists of nonzero complex-valued triple homomorphisms on [V ], which can be lifted to nonzero complex triple homomorphisms on V via the The set ext [V ]∗ SEPARABLY INJECTIVE Cσ-SPACES 7 quotient map and we have where ¯ρ(x + Ja) := ρ(x). Hence we have ext [V ]∗ 1 = { ¯ρ : ρ ∈ K, ρ(Ja) = {0}} X[b] = { ¯ρ : ρ ∈ K, ρ(a) = 0, ρ(eeb) = 1} and ¯ρ ∈ X[b] implies ρ(b) = ¯ρ([b]) 6= 0 and ρ(evb) = 1. A weak* convergent net in [V ]∗ lifts to a weak* convergent net in V ∗ via the quotient map. Considering X[b] ⊂ C0(X[b])∗ = J ∗ [b] and Lemma 2.2, we see that a net (¯ργ) in X[b] converges to ¯ρ ∈ X[b] in the weak* topology of C0(X[b])∗ if and only if (ργ) weak* converges to ρ in K. The homeomorphism ¯ρ ∈ X[b] 7→ ρ ∈ {ρ ∈ K : ρ(a) = 0, ρ(eeb) = 1} enables us to identify these two spaces. We note that TXa ∩ TX[b] ⊂ TXa ∩ ext [V ]∗ (3.2) 1 = ∅ where ρ(a) 6= 0 for all ρ ∈ Xa. In the ℓ1-sum V ∗ = J ∗ (3.3) each ω ∈ V ∗ admits a decomposition ω = ω1 + ω2 in V ∗ with ω2(Ja) = {0} and kωk = kω1Ja k + kω2k, which provides the identification of ω as an element (eω1, eω2) a ⊕1 (V /Ja)∗ = C(Xa)∗ ⊕1 [V ]∗, in the ℓ1-sum, defined by eω1 = ω1Ja ∈ J ∗ a Hence for an extreme point ω ∈ ext V ∗ and eω2([·]) = ω2(·) ∈ (V /Ja)∗. 1 = K, we have eω1 = 0 or eω2 = 0. a , Ja)-topology and the net (eω2 γ) converges to eω1 in the w(J ∗ Given a net (ωγ) in V ∗ weak* converging to a limit ω ∈ V ∗, it can be seen that γ) converges the net (eω1 to eω2 in the weak* topology of (V /Ja)∗. In particular, if the net (ωγ) is in K and γ) to eωj implies that ω ∈ K with eωj 6= 0 (j ∈ {1, 2}), then the convergence of (eωj γ 6= 0 eventually, and hence eωj ′ eωj The closed unit ball of the ℓ1-sum in (3.3) has extreme points γ = 0 for j′ 6= j eventually. (TXa, 0) ∪ (0, ext [V ]∗ 1) := {(ω, 0) : ω ∈ TXa} ∪ {(0, ω) : ω ∈ ext [V ]∗ 1} and in the identification of (3.2), we have the disjoint union K = TXa ∪ ext [V ]∗ 1. Given a net (ωγ) in K weak* converging to some ω ∈ K, and given either ω ∈ TXa or ω ∈ ext [V ]∗ 1, the above observation implies that ωγ belongs to the same set eventually. Observe that [Jb] = J[b] = (Jb + Ja)/Ja and if J[b] 6= [V ], we have the ℓ∞-sum V ∗∗ = J ∗∗ a ⊕∞((Jb+Ja)/Ja)∗∗⊕∞([V ]/J[b])∗∗ = C0(Xa)∗∗⊕∞C0(X[b])∗∗⊕∞([V ]/J[b])∗∗ where the quotient JB*-triple [[V ]] := [V ]/J[b] is separably injective, ρ(va) = 1 for ρ ∈ Xa and ρ(evb) = 1 for ρ ∈ X[b]. The closed unit ball of the ℓ1-sum V ∗ = C0(Xa)∗ ⊕1 C0(X[b])∗ ⊕1 ([V ]/J[b])∗ has extreme points (TXa, 0, 0) ∪ (0, TX[b], 0) ∪ (0, 0, ext [[V ]]∗ 1) and we have the disjoint union K = ext V ∗ 1. Given a net (ωγ) in K weak* converging to some ω ∈ K, if ω belongs to one of the three 1 = TXa ∪ TX[b] ∪ ext [[V ]]∗ 8 CHO-HO CHU AND LEI LI sets above, then repeating the arguments as before, ωγ belongs to the same set eventually. Now transfinite induction together with separable injectivity yields a composition series (Jλ)0≤λ≤µ of closed triple ideals in V , with vλ ∈ V , such that V ∗ is linearly isometric to, and identifies with, the ℓ1-sum ℓ1 M (Jλ+1/Jλ)∗ = λ<µ ℓ1 M λ<µ C0(Xλ)∗ where ρ(vλ) = 1 for each ρ ∈ Xλ, and the pure normal state space Xλ of (Jλ+1/Jλ)∗∗ identifies with the set in which eeλ is the identity of (Jλ+1/Jλ)∗∗. In this identification, we have the disjoint union {ρ ∈ K : ρ(Jλ) = {0}, ρ(eeλ) = 1} K = [ TXλ λ<µ and for a weak* convergent net (ωγ) in K with limit ω ∈ TXλ for some λ, we have (ωγ) in TXλ eventually. As a consequence of Lemma 2.2, in the identification Xλ ⊂ C0(Xλ)∗ and Xλ ⊂ K, the weak* convergence in C0(Xλ)∗ of a net (ργ) in Xλ to ρ ∈ Xλ is the same as the weak* convergence in K. Lemma 3.2. Given that V is separably injective and in the above notation, the subset Xλ ∪ {0} of K ∪ {0} is weak* compact for all λ < µ and also, TXλ is relatively weak* open in K ∪ {0}. Proof. Let (ργ) be a net in Xλ weak* converging to a nonzero limit ω ∈ V ∗. Then ω ∈ K and by the above remark, we must have ω ∈ TXλ, say ω = αρ with α ∈ T and ρ ∈ Xλ. Since Xλ is contained in the weak* compact set {ρ′ ∈ K : ρ′(vλ) = 1}, it follows that α = αρ(vλ) = limγ ργ(vλ) = 1 and ω = ρ ∈ Xλ. This proves that Xλ ∪ {0} is weak* closed in K ∪ {0} and hence weak* compact. For the second assertion, let (ωγ) be a net in (K ∪ {0})\TXλ weak* converging to some ω ∈ K. Then again ω /∈ TXλ for otherwise, the previous remark implies that ωγ belongs to TXλ eventually which is impossible. (cid:3) The above construction enables us to show that a separably injective Cσ-space is isometric to an abelian C*-algebra. Theorem 3.3. Let V be a separably injective Cσ-space. Then V is linearly iso- metric to an abelian C*-algebra. Proof. Let K = ext V ∗ 1 and as shown previously, we have the ℓ1-sum V ∗ = ℓ1 M λ<µ C0(Xλ)∗ with the disjoint union K = Sλ<µ TXλ. For each λ < µ, there is an element vλ ∈ V such that ρ(vλ) = 1 for all ρ ∈ Xλ. We show that V is linearly isometric to the c0-sum C0(Xλ) which would complete the proof. c0M λ<µ We continue to identify V with Cσ(K) in (2.1). By Lemma 3.2, each f ∈ Cσ(K) restricts to a continuous function f Xλ ∈ C0(Xλ). SEPARABLY INJECTIVE Cσ-SPACES 9 We show that the map f ∈ V ≈ Cσ(K) 7→ (f Xλ) ∈ c0M λ<µ C0(Xλ) is a surjective linear isometry. To see that (f Xλ) indeed belongs to the c0-sum, we need to show (kf Xλk) ∈ c0(Λ), where Λ = [0, µ). Let ε > 0. By Lemma 3.2, for each λ ∈ Λ, the set TXλ is relatively weak* open in K ∪ {0}. Since f vanishes at infinity on K, the set Uε = {ω ∈ K : f (ω) < ε} ∪ {0} is a relatively weak* open neighbourhood of 0 in the one-point compactification K ∪ {0} of K. We have K ∪ {0} = [ TXλ ∪ Uε and by weak* compactness, there are finitely many λ1, . . . , λn such that λ∈Λ K ∪ {0} ⊂ TXλ1 ∪ · · · ∪ TXλn ∪ Uε. It follows that kf Xλk ≤ ε for λ /∈ {λ1, . . . , λn} which proves (kf Xλk) ∈ c0(Λ). Since K = ∪λT Xλ and each f ∈ Cσ(K) satisfies f (αω) = αf (ω) for α ∈ T and ω ∈ K, it is evident that the map is a linear isometry. It remains to show that the map is surjective. Let (gλ) ∈ function f : K → C by c0M λ∈Λ C0(Xλ). Define a f (ω) = αgλ(ρλ) for ω = αρλ ∈ TXλ. The function f is well-defined since the sets {TXλ}λ are mutually disjoint and each ω ∈ K has a unique representation ω = αρ ∈ TXλ for if αρ = βσ ∈ TXλ, we have α = αρ(vλ) = βσ(vλ) = β, where Xλ is contained in the weak* compact set {ρ′ ∈ K : ρ′(vλ) = 1}. We complete the proof by showing f ∈ Cσ(K). We have readily f (αω) = αf (ω) for α ∈ T and ω ∈ K. For continuity, let (ωγ) be a net weak* converging to ω ∈ K and say, ω = αρ ∈ TXλ for some λ. By a previous remark, the net (ωγ) is in TXλ eventually. Therefore we have ωγ = αγργ with αγ ∈ T and ργ ∈ Xλ eventually. It follows that eventually αγ = αγργ(vλ) → αρ(vλ) = α and ργ → ρ. Hence we have lim γ f (ωγ) = lim γ αγgλ(ργ) = αgλ(ρ) = f (ω). Finally, for any ε > 0, there are finitely many λ1, . . . , λn such that kgλk < ε for λ /∈ {λ1, . . . , λn}. For each λj with j = 1, . . . , n, there is a weak* compact set Ej ⊂ Xλj such that {ρ ∈ Xλj : gλj (ρ) ≥ ε} ⊂ Ej. This gives T{ρ ∈ Xλj : gλj (ρ) ≥ ε} ⊂ ∪n {ω ∈ K : f (ω) ≥ ε} = ∪n j=1 TEj ⊂ K j=1 where the finite union ∪j TEj is weak* compact and therefore f ∈ C0(K). (cid:3) Finally, by the characterisation of separably injective abelian C*-algebras in [7, Theorem 3.5], together with Theorem 3.3, we conclude with the following main result of the paper. 10 CHO-HO CHU AND LEI LI Theorem 3.4. Let V be a Cσ-space. The following conditions are equivalent. (i) V is separably injective. (ii) V is linearly isometric to the Banach space C0(S) of complex continuous functions vanishing at infinity on a substonean locally compact Hausdorff space S. References [1] E.M. Alfsen and E.G. Effros, Structure in real Banach spaces, Ann. of Math. 96 (1972) 98-173. [2] A. Avil´es, F. Cabello S´anchez, J. M.F. Castillo, M. Gonz´alez and Y. Moreno, On separably injective Banach spaces, Adv. Math. 234 (2013) 192 -- 216. [3] A. Avil´es, F. Cabello S´anchez, J. M.F. Castillo, M. Gonz´alez and Y. Moreno, ℵ-injective Banach spaces and ℵ-projective compacta, Rev. Mat. Iberoamericana, 31 (2015), 575 -- 600. [4] A. Avil´es, F. Cabello S´anchez, J. M.F. Castillo, M. Gonz´alez and Y. Moreno, Separably injective Banach spaces, Lecture Notes in Math. 2132, Springer, 2016. [5] L.J. Bunce and C-H. Chu, Real contractive projections on commutative C*-algebras, Math. Z. 226 (1997) 85-101. [6] C-H. Chu, Jordan structures in geometry and analysis, Cambridge Tracts in Math. 190, Cambridge Univ. Press, Cambridge, 2012. [7] C-H. Chu and L. Li, Separably injective C*-algebras, Arch. Math. 106(2016), 553 -- 559. [8] L. Gillman and M. Jerison, Rings of continuous functions, Van Nostrand, Princeton, 1960. [9] Y. Friedman and B. Russo, Function representation of commutative operator triple systems, J. London Math. Soc. 27(1983), 513-524. [10] Y. Friedman and B. Russo, Solution of the contractive projection problem, J. Funct. Anal. 60 (1985) 56-79. [11] K. Grove and G.K. Pedersen, Sub-Stonean spaces and corona sets, J. Funct. Anal., 56 (1984), 124 -- 143. [12] M. Hasumi, The extension property of complex Banach spaces, Tohoku Mathematical Jour- nal, 10(2) (1958), 135 -- 142. [13] M. Henriksen and R. G. Woods, F-spaces and substonean spaces: general topology as a tool in functional analysis, Papers on general topology and related category theory and topological algebra (New York, 1985/1987), 60 -- 68, Ann. New York Acad. Sci., 552, New York Acad. Sci., New York, 1989. [14] J.L. Kelley, Banach spaces with the extension property, Trans. Amer. Math. Soc. 72 (1952) 323-326. [15] J. Lindenstrauss and D.E. Wulbert, On the classification of the Banach spaces whose duals are L1 spaces, J. Funct. Anal. 4 (1969) 332-349. [16] G.H. Olsen, On the classification of complex Lindenstrauss spaces, Math. Scand. 35 (1974) 237 -- 258. [17] B. Russo, Structure of JB*-triple, in Proc. Oberwolfach Conf. on Jordan algebras, 1992, eds. W. Kaup, K. McCrimmon, H.P. Petersson (Walter de Gruyter, Berlin, 1994) 209-280. [18] G.L. Seever, Measures on F-spaces, Trans. Amer. Math. Soc. 133 (1968), 267-280. [19] H. Upmeier, Symmetric Banach manifolds and Jordan C*-algebras (North Holland Math. Studies 104) North Holland, Amsterdam 1985. [20] M. Zippin, The separable extension problem, Israel J. Math. 26 (1977) 372-387. School of Mathematical Sciences, Queen Mary, University of London, London E1 4NS, UK E-mail address: [email protected] School of Mathematical Sciences and LPMC, Nankai University, Tianjin 300071, China E-mail address: [email protected]
1904.00174
1
1904
2019-03-30T08:48:38
Barrier functions in the subdifferential theory
[ "math.FA" ]
We present a new method for proving Correa-Jofr\'e-Thibault theorem that monotonicity of subdifferential implies convexity of the function. This new method is based on barrier functions. Barrier functions help overcome some of the main technical difficulties when working with lower semicontinuous functions.
math.FA
math
Barrier functions in the subdifferential theory Milen Ivanov∗ Nadia Zlateva† Radiant Life Technologies Ltd. Faculty of Mathematics and Informatics Nicosia, Cyprus Sofia University [email protected] 5, James Bourchier Blvd. 1164 Sofia, Bulgaria [email protected] Abstract We present a new method for proving Correa-Jofr´e-Thibault theo- rem that monotonicity of subdifferential implies convexity of the func- tion. This new method is based on barrier functions. Barrier functions help overcome some of the main technical difficulties when working with lower semicontinuous functions. 2010 Mathematics Subject Classification: 49J52, 47H05, 52A41. Keywords: subdifferential, monotonicity, convex function, barrier function 1 Introduction In 1990's Correa, Jofr´e and Thibault in series of papers proved that a convex lower semicontinuous function can be characterized by monotonicity property of its subdifferential -- in reflexive Banach space for the Clarke subdifferential in [2] and in any Banach space for axiomatically introduced subdifferential in [3] and for more general axiomatic presubdifferential in [4]. ∗Supported by Bulgarian National Scientific Fund under grant KP-06-22/4. †Supported by Scientific Fund of Sofia University under grant for 2019. 1 The main tool for proving this characterization is the Mean Value The- orem of Zagrodny [13] which holds for a lower semicontinuous function in a Banach space for any presubdifferential, see [11]. Jules and Lassonde prove a subdifferential test for optimality, see [7] of Minty type [8] involving a subdifferential satisfying certain axioms. In setting up the axiomatic framework we follow [12], but we pick the apparently minimal set of axioms under which proofs can work. In this way our results are slightly more general. As mentioned below, adding another natural axiom can significantly simplify the presentation. We work in a real Banach space X with dual X ∗. Definition 1 (axioms for subdifferential). Multi-valued operator ∂ which associates to any function f : X → R ∪ {+∞} and any x ∈ X a (possibly empty) subset ∂f (x) ⊂ X ∗ is feasible subdifferential if (P1) ∂f (x) = ∂cf (x) whenever f is a convex and continuous function in a neighbourhood U of x, where ∂c stands for the Fenchel subdifferential, i.e. ∂cf (x) := {x∗ ∈ X ∗ : hx∗, y − xi ≤ f (y) − f (x), ∀y ∈ U}; (P2) For f lower semicontinuous and g convex and continuous in a neigh- bourhood of x ∈ dom f , 0 ∈ ∂g(x) + lim sup ∂f (y) y→x, f (y)→f (x) (1) whenever x is a local minimum point of f + g. In more details (1) means that there are y ∗ n ∈ ∂f (yn) such that yn → x, n converges in the w∗ topology to some y ∗ f (yn) → f (x) and the sequence y ∗ such that −y ∗ ∈ ∂g(x). Note that all presubdifferentials considered by Correa, Jofr´e and Thibault in [2, 3, 4], as well as subdifferentials considered by Jules and Lassonde in [7] are feasible subdifferentials in the sense of Definition 1. In terms of Ioffe's extensive classification, see [6], (P1) is called contiguity, while (P2) is weak-star form of trustworthness. We prove the above mentioned results for feasible subdifferentials and in a different and unified way -- by using barrier functions instead of (a variant of) Zagrodny Theorem. 2 For convenience of notation we often identify the map ∂f : X ⇒ X ∗ with its graph, that is, (x, x∗) ∈ ∂f is a shorthand for x∗ ∈ ∂f (x). The main contribution of this work is a new method, based on [5], see also [10, p.569], for proving the following result of Correa, Jofr´e and Thibault. Theorem 2 (Correa-Jofr´e-Thibault). Let X be a Banach space and let ∂ be a feasible subdifferential. Let f : X → R ∪ {+∞} be a proper lower semicontinuous function. If ∂f is monotone, then f is convex. Recall that monotonicity of ∂f means (xi, x∗ i ) ∈ ∂f, i = 1, 2 ⇒ hx∗ 2 − x∗ 1, x2 − x1i ≥ 0. The routine way of demonstrating the above result can be sketched like this: examining the proof in [11] it is clear that Zagrodny Mean Value The- orem holds for any feasible subdifferential. Using it, one can prove a Minty test for optimality [8] of the following form hx∗, x − x0i ≥ 0, ∀(x, x∗) ∈ ∂f ⇒ 0 ∈ ∂cf (x0), where f is proper and lower semicontinuous and ∂ is feasible subdifferential. If -- on top of (P1) and (P2) -- ∂ also satisfies the natural axiom (P3) ∂(f + x∗) = ∂f + x∗, ∀x∗ ∈ X ∗ (called in [7] stability property which is rather limited form of calculability axiom in [6]), then from Minty test immediately follows that ∂ ∪ ∂c has the somewhat surprising property (first noted by Jules and Lasonde) to be max- 0) is monotonous imal with respect to monotonicity relation. That is, if (x0, x∗ related to ∂f : hx∗ − x∗ 0, x − x0i ≥ 0, ∀(x, x∗) ∈ ∂f, 0) ∈ ∂cf . From the latter Moreau-Rockafellar Theorem about then (x0, x∗ maximal monotonicity of ∂cf for convex, proper and lower semicontinuous f follows immediately, but the surprising fact is that the above is true even if ∂f is not itself monotone. (For precise statement see Theorem 6.) So, in particular if ∂ is feasible and satisfies in addition (P3), then ∂f ⊂ ∂cf, 3 whenever ∂f is monotone. Further the proof can be completed as we do in here presented proof of Theorem 2. Note that the additional axiom (P3) is not really necessary. Our approach is based on a different technique involving barrier func- tions instead of Zagrodny Theorem. We will also make the effort to obtain Theorem 2 for general feasible subdifferental. The paper is organized as follows. In Section 2 we construct and consider the class of barrier functions we use. In Section 3 we show some additional properties of the feasible subdif- ferential linked to (P2) axiom. Finally, in Section 4 we present our proof of Correa-Jofr´e-Thibault theorem. 2 Barrier functions Let U ⊂ X be a open, convex and bounded neighbourhood of 0, i.e. 0 ∈ U. Let µ be the Minkowski functional of U, that is, µ(x) = inf{λ : λ > 0, x ∈ λU} for x 6= 0, µ(0) = 0. It is clear that µ(x) < 1 if x ∈ U, and µ(x) ≥ 1 if x 6∈ U. Let us first list few properties of Minkowski functional, see e.g. [1]: (i) µ has values in [0, +∞); (ii) µ is positively homogeneous, i.e. µ(tx) = tµ(x) for all x ∈ X and t ≥ 0; (iii) µ(x + y) ≤ µ(x) + µ(y) for all x, y ∈ X and hence µ is convex; (iv) {x : µ(x) ≤ 1} = cl U ⇒ {x : µ(x) = 1} = ∂U := cl U \ U, where cl U denotes the topological closure of U and ∂U denotes its boundary. Moreover, (v) There exists c > 0 such that µ(x) ≤ ckxk for all x ∈ X; (vi) There exists b > 0 such that µ(x) ≥ bkxk for all x ∈ X; (vii) µ is Lipschtz continuous. 4 Indeed, let δ > 0 be such that δBX ⊂ U, where BX := {x ∈ X : kxk ≤ 1}. Then δx kxk ∈ δBX ⊂ U, so µ(cid:18) δx kxk(cid:19) ≤ 1, and µ(x) ≤ δ−1kxk, which is (v) with c := δ−1. Further, let U ⊂ sBX , then giving (vi) with b := s−1. sx kxk 6∈ U, so µ(cid:18) sx kxk(cid:19) ≥ 1, and µ(x) ≥ 1 s kxk Finally, µ(x) = µ((x − y) + y) ≤ µ(x − y) + µ(y) by (ii). Hence, µ(x) − µ(y) ≤ µ(x − y) and using (vi) we get µ(x) − µ(y) ≤ bkx − yk which yields µ(x) − µ(y) ≤ bkx − yk for all x, y ∈ X and (vii) holds. For x ∈ U, define the function k(x) := µ(x) 1 − µ(x) = 1 1 − µ(x) − 1. (2) The next lemma shows that k is a barrier function for U, i.e. a continuous function whose values tend to infinity while the arguments tend to ∂U (see e.g. [9]). Lemma 3. The function k defined by (2) has the following properties: (a) k(0) = 0 and for some b > 0, k(x) ≥ µ(x) ≥ bkxk. (b) lim x→∂U k(x) = +∞; (c) k is convex and Lipschitz on each level set {x : k(x) ≤ λ}; (d) the function k(x) := (cid:26) k(x), x ∈ U +∞, x ∈ X \ U convex. Proof. The properties (a) and (b) are clear. is lower semicontinuous and In order to prove that k is convex, it is enough to show that the function g(x) := k(x) + 1 = is convex. 1 1 − µ(x) To this end, fix x, y ∈ U and λ ∈ [0, 1]. From (i) µ(λx + (1 − λ)y) ≤ λµ(x) + (1 − λ)µ(y) ⇔ 1 − µ(λx + (1 − λ)y) ≥ λ(1 − µ(x)) + (1 − λ)(1 − µ(y)). 5 Set α := 1 − µ(x), β := 1 − µ(y), so α, β ∈ (0, 1). Then the latter says g(λx + (1 − λ)y) ≤ 1 λα + (1 − λ)β . We claim that 1 λα + (1 − λ)β ≤ λ α + 1 − λ β . Of course, (4) is equivalent to (3) (4) 0 ≤ λβ(λα + (1 − λ)β) + (1 − λ)α(λα + (1 − λ)β) − αβ = λ(1 − λ)(α2 + β2) + (λ2 + (1 − λ)2 − 1)αβ = λ(1 − λ)(α2 + β2) + (2λ2 − 2λ)αβ = λ(1 − λ)(α − β)2. From (3) and (4) it follows that g(λx + (1 − λ)y) ≤ λg(x) + (1 − λ)g(y) and k is convex. Lipschitz continuity of k on each level set is inherited by the Lipschitz continuity of µ. (d) follows from (b) and (c). 3 Properties of feasible subdifferential For the sake of clarity, we will take two technical parts out of the proof of the main result. Namely, we will show some additional properties of feasible subdifferential linked to (P2) axiom. Lemma 4. Let ∂ be a feasible subdifferential. Let f : X → R ∪ {+∞} be lower semicontinuous and bounded below on the open, convex and bounded set U ⊂ X. Let, moreover, dom f ∩ U 6= ∅. Let g be convex continuous and bounded below barrier function for U. Let ¯x ∈ Xbe fixed. n ∈ ∂f (xn), y ∗ n ∈ Then there exist sequences (xn)∞ n=1, (yn)∞ n=1 ⊂ U and x∗ ∂cg(yn) such that kxn − ynk → 0, f (xn) + g(yn) → inf U (f + g), hx∗ n, xn − ¯xi + hy ∗ n, yn − ¯xi → 0. 6 (5) (6) (7) Proof. Define M := sup{ky − ¯xk : y ∈ U}. (8) Since U is bounded, M ∈ R. Fix a sequence (εn)∞ Pick zn ∈ U such that n=1 such that εn > 0 and εn → 0 as n → ∞. f (zn) + g(zn) < inf U (f + g) + εn. By Ekeland Variational Principle there are yn ∈ U such that f (yn) + g(yn) < inf U (f + g) + εn, (9) and the function x → f (x) + g(x) + εkx − ynk attains its minimum at yn. By (P2) there are uk → yn, as k → ∞, and k ∈ ∂f (uk) such that f (uk) → f (yn) and u∗ w∗− lim u∗ k = u∗ ∈ −∂(g + εnk · −ynk)(yn). By Sum Theorem for the Fenchel subdifferential and the fact that ∂εnk · −ynk = εn∂k · −ynk it follows that there are y ∗ n ∈ ∂g(yn) such that ku∗ + y ∗ nk ≤ εn. (10) Since (uk, f (uk)) → (yn, f (yn)) as k → ∞, there is k1 ∈ N such that kuk − ynk < εn, f (uk) − f (yn) < εn, ∀k > k1. (11) By (9) we have f (uk) + g(yn) − inf U (f + g) ≤ 2εn, ∀k > k1. (12) Note that hu∗ k, uk − ¯xi + hy ∗ n, yn − ¯xi = hu∗ + y ∗ n, yn − ¯xi + hu∗ k − u∗, yn − ¯xi + hu∗ k, uk − yni. n, yn − ¯xi ≤ ku∗ + y ∗ Now, hu∗ + y ∗ k weak-star converges to u∗, we have hu∗ u∗ since by Banach-Steinhaus Theorem the sequence (u∗ nkkyn − ¯xk ≤ Mεn by (8) and (10). Since k − u∗, yn − ¯xi → 0 as k → ∞. Also, k=1 is bounded, and k)∞ 7 uk → yn as k → ∞, we have hu∗ Therefore, there is k2 ∈ N such that k, uk − yni ≤ ku∗ kkkuk − ynk → 0 as k → ∞. hu∗ k, uk − ¯xi + hy ∗ n, yn − ¯xi ≤ (M + 1)εn, ∀k > k2. (13) Set ¯k = max{k1, k2} + 1 and (xn, x∗ n) := (u¯k, u∗ ¯k). From (11), (12) and (13) it follows that so constructed sequences (xn, x∗ and (yn, y ∗ n=1 satisfy (5), (6) and (7). n)∞ n)∞ n=1 Proposition 5. Let ∂ be a feasible subdifferential. Let f : X → R ∪ {+∞} be a proper and lower semicontinuous function. Then dom ∂f is nonempty and f (x) ≤ sup{f (x) + hx∗, x − xi : (x, x∗) ∈ ∂f }, for all x ∈ X. (14) Proof. Fix arbitrary ¯x ∈ X and r < f (¯x). Fix some y ∈ dom f . Since f is lower semicontinuous and the segment [¯x, y] is compact, there is δ > 0 such that f is bounded below on the set C := [¯x, y] + δB ◦ X. Let k be the barrier function defined by (2) for the set U := C − {¯x}. Let ξ > 0 be such that f > r on ¯x + ξBX. Fix a > 0 and such that a > (r − inf U f )/(ξb), where b > 0 and k(x) ≥ µ(x) ≥ bkxk (see Lemma 3 (a)). Then it is immediate that f (x) + ak(x − ¯x) > r, ∀x ∈ U. (15) Apply Lemma 4 to f and the barrier function g = ak(·−x) to get x∗ and y ∗ n ∈ ∂ck(· − ¯x)(yn) such that (5) and (6) are fulfilled and, moreover, n ∈ ∂f (xn) hx∗ n, xn − ¯xi + ahy ∗ n, yn − ¯xi = αn, where lim n→∞ αn = 0. (16) 8 Since y ∗ n ∈ ∂ck(yn − ¯x), we have k(0) ≥ k(yn − ¯x) + hy ∗ n, ¯x − yni ⇐⇒ hy ∗ n, yn − ¯xi ≥ k(yn − ¯x), From the boundedness below of f and (6) it follows that the sequence n=1 is bounded. Since k is Lipschitz on its level sets (see Lemma 3 (k(yn − ¯x))∞ (c)), from (5) it follows that k(yn − ¯x) = k(xn − ¯x) + βn, where lim n→∞ βn = 0. These and (16) give hx∗ −ak(xn − ¯x) + αn − aβn. So, n, xn − ¯xi = αn − ahy ∗ n, yn − ¯xi ≤ αn − ak(yn − ¯x) = hx∗ n, xn − ¯xi ≤ −ak(xn − ¯x) + γn, where lim n→∞ γn = 0. From (15) it follows that hx∗ n, xn − ¯xi < f (xn) − r + γn, or, equivalently, r < f (xn) + hx∗ n, ¯x − xni + γn. Therefore, r ≤ sup{f (xn) + hx∗ n, ¯x − xni : n ∈ N} ≤ sup{f (x) + hx∗, ¯x − xi : (x, x∗) ∈ ∂f }. Since r < f (¯x) were arbitrary, we are done. 4 Monotonicity and convexity We start with the following extension to the case of feasible subdifferential of a result of Jules and Lassonde [7]. Theorem 6. Let ∂ be a feasible subdifferential. Let f : X → R ∪ {+∞} be a 0) ∈ X × X ∗ be in monotone proper lower semicontinuous function. Let (x0, x∗ relation to ∂f , that is, hx∗ − x∗ 0, x − x0i ≥ 0, ∀(x, x∗) ∈ ∂f. (17) Then (x0, x∗ 0) ∈ ∂cf . 9 Proof. Let r ∈ R be such that r < f (x0). (18) Let y ∈ dom f be arbitrary. Since f is lower semicontinuous and the segment [x0, y] is compact, there is δ > 0 such that f is bounded below on Let k be the barrier function defined by (2) for the set U := C − {x0}. C := [x0, y] + δB ◦ X. Let a > 0 be arbitrary. Obviously, g(x) := −hx∗ 0, x − x0i + ak(x − x0) is a convex and continuous barrier for U. So, we can apply Lemma 4 to f and g. Since ∂cg = −x∗ 0 + a∂ck(· − x0), there are x∗ fulfilled and, moreover, n ∈ ∂f (xn) and y ∗ n ∈ ∂ck(· − x0)(yn) such that (5) and (6) are hx∗ n, xn − x0i + h−x∗ 0 + ay ∗ n, yn − x0i → 0. Clearly, hx∗ 0, xn − yni → 0, so (19) is equivalent to hx∗ n − x∗ 0, xn − x0i + ahy ∗ n, yn − x0i → 0. This and (17) give Since y ∗ n ∈ ∂ck(· − x0)(yn), we have lim sup n→∞ hy ∗ n, yn − x0i ≤ 0. k(0) ≥ k(yn − x0) + hy ∗ n, x0 − yni ⇐⇒ hy ∗ n, yn − x0i ≥ k(yn − x0), (19) (20) (21) because k(0) = 0. But k(yn − x0) ≥ bkyn − x0k (cf. Lemma 3 (a)), so (20) and (21) imply that lim sup n→∞ kyn − x0k ≤ 0 ⇐⇒ lim n→∞ yn = x0. From (5) it follows that xn → x0 as well; and from the lower semicontinuity of f and (6) it follows that f (x0) + g(x0) = inf U (f + g). 10 In particular, f (y) − hx∗ 0, y − x0i + ak(y − x0) ≥ f (x0). Since a > 0 was arbitrary, f (y) ≥ f (x0) + hx∗ 0, y − x0i. So, x0 ∈ dom f and, since y ∈ dom f was arbitrary, x∗ 0 ∈ ∂cf (x0). After all the above development, the proof of Correa-Jofr´e-Thibault is now almost immediate. Theorem 2 (Correa-Jofr´e-Thibault). Let X be a Banach space and ∂ be a feasible subdifferential. Let f : X → R ∪ {+∞} be a proper lower semicontinuous function. If ∂f is monotone, then f is convex. Proof. Consider g : X → R ∪ {+∞} defined as g(x) := sup{f (x) + hx∗, x − xi : (x, x∗) ∈ ∂f }. As a supremum of linear functions, g is convex and lower semicontinuous. By (14) we have that f ≤ g. From Theorem 6 and monotonicity of ∂f we have that ∂f ⊂ ∂cf , which implies f ≥ g. Therefore, f = g. References [1] F. Clarke, Functional Analysis, Calculus of Variations and Optimal Con- trol, Graduate Texts in Mathematics 264 (2013) Springer-Verlag, Lon- don. [2] R. Correa, A. Jofr´e, L. Thibault, Characterization of lower semicontin- uous convex functions, PAMS 116 (1992) 67 -- 72. 11 [3] R. Correa, A. Jofr´e, L. Thibault, Subdifferential monotonicity as char- acterization of convex functions, Numer. funct. anal. and Optimiz. 15 (1994) 531 -- 535. [4] R. Correa, A. Jofr´e, L. Thibault, Subdifferential characterization of con- vexity, In: Recent advances in non-smooth optimization, Eds. D.-Z. Du, L. Qi and R. S. Womersly (1995) World Scentific publishing, 18 -- 23. [5] M. Ivanov and N. Zlateva, Maximal Monotonicity of the Subdifferential of a Convex Function: a Direct Proof, Journal of Convex Analysis 24 (2017) 1307 -- 1311. [6] A. Ioffe, On the theory of subdifferentials, Adv. Nonlinear Anal. 1 (2012) 47-120. [7] F. Jules, M. Lassonde, Subdifferential test for optimality, J. Global Op- tim. 59 (2014) 101 -- 106. [8] G. J. Minty: On the monotonicity of the gradient of a convex function, Pacific J. Math. 14 (1964) 243 -- 247. [9] J. Nocedal, S. Wright, Numerical Optimization (1999) Springer, New York. [10] J.-P. Penot, Analysis: From Concepts to Applications (2016) Springer. [11] L. Thibault, A note on the Zagrodny mean value theorem, Optimization 35 (1995) 127 -- 130. [12] L. Thibault, N. Zlateva, Integrability of subdifferentials of directionally Lipschitz functions, PAMS 133 (10) (2005) 2939 -- 2948. [13] D. Zagrodny, Approximate mean value theorem for upper subderiva- tives, Nonlinear Anal. 12 (1988) 1413 -- 1428. 12
1111.2435
1
1111
2011-11-10T10:01:48
A Simple Algorithm for Constructing all Real Hessenberg Unitary Matrices
[ "math.FA" ]
Unitary matrices which are zero below the secondary diagonal (Hessenberg unitary matrices) have many uses in analysis. Given a set of needed conditions on a unitary matrix, this algorithm will give the sparsest unitary matrix. We give an algorithm for constructing all real Hessenberg unitary matrices. The $n\times n$ unitary matrices given by the algorithm have $n-1$ variables which can be chosen to give additional properties needed for a particular application.
math.FA
math
A SIMPLE ALGORITHM FOR CONSTRUCTING ALL REAL HESSENBERG UNITARY MATRICES JANET C. TREMAIN Abstract. Unitary matrices which are zero below the secondary diagonal (Hessenberg unitary matrices) have many uses in analysis. Given a set of needed conditions on a unitary matrix, this algorithm will give the sparsest unitary matrix. We give an algorithm for constructing all real Hessenberg unitary matrices. The n × n unitary matrices given by the algorithm have n − 1 variables which can be chosen to give additional properties needed for a particular application. 1. Introduction Hessenberg matrices have may uses in both pure an applied mathematics. It is a well used fact that every matrix is unitarily equivalent to such a matrix and there are polynomial time algorithms for doing this. In numerical linear algebra, they are used to speed-up eigenvalue computations. Also, real Hessenberg unitary matrices have various uses in analysis. They also have the advantage of being quite sparse. Here we give an algorithm which constructs all such real unitary matrices. The algorithm starts with 2 × 2 matrices and for each new dimension, we drop the top row of the previous dimension, introduce a new variable and then construct two new rows with the new variable added and finally use the rest of the rows of the previous case by adding the first column of zeroes. The n × n unitary matrices have n − 1 variables which can be chosen to give additional properties needed for a particular application. This algorithm was developed at the request of the people working in the Frame Research Center (www.framerc.org) for their work in frame theory where they needed the sparsest unitary matrices. Remark 1.1. Since we can multiply rows and columns of a unitary by −1 and still have a unitary, and if we permute rows or columns while maintaining the Hessenberg property we still have a unitary, we will construct all real triangular unitary matrices up to such changes in sign and permutations of rows or columns. 2. The Construction We will give an algorithm which constructs all real Hessenberg unitary matrices. So we can better see how the algorithm works, we will start by giving some small examples. The author was supported by DTRA/NSF 1042701. 1 2 J. TREMAIN 2.1. 2 × 2 Matrices. This is a unique class and is given by: (cid:20)√1 − z1 √z1 −√z1 √1 − z1(cid:21) Here we must have 0 ≤ z1 ≤ 1. 2.2. 3 × 3 Matrices. This class is given by: p(1 − z1)z2 √1 − z2 √z1z2  −√z2 p(1 − z2)(1 − z1) p(1 − z2)z1  √1 − z1 Here, 0 ≤ z1, z2 ≤ 1. Proposition 2.1. This class of matrices consists of unitary matrices.   0 −√z1 Proof. The proof is done by cases: Case 1: The rows have norm 1. The square sum of the elements of row 1 is The square sum of the elements of row 2 is (1 − z2) + (1 − z1)z2 + z1z2 = 1. z2 + (1 − z2)(1 − z1) + (1 − z2)z1 = z2 + (1 − z2) = 1. It is clear that row 3 is norm 1. Because of the symmetry of the matrix (I.e. switching z1 and z2 in the calcula- tion), it is clear that the row sums and column sums are equal. Case 2: The rows are orthogonal. The inner product of rows 1 and 2 is −p(1 − z2)z2 + (1 − z1)p(1 − z2)z2 + z1p(1 − z2)z2 = p(1 − z2)z2[−1 + (1 − z1) + z1] = 0. −p(1 − z1)z1z2 +p(1 − z1)z1z2 = 0. The inner product of rows 1 and 3 is The inner product of rows 2 and 3 is −p(1 − z2)(1 − z1)z1 +p(1 − z2)(1 − z1)z1 = 0. By symmetry again, the columns are orthogonal. (cid:3) We give some examples of how to use this: Case 1: If z1 = z2 = 0 we get 1 0 0 1 0 0   0 0 1   HESSENBERG UNITARY MATRICES 3 Case 2: If we let z1 = z2 = 1 we get 0 0  0 1 −1 0  0 −1 0   Case 3: If we let z1 = 1 and z2 = 0 we get 0 0  1 0 0 1  0 −1 0   Case 4: If we let z1 = 1 2 and z2 = 1 we get   2 0 q 1 −1 0 −q 1 2 q 1 2 q 1 0 0 2   Case 5: If z1 = 1 2 and z2 = 2 3 we get 3 3 q 1 q 1   3 q 1 −q 2 −q 1 3 q 1 6 q 1 2 q 1 0 6 2   Theorem 2.2. This algorithm gives all 3 × 3 real Hessenberg unitary matrices. Proof. We will just outline this. Given a unitary a11 a12 a13 a21 a22 a23 0 a32 a33     If one of a32 or a33 is zero, it is easily checked that the matrix is a permutation of the rows or columns of the identity multiplied by −1 if necessary and this matrix easily arises from our algorithm. Otherwise, we may assume a32 is negative and 2 32 to get the last row correct given that it square sums to 1. Now, let our z1 = a the last row must be orthogonal to the rows above it so (a22, a23) = c(a12, a13) are the unique vectors (up to length) in R2 which are orthogonal to row 3. Since z1 is given, we can find the unique z2 which makes the last two terms of our first and second rows equal to the given one. The first column is now uniquely determined and must be of our form. (cid:3) 2.3. 4 × 4 Matrices. We construct a unitary matrix by: √1 − z3 √z1z2z3 p(1 − z1)z2z3 −√z3 p(1 − z3)(1 − z2) p(1 − z3)(1 − z1)z2 p(1 − z3)z1z2 p(1 − z2)(1 − z1) p(1 − z2)z1 √1 − z1 p(1 − z2)z3 −√z1 −√z2   0 0 0   We leave it to the reader to check that this constructs all 4 × 4 real Hessenberg unitary matrices. 4 J. TREMAIN 2.4. 5 × 5 Matrices: The Algorithm. Given our general n × n real Hessenberg unitary matrix with n − 1 variables in it, we introduce a new multiplier √zn+1 on the second row of the n × n case and the first entry in the second row of the (n + 1) × (n + 1) matrix is −√zn+1. We also need the sum of the squares of the entries of the two new rows we are adding at the top to equal the squares of the entries of the top row of the n × n matrix. After that, we add the remaining rows of the previous case with a column of zeroes in front. p(1 − z3)z4 p(1 − z2)z3z4 Combining this with the previous case we get: So, to pass from the 4 × 4 case to the 5 × 5 case we create two new rows: (cid:20)√1 − z4 −√z4 p(1 − z4)(1 − z3) p(1 − z4)(1 − z2)z3 p(1 − z4)(1 − z1)z2z3 p(1 − z4)z1z2z3(cid:21) √1 − z4 √z1z2z3z4 p(1 − z2)z3z4  −√z4 p(1 − z4)(1 − z3) p(1 − z4)(1 − z2)z3 p(1 − z4)(1 − z1)z2z3 p(1 − z4)z1z2z3 p(1 − z3)(1 − z2) p(1 − z3)(1 − z1)z2 p(1 − z3)z1z2 p(1 − z2)z1 √1 − z1 p(1 − z2)(1 − z1) p(1 − z1)z2z3z4 p(1 − z1)z2z3z4 p(1 − z3)z4  0 0 0 −√z3 0 0 √z1z2z3z4 −√z2 0 −√z1   Theorem 2.3. The matrix given above is a unitary matrix. Proof. We will outline the proof. Since the last three rows come from the previous unitary matrix, we just need to check that the first two rows are orthogonal to all the others, that they are norm 1 and the column vectors have norm 1. The two new rows are norm 1: The norm of row 1 is: 1 − z4 + (1 − z3)z4 + (1 − z2)z3z4 + (1 − z1)z2z3z4 + z1z2z3z4 = 1 − z3z4 + (1 − z2)z3z4 + z2z3z4 = 1 − z3z4 + z3z4 = 1. The norm of row 2 is: z4 + (1 − z4)(1 − z3) + (1 − z4)(1 − z2)z3 + (1 − z4)(1 − z1)z2z3 + (1 − z4)z1z2z3 = z4 + (1 − z4)[(1 − z3) + (1 − z2)z3 + (1 − z1)z2z3 + z1z2z3] = z4 + (1 − z4)[1 − z2z3 + z2z3] = 1. The column vectors are norm 1: Column 1 square sums to (1 − z4) + z4 = 1. Column 2 square sums to (1 − z3)z4 + (1 − z4)(1 − z3) + z3 = 1 − z3 + z3 = 1. Column 3 square sums to (1 − z2)z3z4 + (1 − z4)(1 − z2)z3 + (1 − z3)(1 − z2) + z2 = (1 − z2)[z3z4 + (1 − z4)z3 + (1 − z3)] + z2 = (1 − z2)[1] + z2 = 1. Column 4 square sums to (1 − z1)z2z3z4 + (1 − z4)(1 − z1)z2z3 + (1 − z3)(1 − z1)z2 + (1 − z2)(1 − z1) + z1 = (1 − z1)[z2z3z4 + (1 − z4)z2z3 + (1 − z3)z2 + (1 − z2)] + z1 = (1 − z1)[z2z3 + 1 − z2z3] + z1 = 1. HESSENBERG UNITARY MATRICES 5 Column 5 square sums to z1z2z3z4 + (1 − z4)z1z2z3 + (1 − z3)z1z2 + (1 − z2)z1 + (1 − z1) = z1z2z3 + z1z2 − z1z2z3 + x1 − z1z2 + 1 − z1 = 1. The rows are orthogonal: The inner product of rows 1 and 2 is: −p(z4(1 − z4) +p(1 − z3)2 z4(1 − z4) +q(1 − z2)2 z 2 3 z4(1 − z4)+ q(1 − z1)2 2 2 z z 2 3 z4(1 − z4) +qz 2 1 z 2 2 z 2 3 z4(1 − z4) = −pz4(1 − z4) + (1 − z3)pz4(1 − z4) + (1 − z2)z3pz4(1 − z4)+ (1 − z1)z2z3pz4(1 − z4) + z1z2z3pz4(1 − z4) = pz4(1 − z4)[−1 + (1 − z3) + (1 − z2)z3 + (1 − z1)z2z3 + z1z2z3] = pz4(1 − z4)[−1 + (1 − z3) + z3 − z2z3 + z2z3 − z1z2z3 + z1z2z3] = 0. (cid:3) The inner product of rows 1 and 3 and taking our squares as above at the same time is −pz3(1 − z3)z4+(1−z2)pz3(1 − z3)z4+(1−z1)z2px3(1 − z3)z4+z1z2pz3(1 − z3)z4 = pz3(1 − z3)z4[−1 + (1 − z2) + (1 − z1)z2 + z1z2] = 0 The inner product of rows 1 and 4 is −pz2(1 − z2)z3z4 + (1 − z1)p(1 − z2)z2z3z4 + z1p(1 − z2)z2z3z4 = The inner product of rows 1 and 5 is pz2(1 − z2)z3z4[−1 + (1 − z1) + z1] = 0. −pz1(1 − z1)z2z3z4 +pz1(1 − z1)z2z3z4] = 0. The inner product of rows 2 and 3 is −p(1 − z4)(1 − z3)z3 + (1 − z2)p(1 − z4)(1 − z3)z3+ (1 − z1)z2p(1 − z4)(1 − z3)z3 + z1z2p(1 − z4)(1 − z3)z3 = p(1 − z4)(1 − z3)z3)[−1 + (1 − z2) + (1 − z1)z2 + z1z2] = 0. The inner product of rows 2 and 4 is −p(1 − z4)z2(1 − z2)z3+(1−z1)p(1 − z4)(1 − z2)z2z3+z1p(1 − z4)(1 − z2)z2z3 = 0. The inner product of rows 2 and 5 is −p(1 − z4)z1(1 − z1)z2z3 +p(1 − z4)z1(1 − z1)z2z3 = 0. 6 J. TREMAIN 3. The General Case For the general case, we take the n × n case and delete the first row and add a first column of zeroes. Then we add two new rows at the top of this matrix by choosing a new variable zn and making row 1 as: √1 − zn p(1 − zn−1)zn p(1 − zn−2)zn−1zn p(1 − zn−3)zn−1zn−2 · · · and making row 2 as p(1 − z1)z2z3 · · · zn . . . √z1z2 · · · zn −√zn p(1 − zn)(1 − zn−1) p(1 − zn)(1 − zn−2)zn−1 p(1 − zn)(1 − zn−3)zn−1zn−2 · · · p(1 − zn)(1 − z1)z2z3 . . . zn−1 p(1 − zn)z1z2z3 . . . zn−1. Remark 3.1. It takes a significant amount of effort to show that we can produce all the required Hessenberg unitary matrices and it does not seem to be important enough to justify the effort since our intention is not to publish this but just post it on the arXiv so it is available to researchers. So we will not address this here. Basically, it can be done by induction on n and a case analysis of the placement of zeroes. References Department of Mathematics, University of Missouri, Columbia, MO 65211-4100 E-mail address: [email protected]
1803.07468
1
1803
2018-03-20T14:56:04
Equiangular tight frames from group divisible designs
[ "math.FA", "math.CO" ]
An equiangular tight frame (ETF) is a type of optimal packing of lines in a real or complex Hilbert space. In the complex case, the existence of an ETF of a given size remains an open problem in many cases. In this paper, we observe that many of the known constructions of ETFs are of one of two types. We further provide a new method for combining a given ETF of one of these two types with an appropriate group divisible design (GDD) in order to produce a larger ETF of the same type. By applying this method to known families of ETFs and GDDs, we obtain several new infinite families of ETFs. The real instances of these ETFs correspond to several new infinite families of strongly regular graphs. Our approach was inspired by a seminal paper of Davis and Jedwab which both unified and generalized McFarland and Spence difference sets. We provide combinatorial analogs of their algebraic results, unifying Steiner ETFs with hyperoval ETFs and Tremain ETFs.
math.FA
math
Equiangular tight frames from group divisible designs Matthew Fickusa, John Jasperb aDepartment of Mathematics and Statistics, Air Force Institute of Technology, Wright-Patterson AFB, OH 45433 bDepartment of Mathematics and Statistics, South Dakota State University, Brookings, SD 57007 Abstract An equiangular tight frame (ETF) is a type of optimal packing of lines in a real or complex Hilbert space. In the complex case, the existence of an ETF of a given size remains an open problem in many cases. In this paper, we observe that many of the known constructions of ETFs are of one of two types. We further provide a new method for combining a given ETF of one of these two types with an appropriate group divisible design (GDD) in order to produce a larger ETF of the same type. By applying this method to known families of ETFs and GDDs, we obtain several new infinite families of ETFs. The real instances of these ETFs correspond to several new infinite families of strongly regular graphs. Our approach was inspired by a seminal paper of Davis and Jedwab which both unified and generalized McFarland and Spence difference sets. We provide combinatorial analogs of their algebraic results, unifying Steiner ETFs with hyperoval ETFs and Tremain ETFs. Keywords: 2010 MSC: 42C15 equiangular tight frames, group divisible designs 1. Introduction Let N ≥ D be positive integers, let F be either R or C, and let hx1, x2i = x∗ product on FD. The Welch bound [56] states that any N nonzero vectors {ϕn}N 1x2 be the dot n=1 in FD satisfy max n6=n′ hϕn,ϕn′ i kϕnkkϕn′ k ≥(cid:2) N −D D(N −1)(cid:3) 1 2 . (1) A > 0 such that Akxk2 = PN n=1 in FD achieve equality in (1) if and It is well-known [49] that nonzero equal-norm vectors {ϕn}N only if they form an equiangular tight frame for FD, denoted an ETF(D, N ), namely if there exists n=1 hϕn, xi2 for all x ∈ FD (tightness), and the value of hϕn, ϕn′i is constant over all n 6= n′ (equiangularity). In particular, an ETF is a type of optimal packing in projective space, corresponding to a collection of lines whose minimum pairwise angle is as large as possible. ETFs arise in several applications, including waveform design for communications [49], compressed sensing [5, 6], quantum information theory [59, 45] and algebraic coding theory [34]. In the general (possibly-complex) setting, the existence of an ETF(D, N ) remains an open problem for many choices of (D, N ). See [23] for a recent survey. Beyond orthonormal bases and regular simplices, all known infinite families of ETFs arise from combinatorial designs. Real ETFs in particular are equivalent to a class of strongly regular graphs (SRGs) [36, 46, 32, 55], and such graphs have been actively studied for decades [11, 12, 15]. This equivalence has been partially generalized to the complex setting in various ways, including approaches that exploit properties of roots of 1 unity [9, 8], abelian distance-regular covers of complete graphs (DRACKNs) [16], and association schemes [33]. Conference matrices, Hadamard matrices, Paley tournaments and quadratic residues are related, and lead to infinite families of ETFs whose redundancy N D is either nearly or exactly two [49, 32, 44, 48]. Harmonic ETFs and Steiner ETFs offer more flexibility in choosing D and N . Harmonic ETFs are equivalent to difference sets in finite abelian groups [54, 49, 58, 18], while Steiner ETFs arise from balanced incomplete block designs (BIBDs) [29, 25]. Recent generalizations of Steiner ETFs have led to new infinite families of ETFs arising from projective planes that contain hyperovals [24] as well as from Steiner triple systems [19], dubbed hyperoval ETFs and Tremain ETFs, respectively. Another new family arises by generalizing the SRG construction of [28] to the complex setting, using generalized quadrangles to produce abelian DRACKNs [22]. Far less is known in terms of necessary conditions on the existence of complex ETF(D, N ). The Gerzon bound implies that N ≤ min{D2, (N − D)2} whenever a complex ETF(D, N ) with N > D > 1 exists [38, 32, 52]. Beyond this, the only known nonexistence result in the complex case is that an ETF(3, 8) does not exist [51], a result proven using computational techniques in algebraic geometry. In quantum information theory, ETF(D, D2) are known as symmetric informationally- complete positive operator-valued measures (SIC-POVMs). It is famously conjectured that such Gerzon-bound-equality ETFs exist for any D [59, 44, 26]. In this paper, we give a new method for constructing ETFs that yields several new infinite families of them. Our main result is Theorem 3.2, which shows how to combine a given initial ETF with a group divisible design (GDD) in order to produce another ETF. In that result, we require the initial ETF to be of one of the following types: Definition 1.1. Given integers D and N with 1 < D < N , we say (D, N ) is type (K, L, S) if D = S K [S(K − 1) + L] = S2 − S(S−L) K , N = (S + L)[S(K − 1) + L], (2) where K and S are integers and L is either 1 or −1. For a given K, we say (D, N ) is K-positive or K-negative when it is type (K, 1, S) or type (K, −1, S) for some S, respectively. We simply say (D, N ) is positive or negative when it is K-positive or K-negative for some K, respectively. When we say that an ETF is one of these types, we mean its (D, N ) parameters are of that type. It turns out that every known ETF(D, N ) with N > 2D > 2 is either a harmonic ETF, a SIC-POVM, or is positive or negative. In particular, every Steiner ETF is positive, while every hyperoval ETF and Tremain ETF is negative. In this sense, the ideas and results of this paper are an attempt to unify and generalize several constructions that have been regarded as disparate. This is analogous to-and directly inspired by-a seminal paper of Davis and Jedwab [17], which unifies McFarland [41] and Spence [47] difference sets under a single framework, and also generalizes them so as to produce difference sets whose corresponding harmonic ETFs have parameters D = 1 3 22J−1(22J+1 + 1), N = 1 3 22J+2(22J − 1), (3) for some J ≥ 1. It is quickly verified that such ETFs are type (4, −1, S) where S = 1 3 (22J+1 +1). As we shall see, combining our main result (Theorem 3.2) with known ETFs and GDDs recovers the existence of ETFs with these parameters, and also provides several new infinite families, including: Theorem 1.2. An ETF(D, N ) of type (K, −1, S) exists whenever: (a) K = 4 and either S ≡ 3 mod 8 or S ≡ 7 mod 60; 2 (b) K = 5 and either S ≡ 4 mod 15 or S ≡ 5, 309 mod 380 or S ≡ 9 mod 280. This result extends the S for which an ETF of type (4, −1, S) is known to exist from a geometric progression to a finite union of arithmetic progressions, with the smallest new ETF having S = 19, namely (D, N ) = (266, 1008), cf. [23]. Meanwhile, the ETFs given by Theorem 1.2 in the K = 5 case seem to be completely new except when S = 4, 5, 9, with (D, N ) = (285, 1350) being the smallest new example. Using similar techniques, we were also able to find new, explicit infinite families of K-negative ETFs for K = 6, 7, 10, 12. The description of these families is technical, and so is given in Theorem 4.4 as opposed to here. More generally, using asymptotic existence results for GDDs, we show that an infinite number of K-negative ETFs also exist whenever K = Q + 2 where Q is a prime power, K = Q + 1 where Q is an even prime power, or K = 8, 20, 30, 42, 56, 342. In certain cases, the new ETFs constructed by these methods can be chosen to be real: Theorem 1.3. (a) There are an infinite number of real Hadamard matrices of size H ≡ 1 mod 35, and a real ETF of type (5, −1, 8H + 1) exists for all such H. (b) There are an infinite number of real Hadamard matrices of size H ≡ 1, 8 mod 21, and a real ETF of type (6, −1, 2H + 1) exists for all sufficiently large such H. (c) There are an infinite number of real Hadamard matrices of size H ≡ 1, 12 mod 55, and a real ETF of type (10, −1, 4H + 1) exists for all sufficiently large such H. (d) There are an infinite number of real Hadamard matrices of size H ≡ 1, 277 mod 345, and a real ETF of type (15, −1, 4H + 1) exists for all sufficiently large such H. These correspond to four new infinite families of SRGs, with the smallest new example being a real ETF(66759, 332640), which is obtained by letting H = 36 in (a). In the next section, we introduce known concepts from frame theory and combinatorial design that we need later on. In Section 3, we provide an alternative characterization of when an ETF is positive or negative (Theorem 3.1), which we then use to help prove our main result (Theorem 3.2). In the fourth section, we discuss how many known ETFs are either positive or negative, and then apply Theorem 3.2 to them along with known GDDs to obtain the new infinite families of negative ETFs described in Theorems 1.2 and 4.4. We conclude in Section 5, using these facts as the basis for new conjectures on the existence of real and complex ETFs. 2. Preliminaries 2.1. Equiangular tight frames For any positive integers N and D, and any sequence {ϕn}N sponding synthesis operator is Φ : FN → FD, Φy := PN n=1 of vectors in FD, the corre- n=1 y(n)ϕn, namely the D × N matrix whose nth column is ϕn. Its adjoint (conjugate transpose) is the analysis operator Φ∗ : FD → FN , which has (Φ∗x)(n) = hϕn, xi for all n = 1, . . . , N . That is, Φ∗ is the D × N matrix whose nth n. Composing these two operators gives the N × N Gram matrix Φ∗Φ whose (n, n′)th row is ϕ∗ We say {ϕn}N entry is (Φ∗Φ)(n, n′) = hϕn, ϕn′i, as well as the D × D frame operator ΦΦ∗ =PN n=1 is a tight frame for FD if there exists A > 0 such that ΦΦ∗ = AI, namely if the rows of Φ are orthogonal and have an equal nontrivial norm. We say {ϕn}N n=1 is equal norm if there exists some C such that kϕnk2 = C for all n. The parameters of an equal norm tight frame n=1 ϕnϕ∗ n. 3 are related according to DA = Tr(AI) = Tr(ΦΦ∗) = Tr(Φ∗Φ) = PN n=1 kϕnk2 = N C. We say n=1 is equiangular if it is equal norm and the value of hϕn, ϕn′i is constant over all n 6= n′. {ϕn}N For any equal norm vectors {ϕn}N n=1 in FD, a direct calculation reveals 0 ≤ Tr[( 1 C ΦΦ∗ − N D I)2] = N N Xn=1 Xn′=1 n′6=n hϕn,ϕn′ i2 C 2 − N (N −D) D ≤ N (N − 1) max n6=n′ hϕn,ϕn′ i2 C 2 − N (N −D) D . Rearranging this inequality gives the Welch bound (1). Moreover, we see that achieving equality in (1) is equivalent to having equality above throughout, which happens precisely when {ϕn}N is a tight frame for FD that is also equiangular, namely when it is an ETF for FD. n=1 If N > D and {ϕn}N n=1 is a tight frame for FD then completing the D rows of Φ to an equal-norm orthogonal basis for FN is equivalent to taking a (N − D) × N matrix Ψ such that Ψ∗Ψ = AI, ΦΨ∗ = 0 and Φ∗Φ + Ψ∗Ψ = AI. The sequence {ψn}N n=1 of columns of any such Ψ is n=1. Since ΨΨ∗ = AI and Ψ∗Ψ = AI − Φ∗Φ, any Naimark called a Naimark complement of {ϕn}N complement of an ETF(D, N ) is an ETF(N − D, N ). Since any nontrivial scalar multiple of an ETF is another ETF, we will often assume without loss of generality that a given ETF(D, N ) and its Naimark complements satisfy 1 2 , ∀n = 1, . . . , N, (4) A = N(cid:2) N −1 D(N −D)(cid:3) 1 2 , kϕnk2 =(cid:2) D(N −1) N −D (cid:3) 1 2 , kψnk2 =(cid:2) (N −D)(N −1) D (cid:3) which equates to having hϕn, ϕn′i = 1 = hψn, ψn′i for all n 6= n′. For positive and negative ETFs in particular (Definition 1.1), we shall see that all of these quantities happen to be integers. Any ETF(D, N ) with N = D + 1 is known as a regular simplex, and such ETFs are Naimark complements of ETFs for F1, namely sequences of scalars that have the same nontrivial modulus. n=1 in FN −1 is a Naimark complement of the all-ones In particular, a sequence of vectors {fn}N sequence in F1 if and only if FF∗ = N I, N Xn=1 fn = F1 = 0, F∗F = N I − J, (5) where 1 and J denote an all-ones column vector and matrix, respectively. Equivalently, the vectors n=1 in FN are equal norm and orthogonal. In particular, for any N > 1, we can always {1 ⊕ fn}N take {1 ⊕ fn}N n=1 to be the columns of a possibly-complex Hadamard matrix of size N . In this case, F satisfies (5) and is also flat, meaning every one of its entries has modulus one. As detailed below, flat regular simplices can be used to construct several families of ETFs, including Steiner ETFs as well as those we introduce in Theorem 3.2. Harmonic ETFs are the best-known class of ETFs [48, 58, 18]. A harmonic ETF(D, N ) is obtained by restricting the characters of an abelian group G of order N to a difference set of cardinality D, namely a D-element subset D of G with the property that the cardinality of {(d, d′) ∈ D × D : g = d − d′} is constant over all g ∈ G, g 6= 0. The set complement G\D of any difference set in G is another difference set, and the two corresponding harmonic ETFs are Naimark complements. In particular, for any abelian group G of order N , the harmonic ETF arising from G\{0} is a flat regular simplex that satisfies (5). 4 2.2. Group divisible designs For a given integer K ≥ 2, a K-GDD is a set V of V > K vertices, along with collections G and B of subsets of V, called groups and blocks, respectively, with the property that the groups partition V, every block has cardinality K, and any two vertices are either contained in a common group or a common block, but not both. A K-GDD is uniform if its groups all have the same cardinality M , denoted in exponential notation as a "K-GDD of type M U " where V = U M . Letting B be the number of blocks, a {0, 1}-valued B × U M incidence matrix X of a K-GDD of type M U has the property that each row of X contains exactly K ones. Moreover, for any v = 1, . . . , V = U M , the vth column of X is orthogonal to M − 1 other columns of X, and has a dot product of 1 with each of the remaining (U − 1)M columns. This implies that the replication number Rv of blocks that contain the vth vertex satisfies (U −1)M = V Xv′=1 v′6=v (X∗X)(v, v′) = X(b, v) B Xb=1 V Xv′=1 v′6=v X(b, v′) = B Xb=1(cid:26) K − 1, X(b, v) = 1 0, X(b, v) = 0(cid:27) = Rv(K −1). As such, this number Rv = R is independent of v. At this point, summing all entries of X gives BK = V R and so B is also uniquely determined by K, M and U . Because of this, the existence of a K-GDD of type M U is equivalent to that of a {0, 1}-valued B × U M matrix X with R = M (U −1) K−1 , B = M U R K = M 2U (U −1) K(K−1) , X1 = K1, X∗X = R I + (JU − IU ) ⊗ JM . (6) In the special case where M = 1, a K-GDD of type 1U is called a BIBD(U, K, 1). In the special case where U = K, a K-GDD of type M K is called a transversal design TD(K, M ), which is equivalent to a collection of K − 2 mutually orthogonal Latin squares (MOLS) of size M . In order for a K-GDD of type M U to exist, the expressions for R and B given in (6) are necessarily integers. Beyond this, we necessarily have U ≥ K since we can partition any given block into its intersections with the groups, and the cardinality of these intersections is at most one. Altogether, the parameters of a K-GDD of type M U necessarily satisfy U ≥ K, M (U −1) K−1 ∈ Z, M 2U (U −1) K(K−1) ∈ Z. (7) Though these necessary conditions are not sufficient [27], they are asymptotically sufficient in two distinct ways: for any fixed K ≥ 2 and M ≥ 1, there exists U0 = U0(K, M ) such that a K-GDD of type M U exists for all U ≥ U0 such that (7) is satisfied [13, 37]; for any fixed U ≥ K ≥ 2, there exists M0 = M0(K, U ) such that a K-GDD of type M U exists for all M ≥ M0 such that (7) is satisfied [43]. In the M = 1 and U = K cases, these facts reduce to more classical asymptotic existence results for BIBDs and MOLS, respectively. Many specific examples of GDDs are formed by combining smaller designs in clever ways. We in particular will make use of the following result, which is a special case of Wilson's approach [57]: Lemma 2.1. If a K-GDD of type M U and a U -GDD of type N V exist, then a K-GDD of type (M N )V exists. Proof. Let X and Y be incidence matrices of the form (6) for the given K-GDD of type M U and U -GDD of type N V , respectively. In particular, taking R and B as in (6), we can write uXu′ = J for any X = (cid:2) X1 · · · XU (cid:3) where each Xu is a B × M matrix with X∗ uXu = RI, and X∗ 5 u 6= u′. We now construct the incidence matrix Z of a K-GDD of type (M N )V in the following manner: in each row of Y, replace each of the U nonzero entries with a distinct matrix Xu, and replace each of the zero entries with a B × M matrix of zeros. This result generalizes MacNeish's classical method for combining MOLS [40]: if a TD(K, M ) and a TD(K, N ) exist, then applying Lemma 2.1 to them produces a TD(K, M N ). We will also use one GDD to "fill the holes" of another: Lemma 2.2. If K-GDDs of type M U and (M U )V exist, then a K-GDD of type M U V exists. Proof. Letting X and Y be incidence matrices of the form (6) for the given K-GDDs of type M U and (M U )V , respectively, it is straightforward to verify that is the incidence matrix of a K-GDD of type M U V . Z =(cid:20) IV ⊗ X Y (cid:21) 2.3. Previously known constructions of ETFs involving BIBDs and MOLS In the next section, we introduce a method for constructing ETFs that uses GDDs. This method makes use of a concept from [19], which we now generalize from BIBDs to GDDs: Definition 2.3. Take a K-GDD of type M U where M ≥ 1 and U ≥ K ≥ 2, and define R, B and an incidence matrix X according to (6). Without loss of generality, write the columns of X as {xu,m}U m=1 have disjoint support. Then, for any u and m, a corresponding embedding operator Eu,m is any {0, 1}-valued B × R matrix whose columns are standard basis elements that sum to xu,m. m=1 where, for each u, the vectors {xu,m}M M u=1, letting {Ev}V v=1 = {Eu,1}U In the special case where M = 1, this concept leads to an elegant formulation of Steiner u=1 be the embedding operators of a BIBD(V, K, 1), and ETFs [19]: letting {1 ⊕ fi}R K−1 + 1, the V (R + 1) vectors {Evfi}V i=0 form an ETF for FB. In [19], this fact is proven using several properties of embedding operators. We now show those properties generalize to the GDD setting; later on, we use these facts to prove our main result: i=0 be the columns of a possibly-complex Hadamard matrix of size R + 1 = V −1 v=1, R Lemma 2.4. If {Eu,m}U u=1,, M m=1 are the embedding operators arising from a K-GDD of type M U , E∗ u,mEu′,m′ =  u = u′, m = m′, u = u′, m 6= m′, I, 0, δrδ∗ r′, u 6= u′. Here, for any u 6= u′ and m, m′, δr and δr′ are standard basis elements in FR whose indices r, r′ depend on u, u′, m, m′. Proof. Each Eu,m is a matrix whose columns are standard basis elements that sum to xu,m, and so is an isometry, that is, E∗ u,mEu′,m′ is a matrix whose entries are nonnegative integers that sum to: u,mEu,m = I. Moreover, for any u, u′, m, m′, E∗ R R Xr=1 Xr′=1 (E∗ u,mEu′,m′)(r, r′) = 1∗E∗ u,mEu′,m′ 1 = hxu,m, xu′,m′i =  6 R, u = u′, m = m′, 0, u = u′, m 6= m′, 1, u′ 6= u. u,mEu′,m′ = 0. If instead u 6= u′ then this implies that u,mEu′,m′ has a single nonzero entry, and that this entry has value 1. This means there exists When u = u′ and m 6= m′, this implies E∗ E∗ some r, r′ = 1, . . . , R, r = r(u, m, u′, m′), r′ = r′(u, m, u′, m′) such that E∗ u,mEu,m′ = δrδ∗ r′. Other Steiner-like constructions of ETFs include hyperoval ETFs [24] and Tremain ETFs [22]. Beyond Steiner and Steiner-like techniques, there are at least two other methods for constructing ETFs that make direct use of the incidence matrix of some kind of GDD. One method leads to the phased BIBD ETFs of [22]: if X is the B × V incidence matrix of a BIBD(V, K, 1), and Φ is any matrix obtained by replacing each 1-valued entry of X with any unimodular scalar, then the columns of Φ are immediately equiangular, and the challenge is to design them so that they form a tight frame for their span. Another method constructs ETFs with (D, N ) = ( 1 2 M (M ± 1), M 2) from MOLS. To elaborate, a TD(K, M ) is a K-GDD of type M K , meaning by (6) that it has an M 2 × KM incidence matrix X that satisfies X1 = K1, X∗X = M I + (JK − IK) ⊗ JM . (8) Here, the columns of X have support M , and are arranged as K groups of M columns apiece, where columns in a common group have disjoint support. Together, these facts imply, in turn, that X(IK ⊗ 1M ) = 1M 21∗ K , (XX∗)2 = M XX∗ + K(K − 1)J. (9) At this point, the traditional approach is to let A = XX∗ − KI be the adjacency matrix of the TD's block graph, and use (8) and (9) to show that this graph is strongly regular with parameters (M 2, K(M − 1), M + K(K − 3), K(K − 1)). In the M = 2K case, applying Theorem 4.4 of [21] to this graph then produces a real ETF with (D, N ) = ( 1 2 M (M − 1), M 2) whose vectors sum to zero, while applying this same result in the M = 2(K − 1) case produces a real ETF with (D, N ) = ( 1 2 M (M + 1), M 2) whose synthesis operator's row space contains the all-ones vector. That said, a careful read of the literature reveals that this construction can be made more explicit, and that doing so has repercussions for coding theory. To elaborate, in [10], MOLS are used to produce quasi-symmetric designs (QSDs) which, via the techniques of [42], yield self- complementary binary codes that achieve equality in the Grey-Rankin bound. In [34], such codes are shown to be equivalent to flat real ETFs. A method for directly converting the incidence matri- ces of certain QSDs into synthesis operators of ETFs was also recently introduced [20]. Distilling these ideas leads to the following streamlined construction: let X be the incidence matrix of a TD(K, M ), let {1 ⊕ fm}M m=1 be the columns of a possibly-complex Hadamard matrix of size M , let m=1, and consider the K(M − 1) × M 2 matrix F be the (M − 1) × M synthesis operator of {fm}M Φ = (IK ⊗ F)X∗. (10) Using (5), (8), and (9) along with the fact that F is flat, it is straightforward to show that Φ is flat and satisfies ΦΦ∗ = M 2I and Φ∗Φ = M XX∗ − KJ. As such, the columns of Φ form a flat two-distance tight frame (TDTF) for FK(M −1) [7]. Moreover, this TDTF is an ETF when M = 2K. In particular, if there exists a TD(K, 2K) and a real Hadamard matrix of size 2K, then there exists a flat real ETF(K(2K − 1), 4K 2). Using the equivalence between flat real ETFs and Grey-Rankin-bound-equality codes given in [34], or alternatively the equivalence between such ETFs and certain QSDs given in [20], this recovers Theorem 1 of [10]. In the K = 6 case, that result gives the only known proof to date of the existence of a flat real ETF(66, 144). 7 For any TD and corresponding flat regular simplex, it is quickly verified that the corresponding TDTF (10) is centered [21] in the sense that Φ1 = 0, namely that the all-ones vector is orthogonal to the row space of Φ. This fact leads to an analogous reinterpretation of the second main result of [10]: in lieu of (10), we instead consider the [K(M − 1) + 1] × M 2 flat matrix Ψ =(cid:20) 1∗ Φ (cid:21) =(cid:20) 1∗ (IK ⊗ F)X∗(cid:21) . (11) Here, the properties of Φ immediately imply ΨΨ∗ = M 2I and Ψ∗Ψ = M XX∗ −(K −1)J, meaning the columns of Ψ form a flat TDTF. However, unlike (10), the columns of (11) are equiangular precisely when M = 2(K − 1). Replacing K with K + 1, this implies in particular that if there exists a TD(K + 1, 2K) and a real Hadamard matrix of size 2K, then there exists a flat real ETF(K(2K + 1), 4K 2). This recovers Theorem 2 of [10] via the equivalences of [34, 20], and gives the only known proof of the existence of a flat real ETF(78, 144). Simply put, if certain TDs exist, then certain ETFs exist. In the next section, we introduce a new method of constructing ETFs from TDs. 3. Constructing equiangular tight frames with group divisible designs In [17], Davis and Jedwab unify McFarland [41] and Spence [47] difference sets under a single framework, and also generalize them so as to produce difference sets with parameters (3). McFar- land's construction relies on nice algebro-combinatorial properties of the set of all hyperplanes in a finite-dimensional vector space over a finite field. Davis and Jedwab exploit these properties to form various types of building sets, which in some cases lead to difference sets. In [34], it is shown that every harmonic ETF arising from a McFarland ETF is unitarily- equivalent to a Steiner ETF arising from an affine geometry. When we applied a similar analysis to the building sets of [17], we discovered that they have an underlying TD-like incidence structure. (We do not provide this analysis here since it is nontrivial and does not help us prove our results in their full generality.) This eventually led us to the ETF construction technique of Theorem 3.2 below. In short, our approach here is directly inspired by that of [17], though this is not apparent from our proof techniques. In particular, the fact that the L parameter in Definition 1.1 is either 1 or −1 is a generalization of Davis and Jedwab's notion of extended building sets with "+" and "−" parameters, respectively. To facilitate our arguments later on, we now consider these types of parameters in greater detail: Theorem 3.1. If 1 < D < N and (D, N ) is type (K, L, S), see Definition 1.1, then where S ≥ 2. Conversely, given (D, N ) such that 1 < D < N , and letting L be either 1 or −1, if the above expressions for S and K are integers then (D, N ) is type (K, L, S). S =(cid:2) D(N −1) N −D (cid:3) 1 2 , K = N S D(S+L) , (12) Moreover, in the case that the equivalent conditions above hold, scaling an ETF {ϕn}N FD so that hϕn, ϕn′i = 1 for all n 6= n′ gives that it and its Naimark complements {ψn}N tight frame constant A = K(S + L) and n=1 for n=1 have kϕnk2 =(cid:2) D(N −1) N −D (cid:3) 1 2 = S, kψnk2 =(cid:2) (N −D)(N −1) D 8 1 2 = S(K − 1) + KL, ∀n = 1, . . . , N. (13) (cid:3) Proof. Whenever (D, N ) is type (K, L, S) we have that L is either 1 or −1 by assumption, at which point the fact that L2 = 1 coupled with (2) gives N − 1 = (S + L)[S(K − 1) + L] − 1 = S2(K − 1) + SKL = S[S(K − 1) + KL], D − 1 = K(S+L)[S(K−1)+L] S[S(K−1)+L] − 1 = K(S+L) S [S(K − 1) + KL]. − 1 = 1 N S Multiplying and dividing these expressions immediately implies that 1 2 = S, (cid:2) D(N −1) N −D (cid:3) (cid:2) (N −D)(N −1) D (cid:3) 1 2 = S(K − 1) + KL. (14) Here, S is an integer by assumption, and is clearly positive. Moreover, if S = 1 then (14) implies D = 1. Since D > 1 by assumption, we thus have S ≥ 2. Continuing, (2) further implies N S D(S+L) = K(S+L)[S(K−1)+L] S[S(K−1)+L] S S+L = K. Conversely, now assume that S and K are defined by (12), where L is either 1 or −1, and that S and K are integers. As before, the fact that D > 1 implies that S ≥ 2 and so K > 0. We solve for N in terms of D, K, and L. Here, (12) gives N S . Since L2 = 1, this implies S = 1 + L DK = S+L Squaring this equation and multiplying the result by N − 1 gives 1 2 = 1 (cid:2) N −D D(N −1)(cid:3) D = (N − 1)(cid:0) N S = L(cid:0) N DK − 1(cid:1)2 = N(cid:0) N DK − 1(cid:1). DK − 1(cid:1)2 − N N D − 1 = N −D Adding 1 to this equation and multiplying by (DK)2 N then leads to a quadratic in N : DK(cid:0) N DK − 2(cid:1) − 1. DK 2 = (N − DK)2 − (N − 2DK) = N 2 − (2DK + 1)N + DK(DK + 2). Applying the quadratic formula then gives (15) (16) (17) (18) N = DK + 1 1 2(cid:9). 2(cid:8)1 ±(cid:2)4DK(K − 1) + 1(cid:3) 2DK(cid:8)1 ±(cid:2)4DK(K − 1) + 1(cid:3) As such, (15) becomes 1 here correspond to L = 1 and L = −1 respectively, that is, S = L(cid:0) N DK − 1(cid:1) = L 2DK(cid:8)1 + L(cid:2)4DK(K − 1) + 1(cid:3) 1 1 S = L 2(cid:9) = 1 2DK(cid:8)L +(cid:2)4DK(K − 1) + 1(cid:3) 2(cid:9). 1 2(cid:9), implying "+" and "−" 1 Moreover, since N is an integer, (16) implies that 4DK(K − 1) + 1 is the square of an odd integer, that is, that 4DK(K −1)+1 = (2J −1)2 = 4J(J −1)+1 or equivalently that DK(K −1) = J(J −1) for some positive integer J. Writing D = J(J−1) K(K−1), (16) and (17) then become N = DK + 1 2(cid:8)1 + L(cid:2)4DK(K − 1) + 1(cid:3) 1 S = 1 1 2DK(cid:8)L +(cid:2)4DK(K − 1) + 1(cid:3) 2(cid:9) = K−1 1 2 [1 + L(2J − 1)], 2(cid:9) = DK + 1 2J(J−1) [L + (2J − 1)] =( K−1 J−1 , L = 1 K−1 J , L = −1) = 2(K−1) 2J−L−1 . 9 That is, J = S(K − 1) + 1 the fact that L2 = 1 then gives the expressions for D and N given in Definition 1.1: 2 (L + 1). Substituting this into D = J(J−1) K(K−1) and (18) and again using [S(K−1)+ 1 2 (L+1)][S(K−1)+ 1 2 (L−1)] D = N = S[S(K − 1) + L] + 1 K(K−1)) = S 2(K−1)2+S(K−1)L K [S(K − 1) + L], 2 {1 + L[2S(K − 1) + L]} = (S + L)[S(K − 1) + L]. K(K−1)) = S Finally, in the case where (D, N ) is type (K, L, S), if {ϕn}N n=1 is an ETF for FD, and is without loss of generality scaled so that hϕn, ϕn′i = 1 for all n 6= n′, then (4) and (14) immediately imply that {ϕn}N n=1 satisfy (13), and that both are tight frames with tight frame constant A = N S n=1 and any one of its Naimark complements {ψn}N D = K(S + L). Theorem 3.1 implies that the (D, N ) parameters of an ETF with N > D > 1 are type (K, L, S) with K = 1 if and only if that ETF is a regular simplex, and moreover that this only occurs when L = 1 and S = D. Indeed, for any D > 1, the pair (D, N ) = (D, D + 1) satisfies (2) when (K, L, S) = (1, 1, D). Conversely, in light of (13), an ETF of type (1, L, S) has a Naimark n=1 with the property that kψnk2 = L = 1 and hψn, ψn′i = 1 for all n 6= n′, complement {ψn}N namely a Naimark complement that is an ETF(1, N ). We also emphasize that it is sometimes possible for the parameters of a single ETF to be In particular, if D > 1 and simultaneously positive and negative for different choices of K. (D, D + 1) is type (K, −1, S), then (12) gives S = D and K = N S D−1 . Since K is an integer, this implies either D = 2 or D = 3. And, letting (K, L, S) be (3, −1, 2) and (2, −1, 3) in (2) indeed gives that (D, N ) is (2, 3) and (3, 4), respectively. D(S+L) = (D+1)D D(D−1) = D+1 D−1 = 1 + 2 In the next section, we provide a much more thorough discussion of positive and negative ETFs, including some other examples of ETFs that are both. For now, we turn to our main result, which shows how to combine a given ETF(D, N ) whose parameters are type (K, L, S) with a certain K-GDD to produce a new ETF whose parameters are type (K, L, S′) for some S′ > S. Here, as with any GDD, we require K ≥ 2. In light of the above discussion, this is not a significant restriction since (D, N ) has type (1, L, S) if and only if L = 1 and S = D, and we already know that ETFs of type (1, 1, D) exist for all D > 1, being regular simplices. Theorem 3.2. Assume an ETF of type (K, L, S) exists where K ≥ 2, and let M = S(K − 1) + L. The necessary conditions (7) on the existence of a K-GDD of type M U reduce to having Moreover, if such a GDD exists, and U has the additional property that U ≥ K, U −1 K−1 ∈ Z, (S−L)U (U −1) K(K−1) ∈ Z. (K−2)(U −1) (S+L)(K−1) ∈ Z, (19) (20) then there exists an ETF of type (K, L, S′) where S′ = S + R = M U −L K−1 , where R = M (U −1) K−1 . In particular, under these hypotheses, W := R given ETF as {ϕm,i}M embedding operators of the GDD (Definition 2.3), letting {δu}U and letting {ei}S+L S + L and W + 1, respectively, then the following vectors form an ETF of type (K, L, S′): S+L ∈ Z, and without loss of generality writing the M m=1 be the u=1 be the standard basis for FU , j=0 be the columns of possibly-complex Hadamard matrices of size S+L i=1 where kϕm,ik2 = S for all m and i, letting {Eu,m}U i=1 and {1 ⊕ fj}W m=1, u=1, {ψu,m,i,j}U u=1, M m=1, S+L i=1, W As special cases of this fact, an ETF of type (K, L, S′) exists whenever either: j=0, ψu,m,i,j := (δu ⊗ ϕm,i) ⊕(cid:0)Eu,m(ei ⊗ fj)(cid:1). (21) 10 (a) U is sufficiently large and satisfies (19) and (20); (b) there exists a K-GDD of type M U , provided we also have that S + L divides K − 2. Proof. Since M = S(K − 1) + L, M K−1 = S + L K−1 = (S + L) − L(cid:0) K−2 K−1(cid:1), M K = S − S−L K . (22) As such, the replication number of any K-GDD of type M U is R = M (U −1) K−1). In particular, such a GDD can only exist when K − 1 necessarily divides U − 1. Moreover, multiplying the expressions in (22) gives that the number of blocks in any such GDD is K−1 = S(U −1)+L( U −1 B = M 2U (U −1) K(K−1) =(cid:0)S − S−L K (cid:1)(cid:0)S + L = S2U (U − 1) + LSU U −1 K−1(cid:1)U (U − 1) K−1 − S(S−L) K U (U − 1) − L (S−L)U (U −1) K(K−1) . Here, since our initial ETF(D, N ) is type (K, L, S), S(S−L) K = S2 −D is an integer, and so the above expression for B is an integer precisely when K(K − 1) divides (S − L)U (U − 1). To summarize, since M = S(K − 1) + L where K divides S(S − L), the necessary conditions (7) on the existence of a K-GDD of type M U reduce to having (19). These necessary conditions are known to be asymptotically sufficient [13, 37]: for this fixed K and M , there exists U0 such that a K-GDD of type M U exists for any U ≥ U0 that satisfies (19). Regardless, to apply our construction below with any given K-GDD of type M U , we only need U to satisfy the additional property that W = R S+L = M (U −1) (S+L)(K−1) = U −1 S+L(cid:2)(S + L) − L(cid:0) K−2 K−1(cid:1)(cid:3) = (U − 1) − L (K−2)(U −1) (S+L)(K−1) is an integer, namely to satisfy (20). Since K − 1 necessarily divides U − 1, this is automatically satisfied whenever S + L happens to divide K − 2, and some of the ETFs we will identify in the next section will have this nice property. Regardless, there are always an infinite number of values of U which satisfy (19) and (20), including, for example, all U ≡ 1 mod (S + L)K(K − 1). Turning to the construction itself, the fact that the given ETF(D, N ) is type (K, L, S) implies D = S K [S(K − 1) + L] = SM K , N = (S + L)[S(K − 1) + L] = (S + L)M. (23) In particular, since N = (S + L)M , the vectors in our initial ETF(D, N ) can indeed be indexed as {ϕm,i}M S+L i=1 . Moreover, m=1, M U = M (U − 1) + M = (K − 1)R + [S(K − 1) + L] = (S + R)(K − 1) + L. (24) As such, the number of vectors in the collection (21) is N ′ = U M (S + L)(W + 1) = (S + L)( R S+L + 1)M U = [(S + R) + L][(S + R)(K − 1) + L]. (25) Also, for each i and j, ei ⊗ fj lies in a space of dimension F(S+L)W = FR. And, for each u and m, Eu,m is a B × R matrix. As such, for any u, m, i and j, ψu,m,i,j is a well-defined vector in FD′ where, by combining (6), (23) and (24), we have D′ = U D + B = U SM K + M U K R = S+R K M U = S+R 11 K [(S + R)(K − 1) + L]. (26) Comparing (25) and (26) against (2), we see that (D′, N ′) is indeed type (K, L, S′) where S′ = S+R. Here, (24) further implies S′ = S + R = M U −L K−1 . Continuing, since (D′, N ′) is type (K, L, S′), Theorem 3.1 gives that the Welch bound for N ′ vec- tors in FD′ , it suffices to prove that kψu,m,i,jk2 = S′ for all u, m, i, j, and that hψu,m,i,j, ψu′,m′,i′,j ′i is unimodular whenever (u, m, i, j) 6= (u′, m′, i′, j′). Here, for any u, u′ = 1, . . . , U , m, m′ = 1, . . . , M , i, i′ = 1, . . . , S + L, j, j′ = 0, . . . , W , S ′ . As such, to show (21) is an ETF for FD′ is 1 hψu,m,i,j, ψu′,m′,i′,j ′i = h(δu ⊗ ϕm,i) ⊕(cid:0)Eu,m(ei ⊗ fj)(cid:1), (δu′ ⊗ ϕm′,i′) ⊕(cid:0)Eu′,m′(ei′ ⊗ fj ′)(cid:1)i = hδu, δu′ihϕm,i, ϕm′,i′i + hEu,m(ei ⊗ fj), Eu′,m′(ei′ ⊗ fj ′)i = hδu, δu′ihϕm,i, ϕm′,i′i + hei ⊗ fj, E∗ u,mEu′,m′(ei′ ⊗ fj ′)i. When u = u′, m = m′, i = i′ and j = j′, (27) indeed becomes (27) kψu,m,i,jk2 = kδuk2kϕm,ik2 + keik2kfjk2 = 1S + (S + L)W = S + R = S′. As such, all that remains is to show that (27) is unimodular in all other cases. For instance, if u 6= u′ u,mEu′,m′ = δrδ∗ then Lemma 2.4 gives that for any m, m′ there exists r, r′ = 1, . . . , R such that E∗ r′ meaning in this case (27) becomes hψu,m,i,j, ψu′,m′,i′,j ′i = 0hϕm,i, ϕm′,i′i + hei ⊗ fj , δrδ∗ r′(ei′ ⊗ fj ′)i = (ei ⊗ fj)(r)(ei′ ⊗ fj ′)(r′), which is unimodular, being a product of unimodular numbers. m 6= m′ then Lemma 2.4 gives E∗ u,mEu,m′ = 0 and so (27) becomes If we instead have u = u′ and hψu,m,i,j, ψu,m′,i′,j ′i = kδuk2hϕm,i, ϕm′,i′i + hei ⊗ fj, 0(ei′ ⊗ fj ′)i = hϕm,i, ϕm′,i′i, m=1, S+L which is unimodular since {ϕm,i}M i=1 is an ETF of type (K, L, S), and has been scaled so that kϕm,ik2 = S for all m and i. Next, if we instead have u = u′ and m = m′ then Lemma 2.4 gives E∗ u,mEu,m = I and so (27) becomes hψu,m,i,j, ψu,m,i′,j ′i = kδuk2hϕm,i, ϕm,i′i + hei ⊗ fj, ei′ ⊗ fj ′i = hϕm,i, ϕm,i′i + hei, ei′ihfj , fj ′i. (28) In particular, when u = u′, m = m′ and i 6= i′, the fact that {ei}S+L is orthogonal implies i=1 that (28) reduces to hϕm,i, ϕm,i′i, which is unimodular for the same reason as the previous case. The final remaining case is the most interesting: when u = u′, m = m′, i = i′ but j 6= j′, we have 0 = h1 ⊕ fj, 1 ⊕ fj ′i = 1 + hfj, fj ′i and so hfj , fj ′i = −1; when combined with the fact that kϕm,ik2 = S and keik2 = S + L, this implies that in this case (28) becomes hψu,m,i,j, ψu,m,i,j ′i = kϕm,ik2 + keik2hfj , fj ′i = S + (S + L)(−1) = −L, where, in Definition 1.1, we have assumed that L is either 1 or −1. The construction of Theorem 3.2 leads to the concept of ETFs of type (K, L, S). To clarify, the construction (21) and the above proof of its equiangularity is valid for any initial ETF(D, N ) and S+L for some L ∈ {−1, 1}, and S +L divides R = M (U −1) any K-GDD of type M U , provided M = N K−1 . However, it turns out that the first U D rows of the corresponding synthesis operator have squared- norm (W + 1) N S D , whereas the last B rows have squared-norm (W + 1)K(S + L). As such, the equiangular vectors (21) are only a tight frame when K = N S D(S+L) . Applying the techniques of the proof of Theorem 3.1 then leads to the expressions for (D, N ) in terms of (K, L, S) given in (2). These facts are not explicitly discussed in our proof above since any equal-norm vectors that attain the Welch bound are automatically tight. 12 Remark 3.3. There is no apparent value to recursively applying Theorem 3.2. To elaborate, given an ETF(D, N ) of type (K, L, S) and a K-GDD of type M U where M = S(K − 1) + L and U satisfies (20), Theorem 3.2 yields an ETF(D′, N ′) of type (K, L, S′) where S′ = M U −L K−1 . The "M " parameter of this new ETF is thus M ′ = S′(K − 1) + L = (M U − L) + L = M U , and we can apply Theorem 3.2 a second time provided we have a K-GDD of type (M U )U ′ where U ′ satisfies the appropriate analog of (20), namely (K−2)(U ′−1) (S ′+L)(K−1) ∈ Z. (29) Doing so yields an ETF of type (K, −1, S′′) where S′′ = M ′U ′−L . However, under these hypotheses, there is a simpler way to construct an ETF of this same type. Indeed, using the first GDD to fill the holes of the second GDD via Lemma 2.2 produces a K-GDD of type M U U ′ . Moreover, U U ′ is a value of "U " that satisfies (20): since S′ = S + R, K−1 = M U U ′−L K−1 S+L = S ′−S S ′+L S+L + 1 = R S+L + 1 = W + 1 ∈ Z, and this together with (20) and (29) imply (K−2)(U U ′−1) (S+L)(K−1) = (K−2)[U (U ′−1)+(U −1)] (S+L)(K−1) = U (S ′+L) (S+L) (K−2)(U ′−1) (S ′+L)(K−1) + (K−2)(U −1) (S+L)(K−1) ∈ Z. As such, we can combine our original ETF of type (K, L, S) with our K-GDD of type M U U ′ Theorem 3.2 to directly produce an ETF of type (K, L, M U U ′−L ). K−1 via We also point out that, in a manner analogous to how every McFarland difference set can be viewed as a degenerate instance of a Davis-Jedwab difference set [17], every Steiner ETF can be regarded as a degenerate case of the construction of Theorem 3.2. Here, in a manner con- sistent with (2), we regard (D, N ) = (0, 1) as being type (K, L, S) = (K, 1, 0) where K ≥ 1 is arbitrary. Under this convention, M = S(K − 1) + L = 1, R = U −1 K−1 and W = R, meaning we need a K-GDD of type 1U , namely a BIBD(V, K, 1) where V = U . When such a BIBD ex- ists, Theorem 3.2 suggests we let {ϕ1,1} be some fictitious ETF for the nonexistent space F0, 1 scaled so that kϕ1,1k2 = 0. It also suggests we let {Ev}V m=1 be the embedding operators of the BIBD, let e1 = 1, and let {1 ⊕ fj}R j=0 be the columns of a possibly-complex Hadamard matrix of size R. Under these conventions, (21) reduces to a collection of V (R + 1) vectors {ψv,j}V R j=0, where for any v = u and j, the fact that ϕ1,1 lies in a "zero-dimensional space" makes it reasonable to regard j=0 = {ψu,m,i,j}U v=1 = {Eu,m}U 1 m=1, 1 i=1, u=1, u=1, v=1, R ψv,j = ψu,1,1,j = (δv ⊗ ϕ1,1) ⊕(cid:0)Eu,1(e1 ⊗ fj)(cid:1) = Eu,1(1 ⊗ fj) = Eufj . Here, Theorem 3.2 leads us to expect that {Evfj}V v=1, This is indeed the case: as discussed in the previous section, {Evfj}V R j=0 is an ETF of type (K, 1, S+R) = (K, 1, R). R j=0 is by definition a v=1, Steiner ETF for FB where B = V R K , and moreover letting (K, L, S) = (K, 1, R) in (2) gives: S K [S(K − 1) + L] = R K [R(K − 1) + 1] = R K V = B, (S + L)[S(K − 1) + L] = (R + 1)[R(K − 1) + 1] = V (R + 1). In particular, every Steiner ETF is a positive ETF. As a degenerate case of Remark 3.3, we further have that when a Steiner ETF is regarded as being positive, applying Theorem 3.2 to it yields an ETF whose parameters match those of 13 another Steiner ETF. Indeed, since a Steiner ETF arising from a BIBD(V, K, 1) is type (K, 1, R) where R = V −1 K−1, we can only apply Theorem 3.2 to it whenever there exists a K-GDD of type M U where M = R(K − 1) + 1 = V and U satisfies (20). In this case, the resulting ETF is type (K, 1, V U −1 K−1 ). However, under these same hypotheses, we can more simply use the BIBD(V, K, 1) to fill the holes of the K-GDD of type V U via Lemma 2.2 to obtain a BIBD(U V, K, 1), and its corresponding Steiner ETF is type (K, 1, V U −1 K−1 ). 4. Families of positive and negative equiangular tight frames In this section, we use Theorems 3.1 and 3.2 to better our understanding of positive and negative ETFs, in particular proving the existence of the new ETFs given in Theorems 1.2 and 4.4. Recall that by (2), any ETF(D, N ) of type (K, L, S) has D = S K , and so K necessarily divides S(S − L). As we shall see, it is reasonable to conjecture that such ETFs exist whenever this necessary condition is satisfied. K [S(K − 1) + L] = S2 − S(S−L) 4.1. Positive equiangular tight frames In light of Definition 1.1 and Theorem 3.1, an ETF(D, N ) with 1 < D < N is positive if and only if there exists integers K ≥ 1 and S ≥ 2 such that D = S K [S(K − 1) + 1] = S2 − S(S−1) K , N = (S + 1)[S(K − 1) + 1], (30) 2 and K = N S D(S+1) are integers. or equivalently that S =(cid:2) D(N −1) N −D (cid:3) 1 As discussed in the previous section, every regular simplex is a 1-positive ETF and vice versa, and moreover, every Steiner ETF arising from a BIBD(V, K, 1) is type (K, 1, R) where R = V −1 K−1 . Here, the fact that K necessarily divides S(S − 1) = R(R − 1) is equivalent to having that D = B is necessarily an integer. Fisher's inequality also states that a BIBD(V, K, 1) can only exist when K ≤ R. These necessary conditions on the existence of a BIBD(V, K, 1) are known to also be sufficient when K = 2, 3, 4, 5, and also asymptotically sufficient in general: for any K ≥ 2, there exists V0 such that for all V ≥ V0 with the property that R = V −1 K(K−1) , a BIBD(V, K, 1) exists [2]. Moreover, explicit infinite families of such BIBDs are known, including affine geometries, projective geometries, unitals and Denniston designs [25], and each thus gives rise to a corresponding infinite family of (positive) Steiner ETFs. K−1 and B = V R K = V (V −1) That said, not every positive ETF is a Steiner ETF. In particular, for any prime power Q, there is a phased BIBD ETF [22] whose Naimark complements are ETF(D, N ) where D = Q3+1 Q+1 , N = Q3 + 1, S =(cid:0) N −1 D −1(cid:1) N 1 2 =(cid:0) Q3 Q (cid:1) 1 2 = Q, K = N S D(S+1) = (Q+1)Q Q+1 = Q, namely an ETF of type (Q, 1, Q). These parameters match those of a Steiner ETF arising from a projective plane of order Q − 1. However, such an ETF can exist even when no such projective plane exists. For example, such an ETF exists when Q = 7 despite the fact that no projective plane of order Q − 1 = 6 exists. Other ETFs of type (K, 1, S) have K > S, and are thus not Steiner ETFs since the underlying BIBD would necessarily violate Fisher's inequality. To elaborate, when K is a prime power, the requirement that K divides S(S − 1) where S and S − 1 are relatively prime implies that either K divides S or S − 1, and in either case K ≤ S. However, when K is not a prime power, we can sometimes choose it to be a divisor of S(S − 1) that is larger than S. 14 For example, taking K = S(S − 1) in (30) gives D = S2 − 1 and N = (S2 − 1)2 = D2. Such an ETF thus corresponds to a SIC-POVM in a space whose dimension is one less than a perfect square. Such SIC-POVMs are known to exist when S = 2, . . . , 7, 18, and are conjectured to exist for all S [26, 31]. Similarly, taking K =(cid:0)S 2(cid:1) in (30) gives D = S2 −2 and N = 1 We refer to such (D, N ) as being of real maximal type since real-valued examples of such ETFs meet the real-variable version of the Gerzon bound, and are known to exist when S = 3, 5. Remarkably, it is known that a real ETF of this type does not exist when S = 7 [23]. We also caution that there 2 (S2 − 2)(S2 − 1) =(cid:0)D+1 2 (cid:1). is a single pair (D, N ) with N =(cid:0)D+1 ETF(D, N ) exists, namely (D, N ) = (3, 6). 2 (cid:1) that is neither positive or negative despite the fact that an We summarize these facts as follows: Theorem 4.1. An ETF of type (K, 1, S) exists whenever: (a) K = 1 and S ≥ 2, (regular simplices); (b) K ≥ 2 and there exists a BIBD(S(K − 1) + 1, K, 1) (Steiner ETFs [25]), including: Q−1 where Q is a prime power and J ≥ 2 (affine geometries); (i) K = 2, 3, 4, 5 and S ≥ K has the property that K divides S(S − 1); (ii) K = Q and S = QJ −1 (iii) K = Q + 1 and S = QJ −1 (iv) K = Q + 1 and S = Q2 where Q is a prime power (unitals); (v) K = 2J1 and S = 2J2 + 1 where 2 ≤ J1 < J2 (Denniston designs); (vi) K ≥ 2 and S is sufficiently large and has the property that K divides S(S − 1); Q−1 where Q is a prime power and J ≥ 2 (projective geometries); (c) K = Q and S = Q whenever Q is a prime power [22]; (d) K = S(S − 1) where S = 2, . . . , 7, 18 (SIC-POVMs [26, 31]); (e) K =(cid:0)S 2(cid:1) where S = 3, 5 (real maximal type [23]). Because so much is already known regarding the existence of positive ETFs, we could not find any examples where Theorem 3.2 makes a verifiable contribution. As we now discuss, much less is known about negative ETFs, and this gives Theorem 3.2 an opportunity to be useful. 4.2. Negative equiangular tight frames By Definition 1.1 and Theorem 3.1, an ETF(D, N ) with 1 < D < N is negative if and only if there exists integers K ≥ 1 and S ≥ 2 such that D = S K [S(K − 1) − 1] = S2 − S(S+1) K , N = (S − 1)[S(K − 1) − 1], (31) 2 and K = N S D(S−1) are integers. Here, since N D > 1 and S S−1 > 1, 2 S(S − 1), (S − 1)2) and such ETFs exist for any S ≥ 3, or equivalently that S =(cid:2) D(N −1) N −D (cid:3) we actually necessarily have that K ≥ 2. When K = 2, (31) becomes (D, N ) = ( 1 1 being the Naimark complements of ETFs of type (2, 1, S − 1). In the K = 3 case, (31) becomes D = S(2S−1) 3 = S2 − S(S+1) 3 , N = (S − 1)(2S − 1). 15 For D to be an integer, we necessarily have S ≡ 0, 2 mod 3. Moreover, for any S ≥ 2 with S ≡ 0, 2 mod 3, an ETF of type (3, −1, S) exists. Indeed, the recent paper [19] gives a way to modify the Steiner ETF arising from a BIBD(V, 3, 1) to yield a Tremain ETF(D, N ) with D = 1 6 (V + 2)(V + 3), N = 1 2 (V + 1)(V + 2), for any V ≥ 3 with V ≡ 1, 3 mod 6. Such an ETF is type (3, −1, S) with S = 1 2 (V + 3). We also note that for every J ≥ 1, there is a harmonic ETF(D, N ) arising from a Spence difference set [47] with (D, N ) = ( 1 2 3J+1(3J+1 − 1)), and such an ETF is type (3, −1, S) where S = 1 2 (3J+1 + 1). It remains unclear whether any Spence ETFs are unitarily equivalent to special instances of Tremain ETFs. 2 3J (3J+1 + 1), 1 In the K = 4 case, for any positive integer J, Davis and Jedwab [17] give a difference set whose harmonic ETF has parameters (3) and so is a type (4, −1, S) ETF where S = 1 3 (22J+1 +1). Beyond these examples, a few other infinite families of negative ETFs are known to exist. In particular, in order for the expression for D given in (31) to be an integer, K necessarily divides S(S + 1), and so it is natural to consider the special cases where S = K and S = K − 1. When S = K − 1, (31) gives that ETFs of type (K, −1, K − 1) have D = (K − 1)(K − 2), N = K(K − 2)2. Remarkably, these are the same (D, N ) parameters as those of an ETF of type (K − 2, 1, K − 1). In particular, every Steiner ETF arising from an affine plane of order Q is both of (positive) type (Q, 1, Q + 1) and (negative) type (Q + 2, −1, Q + 1). More generally, a Steiner ETF arising from a BIBD(V, K, 1) has (D, N ) = (B, V (R + 1)) where R = V −1 K , and so is only negative when K−1 and B = V R D(S−1) = V (R+1)R B(R−1) = K(R+1) R−1 N S (32) is an integer. When R is even, R − 1 and R + 1 are relatively prime, and this can only occur when R − 1 divides K. Here, since Fisher's inequality gives K ≤ R, this happens precisely when either R = K = 2 or R = K + 1, namely when the underlying BIBD is a BIBD(3, 2, 1) or is an affine plane of odd order K, respectively. Meanwhile, when R is odd, R − 1 and R + 1 have exactly one prime factor in common, namely 2, and (32) is an integer precisely when 1 2 (R − 1) divides K. Since K ≤ R, this happens precisely when either R = K = 3, R = K + 1 or R = 2K + 1, namely when the underlying BIBD is the projective plane of order 2, an affine plane of even order, or is a BIBD(V, K, 1) where V = (2K + 1)(K − 1) + 1 = K(2K − 1) for some K ≥ 2, respectively. With regard to the latter, it seems to be an open question whether a BIBD(K(2K − 1), K, 1) exists for every K ≥ 2, though a Denniston design provides one whenever K = 2J for some J ≥ 1, and they are also known to exist when K = 3, 5, 6, 7 [39]. For such ETFs, (32) becomes K + 1, meaning they are type (K ′, −1, 2K ′ − 1) where K ′ = K + 1. Meanwhile, when S = K, (31) gives that ETFs of type (K, −1, K) have D = K 2 − K − 1, N = (K − 1)(K 2 − K − 1). (33) The recently-discovered hyperoval ETFs of [24] are instances of such ETFs whenever K = 2J + 1 for some J ≥ 1. In the K = 4 case, (33) becomes (D, N ) = (11, 33), and this seems to be the smallest set of positive or negative parameters for which the existence of a corresponding ETF remains an open problem. 16 Apart from these examples, it seems that only a finite number of other negative ETFs are known to exist. For example, since K necessarily divides S(S + 1), it is natural to also consider the cases where K = S(S+1) and K =(cid:0)S+1 In particular, every positive SIC-POVM is also negative. Similarly, when K = (cid:0)S+1 N =(cid:0)D+1 2 (cid:1). When K = S(S+1), (31) becomes (D, N ) = (S2−1, (S2−1)2). 2 (cid:1), (31) gives 2 (cid:1) where D = S2 − 2, meaning every positive (D, N ) of real maximal type is also negative. The only other example of a negative ETF that we found in the literature was an ETF(22, 176), which has type (10, −1, 5), and arises from a particular SRG. When searching tables of known ETFs such as [23], it is helpful to note that most positive and negative ETFs have redundancy N D > 2, with the only exceptions being 1-positive ETFs (regular simplices), 2-negative ETFs (Naimark complements of 2-positive ETFs), ETF(2, 3) when regarded as type (3, −1, 2), and ETF(5, 10), which are type (3, −1, 3). Indeed, when K ≥ 2, any K-positive ETF has N > K ≥ 2. Meanwhile, when K ≥ 3, any ETF of type (K, −1, S) only has K(S−1) = N K−2 . Since K necessarily divides S(S + 1) where S ≥ 2, such an ETF can only exist when K = 3 and S = 2, 3. We summarize these previously-known constructions of negative ETFs as follows: D = K(S+1) D ≤ 2 when S ≤ K S S Theorem 4.2. An ETF of type (K, −1, S) exists whenever: (a) K = 2 and S ≥ 3 (Naimark complements of 2-positive ETFs); (b) K = 3 and S ≥ 2 with S ≡ 0, 2 mod 3 (Tremain ETFs [19] and Spence harmonic ETFs [47]); (c) K = 4 and S = 1 3 (22J+1 + 1) for some J ≥ 1 (Davis-Jedwab harmonic ETFs [17]); (d) K = Q + 2 and S = K − 1 where Q is a prime power (Steiner ETFs from affine planes [25]); (e) K = 2J + 1 and S = 2K − 1 where J ≥ 1 (Steiner ETFs from Denniston designs [25]); (f) K = 2J + 1 and S = K where J ≥ 1 (hyperoval ETFs [24]); (g) K = S(S + 1) where S = 2, . . . , 7, 18 (SIC-POVMs [26]); (h) K =(cid:0)S+1 2 (cid:1) where S = 3, 5 (real maximal type [23]); (i) (K, S) = (4, 7), (6, 11), (7, 13), (8, 15), (10, 5) (various other ETFs [23]). From this list, we see that for any K ≥ 5, the existing literature provides at most a finite number of K-negative ETFs. Theorem 3.2(a) implies that many more negative ETFs exist: if an ETF of type (K, −1, S) exists, then an ETF of type (K, −1, M U +1 K−1 ) exists for all sufficiently large U that satisfy (19) and (20). Combining this fact with Theorem 4.2 immediately gives: Corollary 4.3. There exists an infinite number of K-negative ETFs whenever: (a) K = Q + 2 where Q is a prime power; (b) K = Q + 1 where Q is an even prime power; (c) K = 2, 8, 12, 20, 30, 42, 56, 342. In particular, we now know that there are an infinite number of values of K for which an infinite number of K-negative ETFs exist, with K = 14 being the smallest open case. With these asymptotic existence results in hand, we now focus on applying Theorem 3.2 with explicit GDDs. For example, the "Mercedes-Benz" regular simplex ETF(2, 3) is type (K, L, S) = (3, −1, 2) and so has M = S(K − 1) + L = 3. Since S + L = 1 divides K − 2 = 1, Theorem 3.2 can be applied with any 3-GDD of type 3U so as to produce an ETF of type (K, L, M U −L 2 (3U + 1)), and moreover that the necessary conditions (7) on the existence of such a GDD reduce to (19), K−1 ) = (3, −1, 1 17 2 (U − 1) ∈ Z and 1 namely to having U ≥ 3, 1 2 U (U − 1) ∈ Z. In fact, such GDDs are known to exist whenever these necessary conditions are satisfied [27], namely when U ≥ 3 is odd. (This also follows from the fact that such GDDs are equivalent to the incidence structures obtained by removing a parallel class from a resolvable Steiner triple system.) That is, writing U = 2J + 1 for some J ≥ 1, we can apply Theorem 3.2 with an ETF(2, 3) and a known 3-GDD of type 32J+1 to produce an ETF of type (3, −1, 1 2 (3U + 1)) = (3, −1, 3J + 2). In summary, applying Theorem 3.2 to an ETF of type (3, −1, 2) produces ETFs of type (3, −1, S) for any S ≡ 2 mod 3, and so recovers the parameters of "half" of all possible 3-negative ETFs, cf. Theorem 4.2 and [19], including the parameters of all harmonic ETFs arising from Spence difference sets. To instead recover some of the ETFs of type (3, −1, S) with S ≡ 0 mod 3, one may, for example, apply Theorem 3.2 to the well-known ETF(5, 10), which is type (3, −1, 3). In order to obtain ETFs with verifiably new parameters, we turn our attention to applying Theorem 3.2 to known K-negative ETFs with K ≥ 4. Here, the limiting factor seems to be a lack of knowledge regarding uniform K-GDDs: while the literature has much to say when K = 3, 4, 5 [27], we are relegated to well-known simple constructions involving Lemmas 2.1 and 2.2 whenever K > 5. As such, we consider the K = 4, 5 cases separately from those with K > 5: Proof of Theorem 1.2. The ETF(6, 16) is type (4, −1, 3), and so we can apply Theorem 3.2 when- ever there exists a 4-GDD of type 8U where U satisfies (20). By Theorem 3.2, the known necessary conditions (7) on the existence of such GDDs reduce to (19): U ≥ 4, U −1 3 ∈ Z, U (U −1) 3 = 4U (U −1) 4(3) ∈ Z, 3 = 2(U −1) namely to having U ≥ 4 and U ≡ 1 mod 3. These necessary conditions on the existence of 4-GDDs of type 8U are known to be sufficient [27]. Moreover, for any such U , we have (20) is automatically satisfied since U −1 2(3) ∈ Z. Altogether, for any U ≥ 4 with U ≡ 1 mod 3, we can apply Theorem 3.2 to the ETF of type (4, −1, 3) with a 4-GDD of type 8U , and doing so produces an ETF of type (4, −1, 1 3 (8U + 1)). Here, letting U = 1 recovers the parameters of the original ETF. Overall, writing U = 3J + 1 for some J ≥ 0, this means that an ETF of type (4, −1, S) exists whenever S = 1 3 [8(3J + 1) + 1] = 8J + 3 for any J ≥ 0, namely whenever S ≡ 3 mod 8. In particular, for any J ≥ 1 we can take U = 4J−1 to obtain an ETF of type (4, −1, S) where S = 1 3 (22J−1 +1). This means that applying Theorem 3.2 to the ETF of type (4, −1, 3) recovers the parameters (3) of harmonic ETFs corresponding to Davis-Jedwab difference sets. 3 (8U + 1) = 1 3 (8U +1) = 1 In light of Remark 3.3, applying Theorem 3.2 to any ETF of type (4, −1, S) with S ≡ 3 mod 8 simply recovers a subset of the ETF types obtained by applying Theorem 3.2 to the ETF of type (4, −1, 3). As such, to obtain more 4-negative ETFs via Theorem 3.2, we need to apply it to initial ETFs that lie outside of this family. By Theorem 4.2, the existing literature gives one such set of parameters, namely ETFs of type (4, −1, 7), which have (D, N ) = (35, 120). Here, M = 20, and a 4-GDD of type 20U can only exist if U satisfies (19): U ≥ 4, U −1 3 ∈ Z, U (U −1) 3 = 4U (U −1) 4(3) ∈ Z. Moreover, these necessary conditions are known to be sufficient [27], meaning Theorem 3.2 can be applied whenever U also satisfies (20), namely U −1 6(3) ∈ Z. Thus, for any U ≡ 1 mod 9, Theorem 3.2 yields an ETF of type (4, −1, 1 3 (20U + 1)). In particular, an ETF of type (4, −1, S) exists for any S ≡ 7 mod 60. ETFs of type (4, −1, S) with S ≡ 67 mod 120 thus arise from both constructions; in the statement of Theorem 1.2, we elect to not remove the overlapping values of 9 = 2(U −1) 18 S from either family so as to not emphasize one family over the other, and possibly make it easier for future researchers to identify potential patterns. In a similar manner, as summarized in Theorem 4.2, the existing literature provides ETFs of type (5, −1, S) for exactly three values of S, namely ETF(12, 45), ETF(19, 76) and ETF(63, 280) which have S = 4, 5, 9 and so M = 15, 19, 35, respectively. For these particular values of M , the corresponding necessary conditions (7) on the existence of 5-GDDs of type M U are known to be sufficient [27], meaning we only need U to satisfy (19) and (20) for the corresponding value of S, namely to satisfy U ≥ 5, 1 4 (U − 1) ∈ Z and that 5U (U −1) 5(4) ∈ Z, 6U (U −1) 5(4) ∈ Z, 10U (U −1) 5(4) ∈ Z, 3(U −1) 3(4) ∈ Z, when S = 4, 3(U −1) 4(4) ∈ Z, when S = 5, 3(U −1) 8(4) ∈ Z, when S = 9. An ETF of type (5, −1, S) thus exists when S = 1 with U ≡ 1, 65 mod 80, or S = 1 exists when S ≡ 4 mod 15, or S ≡ 5, 309 mod 380, or S ≡ 9 mod 280. 4 (19U + 1) 4 (35U + 1) with U ≡ 1 mod 32. That is, an ETF of type (5, −1, S) 4 (15U + 1) with U ≡ 1 mod 4, or S = 1 In certain special cases, these techniques yield real ETFs: Proof of Theorem 1.3. The ETF (21) constructed in Theorem 3.2 is clearly real when the initial ETF(D, N ) {ϕn}N n=1 and the Hadamard matrices of size S + L and W + 1 are real. In particular, a real ETF(63, 280) exists [23], and such ETFs are type (5, −1, 9). Since a real Hadamard matrix of size S + L = 8 exists, then letting M = S(K − 1) + L = 35, Theorem 3.2 yields a real ETF of type (5, −1, 1 4 (35U + 1)) whenever there exists a 5-GDD of type 35U where U satisfies (20) and there exists a real Hadamard matrix of size H = W + 1 = R S+L + 1 = M (U −1) (S+L)(K−1) + 1 = 35(U −1) 8(4) + 1 = 35U −3 . 32 Here, we recall from the proof of Theorem 1.2 that such GDDs exist whenever U ≡ 1 mod 32, namely whenever H ≡ 1 mod 35. Altogether, since 1 ) + 1] = 8H + 1, Theorem 3.2 yields a real ETF of type (5, −1, 8H + 1) whenever there exists a Hadamard matrix of size H when H ≡ 1 mod 35. An infinite number of such Hadamard matrices exist: since 17 is relatively prime to 140, Dirichlet's theorem implies an infinite number of primes Q ≡ 17 mod 140 exist, and each has the property that Q ≡ 1 mod 4, meaning that Paley's construction yields a real Hadamard matrix of size 2(Q + 1) ≡ 36 mod 280, in particular of size 2(Q + 1) ≡ 1 mod 35. 4 (35U + 1) = 1 4 [35( 32H+3 35 Similarly, a real ETF(7, 28) exists [23], is type (6, −1, 3), and there exists a real Hadamard matrix of size S + L = 2. Letting M = 14, Theorem 3.2 thus yields a real ETF of type (6, −1, 1 5 (14U + 1)) whenever there exists a 6-GDD of type 14U where U satisfies (20) and there exists a real Hadamard matrix of size H = W + 1 = 1 5 (7U − 2). Here, the necessary condi- tions (19) and (20) on the existence of such a GDD reduce to having U ≡ 1, 6 mod 15, namely to having H ≡ 1, 8 mod 21. Since K = 6, these necessary conditions are not known to be sufficient. Nevertheless, they are asymptotically sufficient: there exists U0 such that for all U ≥ U0 with U ≡ 1, 6 mod 15, there exists a 6-GDD of type 14U where U satisfies (20). As such, there exists H0 such that for all H ≥ H0 with H ≡ 1, 8 mod 21, if there exists a real Hadamard matrix of size H, then there exists a real ETF of type (6, −1, 1 5 (14U + 1)) = (6, −1, 2H + 1). As above, an infinite number of such Hadamard matrices exist: since gcd(73, 84) = 1, there are an infinite 19 number of primes Q ≡ 73 mod 84, and each has the property that Q ≡ 1 mod 4, meaning Paley's construction yields a real Hadamard matrix of size 2(Q + 1) ≡ 148 mod 168, in particular of size 2(Q + 1) ≡ 1 mod 21. Applying these same techniques to real ETF(22, 176) and ETF(23, 276) yields the infinite fam- ilies stated in (c) and (d) of the result. We have now seen that the construction of Theorem 3.2 recovers all Steiner ETFs as a degen- erate case, recovers the parameters of "half" of all Tremain ETFs including those of all harmonic ETFs arising from Spence difference sets, and also recovers the parameters (3) of harmonic ETFs arising from the Davis-Jedwab difference sets. This is analogous to how the approach of [17] unifies McFarland, Spence and Davis-Jedwab difference sets with those with parameters (3). From this perspective, the value of the generalization of [17] given in Theorem 3.2 is that it permits weaker conclusions to be drawn from weaker assumptions: while [17] forms new difference sets (i.e., new harmonic ETFs) by combining given difference sets with building sets formed from a collection of hyperplanes, Theorem 3.2 forms new ETFs by combining given ETFs with GDDs. In fact, a careful read of [17] indicates that the building sets used there to produce difference sets with parameters (3) are related to 4-GDDs of type 84J −1 obtained by recursively using TD(4, 22j+1) to fill the holes of TD(4, 22j+3) for every j = 1, . . . , J − 1 via Lemma 2.2. Alternatively, such GDDs can be constructed by using Lemma 2.1 to combine a TD(4, 8) with a BIBD(4J−1, 4, 1) arising from an affine geometry. We now generalize these approaches, using Lemmas 2.1 and 2.2 to produce the GDDs needed to apply Theorem 3.2 to several known K-negative ETFs with K > 5: Theorem 4.4. If an ETF of type (K, −1, S) exists where K−2 M = S(K − 1) − 1, then an ETF of type (K, −1, M U +1 K−1 ) exists when either: S−1 ∈ Z, and a TD(K, M ) exists where (i) U = 1 or U = K; (ii) a BIBD(U, K, 1) exists; (iii) U = K J for some J ≥ 2, provided a TD(K, M K j ) exists for all j = 1, . . . , J − 1. As a consequence, an ETF of type (K, −1, S) exists when either: (a) K = 6, S = 1 (b) K = 6, S = 1 (c) K = 7, S = 1 (d) K = 10, S = 1 (e) K = 12, S = 1 5 (9U + 1) where either U = 6J for some J ≥ 0 or a BIBD(U, 6, 1) exists; 5 (24U + 1) where either U = 6J for some J ≥ 0 or a BIBD(U, 6, 1) exists; 6 (35U + 1) where either U = 7J for some J ≥ 0 or a BIBD(U, 7, 1) exists; 9 (80U + 1) where either U = 10J for some J ≥ 0 or a BIBD(U, 10, 1) exists; 11 (32U + 1) when either U = 12J for some J ≥ 0 or a BIBD(U, 12, 1) exists. Proof. Since K−2 S−1 ∈ Z and M = S(K − 1) − 1, any K-GDD of type M U can be combined with the given ETF of type (K, −1, S) via Theorem 3.2 in order to construct an ETF of type (K, −1, M U +1 K−1 ). (Since M +1 K−1 = S, such an ETF also exists when U = 1, namely the given initial ETF.) For instance, the given TD(K, M ) is a K-GDD of type M K , and so such an ETF exists when U = K. Other examples of such GDDs can be constructed by combining the given TD(K, M ) with any BIBD(U, K, 1)-a K-GDD of type 1U -via Lemma 2.1. In particular, when K is a prime power, we can construct these BIBDs from affine geometries of order K to produce examples of such GDDs with U = K J for any J ≥ 2. GDDs with these parameters also sometimes exist even when K is not a prime power: if a TD(K, M K j) exists for all j = 1, . . . , J − 1, then recursively using 20 TD(K, M K j−1) to fill the holes of TD(K, M K j ) for all j = 1, . . . , J − 1 via Lemma 2.1 gives a K-GDD of type M K J . We now apply these ideas to some known K-negative ETFs with K > 5, organized according to the families of Theorem 4.2. In particular, ETFs of type (Q + 2, −1, Q + 1) exist whenever Q is a prime power, and S − 1 = Q divides K − 2 = Q. For such ETFs, M = Q(Q + 2) and a TD(Q + 2, Q(Q + 2)) equates to Q MOLS of size Q(Q + 2), which is known to occur when Q = 2, 3, 4, 5, 8 [1]. (When Q and Q + 2 are both prime powers, the standard method [40] only produces Q − 1 MOLS of size Q(Q + 2).) Taking Q = 2, 3 recovers a subset of the ETFs produced in Theorem 1.2, and so we focus on Q = 4, 5, 8. In particular, for these values of Q, the above 1 Q+1 [Q(Q + 2)U + 1]) when either U = 1, U = Q + 2, or methods yield an ETF of type (Q + 2, −1, a BIBD(U, Q + 2, 1) exists. Moreover, even when Q + 2 is not a prime power, such an ETF with U = (Q+2)J exists, provided a TD(Q+2, Q(Q+2)j+1) exists for all j = 1, . . . , J −1. When Q = 4, 8, this occurs for any J ≥ 2, since for any j ≥ 1, the number of MOLS of size (4)6j+1 = 2j+33j+1 is at least min{2j+3, 3j+1} − 1 ≥ 7, while the number of MOLS of size (8)10j+1 = 2j+45j+1 is at least min{2j+4, 5j+1} − 1 ≥ 23. Other new explicit infinite families of negative ETFs arise from SIC-POVMs, which by Theo- rem 4.2 are ETFs of type (K, −1, S) where K = S(S + 1). Such ETFs always have the property that S − 1 divides K − 2 = (S − 1)(S + 2), meaning we can apply the above ideas whenever there exists a TD(K, M ) where M = (S − 1)(S + 1)2. Such TDs exist when S = 2, 3, being TD(6, 9) and TD(12, 32), respectively. When S = 2 in particular, the above methods yield an ETF of type (6, −1, 1 5 (9U + 1)) when either U = 1, U = 6, a BIBD(U, 6, 1) exists, or U = 6J for some J ≥ 2. This final family arises from the fact that a TD(6, (9)6j ) exists for all j ≥ 1: there are at least 5 MOLS of size 54 = (9)6, at least 8 MOLS of size 324 = 9(62) (since there are 8 MOLS of size 9, and at least 8 MOLS of size 36) [1], while for any j ≥ 3, the number of MOLS of size (9)6j = 2j3j+2 is at least 2j − 1 ≥ 7. Similarly, taking S = 3 yields an ETF of type (12, −1, 1 11 (32U + 1)) when either U = 1, U = 12, a BIBD(U, 12, 1) exists, or U = 12J for some J ≥ 2: the number of MOLS of size 32(12) = 384 and 32(12)2 = 4608 is at least 15 [1], while for j ≥ 3, the number of MOLS of size 32(12)j = 22j+53j is at least min{22j+5, 3j } − 1 ≥ 26. To be clear, we view the ETFs produced in Theorem 4.4 as a "proof of concept," and believe that GDD experts will be able to find many more examples of new ETFs using Theorems 3.2 and 4.2. For this reason, we have omitted technical cases where an ETF of type (K, −1, S) and a TD(K, M ) exist where M = S(K − 1) − 1, but S − 1 does not divide K − 2, such as when S = K = 9, and when S = 2K − 1 where K = 6, 7, 8, 9, 17, 65537. In such cases, one can still combine the TD with a BIBD(U, K, 1) via Lemma 2.1 to produce a K-GDD of type M U , but (20) is not automatically satisfied. We also point out that though Theorem 3.2 is a generalization of [17], the new ETFs we have found here are disjoint from those produced by another known generalization of this same work. In particular, the parameters (3) of Davis-Jedwab difference sets can be regarded as the Q = 2 case of the more general family: D = Q2J−1( 2Q2J +Q−1 Q+1 ), N = 4Q2J ( Q2J −1 Q2−1 ), (34) where J ≥ 1 and Q ≥ 2 are integers. In particular, in [14], Chen generalizes the theory of [17] in a way that produces difference sets with parameters (34) for any J ≥ 1 and any Q that is either a power of 3 or any even power of an odd prime. For this reason, difference sets with parameters 21 of the form (34) are said to be Davis-Jedwab-Chen difference sets [35]. The inverse Welch bounds for the corresponding harmonic ETFs and their Naimark complements are always integers: However, such ETFs are seldom positive or negative. Indeed, a direct computation reveals (cid:2) (N −D)(N −1) D (cid:3) 1 2 = 2Q2J −Q−1 . Q−1 1 , Q+1 2 = 2Q2J +Q−1 S =(cid:2) D(N −1) N −D (cid:3) D(S+1) = 2 + 2Q(Q2J −2−1) N S (Q−1)(Q2J −1+1) , N S D(S−1) = 2 + 2 Q−1 . 2 (9J + 1)) or (4, −1, 1 Thus, by Theorem 3.1, the only such ETFs that are positive or negative are of type (2, 1, 2Q − 1), (3, −1, 1 3 (22J+1 + 1)), corresponding to the special cases of (34) where J = 1, Q = 3 and Q = 2, respectively. That is, the only overlap between the K-negative ETFs constructed in Theorems 1.2 and 4.4 and the ETFs constructed in [14] are those with parameters (34). 5. Conclusions Theorems 1.2 and 4.4 make some incremental progress towards resolving the ETF existence problem. By comparing the existing ETF literature, as summarized in [23] for example, against Theorems 4.1 and 4.2, we find that every known ETF is either an orthonormal basis, or has the property that either it or its Naimark complement is a regular simplex, has N = 2D or N = 2D±1, is a SIC-POVM, arises from a difference set, or is either positive or negative. In particular, every known ETF(D, N ) with N > 2D > 2 is either a SIC-POVM (N = D2), or is a harmonic ETF, or has N = 2D + 1 where D is odd, or is either positive or negative. This fact, along with the known nonexistence of the ETF(3, 8) [51], and the available numerical evidence [53], leads us to make the following conjecture: Conjecture 5.1. If N > 2D > 2, an ETF(D, N ) exists if and only if N = D2 or D(D−1) (D, N ) is positive or negative (Definition 1.1). N −1 ∈ Z or This is a substantial strengthening of an earlier conjecture made by one of the authors, namely that if N > D > 1 and an ETF(D, N ) exists, then one of the three numbers D, N − D and N − 1 necessarily divides the product of the other two. To help resolve Conjecture 5.1, it would in particular be good to know whether an ETF(9, 25) exists: this is the smallest value of D for which there exists an N such that N − 1 divides D(D − 1) but no known ETF exists. For context, we note that there are numerous pairs (D, N ) for which N − 1 divides D(D − 1) and an ETF(D, N ) is known to exist, despite the fact that a difference set of that size does not exist [30], including (35, 120), (40, 105), (45, 100), (63, 280), (70, 231), (77, 210), (91, 196), (99, 540), (130, 560), (143, 924), (176, 561), (187, 528), (208, 1105), (231, 484), (247, 780), (260, 741). It would also be good to know whether an ETF(11, 33) exists, since such an ETF would be type (4, −1, 4): as detailed in the previous section, an ETF of type (K, L, S) can only exist when K divides S(S − L), and moreover this necessary condition for existence is known to be sufficient when L = 1 and K = 1, 2, 3, 4, 5, as well as when L = −1 and K = 2, 3. The evidence also supports an analogous conjecture in the real case. In fact, comparing the relevant literature [12, 23] against Theorems 4.1 and 4.2, we find that every known real ETF(D, N ) with N > 2D > 2 is either positive or negative. Moreover, when 1 < D < N − 1, N 6= 2D and 22 a real ETF(D, N ) exists, then it and its Naimark complements' Welch bounds are necessarily the reciprocals of odd integers [50]. In particular, any such ETF automatically satisfies one of the two integrality conditions given in Theorem 3.1 that characterize positive and negative ETFs. Conversely, since S(K − 1) + KL is odd whenever S is odd, Theorem 3.1 implies that any ETF of type (K, L, S) satisfies the necessary conditions of [50] when S is odd. That said, real ETF(19, 76), ETF(20, 96) and ETF(47, 1128) do not exist [3, 4, 23], despite the fact that (19, 76) is type (5, −1, 5), (20, 96) is both type (4, 1, 5) and type (6, −1, 5), and (47, 1128) is both type (21, 1, 7) and type (28, −1, 7). These facts suggest the following analog of Conjecture 5.1: Conjecture 5.2. If N > 2D > 2 and a real ETF(D, N ) exists, then (D, N ) is positive or negative. Acknowledgments The views expressed in this article are those of the authors and do not reflect the official policy or position of the United States Air Force, Department of Defense, or the United States Government. This work was partially supported by the Summer Faculty Fellowship Program of the United States Air Force Research Laboratory. References [1] R. J. R. Abel, C. J. Colbourn, J. H. Dinitz, Mutually Orthogonal Latin Squares (MOLS), in: C.J. Colbourn, J.H. Dinitz (Eds.), Handbook of Combinatorial Designs, Second Edition (2007) 160–193. [2] R. J. R. Abel, M. Greig, BIBDs with small block size, in: C.J. Colbourn, J.H. Dinitz (Eds.), Handbook of Combinatorial Designs, Second Edition (2007) 72–79. [3] J. Azarija, T. Marc, There is no (75,32,10,16) strongly regular graph, arXiv:1509.05933. [4] J. Azarija, T. Marc, There is no (95,40,12,20) strongly regular graph, arXiv:1603.02032. [5] W. U. Bajwa, R. Calderbank, D. G. Mixon, Two are better than one: fundamental parameters of frame coherence, Appl. Comput. Harmon. Anal. 33 (2012) 58-78. [6] A. S. Bandeira, M. Fickus, D. G. Mixon, P. Wong, The road to deterministic matrices with the Restricted Isometry Property, J. Fourier Anal. Appl. 19 (2013) 1123–1149. [7] A. Barg, A. Glazyrin, K. A. Okoudjou, W.-H. Yu, Finite two-distance tight frames, Linear Algebra Appl. 475 (2015) 163–175. [8] B. G. Bodmann, H. J. Elwood, Complex equiangular Parseval frames and Seidel matrices containing pth roots of unity, Proc. Amer. Math. Soc. 138 (2010) 4387–4404. [9] B. G. Bodmann, V. I. Paulsen, M. Tomforde, Equiangular tight frames from complex Seidel matrices con- taining cube roots of unity, Linear Algebra Appl. 430 (2009) 396–417. [10] C. Bracken, G. McGuire, H. Ward, New quasi-symmetric designs constructed using mutually orthogonal Latin squares and Hadamard matrices, Des. Codes Cryptogr. 41 (2006) 195–198. [11] A. E. Brouwer, Strongly regular graphs, in: C. J. Colbourn, J. H. Dinitz (Eds.), Handbook of Combinatorial Designs, Second Edition (2007) 852-868. [12] A. E. Brouwer, Parameters of Strongly Regular Graphs, http://www.win.tue.nl/∼aeb/graphs/srg/ [13] K. I. Chang, An existence theory for group divisible designs, Ph.D. Thesis, The Ohio State University, 1976. [14] Y. Q. Chen, On the existence of abelian Hadamard difference sets and a new family of difference sets, Finite Fields Appl. 3 (1997) 234–256. [15] D. Corneil, R. Mathon, eds., Geometry and combinatorics: Selected works of J. J. Seidel, Academic Press, 1991. [16] G. Coutinho, C. Godsil, H. Shirazi, H. Zhan, Equiangular lines and covers of the complete graph, Linear Algebra Appl. 488 (2016) 264–283. [17] J. A. Davis, J. Jedwab, A unifying construction for difference sets, J. Combin. Theory Ser. A 80 (1997) 13–78. [18] C. Ding, T. Feng, A generic construction of complex codebooks meeting the Welch bound, IEEE Trans. Inform. Theory 53 (2007) 4245–4250. [19] M. Fickus, J. Jasper, D. G. Mixon, J. D. Peterson, Tremain equiangular tight frames, J. Combin. Theory Ser. A 153 (2018) 54-66. 23 [20] M. Fickus, J. Jasper, D. G. Mixon, J. D. Peterson, Hadamard equiangular tight frames, submitted, arXiv:1703.05353. [21] M. Fickus, J. Jasper, D. G. Mixon, J. D. Peterson, C. E. Watson, Equiangular tight frames with centroidal symmetry, to appear in Appl. Comput. Harmon. Anal. [22] M. Fickus, J. Jasper, D. G. Mixon, J. D. Peterson, C. E. Watson, Polyphase equiangular tight frames and abelian generalized quadrangles, to appear in Appl. Comput. Harmon. Anal. [23] M. Fickus, D. G. Mixon, Tables of the existence of equiangular tight frames, arXiv:1504.00253 (2016). [24] M. Fickus, D. G. Mixon, J. Jasper, Equiangular tight frames from hyperovals, IEEE Trans. Inform. Theory 62 (2016) 5225–5236. [25] M. Fickus, D. G. Mixon, J. C. Tremain, Steiner equiangular tight frames, Linear Algebra Appl. 436 (2012) 1014–1027. [26] C. A. Fuchs, M. C. Hoang, B. C. Stacey, The SIC question: history and state of play, Axioms 6 (2017) 21. [27] G. Ge, Group divisible designs, in: C. J. Colbourn, J. H. Dinitz (Eds.), Handbook of Combinatorial Designs, Second Edition (2007) 255–260. [28] C. D. Godsil, Krein covers of complete graphs, Australas. J. Combin. 6 (1992) 245–255. [29] J. M. Goethals, J. J. Seidel, Strongly regular graphs derived from combinatorial designs, Can. J. Math. 22 (1970) 597–614. [30] D. Gordon, La Jolla Covering Repository, https://www.ccrwest.org/diffsets.html. [31] M. Grassl, A. J. Scott, Fibonacci-Lucas SIC-POVMs, J. Math. Phys. 58 (2017) 122201. [32] R. B. Holmes, V. I. Paulsen, Optimal frames for erasures, Linear Algebra Appl. 377 (2004) 31–51. [33] J. W. Iverson, J. Jasper, D. G. Mixon, Optimal line packings from nonabelian groups, submitted, arXiv:1609.09836. [34] J. Jasper, D. G. Mixon, M. Fickus, Kirkman equiangular tight frames and codes, IEEE Trans. Inform. Theory 60 (2014) 170-181. [35] D. Jungnickel, A. Pott, K. W. Smith, Difference sets, in: C. J. Colbourn, J. H. Dinitz (Eds.), Handbook of Combinatorial Designs, Second Edition (2007) 419–435. [36] J. H. van Lint, J. J. Seidel, Equilateral point sets in elliptic geometry, Indag. Math. 28 (1966) 335–348. [37] E. R. Lamken, R. M. Wilson, Decompositions of edge-colored complete graphs, J. Combin. Theory Ser. A 89 (2000) 149–200. [38] P. W. H. Lemmens, J. J. Seidel, Equiangular lines, J. Algebra 24 (1973) 494–512. [39] R. Mathon, A. Rosa, 2 − (v, k, λ) designs of small order, in: C. J. Colbourn, J. H. Dinitz (Eds.), Handbook of Combinatorial Designs, Second Edition (2007) 25–58. [40] H. F. MacNeish, Euler Squares, Ann. of Math. 23 (1922) 221–227. [41] R. L. McFarland, A family of difference sets in non-cyclic groups, J. Combin. Theory Ser. A 15 (1973) 1–10. [42] G. McGuire, Quasi-symmetric designs and codes meeting the Grey-Rankin bound, J. Combin. Theory Ser. A 78 (1997) 280–291. [43] Hedvig Moh´acsy, The asymptotic existence of group divisible designs of large order with index one, J. Combin. Theory Ser. A 118 (2011) 1915–1924. [44] J. M. Renes, Equiangular tight frames from Paley tournaments, Linear Algebra Appl. 426 (2007) 497–501. [45] J. M. Renes, R. Blume-Kohout, A. J. Scott, C. M. Caves, Symmetric informationally complete quantum measurements, J. Math. Phys. 45 (2004) 2171–2180. [46] J. J. Seidel, A survey of two-graphs, Coll. Int. Teorie Combin., Atti dei Convegni Lincei 17, Roma (1976) 481–511. [47] E. Spence, A family of difference sets, J. Combin. Theory Ser. A 22 (1977) 103–106. [48] T. Strohmer, A note on equiangular tight frames, Linear Algebra Appl. 429 (2008) 326-330. [49] T. Strohmer, R. W. Heath, Grassmannian frames with applications to coding and communication, Appl. Comput. Harmon. Anal. 14 (2003) 257–275. [50] M. A. Sustik, J. A. Tropp, I. S. Dhillon, R. W. Heath, On the existence of equiangular tight frames, Linear Algebra Appl. 426 (2007) 619–635. [51] F. Szollosi, All complex equiangular tight frames in dimension 3, arXiv:1402.6429. [52] J. A. Tropp, Complex equiangular tight frames, Proc. SPIE 5914 (2005) 591401/1–11. [53] J. A. Tropp, I. S. Dhillon, R. W. Heath, Jr., T. Strohmer, Designing structured tight frames via an alternating projection method, IEEE Trans. Inform. Theory 51 (2005) 188–209. [54] R. J. Turyn, Character sums and difference sets, Pacific J. Math. 15 (1965) 319–346. [55] S. Waldron, On the construction of equiangular frames from graphs, Linear Algebra Appl. 431 (2009) 2228– 2242. 24 [56] L. R. Welch, Lower bounds on the maximum cross correlation of signals, IEEE Trans. Inform. Theory 20 (1974) 397-399. [57] R. M. Wilson, An existence theory for pairwise balanced designs I. Composition theorems and morphisms, J. Combin. Theory Ser. A 13 (1972) 220–245. [58] P. Xia, S. Zhou, G. B. Giannakis, Achieving the Welch bound with difference sets, IEEE Trans. Inform. Theory 51 (2005) 1900–1907. [59] G. Zauner, Quantum designs: Foundations of a noncommutative design theory, Ph.D. Thesis, University of Vienna, 1999. 25
1601.02122
1
1601
2016-01-09T15:29:08
Tensor products and joint spectra for solvable Lie algebras of operators
[ "math.FA" ]
Given two complex Hilbert spaces, $H_1$ and $H_2$, and two complex solvable finite dimensional Lie algebras of operators, $L_1$ and $L_2$, such that $L_i$ acts on $H_i$ (i= 1,2), the joint spectrum of the Lie algebra $L_1\times L_2$, which acts on $H_1\overline\otimes H_2$, is expressed by the cartesian product of $Sp(L_1,H_1)$ and $Sp(L_2,H_2)$.
math.FA
math
TENSOR PRODUCTS AND JOINT SPECTRA FOR SOLVABLE LIE ALGEBRAS OF OPERATORS ENRICO BOASSO Abstract. Given two complex Hilbert spaces, H1 and H2, and two complex solvable finite dimensional Lie algebras of operators, L1 and L2, such that Li acts on Hi (i= 1,2), the joint spectrum of the Lie algebra L1 ×L2, which acts on H1⊗H2, is expressed by the cartesian product of Sp(L1, H1) and Sp(L2, H2). 1. Introduction J. L. Taylor developed in [6] a notion of joint spectrum for an n-tuple a, a = (a1, ..., an), of mutually commuting operators acting on a Banach space E, i.e., ai ∈ L(E), the algebra of all bounded linear operators on E, and [ai, aj] = 0, 1 ≤ i, j ≤ n. This interesting notion, which extends in a natural way the spectrum of a single operator, has many important properties, among then, the projection property and the fact that Sp(a, E) is a compact non empty subset of Cn, where Sp(a, E) denotes the joint spectrum of a in E. One of the most remarkable results of the Taylor joint spectrum is the one related with tensor products of tuples of operators. For example, in [2], Z. Ceaus- escu and F. H. Vasilescu proved the following result. Let Hi, 1 ≤ i ≤ n, be com- plex Hilbert spaces, and ai, 1 ≤ i ≤ n, be bounded linear operators defined on Hi, respectively, 1 ≤ i ≤ n. If we denote by H the completion of the tensor product H1 ⊗ . . . ⊗ Hn with respect to the canonical scalar product, we may consider the n-tuple of operators a, a = (a1, . . . , an), where ai = 1 ⊗ . . . ⊗ 1 ⊗ ai ⊗ 1 ⊗ . . . ⊗ 1, 1 ≤ i ≤ n, and 1 denotes the identity of the corresponding spaces. Then the following identity holds, Sp(a, E) = Sp(a1) × . . . × Sp(an). Furthermore, in [3], Z. Ceausescu and F. H. Vasilescu showed that if H1 (resp. H2) is a complex Hilbert space and a = (a1, . . . , an), (resp. b = (b1, . . . , bm)), is a mutually commuting tuple of operators acting on H1, (resp. H2), then, the commuting tuple (a, b) = (a1 ⊗ 1, . . . , an ⊗ 1, 1 ⊗ b1, . . . , 1 ⊗ bm) in L(H1⊗H2), satisfies the relation, Sp((a, b), H1⊗H2) = Sp(a, H1) × Sp(b, H2), where 1 denotes the identity map of the corresponding Hilbert spaces, and H1⊗H2 is the completion of the tensor product H1 ⊗ H2 with respect to the canonical scalar product, see [3,Theorem 2.2]. In [1] we defined a joint spectum for complex solvable finite dimensional Lie algebras of operators L, acting on a Banach space E, and we denoted it by Sp(L, E). We proved that Sp(L, E) is a compact non empty subset of L∗ and that 1 2 ENRICO BOASSO the projection property for ideals still holds. Besides, when L is a commutative algebra, our spectrum reduces to Taylor joint spectrum in the following sense. If dim L = n, {ai}(1≤i≤n) is a basis of L and we consider the n-tuple a = (a1, . . . , an), then {(f (a1), . . . , f (an)) : f ∈ Sp(L, E)} = Sp(a, E), i.e., Sp(L, E) in terms of the basis of L∗ dual of {ai}(1≤i≤n) coincides with the Taylor joint spectrum of the n-tuple a. Then, the following question arises naturally. If Hi, i = 1, 2, are two complex Hilbert spaces, and Li, i = 1, 2, are two complex solvable finite dimensional Lie algebras of operators such that Li acts on Hi, respectively, i = 1, 2, is there any relation between Sp(L1×L2, H1⊗H2) and Sp(L1, H1)×S(L2, H2). In this paper we answer this question in the affirmative. Moreover, by a re- finement of the argument of Z. Ceausescu and F. H. Vasilescu in [3], we extend the main results of [2] and [3] for complex solvable finite dimensional Lie algebras and its joint spectrum. In order to describe in more detail our main theorem we need to introduce a definition. If Hi and Li are as above, i = 1, 2, we consider the direct product of L1 and L2, i.e., the complex solvable finite dimensional Lie algebra L1 × L2 defined by, L1 × L2 = {x1 ⊗ 1 + 1 ⊗ x2 : xi ∈ Li, i = 1, 2}, where 1 is as above. Then, its is clear that L1 × L2 is a Lie algebra of operators which acts on H1⊗H2, and our main theorem may be stated as follows, Sp(L1 × L2, H1⊗H2) = Sp(L1, H1) × Sp(L2, H2), where the above sets are considered as subsets of (L1 × L2)∗ under the natural identification (L1 × L2)∗ ∼= L∗ 1 × L∗ 2. The paper is organized as follows. In Section 2 we review several definitions and results of [1] and we also prove a proposition which is an important steps to our main result. Finally, in Section 3, we prove our main theorem. 2. Preliminaries We briefly recall several definitions and results related to the spectrum of a complex solvable Lie algebra of operators, see [1]. From now on L denotes a complex solvable finite dimensional Lie algebra and H a complex Hilbert space on which L acts as right continous operators, i.e., L is a Lie subalgebra of L(H) with the opposite product. If dim L = n and f is a character of L, i.e., f ∈ L∗ and f (L2) = 0, where L2 = {[x, y] : x, y ∈ L}, then consider the following chain complex, (H ⊗ ∧L, d(f )), where ∧L denotes the exterior algebra of L, and dp(f ) is as follows, dp(f ) : H ⊗ ∧pL → H ⊗ ∧p−1L, LIE ALGEBRAS OF OPERATORS 3 dp(f )ehx1 ∧ · · · ∧ xpi = k=p X k=1 (−1)k+1e(xk − f (xk))hx1 ∧ . . . ∧ xk ∧ · · · ∧ xpi + X (−1)k+leh[xk, xl] ∧ x1 ∧ . . . ∧ xk ∧ . . . ∧ xl ∧ . . . ∧ xpi, 1≤k<l≤p wheremeans deletion. If p ≤ 0 or p ≥ n + 1, we define dp(f ) = 0. If we denote by H∗((H⊗∧L, d(f ))) the homology of the complex (H⊗∧L, d(f )), we may state our first definition. Definition 2.1. With L and f as above, the set {f ∈ L∗ : f (L2) = 0, H∗((H ⊗ ∧L, d(f ))) 6= 0} is the joint spectrum of L acting on H, and it is denoted by Sp(L, H). As we have said, in [1] we proved that Sp(L, E) is a compact non empty subset of L∗ which reduces to Taylor joint spectrum when L is a commutative algebra, in the sense explained in the introduction. Besides, if I is an ideal of L and π denotes the projection map from L∗ to I ∗, then, Sp(I, H) = π(Sp(L, H)), i.e., the projection property for ideals still holds. With regard to this property, we ought to mention the paper of C. Ott, see [5], who pointed out a gap in the proof of this result, and give another proof of it. In any case, the projection property remains true. We observe that the set H ⊗ ∧L has a natural structure of Hilbert space, so that the sets H ⊗ hxi1 ∧ · · · ∧ xipi, 1 ≤ i1 < . . . < ip ≤ n, 0 ≤ p ≤ n, are orthogonal subspaces of H ⊗ ∧L, and if <, > denotes the inner product of H, < ahxi1 ∧ . . . ∧ xipi, bhxi1 ∧ . . . ∧ xipi >=< a, b >. We shall have occasion to use the direct product of two complex solvable finite dimensional Lie algebras and its action on the tensor product of two complex Hilbert spaces. We recall here the main facts which we need for our work. If Hi, i =1,2, are two complex Hilbert spaces, then H1⊗H2 denotes the completion of the tensor product H1 ⊗ H2 with respect to the canonical scalar product. Now, if Li, i=1,2, are two complex solvable finite dimensional Lie algebras of operators, such that Li acts on Hi, respectively, i=1,2, we consider the algebra L1 × L2, the direct product of L1 and L2, which acts in a natural way on H1⊗H2, and it is defined by, L1 × L2 = {x ⊗ 1 + 1 ⊗ y : x ∈ L1 + y ∈ L2}, where 1 denotes the identity of the corresponding space. It is clear that L1×L2, defined as above, is a complex solvable finite dimensional Lie subalgebra of L(H1⊗H2). Moreover, by the structure of the Lie bracket in L1 × L2, we have two distinguished ideals, L′ 2, which we define as follows, 1 and L′ ′ L 1 = {x ⊗ 1 : x ∈ L1}, ′ L 2 = {1 ⊗ y : y ∈ L2}. 4 ENRICO BOASSO In addition, if we consider the natural identification K : (L1 × L2)∗ ∼= L∗ 1 × L∗ 2, K(f ) = (f ◦ i1, f ◦ i2), where f ∈ (L1 × L2)∗ and ij : Lj → L1 × L2, j = 1, 2, are the canonical inclusions, as (L1 × L2)2 = L2 2, we have that the set of characters of L1 × L2 is the cartesian product of the sets of characters of L1 and L2. 1 × L2 The following proposition is an important step to our main theorem. Proposition 2.2. Let Hi, i = 1, 2, be two complex Hilbert spaces, and L a complex solvable finite dimensional Lie algebra of operators acting on H1. Let L′ 1, (resp. L′ 2), be the Lie algebra of operators {x⊗1 : x ∈ L}, (resp. {1⊗x : x ∈ L}), which acts on H1⊗H2, (resp. H2⊗H1), where 1 denotes the identity map of H2, (resp. H1). Then, (i) Sp(L (ii) Sp(L (iii) Sp(L 1, H1⊗H2) = Sp(L 1, H1⊗H2) ⊆ Sp(L, H1), 2, H2⊗H1) ⊆ Sp(L, H1). 2, H2⊗H1), ′ ′ ′ ′ Proof. It is clear that (iii) is a consequence of (i) and (ii). Let us prove (i). If f is a character of L, we consider C1 (resp. C2), the Koszul complex associ- ated to the Lie algebra L 2) and f , then, ′ ′ 1 (resp. L 1, d1(f )), ′ C1 = (H1⊗H2 ⊗ ∧L C2 = (H2⊗H1 ⊗ ∧L ′ 2, d2(f )), where the maps di(f ), i= 1,2, are as above. In addition, a elementary calculation shows that the map µ, µp : H1⊗H2 ⊗ ∧pL ′ 1 → H2⊗H1 ⊗ ∧pL ′ 2, µp(e1 ⊗ e2hx1, . . . , xpi) = e2 ⊗ e1hx1, . . . , xpi, defines an isomorphism which commutes with di(f ), i=1,2, i.e., µ is an isomor- phism of chain complexes. Then, ′ ′ Sp(L 1, H1⊗H2) = Sp(L 2, H2⊗H1). In order to verify (ii), let us consider the Koszul complex associated to L and f , C = (H1 ⊗ ∧pL, d(f )), and C the following chain complex, C = ((H1 ⊗ ∧L)⊗H2, d(f ) ⊗ 1), where 1 denotes the identity map of H2. We observe that if η is the map defined by, η : (H1⊗H2) ⊗ ∧pL ′ 1 → (H1 ⊗ ∧pL)⊗H2, ηp(e1 ⊗ e2hx1, . . . , xpi) = e1hx1, . . . , xpi ⊗ e2, then an easy calculation shows that η defines an isomorphism of chain complexes between C1 and C. Now, if f does not belong to Sp(L, H1), by [8, Lemma 2.4], the complex C is exact. As η is an isomorphism of chain complexes, if C1 is exact, then f does not belong to Sp(L ′ 1, H1⊗H2). (cid:3) LIE ALGEBRAS OF OPERATORS 5 We now state our main result. 3. The Main Result Theorem 3.1. Let H1 and H2 be two complex Hilbert spaces and L1 and L2 two complex solvable finite dimensional Lie algebras of operators such that Li acts on Hi, respectively, i= 1,2. Let us consider the complex solvable finite dimensional Lie algebra L1 × L2, which acts on H1⊗H2 and it is defined by, L1 × L2 = {x1 ⊗ 1 + 1 ⊗ x2 : xi ∈ Li, i = 1, 2}, where 1 denotes the identity of the corresponding spaces. Then, Sp(L1 × L2, H1⊗H2) = Sp(L1, H1) × Sp(L2, H2), 1 × L∗ 2 under the natural identification K : (L1 × L2)∗ ∼= L∗ where, in the above equality, the set Sp(L1 × L2, H1⊗H2) is considered as a subset of L∗ 2 of Section 2. Proof. In order to prove that Sp(L1 × L2, H1⊗H2) is contained in the cartesian product of Sp(L1, H1) and Sp(L2, H2), let L 2) be the ideal of L1 × L2 defined by {x ⊗ 1 : x ∈ L1} (resp. {1 ⊗ y : y ∈ L2}), where 1 is as above, see Section 2. Then, by the projection property of the joint spectrum, Proposition 1 and the identification K, 1 (resp. L 1 × L∗ ′ ′ Sp(L1 × L2, H1⊗H2) ⊆ Sp(L 1, H1⊗H2) × Sp(L 2, H1⊗H2) ′ ′ ⊆ Sp(L1, H1) × Sp(L2, H2). Let us prove the converse inclusion. We consider the following preliminary facts first. Let fi be a character of Li and Ki the Koszul complex of Li acting on Hi associated to fi, for i = 1, 2, K1 = (H1 ⊗ ∧L1, d1(f1)), K2 = (H2 ⊗ ∧L2, d2(f2)). We observe that there is a natural identification between the spaces (H1⊗H2)⊗ ∧(L1 × L2) and (H1 ⊗ ∧L1) ⊗ (H2 ⊗ ∧L2). If we denote by ψ this identification, ψ is the following map, ψ : (H1 ⊗ H2) ⊗ ∧(L1 × L2) → (H1 ⊗ ∧L1) ⊗ (H2 ⊗ ∧L2), ψ(e1 ⊗ e2hx1 ∧ · · · ∧ xp ∧ y1 ∧ . . . ∧ yqi) = e1hx1 ∧ · · · ∧ xpi ⊗ e2hy1 ∧ · · · ∧ yqi, where e1 ∈ H1, e2 ∈ H2, p ∈ [[1, n]], q ∈ [[1, m]], n = dim(L1), and m = dim(L2). As (H1⊗H2) ⊗ ∧k(L1 × L2) = ⊕p+q=k(H1 ⊗ ∧pL1)⊗(H2 ⊗ ∧qL2), if we consider Hi ⊗∧Li, i=1,2, (H1⊗H2)⊗∧(L1 ×L2) and (H1 ⊗∧L1)⊗(H2 ⊗∧L2) with their natural structure of Hilbert spaces, a straightforward calculation shows that the map ψ may be extended to an isometric isomorphism from (H1⊗H2) ⊗ ∧(L1 × L2) onto (H1 ⊗ ∧L1)⊗(H2 ⊗ ∧L2). Besides, if f is a character of L1 × L2 and if we define fj = f ◦ ij ∈ L∗ j , where ij : Lj → L1 × L2 are the canonical inclusions and j = 1, 2, i.e., if we consider f 6 ENRICO BOASSO decomposed under the natural identification K : (L1 × L2)∗ ∼= L∗ the Koszul complex of L1 × L2 acting on H1⊗H2 associated to f , 1 × L∗ 2, if K is K = ((H1⊗H2) ⊗ ∧(L1 × L2), dk(f )), then an easy calculation shows that, ψdk(f ) = (d1(f1) ⊗ 1 + ξ ⊗ d2(f2))ψ, where ξ is the map, ξ : ⊕n p=0 (H1 ⊗ ∧pL1) → ⊕n p=0(−1)p, ξ = ⊕n p=0(H1 ⊗ ∧pL1), 1 is the identity map of H1 and dk(f ) is the boundary map of the Koszul complex K, equivalently, if we consider the algebraic tensor product of the complexes K1 and K2, K1 ⊗ K2, and its natural completion K1⊗K2, the map ψ provides an isometric isomorphism of chain complexes, from K onto K1⊗K2. Moreover, if Ti, i = 1, 2, and Tk are the maps, Ti = di(fi) + di(fi) ∗ , Tk = dk(f ) + dk(f ) ∗ , as ξ is a selfadjoint map, an easy calculation shows that, ψTk = (T1 ⊗ 1 + ξ ⊗ T2)ψ. Let us return to the proof. If f does not belong to Sp(L1 × L2, H1⊗H2), by [7, Lemma 3.1], the operator Tk is an invertible map. On the other hand, if fi, i = 1, 2, belongs to Sp(Li, Hi), i = 1, 2, as Ti is a selfadjoint operator, i= 1,2, there exist by [7, Lemma 3.1] and by [4, Chapter II, Section 31, Theorem 2] two sequences of unit vectors, (ai n ∈ Hi ⊗ ∧Li, n) → 0 (n → ∞), for i = 1, 2. However, as ψ and ψ−1 are isometric and Ti(ai isomorphisms, by elementary properties of the tensor product in Hilbert spaces, we have that: k a1 n)) → 0 (n → ∞), equivalently, by [4, Chapter II, Section 31, Theorem 2], 0 ∈ Sp(Tk), which is impossible by our assumption. n)n∈N, i = 1, 2, such that k ai n) k= 1 and Tk(ψ−1(a1 n k= 1, ai n ⊗ a2 n ⊗ a2 n k= 1, k ψ−1(a1 n ⊗ a2 (cid:3) References 1. E. Boasso and A. Larotonda, A spectral theory for solvable Lie algebras of operators, Pacific J. Math. 158 (1993), 15-22. 2. Z. Ceausescu and F. H. Vasilescu, Tensor products and Taylor's joint spectrum, Studia Math. 62 (1978), 305-311. 3. Z. Ceausescu and F. H. Vasilescu, Tensor products and the joint spectrum in Hilbert spaces, Proc. Amer. Math. Soc. 72 (1978), 505-508. 4. P. Halmos, Introduction to Hilbert spaces and the theory of spectral multiplicity, Chelsea Publishing Company, 1972. 5. C. Ott, A note on a paper of E. Boasso and A. Larotonda, Pacific J. Math., 173 (1996), 173-179. 6. J. L. Taylor, A joint spectrum for several commuting operators, J. Funct. Anal. 6 (1970), 172-191. 7. F. H. Vasilescu, Analytical perturbations of the δ-operator and integral representation formulas in Hilbert spaces, J. Operator Theory 1 (1979), 187-205. LIE ALGEBRAS OF OPERATORS 7 8. C. Grosu and F. H. Vasilescu, The Kunneth formula for Hilbert complexes, Integral Equa- tions Operator Theory 5 (1982), 1-17. Enrico Boasso E-mail address: enrico [email protected]
1808.01189
1
1808
2018-08-03T13:46:27
On the exponential ultradistribution semigroups in Banach spaces
[ "math.FA" ]
In this paper, we continue our previous research studies of exponential ultradistribution semigroups in Banach spaces. The existence and uniqueness of analytical solutions of abstract fractional relaxation equations associated with the generators of exponential ultradistribution semigroups have been considered. Some other results are also proved.
math.FA
math
ON THE EXPONENTIAL ULTRADISTRIBUTION SEMIGROUPS IN BANACH SPACES MARKO KOSTI ´C, STEVAN PILIPOVI ´C, AND DANIEL VELINOV Abstract. In this paper, we continue our previous research studies of ex- ponential ultradistribution semigroups in Banach spaces. The existence and uniqueness of analytical solutions of abstract fractional relaxation equations as- sociated with the generators of exponential ultradistribution semigroups have been considered. Some other results are also proved. 1. Introduction and preliminaries The class of (exponential) distribution semigroups in Banach spaces was intro- duced by J. L. Lions in [24]. Along with the papers [3] by W. Arendt on integrated semigroups and vector-valued Widder representation theorem, and [6] by G. Lumer and I. Cioranescu on K-convoluted semigroups, the paper [24] has been a motivation for a great number of other authors to study various classes of (ultra-)distribution semigroups in the setting of Banach spaces or, more generally, locally convex spaces (cf. [1], [2], [10]-[12], [17], [22], [25]-[26], [29]-[30] and references cited therein for further information). Following the approaches employed in [23]-[24], the authors of [20] have defined the classses of L-ultradistribution semigroups and ultradistri- bution semigroups with non-densely defined generators. In [21], some structural characterizations of ultradistribution semigroups and exponential ultradistribution semigroups have been proved. Motivated by the ideas from D. Fujiwara's research of exponential distribution semigroups [9], we shall prove some new characteriza- tions of spaces of vector-valued tempered ultradistributions in Proposition 2.1. In Theorem 2.3, we consider the existence and uniqueness of analytical solutions of abstract fractional relaxation equations associated with the generators of exponen- tial ultradistribution semigroups. Some new qualitative properties of solutions of the first order abstract Cauchy problem (ACP ) clarified below have been proved in the case that the operator A generates an exponential ultradistribution semigroup of (Mp)-class and that the initial value x belongs to the abstract Beurling space associated with the operator A. We use the standard terminology throughout the paper. We impose the following conditions (cf. [14]) on the sequence (Mp)p∈N0 of positive real numbers satisfying M0 = 1: (M.1) M 2 (M.2) For some A, H > 0, Mp ≤ AH p min0≤q≤p Mp−qMq, (M.3)′ : p ≤ Mp−1Mp+1 for p ∈ N; p, q ∈ N; Every employment of the condition Xp=1 ∞ Mp−1 Mp < ∞. 1 2 MARKO KOSTI ´C, STEVAN PILIPOVI ´C, AND DANIEL VELINOV (M.3) : ∞ sup p∈N Xq=p+1 Mq−1Mp+1 pMpMq < ∞, K :=(cid:8)ϕ ∈ C∞(Rn) : supp(ϕ) ⊆ K and there exists C > 0 such that kDαϕkC(K) ≤ Ch−αMα, α = 0, 1, 2,···(cid:9). becomes a Ba- := supx∈K,α∈N0 K Mp which is a slightly stronger than (M.3)′, will be explicitly emphasized. The Gevrey sequence Mp = p!s, s > 1 satisfies all the above conditions. The associated function M (ρ) of sequence (Mp) on (0, +∞) is defined by M (ρ) = supp∈N ln ρp , ρ > 0. We know that he function t 7→ M (t), t ≥ 0 is increasing as well as that limλ→∞ M (λ) = ∞ and that the function M (·) vanishes in some open neighborhood of zero. Put mp := Mp/Mp−1, p ∈ N. Then (M.1) implies that the sequence (mp)p∈N is increas- ing. The spaces of ultradifferentiable functions of ∗-class and ultradistributions of ∗-class are defined in the sequel. Let ∅ 6= Ω ⊆ R, and let ∅ 6= K ⋐ Ω denote a compact set in Ω. Set, for every h > 0, D{Mp},h Equipped with the norm k·kD nach space. Define also, as locally convex spaces, D(Mp) D(Mp)(Ω) := lim indK⋐ΩD(Mp) lim indK⋐ΩD{Mp} E {Mp}(Ω) will be understood in the sense of [13]. type, are defined in [27] as duals of the following test spaces , D{Mp},h := lim projh→+∞D{Mp},h , and D{Mp}(Ω) := . The notions of spaces E (Mp)(K), E (Mp)(Ω), E {Mp}(K) and The spaces of tempered ultradistributions of the Beurling, resp., the Roumieu := lim indh→0D{Mp},h K , D{Mp} Dα·(x) h−αMα {Mp},h K K K K K K S(Mp)(R) := lim projh→∞SMp,h(R), resp., S{Mp}(R) := lim indh→0SMp,h(R), where for each h > 0, and SMp,h(R) :=(cid:8)φ ∈ C∞(R) : kφkMp,h < ∞(cid:9) kφkMp,h := sup( hα+β MαMβ(cid:0)1 + t2(cid:1)β/2(cid:12)(cid:12)φ(α)(t)(cid:12)(cid:12) : t ∈ R, α, β ∈ N0). The common notation for symbols (Mp) and {Mp} will be ∗. It is said that a function P (ξ) = Pα∈N0 the class (Mp), resp. {Mp}, if the coefficients aα satisfy the estimate (1.1) aαξα, ξ ∈ R is an ultrapolynomial of aα ≤ for some L > 0 and C > 0, resp. , α ∈ N0, CLα Mα for every L > 0 and some CL > 0. The aαDα is said to be an ultradifferential corresponding operator P (D) = Pα∈N0 operator of the class (Mp), resp. {Mp} (see [13]). Let X and Y be two Hausdorff sequentially complete locally convex spaces over the field of complex numbers. By L(X, Y ) we denote the space consisting of all continuous linear mappings from X into Y ; L(X) ≡ L(X, X). A C0 semigroup (T (t))t≥0 on X (cf. [16] and [28] for the notion) is called locally equicontinuous if the family {T (t) : t ∈ [0, t0]} is equicontinuous for all t0 > 0; (T (t))t≥0 is called (globally) equicontinuous if the family {T (t) : t ≥ 0} is equicontinuous. Recall that any C0-semigroup in a barreled space X is automatically locally equicontinuous ON THE EXPONENTIAL ULTRADISTRIBUTION SEMIGROUPS IN BANACH SPACES 3 ([16]). If for some α ∈ R the C0-semigroup (Tα(t) ≡ e−αtT (t))t≥0 is equicontinuous, then (T (t))t≥0 is called α quasi -- equicontinuous. Henceforth we shall always assume that (E,k·k) is a complex Banach space. Let A be a closed linear operator with domain D(A) and range R(A) contained in E. Since no confusion seems likely, we will identify A with its graph; [D(A)] denotes the Banach space D(A) equipped with the graph norm kxk[D(A)] := kxk + kAxk, x ∈ D(A). As is usually the case, ρ(A) denotes the resolvent set of A and R(λ : A) denotes the operator (λI−A)−1 (λ ∈ ρ(A)), where I stands for the identity operator on E. If F is a linear subspace of E, then we define the part of A in F, AF for short, by AF := {(x, y) ∈ A : x, y ∈ F}. Set D∞(A) := T∞ n=1 D(An) and Σα := {z ∈ C \ {0} : arg(z) < α} (α ∈ (0, π]). We denote by Aα(E) the vector space consisting of all analytic mappings from Σα into E (α ∈ (0, π]). The following system of seminorms k·kn ≡ k·k +··· +kAn·k, n ∈ N0 turns D∞(A) into a Fr´echet space. We define the abstract Beurling space of (Mp) class associated to a closed linear operator A as in [5]. Put E(Mp)(A) := lim projh→+∞E{Mp} (A), where h E{Mp} h (A) :=(x ∈ D∞(A) : kxk{Mp} (A), k · k{Mp} h h ≡ sup p∈N0 hpkApxk Mp < ∞). h h h′ (A) ⊆ E{Mp} ) is a Banach space, E{Mp} Then (E{Mp} (A) if 0 < h < h′ < ∞, E(Mp)(A) is a Fr´echet space, and E(Mp)(A) is a dense subspace of E whenever A is the generator of a regular (Mp)-ultradistribution semigroup ([5]). By D′∗ + (R, L(E)) we denote the space consisting of those L(E)-valued ultradis- tributions G of ∗-class for which supp(G) ⊆ [0,∞) (cf. [17] for the notion) and by D∗ 0 we denote the space consisting of those ultradifferentiable functions ϕ of ∗-class such that ϕ(t) = 0 for t ≤ 0. We refer the reader to [13]-[14] for further informa- tion concerning the convolution of (vector-valued) ultradifferentiable functions and ultradistributions of ∗-class. Definition 1.1. ([20]) Let G ∈ D′∗ ultradistribution semigroup of ∗-class if: φ, ψ ∈ D∗ (U.1) G(φ ∗ ψ) = G(φ)G(ψ), 0; (U.2) N (G) :=Tφ∈D∗ N (G(φ)) = {0}; (U.3) R(G) :=Sφ∈D∗ R(G(φ)) is dense in E; (U.4) For every x ∈ R(G) there exists a function u ∈ C([0,∞), E) satisfying + (R, L(E)). Then it is said that G is an L- 0 0 u(0) = x and G(φ)x = ∞ Z0 φ(t)u(t) dt, φ ∈ D∗ 0. If G ∈ D′∗ + (R, L(E)) satisfies (U.5) G(φ ∗0 ψ) = G(φ)G(ψ), 0 , where t φ, ψ ∈ D∗ Z0 (f ∗0 g)(t) := f (t − s)g(s) ds, t ∈ R, then it is said that G is a pre-(UDSG) of ∗-class. If (U.5) and (U.2) are fulfilled for G, then G is called an ultradistribution semigroup of ∗-class, in short, (UDSG). A pre-(UDSG) G is said to be dense if it additionally satisfies (U.3). 4 MARKO KOSTI ´C, STEVAN PILIPOVI ´C, AND DANIEL VELINOV Definition 1.2. An L-ultradistribution semigroup G of ∗-class is said to be ex- ponential, EL-ultradistribution semigroup of ∗-class in short, if in addition to (U.1) − (U.4), G fulfills: (U.6) (∃a ≥ 0)(e−a·G ∈ S′∗(R, L(E))). Then we say that G is of order a. Conditions (U.6), (U.5), resp., (U.6), (U.5) and (U.2), define an exponential pre-(UDSG), resp., exponential (UDSG), and they are denoted by pre-(EUDSG), resp., (EUDSG). The definition of (infinitesimal) generator A of an ultradistribution semigroup G of ∗-class is defined by A := {(x, y) ∈ E × E : G(−ϕ′)x = G(ϕ)y for all ϕ ∈ D∗ 0}. Since every L-ultradistribution semigroup G of ∗-class is also an ultradistribution semigroup of ∗-class ([20]), the definition of (infinitesimal) generator of such a semigroup is clear. We refer the reader to [20] for the notion of an (exponential) ultradistribution fundamental solution for a closed linear operator A acting on E. For the sake of convenience, we shall remind the reader of the following important result. Theorem 1.3. ([19]) Suppose A is a closed linear operator on E. Then there exists an exponential ultradistribution fundamental solution of ∗-class for A iff there exist a ≥ 0, k > 0 and L > 0, in the Beurling case, resp., there exists a ≥ 0 such that, for every k > 0 there exists Lk > 0, in the Roumieu case, such that: {λ ∈ C : ℜλ > a} ⊆ ρ(A), and kR(λ : A)k ≤ LeM(kλ), λ ∈ C, ℜλ > a, resp., kR(λ : A)k ≤ LkeM(kλ) for all k > 0 and λ ∈ C,ℜλ > a. Before proceeding further, we would like to write down the following facts about (exponential) ultradistribution semigroups of ∗-class. If the operator A generates an (exponential) ultradistribution semigroup G of ∗-class, then there exists an injective operator C ∈ L(E) such that the operator A generates a (an exponentially bounded) [17, Section 3.6]). This implies that C-regularized semigroup (T (t))t≥0 on E (cf. there exist two Fr´echet spaces Y and W such that Y ֒→ X ֒→ W (֒→ means the continuous embedding), as well as that AY generates a C0-semigroup on Y and that A = BX, where B generates a C0-semigroup on W ; without going into full details, it should be only noticed that in (Mp)-case Y may be chosen to be exactly E(Mp)(A), the abstract Beurling space of the operator A (then AE(Mp )(A) ∈ L(E(Mp)(A)); cf. [17, Theorem 3.6.6] for more details). Furthermore, if the ultradistribution semigroup G generated by A is exponential, then Y and W may be chosen to be Banach spaces ([8]). derivative Dα Let α > 0, β ∈ R, γ ∈ (0, 1), and let m = ⌈α⌉. Then the Caputo fractional t u(t) is defined for those functions u ∈ Cm−1([0,∞) : E) for which gm−α ∗ (u −Pm−1 k=0 uk m−1 ·k k! ) ∈ Cm([0,∞) : E), by dtm"gm−α ∗ u − t u(t) = dm Dα k!!#. uk ·k Xk=0 The Mittag-Leffler function Eα,β(z) is defined by Eα,β(z) := P∞ n=0(zn/Γ(αn + β)), z ∈ C. Here we assume that 1/Γ(αn + β) = 0 if αn + β ∈ −N0. Set, for short, Eα(z) := Eα,1(z), z ∈ C. The Wright function Φγ(·) is defined by Φγ(t) := ON THE EXPONENTIAL ULTRADISTRIBUTION SEMIGROUPS IN BANACH SPACES 5 L−1(Eγ(−λ))(t), t ≥ 0, where L−1 denotes the inverse Laplace transform. We refer the reader to [4] and [18, Section 1.3] for more details about Mittag-Leffler and Wright functions. 2. Characterization of exponential ultradistribution semigroups The following proposition is motivated by D. Fujiwara's research of exponential distribution semigroups (cf. [9, Lemma 1]). Proposition 2.1. (i) Let a0 ≥ 0, and let e−a0·G ∈ L(S(Mp)(R), L(E)). Then there exists h′ > 0 such that, for every a > a0 and for every non-empty compact set K ⊆ [0,∞), there exists C > 0 such that for any ϕ, ψ ∈ D(Mp) with supp(ϕ)∪ supp(ψ) ⊆ K, the following estimate holds: 0 (2.1) (cid:13)(cid:13)e−atG(ϕ ∗ ψ)(cid:13)(cid:13)L(E) ≤ CkϕkD {Mp},h′ K kψkL1. K (ii) Let a0 ≥ 0, and let e−a0·G ∈ L(S{Mp}(R), L(E)). Then, for every h′ > 0, for every a > a0, and for every non-empty compact set K ⊆ [0,∞), there exists C > 0 such that for any ϕ, ψ ∈ D{Mp},h′ , the estimate (2.1) holds. Proof. Let ζ ∈ D{Mp}, and let ζ(t) = 1 for all t ∈ [−1, 1]. Put γ1 := Hζ (H is the 1 = δ + Hζ′ in the sense of Roumieu ultradistributions. Heaviside function). Then γ′ So, with ω1 = Hζ′ ∈ D{Mp} , we have δ′ ∗ γ1 = δ + ω1. (2.2) Let the numbers h′ > 0 and a > a0 be fixed, let K be a non-empty compact subset of [0,∞), and let ϕ, ψ ∈ D{Mp},h′ . We will prove that the function e(a0−a)·(ϕ∗ ψ)(·) ∈ SMp,h(R) for all h ∈ (0, h′H/4(1 + a − a0)); here H denotes the constant from the condition (M.2). Set K ′ := (K+supp(γ1)) ∪ (K+supp(ω1)) and K ′′ := K + K ′. Using the product rule, Young's inequality and the condition (M.2), we have that K 0 sup sup = sup ≤ + MαMβ MαMβ : t ∈ R, α, β ∈ N0) (cid:13)(cid:13)e(a0−a)·(ϕ ∗ ψ)(·)(cid:13)(cid:13)Mp,h = sup( hα+β t∈R, α, β∈N0( hα+β(cid:0)1 + t2(cid:1)β/2 t∈R, α, β∈N0( hα+β(cid:0)1 + t2(cid:1)β/2 t∈R, α, β∈N0( hα+β(cid:0)1 + t2(cid:1)β/2 t∈K ′′, α, β∈N0( hα+β(cid:0)1 + t2(cid:1)β/2 t∈K ′′, α, β∈N0( hα+β(cid:0)1 + t2(cid:1)β/2 ≤ eM(cid:0)h√1+(sup K ′′)2(cid:1) t∈K ′′, α∈N0( hα MαMβ(cid:0)1 + t2(cid:1)β/2(cid:12)(cid:12)(cid:12)(cid:10)δ(α) ∗ e(a0−a)·(ϕ ∗ ψ)(·)(cid:11)(t)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:10)δ(α+1) ∗ γ1 ∗ (e(a0−a)·ϕ ∗ e(a0−a)·ψ)(cid:11)(t) −(cid:10)δ(α) ∗ ω1 ∗ (e(a0−a)·ϕ ∗ e(a0−a)·ψ)(cid:11)(t)(cid:12)(cid:12)(cid:12) ZR(cid:12)(cid:12)(cid:12)(cid:16)e(a0−a)·ϕ(·)(cid:17)(α+1) ZR(cid:12)(cid:12)(cid:12)(cid:16)e(a0−a)·ϕ(·)(cid:17)(α) ZK ′(cid:12)(cid:12)(cid:12)(cid:16)e(a0−a)·ϕ(·)(cid:17)(α+1) ZK ′(cid:12)(cid:12)(cid:12)(cid:16)e(a0−a)·ϕ(·)(cid:17)(α) Mα ZK ′(cid:12)(cid:12)(cid:12)(cid:16)e(a0−a)·ϕ(·)(cid:17)(α+1) ds) (t − s)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:0)e(a0−a)·ψ ∗ γ1(cid:17)(s)(cid:12)(cid:12)(cid:12) ds) (t − s)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:0)e(a0−a)·ψ ∗ ω1(cid:17)(s)(cid:12)(cid:12)(cid:12) ds) (t − s)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:0)e(a0−a)·ψ ∗ γ1(cid:17)(s)(cid:12)(cid:12)(cid:12) ds) (t − s)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:0)e(a0−a)·ψ ∗ ω1(cid:17)(s)(cid:12)(cid:12)(cid:12) (t − s)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:0)e(a0−a)·ψ ∗ γ1(cid:17)(s)(cid:12)(cid:12)(cid:12) ds) MαMβ MαMβ MαMβ + sup = sup sup 6 MARKO KOSTI ´C, STEVAN PILIPOVI ´C, AND DANIEL VELINOV ds) α+1 (t − s)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:0)e(a0−a)·ψ ∗ ω1(cid:17)(s)(cid:12)(cid:12)(cid:12) Xj=0(cid:13)(cid:13)ϕ(j)kL∞(K) (cid:13)(cid:13)e(a0−a)·ψ ∗ γ1(cid:13)(cid:13)L1(R)) Xj=0(cid:13)(cid:13)ϕ(j)kL∞(K) (cid:13)(cid:13)e(a0−a)·ψ ∗ ω1(cid:13)(cid:13)L1(R)) α sup Mα Mα 4 A ≤ Mα ZK ′(cid:12)(cid:12)(cid:12)(cid:16)e(a0−a)·ϕ(·)(cid:17)(α) α∈N0( (2(1 + a − a0)h)α α∈N0( (2(1 + a − a0)h)α + eM(cid:0)h√1+(sup K ′′)2(cid:1) t∈K ′′, α∈N0( hα ≤ 2(1 + a − a0)eM(cid:0)h√1+(sup K ′′)2(cid:1) sup + 2(1 + a − a0)eM(cid:0)h√1+(sup K ′′)2(cid:1) sup (1 + a − a0)eM(cid:0)h√1+(sup K ′′)2(cid:1)(cid:13)(cid:13)ψ(cid:13)(cid:13)L1(R)h(cid:13)(cid:13)γ1(cid:13)(cid:13)L1(R) +(cid:13)(cid:13)ω1(cid:13)(cid:13)L1(R)i ) (cid:17)j (2(1 + a − a0)h)α−j Mj (1 + a − a0)eM(cid:0)h√1+(sup K ′′)2(cid:1)(cid:13)(cid:13)ψ(cid:13)(cid:13)L1(R)h(cid:13)(cid:13)γ1(cid:13)(cid:13)L1(R) +(cid:13)(cid:13)ω1(cid:13)(cid:13)L1(R)i (1 + a − a0)eM(cid:0)h√1+(sup K ′′)2(cid:1)(cid:13)(cid:13)ψ(cid:13)(cid:13)L1(R)h(cid:13)(cid:13)γ1(cid:13)(cid:13)L1(R) +(cid:13)(cid:13)ω1(cid:13)(cid:13)L1(R)ikϕkD Xj=0(cid:16) 2(1 + a − a0)h Xj=0(cid:16) 2(1 + a − a0)h eM(2(1+a−a0)h) α∈N0(α+1 sup H jMjMα−j × kϕkD × kϕkD {Mp},h′ K 4 A ≤ {Mp},h′ K (cid:17)j h′ ∞ h′H 4 A ≤ The final conclusion follows from the decomposition he−a·G, ϕ∗ψi = he−a0·G, e(a0−a)·(ϕ∗ ψ)i and the continuity of mapping e−a0·G : SMp,h → L(E), with the number h ∈ (0, h′H/4(1 + a − a0)) chosen arbitrarily. Remark 2.2. (i) Notice that we have not used the structural theorem for the space of tempered vector-valued ultradistributions in the proof (cf. [21, Theorem 2.2]) and that we do not need the condition (M.3) here. Observe also that the proof of Proposition 2.1 works in the case that E is a general sequentially complete locally convex space ([22]). {Mp},h′ K (cid:3) eM(h′H/2). (ii) As mentioned above, Proposition 2.1 is inspired by Lemma 1 of [9]. Re- grettably, it seems that the assertions of [9, Theorem 2-Theorem 3] can- not be so simply reconsidered for exponential ultradistribution semigroups. Speaking-matter-of-factly, in ultradistribution case we cannot prove the ex- istence of number N such that an analog of the estimate [9, (9)] holds. In the remaining part of this section, we shall always assume that the condition (M.3) holds. The following entire function of exponential type zero ([5], [13]) plays an important role in our further work: ∞ ω(Mp)(z) =: Yi=1(cid:0)1 + iz mp(cid:1), z ∈ C. Then ω(Mp)(z) ≥ eM(z), z ∈ C and the properties (P.1)-(P.5) stated in [17, Subsection 3.6.2] hold. Suppose now that A generates an exponential ultradistribution semigroup of (Mp)-class. By Theorem 1.3, it readily follows that there exist numbers a ≥ 0, k > 0 and L > 0 such that {λ ∈ C : ℜλ > a} ⊆ ρ(A) and kR(λ : A)k ≤ LeM(kλ), λ ∈ C, ℜλ > a. Let ¯a > a. Then there exists a sufficiently large number n0 ∈ N such ON THE EXPONENTIAL ULTRADISTRIBUTION SEMIGROUPS IN BANACH SPACES 7 that, for every n ∈ N with n ≥ n0, the operator A generates a global exponentially bounded Cn-regularized semigroup (Sn(t))t≥0 on E, where Sn(t)x := 1 2πi ¯a+i∞ Z¯a−i∞ eλt R(λ : A)x ωn(iλ) dλ, t ≥ 0, n ∈ N, x ∈ E and Cn := Sn(0). cf. the proof of [17, Theorem 3.6.4] (in this result, we have clarified the result on generation of local C-regularized semigroups; observe, however, that in the case of exponential ultradistribution semigroups the resulting Cn-regularized semigroup is global and exponentially bounded because the integration is taken along the line connecting the points ¯a− i∞ and ¯a + i∞, not along the upwards oriented boundary of a suitable ultra-logarithmic region of (Mp)-class). An elementary application of Cauchy's formula, taken together with the consideration in the paragraph contain- ing the equation [17, (316)] and [17, Lemma 3.6.5, Theorem 3.6.6], implies that E(Mp)(A) ⊆ Cn(D∞(A)), n ≥ n0 and that for each x ∈ E(Mp)(A) a unique solution of the abstract Cauchy problem u ∈ C∞([0,∞) : E) ∩ C([0,∞) : [D(A)]), u′(t) = Au(t), t ≥ 0, u(0) = x, (ACP ) :  is given by u(t) = Sn(t)C−1 n x, t ≥ 0, n ≥ n0. Using the same argumentation as in the proof of [17, Theorem 3.6.4], we get that for each x ∈ E(Mp)(A) and h > 0 there exist n ≥ n0 and c > 0 such that (2.3) sup p∈N0 hp(cid:13)(cid:13) dp dtp u(t)(cid:13)(cid:13)Mp ≤ ce¯at(cid:13)(cid:13)C−1 n x(cid:13)(cid:13), t ≥ 0. Applying [18, Theorem 2.4.2] (cf. also [4, Theorem 3.1, Theorem 3.3]) and the Ljubich uniqueness theorem for abstract time-fractional equations [18, Theorem 2.1.34], it readily follows that for each x ∈ E(Mp)(A) and α ∈ (0, 1) there exists a unique solution of the following abstract fractional Cauchy problem v ∈ C([0,∞) : E) ∩ Amin(( 1 Dα v(0) = x, t v(t) = Av(t), t ≥ 0, α −1) π 2 ,π)(E), (ACP )α :  the following theorem. 0 Φα(s)Sn(stα)C−1 n x ds, t ≥ 0 (n ≥ n0). Thus, we have proved given by v(t) := R ∞ Theorem 2.3. Suppose that (M.3) holds for (Mp) and A generates an exponential ultradistribution semigroup of (Mp)-class. Then for each x ∈ E(Mp)(A) there exists a unique solution u(t) of the abstract Cauchy problem (ACP ) satisfying additionally that for each h > 0 there exist n ≥ n0 and c > 0 such that (2.3) holds. Furthermore, for each x ∈ E(Mp)(A) and α ∈ (0, 1) there exists a unique solution of the abstract fractional Cauchy problem (ACP )α. It is worth noting that the assertion of Theorem 2.3 continues to hold, with suitable modifications, in the setting of Fr´echet spaces and that possible applications can be made to differential operators considered in [17, Example 3.5.15] and [22, Example 9.3]. 8 MARKO KOSTI ´C, STEVAN PILIPOVI ´C, AND DANIEL VELINOV References [1] W. Arendt, C. J. K. Batty, M. Hieber and F. Neubrander, Vector-valued Laplace Transforms and Cauchy Problems. Birkhauser Verlag, Basel, 2001. [2] W. Arendt, O. El-Mennaoui and V. Keyantuo, Local integrated semigroups: evolution with jumps of regularity. J. Math. Anal. Appl. 186 (1994), 572 -- 595. [3] W. Arendt, Vector-valued Laplace transforms and Cauchy problems. Israel J. Math. 59 (1987), 327 -- 352. [4] E. Bazhlekova, Fractional Evolution Equations in Banach Spaces. PhD Thesis, Eindhoven University of Technology, Eindhoven, 2001. [5] I. Cioranescu, Beurling spaces of class (M Math. 102 (1978), 167 -- 192. p) and ultradistribution semi-groups. Bull. Sci. [6] I. Cioranescu and G. Lumer, Probl`emes d'´evolution r´egularis´es par un noyan g´en´eral K(t), Formule de Duhamel, prolongements, th´eor`emes de g´en´eration. C. R. Acad. Sci. Paris S´er. I Math. 319 (1995), 1273 -- 1278. [7] G. Da Prato and U. Mosco, Semigruppi distribuzioni analitici. Ann. Scuola Norm. Sup. Pisa 19 (1965), 367-396. [8] R. deLaubenfels, C-semigroups and strongly continuous semigroups. Israel J. Math. 81 (1993), 227 -- 255. [9] D. Fujiwara, A characterisation of exponential distribution semi-groups. J. Math. Soc. Japan. 18 (1966), 267 -- 274. [10] M. Hieber, Integrated semigroups and differential operators on Lp spaces. Math. Ann. 29 (1991), 1- 16. [11] M. Li, F. Huang and Q. Zheng, Local integrated C-semigroups. Studia Math. 145 (2001), 265 -- 280. [12] V. Keyantuo, Integrated semigroups and related partial differential equations. J. Math. Anal. Appl. 212 (1997), 135 -- 153. [13] H. Komatsu, Ultradistributions, I. Structure theorems and a characterization. J. Fac. Sci. Univ. Tokyo Sect. IA Math. 20 (1973), 25 -- 105. [14] H. Komatsu, Ultradistributions, III. Vector valued ultradistributions the theory of kernels. J. Fac. Sci. Univ. Tokyo Sect. IA Math. 29 (1982), 653 -- 718. [15] H. Komatsu, Microlocal Analysis in Gevrey Class and in Complex Domains, Microlocal analysis and applications. (Montecatini Terme, 1989), 161 -- 236, Lecutre Notes in Math., 1494, Springer, Berlin, 1991. [16] T. K¯omura, Semigroups of operators in locally convex spaces, J. Funct. Anal. 2 (1968), 258 -- 296. [17] M. Kosti´c, Generalized Semigroups and Cosine Functions. Mathematical Institute SANU, Belgrade, 2011. [18] M. Kosti´c, Abstract Volterra Integro-Differential Equations. Taylor and Francis Group/CRC Press/Science Publishers, New York, in press. [19] M. Kosti´c, Differential and analytical properties of semigroups of operators. Integral Equa- tions Operator Theory 67 (2010), 499 -- 557. [20] M. Kosti´c and S. Pilipovi´c, Ultradistribution semigroups. Siberian Math. J. 53 (2012), 232 -- 242. [21] M. Kosti´c, S. Pilipovi´c and D. Velinov, Structural theorems for ultradistribution semigroups. To appear in Siberian Math. J. [22] M. Kosti´c, S. Pilipovi´c and D. Velinov, C-distribution semigroups and C-ultradistribution semigroups in locally convex spaces. Preprint. [23] P. C. Kunstmann, Distribution semigroups and abstract Cauchy problems. Trans. Amer. Math. Soc. 351 (1999), 837 -- 856. [24] J. L. Lions, Semi-groupes distributions. Portugal. Math. 19 (1960), 141 -- 164. [25] I. V. Melnikova, A. I. Filinkov, Abstract Cauchy Problems: Three Approaches. Chapman & Hall/CRC, Boca Raton, 2001. [26] F. Neubrander, Integrated semigroups and their applications to the abstract Cauchy problem. Pacific J. Math. 135 (1988), 111 -- 155. [27] S. Pilipovi´c, Tempered ultradistributions, Boll. Un. Mat. Ital. 7 (1988), 235-251. [28] K. Yosida, Functional Analysis. Springer-Verlag, Berlin, 1965. ON THE EXPONENTIAL ULTRADISTRIBUTION SEMIGROUPS IN BANACH SPACES 9 [29] K. Yoshinaga, Values of vector-valued distributions and smoothness of semigroup distribu- tions. Bull. Kyushu Inst. Tech. Math. Nat. Sci. 12 (1965), 1 -- 27. [30] S. Wang, Quasi-distribution semigroups and integrated semigroups. J. Funct. Anal. 146 (1997), 352 -- 381. Faculty of Technical Sciences, University of Novi Sad, Trg D. Obradovi´ca 6, 21125 Novi Sad, Serbia E-mail address: [email protected] Department for Mathematics and Informatics, University of Novi Sad, Trg D. Obradovi´ca 4, 21000 Novi Sad, Serbia E-mail address: [email protected] Department for Mathematics, Faculty of Civil Engineering, Ss. Cyril and Methodius University, Skopje, Partizanski Odredi 24, P.O. box 560, 1000 Skopje, Macedonia E-mail address: [email protected]
1906.10349
1
1906
2019-06-25T06:58:09
Optimal extension of the Fourier transform and convolution operator on compact groups
[ "math.FA" ]
Let $G$ be a compact group (not necessarily abelian) and let $\Phi$ be a Young function satisfying the $\Delta_2$-condition. We determine the optimal domain and the associated extended operator for both Fourier transform and the convolution operator defined on the Orlicz spaces $L^\Phi(G).$
math.FA
math
OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR ON COMPACT GROUPS MANOJ KUMAR AND N. SHRAVAN KUMAR Abstract. Let G be a compact group (not necessarily abelian) and let Φ be a Young function satisfying the ∆2-condition. We determine the optimal domain and the associated extended operator for both Fourier transform and the convolution operator defined on the Orlicz spaces LΦ(G). 1. Introduction Let X be a σ-order continuous Banach function space [13] defined over a finite positive measure space (Ω,A, µ). If Y is any Banach space and T : X → Y is a bounded linear operator, define νT : A → Y as νT (A) = T (χA). It is well known that νT is a σ-additive vector measure. Further, by [13, Pg. 194, Theorem 4.14], X can be embedded continuously inside L1(νT ) and the mapping IνT : L1(νT ) → Y as the integration operator. Also, the space L1(νT ) is optimal in the sense that if given by IνT (f ) =RΩ f dνT is a continuous extension of T. The mapping IνT is called T extends continuously to a space eX with values in Y, then eX can be embedded continuously inside L1(νT ). The determination of the optimal domain and the extension for operators arising from analysis is one of the classical problems of functional analysis. This optimal extension process is known for kernel operators [1] and Sobolev embeddings [2, 5, 8]. In 2007, S. Okada and W. J. Ricker considered the optimal extension problem for the convolution operators arising from measures defined on compact abelian groups. Further, in [13, Chapter 7], the authors have also studied the Fourier transform operator on compact abelian groups from the optimal domain view point. This paper aims to study the optimal domain, extension and their properties for both convolution operator and Fourier transform defined on a compact group. The initial domain for these operators are the Orlicz spaces on a compact group. In this way, these results are new even when the underlying group is abelian. In section 3, we study the Fourier transform operators while in section 4, we study the convolution operator and the compactness properties of the extended operator. 2010 Mathematics Subject Classification. Primary 43A15, 43A30, 46G10; Secondary 46E30, 43A77. Key words and phrases. Compact group, Vector measure, Fourier transform, Convolution, Orlicz space. 1 2 MANOJ KUMAR AND N. SHRAVAN KUMAR 2. Preliminaries 2.1. Fourier analysis on compact groups. Let G be a compact Hausdorff group and let mG denote the normalized positive Haar measure on G. For 1 ≤ p ≤ ∞, Lp(G) will denote the usual pth-Lebesgue space with respect to the measure mG. It is well known that an irreducible unitary representation of a compact group G Let {(Xα,k.kα)}α∈∧ be a collection of Banach spaces. We shall denote by ℓ1- ⊕α∈∧ is always finite-dimensional. Let bG be the set of all unitary equivalence classes of irreducible unitary representations of G. The set bG is called the unitary dual of G and bG is given the discrete topology. Xα : Pα∈∧kxαkα < ∞(cid:27) equipped with the norm Xα, the Banach space (cid:26)(xα) ∈ Π k(xα)k∞ := Pα∈∧kxαkα. Similarly, we shall denote by ℓ∞- ⊕α∈∧ (cid:26)(xα) ∈ Π α∈∧kxαkα < ∞(cid:27) equipped with the norm k(xα)k∞ := sup α∈∧kxαkα. Xα, the space consisting of those Xα, the Banach space Xα : sup α∈∧ α∈∧ In the same lines, we shall also denote by c0- ⊕α∈∧ vectors (xα) from ℓ∞- ⊕α∈∧ is a closed subspace of ℓ∞- ⊕α∈∧ Xα. Xα which goes to 0 as α → ∞. It is clear that c0- ⊕α∈∧ Xα Let π be an irreducible unitary representation of G on the Hilbert space Hπ of dimension dπ. For f ∈ L1(G), the Fourier transform of f at π denoted bf (π), is defined as bf (π) =ZG f (t)π(t)∗ dmG(t), [π] ∈ bG. Note that the Fourier transform operator f 7→ bf maps L1(G) into ℓ∞- ⊕[π]∈ bG B2(Hπ). This operator is injective and bounded. For more details on compact groups, we refer to [6]. 2.2. Orlicz Spaces. Let Φ : [0,∞] → [0,∞] be a convex function. Then Φ is called a Young function if it satisfies Φ(0) = 0 and lim Φ(x) = +∞. If Φ is any Young x→∞ function, then define Ψ as Ψ(y) := sup {xy − Φ(x) : x ≥ 0}, y ≥ 0. Then Ψ is also a Young function and is termed as the complementary function to Φ. Further, the pair (Φ, Ψ) is called a complementary pair of Young functions. A pair (Φ, Ψ) of complementary Young functions is said to be normalized if Φ(1)+Ψ(1) = 1. We say that a Young function Φ satisfies the ∆2-condition if there exists a constant K > 0 and x0 > 0 such that Φ(2x) ≤ KΦ(x) whenever x ≥ x0. OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR 3 The Orlicz space, denoted LΦ(G), is a Banach space consisting of complex mea- surable functions on G for which kfkΦ := inf(cid:26)k > 0 :ZG Φ(cid:18)f k (cid:19) dmG ≤ 1(cid:27) < ∞. The above norm is called as the Luxemburg norm or Gauge norm. If (Φ, Ψ) is a complementary Young pair, then there is a norm on LΦ(G), equivalent to the Luxemburg norm, given by, Φ = sup(ZG f gdmG :ZG kfko Ψ(g)dmG ≤ 1). This norm is called as the Orlicz norm. Note that the Orlicz space is a natural generalisation of the classical Lp spaces. Let Φ be a Young function satisfying the ∆2-condition. Then the space of all continuous functions on G is dense in LΦ(G). Similarly, the space of simple functions is also dense in LΦ(G). Further, the dual of (LΦ(G),k·kΦ) is isometrically isomorphic to (LΨ(G),k · ko Ψ). In particular, if both Φ and Ψ satisfies the ∆2-condition, then LΦ(G) is reflexive [14, Pg. 112, Theorem 10]. Further, by [16, Pg. 595, Theorem 133.3], LΦ(G) is a σ-order continuous Banach function space. The following is an analogue of [7, Pg. 380, Lemma 35.11]. The proof of this follows as in the case of Lp spaces. Lemma 2.1. Let Φ be a Young function satisfying the ∆2-condition and let µ ∈ M(G) such that sup{kµ ∗ fkΦ : kfk1 ≤ 1} < ∞. Then there exists g ∈ LΦ(G) such that dµ = gdmG. For more on Orlicz spaces see [14]. 2.3. Vector measure. Let G be a compact group. Let B(G) denote the σ-algebra consisting of all Borel subsets of G. Further, let M(G) denote the space of all bounded complex Radon measures on G. Let X be a complex Banach space and let ν be a σ-additive X-valued measure called vector measure on G. Let X ′ be the dual of X and let BX ′ be the closed unit ball in X ′. For each x′ ∈ X ′, we shall denote by hν, x′i, the corresponding scalar valued measure for the vector measure ν, which is defined as hν, x′i(A) = hν(A), x′i, A ∈ B(G). A set A ∈ B(G) is said to be ν-null if ν(B) = 0 for every Borel set B ⊂ A. The variation of ν, denoted ν, is a positive measure defined as follows: For a set A ∈ B(G), ν(A) = sup(XE∈ρ kν(E)k : ρ the finite partition of A) . 4 MANOJ KUMAR AND N. SHRAVAN KUMAR The vector measure ν is said to be measure of bounded variation if ν(G) < ∞. x′∈BX ′hν, x′i(A), The semivariation of ν on a set A ∈ B(G) is given by kνk(A) = sup where hν, x′i is the total variation of the scalar measure hν, x′i. Let kνk denote the quantity kνk(G). The vector measure ν is said to be absolutely continuous with respect to a non-negative scalar measure µ if ν(A) = 0, A ∈ B(G). A Banach space X is said to have the Radon-Nikodym Property with respect to (G, B(G), mG) if for each X-valued absolutely continuos vector measure ν of bounded variation there exists a Bochner integrable function f : G → X such that dν = f dmG. A complex valued function f on G is said to be ν-weakly integrable if f ∈ w(ν) the Banach space of all L1(hν, x′i), for all x′ ∈ X ′. We shall denote by L1 ν-weakly integrable functions equipped with the norm lim µ(A)→0 kfkν = sup x′∈BX ′ZG f dhν, x′i. A ν-weakly integrable function f is said to be ν-integrable if for each A ∈ B(G) there exists a unique xA ∈ X such thatRA f dhν, x′i = hxA, x′i, x′ ∈ X ′. The vector xA is denoted byRA f dν. We shall denote by L1(ν) the space of all ν-integrable functions the continuous linear operator Iν : L1(ν) → X given by Iν(f ) =RG f dν, f ∈ L1(ν). and it is also a Banach space when equipped with the k · kν norm. We denote by Iν See [13, Pg. 152]. Note that the space of simple functions on G is dense in L1(ν). Let T : LΦ(G) → X be a bounded linear operator. Define νT : B(G) → X as νT (A) = T (χA). Then νT is a vector measure on G. The measure νT will be called as the vector measure associated with the operator T. See [13, Pg. 185]. For more details on vector measures and integration with respect to vector mea- sures, we refer to [3, 4, 13]. Throughout this paper G will be a compact group and mG will denote the unique normalized Haar measure on G. Further, Φ will always denote a Young function sat- isfying the ∆2-condition and Ψ will be the Young function which is complementary to Φ. Also, LΦ(G) will be considered only with the Luxemburg norm while its dual LΨ(G) will be considered only with the Orlicz norm. 3. Optimal extension of the Fourier transform In this section, we study the Fourier transform from two different view points. Firstly we restrict the Fourier transform operator on L1(G) to LΦ(G) by treating LΦ(G) as a subspace of L1(G). On the other hand, there is also a Fourier transform on the LΦ(G) spaces arising from the Hausdorff-Young inequality. The optimal domain for these operators is studied in this section. The results obtained in this section are motivated by the results presented in [13, Chapter 7]. OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR 5 3.1. Optimal extension of c0-valued Fourier transform. Define FΦ,0 : LΦ(G) → c0- ⊕[π]∈ bG B2(Hπ) as FΦ,0(f ) = bf . Here bf denotes the Fourier transform of f as an element of L1(G), since LΦ(G) is contained in L1(G). Further, it is well known that FΦ,0 is a bounded linear operator. We shall denote by νΦ,0 the associated vector measure. Lemma 3.1. The vector measure νΦ,0 has finite variation and the variation of νΦ,0 coincides with the Haar measure mG. Proof. Let A ∈ B(G). Then, Thus νΦ,0 is finite. Let π0 denote the one dimensional representation given by π0(t) = 1 for all t ∈ G. Let kνΦ,0(A)k∞ =kcχAk∞ ≤ kχAk1 = mG(A). χ{π0}(π) =( 1 [π] = [π0] [π] 6= [π0] if if 0dπ , for [π] ∈ bG. Here 0dπ denotes the dπ × dπ zero matrix. Then, it is clear that χ{π0} ∈ B2(Hπ ). Further, bG Bℓ1- ⊕ [π]∈ hνΦ,0(A), χ{π0}i =hcχA, χ{π0}i =X[π]∈ bG tr(cχA(π)χ{π0}(π)) =tr(cχA(π0)) = cχA(π0) =ZA π0(t−1) dmG(t) = mG(A). Therefore, mG(A) = hνΦ,0(A), χ{π0}i ≤ kνΦ,0(A)k∞. Hence kνΦ,0(A)k∞ = mG(A). Since A ∈ B(G) is arbitrary, it follows that νΦ,0 = mG. (cid:3) As an immediate corollary we have the following. Corollary 3.2. The measures νΦ,0 and mG have same null sets. Remark 3.3. By Corollary 3.2 and [13, Pg. 194, Theorem 4.14], the space L1(νΦ,0) is the largest Banach function space with σ-order continuous norm such that LΦ(G) is continuously embedded into L1(νΦ,0) and the operator FΦ,0 has a continuous ex- tension with values in c0- ⊕[π]∈ bGB2(Hπ). Further, this extension is unique and is the integration operator IνΦ,0. Corollary 3.4. The spaces L1(νΦ,0) and L1(G) coincide. Further, the extended op- erator IνΦ,0 coincides with the Fourier transform on L1(G). 6 MANOJ KUMAR AND N. SHRAVAN KUMAR Proof. By Lemma 3.1, it follows that the spaces L1(G) and L1(νΦ,0) coincide. Fur- ther, the inclusions L1(νΦ,0) ⊆ L1(νΦ,0) ⊆ L1(hνΦ,0, χ{π0}i) are clear, where the first inclusion follows from [13, Pg. 116, Lemma 3.14]. Here π0 and χ{π0} are defined as in proof of Lemma 3.1. Also, it follows from the proof of Lemma 3.1 that L1(hνΦ,0, χ{π0}i) coincides with L1(G). Thus the spaces L1(G) and L1(νΦ,0) coincide. Hence, by the density of the space continuous functions on G in both L1(G) and LΦ(G), it follows that the extended operator is just the Fourier transform on L1(G). (cid:3) Corollary 3.5. Let (Φ, Ψ) be a pair of complementary Young functions satisfying the ∆2-condition. If G is an infinite compact group then the domain of the optimal extension of FΦ,0 contains LΦ(G) as a proper subspace. Proof. Since Φ and Ψ both satisfy the ∆2-condition, the space LΦ(G) is reflexive and hence the proof follows from Corollary 3.4. (cid:3) 3.2. Optimal extension of the Hausdorff-Young inequality. Let (Φ, Ψ) de- note a pair of complementary Young functions which are normalized. Further, we shall denote by ℓΦ- ⊕[π]∈ bGB2(Hπ), the Banach space consisting of (x[π])[π]∈ bG ∈ Π [π]∈ bGB2(Hπ) for which k(x[π])kΦ := inf k > 0X[π]∈ bG d2 πΦ(cid:18)kx[π]kB2(Hπ ) k√dπ (cid:19) ≤ Φ(1) < ∞. The following is the Hausdorff-Young inequality for Orlicz spaces. See [15, Pg. 211, Theorem 2]. Also, see [15] for any undefined notations regarding this theorem. Theorem 3.6. Let (Φ, Ψ) be a complementary pair of Young functions which are continuous and normalized. Suppose that (i) Φ ≺′ Φ0, where Φ0(x) = c1x2 and (ii) Ψ′(x) ≤ c2xr, for some r ≥ 1, c1, c2 > 0. Then the Fourier transform is a bounded linear operator from LΦ(G) into ℓΨ- ⊕[π]∈ bG B2(Hπ). Throughout this subsection, the complementary Young pair (Φ, Ψ) will satisfy the conditions of the above theorem. We shall denote by FΦ,Ψ the operator from LΦ(G) to ℓΨ- ⊕[π]∈ bG B2(Hπ) given by the above theorem. Further, we shall also denote by νΦ,Ψ the associated vector measure. Lemma 3.7. The measures νΦ,Ψ and mG have same null sets. OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR 7 Proof. Let A ∈ B(G). Then mG(A) = 0 if and only ifcχA = 0 if and only if νΦ,Ψ(A) = 0. Hence the proof. Remark 3.8. By Lemma 3.7 and by [13, Pg. 194, Theorem 4.14], the space L1(νΦ,Ψ) is the largest Banach function space with σ-order continuous norm such that LΦ(G) is continuously embedded into L1(νΦ,Ψ) and the operator FΦ,Ψ has a continuous extension with values in ℓΨ- ⊕[π]∈ bGB2(Hπ). Further, this extension is unique and is the integration operator IνΦ,Ψ. (cid:3) Theorem 3.9. Let (Φ, Ψ) be a complementary pair of Young functions satisfying the ∆2-condition. (i) The space LΦ(G) is a dense subspace of L1(νΦ,Ψ) and kfkνΦ,Ψ ≤ kFΦ,ΨkkfkΦ, (ii) The space L1(νΦ,Ψ) is a dense subspace of L1(G) and kfk1 ≤ kfkνΦ,Ψ, f ∈ f ∈ LΦ(G). L1(νΦ,Ψ). (iii) The extended operator IνΦ,Ψ : L1(νΦ,Ψ) → ℓΨ- ⊕[π]∈ bGB2(Hπ) is just the Fourier transform map, i.e., IνΦ,Ψ(f ) = bf , for all f ∈ L1(νΦ,Ψ). Proof. (i) Since L1(νΦ,Ψ) is the optimal domain, it follows that LΦ(G) is contained in L1(νΦ,Ψ). Further, by [13, Pg. 185, Theorem 4.4], it follows that the inclusion of LΦ(G) inside L1(νΦ,Ψ) is continuous with its operator norm equal to kFΦ,Ψk. The conclusion follows from the density of simple functions in both the spaces. (ii) Consider the element χ{π0} from ℓΦ- ⊕[π]∈ bGB2(Hπ). Here π0 and χ{π0} are defined as in the proof of Lemma 3.1. Here, note that χ{π0} ∈ BℓΦ- ⊕ B2(Hπ ). Again, as in the proof of Lemma 3.1, one can show that mG = hνΦ,Ψ, χ{π0}i. Thus, if f ∈ L1(νΦ,Ψ), then [π]∈ bG ZG fdmG =ZG fdhνΦ,Ψ, χ{π0}i ≤ kfkνΦ,Ψ. Thus f ∈ L1(G). Further, the density of L1(νΦ,Ψ) in L1(G), again, follows from the density of the simple functions. (iii) This is again a consequence of the density of simple functions and (ii). (cid:3) 4. Optimal extension of the convolution operator In this section, we consider the convolution operator and determine its optimal domain. Further, we also characterize when the extended operator will be compact. The results of this section are analogues of the results obtained in [11]. Throughout this section, (Φ, Ψ) will denote a pair of complementary Young functions satisfying the ∆2-condition. 8 MANOJ KUMAR AND N. SHRAVAN KUMAR Let µ ∈ M(G). Define TΦ,µ : LΦ(G) → LΦ(G) as TΦ,µ(f ) = µ ∗ f, where µ ∗ f (t) =ZG f (s−1t)dµ(s). Then TΦ,µ is a bounded linear operator and kTΦ,µk ≤ kµk. Let νΦ,µ denote the associated vector measure. If dµ = gdmG, then νΦ,g will denote the measure νΦ,µ. 4.1. Optimal extension. Proposition 4.1. Let g ∈ LΦ(G). Let FΦ,g(t) = τtg, t ∈ G, where τtg(s) = g(st). (i) If f ∈ L1(G), then the function t 7→ f FΦ,g(t) from G to LΦ(G) is integrable (in the sense of Bochner). (ii) If f ∈ LΦ(G), then g ∗ f =RG f FΦ,g dmG. Proof. (i) Since G is compact, it follows that the range of the continuous mapping t 7→ f FΦ,g(t) is compact and hence separable. Thus, by Pettis measurability theorem [3, Pg. 42, Theorem 2], it follows that the function f FΦ,g is strongly measurable. By the fact that LΦ(G) is translation invariant and that the translations are norm (ii) By (i), Fubini's theorem and the fact that the dual of LΦ(G) is LΨ(G), we get that preserving, it follows thatRG kf FΦ,gkΦ dmG < ∞, thus proving (i). f FΦ,g dmG, h(cid:29) , ∀ h ∈ LΨ(G). hg ∗ f, hi =(cid:28)ZG (cid:3) Corollary 4.2. Let g ∈ LΦ(G). Then the LΦ(G)-valued measure νΦ,g can be written as νΦ,g(A) =ZA FΦ,g dmG, where FΦ,g is defined as in Proposition 4.1. Proof. Follows from (ii) of Proposition 4.1 by letting f = χA. (cid:3) Corollary 4.3. Let µ ∈ M(G). If f ∈ LΦ(G) such that dµ = f dmG a.e., then there exists an integrable function (in the sense of Bochner) F : G → LΦ(G) such that for every Borel set A of G, νΦ,µ(A) =ZA F dmG. Proof. By our assumption that the function f is the Radon-Nikodym derivative of µ w.r.t. the Haar measure, it follows that νΦ,µ = νΦ,f . Now the proof follows from Corollary 4.2. (cid:3) The following theorem is the converse to the previous corollary. OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR 9 exists f ∈ LΦ(G) such that dµ = f dmG a.e.. Theorem 4.4. Let µ ∈ M(G). Suppose that an integrable function (in the sense of Bochner) F : G → LΦ(G) satisfies νΦ,µ(A) =RA F dmG, A ∈ B(G). Then there Proof. Let F : G → LΦ(G) such that νΦ,µ(A) =RA F dmG. By [3, Pg. 42, Theorem 2], there exists a Borel set E of G such that mG(E) = 1 and {F (x) : x ∈ E} is separable in LΦ(G). So, without loss of generality, we can assume that F (G) is a separable subset of LΦ(G). Let FΦ,0 denote the Fourier transform defined as in section 3. As F (G) is separable and the map FΦ,0 is continuous, (FΦ,0 ◦ F )(G) is a separable subset of c0- ⊕[π]∈ bGB2(Hπ). Thus, by definition of the c0 direct sum, there (1) i.e., bf (π) = 0dπ ∀ f ∈ F (G) and ∀ [π] ∈ bG \ Γ. exists a countable subset Γ of bG such that if f ∈ F (G) then bf vanishes outside Γ, Let [π] ∈ bG. Now, consider the map eFπ : G → B2(Hπ) given by eFπ(t) = dF (t)(π). Since G is compact, LΦ(G) is continuously embedded in L1(G) and hence the above map is well-defined. Note that the linear maps FΦ,0 : LΦ(G) → c0- ⊕[π]∈ bGB2(Hπ) and Pπ : c0- ⊕[σ]∈ bGB2(Hσ) → B2(Hπ) are continuous, where Pπ is the projection onto B2(Hπ) and hence, by [13, Pg. 149], the composition Pπ ◦ FΦ,0 ◦ F is Bochner integrable. Further, as ZA eFπ(t) dmG(t) =ZA F (t) dmG(t)(cid:19) (Pπ and FΦ,0 are linear maps) it follows that eFπ is Bochner integrable. We now claim that, for every Borel subset A of G, eFπ = Pπ ◦ FΦ,0 ◦ F, (2) (3) where τtµ(A) = µ(At−1). Let A be a Borel subset of G. Then ZA eFπ(t) dmG(t) =ZA cτtµ(π) dmG(t), (Pπ ◦ FΦ,0 ◦ F )(t) dmG(t) (by (2)) =Pπ ◦ FΦ,0(cid:18)ZA =Pπ ◦ FΦ,0 (νΦ,µ(A)) = \νΦ,µ(A)(π) =\µ ∗ χA(π) = cχA(π)bµ(π) =(cid:18)ZA π(t−1) dmG(t)(cid:19)bµ(π) =ZA π(t−1)bµ(π) dmG(t) 10 MANOJ KUMAR AND N. SHRAVAN KUMAR π(st)∗ dµ(s) dmG(t) =ZAZG =ZA cτtµ(π) dmG(t). ZA Let [π] ∈ bG \ Γ. Then, by (1) and (3), we have, for every Borel set A of G, π(t−1)bµ(π) dmG(t) =ZAcτtµ(π) dmG(t) =ZA eFπ(t) dmG(t) = 0dπ . This implies that π(t−1)bµ(π) = 0dπ a.e. t ∈ G. As π is a unitary representation, it follows thatbµ(π) = 0dπ and hence cτtµ(π) = 0dπ , for every t ∈ G. Thus, (4) On the other hand, using the fact that Γ is countable and (3), there exists a Borel set A of G such that mG(A) = 1 and cτtµ(π) = eFπ(t) = 0dπ ∀ t ∈ G. cτtµ(π) = eFπ(t) ∀ t ∈ A and ∀ [π] ∈ Γ. By (4) and (5) we have (5) cτtµ(π) = eFπ(t) = dF (t)(π) ∀ t ∈ A and ∀ [π] ∈ bG and therefore by the uniqueness of the Fourier transform it follows that τtµ = F (t) ∀ t ∈ A. Now the conclusion follows by applying τt−1 on both sides. (cid:3) Lemma 4.5. If µ ∈ M(G), then the semivariation of νΦ,µ satisfies the following: for any Borel set A, µ(G) Ψ−1(1) mG(A) ≤ kνΦ,µk(A) = sup g∈BLΨ(G)ZA eµ ∗ g dmG ≤(cid:20)Φ−1(cid:18) 1 mG(A)(cid:19)(cid:21)−1 kµk. hνΦ,µ, gi(A) = hTΦ,µ(χA), gi = hχA, T ∗ Proof. Note that, for any g ∈ LΨ(G), T ∗ For any g ∈ LΨ(G), we have Thus, by [3, Pg. 46, Theorem 4] and from the definition of the semivariation of a vector measure, it follows that Φ,µ(g) =eµ ∗ g. Let A be a Borel subset of G. Φ,µ(g)i = hχA,eµ ∗ gi =ZAeµ ∗ g dmG. g∈BLΨ(G)ZA eµ ∗ g dmG. Ψ−1(1) χG(cid:13)(cid:13)(cid:13) Note that, by [14, Pg. 78, Corollary 7], it follows that(cid:13)(cid:13)(cid:13) 1 Ψ−1(1)ZA eµ ∗ χG dmG. the above equality, it follows that kνΦ,µk(A) = sup kνΦ,µk(A) ≥ = 1. Thus, from o Ψ 1 OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR11 Now the inequality on the left follows from the fact that eµ ∗ χG = µ(G). Further, the right inequality follows from the Holder's inequality for Orlicz spaces [14, Pg. 62, Remark 1] and [14, Pg. 78, Corollary 7]. (cid:3) Corollary 4.6. If µ ∈ M(G), then the vector measure νΦ,µ is absolutely continuous w.r.t. mG. Further, if µ(G) is non-zero, then mG is absolutely continuous w.r.t. νΦ,µ. Proof. The proof of this follows from Lemma 4.5. (cid:3) Remark 4.7. If µ ∈ M(G) is such that µ(G) 6= 0 then, by Corollary 4.6, the measures νΦ,µ and mG have same null sets. Hence, by [13, Pg. 194, Theorem 4.14], the space L1(νΦ,µ) is the largest Banach function space with σ-order continuous norm such that LΦ(G) is continuously embedded into L1(νΦ,µ) and the operator TΦ,µ has a continuous extension with values in LΦ(G). Further, this extension is unique and is the integration operator IνΦ,µ. Corollary 4.8. Let µ ∈ M(G). Then, there exists f ∈ LΦ(G) such that dµ = f dmG a.e. if and only if νΦ,µ < ∞. Proof. Forward part follows from Corollary 4.3 and [3, Pg. 46, Theorem 4(iv)]. For the converse, note that the space LΦ(G) is reflexive and hence, by [3, Pg. 76, Corol- lary 13], LΦ(G) has the Radon-Nikodym property. Thus the conclusion follows from Corollary 4.6 and Theorem 4.4. (cid:3) 4.2. Properties of the optimal domain. From now onwards µ will denote a complex Radon measure such that µ(G) 6= 0. Proposition 4.9. Let f be a complex-valued Borel measurable function on G. Then, f ∈ L1(νΦ,µ) if and only ifRG feµ ∗ g dmG < ∞ for every g ∈ LΨ(G). Further, if f ∈ L1(νΦ,µ), then (6) kfkνΦ,µ = sup g∈BLΨ(G)ZG feµ ∗ g dmG. (7) Proof. Since Φ satisfies the ∆2-condition, by [14, Pg. 139, Theorem 5], LΦ(G) does not contain a copy of c0 and hence, by [13, Pg. 138], the spaces L1(νΦ,µ) and L1 w(νΦ,µ) coincide. Let g ∈ LΨ(G). Then, by following the steps as in Lemma 4.5 and by [3, Pg. 46, Theorem 4], it follows that hνΦ,µ, gi(A) =ZA eµ ∗ g dmG. Thus, f ∈ L1(νΦ,µ) if and only if f ∈ L1 w(νΦ,µ), i.e.,RG f dhνΦ,µ, gi < ∞ for every g ∈ LΨ(G), i.e.,RG feµ ∗ g dmG < ∞ for every g ∈ LΨ(G). Now the formula for the norm of f ∈ L1(νΦ,µ) follows from the definition and from (7). Theorem 4.10. (cid:3) 12 MANOJ KUMAR AND N. SHRAVAN KUMAR (i) If f ∈ LΦ(G), then kfkνΦ,µ ≤ kfkΦkµk. (ii) The inclusion ı : LΦ(G) → L1(νΦ,µ) is continuous with the operator norm satisfying µ(G) ≤ kık ≤ kµk. Proof. (i) The proof follows from (6) and from the Holder's inequality for Orlicz spaces [14, Pg. 62, Remark 1]. (ii) The continuity of the inclusion map ı and the inequality on the right follows from (i). We now prove the inequality on the left side. By (6), it is clear that kık = sup g∈BLΨ(G)keµ ∗ gko Ψ Ψ for all g ∈ BLΨ(G). In particular, taking g = 1 Ψ−1(1) χG, we (cid:3) and hence kık ≥ keµ ∗ gko get the required inequality. Corollary 4.11. The space LΦ(G) is dense in L1(νΦ,µ). Proof. This follows from the previous theorem and the fact that the space of simple functions is dense in the both of the spaces w.r.t. their respective norms. (cid:3) Corollary 4.12. The extended map IνΦ,µ : L1(νΦ,µ) → LΦ(G) is given by IνΦ,µ(f ) = µ ∗ f. Proof. This follows from [13, Pg. 185, Theorem 4.4] and Corollary 4.11. (cid:3) Theorem 4.13. (i) If f ∈ L1(νΦ,µ), then kfk1 ≤ Ψ−1(1) µ(G) kfkνΦ,µ. (ii) The inclusion  : L1(νΦ,µ) → L1(G) is continuous with the operator norm satisfying (iii) The range of  is a dense and translation invariant subspace of L1(G). 1 kνΦ,µk ≤ kk ≤ Ψ−1(1) µ(G) . particular, taking g = 1 Proof. (i) It follows from (6) that kfkνΦ,µ ≥RG feµ ∗ g dmG for all g ∈ BLΨ(G). In (ii) The continuity of the inclusion map  and the inequality on the right follows from (i). We now prove the inequality on the left side. From the definition of the operator norm, it follows that kk ≥ k(f )k1 for all f ∈ BL1(νΦ,µ). In particular, taking f = 1 Ψ−1(1) χG, we get the required inequality. kνΦ,µk χG, we get the required inequality. Borel set A of G, Borel set A of G, νΦ,µ(A) =ZA νΦ,µ(A) =ZA F1 dmG. F2 dmG. OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR13 (iii) Denseness of the range of  follows from the fact that the space of simple functions is dense in the both of the spaces w.r.t. their respective norms. Further, translation invariance follows from (6) and the fact that translation on LΨ(G) is an isometry. (cid:3) 4.3. Compactness of the extended operator. A function f : G → X is said to be Pettis integrable if hf, x′i ∈ L1(G), x′ ∈ X ′ and if for each A ∈ B(G) there exists a unique xA ∈ X such thatRAhf, x′i dmG = hxA, x′i, x′ ∈ X ′. We now characterize the compactness of the extended operator in terms of the measure νΦ,µ. Theorem 4.14. Let µ ∈ M(G). Then the following are equivalent: (i) The LΦ(G)-valued integral operator IνΦ,µ is compact. (ii) The measure νΦ,µ has finite variation. (iii) There exists a LΦ(G)-valued Bochner integrable function F1 such that for every (iv) There exists a LΦ(G)-valued Pettis integrable function F2 such that for every (v) There exists a Borel set A of G such that νΦ,µ(A) is non-zero and finite. (vi) There exists f ∈ LΦ(G) such that dµ = f dmG. Proof. (i)⇒(ii) is a consequence of [12, Theorem 4]. (ii)⇒(iii) follows from Corollary 4.8 and Corollary 4.3. (iii)⇒(iv) follows from the fact that a Bochner integrable function is Pettis inte- grable [13, Pg. 149]. (iv)⇒(v) is a consequence of the fact that a vector measure defined by using a Pettis integrable function has σ-finite variation [9, Corollary 1]. (v)⇒(vi). Let A be a Borel subset subset of G such that 0 < νΦ,µ(A) < ∞. Then, by [13, Pg. 106] and Lemma 4.5, mG(A) > 0. As LΦ(G) is reflexive, by [3, Pg. 76, Corollary 13] applied to the measure νΦ,µ restricted to A, we get a function F : A → LΦ(G) which is the Radon-Nikodym derivative of νΦ,µ restricted to A w.r.t. the Haar measure mG. Now the argument given for Theorem 4.4 can be modified to obtain the necessary conclusion. (vi)⇒(i). By Corollary 4.8 the measure νΦ,µ has finite variation and by [3, Pg. 46, Theorem 4] and Corollary 4.2, it follows that the variation of νΦ,µ is kfkΦmG. Further, the Radon-Nikodym derivative dνΦ,µ with a compact range. Thus the compactness of the extended operator IνΦ,µ follows from [12, Theorem 1]. (cid:3) dνΦ,µ is FΦ,f kf kΦ 14 MANOJ KUMAR AND N. SHRAVAN KUMAR Our next characterization is in terms of the L1-spaces. Corollary 4.15. Let µ ∈ M(G). Then the following are equivalent: (i) The LΦ(G)-valued integral operator IνΦ,µ is compact. (ii) The spaces L1(G) and L1(νΦ,µ) coincide. (iii) The spaces L1(G) and L1(νΦ,µ) coincide. (iv) The spaces L1(νΦ,µ) and L1(νΦ,µ) coincide. Proof. (i)⇒(iv) follows from [12, Theorem 1]. Conversely, (iv) implies that νΦ,µ is finite as the constant function 1 ∈ L1(νΦ,µ). Hence, by Theorem 4.14, (i) follows. (i)⇒(ii). By Theorem 4.14 there exists f ∈ LΦ(G) such that dµ = f dmG. Further, by [3, Pg. 46, Theorem 4] and Corollary 4.2, it follows that the variation of νΦ,µ is kfkΦmG. Hence (ii). (ii)⇒(iii) follows from [13, Pg. 116, Lemma 3.14] and Theorem 4.13. We now prove (iii)⇒(i). Let f ∈ L1(G). By our assumption, f ∈ L1(νΦ,µ) and hence, by Corollary 4.12, µ ∗ f ∈ LΦ(G). Note that, by [13, Pg. 152], sup f ∈BL1(G)kµ ∗ fkΦ = sup f ∈BL1(νΦ,µ)kIνΦ,µ(f )k = kIνΦ,µk = 1. Now the conclusion follows from Lemma 2.1 and Theorem 4.14. (cid:3) As our final result, we show that the optimal extension is genuine. Corollary 4.16. Let G be an infinite compact group and let µ ∈ M(G) be such that dµ = f dmG for some f ∈ LΦ(G). Then the optimal extension of TΦ,µ is proper, i.e., LΦ(G) is a proper subspace of L1(νΦ,µ). Proof. By Theorem 4.10, LΦ(G) ⊆ L1(νΦ,µ). Further, our assumption implies, by Theorem 4.14 and Corollary 4.15, that LΦ(G) ⊆ L1(G). As Φ and Ψ satisfy the ∆2 condition, LΦ(G) is reflexive and as G is infinite it follows that LΦ(G) is a proper subspace of L1(G) and hence a proper subspace of L1(νΦ,µ). (cid:3) Acknowledgement The first author would like to thank the University Grants Commission, India, for providing the research grant. References [1] G. P. Curbera and W. J. Ricker, Optimal domains for kernel operators via interpolation, Math. Nachr., 244 (2002) 4763. [2] G. P. Curbera and W. J. Ricker, Optimal domains for the kernel operator associated with Sobolevs inequality, Studia Math., 158 (2003) 131152. [3] J. Diestel and J. J. Uhl, Vector Measures, Math. Surveys, vol. 15. Amer. Math. Soc., Providence (1977). OPTIMAL EXTENSION OF THE FOURIER TRANSFORM AND CONVOLUTION OPERATOR15 [4] N. Dinculeanu, Vector measures, VEB Deutscher Verlag der Wissenschaften, Berlin (1966). [5] D. Edmunds, R. Kerman and L. Pick, Optimal Sobolev imbeddings involving rearrangement-invariant quasinorms, J. Funct. Anal., 170 (2000) 307355. [6] G. B. Folland, A Course in Abstract Harmonic Analysis, CRC Press, Boca Raton, 1995. [7] E. Hewitt and K. A. Ross, Abstract harmonic analysis. Vol II: Structure and Analy- sis for Compact Groups Analysis on Locally Compact Abelian Groups, Grundlehren der Math.Wissenschaften, 152, Springer, 1970. [8] R. Kerman and L. Pick, Optimal Sobolev imbeddings, Forum Math., 18 (2006) 535570. [9] K. Musial, The weak Radon-Nikodym property in Banach spaces, Studia Math., 64 (1979) 151-174. [10] S. Okada and W. J. Ricker, Optimal domains and integral representations of convolution operators in Lp(G), Integr. Equ. Oper. Theory, 48 (2004), 525-526. [11] S. Okada and W. J. Ricker, Optimal domains and integral representations of Lp(G)- valued convolution operators via measures, Math. Nachr., 280 (2007), 423-436. [12] S. Okada, W. J. Ricker and L. Rodr´ıguez-Piazza, Compactness of the integration operator associated with a vector measure, Studia Math., 150 (2002) 133-149. [13] S. Okada, W. J. Ricker, and E. A. S´anchez P´erez, Optimal Domain and Integral Extension of Operators acting in Function Spaces, Oper. Theory Adv. Appl., vol. 180. Birkhauser, Basel (2008). [14] M. M. Rao and Z. D. Ren, Theory of Orlicz spaces, Dekker, New York, 1991. [15] M. M. Rao and Z. D. Ren, Applications of Orlicz spaces, Dekker, New York, 2002. [16] A. C. Zaanen Riesz spaces, Volume II, North Holland, 1983. Department of Mathematics Indian Institute of Technology Delhi Delhi - 110016 India E-mail address: [email protected] Department of Mathematics Indian Institute of Technology Delhi Delhi - 110016 India E-mail address: [email protected]
1503.06112
1
1503
2015-03-20T15:31:34
Norm-attaining functionals and proximinal subspaces
[ "math.FA" ]
G. Godefroy asked whether, on any Banach space, the set of norm-attaining functionals contains a 2-dimensional linear subspace. We prove that a recent construction due to C.J. Read provides an example of a space which does not have this property. This is done through a study of the relation between the following two sentences where X is a Banach space and Y is a closed subspace of finite codimension in X: (A) Y is proximinal in X. (B) The annihilator of Y consists of norm-attaining functionals. We prove that these are equivalent if X is the Read's space. Moreover, we prove that any non-reflexive Banach space X with any given closed subspace Y of finite codimension at least 2 admits an equivalent norm such that (B) is true and (A) is false.
math.FA
math
NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES MARTIN RMOUTIL Abstract. G. Godefroy asked whether, on any Banach space, the set of norm- attaining functionals contains a 2-dimensional linear subspace. We prove that a recent construction due to C.J. Read provides an example of a space which does not have this property. This is done through a study of the relation between the following two sentences where X is a Banach space and Y is a closed subspace of finite codimension in X: (A) Y is proximinal in X. (B) Y ⊥ consists of norm-attaining functionals. We prove that these are equivalent if X is the Read's space. Moreover, we prove that any non-reflexive Banach space X with any given closed subspace Y of finite codimension at least 2 admits an equivalent norm such that implication (B) =⇒ (A) fails. 1. Introduction Let X = (X, k · k) be a normed linear space. Then by X ∗ we denote the topo- logical dual of X and by NA(X) = NA(X, k · k) the set of all functionals in X ∗ for which there exists x ∈ SX such that x∗(x) = kxk; its elements are norm-attaining functionals. A set F ⊂ X is said to be proximinal if for each x ∈ X there is y ∈ F with kx − yk = dist(x, F ). Note that proximinal sets are closed. In this paper we investigate the relation of the following two sentences in which X is a Banach space and Y ⊂ X is a closed subspace: (A) Y is proximinal in X. (B) Y ⊥ := {x∗ ∈ X ∗; Y ⊂ Ker(x∗)} ⊂ NA(X). In view of the easy observation that a functional x∗ ∈ X ∗ is norm-attaining if and only if Ker(x∗) is proximinal (see e.g. Corollary 3.4), it is reasonable to ask whether sentences (A) and (B) are also equivalent when the codimension of Y in X is higher than 1. And, indeed, implication (A) =⇒ (B) is true whenever the quotient space X/Y is reflexive (cf. Lemma 3.3). In particular, it holds when the codimension of Y in X is finite (and therefore dim(X/Y ) < ∞) which is the case we are interested in the most. Date: March 6, 2018. 2010 Mathematics Subject Classification. Primary 46B10, 46B20; Secondary 46B03. Key words and phrases. proximinal subspace, norm-attaining functional, lineability, Banach space, non-reflexive space, equivalent norm. The research leading to these results has received funding from the Grant GA CR P201/12/0290, and the European Research Council under the European Union's Seventh Framework Programme (FP/2007-2013) / ERC Grant Agreement n.291497. 1 2 MARTIN RMOUTIL However, implication (B) =⇒ (A) is not necessarily true even if Y has finite codimension. The validity of this implication (and related statements) for vari- ous spaces has been studied by several authors (e.g. A.L. Garkavi, V. Indumathi or G. Godefroy) and counterexamples are known: e.g. Indumathi [12, Section 4.] shows that in any infinite-dimensional space L1(T, ν) (here (T, ν) is a positive mea- sure space) there exists a closed non-proximinal subspace M of finite codimension d ≥ 2 such that all subspaces N ) M are proximinal; in particular, L1(T, ν) fails implication (B) =⇒ (A). In another work [13] of Indumathi, the author studies the relation of (A) and (B) (and similar statements) in more detail and defines a space to be R(1) if implication (B) =⇒ (A) is true for any closed subspace Y ⊂ X of finite codimension (in general, a non-reflexive space is said to be R(n) space if proximinality of its closed subspaces of finite codimension at least n is implied by the proximinality of all n-codimensional subspaces containing it). In Section 5 we show that, given any non-reflexive Banach space X and its closed subspace Y of finite codimension at least 2, there is an equivalent norm · on X such that Y is not proximinal in (X, ·), but Y ⊥ ⊂ NA(X, ·) (see Theorem 5.6). In particular, any non-reflexive Banach space can be renormed so that it is not R(1). The main motivation for the present work is the following Problem of G. Gode- froy together with a recent theorem of C.J. Read below. Problem 1.1 ([9, Problem III.], also e.g. [2] or [1]). Does the set NA(X) contain a 2-dimensional subspace for any Banach space X? The well-known theorem of E. Bishop and R.R. Phelps [4] states that for any Banach space X, the set NA(X) is dense in X ∗; in particular, NA(X) is nonempty and so it clearly contains 1-dimensional subspaces. Since then, a number of papers studying the structure of NA(X) from various viewpoints have been written; the ones which are interesting for us are concerned with the linear structure of NA(X), in particular the question whether NA(X) is lineable or spaceable (for definition see Section 2). For example, in [2] it is shown that if X is an Asplund space enjoying the Dunford-Pettis property, then NA(X) is not spaceable (e.g. C(K) with K scattered (Hausdorff) compactum; however, NA(C(K)) for K infinite is always lineable-see [1]). On the other hand, in [1] the authors prove that if the Banach space X possesses an infinite-dimensional complemented subspace with a Schauder basis, then X can be equivalently renormed so that NA(X) is lineable. Some articles also investigate the lineability of X ∗ \ NA(X), e.g. [1] or [7]. In [7] it is observed that if J denotes the James space, then J ∗ \ NA(J ) is not even 2-lineable (i.e. (J ∗ \ NA(J )) ∪ {0} contains no 2-dimensional subspace). A related problem (again [9, Problem III.]) was to decide whether every Banach space contains a proximinal subspace of codimension 2. This problem was recently solved by C.J. Read: Theorem R (C.J. Read [20]). Let · be the renorming of c0 from Definition 1. Then (c0, ·) contains no proximinal subspaces of finite codimension 2 or larger. In Section 4 we show that this space also works as a counterexample for Prob- lem 1.1. More precisely, our Theorem 4.2 states that NA(c0, ·) contains no 2-dimensional subspaces. This is done by showing that (c0, ·) is R(1) (i.e. im- plication (B) =⇒ (A) holds for it) and then applying Read's result. Among the known sufficient conditions for a Banach space X to be R(1) are the following three: • the dual space X ∗ is (Gateaux) smooth [13, Corollary 2]; NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 3 • X is WLUR (for definition see Section 2.) [13, Corollary 3]; • X is a subspace of c0 [10]. None of these seems to be suitable in the case of (c0, ·), however; our proof has to be more direct, although it is done similarly to our new proof of the sufficiency of X being WLUR (see Corollary 3.6). The essential (but simple) tool we use throughout the paper is a characterization of conditions (A) and (B) in terms of the quotient space X/Y (see parts (i) and (ii) of Lemma 3.3); for example, a closed subspace Y ⊂ X is proximinal if and only if the quotient mapping q : X → X/Y maps BX onto the whole BX/Y (where BZ denotes the closed unit ball in Z), and we have a similar condition characterizing the fact that Y ⊥ ⊂ NA(X). It is not difficult to see that our characterization of proximinality is equivalent to that of Garkavi [8] (see [13, Theorem A]), which is not formulated using the quotient space. The characterization of Garkavi was used in many subsequent articles on the topic, but it seems that our approach using quotient spaces is, at least in some situations, more natural. In Lemma 3.3 (iv) we use our characterizations of (A) and (B) to prove that implication (B) =⇒ (A) holds whenever X/Y is strictly convex. Moreover, in Proposition 3.5 we show that this is the case whenever the norm on X is WLUR (thus obtaining a different proof of this known sufficient condition for R(1)). This is not completely trivial since in Example 3.7 we have a strictly convex Banach space X and its closed subspace Y of codimension 2 such that X/Y is not strictly convex. The proof of Theorem 4.2 (i.e. that Read's space solves Problem 1.1) is done by contradiction: Setting X = (c0, ·) we assume that there is a closed subspace Y ⊂ X of finite codimension at least 2 such that Y ⊥ ⊂ NA(X), and then we prove that X/Y is strictly convex. It follows from Lemma 3.3 that Y is proximinal which is a contradiction with Theorem R. We also use quotient spaces to obtain our renorming of a non-reflexive space X such that the resulting space contradicts implication (B) =⇒ (A). Indeed, in Lemma 5.5 we take any closed subspace Y ⊂ X of finite codimension in X. Given a set A ⊂ X/Y such that UX/Y ⊂ A ⊂ BX/Y satisfying some technical conditions, we find a renorming of X such that q(BX ) = A. Then, using our Lemma 3.3 we show in Theorem 5.6 that if the codimension of Y is at least 2, then A can be chosen in such a way that the resulting renorming satisfies (B), but not (A). 2. Notation and Definitions Let X = (X, k · k) be a normed linear space over the field of real numbers. Then 0 is the origin in X (there is no risk of confusion), X ∗ is its (topological) dual, BX its closed unit ball, UX is the open unit ball and SX denotes the unit sphere. For a functional x∗ ∈ X ∗ we use the notation Ker(x∗) := (x∗)−1(0) and att(x∗) := {x ∈ BX ; x∗(x) = kx∗k}. As explained in the Introduction, NA(X) = {x∗ ∈ X ∗; att(x∗) 6= ∅}. If Y is a normed linear space, by X ∼= Y we mean that X is linearly isomorphic to Y , and X ⊕∞Y is the direct sum of X and Y equipped with the maximum norm, i.e. for (x, y) ∈ X ⊕∞ Y we have k(x, y)k = max{kxk, kyk}. If Y is a subspace of X, we write Y ⊥ for its annihilator {x∗ ∈ X ∗; Y ⊂ Ker(x∗)}. Recall that the quotient space X/Y consists of all cosets of the form x + Y and is equipped with the norm kx + Y k = dist(0, x + Y ). If X is a Banach space and Y is closed, then X/Y is a Banach space and its dimension is equal to the codimension of 4 MARTIN RMOUTIL Y in X. We shall often work with the quotient mapping q : X → X/Y : x 7→ x + Y , which is a linear operator satisfying kqk = 1. Next, if A ⊂ X is any set, then Int(A), A and ∂A stand for the interior, closure w→ x we mean that xn converges to x and boundary of A respectively. By xn weakly in X, and similarly for w∗. If we do not specify which topology we mean, it is usually the norm topology; e.g. xn → x means limn→∞ kxn − xk = 0. We recall that the norm k · k on X is called strictly convex (or rotund ; SC for short) if all the points of SX are extremal points of BX , that is, if there are no nontrivial line segments contained in SX . In that case X is also called strictly convex. xn limn→∞ kx + xnk = 2kxk. Then the space X is called WLUR. The norm k · k is said to be weakly locally uniformly rotund (WLUR for short) if w→ x whenever x ∈ X and (xn)∞ n=1 ⊂ X are such that limn→∞ kxnk = kxk and For any sequence a ∈ ℓ∞ we denote a = (cid:0)aj(cid:1)∞ Let A be a bounded set in a Banach space. We denote by convσ(A) the σ-convex j=1 = (cid:0)a1, a2, . . .(cid:1). hull of A, that is convσ(A) = n ∞ X i=1 λixi; (λi)∞ i=1 ⊂ [0, 1], ∞ X i=1 λi = 1 and (xi)∞ i=1 ⊂ Ao. A subset M of a Banach space is said to be: • n-lineable if M ∪ {0} contains an n-dimensional subspace; • lineable if M ∪ {0} contains an infinite-dimensional subspace; • spaceable if M ∪ {0} contains an infinite-dimensional closed subspace. A comprehensive survey on linear subspaces of various sets is [3]. 3. Observations and lemmas The following observations are obvious and will be applied without reference. Observations 3.1. Let X be a normed linear space. (i) Any proximinal set in X is closed. (ii) Let F be a proximinal set in X and let x ∈ X. Then x + F is also proximinal in X. (iii) A subspace Y ⊂ X is proximinal if and only if for any x ∈ X there is a point z with minimal norm in the coset x + Y . The next facts are well-known; we include them for the sake of completeness. Fact 3.2. n ∈ X ∗. 1, . . . , x∗ i ) ⊂ Ker(y∗). Then (i) Let X be a normed linear space and y∗, x∗ i=1 Ker(x∗ y∗ ∈ span{x∗ n} if and only if Tn 1, . . . , x∗ (ii) Let Z be any normed linear space. Then Z is not strictly convex if and only if there is a functional ϕ ∈ Z ∗ such that att(ϕ) contains more than one point. (iii) Let X be a Banach space, Y ⊂ X be a closed subspace and q : X → X/Y the quotient mapping. Then q∗ : (X/Y )∗ → Y ⊥ is an isometry onto Y ⊥. Proof. (i) and (iii) are standard; for proof of (iii) see e.g. [6, Proposition 2.6]. (ii): Assume that Z is not strictly convex. Take two distinct points u, v ∈ SZ such that for any λ ∈ (0, 1), λu + (1 − λ)v ∈ SZ , and find (using the Hahn–Banach theorem) a functional ϕ ∈ SZ ∗ such that ϕ( u+v 2 ) = 1. Clearly, if ϕ(u) < 1, then NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 5 ϕ(v) > 1, and vice versa. But kϕk = 1, whence ϕ(u) = ϕ(v) = 1. The opposite implication is trivial. (cid:3) The following lemma translates properties (A) and (B) to the language of quo- tient spaces, and is essential in the sequel. Lemma 3.3. Let X be a Banach space, Y ⊂ X be a closed subspace and q : X → X/Y be the quotient mapping. Then: (i) Y is proximinal in X ⇐⇒ q(BX ) = BX/Y . (ii) Y ⊥ ⊂ NA(X) ⇐⇒ (∀ϕ ∈ (X/Y )∗) (∃u ∈ q(BX )) ϕ(u) = kϕk. (iii) If X/Y is reflexive and Y is proximinal, then Y ⊥ ⊂ NA(X). (iv) If X/Y is strictly convex and Y ⊥ ⊂ NA(X), then Y is proximinal. Proof. (i): Assume Y is proximinal. Obviously q(BX ) ⊂ BX/Y ; to prove the other inclusion, take any u ∈ BX/Y . Further, pick z ∈ q−1(u). Then dist(0, z + Y ) = kuk ≤ 1, and by Observation 3.1 (iii) there is a point y ∈ z + Y such that kyk = dist(0, z + Y ), so q(y) = u and y ∈ BX . To prove the opposite implication, assume that q(BX ) = BX/Y and take an arbitrary x ∈ X. By homogeneity we can clearly assume that dist(0, x + Y ) = 1, so q(x) ∈ BX/Y , and our assumption now implies that there exists y ∈ BX such that q(y) = q(x) (which is equivalent to y ∈ x + Y ). By Observation 3.1 (iii) we are done. (ii): Let Y ⊥ ⊂ NA(X) and take an arbitrary ϕ ∈ (X/Y )∗. Then y∗ := ϕ ◦ q ∈ Y ⊥, and so there exists x ∈ BX such that ϕ(q(x)) = y∗(x) = ky∗k = kϕk (by Fact 3.2 (iii)). Hence q(x) ∈ q(BX ) ∩ att(ϕ). Conversely, take any y∗ ∈ Y ⊥ and the corresponding ϕ ∈ (X/Y )∗ such that y∗ = ϕ ◦ q. By the assumption there exists u ∈ q(BX ) ∩ att(ϕ). So there also exists x ∈ BX ∩ q−1(u). Then y∗(x) = ϕ(q(x)) = ϕ(u) = kϕk = ky∗k, and y∗ is therefore norm-attaining. (iii): Trivial from (i), (ii) and the fact that on a reflexive space, all functionals are norm-attaining. (iv): Assume the condition in (ii) is satisfied. Clearly UX/Y ⊂ q(BX ) ⊂ BX/Y , so by (i) it suffices to show that SX/Y ⊂ q(BX ). To that end, take an arbitrary u ∈ SX/Y and use the Hahn-Banach theorem to find ϕ ∈ S(X/Y )∗ such that u ∈ att(ϕ). Since X/Y is strictly convex, it follows by Fact 3.2 that att(ϕ) = {u}, and the condition in (ii) now implies that u ∈ q(BX ). (cid:3) Corollary 3.4. Let X be a Banach space. Then: (a) Let Y ⊂ X be a closed subspace of finite codimension in X. If Y is proximinal, then Y ⊥ ⊂ NA(X). (b) A functional x∗ ∈ X ∗ is norm-attaining if and only if Ker(x∗) is proximinal. Proof. (a): The quotient space X/Y is finite-dimensional and in particular it is reflexive. (b): In addition to being reflexive, the quotient space X/ Ker(x∗) is (for trivial (cid:3) reasons) also strictly convex, and the statement follows. Proposition 3.5. Let X be a Banach space with a WLUR norm, and Y ⊂ X be its closed subspace such that Y ⊥ ⊂ NA(X). Then X/Y is strictly convex. Proof. Aiming for a contradiction, suppose that X is WLUR and X/Y is not strictly convex. Then we can find a functional ϕ ∈ (X/Y )∗ such that kϕk = 1 and att(ϕ) 6 MARTIN RMOUTIL contains more than one point (Fact 3.2 (ii)). Now, set x∗ = ϕ◦q where q : X → X/Y is the quotient mapping. Fact 3.2 (iii) gives us that kx∗k = kϕk = 1 and x∗ ∈ Y ⊥. As Y ⊥ ⊂ NA(X), x∗ attains its norm at a point x ∈ SX and ϕ(q(x)) = x∗(x) = 1, whence q(x) ∈ att(ϕ). Take any u ∈ att(ϕ) \ {q(x)} and any sequence of elements xn ∈ u such that limn→∞ kxnk = 1 = kxk. Then limn→∞ kx + xnk = 2 since 2 = ϕ(q(xn) + q(x)) = x∗(xn + x) ≤ kxn + xk ≤ kxnk + kxk → 2, n → ∞. Our assumption that X is WLUR yields that xn converges to x weakly in X. To obtain the desired contradiction, take any ψ ∈ (X/Y )∗ such that ψ(q(x)) 6= ψ(u) and set y∗ := ψ ◦ q ∈ X ∗. Then for all n ∈ N we have y∗(x) = ψ(q(x)) 6= ψ(u) = y∗(xn), so y∗(xn) does not converge to y∗(x) which is a contradiction. (cid:3) Let us note that these observations provide a new proof for the following result of Indumathi (see [13, Corollary 3]). Corollary 3.6. Let X be a Banach space with a WLUR norm and Y ⊂ X be its closed subspace of finite codimension. Then Y is proximinal if and only if Y ⊥ ⊂ NA(X). Proof. The statement follows immediately from Proposition 3.5, Lemma 3.3 and Corollary 3.4. (cid:3) The following is a simple example of a strictly convex space X with a finite- codimensional closed subspace Y such that X/Y is not strictly convex. Example 3.7. By [0, ω] we mean the compact interval of ordinals with the order topology. Let us define the Banach space X as C([0, ω] × {0, 1}) with the norm k · k defined by ∞ kxk2 = kxk2 ∞ + X k=0 2−k(cid:0)x(k, 0)2 + x(k, 1)2(cid:1). It is easy to check that k · k is indeed a norm on C([0, ω] × {0, 1}), which is moreover equivalent to k · k∞. It is also clear that the dual to (C([0, ω] × {0, 1}), k · k∞) is (ℓ1([0, ω] × {0, 1}), k · k1), so X ∗ is isomorphic to the latter space. Set Y := {x ∈ X; x(ω, 0) = x(ω, 1) = 0}; then Y is a closed subspace of codimension 2 and Y ⊥ = span{δ(ω,0), δ(ω,1)} and Y . We claim that X is strictly convex, but X/Y is not. Let x, y ∈ X be arbitrary points satisfying 2kxk2 + 2kyk2 − kx + yk2 = 0; (1) [6]) it by a well-known (and easy) characterization of strict convexity (see e.g. suffices to prove that x = y. One can rewrite equation (1) using the definition of k · k, and observe that the left hand side is the sum of the term 2kxk2 ∞ + 2kyk2 ∞ − kx + yk2 ∞, and all terms of the form 2−k(cid:0)2x(k, i)2 + 2y(k, i)2 − (x(k, i) + y(k, i))2(cid:1), (k, i) ∈ [0, ω) × {0, 1} which are easily seen to be non-negative. Since the sum of all these terms is equal to 0, it follows that so is each of the terms. In particular, x(k, i) = y(k, i) for all (k, i) ∈ [0, ω) × {0, 1}, and the continuity of x and y now implies that x = y. NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 7 We shall now prove that X/Y is isometrically isomorphic to (R2, k · k∞); in particular, X/Y is not strictly convex. Consider the quotient mapping q : X → X/Y : x 7→ x + Y = q(x). Obviously z ∈ q(x) if and only if z(ω, i) = x(ω, i), i = 0, 1, whence the mapping L : X/Y → R2 : x + Y 7→ (x(ω, 0), x(ω, 1)) is a linear bijection. It remains to show that L : X/Y → (cid:0)R2, k · k∞(cid:1) is an isometry. Let x ∈ X be given; by definition, kx + Y k = dist(0, x + Y ) ≥ max{x(ω, 0), x(ω, 1)}. To prove the opposite inequality, set for all N ∈ N, xN := x · χ[N,ω]×{0,1}, where χA denotes the characteristic function of A ⊂ [0, ω] × {0, 1}; then xN ∈ x + Y as xN (ω, i) = x(ω, i), i = 0, 1. Now, kxN k2 = kxN k2 ∞ + ∞ X k=N 2−k(cid:0)x(k, 0)2 + x(k, 1)2(cid:1), and it follows from the continuity of x that lim N→∞ kxN k = max{x(ω, 0), x(ω, 1)}; hence, dist(0, x + Y ) ≤ max{x(ω, 0), x(ω, 1)}. In view of the last example and Proposition 3.5 it is natural to ask the following. Problem 3.8. Can the assumption that X be WLUR in Proposition 3.5 be weak- ened to SC? Equivalently, can Example 3.7 be done in such a way that we moreover have Y ⊥ ⊂ NA(X)? Remark 3.9. The following classical results are related to the problem above. Let X be a Banach space. (a) Let X be SC and Y be a reflexive subspace of X. Then X/Y is SC (V. Klee, [17]). (b) The following are equivalent (see e.g. M. Day, [5, Chapter VII, §2]): (i) X ∗ is Gateaux smooth (equivalently, at each point of SX ∗ there is only one supporting hyperplane of BX ∗); (ii) every quotient of X is SC; (iii) every 2-dimensional quotient of X is SC. It is interesting to note that having a strictly convex norm on X/Y is not the only known condition which ensures that the implication (B) =⇒ (A) holds. The following theorem can be found in [10] or [9]: Theorem GI (G. Godefroy, V. Indumathi). Let X be a closed subspace of c0 and Y ⊂ X be a closed finite-codimensional subspace of X. Then Y is proximinal in X if and only if Y ⊥ ⊂ NA(X). Instead of showing that X/Y is strictly convex (which is not the case), the authors essentially prove that X/Y is polyhedral (i.e. the unit ball is the convex hull of finitely many points), and then show that each of the extremal points is contained in q(BX ) where q : X → X/Y is the quotient mapping (in the same manner as we show it for each point of the sphere SX/Y ); it then follows from the convexity of q(BX ) that q(BX ) = BX/Y , and therefore Y is proximinal. (In their proof, our 8 MARTIN RMOUTIL Lemma 3.3 (i) is replaced by an essentially equivalent theorem of Garkavi, which is not formulated in terms of quotient spaces.) The next proposition can be regarded as a generalization of Corollary 3.4 (a). This can be seen if we reformulate the corollary as follows: Let X be a Banach space and x∗ ∈ X ∗. Then the hyperplane Ker(x∗) is proxim- inal whenever it contains a finite-codimensional proximinal subspace Y . Proposition 3.10. Let X be a normed linear space and let Y ⊂ Z ⊂ X be linear subspaces such that the codimension of Y in Z is finite. If Y is proximinal (in X), then Z is also proximinal. Proof. First assume that the codimension of Y in Z is 1. Fix a functional x∗ ∈ SX such that Y = Z ∩ Ker(x∗) and any vector v ∈ Z \ Y ; then it is easy to see that Z = [ c∈R (cv + Y ). Now, pick an arbitrary x ∈ X; we aim to find a point y ∈ x + Z with kyk = dist(0, x + Z). To that end, let us consider the function ϕ : R → R defined at c ∈ R by the formula ϕ(c) := dist(0, x + cv + Y ). Assume, for the moment, that we have found a global minimum of ϕ at a point c; then we can use the proximinality of Y to find a point y ∈ x + cv + Y with minimal norm. Clearly y has the minimal norm also in x + Z. Since ϕ is easily seen to be continuous (even kvk-Lipschitz), it is enough to show that there is a compact set I ⊂ R with ϕ(R \ I) ⊂ (ϕ(0), ∞). We claim that it suffices to set I := {c ∈ R; x∗(x + cv) ≤ kxk}. Indeed, if c ∈ R \ I, then x∗(x + cv) > kxk. Since Y ⊂ Ker(x∗), the functional x∗ ∈ SX ∗ is constant on any coset of the form z + Y , z ∈ X. It follows that ϕ(c) = inf{kx + cv + yk; y ∈ Y } ≥ inf{x∗(x + cv + y); y ∈ Y } = x∗(x + cv) > kxk ≥ ϕ(0); In particular, 0 ∈ I. The boundedness of I follows from the fact that v /∈ Ker(x∗). Now assume that we have proved the assertion of the Proposition in case the codimension of Y in Z is equal to n ∈ N, and suppose we have closed subspaces Y and Z of X such that Y ⊂ Z with codimension n+1. This means that there are some x∗ 1, . . . , x∗ i ). Then the codimension of Y in Z1 is 1, so Z1 is proximinal, and the codimension of Z1 in Z is n, so Z is proximinal. (cid:3) n+1 ∈ X ∗ such that Y = Z ∩ Tn+1 i ). Set Z1 := Z ∩ Tn i=1 Ker(x∗ i=1 Ker(x∗ Remark 3.11. Following W. Pollul [19], by Z ⊂p Y we mean that Z is a proxim- inal subspace of the normed linear space Y ; it is then natural to ask under what conditions is the relation ⊂p transitive (in which case the ambient space is said to be a P -space). Indumathi [13, Corollary 5] gives a characterization: A normed linear space X is a P -space if and only if it is an R(1) space and NA(X) is orthogonally linear (for definition see the same article). In particular, X is a P -space if it is an R(1) space and NA(X) is a vector space. Another related question is whether proximinality of subspaces is preserved un- der intersections. This is not always the case either; if a Banach space X has the NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 9 property that the intersection of arbitrary two finite-codimensional proximinal sub- spaces is proximinal, X is said to be lattice-proximinal. A simple characterization was given by Godefroy and Indumathi in [11]: X is lattice-proximinal if and only if it is R(1) and NA(X) is a vector space. An example of such a space is c0; we already know that it is R(1) (Theorem GI), and it is easy to see that norm-attaining functionals on c0 are exactly finitely supported elements of ℓ1. 4. NA(X) need not be 2-lineable The following renorming of c0 is due to C.J. Read [20] and it is the first Banach space known to have no proximinal subspaces of finite codimension strictly larger than 1 (Theorem R). In this section we will show that the set of norm-attaining functionals on this space is not 2-lineable, solving Problem 1.1. Definition 1 (C.J. Read). Let c00(Q) be the terminating sequences with rational coefficients, and let (un)∞ n=1 be a sequence of elements of c00(Q) which lists every element infinitely many times. Further, let (an)∞ fying for each n ∈ N, n=1 be a strictly increasing sequence of positive integers satis- an > max supp un, and an ≥ kunk1. The norm · on c0 is defined as follows: x = kxk∞ + ∞ X n=1 2−a2 nhx, un − eani. Here kxk∞ = supn xn is the usual norm on c0, ej's are the vectors from the usual cannonical basis, and the duality h·, ·i is the hc0, l1i duality. Remark 4.1. Proposition 3.10 implies that if a normed linear space X has no proximinal subspace of codimension 2, then it has no proximinal subspace of any finite codimension larger than 2 either. Hence, in the proof of C.J. Read it is enough to show that the space (c0, ·) constructed there has no proximinal subspace of codimension 2 (he proves it at once for all codimensions, slightly complicating the proof). The following theorem shows that the construction of Read does, in fact, answer Problem 1.1. Theorem 4.2. Let X be the space (c0, ·) from Definition 1. Then NA(X) con- tains no linear subspaces of dimension larger than one. Proof. We shall denote by · the norm on ℓ1 which is dual to · on c0. Assume that Y $ X is a closed subspace of finite codimension and such that Y ⊥ ⊂ NA(X); we aim to prove that the codimension of Y in X is equal to 1. It follows from Theorem R and Lemma 3.3 that it is enough to prove the strict convexity of X/Y ; we shall do so by contradiction. To that end, we can (by Fact 3.2) assume that we have a functional ϕ ∈ S(X/Y )∗ such that att(ϕ) contains more than one point; set x∗ := ϕ ◦ q. Then x∗ ∈ Y ⊥ ⊂ NA(X), and so there is x ∈ att(x∗). Fact 3.2 gives that x∗ = kϕk = 1 and q(x) ∈ att(ϕ). Let us take any point u ∈ att(ϕ)\{q(x)} and choose any sequence (xn)∞ X such that limn→∞ xn = 1 (recall that kuk = 1). Then (xn)∞ n=1 ⊂ u ⊂ n=1 is a bounded 10 MARTIN RMOUTIL sequence and we can without loss of generality assume that there is x ∈ ℓ∞ such that xn Now, w∗ → x in ℓ∞. 2 = ϕ(q(x) + q(xn)) = x∗(x + xn) ≤ x + xn ≤ x + xn n→∞−→ 2, whence x + xn − x + xn −→ 0. Since the norm · is defined by a series of pseudonorms (which satisfy the triangle inequality), we in turn obtain for each k ∈ N: hx, uk − eak i + hxn, uk − eak i − hx + xn, uk − eak i n→∞−→ 0, so, since (xn) converges to x in weak∗ topology, we have that hx, uk − eak i + hx, uk − eak i − hx + x, uk − eak i = 0. (2) Claim: There is λ ≥ 0 such that x = λx. In particular, x ∈ X and xn Assume this is not the case. Then there is z∗ ∈ c00(Q) ⊂ (ℓ1, ·) = X ∗ such w→ x. that z∗(x) < 0 < x(z∗), z∗(x) > kxk∞ and x(z∗) > kxk∞. Indeed, by the Hahn-Banach theorem we can find z∗ x and the weak∗-closed convex set [0, ∞) · x, and z∗ weak∗-closed convex set [0, ∞) · x with 1 ∈ ℓ1 separating the point 2 ∈ ℓ1 separating x and the z∗ 1 (x) < 0 ≤ x(z∗ 1 ), and z∗ 2(x) ≤ 0 < x(z∗ 2). Setting z∗ 3 := z∗ 1 + z∗ 2, one can readily see that z∗ 3(x) < 0 < x(z∗ 3 ). Using the density of c00(Q) in ℓ1 we find a functional z∗ ∈ c00(Q) ⊂ ℓ1 so close to z∗ 3 that, again, z∗(x) < 0 < x(z∗). Without loss of generality we can now assume that z∗(x) > kxk∞ and x(z∗) > kxk∞ because in the opposite case we would simply multiply z∗ by a sufficiently large rational constant. Now, take k ∈ N such that uk = z∗ (see Definition 1); Then hx, uk − eak i = z∗(x) − xak ≤ z∗(x) + kxk∞ < 0 hx, uk − eak i = x(z∗) − xak ≥ x(z∗) − kxk∞ > 0. and But on the other hand it follows from (2) that for each natural k we have hx, uk − eak i · hx, uk − eak i ≥ 0, a contradiction which proves the Claim. Finally, find a functional ψ ∈ (X/Y )∗ such that ψ(q(x)) < 0 < ψ(u) (we can 3 above in this proof). But then, setting do so e.g. using the same idea as with z∗ y∗ := ψ ◦ q, we have that y∗(x) = λy∗(x) = λψ(q(x)) ≤ 0 < ψ(u) = y∗(xn) which contradicts the fact that xn w→ x, concluding the proof. (cid:3) NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 11 5. Non-R(1) renormings The main result of the present section is Theorem 5.6, which shows that the implication (B) =⇒ (A) (see the Introduction) can fail in all non-reflexive spaces if we take a suitable renorming; the purpose of all the preceding lemmas is to prove this theorem. At the end of this section we also prove a slightly stronger version of the key Lemma 5.5 which holds for the space ℓ1. Lemma 5.1. Let x1, x2, . . . , xn ∈ Rd be such that C := conv{x1, x2, . . . , xn} has nonempty interior. Let x = Pn i=1 λi = 1 and λi > 0 for each i. Then x ∈ Int(C). i=1 λixi with Pn Proof. Assume that x ∈ ∂C and take any non-zero functional ϕ ∈ (Rd)∗ such that supC ϕ = ϕ(x) (it exists due to the Hahn-Banach Theorem). The hyperplane H := ϕ−1(ϕ(x)) is nowhere dense, so there is a j ∈ {1, . . . , n} such that xj ∈ C \ H (otherwise C ⊂ H which is impossible). But λj > 0, so λj ϕ(xj ) < λjϕ(x), whence ϕ(x) = ϕ(cid:16) n X i=1 λixi(cid:17) = X i6=j λiϕ(xi) + λj ϕ(xj ) < X i6=j λiϕ(x) + λj ϕ(x) = ϕ(x), which is a contradiction. (cid:3) Lemma 5.2. If C is a convex set in a normed linear space X, y ∈ Int(C) and z ∈ C, then λy + (1 − λ)z ∈ Int(C) for any λ ∈ (0, 1). Proof. If C, y, z and λ are as above, we can find r > 0 such that U (y, r) ⊂ C and a point x ∈ C ∩ U (z, λr 1−λ ). Then the ball U := λU (y, r) + (1 − λ)x = U (λy + (1 − λ)x, λr) is contained in C as it consists of convex combinations of points of U (y, r) and x. But kλy + (1 − λ)z − (λy + (1 − λ)x)k = (1 − λ)kz − xk < λr, and so λy + (1 − λ)z ∈ U . (cid:3) Proposition 5.3. Let A ⊂ Rd be bounded. Then convσ(A) = conv(A). Proof. Take any σ-convex combination x = P∞ i=1 λixi of elements of A. If there are only finitely many non-zero λi's, then x ∈ conv(A) trivially. So we can clearly assume that λi 6= 0 for every i, and we aim to prove that x ∈ conv{xi; i ∈ N} ⊂ conv(A). Now, any convex set in Rd is either contained in a hyperplane, or has a nonempty interior. Thus, setting C := conv{xi; i ∈ N}, we can assume that Int(C) 6= ∅. Take N ∈ N so large that the interior of CN := conv{xi; i = 1, . . . , N } is nonempty and set λ := PN i=1 λi; then by Lemma 5.1, y := 1 i=1 λixi ∈ Int(CN ) ⊂ Int(C). λ PN The point z := 1 1−λ P∞ λ)z, the result follows from Lemma 5.2. i=N +1 λixi is clearly an element of C. Since x = λy + (1 − (cid:3) Lemma 5.4. Let X be a Banach space and let for each j, k ∈ N, f j k=1 ⊂ ℓ1 and (bk)∞ Let (ak)∞ limk→∞ kak − bkk1 = 0. Then k ∈ BX . k=1 ⊂ ℓ1 be two sequences of elements of ℓ1 such that lim k→∞ ∞ X j=1 aj kf j k = lim k→∞ ∞ X j=1 bj kf j k whenever one of the limits exists. 12 MARTIN RMOUTIL Proof. Since X is a Banach space, all the series in the statement, as well as the estimate below, converge because they converge absolutely. Hence, the following suffices to conclude the proof: (cid:13)(cid:13)(cid:13) X ∞ j=1 ≤ ∞ X j=1 aj kf j k − ∞ X j=1 k − bj (cid:12)(cid:12)aj k(cid:12)(cid:12) · (cid:13)(cid:13)f j ∞ j=1 (cid:0)aj X bj kf j = (cid:13)(cid:13)(cid:13) k(cid:13)(cid:13)(cid:13) k − bj (cid:12)(cid:12)aj X k(cid:13)(cid:13) ≤ j=1 ∞ k − bj k(cid:13)(cid:13)(cid:13) k(cid:1)f j k(cid:12)(cid:12) = kak − bkk1 −→ 0. (cid:3) The following lemma is the core result of the present section showing how to construct certain renormings of non-reflexive Banach spaces. The proof is rela- tively complicated from the technical point of view, but except for its use of the James's theorem [14] (stating that on a non-reflexive Banach space there always is a functional which does not attain its norm on the closed unit ball) it is elementary. The proof is divided into 20 paragraphs to make it easier to read and also to relate it to the proof of Proposition 5.7 which follows the same scheme. The main idea of the proof appears already in the second paragraph where we define the renorming by choosing a new unit ball for the space. The rest of the proof is to show that the image of this new unit ball via the quotient mapping is exactly what we want it to be (i.e. the set A). The difficult part is to show that the image contains no extra points, and that follows from the choice of renorming which uses the non- norm-attaining functional and the subsequent existence of a "very non-convergent" bounded sequence of points (a norming sequence for the functional contained in the unit sphere). Lemma 5.5. Let X be a non-reflexive Banach space, d ∈ N, Y ⊂ X be a closed subspace of codimension d in X, and q : X → X/Y be the quotient mapping. Let k · k be an equivalent norm on X/Y and let A ⊂ X/Y be a symmetric convex Fσ-set satisfying U(X/Y,k·k) ⊂ A ⊂ B(X/Y,k·k) for which there exists v ∈ ∂A and r ∈ (0, 1) such that A \ (cid:0)Uk·k(v, r) ∪ Uk·k(−v, r)(cid:1) is closed. Then X admits an equivalent norm · such that q(B(X,·)) = A. Proof. §1: Since Y is a closed subspace of finite codimension d in X, it is comple- mented in X, its (topological) complement is isomorphic to Rd as well as to X/Y , and X ∼= Y ⊕∞ X/Y (see e.g. [6, Section 4.1]). It follows that X ∼= Y ⊕∞ Rd. Note also that Y is non-reflexive as in the opposite case X would be reflexive, be- ing isomorphic to the direct sum of two reflexive spaces, which is easily seen to be reflexive. Our goal is to find a certain equivalent norm on X; we can therefore assume that X = Y ⊕∞ Rd and consider the projection to the second coordinate q : X → Rd instead of the quotient mapping X → X/Y . This is, indeed, correct as the topological and linear structure remains the same, so we can also assume without loss of generality that A ⊂ Rd and that k · k is a norm on Rd (and it is the one satisfying B(Rd,k·k) = A). To avoid any confusion, let us emphasize that we shall use no other norm on Rd in this proof. §2: Now, denote F0 := conv(cid:0)A \ (U (v, r) ∪ U (−v, r))(cid:1); then F0 is a compact subset of A. Clearly, the set A ∩ U (v, r) is of the type Fσ, so there are closed sets Ej, j ≥ 1, such that A ∩ U (v, r) = S∞ j=1 Ej. From the convexity of A ∩ U (v, r) it NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 13 follows that A ∩ U (v, r) = S∞ each j. j=1 Fj where Fj := conv(Ej ); then Fj is also closed for To define an equivalent norm with the desired properties, we need to use the fact that Y is non-reflexive. The James's theorem implies that there exists y∗ ∈ Y ∗ \ NA(Y ), and we can fix a sequence (cid:0)zj(cid:1)∞ j=0 ⊂ SY such that y∗(cid:0)zj(cid:1) ր ky∗k as j → ∞. The following set B clearly is the closed unit ball for an equivalent norm · on X. B := conv(cid:16)BY × 1 2 A ∪ ∞ [ j=0 (cid:0)(cid:8)zj(cid:9) × Fj ∪ (cid:8) − zj(cid:9) × (−Fj)(cid:1)(cid:17) Indeed, B is closed absolutely convex with BY × 1 2 A ⊂ B ⊂ BY × A. §3: We will be done when we prove that q(B) = A. The inclusion q(B) ⊃ A j=0(Fj ∪ (−Fj )), and obviously is easy to see; indeed, A is symmetric so A = S∞ S∞ j=0(Fj ∪ (−Fj)) ⊂ q(B). It remains to prove that q(B) ⊂ A. Note that the sequence (zj)∞ j=0 does not converge; our task would therefore be easy if the definition of B only involved closure instead of the closed convex hull. If that were the case, our argument would be based on the simple observation that a sequence of the form (zjk , fk)∞ k=1 with fk ∈ Fjk converges only if the sequence of indices (jk)∞ k=1 is eventually constant with limk→∞ jk = j∞, whence limk→∞ fk ∈ Fj∞ ⊂ A. The rest of the proof deals with the technical difficulties arising from the presence of convex combinations. §4: To that end, let us fix an arbitrary point u ∈ q(B) and an x ∈ Y such that (x, u) ∈ B. Then (x, u) can be written as the limit of a sequence of convex combinations: (x, u) = lim k→∞(cid:16)tk(cid:16)xk, 1 2 vk(cid:17) + nk X j=0 (cid:0)αj k(cid:0)zj, f j k(cid:1) − βj k(cid:0)zj, gj k(cid:1)(cid:1)(cid:17) = lim k→∞(cid:16)tkxk + nk X j=0 kzj − βj (cid:0)αj kzj(cid:1), 1 2 tkvk + nk X j=0 kf j k − βj (cid:0)αj k(cid:1)(cid:17) kgj (3) k + βj j=0 (cid:0)αj where nk ∈ N, tk + Pnk k(cid:1) = 1 with all the coefficients non-negative, vk ∈ A and f j k ∈ Fj for each k ∈ N and j ∈ N ∪ {0}. Note that all the sets Fj , as well as A, are convex, and so in the convex combinations it is enough to take only one element from each of them instead of a convex combination of finitely many elements. k, gj In paragraphs §5 to §14 of this proof we essentially make a series of simplifying assumptions which will be shown to cause no loss of generality; in this manner we shall reduce the matter enough to be able to formulate the final argument in the last six paragraphs. §5: First, let us formally make all the sums in (3) infinite series by defining for each k ∈ N and each j > nk, αj k := 0, βj k := 0, f j k := f 1 k and gj k := g1 k. Thus we now have (x, u) = lim k→∞(cid:16)tkxk + ∞ X j=0 kzj − βj (cid:0)αj kzj(cid:1), 1 2 tkvk + ∞ X j=0 kf j k − βj (cid:0)αj k(cid:1)(cid:17) kgj (4) (5) 14 MARTIN RMOUTIL where each of the infinite series has only finitely many non-zero summands (for each k ∈ N). In the following we shall often use this fact without recalling it. §6: By passing to a subsequence (using a diagonal argument) we can assume that all the following limits exist for all j ∈ N (recall that all the sequences (f j and (gj k=1 are bounded in Rd): k)∞ k )∞ k=1 αj := lim k→∞ αj k, βj := lim k→∞ f j k βj k, t := lim k→∞ and gj := lim k→∞ f j := lim k→∞ tk, gj k. (6) §7: We now distinguish two cases: Either t = 0 or t > 0. Assume first t > 0; then we shall conclude that u ∈ A as follows. We can find k0 ∈ N such that tk > t 2 for each k ≥ k0. Hence, for each k ≥ k0 we have tkvk + ∞ X j=0 kf j (cid:0)αj tk 2 + ∞ X j=0 k + βj (cid:0)αj k − βj kgj k(cid:1)(cid:13)(cid:13)(cid:13) k(cid:1) = tk + tk 2 vk(cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13) (cid:0)αj X ∞ j=0 + ∞ X j=0 kf j (cid:13)(cid:13)αj k(cid:13)(cid:13) + ∞ X j=0 kgj (cid:13)(cid:13)βj k(cid:13)(cid:13) k + βj k(cid:1) − tk 2 = 1 − tk 2 < 1 − t 4 . 1 2 (cid:13)(cid:13)(cid:13) ≤ Hence, u = lim k→∞(cid:16) 1 2 tkvk + ∞ X j=0 kf j k − βj (cid:0)αj k(cid:1)(cid:17) ∈ (cid:16)1 − kgj t 4(cid:17)B(Rd,k·k) ⊂ U(Rd,k·k) ⊂ A. §8: Till the end of the proof, we shall therefore investigate the case when t = limk→∞ tk = 0. It follows from (5) that u = lim k→∞ (1 − tk) ∞ X j=1 (cid:16) αj k 1 − tk f j k − βj k 1 − tk k(cid:17) = lim gj k→∞ ∞ X j=1 (cid:16) αj k 1 − tk f j k − βj k 1 − tk k(cid:17), gj and similarly for the first coordinate x; the sum on the right is a convex combination for each k ∈ N. Hence, we can assume without loss of generality that tk = 0 for each k, so we have ∞ X j=0 k + βj (cid:0)αj k(cid:1) = 1 for each k, and (x, u) = lim k→∞(cid:16) ∞ X j=0 kzj − βj (cid:0)αj kzj(cid:1), ∞ X j=0 kf j k − βj (cid:0)αj k(cid:1)(cid:17). kgj §9: Next assumption we want to make is α0 k = β0 k = 0 for all k ∈ N. In the present paragraph we show that this causes no loss of generality: Let us first consider the case when α0 + β0 = 1; then lim k→∞ ∞ X j=1 k + βj (cid:0)αj k(cid:1) = lim k→∞(cid:0)1 − (cid:0)α0 k + β0 k(cid:1)(cid:1) = 1 − (cid:0)α0 + β0(cid:1) = 0. It is now easy to see (e.g. using Lemma 5.4) that limk→∞ P∞ whence (by (5) and (6)) u = α0f 0 + β0g0 ∈ F0 ⊂ A. j=1 (cid:0)αj kf j k − βj kgj k(cid:1) = 0, NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 15 Assume now that α0 + β0 < 1. Clearly we can also assume that for each k ∈ N we have α0 k + β0 k < 1, and consider the limit of convex combinations u := lim k→∞ ∞ X j=1 (cid:16) αj k k + β0 1 − (cid:0)α0 k(cid:1) f j k − βj k k + β0 1 − (cid:0)α0 k(cid:1) k(cid:17); gj (7) it is easy to see that the limit exists. Then we have u = (cid:0)α0 + β0(cid:1)(cid:16) α0 α0 + β0 f 0 − + (cid:0)1 − (cid:0)α0 + β0(cid:1)(cid:1) lim = (cid:0)α0 + β0(cid:1)(cid:16) α0 α0 + β0 f 0 − k→∞ β0 α0 + β0 g0(cid:17) αj (cid:16) X k ∞ j=1 1 − (cid:0)α0 + β0(cid:1) β0 α0 + β0 g0(cid:17) + (cid:0)1 − (cid:0)α0 + β0(cid:1)(cid:1)u, f j k − βj k 1 − (cid:0)α0 + β0(cid:1) k(cid:17) gj (8) where the last equality follows from Lemma 5.4 whose assumption is verified as follows: j=1 αj k (cid:17)∞ (cid:13)(cid:13)(cid:13) (cid:16) = (cid:12)(cid:12)(cid:12) 1 − (cid:0)α0 + β0(cid:1) 1 − (cid:0)α0 + β0(cid:1) − αj (cid:17)∞ j=1(cid:13)(cid:13)(cid:13)1 − (cid:16) k k + β0 1 − (cid:0)α0 k(cid:1) 1 (cid:12)(cid:12)(cid:12) k(cid:1)∞ · (cid:13)(cid:13)(cid:0)αj j=1(cid:13)(cid:13)1 −→ 0, k + β0 1 − (cid:0)α0 k(cid:1) k instead of αj k. 1 and similarly with βj k → ∞, In (8) we expressed u as a convex combination of an element of F0 ⊂ A and u; therefore, to see that u ∈ A, it is sufficient to prove u ∈ A. Of course, u is in q(B) as it is the projection of (x, u) ∈ B where x := lim k→∞ ∞ X j=1 (cid:16) αj k k + β0 1 − (cid:0)α0 k(cid:1) zj − βj k k + β0 1 − (cid:0)α0 k(cid:1) zj(cid:17). To avoid introducing more notation, we henceforth assume that α0 k = β0 k = 0 for all k ∈ N, so we now have (x, u) = lim k→∞(cid:16) ∞ X j=1 kzj − βj (cid:0)αj kzj(cid:1), ∞ X j=1 kf j k − βj (cid:0)αj k(cid:1)(cid:17). kgj §10: Next we distinguish two subcases depending on the value of ν := minn lim k→∞ ∞ X j=1 αj k , lim k→∞ ∞ X j=1 ko; βj by passing to a subsequence we can guarantee that both limits exist. §11: Let us first consider the case when ν > 0; we claim that then u ∈ A. To show that, set for k ∈ N, αk := ∞ X j=1 αj k, βk := ∞ X j=1 βj k, fk := 1 αk ∞ X j=1 αj kf j k and gk := 1 βk ∞ X j=1 kgj βj k, and find k0 ∈ N such that for each k ≥ k0, αk, βk ∈ (cid:0) ν k ≥ k0, let us assume for example αk ≥ βk; the opposite case is analogous. 2(cid:1). Now, for a given 2 , 1 − ν 16 MARTIN RMOUTIL Recall that for j ≥ 1, f j k , gj k ∈ Fj ⊂ U (v, r) which implies that fk, gk ∈ U (v, r), and thus kfk − gkk < 2r. We now have ∞ X j=1 (cid:13)(cid:13)(cid:13) kf j k − βj kgj (cid:0)αj k(cid:1)(cid:13)(cid:13)(cid:13) = kαkfk − βkgkk = k(αk − βk)fk + βk(fk − gk)k ≤ αk − βk + 2rβk = 1 − 2βk + 2rβk = 1 − 2βk(1 − r) < 1 − ν(1 − r) < 1. This is true for any k ≥ k0, and consequently u = lim k→∞ ∞ X j=1 kf j k − βj kgj (cid:0)αj k(cid:1) ∈ (1 − ν(1 − r)) · B(Rd,k·k) ⊂ U(Rd,k·k) ⊂ A. §12: We now turn to the case ν = 0. The first step is to observe that we can assume without loss of generality that either αk = 0 for all k, or βk = 0 for all k. Setting ck := (cid:0)α1 k, α2 k, . . .(cid:1) and dk := (cid:0)β1 k, β2 k, . . .(cid:1), we see that either kckk1 = αk → 0 or kdkk1 = βk → 0. Let us examine only the second case as the first one is analogous. Then we have (e.g. by Lemma 5.4) lim k→∞ ∞ X j=1 βj kgj k = 0, and therefore (recall that αk + βk = 1, whence αk → 1) lim k→∞ ∞ X j=1 αj k αk f j k = lim k→∞ ∞ X j=1 αj kf j k = lim k→∞ ∞ X j=1 kf j k − βj kgj (cid:0)αj k(cid:1) = u, from where the observation follows. From now on, we assume that βj k = 0 for all j, k ∈ N, so (x, u) = lim k→∞(cid:16) ∞ X j=1 αj kzj , ∞ X j=1 αj kf j k(cid:17). (9) §13: Again we need to distinguish two further subcases: This time depending on the value of λ := ∞ X j=1 αj . §14: It is easy to see that λ ∈ [0, 1]; the easier situation is when λ = 1: In this case we need to observe that the limits in the definition of u commute, that is: u = lim k→∞ ∞ X j=1 kf j αj k = ∞ X j=1 αjf j. (10) Indeed, let ε > 0 be given and let us find j0 ∈ N such that P∞ and k0 ∈ N such that for each k ≥ k0 we have (cid:13)(cid:13)Pj0 (cid:12)(cid:12)Pj0 j=j0+1 αj < ε 4 , 4 and 4 . Then for each k ≥ k0 we also obtain (using the assumption kf j k − αj f j(cid:1)(cid:13)(cid:13) < ε k − αj(cid:1)(cid:12)(cid:12) < ε j=1 (cid:0)αj j=1 (cid:0)αj NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 17 λ = 1) that (cid:12)(cid:12)P∞ j=j0+1 (cid:0)αj αj kf j X k − X ∞ ∞ j=1 j=1 (cid:13)(cid:13)(cid:13) 4 , and so k − αj(cid:1)(cid:12)(cid:12) < ε αjf j(cid:13)(cid:13)(cid:13) k − αjf j(cid:1)(cid:13)(cid:13)(cid:13) X k + αj ∞ j0 X j=1 ≤ (cid:13)(cid:13)(cid:13) < ε 4 + kf j (cid:0)αj ∞ X j=j0+1 j=j0+1 ∞ X j=j0+1 + (cid:13)(cid:13)(cid:13) αj < ε 4 + (cid:16) kf j αj k(cid:13)(cid:13)(cid:13) + (cid:13)(cid:13)(cid:13) X ∞ j=j0+1 ∞ X j=j0+1 αjf j(cid:13)(cid:13)(cid:13) 4(cid:17) + ε 4 ε αj + < ε. This proves the second equality in (10), which implies that u ∈ convσ(A) (note that f j, gj ∈ Fj ⊂ A for every j ∈ N) and it follows by Proposition 5.3 that u ∈ conv(A) = A. §15: Finally, assume that λ ∈ [0, 1); we will show that this assumption leads to a contradiction. Set xλ := ∞ X j=1 αjzj and x(1−λ) := x − xλ = lim k→∞ ∞ X j=1 (cid:0)αj k − αj(cid:1)zj. Now we aim to prove that x(1−λ) of zj. But it need not suffice to simply divide all the coefficients αj as some of them could be negative. Hence, for any a ∈ R we define 1−λ is the limit of a sequence of convex combinations k − αj by 1 − λ a+ := max{a, 0}, a− := (−a)+, and γj k := (cid:0)αj k − αj(cid:1)+ , and we claim that x(1−λ) = lim k→∞ ∞ X j=1 γj kzj. §16: By Lemma 5.4 it is enough to prove that j=1 − (cid:0)γj k − αj(cid:1)∞ lim k→∞(cid:13)(cid:13)(cid:0)αj k(cid:1)∞ j=1(cid:13)(cid:13)1 = 0. ≤ αj as αj j=N +1 αj < ε 2 2 . Note k ≥ 0. Then for each k ≥ k0 we have k − αj(cid:1)− k − αj(cid:1)∞ Let an arbitrary ε > 0 be given. We can find N ∈ N such that P∞ and k0 ∈ N such that for each natural k ≥ k0 we have PN that (cid:0)αj k(cid:1)∞ (cid:13)(cid:13)(cid:0)αj j=1(cid:13)(cid:13)1 = §17: Now, set σk := P∞ j=1 γj j=1 (cid:0)αj we obtain k −αj(cid:1) = 1−λ also implies that limk→∞ σk = 1−λ. Thus, setting γj k; then the last part together with the fact that k := γ j (cid:12)(cid:12)αj k − αj(cid:12)(cid:12) + k − αj(cid:12)(cid:12) < ε k − αj(cid:1)− j=1 − (cid:0)γj j=1 (cid:12)(cid:12)αj P∞ αj < ε. (cid:0)αj X j=1 X j=1 j=N +1 ∞ X k σk ≤ ∞ N x(1−λ) = lim k→∞ 1 − λ σk ∞ X j=1 γj kzj = (1 − λ) lim k→∞ γj kzj, X j=1 ∞ and from the definition it immediately follows that ∞ X j=1 γj k = 1 for each k, and lim k→∞ γj k = 0 for each j. 18 MARTIN RMOUTIL §18: From these last facts we can readily see that x(1−λ) 1 − λ = lim k→∞ ∞ X j=1 γj kzj ∈ ∞ \ N =1 conv(cid:8)zj; j > N(cid:9). (11) Indeed, pick arbitrary N ∈ N and ε > 0, and find k0 such that for each natural k ≥ k0, γk := N X j=1 γj k < ε 2 . Then for each k ≥ k0, dist(cid:16) kzj, conv(cid:8)zj; j > N(cid:9)(cid:17) γj ∞ ∞ j=1 j=1 X X ≤ (cid:13)(cid:13)(cid:13) ≤ (cid:13)(cid:13)(cid:13) X ≤ γk + (cid:12)(cid:12)(cid:12) j=1 N γj kzj − 1 1 − γk ∞ X j=N +1 γj kzj(cid:13)(cid:13)(cid:13) + (cid:13)(cid:13)(cid:13) 1 1 − 1 − γk (cid:16)1 − (cid:12)(cid:12)(cid:12) γj kzj(cid:13)(cid:13)(cid:13) (cid:17) X ∞ j=N +1 1 1 − γk γj kzj(cid:13)(cid:13)(cid:13) (1 − γk) = γk + γk < ε. This estimate yields that x(1−λ) 1−λ ∈ conv(cid:8)zj; j > N(cid:9). §19: But this is impossible since the intersection in (11) is empty as we will show below. On the other hand, the limit in the definition (9) of x converges by the assumption and since xλ is a well-defined element of the Banach space Y , the limit defining x(1−λ) necessarily converges as well. This is a contradiction showing that the case λ ∈ [0, 1) does not occur; since in all the other cases we have already shown that u ∈ A, the proof will be complete, once we make the observation that C := ∞ \ N =1 conv(cid:8)zj; j > N(cid:9) = ∅. §20: Recall that (cid:0)zj(cid:1)∞ j=1 ⊂ SY was chosen to be a norming sequence of the non-norm-attaining functional y∗. If there existed a point z ∈ C, then z ∈ BY and y∗(z) > y∗(cid:0)zN(cid:1) for each N ∈ N. Hence, y∗(z) = ky∗k which is a contradiction, and the proof is complete. (cid:3) As a corollary we obtain the following result. Theorem 5.6. Let X be a non-reflexive Banach space and Y ⊂ X a closed subspace of finite codimension d ≥ 2. Then X admits an equivalent norm such that Y ⊥ ⊂ NA(X) and Y is not proximinal. Proof. As explained in §1 of the proof of Lemma 5.5, we can assume that X/Y is isomorphic to Rd where d is the codimension of Y in X. Let k · k2 be the Euclidean norm on Rd and let ei, i = 1, . . . , d, be the canonical basis in Rd. Put A := conv(cid:0)Bk·k2(e1, 1) ∪ Bk·k2(−e1, 1)(cid:1) \ {e1 + e2, −e1 − e2}. Clearly A is a symmetric convex Fσ set with nonempty interior satisfying the condi- tion from Lemma 5.5, so there is a renorming · of X such that q(BX ) = A where NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 19 q : X → Rd is the quotient mapping; then A is the closed unit ball in Rd ∼= X/Y with the corresponding norm. In the sequel, by X we mean (X, ·). Since q(BX ) = A 6= BX/Y , Lemma 3.3 yields that Y is not proximinal in X. To prove that Y ⊥ ⊂ NA(X) we will use Lemma 3.3 again. Take any ϕ ∈ (X/Y )∗; we are to show that it attains its norm at a point of q(BX ). Since dim(X/Y ) < ∞, there is a point u ∈ att(ϕ). If u /∈ {e1 + e2, −e1 − e2}, we are done. Suppose u = e1 + e2; of course, the case u = −e1 − e2 is symmetrical. Further, we can clearly assume kϕk = 1, so ϕ(u) = 1. Consider the hyperplane H := {(x1, . . . , xn) ∈ Rd; x2 = 1}; obviously H is a tangent hyperplane to A at the point e1+e2 (i.e. H∩Int(A) = ∅ and e1+e2 ∈ H∩A). But there is only one such hyperplane because A ⊃ Bk·k2(e1, 1) and e1 + e2 lies on the boundary of Bk·k2(e1, 1). It follows that H = ϕ−1(1). But e2 ∈ H ∩ q(BX ) which concludes the proof. (cid:3) Proposition 5.7. Let Y ⊂ ℓ1 be a closed subspace of codimension d ∈ N, and q : ℓ1 → ℓ1/Y be the quotient mapping. Let k · k be an equivalent norm on ℓ1/Y and let A ⊂ ℓ1/Y be a symmetric convex Fσ-set satisfying U(ℓ1/Y,k·k) ⊂ A ⊂ B(ℓ1/Y,k·k). Then there is an equivalent norm · on ℓ1 such that q(B(ℓ1,·)) = A. Proof. The proof is very similar to that of Lemma 5.5, and so we will do it more briefly, emphasizing the necessary changes in each paragraph of the proof. §1: Denote X := ℓ1. The subspace Y ⊂ X is isomorphic to Z := (cid:8)(cid:0)xj(cid:1)∞ j=1 ∈ ℓ1; x1 = x2 = · · · = xd = 0(cid:9) as all subspaces of the same finite codimension in X are isomorphic. But Z is obviously (isometrically) isomorphic to ℓ1, and so we can (similarly as in the proof of Lemma 5.5) assume that X = Y ⊕∞ Rd where Y = ℓ1. Throughout the proof we shall only consider the norm k · k on Rd for which B(Rd,k·k) = A. §2: In a similar (but simpler) way as in 5.5 we can find closed convex sets Fj (j ∈ N) of diameter less than 1 and such that A = S∞ of Y = ℓ1 and we define j=1 Fj. Now, instead of using the James's theorem, we take the canonical basis (ej)∞ j=1 B := conv(cid:16)BY × 1 2 A ∪ ∞ [ j=1 (cid:0)(cid:8)ej(cid:9) × Fj ∪ (cid:8) − ej(cid:9) × (−Fj)(cid:1)(cid:17). Again, B is the closed unit ball for an equivalent norm · on X. §3: We want to prove that q(B) = A; the inclusion q(B) ⊃ A follows immedi- ately from the definition of B and the fact that A = S∞ such that (x, u) ∈ B. As in 5.5, j=1 Fj. §4: To prove the converse inclusion, we again take arbitrary u ∈ q(B) and x ∈ Y (x, u) = lim k→∞(cid:16)tkxk + nk X j=1 kej − βj (cid:0)αj kej(cid:1), 1 2 nk tkvk + kf j X (cid:0)αj k − βj k(cid:1)(cid:17) kgj k , gj k ≥ 0, vk ∈ A and f j j=1 (12) §5, §6: Setting αj where nk ∈ N, tk + Pnk j=1 (cid:0)αj k, βj k, f j k(cid:1) = 1 with αj k for k ∈ N and j > nk as in (4), we can assume again that all the limits in (6) exist. As in 5.5, we now formally consider all the sums in (12) as infinite series. k + βj k and gj k ∈ Fj. k, βj §7, §8: In the same way as in 5.5 we prove that we can assume all tk's to be zero (otherwise u ∈ Int(A)). 20 MARTIN RMOUTIL §9: This step of the proof we skip entirely because in this case we do not work with the set F0. §10: Here we make one of the important changes by defining µ := lim k→∞ ∞ X j=1 min(cid:8)αj k, βj k(cid:9); of course, we can assume that this limit exists, and we distinguish two cases: µ > 0 and µ = 0. §11: Suppose µ > 0. For k ∈ N set ∞ uk := kgj k(cid:1); then u = limk→∞ uk. Further, for all j, k ∈ N set µj k := min(cid:8)αj and for a given j ∈ N assume that αj obtain the same estimate: k ≥ βj k − βj (cid:0)αj X kf j j=1 k(cid:9). Fix k ∈ N k; the opposite case is analogous and we k, βj k − βj kf j (cid:13)(cid:13)αj k(cid:1)f j k(cid:1) + µj k − µj k; here the inequality follows from the fact that f j k , gj Now, by taking the sum over all j ∈ N we obtain kgj k(cid:13)(cid:13) = (cid:13)(cid:13)(cid:0)αj ≤ (cid:0)αj = αj k − µj k − µj k + βj k k + µj k(cid:0)f j k − gj k(cid:1)(cid:13)(cid:13) k ∈ Fj ⊂ BRd and diam(Fj ) < 1. kukk ≤ ∞ X j=1 k − βj kf j (cid:13)(cid:13)αj kgj k(cid:13)(cid:13) ≤ ∞ X j=1 k + βj k − µj (cid:0)αj k(cid:1) = 1 − ∞ X j=1 µj k. It follows after passage to limit as k → ∞ that kuk ≤ 1−µ, and so u ∈ (1−µ)BRd ⊂ URd ⊂ A. §12: For the rest of the proof we shall assume that µ = 0. Similarly as in the proof of Lemma 5.5 we prove that there is no loss of generality in assuming that µk = 0 for each k. In other words, we can assume that for all j, k ∈ N we have αj k = 0 or βj §13: Recalling the notation in (4), we finally distinguish two cases depending k = 0. on the value of λ := ∞ X j=1 (cid:0)αj + βj(cid:1). §14: Again, λ ∈ [0, 1] and we first resolve the case λ = 1: We shall observe that the limit and the series in the definition of u commute, that is: u = lim k→∞ ∞ X j=1 kf j k − βj kgj (cid:0)αj k(cid:1) = ∞ X j=1 (cid:0)αj f j − βjgj(cid:1). (13) Indeed, let ε > 0 be given and let us find j0 ∈ N such that P∞ j=j0+1 (cid:0)αj + βj(cid:1) < ε 4 , and k0 ∈ N such that for each k ≥ k0 we have (cid:13)(cid:13)Pj0 kgj k − αj f j − (cid:0)βj j=1 (cid:0)αj kf j k − βjgj(cid:1)(cid:1)(cid:13)(cid:13) < ε j=1 (cid:0)αj + βj(cid:1)(cid:12)(cid:12) < ε 4 . By the same calculation (cid:0)αj f j − βjgj(cid:1)(cid:13)(cid:13)(cid:13) X 4 and (cid:12)(cid:12)Pj0 (cid:13)(cid:13)(cid:13) X as in the proof of 5.5 we obtain that for each k ≥ k0, k(cid:1) −Pj0 j=1 (cid:0)αj k − βj k + βj k(cid:1) − (cid:0)αj kf j kgj < ε. j=1 j=1 ∞ ∞ NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 21 This proves (13), so u ∈ convσ(A) = conv(A) = A by Proposition 5.3 and the convexity of A. §15: Aiming for a contradiction, suppose that λ ∈ [0, 1) and set xλ := ∞ X j=1 (cid:0)αj − βj(cid:1)ej and x(1−λ) := x − xλ = lim k→∞ ∞ X j=1 (cid:0)αj k − αj − (cid:0)βj k − βj(cid:1)(cid:1)ej. Note that the last limit must exist as it is the difference of an existing limit and xλ. Define γj k := (cid:0)αj and we now claim that k − αj(cid:1)+ and δj k := (cid:0)βj k − βj(cid:1)+ , x(1−λ) = lim k→∞ ∞ X j=1 k − δj (cid:0)γj k(cid:1)ej. §16: By Lemma 5.4 it suffices to prove that This can be done in the same way as in the proof of 5.5. lim k − αj(cid:1)∞ j=1 − (cid:0)γj k→∞(cid:13)(cid:13)(cid:0)αj §17: Now, set σk := P∞ Further, define k(cid:1)∞ j=1(cid:13)(cid:13)1 = 0 j=1 (cid:0)γj γj k σk γj k := and lim j=1 − (cid:0)δj k − βj(cid:1)∞ k→∞(cid:13)(cid:13)(cid:0)βj k(cid:1)∞ j=1(cid:13)(cid:13)1 = 0. k(cid:1) and observe that limk→∞ σk = 1 − λ. k + δj and δj k := δj k σk ; then we have: (i) x(1−λ) = (1 − λ) limk→∞ P∞ j=1 (cid:0)γj (ii) for each k ∈ N, P∞ k + δj j=1 (cid:0)γj k(cid:1) = 1; k = 0 or δj (iii) for all j, k ∈ N, either γj k = 0 and limk→∞ δj (iv) and for each j ∈ N, limk→∞ γj §18, §19, §20: Now, on one hand we can readily see from (i), (ii) and (iii) that k(cid:1)ej; k − δj k = 0; k = 0. kx(1−λ)k = (1 − λ) lim k→∞ ∞ X j=1 k + δj (cid:0)γj k(cid:1) = (1 − λ) lim k→∞ 1 = 1 − λ > 0. On the other hand, (i) and (iv) yield that x(1−λ) = 0. Indeed, (i) says that x(1−λ) is the norm-limit of a certain sequence which by (iv) converges pointwise to 0. (cid:3) Remark 5.8. (a) Let X be a Banach space, Y ⊂ X be a closed subspace and q : X → X/Y be the quotient mapping. Since any functional y∗ ∈ Y ⊥ can be expressed as y∗ = ϕ ◦ q for some ϕ := (X/Y )∗, one can notice that in the proof of Lemma 3.3 (ii) we actually show the following equivalence for each y∗ ∈ Y ⊥: y∗ ∈ NA(X) ⇐⇒ the corresponding ϕ ∈ (X/Y )∗ attains its norm on q(BX ). Consider the situation when X = ℓ1 and Y has codimension d in X; then X/Y ∼= Rd. The set A := U(Rd,k·k2) satisfies the assumptions of Proposi- tion 5.7, so there is an equivalent norm · on ℓ1 such that q(B(ℓ1,·)) = 22 MARTIN RMOUTIL A. From the above equivalence we immediately obtain that Y ⊥ ⊂ (X ∗ \ NA(ℓ1, ·)) ∪ {0}. In other words, given any finite-dimensional subspace Z ⊂ (ℓ1)∗, we can find a renorming such that Z \ {0} is contained in non-norm-attaining functionals. Note that a result of F.J. Garc´ıa-Pacheco [7, Theorem 1.2] implies that ℓ∗ 1 \ NA(ℓ1) is, in fact, spaceable. It is also easy to see that, in the new norm, Y is antiproximinal (i.e. no point x /∈ Y has in Y a closest point). Indeed, the proof of Lemma 3.3 (i) also shows that X is antiproximinal ⇐⇒ q(BX ) = UX/Y . Likewise, given a closed subspace Y of finite codimension in a non-reflexive Banach space X, there exists an equivalent norm on X such that Y is prox- iminal in X. This follows from Lemma 5.5 as the set A := BX/Y satisfies the assumptions. (b) The assumption that Y is finite-codimensional is very important for our ap- proach to both Lemma 5.5 and Proposition 5.7. In the proof of the former, we use this assumption to see that Y is complemented, and therefore non- reflexive as its complement is finite-dimensional; this argument is not neces- sary in the latter case as ℓ1 contains no infinite-dimensional complemented reflexive subspaces-see [6, Section 13.7]. The fact that Y is non-reflexive is needed as any reflexive space is proximinal in any superspace (see [21]; the converse is also true-[19] or [22]). In both cases we rely on Proposition 5.3. (c) It is natural to ask whether the assumption that A is Fσ in Lemma 5.5 and Proposition 5.7 can be relaxed to higher Borel classes or even analytic sets. Related questions have already been studied: R. Kaufman proves in [15] that any non-reflexive Banach space admits an equivalent norm such that the set of norm-attaining functionals is not Borel; in [16] he improves this result showing that such a renorming can be made even Fr´echet smooth. Following Kaufman and using his deep abstract lemmas, O. Kurka proves in his paper [18] that given any separable non-reflexive Banach space X and an ordinal α < ω1, there exists an equivalent strictly convex norm on X such that the corresponding set of norm-attaining functionals is not of the (additive) Borel class α (it is observed in [15] that when X is strictly convex, N A(X) is Borel). It seems likely that Kurka's method could be (at least in some cases) used to prove a version of Lemma 5.5 with weaker assumptions on the descriptive quality of A. Acknowledgment. I would like to thank Bernardo Cascales (who suggested the topic of this article during my visit in Murcia), Richard Aron and Mat´ıas Raja for their helpful comments, moral support and interest in my work. Most of all, I am very grateful to Ondrej Kalenda for many fruitful discussions and careful reading of the manuscript which resulted in further helpful suggestions. References [1] Mar´ıa D. Acosta, Antonio Aizpuru, Richard M. Aron, and Francisco J. Garc´ıa-Pacheco. Functionals that do not attain their norm. Bull. Belg. Math. Soc. Simon Stevin, 14(3):407– 418, 2007. [2] Pradipta Bandyopadhyay and Gilles Godefroy. Linear structures in the set of norm-attaining functionals on a Banach space. J. Convex Anal., 13(3-4):489–497, 2006. NORM-ATTAINING FUNCTIONALS AND PROXIMINAL SUBSPACES 23 [3] Luis Bernal-Gonz´alez, Daniel Pellegrino, and Juan B. Seoane-Sep´ulveda. Linear subsets of nonlinear sets in topological vector spaces. Bull. Amer. Math. Soc. (N.S.), 51(1):71–130, 2014. [4] Errett Bishop and Robert R. Phelps. A proof that every Banach space is subreflexive. Bull. Amer. Math. Soc., 67:97–98, 1961. [5] Mahlon M. Day. Normed linear spaces. Springer-Verlag, New York-Heidelberg, third edition, 1973. Ergebnisse der Mathematik und ihrer Grenzgebiete, Band 21. [6] Mari´an Fabian, Petr Habala, Petr H´ajek, Vicente Montesinos, and V´aclav Zizler. Banach space theory. CMS Books in Mathematics/Ouvrages de Math´ematiques de la SMC. Springer, New York, 2011. The basis for linear and nonlinear analysis. [7] Francisco J. Garc´ıa-Pacheco. Vector subspaces of the set of non-norm-attaining functionals. Bull. Aust. Math. Soc., 77(3):425–432, 2008. [8] Aleksandr L. Garkavi. On best approximation by the elements of infinite-dimensional sub- spaces of a certain class. Mat. Sb. (N.S.), 62 (104):104–120, 1963. [9] Gilles Godefroy. The Banach space c0. Extracta Math., 16(1):1–25, 2001. [10] Gilles Godefroy and V. Indumathi. Proximinality in subspaces of c0. J. Approx. Theory, 101(2):175–181, 1999. [11] Gilles Godefroy and V. Indumathi. Proximinality and renormings: some new examples. J. Approx. Theory, 176:118–128, 2013. [12] V. Indumathi. Proximinal subspaces of finite codimension in general normed linear spaces. Proc. London Math. Soc. (3), 45(3):435–455, 1982. [13] V. Indumathi. On transitivity of proximinality. J. Approx. Theory, 49(2):130–143, 1987. [14] Robert C. James. Characterizations of reflexivity. Studia Math., 23:205–216, 1963/1964. [15] Robert P. Kaufman. Topics on analytic sets. Fund. Math., 139(3):215–229, 1991. [16] Robert P. Kaufman. On smooth norms and analytic sets. Israel J. Math., 116:21–27, 2000. [17] Victor Klee. Some new results on smoothness and rotundity in normed linear spaces. Math. Ann., 139:51–63 (1959), 1959. [18] Ondrej Kurka. Structure of the set of norm-attaining functionals on strictly convex spaces. Canad. Math. Bull., 54(2):302–310, 2011. [19] Walter Pollul. Reflexivitat und Existenz-Teilraume in der linearen Approximationstheorie. Gesellschaft fur Mathematik und Datenverarbeitung, Bonn, 1972. Gesellschaft fur Mathe- matik und Datenverarbeitung, Bonn, Ber. No. 53. [20] Charles J. Read. Banach spaces with no proximinal subspaces of codimension 2. arXiv:1307.7958. [21] Ivan Singer. Best approximation in normed linear spaces by elements of linear subspaces. Translated from the Romanian by Radu Georgescu. Die Grundlehren der mathematischen Wissenschaften, Band 171. Publishing House of the Academy of the Socialist Republic of Romania, Bucharest; Springer-Verlag, New York-Berlin, 1970. [22] Ivan Singer. On normed linear spaces which are proximinal in every superspace. J. Approxi- mation Theory, 7:399–402, 1973. E-mail address, M. Rmoutil: [email protected] University of Warwick, Mathematics Institute, Zeeman Building, Coventry CV4 7AL, United Kingdom Charles University, Faculty of Mathematics and Physics, Department of Mathematical Analysis, Sokolovsk´a 83, 186 75 Prague 8, Czech Republic
1004.2386
1
1004
2010-04-14T12:52:32
The Topological Centers Of Module Actions
[ "math.FA" ]
In this article, for Banach left and right module actions, we will extend some propositions from Lau and $\ddot{U}lger$ into general situations and we establish the relationships between topological centers of module actions. We also introduce the new concepts as $Lw^*w$-property and $Rw^*w$-property for Banach $A-bimodule$ $B$ and we investigate the relations between them and topological center of module actions. We have some applications in dual groups.
math.FA
math
THE TOPOLOGICAL CENTERS OF MODULE ACTIONS KAZEM HAGHNEJAD AZAR Abstract. In this article, for Banach left and right module actions, we will extend some propositions from Lau and U lger into general situations and we establish the relationships between topological centers of module actions. We also introduce the new concepts as Lw∗w-property and Rw∗w-property for Banach A − bimodule B and we investigate the relations between them and topological center of module actions. We have some applications in dual groups. 1.Introduction and Preliminaries As is well-known [1], the second dual A∗∗ of A endowed with the either Arens mul- tiplications is a Banach algebra. The constructions of the two Arens multiplications in A∗∗ lead us to definition of topological centers for A∗∗ with respect both Arens multiplications. The topological centers of Banach algebras, module actions and ap- plications of them were introduced and discussed in [6, 8, 13, 14, 15, 16, 17, 21, 22], and they have attracted by some attentions. Now we introduce some notations and definitions that we used throughout this paper. Let A be a Banach algebra. We say that a net (eα)α∈I in A is a left approximate identity (= LAI) [resp. right approximate identity (= RAI)] if, for each a ∈ A, eαa −→ a [resp. aeα −→ a]. For a ∈ A and a′ ∈ A∗, we denote by a′a and aa′ respectively, the functionals on A∗ defined by < a′a, b >=< a′, ab >= a′(ab) and < aa′, b >=< a′, ba >= a′(ba) for all b ∈ A. The Banach algebra A is embedded in its second dual via the identification < a, a′ > - < a′, a > for every a ∈ A and a′ ∈ A∗. We denote the set {a′a : a ∈ A and a′ ∈ A∗} and {aa′ : a ∈ A and a′ ∈ A∗} by A∗A and AA∗, respectively, clearly these two sets are subsets of A∗. Let A has a BAI. If the equality A∗A = A∗, (AA∗ = A∗) holds, then we say that A∗ factors on the left (right). If both equalities A∗A = AA∗ = A∗ hold, then we say that A∗ factors on both sides. Let X, Y, Z be normed spaces and m : X × Y → Z be a bounded bilinear mapping. Arens in [1] offers two natural extensions m∗∗∗ and mt∗∗∗t of m from X ∗∗ × Y ∗∗ into Z ∗∗ as following: 1. m∗ : Z ∗ × X → Y ∗, given by < m∗(z ′, x), y >=< z ′, m(x, y) > where x ∈ X, y ∈ Y , z ′ ∈ Z ∗, 2. m∗∗ : Y ∗∗ × Z ∗ → X ∗, given by < m∗∗(y′′, z ′), x >=< y′′, m∗(z ′, x) > where x ∈ X, y′′ ∈ Y ∗∗, z ′ ∈ Z ∗, 3. m∗∗∗ : X ∗∗ × Y ∗∗ → Z ∗∗, given by < m∗∗∗(x′′, y′′), z ′ > =< x′′, m∗∗(y′′, z ′) > where x′′ ∈ X ∗∗, y′′ ∈ Y ∗∗, z ′ ∈ Z ∗. The mapping m∗∗∗ is the unique extension of m such that x′′ → m∗∗∗(x′′, y′′) from X ∗∗ into Z ∗∗ is weak∗ − to − weak∗ continuous for every y′′ ∈ Y ∗∗, but the mapping 2000 Mathematics Subject Classification. 46L06; 46L07; 46L10; 47L25. Key words and phrases. Arens regularity, bilinear mappings, Topological center, Second dual, Module action. 1 2 y′′ → m∗∗∗(x′′, y′′) is not in general weak∗ − to − weak∗ continuous from Y ∗∗ into Z ∗∗ unless x′′ ∈ X. Hence the first topological center of m may be defined as following Z1(m) = {x′′ ∈ X ∗∗ : y′′ → m∗∗∗(x′′, y′′) is weak∗ − to − weak∗ − continuous}. Let now mt : Y × X → Z be the transpose of m defined by mt(y, x) = m(x, y) for every x ∈ X and y ∈ Y . Then mt is a continuous bilinear map from Y × X to Z, and so it may be extended as above to mt∗∗∗ : Y ∗∗ × X ∗∗ → Z ∗∗. The mapping mt∗∗∗t : X ∗∗ × Y ∗∗ → Z ∗∗ in general is not equal to m∗∗∗, see [1], if m∗∗∗ = mt∗∗∗t, then m is called Arens regular. The mapping y′′ → mt∗∗∗t(x′′, y′′) is weak∗ − to − weak∗ continuous for every y′′ ∈ Y ∗∗, but the mapping x′′ → mt∗∗∗t(x′′, y′′) from X ∗∗ into Z ∗∗ is not in general weak∗ − to − weak∗ continuous for every y′′ ∈ Y ∗∗. So we define the second topological center of m as Z2(m) = {y′′ ∈ Y ∗∗ : x′′ → mt∗∗∗t(x′′, y′′) is weak∗ − to − weak∗ − continuous}. It is clear that m is Arens regular if and only if Z1(m) = X ∗∗ or Z2(m) = Y ∗∗. Arens regularity of m is equivalent to the following lim i lim j < z ′, m(xi, yj) >= lim j lim i < z ′, m(xi, yj) >, whenever both limits exist for all bounded sequences (xi)i ⊆ X , (yi)i ⊆ Y and z ′ ∈ Z ∗, see [6, 18]. The regularity of a normed algebra A is defined to be the regularity of its algebra multiplication when considered as a bilinear mapping. Let a′′ and b′′ be elements of A∗∗, the second dual of A. By Goldstin,s Theorem [6, P.424-425], there are nets (aα)α and (bβ)β in A such that a′′ = weak∗ − limα aα and b′′ = weak∗ − limβ bβ. So it is easy to see that for all a′ ∈ A∗, and lim α lim β < a′, m(aα, bβ) >=< a′′b′′, a′ > lim β lim α < a′, m(aα, bβ) >=< a′′ob′′, a′ >, where a′′b′′ and a′′ob′′ are the first and second Arens products of A∗∗, respectively, see [6, 14, 18]. The mapping m is left strongly Arens irregular if Z1(m) = X and m is right strongly Arens irregular if Z2(m) = Y . This paper is organized as follows. a) In section two, for a Banach A − bimodule, we have (1) a′′ ∈ ZB∗∗(A∗∗) if and only if π∗∗∗∗ (2) F ∈ ZB∗∗((A∗A)∗) if and only if π∗∗∗∗ (3) G ∈ Z(A∗A)∗(B ∗∗) if and only if π∗∗∗∗ (4) Let B has a BAI (eα)α ⊆ A such that eα r ℓ ℓ (b′, a′′) ∈ B ∗ for all b′ ∈ B ∗. (g, F ) ∈ B ∗ for all g ∈ B ∗. (g, G) ∈ A∗A for all g ∈ B ∗. e∗∗(B ∗∗) = B ∗∗ [ resp. Ze∗∗(B ∗∗) = B ∗∗] and B ∗ factors on the left [resp. right], but not on the right [resp. left], then ZB∗∗(A∗∗) 6= Z t B∗∗(A∗∗). (5) B ∗A ⊆ wapℓ(B) if and only if AA∗∗ ⊆ ZB∗∗(A∗∗). (6) Let b′ ∈ B ∗. Then b′ ∈ wapℓ(B) if and only if the adjoint of the mapping w∗ → e′′. Then if Z t π∗ ℓ (b′, ) : A → B ∗ is weak∗ − to − weak continuous. 3 b) In section three, for a Banach A − bimodule B, we define Lef t − weak∗ − to − weak property [=Rw∗w− property] and Right − weak∗ − to − weak property [=Rw∗w− property] for Banach algebra A and we show that (1) If A∗∗ = a0A∗∗ [resp. A∗∗ = A∗∗a0] for some a0 ∈ A and a0 has Rw∗w− property [resp. Lw∗w− property], then ZB∗∗ (A∗∗) = A∗∗. (2) If B ∗∗ = a0B ∗∗ [resp. B ∗∗ = B ∗∗a0] for some a0 ∈ A and a0 has Rw∗w− property [resp. Lw∗w− property] with respect to B, then ZA∗∗(B ∗∗) = B ∗∗. (3) If B ∗ factors on the left [resp. right] with respect to A and A has Rw∗w− property [resp. Lw∗w− property], then ZB∗∗ (A∗∗) = A∗∗. (4) If B ∗ factors on the left [resp. right] with respect to A and A has Rw∗w− property [resp. Lw∗w− property] with respect B, then ZA∗∗(B ∗∗) = B ∗∗. (5) If a0 ∈ A has Rw∗w− property with respect to B, then a0A∗∗ ⊆ ZB∗∗(A∗∗) and a0B ∗ ⊆ wapℓ(B). (6) Assume that AB ∗ ⊆ wapℓB. If B ∗ strong factors on the left [resp. right], then A has Lw∗w− property [resp. Rw∗w− property ] with respect to B. (7) Assume that AB ∗ ⊆ wapℓB. If B ∗ strong factors on the left [resp. right], then A has Lw∗w− property [resp. Rw∗w− property ] with respect to B. 2. The topological centers of module actions Let B be a Banach A − bimodule, and let πℓ : A × B → B and πr : B × A → B. be the left and right module actions of A on B. Then B ∗∗ is a Banach A∗∗ − bimodule with module actions π∗∗∗ ℓ : A∗∗ × B ∗∗ → B ∗∗ and π∗∗∗ r : B ∗∗ × A∗∗ → B ∗∗. Similarly, B ∗∗ is a Banach A∗∗ − bimodule with module actions πt∗∗∗t ℓ : A∗∗ × B ∗∗ → B ∗∗ and πt∗∗∗t r : B ∗∗ × A∗∗ → B ∗∗. We may therefore define the topological centers of the right and left module actions of A on B as follows: ZA∗∗(B ∗∗) = Z(πr) = {b′′ ∈ B ∗∗ : the map a′′ → π∗∗∗ r (b′′, a′′) : A∗∗ → B ∗∗ is weak∗ − to − weak∗ continuous} ZB∗∗(A∗∗) = Z(πℓ) = {a′′ ∈ A∗∗ : the map b′′ → π∗∗∗ ℓ (a′′, b′′) : B ∗∗ → B ∗∗ is weak∗ − to − weak∗ continuous} Z t A∗∗(B ∗∗) = Z(πt ℓ) = {b′′ ∈ B ∗∗ : the map a′′ → πt∗∗∗ ℓ (b′′, a′′) : A∗∗ → B ∗∗ is weak∗ − to − weak∗ continuous} Z t B∗∗(A∗∗) = Z(πt r) = {a′′ ∈ A∗∗ : the map b′′ → πt∗∗∗ r (a′′, b′′) : B ∗∗ → B ∗∗ is weak∗ − to − weak∗ continuous} We note also that if B is a left(resp. right) Banach A − module and πℓ : A × B → B (resp. πr : B × A → B) is left (resp. right) module action of A on B, then B ∗ is a right (resp. left) Banach A − module. 4 We write ab = πℓ(a, b), ba = πr(b, a), πℓ(a1a2, b) = πℓ(a1, a2b), πr(b, a1a2) = πr(ba1, a2), π∗ ℓ (a1b′, a2) = π∗ b′ ∈ B ∗ when there is no confusion. r (b′, ab), for all a1, a2, a ∈ A, b ∈ B and ℓ (b′, a2a1), π∗ r (b′a, b) = π∗ Theorem 2-1. We have the following assertions. (1) Assume that B is a Left Banach A − module. Then, a′′ ∈ ZB∗∗(A∗∗) if and only if π∗∗∗∗ ℓ (b′, a′′) ∈ B ∗ for all b′ ∈ B ∗. (2) Assume that B is a right Banach A − module. Then, b′′ ∈ ZA∗∗(B ∗∗) if and only if π∗∗∗∗ r (b′, b′′) ∈ A∗ for all b′ ∈ B ∗. Proof. (1) Let b′′ ∈ B ∗∗. Then, for every a′′ ∈ ZB∗∗ (A∗∗), we have < π∗∗∗∗ ℓ =< πt∗∗∗t ℓ (b′, a′′), b′′ >=< b′, π∗∗∗ ℓ (a′′, b′′), b′ >=< πt∗∗∗ (b′, a′′) = πt∗∗ ℓ It follow that π∗∗∗∗ Conversely, let a′′ ∈ A∗∗ and let π∗∗∗∗ all b′′ ∈ B ∗∗, we have ℓ ℓ ℓ (a′′, b′′) >=< π∗∗∗ ℓ (a′′, b′′), b′ > (b′′, a′′), b′ >=< b′′, πt∗∗ (a′′, b′) ∈ B ∗. ℓ (a′′, b′) > . (a′′, b′) ∈ B ∗ for all b′ ∈ B ∗. Then for < π∗∗∗ ℓ (a′′, b′′), b′ >=< b′, π∗∗∗ ℓ (a′′, b′′) >=< π∗∗∗∗ ℓ (b′, a′′), b′′ > =< πt∗∗ ℓ (a′′, b′), b′′ >=< b′′, πt∗∗ (a′′, b′) >=< πt∗∗∗ ℓ ℓ =< πt∗∗∗t (a′′, b′′), b′ > . ℓ Consequently a′′ ∈ ZB∗∗(A∗∗). (2) Prof is similar to (1). (b′′, a′′), b′ > (cid:3) Theorem 2-2. Assume that B is a Banach A−bimodule. Then we have the following assertions. (1) F ∈ ZB∗∗((A∗A)∗) if and only if π∗∗∗∗ (2) G ∈ Z(A∗A)∗(B ∗∗) if and only if π∗∗∗∗ ℓ r (g, F ) ∈ B ∗ for all g ∈ B ∗. (g, G) ∈ A∗A for all g ∈ B ∗. Proof. (1) Let F ∈ ZB∗∗((A∗A)∗) and (b′′ α)α ⊆ B ∗∗ such that b′′ α w∗ → b′′. Then for all g ∈ B ∗, we have < π∗∗∗∗ ℓ (g, F ), b′′ α >=< g, π∗∗∗ ℓ (F, b′′ α) >=< π∗∗∗ ℓ (F, b′′ α), g > →< π∗∗∗ ℓ (F, b′′), g >=< π∗∗∗∗ ℓ (g, F ), b′′ > . (g, F ) ∈ (B ∗∗, weak∗)∗ = B ∗. (g, F ) ∈ B ∗ for F ∈ (A∗A)∗ and g ∈ B ∗. Assume that Thus, we conclude that π∗∗∗∗ Conversely, let π∗∗∗∗ ℓ ℓ α)α ⊆ B ∗∗ such that b′′ b′′ ∈ B ∗∗ and (b′′ α α) >=< π∗∗∗∗ α), g >=< g, π∗∗∗ (F, b′′ (g, F ) >=< π∗∗∗∗ (g, F ) >→< b′′, π∗∗∗∗ w∗ → b′′. Then ℓ α, π∗∗∗∗ < π∗∗∗ =< b′′ (F, b′′ ℓ ℓ ℓ ℓ =< π∗∗∗ It follow that F ∈ ZB∗∗((A∗A)∗). ℓ (2) Proof is similar to (1). ℓ (F, b′′), g > . (g, F ), b′′ α > (g, F ), b′′ > (cid:3) 5 In the proceeding theorems, if we take B = A, we obtain some parts of Lemma 3.1 from [14]. An element e′′ of A∗∗ is said to be a mixed unit if e′′ is a right unit for the first Arens multiplication and a left unit for the second Arens multiplication. That is, e′′ is a mixed unit if and only if, for each a′′ ∈ A∗∗, a′′e′′ = e′′oa′′ = a′′. By [4, p.146], an element e′′ of A∗∗ is mixed unit if and only if it is a weak∗ cluster point of some BAI (eα)α∈I in A. Let B be a Banach A − bimodule and a′′ ∈ A∗∗. We define the locally topological center of the left and right module actions of a′′ on B, respectively, as follows a′′ (B ∗∗) = Z t Z t a′′(πt Za′′ (B ∗∗) = Za′′ (πt ℓ) = {b′′ ∈ B ∗∗ : πt∗∗∗t r) = {b′′ ∈ B ∗∗ : πt∗∗∗t r ℓ (a′′, b′′) = π∗∗∗ (b′′, a′′) = π∗∗∗ ℓ r (a′′, b′′)}, (b′′, a′′)}. Thus we have \ a′′ ∈A∗∗ \ a′′ ∈A∗∗ Z t a′′(B ∗∗) = Z t A(B ∗∗) = Z(πt r), Za′′(B ∗∗) = ZA(B ∗∗) = Z(πr). Definition 2-3. Let B be a left Banach A − module and e′′ ∈ A∗∗ be a mixed unit for A∗∗. We say that e′′ is a left mixed unit for B ∗∗, if π∗∗∗ ℓ (e′′, b′′) = πt∗∗∗t ℓ (e′′, b′′) = b′′, for all b′′ ∈ B ∗∗. The definition of right mixed unit for B ∗∗ is similar. B ∗∗ has a mixed unit if it has left and right mixed unit that are equal. It is clear that if e′′ ∈ A∗∗ is a left (resp. right) unit for B ∗∗ and Ze′′ (B ∗∗) = B ∗∗, then e′′ is left (resp. right) mixed unit for B ∗∗. w∗ → e′′. e∗∗ (B ∗∗) = B ∗∗ [ resp. Ze∗∗ (B ∗∗) = B ∗∗] and B ∗ factors on the left [resp. Theorem 2-4. Let B be a Banach A−bimodule with a BAI (eα)α such that eα Then if Z t right], but not on the right [resp. left], then ZB∗∗(A∗∗) 6= Z t B∗∗(A∗∗). Proof. Suppose that B ∗ factors on the left with respect to A, but not on the right. w∗ → e′′. Thus for all b′ ∈ B ∗ there are Let (eα)α ⊆ A be a BAI for A such that eα a ∈ A and x′ ∈ B ∗ such that x′a = b′. Then for all b′′ ∈ B ∗∗ we have < π∗∗∗ ℓ (e′′, b′′), b′ >=< e′′, π∗∗ ℓ (b′′, b′) >= lim α < π∗∗ ℓ (b′′, b′), eα > = lim α = lim α ℓ (b′, eα) >= lim α ℓ (x′, aeα) >= lim α < b′′, π∗ ℓ (x′a, eα) > < π∗∗ ℓ (b′′, x′), aeα > < b′′, π∗ < b′′, π∗ =< π∗∗ ℓ (b′′, x′), a >=< b′′, b′ > . Thus π∗∗∗ e′′ ∈ ZB∗∗(A∗∗). If we take ZB∗∗(A∗∗) = Z t (e′′, b′′) = b′′ consequently B ∗∗ has left unit A∗∗ − module. It follows that B∗∗(A∗∗). Then the B∗∗(A∗∗), then e′′ ∈ Z t ℓ 6 mapping b′′ → πt∗∗∗t w∗ → e′′, πt∗∗∗t Since eα r (b′′, e′′) is weak∗ − to − weak∗ continuous from B ∗∗ into B ∗∗. (b′′, eα) w∗ (b′′, e′′). Let b′ ∈ B ∗ and (bβ)β ⊆ B such that → πt∗∗∗t r w∗ → b′′. Since Z t e∗∗(B ∗∗) = B ∗∗, we have the following quality r bβ < πt∗∗∗t r (b′′, e′′), b′ >= lim α < πt∗∗∗t r (b′′, eα), b′ >= lim α < πt∗∗∗ r (eα, b′′), b′ > = lim α lim β < πt∗∗∗ r (eα, bβ), b′ >= lim α lim β < πr(bβ, eα), b′ > = lim α lim β < b′, πr(bβ, eα) >= lim β lim α < b′, πr(bβ, eα) > = lim β < b′, bβ >=< b′′, b′ > . Thus πt∗∗∗t that b′′ ∈ B ∗∗ and (bβ)β ⊆ B such that bβ (b′′, e′′) = π∗∗∗ r r (b′′, e′′) = b′′. It follows that B ′′ has a right unit. Suppose w∗ → b′′. Then for all b′ ∈ B ∗ we have < b′′, b′ >=< π∗∗∗ r (b′′, e′′), b′ >=< b′′, π∗∗ r (e′′, b′) >= lim β < π∗∗ r (e′′, b′), bβ > = lim β < e′′, π∗ r (b′, bβ) >= lim β lim α < π∗ r (b′, bβ), eα > = lim β lim α < π∗ r (b′, bβ), eα >= lim β lim α < b′, πr(bβ, eα) > = lim α lim β < π∗∗∗ r (bβ, eα), b′ >= lim α lim β < bβ, π∗∗ r (eα, b′) > = lim α < b′′, π∗∗ r (eα, b′) > . It follows that weak − limα π∗∗ factors on the right that is contradiction. r (eα, b′) = b′. So by Cohen Factorization Theorem, B ∗ (cid:3) Corollary 2-5. Let B be a Banach A − bimodule and e′′ ∈ A∗∗ be a left mixed unit for B ∗∗. If B ∗ factors on the left, but not on the right, then ZB∗∗(A∗∗) 6= Z t B∗∗(A∗∗). In the proceeding corollary, if we take B = A, then it is clear Z t so we obtain Proposition 2.10 from [14]. e∗∗ (A∗∗) = A∗∗, and Theorem 2-6. Suppose that B is a weakly complete Banach space. Then we have the following assertions. (1) Let B be a Left Banach A − module and e′′ be a left mixed unit for B ∗∗. If AB ∗∗ ⊆ B, then B is reflexive. (2) Let B be a right Banach A − module and e′′ be a right mixed unit for B ∗∗. If ZA∗∗(B ∗∗)A ⊆ B, then ZA∗∗(B ∗∗) = B. Proof. (1) Assume that b′′ ∈ B ∗∗. Since e′′ is also mixed unit for A∗∗, there is a BAI (eα)α ⊆ A for A such that eα b′′ in B ∗∗. Since AB ∗∗ ⊆ B, we have π∗∗∗ π∗∗∗ and so B is reflexive. (e′′, b′′) = (eα, b′′) ∈ B. Consequently (e′′, b′′) = b′′ in B. Since B is a weakly complete, b′′ ∈ B, (eα, b′′) w→ π∗∗∗ → π∗∗∗ ℓ ℓ ℓ w∗ → e′′. Then π∗∗∗ (eα, b′′) w∗ ℓ ℓ 7 (2) Since b′′ ∈ ZA∗∗(B ∗∗), we have π∗∗∗ ZA∗∗(B ∗∗)A ⊆ B, π∗∗∗ π∗∗∗ (b′′, e′′) = b′′ in B ∗∗. Since (b′′, eα) w→ (b′′, e′′) = b′′ in B. It follows that b′′ ∈ B, since B is a weakly complete. (cid:3) (b′′, eα) ∈ B. Consequently we have π∗∗∗ → π∗∗∗ r r r r r (b′′, eα) w∗ A functional a′ in A∗ is said to be wap (weakly almost periodic) on A if the mapping a → a′a from A into A∗ is weakly compact. The procceding definition to the equiva- lent following condition, see [6, 14, 18]. For any two net (aα)α and (bβ)β in {a ∈ A : k a k≤ 1}, we have limαlimβ < a′, aαbβ >= limβlimα < a′, aαbβ >, whenever both iterated limits exist. The collection of all wap functionals on A is de- noted by wap(A). Also we have a′ ∈ wap(A) if and only if < a′′b′′, a′ >=< a′′ob′′, a′ > for every a′′, b′′ ∈ A∗∗. Definition 2-7. Let B be a left Banach A − module. Then, b′ ∈ B ∗ is said to be left weakly almost periodic functional if the set {πℓ(b′, a) : a ∈ A, k a k≤ 1} is relatively weakly compact. We denote by wapℓ(B) the closed subspace of B ∗ consisting of all the left weakly almost periodic functionals in B ∗. The definition of the right weakly almost periodic functional (= wapr(B)) is the same. By [18], the definition of wapℓ(B) is equivalent to the following < π∗∗∗ ℓ (a′′, b′′), b′ >=< πt∗∗∗t ℓ (a′′, b′′), b′ > for all a′′ ∈ A∗∗ and b′′ ∈ B ∗∗. Thus, we can write wapℓ(B) = {b′ ∈ B ∗ : < π∗∗∗ ℓ (a′′, b′′), b′ >=< πt∗∗∗t ℓ (a′′, b′′), b′ > f or all a′′ ∈ A∗∗, b′′ ∈ B ∗∗}. Theorem 2-8. Suppose that B is a left Banach A − module. Consider the following statements. (1) B ∗A ⊆ wapℓ(B). (2) AA∗∗ ⊆ ZB∗∗(A∗∗). (3) AA∗∗ ⊆ AZB∗∗((A∗A)∗). Then, we have (1) ⇔ (2) ⇐ (3). Proof. (1) ⇒ (2) Let (b′′ α)α ⊆ B ∗∗ such that b′′ α w∗ → b′′. Then for all a ∈ A and a′′ ∈ A∗∗, we have < π∗∗∗ ℓ (aa′′, b′′ =< a′′, π∗∗ ℓ (b′′ α, b′)a > ℓ (b′′ (a′′, b′′), b′a) > ℓ (b′′ α), b′ >=< aa′′, π∗∗ (a′′, b′′ (aa′′, b′′), b′) > . α, b′a) >=< π∗∗∗ =< π∗∗∗ ℓ ℓ α, b′) >=< a′′, π∗∗ α), b′a) >→< π∗∗∗ ℓ Hence aa′′ ∈ ZB∗∗ (A∗∗). (2) ⇒ (1) Let a ∈ A and b′ ∈ B ∗. Then < π∗∗∗ ℓ (a′′, b′′ α), b′a >=< aπ∗∗∗ ℓ (a′′, b′′ α), b′ >=< π∗∗∗ ℓ (aa′′, b′′ α), b′ > 8 =< πt∗∗∗t It follow that b′a ∈ wapℓ(B). (3) ⇒ (2) Since AZB∗∗((A∗A)∗) ⊆ ZB∗∗(A∗∗), proof is hold. α), b′ >=< πt∗∗∗t (aa′′, b′′ ℓ ℓ (a′′, b′′ α), b′a > . (cid:3) In the proceeding theorem, if we take B = A, then we obtain Theorem 3.6 from [14] and the same as proceeding theorem, we can claim the following assertions: If B is a right Banach A − module, then for the following statements we have (1) ⇔ (2) ⇐ (3). (1) AB ∗ ⊆ wapr(B). (2) A∗∗A ⊆ ZB∗∗(A∗∗). (3) A∗∗A ⊆ ZB∗∗((A∗A)∗)A. The proof of the this assertion is similar to proof of Theorem 2-8. Corollary 2-9. Suppose that B is a Banach A − bimodule. Then if A is a left [resp. right] ideal in A∗∗, then B ∗A ⊆ wapℓ(B) [resp. AB ∗ ⊆ wapr(B)]. Example 2-10. Suppose that 1 ≤ p ≤ ∞ and q is conjugate of p. We know that if G is compact, then L1(G) is a two-sided ideal in its second dual of it. By proceeding Theorem we have Lq(G) ∗ L1(G) ⊆ wapℓ(Lp(G)) and L1(G) ∗ Lq(G) ⊆ wapr(Lp(G)). Also if G is finite, then Lq(G) ⊆ wapℓ(Lp(G))∩wapr(Lp(G)). Hence we conclude that ZL1(G)∗∗(Lp(G)∗∗) = Lp(G) and ZLp(G)∗∗(L1(G)∗∗) = L1(G). Theorem 2-11. We have the following assertions. (1) Suppose that B is a left Banach A − module and b′ ∈ B ∗. Then b′ ∈ wapℓ(B) ℓ (b′, ) : A → B ∗ is weak∗−to−weak if and only if the adjoint of the mapping π∗ continuous. (2) Suppose that B is a right Banach A−module and b′ ∈ B ∗. Then b′ ∈ wapr(B) r (b′, ) : B → A∗ is weak∗−to−weak if and only if the adjoint of the mapping π∗ continuous. Proof. (1) Assume that b′ ∈ wapℓ(B) and π∗ ℓ (b′, )∗ : B ∗∗ → A∗ is the adjoint of ℓ (b′, ). Then for every b′′ ∈ B ∗∗ and a ∈ A, we have π∗ ℓ (b′, a) > . ℓ (b′, )∗b′′, a >=< b′′, π∗ < π∗ Suppose (b′′ α)α ⊆ B ∗∗ such that b′′ α w∗ → b′′ and a′′ ∈ A∗∗ and (aβ)β ⊆ A such that aβ w∗ → a′′. By easy calculation, for all y′′ ∈ B ∗∗ and y′ ∈ B ∗, we have < π∗ ℓ (y′, )∗, y′′ >= π∗∗ ℓ (y′′, y′). Since b′ ∈ wapℓ(B), < π∗∗∗ ℓ (a′′, b′′ α), b′ >→< π∗∗∗ ℓ (a′′, b′′), b′ > . Then we have the following statements lim α < a′′, π∗ ℓ (b′, )∗b′′ α >= lim α < a′′, π∗∗ ℓ (b′′ α, b′) > = lim α < π∗∗∗ ℓ (a′′, b′′ α), b′ >=< π∗∗∗ ℓ 9 (a′′, b′′), b′ > =< a′′, π∗ ℓ (b′, )∗b′′ > . It follow that the adjoint of the mapping π∗ weak continuous. Conversely, let the adjoint of the mapping π∗ weak continuous. Suppose (b′′ for every a′′ ∈ A∗∗, we have ℓ (b′, ) : A → B ∗ is weak∗ − to − α)α ⊆ B ∗∗ such that b′′ α ℓ (b′, ) : A → B ∗ is weak∗ − to − w∗ → b′′ and b′ ∈ B ∗. Then lim α < π∗∗∗ ℓ (a′′, b′′ α), b′ >= lim α < a′′, π∗∗ ℓ (b′′ α, b′) > = lim α < a′′, π∗ ℓ (b′, )∗b′′ α >=< a′′, π∗ ℓ (b′, )∗b′′ >=< π∗∗∗ ℓ It follow that b′ ∈ wapℓ(B). (2) proof is similar to (1). (a′′, b′′), b′ > . (cid:3) Corollary 2-12. Let A be a Banach algebra. Assume that a′ ∈ A∗ and Ta′ is the linear operator from A into A∗ defined by Ta′ a = a′a. Then, a′ ∈ wap(A) if and only if the adjoint of Ta′ is weak∗ − to − weak continuous. So A is Arens regular if and only if the adjoint of the mapping Ta′ a = a′a is weak∗ − to − weak continuous for every a′ ∈ A∗. 3. Lw∗w-property and Rw∗w-property In this section, we introduce the new definition as Lef t − weak∗ − to − weak property and Right−weak∗−to−weak property for Banach algebra A and make some relations between these concepts and topological centers of module actions. As some conclu- sion, we have ZL1(G)∗∗(M (G)∗∗) 6= M (G)∗∗ where G is a locally compact group. If G is finite, we have ZM(G)∗∗ (L1(G)∗∗) = L1(G)∗∗ and ZL1(G)∗∗(M (G)∗∗) = M (G)∗∗. Definition 3-1. Let B be a left Banach A − module. We say that a ∈ A has Lef t − weak∗ − to − weak property (= Lw∗w− property) with respect to B, if for w∗ → 0 implies ab′ α w→ 0. If every a ∈ A has Lw∗w− property all (bα)α ⊆ B ∗, ab′ α with respect to B, then we say that A has Lw∗w− property with respect to B. The definition of the Right − weak∗ − to − weak property (= Rw∗w− property) is the same. We say that a ∈ A has weak∗ − to − weak property (= w∗w− property) with respect to B if it has Lw∗w− property and Rw∗w− property with respect to B. If a ∈ A has Lw∗w− property with respect to itself, then we say that a ∈ A has Lw∗w− property. For proceeding definition, we have some examples and remarks as follows. a) If B is Banach A-bimodule and reflexive, then A has w∗w−property with respect to B. Then i) L1(G), M (G) and A(G) have w∗w−property when G is finite. ii) Let G be locally compact group. L1(G) [resp. M (G)] has w∗w−property [resp. 10 Lw∗w− property ]with respect to Lp(G) whenever p > 1. b) Suppose that B is a left Banach A − module and e is left unit element of A such that eb = b for all b ∈ B. If e has Lw∗w− property, then B is reflexive. c) If S is a compact semigroup, then C+(S) = {f ∈ C(S) : f > 0} has w∗w−property. Theorem 3-2. Suppose that B is a Banach A − bimodule. Then we have the following assertions. (1) If A∗∗ = a0A∗∗ [resp. A∗∗ = A∗∗a0] for some a0 ∈ A and a0 has Rw∗w− property [resp. Lw∗w− property], then ZB∗∗ (A∗∗) = A∗∗. (2) If B ∗∗ = a0B ∗∗ [resp. B ∗∗ = B ∗∗a0] for some a0 ∈ A and a0 has Rw∗w− property [resp. Lw∗w− property] with respect to B, then ZA∗∗(B ∗∗) = B ∗∗. Proof. (1) Suppose that A∗∗ = a0A∗∗ for some a0 ∈ A and a0 has Rw∗w− w∗ → b′′. Then for all a ∈ A and α)α ⊆ B ∗∗ such that b′′ α property. Let (b′′ b′ ∈ B ∗, we have α, π∗ α, b′) w∗ < π∗∗ ℓ (b′′ α, b′), a >=< b′′ ℓ (b′, a) >→< b′′, π∗ ℓ (b′, a) >=< π∗∗ ℓ (b′′, b′), a >, ℓ (b′′ w∗ α, b′)a0 it follow that π∗∗ → ℓ (b′′, b′)a0. ℓ (b′′, b′)a0. Since a0 has Rw∗w− property, π∗∗ π∗∗ Now let a′′ ∈ A∗∗. Then there is x′′ ∈ A∗∗ such that a′′ = a0x′′ consequently we have ℓ (b′′, b′). Also we can write π∗∗ ℓ (b′′ w→ π∗∗ α, b′)a0 → π∗∗ ℓ (b′′ < π∗∗∗ ℓ (a′′, b′′ α), b′ >=< a′′, π∗∗ ℓ (b′′ α, b′) >=< x′′, π∗∗ α, b′)a0 > →< x′′, π∗∗ ℓ (b′′, b′)a0 >=< π∗∗∗ ℓ (a′′, b′′ ℓ (b′′ α), b′ > . We conclude that a′′ ∈ ZB∗∗(A∗∗). Proof of the next part is the same as the proceeding proof. (2) Let B ∗∗ = a0B ∗∗ for some a0 ∈ A and a0 has Rw∗w− property with respect w∗ → a′′. Then for all b ∈ B, we α)α ⊆ A∗∗ such that a′′ α to B. Assume that (a′′ have < π∗∗ r (a′′ α, b′), b >=< a′′ α, π∗∗ r (b′, b) >→< a′′, π∗∗ r (b′, b) >=< π∗∗ r (a′′, b′), b > . α, b′) w∗ r (a′′ → π∗∗ We conclude that π∗∗ π∗∗ r (a′′, b′)a0. Since a0 has Rw∗w− property with respect to B, π∗∗ π∗∗ r (a′′, b′)a0. Now let b′′ ∈ B ∗∗. Then there is x′′ ∈ B ∗∗ such that b′′ = a0x′′. Hence, we have r (a′′, b′) then we have π∗∗ α, b′)a0 α, b′)a0 r (a′′ r (a′′ w∗ → w→ < π∗∗∗ r (b′′, a′′ =< x′′, π∗∗ r (a′′ r (a′′ α), b′ >=< b′′, π∗∗ α, b′)a0 >→< x′′, π∗∗ α, b′) >=< a0x′′, π∗∗ r (a′′ r (a′′, b′)a0 >=< b′′, π∗∗ α, b′) > r (a′′, b′) > =< π∗∗∗ r (b′′, a′′), b′ > . It follow that b′′ ∈ ZA∗∗(B ∗∗). The next part is similar to the proceeding proof. (cid:3) 11 i) Let G be a locally compact group. Since M (G) is a Banach Example 3-3. L1(G)-bimodule and the unit element of M (G) has not Lw∗w− property or Rw∗w− property, by Theorem 2-3, ZL1(G)∗∗(M (G)∗∗) 6= M (G)∗∗. ii) If G is finite, then by Theorem 2-3, we have ZM(G)∗∗(L1(G)∗∗) = L1(G)∗∗ and ZL1(G)∗∗(M (G)∗∗) = M (G)∗∗. Assume that B is a Banach A − bimodule. We say that B factors on the left (right) with respect to A if B = BA (B = AB). We say that B factors on both sides, if B = BA = AB. Theorem 3-4. Suppose that B is a Banach A − bimodule and A has a BAI. Then we have the following assertions. (1) If B ∗ factors on the left [resp. right] with respect to A and A has Rw∗w− property [resp. Lw∗w− property], then ZB∗∗ (A∗∗) = A∗∗. (2) If B ∗ factors on the left [resp. right] with respect to A and A has Rw∗w− property [resp. Lw∗w− property] with respect B, then ZA∗∗(B ∗∗) = B ∗∗. Proof. (1) Assume that B ∗ factors on the left and A has Rw∗w− property. Let α)α ⊆ B ∗∗ such that b′′ α w∗ → b′′. Since B ∗A = B ∗, for all b′ ∈ B ∗ there are (b′′ x ∈ A and y′ ∈ B ∗ such that b′ = y′x. Then for all a ∈ A, we have < π∗∗ =< b′′ ℓ (b′′ α, π∗ α, y′)x, a >=< b′′ α, π∗ ℓ (b′, a) >→< b′′, π∗ ℓ (y′, a)x >=< π∗∗ ℓ (b′, a) >=< π∗∗ α, b′), a > ℓ (b′′, y′)x, a > . ℓ (b′′ Thus, we conclude that π∗∗ property, π∗∗ ℓ (b′′ w→< π∗∗ ℓ (b′′ (a′′, b′′ α, y′)x α), b′ >=< a′′, π∗∗ ℓ (b′′, y′)x. Now let b′′ ∈ A∗∗. Then α, y′)x > α, b′) >=< a′′, π∗∗ ℓ (b′′ < π∗∗∗ ℓ α, y′)x w∗ →< π∗∗ ℓ (b′′, y′)x. Since A has Rw∗w− →< a′′, π∗∗ ℓ (b′′, y′)x >=< π∗∗∗ ℓ ℓ (b′′ (a′′, b′′), b′ > . It follow that a′′ ∈ ZB∗∗(A∗∗) = A∗∗. If B ∗ factors on the right and A has Lw∗w− property, then proof is the same as preceding proof. (2) Let B ∗ factors on the left with respect to A and A has Rw∗w− property with w∗ respect to B. Assume that (a′′ → a′′. Since B ∗A = B, for all b′ ∈ B ∗ there are x ∈ A and y′ ∈ B ∗ such that b′ = y′x. Then for all b ∈ B, we have α)α ⊆ A∗∗ such that a′′ α < π∗∗ r (a′′ α, y′)x, b >=< π∗∗ =< a′′, π∗ r (a′′ r (b′, b) >=< π∗∗ α, b′), b >=< a′′ α, π∗ r (a′′, y′)x, b > . r (b′, b) > Consequently π∗∗ respect to B, π∗∗ have r (a′′ r (a′′ α, y′)x α, y′)x w∗ → π∗∗ w→ π∗∗ r (a′′, y′)x. Since A has Rw∗w− property with r (a′′, y′)x. It follow that for all b′′ ∈ B ∗∗, we < π∗∗∗ r (b′′, a′′ α), b′ >=< b′′, π∗∗ =< π∗∗∗ r (a′′ (b′′, a′′), b′ > . r α, y′)x >→< b′′, π∗∗ r (a′′, y′)x > 12 Thus we conclude that b′′ ∈ ZA∗∗(B ∗∗). The proof of the next assertions is the same as proceeding proof. (cid:3) Theorem 3-5. Suppose that B is a Banach A−bimodule. Then we have the following assertions. (1) If a0 ∈ A has Rw∗w− property with respect to B, then a0A∗∗ ⊆ ZB∗∗(A∗∗) and a0B ∗ ⊆ wapℓ(B). (2) If a0 ∈ A has Lw∗w− property with respect to B, then A∗∗a0 ⊆ ZB∗∗(A∗∗) and B ∗a0 ⊆ wapℓ(B). (3) If a0 ∈ A has Rw∗w− property with respect to B, then a0B ∗∗ ⊆ ZA∗∗(B ∗∗) and B ∗a0 ⊆ wapr(B). (4) If a0 ∈ A has Lw∗w− property with respect to B, then B ∗∗a0 ⊆ ZA∗∗(B ∗∗) and a0B ∗ ⊆ wapr(B). Proof. (1) Let (b′′ α)α ⊆ B ∗∗ such that b′′ α w∗ → b′′. Then for all a ∈ A and b′ ∈ B ∗, we have < π∗∗ ℓ (b′′ α, b′)a0, a >=< π∗∗ ℓ (b′′ α, b′), a0a >=< b′′ α, π∗ ℓ (b′, a0a) > →< b′′, π∗ ℓ (b′, a0a) >=< π∗∗ ℓ (b′′, b′)a0, a > . w∗ → π∗∗ ℓ (b′′ It follow that π∗∗ with respect to B, π∗∗ We conclude that a0a′′ ∈ ZB∗∗(A∗∗) so that a0A∗∗ ∈ ZB∗∗(A∗∗). Since ℓ (b′′, b′)a0 = π∗∗ π∗∗ ℓ (b′′, b′)a0. Since a0 has Rw∗w− property w→ π∗∗ ℓ (b′′, b′a0), a0B ∗ ⊆ wapℓ(B). α, b′)a0 ℓ (b′′ ℓ (b′′, b′)a0. α, b′)a0 (2) proof is similar to (1). (3) Assume that (a′′ α)α ⊆ A∗∗ such that a′′ α w∗ → a′′. Let b ∈ B and b′ ∈ B ∗. Then we have < π∗∗ r (a′′ α, b′)a0, b >=< π∗∗ r (a′′ α, b′), a0b >=< a′′ α, π∗ r (b′, a0b) > →< a′′, π∗ r (b′, a0b) >=< π∗∗ r (a′′, b′)a0, b > . Thus we conclude π∗∗ erty with respect to B, π∗∗ have r (a′′ α, b′)a0 r (a′′ r (a′′, b′)a0. Since a0 has Rw∗w− prop- r (a′′, b′)a0. If b′′ ∈ B ∗∗, then we w→ π∗∗ w∗ → π∗∗ α, b′)a0 < π∗∗∗ r (a0b′′, a′′ α), b′ >=< a0b′′, π∗∗∗ r (a′′ α, b′) >=< b′′, π∗∗ r (a′′ α, b′)a0 > =< b′′, π∗∗ r (a′′ α, b′)a0 >=< π∗∗∗ r (a0b′′, a′′), b′ > . It follow that a0b′′ ∈ ZA∗∗(B ∗∗). Consequently we have a0B ∗∗ ∈ ZA∗∗(B ∗∗). The proof of the next assertion is clear. (4) Proof is similar to (3). (cid:3) 13 Theorem 3-6. Let B be a Banach A − bimodule. Then we have the following assertions. (1) Suppose lim α lim β < b′ β, bα >= lim β lim α < b′ β, bα >, for every (bα)α ⊆ B and (b′ Rw∗w− property with respect to B. β)β ⊆ B ∗. Then A has Lw∗w− property and (2) If for some a ∈ A, lim α lim β < ab′ β, bα >= lim β lim α < ab′ β, bα >, for every (bα)α ⊆ B and (b′ respect to B. Also if for some a ∈ A, β)β ⊆ B ∗, then a has Rw∗w− property with lim α lim β < b′ βa, bα >= lim β lim α < b′ βa, bα >, for every (bα)α ⊆ B and (b′ respect to B. β)β ⊆ B ∗, then a has Lw∗w− property with Proof. (1) Assume that a ∈ A such that ab′ β w∗ → 0 where (b′ β)β ⊆ B ∗. Let b′′ ∈ B ∗∗ and (bα)α ⊆ B such that bα w∗ → b′′. Then lim β < b′′, ab′ β >= lim β lim α < bα, ab′ β >= lim β lim α < ab′ β, bα > = lim α lim β < ab′ β, bα >= 0. We conclude that ab′ β has Rw∗w− property. w→ 0, so A has Lw∗w− property. It also easy that A (2) Proof is easy and is the same as (1). (cid:3) α = b′aα [resp. b′ Definition 3-7. Let B be a left Banach A − module. We say that B ∗ strong factors on the left [resp. right] if for all (b′ α)α ⊆ B ∗ there are (aα)α ⊆ A and b′ ∈ B ∗ such that b′ If B ∗ strong factors on the left and right, then we say that B ∗ strong factors on the both side. It is clear that if B ∗ strong factors on the left [resp. right], then B ∗ factors on the left [resp. right]. α = aαb′] where (aα)α has limit the weak∗ topology in A∗∗. Theorem 3-8. Suppose that B is a Banach A − bimodule. Assume that AB ∗ ⊆ wapℓB. If B ∗ strong factors on the left [resp. right], then A has Lw∗w− property [resp. Rw∗w− property ] with respect to B. Proof. Let (b′ are (aα)α ⊆ A and b′ ∈ B ∗ such that b′ α)α ⊆ B ∗ such that ab′ α w∗ → 0. Since B ∗ strong factors on the left, there α = b′aα. Let b′′ ∈ B ∗∗ and (bβ)β ⊆ B such 14 that bβ w∗ → b′′. Then we have lim α < b′′, ab′ α >= lim α lim β < bβ, ab′ α >= lim α lim β < ab′ α, bβ > = lim α lim β < ab′aα, bβ >= lim α lim β < ab′, aαbβ > = lim β lim α < ab′, aαbβ >= lim β lim α < ab′ α, bβ >= 0 It follow that ab′ α w→ 0. Problems . (cid:3) (1) Suppose that B is a Banach A − bimodule. If B is left or right factors with respect to A, dose A has Lw∗w−property or Rw∗w−property, respectively? (2) Suppose that B is a Banach A − bimodule. Let A has Lw∗w−property with respect to B. Dose ZB∗∗(A∗∗) = A∗∗? References 1. R. E. Arens, The adjoint of a bilinear operation, Proc. Amer. Math. Soc. 2 (1951), 839-848. 2. N. Arikan, A simple condition ensuring the Arens regularity of bilinear mappings, Proc. Amer. Math. Soc. 84 (4) (1982), 525-532. 3. J. Baker, A.T. Lau, J.S. Pym Module homomorphism and topological centers associated with weakly sequentially compact Banach algebras, Journal of Functional Analysis. 158 (1998), 186- 208. 4. F. F. Bonsall, J. Duncan, Complete normed algebras, Springer-Verlag, Berlin 1973. 5. H. G. Dales, A. Rodrigues-Palacios, M.V. Velasco, The second transpose of a derivation, J. London. Math. Soc. 2 64 (2001) 707-721. 6. H. G. Dales, Banach algebra and automatic continuity, Oxford 2000. 7. N. Dunford, J. T. Schwartz, Linear operators.I, Wiley, New york 1958. 8. M. Eshaghi Gordji, M. Filali, Arens regularity of module actions, Studia Math. 181 3 (2007), 237-254. 9. M. Eshaghi Gordji, M. Filali, Weak amenability of the second dual of a Banach algebra, Studia Math. 182 3 (2007), 205-213. 10. K. Haghnejad Azar, A. Riazi, Arens regularity of bilinear forms and unital Banach module space, (Submitted). 11. E. Hewitt, K. A. Ross, Abstract harmonic analysis, Springer, Berlin, Vol I 1963. 12. E. Hewitt, K. A. Ross, Abstract harmonic analysis, Springer, Berlin, Vol II 1970. 13. A. T. Lau, V. Losert, On the second Conjugate Algebra of locally compact groups, J. London Math. Soc. 37 (2)(1988), 464-480. 14. A. T. Lau, A. U lger, Topological center of certain dual algebras, Trans. Amer. Math. Soc. 348 (1996), 1191-1212. 15. S. Mohamadzadih, H. R. E. Vishki, Arens regularity of module actions and the second adjoint of a derivation, Bulletin of the Australian Mathematical Society 77 (2008), 465-476. 16. M. Neufang, Solution to a conjecture by Hofmeier-Wittstock, Journal of Functional Analysis. 217 (2004), 171-180. 17. M. Neufang, On a conjecture by Ghahramani-Lau and related problem concerning topological center, Journal of Functional Analysis. 224 (2005), 217-229. 18. J. S. Pym, The convolution of functionals on spaces of bounded functions, Proc. London Math Soc. 15 (1965), 84-104. 19. A. U lger, Arens regularity sometimes implies the RNP, Pacific Journal of Math. 143 (1990), 377-399. 20. A. U lger, Some stability properties of Arens regular bilinear operators, Proc. Amer. Math. Soc. (1991) 34, 443-454. 15 21. A. U lger, Arens regularity of weakly sequentialy compact Banach algebras, Proc. Amer. Math. Soc. 127 (11) (1999), 3221-3227. 22. A. U lger, Arens regularity of the algebra A ⊗B, Trans. Amer. Math. Soc. 305 (2) (1988) 623-639. 23. P. K. Wong, The second conjugate algebras of Banach algebras, J. Math. Sci. 17 (1) (1994), 15-18. 24. N. J. Young Theirregularity of multiplication in group algebra Quart. J. Math. Soc., 24 (2) (1973), 59-62. 25. Y. Zhing, Weak amenability of module extentions of Banach algebras Trans. Amer. Math. Soc. 354 (10) (2002), 4131-4151. Department of Mathematics, Amirkabir University of Technology, Tehran, Iran Email address: [email protected]
1906.09599
1
1906
2019-06-23T15:55:58
$L_p$ functional Busemann-Petty centroid inequality
[ "math.FA", "math.MG" ]
If $K\subset\mathbb{R}^n$ is a convex body and $\Gamma_pK$ is the $p$-centroid body of $K$, the $L_p$ Busemann-Petty centroid inequality states that $\vol(\Gamma_pK) \geq \vol(K)$, with equality if and only if $K$ is an ellipsoid centered at the origin. In this work, we prove inequalities for a type of functional $r$-mixed volume for $1 \leq r < n$, and establish as a consequence, a functional version of the $L_p$ Busemann-Petty centroid inequality. \keywords{Convex body, Moment body, Busemann-Petty centroid} }
math.FA
math
Lp functional Busemann-Petty centroid inequality J. E. Haddad∗, C. H. Jim´enez†, L. A. Silva‡ Abstract If K ⊂ Rn is a convex body and ΓpK is the p-centroid body of K, the Lp Busemann- Petty centroid inequality states that vol(ΓpK) ≥ vol(K), with equality if and only if K is an ellipsoid centered at the origin. In this work, we prove inequalities for a type of functional r-mixed volume for 1 ≤ r < n, and establish as a consequence, a functional version of the Lp Busemann-Petty centroid inequality. Keywords. Convex body, Moment body, Busemann-Petty centroid 1 Introduction The study of affine isoperimetric inequalities on one side and affine Sobolev inequalities for functions on Rn on the other is connected to a great extent. The equivalence of the classical isoperimetric inequality and the classical L1 Sobolev inequality has been known for quite some time (see for example[2, 33, 10, 6, 29, 24, 9]). Following this path Zhang in [34] established the equivalence of an affine L1 Sobolev inequality with the Petty Projection inequality for convex bodies. Some time after, along with Lutwak and Yang continued in this direction obtaining Lp versions of the mentioned equivalence. These authors developed around the same time a rich theory of geometrical inequalities for centroid bodies and established Lp extensions of many other fundamental parameters in Convex Geometry, such as mixed volume and surface area. On top of the strong connections mentioned above, other geometrical inequalities of isoperimetric flavour like the Busemann-Petty centroid inequality or Blaschke-Santal´o, among others, have been fundamental in the study of several inequalities of Sobolev type, like Lp log-Sobolev, Gagliardo-Nirenberg, Sobolev trace or weighted Sobolev inequalities (e.g [15, 8, 13, 14, 11, 12]). It is important to notice that in many of the works mentioned above, where the Busemann-Petty centroid inequality was used to recover some known results for Sobolev type inequalities, this inequality provided a more direct approach. This approach often went around the use (in their original proofs) of other well known tools in the area of convex geometric analysis like the Minkowski problem or the theory of mixed or dual mixed volumes. ∗Departamento de Matem´atica, ICEx, Universidade Federal de Minas Gerais, 30123-970, Belo Horizonte, Brasil; e-mail: [email protected] †Departamento de Matem´atica, Pontif´ıcia Universidade Cat´olica do Rio de Janeiro, 22451-900 Rio de Janeiro, Brasil [corresponding author]; e-mail: [email protected] ‡Pontif´ıcia Universidade Cat´olica do Rio de Janeiro, Departamento de Matem´atica, 22451-900, Rio de leti- Janeiro, Brasil. Professora do Ensino B´asico, T´ecnico e Tecnol´ogico no IFMG - Campus Bambu´ı; e-mail: [email protected] 1 In this work we continue with this line of research. We obtain a family of inequalities for functions on Rn, inequalities of Sobolev type, and that in particular recover the Lp Busemann- Petty centroid inequality for convex bodies in Rn. Our main inequality is presented in the form of a functional mixed volume inequality. 2 Theorem 1.1. Let f be a C 1 function and g a continuous non-negative function, both with compact support in Rn, then for 1 ≤ r < n, q = nr n−r and λ ∈(cid:16) n n+p , 1(cid:17) ∪ (1, ∞), [(n+p)(λ−1)+p]r ZRn(cid:18)ZRn g(y)h∇f (x), yipdy(cid:19)r/p dx ≥ Cn,p,λg 1 np(λ−1) − λr g λ (λ−1)n f r q. (1) The sharp constant Cn,p,λ is computed in Section 3 and equality is attained if and only if f and g have the following forms g(x) = aGp,λ(Ax2) f (x) = bFr(Ax2) for positive constants a, b, A ∈ GLn(R), Gp,λ : R+ → R defined by Gp,λ(t) =  and (1 + tp) (1 − tp) 1 λ−1 1 λ−1 + Fr(t) = (1 + t if λ ∈(cid:16) n n+p , 1(cid:17) if λ > 1, r r−1 )1− r n . 2 Some notations and tools from Convex Geometry In order to show the intrinsic geometric nature of inequality (1), and in particular, its relation to the Lp Busemann-Petty centroid inequality, let us first recall some basic definitions. A convex body is a convex set K ⊂ Rn which is compact and has non-empty interior. For a convex body K, its support function hK , which uniquely characterizes it, is defined as hK (x) = max{hx, yi : y ∈ K}. If K contains the origin in the interior, then we also have the gauge k · kK and radial rK (·) functions of K defined respectively as kykK := inf{λ > 0 : y ∈ λK} , y ∈ Rn \ {0} , rK(y) := max{λ > 0 : λy ∈ K} , y ∈ Rn \ {0} . Clearly, kykK = 1 rK (y) . For a convex body K ⊂ Rn and p ≥ 1, its Lp-moment and Lp-centroid bodies, denoted by MpK and ΓpK, are defined by their support functions hMpK(x)p =ZK hx, yipdy, and hΓpK(x)p = 1 vol(K)cn,p ZK hx, yipdy, (2) respectively, where cn,p = ωn+p of Rm. The Lp Busemann-Petty centroid inequality states that ω2ωnωp−1 and ωm is the m-dimensional volume of the unit ball B vol(ΓpK) ≥ vol(K) or vol(MpK) ≥ cn/p n,p vol(K) n+p p , (3) 3 in terms of the moment body MpK. Equality holds in (3) if and only if K is a 0-symmetric ellipsoid. Centroid bodies for p = 1 can be found for the first time in a work of Blaschke [3] whereas the respective Busemann-Petty centroid inequality for p = 1 is due to Petty [31]. The Lp version of centroid bodies above was introduced by Lutwak and Zhang [23], while (3) was obtained by Lutwak, Yang and Zhang in [19]. For the history of the Busemann-Petty centroid inequality and a comprehensive introduction on centroid and moment bodies we refer to Chapter 10 in [32]. The theory of mixed volumes, first developed by Minkowski [28, 27], is one of the pillars of the Brunn-Minkowski theory, it provides us with a unified approach to the study of several of the most important parameters in Convex Geometry, such as volume, mean width, surface area, among others. At the same time, it has been fundamental in many other problems ranging from characterization of special families of convex bodies to establish new isoperi- metric inequalities, we refer to [32, 4] for a comprehensive introduction to the theory of mixed volumes. There are several extensions of the concept of mixed volume, in this work we will focus mainly in the dual mixed volume and the Lp extension of the mixed volume, concepts belonging to the dual and Lp Brunn-Minkowski theory respectively. Regarding the latter we have the following Lp extension of mixed volume, for some background on this we refer to [18] and to [22] and the references therein. For r ≥ 1, the Lr-mixed volume Vr(K, L) of convex bodies K and L is defined by Vr(K, L) = r n lim ε→0 vol(K +r ε ·r L) − vol(K) ε , where K +r ε ·r L is the convex body defined by: hK+rε·rL(x)r = hK (x)r + εhL(x)r, ∀x ∈ Rn. One of the main aspects of the mixed volume is that it has an integral representation. As in the classical case for the Lr version it is known (see [18]) that there exists a unique finite positive Borel measure Sr(K, .) on Sn−1 such that for each convex body L. Vr(K, L) = 1 nZSn−1 hL(u)rdSr(K, u), (4) If 1 ≤ r < ∞ and K, L are convex bodies in Rn containing the origin as interior point, we can find also in [18] that Vr(K, L) ≥ vol(K) n−r n vol(L) r n , (5) with equality if and only if K and L are dilates of each other. Combining inequalities (5) and (3), we obtain: Vr(L, MpK) ≥ cr/p n,p vol(L) n−r n vol(K) (n+p)r np . (6) Taking L = MpK in (6), we recover (3), hence (6) is an equivalent formulation for the Lp Busemann-Petty centroid inequality. This and similar geometric inequalities for mixed vol- umes involving centroid and projection bodies were already considered in [17]. The main result, Theorem 1.1 is a functional version of inequality (6), replacing the sets L, K by func- tions f, g. In order to establish a functional version of (6) and considering the integral representation of the geometric Lr mixed volume (4), let us recall the following result obtained by Lutwak, Yang and Zhang, where they introduced the concept of surface area measure of a Sobolev function. The Lr surface area measure of a function f : Rn → R with Lr weak derivative is given by: 4 Lemma 2.1 (Lemma 4.1 of [22]). Given 1 ≤ r < ∞ and a function f : Rn → R with Lr weak derivative, there exists a unique finite Borel measure Sr(f, .) on Sn−1 such that ZRn φ(−∇f (x))rdx =ZSn−1 φ(u)rdSr(f, u), (7) for every non-negative continuous function φ : Rn → R homogeneous of degree 1. If f is not equal to a constant function almost everywhere, then the support of Sr(f, .) cannot be contained in any n − 1 dimensional linear subspace. Conversely, for a convex body L the function fL(x) = F (kxkL) satisfies Sr(f, .) = Sr(L, .) if F is any function F : R+ → R+ satisfying (see [22]). By the Sobolev inequality we have tn−1F ′(t)rdt = 1 Z ∞ 0 fL(x) ZRn nr n−r n−r dx ≤ c s nr (nωn) vol(L) n n−r ωn where cs is the sharp constant in the Sobolev inequality on Rn, and there is equality when F (t) = aFr(t) with a, b > 0, where Fr(t) = (1 + t r r−1 )1− r n . The function F (x2) is an extremal function of the euclidean Lr Sobolev inequality on Rn. In view of identity (7), for any f and L such that Sr(f, .) = Sr(L, .), we have Vr(L, K) = 1 nZRn hK(−∇f (x))rdx. This motivates the following definition. Definition 2.2. Given 1 ≤ r < ∞ and a function f : Rn → R with Lr weak derivative, we define Vr(f, K) = hK (−∇f (x))rdx 1 nZRn The Lp Sobolev inequality for general norms was proved in [7] and [1] and can be stated as a mixed volume inequality for functions as follows: Theorem 2.3. If f is a C 1 function with compact support in Rn and K is an origin- symmetric convex body, then for 1 < r < n and q = nr n−r Vr(f, K) ≥ cr 1kf kr q vol(K) r n , (8) where c1 is the optimal constant and equality holds in (8) if and only if f (x) = aFr(bkxkK ) for some a, b > 0. Taking f (x) = Fr(kxkL) we recover inequality (5). Theorem 2.3 was originally proved using an innovative approach based on optimal trans- portation of mass in [7] and in [1] using Convex Symmetrization. In Section 4 we give an alternative, simpler and elementary proof of this inequality using the tools developed in [20]. Some of the tools we are using here, specially those contained in [22], have been used in the study of Sobolev type inequalities. Their approach is often based on a functional extension of the so-called LYZ body and other known geometric inequalities for projection and polar projection bodies (see Subsection 10.15 in [32] and references therein for more on this). Let us go back to the definition of the moment body (2), it has been noticed that hMpK is a convex function regardless of the set K (see e.g. Chapter 5 in [5]). This observation allows us to make the following definition: 5 Definition 2.4. If g is a non-negative measurable function with compact support, we define the convex body Mpg by hMpg(ξ)p =ZRn g(x)hx, ξipdx. The left-hand side of (1) has then a geometric meaning: Vr(f, Mpg) = 1 nZRn(cid:18)ZRn g(y)h∇f (x), yipdy(cid:19)r/p dx. If K is a convex body and g(x) = G(kxkK ) for any non-negative continous function G : R+ → R with compact support, it is not hard to verify using polar coordinates that Mpg =(cid:18)(n + p)Z ∞ 0 tn+p−1G(t)dt(cid:19)1/p MpK. Our main result (Theorem 1.1) is a consequence of Theorem 2.3, and Theorem 2.5 below: Theorem 2.5. If g is a non-negative function with compact support in Rn, then, for each λ ∈(cid:16) n n+p , 1(cid:17) ∪ (1, ∞), we have that p vol(Mpg) n ≥ cn,pan,p,λg 1 (n+p)(λ−1)+p (λ−1)n − λp g λ (λ−1)n , (9) where an,p,λ is given by the Lemma (3.4). Let Gp,λ : R+ → R be defined by Gp,λ(t) =( (1 + tp) (1 − tp) 1 λ−1 1 λ−1 + if λ < 1 if λ > 1, then taking g(x) = Gp,λ(kxkK ) in (9) we recover (3). Equality holds in (9) if and only if g(x) = aGp,λ(A.x2) for any a > 0 and A ∈ Gln(Rn). Even though Theorem 2.5 contains the geometric core of the main Theorem 1.1, the term vol(Mpg) cannot be expressed in terms of g in an elementary way, as Vr(f, Mpg) does. This is the reason why we need to combine it with Theorem 2.3 to obtain a functional inequality. Let us note that Theorem 1.1 cannot be regarded as a functional mixed volume inequality in full generality since it can only be applied to a function f and the centroid/moment body of another function g. We refer the interested reader to review the works of Milman and Rotem [26, 25] where they have defined a functional extension of mixed volumes and have extended some of their main properties to a functional setting. We should finally also mention other related extension of the Busemann-Petty centroid inequality obtained by Paouris and Pivovarov in [30] where the authors obtained randomized versions of this and other important isoperimetric inequalities. The rest of the paper is organized as follows: In Section 3 we shall prove some preliminary results, including an extension of the Lp Busemann-Petty centroid inequality, to compact domains. Then in Section 4 we prove Theorems 2.3 and 2.5. We hope this work shed some more light into the deep connection between isoperimetric and functional inequalities. 6 3 Preliminary results In order to prove our main result, Theorem 1.1, we consider two cases: r = 1 and 1 < r < n. For r = 1, inequality (5) holds for more general sets. As in [34], a compact domain is the closure of a bounded open set. Lemma 3.1 (Lemma 3.2 of [34]). If M is a compact domain with piecewise C 1 boundary and K a convex body in Rn, then, with equality if and only if M and K are homothetic. V (M, K)n ≥ vol(M )n−1 vol(K), In the same spirit, the next lemma shows that the Lp-Busemann-Petty Centroid inequality remains valid for a compact domain: Lemma 3.2. If M is a compact domain, then vol(ΓpM ) ≥ vol(M ). (10) Equality holds in (10) if and only if M is a 0-symmetric ellipsoid. Proof. For a compact domain M and ξ ∈ Sn−1, we define the set Consider δ(t) = tn n , for t ≥ 0, and the star set SM defined by its radial function Lξ = {t ∈ [0, ∞) : tξ ∈ M }. ρSM (ξ) = δ−1(µ(δ(Lξ))), where µ denotes the one dimensional Lebesgue measure of Lξ. It is easy to see that vol(SM ) = vol(M ). Also, let s = δ(t) = tn n , then ds = tn−1dt. For x ∈ Rn, we have: ZM hx, yipdy =ZSn−1ZLξ =ZSn−1ZLξ =ZSn−1Zδ(Lξ) n ZSn−1 = n p hx, tξiptn−1dtdξ hx, ξiptptn−1dtdξ hx, ξip(ns) p n dsdξ hx, ξipZδ(Lξ) p n dsdξ s On the other hand, we have ZSM 0 hx, yipdy =ZSn−1Z ρSM (ξ) =ZSn−1Z ρSM (ξ) =ZSn−1Z δ(ρSM (ξ)) 0 0 hx, tξiptn−1dtdξ hx, ξiptptn−1dtdξ hx, ξip(ns) p n dsdξ = n p n ZSn−1 hx, ξipZ µ(δ(Lξ )) 0 p n dsdξ. s By the Bathtub principle (see Theorem 1.14, pag. 28 of [16]) we have 7 therefore, p n ds s p 0 s n ds ≥Z µ(δ(Lξ )) Zδ(Lξ) hx, yipdy ≥ZSM ZM Since vol(SM ) = vol(M ), we obtain hΓpM (x)p ≥ hΓpSM (x)p, whence ΓpM ⊃ ΓpSM and vol(ΓpM ) ≥ vol(ΓpSM ). We conclude, hx, yipdy. (11) vol(ΓpM ) ≥ vol(ΓpSM ) ≥ vol(SM ) = vol(M ). If M is a compact domain attaining equality in (10), then equality in (11) implies µ(δ(Lξ)) = δ(Lξ) for a.e ξ, meaning that M is a star body. We conclude the proof recalling the equality case of (3). Let f be a C 1 function with compact support in Rn. For t > 0, consider the level sets of f in Rn: and Nf,t = {x ∈ Rn : f (x) ≥ t} Sf,t = {x ∈ Rn : f (x) = t}. Since f is of class C 1, by Sard's Theorem, Sf,t is a C 1 submanifold which has non-zero normal vector ∇f , for almost all t. Denote by dSt the surface area element of Sf,t. Then the co-area formula relates the area elements dx = ∇f −1dStdt. We present a lemma, whose proof is inside of the proof of Theorem 4.1 of [34]. It will be useful to prove Theorem 1.1 for the case r = 1. Lemma 3.3. If f is a C 1 function with compact support in Rn, then: Z ∞ 0 vol(Nf,t) n−1 n dt ≥ f n n−1 . We observe that the proof of Lemma 3.3 carries over replacing n−1 not for η > 1. We prove an analogous result for η = n+p p > 1. n by any η ∈ (0, 1), but Lemma 3.4. If g is a C 1 function with compact support in Rn and λ ∈(cid:16) n n+p , 1(cid:17) ∪ (1, ∞) n+p vol(Ng,t) n dt ≥ an,p,λg 1 (n+p)(λ−1)+p (λ−1)n − λp g λ (λ−1)n , Z ∞ 0 where with an,p,λ =  A − (n+p)(λ−1)+p n,p,λ p (λ−1)n B (λ−1)n n,p,λ if λ > 1 if λ ∈(cid:16) n n+p , 1(cid:17) An,p,λ = ((λ − 1)n + λp)  λ−1(cid:17) (λp) Γ(cid:16) λ 1 1−λ ((λ − 1)(n + p))− n+p Γ(cid:16) n p + 1 λ−1 + 2(cid:17) p Γ(cid:16) n p + 2(cid:17) (λ−1)p (λ−1)n+λp   8 and Bn,p,λ = λ p n + p(cid:18)λ − n n + p(cid:19) (1−λ)(n+p) p −1  Proof. For λ > 1 and t > 0, let pλ(t) = (1 − tλ−1) (1 − tλ−1s1−λ) n p + and (1 − λ)− n 1−λ p(cid:17) 1−λ − n p −2Γ(cid:16) n p + 2(cid:17) Γ(cid:16) λ Γ(cid:16) λ−2 λ−1(cid:17)   + and l(t) = vol(Ng,t). Then pλ(cid:0) t n p s(cid:1) = ≥ 1 − tλ−1s1−λ. (12) p n pλ(cid:18) t s(cid:19) Multiplying (12) by l(t) and integrating, we obtain: p n 0 Z ∞ l(t)pλ(cid:18) t s(cid:19) g1 ≤Z ∞ dt ≥Z ∞ l(t)pλ(cid:18) t s(cid:19) 0 0 p n 0 l(t)dt − s1−λZ ∞ dt + s1−λZ ∞ 0 l(t)tλ−1dt, l(t)tλ−1dt. whence By Holder, observe that: p n Z ∞ 0 l(t)pλ(cid:18) t s(cid:19) dt ≤(cid:18)Z ∞ 0 l(t) n+p n dt(cid:19) Write u = t/s and dt = sdu. Then: n n+p Z ∞ 0 pλ(cid:18) t s(cid:19) n+p n p n+p . dt! p n Z ∞ 0 l(t)pλ(cid:18) t s(cid:19) dt ≤(cid:18)Z ∞ 0 l(t) n+p n dt(cid:19) Now, observe that: n n+p (cid:18)Z ∞ 0 pλ(u) n+p n du(cid:19) p n+p p n+p . s Z ∞ 0 l(t)tλ−1dt =Z ∞ 0 vol(Ngλ,tλ)tλ−1dt. Write v = tλ, dv = λtλ−1dt, then tλ−1dt = 1 λ dv and: l(t)tλ−1dt = Z ∞ 0 1 λZ ∞ 0 vol(Ngλ,t)dt = 1 λ gλ λ. Hence, g1 ≤ 1 λ gλ where a = 1 λ gλ 0 λs1−λ +(cid:18)Z ∞ λ, b =(cid:16)R ∞ 0 l(t) n 0 n+p n+p (cid:18)Z ∞ n dt(cid:19) n+p (cid:16)R ∞ n dt(cid:17) n 0 pλ(t) n+p l(t) p n+p pλ(t) n+p n dt(cid:19) p n+p = as−α + bsβ, s (13) p n+p , α = λ − 1 e β = p n+p . n+p n dt(cid:17) Notice that the right-hand side of (13) has a unique minimum for s ∈ (0, ∞), then mini- mizing with respect to s ∈ (0, ∞), we obtain: (n+p)(λ−1)+p g1 ≤ An,p,λg λ λp (cid:18)Z ∞ 0 l(t) n+p n dt(cid:19) (λ−1)n (n+p)(λ−1)+p , where An,p,λ is given in the statement of the Lemma. Hence, 9 (n+p)(λ−1)+p g1 ≤ An,p,λg λ λp (cid:18)Z ∞ 0 vol(Ng,t) n+p n dt(cid:19) (λ−1)n (n+p)(λ−1)+p , that proves the statement of Lemma for the case λ > 1. n+p , 1(cid:17), we define qλ(t) = (tλ−1 − 1) n p +. Then, qλ(t) p n ≥ tλ−1 − 1 and p For the case λ ∈ (cid:16) n qλ(cid:0) t s(cid:1) n ≥ tλ−1s1−λ − 1. It follows that l(t)tλ−1dt = 1 l(t)dt = g1, we obtain Z ∞ 0 p n 0 λ gλ dt ≥ s1−λZ ∞ s(cid:19) l(t)qλ(cid:18) t λ and R ∞ λ ≤ g1 +Z ∞ gλ s1−λ λ 0 0 l(t)tλ−1dt −Z ∞ 0 l(t)dt p n dt. l(t)qλ(cid:18) t s(cid:19) Since R ∞ 0 By Holder p n Z ∞ 0 l(t)qλ(cid:18) t s(cid:19) 0 dt ≤(cid:18)Z ∞ =(cid:18)Z ∞ 0 l(t) l(t) n+p n dt(cid:19) n dt(cid:19) n+p n 0 n+p Z ∞ n+p (cid:18)Z ∞ 0 n n+p n dt! qλ(cid:18) t s(cid:19) n du(cid:19) qλ(u) n+p p n+p p n+p s p n+p Hence, 1 λ gλ For λ ∈ (cid:16) n n p n+p l(t) λ ≤ sλ−1g1 +(cid:18)Z ∞ n+p , 1(cid:17), the right-hand side of (14) has a unique minimum s ∈ (0, ∞), then n+p (cid:18)Z ∞ n dt(cid:19) n dt(cid:19) n+p +λ−1. qλ(t) (14) n+p n+p s 0 0 p minimizing with respect to s ∈ (0, ∞), we obtain: gλ λ ≤ Bn,p,λg 1 p (n+p)(λ−1)+p (cid:18)Z ∞ 0 l(t) n+p n dt(cid:19) (1−λ)n p , where Bn,p,λ is given in the statement of the Lemma. Therefore, vol(Ng,t) n+p n dt ≥ B − p c,d,λ (1−λ)n − (n+p)(λ−1)+p g 1 (1−λ)n pλ (1−λ)n g λ . Z ∞ 0 Now, we present other tools for the case 1 < r < n of our main result, introduced by Lutwak, Yang, and Zhang in [20]. Let H 1,r(Rn) denote the usual Sobolev space of real- valued functions of Rn with Lr partial derivatives. If f ∈ H 1,r(Rn) ∩ C ∞(Rn) and Q is a compact convex set that contains the origin in its relative interior, then they define Vr(f, t, Q) = 1 nZSf,t hQ(ν(x))r∇f (x)r−1dSt(x), 10 where ν(x) = ∇f (x) convex body Kt such that, for each origin-symmetric convex body Q ∇f (x) . They prove that for almost every t > 0, there exists an origin-symmetric Vr(Kt, Q) = Vr(f, t, Q). (15) The next Lemma can be deduced from [20], inequalities (6.3), (5.3), (5.4), and (5.1). Lemma 3.5. If r ∈ (1, n), f ∈ H 1,r(Rn) and q = nr n−r , then vol(Kt) n−r n dt ≥ n Z ∞ 0 r−n n cr 2f r q, where c2 = n 1 q (cid:18) n − r r − 1(cid:19) r−1 r Γ(cid:0) n r(cid:1) Γ(cid:0)n + 1 − n r(cid:1) Γ(n) 1 n ! . 4 Proof of the main results We present separate proofs for the cases 1 < r < n and r = 1. 4.1 Case 1 < r < n: Proof of 2.3. By the co-area formula, (15), (5) and Lemma 3.5: Vr(f, K) = hK (−∇f (x))rdx hK(n Sf,t x )r∇f (x)r−1dSf,tdt Vr(f, t, K)dt Vr(Kt, K)dt 1 1 = nZRn nZ ∞ 0 ZSf,t =Z ∞ =Z ∞ ≥Z ∞ =Z ∞ 0 r−n ≥ n n cr 0 0 0 vol(Kt) vol(Kt) 2f r q vol(K) r n n−r n vol(K) r n dt n−r n dt vol(K) r n Proof of 2.5. We may observe that: g(x)hx, ξipdx hx, ξipdxdt hMpg(ξ)p =ZRn =Z ∞ 0 Z{g≥t} =Z ∞ 0 hMpNg,t(ξ)pdt. In this sense, we regard Mpg as a generalized p-sum of sets, where we replace finite p-sums 11 by a p-integral of sets and clearly, for any convex body K, Vp(cid:18)K,Zp We compute: MpNg,tdt Mpg =Zp MpNg,tdt(cid:19) =Z ∞ 0 Vp(K, MpNg,t)dt. vol(Mpg) = Vp(Mpg, Mpg) MpNg,tdt(cid:19) Vp(Mpg, MpNg,t)dt = Vp(cid:18)Mpg,Zp =Z ∞ 0 ≥ vol(Mpg) n−p n Z ∞ 0 vol(Mpg) vol(MpNg,t) p 0 n ≥Z ∞ ≥ cn,pZ ∞ 0 vol(Ng,t) n+p n dt (n+p)(λ−1)+p vol(MpNg,t)p/ndt p n dt − λp g λ (λ−1)n , Then using Lemmas 3.2 and 3.4, it follows that where an,p,λ is given by Lemma (3.4). ≥ cn,pan,p,λg 1 (λ−1)n 4.2 Proof of Theorem 1.1: Case r = 1 Proof. Let V1(f, Mpg) = 1 V1(f, Mpg) = dStdt. We denote ηSt x = ∇f (x) ∇f (x) , as p 1 (cid:28) ∇f (x) ∇f (x) nRRn(cid:0)RRn g(y)h∇f (x), yipdy(cid:1)1/p dx. Then, dy(cid:19)1/p nZ ∞ , y(cid:29)(cid:12)(cid:12)(cid:12)(cid:12) 0 ZSf,t(cid:18)ZRn dy(cid:19)1/p dyds!1/p g(y)(cid:12)(cid:12)(cid:12)(cid:12) g(y)(cid:12)(cid:12)(cid:10)ηSt x , y(cid:11)(cid:12)(cid:12) 0 ZNg,s(cid:12)(cid:12)(cid:10)ηSt x , y(cid:11)(cid:12)(cid:12) x ) =(cid:18)ZRn = Z ∞ p p , hMpg(ηSt it follows that: V1(f, Mpg) = = hMpg(ηSt x )dStdt 1 nZ ∞ 0 ZSf,t 0 ZSf,t Z ∞ nZ ∞ 1 p 0 ZNg,s(cid:12)(cid:12)(cid:10)ηSt x , y(cid:11)(cid:12)(cid:12) dyds!1/p dStdt Write hMpNg,s(ηSt p dy, then: x )p =RNg,s(cid:12)(cid:12)(cid:10)ηSt x , y(cid:11)(cid:12)(cid:12) 0 ZSf,t(cid:18)Z ∞ nZ ∞ V1(f, Mpg) = 1 0 hMpNg,s(ηSt x )pds(cid:19)1/p dStdt By the co-area formula, the Minkowski integral inequality and Lemmas 3.1, 3.2, 3.5 and 3.4: 12 V1(f, Mpg) ≥ 0 1 1 ≥ 0 Z ∞ nZ ∞ 0 Z ∞ n Z ∞ =(cid:18)Z ∞ 0 (cid:18)Z ∞ ≥(cid:18)Z ∞ 0 (cid:18)Z ∞ =(cid:18)Z ∞ 0 (cid:18)Z ∞ =(cid:18)Z ∞ n,p(cid:18)Z ∞ ≥ c 1 p 0 0 0 0 vol(Nf,t) ≥ c1/p n,p f n n−1 C hMpNg,s(ηSt 0 ZSf,t 0 ZSf,t V1(Nf,t, MpNg,s)dt(cid:19)p hMpNg,s(ηSt vol(Nf,t) n−1 n vol(MpNg,s) vol(Nf,t) n−1 n dt(cid:19)p n dt(cid:19)(cid:18)Z ∞ 0 n−1 1 p x )dSt!p ds! x )dSt! dt!p ds(cid:19) n dt(cid:19)p 1 p 1 vol(MpNg,s) vol(MpNg,s) dt 1 p ds! 1 p p 1 p ds(cid:19) n ds(cid:19) n ds(cid:19) n ds(cid:19) n+p 1 p p 1 p vol(Nf,t) n−1 n dt(cid:19)(cid:18)Z ∞ vol(Ng,s) 0 (n+p)(λ−1)+p − (n+p)(λ−1)+p n,p,λ (λ−1)np (λ−1)np g 1 − λ g λ (λ−1)n . Remark 4.1. Let us point out that a simpler proof of Theorem 1.1 for the case r = p can be deduced using the Lp Affine Sobolev inequality [20] and the equivalence between the Lp Busemann-Petty centroid inequality and the Lp Petty projection inequality (see [19]). The well known identity for sets Vp(L, ΓpK) = ωn vol(K) V−p(K, Π◦ pL), where Vp(·, ·) denotes the Lp dual mixed volume and Π◦ can be extended to functions as pL the Lp polar projection body of L, Vp(f, Mpg) = V−p(g, Π◦ pf ) where we define and V−p(g, L) =ZRn pf, ξ)p =ZRn h(Π◦ xp Lg(x)dx h∇f (x), ξ)ipdx. Then an application of the dual mixed volume inequality for functions (Lemma 4.1 in [21]) and the Lp Affine Sobolev inequality (which corresponds to the Lp Petty Projection inequality for functions), gives the result. 13 Acknowledgements The first author was partially supported by Fapemig, Project APQ-01542-18 and CNPQ grant PQ-301203/2017-2. The second and third authors are partially supported by FAPERJ grant JCNE 236508 and CNPQ grant 428076/2018-1. The second author was also partially supported by CNPQ grant PQ 305650/2016-5 and PUC-Rio programa de incentivo a pro- dutividade em pesquisa. The third author acknowledges the support of the IFMG campus Bambui while conducting this work. References [1] Angelo Alvino, Vincenzo Ferone, Guido Trombetti, and Pierre-Louis Lions. Convex symmetrization and applications. In Annales de l'Institut Henri Poincare (C) Non Linear Analysis, volume 14, pages 275 -- 293. Elsevier, 1997. [2] Thierry Aubin. Problemes isop´erim´etriques et espaces de Sobolev. Journal of differential geometry, 11(4):573 -- 598, 1976. [3] Wilhelm Blaschke. Affine geometrie ix: Verschiedene Bemerkungen und aufgaben. Ber. Verh. Sachs. Akad. Wiss. Leipzig Math. -- Phys. Kl, 69:412 -- 420, 1917. [4] Fenchel W. Bonnesen, T. Theorie der Konvexen Korper. Springer, Berlin, 1934. Reprint: Chelsea Publ. Co., New York, 1948. English translation: BCS Associates, Moscow, Idaho, 1987. [5] Silouanos Brazitikos, Apostolos Giannopoulos, Petros Valettas, and Beatrice-Helen Vrit- siou. Geometry of isotropic convex bodies, volume 196. American Mathematical Soc., 2014. [6] Yu.D. Burago and V.A. Zalgaller. Geometric Inequalities. Grundlehren der Mathematis- chen Wissenschaften, 285, Springer-Verlag, Berlin, 1988, MR 89b:52020, Zbl 0633.53002. [7] Dario Cordero-Erausquin, Bruno Nazaret, and C´edric Villani. A mass-transportation approach to sharp Sobolev and Gagliardo-Nirenberg inequalities. Advances in Mathe- matics, 182(2):307 -- 332, 2004. [8] PL De N´apoli, Juli´an Haddad, Carlos Hugo Jim´enez, and Marcos Montenegro. The sharp affine L2 Sobolev trace inequality and variants. Mathematische Annalen, 370(1- 2):287 -- 308, 2018. [9] H. Federer. Geometric Measure Theory. Springer-Verlag, Berlin, 1969. [10] Herbert Federer and Wendell H Fleming. Normal and integral currents. Annals of Mathematics, pages 458 -- 520, 1960. [11] Christoph Haberl and Franz E Schuster. Asymmetric affine Lp Sobolev inequalities. Journal of Functional Analysis, 257(3):641 -- 658, 2009. [12] Christoph Haberl and Franz E Schuster. General Lp affine isoperimetric inequalities. Journal of Differential Geometry, 83(1):1 -- 26, 2009. [13] J Haddad, C Jim´enez, and M Montenegro. Sharp affine weighted Lp Sobolev type inequalities. Transactions of the American Mathematical Society, 2018. [14] Julian Haddad, C Hugo Jimenez, and Marcos Montenegro. Asymmetric Blaschke-Santal´o functional inequalities. arXiv preprint arXiv:1810.02288, 2018. [15] Julian Haddad, Carlos Hugo Jim´enez, and Marcos Montenegro. Sharp affine Sobolev type inequalities via the Lp Busemann-Petty centroid inequality. Journal of Functional Analysis, 271(2):454 -- 473, 2016. 14 [16] Elliott H Lieb and Michael Loss. Analysis, graduate studies in mathematics, vol. 14. American Mathematical Society, Providence, 2001. [17] Erwin Lutwak. Mixed projection inequalities. Transactions of the American Mathemat- ical Society, 287(1):91 -- 106, 1985. [18] Erwin Lutwak. The Brunn-Minkowski-Firey theory. i. Mixed volumes and the Minkowski problem. J. Differential Geom, 38(1):131 -- 150, 1993. [19] Erwin Lutwak, Deane Yang, and Gaoyong Zhang. Lp affine isoperimetric inequalities. J. Differential Geom, 56(1):111 -- 132, 2000. [20] Erwin Lutwak, Deane Yang, and Gaoyong Zhang. Sharp affine Lp Sobolev inequalities. Journal of Differential Geometry, 62(1):17 -- 38, 2002. [21] Erwin Lutwak, Deane Yang, and Gaoyong Zhang. Moment-entropy inequalities. The Annals of Probability, 32(1B):757 -- 774, 2004. [22] Erwin Lutwak, Deane Yang, and Gaoyong Zhang. Optimal Sobolev norms and the Lp Minkowski problem. International Mathematics Research Notices, 2006, 2006. [23] Erwin Lutwak and Gaoyong Zhang. Blaschke-Santal´o inequalities. J. Differential Geom, 47(1):1 -- 16, 1997. [24] Vladimir Gilelevich Maz'ya. Classes of domains and imbedding theorems for function spaces. In Doklady Akademii Nauk, volume 133, pages 527 -- 530. Russian Academy of Sciences, 1960. [25] Vitali Milman and Liran Rotem. α-concave functions and a functional extension of mixed volumes. arXiv preprint arXiv:1302.0823, 2013. [26] Vitali Milman and Liran Rotem. Mixed integrals and related inequalities. Journal of Functional Analysis, 264(2):570 -- 604, 2013. [27] Hermann Minkowski. Theorie der konvexen korper, insbesondere begrundung ihres oberflachenbegriffs. Gesammelte Abhandlungen, 2:131 -- 229, 1911. [28] Hermann Minkowski. Volumen und oberflache. Math. Ann. 57 (1903), 447 -- 495. Gesam- melte Abhandlungen, vol. II, pp. 230 -- 276, Teubner, Leipzig, 1911. [29] Robert Osserman. The isoperimetric inequality. Bulletin of the American Mathematical Society, 84(6):1182 -- 1238, 1978. [30] Grigoris Paouris and Peter Pivovarov. Randomized isoperimetric inequalities. In Con- vexity and Concentration, pages 391 -- 425. Springer, 2017. [31] Clinton M Petty. Centroid surfaces. Pacific Journal of Mathematics, 11(4):1535 -- 1547, 1961. [32] Rolf Schneider. Convex bodies: the Brunn-Minkowski theory. Number 151. Cambridge university press, 2014. [33] Giorgio Talenti. Best constant in Sobolev inequality. Annali di Matematica pura ed Applicata, 110(1):353 -- 372, 1976. [34] Gaoyong Zhang. The affine Sobolev inequality. J. Differential Geom, 53(1):183 -- 202, 1999.
1705.04635
2
1705
2017-06-19T11:05:20
Local approach to order continuity in Ces\`aro function spaces
[ "math.FA" ]
The goal of this paper is to present a complete characterisation of points of order continuity in abstract Ces\`aro function spaces $CX$ for $X$ being a symmetric function space. Under some additional assumptions mentioned result takes the form $(CX)_a = C(X_a)$. We also find simple equivalent condition for this equality which in the case of $I=[0,1]$ comes to $X\neq L^\infty$. Furthermore, we prove that $X$ is order continuous if and only if $CX$ is, under assumption that the Ces\`aro operator is bounded on $X$. This result is applied to particular spaces, namely: Ces\`aro-Orlicz function spaces, Ces\`aro-Lorentz function spaces and Ces\`aro-Marcinkiewicz function spaces to get criteria for OC-points.
math.FA
math
LOCAL APPROACH TO ORDER CONTINUITY IN CES `ARO FUNCTION SPACES TOMASZ KIWERSKI AND JAKUB TOMASZEWSKI Abstract. The goal of this paper is to present a complete characterization of points of order continuity in abstract Ces`aro function spaces CX for X being a symmetric function space. Under some additional assumptions mentioned result takes the form (CX)a = C(Xa). We also find simple equivalent condition for this equality which in the case of I = [0, 1] comes to X 6= L∞. Furthermore, we prove that X is order continuous if and only if CX is, under assumption that the Ces`aro operator is bounded on X. This result is applied to particular spaces, namely: Ces`aro-Orlicz function spaces, Ces`aro-Lorentz function spaces and Ces`aro- Marcinkiewicz function spaces to get criteria for OC-points. 1. Introduction Much has been said about the Ces`aro spaces both from isomorphic and isometric point of view, see for example the papers [1], [2] by Astashkin and Maligranda, [9] by Curbera and Ricker, [10] by Delgado and Soria and [22], [23], [24] by Le´snik and Maligranda and references given there. For a Banach ideal space X the abstract Ces`aro space CX is the space of all functions f such that Cf ∈ X, equipped with the norm kf kCX = kCf kX , where C denotes the continuous Ces`aro operator General considerations for this construction has been initiated in [22]. The abstract Ces`aro spaces CX are neither symmetric nor reflexive. Surprisingly, the descriptions of Kothe duals of CX spaces is different for the case of I = [0, 1] and I = [0, ∞), see [22]. Here we will focus on considering the abstract Ces`aro spaces CX for those function spaces X which are symmetric. We study the local structure of this spaces in the terms of order continuity property. The paper is organized as follows. In Section 2 we collect some necessary preliminaries on Banach ideal spaces, symmetric func- tion spaces, Ces`aro function spaces and order continuity property. Here, we specify Theorems A, B, C, D and E, Fact 1 and Lemma 2 because we will used them often in this article. Section 3 contains the main results of this paper. Curbera and Ricker in [9] proved that (CX)a = C(Xa) for symmetric spaces X 6= L∞[0, 1] on I = [0, 1]. Moreover, Kiwerski and Kolwicz in [15] have shown analogous equality in the case of the Ces`aro-Orlicz function spaces CesΦ := CLΦ, see also Remark 6 for a more accurate discussion. We extend these results to the class of symmetric spaces and we get a full characterization of order continuous points in abstract Ces`aro function spaces CX. In the last Section 4, we show applications of our characterization for particular cases of symmetric spaces. Some results from this section were proved earlier directly for X = Lp by Hassard and Hussein in [12], Shiue in [31], for X = L∞ by Zaanen in [32] and for X = LΦ by Kiwerski and Kolwicz in [15]. 2010 Mathematics Subject Classification: 46B20, 46B42, 46E30. Key words and phrases: Ces`aro function spaces; Ces`aro-Orlicz function spaces; Ces`aro-Lorentz function spaces; Ces`aro-Marcinkiewicz function spaces; order continuity; local structure of a separated point. 1 C : f 7→ Cf (x) := f (t)dt. 1 xZ x 0 2. Notation and preliminaries Denote by m the Lebesgue measure on I and by L0 = L0(I) the space of all classes of real- valued Lebesgue measurable functions defined on I, where I = [0, 1] or I = [0, ∞). Through all the paper when we pick a subset A ⊂ I we assume that A is a Lebesgue measurable set. For a subset A ⊂ I we define the essential infimum infess(A) of A as follows infess(A) := inf{ǫ ≥ 0 : m([0, ǫ) ∩ A) = 0}. A Banach space E = (E, k · k) is said to be a Banach ideal space on I (we write E[0, 1] or E[0, ∞)) if E is a linear subspace of L0(I) and satisfies the condition that if g ∈ E, f ∈ L0 and f ≤ g a.e. on I then f ∈ E and kf k ≤ kgk. Sometimes we write k·kE to be sure in which space the norm has been taken. We say that E is non-trivial if E 6= {0}. For f ∈ L0 we define support as supp(f ) := {t ∈ I : f (t) 6= 0}. Recall that the support supp(X) of Banach ideal space X is defined as measurable subset of I such that m(supp(f ) \ supp(X)) = 0 for every f ∈ X, and for every U ⊂ supp(X) of finite positive measure we have χU ∈ X. For two Banach spaces E and F on I the symbol E ֒→ F means that the embedding E ⊂ F is continuous, i.e. there exists a constant A > 0 (we call it the embedding constant) such that kf kF ≤ A kf kE for all f ∈ E. Recall that for two Banach ideal spaces E and F the embedding E ⊂ F is always continuous. Moreover, E = F (resp. E ≡ F ) means that the spaces are the same as the sets and the norms are equivalent (resp. equal). By E ≃ F we denote the fact that the Banach spaces E and F are isomorphic. For function f ∈ L0 we define distribution functions as df (λ) := m({t ∈ I : f (t) > λ}) for λ ≥ 0. We say that two functions f, g ∈ L0(I) are equimeasurable when they have the same distribution functions, i.e. df ≡ dy. By a symmetric function space (symmetric Banach function space or rearrangement invariant Banach function space) on I we mean a Banach ideal space E = (E, k·kE) with the additional property that for any two equimeasurable functions f, g ∈ L0(I) if f ∈ E then g ∈ E and kf kE = kgkE (we also accept the convention to write "symmetric space" within the meaning of "symmetric function space" because in this paper we focus only on the consideration of this case). In particular, kf kE = kf ∗kE, where f ∗(t) := inf{λ > 0 : df (λ) ≤ t} for t ≥ 0. Note that, if a symmetric function space E on I is non-trivial then supp(E) = I. For the theory of symmetric spaces the reader is referred to [4] and [19]. For a symmetric function space E on I its fundamental function ϕE is defined by the formula for 0 < t ∈ I. Writing ϕE(0+) or ϕE(∞) we understand lim t→0+ ϕE(t) or lim t→∞ ϕE(t), respectively. ϕE(t) :=(cid:13)(cid:13)χ[0,t](cid:13)(cid:13)E , A point f ∈ X is said to have an order continuous norm (f is an OC-point) if for any sequence (fn) ⊂ X with 0 ≤ fn ≤ f and fn → 0 a.e. on I, we have kfnkX → 0. By Xa we denote the subspace of all order continuous elements of X. A Banach ideal space X is called order continuous (we write X ∈ (OC) for short) if every element of X has an order continuous norm, if Xa = X. It is worth to notice that in case of Banach ideal spaces on I, f ∈ Xa if and i.e. only if kf χAnkX → 0 for any sequence (An) satisfying An ↓ ∅ (that is, decreasing sequence of Lebesgue measurable sets An ⊂ I with intersection of measure zero, see [4, Proposition 3.5, p. 15]). The subspace Xa is always closed, see [4, Th. 3.8, p. 16]. Characterization of order continuity given in Theorem A (iii) and (iv) is well known. Theorem A. (i) ([8, Lemma 2.6]) Let E be symmetric space. Then f ∈ E is a point of order continuity if and only if f ∗ is also. 2 (ii) ([3, Lemma 2.5], [8, Lemma 2.5], cf. [19]) Let E be symmetric space. Then f ∈ Ea if and only if for n → ∞. In particular if f is a point of order continuity then f ∗(∞) = 0. (cid:13)(cid:13)(cid:13) f ∗χ[0, 1 → 0 n )(cid:13)(cid:13)(cid:13)E and (cid:13)(cid:13)f ∗χ[n,∞)(cid:13)(cid:13)E → 0 (iii) ([27]) A Banach ideal space E is order continuous if and only if E contains no isomorphic copy of l∞. (iv) ([4, Th. 5.5, p. 27]) A Banach ideal space E is order continuous if and only if is separable. (cid:3) Let X be a Banach ideal space. The closure in X of the set of simple functions is denoted by Xb. It is well known, that the subspace Xb is the closure in X of the set of bounded functions supported in sets of finite measure, cf. [4, Prop. 3.10. p. 17]. Of course, Xb is always non-trivial for non-trivial space X, moreover the subspace Xb is an order ideal of X. We always have Xa ⊂ Xb, see [4, Th. 3.11, p. 18], and the inclusion {0} 6= Xa ⊂ Xb may be proper, cf. [4, Ex. 3, p. 30]. The fact, that this example is based on non-symmetric construction is essential. In the symmetric case we have only two possibilities, more precisely Xa = {0} or Xa = Xb. The subspaces Xa and Xb coincide if and only if the characteriztic functions of the sets of finite measure all have absolutely continuous norms, cf. [4, Th. 3.13, p. 19], [20, Prop. 2.2] and Theorem B. Next theorem describes the opposite extreme, when Xa is trivial. Theorem B. Let E be a symmetric space. The following conditions are equivalent: (i) Ea = {0}, (ii) E ֒→ L∞, (iii) Ea 6= Eb, (iv) ϕE(0+) > 0. In particular, if I = [0, 1] then condition (ii) is equivalent to the statement E = L∞. Proof. Obviously, E ֒→ L∞ is equivalent to E = L∞ if I = [0, 1] because the embedding L∞ ֒→ E holds for each symmetric space E on I = [0, 1], see [4, Corollary 6.7, p. 78]. Equivalence of conditions (i), (iii) and (iv) follows immediately from [4, Th. 5.5, p. 67] and our discussion preceding this theorem. For the implication (ii) ⇒ (iv) it is enough to observe, that ϕE(0+) = lim t→0+(cid:13)(cid:13)χ[0,t](cid:13)(cid:13)E ≥ A lim t→0+(cid:13)(cid:13)χ[0,t](cid:13)(cid:13)L∞ = A > 0, where A is the embedding constant. Finally, suppose ϕE(0+) > 0 and E * L∞. This means, that there is an unbounded element f ∈ E. Then the f ∗ is unbounded in zero and f ∗ ∈ E by the symmetry of E. Therefore we can find a sequence (tn) ⊂ I, tn → 0+ as n → ∞, with f ∗(tn) > n. We have 1 n as n → ∞ contrary to the assumption that ϕE(0+) > 0. (cid:13)(cid:13)χ[0,tn](cid:13)(cid:13)E ≤(cid:13)(cid:13)(cid:13)(cid:13) f ∗χ[0,tn](cid:13)(cid:13)(cid:13)(cid:13)E 1 n ≤ kf ∗kE → 0, (cid:3) For s > 0 the dilation operator Ds is defined, on L0(I), by Dsf (t) := f ( t s )χI ( t s ) = f ( t s )χ[0,min{1,s})(t), for t ∈ I, is bounded in any symmetric space E on I and kDskE→E ≤ max{1, s} (see [4, p. 148]). The lower and upper Boyd indices of E are defined by p(E) := lim s→∞ lns , ln kDskE→E 3 q(E) := lim s→0+ lns ln kDskE→E . They satisfy the inequalities 1 ≤ p(E) ≤ q(E) ≤ ∞. For more details see [6], [5] and [28]. Next result will be used in Section 3. Theorem C. ([26, Prop. 2.b.3, p. 132]) Let E be a symmetric function space. Then for every 1 ≤ p < p(E) and q(E) < q ≤ ∞ we have Lp ∩ Lq ֒→ E ֒→ Lp + Lq. In particular, if we take q = ∞ then for every 1 ≤ p < p(E) we get Lp ∩ L∞ ֒→ E ֒→ Lp + L∞. The continuous Ces`aro operator C : L0(I) → L0(I) is defined by (cid:3) Cf (x) := f (t)dt, 1 xZ x 0 for 0 < x ∈ I. For a Banach ideal space X on I we define an abstract Ces`aro space CX = CX(I) by CX := {f ∈ L0(I) : Cf ∈ X}, with the norm kf kCX = kCf kX (see [9], [10], [22], [23], [24]). Let us note that for non- symmetric space X the space CX need not have a weak unit even if X has it (see [22, Example 2]), so in general supp(CX) ⊂ supp(X). Of course, if we assume that C : X → X then supp(CX) = supp(X), cf. [22, Remark 1]. Moreover, if CX is non-trivial and X is symmetric space on I then supp(CX) = I, see Lemma 2. Let us mention the important result about boundedness of the Ces`aro operator. Theorem D. [17, p. 127] For any symmetric space E on I the operator C : E → E is bounded if and only if the lower Boyd index satisfies p(E) > 1. (cid:3) Note that if (X, k·kX) is a Banach ideal space then the assumption C : X → X is equivalent to the statement C is bounded. In fact, if C : X → X, then X ֒→ CX. This means that there is M > 0 with kf kCX = kCf kX ≤ M kf kX for all f ∈ X, i.e. C is bounded. On the other hand, if C is bounded then for all f ∈ X we have kf kCX = kCf kX ≤ kCk kf kX = kCk kf kX , which means that X ⊂ CX and C : X → X. However, it may happen that X 6֒→ CX (see [10, Proposition 2.1], [16, Example 14]). Moreover, C : CX → X is always bounded (from the definition of CX) and CX is so-called optimal domain of C for X (cf. [10] and [23]). The immediate consequence of Theorem D and the above discussion about boundedness of the Ces`aro operator is a next Theorem E. For any symmetric space E the embedding E ֒→ CE holds if and only if p(E) > 1. In particular, if p(E) > 1 then the space CE is non-trivial. (cid:3) It is worth to notice the following useful observation. Fact 1. (cf. [4, Prop. 3.2, p. 52]) For every function f : I → R we have the following inequalities Cf ≤ Cf ≤ C f ≤ C(f ∗) := f ∗∗. Moreover, (Cf )∗ ≤ C(f ∗). (cid:3) Let us note some basic fact about the non-triviality of the space CX when X is symmetric function space. Lemma 2. Let X be symmetric space on I. If I = [0, 1] then CX 6= {0}. If I = [0, ∞) then the following conditions are equivalent 4 (i) CX 6= {0}, (ii) 1 (iii) 1 x χ(t0,∞) ∈ X for some t0 > 0, x χ(t,∞) ∈ X for all t > 0. Proof. We give an easy proof for the convenience of the reader. Assume that X is symmetric space on I. Because L∞[0, 1] ֒→ X[0, 1], so we have that {0} 6= CL∞[0, 1] := Ces∞[0, 1] ֒→ CX[0, 1], see [4, Corollary 6.7, p. 78] and [15, Remark]. This means that CX[0, 1] is always non-trivial. Equivalence of conditions (i) and (ii) follows from [22, Theorem 1 (a)]. Of course, (iii) implies (ii). Therefore, we will only show that (ii) implies (iii). In fact, take 0 < t < t0 (if t0 < t there is nothing to prove). Then χ(t,t0] + x χ(t0,∞) ∈ X from the assumption and 1 χ(t,∞) = because 1 1 x 1 x 1 x χ(t0,∞) ∈ X, x χ(t,t0] ∈ X since X is symmetric. (cid:3) 3. On the OC-points in Ces`aro function spaces CX Theorem 3. Let X be symmetric space such that Ces`aro operator is bounded on X. Then X ∈ (OC) if and only if CX ∈ (OC). The proof of the above theorem will be given at the end of this section as a consequence of characterization of subspace (CX)a. We will begin with some observations and examples. Remark 4. The implication: X is order continuous then CX is also, follows easily (in fact from the definition) without any assumption, see [24, Lemma 1 (a)]. Moreover, in the case of I = [0, 1] Theorem 3 has been proved already, see [15, Proposition 2]. Leaving the assumption of symmetry Theorem 3 ceases to be true, as the following example shows. Example 5. (see [24, Example 1]) If X = L2[0, 1 but CX = Ces2[0, 1] ∈ (OC), see Proposition 20. 4 ]⊕L∞[ 1 4 , 1 2 ]⊕L2[ 1 2 , 1], then of course X /∈ (OC) Remark 6. Curbera and Ricker showed in [9, Prop. 3.1 (c)], using the methods of vector measures and integral representation, that (adapting into our notation) C(Xa) = CXa whenever X is a symmetric space on [0, 1] and X 6= L∞. The last condition can be expressed equivalently in several ways, cf. Theorem B. Similar results are also known in the case of I = [0, ∞) but only for a certain class of the Orlicz spaces. Namely, suppose that Φ is an Orlicz function with Φ < ∞, i.e. bΦ = ∞ and LΦ is the Orlicz space generated by function Φ (cf. the definitions in the subsection 4.2 and references therein). From [15, Theorem 5 (i)] we have the following equalities C(LΦ a ) = {f : C f ∈ Lϕ a } = {f : IΦ(λC f ) < ∞ for all λ(f ) := λ > 0} = {f : ρΦ(λf ) < ∞ for all λ(f ) := λ > 0} = CΦ, where CΦ is the space of all order continuous elements of space CLΦ := CesΦ, so (CLΦ)a = (CesΦ)a = CΦ = C(LΦ a ). Example 7. The above observation leads naturally to the question - is it true, that C(Xa) = (CX)a for any symmetric space X? Moment of thought show us that in general that is not the case. Taking L∞ we observe that C(L∞ a ) = C{0} = {0} (Ces∞)a := (CL∞)a 6= {0}, 5 since elementary computation shows that, for example χ[ 1 2 ,1) ∈ (CL∞)a. However, in Theorem 16 we will give an answer when this representation is possible and describe the space (CX)a in the other cases. We start with the following lemma. Lemma 8. Let X be symmetric space such that CX 6= {0} and let A ⊂ I. If infess(A) > 0 and m(A) < ∞ then χA ∈ (CX)a. Proof. At the beginning we will show that χ(a,b) ∈ (CX)a, when 0 < a < b < ∞. Let (An) ⊂ I be sequence of measurable sets with An ↓ ∅. Put Bn = An ∩ (a, b). Simply m(Bn) → 0 when n → ∞. We have χBn(t)dt CχBn(x) = = ≤ 1 xZ x 0 χ(a,b)(x) x m(Bn) b Z(a,min{x,b})∩An dt + m(Bn) x χ(b,∞)(x) χ(a,b)(x) + χ(b,∞)(x). m(Bn) x < ∞. Therefore m(Bn) x ) → 0 χ(b,∞)(cid:13)(cid:13)(cid:13)(cid:13)X as n → ∞. From Lemma 2, assumption CX 6= {0} implies (cid:13)(cid:13) (cid:13)(cid:13)χ(a,b)χAn(cid:13)(cid:13)CX = kχBnkCX = kCχBnkX ≤(cid:13)(cid:13)(cid:13)(cid:13) = m(Bn)( 1 b m(Bn) 1 x χ(b,∞)(cid:13)(cid:13)X χ(a,b)(cid:13)(cid:13)(cid:13)(cid:13)X +(cid:13)(cid:13)(cid:13)(cid:13) χ(b,∞)(cid:13)(cid:13)(cid:13)(cid:13)X 1 x b (cid:13)(cid:13)χ(a,b)(cid:13)(cid:13)X +(cid:13)(cid:13)(cid:13)(cid:13) Since An was chosen arbitrary, we get χ(a,b) ∈ (CX)a. Let A ⊂ I be a subset with infess(A) = a > 0 and m(A) < ∞. Take a sequence of measurable sets (An) ⊂ I with An ↓ ∅. Denote Bn = An ∩ A. Then, like above, m(Bn) → 0 when n → ∞. Note that Since χ[a,m(Bn)+a] ≤ χ[a,m(A)+a] and χ[a,m(Bn)+a](t) ↓ 0 as n → ∞ for a.e. t ∈ I thus, from a previous case, we obtain C(χBn) ≤ C(χ[a,m(Bn)+a]). kχAχAnkCX = kC(χBn)kX ≤(cid:13)(cid:13)Cχ[a,m(Bn)+a](cid:13)(cid:13)X =(cid:13)(cid:13)χ[a,m(Bn)+a](cid:13)(cid:13)CX → 0 as n → ∞. Thus χA ∈ (CX)a. (cid:3) Corollary 9. Let X be symmetric space such that CX 6= {0}. If f ∈ (CX)b and infess(A) > 0, where A ⊂ I then f χA ∈ (CX)a. Proof. Let f ∈ (CX)b, (fn) be sequence of simple functions with lim fn = f and A ⊂ I be n→∞ like in the assumptions above. Of course, fnχA → f χA. Moreover, m(supp(fn) ∩ A) < ∞ thus fnχA ∈ (CX)a from Lemma 8. Since C : CX → X is always bounded and (CX)a is closed hence f χA ∈ (CX)a. (cid:3) We will use the notation ∆0 = ∆0(E) := {f ∈ E : lim t→0 space. Cf (t) = 0}, where E is a Banach ideal Lemma 10. Let X be symmetric space. Then (CX)b ∩ ∆0 ⊂ (CX)a. 6 Proof. If CX = {0} there is nothing to prove. Assume that CX 6= {0} and take 0 ≤ f ∈ (CX)b ∩ ∆0. Let (An) ⊂ I be arbitrary sequence of measurable subsets with An ↓ ∅. Take ǫ > 0. Since Cf (0+) = 0 there exist δ > 0 such that supt∈[0,δ) Cf (t) < ǫ. From Corollary 9 we get f χ[δ,∞) ∈ (CX)a. Therefore there exist nδ ∈ N with for every n ≥ nδ. Without loss of generality suppose that δ ≤ 1. Then C(f χAn∩[0,δ)) ≤ sup{Cf (s) : s ∈ [0, δ)}(χ[0,1) + χ[1,∞)) ≤ ǫ(χ[0,1) + 1 t χ[1,∞)), (cid:13)(cid:13)f χ[δ,∞)χAn(cid:13)(cid:13)CX ≤ ǫ, 1 t for n ∈ N. Moreover, (cid:13)(cid:13) have 1 t χ[1,∞)(cid:13)(cid:13)X < ∞ in view of our assumption CX 6= {0} and Lemma 2. We kf χAnkCX ≤(cid:13)(cid:13)(cid:13) f χ[δ,∞)χAn(cid:13)(cid:13)(cid:13)CX ≤ ǫ + ǫ(cid:13)(cid:13)(cid:13)(cid:13) ≤ ǫ(1 + ϕX (1) +(cid:13)(cid:13)(cid:13)(cid:13) +(cid:13)(cid:13)(cid:13) f χ[0,δ)χAn(cid:13)(cid:13)(cid:13)CX χ[1,∞)(cid:13)(cid:13)(cid:13)(cid:13)X χ[1,∞)(cid:13)(cid:13)(cid:13)(cid:13)X χ[0,1) + 1 t 1 t ), for n ≥ nδ. Since ǫ > 0 was arbitrary, we conclude that which ends the proof. (cid:3) kf χAnkCX → 0 as n → ∞, Lemma 11. Let X be Banach ideal function space, then C(Xa) ⊂ (CX)a. Proof. Suppose 0 ≤ f ∈ C(Xa). Let (An) be arbitrary sequence of measurable subsets of I with An ↓ ∅. First, note that f (t)χAn(t) ↓ 0 as n → ∞ for a.e. t ∈ I thus, from Lebesgue dominated convergence theorem, we obtain that C(f χAn)(t) ↓ 0 as n → ∞ for a.e. t ∈ I. Since Cf ∈ Xa and C(f χAn) ≤ Cf for every n ∈ N we have kf χAnkCX = kC(f χAn)kX → 0 as n → ∞, i.e. f ∈ (CX)a. (cid:3) Lemma 12. Let X be symmetric function space such that Ces`aro operator is bounded on X and Xa = Xb. Then (CX)b ⊂ C(Xa). Proof. We divide the proof into three steps. First, we will show that χ[0,a] ∈ C(Xa) for every 0 < a < ∞. Fix a > 0 and note that C(χ[0,a])(t) = χ[0,a](t) + a t χ[a,∞](t), for 0 < t ∈ I. Since Xa = Xb thus χ[0,a] ∈ Xa and we only need to show that f (t) := 1 t χ[a,∞](t) ∈ Xa. Let (An) be sequence of measurable subsets of I with An ↓ ∅. Without loss of generality we can assume that An ⊂ supp(f ) = [a, ∞) for every n ∈ N. Let δ > 0 be arbitrary. From δ for all n ≥ nδ. From Theorem C and our assumption that Ces`aro operator is bounded there exist p > 1 with Lp ∩ L∞ ֒→ X. Since Lp ∈ (OC), we can find n′ Lemma 8 we have χ[a,a+δ] ∈ Xa therefore there exist nδ ∈ N such that (cid:13)(cid:13) a χ[a,a+δ]χAn(cid:13)(cid:13)X ≤ 1 δ ∈ N satisfying 1 (cid:13)(cid:13)(cid:13)(cid:13) 1 t χ[δ,∞]χAn(cid:13)(cid:13)(cid:13)(cid:13)Lp 7 ≤ 1 δ , for every n ≥ n′ δ. Let n ≥ max{nδ, n′ δ} we have 1 a kf χAnkX ≤(cid:13)(cid:13)(cid:13)(cid:13) ≤ 1 δ 1 δ 1 δ = ≤ 1 t 1 t χ[δ,∞]χAn(cid:13)(cid:13)(cid:13)(cid:13)X χ[a,a+δ]χAn(cid:13)(cid:13)(cid:13)(cid:13)X +(cid:13)(cid:13)(cid:13)(cid:13) χ[δ,∞]χAn(cid:13)(cid:13)(cid:13)(cid:13)L∞∩Lp + D(cid:13)(cid:13)(cid:13)(cid:13) χ[δ,∞]χAn(cid:13)(cid:13)(cid:13)(cid:13)Lp ,(cid:13)(cid:13)(cid:13)(cid:13) 1 t (p − 1)− 1 + D max{ + D max{ }, } p 1 δ 1 δ , δ1− 1 p where D > 0 is inclusion constant. Since δ > 0 was arbitrary, we obtain kf χAnkX → 0 as n → ∞. Consequently, f ∈ Xa and hence χ[0,a] ∈ C(Xa). Now let A ⊂ I be arbitrary set of finite measure. From Fact 1 we have C(χA) ≤ C(χ[0,m(A)]). From a previous step of proof we also know that C(χ[0,m(A)]) ∈ Xa thus from ideal property of Xa we get C(χA) ∈ Xa, i.e. χA ∈ C(Xa). Finally, let f ∈ (CX)b and (fn) be a sequence of simple functions with lim n→∞ fn = f pointwise. We already know that Cfn ∈ Xa for every n ∈ N. From continuity of Ces`aro operator C : CX → X we have that Cfn → Cf . Therefore Cf ∈ Xa because Xa is closed and the proof is complete. (cid:3) Remark 13. We can prove more, namely: if X is a symmetric function space with Xa = Xb then (CX)b ⊂ C(Xa) if and only if 1 t χ(λ0,∞)(t) ∈ Xa for some λ0 > 0. In above proof we in fact proved the sufficiency. The necessity is even simpler, if 1 t χ(1,∞)(t) /∈ Xa then C(χ(0,1))(t) /∈ Xa, i.e. χ(0,1)(t) /∈ C(Xa), but of course χ(0,1)(t) ∈ (CX)b. It follows from the proof of remaining lemma that if C : X → X and Xa = Xb then 1 t χ[1,∞) ∈ Xa. Is also worth noting that in the case of I = [0, 1] the assumption X 6= L∞ is equivalent to 1 t χ(λ,∞)(t) ∈ Xa for all 0 < λ < 1, see Theorem B. Lemma 14. Let X be symmetric space with Xa = {0}. Then (CX)a ⊂ ∆0. Proof. If CX = {0} then inclusion is trivial, so we can assume that CX 6= {0}. Take 0 ≤ f /∈ ∆0. Then lim sup Cf (t) = δ > 0. t→0+ We can find a sequence (an) ⊂ (0, 1) such that an → 0 and Cf (an) > δ is continuous function, therefore there exist open neighborhood Un of an with inf t∈Un for n ∈ N. We can chose a subsequence nk ∈ N such that Unk ∩ [0, 1 2 for every n ∈ N. Since Cf {Cf (t)} ≥ δ 4 k ) 6= ∅ for k ∈ N. Now we 8 have f χ[0, 1 C(f χ[0, 1 (cid:13)(cid:13)(cid:13) k )(cid:13)(cid:13)(cid:13)CX =(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:13)(cid:13)(cid:13) Cf (t)χ[0, 1 Cf (t)χ[0, 1 δ 4 χUnk ∩[0, 1 k ) + k ))(cid:13)(cid:13)(cid:13)X k )(cid:13)(cid:13)(cid:13)X k )(cid:13)(cid:13)(cid:13)(cid:13)X Cf ( 1 k ) t χ[ 1 k ,∞)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)X ≥ δ 4 ϕX (0+), (cid:3) which shows that f /∈ (CX)a. Lemma 15. Let X be symmetric space with Xa 6= {0}. Then (CX)a = (CX)b. Proof. Take δ > 0. Let (fn) ⊂ (CX)b be a sequence chosen so that fn → f almost everywhere. Using Corollary 9 we conclude that f χ[δ,∞) ∈ (CX)a. Since C(Xa) ⊂ (CX)a and Xa 6= {0} we have Observe that χ[0,δ) ∈ C(Xa) ⊂ (CX)a. fnχ[0,δ) ≤ f ∗ n(0)χ[0,δ), thus fnχ[0,δ) ∈ (CX)a using the ideal property of (CX)a. From continuity of C : CX → X we have fnχ[0,δ) → f χ[0,δ) in CX. This means that f χ[0,δ) ∈ (CX)a because (CX)a is closed. (cid:3) Theorem 16. Let X be symmetric function space. Then one of following holds: (i) (CX)a = (CX)b if Xa 6= {0}, (ii) (CX)a = (CX)b ∩ ∆0 if Xa = {0}. In particular, if Ces`aro operator is bounded on X and Xa 6= {0} then (CX)a = C(Xa). Proof. Suppose Xa 6= {0}. Then Xa = Xb from Theorem B. This is the case of Lemma 15 so we have (CX)a = (CX)b. If we additionally assume that the Ces`aro operator C is bounded on X, using Lemma 11 and Lemma 12 we have C(Xa) ⊂ (CX)a = (CX)b ⊂ C(Xa). Therefore (CX)a = C(Xa). Assume now that Xa = {0}. From Lemma 14 we have (CX)a ⊂ ∆0. Since we always have that (CX)a ⊂ (CX)b hence (CX)a ⊂ (CX)b ∩ ∆0. Combining this with Lemma 10 we obtain (CX)b ∩ ∆0 ⊂ (CX)a ⊂ (CX)b ∩ ∆0, thus (CX)a = (CX)b ∩ ∆0. (cid:3) Remark 17. The previous theorem can be formulated in a more concise form if we assume that the Ces`aro operator is bounded on symmetric space X. In this case (CX)a = C(Xa) + (CX)b ∩ ∆0. Moreover, it follows from Remark 13 that if X is symmetric space and Xa is non-trivial then (CX)a = C(Xa) if and only if 1 t χ(λ0,∞)(t) ∈ Xa for some λ0 > 0. Additionally, as consequence of Remark 13, we have (CX)a = C(Xa) if and only if X 6= L∞ in the case of I = [0, 1]. Lemma 18. Let X be a symmetric space such that the Ces`aro operator is bounded on X. If f ∈ X \ Xa then f ∗ ∈ CX \ (CX)a. 9 Proof. Take an element f ∈ X \ Xa. From a symmetry of X we get f ∗ ∈ X and C(f ∗) ∈ X since the operator C is bounded on X. Now we have to consider two cases: (i) Xa 6= {0}. Then, from Theorem 16, (CX)a = C(Xa). From Theorem A (i) f ∗ /∈ Xa. From ideal property of Xa and since C(f ∗) ≥ f ∗, we have that C(f ∗) /∈ Xa, i.e. f ∗ /∈ C(Xa) = (CX)a. (ii) Xa = {0}. Observe that 0 6= f , since 0 ∈ Xa thus lim sup t→0+ C(f ∗)(t) ≥ lim sup t→0+ f ∗(t) > 0. From Theorem 16 we obtain f ∗ /∈ ∆0 ∩ (CX)b = (CX)a. (cid:3) It is time to give proof of the Theorem 3 announced at the beginning of this section. Proof of Theorem 3. Necessity. Let X be a symmetric space such that Ces`aro operator is bounded on X and suppose X ∈ (OC). Then X = Xa and from Theorem 16 we get (CX)a = C(Xa) = CX, which means that CX ∈ (OC). Sufficiency. If X /∈ (OC) then there exist an element f /∈ Xa. From Lemma 18 f ∗ /∈ (CX)a and consequently CX /∈ (OC) which completes the proof. (cid:3) 4. Applications Although each of the following spaces belongs to the class of symmetric spaces, in this special cases, the criteria for OC-points become more specified. 4.1. The Ces`aro function spaces Cesp. Remark 19. We have the following characterization of the closure of the set of simple functions in the space L∞ and (L∞[0, 1])b = L∞[0, 1], (L∞[0, ∞))b = {f ∈ L∞[0, ∞) : lim t→∞ f (t) = 0}. Of course, (L∞)a = {0} and p(L∞) > 1. Let us define a set ∆∞ = ∆∞(E) := {f ∈ E : Cf (t) = 0} for E being a Banach ideal space on [0, ∞). With this notation, we get the lim t→∞ following reformulation of Theorem 16 (ii): (Ces∞)a =  ∆0 = {f ∈ Ces∞ : lim t→0+ ∆0 ∩ ∆∞ = {f ∈ Ces∞ : Cf (t) = 0} if I = [0, 1] lim t→0+,∞ Cf (t) = 0} if I = [0, ∞). The space Ces∞[0, 1] is known as Korenblyum-Krein-Levin space K and it is also known that Ka = {f ∈ K : lim h→0+ f dm = 0} = ∆0(K), 1 hZ h 0 see [32, pp. 469-471]. Thus, the above characterization is a generalization of this classical result. It is worth to mention that Ces1[0, 1] ≡ L1(ln(1/t)). From Lebesgue dominated convergence theorem we obtain that L1(ln(1/t) is order continuous. Moreover, Ces1[0, ∞) = {0}, see [2, Th. 3.1 (a)]. Since the spaces Lp are order continuous for 1 ≤ p < ∞ and L∞ a = {0}, so as a consequence of Theorem 16 and Remark 19 we obtain well known result (cf. [2, Th. 3.1 (b)]). 10 Proposition 20. We have the following characterization (Cesp)a =( Cesp {f ∈ Ces∞ : lim t→0+,∞ when 1 ≤ p < ∞ Cf (t) = 0} when p = ∞. In particular, the Ces`aro function space Cesp is order continuous if and only if 1 ≤ p < ∞. 4.2. The Ces`aro-Orlicz function spaces CesΦ. A function Φ : [0, ∞) → [0, ∞] is called an Orlicz function if: (i) Φ is convex, (ii) Φ(0) = 0, Φ(∞) = ∞, (iii) Φ is neither identically equal to zero nor infinity on (0, ∞). For more information about Orlicz functions see [7] and [18]. Let us denote by and aΦ = sup{u ≥ 0 : Φ(u) = 0}, bΦ = sup{u > 0 : Φ(u) < ∞}. We have, aΦ < ∞ and bΦ > 0, since an Orlicz function is neither identically equal to zero nor infinity on [0, ∞). The function ϕ is continuous and non-decreasing on [0, bΦ) and is strictly increasing on [aΦ, bΦ). We say an Orlicz function Φ satisfies the condition ∆2 for large arguments (Φ ∈ ∆2(∞) for short) if there exists K > 0 and u0 > 0 such that Φ(u0) < ∞ and Φ(2u) ≤ KΦ(u) for all u ∈ [u0, ∞). Similarly, we can define the condition ∆2 for small, with Φ(u0) > 0 (Φ ∈ ∆2(0)) or for all arguments (Φ ∈ ∆2(R+)). These conditions play a crucial role in the theory of Orlicz spaces, see [7], [18], [29] and [30]. The Orlicz function space LΦ = LΦ(I) generated by an Orlicz function Φ is defined by LΦ := {f ∈ L0(I) : IΦ(f /λ) < ∞ for some λ = λ(f ) > 0}, where IΦ(f ) = RI Φ(f (t))dt is a convex modular (for the theory of Orlicz spaces and modular spaces see [29] and [30]). The space LΦ is a Banach ideal function space with the Luxemburg- Nakano norm kf kΦ = inf{λ > 0 : IΦ(f /λ) ≤ 1}. Let us recall that LΦ b = {f ∈ CesΦ : ∀λ>0∃sλ∈S0 such that IΦ(λ(f − s)) < ∞}, where S0 is the ideal of bounded functions with support of finite measure. It is also easy to see that b = {f ∈ LΦ : IΦ(λf ) < ∞ for all λ > 0} = LΦ LΦ a , when bΦ = ∞, cf. [11], [29, Th. 3.3 (a), p. 17] and Theorem B. The Ces`aro-Orlicz function space CesΦ = CesΦ(I) is defined by CesΦ(I) := CLΦ(I). Conse- quently, the norm in the space CesΦ is given by the formula kf kCesΦ = inf{λ > 0 : ρΦ(f /λ) ≤ 1}, where ρΦ(f ) = IΦ(Cf ) is a convex modular. Note that CesΦ[0, 1] 6= {0} for any Orlicz function Φ, see Lemma 2, [15, Remark 4]. Moreover, for the equivalent conditions to the non-triviality of CesΦ[0, ∞) see [15, Proposition 3]. Lemma 21. Let Φ be an Orlicz function. Then (CesΦ)b = {f ∈ CesΦ : ∀λ>0∃sλ∈S0 ρΦ( 11 f − sλ λ ) < ∞}. Proof. The idea of this proof comes from [11, Th. 3.1]. Without loss of generality, we can consider only the case when the Ces`aro-Orlicz function space is non-trivial. Like in mentioned article we use the notation H(S0) = {f ∈ CesΦ : ∀λ>0∃sλ∈S0 ρΦ( f − sλ λ ) < ∞}. We claim that: (i) H(S0) is closed, (ii) H(S0) ⊂ clS0, where by clS0 we mean the closure of S0 in CesΦ. Take f ∈ clH(S0) and fix λ > 0. There exist a sequence (fn) ⊂ H(S0) with fn → f . From [29, Th. 1.3, p. 8] there exist n0 ∈ N such that Let s ∈ S0 be such that Since modular ρΦ is convex, we have ρΦ( f − fn0 λ/2 ) < ∞. ρΦ( fn0 − s λ/2 ) < ∞. ρΦ( f − s λ ) = ρΦ(2 (f − fn0 + fn0 − s) 2λ ) ≤ 1 2 (ρΦ( f − fn0 λ/2 ) + ρΦ( fn0 − s λ/2 )) < ∞, and (i) follows. Pick f ∈ H(S0)+ and λ > 0. From the definition of H(S0) there exist 0 ≤ sλ ≤ f such that ρΦ( f − sλ λ ) < ∞. Take a sequence (yn) ⊂ S0 with yn ↑ f − sλ a.e. on I. Since 0 ↓ f − sλ − yn ≤ f − sλ a.e. on I thus using Lebesgue dominated convergence theorem we get C(f − sλ − yn) ↓ 0 a.e. on I. Because L1 ∈ (OC) and for all n ∈ N, so ρΦ( f − sλ − yn ) ≤ ρΦ( f − sλ λ ) < ∞, λ ρΦ( f − sλ − yn λ ) → 0. Therefore there exist n0 ∈ N with ρΦ( f − sλ − yn0 λ ) < λ, i.e. dist(f, S0) ≤ λ, see [11]. Since λ > 0 was arbitrary so f ∈ clS0. This completes the proof of part (ii). It is clear that clS0 = (CesΦ)b and S0 ⊂ H(S0). Combining this observation with (i) and (ii) we conclude that and i.e. (CesΦ)b = H(S0). H(S0) ⊂ clS0 = (CesΦ)b, (CesΦ)b = clS0 ⊂ clH(S0) = H(S0), (cid:3) Next proposition is a generalisation of Theorem 5, Theorem 7 and Corollary 8 from [15]. 12 Proposition 22. Let Φ be an Orlicz function. Then (CesΦ)a =  {f ∈ CesΦ : ρΦ(λf ) < ∞ for all λ > 0} {f ∈ CesΦ : ∀λ>0∃sλ∈S0 ρΦ(λ(f − s)) < ∞ and lim t→0+ {f ∈ CesΦ : Cf (t) = 0} lim t→0+,∞ if bΦ = ∞ Cf (t) = 0} if aΦ = 0, bΦ < ∞ if aΦ > 0, bΦ < ∞ Moreover, if p(LΦ) > 1 then the Ces`aro-Orlicz function space CesΦ is order continuous if and only if ϕ ∈ ∆2 (i.e. if p(LΦ) > 1 then CesΦ ∈ (OC) if and only if LΦ ∈ (OC)). Proof. Suppose CesΦ is non-trivial and (i) bΦ = ∞. Then there exist λ0 > 0 with IΦ( η0 that this inequality holds for all η > 0. There exist γ > 0 such that IΦ( 1 Indeed, taking γ = λ0 η0 we have x χ(λ0,∞)) < ∞ for some η0 > 0. We will show x χ(γ,∞)) < ∞. IΦ( 1 x χ(γ,∞)) =Z ∞ λ0 η0 Φ( 1 x )dx = η−1 0 Z ∞ λ0 Φ( η0 u )du Φ( η0 u χ(λ0,∞)(u))du = η−1 0 IΦ( η0 x χ(λ0,∞)) < ∞. Take η > 0. Then IΦ( η x 0 = η−1 0 Z ∞ χ(λ0,∞)) = ηZ ∞ ≤ ηZ γ ≤ ηZ γ 0 0 0 Φ( Φ( Φ( 1 u η λ0 η λ0 χ ( λ0 η ,∞) (u))du γ )du + ηZ ∞ )du + ηZ ∞ γ Φ( Φ( 1 u 1 u χ( λ0 η ,∞)(x))du χ(γ,∞)(u))du < ∞, which proves the claim. Therefore, 1 x χ(λ0,∞)(x) ∈ LΦ a . Using Remark 13 we get (CesΦ)a = C(LΦ a ) = {f ∈ CesΦ : ρΦ(λf ) < ∞ for all λ > 0}. (ii) aΦ = 0 and bΦ < ∞. In this case LΦ a = {0}, thus from Theorem 16 (ii) and Lemma 21 we get (CesΦ)a = (CesΦ)b ∩ ∆0 = {f ∈ CesΦ : ∀λ>0∃sλ∈S0 ρΦ(λ(f − s)) < ∞ and lim t→0+ Cf (t) = 0}. (iii) aΦ > 0 and bΦ < ∞. Then LΦ = L∞, cf. [29, Ex. 1, p. 98] thus CesΦ = Ces∞ and it is enough to apply Remark 19. Moreover, the Orlicz space LΦ is order continuous if and only if Φ ∈ ∆2, see [29, Remark 1, p. 22]. Therefore, if p(LΦ) > 1 then the Ces`aro-Orlicz function space CesΦ is order continuous if and only if ϕ ∈ ∆2 from Theorem 3. (cid:3) Remark 23. Note that p(LΦ) = αΦ and q(LΦ) = βΦ, where αΦ and βΦ are the so-called lower and upper Orlicz-Matuszewska indices, see [26, Prop. 2.b.5 and Remark 2, p. 140]) and [29], [30] for more information. 4.3. The Ces`aro-Lorentz function spaces CΛϕ. The fundamental function ϕE of every symmetric function space E on I is quasi-concave on I, that is, ϕE(t) = 0 if and only if t = 0, ϕE(t) is increasing on I and ϕE(t)/t is non-increasing for t ∈ (0, m(I)). Moreover, for any quasi-concave function ϕ there is a symmetric function space whose fundamental function is 13 ϕ. The smallest symmetric function space with fundamental function ϕ is called the Lorentz function space Λϕ = Λϕ(I) and is defined as Λϕ(I) := {f ∈ L0(I) :Z m(I) 0 f ∗dϕ < ∞}, with a norm given as kf kΛϕ(I) =R m(I) f ∗dϕ. The fundamental function of a symmetric function space E is not necessary concave but E can be equivalently renormed in such a way that the resulting fundamental function is concave. In this case the Riemann-Stieltjes integral in the definition of the Lorentz function space ΛϕE my be rewritten in the form 0 kf kΛϕE (I) = kf kL∞(I) ϕE(0+) +Z m(I) 0 f ∗(t)φE(t)dt, 0 φE(s)ds. Then we have embedding ΛϕE ֒→ E with embedding constant equal where ϕE(t) =R t 1. Proposition 24. Let ϕ be quasi-concave function. Then (CΛϕ)a = CΛϕ {f ∈ CΛϕ : lim t→∞ {f ∈ (CΛϕ)b : if ϕ(0+) = 0, ϕ(∞) = ∞ if ϕ(0+) = 0, ϕ(∞) < ∞ Cf (t) = 0} if ϕ(0+) > 0, ϕ(∞) = ∞ if ϕ(0+) > 0, ϕ(∞) < ∞ (Cf )∗(t) = 0} lim t→0+ lim Cf (t) = 0} {f ∈ CΛϕ : t→0+,∞   In particular, the Ces`aro-Lorentz function space CΛϕ is order continuous if and only if ϕ(0+) = 0 and ϕ(∞) = ∞ (i.e. CΛϕ ∈ (OC) if and only if Λϕ ∈ (OC)). Proof. We can assume that CΛϕ is non-trivial because if CΛϕ = {0} there is noting to prove. To prove the first part we have to consider the following cases. Suppose (i) ϕ(0+) = 0 and ϕ(∞) = ∞. It follows from [19, Lemma 5.1] and Theorem A (iv) that x χ(1,∞)(x) ∈ (Λϕ)a so using Λϕ ∈ (OC). Moreover, from [21, Corollary 4.13] we have 1 Remark 17 we obtain (CΛϕ)a = C(Λϕ)a = CΛϕ. (ii) ϕ(0+) = 0 and ϕ(∞) < ∞. In this situation Λϕ /∈ (OC) but (Λϕ)a is non-trivial. More precisely, (Λϕ)a = {f ∈ Λϕ : lim t→∞ f ∗(t) = 0}, see [21, Corollary 4.13]. Since C is bounded, using Theorem 16 we get (CΛϕ)a = C(Λϕ)a = {f ∈ CΛϕ : lim t→∞ (Cf )∗(t) = 0}. (iii) ϕ(0+) > 0 and ϕ(∞) = ∞. From Theorem B we get (Λϕ)a = {0} and it is enough to use Theorem 16 (ii). (iv) ϕ(0+) > 0 and ϕ(∞) < ∞. In this case Λϕ = L∞, see [19, p. 108] and we can use Remark 19. The second part is a direct consequence of the foregoing considerations. (cid:3) 4.4. The Ces`aro-Marcinkiewicz function spaces CMϕ. For any quasi-concave function ϕ on I the Marcinkiewicz function space (called also weak Lorentz space) Mϕ = Mϕ(I) is defined as Mϕ(I) := {f ∈ L0(I) : sup t∈I ϕ(t)f ∗∗(t) < ∞}, with a norm kf kMϕ(I) = supt∈I ϕ(t)f ∗∗(t), where f ∗∗(t) := 1 0 f ∗(s)ds is a maximal function. The Marcinkiewicz function space Mϕ is the largest symmetric function space with fundamental function ϕ, i.e. for any symmetric function space E we have E ֒→ MϕE with embedding constant equal 1. t R t 14 Proposition 25. Let ϕ be quasi-concave function. Then {f ∈ CMϕ : lim t→0+,∞ (CMϕ)b {f ∈ (CMϕ)b : {f ∈ CMϕ : lim t→0+ lim t→0+,∞ ϕ(t)(Cf )∗∗(t) = 0} if ϕ(0+) = 0, p(Mϕ) > 1 if ϕ(0+) = 0, p(Mϕ) = 1 if ϕ(0+) > 0, ϕ(∞) = ∞ if ϕ(0+) > 0, ϕ(∞) < ∞ Cf (t) = 0} Cf (t) = 0} (CMϕ)a =   In particular, if we assume that p(Mϕ) > 1 then the non-trivial Ces`aro-Marcinkiewicz function space CMϕ is never order continuous. Proof. Let us consider the following cases: (i) ϕ(0+) = 0 and p(Mϕ) > 1. Operator C is bounded from Theorem D and (Mϕ)a = {f ∈ Mϕ : lim t→0+,∞ ϕ(t)f ∗∗(t) = 0}, see [13, Def. 1.3 and Th. 1.3] so applying Theorem 16 we get (CMϕ)a = C(Mϕ)a = {f ∈ CMϕ : lim t→0+,∞ ϕ(t)(Cf )∗∗(t) = 0}. (ii) ϕ(0+) = 0 and p(Λϕ) = 1. Because (Mϕ)a 6= {0} and the Ces`aro operator C is not bounded on Mϕ we can use Theorem 16 (i). (iii) ϕ(0+) > 0 and ϕ(∞) = ∞. In this situation (Mϕ)a is trivial in view of Theorem B, so we simply use Theorem 16 (ii). (iv) ϕ(0+) > 0 and ϕ(∞) < ∞. It is easy to see that Mϕ = L∞. Indeed, inclusion Mϕ ֒→ L∞ is a consequence of Theorem B. Now, if f ∈ L∞ then f ∗∗ also is a bounded function and kf kMϕ = sup t∈I ϕ(t)f ∗∗(t) ≤ kf ∗∗kL∞(I) kϕkL∞(I) < ∞, since ϕ(∞) < ∞. Therefore also L∞ ֒→ Mϕ. From Remark 19 we get (CMϕ)a = (Ces∞)a = {f ∈ CMϕ : lim t→0+,∞ Cf (t) = 0}, and the proof is finished. Question 26. It seems that non-trivial Ces`aro-Marcinkiewicz function space CMϕ is never order continuous but this is only our conjecture. 4.5. The spaces C(L1 ∩ L∞) and C(L1 + L∞). The spaces L1 ∩ L∞ and L1 + L∞ occupy a special place in the theory of symmetric spaces because they are respectively the largest and the smallest of all symmetric function spaces, i.e. L1 ∩ L∞ ֒→ E[0, ∞) ֒→ L1 + L∞ and L∞ ֒→ E[0, 1] ֒→ L1, cf. [4, Th. 6.6 and Th. 6.7, pp. 77-78], kf kL1∩L∞ = max{kf kL1 , kf kL∞}, and (cid:3) kf kL1+L∞ = inf{kgkL1 + khkL∞ : f = g + h, f ∈ L1 and h ∈ L∞} =Z 1 0 cf. [4, Th. 6.2 and Th. 6.4, p. 74 and 76]. f ∗(t)dt, The case of L1 ∩ L∞ is not very interesting because if I = [0, ∞) then C(L1 ∩ L∞) = {0} and if I = [0, 1] then L∞ ֒→ L1 so L1 ∩ L∞ = L∞ and L1 + L∞ = L1. Therefore for I = [0, 1] we have C(L1 ∩ L∞) = C(L∞) = Ces∞ and C(L1 + L∞) = C(L1) = Ces1 - these cases were already discussed in the Section 4.1. Proposition 27. (C(L1 + L∞))a = {f ∈ C(L1 + L∞) : (Cf )∗(∞) = 0}. 15 Proof. We claim that (4.1) (L1 + L∞)a = {f ∈ L1 + L∞ : f ∗(∞) = 0}. This characterization is well known (see e.g. note in Section 2 in [13] and references therein) but we will give a short proof for the sake of completeness. Inclusion ⊂ is a direct consequence of Theorem A (ii). For the reverse inclusion ⊃ take f ∈ L1 + L∞ and observe that (cid:13)(cid:13)f ∗χ[0,1)(cid:13)(cid:13)L1 =Z 1 =Z 1 0 0 (f ∗χ[0,1))∗(t)dt f ∗(t)dt = kf kL1+L∞ < ∞. Since f ∗χ[0, 1 Lebesgue dominated convergence theorem n ] → 0 almost everywhere and is dominated by integrable function thus, from as n → ∞. Therefore, if f /∈ (L1 + L∞)a then according to Theorem A (ii) we get → 0, f ∗χ[0, 1 (cid:13)(cid:13)(cid:13) f ∗χ[0, 1 =(cid:13)(cid:13)(cid:13) n ](cid:13)(cid:13)(cid:13)L1 n ](cid:13)(cid:13)(cid:13)L1+L∞ (cid:13)(cid:13)f ∗χ[n,∞)(cid:13)(cid:13)L1+L∞ 9 0. f ∗(t)dt =Z 1 0 =(cid:13)(cid:13)f ∗χ[n,∞)(cid:13)(cid:13)L1+L∞ ≥ δ > 0, χ[1,∞)(t) ∈ (L1 + L∞)a. 1 t Passing to subsequence if necessary we can assume that (cid:13)(cid:13)f ∗χ[n,∞)(cid:13)(cid:13)L1+L∞ ≥ δ for some δ > 0. We have f ∗(n) ≥Z n+1 n (f ∗χ[n,∞))∗(t)dt so f ∗(∞) > 0 and the claim follows. Note that from (4.1) Using Remark 13 we obtain (C(L1 + L∞))a = C((L1 + L∞)a) and the proof is complete. (cid:3) 16 References [1] S. V. Astashkin, L. Maligranda, Structure of Ces`aro function spaces, Indag. Math. (N.S.) 20 (3) (2009) 329-379. [2] S. V. Astashkin, L. Maligranda, Structure of Ces`aro function spaces: a survey, Banach Center Publ. 102 (2014) 13-40. [3] W. Arendt, B. de Pagter, Spectrum and asymptotics of the Black-Scholes partial differential equation in (L1, L∞)-interpolation spaces, Pacific J. Math. 202 (1) (2002) 1-36. [4] C. Bennet, R. Sharpley, Interpolation of Operators, Pure Appl. Math., vol. 129, Academic Press, Inc., Boston, San Diego, New York, Berkeley, London, Sydney, Tokyo, Toronto, 1988. [5] D. W. Boyd, Indices for the Orlicz spaces, Pacific J. Math. 38 (2) (1971) 315-323. [6] D. W. Boyd, Indices of function spaces and their relationship to interpolation, Canad. J. Math. 21 (1969) 1245-1254. [7] S. Chen, Geometry of Orlicz spaces, Disserationes Math. (Rozprawy Mat.) CCCLVI (1996). [8] M. Ciesielski, P. Kolwicz, A. Panfil, Local monotonicity structure of symmetric spaces with application, J. Math. Anal. Appl. 409 (2014) 649-662. [9] G. P. Curbera, W. J. Ricker, Abstract Ces`aro spaces: integral representation, J. Math. Anal. Appl. 441 (2016) (1) 25-44. [10] O. Delgado, J. Soria, Optimal domain for the Hardy operator, J. Funct. Anal. 244 (1) (2007) 119-133. [11] A. S. Granero, H. Hudzik, On some proximinal subspaces of modular spaces, Acta Math. Hungar. 85 (1-2) (1999) 59-79. [12] B. D. Hassard, D. A. Hussein, On Ces`aro function spaces, Tamkang J. Math. 4 (1973) 19-25. [13] A. Kami´nska, M. J. Lee, M-ideal properties in Marcinkiewicz spaces, Comment. Math. (2004), 123-144, Tomus specialist in Honorem Juliani Musielak. [14] A. Kami´nska, M. Maligranda, L.-E. Persson, Indices and regularization of measurable functions, in: Function Spaces, The 5th Conference: Proceedings of the Conference at Pozna´n, Poland 2000, 231-246 [15] T. Kiwerski, P. Kolwicz, Isomorphic copies of l∞ in Ces`aro-Orlicz function spaces, Positivity, published online on 3 November 2016, doi: 10.1007/s11117-016-0449-6. [16] T. Kiwerski, P. Kolwicz, Isometric copies of l∞ in Ces`aro-Orlicz function spaces, preprint of 16 pages, 30 Jan- uary 2016, \protect\vrule width0pt\protect\href{http://arxiv.org/abs/1701.08794}{arXiv:1701.08794}v1 at http://arxiv.org/abs/1701.08794. [17] A. Kufner, L. Maligranda, L. E. Persson, The Hardy inequality. About its history and some related results, Vydavatelsky Servis Publishing House, Pilsen 2007. [18] M. A. Krasnosel'skiı, Ya. B. Rutickiı, Convex functions and Orlicz Spaces, P. Noorddhoff Ltd., Groningen, 1961 (translation). [19] S. G. Krein, Yu. I. Petunin, E. M. Semenov, Interpolation of linear operators, Nauka, Moscow, Russia, 1978. [20] P. Kolwicz, Rotundity properties in Calder`on-Lozanovskiı, Houston J. Math., 31 (3) (2005) 883-912. [21] P. Kolwicz, Local structure of symmetrizations E(∗) with applications, J. Math. Anal. Appl. 440 (2) (2016) doi: 10.1016/j.jmaa.2016.03.075 [22] K. Le´snik, M. Maligranda, On abstract Ces`aro spaces. I. Duality, J. Math. Anal. Appl. 424 (2) (2015) 932-951, doi: 10.1016/j.jmaa.2014.11.023. [23] K. Le´snik, M. Maligranda, On abstract Ces`aro spaces. II. Optimal Range, Integral Equations Operator Theory 81 (2) (2015) 227-235. [24] K. Le´snik, M. Maligranda, Interpolation of abstract Ces`aro, Copson and Tandori spaces, Indag. Math. (N.S.) 27 (3) (2016), 764 -- 785, doi:10.1016/j.indag.2016.01.009. [25] J. Lindenstrauss, L. Tzafriri, Classical Banach Spaces I. Sequence Spaces, Springer-Verlag, Berlin, Heidelberg, 1977. [26] J. Lindenstrauss, L. Tzafriri, Classical Banach Spaces II. Function Spaces, Springer-Verlag, Berlin, Heidel- berg, 1979. [27] G. Ya. Lozanovskiı, On isomorphic Banach structures, Sibirsk. Math. J. 10 (1969) 93-98. [28] L. Maligranda, Indices and interpolation, Disserationes Math. (Rozprawy Mat.) CCXXXIV (1985). [29] L. Maligranda, Orlicz Spaces and Interpolation, Sem. Mat., vol. 5, Univ. of Campinas, Campinas SP, Brazil, 1989. [30] J. Musielak, Orlicz Spaces and Modular Spaces, Lecture Notes in Math., vol. 1034, Springer-Verlag, 1983. [31] J. S. Shiue, A note on Ces`aro function space, Tamkang J. Math. 1 (1970) 91-95. [32] A. C. Zaanen, Riesz spaces II, North-Holland Math., Library 30, North-Holland, Amsterdam, 1983. (Tomasz Kiwerski) Faculty of Mathematics, Computer Science and Econometrics, University of Zielona G`ora, prof. Z. Szafrana 4a, 65-516 Zielona G`ora, Poland 17 e-mail address: [email protected] (Jakub Tomaszewski) Institute of Mathematics, Faculty of Electrical Engineering, Pozna´n University of Technology, Piotrowo 3A, 60-965 Pozna´n, Poland e-mail address: [email protected] 18
1901.02796
3
1901
2019-03-26T17:05:31
The pseudo-differential calculus in a Bargmann setting
[ "math.FA" ]
We give a fundament for Berezin's analytic $\Psi$do considered in \cite{Berezin71} in terms of Bargmann images of Pilipovi{\'c} spaces. We deduce basic continuity results for such $\Psi$do, especially when the operator kernels are in suitable mixed weighted Lebesgue spaces and act on certain weighted Lebesgue spaces of entire functions. In particular, we show how these results imply well-known continuity results for real $\Psi$do with symbols in modulation spaces, when acting on other modulation spaces.
math.FA
math
PSEUDO-DIFFERENTIAL CALCULUS IN A BARGMANN SETTING NENAD TEOFANOV AND JOACHIM TOFT Abstract. We give a fundament for Berezin's analytic Ψdo con- sidered in [4] in terms of Bargmann images of Pilipovi´c spaces. We deduce basic continuity results for such Ψdo, especially when the operator kernels are in suitable mixed weighted Lebesgue spaces and act on certain weighted Lebesgue spaces of entire functions. In particular, we show how these results imply well-known continu- ity results for real Ψdo with symbols in modulation spaces, when acting on other modulation spaces. 0. Introduction The aim of the paper is to put a fundament for the theory of an- alytic pseudo-differential operators, considered in [4] by F. Berezin. This is essentially done through a detailed analysis of Bargmann im- ages of the so-called Pilipovi´c spaces of functions and distributions, given in [11, 26]. More precisely, we consider kernels related to integral representations of analytic pseudo-differential operators to deduce their continuity properties. When the corresponding symbols belong to suit- able (weighted) Lebesgue spaces of semi-conjugate analytic functions, we prove the continuity of the analytic pseudo-differential operators when acting between (weighted) Lebesgue spaces of analytic functions. Moreover, by using the relationship between the Bargmann transform and the short-time Fourier transform we show that our results can be used to recover well-known (sharp) continuity properties of (real) pseudo-differential operators with symbols in modulation spaces which act between other modulation spaces, see [23, 25, 28]. We emphasize that our approach here is more general, because we have relaxed the assumptions on the involved weight functions, compared to earlier con- tributions. Analytic pseudo-differential operators, considered in [4] by Berezin are well-designed when considering several problems in analysis and its applications, e. g. in quantum mechanics. In the context of abstract harmonic analysis it follows that any linear and continuous operator be- tween Fourier invariant function and (ultra-)distribution spaces may, in 2010 Mathematics Subject Classification. Primary: 32W25, 35S05, 32A17, 46F05, 42B35 Secondary: 32A25, 32A05. Key words and phrases. Analytic kernels, Berezin operators, Pilipovi´c spaces, modulation spaces, Gelfand-Shilov spaces. 1 a unique way, be transformed into an analytic pseudo-differential oper- ator by the Bargmann transform (see Section 2). An advantage of such reformulations is that all of the involved objects are essentially entire functions and thereby possess several strong and convenient properties. The definition of analytic pseudo-differential operators resembles the definition of real pseudo-differential operators. In fact, let apx, ξq be a suitable function or (ultra-)distribution on the phase space R2d. Then the (real) pseudo-differential operator Oppaq acting on suitable sets of functions or (ultra-)distributions on the configuration space Rd is given by 2żRd fpxq ÞÑ pOppaqfqpxq " p2πq´ d apx, ξqpfpξqeixx,ξy dξ. (0.1) Here the integral in (0.1) should be interpreted in a distributional (weak) sense, if necessary, and we refer to [15] or Section 1 for the notation. Suppose instead that a is a suitable semi-conjugate entire (analytic) function on Cd Cd -- C2d, i. e. pz, wq ÞÑ apz, wq is an entire (analytic) function. Then the analytic pseudo-differential operator OpVpaq acting on suitable entire functions F on Cd is given by Fpzq ÞÑ pOpVpaqFqpzq "żCd apz, wqFpwqepz,wq dµpwq. (0.2) Here dµpwq is the Gauss measure π´de´w2 dλpwq, where dλpwq is the Lebesgue measure on Cd, and pz, wq " řd j"1 zj wj, when z " pz1, . . . , zdq P Cd, and w " pw1, . . . , wdq P Cd. This means that the operator kernel (with respect to dµ) is given by Kpz, wq " Kapz, wq " apz, wqepz,wq. pTKFqpzq "żCd Evidently, pOpVpaqFqpzq is equal to the integral operator Kpz, wqFpwq dµpwq with respect to dµ, when K is given by (0.3). By the analyticity prop- erties of the symbol a it follows that pz, wq ÞÑ Kpz, wq is an entire function on C2d. In [4, 25] several facts of analytic pseudo-differential operators are (0.3) (0.4) deduced. For example, if a and F are chosen such that z ÞÑ apz, qF epz, q is locally uniformly bounded and analytic from Cd to L1pdµq, then OpVpaqF in (0.2) is a well-defined entire function on Cd. In [4, 25] it is also observed that pOpVpzjqFqpzq " zjFpzq and pOpVpwjqFqpzq " pBjFqpzq when F P L1pdµ1q X ApCdq, and dµ1pwq " p1 ` wq dµpwq. 2 (0.5) In such setting we study the mapping properties for complex integral operators and pseudo-differential operators when respectively K " Ka and a above belong to suitable classes of semi-conjugate entire func- tions. In fact, we permit more generally that K and a belong to suitable classes of formal semi-conjugate analytic power series expansions. That is, Kpz, wq and apz, wq are of the forms cKpα, βqeαpzqeβpwq capα, βqeαpzqeβpwq, eαpzq " zα ?α! , and ÿα,β ÿα,β respectively. To set the stage for our study we collect the background material in Section 1. It contains a brief account on weight functions, Gelfand- Shilov spaces, spaces of Hermite functions and power series expansions, modulation spaces, and Bargmann transform and spaces of analytic functions. Especially, we recall basic facts for the spaces and their Bargmann duals A5σpCdq`A0,5σpCdq, AspCdq`A0,spCdq, 0,spCdq, 5σpCdq`A1 spCdq`A1 A1 0,5σpCdq, A1 Fpzq "ÿα cpF, αq À hαα!´ 1 2σ , cpF, αqÀ e´rα cpF, αqeαpzq, (0.6) (0.7) (0.8) 1 2s , when s, σ ą 0. The spaces in (0.6) consist of all formal power series with coefficients satisfying respectively, for some (for every) h, r ą 0, and the spaces in (0.7) consist of all formal power series in (0.8) such that cpF, αq À hαα!` 1 2σ , cpF, αqÀ e`rα 1 2s , respectively, for every (for some) h, r ą 0. to belong to any of the spaces In Section 2 we extend the definition of (0.4) to allow the kernels K and their duals where uA5σpC2dq, uA1 5σpC2dq, uA0,spC2dq, uA1 0,spC2dq, uAspC2dq, uA0,5σpC2dq, uA1 uA1 spC2dq, 0,5σpC2dq, uA5σpC2dq " t K ; pz, wq ÞÑ Kpz, wq P A5σpC2dqu, TK : AspCdq ÞÑ A1 spC2dq, spCdq when K P uA1 3 and similarly for the other spaces in (0.9) and (0.10). In the end we prove that if s ą 0 or s " 5σ, then the integral operators in (0.4), (0.9) (0.10) (0.11) and (0.12) TK : A0,spCdq ÞÑA1 0,spCdq when K P uA1 0,spC2dq, are uniquely defined and continuous, and similarly when the roles of the non-duals in (0.6) and (0.9), and their duals in (0.7) and (0.10) are swapped. We also prove the opposite direction, that any linear and continuous operators between such spaces are given by such kernel operators. These kernel results are given in Propositions 2.2 and 2.3. Due to the Bargmann transform homeomorphisms, these results are also equivalent to Theorems 3.3 and 3.4 in [7] on kernel theorems for Pilipovi´c spaces. (See Subsection 1.5.) σ ą 0, σ ą 1, 2σ 2σ Note that, if s ě 1 1 2 z2´rz 1 2 z2`rz 1 1 σ`1 for some r ą 0 u, 2s for some r ą 0 u 2s for every r ą 0 u 2, then the spaces of power series expansions above can be identified with certain spaces of analytic and semi-conjugate analytic functions. For example we have A5σpCdq " t F P ApCdq ; Fpzq À erz AspCdq " t F P ApCdq ; Fpzq À e A1 spCdq " t F P ApCdq ; Fpzq À e 5σpCdq " t F P ApCdq ; Fpzq À erz A1 A1 51pCdq " ApCdq and similarly for uAspC2dq and uA1 Kpz, wqetpz,wq is homeomorphic on uA1 0,5σpC2dq, spC2dq. In particular, the mappings (0.11) and (0.12) can be formulated in terms of those function spaces. 2s and σ ą 0, and t P C, then Kpz, wq ÞÑ see Theorem 2.6. In particular, (0.3) implies that the mappings (0.11) and (0.12) still hold true with OpVpaq in place of TK. (Cf. Theorems 2.7 and 2.8.) 2, the conditions on a and its kernel Ka of OpVpaq are slightly different. More precisely, these conditions are of the form In the case s ě 1 uA1 spC2dq and σ´1 for every r ą 0 u, and A1 0,51pCdq " Adpt0uq, If instead s P p0, 1 uA1 0,spC2dq, uA1 5σpC2dq, and apz, wq À e 1 2 z´w2`rpz 1 2s `w 1 2s q Kpz, wq À e 1 2 pz2`w2q`rpz 1 2s `w 1 2s q in order for the mappings (0.11) and (0.12) should hold. (Cf. Theorems 2.9 and 2.10.) In Section 3 we consider operators (0.4), where certain linear pull- backs of their kernels obey suitable mixed and weighted Lebesgue norm 4 estimates. We prove that such operators are continuous between ap- propriate (weighted) Lebesgue spaces of entire functions. For example, let ω be a weight on Cd Cd and ω1, ω2 be weights on Cd such that ω2pzq ω1pwq À ωpz, wq and let where Kωpz, wq " e´ 1 If p, q, pj, qj P r1,8s satisfy 1 q1 ´ 1 p2 " 1 p1 ´ GK,ωpz, wq " Kωpz, z ` wq, 2 pz2`w2qKpz, wqωp ?2z,?2wq. 1 q2 " 1 ´ 1 p ´ 1 q and q ď p, E,pω1qpCdq to Ap2 and GK,ω P Lp,qpCd Cdq, then it follows from Theorem 3.3 that TK is E,pω2qpCdq. By slightly modifying the continuous from Ap1 definition of GK,ω we also deduce another similar but different conti- nuity result where the condition q ď p above is removed (cf. Theorem 3.5). We also present some consequences of these results. Theorem 3.4 can be considered as a special case of Theorem 3.3 formulated by ana- lytic pseudo-differential operators instead of integral operators. Theo- rems 3.8 and 3.9 are obtained by imposing conditions on moderateness on ω, ω1 and ω2 above and translating Theorem 3.3 and 3.5 to real pseudo-differential operators via the Bargmann transform and its in- verse. These approaches show that obtained continuity results on an- alytic pseudo-differential or integral operators might be suitable when investigating real pseudo-differential operators. In fact, Theorems 3.8 and 3.9 agree with the sharp results [24, Theorem 3.3], [27, Theorem 3.1] and [28, Theorem 2.2] in the Banach space case. Remark 3.10 in the end of Section 3 shows that our approach can be used to extend the latter results on real pseudo-differential operators to include situations with non-moderate weights. We note that the moderate condition on weights may in some situations be significantly restrictive (cf. Remark 1.14 in Section 1). 1. Preliminaries In this section we recall some facts on involved function and distri- bution spaces as well as on pseudo-differential operators. In Subsection 1.1 we introduce suitable weight classes. Thereafter we recall in Sub- sections 1.2 -- 1.4 the definitions and basic properties for Gelfand-Shilov, Pilipovi´c and modulation spaces. Then we discuss in Subsection 1.5 the Bargmann transform and recall some topological spaces of entire functions or power series expansions on Cd. The section is concluded with a review of some facts on pseudo-differential operators. 5 1.1. Weight functions. A weight on Rd is a positive function ω P L8 locpRdq such that 1{ω P L8 locpRdq. The weight ω on Rd is called moderate if there is a positive locally bounded function v on Rd such that x, y P Rd, ωpx ` yq ď Cωpxqvpyq, (1.1) for some constant C ě 1. If ω and v are weights on Rd such that (1.1) holds, then ω is also called v-moderate. The set of all moderate weights on Rd is denoted by PEpRdq. The weight v on Rd is called submultiplicative, if it is even and (1.1) holds for ω " v. From now on, v always denotes a submultiplicative In particular, if (1.1) holds and v weight if nothing else is stated. is submultiplicative, then it follows by straight-forward computations that ωpxq vpyq À ωpx ` yq À ωpxqvpyq, vpx ` yq À vpxqvpyq and vpxq " vp´xq, (1.2) x, y P Rd. Here and in what follows we write Apθq À Bpθq, θ P Ω, if there is a constant c ą 0 such that Apθq ď cBpθq for all θ P Ω. If ω is a moderate weight on Rd, then by [25] and above, there is a submultiplicative weight v on Rd such that (1.1) and (1.2) hold (see also [13, 25]). Moreover if v is submultiplicative on Rd, then (1.3) for some constant r ą 0 (cf. [13]). In particular, if ω is moderate, then (1.4) 1 À vpxq À erx and e´rx ď ωpxq À erx, ωpx ` yq À ωpxqery x, y P Rd for some r ą 0. 1.2. Gelfand-Shilov spaces. Let 0 ă s P R be fixed. Then the (Fourier invariant) Gelfand-Shilov space SspRdq (ΣspRdq) of Roumieu type (Beurling type) consists of all f P C 8pRdq such that }f}S σ s,h " sup xαBβfpxq hα`βpα! β!qs (1.5) is finite for some h ą 0 (for every h ą 0). Here the supremum should be taken over all α, β P Nd and x P Rd. The semi-norms } }S σ induce an inductive limit topology for the space SspRdq and projective limit topology for ΣspRdq, and the latter space becomes a Fr´echet space under this topology. 2 (s ą 1 2 ). spRdq are the The space SspRdq ‰ t0u (ΣspRdq ‰ t0u), if and only if s ě 1 The Gelfand-Shilov distribution spaces S 1 spRdq and Σ1 dual spaces of SspRdq and ΣspRdq, respectively. s,h 6 We have S1{2pRdq ãÑ ΣspRdq ãÑ SspRdq ãÑ ΣtpRdq tpRdq 1{2pRdq, ãÑ S pRdq ãÑ S 1pRdq ãÑ Σ1 spRdq ãÑ S 1 ãÑ S 1 spRdq ãÑ Σ1 (1.6) 1 2 ă s ă t. Here and in what follows we use the notation A ãÑ B when the topo- logical spaces A and B satisfy A Ď B with continuous embeddings. A convenient family of functions concerns the Hermite functions hαpxq " π´ d 4p´1qαp2αα!q´ 1 2 e 2 pBαe´x2 q, α P Nd. 2 x The set of Hermite functions on Rd is an orthonormal basis for L2pRdq. It is also a basis for the Schwartz space and its distribution space, and for any Σs when s ą 1 2 and their distribution spaces. They are also eigenfunctions to the Harmonic oscillator H " Hd " x2 ´ ∆ and to the Fourier transform F , given by fpxqe´ixx,ξy dx, 2, Ss when s ě 1 2żRd pF fqpξq " pfpξq " p2πq´ d when f P L1pRdq. Here x , y denotes the usual scalar product on Rd. In fact, we have ξ P Rd, spRdq and on Σ1 Hdhα " p2α ` dqhα. The Fourier transform F extends uniquely to homeomorphisms on S 1pRdq, S 1 spRdq. Furthermore, F restricts to homeo- morphisms on S pRdq, SspRdq and on ΣspRdq, and to a unitary oper- ator on L2pRdq. Similar facts hold true when the Fourier transform is replaced by a partial Fourier transform. Gelfand-Shilov spaces and their distribution spaces can also be char- acterized by estimates of short-time Fourier transform, (see e. g. [14, 21, 26]). More precisely, let φ P S pRdq be fixed. Then the short- time Fourier transform Vφf of f P S 1pRdq with respect to the window function φ is the Gelfand-Shilov distribution on R2d, defined by Vφfpx, ξq " Fpf φp ´ xqqpξq, x, ξ P Rd. If f, φ P S pRdq, then it follows that 2żRd Vφfpx, ξq " p2πq´ d fpyqφpy ´ xqe´ixy,ξy dy, x, ξ P Rd. By [25, Theorem 2.3] it follows that the definition of the map pf, φq ÞÑ Vφf from S pRdq S pRdq to S pR2dq is uniquely extendable to a continuous map from S 1 spR2dq, and restricts to a continuous map from SspRdqSspRdq to SspR2dq. The same conclusion holds with Σs in place of Ss, at each place. In the following propositions we give characterizations of Gelfand- Shilov spaces and their distribution spaces in terms of estimates of spRdq to S 1 spRdq S 1 7 the short-time Fourier transform. We omit the proof since the first part follows from [14, Theorem 2.7]) and the second part from [26, Proposition 2.2]. See also [8] for related results. Proposition 1.1. Let s ě 1 2), φ P SspRdqz0 (φ P ΣspRdqz0) and let f be a Gelfand-Shilov distribution on Rd. Then the following is true: 2 (s ą 1 (1) f P SspRdq (f P ΣspRdq), if and only if Vφfpx, ξq À e´rpx s q, for some r ą 0 (for every r ą 0). (2) f P S 1 spRdq (f P Σ1 spRdq), if and only if 1 1 s `ξ Vφfpx, ξq À erpx σ q, for every r ą 0 (for some r ą 0). 1 1 s `ξ x, ξ P Rd, x, ξ P Rd, (1.7) (1.8) 1.3. Spaces of Hermite series and power series expansions. Next we recall the definitions of topological vector spaces of Hermite se- ries expansions, given in [26]. As in [26], it is convenient to use suitable extensions of R` when indexing our spaces. Definition 1.2. The sets R5 and R5 are given by R5 " R`Ťσą0t5σu and R5 " R5Ťt0u. Moreover, beside the usual ordering in R, the elements 5σ in R5 and R5 are ordered by the relations x1 ă 5σ1 ă 5σ2 ă x2, when σ1, σ2, x1 and x2 are positive real numbers such that x1 ă 1 Definition 1.3. Let p P r1,8s, s P R5, r P R, ϑ be a weight on Nd, and let 2 and x2 ě 1 2. ϑr,spαq "$&% Then, 1 erα 2s , rαpα!q when s P R`, 2σ , when s " 5σ, 1 α P Nd. 0pNdq is the set of all sequences tcαuαPNd Ď C on Nd; (1) ℓ1 (2) ℓ0,0pNdq " t0u, and ℓ0pNdq is the set of all sequences tcαuαPNd Ď (3) ℓp C such that cα ‰ 0 for at most finite numbers of α; rϑspNdq is the Banach space which consists of all sequences tcαuαPNd Ď C such that }tcαuαPNd}ℓp rϑs " }tcαϑpαquαPNd}ℓp ă 8; (4) ℓ0,spNdq " Şrą0 ℓp rϑr,sspNdq and ℓspNdq " Ťrą0 jective respective inductive limit topologies of ℓp respect to r ą 0; ℓp rϑr,sspNdq, with pro- rϑr,sspNdq with 8 (5) ℓ1 0,spNdq " Ťrą0 ℓp r1{ϑr,sspNdq and ℓ1 inductive respective projective limit topologies of ℓp with respect to r ą 0. spNdq " Şrą0 ℓp r1{ϑr,sspNdq, with r1{ϑr,sspNdq Let p P r1,8s, and let ΩN be the set of all α P Nd such that α ď N. Then the topology of ℓ0pNdq is defined by the inductive limit topology of the sets tcαuαPNd P ℓ1 0pNdq ; cα " 0 when α R ΩN( with respect to N ě 0, and whose topology is given through the semi- norms (1.9) tcαuαPNd ÞÑ }tcαuαďN}ℓppΩN q, It is clear that these topologies are independent of p. Furthermore, the topology of ℓ1 0pNdq is defined by the semi-norms (1.9). It follows that ℓ1 0pNdq is a Fr´echet space, and that the topology is independent of p. Next we introduce spaces of formal Hermite series expansions f " ÿαPNd F " ÿαPNd cαhα, tcαuαPNd P ℓ1 0pNdq. cαeα, tcαuαPNd P ℓ1 0pNdq. (1.10) (1.11) (1.12) (1.13) cαeα (1.14) and power series expansions which correspond to ℓ0,spNdq, Here eαpzq " We consider the mappings TH : tcαuαPNd ÞÑ ÿαPNd cαhα ℓspNdq, zα ?α! , ℓ1 spNdq and ℓ1 0,spNdq. z P Cd, α P Nd. and TA : tcαuαPNd ÞÑ ÿαPNd between sequences, and formal Hermite series and power series expan- sions. Definition 1.4. If s P R5, then H0,spRdq, HspRdq, H1 0,spRdq, spRdq and H1 (1.15) and A0,spCdq, AspCdq, A1 spCdq and A1 0,spCdq, (1.16) are the images of TH and TA respectively in (1.14) of corresponding spaces in (1.12). The topologies of the spaces in (1.15) and (1.16) are inherited from the corresponding spaces in (1.12). 9 Since locally absolutely convergent power series expansions can be identified with entire functions, several of the spaces in (1.16) are iden- tified with topological vector spaces contained in ApCdq (see Theo- rem 1.9 below and the introduction). Here ApΩ0q is the set of all (complex valued) functions which are analytic in Ω0. (For Ω0 Ď Cd, contain Ω0.) ApΩ0q "Ť ApΩq, where the union is taken over all open Ω Ď Cd which We recall that f P S pRdq if and only if it can be written as (1.10) such that for every N ě 0 (cf. definitions that the inclusions e. g. [19]). In particular it follows from the cα À xαy´N , H0pRdq ãÑ H0,spRdq ãÑ HspRdq ãÑ H0,tpRdq ãÑ S pRdq ãÑ S 1pRdq ãÑ H1 ãÑ H1 0,spRdq ãÑ H1 0,tpRdq ãÑ H1 spRdq 0pRdq, when s, t P R5, s ă t, (1.17) are dense. Remark 1.5. By the definition it follows that TH in (1.14) is a homeo- morphism between any of the spaces in (1.12) and corresponding space in (1.15), and that TA in (1.14) is a homeomorphism between any of the spaces in (1.12) and corresponding space in (1.16). The next results give some characterizations of HspRdq and H0,spRdq when s is a non-negative real number. Proposition 1.6. Let 0 ď s P R and let f P H1 HspRdq (f P H0,spRdq), if and only if f P C 8pRdq and satisfies 0pRdq. Then f P }H N d f}L8 À hN N!2s, (1.18) for some h ą 0 (every h ą 0). Moreover, it holds HspRdq " SspRdq ‰ t0u, H0,spRdq " ΣspRdq ‰ t0u when s P p 1 2 ,8q, HspRdq " SspRdq ‰ t0u, H0,spRdq ‰ ΣspRdq " t0u when s " 1 2 , HspRdq ‰ SspRdq " t0u, H0,spRdq ‰ ΣspRdq " t0u when s P p0, 1 2q, HspRdq ‰ SspRdq " t0u, H0,spRdq " ΣspRdq " t0u when s " 0. We refer to [26] for the proof of Proposition 1.6. Due to the pioneering investigations related to Proposition 1.6 by Pilipovi´c in [17, 18], we call the spaces HspRdq and H0,spRdq Pilipovi´c In fact, in the spaces of Roumieu and Beurling types, respectively. restricted case s ě 1 2, Proposition 1.6 was proved already in [17, 18]. 10 Later on it will also be convenient for us to have the following defi- nition. Here we let Fpz2, z1q and Fpz2, z1q be the formal power series (1.19) and ÿ cpα2, α1qeα2pz2qeα1pz1q, ÿ cpα2, α1qeα2pz2qeα1pz1q respectively, when Fpz2, z1q is the formal power series ÿ cpα2, α1qeα2pz2qeα1pz1q. (1.20) Here zj P Cdj , j " 1, 2, and the sums should be taken over all pα2, α1q P Nd2 Nd1. Definition 1.7. Let d " d2`d1, s P R5, ΘC,1 and ΘC,2 be the operators pΘC,1Fqpz2, z1q " Fpz2, z1q and pΘC,2Fqpz2, z1q " Fpz2, z1q between formal power series in (1.19) and (1.20), zj P Cdj , j " 1, 2. Then uAspCd2 Cd1q, uA1 spCd2 Cd1q, uA0,spCd2 Cd1q, uA1 0,spCd2 Cd1q (1.21) are the images of (1.16) under ΘC,1, and uApCd2Cd1q and uAd2,d1pt0uq " uAd2`d1pt0uq are the images of ApCdq and Ad2`d1pt0uq respectively un- der ΘC,1. The topologies of the spaces in (1.21), uApCd2 Cd1q and uAd2,d1pt0uq are inherited from the topologies in the spaces (1.16), ApCdq the spaces in (1.16) can be considered as special cases of uApCd2 Cd1q and Adpt0uq, respectively. Remark 1.8. By letting d2 " d and d1 " 0, it follows that ApCdq and and the spaces in (1.21). 51pCdq " ApCdq and A1 0,51pCdq " Adpt0uq, it follows that Since A1 uA1 51pCd2 Cd1q " uApCd2 Cd1q, uA1 0,51pCd2 Cd1q " uAd2,d1pt0uq. (1.22) The following results are now immediate consequences of Theorems 4.1, 4.2, 5.2 and 5.3 in [26] and Definition 1.7. Here let and κ1,r,spzq " $''''&''''% κ2,r,spzq "$&% erplogxzyq 1 1´2s , erz 2σ σ`1 , 2 z 2 ´rz 1 s , e erz 2σ σ´1 , e 2 z 2 `rz 1 s , 11 2 s ă 1 s " 5σ, σ ą 0, s ě 1 2 , s " 5σ, σ ą 1, s ě 1 2 , (1.23) (1.24) Theorem 1.9. Let s1, s2 P R5 be such that s2 ą 51, and let κ1,r,s and κ2,r,s be given by (1.23) and (1.24) respectively, when r ą 0. Then the following is true: 1 Cd1q such that K À κ1,r,s1 for some r ą 0 (for every 0 ă r ă 2 ). (1) uAs1pCd2 Cd1q (uA0,s1pCd2 Cd1q) consists of all K P uApCd2 (2) uA1 0,s2pCd2 Cd1q) consists of all K P uApCd2 Cd1q such that K À κ2,r,s2 for every r ą 0 (for some r ą 0). By Remark 1.8 it follows that Theorem 1.9 remains true after the s2pCd2 Cd1q (uA1 spaces in (1.21) are replaced by corresponding spaces in (1.16). 1.4. Modulation spaces. Before giving the definition of a broad fam- ily of modulation spaces, we make a review of mixed normed spaces of Lebesgue types, adapted to suitable bases of the Euclidean space Rd. Let E be the ordered basis te1, . . . , edu of Rd. Then the ordered basis E1 " te1 1, . . . , e1 du (the dual basis of E) satisfies xej, e1 ky " 2πδjk for every j, k " 1, . . . , d. The corresponding parallelepiped, lattice, dual parallelepiped and dual lattice are given by κpEq " t x1e1 ` ` xded ; px1, . . . , xdq P Rd, 0 ď xk ď 1, k " 1, . . . , d u, ΛE " t j1e1 ` ` jded ; pj1, . . . , jdq P Zd u, κpE1q " t ξ1e1 and 1 ` ` ξde1 d ; pξ1, . . . , ξdq P Rd, 0 ď ξk ď 1, k " 1, . . . , d u, E " ΛE1 " t ι1e1 Λ1 1 ` ` ιde1 d ; pι1, . . . , ιdq P Zd u, respectively. Note here that the Fourier analysis with respect to general biorthogonal bases has recently been developed in [20]. We observe that there is a matrix TE such that e1, . . . , ed and e1 are the images of the standard basis under TE and TE1 " 2πpT ´1 spectively. 1, . . . , e1 E qt, re- d In the following we let maxpqq " maxpq1, . . . , qdq and minpqq " minpq1, . . . , qdq when q " pq1, . . . , qdq P r1,8sd. Definition 1.10. Let E be an ordered basis of Rd and p " pp1, . . . , pdq P r1,8sd. If f P L1 locpRdq, then }f}Lp is defined by E where gkpzkq, zk P Rd´k, k " 0, . . . , d ´ 1, are inductively defined as g0px1, . . . , xdq " fpx1e1 ` ` xdedq, px1, . . . , xdq P Rd, }f}Lp E " }gd´1}Lpd pRq 12 and gkpzkq " }gk´1p , zkq}Lpk pRq, EpRdq consists of all f P L1 The space Lp and is called E-split Lebesgue space (with respect to p). zk P Rd´k, k " 1, . . . , d ´ 1. locpRdq such that }f}Lp is finite, E Next we discuss suitable conditions for bases in the phase space R2d. We let σpX, Y q be the standard symplectic form on the phase space, given by σpX, Y q " xy, ξy ´ xx, ηy, X " px, ξq P R2d, Y " py, ηq P R2d. We notice that if is the standard basis of R2d, then te1, . . . , ed, ε1, . . . , εdu (1.25) σpej, ekq " 0, σpej, εkq " ´δj,k, (1.26) when j, k P t1, . . . , du. More generally, a basis in (1.25) for the phase space R2d is called symplectic if (1.26) holds. A symplectic basis (1.25) for R2d is called phase split if e1, . . . , ed and ε1, . . . , εd span and σpεj, εkq " 0, tpx, 0q P R2d ; x P Rd u and tp0, ξq P R2d ; ξ P Rd u, respectively. Next we give the definition of our class of modulation spaces. Definition 1.11. Let E be an ordered basis for R2d, p P r1,8s2d, φpxq " π´ d and let ω be a weight on R2d. Then the modulation space M p 4 e´ 1 E,pωqpRdq consists of all f P H1 51pRdq such that 2 x2 }f}M p E,pωq " }Vφf ω}Lp E (1.27) is finite. 2 x2 4 e´ 1 and f P H1 We remark that if φpxq " π´ d 51pRdq, then px, ξq ÞÑ Vφfpx, ξq is a smooth function (cf. [26]). Furthermore, by [26, Theorem 4.8] we get the following. The proof is omitted. Proposition 1.12. Let E be an ordered basis for R2d, p P r1,8s2d and let ω be a weight on R2d. Then M p E,pωqpRdq is a Banach space with norm given by (1.27). If the weight ω in Definition 1.11 is a moderate weight, then we can E,pωqpRdq. In what follows we let p1 P r1,8s be say more concerning M p the conjugate exponent of p P r1,8s, i. e. 1 Proposition 1.13. Let E be an ordered basis for R2d, p P r1,8s2d and let ω, v P PEpR2dq be such that ω is v-moderate. Then the following is true: p1 " 1. p ` 1 13 (1) Σ1pRdq ãÑ M p E,pωqpRdq ãÑ Σ1 then Σ1pRdq is dense in M p 1pRdq. If in addition maxppq ă 8, E,pωqpRdq; (2) if φ P M 1 pvqpRdqzt0u and f P Σ1 E,pωqpRdq, 1pRdq, then f P M p if and only if the right-hand side of (1.27) is finite. Further- pvqpRdqzt0u in (1.27) give rise more, different choices of φ P M 1 to equivalent norms; (3) M p (4) if p 1, . . . , p1 E,pωqpRdq increases with p1, . . . , p2d and decreases with ω; 1 " pp1 2dq, then the restriction of the L2pRdq scalar product p , q to Σ1pRdq is uniquely extendable to a (semi- E,p1{ωqpRdq. If in conjugate) duality between M p E,pωqpRdq can be iden- addition maxppq ă 8, then the dual of M p tified by M p E,pωqpRdq and M p E,p1{ωqpRdq through the form p , q. 1 1 Proposition 1.13 follows by similar arguments as in Chapters 11 and 12 in [12] (see also [25, 26]). Remark 1.14. In some sense, the variable x at the weight ωpx, ξq in the definition of modulation spaces quantify growth and decay properties for the involved functions or distributions. In the same way the variable ξ quantify regularity or lack of regularity for the involved functions or distributions. 2 x2 4 e´ 1 By the analysis in [26] it follows that there are no bounds on how fast Vφf may grow or decay at infinity when φpxq " π´ d is fixed, x P Rd, and f is taken in the class H1 51pRdq. Since weights in PEpR2dq are bounded by exponential functions, the restrictions of the weights in Proposition 1.13 are significantly stronger compared to what is the case in Proposition 1.12. A question here concerns wether it is possible to extend parts of Proposition 1.13 to larger weight classes than PEpR2dq or not. It seems that the invariance properties (2) in Proposition 1.13 con- cerning the choice of weight function are not possible for weights that are not moderate. On the other hand, (1) and (4) in Proposition 1.13 hold true for certain weights outside PEpR2dq. In fact, in [25], certain weight classes which contain PEpR2dq as well as weights of the form ωpx, ξq " erpx 1 s `ξ 1 s q, x, ξ P Rd, when r ą 0 and s ą 1 2 are introduced. For corresponding (broader) families of modulation spaces it is then proved that Proposition 1.13 (1) and (4) hold true (with some modifications). 1.5. Bargmann transform and spaces of analytic functions. The Bargmann transform Vd is the homeomorphism from the spaces in (1.15) to respective spaces in (1.16), given by TA T ´1 H , where TH and TA are given by (1.14). 14 We notice that if f P LppRdq for some p P r1,8s, then Vdf is the entire function given by pVdfqpzq " π´d{4żRd which can also be formulated as exp´´ pVdfqpzq "żRd or 1 2pxz, zy`y2q`21{2xz, yy¯fpyq dy, z P Cd, Adpz, yqfpyq dy, z P Cd, (1.28) z P Cd, y P Rd. Adpz, yq " π´d{4 exp´ ´ (Cf. [1, 2].) Here pVdfqpzq " xf, Adpz, qy, where the Bargmann kernel Ad is given by 1 z P Cd, 2pxz, zy ` y2q ` 21{2xz, yy¯, dÿj"1 z " pz1, . . . , zdq P Cd xz, wy " and pz, wq " xz, wy zjwj when and w " pw1, . . . , wdq P Cd, and otherwise x , y denotes the duality between test function spaces and their corresponding duals which is clear form the context. We note that the right-hand side in (1.28) makes sense when f P S 1 1{2pRdq and defines an element in ApCdq, since y ÞÑ Adpz, yq can be interpreted as an element in S1{2pRdq with values in ApCdq. It was proved by Bargmann that f ÞÑ Vdf is a bijective and iso- metric map from L2pRdq to the Hilbert space A2pCdq, the set of entire functions F on Cd which fullfils }F}A2 "´żCd Fpzq2dµpzq¯1{2 ă 8. (1.29) dλpzq, where dλpzq is the Lebesgue measure Recall, dµpzq " π´de´z2 on Cd, and the scalar product on A2pCdq is given by pF, GqA2 "żCd pF, Gq ÞÑ xF, GyA2 " xF, GyA2pCdq "żCd FpzqGpzq dµpzq, F, G P A2pCdq. For future references we note that the latter scalar product induces the bilinear form FpzqGpzq dµpzq (1.30) (1.31) on A2pCdq A2pCdq. In [1] it was proved that the orthonormal basis thαuαPNd in L2pRdq of Hermite functions is mapped to the orthonormal basis teαuαPNd in A2pCdq (cf. (1.13)). Furthermore, there is a convenient reproducing 15 In fact, let ΠA be the operator from L2pdµq to Fpwqepz,wq dµpwq, z P Cd. (1.32) formula on A2pCdq. ApCdq, given by pΠAFqpzq "żCd Then it is proved in [1] that ΠA is an orthonormal projection from L2pdµq to A2pCdq. Fourier transform is given by From now on we assume that φ in the definition of the short-time φpxq " π´d{4e´x2{2, x P Rd, (1.33) if nothing else is stated. For such φ, it follows by straight-forward computations that the relationship between the Bargmann transform and the short-time Fourier transform is given by and U ´1 V Vd " Vφ, Vd " UV Vφ, (1.34) where UV is the linear, continuous and bijective operator on D 1pR2dq » D 1pCdq, given by pUVFqpx ` iξq " p2πqd{2epx2`ξ2q{2e´ixx,ξyFp21{2x,´21{2ξq, cf. [25]. x, ξ P Rd, (1.35) Definition 1.15. Let E be an ordered basis for R2d, UV be the oper- ator in (1.35), p P r1,8s2d, φpxq " π´ d and let ω be a weight on R2d. 4 e´ 1 2 x2 (1) The space B p E,pωqpCdq consists of all F P L1 }F}Bp E,pωq " }pU ´1 V Fq ω}Lp E locpCdq such that is finite; (2) The space Ap with topology inherited from B p E,pωqpCdq consists of all F P ApCdqŞ B p E,pωqpCdq. E,pωqpCdq We note that the spaces in Definition 1.15 are normed spaces when minppq ě 1. For conveneincy we set }F}Bp E,pωq " 8, when F R B p E,pωqpCdq is mea- E,pωq " 8, when F P ApCdqzB p surable, and }F}Ap Remark 1.16. In Definitions 1.11 and 1.15, important cases appear when E is the standard basis for R2d and p1 " " pd " p P r1,8s and pd`1 " " p2d " q P r1,8s. For such choices of E and p we set Lp,q " Lp E, M p,q E,pωqpCdq. and pωq " M p E,pωq, Ap,q pωq " Ap E,pωq Bp,q pωq " B p E,pωq. We also set M p pωq " M p,p pωq, Ap pωq " Ap,p pωq 16 and Bp pωq " Bp,p pωq. If in addition ω " 1, then we set M p,q M p E,pωq " M p E, and similarly for Ap E,pωq and B p pωq " M p,q E,pωq spaces. and M p pωq " M p, If instead E " ted`1, . . . , e2d, e1, . . . , edu where e1, . . . , e2d is the stan- dard basis for R2d and p1 " " pd " q P r1,8s and pd`1 " " p2d " p P r1,8s, then we set Lp,q W p,q pωq " M p E,pωq, Ap,q and Bp,q ,pωq " B p E,pωq. " Lp E, ,pωq " Ap E,pωq We notice that the space W p,q pωq in Remark 1.16 is an example of a (weighted) Wiener amalgam space (cf. [9, 10]). For future references we observe that the Bp pωq norm is given by pωq " 2d{pp2πq´d{2żCd e´z2{2Fpzqωp21{2zqp dλpzq1{p }F}Bp " 2d{pp2πq´d{2ijR2de´px2`ξ2q{2Fpx ` iξqωp21{2x,´21{2ξqp dxdξ1{p (1.36) (with obvious modifications when p " 8). Especially it follows that the norm and scalar product in B2 pωqpCdq take the forms }F}B2 pF, GqB2 pωq "żCd Fpzqωp21{2zq2 dµpzq1{2 pωq "żCd FpzqGpzqωp21{2zq2 dµpzq, , F P B2 pωqpCdq F, G P B2 pωqpCdq (cf. (1.29) and (1.30)). By the definitions and (1.34) it follows that the Bargmann transform E,pωqpCdq. In fact, we is an isometric injection from M p have the following refinement. We omit the proof since the result is a special case of Theorem 4.8 in [26]. E,pωqpRdq to Ap Proposition 1.17. Let E be an ordered basis for R2d, p P r1,8s2d, and ω be a weight on R2d. Then the Bargmann transform is an isometric bijection from M p E,pωqpRdq to Ap E,pωqpCdq. Finally, the SCB transform (i. e. the Semi Conjugated Bargmann transform), VΘ,d2,d1 is defined as ΘC,1 Vd2`d1. We also set VΘ,d " VΘ,d,d. Evidently, all properties of the Bargmann transform carry over to analogous properties for the SCB transform. Assume that E is a ba- sis for R2d2 R2d1, p P r1,8s2d2`2d1, p, q P p0,8s and that ω is a weight 17 der the map ΘC,1 with the topology defined by the norm E,pωqpCd2`d1q is the image of Ap E,pωqpCd2`d1q un- on R2d2 R2d1 Then uAp , E,pΘC,1ωq E,pωqpCd2`d1q. E,pωq " }ΘC,1a}Ap }a}uAp The spaces uAp,q uAppCd2Cd1q pωqpCd2Cd1q, and their norms, and the scalar product p , quA2 are defined analo- uAp,qpCd2Cd1q and uAp pωqpCd2Cd1q, a P uAp gously. 1.6. Pseudo-differential operators. Next we recall some properties in pseudo-differential calculus. Let Mpd, Ωq be the set of d d-matrices with entries in the set Ω, a P Σ1pR2dq, and let A P Mpd, Rq be fixed. Then the pseudo-differential operator OpApaq is the linear and contin- uous operator on Σ1pRdq, given by pOpApaqfqpxq " p2πq´dij apx´Apx´yq, ξqfpyqeixx´y,ξy dydξ, x P Rd. (1.37) 1pR2dq, the pseudo-differential operator OpApaq is 1pRdq with distri- For general a P Σ1 defined as the continuous operator from Σ1pRdq to Σ1 bution kernel Ka,Apx, yq " p2πq´d{2pF ´1 x, y P Rd. (1.38) Here F2F is the partial Fourier transform of Fpx, yq P Σ1 1pR2dq with respect to the y variable. This definition makes sense since the map- pings 2 aqpx ´ Apx ´ yq, x´ yq, F2 are homeomorphisms on Σ1 a homeomorphism on Σ1 and Fpx, yq ÞÑ Fpx ´ Apx ´ yq, x ´ yq (1.39) 1pR2dq. In particular, the map a ÞÑ Ka,A is The standard (Kohn-Nirenberg) representation, apx, Dq " Oppaq, and the Weyl quantization Opwpaq of a are obtained by choosing A " 0 and A " 1 2 I, respectively, in (1.37) and (1.38), where I is the identity matrix. 1pR2dq. Remark 1.18. By Fourier's inversion formula, (1.38) and the kernel the- orem [16, Theorem 2.2], [22, Theorem 2.5] for operators from Gelfand- Shilov spaces to their duals, it follows that the map a ÞÑ OpApaq is bijective from Σ1 1pR2dq to the set of all linear and continuous operators from Σ1pRdq to Σ1 1pR2dq. By Remark 1.18, it follows that for every a1 P Σ1 Mpd, Rq, there is a unique a2 P Σ1 By Section 18.5 in [15], the relation between a1 and a2 is given by 1pR2dq and A1, A2 P 1pR2dq such that OpA1pa1q " OpA2pa2q. OpA1pa1q " OpA2pa2q ô a2 " eixpA1´A2qDξ,Dxya1. 18 (1.40) Here we note that the operator eixADξ ,Dxy is homeomorphic on Σ1pR2dq and its dual (cf. [5,6,29]). For modulation spaces we have the following subresult of Proposition 2.8 in [28]. Proposition 1.19. Let s ě 1 Σ1pR2dq and let TA " eixADξ,Dxy. If ω P PEpR4dq and ωApx, ξ, η, yq " ωpx ` Ay, ξ ` Aη, η, yq, 2, A P Mpd, Rq, p, q P p0,8s, φ, a P then TA from Σ1pR2dq to Σ1pR2dq extends uniquely to a homeomor- phism from M p,q pωqpR2dq to M p,q pωAqpR2dq, and }TAa}M p,q pωAq -- }a}M p,q pωq . (1.41) 2. Kernel theorems and analytic pseudo-differential operators s) and mappings with kernels in uAs (uA1 In the first part of the section we show that there is a one to one correspondence between linear and continuous mappings from A1 s to As (As to A1 s) with respect to the measure dµ (cf. Propositions 2.2 and 2.3). Thereafter we deduce in Theorems 2.7 -- 2.10 analogous results for analytic pseudo-differential operators based on Theorem 2.6 which deals with mapping properties of the operator which takes apz, wq into epz,wqapz, wq. Here and in what follows, any extension of the A2-form, p , qA2 from A0pCdq A0pCdq to C is still called A2-form and still denoted by p , qA2. Similar approaches yield extensions of the forms x , yA2 and p , quA2. 0,spNdq are the duals of ℓspNdq and ℓ0,spNdq, respectively, through unique extensions of the ℓ2pNdq form on ℓ0pNdq. Since the spaces in (1.16) are images of the spaces in (1.12) under the map TA in (1.14), the following lemma is an immediate consequence of these duality properties. The result is also implicitly given in [7, 26]. By the definitions, ℓ1 spNdq and ℓ1 Lemma 2.1. Let s P R5. Then the following is true: (1) the form pF, Gq ÞÑ pF, GqA2 from A0pCdq A0pCdq to C is spCdq 0,spCdq to C. Furthermore, the spCdq and uniquely extendable to continuous forms from AspCdq A1 to C, and from A0,spCdq A1 duals of AspCdq and A0,spCdq can be identified by A1 A1 0,spCdq through the form p , qA2; (2) the form pF, Gq ÞÑ xF, GyA2 from A0pCdq A0pCdq to C is spCdq 0,spCdq to C. Furthermore, the spCdq and uniquely extendable to continuous forms from AspCdq A1 to C, and from A0,spCdq A1 duals of AspCdq and A0,spCdq can be identified by A1 A1 0,spCdq through the form x , yA2. 19 The following two propositions follow by applying VΘ,d2,d1 on Theo- rem 3.3 and 3.4 in [7], and using Lemma 2.1. The details are left for the reader. Proposition 2.2. Let s P R5, and let T be a linear and continuous map from A0pCd1q to A1 spCd1q to AspCd2q, 0pCd2q. Then the following is true: (1) if T is a linear and continuous map from A1 then there is a unique K P uAspCd2 Cd1q such that T F "`z2 ÞÑ xKpz2, q, FyA2pCd1 q holds true; (2.1) (2) if T is a linear and continuous map from AspCd1q to A1 spCd2q, spCd2 Cd1q such that (2.1) holds A1 true. then there is a unique K P uA1 The same holds true if As, uAs, A1 0,s and uA1 Proposition 2.3. Let K P uA1 0,s, respectively, at each occurrence. s and uA1 s are replaced by A0,s, uA0,s, 0pCd2 Cd1q, s P R5 and let T be the (2.2) linear and continuous map from A0pCd1q to A1 0pCd2q, given by F ÞÑ T F "`z2 ÞÑ xKpz2, q, FyA2pCd1 q. Then the following is true: continuous map from A1 spCd1q to AspCd2q; continuous map from AspCd1q to A1 spCd2 Cd1q, then T extends uniquely to a linear and (1) if K P uAspCd2 Cd1q, then T extends uniquely to a linear and (2) if K P uA1 s are replaced by A0,s, uA0,s, The same holds true if As, uAs, A1 0,s and uA1 F P A0pCd1q, G P A0pCd2q. Next we recall the definition of analytic pseudo-differential operators. The operator T in (2.2) should be interpreted as T in the formula pT F, GqA2pCd2 q " pK, GbFqA2pCd2 Cd1 q, 0,s, respectively, at each occurrence. s and uA1 spCd2q. A1 (See [25, Definition 6.20] in the case t " 0, as well as [3, 4].) differential operator OpVpaq (AΨDO) with symbol a is given by 51pCd Cdq. Then the analytic pseudo- Definition 2.4. Let a P uA1 pOpVpaqFqpzq "żCd apz, wqFpwqepz,wq dµpwq By the definition it follows that the relation between the operator " pF, apz, qep ,zqqA2pCdq, z P Cd. (2.3) kernel K and the symbol a is given by Kpz, wq " epz,wqapz, wq, z, w P Cd, provided the multiplication on the right-hand side makes sense. 20 This leads to the question about mapping properties of Tt defined by pTtaqpz, wq " etpz,wqapz, wq, z, w P Cd, t P C spC2dq. (2.4) the inverse of Tt is T´t. Hence Tt is well-defined and a homeomorphism when a belongs to a suitable subspace of uA1 First we notice that a P uApC2dq, if and only if Tta P uApC2dq, and that on uA1 If TH is the same as in (1.14) then we shall investigate the map T0,t 51pC2dq. in the commutative diagram: ℓ1 51pN2dq 51pN2dq T0,tÝÝÝÑ ℓ1 TH§§đ uA51pC2dq ÝÝÝÑTt §§đTH uA51pC2dq. 51pC2dq with the expansion cpα, βq cpα, βqeαpzqeβpwq " ÿα,βPNd cpα, βq À rα`βaα!β! tγzγwγ γ! , z, w P Cd, (2.5) zαwβ ?α!β! , z, w P Cd, (2.6) Therefore, let a P uA1 apz, wq " ÿα,βPNd where for every r ą 0. Since we have etpz,wq " ÿγPNd etpz,wqapz, wq " ÿγPNd ÿα,βPNd where ϕt,z,wpα, β, γq, z, w P Cd, t P C, (2.7) ϕt,z,wpα, β, γq " cpα, βqtγzα`γwβ`γ γ!?α!β! , z, w P Cd, t P C, (2.8) We shall prove that the series in (2.7) is locally uniformly convergent with respect to t, z and w. If t ă R, z ă R and w ă R for some fixed R ą 0, then by (2.6) we get ϕt,z,wpα, β, γq À γ! for all α, β, γ P Nd. Since the series rα`βRα`β`2γ ď prRqα`βpdR2qγ γ! ÿα,β,γPNd prRqα`βp2R2qγ γ! is convergent when r is chosen strictly smaller than R´1, the asserted uniform convergence follows from Weierstass' theorem. 21 In particular, we may change the order of summation in (2.7) to obtain pTtaqpz, wq " ÿα,βPNd ÿγPNd " ÿα,βPNd ÿγPNd cpα, βqtγzα`γ wβ`γ γ!?α!β! γ 1{2 γ β ` γ cpα, βqtγα ` γ " ÿα,βPNdpT0,tcqpα, βqeαpzqeβpwq, γβ cpα ´ γ, β ´ γqtγα γ1{2 where pT0,tcqpα, βq " ÿγďα,β eα`γpzqeβ`γpwq z, w P Cd, t P C, , t P C, (2.9) and we have identified T0,t in the diagram (2.5). We have now the following: Proposition 2.5. Let t P C, s, s0 P R5 be such that s ă 1{2 and 0 ă s0 ď 1{2, and let T0,t be the map on ℓ1 0pN2dq given by (2.9). Then T0,t is a continuous and bijective map on ℓ1 0pN2dq with the inverse T0,´t. Furthermore, T0,t restricts to homeomorphism from ℓ1 spN2dq, and from ℓ1 spN2dq to ℓ1 0,s0pN2dq to ℓ1 Proof. The topology on ℓ1 norms 0,s0pN2dq. 0pN2dq can be defined by the family of semi- pNptcpα, βquα,βPNdq " sup 0pN2dq we have αďN Then, for a given c P ℓ1 pNpT0,tpcqq ď sup α,βďN ÿγďα,β βďN cpα, βq, N P N. sup γβ γ1{2 cpα ´ γ, β ´ γqtγα ď pNpcq ÿγďN tγ2N ď 2Np1 ` tqN dpNpcq, 0pN2dq follows. t P C, N P N, and the continuity of T0,t on ℓ1 By straight-forward computations it also follows that T0,´t is the inverse of T0,t, which gives asserted homeomorphism properties of T0,t on ℓ1 0pN2dq. Next we consider the case when s, s0 P R`. Assume that cpα, βq À e 1 h pα1{2s`β1{2sq, α, β P Nd, 22 1 e where for some constant h ą 0. Then h pα´γ1{2s`β´γ1{2sqtγα pT0,tcqpα, βq À ÿγďα,β Ipαq "ÿγďαα h pα´γ1{2sqtγ¸1{2 γe γtγ¸1{2 h α1{2sÿγďαα and similarly for Ipβq. Since Ipαq ď e " e h α1{2s , 1 1 2 we get γβ γ1{2 ď IpαqIpβq, t P C, α P Nd, p1`tqdα{2, t P C, α P Nd, pT0,tcqpα, βq À e 1 h pα1{2s`β1{2sqp1 ` tqdpα`βq{2 À e 2 h pα1{2s`β1{2sq, t P C, α, β P Nd, where the last inequality follows from the fact that s ă 1{2. This gives the continuity assertions for T0,t in the case when s, s0 P R` and s, s0 ă 1{2. For s0 " 1{2 we have 1 pT0,tcqpα, βq À e for some other choice of h1 ą 0 which only depend on t, d and h and the continuity of T0,t on ℓ1 h pα`βqp1`tqdpα`βq{2 " e t P C, α, β P Nd, pα`βq, 1 h1 0, 1 2pN2dq follows. It remains to consider the case when s " s0 " 5σ for some σ ą 0. Assume that cpα, βq ď Crα`β pα!β!q 1 2σ , α, β P Nd, 1 2σ tγα γβ γ1{2 for some constants C, r ą 0. Then rα`β´2γ ppα ´ γq!pβ ´ γq!q pT0,tcqpα, βq ď C ÿγďα,β ď Cr ÿγďα,β ď Crp2rqα`β pα!β!q 1 rα`β pα!β!q 2σ tγ2α`β{2 2σ ÿγďα`β ď Crp2rp1 ` tqdqα`β pα!β!q tγ 1 1 t P C, α, β P Nd, where Cr ą 0 only depends on C and r. This shows that T0,t is contin- uous on ℓ1 2σ , (cid:3) 5σpN2dq and on ℓ1 0,5σpN2dq. We have now the following: 23 0,s0pC2dq. spC2dq to uA1 Proof. By the commutative diagram (2.5) we have 0,1{2pC2dq to uA1 Theorem 2.6. Let t P C, s, s0 P R5 be such that s ă 1 true: 2 and 0 ă s0 ď 1 2, 51pC2dq. Then the following is 0,1{2pC2dq; 0,1{2pC2dq extends uniquely to home- 0,s0pC2dq to and let Tt be given by (2.4) when a P uA1 (1) Tt restricts to a homeomorphism from uA1 (2) Tt from uA1 omorphisms from uA1 uA1 0,1{2pC2dq to uA1 spC2dq and from uA1 a P uA51pC2dq, 1 qa for general a P uA1 and letting Tta " pT1 T0,t T ´1 nuity assertions follow from Proposition 2.5. It remains to prove the uniqueness. Let a P uA1 respectively, and let cTtapα, βq be the coefficients of Tta P uA1 with the corresponding expansion coefficients capα, βq, and cbpα, βq, 0pC2dq, α, β P Nd. Then 0pC2dq, b P uA0pC2dq, Tta " pT1 T0,t T ´1 1 qa, 0pC2dq, the conti- pa, bquA2 " ÿα`βďN capα, βqcbpα, βq, for some N P N depending on b. Now choose a sequence aj P uA0pC2dq such that (2.10) lim jÑ8paj, bquA2 " pa, bquA2, If cajpα, βq denote the coefficients in the expansion of aj, j P N, then it follows from (2.10) for every b P uA0pC2dq. for every α, β P Nd lim jÑ8 cajpα, βq " capα, βq, (2.11) by taking bpz, wq " eαpzqeβpwq. The uniqueness follows if we prove that lim jÑ8ppTtajq, bquA2 " ppTtaq, bquA2, Let the coefficients of Ttaj be denoted by ct,ajpα, βq, α, β P Nd. By (2.9) and (2.11) we get ct,ajpα, βq Ñ ct,apα, βq as j Ñ 8, for every pα, βq P N2d, and (2.12) follows since for every b P uA0pC2dq. (2.12) ppTtajq, bquA2 " ÿα`βďN ct,ajpα, βqcbpα, βq, and ppTtaq, bquA2 " ÿα`βďN where N depends on b only. ct,apα, βqcbpα, βq, b P uA0pC2dq, (cid:3) The following two theorems now follows by combining Propositions 2.2 and 2.3 with Theorem 2.6. The details are left for the reader. 24 A0,s, A1 a P uA1 Theorem 2.8. Let a P uA1 If a P uA1 0,s and uA1 0,s and uA1 s and uA1 s and uA1 Theorem 2.7. Let s P R5 be such that s ă 1 and continuous map from AspCdq to A1 spCd Cdq such that T " OpVpaq. 2, As, A1 The same holds true if s ă 1 2 and let T be a linear spCdq. Then there is a unique s are replaced by s ď 1 2, 0,s, respectively, at each occurrence. 0pCd Cdq and s P R5 be such that s ă 1 2. spCd Cdq, then OpVpaq extends uniquely to a linear and s are replaced by s ď 1 2, continuous map from AspCdq to A1 A0,s, A1 The same holds true if s ă 1 0,s, respectively, at each occurrence. spCdq. 2, As, A1 The analogous results to Theorems 2.7 and 2.8 for larger s are equiv- alent to kernel theorems for Fourier invariant Gelfand-Shilov spaces. Theorem 2.9. Let s ě 1 2 ). Then the following is true: 2 (s ą 1 (1) If T is a linear and continuous map from A1 spCdq to AspCdq (from A1 Cdq such that 0,spCdq to A0,spCdq), then there is a unique a P uApCd 1 apz, wq À e 1 2 z´w2´rpz 1 s `w s q, z, w P Cd, (2.13) for some (for every) r ą 0 and T " OpVpaq; (2) If T is a linear and continuous map from AspCdq to A1 spCdq (from A0,spCdq to A1 Cdq such that 0,spCdq), then there is a unique a P uApCd 1 apz, wq À e 1 2 z´w2`rpz 1 s `w s q, z, w P Cd, (2.14) for every (for some) r ą 0 and T " OpVpaq. Theorem 2.10. Let s ě 1 2 (s ą 1 2 ). Then the following is true: 0, then OpVpaq from A0pCdq to A1 to a linear and continuous map from A1 A1 0,spCdq to A0,spCdq); (1) If a P uApCd Cdq satisfies (2.13) for some (for every) r ą (2) If a P uApCd Cdq satisfies (2.14) for every (for some) r ą 0, then OpVpaq from A0pCdq to A1 to a linear and continuous map from AspCdq to A1 A0,spCdq to A1 0pCdq is uniquely extendable spCdq (from 0pCdq is uniquely extendable spCdq to AspCdq (from 0,spCdq). Proof. We only prove the results in the Roumieu case. The Beurling case follows by similar arguments and is left for the reader. If T is the same as in (2.2) for some K P uApCd Cdq, then T " OpVpaq when apz, wq " e´pz,wqKpz, wq, pz, wq P Cd Cd. Since 1 e´pz,wqe 2 pz2`w2q " e´ 1 25 2 z´w2 , z, w P Cd, Theorem 1.9 gives K P uAspCd Cdq ô Kpz, wq À e ô apz, wq À e for some r ą 0. In the same way, 1 2 pz2`w2q´rpz 1 s `w 1 s q 1 2 pz´w2q´rpz 1 s `w 1 s q, z, w P Cd. 1 s `w 1 1 s q, 2 pz´w2q`rpz spCd Cdq ô apz, wq À e z, w P Cd, for every r ą 0. The results now follows from these relations and Propositions 2.2 and 2.3 K P uA1 Remark 2.11. For strict subspaces of uA1 0,1{2pCdq in Definition 1.7, the estimates imposed on their elements are given by (1.23) or by (1.24) for suitable assumptions on r ą 0. It is evident that in all such cases, these conditions are violated under the action of Tt in Theorem 2.6 when t ‰ 0. Hence, Theorem 2.6 cannot be extended to other spaces in Definition 1.7. In particular, the conditions (2.13) and (2.14) in Theorems 2.9 and 2.10 can not be replaced by the convenient condition that a should belong to e. g. (cid:3) 2 and s2 ą 1 uAspCd Cdq, uA0,spCd Cdq, uA1 s1pCd Cdq or uA1 0,s2pCd Cdq (2.15) when s P R5, s1 ě 1 2. On the other hand, the conditions on a in Theorems 2.9 and 2.10 means exactly that pz, wq ÞÑ epz,wqapz, wq belongs to the spaces in (2.15), depending on the choice between (2.13) and (2.14), and the condition on r. Remark 2.12. Let s P R5 be such that s ď 1 2 . By similar arguments as in the proofs of Theorems 2.9 and 2.10, one may also characterize linear and continuous operators from A1 0,spCdq to A0,spCdq as operators of the form OpVpaq for suitable conditions on a. The details are left for the reader. spCdq to AspCdq, and from A1 3. Operators with kernels and symbols in mixed weighted Lebesgue spaces In this section we focus on operators in the previous section, whose kernels should belong to uApCd Cdq and obey certain mixed norm estimates of Lebesgue types. We deduce continuity properties of such operators when acting between suitable Lebesgue spaces of analytic functions. (See Theorems 3.3 -- 3.5.) Thereafter we show that our results can be used to regain well-known and sharp continuity results in [27] for pseudo-differential operators with symbols in modulation spaces when acting on other modulation spaces. (See Theorems 3.8 and 3.9.) A key step here is to deduce an explicit formula which relates the short-time Fourier transform of the symbol to a real pseudo-differential operator 26 Oppaq with the Bargmann transform of the kernel to Oppaq. Lemma 3.7.) (See We shall consider Lebesgue norm conditions of matrix pull-backs of the involved kernels. Let C l C 3 j,k C 2 j,k j,k C 4 jk P Mpd, Rq, Cjk "C 1 C21 C22¸ P Mp4d, Rq, C "C11 C12 Udpx ` iξq " px, ξq P R2d, j,k¸ P Mp2d, Rq, j, k, l P Z`, and let Ud,dpx1 ` iξ1, x2 ` iξ2q " px1, ξ1, x2, ξ2q P R4d, ?2 z,?2 wq, Kωpz, wq " e´ 1 2 pz2`w2qKpz, wq ωp (3.2) GK,C,ω " Kω U ´1 d,d C Ud,d, z, w P Cd, j " 1, 2. x, xj, ξ, ξj P Rd, E,pω2qpCdq, We will consider continuity of operators from Ap1 when GK,C,ω fullfils suitable Lp,qpC2dq estimates, where the weights full- fil E,pω1qpCdq to Ap2 (3.1) (3.3) (3.4) (3.5) (3.6) pC2dq be the sets of pC2dq " }GUd,d}Lp,q pR4dq, ω2pzq ω1pwq À ωpz, wq, z, w P Cd. Here and in what follows we let Lp,qpC2dq and Lp,q all G P L1 and }G}Lp,q }G}Lp,qpC2dq " }GUd,d}Lp,qpR4dq respectively, are finite. (See also Remark 1.16.) locpC2dq such that The involved Lebesgue exponents should satisfy 1 p1 ´ 1 p2 " 1 ´ 1 p ´ 1 q , q ď p2 ď p, p, q P r1,8s, p1, p2 P r1,8s2d. We need that C and Cjk above should satisfy detpCq detpC11C21q ‰ 0 or In (3.4) and in what follows we use the convention detpCq detpC12C22q ‰ 0. 1 p " 1 p1 , . . . , 1 pd , p0 ď p, p ď p, q0 ă q, q ă q and r " r, 27 when p " pp1, . . . , pdq, q " pq1, . . . , qdq, r " pr1, . . . , rdq belong to r1,8sd and p, q, r, p0, q0, r0 P r1,8s satisfy k P t1, . . . , du. and rk " r, p0 ď pk, Remark 3.1. We notice that (3.1) -- (3.6) implies that C is invertible and that one of the following conditions hold true: q0 ă qk, pk ď p, qk ă q (1) both C11 and C21 are invertible; (2) both C12 and C22 are invertible. If (1) holds, then 11 C12 C21 C22¸ -- detI2d C ´1 detC11 C12 " detI2d 21 C22¸ 11 C12¸ " detpC ´1 11 C12 21 C22 ´ C ´1 I2d C ´1 0 C ´1 C ´1 21 C22 ´ C ´1 11 C12q. Here and in what follows, I " Id is the d d identity matrix. From these computations it follows that C ´1 11 C12 ´ C ´1 11 ´ C ´1 C12C ´1 21 C22, 21 C22, C ´1 11 C12 ´ C22C ´1 11 ´ C22C ´1 C12C ´1 21 21 are invertible when (1) holds, and 22 C21, 12 C11 ´ C ´1 C ´1 12 ´ C ´1 C11C ´1 are invertible when (2) holds. Remark 3.2. Let C0 P Mpd, Cq and let Ud be the same as in (3.2). Then the matrix Ud C0 U ´1 in Mp2d, Rq which corresponds to C0 is given by 12 C11 ´ C21C ´1 C ´1 12 ´ C21C ´1 C11C ´1 22 C21, 22 22 d Obviously, the map which takes C0 P Mpd, Cq into the matrix (3.7) in Mp2d, Rq is injective, but not bijective. In this way we identify Mpd, Cq with the set of all matrices in Mp2d, Rq which are given by (3.7) for some C0 P Mpd, Cq. d Cjk Ud, then GK,C,ω in (3.2) is given by If Cjk P Mp2d, Rq and Tjk " U ´1 GK,C,ωpz, wq " KωpT11pzq ` T12pwq, T21pzq ` T22pwqq, 28 z, w P Cd. RepC0q¸ . RepC0q ´ ImpC0q ImpC0q (3.7) If more restricted, Cjk can be identified as matrices in Mpd, Cq as above, for j, k P t1, 2u, then GK,C,ω in (3.2) is given by GK,C,ωpz, wq " KωpC11z ` C12w, C21z ` C22wq, for such choices of C. z, w P Cd, Theorem 3.3. Let E be an ordered basis for R2d, ω1 and ω2 be weights on Cd, ω be a weight on Cd Cd such that (3.3) holds, and let p1, p2, p and q be as in (3.4). Also let C P Mp4d, Rq be such that (3.1) holds, true: K P uApCd Cdq, and let GK,C,ω be as in (3.2). Then the following is (1) if (3.5) holds and GK,C,ωpz, wq P Lp,qpC2dq, then TK in (2.1) from A51pCdq to ApCdq is uniquely extendable to a continuous mapping from Ap1 E,pω1qpCdq to Ap2 }TKF}Ap2 E,pω2q À }GK,C,ω}Lp,q}F}Ap1 (2) if (3.6) holds and GK,C,ω P Lq,p E,pω1q E,pω2qpCdq, and , F P Ap1 E,pω1qpCdq; (3.8) GK,C,ωpz, wqΦpz, wq dλpzqdλpwq À }GK,C,ω}Lp,q}Φ}Lp1,q1 , 29 pC2dq, then TK in (2.1) from A51pCdq to ApCdq is uniquely extendable to a continuous map- ping from Ap1 E,pω1qpCdq to Ap2 E,pω2qpCdq, and }TKF}Apω2,Lp2 q À }GK,C,ω}Lq,p 2 }F}Ap1 E,pω1q , F P Ap1 E,pω1qpCdq. (3.9) Proof. We only prove (1). The assertion (2) follows by similar argu- ments and is left for the reader. Let GK,C,ω,ppwq " }GK,C,ωp , wq}LppCdq, w P Cd. Then }GK,C,ω}Lp,qpC2dq -- }GK,p,ω}LqpCdq. Also let Kω be as in (3.2), F P Ap1 E,pω1qpCdq and H P B p2 E,p1{ω2qpCdq, and set Fω1pwq " Fpwqe´w2{2ω1p ?2 wq, w P Cd and Hω2pzq " Hpzqe´z2{2{ω2p ?2 zq, z P Cd. By Holder's inequality we get pT F, HqB2 -- żCdpT FqpzqHpzqe´z2 dλpzq Kωpz, wqFω1pwqHω2pzq dλpzqdλpwq ďij C2d "ij C2d where Φpx ` iξ, y ` iηq " Fω1pU ´1 d pC21px, ξq ` C22py, ηqqqHω2pU ´1 d pC11px, ξq ` C12py, ηqqq, x, y, ξ, η P Rd. Here we identify px, ξq P R2d by corresponding 2d 1-matrix` x ξ, as usual. We need to estimate }Φ}Lp1,q1 , and start with reformulating }Φp , wq}Lp1 . For }Φp , wq}Lp1 we take px, ξq ÞÑ C21ppx, ξq ` C ´1 21 C22py, ηqq as new variables of integration, and get }Φp , wq}Lp1 -- }Fω1 Hω2pU ´1 where B1 and B2 are the matrices B1 " C11C ´1 11 C12 ´ C22 P Mp2d, Rq, which are invertible due to Remark 3.1 and the assumptions. Hence, for F 0 21 P Mp2d, Rq and B2 " C21C ´1 d pB1p ´ B2py, ηqqq}Lp1 , w " y ` iη, ω1 " Fω1pU ´1 }Φp , wq}Lp1 -- ´´F 0 ω1p1 If r1 " p1{p1 and r2 " p ω2 " Hω2 U ´1 q and H 0 ω2p1¯pB2py, ηqq¯ 1 H 0 2{p1, then it follows from (3.4) that d p´B1q we have , w " y` iη. (3.10) p1 d 1 1 r1 ` 1 r2 " 1 ` p1 q1 , and r1, r2, q1 p1 ě 1. Hence, by (3.10), that q1{p1 ě 1, the fact that B2 is invertible, and Holder's and Young's inequalities we obtain }Φ}Lp1,q1 À››››´´F 0 ω1p1 H 0 -- ›››F 0 ď´}F 0 ω1p1 p1››››Lq1 ω2p1¯ pB2 q¯ 1 E 1 ω2p1›››Lq1{p1 ω1p1 H 0 E¯ 1 E }H 0 }Lr1 ω2p1 }Lr2 p1 p1 -- }Fω1}Lp1 E }Hω2}L p1 2 E and the right-hand side of (3.8) follows by taking the supremum over all such H with }H}B E,p1{ω2q ď 1. p1 2 The existence of extension now follows from Hahn-Banach's theorem. By these estimates it also follows that pz, wq ÞÑ Kωpz, wqFω1pwqHω2pzq 30 belongs to L1pCd Cdq, and the uniqueness is a straight-forward ap- plication of Lebesgue's theorem. (cid:3) For corresponding pseudo-differential operator with symbol a, the kernel is given by Kpz, wq " epz,wqapz, wq. By straight-forward compu- tations it follows that GK,C,ω,p takes the form GK,C,ω,ppwq " }aωp , wq}LppCdq, w P Cd, when C11 " C12 " C22 " I2d and C21 " 0, or C11 " C12 " C21 " I2d and C22 " 0, where aωpz, wq " e´ 1 2 z2 apz ` w, wqωp or ?2pz ` wq,?2 wq aωpz, wq " e´ 1 apz ` w, zqωp Hence, Theorem 3.3 gives the following. 2 w2 ?2 pz ` wq,?2 zq. (3.11) (3.12) Theorem 3.4. Let ω1 and ω2 be weights on Cd, ω be a weight on Cd Cd such that (3.3) holds, p1, p2, p and q be as in (3.4). Also let If aω P Lp,qpC2dq, then the operator OpVpaq in (2.4) from A51pCdq to ApCdq is uniquely extendable to a continuous mapping from Ap1 E,pω1qpCdq to Ap2 a P uApCd Cdq and let aω be given by (3.11) or by (3.12) for z, w P Cd. E,pω2qpCdq. We also have the following result related to Theorem 3.3. Here the matrix C is given by (3.1) with C11 " C21 " I2d, C12 "0 0 Id¸ and C22 "Id 0 0¸ (3.13) 0 0 which obviously satisfies (3.5). Also again recall Remark 1.16 for no- tations. Theorem 3.5. Let C be given by (3.1) with Cjk given by (3.13), ω1 and ω2 be weights on Cd, ω be a weight on Cd Cd such that (3.3) holds, and let p, q P r1,8s. Also let K P uApCd Cdq and GK,C,ω be as in (3.2). If GK,C,ω P Lp,q pC2dq, then then TK in (2.1) from A51pCdq to ApCdq is uniquely extendable to a continuous mapping from Ap1,q1 pω1qpCdq to Aq,p ,pω2qpCdq, and }TKF}Aq,p ,pω2q À }GK,C,ω}Lp,q }F}Ap1,q1 pω1q , F P Ap1,q1 pω1qpCdq. (3.14) 31 Proof. Let Fω1, F 0 orem 3.3, and let Kω be as in (3.2). Then ω1, Hω2 and H 0 ω2 be the same as in the proof of The- pTKF, HqB2 ďżżżż Kωpx, ξ, y, ηqF 0 "żżżż GK,C,ωpx, ξ, y, ηqqF 0 ω1px ` y, ξqH 0 ω1py, ηqH 0 ω2px, ξq dxdξdydη ω2px, ξ ` ηq dxdξdydη ď }GK,C,ω}Lp,q }Φ0}Lp1 , where Φ0px, ξq "ij F 0 ω1px ` y, ξqH 0 ω2px, ξ ` ηqq1 ω1p , ξq}Lq1}H 0 " }F 0 q1 dydη 1 ω2px, q}Lq1 , x, ξ P Rd. Hence, }Φ0}Lp1 " }Fω1}Lq1,p1 1 }Hω2}Lp1,q1 2 . The continuity assertion now follows from these estimates, an applica- tion of Hahn-Banach's theorem and Lebesgue's theorem (cf. the end of the proof of Theorem 3.3). (cid:3) Remark 3.6. In Theorem 3.5, the matrix C is chosen only as (3.1) and (3.13), while Theorem 3.3 is valid for a whole family of matrices with the only restriction (3.5) or (3.6). On the other hand, by similar arguments, it follows that the conclusions in Theorem 3.5 are still true when, more generally, C P Mp4d, Rq is of the form Id 0 Id Id 0 0 Id Id 0 0 0 Id 0 Id 0 0 0 0 0 0 Id 0 0 Id 0 0 Id 0 Id 0 0 Id , ‹‹‹‹‚ ‹‹‹‹‚ or Id 0 0 Id Id 0 Id 0 0 Id Id 0 0 0 0 Id 0 0 Id 0 0 0 Id 0 0 Id 0 0 Id 0 0 Id. , ‹‹‹‹‚ ‹‹‹‹‚ , for any choice of at each place, provided the mixed Lebesgue condi- tions on GK,C,ω are slightly modified. In order to apply Theorem 3.3 to real pseudo-differential operators we have the following. 32 2 px2`ξ2q, x, ξ P Rd, a P H1 51pR2dq Lemma 3.7. Let φpx, ξq " π´ d 2 eixx,ξye´ 1 and let Ka be the kernel of Oppaq. Then e´ 1 2 pz2`w2qVΘ,dKapz, wq " 2 2 e´ipxx,ξ´2ηy`xy,ηyqpVφaqp ?2x,´ when z " x ` iξ P Cd and w " y ` iη P Cd. Proof. Let φ0px, ξq " π´ d putations and Fourier's inversion formula we get 2 e´ 1 d ?2η,?2pη ´ ξq,?2py ´ xqq 2 px2`ξ2q " e´ixx,ξyφpx, ξq. By formal com- ?2, w{ ?2q p4π3q d 2 e i 2 pxx,ξy`xy,ηyqe´ 1 4 pz2`w2qVΘ,dKapz{ d d d " p2πq " p2πq "¡ apx1, ξ1qeixx1´y1,ξ1yφ0px1 ´ x, y1 ´ yqe´ipxy1,ηy´xx1,ξyq dx1dy1dξ1 2 e´ixy,ηyij apx1, ξ1qφ0px1´x, ξ1`ηqeixx1,ξ1ye´ipxy,ξ1y´xx1,ξyq dx1dξ1 2 eixx´y,ηyij apx1, ξ1qφpx1 ´ x, ξ1 ` ηqe´ipxx1,η´ξy`xy´x,ξ1yq dx1dξ1 2 eixx´y,ηypVφaqpx,´η, η ´ ξ, y ´ xq. (cid:3) We can now use the previous lemma and theorems to obtain mapping properties for pseudo-differential operators with symbols in modulation spaces. For example, we may combine Lemma 3.7 and Theorem 3.3 to deduce the following result, which is the same as [28, Theorem 2.2]. Hence our kernel results on the Bargmann transform side can be used to regain classical mapping properties pseudo-differential operators when acting on modulation spaces. Theorem 3.8. Let E be an ordered basis of R2d, A P MpR, dq, p, q P r1,8s and p1, p2 P r1,8s2d be as in (3.4), ω0 P PEpR4dq and ω1, ω2 P PEpR2dq be such that " p4πq ω2px ´ Ay, ξ ` pI ´ Aqηq ω1px ` pI ´ Aqy, ξ ´ Aηq À ω0px, ξ, η, yq, x, y, ξ, η P Rd, (3.15) and let a P M p,q uniquely to continuous operator from M p1 pω0qpR2dq. Then OpApaq from Σ1pRdq to Σ1 E,pω1qpRdq to M p2 1pRdq extends E,pω2qpRdq, and } OpApaq}M p1 E,pω1qÑM p2 E,pω2q À }a}M p,q pω0q . Proof. By (1.40) and Proposition 1.19 we may assume that A " 0. Let and let ω be given by C11 " C21 " C22 " I2d, C12 " 0 ωpx, ξ, y, ηq " ω0px,´η, ξ ` η, y ´ xq, 33 x, y, ξ, η P Rd, which we identify with ωpz, wq, z " x ` iξ P Cd, w " y ` iη P Cd. Then it follows by straight-forward computations that (3.15) is the same as (3.3). Furthermore, let K0 " VΘ,dKa, where Ka is the kernel of the operator Oppaq. Then it follows from Lemma 3.7 and straight- forward computations that if Ha,ω0px, ξ, η, yq " Vφapx, ξ, η, yq ω0px, ξ, η, yq, , x, y, ξ, η P Rd, then ?2 x,´ ?2pξ ` ηq,?2 η,?2 yq -- GK0,C,ωpz, wq, Ha,ω0p (3.16) z " x ` iξ P Cd, w " y ` iη P Cd. By first applying the Lp-norm on (3.16) with respect to x and ξ, and thereafter applying the Lq-norm with respect to y and η, we get }GK0,C,ω}Lp,qpC2dq -- }Ha,ω0}Lp,qpR4dq -- }a}M p,q pω0q ă 8. Hence, the assumptions in Theorem 3.3 are fullfiled, and we conclude E,pω1qpCdq to that the operator TK0 with kernel K0 is continuous from Ap1 E,pω2qpCdq. The asserted continuity for Oppaq is now a consequence Ap2 of the commutative diagram E,pω1qpRdq ÝÝÝÑ M p2 E,pω2qpRdq M p1 Oppaq (3.17) Vd§§đ E,pω1qpCdq ÝÝÝÑTK0 Ap1 E,pω2qpCdq. (cid:3) Ap2 §§đVd The next result extends [24, Theorem 3.3] and follows by similar arguments as in the previous proof, using Theorem 3.5 instead of The- orem 3.3. The details are left for the reader. Theorem 3.9. Let ω1 and ω2 P PEpR2dq, ω P PEpR4dq be such that ω2px, ξ ` ηq ω1px ` y, ξq À ω0px, ξ, η, yq, , x, y, ξ, η P Rd, let p, q P r1,8s, and let a P W p,q 1pRdq is uniquely extendable to a continuous mapping from M q1,p1 Σ1 to W p,q pω0qpR2dq. Then Op0paq from Σ1pRdq to pω1qpRdqq pω2qpRdqq, and } Oppaqf}W p,q pω2q À }a}W p,q pω0q}f}M q1,p1 pω1q , f P M q1,p1 pω1qpRdqq. Remark 3.10. Let E, p, q, pj, φ0 and φ be the same as in Lemma 3.7, Theorem 3.8 and their proofs. Also let A P MpR, dq and let φA " eixADξ ,Dxyφ. Then the condition on a in Theorem 3.8 is that }Vφ0a ω0}Lp,q ă 8. In view of [5, 6, 29] and Proposition 1.13 (2), the 34 previous condition is the same as }VφAa ω0}Lp,q ă 8 because ω0 is moderate. We observe that all weights in Theorem 3.8 are moderate, while there are no such assumptions or other restrictions on the involved weight functions in Theorem 3.3. Since the latter result is used to prove the former one, a natural question is wether Theorem 3.8 can be extended to broader classes of weight functions. In view of Remark 1.14, it is evident that the imposing moderate conditions on weights might in some context be considered as strong restrictions. The answer on this question is affirmative in the sense that for suit- able modifications, the moderate conditions on the weights in Theorem 3.8 can be removed. In fact, let HA 51pR2dq be the modification of H51pR2dq, given by HA 51pR2dq " t eixADξ ,Dxypeixx,ξyaq ; a P H51pR2dqu, 51q1pR2dq be the dual of HA 51pR2dq, ω1, ω2 be weights on R2d and let pHA ω0 be a weight on R4d such that (3.15) holds. Then it follows from the proof of Theorem 3.8 that the following is true: ‚ if a P pHA ‚ if a P pHA H51pRdq to H1 from M p1 51q1pR2dq, then VφAa makes sense as a smooth function; 51q1pR2dq satisfies }VφAaω0}Lp,q ă 8, then OpApaq from 51pRdq extends uniquely to a continuous operator E,pω1qpRdq to M p2 E,pω2qpRdq. In similar ways, Theorem 3.9 can be extended to permit more general weight classes. References [1] V. Bargmann On a Hilbert space of analytic functions and an associated inte- gral transform, Comm. Pure Appl. Math., 14 (1961), 187 -- 214. [2] V. Bargmann On a Hilbert space of analytic functions and an associated in- tegral transform. Part II. A family of related function spaces. Application to distribution theory., Comm. Pure Appl. Math., 20 (1967), 1 -- 101. [3] W. Bauer Berezin-Toeplitz quantization and composition formulas, J. Funct. Anal., 256 (2007), 3107 -- 3142. [4] F. A. Berezin Wick and anti-Wick symbols of operators, Mat. Sb. (N.S.), 86 (1971), 578 -- 610. [5] M. Cappiello, J. Toft, Pseudo-differential operators in a Gelfand -- Shilov setting, Math. Nachr. 290 (2017), 738 -- 755. [6] E. Carypis, P. Wahlberg, Propagation of exponential phase space singularities for Schrodinger equations with quadratic Hamiltonians, J. Fourier Anal. Appl. 23 (2017), 530 -- 571. [7] Y. Chen, M. Signahl, J. Toft Factorizations and singular value estimates of operators with Gelfand-Shilov and Pilipovi´c kernels, J. Fourier Anal. Appl. 24 (2018), 666 -- 698. [8] E. Cordero, S. Pilipovi´c, L. Rodino, N. Teofanov Quasianalytic Gelfand-Shilov spaces with applications to localization operators, Rocky Mt. J. Math. 40 (2010), 1123-1147. 35 [9] H. G. Feichtinger Banach spaces of distributions of Wiener's type and inter- polation, in: Ed. P. Butzer, B. Sz. Nagy and E. Gorlich (Eds), Proc. Conf. Oberwolfach, Functional Analysis and Approximation, August 1980, Int. Ser. Num. Math. 69 Birkhauser Verlag, Basel, Boston, Stuttgart, 1981, pp. 153 -- 165. [10] H. G. Feichtinger Modulation spaces on locally compact abelian groups. Techni- cal report, University of Vienna, Vienna, 1983; also in: M. Krishna, R. Radha, S. Thangavelu (Eds) Wavelets and their applications, Allied Publishers Private Limited, NewDehli Mumbai Kolkata Chennai Hagpur Ahmedabad Bangalore Hyderbad Lucknow, 2003, pp. 99 -- 140. [11] C. Fernandez, A. Galbis, J. Toft The Bargmann transform and powers of har- monic oscillator on Gelfand-Shilov subspaces, RACSAM 111 (2017), 1 -- 13. [12] K. Grochenig Foundations of Time-Frequency Analysis, Birkhauser, Boston, 2001. [13] K. Grochenig Weight functions in time-frequency analysis in: L. Rodino, M. W. Wong (Eds) Pseudodifferential Operators: Partial Differential Equations and Time-Frequency Analysis, Fields Institute Comm., 52 2007, pp. 343 -- 366. [14] K. Grochenig, G. Zimmermann Spaces of test functions via the STFT J. Funct. Spaces Appl. 2 (2004), 25 -- 53. [15] L. Hormander The Analysis of Linear Partial Differential Operators, vol I -- III, Springer-Verlag, Berlin Heidelberg NewYork Tokyo, 1983, 1985. [16] Z. Lozanov Crvenkovi´c, D. Perisi´c Hermite expansions of elements of Gelfand Shilov spaces in quasianalytic and non quasianalytic case, Novi Sad J. Math. 37 (2007), 129 -- 147. [17] S. Pilipovi´c Generalization of Zemanian spaces of generalized functions which have orthonormal series expansions, SIAM J. Math. Anal. 17 (1986), 477 -- 484. [18] S. Pilipovi´c Tempered ultradistributions, Boll. U.M.I. 7 (1988), 235 -- 251. [19] M. Reed and B. Simon Methods of modern mathematical physics, Academic Press, London New York, 1979. [20] M. Ruzhansky, N. Tokmagambetov Nonharmonic analysis of boundary value problems, Int. Math. Res. Notices 12 (2016), 3548 -- 3615. [21] N. Teofanov Ultradistributions and time-frequency analysis in: P. Boggiatto, L. Rodino, J. Toft, M. W. Wong (eds) Pseudo-differential operators and related topics, Operator Theory: Advances and Applications 164, Birkhauser, Basel, 2006, pp. 173 -- 192. [22] N. Teofanov, Gelfand-Shilov spaces and localization operators, Funct. Anal. Approx. Comput. 7 (2015), 135 -- 158. [23] J. Toft Continuity properties for modulation spaces with applications to pseudo- differential calculus, II, Ann. Global Anal. Geom. 26 (2004), 73 -- 106. [24] J. Toft Pseudo-differential operators with symbols in modulation spaces, in: B.-W. Schulze, M. W. Wong (Eds), Pseudo-Differential Operators: Complex Analysis and Partial Differential Equations, Operator Theory Advances and Applications 205, Birkhauser Verlag, Basel, 2010, pp. 223 -- 234. [25] J. Toft The Bargmann transform on modulation and Gelfand-Shilov spaces, with applications to Toeplitz and pseudo-differential operators, J. Pseudo- Differ. Oper. Appl. 3 (2012), 145 -- 227. [26] J. Toft Images of function and distribution spaces under the Bargmann trans- form, J. Pseudo-Differ. Oper. Appl. 8 (2017), 83 -- 139. [27] J. Toft Continuity and compactness for pseudo-differential operators with sym- bols in quasi-Banach spaces or Hormander classes, Anal. Appl. 15 (2017), 353 -- 389. 36 [28] J. Toft Matrix parameterized pseudo-differential calculi on modulation spaces in: M. Oberguggenberger, J. Toft, J. Vindas, P. Wahlberg (eds) Generalized Functions and Fourier Analysis, Operator Theory: Advances and Applications 260, Birkhauser, Basel Heidelberg NewYork Dordrecht London, pp. 215 -- 235. [29] G. Tranquilli Global normal forms and global properties in function spaces for second order Shubin type operators PhD Thesis, 2013. Department of Mathematics and Informatics, University of Novi Sad, Novi Sad, Serbia E-mail address: [email protected] Department of Computer science, Mathematics and Physics, Linnaeus University, Vaxjo, Sweden E-mail address: [email protected] 37
1812.00204
1
1812
2018-12-01T13:24:18
On compressions of self-adjoint extensions of a symmetric linear relation
[ "math.FA" ]
Let $A$ be a symmetric linear relation in the Hilbert space $\gH$ with equal deficiency indices $n_\pm (A)\leq\infty$. A self-adjoint linear relation $\wt A\supset A$ in some Hilbert space $\wt\gH\supset \gH$ is called an exit space extension of $A$; such an extension is called finite-codimensional if $\dim (\wt\gH\ominus\gH)< \infty$. We study the compressions $C (\wt A)=P_\gH\wt A\up\gH$ of exit space extensions $\wt A=\wt A^*$. For a certain class of extensions $\wt A$ we parameterize the compressions $C (\wt A)$ by means of abstract boundary conditions. This enables us to characterize various properties of $C (\wt A)$ (in particular, self-adjointness) in terms of the parameter for $\wt A$ in the Krein formula for resolvents. We describe also the compressions of a certain class of finite-codimensional extensions. These results develop the results by A. Dijksma and H. Langer obtained for a densely defined symmetric operator $A$ with finite deficiency indices.
math.FA
math
ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS OF A SYMMETRIC LINEAR RELATION V.I. MOGILEVSKII Abstract. Let A be a symmetric linear relation in the Hilbert space H with equal deficiency indices n±(A) ≤ ∞. A self-adjoint linear relation eA ⊃ A in some Hilbert space eH ⊃ H is called an exit space extension of A; such an extension is called finite- codimensional if dim(eH ⊖ H) < ∞. We study the compressions C(eA) = PHeA ↾ H of exit space extensions eA = eA∗. For a certain class of extensions eA we parameterize the compressions C(eA) by means of abstract boundary conditions. This enables us to characterize various properties of C(eA) (in particular, self-adjointness) in terms of the parameter for eA in the Krein formula for resolvents. We describe also the compressions of a certain class of finite-codimensional extensions. These results develop the results by A. Dijksma and H. Langer obtained for a densely defined symmetric operator A with finite deficiency indices. 1. Introduction Assume now that A is a not necessarily densely defined symmetric operator in a Hilbert Self(A) = Self 0(A) if and only if A is densely defined. problem in the extension theory of symmetric operators. In the paper by A.V. Shtraus were recently studied in [2, 3, 4, 11, 12, 20]. In particular, it was shown in [3] that a Let H be a subspace in a Hilbert space eH and let eA be a linear relation (in particular operator) in eH. Recall that a linear relation C(eA) in H given by C(eA) = PHeA ↾ H is called a compression of eA; moreover, the relation eA and its compression C(eA) are called finite-codimensional if dim(eH ⊖ H) < ∞. Compressions of linear operators or relations finite-codimensional compression C(eA) of a self-adjoint linear relation eA is self-adjoint (for operators eA = eA∗ this fact was established earlier in [24]). space H. A self-adjoint linear relation eA ⊃ A in a Hilbert space eH ⊃ H is called an exit space extension of A. Denote by Self(A) the set of all exit space extensions eA = eA∗ of A and by Self 0(A) the set of all eA ∈ Self(A) such that eA is an operator. As is known If eA ∈ Self(A), then the compression C(eA) of eA is a symmetric extension of A. A description of all extensions eA ∈ Self(A) and their compressions C(eA) is an important [23] all extensions eA ∈ Self 0(A) of an operator A with arbitrary (equal or unequal) eA ∈ Self(A). In [22, 25] the compressions C(eA) of the extensions eA ∈ Self 0(A) we a bijective correspondence eA = eAτ between all relation-valued Nevanlinna functions deficiency indices n±(A) ≤ ∞ are parameterized by means of holomorphic operator- functions F (λ), λ ∈ C+, whose values are contractions from Ni to N−i (here N±i = ker (A∗∓i) are defect subspaces of A). In [13] the result of [23] was extended to extensions In the case n+(A) = n−(A) another parametrization of the set Self(A) is given by the Krein formula for generalized resolvents [15, 17]. This formula (see (2.36)) establishes τ = τ (λ) (λ ∈ C \ R) (τ (λ) is a linear relation in an auxiliary Hilbert space H) and all described in terms of the asymptotic behaviour of the corresponding parameter F (λ). 2010 Mathematics Subject Classification. 47A06, 47A20, 47A56, 47B25. Key words and phrases. Symmetric and self-adjoint linear relation (operator), exit space self-adjoint extension, compression, boundary triplet. 1 (1.1) τ (λ) = A + λB + Aj, λ ∈ C \ R, 1 αj − λ lXj=1 2 V.I. MOGILEVSKII extensions eA ∈ Self(A). The canonical self-adjoint extension A0 of A in (2.36) is called a basic extension. An operator-valued parameter τ = τ (λ) in (2.36) is called rational if where αj ∈ R and A = A∗, B ≥ 0 and Aj ≥ 0 are bounded operators in H. In the case parameter τ is rational. n+(A) = n−(A) < ∞ an extension eAτ ∈ Self(A) is finite-codimensional if and only if a In the paper by H. Langer and A. Dijksma [12] the compressions C(eAτ ) of extensions eAτ are investigated in terms of the parameter τ from the Krein formula (2.36). The main results of [12] can be formulated in the form of the following theorem. Theorem 1.1. Assume that A is a densely defined, closed symmetric operator in H with finite deficiency indices n+(A) = n−(A) < ∞ and A0 = A∗ 0 is the basic extension in the Krein formula (2.36). Let for simplicity a parameter τ = τ (λ), λ ∈ C \ R, in (2.36) be an operator-valued function, let eAτ ∈ Self 0(A) be the corresponding extension of A and let C(eAτ ) be the compression of eAτ . Assume also that Bτ ≥ 0 is an operator in H given by Bτ = lim y→∞ 1 iy τ (iy). Then: (i) If (1.2) lim y→∞ yIm (τ (iy)h, h) = ∞, h ∈ H, h 6= 0, then C(eAτ ) ⊂ A0. (ii) If (1.2) holds, then C(eAτ ) = A if and only if Bτ = 0. (iii) If Bτ > 0, then C(eAτ ) = A0. If (1.2) holds and C(eAτ ) = A0, then Bτ > 0. Moreover, if τ is a rational parameter (1.1), then C(eAτ ) = A0 ⇐⇒ Bτ > 0. (iv) If τ is the rational parameter (1.1), then the extension eAτ is finite-codimensional and the canonical self-adjoint extension C(eAτ ) of A corresponds in the Krein formula to the self-adjoint linear relation τ∞ in H given by τ∞ = {{h, Pker BAh + h′} : h ∈ ker B, h′ ∈ ran B}. (1.3) It is also shown in [12] that τ∞ admits the representation (1.4) τ∞ = {{h, h′} ∈ H2 : ∃h(λ) ∈ H : h = lim λ→∞ h(λ), h′ = lim λ→∞ τ (λ)h(λ)}. In the present paper we study compressions of extensions eA ∈ Self(A) of a symmetric linear relation (in particular, not necessarily densely defined symmetric operator) A with possibly infinite deficiency indices n+(A) = n−(A) ≤ ∞. Our approach is based on the theory of boundary triplets and their Weyl functions (see [9, 14, 18] and references therein). Recall [14, 18] that a collection Π = {H, Γ0, Γ1} consisting of an auxiliary Hilbert space H and linear mappings Γj : A∗ → H, j ∈ {0, 1}, is called a boundary triplet for A∗ if the mapping (Γ0, Γ1)⊤ is surjective and the abstract Green identity (2.15) is valid (here A∗ is the adjoint relation). If Π = {H, Γ0, Γ1} is a boundary triplet for A∗, then the equality (the abstract boundary conditions) (1.5) θ → Aθ := { f ∈ A∗ : {Γ0 f , Γ1 f } ∈ θ} gives a parametrization of all proper extensions eA of A (i.e., all linear relations eA in H with A ⊂ eA ⊂ A∗) in terms of linear relations θ in H. Moreover, it was shown in [9, 18] that each boundary triplet Π for A∗ gives rise to the Krein formula (2.36) with coefficients A0, γ(λ) and M(λ) naturally defined in terms of Π. ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 3 Assume that A is a closed symmetric linear relation in H with equal deficiency indices n+(A) = n−(A) ≤ ∞ and let Π = {H, Γ0, Γ1} be a boundary triplet for A∗. In the paper linear relation θc, which in a certain sense is a limit value of τ (iy) when y → ∞. This we consider extensions eAτ ∈ Self(A) corresponding to a certain subclass of Nevanlinna parameters τ (λ) in (2.36). We show that the compression C(eAτ ) of eAτ being a symmetric extension of A admits the representation C(eAτ ) = Aθc (see (1.5)) with a symmetric result enables us to characterise various properties of C(eAτ ) (self-adjointness, inclusion C(eAτ ) ⊂ A0, equalities C(eAτ ) = A and C(eAτ ) = A0) in terms of the parameter τ . In the case n±(A) < ∞ the obtained results become valid for any Nevanlinna parameter τ . This fact enables us to prove the following theorem which strengthens essentially statements (i) -- (iii) of Theorem 1.1. Theorem 1.2. Assume that A is a closed symmetric linear relation in H with finite deficiency indices n+(A) = n−(A) < ∞ and let all other assumptions of Theorem 1.1 be satisfied. Then: ficiency indices n±(A). 2. Preliminaries 2.1. Notations. The following notations will be used throughout the paper: H, H de- note separable Hilbert spaces; B(H1, H2) is the set of all bounded linear operators defined on H1 with values in H2; A ↾ L is a restriction of the operator A ∈ B(H1, H2) onto the linear manifold L ⊂ H1; PL is the orthoprojection in H onto the subspace L ⊂ H; C+ (C−) is the open upper (lower) half-plane of the complex plane. Recall that a linear manifold T in the Hilbert space H0 ⊕ H1 (H ⊕ H) is called a linear in H). The set of all closed linear relations from H0 to relation from H0 to H1 (resp. H1 (in H) will be denoted by eC(H0, H1) (resp. eC(H)). Clearly for each linear operator T : dom T → H1, dom T ⊂ H0, its grT = {{f, T f } : f ∈ dom T } is a linear relation from H0 to H1. This fact enables one to consider an operator as a linear relation. For a linear relation T from H0 to H1 we denote by dom T := {h0 ∈ H0 : ∃h1 ∈ H1 {h0, h1} ∈ T }, ker T := {h0 ∈ H0 : {h0, 0} ∈ T } ran T := {h1 ∈ H1 : ∃h0 ∈ H0 {h0, h1} ∈ T }, mul T := {h1 ∈ H1 : {0, h1} ∈ T } yIm (τ (iy)h, h) < ∞ for all h ∈ ker Bτ . y→∞ Theorem 1.3. Let the assumptions be the same as in Theorem 1.2. Then: (i′) C(eAτ ) ⊂ A0 if and only if (1.2) holds. (ii′) C(eAτ ) = A if and only Bτ = 0 and (1.2) holds. (iii′) C(eAτ ) = A0 if and only if Bτ > 0. We prove also the following theorem which characterizes extensions eA ∈ Self(A) with self-adjoint compression C(eA). (i) C(eAτ ) is self-adjoint if and only if lim (i) C(eAτ ) is self-adjoint and transversal with A0 (that is, C(eAτ ) ∩ A0 = A and C(eAτ )b+A0 = A∗) if and only if the corresponding parameter for C(eAτ ) in the Krein formula is the operator N ′ τ (iy)h, h ∈ H. Moreover, in the sense of (1.5) C(eAτ ) = ANτ with extensions eA ∈ Self(A) of a symmetric linear relation A with possibly infinite equal de- finite-codimensional extension eA ∈ Self(A). In the case n+(A) = n−(A) < ∞ this result holds for any given by N ′ τ h = lim y→∞ Nτ = −N ′ τ . lim y→∞ yIm (τ (iy)h, h) < ∞ for all h ∈ H. In this case Finally, we extend statement (iv) of Theorem 1.1 to a certain class of finite-codimensional τ in H 4 V.I. MOGILEVSKII the domain, kernel, range and multivalued part of T respectively; moreover, we let Denote also by T −1 and T ∗ the inverse and adjoint linear relations of T respectively. For an operator T = T ∗ ∈ B(H) we write T ≥ 0 if (T f, f ) ≥ 0, f ∈ H, and T > 0 if dmul T = {0} ⊕ mul T = {{0, h1} : h1 ∈ mul T }. T − αI ≥ 0 with some α > 0. For linear relations T1 and T2 in H we let T1b+T2 = {bf +bg : bf ∈ T1,bg ∈ T2}. 2.2. Nevanlinna functions. Recall that a holomorphic operator function M : C \ R → B(H) is called a Nevanlinna function if Im λ·Im M(λ) ≥ 0 and M ∗(λ) = M(λ), λ ∈ C\R. The class of all Nevanlinna B(H)-valued functions will be denoted by R[H]. Definition 2.1. The operator-function M ∈ R[H] is referred to the class: (i) Rc[H], if ran Im M(λ) is closed for all λ ∈ C \ R; (ii) Ru[H], if (Im M(λ))−1 ∈ B(H) for all λ ∈ C \ R or equivalently if Im λ · Im M(λ) > 0, λ ∈ C \ R. Clearly Ru[H] ⊂ Rc[H] ⊂ R[H]; moreover in the case dim H < ∞ one has Rc[H] = R[H]. The following proposition is well known (see e.g. [18]). Proposition 2.2. If M ∈ R[H], then the equality (2.1) BM = s- lim y→∞ 1 iy M(iy) defines the operator BM ∈ B(H) such that BM ≥ 0. Moreover, the equality (2.2) dom NM = {h ∈ H : lim y→∞ yIm (M(iy)h, h) < ∞} defines the (not necessarily closed) linear manifold dom NM ⊂ H and for each h ∈ dom NM there exists the limit (2.3) NM h := lim y→∞ M(iy)h, h ∈ dom NM . Hence the equalities (2.2) and (2.3) define the linear operator NM : dom NM → H. Proposition 2.3. Let τ ∈ R[H]. Then the subspace H′′ := ker Im τ (λ) ⊂ H does not depend on λ ∈ C \ R and the block representation holds with H′ = H ⊖ H′′, B2 = B∗ 2 ∈ B(H′′), B1 ∈ B(H′′, H′) and the operator-function τ1(·) ∈ R[H′] such that ker Im τ1(λ) = {0}, λ ∈ C \ R. Moreover, τ ∈ Rc[H] if and only if τ1 ∈ Ru[H′]. Proof. As is known the Cayley transform K(λ) = (τ (λ) − i)(τ (λ) + i)−1, λ ∈ C+, is a holomorphic contractive operator-function on C+ and for each λ ∈ C+ the operator τ (λ) admits the representation (2.5) f = i(K(λ) − I)h, τ (λ)f = (K(λ) + I)h, h ∈ H. It follows from (2.5) that (2.6) Im (τ (λ)f, f ) = h2 − K(λ)h2 for all λ ∈ C+, h ∈ H and f = i(K(λ) − I)h. According to [19] the subspace H0 = {h ∈ H : K(λ)h = h} and the operator V = K(λ) ↾ H0(∈ B(H0, H)) do not depend on λ ∈ C+. Moreover, by (2.5) and (2.6) ker Im τ (λ) = (V − I)H0 and hence the subspace (2.4) τ (λ) =(cid:18)τ1(λ) B1 1 B2(cid:19) : H′ ⊕ H′′ → H′ ⊕ H′′, B∗ λ ∈ C \ R ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 5 H′′ := ker Im τ (λ) does not depend on λ ∈ C+. Note also that by (2.5) the operator τ (λ) ↾ H′′ is defined by f = i(V − I)h0, τ (λ)f = (V + I)h0, h0 ∈ H0. Therefore B = τ (λ) ↾ H′′(∈ B(H′′, H)) does not depend on λ ∈ C+ and, consequently, τ (λ) admits the block representation (2.7) This implies that τ (λ) =(cid:18)τ1(λ) B1 τ2(λ) B2(cid:19) : H′ ⊕ H′′ → H′ ⊕ H′′, λ ∈ C+ Im τ (λ) =(cid:18) Im τ1(λ) 2i (τ2(λ) − B∗ 1) 1 1 2i (B1 − τ ∗ Im B2 2 (λ) (cid:19) : H′ ⊕ H′′ → H′ ⊕ H′′, λ ∈ C+ 1. Therefore by (2.7) the equality (2.4) holds with and hence Im B2 = 0, τ1(λ) = B∗ B2 = B∗ 2, which implies that (2.8) Im τ (λ) =(cid:18)Im τ1(λ) 0 0(cid:19) : H′ ⊕ H′′ → H′ ⊕ H′′, 0 λ ∈ C+. Hence ker Im τ1(λ) = {0} and the required statements are proved for λ ∈ C+. The same statements for λ ∈ C− are implied by τ (λ) = τ ∗(λ), λ ∈ C−. (cid:3) Nevanlinna relation valued functions if: As is known (see e.g. [8]) a function τ : C \ R → eC(H) is referred to the class eR(H) of (i) Im (f ′, f ) ≥ 0, (ii) (τ (λ) + i)−1 ∈ B(H), λ ∈ C+, and (τ (λ) + i)−1 is a holomorphic operator-function {f, f ′} ∈ τ (λ), λ ∈ C+; in C+; (iii) τ ∗(λ) = τ (λ), λ ∈ C \ R. (2.9) λ ∈ C \ R H = H0 ⊕ K, the operator part of τ . τ (λ) does not depend on λ ∈ C \ R and the decompositions τ (λ) = τ0(λ) ⊕ bK, mul τ (λ) = 0, λ ∈ C \ R. If τ (·) ∈ R(H), then dom τ (λ) = H, λ ∈ C \ R. A function τ ∈ eR(H) is referred to the class R(H) if its values are operators, i.e., if It is clear that R[H] ⊂ R(H) ⊂ eR(H). According to [15] for each function τ ∈ eR(H) the multivalued part K := mul τ (λ) of hold with τ0 ∈ R[H0] and bK = dmul τ (λ) = {0} ⊕ K. The operator function τ0 is called Definition 2.4. A relation valued function τ ∈ eR(H) is referred to the class eRc(H) (eRu(H)) if decompositions (2.9) hold with τ0 ∈ Rc[H0] (resp. with τ0 ∈ Ru[H0]). Clearly eRu(H) ⊂ eRc(H) ⊂ eR(H) and in the case dim H < ∞ one has eRc(H) = eR(H). Corollary 2.5. The relation valued function τ : C \ R → eC(H) belongs to the class eRc(H) if and only if there exist a decomposition 1 −B2(cid:19) : H′ ⊕ H′′ {z } operators B1 ∈ B(H′′, H′), B2 = B∗ such that τ (λ) admits the representation (2.9) with the operator function τ0(λ) =(cid:18)τ1(λ) −B1 2 ∈ B(H′′) and an operator-function τ1 ∈ Ru[H′] The following corollary is immediate from Proposition 2.3. H = H0 ⊕ K = H′ ⊕ H′′ → H′ ⊕ H′′ λ ∈ C \ R. H0 {z , } H0 {z } H0 (2.10) (2.11) −B∗ ⊕K, 6 V.I. MOGILEVSKII This means that (2.12) τ (λ) = {{h ⊕ ϕ, (τ1(λ)h − B1ϕ) ⊕ (−B∗ 1 h − B2ϕ) ⊕ k} : h ⊕ ϕ ∈ H′ ⊕ H′′, k ∈ K}. Remark 2.6. In the following for the relation valued function τ ∈ eRc(H) represented in accordance with Corollary 2.5 by (2.10) and (2.12) we will use the notation τ = {H′ ⊕ H′′ ⊕ K, B1, B2, τ1}. 2.3. Boundary triplets and Weyl function. Recall that a linear relation T in H is called: (i) symmetric if T ⊂ T ∗ or, equivalently, if (2.13) (f ′, g) − (f, g′) = 0, {f, f ′}, {g, g′} ∈ T ; (ii) self-adjoint if T = T ∗ (and hence T ∈ eC(H)). Then Assume that θ is a symmetric relation in H with a closed multivalued part mul θ. (2.14) H = H0 ⊕ K, where K = mul θ, H0 = H ⊖ mul θ and B is a symmetric operator in H0 (the operator part of θ) with dom B = dom θ ⊂ H0. Clearly, θ is closed (self-adjoint) if and only if B is closed (self-adjoint). θ = B ⊕ dmul θ = {{h, Bh + k} : h ∈ dom θ, k ∈ mul θ}, The following lemma will be useful in the sequel. Lemma 2.7. Let θ be a symmetric linear relation in H. Then: (i) dom θ ⊂ H ⊖ mul θ. (ii) If mul θ = mul θ and H ⊖ mul θ ⊂ dom θ, then θ∗ = θ. (iii) If dim H < ∞, then the condition H ⊖ mul θ ⊂ dom θ is equivalent to θ∗ = θ. Proof. (i) Statement (i) directly follows from (2.13) (ii) Assume that mul θ = mul θ and H ⊖ mul θ ⊂ dom θ. Then in the representation (2.14) of θ one has dom B = dom θ = H0. Therefore B = B∗ ∈ B(H0) and, consequently, θ∗ = θ. (iii) Assume that dim H < ∞. Then mul θ is closed and hence decompositions (2.14) hold. Moreover, dim H0 < ∞ and therefore θ∗ = θ if and only if dom B = H0 or, equivalently, dom θ = H ⊖ mul θ. This yields statement (ii). (cid:3) In the following we denote by A a closed symmetric linear relation (in particular closed not necessarily densely defined symmetric operator) in a Hilbert space H. Let Nλ(A) = let N(A) = H ⊖ dom A(= mul A∗) and let n±(A) := dim Nλ(A) ≤ ∞, λ ∈ C±, be deficiency indices of A. Denote by ext(A) the set of all proper extensions of A (i.e., ker (A∗ − λ) (λ ∈ C \ R) be a defect subspace of A, let bNλ(A) = {{f, λf } : f ∈ Nλ(A)}, the set of all relations eA in H such that A ⊂ eA ⊂ A∗) and by ext(A) the set of closed extensions eA ∈ ext(A). Recall that two extensions eA1, eA2 ∈ ext(A) are called transversal if eA1 ∩ eA2 = A and eA1b+eA2 = A∗. Definition 2.8. [14, 5] A collection Π = {H, Γ0, Γ1} consisting of a Hilbert space H and linear mappings Γj : A∗ → H, j ∈ {0, 1}, is called a boundary triplet for A∗, if the mapping Γ = (Γ0, Γ1)⊤ from A∗ into H ⊕ H is surjective and the following Green's identity holds: (2.15) (f ′, g) − (f, g′) = (Γ1bf , Γ0bg) − (Γ0bf , Γ1bg), [14, 18] If Π = {H, Γ0, Γ1} is a boundary triplet for A∗, then n+(A) = Proposition 2.9. n−(A) = dim H. Conversely, for each symmetric relation A in H with equal deficiency indices n+(A) = n−(A) there exists a boundary triplet for A∗. bf = {f, f ′}, bg = {g, g′} ∈ A∗. ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 7 Proposition 2.10. [14, 18] Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗. Then: (i) ker Γ = A and Γ is a bounded operator from A∗ onto H ⊕ H. (ii) The mapping (2.16) θ → Aθ := { f ∈ A∗ : {Γ0 f , Γ1 f } ∈ θ} establishes a bijective correspondence between all linear relations θ in H and all extensions (iii) (Aθ)∗ = Aθ∗ eA = Aθ ∈ ext(A). In the case θ ∈ B(H) one has Aθ = { f ∈ A∗ : Γ1 f = θΓ0 f } (iv) Aθ = Aθ and hence Aθ ∈ ext(A) if and only if θ ∈ eC(H) (v) Aθ is symmetric (self-adjoint) if and only if θ is symmetric (resp. self-adjoint) (vi) The equality (2.17) A0 := ker Γ0 = { f ∈ A∗ : Γ0 f = 0} defines a self-adjoint extension A0 of A. (vii) The extensions A0 and Aθ ∈ ext(A) are transversal if and only if θ ∈ B(H). The following lemma will be useful in the sequel. Lemma 2.11. Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗, (2.18) H = H′ ⊕ H′′ ⊕ K Γ11 Γ12 Γ13  : A∗ → H′ ⊕ H′′ ⊕ K. Moreover, let B ∈ B(H′′, H′ ⊕ H′′) be the operator with the block representation (2.19) (2.20) and the block representations of Γ0 and Γ1 are Γ01 Γ02 Γ03 Γ0 =  : A∗ → H′ ⊕ H′′ ⊕ K, B =(cid:18)B1 Γ1 = B2(cid:19) : H′′ → H′ ⊕ H′′ the domain dom B = H′′ )and let θ0 ∈ eC(H) be given by with bK = dmul θ0 = {0} ⊕ K. Then: such that B2 = B∗ (i)The equalities (2.21) θ0 = B ⊕ bK = {{h, B1h ⊕ B2h ⊕ k} : h ∈ H′′, k ∈ K} (2.22) S := Aθ0 = {bf ∈ A∗ : Γ01bf = 0, Γ03bf = 0, Γ11bf = B1Γ02bf , Γ12bf = B2Γ02bf } (2.23) define a symmetric extension S ∈ ext(A) and its adjoint S∗. (ii) The collection Π′ = {H′, Γ′ j : S∗ → H′ defined by 2 (this means that B is a bounded symmetric operator in H′ ⊕ H′′ with (2.24) is a boundary triplet for S∗. (iii) Let θ be a linear relation in H′ defined by Γ′ 0, Γ′ 1} with operators Γ′ Γ′ 1bf = Γ11bf − B1Γ02bf , S∗ = {bf ∈ A∗ : Γ03bf = 0, Γ12bf = B∗ 0bf = Γ01bf , 1Γ01bf + B2Γ02bf } bf ∈ S∗ θ = θ′ ∔ dmul θ = {{h, θ′h + h′} : h ∈ dom θ′, h′ ∈ mul θ}, eθ′ =(cid:18) 1 ↾ dom θ′ B2(cid:19) : dom θ′ ⊕ H′′ → H′ ⊕ H′′ B∗ B1 θ′ with a linear operator θ′ : dom θ′ → H′ (dom θ′ ⊂ H′) and let the extension eA ∈ ext(S) be given by eA = Sθ (in the triplet Π′). Moreover, let eθ′ : dom θ′ ⊕ H′′ → H′ ⊕ H′′ be a linear operator, defined by 8 V.I. MOGILEVSKII and let eθ be a linear relation in H given by (2.25) eθ :=eθ′ ∔ dmuleθ = {{h ⊕ ϕ, (θ′h + B1ϕ + h′) ⊕ (B∗ muleθ = mul θ ⊕ K (2.26) Then eA ∈ ext(A) and eA = Aeθ (in the triplet Π). If in addition θ ∈ B(H′), then h ⊕ ϕ ∈ dom θ′ ⊕ H′′, h′ ∈ mul θ, k ∈ K}. 1h + B2ϕ) ⊕ k} : and (2.26) takes the form eθ′ =(cid:18) θ B1 B∗ 1 B2(cid:19) : H′ ⊕ H′′ → H′ ⊕ H′′. (2.27) eθ =eθ′ ⊕ bK = {h ⊕ ϕ, (θ + B1ϕ) ⊕ (B∗ 1 h + B2ϕ) ⊕ k} : h ⊕ ϕ ∈ H′ ⊕ H′′, k ∈ K} Proof. (i) Clearly the relation θ0 is symmetric and by Proposition 2.10, (v) S := Aθ0 ∈ ext(A) is symmetric. Moreover, by (2.19) the second equality in (2.22) holds. Since obviously θ∗ 0 = {{h1 ⊕ h2, h′ ⊕ (B∗ 1 h1 + B2h2) ⊕ k} : h1 ⊕ h2 ∈ H′ ⊕ H′′, h′ ∈ H′, k ∈ K} and by Proposition 2.10, (iii) S∗ = Aθ∗ 0 , the equality (2.23) is valid. 0 and Γ′ 1. 1bf , Γ′ Next assume that h′ 1, h′ Π) and (2.23) one obtains 0bf , Γ′ 1bg). This proves Green's identity (2.15) for operators Γ′ 2 ∈ H′. Since the mapping Γ = (Γ0, Γ1)⊤ is surjective, there is (ii) Let bf = {f, f ′},bg = {g, g′} ∈ S∗. By using Green identity (2.15) (for the triplet (f ′, g) − (f, g′) = (Γ11bf , Γ01bg) − (Γ01bf , Γ11bg) + (Γ12bf , Γ02bg) − (Γ02bf , Γ12bg)+ (Γ13bf , Γ03bg) − (Γ03bf , Γ13bg) = (Γ11bf , Γ01bg) − (Γ01bf , Γ11bg) + (Γ01bf , B1Γ02bg)+ (B2Γ02bf , Γ02bg) − (B1Γ02bf , Γ01bg) − (Γ02bf , B2Γ02bg) = ((Γ11 − B1Γ02)bf , Γ01bg)− (Γ01bf , (Γ11 − B1Γ02)bg) = (Γ′ bf ∈ A∗ such that Γ01bf = h′ 1, Γ02bf = 0, Γ03bf = 0, Γ11bf = h′ Hence Γ12bf = B∗ 1Γ01bf + B2Γ02bf and by (2.23) bf ∈ S∗. Moreover, in view of (2.24) one 1bf = h′ 0bf = h′ (iii) It follows from (2.23) and (2.24) that bf ∈ eA if and only if bf ∈ A∗ and Γ01bf ∈ dom θ, Γ11bf − B1Γ02bf = θ′Γ01bf + h′, Γ12bf = B∗ 1Γ01bf + B2Γ02bf , Γ03bf = 0 with some h′ ∈ mul θ. In turn, these conditions are equivalent to inclusions bf ∈ A∗ and {Γ0bf , Γ1bf } ∈eθ. Hence eA = Aeθ (in the triplet Π). Lemma 2.12. Let L1 and L2 be closed subspaces in a Hilbert space H such that L1∩L2 = {0}. Then L1 ∔ L2 = L1 ∔ L2 if and only if there exists a closed subspace L′ 1 ⊃ L1 in H such that H = L′ 0bg) − (Γ′ 2, Γ12bf = B∗ 2, which proves surjectivity of the mapping Γ′ = (Γ′ 1, Γ13bf = 0. has Γ′ 1, Γ′ 0, Γ′ 1)⊤. 1h′ (cid:3) 1 ∔ L2. 1 ⊕ L2 with a closed subspace L′ Proof. Let L1 ∔ L2 = L1∔L2. Then H = (L1∔L2)⊕H′ with H′ = H⊖(L1∔L2) and hence H = L′ 1 ⊃ L1 be a closed subspace in H such that H = L′ 1) and π2 ∈ B(H, L2) be the corresponding skew projections onto L′ 1 and L2 respectively. Assume that fn ∈ L1 ∔ L2 and fn → f . Then π1fn ∈ L1, π1fn → π1f and hence π1f ∈ L1. Since π2f ∈ L2, it follows that f (= π1f + π2f ) ∈ L1 ∔ L2. Therefore L1 ∔ L2 is closed. (cid:3) 1 = L1 ⊕ H′. Conversely, let L′ 1 ∔ L2 and let π1 ∈ B(H, L′ ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 9 Proposition 2.13. Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗, let A0 = ker Γ0 (2.28) 2.10, (i) that and only if dom θ = dom θ). linear operator B : dom θ → H0, dom θ ⊂ H0 (in particular, this assumption is satisfied and let θ ∈ eC(H) be a linear relation such that decompositions (2.14) hold with a (closed) for symmetric θ). Then Aθb+A0 = Aθb+A0 if and only if B ∈ B(dom θ, H0) (that is, if Proof. Let bH = {0} ⊕ H and bH0 = {0} ⊕ H0. Since A0 = A bH, it follows from Proposition Moreover, B ∩ bH0 = {0} and bH = bH0 ⊕ bK, which in view of (2.14) implies that θb+bH = (B ∔ bH0) ⊕ bK. Therefore Next, by Lemma 2.12 B ∔ bH0 = B ∔ bH0 if and only if there exists B′ ∈ eC(H0) such that B ⊂ B′ and B′ ⊕ bH0 = H2 Aθb+A0 = Aθb+A0 ⇐⇒ θb+bH = θb+bH. θb+bH = θb+bH ⇐⇒ B ∔ bH0 = B ∔ bH0. 0. Since the last equality is equivalent to the inclusion B′ ∈ B(H0), the following equivalence holds: (2.29) (2.30) Now combining (2.28), (2.29) and (2.30), one obtains the required statement. Corollary 2.14. Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗, let A0 = ker Γ0 and (cid:3) B ∔ bH0 = B ∔ bH0 ⇐⇒ B ∈ B(dom θ, H0). let θ ∈ eC(H). Then Aθ ∩ A0 = A and Aθb+A0 = Aθb+A0 if and only if θ ∈ B(dom θ, H). Proof. Since A0 = A{0}⊕H, it follows from Proposition 2.10, (i) that Aθ ∩ A0 = A if and only if θ ∩ ({0} ⊕ H) = {0}, that is if and only if θ is an operator. Thus Aθ ∩ A0 = A if and only if decompositions (2.9) hold with K = {0}. This and Proposition 2.13 yield (cid:3) the result. In the rest of this subsection we recall some definitions and results from [9, 18, 10]. Proposition 2.15. Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗ and let A0 = A∗ 0 ∈ ext(A) be given by (2.17). Moreover, let π1 be the orthoprojection in H ⊕ H onto H ⊕ {0}. Then the operator Γ0 ↾ bNλ(A), λ ∈ C \ R, isomorphically maps bNλ(A) onto H and hence the equalities γ(λ) = π1(Γ0 ↾ bNλ(A))−1, Γ1 ↾ bNλ(A) = M(λ)Γ0 ↾ bNλ(A), correctly define the operator functions γ(·) : C\ R → B(H, H) and M(·) : C\ R → B(H). Moreover, γ(·) and M(·) satisfy the identities λ ∈ C \ R (2.31) (2.32) γ(λ) = γ(z) + (λ − z)(A0 − λ)−1γ(z) M(z) − M ∗(λ) = (z − λ)γ∗(λ)γ(z), z, λ ∈ C \ R which imply that γ(·) and M(·) are holomorphic in C \ R and M(·) ∈ R[H]. Definition 2.16. The operator-functions γ(·) and M(·) defined in Proposition 2.15 are called the γ-field and the Weyl function of the triplet Π respectively. Remark 2.17. It follows from (2.31) and (2.32) that γ(·) and M(·) are the γ-field and the Q-function of the pair (A, A0) respectively in the sense of [16, 17]. Definition 2.18. The (not necessarily closed) linear relation F in H defined by is called a forbidden relation of the boundary triplet Π = {H, Γ0, Γ1} for A∗. F := Γdmul A∗ = {{Γ0{0, n}, Γ1{0, n}} : n ∈ N(A)} 10 V.I. MOGILEVSKII In the following theorem the forbidden relation F is characterized in terms of the asymptotic behavior of the Weyl function. Theorem 2.19. Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗, let M(·) be the Weyl function of Π and let BM and NM be the operators defined by (2.1) and (2.2), (2.3) respectively. Moreover, let F be the forbidden relation of Π. Then (2.33) ran BM ⊂ mul F , ran BM = mul F , and F admits the representation (2.34) 2.4. Exit space extensions and formula for generalized resolvents. As is known F = NM ∔ dmul F = {{h, NM h + h′} : h ∈ dom NM , h′ ∈ mul F }. (2.35) λ ∈ C \ R a linear relation eA = eA∗ in a Hilbert space eH ⊃ H is called an exit space extension of A if A ⊂ eA and the minimality condition span{H, (eA − λ)−1H : λ ∈ C \ R} =eH is satisfied. For an exit space extension eA ∈ eC(eH) of A the compressed resolvent is called a generalized resolvent of A (here PH is the orthoprojection ineH onto H). If two exit space extensions eA1 ∈ eC(eH1) and eA2 ∈ eC(eH2) of A generates the same generalized resolvent R(λ), then eA1 and eA2 are equivalent. The latter means that there exists a unitary operator V ∈ B(eH1 ⊖ H,eH2 ⊖ H) such that eA2 = eUeA1 with the unitary operator 1,eH2 eU = (IH ⊕ V ) ⊕ (IH ⊕ V ) ∈ B(eH2 2). Hence each exit space extension eA of A is defined R(λ) = PH(eA − λ)−1 ↾ H, by the generalized resolvent (2.35) uniquely up to the equivalence. [9, 18] Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗, Theorem 2.20. A0 = ker Γ0 and γ(·) and M(·) are the γ-field and the Weyl function of Π respectively. Then: (i) the equality (Krein formula for generalized resolvents) (2.37) (2.36) λ ∈ C \ R λ ∈ C \ R; PH(eAτ − λ)−1 ↾ H = (A0 − λ)−1 − γ(λ)(τ (λ) + M(λ))−1γ∗(λ), PH(eAτ − λ)−1 ↾ H = (A−τ (λ) − λ)−1, establishes a bijective correspondence eA = eAτ between all relation valued functions τ = τ (λ) ∈ eR(H) and all exit space self-adjoint extensions eA of A. Moreover, for each τ ∈ eR(H) the following equality holds: (ii) an extension eAτ is canonical (that is, eAτ ∈ eC(H)) if and only if τ (λ) ≡ θ(= θ∗), λ ∈ C \ R. In this case eAτ = A−θ (in the sense of Proposition 2.10, (ii)). It follows from Theorem 2.20 that Krein formula (2.36) gives a parametrization eA = eAτ of all exit space extensions eA = eA∗ of A in terms of functions τ (·) ∈ eR(H). Without 3.1. Parametrization of compressions. Assume thateH ⊃ H is a Hilbert space, Hr := eH⊖H, PH is the orthoprojection ineH onto H and eA = eA∗ ∈ eC(eH) is an exit space extension connection with boundary triplets formula (2.36) was originally proved in [15, 17]. 3. Compressions of exit space self-adjoint extensions of A. In the following we let (3.1) (3.2) (3.3) S(eA) := eA ∩ H2 = {bf ∈ H2 : bf ∈ eA} T (eA) := {{PHf, PHf ′} : {f, f ′} ∈ eA} C(eA) := PHeA ↾ H = {{f, f ′} ∈ H2 : {f, f ′ ⊕ f ′ r} ∈ eA with some f ′ r ∈ Hr} ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 11 Proof. According to (2.37) λ ∈ C \ R, (3.4) (3.5) Π). (3.6) Π). The following theorem directly follows from the results of [6, 8]. −τ (λ) = {h ⊕ ϕ, (−τ1(λ)h + B1ϕ) ⊕ (B∗ 1 h + B2ϕ) ⊕ k} : h ⊕ ϕ ∈ H′ ⊕ H′′, k ∈ K}. Lemma 3.3. Let the assumptions of Lemma 2.11 be satisfied. Moreover, let S ∈ ext(A) be symmetric extension (2.22) of A, let Π′ = {H′, Γ′ 1} be boundary triplet (2.24) 0, Γ′ Clearly, S(eA) is a closed symmetric relation in H, C(eA) is a symmetric relation in H, T (eA) is a linear relation in H and A ⊂ S(eA) ⊂ C(eA) ⊂ T (eA) ⊂ A∗. Moreover, S(eA) ⊂ eA and according to [21, 7] (S(eA))∗ = T (eA). Definition 3.1. The linear relation C(eA) is called the compression of eA. Theorem 3.2. Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗, let τ ∈ eR(H) and let eAτ = eA∗ τ be the corresponding exit space extension of A. Then the equalities S(eA) = A and T (eA) = T (eA) hold if and only if τ ∈ Ru[H]. for S∗, let τ1(·) ∈ Ru[H′] and let eA = eSτ1 be the corresponding exit space self-adjoint extension of S. Assume also that τ = {H′ ⊕ H′′ ⊕ K, B1, B2, τ1} ∈ eRc(H) (see Remark 2.6). Then eA ⊃ A and eA = eAτ (in the triplet Π). PH(eA − λ)−1 ↾ H = (eA(λ) − λ)−1, where eA(λ) = S−τ1(λ) (in the triplet Π′). Moreover, Comparing this equality with (2.27) we obtain from Lemma 2.11, (iii) that eA(λ) ∈ ext(A) and eA(λ) = A−τ (λ) (in the triplet Π). Therefore by (3.5) and (2.37) eA = eAτ (in the triplet Proposition 3.4. Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗, let τ ∈ eRc(H) and let eAτ = eA∗ (see Remark 2.6), Γ0 and Γ1 have the block representations (2.19) and θ0 ∈ eC(H) is given by (2.21). Then the extension S = S(eAτ ) admits the representation (2.22) (in the triplet 1} be boundary triplet (2.24) for S∗. Assume that eA := eSτ1 (in the triplet Π′). Since τ1 ∈ Ru[H′], it follows from Theorem 3.2 that S(eA) = S. Moreover, according to Lemma 3.3 A ⊂ eA and eA = eAτ (in the triplet Π), which proves the required let τ ∈ eRc(H). Then S(eAτ ) ∩ A0 = A if and only if τ ∈ Rc[H]. S(eAτ ) = Aθ0 with θ0 given by (2.21). Since obviously Aθ0 ∩ A0 = A bK, the equivalences S(eAτ ) ∩ A0 = A ⇔ K = {0} ⇔ τ ∈ Rc[H] are valid. eAτ corresponding to τ ∈ eRc(H). Proof. Clearly the assumptions of Lemma 2.11 are satisfied with the subspaces H′, H′′, K and the operator B = (B1, B2)⊤. Let S ∈ ext(A) be symmetric extension (2.22) and let Π′ = {H′, Γ′ statement. Corollary 3.5. Let Π = {H, Γ0, Γ1} be a boundary triplet for A∗, let A0 = ker Γ0 and Proof. Assume that τ is represented as in (3.6). Then according to Proposition 3.4 In the following theorem we give a geometric characterization of self-adjoint extensions τ be the corresponding exit space extension of A. Assume also that (cid:3) (cid:3) (cid:3) τ = {H′ ⊕ H′′ ⊕ K, B1, B2, τ1} 0, Γ′ 12 V.I. MOGILEVSKII Theorem 3.6. Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗ and A0 = 0, Γ′ A. Then: H = H0 ⊕ K, decompositions only if τ ∈ Rc[H]. T = T . Since S ∈ ext(A) and S ⊂ S∗, it follows from Proposition 2.10, (ii) and (v) that ker Γ0. Let τ ∈ eR(H) and let eAτ be the corresponding exit space self-adjoint extension of (i) The linear relations S(eAτ )b+A0 and T (eAτ ) are closed if and only if τ ∈ eRc(H). (ii) S(eAτ ) ∩ A0 = A and the linear relations S(eAτ )b+A0 and T (eAτ ) are closed if and Proof. (i) We put S = S(eAτ ) and T = T (eAτ ). Assume that Sb+A0 = Sb+A0 and S = Aθ0 with some symmetric relation θ0 ∈ eC(H). Moreover, by Proposition 2.13 the holds with K = mul θ0, bK = dmul θ0 and B ∈ B(dom θ0, H0), where dom θ0 ⊂ H0 is closed. Let H′′ = dom θ0 and H′ = H0 ⊖ H′′, so that H0 = H′ ⊕ H′′. Then decomposition (2.18) of H holds, the operator B admits the block representation (2.20) and θ0 can be written in the form (2.21). Moreover, since B is symmetric, the equality B2 = B∗ 2 holds. Therefore by Lemma 2.11, (ii) the equalities (2.24) define the boundary triplet Π′ = {H′, Γ′ θ0 = B ⊕ bK Π′ = {H′, Γ′ the triplet Π′) and in view of Theorem 3.2 one has T = T . of H, an operator B = (B1, B2)⊤ ∈ B(H′′, H′ ⊕ H′′) and a function τ1 ∈ Ru[H′] such that τ = {H′ ⊕ H′′ ⊕ K, B1, B2, τ1}. It follows from Proposition 3.4 that S = Aθ0 with 1} for S∗. Since eA = eAτ is an extension of S, it follows from Theorem 3.2 that eA = eAτ1 (in the triplet Π′) with some τ1 ∈ Ru[H′]. Therefore by Lemma 3.3 τ ∈ eRc(H). Conversely, let τ ∈ eRc(H). Then by Corollary 2.5 there exist a decomposition (2.10) θ0 given by (2.21). Therefore by Proposition 2.13 Sb+A0 = Sb+A0. Next assume that 1} is a boundary triplet (2.24) for S∗. Then by Lemma 3.3 eAτ = eSτ1 (in τ is the corresponding exit space extension of A. Let eAτ ∈ eC(eH) with a and eAτ = eA∗ Hilbert space eH ⊃ H and let Hr =eH ⊖ H. Then: (i) There exist a symmetric relation Ar ∈ eC(Hr) and a boundary triplet Πr = {H, Γr r : Γ0bf = Γr (ii) C(eAτ ) = A−Fr (in the triplet Π), were Fr is the forbidden relation of Πr. Proposition 3.7. Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗, τ ∈ Ru[H] eAτ = {bf ⊕ bfr ∈ A∗ ⊕ A∗ (ii) Combining statement (i) with Corollary 3.5 we arrive at statement (ii). 1bfr}. 0bfr, Γ1bf = −Γr Proof. Statement (i) directly follows from the results of [6, 8]. r such that τ is the Weyl function of Πr and (i) According to Proposition 2.10, (ii) statement (ii) is equivalent to the equality for A∗ 0, Γ′ (cid:3) 0, Γr 1} (3.7) (3.8) (3.9) and hence f ′ C(eAτ ) = {bf ∈ A∗ : {Γ0bf , −Γ1bf} ∈ Fr}. r. Moreover, by (3.8) {Γ0bf , −Γ1bf} = {Γr Let bf = {f, f ′} ∈ C(eAτ ). Then bf ∈ A∗ and by (3.3) there exists f ′ r} ∈ eAτ . Letting bfr = {0, f ′ r one getsbg = bf ⊕bfr. Therefore bfr ∈ A∗ bg := {f, f ′ ⊕ f ′ 1bfr} ∈ Fr. 0bfr, Γr Conversely, let bf = {f, f ′} ∈ A∗ and {Γ0bf , −Γ1bf } ∈ Fr. Then there exists f ′ r} = −Γ1bf . Letting bfr = {0, f ′ r} = Γ0bf and Γr from (3.8) thatbg := bf ⊕bfr = {f, f ′ ⊕ f ′ r} ∈ eAτ and therefore bf = {f, f ′} ∈ C(eAτ ). This r ∈ mul A∗ r r we obtain r ∈ Hr such that such that Γr r ∈ mul A∗ proves (3.9). r} ∈ A∗ r} ∈ H2 1{0, f ′ 0{0, f ′ (cid:3) r ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 13 In the following theorem we parameterize in terms of τ the compressions C(eAτ ) of exit space extensions eAτ with τ ∈ eRc(H). Theorem 3.8. Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗, τ ∈ eRc(H), eAτ = eA∗ τ is the corresponding exit space extension of A and C(eAτ ) is the compression of eAτ . Moreover, let τ0 ∈ Rc[H0] and K be the operator and multivalued parts of τ be operators corresponding to τ0 in accordance with Proposition 2.2. Then C(eAτ ) = Aθc respectively (see (2.9)) and let Bτ0 ∈ B(H0) and Nτ0 : dom Nτ0 → H0 (dom Nτ0 ⊂ H0) (in the triplet Π) with the symmetric linear relation θc in H given by θc = −Nτ0 ∔ dmul θc = {{h, −Nτ0h + h′} : h ∈ dom Nτ0, h′ ∈ mul θc}. (3.10) ran Bτ0 ⊕ K ⊂ mul θc, mul θc = ran Bτ0 ⊕ K. Moreover, (3.11) If in addition ran Bτ0 is closed, then mul θc = mul θc = ran Bτ0 ⊕ K and hence θc = {{h, −Nτ0h + Bτ0ψ + k} : h ∈ dom Nτ0, ψ ∈ H0, k ∈ K}. (3.12) Proof. Assume that τ = {H′⊕H′′⊕K, B1, B2, τ1} with H′⊕H′′ = H0 and τ1 ∈ Ru[H′] (see Corollary 2.5 and Remark 2.6). Moreover, let Γ0 and Γ1 have the block representations admits the representation (2.22) and by Lemma 2.11, (ii) the equalities (2.24) define a boundary triplet Π′ = {H′, Γ′ (in the triplet Π′) and by Proposition 3.7 there exist a Hilbert space Hr, a symmetric r such that τ1 is the (2.19) and let θ0 ∈ eC(H) be given by (2.21). Then according to Proposition 3.4 S = S(eAτ ) 1} for S∗. Moreover, according to Lemma 3.3 eAτ = eSτ1 relation Sr ∈ eC(Hr) and a boundary triplet Πr = {H′, Γr Weyl function of Πr and C(eAτ ) = S−Fr (in the triplet Π′), where Fr is the forbidden relation of Πr. It follows from Theorem 2.19 that 1} for S∗ 0, Γr 0, Γ′ ran Bτ1 ⊂ mul Fr, mul Fr = ran Bτ1 (3.13) (3.14) where Nτ1 : dom Nτ1 → H′ is the operator given by (3.15) −Fr = −Nτ1 ∔ dmul Fr = {{h, −Nτ1h + h′} : h ∈ dom Nτ1, h′ ∈ mul Fr}, yIm (τ1(iy)h, h) < ∞} dom Nτ1 = {h ∈ H′ : lim y→∞ τ1(iy)h, Nτ1h = lim y→∞ c : dom θ′ h ∈ dom Nτ1. Let dom θ′ c := dom Nτ1 ⊕ H′′ and θ′ c → H0 be the operator given by θ′ c =(cid:18) −Nτ1 1 ↾ dom Nτ1 B2(cid:19) : dom Nτ1 ⊕ H′′ → H′ ⊕ H′′. B1 B∗ Then by Lemma 2.11 C(eAτ ) = Aθc (in the triplet Π) with the linear relation θc in H given by mul θc = mul Fr ⊕ K (3.16) (3.17) (3.18) (3.19) It follows from (2.8) (with τ0(λ) in place of τ (λ)) that θc = θ′ c ∔ dmul θc = {{h, θ′ ch + h′} : h ∈ dom θ′ c, h′ ∈ mul θc} yIm (τ0(iy)h, h) = yIm (τ1(iy)PH′h, PH′h), h ∈ H0. Therefore by (3.15) dom θ′ for each h = h′ ⊕ h′′ ∈ dom θ′ c one obtains c = dom Nτ0. Moreover, combining (3.17), (3.16) and (2.11) θ′ ch = (−Nτ1h′ + B1h′′) ⊕ (B∗ 1 h′ + B2h′′) = − lim y→∞ ((τ1(iy)h′ − B1h′′) ⊕ (−B∗ 1 h′ − B2h′′)) = − lim y→∞ τ0(iy)h = −Nτ0h. 14 V.I. MOGILEVSKII This and (3.19) yield (3.10). Next, by (2.11) Bτ1 = Bτ0 PH′ and hence ran Bτ1 = ran Bτ0. Therefore by (3.13) and (3.18) the relations (3.11) are valid. Finally, θc is symmetric in view Proposition 2.10, (cid:3) (v). Theorem 3.9. Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗, A0 = ker Γ0 (that is A0 is a fixed canonical self-adjoint extension of A in the Krein formula (2.36)), statements are equivalent: τ ∈ eR(H) and τ0 ∈ R[H0] is the operator part of τ (see (2.14)). Then the following (i) T (eAτ ) is closed and C(eAτ ) ⊂ A0; (ii) τ ∈ eRc(H) and τ0 satisfies yIm (τ0(iy)h, h) = ∞, h ∈ H0, h 6= 0. lim y→∞ (3.20) If statement (i) (or equivalently (ii)) is valid then (3.21) (3.22) C(eAτ ) = {bf ∈ A∗ : Γ0bf = 0, Γ1bf ∈ ran Bτ0 ⊕ K}. If in addition ran Bτ0 is closed, then C(eAτ ) is closed and C(eAτ ) = {bf ∈ A∗ : Γ0bf = 0, Γ1bf = Bτ0 h ⊕ k with some h ∈ H0 and k ∈ K}. Proof. Assume statement (i). Then by (3.4) S(eAτ ) ⊂ A0 and hence S(eAτ )b+A0 = A0. Therefore S(eAτ )b+A0 is closed and by Theorem 3.6, (i) τ ∈ eRc(H). This in view of Theorem 3.8 implies that C(eAτ ) = Aθc with θc given by (3.10). It follows from (3.10) that C(eAτ ) ⊂ A0 if and only if dom Nτ0 = {0}, which yields (3.20). Thus (ii) holds. Conversely, let (ii) holds. Then by Theorem 3.6, (i) T (eAτ ) is closed. Moreover, by Theorem 3.8 C(eAτ ) = Aθc with θc of the form (3.10) and (3.20) shows that dom Nτ0 = {0}. Hence θc = {0} ⊕ mul θc and therefore C(eAτ ) ⊂ A0, which yields statement (i). Moreover, by Proposition 2.10, (iv) C(eAτ ) = Aθc, where in view of (3.11) θc = {0} ⊕ mul θc = {0} ⊕ Hc with Hc = ran Bτ0 ⊕ K. This yields (3.21). Finally, the last statement (cid:3) of the theorem follows from the previous one. Corollary 3.10. Let the assumptions be the same as in Theorem 3.9. Then the following statements are equivalent: H0 ⊕ K. Hence ran Bτ0 = H0 or, equivalently, ker Bτ0 = {0}. Thus (ii) is valid. If in addition ran Bτ0 is closed, then statement (ii) is equivalent to the following one: (i) T (eAτ ) is closed and C(eAτ ) = A0; (ii) τ ∈ eRc(H) and ker Bτ0 = {0}. (i ′) T (eAτ ) is closed and C(eAτ ) = A0. Proof. Assume that (i) holds. Then by Theorem 3.9 τ ∈ eRc(H) and C(eAτ ) = A{0}⊕Hc with Hc = ran Bτ0 ⊕ K. Since A0 = A{0}⊕H, it follows from C(eAτ ) = A0 that Hc = H = Conversely, let (ii) holds. Then by Theorem 3.6, (i) T (eAτ ) is closed. Moreover, by Theorem 3.8 C(eAτ ) = Aθc with θc given by (3.10) and satisfying (3.11). Since ran Bτ0 = in view of Proposition 2.10, (iv) C(eAτ ) = A{0}⊕H = A0. This proves statement (i). versely, let (ii) holds. Then (i) is valid and hence C(eAτ ) = Aθc with θc = {0} ⊕ mul θc. C(eAτ ) = A{0}⊕H = A0, which yields (i ′). Moreover, ran Bτ0 = H0 and by Theorem 3.8 mul θc = H0 ⊕ K = H. Therefore (cid:3) Assume now that ran Bτ0 is closed. Then (i ′) implies (i) and, consequently, (ii). Con- H0, it follows from (3.11) that mul θc = H. Hence by Lemma 2.7, (i) θc = {0} ⊕ H and ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 15 Corollary 3.11. Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗ and τ ∈ (3.23) lim y→∞ h ∈ H, h 6= 0. yIm (τ (iy)h, h) = ∞, operator-function τ0 ∈ Rc[H0] in decomposition (2.9) of τ satisfies (3.20) and ran Bτ0 ⊕ K = {0}. Hence K = {0} and, consequently, H0 = H and τ = τ0. This implies that τ ∈ Rc[H] and (3.23) holds. Moreover, ran Bτ0 = {0} and therefore Bτ = Bτ0 = 0. eR(H). Then T (eAτ ) is closed and C(eAτ ) = A if and only if τ ∈ Rc[H], Bτ = 0 and Proof. Let T (eAτ ) be closed and C(eAτ ) = A. Then by Theorem 3.9 τ ∈ eRc(H), the Conversely, let τ ∈ Rc[H], Bτ = 0 and (3.23) is satisfied. Then by Theorem 3.9 T (eAτ ) is closed and C(eAτ ) is given by (3.21) with ran Bτ0 = ran Bτ = {0} and K = {0}. Hence C(eAτ ) = C(eAτ ) = ker Γ and by Proposition 2.10, (i) C(eAτ ) = A. each exit space extension eA = eA∗ of A is a densely defined operator and according to M.A. Naimark (see e.g. [1]) an extension eA of A is said to be of the second kind if dom eA∩ H = dom A or equivalently if C(eA) = A. Clearly, in the case dom A = H Corollary 3.11 gives a parametrisation of all extensions eA of the second kind satisfying T (eA) = T (eA). Note that this result follows from the results of [8], where a parametrisation of all extensions eA ⊃ A of the second kind of a densely defined A was obtained in terms of the parameter τ from the Krein formula (2.36). Observe also that a somewhat other parametrization of the second kind extensions can be found in [22, 25]. Remark 3.12. Assume that A is a closed densely defined symmetric operator in H. Then (cid:3) 3.2. Exit space self-adjoint extensions with a self-adjoint compression. In the following proposition we provide a sufficient condition on the parameter τ for self- adjointness of the compression C(eAτ ). Proposition 3.13. Assume that Π = {H, Γ0, Γ1} is a boundary triplet for A∗, τ ∈ eRc(H) and τ0 ∈ Rc[H0] is the operator part of τ (see (2.9)). If ran Bτ0 is closed and (3.24) lim y→∞ yIm (τ0(iy)h, h) < ∞, h ∈ ker Bτ0, Proof. Assume that ran Bτ0 is closed and (3.24) holds. Then according to Theorem 3.8 then C(eAτ ) is self-adjoint. C(eAτ ) = Aθc with a symmetric linear relation θc in H given by (3.10) and satisfying mul θc = mul θc = ran Bτ0 ⊕ K. Moreover, by (3.24) ker Bτ0 ⊂ dom Nτ0 = dom θc. Therefore H ⊖ mul θc = H0 ⊖ ran Bτ0 = ker Bτ0 ⊂ dom θc and by Lemma 2.7, (ii) θc = θ∗ 2.10, (v). c . Now the required statement follows from Proposition (cid:3) Corollary 3.14. Let the assumptions be the same as in Proposition 3.13. If ran Bτ0 is closed and there is a compact interval [a, b] ⊂ R such that τ0 admits a holomorphic continuation onto R \ [a, b], then C(eAτ ) is self-adjoint. Proof. If the function τ0 is holomorphic in C\[a, b], then it admits the Laurent expansion τ0(λ) = C0 + λBτ0 + 1 λk ∞Xk=1 Ck, λ > R > 0 16 V.I. MOGILEVSKII with operator-valued coefficients Ck = C ∗ k ∈ B(H0), k = 0, 1, 2, . . . . Therefore Im (τ0(iy)h, h) = and,consequently, (−1)m y2m−1 (C2m−1h, h), ∞Xm=1 y ∈ R, h ∈ ker Bτ0 lim y→∞ yIm (τ0(iy)h, h) = −(C1h, h) < ∞, h ∈ ker Bτ0. Thus (3.24) holds and the required statement follows from Proposition 3.13. (cid:3) Theorem 3.15. Assume the conditions of Theorem 3.9. Then the following statements are equivalent: In the following theorem we describe exit space self-adjoint extensions eA of A such that the compression C(eA) of eA is self-adjoint and transversal with A0. (i) T (eAτ ) is closed, C(eAτ )∗ = C(eAτ ) and the extensions C(eAτ ) and A0 are transversal; (ii) τ ∈ Rc[H] and yIm (τ (iy)h, h) < ∞, h ∈ H. (3.25) lim y→∞ Moreover, if (i) (equivalently (ii)) is satisfied, then (3.26) where Nτ ∈ B(H) is the operator given by C(eAτ ) = A−Nτ = {bf ∈ A∗ : Γ1bf = −Nτ Γ0bf}, (3.27) Nτ = s- lim y→∞ τ (iy). Proof. Let statement (i) holds. Since S(eAτ ) ⊂ C(eAτ ), it follows from Lemma 2.12 that the linear relation S(eAτ )b+A0 is closed. Moreover, by (3.4) S(eAτ ) ∩ A0 = A. Therefore by Theorem 3.6, (ii) τ ∈ Rc[H]. Next, according to Theorem 3.8 C(eAτ ) = Aθc, where θc is given by (3.10) with τ0 = τ , and by Proposition 2.10, (vii) θc ∈ B(H). Therefore by (3.10) dom Nτ = dom θc = H, which in view of (2.2) yields (3.25). Thus statement (ii) is valid. Moreover, since mul θc = {0}, it follows from (3.10) that θc = −Nτ and in view of definition (2.3) of Nτ the equality (3.27) holds. This yields (3.26). transversal. c ∈ B(H). Therefore by dom Nτ = H. Since θc is symmetric, this implies that θc = θ∗ Conversely, assume (ii). Then by Theorem 3.6 T (eAτ ) is closed. Moreover, by Theorem 3.8 C(eAτ ) = Aθc, where θc is given by (3.10) with τ0 = τ and (3.25) yields dom θc = Proposition 2.10, (v) and (vii) C(eAτ )∗ = C(eAτ ) and the extensions C(eAτ ) and A0 are 3.3. Finite-codimensional exit space extensions. LeteH ⊃ H be a Hilbert space, let Hr = eH ⊖ H and let eA = eA∗ ∈ eC(eH) be an exit space extension of A. Such an extension Proposition 3.16. Let A ∈ eC(H) be a symmetric linear relation with finite deficiency (hence dim H = d < ∞). Moreover, let τ ∈ eR(H) and let τ0 ∈ R[H0] be the operator part of τ (see (2.9)). Then eAτ is finite-codimensional if and only if indices n+(A) = n−(A) =: d < ∞ and let Π = {H, Γ0, Γ1} be a boundary triplet for A∗ The following proposition is well known (see e.g. [12]). is called finite-codimensional, if nr := dim Hr < ∞. (cid:3) (3.28) τ0(λ) = A + λB + Aj, λ ∈ C \ R 1 αj − λ lXj=1 ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 17 where (3.29) αj ∈ R, A, B, Aj ∈ B(H0), A = A∗, B ≥ 0, Aj ≥ 0, Aj 6= 0, j = 1, 2, . . . , l. Moreover, (3.30) nr = dim ran B + dim ran Aj. lXj=1 In the following two theorems we extend Proposition 3.16 to a ceratin class of self- ∞, j = 1, 2, . . . , l. satisfies (3.30). given by (ii) τ0 is the rational function (3.28), (3.29) such that dim ran B < ∞ and dim ran Aj < indices n+(A) = n−(A) ≤ ∞, let Π = {H, Γ0, Γ1} be a boundary triplet for A∗, let space extension of A (see Theorem 2.20). Assume also that τ0 ∈ R[H0] is the operator part of τ (see (2.9)). Then the following statements are equivalent: adjoint extensions eA of a symmetric relation A with possibly infinite deficiency indices n+(A) = n−(A) ≤ ∞ and characterise compressions C(eA) of such extensions. Theorem 3.17. Let A ∈ eC(H) be a symmetric linear relation in H with equal deficiency τ ∈ eC(eH) (eH ⊃ H) be the corresponding exit A0 = ker Γ0, let τ ∈ eR(H) and let eAτ = eA∗ (i) eAτ is finite-codimensional and the linear relation S(eAτ )b+A0 is closed; Moreover, if statement (ii) (and hence (i)) is valid, then the dimension nr of Hr =eH⊖H Proof. Assume statement (i) and let Sr = Sr(eAτ ) and Tr(eAτ ) be linear relations in Hr Since dim Hr < ∞, it follows that Tr(eAτ ) is closed and according to [7, Proposition 2.15] T (eAτ ) is closed as well. Therefore by Theorem 3.6 τ ∈ eRc(H). Let in accordance S = S(eAτ ) admits the representation (2.22). Let Π′ = {H′, Γ′ (2.24) for S∗. Then by Lemma 3.3 eAτ = eSτ1 (in the triplet Π′). Moreover, according to with Corollary 2.5 τ0 be decomposed as in (2.11) with τ1 ∈ Ru[H′], that is τ = {H′ ⊕ H′′ ⊕ K, B1, B2, τ1} in the sense of Remark 2.6. Then by Proposition 3.4 the relation 1} be boundary triplet [7, Lemma 2.14] n±(S) = n±(Sr) and, consequently, n+(S) = n−(S) < ∞. Therefore dim H′ < ∞ and by Proposition 3.16 r, Sr = eAτ ∩ H2 Tr(eAτ ) = {{PHrf, PHrf ′} : {f, f ′} ∈ eAτ }. 0, Γ′ (3.31) τ1(λ) = A0 + λB0 + 1 αj − λ lXj=1 A0j, λ ∈ C \ R, where αj ∈ R and A0, B0, A0j ∈ B(H′) are the operators satisfying A0 = A∗ 0, B0 ≥ 0 and A0j ≥ 0, A0j 6= 0, j = 1, 2, . . . , l.This and the block representation (2.11) of τ0(λ) imply that τ0(λ) is of the form (3.28) with (3.32) A =(cid:18) A0 −B1 1 −B2(cid:19) , −B∗ B =(cid:18)B0 0 0(cid:19) , 0 Hence A = A∗, B ≥ 0, Aj ≥ 0, Aj 6= 0 and Aj =(cid:18)A0j 0 0(cid:19) , j = 1, 2, . . . , l. 0 (3.33) dim ran B = dim ran B0 < ∞, dim ran Aj = dim ran A0j < ∞, which yields statement (ii). 18 V.I. MOGILEVSKII Conversely, let statement (ii) holds. Put ϕ0(λ) = Im λ and ϕj(λ) = Im 1 αj−λ. Then by (3.28) (3.34) Im τ0(λ) = ϕ0(λ)B + lXj=1 ϕj(λ)Aj, λ ∈ C \ R. Since Im λ · Im ϕj(λ) > 0, j = 0, 1, . . . , l, it follows from (3.34) that (3.35) H′′ := ker Im τ0(λ) = ker B ∩ l\j=1 ker Aj! and, consequently, H′ := H0 ⊖ H′′ = ran B + ran A1 + · · ·+ ran Al. Therefore dim H′ < ∞ and in view of (3.35) block representations (3.32) hold (with respect to the decomposition H0 = H′ ⊕H′′). Combining (3.32) with (3.28) one obtains the block representation (2.11) of τ0(λ) with τ1 ∈ R[H′] given by (3.31). Since dim H′ < ∞, it follows that τ1 ∈ Ru[H′] and according to Proposition 2.3 τ0 ∈ Rc[H0]. Therefore τ ∈ eRc(H) and by Theorem 3.6 S(eAτ )b+A0 is closed. Now it remains to show that eAτ is finite-codimensional and (3.30) holds. Since τ = {H′ ⊕ H′′ ⊕ K, B1, B2, τ1}, it follows from Proposition 3.4 that S = S(eAτ ) admits the by Lemma 3.3 eAτ = eSτ1 (in the triplet Π′). Thus in accordance with Proposition 3.16 eAτ is finite-codimensional and representation (2.22) and by Lemma 2.11 the equalities (2.24) define a boundary triplet Π′ = {H′, Γ′ 1} for S∗. Therefore by Proposition 2.9 n±(S) = dim H′ < ∞. Moreover, 0, Γ′ nr = dim ran B0 + dim ran A0j. lXj=1 Combining this equality with (3.33) we arrive at (3.30). (cid:3) Theorem 3.18. Let the assumptions of Theorem 3.17 be satisfied and let claim (ii) (equivalently (i)) of this theorem be valid. Moreover, let A′ = Pker BA ↾ ker B ∈ B(ker B), let K be the multivalued part of τ (see (2.9)) and let θc be a self-adjoint linear relation in H given by (3.36) θc = {{h, −A′h ⊕ h′ ⊕ k} : h ∈ ker B, h′ ∈ ran B, k ∈ K} (this means that −A′ is the operator part of θc and mul θc = ran B ⊕ K). Then the Moreover, the following holds: extension eAτ is finite-codimensional, the compression C(eAτ ) of eAτ is self-adjoint and C(eAτ ) = Aθc (in the triplet Π). (i) C(eAτ ) = A0 if and only if ker B = 0 (that is B > 0); (ii) C(eAτ ) and A0 are transversal if and only if τ ∈ R[H] and B = 0, in which case Proof. It follows from Theorem 3.17 that eAτ is finite-codimensional. Moreover, by The- orem 3.8 C(eAτ ) = θc with symmetric relation θc in H given by (3.10). Since in view of θc = −A. (3.28) yIm (τ0(iy)h, h) = y2(Bh, h) + lXj=1 y2 j + y2 (Ajh, h), α2 h ∈ H0, ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 19 it follows that lim y→∞ yIm (τ0(iy)h, h) < ∞ if and only if h ∈ ker B. Therefore by definitions (2.2) and (2.3) dom Nτ0 = ker B and Nτ0h = lim y→∞ τ0(iy)h = Ah, h ∈ ker B(= dom Nτ0) Moreover, by (2.1) Bτ0 = B and the finite-dimensionality of ran B yields ran Bτ0 = ran Bτ0 = ran B. Therefore by Theorem 3.8 mul θc = ran B ⊕ K and (3.10) can be written as (3.37) θc = {{h, −Ah + (h′ ⊕ k)} : h ∈ ker B, h′ ∈ ran B, k ∈ K} Since H0 = ker B ⊕ ran B and h′ ∈ ran B in (3.37) is arbitrary, the equality (3.37) is equivalent to (3.36). Finally, statements (i) and (ii) are implied by Corollary 3.10 and (cid:3) Theorem 3.15. 3.4. The case of finite deficiency indices. In this subsection we suppose that the following assumption (a) is satisfied: (a) A is a closed symmetric linear relation in H with finite deficiency indices n+(A) = adjoint. (3.40) (3.38) lim y→∞ yIm (τ0(iy)h, h) = ∞, h ∈ H0, h 6= 0. τ is the corresponding exit space extension of A Π) with the symmetric linear relation θc in H defined by (3.12). Theorem 3.20. Let the assumption (a) be satisfied and let A0 = ker Γ0. Then: and ran Bτ = ran Bτ for any subspace H′ ⊂ H and operator-function τ ∈ R[H′]. These facts together with Theorems 3.8, 3.9 and Corollaries 3.10, 3.11 yield the following two theorems. n−(A) < ∞, Π = {H, Γ0, Γ1} is a boundary triplet for A∗, τ ∈ eR(H), τ0 ∈ R[H0] is the operator part of τ (see (2.9)), eAτ = eA∗ and C(eAτ ) is the compression of eAτ . Then ext(A) = ext(A) and hence for each exit space extension eA = eA∗ of A the linear relations C(eA), T (eA) and S(eA)b+A0 are closed (here A0 = ker Γ0). Moreover, under the assumption (a) dim H < ∞, which implies that eR(H) = eRc(H), R[H′] = Rc[H′] Theorem 3.19. Let the assumption (a) be satisfied. Then C(eAτ ) = Aθc (in the triplet (i) C(eAτ ) ⊂ A0 if and only if the operator function τ0 satisfies In this case C(eAτ ) is defined by abstract boundary conditions (3.22). (ii) C(eAτ ) = A0 if and only if ker Bτ0 = 0 (that is Bτ0 > 0 ). (iii) C(eAτ ) = A if and only if τ ∈ R[H], Bτ = 0 and In the following theorem we describe all exit space extensions eA = eA∗ of a symmetric relation A with finite deficiency indices such that the compression C(eA) of eA is self- (i) C(eAτ ) is self-adjoint if and only if If in particular τ0 is holomorphic out of a compact interval [a, b] ⊂ R, then C(eAτ ) is Theorem 3.21. Let the assumption (a) be satisfied and let A0 = ker Γ0. Then: lim y→∞ yIm (τ0(iy)h, h) < ∞, h ∈ ker Bτ0. yIm (τ (iy)h, h) = ∞, h ∈ H, h 6= 0. self-adjoint. (3.39) lim y→∞ 20 V.I. MOGILEVSKII (ii) C(eAτ ) is self-adjoint and transversal with A0 if and only if τ ∈ R[H] and yIm (τ (iy)h, h) < ∞, h ∈ H. lim y→∞ In this case where Nτ = N ∗ C(eAτ ) = A−Nτ = {bf ∈ A∗ : Γ1bf = −Nτ Γ0bf}, τ ∈ B(H) is the operator given by Nτ = lim y→∞ τ (iy). Proof. (i) According to Theorem 3.19 C(eAτ ) = Aθc with a symmetric linear relation θc in H defined by (3.12). It follows from (3.12) that mul θc = ran Bτ0 ⊕ K and therefore H ⊖ mul θc = H0 ⊖ ran Bτ0 = ker Bτ0. Moreover by (3.12) dom θc = dom Nτ0 and hence the inclusion H ⊖ mul θc ⊂ dom θc is equivalent to (3.40). Therefore by Lemma 2.7, (iii) (3.40) is equivalent to self-adjointness of θc, which by Proposition 2.10, (v) is equivalent of the statement (i) is implied by Corollary 3.14. to C(eAτ )∗ = C(eAτ ). Thus C(eAτ )∗ = C(eAτ ) if and only if (3.40) holds. The second part (ii) Statement (ii) directly follows from Theorem 3.15 and arguments before Theorem (cid:3) 3.19. In the following theorem we characterize compressions of finite-codimensional exten- relation θc = θ∗ (3.18) are valid. this theorem follow from Theorem 3.18 and arguments before Theorem 3.19. Theorem 3.22. Let under the assumption (a) τ0 be a rational function (3.28), (3.29) (according to Proposition 3.16 this assumption is equivalent to the finite-codimensionality sions eA = eA∗ of a symmetric relation A with finite deficiency indices. The statements of of eAτ ). Moreover, let A′ = Pker BA ↾ ker B and let K be the multivalued part of τ (see (2.9)). Then C(eAτ ) is a self-adjoint linear relation in H and C(eAτ ) = Aθc with the c ∈ eC(H) given by (3.36). Moreover, statements (i) and (ii) of Theorem Corollary 3.23. Let eH be a Hilbert space, let H be a subspace of a finite codimension in eH (this means that dim(eH ⊖ H) < ∞) and let eA = eA∗ be a linear relation in eH. Then the compression C(eA) = PHeA ↾ H of eA (see (3.3)) is a self-adjoint linear relation in H. Proof. Clearly, eA is an exit space extension of the symmetric relation S = S(eA) (see Therefore by Theorem 3.22 C(eA) is self-adjoint. case of an operator eA = eA∗ see [24]). (3.1)). Moreover, it was shown in the proof of Theorem 3.17 that n+(S) = n−(S) < ∞. (cid:3) Remark 3.24. Statement of Corollary 3.23 was proved by another method in [3] (for the References [1] N. I. Akhiezer, I. M.Glazman, Theory of linear operators in Hilbert space, vol. I and II, Pitman, Boston-London-Melbourne, 1981. [2] T.Ya. Azizov, A. Dijksma, Closedness and adjoints of products of operators, and compressions, Int. Equ. Oper. Theory 74 (2012), 259-269. [3] T.Ya. Azizov, A. Dijksma, and G. Wanjala, Compressions of maximal dissipative and self-adjoint linear relations and of dilations, Linear Algeb. Appl. 439 (2013), 771- 792. [4] T.Ya. Azizov, B. ´Curgus, and A. Dijksma, Finite-codimensional compressions of symmetric and self-adjoint linear relations in Krein spaces, Int. Equ. Oper. Theory 86 (2016), 71-95. [5] V. M. Bruk, Extensions of symmetric relations, Math. Notes 22 (1977), no.6, 953 -- 958. [6] V.A. Derkach, S. Hassi, M.M. Malamud, and H.S.V. de Snoo, Generalized resolvents of symmetric operators and admissibility, Methods of Functional Analysis and Topology 6 (2000), no. 3 , 24 -- 55. [7] V. A. Derkach, S. Hassi, M. Malamud, and H. S.V. de Snoo,Boundary Relations and their Weyl families, Trans. Amer. Math. Soc. 358 (2006), no. 12, 5351 -- 5400. ON COMPRESSIONS OF SELF-ADJOINT EXTENSIONS 21 [8] V.A. Derkach, S. Hassi, M.M. Malamud, and H.S.V. de Snoo, Boundary relations and generalized resolvents of symmetric operators, Russian J. Math. Ph. 16 (2009), no. 1, 17 -- 60. [9] V.A. Derkach and M.M. Malamud, Generalized resolvents and the boundary value problems for Hermitian operators with gaps, J. Funct. Anal. 95 (1991),1 -- 95. [10] V.A. Derkach and M.M. Malamud, The extension theory of Hermitian operators and the moment problem, J. Math. Sciences 73 (1995), no. 2, 141-242. [11] A. Dijksma and H. Langer, Finite-dimensional self-adjoint extensions of a symmetric operator with finite defect and their compressions, In: Advances in complex analysis and operator theory, Festschrift in honor of Daniel Alpay, Birkhauser, Basel, (2017), 135-163. [12] A. Dijksma and H. Langer,Compressions of self-adjoint extensions of a symmetric operator and M.G. Kreins resolvent formula, Integr. Equ. Oper. Theory 90:41 (2018). [13] A. Dijksma and H.S.V. de Snoo, Self-adjoint extensions of symmetric subspaces, Pasif. J. Math. 54 (1974), no. 1, 71 -- 100. [14] V.I. Gorbachuk and M.L. Gorbachuk, Boundary problems for differential-operator equations, Kluver Acad. Publ., Dordrecht-Boston-London, 1991. (Russian edition: Naukova Dumka, Kiev, 1984). [15] M.G. Krein and H. Langer,On defect subspaces and generalized resolvents of a Hermitian operator in the space Πκ, Funct. Anal. Appl. 5 (1971/1972), 136 -- 146, 217 -- 228 [16] M.G Krein and H. Langer, Uber die Q-functions eines π-hermiteschen operators in raume Πκ, Acta. Sci. Math. (Szeged) 34 (1973), 191 -- 230. [17] H. Langer and B. Textorious, On generalized resolvents and Q-functions of symmetric linear relations (subspaces) in Hilbert space, Pacif. J. Math. 72(1977), no. 1 , 135 -- 165. [18] M. M. Malamud, On the formula of generalized resolvents of a nondensely defined Hermitian operator, Ukr. Math. Zh. 44(1992), no. 12, 1658 -- 1688. [19] B. Sz.-Nagy, C. Foias, Harmonic Analysis of Operators in Hilbert Space, Paris and Akad..Kiado, Budapest, 1967. [20] M.A. Nudelman, A generalization of Stengers lemma to maximal dissipative operators. Integr. Equ. Oper. Theory 70 (2011), 301-305. [21] A.V.Shtraus, On selfadjoint operators in the orthogonal sum of Hilbert spaces, Dokl. Akad. Nauk SSSR, 144 (1962), no.3, 512-515. [22] A. V. Shtraus, On one-parameter families of extensions of a symmetric operator, Izv. Akad. Nauk SSSR, Ser. Mat. 30 (1966), 1325-1352. [23] A. V. Shtraus, Extensions and generalized resolvents of a symmetric operator which is not densely defined, Izv. Akad. Nauk SSSR. Ser. Mat. 34 (1970),no.1, 175-202. (Russian); English translation: Mathematics of the USSR-Izvestiya, 4 (1970), no. 1, 179 -- 208. [24] W. Stenger, On the projection of a selfadjoint operator, Bull. Am. Math. Soc. 74(1968), 369-372. the Shtraus [25] S.M. Zagorodnyuk, Generalized resolvents of symmetric and isometric operators: approach, Ann. Funct. Anal. 4 (2013), no. 1, 175 -- 285. Department of Mathematical Analysis and Informatics, Poltava National V.G. Ko- rolenko Pedagogical University, Ostrogradski Str. 2, 36000 Poltava, Ukraine E-mail address: [email protected]
1504.05242
1
1504
2015-04-20T21:43:57
Criterion for $\mathbb{Z}_d$--symmetry of a Spectrum of a Compact Operator
[ "math.FA", "math.SP" ]
If $A$ is a compact operator in a Banach space and some power $A^q$ is nuclear we give a criterion of $\mathbb{Z}_{d}$ -- symmetry of its spectrum $\sigma{A}$ in terms of vanishing of the traces $\mathop{\mathit{Trace}} A^n$ for all $n$, $n \geq 0$, $n \neq 0 \mod d$, sufficiently large.
math.FA
math
CRITERION FOR Zd -- SYMMETRY OF A SPECTRUM OF A COMPACT OPERATOR BORIS S. MITYAGIN Abstract. If A is a compact operator in a Banach space and some power Aq is nuclear we give a criterion of Zd -- symmetry of its spectrum σ(A) in terms of vanishing of the traces Trace An for all n, n ≥ 0, n 6= 0 mod d, sufficiently large. In the case of matrices, or linear operators T : X → X in a finite-dimensional space, one can check (prove) that the following conditions are equivalent. (a) The spectrum of T is symmetric, or Z2-symmetric, i.e., λ ∈ σ(T ) → −λ ∈ σ(T ) and their algebraic multiplicities m(λ), m(−λ) are equal; (b) Trace T p = 0 for all odd p ∈ N. M. Zelikin [Zel08] observed and proved that this claim could be extended to S1, the trace- class operators in a Hilbert space. We will show that such claims could be made (i) in general Banach spaces; (ii) for Zd symmetry of a spectrum, d ≥ 2. Of course, we need to make sure that Trace is well -- defined if we write conditions like (b). Then, the formula for the trace Trace A = Pj λj(A) should be properly explained if we use it. We now recall a few notions and facts about nuclear operators (see more in [Kon86]). An operator A : X → Y between two Banach spaces is called nuclear if it has represen- tation (1) where (2) Ax = qXk=1 akfk(x)yk, q ≤ ∞ ak > 0, kfkX ′k ≤ 1, a∗ =X ak < ∞, kykY k ≤ 1, 1 and ∀k. 2 BORIS S. MITYAGIN A linear space of nuclear operators X → Y is a Banach space N (X; Y ) with the norm (3) kAk1 = inf{a∗ : (1), (2)}. A linear functional Trace is well-defined on N (X; X) (for any Banach Space X) by (4) Of course, (5) and kTracek = 1. Trace A = akfk(yk). qXk=1 kTrace Ak ≤ kAk1, A. Grothendieck [Gro95] showed that for operators (1) [X = Y ], (6) (7) (8) if qXk=1 a2/3 k < ∞ then Xλj(A) < ∞ Trace A =X λj(A). where points of the spectrum σ(A) are enumerated with their multiplicity, and The presentation (4) with (1) to (3) gives a factorization (9) where (10) (11) with (12) A = J F, X F−→ ℓ2(N) J−→ X, F x = J ξ = ∞X1 ∞X1 a1/2 k fk(x)ek, and a1/2 k ξkyk, kF k ≤ (a∗)1/2, kJk ≤ (a∗)1/2 Moreover, the product F J is a Hilbert-Schmidt operator, or of the Schatten class S2 in a Hilbert space ℓ2(N); see more in [GK69], [Sim79]. Indeed, (13) hF J ek, emi = a1/2 k a1/2 m fm(yk) CRITERION FOR Zd -- SYMMETRY 3 and (14) so kF Jk2 ≤ a∗. hF J ek, emi2 = ∞Xk,m=1 ∞Xk,m=1 akamfm(yk)2 ≤ (a∗)2 By Holder inequality for Schatten classes ([GK69] or [Sim79]), (15) kBCDk2/3 ≤ kBk2kCk2kDk2 so (F J)3 ∈ S2/3 and has a representation (16) (F J)3 = where kfkk, khkk ≤ 1 and (17) Therefore, ∞Xk=1 ∞Xk=1 ckh·, fkihk, c > 0, c2/3 k < ∞. (18) has 2 3 (19) with (20) A4 = J(F J)3F = ckhF (·), fkiJ hk ∞Xk=1 -property (6) and λj(Aq) < ∞ for all q ≥ 4, ∞Xj=1 Trace Aq = λj(Aq) ∞Xj=1 More careful geometric analysis, based on approximative characteristics of operators [MP66], [Pie87] -- if we use [Kon80], or [Kon86, Theorem 4.a.6, p. 227] -- shows that we can lower q in (19), (20) to 3. Indeed, (F J)2 is in S1(ℓ2(N)), so there are finite-dimensional operators Gn, Rank Gn ≤ n, such that (21) Then (22) αn < ∞, where αn := k(F J)2 − Gnk Xn kA3 − J GnF k ≤ a∗ · αn 4 BORIS S. MITYAGIN and by [Kon86, Theorem 4.a.6] (23) (24) and (25) A3 is nuclear, λj(A3) ≤ 2a∗Xn Xj αn < ∞, Trace A3 =Xj λj(A3). But this remark will not improve our Theorem 1 (below) in an essential way (just in (41) we can say p ≥ p∗ ≥ 3q∗). In a Hilbert space X = H by Lisdkiı Theorem [Lid59], for any trace-class operator C ∈ S1, (26) and (27) λj(C) < ∞ ∞Xj=1 Trace C = λj(C). ∞Xj=1 Maybe, talking just about nuclear operators, M. Zelikin considered in [Zel08, Thm. 2] only Hilbert spaces. Before stating our main result let us recall [DS58, Chapter VII, Sections 3 and 4] elements of Riesz theory of compact operators. If T : X → X is compact its spectrum σ(T ) is discrete with 0 being the only accumulation point, and it has the following properties (i) for any ρ > 0, σ(T ) ∩ {z : z ≥ ρ} is a finite set; (ii) if (28) (29) δ(α) = 1 2 min{α − λ : λ ∈ σ(T ), λ 6= α} [so δ(α) > 0 for any α ∈ C \ 0] and P (α) = 1 2πi Z z−α=δ(α) (z − T )−1 dz, CRITERION FOR Zd -- SYMMETRY 5 then with (30) (31) m(α) = Rank P (α) < ∞, α ∈ C \ {0} m(α) = 0 if and only if α 6∈ σ(T ). For α ∈ σ(T ) \ 0, m(α) is an algebraic multiplicity of an eigenvalue α. The operational calculus [DS58, Chapter VII, Sections 3 and 4] explains that for any ρ > 0 such that (32) we have (33) σ(T ) ∩ {z = ρ} = ∅ T = Xα>ρ T (α) + S, where T (α) = 1 2πi Z z−α=δ(α) z(z − T )−1 dz is an operator of rank m(α) with (34) and (35) σ(T (α)) = {α}, S = 1 2πi Zz=ρ z(z − T )−1 dz. Moreover, for any entire function F (z), say, for polynomials, (36) F (T ) = Xα>ρ F (T (α)) + F (S), where by the Riesz-Cauchy formulae, 1 2πi Z z−α=δ(α) (37) F (T (α)) = It follows that F (z)(z − T )−1 dz, F (S) = 1 2πi Zz=ρ F (z)(z − T )−1 dz. (38) (39) Trace F (T (α)) = F (α) · m(α). F (T (α)) = 0 if F (j)(α) = 0, 0 ≤ j ≤ m(α). Now we are ready to prove 6 BORIS S. MITYAGIN Theorem 1. Let T be a compact operator in a Banach space X, and some power T q∗ is a nuclear operator. Then σ(T ) is Zd -- symmetric, i.e., for any β ∈ C \ {0}, (40) if and only if (41) m(βωk) = m(β) for all k = 0, 1, . . . d − 1, ω = exp(cid:18)i 2π d (cid:19) Trace T dp+r = 0, 1 ≤ r ≤ d − 1, for all sufficiently large p, say p ≥ p∗ ≥ 4q∗. Of course, if d = 2, this is an extension of [Zel08] , Thm. 2, to a Banach case. Proof. Part 1: (40) ⇒ (41). This is an "algebraic" claim although first we notice: the assumption p ≥ 4q∗ guarantees that all operators T n, n = dp+r, in (41) satisfy 2 3 -- condition so by Grothendieck theorem (42) Trace T n = λj(T n) ∞Xj=1 and the absolute convergence permits to rearrange the terms of the right sum as we wish to write (43) With (44) Trace T n =X µ · m(µ; T n) m(µ; T n) = 0 for µ 6∈ σ(T n) we can "add" the terms with µ 6∈ σ(T n) and this does not change the right side in (43). For (45) so (46) n = dp + r ∈ (41) define g = gcd{r, d} r = ag, d = bg, (a, b) = 1 and with r ≤ d − 1 we have 1 ≤ a < b. For any µ ∈ C \ {0} take its Zb-orbit, i.e., (47) eµ = {µ · τ j : 0 ≤ j < b}, τ = ωq = exp(cid:18)i 2π b (cid:19). CRITERION FOR Zd -- SYMMETRY 7 The sum in (43) could be written as (48) XZb−orbits b−1Xj=0 µτ j · m(µτ j; T n) where for certainty µ in the orbit (47) is chosen as µ = µeiϑ, 0 ≤ ϑ < 2π b . Now we will show that the sum in (48) over each orbit is equal to zero. With numbers as in (45) put κ = exp(cid:18)i 2π n(cid:19) so κn = 1 and notice that if µ = λn, we choose λ = µ1/neiϑ′ , ϑ′ = ϑ n , (cid:16)λωk(cid:17)n = µωk(dp+r) = µωkr = µτ ak µτ j m(µτ j; T n) = (1) = (49) then (50) and (51) b−1Xj=0 m(λωkκs; T ) (2) = m(λκs · ωk; T ) (3) = = = 1 g 1 g µτ akm(µτ ak; T n) d−1Xk=0 d−1Xk=0(cid:16)λωk(cid:17)n n−1Xs=0 d−1Xk=0(cid:16)λωk(cid:17)n n−1Xs=0 d−1Xk=0 n−1Xs=0 m(λκs; T )! µ n−1Xs=0 b−1Xj=0 m(λκs; T )µ 1 g 1 g = τ αk (4) = τ j (5) = 0 The steps in (51) are justified in the following way. (1) comes from (50). (2) is just (3) uses in essential way the theorem's the change of order of the double summation. assumption (40) on m(βωk) being independent on k. (4) is bases on the properties of the roots ω, τ , ωd = 1, τ = ωg under (46). Of course, in (5) times over {τ j}b−1 j=0. Part (40) ⇒ (41) is proven. b−1Xj=0 τ j = 0, and {τ ak}d−1 k=0 runs g (cid:3) 8 BORIS S. MITYAGIN Proof. Part 2: (41) ⇒ (40). Take λ 6= 0 and as before (52) n = dp∗ + dp + r, 1 ≤ r ≤ d − 1, p ≥ 0 and 0 < ρ < λ is such that σ(T ) ∩ {z ∈ C : z = ρ} = ∅, (53) with (54) (55) where (56) (57) Then by (39) (58) (59) being the Zd-orbit of λ. Now we use (36) for the special choice F = Fpr with eλ = {λωk : 0 ≤ k ≤ d − 1} Fpr(z) =(cid:16) z λ(cid:17)dp∗+dp+r (cid:18) zd − αd λd − αd(cid:19)m(α) = ϕ(z), ϕ(z) = Yα≥ρ α∈σ(T ) α6∈eλ = ψ(zd), and ψ is a polynomial. ϕ(T (α)) = 0, Fpr(T (α)) = 0, α > ρ ∀α 6∈eλ, Trace Fpr(T (β)) = m(β)Fpr(β) = m(λωk)ωkr. but for β ∈eλ, i.e., β = λωk, (60) Therefore, (61) where (62) Put (63) ωkrm(λωk) + Trace Fpr(S) Trace Fpr(T ) = Fpr(S) =(cid:18) T λ(cid:19)dp∗ · d−1Xk=0 2πi Zz=ρ (cid:16) z 1 λ(cid:17)dp+r ϕ(z)(z − T )−1 dz. Φ = max{ϕ(z) : z ≤ ρ} CRITERION FOR Zd -- SYMMETRY 9 and with (53) (64) Then (65) where (66) and (67) M = max{kR(z; T )k : z = ρ} < ∞ kFpr(S)k1 ≤ Ctp, any r, 1 ≤ r ≤ d − 1, C = Φ · M · ρ · kT dp∗k1 λdp∗ λ(cid:19)d t =(cid:18) ρ < 1. Now by (41) and (61) (68) 0 = d−1Xk=0 ωkrm(λωk) + ξpr for any p ≥ 1 and r, 1 ≤ r ≤ d − 1. d−1Xk=0 The sum (69) does not depend on p but the remainder by (65) to (67) have estimates ξpr ≤ Ctp so ξpr → 0 (p → ∞) This implies by (68) (70) or (71) ωkrm(λωk) = 0, ∀r, 1 ≤ r ≤ d − 1 d−1Xk=0 yk = m(λωk), 1 ≤ k ≤ d − 1. is a solution of the system (72) d−1Xk=1 ωkryk = −y0, 1 ≤ r ≤ d − 1. Its determinant is of Vandermonde type so (73) det{ωkr}d−1 k,r=1 6= 0, 10 BORIS S. MITYAGIN and the identities (74) show that by (72) (ωr)k = 0, d−1Xk=0 ∀r, 1 ≤ r ≤ d − 1 (75) yk = y0, i.e., m(λωk) = m(ω), ∀k, 1 ≤ k ≤ d − 1. This proves that the multiplicity function m is constant on Zd -- orbits in C \ {0}, and (40) is proven. (cid:3) It is worth to notice that the proof of Part II does not use any form of Grothendick or Lidskii thoerem but it uses only properties of a linear function Trace on N (X; X) and an elementary formula for Trace K when K is an operator of finite rank. Acknowledgements I thank M. Zelikin who told me his results [Zel08] in September 2007 during the 18th Crimean Autumn Mathematical School -- Symposium, in Laspi -- Batiliman, Crimea. At that time I realized that his result could be extended to the Banach space case. In the late 2007 I had useful discussions with P. Djakov, E. Gorin, H. Konig, M. Solomyak. Recently, during the Aleksander Pelczynski Memorial Conference, July 2014, Bedlewo, Poland, O. Reinov brought my attention to his papers [Rei12], [RL13]. I thank C. Baker and O. Reinov for recent discussions. References [DS58] Nelson Dunford and Jacob T. Schwartz. Linear Operators. I. General Theory. With the assistance of W. G. Bade and R. G. Bartle. Pure and Applied Mathe- matics, Vol. 7. Interscience Publishers, Inc., New York; Interscience Publishers, Ltd., London, 1958, pp. xiv+858. [GK69] I. C. Gohberg and M. G. Kreın. Introduction to the theory of linear nonselfadjoint operators. Translated from the Russian by A. Feinstein. Translations of Mathe- matical Monographs, Vol. 18. American Mathematical Society, Providence, R.I., 1969, pp. xv+378. [Gro95] A. Grothendieck. "Produits tensoriels topologiques et espaces nucl´eaires". In: S´eminaire Bourbaki, Vol. 2. Soc. Math. France, Paris, 1995, Exp. No. 69, 193 -- 200. REFERENCES 11 [Kon80] Hermann Konig. "s-numbers, eigenvalues and the trace theorem in Banach spaces". In: Studia Math. 67.2 (1980), pp. 157 -- 172. issn: 0039-3223. [Kon86] Hermann Konig. Eigenvalue distribution of compact operators. Vol. 16. Operator Theory: Advances and Applications. Birkhauser Verlag, Basel, 1986, p. 262. isbn: 3-7643-1755-8. [Lid59] V. B. Lidskiı. "Non-selfadjoint operators with a trace". In: Dokl. Akad. Nauk SSSR 125 (1959), pp. 485 -- 487. issn: 0002-3264. [MP66] Boris S. Mityagin and A. Pelczy´nski. "Nuclear Operators and Approximative Dimension". In: Proc. Inter. Congr. Math. Moscow, 1966, pp. 366 -- 372. [Pie87] A. Pietsch. Eigenvalues and s-numbers. Vol. 43. Mathematik und ihre Anwen- dungen in Physik und Technik [Mathematics and its Applications in Physics and Technology]. Akademische Verlagsgesellschaft Geest & Portig K.-G., Leipzig, 1987, p. 360. isbn: 3-321-00012-1. [Rei12] Oleg Reinov. "Some more remarks on Grothendieck-Lidskiı trace formulas". In: J. Prime Res. Math. 8 (2012), pp. 5 -- 11. issn: 1817-2725. [RL13] Oleg Reinov and Qaisar Latif. "Grothendieck-Lidskiı theorem for subspaces of Lp-spaces". In: Math. Nachr. 286.2-3 (2013), pp. 279 -- 282. issn: 0025-584X. doi: 10.1002/mana.201100112. url: http://dx.doi.org/10.1002/mana.201100112. [Sim79] Barry Simon. Trace ideals and their applications. Vol. 35. London Mathematical Society Lecture Note Series. Cambridge University Press, Cambridge-New York, 1979, pp. viii+134. isbn: 0-521-22286-9. [Zel08] Mikhail I. Zelikin. "A criterion for the symmetry of a spectrum". In: Dokl. Akad. Nauk 418.6 (2008), pp. 737 -- 740. issn: 0869-5652. doi: 10.1134/S1064562408010328. url: http://dx.doi.org/10.1134/S1064562408010328. 231 West 18th Avenue, The Ohio State University, Columbus, OH 43201 E-mail address: [email protected]
1112.4985
2
1112
2012-01-19T16:08:02
Multipliers and Wiener-Hopf operators on weighted L^p spaces
[ "math.FA" ]
We study the operators T on the weighted space L^p commuting either with the right translations St or left translations P^+S_{-t} and we establish the existence of a symbol of T. We characterize completely the spectrum of St. We obtain a similar result for the spectrum of P^+S_{-t} and some spectral results for the bounded operators commuting with (St), t>0 or with (P^+St), t<0.
math.FA
math
MULTIPLIERS AND WIENER-HOPF OPERATORS ON WEIGHTED Lp SPACES VIOLETA PETKOVA Abstract. We study the multipliers M (bounded operators commuting with the trans- lations) on weighted spaces Lp ω(R). We establish the existence of a symbol µM for M and some spectral results for the translations St and the multipliers. We also study the operators T on the weighted space Lp ω(R+) commuting either with the right translations St, t ∈ R+, or left translations P +S−t, t ∈ R+, and we establish the existence of a sym- bol µ of T . We characterize completely the spectrum σ(St) of the operator St proving that σ(St) = {z ∈ C : z ≤ etα0}, where α0 is the growth bound of (St)t≥0. We obtain a similar result for the spectrum of (P +S−t), t ≥ 0. Moreover, for an operator T commuting with St, t ≥ 0, we establish the inclusion µ(O) ⊂ σ(T ), where O = {z ∈ C : Im z < α0}. 1. Introduction Let E be a Banach space of functions on R. For t ∈ R, define the translation by t on E by Stf (x) = f (x − t), a.e., ∀f ∈ E. We call a multiplier on E, every bounded operator on E commuting with St for every t ∈ R. For the multipliers on a Hilbert space we have the existence of a symbol and some spectral results concerning the translations and the multipliers are obtained by using this property of the multipliers (see [7], [8]). In the arguments exploited in [7], [8] the spectral mapping theorem of Gearhart [3] for semigroups in Hilbert spaces plays an essential role. The first purpose of this paper is to extend the main results in [8], [7] concerning the existence of the symbol of a multiplier as well as the spectral results in the case where E is a weighted Lp ω(R) space. For general Banach spaces the characterization of the spectrum of the semigroup V (t) = etG by the resolvent of its generator G is much more complicated than for semigroups in Hilbert spaces (see for instance [4]). In particular, the statements of Lemma 1, 2 and 3 (see Section 2) are rather difficult to prove and for general Banach spaces this problem remains open. In this paper we restrict our attention to Lp ω(R), 1 ≤ p < ∞, weighted spaces. The advantage that we take account is that the semigroup of the translations (St) preserves the positive functions. For semigroups having this special property in the spaces Lp ω(R) we have a spectral mapping theorem (see [1], [12], [13]). We obtain Theorems 1-4 for multipliers on Lp ω(R) and in this work we explain only these parts of the proofs which are based on spectral mapping techniques and which are different from the arguments used to establish Theorems 1-4 in the particular case 1 2 VIOLETA PETKOVA p = 2 (see for more details [8], [7]). For a Banach space E denote by E ′ the dual space of E. For f ∈ E, g ∈ E ′, denote by < f, g > the duality. Let p ≥ 1, and let ω be a weight on R. More precisely, ω is a positive, continuous function such that ω(x + t) ω(x) sup x∈R < +∞, ∀t ∈ R. Let Lp ω(R) be the set of measurable functions on R such that kf kp,ω =(cid:16)ZR f (x)pω(x)pdx(cid:17)1/p < +∞, 1 ≤ p < +∞. Let Cc(R) (resp. Cc(R+) ) be the space of continuous functions on R (resp. R+) with compact support in R (resp. R+). Notice that Cc(R) is dense in Lp In the following we set E = Lp for 1 ≤ p < +∞. In this case ω(R) and we consider only Banach spaces having this form ω(R). hf, gi =ZR f (x)¯g(x)ω2(x)dx and hf, gi ≤ kf kp,ωkgkq,ω, for 1 < p < +∞, where 1 p + 1 q = 1. For p = 1, we have E′ = L∞ ω (R) = {f is measurable : f (x)ω(x) < ∞, a.e.} and If M is a multiplier on E then, there exists a distribution µ such that kgk∞,ω = esssup {f (x)ω(x), x ∈ R}. For φ ∈ Cc(R+), the operator Mf = µ ∗ f, ∀f ∈ Cc(R+). Mφ : Lp ω(R) ∋ f −→ φ ∗ f is a multiplier on E. Introduce α0 = lim t→+∞ ln kStk 1 t , α1 = lim t→+∞ ln kS−tk 1 t . It is easy to see that α1 + α0 ≥ 0. Consider U = {z ∈ C, Im z ∈ [−α1, α0]}. For an operator T denote by ρ(T ) the spectral radius of T and by σ(T ) the spectrum of T . It is well known that ρ(St) = eα0t, for t ≥ 0. Given a function f and a ∈ C, denote by (f )a the function R ∋ x −→ f (x)eax WIENER-HOPF OPERATORS 3 and denote by M the algebra of the multipliers on E. We note by g the Fourier transform of a function g ∈ L2(R). Our first result is a theorem saying that every multiplier on E has a representation by a symbol. Theorem 1. Let M be a multiplier on E. Then 1) For a ∈ [−α1, α0], we have (Mf )a ∈ L2(R), for every f ∈ E such that (f )a ∈ L2(R). 2) For a ∈ [−α1, α0], there exists a function νa ∈ L∞(R) such that \(Mf )a(x) = νa(x)d(f )a(x), ∀f ∈ E, with (f )a ∈ L2(R), a.e. Moreover, we have kνak∞ ≤ CkMk, ∀a ∈ [−α1, α0]. 3) If ◦ U 6= ∅, there exists a function ν ∈ H∞( ◦ U ) such that dMf (z) = ν(z) f (z), z ∈ ◦ U, ∀f ∈ C ∞ c (R), where dMf (ia + x) = \(Mf )a(x), for a ∈] − α1, α0[, f ∈ C ∞ c (R). The function ν is called the symbol of M. The above result is similar to that established in [8], [7] and the novelty is that we treat Banach spaces Lp ω(R) and not only Hilbert spaces. Define A as the closed Banach algebra generated by the operators Mφ, for φ ∈ Cc(R). Notice that A is a commutative algebra. Our second result concerns the spectra of St and M ∈ M. Theorem 2. We have i) σ(St) = {z ∈ C, e−α1t ≤ z ≤ eα0t}, ∀t ∈ R. Let M ∈ M and let µM be the symbol of M. ii) We have iii) If M ∈ A, then we have µM (U) ⊂ σ(M). µM (U) = σ(M). (1.1) (1.2) (1.3) The equality (1.3) may be considered as a weak spectral mapping property (see [2]) for operators in the Banach algebra A. On the other hand, it is important to note that if M ∈ M, but M /∈ A, in general we have µM (U) 6= σ(M). For the space E = L1(R), there exists a counter-example (see section 2 and [2]). Thus the inclusion in (1.2) could be strict. In section 3, we obtain similar results for Wiener-Hopf operators on weighted Lp ω(R+) spaces. In the analysis of Wiener-Hopf operators some new difficulties appear in com- parison with the case of multipliers. Let E be a Banach space of functions on R+. Let p ≥ 1 and let ω be a weight on R+. It means that ω is a positive, continuous function such that 0 < inf x≥0 ω(x + t) ω(x) ≤ sup x≥0 ω(x + t) ω(x) < +∞, ∀t ∈ R+. 4 VIOLETA PETKOVA Let Lp ω(R+) be the set of measurable functions on R+ such that Z ∞ f (x)pω(x)pdx < +∞. 0 Notice that Cc(R+) is dense in Lp ω(R+). Let P + be the projection from L2(R−) ⊕ Lp denote by Sa the restriction of Sa on Lp denoted by S. Let I be the identity operator on Lp ω(R+) into Lp ω(R+). From now we will ω(R+) for a ≥ 0 and, for simplicity, S1 will be ω(R+). Definition 1. A bounded operator T on Lp ω(R+) is called a Wiener-Hopf operator if P +S−aT Saf = T f, ∀a ∈ R+, f ∈ Lp ω(R+). As in [5] we can show that every Wiener-Hopf operator T has a representation by a convolution. More precisely, there exists a distribution µ such that T f = P +(µ ∗ f ), ∀f ∈ C ∞ c (R+). If φ ∈ Cc(R), then the operator Lp ω(R+) ∋ f −→ P +(φ ∗ f ) is a Wiener-Hopf operator and we will denote it by Tφ. Moreover, we have but it is obvious that (P +S−aSa)f = f, ∀f ∈ Lp ω(R+), (SaP +S−a)f 6= f, for all f ∈ Lp ω(R+) with a support not included in ]a, +∞[. The fact that Sa is not invertible leads to many difficulties in contrast to the case when we deal with the space Lp ω(R). Let E be the space Lp ω(R+). As above define a0 = lim t→+∞ ln kStk 1 t , a1 = lim t→+∞ ln kS−tk 1 t and set J = [−a1, a0]. The next theorem is similar to Theorem 1. Theorem 3. Let a ∈ J and let T be a Wiener-Hopf operator. Then for every f ∈ Lp such that (f )a ∈ L2(R+), we have ω(R+) with ha ∈ L∞(R) and (T f )a = P +F −1(had(f )a) khak∞ ≤ CkT k, (1.4) where C is a constant independent of a. Moreover, if a1 + a0 > 0, the function h defined on U = {z ∈ C : Im z ∈ J} by h(z) = hIm z(Re z) is holomorphic on ◦ U. Definition 2. The function h defined in Theorem 3 is called the symbol of T . WIENER-HOPF OPERATORS 5 We are able to examine the spectrum of the operators in the space W of bounded operators on E commuting with (St)t≥0 or (P +S−t)t≥0. Let O = {z ∈ C, Im z < a0} and V = {z ∈ C, Im z < a1}. Theorem 4. We have i) σ(St) = {z ∈ C, z ≤ ea0t}, ∀t > 0. ii) σ(P +S−t) = {z ∈ C, z ≤ ea1t}, ∀t > 0. Let T ∈ W and let µT be the symbol of T . iii) If T commutes with St, ∀t ≥ 0, then we have iv) If T commutes with P +S−t, ∀t ≥ 0, then we have µT (O) ⊂ σ(T ). µT (V) ⊂ σ(T ). (1.5) (1.6) (1.7) (1.8) The equalities (1.5),(1.6) generalize the well known results for the spectra of the right and left shifts in the space of sequences l2 (see for instance, [10]). However, our proofs are based heavily on the existence of symbols for Wiener-Hopf operators and having in mind Theorem 3, we follow the arguments in [9]. In section 4, we obtain a sharp spectral result for Wiener-Hopf operators having the ω(R+). ω(R+) our result form Tφ with φ ∈ Cc(R). This result is established here for operators in spaces Lp It is important to note that even for p = 2 and for the Hilbert space L2 below is new. Theorem 5. Let φ ∈ Cc(R). Then i) if supp (φ) ⊂ R+, we have ii) if supp (φ) ⊂ R−, we have φ(O) = σ(Tφ). φ(V) = σ(Tφ). The above result yields a weak spectral mapping property and can be compared with the equality (1.3) in Theorem 2, however the proof is more complicated. Recall that we use the notation E = Lp ω(R). We start with the following 2. Multipliers on Lp ω(R) Lemma 1. Let λ ∈ C be such that eλ ∈ σ(S) and let Re λ = α0. Then there exists a sequence (fn)n∈N of functions of E and an integer k ∈ Z so that lim n→∞(cid:13)(cid:13)(cid:16)etA − e(λ+2πki)t(cid:17)fn(cid:13)(cid:13) = 0, ∀t ∈ R, kfnk = 1, ∀n ∈ N. (2.1) 6 VIOLETA PETKOVA Proof. Let A be the generator of the group (St)t∈R. It is clear that the group (St)t∈R ω(R) the results of [12], [13] say that the spectral preserves positive functions. Since E = Lp mapping theorem holds and σ(etA) \ {0} = etσ(A) = {etλ : λ ∈ σ(A)}. In particular, for the spectral bound s(A) of A we get s(A) := sup{Re z : z ∈ σ(A)} = α0. Thus eλ ∈ σ(S) \ {0} = eσ(A) yields λ + 2πki = λ0 ∈ σ(A) for some k ∈ Z. On the other hand, Re λ0 = α0, and we deduce that λ0 is on the boundary of the spectrum of A. By a well known result, this implies that λ0 is in the approximative point spectrum of A. Let µn be a sequence such that µn →n→∞ λ0, Re µn > λ0, ∀n ∈ N. Then hence k(µnI − A)−1k → ∞. Applying the uniform boundedness principle and passing to a subsequence of µn (for simplicity also denoted by µn), we may find f ∈ E such that (cid:13)(cid:13)(µnI − A)−1(cid:13)(cid:13) ≥ (dist (µn, σ(A)))−1, lim n→∞(cid:13)(cid:13)(µnI − A)−1f(cid:13)(cid:13) → ∞. Introduce fn ∈ D(A) defined by fn = (µnI − A)−1f k(µnI − A)−1f k . The identity (λ + 2πki − A)fn = (λ0 − µn)fn + (µn − A)fn implies that (λ + 2πki − A)fn → 0 as n → ∞. Then the equality (etA − et(λ+2πki))fn =(cid:16)Z t 0 e(λ+2πki)(t−s)eAsds(cid:17)(A − λ − 2πki)fn yields (2.1). (cid:3) Now we prove the following important lemma. Lemma 2. For all φ ∈ C ∞ c (R) and λ such that eλ ∈ σ(S) with Re λ = α0 we have φ(iλ + a) ≤ kMφk, ∀a ∈ R. (2.2) Proof. Let λ ∈ C be such that eλ ∈ σ(S) and Re λ = α0 and let (fn)n∈N be the sequence constructed in Lemma 1. We have 1 = kfnk = sup g∈E ′ , kgk E ′ ≤1 < fn, g > . Then, there exists gn ∈ E ′ such that < fn, gn > −1 ≤ 1 n WIENER-HOPF OPERATORS 7 and kgnkE ′ ≤ 1. Fix φ ∈ C ∞ c (R) and consider φ(iλ + a) ≤ φ(iλ + a)hfn, gni + φ(iλ + a) 1 n The first two terms on the right side of the last inequality go to 0 as n → ∞ since by Lemma 1 we have On the other hand, φ(iλa) 1 n ≤(cid:12)(cid:12)(cid:12)ZR(cid:10)φ(t)(cid:16)e(λ+2πki)t − St(cid:17)e−i(a+2πk)tfn, gn(cid:11)dt(cid:12)(cid:12)(cid:12) + hφ(t)Ste−i(a+2πk)tfn, gnidt(cid:12)(cid:12)(cid:12). +(cid:12)(cid:12)(cid:12)ZR (cid:13)(cid:13)e−i(a+2πk)t(cid:16)e(λ+2πki)t − St(cid:17)fn(cid:13)(cid:13) −→n→∞ 0. In =(cid:12)(cid:12)(cid:12)ZR < φ(t)Ste−i(a+2πk)tfn, gn > dt(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)hhZR φ(t)e−i(a+2πk)tfn(. − t)dti, gni(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)hZR (φ(. − y)ei(a+2πk)yfn(y)dy, ei(a+2πk).gni(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)h(cid:16)Mφ(ei(a+2πk).fn)(cid:17), ei(a+2πk).gni(cid:12)(cid:12)(cid:12) and In ≤ kMφkkfnkkgnkE ′ ≤ kMφk. Consequently, we deduce that (cid:3) Notice that the property (2.2) implies that φ(iλ + a) ≤ kMφk. Lemma 3. Let φ ∈ C ∞ Then we have φ(λ) ≤ kMφk, ∀λ ∈ C, provided Im λ = α0. c (R) and let λ be such that e−¯λ ∈ σ((S−1)∗) with Re λ = −α1. φ(iλ + a) ≤ k(Mφ)k, ∀a ∈ R. (2.3) t∈R acting on E ′. Let λ ∈ C be such that e−¯λ ∈ Proof. Consider the group (S−t)∗ σ((S−1)∗) and e−¯λ = ρ(S−1) = ρ((S−1)∗) = eα1. The group (S−t)∗ preserves positive functions. To prove this, assume that g(x) ≥ 0, a.e. is a positive function and let h ∈ E be such that h(x) ≥ 0, a.e. Then hh, (S−t)∗gi = hS−th, gi ≥ 0. If F (x) = ((S−t)∗g)(x) < 0 for x ∈ Λ ⊂ R and Λ has a positive measure, we choose h(x) = 1Λ(x). Then h1Λ, F i ≤ 0 and we conclude that F (x) = 0 a.e. in Λ which is a contradiction. For the group (S−t)∗ the spectral mapping theorem holds and, by the and we deduce that for every t ∈ R we have ≤ k(St)∗kE ′→E ′(cid:13)(cid:13)(cid:16)e(−¯λ+2πmi)t − (S−t)∗(cid:17)gk(cid:13)(cid:13)E ′ k→∞(cid:13)(cid:13)(cid:16)(St)∗ − e(¯λ−2πmi)t(cid:17)gk(cid:13)(cid:13)E ′ = 0. lim 8 VIOLETA PETKOVA same argument as in Lemma 1, we prove that there exists a sequence (gk)k∈N of functions of E ′ and an integer m so that for all t ∈ R, k(etB − e(−¯λ+2πmi)t))gkkE ′ = 0 lim k→∞ and kgkkE ′ = 1. Since S−tSt = I, we have (St)∗(S−t)∗ = I. This implies that (cid:13)(cid:13)(St)∗gk − e(¯λ−2πmi)tgk(cid:13)(cid:13)E ′ =(cid:13)(cid:13)(cid:16)(St)∗ − e(¯λ−2πmi)t(St)∗(S−t)∗(cid:17)gk(cid:13)(cid:13)E ′ For 1 < p < +∞ the space E = Lp with E. Consequently, since kgkkE ′ = 1, there exists fk ∈ E such that ω(R) is reflexive and the dual to E′ can be identified < fk, gk > −1 ≤ 1 k , kfkkE ≤ 1. (2.4) For p = 1 the space L1 In this case the dual to L1 the set ω(R) is not reflexive and to arrange (2.4), we use another argument. ω (R) = 1. Fix 0 < ǫ < 1 and consider ω (R). Let kgkL∞ ω(R) is L∞ Mǫ,m = {x ∈ R : g(x)ω(x) ≥ 1 − ǫ, m ≤ x < m + 1}, m ∈ Z. If µ(Mǫ,m) (the Lebesgue measure of Mǫ,m) is zero for all m ∈ Z, we obtain a contra- ω (R) = 1. Thus there exists r ∈ Z such that µ(Mǫ,r) > 0. Now we diction with kgkL∞ take f (x) = 1Mǫ,r(x)ei arg(g(x)) µ(Mǫ,r)ω2(x) . Then 1 ≥ hf, gi =Z r+1 r f (x)¯g(x)ω2(x)dx ≥ 1 − ǫ and we can obtain (2.4) choosing ǫ = 1/k. Passing to the proof of (2.3), we get φ(iλ + a) ≤ φ(iλ + a) < fk, gk > + φ(iλ + a) 1 k ≤(cid:12)(cid:12)(cid:12)ZR < φ(t)e−i(a+2πm)tfk,(cid:16)e < φ(t)St(cid:16)e−i(a+2πm)tfk(cid:17), gk > dt(cid:12)(cid:12)(cid:12) + 1 k ¯(λ−2πmi)t − (St)∗(cid:17)gk > dt(cid:12)(cid:12)(cid:12) φ(iλ + a) = J ′ 1 k k → 0 as k → ∞. For I ′ k + I ′ k + From the argument above we deduce that J ′ +(cid:12)(cid:12)(cid:12)ZR same argument as in the proof of Lemma 2 and we deduce φ(iλ + a). k we apply the φ(iλ + a) ≤ kMφk. (cid:3) WIENER-HOPF OPERATORS 9 For the proof of Theorem 1 we apply the argument in [7] and Lemmas 2-3. There exists eλ0 ∈ σ(S) such that Re λ0 = α0. Then for every z ∈ C with Im z = α0 we have ϕ(z) ≤ kMϕk. Also there exists e−λ1 ∈ σ((S−1)∗) with Re λ1 = −α1 and for every z ∈ C with Im z = −α1 we have ϕ(z) ≤ kMϕk. Applying Phragmen-Lindeloff theorem for the Fourier transform of ϕ ∈ C ∞ domain {z ∈ C : −α1 ≤ Im z ≤ α0}, we deduce c (R) in the ϕ(z) ≤ kMϕk for z ∈ U. Next we exploit the fact that M can be approximated by Mϕ with respect to the strong operator topology (see [6] for a very general setup covering our case). We complete the proof repeating the arguments from [6], [7] and since this leads to minor modifications, we omit the details. To obtain Theorem 2 we follow the same argument as in [8] and the proof is omitted. To see that in (1.2) the inclusion may be strict, consider a measure η on R such that the operator Mη : f −→ZR Sx(f )dη(x) is bounded on L1(R). For this it is enough to have RR dη(x) < ∞. Then Mη is a multiplier on L1(R) with symbol η(t) =ZR e−ixtdη(x). On the other hand, there exists a bounded measure η on R such that η(R) 6= σ(Mη) (see for details [2]). In L1(R) we have α0 = α1 = 0 and U = R. So we have not the property (1.2) in Theorem 2 for every multiplier even in the case L1(R). We need the following lemmas. 3. Wiener-Hopf operators Lemma 4. Let φ ∈ Cc(R+). The operator Tφ commutes with St, ∀t > 0, if and only if the support of φ is in R+. Proof. Consider φ ∈ Cc(R+) and suppose that Tφ commutes with St, t ≥ 0. We write φ = φχR− + φχR+. 10 VIOLETA PETKOVA If Tφ commutes with St, t ≥ 0, then the operator TφχR− commutes too. Let ψ = φχR− and fix a > 0 such that ψ has a support in [−a, 0]. Setting f = χ[0,a], we get Saf = χ[a,2a]. For x ≥ 0 we have P +(ψ ∗ Saf )(x) =Z 0 ψ(t)χ{a≤x−t≤2a}dt =Z min(x−a,0) −a ψ(t)dt. max(−a,−2a+x) Since P +(ψ ∗ Saf ) = SaP +(ψ ∗ f ), for x ∈ [0, a], we deduce P +(ψ ∗ Saf )(x) = 0 and Z x−a −a ψ(t)dt = 0, ∀x ∈ [0, a]. This implies that ψ(t) = 0, for t ∈ [−a, 0] hence supp(φ) ⊂ R+. (cid:3) Next we establish the following Lemma 5. Let Tφ, φ ∈ Cc(R). Then Tφ commutes with P +(S−t), ∀t > 0 if and only if supp(φ) ⊂ R−. Proof. For φ ∈ Cc(R), suppose that Tφ commutes with P +(S−t), ∀t > 0. Set ψ = φχR+. There exists a > 0 such that supp(ψ) ⊂ [0, a]. We have P +(ψ ∗ P +S−aχ[0,a]) = 0 and then P +S−a(P +ψ ∗ χ[0,a]) = 0. This implies that On the other hand, for x > a we have (ψ ∗ χ[0,a])(x) = 0, ∀x > a. (ψ ∗ χ[0,a])(x) =ZR ψ(t)χ[0,a](x − t)dt =Z min(a,x) max(0,x−a) ψ(t)dt =Z a x−a ψ(t)dt. ǫ ψ(t)dt = 0, ∀a > ǫ > 0 and we get ψ = 0. Thus we conclude that supp(φ) ⊂ R−. HenceR a (cid:3) It is clear that (St)t≥0 and (P +(S−t))t≥0 form continuous semigroups and these semi- groups preserve positive functions. Moreover, by using the equality h(P +St)h, gi = hh, (P +S−t)∗gi, we conclude that the semigroup (P +S−t)∗ preserve positive functions. The issue is that for St and (P +S−t)∗ the spectral mapping theorem holds and we may repeat the arguments used in section 2. Thus we obtain the following Lemma 6. 1) For all φ ∈ C ∞ and Re λ = a0, we have c (R) such that supp(φ) ⊂ R+, for λ such that eλ ∈ σ(S) φ(iλ + a) ≤ kTφk, ∀a ∈ R. 2) For all φ ∈ C ∞ Re λ = −a1, we have c (R) such that supp(φ) ⊂ R− and for λ such that e−¯λ ∈ σ((S−1)∗) and φ(iλ + a) ≤ kTφk, ∀a ∈ R. WIENER-HOPF OPERATORS 11 Proof. Let A be the generator of the semi-group (St)t≥0. First we obtain using the same arguments as in the proof of Lemma 1 that for λ such that eλ ∈ σ(S) and Re λ = a0, there exists a sequence (fn) of functions of E and an integer k ∈ Z so that lim n→∞(cid:13)(cid:13)(cid:16)etA − e(λ+2kπi)t(cid:17)fn(cid:13)(cid:13) = 0, ∀t ∈ R+, kfnk = 1, ∀n ∈ N. Then we notice that Thus So we have k(P +S−t − e−(λ+2kπi)t)fnk = k(P +S−t − e−(λ+2kπi)tP +S−tSt)fnk ≤ kP +S−tke−(λ+2kπi)tk(e(λ+2kπi)t − St)fnk, ∀t ∈ R+. lim n→+∞ k(P +S−t − e−(λ+2kπi)t)fnk = 0, ∀t ∈ R+. lim n→+∞ k(cid:16)P +St − e(λ+2kπi)t(cid:17)fnk = 0, ∀t ∈ R. Using the same arguments as in the proof of Lemma 2, we obtain φ(iλ + a) ≤ kTφk, ∀a ∈ R, ∀φ ∈ C ∞ c (R) and λ such that eλ ∈ σ(S) and Re λ = a0. semi-group ((P +S−t)∗)t≥0. (cid:3) In the same way we prove 2) using the To establish Theorem 3, we use Lemma 6 and we follow with trivial modifications the arguments in [5], [7], [8]. We omit the details. For the proof of Theorem 4 we repeat the arguments in [9]. Now we pass to the proof of Theorem 5. Proof of Theorem 5. Let A be the commutative algebra generated by Tφ for all φ in Cc(R+) with support in R+ and Sx, for all x ∈ R+. Denote by bA the set of the characters on A. Let β ∈ σ(Tφ) \ {0}. Then there exists γ ∈ bA such that β = γ(Tφ). We will prove the following equality This result il not trivial because we cannot commute γ with the Bochner integralRR+ φ(x)Sxdx. γ(Tφ) =ZR+ φ(x)γ(Sx)dx. Set θγ(x) = γ(Sx) = γ(Sx ◦ Tφ) γ(Tφ) , ∀x ∈ R+. Let ψ ∈ Cc(R+) and let supp(ψ) ⊂ K, where K is a compact subset of R+. Suppose that (ψn)n≥0 ⊂ CK(R+) is a sequence converging to ψ uniformly on K. For every g ∈ E, we get kTψng − Tψgk ≤ kψn − ψk∞ sup y∈K kSykkgk 12 VIOLETA PETKOVA and this implies that limn→+∞ kTψn − Tψk = 0. This shows that the linear map ψ −→ Tψ is sequentially continuous and hence it is continuous from Cc(R+) into A. Since the map is continuous from R+ into Cc(R+), we conclude that the map x −→ Sx ◦ Tφ = TSx(φ) x −→ Sx(ψ) is continuous from R+ into A. Consequently, the function θγ is continuous on R+. Intro- duce η : Cc(R+) ∋ ψ −→ γ(Tψ). The map η is a continuous linear form on Cc(R+) and applying Riesz representation theorem, there exists some Borel measure µ (see for instance, [11]) such that η(ψ) =ZR+ ψ(x)dµ(x), ∀ψ ∈ Cc(R+). This implies that for all f , ψ ∈ Cc(R+), we have (ψ ∗ f )(t)dµ(t) ψ(x)f (t − x)dx(cid:17)dµ(t). f (t − x)dµ(t)(cid:17)dx =ZR+ γ(Tψ ◦ Tf ) =ZR+ =ZR+(cid:16)ZR+ ψ(x)(cid:16)ZR+ γ(Tφ) =ZR+ Using the Fubini theorem, we obtain γ(Tψ ◦ Tf ) =ZR+ and replacing f and ψ by φ, we get ψ(x)γ(Sx ◦ Tf )dx φ(x)θγ(x)dx, ∀φ ∈ Cc(R+). (3.1) Notice that θγ(x + y) = θγ(x)θγ(y), ∀x, y ∈ R+ . We will prove that θγ(x) 6= 0, ∀x ∈ n ) = = 0 and θγ( x0 R+. Suppose θγ(x0) = 0, for x0 > 0. Then γ(Sx0) = (cid:16)γ(S x0 γ(cid:16)S x0 n(cid:17) = 0 for every n ∈ N. Since θγ is continuous on R+, n(cid:17) = θγ(0) = 1 θγ(cid:16)x0 lim n→+∞ )(cid:17)n n and we obtain a contradiction. Consequently, we have θγ(x) = γ(Sx) 6= 0, for all x ∈ R+. Now define θγ(−x) = 1 θγ(x) , ∀x ∈ R+. It is easy to check that θγ is a morphism on R. It is clear that θγ(x + y) = θγ(x)θγ(y), for (x, y) ∈ R+ × R+ and for (x, y) ∈ R− × R−. Suppose that x > y > 0, θγ(x − y) = γ(SxS−y) = γ(SxS−ySy) γ(Sy) = θγ(x) θγ(y) = θγ(x)θγ(−y). WIENER-HOPF OPERATORS 13 Moreover, θγ(y − x) = 1 θγ(x − y) = 1 θγ(x)θγ(−y) = θγ(y)θγ(−x). Since θγ satisfies θγ(x + y) = θγ(x)θγ(y), for all (x, y) ∈ R2, it is well known that this implies that there exists λ ∈ C such that θγ(x) = eλx, for all x ∈ R. On the other hand, we have γ(Sx) ∈ σ(Sx) and γ(S1) = eλ ∈ σ(S). Thus (3.1) implies with λ ∈ O. We conclude that β = γ(Tφ) = φ(−iλ) σ(Tφ) \ {0} ⊂ φ(O). Now, suppose that supp(φ) ⊂ R−. Let B be the commutative Banach algebra generated by Tψ for all ψ ∈ Cc(R−) and by P +S−x, for all x ∈ R+. Let κ ∈ σ(Tφ). Using the same arguments as above, and the set of characters bB of B, we get φ(x)eδxdx, κ =ZR− with −iδ ∈ V. This completes the proof of Theorem 5. (cid:3) 4. Comments and open problems Following the general schema of the proof of the existence of symbols for multipliers developed in [6] for locally compact abelian groups, it is natural to conjecture that an analog of Theorem 1 holds for general Banach spaces of functions under some hypothesis as we have proved this for general Hilbert space of functions in [7], [9]. Using the notations of Section 2, the crucial point is the inequality ϕ(z) ≤ kMϕk, ∀ϕ ∈ C ∞ c (R), Im z = α0. (4.1) and a similar inequality for Im z = −α1. To establish (4.1), we introduced the factor hfk, gki (see proof of Lemma 1) close to 1 and we want to estimate ϕ(z)hfk, gki. Here the sequence fk, kfkk = 1, must be chosen so that for some integers nk ∈ Z and eλ ∈ σ(S), Re λ = α0, we have lim k→∞ k(St − e(λ+2πnki)t)fkk = 0, ∀t ∈ R. (4.2) If the spectral mapping theorem is true for the group St = eAt, we have s(A) = α0 and (4.2) can be obtained as in Section 2. On the other hand, if s(A) < α0, we may construct (fk) assuming that k(A − α0 − 2πmi)−1k = +∞. sup m∈R (4.3) For Hilbert spaces (4.3) holds (see [3], [4], [1]) and author has exploited this property in [8], [7] to complete the proof of (4.2). For semigroups in Banach spaces s(A) < α0 does not implies in general (4.3) 14 VIOLETA PETKOVA (see a counter-example in Chapter V in [1] and the relation between the resolvent of A and the spectrum of St in [4]). Consequently, it is not possible to use (4.3) and to construct a sequence fk for which (4.2) holds. Of course another proof of (4.1) could be possible, and in Banach spaces of functions for which s(A) < α0 this is an open problem. References [1] K. J. Engel and R. Nagel, A short course on operator semigroups, Springer, Berlin, 2006. [2] E. Fasangov´a and P.J. Miana, Spectral mapping inclusions for Phillips functional calculus in Banach spaces and algebras, Studia Math. 167 (2005), 219-226. [3] L. Gearhart, Spectral theory for contraction semigroups on Hilbert space, Trans. AMS, 236 (1978), 385-394. [4] Y. Latushkin and S. Montgomery-Smith, Evolutionary Semigroups and Lyapunov Theorems in Ba- nach Spaces, J. Funct. Anal. 127 (1995), no. 1, 173-197. [5] V. Petkova, Wiener-Hopf operators on L2 [6] V. Petkova, Multipliers on Banach spaces of functions on a locally compact abelian group, J. London ω(R+), Arch. Math.(Basel), 84 (2005), 311-324. Math. Soc. 75 (2007), 369-390. [7] V. Petkova, Multipliers on a Hilbert space of functions on R, Serdica Math. J. 35 (2009), 207-216. [8] V. Petkova, Spectral theorem for multipliers on L2 [9] V. Petkova, Spectrum of the translations and Wiener-Hopf operators on L2 ω(R), Arch. Math. (Basel), 93 (2009), 357-368. ω(R+), Preprint 2011, (arXiv.math: 1106.4769). [10] W. C. Ridge, Approximative point spectrum of a weighted shift, Trans. AMS, 147 (1970), 349-356. [11] W. Rudin, Fourier analysis on groups, Interscience, New York, 1962. [12] L. Weis, The stability of positive semigroups on Lp-spaces, Proc. Amer. Math. Soc. 123 (1995), 3089-3094. [13] L. Weis, A short proof for the stability theorem for positive semigroups on Lp(µ), Proc. Amer. Math. Soc. 126 (1998), 325-3256. LMAM, Universit´e de Metz, UMR 7122,Ile du Saulcy 57045, Metz Cedex 1, France. E-mail address: [email protected]
1909.13536
1
1909
2019-09-30T09:01:26
Lebesgue-type inequalities in greedy approximation
[ "math.FA" ]
We present new results regarding Lebesgue-type inequalities for the Weak Chebyshev Greedy Algorithm (WCGA) in uniformly smooth Banach spaces. We improve earlier bounds in Temlyakov (Forum Math Sigma 2014), for dictionaries satisfying a new property introduced here. We apply these results to derive optimal bounds in two natural examples of sequence spaces. In particular, optimality is obtained in the case of the multivariate Haar system in Lp with 1<p<2, under the Littlewood-Paley norm.
math.FA
math
LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV Abstract. We present new results regarding Lebesgue-type inequalities for the Weak Chebyshev Greedy Algorithm (WCGA) in uniformly smooth Banach spaces. We improve earlier bounds in [19] for dictionaries satisfying a new property in- troduced here. We apply these results to derive optimal bounds in two natural examples of sequence spaces. In particular, optimality is obtained in the case of the multivariate Haar system in Lp with 1 < p ≤ 2, under the Littlewood-Paley norm. 1. Introduction This paper is devoted to theoretical aspects of sparse approximation. The main motivation for the study of sparse approximation is that many real world signals can be well approximated by sparse ones. In a general setting we are working in a Banach space X with a redundant system of elements (dictionary) D. There is a solid justification of the importance of a Banach space setting in numerical analysis in general, and in sparse approximation in particular; see, for instance, [18, Preface]. An element (function, signal) f ∈ X is said to be N-sparse with respect to D if it has j=1 cjgj, where gj ∈ D and cj is a scalar, j = 1, . . . , N. The set of all N-sparse elements is denoted by ΣN (D). For a given element f we introduce the error of best N-term approximation a representation f = PN σN (f, D) := inf a∈ΣN (D) kf − ak. In a general setting one studies algorithms (approximation methods) A = {AN (·, D)}∞ with respect to a given dictionary D. These mappings must satisfy that AN (f, D) ∈ ΣN (D), for all f ∈ X; in other words, AN provides an N-term approximant with respect to D. It is clear that for any f ∈ X and any N we have kf − AN (f, D)k ≥ σN (f, D). We are interested in such pairs (D, A ) for which the algorithm A provides approximation close to the best N-term approximation. We introduce the correspond- ing definition (see [21], p.423). Let ϑ(u) be a function such that ϑ(u) ≥ 1. N =1 DEFINITION 1.1. We say that D is a ϑ-greedy dictionary with respect to A if there exists a constant C such that for any f ∈ X and all N ∈ N we have (1.2) kf − Aϑ(N )N (f, D)k ≤ CσN (f, D). 2010 Mathematics Subject Classification. 41A65, 41A25, 41A46, 46B15, 46B20. Key words and phrases. Non-linear approximation, weak Chebyshev greedy algorithm, threshold- ing greedy algorithm, uniformly smooth Banach space, greedy basis. 1 2 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV In the case ϑ(u) = C0 ≥ 1 is a constant we call D an almost greedy dictionary with respect to A , and in the case C0 = 1 we call D a greedy dictionary. If D = Ψ is a basis then in the above definitions we replace dictionary by basis. Inequalities of the form (1.2) are called Lebesgue-type inequalities. In the case that A = {GN (·, Ψ)}∞ N =1 is the Thresholding Greedy Algorithm (TGA), the theory of greedy and almost greedy bases is well developed (see [18] and [20]). We remind that if Ψ = {ψk}∞ k=1 ckψk ∈ X, then the TGA at the Nth iteration gives an approximant k=1 is a normalized basis in X, and f =P∞ GN (f, Ψ) := N Xj=1 ckj ψkj , where ck1 ≥ ck2 ≥ . . . . In particular, it is known (see [18], p.17) that the univariate Haar system is a greedy basis with respect to TGA in Lp, for all 1 < p < ∞. Also, it is known that the TGA does not work well with respect to the trigonometric system (see, for instance, [21, Ch. 8]). It was demonstrated in the paper [19] (see also [21], Ch.8) that the Weak Chebyshev Greedy Algorithm (WCGA), which we define momentarily, works very well for a special class of dictionaries, which includes the trigonometric system. In this paper we further develop recent results from [19] concerning Lebesgue-type inequalities for the WCGA in the context of uniformly smooth Banach spaces. We mostly concentrate on the case when WCGA is applied with respect to a basis. We also emphasize that, although the theory of these algorithms is typically stated for real Banach spaces, our results are actually valid for both, real or complex Banach spaces; in particular they can be applied to the standard (complex) trigonometric basis. We now recall how this algorithm is defined; see [17] or [18, Chapter 6.2]. Let (X, k · k) be a Banach space over K = R or C. Given f ∈ X \ {0}, let Ff be an associated norming functional in X∗, that is, (1.3) kFf kX∗ = 1, and Ff (f ) = kf k. Such functionals always exist by the Hahn-Banach theorem, and are unique provided the norm k · k is smooth. Here we shall assume the stronger property that k · k is uniformly smooth of power type, that is, there exists q > 1 and a constant γ > 0 such that (1.4) ρ(t) ≤ γ tq, t > 0, where ρ(t) is the associated modulus of smoothness, given by (1.5) 2ρ(t) = sup kf k=kgk=1(cid:16)kf + tgk + kf − tgk − 2kf k(cid:17), t ∈ R. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 3 Let D = {ϕi}i∈I ⊂ X be a fixed dictionary in X, that is, a subset of unit vectors with dense span. We also fix a weakness parameter τ ∈ (0, 1]. The WCGA associated with (X, k · k, D, τ ) is defined as follows. Weak Chebyshev Greedy Algorithm (WCGA). Given f ∈ X \ {0}, we let f0 := f and define inductively vectors ϕi1, . . . , ϕin in D and f1, . . . , fn ∈ X by the following procedure: at step n + 1 we pick any ϕin+1 ∈ D such that Ffn(ϕin+1) ≥ τ sup ϕ∈D Ffn(ϕ), and let Gn+1(f ) be any element in [ϕi1, . . . , ϕin+1] such that kf − Gn+1(f )k = dist(cid:0)f, [ϕi1, . . . , ϕin+1](cid:1). Then we set fn+1 = f − Gn+1(f ), and iterate the process (indefinitely, or until the remainder fn+1 = 0). It is known from [17] that kf − Gn(f )k → 0 as n → ∞, for all f ∈ X, provided that (X, k · k) is uniformly smooth. More recently, it has been shown in [19] that Lebesgue-type inequalities hold for this algorithm. In this paper it is convenient to formulate them as follows: we search for functions φ : N → N and constants C > 0 such that, for all N ≥ 1 it holds (1.6) (cid:13)(cid:13)f − Gφ(N )(f )k ≤ C σN (f ), ∀ f ∈ X. Note, in particular, that (1.6) implies exact recovery for each f ∈ ΣN after φ(N) steps. Since the dictionary D is fixed we omit it from the notation, that is, ΣN = ΣN (D) and σN (f ) = σN (f, D). We now state a result from [19] which gives bounds for such functions φ in terms of suitable parameters depending on (X, k · k, D); see also [21, Section 8.7]. These parameters are quantified by the following properties. Property A2. We say that (X, D) satisfies property A2(U), with parameter U ≥ 1, if (cid:13)(cid:13)Xi∈A aiϕi(cid:13)(cid:13) ≤ U(cid:13)(cid:13)Xi∈B aiϕi(cid:13)(cid:13), for all finite sets A ⊂ B, and all scalars ai ∈ K. When (1.7) holds only for sets with A ≤ K, we say that ΣK has the property A2(U). Property A3. We say that (X, D) satisfies property A3(r, V ), with parameters r ∈ (0, 1] and V ≥ 1, if (1.7) (1.8) Xi∈A ai ≤ V Ar(cid:13)(cid:13)Xi∈B aiϕi(cid:13)(cid:13), for all finite sets A ⊂ B, and all scalars ai ∈ K. When (1.8) holds only for sets with A ≤ K, we say that ΣK has the property A3(r, V ). 4 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV A brief discussion on the meaning of these properties and some examples, is given in [19]; see also [21, section 8.7] or §2.2 below. The next result is a special case of Theorem 2.8 in [19]; see also [21, Theorem 8.7.18]. THEOREM 1.9. Let N ≥ 1 and assume that (X, k · k, D) has the following properties (i) ρ(t) ≤ γtq, for some γ > 0 and q ∈ (1, 2]; (ii) ΣN satisfies property A3(r, V ), for some r ∈ (0, 1] and V > 0; (iii) ΣN satisfies property A2(U), for some U ≥ 1. Then, there is a universal constant C ≥ 1 such that, for every τ ∈ (0, 1], the Lebesgue- type inequality in (1.6) holds with (1.10) where C1(τ, γ, q) = C2(q) γq′/q τ −q′. φ(N) = (cid:4)C1(τ, γ, q)(cid:0) log(U + 1)(cid:1) V q′ N rq′(cid:5), REMARK 1.11. A similar result holds with U replaced by V N in (1.10), since A3(V, r) for ΣN implies A2(U) with U = V N; see [19, Theorem 2.7]. In this paper we elaborate further on these results in the following directions. First we replace the condition (i) on (X, k · k), by a less demanding condition involving as well the system D. DEFINITION 1.12. We say that (X, k · k, D) has the property D(s, c1), for some parameters c1 > 0 and s > 1, if (1.13) for all f ∈ X \ {0} and all ϕ ∈ D. dist(f, [ϕ]) ≤ kf k(cid:0)1 − c1Ff (ϕ)s(cid:1), We shall show in Proposition 2.6 below that ρ(t) ≤ γtq implies (1.13) with s = q′ and some c1 = c1(γ, q), actually for all kϕk = 1. However, for certain (X, D) it may happen D(s, c1) holds with a better (smaller) value of s. For example, if X = ℓp, 1 < p < ∞, and D is the canonical basis, then D(s, c1) holds with s = p′, while the modulus of smoothness has power type q = min{p, 2}. This, and the more general example X = ℓp(ℓq), will be discussed in §4 below. Using this new concept we shall show the following improvement over Theorem 1.9. THEOREM 1.14. Let N ≥ 1 and assume that (X, k·k, D) has the following properties (i) (X, D) satisfies property D(s, c1), for some s > 1 and c1 > 0; (ii) ΣN satisfies property A3(r, V ), for some r ∈ (0, 1] and V > 0; (iii) ΣN satisfies property A2(U), for some U ≥ 1. Then, there is a universal constant C ≥ 1 such that, for every τ ∈ (0, 1], the Lebesgue- type inequality in (1.6) holds with (1.15) φ(N) = (cid:4)C ′ 1(τ, c1, s)(cid:0) log(U + 1)(cid:1) V s N rs(cid:5), LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 5 where C ′ 1(τ, c1, s) = C ′ 2(s) c−1 1 τ −s. As an application, we shall find optimal Lebesgue type inequalities for some special pairs (X, D). We remark that no lower bounds for functions φ satisfying (1.6) seemed to appear earlier in the literature (besides the trivial φ(N) ≥ N). Our first example concerns the space X = ℓp(ℓq) with D the canonical basis. THEOREM 1.16. Let X = ℓp(ℓq), with 1 < p, q < ∞, and D be the canonical basis. Let Then, there exists c = c(p, q, τ ) > 0 such that the WCGA satisfies (1.6) with q′ p′o. q′ , β = β(p, q) = maxn p′ φ(N) = (cid:4)c N β(cid:5). Moreover, suppose that (1.17) kx − Gψ(N )xk ≤ C σN (x), ∀ N ≥ 1, x ∈ X. Then ψ(N) ≥ c′ N β, for some c′ > 0. Concerning this example, we shall also compare the WCGA with the usual TGA, showing that the former performs better when p > q′; see Figure 4.1 below. Our second application of Theorem 1.14 regards the d-variate Haar system in X = Lp([0, 1]d). Here, however, we must renorm the space to obtain new results. We consider the Littlewood-Paley renorming, defined as follows. Let Hd p = Hp×. . .×Hp = {Hn}n≥0 denote the Lp-normalized d-variate Haar system (ie, the tensor product of (1.18) the 1-dimensional Haar basis Hp in [0, 1]). Given f =Pn≥0 cn(f )Hn ∈ Lp, we let 1 " "f " "p :=(cid:13)(cid:13)S(f )(cid:13)(cid:13)Lp, where S(f ) =(cid:16)Xn≥0 cn(f )Hn2(cid:17) 2 . This defines an equivalent norm in Lp([0, 1]d), provided 1 < p < ∞. Our next result gives the optimal growth for the Lebesgue type functions when 1 < p ≤ 2. THEOREM 1.19. Let X = Lp([0, 1]d), 1 < p ≤ 2, endowed with the norm " let D = Hd p be the Haar basis. Let " · " "p, and Then, the WCGA satisfies (1.6) with 1 1 2 . − h(p, d) = (d − 1)(cid:12)(cid:12)(cid:12) p(cid:12)(cid:12)(cid:12) φ(N) = (cid:4)c (1 + log N)p′h(p,d) N(cid:5), for some c = c(p, d, τ ) > 0. Moreover, suppose that (1.20) " "f − Gψ(N )f " "p ≤ C σN (f ), ∀ N ≥ 1, f ∈ X. Then ψ(N) ≥ c′ (1 + log N)p′h(p,d) N, for some c′ > 0. 6 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV REMARK 1.21. In the case 2 < p < ∞, it is known that an analog of (1.20), with (cid:129) · (cid:129)p replaced by the standard k · kLp, holds with ψ(N) ≤ c N 2/p′ ; see [19, Example 4]. However, we do not know whether this power is optimal. The proof of Theorem 1.19, presented in §6, is obtained as the special case q = 2 of a more general statement for a class of sequence spaces fp,q. These are defined by norms as in (1.18), but with the square function Sf replaced by an ℓq-function Sqf ; see §5 below for precise statements. The spaces ℓp(ℓq) and fp,q are discrete analogs of the Besov and Triebel-Lizorkin spaces, so our results can be transferred as well to these settings, see Remarks 6.1 and 6.2 below. Finally, in the last part of the paper we discuss the significance of property A3 for suitable classes of bases D = Ψ = {ψn}n≥1 in X = Lp. As a consequence of our results we shall obtain the following. THEOREM 1.22. Let X = Lp[0, 1], 1 < p < ∞, and let D = Ψ be a greedy basis with respect to the TGA. Let (1.23) α(p) = 1 if 1 < p ≤ 2, and α(p) = 2/p′ if 2 < p < ∞. Then, there exists a universal constant C ≥ 1 such that for some c = c(p, τ ) > 0. (cid:13)(cid:13)f − G cN α(p)f(cid:13)(cid:13)p ≤ C σN (f ), ∀ N ≥ 1, f ∈ Lp, In particular, when 1 < p ≤ 2, greedy bases with respect to the TGA are almost greedy with respect to the WCGA (in the sense of Definition 1.1). For p > 2, however, we do not know whether the power α(p) = 2/p′ may be improved. These resuls are discussed in §7, together with various additional statements under weaker assump- tions in Ψ (such as almost-greedy, quasi-greedy, etc...), where the best exponents r in property A3 are found for each class of bases. We remark that, in the case of X = Lp with p > 2, we do not know whether, for some basis (or dictionary) D, the Lebesgue inequality for the WCGA may hold with φ(N) = O(N). 2. Preliminaries 2.1. Norming functionals and distance to subspaces. Let (X, k · k) be a Banach space over K = R or C. Let ρ(t) be the associated modulus of smoothness, given in (1.5). Recall that k · k is called uniformly smooth if ρ(t) = o(t) as t → 0. Recall also from (1.3) that Ff ∈ X∗ denotes the norming functional of a vector f ∈ X \ {0}. In all our proofs below, the functional ρ(t) will only appear via the following inequality. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 7 PROPOSITION 2.1. For all f, g ∈ X with kf k = kgk = 1 it holds (2.2) 0 ≤ kf + tgk − kf k − tℜ[Ff (g)] ≤ 2ρ(t), t ∈ R. In particular, if ρ(t) = o(t), then ℜ[Ff (·)] is the Fr´echet derivative of k · k at f . Proof. The left inequality in (2.2) follows easily from (2.3) kf + tgk ≥ Ff (f + tg) ≥ ℜ[Ff (f + tg)] = kf k + t ℜ[Ff (g)], t ∈ R. Define σ(t, f, g) := kf + tgk + kf − tgk − 2kf k ≥ 0, ∀ t ∈ R. Then, using (2.3) with t replaced by −t we have kf + tgk = 2kf k − kf − tgk + σ(t, f, g) ≤ 2kf k − kf k + tℜ[Ff (g)] + σ(t, f, g), and therefore kf + tgk − kf k − tℜ[Ff (g)] ≤ σ(t, f, g) ≤ 2ρ(t), t ∈ R. This establishes the upper bound in (2.2). (cid:3) By homogeneity one deduces COROLLARY 2.4. For every (non-null) f, g ∈ X with kgk = 1 it holds (2.5) 0 ≤ kf + tgk − kf k − tℜ[Ff (g)] ≤ 2kf kρ(t/kf k), t ∈ R. An interesting application of (2.5) gives the following result. PROPOSITION 2.6. Suppose that 2ρ(t) ≤ γ tq, for some γ > 0 and q > 1. Then, there exists c1 = c1(γ, q) > 0 such that for all f, ϕ ∈ X \ {0} with kϕk = 1. dist(f, [ϕ]) ≤ kf k(cid:0)1 − c1Ff (ϕ)q′(cid:1), Proof. Let ν = sign Ff (ϕ) and λ > 0. Then Corollary 2.4 implies that kf − λνϕk ≤ kf k − λℜ[Ff (νϕ)] + γkf k(λ/kf k)q (2.7) (2.8) = kf k(cid:16)1 − λ kf kFf (ϕ) + γ( λ kf k )q(cid:17). Now, it is easily checked that the 1-dimensional function h(λ) = γλq − λF reaches a minimum at λ1 = [F/(γq)]q′−1, and that h(λ) = −c1F q′ , with c1 = min λ>0 1 q′(γq)q′−1 . So, minimizing over λ > 0 in (2.8) one obtains (2.7). (cid:3) Thus, recalling Definition 1.12, we recover the following assertion from §1. 8 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV COROLLARY 2.9. Suppose that ρ(t) ≤ γ tq, for some γ > 0 and q > 1. Then, for every dictionary D, the pair (X, D) has the property D(s, c1) with s = q′ and some c1 = c1(γ, q) > 0. REMARK 2.10. Observe that ρ(t) ≤ γ tq can only hold for q ≤ 2, and hence q′ ≥ 2. However, property D(s, c1) may hold in some cases with s close to 1. For instance, if X = ℓp with p > 2 and D is the canonical basis, then it is not hard to verify that (X, D) has property D(s, c1) with s = p′ and c1 = 1/p; see Proposition 4.12 below for a more general class of such examples. REMARK 2.11. Observe that the property D(s, c1) can also be written as One may ask whether this property could hold for some non-uniformly smooth Banach space (X, k · k) and some dictionary D. , ∀ kf k = 1, ϕ ∈ D. ≤ c−1 1 (cid:16)1 − dist(f, [ϕ])(cid:17)1/s (cid:12)(cid:12)(cid:12) Ff (ϕ)(cid:12)(cid:12)(cid:12) Below we shall also use the following known lemma. LEMMA 2.12. Assume that the norm k · k is Gateaux differentiable in X. Consider a vector f ∈ X, a finite dimensional subspace Y ⊂ X, and an element g ∈ Y such that kf − gk = dist(f, Y) > 0. Then (2.13) Ff −g(h) = 0, ∀ h ∈ Y. Proof. Let h ∈ Y and let ν = sign Ff −g(h). Then, for all λ > 0 we have dist(f, Y) ≤ kf − g − λνhk ≤ kf − gk − λℜ[Ff −g(νh)] + o(λ) = dist(f, Y) − λFf −g(h) + o(λ). Thus, Ff −g(h) ≤ o(λ)/λ, which letting λ ց 0 implies that Ff −g(h) = 0. (cid:3) REMARK 2.14. As a consequence of Lemma 2.12, the vectors ϕi1, ϕi2, ... chosen by the WCGA are always linearly independent. 2.2. Properties A2 and A3 for bases. Suppose that D = Ψ = {ψn}∞ n=1 is a (normal- ized) Schauder basis in X. We denote its dual system by Ψ∗ = {ψ∗ n}n≥1. We briefly discuss the meaning of properties A2 and A3 in this case, and relate them with more familiar concepts from the theory of thresholding greedy algorithms. In the first result we use the standard notation SA(f ) :=Xn∈A ψ∗ n(f )ψn, f ∈ X, for each finite set A ⊂ N. The next lemma is immediate from the definitions. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 9 LEMMA 2.15. Let Ψ be a (normalized) Schauder basis in X. Then, for each N ≥ 1, ΣN satisfies the property A2 with parameter (2.16) U = kN := sup A≤N kSAk. In particular, (X, Ψ) has the property A2(U) if and only if Ψ is an unconditional basis, with suppression unconditionality constant U. The second result relates property A3 with the right-democracy function of the dual basis Ψ∗. We use the convenient notation when A ⊂ N is a finite set, and ε = (εn)∞ n=1 ⊂ K is such that εn = 1 for all n. 1 ∗ εA :=Xn∈A εnψ∗ n, LEMMA 2.17. Let Ψ be a (normalized) Schauder basis in X. Then (X, Ψ) satisfies property A3(r, V ) if and only if (2.18) for all finite A ⊂ N, and all ε = (εn)∞ (cid:13)(cid:13)1 ∗ εA(cid:13)(cid:13)X∗ ≤ V Ar n=1 ⊂ K with εn = 1. Proof. Assume first (2.18), and take sets A ⊂ B and scalars an ∈ K. If εn = sign an then Xn∈A ∗ an = 1 εA(cid:0)Pn∈B anψn(cid:1) ≤ k1 ∗ εAkX∗(cid:13)(cid:13)Pn∈B anψn(cid:13)(cid:13) ≤ V Ar(cid:13)(cid:13)Pn∈B anψn(cid:13)(cid:13). If f ∈ X has a finite Conversely, assume property A3(r, V ), and pick A and ε. expansion with respect to Ψ we have ∗ 1 εA(f ) ≤Xn∈A ψ∗ n(f ) ≤ V Ar(cid:13)(cid:13) εAkX∗ ≤ V Ar, and (2.18) holds. ∗ Therefore, k1 ∞ Xn=1 ψ∗ n(f )ψn(cid:13)(cid:13) = V Ar kf k. (cid:3) 3. Proof of Theorem 1.14 We use the definition of WCGA given in §1. We shall follow closely the arguments given in [19, §3]; see also [21, §8.7]. As in that paper, the proof is split into two main steps. 3.1. The iteration inequality. In the first step we prove the following variant of [19, Theorem 2.3]; see also [21, Theorem 8.7.12]. The proof makes use of the property D(s, c1), and has been rewritten to simplify some steps from [19]. THEOREM 3.1. Let K ≥ 1 and assume that (i) (X, D) has property D(s, c1), for some s > 1 (ii) ΣK has property A3(r, V ), for some r ≤ 1. 10 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV Then, for all f ∈ X, Φ ∈ ΣK and all m, M ≥ 0 it holds (3.2) kfm+M k ≤ e− c2 M Krs kfmk + 2 kΦ − f k, where c2 = c(s) c1 τ s V −s, for some c(s) > 0. Proof. The case fm+M = 0 is trivial, so we assume kfm+M k > 0. If n ∈ [m, m + M) then also kfnk > 0. From the definition of the algorithm and property D(s, c1), (3.3) kfn+1k ≤ dist(fn, [ϕin+1]) ≤ kfnk(cid:16)1 − c1Ffn(ϕin+1)s(cid:17). When Φ =PT aiϕi ∈ ΣK, we use the notation aiϕi, for each A ⊂ T. ΦA :=XA We also denote (3.4) Γn := supp Gn(f ) = {i1, . . . , in}, and Tn = T \ Γn. By Lemma 2.12 and the definition of the algorithm we have (3.5) By property A3, τ Ffn(ΦA) = τ Ffn(ΦA∩Tn) ≤ XA∩Tn ai Ffn(ϕin+1). XA∩Tn ai ≤ V Ar kΦ − Gnf k ≤ V Ar(cid:0)kΦ − f k + kfnk(cid:1). Therefore, inserting these estimates into (3.3) we obtain (3.6) kfn+1k ≤ kfnkh1 − c1(cid:16) τ Ffn(ΦA) V Ar (kΦ − f k + kfnk)(cid:17)si, for all (non-empty) sets A ⊂ T . One wants to pick the best possible set A in (3.6). For later estimates we shall allow sets A ⊂ Tk, for a fixed k ∈ [0, m], and define B = Tk \ A. In this theorem it will suffice to consider k = 0 and A = T (so B = ∅), but we keep on with the general case. Since Γk ⊂ Γn, using Lemma 2.12 we have Ffn(ΦA) = Ffn(ΦA + ΦΓk∩A − Gnf ) = Ffn(Φ − ΦB − f + fn) ≥ kfnk − kΦ − f k − kΦBk. Inserting the above inequalities into (3.6), and denoting c3 = c1(τ /(2V ))s, we obtain kfn+1k ≤ kfnk(cid:16)1 − c3 ≤ kfnk(cid:16)1 − c3 [kfnk − kΦ − f k − kΦBk]s + [2−1Ar (kΦ − f k + kfnk)]s(cid:17) (cid:17), [kfnk − kΦ − f k − kΦBk]s + [Ar kfnk]s LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 11 where the last inequality is trivial when kfnk ≤ kΦ−f k+kΦBk, and in the complement case follows from kfnk ≥ kΦ − f k. We now use the elementary inequality (1 − u)s + ≥ δs(1 − 2u), u > 0, which holds for some δs > 0. Applying this with u = (kΦ − f k + kΦBk)/kfnk, and letting c2 = c3δs, we obtain (3.7) kfn+1k ≤ kfnk(cid:16)1 − c2 Ars(cid:17) + 2 c2 Ars (kΦ − f k + kΦBk). This expression can be rewritten as Denoting β = c2/Ars, one also has kfn+1k − 2(kΦ − f k + kΦBk) ≤ (cid:16)kfnk − 2 (kΦ − f k + kΦBk)(cid:17)(cid:16)1 − (cid:16)kfn+1k − 2(kΦ − f k + kΦBk)(cid:17)+ Ars(cid:17). ≤ (1 − β)(cid:16)kfnk − 2 (kΦ − f k + kΦBk)(cid:17)+ after checking the trivial negative cases. One can now iterate this inequality for m ≤ n < m + M to obtain c2 , kfm+M k − 2(kΦ − f k + kΦBk) ≤ (1 − β)M(cid:16)kfmk − 2 (kΦ − f k + kΦBk)(cid:17)+ ≤ (1 − β)M kfmk, and therefore (3.8) kfm+M k ≤ (1 − β)M kfmk + 2(kΦ − f k + kΦBk). Since 1 − β ≤ e−β and β ≥ c2/K rs, letting B = ∅ in (3.8) gives (3.2). (cid:3) REMARK 3.9. The above proof actually shows the validity of the following more general inequality (3.10) kfm+M k ≤ e−c2M /Ars kfmk + 2(kΦ − f k + kΦBk), for any sets A ⊂ Tk and B = Tk \ A, and any 0 ≤ k ≤ m. This is the analog of the corresponding expression in [19, (3.7)]; see also [21, (8.7.22)]. Observe that (3.10) also makes sense for A = ∅, in which case the inequality follows trivially from kfkk ≤ kf − ΦΓkk = kf − Φ + ΦBk ≤ kΦ − f k + kΦBk. 3.2. Proof of Theorem 1.14. The second part of the proof of Theorem 1.14 is then carried exactly as in the paper [19]; see also [21, pp. 433-440]. It is based on appropriate choices of sets Aj and Bj in inequality (3.10), at various inductive steps. We only remark that the quantities Ajrq′ that appear in [19] can all be replaced by Ajrs, according to inequality (3.10), as it is only via this inequality that the uniform smoothness of the norm (or the property D(s, c1)) is used in this proof. All the arguments in [19] can then be applied verbatim, so we do not write down the details. 12 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV (cid:3) 4. The WCGA in the spaces ℓp(ℓq) In this section we illustrate the performance of the WCGA in the space X = ℓp(ℓq), when 1 < p, q < ∞. This is the set of all sequences x = (xj,k)∞ j,k=1 ⊂ K such that (4.1) kxk :=h ∞ Xj=1(cid:16) ∞ Xk=1 xj,kq(cid:17) p qi 1 p < ∞. Throughout this section we let D = {ej,k}j,k≥1, the canonical basis in the space X. 4.1. Norming functionals in ℓp(ℓq). It will be convenient to use the following no- tation ∆j(x) :=(cid:16) ∞ Xk=1 1 q , xj,kq(cid:17) so that kxk = (Pj≥1 ∆j(x)p)1/p. LEMMA 4.2. When x ∈ ℓp(ℓq) \ {0}, the norming functional Fx is given by (4.3) Fx(y) = 1 kxkp−1 ∆j(x)p−q ∞ Xj=1 ∞ Xk=1 xj,kq−2¯xj,kyj,k. Proof. It is straightforward to check that Fx(x) = kxk. So we only need to verify that Fx(y) ≤ kyk. This follows from a double use of Holder's inequality Fx(y) ≤ kxk−(p−1)Xj = kxk−(p−1)Xj ≤ kxk−(p−1)(cid:16)Xj ∆j(x)p−q(cid:16)Xk ∆j(x)p−1 ∆j(y) xj,k(q−1)q′(cid:17)1/q′ ∆j(y) ∆j(x)(p−1)p′(cid:17)1/p′ kyk = kyk. (cid:3) We shall apply the functionals Fx to the elements ej,k of the canonical basis. COROLLARY 4.4. If x ∈ ℓp(ℓq) \ {0} and j, k ∈ N, then (4.5) As special cases notice that ∆j(x)p−q xj,kq−1 kxkp−1 . (cid:12)(cid:12)Fx(ej,k)(cid:12)(cid:12) = • if supp x ⊂ {j0} × N, then Fx(ej,k) = δj,j0 xj0,kq−1/kxkq−1. • if supp x ∩ ({j} × N) ≤ 1, for all j, then Fx(ej,k) = xj,kp−1/kxkp−1. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 13 In particular, for vectors x in these two cases, the WCGA (with τ = 1) coincides with the usual TGA; that is, after N steps of the algorithm it picks the N largest coefficients, and the remainder xN is the projection of x on the subspace spanned by the remaining basis vectors. 4.2. Performance of WCGA in ℓp(ℓq). Here we compute the relevant parameters in (i)-(iii) of Theorems 1.9 and 1.14. Since the canonical basis is unconditional we see that property A2 holds with U = 1. Property A3 is given by the next lemma, which follows immediately from the inclusion X ֒→ ℓmax{p,q}. LEMMA 4.6. For every finite set A ⊂ N × N and every x ∈ X we have xj,k ≤ Ar kxk, with r = max{ 1 p′ , 1 q′ }. X(j,k)∈A In particular, property A3 holds with parameters r and V = 1. The power of uniform smoothness for the standard norm (4.1) in ℓp(ℓq) can be estimated as follows; see [5, Proposition 17] or [13, Theorem 3.5]. LEMMA 4.7. If 1 < p, q < ∞, then ρ(t) ≤ γ tσ with σ = min{p, q, 2}. If at this point one applies Theorem 1.9, a simple computation easily leads to the following. COROLLARY 4.8. Let β = max{ p′ (4.9) p′ }. Assume also that q′ , q′ min{p, q} ≤ 2. Then there are constants C1 and C2 such that (4.10) kx − GC1 N β (x)k ≤ C2 σN (x), ∀ N ≥ 1, x ∈ X. REMARK 4.11. If one attempts to use Theorem 1.9 in the case min{p, q} > 2, then σ = 2 in Lemma 4.7, and one only obtains kx − G C1 N β1 (x)k ≤ C2 σN (x) with β1 = max{ 2 p′ , 2 q′ } > max{ p′ q′ , q′ p′ }. As we shall see below, this result can be improved using instead Theorem 1.14. 4.3. An improvement of the previous bound. We begin by establishing the va- lidity of property D(s, c1). PROPOSITION 4.12. Let 1 < p, q < ∞, and s = max{p′, q′}. Then there exists c1 = c1(p, q) > 0 such that, (4.13) dist(x, [ej,k]) ≤ kxk(cid:0)1 − c1Fx(ej,k)s(cid:1), for all (j, k) ∈ N × N and all x ∈ X = ℓp(ℓq) \ {0}. In particular, the canonical basis in ℓp(ℓq) has the property D(s, c1) with s = max{p′, q′}. 14 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV Proof. We may assume that kxk = 1 and that j = k = 1. Using the notation ∆j = ∆j(x) = (Pk xj,kq)1/q, we can write j +(cid:16)∆q dist(x, [e1,1])p = (4.14) ∆p Xj=2 ∞ 1 − x1,1q(cid:17) p q = 1 − ∆p 1 + ∆p 1(cid:16)1 − x1,1 ∆1 p q . q(cid:17) We shall often make use of the following elementary inequality. LEMMA 4.15. Let α > 0. Then (cid:0)1 − u(cid:1)α ≤ 1 − min{1, α}u, u ∈ [0, 1]. In the last expression of (4.14) we apply Lemma 4.15, with u = x1,1/∆1q and α = p/q, so if c = min{1, p/q} we obtain (4.16) dist(x, [e1,1])p ≤ 1 − c ∆p 1 x1,1 ∆1 q = 1 − c Fx(e1,1) x1,1, the last equality due to (4.5). Letting s = max{p′, q′}, note that min{p − 1, q − 1} = s′ − 1. So, using that both ∆1 and x1,1/∆1 are ≤ 1, it follows that p−1 Fx(e1,1)s−1 = ∆ s′−1 1 Therefore, we have shown q−1 s′ −1 ≤ x1,1. x1,1 ∆1(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12) dist(x, [e1,1])p ≤ 1 − cFx(e1,1)s. Finally, using again Lemma 4.15 we obtain (4.13) with c1 = c/p = min{ 1 p, 1 q }. (cid:3) Thus, Theorem 1.14 will produce the following improvement over Corollary 4.8. COROLLARY 4.17. Let 1 < p, q < ∞, and let β = max{ p′ constants C1 and C2 such that q′ , q′ p′ }. Then there are (4.18) kx − GC1N β (x)k ≤ C2σN (x), ∀ N ≥ 1, x ∈ X. This proves the first assertion in Theorem 1.16. 4.4. Optimality of the bound in (4.18). We now prove the last assertion in Theo- rem 1.16. The lower bounds will be obtained by testing with suitable examples. For simplicity we assume τ = 1 and C = 1. PROPOSITION 4.19. Let 1 < p, q < ∞. Suppose that (4.20) kx − Gψ(N )xk ≤ σN (x), ∀ N ≥ 1, x ∈ X. Then, there exists cp,q > 0 such that (4.21) ψ(N) ≥ cp,q N β, with β = max{ p′ q′ , q′ p′ }. Proof. Pick x = 1A + nα 1B, where • A = m and A ∩ ({j} × N) = 1, for 1 ≤ j ≤ m, and 0 otherwise. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 15 • B = n and B ⊂ {m + 1} × N. • α = p′( 1 q′ − 1 p′ ). Observe that with this choice we have (4.22) Fx(ej,k) = kxk−(p−1), (j, k) ∈ A ∪ B. Indeed, if (j, k) ∈ A then xj,k = ∆j(x) = 1 and (4.22) follows from (4.5). If (j, k) ∈ B then j = m + 1 and ∆m+1(x)p−qxm+1,kq−1 = nα(p−q) n p−q q nα(q−1) = 1. In the case p′ ≥ q′, for each 1 ≤ ℓ ≤ m we shall pick Gℓx = 1Aℓ, where Aℓ−1 ⊂ Aℓ ⊂ A with Aℓ = ℓ. Observe that this is possible because the equality in (4.22) continues to hold when x is replaced by each remainder xℓ = x − Gℓx. So if we let m = ψ(n), then (4.20) implies nαk1Bk = kx − Gmxk ≤ σn(x) ≤ kx − nα 1Bk = k1Ak = m1/p. Since k1Bk = n1/q, this implies (4.23) ψ(n) = m ≥ np(α+ 1 q ) = n p′ q′ , where in the last equality we use the expression of α. In the case q′ ≥ p′ we use the same example, except that we take B = 2n. For 1Bℓ where Bℓ−1 ⊂ Bℓ ⊂ B with Bℓ = ℓ. each 1 ≤ ℓ ≤ n, this time we pick Gℓx = nα To justify this choice at each step, observe that, if (j, k) ∈ B \ Bℓ then, kxℓkp−1Fxℓ(ej,k) = ∆m+1(xℓ)p−q nα(q−1) = nα(p−1) (2n − ℓ) p−q q = (cid:16) 2n − ℓ n (cid:17) p q −1 ≥ 1, while for (j, k) ∈ A we have kxℓkp−1Fxℓ(ej,k) = 1. So, if we let n = ψ(m), from (4.20) we have m1/p = k1Ak ≤ kx − Gnxk ≤ σm(x) ≤ kx − 1Ak = nαk1Bk = 2 1 q nα+ 1 q , which implies (4.24) ψ(m) = n ≥ cp,q m 1 p(α+1/q) = cp,q m q′ p′ . Combining (4.23) and (4.24) one obtains (4.21). PROOF of Theorem 1.16: Combine Corollary 4.17 and Proposition 4.19. (cid:3) (cid:3) 16 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV 4.5. WCGA vs TGA in ℓp(ℓq). Let Gn(x) denote the usual thresholding greedy algorithm (TGA) applied to x, with regard to the canonical basis in X = ℓp(ℓq). The following result is a consequence of [12, Theorem 5]. PROPOSITION 4.25. Let 1 < p, q < ∞. Then there are constants C1, C2 such that (4.26) kx − GC1N b(x)k ≤ C2 σN (x), ∀ N ≥ 1, x ∈ X, with b = max{ p q , q p}. REMARK 4.27. The exponent b = max{ p [12, Theorem 3.2]. q , q p } in (4.26) is best possible, in view of So one may pose the question of when WCGA performs better than TGA, in the sense that the power β in Corollary 4.17 is smaller than the power b in (4.26). As- suming p > q, this is equivalent to β = q′ p′ ≤ p q = b ⇐⇒ qq′ ≤ pp′ ⇐⇒ q ≥ p′. If p < q, then β ≤ b iff p ≥ q′. Thus, overall the WCGA performs better than the TGA, in the sense that β ≤ b, if and only if q ≥ p′ (or q = p). 1 q TGA N p q WCGA q′ p′ N TGA q p N WCGA N p′ q′ 1 p Figure 4.1. Number of iterations (modulo constants) of TGA or WCGA to reach σN (x) in ℓp(ℓq). 5. The WCGA in the spaces fp,q In this section we consider a class of sequence spaces fp,q related with the family of Triebel-Lizorkin spaces. In the next section we shall specialize to the case q = 2, and deduce the results for the Haar system in (Lp, (cid:129) · (cid:129)p) asserted in Theorem 1.19. Throughout this section we make use of the following notation. We fix d ≥ 1 and let Rd be the set of all dyadic rectangles I ⊂ [0, 1]d. That is, I = I1 × . . . × Id, with each Ii = 2−ji(ki + [0, 1)), for some ji, ki ∈ N0 and 0 ≤ ki < 2ji, when i = 1, . . . , d. When d = 1 we will write R1 = D, so that Rd = D × . . . × D. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 17 For fixed d ≥ 1 and 1 < p, q < ∞, we consider the space X = fp,q = fp,q(Rd), defined as the set of all sequences x = (xI)I∈Rd such that (5.1) where 1I,p = I−1/p use the notation kxk = kxkfp,q :=(cid:13)(cid:13)(cid:13)(cid:0) XI∈Rd(cid:12)(cid:12)xI 1I,p(cid:12)(cid:12) q(cid:1)1/q(cid:13)(cid:13)(cid:13)Lp([0,1]d) , 1I is the Lp-normalized characteristic function of I. We shall often (Sqx)(u) :=(cid:0) XI∈Rd(cid:12)(cid:12)xI 1I,p(u)(cid:12)(cid:12) q(cid:1)1/q, u ∈ [0, 1]d. Throughout this section, D = {eI}I∈Rd will be the canonical basis in fp,q. 5.1. Norming funcionals in fp,q. LEMMA 5.2. When x ∈ fp,q \ {0}, the norming functional Fx is given by (5.3) Fx(y) = 1 kxkp−1 Z [(Sqx)(u)]p−q(cid:16) XI∈Rd xIq−2 xI yI 1I,p(u)q(cid:17)du . Proof. It is easy to see that Fx(x) = kxk, so we only need to verify that Fx(y) ≤ kyk. This follows easily from a double use of Holder's inequality kxkp−1Fx(y) ≤ Z (Sqx)p−qhXI q′ hXI = Z (Sqx)p−1 (Sqy) du ≤ kxkp−1 kyk . xI(q−1)q′ q I,pi 1 1 yIq 1 1 q du q I,pi (cid:3) If we specialize Fx to the elements eI of the canonical basis we obtain. COROLLARY 5.4. If x ∈ fp,q \ {0} and I ∈ Rd, then (5.5) xIq−1 kxkp−1 1 Iq/pZI Sqxp−q. (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) = REMARK 5.6. As a special case, if supp x consists of pairwise disjoint I's, one has So, in this case WCGA and TGA coincide. (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) = xIp−1/kxkp−1. 5.2. Distance to subspaces. We study the property D(s, c1) in the spaces fp,q(Rd). This will make unnecessary to compute the modulus of smoothness for k · kfp,q (which seems not to appear in the literature). LEMMA 5.7. Let 1 < p, q < ∞, and let s = s(p, q) = max{p′, q′}. Then, for all x ∈ fp,q \ {0} it holds (5.8) s(cid:17), dist(x, [eI]) ≤ kxk(cid:16)1 − cp,q(cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) ∀ I ∈ Rd, 18 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV with cp,q = min{1/p, 1/q}. In particular, the canonical basis in fp,q has the property D(s, c1) with s = max{p′, q′}. Proof. We may assume that kxk = 1. Then, dist(x, [eI])p = kx − xIeIkp =Z (Sqx)(u)ph1 − xI 1I,p(u)q (Sqx)(u)q ip/q du, where it is understood that the integral is taken over the set {u : Setting c := min{1, p/q} and applying Lemma 4.15 we obtain (Sqx)(u) 6= 0}. (5.9) dist(x, [eI])p ≤Z (Sqx)ph1 − c(cid:12)(cid:12)(cid:12) xI 1I,p Sqx (cid:12)(cid:12)(cid:12) qi du = 1 − c xI(cid:12)(cid:12)Fx(eI)(cid:12)(cid:12), using in the last step Corollary 5.4 (and kxk = 1). We now separate two cases: a) Case 1 < p ≤ q. Since xI 1I,p(u) ≤ (Sqx)(u) and p − q ≤ 0, we have (Sqx)(u)p−q ≤ xI 1I,p(u)p−q . Integration over I gives Z (Sqx)p−q Then, using (5.5) we see that 1 q I,p ≤ xIp−q Z 1 p I,p = xIp−q. Fx(eI) = xIq−1 Z (Sqx)p−q q I,p ≤ xIp−1. 1 Thus, we have obtained a lower bound for xI, which inserted into (5.9) gives Finally, since c = p/q, a last use of Lemma 4.15 (with α = 1/p < 1) gives dist(x, [eI])p ≤ 1 − c (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) p−1 = 1 − c (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) p′ . 1+ 1 b) Case q < p < ∞. Holder's inequality with exponents (p/(p − q), p/q) gives (5.10) Then, from (5.5), p′ . dist(x, [eI]) ≤ 1 − q−1 (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) p hZ 1 I,pi I,p ≤ hZ (Sqx)pi p−q p q 1 Z (Sqx)p−q q p = 1. Fx(eI) ≥ Fx(eI) xIq−1 = R (Sqx)p−q dist(x, [eI])p ≤ 1 − (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) q I,p 1 1+ 1 q−1 = 1 − (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) q′ . Inserted into (5.9) (and using c = 1) it gives Hence, a last use of Lemma 4.15 (with α = 1/p < 1) implies (5.11) dist(x, [eI]) ≤ 1 − p−1 (cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) q′ . (cid:3) LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 19 REMARK 5.12. Testing with explicit examples, it is possible to show that the power s = max{p′, q′} in (5.8) cannot be replaced by any smaller number. 5.3. Property A3. Here, ΣN is the set of N-term combinations from the canonical basis of X = fp,q(Rd). We define the exponent h = h(p, q; d) = (d − 1)(cid:16)1 p − . 1 q(cid:17)+ LEMMA 5.13. Let 1 < p, q < ∞ and N ≥ 1. Then ΣN satisfies property A3(r, V ) with V = c(log N)h and r = 1/p′, for some constant c = c(p, q, d) > 0. That is (5.14) XI∈A xI ≤ c (1 + log N)(d−1)( 1 p − 1 q )+ A1/p′ xI eIk. kXI∈B for all sets A ⊂ B ⊂ Rd with A ≤ N and B finite. Proof. In view of Lemma 2.17 it suffices to estimate Since (fp,q)∗ = fp′,q′, then (5.14) reduces to show that (5.15) ≤ c (1 + log N)(d−1)( 1 q − 1 p )+ N 1 p . eI(cid:13)(cid:13)(fp,q)∗ . sup A≤N(cid:13)(cid:13)Xn∈A eI(cid:13)(cid:13)fp,q sup A≤N(cid:13)(cid:13)Xn∈A When q = 2 this was proved in [23, Proposition 10], see also [15, Theorem A]. A small modification of those proofs gives also the case q 6= 2. For completeness, we sketch the arguments in Appendix 1 below. (cid:3) 5.4. The WCGA in fp,q. A direct application of Theorem 1.14, with the parameters obtained in Lemmas 5.7 and 5.13, gives the following. THEOREM 5.16. Let X = fp,q(Rd), 1 < p, q < ∞, and let D be the canonical basis. Define h(p, q; d) = (d − 1)(cid:16)1 p − , 1 q(cid:17)+ Then, the WCGA satisfies (1.6) with and α(p, q) =( 1 q′/p′ if p ≤ q q ≤ p. if (5.17) for some c = c(p, q, d, τ ) > 0. φ(N) = (cid:4)c (1 + log N)p′h(p,q;d) N α(p,q)(cid:5), REMARK 5.18. As in Remark 4.11 above, if in the case min{p, q} > 2 we had used Theorem 1.9 rather than Theorem 1.14, then one would have obtained a power 2/p′, which is worse than α(p, q) (assuming that the estimate for the modulus of smoothness in Lemma 4.7 also applies when X = fp,q). 20 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV 5.5. A lower bound for the WCGA in fp,q when p ≤ q. We show in this sub- section that, when p ≤ q, the function φ(N) = ⌊c (1 + log N)p′(d−1)( 1 q ) N⌋ in (5.17) cannot be replaced by a slower growing one. p − 1 THEOREM 5.19. In the conditions of Theorem 5.16, suppose that p ≤ q and that (5.20) ∀ N ≥ 1, x ∈ fp,q(Rd). Then ψ(N) ≥ c′ (1 + log N)p′(d−1)( 1 q ) N, for some c′ > 0. (cid:13)(cid:13)x − Gψ(N )x(cid:13)(cid:13) ≤ C σN (x), p − 1 Proof. Fix n ∈ N and ~m = (m1, . . . , md) ∈ Nd, and let m := m1 + . . . + md. Consider the following sets An =nI ∈ Rd : I ⊂ [0, 1/2]d and I = 2−no and B ~m =nI1 × . . . × Id ∈ Rd : Ii ⊂ [1/2, 1] and Ii = 2−mi, i = 1, . . . , do. Observe that (5.21) An ≈ nd−1 2n ≈ n d−1 q 2n/p, u ∈ (0, 1/2)d, and (cid:16) XI∈An 1 q I,p(u)(cid:17)1/q I,p(u)(cid:17)1/q q 1 B ~m = 2m−1 and (cid:16) XI∈B ~m while (5.22) ≈ 2m/p, u ∈ (1/2, 1)d, We build a vector x = xAn + xB ~m := a XI∈An eI + b XI∈B ~m eI, for suitable a, b > 0 to be chosen later. Using the expression for the fp,q-norming functionals in (5.5), and the estimates in (5.21), (5.22), one sees that (5.23) kxkp−1(cid:12)(cid:12)Fx(eI)(cid:12)(cid:12) = xIq−1 ZI We now pick b such that (Sqx)p−q 1 q I,p ≈ ( ap−1n bp−1, (d−1)(p−q) q , if if I ∈ An I ∈ B ~m. (5.24) bp−1 ≈ ap−1 n(d−1)(p−q)/q. The constants can be adjusted so that the WCGA always chooses Gk(x) = b k Xj=1 eIj , when k ≤ B ~m, for some enumeration I1, I2, . . . of B ~m. At this point one should notice that (5.23) continues to hold with x replaced by each remainder x − Gk(x), as long as k < B ~m (so one can still pick elements I from B ~m \ {I1, . . . , Ik}). LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 21 Finally, given N ≫ 1, we select the largest n and the smallest m such that (5.25) An ≤ N and B ~m ≥ ψ(N). Then, N ≈ nd−12n and ψ(N) ≈ 2m. The assumption in (5.20) gives kxAnk ≤ kx − Gψ(N )(x)k ≤ C σN (x) ≤ C kxB ~mk. By (5.21) and (5.22) this implies a n d−1 q 2n/p ≈ kxAnk ≤ C kxB ~mk ≈ b 2m/p ≈ b ψ(N)1/p. Recalling (5.24) this amounts to ψ(N) & (a/b)p n (d−1)p q 2n ≈ n (d−1)p′ q′ 2n ≈ [log N](d−1)( p′ q′ −1) N, using in the last step the relation N ≈ nd−12n. This proves the theorem since the exponent in the power of the log can also be written as p′/q′ − 1 = p′( 1 p − 1 q ). (cid:3) 6. The WCGA for the d-variate Haar basis in Lp We recall the definition of the Haar system. Consider the 1-dimensional functions h0 = 1[0,1) and h = 1[0,1/2) − 1[1/2,1). For every dyadic interval I = 2−j(k + [0, 1)) in [0, 1], define hI,p(x) = 2j/ph(2jx − k), and h0,p(x) = h0(x). Then, if 1 < p < ∞, the system Hp = {h0, hI,p}I∈D is a (normalized) unconditional basis of Lp[0, 1]. With a slight abuse of notation we write Hp = {hI,p}I∈ ¯D with ¯D = D ∪ {0}. The d-variate Haar system is then Hd p = Hp × . . . × Hp = {HI,p}I∈ ¯Rd, with ¯Rd = ¯D × . . . × ¯D, that is, HI,p(x1, . . . , xd) = d Yi=1 hIi,p(xi), if I = I1 × . . . × Id ∈ ¯Rd. The system D = Hd p is an unconditional basis of Lp([0, 1]d), and moreover, for every function f =PI cI(f )HI,p ∈ Lp, the expression " "f " "p :=(cid:13)(cid:13)S(f )(cid:13)(cid:13)Lp, where S(f ) =(cid:16) XI∈ ¯Rd 1 2 , cI(f )HI,p2(cid:17) defines an equivalent norm in Lp([0, 1]d), provided 1 < p < ∞. Here, the coefficients are given by cI(f ) := h f , HI,p′ i, I ∈ ¯Rd. 22 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV Clearly, (Lp, (cid:129) · (cid:129)p) is related with the spaces fp,2 from the last section. More precisely, the coefficient map defines an isometric isomorphism (cid:0)Lp([0, 1]d), (cid:129)·(cid:129)p(cid:1) −→ fp,2( ¯Rd) f 7−→ (cid:0)cI(f )(cid:1)I∈ ¯Rd with the latter space defined as in §5, with the only minor change that Rd is replaced by ¯Rd (so a few additional terms appear in the norm (5.1)). This change does not affect the proofs, and all the results in §5 continue to hold for the spaces fp,q( ¯Rd). In particular, Theorem 1.19 from the introduction becomes a corollary of Theorems 5.16 and 5.19, in the special case q = 2. REMARK 6.1. One can use the previous isometric isomorphism to transfer the results for X = fp,q in §5 into results for (univariate) wavelet bases Ψ = {ψI} in the class of Triebel-Lizorkin spaces F r p,q. Indeed, one can define equivalent norms in the latter of the form kf kF r p,q :=(cid:13)(cid:13)(cid:13)(cid:0)XI cI(f )1I,pq(cid:1) , 1 q(cid:13)(cid:13)(cid:13)Lp where cI(f ) = h f , ψ∗ I i = I−r h f , ψI,p′ i, with ψI,p′ = ψI /kψIkp′. Just remark that in §5 one should modify the index set D acccording to the underlying space (say, over R rather than [0, 1]). In the d-variate case, the use of tensor wavelet bases, gives rise to an isometry between fp,q(Rd) and Triebel-Lizorkin classes with dominating mixed smoothness (sometimes denoted Sr p,qF ; see e.g. [22, Th. 1.12]). REMARK 6.2. In a similar fashion, one may transfer the results for X = ℓp(ℓq) in §4 into results for wavelet bases in the class of Besov spaces Br q,pB in the d-variate case), using the equivalent norms q,p (or Sr kf kBr q,p :=hXj (cid:16)Xk cI(j,k)(f )q(cid:17) , p qi1/p with I(j, k) = 2−j(k + [0, 1)), j, k ∈ Z. 7. Property A3 for classes of bases in Lp In this section we find optimal exponents for the property A3 in certain classes of bases of Lp, 1 < p < ∞. These classes (greedy, almost greedy, quasi-greedy,...) behave well with respect to the TGA, and we study the performance with respect to the WCGA. Results of this section complement the corresponding results from [19] (see also [21], Section 8.7.4). Below we shall use the following known lemma, which is valid for Lp over an arbi- K the subspace of K-valued functions trary (σ-finite) measure space. We denote by Lp in Lp. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 23 LEMMA 7.1. Let 1 < p < ∞, and let X = Lp = Lp such that C. Then, there exists γ = γ(p) > 0 ρLp(t) := sup kf kp=kgkp=1(cid:16)kf + tgkp + kf − tgkp − 2kf kp(cid:17)/2 ≤ γ tq, with q = min{p, 2}. ∀ t > 0, Proof. This result is well-known when X = Lp that in the complex-valued case one has R; see e.g. [14, Vol II, p. 63]. We remark ρLp C (t) = ρLp R (t) ≤ γ tq, ∀ t > 0 , since Lp C can be isometrically embedded into some Lp R; see [3, Example 4.6]. (cid:3) 7.1. Greedy bases. The univariate Haar system Hp is a greedy basis of Lp[0, 1], 1 < p < ∞, and it was shown in [19, Example 4] that a Lebesgue inequality holds with φ(N) = O(N) if 1 < p ≤ 2, and φ(N) = O(N 2/p′) if 2 ≤ p < ∞. Here we generalize this result to any greedy basis Ψ = {ψn}n≥1 of Lp, as stated in Theorem 1.22 from §1. We begin with a computation of the parameters for the A3 property. We assume that Ψ is normalized in Lp, and recall that a greedy basis is unconditional and democratic. LEMMA 7.2. If Ψ is a greedy basis in Lp, then A3(r, V ) holds with r = 1/p′ and some constant V > 0. Proof. Let Ψ∗ = {ψ∗ suffices to show that n}∞ n=1 be the dual system to Ψ in Lp′. In view of Lemma 2.17 it (7.3) (cid:13)(cid:13)Xn∈Λ ψ∗ n(cid:13)(cid:13)Lp′ ≤ C Λ1/p′ , for all finite sets Λ ⊂ N. We shall use a result from functional analysis [11, Theorem 4b] which says that for any (seminormalized) unconditional basis B = {bk}k≥1 of Lq, 1 < q < ∞, there is a subsequence kj, j = 1, 2, . . . , such that In particular, if B is unconditional and democratic, then necessarily (7.4) bkkq ≍ Λ1/q, ∞ Xj=1 1 q . αkj q(cid:17) ∞ Xj=1 (cid:13)(cid:13) αkj bkj(cid:13)(cid:13)Lq ≍(cid:16) kXk∈Λ with the constants of equivalency depending at most on B and q. Now, if Ψ is a greedy basis in Lp, 1 < p < ∞, then it was shown in [2, Theorem 5.1] that Ψ∗ is also a greedy basis in the dual space (Lp)∗ = Lp′. Therefore (7.3) is just a consequence of (7.4) with q = p′. (cid:3) 24 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV PROOF of Theorem 1.22: The proof is now a direct application of Theorem 1.9. Indeed, A2(U) holds because Ψ is unconditional, while A3(r = 1/p′, V ) was proved in the previous lemma. By Lemma 7.1 we also have ρ(t) ≤ γtq with q = min{p, 2}. Therefore, Theorem 1.9 implies (7.5) kf − G C(t,p,Ψ)mα(p)(f )kp ≤ Cσm(f, Ψ)p, with α(p) = rq′, which agrees with (1.23). (cid:3) 7.2. Tensor products of greedy bases. We now extend the first part of Theorem 1.19 from the multivariate Haar system to an arbitrary tensor product of univariate greedy bases. Let Ψ be a normalized basis in Lp([0, 1)). In the space Lp([0, 1)d) we define the system Ψd := Ψ × · · · × Ψ =nψn(x) = ψn1(x1) · · · ψnd(xd) : n = (n1, . . . , nd) ∈ Ndo. Clearly, if Ψ is unconditional, so is Ψd. Democracy, however, does not transfer, but one has the following result from [15] (see also [21, Ch. 8]). PROPOSITION 7.6. Let 1 < q < ∞ and let Ψ be a greedy basis for Lq[0, 1]. Then for any Λ with Λ = m ≥ 2, we have: for 2 ≤ q < ∞ C 1 q,d m1/q min n∈Λ and for 1 < q ≤ 2 cn ≤ kXn∈Λ cnψnkq ≤ C 2 q,d (log m)h(q,d) m1/q max n∈Λ cn, C 3 q,dm1/q(log m)−h(q,d) min n∈Λ where h(q, d) := (d − 1)1/2 − 1/q. cn ≤ kXn∈Λ cnψnkq ≤ C 4 q,dm1/q max n∈Λ cn Note that h(p, d) = h(p′, d). We now derive the property A3 for Ψd. LEMMA 7.7. Consider the system Ψd in Lp([0, 1]d) defined above, where Ψ is a uni- variate greedy basis in Lp[0, 1]. Then, for each N ≥ 1, the set ΣN (Ψd) satisfies property A3(r, V ) with r = 1/p′ and V = C(p, d) (1 + log N)(d−1)( 1 p − 1 2 )+ . Proof. By Lemma 2.17, it suffices to compute sup Λ≤N(cid:13)(cid:13)Xn∈Λ ψ∗ n(cid:13)(cid:13)(cid:13)Lp′ , n = ψ∗ n1 ⊗ . . . ⊗ ψ∗ where ψ∗ nd, n ∈ Nd, are the elements of the dual system (Ψd)∗. Since Ψ∗ is a univariate greedy basis (by [2, Thm 5.1]), we can apply Proposition 7.6 with q = p′, and obtain sup Λ≤N(cid:13)(cid:13)Xn∈Λ ψ∗ n(cid:13)(cid:13)(cid:13)Lp′ ≤ Cp,d (1 + log N)(d−1)( 1 p − 1 2 )+ N 1/p′ . LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 25 (cid:3) THEOREM 7.8. Let 1 < p < ∞. Consider the system Ψd in Lp([0, 1]d) defined above, where Ψ is a univariate greedy basis in Lp[0, 1]. Then, the WCGA in Lp([0, 1]d) applied to Ψd satisfies (7.9) with φ(m) = C(t, p, d, Ψ) (1 + log m)(d−1)p′( 1 p − 1 kf − Gφ(m)(f )kp ≤ Cσm(f )p, 2 )+ mα(p), and with α(p) as in (1.23). Proof. This is again a direct application of Theorem 1.9 and the previous lemma. (cid:3) 7.3. Nikol'skii ℓ1X property. When Ψ = {ψk}∞ k=1 is a basis in Lp := Lp([0, 1)d), 1 < p < ∞, then the A3 property essentially amounts to compute the following operator norms kSΛkp,A := kSΛkLp→A := sup f 6=0 kSΛ(f )kA/kf kp , where we denote, for each finite set Λ ⊂ N, SΛ(f ) := SΛ(f, Ψ) :=Xk∈Λ ck(f )ψk, kSΛ(f )kA :=Xk∈Λ ck(f ), and as usual ck = h f , ψ∗ k i. Then A3(r, V ) holds if and only if kSΛkp,A ≤ V Λr. The notation refers to A = {f ∈ L1 : Pk∈Λ ck(f ) < ∞}, which for the trigono- metric system is the usual Wiener algebra. 7.4. General lower bounds. We say that a Banach space X has type t if there exists a universal constant C such that for fk ∈ X and all n ≥ 1, (7.10) Aveεk=±1k εkfkkt!1/t n Xk=1 ≤ C n Xk=1 kfkkt!1/t . So, given Λ ⊂ N, if we set fk = ψk, k ∈ Λ, then there exists a function f =Pk∈Λ εkψk with kf kp ≤ CΛ1/t. Thus, kSΛf kA = Λ = Λ1/t′ Λ1/t ≥ C −1 Λ1/t′ kf kp. For X = Lp it is well-known that t = min{p, 2} if 1 < p < ∞. Therefore, the following lower bounds always hold (7.11) kSΛkp,A & Λ1/p′ if 1 < p ≤ 2, and kSΛkp,A & Λ1/2 if p > 2. As we shall see below, there are bases Ψ for which the symbols "&" in (7.11) can be replaced by "≈" for all Λ. Notice, from Theorem 1.9, that the WCGA will be most effective in those situations, since the exponents will satisfy rq′ = 1, and therefore we have φ(N) = c(t, p) (log U) N. 26 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV 7.5. Schauder bases. Let Ψ be a (normalized) Schauder basis in Lp, 1 < p < ∞. Then, the following general upper bound holds. PROPOSITION 7.12. There exists ε = ε(p, Ψ) > 0 and c > 0 such that kSΛkp,A ≤ c Λ1−ε, ∀ Λ ⊂ N. Proof. Indeed, by a well-known theorem of Gurari-Gurari [10], there exists s < ∞ and c > 0 such that k ∞ Xn=1 anψnkp ≥ 1 c (cid:16) ∞ Xn=1 ans(cid:17)1/s , for all finite sequences of scalars an. So if f =Pn anψn we have ∞ kSΛf kA =Xn∈Λ an ≤ Λ1/s′(cid:16) ans(cid:17)1/s Xn=1 ≤ cΛ1− 1 s kf kp. Thus, the result holds with ε = 1/s. (cid:3) REMARK 7.13. When no further assumption is made on the basis, the bound in the proposition cannot be improved, even if p = 2. Indeed, for every α < 1, consider the basis Ψ in L2 constructed in [8, Proposition 3.10]. This basis satisfies the following property: if f =P2N n=1 ψn and Λ = {1, 3, . . . , 2N − 1}, then kSΛf k2 ≥ cN α kf k2. Since kSΛf k2 ≤ kSΛf kA (by the triangular inequality), we conclude that kSΛk2,A ≥ c Λα, ∀ Λ ⊂ 2N − 1. 7.6. Uniformly bounded orthogonal bases. Suppose that Ψ = {ψk}k≥1 is an orthonormal system in L2([0, 1]d) such that cΨ := supk≥1 kψkkL∞ < ∞. These systems satisfy c−1 Ψ ≤ kψkkp ≤ cΨ, for all k ≥ 1 and all 1 ≤ p ≤ ∞. 7.6.1. Case 2 ≤ p < ∞. PROPOSITION 7.14. If Ψ is a uniformly bounded orthogonal basis of Lp[0, 1], 2 ≤ p < ∞, then kSΛ(·, Ψ)kLp→A ≍ Λ1/2, ∀ Λ. Proof. The lower bound was shown in (7.11). The upper bound follows from an elementary argument, as in [19, Example 1]. (cid:3) LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 27 7.6.2. Case 1 < p ≤ 2. In this case the following holds (7.15) c1 Λ1/2 ≤ kSΛkp,A ≤ c2 Λ1/p, ∀ Λ, and some constants ci(p, Ψ) > 0. The lower bound is due to kSΛkp,A ≥ kSΛk2,A = Λ1/2. The upper bound follows from the argument in [19, Example 1q]; see also [21, Ch. 8]. The example of the univariate trigonometric system shows that the bounds in (7.15) cannot be improved for this class of bases. REMARK 7.16. It is known that uniformly bounded orthogonal bases cannot be unconditional in Lp, for any p 6= 2; see [6]. Thus, when p > 2 we cannot hope to obtain φ(N) = O(N) from Theorem 1.9 using these examples. However, it is possible to construct such bases with φ(N) = O(N log log N); see Proposition 7.25 below. 7.7. Unconditional bases. Recall that a Banach space X has cotype q if there exists c > 0 such that (7.17) Aveεk=±1 hk N Xn=1 εnfnki ≥ ch N Xn=1 1 q . kfnkqi If Ψ is an unconditional basis, and we let fn = anψn, then we have (7.18) k N Xn=1 anψnk ≥ c2h N Xn=1 1 q , anqi with a constant c2 = c2(Ψ, c) > 0. Thus, if f =Pn anψn, for a finite sequence an, we have (7.19) kSΛf kA =Xn∈Λ an ≤ Λ1/q′hXn anqi 1 q ≤ c−1 2 Λ1/q′ kf k. Now, specializing to the case when X = Lp, we know that q = max{2, p}. Therefore, from (7.19) and (7.11) we obtain the following Case 1 < p ≤ 2. (7.20) c1 Λ1/p′ ≤ kSΛkp,A ≤ c2 Λ1/2, ∀ Λ, Case 2 ≤ p < ∞. (7.21) c1 Λ1/2 ≤ kSΛkp,A ≤ c2 Λ1/p′ , ∀ Λ, These inequalities cannot be improved for the whole class of unconditional bases. Indeed, when Ψ is the univariate Haar basis we have, for all 1 < p < ∞, (7.22) kSΛkp,A ≍ Λ1/p′ , ∀ Λ. On the other hand, it is known that Lp ≈ Lp ⊕ ℓ2, when 1 < p < ∞. Consider n=1 in Lp ⊕ ℓ2, where {ϕ2n−1} is the Haar system in Lp, and {ϕ2n} is the basis {ϕn}∞ 28 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV the canonical basis in ℓ2. Then, the above isomorphim produces a (seminormalized) unconditional basis Ψ in Lp, with the property (7.23) kSΛkp,A ≈ Λ1/2, ∀ Λ ⊂ 2N. 7.8. Quasi-greedy bases. If Ψ is a (normalized) quasi-greedy basis in a Banach space X of cotype q, then letting fn = ψn in (7.17) we obtain ψnk ≥ c1 Λ1/q, kXn∈Λ for some c1(Ψ) > 0. This is weaker than (7.18), but using [9, Lemma 4.1] we have k(an)kℓq,∞ ≤ C kXn If f =Pn anψn, for a finite sequence (an), and (a∗ ment, then anψnk. j ) denotes its decreasing rearrange- kSΛf kA =Xn∈Λ (7.24) an ≤ a∗ j ≤ k(an)kℓq,∞ Λ Xj=1 ≤ C ′ Λ1/q′ kf k. j− 1 q Λ Xj=1 When X = Lp, using the corresponding value for the cotype q, one obtains again the bounds in (7.20) and (7.21). From the examples in (7.22) and (7.23) it is clear that these bounds cannot be improved in the class of all quasi-greedy bases. Notice further that within this class one can replace the example in (7.23) by the following construction. PROPOSITION 7.25. There exists a uniformly bounded orthonormal system Ψ = {ψk}∞ k=1, consisting of trigonometric polynomials, which is a quasi-greedy basis in Lp[0, 1], for all 1 < p < ∞. Moreover, Ψ is democratic with kPk∈Λ ψkkp ≍ Λ1/2, for all finite Λ ⊂ N. For details on this construction, we refer to [16] and [4] (see also [20, Ch 3] and [21, Ch 3]). The example Ψ in Proposition 7.25 produces the bounds (7.26) kSΛkp,A ≈ Λ1/2, ∀ Λ ⊂ N. Also, quasi-greedness implies that ΣN has property A2 with U . log N; see [1, Lemma 8.2]. In particular, when p > 2, the Lebesgue inequality for the WCGA which one obtains from Theorem 1.9 holds with φ(N) = O(N log log N); see [19] or [21, p. 445]. So far, we do not know any example of a basis (or even a dictionary) in Lp, p > 2, with φ(N) = O(N). LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 29 7.9. Almost greedy bases. This class is a subset of the previous case, and therefore kSΛkp,A satisfies the same bounds (7.20) and (7.21). The examples in (7.22) and (7.26) show that the assumption that Ψ is additionally democratic does not imply any improvement in those bounds. 7.10. Greedy bases. This class is also a subset of the previous three cases, but this time we obtain the following improvement. The proof follows easily from the same arguments we already gave in (7.4). PROPOSITION 7.27. Let Ψ be a greedy basis of Lp, 1 < p < ∞. Then (7.28) kSΛkp,A ≍ Λ1/p′ , ∀ Λ ⊂ N. 8. Appendix 1 In this section we prove the following result, which was asserted in (5.15). PROPOSITION 8.1. Let X = fp,q(Rd), 1 < p, q < ∞, with norm defined as in (5.1). Then sup A≤N(cid:13)(cid:13)Xn∈A eI(cid:13)(cid:13)fp,q(Rd) ≤ c N 1 p (1 + log N)(d−1)( 1 q − 1 p )+ . For q = 2 this was proved in [23, 15]. Here we adapt the arguments to the case of a general q. We split the proof into several lemmas, which have an independent interest. Recall that, for a sequence x = (xλ)λ∈Λ, over a set of indices Λ, we define its support as supp x = {λ ∈ Λ : xλ 6= 0}. LEMMA 8.3. Let p ≥ q. If x0, x1, x2, . . . have disjoint supports, then (8.2) (8.4) Proof. The support condition implies that Then, the definition of norm in (5.1) and Minkowski's inequality (since p/q ≥ 1) give . q fp,q(cid:17)1/q (cid:13)(cid:13)(cid:13)Pn≥0 xn(cid:13)(cid:13)(cid:13)fp,q ≤ (cid:16)Pn≥0(cid:13)(cid:13)xn(cid:13)(cid:13) Sq(Pn≥0 xn)q =Pn≥0 Sq(xn)q. = (cid:13)(cid:13)(cid:13) Sq(Pn≥0 xn)(cid:13)(cid:13)(cid:13)Lp ≤ (cid:16)Pn≥0(cid:13)(cid:13)Sq(xn)q(cid:13)(cid:13)Lp/q(cid:17)1/q 1/q Lp/q = (cid:13)(cid:13)(cid:13)Pn≥0 Sq(xn)q(cid:13)(cid:13)(cid:13) = (cid:16)Pn≥0(cid:13)(cid:13)xn(cid:13)(cid:13) q fp,q(cid:17)1/q . (cid:3) (cid:13)(cid:13)(cid:13)Pn≥0 xn(cid:13)(cid:13)(cid:13)fp,q In the next two lemmas we shall assume that d = 1, so that R1 = D. We denote by ℓp,q = ℓp,q(D) the discrete Lorentz space indexed by D. We shall use the follow- j = xIj )j≥1 is its decreasing ing (equivalent) quasi-norm: if x = (xI)I∈D and if (x∗ 30 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV rearrangement, then (cid:13)(cid:13)x(cid:13)(cid:13)ℓp,q := (cid:16)Xj≥0 2j/px∗ 2j q(cid:17)1/q . LEMMA 8.5. Let d = 1 and p ≥ q. Then, ℓp,q(D) ֒→ fp,q(D) ֒→ ℓp(D), that is (8.6) kxkℓp ≤ kxkfp,q ≤ c kxkℓp,q. Proof. The left inequality is trivial, since q ≤ p implies fp,q ֒→ fp,p = ℓp(D). We prove the right inequality. Let x = (xI )I∈D and let xI1 ≥ xI2 ≥ . . . be its decreasing rearrangement. Define xn = X2n≤j<2n+1 xIj eIj , so that x = Pn≥0 xn and the vectors xn have disjoint supports. Then, (8.4) holds. Moreover, for each n ≥ 0 kxnkfp,q ≤ x∗ 2n(cid:13)(cid:13)P2n≤j<2n+1 eIj(cid:13)(cid:13)fp,q ≤ cp,q x∗ 2n 2n/p, where the last inequality is due to the p-democracy of the canonical basis in fp,q(D) when d = 1; see [7, Prop. 3.2]. Thus, kxkfp,q ≤ (cid:16)Xn≥0 kxnkq fp,q(cid:17)1/q . (cid:16)Xn≥0 2n/px∗ 2nq(cid:17)1/q = kxkℓp,q. (cid:3) LEMMA 8.7. Let d = 1 and 1 < q < p < ∞. Then, for all A ⊂ D with A ≤ N, it holds (cid:13)(cid:13)(cid:13)XI∈A xIeI(cid:13)(cid:13)(cid:13)fp,q(D) ≤ c′(cid:2)1 + log N(cid:3) 1 q − 1 p (cid:16)XI∈A xIp(cid:17)1/p . Proof. By (8.6) and Holder's inequality we have (8.8) kxkfp,q . kxkℓp,q ≤ kxk1−θ ℓp,1 kxkθ ℓp,p, with θ = (1 − 1 q )/(1 − 1 p ). Another use of Holder's inequality gives kxkℓp,1 = ⌊log2 A⌋ Xj=0 2j/px∗ 2j ≤ (1 + log2 A)1/p′ kxkℓp,p. Inserting this into (8.8) and using that kxkℓp,p ≈ kxkℓp the result follows. (cid:3) PROOF of Proposition 8.1: and The proof for p ≤ q is trivial, since ℓp = fp,p ֒→ fp,q, (cid:13)(cid:13)Xn∈A eI(cid:13)(cid:13)fp,q ≤(cid:13)(cid:13)Xn∈A eI(cid:13)(cid:13)ℓp = A1/p. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 31 So from now on we consider p > q. The result is known for d = 1 by the p-democracy of fp,q(D); see [7, Prop. 3.2]. So we proceed by induction, and will prove (8.2) assuming its validity with d replaced by d − 1. Let A ⊂ Rd with A ≤ N, and define A1 = {I1 ∈ D : ∃ I ′ ∈ Rd−1 s.t. I1 × I ′ ∈ A} and for each I1 ∈ A1, A′(I1) = {I ′ ∈ Rd−1 : I1 × I ′ ∈ A}. Then (cid:13)(cid:13)XI∈A eIkp fp,q(Rd) = Zu′Zu1(cid:12)(cid:12)(cid:12) XI1∈A1(cid:0) XI ′∈A′(I1) (1 + log N)( 1 q − 1 . Zu′ p/q 1 q q I ′,p(cid:1)1 I1,p(cid:12)(cid:12)(cid:12) p )p XI1∈A1(cid:12)(cid:12)(cid:12) XI ′∈A′(I1) du1 du′ p/q du′, 1 q I ′,p(cid:12)(cid:12) using Lemma 8.7 in the inner integral, since A1 ≤ N. The last displayed expression equals (1 + log N)( 1 q − 1 p/q du′ 1 q . (1 + log N)( 1 p )p XI1∈A1Zu′(cid:12)(cid:12)(cid:12) XI ′∈A′(I1) I ′,p(cid:12)(cid:12) p )p XI1∈A1 = h(1 + log N)(d−1)( 1 q − 1 q − 1 p )A1/pip , (1 + log N)(d−2)( 1 q − 1 p )pA′(I1) using in the middle step the induction hypothesis (since A′(I1) ≤ N). (cid:3) Acknowledgments The authors would like to thank the Isaac Newton Institute for Mathematical Sci- ences, Cambridge, for support and hospitality during the program Approximation, Sampling and Compression in Data Science where some work on this paper was un- dertaken; this work was supported by EPSRC grant no EP/K032208/1. G.G. was supported in part by grants MTM2016-76566-P, MTM2017-83262-C2- 2-P and Programa Salvador de Madariaga PRX18/451 from Micinn (Spain), and grant 20906/PI/18 from Fundaci´on S´eneca (Regi´on de Murcia, Spain). E.H. was supported by grant MTM2016-76566-P (Spain), and by the European Union's Horizon 2020 research and innovation programme under the Marie Sk lodowska-Curie grant agreement No 777822. D.K. was supported by Simons Foundation Collaborative Grant No 636954. V.T. was supported by the Russian Federation Government Grant No. 14.W03.31.0031. 32 S. DILWORTH, G. GARRIG ´OS, E. HERN ´ANDEZ, D. KUTZAROVA, AND V. TEMLYAKOV References [1] S.J. Dilworth, N.J. Kalton, D. Kutzarova, On the existence of almost greedy bases in Banach spaces, Studia Math. 159 (1) (2003), 67 -- 101. [2] S.J. Dilworth, N.J. Kalton, Denka Kutzarova, and V.N. Temlyakov, The Thresholding Greedy Algorithm, Greedy Bases, and Duality, Constr. Approx. 19 (4) (2003), 575 -- 597. [3] S.J. Dilworth, D. Kutzarova, K. Shuman, V. Temlyakov, and P. Wojtaszczyk, Weak convergence of greedy algorithms in Banach spaces. J. Fourier Anal. Appl. 14 (2008), no. 5-6, 609 -- 628. [4] S.J. Dilworth, M. Soto-Bajo, and V.N. Temlyakov, Quasi-greedy bases and Lebesgue- type inequalities, Studia Math. 211 (1) (2012), 41 -- 69. [5] T. Figiel, On the moduli of convexity and smoothness. Studia Math. 56 (2) (1976), 121 -- 155. [6] V.F. Gaposkin, On unconditional bases in Lp (p > 1) spaces. Uspehi Mat. Nauk 13 (1958) no. 4 (82), 179184. [7] G. Garrig´os, E. Hern´andez, Sharp Jackson and Bernstein Inequalities for n-term Approxi- mation in Sequence spaces with Applications, Indiana Univ. Math. J. 53 (6) (2004), 1739 -- 1762. [8] G. Garrig´os, P. Wojtaszczyk, Conditional quasi-greedy bases in Hilbert and Banach spaces. Indiana Univ. Math. J. 63 (4) (2014), 1017 -- 1036. [9] R. Gribonval, M. Nielsen, Some remarks on non-linear approximation with Schauder bases. East J. Approx. 7 (3) (2001), 267 -- 285. [10] V.I. Gurarii and N.I. Gurarii, On bases in uniformly convex and uniformly smooth spaces. Izv. Akad. Nauk SSSR Ser. Mat. 35 (1971), 210 -- 215. [11] M.I. Kadec and A. Pelczynski, Bases, lacunary sequences, and complemented subspaces in the spaces Lp. Studia Math. 21 (1962), 161-176. [12] A. Kamont, V.N. Temlyakov, Greedy approximation and the multivariate Haar system. Studia Math. 161 (3) (2004), 199 -- 223. [13] K. Kazimierski, On the smoothness and convexity of Besov spaces. J. Inverse Ill-Posed Probl. 21 (3) (2013), 411 -- 429. [14] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I,II. Springer-Verlag, Berlin, 1979. [15] G. Kerkyacharian, D. Picard, V.N. Temlyakov, Some inequalities for the tensor product of greedy bases and weight-greedy bases. East J. Approx. 12 (1) (2006), 103 -- 118. [16] M. Nielsen, An example of an almost greedy uniformly bounded orthonormal basis for Lp(0, 1). J. Approx. Theory, 149 (2) (2007), 188 -- 192. [17] V.N. Temlyakov, Greedy algorithms in Banach spaces. Adv. Comput. Math. 14 (3) (2001), 277 -- 292. [18] V.N. Temlyakov. Greedy Approximation. Cambridge University Press, Cambridge, 2011. [19] V.N. Temlyakov, Sparse approximation and recovery by greedy algorithms in Banach spaces. Forum Math, Sigma 2 (12) (2014), 26 p. [20] V.N. Temlyakov, Sparse Approximation with Bases, Advanced Courses in Mathematics CRM Barcelona, Birkhauser, Springer Basel 2015. [21] V.N. Temlyakov. Multivariate Approximation. Cambridge University Press, Cambridge, 2018. [22] H. Triebel. Function spaces with dominating mixed smoothness. European Mathematical So- ciety, Zurich, 2019. [23] P. Wojtaszczyk, Greedy Algorithm for General Biorthogonal Systems, J. Approx. Theory 107 (2) (2000), 293-314. LEBESGUE-TYPE INEQUALITIES IN GREEDY APPROXIMATION 33 S. J. Dilworth, Department of Mathematics, University of South Carolina, Columbia, SC, USA E-mail address: [email protected] Gustavo Garrig´os, Departamento de Matem´aticas, Universidad de Murcia, 30100 Murcia, Spain E-mail address: [email protected] Eugenio Hern´andez, Departamento de Matem´aticas, Universidad Aut´onoma de Madrid, 28049 Madrid, Spain E-mail address: [email protected] D. Kutzarova, Department of Mathematics, University of Illinois Urbana-Champaign, Urbana, IL 61807, USA; Institute of Mathematics and Informatics, Bulgarian Acad- emy of Sciences, Sofia, Bulgaria E-mail address: [email protected] V. N. Temlyakov, Department of Mathematics, University of South Carolina, Columbia, SC, USA; Steklov Institute of Mathematics, Moscow, Russia; Lomonosov Moscow State University, Moscow, Russia E-mail address: [email protected]
1901.05245
1
1901
2019-01-16T12:03:12
c-numerical range of operator products on B(H)
[ "math.FA" ]
Let H be a complex Hilbert space of dimension no less than 2 and B(H) be the algebra of all bounded linear operators on H. We give the form of surjective maps on B(H) preserving c-numerical range of operator products when the maps satisfy preserving weak zero products. As a result, we obtain the characterization of surjective maps on Mn(C) preserving c-numerical range of operator products. The proof of the results depends on some propositions of operators in B(H), which are of different interest.
math.FA
math
c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) YANFANG ZHANG,1 and XIAOCHUN FANG2∗ Abstract. Let H be a complex Hilbert space of dimension ≥ 2 and B(H) be the algebra of all bounded linear operators on H. We give the form of surjective maps on B(H) preserving c-numerical range of operator products when the maps satisfy preserving weak zero products. As a result, we obtain the characterization of surjective maps on Mn(C) preserving c-numerical range of operator products. The proof of the results depends on some propositions of operators in B(H), which are of different interest. 1. Introduction and preliminaries Let H be a complex Hilbert space with inner product h·, ·i, B(H) and Bs(H) be the algebra of all bounded linear operators and the real linear space of all self- adjoint operators on H, respectively. For c = (c1, ..., ck) ∈ Rk and k ≤ dim H, the c-numerical range and c-numerical radius of A ∈ B(H) are respectively defined as Wc(A) =( kXj=1 cjhAej, eji : {e1, ..., ek} is an orthonormal subset in H) , rc(A) = sup{λ : λ ∈ Wc(A)}. Obviously, if k = 1, we get the classical numerical range W (A) and the nu- merical radius w(A) of A. If (c1, ..., ck) = (1, ..., 1), Wc(A) and rc(A) reduce to k-numerical range and k-numerical radius of A, respectively ( [4, 8, 10]). Numer- ical range (radius) and c-numerical range (radius) are important concepts and have many applications in pure and applied mathematics, especially in quantum control and quantum information. Readers can refer to [8, 21] for more informa- tion. It is always of interest to characterize mappings with some special properties such as leaving certain functions, subsets or relations invariant, which are called preserving problems. There is a considerable interest about preservers of numer- ical range and generalized numerical range (radius)[3, 6, 15, 16]. Li gave the form of linear map on Mn(C) preserving c-numerical range (resp. radius) [15]. It is quite different from linear maps preserving numerical range (resp. radius)on B(H)([1]). Hou and Di [13] gave the characterization of surjective maps pre- serving numerical range of operator products on B(H) and Bs(H) and showed that such maps have the form A → ±U AU∗ for some unitary U ∈ B(H). It is 2010 Mathematics Subject Classification. Primary 47B49; Secondary 47A12. Key words and phrases. Preserver, c-numerical range, and elliptical ranges. 1 2 YANFANG ZHANG, and XIAOCHUN FANG natural to ask what is the form of surjective maps preserving c-numerical range of operator products. Let A, B be two algebras and Φ : A → B is a map. For A, B ∈ A, AB = 0 ⇒ Φ(A)Φ(B) = 0 (AB = 0 ⇔ Φ(A)Φ(B) = 0), Φ is called a map preserving zero product (preserving zero products on both sides). The characterization of surjec- tive maps preserving zero products on both sides on standard operator algebras was given in [5, Lemma 2.2]. If c-numerical radius is a norm, maps preserv- ing c-numerical radius of operator products on B(H) are maps preserving zero products on both sides. Then by [5, Theorem 2.3], the form of maps preserving c-numerical radius of operator products for dim(H) ≥ 3 can get when rc(·) is a norm. The same result can also be obtained from [2, Theorem 3.2 and Theorem 3.3 ]. From these, it is possible to get the form of surjective maps preserving c-numerical range of operator products on B(H) only when rc(·) is a norm and the dimension of H is no less than 3. In this article, we give the characterization of surjective maps on Mn(C) pre- serving c-numerical range of matrix products. Moreover, we give the definition of maps preserving weak zero products (on both sides). Then we obtain the form of surjective maps on B(H) preserving c-numerical range of operator products when the maps satisfy preserving weak zero products. From this result, the char- acterization of surjective maps on B(H) preserving c-numerical range of operator products can get easily when rc(·) is a norm. Here dim H just needs to be greater than 1. Then the characterization of surjective maps preserving numerical range of operator products in [13] is just a special case of this result. It is worth noting that we calculate the c-numerical range of some special operators in B(H) and show a type of operators with symmetric c-numerical range whenever H is of finite or infinite dimension. They are of different interest. Next we introduce some notations and state the assumptions in the paper. (1) Assume c = (c1, ..., ck) ∈ Rk and c1 ≥ ... ≥ ck with not all ci's equal. (2)To avoid trivial considerations, we always assume k ≥ 2. (3) If dim H = n < ∞, we regard B(H) as Mn(C), which is the set of all complex n × n matrices. Also we assume n ≥ k. (4) F1(H), S and T are respectively the set of rank-1 operators, the set of operators with symmetrical c-numerical range and the set satisfying {A ∈ S : AB ∈ S and BA ∈ S f or B ∈ B(H)} in B(H). The following are our main results. Theorem 1.1. Let Mn(C) be the set of all n × n complex matrices and Φ : Mn(C) → Mn(C) be a surjective map. Then Wc(Φ(A)Φ(B)) = Wc(AB) for all A, B ∈ Mn(C) if and only if Φ has one of the following forms: (i) when c1 + ck 6= 0, then there is a unitary matrix U ∈ Mn(C) such that Φ(A) = ±U AU∗ for all A ∈ Mn(C); (ii) when there exists an integer p (1 < p < k) satisfying cp + ck+1−p 6= 0 and cj + ck+1−j = 0 for j = 1, ..., p − 1, then there are a unitary matrix U ∈ Mn(C), functions φ : T → {−1, 1} and φ′ : Mn(C) → {−1, 1} satisfying φ′(A)φ′(B) = such that sgn(AB) =(cid:26) {1} Φ(A) = φ(A)U AU∗ φ′(A)U AU∗ ±U AU∗ if AB 6∈ S {−1, 1} if AB ∈ S if A ∈ T if A ∈ S \ T if A ∈ Mn(C) \ S for all A ∈ Mn(C). (iii) when cj + ck+1−j = 0 for j = 1, ..., k, then there are a unitary matrix U ∈ Mn(C) and a function g : Mn(C) → {1, −1} such that or Φ(A) = g(A)U AU∗ for all A ∈ Mn(C) Φ(A) = ig(A)U AU∗ for all A ∈ Mn(C). Next result is about surjective maps on B(H) preserving c-numerical range of operator products when the maps satisfy preserving weak zero products. Theorem 1.2. Let B(H) be the algebra of all bounded linear operators on complex Hilbert space H. Suppose Φ : B(H) → B(H) is a surjective map preserving weak zero products on both sides. Then Wc(Φ(A)Φ(B)) = Wc(AB) for all A, B ∈ B(H) if and only if Φ has one of the following forms: (i) when c1 + ck 6= 0, there is a unitary operator U ∈ B(H) such that Φ(A) = ±U AU∗ for all A ∈ B(H); (ii) When there exists an integer p (1 < p < k) satisfying cp + ck+1−p 6= 0 and cj + ck+1−j = 0 for j = 1, ..., p − 1, then there are a unitary operator U ∈ B(H), functions φ : T → {−1, 1} and φ′ : B(H) → {−1, 1} satisfying φ′(A)φ′(B) = sgn(AB) for any A, B ∈ B(H) where c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) 3 sgn(AB) for any A, B ∈ Mn(C) where sgn(AB) =(cid:26) {1} Φ(A) = φ(A)U AU∗ φ′(A)U AU∗ ±U AU∗ if AB 6∈ S {−1, 1} if AB ∈ S if A ∈ T if A ∈ S \ T if A ∈ B(H) \ S such that for all A ∈ B(H). (iii) When cj + ck+1−j = 0 for j = 1, ..., k, then there are a unitary operator U ∈ B(H) and a function g : B(H) → {1, −1} such that Φ(A) = g(A)U AU∗ for all A ∈ B(H) Φ(A) = ig(A)U AU∗ for all A ∈ B(H). ci 6= 0, then rc(·) is a norm, the following result can be obtained imme- or If nPk=1 diately. 4 YANFANG ZHANG, and XIAOCHUN FANG Corollary 1.3. Suppose Φ : B(H) → B(H) is a surjective map satisfying Wc(Φ(A)Φ(B)) = Wc(AB) for all A, B ∈ B(H). If is either (i) or (ii) in Theorem 1.2. nPk=1 ci 6= 0, the form of Φ The paper is organized as follows. In Section 2, we show some properties of c-numerical range, which are true whether H is finite or not. We calculate c- numerical range of rank-1, some rank-2 operators in B(H) and show a type of operators with symmetrical c-numerical range. In Section 3, the proof of main results will be given. 2. c-numerical range of operators on B(H) In [15], Li showed the properties of c-numerical range and c-numerical radius on Mn(C). First we list some of them which are also true in B(H) when H is of infinite dimension. Proposition 2.1. Let c = (c1, ..., ck) ∈ Rk. For A ∈ B(H), then (1) [20]The set Wc(A) is convex for all A ∈ B(H). (2) Wc(U AU∗) = Wc(A) for any unitary U ∈ B(H). (3) Wc(λA) = λWc(A) for any λ ∈ C. (4) Wc(λI + A) = λ ci + Wc(A) for any λ ∈ C. kPi=1 (5) Wβc(A) = βWc(A) for any β ∈ R. (6) Suppose ci's are not equal. Then Wc(A) is a singleton if and only if A is a scalar multiple of the identity. (7) The c-numerical radius rc(A) is a norm on B(H) if and only if and not all ci's are equal. ci 6= 0 kPi=1 (8) Suppose ci's are not equal. Then Wc(A) ⊂ R if and only if A is hermitian. In [17], Li described the c-numerical range for self-adjoint operators especially when H is infinite dimensional. Let S be a self-adjoint operator in B(H). If dimH = k, denote by λ1(S) ≥ ... ≥ λk(S) the eigenvalues of S. If H is infinite dimensional, define λm(S) = sup{λm(X∗SX) : X∗X = Im}. Also λm of an infinite dimensional operator S can be defined as follows. Let σe(A) = ∩{σ(A + F ) : F ∈ B(H) has finite rank} be the essential spectrum of A ∈ B(H), and let which also equals the supremum of the set λ∞(S) = sup σe(S), σ(S)\{S − µI has a non − trivial finite dimensional null space}. Then S = σ(S)∩(λ∞(S), ∞) has only isolated points, and we arrange the elements in descending order, say, λ1 ≥ λ2 ≥ ... counting multiplicities. If S is infinite, then λj(S) = λj for each positive integer j. If S has m elements, then λj(S) = λj c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) 5 for j = 1, ..., m, and λj = λ∞ for j > m. Then the following property is about c-numerical range of self-adjoint operators in B(H). Proposition 2.2. (See [15, 17]) Let S ∈ Bs(H). (1) If dim H = k, Wc(S) =" kPj=1 (2) If H is of infinite dimension, cl(Wc(S)) = [mc(S), Mc(S)], where cjλk+1−j(S), cjλj(S)#. kPj=1 ck+1−jλj(S) : 0 ≤ l ≤ k) k−lXj=1 ck+1−jλj(−S) : 0 ≤ l ≤ k) . k−lXj=1 cjλj(−S) + cjλj(S) − cj if dim H = k max{cj, 0} if dim H > k cj if dim H = k min{cj, 0} if dim H > k and mc(S) = inf(− lXj=1 Mc(S) = sup( lXj=1 ¯cj =(cid:26) ecj =(cid:26) Next we denote and for j = 1, .., k. This will be used to describe c-numerical range of some operators in B(H). Proposition 2.3. Suppose that T is a rank-1 operator in B(H), then Wc(T ) is an elliptical disk with foci ¯c1tr(T ),ecktr(T ) and minor axis (¯c1−eck)pk T k2 − tr(T ) 2 or a line segment with end points ¯c1tr(T ) andecktr(T ). Proof. If dim H = k, the conclusion can be found in [6]. If dim H = n > k, adding zero to c, we get d = (d1, ..., dn) with d1 ≥ ... ≥ dn and Wc(A) = Wd(A) for all A ∈ B(H). Obviously, d1 = max{c1, 0} and dn = min{ck, 0}. The conclusion also can be obtained. Next we prove it when H is infinite dimensional. Assume T = x ⊗ f with k x k= 1, then there is unit x′ in [x]⊥ and complex a, d such that f = ax + dx′. then with respect to a suitable orthogonal basis in H = [x] ⊕ [x′] ⊕ [x, x′]⊥, T =(cid:18) a d 0 0 (cid:19) ⊕ 0 with k T k=p a 2 + d 2 and tr(T ) = a. 2(eiθT + e−iθT ∗). For 0 ≤ θ < 2π, The real part of eiθT is denoted by Re(eiθT ) = 1 We calculate the non-zero eigenvalues of Re(eiθT ), that is and λ1 = 1 2 Re(eiθa) + λ2 = 1 2 Re(eiθa) − 1 2p(Re(eiθa))2 + d2 > 0 2p(Re(eiθa))2 + d2 < 0. 1 6 YANFANG ZHANG, and XIAOCHUN FANG From Proposition 2.2(2), we know Wc(Re(eiθT )) = [mc(Re(eiθT ), Mc(Re(eiθT )], ckλ2 c1λ1 where if c1 ≤ 0 c1λ1 + ckλ2 if c1 > 0 and ck < 0 Writing a as a = aeiφ, then Mc(Re(eiθT )) = With the notation of ¯cj andecj, it can be written Mc(Re(eiθT )) = ¯c1λ1 +eckλ2. Mc(Re(eiθT )) = ¯c1λ1 +eckλ2 2 pa2 cos2(θ + φ) + d2 2(¯c1 +eck)Re(eiθa) + (¯c1−eck) 2(¯c1 +eck)a cos(θ + φ) + (¯c1−eck) It is shown in ([6, 7]) that Ω is an elliptical disc centered at peiφ with foci 2 p(Re(eiθa))2 + d2 (p − β)eiφ and (p + β)eiφ, and semi-minor axis α if and only if = 1 = 1 if ck ≥ 0. (2.1) Mc(Re(eiθΩ)) = p cos(θ + φ) +pβ2 cos2(θ + φ) + α2. 1 (cid:3) Proposition 2.4. Let A ∈ B(H) be rank-2 and there is a suitable orthonormal 0 b (cid:19) ⊕ 0 with a, b, d ∈ C. Then we can get Wc(T ) is an ellipse with ¯c1tr(T ) andecktr(T ) as foci, and 2(¯c1 −eck)pk (T ) k2 −tr(T )2 as semi-minor axis. Specially, if tr(T ) =k T k, Wc(T ) degenerates to a line segment with end points ¯c1tr(T ) andecktr(T ). basis such that A =(cid:18) a d (1) If a ≥ b > 0 and d = 0, then Wc(A) is the line segment [ecka +eck−1b, ¯c1a + (2) If a > 0 > b and d = 0, then Wc(A) is the line segment [ecka+¯c1b, ¯c1a+eckb]. (3) If 0 > a ≥ b and d = 0, then Wc(A) is the line segment [¯c1b + ¯c2a,eckb + eck−1a]. ¯c1a +eckb and ¯c1b +ecka and with minor axis (¯c1 −eck)d. (5) If dim H ≥ 3 and c satisfying cj + ck+1−j = 0 for j = 1, 2, and a, b ∈ R+ (4) If a, b ∈ R and d2 > 4ab, then Wc(A) is the elliptical disk with foci with 0 < d2 < 4ab, then Wc(A) is not an elliptical disk. ¯c2b]. Proof. (1) For the case dim H < ∞, the conclusion can get from Proposition 2.2 easily. If dim H = ∞, A is self-adjoint and Wc(A) = [m, M], where M = max{c1a + c2b, c1a, 0} and m = min{cka + ck−1b, cka, 0}. That is c1a + c2b if c2 ≥ 0 c1a 0 if c1 > 0 and c2 < 0 if c1 ≤ 0 cka + ck−1b cka 0 if ck−1 ≤ 0 if ck−1 > 0 and ck < 0 if ck ≥ 0. and M = m = There are 9 combinations of m, M, five of which can be possible with c1 ≥ c2 ≥ ... ≥ ck as follows: (a) if ck ≥ 0, then m = 0, M = c1a + c2b; c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) 7 (b) if c2 > 0 and ck−1 < 0, then m = cka + ck−1b and M = c1a + c2b; (c) if c1 ≤ 0, then m = cka + ck−1b and M = 0; (d) if c1 > 0 and c2 < 0, then m = cka + ck−1b and M = c1a; (e) if ck−1 > 0 and ck < 0, then m = cka and M = c1a + c2b. (2),(3) can get as the proof of (1) similarly. (4) Assume a, b ∈ R and d2 > 4ab. For 0 ≤ θ < 2π, the real part of eiθA is 2 (eiθA + e−iθA∗). We calculate the non-zero eigenvalues of Re(eiθA) Re(eiθA) = 1 denoting λ1, λ2. Then From the cases, we get m =ecka +eck−1b and M = ¯c1a + ¯c2b. (a + b) cos θ +p(a − b) cos2 θ + d2 (a + b) cos θ −p(a − b) cos2 θ + d2 (¯c1 +eck)(a + b) cos θ Mc(Re(eiθA)) = λ1 = λ2 = 2 2 + 2 (¯c1 −eck)p(a − b)2 cos2 θ + d2 2 Together with Equation (2.1), it follows that Wc(A) is an elliptical disk with foci ¯c1a +eckb and ¯c1b +ecka and minor axis (¯c1 −eck)d. (5) Without loss of generality, suppose a ≥ b. For 0 ≤ θ < 2π, the non-zero eigenvalues of Re(eiθA) is same as that (2.2) and (2.3). Besides, for d2 < 4ab, there must be φ ∈ (0, π 2 ) such that cos2 φ = d2 4ab . Then Obviously, λ1 > 0 > λ2. So we have Mc(Re(eiθA)) = ¯c1λ1 + ckλ2. Then (2.2) (2.3) . where S1 = [2π − φ, 2π) ∪ [0, φ], S2 = (φ, π − φ) ∪ (π + φ, 2π − φ) and S3 = [π − φ, π + φ]. We know W (Re(eiθA)) is interval [m(θ), M(θ)]. From the results in (1), (2) and (3) above, we get λ1 ≥ λ2 > 0 if θ ∈ S1 λ1 > 0 > λ2 if θ ∈ S2 if θ ∈ S3. 0 > λ1 ≥ λ2 ¯c1λ1 + ¯c2λ2 ¯c1λ1 +eckλ2 eck−1λ1 +eckλ2 c1λ1 c1(λ1 − λ2) −c1λ2 c1λ1 + c2λ2 c1(λ1 − λ2) −c2λ1 − c1λ2 if θ ∈ S1 if θ ∈ S2 if θ ∈ S3. if θ ∈ S1 if θ ∈ S2 if θ ∈ S3. if θ ∈ S1 if θ ∈ S2 if θ ∈ S3.  M(θ) = M(θ) = M(θ) = With c1 + ck = 0 and c2 + ck−1 = 0, we get when k = 2 and dim H ≥ 3; or when k ≥ 3. 8 YANFANG ZHANG, and XIAOCHUN FANG Denote M(θ) can be written as M(θ) = if k ≥ 3 if k = 2. 2 c1 2 2 c1 2 if k ≥ 3 if k = 2 and c =(cid:26) c1−c2 c =(cid:26) c1+c2 c(a + b) cos θ + cp(a − b)2 cos2 θ + d2 − c(a + b) cos θ + cp(a − b)2 cos2 θ + d2 M(θ) = p cos(θ + ω) +pq2 cos2(θ + ω) + t2. ( c + c)p(a − b)2 cos2 θ + d2 if θ ∈ S1 if θ ∈ S2 if θ ∈ S3. If Wc(A) is an ellipse, then M(θ) has the form as (2.1), i.e. there is p, q, t ∈ R and ω ∈ [0, 2π) such that From (2.4), we get M(θ) = M(θ + π). Then it can be deduced p = 0 in (2.5). So (2.5) can be written as (2.3′) Next we show ω = 0. For θ ∈ (φ, π − φ) ∪ (π + φ, 2π − φ), (2.4) and (2.3′) tell M(θ) =pq2 cos2(θ + ω) + t2. that q2 cos2(θ + ω) + t2 = ( c + c)2(a − b)2 cos2 θ + ( c + c)2d2. (2.6) Taking the derivative about θ on both sides of (2.6), we get q2 sin(2θ + 2ω) = ( c + c)2(a − b)2 sin 2θ. (2.7) Let θ = π 2 , (2.7) shows sin 2ω = 0 i.e. ω = 0 or π 2 . Again take the derivative on both sides of (2.6) and let θ = π 2 , we get q2 cos 2ω = ( c + c)2(a − b)2. So cos 2ω has to be non-negative and then ω = 0. Then (2.3′) can be written as 2 or 3π M(θ) = ( c + c)p(a − b)2 cos2 θ + d2. Let θ = 0 in (2.4) and (2.3′′), we have c(a + b) + cp(a − b)2 + d2 = ( c + c)p(a − b)2 + d2. which get d2 = 4ab, a contradiction. So Wc(A) is not an ellipse. Proposition 2.5. Suppose T, S ∈ B(H) are rank-1 with Wc(T ) = Wc(S), then (1) If c1 + ck 6= 0, tr(T ) = tr(S); (2) If c1 + ck = 0, tr(T ) = tr(S) or tr(T ) = −tr(S). Proof. By Proposition 2.3, Wc(T ) = Wc(S) implies one of the following equations true: (2.4) (2.5) (2.3′′) (cid:3) (2.8) (2.9) (cid:26) ¯c1tr(T ) = ¯c1tr(S) ecktr(T ) =ecktr(S) (cid:26) ¯c1tr(T ) =ecktr(S) ecktr(T ) = ¯c1tr(S) With the fact that c1 and ck are not 0 at the same time, we can get tr(T ) = tr(S) from (2.8). For (2.9), it shows tr(T ) = tr(S) = 0 when c1 + ck 6= 0 and tr(T ) = −tr(S) when c1 + ck = 0. Hence, when c1 + ck 6= 0, tr(T ) = tr(S). When c1 + ck = 0, either tr(T ) = tr(S) or tr(T ) = −tr(S) holds. (cid:3) c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) 9 Next we consider the symmetry of c-numerical range for operators in B(H). Proposition 2.6. Let A ∈ B(H) be a self-adjoint operator with finite ranks. For c = (c1, ..., ck), there is an integer p (1 < p < k) such that cp + ck+1−p 6= 0 and cj + ck+1−j = 0 for j = 1, ..., p − 1. (1) If rank A < p, then Wc(A) = −Wc(A); (2) If rank A = p and A or −A is positive, then Wc(A) 6= −Wc(A). Proof. Assume A ∈ B(H) is self-adjoint with rank A = a (a ≤ p). Let λ1 ≥ ... ≥ λl be the positive eigenvalues of A and λk ≤ ... ≤ λk+1−(a−l) be the negative ones. From Proposition 2.3, there is ξ, η ∈ R such that Wc(A) = [ξ, η]. (1) If 1 ≤ a < p, we can calculate ξ = ck+1−jλj + cjλa+1−j and η = lPj=1 a−lPj=1 cjλj + ck+1−jλa+1−j. Obviously, ξ + η = (cj + ck+1−j)λj + ck+1−j)λa+1−j = 0. So Wc(A) = −Wc(A). (2) If a = p and A is positive, the eigenvalues of A, saying λ1, ..., λp, are all (cj + a−lPj=1 positive. When dim H = k, then ξ = ck+1−jλj and η = cjλj. Easily, a−lPj=1 lPj=1 lPj=1 pPj=1 ξ + η = ck+1−jλj + cjλj 6= 0. When H is infinite dimensional, we know pPj=1 ξ = and pPj=1  pPj=1 k−tPj=1 pPj=1 pPj=1 pPj=1 tPj=1  η = ck+1−jλj ck+1−jλj ck+1−jλj if p ≤ t and p < k − t if p ≤ t and p > k − t if p > t and p ≤ k − t cjλj if p ≤ t and p < k − t cjλj if p ≤ t and p > k − t cjλj if p > t and p ≤ k − t pPj=1 p−1Pj=1 k−tPj=1 where t is the integer satisfying 1 ≤ t ≤ k and ct ≥ 0 > ct+1. Next we show ξ + η 6= 0 according to different cases of t: (a) If p ≤ t and p < k − t, then ξ + η = (cj + ck+1−j)λj + (cp + ck+1−p)λp 6= 0. (b) If p ≤ t and p > k − t, then ξ + η = (cj + ck+1−j)λj + (ck−t+1λk−t+1 + ... + cpλp) = ck−t+1 + ... + cp. If we assume ξ + η = 0, then ck−t+1λk−t+1 + ... + cpλp = 0, and with ck−t+1 ≥ ... ≥ cp ≥ ct ≥ 0, we get ck−t+1 = ... = cp = ... = ct = 0. Also cp + ck+1−p 6= 0, so ck+1−p 6= 0. By p > k − t, we can get k + 1 − p ≤ t. That is 10 YANFANG ZHANG, and XIAOCHUN FANG ck+1−p ≥ ct = 0. But from the assumption p ≤ k +1−p, it comes ck+1−p ≤ cp = 0. We get a contradiction and ξ + η 6= 0. (c) If p > t and p ≤ k − t, then ξ + η = (cj + ck+1−j)λj + (ck+1−(t+1)λt+1 + ... + ck+1−pλp) = ck+1−(t+1)λt+1 + ... + ck+1−pλp < 0, so ξ + η 6= 0. This means Wc(A) 6= −Wc(A) if a = p. As to −A is positive, similar results can get. (cid:3) tPj=1 For the self-adjoint operators with rank more than p, the symmetry of c- numerical range is associated with cj and ck+1−j for j > p, hence we failed to give a fixed conclusion. 3. Maps preserve c-numerical range of operator products on B(H) In this section, first we give the definition of maps preserving weak zero prod- ucts. Definition 3.1. Let A, B be two subalgebras of B(H) containing F1(H) and Φ : A → B be a map. Φ is said to preserve weak zero products if for any A and rank-1 T ∈ A, AT = 0 implies Φ(A)Φ(T ) = 0. Φ is said to preserve weak zero products on both sides if it preserves weak zero products and for any B and rank-1 S ∈ B, BS = 0 implies B S = 0 where Φ( B) = B and Φ( S) = S. Obviously, a map preserving zero products must be one preserving weak zero products. Next we give an example to show the inverse is not true. Example 3.2. Let H = H1 ⊕ H2 with dim H1 = 2 and U =(cid:18) 0 1 1 0 (cid:19) ⊕ I be an unitary operator. Define Φ : B(H) → B(H) by Φ(A) =(cid:26) U AU∗ AU∗ if rankA ≤ 1 if rank A ≥ 2 It is easy to check that Φ is preserving left-weak zero product. Take non-zero B0 ∈ B(H) and let A =(cid:18) 0 1 However, Φ(A)Φ(B) = U(AU∗B)U∗ = U((cid:18) 1 0 0 0 (cid:19) ⊕ 0 and B =(cid:18) 1 0 0 0 (cid:19) ⊕ 0)U∗ 6= 0. 0 0 (cid:19) ⊕ B1, then AB = 0. From Theorem 1.2, it can be see that for maps preserving c-numerical range of operator products on B(H), they are equivalent between preserving weak zero products and preserving zero products. Next we show some lemmas which are useful to the proof of main theorems . Lemma 3.3. Let A ∈ B(H). The following conditions are equivalent. (1) A is a rank-1 operator. (2) For any B ∈ B(H) with AB 6= 0, there is tB ∈ C such that Wc(AB) is an elliptical disk (a line segment) with ¯c1tB andecktB as foci (end points). Proof. Since(1)⇒(2) is clear, next we show (2)⇒(1). Suppose rank A ≥ 2, then there are linear independent x1,x2 ∈ H such that Ax1⊥Ax2 and kAx1k = kAx2k = 1. We finish the proof by three cases: c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) 11 (a): If c1 + ck 6= 0, we let B1 = αx1 ⊗ Ax1 + βx2 ⊗ Ax2, where α > 0 > β. AB1 is a self-adjoint operator in B(H). By proposition 3(2), Wc(AB1) = [¯c1β + (cid:26) ¯c1β +eckα =ecktB1 ¯c1α +eckβ = ¯c1tB1 (b): If c1 + ck = 0 and c2 + ck−1 6= 0, we let B2 = ξx1 ⊗ Ax1 + ηx2 ⊗ Ax2, where ξ > η > 0 . AB is a self-adjoint operator in B(H). By proposition 3(1), eckα, ¯c1α +eckβ]. On the other hand, there is a complex tB1 such that Wc(AB1) is a line segment with end points ¯c1tB1 andecktB1. So we get (cid:26) ¯c1β +eckα = ¯c1tB1 ¯c1α +eckβ =ecktB1 It deduces ¯c1 +eck = 0, impossible. Wc(AB2) = [eckξ +eck−1η, ¯c1ξ +¯c2η]. Also there is a complex tB2 such that Wc(AB2) is a line segment with end points ¯c1tB2 andecktB2. So (cid:26) eckξ +eck−1η = ¯c1tB2 ¯c1ξ + ¯c2η =ecktB2 It deduces ¯c2 +eck−1 = 0, impossible. (c): If c1 + ck = 0 and c2 + ck−1 = 0, we let B3 = x1 ⊗ Ax1 + x2 ⊗ Ax2 when dim H = 2. Then Wc(AB3) is a singleton, a contradiction. When dim H ≥ 3, let B4 = ax1 ⊗ Ax1 + εx2 ⊗ Ax1 + bx2 ⊗ Ax2 where a > b > 0 and 0 < ε2 < 4ab, then we know Wc(AB4) is not an ellipse from proposition (5), a contradiction. So A must be rank-1. (cid:3) or or (cid:26) eckξ +eck−1η =ecktB2 ¯c1ξ + ¯c2η = ¯c1tB2 Lemma 3.4. Let T ∈ B(H), there is a function h : H × H → {1, −1} satisfying hT x, f i = h(x, f )hx, f i for each x, f ∈ H, then T = ±I. Proof. Firstly, we will show hT x, xi = hx, xi for any unit x ∈ H or hT x, xi = −hx, xi for any unit x ∈ H. If not, there are linearly independent unit x1, x2 ∈ H such that hT x1, x1i = hx1, x1i and hT x2, x2i = −hx2, x2i, then hT (x1 + x2), x1 + x2i = hT x1, x1i + hT x2, x2i + hT x1, x2i + hT x2, x2i = hx1, x1i + hx1, x2i − hx2, x2i − hx2, x1i. We know either hT (x1 + x2), x1 + x2i = hx1 + x2, x1 + x2i or hT (x1 + x2), x1 + x2i = −hx1 + x2, x1 + x2i. Hence, we get hx1, x2i = kx1k2 or hx2, x1i = kx2k2, a contradiction with the linear independence of x1, x2. Assume hT x, xi = hx, xi for x ∈ H. For each f ∈ H, there is x ∈ [x] and f ∈ [x]⊥ such that f = x + f , where [x] is the subspace spanned by x and [x]⊥ is orthogonal subspace of [x]. Then hT x, f i = hT x, x + f i = hx, xi. So hT x, f i = hx, f i for any x, f ∈ H. Then T = I. If hT x, xi = −hx, xi holds for x ∈ H, T = −I can be get similarly. (cid:3) Lemma 3.5. [18, Wigner Theorem] Let H be a complex Hilbert space and T : H → H be a bijective satisfying hT x, T yi = hx, yi for any x, y ∈ H, then there are a unitary or anti-unitary operator U : H → H and a function θ : H → C with θ(x) ≡ 1 such that T x = θ(x)U x for any x ∈ H. Proof of Theorem 3.4. The sufficiency is obvious, here we only show the necessity, i.e. Φ is a surjective map preserving weak zero products on both sides and Wc(Φ(A)Φ(B)) = Wc(AB) for all A, B ∈ B(H). 12 YANFANG ZHANG, and XIAOCHUN FANG Claim 1. Φ preserving rank-1 on both sides. Assume T ∈ B(H) is rank-1 and T satisfying Φ( T ) = T . For any A ∈ B(H) with AΦ(T ) 6= 0, there is A such that Φ( A) = A. Then Wc(Φ( A)Φ(T )) = Wc(AΦ(T )) = Wc( AT ). Because Φ preserves weak zero products on both sides, then AT 6= 0. Thus Wc(AΦ(T )) is an ellipse (line segment) with foci(end points) rank-1. Next we will finish the proof by two cases. ¯c1tr( AT ) and ecktr( AT ). By Lemma 3.3, Φ(T ) is rank-1. Similarly, T is also Case I when c1 + ck 6= 0. Proof of this case will be finished by 3 steps. Step 1.1. Φ is injective. Assume A1, A2 ∈ B(H) satisfying Φ(A1) = Φ(A2). From Wc(Φ(Ai)Φ(x⊗f )) = Wc(Aix⊗f ) for i = 1, 2, we can get Wc(A1x⊗f ) = Wc(A2x⊗f ) for any x, f ∈ H. With c1 + ck 6= 0 and Proposition ??, hA1x, f i = hA2x, f i for any x, f ∈ H. So A1 = A2 and Φ is injective. Step 1.2. Φ is linear. Let A1 and A2 in B(H) and T ∈ B(H) be rank-one, we know tr(Φ(A1 + A2)Φ(T )) = tr((A1 + A2)T ). So tr(Φ(A1 + A2)Φ(T )) = tr(A1T ) + tr(A2T ) = tr((Φ(A1) + Φ(A2))Φ(T )). By the surjection of Φ, Φ(T ) can run over all rank- one operators. So Φ(A1 + A2) = Φ(A1) + Φ(A2). Similarly, we can check Φ is homogeneous. Step 1.3. There is a unitary operator U ∈ B(H) such that Φ(A) = U AU∗ for any A ∈ B(H) or Φ(A) = −U AU∗ for any A ∈ B(H) . Now we get that Φ is a linear bijection preserving rank-one operators in both directions. By [11, Lemma 1.2], Φ has one of the following forms: (i) There exist bijective linear operators U and V such that Φ(x⊗f ) = U x⊗V f for any x, f ∈ H. (ii) There exist bijective conjugate operators U and V such that Φ(x ⊗ f ) = U f ⊗ V x for any x, f ∈ H. If (i) holds, that is, there exist bijective linear operators U and V such that Φ(x ⊗ f ) = U x ⊗ V f for any x, f ∈ H. By considering Wc((x ⊗ f )2) and Proposition ??2.5(1), we get tr((x ⊗ f )2) = tr(Φ(x ⊗ f )2). That is hx, f i2 = hU x, V f i2 for any x, f ∈ H. It follows that U, V are bounded. With Lemma 3.4, V ∗U = ±I. For any A ∈ B(H), Wc(Ax⊗f ) = Wc(Φ(A)Φ(x⊗f )) implies hV ∗Φ(A)U x, f i = hAx, f i. Hence Φ(A) = ±U AU−1. For any unit x ∈ H, Wc((x ⊗ x)2) = W ((U x ⊗ (U−1)∗x)2) implies U x ⊗ (U−1)∗x is self-adjoint. So there is a positive number r such that U−1x = rU x and kxk2 = rkU xk2 for unit x. Thus we get U U∗ = rI. Let U1 = 1√r U and U1 is unitary such that Φ(A) = ±U1AU∗1 for any A ∈ B(H). If (ii) holds, we can get there is conjugate unitary U such that Φ(A) = ±U A∗U∗ for all A ∈ B(H) similarly as (i). Pick an orthogonal basis {ejj ∈ J} and define a conjugate unitary operator J : H → H by Jx =Pj∈J ¯ξjej if x =Pj∈J ξjej. Clearly, J 2 = I, J∗ = J and A∗ = JAtJ, where At is the transpose of A respect to the basis {ej : j ∈ J}. c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) 13 Let U1 = U J, Then U1 is a unitary operator and Φ(A) = ±U1AtU∗1 for all A ∈ B(H). Thus for any A, B ∈ B(H), we have Wc(AB) = Wc(Φ(A)Φ(B)) = Wc(U1AtBtU∗1 ) = Wc((BA)t) = Wc(BA), which is impossible. Case II when c1 + ck = 0. Step 2.1. For any x, y ∈ H, there are ux, uy and λx,y ∈ T = {λ : λ = 1} such that Φ(x ⊗ y) = λx,yux ⊗ uy and hx, yi = hux, uyi. Specially, λx,x ∈ {−1, 1} for any x ∈ H or λx,x ∈ {−i, i} for any x ∈ H. For any x ∈ H, assume Φ(x ⊗ x) = ux ⊗ vx with kuxk = kxk. By considering x ⊗ x, we get kxk2Wc(x ⊗ x) = hux, vxiWc(ux ⊗ vx). Hence kxk4 = hux, vxi2or − hux, vxi2 and hux, vxiux ⊗ vx is self-adjoint. So there exists t ∈ C such that ux = tvx. Then kxk4 = hux, vxi2 = ¯t−2kuxk4 i.e. t ∈ {−1, 1, −i, i}. Then there is λx,x ∈ {−1, 1, −i, i} such Φ(x ⊗ x) = λx,xux ⊗ ux for all x ∈ H. For any y ∈ H, next we will show there is λx,y ∈ T such that Φ(x ⊗ y) = λx,yux ⊗ uy and hux, uyi = hx, yi. Obviously, Φ preserves orthogonality of rank-1 operators in both directions. So the ranges of Φ(x ⊗ y) and Φ(x ⊗ y)∗ contain in span[ux, uy], which is the linear subspace spanned by ux, uy. Assume Φ(x ⊗ y) = vx,y ⊗ wx,y. From Wc(Φ(x ⊗ y)Φ(x ⊗ x)) = Wc((x ⊗ y)(x ⊗ x)), we If x⊥y, obviously ux⊥wx,y, get λx,xhux, wx,yiWc(vx,y ⊗ ux) = hx, yiWc(x ⊗ x). then wx,y is linearly dependant with uy. vx,y ⊗ ux is self-adjoint, then vx,y is also linearly dependant with ux. Similarly by considering (x ⊗ y)(x ⊗ x), we get wx,y is linearly dependant with uy. So there is a complex λx,y such that Φ(x ⊗ y) = λx,yux ⊗ uy. If hx, yi 6= 0, λx,xhux,wx,yi hx,yi If x ⊥ y, hux, uyi = hx, yi = 0. For Wc(ux⊗uy) is an circle disc with diameter λx,y = 1 and hux, uyi = hx, yi. x)). If x 6⊥ y, considering Φ(x ⊗ y)Φ(x ⊗ x), we know λx,yλx,xhux, uyi = hx, yior − ( ¯c1 −eck)kuxkkuyk, we get λx,y = 1 from Wc(Φ(x ⊗ y)Φ(x ⊗ x)) = Wc((x ⊗ y)(x ⊗ hx, yi and pkxk2kyk2 − hx, yi2 = pλx,y2kuxk2kuyk2 − λx,yhux, uyi2. Hence In fact, between λx,x ∈ {−1, 1} and λx,x ∈ {i, −i}, only one holds for any x ∈ H. If not, assume there are x0 and y0 satisfying Φ(x0 ⊗ x0) = iux0 ⊗ ux0 and Φ(y0 ⊗ y0) = uy0 ⊗ uy0. Take z0 ∈ span[x0, y0]. Considering Wc(Φ(x0 ⊗ x0)Φ(z0 ⊗z0)), we get λz0 ∈ {i, −i}. However, we get λz0 ∈ {1, −1} by considering Wc(Φ(y0 ⊗ y0)Φ(z0 ⊗ z0)). That is a contradiction. Step 2.2. There are a unitary operator U : H → H and a function g : B(H) → {−1, 1} such that Φ(A) = g(A)U AU∗ for any A ∈ B(H) or Φ(A) = ig(A)U AU∗ for any A ∈ B(H). By Step 2.1, there are ux with kuxk = kxk such that Φ(x ⊗ x) = λx,xux ⊗ ux for any x ∈ H. Then define a map V : H → H by V x = ux and V (eiθx) = eiθV x for any θ ∈ [0, 2π). Next we show V is injective. Assume there is x, y such that V x = V y. By considering (x⊗x)(y⊗y), we get hy, xiWc(x⊗y) = hV y, V xiWc(V x⊗V y). then x = ty and t2 = 1 or t2 = −1. Since V (tx) = tV x, it deduces V x 6= V y for t 6= 1. So x = y. Because Φ is surjective and preserves rank-1 on both sides, we get V is surjective. Then there is a bijective map V with hV x, V yi = hx, yisuch that Φ(x ⊗ y) = λx,yV x ⊗ V y for any x, y ∈ H where λx,y ∈ T may be not same as above and we still use the same symbol. By Lemma 3.5, there are a unitary 14 YANFANG ZHANG, and XIAOCHUN FANG or anti-unitary operator U : H → H and θ : H → C with θ(x) = 1 such that V x = θ(x)U x. So Φ(x ⊗ y) = λx,yθ(x)θ(y)U x ⊗ U x for any x, y ∈ H. Let h(x, y) = θ(x)θ(y)λx,y, then Φ(x⊗y) = h(x, y)U x⊗U x for any x, y ∈ H. It is obvious that h(x, x) = λx,x. Next we show h(x, y) ∈ {−1, 1} for any x, y ∈ H or h(x, y) ∈ {−i, i} for any x, y ∈ H. Take z ∈ span[x, y] with hz, xi 6= 0 and hz, yi 6= 0. From Wc(Φ(z ⊗ z)Φ(x ⊗ y)) = Wc((z ⊗ z)(x ⊗ y)), we get h(z, z)h(x, y)Wc(U z ⊗ U x) = Wc(z ⊗ x). For Wc(z ⊗ x) is an ellipse but not a circle, h(z, z)h(x, y) = 1 or − 1. With the fact h(x, x) ∈ {−1, 1} for all x ∈ H or h(x, x) ∈ {−1, 1} for all x ∈ H, we get h(x, y) ∈ {−1, 1} for any x, y ∈ H or h(x, y) ∈ {−i, i} for any x, y ∈ H. Define a function f : F1(H) → {−1, 1} with f (x ⊗ y) = h(x, y), then either Φ(x ⊗ y) = f (x ⊗ y)U x ⊗ U y for any x, y ∈ H (3.1) or for any x, y ∈ H. Φ(x ⊗ y) = if (x ⊗ y)U x ⊗ U y (3.2) If (3.1) holds, for any A with rankA ≥ 2 and x 6∈ ker A, Wc((Φ(A)U x⊗U Ax) = Wc(Ax ⊗ Ax). It implies f (x ⊗ Ax)Φ(A)U x ⊗ U Ax is self-adjoint. Then there is tx,A ∈ {−1, 1} such that U∗Φ(A)U x = tx,AAx. In fact we can denote tx,A = tA because it is independent with x. For any x, y 6∈ ker A, U∗Φ(A)U x = tx,AAx and U∗Φ(A)U y = ty,AAy. On the other hand, U∗Φ(A)U(x + y) = tx+y,AA(x + y). So (ty,A − tx+y,A)Ay = (tx+y,A − tx,A)Ax. tx,A 6= ty,A will deduce Ax = 0 or Ay = 0, impossible. Define a function g : B(H) → {−1, 1} with g(A) =(cid:26) f (A) tA rankA = 1 rankA ≥ 2, then Φ(A) = g(A)U AU∗ for any A ∈ B(H). If (3.2) holds, −iΦ is accord with (3.1). Then Φ(A) = ig(A)U AU∗ for any A ∈ B(H). Similarly as Step 1.3 of the proof in Case I, it is impossible that U is anti- unitary. Step 2.3. Φ has the form in Theorem 1.2. When there exists an integer p (1 < p < k) satisfying cp + ck+1−p 6= 0 and cj + ck+1−j = 0 for j = 1, ..., p − 1, the form Φ(A) = ig(A)U AU∗ for A ∈ B(H) is impossible. If not, let D ∈ B(H) is a project with rank D = p. Wc(Φ(D)2) = Wc(D2) deduces −Wc(D) = Wc(D), which is contradict with Proposition 2.6. Remembering S = {A ∈ B(H) : Wc(A) = −Wc(A)} and T = {A ∈ S : AB ∈ S and BA ∈ S for any B ∈ B(H)}, we know Wc(Φ(I)Φ(B)) = Wc(B) for any operator B 6∈ S. Thus g(B) = g(I) = 1 or − 1 for any operator B ∈ B(H) \ S. For any T ∈ T , define the function φ : T → {−1, 1} with φ(T ) = g(T ). For any S ∈ S \ T , define the function φ′ : B(H) → {−1, 1} with φ′(S) = g(S). Because Wc(Φ(S)Φ(A)) = Wc(SA) for any A ∈ B(H), φ′ must satisfy φ′(S)φ′(A) = sgn(SA) for any A ∈ B(H) where sgn(SA) =(cid:26) {1} SA 6∈ S {−1, 1} SA ∈ S. The proof is finished. (cid:3) c-NUMERICAL RANGE OF OPERATOR PRODUCTS ON B(H) 15 Remark 3.6. For the form (iii) in Theorem 1.2, it is not easy to show all the elements in S and T . Here we show some subsets of them. Lemma 3.7. For c = (c1, ..., ck), if there exists an integer p (1 < p < k) sat- isfying cp + ck+1−p 6= 0 and cj + ck+1−j = 0 for j = 1, ..., p − 1. Let V1 = {T : T is a positive or non-positive operator with rank T = p } and V2 = {T : T is a self-adjoint operator with rank T < p}, then V1 ⊂ B(H) \ S, V2 ⊂ S, F1(H) ⊂ T and T 6= S. Proof. From Proposition 2.6(1),(2), we know V1 ⊂ B(H) \ S, V2 ⊂ S. For each rank-1 operator T and any operator A ∈ B(H), it comes Wc(AT ) = −Wc(AT ). Then F1(H) ⊂ T . Let D = Ip−1 ⊕ (−1) ⊕ 0 and G = Ip−1 ⊕ (−2) ⊕ 0. Wc(D) = cj −ck]. Obviously, Wc(D) = −Wc(D). Then D ∈ S. However [ ck+1−j −c1, DG is an positive operator with rank (DG) = p, then Wc(DG) 6= −Wc(DG). So D 6∈ T and T 6= S. (cid:3) p−1Pj=1 At the end, we finish the proof of Theorem 1.1. Proof of Theorem 1.1. Claim 1. Φ(A) = 0 ⇔ A = 0. If cj 6= 0, then Wc(A) = {0} if and only if A = 0 and the claim is true. If p−1Pj=1 kPj=1 kPj=1 kPj=1 cj = 0, Wc(T ) = {0} implies T is multiple of the identity. Assume Φ(A) = 0, then AB is a multiple of the identity for any B ∈ Mn(C). Let B run over all rank-1 matrices, we get AD = 0 for any rank-1 matrices. Because any matrix in Mn(C) is a combination of rank-1 matrices. Thus AB = 0 for all B ∈ Mn(C). Then A = 0. Claim 2. Φ preserves zero products on both sides. If cj 6= 0, obviously Φ preserves zero products. Next we show if cj = 0, Φ kPj=1 also preserves zero products. When cj = 0, On contrary assuming that there are non-zero A, B ∈ Mn(C) with AB = 0 such that Φ(A)Φ(B) = tI 6= 0. Then Φ(A) and Φ(B) are both invertible. Taking x, f ∈ Cn such that x ∈ ker A and f 6∈ ker A∗, we see Ax ⊗ f = 0 and x ⊗ f A 6= 0. This deduces Φ(A)Φ(x ⊗ f ) = sI where s ∈ C. If s 6= 0, then Φ(x ⊗ f )Φ(A) = tI and x ⊗ f A = 0, impossible. So AB = 0 implies Φ(A)Φ(B) = 0. If s = 0, we get Φ(x ⊗ f ) = 0, impossible. So from Theorem 1.2, Φ has the form in Theorem 1.1. (cid:3) References 1. Z.F. Bai, J.C. Hou, Numerical radius distance preserving maps on B(H), Proceedings of the American Mathematical Society, 132 (2004), no. 5, 1453 -- 1461. 2. M. Bendaoud, A. Benyouness, M. Sarih, et al, Preservers of radial unitary similarity func- tions on products of operators, Linear Algebra and Its Applications, 542 (2017), no. 5, 484-500. kPj=1 16 YANFANG ZHANG, and XIAOCHUN FANG 3. K. Chan, c -Numerical radius isometries on matrix algebras and triangular matrix algebras, Linear Algebra and Its Applications, 466 (2015), no. 2, 160-181. 4. C.Q. Chen, F.F. Lu, Maps that preserve higher-dimensional numerical ranges with operator Jordan products, Linear and Multilinear Algebra, 132 (2017), no. 12, 2530C2537. 5. J.L. Cui, J.C. Hou, Maps leaving functional values of operator products invariant, Linear Algebra and Its Applications, 428 (2008), no. 7, 1649-1663. 6. M.T. Chien, The c -numerical range of a rank-one matrix, Applied Mathematics Letters, 14 (2001), no.2, 167-170. 7. M.T. Chien, et al, On geometric properties of the numerical range, Linear Algebra and Its Applications, 274 (1998), no. 1, 389-410. 8. K.E. Gustafson, D.K.M. Rao, Numerical Range, Springer New York, 1997. 9. H.L. Gau, C.K. Li, C*-isomorphisms, Jordan isomorphisms, and numerical range preserv- ing maps, Proceedings of the American Mathematical Society, 135 (2007), no. 9, 2907-2914. 10. P.R. Halmos, A Hilbert Space Problem Book, Grad,Texts in Math., Spring-Verlag, New York, 1982. 11. J.C.Hou, Rank-preserving linear maps on B(H), Sci. in China (Ser.A), 32 (1989), no. 15, 929-940. 12. J.C. Hou, C.K. Li, X.F. Qi, Numerical Range of Lie Product of Operators, Integral Equa- tions and Operator Theory, 83 (2015), no. 4, 497-516. 13. J.C. Hou, Q.H. Di, Maps preserving numerical ranges of operator products, Proceedings of the American Mathematical Society, 134 (2006), no. 5, 1435-1446. 14. L.L. Liu, Linear transformations preserving log-convexity, Ars Combinatoria -Waterloo then Winnipeg-, 100 (2011), no. 11, 473-483. 15. C.K. Li, N.K.Tsing, Linear operators that preserve the c-numerical range or radius of ma- trices, Linear and Multilinear Algebra, 23 (1988), no. 1 , 27-46. 16. C.K. Li, The c-spectral, c-radial and c-convex matrices, Linear and Multilinear Algebra, 15 (1986), no. 1, 5-15. 17. C.K. Li, N.K. Tsing, N.S.Sze , Elliptical range theorems for generalized numerical ranges of quadratic operators, Rocky Mountain Journal of Mathematics, 41 (2011), no.3, 813-832. 18. L. Molnr, A generalization of Wigner's unitary-antiunitary theorem to Hilbert modules, Journal of Mathematical Physics, 40 (1999), no. 11, 5544-5555. 19. Y.T.Poon, Another proof of a result of Westwick, Linear and Multilinear Algebra, 9 (1980), no. 1, 181-186. 20. Westwick R, A theorem on numerical range, Linear and Multilinear Algebra, textbf2 (1975), no. 4, 311-315. 21. T. Schulte-herbrggen, Gunther Dirr, Uwe Helmke, et al, The Significance of the C- Numerical Range and the Local C-Numerical Range in Quantum Control and Quantum Information, Linear and Multilinear Algebra, 56 (2008), no. 1, 3-26. 1School of Mathematical Sciences, Tongji University, Shanghai 200092, P.R. China. E-mail address: [email protected] 2School of Mathematical Sciences, Tongji University, Shanghai 200092, P.R. China. E-mail address: [email protected]
1807.04012
1
1807
2018-07-11T09:05:52
On the existence of vectors dual to a set of linear functionals
[ "math.FA" ]
We give a simple proof of a crucial lemma that is established in [1, Lemma 2.1] by induction, and plays important roles in that paper and [2].
math.FA
math
ON THE EXISTENCE OF VECTORS DUAL TO A SET OF LINEAR FUNCTIONALS SHIBO LIU ABSTRACT. We give a simple proof of a crucial lemma that is established in [1, Lemma 2.1] by induction, and plays important roles in that paper and [2]. In a recent paper [1], Brezhneva and Tret'yakov give an elementary proof of the Karush -- Kuhn -- Tucker Theorem (that is, [1, Theorem 1.1]) in normed linear spaces. This theorem is about the minimization problems with a finite number of inequality constraints. In their proof of [1, Theorem 1.1], an important step is to establish the following lemma. Lemma 0.1 ([1, Lemma 2.1]). Let X be a linear space. Let ξi : X → R, i = 1, . . . , n, be linear functionals which are linearly independent. Then there exists a set {ηi} of n linearly independent vectors in X such that the matrix A =  · · · 〈ξ1, η1〉 ... 〈ξ1, ηn〉 ... 〈ξn, η1〉 · · · 〈ξn, ηn〉   is invertible. This lemma also plays a crucial role in [2], where an elementary proof of the La- grange multiplier theorem in normed linear spaces is given. We emphasize that 0.1 is equivalent to the following seemingly stronger result. Theorem 0.2. Let X be a vector space. Let ξi : X → R, i = 1, . . . , n, be linear functionals which are linearly independent. Then there exist ǫ1, . . . , ǫn in X such that 〈ξi, ǫ j〉 = δi j, where δi j is the Kronecker delta. In fact, let Then AA−1 = In, the n × n identity matrix, means A−1 =  〈ξi, ǫ j〉 =ξi, b1 1 ... bn 1 · · · · · · b1 n bn n   , n Xk=1 bk j ηk = n Xk=1 ǫ j = n Xk=1 bk j ηk 〈ξi, ηk〉bk j =AA−1i j = δi j. Therefore, Lemma 0.1 and Theorem 0.2 are equivalent. In [1], Lemma 0.1 is proved by induction on n. In this note, we present a much simple proof of Theorem 0.2. Proof of Theorem 0.2. We define a linear map ϕ : X → Rn by ϕ(x ) =ξ1(x ), . . . , ξn(x ) . 1991 Mathematics Subject Classification. 15A03, 15A04. Key words and phrases. linear functionals, linearly independent, linear map, perpendicular. This work was supported by NSFC (11671331). It was done while the author was visiting the Abdus Salam International Centre for Theoretical Physics. The author is grateful to ICTP for its hospitality. 1 2 SHIBO LIU We claim that ϕ is surjective. Otherwise, ϕ(X ) is a proper subspace of Rn, there exists λ ∈ Rn\ {0} which is perpendicular to ϕ(X ). Let λ = (λ1, . . . , λn), then for all x ∈ X , 0 = λ · ϕ(x ) = λiξi(x ). n Xi=1 Hence the set of linear functionalsξi is linearly dependent, a contradiction. Let e j = (0, . . . , 1, . . . , 0) be the j-th standard base vector of Rn and take ǫ j ∈ ϕ−1(e j), then It follows that ξi(ǫ j) = δi j. ξ1(ǫ j), . . . , ξn(ǫ j) = ϕ(ǫ j) = e j. REFERENCES (cid:3) [1] O. Brezhneva, A. A. Tret'yakov, An elementary proof of the Karush-Kuhn-Tucker theorem in normed linear spaces for problems with a finite number of inequality constraints, Optimization 60 (2011) 613 -- 618. [2] O. Brezhneva, A. A. Tret'yakov, An elementary proof of the Lagrange multiplier theorem in normed linear spaces, Optimization 61 (2012) 1511 -- 1517. SCHOOL OF MATHEMATICAL SCIENCES, XIAMEN UNIVERSITY, XIAMEN 361005, CHINA E-mail address: [email protected]
1703.02792
2
1703
2017-05-29T08:27:58
On sum of two subnormal kernels
[ "math.FA" ]
We show, by means of a class of examples, that if $K_1$ and $K_2$ are two positive definite kernels on the unit disc such that the multiplication by the coordinate function on the corresponding reproducing kernel Hilbert space is subnormal, then the multiplication operator on the Hilbert space determined by their sum $K_1+K_2$ need not be subnormal. This settles a recent conjecture of Gregory T. Adams, Nathan S. Feldman and Paul J. McGuire in the negative. We also discuss some cases for which the answer is affirmative.
math.FA
math
ON SUM OF TWO SUBNORMAL KERNELS SOUMITRA GHARA AND SURJIT KUMAR Abstract. We show, by means of a class of examples, that if K1 and K2 are two positive definite kernels on the unit disc such that the multiplication by the coordinate function on the corresponding reproducing kernel Hilbert space is subnormal, then the multiplication operator on the Hilbert space determined by their sum K1 + K2 need not be subnormal. This settles a recent conjecture of Gregory T. Adams, Nathan S. Feldman and Paul J. McGuire in the negative. We also discuss some cases for which the answer is affirmative. 1. Introduction Let H be a complex separable Hilbert space. Let B(H) denote the Banach algebra of bounded linear operators on H. Recall that an operator T in B(H) is said to be subnormal if there exists a Hilbert space K ⊃ H and a normal operator N in B(K) such that N (H) ⊂ H and NH = T. For the basic theory of subnormal operators, we refer to [12]. Completely hyperexpansive operators were introduced in [6]. An operator T ∈ B(H) is said to be completely hyperexpansive if n (−1)j(cid:18)n Xj=0 j(cid:19)T ∗jT j ≤ 0 (n ≥ 1). The theory of subnormal and completely hyperexpansive operators are closely related with the theory completely monotone and completely alternating sequences (cf. [5], [6] ). Let Z+ denote the set of non-negative integers. A sequence {ak}k∈Z+ of positive real numbers is said to be a completely monotone if n (−1)j(cid:18)n Xj=0 j(cid:19)am+j ≥ 0 (m, n ≥ 0). It is well-known that a sequence {ak}k∈Z+ is completely monotone if and only if the sequence {ak}k∈Z+ is a Hausdorff moment sequence, that is, there exists a positive measure ν supported in [0, 1] such that ak =R[0,1] xkdν(x) for all k ∈ Z+ (cf. [9]). Similarly, a sequence {ak}k∈Z+ of positive real numbers is said to be completely alternating if n (−1)j(cid:18)n Xj=0 j(cid:19)am+j ≤ 0 (m ≥ 0, n ≥ 1). 2010 Mathematics Subject Classification: Primary 46E20, 46E22; Secondary 47B20, 47B37 Key words and phrases: Completely alternating, Completely hyperexpansive, Completely monotone, Positive definite kernel, Spherically balanced spaces, Subnormal operators Work of the first author was supported by CSIR SPM Fellowship and work of the second author was supported by Inspire Faculty Fellowship. 1 ON SUM OF TWO SUBNORMAL KERNELS 2 Note that {ak}k∈Z+ is completely alternating if and only if the sequence {∆ak}k∈Z+ is completely monotone, where ∆ak := ak+1 − ak. open unit disc D := {z ∈ C : z < 1} with reproducing kernel K(z, w) = Pk∈Z+ Let H(K) be a reproducing kernel Hilbert space consisting of holomorphic functions on the ak(z ¯w)k. Thus K is holomorphic in the first variable and anti-holomorphic in the second. We make the assumption that a kernel function is always holomorphic in the first variable and anti- holomorphic in the second throughout this note. Consider the operator Mz of multiplication by the coordinate function z on H(K). As is well-known, such a multiplication operator is unitarily equivalent to a weighted shift operator W on the sequence space xk2 ak ℓ2(a) := {x := (x0, x1, . . .) : kxk2 = < ∞} ∞ Xk=0 an+1 with W en = q an en+1, where en is the standard unit vector. Assume that Mz is bounded. Then Mz is a contractive subnormal if and only if the sequence { 1 ak }k∈Z+ is a Hausdorff moment sequence. On the other hand, Mz on H(K) is completely hyperexpansive if and only if the sequence { 1 ak }k∈Z+ is completely alternating (cf. [6, Proposition 3]). For any two positive definite kernels K1 and K2, their sum K1 +K2 is again a positive definite kernel and therefore determines a Hilbert space H(K1 + K2) of functions. It was shown in [4] that H(K1 + K2) = {f = f1 + f2 : f1 ∈ H(K1), f2 ∈ H(K2)}, H(K1+K2) := infnkf1k2 is a Hilbert space with the norm given by kfk2 H(K1) + kf2k2 Sum of two kernel functions is also discussed by Salinas in [19]. He proved that if K1 and K2 are generalized Bergman kernels (for definition, refer to [16] ), then so is K1 + K2. Although not explicitly stated in [4], it is not hard to verify that the multiplication operator Mz on H(K1 + K2) is unitarily equivalent to the operator PM⊥(Mz,1 ⊕ Mz,2)M⊥ , where Mz,i is the operator of multiplication by the coordinate function z on H(Ki) and H(K2) : f = f1 + f2, f1 ∈ H(K1), f2 ∈ H(K2)o . M = {(g,−g) ∈ H(K1) ⊕ H(K2) : g ∈ H(K1) ∩ H(K2)} ⊆ H(K1) ⊕ H(K2). Evidently, if Mz,1 and Mz,2 are subnormal, then so is Mz,1 ⊕ Mz,2. Here, we discuss the subnor- mality of the compression PM⊥(Mz,1 ⊕ Mz,2)M⊥ for a class of kernels. In particular, we show that the subnormality of Mz,1 and Mz,2 need not imply PM⊥(Mz,1 ⊕ Mz,2)M⊥ is subnormal. A similar question on subnormality involving the point-wise product of two positive definite kernels was raised in [19]. Recall that the product K1K2 of two positive definite kernels defined on, say the unit disc D, is also a positive definite kernel on D. Indeed, if H(K1) ⊗ H(K2) ⊆ Hol(D × D) be the usual tensor product of the two Hilbert spaces H(K1) and H(K2), and N is the subspace {h ∈ H(K1)⊗H(K2) : h(z, z) = 0, z ∈ D}, then the operator Mz,1⊗I acting on the Hilbert space H(K1) ⊗ H(K2) compressed to N⊥ is unitarily equivalent to the multiplication operator Mz on the Hilbert space H(K1K2). If Mz,1 is subnormal on H(K1), then so is the operator Mz,1 ⊗ I. ON SUM OF TWO SUBNORMAL KERNELS 3 The answer to the question of subnormality, both in the case of the sum as well as the product, is affirmative in several examples. For over thirty years, the question of whether the compression to N⊥ is subnormal had remained open. Recently, a counter-example has been found, see [3, Theorem 1.5]. The conjecture below is similar except that it involves the sum of two kernels. Conjecture 1.1. ([1, pp. 22]). Let K1(z, w) =Pk∈Z+ be any two reproducing kernels satisfying: ak(z ¯w)k and K2(z, w) =Pk∈Z+ bk(z ¯w)k = lim bk bk+1 (a) lim ak ak+1 (b) lim ak = lim bk = ∞ (c) 1 ak = 1 =R[0,1] tkdν1(t) and 1 bk =R[0,1] tkdν2(t) for all k ∈ Z+. Then the multiplication operator Mz on H(K1 + K2) is a subnormal operator. An equivalent formulation, in terms of the moment sequence criterion, of the conjecture is bk }k∈Z+ are Hausdorff moment sequences, does it necessarily ak }k∈Z+ and { 1 ak+bk }k∈Z+ is also a Hausdorff moment sequence? the following. If { 1 follow that { 1 Like the case of the product of two kernels, here we give a class of counter examples to the conjecture stated above. Paul McGuire, in a private communication, has informed the authors of an example that they had found. In fact, he says that their example is one of the examples discussed in this note. These two cases suggest that it may be fruitful to ask when the compression of a subnormal operator to an invariant subspace is again subnormal. The paper is organized as follows. In section 2, we provide a class of counter-examples which settles the Conjecture 1.1. We also discuss some cases for which answer to this conjecture is affirmative. In the last section, we try to answer analogously in a certain class of weighted multi-shifts. 2. Sum of two subnormal reproducing kernels need not be subnormal For the construction of counter-examples to the conjecture, we make use of the following result, borrowed from [2, Proposition 4.3]. Proposition 2.1. For distinct positive real numbers a0, ..., an and non-zero real numbers b0, ..., bn, consider the polynomial p(x) = Πn p(l)}l∈Z+ is never a Hausdorff moment sequence. k=0(x + ak + ibk)(x + ak − ibk). Then the sequence { 1 For r > 0, let Kr be a positive definite kernel given by Kr(z, w) := Xk∈Z+ k + r r (z ¯w)k (z, w ∈ D). The case r = 1 corresponds to the Bergman kernel. It is easy to see that the multiplication operator Mz on H(Kr) is a contractive and subnormal and the representing measure is rxr−1dx. ON SUM OF TWO SUBNORMAL KERNELS 4 For s, t > 0, consider the multiplication operator Mz on H(Ks,t), where Ks,t(z, w) := Xk∈Z+ (k + s)(k + t) st (z ¯w)k (z, w ∈ D). The case s = 1 and t = 2, corresponds to the kernel (1 − z ¯w)−3. Note that Mz is a contractive subnormal with the representing measure ν is given by dν(x) = (cid:26) −s2xs−1 log x dx if s = t if s 6= t. st xt−1−xs−1 dx s−t One easily verifies that Kr and Ks,t both satisfy all the conditions (a), (b) and (c) of the Conjecture 1.1. But the multiplication operator on their sum need not be subnormal for all possible choices of s, t > 0. This follows from the following theorem. Theorem 2.2. The multiplication operator Mz on H(Kr + Ks,t) is subnormal if and only if (2.1) Proof. Notice that (rs + st + tr)2 ≥ 8r2st. (Kr + Ks,t)(z, w) = Xk∈Z+ k2 + (s + t + st r )k + 2st st (z ¯w)k (z, w ∈ D). The roots of the polynomial x2 + (s + t + st r )x + 2st are := −(s + t + st := −(s + t + st r ) +q(s + t + st r ) −q(s + t + st 2 2 r )2 − 8st r )2 − 8st and . x1 x2 Suppose that (rs + st + tr)2 ≥ 8r2st. Then the kernel Kr + Ks,t will be of the form 2Ks′,t′ where s′ = −x1 and t′ = −x2. Hence, Mz on H(Kr + Ks,t) is a subnormal operator. Conversely, assume that (rs + st + tr)2 < 8r2st. By Proposition 2.1, it follows that Mz on (cid:3) H(Kr + Ks,t) can not be subnormal. Remark 2.3 : If we choose s = 1, t = 2 and r > 2, then the inequality (2.1) is not valid. We also point out that if K1 and K2 are any two reproducing kernels such that the multi- plication operators on H(K1) and H(K2) are hyponormal, then the multiplication operator on H(K1 + K2) need not be hyponormal. An example illustrating this is given below. Example 2.4. For any s, t > 0, consider the reproducing kernel K s,t given by K s,t(z, w) := 1 + sz ¯w + s2(z ¯w)2 + t (z ¯w)3 1 − z ¯w . Note that K s,t defines a reproducing kernel on the unit disc D and the multiplication operator t , 1, 1,··· ). Mz on H(K s,t) can be seen as a weighted shift operator with weights (q 1 s ,q 1 Thus, it follows that Mz on H(K s,t) is hyponormal if and only if s2 ≤ t ≤ s3. s ,q s2 ON SUM OF TWO SUBNORMAL KERNELS 5 Observe that Mz on H(K s,t + K s′,t′ s+s′ ,q s+s′ (q 2 necessary that s2+s′2 ,q s2+s′2 s+s′ ≤ s+s′ 2 ) can be realized as a weighted shift operator with weights , 1, 1,··· ). For the hyponormality of this weighted shift operator, it is t+t′ s2+s′2 , which is true only when s = s′. We remark that this is different from the case of the product of two kernels, where, the hyponormality of the multiplication operator on the Hilbert space H(K1K2) follows as soon as we assume they are hyponormal on the two Hilbert spaces H(K1) and H(K2), see [10]. If T ∈ B(H) is left invertible then the operator T ′ given by T ′ = T (T ∗T )−1 is said to be the operator Cauchy dual to T. The following result has been already recorded in [6, Proposition 6], which may be paraphrased as follows: Theorem 2.5. Let K(z, w) =Pk∈Z+ ak(z ¯w)k be a positive definite kernel on D and Mz be the multiplication operator on H(K). Assume that Mz is left invertible. Then the followings are equivalent: (i) {ak}k∈Z+ is a completely alternating sequence. (ii) The Cauchy dual M′z of Mz is completely hyperexpansive. (iii) For all t > 0, { (iv) For all t > 0, the multiplication operator Mz on H(tK + (1− t)S) is contractive subnor- t(ak−1)+1}k∈Z+ is a completely monotone sequence. 1 mal, where S is the Szego kernel on D. Remark 2.6 : If {ak}k∈Z+ is a completely alternating sequence then by putting t = 1 in part (iii) of Theorem 2.5, it follows that { 1 bk(z ¯w)k be any two Corollary 2.7. Let K1(z, w) = Pk∈Z+ reproducing kernels such that {ak}k∈Z+ and {bk}k∈Z+ are completely alternating sequences, then the multiplication operator Mz on H(K1 + K2) is subnormal. ak(z ¯w)k and K2(z, w) = Pk∈Z+ ak }k∈Z+ is a completely monotone sequence. Proof. It is easy to verify that the sum of two completely alternating sequences is completely alternating. The desired conclusion follows immediately from Remark 2.6. (cid:3) Remark 2.8 : Note that { k+r { (k+s)(k+t) r }k∈Z+ is a completely alternating sequence but the sequence }k∈Z+ is not completely alternating. So, the reproducing kernels Kr and Ks,t dis- st cussed in Theorem 2.2 does not satisfy the hypothesis of Corollary 2.7. Proposition 2.9. Let K(z, w) = Pk∈Z+ ak(z ¯w)k be any positive definite kernel such that the multiplication operator Mz on H(S + K) is subnormal. Then the multiplication operator on H(K) is subnormal. Proof. From subnormality of Mz on H(S + K), it follows that { 1 monotone sequence. Thus, {1 − 1 1+ak}k∈Z+ is completely alternating. Note that 1+ak }k∈Z+ is a completely (ak)−1 = (1 + ak)−1(1 − 1 1+ak )−1 = ∞ Xj=1(cid:16) 1 1+ak(cid:17)j . ON SUM OF TWO SUBNORMAL KERNELS 6 Observe that { (1+ak)j }k∈Z+ is a completely monotone sequence for all j ≥ 1. Now, being the limit of completely monotone sequences, {a−1 k }k∈Z+ is completely monotone. 1 (cid:3) Remark 2.10 : We have following remarks: (i) The converse of the Proposition 2.9 is not true (see the example discussed in part (ii) of the Remark 2.13). (ii) If we replace Szego kernel S by Bergman kernel then the conclusion of the Proposition 2.9 need not be true. For example, by using Proposition 2.1, one may choose α > 0 k2+αk+1}k∈Z+ is not completely monotone but the sequence such that the sequence { { k2+(α+1)k+2}k∈Z+ is completely monotone. 1 1 We use the convenient Pochhammer symbol given by (x)n := Γ(x+n) Γ(x) , where Γ denotes the gamma function. For λ, µ > 0, consider the positive definite kernel Kλ,µ(z, w) = Xk∈Z+ (λ)k (µ)k (z ¯w)k (z, w ∈ D). It is easy to see that the case µ = 1 corresponds to the kernel (1 − z ¯w)−λ. Note that the multiplication operator Mz on H(Kλ,µ) may be realized as a weighted shift operator with weight sequence {q k+µ First part of the following theorem is proved in [15] and the representing measure is given in k+λ}k∈Z+. [13, Lemma 2.2]. Here, we provide a proof for the second part only. Theorem 2.11. The multiplication operator Mz on H(Kλ,µ) is (i) subnormal if and only if λ ≥ µ. In this case, the representing measure ν of Mz is given by dν(x) = ( Γ(λ) Γ(µ)Γ(λ−µ) xµ−1(1 − x)λ−µ−1dx if λ > µ if λ = µ, δ1(x)dx where δ1 is the Dirac delta function. (ii) completely hyperexpansive if and only if λ ≤ µ ≤ λ + 1. Proof. The multiplication operator Mz on H(Kλ,µ) is completely hyperexpansive if and only if the sequence { (µ)k (λ)k }k∈Z+ is completely alternating. Here ∆(cid:18) (µ)k (λ)k(cid:19) = (µ)k+1 (λ)k+1 − (µ)k (λ)k = = (µ)k+1 − (µ)k(λ + k − 1) (λ)k+1 (µ)k . (λ + 1)k µ − λ λ By first part of this theorem, { µ−λ if λ ≤ µ ≤ λ + 1. This completes the proof. λ (µ)k (λ+1)k }k∈Z+ is a completely monotone sequence if and only (cid:3) ON SUM OF TWO SUBNORMAL KERNELS 7 The following proposition gives a sufficient condition for the subnormality of multiplication operator on Hilbert space determined by sum of two kernels belonging to the class Kλ,µ. Proposition 2.12. Let 0 < µ ≤ λ′ ≤ λ ≤ λ′ + 1. Then the multiplication operator Mz on H(Kλ,µ + Kλ′,µ) is contractive subnormal. Proof. Observe that (Kλ,µ + Kλ′,µ)(z, w) = Xk∈Z+ (λ)k + (λ′)k (µ)k (z ¯w)k (z, w ∈ D) and (µ)k (λ)k + (λ′)k = (µ)k (λ′)k 1 1 + (λ)k (λ′)k . Since µ ≤ λ′, it follows from part (i) of Theorem 2.11 that { (µ)k If λ′ ≤ λ ≤ λ′ + 1, then by part (ii) of Theorem 2.11, the sequence { (λ)k (λ′)k}k∈Z+ is completely monotone. (λ′)k}k∈Z+ is a completely alternating sequence. So is the sequence {1 + (λ)k (λ′)k }k∈Z+. Hence, by Remark 2.6,( 1 1+ (λ′)k)k∈Z+ (λ)k is a completely monotone sequence. Thus, being a product of two completely monotone se- quences, { Remark 2.13 : Here are some remarks: (λ)k+(λ′)k}k∈Z+ is a completely monotone sequence. This completes the proof. (µ)k (cid:3) (i) The case when µ < λ′ and λ = λ′ + 1. The representing measure for the sequence ( 1 1+ (λ′)k)k∈Z+ (λ)k is λ′x2λ′−1dx. The representing measure for the sequence { (µ)k (λ′)k }k∈Z+ is given in part (i) of Theorem 2.11. Thus, using Remark 2.4 of [2], one may obtain the representing measure for Mz on H(Kλ,µ + Kλ′,µ) to be given by dν(x) = λ′Γ(λ′) Γ(µ)Γ(λ′ − µ) x2λ′−1(cid:18)Z 1−x 0 tλ′−µ−1(1 − t)µ−2λ′−1dt(cid:19) dx. But in general, when λ < λ′ + 1, we do not know the representing measure for the sequence ( 1 1+ (λ′)k)k∈Z+ (λ)k as well as for the sequence n (µ)k (ii) Consider the kernel (K1,1+K3,1)(z, w) =Pk∈Z+ from Proposition 2.1 that the sequence n k2+3k+4ok∈Z+ 2 2 k2+3k+4 (λ)k+(λ′)kok∈Z+. (z ¯w)k for all z, w ∈ D. It follows is not completely monotone. Consequently, the multiplication operator Mz on H(K1,1 + K3,1) is not subnormal. (1)k For λ > 1, consider the kernel Kλ,1 + K3,1. We claim that there exists a λ0 > 1 such is not completely monotone. If not, assume that it is a completely (λ0)k+(3)kok∈Z+ thatn monotone sequence for all λ > 1. As λ goes to 1, one may get that n completely monotone, which is a contradiction. Therefore, we conclude that there exists a λ0 > 1 such that the multiplication operator Mz on H(Kλ0,1 + K3,1) is not subnormal. By using properties of gamma function, one may verify that Kλ0,1 and K3,1 both satisfy k2+3k+4ok∈Z+ is 2 ON SUM OF TWO SUBNORMAL KERNELS 8 (a), (b) and (c) of the Conjecture 1.1. This also provides a class of counterexamples for the Conjecture 1.1. ap k(z ¯w)k and K2(z, w) = aq k(z ¯w)k are any two reproducing kernels such that {ak}k∈Z+ is a completely alternating Proposition 2.14. Let 0 < p ≤ q ≤ p + 1. Suppose K1(z, w) =Pk∈Z+ Pk∈Z+ sequence. Then the multiplication operator Mz on H(K1 + K2) is subnormal. Proof. Note that 1 ap k + aq k = 1 ap k(1 + aq−p k ) . k }k∈Z+ is also completely alternating. Thus so is {1 + aq−p Since 0 ≤ q − p ≤ 1 and {ak}k∈Z+ is completely alternating, it follows from [7, Corollary 1] that {aq−p }k∈Z+ . Hence, by Remark 2.6, {(1 + aq−p k }k∈Z+ is completely monotone. Now the proof follows as the product of two completely monotone sequences is also completely monotone. )−1}k∈Z+ is completely monotone. By [8, Corollary 4.1], {a−p (cid:3) k k Example 2.15. For any p > 0, let Kp(z, w) be the positive definite kernel given by Kp(z, w) := Xk∈Z+ (k + 1)p(z ¯w)k (z, w ∈ D). Then it is known that the multiplication operator Mz on H(Kp) is subnormal with the repre- senting measure dν(x) = (− log x)p−1 dx (cf. [14, Theorem 4.3]). By Proposition 2.14, it follows that Mz on H(Kp + Kq) is subnormal if p ≤ q ≤ p + 1. Γ(p) The next result also provides a class of counter-examples to the Conjecture 1.1. Theorem 2.16. Consider the positive definite kernel Kp given in Example 2.15. Then the multiplication operator Mz on H(Kp + Kp+2) is subnormal if and only if p ≥ 1. Proof. For x ∈ (0, 1], let g(x) := 1 1 (− log x − y)p−1 sin y dy and dν(x) = g(x)dx. Then 0 xkdν(x) = Z 1 0 Γ(p)R − log x Γ(p)Z ∞ Γ(p)Z ∞ Γ(p)Z ∞ y=0 1 1 1 x=0 y=0Z e−y e−(k+1)yZ ∞ e−(k+1)y Γ(p) u=0 y=0 (k + 1)p (k + 1)2 + 1 1 . = = = 1 xk(− log x − y)p−1dx sin y dy e−(k+1)uup−1du sin y dy (k + 1)p sin y dy 1 (k+1)p Thus the sequence { (k+1)2+1}k∈Z+ is completely monotone if and only if the function g(x) is non-negative a.e. Note that the function g(x) is non-negative on (0, 1] a.e. if and only if the function h(x) := g(e−x) is non-negative a.e. on (0,∞). Now Γ(p)Z 1 Γ(p)Z x h(x) = xp 1 0 0 By [17, Chapter 3, pp 439], we have h(x) = function of first kind. Thus, the sequence { (k+1)p 1 2 1 2 , 1 (x) is the Lommel's p− (k+1)2+1}k∈Z+ being completely monotone is p− 1 2 , 1 2 (x − y)p−1 sin y dy = √x Γ(p) s 1 (1 − y)p−1sin(xy) dy. (x), where s ON SUM OF TWO SUBNORMAL KERNELS 9 equivalent to the non-negativity of the function s (x). If p ≥ 1 then by [20, Theorem A], (x) ≥ 0 for all x > 0. The converse follows from [20, Theorem 2], which p− 1 2 , 1 2 we get that s 1 2 , 1 2 p− completes the proof. 3. Multi-variable case (cid:3) Let Zd + denote the cartesian product Z+ × ··· × Z+ (d times). Let α = (α1,··· , αd) ∈ Zd +, we write α := α1 + ··· + αd and α! = α1!··· αd!. If T = (T1,··· , Td) is a d-tuple of commuting bounded linear operators Tj (1 ≤ j ≤ d) on H then we set T ∗ to be (T ∗1 ,··· , T ∗d ) and T α to be T α1 1 ··· T αd d . Given a commuting d-tuple T of bounded linear operators T1,··· , Td on H, set QT (X) := d Xi=1 T ∗i XTi (X ∈ B(H)). For X ∈ B(H) and k ≥ 1 , one may define Qk Recall that T is said to be T (X) := QT (Qk−1 T (X)), where Q0 T (X) = X. (i) spherical contraction if QT (I) ≤ I. (ii) jointly left invertible if there exists a positive number c such that QT (I) ≥ cI. For a jointly left invertible T, the spherical Cauchy dual T s of T is the d-tuple (T s 1 , T s d ), 2 ··· , T s i := Ti(QT (I))−1 (i = 1, 2,··· , d). We say that T is a joint complete hyperexpansion if where T s Bn(T ) := n (−1)k(cid:18)n Xk=0 k(cid:19)Qk T (I) ≤ 0 (n ≥ 1). Throughout this section B denotes the open unit ball {z ∈ Cd : z12 + ··· + zd2 < 1} and be a multi-sequence of positive numbers. Consider the Hilbert space H 2(β) of ∂B denotes the unit sphere {z ∈ Cd : z12 + ··· + zd2 = 1} in Cd. Let {βα}α∈Zd formal power series f (z) =Pα∈Zd kfk2 f (α)zα such that f (α)2β2 α < ∞. + + H 2(β) = Xα∈Zd + The Hilbert space H 2(β) is said to be spherically balanced if the norm on H 2(β) admits the slice representation [ν, H 2(γ)], that is, there exist a Reinhardt measure ν and a Hilbert space H 2(γ) of formal power series in one variable such that kfk2 H 2(β) =Z∂B kfzk2 H 2(γ)dν(z) (f ∈ H 2(β)), where γ = {γk}k∈Z+ is given by the relation βα = γαkzαkL2(∂B,ν) for all α ∈ Zd +. Here, by the Reinhardt measure, we mean a Td-invariant finite positive Borel measure supported in ∂B, where Td denotes the the unit d-torus {z ∈ Cd : z1 = 1,··· ,zd = 1}. For more details on spherically balanced Hilbert spaces, we refer to [11]. ON SUM OF TWO SUBNORMAL KERNELS 10 The following lemma has been already recorded in [11, Lemma 4.3]. We include a statement for ready reference. Lemma 3.1. Let H 2(β) be a spherically balanced Hilbert space and let [ν, H 2(γ)] be the slice rep- resentation for the norm on H 2(β). Consider the d-tuple Mz = (Mz1,··· , Mzd) of multiplication by the co-ordinate functions z1,··· , zd on H 2(β). Then for every n ∈ Z+ and α ∈ Zd +, L2(∂B,ν). n n hBn(Mz)zα, zαi = Mz (I)zα, zαi = (−1)k(cid:18)n Xk=0 k(cid:19)γ2 k+αkzαk2 (−1)k(cid:18)n k(cid:19)hQk Xk=0 If the interior of the point spectrum σp(M∗z ) of M∗z is non-empty then H 2(β) may be realized as a reproducing kernel Hilbert space H(K) [18, Propositions 19 and 20], where the reproducing kernel K is given by K(z, w) = Xα∈Zd + zα ¯wα β2 α (z, w ∈ σp(M∗z )). This has lead to the following definition. Definition 3.2 : Let H(K) be a reproducing kernel Hilbert space defined on the open unit ball aαzα ¯wα for all z, w ∈ B. We say that K is balanced kernel if H(K) is a spherically balanced Hilbert space. Further, the multiplication d-tuple Mz on H(K) may be called as balanced multiplication tuple. B with reproducing kernel K(z, w) =Pα∈Zd + Remark 3.3 : The spherical Cauchy dual M s tuple Mz can be seen as a multiplication d-tuple M s co-ordinate functions z1,··· , zd on H 2(β s), where kzαkL2(∂B,ν) β s α = 1 γα (α ∈ Zd +). z of a jointly left invertible balanced multiplication zd) of multiplication by the z = (M s z1,··· , M s In other words, the norm on H 2(β s) admits the slice representation [ν, H 2(γ′)], where γ′k = 1/γk for all k ∈ Z+. bαzα ¯wα are any two Proposition 3.4. If K1(z, w) = Pα∈Zd balanced kernels with the slice representations [ν, H 2(γ1)] and [ν, H 2(γ2)] respectively. Then K1 + K2 is a balanced kernel with the slice representation [ν/2, H 2(γ)], where γ = {γk} is given by the relation aαzα ¯wα and K2(z, w) = Pα∈Zd + + γk = √2γk,1γk,2 (γ2 k,1 + γ2 k,2)1/2 (k ∈ Z+). Proof. For every α ∈ Zd +, we have aα + bα = Therefore 1 γ2 α,1kzαk2 L2(∂B,ν) + 1 γ2 α,2kzαkL2(∂B,ν) = α,1 + γ2 (γ2 α,2kzαk2 γ2 α,1γ2 α,2) L2(∂B,ν) . kzαk2 H(K1+K2) = α,1γ2 2γ2 α,2 α,1 + γ2 (γ2 α,2)kzαk2 L2(∂B,ν/2) = γ2 αkzαk2 L2(∂B,ν/2) ON SUM OF TWO SUBNORMAL KERNELS 11 for all α ∈ Zd immediately. +. Since {zα}α∈Zd + forms an orthogonal subset of L2(∂B, ν/2), the conclusion follows (cid:3) Remark 3.5 : The conclusion of the Proposition 3.4 still holds even if we chose two different Reinhardt measures ν1 and ν2 in the slice representations of K1 and K2, such that for some sequence of positive real numbers {hk}k∈Z+, kzαkL2(∂B,ν1) = hαkzαkL2(∂B,ν2) for all α ∈ Zd +. For every j = 1, 2, it is easy to verify that d Xi=1 kzα+εik2 kzαk2 L2(∂B,νj) L2(∂B,νj ) = 1. This implies that {hk}k∈Z+ is a constant sequence, say c. Now, a routine argument, using the Stone-Weierstrass theorem, we conclude that µ1 = c2µ2. Let Kν denote the class of all balanced kernels with the following properties: (i) For all K ∈ Kν, the norm on H(K) admits the slice representations with fixed Reinhardt measure ν. (ii) For every member K of Kν, the multiplication operator Mz defined on H(K) is jointly left invertible. (iii) The Cauchy dual tuple M s z of is a joint complete hyperexpansion. Lemma 3.6. For every member K of Kν, the multiplication operator tuple Mz defined on H(K) is a subnormal spherical contraction. Proof. Let K ∈ Kν and [ν, H 2(γ)] be the slice representation for the norm on H(K). Note z of Mz is a balanced multiplication tuple with slice representation that the Cauchy dual M s [ν, H 2(1/γ)] (see Remark 3.3). Since M s z is a joint complete hyperexpansion. It follows from Lemma 3.1 that {1/γ2 k}k∈Z+ is a completely alternating sequence. Therefore, by Remark 2.6, {γ2 k}k∈Z+ is completely monotone sequence. Now again by applying Lemma 3.1, we conclude that the multiplication operator Mz is a subnormal spherical contraction. (cid:3) Theorem 3.7. If K1 and K2 are any two members of Kν then the multiplication operator Mz on H(K1 + K2) is a subnormal spherical contraction. k,1γ2 γ2 k,2 k,1+γ2 γ2 Proof. Note that the norm on H(K1 + K2) admits the slice representation [ν/2, H 2(γ)], where γ2 for all k ∈ Z+ (see Proposition 3.4). It follows from the proof of Lemma 3.6 that k = 2 {1/γ2 k}k∈Z+ is a completely alternating sequence. Now the conclusion follows by imitating the argument given in Lemma 3.6. k,2}k∈Z+ are completely alternating. So their sum, that is, {1/γ2 k,1}k∈Z+ and {1/γ2 (cid:3) k,2 For λ > 0, consider the positive definite kernel Kλ given by Kλ(z, w) = 1 (1 − hz, wi)λ The norm on H(Kλ) admits the slice representation [σ, H 2(γ)], where σ denotes the normalized surface area measure on ∂B and γ2 for all k ∈ Z+. It is well known that the multiplication k = (d)k (λ)k (z, w ∈ B). ON SUM OF TWO SUBNORMAL KERNELS 12 operator Mz,λ on H(Kλ) is a subnormal contraction if and only if λ ≥ d. The same can also be verified by using Lemma 3.1 and part (i) of Theorem 2.11. Similarly, by using Lemma 3.1 and part (ii) of Theorem 2.11, one may conclude that the Cauchy dual tuple M s z,λ is a joint complete hyperexpansion if and only if d ≤ λ ≤ d + 1. Thus, if we choose λ and λ′ are such that d ≤ λ, λ′ ≤ d + 1. Then Kλ and Kλ′ ∈ Kσ. It now follows from Theorem 3.7 that the multiplication operator Mz on H(Kλ + Kλ′) is subnormal. This is also included in the following example. Example 3.8. Let 0 < d ≤ λ′ ≤ λ ≤ λ′ + 1. Note that the norm on H(Kλ + Kλ′) admits the slice representation [σ/2, H 2(γ)], where γ2 for all k ∈ Z+. From the proof of Proposition 2.12, it is clear that {γ2 k}k∈Z+ is completely monotone. Hence, the multiplication operator Mz on H(Kλ + Kλ′) is subnormal. k = 2(d)k (λ)k+(λ′)k A d-tuple S = (S1,··· , Sd) of commuting bounded linear operators S1,··· , Sd in B(H) is a spherical isometry if S∗1 S1 + ··· + S∗dSd = I. In other words, QS(I) = I. The most interesting example of a spherical isometry is the Szego d-shift; that is, the d-tuple Mz of multiplication operators Mz1,··· , Mzd on the Hardy space H 2(∂B) of the unit ball. Let ν be a Reinhardt measure. Consider the multiplication d-tuple Mz on a reproducing kernel Hilbert space H(K ν ) determined by the reproducing kernel (3.2) K ν(z, w) = Xα∈Zd + zα ¯wα kzαk2 L2(∂B,ν) (z, w ∈ B). Note that Mz is a spherical isometry. In this case, the norm on H(K ν ) admits the slice repre- sentation [ν, H 2(D)], where H 2(D) is the Hardy space of the unit disc. Theorem 3.9. Let K ν be the reproducing kernel given as in equation (3.2) and K be any balanced kernel with the slice representation [ν, H 2( γ)]. Assume that the multiplication operator Mz on H(K ν + K) is subnormal. Then the multiplication operator on H( K) is subnormal. Proof. Observe that the norm on H(K ν + K) admits the slice representation [ν/2, H 2(γ)], k )−1 for all k ∈ Z+. Since Mz on H(K ν + K) is subnormal, it follows from k = 2(1 + 1/γ2 where γ2 k)−1}k∈Z+ is k}k∈Z+ is a completely monotone sequence. Hence, {(1 + 1/γ2 Lemma 3.1 that {γ2 a completely monotone sequence. If we replace ak by 1/γ2 k in the proof of the Proposition 2.9, we get that {γ2 k}k∈Z+ is completely monotone. Now, by applying Lemma 3.1, we conclude that the multiplication operator on H( K) is subnormal. (cid:3) We conclude the paper with the following questions: Question 3.10. In view of Proposition 2.12 and Theorem 2.16, it is natural to ask that (i) what is the necessary and sufficient condition for the multiplication operator Mz on H(Kλ,µ + Kλ′,µ) to be subnormal? (ii) what is the necessary and sufficient condition for the multiplication operator Mz on H(Kp + Kq) to be subnormal? ON SUM OF TWO SUBNORMAL KERNELS 13 Question 3.11. Let K ν be the reproducing kernel given as in equation (3.2) and K be any positive definite kernel given by K(z, w) := Xα∈Zd + aαzα ¯wα (z, w ∈ B). Assume that the d-tuple Mz = (Mz1,··· , Mzd) of multiplication by the co-ordinate functions z1,··· , zd on H(K ν + K) is subnormal. Is it necessary that the multiplication operator on H( K) subnormal? Acknowledgments. We express our sincere thanks to Prof. G. Misra for many fruitful conver- sations and suggestions in the preparation of this paper. We would also like to thank Prof. S. Chavan for his many useful comments and careful reading of the manuscript. References [1] G. T. Adams, N. S. Feldman and P. J. McGuire, Tridiagonal reproducing kernels and subnormality, J. Operator Theory 70, no. 2 (2013), 477-494. [2] A. Anand and S. Chavan, A moment problem and joint q-isometry tuples, Complex Anal. Oper. Theory 11 (2017), 785-810. [3] A. Anand and S. Chavan, Module tensor product of subnormal modules need not be subnormal, J. Funct. Anal. 272 (2017), 4752-4761. [4] N. Aronszajn, Theory of reproducing kernels, Trans. Amer. Math. Soc. 68 (1950), 337-404. [5] A. Athavale, Holomorphic kernels and commuting operators, Trans. Amer. Math. Soc. 304, no. 1 (1987), 101-110. [6] A. Athavale, On completely hyperexpansive operators, Proc. Amer. Math. Soc. 124 (1996), 3745-3752. [7] A. Athavale and A. Ranjekar, Bernstein functions, complete hyperexpansivity and subnormality-I, Integr. Equ. Oper. Theory 43, no. 3 (2002), 253-263. [8] C. Benhida, R. Curto and G. Exner, Moment infinitely divisible weighted shift, http://homepage.divms.uiowa.edu/∼rcurto/MID.pdf [9] C. Berg, J. P. R. Christensen and P. Ressel, Harmonic Analysis on Semigroups, Springer-Verlag, Berlin 1984. [10] M. Badri and P. Szeptycki,Cauchy products of positive sequences Rocky Mountain J. Math. 20 (1990), 351-357. [11] S. Chavan and S. Kumar, Spherically balanced Hilbert spaces of formal power series in several variables-I, J. Operator Theory 72 (2014), 405-428. [12] J. Conway, The Theory of Subnormal Operators, Math. Surveys Monographs, vol 36, Amer. Math. Soc. Providence, RI 1991. [13] J. Cui and Y. Duan, Berger measure for S(a,b,c,d), J. Math. Anal. Appl. 413 (2014), 202-211. [14] R. Curto and G. Exner, Berger measure for some transformations of subnormal weighted shifts, Integr. Equ. Oper. Theory 84, no. 3 (2016), 429-450. [15] R. Curto, Y. Poon, and J. Yoon, Subnormality of Bergman-like weighted shifts, J. Math. Anal. Appl. 308 (2005), 334-342. [16] R. Curto and N. Salinas, Generalized Bergman Kernels and the Cowen-Douglas Theory, Amer. J. Math. 106, no. 2 (1985), 447-488. [17] I.S. Gradshteyn and I.M. Ryzhik, Table of integrals, series, and products, 7th edition, Elsevier Academic Press, New York, 2007. [18] N. Jewell and A. Lubin, Commuting weighted shifts and analytic function theory in several variables, J. Operator Theory 1 (1979), 207-223. [19] N. Salinas, Products of kernel functions and module tensor products, Topics in operator theory, 219-241, Oper. Theory Adv. Appl. 32, Birkhuser, Basel, 1988. [20] J. Steinig, The sign of Lommel's function, Trans. Amer. Math. Soc. 163 (1972), 123-129. ON SUM OF TWO SUBNORMAL KERNELS 14 (S. Ghara) Department of Mathematics, Indian Institute of Science, Bangalore 560012, India E-mail address: [email protected] (S. Kumar) Department of Mathematics, Indian Institute of Science, Bangalore 560012, India E-mail address: [email protected]
1001.0755
1
1001
2010-01-05T20:09:24
Sharp estimates for the commutators of the Hilbert, Riesz transforms and the Beurling-Ahlfors operator on weighted Lebesgue spaces
[ "math.FA", "math.CA" ]
We prove that the operator norm on weighted Lebesgue space L2(w) of the commutators of the Hilbert, Riesz and Beurling transforms with a BMO function b depends quadratically on the A2-characteristic of the weight, as opposed to the linear dependence known to hold for the operators themselves. It is known that the operator norms of these commutators can be controlled by the norm of the commutator with appropriate Haar shift operators, and we prove the estimate for these commutators. For the shift operator corresponding to the Hilbert transform we use Bellman function methods, however there is now a general theorem for a class of Haar shift operators that can be used instead to deduce similar results. We invoke this general theorem to obtain the corresponding result for the Riesz transforms and the Beurling-Ahlfors operator. We can then extrapolate to Lp(w), and the results are sharp for 1 < p < 1.
math.FA
math
Sharp estimates for the commutators of the Hilbert, Riesz transforms and the Beurling-Ahlfors operator on weighted Lebesgue spaces Daewon Chung June 8, 2018 Abstract We prove that the operator norm on weighted Lebesgue space L2(w) of the commu- tators of the Hilbert, Riesz and Beurling transforms with a BM O function b depends quadratically on the A2-characteristic of the weight, as opposed to the linear dependence known to hold for the operators themselves. It is known that the operator norms of these commutators can be controlled by the norm of the commutator with appropriate Haar shift operators, and we prove the estimate for these commutators. For the shift operator corresponding to the Hilbert transform we use Bellman function methods, however there is now a general theorem for a class of Haar shift operators that can be used instead to deduce similar results. We invoke this general theorem to obtain the corresponding result for the Riesz transforms and the Beurling-Ahlfors operator. We can then extrapolate to Lp(w), and the results are sharp for 1 < p < ∞. 1 2 1 Introduction We say w is a weight if it is positive almost everywhere and locally integrable. The norm of f ∈ Lp(w) is kfkLp(w) :=(cid:18)ZR f (x)pw(x)dx(cid:19)1/p . Helson and Szegö first found necessary and sufficient conditions for the boundedness of the Hilbert transform Hf (x) = p.v. 1 πZ f (y) x − y dy in weighted Lebesgue spaces in [11]. Hunt, Muckenhoupt, and Wheeden showed in [13] that a new necessary and sufficient condition for boundedness of the Hilbert transform in Lp(w) is that the weight satisfies the Ap condition, namely: [w]Ap := sup I hwiIhw−1/(p−1)ip−1 I < ∞ , (1.1) where we denote the average over the interval I by h·iI , and we take the supremum over all intervals in R . After one year, Coifman and Fefferman extended in [5] the result to a larger class of convolution singular integrals with standard kernels. Recently, many authors have been interested in finding the 1Key words and phrases: Operator-weighted inequalities, Commutator, Hilbert transform, Paraproduct 22000 Mathematics Subject Classification. Primary 42A50; Secondary 47B47. 1 sharp bounds for the operator norms in terms of the Ap-characteristic [w]Ap of the weight. That is, one looks for a function φ(x), sharp in terms of its growth, such that kT fkLp(w) ≤ Cφ(cid:0)[w]Ap(cid:1)kfkLp(w) . For T = M, the Hardy-Littlewood maximal function, S. Buckley [3] showed that φ(x) = x1/(p−1) is the sharp rate of growth for all 1 < p < ∞ . He also showed in [3] that φ(x) = x2 works for the Hilbert transform in L2(w) . S. Petermichl and S. Pott improved the result to φ(x) = x3/2 , for the Hilbert transform in L2(w) in [28]. More recently, S. Petermichl proved in [26] the linear dependence, φ(x) = x, for the Hilbert transform in L2(w) kHfkL2(w) ≤ C[w]A2kfkL2(w) , by estimating the operator norm of the dyadic shift. Linear bounds in L2(w) were also obtained by O. Beznosova for the dyadic paraproduct [2]. Most recently there are new proofs for the linear estimates in L2(w) for some operators including the Hilbert transform and the dyadic paraproduct, [17] and [7]. The conjecture is that all the Calderón-Zygmund singular integral operator obey linear bounds in weighted L2 . So far this is known only for the Beurling-Ahlfors operator [9], [30], the Hilbert transform [26], Riesz transforms [27], the martingale transform [32], the square function [12], [24] and well localized dyadic operators [15], [17], [7]. It is now also known for Calderón-Zygmund convolutions operators that are smooth averages of well localized operators [31]. In this paper, we are interested in obtaining sharp weight inequalities for the commutators of the Hilbert, Riesz, and Beurling transforms with multiplication by a locally integrable function b ∈ BM O . 1.1 Commutators: main results Commutator operators are widely encountered and studied in many problems in PDEs, and Har- monic Analysis. One classical result of Coifman, Rochberg, and Weiss states in [6] that, for the Calderón-Zygmund singular integral operator with a smooth kernel, [b, T ]f := bT (f ) − T (bf ) is a bounded operator on Lp Rn, 1 < p < ∞ , when b is a BMO function. Weighted estimates for the commutator have been studied in [1], [20], [21], and [23]. Note that the commutator [b, T ] is more singular than the associated singular integral operator T , in particular, it does not satisfy the cor- responding weak (1, 1) estimate. However one can find a weaker estimate in [21]. In 1997, C. Pérez [21] obtained the following result concerning commutators of singular integrals, for 1 < p < ∞ , k[b, T ]fkLp(w) ≤ CkbkBM O[w]2 A∞kM 2fkLp(w) , where M 2 = M ◦ M denotes the Hardy-Littlewood maximal function iterated twice. With this result and Buckley's sharp estimate for the maximal function [3] one can immediately conclude that 2 k[b, T ]kLp(w)→Lp(w) ≤ C[w]2 A∞ [w] p−1 Ap kbkBM O . In this paper we show that for T the Hilbert, Riesz, Beurling transform, for 1 < p ≤ 2 one can drop [w]A∞ term, in the above estimate, and this is sharp. However for p > 2 , the Lp(w)-norm of [b, T ] is bounded above by kbkBM Od[w]2 . For T = H the Hilbert transform we prove, using Bellman function techniques similar to those used in [2], [26], the following Theorem. A2 p 2 Theorem 1.1. There exists a constant C > 0, such that k[b, H]fkLp(w)→Lp(w) ≤ CkbkBM O[w] 2 max{1, 1 Ap p−1 } kfkLp(w) , and this is sharp for 1 < p < ∞. Most of the work goes into showing the quadratic estimate for p = 2, sharp extrapolation [8] then provides the right rate of growth for p 6= 2. An example of C. Pérez shows the rates are sharp. Our method involves the use of the dyadic paraproduct πb and its adjoint π∗ b , both of which obey linear estimates in L2(w), see [2], like the Hilbert transform. It also uses Petermichl description of the Hilbert transform as an average of dyadic shift operators S, [25] and reduces the estimate to obtaining corresponding estimates for the commutator [b, S]. After we decompose this commutator in three parts: [b, S]f = [πb, S]f + [π∗ b , S]f + [λb, S]f , we estimate each commutator separately. This decomposition has been used before to analyze the commutator, [25], [15], [16]. For precise definitions and detail derivations, see Section 1.2. The first two commutators immediately give the desired quadratic estimates in L2(w) from the known linear bounds of the operators commuted. For the third commutator we can prove a better than quadratic bound, in fact a linear bound. The following Theorem will be the crucial part of the proof. Theorem 1.2. There exists a constant C > 0, such that k[λb, S]kL2(w)→L2(w) ≤ C[w]Ad for all b ∈ BM O d and w ∈ Ad 2. 2kbkBM O d , (1.2) This theorem is an immediate consequence of results in [14], [17] and [7], since the operator [λb, S] belongs to the class of Haar shift operators for which they can prove linear bounds. We present a different proof of this result and others, using Bellman function techniques and bilinear Carleson embedding theorems, very much in the spirit of [25] and [2]. These arguments were found independently by the author, and we think they can be of interest. We then observe that for any Haar shift operator T as defined in [17] the commutator [λb, T ] is again a Haar shift operator, and therefore it obeys linear bounds in A2-characteristic of the weight as in Theorem 1.2. As a consequence, we obtain quadratic bounds for all commutators of Haar shift operators and BMO function b . In particular, this holds true for Haar shift operators in Rn whose averages recover the Riesz transforms [27] and for martingale transforms in R2 whose averages recover the Beurling-Ahlfors operator [30], [9]. Extrapolation will provide Lp(w) bounds which turn out to be sharp for the Riesz transforms and Beurling-Ahlfors operators as well. The following Theorem holds Theorem 1.3. Let Tτ be the first class of Haar shift operators of index τ . Its convex hull include the Hilbert transform, Riesz transforms, the Beurling-Ahlfors operator and so on. Then, there exists a constant C(τ, n, p) which only depend on τ , n and p such that k[b, Tτ ]kLp(w)→Lp(w) ≤ C(τ, n, p)[w] 2 max{1, 1 Ap p−1 } kbkBM O See Section 10 for definitions and precise statements of these results. 3 1.2 The Hilbert transform case The bilinear operator f H(g) + H(f )g maps L2 × L2 into H 1, here H is the Hilbert transform and H 1 is the real Hardy space defined by H 1(R) := {f ∈ L1(R) : Hf ∈ L1(R)} with norm Because the dual of H 1 is BM O, we will pair with a BM O function b . Using that H ∗ = −H, we obtain that kfkH 1 = kfkL1 + kHfkL1 . h f H(g) + H(f )g, bi = h f, H(g)b − H(gb)i . Hence the operator g 7→ H(g)b − H(gb) should be L2 bounded. This expression H(g)b − H(gb) is called the commutator of H with the BM O function b . More generally, we define as follows. Definition 1.4. The commutator of the Hilbert transform H with a function b is defined as [b, H](f ) = bH(f ) − H(bf ) . Our main concern in this paper is to prove that the commutator [b, H], as an operator from R(w) is bounded by the square of the A2-characteristic, [w]A2 , of the weight times the R(w) into L2 L2 BM O norm, of b , kbkBM O , where kbkBM O := sup I 1 IZI b(x) − hbiIdx . The supremum is taken over all intervals in R . Note that when we restrict the supremum to dyadic intervals this will define BM O d and we denote this dyadic BM O norm by k·kBM O d . We now state our main results. Theorem 1.5. There exists C such that for all w ∈ A2 , k[b, H]kL2(w)→L2(w) ≤ C[w]2 A2kbkBM O , for all b ∈ BM O . Once we have boundedness and sharpness for the crucial case p = 2, we can carry out the power of the Ap-characteristic, for any 1 < p < ∞ , using the sharp extrapolation theorem [8] to obtain Theorem 1.1. Furthermore, an example of C. Pérez [22] shows this quadratic power is sharp. In [25], S. Petermichl showed that the norm of the commutator of the Hilbert transform is bounded by the supremum of the norms of the commutator of certain shift operators. This result follows after writing the kernel of the Hilbert transform as a well chosen averages of certain dyadic shift operators discovered by Petermichl. More precisely, S. Petermichl showed there is a non zero constant C such that k[b, H]k ≤ C sup α,r k[b, Sα,r]k , (1.3) where the dyadic shift operator Sα,r is defined by Sα,rf = XI∈Dα,rhf, hIi(hI− − hI+) . 4 π∗ b (f )(x) :=XI∈D πb(f )(x) :=XI∈D λb(f )(x) :=XI∈D hb, hIihf, hIih2 I (x) , hb, hIihfiI hI (x) , hbiIhf, hIihI (x) . S(f ) =XI∈D hf, hIi(hI− − hI+) Thus, we have where [b, S] = [π∗ b , S] + [πb, S] + [λb, S] , (1.5) Where hI denotes the Haar function associated to the interval I, I± denote the left and right halves of I, and Dα,r is a shifted and dilated system of dyadic intervals. Denote by S the shift op- erator corresponding to the standard dyadic intervals D . See the next section for a precise definition. Let us consider a compactly supported b ∈ BM O d and f ∈ L2 . Expanding b and f in the Haar system associated to the dyadic intervals D , b(x) =XI∈D hb, hIihI (x), f (x) = XJ∈D hf, hJihJ (x) ; formally, we get the multiplication of b and f to be broken into three terms, bf = π∗ b (f ) + πb(f ) + λb(f ) , (1.4) where πb is the dyadic paraproduct, π∗ b is its adjoint and λb(·) = π(·)b , defined as follows and we can estimate each term separately. Notice that both πb and π∗ b are bounded operators in Lp for b ∈ BM O , despite the fact that multipication by b is not a bounded operator in L2 unless b is bounded (L∞) . Therefore, λb can not be a bounded operator in Lp . However [λb, S] will be bounded on Lp(w) and will be better behaved than [b, S] . Decomposition (1.5) was used to analyze the commutator with the shift operator first by Petermichl in [25], but also Lacey in [15] and authors in [16] to analyze the iterated commutators. Since all estimates are independent on the dyadic grid, through out this paper we only deal with the dyadic shift operator S associated to the standard dyadic grid D . For a single shift operator the hypothesis required on b and w are that they belong to dyadic BM O d and Ad 2 with respect to the underlying dyadic grid defining the operator. However since ultimately we want to average over all grids, we will need b and w belonging to BM O d and Ad 2 for all shifted and scaled dyadic grids, that we will have if b ∈ BM O and w ∈ A2 , non-dyadic BM O and A2 . Beznosova has proved linear bounds for πb and π∗ b , [2], together with Petermichl's [26] linear bounds for S , this immediately provides the quadratic bounds for [πb, S] and [π∗ b , S]. Theorem 1.5 will be proved once we show the quadratic estimate holds for [λb, S]. We can actually obtain a better linear estimate as in Theorem 1.2. Some terms in (1.5) do also obey linear bounds. Theorem 1.6. There exists C such that kπ∗ b SkL2(w) + kSπbkL2(w) ≤ C[w]Ad 2kbkBM Od . for all b ∈ BM Od . 5 Note the three operators [λb, S] , π∗ b S and Sπb are generalized Haar shift operators for which there are now two different proofs of linear bounds on L2(w) with respect to [w]Ad , [17] and [7], and in this paper we present a third proof. We are now ready to explain the organization of this paper. In Section 2 we will introduce notation and discuss some useful results about weighted Haar systems. In Section 3 we will start our discussion on how to find the linear bound for the term [λb, S] , most of which will be very similar to calculations performed in [26]. In Section 4 we will introduce a number of Lemmas and Theorems that will be used. In Section 5 we will finish the linear estimate for the term [λb, S] , and prove Theorem 1.5. In Section 6 we prove the linear bound for π∗ b S . In Section 7 we reduce the proof of the linear bound for Sπb to verifying three embedding conditions, two are proved in this section, the third is proved in Section 8 using a Bellman function argument. In Section 9 we present the Lp(w) estimate of the commutator with a Haar shift operator, Theorem 1.3. Finally, in Section 10 we provide the sharpness for the commutators of Hilbert, Riesz transforms and Beurling-Ahlfors operators. 2 2 Preliminaries and Notation Let us now introduce the notation which will be used frequently through this paper. Even though the Ap conditions have already been introduced in (1.1), we will state the special case of this condition when p = 2 , namely Ad 2 since we will refer repeatedly to this. We say w belongs to Ad 2 class, if [w]Ad 2 := sup I∈DhwiIhw−1iI < ∞ . (2.1) Here we take the supremum over all dyadic interval in R . Note that if w ∈ A2 then w ∈ Ad 2 and 2 ≤ [w]A2 . Intervals of the form [k2−j , (k + 1)2−j ) for integers j, k are called dyadic intervals. [w]Ad Again, let us denote D the collection of all dyadic intervals, and let us denote D(J) the collection of all dyadic subintervals of J . For any interval I ∈ D, there is a Haar function defined by hI (x) = 1 I1/2 (χI+(x) − χI−(x)) , where χI denotes the characteristic function of the interval I , χI (x) = 1 if x ∈ I , χI (x) = 0 otherwise. It is a well known fact that the Haar system {hI}I∈D is an orthonormal system in L2 R . We also consider the different grids of dyadic intervals parametrized by α, r, defined by D α,r = {α + rI : I ∈ D} , for α ∈ R and positive r . For each grid D α,r of dyadic intervals, there are corresponding Haar func- tions hα,r R . Let us introduce a proper orthonormal system for L2 , I ∈ D α,r that are an orthonormal system in L2 I R(w) defined by hw I := 1 w(I)1/2(cid:20) w(I−)1/2 w(I+)1/2 χI+ − w(I+)1/2 w(I−)1/2 χI−(cid:21) , where w(I) =RI w . We define the weighted inner product by hf, giw =RR f gw . Then, every function f ∈ L2(w) can be written as hf, hw I iwhw I , where the sum converges a.e. in L2(w) . Moreover, f =XI∈D L2(w) =XI∈D kfk2 6 hf, hw I iw2 . (2.2) Again D can be replaced by D α,r and the corresponding weighted Haar functions are an orthonormal system in L2(w) . For convenience we will observe basic properties of the disbalanced Haar system. First observe that hhK , hw I iw could be non-zero only if I ⊇ K, moreover, for any I ⊇ K, hhK , hw Here is the the calculation that provides (2.3), I iw ≤ hwi1/2 K . (2.3) 1 hhK , hw I iw =(cid:12)(cid:12)(cid:12)(cid:12)Z K1/2w(I)1/2 (χK+(x) − χK−(x))(cid:20) w(I−)1/2 K1/2w(I)1/2ZK(cid:12)(cid:12)(cid:12)(cid:12) {z If K ⊂ I+, then A ≤ w(I−)1/2w(I+)−1/2w(K). Thus w(I−)1/2 w(I+)1/2 χI+(x) + w(I+)1/2 χI+(x) − w(I+)1/2 w(I−)1/2 w(I+)1/2 w(I−)1/2 χI−(x)(cid:21)w(x)dx(cid:12)(cid:12)(cid:12)(cid:12) χI−(x)(cid:12)(cid:12)(cid:12)(cid:12)w(x)dx } ≤ 1 A . hhK , hw I iw ≤ hwi1/2 K s w(K)w(I−) w(I+)w(I) ≤ hwi1/2 K . Similarly, if K ⊂ I− . If K = I, then A ≤ 2 w(K−)1/2w(K+)1/2 . Thus 2pw(K−)w(K+) I iw = hhK , hw Kiw ≤ hwi1/2 hhK , hw w(K) K ≤ hwi1/2 K . Estimate (2.3) implies that hh J , hw−1 J iw−1hhJ , hw J iw ≤ √2[w]1/2 Ad 2 , where J is the parent of J , hh J , hw−1 J iw−1hhJ , hw w(cid:19)1/2 J (cid:18) 1 JZJ J = hw−1i1/2 J iw ≤ hw−1i1/2 J hwi1/2 J (cid:18) 2 w(cid:19)1/2 JZJ √2hw−1i1/2 = hw−1i1/2 J hwi1/2 √2 [w]1/2 ≤ ≤ Ad 2 J . Also, one can deduce similarly the following estimate hhJ , hw−1 J iw−1hhJ−, hw J iw ≤ √2[w]1/2 Ad 2 . (2.4) (2.5) For I ) J, hw value of hw J J I is constant on J. We will denote this constant by hw on J and hw (J) ≤ w(J)−1/2 , as can be seen by the following estimate, (J) =(cid:26) w( J−)1/2/w( J )1/2w( J+)1/2 ≤ w(J)−1/2 if J = J+ w( J+)1/2/w( J )1/2w( J−)1/2 ≤ w(J)−1/2 if J = J− . I (J) . Then hw J hw J Let us define the weighted averages, hgiJ,w := w(J)−1RJ g(x)w(x)dx . As with the standard Haar system, we can write the weighted averages (J) is the constant (2.6) hgiJ,w = XI∈D:I)J 7 hg, hw I iwhw I (J) . (2.7) In fact, here is the derivation of (2.7). hgiJ,w = = 1 w(J)ZJXI∈D w(J)ZJ hg, hw I iwhw w(x)dx XI∈D:I)J 1 1 w(J)ZJXI∈D I (J) = XI∈D:I)J hg, hw I iwhw hg, hw I iwhw I (J) . I (x)w(x)dx = hg, hw I iwhw I (J)w(x)dx Also, we will be using system of functions {H w I }I∈D defined by H w I = hIpI − Aw I χI where Aw I = hwiI+ − hwiI− . (2.8) 2hwiI I } is orthogonal in L2 with norms satisfying the inequality kw1/2H w Then, {w1/2H w refer to [2]. Moreover, by Bessel's inequality we have, for all g ∈ L2 , L2 . 1 IhwiI hg, w1/2H w I i2 ≤ kgk2 XI∈D I kL2 ≤pIhwiI , (2.9) 3 Linear bound for [λb, S] part I In general, when we analyse commutator operators, a subtle cancelation delivers the result one wants to find. In the analysis of the commutator [b, S], the part [λb, S] will allow for certain cancelation. First, let us rewrite [λb, S] . [λb, S](f ) = λb(Sf ) − S(λbf ) hλbf, hJi(hJ− − hJ+) hbiIhSf, hIihI −XJ∈D hbiIhf, hJihhJ− − hJ+, hIihI −XJ∈DXI∈D hbiJ+hf, hJihJ+ −XJ∈D hbiJ−hf, hJihJ− −XJ∈D hbiJ− − hbiJ+ 2 hf, hJihJ− −XJ∈D hbiJ+ − hbiJ− 2 ∆J bhf, hJi(hJ+ + hJ−) , =XI∈D =XI∈DXJ∈D =XJ∈D =XJ∈D = −XJ∈D hbiIhf, hIihhI , hJi(hJ− − hJ+) hbiJ+ + hbiJ− 2 hf, hJi(hJ− − hJ+) hf, hJihJ+ recall the notation ∆J b = (hbiJ+ −hbiJ−)/2 . To find the L2(w) operator norm of [λb, S], it is enough to deal with the operator Sb(f ) =XI∈D ∆I bhf, hIihI− . We shall state the weighted operator norm of Sb as a Theorem and give a detailed proof. Theorem 1.2 is a direct consequence of the following Theorem. We will prove the following Theorem by the technique used in [26]. Theorem 3.1. There exists a constant C > 0, such that for all b ∈ BM O d and w ∈ Ad kSbkL2(w)→L2(w) ≤ C[w]Ad 2 for all f ∈ L2(w). 2kbkBM O d 8 (3.1) One can easily check that choosing b = I1/2hI , yields kSkL2(w)→L2(w) ≤ C[w]Ad (3.1) is equivalent to the following inequality for any positive function f, g 2 . Inequality hSb,w−1f, giw ≤ C[w]Ad 2kbkBM O dkfkL2(w−1)kgkL2(w) , (3.2) where Sb,w−1(f ) = Sb(w−1f ) . Expanding f and g in the disbalanced Haar systems respectively for L2(w−1) and L2(w) yields for (3.2), hSb,w−1f, giw =(cid:12)(cid:12)(cid:12)(cid:12)Z Sb,w−1(cid:0)XI∈D =(cid:12)(cid:12)(cid:12)(cid:12)XI∈DXJ∈D =(cid:12)(cid:12)(cid:12)(cid:12)XI∈DXJ∈D I hf, hw−1 I hf, hw−1 I iw−1hg, hw hf, hw−1 iw−1hg, hw hg, hw I (cid:1)(cid:0)XJ∈D iw−1hw−1 J iwZ hw J iwhSb,w−1hw−1 I J Sb,w−1(hw−1 I J iwhw J(cid:1)wdx(cid:12)(cid:12)(cid:12)(cid:12) )wdx(cid:12)(cid:12)(cid:12)(cid:12) J iw(cid:12)(cid:12)(cid:12)(cid:12) . , hw (3.3) In (3.3), hSb,w−1hw−1 I , hw ∆LbhhL, w−1hw−1 J iw =DXL∈D JEw i hL− , hw iw−1 6= 0, only when L ⊆ I and hhL−, hw I = XL∈D ∆LbhhL, hw−1 I iw−1hhL−, hw J iw I Since hhL, hw−1 J iw 6= 0 only when L− ⊆ J , then we have non-zero terms if I ⊆ J or J ⊆ I in the sum of (3.3). Thus we can split the sum into four parts, PI=J , PI= J , P J (I , andPI(J . Let us now introduce the truncated shift operator SI b(f ) := XL∈D(I) ∆Lbhf, hLihL− , and its composition with multiplication by w−1, SI b,w−1(f ) := XL∈D(I) ∆Lbhw−1f, hLihL− . We will see that the weighted norm kSI b,w−1χIkL2(w), plays a main role in our estimate for hS b,w−1hw−1 I , hw J iw . 4 Theorems and Lemmas To prove inequality (3.2) we need several theorems and lemmas. One can find the proof in the indicated references. First we recall some embedding theorems. Another main result in [26] is a two-weighted bilinear embedding theorem, which was proved by a Bellman function argument. Theorem 4.1 (Petermichl's Bilinear Embedding Theorem). Let w and v be weights so that hwiIhviI ≤ Q for all intervals I and let {αI} be a non-negative sequence so that the three estimates below hold for all J XI∈D(J) XI∈D(J) αI hwiI ≤ Q v(J) hviI ≤ Q w(J) αI 9 (4.1) (4.2) XI∈D(J) αI ≤ QJ . (4.3) Then there is c such that for all f ∈ L2(w) and g ∈ L2(v) XI∈D αIhfiI,whgiI,v ≤ cQkfkL2(w)kgkL2(v) . Corollary 4.2 (Bilinear Embedding Theorem). Let w and v be weights. Let {αI} be a sequence of nonnegative numbers such that for all dyadic intervals J ∈ D the following three inequalities hold with some constant Q > 0 , XI∈D(J) XI∈D(J) XI∈D(J) αIhviI I ≤ Q v(J) αIhwiI I ≤ Q w(J) αIhwiIhviI ≤ QJ . (4.4) (4.5) (4.6) Then for any two nonnegative function f, g ∈ L2 XI∈D αIhf v1/2iIhgw1/2iI I ≤ CQkfkL2kgkL2 holds with some constant C > 0 . Both Bilinear Embedding Theorems are key tools in our estimate. One version of such a theorem appeared in [19]. The original version of the next lemma also appeared in [26]. Using the fact that, for all I ∈ D , ∆I b ≤ 2kbkBM Od , one can prove the following lemma similarly to its original proof. w−1(I)1/2 for all Lemma 4.3. There is a constant c such that kSI intervals I and weights w ∈ Ad 2 . b,w−1χIkL2(w) ≤ ckbkBM O d[w]Ad 2 We will also need the Weighted Carleson Embedding Theorem from [19], and some other in- equalities for weights. Theorem 4.4 (Weighted Carleson Embedding Theorem). Let {αJ} be a non-negative sequence such that for all dyadic intervals I XJ∈D(I) αJ ≤ Qw−1(I) . Then for all f ∈ L2(w−1) XJ∈D αJhfi2 J,w−1 ≤ 4Qkfk2 L2(w−1) . Theorem 4.5 (Wittwer's sharp version of Buckley's inequality). There exist a positive constant C such that for any weight w ∈ Ad 2 and dyadic interval I ∈ D , 1 J XI∈D(J)(cid:0)hwiI+ − hwiI−(cid:1)2 hwiI I ≤ C[w]Ad 2hwiJ . 10 We refer to [32] for the proof. You can find extended versions of Theorem 4.4 and 4.5 to the doubling positive measure σ in [24]. One can find the Bellman function proof of the following three Lemmas in [2]. Lemma 4.6. For all dyadic interval J and all weights w . 1 J XI∈D(J) I∆I w2 1 hwi3 I ≤ hw−1iJ . Lemma 4.7. Let w be a weight and {αI} be a Carleson sequence of nonnegative numbers. If there exist a constant Q > 0 such that then ∀ J ∈ D , 1 J XI∈D(J) αI ≤ Q , ∀ J ∈ D , 1 J XI∈D(J) αI hw−1iI ≤ 4QhwiJ and therefore if w ∈ Ad 2 the for any J ∈ D we have Lemma 4.8. If w ∈ Ad ∀J ∈ D , 1 J XI∈D(J) hwiI αI ≤ 4Q[w]Ad 2hwiJ . 2 then there exists a constant C > 0 such that 1 J XI∈D(J)(cid:18)hwiI+ − hwiI− hwiI (cid:19)2 IhwiIhw−1iI ≤ C[w]Ad 2 . 5 Linear bound for [λb, S] part II We will continue to estimate the sum (3.3) in four parts. 5.1 PI= J For this case, it is sufficient to show that hSb,w−1hw−1 J , hw J iw ≤ ckbkBM O d[w]Ad 2 Since hhk, hw hSb,w−1hw−1 J I iw could be non-zero only if K ⊆ I, , hw ∆Lbhw−1hw−1 L , hLihL−, hw has non-zero term only when L ⊆ J. Thus ∆Lbhhw−1 hSb,w−1hw−1 , hw J L , hLiw−1hhL−, hw J iw =(cid:12)(cid:12)(cid:12)DXL∈D J iw =(cid:12)(cid:12)(cid:12) XL∈D( J ) =(cid:12)(cid:12)(cid:12) XL∈D(J) ∆Lbhhw−1 + ∆ J bhhw−1 J b,w−1hw−1 ≤ hSJ J J , hLiw−1hhL−, hw , h Jiw−1hh J−, hw J iw J iw + ∆ J bhhw−1 , hw J 11 . ∆Lbhhw−1 J , hLiw−1hhL−, hw J iw(cid:12)(cid:12)(cid:12) ∆Lbhhw−1 J , hLiw−1hhL−, hw J iw(cid:12)(cid:12)(cid:12) JEw(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)XL∈D J iw(cid:12)(cid:12)(cid:12) J iw(cid:12)(cid:12)(cid:12) +(cid:12)(cid:12)(cid:12) XL∈D(J s) , h Jiw−1hh J−, hw J iw in the second equality, J s denotes the sibling of J, so for all L ⊆ J s, hhL−, hw (2.4), J iw = 0 . Then, by ∆ J bhhw−1 J , h Jiw−1hh J− , hw J iw ≤ So for the remaining part: √2kbkBM O d[w]1/2 Ad 2 . (5.1) (J)hSJ here the last inequality uses (2.6) and Lemma 4.3. J iw = hw−1 , hw b,w−1hw−1 hSJ J J b,w−1χJ , hw J iw ≤ ckbkBM O d[w]Ad 2 , (5.2) 5.2 PI=J In this case, the argument is similar to the argument in Section 5.1. We have hSb,w−1hw−1 J , hw J iw = XL∈D ∆Lbhhw−1 J , hLiw−1hhL−, hw J iw here we have zero summands, unless L ⊆ J. Thus, hSb,w−1hw−1 J J J iw = hSJ b,w−1hw−1 b,w−1hw−1 , hw , hw J iw b,w−1hw−1 ≤ hSJ+ J iw + hSJ− , hw ≤ ckbkBM O d[w]Ad J J . 2 J iw + ∆J bhhw−1 , hw J , hJiw−1hhJ−, hw J iw In the last inequality, we use same arguments as in (5.2) for the first two term, and (2.5) for the last term. To obtain our desired results, we need to understand the supports of Sb(w−1hw−1 Since I ) and S∗ b (whw J ) . 5.3 P J (I and PI(J and hhw−1 I I Sb(w−1hw−1 ) = XL∈D ∆Lbhw−1hw−1 , hLihL− = XL∈D , hLi can be non-zero only when L ⊆ I , Sb(w−1hw−1 b (whw ), hw , S∗ I I hSb(w−1hw−1 I J iw = hhw−1 I J )iw−1 , ∆Lbhhw−1 I , hLiw−1hL− , ) is supported by I . Also, (5.3) 12 J , hL−ihL is supported by J . Let us now consider the sum yield that S∗ J ( I . Then b (whw J ) = PL∈D ∆Lbhwhw (cid:12)(cid:12)(cid:12)(cid:12) XI,J: J(I iw−1hg, hw hf, hw−1 I J iwhSb,w−1hw−1 I , hw J iw(cid:12)(cid:12)(cid:12)(cid:12) hf, hw−1 I iw−1hg, hw J iwhhw−1 I , S∗ b (whw = (cid:12)(cid:12)(cid:12)(cid:12) XI,J: J(I = (cid:12)(cid:12)(cid:12)(cid:12)XJ∈D XI:I) J = (cid:12)(cid:12)(cid:12)(cid:12)XJ∈D ≤ kgkL2(w)(cid:18)XJ∈D hf, hw−1 I iw−1hg, hw J iwhw−1 I ( J)hSb,w−1χ J , hw hfi J,w−1hg, hw J iwhSb,w−1χ J , hw J iw(cid:12)(cid:12)(cid:12)(cid:12) w(cid:19)1/2 J i2 J,w−1hSb,w−1χ J , hw , hfi2 J )iw−1(cid:12)(cid:12)(cid:12)(cid:12) J iw(cid:12)(cid:12)(cid:12)(cid:12) (5.4) (5.5) (5.6) here (5.3) and the fact that S∗ (2.7) and (5.3). If we show that b (whw J ) is supported by J are used for equality (5.4), and (5.5) uses XJ∈D then we have hfi2 J,w−1hSb,w−1χ J , hw J i2 w ≤ ckbk2 BM O d[w]2 Ad 2kfk2 L2(w−1) , (5.7) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) XI,J: J(I hf, hw−1 I iw−1hg, hw J iwhSb,w−1hw−1 I , hw 2kfk2 L(w−1)kgkL2(w) . J iw(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CkbkBM O d[w]Ad To prove the inequality (5.7), we apply the Theorem 4.4. The embedding condition becomes XJ∈D:J (I hSb,w−1χ J , hw J i2 w ≤ ckbk2 BM O d[w]2 Ad 2 w−1(I) after shifting the indices. Since hhL−, hw J ( I, we can write J iw = 0 unless L ⊆ J , and we will sum over J such that ∆Lbhw−1χ J , hLihhL−, hw J iw b,w−1χI, hw J iw . hSb,w−1χ J , hw J iw = XL∈D( J ) J iw = hSI ∆Lbhw−1χ J , hLihhL− , hw J iw = XL∈D = XL∈D(I ) w = XJ∈D:J (I J i2 hSb,w−1χ J , hw XJ∈D:J (I ∆Lbhw−1χI , hLihhL−, hw Thus, b,w−1χIk2 the sum J ( I. The partPI(J is similar toP J (I . One uses that Sb(w−1hw−1 last inequality due to (2.2). By Lemma 4.3, the embedding condition holds. Hence we are done for ) is supported by I b,w−1χI , hw J i2 and Theorem 4.4. w ≤ kSI L2(w) , hSI I 13 5.4 Proof of Theorem 1.5 We refer to [2] for following theorem. Theorem 5.1. The norm of dyadic paraproduct on the weighted Lebesgue space L2(w) is bounded from above by a constant multiple of the product of the Ad 2 characteristic of the weight w and the BM O d norm of b. To break [b, S] into three parts, as in (1.5), we assumed that b ∈ BM O d is compactly supported. However, we need to replace such a b with a general BM O d function. In order to pass from a compactly supported b to general b ∈ BM O d , we need the following lemma which is suggested in [10] . Lemma 5.2. Suppose φ ∈ BM O . Let eI be the interval concentric with I having length eI = 3I . Then there is ψ ∈ BM O such that ψ = φ on I, ψ = 0 on R \eI and kψkBM O ≤ ckφkBM O . Proof. Without loss of generality, we assume hφiI = 0 . Write I =S ∞ Jn, as in following figure. n=0 Jn where dist(Jn, ∂I) = J3 J1 J0 J2 J4 K1 {z I K2 } Then J0 is the middle third of I . For n > 0, let Kn be the reflection of Jn across the nearest endpoint of I and set This construction of ψ satisfies Lemma 5.2. ψ(x) = φ(x) if x ∈ I hφiJn if x ∈ Kn 0 otherwise . By Theorem 3.1, Corollary 1.2, Theorem 5.1, Lemma 5.2 and using the fact kπbk = kπ∗ bk, we can prove Theorem 1.5. Proof of Theorem 1.5. For any compactly supported b ∈ BM O , k[b, H]kL2(w)→L2(w) ≤ C sup ≤ C sup α,r k[b, S α,r]kL2(w)→L2(w) α,r (cid:0)k[πb, S α,r]kL2(w)→L2(w) + k[π∗ b , S α,r]kL2(w)→L2(w) + k[λb, S α,r]kL2(w)→L2(w)(cid:1) ≤ C(cid:0)4kπbkL2(w)→L2(w) sup ≤C[w]2 A2kbkBM O . α,r kS α,rkL2(w)→L2(w) + C[w]A2kbkBM O(cid:1) For fixed b, we consider the sequence of intervals Ik = [−k, k] and the sequence of BM O functions bk which are constructed as in Lemma 5.2. Then, there is a constant c, which does not depend on k , such that kbkkBM O ≤ ckbkBM O . Furthermore, there is a uniform constant C such that k[bk, H]kL2(w)→L2(w) ≤ C[w]2 (5.8) Therefore, for some subsequence of integers kj and f ∈ L2(w) , [bkj , H](f ) converges to [b, H](f ) almost everywhere. Letting j → ∞ and using Fatou's lemma, we deduce that (5.8) holds for all b ∈ BM O . A2kbkBM O . 14 6 Linear bound for π∗b S It might be useful to know what is the adjoint operator of S. Let us define sgn(I) = ±1 if I = I∓ . Then, for any function f, g ∈ L2 , hf, hIihg, hJihhI− − hI+, hJi hf, hIihg, hI+i hf, hIihg, hI−i −XI∈D hf, hIi(cid:0)sgn(I−)hg, hI−i + sgn(I+)hg, hI+i(cid:1) sgn(I)hg, hIih IE = hf, S∗gi . hf, h Iisgn(I)hg, hIi hSf, gi =XI∈DXJ∈D =XI∈D =XI∈D =XI∈D =Df,XI∈D S∗f (x) =XI∈D sgn(I)hf, hIih I (x) . Now, we see the adjoint operator of dyadic shift operator S is The following lemma provides the bound we are looking for the term π∗ Lemma 6.1. Let w ∈ Ad 2 and b ∈ BM O d. Then, there exists C so that b S . kπ∗ b SkL2(w)→L2(w) ≤ C[w]Ad 2kbkBM O d . Proof. In order to prove Lemma 6.1 it is enough to show that for any positive square integrable function f, g Using the system of functions {H w b S(f w−1/2), gw1/2i ≤ C[w]Ad hπ∗ (6.1) I }I∈D defined in (2.8), we can rewrite the left hand side of (6.1) 2kbkBM O dkfkL2kgkL2 . b S(f w−1/2), gw1/2i = hS(f w−1/2), πb(gw1/2)i =XI∈D hπ∗ hgw1/2iIhb, hIihS(f w−1/2), hIi =XI∈D =XI∈D hgw1/2iIhb, hIi sgn(I)hf w−1/2, h Ii sgn(I)hgw1/2iIhb, hIihf w−1/2, H w−1 +XI∈D q I sgn(I)hgw1/2iIhb, hIihf w−1/2, Aw−1 i 1 I I . (6.2) χ Ii 1 q I Our claim is that both sums in (6.2) are bounded by [w]Ad 2kbkBM O dkfkL2kgkL2 , i.e. sgn(I)hgw1/2iIhb, hIihf w−1/2, H w−1 I ≤ C[w]Ad 2kbkBM O dkfkL2kgkL2 (6.3) XI∈D and XI∈D sgn(I)hgw1/2iIhb, hIihf w−1/2, Aw−1 I χ Ii ≤ C[w]Ad 2kbkBM O dkfkL2kgkL2 . (6.4) 1 i q I q I 1 15 First let us verify the bound for (6.3). Using Cauchy-Schwarz inequality, XI∈D sgn(I)hgw1/2iIhb, hIihf w−1/2, H w−1 I 1 i q I Ihb, hIi2hw−1i I(cid:19)1/2(cid:18)XI∈D Ihb, hIi2hw−1i I(cid:19)1/2 hgw1/2i2 hgw1/2i2 ≤(cid:18)XI∈D ≤ kfkL2(cid:18)XI∈D hf, w−1/2H w−1 I i2(cid:19)1/2 1 Ihw−1i I . Thus, for (6.3), it is enough to show that XI∈D hgw1/2i2 Ihb, hIi2hw−1i I ≤ C[w]2 Ad 2kbk2 BM O dkgk2 L2 . It is clear that 2hwi I ≥ hwiI , thus XI∈D hgw1/2i2 Ihb, hIi2hw−1i I =XI∈D Ihb, hIi2hw−1i IhwiIhwi−1 hgw1/2i2 Ihb, hIi2hwi−1 . I I hgw1/2i2 2XI∈D ≤ 2[w]Ad If we show for all J ∈ D , 1 J XI∈D(J) hb, hIi2hwi−1 I hwi2 I = 1 J XI∈D(J) hb, hIi2hwiI ≤ [w]Ad 2kbk2 BM O dhwiJ , (6.5) (6.6) (6.7) then by Weighted Carleson Embedding Theorem 4.4 with w instead of w−1 , we will have (6.6) . Since b ∈ BM O d , {hb, hIi2}I∈D is a Carleson sequence with constant kbk2 BM O d that is 1 J XI∈D(J) hb, hIi2 ≤ kbk2 BM O d . Applying Lemma 4.7 with αI = hb, hIi we have inequality (6.7). We now concentrate on the estimate (6.4), we can estimate the left hand side of (6.4) as follows. XI∈D 1 sgn(I)hgw1/2iIhb, hIihf w−1/2, Aw−1 I χ Ii hgw1/2iI hb, hIihf w−1/2i I Aw−1 q I I q I sgn(I)hgw1/2iIhb, hIihf w−1/2i I Aw−1 q I q I hgw1/2i I hb, hIihf w−1/2i I Aw−1 hgw1/2iI(cid:0)hb, hI−i + hb, hI+i(cid:1) hf w−1/2iI Aw−1 =XI∈D ≤XI∈D ≤ 2XI∈D = 2XI∈D I I I pI . 16 By Bilinear Embedding Theorem, inequality (6.4) holds provided the following three inequalities hold, ∀ J ∈ D, ∀ J ∈ D, ∀ J ∈ D, I 1 1 J XI∈D(J)(cid:0)hb, hI−i + hb, hI+i(cid:1)Aw−1 J XI∈D(J)(cid:0)hb, hI−i + hb, hI+i(cid:1)Aw−1 J XI∈D(J)(cid:0)hb, hI−i + hb, hI+i(cid:1)Aw−1 1 I I 2 , pIhw−1iIhwiI ≤ CkbkBM O d [w−1]Ad pIhw−1iI ≤ CkbkBM O d [w−1]Ad 2hw−1iJ , pIhwiI ≤ CkbkBM O d [w−1]Ad 2hwiJ . (6.8) (6.9) (6.10) For (6.8), by Cauchy-Schwarz inequality I 1 J XI∈D(J)(cid:0)hb, hI−i + hb, hI+i(cid:1)Aw−1 pIhw−1iIhwiI ≤(cid:18) 1 J XI∈D(J)(cid:0)hb, hI−i + hb, hI+i(cid:1)2hw−1iIhwiI(cid:19)1/2(cid:18) 1 (hb, hI−i + hb, hI+i)2 ≤ 3XI∈D XI∈D Since 1 J XI∈D(J) and by Lemma 4.8, 1 J XI∈D(J) (Aw−1 I (hb, hI−i + hb, hI+i)2hw−1iIhwiI ≤ C[w−1]Ad 2 hb, hIi2 ≤ C[w−1]Ad 2kbk2 BM O d , )2Ihw−1iIhwiI(cid:19)1/2 . (6.11) (Aw−1 I J XI∈D(J) hb, hIi2 , J XI∈D(J) )2Ihw−1iIhwiI ≤ C [w−1]Ad 1 2 . Thus embedding condition (6.8) holds. For (6.9), by Cauchy-Schwarz inequality and (6.11) we have 1 J XI∈D(J) I (hb, hI−i + hb, hI+i)Aw−1 ≤ C(cid:18) 1 J XI∈D(J) pIhw−1iI hb, hIi2hw−1iI(cid:19)1/2(cid:18) 1 J XI∈D(J) J XI∈D(J)(cid:18)hw−1iI+ − hw−1iI− (cid:19)2 2hw−1iI )2Ihw−1iI = 1 By Theorem 4.5, 1 J XI∈D(J) (Aw−1 I (Aw−1 I )2Ihw−1iI(cid:19)1/2 . Ihw−1iI ≤ C[w−1]Ad 2hw−1iJ . Similarly with (6.7), we have 1 J XI∈D(J) hb, hIi2hw−1iI ≤ [w−1]A2kbk2 BM O dhw−1iJ . To finish, we must estimate (6.10). In a similar way with (6.9), we need to estimate (cid:18) 1 J XI∈D(J) hb, hIi2hwiI(cid:19)1/2(cid:18) 1 J XI∈D(J) (Aw−1 I )2IhwiI(cid:19)1/2 . 17 By Lemma 4.6, we have 1 J XI∈D(J) (Aw−1 I )2IhwiI ≤ [w−1]Ad 2 = [w−1]Ad 2 I 1 (Aw−1 )2Ihw−1i−1 J XI∈D(J) J XI∈D(J)(cid:18)hw−1iI+ − hw−1iI− hw−1i3 1 (cid:19)2 I I I ≤ C[w−1]Ad 2hw−1iJ . This completes the proof of Lemma 6.1. Due to the almost self adjoint property of the Hilbert transform, a certain bound for π∗ b H immediately returns the same bound for Hπb. However we have to prove the boundedness of Sπb independently because S is not self adjoint. 7 Linear bound for Sπb Lemma 7.1. Let w ∈ Ad 2 and b ∈ BM O d. Then, there exists C so that kSπbkL2(w)→L2(w) ≤ C[w]Ad 2kbkBM O d . Proof. We are going to prove Lemma 7.1 by showing hSπb(w−1f ), giw ≤ C[w]Ad 2kbkBM O dkfkL2(w−1)kgkL2(w) , (7.1) for any positive function f, g ∈ L2 . Since hSπb(f ), hIi = sgn(I)hπb(f ), h Ii = sgn(I)hfi Ihb, h Ii , We have By expanding g in the disbalanced Haar system for L2(w) , Sπb(f ) =XI∈D sgn(I)hfi Ihb, h IihI . hw−1fi Ihb, h Iisgn(I)hhI , giw hSπb(w−1f ), giw =XI∈D =XI∈DXJ∈D sgn(I)hw−1i Ihfi I,w−1hb, h Iihg, hw J iwhhI , hw J iw . Since hhI , hw J iw could be non zero only if J ⊇ I , we can split above sum into three parts, and XI∈D XI∈D XI∈D XJ:J ) I sgn(I)hw−1i Ihfi I,w−1hb, h Iihg, hw I iwhhI , hw I iw, sgn(I)hw−1i Ihfi I,w−1hb, h Iihg, hw I iwhhI , hw I iw, sgn(I)hw−1i Ihfi I,w−1hb, h Iihg, hw J iwhhI , hw J iw . 18 (7.2) (7.3) (7.4) We claim that all sums, (7.2), (7.3), and (7.4), can be bounded with a bound that depends on [w]Ad 2kbkBM O d at most linearly. Since hhI , hw I iw ≤ hwi1/2 , we can estimate (7.2) I (cid:12)(cid:12)(cid:12)(cid:12)XI∈D sgn(I)hw−1i Ihfi I,w−1hb, h Iihg, hw ≤(cid:18)XI∈D hw−1i2 Ihfi2 ≤ CkgkL2(w)[w]1/2 Ad I iwhhI , hw I iw(cid:12)(cid:12)(cid:12)(cid:12) I,w−1hb, h Ii2hwiI(cid:19)1/2(cid:18)XI∈D 2(cid:18)XI∈D hg, hw I,w−1hb, hIi2hw−1iI(cid:19)1/2 hfi2 . I i2(cid:19)1/2 By Weighted Carleson Embedding Theorem 4.4, XI∈D is provided by hfi2 I,w−1hb, hIi2hw−1iI ≤ C[w]Ad 2kbk2 BM O dkfk2 L2(w−1) 1 J XI∈D(J) hb, hIi2hw−1iI ≤ [w]Ad 2kbk2 BM O dhw−1iJ which we already have in (6.7) . Thus, we have sgn(I)hw−1i Ihfi I,w−1hb, h Iihg, hw XI∈D Similarly to (7.2), we can estimate (7.3) using hhI , hw I iwhhI , hw I iw ≤ hwi1/2 I ≤ √2hwi1/2 I I iw ≤ C[w]Ad 2kbkBM O dkfkL2(w−1)kgkL2(w) . (7.5) (cid:12)(cid:12)(cid:12)(cid:12)XI∈D sgn(I)hw−1i Ihfi I,w−1hb, h Iihg, hw I iw(cid:12)(cid:12)(cid:12)(cid:12) I iwhhI , hw hw−1i Ihfi I,w−1 hb, h Iihg, hw hw−1iIhfiI,w−1 hb, hIihg, hw I I iwhwi1/2 I iwhwi1/2 I ≤ √2XI∈D = 2√2XI∈D ≤ 2√2(cid:18)XI∈D hw−1i2 Ihfi2 2(cid:18)XI∈D ≤ CkgkL2(w)[w]1/2 ≤ C[w]Ad Ad I,w−1hb, hIi2hwiI(cid:19)1/2(cid:18)XI∈D hg, hw I,w−1hb, hIi2hw−1iI(cid:19)1/2 hfi2 I i2(cid:19)1/2 2kbkBM O dkfkL2(w−1)kgkL2(w) . Since hw J is constant on I, for J ) I and we denote this constant by hw J ( I) . Then we know by (2.7), XJ:J ) I J iw = XJ:J ) I hg, hw J iwhhI , hw hg, hw J iwhw J ( I)hhI , wi = hgi I,whhI , wi . 19 Thus, we can rewrite (7.4) (cid:12)(cid:12)(cid:12)(cid:12)XI∈D sgn(I)hw−1i Ihfi I,w−1hb, h Iihgi I,whhI , wi(cid:12)(cid:12)(cid:12)(cid:12) ≤XI∈D =XI∈D hw−1i Ihfi I,w−1 hb, h Iihgi I,whhI , wi hw−1iI hb, hIi (hhI− , wi + hhI+, wi)hfiI,w−1hgiI,w . (7.6) (7.7) 2kbkBM O dkfkL2(w−1)kgkL2(w) . We are going to prove We claim the sum (7.7) is bounded by [w]Ad it using Petermichl's Bilinear Embedding Theorem 4.1. Thus, we need to show that the following three embedding conditions hold, ∀ J ∈ D, ∀ J ∈ D, ∀ J ∈ D, 1 1 J XI∈D(J) J XI∈D(J) J XI∈D(J) 1 hb, hIihw−1iI (hhI−, wi + hhI+, wi) hb, hIihw−1iI (hhI−, wi + hhI+, wi) 1 hwiI ≤ C[w]Ad hw−1iI ≤ C[w]Ad 1 2kbkBM O d hw−1iJ , (7.8) 2kbkBM O d hwiJ , (7.9) hb, hIihw−1iI (hhI−, wi + hhI+, wi) ≤ C[w]Ad 2kbkBM O d . (7.10) After we split the sum in (7.8): 1 J XI∈D(J) hb, hIihw−1iIhhI−, wi 1 hwiI + 1 J XI∈D(J) hb, hIihw−1iIhhI+, wi 1 hwiI , we start with Cauchy-Schwarz inequality to estimate the first sum of embedding condition (7.8), 1 J XI∈D(J) hb, hIihw−1iIhhI−, wi ≤(cid:18) 1 J XI∈D(J) hb, hIi2hw−1i2 A2(cid:18) 1 J XI∈D(J) ≤ C[w]1/2 2kbkBM O dhw−1iJ . ≤ C[w]Ad 1 = 1 hwiI J XI∈D(J) IhwiI(cid:19)1/2(cid:18) 1 hb, hIi2hw−1i(cid:19)1/2(cid:18) 1 hb, hIihw−1iIpI−∆I−w hwiI I(cid:19)1/2 J XI∈D(J) I−∆I−w2 1 hwi3 I−(cid:19)1/2 J XI∈D(J) 1 hwi3 I−∆I−w2 (7.11) Inequality (7.11) due to Lemma 4.6 and (6.7) . Also, the other sum can be estimated by exactly the same method. Thus we have the embedding condition (7.8). To see the embedding condition (7.9), it is enough to show 1 J XI∈D(J) hb, hIihhI−, wi ≤ C[w]Ad 2kbkBM O dhwiJ , 20 as we did above. We use Cauchy-Schwarz inequality for embedding condition (7.9), then 1 J XI∈D(J) hb, hIihhI−, wi = 1 J XI∈D(J) ≤(cid:18) 1 J XI∈D(J) ≤ CkbkBM O dhwi1/2 hb, hIipI−∆I−w hw−1iI(cid:19)1/2(cid:18) 1 hb, hIi2 J (cid:18)[w]A2 J XI∈D(J) 1 J XI∈D(J) I−∆I−w2 I−∆I−w2hw−1iI(cid:19)1/2 hwiI−(cid:19)1/2 1 (7.12) ≤ C[w]Ad 2kbkBM O dhwiJ . (7.13) Here inequality (7.12) uses Lemma 4.7, and inequality (7.13) uses the fact that hwi−1 and Theorem 4.5 after shifting the indices. I ≤ 2hwi−1 I− If we show the embedding condition (7.10), then we can immediately finish the estimate for (7.7) 2kbkBM O dkfkL2(w−1)kgkL2(w) . Combining this and (7.5) will give us our desire with bound C[w]Ad result. 8 Proof for embedding condition (7.10) The following lemma lies at the heart of the matter for the proof of the embedding condition (7.10) . Lemma 8.1. There is a positive constant C so that for all dyadic interval J ∈ D J hw−1i1/4 I (cid:18)∆I+w + ∆I−w ≤ Chwi1/4 Ihwi1/4 hw−1i1/4 (cid:19)2 J I hwiI 1 J XI∈D(J) , (8.1) whenever w is a weight. Moreover, if w ∈ Ad 2 then for all J ∈ D 1 J XI∈D(J) IhwiIhw−1iI(cid:18)∆I+w + ∆I−w hwiI (cid:19)2 ≤ C [w]Ad 2 . Proof of condition (7.10). By using Cauchy-Schwarz inequality and Lemma 8.1, we have: 1 J XI∈D(J) 1 hb, hIihw−1iI (hhI−, wi + hhI+, wi) J XI∈D(J) √2(cid:18) 1 hb, hIihw−1iIrI J XI∈D(J) hb, hIi2hw−1iIhwiI(cid:19)1/2(cid:18) 1 1 2 (cid:0)hwiI−+ − hwiI−− + hwiI++ − hwiI+−(cid:1) Ihw−1iIhwi−1 J XI∈D(J) = ≤ I (∆I+w + ∆I−w)2(cid:19)1/2 ≤ C[w]Ad 2kbkBM O d . 21 We turn to the proof of Lemma 8.1. In the first place, we need to revisit some properties of function B(u, v) := 4√uv on the domain D0 which is given by It is known, we refer to [2], that B(u, v) satisfies the following differential inequality in D0 {(u, v) ∈ R2 + : uv ≥ 1/2} . − (du, dv)d2B(u, v)(du, dv)t ≥ 1 8 v1/4 u7/4du2 . Furthermore, this implies the following convexity condition. For all (u, v), (u±, v±) ∈ D0 , B(u, v) − B(u+, v+) + B(u−, v−) 2 v1/4 u7/4 ≥ C1 (u+ − u−)2 , where u = (u+ + u−)/2 and v = (v+ + v−)/2 . Let us define A(u, v, ∆u) := aB(u, v) + B(u + ∆u, v) + B(u − ∆u, v) , (8.2) (8.3) on the domain D1 with some positive constant a > 0 . Here (u, v, ∆u) ∈ D1 means all pairs (u, v), (u + ∆u, v), (u − ∆u, v) ∈ D0 . Then A has the size property and the convexity property: if (u, v, ∆u) ∈ D1, then 0 ≤ A(u, v, ∆u) ≤ (a + 2) 4√uv , (8.4) (8.5) and A(u, v, ∆u) − 1 2(cid:2)A(u+, v+, ∆u1) + A(u−, v−, ∆u2)(cid:3) ≥ C2 v1/4 u7/4 (∆u2 1 + ∆u2 2) , where u = (u+ + u−)/2 , v = (v+ + v−)/2, and ∆u = (u+ − u+)/2 . The property (8.5) is directly from the definition of function B(u, v) . At the end, ∆u will play the role of ∆Iw, ∆u1 is ∆I+w, and ∆u2 is ∆I−w . We can rewrite left hand side of the inequality (8.5) as follows 1 A(u, v, ∆u) − = aB(u, v) + B(u + ∆u, v) + B(u − ∆u, v) − 2(cid:2)A(u+, v+, ∆u1) + A(u−, v−, ∆u2)(cid:3) 1 2(cid:2)aB(u+, v+) + B(u+ + ∆u1, v+) + B(u+ − ∆u1, v+) = aB(u, v) − + aB(u−, v−) + B(u− + ∆u1, ∆u−) + B(u− − ∆u1, v−)(cid:3) 2hB(u + ∆u + ∆u1, v + ∆v) + B(u + ∆u − ∆u1, v + ∆v) + B(u − ∆u + ∆u2, v − ∆v) + B(u − ∆u − ∆u2, v − ∆v)i . (B(u+, v+) + B(u−, v−)) + B(u+, v) + B(u−, v) − a 2 1 (8.6) Using Taylor's theorem: B(u + u0, v + v0) = B(u, v) +∇B(u, v)(u0, v0)t +Z 1 0 (1− s)(u0, v0)d2B(u + su0, v + sv0)(u0, v0)tds , 22 and the convexity condition of B(u, v), we are going to estimate the lower bounds of (8.6). B(u + ∆u + ∆u1, v + ∆v) 1 − 2 = − 0 1 1 1 2(cid:16)B(u, v) + ∇B(u, v)(∆u + ∆u1, ∆v)t(cid:17) −Z 1 (1 − s)(∆u + ∆u1, ∆v)d2B(u + s(∆u + ∆u1), v + s∆v)(∆u + ∆u1, ∆v)tds 2(cid:16)B(u, v) + ∇B(u, v)(∆u + ∆u1, ∆v)t(cid:17) + 2(cid:16)B(u, v) + ∇B(u, v)(∆u + ∆u1, ∆v)t(cid:17) + 2(cid:16)B(u, v) + ∇B(u, v)(∆u + ∆u1, ∆v)t(cid:17) + 2(cid:16)B(u, v) + ∇B(u, v)(∆u + ∆u1, ∆v)t(cid:17) + (1 − s)(v + s∆v)1/4ds Z 1 ∆v v (1 − s) (∆u + ∆u1)2 0 (∆u + ∆u1)2v1/4 8(4u)7/4 v1/4 1 u7/4 (u + s(∆u + ∆u1))7/4 (1 − s)(1 + s 8Z 1 (∆u + ∆u1)2 . (v + s∆v)1/4 72 · 43/4 Z 1 8(4u)7/4 1 1 1 0 0 ≥ − ≥ − ≥ − ≥ − Inequality (8.7) is due to the following inequalities (∆u + ∆u1)2ds )1/4ds (8.7) (8.8) ∆u = u+ − u− 2 ≤ u+ + u− 2 = u and ∆u1 = u++ − u+− 2 ≤ u+ ≤ 2u . Since (1 − s)1/4 ≤ (1 − βs)1/4 ≤ (1 + βs)1/4 for any β < 1, it is clear that Z 1 0 (1 − s)(1 + βs)1/4ds ≥Z 1 0 (1 − s)5/4ds = 4 9 , and this allows the inequality (8.8). With the same arguments, we also estimate the following lower bounds: − 1 2hB(u + ∆u − ∆u1, v + ∆v) + B(u − ∆u + ∆u2, v − ∆v) + B(u − ∆u − ∆u2, v − ∆v)i ≥ − − − 1 1 2(cid:16)B(u, v) + ∇B(u, v)(∆u − ∆u1, ∆v)t(cid:17) + 2(cid:16)B(u, v) + ∇B(u, v)(−∆u + ∆u2,−∆v)t(cid:17) + 2(cid:16)B(u, v) + ∇B(u, v)(−∆u − ∆u2,−∆v)t(cid:17) + 1 1 72 · 43/4 1 1 72 · 43/4 72 · 43/4 v1/4 u7/4 (∆u − ∆u1)2 v1/4 u7/4 v1/4 u7/4 (−∆u + ∆u2)2 (∆u + ∆u2)2 . (8.9) We can have the following inequality by combining (8.8), (8.9) and (8.6), A(u, v, ∆u) − 1 2(cid:2)A(u+, v+, ∆u1) + A(u−, v−, ∆u2)(cid:3) a 2 1 ≥ (a − 2)B(u, v) − v1/4 u7/4 (2∆u2 + ∆u2 (∆u2 2) . 36 · 43/4 1 1 + ∆u2 + v1/4 u7/4 ≥ 36 · 43/4 1 + ∆u2 2) (B(u+, v+) + B(u−, v−)) + B(u+, v) + B(u−, v) (8.10) 23 To see the inequality (8.10), using convexity condition of B(u, v) = 4√uv and inequality: (1− s)u ≤ u − s∆u ≤ u + s∆u , a (a − 2)B(u, v) − 1 2 0 2(cid:16)B(u+, v+) + B(u−, v−)(cid:17) + B(u+, v) + B(u−, v) = a(cid:18)B(u, v) − −(cid:18) 3 ∆u2Z 1 ∆u2Z 1 ∆u2 − ∆u2 − 2(cid:19) v1/4 3 (B(u+, v+) + B(u−, v−)(cid:19) (1 − s)v1/4(u + s∆u)−7/4ds + 6 16 ∆u2 v1/4 3 u7/4 2 =(cid:18)aC1 − u7/4Z 1 16 v1/4 u7/4 v1/4 u7/4 (1 − s)−3/4ds 0 ∆u2 v1/4 ≥ aC1 = aC1 ∆u2 . u7/4 3 16 0 (1 − s)v1/4(u − s∆u)−7/4ds(cid:19) (8.11) Choosing a constant a sufficiently large so that aC1 > 3/2, quantity in (8.11) remains positive. This observation and discarding nonnegative terms yield inequality (8.10). Choosing the constant C2 = 1/(36 · 43/4) in (8.5) completes the proof of the concavity property of A(u, v, ∆u) . We now turn to the proof of Lemma 8.1. Proof of Lemma 8.1. Let uI := hwiI , vI := hw−1iI , u± = uI±, v± = vI±, ∆uI = ∆I w, ∆u1 = ∆uI+, and ∆u2 = ∆uI− . Then by Hölder's inequality (u, v, ∆u), (u+, v+, ∆u1), and (u−, v−, ∆u2) belong to D1 . Fix J ∈ D , by properties (8.4) and (8.5) (a + 2)J 4phwiJhw−1iJ ≥ JA(uJ , vJ , ∆uJ ) ≥ 1 2(cid:18)J+A(u+, v+, ∆u1) + J−A(u−, v−, ∆u2)(cid:19) + JC hw−1iJ hwi7/4 J Since A(u, v, ∆u) ≥ 0, iterating the above process will yield (∆J+u2 + ∆J−u2) . J 4phwiJhw−1iJ ≥ C XI∈D(J) Ihw−1i1/4 I hwi−7/4 I (∆I+w2 + ∆I−w2) . (8.12) Also, one can easily have and J 4phwiJhw−1iJ ≥ C XI∈D(J) J 4phwiJhw−1iJ ≥ C XI∈D(J) Ihw−1i1/4 I hwi−7/4 I ∆I+w2 , Ihw−1i1/4 I hwi−7/4 I ∆I−w2 . (8.13) (8.14) 24 Then, 1 J XI∈D(J) Ihw−1i1/4 I hwi−7/4 I (∆I+w + ∆I−w)2 = ≤ ≤ I 1 1 J(cid:18) XI∈D(J) Ihw−1i1/4 + ∆I−w2) + 2 XI∈D(J) J XI∈D(J) Ihw−1i1/4 + 2(cid:18) XI∈D(J) 4phwiIhw−1iI . Ihw−1i1/4 3 C I I hwi−7/4 I (∆I+w2 Ihw−1i1/4 I hwi−7/4 I (∆I+w∆I−w)(cid:19) hwi−7/4 I (∆I+w2 + ∆I−w2) hwi−7/4 I ∆I+w2(cid:19)1/2(cid:18) XI∈D(J) Ihw−1i1/4 I I ∆I−w2(cid:19)1/2! hwi−7/4 9 Recently developed tools and their applications The dyadic shift operator was first introduced in [20] to replace the weighted norm estimate for the Hilbert transform. It was also encountered in [29], so Riesz transforms can be obtained as the result of averaging some dyadic shift operator. Recently, in [17] and [7], a more general class of dyadic shift operators, so called the Haar shift operators were introduced. The Hilbert transform, Riesz transforms, and Beurling-Ahlfors operator are in the convex hull of this class, as they can be written as appropriate averages of Haar shift operators. Let Dn denote the collection of dyadic cubes in Rn , Dn(Q) denotes dyadic subcubes of Q , and Q denotes the volume of the dyadic cube Q . We start with some definitions. Definition 9.1. A Haar function on a cube Q ⊂ Rn is a function HQ such that (a) HQ is supported on Q, and is constant on Dn(Q) . (b) kHQk∞ ≤ Q−1/2 . (c) HQ has a mean zero. Definition 9.2. Given an integer τ > 0 , we say an operator of the following form is in the first class of Haar shift operators of index τ Tτ f (x) = XQ∈Dn XQ′,Q′′∈D(Q) 2−τ nQ≤Q′,Q′′ aQ′,Q′′hf, HQ′iHQ′′(x) , where the constant aQ′,Q′′ satisfy the following size condition: aQ′,Q′′ ≤ C(cid:18)Q′ Q · Q′′ Q(cid:19)1/2 25 . (9.1) Note that once a choice of Haar functions has been made {HQ}Q∈Dn , then this is an orthogonal family, such that kHQkL2 ≤ 1 , so one could normalize in L2 . Note that one can easily see that the dyadic shift operator S belongs to the first class of a Haar shift operator of index τ = 1 with aI ′,I ′′ =(cid:26) ±1 for I ′ = I, I ′′ = I∓ otherwise . 0 One of the main result in [17] and [7] is the following Theorem 9.3 ([17], [7]). Let T be in the first class of Haar shift operators of index τ . Then for all w ∈ Ad 2 , there exists C(τ, n) which only depends on τ and n such that kTkL2(w)→L2(w) ≤ C(τ, n)[w]Ad 2 . As a consequence of this Theorem, linear bounds for the Hilbert transform, Riesz transforms, and the Beurling-Ahlfors operator are recovered. There are now two different proofs of Theorem (9.3) in [17] and [7]. The commutator [λb, S] is also in the first class of Haar shift operators of index τ = 1 . Recall the observation in the section 4, Then we can see [λb, S](f ) = −XI∈D aI ′,I ′′ =(cid:26) −∆I b 0 ∆Ibhf, hIi(hI+ + hI−) . for I ′ = I, I ′′ = I± otherwise , moreover aI ′,I ′′ = ∆I b ≤ 2kbkBM O d this means the constant aI ′,I ′′ satisfy the size condition (9.1) with C = 2√2kbkBM O d . These observations, Theorem 5.1, and Theorem 9.3 immediately recover the quadratic bound for the commutator of the Hilbert transform which was proved in this paper. We now define the second class of Haar shift operators of index τ . Definition 9.4. Given an integer τ > 0 , we say an operator T of the form in Definition 9.2 is in the second class of Haar shift operators of index τ , if T is bounded on L2 and the function HQ satisfy the condition (a) and (b) in Definition 9.2. The second class of Haar shift operators is more general than the first class. One can easily observe that the operators πb, Sπb and π∗ b S do not satisfy the condition (c) on Definition 9.1, however these operators satisfy the conditions of Definition 9.4. Note that the n-variable paraproduct is a sum of 2n − 1 of the second class of Haar shift operator of index 1, the restricted n-variable dyadic paraproduct πbf = XQ∈DnhfiQhb, HQiHQ . In [14], the linear estimate for the maximal truncations of these operators is presented. This also recovers our linear bound estimates for Sπb and π∗ b S. On the other hand, authors in [7] also reproduce the linear estimate for the dyadic paraproduct with different technique. Lemma 9.5. Let Tτ a Haar shift operator of the first class, then [λb, Tτ ] is an operator of the same class. Proof. We are going to use the restricted multi-variable λb operator which is λbf = XQ∈DnhbiQhf, HQiHQ . 26 One can get the n-variable λb operator by summing over 2n − 1 of restricted λb operator. Observe that, [λb, Tτ ]f = λb(Tτ f ) − Tτ (λbf ) = XQ∈Dn XQ′,Q′′∈D(Q) = XQ∈Dn XQ′,Q′′∈D(Q) 2−τ nQ≤Q′,Q′′ 2−τ nQ≤Q′,Q′′ aQ′,Q′′hbiQ′′hf, HQ′iHQ′′ − XQ∈Dn XQ′,Q′′∈D(Q) 2−τ nQ≤Q′,Q′′ aQ′,Q′′hbiQ′hf, HQ′iHQ′′ aQ′,Q′′(hbiQ′′ − hbiQ′)hf, HQ′iHQ′′ . Since [λb, Tτ ] remains in the same class of Tτ . (cid:12)(cid:12) aQ′,Q′′(hbiQ′′ − hbiQ′)(cid:12)(cid:12) ≤ kbkBM O aQ′,Q′′ , Theorem 9.3 and Lemma 9.5 allow to extend our result to more general class of commutators including the Riesz transforms and the Beurling-Ahlfors operator as in Theorem 1.3. 10 Sharp bounds In this section, we start proving that the quadratic estimate in Theorem 1.5 is sharp, by showing an example which returns quadratic bound. This example was discovered by C. Peréz [22] who is kindly allowing us to reproduce it in this paper. The same calculations show that the bounds in Theorem 1.1 are also sharp for p 6= 2 and 1 < p < ∞ . Variations over this example will then show that the bounds in Theorem 1.3 are sharp for the Riesz transforms and the Beurling-Ahlfors operator as well. 10.1 The Hilbert transform Consider the weight, for 0 < δ < 1 : It is well known that w is an A2 weight and w(x) = x1−δ . [w]A2 ∼ 1 δ . We now consider the function f (x) = x−1+δχ(0,1)(x) and BMO function b(x) = log x . We claim that For 0 < x < 1 , we have k[b, H]f (x) ≥ 1 δ2 f (x) . [b, H]f (x) =Z 1 0 log x − log y x − y = x−1+δZ 1/x 0 y−1+δdy =Z 1 t(cid:17) log(cid:16) 1 t−1+δdt . 0 1 − t log(x/y) x − y y−1+δdy 27 Now, 0 and since log(1/t) 1−t Z 1/x log(1/t) 1 − t But since Z 1 0 log(1/t) 1 − t our claim follows and is positive for (0, 1) ∪ (1,∞) we have for 0 < x < 1 t−1+δdt . log(1/t) 1 − t 0 log(1/t) 1 − t t−1+δdt =Z 1 [b, H]f (x) > x−1+δZ 1 t−1+δdt >Z 1 0 0 k[b, H]fkL2(w) ≥ t−1+δdt +Z 1/x 1 log(1/t) 1 − t t−1+δdt , log(1/t)t−1+δdt =Z ∞ 1 δ2kfkL2(w) ∼ [w]2 0 A2kfkL2(w) . se−sδds = 1 δ2 , (10.1) (10.2) A first approximation of what the bounds in Lp(w) is given by an application of the sharp extrap- olation theorem for the upper bound, paired with the knowledge of the sharp bound on L2(w) to obtain a lower bound. Proposition 10.1. For 1 < p < ∞ there exist constants c and C only depending on p such that 2 min{1, 1 Ap p−1 } kbkBM O ≤ k[b, H]kLp(w)→Lp(w) ≤ C[w] c[w] for all b ∈ BM O . Proof. Because the upper bound in (10.3) is the direct consequence of the quadratic bound in the Theorem 1.5 and sharp extrapolation theorem, we will only prove the lower bound. Let us assume that, for 1 < r < 2 and α < 1 , kbkBM O , (10.3) 2 max{1, 1 Ap p−1 } This and the sharp extrapolation theorem return k[b, H]kLr (w)→Lr(w) ≤ C[w]2α ArkbkBM O . This contradicts to the sharpness (p = 2) . Similarly, one can conclude for p > 2 . k[b, H]kL2(w)→L2(w) ≤ C[w]2α A2kbkBM O . We now consider the weight w(x) = x(1−δ)(p−1) then w is an Ap weight with [w]Ap ∼ δ1−p . By (10.1) and (10.2) we have k[b, H]fkLp(w) ≥ 1 δ2 kfkLp(w) = (δ1−p) 2 p−1kfkLp(w) ∼ [w] p−1 Ap kfkLp(w) . 2 This shows the upper bound in (10.3) is sharp for 1 < p ≤ 2 . We use the duality argument to see the sharpness of the quadratic estimate for p > 2 . Note that the commutator is a self-adjoint operator: hbH(f ) − H(bf ), gi = hf, H ∗(bg)i − hf, bH ∗(g)i = hf, bH(g) − H(bg)i . Consider 1 < p ≤ 2 and set u = w1−p′ , then k[b, H]kLp′ (u)→Lp′ (u) = k[b, H]kLp′ (w1−p′ )→Lp′ (w1−p′ ) = k[b, H]∗kLp′ 2 (w1−p′ )→Lp′ (w1−p′ ) = k[b, H]kLp(w)→Lp(w) ≤ CkbkBM Od[w] = CkbkBM O[w1−p′ Ap′ = CkbkBM O[u]2 ]2 1−p Ap Ap′ . (10.4) Since the inequality in (10.4) is sharp, we can conclude that the result of Theorem 1.1 is also sharp for p > 2 . 28 10.2 Beurling-Ahlfors operator Recall the Beurling-Ahlfors operator B is given by convolution with the distributional kernel p.v.1/z2: Bf (x, y) = p.v. f (x − u, y − v) (u + iv)2 dudv . Then the commutator of the Beurling-Ahlfors operator can be written: [b,B]f (x, y) = p.v. b(x, y) − b(s, t) ((x − s) + i(y − t))2 f (s, t) dsdt . It was observed, in [9], that the linear bound for the Beurling-Ahlfors operator is sharp in L2(w) , with weights w(z) = zα and functions f (z) = z−α where α < 2 . Similarly, we consider weights w(z) = z2−δ where 0 < δ < 1. Note that w(z) = z2−δ : C → [0,∞) is a A2-weight with [w]A2 ∼ δ−1 . We also consider a BMO function b(x) = log z . Let E = {(r, θ) 0 < r < 1, 0 < θ < π/2} and Ω = {(r, θ) 1 < r < ∞, π < θ < 3π/2} We are going to estimate [b,B]f (z) for z ∈ Ω with a function f (z) = zδ−2χE(z) . Let z = x + iy and ζ = s + ti . Then, for z ∈ Ω , 1 πZR2 πZR2 1 [b,B]f (z) = = = 1 1 π(cid:12)(cid:12)(cid:12)(cid:12)ZE π(cid:12)(cid:12)(cid:12)(cid:12)ZE π(cid:12)(cid:12)(cid:12)(cid:12)ZE 1 dζ(cid:12)(cid:12)(cid:12)(cid:12) (b(z) − b(ζ))f (ζ) (z − ζ)2 b(x, y) − b(s, t) ((x − s) + i(y − t))2 f (s, t)dsdt(cid:12)(cid:12)(cid:12)(cid:12) (log z − log ζ) ζδ−2((x − s)2 − (y − t)2) + iZE ((x − s)2 + (y − t)2)2 (log z − log ζ) ζδ−2(2(x − s)(y − t)) ((x − s)2 + (y − t)2)2 dsdt dsdt(cid:12)(cid:12)(cid:12)(cid:12) . 4 For z ∈ Ω and ζ ∈ E , we have (x− s)(y− t) ≥ xy and by triangle inequality ((x− s)2 + (y− t)2)2 = z − ζ4 ≤ ( z + ζ)4 . After neglecting the positive term (real part), we get drdθ(cid:19)2 zdt(cid:19)2 π2(cid:18)xyZ π/2 Z 1 z4Z 1/z log(1/t)( zt)δ−1 ( z + ζ)4 log( z/r)rδ−1 log( z/ ζ) ζδ−2 log( z/r)rδ−2r [b,B]f (z)2 ≥ π2(cid:18)xyZE = x2y2(cid:18) 1 z4Z 1 = x2y2(cid:18) 1 z4−δ Z 1/z dsdt(cid:19)2 (1 + r/ z)4 dr(cid:19)2 = x2y2(cid:18) 1 dt(cid:19)2 log(1/t)tδ−1 ( z + r)4 (1 + t)4 (1 + t)4 = 4 0 0 0 0 0 . Since z/( z + 1) ≤ 1/(1 + t) , for t < 1/ z , we have [b,B]f (z)2 ≥ x2y2(cid:18) 1 z−δ( z + 1)4Z 1/z 0 log(1/t)tδ−1 dt(cid:19)2 x2y2 z−2δ( z + 1)8(cid:18) z−δ(1 + δ log z) δ2 (cid:19)2 = = x2y2 ( z + 1)8 (1 + δ log z)2 . δ4 29 Then, we can estimate the L2(w)-norm as follows. k[b,B]fk2 L2(w) ≥ = = = = r4 cos2 θ sin2 θ(1 + δ log r)2 z2−δ dxdy (r + 1)8 π 1 1 π ( z + 1)8 δ4ZΩ x2y2(1 + δ log z)2 δ4Z ∞ 1 Z 3π/2 δ416Z ∞ δ4212Z ∞ δ4212(cid:18) 1 δ3 (cid:19) = 2δ δ2 + (r + 1)8 2δ2 + π π δ 1 1 r7−δ(1 + δ log r)2 dr ≥ (1 + δ log r)2r−1−δ dr 5π 212 · 1 δ5 . r3−δdrdθ π δ416Z ∞ 1 r7−δ(1 + δ log r)2 (2r)8 dr L2(w) = π/2δ , we have that k[b,B]fkL2(w)/kfkL2(w) ∼ δ−2 , which allows to Combining with kfk2 conclude that the quadratic bound for the commutator with the Beurling-Ahlfors operators is sharp in L2(w) . Same calculations with weights w(z) = z(2−δ)(p−1) and functions f (z) = z(δ−2)(p−1) will provide the sharpness for 1 < p ≤ 2 , and it is sufficient to conclude for all 1 < p < ∞ because the Beurling-Ahlfors operator is essentially self adjoint operator (B∗ = eiφB) , so the commutator of the Beurling-Ahlfors operator is also self adjoint. 10.3 Riesz transforms Consider weights w(x) = xn−δ and functions f (x) = xδ−nχE(x) where E = {x x ∈ (0, 1)n ∩ B(0, 1)} , and a BMO function b(x) = log x . It was observed that xn−δ is an A2-weight in Rn with [w]A2 ∼ δ−1 . We are going to estimate [b, Rj]f over the set Ω = {y ∈ B(0, 1)c yi < 0 for all i = 1, 2, ..., n} , where Rj stands for the j-th direction Riesz transform on Rn and is defined as follows: Rjf (x) = cnp.v.ZRn yj yn+1 f (x − y) dy , 1 ≤ j ≤ n , where cn = Γ((n + 1)/2)/π(n+1)/2 . One can observe that, for all x ∈ E and fixed y ∈ Ω , yj − xj ≥ yj and y − x ≤ y + x . Cn denotes a constant depending only on the dimension that may change from line to line. Then, [b, Rj]f (y) =(cid:12)(cid:12)(cid:12)(cid:12)ZE (yj − xj)(log y − log x) xδ−n y − xn+1 log( y/r)rδ−nrn−1 ≥ yjZE∩Sn−1Z 1 yn+1−δZ 1/y y−δ( y + 1)n+1(cid:18) y−δ(1 + δ log y) ( y + r)n+1 log(1/t)tδ−1 (1 + t)n+1 dt ≥ Cn yj Cn yj δ2 = = 0 0 dx log( y/ x) xδ−n ( y + x)n+1 dx(cid:12)(cid:12)(cid:12)(cid:12) ≥ yjZE drdσ = Cn yjZ 1/y yn+1−δ(cid:18) y y + 1(cid:19)n+1Z 1/y Cn yj 0 0 log(1/t)(t y)δ−1 y ( y + yt)n+1 dt log(1/t)tδ−1 dt (cid:19) = Cn yj ( y + 1)n+1 1 + δ log y . δ2 30 We now can bound from below the L2(w)-norm as follows. y2 j (1 + δ log y)2 ( y + 1)2n+2 yn−δdy j r2(1 + δ log r)2rn−δrn−1 γ2 (r + 1)2n+2 drdσ(γ) k[b, Rj ]f (x)k2 L2(w) > ≥ ≥ = Cn Cn Cn Cn δ4 ZΩ δ4 ZΩ∩Sn−1Z ∞ δ4 Z ∞ δ4 Z ∞ 1 1 1 (1 + δ log r)2r2n−δ+1 dr r2n+2 (1 + δ log r)2r−δ−1 dr = Cn δ5 , which establishes sharpness for the commutator of the Riesz transform when p = 2 . Since R∗ j = −Rj, one can easily check that the commutator of Riesz transforms are also self-adjoint operators. Furthermore, choosing weight w(x) = x(n−δ)(p−1), we will obtain the sharpness for 1 < p < ∞ by the same argument we used in the case of the Hilbert transform. 11 Acknowledgments This work is part of the author's Ph.D dissertation [4]. The author like to thank Professor Carlos Pérez for his comments and useful suggestions. Finally, the author is also very grateful to his graduate adviser María Cristina Pereyra for her suggestions and helpful interaction. References [1] J. Alvarez, R. J. Bagby, D. S. Kurtz and C. Pérez, Weighted estimates for commutators of linear operators Studia Math. 104(2) (1993) 195-209. [2] O. Beznosova, Linear bound for dyadic paraproduct on weighted Lebesgue space L2(w) J. of Fun. Anal. 255 4 (2008), 994-1007. [3] S. Buckley, Summation condition on Weights, Michigan Math. J., 40(1) (1993) 153-170. [4] D. Chung, Ph.D Dissertation, University of New Mexico, (2010). [5] R. Coifman and C. Fefferman, Weighted norm inequalities for maximal functions and singular integrals Studia Math. 51 (1974), 241-250. [6] R. Coifman, R. Rochberg, and G. Weiss, Factorization theorems for Hardy spaces in several variables Ann. of Math. (2) 103 (1976), 611-635. [7] D. Cruz-Uribe, SFO, J. Martell and C. Pérez, Sharp weighted estimates for classical operators Preprint. [8] O. Dragicević, L. Grafakos, M. C. Pereyra, and S. Petermichl, Extrapolation and sharp norm estimates for classcical operators on weighted Lebesgue spaces Publ. Mat. 49 (2005), 73-91. [9] O. Dragicević and A. Volberg, Sharp estimate of the Ahlfor-Beurling operator via averaging Martingale transforms, Michigan Math. J. 51 (2003) [10] J. B. Garnett, Bounded analytic functions Acad. Press, NY, 1981. 31 [11] H. Helson and G. Szegö, A problem in prediction theory, Ann. Math. Pura. Appl. 51 (1960), 107-138. [12] S. Hukovic, S. Treil, A. Volberg, The Bellman functions and the sharp weighted inequalities for square functions Operator Theory : Advances and Applications, the volume in memory of S. A. Vinogradov, v. 113, Birkhauser Verlag, (2000) [13] R. Hunt, B. Muckenhoupt, and R. Wheeden, Weighted norm inequalities for conjugate function and the Hilbert transform Trans. Amer. Math. Soc. 176 (1973), 227-251. [14] T. Hytönen, M. Lacey, M. Reguera, and A. Vagharshakyan, Weak and Strong-type esti- mates for Haar Shift Operators: Sharp power on the Ap characteristic (2009), available at http://arxiv.org/abs/0911.0713 [15] M. Lacey, Haar shifts, Commutators, and Hankel operators CIMPA-UNESCO conference on Real Analysis and it Applications, La Falda Argentina, (2008). [16] M. Lacey, S. Petermichl, J. Piper and B. Wick, Iterated Riesz Commutators: A simple proof of boundedness Proc. of El Escorial (2008). [17] M. Lacey, S. Petermichl and M. Reguera, Sharp A2 inequality for Haar shift operators Math. Ann. to appear. [18] B. Muckenhoupt, Weighted norm inequalities for the Hardy maximal function Trans. Amer. Math. Soc. 165 (1972), 207-226. [19] F. Nazarov, S. Treil, and A. Volberg, The Bellman functions and two-weight inequalities for Haar multipliers J. Amer. Math. Soc. 12 (1999), 909-928. [20] C. Pérez, Endpoint estimates for commutators of singular integral operators J. Funct. Anal. 128 (1995), 163-185. [21] C. Pérez, Sharp estimats for commutators of singular integrals via iterations of the Hardy- Littlewood maximal functions J. Fourier Anal. Appl. 3 (1997), 743-756. [22] C. Pérez, Personal communication (2009). [23] C. Pérez and G. Pradolini, Sharp weighted Endpoint Estimates for commutators of singular integrals Michigan Math. H. 49 (2001) 23-37. [24] M. C. Pereyra, Haar Multipliers meet Bellman Functions Revista Mat. Iber. 25 3 (2009), 799- 840. [25] S. Petermichl, Dyadic shifts and a logarithmic estimate for Hankel operators with matrix symbol C. R. Acad. Sci. Paris 330 (2000), 455-460. [26] S. Petermichl, The sharp bound for the Hilbert transform on weighted Lebesgue spaces in terms of the classical Ap Characteristic, Amer. J. of Math. 129 (2007) 1355-1375. [27] S. Petermichl, The sharp weighted bound for the Riesz transforms Proc. Amer. Math. Soc. 136 (2008), 1237-1249. [28] S. Petermichl and S. Pott, An estimate for weighted Hilbert transform via square functions Trans. Amer. Math. Soc. 354 (2002), 1699-1703. 32 [29] S. Petermichl, S. Treil and A. Volberg, Why the Riesz transforms are averages of the dyadic shifts? Proc. of the 6th Inter. Conf. on Harmonic Analysis and Partial Differential Equations (El Escorial, 2000) (2002) 209-228. [30] S. Petermichl and A. Volverg, Heating of the Ahlfors-Beurling operator: Weakly quasiregular maps on the plane are quasiregular Duke Math J. 112 (2002), 281-305. [31] A. Vagharshakyan Recovering Singular Integrals from Haar Shifts (2009), available at http://arxiv.org/abs/0911.4968 [32] J. Wittwer, A shap estimate on the norm of Martingale transform Math. Res. Lett. 7 (2000) 1-12. Department of Mathematics and Statistics, 1 University of New Mexico, Albu- querque, NM 87131-0001, E-mail: [email protected] and [email protected] 33
1511.04568
1
1511
2015-11-14T15:11:21
Reducibility of invertible tuples to the principal component in commutative Banach algebras
[ "math.FA" ]
Let $A$ be a complex, commutative unital Banach algebra. We introduce two notions of exponential reducibility of Banach algebra tuples and present an analogue to the Corach-Su\'arez result on the connection between reducibility in $A$ and in $C(M(A))$. Our methods are of an analytical nature. Necessary and sufficient geometric/topological conditions are given for reducibility (respectively reducibility to the principal component of $U_n(A)$) whenever the spectrum of $A$ is homeomorphic to a subset of $\mathbb C^n$.
math.FA
math
REDUCIBILITY OF INVERTIBLE TUPLES TO THE PRINCIPAL COMPONENT IN COMMUTATIVE BANACH ALGEBRAS RAYMOND MORTINI AND RUDOLF RUPP Abstract. Let A be a complex, commutative unital Banach algebra. We introduce two notions of exponential reducibility of Banach algebra tuples and present an analogue to the Corach-Su´arez result on the con- nection between reducibility in A and in C(M (A)). Our methods are of an analytical nature. Necessary and sufficient geometric/topological conditions are given for reducibility (respectively reducibility to the prin- cipal component of Un(A)) whenever the spectrum of A is homeomorphic to a subset of Cn. 1. Introduction The concepts of stable ranges and reducibility of invertible tuples in rings originate from Hyman Bass's work [2] treating problems in algebraic K- theory. Later on, due to work of L. Vasershtein [29], these notions also turned out to be very important in the theory of function algebras and topology because of their intimate relations to extension problems. This direction has further been developed by Corach and Su´arez, [4], [5]. Function theorists have also been interested in this subject and mainly computed the stable ranks for various algebras of holomorphic functions. For example, P.W. Jones, D. Marshall and T. Wolff [9] determined the stable rank of the disk algebra A(D), and Corach and Su´arez [7] the one for the polydisk and ball algebras. The whole culminated in S. Treil's work on the stable rank for the algebra H ∞ of bounded analytic functions on the unit disk [28]. Recent work includes investigations of stable ranks for real-symmetric function algebras (see for instance [17] and [15]). The subject of the present paper is linked to the theory developped by Corach and Su´arez and provides a detailed analysis of the fine structure of the set Un(A) of invertible tuples within the realm of commutative Banach algebras. The main intention is the introduction of a new concept, the exponential reducibility of n-tuples, and to present a new view on the structure of the connected components of Un(A). In contrast to the work of Corach and Su´arez (and Lin), we use an ana- lytic framework (and methods) instead of the powerful algebraic-topological setting. We think that this makes the theory accessible to a larger reader- ship. 1 2 RAYMOND MORTINI AND RUDOLF RUPP 1.1. Notational background and scheme of the paper. Let A be a commutative unital Banach algebra over K = R or K = C, the identity element (or multiplicatively neutral element) being denoted by 1. Then the spectrum (=set of nonzero, multiplicative K-linear functionals on A) of A is denoted by M (A), and the set of all n×n-matrices over A by Mn(A). If f ∈ C(X, K), the space of all K-valued continuous functions on the topological space X, then Z(f ) := {x ∈ X : f (x) = 0}. If f = (f1, . . . , fn) ∈ C(X, Kn), j=1 Z(fj) is the joint zero-set. Moreover, if f ∈ An, then j=1 fjgj and, when viewed as an j=1 fj2, hf , gi := f · g := Pn then Z(f ) := Tn f = qPn element in An, e1 := (1, 0, . . . , 0). Finally, for f ∈ C(X, K), f ∞ = f X = sup{f (x) : x ∈ X}. Let us begin with the pertinent definitions. Definition 1.1. • An n-tuple (f1, . . . , fn) ∈ An is said to be invertible (or unimodular), if there exists (x1, . . . , xn) ∈ An such that the B´ezout equation j=1 xjfj = 1 is satisfied. The set of all invertible n-tuples is denoted by Un(A). Note that U1(A) = A−1. Pn • An (n + 1)-tuple (f1, . . . , fn, g) ∈ Un+1(A) is called reducible (in A) if there exists (a1, . . . , an) ∈ An such that (f1 + a1g, . . . , fn + ang) ∈ Un(A). • The Bass stable rank of A, denoted by bsr A, is the smallest integer n such that every element in Un+1(A) is reducible. If no such n exists, then bsr A = ∞. The following two results due to Corach and Su´arez are the key to the theory of stable ranks. Lemma 1.2. ([4, p. 636] and [6, p. 608]). Let A be a commutative, unital Banach algebra over K.Then, for g ∈ A, the set Rn(g) := {f ∈ An : (f , g) is reducible} is open-closed inside the open set In(g) := {f ∈ An : (f , g) ∈ Un+1(A)}. In particular, if φ : [0, 1] → In(g) is a continuous map and (φ(0), g) is reducible, then (φ(1), g) is reducible. Moreover, Rn(g) = gAn + Un(A). The next assertion, which gives us a relation between reducibility in a Banach algebra A and the associated uniform algebra C(M (A)) of all con- tinuous complex-valued functions on the spectrum M (A) of A, actually is one of the most important theorems in the theory of the Bass stable rank: Theorem 1.3 (Corach-Su´arez). ([5, p. 4]). Let A be a commutative uni- tal complex Banach algebra and suppose that (f1, . . . , fn, g) is an invertible (n + 1)-tuple in A. Then (f1, . . . , fn, g) is reducible in A if and only if (bf1, . . . , bfn,bg) is reducible in C(M (A)). REDUCIBILITY TO THE PRINCIPAL COMPONENT 3 Here is now the scheme of the paper. In Section two we have a look at the principal components of Mn(A) and Un(A) and in Section three we give a connection between reducibility and the extension of invertible rows to invertible matrices in the principal component of Mn(A). In the forth section of our paper we are concerned with the analogues of the results quoted above for our new notion of "reducibility of (n + 1)-tuples in A to the principal component of Un(A)" (see below for the definition). In the fifth section we apply these results and give geometric/topological conditions under which (n + 1)-tuples in C(X, K) for X ⊆ Kn are reducible, respectively reducible to the principal component of Un(C(X, K)). Let us point out that due to Vasershtein's work, the Bass stable rank of C(X, K) is less than or equal to n + 1; hence every invertible (n + 2)-tuple in C(X, K) is reducible, but in general, not every tuple having length less than n + 1 is reducible. In the sixth section we apply our results to the class of Euclidean Banach algebras. In Section 8 we give a simple proof of a result by V. Ya. Lin telling us that a left-invertible matrix L over A can be complemented to an invertible matrix over A if and only if the matrix L of its Gelfand transforms can be complemented in the algebra C(M (A)). 2. The principal components of Mn(A) and Un(A) In this section we expose for the reader's convenience several results nec- essary to develop our theory, and whose proofs we could not locate in the literature (in particular for the case of real algebras). First, let us recall that if A = (A, · ) is a commutative unital Banach algebra over K, then the principal component Exp Mn(A) (=the connected component of the identity matrix In) of the group of invertible n × n-matrices over A is given by Exp Mn(A) = {eM1 · · · eMk : Mj ∈ Mn(A)}. (see [19, p. 201]) Our description of the connected components of the set Un(A), viewed as a topological subspace of An, is based on the following classical result giving a relation between two invertible tuples that are close to each other. An elementary proof of that result (excepted the addendum) is given in [26]. Theorem 2.1. Let A = (A, · ) be a commutative unital Banach algebra over K. Suppose that f = (f1, . . . , fn) is an invertible n-tuple in A. Then there exists ε > 0 such that the following is true: j=1 gj − fj < ε there is a matrix H ∈ Mn(A) such that For each g = (g1, . . . , gn) ∈ An satisfying Pn   .  = (exp H)     g1 ... gn f1 ... fn 4 RAYMOND MORTINI AND RUDOLF RUPP In particular, g itself is an invertible n-tuple. Moreover, if u · f t = 1, then ue−H · gt = 1. Addendum: if fn = gn, then H can be chosen so that its last row is the zero vector and eH =(cid:18) eM ∗ 0n−1 1(cid:19) for some matrix M ∈ Mn−1(A). Theorem 2.2. Let A be a commutative unital Banach algebra over K. If f = (f1, . . . , fn) ∈ Un(A), then the connected component, C(f ), of f in Un(A) equals the set In particular, C(f ), is path-connected. f · Exp Mn(A). Proof. Let C = f · Exp Mn(A); that is C =  f ·(cid:16) kYj=1 exp Mj(cid:17) : Mj ∈ Mn(A), k ∈ N  . We first note that C is path-connected. To see this, let g = f · exp(M1) · · · exp(Mk), for some Mj ∈ Mn(A). Then the map φ : [0, 1] → Un(A), given by φ(t) := f · exp(tM1) · · · exp(tMk), is a continuous path joining f to g within Un(A). We claim that C is open and closed in Un(A). In fact, let g ∈ C. According to Theorem 2.1, there is ε > 0 so that for every h ∈ An with nXj=1 gj − hj < ε, there is matrix M ∈ Mn(A) such that ht = (exp M ) gt. That is h = g · exp M t. Therefore h ∈ C. Hence g is an interior point of C. Thus C is open in An. Since Un(A) itself is open in An, we conclude that C is open in Un(A). Theorem 2.1 again, there is ε > 0 so that for every h ∈ An withPn To show that C is (relatively) closed in Un(A), we take a sequence (gk) in C that converges (in the product topology of An) to g ∈ Un(A). Applying j=1 gj − hj < ε, there is matrix M ∈ Mn(A), depending on h, such that h = g · exp M t. This holds in particular for h = gk, whenever k is large. Thus g = gk · exp Mk for some Mk ∈ Mn(A), from which we conclude that g ∈ C. Hence C is closed in Un(A). Being open-closed and connected now implies that C is the maximal con- nected set containing itself. Hence, with f ∈ C, we deduce that C(f ) = C. (cid:3) REDUCIBILITY TO THE PRINCIPAL COMPONENT 5 We are now able to define the main object of this paper: Definition 2.3. Let A be a commutative unital Banach algebra over K. Then the principal component of Un(A) is the connected component of e1 in Un(A) and is given by the set P(Un(A)) := e1 · Exp Mn(A). Remark. • If n = 1, then U1(A) = A−1 and P(Un(A)) = exp A := {ea : a ∈ A}. • Let us note that in the representation e1 · Exp Mn(A) of the principal component of Un(A) any other "canonical" element ej, with ej := (0, . . . , 0, 1{z}j is admissible, too. In fact, for i 6= j, , 0, . . . , 0), is a path in Un(A) joining ei with ej, because γi,j(t) := (1 − t)ei + tej D(1 − t)ei + t ej , ei + ejE = 1. Hence ei and ej belong to the same connected component of Un(A). 3. Extension of invertible rows to the principal component An interesting connection between reducibility and extension of rows to invertible matrices (resp. to matrices in the principal component) is given in the following theorem (see for example [21, p. 311] and [20, p. 1129]). The additional property of being extendable to finite products of exponential matrices (hence to the principal component of Mn(A)), seems not to have been considered in the literature before (as far as we know). Theorem 3.1. Let A be a commutative unital Banach algebra over K with unit element 1. Suppose that u := (f1, . . . , fn, g) ∈ Un+1(A) is reducible. Then there is an invertible matrix W ∈ Mn+1(A) with determinant 1 and which is a finite product of exponential matrices such that u W = e1. In other words, u ∈ P(Un(A)). Moreover, if M = W −1, then u is the first row of M . Proof. Let f = (f1, . . . , fn). Since u = (f , g) is reducible, there exists x = (x1, . . . , xn) ∈ An with f + g x ∈ Un(A). Hence there is y = (y1, . . . , yn) ∈ An such that 1 y · (f + gx)t = 1 − g. 1 Note that we do want the element 1 − g here on the right-hand side. 6 RAYMOND MORTINI AND RUDOLF RUPP Consider the matrices   1 1 0 ... · · · 0 x1 x2 1 . . . · · · 1 . . . xn 1 y1 y2 ... yn y · xt + 1   . . . −(f1 + gx1) · · · −(fn + gxn) 1 W1 = W2 =       Since W1 =   1 0 · · · 1 0 ... · · · 0 x1 x2 . . . · · · . . . · · · 0 ... ... 0 1 xn 1   ·   1 0 ... 0 . . . · · · . . . 0 · · · · · · y1 y2 ... 1 yn 0 1 =: M1M2, it is easy to see that W1 and W2 are invertible matrices in Mn+1(A) with determinant 1 satisfying 2 (f1, . . . , fn, g) W1W2 = (0, . . . , 0, 1) ∈ An+1. Let W3 = h et (−1)net 1 (when identifying R · 1 with R), n+1 n i et 2 . . . et ∈ Mn+1(A); that is W3 =   0 (−1)n 0 ... ... 0 1 0 ... ... 0 0 0 1 0 0 . . . . . . . . . . . . . . . . . . . . . 0 0 1 0   . Note that en+1W3 = e1 and det W3 = 1. If we put W = W1W2W3, then W is invertible in Mn+1(A), det W = 1, and uW = e1, where e1 ∈ An+1. Write W −1 =(cid:20) w V (cid:21), where w is the first row. It is easy to see that u = w. Thus (cid:20) u V (cid:21) W = In+1. Hence the row u has been extended by V to an invertible matrix M := W −1. 2 The matrix multiplication here is preferably done from the left to the right: first multiply the one-row matrix with W1, then go on. REDUCIBILITY TO THE PRINCIPAL COMPONENT 7 Note that M1 and M2 in the decomposition W1 = M1M2, as well as W2, have the form In+1 + N , where N is a nilpotent matrix. Hence, In+1 + N = eB for some B ∈ Mn+1(A) (just use an appropriate finite section of the power series expansion of the real logarithm log(1 + x)). Moreover, W3 ∈ Mn+1(R) and det W3 > 0. Thus, W3 is a product of exponential matrices over R 3. Consequently, W = W1W2W3 is a finite product of exponential matrices over A. (cid:3) 4. Reducibility to the principal component Definition 4.1. Let A be a commutative unital Banach algebra over K = R or K = C. An invertible pair (f, g) ∈ U2(A) is said to be reducible to the principal component exp A of A−1 if there exists u, v ∈ A such that f + ug = ev. It is clear that if A−1 = U1(A) is connected, then the notions of "re- ducibility of pairs" and "reducibility of pairs to the principal component" coincide. Our favourite example is the disk algebra A(D). If U1(A) is dis- connected, as it is the case for the algebra C(T, C) for example, then for every f ∈ A−1 \ exp A, the pair (f, 0) is reducible, but not reducible to the principal component of A−1. This notion seems to have appeared for the first time in Laroco's work [10] in connection with the stable rank of H ∞. Criteria for various function algebras have been established by the second author of this note in [22, 23, 24, 25]. Now we generalize this notion to tuples, a fact that never before has been considered. We propose two different settings. Here is the first one (the second one will be dealt with in Section 7). Definition 4.2. Let A be a commutative unital Banach algebra over K. An invertible (n + 1)-tuple (f , g) ∈ Un+1(A) is said to be reducible to the principal component of Un(A) if there exists h ∈ An such that f + g h ∈ P(Un(A)). The following Proposition is pretty clear in the case of complex Banach algebras, since every permutation matrix P ∈ Mn(C) has a complex loga- rithm in Mn(C) (see [16]). So what does matter here, is that we consider real algebras, too. Proposition 4.3. Let A be a commutative unital Banach algebra over K and let (f , g) ∈ Un+1(A) be an invertible (n + 1)-tuple in A which is re- ducible to the principal component P(Un(A)) of Un(A). Suppose that f is a permutation of f . Then also the tuple ( f , g) is reducible to the principal component of Un(A). 3Actually, two exponentials will suffice; see [16]. 8 RAYMOND MORTINI AND RUDOLF RUPP Proof. Without loss of generality, n ≥ 2. Case 1 Let f = (f1, . . . , fn), f = (fn, f2, . . . , fn−1, f1) and 0 · · · 1 . . . S = 1   . 1 · · · 0   1 0 0 Note that S = S−1 and det S = −1. The action of S in A 7→ AS is to interchange the first and last column. Let W ∈ Mn(R) be given by (−1)n−1 W =   · · · 0 1 1 0 0 . . . . . . 1   . Then det W = 1 and en = e1W . In particular W ∈ Exp Mn(R). Now, by assumption, f + g x = e1M for some M ∈ Exp Mn(A). Hence with x := x S, f + g x = (f + g x)S = e1 M S = (enS)(M S) = en(SM S) = (e1W ) (SM S) = e1(W SM S) ∈ e1 · Mn(A) = P(Un(A)), where we have used that M ∈ Exp Mn(A) if and only if S−1M S ∈ Exp Mn(A) for every invertible matrix S; just observe that S−1 (cid:0) kYj=1 eMj(cid:1)S = kYj=1 (S−1eMj S) = kYj=1 eS−1Mj S. Case 2 Let f be an arbitrary permutation of f . Hence f = f P for some permutation matrix P . If det P > 0 then, P = eP1eP2 for some matrices Mj ∈ Mn(R) (see for example [16]). If det P < 0, then we aditionally interchange via S the first coordinate with the last one in f . Let us call this new n-tuple F . Then F = f Q for some permutation matrix Q with det Q > 0, and again Q = eQ1eQ2 for some Qj ∈ Mn(R). Now, by assumption, there exists x ∈ An such that f + g x = e1 eM1 . . . eMk for some matrices Mj ∈ Mn(A). Hence, by multiplying at the right with Q, f Q + g xQ = e1 eM1 . . . eMk eQ1eQ2. Thus (F , g) is reducible to the principal component of Un(A). The first case now implies that the same holds for ( f , g). (cid:3) REDUCIBILITY TO THE PRINCIPAL COMPONENT 9 A sufficient condition for reducibility to the principal component is given in the following technical result: Lemma 4.4. Let A be a commutative unital Banach algebra over K. Sup- pose that (f , g) ∈ Un+1(A) and n ≥ 2. Then (f , g) is reducible to the prin- cipal component of Un(A) if there exist two vectors x ∈ An and v ∈ Un(A) such that v is reducible itself with respect to some of its coordinates 4 and (f + x g) · vt = 1. Proof. Suppose that i0 6= n. Then we interchange the i0-th coordinate with the n-th coordinate in the three vectors f , x and v appearing here. The new vectors f , x and v still satisfy the B´ezout equation (f + x g) · vt = 1. Since (f , g) is reducible to the principal component of Un(A) if and only if (f , g) does (Proposition 4.3), we may assume, right at the beginning, that v is reducible with respect to its last coordinate. By Theorem 3.1, the reducibility of the row vector v implies the existence of a finite product P1 of exponential matrices over A such that vt is the first column of a matrix P1 ∈ Exp Mn(A). Hence (as matricial products) (f + x g) P1 = (f + x g) (vt ∗ ∗ ∗) = (1, x2, . . . , xn) for some xj ∈ A. If we let P2 =   1 −x2 . . . −xn 1 . . . 1   , then P2 ∈ Exp Mn(A) (because it has the form In + N , where N is nilpo- tent), and (f + x g) P1P2 = (1, x2, . . . , xn)P2 = e1. Hence f + x g ∈ e1 · Exp Mn(A) = P(Un(A)). (cid:3) In order to study the reducibility to the principal component, we introduce a certain equivalence relation on the set of n-tuples, reminiscent of that in [4]. Corach and Su´arez considered diagonal matrices M all of whose diagonal g ⇐⇒ f − g M ∈ a An for entries were invertible elements in A: f such a matrix M . The equivalence classes of that relation, though, do not seem to be compatible with the connected components of In(a); openness for example fails. CS∼ a 4 This means that there exists i0 and aj ∈ A, (j = 1, . . . , n − 1), such that for v = (v1, . . . , vi0 , . . . , vn), the vector (v1 + a1vi0 , . . . , vi0−1 + ai0−1vi0 , vi0 +1 + ai0+1vi0 , . . . , vn + anvi0 ) belongs to Un−1(A). 10 RAYMOND MORTINI AND RUDOLF RUPP Theorem 4.5. Let A be a commutative unital Banach algebra over K, a ∈ A, and consider the open set In(a) := {f ∈ An : (f , a) ∈ Un+1(A)}. Given f , g ∈ An, define the relation f exp ∼ a Then (1) (2) g ⇐⇒ ∃ x ∈ An, ∃B1, . . . , Bk ∈ Mn(A) : f + a x = g eB1 . . . eBk . is an equivalence relation on An. exp ∼ a If f ∈ In(a), then [f ] ⊆ In(a), where exp ∼ a [f ] := {h ∈ An : h f } (3) is the equivalence class associated with f . If f ∈ In(a), then (3i) (3ii) (3iii) [f ] is open in An, [f ] is a closed-open subset of In(a), [f ] is a (path)-connected set within In(a). (4) The connected components of In(a) are the equivalence classes [f ], where f ∈ In(A). Proof. (1) • exp ∼ a is reflexive: just take x = 0 and Bj = O. is symmetric: if f + a x = g eB1 . . . eBk , then • exp ∼ a g − a(cid:0)x e−Bk . . . e−B1(cid:1) = f e−Bk . . . e−B1. • exp ∼ a is transitive (here we use that in the definition of the relation exp ∼ a products of exponential matrices appear; a single exponential matrix would f 3, then there exist xj ∈ An and not be sufficient): let f 1 Ej ∈ Exp Mn(A) such that f 2 and f 2 exp ∼ a exp ∼ a f 1 + a x1 = f 2 E1 = (f 3 E2 − a x2)E1. f 1 + a(x1 + x2E1) = f 3 E2E1. f 3. Then Hence f 1 exp ∼ a (2) Let f ∈ In(a). Then there is x ∈ An such that f + a x ∈ Un(A). Now if f ∈ [f ] then, for some E ∈ Exp Mn(A). Hence f + a x = f E f + a (x + xE) = (f + a x)E ∈ Un(A), from which we conclude that ( f , a) ∈ Un+1(A). In other words, f ∈ In(a). Thus [f ] ⊆ In(a). REDUCIBILITY TO THE PRINCIPAL COMPONENT 11 (3i) To show the openness of [f ] whenever f ∈ In(a), let h ∈ [f ]. By (2), (h, a) ∈ Un+1(A). We claim that there is ε > 0 such that every h′ ∈ An with h′ − h < ε is equivalent to f . To see this, choose according to Theorem 2.1, ε > 0 so small that (h′, a) = (h, a) eH for some H ∈ Mn+1(A). In particular h′ ∈ In(a). By that same Theorem, H may be chosen so that the last column of H is zero and that for some K ∈ Mn(A) and x ∈ An. Since x eH =(cid:18)eK 0t 1(cid:19) (h′, a) = (h, a) (cid:18)eK 0t 1(cid:19) , x n n we conclude that h′ = h eK + a x. In other words, h′ ∈ [h] = [f ]. Hence [f ] is open in An. (3ii) Let (hj) be a sequence in [f ] ⊆ In(a) converging to some h′ ∈ In(a). As in the previous paragraph, if n is sufficently large, we may conclude that h′ ∈ [hj] = [f ] for j ≥ j0. Hence [f ] is (relatively) closed in In(a). Furthermore, since [f ] ⊆ In(a), we deduce from (3i) that [f ] is also open in In(a). (3iii) Let f ∈ [f ]; say f + a x = f eM1 · · · eMk for some x ∈ An and Mj ∈ Mn(A). Then the map H : [0, 1] → An given by H(t) = f etM1 · · · etMk − ta x is a continuous path joining f with f . By definition of equivalent to f ; that is H(t) ∈ [f ]. Thus [f ] is path connected. exp ∼ a , each H(t) is (4) This follows immediately from (3i)-(3iii). (cid:3) Here is the counterpart to Lemma 1.2. Theorem 4.6. Let A be a commutative unital Banach algebra over K. Then, for g ∈ A, the set Rexp := {f ∈ An : (f , g) is reducible to the principal component of Un(A)} = g An + P(Un(A)) n (g) is open-closed inside In(g). In particular, if F : [0, 1] → In(g) is a contin- uous map for which (F (0), g) is reducible to the principal component, then (F (1), g) is reducible to the principal component, too. Proof. We first note that, by definition, Rexp n (g) ⊆ In(g) and that the re- ducibility of (f , g) ∈ Un+1(A) to the principal component of Un(A) is equiv- alent to the assertion that f e1. Thus exp ∼ g (4.1) f ∈ Rexp n (g) ⇐⇒ f ∈ [e1] ⇐⇒ [f ] = [e1]. 12 RAYMOND MORTINI AND RUDOLF RUPP In other words, Rexp 4.5. n (g) = [e1]. The assertion then follows from Theorem Now if F : [0, 1] → In(g) is a curve in In(g), then C := F ([0, 1]) is (cid:3) n (g), we deduce that C ⊆ Rexp connected. Since F (0) ∈ Rexp n (g). The following corollaries are immediate (the second one is originally due to Corach and Su´arez [4]). Corollary 4.7. Let A be a commutative unital Banach algebra over K and g ∈ A. Then the following assertions are equivalent: (1) Every invertible (n + 1)-tuple (f , g) ∈ Un+1(A) is reducible to the principal component of Un(A); that is In(g) = Rexp n (g). (2) In(g) is connected. Proof. Just note that by equation (4.1), Rexp n (g) = [e1] and that [e1] is a connected set which is contained in In(g) for every g ∈ A. The result now follows from Theorem 4.5. (cid:3) Corollary 4.8. Let A be a commutative unital Banach algebra over K and g ∈ A. Then the following assertions are equivalent: (1) In(g) = Rn(g); (2) Each component of In(g) meets Un(A). Proof. Since the connected components of In(g) are the equivalence classes with f ∈ In(g) (Theorem 4.5), we have the following equivalent [f ] for exp ∼ g assertions for a given f ∈ In(g): (i) (ii) (iii) [f ] ∩ Un(A) 6= ∅, there exists u ∈ Un(A) such that u there exists u ∈ Un(A), x ∈ An, and B1, . . . , Bk ∈ Mn(A) such that exp ∼ g f , (iv) (f , g) is reducible. f + g x = u eB1 · · · eBk , Note that we actually proved a stronger result than stated, because the assertions (i)-(iv) are valid for each individual f . The following two Lemmas are very useful to check examples upon re- ducibility. They roughly say that the reducibility of (f , g) depends only on the behaviour of the Gelfand transforms of the coordinates fj of f on the (cid:3) Lemma 4.9. Let A be a commutative unital Banach algebra over C. Sup- zero set of bg. We use the following notation: bf := (bf1, . . . , bfn). pose that (f , g) ∈ Un+1(A). Let E := Z(bg). If for some u ∈ Un(A) and matrices Mj ∈ Mn(A) sup x∈E(cid:12)(cid:12)bf (x) −bu(x) expcM1(x) · · · expcMm(x)(cid:12)(cid:12) < ε, REDUCIBILITY TO THE PRINCIPAL COMPONENT 13 where ε is sufficiently small, then (f , g) is reducible in A. Proof. Let δ := min{bf (x) : x ∈ E}. Since (f , g) ∈ Un(A), we have δ > 0. Fix ε ∈ ]0, δ/2], and let b := u exp M1 · · · exp Mm ∈ Un(A) be chosen so that Consider the path ψ : [0, 1] → An given by sup x∈E bf (x) −bb(x) < ε. ψ(t) = (1 − t) f + t b. On E we then have the following estimates: (cid:12)(cid:12)(1 − t)bf + tbb(cid:12)(cid:12) = (cid:12)(cid:12)t(bb −bf ) +bf(cid:12)(cid:12) ≥ bf − t bb −bf ≥ δ − δ/2 = δ/2. Hence the tuples (ψ(t), g) are invertible in A for every t. Since for t = 1, ψ(1) = b ∈ Un(A), the tuple (ψ(1), g) is reducible in A. By Lemma 1.2, (ψ(0), g) then is reducible which in turn implies the reducibility of (f , g). (cid:3) Lemma 4.10. Let A be a commutative unital Banach algebra over C. If for some matrices Suppose that (f , g) ∈ Un+1(A). Let E := Z(bg). Mj ∈ Mn(A) sup x∈E(cid:12)(cid:12)bf (x) − e1 · expcM1(x) · · · expcMm(x)(cid:12)(cid:12) < ε, where ε is sufficiently small, then (f , g) is reducible to the principal compo- nent P(Un(A)) of Un(A). Proof. Consider the path ψ : [0, 1] → An given by ψ(t) = (1 − t) f + t b, where b := e1 · exp M1 · · · exp Mm. If then (ψ(t), g) ∈ Un+1(A) for every t ∈ [0, 1]. Now (ψ(1), g) = (b, g) is reducible to the principal component of Un(A) since 0 < ε < (1/2) min{bf (x) : x ∈ Z(bg)}, b + 0 · g = e1 · exp M1 · · · exp Mm ∈ P(Un(A)). Hence, by Theorem 4.6, (ψ(0), g) = (f , g) is reducible to the principal com- ponent of Un(A), too. (cid:3) We close this section with our main theorem, which is the analogue to the Corach-Su´arez result Theorem 1.3 ([5]). It is based on the Arens- Novodvorski-Taylor theorem ([1],[18], [27]), a version of which we recall here. Theorem 4.11 (Arens-Novodvorski-Taylor). Let A be a commutative unital complex Banach algebra, X := M (A) its spectrum and Mn(A) the Banach algebra of n × n matrices over A. 14 RAYMOND MORTINI AND RUDOLF RUPP (1i) Suppose that for some M ∈ Mn(A) there is M ∈ Mn(C(X)) such 5 Then M = exp L1 · · · exp Lm for some Lj ∈ that cM = exp M . Mn(A). (1ii) Let M ∈ Mn(A)−1. If cM belongs to the principal component of Mn(C(X))−1, then M already belongs to the principal component of Mn(A)−1. (2) Let f ∈ Un(C(X)). Then there exist g ∈ Un(A) and G1, . . . , Gm ∈ (3) Let u and v be in Un(A). Suppose that there are matrices Gj ∈ Mn(C(X)) such that f = bg exp G1 · · · exp Gm. Mn(C(X)) such that bu = bv exp G1 · · · exp Gm. Then u and v be- (4) The Gelfand transform induces a group isomorphism between the long to the same connected component of Un(A). quotient groups Mn(A)−1/Exp Mn(A) and Mn(C(X))−1/Exp Mn(C(X)). Item (2), in particular, says that every connected component of Un(C(X)) path in Un(C(X)), then u and v can be joined by a path in Un(A). Item (4) also says that every element in Mn(C(X))−1 is homotopic in Mn(C(X))−1 contains an element of the form bf := (bf1, . . . , bfn), where f ∈ Un(A). More- over, (3) is equivalent to the assertion that if bu and bv can be joined by a to cM for some M ∈ Mn(A)−1. Theorem 4.12. Let A be a commutative unital Banach algebra over C. Given an invertible tuple (f , g) ∈ Un+1(A), the following assertions are equivalent: (1) (f , g) is reducible to the principal component of Un(A); (3) (2) bf Z(bg) belongs to the principal component of Un(C(Z(bg))). (bf ,bg) is reducible to the principal component of Un(C(M (A))). Proof. (1) =⇒ (2) Let E := Z(bg). By assumption there is h ∈ Un(A) such that u := f + g h ∈ P(Un(A)). That is, there are matrices Mj ∈ Mn(A) such that u = e1 · exp M1 · · · exp Mm. If we apply the Gelfand transform and restrict to Z(bg), then Hence bf E ∈ P(Un(C(E))). bf E = e1 ·(cid:0) expcM1 · · · expcMm(cid:1)E. (2) =⇒ (1) By assumption, there exist matrices Cj ∈ Mn(C(E)) such that 5 Here cM is the matrix whose entries are the Gelfand transforms of the entries of the bf E = e1 · exp C1 · · · exp Ck. matrix M . REDUCIBILITY TO THE PRINCIPAL COMPONENT 15 ·∞ belongs to the principal component P(Un(C(E))) of Un(C(E)), which is "generated" by e1, we conclude from the Arens-Novodvorski-Taylor The- be the uniform closure of the restriction algebra bAE in Let AE = bAE C(E). Since Z(g) is A-convex, M (AE) = E ([8]). Because bf E ∈ (AE)n orem 4.11(3) that bf E belongs to the same component of Un(AE) as e1; namely the principal component P(Un(AE)) of Un(AE). Hence there are matrices Bj ∈ Mn(AE) such that bf E = e1 · exp B1 · · · exp Bm. Now, we uniformly approximate on E the matrices Bj by matrices cMj with Mj ∈ Mn(A); say sup x∈E(cid:12)(cid:12)(cid:12)bf (x) − e1 · expcM1(x) · · · expcMm(x)(cid:12)(cid:12)(cid:12) < ε By Lemma 4.10, (f , g) is reducible to the principal component P(Un(A)) of Un(A). (2) =⇒ (3) follows from Lemma 4.10 and (3) =⇒ (2) is clear. (cid:3) If n = 2, then the previous result reads as follows: Corollary 4.13. Let A be a commutative unital Banach algebra over C. Given an invertible pair (f, g) ∈ U2(A), the following assertions are equiva- lent: (1) There exist a, h ∈ A such that f + ag = eh; (2) bf Z(bg) = ev for some v ∈ C(Z(bg)). 5. Reducibility in C(X, K) with X ⊆ Kn In this section we study the reducibility in C(X, K) with X ⊆ Kn in detail. Definition 5.1. a) Let K ⊆ Rn be compact. A bounded connected component of Rn \ K is called a hole of K. b) Let K, L be two compact sets in Rn with K ⊆ L. The pair (K, L) is said to satisfy the hole condition if every hole of K contains a hole of L. The following concepts were introduced in C by the second author in [22]. Definition 5.2. Let K ⊆ Rn be compact and g ∈ C(K, K). Then g is said to satisfy the boundary principle if for every nonvoid open set G in Rn with G ⊆ K the following condition holds: (B1) If g ≡ 0 on ∂G, then g ≡ 0 on G. Proposition 5.3. Condition (B1) is equivalent to the following assertion: (B2) If G is an open set in Rn such that G ⊆ K \ Z(g), then there exists x0 ∈ ∂G such that g(x0) 6= 0. 16 RAYMOND MORTINI AND RUDOLF RUPP Proof. Assume that g satisfies (B1) and let G ⊆ K \ Z(g) be open. Then g cannot vanish identically on ∂G since otherwise (B1) would imply that g ≡ 0 on G. A contradiction to the assumption that G ∩ Z(g) = ∅. Hence g satisfies (B2). Conversely, let g satisfy (B2). Suppose, to the contrary, that g does not satisfy condition (B1). Then there is an open set G in Rn with G ⊆ K, g ≡ 0 on ∂G, but such that g does not vanish identically on G. Hence U := {x ∈ G : g(x) 6= 0} is an open, nonvoid set in Rn which is contained in K \Z(g). But ∂U ⊆ Z(g), because ∂U = U \ U ⊆ G \ U ⊆ (G \ U ) ∪ ∂G ⊆ Z(g). This contradicts condition (B2) for U . Hence such a set G cannot exist and we deduce that g has property (B1). (cid:3) The following result gives an interesting connection between the hole con- dition (Definition 5.1) and the boundary principle (Definition 5.2). Theorem 5.4. Let K ⊆ Rn be compact and g ∈ C(K, K). The following assertions are equivalent: (1) g satisfies the boundary principle (B1). (2) (Z(g), K) satisfies the hole condition; that is, every hole of Z(g) contains a hole of K 6. Proof. (1) =⇒ (2) Suppose that there is a component Ω of Rn \ Z(g) with Ω ⊆ K. Then Ω is open in Rn and ∂Ω ⊆ Z(g). Condition (B1) now implies that g ≡ 0 on Ω (note that by assumption Ω ⊆ K). Hence Ω ⊆ Z(g). A contradiction. (2) =⇒ (1) We show that the equivalent condition (B2) is satisfied. So let G be open in Rn and assume that G ⊆ K \ Z(g). Let Ω be a component of G. Then Ω is bounded and open in Rn. In view of achieving a contradiction, suppose that g ≡ 0 on ∂Ω. Then ∂Ω ⊆ Z(g). Since Ω ∩ Z(g) = ∅, we deduce that Ω belongs to a connected component C of Rn \ Z(g). If Ω is a proper subset of C, then a path connecting z0 ∈ Ω and w ∈ C \ Ω within C would pass through a boundary point z1 of Ω. But then g(z1) = 0, contradicting z1 ∈ C. Hence Ω = C. Since Ω is bounded, we conclude that Ω is a hole of Z(g). But Ω ⊆ G ⊆ K; thus (Z(g), K) cannot satisfy the hole condition (2). This is a contradiction. Hence there is x0 ∈ ∂Ω such that g(x0) 6= 0. Since ∂Ω ⊆ ∂G (note that Ω is supposed to be a component of G), we are done. (cid:3) Let us emphasize that for Z(g) ⊆ K ⊆ Rn the pair (Z(g), K) automat- ically satisfies the hole condition (and equivalently the boundary principle (B1)) if K ◦ = ∅. An important class of zero-sets satisfying the equivalent conditions (1) and (2) is given in the following example: 6 or which is the same, no hole of Z(g) is entirely contained in K. REDUCIBILITY TO THE PRINCIPAL COMPONENT 17 Example 5.5. Let K ⊆ Rn be compact and g ∈ C(K, K). Then Z(g) has property (B1) if: (i) Rn \ Z(g) is connected whenever n ≥ 2; (ii) R \ Z(g) has exactly two components whenever n = 1. The next two theorems give nice geometric/topological conditions for re- ducibility of n-tuples, respectively for reducibility to the principal compo- nent of Un(C(X, K)). They generalize the corresponding facts for pairs developed by the second author in [22] for certain planar compacta. Theorem 5.6. Let K ⊆ Kn be compact and g ∈ C(K, K). The following assertions are equivalent: (1) C(K, K) is n-stable at g, that is (f , g) is reducible for every f = (f1, . . . , fn) ∈ C(K, Kn) such that (f , g) ∈ Un+1(C(K, K)). (2) f Z(g) admits a zero-free extension to K for all f ∈ C(K, Kn) with Z(f ) ∩ Z(g) = ∅. (3) g satisfies the boundary principle (B1). (4) (Z(g), K) satisfies the hole condition. Proof. i) The equivalence of (1) with (2) is well-known (see [5] or [26]). The equivalence of (2) with (4) is [14, Theorem 5.6], provided we identify (u1 + iv1, . . . un + ivn) in the complex-valued case with the real-valued (2n)- tuple (u1, v1, . . . , un, vn). The equivalence of (3) with (4) is Theorem 5.4. (cid:3) Theorem 5.7. Let K ⊆ Kn be compact and g ∈ C(K, K). The following three assertions are equivalent: (1) (2) (f , g) is reducible to the principal component of Un(C(K, K)) for every f ∈ C(K, Kn) such that (f , g) ∈ Un+1(C(K, K)); there exist matrices Bj ∈ Mn(C(Z(g), K)) such that f Z(g) = e1 · eB1 · · · eBk for every f ∈ C(K, Kn) with Z(f ) ∩ Z(g) 6= ∅; (3) Z(g) has no holes in Kn. Proof. First we note that (1) and (2) are equivalent in view of Theorem 4.12. Since we have to deal here only with the special case C(K, K), the following simple proof is available: (1) =⇒ (2) If f + g h ∈ P(Un(C(K, K))) for some h ∈ C(K, Kn) then, by using the representation of the principal component, f + g h = e1 · eB1 · · · eBk for some matrices Bj ∈ Mn(C(K, K)). Restricting this identity to Z(g) yields the assertion (2). (2) =⇒ (1) This follows from Lemma 4.10. (2) =⇒ (3) Suppose, to the contrary, that Z(g) admits a bounded com- plementary component G ⊆ Kn. Then ∂G ⊆ Z(g). Let a ∈ G and f (z) = 18 RAYMOND MORTINI AND RUDOLF RUPP z − a, z ∈ Z(g). Since (f , g) ∈ Un+1(C(Z(g), K)), there exist by hypothesis (2) matrices Bj ∈ Mn(C(Z(g), K)) such that f Z(g) = e1 · eB1 · · · eBk . Extending via Tietze's result the matrices Bj continuously to Kn, would yield a zero-free extension of the n-tuple (z − a)Z(g) to G. This contradicts a corollary to Brouwer's fixed point theorem (see [3, Chapter 4]). (3) =⇒ (2) By a standard result in vector analysis (see for example [14, Corollary 5.8]), the connectedness of Kn \ Z(g) implies that the invertible tuple f Z(g) ∈ Un(C(Z(g), K)) admits a zero-free extension F to Kn. Let B ⊆ Kn be a closed ball whose interior contains X. Note that B is a contractible Hausdorff space. Hence, the set Un(C(B, K)) is connected. eB1 · · · eBk for some matrices Bj ∈ Mn(C(B, K)). Thus, F B = e1 Restricting to Z(g) yields the assertion (2). (cid:3) · 6. Reducibility in Euclidean Banach algebras Let us call a complex commutative unital Banach algebra A a Euclidean Banach algebra if the spectrum M (A) of A is homeomorphic to a compact set in Cn. This class of algebras includes every finitely generated Banach algebra over C, for example the algebras and certain algebras 7 of type P (K) = C[z1, . . . , zn](cid:12)(cid:12)K ·K A(K) = {f ∈ C(K, C) : f holomorphic in K ◦}, K ⊆ Cn compact (for example K = D or K = Bn, the closed unit ball in Cn). Using the Arens-Novodvorski-Taylor theorem we can now generalize Theorems 5.6 and 5.7 to Euclidean Banach algebras. n Theorem 6.1. Let A be a Euclidean Banach algebra with spectrum X ⊆ Cn and let g ∈ A. Then the assertions (1)-(3), respectively (4)-(5), are equivalent: (f , g) is reducible for every n-tuple f ∈ An with (f , g) ∈ Un+1(A). (1) (2) (4) (Z(bg), X) satisfies the hole condition. (f , g) is reducible to the principal component of Un(A) for every n-tuple f ∈ An with (f , g) ∈ Un+1(A). (3) Z(bg) satisfies the boundary principle in Cn. (5) Z(bg) has no holes in Cn. need to show that for every f ∈ C(X, Cn) with Z(f )∩ Z(bg) = ∅, the (n + 1)- tuple (f ,bg) is reducible in C(X, C) (resp. reducible to P(Un(C(X, C))) ). (1) =⇒ (2) (resp. (4) =⇒ (5) ) By Theorem 5.6 (resp. Theorem 5.7) we Proof. First we note that by Theorem 5.4, (2) and (3) are equivalent. 7 We do not know whether every algebra A(K) is finitely generated. REDUCIBILITY TO THE PRINCIPAL COMPONENT 19 ·E Let E := Z(bg). Consider the algebra B := bAE closure of the restriction algebra bAE to E. Since E is the zero set of the Gelfand transform of a function in A, E is A-convex and so the spectrum, M (B), of B coincides with E (see [8]). Note that f E ∈ Un(C(E, C)). By the Arens-Novodvorski-Taylor Theorem 4.11 (2), there exist h ∈ Un(B) and G1, . . . , Gm ∈ Mn(C(E, C)) such that , that is the uniform f E = h exp G1 · · · exp Gm. In particular h ≥ δ > 0 on E. By Tietze's extension theorem, we may assume that Gj ∈ Mn(C(X, C)). Using the definition of B, choose a ∈ An so that 8 h(x) −ba(x) < ε M HS , (6.1) sup x∈E where M := exp G1 · · · exp Gm and where ε is so small that Z(a) ∩ Z(bg) = Z(a) ∩ E = ∅. Thus (a, g) ∈ Un+1(A). The hypothesis (1) (resp. (4) implies that (a, g) is reducible in A (resp. reducible to the principal component of Un(A)). That is, there is x ∈ An such that u := a + x g ∈ Un(A) (resp. u ∈ P(Un(A))). In particular, bu = ba on E. Since vM ≤ v · M HS for every vector v, we have the following estimates on E: (cid:12)(cid:12)f E −bu exp G1 · · · exp Gm(cid:12)(cid:12) = (cid:12)(cid:12)f E −ba exp G1 · · · exp Gm(cid:12)(cid:12) = (cid:12)(cid:12)(h −ba) exp G1 · · · exp Gm(cid:12)(cid:12) ≤ h −ba M HS < ε. from Lemma 4.9 (resp. Lemma 4.10), applied to the algebra C(X, C), that (f , g) is reducible in C(X, C) (resp. reducible to the principal component of C(X, C) (whenever ε > 0 is small). Since bu ∈ Un(C(X, C)) (resp. bu ∈ P(Un(C(X, C)))), we deduce (2) =⇒ (1) (resp. (5) =⇒ (4) ) Let (f , g) ∈ Un+1(A). Hence bf andbg have (2) (resp. (5)) and the assumption M (A) = X imply that (bf ,bg) is reducible in C(X, C) (resp. reducible to the principal component of Un(C(X, C))). By the Corach-Su´arez Theorem 1.3 (resp. Theorem 4.12) (f , g) is reducible in A (resp. reducible to the principal component of Un(A)). (cid:3) no common zeros on X. By Theorem 5.6, (resp. Theorem 5.7) hypothesis 7. Exponential reducibility II Here we introduce our second notion of exponential reducibility. 8Here M HS is the Hilbert-Schmidt norm. 20 RAYMOND MORTINI AND RUDOLF RUPP Definition 7.1. Let A be a commutative unital Banach algebra over K with identity element 1. Given (a, g) := (a1, . . . , an, g) ∈ Un+1(A), we call (a, g) exponentially reducible if there exists xj, bj ∈ A such that nXj=1 exj (aj + bjg) = 1. Observation 7.2. Let A be a commutative unital Banach algebra over K such that (1) bsr A = 1, (2) U1(A) is connected. Then every invertible pair (a, g) ∈ A is exponentially reducible. Proof. By (1), a + bg ∈ U1(A) for some b ∈ A. Since U1(A) = exp A, we arrive at a + bg = ex for some x ∈ A. This is of course equivalent to say that e−x(a + bg) = 1. (cid:3) The following result gives a relation between exponential reducibility and reducibility to the principal component. Proposition 7.3. Let A be a commutative unital Banach algebra over K and let (a, g) ∈ Un+1(A). Suppose that (a, g) is exponentially reducible. Then (a, g) is reducible to the principal component of Un(A). Proof. By assumption,Pn Un(A) and b := (b1, . . . , bn) ∈ An; that is j=1 exj (aj+bjg) = 1 for some v := (ex1 , . . . , exn) ∈ (a + b g) · vt = 1. Since v ∈ Un(A) is reducible in A (if n ≥ 2), we deduce from Lemma 4.4 that a + b g ∈ e1 · Exp Mn(A) = P(Un(A)). In other words, (a, g) is reducible to the principal component of Un(A). The case n = 1 is obvious. (cid:3) Remark Whereas for invertible pairs both notions coincide, exponential reducibility of tuples of length at least three is, in general, a much stronger requirement than reducibility to the principal component. As an example, we take the disk algebra A(D). Let (z, f ) ∈ U2(D) be an invertible pair that is not totally reducible (this means that there do not exist two invertible functions u and v in A(D) such that uz + vf = 1, see [13] for the existence). Then (z, f, 0) ∈ U3(A(D)). Since bsr A(D) = 1, U2(A(D)) is connected ([5]), and so U2(A(D)) coincides with its principal component. In particular every invertible triple in A(D) is reducible to the principal component of U2(A(D)). On the other hand, (z, f, 0) cannot be exponentially reducible, since otherwise ea1(z)(z + b1(z) · 0) + ea2(z)(f (z) + b2(z) · 0) = 1 REDUCIBILITY TO THE PRINCIPAL COMPONENT 21 for some functions aj, bj ∈ A(D). But an equation of the form ea1(z)z + ea2(z)f (z) = 1 is not possible because by our choice, (z, f ) is not totally reducible. Examples of exponentially reducible tuples appeared in [10] (for pairs) and [12] (for tuples): Example 7.4. Let fj ∈ H ∞ and let b be an interpolating Blaschke product. Suppose that (f1, . . . , fn, b) ∈ Un+1(H ∞). Then (f1, . . . , fn, b) is exponen- tially reducible. We do not know yet a characterization of the exponentially reducible tuples (in none of the standard algebras). 8. Complementing left-invertible matrices Based on the Arens-Novodvorski-Taylor theorem, we conclude our paper by giving a simple analytic proof of a result by V. Ya. Lin [11, p. 127] concerning extension of left invertible matrices. Although this proof seems to be known among the specialists in the field, it never appeared explicitely in print (see also the footnote in [11, p. 127]). One may view this result as a companion result to Theorem 4.12 and Theorem 3.1. Theorem 8.1 (Lin). Let A be a commutative unital Banach algebra over C. Then a left-invertible matrix L over A can be complemented/extended to an invertible matrix over A if and only if L 9 can be complemented in the algebra C(M (A)). Proof. It is sufficient to consider invertible rows (see [30, p. 345/346]). So let a ∈ Un(A). Suppose that there exists an invertible matrix M ∈ By the Arens-Novodvorski-Taylor Theorem 4.11 (4), there is Q ∈ Mn(A)−1 Mn(C(M (A))) such that ba = e1M ; that is, ba is the first row of M . such that bQ is homotopic in Mn(C(M (A)))−1 to M . Hence, for some Gj ∈ Mn(C(M (A))), and so M = bQ eG1 . . . eGk , ba = (e1bQ) eG1 . . . eGk . Let b be the first row of Q; that is b = e1Q. Then bb = e1bQ. Since b ∈ Un(A), we see that ba and bb belong to the same component of Un(C(M (A))). Hence, by another application of the Arens-Novodvorski Theorem 4.11 (3), a and b belong to the same component of Un(A). Thus, there exist Hj ∈ Mn(A) such that a = b eH1 · · · eHk . 9 Here L = (dai,j) is the matrix formed with the Gelfand-transforms of the entries of L. 22 RAYMOND MORTINI AND RUDOLF RUPP Consequently, a = (e1Q) eH1 · · · eHk = e1 R for some R ∈ Mn(A)−1. (cid:3) Acknowledgements We thank Alexander Brudnyi for valuable com- ments on the Arens-Novodvorski-Taylor Theorem 4.11 and the referee for his suggestions. References [1] Arens, R., To what extent does the space of maximal ideals determine the algebras? in: Function Algebras, Proc. Internat. Sympos. on Function Algebras, Tulane Univ., (1965-1966), 164 -- 168. 13 [2] Bass, H., K-theory and stable algebra, Publications Math´ematiques de L'I.H. ´E.S., 22 (1964), 5 -- 60. 1 [3] Burckel, R.B., An Introduction to Classical Complex Analysis, Birkhauser, Basel+Stuttgart, 1979. 18 [4] Corach, G. and Su´arez, F. D., Stable rank in holomorphic function algebras, Illinois J. Math. 29 (1985), 627 -- 639. 1, 2, 9, 12 [5] Corach, G. and Su´arez, F. D., Extension problems and stable rank in commutative Banach algebras, Topology and its Applications 21 (1985), 1 -- 8. 1, 2, 13, 17, 20 [6] Corach, G. and Su´arez, F. D., On the stable range of uniform algebras and H ∞, Proc. Amer. Math. Soc. 98 (1986), 607 -- 610. 2 [7] Corach, G. and Su´arez, F. D., Dense morphisms in commutative Banach algebras, Trans. Amer. Math. Soc. 304 (1987), 537 -- 547. 1 [8] Gamelin, T.W., Uniform algebras, Chelsea, New York, 1984. 15, 19 [9] Jones, P.W., Marshall, D., and Wolff, T.H., Stable rank of the disc algebra, Proc. Amer. Math. Soc. 96 (1986), 603-604. 1 [10] Laroco, L. A., Stable rank and approximation theorems in H ∞, Trans. Amer. Math. Soc. 327 (1991), 815 -- 832. 7, 21 [11] Lin, V. Ya., Holomorphic fiberings and multivalued functions of elements of a Banach algebra, Funct. Anal. Appl. 7 (1973), 122 -- 128; translation from Funkts. Anal. Prilozh. 7 (1973), 43 -- 51. 21 [12] Mortini, R., An example of a subalgebra of H ∞ on the unit disk whose stable rank is not finite, Studia Math. 103 (1992), 275 -- 281. 21 [13] Mortini, R. and Rupp, R., Totally reducible elements in rings of analytic functions, Comm. Algebra 20 (1992), 1705 -- 1713. 20 [14] Mortini, R. and Rupp, R., Mappings into the Euclidean sphere, Amer. Math. Monthly 119 (2012), 485 -- 494. 17, 18 [15] Mortini, R. and Rupp, R., The Bass stable rank for the real Banach algebra A(K)sym, J. Funct. Anal. 261 (2011), 2214 -- 2237. 1 [16] Mortini, R. and Rupp, R., Logarithms and exponentials in Banach algebras, Banach J. Math. Anal. 9 (2015), 164 -- 172. 7, 8 [17] Mortini, R. and Wick, B., The Bass and topological stable ranks of H ∞ R (D) and AR(D), Crelle Journal (Journal fur Reine und Angewandte Mathematik) 636 (2009), 175 -- 191. 1 [18] Novodvorski, M.E., Certain homotopical invariants of spaces of maximal ideals, Mat. Zametki 1 (1967), 487 -- 494. 13 [19] Palmer, T.W., Banach algebras and the general theory of ∗-algebras, Vol 1, Cam- bridge Univ. Press, London, 1994. 3 REDUCIBILITY TO THE PRINCIPAL COMPONENT 23 [20] Quadrat, A. and Robertz, D., Computation of bases free modules over the Weyl algebras, J. Symbolic Comp. 42 (2007), 1113 -- 1141. 5 [21] Rieffel, M., Dimension and stable rank in the K-theory of C ∗-algebras, Proc. Lon- don Math. Soc. 46 (1983), 301 -- 333. 5 [22] Rupp, R., Stable rank and boundary principle, Topology Appl. 40 (1991), 307 -- 316. 7, 15, 17 [23] Rupp, R., Stable rank and the ∂-equation, Canad. Math. Bull. 34 (1991), 113 -- 118. 7 [24] Rupp, R., Stable rank in Banach algebras, In: Function spaces (Edwardsville, IL, 1990), Lecture Notes in Pure and Appl. Math., 136, Dekker, New York, 1992, 357 -- 365. 7 [25] Rupp, R., Analytic functions on circular domains, Travaux mathmatiques, IV, 1 -- 81, S´em. Math. Luxembourg, Centre Univ. Luxembourg, Luxembourg, 1992. 7 [26] Rupp, R., Zerofree extension of continuous functions on a compact Hausdorff space, Topology Appl. 93 (1999), 65 -- 71. 3, 17 [27] Taylor, J.L., Topological invariants of the maximal ideal space of a Banach algebra, Advances in Math. 19 (1976), 149 -- 206. 13 [28] Treil, S., The stable rank of the algebra H ∞ equals 1, J. Funct. Anal. 109 (1992), 130 -- 154. 1 [29] Vasershtein, L., Stable rank of rings and dimensionality of topological spaces, Funct. Anal. Appl. 5 (1971), 102 -- 110; translation from Funkts. Anal. Prilozh. 5 (1971), No.2, 17 -- 27. 1 [30] Vidyasagar, M., Control System Synthesis: A Factorization Approach, MIT Press Series in Signal Processing, Optimization, and Control, 7, MIT Press, Cambridge, MA, 1985. 21 Raymond Mortini, D´epartement de Math´ematiques et Institut ´Elie Cartan de Lorraine, UMR 7502, Universit´e de Lorraine, Metz, France, raymond.mortini@univ- lorraine.fr and Rudolf Rupp, Fakultat fur Angewandte Mathematik, Physik und Allgemeinwissenschaften, TH-Nurnberg, Germany, [email protected]
1709.07060
1
1709
2017-09-20T20:05:24
Continuity and growth of free multiplicative convolution semigroups
[ "math.FA" ]
Let $\mu$ be a compactly supported probability measure on the positive half-line and let $\mu^{\boxtimes t}$ be the free multiplicative convolution semigroup. We show that the support of $\mu^{\boxtimes t}$ varies continuously as $t$ changes. We also obtain the asymptotic length of the support of these measures.
math.FA
math
CONTINUITY AND GROWTH OF FREE MULTIPLICATIVE CONVOLUTION SEMIGROUPS XIAOXUE DENG AND PING ZHONG Abstract. Let µ be a compactly supported probability measure on the positive half-line and let µ⊠t be the free multiplicative convolution semigroup. We show that the support of µ⊠t varies continuously as t changes. We also obtain the asymptotic length of the support of these measures. 7 1 0 2 p e S 0 2 ] . A F h t a m [ 1 v 0 6 0 7 0 . 9 0 7 1 : v i X r a 1. Introduction Let µ and ν be probability measures on [0, ∞). The free convolution µ ⊠ ν represents the distribution of product of two positive operators in a tracial W ∗-probability space whose distributions are µ and ν respectively. We refer to [14] for an introduction to free probability theory. Given a probability measure µ on [0, ∞) not being a Dirac measure, it is known [4] that, for any t > 1, the fractional free convolution power µ⊠t is defined appropriately, such that it interpolates the discrete convolution semigroup {µ⊠n}n∈Z+ , where µ⊠n = µ ⊠ · · · ⊠ µ is the n-fold free convolution. This is the multiplicative analogue of Nica-Speicher semigroup [12] defined firstly for the free additive convolution. The free convolution semigroups obey many regularity properties and has been studied extensively. See [15] for a survey on free convolutions and other topics in free probability theory. Denote by supp(µ) the support of the measure µ. We will prove the following result about continuity of the support of µ⊠t, which is the analogue of the work [16] for free multiplicative convolution on the positive half line. Theorem 1.1. Let µ be a compactly supported probability measure on [0, ∞). Then the supports of measures {supp(µ⊠t} change continuously in the Hausdorff metric with respect to the parameter t. We then study the asymptotic size of the support. Theorem 1.2. Let µ be a compactly supported probability measure on [0, ∞) with mean m1(µ) = R ∞ 0 s dµ(s) = 1. We denote by µ⊠t = max{m : m ∈ supp(µ⊠t)}. Then µ⊠t/t = eV. lim t→∞ where V is the variance of µ. Theorem 1.2 generalizes Kargin's work [11] (see also [2]) to continuous semigroup. Our proof is different from Kargin's proof, but uses the density formula for free convolution semigroups [10, 18]. The free multiplicative convolution on the unit circle is usually studied together with the positive half line case. It was shown in [1] that many results can be 1 2 deduced from results on free additive convolutions. Since analogue results were known in additive case, a separate work on free multiplicative convolution on the unit circle become unnecessary. Hence we focus on measures on the positive half line in our article. An estimation for the size of support of free additive convolution semigroups was obtained in [3, 8]. In light of the proof of Theorem 1.2, it is very likely higher order asymptotic expansion can be obtained for free additive convolution semigroups. We plan to investigate it in a forthcoming work as it may have some applications in quantum information theory. The paper is organized as follows. In Section 2, we collect some known regularity results about free multiplicative convolution semigroups. In Section 3, we give the proof for The- orem 1.1. We study the asymptotic behavior of free multiplicative convolution semigroups and give the proof for Theorem 1.2 in Section 4. 2. Free convolution on the positive half line Let µ be a probability measure on [0, ∞). The ψ-transform of µ is the moment generating function of µ defined as ψµ(z) = Z ∞ 0 zs 1 − zs dµ(s), which is analytic on Ω = C\[0, ∞). The η-transform of µ is defined as ηµ = ψµ/(1 + ψµ) on the same domain as the ψ-transform. Any probability measure µ on [0, ∞) can be recovered from its η-transform by Stieltjes inversion formula. Indeed, we have the identity (2.1) Gµ(cid:18) 1 z(cid:19) = z 1 − ηµ(z) , z ∈ Ω, where Gµ is the Cauchy transform of µ. µ(z) > 0 for z < 0, and therefore ηµ(−∞, 0) is invertible. Let η−1 µ be the inverse of ηµ and set Σµ(z) = η−1 µ (z)/z, where z < 0 is sufficiently small. The free convolution of two such probability measures µ and ν is determined by Σµ⊠ν(z) = Σµ(z)Σν (z). In particular, the n-th order free multiplicative convolution power µ⊠n of µ satisfies the identity If µ is not a Dirac measure, then η′ (2.2) Σµ⊠n (z) = Σn µ(z), where z < 0 is sufficiently small. We now briefly recall the construction of µ⊠t which interpolates the relation (2.2) as follows. Let κµ(z) = z/ηµ(z) for z ∈ Ω and one can write κµ(z) = exp(u(z)), where u is an analytic function on Ω and can be expressed as (2.3) u(z) = a +Z ∞ 0 1 + zs z − s dρ(s), where a = − log ηµ(i) and ρ is a finite positive Borel measure on [0, ∞) following [10, Proposition 4.1]. We define Φt(z) := z exp[(t − 1)u(z)]. It turned out that the function Φt is the right inverse of Voiculescu subordination function ωt [4]. More precisely, we have Φt(ωt(z)) = z, and ηµ⊠t (z) = ηµ(ωt(z)) 3 for all z ∈ Ω and t > 1. It turns out that the function ωt can be regarded as the η-transform of a ⊠-infinitely divisible measure on [0, ∞) and the function ηµ⊠t can be retrieved from ωt. We refer to [4, 10] for more details. The following result was proved in [4]. Theorem 2.1. Let µ be a probability measure on [0, ∞) and t > 1. (1) A point x ∈ (0, ∞) satisfies ηµ⊠t (x) = 1 if and only if x−1/t is an atom of µ with mass µ({x−1/t}) ≥ (t − 1)/t. If µ({x−1/t}) > (t − 1)/t, then 1/x is an atom of µ⊠t, and µ⊠t ({1/x}) = tµ({x−1/t}) − (t − 1). (2) The nonatomic part of µ⊠t is absolutely continuous and its density is continuous except at the finitely many points x such that η⊠t(x) = 1. (3) The density of µ⊠t is analytic at all points where it is different from zero. The study of regularity property of free convolutions relies on Voiculescu's subordination result [4, 5, 7, 13]. By a careful study of boundary behavior of subordination functions, we were able to give a formula for the density function of absolutely continuous part of µ⊠t in [10]. To describe our result, we need some auxiliary functions studied in [10]. Let g be a function defined on (0, ∞) × (0, π) by (2.4) g(r, θ) = − ℑu(reiθ) θ = r sin θ θ Z ∞ 0 s2 + 1 r2 − 2rs cos θ + s2 dρ(s). The function g(r, θ) is decreasing on (0, π) for any r ∈ (0, ∞) fixed and limθ→π− g(r, θ) = 0. We then set g(r) = lim θ→0 g(r, θ) = Z ∞ 0 At(r) = inf (cid:26)θ ∈ (0, π) : g(r, θ) < and r(s2 + 1) (r − s)2 dρ(s) t − 1(cid:27) . 1 It is clear that if At(r) > 0, then g(r, At(r)) = 1/(t − 1). We further let ht(r) := Φt(reiAt(r)) and t = (cid:26)r ∈ (0, ∞) : g(r) > V + 1 t − 1(cid:27) . The function ht is a homeomorphism of [0, ∞) and limr→∞ ht(r) = ∞. The functions defined above can be used to describe the image set Ωt = ωt(C+). The set t (C+ ∪ (−∞, 0)) having the negative half plane Ωt is in fact the connected component of Φ−1 as part of its boundary. Moreover, we proved that and Ωt = {reiθ : At(r) < θ < π, r ∈ (0, ∞)}, ∂Ωt = (−∞, 0] ∪ {reiθ : θ = At(r), r ∈ (0, ∞)}. The following result is one of main results in [10]. Theorem 2.2. Suppose that µ is a probability measure on [0, ∞) not being a Dirac measure and t > 1. Let St = (cid:8)1/ht(r) : r ∈ V + (1) The measure (µ⊠t)ac is concentrated on the closure of St. t (cid:9). Then the following statements hold. 4 (2) The density of (µ⊠t)ac is analytic on the set St and is given by d(µ⊠t)ac dx (cid:18) 1 ht(r)(cid:19) = 1 π ht(r)lt(r) sin θt(r) 1 − 2lt(r) cos θt(r) + l2 t (r) , r ∈ V + t , where lt(r) = r exp ℜu(reiAt(r)) and θt(r) = tAt(r)/(t − 1) for r ∈ V + t . (3) The number of components in supp(µ⊠t)ac is a decreasing function of t. 3. Continuity of free convolution semigroups In this section, we assume that the probability measure µ on [0, ∞) is compactly sup- ported. Lemma 3.1. Let µ be a probability measure on [0, ∞). Set κµ(z) = z/ηµ(z) and write κµ(z) = exp[u(z)], where u is give by (2.3). Then limx→0− κ(x) = 1/m1(µ) and 1 s Moreover, when m1(µ) = 1, we have Z ∞ 0 dρ(s) = log m1(µ) + a. Z ∞ 0 1 + s2 s2 dρ(s) = V, where V is the variance of µ. Proof. Observe that limr→0− ηµ(r)/r = m1(µ) by the definition of ηµ. We write u as u(z) = a − zZ ∞ and hence limr→0− u(r) = a −R ∞ We calculate 0 1 dρ(s) − (z2 + 1)Z ∞ 0 1 s − z dρ(s), 0 1/s dρ(s). We then deduce the first equation. u′(r) = −Z ∞ s2 + 1 (r − s)2 dρ(s), 0 for all r < 0 and, by Monotone Convergence Theorem, s2 + 1 lim r→0− u′(r) = −Z ∞ 0 dρ(s). s2 On the other hand, when m1(µ) = 1, we have u(z) = − ln (ηµ(z)/z) = − ln(1 + V z + o(z)), which yields that u′(0−) = −V . Therefore, we deduce that V = Z ∞ 0 s2 + 1 s2 dρ(s). This finishes the proof. The following result is from [10]. (cid:3) Lemma 3.2. Let I be an open interval contained in (V + convex on I. t )c, then ρ(I) = 0 and g is strictly Proposition 3.3. Let µ be a compactly supported probability measure on [0, ∞) and 1 < α < β. Write κµ(z) = exp[u(z)], where u is an analytic function given by (2.3). (1) Then there exists a > 0 such that [0, a)∩supp(ρ) = ∅. Moreover, the set V + t bounded away from zero and increasing for all t ∈ (α, β). 5 is uniformly (2) Given any b > 0, the sets V + t ∩ [0, b] are Hausdorff continuous with respect to t. Proof. By the definition of η-transform, we have ηµ(0) = 0 and η′ identity µ(0) = m1(µ) < ∞. The z 1 − ηµ(z) = Gµ(cid:18) 1 z(cid:19) = Z ∞ 0 1 1/z − x dµ(x), implies that ηµ is real on some interval [0, a) under the assumption that µ is compactly supported. It follows that u is also real on [0, a) and hence ρ([0, a)) = 0 by Stieltjes inversion formula. Hence, the function g is finite in a neighborhood of zero and g(r) = is bounded away from 0. The definition of 0 V + t r(s2+1) (r−s)2 dρ(s) → 0 as r → 0. We see that V + immediately implies that V + R ∞ t1 ⊂ V + Assume now tn → t ∈ (α, β) and (V + t t ∩ [0, b]) is tn ∩ [0, b])\Bǫ(V + a ǫ-neighborhood of the set V + t ). We may assume that rn → r by passing to a subsequence if necessary. Lemma 3.3 implies that supp(ρ) ⊂ Bǫ(V + t ∩ [0, b]. We then have a series rn ∈ (V + t ) and hence we can take the limit t2 if t1 < t2. This proves the first assertion. t ∩ [0, b]), where Bǫ(V + tn ∩ [0, b]) 6⊂ Bǫ(V + g(r) = lim n→∞ g(rn) = lim n→∞ 1 tn − 1 . On the other hand, g ≤ 1 (V + t )c. This contradiction proves the second claim. t−1 and g is strictly convex on any open interval contained in (cid:3) Proposition 3.4. Given b > 0, the graphs {reiAt(r) : r ∈ V + Hausdorff metric for t ∈ (1, ∞). t ∩ (0, b)} are continuous in the Proof. For 0 < c < π, we define V + prove that there exists δ > 0 such that t,c = {r ∈ V + t : At(r) ≥ c}. Given ǫ > 0, we will start to (3.1) At(r) < As(r) ≤ (1 + ǫ)At(r) for all r ∈ V + t,c ∩ (0, b) if 0 < s − t < δ. The first inequality follows from the fact that the function g(r, θ) is a decreasing function of θ. To prove the second inequality by contradiction, we assume that there exists a series tn > t, tn → t and rn, r ∈ V + t,c ∩ (0, b) such that rn → r and Atn (rn) > (1 + ǫ)At(rn). We then have g(rn, (1 + ǫ)At(rn)) ≥ g(rn, Atn(rn)) = 1 tn − 1 . As At(rn) ≥ c, we can take the limit and obtain g(r, (1 + ǫ)At(r)) ≥ 1 t − 1 , due to the fact that the integrand in (2.4) is bounded away from zero and At is continuous. On the other hand, we have g(r, At(r)) = 1 t−1 and g(r, θ) is a strictly decreasing function of θ. This contradiction yields the second inequality in (3.1). 6 Given ǫ > 0, using the similar argument as above, we can prove that there exists δ > 0 such that (3.2) (1 − ǫ)At(r) ≤ As(r) < At(r) for all r ∈ V + t,c ∩ (0, b) if −δ < s − t < 0. We claim that sup{As(r) : r ∈ (0, b)\V + t,c} ≤ 2c if s − t is small enough. Assume that is not the case, then there exists a series tn → t and rn → r, where rn /∈ V + t,c ∩ (0, b), such that Atn(rn) > 2c. We have g(rn, 2c) ≥ g(rn, Atn(rn)) = 1 tn − 1 . Taking the limit, we have g(r, 2c) ≥ 1/(t − 1), which implies that At(r) ≥ 2c. As the cluster set of rn /∈ V + t,c, it must satisfy At(r) ≤ c. This contradiction proves our claim. t,c or an end point of V + t,c, r is either not in V + The desired assertion follows by above results and applying Proposition 3.4. (cid:3) Proposition 3.5. Let µ be a compactly supported probability measure on [0, ∞). If µ({0}) > 1, then µ⊠t({0}) = µ({0}) for all t > 1. If 0 ∈ supp((µ⊠s)ac) for some s > 1, then 0 ∈ supp((µ⊠t)ac) for all t > 1. Proof. It follows from [6] that rGµ(r) → µ({0}) as r → 0− on the negative half line. Hence, µ({0}) = 1 + limr→−∞ ψµ(r). For any t > 0, as limr→−∞ ωt(r) = −∞, we have µ⊠t({0}) = 1 + lim r→−∞ ψµ(ωt(r)) = 1 + lim r→−∞ ψµ(r) = µ({0}). Assuming that 0 ∈ supp((µ⊠s)ac) for some s > 1, we claim that ρ is not compactly supported. Indeed, if ρ is compactly supported, then limr→∞ g(r) = 0 and hence the set As is compact and the set supp((µ⊠s)ac) is bounded away from zero by Theorem 2.2, which contradicts to the assumption. The fact that the measure ρ is not compactly supported in turn implies that Vt is not compact and 0 ∈ supp((µ⊠t)ac) for all t > 1 by according to Theorem 2.2. (cid:3) We are now in a position to prove the main result of this section. Proof of Theorem 1.1. We first focus on the absolutely continuous part. Given t0 > 1, δ ∈ (0, t0 − 1) and choose a such that V + t0+δ ⊂ [a, ∞). For s ∈ (t0 − δ, t0 + δ), we will prove that (cid:12)(cid:12)(cid:12)(cid:12) ht0 (x) hs(y) (cid:12)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Φt0(xeiAt0 (x)) Φs(yeiAs(y)) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = O(cid:18) x y(cid:19) , (3.3) for x ∈ V + t0 and y ∈ V + s ∩ V + t0 . We fix x0 ∈ V + t0 and write Φt0(x0eiAt(x0)) Φs(yeiAs(y)) = Φt0(yeiAs(y)) Φs(yeiAs(y)) Φt0(x0eiAt(x0)) Φt0(yeiAs(y)) . We observe that the function (s, x) → As(x) is continuous in a neighborhood of (t0, x0) such that x0 ∈ V + t0 by the definition (2.4). When δ is sufficiently small, we thus have 7 = exp[(t0 − s)u(yeiAs(y))] ≤ exp[(t0 − s)C], where C is a constant depending on δ and y. To estimate the second factor, denote z1 = x0eiAt(x0) and z2 = yeiAs(y). We have = (cid:18) x0 y (cid:19) · exp[(t0 − 1)(u(z1) − u(z2))] Φt0(z1) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Φt0(yeiAs(y)) Φs(yeiAs(y)) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) Φt0(z2)(cid:12)(cid:12)(cid:12)(cid:12) and u′(z) = −Z ∞ 0 1 + s2 (z − s)2 dσ(s). For any z = reiθ ∈ Ωs ∩ Ωt0, we have that r sin(θ) θ Z ∞ 0 s2 + 1 z − s2 dσ(s) < 1 t0 − δ − 1 , which yields that u′(z) ≤ Z ∞ 0 1 + s2 z − s2 dz 1 θ · ≤ r sin(θ) t0 − δ − 1 . We now choose a curve γ ⊂ Ωs ∩ Ωt0 connecting z1 and z2, and obtain u(z1) − u(z2) = (cid:12)(cid:12)(cid:12)(cid:12) Zγ u′(z)dz(cid:12)(cid:12)(cid:12)(cid:12) ≤ 1 t0 − δ − 1 Zγ θ r sin(θ) dz . The curve γ can be chosen such that θ/(r sin(θ)) is bounded along γ when δ is chosen to be small thanks to Proposition 3.4. The length of γ can be arbitrarily small if x0/y is. We then deduce (3.3). Moreover, this estimation is uniform over compact subsets of V + t0 ∩ [a, ∞]. Hence, for any point ht0(x) = Φt0(xeiAt0 (x)) with x ∈ V + there exists y ∈ V + and choose b such that ǫ > max{1/hs(x) : x ∈ V + t0 > 1, we thus have t0 ∩ [a, b] and δ small enough, s such that ht0 (x)/hs(y) is sufficiently small. Let ǫ > 0 be a small number t0+δ ∩ [a, b], s − t0 < δ}. Given ǫ > 0 and and supp((µ⊠t0)ac) ∩ [ǫ, ∞) ⊂{1/x : x = ht0 (r), r ∈ V + t0 ∩ [a, b]} ⊂ Bǫ(supp(µ⊠s)ac), supp((µ⊠s)ac) ∩ [ǫ, ∞) ⊂{1/x : x = hs(r), r ∈ V + s ∩ [a, b]} ⊂ Bǫ(supp(µ⊠t0)ac), if δ is sufficiently small. To discuss the behavior of supp((µ⊠t)ac) close to zero, we have two cases. supp((µ⊠t)ac), then 0 ∈ supp((µ⊠s)ac) by Proposition 3.5. If 0 ∈ If 0 /∈ supp((µ⊠t)ac), then ρ 8 is bounded away from ∞ and thus supp((µ⊠t)ac) is bounded away zero. above inclusions imply the sets {supp(µ⊠t)ac} is Hausdorff continuous. In both cases, Finally, atoms of µ⊠t change continuously as time evloves by Theorem 2.1 and Proposition (cid:3) 3.5. This completes the proof. 4. Estimation of norm of free multiplicative convolution semigroups We give an estimation of the size of the support of µ⊠t for a compactly supported prob- ability measure on [0, ∞). Proof of Theorem 1.2. Let a > 0 be such that supp(ρ) ⊂ (a, ∞). As limr→0 g(r) = 0, the set V + t } > 0. By Theorem 1.2, we t have is bounded away from zero. Let αt = min{r : r ∈ V + By the choice of αt, we have At(αt) = 0, µ⊠t = 1 . ht(αt) and g(αt) = Z ∞ 0 αt(s2 + 1) (αt − s)2 dρ(s) = 1 t − 1 ht(αt) = Φt(αt) = αt exp[(t − 1)u(αt)]. By (2.3) and the assumption m1(µ) = 1, we have u(αt) = a +Z ∞ 0 (cid:18) 1 s 0 1 + αts αt − s dρ(s) + 1 + αts αt − s (cid:19) dρ(s) αt(1 + s2) (αt − s)2 αt − s s dρ(s) = Z ∞ = Z ∞ = −Z ∞ 0 0 1 = − t − 1 αt(1 + s2) (αt − s)2 dρ(s) + αtZ ∞ + αtZ ∞ αt(1 + s2) (αt − s)2 1 s 0 0 dρ(s). αt(1 + s2) (αt − s)2 1 s dρ(s) It is clear that limt→∞ αt = 0 and hence 1 = lim t→∞ g(αt)(t − 1) = lim t→∞ (tαt) ·(cid:18)Z ∞ 0 s2 + 1 s2 dρ(s)(cid:19) , which yields that limt→∞(tαt) = 1/V thanks to Lemma 3.1. We then obtain lim t→∞ u(αt)(t − 1) = −1 + lim t→∞ [(t − 1)αt] · αtZ ∞ 0 1 + s2 (αt − s)2 1 s dρ(s) as supp(ρ) is bounded below from zero. = −1 Finally, we have µ⊠t t lim t→∞ = lim t→∞ 1 tht(αt) = lim t→∞(cid:18) 1 = e ·Z ∞ 0 t · αt s2 + 1 s2 · exp[−(t − 1)u(αt)](cid:19) dρ(s) This proves the desired result. = eV. 9 (cid:3) Proposition 4.1. Let µ be a probability measure supported on [c, d] not being a Dirac measure with c, d > 0. There exists T such that the sets V + t and supp(µ⊠t) have only one connected component for all t > T . Proof. The measure ρ determined by (2.3) is compactly supported since supp(µ) ⊂ [c, d]. Hence limr→0 g(r) = limr→∞ g(r) = 0 and V + t is bounded away from zero and ∞. t0 , ρ(I) = 0 by Lemma 3.2. Hence the function g(r) = R ∞ Let [a, b] ⊂ (0, ∞) be a finite interval such that supp(ρ) ⊂ [a, b]. Fix t0 > 1. For any open r(s2+1) (r−s)2 dρ(s) t0 . Therefore, we can choose T large enough, so T . The set V + T has only one connected component as supp(ρ) ⊂ [a, b]. It then t and hence supp(µ⊠t) have only one connected component for all t > T by (cid:3) interval I ⊂ [a, b]\V + has a positive local minimum at [a, b]\V + that [a, b] ⊂ V + follows that V + Theorem 2.2. 0 Proposition 4.2. Let µ be a probability measure on [c, d] not being a Dirac measure for some c, d > 0. Define and Then at = sup{a : a < x for all x ∈ supp(µ⊠t)}, bt = inf{b : b > x for all x ∈ supp(µ⊠t)}. lim t→∞ (at)1/t = (cid:18)Z ∞ 0 x−1dµ(x)(cid:19)−1 , and lim t→∞ (bt)1/t = Z ∞ 0 x dµ(x). Proof. The set V + t βt = max{r : r ∈ V + By adapting the argument in the proof of Theorem 1.2, we have is bounded away from zero and ∞. Let αt = min{r : r ∈ V + t } > 0 and t } < ∞, we have limt→∞ αt = 0 and limt→∞ βt = 0 by Theorem 2.2. lim t→∞ tαt = m1(µ)/V and lim t→∞ (t/βt)Z ∞ 0 (s2 + 1) dρ(s) = 1. On the other hand, we have and lim t→∞ exp[u(αt)] = lim z→0 exp[u(z)] = lim z→0 z/ηµ(z) = 1/m1(µ) lim t→∞ exp[u(βt)] = lim z→∞ exp[u(z)] = lim z→∞ z/ηµ(z) = Z ∞ 0 x−1dµ(x). 10 Therefore, we have lim t→∞ (at)1/t = lim t→∞ (ht(βt))−1/t {βt exp[(t − 1)u(βt)]}−1/t = lim t→∞ = lim t→∞ exp[−u(βt)] and = (cid:18)Z ∞ 0 x−1dµ(x)(cid:19)−1 ; lim t→∞ (bt)1/t = lim t→∞ (ht(αt))−1/t = lim t→∞ = lim t→∞ {αt exp[(t − 1)u(αt)]}−1/t exp[−u(αt)] = m1(µ). This finishes the proof. (cid:3) Remark 1. Proposition 4.2 should be compared with the main result in [9], where weak limit of rescaled discrete semigroups were studied. Our method is not suitable to study weak limit, but works to estimate the asymptotic bound of free multiplicative convolution semigroups. The interested reader can easily generalize results in [9] to continuous free multiplicative convolution semigroups using their method. Remark 2. Free multiplicative convolution semigroup can also be defined for measures on the unit circle [4]. Let µ be a probability measure on the unit circle not being a Dirac measure, it is known in [17, Proposition 3.26] that µ⊠t converges to the uniform measure on the unit circle as t → ∞. 5. Acknowledgements The second author wants to thank Hari Bercovici, Alexandru Nica, Jiun-Chau Wang and John Williams for useful discussions. This work was partially supported by NSFC no. 11501423, 11431011. References [1] Michael Anshelevich and Octavio Arizmendi, The exponential map in non-commutative probability, Int. Math. Res. Notices 17 (2017), 5302 -- 5342. [2] Octavio Arizmendi and Carlos Vargas, Products of free random variables and k-divisible non-crossing partitions, Electron. Commun. Probab. 17 (2012), no. 11, 13. MR 2892410 [3] Serban Belinschi, Benoıt Collins, and Ion Nechita, Eigenvectors and eigenvalues in a random subspace of a tensor product, Invent. Math. 190 (2012), no. 3, 647 -- 697. MR 2995183 [4] Serban T. Belinschi and Hari Bercovici, Partially defined semigroups relative to multiplicative free con- volution, Int. Math. Res. Not. (2005), no. 2, 65 -- 101. MR 2128863 [5] , A new approach to subordination results in free probability, J. Anal. Math. 101 (2007), 357 -- 365. MR 2346550 [6] Hari Bercovici and Dan Voiculescu, Regularity questions for free convolution, Nonselfadjoint operator algebras, operator theory, and related topics, Oper. Theory Adv. Appl., vol. 104, Birkhauser, Basel, 1998, pp. 37 -- 47. MR 1639647 11 [7] Philippe Biane, Processes with free increments, Math. Z. 227 (1998), no. 1, 143 -- 174. MR 1605393 [8] Benoıt Collins, Motohisa Fukuda, and Ping Zhong, Estimates for compression norms and additivity violation in quantum information, Internat. J. Math. 26 (2015), no. 1, 1550002, 20. MR 3313648 [9] Uffe Haagerup and Soren Moller, The law of large numbers for the free multiplicative convolution, Operator algebra and dynamics, Springer Proc. Math. Stat., vol. 58, Springer, Heidelberg, 2013, pp. 157 -- 186. MR 3142036 [10] Hao-Wei Huang and Ping Zhong, On the supports of measures in free multiplicative convolution semi- groups, Math. Z. 278 (2014), no. 1-2, 321 -- 345. MR 3267581 [11] Vladislav Kargin, On asymptotic growth of the support of free multiplicative convolutions, Electron. Commun. Probab. 13 (2008), 415 -- 421. MR 2424965 [12] Alexandru Nica and Roland Speicher, On the multiplication of free N -tuples of noncommutative random variables, Amer. J. Math. 118 (1996), no. 4, 799 -- 837. MR 1400060 [13] Dan Voiculescu, The analogues of entropy and of Fisher's information measure in free probability theory. I, Comm. Math. Phys. 155 (1993), no. 1, 71 -- 92. MR 1228526 [14] Dan Voiculescu, Ken Dykema, and Alexandru Nica, Free random variables, CRM Monograph Series, vol. 1, American Mathematical Society, Providence, RI, 1992, A noncommutative probability approach to free products with applications to random matrices, operator algebras and harmonic analysis on free groups. MR 1217253 (94c:46133) [15] Dan Voiculescu, Nicolai Stammeier, and Moritz Weber, Free probability and operator algebras, Munster Lectures in Mathematics, European Mathematical Society, 2016. [16] John Williams, On the hausdorff continuity of free L´evy processes and free convolution semigroups, J. Math. Anal. Appl. (online 2016). [17] Ping Zhong, Free Brownian motion and free convolution semigroups: multiplicative case, Pacific J. Math. 269 (2014), no. 1, 219 -- 256. MR 3233917 [18] , On the free convolution with a free multiplicative analogue of the normal distribution, J. Theoret. Probab. 28 (2015), no. 4, 1354 -- 1379. MR 3422934 Xiaoxue Deng: School of Mathematics and Statistics, Wuhan University, No. 299 BaYi Road, Wuhan, Hubei 430074, China E-mail address: [email protected] Ping Zhong: School of Mathematics and Statistics, Wuhan University, No. 299 BaYi Road, Wuhan, Hubei 430074, China; and Department of Pure Mathematics, University of Waterloo, Waterloo, ON, Canada E-mail address: [email protected]
1806.06691
1
1806
2018-06-18T13:54:25
Improved Ingham-type result on $\mathbb R^d$ and on connected, simply connected nilpotent Lie Groups
[ "math.FA" ]
In \cite{BRS} we have characterized the existance of a non zero function vanishing on an open set in terms of the decay of it's Fourier transform on the $d$-dimensional Euclidean space, the $d$-dimensional torus and on connected, simply connected two step nilpotent Lie groups. In this paper we improved these results on $\mathbb R^d$ and prove analogus results on connected, simply connected nilpotent Lie groups.
math.FA
math
IMPROVED INGHAM-TYPE RESULT ON Rd AND ON CONNECTED, SIMPLY CONNECTED NILPOTENT LIE GROUPS MITHUN BHOWMIK Abstract. In [3] we have characterized the existance of a non zero function vanishing on an open set in terms of the decay of it's Fourier transform on the d-dimensional Euclidean space, the d-dimensional torus and on connected, simply connected two step nilpotent Lie groups. In this paper we improved these results on Rd and prove analogus results on connected, simply connected nilpotent Lie groups. 1. Introduction It is a well known fact in harmonic analysis that if the Fourier transform of an integrable function on the real line is very rapidly decreasing then the function can not vanish on a nonempty open set unless it vanishes identically. For instance if, let f ∈ L1(R) and a > 0 be such that its Fourier transform satisfies the estimate bf (ξ) ≤ Ce−aξ, for all ξ ∈ R. If f vanishes on a nonempty open set then f is identically zero. This initial observation motivates one to endeavour for a more optimal decay of the Fourier transform bf for such a conclusion to hold. For instance we may ask: if bf decays faster than 1/(1 + · )n for all n ∈ N but slower than the function e−a·, can f vanish on a nonempty open set without being identically zero? A more precise question could be: is there a nonzero integrable function f on R vanishing on a nonempty open set such that its Fourier transform satisfies the following estimate (1.1) log ξ , for large ξ? bf (ξ) ≤ Ce− ξ The answer to the above question is in the negative and follows from a classical result due to Ingham [7]. Analogous results were also obtained by Paley-Wiener, Levinson and Hirschman [11, 10, 9, 6]. In [3], we have obtained analogues results on d-dimensioan Euclidean spaces, d-dimensional torus and on connected, simply connected two step nilpotent Lie groups. In the context of Euclidean space we have the following theorem [3]. For f ∈ L1(Rd), we shall define its Fourier transform bf by f (x) e−2πix·ξ dx, for ξ ∈ Rd. For 1 < p ≤ 2, the definition of Fourier transform of f ∈ Lp(Rd) is extended in the usual way. bf (ξ) =ZRd 2010 Mathematics Subject Classification. Primary 22E25; Secondary 22E30, 43A80. Key words and phrases. nilpotent Lie group. This work was supported by Indian Statistical Institute, India (Research fellowship). 1 2 MITHUN BHOWMIK Theorem 1.1. Let f ∈ Lp(Rd), for some p with p ∈ [1, 2] and ψ : [0, ∞) → [0, ∞) be a locally integrable function. Suppose that the Fourier transform bf of f satisfies the following estimate (1.2) for almost every ξ ∈ Rd, where P (ξ) is a polynomial and we define bf (ξ) ≤ CP (ξ)e−ψ(kξk), I =Z ∞ ψ(t) t2 dt. 1 (a) Let ψ(t) = θ(t)t, for t ∈ [0, ∞), where θ : [0, ∞) → [0, ∞) is a decreasing function which decreases to zero as t goes to infinity. If f vanishes on any nonempty open set in Rd and I is infinite then f is zero on Rd. Conversely, if I is finite, then given any positive real number l there exists a nonzero even function f ∈ Cc(Rd) supported in ball B(0, l) of radius l, centered at zero satisfying (1.2). (b) Let ψ be an increasing function on [0, ∞). If f vanishes on any nonempty open set in Rd and I is infinite then f is zero on Rd. Conversely, if I is finite, given any positive real number l there exists a nonzero f ∈ Cc(Rd) supported in B(0, l) satisfying (1.2). (c) If supp f ⊆ {x ∈ Rd x · η ≤ s}, and I is infinite then f is zero on Rd. Conversely, if I is finite then there exists a nonzero f ∈ L2(Rd) vanishing on the set {x ∈ Rd : x · η ≥ x0}, for some x0 ∈ R and η ∈ Sd−1 satisfying (1.2). In [3] we have shown that the original statements of the results of Ingham, Levinson and Paley-Wiener on R are follows form the theorem above. We shall show in this paper that it is possible to prove analogues of Theorem 1.1 under the following variant of (1.2) (1.3) ZRd bf (ξ)qeqψ(kξk) (1 + kξk)N dξ < ∞, where q ∈ [1, ∞] (with obvious modification for q = ∞) and N is a positive real number. Our next main result in this paper is an analogue of Theorem 1.1, c) on the connected, simply connected nilpotent Lie group G. We shall use the following standard notation in this paper: supp f denotes the support of the function f and C denotes a constant whose value may vary. For a finite set A we shall use the symbol #A to denote the number of elements in A. B(0, r) denotes the open ball of radius r centered at 0 in Rd and ¯B(0, r) denotes its closure. For x, y ∈ Rd, we shall use kxk to denote the norm of the vector x and x · y to denote the Euclidean inner product of the vectors x and y. A function ψ : Rd → [0, ∞) is said to be radially increasing (or radially decreasing) if ψ is radial and satisfies the condition ψ(x) ≥ ψ(y) (or ψ(x) ≤ ψ(y)) whenever x ≥ y. We shall often consider a radial function on Rd as an even function on R or equivalently, as a function on [0, ∞). 2. Results of Ingham, Levinson and Paley-Wiener on Rd In this section our aim is to improve Theorem 1.1 in the following way. Instead of the pointwide decay (1.2) on the Fourier transform side we will consider the weighted Lq estimate for 1 ≤ q ≤ ∞ of the form (1.3). Using convolution techniques we will be able to show that it is enough to prove the result for q = 1 and N = 0 and rest of the proof goes like the proof of Theorem 1.1. First we state and prove an analogue of Theorem 1.1 (a) on Rd. IMPROVED INGHAM-TYPE RESULT ON RD AND ON NILPOTENT LIE GROUPS 3 Theorem 2.1. Let θ : Rd → [0, ∞) be a radially decreasing measurable function with limkξk→∞ θ(ξ) = 0 and I =Zkξk≥1 θ(ξ) kξkd dξ. (a) Let f ∈ Lp(Rd), p ∈ [1, 2], be such that ZRd bf (ξ)q eqθ(ξ)kξk (1 + kξk)N for some q ∈ [1, ∞) and some N ≥ 0 or dξ < ∞, (2.1) (2.2) (2.3) bf (ξ) ≤ C P (ξ) e−θ(ξ)kξk, for almost every ξ ∈ Rd, where P is a polynomial. If f vanishes on a nonempty open subset in Rd and I is infinite, then f is zero almost everywhere on Rd. (b) If I is finite then there exists a nontrivial f ∈ C∞ c (Rd) satisfying the estimate (2.2) (or (2.3) if q = ∞). Proof. We shall first prove the assertion (b). For q = ∞ it follows from Theorem 1.1, (a). For q ∈ [1, ∞) we choose φ1 ∈ C∞ c (Rd) with supp φ1 ⊆ B(0, l/2) and consider the function f = f0 ∗ φ1. Clearly, support of the function f is contained in B(0, l) and This, in particular, proves (b). It now remains to prove (a). ZRd bf (ξ)q eqθ(ξ)kξk (1 + kξk)N dξ ≤ CZRd cφ1(ξ)q dξ < ∞. For part (a), we first show that it suffices to prove the case q = 1, N = 0. Suppose f ∈ Lp(Rd) vanishes on an open set U ⊆ Rd and satisfies (2.2), for some q > 1 and N ∈ N. Since the condition (2.2) is invariant under translation of f we can assume that f vanishes on B(0, l), for some positive l. If we choose φ ∈ C∞ c (Rd) supported in B(0, l/2) then f ∗ φ vanishes on the ball B(0, l/2). In fact, if kxk < l/2 then ZRd f (x − y)φ(y) dy =ZB(0, l 2 ) f (x − y)φ(y) dy = 0, as kx − yk ≤ kxk + kyk < l. By using Holder's inequality we get ZRd ≤ ZRd < ∞, \(f ∗ φ)(ξ) eθ(ξ)kξk dξ (1 + kξk)N ! 1 bf (ξ)q eqθ(ξ)kξk q k(1 + kξk)Nbφ(ξ)kLq′ (Rd) as N/q is smaller than N . Here q′ satisfies the relation 1 q + 1 q′ = 1. 4 MITHUN BHOWMIK Hence, by the case q = 1, N = 0 it follows that f ∗ φ vanishes identically. As φ ∈ C∞ c (Rd) we have that bφ is nonzero almost everywhere. This implies that bf vanishes almost everywhere and so does f . The same technique can be applied to reduce the case q = 1 and N ∈ N to the case q = 1 and N = 0 by using Holder's inequality. For the case q = ∞, we get from (2.3) that ZRd [f ∗ φ(ξ) eθ(ξ)kξk dξ ≤ZRd P (ξ) bφ(ξ) dξ < ∞. So, without loss of generality, we assume that f ∈ Lp(Rd) such that bf satisfies the condition (2.4) ZRd bf (ξ) eθ(ξ)kξk dξ = C′ < ∞. Rest of the proof goes similarly as the proof of Theorem 2.2 in [3]. (cid:3) As in Theorem 2.1 we will have the following version of Theorem 1.1, (b) on Rd. Theorem 2.2. Let ψ : Rd → [0, ∞) be a radially increasing function and I =Z ∞ 1 ψ(t) t2 dt. (a) Let f ∈ Lp(Rd), p ∈ [1, 2], be such that supp f ⊆ {x ∈ Rd x · η ≤ s}, for some η ∈ Sd−1 and s ∈ R and bf satisfies the estimate dξ < ∞. ZRd (1 + kξk)N bf (ξ)q eqψ(ξ·η) bf (ξ) ≤ C (1 + kξk)N e−ψ(ξ·η), If the integral I is infinite then f is the zero function. for some q ∈ [1, ∞) and some N ≥ 0, or for almost every ξ ∈ Rd. (2.7) (2.8) ψ(ξ) kξkd+1 dξ. I =Zkξk≥1 ZRd bf (ξ)q eqψ(ξ) (1 + kξk)N dξ < ∞, (a) Let f ∈ Lp(Rd), p ∈ [1, 2], be a function satisfying the estimate (2.5) (2.6) for some q ∈ [1, ∞) and some N ≥ 0 or bf (ξ) ≤ C P (ξ) e−ψ(ξ), for almost every ξ ∈ Rd, where P is a polynomial. If f vanishes on a nonempty open subset in Rd and I is infinite, then f is zero almost everywhere on Rd. (b) If I is finite then there exists a nontrivial f ∈ C∞ c (Rd) satisfying the estimate (2.5) (or (2.6) if q = ∞). In the following, we shall improved Theorem 1.1, c) which will be used in the next section to prove an analogous result for connected simply connected nilpotent Lie groups (see Theorem 3.1). Theorem 2.3. Let ψ : [0, ∞) → [0, ∞) be a locally integrable function and IMPROVED INGHAM-TYPE RESULT ON RD AND ON NILPOTENT LIE GROUPS 5 (b) If ψ is non-decreasing and I is finite then there exists a nontrivial f ∈ Cc(Rd) satisfying the estimate (2.7), for some q ∈ [1, ∞) and all η ∈ Sd−1 or (2.8), for q = ∞ and all η ∈ Sd−1. Proof. We shall first prove a) for the case p = 2. Then for all p ∈ [1, 2) the result will follow by reducing it to the case p = 2 as was done in the proof of Theorem 1.1, (c). As in the proof of Theorem 2.1 it suffices to prove the case q = 1 and N = 0. Now, by using translation and rotation of the function f , we can assume without loss of generality that η = e1 = (1, 0, · · · , 0) and s = 0. Then, by writing ξ = (ξ1, ξ2, · · · , ξd) the hypothesis (2.7) becomes (2.9) For y ∈ Rd−1, we now define gy by ZRd bf (ξ) eψ(ξ1) dξ < ∞. gy(x) = Fd−1f (x, y), for almost every x ∈ R. It then follows that for almost every y ∈ Rd−1, gy ∈ L2(R) with supp gy ⊆ {x ∈ R x ≤ 0}, and by (2.9) (2.10) As y varies over a set of full (d − 1) dimensional Lebesgue measure, we just need to prove that gy is the zero function. By the Paly-Wiener theorem (Theorem 2.7, [3]) it suffices to show that ZR 1 + t2 1 + t2 dt = ∞. bgy(t) eψ(t) dt < ∞. ZR log(bgy(t)) ZR log(bgy(t)eψ(t)) dt = ZR ≥ ZR ZR log(bgy(t)) 1 + t2 dt dt < ∞ dt 1 + t2 log(bgy(t)eψ(t)) − ψ(t) 1 + t2 dt −ZR log(bgy(t)eψ(t)) 1 + t2 ψ(t) dt. If (2.11) then ZR log(bgy(t)) 1 + t2 As I is infinite, it follows from (2.11) that is divergent. Hence, by the Paly-Wiener theorem (Theorem 2.7, [3]) it follows that gy is the zero function. Now, suppose (2.12) ZR log(bgy(t)eψ(t)) 1 + t2 dt = ∞. 6 MITHUN BHOWMIK For a measurable function F on Rd, we define and hence log+ F (x) = max{log F (x), 0} log− F (x) = − min{log F (x), 0}, log F (x) = log+ F (x) + log− F (x). As log+ F (x) is always smaller than F (x) we get that 1 + t2 1 + t2 dt = ∞. ZR bgy(t)eψ(t) 1 + t2 log+(cid:0)bgy(t)eψ(t)(cid:1) dt ≤ZR ZR log−(bgy(t)eψ(t)) log−(cid:0)bgy(t)eψ(t)(cid:1) dt = Z(cid:8)t cgy(t)eψ(t)≤1(cid:9) ≤ ZR log− bgy(t) bgy(t) ≤ bgy(t)eψ(t) ≤ 1, log− (bgy(t)) ≥ log−(cid:16)bgy(t)eψ(t)(cid:17) . ZR ZR log− (bgy(t)) log (bgy(t)) dt dt dt, 1 + t2 1 + t2 1 + t2 as on the set {t bgy(t)eψ(t) ≤ 1} we have and hence Therefore the integral is divergent. Hence, the integral by (2.10). From (2.12) we now conclude that But ZR 1 + t2 dt < ∞, log−(cid:0)bgy(t)eψ(t)(cid:1) 1 + t2 dt is divergent. By the Paly-Wiener theorem (Theorem 2.7, [3]) it now follows that gy is the zero function. This completes the proof of part a). To prove Part b) we observe that since ψ is non decreasing ψ(ξ · η) ≤ ψ(kξk), for ξ ∈ Rd. Therefore the proof follows from Theorem 2.2, (b). (cid:3) Remark 2.4. (1) It is clear from the statement of Theorem 2.3 that if f is compactly supported con- tinuous function then η ∈ Sd−1 can be taken to be arbitrary. (2) If ψ is assumed to be a radial increasing function then Theorem 1.1, (c) follows from the case p = ∞ of Theorem 2.3 and also from Theorem 2.2. In this section, our aim is to prove an analogue of Theorem 2.3 in the context of connected, simply connected nilpotent Lie groups. 3. Nilpotent Lie Groups IMPROVED INGHAM-TYPE RESULT ON RD AND ON NILPOTENT LIE GROUPS 7 3.1. Preliminaries on Nilpotent Lie Groups. In this section, we shall discuss the preliminaries and notation related to connected, simply connected nilpotent Lie groups. These are standard and can be found, for instance, in [1, 2, 5]. Let G be a connected, simply connected nilpotent Lie group with Lie algebra g and let g ∗ be the vector space of real-valued linear functionals on g. In this case, the exponential map exp : g → G becomes an analytic diffeomorphism, which enables us to identify G with g ∼= Rd, where d = dim g. Let (3.1) {0} = g0 ⊂ g1 ⊂ · · · ⊂ gd = g be a Jordan-Holder series of ideals of the nilpotent Lie algebra g such that dim gj = j, for j = 0, · · · , d and ad(X)gj ⊆ gj−1, for j = 1, · · · , d and for all X ∈ g ([5], Theorem 1.1.9). We choose from this sequence a Jordan-Holder basis (3.2) for j = 1, · · · , d. In particular, RX1 is contained in the center of g. Let {X ∗ dual to {X1, · · · , Xd}. We define coordinates on G by {X1, · · · , Xd}, where Xj ∈ gj\gj−1, 1 , · · · , X ∗ d } be the basis of g ∗ (x1, · · · , xd) ∼= exp(x1X1 + · · · + xdXd). By the Baker-Campbell-Hausdorff formula exp(X) exp(Y ) = exp(Z), where (3.3) Z = X + Y + 1 2 [X, Y ] + 1 12 [X, [X, Y ]] − 1 12 [Y, [X, Y ]] + · · · ([5], P.12), which is a finite sum as G is a nilpotent Lie group. We introduce a 'norm function' on G defined by The composite map kxk =(cid:0)x2 d(cid:1) 1 1 + · · · + x2 2 , x = exp(x1X1 + · · · + xdXd) ∈ G, xj ∈ R. Rd −→ g −→ G, (x1, · · · , xd) 7→ dXj=1 xjXj 7→ exp( dXj=1 xjXj), is a diffeomorphism and maps the Lebesgue measure on Rd to a Haar measure on G. We shall also identify g ∗ with Rd via the map ν = (ν1, · · · , νd) → νjX ∗ j , νj ∈ R, dXj=1 where {X ∗ 1 , · · · , X ∗ d } is the dual basis of g ∗. On g ∗ we define the Euclidean norm relative to this basis by k dXj=1 νjX ∗ j k =(cid:0)ν2 d(cid:1) 1 1 + · · · + ν2 2 = kνk. For ν ∈ g ∗ we have a natural bilinear map This map is antisymmetric and the radical of this map is (X, Y ) 7→ Bν(X, Y ) = ν([X, Y ]). {X ∈ g Bν(X, Y ) = 0, for all Y ∈ g} = rν . 8 MITHUN BHOWMIK Clearly, Bν descends to nondegenerate, antisymmetric bilinear map Bν : g/rν × g/rν → R. Hence, rν is of even codimension in g ([5], Lemma 1.3.2). Since the coadjoint orbit Oν of ν is diffeomorphic to G/Rν, it follows that Oν is even dimensional. An index j ∈ {1, 2, · · · , d} is called a jump index for ν ∈ g ∗ if rν + gj 6= rν + gj−1, where the Jordan-Holder sequence gj are as in (3.2) (see [1]). Let e(ν) be the set of jump indices for ∗. This set contains dim(Oν ) indices, which is an even number. Moreover, there are disjoint sets ν ∈ g ∗ of indices P and Q such that P ∪ Q = {1, · · · , d} and a G-invariant nonempty Zariski open set U ⊆ g such that e(ν) = P , for all ν ∈ U ([5], Theorem 3.1.2; [1]). The elements of U are called generic linear functionals. For ν ∈ U, we define Then the Pfaffian P f (ν) is given by Bν,P = ν([Xi, Xj])i,j∈P . (3.4) We define P f (ν)2 = det Bν,P , for all ν ∈ U. VP = span {X ∗ i : i ∈ P }, and VQ = span{X ∗ i : i ∈ Q}. Let dν be the Lebesgue measure on VQ normalized so that the unit cube spanned by {X ∗ i : i ∈ Q} has ∗ = VQ ⊕ VP and W = U ∩ VQ is a cross-section for the coadjoint orbits through points volume 1. Then g in U. For φ ∈ L1(G) we define its operator-valued Fourier transform by π(φ) =ZG φ(g)π(g−1) dg, for π ∈ bG, where dg denotes the Haar measure of G and bG is the unitary dual of G. If φ ∈ L1(G) ∩ L2(G) then π(φ) is a Hilbert-Schmidt operator and kπ(φ)kHS will denote the Hilbert-Schmidt norm of π(φ). If dν denotes the element of Lebesgue measure on W, then µ is the Plancherel measure for G, where dµ = P f (ν)dν. The Plancherel formula is thus given by kφk2 2 =ZG φ(g)2 dg =ZW kπν(φ)k2 HS dµ(ν), for all φ ∈ L1(G) ∩ L2(G). We shall now state and prove the following analogue of Theorem 1.1, (c) on a connected, simply connected nilpotent Lie group G. Theorem 3.1. Let ψ : [0, ∞) → [0, ∞) be a locally integrable function and I =Z ∞ 1 ψ(t) t2 dt. Suppose f ∈ Cc(G) is such that (3.5) ZW kπν(f )k2 HS e2ψ(ν1) P f (ν) dν < ∞, where ν1 = ν(X1) and X1 is given by (3.2). If I is infinite then f vanishes identically on G. Conversely, if ψ is nondecreasing and I is finite, then there exists a nontrivial f ∈ Cc(G) such that (3.5) holds. IMPROVED INGHAM-TYPE RESULT ON RD AND ON NILPOTENT LIE GROUPS 9 To prove the theorem we need some more notation. If r is the dimension of coadjoint orbits Oν , ν ∈ U, then Q has d − r elements and so VQ can be identified with Rd−r. In an abuse of notation we write VQ = RX ∗ 1 ⊕ Rd−r−1 and let p∗ : VQ → RX ∗ 1 , ν 7→ ν1X ∗ 1 denote the canonical projection. As W is a Zariski open set in VQ, p∗(W) = O is also a nonempty Zariski open subset of R. In particular, this is a subset in R with full Lebesgue measure. So, it will be convenient to write elements ν ∈ W as (ν1, ν′), where ν1 ∈ O and ν′ is in the set Wν1 defined by Wν1 = {ν′ ∈ Rn−d−1 : (ν1, ν′) ∈ W}. It turns out that Wν1 is also a Zariski open subset of Rn−d−1, for each fixed ν1 ∈ O. Given f ∈ Cc(G) and the Jordan-Holder basis {X1, · · · , Xd} given in (3.2) we define for each y ∈ Rd−1 and f ∗ y (t) = fy(−t). We now consider the function fy(t) = fexptX1 + g(t) =ZRd−1 dXj=2 yjXj , for t ∈ R, (fy ∗ f ∗ y )(t)dy, for t ∈ R. Lemma 3.2. ([2], Lemma 3.2) For all ν1 ∈ O, bg(ν1) =ZWν1 kπν(f )k2 HS P f (ν) dν′, wherebg is the one dimensional Fourier transform of g. We now present the proof of the theorem. Proof of Theorem 3.1. It follows from the definition of g that Since f ∈ Cc(G) it follows that g ∈ Cc(R). In particular, g ∈ L1(R). We now have the following lemma. ξ ∈ R, bfy(ξ)2dy, fy(ξ)2dy dξ = kf k2 2. and consequentlybg ≥ 0 with bg(ξ) =ZRd−1 ZRbg(ξ) =ZRd ZRbg(ν1) eψ(ν1) dν1 = ZOZWν1 = ZW kπν(f )k2 HS eψ(ν1) P f (ν) dν′ dν1 kπν(f )k2 HS eψ(ν1) P f (ν) dν. Therefore the function f is identically zero on G if and only if g is zero on R. We now complete the proof by showing that g vanishes identically. In order to do this we shall apply Theorem 2.3 to the function g. By Lemma 3.2, it follows that By the hypothesis (3.5) the right-hand side integral is finite. Therefore by Theorem 2.3 we conclude that g is identically zero on R and hence f vanishes identically. 10 MITHUN BHOWMIK Conversely, if ψ is non decreasing and the integral I is finite then by Theorem 1.1, (b), for a given δ positive there exists g ∈ Cc(R) with supp g ⊂ [−δ, δ] satisfying the estimate (3.6) y ∈ R. Let Z = exp(RX1) denote a central subgroup of G and we identify Z with R. Let h ∈ Cc(G) with supp h ⊂ V where V is a symmetric neighbourhood of the identity element in G. We now define a function f on G by bg(y) ≤ Ce−ψ(y), f (x) =ZZ g(t) h(t−1x) dt, x ∈ G. Clearly, f is continuous. Moreover, f ∈ Cc(G), which follows from the fact that as t ∈ Z. It can be shown ([8], P. 490) that kt−1xk ≥ kxk − ktk (3.7) kπν (f )k2 HS, for all ν ∈ W. Therefore, from (3.6) and (3.7) it follows that kπν(f )k2 HS e2ψ(ν1) P f (ν) dν HS = bg(ν1)2 kπν(h)k2 ZW = ZW ≤ CZW bg(ν1)2 kπν(h)k2 kπν(h)k2 = C khkL2(G) < ∞. HS P f (ν) dν HS e2ψ(ν1) P f (ν) dν This completes the proof. (cid:3) If the function ψ is assumed to be nondecreasing then we have the following corollary. Corollary 3.3. Let ψ : [0, ∞) → [0, ∞) be a nondecreasing function and I =Z ∞ 1 ψ(t) t2 dt. Suppose f ∈ Cc(G) is such that ZW kπν(f )k2 HS e2ψ(kνk) P f (ν) dν < ∞. If the integral I is infinite then f vanishes identically on G. Proof. Since ψ is nondecreasing ψ(kνk) ≥ ψ(ν1). Therefore, the result follows from Theorem 3.1. (cid:3) Acknowledgement. We would like to thank Swagato K. Ray and Suparna Sen for the many useful discussions during the course of this work. IMPROVED INGHAM-TYPE RESULT ON RD AND ON NILPOTENT LIE GROUPS 11 References [1] Baklouti, A; Ben Salah, Nour . On theorems of Beurling and Cowling-Price for certain nilpotent Lie groups, Bull. Sci. Math. 132 (2008), no. 6, 529-550. MR2445579 (2009f:22007) [2] Baklouti, A; Thangavelu, S. Variants of Miyachi's theorem for nilpotent Lie groups, J. Aust. Math. Soc. 88 (2010), no. 1, 1-17. MR2770922 (2012a:22012) [3] Bhowmik, M.; Ray, S. K.; Sen, S. Around theorems of Ingham-type regarding decay of Fourier transform on Rn, Tn and two step nilpotent Lie Groups, to appear in Bull. Sci. Math. [4] Bochner, S. Quasi-analytic functions, Laplace operator, positive kernels Ann. of Math. (2) 51 (1950), 68-91. MR0032708 (11,334g) [5] Corwin,L. J.; Greenleaf, F. P. Representations of nilpotent Lie groups and their applications Cambridge University Press, Cambridge, 1990. MR1070979 (92b:22007) [6] Hirschman, I. I. On the behaviour of Fourier transforms at infinity and on quasi-analytic classes of functions Amer. J. Math. 72 (1950), 200-213. MR0032816 (11,350f) [7] Ingham, A. E. A Note on Fourier Transforms J. London Math. Soc. 9 (1934) no. 1 , 29-32. MR1574706 [8] Kaniuth, E.; Kumar, A. Hardy's theorem for simply connected nilpotent Lie groups Math. Proc. Cambridge Philos. Soc. 131 (2001), no. 3, 487-494. MR1866390 (2002j:22007) [9] Levinson, N. Gap and Density Theorems American Mathematical Society Colloquium Publications, v. 26. American Mathematical Society, New York, 1940. MR0003208 (2,180d) [10] Paley, R. E. A. C.; Wiener, N. Fourier transforms in the complex domain (Reprint of the 1934 original) American Mathematical Society Colloquium Publications, 19. American Mathematical Society, Providence, RI, 1987. MR1451142 (98a:01023) [11] Paley, R. E. A. C.; Wiener, N. Notes on the theory and application of Fourier transforms. I, II. Trans. Amer. Math. Soc. 35 (1933), no. 2, 348355. MR1501688 Stat-Math Unit, Indian Statistical Institute, 203 B. T. Road, Kolkata - 700108, India. E-mail address: [email protected]
1607.02161
1
1607
2016-07-07T20:21:28
Unconditionally $p$-converging operators and Dunford-Pettis Property of order $p$
[ "math.FA" ]
In the present paper we study unconditionally $p$-converging operators and Dunford-Pettis property of order $p$. New characterizations of unconditionally $p$-converging operators and Dunford-Pettis property of order $p$ are established. Six quantities are defined to measure how far an operator is from being unconditionally $p$-converging. We prove quantitative versions of relationships of completely continuous operators,unconditionally $p$-converging operators and unconditionally converging operators. We further investigate possible quantifications of the Dunford-Pettis property of order $p$.
math.FA
math
UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI Abstract. In the present paper we study unconditionally p-converging operators and Dunford-Pettis property of order p. New characterizations of unconditionally p-converging operators and Dunford-Pettis property of order p are established. Six quantities are defined to measure how far an operator is from being unconditionally p-converging. We prove quantitative versions of relationships of completely continu- ous operators,unconditionally p-converging operators and unconditionally converg- ing operators. We further investigate possible quantifications of the Dunford-Pettis property of order p. 1. Introduction and notations Throughout the paper, p∗ denotes the conjugate number of p for 1 ≤ p < ∞; if p = 1, lp∗ plays the role of c0. X, Y will denote real (or complex) Banach spaces and L(X, Y ) the space of all the operators (=continuous linear maps) between X and Y . K(X, Y ) denotes the space of all the compact operators between X and Y . Let X be a Banach space, 1 ≤ p < ∞ and we denote lp(X) by the set of all p-summable p (X) be the p (X) is a Banach space with sequences in X with the natural norm k(xn)nkp = (P∞n=1 kxnkp) set of all weakly p-summable sequences in X. Then lw the norm p . Let lw 1 k(xn)nkw p = sup{( ∞X n=1 < x∗, xn > p) 1 p : x∗ ∈ BX ∗}, ∀ (xn)n ∈ lw p (X). It is a well-known result of A. Grothendieck ([15],[12,Proposition 2.2])that the canon- ical correspondence T 7→ (T en)n provides an isometric isomorphism of L(lp∗, X) onto lw p (X). A sequence (xn)n ∈ lw p (X) is unconditionally p-summable if sup{( ∞X n=m < x∗, xn > p) 1 p : x∗ ∈ BX ∗} → 0 as m → ∞. Date: August 27, 2018. Dongyang Chen's project was supported by the Natural Science Foundation of Fujian Province of China(No. 2015J01026). Lei Li is the corresponding author and his research is partly supported by the NSF of China (11301285). 1 2 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI We denote the set of all unconditionally p-summable sequences on X by lu p (X). It is obvious that (xn)n is unconditionally 1-summable if and only if (xn)n is uncondi- tionally summable. J. H. Fourie and J. Swart proved that the same correspondence T 7→ (T en)n provides an isometric isomorphism of K(lp∗, X) onto lu p (X) (see [14]). Let us recall that an operator T : X → Y is unconditionally converging if T takes weakly 1-summable sequences to unconditionally 1-summable sequences. For p = ∞, the space lu ∞(X) is identical to c0(X), the space of all norm null sequences in X. Hence- forth, for p = ∞, we refer to consider the space cw 0 (X) of weakly null sequences in X, instead of lw ∞(X) = l∞(X). Recall that an operator T : X → Y is completely contin- uous if T takes weakly null sequences to norm null sequences. It is well-known that p-summing operators are precisely those operators which take weakly p-summable se- quences(unconditionally p-summable sequences) to p-summable sequences. A natural question arises: what are operators which take weakly p-summable sequences to un- conditionally p-summable sequences? This is the starting point of our investigation. The paper is organized as follows: In Section 2, we introduce the concept of unconditionally p-converging operators(1 ≤ p ≤ ∞), which is the extension of unconditionally converging operators and com- pletely continuous operators. It is proved that unconditionally p-converging opera- tors coincide with the p-converging operators introduced by J. M. F. Castillo and F. S´anchez in [7] although their original definitions are different. New concepts of weakly p-Cauchy sequences and weakly p-limited sets are introduced to characterize unconditionally p-converging operators. We establish characterizations of weakly p- limited sets and investigate connections between weakly p-limited sets and relatively norm compact sets. A counterexample is constructed to show that an operator is unconditionally p-converging not precisely when its second adjoint is. Section 3 is concerned with Dunford-Pettis property of order p (DP Pp for short) introduced in [7], which is a generalization of the classical Dunford-Pettis property. It turns out that many classical spaces failing Dunford-Pettis property enjoy DP Pp, such as Hardy space H 1 and Lorentz function spaces Λ(W, 1). In this section, we use weakly p-Cauchy sequences and weakly p-limited sets to characterize DP Pp. New characterizations of DP Pp in dual spaces are obtained. We also introduce the notion of hereditary Dunford-Pettis property of order p and establish its characterizations. In particular, we prove that a Banach space X has the hereditary DP Pp if and only if every weakly p-summable sequence in X admits a weakly 1-summable subsequence. UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p3 Finally, the surjective Dunford-Pettis property of order p, a formally weaker property than DP Pp, is introduced and its characterizations are obtained. In the last two sections of the present paper we investigate possibilities of quantify- ing unconditionally p-converging operators and the Dunford-Pettis property of order p. This is inspired by a large number of recent results on quantitative versions of various theorems and properties of Banach spaces (see [1,3,13,17,18,19]). Section 4 contains quantitative versions of the implications among three classes of operators- completely continuous,unconditionally p-converging and unconditionally converging ones. M. Kacena, O. F. K. Kalenda and J. Spurn´y have already defined a quan- tity measuring how far an operator is from being completely continuous in [17]. In this section, we define another equivalent quantity measuring complete continuity of an operator. We further define six quantities measuring how far an operator is from being unconditionally p-converging. Moreover, we show that one of the six new quantities is equal to the quantity defined in [20] to measure how far an operator is unconditionally converging in case of p = 1. In Section 5 we introduce a new locally convex topology and give two topolog- ical characterizations of Dunford-Pettis property of order p. Using the introduced quantity measuring unconditional p-convergence of an operator and the new locally convex topology, we show that the Dunford-Pettis property of order p is automati- cally quantitative in a sense. We also define two quantities measuring how far a set is weakly p-limited. One of the two new quantities is used to quantify the Dunford- Pettis property of order p. The other is used to define a stronger quantitative version of Dunford-Pettis property of order p. Several characterizations of this quantitative version of Dunford-Pettis property of order p are established. The reader is referred to [12] and [22] for any unexplained notation or terminology. 2. Unconditionally p-converging operators Definition 2.1. Let 1 ≤ p ≤ ∞. We say that an operator T : X → Y is uncondi- tionally p-converging if T takes a weakly p-summable sequence (xn)n ∈ lw p (X)((xn)n ∈ cw 0 (X) for p = ∞) to an unconditionally p-summable sequence (T xn)n ∈ lu p (Y )((xn)n ∈ c0(Y ) for p = ∞). We begin with a simple, but extremely useful, characterization of unconditionally p-converging operators. 4 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI Theorem 2.1. Let 1 ≤ p < ∞. The following are equivalent for an operator T : X → Y : (1) T is unconditionally p-converging; (2) T S is compact for any operator S ∈ L(lp∗, X)(L(c0, X) for p = 1). Proof. (1) ⇒ (2). Let S ∈ L(lp∗, X)(1 < p < ∞)(L(c0, X) for p = 1). By the ideal property of unconditionally p-converging operators, T S is unconditionally p- converging. Since (en)n is weakly p-summable in lp∗(1 < p < ∞)(c0 for p = 1), (T Sen)n is unconditionally p-summable. Then there exists a compact operator R : lp∗ → X such that Ren = T Sen(n = 1, 2, ...). Thus T S is compact. (2) ⇒ (1). Let (xn)n ∈ lw p (X). Then there exists an operator S : lp∗ → X(1 < p < ∞)(S : c0 → X for p = 1) such that Sen = xn(n = 1, 2, ...). By (2), we get (T Sen)n is unconditionally p-summable. Thus T S is unconditionally p-converging. (cid:3) Before another frequently useful characterization of unconditionally p-converging operators is given, we recall the notion of weakly p-convergent sequences introduced in [8]. A sequence (xn)n in a Banach space X is said to be weakly p-convergent to x ∈ X(1 ≤ p ≤ ∞) if the sequence (xn − x)n is weakly p-summable in X. Weakly ∞-convergent sequences are simply the weakly convergent sequences. It is natural to generalize weakly Cauchy sequences to the general case 1 ≤ p ≤ ∞. Definition 2.2. Let 1 ≤ p ≤ ∞. We say that a sequence (xn)n in a Banach space X is weakly p-Cauchy if for each pair of strictly increasing sequences (kn)n and (jn)n of positive integers, the sequence (xkn − xjn)n is weakly p-summable in X. Obviously, every weakly p-convergent sequence is weakly p-Cauchy, and the weakly ∞-Cauchy sequences are precisely the weakly Cauchy sequences. Theorem 2.2. Let 1 ≤ p ≤ ∞. The following statements about an operator T : X → Y are equivalent: (1) T is unconditionally p-converging; (2) T sends weakly p-convergent sequences onto norm convergent sequences; (3) T sends weakly p-Cauchy sequences onto norm convergent sequences. Proof. (1) ⇒ (2). Suppose that (xn)n is weakly p-convergent in X. We may assume that (xn)n is weakly p-summable. Then there exists an operator S : lp∗ → X, 1 < p < ∞(S : c0 → X for p = 1) such that Sen = xn(n = 1, 2, ...). By Theorem 2.1, T S is UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p5 compact and hence (T Sen)n is relatively compact. Consequently, limn→∞ kT Senk = 0. (2) ⇒ (3). Let (xn)n be a weakly p-Cauchy sequence in X. By (2), for each pair of strictly increasing sequences (kn)n and (jn)n of positive integers, the sequence (T xkn − T xjn)n converges to 0 in norm and hence (T xn)n converges in norm. (3) ⇒ (1). Suppose that T is not unconditionally p-converging. By Theorem 2.1, the operator T S is non-compact for some operator S ∈ L(lp∗, X)(1 < p < ∞)(L(c0, X) for p = 1). Then there exists a weakly null sequence (zn)n in lp∗(1 < p < ∞)(c0 for p = 1) such that kT Sznk > ǫ0 > 0(n = 1, 2, ...). By passing to subsequences, we may assume that the sequence (zn)n is equivalent to the unit vector basis (en)n in lp∗. Let R : lp∗ → lp∗ be an isomorphic embedding with Ren = zn(n = 1, 2, ...). Let xn = SRen. Then (xn)n is weakly p-summable in X and hence weakly p-Cauchy. By the assumption, (T xn)n converges to 0 in norm, but kT xnk > ǫ0 > 0(n = 1, 2, ...), which is a contradiction. (cid:3) It should be noted that Theorem 2.2(2) is the definition of the so called p-converging operators defined by J. M. F. Castillo and F. S´anchez in [7]. In this note, we use the terminology unconditionally p-converging operators instead of p-converging operators. Recall that a subset K of a Banach space X is relatively weakly p-compact (1 ≤ p < ∞) if K is contained in S(Blp∗ ) for 1 < p < ∞(S(Bc0) for p = 1) for some operator S from lp∗(c0 for p = 1) into X (see [25]). A subset K of a Banach space X is said to be relatively weakly p-precompact if every sequence in K admits a weakly p < ∞. But the converse needs not to be true. Let X = (P∞n=1 ln p-convergent subsequence (see [6]). Bessaga-Pe lczy´nski Selection Principle yields that every relatively weakly p-compact set is relatively weakly p-precompact for any 1 < 1 )p∗(1 < p < ∞). It follows from Bessaga-Pe lczy´nski Selection Principle that BX is relatively weakly p- precompact. But BX is not relatively weakly p-compact because X is not isomorphic to a quotient of lp∗. Another counterexample is Lp(1 < p < ∞, p 6= 2). For each 1 < p < ∞, p 6= 2, BLp is relatively weakly r-precompact, where r = max(p∗, 2), but is not relatively weakly r-compact because such Lp is not isomorphic to a quotient of lr∗. By using the weakly p-Cauchy sequences, we can correspondingly define the con- ditionally weakly p-compact sets as follows: 6 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI Definition 2.3. Let 1 ≤ p ≤ ∞. We say that a subset K of a Banach space X is conditionally weakly p-compact if every sequence in K admits a weakly p-Cauchy subsequence. The following result,which follows from Theorem 2.2, says that unconditionally p-converging operators are precisely those operators that send conditionally weakly p-compact subsets onto relatively norm compact subsets. Theorem 2.3. Let T ∈ L(X, Y ) and 1 ≤ p < ∞. The following statements are equivalent: (1) T is unconditionally p-converging; (2) T maps relatively weakly p-precompact subsets onto relatively norm compact sub- sets; (3) T maps conditionally weakly p-compact subsets onto relatively norm compact sub- sets; (4) T maps relatively weakly p-compact subsets onto relatively norm compact subsets. Definition 2.4. Let X be a Banach space and 1 ≤ p < ∞. We say that a bounded subset K of X∗ is weakly p-limited if limn→∞ supx∗∈K < x∗, xn > = 0 for every (xn)n ∈ lw p (X). The following result, an immediate consequence of Theorem 2.2, is a characteriza- tion of unconditionally p-converging operators in terms of weakly p-limited subsets. Theorem 2.4. Let 1 ≤ p < ∞. The following are equivalent for an operator T : X → Y : (1) T is unconditionally p-converging; (2) T ∗ maps bounded subsets of Y ∗ onto weakly p-limited subsets of X∗. J.M.F.Castillo and F.S´anchez said that a Banach space X ∈ Wp(1 ≤ p < ∞) if any bounded sequence in X admits a weakly p-convergent subsequence (see [8]). We use this notion to characterize weakly p-limited sets. Theorem 2.5. Let 1 < p < ∞ and X be a Banach space. The following statements are equivalent about a bounded subset K of X∗: (1) K is weakly p-limited; (2) For all spaces Y ∈ Wp and for every operator T from Y into X, the subset T ∗(K) is relatively norm compact; UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p7 (3) For every operator T from lp∗ into X, the subset T ∗(K) is relatively norm compact. Proof. (1) ⇒ (2). Let T be an operator from Y ∈ Wp into X such that T ∗(K) is not relatively norm compact. Then there exists a sequence (x∗n)n in K such that (T ∗x∗n)n admits no norm convergent subsequences. Since Y ∗ is reflexive, by passing to a subsequence if necessary we may assume that (T ∗x∗n)n converges weakly to some y∗ ∈ Y ∗ and kT ∗x∗n − y∗k > ǫ0 for some ǫ0 > 0 and for all n ∈ N. For each n, choose yn with kynk ≤ 1 such that < T ∗x∗n − y∗, yn > > ǫ0. Since Y ∈ Wp, by passing to a subsequence again if necessary one can assume that the sequence (yn)n is weakly p-convergent to some y ∈ Y . Thus, by hypothesis, we get limn→∞ supx∗∈K < x∗, T yn − T y > = 0. Note that, for each n ∈ N, < T ∗x∗n−y∗, yn > ≤ < x∗n, T yn−T y > + < x∗n, T y > − < y∗, y > + < y∗, y−yn > . This implies that limn→∞ < T ∗x∗n − y∗, yn >= 0, which is a contradiction. (2) ⇒ (3) is immediate because lp∗ ∈ Wp; (3) ⇒ (1). Let (xn)n ∈ lw p (X). Then there exists an operator T from lp∗ into X such that T en = xn for all n ∈ N. It follows from (3) that T ∗(K) is relatively norm compact. By the well-known characterization of relatively norm compact subsets of lp, one can derive that limn→∞ supx∗∈K < x∗, xn > = 0. (cid:3) By Theorem 2.5, we see that relatively norm compact sets are weakly p-limited. But Theorem 2.4 demonstrates that there are many weakly p-limited sets which are not relatively norm compact. Indeed, for each 1 < p < ∞ and for each 1 < r < p∗, the identity map Ir on lr is unconditionally p-converging and hence the unit ball Blr∗ of lr∗ is weakly p-limited. In the following result, we use biorthogonal sequences to characterize weakly p-limited sets which are not relatively norm compact. Theorem 2.6. Suppose that X is reflexive and K is a weakly p-limited subset of X∗. If K is not relatively norm compact, then there exits a seminormalized biorthogonal sequence (xn, x∗n)n in X × (K − K) such that (x∗n)n is a basic sequence and (xn)n has no weakly p-Cauchy subsequence. Proof. Suppose that K is not relatively norm compact, and let (fn)n be a sequence in K with no norm convergent subsequence. Since X is reflexive, we may assume that the sequence (fn)n converges weakly. Then there exist two strictly increasing sequences (kn)n and (jn)n of positive integers and ǫ0 > 0 such that kfkn − fjnk > ǫ0 8 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI for all n ∈ N. Let x∗n = fkn − fjn ∈ (K − K). Then (x∗n)n is weakly null. By Bessaga- Pe lczy´nski Selection Principle, we can assume that (x∗n)n is a basic sequence. Let (x∗∗n )n be the associated sequence of coefficient functionals, and for each n ∈ N, let xn ∈ X be a Hahn-Banach extension of x∗∗n to all of X∗. Then the sequence (xn, x∗n)n is seminormalized and biorthogonal. It remains to show that (xn)n has no weakly p-Cauchy subsequence. If (yn)n is a weakly p-Cauchy subsequence of (xn)n, then (yn+1 − yn)n is weakly p-summable. Since K is weakly p-limited, the subset K − K is also weakly p-limited, which implies that limn→∞ supk < x∗k, yn+1 − yn > = 0. This is impossible because (xn, x∗n)n is biorthogonal. (cid:3) A consequence of Theorem 2.6 is that for any 1 < p < ∞, there exists a relatively weakly compact sequence that admits no weakly p-Cauchy subsequence. Moreover, it should be noted that the converse of Theorem 2.6 is true. Actually, it is easy to verify that if K is a subset of X∗ and the sequence (xn, x∗n)n in X × (K − K) is biorthogonal with supn kxnk < ∞, then K is not relatively norm compact. The following result shows that an operator is unconditionally p-converging not precisely when its second adjoint is. Theorem 2.7. (1) Let T ∈ L(X, Y ) and 1 ≤ p ≤ ∞. If T ∗∗ is unconditionally p-converging, then T is unconditionally p-converging; (2) For each 1 ≤ p ≤ ∞, there exists an unconditionally p-converging operator T , but T ∗∗ is not unconditionally p-converging. Proof. (1). By the ideal property of unconditionally p-converging operators, JY T is unconditionally p-converging, where JY : Y → Y ∗∗ is the canonical mapping. Let S ∈ L(lp∗, X)(1 < p < ∞)(L(c0, X) for p = 1). By Theorem 2.1, JY T S is compact and hence T S is compact. Again by Theorem 2.1, T is unconditionally p-converging. (2). J. Bourgain and F. Delbaen (see [5]) constructed a Banach space XBD such that XBD has the Schur property and X∗∗BD is isomorphically universal for separable Banach spaces. Since XBD has the Schur property, every operator from lp(1 < p < ∞) and from c0 into XBD is compact. By Theorem 2.1, every operator with domain XBD is unconditionally p-converging for each 1 ≤ p < ∞. In particular, the identity map IXBD on XBD is unconditionally p-converging. But since X∗∗BD is isomorphically UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p9 universal for separable Banach spaces, there exists a closed subspace Xp∗(X0 for p = 1) of X∗∗BD such that Xp∗ is isomorphic to lp∗ for 1 < p < ∞ (X0 is isomorphic to c0 for p = 1). This implies that I∗∗XBD = IX ∗∗ is not lp∗-strictly singular for 1 < p < ∞ (c0- strictly singular for p = 1). Thus I∗∗XBD = IX ∗∗ is not unconditionally p-converging. For p = ∞, the identity map IXBD is obviously completely continuous, but I∗∗XBD = IX ∗∗ (cid:3) is not completely continuous because X∗∗BD has not the Schur property. BD BD BD 3. Dunford-Pettis Property of order p Let us recall that a Banach space X has the Dunford-Pettis property (in short, DPP) if for every Banach space Y , every weakly compact operator T : X → Y is completely continuous (see [16]). An operator T : X → Y is said to be weakly compact if T BX is relatively weakly compact in Y . J. M. F. Castillo and F. S´anchez extended the classical Dunford-Pettis property to the general case for 1 ≤ p ≤ ∞ in [7]. Let 1 ≤ p ≤ ∞. A Banach space X is said to have the Dunford-Pettis property of p (in short, DP Pp) if for every Banach space Y , every weakly compact operator T : X → Y is unconditionally p-converging. Many classical spaces failing the DPP enjoy the DP Pp. A simple observation is that if a Banach space X has cotype q < ∞, then X has the DP Pp for any 1 < p < q∗. Thus, the classical Hardy space H 1, which fails the DPP (see [10]), has the DP Pp for any 1 < p < 2. It is known that all the Lorentz function spaces Λ(W, 1)'s fail the DPP (see [10]). But there are certain positive results for DP Pp. For example, if we take W (t) = 1 2√t , t ∈ (0, 1], then the space Λ(W, 1) has the DP Pp for some 1 < p ≤ 2. Another non-reflexive space failing the DPP is the interesting space L built in [21]. Indeed, it was shown in [4]that even duals of L fail the DPP and odd duals of L fail the surjective DPP, which is genuinely weaker than the DPP. Moreover, F. Bombal, P. Cembranos and J. Mendoza proved that for any 1 ≤ p < ∞,every operator from L into lp is compact (see [4]). This means that L∗ has the DP Pp for any 1 < p < ∞. More examples can be found in [7]. Let us start with a characterization of the DP Pp by means of weakly p-limited sets. Theorem 3.1. Let 1 < p < ∞. A Banach space X has the DP Pp if and only if each relatively weakly compact subset of X∗ is weakly p-limited. Proof. The sufficient part follows immediately from Theorem 2.4. On the other hand, let K be a relatively weakly compact subset of X∗. By the Davis-Figiel-Johnson- Pe lczy´nski factorization lemma (see [9]), there exists a reflexive space Z, which is a 10 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI linear subspace of X∗, such that the inclusion map J : Z → X∗ is bounded and the unit ball BZ of Z contains K. Since Z is reflexive, there is an operator T : X → Z∗ such that T ∗ = J. By the assumption, T is unconditionally p-converging. By Theorem 2.4, the set T ∗(BZ) = J(BZ) = BZ is weakly p-limited in X∗. Thus K is also weakly p-limited. (cid:3) Let us remark that for each 1 < p < ∞, there exists a weakly p-limited set which is not relatively weakly compact. Indeed, we take X = L∗, where the space L is built in [21]. As mentioned above, the identity IX on X is unconditionally p-converging for each 1 < p < ∞. It follows from Theorem 2.4 that the unit ball BX ∗ is weakly p-limited, but it is not weakly compact because the space L is non-reflexive. The following result is an internal characterization of the DP Pp. It is a refinement of [7,Proposition 3.2]. Theorem 3.2. Let 1 < p < ∞ and X be a Banach space. The following are equiva- lent: (1) X has the DP Pp; (2) Every weakly compact operator T from X into c0 is unconditionally p-converging; (3) limn→∞ < x∗n, xn >= 0, for every weakly p-Cauchy sequence (xn)n in X and every weakly null sequence (x∗n)n in X∗; (4) limn→∞ < x∗n, xn >= 0, for every (xn)n ∈ lw (x∗n)n in X∗; (5) limn→∞ < x∗n, xn >= 0, for every (xn)n ∈ lw (x∗n)n in X∗. p (X) and every weakly Cauchy sequence p (X) and every weakly null sequence Proof. (1) ⇒ (2) is trivial. (2) ⇒ (3). Given a weakly p-Cauchy sequence (xn)n in X and a weakly null sequence (x∗n)n in X∗. Define an operator T : X → c0 by T x = (< x∗n, x >)n. Since (x∗n)n converges to 0 weakly, T ∗ is weakly compact and so is T . By (2), T is unconditionally p-converging. By Theorem 2.2, (T xn)n converges to some ξ = (ξk)k ∈ c0 in norm. Let ǫ > 0. There exists a positive integer N1 such that kT xn − ξk < ǫ 2 for all n > N1. Choose another positive integer N2 such that ξk < ǫ 2 for all k > N2. By the definition of T , we have < x∗n, xn > < ǫ for all n > max(N1, N2). Thus limn→∞ < x∗n, xn >= 0. (3) ⇒ (4) is trivial. (4) ⇒ (5). If (xn)n is weakly p-summable in X and (x∗n)n is weakly Cauchy in X∗, yet (< x∗n, xn >)n does not converge to 0. By passing to subsequences, we may UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p11 assume that < x∗n, xn > > ǫ0 for some ǫ0 > 0 and all n ∈ N. Since (xn)n is weakly p-summable and in particular weakly null, there exists a subsequence (xkn)n of (xn)n such that < x∗n, xkn > < ǫ0 2 for all n ∈ N. Since (x∗n)n is weakly Cauchy, we see that (x∗kn − x∗n)n is weakly null. By (3), limn→∞ < x∗kn − x∗n, xkn >= 0. This implies that < x∗kn − x∗n, xkn > < ǫ0 3 for n large enough. But for such n's, we have ǫ0 < < x∗kn, xkn > ≤ < x∗kn − x∗n, xkn > + < x∗n, xkn > < 5ǫ0 6 . (5) ⇒ (1). Let T : X → Y be a weakly compact operator. Let us suppose that T is not unconditionally p-converging. Appealing again to Theorem 2.2, we obtain a weakly p-summable sequence (xn)n in X and ǫ0 > 0 such that kT xnk > ǫ0(n = 1, 2, ...). Pick y∗n ∈ Y ∗ such that < y∗n, T xn >= kT xnk and ky∗nk = 1 for all n ∈ N. Since T is weakly compact, so is T ∗. Hence there is a subsequence (y∗kn)n of (y∗n)n such that the sequence (T ∗y∗kn)n converges weakly and hence is weakly Cauchy. The assumption ensures that the sequence (< T ∗y∗kn, xkn >)n = (kT xknk)n converges to 0, which is a contradiction. (cid:3) Corollary 3.3. Let 1 < p < ∞. If X∗∗ has the DP Pp, then so is X. The converse of Corollary 3.3 is not true. In fact, the Banach space X = (Pn ln 2 )c0 enjoys the DPP, but X∗∗ = (Pn ln 2 )l∞ contains a complemented copy of l2. Since l2 fails the DP Pp for any 2 ≤ p < ∞, X∗∗ also fails the DP Pp for any 2 ≤ p < ∞. In the case of the classical DPP, there is a result better than Corollary 3.3: If X∗ has the DPP, then X has the DPP too (see [10]). The analogous result is not true for the DP Pp: for each 1 < p < ∞, every operator from lp into Tsirelson's space T is compact, hence T has the DP Pp for any 1 < p < ∞. But, for each 1 < p < ∞, there is a non-compact operator from lp into T ∗. Thus, for each 1 < p < ∞, T ∗ fails the DP Pp. Corollary 3.4. Suppose that a Banach space X contains no copy of l1 and let 1 < p < ∞. The following statements are equivalent: (1) X∗ has the DP Pp; (2) For all Banach spaces Y , every weakly compact operator T : Y → X has the unconditionally p-converging adjoint; (3) limn→∞ < x∗n, xn >= 0, for every (x∗n)n ∈ lw quence (xn)n in X; p (X∗) and every weakly Cauchy se- 12 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI (4) limn→∞ < x∗n, xn >= 0, for every weakly p-Cauchy sequence (x∗n)n in X∗ and every weakly null sequence (xn)n in X; (5) limn→∞ < x∗n, xn >= 0, for every (x∗n)n ∈ lw (xn)n in X. p (X∗) and every weakly null sequence Proof. We only prove (2) ⇒ (3) and (5) ⇒ (1). (2) ⇒ (3). Assuming the contrary, we can find (x∗n)n ∈ lw p (X∗) and a weakly Cauchy sequence (xn)n in X such that < x∗n, xn > > ǫ0 for some ǫ0 > 0 and all n ∈ N. Since (x∗n)n is weakly null, there exists a subsequence (x∗kn)n of (x∗n)n such that < x∗kn, xn > < ǫ0 2 for all n ∈ N. Define an operator S : X∗ → c0 by 2 for all n ∈ N. Thus < x∗kn, xn − xkn > > ǫ0 Sx∗ = (< x∗, xn − xkn >)n, x∗ ∈ X∗. It is easy to check that S∗en = xn − xkn(n = 1, 2, ...), where (en)n is the unit vector basis of l1. Thus the operator S∗ maps l1 into X and is weakly compact. By (2), the operator S∗∗ is unconditionally p-converging. Moreover, an easy verification shows that S∗∗ = S. By Theorem 2.2, we get limn→∞ kSx∗knk = 0. It follows from the definition of the operator S that limn→∞ < x∗kn, xn − xkn > = 0, which is a contradiction. (5) ⇒ (1). By Theorem 3.2, it is enough to verify that for every (x∗n)n ∈ lw p (X∗) and every weakly null sequence (x∗∗n )n in X∗∗, the sequence (< x∗∗n , x∗n >)n converges to 0. Now we suppose that it is false. Then, by passing to subsequences, we may assume that < x∗∗n , x∗n > > ǫ0 for some ǫ0 > 0 and all n ∈ N. Of course, we may also assume that kx∗∗n k ≤ 1 for all n ∈ N. It follows from Goldstine's Theorem that for each n ∈ N, there exists an xn ∈ BX such that < xn − x∗∗n , x∗n > < ǫ0 2 . Then < x∗n, xn > > ǫ0 2 for all n ∈ N. By Rosenthal's Theorem, (xn)n has a weakly Cauchy subsequence, which is still denoted by (xn)n. Then there exists a subsequence (x∗kn)n of (x∗n)n such that < x∗kn, xn > < ǫ0 3 for all n ∈ N. By (5), we get limn→∞ < x∗kn, xkn − xn >= 0, which implies that < x∗kn, xkn − xn > < ǫ0 for n large enough. It is easy to verify that for such n's, < x∗kn, xkn > < ǫ0 2 . This contradiction completes the proof. 6 Definition 3.1. Let 1 < p < ∞. We say that a Banach space X has the heredi- tary Dunford-Pettis property of order p (in short, hereditary DP Pp)if every (closed) subspace of X has the DP Pp. (cid:3) UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p13 We present a useful characterization of hereditary DP Pp. We need a J. Elton's result that can be found in [11]. Lemma 3.5. [11] If (xn)n is a normalized weakly null sequence of a space X such that no subsequence of it is equivalent to the unit vector basis (en)n of c0, then (xn)n has a subsequence (yn)n for which given any subsequence (zn)n of (yn)n and any sequence (αn)n∈c0 we have supn kPn k=1 αkzkk = +∞. Theorem 3.6. Let X be Banach space and 1 < p < ∞. The following are equivalent: (1) X has the hereditary DP Pp; (2) Every normalized weakly p-summable sequence in X admits a subsequence that is equivalent to the unit vector basis of c0; (3) Every weakly p-summable sequence in X admits a weakly 1-summable subsequence; (4) Every weakly p-summable sequence in X admits a subsequence (yn)n such that supN kPN n=1 ynk < ∞. Proof. (1) ⇒ (2). Let (xn)n be a normalized weakly p-summable sequence in X such that it admits no subsequence that is equivalent to the unit vector basis (en)n of c0. It follows from Lemma 3.5 that (xn)n has a subsequence (yn)n as stated in Lemma 3.5. By Bessaga-Pe lczy´nski Selection Principle, we may assume that (yn)n is a basic sequence. Let X0 = span{yn : n = 1, 2, ...}. Let (y∗n)n ⊂ X∗0 be the coefficient functionals of the basic sequence (yn)n. For each N, define a projection PN : X0 → X0 by PN (y) = NX n=1 < y∗n, y > yn, y ∈ X0. Then the projection PN 's are uniformly bounded in operator norm. An easy verifi- cation shows that P ∗∗N y∗∗ = PN n=1 < y∗∗, y∗n > yn for all y∗∗ ∈ X∗∗0 . Lemma 3.5 and the uniform boundedness of the projection PN 's imply that (< y∗∗, y∗n >)n ∈ c0 for all y∗∗ ∈ X∗∗0 , that is, (y∗n)n is weakly null. Since < y∗n, yn >= 1 for all n ∈ N, it follows from Theorem 3.2 again that X0 fails the DP Pp. (2) ⇒ (3) and (3) ⇒ (4) are obvious. (4) ⇒ (1). Take a subspace X0 of X that fails the DP Pp. Appealing to Theorem 3.2, we obtain a weakly compact operator T : X0 → c0 which is not unconditionally p-converging. Applying Theorem 2.2, we get a normalized weakly p-summable se- quence (xn)n in X such that kT xnk ≥ ǫ0 for all n ∈ N. Bessaga-Pe lczy´nski Selection 14 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI Principle allows us to assume that the sequence (T xn)n is equivalent to the unit vec- tor basis (en)n of c0. By the weak compactness of T , the sequence (xn)n admits no subsequence equivalent to the unit vector basis (en)n. By Lemma 3.5, the sequence (xn)n admits a subsequence (yn)n for which given any subsequence (zn)n of (yn)n, one has supN kPN n=1 znk = ∞. A direct consequence of Theorem 3.6 is the following corollary: (cid:3) Corollary 3.7. If a Banach space X has the hereditary DP Pp, then each weakly p- summable sequence in X admits a subsequence (xn)n such that limn→∞ kPn k=1 xkk/n 1 p∗ = 0. We close this section with the surjective DP Pp, a formally weaker property than the DP Pp. By the Davis-Figiel-Johnson-Pe lczy´nski's factorization theorem (see [9]), a Banach space X has the DP Pp if and only if for all reflexive spaces Y , every operator from X into Y is unconditionally p-converging. We introduce the surjective DP Pp by imposing that every surjective operator from X onto the reflexive space Y is unconditionally p-converging. The motivation for introducing the surjective DP Pp was to extend the surjective DPP introduced in [21]. Definition 3.2. Let 1 < p < ∞. We say that a Banach space X has the surjec- tive DP Pp if for all reflexive spaces Y , every surjective operator from X onto Y is unconditionally p-converging. The following are the internal characterizations of the surjective DP Pp. Theorem 3.8. The following are equivalent for a Banach space X and 1 < p < ∞: (1) X has the surjective DP Pp; (2) limn→∞ < x∗n, xn >= 0, for every weakly p-Cauchy sequence (xn)n in X and every weakly null sequence (x∗n)n in X∗ such that span{x∗n : n = 1, 2, ...} is reflexive; (3) limn→∞ < x∗n, xn >= 0, for every (xn)n ∈ lw (x∗n)n in X∗ such that span{x∗n : n = 1, 2, ...} is reflexive; (4) limn→∞ < x∗n, xn >= 0, for every (xn)n ∈ lw (x∗n)n in X∗ such that span{x∗n : n = 1, 2, ...} is reflexive. p (X) and every weakly Cauchy sequence p (X) and every weakly null sequence Proof. (1) ⇒ (2). Let (xn)n ⊂ X and (x∗n)n ⊂ X∗ be as in (2). Let Z = span{x∗n : n = 1, 2, ...}. Then (Z⊥)⊥ = Z, where Z⊥ := {x ∈ X :< x∗, x >= 0 for all x∗ ∈ Z} UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p15 and (Z⊥)⊥ := {x∗ ∈ X∗ :< x∗, x >= 0 for all x ∈ Z⊥}. Let Q : X → X/Z⊥ be the natural quotient. Then Q∗ : (X/Z⊥)∗ → Z is a surjective isometrical isomorphism. Let Q∗fn = x∗n, fn ∈ (X/Z⊥)∗ for all n ∈ N. By (1), the quotient Q is unconditionally p-converging. By Theorem 2.2, the sequence (Qxn)n converges in norm to Qx for some x ∈ X. Thus < x∗n, xn − x > = < fn, Qxn − Qx > ≤ (sup n kfnk)kQxn − Qxk → 0 (n → ∞). Since (x∗n)n is weakly null, x∗n, xn >= 0. limn→∞ < x∗n, x >= 0.Therefore we have limn→∞ < (2) ⇒ (3) is obvious. (3) ⇒ (4). Suppose that (4) is false. Then there exist a sequences (xn)n ∈ lw p (X) and a weakly Cauchy sequence (x∗n)n in X∗ such that span{x∗n : n = 1, 2, ...} is reflexive so that < x∗n, xn > > ǫ0 > 0 for all n ∈ N. Since the sequence (xn)n converges to 0 weakly, there is a subsequence (xkn)n of (xn)n such that < x∗n, xkn > < ǫ0 for all n ∈ N. Since the space span{x∗n : n = 1, 2, ...} is reflexive, the space span{x∗n − x∗kn : n = 1, 2, ...} is reflexive too. By the hypothesis, limn→∞ < x∗n − x∗kn, xkn >= 0. Thus, < x∗n − x∗kn, xkn > < ǫ0 2 for n large enough, which implies that for such n's, < x∗kn, xkn > < ǫ0, a contradiction. 2 (4) ⇒ (1). Suppose that X fails the surjective DP Pp. Then there exists a surjective operator T from X onto a reflexive space Y such that T is not unconditionally p- converging. By Theorem 2.2, there exists a normalized weakly p-summable sequence (xn)n in X such that kT xnk > ǫ0 for all n ∈ N. For each n, choose y∗n ∈ Y ∗ with ky∗nk = 1 such that < y∗n, T xn >= kT xnk. By the reflexivity of Y , we may assume that the sequence (y∗n)n converges to 0 weakly by passing to subsequences if necessary. Let x∗n = T ∗y∗n. Then the sequence (x∗n)n converges to 0 weakly too. Since T is surjective, the operator T ∗ : Y ∗ → X∗ is an isomorphic embedding. This implies that the space span{x∗n : n = 1, 2, ...} is contained in T ∗(span{y∗n : n = 1, 2, ...}) and hence is reflexive. By (4), limn→∞ < x∗n, xn >= 0, a contradiction because < x∗n, xn >> ǫ0 for all n ∈ N. This concludes the proof. (cid:3) An immediate consequence of Theorem 3.8 is the following: Corollary 3.9. Let 1 < p < ∞. If X∗∗ has the surjective DP Pp, then so is X. 16 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI We also use the space X = (Pn ln 2 )c0 to show that the converse of Corollary 3.9 is not true. The same argument shows that the space X = (Pn ln 2 )c0 enjoys the surjective DP Pp for any 1 < p < ∞, but X∗∗ also fails the surjective DP Pp for any 2 ≤ p < ∞. The following result analogous to Theorem 3 in [4] shows that the surjective DP Pp and the DP Pp coincide for certain classes of Banach spaces. Theorem 3.10. If a Banach space X contains a complemented copy of l1, then X has the DP Pp if and only if X has the surjective DP Pp. 4. Quantifying unconditionally p-converging operators As discussed above, we see that unconditionally p-converging operators are in- termediate between completely continuous operators and unconditionally converging operators. Precisely, we have the following implications: T completely continuous ⇒ T unconditionally p-converging ⇒ T unconditionally converging. In this section, we quantify these implications. We need some necessary quantities. Let (xn)n be a bounded sequence in a Banach space X. Set ca((xn)n) = inf n sup{kxk − xlk : k, l ≥ n}. This quantity is a measure of non-Cauchyness of the sequence (xn)n. More precisely, ca((xn)n) = 0 if and only if (xn)n is norm Cauchy. In [17], an important quantity measuring how far an operator T : X → Y is from being completely continuous, denoted as cc(T ), is defined by cc(T ) = sup{ca((T xn)n) : (xn)n ⊂ BX weakly Cauchy }. Obviously, T is completely continuous if and only if cc(T ) = 0. In this note, we define another equivalent quantity measuring the complete continuity of an operator T : X → Y as follows: ccn(T ) = sup{lim supn kT xnk : (xn)n ⊂ BX weakly null }. Obviously, T is completely continuous if and only if ccn(T ) = 0. The following theorem demonstrates these two quantities are equivalent. Theorem 4.1. Let T ∈ L(X, Y ). Then ccn(T ) ≤ cc(T ) ≤ 2ccn(T ). To prove Theorem 4.1, we need the following lemma. UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p17 Lemma 4.2. Let X be a Banach space and (xn)n be a weakly null sequence in BX . Let ǫ > 0 be such that kxnk > ǫ for all n ∈ N. Then, for every δ > 0, there is a subsequence (xkn)n of (xn)n such that ca((xkn)n) ≥ ǫ − δ. Proof. We set xk1 = x1. Choose x∗1 ∈ SX ∗ such that < x∗1, xk1 >= kxk1k. Since (xn)n is weakly null, there exists k2 > k1 such that < x∗1, xk2 > < δ. Then kxk1 − xk2k ≥ < x∗1, xk1 − xk2 > ≥ < x∗1, xk1 > − < x∗1, xk2 > ≥ ǫ − δ. Suppose that we have obtained {xk1, xk2, ..., xkn} such that kxki − xknk ≥ ǫ − δ for i = 1, 2, ..., n − 1. Let Yn = span{xk1, xk2, ..., xkn}. Pick a c-net {z1, z2, ..., zm} ⊂ SYn for SYn, where 0 < c < δ 2 . Choose z∗1, z∗2, ..., z∗m in SX ∗ such that < z∗i , zi >= 1 for i = 1, 2, ..., m. Since (xn)n is weakly null, there exists kn+1 > kn such that < z∗i , xkn+1 > < c for all i = 1, 2, ..., m. Then, for each 1 ≤ j ≤ n, there exists 1 ≤ i ≤ m such that k − zik < c. Thus xkj kxkj k kxkj − xkn+1k ≥ < z∗i , xkj − xkn+1 > ≥ 1 − < z∗i , xkn+1 > − < z∗i , xkj − zi > ≥ 1 − c − kxkj − zik ≥ 1 − c − (1 + c − ǫ) = ǫ − 2c ≥ ǫ − δ By induction, we get a subsequence (xkn)n such that kxkn − xkmk ≥ ǫ − δ(n 6= m, n, m = 1, 2, ...). This yields that ca((xkn)n) ≥ ǫ − δ. (cid:3) Proof of Theorem 4.1. Step 1. cc(T ) ≤ 2ccn(T ). We may suppose that cc(T ) > 0 and fix any c > 0 satisfying cc(T ) > c. Then there is a weakly Cauchy sequence (xn)n in BX such that ca((T xn)n) > c. It follows that there exist two strictly increasing sequences (kn)n, (ln)n of positive integers such that kT xkn − T xlnk > c for all n ∈ N. Set zn = (xkn − xln)/2. Then (zn)n is a weakly null sequence in BX and kT znk > c/2 for each n ∈ N. Hence lim supn kT znk ≥ c/2 and then ccn(T ) ≥ c/2. Since c < cc(T ) is arbitrary, we get cc(T ) ≤ 2ccn(T ). Step 2. ccn(T ) ≤ cc(T ). We may suppose that kT k = 1 and ccn(T ) > 0. Suppose that ccn(T ) > ǫ > 0. Then there is a weakly null sequence (xn)n in BX such that lim supn kT xnk > ǫ. This 18 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI yields a subsequence of (xn)n, still denoted by (xn)n, so that kT xnk > ǫ for each n ∈ N. By Lemma 4.2, for every δ > 0, there is a subsequence (xkn)n of (xn)n such that ca((T xkn)n) ≥ ǫ − δ. This means that cc(T ) ≥ ǫ − δ. Since δ > 0 is arbitrary, we get cc(T ) ≥ ǫ. By the arbitrariness of ǫ < ccn(T ), we obtain ccn(T ) ≤ cc(T ). This completes the proof of Theorem 4.1. (cid:3) To quantify unconditionally p-converging operators, we will need two measures of non-compactness. Let us fix some notations. If A and B are nonempty subsets of a Banach space X, we set d(A, B) = inf{ka − bk : a ∈ A, b ∈ B}, bd(A, B) = sup{d(a, B) : a ∈ A}. Thus, d(A, B) is the ordinary distance between A and B, and bd(A, B) is the non- symmetrized Hausdorff distance from A to B. Let A be a bounded subset of a Banach space X. The Hausdorff measure of non- compactness of A is defined by χ(A) = inf{bd(A, F ) : F ⊂ X finite}, χ0(A) = inf{bd(A, F ) : F ⊂ A finite}. Then χ(A) = χ0(A) = 0 if and only if A is relatively norm compact. It is easy to verify that (4.1) χ(A) ≤ χ0(A) ≤ 2χ(A). Now we define five quantities which measure how far an operator is from being unconditionally p-converging. Let T ∈ L(X, Y ) and 1 ≤ p < ∞. We set uc1 p(T ) = sup{lim supn kT xnk : (xn)n ∈ lw p(T ) = sup{ca((T xn)n) : (xn)n ⊂ BX weakly p-Cauchy }, p (X), (xn)n ⊂ BX}, uc2 uc3 p(T ) = sup{ca((T xn)n) : (xn)n ⊂ BX weakly p-convergent }, uc4 p(T ) = sup{χ0(T L) : L ⊂ BX relatively weakly p-compact }, uc5 p(T ) = sup{χ0(T L) : L ⊂ BX relatively weakly p-precompact }. Clearly, uc1 p(T ) = 0 if and only if T is un- conditionally p-converging. It turns out that the above five quantities are equivalent. p(T ) = uc3 p(T ) = uc4 p(T ) = uc2 p(T ) = uc5 Theorem 4.3. Let T ∈ L(X, Y ) and 1 < p < ∞. Then uc5 p(T ) ≤ uc3 p(T ) ≤ uc2 p(T ) ≤ 2uc1 p(T ) ≤ 2uc4 p(T ) ≤ 2uc5 p(T ). UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p19 Proof. Step 1. uc5 p(T ) ≤ uc3 p(T ). We may assume that uc5 p(T ) > 0. Let us fix any 0 < c < uc5 p(T ). Then there exists a relatively weakly p-precompact subset L ⊂ BX such that χ0(T L) > c. By induction, we can construct a sequence (xn)n in L such that kT xn − T xmk > c, n 6= m, n, m = 1, 2, ... Since L is relatively weakly p-precompact, the sequence (xn)n admits a weakly p-convergent subsequence that is still denoted by (xn)n. Thus we get ca((T xn)n) ≥ c, which yields uc3 p(T ) ≥ c. By the arbitrariness of c, we get uc5 p(T ) ≤ uc3 p(T ). Step 2. uc2 We assume that uc2 p(T ) ≤ 2uc1 p(T ). p(T ) > 0 and fix any 0 < c < uc2 p(T ). Then there is a weakly p- Cauchy sequence (xn)n in BX such that ca((T xn)n) > c. By induction, there exist two strictly increasing sequences (kn)n, (ln)n of positive integers such that kT xkn −T xlnk > c for all n ∈ N. Set zn = (xkn − xln)/2. Then (zn)n is a weakly p-summable sequence in BX and kT znk > c/2 for each n ∈ N. Hence uc1 p(T ) ≥ c/2. Since c is arbitrary, we get Step 2. Step 3. uc1 Suppose uc1 p(T ) ≤ uc4 p(T ) > c > 0. Then there exists a weakly p-summable sequence (xn)n in BX such that kT xnk > c for all n ∈ N. We claim that χ0((T xn)n) ≥ c. If this is false, p(T ). we can find a finite subset F of (T xn)n such that bd((T xn)n, F ) < c. Since F is finite, there exist y ∈ F and a subsequence (T xkn)n of (T xn)n such that kT xkn − yk ≤ c for each n ∈ N. Since the sequence (T xkn)n is weakly null, we get kyk ≤ c. This contradiction completes the proof Step 3. p(T ) ≤ uc2 The remaining inequalities uc3 p(T ) are immediate. p(T ) ≤ uc5 p(T ), uc4 It should be mentioned that a quantity is defined in [20] to measure how far an operator is unconditionally converging as follows: (cid:3) uc(T ) = sup{ca(( nX i=1 T xi)n) : (xn)n ∈ lw 1 (X), k(xn)nkw 1 ≤ 1}. Obviously, uc(T ) = 0 if and only if T is unconditionally converging. Inspired by this quantity, we define the sixth quantity measuring how far an operator is uncondition- ally p-converging as follows: uc6 p(T ) = sup{lim sup n kT xnk : (xn)n ∈ lw p (X), k(xn)nkw p ≤ 1}. 20 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI It is obvious that uc6 p(T ) = 0 if and only if T is unconditionally p-converging. This new quantity will be used in next section to prove a quantitative version of the Dunford-Pettis property of order p. Theorem 4.4. Let T ∈ L(X, Y ). Then uc6 1(T ) = uc(T ). Proof. Step 1. uc6 1(T ) ≤ uc(T ). i=1 T xi)n). 1 ≤ 1. It aims to show lim supn kT xnk ≤ ca((Pn 1 (X) with k(xn)nkw i=1 T xi)n). Then there exists n ∈ N such that kPk Let (xn)n ∈ lw Let c > ca((Pn i=1 T xi−Pl i=1 T xik < c for all k, l ≥ n. In particular, we have kT xkk = kPk i=1 T xik < c for all k ≥ n + 1. Thus one can derive that lim supn kT xnk ≤ c. Since c > ca((Pn i=1 T xi)n) is arbitrary, we get lim supn kT xnk ≤ ca((Pn i=1 T xi −Pk−1 i=1 T xi)n). 1(T ). 1 (X) with k(xn)nkw 1 ≤ 1 such that ca((Pn Step 2. uc(T ) ≤ uc6 We can suppose that uc(T ) > 0 and fix an arbitrary 0 < c < uc(T ). Then there exists (xn)n ∈ lw i=1 T xi)n) > c. By induction, we can find two strictly increasing sequences (kn)n, (ln)n, ln < kn of positive integers i=ln+1 xi(n = 1, 2, ...). It is 1 ≤ 1 such that kT znk > c for 1(T ) ≥ c and the proof of Step such that kPkn easy to see that (zn)n belongs to lw 1 (X) with k(zn)nkw all n ∈ N, which yields lim supn kT znk ≥ c. Hence uc6 2 is completed. i=ln+1 T xik > c for all n ∈ N. Let zn = Pkn (cid:3) Combining Theorem 4.1 with Theorem 4.4, we get the promised quantitative ver- sions of the above implications. Theorem 4.5. Let T ∈ L(X, Y ) and 1 ≤ p < ∞. Then uc(T ) ≤ uc6 p(T ) ≤ cc(T ). 5. Quantifying Dunford-Pettis property of order p Let X be a Banach space and let F be the family of all weakly compact subsets of BX ∗. For F ∈ F , define a semi-norm qF on X∗∗ by qF (x∗∗) = sup x∗∈F < x∗∗, x∗ > , x∗∗ ∈ X∗∗. The locally convex topology generated by the family of semi-norms {qF : F ∈ F } is called the Mackey topology, denoted by τ (X∗∗, X∗). The restriction to X of the Mackey topology τ (X∗∗, X∗) is called the Right topology in [23]. This topology is denoted by ρX or simply ρ when X is obvious. UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p21 In this section, we introduce a new locally convex topology. Let X be a Banach space and let 1 ≤ p < ∞. Let Fp be the family of all relatively weakly p-compact subsets of X. For F ∈ Fp, we define a semi-norm qF on X∗ by qF (x∗) = sup x∈F < x∗, x > , x∗ ∈ X∗. The locally convex topology generated by the family of semi-norms {qF : F ∈ Fp} is denoted by ρ∗p when X is obvious. Applying Grothendieck's Completeness Theo- rem([24, p.148]), we obtain that the space (X∗, ρ∗p) is complete. Hence, a bounded subset A of X∗ is relatively ρ∗p-compact if and only if A is totally bounded, equiva- lently, the set AF = {x∗F : x∗ ∈ A} is totally bounded in l∞(F ) for each relatively weakly p-compact subset F ⊂ BX . So, if we set χp m(A) = sup{χ0(AF ) : F ∈ Fp, F ⊂ BX}, then A is relatively ρ∗p-compact if and only if χp m(A) = 0. The following result, which is immediate from [19, Lemma 4.4], implies that an operator T : X → Y is unconditionally p-converging if and only if T ∗BY ∗ is relatively ρ∗p-compact. Theorem 5.1. Let T ∈ L(X, Y ) and 1 ≤ p < ∞. Then 1 2uc4 p(T ). 2 uc4 p(T ) ≤ χp m(T ∗BY ∗) ≤ Let (x∗n)n be a bounded sequence in X∗. We set caFp((x∗n)n) = sup F∈Fp,F⊂BX inf n sup{qF (x∗k − x∗l ) : k, l ≥ n}, and ecaFp((x∗n)n) = inf{caFp((x∗kn)n) : (x∗kn)n is a subsequence of (x∗n)n}. The quantity caFp measures how far the sequence (x∗n)n is from being ρ∗p-Cauchy. In particular, caFp((x∗n)n) = 0 if and only if the sequence (x∗n)n is ρ∗p-Cauchy. The following result contains two topological characterizations of DP Pp. Theorem 5.2. The following are equivalent about a Banach space X and 1 < p < ∞: (1) X has the DP Pp; (2) Every weakly p-summable sequence in X is ρ-null; (3) Every weakly convergent sequence in X∗ is ρ∗p-convergent. Proof. The equivalence of (1) and (2) is essentially Theorem 3.1. The implication (3) ⇒ (1) follows from Theorem 3.2. It remains to prove (1) ⇒ (3). 22 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI Let (x∗n)n be weakly null in X∗. Define an operator T : X → c0 by T x = (< x∗n, x >)n, x ∈ X. Since (x∗n)n is weakly null, T is weakly compact. By (1), we get T is unconditionally p-converging. Let F ∈ Fp. It follows from Theorem 2.3 that T F is relatively norm compact in c0. By the well-known characterization of relatively norm compact subsets of c0, we get lim n→∞ qF (x∗n) = lim n→∞ which implies that (x∗n)n is ρ∗p-null. < x∗n, x > = 0, sup x∈F (cid:3) To quantify the DP Pp, we will need several measures of weak non-compactness. Let A be a bounded subset of a Banach space X. The de Blasi measure of weak non-compactness of A is defined by ω(A) = inf{bd(A, K) : ∅ 6= K ⊂ X is weakly compact }. Then ω(A) = 0 if and only if A is relatively weakly compact. It is easy to verify that ω(A) = inf{ǫ > 0 : there exists a weakly compact subset K of X such that A ⊂ K + ǫBX}. Other commonly used quantities measuring weak non-compactness are: wkX(A) = bd(A w∗ , X), where A w∗ denotes the weak∗ closure of A in X∗∗. wckX(A) = sup{d(clustX ∗∗((xn)n), X) : (xn)n is a sequence in A}, where clustX ∗∗((xn)n) is the set of all weak∗ cluster points in X∗∗ of (xn)n. γ(A) = sup{ limn limm < x∗m, xn > − limm limn < x∗m, xn > : (xn)n is a sequence in A, (x∗m)m is a sequence in BX ∗ and all the involved limits exist}. It follows from [1,Theorem 2.3] that for any bounded subset A of a Banach space X we have wckX(A) ≤ wkX(A) ≤ γ(A) ≤ 2wckX(A), wkX(A) ≤ ω(A). For an operator T , ω(T ), wkY (T ), wckY (T ), γ(T ) denote ω(T BX), wkY (T BX), wckY (T BX) and γ(T BX),respectively. C. Angosto and B. Cascales([1])proved the following in- equality: γ(T ) ≤ γ(T ∗) ≤ 2γ(T ), for any operator T . UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p23 So,putting these inequalities together, we get,for any operator T , (5.1) 1 2 wkY (T ) ≤ wkX ∗(T ∗) ≤ 4wkY (T ). Let X be a Banach space and A be a bounded subset of X∗. For 1 ≤ p < ∞, we set ιp(A) = sup{lim sup n ηp(A) = sup{lim sup n sup x∗∈A sup x∗∈A < x∗, xn > : (xn)n ∈ lw p (X), (xn)n ⊂ BX}, < x∗, xn > : (xn)n ∈ lw p (X), k(xn)nkw p ≤ 1}. These two quantities measure how far A is weakly p-limited. Obviously, ηp(A) = in ιp(A) = 0 if and only if A is weakly p-limited. The following theorem says, particular, that weakly p-limited sets coincide with relatively ρ∗p-compact sets. Its proof is similar to [19, Lemma 5.6]. Theorem 5.3. Let X be a Banach space, 1 ≤ p < ∞ and A be a bounded subset of X∗. Then 1 8 χp m(A) ≤ ιp(A) ≤ χp m(A). In the following theorem, we quantify the DP Pp by using the quantities ω(·), ιp(·), ecaFp(·) and χp m(·). Theorem 5.4. Let X be a Banach space and 1 < p < ∞. The following are equiva- lent: (1) X has the DP Pp; (2) uc1 (3) ιp(A) ≤ ω(A) for every bounded subset A of X∗; (4) ecaFp((x∗n)n) ≤ 2ω((x∗n)n) whenever (x∗n)n is a bounded sequence in X∗; p(T ) ≤ ω(T ∗) for every operator T from X into any Banach space Y ; m(A) ≤ 2ω(A) for every bounded subset A of X∗. (5) χp Proof. (2) ⇒ (1) is obvious. (3) ⇒ (1) and (5) ⇒ (1) follow from Theorem 3.1. (1) ⇒ (2). Let Y be a Banach space and let T ∈ L(X, Y ). Let ǫ > 0 be such that T ∗BY ∗ ⊂ K + ǫBX ∗, K ⊂ X∗ is weakly compact. Let (xn)n ∈ lw p (X) and (xn)n ⊂ BX . Since X has the DP Pp, it follows from Theorem 3.1 that limn→∞ supx∗∈K < x∗, xn > = 0. Let c > 0. Then there exists a positive integer N such that supx∗∈K < x∗, xn > < c for each n ≥ N. For each n ∈ N, pick y∗n ∈ BY ∗ with kT xnk =< y∗n, T xn >. Since T ∗BY ∗ ⊂ K + ǫBX ∗, then, for each n ∈ N, there exists x∗n ∈ K such that 24 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI kT ∗y∗n − x∗nk ≤ ǫ. Then, for n ≥ N, we get kT xnk =< T ∗y∗n, xn > ≤ ǫ + < x∗n, xn > ≤ ǫ + sup x∗∈K < x∗, xn > ≤ ǫ + c. This yields lim supn kT xnk ≤ ǫ+c. Since c > 0 is arbitrary, we obtain lim supn kT xnk ≤ ǫ and hence uc1 p(T ) ≤ ǫ. This proves uc1 p(T ) ≤ ω(T ∗). (1) ⇒ (3). Let (xn)n be a weakly p-summable sequence in BX . Let ǫ > 0 be such that A ⊂ K +ǫBX ∗, K ⊂ X∗ is weakly compact. For each x∗ ∈ A, there exists z∗ ∈ K such that kx∗ − z∗k ≤ ǫ. This yields < x∗, xn > ≤ ǫ + sup x∗∈K < x∗, xn > (n = 1, 2, ...). Since X has the DP Pp, it follows from Theorem 3.1 that limn→∞ supx∗∈K < x∗, xn > = 0. Thus we get lim supn supx∗∈A < x∗, xn > ≤ ǫ, which completes the proof (1) ⇒ (3). (1) ⇒ (4). Let (x∗n)n be a bounded sequence in X∗. Let ǫ > 0 be such that (x∗n)n ⊂ K + ǫBX ∗, K ⊂ X∗ is weakly compact. For each x∗n, there exists z∗n ∈ K such that kx∗n − z∗nk ≤ ǫ. Since K is weakly compact, there exists a weakly convergent subsequence (z∗kn)n of (z∗n)n. By Theorem 5.2, we see that the sequence (z∗kn)n is ρ∗p-convergent and hence caFp((z∗kn)n) = 0. Note that for any F ∈ Fp, F ⊂ BX, we have qF (x∗ki − x∗kj ) ≤ qF (x∗ki − z∗ki) + qF (z∗ki − z∗kj ) + qF (z∗kj − x∗kj ) ≤ 2ǫ + qF (z∗ki − z∗kj ), i, j = 1, 2, ... This yields caFp((x∗kn)n) ≤ 2ǫ + caFp((z∗kn)n) = 2ǫ. (4) ⇒ (1). Let (xn)n ∈ lw Hence, we get ecaFp((x∗n)n) ≤ 2ǫ and then ecaFp((x∗n)n) ≤ 2ω((x∗n)n). p (X) and let (x∗n)n be weakly null in X∗. By (4), we get ecaFp((x∗n)n) = 0. A classical diagonal argument yields a subsequence (x∗kn)n of (x∗n)n which is ρ∗p-Cauchy. By the completeness of the topology ρ∗p, we see that the subsequence (x∗kn)n is ρ∗p-convergent. Since (x∗n)n is weakly null, (x∗kn)n is ρ∗p-null. UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p25 Since (xn)n is weakly p-summable, one has < x∗kn, xkn > ≤ sup i < x∗kn, xi > → 0 (n → ∞). Then Theorem 3.2 gives (1). (1) ⇒ (5). Let c > ω(A). Then there exists a weakly compact subset K of X∗ such that bd(A, K) < c. Since X has the DP Pp, it follows from Theorem 3.1 that χp m(K) = 0. Let ǫ > 0 and L ∈ Fp, L ⊂ BX. Then there exists a finite subset F ⊂ K such that bd(KL, F L) < ǫ, so χ(AL) ≤ c + ǫ. Since ǫ > 0 is arbitrary, we get χ(AL) ≤ c. By (4.1), we get χ0(AL) ≤ 2c. This implies that χp m(A) ≤ 2c, which completes the proof. (cid:3) The following quantitative version obviously strengthens the Dunford-Pettis prop- erty of order p. p(T ) ≤ C · wkX ∗(T ∗) for every operator T from X into Theorem 5.5. Let X be a Banach space and 1 < p < ∞. The following are equiva- lent: (1) There is C > 0 such that uc6 any Banach space Y ; (2) There is C > 0 such that uc6 l∞; (3) There is C > 0 such that ηp(A) ≤ C · wkX ∗(A) for each bounded subset A of X∗; (4) There is C > 0 such that uc6 p(T ) ≤ C · wkY (T ) for every operator T from X into any Banach space Y ; (5) There is C > 0 such that uc6 l∞. p(T ) ≤ C · wkX ∗(T ∗) for every operator T from X into p(T ) ≤ C · wkl∞(T ) for every operator T from X into Proof. The implication (1) ⇒ (2) is trivial with the same constant. (2) ⇒ (3). Assume that there is C > 0 such that uc6 p(T ) ≤ C · wkX ∗(T ∗) for every operator T from X into l∞. We'll show that (3) holds with the constant 32C. Let A be a bounded subset of X∗. We may assume that ηp(A) > 0. Let us fix any 0 < ǫ < ηp(A). By the definition of ηp(A), there exist a sequence (x∗n)n in A and a p ≤ 1 such that < x∗n, xn > > ǫ for each sequence (xn)n in lw n ∈ N. Let us define an operator S : l1 → X∗ by p (X) with k(xn)nkw S((αn)n) = X αnx∗n, n (αn)n ∈ l1. 26 DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI As in the proof of Theorem 5.4 in [17], the set S(Bl1) is contained in the closed absolutely convex hull of (x∗n)n and so wkX ∗(S) ≤ 2wkX ∗((x∗n)n). Let T = S∗JX : X → l∞. By (2) and (5.1), we get uc6 p(T ) ≤ C · wkX ∗(T ∗). Thus ǫ ≤ lim sup < x∗n, xn > ≤ lim sup n kT xnk n ≤ uc6 p(T ) ≤ C · wkX ∗(T ∗) ≤ 4C · wkl∞(T ) ≤ 4C · wkl∞(S∗) ≤ 16C · wkX ∗(S) ≤ 32C · wkX ∗((x∗n)n) ≤ 32C · wkX ∗(A) Since ǫ < ηp(A) is arbitrary, we get the assertion (3). (3) ⇒ (1). Let us suppose that (3) holds with a constant C > 0. Let T ∈ L(X, Y ). p ≤ 1. For each n ∈ N, pick y∗n ∈ BY ∗ so that Let (xn)n ∈ lw kT xnk =< y∗n, T xn >. Applying (3) to A = (T ∗y∗n)n, we get p (X) with k(xn)nkw lim sup kT xnk = lim sup n n ≤ lim sup n sup x∗∈A ≤ C · wkX ∗(A) ≤ C · wkX ∗(T ∗), < T ∗y∗n, xn > < x∗, xn > ≤ ηp(A) which yields uc6 p(T ) ≤ C · wkX ∗(T ∗). Finally, the equivalences of (1) ⇔ (4) and (2) ⇔ (5) follow from estimate (5.1). (cid:3) It should be mentioned that the assertion (3) of Theorem 5.5 is a quantitative version of Theorem 3.1. Definition 5.1. We say that a Banach space X has the quantitative Dunford-Pettis property of order p if X satisfies the equivalent conditions of Theorem 5.5. The following Theorem 5.7 is a quantitative version of Corollary 3.3. To prove it, we need a simple lemma. Lemma 5.6. Let X be a closed subspace of a Banach space Y and let A be a bounded subset of X. Then (5.2) wkY (A) ≤ wkX(A) ≤ 2wkY (A). UNCONDITIONALLY p-CONVERGING OPERATORS AND DUNFORD-PETTIS PROPERTY OF ORDER p27 Proof. We can identify X∗∗ with X⊥⊥ ⊂ Y ∗∗. Under this identification, the weak∗ closure of A in X∗∗ is equal to the weak∗ closure of A in Y ∗∗. This yields the left inequality immediately. To prove the right inequality of (5.2), let us fix any . Then there exists y ∈ Y such that ky∗∗ − yk ≤ c. c > wkY (A). Take any y∗∗ ∈ A Choose y∗ ∈ X⊥ with ky∗k = 1 so that d(y, X) = < y∗, y > . Then we get w∗ d(y∗∗, X) ≤ ky∗∗ − yk + d(y, X) ≤ c + < y∗, y > = c + < y∗, y∗∗ − y > ≤ 2c. Thus wkX(A) ≤ 2c. By the arbitrariness of c > wkY (A), we obtain wkX(A) ≤ 2wkY (A). (cid:3) Theorem 5.7. If X∗∗ has the quantitative Dunford-Pettis property of order p, then so is X. More precisely, (a) If X∗∗ satisfies one of the conditions (1),(2),(4) and (5) of Theorem 5.5 with a given constant C, then X satisfies the respective condition of Theorem 5.5 with 16C; (b) If X∗∗ satisfies the condition (3) of Theorem 5.5 with a given constant C, then X satisfies the respective condition (3) of Theorem 5.5 with C. Proof. The assertion (a) follows immediately from the inequality (5.1) and the easy p(T ∗∗) for each operator T . The assertion (b) is a direct conse- fact that uc6 quence of (5.2). (cid:3) p(T ) ≤ uc6 Acknowledgements. This work is done during the first author's visit to Depart- ment of Mathematics, Texas A&M University. We would like to thank Professor W. B. Johnson for helpful discussions and comments. References [1] [2] [3] [4] [5] [6] [7] C. Angosto and B. Cascales, Measures of weak non-compactness in Bananch spaces, Topol- ogy Appl.156(2009), 1412-1421. F. Albiac and N. J. Kalton, Topics in Banach space theory, Springer, 2005. H. Bendov´a, O. F. K. Kalenda and J. Spurn´y, Quantification of the Banach-Saks property, J.Funct. Anal.268(2015), 1733-1754. F. Bombal, P. Cembranos and J. Mendoza, On the surjective Dunford-Pettis property, Math.Z.204 (1990), 373-380. J. Bourgain and F. Delbaen, A class of special L∞ spaces, Acta Math.145(1980),no.3 -- 4, 155 -- 176. J. M. F. Castillo and F. S´anchez, Remarks on some basic properties of Tsirelson's space, Note di Matematica.XIII(1993), 117-122. J. M. F. Castillo and F. S´anchez, Dunford-Pettis-like properties of continuous vector func- tion spaces, Revista Matematica.6(1993), 43-59. 28 [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] DONGYANG CHEN, J. ALEJANDRO CH ´AVEZ-DOM´INGUEZ, AND LEI LI results related to the Dunford-Pettis property, Con- J. M. F. Castillo and F. S´anchez, Weakly p-compact, p-Banach-Saks and super-reflexive Banach spaces, J.Math.Anal.Appl.185(1994), 256-261. W.J. Davis, T. Figiel, W. B. Johnson, and A. Pe lczy´nski, Factoring weakly compact opera- tors, J. Funct. Anal. 7(1974), 311-327. J. Diestel, A survey of temp.Math.2(Amer.Math. Soc.,Providence,RI,1980), 15-60. J. Diestel, Sequences and Series in Banach Spaces, Springer, New York, 1984. J. Diestel, H. Jarchow and A. Tonge, Absolutely summing operators, Cambridge Studies in Adv. Math., Vol.43, Cambridge Univ. Press, Cambridge, 1995. M. Fabian, P. H´ajek, V. M. Santaluc´ıa and V. Zizler, A quantitative version of Krein's theorem, Rev.Mat.Iberoam.21(2005), 237-248. J. H. Fourie and J. Swart, Banach ideals of p-compact operators, Manuscripta Math.26(1979), 349-362. A. Grothendieck, Sur certaines classes des suites dans les espaces de Banach, et le th´eor´eme de Dvoretzky-Rogers, Bol.Soc.Mat.Sao Paulo.8(1956), 81-110. A. Grothendieck, Sur les applications lin´eaires faiblement compactes d'espaces du type C(K), Canad J.Math.5(1953), 129-173.(French) M. Kacena, O. F. K. Kalenda and J. Spurn´y, Quantitative Dunford-Pettis property, Adv.Math. 234 (2013), 488-527. O. F. K. Kalenda, H. Pfitzner and J. Spurn´y, On quantifications of weak sequential com- pleteness, J.Funct. Anal. 260(2011), 2986-2996. O. F. K. Kalenda and J. Spurn´y, Quantifications of the reciprocal Dunford-Pettis property, Studia Math.210(2012), 261-278. H. Krulisov´a, Quantification of Pe lczy´nski's property (V), to appear. D. Leung, Uniform convergence of operators and Grothendieck spaces with the Dunford- Pettis property, Math.Z.197(1988), 21-32. J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I, Sequence Spaces, Springer, Berlin, 1977. A. M. Peralta, I. Villanueva, J. D. M. Wright and K. Ylinen, Topological characterisation of weakly compact operators, J.Math.Anal.Appl.325(2007), 968-974. H. H. Schafer, Topological vector spaces, New York: Springer 1971. D. P. Sinha and A. K. Karn, Compact operators whose adjoints factor through subspaces of lp, Studia Math.150(2002), 17-53. School of Mathematical Sciences, Xiamen University, Xiamen,361005,China E-mail address: [email protected] Department of Mathematics, University of Oklahoma, Norman, OK 73019-3103,USA E-mail address: [email protected] School of Mathematical Sciences and LPMC, Nankai University, Tianjin, 300071, China E-mail address: [email protected]
1511.08929
1
1511
2015-11-28T21:02:14
On ergodic operator means in Banach spaces
[ "math.FA" ]
We consider a large class of operator means and prove that a number of ergodic theorems, as well as growth estimates known for particular cases, continue to hold in the general context under fairly mild regularity conditions. The methods developed in the paper not only yield a new approach based on a general point of view, but also lead to results that are new, even in the context of the classical Cesaro means.
math.FA
math
ON ERGODIC OPERATOR MEANS IN BANACH SPACES ALEXANDRU ALEMAN AND LAURIAN SUCIU Abstract. We consider a large class of operator means and prove that a number of ergodic theorems, as well as growth estimates known for particular cases, continue to hold in the general context under fairly mild regularity conditions. The methods developed in the paper not only yield a new approach based on a general point of view, but also lead to results that are new, even in the context of the classical Ces`aro means. 1. Introduction Let X and Y be two Banach spaces, and B(X , Y) be the Banach space of all bounded linear operators from X into Y. Consider B(X ) = B(X , X ) as a Banach algebra and I ∈ B(X ) the identity operator on X . The range and the kernel of T ∈ B(X ) will be denoted by R(T ) and N (T ), respectively. Also, σ(T ) and σp(T ) stand for the spectrum and the point spectrum of T , and we denote by X ∗ and T ∗ the dual space of X and the adjoint operator of T . The present paper is concerned with sequences {Tn} in the closed convex hull of the powers of a fixed operator T ∈ B(X ) of the form (1.1) tnjT j, Tn =Xj≥0 tnj ≥ 0 , Xj≥0 tnj = 1, and if tnj > 0 for infinitely many values of j, the above sum converges in norm. We shall denote from now on by κ(T ) the set of such, possibly infinite, convex combinations of powers of T and we refer to sequences in κ(T ). Actually, we formally apply an infinite matrix to the operator sequence {T n}. We let (1.2) and (1.3) X0 = {x ∈ X : kTnxk → 0, n → ∞}, Xc = {x ∈ X : {Tnx} converges in norm}. This general type of operator mean has already been considered in the literature, for example in [C], [Eb], [Kr], and more recently, in [R] and [Sc], but only under the additional assumption that T is power bounded, that is {T n} is bounded in norm. 2010 Mathematics Subject Classification. Primary 47A10, 47A35; Secondary 47B20. Key words and phrases. Cesaro mean, supercyclic operator, Kreiss bounded operator. 1 2 A. ALEMAN AND L. SUCIU Let us notice that in [C] such sequences {Tn} are defined with respect to a regular matrix {tnj} which implies, in particular, that limn→∞ tnj = 0 for j = 0, 1, .... The basic idea behind the results presented in this paper is that a great deal of estimates, or ergodic theorems for specific operator means, continue to hold in a fairly general context than the power boundedness, and they can be deduced from either: 1) A regularity condition for the sequence {Tn} (which does not require the regularity of the matrix {tnj}), or, 2) Conditions invariant to rotations. The regularity condition considered here is that for some fixed, nonnegative integer n0, we have T Tn − Tn+n0 → 0 either strongly, or in the operator norm. It is used in order to exhibit useful isometries related to {Tn} that are acting on quotient spaces of X. This is a method that appears also in the works of [K´e] (concerning the generalized Banach limits), and is a powerful tool in the study of asymptotic behavior (see for instance [Al], [V]). By a rotational invariant condition we mean a growth restriction obeyed not only by the sequence {Tn}, but also by all ”rotated” sequences Tnλ obtained by applying the same convex combinations to the operator λT . More precisely, for Tn as in (1.1) we set (1.4) tnjλjT j, λ = 1. Tnλ =Xj≥0 This type of condition brings complex analysis in the picture, a powerful tool in the study of bounded linear operators. The paper is organized as follows. Section 2 serves to an introductory purpose. We present some general technical tools, including properties of X0, as well as a list of examples of sequences in κ(T ), satisfying the regularity conditions. Moreover, we introduce and discuss the backward iterate sequence {T (−1) } of a sequence {Tn} in κ(T ), which is inspired by the well known relations between Ces´aro means of order p ≥ 1 and the means of order p + 1. The n backward iterate is determined by the identity T (−1) n (T − I) = Xj≥1 jtnj  −1 (Tn − I) , and plays an important role in our considerations. In Section 3 we investigate sequences in κ(T ) for supercyclic operators T and we refer to the Ansari-Bourdon theorem [AB], which asserts that for a power bounded and supercyclic operator T , the powers must converge strongly to zero. The result may fail for general sequences in κ(T ), because of the presence of eigenvalues on the unit circle of the adjoint T ∗. However, we are able to give a fairly complete picture for the subspaces X0 and Xc defined above (Theorem 3.3), under the regularity condition. Our general method based ON ERGODIC OPERATOR MEANS IN BANACH SPACES 3 on the intertwining of T with an isometry gives even in the power boundedness context, an improvement of Ansari-Bourdon’s result, namely the fact that if T is supercyclic and power bounded, then every sequence in κ(T ) which satisfies the regularity condition strongly converges to zero. Concrete examples of supercyclic operators which actually are Ces´aro bounded but not power bounded, are deferred to Section 5. In particular, this answers a question raised by J. Bonet in a recent public communication concerning the existence of a Ces`aro bounded hypercyclic operator. Section 4 is devoted to conditions invariant to rotations. Our first main theorem (Theorem 4.1) is a far-reaching generalization of a result of O. Nevanlinna [N]. More precisely, for operators T ∈ B(X ) with peripheral spectrum σ(T ) ∩ T of arclength measure zero we derive an estimate for the powers T n, under the assumption that for some sequence {Tn} in κ(T ), both sequences {Tn} and {T (−1) } satisfy uniform rotational invariant growth restrictions. To avoid technicalities we shall only mention some applications of this theorem. For example, n when applied to (discretized) Abel means the result yields the following Nevanlinna-type theorem: If u : (1, ∞) → (0, ∞) is decreasing, then such a behavior of the resolvent implies k(T − λI)−1k ≤ u(λ) λ − 1 , λ > 1 , kT nk = o(cid:18)nu(cid:18) n n − 1(cid:19)(cid:19) , n → ∞ . The result applies to other means as well, in particular, to other functions of T . We pay special attention to the Ces`aro means. We prove that certain rotational invariant growth conditions for Ces`aro means of any order p ≥ 2 are actually equivalent to a Kreiss-type condition (Theorem 4.7). When p = 2, a special case of this result was proved by J. Strikwerda and B. Wade in [SW]. Other similar considerations on the Kreiss resolvent condition were obtained by E. Berkson and T. A. Gillespie [BG] for trigonometrically well-bounded operators, and by O. El-Fallah and T. Ransford [FR] for operators satisfying a generalized Kreiss-type condition with respect to a compact subset of the unit circle. In the case p = 1, our conditions are equivalent to uniform estimates for the partial sums of the Taylor expansion at infinity of the resolvent function of T (Theorem 4.7). This is related to the uniform Kreiss boundedness condition, a concept introduced recently in [MSZ]. For such conditions a weaker conclusion (in the sense of strong convergence) holds under an additional technical assumption (Theorem 4.8), and this can be applied to the supercyclic operators, for instance. Section 5 concerns examples which require more work. In addition to the construction of a Ces`aro bounded hypercyclic operators already mentioned above, we have even a Ces`aro bounded 3-isometry (in the sense of [AS]). This operator has the remarkable property that for any x 6= 0, the Ces`aro means of T converge weakly, but not strongly, to zero. This is a 4 A. ALEMAN AND L. SUCIU somewhat surprising fact which is also related to a well-known mean ergodic theorem (see [Kr]). It shows that even if the regularity condition holds, weak convergence of the means fails to imply their strong convergence in a fairly dramatic way. Finally, by modifying an example of A. L. Shields [Sh] we show that the conclusion of the generalized O. Nevanlinna theorem (Theorem 4.1) may fail if the peripheral spectrum of T has positive arclength measure on the unit circle. Let us notice that at this moment we know only an example, which appear also in [Sh], of a uniformly Kreiss bounded operator on a Hilbert space, which is not power bounded. The operator in Subsection 5.1 is another such example which is also hypercyclic. 2. Preliminaries and terminology 2.1. General considerations. Given T ∈ B(X ), we begin with two observations about sequences {Tn} in κ(T ). Proposition 2.1. Assume that the sequence {Tn} in κ(T ) is bounded on R(T − I)m, for some nonnegative integer m. Then X0 is closed and R(T − I)k , X0 ⊂ \k≥0 Proof. Clearly, the closure Y of X0 is invariant for T, Tn and if S denotes the restriction of T to Y, it is easy to see that S −I has dense range. Indeed, if f ∈ R(S −I)⊥, then f (T x) = f (x) for all x ∈ Y, so that from the T -invariance and the fact that Tn ∈ κ(T ), we deduce that f (Tnx) = f (x) , for all x ∈ Y and all n ∈ N. But then f vanishes on the dense subspace X0, i.e. f = 0. Since S is bounded on Y it follows that (S − I)k has dense range for all k ∈ N, and we obtain X0 ⊂ (S − I)kY ⊂ R(T − I)k . In particular, this gives X0 ⊂ R(T − I)m . and by the uniform boundedness principle we conclude that X0 = Y. (cid:3) A direct consequence of this result is that if X0 = R(T − I)k for some nonnegative integer k then (2.1) R(T − I)p = R(T − I)k , p ≥ k . A simple additional condition that determines X0 is given below. Corollary 2.2. Assume that there exist nonnegative integers k, m with k ≥ m, such that the sequence {Tn} in κ(T ) is bounded on R(T − I)m, and that {Tn(T − I)k} converges strongly to zero. Then X0 = R(T − I)k . ON ERGODIC OPERATOR MEANS IN BANACH SPACES 5 Our second observation provides a tool to estimate X0, and it applies to more general sequences {Tn}. We shall use the following notations. Given a continuous seminorm γ on X we denote its kernel by N (γ), and consider the quotient space X/N (γ) as a normed space with k[x]k = γ(x), [x] = x + N (γ) . We denote by Xγ the completion of this space, and by Qγ the quotient map from X into Xγ. Obviously, Qγ ∈ B(X , Xγ) with N (Qγ) = N (γ) , R(Qγ) = Xγ . Proposition 2.3. Assume that {Tn} ⊂ B(X ) is bounded on R(T −I)m, for some nonnegative integer m, and that each Tn commutes with T ∈ B(X ). Let (2.2) γ(x) = lim sup n→∞ kTn(T − I)mxk, x ∈ X . Then there exists a unique operator Vγ ∈ B(Xγ), such that QγT = VγQγ. Moreover, σ(Vγ) ⊂ σ(T ) and if F is analytic in a neighborhood of σ(T ) and satisfies F (Vγ) = 0, then X0 ⊂ R(F (T )(T − I)m). Proof. The statement is almost self-explanatory. The equality QγT = VγQγ defines Vγ on a dense subspace, and since N (Qγ) = N (γ) is invariant for T , it follows that Vγ is bounded on this subspace, hence it has a unique bounded extension to Xγ. In a similar way we obtain bounded linear operators Vλ, λ ∈ C \ σ(T ) with Qγ(T − λI)−1 = VλQγ, which implies that Vλ = (Vγ − λI)−1. Finally, for functions F as in the statement we have QγF (T ) = F (Vγ)Qγ, hence QγF (T ) = 0 whenever F (Vγ) = 0. This also gives the required inclusion, which finishes the proof. (cid:3) Due to the last assertion above, we are of course interested in situations where Vγ is as simple as possible, in particular, when it has a large functional calculus algebra. An immediate application for sequences in κ(T ) is given below. Corollary 2.4. Assume that the sequence {Tn} in κ(T ) is bounded on R(T − I)m, for some nonnegative integer m, and let γ be the seminorm given by (2.2). Then X0 = R(T − I)k for some k ≥ m if and only if (Vγ − I)k−m = 0. Proposition 2.3 allows us to extend to the ergodic context the method developed in [Al], [K´e], [V] and [SZ] in the study of asymptotic properties of scalar weighted powers of an operator in B(X ). 6 A. ALEMAN AND L. SUCIU 2.2. A regularity assumption. Given T ∈ B(X ) we shall consider sequences {Tn} in B(X ) which are bounded on R(T − I)m, and satisfy the condition (2.3) lim n→∞ k(T Tn − Tn+n0)xk = 0 , x ∈ R(T − I)m . for some non-negative integers m, n0. A sequence {Tn} satisfying the condition (2.3) will be shortly called (n0, m)-regular. In particular, we say that {Tn} is regular when n0 = 1 and m = 0, and {Tn} is ergodic (according to [Kr]) if it verifies (2.3) with n0 = m = 0. For different values of m, n0, these conditions of regularity are in general not related. In particular, ergodicity cannot be obtained from the regularity, as we shall see in the following examples (including the one constructed in Subsection 5.2). Some direct consequences of the regularity condition are listed below. Theorem 2.5. Assume that {Tn} ⊂ B(X ) is (n0, m)-regular. Then (i) If {Tn} ⊂ κ(T ) is bounded on R(T − I)m, then X0 = R(T − I)k for some k > m if and only if lim n→∞ k−m Xl=0 (−1)l(cid:18)k − m l (cid:19)Tn+ln0x = 0 , x ∈ R(T − I)m . Moreover, Xc ∩ R(T − I)m is the direct topological sum Xc ∩ R(T − I)m = X0 ⊕ (N (T − I) ∩ R(T − I)m) . (ii) If each Tn commutes with T , {Tn} is bounded on R(T − I)m and γ is the seminorm given by (2.2), then the operator Vγ from Proposition 2.3 is an isometry. Proof. (i) A repeated application of (2.3) gives for x ∈ R(T − I)m k−m (T − I)k−mTnx − Xl=0 (−1)l(cid:18)k − m l (cid:19)Tn+ln0x → 0 , n → ∞ , which proves the first part. Also from (2.3) we see that if x ∈ Xc then lim n→∞ Tnx ∈ N (T − I) , which together with Proposition 2.1 proves that Xc ∩ R(T − I)m is the algebraic direct sum of X0 and N (T − I) ∩ R(T − I)m. If x ∈ X0, y ∈ N (T − I) ∩ R(T − I)m then we can choose n such that Since Tny = y we obtain kTnxk < kx − yk . kyk ≤ kTnxk + kTn(y − x)k < ky − xk(1 + sup n kTnR(T − I)mk) , and (i) follows. ON ERGODIC OPERATOR MEANS IN BANACH SPACES 7 To see (ii) just apply again (2.3) to conclude that γ(T x) = γ(x), and this assures that Vγ (cid:3) is an isometry. One motivation for the condition (2.3) is a result in [Ku] which asserts that for m = 0, n0 = 0, 1, this condition is necessary for the strong convergence of {Tn} to the ergodic projection PT of T , that is PT ∈ B(X ) with N (T − I) = R(PT ) and R(T − I) = N (PT ). More precisely, the result in [Ku] states that Tn → PT strongly if and only if {Tn} satisfies (2.3) with m = 0, n0 = 0, 1 and (2.4) lim k→∞ sup n∈N k 1 k + 1 k Xj=0 Tn+jx − PT xk = 0, x ∈ X . The last condition is called almost convergence and was introduced by G. G. Lorentz in [L]. One can obtain similar characterizations of the strong convergence of {Tn} for m = 0 and n0 > 1, by using (2.3) and an appropriate almost convergence of step n0 > 1. 2.3. Some examples. A large variety of examples of sequences satisfying (2.3) arise from matrix summability methods (see [B]), more precisely, by applying such matrices to the sequence {T n}. The simplest choice of such matrices is the identity which gives the sequence {T n}. It satisfies (2.3) with n0 = 1. The so-called Zweier matrix (see [B]) leads to the sequence {Tn} with (2.5) Tn = 1 2 (T n−1 + T n), n ≥ 1 which also satisfies (2.3) with n0 = 1, that is {Tn} is regular. Remark that {Tn} is ergodic if and only if {T n} converges strongly on R(T 2 − I). The most common examples considered in this area are the Ces`aro means of order p ∈ N of 0 (T ) = I, n (T )defined for n ∈ N by : M (p) an operator T ∈ B(X ), which are the operators M (p) M (0) n (T ) = T n, while for n, p ≥ 1 as (2.6) M (p) n (T ) = = = p (n + 1)...(n + p) p (n + 1)...(n + p) (j + p − 1)! j! M (p−1) j (T ) (n − j + p − 1)! (n − j)! T j n n Xj=0 Xj=0 j n + k )T j p n + p n p−1 Xj=0 (1 − Yk=1 In the case p = 1 we write Mn(T ) := M (1) n (T ) = 1 n + 1 T j n Xj=0 and it is usually called the Ces`aro mean of T . 8 A. ALEMAN AND L. SUCIU If {M (p) n (T )} is bounded or it strongly converges in B(X ), we say that T is p-Ces`aro bounded (ergodic), respectively T is Ces`aro bounded (ergodic) when p = 1. These operator means have a considerably richer algebraic structure. They satisfy the following recurrent identities valid for n, p ≥ 1: (2.7) (2.8) (2.9) M (p) n (T )(T − I) = p n + 1 (M (p−1) n+1 (T ) − I) , T M (p) n (T ) = n + p + 1 n + 1 M (p) n+1(T ) − p n + 1 I , n + p + 1 n + 1 M (p) n+1(T ) − M (p) n (T ) = p n + 1 M (p−1) n+1 (T ) . It is easy to see by the relations (2.7) and (2.8) before that the regularity of {M (p) means the ergodicity of {M (p+1) (by using (2.9)), hence each bounded mean {M (p) n (T )} just n (T )} is ergodic it is also regular n (T )} satisfies (2.3) with m = 0 and n0 = 1. It is interesting to remark that these means satisfy the regularity assumption in the operator (T )}. In addition, if {M (p) n norm when they are bounded on R(T − I)m. Another example of a sequence {Tn} in κ(T ) is given by the binomial means (2.10) Tn = 2−n n Xk=0(cid:18)n k(cid:19)T k = Sn. where S = 1 2 (T + I). Clearly, R(T − I)m = R(S − I)m for m ≥ 0, and {Tn} satisfies (2.3) for some integer n0 ≥ 1 if and only if it is ergodic. For example, as it is proved in [DyS], every contraction on a Hilbert space will satisfy (2.3). Clearly, one may consider also continuous summability methods, for example, the Abel means defined as (2.11) A(T, r) = (1 − r)Xj≥0 rjT j = (1 − r)(I − rT )−1 , 0 < r < 1 , where T ∈ B(X ) has spectral radius at most one. Using the equality (I − rT )(I − ρT )−1 = I + (ρ − r)(I − ρT )−1 , it is easy to verify that the above sums are bounded if and only if the sequence {Tn} with Tn = A(T, 1 − 1 n ) is bounded. Moreover, since A(T,r) 1−r = (1 − rT )−1 we have 1 n − 1 Tn(T − I) = (Tn − I) and it is again easy to see that T Tn − Tn+1 = 1 (n + 1)2 (cid:18) n(n + 3) n − 1 Tn + I(cid:19) . ON ERGODIC OPERATOR MEANS IN BANACH SPACES 9 Hence the sequence {Tn} is ergodic if and only if it is regular, and this equivalently means that 1 n Tn → 0 strongly on X . In this case {Tn} also satisfies (2.3) for any positive integer n0 (with m = 0) even in operator norm if it is bounded. Finally, one can deal with more complicated continuous summability methods as well. If F is analytic in the unit disc D and has positive Taylor coefficients, then for any operator T ∈ B(X ) with spectral radius at most 1, and any strictly increasing sequence {rn} tending to 1, the identity (2.12) Tn = 1 F (rn) F (rnT ) defines a sequence in κ(T ). Of course, the regularity of such sequences is more involved and depends on the Taylor coefficients of F , as well as on the choice of {rn}. 2.4. The backward iterate. Given a sequence {Tn} in κ(T ) with Tn =Pj≥0 tnjT j, where (2.13) jtnj < ∞ tn0 6= 1 , Xj≥1 for all n ∈ N, its backward iterate {T (−1) n } is defined by (2.14) Then T (−1) n (2.15) T (−1) n =Xk≥0 snkT k , where snk = Pj≥k+1 tnj Pj≥1 jtnj . belongs to κ(T ) as well, and satisfies the identity T (−1) n −1 (T − I) = Xj≥1 jtnj  (Tn − I) . It is a simple exercise to verify that if Tn = M (p) n (T ) then by (2.7), for n > 1, we have T (−1) n = M (p+1) n−1 (T ). Another direct computation reveals that for the discretized Abel means introduced before T (−1) n coincides with Tn. Also, if Tn is a sequence of the form (2.12) then T (−1) F (rnz) − F (rn) rnF ′(rn)(z − 1) Gn(z) = n . (2.16) = Gn(T ), where Similarly, for the sequence {Tn} given by (2.5) we have for n ≥ 2, (2.17) T (−1) n = = 2 2n − 1 n−2 Xk=0 2 2n − 1 T k + 1 2 T n−1! (nMn−1(T ) − T n−1). 1 2 Also, it is easy to verify that T (−1) n to zero on R(T 2 − I), and it is regular if and only if 1 case, one can see that if {T (−1) is ergodic if and only if {Mn(T )} converges strongly n → 0 strongly on X . In this In } is ergodic then it is also regular, but not conversely. n T (−1) n 10 A. ALEMAN AND L. SUCIU n (T )}. In fact, {Mn(T )} is ergodic if and only if {T (−1) n n n particular, the ergodicity, or regularity of {T (−1) respectively of {M (2) n T n(T − I)x → 0 for x ∈ X . On the other hand, {M (2) 1 is regular and 1 of ergodicity or regularity of {T (−1) bounded, respectively. } are assured by the ergodicity of {Mn(T )}, } is ergodic and n (T )} is ergodic if and only if {T (−1) } n2 T nx → 0 for x ∈ X . Remark that in these last remarks, the conditions n (T ) are } are superfluous, if the means Mn(T ) or M (2) n Finally, we note that in certain cases the regularity condition can imply ergodicity. Such an example is the sequence {Tn} with Tn = Mn(T )2. We omit the details. 3. Operators from linear dynamics In this section we refer to two classes of operators related to linear dynamics, which have been considered in [AB], [BM], [BC], [GLM], [He] and [K´e]. Recall that an operator T ∈ B(X ) is called supercyclic if there exists x ∈ X such that the set {λT nx : λ ∈ C, n ∈ N} is dense in X . In this case, any such vector x is called a supercyclic vector for T . According to the terminology of [BC] we say that T ∈ B(X ) is (weakly) hypercyclic with support N (N ∈ N \ {0}) if there exists x0 ∈ X such that the set (3.1) {T k1x0 + T k2x0 + ... + T kN x0 : k1, ..., kN ≥ 0} is (weakly) dense in X . Clearly, these concepts only make sense when X is separable. The motivation for considering such conditions is the Ansari-Burdon theorem ([AB], Theorem 2.2) which asserts that any supercyclic power bounded operator T must satisfy the condition kT nxk = 0 , x ∈ X . lim n→∞ Let us start with the weakly hypercyclic case and record the following simple observation. Lemma 3.1. Let Y be a separable Banach space and let V ∈ B(Y) be a contraction. If λV is weakly hypercyclic with support N , for some λ ∈ C with λ ≤ 1, then Y = {0}. Proof. If x0 ∈ Y, the set in (3.1) is contained in the ball centered at the origin and of radius N kx0k (as kV k ≤ 1), hence it cannot be weakly dense in Y, unless Y = {0}. (cid:3) Throughout in what follows, for a bounded sequence {Tn} in κ(T ) we let γ the seminorm from (2.2) with m = 0, and consider the operator Vγ obtained by an application of Proposition 2.3 to this seminorm. As a consequence of Lemma 3.1 we have Corollary 3.2. Assume that λT is weakly hypercyclic with support N , for some λ ∈ C with λ ≤ 1. If the sequence {Tn} in κ(T ) is bounded and the operator Vγ is a contraction then {Tn} converges strongly to zero on X . In particular, this holds when {Tn} satisfies (2.3) with m = 0 and n0 ≥ 0. ON ERGODIC OPERATOR MEANS IN BANACH SPACES 11 Proof. The assumption on λT assures for the operators Vγ and Qγ given by (the proof of) Proposition 2.3 that if x0 is the vector in (3.1), the set {Qγ(λT )k1x0 + Qγ(λT )k2x0 + ... + Qγ(λT )kN x0 : k1, ..., kN ≥ 0} = {λVγ)k1Qγx0 + (λVγ)k2Qγx0 + ... + (λVγ)kN Qγx0 : k1, ..., kN ≥ 0} is weakly dense in Xγ, because Q∗ γ into X ∗. Thus λVγ is hypercyclic with support N , and by the previous lemma we must have Xγ = {0} which means Tnx → 0 for x ∈ X . (cid:3) γ maps X ∗ Let us now turn to supercyclic operators T , where the situation is more subtle. This is due to the fact that for a supercyclic operator T , its adjoint T ∗ may have eigenvalues. A simple consequence of the definition (see also [BM]) shows that such an eigenvalue is unique, and the corresponding eigenspace has dimension one. Due to this fact the Ansari-Bourdon theorem does not extend directly to more general means, in particular Corollary 3.2 does not necessarily hold for supercyclic operators. In fact, in Subsection 5.1 we shall construct a Ces`aro (even uniformly Kreiss) bounded and supercyclic operator T such that both T and T ∗ have the eigenvalue 1, hence {Mn(T )} does not converge strongly to zero. It could also happens that T ∗ has an eigenvalue, but T does not. The following theorem describes the possibilities that can occur in the general context, where the condition (2.3) is essential. Theorem 3.3. Let T ∈ B(X ) be supercyclic and assume that the sequence {Tn} ⊂ κ(T ) is bounded and satisfies condition (2.3) for some n0 ≥ 0. If T ∗ has no eigenvalues on the unit circle then {Tn} converges strongly to zero on X , that is X0 = X . If T ∗ has the (unique) eigenvalue µ ∈ T then : (i) {Tny} converges to zero for some y ∈ X \ R(T − µI) if and only if X0 = X . In this case µ 6= 1. (ii) {Tny} converges to a nonzero limit for some y ∈ X \ R(T − µI) if and only if X0 6= Xc = X . This is further equivalent to the fact that both T and T ∗ have the eigenvalue 1. (iii) {Tny} diverges for some (and hence for all) y ∈ X \ R(T − µI) if and only if X0 = Xc 6= X . Proof. Let γ be the seminorm given by (2.2) with m = 0. From the fact that T is supercyclic it follows immediately that the isometry Vγ on Xγ given by condition (2.3) and Theorem 2.5 is supercyclic, as well. Thus, by the Ansari-Bourdon theorem mentioned before, we have dim Xγ ≤ 1. Clearly, Xγ = {0} means X0 = X and this certainly occurs when T ∗ has not eigenvalues on T, as we see below. Suppose that dim Xγ = 1 . Then there exists µ ∈ T such that Vγ[x] = µ[x] , [x] = x + N (γ), x ∈ X , 12 A. ALEMAN AND L. SUCIU and since N (γ) = X0, we obtain that {Tn(T − µI)} converges strongly to zero, i.e. {Tn} con- verges strongly to zero on R(T − µI). If T ∗ has no eigenvalues on T we have R(T − µI) = X and this is a contradiction which shows that X0 = X . Suppose that T ∗ has the eigenvalue µ ∈ T. As pointed out above, the corresponding eigenspace has dimension one. Thus for every y ∈ X \ R(T − µI) there exists f ∈ X∗, T ∗f = µf 6= 0, such that x − f (x)y ∈ R(T − µI) , x ∈ X . With respect to the direct sum decomposition X = R(T − µI) ⊕ Cy the operator T has the matrix representation T = T R(T − µI) 0 (T − µI)Cy µI2 ! , where I2 denotes the identity on Cy. A simple computation shows that T n = T nR(T − µI) 0 (T n − µnI)Cy µnI2 ! , so that, if and we denote we obtain tnjT j , tnjµj , Tn =Xj≥0 τn(µ) =Xj≥0 Tn = TnR(T − µI) 0 (Tn − τn(µ)I)Cy τn(µ)I2 ! . (i) If we can choose y as above such that Tny → 0, then from f (Tny) = τn(µ)f (y), we see that τn(µ) → 0 and we obtain that {Tn} converges strongly to zero on X . The converse is obvious. Note also that since τn(µ) → 0 and τn(1) = 1, we must have µ 6= 1. (ii) We show the equivalence of the first and the third assertion. If we can choose y such that Tny → y0 6= 0, then by (2.3) we have T y0 = lim n→∞ TnT y = lim n→∞ Tn+n0y = y0 . from the fact that Tn(T − µI)y → 0 we deduce that T y0 = µy0, hence µ = 1 and from above, 1 is an eigenvalue for T . Conversely, if µ = 1 and (T − I)y0 = 0, y0 6= 0, then i.e. y0 ∈ X \ R(T − µI) and {Tny0} converges to y0. Moreover, this assumption implies Tny0 = τn(1)y0 = y0 , Tn = TnR(T − I) 0 0 I2! → 0 0 I2! , 0 ON ERGODIC OPERATOR MEANS IN BANACH SPACES 13 in the strong operator topology, that is X0 6= Xc = X . If the last relation holds it is obvious that we can choose y as in the statement. (iii) follows immediately from the fact that {Tn} converges strongly to zero on R(T − µI) and that this space has codimension one. (cid:3) If one of the alternatives (i) or (ii) occur then {Tn(T − I)} converges strongly to zero, that is {Tn} satisfies the regularity assumption (2.3) with m = n0 = 0. Some interesting examples appear in the study of rotational invariant conditions, like Kreiss, or uniform Kreiss boundedness, and will be discussed in the last section. Within this class of operators the situation becomes more transparent, as the following result shows. Corollary 3.4. Let T ∈ B(X ) be supercyclic and we assume that the sequence {Tn} ⊂ κ(T ) is bounded and ergodic. If T ∗ has no eigenvalues on T then {Tn} converges strongly to zero on X that is X0 = X . If T ∗ has the (unique) eigenvalue µ ∈ T then : (i) X0 = X if and only if µ 6= 1. (ii) X0 6= Xc = X if and only if both T and T ∗ have the eigenvalue 1. (iii) X0 = Xc 6= X if and only if σp(T ∗) \ σp(T ) = {1}. Proof. In view of Theorem 3.3 we only need to prove that if T ∗ has the unique eigenvalue µ ∈ T \{1} then X0 = X . This is however obvious, since in this case R(T − I) = X = X0. (cid:3) One reason why the Ansari-Bourdon theorem holds is a result in [BM] which asserts that if T ∗ has the eigenvalue µ then T R(T − µI) is hypercyclic, hence T cannot be power bounded. However, even in the context of power boundedness our method based on the intertwining of T with Vγ (as in the previous proof) leads to an improvement of the Ansari-Bourdon result. For a supercyclic power bounded operator T it shows that every sequence in κ(T ) which satisfies (2.3) must converge strongly to zero. Despite the problems that occur in the general context considered here, one can prove the following interesting extension of the Ansari-Bourdon result. Theorem 3.5. Let T ∈ B(X ) be supercyclic and assume that the sequence {Tn} ⊂ κ(T ) is bounded and satisfies the regularity assumption (2.3) for some integer n0 ≥ 0. If there exists a supercyclic vector x0 ∈ X satisfying one of the following conditions (i) or (ii) kT nx0k = 0 , inf n∈N kT nx0k < ∞ , sup n∈N then {Tn} converges strongly to zero on X . 14 A. ALEMAN AND L. SUCIU Proof. If (i) holds, there is a subsequence {mj} ⊂ N with T mj x0 → 0 as j → ∞. If Qγ and Vγ are the operators associated to the seminorm γ given by (2.2) with m = 0, from the relation V mj γ Qγx0 = QγT mj x0 → 0 it follows that Qγx0 = 0 (Vγ being an isometry). This gives Qγ = 0, since x0 is a supercyclic vector for T . Hence X0 = N (Qγ) = X which means that Tn → 0 strongly on X . If (ii) holds, we can use (i) to assume, without loss of generality, that α := inf n∈N kT nx0k > 0. Let 0 6= x ∈ X and {λk} ⊂ C \ {0}, {nk} ⊂ N be such that x = limk→∞ λkT nkx0. Then kxk = lim k→∞ λkkT nk x0k ≥ α lim sup k→∞ λk, therefore the sequence {λk} is bounded, and so it contains a convergent subsequence, let’s say λkj → λ, j → ∞. Obviously, one has λ 6= 0 because {T nkj x0} is bounded and limj→∞ λkj T nkj x0 = x 6= 0. Then for m ∈ N we have that kT mxk = lim j→∞ λkj kT mT nkj x0k ≤ cλ. Hence T is power bounded and so T m → 0 strongly. In turn, this gives kTnxk = lim j→∞ kλkj TnT nkj x0k ≤ λ sup m∈N kTmk lim j→∞ kT nkj x0k = 0, and consequently Tnx → 0 for any x ∈ X . (cid:3) 4. Conditions invariant to rotations Given a sequence {Tn} in κ(T ) and λ ∈ C, we denote by Tnλ the sequence obtained by applying the same convex combinations to the operator λT , that is (4.1) Tn =Xj≥0 tnjT j =⇒ Tnλ =Xj≥0 tnjλjT j , where tnj ≥ 0, Pj≥0 tnj = 1. We are interested in the effect of growth restrictions for Tnλ, that are uniform in λ ∈ T. Note that if Tn = T n, then Tnλ = λnTn, λ ∈ T. If Tn = 1 2 (T n−1 + T n), it easy to verify that {Tnλ} are uniformly bounded on the unit circle, if and only if T is power bounded. The situation become much more interesting for other sequences. For example, if {Tn} is the sequence of discretized Abel sums (4.2) Tn = 1 n ∞ Xj=0 (1 − 1 n )jT j , then the uniform boundedness of the sequences {Tnλ}, λ ∈ T, is equivalent to the Kreiss boundedness condition k(T − λI)−1k ≤ C λ − 1 , λ > 1 , for some constant C > 0. ON ERGODIC OPERATOR MEANS IN BANACH SPACES 15 4.1. A generalized Nevanlinna theorem. O. Nevanlinna [N] proved the interesting result that for operators T ∈ B(X ) satisfying the Kreiss boundedness condition, and with σ(T ) ∩ T of arclength measure zero, one has kT nk = o(n), n → ∞ . The aim of the present subsection is to show that this type of estimate holds in a very general context and can be derived from rotational invariant conditions. This requires a different approach which makes use of the backward iterate T (−1) defined in (2.14). n Theorem 4.1. Let {Tn} be a sequence in κ(T ) with Tn = Pj≥0 tnjT j, that satisfies (2.13) and jtnj = ∞ . lim n→∞Xj≥1 If {wn} is an increasing sequence of positive numbers such that kTnλk + kT (−1) nλ k ≤ wn , for all λ ∈ T, and σ(T ) ∩ T has arclength measure zero, then for every strictly increasing sequence of positive integers {mn}, we have Proof. We obviously have for λ ∈ T, kT nkPj≥n+1 tmnj Pj≥1 jtmnj nλ (λT − I) = Xj≥1 T (−1) = o(wmn) , n → ∞ . −1 jtnj  (Tnλ − I) , hence, by assumption we obtain kT (−1) nλ R(λT − I)k = o(wn), n → ∞ . Since R(λT − I) = X a.e. on T with respect to arclength measure, it follows that (4.3) kT (−1) nλ k = o(wn) , a.e. on T. With the notation in (2.14) we also have w−1 mnsmnnT n = w−1 mneite−int dt T (−1) , 2π 0 mnZ 2π mnZ 2π 0 kT (−1) mneitk dt 2π . in the sense of Bochner integral, so that (4.4) w−1 mnsmnnkT nk ≤ w−1 Again by assumption we have that sup t∈[0,2π] mnkT (−1) w−1 mneitk = O(1), n → ∞ , 16 A. ALEMAN AND L. SUCIU and by (4.3) we can apply the dominated convergence theorem in (4.4) to obtain the conclu- sion. (cid:3) Remark 4.2. The reason for considering arbitrary increasing sequences of integers {mn} in Theorem 4.1 is that for a general sequence {Tn} in κ(T ), the n-th power T n may not appear in the n-term Tn. Moreover, different choices of {mn} may lead to different estimates of kT nk. Both aspects occur in the case of Ces`aro means. A very natural class of sequences {Tn} for which Theorem 4.1 applies directly, are those generated by power series with positive coefficients, as defined in (2.12), since their backward iterates are easily determined by (2.16). However, in order to avoid technicalities we shall only consider the particular case of the discretized Abel sums for which we have T (−1) = Tn. As a direct consequence we find the following extended version of Nevanlinna’s theorem. n Corollary 4.3. Let u : (1, ∞) → (0, ∞) be a decreasing function. If T ∈ B(X ) satisfies the condition k(T − λI)−1k ≤ u(λ) λ − 1 , λ > 1 , and σ(T ) ∩ T has arclength measure zero, then kT nk = o(cid:18)nu(cid:18) n n − 1(cid:19)(cid:19) , n → ∞. Proof. We have that the Abel means (4.2) are in κ(T ) and satisfy kTnk ≤ u(cid:18) n n − 1(cid:19) . By the observation preceding the corollary, T (−1) follows by an application of Theorem 4.1. n satisfy the same inequality and the result (cid:3) With a specific choice of the function u above we obtain : Corollary 4.4. If T ∈ B(X ) satisfies for some r ≥ 0 the condition (λ − 1)r+1 λr sup λ>1 k(T − λI)−1k < ∞ , and σ(T ) ∩ T has arclength measure zero, then kT nk = o(nr+1) , n → ∞ . The next corollary provides only spectral conditions (on the resolvent function of T ) in order for {M (r) n (T )} to be uniformly convergent to the ergodic projection, for r ≥ 1. Corollary 4.5. Let T ∈ B(X ). If 1 is a simple pole of the resolvent of T , σ(T ) ∩ T has arclength measure zero and for an integer r ≥ 1 we have (λ − 1)r λr−1 k(T − λI)−1k < ∞, sup λ>1 (λ − 1)k(T − λI)−1k < ∞ , sup λ>1 ON ERGODIC OPERATOR MEANS IN BANACH SPACES 17 then kM (r) n (T ) − PT k → 0, as n → ∞. Proof. This is obtained by using Corollary 4.4, Theorem 6 in [Hi], Theorem 1 [Ed] and the uniform Abel ergodic theorem in [LSS]. (cid:3) We should point out here that the hypothesis that σ(T ) ∩ T has arclength measure zero cannot be relaxed, in general (see [N] and Subsection 5.3 below). 4.2. Ces`aro means and resolvent estimates. As mentioned in Section 2, if Tn = M (p) then T (−1) n−1(T ). Then Theorem 4.1 applies to Ces`aro means in the following way. = M p+1 n n (T ), Corollary 4.6. Let {wn} be an increasing sequence of positive numbers. If T ∈ B(X ) satisfies the condition w−1 n kM (p) n (λT )k < ∞ , sup n∈N λ∈T for some integer p ≥ 1, and σ(T ) ∩ T has arclength measure zero, then kT nk = o(nwpn) , n → ∞ . Proof. By (2.6) we have that sup n∈N λ∈T w−1 n kM (p+1) n (λT )k < ∞ , hence, by the remarks above, we can apply Theorem 4.1 to Tn = M (p) From the second equality in (2.6) we have n (T ) with mn = pn. smnn = Pj≥n+1 tmnj Pj≥1 jtmnj = p + 1 p (cid:19)p pn + 1(cid:18) p − 1 αn , where αn → 1+. The conclusion follows from Theorem 4.1. (cid:3) For the special choice wn = nr, n ≥ 1, with r ≥ 0 fixed, Corollary 4.6 yields the implication (4.5) sup n≥1 λ∈T n−rkM (p) n (λT )k < ∞ =⇒ kT nk = o(nr+1), n → ∞ , whenever p ≥ 1 and σ(T ) ∩ T has arclength measure zero. It turns out that such ”regular” growth restrictions for the Ces`aro means can be reformulated in terms of resolvent estimates. This has been first proved by Strikwerda and Wade [SW] in the case when p = 2 and r = 0. The Ces`aro means of order one are related to the partial sums of the Taylor expansion at infinity of the resolvent function of T . The uniform boundedness of these partial sums in the exterior of the unit disc is called the uniform Kreiss boundedness condition (see [MSZ] and [GH] for the boundedness only on the real line) and is equivalent to the uniform boundedness of {Mn(λT )} on the unit circle. These last two conditions are more restrictive. The result below provides an extension of the theorems we just mentioned. 18 A. ALEMAN AND L. SUCIU Theorem 4.7. Given T ∈ B(X ) we have: (i) If p ≥ 2 and r ≥ 0 then n−rkM (p) n (λT )k < ∞ ⇐⇒ sup λ>1 sup n≥1 λ∈T (λ − 1)r+1 λr k(T − λI)−1k < ∞ , (ii) If p = 1 and r ≥ 0 then sup n≥1 λ∈T n−rkMn(λT )k < ∞ ⇐⇒ sup n∈N λ>1 (λ − 1)r+1 λr < ∞ . λ−k−1T k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=0 p (cid:19)M (p) Xn=0(cid:18)n + p Proof. (=⇒) The argument uses the ideas in [SW]. We start with the identity proved in [SW], formula (6.3), for 0 < ρ < 1 and λ ∈ T, namely ∞ (I − ρλT )−1 = (1 − ρ)p n (λT )ρn . For p ≥ 2 the resolvent estimate follows directly from ∞ Xn=0(cid:18)n + p p (cid:19)nrρn ≤ C(1 − ρ)−p−r−1 . By a straightforward argument we see that the above equality holds for all ρ ∈ D and by comparing the coefficients of ρk we obtain for p = 1, n n−1 ρkλkT k = (1 − ρ) (k + 1)Mk(λT )ρk + (n + 1)Mn(λT )ρn . Xk=0 In this case the desired estimate follows from the inequalities (k + 1)r+1ρk ≤ C(1 − ρ)−r−2 , (1 − ρ)r+1ρn ≤ C ′(n + 1)−r−1 , Xk=0 n−1 Xn=0 for 0 < ρ < 1. λ ∈ T, so that (n + 1)(n + 2) 2 (⇐=) We begin with (i) and assume first that p = 2. Use (2.6) to write for 0 < ρ < 1, (n + 1)(n + 2) 2 (cid:13)(cid:13)(cid:13)(cid:13) M (2) 0 n (λT ) =Z 2π n (λT )(cid:13)(cid:13)(cid:13)(cid:13) M (2) ≤ = (I − ρλeitT )−1 Xj=0 (1 − ρ)r+1 ρ−nZ 2π C 0 C (1 − ρ)r+1(1 − ρ2) (1 − ρ)r+2 ρ−n . n we obtain the estimate for M (2) ≤ C n (n − j + 1)e−ijtρ−j dt 2π , 1 1 − ρe−it2 dt 2π ρ−n If we now choose ρ = 1 − 1 this follows immediately from the first equality in (2.6). n (T ) in (i) for all r ≥ 0. For p > 2, ON ERGODIC OPERATOR MEANS IN BANACH SPACES 19 Let us turn to the case when p = 1. For n ≥ 1, µ ∈ T and λ ∈ C with λ > 1 we have by the Abel summation formula Mn(µT ) = = 1 n + 1 n + 1 λn+1 λn+1µ 1 n Xk=0 Xk=0 n = From the assumption we obtain kMn(µT )k ≤ C and if λ = 1 + 1 n , n 1 n + 1 n 1 n−1 n−1 n + 1 Xk=0 (µT )k = λ−k−1(µT )k + (1 − λ) (λµ)−k−1T k + (1 − λ) Xk=0 Xk=0 Xk=0 λ−k−1T k(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (λ − 1)r+1(n + 1)(cid:18)λn + (λ − 1)r+1 Xk=0 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) λr+1 λr n sup n∈N λ>1 λk+1λ−k−1(µT )k λk+1 λk+1µ k Xj=0 Xj=0 k λ−j−1(µT )j  (λµ)−j−1T j  . < ∞ , 1 − λ(λn − 1) λ − 1 (cid:19) , kMn(µT )k ≤ 2r(2e − 1)Cnr for n ≥ 1 and µ ∈ T, which concludes the proof. (cid:3) In Subsection 5.3 below we construct examples which show that the conditions in part (ii) are stronger than those in part (i) of the theorem. In fact, by a similar argument as in the previous proof one can show that if one (any) of the conditions in (i) is satisfied then (4.6) kMn(λT )k = O(nr log(n)), n → ∞. sup λ∈T On the other hand, the conditions in (ii) do not imply a similar growth condition for the powers T n. Indeed, let us consider T = I − V , where V is the classical Volterra operator (V f )(t) :=Z t 0 f (s)ds, f ∈ Lp[0, 1], 1 ≤ p ≤ ∞, p 6= 2. It is known from [MSZ] that T is uniformly Kreiss bounded and kT nk ∼ n 1 stands for ”comparable”. In particular this gives that { T n log(n) } is unbounded. 4 − 1 2p , where ∼ 4.3. The case p = 1. We want to investigate the growth restrictions (4.7) n−rkMn(λT )k < ∞ , sup n≥1 λ∈T in more detail. The somewhat surprising result we are going to prove below asserts that under these conditions, the conclusion of Theorem 4.1 in the sense of strong convergence only holds under an additional assumption. In particular, our theorem provides a partial answer to Question 3 in [MSZ]. 20 A. ALEMAN AND L. SUCIU Theorem 4.8. Let T ∈ B(X ) satisfying for some integer r ≥ 0 the condition (4.7), and let Qγ be the operator induced by the seminorm γ given by (2.2) for Tn = M (r+1) m = r + 1. If R(Qγ) is closed then (T ) and n kT nxk = o(nr+1) , n → ∞ for x ∈ X . Proof. Note first that a direct application of (2.7) shows that under our hypothesis we have that {n−r−1T n} is bounded, so one can define the seminorm on X γ(x) = lim sup n→∞ n−r−1kT nxk = lim sup n→∞ kTn(T − I)r+1xk, x ∈ X , n (T ). By (4.7) we have 1 where Tn = M (r+1) n Tnx → 0 for x ∈ R(T − I)r, while by (2.8) this means that {Tn} satisfies (2.3) with n0 = 1 and m = r. By applying Theorem 2.5 (ii) and Proposition 2.3 to Tn and γ, one obtains an isometry Vγ on the corresponding quotient space Xγ satisfying VγQγ = QγT , where Qγ is the corresponding quotient map of X into Xγ = QγX . The following observation reveals a remarkable property of these operators. Namely, we have (4.8) Indeed, if α = sup n≥1 n λ∈T(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=0 then for j ≥ 1 we have sn(λT ) = λkT k , < ∞ . λkV k γ Qγ(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=0 n T jsn(λT ) = λ j [(n + j + 1)Mj+n(λT ) − jMj−1(λT )] . Divide both sides by jr+1 to obtain for x ∈ X and n ≥ 1 fixed kQγsn(λT )xk = γ(sn(λT )x) ≤ 2 lim sup m→∞ m−rkMm(λT )kkxk , which gives (4.8). We claim that (4.8) implies Qγ = 0, that is Xγ = {0}, under our assumption. To see this, assume that R(Qγ) is closed and Qγ 6= 0. Since Vγ is a non zero isometry it must have an approximate eigenvalue ζ ∈ T. Then there exist sequences {xj} in Xγ, {lj} in X ∗ γ , with and kxjkγ = 1 , lim j→∞ (Vγ − ζI)xj = 0 , kljk = 1 , lj(xj) > 1 2 . ON ERGODIC OPERATOR MEANS IN BANACH SPACES 21 Now by (4.8) we have for j, n ≥ 1 Z 2π 0 and Parseval’s formula gives Since for fixed n we have lim j→∞ by letting j → ∞ above we obtain 2 dt 2π ≤ α2 , lj(eiktV k γ xj)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) n (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=0 n lj(V k γ xj)2 ≤ α2 . Xk=0 kV k γ xj − ζ kxjk2 = 0 , n Xk=0 lim sup j→∞ n Xk=0 lj(xj)2 ≤ α2 , and we arrive at the inequality n ≤ 4α2 for all positive integers n, which gives a contradiction. Thus Qγ = 0 and consequently, Xγ = {0}, that is kT nxk = o(nr+1) , n → ∞ , for all x ∈ X . This ends the proof. (cid:3) The examples in the next section show that the exponent of n cannot be improved. As an application of this theorem, let us remark that if T is supercyclic and it satisfies (4.7) then Vγ is also supercyclic, hence dim(Xγ) ≤ 1 because Vγ is an isometry. So R(Qγ) = Xγ, and by Theorem 4.8 we have Xγ = {0}. In particular, if r = 0 in (4.7) then one obtains kT nxk = o(n) as n → ∞ for x ∈ X . Thus we infer by Corollary 3.4 the following extension of the Ansari-Bourdon theorem Corollary 4.9. If T ∈ B(X ) is supercyclic and uniformly Kreiss bounded such that 1 /∈ σp(T ∗) \ σp(T ), then for every λ ∈ T the Ces`aro means Mn(λT ) converge strongly on X . Notice that in reflexive spaces, the spectral condition in this corollary is superfluous (by Corollary 3.4), therefore the sequence {Mn(λT )} always converges strongly in this case, the limit being non zero if and only if 1 ∈ σp(T ) ∩ σp(T ∗). This last spectral condition can be satisfied, as we shall see in Subsection 5.1. Thus a version of Lorch’s theorem [Kr, Theorem 2.1.2, p. 73] on one hand, and of Ansari-Bourdon on the other hand, is obtained for supercyclic and uniformly Kreiss bounded operators, in reflexive spaces. A question arises which is suggested by Theorem 4.8. Question. For T ∈ B(X ) uniformly Kreiss bounded does k 1 n T nk converge to zero as n → ∞? 22 A. ALEMAN AND L. SUCIU 5. Examples 5.1. Uniformly Kreiss bounded hypercyclic and supercyclic operators. Consider the weighted Dirichlet spaces Dα, α > −1 consisting of analytic functions f (z) = Pn≥0 fnznin the unit disc D with the property that (5.1) (n + 1)1−αfn2 < ∞ , kf k2 =Xn≥0 Clearly, the monomials en(z) = (n + 1)(1−α)/2zn form an orthonormal basis in this space. Therefore, the operator Mz of multiplication by the independent variable is a forward weighted shift of the form Its adjoint M ∗ z is a hypercyclic backward shift, and for every µ ∈ T the matrix operator . Mzen =(cid:18) n + 2 Tµ = µM ∗ n + 1(cid:19)(1−α)/2 µIC! 0 0 z is supercyclic on Dα ⊕ C, with σp(Tµ) ∩ σp(T ∗ µ ) = {µ} (see [BM, Theorem 1.40]). The following result is well known. For example, it can be deduced from the results in z Dα, α > 0, and hence all corresponding operators [AP]. A direct consequence of it is that M ∗ Tµ are uniformly Kreiss bounded. Proposition 5.1. If α > 0 then the operator Mz is uniformly Kreiss bounded on Dα. Proof. A straightforward computation based on Parseval’s formula yields kf k2 ∼ f (0)2 +ZD f ′(z)2(1 − z2)αdA(z) , f ∈ Dα , where A denotes the area measure. Then for λ = 1 we have Since the functions Mn(λMz)f (ζ) = 1 − (λζ)n+1 (n + 1)(1 − λζ) f (ζ) . Fn(λ, ζ) = 1 − (λζ)n+1 (n + 1)(1 − λζ) are uniformly bounded in D × D and satisfy for some C > 0 and all λ, ζ ∈ D and all positive integers n, we obtain kMn(λMz)f k2 ∼ 1 − λζ C ≤ (cid:12)(cid:12)(cid:12)(cid:12) ∂ ∂ζ Fn(λ, ζ)(cid:12)(cid:12)(cid:12)(cid:12) (n + 1)2 (cid:18)f (0)2 +ZD (n + 1)2 f (0)2 +ZD f (z)2 +ZD 1 − λz2 (1 − z2)αdA . . 1 1 (Mn(λMz)f )′2(1 − z2)αdA(cid:19) f (z)′2(1 − z2)αdA ON ERGODIC OPERATOR MEANS IN BANACH SPACES Thus, the statement will follow if the operator is bounded on Dα. As pointed out before, a proof of this fact can be found in [AP]. f 7→Z z 0 f (ζ)dζ 1 − λζ , 23 (cid:3) This example shows that one can has in view some non-trivial ergodic properties (as in Section 3) for supercyclic operators T which are not power bounded, but with some bounded sequences in κ(T ), as well the means M (p) n (T ), p ≥ 1. 5.2. A Ces`aro bounded 3-isometry. Given a Hilbert space H and T ∈ B(H), we say that T is an N -isometry, N ∈ N, (see [AS]), if it satisfies N (−1)k(cid:18)N Xk=0 k(cid:19)kT kxk2 = 0 , x ∈ H . We shall be concerned with the cases when N = 2, 3. Our aim is to construct a Ces`aro bounded 3-isometry T on the Hilbert space H for which the sequence { 1 n T nx} is bounded below for all x ∈ H \ {0}. In particular, {Mn(T )x} diverges for each f ∈ H \ {0}. Such an example is unknown in literature. In addition, we shall see that for any x the sequence {Mn(T )x} converges weakly to zero. We start with the following simple method of constructing a 3-isometry using a given 2-isometry. Throughout in what follows we shall denote for a polynomial p(z) =Pn≥0 pnzn p(z)2dm(z) , kpk2 2 =Xn≥0 pn2 =ZT where m is the normalized arclength measure on the unit circle and Lp(z) =Xn≥1 pnzn−1 = p(z) − p(0) z . Proposition 5.2. Let T be a 2-isometry on the Hilbert space H. Fix x0 ∈ H \ {0} and let H1 be the completion of the space of analytic polynomials with respect to the norm kpk2 1 = kpk2 2 +Xn≥0 kLnp(T )x0k2 . Then the operator of multiplication by the independent variable Mzp = zp extends to a bounded operator on H1 which is a 3-isometry with the property that kMzpk2 1 − kpk2 1 = kp(T )x0k2 . Proof. A straightforward computation gives kMzpk2 1 = kpk2 2 + kT p(T )x0k2 +Xn≥0 kLnp(T )x0k2 , hence kMzpk2 1 ≤ (1 + kT k2)kp(T )x0k2 24 A. ALEMAN AND L. SUCIU for all polynomials p. Moreover the equality kMzpk2 1 = kpk2 1 + kT p(T )x0k2 follows directly from above. Since T is a 2-isometry, it satisfies for every polynomial p, kT 2p(T )x0k2 − 2kT p(T )x0k2 + kp(T )x0k2 = 0 . This immediately implies that kM 3 z pk2 1 − 3kM 2 z pk2 1 + 3kMzpk2 1 − kMzpk2 1 = 0 , and the result follows. (cid:3) We shall work with a specific choice of the 2-isometry T , namely the weighted shift Mz on the Dirichlet space D0 as defined in the previous subsection. Now let x0(z) = 1 − z, z ∈ D. We then have the following description of the space H1 considered in Proposition 5.2. Lemma 5.3. With the above notations we have kpk2 1 ∼ kpk2 ∗ = kpk2 2 +ZT p′(z)(1 − z)2dm(z) , for all analytic polynomials p Proof. Note that for every polynomial p(z) =Pk pkzk we have k(1 − Mz)Lnpk2 pk − pk−12(k − n + 1) , 1 = pn2 + Xk≥n+1 so that (5.2) k(1 − Mz)pk2 1 = 2kpk2 pk − pk−12(k − n + 1) 2 +Xn≥0 Xk≥n+1 2 +Xk≥1 ∼ kpk2 k2pk − pk−12 . By Parseval’s formula kpk2 ∗ = kpk2 2 +ZT p′(z)(1 − z)2dm(z) = p12 +Xk≥2 kpk − (k − 1)pk−12 and the result follows from the elementary inequalities k2pk − pk−12 − 2kpk − pk−1pk−1 ≤ kpk − (k − 1)pk−12 ≤ 2k2pk − pk−12 + 2pk−12. (cid:3) ON ERGODIC OPERATOR MEANS IN BANACH SPACES 25 An immediate consequence of the lemma is that H1 is continuously contained in the Hardy space H 2, that is kf k2 ≤ Ckf k1 , for some constant C > 0 and all f ∈ H1. With the lemma in hand we can prove the following result. Theorem 5.4. The operator Mz on H1 is Ces`aro bounded and satisfies for every f ∈ H1 kM n z f k1 ≥ k(1 − Mz)f k2rn(n − 1) 2 . Moreover, the sequences {Mn(Mz)f } converges weakly to zero for every f ∈ H1. Proof. For every polynomial p we have that Mn(Mz)p(z) = 1 − zn+1 (n + 1)(1 − z) p(z) , and the inequality 1 − zn+1 ≤ (n + 1)1 − z, z ∈ T easily implies that kMn(Mz)pk2 ∗ ≤ kpk2 2 + 2ZT p′(z)(1 − z)2dm(z) + 2ZT p(z)2dm(z) . By Lemma 5.3 and the remark following it, we obtain that MzH1 is Ces`aro bounded. More- over, it is well known and easy to prove that the norm on D0 satisfies kM n z pk2 − kpk2 = nkpk2 2 hence by Proposition 5.2 we have for f ∈ H1 n kM k z f k2 1 − kM k−1 z f k2 1 kM k−1 z (1 − Mz)f k2 1 (k − 1)k(1 − Mz)f k2 2 kM n z f k2 1 − kf k2 1 = = ≥ = n Xk=1 Xk=1 Xk=1 n n(n − 1) 2 k(1 − Mz)f k2 2 and the required inequality is proved. Now, by using the boundedness of {Mn(Mz)}, the weak convergence to zero of {Mn(Mz)} in B(H1) one reduces to the convergence to zero of the scalar products hMn(Mz)p, qi for all analytical polynomials p, q. But this easily follows from the above expression of Mn(Mz)p and Lemma 5.3. The proof is complete. (cid:3) 26 A. ALEMAN AND L. SUCIU 5.3. Modified Shields examples. This is a simple modification of the example in [Sh] showing that the hypothesis that σ(T ) ∩ T has arclength measure zero cannot be removed form Corollary 4.4. Let r be a nonnegative integer and let Xr be the Banach space of analytic functions f in D with the property that f (r+1) belongs to the Hardy space H 1. The norm on Xr is given by kf kr = r Xj=0 f (j)(0) +ZT f (r+1)dm . Then it is well known (see for example [Du]) that there exists C > 0 such that f (j)dm ≤ Ckf kr and kf k∞ ≤ Ckf kr , ZT for all j ≤ r and all f ∈ Xr. Proposition 5.5. The operator T = MzXr has the following properties: (i) For p ≥ 2 we have that the sequences {n−rM p (ii) {n−rMn(T )} is unbounded, (iii) {n−r−1T n} does not converge strongly to zero. n(λT )} are uniformly bounded in λ ∈ T, Proof. To see (i) we use another standard estimate (see again [Du]), to obtain for λ < 1, k(I − λT )−1f kr . . . . r r+1 f (j)(0) + Xj=0ZT Xj=0 (1 − λ)r+1 +ZT (1 − λ)r+1 + kf k∞ZT (1 − λ)r+1 . kf kr kf kr f (j)(z) 1 − λzj+1 dm(z) kf kr f (z) 1 − λzr+2 dm(z) 1 1 − λzr+2 dm(z) Then the assertion follows by an application of Theorem 4.7. (ii) We have Mn(T )f = Fnf , where Fn(z) = 1 − zn+1 (n + 1)(1 − z) , z ∈ D . We claim that for f (z) = 1, n−rMn(T )f = n−rFn are unbounded in Xr. Using the Leibniz rule and the triangle inequality we obtain n > r + 1 n−rF (r+1) n (z) ≥ ar 1 − z − br n−j 1 − zj+1 , r+1 Xj=1 ON ERGODIC OPERATOR MEANS IN BANACH SPACES 27 with ar, br > 0 independent of n and z. The same standard estimates for integrals mentioned above yield for ρn = 1 − 1 n n−rZ 2π 0 F (r+1) n (ρneit)dt ≥ a′ r log n − b′ r , r, b′ with a′ (iii) For n > r + 1 and f (z) = 1 we have r > 0 independent of n, and the claim together with (ii) follow. nr+1 . kT nf kr = n(n − 1) . . . (n − r) , which gives the desired conclusion. (cid:3) In particular, without additional assumptions on the peripheral spectrum the conclusion of Corollary 4.4 may fail. Acknowledgements. The authors are grateful to the referee for his suggestions and useful comments which improve the original version. The second named author would like to express his gratitude to the University of Lund, Centre for Mathematical Sciences, Sweden for hospitality and excellent working conditions. He was also supported by a Project financed from Lucian Blaga University of Sibiu research grants LBUS-IRG-2015-01. References [AS] J. Agler and M. Stankus, m-Isometric transformations of Hilbert space, I, Integr. Equat. Oper. Th. Vol. 21 (1995), 383-429. [AP] A. Aleman and A-M. Persson, Resolvent estimates and decomposable extensions of generalized Ces`aro operators, J. Funct. Anal. 258 (2010), no. 1, 67-98. [Al] G. R. Allan, Power-bounded elements and radical Banach algebras, Banach Center Publ. 38, Institute of Mathematics, Polish Academy of Sciences, Warszawa, 1997, 9-16. [AB] S. I. Ansari and P. S. Bourdon, Some properties of cyclic operators, Acta Sci. Math. (Szeged) 63 (1997), 195-207. [BC] F. Bayart and G. Costakis, Cyclic operators with finite support, Israel Journal of Mathematics, (2012), 1-37. [BM] F. Bayart and E. Matheron, Dynamics of linear operators, Cambridge University Press, 2009. [BG] E. Berkson and T. A. Gillespie Spectral decompositions, ergodic averages, and the Hilbert transform, Studia Math., 144 (1), 2001, 39-61. [B] J. Boos, Classical and modern methods in summability Assisted by Peter Cass. Oxford Mathematical Monographs. Oxford Science Publications. Oxford University Press, Oxford, 2000. [C] L. W. Cohen, On the Mean Ergodic Theorem, Ann. of Math. Second Series, Vol. 41, No. 3, (1940), 505-509. 28 A. ALEMAN AND L. SUCIU [Du] P.L. Duren, Theory of H p spaces, Pure and Applied Mathematics, Vol. 38 Academic Press, New York-London 1970 [DyS] K. Dykema and H. Schultz, Brown measure and iterates of the Aluthge transform for some operators arising from measurable actions, Trans. Amer. Math. Soc. 361 (2009), 6583-6593. [Eb] W. F. Eberlein, Abstract Ergodic Theorems and Weak Almost Periodic Functions, Trans. Amer. Math. Soc. 67 (1949), 217-240. [Ed] E. Ed-Dari, On the (C, α) uniform ergodic theorem, Studia Math. 156 (1), 2003, 3-13. [FR] O. El-Fallah and T. Ransford, Extremal growth of powers of operators satisfying resolvent conditions of Kreiss-Ritt type, J. Funct. Anal. 196 (2002), 135-154. [GLM] M. Gonzalez, F. Le´on-Saavedra, A. Montes-Rodr´ıguez, Semi-Fredholm theory: hy- percyclic and supercyclic subspaces, Proc. London Math Soc. (3) 81 (2000), no. 1, 169-189. [GH] J. J. Grobler and C. B. Huijsmans, Doubly Abel bounded operators with single spec- trum, Questiones Mathematicae, 18 (1995), 397-406. [He] D. A. Herrero, Limits of hypercyclic and supercyclic operators, J. Funct. Anal., 99 (1991), 179-190. [Hi] E. Hille, Remarks on ergodic theorems, Trans. Amer. Math. Soc. 57 (1945), 246-269. [K´e] L. K´erchy, Operators with regular norm-sequences, Acta Sci. Math. (Szeged) 63 (1997), 571-605. [Kr] U. Krengel, Ergodic Theorems, Walter de Gruyter Studies in Mathematics 6, Berlin - New York, 1985. [Ku] M. K. Kuo, Tauberian conditions for almost convergence, Positivity 13 (2009), 611- 619. [L] G. G. Lorentz, A contribution to the theory of divergent sequences, Acta Math., 80 (1948), 167-190. [LSS], M. Lin, D. Shoikhet and L. Suciu, Remarks on uniform ergodic theorems, Acta Sci Math. (Szeged), 81 (2015), 251-283. [MSZ] A. Montes-Rodr´ıguez, J. S´anchez- ´Alvarez, and J. Zem´anek, Uniform Abel-Kreiss boundedness and the extremal behaviour of the Volterra operator, Proc. London Math. Soc. 91 (2005), 761-788. [N] O. Nevanlinna, Resolvent conditions and powers of operators, Studia Math. 145 (2) (2001), 113-134. [R] A. V. Romanov, Weak convergence of operator means Izvestiya: Mathematics 75:6, 2011, 1165-1183. [Sc] M. Schreiber, Uniform families of ergodic operator nets, Semigroup Forum, Vol. 86, Issue 2, (2013), 321-336. [Sh] A. L. Shields, On Mobius bounded operators, Acta Sci. Math. (Szeged) 40 (1978), 371-374. ON ERGODIC OPERATOR MEANS IN BANACH SPACES 29 [SW] J. C. Strikwerda and B. A. Wade, A survey of the Kreiss matrix theorem for power bounded families of matrices and its extensions, Banach Center Publ. 38, Institute of Math- ematics, Polish Academy of Sciences, Warszawa, 1994, 339-360. [SZ], L. Suciu and J. Zem´anek, Growth conditions and Ces`aro means of higher order, Acta Sci Math. (Szeged), 79 (2013), 545-581. [V] Q. P. Vu, A short proof of the Y. Katznelson’s and L. Tzafriri’s theorem, Proc. Amer. Math. Soc., 115 (1992), 1023-1024. Lund University, Mathematics, Faculty of Sciences, P.O. Box 118 S-221 00, Lund, Sweden E-mail address: [email protected] Department of Mathematics, ”Lucian Blaga” University of Sibiu, Dr. Ion Rat¸iu 5-7, Sibiu, 550012, Romania E-mail address: [email protected]
1506.03260
1
1506
2015-06-10T11:35:36
Estimates for entropy numbers of embedding operators of function spaces on sets with tree-like structure: some limiting cases
[ "math.FA", "math.CA" ]
In this paper we obtain order estimates for entropy numbers of embeddings of weighted Sobolev spaces into weighted Lebesgue spaces and of weighted summation operators on trees. Here we consider some critical conditions on the parameters.
math.FA
math
Estimates for entropy numbers of embedding operators of function spaces on sets with tree-like structure: some limiting cases A.A. Vasil'eva 1 Introduction In [37] order estimates for entropy numbers of the embedding operator of a weighted Sobolev space on a John domain into a weighted Lebesgue space were obtained, as well as estimates for entropy numbers of a two-weighted summation operator on a tree. Here we consider some critical cases. Recall the definition of entropy numbers (see, e.g., [6, 9, 32]). Definition 1. Let X, Y be normed spaces, and let T : X → Y be a linear continuous operator. Entropy numbers of T are defined by ek(T ) = infnε > 0 : ∃y1, . . . , y2k−1 ∈ Y : T (BX ) ⊂ ∪2k−1 i=1 (yi + εBY )o , k ∈ N. Kolmogorov, Tikhomirov, Birman and Solomyak [3, 16, 34] studied properties of ε-entropy (this magnitude is related to entropy numbers of embedding operators). Denote by lm p (1 6 p 6 ∞) the space Rm with norm k(x1, . . . , xm)klm p =(cid:26) (x1p + · · · + xmp)1/p for p < ∞, max{x1, . . . , xm} for p = ∞. Estimates for entropy numbers of the embedding operator of lm q were obtained in the paper of Schutt [33] (see also [9]). Later Edmunds and Netrusov [7], [8] generalized this result for vector-valued sequence spaces (in particular, for sequence spaces with mixed norm). p into lm Haroske, Triebel, Kuhn, Leopold, Sickel and Skrzypczak [10 -- 15, 17 -- 24] studied the problem of estimating entropy numbers of embeddings of weighted sequence spaces or weighted Besov and Triebel -- Lizorkin spaces. Triebel [35] and Mieth [31] studied the problem of estimating entropy numbers of embedding operators of weighted Sobolev spaces on a ball with weights that have singularity at the origin. Lifshits and Linde [28] obtained estimates for entropy numbers of two-weighted Hardy-type operators on a semiaxis (under some conditions on weights). The similar 1 problem for one-weighted Riemann-Liouville operators was considered in the paper of Lomakina and Stepanov [29]. In addition, Lifshits and Linde [25 -- 27] studied the problem of estimating entropy numbers of two-weighted summation operators on a tree. This paper is organized as follows. In §2 we obtain the general result about upper estimates for entropy numbers of embedding operators of function spaces on a set with tree-like structure. The properties of such spaces are almost the same as properties of function spaces defined in [36, 37] (see Assumptions 1, 2, 3), but there are some differences. In particular, here we suppose that (6) holds; this condition cannot be directly derived from known results for weighted Sobolev and Lebesgue spaces. Therefore, first we consider some particular cases of function spaces satisfying Assumptions A -- C or A, B, D (see §3) and we prove that these spaces satisfy Assumptions 1 -- 3. In §4 we obtain order estimates for entropy numbers of embedding operators of weighted Sobolev spaces; to this end, we prove that under given conditions on weights Assumptions A -- C or A, B, D hold. In §5 we obtain order estimates for entropy numbers of two-weighted summation operators on a tree in critical cases. 2 Upper estimates for entropy numbers of embedding operators of function spaces on sets with tree-like structure Let us give some notations. Let (Ω, Σ, mes) be a measure space. We say that sets A, B ⊂ Ω are disjoint if mes(A ∩ B) = 0. Let E, E1, . . . , Em ⊂ Ω be measurable sets, and let m ∈ N ∪ {∞}. We say that {Ei}m i=1 is a partition of E if the sets Ei are pairwise disjoint and mes ((∪m i=1Ei) △ E) = 0. Denote by χE(·) the indicator function of a set E. Let G be a graph containing at most countable number of vertices. We shall denote by V(G) the vertex set of G. Two vertices are called adjacent if there is an edge between them. Let ξi ∈ V(G), 1 6 i 6 n. The sequence (ξ1, . . . , ξn) is called a path if the vertices ξi and ξi+1 are adjacent for any i = 1, . . . , n − 1. If all the vertices ξi are distinct, then such a path is called simple. Let (T , ξ0) be a tree with a distinguished vertex (or a root) ξ0. We introduce a partial order on V(T ) as follows: we say that ξ′ > ξ if there exists a simple path (ξ0, ξ1, . . . , ξn, ξ′) such that ξ = ξk for some k ∈ 0, n. In this case, we set ρT (ξ, ξ′) = ρT (ξ′, ξ) = n + 1 − k. In addition, we denote ρT (ξ, ξ) = 0. If ξ′ > ξ or ξ′ = ξ, then we write ξ′ > ξ. This partial order on T induces a partial order on its subtree. Let G be a disjoint union of trees (Tj, ξj), 1 6 j 6 k. Then the partial order on each tree Tj induces the partial order on G. 2 Given j ∈ Z+, ξ ∈ V(T ), we denote Vj(ξ) := VT j (ξ) := {ξ′ > ξ : ρT (ξ, ξ′) = j}. For ξ ∈ V(T ) we denote by Tξ = (Tξ, ξ) the subtree in T with vertex set {ξ′ ∈ V(T ) : ξ′ > ξ}. Let G be a subgraph in T . Denote by Vmax(G) and Vmin(G) the sets of maximal and minimal vertices in G, respectively. Let W ⊂ V(T ). We say that G ⊂ T is a maximal subgraph on the vertex set W if V(G) = W and any two vertices ξ′, ξ′′ ∈ W adjacent in T are also adjacent in G. Given subgraphs Γ1, Γ2 ⊂ T , we denote by Γ1 ∩ Γ2 the maximal subgraph in T on the vertex set V(Γ1) ∩ V(Γ2). Let P = {Tj}j∈N be a family of subtrees in T such that V(Tj) ∩ V(Tj′) = ∅ for j 6= j′ and ∪j∈NV(Tj) = V(T ). Then {Tj}j∈N is called a partition of the tree T . Let ξj be the minimal vertex of Tj. We say that the tree Ts succeeds the tree Tj (or Tj precedes the tree Ts) if ξj < ξs and {ξ ∈ T : ξj 6 ξ < ξs} ⊂ V(Tj). If Γ ⊂ T is the maximal subgraph on the vertex set W, we set PΓ = {Γ ∩ Tj}j∈N. We consider the function spaces on sets with tree-like structure from [36, 37]. Let (Ω, Σ, mes) be a measure space, let Θ be a countable partition of Ω into measurable subsets, let (A, ξ0) be a tree such that ∃c1 > 1 : card VA 1 (ξ) 6 c1, ξ ∈ V(A), (1) and let F : V(A) → Θ be a bijective mapping. Throughout we consider at most countable partitions into measurable subsets. Let 1 6 p, q < ∞ be arbitrary numbers. We suppose that, for any measurable subset E ⊂ Ω, the following spaces are defined: • the space Xp(E) with seminorm k · kXp(E), • the Banach space Yq(E) with norm k · kYq(E), which all satisfy the following conditions: 1. Xp(Ω) ⊂ Yq(Ω); 2. Xp(E) = {f E : f ∈ Xp(Ω)}, Yq(E) = {f E : f ∈ Yq(Ω)}; 3. if mes E = 0, then dim Yq(E) = dim Xp(E) = 0; 3 4. if E ⊂ Ω, Ej ⊂ Ω (j ∈ N) are measurable subsets, E = ⊔j∈NEj, then kf kXp(E) =(cid:13)(cid:13)(cid:13)(cid:8)kf Ej kXp(Ej )(cid:9)j∈N(cid:13)(cid:13)(cid:13)lp kf kYq(E) =(cid:13)(cid:13)(cid:13)(cid:8)kf Ej kYq(Ej )(cid:9)j∈N(cid:13)(cid:13)(cid:13)lq 5. if E ∈ Σ, f ∈ Yq(Ω), then f · χE ∈ Yq(Ω). , , f ∈ Xp(E), f ∈ Yq(E); (2) (3) Let P(Ω) ⊂ Xp(Ω) be a subspace of finite dimension r0 and let kf kXp(Ω) = 0 for any f ∈ P(Ω). For each measurable subset E ⊂ Ω we write P(E) = {P E : P ∈ P(Ω)}. Let G ⊂ Ω be a measurable subset and let T be a partition of G. We set ST (Ω) = {f : Ω → R : f E ∈ P(E), f Ω\G = 0}. (4) If T is finite, then ST (Ω) ⊂ Yq(Ω) (see property 5). For any finite partition T = {Ej}n j=1 of the set E and for each function f ∈ Yq(Ω) we put kf kp,q,T = n Xj=1 kf Ejkσp,q Yq(Ej )! 1 σp,q (5) with σp,q = min{p, q}. Denote by Yp,q,T (E) the space Yq(E) with the norm k · kp,q,T . Notice that k · kYq(E) 6 k · kp,q,T . For each subtree A′ ⊂ A we set ΩA′ = ∪ξ∈V(A′) F (ξ). In [37] upper estimates for entropy numbers of the embedding operator of the space Xp(Ω) into Yq(Ω) were obtained under some conditions on these spaces (the space Xp(Ω) ∼= Xp(Ω)/P(Ω) will be defined later). Some limiting relations between the parameters were not considered. Here we investigate one of those critical cases. Throughout we assume that 1 < p < q < ∞ and the following conditions hold. Assumption 1. There exist a partition {At,i}t>t0, i∈ Jt of the tree A and a number c > 1 with the following properties. 1. If the tree At′,i′ follows the tree At,i, then t′ = t + 1. 2. For each vertex ξ∗ ∈ V(A) there exists a linear continuous projection Pξ∗ : Yq(Ω) → P(Ω) such that for any function f ∈ Xp(Ω) and for any subtree D ⊂ A rooted at ξ∗ kf − Pξ∗f kq Yq(ΩD) 6 c ∞ Xt=t0 Xi∈Jt,D here 2(1− q p )tkf kq Xp(ΩDt,i ); Jt,D = {i ∈ Jt : V(At,i) ∩ V(D) 6= ∅}, Dt,i = At,i ∩ D. 4 (6) (7) Assumption 2. There exist numbers δ∗ > 0 and c2 > 1 such that for each vertex ξ ∈ V(A) and for any n ∈ N, m ∈ Z+ there exists a partition Tm,n(G) of the set G = F (ξ) with the following properties: 1. card Tm,n(G) 6 c2 · 2mn. 2. For any E ∈ Tm,n(G) there exists a linear continuous operator PE : Yq(Ω) → P(E) such that for any function f ∈ Xp(Ω) kf − PEf kYq(E) 6 (2mn)−δ∗2( 1 q − 1 p)tkf kXp(E), where t > t0 is such that ξ ∈ ∪j∈ Jt V(At,j). 3. For any E ∈ Tm,n(G) card {E′ ∈ Tm±1,n(G) : mes(E ∩ E′) > 0} 6 c2. (8) (9) Assumption 3. There exist numbers γ∗ > 0, c3 > 1 and an absolutely continuous function ψ∗ : (0, ∞) → (0, ∞) such that lim card V(At,i) y→∞ ∗(y) yψ′ ψ∗(y) = 0 and for νt := Pi∈ Jt the following estimate holds: νt 6 c3 · 2γ∗2t ψ∗(22t ) =: c3νt, t > t0. (10) Assumption 1 together with the inequality p < q implies that for any t > t0 and for each vertex ξ∗ ∈ VA t−t0(ξ0) there exists a linear continuous projection Pξ∗ : Yq(Ω) → P(Ω) such that for any function f ∈ Xp(Ω) and for any subtree D ⊂ A rooted at ξ∗ kf − Pξ∗f kYq(ΩD) 6 c · 2( 1 q − 1 p )tkf kXp(ΩD). (11) Hence, Assumptions 1 -- 3 from [36, 37] hold with λ∗ = µ∗ = 1 q and u∗ ≡ 1. In particular, there exist a linear continuous projection P : Yq(Ω) → P(Ω) and a number M > 0 such that for any function f ∈ Xp(Ω) the following estimate holds: p − 1 kf − P f kYq(Ω) 6 Mkf kXp(Ω). As P we take the operator Pξ0 (recall that ξ0 is the root of A). Similarly as in [37] we set Xp(Ω) = {f − P f : f ∈ Xp(Ω)} and denote by I the embedding operator of Xp(Ω) into Yq(Ω). We set Z0 = (p, q, c1, c2, c3, c, δ∗, γ∗, ψ∗). We use the following notations for order inequalities. Let X, Y be sets, and let f1(x, y)) if for f1, f2 : X × Y → R+. We write f1(x, y) . y f2(x, y) (or f2(x, y) & y any y ∈ Y there exists c(y) > 0 such that f1(x, y) 6 c(y)f2(x, y) for any x ∈ X; f1(x, y) ≍ y f2(x, y) and f2(x, y) . y f2(x, y) if f1(x, y) . y f1(x, y). 5 Theorem 1. Suppose that Assumptions 1 -- 3 hold. Then en(I : Xp(Ω) → Yq(Ω)) . Z0 1 q − 1 p . n (12) Similarly as in [36], [37] we introduce some more notation. • ξt,i is the minimal vertex of the tree At,i. • Gt = ∪ξ∈V(Γt) F (ξ) = ∪i∈ JtΩAt,i. • Γt is the maximal subgraph on the vertex set ∪j>tV(Γj), t ∈ N. • { At,i}i∈J t is the set of connected components of the graph Γt. • Ut,i = ∪ξ∈V( At,i) F (ξ). • Ut = ∪i∈J t Ut,i = ∪ξ∈V(Γt) F (ξ). If t > t0, then Vmin(Γt) = Vmin(Γt) = { ξt,i}i∈ Jt (see [36]); hence, we may assume that J t = Jt, t > t0. (13) (14) The set Jt0 is a singleton. Denote {i0} = Jt0. In [37] the operators Qt and Pt,m were defined as follows. Definition of the operator Qt. For each t > t0, i ∈ Jt (14) = J t there exists a linear continuous operator Pt,i : Yq(Ω) → P(Ω) such that for any function f ∈ Xp(Ω) and for any subtree A′ ⊂ A rooted at ξt,i kf − Pt,if kYq(ΩA′ ) (11) . Z0 2−( 1 p − 1 q )tkf kXp(ΩA′ ). (15) As Pt,i we take P ξt,i it follows that Pt0,i0 Xp(Ω) = 0 (see [37]). We set . From the definition of the space Xp(Ω) and of the operator P Qtf (x) = Pt,if (x) = P ξt,i f (x) for x ∈ Ut,i, i ∈ J t, Qtf (x) = 0 for x ∈ Ω\ Ut, Tt = {Ut+1,i}i∈J t+1. 6 (16) (17) Since p < q, we have for any f ∈ B Xp(Ω) kf − Qtf kYq( Ut) 6 kf − Qtf kYp,q,Tt−1 ( Ut) (15) . Z0 2−( 1 p − 1 q )t. Notice that if t < t0, then Qtf = Qt+1f = 0 (since Pt,i0 = 0 for t < t0). Throughout we set log x := log2 x. Definition of the operators Pt,m. For t > t0 we set mt = ⌈log νt⌉. (18) (19) In [36] for each m ∈ Z+ the set Gm,t ⊂ Gt, the partition Tt,m of the set Gm,t and the linear continuous operator we constructed. Here the following properties hold: Pt,m : Yq(Ω) → S Tt,m(Ω) 1. Gm,t ⊂ Gm+1,t, Gmt,t = Gt; 2. for any m ∈ Z+ card Tt,m . Z0 2m; 3. for any function f ∈ Xp(Ω) and for any set E ∈ Tt,m kf − Pt,mf kYq(E) . Z0 2−( 1 p − 1 q )tkf kXp(E), m 6 mt, kf − Pt,mf kYq(E) . Z0 4. for any set E ∈ Tt,m 2−( 1 p − 1 q )t · 2−δ∗(m−mt)kf kXp(E), m > mt; card {E′ ∈ Tt,m±1 : mes(E ∩ E′) > 0} . Z0 1. Moreover, we may assume that Tmt,t = { F (ξ)}ξ∈V(Γt), Pt,mtf F (ξ) = Pξf F (ξ), ξ ∈ V(Γt) (it follows from the construction in [36, p. 37 -- 40]). 7 (20) (21) (22) (23) (24) (25) (26) Let Then t∗(n) = min{t ∈ N : ν t > n}, t∗∗(n) = min{t ∈ N : ν t > 2n}. 2t∗(n) ≍ Z0 log n, 2t∗∗(n) ≍ Z0 n (see [37, formula (49)]). For any f ∈ Xp(Ω) the following equality holds: t∗(n)−1 f = (Qt+1f − Qtf )χ Ut+1+ Pt=t0 t∗(n)−1 Pt=t0 ∞ Pm=0 + (Pt,m+1f − Pt,mf )χGm,t + (f − Qt∗(n)f )χ Ut∗(n) (it can be proved similarly as formula (82) in [36]). We set Qf Gt = Pt,mtf Gt, t > t0, Qn,mf Gt = Pt,mt+mf Gt Qn,mf Gt = 0 for for t∗(n) 6 t < t∗∗(n), m ∈ Z+, t < t∗(n) or t > t∗∗(n). (22),(25) q )tkf kXp( F (ξ)) for any ξ ∈ V(Γt) and the Since kf − Pt,mtf kYq( F (ξ)) space Yq(Ω) is Banach, we get from the inequality p < q that for any f ∈ Xp(Ω) the inclusion Qf ∈ Yq(Ω) holds and 2−( 1 p − 1 . Z0 kf − Qf kYq( Ut∗∗(n)) . Z0 Further, 2( 1 q − 1 p )t∗∗(n)kf kXp(Ω) (29) . Z0 n 1 q − 1 p kf kXp(Ω). (27) (28) (29) (30) (31) (32) (33) (34) (f − Qt∗(n)f )χ Ut∗(n) + = ( Qf − Qt∗(n)f )χ Ut∗(n) + ( Qn,m+1f − Qn,mf ) + (f − Qf )χ Ut∗∗(n) ∞ ; Pm=0 indeed, from (23), (31) and (32) it follows that Qf )χ Ut∗(n)\ Ut∗∗(n) . 8 ( Qn,m+1f − Qn,mf ) = (f − ∞ Pm=0 Lemma 1. There exists a sequence {kt}t∗(n)−1 t=t0 ⊂ N such that t∗(n)−1 Xt=t0 (kt − 1) . Z0 n, t∗(n)−1 Xt=t0 ekt(Qt+1 − Qt : Xp(Ω) → Yq( Ut+1)) . Z0 1 q − 1 p . n (35) (36) Lemma 2. There exists a sequence {kt,m}t06t<t∗(n),m∈Z+ ⊂ N such thatPt,m n, (kt,m−1) . Z0 t∗(n)−1 ∞ Xt=t0 Xm=0 ekt,m(Pt,m+1 − Pt,m : Xp(Ω) → Yq(Gm,t)) . Z0 1 q − 1 p . n Lemmas 1 and 2 are proved similarly as Lemmas 6 and 7 in [37]. Lemma 3. We have en(( Q − Qt∗(n))χ Ut∗(n) : Xp(Ω) → Yq(Ω)) . Z0 1 q − 1 p . n Notice that for t∗∗(n) 6 t0 this estimate follows from (33). Hence, throughout we assume that t∗∗(n) > t0. We need some auxiliary assertions. For any ν ∈ N we denote by Iν the identity operator on Rν. Theorem A. [9, 33]. Let 1 6 p 6 q 6 ∞. Then ek(Iν : lν p → lν q ) ≍ p,q 1, 1 6 k 6 log ν, k ) (cid:18) log(1+ ν 2− k ν ν k q − 1 p , 1 (cid:19) 1 p − 1 q , log ν 6 k 6 ν, ν 6 k.   The following properties of entropy numbers are well-known (see, e.g., [9, 32]): 1. if T : X → Y , S : Y → Z are linear continuous operators, then ek+l−1(ST ) 6 ek(S)el(T ); 2. if T, S : X → Y are linear continuous operators, then ek+l−1(S + T ) 6 ek(S) + el(T ). In particular, from the first property it follows that ek(ST ) 6 kSkek(T ), ek(ST ) 6 kT kek(S). (37) (38) 9 Theorem B. [25]. Let X, Y be normed spaces, and let V ∈ L(X, Y ), {Vν}ν∈N ⊂ L(X, Y ). Then for any n ∈ N en+[log2 N ]+1(V ) 6 sup ν∈N en(Vν) + sup x∈BX inf ν∈N kV x − VνxkY . Lemma 4. [39]. Let (T , ξ∗) be a tree with finite vertex set, let card V1(ξ) 6 k for any vertex ξ ∈ V(T ), (39) and let the mapping Φ : 2V(T ) → R+ satisfy the following conditions: Φ(V1 ∪ V2) > Φ(V1) + Φ(V2), V1, V2 ⊂ V(T ), V1 ∩ V2 = ∅, (40) Φ(V(T )) > 0. Then there is a number C(k) > 0 such that for any n ∈ N there exists a partition Sn of the tree T into at most C(k)n subtrees Tj, which satisfies the following conditions: 1. Φ(V(Tj)) 6 (k+2)Φ(V(T )) n for any j such that card V(Tj) > 2; 2. if m 6 2n, then each element of Sn intersects with at most C(k) elements of Sm. Lemma 5. [36]. Let T be a finite partition of a measurable subset G ⊂ Ω, ν = dim ST (Ω) (see (4)). Then there exists a linear isomorphism A : ST (Ω) → Rν such that kAkYp,q,T (G)→lν 1, kA−1klν 1. σp,q . σp,q, r0 q →Yq(G) . q, r0 Lemma 6. [44, formula (60)]. Let Λ∗ : (0, +∞) → (0, +∞) be an absolutely continuous function such that yΛ′ Λ∗(y) = 0. Then for any ε > 0 ∗(y) lim y→+∞ t−ε . ε,Λ∗ Λ∗(ty) Λ∗(y) tε, . ε,Λ∗ 1 6 y < ∞, 1 6 t < ∞. (41) Given t∗(n) 6 t 6 t∗∗(n), i ∈ Jt, we denote At,i =(cid:26) At,i for At∗∗(n),i t < t∗∗(n), for t = t∗∗(n), Γt = Γt for t < t∗∗(n), Γt∗∗(n) = Γt∗∗(n). Let D be a subtree in Γt∗(n). We set Jt,D = {i ∈ Jt : V(At,i) ∩ V(D) 6= ∅}, Dt,i = At,i ∩ D. (42) Throughout we take as ε = ε(Z0) > 0 a sufficiently small number (it will be chosen later by Z0). 10 Proof of Lemma 3. We set t′ ∗(n) = max{t∗(n), t0}. Step 1. Given t < t∗∗(n), we denote by At the subtree in A with vertex set V(A)\V(Γt+1). 2V( At) → R+ by Let f ∈ B Xp(Ω). For each t′ ∗(n) 6 t < t∗∗(n) we define the mapping Φf,t : Φf,t(W) = Xξ∈W∩V(Γt) kf kp Xp( F (ξ)) . Then for any disjoint sets W1, W2 we have Φf,t(W1 ⊔ W2) = Φf,t(W1) + Φf,t(W2). For each t′ ∗(n) 6 t < t∗∗(n) we set kf kp Xp( F (ξ)) ; εt = Pξ∈V(Γt) if nt = ⌈n · 2−tεt⌉ εt > 0; nt = 1 if εt = 0. Then t∗∗(n)−1 Xt=t′ ∗(n) εt 6 1, t∗∗(n)−1 Xt=t′ ∗(n) 2tnt 6 t∗∗(n)−1 Xt=t′ ∗(n) nεt + t∗∗(n)−1 Xt=t′ ∗(n) (29),(45) 6 2t C(Z0)n with C(Z0) ∈ N. By (44), Φf,t(V( At)) = εt. (43) (44) (45) (46) (47) Let f 6= 0. It follows from Lemma 4 and (1) that there exists a number C(Z0) ∈ N and a family of partitions {Tf,t,l}06l6log nt of the tree At, which satisfy the following conditions: card Tf,t,l 6 C(Z0)2−lnt, card {A′′ ∈ Tf,t,l±1 : V(A′) ∩ V(A′′) 6= ∅} . Z0 1, A′ ∈ Tf,t,l, and for any subtree A′ ∈ Tf,t,l such that card V(A′) > 2 Φf,t(V(A′)) (47) 6 C(Z0)2ln−1 t εt. 11 (48) (49) (50) Moreover, we may assume that For f ≡ 0 we set Tf,t,l = { At}. Given t < t∗∗(n), we denote Tf,t,⌊log nt⌋ = { At}. Wt =nξ ∈ V(Γt) : {ξ} ∈ Tf,t,0, kf kp Xp( F (ξ)) > C(Z0)n−1 t εto , Wt = {ξ ∈ Wt : VA 1 (ξ) ∩ { ξt+1,j}j∈ Jt+1 6= ∅}, St = {ξ ∈ V(Γt) : ∃D ∈ Tf,t,0 : ξ ∈ Vmin(D)}, St = St ∪(cid:16)∪ξ∈Wt−1[VA St∗∗(n) = ∪ξ∈W 1 (ξ) ∩ V(Γt)](cid:17) for [VA 1 (ξ) ∩ V(Γt∗∗(n))], t∗∗(n)−1 t < t∗∗(n), S =(cid:16)∪t∗∗(n) t=t′ ∗(n) Then St(cid:17) ∪ {ξt′ ∗(n),j}j∈ Jt′ . ∗(n) S\ ∪t∗∗(n)−1 ∗(n){ ξt,j : j ∈ Jt}, Wt ⊂ St. For each vertex ξ ∈ S we denote by D(ξ) the tree with vertex set ∗(n) St ⊂ ∪t∗∗(n) t=t′ t=t′ V(D(ξ)) = {ξ′ > ξ : [ξ, ξ′] ∩ S = {ξ}}. We set Tf = {D(ξ) : ξ ∈ S}. (51) (52) (53) (54) (55) (56) (57) (58) (59) Then Tf is a partition of the graph Γt′ ∗(n) into subtrees. Step 2. We say that D ∈ Tf if D ∈ Tf and for any ξ ∈ Wt we have D 6= {ξ}. Let (D, ξ∗) ∈ Tf . Then ξ∗ ∈ V(Γt′ ∗(n)). From Assumption 1 and the inequality p < q it follows that kf − Pξ∗f kq Yq(ΩD) (6),(42) . Z0 t∗∗(n) Xt=t′ ∗(n) 2t(1− q p ) Xj∈ Jt,D kf kq Xp(ΩDt,j ), kf − Qf kq Yq(ΩD) (31) = kf − Pt,mtf kq Yq(ΩDt,j (22) . Z0 ) t∗∗(n) Xt=t′ ∗(n) Xj∈ Jt,D 12 . t∗∗(n) Xt=t′ ∗(n) 2t(1− q p ) Xj∈ Jt,D kf kq Xp(ΩDt,j ). Hence, X(D, ξ∗)∈ Tf kf − Pξ∗f kq Yq(ΩD) . Z0 t∗∗(n) Xt=t′ ∗(n) 2t(1− q p ) X(D, ξ∗)∈ Tf Xj∈ Jt,D kf kq Xp(ΩDt,j ), (60) X(D, ξ∗)∈ Tf kf − Qf kq Yq(ΩD) . Z0 t∗∗(n) Xt=t′ ∗(n) 2t(1− q p ) X(D, ξ∗)∈ Tf Xj∈ Jt,D kf kq Xp(ΩDt,j ). (61) Denote by Tf,t the family of trees D ∈ Tf,t,0 such that Xξ∈V( D)∩V(Γt) kf kp Xp( F (ξ)) 6 C(Z0)n−1 t εt. (62) Assertion 1. Let D ∈ Tf , t′ T ∈ Tf,t such that Dt,j = Tt,j. ∗(n) 6 t < t∗∗(n), j ∈ Jt,D. Then there exists a tree Proof of Assertion 1. Let D = D(ξ∗), ξ∗ ∈ S. Since j ∈ Jt,D, the vertices ξ∗ and ξt,j are comparable. There exists a tree T ∈ Tf,t,0 such that max{ξ∗, ξt,j} ∈ V(T ). (63) We claim that V(Dt,j) ⊂ V(T ). Indeed, let ξ ∈ V(Dt,j)\V(T ). Then ξ > max{ ξt,j, ξ∗}. We set η∗ = min{η ∈ [max{ξ∗, ξt,j}, ξ] : η /∈ V(T )} > ξ∗. Then η∗ ∈ V(Γt) and there exists a tree T ′ ∈ Tf,t,0 such that η∗ is the minimal vertex of T ′. Hence, η contradiction. (54) ∈ St (55),(56) ⊂ S; therefore, ξ (58) /∈ V(D), which leads to a Thus, V(Tt,j) ⊃ V(Dt,j) 6= ∅. Let us show that V(Tt,j) ⊂ V(Dt,j). Denote by ξ the minimal vertex of the tree T . Since V(Tt,j) 6= ∅, the vertices ξ and ξt,j are comparable. Let us show that max{ξ∗, ξt,j} = max{ ξ, ξt,j}. Indeed, by (63) we have max{ξ∗, ξt,j} > max{ ξ, ξt,j}. If the inequality is strict, then ξ∗ > ξt,j. In addition, ξ∗ ∈ V(At,j) ∩ S, and by (54) and (57) we get ξ∗ ∈ St; i.e., ξ∗ is the minimal vertex of some tree from the partition Tf,t,0. By (63), ξ∗= ξ. Thus, max{ ξ, ξt,j} = max{ξ∗, ξt,j} ∈ V(Dt,j). (64) 13 Suppose that there exists a vertex ξ ∈ V(Tt,j)\V(D). We set η∗∗ = min(cid:16)hmax{ ξ, ξt,j}, ξi \V(D)(cid:17) . (65) Then η∗∗ > max{ ξ, ξt,j} and η∗∗ ∈ V(At,j). Hence, η∗∗ /∈ ∪t∗∗(n) i ∈ Jt′}. t′=t′ Further, η∗∗ ∈ V(Tt,j); therefore, η∗∗ /∈ ∪t∗∗(n) ∗(n)St′ by (54). From (57) it follows that t′=t′ η∗∗ /∈ S. This together with (58), (64), (65) yields that [ξ∗, η∗∗] ∩ S = {ξ∗}; applying (58) once again, we get that η∗∗ ∈ V(D), which leads to a contradiction. ∗(n){ ξt′,i : It remains to check that T ∈ Tf,t. If T /∈ Tf,t, then V(T ) (50),(62) = {ξ}, kf kXp( F (ξ)) > C(Z0)n−1 t εt; i.e., ξ (52) ∈ Wt. Then either ξ ∈ Wt (in this case, VA 1 (ξ) that V(D) (54),(55) St ∪ St+1) or VA ⊂ (58) = {ξ} and D /∈ Tf . This completes the proof of Assertion 1. ⋄ 1 (ξ) 1 (ξ) (54) ⊂ St. Hence, VA (56) ⊂ S. This implies By (54), the first inclusion of (57), (58) and (59), for any tree D ∈ Tf we have Dt∗∗(n),j = At∗∗(n),j, j ∈ Jt∗∗(n). This together with Assertion 1 yields that t∗∗(n) Xt=t′ ∗(n) 2t(1− q p ) X(D, ξ∗)∈ Tf Xj∈ Jt,D kf kq Xp(ΩDt,j ) 6 t∗∗(n)−1 Xt=t′ ∗(n) 2t(1− q p ) XT ∈ Tf,t Xj∈Jt,T kf kq Xp(ΩTt,j )+ +2t∗∗(n)(1− q + . n1− q p + i.e., t∗∗(n)−1 ∗(n) t∗∗(n)−1 Xt=t′ Xt=t′ ∗(n) 2t(1− q 2t(1− q (29) . Z0 ) n1− q p + Xp(Ω At∗∗(n),j kf kq p ) Xj∈ Jt∗∗(n)  p ) XT ∈ Tf,t  Xξ∈V(T )∩V(Γt) p ) XT ∈ Tf,t (48) . Z0 − q p t n t ε q p kf kp Xp( F (ξ))  t∗∗(n)−1 Xt=t′ ∗(n) q p (62) . Z0 n1− q p + 2t(1− q p )ε q p 1− q p t n t ; t∗∗(n) Xt=t′ ∗(n) 2t(1− q p ) X(D, ξ∗)∈ Tf Xj∈ Jt,D kf kq Xp(ΩDt,j ). Z0 n1− q p + 2t(1− q p )ε t∗∗(n)−1 Xt=t′ ∗(n) If n · 2−tεt > 1, then nt (44) ≍ n · 2−tεt; hence, q p t n 1− q p t =: A. (66) 2t(1− q p )ε q p t n 1− q p t ≍ p,q 2t(1− q p )ε q p t n1− q p 2−t(1− q p )ε 1− q t = n1− q p p εt. 14 If n · 2−tεt < 1, then nt = 1. From (45) it follows that εt 6 1. Therefore, n1− q p + A . Z0 t∗∗(n)−1 Xt=t′ ∗(n) n1− q p εt + t∗∗(n)−1 Xt=t′ ∗(n) p ) 2t(1− q (45) . Z0 From (60), (61), (66), (67) we get that n1− q p ) p + 2t∗∗(n)(1− q (29) . Z0 n1− q p . (67) X(D, ξ∗)∈ Tf this yields (kf − Pξ∗f kq Yq(ΩD) + kf − Qf kq Yq(ΩD)) . Z0 n1− q p ; X(D, ξ∗)∈ Tf k Qf − Pξ∗f kq Yq(ΩD) . Z0 n1− q p . (68) Step 3. Let T be a partition of Γt′ ∗(n) into subtrees. For each (D, ξ∗) ∈ T we set In addition, we put Then PTf ΩD = Pξ∗f − Qt′ ∗(n)f. PTf Ω\ Ut′ ∗(n) = 0. (69) (70) k Qf − Qt′ ∗(n)f − PTf kq Yq( Ut′ ∗(n)) k Qf − Pξ∗f kq Yq(ΩD) (26),(31) = = X(D, ξ∗)∈T = X(D, ξ∗)∈T, card V(D)>2 k Qf − Pξ∗f kq Yq(ΩD). In particular, from (68) and the definition of Tf (see the beginning of Step 2) it follows that k Qf − Qt′ ∗(n)f − PTf f kq Yq( Ut′ ∗(n)) n1− q p . . Z0 (71) Step 4. Let us define the family N of partitions of the graph Γt′ ∗(n) into subtrees. Each of these partitions is constructed as follows. Let C(Z0), C(Z0) be as defined at Step 1. 1. We choose the sequence {nt}t∗∗(n)−1 ∗(n) ⊂ N such that t=t′ t∗∗(n)−1 Xt=t′ ∗(n) 2tnt 6 C(Z0)n. (72) 15 2. For each t ∈ {t′ ∗(n), . . . , t∗∗(n) − 1} we take C(Z0)nt vertices in V(Γt) (some of them may coincide); we denote this set by Ut. 3. For each t ∈ {t′ ∗(n), . . . , t∗∗(n) − 1} we choose an arbitrary subset Vt of {ξ ∈ Ut : VA 1 (ξ) ∩ V(Γt+1) 6= ∅}. 4. Let U = ∪j∈ Jt′ ∗(n) { ξt′ ∗(n),j} ∪(cid:16)∪t∗∗(n)−1 ∗(n) t=t′ Ut(cid:17) ∪(cid:16)∪t∗∗(n)−1 t=t′ ∗(n) ∪ξ∈ Vt [VA This vertex set generates the desired partition T of the graph Γt′ subtrees: 1 (ξ) ∩ V(Γt+1)](cid:17) . ∗(n) into T = {D′ (ξ) : ξ ∈ U}, V(D′ (ξ)) = {ξ′ > ξ : [ξ, ξ′] ∩ U = {ξ}}. (73) Let us estimate the value N . 1. First we estimate the number of choices of {nt}t∗∗(n)−1 ∗(n) (we denote this value by N1). Let 1 6 l 6 C(Z0)n. The number of choices of nt ∈ N such that t∗∗(n)−1 t=t′ 2tnt = l can be estimated from above by the number of choices of numbers nt ∈ N such that nt = l. The last magnitude can be estimated from above by the number of partitions of {1, . . . , l} into t∗∗(n) − t′ intervals. This value does not exceed (l + t∗∗(n) − t′ ∗(n) ∗(n)−1. Hence, ∗(n))t∗∗(n)−t′ t∗∗(n)−1 Pt=t′ ∗(n) Pt=t′ ∗(n) N1 6 X16l6 C(Z0)n Xk=1 C(Z0)n+t∗∗(n) 6 (l + t∗∗(n) − t′ ∗(n))t∗∗(n)−t′ ∗(n)−1 6 kt∗∗(n)−1 . Z0 ( C(Z0)n + t∗∗(n))t∗∗(n) =: N ′ 1. 2. Given the sequence {nt}t∗∗(n)−1 ∗(n) , we estimate the number of choices of a set t=t′ ∪t∗∗(n)−1 ∗(n) t=t′ Ut (we denote this magnitude by N2). We have (10) 6 N2 t∗∗(n)−1 Yt=t′ ∗(n) (c3ν t)C(Z0)nt =: N ′ 2. 3. Let us estimate the number of choices of Vt, t′ ∗(n) 6 t 6 t∗∗(n) − 1 (denote this value by N3). Since card Ut 6 C(Z0)nt, we have 2C(Z0)nt =: N ′ 3. N3 6 t∗∗(n)−1 Yt=t′ ∗(n) 16 Thus, N 6 N ′ 1N ′ 2N ′ 3, which yields log N 6 t∗∗(n) log( C(Z0)n + t∗∗(n)) + t∗∗(n)−1 Xt=t′ ∗(n) C(Z0)nt log(c3νt)+ + t∗∗(n)−1 Xt=t′ ∗(n) (10),(29),(41) C(Z0)nt . Z0 (log n)2 + t∗∗(n)−1 Xt=t′ ∗(n) 2tnt (72) . Z0 n; i.e., log N . Z0 n. (74) Step 5. Denote by N ′ the family of partitions Tf , f ∈ B Xp(Ω) (see (59)). Then N ′ ⊂ N (it follows from (52) -- (56), (58), (73) and from the estimates (46), (48)). By (74) and Theorem B, there exists l∗ = l∗(Z0) ∈ N such that el∗n( Q − Qt′ ∗(n) : Xp(Ω) → Yq( Ut′ ∗(n))) 6 6 sup f ∈B Xp(Ω) inf T∈N ′ k Qf − Qt′ ∗(n)f − PTf kYq( Ut′ ∗(n)) + sup T∈N ′ en(PT : Xp(Ω) → Yq( Ut′ ∗(n))) 6 6 sup f ∈B Xp(Ω) k Qf − Qt′ ∗(n)f − PTf f kYq( Ut′ ∗(n)) + sup T∈N ′ en(PT : Xp(Ω) → Yq( Ut′ ∗(n))) . Z0 (71) . Z0 n 1 q − 1 p + sup T∈N ′ en(PT : Xp(Ω) → Yq( Ut′ ∗(n))). Step 6. It remains to prove that en(PTf : Xp(Ω) → Yq( Ut′ ∗(n))) . Z0 1 q − 1 p , n (75) where f ∈ B Xp(Ω). Recall that Tf = {D(ξ) : ξ ∈ S}, where S is defined by (56) and D(ξ) is defined by formula (58). Also we observe that if ξ = ξt′ ∗(n),j for some j ∈ Jt′ ∗(n), then PTf ΩD(ξ) (69) = (P ξt′ ∗(n),j − Qt′ ∗(n))ΩD (16) = 0. ( ξ t′ ∗(n),j ) (76) Let t′ ∗(n) 6 t < t∗∗(n). Recall that the sets St and Wt are defined by formulas (54) and (53), respectively. We set Vf,t = St ∪ {ξ ∈ V(Γt+1)\St+1 : ∃η ∈ Wt : ξ ∈ VA 1 (η)}. (77) 17 Notice that Vf,t ⊂ S by (55), (56). Denote by Γf,t the maximal subgraph in A on vertex set V(Γf,t) = ∪ξ∈Vf,t V(D(ξ)). (78) We set Denote Ω(f,t) = ∪ξ∈V(Γf,t) F (ξ), P(f,t)h = PTf h · χΩ(f,t), h ∈ Yq(Ω). (79) T′ f,t = {D(ξ)}ξ∈Vf,t. Then T′ f,t is a partition of the graph Γf,t. We claim that PTf = t∗∗(n)−1 Xt=t′ ∗(n) P(f,t). (80) (81) Indeed, let h ∈ Yq(Ω). If x ∈ ∪t∗∗(n)−1 t∗∗(n)−1 t=t′ ∗(n) ∪ξ∈Vf,t ΩD(ξ), then by (79) we get PTf h(x) = t∗∗(n)−1 Pt=t′ ∗(n) Pt=t′ ∗(n) P(f,t)h(x). In other cases we have P(f,t)h(x) = 0. On the other hand, if PTf h(x) 6= 0, then x ∈ ΩD(ξ) for some ξ ∈ S by (59) and (70). From (55), (56), (76) and (77) it follows that ξ ∈ Vf,t for some t ∈ {t′ ∗(n), . . . , t∗∗(n) − 1}. This completes the proof of (81). Let us prove that there exists a sequence {kt}t′ ∗(n)6t<t∗∗(n) such that t∗∗(n)−1 Xt=t′ ∗(n) (kt − 1) . Z0 n, t∗∗(n)−1 Xt=t′ ∗(n) ekt(P(f,t) : Xp(Ω) → Yq(Ω)) . Z0 1 q − 1 p . n (82) Let ξ ∈ St. If ξ ∈ St\Wt, then we set D(ξ) = D(ξ), P(f,t)Ω D(ξ) = P(f,t)ΩD(ξ) (69),(79) = (Pξ − Qt′ ∗(n))ΩD(ξ) . (83) If η ∈ Wt, then we denote by D(η) the tree with vertex set V( D(η)) = {η} ∪(cid:16)∪ξ∈VA 1 (η)∩V(Γt+1)\St+1 V(D(ξ))(cid:17) (84) 18 and put P(f,t)Ω D(η) = (Pη − Qt′ ∗(n))Ω D(η) . Observe that by (78) we have T′′ f,t := { D(ξ)}ξ∈St is the partition of Γf,t. Finally, we set Let P(f,t)Ω\Ω(f,t) = 0. T ′ f,t = {ΩD : D ∈ T′ By (69), (79), (83), (85), (87) we have f,t}, T ′′ f,t = {ΩD : D ∈ T′′ f,t}. P(f,t) ∈ ST ′ f,t (Ω), P(f,t) ∈ ST ′′ f,t (Ω). Moreover, the partition T ′ Let h ∈ B Xp(Ω). Then f,t refines the partition T ′′ f,t. (85) (86) (87) (88) kP(f,t)h − P(f,t)hkp p,q,T ′ f,t (5),(69),(80),(83),(84),(85) = = Xη∈Wt X ξ∈VA 1 (η)∩V(Γt+1)\St+1 X ξ∈VA 1 (η)∩V(Γt+1)\St+1 kPξh − Pηhkp Yq(ΩD(ξ) ) . Z0 (kh − Pξhkp Yq(ΩD(ξ) ) + kh − Pηhkp Yq(ΩD(ξ) (11) . Z0 )) 2−t(1− p q ); . Xη∈Wt i.e., kP(f,t)h − P(f,t)hkp,q,T ′ f,t . Z0 2−t( 1 p − 1 q ). (89) We set k′ t = ⌈n · 2−ε(t−t∗(n))⌉, where ε > 0 is a sufficiently small number (it is chosen by Z0). For any t 6 t∗∗(n) we have card T′ f,t (80) = card Vf,t (1),(52),(53),(54),(77) . Z0 card St (54) . Z0 (44),(45),(48) card Tf,t,0 . Z0 ⌈2−tn⌉ (29) ≍ Z0 2−tn. This together with (38), (88), (89), Theorem A and Lemma 5 implies that for some c∗ = c∗(Z0) > 0 ∗(n) t∗∗(n)−1 Pt=t′ Pt=t∗(n) . t∗∗(n)−1 ek′ t(P(f,t) − P(f,t) : Xp(Ω) → Yq(Ω)) . Z0 2−t( 1 p − 1 q )(2−tn) 1 q − 1 p · 2−c∗·2t−ε(t−t∗(n)) . Z0 1 q − 1 p . n 19 (90) We claim that there exists a sequence {k′′ t }t′ ∗(n)6t<t∗∗(n) such that t∗∗(n)−1 Xt=t′ ∗(n) (k′′ t − 1) . Z0 n, t∗∗(n)−1 Xt=t′ ∗(n) ek′′ t ( P(f,t) : Xp(Ω) → Yq(Ω)) . Z0 1 q − 1 p . n (91) For each t′ ∗(n) 6 t < t∗∗(n), 0 6 l 6 log nt we consider the partition Tf,t,l as defined at Step 1. Let D ∈ Tf,t,l. We set V∗ t,D = {ξ ∈ V(Γt) ∩ V(D) : VA 1 (ξ) ∩ V(Γt+1) 6= ∅} and denote by D+ the subtree in A with vertex set Let By (48), V(D+) = V(D) ∪(cid:16)∪ξ∈V∗ t,D ∪ξ′∈VA 1 (ξ)∩V(Γt+1) V(Aξ′)(cid:17) . Tf,t,l = {D+ : D ∈ Tf,t,l}. card Tf,t,l . Z0 2−lnt. We claim that for any tree D ∈ Tf,t,l (92) (93) (94) card {D′ + : D′ ∈ Tf,t,l±1 : V(D+) ∩ V(D′ +) 6= ∅} . Z0 1. (95) Indeed, by (49), it is sufficient to check that if V(D+) ∩ V(D′ V(D′) 6= ∅. Let ξ ∈ V(D+)∩ V(D′ V(D′)) or ξ ∈ V(Aη), where η ∈ V(Γt) ∩ V(D) ∩ V(D′), VA +) 6= ∅, then V(D) ∩ +). Then either ξ ∈ V( At) (therefore, ξ ∈ V(D)∩ 1 (η) ∩ V(Γt+1) 6= ∅. Assertion 2. We have Tf,t,0Γf,t = T′′ f,t. If (T , ξ) ∈ Tf,t,0, ξ /∈ V(Γt), then V(T+) ∩ V(Γf,t) = ∅. Proof of Assertion 2. 1. Let (T , ξ) ∈ Tf,t,0. We claim that either there exists a tree D ∈ T′′ f,t such that V(T+) ∩ V(Γf,t) = V(D) or V(T+) ∩ V(Γf,t) = ∅. Case ξ ∈ V(Γt). By (54), we have ξ ∈ St. Let us show that V(T+) ∩ V(Γf,t) = V( D(ξ)) (96) and apply (86). 20 • We claim that V(T+) ∩ V(Γf,t) ⊃ V( D(ξ)). Since T′′ f,t is a partition of the graph Γf,t, we have V( D(ξ)) ⊂ V(Γf,t). Let us prove that V( D(ξ)) ⊂ V(T+). Indeed, let ξ ∈ V( D(ξ))\V(T+). We set η∗ = min{η ∈ [ ξ, ξ] : η /∈ V(T+)} > ξ. Denote by ζ∗ the direct predecessor of η∗. Then ζ∗ /∈ V(Aζ ′) (otherwise, η ∈ V(T+)). Hence, ζ∗ ∈ ∪ζ∈V∗ V(Γt). If η∗ /∈ V(Γt), then η∗ ∈ V(Γt+1) (by Assumption 1, condition 1); once again, we get η∗ ∈ V(T+). Thus, η∗ ∈ V(Γt) and η∗ is the minimal vertex of some tree T ∈ Tf,t,0. Therefore, η∗ S, which (54) ∈ St 1 (ζ)∩V(Γt+1) ∪ζ ′∈VA t,T (55),(56) ⊂ (58),(83),(84) implies η∗ which leads to a contradiction. /∈ V( D(ξ)). On the other hand, η∗ ∈ [ ξ, ξ] ⊂ V( D(ξ)), • We claim that V(T+)∩V(Γf,t) ⊂ V( D(ξ)). We have V(Γf,t) (86) = ∪η∈St V( D(η)). If ζ ∈ V( D(η)) ∩ V(T+), then the vertices ξ and η are comparable. Since ξ ∈ St ⊂ S, the case η < ξ is impossible by (58), (83), (84). Consequently, η ∈ [ ξ, ζ] ⊂ V(T+); in addition, η ∈ St. By (54), η ∈ V(Γt) is the minimal vertex of some tree in Tf,t,0. Therefore, η = ξ; i.e., ζ ∈ V( D(ξ)). This completes the proof of (96). Case ξ /∈ V(Γt). Then ξ ∈ V(Γt′), t′ < t. Let us check that V(T+) ∩ V(Γf,t) = (86) ∈ V( D(η)) for some η ∈ St; i.e., ∅. Indeed, let ξ ∈ V(T+) ∩ V(Γf,t). Then ξ (54) ∈ V(Γt) is the minimal vertex of some tree from Tf,t,0 (this tree does not η coincide with T since ξ /∈ V(Γt) is the minimal vertex of T ). In addition, the vertices ξ and η are comparable. The case η < ξ is impossible by Assumption 1 (see condition 1). Hence, η ∈ [ ξ, ξ] ⊂ V(T+); i.e., η ∈ V(T ), which leads to a contradiction. 2. Let D(ξ) ∈ T′′ f,t. We claim that there exists a tree T ∈ Tf,t,0 such that V( D(ξ)) = V(T+) ∩ V(Γf,t). Indeed, since ξ ∈ St by (86), we have ξ ∈ V(Γt) and ξ is the minimal vertex of some tree T ∈ Tf,t,0 (see (54)). By (96), V(T+) ∩ V(Γf,t) = V( D(η)) for some η ∈ St. In addition, ξ ∈ V(T+) ∩ V(Γf,t); i.e., ξ ∈ V( D(η)) and η = ξ. This completes the proof of Assertion 2. ⋄ Let D ∈ Tf,t,0, t′ ∗(n) 6 t < t∗∗(n). If the minimal vertex of D does not belong to ∗(n)), by Assertion 2 we have V(D)∩V(Γf,t) = ∅. Hence, if V(D)∩V(Γf,t) 6= ∅, V(Γt′ then Γt′ ∗(n) ∩ D is a tree. Denote by Tn ∗(n) ∩ D, D ∈ Tf,t,l. We have f,t,l the partition of Γt′ Γt′ ∗(n) formed by connected components of graphs P(f,t)h = P Tn f,t,0 h · χΩ(f,t), h ∈ Yq(Ω). (97) 21 It follows from (69), (79), (83), (85), (87), (88) and Assertion 2. Given 0 6 l 6 log nt, we set kt,l = ⌈n · 2−ε(t−t∗(n)+l)⌉, k′′ (kt,l − 1), t∗(n) 6 t < t∗∗(n), t′ ∗(n) 6 t < t∗∗(n). (98) t = 1 + P06l6log nt Pt=t′ ∗(n) t∗∗(n)−1 n · 2−ε(t−t∗(n)), Then k′′ t − 1 . Z0 (k′′ t − 1) . Z0 n. From (51), (92) and (93) it follows that Tf,t,⌊log nt⌋ = {A}, Tn f,t,⌊log nt⌋ = {At′ ∗(n),j}j∈ Jt′ . ∗(n) From (16) and (69) we get P Tn f,t,⌊log nt ⌋ = 0. Hence, by (97) and (37), ek′′ ekt,l(P Tn t ( P(f,t) : Xp(Ω) → Yq(Ω)) 6 − P Tn f,t,l+1 f,t,l : Xp(Ω) → Yq(Ω(f,t))). (99) 6 P06l<⌊log nt⌋ Denote by Tn f,t,l+1, and either D′ ∈ Tf,t,l or D′′ ∈ Tf,t,l+1. We set f,t,l the partition formed by trees D′ ∩ D′′, where D′ ∈ Tn Tn f,t,l, D′′ ∈ T n f,t,l = {ΩD}D∈ Tn f,t,l , T n f,t,l = {E = ΩD ∩ Ω(f,t) : mes E > 0}D∈ Tn f,t,l . (100) Let h ∈ B Xp(Ω). We show that (P Tn f,t,l h − P Tn f,t,l+1 h)χΩ(f,t) ∈ S T n f,t,l (Ω). (101) Indeed, if D = D′ ∩ D′′, D′ ∈ Tn f,t,l, D′′ ∈ Tn f,t,l+1, then P Tn f,t,l (69) ∈ P(ΩD′), hΩD′ P Tn f,t,l+1 hΩD′′ (69) ∈ P(ΩD′′); therefore, (P Tn f,t,l h − P Tn f,t,l+1 h)ΩD∩Ω(f,t) ∈ P(ΩD ∩ Ω(f,t)). (102) We show that if D′ /∈ Tf,t,l and D′′ /∈ Tf,t,l+1, then (P Tn h)ΩD∩Ω(f,t) = 0. Indeed, in this case minimal vertices of the trees D′ and D′′ are equal and coincide with ξt∗(n),j for some j ∈ Jt∗(n). From (69) it follows that (P Tn h)ΩD = 0. From (4) and (102) we get (101). We claim that for any E′ ∈ T n h−P Tn h−P Tn f,t,l+1 f,t,l, E′′ ∈ T n f,t,l+1 f,t,l+1 f,t,l f,t,l card {E ∈ T n f,t,l : E ⊂ E′} . Z0 1, card {E ∈ T n f,t,l : E ⊂ E′′} . Z0 1. (103) Let us check the first inequality (the second one is proved similarly). Let E = ΩD′, f,t,l, E ∈ T n D′ ∈ Tn f,t,l is a partition, we have E = ΩD′∩D′′ ∩Ω(f,t), f,t,l+1, V(D′) ∩ V(D′′) 6= ∅. There exist trees T ′ ∈ Tf,t,l and T ′′ ∈ Tf,t,l+1 D′′ ∈ Tn f,t,l, E ⊂ E′. Since Tn 22 ∗(n), respectively. Observe that the connected component of the graph T ′′ ∩ Γt′ such that D′ and D′′ are connected components of the graphs T ′ ∩ Γt′ ∗(n) and T ′′ ∩ Γt′ ∗(n) whose vertex set intersects with V(D′) is unique. This together with (95) implies (103). From (94), (103) and the definition of Tn f,t,l we get that card T n f,t,l . Z0 2−lnt. (104) For any h ∈ B Xp(Ω) we have k(P Tn f,t,l h − P Tn f,t,l+1 h)χΩ(f,t)kp,q, T n f,t,l (5),(103) . Z0 . k(h − Qt′ ∗(n)h − P Tn f,t,l h)χ Ut′ ∗(n) kp,q, T n f,t,l + +k(h − Qt′ ∗(n)h − P Tn f,t,l+1 h)χ Ut′ ∗(n) kp,q, T n f,t,l+1 (5),(69),(100) = (11) . Z0 (105) (106) .  X(D, ξ)∈ Tn f,t,l i.e., kh − Pξhkp 1/p Yq(ΩD)  ∗(n) . Z0 +  X(D, ξ)∈ Tn p )t∗(n) (29) ≍ Z0 . 2( 1 q − 1 p )t′ 2( 1 q − 1 (log n) 1 q − 1 p ; kh − Pξhkp f,t,l+1 1/p Yq(ΩD)  k(P Tn f,t,l h − P Tn f,t,l+1 h)χΩ(f,t)kp,q, T n f,t,l (log n) 1 q − 1 p . . Z0 From (38), (101), (104), (105) and Lemma 5 we get that ekt,l(P Tn f,t,l − P Tn f,t,l+1 : Xp(Ω) → Yq(Ω(f,t))) . Z0 1 q − 1 p ekt,l(Ist,l : lst,l p → lst,l q ) =: At,l, . (log n) where st,l ∈ N, st,l 6 C∗ · 2−lnt (44),(45) 6 C∗ · 2−l⌈n · 2−t⌉, C∗ = C∗(Z0) > 1. For each t 6 t∗∗(n) we have (98) 6 st,l kt,l C∗ · 2−l⌈n · 2−t⌉ n · 2−ε(l+t−t∗(n)) =: σ′ t,l, 23 σ′ t,l (29) ≍ Z0 2−ln · 2−t n · 2−ε(l+t−t∗(n)) = 2−l(1−ε)−t(1−ε)−εt∗(n) 6 1. (107) together with Theorem A implies that there exists γ0 = γ0(Z0) > 0 such that decreases not slower than some geometric progression. This The sequence (cid:8)σ′ t,l(cid:9)l∈Z+ t∗∗(n)−1 Xt=t′ ∗(n) X06l<log nt At,l (106) . Z0 (log n) 1 q − 1 p . (log n) 1 q − 1 p 1 q − 1 t,l p k t∗∗(n)−1 t∗∗(n)−1 Xt=t′ ∗(n) X06l<log nt Xt=t∗(n) q − 1 t,0 k 1 p t∗∗(n)−1 Xt=t′ ∗(n) X06l<log nt s 1 q − 1 t,l 2 p − kt,l st,l . Z0 · (σ′ t,l) 1 q − 1 p · 2 − 1 σ′ t,l (98),(107) . Z0 . (log n) 1 q − 1 p · 2(t(1−ε)+εt∗(n))( 1 q − 1 p ) · 2−γ0·2t(1−ε)+εt∗(n) (98) . Z0 . (log n) i.e., 1 q − 1 p · n 1 q − 1 p · 2( 1 q − 1 p )t∗(n) · 2−γ0·2t∗(n) (29) . Z0 1 q − 1 p ; n t∗∗(n)−1 Xt=t′ ∗(n) X06l<log nt 1 q − 1 p . n At,l . Z0 (108) From (99), (106) and (108) we get (91). This together with (90) yields (82). Applying (81), we have (75). Taking into account the estimate obtained at Step 5, we complete the proof of Lemma 3. It remains to prove that en ∞ Xm=0 ( Qn,m+1 − Qn,m) : Xp(Ω) → Yq(Ω)! . Z0 1 q − 1 p . n (109) In [23] there were obtained order estimates for entropy numbers of diagonal operators with weights of logarithmic type. First we give some notations. Denote by Φ0 the class of non-decreasing functions ϕ : [1, ∞) → (0, ∞) that satisfy the following condition: there exist c > 0 and α > 0 such that for any 1 6 s 6 t < ∞ ϕ(t) ϕ(s) 1 + log s(cid:19)α 6 c(cid:18) 1 + log t . (110) We set wϕ(s) = 1 for 0 6 s 6 1 and wϕ(s) = ϕ(s) for s > 1. 24 Let δ > 0, w : R+ → R+ be a continuous function, and let 0 < r, p 6 ∞. For x = (xm,k)m,k∈Z+ we set ∞ kxlr(2δmlp(w))k := Xm=0  2mδr Xk∈Z+ xm,kw(2−mk)p  1 r r p  (appropriately modified if p = ∞ or r = ∞). By lr(2δmlp(w)) we denote the space of sequences x such that kxlr(2δmlp(w))k < ∞. Theorem C. [23, p. 11]. Let 0 < p < q 6 ∞, 0 < r, s 6 ∞, δ > 0, ϕ ∈ Φ0, and let (110) hold with α = 1 p − 1 q . Then en(id : lr(2δmlp(wϕ)) → ls(lq)) ≍ p,q,r,s,ϕ 1 ϕ(2n) . Let us prove (109). Given m > mt, we denote by Tt,m the partition of the set Gt formed by E′ ∩ E′′, t,m = E′ ∈ Tt,m, E′′ ∈ Tt,m+1 (the partitions Tt,m are defined at page 7). Let s′′ dim S Tt,m(Ω). Then there exists M = M(Z0) > 1 such that (10),(19),(21),(24) 6 s′′ t,m M · 2m−mt · 2γ∗2t ψ∗(22t ). (111) Given t > 0, m′ > 0, we set st,m′ = ⌈M · 2m′ · 2γ∗2t ψ∗(22t )⌉, s∗ t,m′ = sl,m′. t−1 Xl=0 Denote ϕ(x) = (log(2 + x)) 1 p − 1 q , x > 0. Let k = s∗ t,m′ + j, 1 6 j 6 st,m′. Then 2γ∗2t−1 ψ∗(22t−1 ) . Z0 2−m′ k . Z0 2γ∗2t ψ∗(22t ), t > 1, 2−m′k . Z0 1 if t = 0. Hence, wϕ(2−m′ k) (41) ≍ Z0 2( 1 p − 1 q )t, k = s∗ t,m′ + j, 1 6 j 6 st,m′. (112) From Lemma 5 it follows that there is an isomorphism At,m : S Tt,m(Ω) → Rs′′ t,m such that kAt,mk s′′ t,m (Gt)→l p 1, . Z0 Yp,q, Tt,m kA −1 t,mk s′′ t,m q →Yq(Gt) l 1. . Z0 (113) 25 Let us define the operator A : Xp(Ω) → l∞(2δ∗m′lp(wϕ)) as follows. Consider a function f ∈ Xp(Ω). Then ( Qn,m′+1f − Qn,m′f )Gt (20),(32) ∈ S Tt,mt+m′ (Ω). Let At,mt+m′(( Qn,m′+1f − Qn,m′f )Gt) = (cm′,t,j) s′′ t,mt+m′ j=1 . We set Then (Af )m′,s∗ t,m′ +j =(cid:26) cm′,t,j 0 for for 1 6 j 6 s′′ s′′ t,mt+m′ + 1 6 j 6 st,m′ or t < t0. t,mt+m′, t > t0, kAf kl∞(2δ∗m′ lp(wϕ)) = sup m′>0 wϕ(2−m′ . sup m′>0 2δ∗m′ Xt>t0 2δ∗m′ Xk∈Z+ Xj=1 2(1− p q )t s′′ t,mt+m′ cm′,t,jp  1 p (112) . Z0 k)(Af )m′,kp  1/p =: N. From (22), (23) and (24) it follows that for m > mt kPt,m+1 − Pt,mk Xp(Gt)→Yp,q, Tt,m (Gt) . Z0 Hence, by (32), (113) we get 2−δ∗(m−mt) · 2( 1 q − 1 p )tkf kXp(Gt). s′′ t,mt+m′ Xj=1 cm′,t,jp . Z0 2−δ∗pm′ 2( p q −1)tkf kp Xp(Gt); therefore, Thus, N . Z0 Xt>t0 kf kp Xp(Gt)!1/p = kf kXp(Ω). kAk Xp(Ω)→l∞(2δ∗m′ lp(wϕ)) . Z0 1. (114) Let us define the operator K : l1(lq) → Yq(Ω) by formula K((cm′,s∗ t,m′ +j)m′∈Z+,t∈Z+,16j6st,m′ ) = Xm′∈Z+Xt>t0 A −1 t,mt+m′((cm′,s∗ t,m′ +j) s′′ t,mt+m′ j=1 ). 26 Since the sets Gt do not overlap pairwise, we have kK((cm′,s∗ t,m′ +j)m′∈Z+,t∈Z+,16j6st,m′ )kYq(Ω) 6 ∞ 6 Xm′=0 ∞ Xt=t0  Xj=1 Xt∈Z+ st,m′ cm′,s∗ t,m′ +jq  ∞ Xm′=0 . Hence, kA −1 t,mt+m′((cm′,s∗ t,m′ +j) s′′ t,mt+m′ j=1 )kq Yq(Gt)!1/q (113) . Z0 1/q = k(cm′,s∗ t,m′ +j)m′∈Z+,t∈Z+,16j6st,m′ kl1(lq). kKkl1(lq)→Yq(Ω) . Z0 1. (115) Let id : l∞(2δ∗m′lp(wϕ)) → l1(lq) be the identity operator. Then en ∞ Xm=0 ( Qn,m+1 − Qn,m) : Xp(Ω) → Yq(Ω)! = = en(K ◦ id ◦ A : Xp(Ω) → Yq(Ω)) (38),(114),(115) . Z0 . en(id : l∞(2δ∗m′ lp(wϕ)) → l1(lq)) =: En. From Theorem C we get En . Z0 1 q − 1 p . n Applying Lemmas 1, 2, 3 together with (30), (33), (34), (109), we complete the proof of Theorem 1. Remark 1. Suppose that instead of Assumption 2 the following condition holds: for any ξ ∈ V(A) the set F (ξ) is the atom of mes. Then the assertion of Theorem 1 holds as well. 3 Some particular cases Let (T , ξ0) be a tree, and let u, w : V(T ) → R+. We define the summation operator Su,w,T by u(ξ′)f (ξ′), ξ ∈ V(T ), f : V(T ) → R. Let 1 6 p, q 6 ∞. By Sp,q T ,u,w we denote the minimal constant C in the inequality Su,w,T f (ξ) = w(ξ)Xξ′6ξ q u(ξ′)f (ξ′)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) wq(ξ)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xξ′6ξ    Xξ∈V(T ) 1/q 6 C  Xξ∈V(T ) 27 1/p f (ξ)p  , f : V(T ) → R (appropriately modified for p = ∞ or q = ∞). j=0 ∪N j (ξ0). Given a tree (T , ξ0), N ∈ Z+, we denote by [T ]6N a tree with vertex set VT Let (A, ξ0) be a tree, let (1) hold, and let the measure space (Ω, Σ, mes), the partition Θ, the bijection F : V(A) → Θ and the spaces Xp(Ω), Yq(Ω), P(Ω) be as defined at the page 3. Assumptions 1 -- 3 will be replaced by the following conditions. Assumption A. There exist functions u, w : V(A) → (0, ∞) and a constant c2 > 1 with the following property: for any vertex ξ∗ ∈ V(A) there exists a linear continuous projection Pξ∗ : Yq(Ω) → P(Ω) such that for any vertex ξ > ξ∗ and for any function f ∈ Xp(Ω) kf − Pξ∗f kYq( F (ξ)) 6 c2w(ξ) Xξ∗6ξ′6ξ u(ξ′)kf kXp( F (ξ′)). (116) Assumption B. There exists a number δ∗ > 0 such that for any vertex ξ ∈ V(A) and for any n ∈ N, m ∈ Z+ there exists a partition Tm,n(G) of the set G = F (ξ) with the following properties: 1. card Tm,n(G) 6 c2 · 2mn. 2. For any E ∈ Tm,n(G) there exists a linear continuous operator PE : Yq(Ω) → P(E) such that for any function f ∈ Xp(Ω) kf − PEf kYq(E) 6 c2(2mn)−δ∗u(ξ)w(ξ)kf kXp(E). (117) 3. For any E ∈ Tm,n(G) card {E′ ∈ Tm±1,n(G) : mes(E ∩ E′) > 0} 6 c2. (118) Let V(A) = {ηj,i}j>jmin, i∈ Ij , where jmin > 0, Ijmin = {1}. Suppose that ηjmin,1 is the minimal vertex of A and VA j−jmin(ηjmin,1) = {ηj,i}i∈ Ij for any j > jmin. Assumption C. There exist numbers θ > 0, γ ∈ R, κ > θ q , αu, αw ∈ R, c3 > 1, m∗ ∈ N and an absolutely continuous function τ : (0, ∞) → (0, ∞) such that lim t→+∞ tτ ′(t) τ (t) = 0 and the following conditions hold. 1. If κ = θ q , then αw > 1−γ q . 2. If κ > θ q , then αu + αw = 1 p − 1 q ; if κ = θ q , then αu + αw = 1 p . 3. For any j′ > j > jmin and for any vertex ξ ∈ VA j−jmin(ηjmin,1) card VA j′−j(ξ) 6 c3 · 2θm∗(j′−j) jγτ (m∗j) j′γτ (m∗j′) . (119) 28 4. For any j > jmin, i ∈ Ij u(ηj,i) = uj = 2κm∗j(m∗j)−αu, w(ηj,i) = wj = 2−κm∗j(m∗j)−αw . (120) 5. Let κ = θ q . Then there exists a tree A with the minimal vertex ζ0, which satisfies the following conditions. (a) For any j′ > j > jmin and for any vertex ξ ∈ V A j−jmin(ζ0) c−1 3 · 2θm∗(j′−j) jγτ (m∗j) j′γτ (m∗j′) 6 card V A j′−j(ξ) 6 c3 · 2θm∗(j′−j) jγτ (m∗j) j′γτ (m∗j′) . (121) (b) Let {uj}j>jmin ⊂ (0, ∞), {wj}j>jmin ⊂ (0, ∞) be arbitrary sequences. Define the functions u, w : V(A) → (0, ∞), u, w : V( A) → (0, ∞) by uVA j−jmin (ηjmin ,1) ≡ uj, uV A j−jmin (ζ0) ≡ uj, wVA j−jmin (ηjmin ,1) ≡ wj, wV A j−jmin (ζ0) ≡ wj, j > jmin. Then for each N > j > jmin, i ∈ Ij there exists a vertex ξ ∈ V such that Sp,q [Aηj,i ]6N−j ,u, w 6 c3Sp,q [ A ξ]6N−j ,u, w . A j−jmin(ζ0) Denote Z0 = (p, q, c1, c2, c3, θ, γ, κ, αu, αw, m∗, δ∗, τ ). Let ξ∗ ∈ V(A), and let D ⊂ A be a subtree with the minimal vertex ξ∗. Then from Assumption A it follows that kf − Pξ∗f kYq(ΩD) 6 c2Sp,q D,u,wkf kXp(ΩD). (122) From Theorem F in [36], Theorem 3.6 in [38] and Assumption C it follows that if ξ∗ ∈ VA j−jmin(ξ0), then in the case κ > θ q Sp,q D,u,w . Z0 sup s>j usws ≍ Z0 (m∗j)−αu−αw = (m∗j) 1 q − 1 p , (123) and in the case κ = θ q Sp,q D,u,w . Z0 sups>j s Pi=j (m∗i)−p′αu · 2p′κm∗i! 1 p′ (cid:18)Pi>s (m∗i)−αwq · sγτ (m∗s) iγτ (m∗i) · 2−qκm∗s(cid:19) 1 q . Z0 . (m∗j)−αu−αw+ 1 q = (m∗j) 1 q − 1 p . Notice that the proof of (124) depends on condition 5 of Assumption C. 29 (124) From Assumption B and conditions 2, 4 of Assumption C it follows that if ξ ∈ j−jmin(ξ0), G = F (ξ), E ∈ Tm,n(G), then VA kf − PEf kYq(E) 6 c2 · (2mn)−δ∗(m∗j)−αu−αwkf kXp(E) 6 6 c2 · (2mn)−δ∗(m∗j) 1 q − 1 p kf kXp(E). (125) We construct the partition {At,i}t>t0, i∈ Jt as follows. Let t0 = min{t ∈ Z+ : 2t > m∗jmin}. Given t > t0, we denote by Γt the maximal subgraph in A on the vertex set V(Γt) = {ηj,s : 2t−1 6 m∗j < 2t, s ∈ Ij}, (126) and by At,i, i ∈ Jt, the connected components of the graph Γt. By ξt,i we denote the minimal vertex of the tree At,i. Then card V(Γt) (119) . Z0 2θ·2t 2−γtτ −1(2t). (127) Thus, Assumption 3 holds with νt = 2θ·2t2−γtτ −1(2t). From (125) we obtain that Assumption 2 holds. Recall the notations Jt,D = {i ∈ Jt : V(At,i) ∩ V(D) 6= ∅}, Dt,i = D ∩ At,i, where D ⊂ A is a subtree. In order to obtain Assumption 1, it is sufficient to prove the following assertion. Lemma 7. Let D be a subtree in A rooted at ξ∗. Then kf − Pξ∗f kq Yq(ΩD) . Z0 ∞ Xt=t0 2(1− q p )t Xi∈Jt,D kf kq Xp(ΩDt,i ). (128) Proof. By (116), kf − Pξ∗f kq Yq(ΩD) . Z0 Xξ∈V(D) wq(ξ) Xξ∗6ξ′6ξ u(ξ′)kf kXp( F (ξ′))!q . Let κ > θ proof of Lemma 5.1 in [41], we get that q . Then αu + αw = 1 p − 1 q (see condition 2 of Assumption C). Repeating the kf − Pξ∗f kq Yq(ΩD) . Z0 Xξ∈V(D) wq(ξ)uq(ξ)kf kq Xp( F (ξ)) . If ξ = ηj,i, 2t−1 6 m∗j < 2t, then uq(ξ)wq(ξ) together with the condition p < q implies (128). (120) = (m∗j)− q p +1 ≍ Z0 2(1− q p )t. This 30 Now we consider the case κ = θ q . Let ξ∗ ∈ V(At,j0), and let ξ ∈ V(Dt+l,jl). Then there exists a sequence { ξt+s,js}l s=1 such that ξ∗ < ξt+1,j1 < ξt+2,j2 < · · · < ξt+l,jl. Denote ξt,j0 = ξ∗, ξt+s,js = ξt+s,js, 1 6 s 6 l. (129) We have u(ξ′)f (ξ′) = Xξ∗6ξ′6ξ l−1 Xs=0 Xξt+s,js 6ξ′< ξt+s+1,js+1 u(ξ′)kf kXp( F (ξ′)) + Xξt+l,jl 6ξ′6ξ u(ξ′)kf kXp( F (ξ′)). (130) (131) Let αw = αw − 1 q . By condition 2 of Assumption C, We have αu + αw = 1 p − 1 q . wq(ξ) Xξ∈V(Dt+l, jl ) (119),(120),(126) . Z0 . X2t+l−16m∗j<2t+l m∗j !γ 2−θm∗j(m∗j)−αwq · 2θ(m∗j−2t+l−1) 2t+l−1 τ (2t+l−1) τ (m∗j) (41) . Z0 . 2−κq·2t+l−1 · 2(t+l)(−αwq+1) = 2−κq·2t+l−1 · 2−(t+l)αwq, which implies kf − Pξ∗f kq Yq(ΩDt+l,jl ) (116) . Z0 Xξ∈V(Dt+l,jl ) wq(ξ) Xξ∗6ξ′6ξ . Xξ∈V(Dt+l,jl ) + Xξ∈V(Dt+l,jl . 2−qκ·2t+l−1 u(ξ′)f (ξ′)!q u(ξ′)kf kXp( F (ξ′))  q q (130) . Z0 + 6ξ′< ξt+s+1,js+1 l−1 wq(ξ) Xs=0 Xξt+s,js  wq(ξ)  Xξt+l,jl · 2−q αw(t+l) Xs=0  6ξ′6ξ l−1 ) Xξt+s,js 31 . Z0 (124),(126) u(ξ′)kf kXp( F (ξ′))  u(ξ′)kf kXp( F (ξ′))  6ξ′< ξt+s+1,js+1 q + +2(1− q p )(t+l)kf kq Xp(Dt+l,jl ); i.e., kf − Pξ∗f kq Yq(ΩDt+l,jl +2−qκ·2t+l−1 · 2−q αw(t+l) Ps=0  l−1 Denote 2(1− q p )(t+l)kf kq Xp(Dt+l,jl )+ ) . Z0 ξt+s,js P 6ξ′< ξt+s+1,js+1 q u(ξ′)kf kXp( F (ξ′))  . {ζt,i : i ∈ I ′ t} = {ζ ∈ Vmax(D ∩ Γt) : VD 1 (ζ) 6= ∅}. By (126), ∀i ∈ I ′ t ∃s ∈ I⌈2t/m∗⌉−1 : ζt,i = η⌈2t/m∗⌉−1,s. (132) (133) (134) Let us define the tree D with vertex set {ξ∗} ∪(cid:0)∪t>t{ζt,i : i ∈ I ′ order on V( D) is defined as follows. We set t}(cid:1). The partial V D 1 (ζt,i) = {ζt+1,j : ζt+1,j ∈ Dζt,i}. If ξ∗ /∈ {ζt,i : i ∈ I ′ t}, then we set V D 1 (ξ∗) = {ζt,i : i ∈ I ′ t}. Further, we define the functions ϕ, u, w : V( D) → R+. We set ϕ(ζt,i) = Xξ′∈V(Γt)∩V(D), ξ′6ζt,i u(ξ′) u(ζt,i) kf kXp( F (ξ′)), (135) w(ζt,i) = 2−κ·2t · 2− αwt, u(ζt,i) = 2κ·2t ≍ Z0 t}, then we set ϕ(ξ∗) = w(ξ∗) = u(ξ∗) = 0. · 2−αut (120),(134) If ξ∗ /∈ {ζt,i : i ∈ I ′ u(ζt,i). (136) Let t > t, j ∈ Jt,D. If t > t, then ξt,j ∈ V(D). We take i ∈ I ′ t−1 such that ξt,j ∈ VD 1 (ζt−1,i). Then kf − Pξ∗f kq Yq(Dt,j) (132),(136) . Z0 2(1− q p )tkf kq Xp(Dt,j) + wq(ζt−1,i)  Xζ ′∈V( D), ζ ′6ζt−1,i q u(ζ ′)ϕ(ζ ′)  (137) . Notice that card VD 1 (ζt,i) (119) . Z0 1. Summing (137) over all t > t, j ∈ Jt,D, we obtain kf − Pξ∗f kq ∞ 2(1− q Pt=t Yq(ΩD ) . Z0 p )t Pj∈Jt,D wq(ζt,i) Pζ ′∈V( D), ζ ′6ζt,i kf kq Xp(ΩDt,j )+ u(ζ ′)ϕ(ζ ′)!q + ∞ Pt=tPi∈I ′ t (138) . 32 Let us estimate the second summand (we denote it by L). Let u(ζ) = u1(ζ)u2(ζ), u1(ζt,i) = 2εt, where ε = ε(Z0) > 0 is sufficiently small t}, then we set u1(ξ∗) = u2(ξ∗) = 0. By (it will be chosen later); if ξ∗ /∈ {ζt,i : i ∈ I ′ Holder's inequality, L = Xζ∈V( D) 6 u1(ζ ′)u2(ζ ′)ϕ(ζ ′)!q wq(ζ) Xζ ′6ζ 1 (ζ ′)! wq(ζ) Xζ ′6ζ 2(ζ ′)ϕq(ζ ′) = Xζ ′∈V( D) Xζ ′6ζ 2(ζ ′)ϕq(ζ ′)Xζ>ζ ′ uq′ uq uq q q′ uq 2(ζ ′)ϕq(ζ ′) . Z0 6 Xζ∈V( D) 1(ζ)Xζ ′6ζ wq(ζ)uq . Xζ∈V( D) wq(ζ)uq 1(ζ) =: M. By (134), there exists s ∈ I⌈2t/m∗⌉−1 such that card V D l (ζt,i) = card {i′ ∈ I ′ t+l : ζt+l,i′ > ζt,i} 6 6 card VA ⌈2t+l/m∗⌉−⌈2t/m∗⌉(η⌈2t/m∗⌉−1,s) (119) . Z0 2θ·2t+l2−γ(t+l)τ −1(2t+l) 2θ·2t2−γtτ −1(2t) . This together with relations κ = θ yields that for sufficiently smalll ε > 0 q , αw > 1−γ q (see condition 1 of Assumption C) wq(ζ)uq 1(ζ) Xζ>ζt,i (136) . Z0 Xl>0 2−qκ·2t+l · 2−q αw(t+l) · 2qε(t+l) 2θ·2t+l2−γ(t+l)τ −1(2t+l) 2θ·2t2−γtτ −1(2t) = = 2−θ·2t Thus, · 2γtXl>0 M . Z0 Xζ ′∈V( D) i.e., ∞ Xt=t Xi∈I ′ t 2−q(αw+ γ q −ε)(t+l) τ (2t) τ (2t+l) (41) . Z0 2−qκ·2t · 2−q(αw−ε)t (136) = wq(ζt,i)uq 1(ζt,i). wq(ζ ′)uq 1(ζ ′)uq 2(ζ ′)ϕq(ζ ′) = Xζ ′∈V( D) p )t Xi∈I ′ 2(1− q t: ζt,i∈V(Dt,j ) ∞ = Xt=t Xj∈Jt,D wq(ζ ′)uq(ζ ′)ϕq(ζ ′) (131),(136) = ϕq(ζt,i); wq(ζt,i)  Xζ ′∈V( D), ζ ′6ζt,i q u(ζ ′)ϕ(ζ ′)  ∞ . Z0 Xt=t Xj∈Jt,D 33 2(1− q p )t Xi∈I ′ t: ζt,i∈V(Dt,j) ϕq(ζt,i). (139) It remains to prove that for any t > t, j ∈ Jt,D Xi∈I ′ t: ζt,i∈V(Dt,j ) ϕq(ζt,i) . Z0 kf kq Xp(ΩDt,j ). Then (138), (139), (140) imply (128). By (129) and (135), Xi∈I ′ t: ζt,i∈V(Dt,j ) t: ζt,i∈V(Dt,j) ϕq(ζt,i) = Xi∈I ′ u−q(ζt,i) = Xi∈I ′ t: ζt,i∈V(Dt,j ) q kf kXp( F (ξ′))  q =: S. u(ξ′) u(ζt,i)   Xξt,j 6ξ′6ζt,i u(ξ′)kf kXp( F (ξ′))   Xξt,j 6ξ′6ζt,i From (133) it follows that ζt,i ∈ Vmax(At,j). We denote A′ . Then V(Dt,j) ⊂ V(A′ t,j = (At,j) ξt,j t,j). Given σ > 0, we set (140) = Then σ for w(σ)(ξ) =(cid:26) u−1(ξ) S 6hSp,q A′ t,j,u,w(σ)iq for ξ ∈ V(A′ ξ ∈ Vmax(A′ t,j), t,j)\Vmax(A′ t,j). kf kXp(ΩDt,j ). We claim that for sufficiently small σ > 0 the estimate Sp,q A′ t,j,u,w(σ) 1 holds. This . Z0 implies (140). Let ξt,j ∈ VA s∗−jmin(ηjmin,1). From (126) and (129) it follows that 2t−1 6 m∗s∗ < 2t. (141) t,j = [A ξt,j ]6⌈2t/m∗⌉−1−s∗ (since ξt+1,i ∈ VA (ηjmin,1), i ∈ Jt). By Moreover, A′ A s∗−jmin(ζ0) such that for condition 5b of Assumption C, there exists a vertex ζ ∈ V the tree At,j := [ Aζ]6⌈2t/m∗⌉−1−s∗ and for the functions u, w(σ) : V( At,j) → (0, ∞) defined by ⌈2t/m∗⌉−jmin u(ξ) = us, ξ ∈ V A s−s∗(ζ0), w(ξ) =( u−1 ⌈2t/m∗⌉−1, ξ ∈ V σ, ξ ∈ V A s−s∗(ζ0), A ⌈2t/m∗⌉−1−s∗(ζ0), s < ⌈2t/m∗⌉ − 1, (142) the inequality Sp,q A′ t,j ,u,w(σ) Sp,q At,j,u,w(σ) . Z0 holds. From (121) it follows that we can apply Theorem 3.6 in [38] and estimate Sp,q At,j,u,w(σ) from above. For sufficiently small σ = σ(Z0, t) > 0 we get Sp,q At,j,u,w(σ) (121),(142) . Z0 sup s∗6s<⌈2t/m∗⌉ s Xl=s∗ up′ l !1/p′ × 34 ×  . ⌈2t/m∗⌉−2 Xl=s s∗6s6⌈2t/m∗⌉−1 s Xl=s∗ sup Since θ = qκ, we have σq · 2θm∗(l−s) sγτ (m∗s) lγτ (m∗l) + u−q (120),(141) 1/q 2γtτ (2t)  ⌈2t/m∗⌉−1 · 2θ(2t−m∗s) (m∗s)γτ (m∗s)  (cid:16)2−qκ·2t · 2qαut · 2θ(2t−m∗s)(cid:17)1/q . Z0 . 2p′κm∗l · (m∗l)−αup′!1/p′ Sp,q At,j,u,w(σ) . Z0 sup s∗6s6⌈2t/m∗⌉−1 2κm∗s · (m∗s)−αu · 2−κm∗s · 2αut (141) 1. . Z0 This completes the proof. Thus, Assumptions 1, 2, 3 hold, and by Theorem 1 we get (12). Now we suppose that instead of Assumption C the following condition holds. Assumption D. There exist numbers m∗ ∈ N, κ > 0, γ 6 0, ν ∈ R, α∗ ∈ R, λu ∈ R, λw ∈ R such that λu + λw = 1 q and the following assertions hold. p − 1 1. For any j′ > j > jmin and for any vertex ξ ∈ VA card VA j′−j(ξ) 6 c3 2. For any j > jmin, i ∈ Ij j−jmin(ξ0) (m∗j)γ log(m∗j)ν (m∗j′)γ log(m∗j′)ν . u(ηj,i) = uj = 2κm∗j(m∗j)α∗ log(m∗j)−λu, w(ηj,i) = wj = 2−κm∗j(m∗j)−α∗ log(m∗j)−λw. The partition {At,i}t>t0, i∈ Jt is defined as follows. Let t0 = min{t ∈ Z+ : 22t > m∗jmin}. (143) (144) Given t > t0, we denote by Γt the maximal subgraph in A on the vertex set {ηj,s : 22t−1 6 m∗j < 22t, s ∈ Ij}, and by At,i, i ∈ Jt, the connected components of Γt. Then card V(Γt) . Z0 2(1−γ)2t 2−νt. (145) Repeating the proof of Lemma 5.1 in [41] and taking into account that p < q, we get the following assertion. Lemma 8. Let D be a subtree in A, and let ξ∗ be its minimal vertex. Then kf − Pξ∗f kq Yq(D) . Z0 ∞ Xt=t0 2(1− q p )t Xi∈Jt,D 35 kf kq Xp(ΩDt,i ). (146) From Assumption B, (145) and (146) we obtain Assumptions 1, 2, 3. Remark 2. Suppose that Assumption B is replaced by the following condition: for any ξ ∈ V(A) the set F (ξ) is the atom of mes. Then the assertion of Theorem 1 holds as well (see Remark 1). 4 Estimates for entropy numbers of embeddings of weighted Sobolev spaces Let us define the weighted Sobolev class W r Lq,v(Ω). p,g(Ω) and the weighted Lebesgue space Let Ω ⊂ Rd be a bounded domain, and let g, v : Ω → (0, ∞) be measurable functions. For each measurable vector-valued function ψ : Ω → Rl, ψ = (ψk)16k6l, and for each p ∈ [1, ∞), we put max 16k6l kψkLp(Ω) =(cid:13)(cid:13)(cid:13) = ZΩ f defined on Ω we write ∇rf = (cid:16)∂rf /∂xβ(cid:17)β=r Let β = (β1, . . . , βd) ∈ Zd ψk(cid:13)(cid:13)(cid:13)p max 16k6l 1/p . ψk(x)p dx  + := (N ∪ {0})d, β = β1 + . . . + βd. For any distribution (here partial derivatives are taken in the sense of distributions), and denote by lr,d the number of components of the vector-valued distribution ∇rf . We set W r p,g(Ω) =(cid:8)f : Ω → R(cid:12)(cid:12) ∃ψ : Ω → Rlr,d : kψkLp(Ω) 6 1, ∇rf = g · ψ(cid:9) ∇rf (cid:16)we denote the corresponding function ψ by g (cid:17), kf kLq,v(Ω)=kf kq,v=kf vkLq(Ω), Lq,v(Ω) = {f : Ω → R kf kq,v < ∞} . We call the set W r ∇rf ∈ Lloc 1 (Ω). p,g(Ω) a weighted Sobolev class. Observe that if g ∈ Lloc p′ (Ω), then For x ∈ Rd and a > 0 we shall denote by Ba(x) the closed Euclidean ball of radius a in Rd centered at the point x. Definition 2. Let Ω ⊂ Rd be a bounded domain, and let a > 0. We say that Ω ∈ FC(a) if there exists a point x∗ ∈ Ω such that, for any x ∈ Ω, there exist a number T (x) > 0 and a curve γx : [0, T (x)] → Ω with the following properties: 1. γx ∈ AC[0, T (x)], (cid:12)(cid:12)(cid:12) 2. γx(0) = x, γx(T (x)) = x∗, dγx(t) dt (cid:12)(cid:12)(cid:12) = 1 a.e., 36 3. Bat(γx(t)) ⊂ Ω for any t ∈ [0, T (x)]. Definition 3. We say that Ω satisfies the John condition (and call Ω a John domain) if Ω ∈ FC(a) for some a > 0. For a bounded domain the John condition is equivalent to the flexible cone condition (see the definition in [2]). As examples of such domains we can take 1. domains with Lipschitz boundary; 2. the Koch's snowflake; 3. domains Ω = ∪0<t6T int Bct(γ(t)), where γ : [0, T ] → Rd is a curve with natural parametrization and c > 0. Domains with zero inner angles do not satisfy the John condition. We denote by H the set of all nondecreasing positive functions defined on (0, 1]. Definition 4. (see [5]). Let Γ ⊂ Rd be a nonempty compact set and h ∈ H. We say that Γ is an h-set if there are a constant c∗ > 1 and a finite countably additive measure µ on Rd such that supp µ = Γ and c−1 ∗ h(t) 6 µ(Bt(x)) 6 c∗h(t) (147) for any x ∈ Γ and t ∈ (0, 1]. Example 1. Let Γ ⊂ Rd be a Lipschitz manifold of dimension k, 0 6 k < d. Then Γ is an h-set with h(t) = tk. Example 2. Let Γ ⊂ R2 be the Koch curve. Then Γ is an h-set with h(t) = tlog 4/ log 3 (see [30, p. 66 -- 68]). Let Ω ∈ FC(a) be a bounded domain, and let Γ ⊂ ∂Ω be an h-set. Below we consider a function h ∈ H which has the following form near zero: h(t) = tθ log tγτ ( log t), 0 6 θ < d, where τ : (0, +∞) → (0, +∞) is an absolutely continuous function such that tτ ′(t) τ (t) → t→+∞ 0. (148) (149) Let 1 < p 6 ∞, 1 6 q < ∞, r ∈ N, δ := r + d q − d p > 0, βg, βv ∈ R, g(x) = ϕg(dist·(x, Γ)), v(x) = ϕv(dist·(x, Γ)), ϕg(t) = t−βg log t−αg ρg( log t), ϕv(t) = t−βv log t−αvρv( log t), (150) 37 where ρg and ρv are absolutely continuous functions such that tρ′ g(t) ρg(t) → t→+∞ 0, tρ′ v(t) ρv(t) → t→+∞ 0. Moreover, assume that βv < d − θ q or βv = d − θ q , αv > 1 − γ q . (151) (152) Without loss of generality we may consider Ω ⊂(cid:0)− 1 Denote β = βg + βv, α = αg + αv, 2 , 1 2(cid:1)d. ρ(y) = ρg(y)ρv(y), Z = (r, d, p, q, g, v, h, a, c∗), Z∗ = (Z, R), where c∗ is the constant from Definition 4 and R = diam Ω. Denote by Pr−1(Rd) the space of polynomials on Rd of degree not exceeding r−1. For a measurable set E ⊂ Rd we put Pr−1(E) = {f E : f ∈ Pr−1(Rd)}. Observe that W r p,g(Ω) ⊃ Pr−1(Ω). In Theorems 2, 3 the conditions on weights are such that W r p,g(Ω) ⊂ Lq,v(Ω) and there exist M > 0 and a linear continuous operator P : Lq,v(Ω) → Pr−1(Ω) such that for any function f ∈ W r p,g(Ω) (153) ∇rf g (cid:13)(cid:13)(cid:13)(cid:13)Lp(Ω) g (cid:13)(cid:13)(cid:13)Lp(Ω) p,g(Ω), W r kf − P f kLq,v(Ω) 6 M(cid:13)(cid:13)(cid:13)(cid:13) p,g(Ω) :=(cid:13)(cid:13)(cid:13) α0 :=(cid:26) α for βv < d−θ α − 1 q ∇rf for βv = d−θ q . q , (see [40 -- 43]). p,g(Ω) = span W r Denote W r p,g(Ω) be equipped with norm kf k W r W r p,g(Ω) = {f − P f : f ∈ W r . Denote by I : W r p,g(Ω)}. Let p,g(Ω) → Lq,v(Ω) the embedding operator. From (153) it follows that it is continuous. First we consider the case 0 < θ < d. We set In [37] the estimates for entropy numbers en(I : W r under the following conditions. In the case δ − β > θ(cid:16) 1 d 6= δ−β Now we obtain estimates for p < q, β − δ = 0, α0 = 1 θ . In the case δ − β = θ(cid:16) 1 p(cid:17)+ q − 1 δ p − 1 q . p,g(Ω) → Lq,v(Ω)) were obtained we assumed that q − 1 p(cid:17)+ we assumed that α0 6= 1 p − 1 q for p < q. 38 Theorem 2. Suppose that the conditions (148) -- (152) hold, ρg ≡ 1, ρv ≡ 1. Let 0 < θ < d, p < q, β = δ, and let α0 = 1 p − 1 q . Then en(I : W r p,g(Ω) → Lq,v(Ω)) ≍ Z∗ 1 q − 1 p . n Now we consider the case θ = 0, βv < d q . We assume that ρg(t) = log t−λg, ρv = log t−λv, τ (t) = log tν. (154) Denote λ = λg + λv. In [37] estimates of entropy numbers were obtained in the following cases: q − 1 , 1. δ − β > θ(cid:16) 1 2. δ − β = θ(cid:16) 1 3. δ − β = θ(cid:16) 1 p(cid:17)+ p(cid:17)+ p(cid:17)+ q − 1 q − 1 , α > (1 − γ)(cid:16) 1 , α = (1 − γ)(cid:16) 1 q − 1 q − 1 p(cid:17)+ p(cid:17)+ d 6= α , δ 1−γ , , and if p < q, then λ 6= 1 p − 1 q . Here we obtain the estimates for p < q, δ − β = 0, α = 0, λ = 1 p − 1 q . Theorem 3. Let the conditions (148), (150), (154) hold, and let θ = 0, β − δ = 0, βv < d q , α = 0, p < q, λ = 1 q . Then p − 1 en(I : W r p,g(Ω) → Lq,v(Ω)) ≍ Z∗ 1 q − 1 p . n Proof of Theorems 2 and 3. The lower estimates are proved similarly as in [37]. Let us obtain upper estimates. To this end, we apply Theorem 1. Consider the tree A with vertex set {ηj,i}j>jmin,i∈ Ij and the partition of the domain Ω into subdomains Ω[ηj,i], as defined in [40], [41]. We set F (ηj,i) = Ω[ηj,i]. Let the number s = s(a, d) > 4 be as defined in [40]. Repeating arguments from the proof of Theorem 2 in [41] (without summation over vertices ξ ∈ V(Aξ∗)), we obtain that Assumption A holds with u(ηj,i) = ϕg(2−sj) · 2−(r− d q . Assumption B holds with δ∗ = δ d (see [39], [36, Lemma 8], [1]). If θ > 0, then Assumption C holds; condition 5 of this assumption follows from Lemma 2 in [42]. If θ = 0, then Assumption D holds. In both cases we have κ = d q − βv, αu = αg, αw = αv. In the case θ = 0 we have λu = λg, λw = λv. By Theorem 1, we obtain the desired upper estimates of entropy numbers. p )sj, w(ηj,i) = ϕv(2−sj) · 2− dsj 39 5 Estimates for entropy numbers of two-weighted summation operators on a tree Applying Remark 2, we obtain estimates for entropy numbers of weighted summation operators on trees. Let G be a graph. Given a function f : V(G) → R, we set 1/p kf klp(G) =  Xξ∈V(G) f (ξ)p  for 1 6 p < ∞, kf kl∞(G) = sup ξ∈V(G) f (ξ). (155) Denote by lp(G) the space of functions f : V(G) → R with finite norm kf klp(G). Let A be a tree, V(A) = {ηj,i : j ∈ Z+, i ∈ Ij}, let η0,1 be the minimal vertex j (η0,1) = {ηj,i}i∈Ij for any j ∈ Z+. Suppose that for some c∗ > 1, of A, and let VA m∗ ∈ N c−1 ∗ h(2−m∗j) h(2−m∗(j+l)) 6 VA l (ηj,i) 6 c∗ h(2−m∗j) h(2−m∗(j+l)) , j, l ∈ Z+, where the function h ∈ H is defined by (148) near zero. Let u, w : V(A) → (0, ∞), u(ηj,i) = uj, w(ηj,i) = wj, j ∈ Z+, i ∈ Ij, uj = 2κm∗j(m∗j + 1)−αu, wj = 2−κm∗j(m∗j + 1)−αw. (156) In addition, we suppose that κ > θ q or κ = θ q , αw > 1 − γ q . (157) 1 q for κ = θ q . q , α = αu + αw − 1 We set Z = (p, q, u, w, h, m∗, c∗). Denote α = αu + αw for κ > θ In [37] estimates for en(Su,w,A : lp(A) → lq(A)) were obtained in the case α 6= q . The case α = 1 q , q = ∞ was considered by Lifshits and Linde [25, 27]. To u,w,A : l1(A) → lp′(A) p − 1 this end estimates for entropy numbers of the dual operator S∗ were obtained and the result of the paper [4] was applied. Theorem 4. Let θ > 0, 1 < p < q < ∞, α = 1 q . Then q − 1 p . p − 1 p − 1 en(Su,w,A : lp(A) → lq(A)) ≍ Z 1 n Now we suppose that θ = 0, κ > 0, and instead of (156) the following condition holds: uj = 2κm∗j(m∗j + 1)α[log(m∗j + 1)]−λu, wj = 2−κm∗j(m∗j + 1)−α[log(m∗j + 1)]−λw (158) with α ∈ R. Theorem 5. Let θ = 0, γ 6 0, 1 < p < q < ∞, κ > 0, λu + λw = 1 p − 1 q . Then en(Su,w,A : lp(A) → lq(A)) ≍ Z 1 q − 1 p . n 40 REFERENCES [1] O.V. Besov, "Kolmogorov widths of Sobolev classes on an irregular domain", Proc. Steklov Inst. Math., 280 (2013), 34-45. [2] O.V. Besov, V.P. Il'in, S.M. Nikol'skii, Integral representations of functions, and imbedding theorems. "Nauka", Moscow, 1996. [Winston, Washington DC; Wiley, New York, 1979]. [3] M.Sh. Birman and M.Z. Solomyak, "Piecewise polynomial approximations of functions of classes W α p ", Mat. Sb. 73:3 (1967), 331- -- 355. [4] J. Bourgain, A. Pajor, S. Szarek, N. Tomczak-Jaegermann, "On the duality problem for entropy numbers of operators", Geometric Aspects of Functional Analysis, Lecture Notes in Mathematics 1376, 50 -- 63. [5] M. Bricchi, "Existence and properties of h-sets", Georgian Mathematical Journal, 9:1 (2002), 13(cid:22)32. [6] B. Carl, I. Stephani, Entropy, Compactness, and the Approximation of Operators. Cambridge Tracts in Mathematics, V. 98. Cambridge: Cambridge University Press, 1990. [7] D.E. Edmunds, Yu.V. Netrusov, "Entropy numbers of operators acting between vector-valued sequence spaces", Math. Nachr., 286:5 -- 6 (2013), 614 -- 630. [8] D.E. Edmunds, Yu.V. Netrusov, "Schutt's theorem for vector-valued sequence spaces", J. Approx. Theory, 178 (2014), 13 -- 21. [9] D.E. Edmunds, H. Triebel, Function spaces, entropy numbers, differential in Mathematics, 120 (1996). Cambridge operators. Cambridge Tracts University Press. [10] D.D. Haroske, "Entropy numbers in weighted function spaces and eigenvalue distributions of some degenerate pseudodifferential operators. I", Math. Nachr., 167 (1994), 131 -- 156. [11] D.D. Haroske, "Entropy numbers in weighted function spaces and eigenvalue distributions of some degenerate pseudodifferential operators. II", Math. Nachr., 168 (1994), 109 -- 137. [12] D.D. Haroske, H. Triebel, "Wavelet bases and entropy numbers in weighted function spaces", Math. Nachr., 278:1 -- 2 (2005), 108 -- 132. [13] D.D. Haroske, L. Skrzypczak, "Entropy and approximation numbers of function spaces with Muckenhoupt weights", Rev. Mat. Complut., 21:1 (2008), 135 -- 177. "Entropy and approximation numbers of embeddings of function spaces with Muckenhoupt weights, II. General weights", Ann. Acad. Sci. Fenn. Math., 36:1 (2011), 111 -- 138. [14] D.D. Haroske, L. Skrzypczak, [15] D.D. Haroske, L. Skrzypczak, "Entropy numbers of embeddings of function spaces with Muckenhoupt weights, III. Some limiting cases," J. Funct. Spaces Appl. 9:2 (2011), 129 -- 178. 41 [16] A.N. Kolmogorov, V.M. Tikhomirov, "ε-entropy and ε-capacity of sets in function spaces" (Russian) Uspehi Mat. Nauk, 14:2(86) (1959), 3 -- 86. [17] T. Kuhn, "Entropy numbers of diagonal operators of logarithmic type", Georgian Math. J. 8:2 (2001), 307-318. [18] T. Kuhn, "A lower estimate for entropy numbers", J. Appr. Theory, 110 (2001), 120 -- 124. [19] T. Kuhn, "Entropy numbers of general diagonal operators", Rev. Mat. Complut., 18:2 (2005), 479 -- 491. [20] Th. Kuhn, H.-G. Leopold, W. Sickel, and L. Skrzypczak. "Entropy numbers of embeddings of weighted Besov spaces", Constr. Approx., 23 (2006), 61 -- 77. [21] Th. Kuhn, H.-G. Leopold, W. Sickel, L. Skrzypczak, "Entropy numbers of embeddings of weighted Besov spaces II", Proc. Edinburgh Math. Soc. (2) 49 (2006), 331 -- 359. [22] T. Kuhn, "Entropy Numbers in Weighted Function Spaces. The Case of Intermediate Weights", Proc. Steklov Inst. Math., 255 (2006), 159 -- 168. [23] Th. Kuhn, H.-G. Leopold, W. Sickel, and L. Skrzypczak, "Entropy numbers of embeddings of weighted Besov spaces III. Weights of logarithmic type", Math. Z., 255:1 (2007), 1 -- 15. [24] T. Kuhn, "Entropy numbers in sequence spaces with an application to weighted function spaces", J. Appr. Theory, 153 (2008), 40 -- 52. [25] M.A. Lifshits, "Bounds for entropy numbers for some critical operators", Trans. Amer. Math. Soc., 364:4 (2012), 1797 -- 1813. [26] M.A. Lifshits, W. Linde, "Compactness properties of weighted summation operators on trees", Studia Math., 202:1 (2011), 17 -- 47. [27] M.A. Lifshits, W. Linde, "Compactness properties of weighted summation operators on trees (cid:22) the critical case", Studia Math., 206:1 (2011), 75 -- 96. [28] M.A. Lifshits, W. Linde, "Approximation and entropy numbers of Volterra operators with application to Brownian motion", Mem. Amer. Math. Soc., 157:745, Amer. Math. Soc., Providence, RI, 2002. [29] E.N. Lomakina, V.D. Stepanov, "Asymptotic estimates for the approximation and entropy numbers of the one-weight Riemann -- Liouville operator", Mat. Tr., 9:1 (2006), 52 -- 100 [Siberian Adv. Math., 17:1 (2007), 1 -- 36]. [30] P. Mattila, Geometry of sets and measures in Euclidean spaces. Cambridge Univ. Press, 1995. [31] T. Mieth, "Entropy and approximation numbers of embeddings of weighted Sobolev spaces", J. Appr. Theory, 192 (2015), 250 -- 272. [32] A. Pietsch, Operator ideals. Mathematische Monographien [Mathematical Monographs], 16. Berlin, 1978. 451 pp. [33] C. Schutt, "Entropy numbers of diagonal operators between symmetric Banach spaces", J. Appr. Theory, 40 (1984), 121 -- 128. 42 [34] V.M. Tikhomirov, "The ε-entropy of certain classes of periodic functions" (Russian) Uspehi Mat. Nauk 17:6 (108) (1962), 163 -- 169. [35] H. Triebel, "Entropy and approximation numbers of limiting embeddings, an approach via Hardy inequalities and quadratic forms", J. Approx. Theory, 164:1 (2012), 31 -- 46. [36] A.A. Vasil'eva, "Widths of function classes on sets with tree-like structure", J. Appr. Theory, 192 (2015), 19 -- 59. [37] A.A. Vasil'eva, "Entropy numbers of embedding operators of weighted Sobolev spaces with weights that are functions of distance from some h-set", arXiv:1503.00144. [38] A.A. Vasil'eva, "Estimates for norms of two-weighted summation operators on a tree under some restrictions on weights", Math. Nachr. 1 -- 24 (2015) /DOI 10.1002/mana.201300355. [39] A.A. Vasil'eva, "Widths of weighted Sobolev classes on a John domain", Proc. Steklov Inst. Math., 280 (2013), 91 -- 119. [40] A.A. Vasil'eva, "Embedding theorem for weighted Sobolev classes on a John domain with weights that are functions of the distance to some h-set", Russ. J. Math. Phys., 20:3 (2013), 360 -- 373. [41] A.A. Vasil'eva, "Embedding theorem for weighted Sobolev classes on a John domain with weights that are functions of the distance to some h-set", Russ. J. Math. Phys. 21:1 (2014), 112 -- 122. [42] A.A. Vasil'eva, "Some sufficient conditions for embedding a weighted Sobolev class on a John domain", Sib. Mat. Zh., 56:1 (2015), 65 -- 81. [43] A.A. Vasil'eva, "Widths of weighted Sobolev classes with weights that are functions of the distance to some h-set: some limit cases", Russ. J. Math. Phys., 22:1 (2015), 127 -- 140. [44] A.A. Vasil'eva, "Kolmogorov and linear widths of the weighted Besov classes with singularity at the origin", J. Appr. Theory, 167 (2013), 1 -- 41. 43
1605.01543
1
1605
2016-05-05T10:06:45
Grothendieck's inequality in the noncommutative Schwartz space
[ "math.FA", "math.OA" ]
In the spirit of Grothendieck's famous inequality from the theory of Banach spaces, we study a sequence of inequalities for the noncommutative Schwartz space, a Fr\'echet algebra of smooth operators. These hold in non-optimal form by a simple nuclearity argument. We obtain optimal versions and reformulate the inequalities in several different ways.
math.FA
math
GROTHENDIECK'S INEQUALITY IN THE NONCOMMUTATIVE SCHWARTZ SPACE RUPERT H. LEVENE AND KRZYSZTOF PISZCZEK Abstract. In the spirit of Grothendieck's famous inequality from the theory of Banach spaces, we study a sequence of inequalities for the noncommutative Schwartz space, a Fr´echet algebra of smooth operators. These hold in non-optimal form by a simple nuclear- ity argument. We obtain optimal versions and reformulate the inequalities in several different ways. 1. Introduction The noncommutative Schwartz space S is a weakly amenable m- convex Fr´echet algebra whose properties have been investigated in sev- eral recent papers, see e.g. [2,3,13,14]. It is not difficult to see that as a Fr´echet space, S is nuclear. From this, we can easily deduce the follow- ing analogue of Grothendieck's inequality, which we call Grothendieck's inequality in S: there exists a constant K > 0 so that for any contin- uous bilinear form u : S × S → C and any n ∈ N, there exists k ∈ N such that for every m ∈ N and any x1, . . . , xm, y1, . . . , ym ∈ S, we have (1) (cid:12)(cid:12)(cid:12) u(xj, yj)(cid:12)(cid:12)(cid:12) 6 Kkuk∗ n k(xj)kRC k k(yj)kRC k m X j=1 The norms on the right hand side arise naturally from the definition of S, as explained in Section 2 below. Our goal in this note is to show that in fact k = 2n + 1 always suffices, and that this is best possible. This appears to be the first result concerning Grothendieck's inequal- ity in the category of Fr´echet algebras; to the best of our knowledge, 2010 Mathematics Subject Classification. Primary: 47A30, 47A07, 47L10. Sec- ondary: 47A63. Key words and phrases. noncommutative Grothendieck inequality, noncommu- tative Schwartz space, m-convex Fr´echet algebra, bilinear form, state. The research of the second-named author has been supported by the National Center of Science, Poland, grant no. UMO-2013/10/A/ST1/00091. The authors are grateful to the anonymous referee, whose comments greatly improved the presentation of this paper. 1 2 RUPERT H. LEVENE AND KRZYSZTOF PISZCZEK all previous results along these lines concern Banach spaces (includ- ing C∗-algebras, general Banach algebras and operator spaces). For Fr´echet algebras, Grothendieck's inequality seems to have a specific flavour. Every Fr´echet space (and a fortiori, every Fr´echet algebra) which appears naturally in analysis is nuclear, meaning that all tensor product topologies are equal. Since Grothendieck's inequality can be understood as the equivalence of two tensor products, it seems that we can take inequality (1) for granted. The interesting question that remains is then optimality. This paper is divided into four sections. In the remainder of this section, we recall a C∗-algebraic version of Grothendieck's inequality due to Haagerup, and then review the definition and the basic prop- erties of S which we require. In Section 2 we explain how nuclearity gives Grothendieck's inequality in S, and we estimate the constants K and k. Section 3 then settles the optimality question for k via a matricial construction. We conclude with a short section containing several reformulations of the inequality. 1.1. Grothendieck's inequality. Pisier's survey article [12] is a com- prehensive reference for Grothendieck's inequality. This presents many equivalent formulations and applications of this famous result, and re- counts its evolution from 'commutative' [5] to 'noncommutative'. Of these reformulations and extensions, Haagerup's noncommutative ver- sion most closely resembles (1), and we state it here for the convenience of the reader. Theorem 1 ([6], [12, Theorem 7.1]). Let A and B be C∗-algebras. For any bounded bilinear form u : A × B → C and any finite sequence (xj, yj) in A × B, we have X u(xj, yj)(cid:12)(cid:12)(cid:12) where k(xj)kRC := max(cid:8)(cid:13)(cid:13)P x∗ 6 2kuk k(xj)kRC k(yj)kRC j xj(cid:13)(cid:13) (cid:12)(cid:12)(cid:12) 2(cid:9). 1.2. The noncommutative Schwartz space. Let j(cid:13)(cid:13) 2 ,(cid:13)(cid:13)P xjx∗ 1 1 s = nξ = (ξj)j∈N ∈ CN : ξn := (cid:16) +∞ X j=1 1 2 ξj2j2n(cid:17) < +∞ for all n ∈ No denote the so-called space of rapidly decreasing sequences. This space becomes Fr´echet when endowed with the sequence ( · n)n∈N of norms defined above. The basis (Un)n∈N of zero neighbourhoods of s is defined by Un := {ξ ∈ s : ξn 6 1}. The topological dual s′ of s is the so-called GROTHENDIECK'S INEQUALITY IN S 3 space of slowly increasing sequences, namely +∞ X j=1 1 2 n := (cid:16) ηj2j −2n(cid:17) < +∞ for some n ∈ No nη = (ηj)j∈N ∈ CN : η′ where the duality pairing is given by hξ, ηi := Pj∈N ξjηj for ξ ∈ s, η ∈ s′. The noncommutative Schwartz space S is the Fr´echet space L(s′, s) of all continuous linear operators from s′ into s, endowed with the topology of uniform convergence on bounded sets. The formal identity map ι : s ֒→ s′ is a continuous embedding and defines a product on S by xy := x ◦ ι ◦ y for x, y ∈ S. There is also a natural involution on S given by hx∗ξ, ηi := hξ, xηi for x ∈ S, ξ, η ∈ s′. With these operations, S becomes an m-convex Fr´echet ∗-algebra. The inclusion map S ֒→ K(ℓ2) is continuous, and in fact it is a spectrum-preserving ∗-homomorphism [3]. Moreover [13, Proposition 3], an element x ∈ S is positive (i.e., x = y∗y for some y ∈ S), if and only if the spectrum of x is contained in [0, +∞), or equivalently hxξ, ξi ≥ 0 for all ξ ∈ s′. On the other hand, by [3, Cor. 2.4] and [4, Theorems 8.2, 8.3], the topology of S cannot be given by a sequence of C∗-norms. This causes some technical inconvenience (e.g. there is no bounded approximate identity in S) meaning we cannot apply C∗-algebraic techniques directly. 2. The inequality Let (k · kn)n∈N be a non-decreasing sequence of norms which gives the topology of S. For u : S × S → C a continuous bilinear form, we write kuk∗ n := sup{u(x, y) : x, y ∈ Un} where Un = {x ∈ S : kxkn 6 1}; similarly, for a functional φ ∈ S ′, we write kφk∗ n := sup{φ(x) : x ∈ Un}. Following Pisier [11, p. 316], for k ∈ N and x1, x2, . . . , xm ∈ S, we write k(xj)kRC k = maxn(cid:13)(cid:13)(cid:13) m X j=1 1 2 k x∗ j xj(cid:13)(cid:13)(cid:13) ,(cid:13)(cid:13)(cid:13) m X j=1 xjx∗ 1 2 j(cid:13)(cid:13)(cid:13) ko. Relative to our choice of norms k·kn, we have now defined each term in our hoped-for inequality (1). We will now reformulate it using tensor products. For C∗-algebras, such a reformulation is standard. Indeed, by [6, Theorem 1.1] (formulated along the lines of [8, Theorem 2.1]), Haagerup's noncommutative Grothendieck inequality entails the existence of a K > 4 RUPERT H. LEVENE AND KRZYSZTOF PISZCZEK 0 such that for any C∗-algebras A, B and z in the algebraic tensor prod- uct A⊗ B, we have kzkπ 6 Kkzkah where k·kπ is the projective tensor norm and k · kah is the absolute Haagerup tensor norm [8, p. 164] on A ⊗ B, given by m X j=1 xj2(cid:13)(cid:13)(cid:13) 1 2(cid:13)(cid:13)(cid:13) m X j=1 1 2 . yj2(cid:13)(cid:13)(cid:13) kzkah = inf (cid:13)(cid:13)(cid:13) 2(x∗x + xx∗)(cid:1) 1 Here x = (cid:0) 1 2 for x an element of a C∗-algebra, and the infimum is taken over all representations z = Pm j=1 xj ⊗ yj where (xj, yj) ∈ A × B. We proceed similarly for S. For x ∈ S, let x2 = 1 2(x∗x + xx∗) ∈ S and consider the sequence of absolute Haagerup tensor norms (k · kah,n)n∈N on the algebraic tensor product S ⊗ S given by kzkah,n := inf (cid:13)(cid:13)(cid:13) m X j=1 1 2 n(cid:13)(cid:13)(cid:13) xj2(cid:13)(cid:13)(cid:13) m X j=1 1 2 n yj2(cid:13)(cid:13)(cid:13) where the infimum runs over all ways to represent z = Pm j=1 xj ⊗ yj in S ⊗ S. As usual, we write (k · kπ,n)n∈N for the sequence of projective tensor norms on S ⊗ S. Just as in the C∗-algebra case, inequality (1) will follow once we show that the sequences of projective and absolute Haagerup tensor norms are equivalent on S ⊗S. In fact, the equivalence of these norms follows immediately from the nuclearity of S (see [9, Theorem 28.15] and [7, Ch. 21, §2, Theorem 1] for details). On the other hand, the optimal values of k and K (depending on n and our choice of norms (k·kn)n∈N) for which (1) hold are not given by such general considerations. These optimal parameters will be denoted by κ(n) := kbest and Kn := Kbest. Henceforth, we focus only on the sequence of norms (k·kn)n∈N where n = {ξ ∈ s′ : kxkn := sup{xξn : ξ ∈ U ◦ n}, and U ◦ In other words, kxkn is the norm n := ℓ2((j −n)j) of x ∈ S, considered as a Hilbert space operator from H ′ to Hn := ℓ2((jn)j). This sequence does indeed induce the topology of S. In this context, we will estimate Kn and compute the exact values of κ(n). n ∈ N, x ∈ S ξ′ n 6 1}. We start with the following result, which can be compared with [13, Lemma 8]. To fix some useful notation, for n ∈ N we define an infinite diagonal matrix dn := diag(1n, 2n, 3n, 4n, . . . ) which we consider as an isometry dn : ℓ2 → H ′ n and simultaneously as an isometry dn : Hn → ℓ2. GROTHENDIECK'S INEQUALITY IN S 5 Proposition 2. Let n ∈ N. We have (i) kxkn = sup{hxξ, ξi : ξ ∈ U ◦ (ii) kxk2 (iii) kxk2 n 6 kx2k2n for every self-adjoint x ∈ S; and n 6 kx∗xk 2n for every x ∈ S. Moreover, inequalities (ii) and (iii) are sharp. n} for every positive x ∈ S; 2nkxx∗k 1 2 1 2 Proof. (i) Observe that kxkn = kdnxdnkB(ℓ2). Furthermore, since x is positive, dnxdn is positive and we have kxkn = kdnxdnkB(ℓ2) = sup{hxdnξ, dnξi : ξℓ2 6 1} = sup{hxξ, ξi : ξ′ (ii) For x self-adjoint, we have n 6 1}. kx2k2n = kd2nx2d2nkB(ℓ2) = kd2nxk2 B(ℓ2), and by [1, Proposition II.1.4.2], kxkn = kdnxdnkB(ℓ2) = ν(dnxdn) = ν(d2nx) 6 kd2nxkB(ℓ2), where ν(·) denotes the spectral radius. This gives the desired inequal- ity. (iii) Since S ֒→ B(ℓ2), any x ∈ S is also a Hilbert space operator, (x∗x)1/2 i is positive in B(ℓ2⊕ ℓ2) and the block-matrix operator h (xx∗)1/2 x∗ (see e.g. [10, p. 117]). Equivalently, x ∀ ξ, η ∈ ℓ2. hxξ, ηi2 6 h(xx∗)1/2η, ηih(x∗x)1/2ξ, ξi (2) For m ∈ N, let us write pm := [ Im 0 matrix. Now fix n ∈ N and choose ξ, η ∈ H ′ all m ∈ N and (2) gives 0 0 ] where Im ∈ Mm is the identity n. Then pmξ, pmη ∈ ℓ2 for hpmxpmξ, ηi2 6 hpm(xx∗)1/2pmη, ηihpm(x∗x)1/2pmξ, ξi. Since (pm)m∈N is an approximate identity in S (see [13, Proposition 2]), we obtain hxξ, ηi2 6 h(xx∗)1/2η, ηih(x∗x)1/2ξ, ξi. Taking the supremum over all ξ, η in the unit ball of H ′ n we get kxk2 n 6 k(xx∗)1/2knk(x∗x)1/2kn. Applying (ii) to the positive operators (xx∗)1/2 and (x∗x)1/2 we con- clude that kxk2 then we have equality in both (ii) and (iii). For sharpness, observe that if x is a diagonal rank one matrix unit (cid:3) n 6 kx∗xk 2nkxx∗k 2n. 1 2 1 2 6 RUPERT H. LEVENE AND KRZYSZTOF PISZCZEK Proposition 3. For any n, m ∈ N and x1, . . . , xm, y1, . . . , ym ∈ S, we have m 6 1 4 1 4 m m m j=1 j=1 x∗ π2 X X X kxjknkyjkn j yj(cid:13)(cid:13)(cid:13) 2n+1(cid:13)(cid:13)(cid:13) j xj(cid:13)(cid:13)(cid:13) 6 (cid:13)(cid:13)(cid:13) 6 and let p ∈ N. We claim that kxk(cid:13)(cid:13)(cid:13)p+1 X kxkkp 6 C(cid:13)(cid:13)(cid:13) kx∗ j(cid:13)(cid:13)(cid:13) 2n+1(cid:13)(cid:13)(cid:13) X X xjx∗ x∗ y∗ k=1 k=1 j=1 j=1 m m 1 4 Proof. Let C := π2 . 2n+1(cid:13)(cid:13)(cid:13) m X j=1 yjy∗ 1 4 j(cid:13)(cid:13)(cid:13) . 2n+1 By the Cauchy -- Schwarz inequality and Proposition 2(iii) this will then imply the desired inequality. To establish the claim, let ξ1, . . . , ξm ∈ U ◦ p and let us write (ej)j∈N for the standard basis vectors in ℓ2. We have m X k=1 m +∞ hx∗ kxkξk, ξki = X k=1 X i,j=1 hx∗ kxkej, eii(ij)pξk i i−pξk j j −p. Applying the Cauchy -- Schwarz inequality to summation over i, j ∈ N gives m X k=1 hx∗ kxkξk, ξki 6 m X k=1 (cid:16) +∞ X i,j=1 hx∗ kxkej, eii2(ij)2p(cid:17) 1 2 . Since x∗x is positive for any x ∈ S, and for positive operators y ∈ S we have yij2 6 yiiyjj (where yij := hyej, eii), this implies that m X k=1 hx∗ kxkξk, ξki 6 X j=1 +∞ +∞ D(cid:16) m X k=1 kxk(cid:17)jpej, jpejE x∗ m By Proposition 2(i), for any ε > 0 there are ξ1, . . . , ξm ∈ U ◦ kxk(cid:13)(cid:13)(cid:13)p+1 + ε. kxkξk, ξki + ε < C(cid:13)(cid:13)(cid:13) hx∗ kxkkp < kx∗ X X X x∗ k=1 k=1 k=1 m m p with Taking the infimum over ε > 0 yields the claim. (cid:3) kxk(cid:17)ip+1ei, ip+1eiE x∗ j −2 sup 6 j=1 X 6 C(cid:13)(cid:13)(cid:13) m X k=1 m k=1 i∈N D(cid:16) X kxk(cid:13)(cid:13)(cid:13)p+1 . x∗ GROTHENDIECK'S INEQUALITY IN S 7 As a straightforward consequence of Proposition 3, we obtain: Theorem 4 (Grothendieck's inequality in S). There is a constant K 6 π2 6 such that kzkπ,n 6 2Kkzkah,2n+1 for any n ∈ N and z ∈ S ⊗ S. Moreover, every continuous bilinear form u : S × S → C satisfies inequality (1) with k = 2n + 1, for any n, m ∈ N and any x1, . . . , xm, y1, . . . , ym ∈ S. In particular, taking u(x, y) := φ(x)φ(y) where φ ∈ S ′, we obtain (3) φ(xj)2 6 K(kφk∗ n k(xj)kRC 2n+1)2 X m j=1 Remark. This shows that κ(n) 6 2n + 1. On the other hand, it is easy to show that κ(n) > 2n − 1. Indeed, if not, then (3) would hold with 2n + 1 replaced by some ℓ 6 2n − 1. Take m ∈ N, define ξm := Pm j=1 jnej and φm ∈ S ′ by φm(x) := hxξm, ξmi. Then for xj := ejj, j = 1, . . . , m we get k(xj)kRC n = m. On the other hand, Pm j=1 φm(xj)2 is equivalent (up to a constant) to m4n+1. Therefore (3) takes the form m4n+1 6 Cm2ℓ+2 for some constant C (independent of m). Letting m tend to infinity, we obtain ℓ > 2n − 1 2, a contradiction. Hence κ(n) ∈ {2n, 2n + 1}. 3. Optimality ℓ = mℓ and kφmk∗ We will now show that κ(n) = 2n+1. For this, we will use the tensor product formulation, noting that κ(n) = minnk ∈ N : supn kzkπ,n kzkah,k : z ∈ S ⊗ S, z 6= 0o < ∞o. Recall the diagonal operator dn defined on page 4 above. Since every x−→ x ∈ S is an operator on ℓ2 via the canonical inclusions ℓ2 dn֒→ ℓ2, it is clear that if x ∈ S, then dnx and xdn are both operators Hn on ℓ2. This leads to the following observation. Proposition 5. If z = Pk kzkπ,n = (cid:13)(cid:13)(cid:13) j=1 xj ⊗ yj ∈ S ⊗ S, then dnxjdn ⊗ dnyjdn(cid:13)(cid:13)(cid:13)π dn֒→ H ′ X j=1 n . k Proof. Write ∆nz = k X j=1 dnxjdn ⊗ dnyjdn ∈ B(ℓ2) ⊗ B(ℓ2). 8 RUPERT H. LEVENE AND KRZYSZTOF PISZCZEK If ∆nz = Pm for some ε > 0, then z = Pm n knkd−1 kzkπ,n 6 l=1 al ⊗ bl ∈ B(ℓ2)⊗B(ℓ2) and Pm n ⊗ d−1 X n kn = n ald−1 n bld−1 kd−1 l=1 d−1 n ald−1 X m l=1 l=1 kalkkblk < k∆nzkπ + ε n bld−1 m n and kalkkblk < k∆nzkπ + ε. l=1 This gives kzkπ,n 6 k∆nzkπ. The reverse inequality is proved similarly. (cid:3) We also need the following well-known fact. Proposition 6. If H is a Hilbert space and x1, . . . , xm ∈ B(H), then m X j=1 (cid:13)(cid:13)(cid:13) j(cid:13)(cid:13)(cid:13)h xj ⊗ x∗ = (cid:13)(cid:13)(cid:13) m X j=1 xjx∗ . j(cid:13)(cid:13)(cid:13) j=1 xjax∗ Proof. By [15, Theorem 4.3], the Haagerup norm on the left hand side is equal to the completely bounded norm of the map on B(H) given by a 7→ Pm j , which is completely positive, so attains its com- pletely bounded norm at the identity operator. (cid:3) Theorem 7. For every n ∈ N, we have κ(n) = 2n + 1. Proof. By Theorem 4, it only remains to show that κ(n) > 2n. Choose kn ∈ N sufficiently large that k + 1 6 2k(k 4n − 1) for all k > kn. This inequality ensures that for every k > kn, if we define 1 i1 = 2k, ik+1 = ⌊k 1 4n 2k⌋, ij = ik+1 + j − (k + 1), 2 6 j 6 k, then i1 < i2 < · · · < ik+1. Denote by (eij)i,j∈N the standard matrix units, and for j = 2, . . . , k + 1, consider the self-adjoint operators Let xj := ei1,ij + eij ,i1 ∈ Mik+1 ⊂ S. zk := k+1 X j=2 xj ⊗ xj. Since dnxjdn = in by Propositions 5 and 6 we obtain 1 in j (ei1,ij + eij ,i1) and (dnxjdn)2 = i2n 1 i2n j (ei1,i1 + eij ,ij ), kzkkπ,n = (cid:13)(cid:13)(cid:13) > (cid:13)(cid:13)(cid:13) k+1 X j=2 k+1 X j=2 dnxjdn ⊗ dnxjdn(cid:13)(cid:13)(cid:13)π dnxjdn ⊗ dnxjdn(cid:13)(cid:13)(cid:13)h k+1 X j=2 = (cid:13)(cid:13)(cid:13) (dnxjdn)2(cid:13)(cid:13)(cid:13) = i2n 1 k+1 X j=2 i2n j . GROTHENDIECK'S INEQUALITY IN S 9 On the other hand, xj2 = x2 j = ei1,i1 + eij ,ij and k+1 X j=2 Therefore d2nx2 j d2n = i4n 1 kei1,i1 + k+1 X j=2 i4n j eij ,ij . kzkkah,2n 6 (cid:13)(cid:13)(cid:13) 6 i4n k+1 xj2(cid:13)(cid:13)(cid:13)2n X j=2 1 k + i4n k+1. k+1 X j=2 = (cid:13)(cid:13)(cid:13) d2nx2 j d2n(cid:13)(cid:13)(cid:13) = max{i4n 1 k, i4n k+1} Hence kzkkπ,n kzkkah,2n > 1 Pk+1 j=2 i2n i2n j i4n 1 k + i4n k+1 > i−2n 1 i2n 2 1 + k−1i−4n 1 k+1 → ∞ as k → ∞, i4n by our choice of i1, . . . , ik+1. So κ(n) > 2n as required. (cid:3) 4. Reformulations of the inequality Here we give several different ways of stating our inequality; in each case, an analogous result may be found in [12]. The methods here are fairly standard, so full proofs are often omitted. Throughout, we write K = supn∈N Kn 6 π2/6. 4.1. Grothendieck's inequality with states. Given ξ ∈ U ◦ n, let φξ ∈ S ′ be given by φξ(x) = hxξ, ξi, x ∈ S. We call an element of the closed convex hull of {φξ : ξ ∈ U ◦ n} an n-state on S. Note that by Proposition 2(i), for any positive element x ∈ S we have kxkn = sup{φ(x) : φ ∈ Vn}, where Vn ⊆ S ′ is the set of all n-states on S. The next result may be deduced from Theorem 4 by closely following the Hahn -- Banach Separation argument of [12, §23]. Theorem 8. For any continuous bilinear form u : S × S → C and n ∈ N, there are (2n + 1)-states φ1, φ2, ψ1, ψ2 on S with u(x, y) 6 Kkuk∗ n(cid:0)φ1(x∗x) + φ2(xx∗)(cid:1) 1 2(cid:0)ψ1(y∗y) + ψ2(yy∗)(cid:1) 1 2 for all x, y ∈ S. 4.2. 'Little' Grothendieck inequality. As a consequence we ob- tain the following 'little' Grothendieck inequality in S. Recall that if T : X → Y is a linear map between Fr´echet spaces, then kTkn,k := sup{kT xkk : kxkn 6 1}. 10 RUPERT H. LEVENE AND KRZYSZTOF PISZCZEK Theorem 9. For any Fr´echet-Hilbert space H, if u1, u2 : S → H are continuous linear maps, k, m, n ∈ N and x1, . . . , xm, y1, . . . , ym ∈ S, then m X j=1 (cid:12)(cid:12)(cid:12) hu1(xj), u2(yj)ik(cid:12)(cid:12)(cid:12) 6 Kku1kn,k ku2kn,k k(xj)kRC 2n+1 k(yj)kRC 2n+1. Equivalently, for any k, n ∈ N there are (2n + 1)-states φ1, φ2, ψ1, ψ2 such that for all x, y ∈ S we have hu1(x), u2(y)ik 6 Kku1kn,kku2kn,k 1 × (cid:0)φ1(x∗x) + φ2(xx∗)(cid:1) 2(cid:0)ψ1(y∗y) + ψ2(yy∗)(cid:1) 1 2 . Proof. Apply Theorems 4 and 8 to uk(x, y) := hu1(x), u2(y)ik for k ∈ N. (cid:3) Using the same argument as in the proof of Theorem 8 we can obtain an equivalent version of the 'little' Grothendieck inequality. Theorem 10. For any Fr´echet-Hilbert space H, if u : S → H is a continuous linear map and k, n ∈ N, then there exist (2n + 1)-states φ1, φ2 on S such that for all x ∈ S we have kuxkk 6 √Kkukn,k(cid:0)φ1(x∗x) + φ2(xx∗)(cid:1) 1 2 . References [1] B. Blackadar. Operator algebras, volume 122 of Encyclopaedia of Mathematical Sciences. Springer-Verlag, Berlin, 2006. [2] T. Cia´s. On the algebra of smooth operators. Studia Math., 218(2):145 -- 166, 2013. [3] P. Doma´nski. Algebra of smooth operators. http://main3.amu.edu.pl/~domanski/salgebra1.pdf, 2012. [4] M. Fragoulopoulou. Topological algebras with involution. Elsevier Science B.V., Amsterdam, 2005. [5] A. Grothendieck. R´esum´e de la th´eorie m´etrique des produits ten- soriels topologiques. Resenhas, 2(4):401 -- 480, 1996. Reprint of Bol. Soc. Mat. Sao Paulo 8 (1953), 1 -- 79. [6] U. Haagerup. The Grothendieck inequality for bilinear forms on C ∗-algebras. Adv. in Math., 56(2):93 -- 116, 1985. [7] H. Jarchow. Locally convex spaces. B. G. Teubner, 1981. GROTHENDIECK'S INEQUALITY IN S 11 [8] S. Kaijser and A.M. Sinclair. Projective tensor products of C ∗- algebras. Math. Scand., 55(2):161 -- 187, 1984. [9] R. Meise and D. Vogt. Introduction to functional analysis, vol- ume 2 of Oxford Graduate Texts in Mathematics. Oxford Univer- sity Press, 1997. [10] V. Paulsen. Completely bounded maps and operator algebras. Cam- bridge University Press, Cambridge, 2002. [11] G. Pisier. Introduction to operator space theory, volume 294 of Lon- don Mathematical Society Lecture Note Series. Cambridge Univer- sity Press, Cambridge, 2003. [12] G. Pisier. Grothendieck's theorem, past and present. Bull. Amer. Math. Soc. (N.S.), 49(2):237 -- 323, 2012. [13] K. Piszczek. Automatic continuity and amenability in the non- commutative Schwartz space. J. Math. Anal. Appl., 432(2):954 -- 964, 2015. [14] K. Piszczek. A Jordan-like decomposition in the noncommutative Schwartz space. Bull. Aust. Math. Soc., 91(2):322 -- 330, 2015. [15] R. R. Smith. Completely bounded module maps and the Haagerup tensor product. J. Funct. Anal., 102(1):156 -- 175, 1991. School of Mathematics and Statistics, University College Dublin, Belfield, Dublin 4, Ireland E-mail address: [email protected] Faculty of Mathematics and Comp. Sci., A. Mickiewicz University in Pozna´n, Umultowska 87, 61-614 Pozna´n, Poland E-mail address: [email protected]
1501.04513
1
1501
2015-01-19T15:10:25
Integral inequalities for infimal convolution and Hamilton-Jacobi equations
[ "math.FA" ]
Let $f,g:\Bbb{R}^{N}\rightarrow (-\infty ,\infty ]$ be Borel measurable, bounded below and such that $\inf f+\inf g\geq 0.$ We prove that with $ m_{f,g}:=(\inf f-\inf g)/2,$ the inequality $||(f-m_{f,g})^{-1}||_{\phi }+||(g+m_{f,g})^{-1}||_{\phi }\leq 4||(f\Box g)^{-1}||_{\phi }$ holds in every Orlicz space $L_{\phi },$ where $f\Box g$ denotes the infimal convolution of $f$ and $g$ and where $||\cdot ||_{\phi }$ is the Luxemburg norm (i.e., the $L^{p}$ norm when $L_{\phi }=L^{p}$). Although no genuine reverse inequality can hold in any generality, we also prove that such reverse inequalities do exist in the form $||(f\Box g)^{-1}||_{\phi }\leq 2^{N-1}(||(\check{f}-m_{f,g})^{-1}||_{\phi }+||(\check{ g}+m_{f,g})^{-1}||_{\phi }),$ where $\check{f}$ and $\check{g}$ are suitable transforms of $f$ and $g$ introduced in the paper and reminiscent of, yet very different from, nondecreasing rearrangement. Similar inequalities are proved for other extremal operations and applications are given to the long-time behavior of the solutions of the Hamilton-Jacobi and related equations.
math.FA
math
INTEGRAL INEQUALITIES FOR INFIMAL CONVOLUTION AND HAMILTON-JACOBI EQUATIONS PATRICK J. RABIER This paper is dedicated to the memory of Jean Jacques Moreau Abstract. Let f, g : RN → (−∞, ∞] be Borel measurable, bounded below and such that inf f + inf g ≥ 0. We prove that with mf,g := (inf f − inf g)/2, the inequality (f − mf,g )−1φ + (g + mf,g )−1φ ≤ 4(f (cid:3)g)−1φ holds in every Orlicz space Lφ, where f (cid:3)g denotes the infimal convolution of f and g and where · φ is the Luxemburg norm (i.e., the Lp norm when Lφ = Lp). Although no genuine reverse inequality can hold in any generality, we also prove that such reverse inequalities do exist in the form (f (cid:3)g)−1φ ≤ 2N−1(( f − mf,g )−1φ + (g + mf,g )−1φ), where f and g are suitable trans- forms of f and g introduced in the paper and reminiscent of, yet very different from, nondecreasing rearrangement. Similar inequalities are proved for other extremal operations and applica- tions are given to the long-time behavior of the solutions of the Hamilton- Jacobi and related equations. 1. Introduction If f, g : RN → (−∞, ∞], the infimal convolution f (cid:3)g : RN → [−∞, ∞], first introduced by Fenchel [9] and Moreau [24], [25], [26], is defined by the formula (f (cid:3)g)(x) := inf y∈RN (f (x − y) + g(y)). Since then, this operation and its extension to general vector spaces have found an ever growing variety of applications, including convex functions [13], [29], extension of Lipschitz functions [12], solutions of the Hamilton-Jacobi equations [2], [20], [31] and much more (even a proof of the Hahn-Banach theorem [11]). In fact, there are by now several thousands publications using infimal convolution in areas as diverse as image processing, economics and finance, information theory, probabilities and statistics, etc. For a glimpse into some of these problems, see the excellent recent survey by Lucet [21]. In this paper, we investigate the mathematical properties of infimal convolution in a new direction, by exploring the existence of integral inequalities involving f, g and f (cid:3)g. The remark that f (cid:3)g = 0 whenever f ≥ 0 and g ≥ 0 are integrable could cast serious doubts on the value of this program, but they are quickly dispelled by the rebuttal that no similar triviality arises from the integrability of f −1 and g−1. Here and everywhere else, f −1 := 1/f, g−1 := 1/g, etc. This notation will not be used to denote any set-theoretic inverse. 1991 Mathematics Subject Classification. 26D15, 46E30, 35F25, 49L25. Key words and phrases. Brunn-Minkowski inequality, enclosing ball, Hamilton-Jacobi equa- tions, infimal convolution, Orlicz space, rearrangement. 1 2 PATRICK J. RABIER Omitting technicalities to which we shall return shortly, the first batch of in- equalities will relate the (Luxemburg) norm (f (cid:3)g)−1φ in any Orlicz space Lφ, to the norms (f − z)−1φ and (g + z)−1φ for a suitable constant z independent of φ, to be defined in due time. The only restrictions are that f and g must be Borel measurable, bounded below and that f (cid:3)g ≥ 0. The proofs depend crucially upon (a slightly weaker form of) the Brunn-Minkowski inequality. The setting of Orlicz spaces instead of just the classical Lp spaces introduces only mild additional technicalities, is more natural in many respects and, as we shall see in the examples of Section 7, is useful in some applications. It does not even require any knowledge of Orlicz spaces beyond the definitions of Young functions and of the Luxemburg norm, which will both be reviewed. This being said, a simple special case asserts that if f, g ≥ 0 are Borel measurable and inf f = inf g (see Theorem 3.4 for a full and much more general statement) f −1p + g−1p ≤ 4(f (cid:3)g)−1p, (1.1) for every 1 ≤ p ≤ ∞, where · p is the norm of Lp := Lp(RN ). The constant 4 is best possible among all constants independent of p, as is readily seen when f = g = 1 and p = ∞. The Borel measurability requirement has to do with the measurability of f (cid:3)g, without which (1.1) cannot make sense. Curiously, we were unable to find a dis- cussion of the measurability properties of the infimal convolution in the classical literature, but the evidence points to the fact that f and g Lebesgue measurable does not suffice for the measurability of f (cid:3)g. Indeed, as is well-known, the strict epigraph of f (cid:3)g is the (vector, also called Minkowski) sum of the strict epigraphs of f and g and Sierpi´nski [30] showed, almost a century ago, that the sum of two Lebesgue measurable sets need not be Lebesgue measurable. In contrast, the sum of two Borel sets is always Lebesgue measurable (but not always a Borel set). See Section 2 for further details. A peculiar feature of (1.1) and of more general similar inequalities is that only the left-hand side is unchanged by modifications of f and g on null sets, as long as Borel measurability and inf f = inf g > −∞ are preserved. Most of the paper is actually devoted to perhaps more important -and definitely more delicate- reverse inequalities which, in a simpler world, would read (1.2) (f (cid:3)g)−1p ≤ C(f −1p + g−1p), with C > 0 independent of f and g in some suitable class of nonnegative functions. Unfortunately, the main obstacle to (1.2) is that no remotely general converse of the Brunn-Minkowski inequality holds in any form, even for convex sets. Such a converse is actually trivially true for Euclidean balls, but a direct application of this remark only yields (1.2) for a narrow subclass of radially symmetric functions. To take advantage of the converse of the Brunn-Minkowski inequality for balls in a much broader setting, we introduce a new function transform, strongly remi- niscent of, yet very different from, nonincreasing rearrangement. The difference is that the upper level sets are rounded before being rearranged, the rounding being performed by using the concept of enclosing ball (see Section 4). To each function f : RN → [−∞, ∞] (no measurability needed), the aforemen- tioned transform associates a measurable radially symmetric function f , which in turn produces another measurable radially symmetric function f := −(−f). In the special case when f, g ≥ 0 are Borel measurable and inf f = inf g (see Theorem INEQUALITIES FOR INFIMAL CONVOLUTION 3 6.1 for a full and much more general statement), the reverse inequality in Lp reads (compare with (1.1) under the same assumptions) (1.3) (f (cid:3)g)−1p ≤ 2N −1( f −1p + g−1p). Such an inequality breaks down completely if f and g are replaced with f and g, respectively, even if both functions are radially symmetric. For example, if N = 1, f (x) = x2 + 1 and g(x) = x2 + 1 when x /∈ Q, g(x) = 1 if x ∈ Q, then f and g are Borel measurable and inf f = inf g = 1. But (f (cid:3)g)−1 = 1/2 is in no Lp space with p < ∞, whereas f −1 and g−1 = f −1 a.e. are in all of them. (In this example, it turns out that f = f but g = 1. ) This example also shows that, unlike in (1.1), neither side of (1.3) is independent of modifications of f or g on null sets that do not affect Borel measurability or inf f = inf g. When not trivial (i.e., f = f a.e.), the explicit calculation of f is generally not possible. Nevertheless, the inequality (1.3) is useful because some simple and general conditions about f and g ensure the finiteness of the right-hand side (Lemma 6.3). There is certainly more to be discovered in that regard. The proofs of the inequalities involve two other classical extremal operations (f ⊼ g)(x) := sup y min{f (x − y), g(y)} and (f ⊻ g)(x) := inf y max{f (x − y), g(y)}. Either of these operations fully determines the other, but both notations will be useful. In general, f ⊻ g = −(−f ) ⊼ (−g) and, for nonnegative functions, f ⊻ g = (f −1⊼g−1)−1 will be important. We also prove inequalities similar to (1.1) and (1.3) for the operations ⊼ and ⊻ (both being often referred to as "level sum" operations in the literature). In fact, a good part of the work will consist in proving integral inequalities for ⊼, from which those for (cid:3) and ⊻ will be derived. In the last section, the inequalities are used to obtain Lp (and other) estimates for the inverses of solutions of the Hamilton-Jacobi equations and variants thereof. Throughout the paper, µN denotes the N -dimensional Lebesgue measure and, without a qualifier, measurability always means Lebesgue measurability. 2. Background The purpose of this short section is to review the basic properties of the opera- tions mentioned in the Introduction, to set the notation used in future sections and to settle basic measurability issues. Recall that if X and Y are subsets of RN , their sum X + Y is defined by X + Y := (cid:26) {x + y : x ∈ X, y ∈ Y } if X 6= ∅ and Y 6= ∅, ∅ if X = ∅ or Y = ∅. The following key lemma is well-known. The "proof" below merely makes the connection with the deep property behind it. Lemma 2.1. If X and Y are Borel subsets of RN , their sum X + Y is measurable1. Proof. In Euclidean space (any dimension), the continuous image of a Borel set is Lebesgue measurable; see Federer [8, p. 69]. Since X × Y is a Borel subset of RN × RN and the addition is continuous on RN , the result follows. (cid:3) 1Even a Suslin set, but not necessarily a Borel set. 4 PATRICK J. RABIER Given two functions f, g : RN → [−∞, ∞] and ξ ∈ R, call F + ξ and G+ ξ the upper level sets (2.1) F + ξ := {x ∈ RN : f (x) > ξ}, G+ ξ := {x ∈ RN : g(x) > ξ} and call W + ξ (2.2) the corresponding upper level set of f ⊼ g : W + ξ := {x ∈ RN : (f ⊼ g)(x) > ξ}. It is a standard elementary property that ξ = F + (2.3) W + ξ + G+ ξ . By Lemma 2.1 and since f ⊻ g = −(−f ) ⊼ (−g), it follows at once from (2.3) that: Lemma 2.2. If f, g : RN → [−∞, ∞] are Borel measurable, then f ⊼ g and f ⊻ g are measurable. If f : RN → [−∞, ∞] we set (2.4) Mf := sup f, mf := inf f. The next relations are elementary, but important (2.5) Mf ⊼g = min{Mf , Mg}, mf ⊼g ≥ min{mf , mg}, Mf ⊻g ≤ max{Mf , Mg}, mf ⊻g = max{mf , mg}. We now turn to infimal convolution. Given a function f : RN → (−∞, ∞], we denote by Ef := {(x, ξ) ∈ RN × R : f (x) < ξ} the strict epigraph of f. It is also a simple well-known property that if g : RN → (−∞, ∞] is another function, then Ef (cid:3)g = Ef + Eg. Since a function is Borel measurable (measurable) if and only if its strict epigraph is a Borel set (measurable), it follows from Lemma 2.1 that Lemma 2.3. If f, g : RN → (−∞, ∞] are Borel measurable, then f (cid:3)g is measur- able. For future use, we also note that if z ∈ R, (2.6) f (cid:3)g = (f − z)(cid:3)(g + z). 3. First integral inequalities The Brunn-Minkowski inequality (see e.g. Gardner's survey [10]) asserts that if X, Y are nonempty measurable subsets of RN and if X + Y is measurable, then µN (X + Y )1/N ≥ µN (X)1/N + µN (Y )1/N . Obviously, it fails if X or Y is empty and the other has positive measure. We shall only need the less sharp form (3.1) µN (X + Y ) ≥ µN (X) + µN (Y ), if X, Y and X + Y are measurable and X 6= ∅, Y 6= ∅. Lemma 3.1. If f, g : RN → [0, ∞] are measurable and Mf = Mg (possibly ∞; see (2.4)) and if f ⊼ g is measurable, then (3.2) ZRN f +ZRN g ≤ ZRN f ⊼ g. INEQUALITIES FOR INFIMAL CONVOLUTION 5 Proof. Set M := Mf = Mg ≤ ∞. If M = 0, then f = g = f ⊼ g = 0 and (3.2) is trivial. In what follows, M > 0. Since f ≥ 0, it is well-known (see (2.1)) that RRN f = R ∞ ξ = ∅ when ξ ≥ M, this reads RRN f = R M 0 µN (F + ing F + R M 0 (µN (F + ξ ) + µN (G+ 0 µN (G+ By (2.2) and (2.3) and since f ⊼ g is measurable by hypothesis, F + ξ )dξ, so that RRN f +RRN g = R M ξ ))dξ. 0 µN (F + ξ )dξ. By us- ξ )dξ. Likewise, RRN g = ξ = W + ξ ξ 6= ∅ by definition of ξ ). Therefore, 0 µN (W + ξ )dξ (by (2.5), the second inequality ξ )dξ = RRN f ⊼ g since f ⊼ g ≥ 0 and the proof 6= ∅ and G+ ξ + G+ ξ ) = µN (W + ξ ) ≤ µN (F + ξ + G+ (cid:3) is measurable for every ξ. If ξ < M, then F + ξ M and so, by (3.1), µN (F + RRN f +RRN g ≤ R M 0 µN (W + is even an equality). Now, R ∞ ξ ) + µN (G+ ξ )dξ ≤ R ∞ 0 µN (W + is complete. Of course, (3.2) does not follow from a pointwise inequality. The condition Mf = Mg cannot be dropped. For example, if f > 0 and g = 0, then f ⊼ 0 = 0 and (3.2) fails. Lemma 3.1 is just the stepping stone for much more general inequalities. Recall that in the theory of Orlicz spaces, a nonconstant function φ : [0, ∞] → [0, ∞] is called a Young function if φ(0) = 0 and φ is nondecreasing, convex and left continuous ([27]; see also [1], [18] for a simplified treatment limited to N -functions). In particular, φ(∞) = ∞. Remark 3.1. If φ is a Young function and h : RN → [0, ∞] is measurable, the monotonicity of φ shows at once that φ(h) is measurable. If φ is a Young function, the corresponding Orlicz space Lφ consists of all the measurable functions h on RN such that RRN φ (λh) < ∞ for some λ > 0 (this makes sense by Remark 3.1). It is a (complete) normed space for the Luxemburg norm · φ defined by (3.3) hφ := inf (cid:26)r > 0 : ZRN φ(r−1h) ≤ 1(cid:27) . Since the right-hand side of (3.3) is finite if and only if h ∈ Lφ, it will always be understood that hφ = ∞ when h is measurable and h /∈ Lφ. Thus, h ∈ Lφ is equivalent to hφ < ∞. Furthermore, it is readily checked that (3.4) hφ ≤ kφ if h ≤ k and, by the left-continuity of φ and monotone convergence2, that if h ∈ Lφ, (3.5) ZRN φ(h−1 φ h) ≤ 1. If φ(τ ) := τ p for some 1 ≤ p < ∞, then hφ = hp. On the other hand, hφ = h∞ when φ is the indicator function of [0, 1] (φ = 0 in [0, 1] and ∞ outside). Lemma 3.2. If φ is a Young function and if h : RN → [0, ∞], then (see (2.4)) Mφ(h) = φ(Mh). 2This is of course a well-known inequality. 6 PATRICK J. RABIER Proof. It is plain that h ≤ Mh implies φ(h) ≤ φ(Mh), so that Mφ(h) ≤ φ(Mh). It only remains to show that φ(Mh) ≤ Mφ(h), which is trivial if Mh = 0. We henceforth assume Mh > 0. By the monotonicity of φ and φ(0) = 0, there is τ 1 ∈ [0, ∞] such that φ = ∞ on (τ 1, ∞] and that φ < ∞ on [0, τ 1). Specifically, τ 1 = sup{τ ≥ 0 : φ(τ ) < ∞}. If τ 1 ∈ (0, ∞), then φ(τ 1) may be finite or infinite. We split the proof into three cases. (i) Mh > τ 1. If so, τ 1 < ∞ and then φ(Mh) = ∞. The set {x ∈ RN : h(x) > τ 1} is not empty and φ(h(x)) = ∞ for every x in that set. Thus, {x ∈ RN : φ(h(x)) = ∞} 6= ∅, so that Mφ(h) = ∞ = φ(Mh). (ii) Mh = τ 1 = ∞. Then, φ(Mh) = φ(∞) = ∞. If τ > 0 is finite, ∅ 6= {x ∈ RN : h(x) > τ } ⊂ {x ∈ RN : φ(h(x)) ≥ φ(τ )}(by the monotonicity of φ). As a result, Mφ(h) ≥ φ(τ ). By letting τ → ∞ = τ 1 and since limτ→∞ φ(τ ) = φ(∞) = ∞ by the left continuity of φ, it follows that Mφ(h) = ∞ = φ(Mh). (iii) 0 < Mh ≤ τ 1. If Mh = ∞, then τ 1 = ∞ and (ii) above applies. Assume now Mh < ∞. For every ε > 0, S := {x ∈ RN : h(x) > Mh − ε} 6= ∅. If ε is small enough, then Mh − ε > 0 and S ⊂ {x ∈ RN : φ(h(x)) ≥ φ(Mh − ε)} by the monotonicity of φ. Hence, Mφ(h) ≥ φ(Mh − ε). Since φ is left continuous, Mφ(h) ≥ φ(Mh). (cid:3) From Lemma 3.1 and Lemma 3.2, we obtain: Lemma 3.3. Let φ : [0, ∞] → [0, ∞] be a Young function. If f, g : RN → [0, ∞] are Borel measurable and if Mf = Mg (possibly ∞), then f ⊼ g is measurable and (3.6) max{f φ, gφ} ≤ f ⊼ gφ. Proof. By Lemma 2.2, f ⊼ g is measurable and so, by Remark 3.1, φ(f ), φ(g) and φ(f ⊼ g) are measurable. Since φ(min{f (y), g(x − y)}) = min{φ(f (y)), φ(g(x − y))} by the monotonicity of φ, we infer that supy φ(min{f (y), g(x − y)}) = (φ(f ) ⊼ φ(g))(x). By Lemma 3.2 with h(y) := min{f (y), g(x − y)}, the left-hand side is φ((f ⊼g)(x)), so that φ(f ⊼g) = φ(f ) ⊼φ(g). In particular, φ(f ) ⊼φ(g) is measurable. Since Mf = Mg, then Mφ(f ) = Mφ(g), once again by Lemma 3.2. Thus, from the above and from Lemma 3.1 with f and g replaced with φ(f ) and φ(g), respectively, (3.7) ZRN φ(f ) +ZRN φ(g) ≤ ZRN φ(f ⊼ g). If r > 0, then r−1(f ⊼ g) = r−1f ⊼ r−1g and Mr−1f = Mr−1g. Thus, (3.7) for r−1f and r−1g yields RRN φ(r−1f ) ≤ RRN φ(r−1(f ⊼ g)), so that f φ ≤ f ⊼ gφ by (3.3). Likewise, gφ ≤ f ⊼ gφ and (3.6) follows. (cid:3) Of course, when Lφ = L1, Lemma 3.1 yields the stronger f 1 + g1 ≤ f ⊼ g1 but (3.6) is optimal when φ is arbitrary (let f = g = 1 and Lφ = L∞). We are now in a position to prove our first main integral inequality for infimal convolution. Recall once more the notation (2.4). Theorem 3.4. Suppose that f, g : RN → (−∞, ∞] are Borel measurable, that mf , mg ∈ R and that mf + mg ≥ 0. Set (3.8) mf,g = (mf − mg)/2. Then, f − mf,g, g + mf,g and f (cid:3)g are measurable and nonnegative and (3.9) (f − mf,g)−1φ + (g + mf,g)−1φ ≤ 4(f (cid:3)g)−1φ, INEQUALITIES FOR INFIMAL CONVOLUTION 7 for every Young function φ. Proof. The measurability of f (cid:3)g was established in Lemma 2.3. Next, inf(f − mf,g) = inf(g +mf,g) = (mf +mg)/2 ≥ 0, whence f (cid:3)g = (f −mf,g)(cid:3)(g +mf,g) ≥ 0 by (2.6). This also implies (f −mf,g)−1 ≥ 0, (g+mf,g)−1 ≥ 0 and sup(f −mf,g)−1 = 2(mf + mg)−1 = sup(g + mf,g)−1. Therefore, the inequality (3.6) is applicable in the form (f − mf,g)−1φ + (g + mf,g)−1φ ≤ 2(f − mf,g)−1 ⊼ (g + mf,g)−1φ and (3.9) follows from 0 ≤ h−1 ⊼ k−1 = (h ⊻ k)−1 ≤ 2(h(cid:3)k)−1 when h and k are nonnegative, from f (cid:3)g = (f − mf,g)(cid:3)(g + mf,g) ≥ 0 and from (3.4). (cid:3) It was noted in the Introduction that the constant 4 in (3.9) is already best possible when Lφ = L∞. Remark 3.2. Theorem 3.4 gives a simple necessary condition for the existence of solutions of infimal convolution equations (see [23], [21] and the references therein): Suppose that h ≥ 0 is measurable and that g is Borel measurable and bounded below. If h−1 ∈ Lφ for some Orlicz space Lφ and (g − mg + mh/2)−1φ > 4h−1φ (in particular, if (g − mg + mh/2)−1 /∈ Lφ), the equation f (cid:3)g = h has no Borel measurable solution f. Indeed, if f exists, then mf = mh − mg ∈ R and (3.9) cannot hold. If z 6= 0 is a constant, there is no simple pointwise relationship between (f − z) ⊻ (g + z) and f ⊻ g. As a result, the method of proof of Theorem 3.4 does not yield a variant of (3.5) or (3.9) with f (cid:3)g replaced with f ⊻ g. However, if f, g ≥ 0, then 0 ≤ f ⊻ g ≤ f (cid:3)g and such a variant can be obtained as a straightforward corollary of Theorem 3.4: Corollary 3.5. Suppose that f, g : RN → [−∞, ∞] are Borel measurable, that g ≥ 0 and that f 6 ≡∞ and g 6 ≡∞, so that 0 ≤ mf + + mg < ∞, where f+ := max{f, 0}. Then, f+ − mf+,g, g + mf+,g (see (3.8 )) and f ⊻ g are measurable and nonnegative and (3.10) (f+ − mf +,g)−1φ + (g + mf+,g)−1φ ≤ 4(f ⊻ g)−1φ, for every Young function φ. Proof. Since g ≥ 0, it follows that 0 ≤ f ⊻ g = f+ ⊻ g ≤ f+(cid:3)g. By Lemma 2.2, f ⊻ g = f+ ⊻ g is measurable. Therefore, the corollary follows from (3.4) and from Theorem 3.4 for f+ and g. (cid:3) The constant 4 is also best possible in (3.10) (among constants independent of φ): If f = 1 and g = ℓ > 1 is constant, the inequality for the L∞ norm is 4(ℓ + 1)−1 ≤ 4ℓ−1. By letting ℓ → ∞, it follows that, in the right-hand side, 4 cannot be lowered. 4. The radial transforms f and f The proof of Lemma 3.1 shows that the existence of a converse of the inequality (3.2), that is, ZRN f ⊼ g ≤ C(cid:18)ZRN f +ZRN g(cid:19) , 8 PATRICK J. RABIER with C > 0 independent of f and g would require µN (F + ξ ) + µN (G+ ξ )) for every ξ > 0. However, as pointed out in the Introduction, no converse of the Brunn-Minkowski inequality or its weaker form (3.1) holds in any generality. The transforms defined in this section will enable us (in the next section) to take advantage of the fact that such a converse trivially exists when X and Y are Euclidean balls. The thought that this case is so special that it cannot have any broad value would result in a serious oversight. ξ ) ≤ C(µN (F + ξ + G+ By a classical theorem of Jung [8, p. 200], [17], every nonempty bounded subset X of RN is contained in a unique closed ball BX with minimal diameter among all closed balls containing X, called the enclosing ball of X. If X is unbounded, no closed ball contains X and we set BX := RN . Lastly, if X = ∅, every singleton {x} satisfies the "minimal diameter" requirement, whence uniqueness, but not existence, is lost. For definiteness, we arbitrarily set B∅ := {0}. Evidently, X ⊂ BX in all cases. Jung's theorem also provides the estimate diam(X) ≤ diam(BX ) ≤ p2N/(N + 1) diam(X), but we shall only make use of the (trivial) first one. Remark 4.1. An easily overlooked aspect of enclosing balls is that X ⊂ Y implies only µN (BX ) ≤ µN (BY ) but not BX ⊂ BY , unless N = 1. The following property of enclosing balls will be important. Lemma 4.1. If Xn is a nondecreasing sequence of subsets of RN , then µN (B∪Xn ) = lim µN (BXn ) = sup µN (BXn ). Proof. Set X := ∪Xn. If X is unbounded, then BX = RN and so µN (BX ) = ∞. Since X := ∪Xn and Xn ⊂ Xn+1, the diameter of Xn and, hence, that of BXn , tends to ∞. Accordingly, lim µN (BXn ) = ∞. Suppose now that X is bounded, so that BX is a ball. Since Xn ⊂ X im- plies µN (BXn ) ≤ µN (BX ) and since µN (BXn ) is nondecreasing, it is plain that lim µN (BXn ) ≤ µN (BX ). To prove the converse, call rn ≥ 0 the radius of BXn . The sequence rn is nondecreasing and bounded above (by the radius of BX ) and so it has a limit r ≥ rn for every n. As a result, lim µN (BXn ) is the measure of any ball with radius r. Next, call xn the center of BXn . By a simple contradiction argument, the se- quence xn is bounded (since xn might not be in BX -see Remark 4.1- this is not totally trivial). After extracting a subsequence, assume that xn → x ∈ RN . Ev- ery y ∈ X is in Xn for n large enough. Since BXn ⊂ B(xn, r), it follows that y ∈ B(x, r). Thus, X ⊂ B(x, r), whence µN (B(x, r)) ≥ µN (BX ) by definition of BX . Since r = lim rn amounts to µN (B(x, r)) = lim µN (BXn ), it follows that lim µN (BXn ) ≥ µN (BX ). (cid:3) Given any function f : RN → [−∞, ∞], we now proceed to constructing a measurable radially symmetric function f : RN → [−∞, ∞] whose upper level sets F + ) for every ξ. The construction follows that of the nonincreasing rearrangement of f. ξ have measure equal to µN (BF + ξ Lemma 4.2. The function µN (BF + [−∞, ∞]. ξ ) is nonincreasing and right-continuous on INEQUALITIES FOR INFIMAL CONVOLUTION 9 η ⊂ F + Proof. If ξ < η, then F + continuity, let ξn ց ξ, so that F + lim µN (BF + ). ξn ξ , so that µN (BF + ξ = ∪F + ξn η ) ≤ µN (BF + ξ ). For the right and then, by Lemma 4.1, µN (BF + ξ ) = (cid:3) Call ρ+ f (ξ) the radius of BF + follows from Lemma 4.2 that ρ+ ξ . Since ρ+ f (ξ) is proportional to µN (BF + )1/N , it ξ f is nonincreasing and right-continuous. Therefore, (4.1) γ+ f (t) := inf{ξ : ρ+ f (ξ) ≤ t}, is a nonincreasing and right-continuous function on [0, ∞) and (4.2) {t ≥ 0 : γ+ f (t) > ξ} = [0, ρ+ f (ξ) is replaced with µN (F + f (ξ)). ξ ), γ+ Indeed, when f ≥ 0 and ρ+ f becomes the nonincreas- ing rearrangement of f and these properties follow uniquely from the monotonicity and right-continuity of µN (F + ξ ); see for instance [33, pp. 26-27]. We also point out that in most modern expositions, the nonincreasing rearrangement of a function f is defined to be that of f . This has not always been the case (see Day [7] or Lux- emburg [22]) and the monotonicity and right-continuity properties of nonincreasing rearrangements are independent of whether f or f is used in their definition. We now set (4.3) f (x) := γ+ f (x). Some basic properties of f are summarized in the next theorem. ξ ξ ) = µN (BF + ξ denotes the upper ξ- ) for every ξ ∈ [−∞, ∞], where F + Theorem 4.3. Given f : RN → [−∞, ∞], the function f has the following prop- erties: (i) f is measurable and f = f a.e. if and only if f (x) is a.e. equal to a nonincreas- ing function of x. If also f (x) is a right-continuous function of x, then f = f. (ii) µN ( F + level set of f . (iii) M f ≤ Mf and m f ≥ mf (in particular, f ≥ 0 ⇒ f ≥ 0). Furthermore, M f = ess sup f and m f = ess inf f . (iv) (f + z) = f + z for z ∈ R and (f (c·)) = f (c·) for c ∈ R\{0}. (v) (cf) = c f for every c ≥ 0. (vi) If h : RN → [−∞, ∞] and h ≤ f, then h ≤ f . (vii) If f is bounded below on bounded subsets and limx→∞ f (x) = −∞ and if h : RN → [−∞, ∞] satisfies h(x) ≤ f (x) for x large enough, then h(x) ≤ f (x) for x large enough. Furthermore, if f (x) is a strictly decreasing function of x and if h(x) ≤ f (x) when x /∈ B for some open ball B centered at the origin, then h(x) ≤ f (x) (= f (x) by (i)) for every x /∈ B. Proof. (i) The measurability of f follows at once from the monotonicity of γ+ f and the necessity of the given conditions for f = f a.e. is obvious. Conversely, if f (x) = γ(x) with γ : [0, ∞) → [−∞, ∞] nonincreasing, the upper level sets of f are balls centered at the origin (possibly RN ) and ρ+ f is the distribution function of γ, so that γ+ f = γ∗, the nonincreasing rearrangement of γ. Since γ∗ = γ except perhaps at the countably many points of discontinuity of γ, it follows that f = f 10 PATRICK J. RABIER a.e. Clearly, this remains true if f (x) = γ(x) a.e. If γ is right-continuous, then γ∗ = γ and f = f. (ii) Just notice that, by (4.2) and (4.3), F + ξ is the open ball with center 0 and radius ρ+ f (ξ). f = γ+ f (0) = inf{ξ : ρ+ (iii) With no loss of generality, assume Mf < ∞. Since γ+ f (ξ) = 0}. If ξ ≥ Mf , then F + f is nonincreasing, = {0} ξ ) = 0} ≤ Mf . On the other hand, f = M f . This shows that M f ≤ Mf . Furthermore, M f = ess sup f f (ξ) = 0. Thus, max γ+ max γ+ and so ρ+ by (4.3), max γ+ by (4.3) and the right-continuity of γ+ f . f = inf{ξ : µN (F + ξ = ∅, whence BF + ξ That m f ≥ mf is obvious if mf = −∞, or if mf = ∞ (for then f = f = ∞). If f ≥ mf by (4.1) and so f ≥ mf , mf ∈ R and ξ < mf , then ρ+ whence m f ≥ mf . That m f = ess inf f follows from the monotonicity of γ+ f . f +z(ξ) = ρ+ f (ξ − z), f +z, i.e., (f +z) = f +z. The proofs that f (c·) = f (c·) (iv) Since {x ∈ RN : f (x) + z > ξ} = F + ξ−z, it follows that ρ+ f (ξ) = ∞. Thus, γ+ which in turn yields γ+ if c ∈ R\{0} is equally straightforward. f +z = γ+ (v) Since this is trivial when c = 0, assume c > 0. Then, ρ+ cf (ξ) = ρ+ f (ξ/c), whence γ+ cf = cγ+ f , i.e. (cf) = c f . (vi) If h ≤ f, then (with a self-explanatory notation) H + h ≤ γ+ f (ξ) and, by (4.1), γ+ ). Hence, ρ+ h (ξ) ≤ ρ+ ξ ⊂ F + f , so that h ≤ f . µN (BF + ξ and so µN (BH+ ) ≤ ξ ξ (vii) Choose an open ball B centered at the origin such that h(x) ≤ f (x) when x /∈ B. Since f is bounded below on bounded subsets, inf B f is finite and, if for every ξ ≤ ξ0. By (vi), h is increased when h is ξ0 < inf B f, then B ⊂ F + ξ increased. Thus, if it can be shown that h ≤ f after increasing h, this inequality also holds before h is increased. In particular, we may increase h on B so that ξ0 < inf B h and then B ⊂ H + for every ξ ξ ≤ ξ0. On the other hand, {x /∈ B : h(x) > ξ} ⊂ {x /∈ B : f (x) > ξ} since h ≤ f on RN \B. Altogether, if ξ ≤ ξ0, then H + f (ξ). As a result, for every ξ ≤ ξ0. Thus, B ⊂ H + ξ and so ρ+ h (ξ) ≤ ρ+ ξ ⊂ F + ξ ∩ F + ξ (4.4) γ+ h (t) := inf{ξ : ρ+ h (ξ) ≤ t} ≤ inf{ξ ≤ ξ0 : ρ+ h (ξ) ≤ t} ≤ Since limx→∞ f (x) = −∞, the level set F + ξ0 f (ξ) ≤ t}. inf{ξ ≤ ξ0 : ρ+ is bounded, whence ρ+ f (ξ0) < ∞. f (ξ0) ≤ t by the monotonicity f (ξ) ≤ t}. By (4.4), f (ξ0), it follows that h(x) ≤ f (x) Choose any t ≥ ρ+ of ρ+ f , so that γ+ γ+ h (t) ≤ γ+ when x ≥ ρ+ f (ξ0). f (ξ0). If ξ > ξ0, then ρ+ f (t) := inf{ξ : ρ+ f (ξ) ≤ ρ+ f (ξ) ≤ t} = inf{ξ ≤ ξ0 : ρ+ f (t). Since this is true for every t ≥ ρ+ To complete the proof, assume in addition that f (x) is a strictly decreasing function of x. We show that h(x) ≤ f (x) when x /∈ B. If B = ∅, the result follows from (v). From now on, assume B 6= ∅ (hence B 6= {0} as well since B is open). By the monotonicity of f in x, inf tB f < inf B f if t > 1. It follows that F + ξ0 ⊂ tB f (ξ0) ≤ tρ if t > 1 and ξ0 above is close enough to inf B f. Thus, BF + ξ0 where ρ is the radius of B. From the above, h(x) ≤ f (x) when x ≥ tρ and, hence, when x ≥ ρ by first letting t → 1 (which gives only x > ρ) and next using the ⊂ tB, so that ρ+ INEQUALITIES FOR INFIMAL CONVOLUTION 11 right-continuity of f and h with respect to x. Since ρ is the radius of B and B is centered at the origin, this means that h(x) ≤ f (x) when x /∈ B. (cid:3) Even when f ≥ 0, it is not true that f = 0 implies f = 0. The following characterization is important for the proof of the reverse inequalities (specifically, of Theorem 5.5 later). Lemma 4.4. If f : RN → [0, ∞] and f = 0, then either f = 0 or there are x0 ∈ RN and 0 < z ≤ ∞ such that f (x) = 0 if x 6= x0 and f (x0) = z. f is nonincreasing, f = 0 means ρ+ has radius 0, which happens only when F + Proof. By (4.1) and (4.3) and since ρ+ f (0) = 0, 0 = ∅ or i.e., that the enclosing ball BF + 0 = {x0} is a singleton. If F + when F + 0 = ∅, then f = 0 since f ≥ 0. Suppose now that F + 0 = {x0}. This means that x0 is the only point where f is positive, so that f (x0) = z > 0 (possibly ∞) and that f (x) ≤ 0 when x 6= x0. Since f ≥ 0, it follows that f (x) = 0 when x 6= x0. (cid:3) 0 Remark 4.2. By the argument of the above proof, the inequality M f ≤ Mf in Theorem 4.3 (iii) can be made precise: M f < Mf if and only if some upper level set of f is a singleton. Put differently, M f = Mf if and only if there is a sequence of distinct points xn such that f (xn) → Mf . We shall also need the transform defined by f := −(−f), (4.5) directly given by the formula (4.6) where (4.7) f (x) = γ− f (x), γ− f (t) := sup{ξ : ρ− f (ξ) ≤ t} and ρ− f (ξ) is the radius of BF − ξ with, of course, F − ξ = {x ∈ RN : f (x) < ξ}. ξ ) = µN (BF − ξ ) for every ξ ∈ [−∞, ∞], where F − Theorem 4.5. Given f : RN → [−∞, ∞], the function f has the following prop- erties: (i) f is measurable and f = f a.e. if and only if f (x) is a.e. equal to a nondecreas- ing function of x. If also f (x) is a right-continuous function of x, then f = f . (ii) µN ( F − ξ denotes the lower ξ- level set of f . (iii) M f ≤ Mf and m f ≥ mf (in particular, f ≥ 0 ⇒ f ≥ 0). Furthermore, M f = ess sup f and m f = ess inf f . (iv) (f + z) = f + z for every z ∈ R and (f (c·)) = f (c·) for c ∈ R\{0}. (v) (cf) = c f for every c ≥ 0 (and (cf) = c f for every c < 0). (vi) If h : RN → [−∞, ∞] and h ≤ f, then h ≤ f . (vii) If h : RN → [−∞, ∞] is bounded above on bounded subsets, limx→∞ h(x) = ∞ and h(x) ≤ f (x) for x large enough, then h(x) ≤ f (x) for x large enough. Fur- thermore, if h(x) is a strictly increasing function of x and if h(x) ≤ f (x) when x /∈ B for some open ball B centered at the origin, then h(x) (= h(x) by (i)) ≤ f (x) for every x /∈ B. 12 PATRICK J. RABIER (viii) If f ≥ 0, then (f −1) = ( f )−1. (ix) (f+ ) = f+. Proof. Parts (i) to (vii) follow at once from (4.5) and from the corresponding prop- erties in Theorem 4.3. f (ξ) = ∞ and ρ− f (t) := inf{ξ ≥ 0 : ρ+ (viii) First, note that if f ≥ 0 and ξ < 0, then ρ+ f (ξ) = 0. Hence, by (4.1) and (4.7), γ+ f (t) := sup{ξ ≥ 0 : ρ− f (ξ) ≤ t} ≥ 0. Upon replacing f with f −1 in the former formula, we get (since passing from f to f −1 changes upper level sets into lower ones) γ+ f −1 (t) := inf{ξ ≥ f (ξ−1) ≤ t}. On the other hand, since inf(S−1) = (sup S)−1 for every subset 0 : ρ− f (ξ−1) ≤ S ⊂ [0, ∞], we have (γ− t}. This shows that γ+ f (t))−1 = inf{ξ−1 ≥ 0 : ρ− f −1 = (γ− f )−1 and the result follows from (4.3) and (4.6). f (ξ) ≤ t} = inf{ξ ≥ 0 : ρ− f (ξ) ≤ t} and γ− (ix) The lower level sets {x ∈ RN : f+(x) < ξ} are empty if ξ ≤ 0 and coincide f (ξ) if ξ > 0. Therefore, (ξ) = 0 if ξ ≤ 0 and ρ− f+ if ξ > 0. Thus, ρ− f+ (ξ) = ρ− with F − ξ by (4.7), (4.8) γ− f+ (t) = max{sup{ξ ≤ 0 : ρ− f+ (ξ) ≤ t}, sup{ξ > 0 : ρ− f+ (ξ) ≤ t}} = max{0, sup{ξ > 0 : ρ− f (ξ) ≤ t}}. Suppose first that γ− f (t) = sup{ξ : ρ− f (ξ) ≤ t} ≤ 0. Accordingly, {ξ > 0 : ρ− f (ξ) ≤ t} = −∞. By (4.8), γ− f+ t} = ∅ and so sup{ξ > 0 : ρ− 0 = (γ− sup{ξ > 0 : ρ− = (γ− that γ− f+ f (t) = sup{ξ : ρ− f )+(t). Suppose next that γ− (t) = max{0, γ− f (ξ) ≤ t}. By (4.8), γ− f+ f )+, whence (f+ ) = f+ by (4.6). f (ξ) ≤ (t) = max{0, −∞} = f (t) = f )+(t). This shows (cid:3) f (ξ) ≤ t} > 0, so that γ− f (t)} = (γ− Remark 4.3. By (4.5) and Remark 4.2, m f = mf if and only if there is a sequence of distinct points xn such that f (xn) → mf . 5. Reverse inequalities for f ⊼ g We begin with the (trivial) converse of the Brunn-Minkowski inequality for Eu- clidean balls. Lemma 5.1. If B1 and B2 are Euclidean balls in RN , then µN (B1 + B2) ≤ 2N −1(µN (B1) + µN (B2)). Proof. Call ri the radius of Bi, i = 1, 2. It is readily checked that B1 + B2 is a ball with radius r1 + r2 and the inequality simply follows from (r1 + r2)N ≤ 2N −1(rN (cid:3) 1 + rN 2 ). In the next lemma, Mf = Mg is not needed (compare with Lemma 3.1). Lemma 5.2. If f, g : RN → [0, ∞] are Borel measurable, then ZRN f ⊼ g ≤ 2N −1(cid:18)ZRN f +ZRN g(cid:19) . Proof. By (2.2) and (2.3) and since f ⊼ g ≥ 0 is measurable (Lemma 2.2), it follows ξ + )), where Lemma 5.1 was used. thatRRN f ⊼g = R ∞ ξ )dξ. Next, F + and so µN (F + 0 µN (F + +BG+ ξ +G+ )+µN (BG+ ) ≤ 2N −1(µN (BF + ξ ) ≤ µN (BF + ξ ⊂ BF + ξ +G+ G+ +BG+ ξ ξ ξ ξ ξ ξ INEQUALITIES FOR INFIMAL CONVOLUTION 13 This yieldsRRN f ⊼g ≤ 2N −1(cid:16)R ∞ (ii), the right-hand side is 2N −1(cid:16)R ∞ equals 2N −1(cid:16)RRN 0 µN (BF + ξ 0 µN ( F + f +RRN g(cid:17) because f , g ≥ 0 by Theorem 4.3 (iii). )dξ +R ∞ ξ ))dξ +R ∞ 0 µN (BG+ 0 µN ( G+ ξ )dξ(cid:17) . By Theorem 4.3 ξ ))dξ(cid:17) , which in turn (cid:3) No variant of Lemma 5.2 is true if f (or g) is replaced with f (or g): Example 5.1. With N = 1, let 0 < f ≤ 1 be integrable with f (n) = 1 for every n ∈ Z and let g = χ(−1,1) (= g). Then, f ⊼ g = 1, whence RR f ⊼ g = ∞ but f, g ∈ L1. If N = 1, f, g ≥ 0 are even and nonincreasing on [0, ∞) and Mf = Mg. Then, Is there a RR f ⊼ g = RR f + RR g by Theorem 4.3 (i) and Lemmas 3.1 and 5.2. different proof ? (If f = g, this follows from (f ⊼ f )(x) = f (x/2).) To go further, we need a simple property of Young functions. Lemma 5.3. If φ is a Young function and f : RN → [0, ∞], then (φ(f )) = φ( f ). Proof. For brevity, we only give the proof in the more important case when 0 < φ < ∞ on (0, ∞), so that φ is continuous on [0, ∞] and has an inverse ψ. The general case involves extra technicalities that lengthen the exposition. Recall that ρ+ f (ξ) denotes the radius of BF + . Since φ(f ) ≥ 0, it follows that ρ+ φ(f (x)) > ξ} = {x ∈ RN : f (x) > ψ(ξ)} = F + Thus, by (4.1) and (4.3), (φ(f ))(x) = inf{ξ ≥ 0 : ρ+ ρ+ f (η) ≤ x} = φ(inf{η ≥ 0 : ρ+ by (4.1) and (4.3) because f ≥ 0 implies ρ+ φ(f )(ξ) = ∞ if ξ < 0. If ξ ≥ 0, then {x ∈ RN : f (ψ(ξ)). f (ψ(ξ)) ≤ x} = inf{φ(η) : f (η) ≤ x} = f (x) ψ(ξ), so that ρ+ f (η) ≤ x}). Now, inf{η ≥ 0 : ρ+ φ(f )(ξ) = ρ+ f (η) = ∞ if η < 0. (cid:3) ξ Lemma 5.4. If φ is a Young function and f : RN → [0, ∞] is measurable, then f φ ≤ f φ. ξ ξ ) ≤ µN (BF + ) that RRN f = R ∞ Proof. Since f ≥ 0 implies f ≥ 0, it follows from Theorem 4.3 (ii) and from f . Upon µN (F + replacing f by φ(f ) in this inequality, it follows from Lemma 5.3 that RRN φ(f ) ≤ RRN φ( f ). Now, replace f with r−1f where r > 0 and use Theorem 4.3 (v) to get RRN φ(r−1f ) ≤ RRN φ(r−1 f ) for every r > 0. By (3.3), this implies f φ ≤ fφ. ξ )dξ ≤ R ∞ ξ )dξ = RRN 0 µN ( F + 0 µN (F + (cid:3) Theorem 5.5. If f, g : RN → [0, ∞] are Borel measurable and φ is a Young func- tion, then φ(f ), φ(g) and φ(f ⊼ g) are measurable and nonnegative and (5.1) f ⊼ gφ ≤ 2N −1( f φ + gφ). Proof. For the measurability of φ(f ), φ(g) and φ(f ⊼g), see Lemma 3.3 and Remark 3.1. That all three are nonnegative is trivial. In the proof of Lemma 3.3, we already 14 PATRICK J. RABIER established that φ(f ⊼ g) = φ(f ) ⊼ φ(g). Therefore, by Lemma 5.2 for φ(f ) and φ(g) and by Lemma 5.3, we get (5.2) ZRN φ(f ⊼ g) ≤ 2N −1(cid:18)ZRN φ( f ) +ZRN φ(g)(cid:19) . Suppose first that f φ > 0 and gφ > 0. Since (5.1) is trivial otherwise, we may and shall assume f φ < ∞ and gφ < ∞. If so, 0 < r := f φ + gφ < ∞ and the inequality (5.2) for r−1f and r−1g is (use Theorem 4.3 (v)) 21−N ZRN φ(r−1(f ⊼ g)) ≤ ZRN φ(r−1 f ) +ZRN φ(r−1g). With λ := r−1 f φ ∈ (0, 1), so that r−1 f = λ f −1 φ λ)g−1 f and that r−1g = (1 − φ g, this reads 21−N ZRN (5.3) φ(r−1(f ⊼ g)) ≤ ZRN φ(cid:16)λ f−1 φ f(cid:17) +ZRN φ(cid:16)(1 − λ)g−1 φ g(cid:17) . Since φ is convex and φ(0) = 0, then φ(µτ ) ≤ µφ(τ ) when τ ≥ 0 and 0 ≤ µ ≤ 1. The choices µ = 21−N in the left-hand side of (5.3) and, next, µ = λ and µ = 1 − λ in its right-hand side, yield ZRN φ(cid:0)21−N r−1(f ⊼ g)(cid:1) ≤ λZRN φ(cid:16) f −1 φ f(cid:17) + (1 − λ)ZRN φ(cid:16)g−1 φ g(cid:17) . By (3.5), it follows that RRN φ(cid:0)21−N r−1(f ⊼ g)(cid:1) ≤ 1 and so, by (3.3), f ⊼ gφ ≤ 2N −1r = 2N −1( f φ + gφ), as claimed in (5.1). To complete the proof, suppose now that f φ = 0 or gφ = 0. By symmetry, we may and shall assume that f φ = 0, whence f = 0 a.e. Since f (x) is a nonincreasing and right-continuous function of x, it follows that f = 0. Thus, by Lemma 4.4, either f = 0 or there are x0 ∈ RN and 0 < z ≤ ∞ such that f (x0) = z and f (x) = 0 when x 6= x0. If f = 0, then f ⊼ g = 0 since g ≥ 0 and (5.1) is trivial. If f (x0) = z > 0 and f (x) = 0 when x 6= x0, a straightforward calculation shows that (f ⊼ g)(x) = min{g(x − x0), z}. In particular, 0 ≤ (f ⊼ g) ≤ g(· − x0), whence f ⊼ gφ ≤ g(· − x0)φ = gφ ≤ gφ, the latter by Lemma 5.4. This proves (5.1). (cid:3) Since (f ⊼f )(x) ≥ f (x/2), equality holds in (5.1) if Lφ = L1 and 0 ≤ g = f = f ∈ L1. Thus, 2N −1 is best possible among all the constants independent of φ. On the other hand, even when Lφ = L∞, (5.1) is trivial only under additional assumptions, namely, Mf = ess sup f, Mg = ess sup g and min{Mf , Mg} = ess sup(f ⊼ g) (which is not implied by the former two, see Example 5.2 below). If so, f ⊼ g∞ = min{f ∞, g∞}, (5.4) which is much better than (5.1) since f ∞ ≤ f∞ and g∞ ≤ g∞ by Lemma 5.4. However, (5.4) is false without the extra assumptions mentioned above and then (5.1) is no longer trivial in L∞. How badly (5.4) may fail is shown in: Example 5.2. Let C ⊂ [0, 1] be the Cantor set. It is not hard to see that (x−C)∩C 6= ∅ for every x ∈ [0, 2] (notice that (x − Ck) ∩ Ck 6= ∅ for every k ∈ N, where C1 = [0, 1] and Ck+1 ⊂ Ck is obtained by removing the open middle thirds of the intervals in Ck). As a result, if f = ∞ on C and 0 outside, then f is Borel measurable, the INEQUALITIES FOR INFIMAL CONVOLUTION 15 right-hand side of (5.4) with g = f is 0, but since f ⊼ f = ∞ on [0, 2], its left-hand side is ∞. 6. Reverse inequalities for f (cid:3)g and f ⊻ g With the help of Theorem 5.5, it is a simple matter to prove a converse of Theorem 3.4. Theorem 6.1. Suppose that f, g : RN → (−∞, ∞] are Borel measurable, that mf , mg ∈ R and that mf + mg ≥ 0. Then f (cid:3)g ≥ 0, f − mf,g ≥ 0, g + mf,g ≥ 0 and (6.1) (f (cid:3)g)−1φ ≤ 2N −1(( f − mf,g)−1φ + (g + mf,g)−1φ), for every Young function φ. Proof. Use f − mf,g ≥ 0, g + mf,g ≥ 0 along with f (cid:3)g = (f − mf,g)(cid:3)(g + mf,g) ≥ (f − mf,g) ⊻ (g + mf,g) ≥ 0 to get f − mf,g ≥ 0 and g + mf,g ≥ 0 (by Theorem 4.5 (iii) and (iv)) as well as 0 ≤ (f (cid:3)g)−1 ≤ ((f − mf,g) ⊻ (g + mf,g))−1 = (f − mf,g)−1 ⊼ (g + mf,g)−1. This yields (f (cid:3)g)−1φ ≤ (f − mf,g)−1 ⊼ (g + mf,g)−1φ ≤ 2N −1(((f − mf,g)−1)φ + ((g + mf,g)−1)φ, where Theorem 5.5 was used for the second inequality. Next, ((f − mf,g)−1) = ( f − mf,g)−1 and ((g + mf,g)−1) = (g + mf,g)−1 by (cid:3) Theorem 4.5 (viii) and (iv) (in that order), which proves (6.1). The comments after Theorem 5.5 may be repeated: Without extra assumptions, (6.1) is not trivial even when Lφ = L∞. The proof of Theorem 6.1 shows that, more generally, if z ∈ R and −mg ≤ z ≤ mf (whence mf + mg ≥ 0), (f (cid:3)g)−1φ ≤ 2N −1(( f − z)−1φ + (g + z)−1φ). (6.2) However, (6.1) is optimal (to prove (f (cid:3)g)−1 ∈ Lφ) among all the inequalities (6.2). Indeed, if the right-hand side of (6.2) is finite for some z as above, the right- hand side of (6.1) is also finite3. To see this, assume first mf + mg = 0. Then, z = mf = −mg = mf,g is the only possible choice in (6.2) and (6.1) is optimal by default. Suppose now that mf +mg > 0 and note that, by Theorem 4.5 (iii), f −mf,g and g + mf,g are both bounded below by (mf + mg)/2 > 0. Thus, ( f − mf,g)−1 ∈ L∞ and so, if f − z ≥ 0 and ( f − z)−1 ∈ Lφ for some z, then ( f − mf,g)−1 = [1 − (z − mf,g)( f − mf,g)−1]( f − z)−1 ∈ Lφ since 1 − (z − mf,g)( f − mf,g)−1 ∈ L∞. Likewise, if g + z ≥ 0 and (g + z)−1 ∈ Lφ, then (g + mf,g)−1 ∈ Lφ. It is not clear whether 2N −1 is best possible in (6.1), among all the constants independent of φ. The remark that (f (cid:3)f )(x) ≤ 2f (x/2) and the choice g = f = f ≥ 0 with f −1 ∈ L1 = Lφ only shows that the best constant is at least 2N −2. There is also a converse of Corollary 3.5: Corollary 6.2. Suppose that f, g : RN → [−∞, ∞] are Borel measurable, that g ≥ 0 and that f 6 ≡∞, g 6 ≡∞. Then, f ⊻ g ≥ 0 and (6.3) (f ⊻ g)−1φ ≤ 2N (cid:0)( f+ − mf+,g)−1φ + (g + mf+,g)−1φ(cid:1) . 3While mostly true, the converse may fail when mf + mg > 0 and z = mf = m f or z = −mg = −mg . 16 PATRICK J. RABIER for every Young function φ. Proof. Since g is nonnegative, f ⊻ g = f+ ⊻ g ≥ (f+(cid:3)g)/2 = (f+/2)(cid:3)(g/2) ≥ 0. Also, 0 ≤ mf+ , mg < ∞ since f 6 ≡∞ and g 6 ≡∞. Thus, it suffices to use Theorem 6.1 with f+/2 and g/2 along with Theorem 4.5 (v) and (ix). (cid:3) In many cases, Theorem 6.1 and Corollary 6.2 can be used to prove that (f (cid:3)g)−1 ∈ Lφ or (f ⊻ g)−1 ∈ Lφ without any calculation of f or g, because there are easily verifiable sufficient conditions for the finiteness of the right-hand sides of (6.1) and (6.3). The simplest one is given in the following lemma. Lemma 6.3. Given f : RN → (−∞, ∞] with mf > −∞ (i.e., f bounded below), suppose that there are constants c > 0 and α > 0 such that f (x) ≥ h(x) := cxα for x large enough. Then, for every z < mf , ( f − z)−1φ ≤ 2h−1φ,RN \B + (mf − z)−1χBφ. (6.4) for every open ball B centered at the origin such that4 f ≥ h ≥ 2z outside B and every Young function φ. If φ is invertible with inverse ψ, this also reads ( f − z)−1φ ≤ 2h−1φ,RN \B + (mf − z)−1[ψ(µN (B)−1)]−1. (6.5) In particular, if h−1 ∈ Lφ(RN \B), then ( f − z)−1 ∈ Lφ. (Case in point: Since h is explicitly known, h−1 ∈ Lφ(RN \B) can often be checked by a calculation.) Proof. Let B be as in the theorem. By the "furthermore" part of Theorem 4.5 (vii), f ≥ h ≥ 2z outside B. Thus, h ≤ 2( f − z), whence ( f − z)−1 ≤ 2h−1, outside B. Meanwhile, by Theorem 4.5 (iii), ( f − z)−1 ≤ (m f − z)−1 < (mf − z)−1 < ∞ in B (and everywhere else) and (6.4) follows. To get (6.5) when φ is invertible with inverse ψ, just notice that χBφ = [ψ(µN (B)−1)]−1 by (3.3). (cid:3) From the proof of Lemma 6.3, xα can be replaced with a function h(x) satisfying general conditions. For example, if Lφ = Lp with p ≥ 1, Lemma 6.3 yields ( f − z)−1 ∈ Lp if α > N/p but, if f (x) ≥ cxα(log x)β for large x, the choice of any continuous strictly increasing function h of x that coincides with cxα(log x)β for x ≥ 1 (say) shows that ( f − z)−1 ∈ Lp in the limiting case α = N/p if β > p−1. Remark 6.1. The proof of Lemma 6.3 also shows that, more generally, (6.4) is true with mf replaced with m f and that this requires only m f > −∞. This may occasionally be useful, but rarely (see Remark 4.3). It is more delicate to extend Lemma 6.3 when z = mf = m f . The extra difficulty loc (because m f = ess inf f ; see Theorem 4.5 (iii)), so is that ( f − mf )−1 /∈ L∞ that local integrability becomes an issue. This requires further investigation. We only mention without proof (and will not use later) that if mf is a unique and nondegenerate minimum of f (plus a mild technical condition) then ( f − mf )−1 ∈ Lp loc if N ≥ 3 and 1 ≤ p < N/2. A direct application of Lemma 6.3 yields the following sample result. Theorem 6.4. Given f, g : RN → (−∞, ∞], suppose that there are constants c > 0 and α > 0 such that f (x) ≥ cxα and g(x) ≥ cxα for x large enough. Then: (i) If mf , mg ∈ R and mf + mg > 0, then (f (cid:3)g)−1 ∈ Lp if p ≥ 1 and p > N/α. (ii) If g ≥ 0 and mf+ + mg > 0, then (f ⊻ g)−1 ∈ Lp if p ≥ 1 and p > N/α. 4Since limx→∞ h(x) = ∞, the existence of B is not an issue. INEQUALITIES FOR INFIMAL CONVOLUTION 17 Proof. (i) Just notice that mf,g < mf (≤ m f ) and −mf,g < mg (≤ mg) since mf + mg > 0 and use Lemma 6.3 with Lφ = Lp and Theorem 6.1. (ii) If mf+ = ∞ or mg = ∞, then f ⊻ g = ∞ and the result is trivial. From now on, mf+ , mg < ∞, whence mf+ , mg ∈ R (because f+, g ≥ 0). By Theorem 4.5 (ix), ), whereas mf+,g < mf+ and −mf+,g < mg since mf+ + mg > 0. Now, use f+ = (f+ Lemma 6.3 and Corollary 6.2 with Lφ = Lp. (cid:3) Lemma 6.3 and its aforementioned variants yield generalizations of Theorem 6.4 to all Orlicz spaces. The proof of Theorem 6.4 does not use the estimate (6.4) but the next theorem, relevant to the results in the next section, does. As explained after the proof, there is a good reason to confine attention to Lp spaces. Theorem 6.5. Suppose that f, g : RN → (−∞, ∞] are Borel measurable, that mf ≥ 0, mg > −∞ and that mg > 0 if mf = 0. Given 1 ≤ p ≤ ∞, suppose also that for some α > N/p, there are constants c > 0 and α > 0 such that f (x) ≥ cxα for x large enough. For t > 0, set ft(x) := tf (t−1x), so that mft = tmf . Lastly, assume (g−z)−1 ∈ Lp when z < mg (e.g., if g(x) ≥ cxα for x large enough by Lemma 6.3, but this is not necessary). Then, (ft(cid:3)g)−1 ∈ Lp when tmf + mg > 0 (i.e., t > 0 if mf = 0 and t > −m−1 if mf > 0) and: (i) If mf = 0, then (ft(cid:3)g)−1p = O(tN/p) as t → ∞. (ii) If mf > 0 and p > N, then limt→∞ (ft(cid:3)g)−1p = 0. (iii) If mf > 0 and p ≥ N, then (ft(cid:3)g)−1p = O(t−1+N/p) as t → ∞. f mg Proof. As in Lemma 6.3, set h(x) := cxα and, in (6.4), let B be a ball centered at the origin such that f ≥ h and h ≥ mf + mg ≥ 2mf,g = mf − mg outside B. Evidently, ( f )t ≥ ht (:= th(t−1·)) outside tB and, by Theorem 4.5 (iv) and ) = ( f )t. Thus, ( f )t ≥ ht outside tB. Furthermore, if t ≥ 1, then ht ≥ (v), (ft tmf + mg ≥ 2mft,g = tmf − mg outside tB. Since ht(x) = t1−αcxα, it follows from the above that the estimate (6.4) can be used with f, h and B replaced with ft, ht and tB, respectively, and with z = mft,g = (tmf − mg)/2, provided that t ≥ 1 and that tmf + mg > 0 (so that mft,g < mft = tmf ), which holds for large t. Accordingly, (6.6) ( ft − mft,g)−1p ≤ 2t−1+N/ph−1p,RN \B + 2(tmf + mg)−1tN/pµN (B)1/p, where χtBp = tN/pµN (B)1/p was used. Since h−1p,RN \B < ∞ by the choice α > N/p, it follows that ( ft − mft,g)−1 ∈ Lp. Also, (g + mft,g)−1 ∈ Lp since −mft,g = (mg − tmf )/2 < mg when tmf + mg > 0 and since it is assumed that (g − z)−1 ∈ Lp when z < mg. Thus, by (6.1), (ft(cid:3)g)−1 ∈ Lp. The estimates (i), (ii) and (iii) follow from (6.1) and (6.6) and from the remarks that (a) if mf = 0, then (mg > 0 and) (g + mft,g)−1 = (g − mg/2)−1 is independent of t and (b) if mf > 0, then limt→∞ (g +mft,g)−1p = 0 by dominated convergence if p < ∞ and by g + mft,g ≥ (tmf + mg)/2 if p = ∞. (cid:3) Similar estimates hold when f is replaced with ft in Corollary 6.2 and estimates can also be worked out in other spaces Lφ, but the technicalities depend on φ. For 18 PATRICK J. RABIER instance, while Theorem 6.5 remains true if Lp is replaced with L1 + Lp, the proof is substantially more demanding (recall that L1 + Lp = Lφ with φ(τ ) = τ p in [0, 1] and φ(τ ) = pτ + 1 − p in (1, ∞) if 1 ≤ p < ∞ and L1 + L∞ = Lφ with φ(τ ) = 0 in [0, 1] and φ(τ ) = t−1 in (1, ∞); see e.g. [15])). Choices of h other than h(x) = cxα often lead to challenging calculations. 7. Application to the Hamilton-Jacobi equations We shall now apply the results of the previous sections to the Hamilton-Jacobi equations in their simplest form (see Subsection 8.1 for a variant) (7.1) (cid:26) ut + H(∇u) = 0 on (0, ∞) × RN , u(0, ·) = g on RN , where the Hamiltonian H and the initial value g are given functions on RN . Roughly speaking, when the Hamiltonian H (initial condition g) is convex, the Hopf-Lax formula (Hopf formula) provides a solution of (7.1 ). In both cases, various additional conditions are required of H and g and, as always, what constitutes a solution is somewhat flexible. While the more recent work focuses on viscosity solutions, other definitions exist as well. Throughout this section, we assume that g, H : RN → (−∞, ∞], that mg ∈ R (hence g 6 ≡∞), H 6 ≡∞ and that g is Borel measurable. Further assumptions will be introduced when needed. It is once and for all understood that t > 0. We denote by (tH)∗ the Legendre-Fenchel conjugate of tH, that is, (tH)∗(x) := sup y∈RN (x · y − tH(y)) = tH ∗(t−1x) Since (tH)∗ is always lsc, it is Borel measurable. 7.1. Solutions by the Hopf-Lax formula. In this subsection, H is convex and H ∗∗(0) ∈ R. The Hopf-Lax formula (Hopf [14], Lax [19]) (7.2) u(t, ·) = (tH)∗(cid:3)g, is known to give a solution of (7.1) under various conditions about g. That g is real- valued and continuous is a common assumption; see Bardi and Faggian [4] and the references therein. The case when g is lsc and not everywhere finite was considered by Imbert [16] and Stromberg [32]. Chen and Su [6] show that (7.2) is a solution when g is real-valued, a.e. continuous and satisfies a condition weaker than upper semicontinuity. Undoubtedly, other options can be found in the literature. The inequality (3.9) in Theorem 3.4 can be used with f = (tH)∗ if m(tH)∗ +mg ≥ 0. That mg > −∞ was assumed earlier, whereas m(tH)∗ = −tH ∗∗(0) ∈ R. As a result, the condition m(tH)∗ + mg ≥ 0 is simply (7.3) − tH ∗∗(0) + mg ≥ 0, so that m(tH)∗,g (see (3.8)) is given by m(tH)∗,g = −(tH ∗∗(0) + mg)/2. Thus, assuming (7.3), the corresponding inequality (3.9) (7.4) (2(tH)∗ + tH ∗∗(0) + mg)−1φ + (2g − tH ∗∗(0) − mg)−1φ ≤ 2u(t, ·)−1φ INEQUALITIES FOR INFIMAL CONVOLUTION 19 and the reverse inequality (6.1) of Theorem 6.1 (7.5) u(t, ·)−1φ ≤ 2N −1(cid:16)(2((tH)∗) + tH ∗∗(0) + mg)−1φ + (2g − tH ∗∗(0) − mg)−1φ(cid:17) , hold for every Young function φ. Given α > 1, call α′ := α/(α − 1) > 1 the Holder conjugate of α. It is easily checked and certainly folklore that if there is a constant d > 0 such that H(x) ≤ dxα′ for x large enough, then H ∗(x) ≥ cxα for x large enough. Consistent with Theorem 6.4, it follows that if also g(x) ≥ cxα for x large enough, then u(t, ·)−1 ∈ Lp for every p ≥ 1, p > N/α. Furthermore, since (tH)∗ = tH ∗(t−1·), Theorem 6.5 gives estimates for u(t, ·)−1p as t → ∞ if H ∗∗(0) < 0 or if H ∗∗(0) = 0 and mg > 0. (If H ∗∗(0) > 0, then (7.5) breaks down when t > mg/H ∗∗(0).) The accuracy (or possible lack thereof) of these estimates can be evaluated by using the inequality (7.4) with · φ = · p. We now look at two classical examples in more detail. In both cases, H ∗∗(0) = H(0) = 0 will make the inequalities simpler, but confines the discussion to mg ≥ 0. Example 7.1. Suppose that H(x) := x2/2, so that (tH)∗(x) = x2/2t and (7.3) boils down to mg ≥ 0 since H ∗∗ = H. If 1 ≤ p < ∞, a quick calculation shows that (2(tH)∗ + mg)−1p = A(p, mg)tN/2p where A(p, mg) > 0 is a constant which is finite if and only if mg > 0 and p > N/2. Accordingly, from (7.4), u(t, ·)−1 /∈ Lp for any t > 0 if mg = 0 or if N ≥ 2 and 1 ≤ p ≤ N/2, even if u(0, ·)−1 = g−1 ∈ Lp. Assume now mg > 0 and p > N/2. By (7.4) with H ∗∗(0) = 0 and since 0 ≤ g ≤ 2g − mg ≤ 2g, (7.6) u(t, ·)−1p ≥ c(tN/2p + g−1p), for some constant c > 0 depending only upon N, p and mg. Thus, once again, u(t, ·)−1 /∈ Lpif g−1 /∈ Lp. Conversely, since ((tH)∗) = (tH)∗ by Theorem 4.5 (i), it follows from (7.5) and from 0 ≤ g ≤ 2g − mg ≤ 2g (since mg ≥ mg; see Theorem 4.5 (iii)) that u(t, ·)−1 ∈ Lp for every t > 0 if g−1 ∈ Lp and that u(t, ·)−1p ≤ C(tN/2p + g−1p), where C > 0 depends only upon N, p and mg. In particular, u(t, ·)−1p = O(tN/2p) as t → ∞ (which is sharp because of (7.6)). Note that the general estimate of Theorem 6.5 gives only the less precise O(tN/p). Even though u(t, ·)−1 /∈ Lp if mg = 0, this does not preclude u(t, ·)−1 ∈ Lφ for Orlicz spaces outside the Lp scale. For instance, if Lφ = L1 + Lp, a simple calculation5 shows that (2(tH)∗ + mg)−1 ∈ L1 + Lp if p > N/2, with the restriction N ≥ 3 if mg = 0. If so, by (7.5), u(t, ·)−1 ∈ L1 + Lp if g−1 ∈ L1 + Lp and, by another calculation, u(t, ·)−1L1+Lp = O(t) as t → ∞ (optimal by (7.4)). In particular, this holds if g−1 ∈ Lq with 1 ≤ q ≤ p and p > N/2. This complements the Lp discussion above, even when mg > 0. Example 7.2. Suppose that H(x) = x, so that (tH)∗ is the indicator function of the closed ball B(0, t). Thus, the formula (7.2 ) is simply u(t, x) = inf y≤t g(x − y). Once again, H ∗∗ = H, so that (7.3) amounts to mg ≥ 0 and, if φ is a Young function, then (with ωN := µN (B(0, 1)) (7.7) (2(tH)∗ + mg)−1φ = inf (cid:8)r > 0 : φ(cid:0)r−1m−1 g (cid:1) ωN tN ≤ 1(cid:9) , 5The formula for φ was given at the end of the previous section. 20 PATRICK J. RABIER which is ∞ if mg = 0. Thus, by (7.4), u(t, ·)−1 /∈ Lφ for any t > 0 and any φ if mg = 0, even if u(0, ·)−1 = g−1 ∈ Lφ. (Since u(t, x)−1 = supy≤t g(x − y)−1, this also follows from the remark that for every ε > 0, u(t, ·)−1 ≥ ε−1 on some ball of radius t.) Assuming from now on that mg > 0, it follows from (7.7) that if φ has an inverse ψ on [0, ∞], then (2(tH)∗ + mg)−1φ = m−1 (where, as usual, ψ−1 = 1/ψ, not φ). In the simple case when φ(τ ) = τ p with p ≥ 1, this yields N t−N(cid:1)−1 g ψ(cid:0)ω−1 (7.8) (2(tH)∗ + mg)−1p = m−1 g ω1/p N tN/p. Once again, ((tH)∗) = (tH)∗ by Theorem 4.5 (i). Thus, if g−1 ∈ Lp (equivalent to (2g − mg)−1 ∈ Lp), it follows from (7.5) and (7.8) that u(t, ·)−1 ∈ Lp and that u(t, ·)−1p = O(tN/p) as t → ∞ (which is sharp because of (7.4)). This is the general estimate in Theorem 6.5 (i) which, in this example, is therefore optimal. In Example 7.2, g−1 ∈ Lp is not enough to get u(t, ·)−1 ∈ Lp : If N = 1 and if g = f −1/p with p < ∞ and f from Example 5.1, then mg = 1, g−1 = f 1/p ∈ Lp, but u(t, x) = inf y≤t f −1/p(x − y) = 1 if t ≥ 1, so that u(t, ·)−1 = 1 /∈ Lp. 8. Solutions by the Hopf formula The Hopf formula (Hopf [14]) u(t, ·) = (tH + g∗)∗ gives a solution of (7.1) when g is convex and various other technical assumptions are satisfied. See for instance Bardi and Evans [3]. In Penot and Volle [28], g and H can be extended real-valued. Below, we assume that g is lsc. Since u(t, ·) is convex, there is no measurability issue. First, −g∗(0) = inf g = mg (finite, as assumed above) by definition of g∗. Next, if h, k : RN → (−∞, ∞] and h 6 ≡∞, k 6 ≡∞ (so that h∗ and k∗ are proper), it is well- known and easily checked that (h + k)∗ ≤ h∗(cid:3)k∗. As a result, u(t, ·) = (tH + g∗)∗ ≤ (tH)∗(cid:3)g∗∗ = (tH)∗(cid:3)g. On the other hand, by using "inf sup ≥ sup inf", we get inf x(tH +g∗)∗(x) ≥ supy inf x(x·y−(tH +g∗)(y)) = −tH(0)−g∗(0) = −tH(0)+mg. Indeed, inf x(x·y −(tH +g∗)(y)) = −∞ if y 6= 0 because tH and g∗ are proper. This shows that if −tH(0) + mg ≥ 0, then 0 ≤ u(t, ·) ≤ (tH)∗(cid:3)g. Furthermore, since (tH)∗ = (tHC )∗ where HC = H ∗∗ is the closed convex hull of H, it follows that m(tH)∗ = −tHC(0) ≥ −tH(0). To ensure that m(tH)∗ ∈ R, i.e., that HC (0) > −∞, it must be assumed that H is bounded below by an affine function. If so, it follows from Theorem 3.4 with f = (tH)∗ that if −tH(0)+ mg ≥ 0 (hence −tHC(0)+ mg ≥ 0), then (8.1) (2(tH)∗ + tHC (0) + mg)−1φ + (2g − tHC (0) − mg)−1φ ≤ 2u(t, ·)−1φ, for every Young function φ. Since the inequality (8.1) depends only upon HC , it remains true when H is replaced with any proper closed convex function K ≤ H (so that K = KC) under the same assumption −tH(0) + mg ≥ 0 as above (still needed to ensure u ≥ 0) because this substitution decreases the left-hand side. This is less accurate, but often more convenient for practical evaluation. INEQUALITIES FOR INFIMAL CONVOLUTION 21 As an illustration of this point, it follows from the discussion in Example 7.1 that u(t, ·)−1 /∈ Lp if H(x) ≥ x2/2 and either 0 < p ≤ N/2 or mg = 0. Alternatively, from Example 7.2, u(t, ·)−1 /∈ Lφ for any φ if H(x) ≥ x and mg = 0. In the opposite direction, if L ≥ H is any lsc convex function, then (tH + g∗) ≤ (tL+g∗) and, since both tL and g∗ are proper and lsc, (tL+g∗)∗ = (tL)∗(cid:3)g as soon as the relative interiors of dom L and dom g∗ have nonempty intersection ([29, p. 145]). If so, (tL)∗(cid:3)g ≤ u(t, ·), so that the reverse inequalities of Theorem 6.1 can be used with f = (tL)∗ if −tL(0)+mg ≥ 0. Since 0 ≤ −tL(0)+mg = m(tL)∗(cid:3)g ≤ u(t, ·), it follows that (8.2) u(t, ·)−1φ ≤ 2N −1(cid:16)(2((tL)∗) + tL(0) + mg)−1φ + (2g − tL(0) − mg)−1φ(cid:17) , for every Young function φ. Unlike (8.1), the inequality (8.2) does not require H to be bounded below by an affine function. For instance, by Example 7.1, u(t, ·)−1p = O(tN/2p) as t → ∞ if H(x) ≤ L(x) := x2/2, mg > 0 and g−1 ∈ Lp, p > N/2. Also, by Example 7.2, u(t, ·)−1p = O(tN/p) as t → ∞ if H(x) ≤ L(x) := x, mg > 0 and g−1 ∈ Lp, p ≥ 1. 8.1. Explicit solutions of related problems. Various explicit formulas for the solution of (8.3) (cid:26) ut + H(u, ∇u) = 0 on (0, ∞) × RN , u(0, ·) = g on RN , when the Hamiltonian depends upon u, have been obtained under suitable (but restrictive) conditions. The result most directly relevant to this paper can be found in the work of Barron et al. [5], complemented and generalized in [2]. It is shown in [2, Theorem 6.11] that if H is continuous on RN +1, with H(s, x) nondecreasing in s ∈ R, convex and positively homogeneous of degree 1 in x ∈ RN , and if g : RN → (−∞, ∞] is lsc, then (for t > 0) (8.4) u(t, ·) := h[t] ⊻ g, is the minimal lsc supersolution of (8.3), where6 h[t](x) := h(cid:0)t−1x(cid:1) and h(x) := inf{s ∈ R : H(s, ·)∗(x) ≤ 0}. (8.5) It is shown in [5] that h is quasiconvex and lsc and that mh = −∞. ⊻ g(= h[t] From now on, we assume g ≥ 0, so that u(t, ·) = h[t]+ ⊻ g). Note that h[t]+ = h+[t]. Since mh = −∞ implies mh[t] = −∞, it follows that mh[t]+ = 0, so that mh[t]+,g = −mg/2 ≤ 0. Corollary 3.5 and Corollary 6.2 with f = h[t] can be used to evaluate u(t, ·)−1φ. Specifically, (2h[t]+ + mg)−1φ + (2g − mg)−1φ ≤ 2u(t, ·)−1φ (8.6) and (8.7) u(t, ·)−1φ ≤ 2N +1(cid:16)(2(h[t]+ ) + mg)−1φ + (2g − mg)−1φ(cid:17) , for every Young function φ. 6This differs from ht previously defined by th(t−1·). 22 PATRICK J. RABIER Example 8.1. Let H(s, x) = (s+)αx with α > 0. When g ≥ 0, the (nonnegative) solution (8.4) of (8.3) actually solves ut + uα∇u = 0. By (8.5) and a straightfor- ward calculation, h(0) = −∞ and h(x) = x1/α if x 6= 0. If mg = 0, then u(t, ·)−1 /∈ Lp for any p ≥ 1 by (8.6) since h−1 [t]+ = t1/α·−1/α /∈ Lp. However, it is readily checked that h−1 [t]+ ∈ L1 + Lp if 1 < N α < p. If so, it fol- lows from (8.7) with mg = 0 that u(t, ·) ∈ L1 + Lp if g−1 ∈ L1 + Lp (hence g−1 ∈ L1 + Lp by (8.6)) and the calculation of the estimates (8.6) and (8.7) is ) by Theorem 4.5 (i). trivial since mg = 0 and since h[t]+ = t−1/α · 1/α = (h[t]+ ))−1L1+Lp = Ct1/α with C := · −1/α L1+Lp, Thus, (h[t]+)−1L1+Lp = ((h[t]+ which yields u(t, ·)−1L1+Lp = O(t1/α) as t → ∞ (optimal). If mg > 0, then (2h[t]+ + mg)−1 ∈ Lp if and only if p > N α. If so and if g−1 ∈ Lp (equivalent to (2g − mg)−1 ∈ Lp), it follows from (8.7) that u(t, ·)−1 ∈ Lp with u(t, ·)−1p = O(tN/p) as t → ∞ (optimal by (8.6)). Example 7.2 is recovered when α = 0 in Example 8.1. If so, h(x) = −∞ on the closed unit ball and h(x) = ∞ outside, so that h+ is the indicator function of the closed unit ball. Even though the formulas (7.2) and (8.4) look different, they both provide the same solution u(t, x) = inf y≤t g(x − y). Example 8.2. Let H(s, x) = esx, so that, by (8.5), h(0) = −∞ and h(x) = ln x if x 6= 0. Thus, h+(x) = ln+ x is continuous, radially symmetric and nondecreasing ) = h[t]+ by Theorem 4.5 (i). As always in this in x, so that, once again, (h[t]+ subsection, g ≥ 0. Since h[t]+ = 0 on the ball with center 0 and radius t, it follows from (8.6) that u(t, ·)−1 /∈ Lφ for any t > 0 and any Young function φ if mg = 0. In addition, the slow growth of ln x reveals that u(t, ·)−1 /∈ Lp, 1 ≤ p < ∞, even if mg > 0. However, if mg > 0, then (2h[t]+ + mg)−1 ∈ Lφ if φ vanishes fast enough at the origin. Aside from L∞, one of the simplest examples is given by the Young function φ(τ ) := eτ −τ −2 . (The growth of φ at infinity could be damped considerably to enlarge the space Lφ.) Thus, if mg > 0 and g−1 ∈ Lφ (equivalent to (2g − mg)−1 ∈ Lφ), it follows from (8.7) that u(t, ·)−1 ∈ Lφ for every t > 0. We did not attempt to estimate u(t, ·)−1φ as t → ∞. References [1] Adams, R. A. and Fournier, J. F., Sobolev spaces, 2nd edition, Academic Press, Boston 2003. [2] Alvarez, O., Barron, E. N. and Ishii, H., Hopf-Lax formulas for semicontinuous data, Indiana Univ. Math. J. 48 (1999) 993-1035. [3] Bardi, M. and Evans, L. C., On Hopf's formulas for solutions of Hamilton Jacobi equations, Nonlinear Anal. TMA 8 (1984) 1373-1381. [4] Bardi, M. and Faggian, S., Hopf-type estimates and formulas for nonconvex-nonconcave Hamilton-Jacobi equations, SIAM J. Math. Anal. 29 (1998) 1067-1086. [5] Barron, E. N., Jensen, R. and Liu, W., Hopf Lax-type formula for ut + H(u, Du) = 0, J. Differ. Equations 126 (1996) 49-61. [6] Chen, G.-Q. and Su, B., Discontinuous solutions for Hamilton-Jacobi equations: uniqueness and regularity, Discrete Contin. Dyn. Syst. 9 (2003) 167-192. [7] Day, P. W., Rearrangement inequalities, Canadian J. Math. 24 (1972) 930-943. [8] Federer, H., Geometric measure theory, Springer-Verlag, Berlin (1996). [9] Fenchel, W., Convex cones, sets and functions, lecture notes, Princeton Univ., Princeton, N.J. 1953. [10] Gardner, R. J., The Brunn-Minkowski inequality, Bull. Amer. Math. Soc. 39 (2002) 355-405. INEQUALITIES FOR INFIMAL CONVOLUTION 23 [11] Glavosits, T. and Sz´az, ´A., The infimal convolution can be used to easily prove the classical Hahn-Banach theorem, Rostock. Math. Kolloq. 65 (2010) 71-83. [12] Hiriart-Urruty, J.-B., Extension of Lipschitz functions, J. Math. Anal. Appl. 77 (1980) 539- 554. [13] Hiriart-Urruty, J.-B. and Lemar´echal, C., Convex analysis and minimization algorithms (2 volumes), Springer-Verlag, Berlin 1996. [14] Hopf, E., Generalized solutions of non linear equations of first-order, J. Math. Mech. 14 (1965) 951-973. [15] Hudzik, H., Intersections and algebraic sums of Musielak-Orlicz spaces, Portugaliae Math. 40 (1981) 287-296. [16] Imbert, C., Convex analysis techniques for Hopf-Lax formulae in Hamilton-Jacobi equations, J. Nonlinear Convex Anal. 2 (2001) 333-343. [17] Jung, H. W. E, Uber die kleinste Kugel, die eine raumliche Figur einschliesst, J. Reine Angew. Math. 123 (1901) 241-257. [18] Krasnosel'skii, M. A. and Rutickii, Ya. B., Convex functions and Orlicz spaces, P. Noordhoff Ltd, Groningen 1961. [19] Lax, P. D., Hyperbolic systems of conservation laws II, Commun. Pure Appl. Math. 10 (1957) 537-566. [20] Lions, P.-L., Generalized Solutions of Hamilton-Jacobi Equations, Pitman, London, 1982. [21] Lucet, Y., What shape is your conjugate? A survey of computational convex analysis and its applications, SIAM Rev. 52 (2010) 505-542. [22] Luxemburg, W. A. J., Rearrangement invariant Banach function spaces, Queen's Papers on Pure and Applied Math. 10 (1967) 83-144. [23] Mazure, M.-L., Equations de convolution et formes quadratiques, Ann. Mat. Pura Appl. 158 (1991) 75-97. [24] Moreau, J. J., Fonctions `a valeurs dans [−∞, ∞]; notions alg´ebriques, Universit´e de Mont- pellier, Montpellier 1963. [25] Moreau, J. J., Inf-convolution des fonctions num´eriques sur un espace vectoriel, C. R. Acad. Sci. Paris 256 (1963) 5047-5049. [26] Moreau, J. J., Inf-convolution, sous-additivit´e, convexit´e des fonctions num´eriques, J. Math. Pures Appl. 49 (1970) 109-154. [27] O'Neil, R., Fractional integration in Orlicz spaces, Trans. Amer. Math. Soc. 115 (1965) 300-328. [28] Penot, J.-P. and Volle, M., Explicit solutions to Hamilton-Jacobi equations under mild con- tinuity and convexity assumptions. J. Nonlinear Convex Anal. 1 (2000) 177-199. [29] Rockafellar, R. T., Convex analysis, Princeton University Press, Princeton 1970. [30] Sierpi´nski, W., Sur la question de la mesurabilit´e de la base de M. Hamel, Fund. Math. 1 (1920) 105-111. [31] Stromberg, T., The operation of infimal convolution, Dissertationes Math. 352 (1996). [32] Stromberg, T., The Hopf-Lax formula gives the unique viscosity solution, Differential Integral Equations 15 (2002) 47-52. [33] Ziemer, W. P., Weakly differentiable functions, Springer-Verlag, New York 1989. Department of mathematics, University of Pittsburgh, Pittsburgh, PA 15260 E-mail address: [email protected]
1306.0358
2
1306
2013-09-17T10:01:40
Best proximity pair results for relatively nonexpansive mappings in geodesic spaces
[ "math.FA" ]
Given $A$ and $B$ two nonempty subsets in a metric space, a mapping $T : A \cup B \rightarrow A \cup B$ is relatively nonexpansive if $d(Tx,Ty) \leq d(x,y) \text{for every} x\in A, y\in B.$ A best proximity point for such a mapping is a point $x \in A \cup B$ such that $d(x,Tx)=\text{dist}(A,B)$. In this work, we extend the results given in [A.A. Eldred, W.A. Kirk, P. Veeramani, Proximal normal structure and relatively nonexpansive mappings, Studia Math., 171 (2005), 283-293] for relatively nonexpansive mappings in Banach spaces to more general metric spaces. Namely, we give existence results of best proximity points for cyclic and noncyclic relatively nonexpansive mappings in the context of Busemann convex reflexive metric spaces. Moreover, particular results are proved in the setting of CAT(0) and uniformly convex geodesic spaces. Finally, we show that proximal normal structure is a sufficient but not necessary condition for the existence in $A \times B$ of a pair of best proximity points.
math.FA
math
Best proximity pair results for relatively nonexpansive mappings in geodesic spaces Aurora Fern´andez-Le´on1 , Adriana Nicolae2,3 1Dpto. de An´alisis Matem´atico, Universidad de Sevilla, P.O. Box 1160, 41080-Sevilla, Spain 2 Department of Mathematics, Babe¸s-Bolyai University, Kogalniceanu 1, 400084 Cluj-Napoca, Romania 3 Simion Stoilow Institute of Mathematics of the Romanian Academy, Research group of the pro ject PD-3-0152, P.O. Box 1-764, RO-014700 Bucharest, Romania E-mail: [email protected], [email protected] Abstract Given A and B two nonempty subsets in a metric space, a mapping T : A ∪ B → A ∪ B is relatively nonexpansive if d(T x, T y ) ≤ d(x, y ) for every x ∈ A, y ∈ B . A best proximity point for such a mapping is a point x ∈ A ∪ B such that d(x, T x) = dist(A, B ). In this work, we extend the results given in [A.A. Eldred, W.A. Kirk, P. Veeramani, Proximal normal structure and relatively nonexpansive mappings, Studia Math. 171, 283–293 (2005)] for relatively nonex- pansive mappings in Banach spaces to more general metric spaces. Namely, we give existence results of best proximity points for cyclic and noncyclic relatively nonexpansive mappings in the context of Busemann convex reflexive metric spaces. Moreover, particular results are proved in the setting of CAT(0) and uniformly convex geodesic spaces. Finally, we show that proximal normal structure is a sufficient but not necessary condition for the existence in A × B of a pair of best proximity points. MSC: Primary 54E40, 47H10. Keywords: Relatively nonexpansive mapping, best proximity pair, best proximity point, proxi- mal normal structure, Busemann convexity. 1 Introduction Although metric fixed point theory is primary concerned with the existence of fixed points of map- pings that satisfy certain restrictions, there exist many other related problems that have attracted a high amount of interest from researchers in the area. One of such problems consists in studying the existence of approximate solutions of the equation x = T x in the absence of fixed points of the mapping T . A point x ∈ X is said to be an approximate solution of the equation x = T x if x is “close” to T x is some sense. Depending on the considered closeness condition between x and T x, results of different nature have been obtained in the literature. One classical result in this direction due to Ky Fan [16] states that if A is a compact, convex and nonempty subset of a locally convex Hausdorff topological vector space X and T is a continuous mapping from A to X , then there exists a point x ∈ A such that d(x, T x) = d(T x, A), where d is the semi-metric induced by a continuous semi-norm defined on X . If, instead of considering the condition d(x, T x) = d(T x, A), one requires that x is an absolute optimal approximate solution, that is, d(x, T x) = dist(A, B ) either for non-self mappings T : A → B or for mappings T : A ∪ B → A ∪ B such that T (A) ⊆ B , T (B ) ⊆ A or T (A) ⊆ A, T (B ) ⊆ B , existence, uniqueness and convergence results for such points are known as 1 best proximity point theorems. Note that the notion of best proximity point also refers to such a type of approximate solution. In the present work we mainly focus on the study of best proximity points for certain self-mappings T : A ∪ B → A ∪ B satisfying the above inclusion relations. The first results concerning such mappings were given by Kirk, Srinivasan and Veeramani [21] in 2003. More precisely, it was proved that if A and B are two nonempty and closed subsets of a complete metric space, T : A ∪ B → A ∪ B is such that T (A) ⊆ B , T (B ) ⊆ A and there exists k ∈ (0, 1) such that d(T x, T y) ≤ kd(x, y) for every x ∈ A, y ∈ B , then A ∩ B contains a fixed point of T . In the last years, many generalizations of this problem have appeared under the assumption A ∩ B = ∅. In this respect, weaker metric conditions have been considered for the mapping T . This is, for instance, the case of cyclic contractions [9, 29, 13], cyclic Meir-Keeler contractions [8, 27] or relatively nonexpansive mappings [10, 11, 28]. Relatively nonexpansive mappings were introduced by Eldred, Kirk and Veeramani [10] in the following way: a self-mapping T : A ∪ B → A ∪ B is relatively nonexpansive if d(T x, T y) ≤ d(x, y) for every x ∈ A, y ∈ B . If, in addition, T (A) ⊆ B and T (B ) ⊆ A then T is said to be cyclic. Likewise, if T (A) ⊆ A and T (B ) ⊆ B , then T is called noncyclic. In [10], several best proximity point results were given in Banach spaces for cyclic and noncyclic relatively nonexpansive mappings. While many generalizations of best proximity point results from the linear setting to metric spaces have appeared in the literature (see, among others, the works [13, 17] on cyclic contractions in metric spaces), no result has been given for relatively nonexpansive mappings in a nonlinear setting. Here we address this problem extending results proved in [10] in the context of Banach spaces to Busemann convex reflexive metric spaces. We also give more particular results in the setting of CAT(0) and uniformly convex geodesic spaces. Furthermore, we prove an analogue of a result due to Karlovitz [20] showing that proximal normal structure is a sufficient but not necessary condition for the existence in A × B of a pair of best proximity points. 2 Preliminaries In this section we compile the main concepts and results we will work with along this paper. We begin with some basic definitions and notations that are needed. Let (X, d) be a metric space and consider A and B two subsets of X . Define d(x, A) = inf {d(x, y) : y ∈ A}; dist(A, B ) = inf {d(x, y) : x ∈ A, y ∈ B }; δ(x, A) = sup{d(x, y) : y ∈ A}; δ(A, B ) = sup{d(x, y) : x ∈ A, y ∈ B }. From now on, B (a, r) denotes the closed ball in the space X centered at a ∈ X with radius r > 0. The metric projection PA onto A is the mapping PA (x) = {z ∈ A : d(x, z ) = dist(x, A)} for every x ∈ X. When this mapping is well-defined and singlevalued we use the same notation PA (x) to denote the unique point belonging to this set. 2 In the sequel, we say that a pair of sets (A, B ) has a property if each of the sets A and B has this property. For instance, we say that the pair (A, B ) is closed and bounded if A and B are both closed and bounded. A very important property in this paper for a pair of sets is the one of proximity. Definition 2.1. A pair (A, B ) of subsets of a metric space is said to be proximal if for each (a, b) ∈ A × B there exists (a′ , b′ ) ∈ A × B such that d(a, b′ ) = d(a′ , b) = dist(A, B ). In this context, given a pair of sets (A, B ) in a metric space, we say that the point p ∈ A is a proximal point of q ∈ B (with respect to A and B ) if d(p, q) = dist(A, B ). Then, (p, q) is also called pair of proximal points. In this paper we will mainly work with geodesic spaces. A metric space (X, d) is said to be a (uniquely) geodesic space if every two points x and y of X are joined by a (unique) geodesic, i.e, a map c : [0, l] ⊆ R → X such that c(0) = x, c(l) = y , and d(c(t), c(t′ )) = t − t′ for all t, t′ ∈ [0, l]. The image c([0, l]) of such a geodesic forms a geodesic segment which joins x and y and it is not necessarily unique. If no confusion arises, we use [x, y ] to denote a geodesic segment joining x and y . A point z in X belongs to a geodesic segment [x, y ] if and only if there exists t ∈ [0, 1] such that d(x, z ) = td(x, y) and d(y , z ) = (1 − t)d(x, y) and we write z = (1 − t)x + ty for simplicity. Notice 2 , we often use the notation x+y to denote 1 2 x + 1 that this point may not be unique. When t = 1 2 y . 2 Any Banach space is a geodesic space with usual segments as geodesic segments. A subset A of a geodesic space X is said to be convex if any geodesic segment that joins each pair of points x and y of A is contained in A. A geodesic triangle △(x, y , z ) in X consists of three points x, y , z ∈ X (the vertices of △) and three geodesic segments joining each pair of vertices (the edges of △). For more about geodesic spaces the reader can check [3, 5, 26]. A metric d : X × X → R is said to be convex if for any x, y , z ∈ X one has d(x, (1 − t)y + tz ) ≤ (1 − t)d(x, y) + td(x, z ) for all t ∈ [0, 1]. A geodesic space (X, d) is Busemann convex (introduced in [6]) if given any pair of geodesics c1 : [0, l1 ] → X and c2 : [0, l2 ] → X one has d(c1 (tl1 ), c2 (tl2 )) ≤ (1 − t)d(c1 (0), c2 (0)) + td(c1 (l1 ), c2 (l2 )) for all t ∈ [0, 1]. It is well-known that Busemann convex spaces are uniquely geodesic and with convex metric. Given two geodesic segments [x, z ] and [y , w] in a uniquely geodesic space (X, d), we say that [x, z ] is paral lel to [y , w], and we denote it by [x, z ]k[y , w], if d(x, y) = d(m1 , m2 ) = d(z , w), where 2 and m2 = y+w m1 and m2 are the midpoints of [x, z ] and [y , w], respectively (that is, m1 = x+z 2 ). The following property was given by Busemann in [7]. For the convenience of the reader we include a proof of this fact. Proposition 2.2. Let x, y , z , w be four points in a Busemann convex geodesic space. Suppose [x, z ]k[y , w]. Then, [x, y ]k[z , w]. 2 , m3 = x+y 2 , m2 = y+w Proof. Let m1 = x+z 2 and m4 = z+w 2 . Let r = d(x, y) = d(z , w) = d(m1 , m2 ). By using the Busemann convexity of the space we have that d(m1 , y+z 2 ) ≤ r/2, d(m1 , w+x 2 ) ≤ r/2, d(m2 , y+z 2 ) ≤ r/2 and d(m2 , w+x 2 ) ≤ r/2, which implies r = d(m1 , m2 ) ≤ d(cid:18)m1 , 2 (cid:19) + d(cid:18) y + z , m2(cid:19) ≤ y + z 2 2 (cid:19) + d(cid:18) w + x , m2(cid:19) ≤ r = d(m1 , m2 ) ≤ d(cid:18)m1 , w + x 2 3 = r. and + = r r 2 r 2 r 2 + r 2 Consequently, y+z 2 . Let m5 = m1+m2 2 = m1+m2 = w+x . Again by the Busemann convexity, 2 2 d(m4 , m5 ) ≤ 1/2d(x, z ) and d(m4 , m5 ) ≤ 1/2d(w, y), which implies d(m4 , m5 ) ≤ 1/2 min{d(x, z ), d(w, y)}. Since m5 is also the midpoint between m3 and m4 , d(m3 , m4 ) ≤ min{d(x, z ), d(w, y)}. Suppose now that d(x, z ) < d(y , w). Then d(m3 , m4 ) ≤ d(x, z ) < d(y , w). Since [m3 , m4 ]k[y , w], we can proceed similarly to get p1 = m3+y and s1 = m4+w such that d(p1 , s1 ) ≤ d(x, z ) < d(y , w). By 2 2 repeating the process, we obtain the sequences {pn}, {sn } ⊆ X , where, for n ≥ 2, pn = y+pn−1 and 2 sn = w+sn−1 , with d(sn , w) = d(pn , y) = r/2n+1 and d(pn , sn ) ≤ d(x, z ) < d(y , w). 2 Since d(y , w) ≤ d(y , pn ) + d(pn , sn ) + d(sn , w) for every n ∈ N, we may take superior limit in the previous inequality to get d(y , w) ≤ d(x, z ) < d(y , w), which is a contradiction and the result follows. In the sequel, we will also need the notion of uniformly convex geodesic space (see also [18, pg. 107]). A geodesic metric space (X, d) is said to be uniformly convex if for any r > 0 and any ε ∈ (0, 2] there exists δ ∈ (0, 1] such that for all a, x, y ∈ X with d(x, a) ≤ r , d(y , a) ≤ r and d(x, y) ≥ εr , d(m, a) ≤ (1 − δ)r, where m stands for a midpoint of x and y . A mapping δ : (0, +∞) × (0, 2] → (0, 1] providing such a δ = δ(r, ε) for a given r > 0 and ε ∈ (0, 2] is called a modulus of uniform convexity of X . If δ decreases with r (for each fixed ε) we say that δ is a monotone modulus of uniform convexity of X (introduced in [23]). If δ is lower semicontinuous from the right with respect to r (for each fixed ε), then we say δ is a lower semicontinuous from the right modulus of uniform convexity of X . If in the above definition we drop the uniformity conditions then we find the notion of strict convexity in metric spaces. Consequently, every uniformly convex geodesic space is strictly convex. Moreover, it is easy to see that every Busemann convex metric space is strictly convex [15] and that strictly convex metric spaces are uniquely geodesic. A very important class of geodesic spaces are CAT(0) spaces, that is, metric spaces of non- positive curvature in the sense of Gromov. These spaces play an essential role in several areas of mathematics [3] and find applications in other branches of science such as biology and com- puter science [1, 25]. CAT(0) spaces are defined in terms of comparison with E2 , the Euclidean plane, as follows: given (X, d) a geodesic metric space, a comparison triangle for a geodesic tri- angle △(x1 , x2 , x3 ) in (X, d) is a triangle △( ¯x1 , ¯x2 , ¯x3 ) in E2 such that dE2 ( ¯xi , ¯xj ) = d(xi , xj ) for i, j ∈ {1, 2, 3}. Such a comparison triangle always exists in E2 and is unique up to isometry. A geodesic triangle △ in X is said to satisfy the CAT(0) inequality if, given ¯△ a comparison triangle in E2 for △, for all x, y ∈ △ d(x, y) ≤ dE2 ( ¯x, ¯y ), where ¯x, ¯y ∈ ¯△ are the comparison points of x and y , respectively. A geodesic space X is a CAT(0) space if all its geodesic triangles satisfy the CAT(0) inequality. The following four point condition was used by Berg and Nikolaev [2] to characterize CAT(0) spaces. Theorem 2.3. Let (X, d) be a geodesic space. X is a CAT(0) space if and only if for every x, y , z , p ∈ X , d(x, z )2 + d(y , p)2 ≤ d(x, y)2 + d(y , z )2 + d(z , p)2 + d(p, x)2 . In complete CAT(0) spaces, the metric pro jection onto closed and convex subsets behaves as in Hilbert spaces in a certain sense. Proposition 2.4 ([3], Proposition 2.4, p. 176). Let X be a complete CAT(0) space, x ∈ X and C ⊂ X nonempty closed and convex. Then the fol lowing facts hold: 4 1. The metric projection PC (x) of x onto C is a singleton. 2. If y ∈ [x, PC (x)], then PC (x) = PC (y). 3. If x /∈ C and y ∈ C with y 6= PC (x) then ∠ PC (x) (x, y) ≥ π/2. 4. The mapping PC is a nonexpansive retraction from X onto C . Further, the mapping H : X × [0, 1] → X associating to (x, t) the point at distance td(x, PC (x)) on the geodesic [x, PC (x)] is a continuous homotopy from the identity map of X to PC . For a thorough treatment of CAT(0) spaces and related topics the reader can check [3, 19]. In the next section we will also work with reflexive metric spaces which extend the notion of reflexivity from Banach to metric spaces. A geodesic metric space X is said to be reflexive if for every decreasing chain {Cα } ⊂ X with α ∈ I such that Cα is closed convex bounded and nonempty for all α ∈ I we have that \α∈I Cα 6= ∅. Notice that every complete uniformly convex metric space with either a monotone or lower semicontinuous from the right modulus of uniform convexity is reflexive (see [24, 12]). Also note that a reflexive and Busemann convex geodesic space is complete (see [14, Lemma 4.1]). Moreover, in such a context the metric pro jection onto closed and convex subsets is well-defined and singlevalued. Next we give the definition of relatively nonexpansive mapping on the union of two sets. Definition 2.5. Suppose A and B are two nonempty subsets of a metric space X . A mapping T : A ∪ B → A ∪ B is relatively nonexpansive if d(T x, T y) ≤ d(x, y) for al l x ∈ A and y ∈ B . We say that a relatively nonexpansive mapping T is cyclic if T (A) ⊆ B and T (B ) ⊆ A and noncyclic if T (A) ⊆ A and T (B ) ⊆ B . In [10], the notion of proximal normal structure was introduced as a counterpart of the well- known concept of normal structure. This concept plays the same role for relatively nonexpansive mappings as normal structure plays for nonexpansive mappings. We state below this notion in the setting of geodesic metric spaces. Definition 2.6. A convex pair (K1 , K2 ) in a geodesic space is said to have proximal normal struc- ture if for any closed bounded convex proximal pair (H1 , H2 ) ⊆ (K1 , K2 ) for which dist(H1 , H2 ) = dist(K1 , K2 ) and δ(H1 , H2 ) > dist(H1 , H2 ), there exists (x1 , x2 ) ∈ H1 × H2 such that δ(x1 , H2 ) < δ(H1 , H2 ) and δ(x2 , H1 ) < δ(H1 , H2 ). As in the linear case, a pair (K, K ) has proximal normal structure if and only if K has normal structure in the sense of Brodski and Milman [4]. 3 Main results Given a pair of sets (A, B ) in a metric space X , let A0 and B0 be the subsets defined as follows: A0 = {x ∈ A : d(x, y ′ ) = dist(A, B ) for some y ′ ∈ B }, B0 = {y ∈ B : d(x′ , y) = dist(A, B ) for some x′ ∈ A}. Proposition 3.1. Let X be a reflexive and Busemann convex metric space and let (A, B ) be a nonempty closed convex pair of subsets in X . Suppose additional ly B is bounded. Then A0 and B0 are closed, convex, bounded and nonempty. 5 Proof. First we see that B0 is closed, convex, bounded and nonempty. Given any real number ε > 0, consider the set A′ ε = {x ∈ X : d(x, A) ≤ dist(A, B ) + ε}. ε is nonempty and closed. Moreover, A′ It is easy to see that A′ ε is convex. Let x and y be two points in A′ ε and m the midpoint between them. Since the space is Busemann convex, we have that (cid:19) d(m, A) = d(m, PA (m)) ≤ d(cid:18)m, PA (x) + PA (y) 2 ≤ max{d(x, A), d(y , A)} ≤ dist(A, B ) + ε. Thus m ∈ A′ ε . Now consider the set Aε = A′ ε ∩ B . It is immediate that Aε is closed, convex and bounded. Moreover, by definition of A′ ε , Aε is also nonempty. Then, by means of the reflexivity of the space, we conclude that ∩ε>0Aε 6= ∅. Since B0 = ∩ε>0Aε , we see that B0 is closed, convex, bounded and nonempty. Notice that A0 is bounded and nonempty since B0 is so. The fact that A0 is also closed and convex follows by a straightforward verification. Remark 3.2. Notice that in the previous result we just need one of the sets A and B to be bounded, no matter which of them. Theorem 3.3. Let X be a reflexive and Busemann convex metric space and let (A, B ) be a nonempty closed convex pair of subsets of X such that A is bounded. Let T : A ∪ B → A ∪ B be a cyclic relatively nonexpansive mapping. Suppose (A, B ) has proximal normal structure. Then there exists a pair (x, y) ∈ A × B such that d(x, T x) = d(y , T y) = dist(A, B ). Proof. This result follows by applying similar patterns as in the proof of Theorem 2.1 in [10]. However, in this more general setting, several changes and new techniques must be considered to get the result. From Proposition 3.1 we have that (A0 , B0 ) is closed, convex, bounded and nonempty. Moreover, we may notice that this pair is also proximal and satisfies dist(A0 , B0 ) = dist(A, B ). It is easy to see that T (A0 ) ⊆ B0 and T (B0 ) ⊆ A0 . Since (A, B ) has proximal normal structure, so does (A0 , B0 ). Consider the family Γ of sets F ⊆ A0 ∪ B0 such that F ∩ A0 and F ∩ B0 are closed, convex and nonempty and satisfy T (F ∩A0 ) ⊆ F ∩B0 and T (F ∩B0 ) ⊆ F ∩A0 , dist(F ∩A0 , F ∩B0 ) = dist(A0 , B0 ) and the pair (F ∩ A0 , F ∩ B0 ) is proximal. Since A0 ∪ B0 ∈ Γ, we have that Γ 6= ∅. Let {Fα }α∈I be a decreasing chain in Γ. We see that F0 = ∩α∈I Fα ∈ Γ. Since F0 ∩ A0 = ∩α∈I (Fα ∩ A0 ) and X is reflexive, we have that F0 ∩ A0 is closed, convex and nonempty. Similarly F0 ∩ B0 is closed, convex and nonempty. It can be easily proved that T (F0 ∩ A0 ) ⊆ F0 ∩ B0 and T (F0 ∩ B0 ) ⊆ F0 ∩ A0 . Thus, to see that F0 ∈ Γ, it remains to prove that the pair (F0 ∩ A0 , F0 ∩ B0 ) is proximal and dist(F0 ∩ A0 , F0 ∩ B0 ) = dist(A0 , B0 ). Let p ∈ F0 ∩ A0 . By definition, p ∈ Fα ∩ A0 for every α ∈ I . Since (A0 , B0 ) is proximal and X is strictly convex, there exists a unique point q ∈ B0 such that d(p, q) = dist(A0 , B0 ). Moreover, by using the proximity of (Fα ∩ A0 , Fα ∩ B0 ), we have that there exists a point qα ∈ Fα ∩ B0 such that d(p, qα ) = dist(Fα ∩ A0 , Fα ∩ B0 ) = dist(A0 , B0 ) for every α ∈ I . However, since qα ∈ B0 , we have that qα = q for every α ∈ I and therefore q ∈ F0 ∩ B0 . Consequently, F0 ∈ Γ and applying Zorn’s Lemma we obtain a minimal element K in Γ. Let K1 = K ∩ A0 and K2 = K ∩ B0 . If δ(K1 , K2 ) = dist(K1 , K2 ), then d(p, q) = dist(K1 , K2 ) for every pair (p, q) ∈ K1 × K2 . In particular, d(x, T x) = dist(K1 , K2 ) = dist(A0 , B0 ) for every x ∈ K and the result holds. Suppose now δ(K1 , K2 ) > dist(K1 , K2 ). Since (A0 , B0 ) has proximal normal structure, there exists (y1 , y2 ) ∈ K1 × K2 and λ ∈ (0, 1) such that δ(y1 , K2 ) ≤ λδ(K1 , K2 ) 6 d(z , y1 ) + and δ(y2 , K1 ) ≤ λδ(K1 , K2 ). Since K ∈ Γ, (K1 , K2 ) is proximal. In fact, we may notice by the strict convexity of X that for every point p ∈ K there exists only one point q ∈ K such that 2 ) = d(y ′ 2 ) ∈ K1 × K2 such that d(y1 , y ′ 1 , y ′ d(p, q) = dist(K1 , K2 ). Let (y ′ 1 , y2 ) = dist(K1 , K2 ). 1 and y2 ,y ′ Denote by x1 and x2 the midpoints of y1 ,y ′ 2 , respectively. By the Busemann convexity it follows that d(x1 , x2 ) = dist(K1 , K2 ). Since the metric in X is convex, for every z ∈ K2 we have 1 1 1 d(z , y ′ 1 ) ≤ d(z , x1 ) ≤ 2 2 2 (1 + λ) ≤ 2 Similarly, for every z ∈ K1 we have d(z , x2 ) ≤ (1+λ) 2 points (x1 , x2 ) ∈ K1 × K2 and α ∈ (0, 1) satisfying δ(x1 , K2 ) ≤ αδ(K1 , K2 ) and δ(x2 , K1 ) ≤ αδ(K1 , K2 ). Now consider the sets L1 ⊆ K1 and L2 ⊆ K2 defined as L1 = {x ∈ K1 : δ(x, K2 ) ≤ αδ(K1 , K2 ) and for its proximal point y ∈ K2 , δ(y , K1 ) ≤ αδ(K1 , K2 )}, δ(K1 , K2 ). Thus, there exists a pair of proximal (δ(y1 , K2 ) + δ(K1 , K2 )) δ(K1 , K2 ). L2 = {y ∈ K2 : δ(y , K1 ) ≤ αδ(K1 , K2 ) and for its proximal point x ∈ K1 , δ(x, K2 ) ≤ αδ(K1 , K2 )}. Since x1 ∈ L1 and x2 ∈ L2 , Li 6= ∅ for i = 1, 2. Next we show that Li is closed and convex for i = 1, 2. We just give the details for L1 since for L2 the proof follows similar patterns. Let {vn } ⊆ L1 be a sequence that converges to a point v ∈ K1 . Since d(vn , z ) ≤ αδ(K1 , K2 ) for every n ∈ N and z ∈ K2 , we get δ(v , K2 ) ≤ αδ(K1 , K2 ). The fact that vn ∈ L1 implies δ(wn , K1 ) ≤ αδ(K1 , K2 ), where wn ∈ K2 is the proximal point of vn ∈ K1 . Let w ∈ K2 such that d(v , w) = dist(K1 , K2 ). By the Busemann convexity, we get 2 (cid:19) = dist(K1 , K2 ). d(cid:18) vn + v wn + w 2 Now, by Proposition 2.2, we have d(wn , w) = d(vn , v) for n ∈ N, from where wn → w. Taking limit in (3.1), we may conclude δ(w, K1 ) ≤ αδ(K1 , K2 ). Consequently v ∈ L1 and then L1 is closed. Let p1 , q1 ∈ L1 . Next we see that m1 = p1+q1 2 ∈ L1 . Let p2 , q2 ∈ K2 be the proximal points of p1 and q1 , respectively. Consider m2 = p2+q2 . Since (K1 , K2 ) is proximal and the space is Busemann 2 convex, d(m1 , m2 ) = dist(K1 , K2 ). Let z ∈ K2 . The convexity of the metric implies 1 1 d(q1 , z ) ≤ αδ(K1 , K2 ). d(z , m1 ) ≤ 2 2 Thus, δ(m1 , K2 ) ≤ αδ(K1 , K2 ). The fact that δ(m2 , K1 ) ≤ αδ(K1 , K2 ) follows similarly since δ(p2 , K1 ) and δ(q2 , K1 ) are both ≤ αδ(K1 , K2 ). Then m1 ∈ L1 and so L1 is convex. From d(x1 , x2 ) = dist(K1 , K2 ) we get dist(L1 , L2 ) = dist(A0 , B0 ). Moreover, by the definition of the sets Li with i = 1, 2, it is immediate that (L1 , L2 ) is a proximal pair. d(p1 , z ) + (3.1) , 7 In the sequel we see that T (L1 ) ⊆ L2 and T (L2 ) ⊆ L1 . Let x ∈ L1 and y ∈ L2 such that d(x, y) = dist(L1 , L2 ). We prove that T x ∈ L2 . Let z ∈ K2 . Since d(T x, T z ) ≤ d(x, z ) ≤ δ(x, K2 ) ≤ αδ(K1 , K2 ), we get T (K2 ) ⊆ B (T x, αδ(K1 , K2 )) ∩ K1 := K ′ 1 . 1 is closed, convex and nonempty. Let K ′ Then K ′ 2 ⊆ K2 be the set defined as K ′ 2 = {y ′ ∈ K2 : there exists x′ ∈ K ′ 1 with d(x′ , y ′ ) = dist(K1 , K2 )}. Similarly as we proved before that L1 is closed and convex, we get that K ′ 2 is closed, convex and nonempty. 2 ) ⊆ K ′ 1 . The fact that T (K ′ 2 ) ⊆ K ′ 2 and T (K ′ 1 ) ⊆ K ′ Now we see that T (K ′ 1 is immediate. 1 , K ′ 2 ). Then d(T p, T q) = dist(K ′ 1 , K ′ 2 such that d(p, q) = dist(K ′ 1 and q ∈ K ′ Let p ∈ K ′ 2 ). Since 1 ) ⊆ K ′ 2 . Consequently, T (K ′ 1 and therefore T p ∈ K ′ 2 , we have T q ∈ K ′ q ∈ K ′ 2 . Notice that, 1 , K ′ 2 ) is also proximal and satisfies dist(K ′ 1 , K ′ by definition, the pair (K ′ 2 ) = dist(K1 , K2 ). Thus, 1 = K1 and K ′ 2 ∈ Γ and by minimality of K it follows that K ′ 1 ∪ K ′ K ′ 2 = K2 . Consequently, K1 ⊆ B (T x, αδ(K1 , K2 )) and therefore δ(T x, K1 ) ≤ αδ(K1 , K2 ). To conclude that T x ∈ L2 it remains to see that the proximal point z ∈ K1 of T x ∈ K2 satisfies δ(z , K2 ) ≤ αδ(K1 , K2 ). Since T is relatively nonexpansive, we have that z = T y . Thus, we only need to show that δ(T y , K2 ) ≤ αδ(K1 , K2 ). However, notice that this inequality holds if we repeat the previous construction of K ′ 1 and K ′ 2 starting from the point y ∈ L2 and considering any point z ∈ K1 . Thus, we have T x ∈ L2 and therefore T (L1 ) ⊆ L2 . In a similar way, we may see that T (L2 ) ⊆ L1 . As a consequence, L1 ∪ L2 ∈ Γ. Since, δ(L1 , L2 ) ≤ αδ(K1 , K2 ), we get a contradiction with the minimality of K . Theorem 3.4. Let X be a reflexive and Busemann convex metric space and let (A, B ) be a nonempty closed convex pair of subsets of X such that A is bounded. Let T : A ∪ B → A ∪ B be a noncyclic relatively nonexpansive mapping. Suppose (A, B ) has proximal normal structure. Then there exists a pair (x, y) ∈ A × B such that x = T x, y = T y and d(x, y) = dist(A, B ). Proof. Proceeding as in the previous theorem, we may see that (A0 , B0 ) is proximal, closed, convex and nonempty. Moreover, it is also immediate that dist(A0 , B0 ) = dist(A, B ), T (A0 ) ⊆ A0 and T (B0 ) ⊆ B0 . Now let Γ be the collection of sets F ⊆ A0 ∪ B0 such that F ∩ A0 and F ∩ B0 are closed, convex and nonempty and satisfy T (F ∩ A0 ) ⊆ F ∩ A0 and T (F ∩ B0 ) ⊆ F ∩ B0 , dist(F ∩ A0 , F ∩ B0 ) = dist(A0 , B0 ) and the pair (F ∩ A0 , F ∩ B0 ) is proximal. Since A0 ∪ B0 ∈ Γ, Γ 6= ∅. Let {Fα }α∈I be a decreasing chain in Γ. Following similar patterns as in the previous proof, we get that F0 = ∩α∈I Fα ∈ Γ. Then, applying Zorn’s Lemma, we find a minimal element K in Γ. Let K1 = K ∩ A0 and K2 = K ∩ B0 . If δ(K1 , K2 ) = dist(K1 , K2 ), then d(p, q) = dist(K1 , K2 ) for every pair (p, q) ∈ K1 × K2 . Let (p, q) ∈ K1 × K2 . Since T is relatively nonexpansive, d(T p, T q) = dist(K1 , K2 ) = dist(A0 , B0 ). Let m ∈ K2 be the midpoint between q and T q . Then, since d(m, T p) = d(m, p) = dist(A0 , B0 ) and X is strictly convex, we get that T p = p and T q = q with d(p, q) = dist(A0 , B0 ), so that the result holds. Suppose now δ(K1 , K2 ) > dist(K1 , K2 ). Repeating the reasoning of the previous theorem, we may find a pair of proximal points (x1 , x2 ) ∈ K1 × K2 such that δ(x1 , K2 ) ≤ αδ(K1 , K2 ) and δ(x2 , K1 ) ≤ δ(K1 , K2 ). We consider now the sets L1 ⊆ K1 and L2 ⊆ K2 defined as L1 = {x ∈ K1 : δ(x, K2 ) ≤ αδ(K1 , K2 ) and for its proximal point y ∈ K2 , δ(y , K1 ) ≤ αδ(K1 , K2 )}, 8 L2 = {y ∈ K2 : δ(y , K1 ) ≤ αδ(K1 , K2 ) and for its proximal point x ∈ K1 , δ(x, K2 ) ≤ αδ(K1 , K2 )}. Since the definition of these sets is as in Theorem 3.3, we have that (L1 , L2 ) is closed, convex, nonempty, proximal and satisfies dist(L1 , L2 ) = dist(K1 , K2 ). To see that T (L1 ) ⊆ L1 and T (L2 ) ⊆ L2 we may follow a similar reasoning to the one considered in Theorem 3.3 where the cyclic inclusion is proved. Although we omit some technical details, we include the proof for completeness. Let x ∈ L1 and y ∈ L2 such that d(x, y) = dist(L1 , L2 ). We prove that T x ∈ L1 . Let z ∈ K2 . Since d(T x, T z ) ≤ d(x, z ) ≤ δ(x, K2 ) ≤ αδ(K1 , K2 ), we get T (K2 ) ⊆ B (T x, αδ(K1 , K2 )) ∩ K2 := K ′ 2 . 2 is closed, convex and nonempty. Let K ′ Then K ′ 1 ⊆ K1 be the set K ′ 1 = {x ∈ K1 : there exists y ∈ K ′ 2 with d(x, y) = dist(K1 , K2 )}. 1 , K ′ 2 ) is proximal and satisfies dist(K ′ 1 , K ′ 2 is closed, convex and nonempty. Moreover, (K ′ Then K ′ 2 ) 1 ∪ K ′ 2 . Therefore, K ′ 2 ) ⊆ K ′ 1 and T (K ′ 1 ) ⊆ K ′ = dist(K1 , K2 ) and T (K ′ 2 ∈ Γ and by minimality of K it follows that K2 ⊆ B (T x, αδ(K1 , K2 )) and therefore δ(T x, K2 ) ≤ αδ(K1 , K2 ). Proceeding similarly, we may see that δ(T y , K1 ) ≤ αδ(K1 , K2 ). Since T y ∈ K2 is the proximal point of T x ∈ K1 , we conclude that T x ∈ L1 and therefore T (L1 ) ⊆ L1 . Similarly, T (L2 ) ⊆ L2 . As a consequence, L1 ∪ L2 ∈ Γ. Since, δ(L1 , L2 ) ≤ αδ(K1 , K2 ), we get a contradiction with the minimality of K . As a consequence of any of the two previous results, we get Kirk’s fixed point theorem in the setting of reflexive and Busemann convex metric spaces when dist(A, B ) = 0. Notice that, in this particular case, the fact that (A, B ) has proximal normal structure implies that A ∩ B has normal structure in the sense of Brodski and Milman [4]. Proposition 3.5. Every closed convex pair in a uniformly convex metric space X has proximal normal structure. Proof. Let (H1 , H2 ) be a closed convex bounded proximal pair in X with δ(H1 , H2 ) > dist(H1 , H2 ). Let x, y ∈ H1 such that d(x, y) > 0. Consider the points x′ , y ′ ∈ H2 such that d(x, x′ ) = d(y , y ′ ) = 2 ∈ H1 and m′ = x′+y ′ dist(H1 , H2 ). Let m = x+y 2 ∈ H2 . Since X is strictly convex, d(x′ , y ′ ) > 0. Let ε = min{d(x, y), d(x′ , y ′ )} and z ∈ H2 . Denote by δX a modulus of uniform convexity of X . Then, δ(H1 , H2 ) (cid:19)(cid:19)δ(H1 , H2 ). d(z , m) ≤ (cid:18)1 − δX (cid:18)δ(H1 , H2 ), ε Similarly, if we take z ∈ H1 , we get δ(H1 , H2 ) (cid:19)(cid:19)δ(H1 , H2 ). d(z , m′ ) ≤ (cid:18)1 − δX (cid:18)δ(H1 , H2 ), ε δ(H1 ,H2 ) (cid:19). Thus, δ(m, H2 ) ≤ αδ(H1 , H2 ) and δ(m′ , H1 ) ≤ αδ(H1 , H2 ), for α = 1 − δX (cid:18)δ(H1 , H2 ), ε Corollary 3.6. Let (A, B ) be a closed convex pair in a complete Busemann convex metric space X . Suppose that X is uniformly convex with a monotone or lower semicontinuous from the right modulus of uniform convexity and B is bounded. Let T : A ∪ B → A ∪ B be a cyclic relatively nonexpansive mapping. Then there exists a pair (x, y) ∈ A × B such that d(x, T x) = d(y , T y) = dist(A, B ). 9 Corollary 3.7. Let (A, B ) be a closed convex pair in a complete Busemann convex metric space X . Suppose that X is uniformly convex with a monotone or lower semicontinuous from the right modulus of uniform convexity and B is bounded. Let T : A ∪ B → A ∪ B be a noncyclic relatively nonexpansive mapping. Then there exists a pair (x, y) ∈ A × B such that x = T x, y = T y and d(x, y) = dist(A, B ). Proposition 3.8. Let (A, B ) be a closed convex pair in a complete Busemann convex metric space X . Suppose that X is uniformly convex with δX being a monotone or lower semicontinuous from the right modulus of uniform convexity and B is bounded. Let T : A ∪ B → A ∪ B be a noncyclic relatively nonexpansive mapping. Let x0 ∈ A0 and define xn+1 = xn+T xn for every n ≥ 1. Then 2 limn d(xn , T xn ) = 0. Moreover, if T (A) ⊆ C , where C is a compact set in X , then {xn } converges to a fixed point of T . Proof. By Corollary 3.7 we can find a point y ∈ B0 such that y = T y . Since the metric of the space is convex, we get that {d(y , xn )} is nonincreasing and so convergent to some d ≥ 0. Suppose first that d = 0. In this case, the result is immediate since {T xn} also converges to y . Now we consider d > 0. Suppose that there exists a subsequence {xnk } of {xn } such that d(xnk , T xnk ) ≥ ε > 0 for every k ≥ 0. d+1 ) , 1(cid:27). Then, from the uniform Suppose first δX is monotone. Let 0 < ρ < min (cid:26) dδX (d+1, ε d+1 ) 1−δX (d+1, ε convexity of the space, there exists k0 ∈ N such that d + 1 (cid:19)(cid:19)(d + ρ) for every k ≥ k0 . d(y , xnk +1 ) ≤ (cid:18)1 − δX (cid:18)d + ρ, ε By the definition of ρ, we have d(y , xnk +1 ) < d for every k ≥ k0 , which is a contradiction. In this case, let ε∗ > 0 such that Suppose now δX is lower semicontinuous from the right. ε ε ε d+1 ). For such ε∗ > 0, consider µ(ε∗ ) > 0 such that δX (d, d+1 ) + ε∗ 0 < ε∗ < δX (d, d+1 ) ≤ δX (r, for every r ∈ (d, d + µ(ε∗ )). Let 0 < ρ < min (cid:26) dδX (d, ε d+1 )+ε∗ , 1, µ(ε∗ )(cid:27). By using the uniform d+1 )−dε∗ 1−δX (d, ε convexity as before, there exists k0 ∈ N such that d(y , xnk +1 ) ≤ (cid:18)1 − δX (cid:18)d + ρ, d + 1 (cid:19)(cid:19)(d + ρ) for every k ≥ k0 . ε Similarly we get a contradiction. The rest of the proof follows similar patterns to those given in [10, Proposition 2.3]. Next we provide a bound for the existence of approximate fixed points for the mapping T . Recall that having a metric space (X, d), a mapping T : X → X and {xn} ⊆ X , a mapping Φ : (0, ∞) → N is called an approximate fixed point bound for {xn} (see also [22]) if ∀ε > 0, ∃n ≤ Φ(ε) such that d(xn , T xn ) ≤ ε. We don’t include the proof of the result below since it can be obtained by following a similar reasoning as in the main result of [23]. Proposition 3.9. Let (A, B ) be a closed convex pair in a complete Busemann convex metric space X . Suppose that X is uniformly convex with δX being a monotone modulus of uniform convexity and B is bounded. Let T : A ∪ B → A ∪ B be a noncyclic relatively nonexpansive mapping. Let 10 x0 ∈ A0 and b > 0 such that there exists y ∈ B0 with y = T y for which d(x0 , y) ≤ b. Define xn+1 = xn+T xn for every n ≥ 1. Then Φ : (0, ∞) → N, 2 # Φ(ε) = " 2b εδX (cid:0)b, ε b (cid:1) is an approximate fixed point bound for {xn }. Proposition 3.10. Let (A, B ) be a compact convex pair in a geodesic space with convex metric. Then (A, B ) has proximal normal structure. Proof. Let (H1 , H2 ) ⊆ (A, B ) be a closed convex bounded and proximal pair in X with δ(H1 , H2 ) > dist(H1 , H2 ). Suppose δ(x, H2 ) = δ(H1 , H2 ) for every x ∈ H1 . Let x0 ∈ H1 . Then there exists y0 ∈ H2 such that δ(x0 , H2 ) = δ(H1 , H2 ) = d(x0 , y0 ). Let x1 ∈ H1 such that d(x1 , y0 ) = dist(H1 , H2 ). Then d(x1 , x0 ) ≥ d(x0 , y0 ) − d(x1 , y0 ) = δ(H1 , H2 ) − dist(H1 , H2 ). Let y1 ∈ H2 such that , y1(cid:19) = δ(H1 , H2 ). d(cid:18) x1 + x0 2 Since the metric is convex, the fact that 1 x1 + x0 ) ≤ 2 2 implies d(y1 , x0 ) = d(y1 , x1 ) = δ(H1 , H2 ). Let m0,1 = x1+x0 . Take x2 ∈ H1 such that d(y1 , x2 ) = 2 3 m0,1 ) = δ(H1 , H2 ). Let m1,2 = 1 3 x2 + 2 3 x2 + 2 dist(H1 , H2 ) and y2 ∈ H2 such that d(y2 , 1 3 m0,1 . By using again the convexity of the metric, we obtain δ(H1 , H2 ) = d(y2 , x2 ) = d(y2 , m0,1 ) = d(y1 , m0,1 ). Moreover, this last equality also implies (d(y1 , x0 ) + d(y1 , x1 )) δ(H1 , H2 ) = d(y1 , d(y2 , x0 ) = d(y2 , x1 ) = δ(H1 , H2 ). Suppose we have {x1 , . . . , xn} in H1 , {m0,1 , m1,2 , . . . , mn−1,n} in H1 , where i 1 mi−2,i−1 xi + mi−1,i = i + 1 i + 1 for every i ≥ 2, and {y1 , . . . , yn−1} in H2 such that d(xi+1 , yi ) = dist(H1 , H2 ) for every i = 1 . . . n − 1, d(yi , mi−1,i ) = δ(H1 , H2 ) for every i = 1 . . . n − 1 and d(yi , xj ) = δ(H1 , H2 ) for every i = 0 . . . n − 1 and 0 ≤ j ≤ i. Now consider the point yn ∈ H2 such that d(yn , mn−1,n ) = δ(H1 , H2 ). Take xn+1 ∈ H1 such that d(xn+1 , yn ) = dist(H1 , H2 ). By using again the convexity of the metric, we may see that d(yn , xi ) = δ(H1 , H2 ) for every i = 0 . . . n. As a consequence, d(xn+1 , xi ) ≥ d(xi , yn ) − d(xn+1 , yn ) = δ(H1 , H2 ) − dist(H1 , H2 ) for every n ∈ N and for every i = 1 . . . n. Finally, by considering a convergent and, therefore, Cauchy subsequence of {xn } we get a contradiction. 11 4 CAT(0) spaces Proposition 4.1. Let (A, B ) be a closed convex pair in a CAT(0) space. Consider the mapping P : A ∪ B → A ∪ B defined as P (x) = (cid:26) PB (x) x ∈ A, PA (x) x ∈ B . Then P is a cyclic relatively nonexpansive mapping. Proof. The fact that P is cyclic is immediate. Let x ∈ A and y ∈ B . For simplicity, denote a = d(x, PA y), b = d(y , PB x), c = d(y , PA y), e = d(x, PB x), h = d(x, y) and r = d(PA y , PB x). Next we prove that r ≤ h. Let α = ∠PB x (x, y) and β = ∠PAy (x, y). By Proposition 2.4, cos α, cos β ≤ 0. By the Cosine Law in CAT(0) spaces, we get d(x, y)2 = h2 ≥ e2 + b2 and d(x, y)2 = h2 ≥ a2 + c2 . If we apply Theorem 2.3 to the four points {x, PB x, y , PA y} and consider the two previous inequal- ities, we obtain h2 + r2 ≤ a2 + b2 + c2 + e2 ≤ 2h2 , and the result follows. As a consequence of the previous result, we may reason as in [10] to conclude that in the setting of CAT(0) spaces Theorem 3.4 is a consequence of Theorem 3.3. Since every pair of closed and convex sets in a CAT(0) space satisfies property U C (see [13] for more details on this property), we may assert that every noncyclic or cyclic relatively nonexpansive mapping is also continuous if the pair (A, B ) is in addition proximal. Next we see that, as it happens in Hilbert spaces [10, Proposition 3.2], a noncyclic relatively nonexpansive mapping is even nonexpansive if the pair (A, B ) is proximal. Proposition 4.2. Let (A, B ) be a closed convex bounded proximal pair of sets in a CAT(0) space. Let T : A ∪ B → A ∪ B be a noncyclic relatively nonexpansive mapping. Then T is nonexpansive. Proof. Let x1 , x2 ∈ A. Let d1 = d(x1 , PB x2 ), d2 = d(x2 , PB x1 ), α = ∠PB T x1 (T x1 , T PB x2 ) and β = ∠PB T x2 (T x2 , T PB x1 ). Note that d(x1 , x2 ) = d(PB x1 , PB x2 ), d(T x1 , T x2 ) = d(PB T x1 , PB T x2 ), d(x1 , PB x1 ) = d(x2 , PB x2 ) = dist(A, B ) and α, β ≥ π/2. Let △1 = △(T PB x2 , T PB x1 , T x1 ) and △2 = △(T PB x1 , T PB x2 , T x2 ). Since d(T x1 , T PB x1 ) ≤ d(x1 , PB x1 ) it follows by Proposition 2.4, (1) that T PB x1 = PB T x1 . Similarly, T PB x2 = PB T x2 . If we apply the Cosine Law in CAT(0) spaces to △1 and △2 , we obtain d(T x1 , T x2 )2 + dist(A, B )2 = d(PB T x1 , PB T x2 )2 + dist(A, B )2 ≤ d(T x1 , T PB x2 )2 ≤ d2 1 and d(T x1 , T x2 )2 + dist(A, B )2 = d(PB T x1 , PB T x2 )2 + dist(A, B )2 ≤ d(T x2 , T PB x1 )2 ≤ d2 2 . 12 Thus, 1 , d2 d(T x1 , T x2 )2 + dist(A, B )2 ≤ min{d2 2 }. If we apply Theorem 2.3 to the four points {x1 , PB x1 , x2 , PB x2 }, we obtain 2 min{d2 1 , d2 2 } ≤ d2 1 + d2 2 ≤ 2(d(x1 , x2 )2 + dist(A, B )2 ). By using (4.2) and (4.3), we get that T is nonexpansive on A. In a similar way, we get that T is nonexpansive on B and then the result holds. (4.2) (4.3) Next we see that a similar result also holds for cyclic relatively nonexpansive mappings. Proposition 4.3. Let (A, B ) be a closed convex bounded proximal pair of sets in a CAT(0) space. Let T : A ∪ B → A ∪ B be a cyclic relatively nonexpansive mapping. Then T is nonexpansive. Proof. Let x1 , x2 ∈ A. Let d1 = d(x1 , PB x2 ), d2 = d(x2 , PB x1 ) α = ∠PAT x1 (T x1 , T PB x2 ) and β = ∠PAT x2 (T x2 , T PB x1 ). Note that d(x1 , x2 ) = d(PB x1 , PB x2 ), d(T x1 , T x2 ) = d(PB T x1 , PB T x2 ), d(x1 , PB x1 ) = d(x2 , PB x2 ) = dist(A, B ) and α, β ≥ π/2. Let △1 = △(T PB x2 , T PB x1 , T x1 ) and △2 = △(T PB x1 , T PB x2 , T x2 ). By Proposition 2.4, (1), we get that T PB x1 = PAT x1 and T PB x2 = PAT x2 . By applying now the Cosine Law to △1 and △2 and Theorem 2.3 to {x1 , PB x1 , x2 , PB x2 } as in the previous theorem we get the result. 5 Proximal normal structure: a sufficient but not necessary con- dition In 1979, Karlovitz [20] proved that the normal structure of the domain of a nonexpansive self- mapping T is a sufficient but not necessary condition to guarantee existence of fixed points of such a mapping in the context of reflexive Banach spaces. For this aim, a very specific family of reflexive spaces which originated with R. C. James was considered. In the same setting we see now that the proximal normal structure behaves similarly with respect to relatively nonexpansive mappings. First we give an example of a closed convex bounded pair of sets in a reflexive Banach space that does not have proximal normal structure. Example 5.1. Let X denote the Banach space given by the set ℓ2 endowed with the norm kxk = max{kxk∞ , kxk2 /√2}. Let and A = B (θ , 1) ∩ {x = {xn}n≥1 ∈ X : x1 = 1, xi ≥ 0 for every i ≥ 1} B = B (θ , 2) ∩ {x = {xn}n≥1 ∈ X : x1 = 2, xi ≥ 0 for every i ≥ 1}, where θ denotes the origin of the space ℓ2 . It is easy to see that (A, B ) is closed, convex, bounded and proximal. From the definition we have kx − yk ≥ 1 for every x ∈ A and y ∈ B . Notice that e1 ∈ A, 2e1 ∈ B and ke1 − 2e1 k = 1. Then dist(A, B ) = 1. We claim δ(A, B ) = 2. Let x = {xi} ∈ A and y = {yi} ∈ B . Since x ∈ A, 0 ≤ xi ≤ 1 for every i ≥ 1. Equal ly, y ∈ B implies 0 ≤ yi ≤ 2. i − 2(x1 y1 ) − 2 P∞ i + P∞ 2 = P∞ i=1 x2 i=1 y2 Thus, kx − yk∞ ≤ 2. Note that kx − yk2 i=2 xiyi ≤ 6. Then kx − yk2 /√2 ≤ √3 < 2 and therefore kx − yk ≤ 2. It is easy to see that {e1 + en }n≥2 ⊆ A and 13 {2e1 + 2en }n≥2 ⊆ B . Since d(e1 + en , 2e1 + 2em ) = 2 for n 6= m, we have δ(A, B ) = 2. On the other hand, we have δ(x, B ) = 2 for every x = {xi }i≥1 ∈ A . This is a consequence of d(x, 2e1 + 2en ) ≥ lim n→∞ xn − 2 = 2. lim n→∞ Then (A, B ) does not have proximal normal structure. Remark 5.2. Let A∗ = conv({e1 + en : n ≥ 2}) and B ∗ = conv({2e1 + 2en : n ≥ 2}). Note that the pair (A∗ , B ∗ ) ⊆ (A, B ) does not have proximal structure either. Next we see that proximal normal structure is a sufficient but not necessary condition to obtain the existence of best proximity points in Theorem 2.1 in [10] and therefore also in Theorem 3.3. Proposition 5.3. Let (A, B ) be the pair of sets considered in the previous example and suppose T : A ∪ B → A ∪ B is a cyclic relatively nonexpansive mapping. Then there exists a pair (x, y) ∈ A × B such that d(x, T x) = d(y , T y) = dist(A, B ). Proof. Consider the point 2e1 ∈ B . Notice that d(2e1 , A) = δ(2e1 , A) = 1. Thus, 1 ≤ d(T (2e1 ), T 2 (2e1 )) ≤ d(2e1 , T (2e1 )) = 1 and the result follows. Remark 5.4. Notice that the previous result also holds whenever we have a cyclic relatively non- expansive mapping defined on a pair (A, K ) ⊆ (A, B ) with 2e1 ∈ K . Acknowledgements Aurora Fern´andez-Le´on was partially supported by the Plan Andaluz de Investigaci´on de la Junta de Andaluca FQM-127 and Grant P08-FQM-03543, and by MEC Grant MTM2009-10696-C02-01. Part of this work was carried out while she was visiting the Babe¸s-Bolyai University in Cluj-Napoca. She acknowledges the kind hospitality of the Department of Mathematics. Adriana Nicolae was supported by a grant of the Romanian National Authority for Scientific Research, CNCS-UEFISCDI, pro ject number PN-II-ID-PCE-2011-3-0383. References [1] I. Bartolini, P. Ciaccia, and M. Patella (2002). String matching with metric trees using an approximate distance. In, String processing and information retrieval (A. H. F. Laender and A. L. Oliveira, eds.). Lecture Notes in Computer Science vol. 2476, Springer, Berlin, pp. 271– 283. [2] I. D. Berg and I. G. Nikolaev (2008). Quasilinearization and curvature of Alexandrov spaces. Geom. Dedicata 133:195–218. [3] M. R. Bridson and A. Haefliger (1999). Metric Spaces of Non-positive Curvature. Springer- Verlag, Berlin Heidelberg. [4] M. S. Brodski and D. P. Milman (1948). On the center of a convex set. Dokl. Akad. Nauk SSSR (N.S.) 59:837–840 (in Russian). [5] D. Burago, Y. Burago, and S. Ivanov (2001). A Course in Metric Geometry. Amer. Math. Soc., Providence, RI. 14 [6] H. Busemann (1955). Geometry of Geodesics. Academic Press, New York. [7] H. Busemann (1948). Spaces with non-positive curvature. Acta Math. 80:259–310. [8] C. Di Bari, T. Suzuki, and C. Vetro (2008). Best proximity points for Meir-Keeler contractions. Nonlinear Anal. 69:3790–3794. [9] A. A. Eldred and P. Veeramani (2006). Existence and convergence of best proximity points. J. Math. Anal. Appl. 323:1001–1006. [10] A. A. Eldred, W. A. Kirk, and P. Veeramani (2005). Proximal normal structure and relatively nonexpansive mappings. Studia Math. 171:283–293. [11] R. Esp´ınola (2008). A new approach to relatively nonexpansive mappings. Proc. Amer. Math. Soc. 136:1987–1995. [12] R. Esp´ınola, A. Fern´andez-Le´on, and B. Pi¸atek (2010). Fixed points of single- and set-valued mappings in uniformly convex metric spaces with no metric convexity. Fixed Point Theory Appl. 2010: Article ID 169837, 16 pages. [13] R. Esp´ınola and A. Fern´andez-Le´on (2011). On best proximity points in metric and Banach spaces. Can. J. Math. 63:533–550. [14] R. Esp´ınola and A. Nicolae (2012). Mutually nearest and farthest points of sets and the Drop Theorem in geodesic spaces. Monatsh. Math. 165:173–197. [15] R. Esp´ınola and B. Pi¸atek, Fixed point property and unbounded sets in CAT(0) spaces. J. Math. Anal. Appl. doi:10.1016/j.jmaa.2013.06.038 (in press). [16] K. Fan (1969). Extensions of two fixed point theorems of F.E. Browder. Math. Z. 112:234–240. [17] A. Fern´andez-Le´on (2010). Existence and uniqueness of best proximity points in geodesic metric spaces. Nonlinear Anal. 73:915–921. [18] K. Goebel and S. Reich (1984). Uniform Convexity, Hyperbolic Geometry, and Nonexpansive Mappings. Marcel Dekker, New York and Basel. [19] M. Gromov (1984). Metric Structure for Riemannian and non-Riemannian Spaces, Birkhauser, Boston. [20] L. A. Karlovitz (1976). Existence of fixed points of nonexpansive mappings in a space without normal structure. Pacific J. Math. 66:153–159. [21] W. A. Kirk, P. S. Srinivasan, and P. Veeramani (2003). Fixed points for mappings satisfying cyclical contractive conditions. Fixed Point Theory 4:79–89. [22] U. Kohlenbach (2005). Some computational aspects of metric fixed point theory. Nonlinear Anal. 61:823–837. [23] L. Leu¸stean (2007). A quadratic rate of asymptotic regularity for CAT(0)-spaces. J. Math. Anal. Appl. 325:386–399. [24] L. Leu¸stean (2010). Nonexpansive iterations in uniformly convex W -hyperbolic spaces. Con- temp. Math. 513:193–209. 15 [25] M. Owen and S. Provan (2011). A fast algorithm for computing geodesic distances in tree space. IEEE/ACM Trans. Computational Biology and Bioinformatics 8:2–13. [26] A. Papadopoulus (2005). Metric Spaces, Convexity and Nonpositive Curvature. European Math. Soc., Zurich. [27] B. Pi¸atek (2011). On cyclic Meir-Keeler contractions in metric spaces. Nonlinear Anal. 74:35– 40. [28] V. Sankar Ra j and P. Veeramani (2009). Best proximity pair theorems for relatively nonex- pansive mappings. Appl General Topol. 10:21–28. [29] T. Suzuki, M. Kikkawa, and C. Vetro (2009). The existence of the best proximity points in metric spaces with the property UC. Nonlinear Anal. 71:2918–2926. 16
1806.03596
1
1806
2018-06-10T07:21:31
Weighted Riesz bases in g-fusion frames and their perturbation
[ "math.FA" ]
In this paper, we introduce orthonoramal and Riesz bases for g-fusion frames and will show that the weights have basic roles. Next, we prove an effective theorem between frames and g-fusion frames by using an operator. Finally, perturbations of g-fusion frames will be presented.
math.FA
math
WEIGHTED RIESZ BASES IN G-FUSION FRAMES AND THEIR PERTURBATION GHOLAMREZA RAHIMLOU, VAHID SADRI, AND REZA AHMADI Abstract. In this paper, we introduce orthonoramal and Riesz bases for g-fusion frames and will show that the weights have basic roles. Next, we prove an effective theorem between frames and g-fusion frames by using an operator. Finally, perturbations of g-fusion frames will be presented. 1. Introduction Bases play a prominent role in discrete frames and their studying can extract interesting properties from the frames. One of the most important types of bases are orthonormal bases and also, as their special case, the Riesz basis. The Riesz basis has been defined in [6] by the image of orthonormal bases with a bounded bijective operator. Afterwards, this basis has been equalized in a property with complete sequences and inequalities for the synthesis operator. Sun in [17] could introduce a Riesz basis for g-frames by using that property and we will continue his method in Section 3 for g-fusion frames. In Section 4, we present a useful operator for characterizations of these frames and finally in Section 5, a perturbation of these frames will be studied. 2. Preliminaries Throughout this paper, H and K are separable Hilbert spaces and B(H, K) is the collection of all bounded linear operators of H into K. If K = H, then B(H, H) will be denoted by B(H). Also, πV is the orthogonal projection from H onto a closed subspace V ⊂ H and {Hj}j∈J is a sequence of Hilbert spaces, where J is a subset of Z. It is easy to check that if u ∈ B(H) is an invertible operator, then πuV uπV = uπV . We define the space H2 := (Pj∈J ⊕Hj)ℓ2 by H2 = (cid:8){fj}j∈J : fj ∈ Hj, Xj∈J (1) kfjk2 < ∞(cid:9), 2010 Mathematics Subject Classification. Primary 42C15; Secondary 46C99, 41A58. Key words and phrases. g-fusion frame, Dual g-fusion frame, gf-complete, gf- orthonormal basis, gf-Riesz basis. 1 2 GH. RAHIMLOU, V. SADRI, AND R. AHMADI with the inner product defined by h{fj},{gj}i = Xj∈J hfj, gji. It is clear that H2 is a Hilbert space with pointwise operations. Definition 1. Let W = {Wj}j∈J be a collection of closed subspaces of H, {vj}j∈J be a family of weights, i.e. vj > 0 and Λj ∈ B(H, Hj) for each j ∈ J. We say Λ := (Wj, Λj , vj) is a generalized fusion frame (or g-fusion frame) for H if there exists 0 < A ≤ B < ∞ such that for each f ∈ H (2) v2 jkΛj πWj fk2 ≤ Bkfk2. Akfk2 ≤ Xj∈J We call Λ a Parseval g-fusion frame if A = B = 1. When the right hand side of (2) holds, Λ is called a g-fusion Bessel sequence for H with bound B. Throughout this paper, Λ will be a triple (Wj , Λj , vj) with j ∈ J unless otherwise noted. The synthesis and analysis operators in the g-fusion frames are defined by TΛ : H2 −→ H TΛ({fj}j∈J) = Xj∈J vj πWj Λ∗j fj , , T ∗Λ : H −→ H2 T ∗Λ(f ) = {vj ΛjπWj f}j∈J. Thus, the g-fusion frame operator is given by SΛf = TΛT ∗Λf = Xj∈J hSΛf, fi = Xj∈J j πWj Λ∗j Λj πWj f v2 v2 jkΛj πWj fk2. and (3) Therefore AI ≤ SΛ ≤ BI. This means that SΛ is a bounded, positive and invertible operator (with adjoint inverse). So, we have the reconstruction formula for any f ∈ H: (4) j πWj Λ∗j ΛjπWj S−1 v2 j S−1 v2 Λ πWj Λ∗j Λj πWj f. f = Xj∈J Λ f = Xj∈J For the proof of followings, see [16]. Theorem 2.1. Λ is a g-fusion Bessel sequence for H with bound B if and only if the operator TΛ is well-defined and bounded operator with kTλk ≤ √B. Theorem 2.2. Λ is a g-fusion frame for H if and only if TΛ : H2 −→ H, TΛ({fj}j∈J) = Xj∈J vj πWj Λ∗j fj WEIGHTED RIESZ BASES IN G-FUSION FRAMES AND THEIR PERTURBATION 3 is a well-defined, bounded and surjective. Definition 2. A g-fusion frame Λ := (S−1 frame operator S Λ = T ΛT ∗Λ Λ. Λ , vj) with g-fusion is called the (canonical) dual g-fusion frame of Λ Wj , Λj πWj S−1 v2 j π Wj Λj∗Λj πWj f, Now, we can obtain j πWj Λ∗j v2 (5) Λjπ Wj f = Xj∈J Λ Wj , Λj := Λj πWj S−1 Λ . f = Xj∈J where Wj := S−1 Definition 3. Let Λj ∈ B(H, Hj) for each j ∈ J. Λ = (Wj , Λj , vj) is called gf-complete, if It is easy to check that Λ is gf-complete if and only if span{πWj Λ∗j Hj} = H. {f : Λj πWj f = 0, j ∈ J} = {0}. Proposition 1. If Λ = (Wj , Λj , vj) is a g-fusion frame for H, then Λ is a complete. Proof. See [16]. (cid:3) 3. gf-Riesz and Orthonormal Bases Definition 4. Let W = {Wj}j∈J be a collection of closed subspaces of H and j ∈ J. We say that (Wj , Λj) is a gf-orthonormal bases for H with respect to {vj}j∈J, if (6) hviπWiΛ∗i gi, vj πWj Λ∗j gji = δi,jhgi, gji , i, j ∈ J , gi ∈ Hi , gj ∈ Hj (7) jkΛj πWj fk2 = kfk2 , v2 f ∈ H. Xj∈J Definition 5. Λ = (Wj , Λj , vj) is called a gf-Riesz basis for H if (1) Λ is gf-complete, (2) There exist 0 < A ≤ B < ∞ such that for each finite subset I ⊆ J and gj ∈ Hj, j ∈ I, (8) AXj∈I kgjk2 ≤ (cid:13)(cid:13)Xj∈I vjπWj Λ∗j gj(cid:13)(cid:13) 2 ≤ BXj∈I kgjk2. It is easy to check that if Λ is a gf-Riesz bases for H, then the operator TΛ which is defined by is injetive. TΛ : H2 −→ H TΛ({gj}j∈J) = Xj∈J vj πWj Λ∗j gj , 4 GH. RAHIMLOU, V. SADRI, AND R. AHMADI Proposition 2. Let Λ = (Wj, Λj , vj) be a g-fusion frame for H and suppose that (6) holds. Then Λ is a gf-orthonormal basis for H. Proof. Assume that SΛ is the g-fusion frame operator of Λ and M := {f ∈ H : SΛf = f}. It is clear that M is a non-empty closed subspace of H. Let f ∈ H and k ∈ J. Since vjvkπWj Λ∗j ΛjπWj πWk Λ∗kΛkπWkf = δj,kπWk Λ∗kΛkπWk f, then πWk Λ∗kΛkπWkf ∈ M . So, for any h ∈ M⊥ and g ∈ M we have, for all j ∈ J hπWj Λ∗j Λj πWj h, gi = hh, πWj Λ∗j Λj πWj gi = 0. Thus πWj Λ∗j Λj πWj h = 0 for each j ∈ J, and so kΛj πWj hk = 0. By definition of g-fusion frame, we obtain h = 0. Therefore, M⊥ = {0} and we conclude H = M . So, SΛ = idH and the proof is completed. (cid:3) Theorem 3.1. Λ is a gf-orthonormal bases for H if and only if (I) vj πWj Λ∗j is isometric for any j ∈ J; (II) Lj∈J vj πWj Λ∗j (Hj) = H. Proof. Suppose that Λ is a gf-orthonormal basis for H. If j ∈ J and g ∈ Hj, we have hvj πWj Λ∗j g, vj πWj Λ∗j gi = δj,jhg, gi = hg, gi. Then, vjπWj Λ∗j is isometric for any j ∈ J. So, vjπWj Λ∗j (Hj) is a closed subspace of H. Therefore, for each j ∈ J and f ∈ H, we have from (7) hf, fi = Xj∈J = hXj∈J v2 jhΛj πWj f, Λj πWj fi j πWj Λ∗j Λj πWj f, fi. v2 Thus, for each f ∈ H, j πWj Λ∗j ΛjπWj f. v2 f = Xj∈J By letting gj := vjΛj πWj f , we obtain f = Pj∈J vjπWj Λ∗j gj and Xj∈J v2 jkπWj Λ∗j gjk2 = Xj∈J kgjk2 = Xj∈J v2 jkΛj πWj fk2 = kfk2. So, Lj∈J vj πWj Λ∗j (Hj) = H. Conversely, if (I) , (II) are satisfied, then (6) is clear. Indeed, for each i 6= j, viπWiΛ∗i ⊥ vj πWj Λj and vjπWj Λ∗j is isometric. Let f ∈ H, we get from (II) for any j ∈ J and some gj ∈ Hj vj πWj Λ∗j gj f = Xj∈J WEIGHTED RIESZ BASES IN G-FUSION FRAMES AND THEIR PERTURBATION 5 and v2 kgjk2. jkπWj Λ∗j gjk2 = Xj∈J kfk2 = Xj∈J Now, let i ∈ J, then for each f, h ∈ Hj, hvjΛiπWif, hi = hXj∈J j ΛiπWiπWj Λ∗j gj , hi v2 = Xj∈J = hviπWiΛ∗i gi, viπWiΛ∗i hi = hgi, hi. jhπWj Λ∗j gj , πWiΛ∗i hi v2 Hence gi = viΛiπWif for each i ∈ J. So, f = Pj∈J v2 f ∈ H and (7) is proved. Corollary 1. Every gf-orthonormal basss for H is a gf-Riesz bases for H with bounds A = B = 1. j πWj Λ∗j Λj πWj f for all (cid:3) Theorem 3.2. Let Θ = (Wj , Θj) be a gf-orthonormal bases with respect to {vj}j∈J and Λ = (Wj , Λj , vj) be a g-fusion frame for H with same weights. Then, there exists a operator V ∈ B(H) and surjective such that ΛjπWj = ΘjπWj V ∗ for all j ∈ J. Proof. Let Then, V is well-defined and bounded. Indeed, for each finite subset I ⊆ J and f ∈ H, V : H −→ H, V f = Xj∈J j πWj Λ∗j Θj πWj f. v2 kV fk = sup khk=1(cid:12)(cid:12)(cid:10)Xj∈J √B(cid:0)Xj∈J ≤ = √Bkfk, v2 v2 j πWj Λ∗j ΘjπWj f, h(cid:11)(cid:12)(cid:12) jkΘj πWj fk2(cid:1) 1 2 where B is an upper g-fusion frame bound for Λ. Therefore, the series is weakly unconditionally Cauchy and so unconditionally convergent in H (see [7] page 58) and also kV k ≤ √B. Since Θ is a gf-orthonormal bases, then ΘjπWj πWiΘ∗i g = (vivj)−1δi,jg and j πWj Λ∗j ΘjπWj πWiΘ∗i g v2 V πWiΘ∗i g = Xj∈J = πWiΛ∗i g, 6 GH. RAHIMLOU, V. SADRI, AND R. AHMADI for all g ∈ Hj and i ∈ J. Thus ΛjπWj = ΘjπWj V ∗. Now, we show that V is surjective. Assume that f ∈ H. By Theorem 2.2, there is {gj}j∈J ∈ H2 such that Pj∈J vjπWj Λ∗j gj = f . Let g := TΘ({gj}j∈J), thus vj πWj Λ∗j gj = f V g = Xj∈J V vjπWj Θ∗j gj = Xj∈J and V is surjective. (cid:3) Corollary 2. If Λ is a Parseval g-fusion frame for H, then V ∗ is isometric. Corollary 3. If Λ is a gf-Riesz bases for H, then V is invertible. Proof. Let V f = 0 and f ∈ H. Since TΛ is injective and j πWj Λ∗j Θj πWj f = TΛT ∗Θf, v2 V f = Xj∈J therefore, T ∗Θf = 0. So, kfk2 = kTΘfk2 = 0, hence, f = 0. Corollary 4. If (Wj, Λj) is a gf-orthonormal bases for H with respect to {vj}j∈J, then V is unitary. Proof. By Corollaries 1 and 3, the operator V is invertible. Let f ∈ H, we obtain (cid:3) kfk2 = Xj∈J v2 jkΛj πWj fk2 = Xj∈J kΘj πWj V ∗fk2 = kV ∗fk2. Thus, V V ∗ = idH and this means that V is unitary. (cid:3) 4. Characterizations of g-fusion frames, gf-Riesz and gf-orthonormal bases Sun in [17] showed that each g-frame for H induces a sequence in H dependent on the g-frame and he proved a useful theorem about them. In this section, we are going to express Sun's method for g-fusion frames. Let W = {Wj}j∈J be a family of closed subspaces of H, {vj}j∈J be a family of weights, Λj ∈ B(H, Hj) for each j ∈ J and {ej,k}k∈Kj be an orthonormal basis for Hj, where Kj ⊆ Z and j ∈ J. Suppose that ϕ :h −→ C, ϕ(f ) = hvjΛjπWj f, ej,ki. We have therefore, ϕ is a bounded linear functional on H. Now, we can write kϕfk ≤ vjkΛjkkfk, So, if hvjΛjπWj f, ej,ki = hf, vjπWj Λ∗j ej,ki. (9) then hf, uj,ki = hvjΛj πWj f, ej,ki for all f ∈ H. uj,k := vj πWj Λ∗j ej,k, j ∈ J, k ∈ Kj WEIGHTED RIESZ BASES IN G-FUSION FRAMES AND THEIR PERTURBATION 7 Remark 1. Using (9), we get for each f ∈ H (10) vjΛjπWj f = Xk∈Kj Xk∈Kj hf, uj,ki2 ≤ v2 hf, uj,kiej,k. jkΛjk2kfk2. But Thus, {uj,k}k∈Kj is a Bessel sequence for H. It follows that for each f ∈ H and g ∈ Hj hf, vj πWj Λ∗j gi = hvjΛjπWj f, gi hvj πWj Λj f, ej,kihej,k, gi = Xk∈Kj = Xk∈Kj = Df, Xk∈Kj hf, uj,kihej,k, gi hg, ej,kiuj,kE. Therefore, (11) for all g ∈ Hj. vj πWj Λ∗j g = Xk∈Kj hg, ej,kiuj,k We say {uj,k : j ∈ J, k ∈ Kj} the sequence induced by Λ. Theorem 4.1. Let Λ = (Wj , Λj , vj) and uj,k be defined by (9). Then we get the followings. (I) Λ is a g-fusion frame (resp. g-fusion Bessel sequence, Parseval g- fusion frame, gf-Riesz basis, gf-orthonormal basis) for H if and only if {uj,k : j ∈ J, k ∈ Kj} is a frame (resp. Bessel sequence, Parseval frame, Riesz basis, orthonormal basis) for H. (II) The g-fusion operator for Λ coincides with the frame operator for {uj,k : j ∈ J, k ∈ Kj}. Proof. (I). By (10), Λ is a g-fusion frame (resp. g-fusion Bessel sequence, : j ∈ J, k ∈ Kj} is a Parseval g-fusion frame) for H if and only if {uj,k frame (resp. Bessel sequence, Parseval frame) for H. Assume that Λ is a gf-Riesz basis for H and gj ∈ Hj. Thus, gj = Xk∈Kj where cj,k ∈ ℓ(Kj). Note that, by (10), cj,kej,k, {f : ΛjπWj f = 0, j ∈ J} = {f : hf, uj,ki = 0, j ∈ J, k ∈ Kji} 8 and GH. RAHIMLOU, V. SADRI, AND R. AHMADI AXj∈I kgjk2 ≤ (cid:13)(cid:13)Xj∈I cj,k2 ≤ (cid:13)(cid:13)Xj∈I Xk∈Kj is equivalent to AXj∈I Xk∈Kj vj πWj Λ∗j gj(cid:13)(cid:13) 2 ≤ BXj∈I kgjk2 2 cj,kuj,k(cid:13)(cid:13) ≤ BXj∈I Xk∈Kj cj,k2 for any finite I ⊆ J. Thus, Λ is a gf-Riesz basis if and only if {uj,k : j ∈ J, k ∈ Kj} is a Riesz basis. Now, let (Wj, Λj) be a gf-orthonormal basis for H with respect to {vj}j∈J. We get for any j1, j2 ∈ J, k1 ∈ Kj1 and k2 ∈ Kj2 huj1,k1 , uj2,k2i = hvj1πWj1 Λ∗j1 ej1,k1, vj2 πWj2 Λ∗j2ej2,k2i = δj1,k1δj2,k2. So, {uj,k : j ∈ J, k ∈ Kj} is an orthonormal sequence. Moreovere kfk2 = Xj∈J v2 jkΛj πWj fk2 = Xj∈J Xk∈Kj hf, uj,ki2 for any f ∈ H. Hence {uj,k : j ∈ J, k ∈ Kj} is an orthonormal basis. For the opposite implication, we need only to prove that (6) holds. We have by (11) for each j1 6= j2 ∈ J, gj1 ∈ Hj1 and gj2 ∈ Hj2, hvj1 πWj1 Λ∗ j1 gj1 , vj2 πWj2 Λ∗ hgj1 , ej1,k1iuj1,k1 , Xk2 ∈Kj2 j2 gj2i = (cid:10) Xk1∈Kj1 hgj2 , ej2,k2iuj2,k2(cid:11) = 0 and for all g1, g2 ∈ Hj hvjπWj Λ∗ j g1, vjπWj Λ∗ j g2i = (cid:10) Xk1 ∈Kj hg1, ej,k1iuj,k1 , Xk2 ∈Kj hg2, ej,k2iuj,k2(cid:11) = hg1, g2i. (II). By (10) and (11) we have for any f ∈ H Xj∈J v2 j πWj Λ∗j Λj πWj f = Xj∈J Xk∈Kj hvjΛj πWj f, ej,kiuj,k hf, uj,k′ej,k′i, ej,kEuj,k = Xj∈J Xk∈KjD Xk′∈Kj = Xj∈J Xk∈Kj hf, uj,kiuj,k. 5. Perturbation of g-Fusion Frames In this section, we present some perturbation of g-fusion frames and re- view some results about them. First, we need the following which is proved in [3]. (cid:3) WEIGHTED RIESZ BASES IN G-FUSION FRAMES AND THEIR PERTURBATION 9 Lemma 1. Let U be a Linear operator on a Banach space X and assume that there exist λ1, λ2 ∈ [0, 1) such that kx − U xk ≤ λ1kxk + λ2kU xk for all x ∈ X. Then U is bounded and invertible. Moreovere, 1 − λ1 1 + λ2kxk ≤ kU xk ≤ 1 + λ1 1 − λ2kxk and 1 − λ2 1 + λ1kxk ≤ kU−1xk ≤ 1 + λ2 1 − λ1kxk for all x ∈ X. Theorem 5.1. Let Λ := (Wj , Λj , vj) be a g-fusion frame for H with bounds A, B and {Θj ∈ B(H, Hj)}j∈J be a sequence of operators such that for any finite subset I ⊆ J and for each f ∈ H kXj∈I j ΛjπWj f−πWj Θ∗ + µkXj∈I , µ} < 1, then Θ := (Wj , Θj, vj) is a g-fusion frame j ΘjπWj f(cid:1)k ≤ λkXj∈I j ΘjπWj fk + γ(Xj∈I jkΛjπWj fk2) v2 j(cid:0)πWj Λ∗ j ΛjπWj fk where 0 ≤ max{λ + γ√A for H with bounds v2 j πWj Θ∗ v2 j πWj Λ∗ v2 1 2 , 1 −(cid:0)λ + γ√A(cid:1) 1 + µ A and B 1 + λ + γ√B 1 − µ Proof. Assume that I ⊆ J is a finite subset and f ∈ H. We have kXj∈I j ΘjπWj f(cid:1)k j ΘjπWj fk ≤ kXj∈I j(cid:0)πWj Λ∗ v2 j πWj Θ∗ v2 j ΛjπWj f − πWj Θ∗ + kXj∈I v2 j πWj Λ∗ v2 j πWj Λ∗ j ΛjπWj fk j ΛjπWj fk + µkXj∈I jkΛjπWj fk2) v2 2 . 1 + γ(Xj∈I ≤ (1 + λ)kXj∈I v2 j πWj Θ∗ j ΘjπWj fk Then kXj∈I v2 j πWj Θ∗ j ΘjπWj fk ≤ 1 + λ 1 − µkXj∈I v2 j πWj Λ∗ j ΛjπWj fk + γ 1 − µ (Xj∈I jkΛjπwj fk2) v2 1 2 . Let SΛ be the g-fusion frame operator of Λ, then j πWj Λ∗j Λj πWj fk ≤ kSΛfk ≤ Bkfk. v2 kXj∈I 10 GH. RAHIMLOU, V. SADRI, AND R. AHMADI j πWj Θ∗j Θj πWj fk ≤ ( v2 Therefore, for all f ∈ H kXj∈I So, Pj∈J v2 √B + 1 + λ 1 − µ γ 1 − µ )(Xj∈I v2 jkΛj πWj fk2) 1 2 < ∞. j πWj Θ∗j ΘjπWj f is unconditionally convergent. Let SΘ : H −→ H SΘ(f ) = Xj∈J j πWj Θ∗j Θj πWj f. v2 SΘ is well defined and bounded operator with kSΘk ≤ and for each f ∈ H, we have 1 + λ 1 − µ B + γ√B 1 − µ jkΘj πWj fk2 = hSΘf, fi ≤ kSΘkkfk2. v2 Xj∈J It follows that Θ := (Wj , Θj , vj) is a g-fusion Bessel sequence for H. Thus, we obtain by the hypothesis kSΛf − SΘfk ≤ λkSΛfk + µkSΘfk + γ(Xj∈I jkΛj πwj fk)2) v2 1 2 . Therefore, by (3) kf − SΘS−1 Λ fk ≤ λkfk + µkSΘS−1 Λ fk + γ(cid:0)Xj∈J √A(cid:17)kfk + µkSΘS−1 Λ fk. γ ≤ (cid:16)λ + JkΛj πWj S−1 v2 Λ fk2(cid:1) 1 2 Since 0 ≤ max{λ + γ√A SΘ is invertible and we get , µ} < 1, then by Lemma 1, SΘS−1 Λ and consequently kS−1 Θ k ≤ kS−1 Λ kkSΛS−1 Θ k ≤ So, the proof is completed. 1 + µ A(cid:0)1 − (λ + γ√A )(cid:1) . (cid:3) Corollary 5. The optimal lower and upper bounds of Θ which defined in Theorem 5.1 are kS−1 Corollary 6. Let Λ be a g-fusion frame for H with bounds A, B and {Θj ∈ B(H, Hj)}j∈J be a sequence of operators. If there exists a constant 0 < R < A such that Θ k−1 and kSΘk, respectively. jkπWj Λ∗j Λj πWj f − πWj Θ∗j ΘjπWj fk ≤ Rkfk v2 Xj∈J WEIGHTED RIESZ BASES IN G-FUSION FRAMES AND THEIR PERTURBATION 11 for all f ∈ H, then Θ := (Wj, Θj , vj) is a g-fusion frame for H with bounds A − R and min{B + Rr B A , R + √B}. Proof. It is easy to check that Pj∈J v2 f ∈ H. Thus, we obtain for each f ∈ H j πWj Θ∗j ΘjπWj f is converges for any Xj∈J jkπWj Λ∗j Λj πWj f − πWj Θ∗j Θj πWj fk ≤ Rkfk v2 ≤ R √A(cid:0)Xj∈J v2 jkΛj πWj fk2(cid:1) 1 2 and also v2 Xj∈J jkπWj Θ∗j ΘjπWj fk ≤ Rkfk +Xj∈J ≤ (R + √B)kfk. By using Theorem 5.1 with λ = µ = 0 and γ = R√A jkπWj Λ∗j ΛjπWj fk v2 , the proof is completed. (cid:3) The following is another version of perturbation of g-fusion frames. Theorem 5.2. Let Λ be a g-fusion frame for H with bounds A, B and {Θj ∈ B(H, Hj)}j∈J be a sequence of operators such that for any finite subset I ⊆ J and for each {fj}j∈J ∈ H2, kXj∈I vj(cid:0)πWj Λ∗j fj − πWj Θ∗j fj(cid:1)k ≤ λkXj∈I vj(πWj Λ∗j fj)k+ + µkXj∈I vj(πWj Θ∗j fj)k + γ(Xj∈I kfjk2) 1 2 , where 0 ≤ max{λ + γ√A for H with bounds , µ} < 1, then Θ := (Wj, Θj , vj) is a g-fusion frame A(cid:16) 1 −(cid:0)λ + γ√A(cid:1)2 1 + µ Proof. Let {fj}j∈J ∈ H2, then j fjk ≤ kXj∈I kXj∈I vjπWj Θ∗ (cid:17) and B(cid:16) 1 + λ + γ√B 1 − µ . (cid:17)2 vj(cid:0)πWj Λ∗ j fj − πWj Θ∗ j fj(cid:1)k + kXj∈I vjπWj Λ∗ j fjk ≤ (1 + λ)kXj∈I vjπWj Λ∗ j fjk+ + µkXj∈I vjπWj Θ∗ j fjk + γ(cid:0)Xj∈I 1 2 . kfjk2(cid:1) 12 GH. RAHIMLOU, V. SADRI, AND R. AHMADI Hence, kXj∈I Let vjπWj Θ∗j fjk ≤ 1 + λ 1 − µkXj∈I √B + ≤ (cid:16) 1 + λ 1 − µ vj πWj Λ∗j fjk + 1 − µ(cid:17)(cid:16)Xj∈I γ γ 1 − µ(cid:0)Xj∈I kfjk2(cid:17) . 1 2 1 2 kfjk2(cid:1) TΘ : H −→ H2 TΘ{fj}j∈J = Xj∈I vj πWj Θ∗j fj. Therefore, TΘ is a well-defined and bounded. Then, by Theorem 2.1, Θ is a g-fusion Bessel sequence. Suppose that G := TΘT ∗ΛS−1 Λ . Then, we get by the hypotesis and (9) for fj := ΛjπWj S−1 Λ f , kf − Gfk ≤ λkfk + µkGfk + γ(cid:0)Xj∈J √A(cid:1)kfk + µkGfk. ≤ (cid:0)λ + γ jkΛj πWj S−1 v2 Λ fk2(cid:1) 1 2 Since 0 ≤ max{λ + γ√A invertible and , µ} < 1, by Lemma 1, G and consequently TΘT ∗Λ is kG−1k ≤ 1 + µ 1 − (λ + γ√A ) . Now, let f ∈ H and we have kfk4 = hGG−1f, fi2 2 Λ (G−1f ), f(cid:11)(cid:12)(cid:12) jkΘj πWj fk2 v2 j πWj Θ∗j ΛjπWj S−1 v2 = (cid:12)(cid:12)(cid:10)Xj∈J ≤ Xj∈J kΛj πWj S−1 ≤ A−1kG−1fk2Xj∈J Λ (G−1f )k2.Xj∈J jkΘj πWj fk2 v2 and this copmletes the proof. (cid:3) Theorem 5.3. Let Λ be a g-fusion frame for H with bounds A, B and {Θj ∈ B(H, Hj)}j∈J be a sequence of operators. If there exists a constant 0 < R < A such that for all f ∈ H, then Θ := (Wj, Θj , vj) is a g-fusion frame for H with bounds √R)2 and (√R + √B)2. jkΛj πWj f − ΘjπWj f(cid:1)k2 ≤ Rkfk2 v2 Xj∈J (√A − WEIGHTED RIESZ BASES IN G-FUSION FRAMES AND THEIR PERTURBATION 13 Proof. Let f ∈ H. We can write k{vj ΘjπWj f}j∈Jk2 ≤ k{vjΘj πWj f − vjΛjπWj f}j∈Jk2 + k{vjΛj πWj f}j∈Jk2 ≤ (√R + √B)2kfk2 and also k{vj ΘjπWj f}j∈Jk2 ≥ k{vjΛj πWj f}j∈Jk2 − k{vjΘjπWj f − vjΛj πWj f}j∈Jk2 ≥ (√A − √R)2kfk2. Thus, these complete the proof. (cid:3) 6. Conclusions In this paper, we could transfer some common properties of g-frames to g-fusion frames. First, we reviewed gf-Riesz and orthonormal bases and showed that the weights have basic roles in their definitions. After- wards, characterizations of these frames were presented in Section 4 by Sun's method by using the operator (9). Finally, in Section 5, we introduced the perturbation of g-fusion frames and reviewed some useful statements. References [1] H. Blocsli, H. F. Hlawatsch, and H. G. Fichtinger, Frame-Theoretic analysis of over- sampled filter bank, IEEE Trans. Signal Processing 46 (12), 3256- 3268, 1998. [2] E. J. Candes, and D. L. Donoho, New tight frames of curvelets and optimal represen- tation of objects with piecwise C 2 singularities. Comm. Pure and App. Math. 57(2), 219-266, 2004. [3] P. G. Casazza, and O. Christensen, Perturbation of Operators and Application to Frame Theory. J. Fourier Anal. Appl. 3, 543-557, 1997. [4] P. G. Casazza, and G. Kutyniok, Frames of Subspaces. Contemp. Math. 345, 87-114, 2004. [5] P. G. Casazza, G. Kutyniok, and S. Li, Fusion Frames and distributed processing. Appl. comput. Harmon. Anal. 25(1), 114-132, 2008. [6] O. Christensen, Frames and Bases: An Introductory Course (Applied and Numerical Harmonic Analysis). Birkhauser, 2008. [7] J. Diestel, Sequences and series in Banach spaces. Springer-Verlag, New York, 1984. [8] R.J. Duffin, and A. C. Schaeffer, A class of nonharmonik Fourier series. Trans. Amer. Math. Soc. 72(1), 341-366, 1952. [9] M. H. Faroughi, and R. Ahmadi, Some Properties of C-Frames of Subspaces. J. Non- linear Sci. Appl. 1(3), 155-168, 2008. [10] H. G. Feichtinger, and T. Werther, Atomic Systems for Subspaces. Proceedings SampTA. Orlando, FL, 163-165, 2001. [11] L. Gavruta, Frames for Operators. Appl. Comp. Harm. Annal. 32, 139-144, 2012. [12] S. B. Heineken, P. M. Morillas, A. M. Benavente, and M. I. Zakowich, Dual Fusion Frames. arXiv: 1308.4595v2 [math.CA] 29 Apr 2014. [13] H. Heuser, Functional Analysis. John Wiley, New York, 1991. [14] M. Khayyami, and A. Nazari, Construction of Continuous g-Frames and Continuous Fusion Frames. SCMA 4(1), 43-55, 2016. [15] A. Najati, M. H. Faroughi, and A. Rahimi, g-frames and stability of g-frames in Hilbert spaces. Methods of Functional Analysis and Topology. 14(3), 305–324, 2008. 14 GH. RAHIMLOU, V. SADRI, AND R. AHMADI [16] V. Sadri, Gh. Rahimlou, R. Ahmadi and R. Zarghami Farfar, Generalized Fusion Frames in Hilbert Spaces. Submited 2018. [17] W. Sun, G-Frames and G-Riesz bases, J. Math. Anal. Appl.326 , 437-452, 2006. Department of Mathematics, Faculty of Tabriz Branch, Technical and Vo- cational University (TUV), East Azarbaijan , Iran, E-mail address: [email protected] Department of Mathematics, Faculty of Tabriz Branch, Technical and Vo- cational University (TUV), East Azarbaijan , Iran E-mail address: [email protected] Institute of Fundamental Sciences, University of Tabriz, , Iran, E-mail address: [email protected]
1309.7469
1
1309
2013-09-28T15:55:19
Operators on Banach Spaces of Bourgain-Delbaen Type
[ "math.FA" ]
We begin by giving a detailed exposition of the original Bourgain-Delbaen construction and the generalised construction due to Argyros and Haydon. We show how these two constructions are related, and as a corollary, are able to prove that there exists some $\delta > 0$ and an uncountable set of isometries on the original Bourgain-Delbaen spaces which are pairwise distance $\delta$ apart. We subsequently extend these ideas to obtain our main results. We construct new Banach spaces of Bourgain-Delbaen type, all of which have $\ell_1$ dual. The first class of spaces are HI and possess few, but not very few operators. We thus have a negative solution to the Argyros-Haydon question. We remark that all these spaces have finite dimensional Calkin algebra, and we investigate the corollaries of this result. We also construct a space with $\ell_1$ Calkin algebra and show that whilst this space is still of Bourgain-Delbaen type with $\ell_1$ dual, it behaves somewhat differently to the first class of spaces. Finally, we briefly consider shift-invariant $\ell_1$ preduals, and hint at how one might use the Bourgain-Delbaen construction to produce new, exotic examples.
math.FA
math
Operators on Banach Spaces of Bourgain -- Delbaen Type 3 1 0 2 p e S 8 2 ] . A F h t a m [ 1 v 9 6 4 7 . 9 0 3 1 : v i X r a Matthew Tarbard St John's College University of Oxford A thesis submitted for the degree of Doctor of Philosophy Michaelmas 2012 To my family, for all their love and support. Acknowledgements Firstly, I would like to thank my supervisor, Professor Richard Haydon, whose invaluable suggestions, help and support made the work in this thesis possible. I especially thank the hard work of my examiners, Professor Charles Batty and Dr Matthew Daws, who provided a very intellectually stimulating viva. Their careful reading of the the- sis also led to the correction of several errors. I have been fortunate enough to present some of the results in this thesis at various semi- nars and conferences. These opportunities, and the people I have met through these events, helped shape my thoughts and ideas, which led to some of the results in this thesis. I would therefore like to thank Dr Bunyamin Sari for inviting me to the BIRS Banach Space Theory Workshop 2012, where I learnt a great deal about Bourgain-Delbaen spaces and other interesting problems. I would also like to thank Professor William B. Johnson, who brought to my attention at the BIRS workshop a corollary of my result that I had not originally spotted. I would also like to thank Professor Charles Batty, Dr Matthew Daws and Dr Niels Laustsen for inviting me to speak at seminars at the Universities of Oxford, Leeds and Lancaster respectively. Special thanks also go to Dr Richard Earl, who has provided me with fantastic teaching opportunities at Worcester College throughout my DPhil, as well as introducing me to a number of great people, notably, Dr Brian King and Dr Martin Galpin. I am thankful to Richard, Brian and Martin for introducing me to bridge, and for the considerable generosity they have shown me (particularly at the Worcester College Bar)! I am of course very grateful to all my friends and family, who have supported me through- out my DPhil. I am particularly thankful to Victoria Mason, who has provided endless encouragement and support whilst writing this thesis, and contributed significantly to the non-academic aspect of my life. Special thanks also go to Stephen Belding, David Hewings, Tanya Gupta and Carly Leighton for always being around to socialise with and making Oxford a more interesting place to live. Finally, I'd like to acknowledge that my doctoral studies were funded by the EPSRC (En- gineering and Physical Sciences Research Council). Abstract The research in this thesis was initially motivated by an outstanding problem posed by Argyros and Haydon. They used a generalised version of the Bourgain-Delbaen construction to construct a Banach space XAH for which the only bounded linear operators on XAH are compact perturbations of (scalar multiples of) the identity; we say that a space with this property has very few operators. The space XAH possesses a number of additional interesting properties, most notably, it has (cid:96)1 dual. Since (cid:96)1 possesses the Schur property, weakly compact and norm compact operators on XAH coincide. Combined with the other properties of the Argyros-Haydon space, it is tempting to conjecture that such a space must necessarily have very few operators. Curiously however, the proof that XAH has very few operators made no use of the Schur property of (cid:96)1. We therefore arrive at the following question (originally posed in [2]): must a HI, L∞, (cid:96)1 predual with few operators (every operator is a strictly singular perturbation of λI) necessarily have very few operators? We begin by giving a detailed exposition of the original Bourgain-Delbaen construction and the generalised construction due to Argyros and Haydon. We show how these two constructions are related, and as a corollary, are able to prove that there exists some δ > 0 and an uncountable set of isometries on the original Bourgain-Delbaen spaces which are pairwise distance δ apart. We subsequently extend these ideas to obtain our main results. We construct new Banach spaces of Bourgain-Delbaen type, all of which have (cid:96)1 dual. The first class of spaces are HI and possess few, but not very few operators. We thus have a negative solution to the Argyros-Haydon question. We remark that all these spaces have finite dimensional Calkin algebra, and we investigate the corollaries of this result. We also construct a space with (cid:96)1 Calkin algebra and show that whilst this space is still of Bourgain-Delbaen type with (cid:96)1 dual, it behaves somewhat differently to the first class of spaces. Finally, we briefly consider shift-invariant (cid:96)1 preduals, and hint at how one might use the Bourgain-Delbaen construction to produce new, exotic examples. Contents 1 Introduction 1.1 Historical background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Overview of the thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.3 Notation and Elementary Definitions . . . . . . . . . . . . . . . . . . . . . . 1.4 Preliminary Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4.1 Basic sequence techniques . . . . . . . . . . . . . . . . . . . . . . . . 1.4.2 Separation Theorems . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4.3 Hereditary Indecomposability . . . . . . . . . . . . . . . . . . . . . . 1.4.4 Elementary Results from Operator Theory . . . . . . . . . . . . . . 1.4.5 Complexification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.4.6 Strictly Singular Operators . . . . . . . . . . . . . . . . . . . . . . . 2 The Bourgain-Delbaen Construction 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2.2 The generalised Bougain-Delbaen Construction . . . . . . . . . . . . . . . . 2.3 The Bourgain-Delbaen Construction . . . . . . . . . . . . . . . . . . . . . . 2.4 Connecting the two constructions . . . . . . . . . . . . . . . . . . . . . . . . 2.5 The operator algebras for the original Bourgain-Delbaen spaces . . . . . . . 2.5.1 Construction of the basic operator . . . . . . . . . . . . . . . . . . . 2.5.2 Non-separability of L(X) . . . . . . . . . . . . . . . . . . . . . . . . 3 Spaces with few but not very few operators 3.1 The Main Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2 Corollaries of the Main Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.1 On the structure of the closed ideals in L(Xk). 3.2.2 The Calkin algebra L(Xk)/K(Xk). 3.2.3 Commutators in Banach Algebras . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.2.4 Invariant subspaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.3 The Basic Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . i 1 1 3 5 6 6 9 10 12 16 17 28 28 29 37 39 48 49 54 60 60 63 63 64 65 66 66 3.4 Rapidly Increasing Sequences and the operator S : Xk → Xk . . . . . . . . . 3.5 Operators on the Space Xk . . . . . . . . . . . . . . . . . . . . . . . . . . . 3.6 Strict Singularity of S : Xk → Xk 3.7 The HI Property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 A Banach space with (cid:96)1 Calkin algebra 4.1 The Main Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 The Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.3 L(X∞)/K(X∞) is isometric to (cid:96)1(N0) . . . . . . . . . . . . . . . . . . . . . . 5 Shift invariant (cid:96)1 preduals 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.2 Connection to the BD construction . . . . . . . . . . . . . . . . . . . . . . . 5.3 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References 78 88 95 98 100 100 102 119 121 121 127 133 134 ii Chapter 1 Introduction 1.1 Historical background Knowledge about the types of bounded linear operator that exist from a Banach space into itself can reveal much about the structure of the underlying Banach space. In particular, it is possible to infer a great deal about the structure of the space X when its operator algebra, L(X), is 'small'. The first substantial results in this direction are those of Gowers and Maurey, presented in [16] and [15]. As a motivational example, we consider the space Xgm constructed by Gowers and Maurey in [16]. Here it was shown that all (bounded linear) operators defined on a subspace Y of Xgm (and mapping into Xgm) are strictly singular perturbations of the inclusion operator iY : Y → Xgm. More precisely, every such operator is expressible in the form λiY + S, where λ is a scalar and S : Y → Xgm is a strictly singular operator. We shall give a precise definition of a strictly singular operator in Section 1.4.6. For now, it suffices to think of the strictly singular operators as those which are, in some sense, small. Indeed, it is a well known result of Fredholm theory that strictly singular perturbations of Fredholm operators are still Fredholm, with the same index. The representation of operators on subspaces of Xgm just discussed allows us to infer some remarkable structural properties of the space Xgm. We obtain the following: 1. Xgm is not decomposable, that is, it cannot be written as a (topological) direct sum of two of its infinite dimensional subspaces. This is because a non-trivial projection is not expressible as a strictly singular perturbation of the identity. 2. In fact, we conclude by the same argument that Xgm is hereditarily indecomposable, It that is to say, no closed infinite dimensional subspace of Xgm is decomposable. follows that no subspace of Xgm has an unconditional basis, i.e. Xgm has no un- conditional basic sequence. Indeed, if (ei)∞ i=1 were an unconditional basis for some 1 subspace Y ⊆ Xgm, then we could decompose Y as Y = [e2i]∞ i=1. (We remark, and contrast this to, the well known fact that every Banach space contains a i=1 ⊕ [e2i−1]∞ basic sequence.) 3. Xgm is not isomorphic to any of its proper subspaces. Indeed, it follows from the operator representation and elementary results from Fredholm theory that every op- erator from Xgm to itself is either strictly singular (and thus not an isomorphism), In the latter case, if Xgm were isomorphic to a proper or Fredholm with index 0. subspace Y , then there would be an isomorphic embedding T : Xgm → Xgm that maps onto Y (we simply take an isomorphism from Xgm → Y and then compose with the inclusion operator sending Y into Xgm). In particular T is not strictly singular and so must be Fredholm with index 0. However, this is clearly not the case as T is injective but not onto Xgm. One could of course consider the relationship between a Banach space and its operator algebra from a different perspective to that just described. Instead of assuming we have some well behaved properties of the operator algebra and asking what consequence this has for the structure of the underlying Banach space, we may choose to impose some kind of structural conditions on a Banach space and see what affect this has on the associated operator algebra. In 1980, Bourgain and Delbaen [7, 6] introduced two classes of separable Banach spaces which have 'well-behaved' finite dimensional structure, specifically they are L∞-spaces (we refer the reader to Definition 2.2.1 for more details). These spaces have many interesting properties and recently, Argyros and Haydon (see [2]) have managed to modify the original construction of Bourgain and Delbaen to construct a space, XAH, which solves the scalar- plus-compact problem. More precisely XAH is a (hereditarily indecomposable) L∞ space with a Schauder basis, (cid:96)1 dual and L(XAH) = RI ⊕ K(XAH) (where, as usual, K(XAH) denotes the subspace of L(XAH) of compact operators). The proof that all operators are compact perturbations of the identity made essential use of the finite dimensional structure of the space and the specific structure of Bourgain-Delbaen spaces, which embeds some very explicit finite dimensional (cid:96)∞−spaces into XAH. Interestingly, the Schur property of (cid:96)1 plays no role in the proof that XAH has the scalar-plus-compact property; this will be the main subject of Chapter 3, and so we will defer any further discussion of this until then. As well as solving the scalar-plus-compact problem, the space XAH is interesting for many reasons. It is well known that the space of compact operators K(X) is a separable subspace of L(X) whenever X is a Banach space which has the approximation property and a separable dual space. It follows that the operator algebra L(XAH) is separable. Moreover, 2 it is elementary to show that the compact operators on XAH are a complemented subspace of L(XAH). Another interesting property of the space XAH is that every operator in L(XAH) has a proper closed invariant subspace, the first space for which such a result is known. Indeed, Aronszajn and Smith showed in [3] that every compact operator on a Banach space has a proper closed invariant subspace. Clearly if a subspace is invariant under some operator, then it is invariant under that operator perturbed by some scalar multiple of the identity. In particular, every operator of the form λI + K, with K compact admits a proper, closed invariant subspace and the claim about XAH follows. (One may also prove the claim by appealing to a result of Lomonosov, [23], which states that every operator that commutes with a non-zero compact operator must have a proper invariant subspace.) 1.2 Overview of the thesis In this thesis, we will continue to investigate the interplay between the structure of a Banach space and its associated operator algebra. The question originally motivating the research that follows was whether or not the (cid:96)1 duality of the Argyros-Haydon space XAH forces the scalar-plus-compact property. Stated slightly more precisely, must every strictly singular operator on a space with (cid:96)1 dual necessarily be compact if the operator algebra of the space is small? Standard facts about basic sequences and strictly singular operators will play a prominent role in this work, so we begin by recalling in Section 1.4 some well known results from Banach space theory and operator theory which will be used throughout the thesis. Whilst we present nothing new in this chapter, many of the results and proofs are difficult to find in the existing literature. For completeness, results we will frequently rely on are stated and we either provide references to their proofs or outline them here. Chapter 2: Given the success of Argyros and Haydon in modifying the Bourgain-Delbaen construction to solve the scalar-plus-compact problem, the spaces of 'Bourgain-Delbaen type' are a sensible starting place to look for examples of exotic Banach spaces having (cid:96)1 dual and well behaved operator algebras. The construction employed by Argyros and Haydon in [2] appears to be somewhat different to the original construction of Bourgain and Delbaen in, for example, [7, 6]. We begin Chapter 2 by investigating the relationship between the two constructions. Specifically, we show that the Argyros-Haydon construction in [2] is essentially a generalised 'dualised version' of the original Bourgain-Delbaen construction. Of course, this is certainly not a new result, however, to the best knowledge of the author, the precise details of this connection cannot be found in any of the existing literature. The Argyros-Haydon approach to the Bourgain-Delbaen construction enables us to see how we might go about constructing interesting operators on spaces of Bourgain-Delbaen 3 type. Having understood how the original Bourgain-Delbaen construction fits into that of the Argyros-Haydon construction, we present our first new result at the end of Chapter 2: for any of the original Bourgain-Delbaen spaces, X, constructed in [6], there exists a constant c > 0 and an uncountable set of isometries from X to itself, such that any two (non-equal) isometries are separated by distance at least c with respect to the operator norm. In particular, L(X) is non-separable for all of the original Bourgain-Delbaen spaces. This answers a question of Beanland and Mitchell ([4]). Chapter 3: In Chapter 3, we address a question left over from the work of Argyros and Haydon in [2], namely, must the strictly singular and compact operators on a space with (cid:96)1 dual coincide if every operator on the space is a strictly singular perturbation of the identity? The main result of this chapter is a negative solution to this question. In fact, we exhibit for each natural number k, a Banach space Xk with (cid:96)1 dual such that the Calkin algebra L(Xk)/K(Xk) is k-dimensional, with basis {I + K(Xk), S + K(Xk), . . . Sk−1 + K(Xk)}. Here S is some strictly singular operator on Xk. The results are, in some sense, continuations of the work of Gowers and Maurey and Argyros and Haydon already mentioned; indeed, for each space Xk, we have a representation of a general operator as some polynomial of S, possibly perturbed by a compact operator. We immediately obtain from this an interesting corollary concerning the norm-closed ideal structure of L(Xk); it is a finite, totally ordered chain of ideals. We also observe that the space X2 provides a counterexample for an open conjecture of Johnson concerning the form of commutators in the operator algebra. The material in this chapter is an expanded version of the published paper [28]. Chapter 4 : Having shown it is possible to construct Banach spaces with Calkin algebras of any finite dimension, the natural generalisation is to consider what infinite dimensional Calkin algebras can be obtained. In Chapter 4 we exhibit a Banach space, X∞, which has Calkin algebra isometric (as a Banach algebra) to the algebra (cid:96)1(N0). This generalises a result of Gowers ([15]) where a Banach space X was constructed with the property that the quotient algebra L(X)/SS(X) is isomorphic as a Banach algebra to (cid:96)1(Z). (Here SS(X) denotes the strictly singular operators on X.) The ideas used to construct X∞ are very similar to those of Chapter 3, though the proof is a little harder, requiring a combination of the arguments from Chapter 3 and the Gowers' paper [15]. Chapter 5 : The reader will notice that all the Banach spaces constructed in this thesis have (cid:96)1 dual. We conclude the thesis by discussing the exoticness of (cid:96)1 preduals and some open research problems. This leads us to discuss some recently published work by Daws, Haydon, Schlumprecht and White on so-called 'shift-invariant preduals of (cid:96)1' (see [9] for more details). We show that one of the spaces constructed in the aforementioned paper can in fact be considered (in some sense) as being obtained from a specific Bourgain-Delbaen 4 construction. It is hoped by the author that this insight may eventually lead to new, exotic (cid:96)1 preduals. 1.3 Notation and Elementary Definitions Generally notation will be introduced as and when needed. Nevertheless, in this section we remind the reader of some standard notation and state the conventions and definitions that will be used throughout this thesis without further explanation. • All Banach spaces are assumed to be over the real scalar field unless explicitly stated. Given a subset of vectors of a Banach space X, {xi}i∈I , we denote by [xi]i∈I the smallest closed subspace of X generated by the (xi)i∈I . In other words, [xi]i∈I is the closed linear span of the (xi)i∈I . • The continuous dual space of a Banach space X is denoted by X∗ as usual, unless explicitly stated otherwise. (In Chapters 3 and 4 we introduce a 'star notation' in an attempt to be consistent with this notation.) • We occasionally make use of the 'angle bracket notation': for vectors x∗ ∈ X∗ and x ∈ X,(cid:104)x∗, x(cid:105) = x∗(x), the evaluation of the functional x∗ at x. • Given a closed subspace M of a Banach space X, the quotient space X/M is the Banach space which as a set consists of the cosets {x + M}. We equip it with the obvious vector space operations and use the norm (cid:107)x + M(cid:107) = dist(x, M ) = inf{(cid:107)x − m(cid:107) : m ∈ M}. • All operators between Banach spaces are assumed to be continuous and linear. For an operator T : X → Y between Banach spaces (X,(cid:107)(cid:107)X ) and (Y,(cid:107)(cid:107)Y ), we denote by (cid:107)T(cid:107) the usual operator norm of T , that is, (cid:107)T(cid:107) = sup{(cid:107)T x(cid:107)Y : (cid:107)x(cid:107)X ≤ 1}. • With the same notation as above, T ∗ : Y ∗ → X∗ denotes the dual operator defined by T ∗y∗(x) = y∗(T x) for all x ∈ X and y∗ ∈ Y ∗. • An operator T : X → Y is said to be an isomorphism if there exist strictly positive constants a, b > 0, such that a(cid:107)x(cid:107)X ≤ (cid:107)T x(cid:107)Y ≤ b(cid:107)x(cid:107)X Note we do not require an isomorphism to be onto. When we wish to be explicit about an isomorphism that is not surjective, we use the term isomorphic embedding . 5 • An onto operator T : X → Y between normed spaces X and Y is said to be a quotient operator if the (bounded) linear operator T : X/ Ker T → Y , defined by T (x+Ker T ) = T x is an isomorphism, i.e. it has a continuous inverse. It is easy to check that this is equivalent to the condition 'there exists a constant M > 0 such that BY (0; 1) ⊂ M T (BX (0; 1))'. 1.4 Preliminary Results 1.4.1 Basic sequence techniques It would be impossible to include all the known results about Schauder bases and basic sequences. Nevertheless, the notion of a Schauder basis and selected results from basic sequence techniques in Banach space theory feature prominently in this thesis. In this section, we state the results we will need throughout the rest of this work. Definition 1.4.1. Let X be an infinite dimensional Banach space. • A sequence (xn)∞ n=1 ⊂ X is said to be a Schauder basis for X if whenever x ∈ X, n=1 αnxn (where n=1 is a basic sequence in X if there exists a unique sequence of scalars (αn)∞ the sum converges in the norm topology). We say (xn)∞ it is a Schauder basis for the closed linear span [xn : n ∈ N]. n=1 such that x =(cid:80)∞ • More generally, a sequence (Fn)∞ n=1 of non-trivial subspaces of X is said to form a Schauder decomposition of X, written X = ⊕n∈NFn, if every x ∈ X can be written uniquely as(cid:80)∞ n=1 xn with xn ∈ Fn for all n. In a Banach space X we say that vectors yj are successive linear combinations, or that (yj) is a block sequence of a basic sequence (xi) if there exist 0 = q1 < q2 < ··· such that, for all j ≥ 1, yj is in the linear span [xi : qj−1 < i ≤ qj]. If we may arrange that yj ∈ [xi : qj−1 < i < qj] we say that (yj) is a skipped block sequence. More generally, if X n∈N Fn we say that (yj) is a block sequence (resp. a skipped block sequence) with respect to (Fn) if there exist 0 = q0 < q1 < ··· such that yj is Fn). A block subspace is the closed subspace generated has a Schauder decomposition X =(cid:76) in(cid:76) Fn (resp. (cid:76) qj−1<n≤qj qj−1<n<qj by a block sequence. Finally, a basic sequence (xn) is called unconditional if every permutation of the sequence is a Schauder basis for [xn : n ∈ N]. It is well known (see, for example, [1, Proposition 1.1.9]) that a sequence (xn)∞ n=1 of non-zero vectors in a Banach space X is a basic sequence if, and only if, there exists a 6 i=1 are scalars, (cid:107) m(cid:88) constant M ≥ 1 such that whenever m < n are natural numbers and (λi)n λixi(cid:107) ≤ M(cid:107) n(cid:88) defined by Pk : [xn : n ∈ N] → [xn : n ≤ k], (cid:80)∞ n : [xn : n ∈ N] → K defined by (cid:80)∞ From this, it follows easily that there exist uniformly bounded linear projections (Pk)∞ k=1 n=1 anxn; the constant, supk∈N (cid:107)Pk(cid:107), is known as the basis constant. We note that the co-ordinate functionals, m=1 amxm (cid:55)→ an (also known in the literature as x∗ biorthogonal functionals) are continuous and refer the reader to [25, Theorem 4.1.14] for n=1 anxn (cid:55)→ (cid:80)k λixi(cid:107). i=1 i=1 further details of these facts. When the Banach space has an unconditional basis, (xn)∞ n=1, it admits a number of non-trivial projections and the associated operator algebra has a rich structure. Indeed, it is known that for any (ξn)∞ n=1 ξnanxn is a bounded operator. In particular, for every A ⊂ N, the linear projection PA : X → X n∈A anxn is bounded; in fact, supA⊂N (cid:107)PA(cid:107) < ∞. More details n=1 ∈ (cid:96)∞, T(ξn) : X → X defined by(cid:80)∞ n=1 anxn (cid:55)→(cid:80)∞ n=1 anxn (cid:55)→(cid:80) defined by(cid:80)∞ and proofs of these facts can be found in [1, Proposition 3.1.3, Remark 3.1.5]. Frequently we will have some condition that needs to be checked for all closed infinite dimensional subspaces of a Banach space. When the Banach space has a Schauder basis, it is often possible to show that it is sufficient to check the condition only on block subspaces. We will make use of this idea later in the thesis. The proofs of such results generally rely on the following technical result. Proposition 1.4.2. Let X be a Banach space with Schauder basis (en)∞ n=1 and suppose Y is a closed, infinite dimensional subspace of X. Denote by K the basis constant of the basis n=1 in Y with (cid:107)yn(cid:107) = 1 for all n (en)∞ and a block sequence (xn)∞ n=1. For every 0 < θ < 1, we can find a sequence (yn)∞ n=1 in X satisfying ∞(cid:88) n=1 2K (cid:107)yn − xn(cid:107) (cid:107)xn(cid:107) < θ n=1,(cid:80)∞ n=1 is a basic sequence equivalent to (xn)∞ n=1 anyn converges if and only if(cid:80)∞ The sequence (yn)∞ scalars, (an)∞ Proof. Note that for any n ∈ N, there must exist a non-zero vector y ∈ Y with e∗ m(y) = 0 for all m ≤ n (here, the e∗ m are the biorthogonal functionals of the basis (en)). Indeed, if not, the (bounded) projection Pn : Y → [em : m ≤ n] is injective; this contradicts the fact that Y is infinite dimensional. n=1, i.e. given a sequence of n=1 anxn converges. It follows from the above observation that we can pick a sequence (y(cid:48) n(cid:107) = 1 for every n and e∗ n=1 ⊆ Y with n)∞ n) = 0 for all m ≤ n. By the proof of the Bessaga-Pelczy´nski m(y(cid:48) (cid:107)y(cid:48) 7 Selection Principle given in [1, Proposition 1.3.10], there is a subsequence of the y(cid:48) say, and a block basic sequence (xk) (with respect to the basis (en)) satisfying n, (y(cid:48) ) nk ∞(cid:88) k=1 2K (cid:107)y(cid:48) nk − xk(cid:107) (cid:107)xk(cid:107) < θ The first part of the proposition is proved by simply setting yk = y(cid:48) block basic sequence (xk)∞ k=1 as above. nk To see that (yn)∞ n=1, it is enough, by standard results, to show that there is an isomorphism from X to X which n=1 is a basic sequence equivalent to the basic sequence (xn)∞ and choosing the sends xn to yn for every n. This is proved in [1, Theorem 1.3.9]. The Banach spaces constructed later in this thesis rely heavily on the following well- known proposition: Proposition 1.4.3. Suppose that (x∗ sociated to a basis (xn)∞ (with basis constant no bigger than that of (xn)∞ n=1). n=1 of a Banach space X. Then (x∗ n)∞ n=1 is the sequence of biorthogonal functionals as- n)∞ n=1 is a basic sequence in X∗ If we denote by H := [x∗ n : n ∈ N] the closed subspace of X∗ formed by taking the closed linear span of the biorthogonal vectors, then X embeds isomorphically into H∗ under the natural mapping j : X → H∗, x (cid:55)→ j(x)H (where j : X → X∗∗ is the canonical embedding of X into its bidual). Proof. See [1, Proposition 3.2.1, Lemma 3.2.3 and Remark 3.2.4]. We conclude this brief overview of basic sequences by recalling the notions and some basic facts about shrinking and boundedly-complete bases. Definition 1.4.4. Let X be a Banach space. tionals (x∗ • A basis (xn)∞ n=1 of X is said to be shrinking if the sequence of its biorthogonal func- n=1 is a basis for X∗. n)∞ • A basis (xn)∞ n=1 of X is said to be boundedly-complete if whenever (an)∞ n=1 is a sequence of scalars such that the series(cid:80)∞ n=1 anxn converges. (cid:107) sup N anxn(cid:107) < ∞ We will use the following results repeatedly. Proposition 1.4.5. A basis (xn)∞ bounded block basic sequence of (xn)∞ n=1 is weakly null. ∞(cid:88) n=1 8 n=1 of a Banach space X is shrinking if and only if every n=1 be a basis for a Banach space X with biorthogonal functionals Proof. See [1, Proposition 3.2.7]. Theorem 1.4.6. Let (xn)∞ n)∞ (x∗ n=1. The following are equivalent. (i) (xn)∞ n=1 is a boundedly-complete basis for X, (ii) (x∗ n)∞ n=1 is a shrinking basis for H := [x∗ (iii) The map j : X → H∗ defined by j(x)(h) = h(x), for all x ∈ X and h ∈ H, is an onto n : n ∈ N], isomorphism. Proof. See [1, Theorem 3.2.10]. 1.4.2 Separation Theorems There are a number of separation theorems in the existing literature. Whilst we do not feel it is necessary to present the Hahn-Banach Theorem as found in almost any introductory textbook on functional analysis, we find it convenient to state two separation theorems. These results are sufficiently general to cover all cases we will need in this thesis. Theorem 1.4.7 (Hahn, Banach). Let C be a closed convex set in a Banach space X. If x0 /∈ C then there is f ∈ X∗ such that Re(f (x0)) > sup{Re(f (x)) : x ∈ C}. Proof. See [11, Theorem 2.12]. Theorem 1.4.8. Let X be a topological vector space (TVS) and let C1 and C2 be non-empty convex subsets of X such that C2 has non-empty interior. If C1 ∩ C◦ 2 = ∅ then there is a continuous linear functional x∗ ∈ X∗ and a real number s such that (1) Re x∗x ≥ s for each x ∈ C1; (2) Re x∗x ≤ s for each x ∈ C2; (3) Re x∗x < s for each x ∈ C◦ 2 . Proof. See [25, Theorem 2.2.26]. 9 1.4.3 Hereditary Indecomposability We have already discussed hereditarily indecomposable (HI) spaces in the introduction. However, since we will want to prove that the spaces constructed in Chapter 3 are HI, we take this opportunity to formally state the HI condition. A Banach space X is indecomposable if there do not exist infinite-dimensional closed subspaces Y and Z of X such that X can be written as the topological direct sum X = Y ⊕Z; that is to say the bounded operator Y ⊕ext Z → Y ⊕ Z (y, z) (cid:55)→ y + z fails to have continuous inverse. X is hereditarily indecomposable (HI) if every closed subspace is indecomposable. It follows that X is HI if and only if, whenever Y and Z are infinite dimensional, closed subspaces of X, there does not exist δ > 0 such that (cid:107)y + z(cid:107) ≥ δ((cid:107)y(cid:107) + (cid:107)z(cid:107)) for all y ∈ Y, z ∈ Z. δ > 0 to exist, it would certainly have to be the case that Y ∩ Z = {0}.) Schauder basis, it is sufficient to check the condition on block subspaces. More precisely (We note in particular that were such a If X has a Proposition 1.4.9. Let X have a Schauder basis. Then X is HI if and only if whenever Y, Z are block subspaces, there does not exist δ > 0 such that (cid:107)y + z(cid:107) ≥ δ((cid:107)y(cid:107) + (cid:107)z(cid:107)) for all y ∈ Y, z ∈ Z. Proof. Clearly if X is HI then the condition on block subspaces holds. Conversely, suppose the condition on block subspaces holds and for contradiction that there exist infinite dimen- sional, closed subspaces Y , Z, with Y ∩ Z = {0} and such that Y ⊕ Z is a topological direct sum. Consequently, there exists δ > 0 such that (cid:107)y + z(cid:107) ≥ δ((cid:107)y(cid:107) +(cid:107)z(cid:107)) for all y ∈ Y, z ∈ Z. (Clearly for this to hold, we must have δ ≤ 1.) We choose 0 < ε < 1/4 and such that n=1 ⊂ Y , (zn)∞ n=1 ⊂ Z and block basic sequences (y(cid:48) 4ε 1−4ε < δ. By Proposition 1.4.2 we can find norm n)∞ n=1 with the n)∞ n=1, (z(cid:48) 1 sequences (yn)∞ property that 2K and ∞(cid:88) ∞(cid:88) n=1 (cid:107)yn − y(cid:48) n(cid:107) (cid:107)y(cid:48) n(cid:107) < ε (cid:107)zn − z(cid:48) n(cid:107) n(cid:107) < ε where K is the basis constant of the Schauder basis of X. (cid:107)z(cid:48) 2K n=1 Let Y (cid:48) be the block subspace generated by (y(cid:48) n)∞ n=1. Note the basis constant of the basic sequence (y(cid:48) n)∞ n=1 and Z(cid:48) the block subspace generated n) is at most K as it is a block by (z(cid:48) 10 basic sequence. It follows that if y(cid:48) :=(cid:80)∞ Consequently, if y(cid:48) :=(cid:80)∞ (we recall that (yn) and (y(cid:48) n=1 any(cid:48) n ∈ Y (cid:48), then y :=(cid:80)∞ n ∈ Y (cid:48) then an ≤ 2K(cid:107)y(cid:48) ∞(cid:88) n=1 any(cid:48) ∞(cid:88) n) are equivalent) and that n(cid:107)(cid:107)y(cid:48)(cid:107) for every n. n=1 anyn is a well-defined vector in Y (cid:107)y − y(cid:48)(cid:107) = (cid:107) an(yn − y(cid:48) n)(cid:107) ≤ 2K(cid:107)y(cid:48)(cid:107) n=1 n=1 n(cid:107) (cid:107)yn − y(cid:48) n(cid:107) < ε(cid:107)y(cid:48)(cid:107). (cid:107)y(cid:48) An analogous result holds for z(cid:48) ∈ Z(cid:48); we get a corresponding z ∈ Z with (cid:107)z − z(cid:48)(cid:107) < ε(cid:107)z(cid:48)(cid:107). We now suppose that y(cid:48) ∈ Y (cid:48) has norm 1, so that the corresponding vector y ∈ Y satisfies (cid:107)y(cid:107) > 1 − ε. If z(cid:48) ∈ Z(cid:48) is such that (cid:107)z(cid:48)(cid:107) ≤ 1 + δ then (cid:107)y(cid:48) − z(cid:48)(cid:107) = (cid:107)(y − z + y(cid:48) − y) − (z(cid:48) − z)(cid:107) ≥ (cid:107)(y − z) − (y − y(cid:48))(cid:107) − (cid:107)z(cid:48) − z(cid:107) ≥ (cid:107)y − z(cid:107) − (cid:107)y − y(cid:48)(cid:107) − (cid:107)z(cid:48) − z(cid:107) ≥ δ((cid:107)y(cid:107) + (cid:107)z(cid:107)) − ε(cid:107)y(cid:48)(cid:107) − ε(cid:107)z(cid:48)(cid:107) ≥ δ(cid:107)y(cid:107) − ε − ε(1 + δ) ≥ δ(1 − ε) − 2ε − εδ > δ 2 . where the final inequality follows by choice of ε. On the other hand, if z(cid:48) ∈ Z(cid:48) is such that (cid:107)z(cid:48)(cid:107) > 1 + δ, then (cid:107)y(cid:48) − z(cid:48)(cid:107) ≥ (cid:107)z(cid:48)(cid:107) − (cid:107)y(cid:48)(cid:107) > 1 + δ − 1 > δ 2 . We have therefore shown that for any y(cid:48) ∈ Y (cid:48) with (cid:107)y(cid:48)(cid:107) = 1 and z(cid:48) ∈ Z(cid:48), (cid:107)y(cid:48) − z(cid:48)(cid:107) > δ 2 . 2(cid:107)y(cid:48)(cid:107). By symmetry of 2(cid:107)z(cid:48)(cid:107). Consequently, 4 ((cid:107)y(cid:48)(cid:107) + (cid:107)z(cid:48)(cid:107)); this contradictions By scaling, this implies that whenever y(cid:48) ∈ Y (cid:48), z(cid:48) ∈ Z(cid:48), (cid:107)y(cid:48) + z(cid:48)(cid:107) ≥ δ the argument, we also have that whenever y(cid:48) ∈ Y (cid:48), z(cid:48) ∈ Z(cid:48), (cid:107)y(cid:48) + z(cid:48)(cid:107) ≥ δ for all y(cid:48) ∈ Y (cid:48), z(cid:48) ∈ Z(cid:48), (cid:107)y(cid:48) + z(cid:48)(cid:107) ≥ δ the assumed property satisfied by block subspaces. 2 max((cid:107)y(cid:48)(cid:107),(cid:107)z(cid:48)(cid:107)) ≥ δ Whilst there are a number of additional results known about HI spaces, the only other result we will need for the purposes of this thesis is the following proposition (which can be found in [2]). Proposition 1.4.10. Let X be an infinite dimensional Banach space. Then X is HI if and only if, for every pair Y, Z of closed, infinite-dimensional subspaces, and every  > 0, there exist y ∈ Y and z ∈ Z with (cid:107)y − z(cid:107) < ε(cid:107)y + z(cid:107). If X has a Schauder basis it is enough that the previous condition should hold for block subspaces. Proof. We claim that whenever Y and Z are infinite dimensional, closed subspaces, Y + Z fails to be a topological direct sum if and only if ∀ε > 0,∃y ∈ Y, z ∈ Z with (cid:107)y−z(cid:107) < ε(cid:107)y+z(cid:107). This clearly proves the first part of the proposition. Moreover, once we have proved this 11 claim, we can deduce that if the condition holds for block subspaces, then no two block subspaces form a topological direct sum. The previous proposition then implies that X is HI. Therefore, it only remains to prove the claim made at the beginning of the proof. Suppose first that Y and Z are closed infinite dimensional subspaces for which Y + Z fails to be a topological direct sum. We assume for contradiction that there exists ε > 0 such that (cid:107)y − z(cid:107) ≥ ε(cid:107)y + z(cid:107) whenever y ∈ Y, z ∈ Z. It follows that (cid:107)y + z(cid:107) ≥ ε(cid:107)y − z(cid:107) for all y ∈ Y, z ∈ Z. Consequently, (cid:107)y(cid:107) = 1 2 ((cid:107)y + z(cid:107) + (cid:107)y − z(cid:107)) ≤ ε )(cid:107)y + z(cid:107). 2 ((cid:107)y + z(cid:107) + 1 ε )(cid:107)y + z(cid:107). A similar calculation yields (cid:107)z(cid:107) ≤ 1 2(cid:107)y + z + y − z(cid:107) ≤ 1 ε(cid:107)y + z(cid:107)) = 1 2 (1 + 1 2 (1 + 1 1 So (cid:107)y + z(cid:107) ≥ 2(1 + 1 ε )−1 max((cid:107)y(cid:107),(cid:107)z(cid:107)) ≥ (1 + )−1((cid:107)y(cid:107) + (cid:107)z(cid:107)) 1 ε from which it follows that Y ⊕ Z is a direct sum, giving us the required contradiction. Conversely, assume that Y and Z satisfy the condition that ∀ε > 0,∃y ∈ Y, z ∈ Z with (cid:107)y − z(cid:107) < ε(cid:107)y + z(cid:107). We must see that Y + Z fails to be a topological direct sum. We again argue by contradiction, assuming that Y + Z is topological. Consequently, there exists δ > 0 such that (cid:107)y + z(cid:107) ≥ δ((cid:107)y(cid:107) + (cid:107)z(cid:107)) whenever y ∈ Y, z ∈ Z. we note that this implies that (cid:107)y + z(cid:107) ≥ δ(cid:107)z(cid:107) for all y ∈ Y, z ∈ Z. (cid:107)y − z(cid:107) = (cid:107)y + z − 2z(cid:107) ≤ (cid:107)y + z(cid:107) + 2(cid:107)z(cid:107) ≤ (1 + 2 the other hand, taking ε = 1 by the hypothesis (with z replaced by −z). (cid:107)y − z(cid:107) ≤ (1 + 2 2(cid:107)y − z(cid:107). This only fails to be a contradiction δ )(cid:107)y − z(cid:107) = 1 if (cid:107)y − z(cid:107) = 0, which, since it is assumed that Y and Z are topological, can only happen if y = z = 0. But, recalling that y, z were chosen so as to satisfy (cid:107)y + z(cid:107) < ε(cid:107)y − z(cid:107), we clearly cannot have y = z = 0 and we once again have the contradiction we seek. δ )(cid:107)y + z(cid:107) for all y ∈ Y, z ∈ Z. On δ )−1, there exist y ∈ Y, z ∈ Z with (cid:107)y + z(cid:107) < ε(cid:107)y − z(cid:107) It follows that for this choice of y and z, δ )(cid:107)y + z(cid:107) ≤ ε(1 + 2 It follows that In particular, 2 (1 + 2 1.4.4 Elementary Results from Operator Theory We will make repeated use of the following duality results when constructing operators in later chapters of this thesis. Lemma 1.4.11. Let X be a Banach space, Y a normed space, T : X → Y a bounded linear operator. Then T is a quotient operator if, and only if, T ∗ : Y ∗ → X∗ is an isomorphic embedding. Proof. Suppose first that T is a quotient operator. Then there is some M > 0 such that BY (0; 1) ⊆ M T(cid:0)BX (0; 1)(cid:1). It follows that M(cid:107)T ∗y∗(cid:107) = sup x∈BX (0;1) y∗(M T x) ≥ sup y∈BY (0;1) y∗y = (cid:107)y∗(cid:107) 12 i.e. that T ∗ is an isomorphic embedding. Suppose conversely that T ∗ is an isomorphic embedding and let M > 0 be such that M(cid:107)T ∗y∗(cid:107) ≥ (cid:107)y∗(cid:107). We claim that M T(cid:0)BX (0; 1)(cid:1) ⊇ BY (0; 1). Since X is a Banach space, this is sufficient to deduce that T : X → Y is a quotient operator. Let's suppose by contradiction that M T(cid:0)BX (0; 1)(cid:1) = (cid:92) M T(cid:0)BX (0; 1)(cid:1) + εB◦ ε>0 Y and ε > 0 such that y0 /∈ M T(cid:0)BX (0; 1) + εB◦ Y (0; 1) So there is some y0 ∈ B◦ Separation Theorem 1.4.8, we can find a non-zero y∗ ∈ Y ∗ with Y (0; 1) (cid:43) B◦ Y . Applying M(cid:107)T ∗y∗(cid:107) + ε(cid:107)y∗(cid:107) = M sup x∈BX (0;1) Re y∗T x + ε sup y∈B◦ Y Re y∗y ≤ Re y∗(y0) ≤ (cid:107)y∗(cid:107) But this implies that (1 + ε)(cid:107)y∗(cid:107) ≤ (cid:107)y∗(cid:107) which is clearly a contradiction. Lemma 1.4.12. Let X and Y be Banach spaces and suppose T : X∗ → Y ∗ is a bounded linear operator which is also weak∗ to weak∗ continuous. Then there is a bounded linear operator S : Y → X such that S∗ = T . Furthermore, if for each n ∈ N, Tn : X∗ → X∗ is a bounded linear operators which is weak* to weak* continuous and T : X∗ → X∗ is a bounded linear operator with (cid:107)Tn−T(cid:107) → 0 (i.e. Tn → T in the operator norm topology), then T is also weak* to weak* continuous. Proof. Let JX : X → X∗∗, JY : Y → Y ∗∗ denote the canonical embeddings. We first see that T ∗JY maps Y into JX (X). It is enough to see that for every y ∈ Y , the map T ∗JY y : X∗ → K is weak∗ continuous. To this end, we suppose (x∗ α) is a net converging in the weak∗ x∗). By weak∗-weak∗ continuity of T , we have topology to some x∗ (we write x∗ T x∗ T x∗ so that α (cid:42)w∗ α (cid:42)w∗ (T ∗JY y)x∗ α = JY y(T x∗ α) = T x∗ α(y) → T x∗(y) = (T ∗JY y)x∗ as required. We can thus define a bounded linear map S : Y → X by S := J−1 note that for all x∗ ∈ X∗, y ∈ Y X T ∗JY . We (S∗x∗)y = x∗(Sy) = x∗(J−1 X T ∗JY y) = T ∗JY y(x∗) = (T x∗)y so that S∗ = T as required. To prove the second part of the lemma, note that by the first part of the proof, each Tn is the dual operator of an operator Sn : X → X. Precisely, Sn = J−1 n JX . Since Tn → T with respect to the operator norm, T ∗ n=1 is a Cauchy sequence in L(X), so there is some bounded linear operator S such that Sn → S. It n → T ∗ in operator norm. It follows that (Sn)∞ X T ∗ 13 n → S∗ in L(X∗). However, S∗ n = Tn for all n, and since Tn → T , uniqueness follows that S∗ of limits gives that T = S∗. Since T is the dual of some operator, it is certainly weak* continuous. Lemma 1.4.13. Let X and Y be Banach spaces, T : X → Y a bounded linear operator. If T ∗ is one to one and has closed range then im T = Y . Proof. Since T ∗ is one to one and has closed range in the Banach space Y , T ∗ is an isomor- phism onto its image and thus T ∗ : Y ∗ → X∗ is an isomorphic embedding. It follows from Lemma 1.4.11 that T is a quotient operator. In particular T is onto. Theorem 1.4.14 (Closed Range Theorem). Let X and Y be Banach spaces, T : X → Y a bounded linear operator. Then T has closed range if and only if T ∗ has closed range. Proof. The proof is essentially taken from [10]. 1 : T [X]∗ → X∗ is an isomorphic embedding and consequently that im T ∗ Let us suppose first that T has closed range. Since Y is a Banach space and im T is closed, T1 : X → T [X] defined by T1x = T x is a quotient operator. It follows by Lemma 1.4.11 that T ∗ 1 is closed. We claim that im T ∗ 1 = im T ∗ and thus the image of T ∗ is closed as required. Indeed suppose z∗ ∈ T [X]∗ and let y∗ ∈ Y ∗ be an extension of z∗ to Y (the existence of which is of course guaranteed by the Hahn Banach Theorem). Then for x ∈ X, 1 z∗)x = z∗T1x = y∗T x = (T ∗y∗)x (T ∗ 1 ⊆ im T ∗. Conversely, if y∗ ∈ Y ∗, we let z∗ ∈ T [X]∗ be the 1 . It 1 z∗ and so im T ∗ ⊆ im T ∗ 1 z∗ = T ∗y∗ and im T ∗ So T ∗ restriction of y∗ to T [X]. It is easy to see that T ∗y∗ = T ∗ follows that im T ∗ 1 = im T ∗ as required. 1 is one-to-one. If x∗ ∈ X∗ is in the closure of T ∗ Conversely, suppose T ∗ has closed range. We let Z ⊆ Y be the Banach space T [X] and consider the map T1 : X → Z defined by T1x = T x. Since T1 has dense range, it follows that T ∗ n where n ∈ Y ∗ of the z∗ n are in Z∗. We apply the Hahn Banach Theorem obtaining extensions y∗ n and since T ∗ has closed range, x∗ = T ∗y∗ for some z∗ n. It follows that x∗ = limn→∞ T ∗y∗ y∗ ∈ Y ∗. If z∗ is the restriction of y∗ to Z, then x∗ = T ∗ 1 is one-to-one and has closed range. It follows by Lemma 1.4.13 that im T1 = im T = Z and so T has closed range as required. 1 (Z∗) then x∗ = limn→∞ T ∗ 1 z∗ 1 z∗. Hence T ∗ As well as the above duality results, we will make use of the following basic sequence technique which provides a sufficient condition for an operator to be compact. 14 Proposition 1.4.15. Let X be a Banach space with a Schauder basis and T : X → X a bounded linear operator. If T xk → 0 for all bounded block basic sequences (xk)∞ k=1, then T is compact. If we demand that the basis of X is shrinking, then the converse is also true. Proof. Denote by (en)∞ n=1 the Schauder basis of X. We noted in Section 1.4.1 that the j=1 ajej are bounded. We will show that (cid:107)T − T Pn(cid:107) → 0 (as n → ∞); this shows that T is a uniform limit of finite rank operators and consequently that T is compact. projections Pn : X → [ej : j ≤ n] defined by(cid:80)∞ Suppose for contradiction that (cid:107)T − T Pn(cid:107) (cid:88)→ 0. It follows that we can find δ > 0, a j=1 ajej (cid:55)→(cid:80)n j=1 of natural numbers, and a sequence of norm 1 vectors strictly increasing sequence (Nj)∞ (xj)∞ j=1 such that (cid:107)(T − T PNj )xj(cid:107) > δ n=1 is a Schauder basis, we can find M1 > N1 such that (cid:107)x1 − PM1x1(cid:107) < δ/2(cid:107)T(cid:107). Since (en)∞ We set y1 = (PM1 − PN1)x1 noting that (cid:107)y1(cid:107) ≤ 2K (where K is the basis constant of the basis (en)∞ n=1). Moreover, (cid:107)T y1(cid:107) = (cid:107)T x1 − T PN1x1 − (T x1 − T PM1x1)(cid:107) ≥ (cid:107)(T − T PN1)x1(cid:107) − (cid:107)T ◦ (I − PM1)x1(cid:107) > δ 2 Now, we can find Nj2 > M1 and M2 > Nj2 such that (cid:107)xj2 − PM2xj2(cid:107) < δ/2(cid:107)T(cid:107). We set )xj2. Estimating as before yields (cid:107)y2(cid:107) ≤ 2K and (cid:107)T y2(cid:107) > δ/2. Continuing n=1 with (cid:107)T yn(cid:107) > δ/2 for all n. n=1 is a bounded block basic y2 = (PM2 −PNj2 in this way, we obtain a bounded block basic sequence (yn)∞ This contradicts the hypothesis that T xn → 0 whenever (xn)∞ sequence, completing the first part of the proof. The converse of the theorem does not hold in general. However, in our statement of a partial converse, we demand that the basis is shrinking. Let (xn)∞ n=1 be a bounded block basic sequence. We are required to show that if T is compact, T xn → 0. Arguing by contradiction we assume that, after choosing a subsequence and relabelling if necessary, there is δ > 0 such that (cid:107)T xn(cid:107) ≥ δ for all n. We note that (xn)∞ Proposition 1.4.5. Consequently (T xn)∞ n=1 is weakly null by n=1 converges weakly to 0. Since T is compact, some subsequence (T xnj )∞ j=1 converges in norm to x ∈ X. It follows that T xnj → x in the weak topology. By the above argument and uniqueness of weak limits j=1 converges in norm to x, it follows that (cid:107)T xnj(cid:107) → 0, we have that x = 0. Since (T xnj )∞ contradicting the fact that (cid:107)T xnj(cid:107) ≥ δ for all j. This completes the proof. 15 1.4.5 Complexification We will require results from spectral theory. Of course, many of these results assume that the Banach algebras in question are over the field of complex scalars. In order to pass to the real case, we recall some basic complexification arguments. If X is a real Banach space, the complexified space is defined as XC := X ⊕ X with (complex) scalar multiplication and vector addition defined by (x, y) + (u, v) = (x + u, y + v) (α + iβ)(x, y) = (αx − βy, βx + αy) ∀α, β ∈ R, x, y ∈ X. ∀x, y, u, v ∈ X Obviously we can identify X as a (real) subspace of the complexification under the linear injection given by j : X → XC, x (cid:55)→ (x, 0). Noting that i(y, 0) = (0, y), we can write the vector (x, y) ∈ XC as j(x) + ij(y) and it is obvious that this representation is unique. We will often find it convenient to suppress the use of the embedding j and simply write z ∈ XC by z = x + iy, where x, y ∈ X ⊆ XC. Consequently, we write XC = X ⊕ iX. There are many ways to define a complex norm on XC. For the purposes of this thesis, we shall work with the norm defined by (cid:107)x + iy(cid:107)XC := supt∈[0,2π] (cid:107)x cos t − y sin t(cid:107)X . is a trivial exercise to check that this defines a norm and we only give the proof of the homogeneity. Note that if cos θ + i sin θ ∈ C, (cid:107)(cos θ + i sin θ)(x + iy)(cid:107)XC = sup t∈[0,2π] = sup t∈[0,2π] = sup t∈[0,2π] = (cid:107)x + iy(cid:107)XC (cid:107)(x cos θ − y sin θ) cos t − (x sin θ + y cos θ) sin t(cid:107)X (cid:107)(cos θ cos t − sin θ sin t)x − (sin θ cos t + cos θ sin t)y(cid:107)X (cid:107) cos(θ + t)x − sin(θ + t)y(cid:107)X It so that λ(cid:107)z(cid:107)XC = (cid:107)λz(cid:107)XC for all λ ∈ C, z ∈ XC. We remark that whenever x, y ∈ X, we have (cid:107)j(x)(cid:107)XC = (cid:107)x(cid:107)X and (cid:107)x + iy(cid:107)XC = (cid:107)x − iy(cid:107)XC. Using these observations we obtain the following lemma. Lemma 1.4.16. Let x, y ∈ X. Then (cid:107)x(cid:107)X = (cid:107)x(cid:107)XC ≤ (cid:107)x + iy(cid:107)XC and (cid:107)y(cid:107)X = (cid:107)y(cid:107)XC ≤ (cid:107)x + iy(cid:107)XC Proof. Just note that 2(cid:107)x(cid:107)XC ≤ (cid:107)x + iy(cid:107)XC + (cid:107)x − iy(cid:107)XC = 2(cid:107)x + iy(cid:107)XC. The second part of the lemma is similar. If T is a (real) operator on X, then we can uniquely extend it to a (complex) operator TC on XC by TC(x + iy) = T x + iT y. We have the following lemma: 16 Lemma 1.4.17. If T ∈ L(X) then TC ∈ L(XC) and (cid:107)T(cid:107)X→X = (cid:107)TC(cid:107)XC→XC. Moreover, if T is compact, then so is TC. Proof. It is easy to see that TC is a linear operator. Moreover, it is clear that (cid:107)TC(cid:107) ≥ (cid:107)T(cid:107) since TC extends T (and (cid:107)j(x)(cid:107)XC = (cid:107)x(cid:107)X ). Note, (cid:107)TC(x + iy)(cid:107)XC = sup t∈[0,2π] (cid:107)(T x) cos t − (T y) sin t(cid:107)X = sup t∈[0,2π] (cid:107)T (x cos t − y sin t)(cid:107)X ≤ (cid:107)T(cid:107) sup t∈[0,2π] (cid:107)x cos t − y sin t(cid:107)X = (cid:107)T(cid:107)(cid:107)x + iy(cid:107)XC so that (cid:107)TC(cid:107) ≤ (cid:107)T(cid:107). To see that TC is compact when T is, note that if (xn + iyn)∞ n=1 is a bounded sequence in XC then the sequences (xn)∞ n=1 are bounded sequences in X by Lemma 1.4.16. By compactness of T , we can choose subsequences (xnk )∞ k=1 such that both (T xnk ) and (T ynk ) converge. It follows easily that TC(xnk + iynk ) converges in XC, so TC is compact as required. k=1 and (ynk )∞ n=1 and (yn)∞ 1.4.6 Strictly Singular Operators The majority of this thesis is concerned with the interplay between the compact and strictly singular operators. We recall in this section the definition and some some of the elementary results on strictly singular operators which will be used throughout the thesis. We recall the following definition Definition 1.4.18. Let X, Y be Banach spaces and T : X → Y a bounded linear operator. T is strictly singular if whenever Z ⊆ X is a subspace such that there exists some δ > 0 with (cid:107)T z(cid:107) ≥ δ(cid:107)z(cid:107) for all z ∈ Z, then Z is finite dimensional. In other words, a strictly singular operator is one which is not an isomorphic embedding on any infinite dimensional (closed) subspace. The following characterisation of strictly singular operators on a Banach space with a Schauder basis will be particularly useful to us later on in this thesis. n=1 and suppose T : X → X Proposition 1.4.19. Let X be a Banach space with a basis (en)∞ is a bounded linear operator on X. Then T is strictly singular ⇐⇒ whenever [yn : n ∈ N] is a block subspace of X, the restriction, T[yn] : [yn] → X is not an isomorphic embedding. Proof. Clearly when T is strictly singular, we have the condition about block subspaces. To complete the proof, we show that if T is an isomorphism on some closed infinite dimensional subspace Y of X, then T must be an isomorphism on some block subspace of X. 17 Indeed, suppose for some infinite dimensional, closed subspace Y of X, there is δ > 0 such that for every y ∈ Y,(cid:107)T y(cid:107) ≥ δ(cid:107)y(cid:107). Choose any 0 < θ < 1 By Proposition 1.4.2, there exists a sequence (yk)∞ block basic sequence (xk) (with respect to the basis (en)) satisfying 2 such that (cid:107)T(cid:107)θ( 1 2 −θ)−1 ≤ δ. k=1 ⊆ Y with (cid:107)yk(cid:107) = 1 for every k and a ∞(cid:88) k=1 2K (cid:107)yk − xk(cid:107) (cid:107)xk(cid:107) < θ where K is the basis constant of the basis (en)∞ n=1. We recall that the sequences (xk) and (yk) are equivalent basic sequences. Moreover, that T is an isomorphism on the block subspace [xk]. To see this, let(cid:80) the sequence (xk) as a block sequence of the (en) has basis constant at most K. We claim k akxk be a norm 1 vector in [xk], and observe that this implies in particular that ak ≤ 2K(cid:107)xk(cid:107) since (xk) is a basic sequence with basis constant at most K. Then akyk + akyk (cid:88) ak(xk − yk)(cid:1)(cid:13)(cid:13) (cid:1)(cid:13)(cid:13) −(cid:13)(cid:13)T(cid:0)(cid:88) (cid:88) k k ak(xk − yk)(cid:1)(cid:13)(cid:13) (cid:107)xk − yk(cid:107) (cid:107)xk(cid:107) k akyk(cid:107) − (cid:107)T(cid:107)2K akyk(cid:107) − (cid:107)T(cid:107)θ. ak(yk − xk)(cid:107) ≥ 1 − 2K (cid:88) k (cid:107)yk − xk(cid:107) (cid:107)xk(cid:107) ≥ 1 − θ. (cid:107)T(cid:0)(cid:88) k akxk k k (cid:1)(cid:107) =(cid:13)(cid:13)T(cid:0)(cid:88) ≥(cid:13)(cid:13)T(cid:0)(cid:88) ≥ δ(cid:107)(cid:88) ≥ δ(cid:107)(cid:88) akxk(cid:107) − (cid:107)(cid:88) (cid:1)(cid:107) ≥ δ(cid:107)(cid:88) k k k k akyk(cid:107) as follows: (cid:107)(cid:88) We can estimate (cid:107)(cid:80) akyk(cid:107) ≥ (cid:107)(cid:88) (cid:107)T(cid:0)(cid:88) So, k k akxk k k akyk(cid:107) − (cid:107)T(cid:107)θ ≥ δ(1 − θ) − (cid:107)T(cid:107)θ ≥ δ 2 with the final inequality resulting in our choice of θ. We have thus seen that any norm 1 vector x ∈ [xk] has (cid:107)T x(cid:107) ≥ δ 2 > 0. Thus T is an isomorphism on [xk] as claimed. There are two further observations about strictly singular operators that are important in relation to the work of this thesis. The first is that the strictly singular operators on a Banach space X form a closed ideal of the operator algebra L(X). The second is that the the essential spectrum of a strictly singular operator on a complex Banach space is just {0}. (We will actually give a more formal statement of this fact that also holds for real Banach spaces.) In the remainder of this section, we provide proofs of these facts. We find it convenient to introduce another class of operators. 18 Definition 1.4.20. Let T : X → X be a bounded linear operator on a Banach space X. We say T is upper semi-Fredholm if the kernel of T is finite dimensional and T has closed range. Remark 1.4.21. 1. We recall that if T : X → X has finite dimensional kernel and finite co-dimensional image, then T is said to be a Fredholm operator. The range of T is automatically closed in this case, thus all Fredholm operators are, in particular, upper semi-Fredholm. Moreover, for Fredholm operators, we can define the index of T by ind T := dim Ker T − codim imT. It is well known (see, e.g. [26]) that ind : F → Z is a continuous map (here F ⊆ L(X) denotes the subset of Fredholm operators). For upper semi-Fredholm operators T , we can define the generalised index by the same formula as above; this will give a finite integer if T is in fact Fredholm. Otherwise it must be the case that the image of T has infinite co-dimension and the generalised index is set equal to −∞. 2. Suppose T : X → X is upper semi-Fredholm. Since finite dimensional subspaces are always complemented, we can write X as the topological direct sum X = Ker T ⊕ X(cid:48) for some closed, finite co-dimensional subspace X(cid:48) ⊂ X. Since also Im T = T (X(cid:48)) is closed, it is easy to see that T : X(cid:48) → X is an isomorphic embedding. In other words, we have seen that an upper semi-Fredholm operator is an isomorphism on some finite co-dimensional, closed subspace. Conversely, if T is an isomorphism on some closed, finite co-dimensional subspace of X, then T is upper semi-Fredholm. Indeed, Let X1 be a finite co-dimensional subspace on which T is an isomorphism. It follows that Ker T ∩ X1 = {0} and so we can write X as the (algebraic) direct sum X = X1 ⊕ Ker T ⊕ F for some (necessarily finite dimensional) subspace F . Since X1 is finite co-dimensional, dim Ker T ≤ dim (Ker T ⊕ F ) = codim X1 < ∞. The image of T is closed since T (X) = T (X1)⊕ T (F ) and the sum is easily seen to be topological (T (X1) is closed since T is an isomorphism on X1 and T (F ) is finite dimensional). We have therefore proved that an operator is upper semi-Fredholm if and only if it is an isomorphism on some finite co-dimensional subspace. In some literature (e.g. [16], [22]) one defines an operator to be finitely singular if it is an isomorphism on some finite co-dimensional subspace. We have therefore proved that the class of finitely singular operators is exactly the same as class of upper semi-Fredholm operators. 19 3. We also note that every strictly singular operator on an infinite dimensional Banach space fails to be upper semi-Fredholm. Later we will require the following proposition which extends the well known result that the index of a Fredholm operator defines a continuous map into Z, to the case where the operators are upper semi-Fredholm. Proposition 1.4.22. If T : X → Y is an upper semi-Fredholm operator, there exists a neighbourhood U of 0 ∈ L(X, Y ) such that T + U is upper semi-Fredholm for all U ∈ U and moreover, the (generalised) index of T + U is equal to the (generalised) index of T . Proof. When T is a Fredholm operator, the result follows from the well known result that the index map, defined on Fredholm operators, is continuous. In the case where T is upper semi- Fredholm but not Fredholm, the result follows easily from the following proposition. (We remark that the following proposition can be found in [24, Proposition 3.1].) Proposition 1.4.23. Let T : X → Y be an isomorphic embedding. Then there exists a neighbourhood of 0 ∈ L(X, Y ) such that for every S is this neighbourhood, T + S is an isomorphic embedding and codim(T + S)X = codim T X (finite or +∞). To prove the proposition, we need: Lemma 1.4.24. Let T ∈ L(X, Y ) be an isomorphic embedding and k ≥ 0 an integer. 1. If codim T X ≥ k, there exists c > 0 such that codim(T + S)X ≥ k whenever (cid:107)S(cid:107) < c. 2. If codim T X = k, there exists c > 0 such that codim(T + S)X = k whenever (cid:107)S(cid:107) < c. Proof. Since T is an isomorphic embedding, there exists a δ > 0 such that (cid:107)T x(cid:107) ≥ δ(cid:107)x(cid:107) for all x ∈ X. Now, if codim T X ≥ k, there exists a subspace F ⊂ Y such that dim F = k and T X ∩ F = {0}. Let πF : Y → Y /F denote the usual quotient map. We claim that πF ◦ T is an isomorphic embedding from X into Y /F . Indeed, if not there exists a sequence (xn)∞ of vectors in X with (cid:107)xn(cid:107) = 1 for all n and (cid:107)T xn + F(cid:107) → 0. It follows that there exists a n=1 in F such that (cid:107)T xn − fn(cid:107) → 0. It is clear that the sequence sequence of vectors (fn)∞ (fn)∞ n=1 is bounded and since F is finite dimensional, an easy compactness argument yields that (without loss of generality, after passing to a subsequence and relabelling if necessary) fn → f ∈ F . It follows that T xn → f ∈ F . Since T is an isomorphic embedding, T (X) is closed, so limn T xn = f ∈ T X ∩ F = {0}. On the other hand, (cid:107)f(cid:107) = limn (cid:107)T xn(cid:107) ≥ δ. So πF ◦ T is an isomorphic embedding as claimed. n=1 20 It is easy to see that the property of being an isomorphic embedding is closed under small norm perturbation. Since we know πF ◦ T is an isomorphic embedding, it follows that if we take c > 0 sufficiently small, πF ◦ (T + S) is also an isomorphic embedding if (cid:107)S(cid:107) < c. This implies that F ∩ (T + S)X = {0}, hence codim(T + S)X ≥ k, proving the first claim. The proof of the second case is similar; in this case we can select the subspace F such that Y = T X ⊕ F and dim F = k. It follows (using the same argument as before) that πF ◦ T is an onto isomorphism. Consequently, there exists c > 0 such that if (cid:107)S(cid:107) < c then πF ◦ (T + S) is an onto isomorphism. In particular, F ∩ (T + S)X = {0}. Moreover, since πF ◦ (T + S) is an onto isomorphism, for every y ∈ Y there exists x ∈ X such that πF (y) = πF ((T + S)x). Consequently, y − (T + S)x ∈ F , showing that Y = F + (T + S)X. So Y = F ⊕ (T + S)X and codim(T + S)X = k as required. Proof of Proposition 1.4.23. Let c > 0 be such that T + S is an isomorphic embedding whenever S ∈ Bc := {U ∈ L(X, Y ) : (cid:107)U(cid:107) < c} For each positive integer k ≥ 0, the set Dk = {U ∈ Bc : codim(T + U )X = k} is open in Bc by the previous lemma. Moreover, Dk = {U ∈ Bc : codim(T + U )X ≥ k + 1}C(cid:92)∩k−1 j=0{U ∈ Bc : codim(T + U )X = j}C, thus Dk is closed in Bc as an intersection of closed sets (each set in the intersection is a complement of an open set by the previous lemma). Since Bc is connected, each Dk is either empty or equal to Bc. The result now follows in the case where codim T X = k for some positive integer k ≥ 0. In the case where codim T X = +∞, we note that by the preceding argument, Dk has to be empty for every k ≥ 0. Consequently, codim(T + U )X = +∞ for every U ∈ Bc and the result is again proved. Lemma 1.4.25. Let X be a Banach space (real or complex) and suppose T : X → X is a bounded linear operator which fails to be upper semi-Fredholm. Then, for all ε > 0, there exists an infinite dimensional (closed) subspace Yε ⊆ X with (cid:107)TYε(cid:107) < ε. The restriction of T to Yε is a compact operator. Proof. The proof given here follows closely that given in [22]. By Remark 1.4.21, we note that T is not an isomorphism on any finite co-dimensional subspace of X. We begin the proof by fixing a sequence of real numbers an > 0 such that(cid:81)∞ 2. We will construct inductively a normalised, basic sequence (xn)∞ properties n=1(1+an) ≤ n=1 with the following 21 1. If m < n are natural numbers, (λi)n i=1 are scalars, then (cid:107) m(cid:88) λixi(cid:107) ≤ (cid:32) n(cid:89) (cid:33) (cid:107) n(cid:88) (1 + ai) λixi(cid:107) i=1 i=m+1 i=1 2. (cid:107)T xn(cid:107) ≤ ε/2n+2 for all n. at most(cid:81)∞ i=1(1 + ai) ≤ 2. Note that as a consequence of the first property, (xn) is a basic sequence with basis constant To begin the induction, since in particular T is not an isomorphism, we can choose x1 ∈ X with (cid:107)x1(cid:107) = 1 and (cid:107)T x1(cid:107) ≤ ε/8. Inductively, suppose we have constructed x1, . . . xn ∈ X with (cid:107)xi(cid:107) = 1 and satisfying the two properties above. It is enough to show that we can find xn+1 ∈ X with (cid:107)xn+1(cid:107) = 1, (cid:107)T xn+1(cid:107) ≤ ε/2n+3 and having the property that whenever y ∈ E := span{x1, . . . xn}, λ is a scalar, (1 + an+1)(cid:107)y + λxn+1(cid:107) ≥ (cid:107)y(cid:107). To this end, let y1, y2, . . . ym be a (finite) δ-net for the unit sphere SE of the finite dimensional subspace E, where 0 < δ < 1 is such that 1 + an+1 ≥ 1/(1 − δ). By the Hahn- Banach Theorem, we can choose norm 1 linear functionals y∗ i (yi) = 1 for all i = 1, . . . m. By hypothesis of the lemma, T is not an isomorphism when restricted to the finite co -- dimensional subspace Z := ∩m i , so we may choose a norm 1 vector, xn+1 ∈ Z with (cid:107)T xn+1(cid:107) ≤ ε/2n+3. If y ∈ SE, λ a scalar, there exists some i ∈ {1, . . . m} with (cid:107)y − yi(cid:107) ≤ δ, and m ∈ SX∗ with y∗ i=1 Ker y∗ 1, . . . y∗ (1 + an+1)(cid:107)y + λxn+1(cid:107) = (1 + an+1)(cid:107)yi + λxn+1 + y − yi(cid:107) i (yi + λxn+1 + y − yi) ≥ (1 + an+1) y∗ ≥ (1 + an+1) (1 − y∗ ≥ (1 + an+1) (1 − (cid:107)y − yi(cid:107)) ≥ (1 + an+1)(1 − δ) ≥ 1 i (y − yi)) We set Yε := [xn]∞ from which we easily conclude that (cid:107)y + λxn+1(cid:107) ≥ (cid:107)y(cid:107) for any y ∈ E and scalar λ as required. n=1, an infinite dimensional subspace of X with Schauder basis (xn)∞ If y = (cid:80)∞ (cid:107)(cid:80)∞ n=1 αnT xn(cid:107) ≤ 4ε(cid:80)∞ TN : Yε → X, defined by(cid:80)∞ n=1. n=1 αnxn ∈ Yε with (cid:107)y(cid:107) = 1, it follows that αn ≤ 4 for all n and (cid:107)T y(cid:107) = n=1 2−(n+2) = ε, giving the desired estimate for (cid:107)TYε(cid:107). It is easy to see (by a similar estimation) that T is the uniform limit of the finite rank operators n=1 αnT xn, and thus T : Yε → X is compact as n=1 αnxn (cid:55)→(cid:80)N claimed. 22 Corollary 1.4.26. An operator T : X → Y is strictly singular ⇐⇒ whenever Z ⊂ X is an infinite dimensional subspace of X, there exists a further infinite dimensional subspace Z(cid:48) ⊂ Z such that the restriction of T to Z(cid:48) is compact. Proof. Suppose first T is strictly singular and let Z be an infinite dimensional subspace of X. We consider the operator T1 : Z → X where T1 is simply the restriction of T to Z. It is easily verified that T1 is strictly singular because T is, and therefore T1 fails to be upper semi-Fredholm. By the previous lemma, we can find an infinite dimensional subspace Z(cid:48) ⊂ Z such that the restriction of T1 to Z(cid:48) is compact. Obviously by definition of T1, this implies that the restriction of T to Z(cid:48) is compact. Conversely, suppose for contradiction that Z is an infinite dimensional subspace of X on which T is an isomorphism. By the assumed property, there exists an infinite dimensional subspace Z(cid:48) of Z on which T is both a compact operator and an isomorphism. However, these two facts imply that Z(cid:48) has a compact unit ball, contradicting the fact that Z(cid:48) is infinite It follows that T fails to be an isomorphism on every infinite dimensional dimensional. subspace, i.e. T is strictly singular. Corollary 1.4.27. The set of strictly singular operators on a Banach space X forms a closed ideal of L(X). Proof. Suppose first that T, S : X → X are strictly singular operators on X. We will use Corollary 1.4.26 to show T + S is strictly singular. To this end, let Z be an infinite dimensional subspace of X. Since T is strictly singular, there exists an infinite dimensional subspace Z(cid:48) ⊂ Z on which T is compact. We now apply Corollary 1.4.26 to S and the subspace Z(cid:48), obtaining an infinite dimensional subspace Z(cid:48)(cid:48) ⊂ Z(cid:48) on which both S and T are compact. It follows that T + S is compact when restricted to Z(cid:48)(cid:48), and consequently T + S is strictly singular by the corollary. To prove that the strictly singular operators are an ideal, it remains to prove that if P is strictly singular and R, Q ∈ L(X) then RP Q is a strictly singular operator. Suppose Z is an infinite dimensional subspace of X and fix ε > 0. We will exhibit a norm 1 vector, z ∈ Z with (cid:107)RP Qz(cid:107) < ε. Note that if Q(Z) is finite dimensional, then since Z is infinite dimensional, we must have ker Q ∩ Z (cid:54)= {0}. In this case, we just take z to be any norm 1 vector in ker Q∩ Z and the proof is trivial. So we assume that Q(Z) is infinite dimensional. Since P is assumed strictly singular, P is not an isomorphism on the infinite dimensional subspace Q(Z). Consequently, we can find a norm 1 vector x = Q(z0) (for some z0 ∈ Z) with (cid:107)P x(cid:107) = (cid:107)P Qz0(cid:107) < ε/(cid:107)Q(cid:107)(cid:107)R(cid:107). We take z = z0/(cid:107)z0(cid:107) noting that 1/(cid:107)z0(cid:107) ≤ (cid:107)Q(cid:107) since 1 = (cid:107)x(cid:107) = (cid:107)Qz0(cid:107), so that (cid:107)P Qz(cid:107) < (cid:107)Q(cid:107)ε/(cid:107)Q(cid:107)(cid:107)R(cid:107). Therefore, (cid:107)RP Qz(cid:107) ≤ (cid:107)R(cid:107)(cid:107)P Qz(cid:107) < ε as required. 23 Finally, note if Tn is a sequence of strictly singular operators on X with (cid:107)Tn − T(cid:107) → 0 then T is strictly singular, so that the ideal of strictly singular operators is closed. Indeed, if Z is an infinite dimensional subspace of X, ε > 0, then we find N such that (cid:107)TN −T(cid:107) < ε/2. Since TN is strictly singular, there exists z ∈ Z with (cid:107)z(cid:107) = 1 and (cid:107)TN z(cid:107) < ε/2. It follows that (cid:107)T z(cid:107) ≤ (cid:107)TN z(cid:107) + (cid:107)(T − TN )z(cid:107) < ε. We conclude this section by stating and proving a result about strictly singular operators that we will need in Chapter 4. Theorem 1.4.28. Let X be a Banach space (real or complex). If T ∈ L(X) is strictly singular then r(T ) := limn→∞ (cid:107)T n + K(X)(cid:107) 1 n = 0. In the case where X is a complex Banach space, the above result just says that the essential spectrum of a strictly singular operator on a complex Banach space is {0}. To prove the result in the real case, we will make use of a complexification argument and will consequently need the following technical lemma; it is essentially a generalisation of Lemma 1.4.25, with the only real difficulty being to make sure we extract a real subspace of the complexified space. Lemma 1.4.29. Let T : X → X be a bounded linear operator on a real Banach space X, and denote by XC = X ⊕ iX the complexification of X, TC the complex extension of T to XC (as in Section 1.4.5). Suppose there exists λ ∈ C such that TC − λIdXC fails to be upper semi-Fredholm. Then the operator Tλ on X defined by Tλ = T 2 − 2ReλT + λ2IdX has the property that for all ε > 0, there exists a real infinite dimensional Yε ⊆ X with (cid:107)TλYε(cid:107) < ε. Proof. The proof is very similar to that of Lemma 1.4.25 and we continue the notation n=1 ⊆ X with from that proof. We will construct inductively a normalised sequence (xn)∞ the following properties: 1. If m < n are natural numbers, (λi)n i=1 are real scalars, then (cid:107) m(cid:88) λixi(cid:107) ≤ (cid:32) n(cid:89) (cid:33) (cid:107) n(cid:88) (1 + ai) λixi(cid:107) i=1 i=m+1 i=1 2. (cid:107)Tλxn(cid:107) ≤ ε/2n+2 for all n. a real basic sequence with basis constant at most (cid:81)∞ As in the proof of Lemma 1.4.25 , it is a consequence of the first property that (xn) is i=1(1 + ai) ≤ 2. Given such a basic n=1 ⊆ X to complete the proof (which is then identical to the sequence, we take Yε := [xn]∞ proof of the previous lemma). 24 Observe that whenever x, y ∈ X, (cid:0)TC − λIdXC (cid:1) (TC − λIdXC) (x + iy) = Tλx + iTλy = (Tλ)C(x + iy). This, combined with the fact that TC − λIdC is not an isomorphism on any finite co-dimensional subspace of XC (see Remark 1.4.21), implies that (Tλ)C is not an isomorphism on any finite co-dimensional subspace of XC. In particular, we can choose z1 ∈ XC with (cid:107)z1(cid:107) = 1 and (cid:107)(Tλ)C z1(cid:107) ≤ ε/16. Since (cid:107)z1(cid:107) = 1, writing z1 uniquely as z1 = w1 + iy1 with w1, y1 ∈ X, we must have either (cid:107)w1(cid:107) ≥ 1 2 , in which case we define x1 := w1/(cid:107)w1(cid:107), or (cid:107)y1(cid:107) ≥ 1 2 (in which case we define x1 = y1/(cid:107)y1(cid:107)). In either case, (cid:107)Tλx1(cid:107) ≤ ε/8. Indeed, considering the case where x1 = w1/(cid:107)w1(cid:107), we have (cid:107)w1(cid:107)(cid:107)(Tλ)C z1(cid:107) ≤ ε (cid:107)w1(cid:107)(cid:107)Tλw1 + iTλy1(cid:107) = (cid:107)Tλx1(cid:107) ≤ 1 1 8 . (The first inequality above follows from Lemma 1.4.16, and the final inequality from the fact that (cid:107)w1(cid:107) ≥ 1 2 .) The other case is dealt with similarly. Inductively, suppose we have constructed x1, . . . xn ∈ X with (cid:107)xi(cid:107) = 1 and satisfying the two properties above. As in the proof of Lemma 1.4.25, it is enough to show that we can find xn+1 ∈ X with (cid:107)xn+1(cid:107) = 1, (cid:107)Tλxn+1(cid:107) ≤ ε/2n+3 and having the property that whenever y ∈ E := spanR{x1, . . . xn}, λ ∈ R, (1 + an+1)(cid:107)y + λxn+1(cid:107) ≥ (cid:107)y(cid:107). 1, . . . y∗ 1)C, . . . (y∗ To this end, let y1, y2, . . . ym be a (finite) δ-net for the unit sphere SE of the finite dimensional subspace E, where 0 < δ < 1 is such that 1 + an+1 ≥ 1/(1 − δ). By the m ∈ SX∗ with Hahn-Banach Theorem, we can choose norm 1 linear functionals y∗ y∗ i (yi) = 1 for all i = 1, . . . m. Then (y∗ m)C are norm 1 linear functionals on XC with (y∗ i )C(yi) = 1 for all i. By our earlier observation, (Tλ)C is not an isomorphism when restricted to the finite co -- dimensional subspace Z := ∩m i )C ⊆ XC. It follows that we can choose a norm 1 vector, zn+1 ∈ Z with (cid:107)(Tλ)C zn+1(cid:107) ≤ ε/2n+4. Writing zn+1 = wn+1 + iyn+1 with wn+1, yn+1 ∈ X, we must have (for the same reason as earlier) either (cid:107)wn+1(cid:107) ≥ 1 2 . As before, we set xn+1 to be either wn+1/(cid:107)wn+1(cid:107) or yn+1/(cid:107)yn+1(cid:107) (depending on which of wn+1, yn+1 has norm at least 1 2 ). The same argument as before yields (cid:107)Tλxn+1(cid:107) ≤ ε/2n+3. Note also that for j ∈ {1, . . . m}, 0 = (y∗ j )C zn+1 = j (wn+1) + iy∗ y∗ j (xn+1) = 0 for all j. Finally, if y ∈ SE, λ ∈ R there is some i ∈ {1, . . . m} with (cid:107)y − yi(cid:107) ≤ δ, and j (yn+1). It follows that y∗ j (yn+1) and thus y∗ j (wn+1) = 0 = y∗ 2 or (cid:107)yn+1(cid:107) ≥ 1 i=1 Ker(y∗ (1 + an+1)(cid:107)y + λxn+1(cid:107) = (1 + an+1)(cid:107)yi + λxn+1 + y − yi(cid:107) i (yi + λxn+1 + y − yi) ≥ (1 + an+1) y∗ ≥ (1 + an+1) (1 − y∗ ≥ (1 + an+1) (1 − (cid:107)y − yi(cid:107)) ≥ (1 + an+1)(1 − δ) ≥ 1 i (y − yi)) 25 from which we easily conclude that (cid:107)y + λxn+1(cid:107) ≥ (cid:107)y(cid:107) for any y ∈ E, λ ∈ R as required. n = limn→∞ (cid:107)T n + K(X)(cid:107) 1 Proof of Theorem 1.4.28. Note that the limit certainly exists (in fact, it is a classical result that inf n (cid:107)T n + K(X)(cid:107) 1 is a complex Banach space and argue by contradiction, assuming T is a strictly singular operator with r(T ) > 0. Note it is enough to show that there exists some non-zero λ ∈ C, 0 < ε < λ and an infinite dimensional subspace Yε ⊆ X with (cid:107)T y − λy(cid:107) ≤ ε(cid:107)y(cid:107) whenever y ∈ Yε, since this contradicts the strict singularity of T . The argument is identical to that used by Gowers and Maurey to show that every operator on a complex HI space is a strictly n ). We consider first the case where X singular perturbation of the identity. For completeness we include it here. Denote by ΩT the set of all µ ∈ C such that T − µI is upper semi-Fredholm; we recall that the generalised index of such an operator is equal to some finite integer or −∞. We show that if r(T ) > 0, there is a non-zero λ ∈ C \ ΩT . It is a well known theorem of Atkinson that an operator U ∈ L(X) is Fredholm if and only if the class [U ] of U in the Calkin algebra is invertible. It follows that the essential spectrum of T, σess(T ) := {λ ∈ C : T −λI is not Fredholm} is precisely the spectrum, σ([T ]), of the element [T ] ∈ L(X)/K(X). It is well known that σ([T ]) (cid:54)= ∅ (since we are working over complex scalars) and moreover, it follows from well known results about the spectral radius that there exists λ ∈ σ([T ]) with λ = maxµ∈σ([T ]) µ = r(T ) > 0. We claim that λ /∈ ΩT . Note that if µ ∈ C with µ > λ then µ /∈ σ([T ]) and consequently T − µI is Fredholm. Moreover, by continuity of the index, T − µI is Fredholm with index 0 for all µ > λ (because this is true when µ is large enough to make T − µI invertible, i.e. when µ > (cid:107)T(cid:107)). Now if λ ∈ ΩT , we know that T − λI must be upper semi-Fredholm with generalised index −∞ (since T − λI is not Fredholm by choice of λ). It follows from Proposition 1.4.22 that the generalised index of (T − µI) = −∞ for all µ close to λ, but this contradicts the fact that T − µI is Fredholm with index 0 for all µ > λ. Thus 0 (cid:54)= λ ∈ C\ ΩT as required. Since λ /∈ ΩT , it follows by Lemma 1.4.25 that, given ε > 0, we can find an infinite dimensional subspace Yε of X with (cid:107)T y − λy(cid:107) ≤ ε(cid:107)y(cid:107) for every y ∈ Yε. Taking any ε < λ completes the proof of the lemma in the complex case. To prove the real case, we use a complexification argument that appeared in [16], though is attributed to Haydon (see [16, Lemma 20]). As in Section 1.4.5, we let XC = X ⊕ iX denote the complexification of X with the norm (cid:107)x + iy(cid:107)XC := supt∈[0,2π] (cid:107)x cos t− y sin t(cid:107)X . We recall that when L ∈ L(X), there is a unique complex linear extension of L, LC ∈ L(XC), defined by LC(x+iy) = Lx+iLy, Moreover, (cid:107)LC(cid:107) = (cid:107)L(cid:107) and if K ∈ K(X), then the complex linear extension KC ∈ K(XC). We refer the reader back to Lemma 1.4.17 for the details. From these observations, it easily follows that (cid:107)L + K(X)(cid:107) = (cid:107)LC + K(XC)(cid:107). 26 Now assume as before that T is a strictly singular operator with r(T ) > 0. It follows by n = limn→∞ (cid:107)T n + K(X)(cid:107) 1 the preceding remarks that limn→∞ (cid:107)(TC)n + K(XC)(cid:107) 1 n > 0. So by the proof just given for the complex case, there exists a non-zero λ ∈ C such that the operator TC−λIXC : XC → XC fails to be upper semi-Fredholm. It follows by Lemma 1.4.29 that for all ε > 0, there is an infinite dimensional subspace Yε ⊂ X such that (cid:107)TλYε(cid:107) < ε where we recall that Tλ is the operator T 2 − 2ReλT + λ2I ∈ L(X). If we choose ε < λ2, then it follows that T 2 − 2ReλT is an isomorphism on Yε. But, if T is strictly singular then so is T 2 − 2ReλT by Corollary 1.4.27, so we once again have a contradiction. 27 Chapter 2 The Bourgain-Delbaen Construction Introduction 2.1 In 1980, Bourgain and Delbaen [7, 6] introduced two classes of separable L∞-spaces, X and Y. We shall look at these classes in detail and give a precise definition of an L∞ space a little later in this chapter. For now, we simply remark that these spaces were shown to have a number of interesting properties, providing counterexamples for many previously unanswered conjectures in Banach space theory. For instance, none of these spaces have a subspace isomorphic to c0, giving the first set of examples of L∞ spaces with no c0 isomorph. In fact, it was shown in [7] that if Y ∈ Y, then every infinite dimensional subspace of Y contains a further infinite dimensional subspace which is reflexive. More recently, Argyros and Haydon [2] have managed to modify the original construc- tion to exhibit a space which solves the scalar-plus-compact problem. Although it is a question remaining from Argryos' and Haydon's solution to this problem that really mo- tivates our interest in the Bourgain-Delbaen spaces, we concern ourselves in this chapter with investigating the original Bourgain-Delbaen construction and the newer generalisa- tion used by Arygros and Haydon. We are a little relaxed with our terminology, referring to a Banach space constructed using either of the methods just mentioned as a space of Bourgain-Delbaen type. As is remarked in [2], it is interesting (and worth emphasising) that Bourgain-Delbaen constructions are different from the majority of other Banach space con- structions that occur in the literature. It is usual to start with the vector space of finitely supported scalar sequences, c00, and complete with respect to some exotic norm. We shall shortly see that spaces of Bourgain-Delbaen type are 'exotic subspaces' of (cid:96)∞, i.e. the norm is just the usual (cid:107) · (cid:107)∞ norm. We shall begin the chapter by looking at the Argyros-Haydon generalisation in Section 2.2. Here spaces of Bourgain-Delbaen type are subspaces of (cid:96)∞, obtained as the closed linear 28 span of the biorthogonal functionals of a special kind of Schauder basis for (cid:96)1. We will show that we can make a further, minor modification to the Argyros-Haydon construction. The new generalisation that we obtain will be an essential tool in proving the results of later chapters and as such, we choose to look at it in detail. Following this, we briefly recall the original Bourgain-Delbaen construction in Section 2.3, though we present it via the notation used by Haydon in [17]. This really is just notational convenience; it is clear that if one enumerates the set Γ (defined in Section 2.3) in a natural way, we get an obvious isometry from the subspace of (cid:96)∞(Γ) appearing in Section 2.3, and the subspace of (cid:96)∞(N) defined in the original work of Bourgain and Delbaen in [7]. Having looked in detail at the two different constructions, we will then, in some sense, unify them in Section 2.4 by showing that the Argyros-Haydon construction is a genuine generalisation of that due to Bourgain and Delbaen. Precisely, we will show the original Bourgain-Delbaen construction can be obtained as a special case of the Argyros-Haydon construction. Of course, this is not a new result. However, it is, to the author's best knowledge, the only documented exposition describing in detail the relationship between the two constructions. One of the advantages of the Arygros-Haydon generalisation is that it gives us a potential way to construct interesting operators on a space of Bourgain-Delbaen type, B. The idea is to define an operator on (cid:96)1, take the dual operator, and restrict to B. Of course, some care needs to be taken to ensure that the restriction also maps into B. Having shown how to describe the original spaces of Bourgain and Delbaen in the new framework of the Argyros-Haydon construction, we use the preceding idea to conclude the chapter (Section 2.5) by providing a partial answer to a question of Beanland and Mitchell, [4], namely if Y ∈ Y, is L(Y ) separable? Precisely, we will show that for any X ∈ X ∪ Y, there exists a constant CX > 0 and an uncountable collection of isometries on X which are pairwise distance CX apart with respect to the operator norm. Clearly this shows L(X) is non-separable for any X ∈ X ∪Y. We will note in Section 2.5 that there is in fact a shorter argument to see that L(X) is non-separable for X ∈ X and also defer to this section a short discussion of why the author finds this problem interesting. 2.2 The generalised Bougain-Delbaen Construction Before continuing any further, we recall that what we want to be able to do is construct interesting L∞ spaces. After all, it was seen in [2] that the finite-dimensional subspace structure possessed by such spaces is what proves to be so fundamental in showing that the Argyros-Haydon space solves the scalar-plus-compact problem. Moreover, we will exploit 29 similar techniques to prove results about the operator algebras of the new spaces we con- struct in subsequent chapters of this thesis. It is about time we define precisely what we mean by an L∞ space. sequence (Fn)n∈N of finite dimensional subspaces of X such that the union(cid:83) Definition 2.2.1. A separable Banach space X is an L∞,λ-space if there is an increasing n∈N Fn is dense . We say a Banach space is an L∞ space in X and, for each n, Fn is λ-isomorphic to (cid:96)dim Fn∞ if it is a L∞,λ space for some λ. The constructions of Bourgain and Delbaen, and the generalisation due to Argyros and Haydon, provide a way of constructing interesting classes of L∞ spaces. Consequently, we begin by looking in detail at these constructions, starting with the Argyros-Haydon generalisation. The content of this section follows very closely the work in [2] and we make no claim to originality unless explicitly stated. As remarked in the introduction to this chapter, the idea of the Argyros-Haydon con- struction is to construct a particular kind of Schauder basis for the space (cid:96)1 and to study the subspace X of (cid:96)∞ spanned by the biorthogonal elements. It follows by Proposition 1.4.3 that (cid:96)1 naturally embeds into X∗. (Moreover, by Theorem 1.4.6, this embedding is onto X∗ precisely when the Schauder basis of (cid:96)1 is boundedly complete.) Consequently, we will think of elements of (cid:96)1 as functionals and, as in [2], denote them using a star notation, b∗, c∗ and so on. In accordance with this notation, we denote the canonical basis of (cid:96)1(N) by (e∗ n)∞ n=1. It is perhaps easiest to understand the construction by working first with (cid:96)1(N) and this is indeed what is done in [2]. The special kind of Schauder basis of (cid:96)1(N) is obtained by considering a sequence (d∗ 1 = 0 n ⊆ {1, 2, . . . , (n − 1)} for n ≥ 2. An easy induction argument yields that the and supp c∗ n=1 are linearly independent and that moreover, the linear span [d∗ 2, . . . , d∗ n)∞ vectors (d∗ n] is the same as [e∗ n=1 ⊆ (cid:96)1(N) where each d∗ n)∞ n has the form d∗ n with c∗ n], so the closed linear span, [d∗ The clever part of the construction is to choose the c∗ 2, . . . , e∗ 1, e∗ n form a Schauder basis for (cid:96)1. In fact, things are a little more subtle than this, in the sense that the sequence (d∗ n)∞ n=1 will actually be a basic sequence equivalent to the canonical basis of (cid:96)1 if the c∗ n have sufficiently small norm. This case certainly won't be useful in producing any interesting Banach spaces, though we defer further discussion of this to Lemma 2.2.4 n = e∗ n−c∗ 1, d∗ n : n ∈ N] is the whole of (cid:96)1. n in such a way that the d∗ so as not to detract from the point at hand. Argyros and Haydon showed that the d∗ n will always form a Schauder basis for (cid:96)1 if the c∗ n assume a certain form (we refer the reader to [2] for more details). For this reason, it turns out that it is more convenient to work with the space (cid:96)1(Γ) for some suitably 30 defined countable set Γ; the elements γ ∈ Γ can then be used to code the form of the γ ∈ (cid:96)1(Γ). We choose to immediately state (and prove) the main result of this vector c∗ section, working with this more convenient notation, and refer the reader to [2] for a more comprehensive introduction. (cid:83) 1≤p≤q ∆p, Γ = (cid:83) Theorem 2.2.2. Let (∆q)q∈N be a disjoint sequence of non-empty finite sets; write Γq = 2 and a mapping τ defined on p∈N ∆p. Assume that there exists θ < 1 Γ \ ∆1, assigning to each γ ∈ ∆q+1 a tuple of one of the following forms: 0. (ε, α, ξ) with ε = ±1, 0 ≤ α ≤ 1 and ξ ∈ Γq; 1. (p, β, b∗) with 0 ≤ p < q, 0 < β ≤ θ and b∗ ∈ ball (cid:96)1 (Γq \ Γp); 2. (ε, α, ξ, p, β, b∗) with ε = ±1, 0 < α ≤ 1, 1 ≤ p < q, ξ ∈ Γp, 0 < β ≤ θ and γ − c∗ γ ∈ (cid:96)1(Γ) and projections P ∗ (0,q] on (cid:96)1(Γ) uniquely determined b∗ ∈ ball (cid:96)1 (Γq \ Γp). γ = e∗ Then there exist d∗ by the following properties: (A) P ∗ (0,q]d∗ γ = (B) c∗ γ = d∗ 0 γ (cid:40)  if γ ∈ Γq if γ ∈ Γ \ Γq if γ ∈ ∆1 if τ (γ) = (ε, α, ξ) if τ (γ) = (p, β, b∗) if τ (γ) = (ε, α, ξ, p, β, b∗). ξ 0 εαe∗ β(I − P ∗ εαe∗ (0,p])b∗ ξ + β(I − P ∗ (0,p])b∗ γ)γ∈Γ is a basis for (cid:96)1(Γ) with basis constant at most M = (1 − 2θ)−1. The The family (d∗ norm of each projection P ∗ (0,q] is at most M . The biorthogonal vectors dγ generate a L∞,M - subspace X(Γ, τ ) of (cid:96)∞(Γ); if the basis (dγ)γ∈Γ of X(Γ, τ ) is shrinking, then X∗ is naturally isomorphic to (cid:96)1(Γ). For each q and each u ∈ (cid:96)∞(Γq), there is a unique vector in [dγ : γ ∈ Γq] whose restriction to Γq is u; therefore there exists a unique extension operator iq : (cid:96)∞(Γq) → X(Γ, τ )∩ [dγ : γ ∈ Γq] and this operator has norm at most M . The subspaces Mq = [dγ : γ ∈ ∆q] = iq[(cid:96)∞(∆q)] form a finite-dimensional decomposition (FDD) for X. Remark. Strictly speaking, a Schauder basis is, in particular, a sequence indexed by the natural numbers, so there is a natural ordering of the vectors in the sequence. Therefore when we talk of (d∗ γ)γ∈Γ being a basis of (cid:96)1(Γ), or (dγ)γ∈Γ being a basis of X(Γ, τ ), we really need to enumerate Γ in some way. The enumeration we have in mind will always be the one that is described in the proof of this theorem, and we will henceforth speak of (for example) (dγ) being a basis without mentioning the explicit enumeration. 31 We observe that the statement here is almost identical to that of Theorem 3.5 in [2]. We have added an extra degree of freedom to tuples of form (0) and (2) appearing above via the addition of the parameter ε which takes values in {±1}. This means, for example, we can have vectors c∗ of the form −αe∗ (0,q])b∗. This change will only be useful for when we come to examine the connection between this construction and the original ξ + β(I − P ∗ Bourgain-Delbaen construction. In later chapters, we will always set ε = 1. The more important change we have made is that we do not demand the set ∆1 have only one element. This extra freedom will allow us to construct some non-trivial operators on the spaces we construct in later chapters. Since this theorem is so essential to this thesis, we include the proof, though all we are really doing is translating the arguments given in [2] into the notation used throughout this thesis. Proof of Theorem 2.2.2. The construction is recursive and to begin with, we work only γ = e∗ γ for γ ∈ ∆1. Recursively, assume that c∗ algebraically, not worrying about the continuity of the projection maps we define. We start by setting d∗ γ and d∗ γ have been defined for all γ ∈ Γn in accordance with the theorem. It follows from the recursive construction that the vectors (d∗ γ)γ∈Γn are linearly independent and that their span is all of (cid:96)1(Γn). Consequently, the (restricted) (0,q](cid:96)1(Γn) for q ≤ n are uniquely determined on (cid:96)1(Γn) by the formula given projections P ∗ in (A). For γ ∈ Γn+1, we can therefore define vectors c∗ γ − c∗ γ as in (B) and set d∗ γ. γ ⊆ ∆n+1, we see that the vectors γ has support in Γn and γ ∈ ∆n+1 so that supp e∗ Since c∗ (d∗ γ(cid:48))γ(cid:48)∈Γn+1 are linearly independent and that their span is (cid:96)1(Γn+1). We can thus extend the projections P ∗ (0,n+1] on (cid:96)1(Γn+1) in the same way. This completes the recursive construction; we obtain vectors (0,q] defined on c00(Γ) ⊆ (cid:96)1(Γ), satisfying γ and d∗ c∗ the conclusions of the theorem. (0,q] for q ≤ n to all of (cid:96)1(Γn+1) in accordance with (A) and define P ∗ γ for each γ ∈ Γ and projections P ∗ γ = e∗ γ = e∗ γ−c∗ We now show that (d∗ γ)γ∈Γ is a Schauder basis for (cid:96)1(Γ). In the process, we shall see that the projections P ∗ (0,q] defined on c00(Γ) are (uniformly) bounded by M , and thus extend to projections on all of (cid:96)1(Γ). Let us be a little more precise; for p ≥ 1 let kp = #Γp, set k0 = 0 and let n (cid:55)→ γ(n) : N → Γ be a bijection with the property that for each q ≥ 1, ∆q = {γ(n) : kq−1 < n ≤ kq}. We will show (d∗ n=1 is a Schauder basis. It is clear from the recursive construction that this sequence has dense linear span, so it is enough to see that the (densely defined) basis projections, P ∗ γ(n))∞ (cid:40) m defined by if j ≤ m otherwise d∗ 0 γ(j) γ(j) (cid:55)→ d∗ 32 are uniformly bounded. We show (cid:107)P ∗ immediately get (cid:107)P ∗ for all j ≤ m and 0 otherwise since span{d∗ obvious from the recursive construction. (0,q] = P ∗ (0,q](cid:107) ≤ M as claimed in the theorem. We also note that P ∗ m(cid:107) ≤ M for all m. Since P ∗ kq for all q, we γ(j) = e∗ me∗ γ(j) γ(j) : j ≤ m} = span{e∗ γ(j) : j ≤ m}; this is again me∗ me∗ γ(j) = e∗ Note that for j ≤ k1, d∗ γ(j). Consequently, for j ≤ k1, P ∗ every m and n in N. We shall prove by induction on n that (cid:107)P ∗ Since we are working on the space (cid:96)1(Γ) it is enough to show that (cid:107)P ∗ γ(n)(cid:107) ≤ M for γ(j)(cid:107) ≤ M for all m, j ≤ n. γ(j) is either equal to e∗ γ(j) or 0. In particular, there is nothing to prove for our inductive statement in the case when 1 ≤ n ≤ k1. Suppose inductively that (cid:107)P ∗ γ(j)(cid:107) ≤ M holds for all m, j ≤ n (w.l.o.g. n ≥ k1). We must see that this holds for all m, j ≤ n + 1. If j ≤ m, then, as observed in the previous paragraph, P ∗ γ(j) so there is nothing to prove. We therefore assume j > m. If, in addition, m < j ≤ n we are done by the inductive hypothesis. So we just need to see that (cid:107)P ∗ γ(n+1)(cid:107) ≤ M for all m ≤ n. Since we assume n ≥ k1, there exists a q ≥ 1 with γ(n + 1) ∈ ∆q+1. Note that this implies γ(j) = e∗ me∗ me∗ me∗ me∗ Γq ⊆ {γ(1), γ(2), . . . , γ(n)}. (2.1) We now make use of the fact that γ(n+1) = d∗ e∗ γ(n+1) + c∗ γ(n+1), where supp c∗ described in the theorem. We consider only the case where γ(n+1) ⊆ Γq ⊆ {γ(1), γ(2), . . . , γ(n)} and moreover c∗ γ(n+1) has one of the forms c∗ γ(n+1) = εαe∗ ξ + β(I − P ∗ (0,p])b∗, where ε = ±1, 1 ≤ p < q, ξ ∈ Γp, b∗ ∈ ball (cid:96)1(Γq \ Γp) and α, β are as in the theorem, with β ≤ θ by our hypothesis. The other cases are similar. Now, because n + 1 > m we have P ∗ γ(n+1) = 0. We can also write ξ = γ(j) where j ≤ kp < kq ≤ n (this follows from the fact that ξ ∈ Γp with p < q and Equation 2.1). As observered earlier P ∗ . Combining these observations, we get md∗ kp (0,p] = P ∗ P ∗ me∗ γ(n+1) = εαP ∗ me∗ γ(j) + β(P ∗ m − P ∗ min{m,kp})b∗. If kp ≥ m the second term vanishes so that γ(n+1)(cid:107) = α(cid:107)P ∗ me∗ me∗ (cid:107)P ∗ γ(j)(cid:107) ≤ (cid:107)P ∗ me∗ γ(j)(cid:107), which is at most M by our inductive hypothesis. 33 If, on the other hand, kp < m, we certainly have j < m (since recall j ≤ kp) so that γ(j) = e∗ me∗ P ∗ γ(j), leading to the estimate γ(n+1)(cid:107) ≤ α(cid:107)e∗ γ(j)(cid:107) + β(cid:107)P ∗ Now b∗ is a convex combination of functionals ±e∗ hypothesis is applicable to all of these. We thus obtain me∗ (cid:107)P ∗ kpb∗(cid:107). mb∗(cid:107) + β(cid:107)P ∗ γ(l) with l ≤ kq ≤ n, and our inductive (cid:107)P ∗ me∗ γ(n+1)(cid:107) ≤ α + 2M β ≤ 1 + 2M θ = M, by the definition of M = 1/(1 − 2θ) and the assumption that 0 ≤ β ≤ θ. We conclude that (d∗ γ(n))∞ n=1 is a Schauder basis for (cid:96)1(Γ) as claimed. It follows from Proposition 1.4.3 that the biorthogonal functionals (dγ)γ∈Γ form a Schauder basis for the Banach space X := X(Γ, τ ) = [dγ : γ ∈ Γ] ⊆ (cid:96)∞(Γ), or more precisely that the sequence (dγ(n))∞ n=1 is a Schauder basis for X. We easily deduce from this that the finite dimensional subspaces Mq := [dγ : γ ∈ ∆q] form a finite dimensional decomposition of X. Moreover, standard results in Banach space theory yield that X∗ is naturally isomorphic to (cid:96)1 in the case that the basis (dγ)γ∈Γ is shrinking (see Theorem 1.4.6). We must also show that X is a L∞,M space. To see this, we consider the projections P ∗ (0,n] defined in the theorem. We have just shown these are well-defined and moreover (0,n](cid:107) ≤ M . If we modify P ∗ (cid:107)P ∗ (0,n] = (cid:96)1(Γn), rather than the whole of (cid:96)1(Γ), what we have is a quotient operator, which we shall denote qn, of norm at most M . The dual of this quotient operator is an isomorphic embedding in : (cid:96)∞(Γn) → (cid:96)∞(Γ), also of norm at most M . Of course, this is immediate by Lemma 1.4.11. However, we can also show it explicity; if u ∈ (cid:96)∞(Γn), γ ∈ Γn, we have (0,n] by taking the codomain to be the image im P ∗ (inu)(γ) = (cid:104)inu, e∗ γ(cid:105) = (cid:104)u, e∗ So in is an extension operator (cid:96)∞(Γn) → (cid:96)∞(Γ) and we have (cid:107)u(cid:107)∞ ≤ (cid:107)inu(cid:107)∞ ≤ M(cid:107)u(cid:107)∞ γ(cid:105) = (cid:104)u, qne∗ γ(cid:105) = u(γ). for all u ∈ (cid:96)∞(Γn). We claim that the image of in is precisely [dγ : γ ∈ Γn] and that for each γ ∈ ∆n, ineγ = dγ so that in((cid:96)∞(∆n)) = Mn as claimed. We prove these facts in the following lemma (Lemma 2.2.3) as we shall use them again later. Now, since the image of in is M -isomorphic to (cid:96)∞(Γn) ≡ (cid:96)kn∞ and ∪n∈N[dγ : γ ∈ Γn] is dense in X, it follows that X is a L∞,M -space. To complete the proof, it remains to prove that for each n ∈ N and u ∈ (cid:96)∞(Γn) there exists a unique vector in [dγ : γ ∈ Γn] which when restricted to Γn is precisely the vector u. In fact, we have already shown the existence of this vector; in(u) satisfies this property by 34 the above argument. So we only need to prove uniqueness. Suppose v(u) =(cid:80) γ ⊂ Γn. Thus the αγ are uniquely determined by u and so v(u) is unique. is such a vector. It follows that αγ = v(u)(d∗ because supp d∗ Lemma 2.2.3. The image of the map in : (cid:96)∞(Γn) → (cid:96)∞(Γ) defined in the proof of the above theorem is precisely [dγ : γ ∈ Γn]. Moreover, for each γ ∈ ∆n, ineγ = dγ. γ) = u(d∗ αγdγ γ); the final equality must be true γ∈Γn Proof. We saw that the operator in is an isomorphic embedding, so its image is kn-dimensional, where as in the previous proof, kn = #Γn. Since the vectors (dγ)γ∈Γn ⊆ (cid:96)∞(Γ) are linearly independent, it is enough to show that for each γ ∈ Γn there is a u ∈ (cid:96)∞(Γn) such that θ : θ ∈ Γ] = (cid:96)1(Γ), it would be enough to see there is a u ∈ (cid:96)∞(Γn) such inu = dγ. Since [d∗ that (cid:104)inu, d∗ θ(cid:105) = δθ,γ. This ends up being an easy exercise in linear algebra. θ(cid:105) = (cid:104)dγ, d∗ We choose to give an alternative proof. Fix γ ∈ Γn. We will find u ∈ (cid:96)∞(Γn) such that inu(θ) = dγ(θ) for all θ ∈ Γ. Since in is an extension operator, if such a u exists, we must have ∈ (cid:96)∞(Γn). All that remains to be u(θ) = inu(θ) = dγ(θ) whenever θ ∈ Γn. Consequently we set u = (dγ(θ))θ∈Γn seen is that inu(θ) = dγ(θ) for all θ ∈ Γ \ Γn. To this end we note that (cid:88) dγ(ξ)[P ∗ (0,n]e∗ θ(cid:105) = (cid:104)dγ, P ∗ (0,n]e∗ θ(cid:105) = (cid:104)dγ, e∗ = dγ(θ). inu(θ) = (cid:104)inu, e∗ θ(cid:105) = (cid:104)u, P ∗ (0,n]e∗ θ(cid:105) = ξ∈Γn θ](ξ) The penultimate equality follows from the fact that dγ(x∗) = dγ(P ∗ and x∗ ∈ (cid:96)1(Γ). This is a consequence of the biorthogonalilty of the sequences (d∗ and (dγ) ⊆ (cid:96)∞(Γ). (0,n]x∗) for all γ ∈ Γn γ) ⊆ (cid:96)1(Γ) Finally, note that if γ ∈ ∆n and θ ∈ Γn then θ(cid:105) = (cid:104)dγ, d∗ dγ(θ) = (cid:104)dγ, e∗ θ(cid:105) + (cid:104)dγ, c∗ θ(cid:105) = (cid:104)dγ, d∗ θ(cid:105) θ ∈ (cid:96)1(Γn−1) so (cid:104)dγ, c∗ since c∗ dγ(θ) = 1. In other words, when γ ∈ ∆n, u = eγ, proving that ineγ = dγ as claimed. θ(cid:105) = 0. Therefore dγ(θ) (cid:54)= 0 only when γ = θ, in which case Having proved the main theorem of this section, we return to a remark that we noted γ)γ∈Γ of (cid:96)1(Γ) obtained in the Bourgain-Delbaen con- γ)γ∈Γ. To obtain any interesting γ's has to be earlier. One can think of the basis (d∗ struction as a perturbation (by c∗ Banach spaces, we note that the norms of the perturbing vectors, i.e. the c∗ γ) of the canonical basis (e∗ 35 n ∈ (cid:96)1(N) with d∗ 'large'. If not, then the basis (d∗ γ)γ∈Γ will be equivalent to the canonical basis of (cid:96)1 and the corresponding Bourgain-Delbaen space will simply be an isomorph of c0. More precisely, we have the following lemma. Lemma 2.2.4. Let (d∗ each n ∈ N there exists a vector c∗ exists θ < 1 such that (cid:107)c∗ basis (e∗ [dn : n ∈ N], is isomorphic to c0. n)∞ n=1 be a sequence of vectors in (cid:96)1(N) with the property that for n − c∗ n = e∗ n. Suppose further that there n(cid:107)(cid:96)1 ≤ θ for all n. Then (d∗ n)∞ n=1 is equivalent to the canonical n)∞ n=1 of (cid:96)1 and the closed subspace of (cid:96)∞ generated by the biorthogonal vectors, Proof. It is easily checked that T : (cid:96)1 → (cid:96)1, T ((cid:80)∞ it is enough to show that (cid:107)I − T(cid:107) < 1. This is easy, since (cid:107)(I − T )(cid:80)∞ (cid:107)(cid:80)∞ j defines a bounded operator on (cid:96)1 (with norm at most 1 + θ). To see the sequence (d∗ n) is equivalent to the canonical basis, it is enough to see that T has continuous inverse. By standard results, j(cid:107)(cid:96)1 = j(cid:107) ≤ θ(cid:107)(aj)(cid:107)(cid:96)1 by the assumed norm condition on the c∗ j )(cid:107) = (cid:107)(cid:80)∞ j ) =(cid:80)∞ j=1 aj(e∗ j=1 ajd∗ j=1 aje∗ j=1 aje∗ j=1 ajc∗ j − d∗ n vectors. It follows that (cid:107)I − T(cid:107) ≤ θ < 1. The fact that X := [dn : n ∈ N] is isomorphic to c0 follows by taking the dual of T and restricting to X. It is straightforward to check that T ∗ maps the vector dn ∈ (cid:96)∞ to the vector en ∈ (cid:96)∞. Moreover, since T is an isomorphism, so is T ∗, and consequently T ∗(X) is closed. The previous two observations imply that T ∗ restricted to X is an isomorphism onto T ∗(X) = [en : n ∈ N] = c0. We conclude this section by proving that the basis, (dγ)γ∈Γ, of any Bourgain-Delbaen space, X(Γ, τ ), is normalised. Lemma 2.2.5. Continuing with the notation as above, if X(Γ, τ ) is a Bourgain-Delbaen space as in Theorem 2.2.2, then the basis (dγ)γ∈Γ is normalised. Proof. Fix ν ∈ Γ. We prove by induction on n that if γ ∈ ∆n then dν(e∗ γ ∈ ∆1, e∗ some k and choose γ ∈ ∆k+1. We make use of the decomposition e∗ that c∗ the other possible forms of c∗ γ) ≤ 1. When γ and the proof is obvious. So assume the estimate holds for all γ(cid:48) ∈ Γk for γ and assume 2 , ξ ∈ Γl and b∗ ∈ ball (cid:96)1(Γk \ Γl); (l,k]b∗ for some l < k, α ≤ 1, β ≤ θ < 1 γ = εαe∗ ξ + βP ∗ γ = d∗ γ = d∗ γ + c∗ γ) = (cid:104)dν, d∗ We consider two possible cases. If dν(d∗ ν ∈ ∆k+1 and dν(e∗ holds. γ) = dν(d∗ γ can be treated similarly and are in any case easier. Now, dν(e∗ γ + εαe∗ γ) (cid:54)= 0, then we must have ν = γ. In particular γ ⊆ (cid:96)1(Γk). Consequently the inequality γ) = 1 since supp c∗ ξ + βP ∗ (l,k]b∗(cid:105) 36 The remaining case is when dν(d∗ If, in addition, ν ∈ Γl then it follows that dν(e∗ αdν(e∗ ξ) ≤ 1 by the inductive hypothesis and the fact that α ≤ 1. γ) = 0. We divide this case into two further possibilities. γ) ≤ ξ) and consequently, dν(e∗ γ) = dν(εαe∗ Finally, if dν(d∗ γ) = 0 and ν ∈ Γ \ Γl, dν(e∗ (l,k]b∗(cid:105) = β(cid:104)P(l,k]dν, b∗(cid:105). This last expression is either 0 (if ν ∈ Γ \ Γk) or β(cid:104)dν, b∗(cid:105) otherwise. In either case, it is again easily seen using the inductive hypothesis (and that b∗ ∈ ball (cid:96)1(Γk)) that dν(e∗ γ) ≤ 1 as required. γ) = (cid:104)dν, βP ∗ It follows that (cid:107)dν(cid:107) ≤ 1. On the other hand, it is easily seen that dν(e∗ ν) = (cid:104)dν, d∗ ν +c∗ ν(cid:105) = 1 so that (cid:107)dν(cid:107) ≥ 1. It follows that (cid:107)dν(cid:107) = 1 as required. 2.3 The Bourgain-Delbaen Construction We will return to the construction just described shortly. For now, we move on to look at a few details of the original Bourgain-Delbaen construction so that we can explain the relationship between the two constructions in the following section. The notation that follows is that used by Haydon in [17]. We begin by setting ∆1 to be the set consisting of just one element. Inductively, if we have sets ∆k defined for k = 1, 2, . . . , n, we set Γn = ∪n ∆n+1 = {n + 1} × (cid:91) 1≤k<n k=1∆k and define {k} × Γk × Γn × {±1} × {±1}. We let Γ = ∪n≥1Γn. We make two observations. First, ∆2 = ∅ so that Γ1 = Γ2. Second, an element γ ∈ ∆n+1 is a 6-tuple of the form γ = (n + 1, k, ξ, η, ε, ε(cid:48)). We shall use the terminology introduced by Haydon and say that such a γ has rank n + 1. For fixed real numbers a, b > 0, we define inductively, linear maps un : (cid:96)∞(Γn) → (cid:96)∞(∆n+1) (for n ≥ 2) and im,n : (cid:96)∞(Γm) → (cid:96)∞(Γn) (for all n > m, n, m ∈ N). We denote by in the map in,n+1. We set i1 = i1,2 := id(cid:96)∞(Γ1). Suppose, inductively, that ik has been defined for all k < n. Let ik,n = in−1 ◦ ··· ◦ ik : (cid:96)∞(Γk) → (cid:96)∞(Γn). For x ∈ (cid:96)∞(Γ) we denote by πkx the obvious restriction of x onto coordinates in Γk. We define un : (cid:96)∞(Γn) → (cid:96)∞(∆n+1) by (unx)(n + 1, k, ξ, η, ε, ε(cid:48)) = εax(ξ) + ε(cid:48)b[x(η) − (ik,nπkx)(η)] and in : (cid:96)∞(Γn) → (cid:96)∞(Γn+1) by x(γ) (unx)(γ) γ ∈ ∆n+1. It is easily seen that for m < n < p and x ∈ (cid:96)∞(Γm) we have inx(γ) = γ ∈ Γn (cid:40) (im,px)Γn = im,nx 37 and so we can well-define a linear mapping jm : (cid:96)∞(Γm) → RΓ by setting (jmx)(δ) = (im,nx)(δ) whenever δ ∈ Γn. Bourgain and Delbaen supposed the numbers a and b are chosen subject to one of the following conditions: (1) 0 < b < 1 2 < a < 1 and a + b > 1 (in which case we define λ := a 1−2b ) (2) a = 1 and 0 < b < 1 2 such that 1 + 2bλ ≤ λ for some λ > 1. In either of these cases, we obtain the following lemma: Lemma 2.3.1. The maps jm take values in (cid:96)∞(Γ). Moreover, for each m, jm is a linear isomorphic embedding of (cid:96)∞(Γm) into (cid:96)∞(Γ) satisfying (cid:107)x(cid:107) ≤ (cid:107)jm(x)(cid:107) ≤ λ(cid:107)x(cid:107) for all x ∈ (cid:96)∞(Γm) (where λ is as above). If m < n then im jm ⊆ im jn. Proof. It is clear from the definition that jm is an extension operator from (cid:96)∞(Γm) to (cid:96)∞(Γ) and consequently (cid:107)x(cid:107) ≤ (cid:107)jm(x)(cid:107) for all x ∈ (cid:96)∞(Γm). To complete the first part of the lemma, it is clearly enough to see that (cid:107)im,n(cid:107) ≤ λ for all m < n. The proof was given in [7] and is an easy induction argument. Indeed, since i1,2 = id(cid:96)∞(Γ1), we have (cid:107)i1,2(cid:107) = 1 ≤ λ. Suppose inductively (cid:107)ik,m(cid:107) ≤ λ for all k < m ≤ n. We consider an x ∈ (cid:96)∞(Γn) and show first that (cid:107)in,n+1(cid:107) ≤ λ. It is enough to show that un(x)(γ) ≤ λ(cid:107)x(cid:107) for γ ∈ ∆n+1. To this end, write γ = (n + 1, k, ξ, η, ε, ε(cid:48)). We have unx(γ) = aεx(ξ) + bε(cid:48)[x(η) − ik,nπkx(η)] ≤ a(cid:107)x(cid:107) + b((cid:107)x(cid:107) + λ(cid:107)x(cid:107)) ≤ (a + 2bλ)(cid:107)x(cid:107) where we have again made use of the fact that λ ≥ 1 in the final inequality. We are now done, since in either of the cases we have a + 2bλ ≤ λ. It remains to consider (cid:107)im,n+1(cid:107) when m < n. Let x ∈ (cid:96)∞(Γm). If γ ∈ Γn, we have im,n+1x(γ) = in,n+1 ◦ im,nx(γ) = im,nx(γ) ≤ λ(cid:107)x(cid:107) by the inductive hypothesis. So it once again remains to consider those γ = (n + 1, k, ξ, η, ε, ε(cid:48)) in ∆n+1. By a similar argument we get im,n+1x(γ) = aεim,nx(ξ) + bε(cid:48)[im,nx(η) − ik,nπkim,nx(η)]. We consider two separate cases. Firstly, if k ≤ m, since ξ ∈ Γk, we have im,nx(ξ) = x(ξ) and also, from construction of the map im,n we have πkim,nx = πkx. We therefore have im,n+1x(γ) = aεx(ξ) + bε(cid:48)[im,nx(η) − ik,nπkx(η)] and estimating as before yields this is at most (a + 2bλ)(cid:107)x(cid:107) ≤ λ(cid:107)x(cid:107). 38 Otherwise n > k > m and it is easily verified that πkim,nx = im,kx. So ik,nπkim,nx = ik,n ◦ im,kx = im,nx. Consequently im,n+1x(γ) = aεim,nx(ξ) ≤ (cid:107)im,nx(cid:107) ≤ λ(cid:107)x(cid:107). Here we made use of the fact that a ≤ 1 and the inductive hypothesis. This completes the induction and the first part of the proof. It remains to see that if m < n, im jm ⊆ im jn. We will show that for x ∈ (cid:96)∞(Γm), jmx = jn(πnjmx). It is easy to see that πnjmx = im,nx so it suffices to show that jmx = jn(im,nx). To see this, note that if δ ∈ Γ, we can choose p > n > m such that δ ∈ Γp. We then have jn(im,nx)(δ) = in,p(im,nx)(δ) = im,px(δ) = jmx(δ) as required. It follows from the lemma that X = Xa,b = (cid:91) n≥1 im(jn) is a closed subspace of (cid:96)∞(Γ). Moreover, the lemma shows that the image, im(jn) is λ- isomorphic to (cid:96)∞(Γn) for every n so that X is a L∞,λ space. In Bourgain's notation, the spaces of class X are the spaces Xa,b for which a = 1 and 0 < b < 1 2 , whilst those of class Y are the spaces Xa,b with 0 < b < 1 2 < a < 1 and a + b > 1. Remark 2.3.2. We conclude this section with a remark. It is shown in [17] that each of the spaces, Xa,b, described above has a natural Schauder basis, and finite dimensional decomposition. More precisely, for γ ∈ Γ, let n = rank (γ) and eγ be the usual unit vector (i.e. eγ(δ) = 1 if δ = γ and 0 otherwise). Define dγ := jn(eγ). The sequence (dγ)γ∈Γ (under a suitable enumeration) is a Schauder basis for X. In fact, it is observed in [17] that this basis is normalized, i.e. (cid:107)dγ(cid:107) = 1 for all γ. Moreover, if we let Mn := span{dγ : γ ∈ ∆n}, then (Mn)∞ n=1 is a FDD for X. We will see in the next section that once we have shown that we can describe the original Bourgain-Delbaen construction in the Argyros-Haydon framework, these facts are immediate. 2.4 Connecting the two constructions As explained earlier, we wish to formulate the original Bourgain-Delbaen construction just discussed within the framework of the generalised construction discussed in Section 2.2. We continue with the same notation from the previous section; in particular, we work in this section with the set Γ just described in the previous section. We continue to denote vectors of (cid:96)1 with a star notation. This notation allows us to distinguish between eγ (the standard unit vector in (cid:96)∞(Γ)) and e∗ γ (the standard unit vector in (cid:96)1(Γ)) for example. Our aim is to 39 construct a suitable τ mapping on Γ as in Theorem 2.2.2 so that we can apply this theorem, obtaining a space of Bourgain-Delbaen type, which is precisely the same space obtained in the previous section. We note that the extension operator jn : (cid:96)∞(Γn) → (cid:96)∞(Γ) defined in the previous section could really be thought of as an isomorphic embedding from (cid:96)1(Γn)∗ to (cid:96)1(Γ)∗ under the canonical identifications, (cid:96)1(Γn)∗ ≡ (cid:96)∞(Γn) and (cid:96)1(Γ)∗ ≡ (cid:96)∞(Γ). Thought of as a map on these dual spaces, it is easy to see that jn is weak*-norm, and thus weak*-weak* continuous. Indeed, this follows since the weak* and norm topologies coincide on finite dimensional spaces, and we saw in the previous section that jn is norm-norm continuos. It follows by Lemma 1.4.11 that jn can be obtained as the dual of some operator from (cid:96)1(Γ) to (cid:96)1(Γn). We will show that under the correct choice of τ mapping on Γ, this dual operator is exactly the operator P ∗ (0,n] appearing in Theorem 2.2.2. If we are attempting to find the operator which has dual jn, it is sensible to consider the dual of jn since, roughly speaking, jn and j∗∗ n can be considered 'the same' operator. More precisely, it is well known that for any operator T : X → Y , the operator T ∗∗JX maps into JY (Y ) and moreover, T = J−1 Y T ∗∗JX . Thinking about how the operators jn are defined, to compute j∗ n, we must first compute the dual of the operators im,n. We begin by computing the dual operator of in, where we assume n ≥ 2, since i1 = Id(cid:96)∞(Γ1). Throughout this section, we will make the usual identifications of (cid:96)1(Γn)∗ with (cid:96)∞(Γn) and (cid:96)∞(Γn)∗ with (cid:96)1(Γn). Under these identifications, we compute i∗ n, obtaining n : (cid:96)1(Γn+1) → (cid:96)1(Γn). For x∗ ∈ (cid:96)1(Γn+1), γ ∈ Γn, i∗ (cid:88) n(x∗)](γ) = x∗(in(eγ)) [i∗ = x∗(γ) + x∗(ξ)[in(eγ)](ξ). ξ∈∆n+1 nx∗ = x∗ and we see that i∗ We observe that if supp x∗ ⊆ Γn then all the terms appearing in the above sum are zero n : (cid:96)1(Γn+1) → except the x∗(γ) term. Thus, for such vectors, i∗ (cid:96)1(Γn) is a projection onto (cid:96)1(Γn). Furthermore, we note that this is not the usual projection θ to itself when θ ∈ Γn and 0 otherwise). To see this, we on (cid:96)1(Γn) (which would send e∗ exhibit a γ ∈ Γ with rank γ = n + 1 but i∗ γ (cid:54)= 0. We take γ = (n + 1, k, ξ, η, 1, 1) with ne∗ k < n and both ξ and η belonging to Γk. Setting x∗ = e∗ γ in the previous formula we see γ)(ξ) = in(eξ)(γ) = aeξ(ξ) + b[eξ(η) − (ik,nπkeξ)(η)] = a (cid:54)= 0. (Here, we made use that i∗ of the fact that ik,nπkeξ(η) = eξ(η) since ξ and η are in Γk.) n(e∗ Given a γ with rank γ = n + 1, we can obviously decompose e∗ projection, i∗ ne∗ γ, of e∗ γ, and a vector in ker i∗ n. Precisely, we have γ into a sum of the e∗ γ = i∗ n(e∗ γ) + [e∗ γ − i∗ n(e∗ γ)]. 40 Recalling that we are attempting to obtain the original Bourgain-Delbaen spaces via the n(e∗ γ := 0. γ := i∗ γ) and d∗ generalised construction of Argyros-Haydon and comparing this formula with Theorem γ (note we are still assuming n ≥ 2). For 2.2.2, we define c∗ γ ∈ Γ1 define c∗ Remark 2.4.1. Note the vectors d∗ γ just defined should not be confused with the vectors appearing in Thorem 2.2.2; in any case, we haven't even defined a mapping τ on Γ yet γ and c∗ γ = e∗ γ − c∗ for the theorem to be applied in any sensible way. The vectors are defined by the formulae just stated and we work with these vectors throughout this section unless explicitly stated otherwise. The reason for this (perhaps slightly confusing) notation is the following. We appearing in Theorem 2.2.2. We can thus apply the theorem, and obtain vectors (cid:98)c∗ will show that for each γ ∈ Γ, it is possible to define a tuple, τ (γ), having one of the forms (cid:99)d∗ γ and γ = e∗ γ. Moreover, it will turn out that the space of Bourgain-Delbaen type constructed via Theorem 2.2.2 under γ. We will eventually show that (cid:98)c∗ γ and consequently(cid:99)d∗ γ −(cid:98)c∗ γ = d∗ γ = c∗ this τ mapping is precisely the original Bourgain-Delbaen space of the previous section (for a given pair of real numbers a, b satisfying the conditions already discussed). It remains to see how to construct such a τ map. To this end, we write γ = (n + 1, k, ϕ, η, ε, ε(cid:48)) and for γ ∈ Γn, we look at [i∗ n(e∗ γ)](γ) = e∗ γ(γ) + (cid:88) ξ∈∆n+1 e∗ γ(ξ)[in(eγ)](ξ) = ineγ(γ) = uneγ(γ) = εaeγ(ϕ) + ε(cid:48)b[eγ − ik,nπkeγ](η). The second term appearing in the above sum is clearly the harder term to deal with so we look at this first. There is one situation in which this term is easy to compute, namely if η ∈ Γk. In this case ik,nπkeγ(η) = πkeγ(η), so [eγ − ik,nπkeγ](η) = [eγ − πkeγ](η). We then consider cases: (1) γ (cid:54)= η. In this case, we clearly have that eγ(η) = 0. Since the only possible vectors that πkeγ can be are 0 and eγ, we see that πkeγ(η) = 0 also. Thus [eγ − ik,nπkeγ](η) = 0. (2) γ = η ∈ Γk. In this case, πkeγ = eγ and we again find [eγ − ik,nπkeγ](η) = 0. To obtain a useful expression for [eγ − ik,nπkeγ](η) when η ∈ Γn \ Γk, it turns out to m,n : (cid:96)1(Γn) → be useful to look at the dual operator of im,n : (cid:96)∞(Γm) → (cid:96)∞(Γn), giving i∗ 41 (cid:96)1(Γm). For x∗ ∈ (cid:96)1(Γn), γ ∈ Γm [i∗ m,n(x∗)](γ) = x∗(im,neγ) = x∗(γ) + x∗(ξ)[im,neγ](ξ). (cid:88) ξ∈Γn\Γm We take this opportunity to document a special case of this formula which will be particularly useful to us later. We note that for η ∈ Γn \ Γm, γ ∈ Γm, we have [i∗ m,n(e∗ As before, it is easy to see that each i∗ η)](γ) = im,neγ(η) (2.2) m,n is a projection onto (cid:96)1(Γm). Since ∪n(cid:96)1(Γn) = (0,k] : (cid:96)1(Γ) → c00(Γ) is a dense subspace of (cid:96)1(Γ), it is natural to attempt to define projections P ∗ (cid:96)1(Γk) by (0,k](x) = i∗ P ∗ k,n(x) whenever x ∈ (cid:96)1(Γn), (n > k). Assuming this is well-defined, the formula defines (linear) projections P ∗ (0,k] from the dense subspace c00(Γ) ⊆ (cid:96)1(Γ), onto (cid:96)1(Γk). Moreover, it follows from the proof of Lemma k,n(cid:107) ≤ λ for all k < n. Consequently 2.3.1 (and standard facts about dual operators) that (cid:107)i∗ P ∗ (0,k] is bounded (with norm at most λ) and thus extends to a bounded projection defined on (cid:96)1(Γ) as required. So it remains to check that we can well define P ∗ (0,k] by the above formula. Suppose x∗ ∈ (cid:96)1(Γn) ⊂ (cid:96)1(Γn(cid:48)) (n(cid:48) > n > k). We saw earlier that for γ ∈ Γk, k,n(cid:48)(x∗)](γ) = x∗(γ) + [i∗ = x∗(γ) + x∗(ξ)[ik,n(cid:48)eγ](ξ) x∗(ξ)[ik,n(cid:48)eγ](ξ) (cid:88) (cid:88) ξ∈Γn(cid:48)\Γk ξ∈Γn\Γk where the final equality is because x∗ ∈ (cid:96)1(Γn) so x∗(ξ) = 0 for ξ ∈ Γn(cid:48) \ Γn. It is immediate from the definition of the operators im,n that ik,n(cid:48) = in,n(cid:48) ◦ ik,n. So for ξ ∈ Γn, ik,n(cid:48)eγ(ξ) = in,n(cid:48) ◦ ik,neγ(ξ) = ik,neγ(ξ). Combining these observations, we find that [i∗ k,n(cid:48)(x∗)](γ) = x∗(γ) + (cid:88) (cid:88) k,n(x∗)](γ) ξ∈Γn\Γk ξ∈Γn\Γk = x∗(γ) + = [i∗ x∗(ξ)[ik,n(cid:48)eγ](ξ) x∗(ξ)[ik,neγ](ξ) so that P ∗ (0,k] is well defined. We recall what we set out to do. For γ = (n + 1, k, ϕ, η, ε, ε(cid:48)) ∈ ∆n+1 and γ ∈ Γn, we γ)](γ). We have seen already that γ(γ) := [i∗ n(e∗ were attempting to find an expression for c∗ γ(γ) = εaeγ(ϕ) + ε(cid:48)b[eγ − ik,nπkeγ](η). c∗ We consider 2 separate cases: 42 (1) η ∈ Γk. We have seen that the second term in the above expression is just 0. We thus see that c∗ γ(γ) = εaeγ(ϕ) = εae∗ ϕ(γ). In other words, c∗ γ = εae∗ ϕ. (2) η ∈ Γn \ Γk. If in addition, rank (γ) ≤ k, it follows that πkeγ = eγ. Combining this observation with equation (2.2) and the definition of P ∗ (0,k], we get (0,k](e∗ [P ∗ η)](γ) = [i∗ k,n(e∗ η)](γ) = ik,neγ(η) = ik,nπkeγ(η). Thus γ(γ) = εaeγ(ϕ) + ε(cid:48)b[eγ − ik,nπkeγ](η) c∗ η)](γ) (cid:110) η − [P ∗ ϕ(γ) + ε(cid:48)b[e∗ [(I − P ∗ ϕ(γ) + ε(cid:48)b (0,k](e∗ (0,k])(e∗ = εae∗ = εae∗ η)](γ) (cid:111) . If on the other hand rank (γ) > k, πkeγ = 0 so γ(γ) = εaeγϕ + ε(cid:48)b[eγ − ik,nπkeγ](η) c∗ = εae∗ = εae∗ (cid:110) ϕ(γ) + ε(cid:48)be∗ ϕ(γ) + ε(cid:48)b η(γ) [(I − P ∗ (0,k])(e∗ η)](γ) where the final equality follows since rank (γ) > k =⇒ [P ∗ x∗ ∈ (cid:96)1(Γ). We thus conclude that (cid:111) (0,k](x∗)](γ) = 0 for any c∗ γ = εae∗ ϕ + ε(cid:48)b(I − P ∗ (0,k])(e∗ η). We make one more observation before coming to the main result of this section. Remark 2.4.2. An easy induction argument yields that span{d∗ γ : rank γ ≤ n} = span{e∗ γ : rank γ ≤ n} for every n. Indeed, the base case is immediate since by definition, d∗ γ = e∗ γ for rank γ ≤ 2 (recalling that ∆2 = ∅ so Γ1 = Γ2). The inductive step follows from the relation e∗ γ = d∗ γ ⊆ Γn when rank γ = n + 1. γ for all γ and supp c∗ γ + c∗ If follows that [d∗ γ : γ ∈ Γ] = (cid:96)1(Γ). Consequently the operators P ∗ are completely determined by their action on the d∗ (0,k] previously defined (cid:40) d∗ 0 γ γ. In fact rank (γ) ≤ k otherwise. P ∗ (0,k](d∗ γ) = To see this, we note the formula certainly holds when rank (γ) ≤ k since we have already observed that the maps P ∗ γ) = 0 when rank (γ) > k. We consider the case where k ≥ 2. Suppose first rank (γ) = k + 1. Then P ∗ γ = 0. (We have of course (0,k] are projections onto (cid:96)1(Γk). It remains to see that P ∗ γ) = i∗ γ) = i∗ γ) = i∗ γ) − c∗ (0,k](d∗ γ − c∗ γ = c∗ γ − c∗ (0,k](d∗ k(d∗ k(e∗ k(e∗ 43 γ as being equal to i∗ ke∗ made use of the definition of the vector c∗ γ since rank γ = k + 1). If γ) = i∗ n−1,n(d∗ γ) = i∗ k,n(d∗ rank (γ) = n > k + 1, P ∗ k,n−1(0) = 0 where the penultimate equality follows from the previous argument with k replaced by n − 1. Note we also made use of the following easily verified fact: if k < m < n then i∗ m,n. The proof for the case k = 1 is similar. k,n−1 ◦ i∗ k,n = i∗ k,m ◦ i∗ (0,k](d∗ γ) = i∗ We are finally ready to present the main result of this section. (cid:40) τ (γ) = Proposition 2.4.3. Let (∆q)∞ q=1, Γn and Γ be the sets defined as in the original Bourgain- Delbaen construction and let a, b be real numbers satisfying either of the two conditions assumed by Bourgain and Delbaen (see the previous section). We define τ on Γ \ Γ1 as follows. Given γ ∈ ∆n+1, write γ = (n + 1, l, ξ, η, ε, ε(cid:48)) and set if η ∈ Γl if η ∈ Γn \ Γl. (ε, a, ξ) (ε, a, ξ, l, b, ε(cid:48)e∗ η) (We clarify that the slightly clumsy notation ε(cid:48)e∗ interpreted as the scalar ε(cid:48) multiplied by the vector e∗ BD setups, a ≤ 1 and b < 1 projections (cid:91)P ∗ 2 , Theorem 2.2.2 yields that there exist(cid:99)d∗ η appearing in the final tuple above is to be η). Noting that in either of the original γ ∈ (cid:96)1(Γ) and γ −(cid:98)c∗ (0,q](cid:99)d∗ (cid:98)c∗ (cid:16)(cid:99)d∗ γ = e∗ (0,q] on (cid:96)1(Γ) uniquely determined by the following properties: (cid:40)(cid:99)d∗ 0 elements dγ of the(cid:99)d∗ With the notation that has been used throughout this section, we have (cid:98)c∗ γ generate a L∞,λ subpace of (cid:96)∞(Γ), X(Γ) := X(Γ, τ ). the (cid:99)d∗ (cid:16) (cid:17)∗ γ) coincide with the c∗ The family ous section). Moreover, the norm of each projection (cid:91)P ∗ γ ∈ ∆1 γ = (n + 1, l, ξ, η, ε, ε(cid:48)) and η ∈ Γl γ = (n + 1, l, ξ, η, ε, ε(cid:48)) and η ∈ Γn \ Γl. γ (and hence (0,k]. Moreover, the mappings : (cid:96)∞(Γk) → (cid:96)∞(Γ) coincide with the mappings jk of the original Bourgain-Delbaen γ (respectively d∗ γ) and (cid:91)P ∗ is a basis for (cid:96)1(Γ) with basis constant at most λ (as defined in the previ- (0,q] is at most λ and the biorthogonal ξ + ε(cid:48)b(I − (cid:91)P ∗ (0,l])e∗ γ ∈ Γq γ ∈ Γ \ Γq (0,k] = P ∗ εae∗ εae∗ ξ γ γ∈Γ (cid:91)P ∗ γ = γ = (cid:17) γ 0 η P ∗ construction. (0,k] Before giving the proof we make a few observations that follow from the proposition. Note we have the following important corollary. Corollary 2.4.4. The subspace X(Γ) of the above proposition is precisely the original Bourgain-Delbaen space Xa,b. 44 Proof. It follows from the proof of Theorem 2.2.2 and Lemma 2.2.3 that So, by the proposition, X(Γ) = X(Γ) = (cid:17)∗ . (cid:17)∗ (0,n] (cid:16) (cid:91)P ∗ (cid:16) P ∗ (0,n] im im (cid:91) n∈N (cid:91) (cid:91) n∈N im jn = n∈N = Xa,b. Thus we can think of the original Bourgain-Delbaen construction in terms of the Argyros- Haydon construction as we wanted. Moreover, we see from the generalised construction, i.e. by Theorem 2.2.2, that the biorthogonal elements dγ of the d∗ γ are a basis of X(Γ) = Xa,b. It follows by Lemma 2.2.3 and the proposition that for γ ∈ ∆n, dγ = jneγ. The fact this basis is normalised follows from Lemma 2.2.5. Thus we obtain all the results γ = (cid:99)d∗ stated in Remark 2.3.2 as immediate corollaries. γ for all It is γ. Thus by Remark 2.4.2 and γ. So (0,k]d∗ (0,k](cid:99)d∗ It remains for us to prove the proposition. Proof of Proposition 2.4.3. We prove by induction that the statement " (cid:99)d∗ γ, (cid:99)d∗ We recall that the vectors (cid:98)c∗ (0,k](cid:96)1(Γn)" holds for every n ∈ N. immediate from the definitions that when γ ∈ Γ1,(cid:99)d∗ γ and operators (cid:91)P ∗ γ = d∗ γ ∈ Γn and for all k ∈ N, P ∗ (0,k](cid:96)1(Γn) = (cid:91)P ∗ (0,k] are defined inductively. γ = e∗ γ = d∗ (0,k](cid:96)1(Γ1) all γ ∈ Γn and P ∗ (0,k](cid:96)1(Γ1) = (cid:91)P ∗ P ∗ the way in which the (cid:91)P ∗ ∀k ∈ N (0,k] are constructed, we have that for γ ∈ Γ1, P ∗ γ = (cid:91)P ∗ and the statement is true for n = 1. Suppose inductively that for some n ≥ 1,(cid:99)d∗ (0,k](cid:96)1(Γn) = (cid:91)P ∗ (0,k](cid:96)1(Γn). Then for γ = (n + 1, l, ξ, η, ε, ε(cid:48)) ∈ ∆n+1, (cid:40) εae∗ (cid:40) εae∗ εae∗ = εae∗ = c∗ γ. η ∈ Γl η ∈ Γn \ Γl η ∈ Γl η ∈ Γn \ Γl ξ + ε(cid:48)b(I − (cid:91)P ∗ ξ + ε(cid:48)b(I − P ∗ (0,l])e∗ η (0,l])e∗ η (cid:98)c∗ γ = ξ ξ The first of the above equalities holds due to the inductive construction of the (cid:98)c∗ γ, the second by the inductive hypothesis and the final equality is a consequence of the very definition of γ = d∗ γ for 45 γ − (cid:98)c∗ (0,k](cid:96)1(Γn+1) = (cid:91)P ∗ (0,k] and all our previous calculations. It follows that (cid:99)d∗ the P ∗ γ. We again (0,k](cid:96)1(Γn+1) for all k ∈ N as a consequence of Remark 2.4.2 and get that P ∗ the way in which the (cid:91)P ∗ (0,k] are constructed. The first part of the proof is thus complete (cid:17)∗ by induction and an elementary density argument. As a consequence we now drop the 'hat (cid:104)(cid:16) = jk : (cid:96)∞(Γk) → (cid:96)∞(Γ). For x ∈ (cid:96)∞(Γk), γ ∈ Γ It remains to prove γ = d∗ γ = e∗ notation'. (cid:17)∗ (cid:16) P ∗ (cid:105) (0,k] P ∗ (0,k] (x) (γ) = (cid:104)x, P ∗ γ)(cid:105) (0,k](e∗ where we are, of course, thinking of x as an element of (cid:96)1(Γk)∗ (under the canonical iden- tification) in the RHS of the equality. If rank (γ) ≤ k, P ∗ γ from which it follows that (γ) = x(γ) = [jk(x)](γ) as required. Obviously we must see that this γ = e∗ (0,k]e∗ (cid:104)(cid:16) (cid:17)∗ P ∗ (cid:105) (x) (0,k] equality also holds when rank γ > k. We will need the following lemma. Lemma 2.4.5. If x ∈ (cid:96)∞(Γn) (which we identify in the canonical way with (cid:96)1(Γn)∗) and γ ∈ Γ with rank (γ) = n + 1 then x(c∗ γ) = [jnx](γ). In order not to detract from the proof at hand, we defer the proof of this lemma to the end. Suppose now that rank (γ) = k + 1. Then P ∗ ke∗ γ = c∗ γ = i∗ γ. So (cid:104)(cid:16) (cid:17)∗ (cid:105) (0,k]e∗ (γ) = x(c∗ P ∗ (0,k] (x) γ) = [jkx](γ) where we recall x ∈ (cid:96)∞(Γk) and appeal to Lemma 2.4.5 to obtain the final equality. Finally, we assume rank (γ) = n > k + 1. We have (cid:104)(cid:16) (cid:17)∗ (cid:105) P ∗ (0,k] (x) (γ) = (cid:104)x, P ∗ = (cid:104)x, i∗ = (cid:104)x, i∗ = (cid:104)x ◦ i∗ = (cid:104)x ◦ i∗ γ)(cid:105) (0,k](e∗ γ)(cid:105) k,n(e∗ γ)(cid:105) n−1,n(e∗ k,n−1i∗ γ(cid:105) k,n−1, i∗ n−1,ne∗ γ(cid:105). k,n−1, c∗ We now think of x ◦ i∗ again to Lemma 2.4.5 we see k,n−1 ∈ (cid:96)1(Γn−1)∗ as an element of (cid:96)∞(Γn−1). Appealing once (cid:105) (cid:17)∗ (cid:104)(cid:16) P ∗ γ) = [jn−1(x ◦ i∗ So to complete the proof it is enough to see that [jn−1(x ◦ i∗ (γ) = x ◦ i∗ k,n−1(c∗ (x) (0,k] k,n−1)](γ) = [jk(x)](γ). But we recall from the original B.D construction that for any x ∈ (cid:96)∞(Γk), jk(x) = jn−1(ik,n−1x). k,n−1)](γ). 46 (We saw this in the proof of Lemma 2.3.1.) So in fact, all we need to see is that when ik,n−1x is thought of as an element of (cid:96)1(Γn−1)∗ under the canonical identification, it is precisely the element x ◦ i∗ k,n−1 ∈ (cid:96)1(Γn−1)∗. To this end, let y∗ be an element of (cid:96)1(Γn−1). So x ◦ i∗ k,n−1(y∗) = x ◦ i∗ k,n−1(e∗ θ)y∗(θ) (2.3) where we are, of course, still thinking of x as an element of (cid:96)1(Γk)∗. Really x ∈ (cid:96)∞(Γk), so we can compute x ◦ i∗ k,n−1(e∗ θ) = x(ν)[i∗ k,n−1(e∗ θ)](ν). (2.4) Fortunately, we computed earlier that for ν ∈ Γk, [i∗ k,n−1(e∗ θ)](ν) = e∗ θ(ν) + e∗ θ(ε)[ik,n−1eν](ε). (cid:88) θ∈Γn−1 (cid:88) ν∈Γk (cid:88) ε∈Γn−1\Γk If θ ∈ Γn−1 \ Γk then this is just [ik,n−1eν](θ), whilst for θ ∈ Γk, we obtain e∗ So by Equation (2.4) we find θ(ν) = eν(θ). x ◦ i∗ k,n−1(e∗ θ) = x(ν)eν(θ) = x(θ) x(ν)[ik,n−1eν](θ) = ik,n−1x(θ) if θ ∈ Γk if θ ∈ Γn−1 \ Γk. ν∈Γk ν∈Γk (cid:40)(cid:80) (cid:80) Substituting this back into Equation (2.3) we get (cid:88) (cid:88) θ∈Γk x ◦ i∗ k,n−1(y∗) = = (cid:88) x(θ)y∗(θ) + [ik,n−1x](θ)y∗(θ) θ∈Γn−1\Γk [ik,n−1x](θ)y∗(θ) θ∈Γn−1 = [ik,n−1x](y∗) where we think of ik,n−1x as an element of (cid:96)1(Γn−1)∗. The second equality above comes from observing that when θ ∈ Γk, [ik,n−1x](θ) = x(θ). Proof of Lemma 2.4.5. We use the same notation as that used for the statement of the lemma. We consider only the more complex case where γ = (n+1, l, ξ, η, ε, ε(cid:48)) and η ∈ Γn\Γl since the other case is similar (and easier). So, (cid:88) (cid:88) θ∈Γn θ∈Γn x(c∗ γ) = = x(θ)c∗ γ(θ) (cid:88) θ∈Γn x(θ)εae∗ ξ(θ) + x(θ)ε(cid:48)b[(I − P ∗ (0,l])e∗ η](θ) = εax(ξ) + x(θ)ε(cid:48)b[(I − P ∗ (0,l])e∗ η](θ) (cid:88) θ∈Γn 47 ξ(θ) can only be non-zero when θ = ξ and since ξ ∈ Γl, l < n, this happens precisely since e∗ once. If rank (θ) ≤ l, i.e. θ ∈ Γl, since we also assume η ∈ Γn \ Γl, the definition of P ∗ given earlier and Equation (2.2) yields l,n(e∗ (cid:40) where the final equality holds since it is assumed rank (θ) ≤ l. It is then easily seen that η)](θ) = [il,neθ](η) = [il,nπleθ](η) η](θ) = [i∗ [P ∗ (0,l]e∗ (0,l] [(I − P ∗ (0,l])e∗ η](θ) = e∗ η(θ) = eθ(η) [eθ − il,nπleθ](η) if rank (θ) > l if rank (θ) ≤ l However, it is clear that [eθ − il,nπleθ](η) = eθ(η) if rank (θ) > l. So in fact, we have shown [(I − P ∗ η](θ) = [eθ − il,nπleθ](η). Consequently, (cid:88) η](θ) = ε(cid:48)b x(θ)[(I − P ∗ (0,l])e∗ ε(cid:48)b (0,l])e∗ (cid:88) since clearly x =(cid:80) θ∈Γn x(θ)[eθ − il,nπleθ](η) θ∈Γn = ε(cid:48)b{(x − il,nπlx)(η)} θ∈Γn x(c∗ x(θ)eθ and all the maps are linear. So, we finally have γ) = εax(ξ) + ε(cid:48)b{(x − il,nπlx)(η)} = [jn(x)](γ) as claimed. 2.5 The operator algebras for the original Bourgain-Delbaen spaces We conclude this chapter by exhibiting an uncountable set of isometries on the original Bourgain-Delbaen spaces which are pairwise distance C apart (with respect to the usual operator norm) for some constant C > 0. In particular, this establishes that the operator algebra L(X) is non-separable for any of the original Bourgain-Delbaen spaces X. We take this opportunity to note that it is in fact easy to show K(X) is non-separable when X is a Bourgain-Delbaen space of class X , so certainly L(X) is non-separable. Indeed, it was shown in [7, Theorem 4.4] that in this case, every infinite dimensional subspace of X has a subspace isomorphic to (cid:96)1. In particular, (cid:96)1 embeds into X. Since the dual operator of an isomorphic embedding is a quotient operator, it follows that there is a quotient operator Q : X∗ → (cid:96)∞. Consequently, X∗ cannot be separable and therefore neither is K(X) since X∗ always embeds isometrically into K(X). The non-separability is interesting because, as commented by Beanland and Mitchell in [4], Emmanule showed that (cid:96)∞ does not embed into L(X) when X ∈ X , thus giving an example of a non-separable Banach space containing no isomorphic copy of (cid:96)∞. Similar 48 questions about the separability of L(Y ) when Y ∈ Y remained open however. This section addresses this problem; we remark that at the time of writing, it remains unknown if (cid:96)∞ embeds into L(Y ). 2.5.1 Construction of the basic operator For the remainder of this chapter, we fix real numbers a, b satisfying either of the conditions of Bourgain and Delbaen (see Section 2.3) and work with the Bourgain-Delbaen space X = Xa,b = X(Γ, τ ) (for some suitable τ mapping, as shown in the previous section). We begin by inductively constructing an isometry S∗ : (cid:96)1(Γ) → (cid:96)1(Γ). We will show that the dual of this operator restricts to give an isometry S : X → X. In the following section, we will then take appropriate compositions of operators of this type to obtain the uncountable set of isometries we seek. We continue with the notation used in the preceding section. To construct S∗, we fix a γ = (n + 1, k, ξ, η, ε, ε(cid:48)) ∈ ∆n+1 (where n ≥ 2) and aim to construct (inductively) a bijective mapping F : Γ → Γ with the following properties (1) F (γ) = (n + 1, k, ξ, η,−ε,−ε(cid:48)) =: γ The operator S∗ : (cid:96)1(Γ) → (cid:96)1(Γ) will map d∗ τ to d∗ F (τ ) for all τ /∈ {γ, γ} and d∗ τ to −d∗ (2) F (γ) = γ (3) F (τ ) = τ if rank (τ ) ≤ rank (γ) and τ /∈ {γ, γ} (4) rank (F (τ )) = rank (τ ) ∀ τ ∈ Γ. We shall say F is 'rank-preserving'. Note that we must assume n ≥ 2 in the above since we recall that ∆2 = ∅ and Γ1 = Γ2 is simply a singleton set. F (τ ) for τ ∈ {γ, γ}. (The reason for this sign change in the special case where τ ∈ {γ, γ} will shortly become apparent.) The important point here is that we know how S∗ acts on the d∗ τ vectors; it is this property that will ensure the dual map restricts to give a map from X into X. Of course, the property just described only defines a linear map on the dense subspace sp{d∗ τ : τ ∈ Γ} of (cid:96)1(Γ). We also need to show that this map is bounded, so that it can be extended to a bounded operator on (cid:96)1(Γ). Note that since we are working with the (cid:96)1 norm, in order to prove S∗ is bounded, we only need to be able to control (cid:107)S∗(e∗ τ )(cid:107). More precisely, if there is some M ≥ 0 such that (cid:107)S∗(e∗ τ )(cid:107) ≤ M for every τ ∈ Γ, then it is elementary to check that S∗ is bounded with norm at most M . 49 To begin the inductive constructions of S∗ and F we define F : Γn+1 → Γn+1 and S∗ : (cid:96)1(Γn+1) → (cid:96)1(Γn+1) by γ (cid:40) F (τ ) = S∗(d∗ τ ) = if τ = γ if τ = γ τ /∈ {γ, γ} γ τ d∗ −d∗ θ : θ ∈ Γn} = sp{d∗ if τ /∈ {γ, γ} if τ ∈ {γ, γ}. F (τ ) τ Recall that if rank τ ≤ n + 1, c∗ the linearity of S∗, it is easily seen that for all τ ∈ Γ with rank τ ≤ n + 1, S∗c∗ Therefore, S∗e∗ τ ∈ sp{e∗ τ = S∗(c∗ θ : θ ∈ Γn}. Consequently, using τ = c∗ τ . τ for all τ ∈ Γn+1 \ {γ, γ}. We also have γ − d∗ γ − d∗ γ + c∗ γ) γ + (c∗ γ τ ) + S∗(d∗ S∗(e∗ τ ) = c∗ γ) = S∗(c∗ τ + d∗ γ + d∗ τ = e∗ γ) = c∗ = −c∗ = −e∗ γ since if we look at the possible values of c∗ (We remark that this computation works due to the 'sign change', S∗d∗ earlier.) In a similar way we obtain S∗(e∗ summarised by the following formulae which are valid for all τ ∈ Γn+1. γ obtained in Section 2.4 we see c∗ γ = −d∗ γ + c∗ γ = 0. γ, noted γ. The above discussion can be succinctly γ) = −e∗ γ and c∗ (cid:40) (cid:40) S∗(e∗ τ ) = e∗ τ = e∗ −e∗ F (τ ) F (τ ) if τ /∈ {γ, γ} if τ ∈ {γ, γ}. Since also c∗ τ = e∗ τ − d∗ τ for any τ ∈ Γ it is easily checked that c∗ τ = c∗ τ = −c∗ c∗ (We once again made use of the fact that c∗ S∗(c∗ τ ) = F (τ ) F (τ ) γ + c∗ if τ /∈ {γ, γ} if τ ∈ {γ, γ}. γ = 0 in the above formula). It is now clear how the inductive construction proceeds. We assume inductively that for some k ∈ N we have defined F : Γn+k → Γn+k and a linear map S∗ : (cid:96)1(Γn+k) → (cid:96)1(Γn+k) satisfying: (1) F : Γn+k → Γn+k is a bijection, satisfying the four desired properties of F given at the beginning of this section. (2) For each τ ∈ Γn+k, (i) S∗(e∗ τ ) = (cid:40) e∗ −e∗ F (τ ) F (τ ) 50 if τ /∈ {γ, γ} if τ ∈ {γ, γ} (ii) (iii) S∗(d∗ τ ) = S∗(c∗ τ ) = (cid:40) (cid:40) d∗ −d∗ F (τ ) F (τ ) c∗ −c∗ F (τ ) F (τ ) if τ /∈ {γ, γ} if τ ∈ {γ, γ} if τ /∈ {γ, γ} if τ ∈ {γ, γ}. noting that the previous discussion gives us the case k = 1. We will show that we can extend the maps F and S∗ to F : Γn+k+1 → Γn+k+1 and S∗ : (cid:96)1(Γn+k+1) → (cid:96)1(Γn+k+1) such that the above properties are still satisfied. Let τ = (n + k + 1, l, θ, ϕ, , (cid:48)). Recalling that c∗ In fact, by Proposition 2.4.3, τ ) is already defined by the inductive hypothesis. τ ∈ (cid:96)1(Γn+k), S∗(c∗ (cid:40)aS∗(e∗ θ) aS∗(e∗ θ) + (cid:48)bS∗(cid:16) S∗(c∗ τ ) = (cid:17) if ϕ ∈ Γl if ϕ ∈ Γn+k \ Γl. (I − P ∗ (0,l])e∗ ϕ ϕ) are given by the inductive hypothesis and property (2i) of S∗. The ϕ) where ϕ ∈ Γn+k \ Γl. We observe that since (0,l]e∗ θ) and S∗(e∗ S∗(e∗ slightly harder term to deal with is S∗(P ∗ ϕ ∈ (cid:96)1(Γn+k) we can write e∗ e∗ ϕ = (cid:88) By the inductive hypothesis, we get ω∈Γn+k\{γ,γ} (cid:88) αωd∗ ω + αγd∗ γ + αγd∗ γ S∗(e∗ ϕ) = αωd∗ F (ω) − αγd∗ γ − αγd∗ γ ω∈Γn+k\{γ,γ} and, using the fact that F is rank preserving, we find that P ∗ (0,l](S∗(e∗ ϕ)) = In either case, it is now easily seen by a similar computation that P ∗ From this computation and property 2(i) of the inductive hypothesis, we find that (0,l]e∗ ϕ). F (ω) − αγd∗ γ − αγd∗ γ if l < n + 1 if n + 1 ≤ l < n + k. ϕ)) = S∗(P ∗ (0,l](S∗(e∗ F (ω) αωd∗ (cid:40)(cid:80) (cid:80) ω∈Γl ω∈Γl\{γ,γ} αωd∗ (cid:40) S∗P ∗ (0,l]e∗ ϕ = P ∗ (0,l](e∗ F (ϕ)) (0,l](−e∗ P ∗ F (ϕ)) if ϕ /∈ {γ, γ} if ϕ ∈ {γ, γ}. Note that since θ ∈ Γl, it not possible to have ϕ ∈ Γn+k \ Γl and both θ and ϕ in {γ, γ}. So when ϕ ∈ Γn+k \ Γl, it is apparent from our computations that there are precisely 3 51 possibilities to consider when computing S∗(c∗ (0,l])e∗ S∗(c∗ ϕ τ ) = S∗(cid:16)  = (cid:110) θ + (cid:48)b(I − P ∗ ae∗ (cid:110) F (θ) + (cid:48)b (cid:110) F (θ) + (cid:48)b F (θ) + (cid:48)b (I − P ∗ (I − P ∗ (I − P ∗ ae∗ −ae∗ ae∗ In the easier case where ϕ ∈ Γl, we have τ ) = aS∗(e∗ S∗(c∗ θ) = τ ). We obtain (cid:111) (cid:17) (0,l])(e∗ F (ϕ)) (0,l])(e∗ (0,l])(−e∗ (cid:40) ae∗ −ae∗ F (θ) F (θ) F (ϕ)) F (ϕ)) (cid:111) (cid:111) θ, ϕ /∈ {γ, γ} θ ∈ {γ, γ}, ϕ /∈ {γ, γ} θ /∈ {γ, γ}, φ ∈ {γ, γ}. if θ /∈ {γ, γ} if θ /∈ {γ, γ}.  Finally, we note that since F is rank-preserving, all of the above formulae are equal to Ξ for some suitably chosen Ξ ∈ ∆n+k+1. We can therefore define F on ∆n+k+1 so that c∗ S∗(c∗ F (τ ). Our above calculations show that in order to achieve this, the correct definition of F is F ((n + k + 1, l, θ, ϕ, , (cid:48))) = τ ) = c∗ (n + k + 1, l, F (θ), ϕ, , (cid:48)) (n + k + 1, l, F (θ), ϕ,−, (cid:48)) (n + k + 1, l, F (θ), F (ϕ), , (cid:48)) (n + k + 1, l, F (θ), F (ϕ),−, (cid:48)) (n + k + 1, l, F (θ), F (ϕ), ,−(cid:48)) if ϕ ∈ Γl and θ /∈ {γ, γ} if ϕ ∈ Γl and θ ∈ {γ, γ} if ϕ ∈ Γn+k \ Γl and θ, ϕ /∈ {γ, γ} if ϕ ∈ Γn+k \ Γl and θ ∈ {γ, γ}, ϕ /∈ {γ, γ} if ϕ ∈ Γn+k \ Γl and θ /∈ {γ, γ}, ϕ ∈ {γ, γ}. F : Γn+k+1 → Γn+k+1 is easily seen to be a bijection. We now extend the definition of S∗ by defining S∗(d∗ for any τ ∈ ∆n+k+1. Thus we have extended F and S∗ and both maps have all the desired properties that were stated earlier. F (τ ) for τ ∈ ∆n+k+1. Since e∗ τ it is clear that S∗(e∗ τ ) := d∗ τ ) = e∗ τ = c∗ τ + d∗ By induction, we have thus succeeded in defining a linear map S∗ : c00(Γ) → (cid:96)1(Γ). It F (τ ) for all τ ∈ Γ (with a - sign only when is bounded since, by construction, S∗(e∗ τ ∈ {γ, γ}). So (cid:107)S∗(e∗ τ )(cid:107) = 1 for all τ ∈ Γ and therefore S∗ is bounded by our earlier remark. It follows that S∗ extends to a bounded linear map on (cid:96)1(Γ). In fact, S∗ is an F (τ ) for all τ ∈ Γ, S∗ isometry onto (cid:96)1(Γ). Indeed, since F is a bijection, and S∗(e∗ τ ∈ (cid:96)1(Γ), then τ ) = ±e∗ τ ) = ±e∗ F (τ ) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) aτ e∗ F (τ ) − aγe∗ γ − aγe∗ γ aτ + − aγ + − aγ aτ e∗ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)S∗ (cid:32)(cid:88) is certainly onto. Moreover, suppose(cid:80) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:33)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) = τ∈Γ aτ e∗ (cid:88) (cid:88) (cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:88) τ∈Γ\{γ,γ} aτ τ∈Γ τ∈Γ = = = τ aτ e∗ τ τ∈Γ\{γ,γ} τ∈Γ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 52 where we have repeatedly made use of the fact that F : Γ → Γ is a bijection, so e∗ F (τ(cid:48)) when τ (cid:54)= τ(cid:48). It follows that the dual operator (S∗)∗ : (cid:96)1(Γ)∗ → (cid:96)1(Γ)∗ is also an isometry. For convenience, we drop the brackets and just write S∗∗ for (S∗)∗. F (τ ) (cid:54)= e∗ Recall that dτ ∈ (cid:96)1(Γ)∗ denotes the usual biorthogonal vector (as in Section 2.3) of the basis (d∗ θ)θ∈Γ of (cid:96)1(Γ). Using the fact that d∗ −d∗ S∗(d∗ τ ) = F (τ ) F (τ ) (cid:40) (cid:40) it is elementary to check that S∗∗(dω) = dF −1(ω) −d∗ F −1(ω) if τ /∈ {γ, γ} if τ ∈ {γ, γ} if ω /∈ {γ, γ} if ω ∈ {γ, γ} Recalling from Section 2.3 that the Bourgain-Delbaen space is precisely X = [dγ : γ ∈ Γ], we see that S∗∗ restricts to give an onto isometry S := S∗∗ : X → X. We shall call S a basic operator. By the previous observation, we get Lemma 2.5.1. Let(cid:80) ω αωdω ∈ X. Then S where (cid:32)(cid:88) ω αω = αωdω (cid:40) αω −αω (cid:33) (cid:88) ω = αωdF −1(ω) if ω /∈ {γ, γ} if ω ∈ {γ, γ} Of course, the definition of S∗, and consequently S, depends on the initial choice of γ ∈ ∆n+1. Moreover, the construction of S∗ is very much dependent on the bijection F : Γ → Γ, which also depends on our initial choice of γ ∈ ∆n+1. We define some terminology that will be useful later. Definition 2.5.2. Suppose S is constructed as above by fixing γ ∈ ∆n+1. We shall say that S has rank n and that γ is the base of S. We write base (S) = γ and rank (S) = n. We shall say the bijection F : Γ → Γ, used in the construction and definition of S∗, is the underlying function on Γ. We will usually denote a rank n basic operator by Sn and its underlying function on Γ by Fn. Lemma 2.5.3. Let m < n and suppose Sm, Sn : X → X are basic operators of ranks θ)θ∈Γ of (cid:96)1(Γ) such that m and n respectively. X = [dθ : θ ∈ Γ], then If λ is the basis constant of the basis (d∗ (cid:107)Sn − Sm(cid:107) ≥ 1 2λ 53 Proof. Let γ = base (Sm) (so in particular rank (γ) = m + 1) and let Fn : Γ → Γ be the the underlying function on Γ for Sn. So by Lemma 2.5.1, Sn(dγ) = dF −1 n (γ) = dγ since FnΓn = IdΓn, and Sm(dγ) = −dγ. We recall that (cid:107)dτ(cid:107) = 1 for all τ ∈ Γ (see Lemma 2.2.5 and Remark 2.3). (This was also proved using a different method in [17].) Using this fact, we can estimate as follows. (cid:107)Sn − Sm(cid:107) ≥ (cid:107)Sn(dγ) − Sm(dγ)(cid:107) = (cid:107)dγ + dγ(cid:107) ≥ dγ((cid:107)d∗ = (cid:107)d∗ γ(cid:107)−1d∗ γ(cid:107)−1 ≥ 1 2λ γ(cid:107)−1d∗ γ) γ) + dγ((cid:107)d∗ . To obtain the final inequality, we have made use of the fact that for all τ ∈ Γ, (cid:107)d∗ (cid:107)e∗ τ(cid:107) ≤ 1 + λ ≤ 2λ. τ − P ∗ (0,rankτ )e∗ τ(cid:107) = We therefore have at least a countably infinite number of operators on X, all of which are pairwise distance at least 1 apart. To establish the non-separability of L(X) we need to work a bit harder. 2.5.2 Non-separability of L(X) In the final section of this chapter, we establish the non-separability of L(X). Throughout this section, λ is the same constant as appearing in the previous lemma. The idea is to construct an uncountable set of operators on X, all of which are pairwise distance 1/2λ apart by taking suitable compositions of basic operators of increasing ranks. For each n ∈ N, n ≥ 2, we fix a γn ∈ ∆n+1 and let Sn be the basic operator (as constructed in Section 2.5.1) with base (Sn) = γn. We work with these operators for the remainder of the section. The underlying function on Γ for Sn will be denoted by Fn. We recall that each n : (cid:96)1(Γ) → (cid:96)1(Γ) and restricting it to Sn is obtained by taking the dual of some operator S∗ X. Before progressing any further, we need Proposition 2.5.4. Let (nj)∞ n1 ≥ 2. For p ≤ m, (p, m ∈ N) denote by T ∗ nm ◦ ··· ◦ S∗ S∗ Proof. It is enough to show (T ∗ np+1 ◦ S∗ 1,nx∗)∞ j=1 be a strictly increasing sequence of natural numbers with p,m := p,m the composition of basic operators T ∗ np. For any x∗ ∈ (cid:96)1(Γ), the sequence (T ∗ 1,nx∗)∞ n=1 converges in (cid:96)1(Γ). n=1 is a Cauchy sequence. We note that for k > l, (cid:13)(cid:13)T ∗ 1,lx∗(cid:13)(cid:13) =(cid:13)(cid:13)T ∗ 1,kx∗ − T ∗ 1,lx∗(cid:13)(cid:13) l+1,k(T ∗ 1,lx∗) − T ∗ 54 We recall that for any p ∈ N, S∗ is strictly increasing, it is easily seen that T ∗ (Schauder) basis for (cid:96)1(Γ), we can write T ∗ 1,lx∗ = So (cid:13)(cid:13)T ∗ 1,lx∗(cid:13)(cid:13) = 1,kx∗ − T ∗ p(cid:96)1(Γp) = Id(cid:96)1(Γp). Therefore, since the sequence (nj)∞ j=1 ω)ω∈Γ is a ω ω∈Γ ω d∗ α(l) l+1,k(cid:96)1(Γnl+1 ) = Id(cid:96)1(Γnl+1 ). Since (d∗ (cid:88) (cid:17) −(cid:88) (cid:16) (cid:88) (cid:13)(cid:13)(cid:13) (cid:17) −(cid:88) ω∈Γ\Γnl+1 ω d∗ α(l) ω d∗ α(l) (cid:13)(cid:13)(cid:13) ω d∗ α(l) ω∈Γ ω∈Γ l+1,k ω ω ω (cid:13)(cid:13)(cid:13) l+1,k (cid:13)(cid:13)(cid:13)T ∗ (cid:16)(cid:88) (cid:13)(cid:13)(cid:13) (cid:88) (cid:13)(cid:13)(cid:13) (cid:88) ω∈Γnl+1 ω∈Γ\Γnl+1 = ≤ 2 ω ω d∗ α(l) ω + T ∗ ω∈Γ ω d∗ α(l) ω d∗ α(l) ω 1,l − P ∗ (0,nl+1]T ∗ where the final inequality is obtained from the triangle inequality and the fact that T ∗ l+1,k is a composition of isometries, therefore has norm 1. To prove the proposition, it is enough to show (cid:107)(T ∗ To see this, we consider z∗ = (cid:80) 1,l)x∗(cid:107) → 0 as l → ∞. ω∈Γ δωd∗ ω to be any vector in (cid:96)1(Γ). For any j, m ∈ N, since Fm preserves the rank of all elements in Γ, an easy computation yields that P ∗ mz∗ = S∗ (0,j]S∗ m operators, the same equality l,k. for any l, k ∈ N. Using this observation, it is now easy to holds with S∗ see that (cid:107)(T ∗ (0,j]z∗. Consequently, as a composition of S∗ 1,l)x∗(cid:107) → 0 as l → ∞. Indeed, 1,l)x∗(cid:107) = (cid:107)T ∗ 1,l(I − P ∗ (0,nl+1])x∗(cid:107) = (cid:107)(I − P ∗ (0,nl+1])x∗(cid:107). mP ∗ m replaced by T ∗ 1,l − P ∗ 1,l − P ∗ (cid:107)(T ∗ (0,nl+1]T ∗ (0,nl+1]T ∗ We have made use of the fact that the operators T ∗ 1,l are isometries on (cid:96)1(Γ) to obtain the final equality. The right-hand-side of the above expression clearly converges to 0 as l → ∞ since the sequence (nl)∞ l=1 → ∞ and the P ∗ (0,nl+1] operators are basis-projections. Using the above observation, it is easy to prove the following corollary: Corollary 2.5.5. Let (nj)∞ n1 ≥ 2, and let T ∗ linear operator T ∗ j=1 be a strictly increasing sequence of natural numbers with 1,n (n ∈ N) be defined as in Proposition 2.5.4. Then we may well-define a (nj )∞ j=1 : (cid:96)1(Γ) → (cid:96)1(Γ) by T ∗ (x∗) = lim (nj )∞ j=1 j→∞ T ∗ 1,j(x∗) (x∗ ∈ (cid:96)1(Γ)) Moreover, T ∗ (nj )∞ j=1 : (cid:96)1(Γ) → (cid:96)1(Γ) is an onto isometry. We are finally ready to prove the main result of this chapter. 55 Theorem 2.5.6. Given a strictly increasing sequence of natural numbers (nj)∞ 2, the dual operator of the operator T ∗ X. Moreover, if (nj)∞ j=1 with n1 ≥ : X → (nj )∞ j=1 are two different strictly increasing sequences, with restricts to give an onto isometry, T(nj )∞ j=1 j=1 j=1 and ((cid:102)nj)∞ nk =(cid:102)nk for k = 1, 2 and n1 ≥ 2 then(cid:13)(cid:13)(cid:13)T(nj )∞ j=1 (cid:13)(cid:13)(cid:13) ≥ 1 2λ − T((cid:102)nj )∞ j=1 Since the set of all strictly increasing sequences of natural numbers with the first two terms being fixed is uncountable, it follows that L(X) is non-separable. Proof. The dual operator of T ∗ is an onto isometry of (cid:96)∞(Γ) by standard duality arguments. So for the first part of the proof, we must only see that it maps X into (and onto) X. We fix a θ ∈ Γ and show (T ∗ (nj ))∗(dθ)](d∗ (nj ))∗(dθ) ∈ [dγ : γ ∈ Γ]. For τ ∈ Γ, we have (cid:0)T ∗ (nj )(d∗ τ ) = dθ [(T ∗ (nj )∞ j=1 τ )(cid:1) (cid:0)T ∗ (cid:0)S∗ τ )(cid:1) 1,j(d∗ nj ◦ ··· ◦ S∗ n1(d∗ τ )(cid:1) = lim j→∞ dθ j→∞ dθ = lim We let m be mininal such that τ ∈ Γnm (i.e. τ /∈ Γnm−1, so in particular rank (τ ) > nm−1) nj ◦···◦ S∗ and consider 3 possibilities: n1(d∗ (1) m = 1. In this case S∗ (2) m = 2. Now we have that ∀k ≥ 1, S∗ n1d∗ τ ) Fn1 (τ )) (cid:40)dθ(d∗ (nj ))∗(dθ)](d∗ τ ) = dθ(S∗ [(T ∗ τ ) = d∗ nk = −dθ(d∗ Fn1 (τ )) τ so it follows that [(T ∗ ◦ ··· ◦ S∗ n1(d∗ τ ) = S∗ n1(d∗ (nj ))∗(dθ)](d∗ τ ) = dθ(d∗ τ ). τ ). It follows that if τ (cid:54)= base S∗ otherwise. n1 or Fn1(base S∗ n1) (3) m > 2. It is seen that ∀k ≥ m − 1, S∗ (nj ))∗(dθ)](d∗ [(T ∗ nk τ ) = dθ (cid:16) ◦ ··· ◦ S∗ n1(d∗ τ ) = S∗ nm−1 ◦ ··· ◦ S∗ S∗ nm−1 ◦ ··· ◦ S∗ n1(d∗ τ ) . n1(d∗ τ ). So τ ) = d∗ Since rank (τ ) > nm−1 > nm−2 ··· > n1 and we are assuming m > 2, S∗ S∗ n1(d∗ We consider two sub-cases: (a) Either base (S∗ nm−1) = Fnm−2 ◦ ··· ◦ Fn1(τ ) or Fnm−1 Fnm−2◦···◦Fn1 (τ ) and so dθ nm−1◦···◦S∗ nm−1(d∗ n1(d∗ nm−1 nm−2 ◦ ··· ◦ Fnm−2◦···◦Fn1 (τ ))(cid:1). (cid:1) = Fnm−2 ◦ ··· ◦ (cid:0)S∗ (cid:17) (cid:0)S∗ τ )(cid:1) = dθ (cid:0)base S∗ Fn1(τ ). We note that in particular this implies rank (τ ) = nm−1 + 1. In this case, nm−1(d∗ (nj ))∗(dθ)](d∗ S∗ −dθ(d∗ Fnm−1◦Fnm−2◦···◦Fn1 (τ ). It follows that [(T ∗ Fnm−2◦···◦Fn1 (τ )) = −d∗ Fnm−1◦Fnm−2◦···◦Fn1 (τ )). τ ) = 56 (b) Both base Sn∗ m−1 and Fnm−1(base S∗ nm−1) are not equal to Fnm−2◦···◦Fn1(τ ). An ar- gument similar to the previous case shows [(T ∗ (nj ))∗(dθ)](d∗ We note that in all possible cases, [(T ∗ ω ∈ (cid:96)1(Γ) we have given x∗ =(cid:80) ω∈Γ αωd∗ τ ) = dθ(d∗ (nj ))∗(dθ)](d∗ τ ) (cid:54)= 0 only if rank (θ) = rank (τ ). So Fnm−1◦Fnm−2◦···◦Fn1 (τ )) [(T ∗ (nj ))∗(dθ)](x∗) = [(T ∗ (nj ))∗(dθ)](cid:0)(cid:88) ω∈Γ αωd∗ ω (cid:1) = = (cid:88) (cid:88) ω∈Γ rank (ω)=rank (θ) ω∈Γ rank (ω)=rank (θ) αω[(T ∗ (nj ))∗(dθ)](d∗ ω) λωdω(x∗) (nj ))∗(dθ)](d∗ (nj ))∗(dθ) ∈ X as it is a finite linear combination of the dγ. Thus (T ∗ ω) and we have made use of the obvious fact that αω = dω(x∗). (nj ))∗ where λω := [(T ∗ It follows that (T ∗ does indeed restrict to give an isometry from X into X. One can do similar computations to verify that (T ∗ in the theorem, is an onto isometry as required. (nj ))∗ maps X onto X. It follows that the operator T(nj ) : X → X, defined the same first two terms and such that n1 ≥ 2. We choose k minimal s.t. nk (cid:54)= (cid:102)nk and w.lo.g. assume nk >(cid:102)nk. We note that by the minimality of k we have nj =(cid:102)nj whenever j=1 are two strictly increasing sequences which have j=1 and ((cid:102)nj)∞ We suppose now (nj)∞ j < k and that by assumptions on the sequences, k > 2. We note that for any θ, τ ∈ Γ (cid:13)(cid:13)(cid:13)T(nj ) − T((cid:102)nj ) (cid:13)(cid:13)(cid:13) ≥(cid:12)(cid:12)(cid:12)[T(nj )(dθ)]((cid:107)d∗ τ(cid:107)−1d∗ τ ) − [T((cid:102)nj )(dθ)]((cid:107)d∗ τ(cid:107)−1d∗ τ ) (cid:12)(cid:12)(cid:12). We will find specific θ, τ ∈ Γ such that the right hand side of the above inequality is equal to (cid:107)d∗ τ(cid:107)−1, which will complete the proof. Our previous calculations show that [T(nj )(dθ)](d∗ τ ) = dθ(±d∗ Fnm−1◦···◦Fn1 (τ )) where m ∈ N is minimal such that τ ∈ Γnm (and it is assumed that τ is chosen s.t. τ is chosen in Γ \ Γn2). Moreover we have a minus sign precisely when m > 2, i.e. base Snm−1 = Fnm−2 ◦ ··· ◦ Fn1(τ ) or Fnm−1 (cid:1) = Fnm−2 ◦ ··· ◦ Fn1(τ ). Similarly, (cid:0)base Snm−1 [T((cid:102)nj )(dθ)](d∗ τ ) = dθ(±d∗ ◦···◦F(cid:102)n1 (τ )) F(cid:94)np−1 (cid:1) = F(cid:94)np−2 2). Moreover we have a minus sign precisely when base S(cid:94)np−1 F(cid:94)np−1 where p ∈ N is minimal such that τ ∈ Γ(cid:102)np (and it is assumed that τ is chosen s.t. p > (cid:0)base S(cid:94)np−1 ◦ ··· ◦ F(cid:102)n1(τ ) or We will in fact choose τ such that rank (τ ) =(cid:102)nk + 1. It is easily seen that with a choice above must in fact be equal to k since(cid:102)nk < nk =⇒ nk ≥(cid:102)nk + 1 and thus τ ∈ Γ(cid:102)nk+1 ⊆ Γnk , of τ like this, the 'p' above must be equal to k + 1. On the other hand, the choice of m ◦ ··· ◦ F(cid:102)n1(τ ). = F(cid:94)np−2 57 whilst τ /∈ Γnk−1 = Γ(cid:94)nk−1 . Since we have observed k > 2 our above calculations remain valid, and we find that for such a choice of τ we have [T((cid:102)nj )(dθ)](d∗ [T(nj )(dθ)](d∗ τ ) = dθ(±d∗ F(cid:102)nk τ ) = dθ(±d∗ Fnk−1◦···◦Fn1 (τ )) ◦···◦F(cid:102)n1 (τ )) (cid:0) − d∗ where we have a minus sign precisely in the cases described above. Since all the Fm's are rank preserving, a perfectly good choice of τ fulfulling the condition that rank (τ ) =(cid:102)nk + 1 is τ = F −1(cid:102)n1 (base S(cid:102)nk ). If we now choose θ = F(cid:102)nk (base S(cid:102)nk ) we see that [T((cid:102)nj )(dθ)](d∗ [T(nj )(dθ)](d∗ τ ) = dθ τ ) = dθ(±d∗ (cid:1) = −1 ◦ ··· ◦ F −1 (cid:94)nk−1 where to establish the final equalities we made use of the facts that(cid:102)nj = nj for all j < k and θ = F(cid:102)nk (base S(cid:102)nk ) (cid:54)= base S(cid:102)nk by the very construction of F(cid:102)nk . So, for these choices of θ, τ we finally get(cid:13)(cid:13)(cid:13)T(nj ) − T((cid:102)nj ) (cid:13)(cid:13)(cid:13) ≥(cid:12)(cid:12)(cid:12)[T(nj )(dθ)]((cid:107)d∗ τ ) − [T((cid:102)nj )(dθ)]((cid:107)d∗ τ(cid:107)−1d∗ τ ) τ(cid:107)−1d∗ F(cid:102)nk base S(cid:102)nk (cid:12)(cid:12)(cid:12) (base S(cid:102)nk ) ) = 0 τ(cid:107)−1 = (cid:107)d∗ ≥ 1 2λ as required. We conclude this chapter by observing that we can use the non-separability result just obtained to provide a short proof that no Bourgain-Delbaen space of class Y can have the scalar-plus-compact property. Corollary 2.5.7. There exists no X ∈ Y for which every bounded operator on X is a scalar multiple of the Identity plus a compact operator. Proof. It was shown in [7] that for any Bourgain-Delbaen space X ∈ Y, X∗ is isomorphic to (cid:96)1. Moreover, we have seen that all spaces of Bourgain-Delbaen type have a Schauder basis, and consequently have the approximation property. For any Banach space Y with separable dual and approximation property, the space of compact operators on Y , K(Y ) is separable, and thus so also is the set of all operators of the form λIY +K (λ ∈ R, K ∈ K(Y )). Therefore if X ∈ Y had the scalar-plus-compact property, its operator algebra would be separable. This contradicts Theorem 2.5.6. 58 Whilst our non-separability result allows us to give a nice proof that the spaces in Y do not have the scalar-plus-compact property, we remark that the ideas in this section can be suitably modified to produce much stronger results. Indeed, one can use similar ideas to construct non-trivial projections on any of the Bourgain-Delbaen spaces X in X ∪ Y. (A non-trivial projection is one which has both infinite dimensional kernel and image.) It follows that X does not have the property that every bounded linear operator on X is a strictly singular perturbation of a scalar operator. Obviously this implies that the previous corollary holds for any of the original Bourgain-Delbaen spaces, not just those of class Y. 59 Chapter 3 Spaces with few but not very few operators 3.1 The Main Theorem Having looked in detail at the Bourgain-Delbaen construction in the previous chapter, we move on to consider a problem posed by Argyros and Haydon in [2]. We begin by introducing the following definition. Definition 3.1.1. Let X be a Banach space. We will say X has few operators if every operator from X to itself is a strictly singular perturbation of the identity, that is, expressible as λI + S for some strictly singular operator S : X → X. We say X has very few operators if every operator from X to itself is a compact perturbation of the identity. The fact that Banach spaces with few operators exist was first established in 1993 by Gowers and Maurey in [16]. The existence of a Banach space with very few operators was shown much more recently in [2]. In addition to their obvious intrinsic interest, such Banach spaces exhibit remarkable Banach space structure, as discussed in the introduction to this thesis. The motivation behind this chapter comes from a question that naturally arises from the work of Argyros and Haydon in [2]. We recall once again that the Banach space XAH, constructed by Argyros and Haydon in [2], is done so using the generalised Bourgain-Delbaen construction, as discussed in the previous chapter. Consequently, it has a Schauder basis and is a separable L∞ space. It was seen in [2] that the control over the finite dimensional subspace structure provided by the L∞ property was essential in proving that XAH has very few operators and we will use the same type of argument later in this chapter. Moreover, one can exploit the L∞ structure to show that the dual space of XAH is (cid:96)1. In light of the previous discussion, it is natural to conjecture that the ideals of strictly singular and compact operators on a separable L∞ space with (cid:96)1 dual coincide. Clearly this 60 must be true for the space XAH. Moreover, if one considers the space c0, the most obvious example of a separable L∞ space with (cid:96)1 dual, it is once again found that the conjecture holds. Indeed, it is well known (see, e.g. [1, Theorem 2.4.10]) that a bounded linear operator defined on c0 is compact ⇐⇒ it is weakly compact ⇐⇒ it is strictly singular. More generally, for spaces X with (cid:96)1 dual, the weakly compact and norm compact operators from X to itself must coincide as a consequence of (cid:96)1 having the Schur property; this provides further hope that the conjecture holds if we demand (cid:96)1 duality. On the other hand, if we instead look at operators defined on X, a separable L∞ space with (cid:96)1 dual, but mapping into a different target space, it is certainly possible for us to construct strictly singular operators that are not weakly compact (and hence not compact). To see this we need the following lemma. Lemma 3.1.2. Let X be a separable Banach space and suppose (cid:96)1 embeds isomorphically into X∗. Then there is a quotient operator Q : X → c0. Proof. Since (cid:96)1 embeds into X∗, we can find a sequence (y∗ n) in the unit ball, BX∗, of X∗ equivalent to the canonical basis (en) of (cid:96)1. Since X is separable, the weak* topology restricted to the BX∗ is metrizable. It follows from this (and weak∗ compactness of BX∗) that we may assume (passing to a subsequence of the (y∗ n) if necessary) that the sequence 2n−1 is weak∗ n) is weak∗ convergent. Now the sequence (x∗ (y∗ n) ∼ (en) ⊆ (cid:96)1 (recall every seminormalized block basic sequence of null and moreover, (x∗ (cid:96)1 is equivalent to the canonical basis of (cid:96)1, see, e.g. [1, Lemma 2.1.1, Remark 2.1.2]). Let 0 ≡ (cid:96)1 → X∗ be the isomorphic embedding which maps the basis vector en of (cid:96)1 to T : c∗ n. We claim that T is weak∗-weak∗ continuous. Indeed, suppose ((aβ x∗ n=1)β∈I is a weak* (cid:80)∞ nξn → convergent net in c∗ n=1 aβ n=1 αnξn. Now, for any x ∈ X, the sequence (x∗ nx) ∈ c0 since the sequence (x∗ n) is weak∗ ∞(cid:88) n=1 ∈ c0, (cid:80)∞ n) = (αn), i.e. whenever (ξn)∞ n) ⊆ X∗ where x∗ 0 with limβ(aβ null. It follows that n := y∗ 2n − y∗ nx = T(cid:0)(αn)(cid:1)x αnx∗ n)∞ ∞(cid:88) T(cid:0)(aβ n)(cid:1)x = T(cid:0)(αn)(cid:1) as required. n=1 nx∗ aβ nx → n=1 i.e. T(cid:0)(aβ n)(cid:1) (cid:42)w∗ 0 → X∗ is the dual of some operator Q : X → c0. Moreover, by Lemma 1.4.11, Q is a quotient operator. It follows by Lemma 1.4.12 that T : c∗ Corollary 3.1.3. Suppose X is an L∞ space of Bourgain-Delbaen type, i.e. obtained from one of the constructions discussed in the previous chapter. If X has no subspace isomorphic to c0 then there exists a strictly singular operator Q : X → c0 which fails to be weakly compact. 61 Proof. We have seen in the previous chapter that all the L∞ spaces X of Bourgain-Delbaen type are separable (they have a Schauder basis) and moreover, that (cid:96)1 embeds into the dual space. Consequently, by the above lemma, there is a quotient operator Q : X → c0. Suppose for contradiction that Q is weakly compact. Since Q is a quotient operator, it X ) ⊆ so that Bc0 is weakly compact (it is a ). This contradicts the fact that follows from the Open Mapping Theorem, that there is an M such that B◦ M Q(BX ). Therefore, Bc0 = B◦ weakly closed subset of the weakly compact set M Q(BX ) c0 is not reflexive, so Q cannot be weakly compact. c0 ⊆ M Q(B◦ w w ⊆ M Q(BX ) w c0 Since X has no subspace isomorphic to c0 it follows that Q is strictly singular. To see this, suppose there is an infinite dimensional subspace Y on which Q is an isomorphism. Then, by [1, Proposition 2.2.1] the subspace Q(Y ) contains a subspace Z isomorphic to c0. The image (QY →Q(Y ))−1(Z) is then a subspace of X isomorphic to c0, giving us a contradiction. We remark that there do exist L∞ spaces of Bourgain-Delbaen type which have no subspace isomorphic to c0 and therefore satisfy the hypotheses of the previous corollary. Indeed, it is immediate from the HI property that the Arygros-Haydon space is one such example. In fact, it was shown in [7], that if X ∈ X , Y ∈ Y, where X ,Y are the original classes of Bourgain-Delbaen spaces, then every infinite dimensional subspace of X contains a subspace isomorphic to (cid:96)1 and every infinite dimensional subspace of Y contains an infinite dimensional reflexive subspace. It follows from these facts and elementary results in Banach space theory that the original Bourgain-Delbaen spaces also have no subspace isomorphic to c0. Of course, the previous corollary does not give us a counterexample to the conjecture posed at the beginning of this chapter; the proof relies on exploiting properties of the space c0 which features as the co-domain of the operator exhibited. We are interested in operators from a space to itself. The purpose of this chapter is to exhibit a class of Banach spaces which are genuine counterexamples to the previously stated conjecture. In fact, we provide counterexamples to the following, stronger version of the original conjecture; 'must a separable L∞ space with few operators and (cid:96)1 dual necessarily have very few operators?'. This question was originally posed by Argyros and Haydon in [2, Problem 10.7]. Our main result is the following: Theorem 3.1.4. Given any k ∈ N, k ≥ 2, there is a separable L∞ space, Xk, with the following properties:- 1. Xk is hereditarily indecomposable (HI) and X∗ k = (cid:96)1. 62 2. There is a non-compact bounded linear operator S : Xk → Xk. S is nilpotent of degree k, i.e. Sj (cid:54)= 0 for 1 ≤ j < k and, Sk = 0. 3. Moreover, Sj (0 ≤ j ≤ k − 1) is not a compact perturbation of any linear combination of the operators Sl, l (cid:54)= j. 4. The operator S : Xk → Xk is strictly singular (and consequently Sj is strictly singular for all j ≥ 1). 5. Every operator T on Xk can be uniquely represented as T =(cid:80)k−1 λi ∈ R, (0 ≤ i ≤ k − 1) and K is a compact operator on Xk. i=0 λiSi + K, where It is immediate from the above properties (and the fact that the strictly singular oper- ators are a closed ideal in the operator algebra - see Corollary 1.4.27) that the spaces Xk have few operators but not very few operators. In other words, RI + K(Xk) (cid:36) L(Xk) ⊆ RI + SS(Xk) (where SS(Xk) is the space of strictly singular operators on Xk ). We thus have a negative solution to Problem 10.7 of Argyros and Haydon ([2]). In addition to solving the previously discussed problem, it turns out that there are a number of other interesting consequences of these constructions, specifically concerning the Calkin algebras and the structure of closed ideals in L(Xk). Since the details of the proof of the main result are fairly long and technical, we choose to present these corollaries first. 3.2 Corollaries of the Main Theorem 3.2.1 On the structure of the closed ideals in L(Xk). The structure of norm-closed ideals in the algebra L(X) of all bounded linear operators on an infinite dimensional Banach space X is generally not well understood. For example, classifying all the norm-closed ideals in L((cid:96)p ⊕ (cid:96)q) with p (cid:54)= q remains an open problem (though some progress has been made in [27]). What is known, is that for the (cid:96)p spaces, 1 ≤ p < ∞, and c0, there is only one non-trivial closed ideal in L(X), namely the ideal of compact operators. This was proved by Calkin, [8], for (cid:96)2 and then extended to (cid:96)p and c0 by Gohberg et al., [14]. More recently, the complete structure of closed ideals in L(X) was described in [20] for X = (⊕∞ 2 )(cid:96)1. In both cases, there are exactly two nested proper closed ideals. Until the space constructed by Argyros and 2 )c0 and in [21] for X = (⊕∞ n=1(cid:96)n n=1(cid:96)n Haydon, [2], these were the only known separable, infinite dimensional Banach spaces for which the norm-closed ideal structure of the operator algebra is completely known. Clearly the space XAH of Argyros and Haydon provides another example of a separable Banach space for which the ideal of compact operators is the only (proper) closed ideal in the operator algebra. The spaces constructed in this chapter allow us to add to the list 63 of spaces for which the ideal structure of L(X) is completely known. In fact, we see that we can construct Banach spaces for which the ideals of the operator algebra form a finite, totally ordered lattice of arbitrary length. More precisely, the following is an immediate consequence of our main theorem. Theorem 3.2.1. There are exactly k norm-closed, proper ideals in L(Xk). The lattice of closed ideals is given by K(Xk) (cid:40) (cid:104)Sk−1(cid:105) (cid:40) (cid:104)Sk−2(cid:105) . . .(cid:104)S(cid:105) (cid:40) L(Xk). Here, if T is an operator on Xk, (cid:104)T(cid:105) is the smallest norm-closed ideal in L(Xk) containing T. 3.2.2 The Calkin algebra L(Xk)/K(Xk). We note that as a consequence of properties (3) - (5) of the main theorem, the Calkin algebra L(Xk)/K(Xk) is k dimensional with basis {I +K(Xk), S +K(Xk), . . . , Sk−1 +K(Xk)}. More precisely, it is isomorphic as an algebra to the subalgebra A of k × k upper-triangular- Toeplitz matrices, i.e. A is the subalgebra of Mat(k × k) generated by 0 1 0   j  1 . . . . . . . . . 1 0  : 0 ≤ j ≤ k − 1 An explicit isomorphism is given by ψ : L(Xk)/K(Xk) → A ∼= R[X]/(cid:104)xk(cid:105), k−1(cid:88) j=0 λjSj + K(Xk) (cid:55)→  λ0 λ1 λ2 0 ··· λ0 λ1 λ2 0 ... ... 0 0 ... ... 0 λ0 λ1 . . . 0 ... 0 ··· ··· λk−1 ··· λk−2 ... . . . ... . . . ... . . . λ0 0  We remark that the quotient algebra L(X)/SS(X) (where SS(X) denotes the strictly singular operators on X) was studied by Ferenczi in [12] for X a HI, or more generally, a HDn space. Here it was shown that for a complex HDn space, dim(L(X)/SS(X)) ≤ n2. The above remarks show that the Calkin algebra of a HI space can behave somewhat differently however; indeed any finite dimension for the Calkin algebra can be achieved. In the next chapter, we will extend these results further still, and show that it is possible to obtain a Banach space X∞ which has Calkin algebra isometric as a Banach algebra to (cid:96)1(N0). 64 3.2.3 Commutators in Banach Algebras Rather unexpectedly, the space X2 also provides a counterexample to a conjecture of Pro- fessor William B. Johnson; the author is thankful to Professor Johnson for bringing this to his attention. To motivate the conjecture we recall the following classical theorem of Wintner. Theorem 3.2.2 (Wintner). The identity in a unital Banach algebra, A, is not a commu- tator, that is, there do not exist x, y ∈ A such that 1 = [x, y] := xy − yx. Proof. Suppose for contradiction the theorem is false and choose x, y ∈ A such that 1 = xy − yx. We claim that xyn − ynx = nyn−1 for all n ∈ N. The proof is a simple induction; clearly it is true for n = 1. Suppose that xyn − ynx = nyn−1 holds for some n ∈ N. Then xyn+1 − yn+1x = (xyn − ynx)y + yn(xy − yx) = nyn−1 · y + yn · 1 = (n + 1)yn completing the induction. Observe that yn (cid:54)= 0 for any n, since otherwise there is a smallest value of n with yn = 0. However, we now find that 0 = xyn − ynx = nyn−1, contradicting the minimality of n. Finally, we have for any n n(cid:107)yn−1(cid:107) = (cid:107)xyn − ynx(cid:107) ≤ 2(cid:107)x(cid:107)(cid:107)y(cid:107)(cid:107)yn−1(cid:107) and since (cid:107)yn−1(cid:107) (cid:54)= 0 for all n, we have n ≤ 2(cid:107)x(cid:107)(cid:107)y(cid:107) for all n. This obvious contradiction completes the proof. If X is a Banach space and M is a proper norm-closed ideal of L(X), applying the above theorem to the unital Banach algebra L(X)/M, it is easily seen that no operator of the form λI + M with λ (cid:54)= 0 and M ∈ M is a commutator in the algebra L(X). In fact, as is commented in [18], this is in general the only known obstruction for an operator in L(X) to fail to be a commutator when X is any infinite dimensional Banach space. Consequently the authors of [18] define a Wintner space as a Banach space X such that every non commutator in L(X) is of the form λI + M where λ (cid:54)= 0 and M lies in a proper norm-closed ideal. Johnson conjectured that every infinite dimensional Banach space is a Wintner space. Until the existence of the spaces constructed in this thesis, all infinite dimensional Banach spaces X for which the commutators in L(X) could be classified satisfied this conjecture. However, we can easily see that X2 fails to be a Wintner space. Lemma 3.2.3. X2 is not a Wintner space. 65 Proof. We note that the operator S ∈ L(X2) fails to be of the form λI + M with λ (cid:54)= 0 and M lying in a proper norm-closed ideal of L(X2). Indeed, if S has such a form, we can write S = λI + µS + K where λ (cid:54)= 0, since by Theorem 3.2.1 the only proper norm-closed ideals in X2 are K(X2) and (cid:104)S(cid:105) = {µS + K : µ ∈ R, K ∈ K(X2)}. However, this contradicts the fact that the vectors I + K(X2) and S + K(X2) are linearly independent in L(X2)/K(X2) (see property (3) of the main theorem). To complete the proof, it suffices to see that S is not a commutator; this is easy using the operator representation, property (5), of the main theorem. Indeed suppose for con- tradiction that there exist T1, T2 ∈ L(X2) with S = [T1, T2]. For i = 1, 2, we can write Ti = λiI + µiS + Ki where λi, µi are scalars and Ki are compact operators. We then have S = T1T2 − T2T1 = (λ1I + µ1S + K1)(λ2I + µ2S + K2) − (λ2I + µ2S + K2)(λ1I + µ1S + K1) = K(cid:48) for some compact operator K(cid:48). However, we know by property (2) of the main theorem that S is not compact.Therefore S cannot be a commutator in L(X2). 3.2.4 Invariant subspaces. Finally, we remark that it was shown in [2] that all operators on the Argyros-Haydon space admit non-trivial, closed, invariant subspaces. The same is true for all the spaces Xk. Indeed, by a result of Lomonosov (see [23]), if an operator T commutes with a non-zero compact operator, then T has a proper, closed, invariant subspace. In particular, if there is some polynomial of T which is compact and non-zero, then certainly T has a proper, closed, invariant subspace. For an operator T : Xk → Xk, T =(cid:80)k−1 j=0 λjSj + K, we consider the polynomial of T , given by P(T ) := (T − λ0I)k. It follows (by the ring isomorphism of the Calkin algebra with the ring R[X]/(cid:104)xk(cid:105)) that P(T ) is a compact operator. So if P(T ) (cid:54)= 0 then we are done by the result of Lomonosov. Otherwise it is clear that λ0 is an eigenvalue of T , so it has a one dimensional invariant subspace. 3.3 The Basic Construction The rest of this chapter is devoted to proving the main theorem. Before continuing any further, we define the notation and terminology that shall be used throughout this chapter. We construct our spaces Xk by modifying the ideas used by Argyros and Haydon in their 66 construction of the space XAH. Consequently, it is convenient for us to work with the same notation and terminology that they introduced in [2]. We will be working with two strictly increasing sequences of natural numbers (mj) and (nj) which satisfy the following assumptions. Assumption 3.3.1. We assume that (mj, nj)j∈N satisfy the following: 1. m1 ≥ 4; 2. mj+1 = m2 j ; 3. n1 ≥ m2 1; 4. nj+1 ≥ (16nj)log2 mj+1 = m2 j+1(4nj)log2 mj+1. We note that these are almost the same assumptions as in [2]; the only difference is that j for all j ∈ N. (In [2], it was only assumed we demand the stronger condition that mj+1 = m2 that mj+1 ≥ m2 j for all j.) Consequently, we have no problems drawing upon results from [2]. In fact, for the purposes of this chapter, asking that mj+1 ≥ m2 j for all j ∈ N would have been sufficient, but we need the slightly stronger condition in the following chapter. Like the Argyros-Haydon space, our spaces Xk will be obtained from the generalised Bourgain-Delbaen construction, described in the previous chapter (Theorem 2.2.2). Con- I ⊆ N we define the projection PI : X →(cid:76) tinuing with the same notation used in Theorem 2.2.2, we recall that the subspaces Mn = [dγ : γ ∈ ∆n] form a finite-dimensional decomposition for X = X(Γ, τ ). For each interval n∈I Mn in the natural way; this is consistent with our use of P ∗ (0,n] in Theorem 2.2.2. As in [2], many of the arguments will involve sequences of vectors that are block sequences with respect to this FDD. It will therefore be useful to make the following definition; for x ∈ X, we define the range of x, denoted ran x, to be the smallest interval I ⊆ N such that x ∈(cid:76) n∈I Mn. We recall from Theorem 2.2.2 that there is an 'ε parameter' appearing in the admissible tuples of τ (γ) in the cases (0) and (2). Whilst this extra degree of freedom was required in the previous chapter in order for us to unite the original and generalised Bourgain-Delbaen constructions, we will not require it for the rest of this thesis; ε shall always be set equal to 1. Consequently, we choose to simplify our notation and simply omit the ε appearing in these tuples. This is also consistent with the notation from [2]. With this new notation, if γ ∈ ∆n+1, we say the corresponding vector c∗ τ (γ) = (α, ξ) (in the notation of Theorem 2.2.2 this would have been τ (γ) = (1, α, ξ)). Similarly, c∗ γ is called a Type 1 BD-functional if τ (γ) = (p, β, b∗) or a Type 2 BD-functional if τ (γ) = (α, ξ, p, β, b∗) (see Theorem 2.2.2). We will call γ a Type 0, 1 or 2 element in each of the respective cases. γ ∈ (cid:96)1(Γ) is a Type 0 BD-functional if 67 Our space will be a specific Bourgain-Delbaen space very similar to the space XAH constructed in [2]. We adopt the same notation used in [2], in which elements γ of ∆n+1 automatically code the corresponding BD-functionals. Consequently, we can write X(Γ) rather than X(Γ, τ ) for the resulting L∞ space. To be more precise, an element γ of ∆n+1 will be a tuple of one of the following forms: 1. γ = (n + 1, p, β, b∗), 2. γ = (n + 1, ξ, β, b∗) in which case τ (γ) = (1, ξ, rank ξ, β, b∗). in which case τ (γ) = (p, β, b∗); In each case, the first co-ordinate of γ tells us what the rank of γ is, that is to say to which set ∆n+1 it belongs, while the remaining co-ordinates specify the corresponding BD-functional. We observe that BD-functionals of Type 0 do not arise in this construction. Moreover, in the definition of a Type 2 functional, the scalar α that occurs is always 1 and p is always equal to rank ξ. As in the Argyros-Haydon construction, we shall make the further restriction that the β parameter must be of the form m−1 (nj) satisfy Assumption 3.3.1. We shall say that the element γ has weight m−1 of a Type 2 element γ = (n + 1, ξ, m−1 m−1 , where the sequences (mj) and . In the case , b∗) we shall insist that ξ be of the same weight, , as γ. j j j j To ensure that the sets ∆n+1 are finite we shall admit into ∆n+1 only elements of weight mj with j ≤ n + 1. A further restriction involves the recursively defined function called "age" (also defined in [2]). For a Type 1 element, γ = (n + 1, p, β, b∗), we define age γ = 1. For a Type 2 element, γ = (n + 1, ξ, m−1 , b∗), we define age γ = 1 + age ξ, and further restrict the elements of ∆n+1 by insisting that the age of an element of weight m−1 j may not exceed nj. Finally, we shall restrict the functionals b∗ that occur in an element of ∆n+1 by requiring them to lie in some finite subset Bn of (cid:96)1(Γn). It is convenient to fix an increasing sequence of natural numbers (Nn) and take Bp,n to be the set of all linear combinations η aη ≤ 1 and each aη is a rational number with denominator dividing Nn!. We may suppose the Nn are chosen in such a way that Bp,n is a 2−n-net in the unit ball of (cid:96)1(Γn \ Γp). The above restrictions may be summarized as follows. η, where(cid:80) b∗ =(cid:80) η∈Γn\Γp aηe∗ j , b∗) : 0 ≤ p < n, b∗ ∈ Bp,n (cid:111) (n + 1, p, m−1 j Assumption 3.3.2. ∆n+1 ⊆ n+1(cid:91) ∪ n−1(cid:91) j=1 (cid:110) p(cid:91) (cid:110) (n + 1, ξ, m−1 j , b∗) : ξ ∈ ∆p, weight ξ = m−1 j , age ξ < nj, b∗ ∈ Bp,n p=1 j=1 As in [2] we shall also assume that ∆n+1 contains a rich supply of elements of "even weight", more exactly of weight m−1 j with j even. 68 (cid:111) ∆n+1 ⊇ (n + 1, p, m−1 2j , b∗) : 0 ≤ p < n, b∗ ∈ Bp,n (cid:111) Assumption 3.3.3. (cid:110) (cid:98)(n+1)/2(cid:99)(cid:91) (cid:110) (cid:98)p/2(cid:99)(cid:91) ∪ n−1(cid:91) j=1 n+1(cid:91) ∪ n−1(cid:91) j=1 (cid:110) p(cid:91) (cid:110) (cid:111) (cid:111) . (n + 1, ξ, m−1 2j , b∗) : ξ ∈ ∆p, weight ξ = m−1 2j , age ξ < n2j, b∗ ∈ Bp,n p=1 j=1 For the main construction, there will be additional restrictions on the elements with "odd weight" m−1 2j−1, though we will come to these later. To begin with, we shall work with a different class of Bourgain-Delbaen spaces, and therefore choose to slightly change our notation from the above, reserving the above notation for the main construction. In j=1Λj. what follows, Υ will take the role of Γ above, and Λn the role of the ∆n's. In particular, Υ = ∪∞ n=1Λn and Υn = ∪n Now, we fix a k ∈ N, k ≥ 2 and define the space X(Υ) to be the Bourgain-Delbaen n=1Λn and the Λn are finite sets defined by space given as in Theorem 2.2.2, where Υ = ∪∞ recursion: Definition 3.3.4. We define Υ by the recursion Λ1 = {0, 1, . . . (k − 1)}, Λn+1 = (n + 1, p, m−1 , b∗) : 0 ≤ p < n, b∗ ∈ Bp,n j (cid:111) (n + 1, ξ, m−1 j , b∗) : ξ ∈ Λp, weight ξ = m−1 j , age ξ < nj, b∗ ∈ Bp,n p=1 j=1 Remark 3.3.5. 1. Note that the cardinality of Λ1 depends on the choice of k ∈ N. In this way, Υ (and consequently the BD space X(Υ)) really depend on the chosen k ∈ N. In an attempt to avoid even more complicated notation, we consider k ∈ N, k ≥ 2 to be fixed for the remainder of the chapter. 2. Later on, we will want to take a suitable subset of Υ. To avoid any ambiguity in nota- the set of all linear combinations b∗ =(cid:80) tion, in the above definition, and throughout the rest of the chapter, Bp,n will denote η aη ≤ 1 η, where, as before,(cid:80) aηe∗ η∈Υn\Υp and each aη is a rational number with denominator dividing Nn!. Eventually, we want to have a non-compact, bounded linear operator S on our space. The ideas used to construct this operator are very similar, but more involved, than those used in Section 2.5. We will make use of a single element set which is disjoint from Υ, and label the element 'undefined'. We obtain the following theorem. 69 Theorem 3.3.6. There is a map G : Υ → Υ ∪ {undefined} (we say G(γ) is undefined if G(γ) = undefined, otherwise we say G(γ) is defined) and a norm 1, linear mapping R∗ : (cid:96)1(Υ) → (cid:96)1(Υ) satisying: 1. G(j) = j − 1 for 1 ≤ j ≤ k − 1 and G(0) is undefined (recall Λ1 = {0, 1, . . . k − 1}). 2. For elements γ ∈ Υ\ Λ1 such that G(γ) is defined, rank γ = rank G(γ) and weight γ = weight G(γ) (i.e. G preserves weight and rank). Moreover, age G(γ) ≤ age γ (G doesn't increase age). (cid:40) (cid:40) e∗ G(γ) 0 d∗ 0 G(γ) if G(γ) is defined otherwise. if G(γ) is defined otherwise. 3. 4. R∗(e∗ γ) = R∗(d∗ γ) = Proof. We will construct the maps G and R∗ inductively. We note that since R∗ will be a linear operator on (cid:96)1(Υ), in order to ensure it is bounded, we only need to be able to control (cid:107)R∗(e∗ γ)(cid:107) ≤ M for every γ ∈ Υ, then it is elementary to check that R∗ is bounded with norm at most M . In particular, if property (3) of Theorem 3.3.6 holds, it follows that (cid:107)R∗(cid:107) = 1. γ)(cid:107) (for γ ∈ Υ). More precisely, if there is some M ≥ 0 s.t. (cid:107)R∗(e∗ To begin the inductive constructions of R∗ and G we define G : Λ1 → Λ1 by setting γ = d∗ j−1 for j ≥ 1. G(j) = j − 1 for 1 ≤ j ≤ k − 1 and declaring that G(0) is undefined. Noting that e∗ for γ ∈ Λ1, we define R∗(e∗ j−1 = d∗ We observe that this definition is consistent with the properties (1) - (4) above. 0) = 0 and R∗(e∗ j ) = R∗(d∗ 0) = R∗(d∗ j ) = e∗ Suppose that we have defined G : Υn → Υn and R∗ : (cid:96)1(Υn) → (cid:96)1(Υn) satisfying properties (1) - (4) above. We must extend G to Υn+1 and R∗ to a map on (cid:96)1(Υn+1). We consider a γ ∈ Λn+1 and recall that (see Theorem 2.2.2) e∗ γ. We wish to define R∗d∗ γ and R∗e∗ γ however we like. γ ∈ (cid:96)1(Υn)), and we want R∗ to be linear, once we However, since R∗c∗ have defined R∗d∗ γ. By linear independence, we are free to define R∗d∗ γ is already defined (c∗ γ, in order to have linearity we must have R∗e∗ γ + R∗d∗ γ. γ = R∗c∗ γ = c∗ γ + d∗ γ Let us consider R∗c∗ γ. We suppose first that age γ = 1 so that we can write γ = (n + 1, p, β, b∗) where b∗ ∈ (cid:96)1(Υn \ Υp). Consequently (cid:16) (0,p]b∗(cid:17) c∗ γ = β I − P ∗ = βP ∗ (p,∞)b∗. 70 αδd∗ We claim that R∗P ∗ unique choice of αδ). It follows from property (4) and the inductive hypothesis that (p,∞)b∗ = P ∗ δ ∈ Υn δ (for a (p,∞)R∗b∗. Indeed, we can write b∗ = (cid:80) (p,∞)b∗ = R∗(cid:0) (cid:88) R∗P ∗ (cid:1) = αδd∗ δ δ∈Υn\Υp (cid:88) αδd∗ G(δ) δ∈Υn\Υp∩ {η ∈ Υn : G(η) is defined} and it is easily checked (by a similar calculation) that we obtain the same expression for P ∗ (p,∞)R∗b∗. It follows that R∗c∗ γ = βP ∗ (p,∞)R∗b∗. We define G(γ) by G(γ) = (cid:40) undefined (n + 1, p, β, R∗b∗) (p,∞)R∗b∗ = 0 if P ∗ otherwise where it is a simple consequence of the facts that R∗ : (cid:96)1(Υn) → (cid:96)1(Υn) has norm 1 and satisfies property (3) that the element (n+1, p, β, R∗b∗) ∈ Υ. In the case where P ∗ (p,∞)R∗b∗ (cid:54)= 0, G(γ) is defined (with the definition as above) and it is evident that R∗c∗ G(γ). We γ = c∗ can define R∗d∗ γ) that we have R∗e∗ γ = e∗ γ = 0, so we can set R∗d∗ γ = d∗ G(γ) as required. Otherwise, when P ∗ γ = 0. G(γ) and it follows by linearity (and the fact that e∗ (p,∞)R∗b∗ = 0, R∗c∗ γ = 0 and again by linearity we get that R∗e∗ Now if γ has age > 1, we can write γ = (n + 1, ξ, β, b∗) and c∗ (rank ξ,∞)b∗. Making use of the inductive hypothesis yet again and the preceding argument, we have that γ = c∗ γ + d∗ ξ + βP ∗ γ = e∗ R∗c∗ (rank ξ,∞)R∗b∗ (rank ξ,∞)R∗b∗ (cid:40) ξ) + βP ∗ γ = R∗(e∗ G(ξ) + βP ∗ e∗ βP ∗ (rank ξ,∞)R∗b∗ It follows that if G(ξ) is undefined and P ∗ declare G(γ) to be undefined. Otherwise, there are two remaining possiblities (i) G(ξ) is undefined but P ∗ (rank ξ,∞)R∗b∗ = 0 then R∗c∗ if G(ξ) is defined otherwise = element G(γ) := (n + 1, rank ξ, β, R∗b∗) ∈ Υ. (rank ξ,∞)R∗b∗ (cid:54)= 0. In this case, it is easily verified that the γ = 0. In this case we (ii) G(ξ) is defined. It is again easily checked that the element G(γ) := (n+1, G(ξ), β, R∗b∗) ∈ Υ (here we note that in addition to the above arguments, we also need the inductive hypothesis that G does not increase the age of an element). In either of these cases, we see that R∗c∗ before, we then necessarily have R∗e∗ γ = e∗ G(γ). We can define R∗d∗ γ = c∗ G(γ) (in order that R∗ be linear). γ to be d∗ We have thus succeeded in extending the maps G and R∗. By induction, we therefore obtain maps G : Υ → Υ ∪ {undefined} and R∗ : (c00(Υ),(cid:107) · (cid:107)1) → (c00(Υ),(cid:107) · (cid:107)1) satisfying G(γ) and as 71 the four properties above (here (c00(Υ),(cid:107) · (cid:107)1) is the dense subspace of (cid:96)1(Υ) consisting of all finitely supported vectors). It follows from property (3), and the argument above, that R∗ is continuous, with (cid:107)R∗(cid:107) = 1. It therefore extends (uniquely) to a bounded linear map R∗ : (cid:96)1(Υ) → (cid:96)1(Υ) also having norm 1. This completes the proof. We make some important observations about the mappings G and R∗. Lemma 3.3.7. The dual operator of R∗, which we denote by (R∗)(cid:48) : (cid:96)1(Υ)∗ → (cid:96)1(Υ)∗ restricts to give a bounded linear operator R := (R∗)(cid:48)X(Υ) : X(Υ) → X(Υ) of norm at most 1. Proof. It is a standard result that the dual operator (R∗)(cid:48) is bounded with the same norm as R∗. It follows that the restriction of the domain to X(Υ) is a bounded, linear operator into (cid:96)1(Υ)∗ with norm at most 1. It only remains to see that this restricted mapping actually maps into X(Υ). Since the family (dγ)γ∈Υ is a basis for X(Υ), it is enough to see that the image of dγ under (R∗)(cid:48) lies in X(Υ). It is therefore sufficient for us to prove that (cid:88) (R∗)(cid:48)dδ = dγ γ∈G−1({δ}) where of course G−1({δ}) denotes the pre-image of {δ} under G; we remark that G fails to be injective and so is certainly not invertible. Since (d∗ γ)γ∈Υ is a basis for (cid:96)1(Υ) it is enough to show that for every θ ∈ Υ(cid:0)(R∗)(cid:48)dδ θ =(cid:0) (cid:88) (cid:1)d∗ (cid:1)d∗ θ. dγ γ∈G−1({δ}) The right hand side of this expression is easy to evaluate; it is only non-zero if G(θ) = δ, in which case it is equal to 1. In particular, if G(θ) is undefined, then the right hand side of the expression is certainly 0, as is(cid:0)(R∗)(cid:48)dδ (cid:1)d∗ θ = dδ(R∗d∗ θ) = dδ(0). If G(θ) is defined, the G(θ)) which is clearly 1 if G(θ) = δ and 0 otherwise. left hand side of the expression is dδ(d∗ So the expressions are indeed equal, as required. Lemma 3.3.8. For every γ ∈ Υ, there is a unique l, 1 ≤ l ≤ k such that Gj(γ) is defined whenever 1 ≤ j < l but Gl(γ) is undefined. Proof. The uniqueness is easy; if G(γ) is defined, l is the maximal j ∈ N such that Gj−1(γ) is defined. Otherwise we must have l = 1. So we only have to prove existence of such l. We prove by induction on n that if rank γ = n there is some 1 ≤ l ≤ k such that Gl(γ) is undefined, but Gj(γ) is defined if j < l. The case where n = 1 is clear from the construction of the map G. So, inductively, we assume the statement holds whenever rank γ = m ≤ n. Let γ ∈ Λn+1 and consider 2 cases: 72 (i) age γ = 1. We write γ = (n + 1, p, β, b∗). Now b∗ ∈ (cid:96)1(Υn) and by the inductive hypothesis, for every θ ∈ Υ with rank θ ≤ n, there is some l ≤ k such that Gl(θ) is undefined. It follows that we must have (R∗)lb∗ = 0 for some 1 ≤ l ≤ k. So it is (p,∞)(R∗)lb∗ = 0 for some 1 ≤ l ≤ k. It follows by construction of certainly true that P ∗ the map G that the 'l' we seek is the minimal l (1 ≤ l ≤ k) such that P ∗ (p,∞)(R∗)lb∗ = 0. (ii) age γ > 1. We write γ = (n + 1, ξ, β, b∗). If G(γ) is undefined we are done; we must have l = 1. Otherwise, G(γ) ∈ Υ and it follows from the construction of the map G (see the proof of Theorem 3.3.6) that there are two possibilities to consider; either G(γ) = (n + 1, rank ξ, β, R∗b∗) or G(γ) = (n + 1, G(ξ), β, R∗b∗). In the first of these two possibilities, the same argument as in the previous case shows that the l we seek is the minimal l (2 ≤ l ≤ k) such that P ∗ (p,∞)(R∗)lb∗ = 0. In the latter case, G(ξ) is defined. But, since rank ξ < n, we know by the inductive hypothesis that there exists some l0 ∈ N with 2 ≤ l0 ≤ k and Gl0(ξ) undefined but Gj(ξ) defined for j < l0. Now, if P ∗ (rank ξ,∞)(R∗)l0b∗ = 0, then it follows from construction of G that Gl0(γ) is undefined and Gj(γ) is defined for j < l0 so we are done. Otherwise, it follows from an argument above that l0 < k, and there is some (minimal) l, l0 < l ≤ k with P ∗ (rank ξ,∞)(R∗)lb∗ = 0. Once again, this is the desired l. Corollary 3.3.9. The maps R∗ : (cid:96)1(Υ) → (cid:96)1(Υ) and R : X(Υ) → X(Υ) satisfy (R∗)k = 0 and Rk = 0. Proof. It is clear from Lemma 3.3.8 and construction that the restriction of (R∗)k to c00(Υ) is the zero map. It follows by density and continuity that (R∗)k = 0. The other claims are immediate from the definition of R as the restriction of the dual operator of R∗. To obtain the extra constraints that we place on "odd-weight" elements we need to introduce a 'coding function' σ : Υ → N. This is similar to the coding function used in the Argyros-Hadyon construction, [2], and analogous to the coding function used in the Gowers- Maurey construction, [16]. We shall demand that σ satisfies the following properties: (1) σ is injective (2) σ(γ) > rank γ ∀ γ ∈ Υ (3) for γ ∈ Λn+1 (i.e. rank γ = n + 1), σ(γ) > max{σ(ξ) : ξ ∈ Υn}. 73 Such a σ can be constructed recursively as Υ is constructed. Now, for each γ ∈ Υ we can well-define a finite set Σ(γ) by Σ(γ) := {σ(γ)} ∪ k−1(cid:91) (cid:91) {σ(δ)} j=1 δ∈G−j (γ) where G−j(γ) := {θ ∈ Υ : Gj(θ) = γ}. (In particular, if θ ∈ G−j(γ) then Gl(θ) ∈ Υ for every l ≤ j.) Before giving our main construction, we document some important observations. Lemma 3.3.10. If γ ∈ Υ is such that G(γ) is defined then Σ(γ) ⊆ Σ(G(γ)). Proof. Certainly γ ∈ G−1(G(γ)) so σ(γ) ∈ Σ(G(γ)). Suppose σ(δ) ∈ Σ(γ), δ (cid:54)= γ. So, there is some 1 ≤ j ≤ k− 1 with δ ∈ G−j(γ), i.e. there is some δ such that Gj(δ) = γ. Since G(γ) is defined, we must have Gj+1(δ) = G(γ) ∈ Υ. In particular, by Lemma 3.3.8, we must in fact have had j < k − 1 so that j + 1 ≤ k − 1. Thus δ ∈ G−(j+1)(G(γ)) and σ(δ) ∈ Σ(G(γ)), as required. Lemma 3.3.11. If γ, γ(cid:48) ∈ Υ, rank γ > rank γ(cid:48) then Σ(γ) > Σ(γ(cid:48)), i.e. max{k : k ∈ Σ(γ(cid:48))} < min{k : k ∈ Σ(γ)}. Proof. This follows immediately from the definition of the sets Σ(γ) and Σ(γ(cid:48)), the fact that G is rank preserving and the assumption that for γ ∈ Λn+1, σ(γ) > max{σ(ξ) : ξ ∈ Υn}. Lemma 3.3.12. Suppose γ, δ ∈ Υ. If σ(γ) ∈ Σ(δ) then either γ = δ or there is some 1 ≤ j ≤ k − 1 such that Gj(γ) = δ. Proof. Since σ(γ) ∈ Σ(δ) there are two possibilities. Either σ(γ) = σ(δ) or there is some 1 ≤ j ≤ k − 1 and θ ∈ Υ with Gj(θ) = δ and σ(γ) = σ(θ). By injectivity of σ, this implies that either γ = δ, or that θ = γ and Gj(γ) = δ as required. We are finally in a position to describe the main construction. We will take a subset Γ ⊂ Υ by placing some restrictions on the elements of odd weight we permit. As a consequence of imposing these additional odd weight restrictions, we are also forced to (roughly speaking) remove those elements (n + 1, p, β, b∗) and (n + 1, ξ, β, b∗) of Υ for which the support of b∗ is not contained in Γ, in order that we can apply the Bourgain-Delbaen construction to obtain a space X(Γ). Note that the subset Γ will also be constructed inductively, so it will be well defined and consistent with the Bourgain-Delbaen method previously described. We will denote by ∆n the set of all elements in Γ having rank n, and denote by Γn the union Γn = ∪j≤n∆j. The permissible elements of odd weight will be as follows. For an age 1 element of odd weight, γ = (n + 1, p, m−1 2j−1, b∗), we insist that either b∗ = 0 or b∗ = e∗ η 74 4i < n−2 2j−1, b∗), we insist that either b∗ = 0 or b∗ = e∗ where η ∈ Γn \ Γp and weight η = m−1 γ = (n + 1, ξ, m−1 weight η = m−1 Definition 3.3.13. We define recursively sets ∆n ⊆ Λn. Then Γn := ∪j≤n∆j and Γ := ∪n∈N∆n ⊆ Υ. To begin the recursion, we set ∆1 = Λ1. Then 2j−1. For an odd weight element of age > 1, η where η ∈ Γn \ Γrank ξ and 2j−1, k ∈ Σ(ξ). Let us be more precise: 4k < n−2 ∆n+1 = (n + 1, p, m−1 2j , b∗) : 0 ≤ p < n, b∗ ∈ Bp,n ∩ (cid:96)1(Γn) (cid:111) j=1 (cid:110) (cid:98)(n+1)/2(cid:99)(cid:91) (cid:110) ∪ n−1(cid:91) (cid:98)p/2(cid:99)(cid:91) (cid:110) (cid:98)(n+2)/2(cid:99)(cid:91) p=1 j=1 ∪ j=1 ∪ n−1(cid:91) (cid:98)(p+1)/2(cid:99)(cid:91) p=1 j=1 (cid:110) (n + 1, ξ, m−1 2j , b∗) : ξ ∈ ∆p, w(ξ) = m−1 2j , age ξ < n2j, b∗ ∈ Bp,n ∩ (cid:96)1(Γn) (n + 1, p, m−1 2j−1, b∗) : 0 ≤ p < n, b∗ = 0 or b∗ = e∗ η, η ∈ Γn \ Γp (cid:111) 4i < n−2 2j−1 and w(η) = m−1 (n + 1, ξ, m−1 2j−1, b∗) : ξ ∈ ∆p, w(ξ) = m−1 2j−1, age ξ < n2j−1 , b∗ = 0 or b∗ = e∗ η with η ∈ Γn \ Γp, w(η) = m−1 4k < n−2 2j−1, k ∈ Σ(ξ) (cid:111) (cid:111) . Here the Bp,n are defined as in Remark 3.3.5, and, for brevity, we temporarily write w(ξ) for weight ξ. We define Xk to be the Bourgain-Delbaen space X(Γ) where Γ is the subset of Υ just defined. For the rest of this chapter we will work with the space Xk. As in [2], the structure of the space Xk is most easily understood in terms of the basis (dγ)γ∈Γ and the biorthogonal functionals d∗ γ in order to estimate norms. To this end, we have the following proposition. γ. However, we will need to work with the evaluation functionals e∗ Proposition 3.3.14. Let n be a positive integer and let γ be an element of ∆n+1 of weight m−1 j and age a ≤ nj. Then there exist natural numbers p0 < p1 < ··· < pa = n+1, elements ξ1, . . . , ξa = γ of weight m−1 that j with ξr ∈ ∆pr and functionals b∗ (cid:0)Γpr−1 \ Γpr−1 (cid:1) such r ∈ ball (cid:96)1 a(cid:88) a(cid:88) r=1 e∗ γ = = a(cid:88) a(cid:88) r=1 ξr + m−1 d∗ j ξr + m−1 d∗ j P ∗ (pr−1,∞)b∗ r P ∗ (pr−1,pr)b∗ r. r=1 r=1 75 If 1 ≤ t < a we have a(cid:88) ξr + m−1 d∗ j a(cid:88) r=t+1 r=t+1 e∗ γ = e∗ ξt + (pr−1,∞)b∗ P ∗ r . Proof. The proof is an easy induction on the age a of γ. We omit the details because the argument is the same as in [2] except our p0 is not necessarily 0. Remark 3.3.15. As in [2], we shall refer to any of the above identities as the evaluation analysis of the element γ, and the data (p0, (pr, b∗ r, ξr)1≤r≤a) as the analysis of γ. We will omit the p0 when p0 = 0. We will now construct the operator S : Xk → Xk. We need: Proposition 3.3.16. Γ is invariant under G. More precisely, if γ ∈ Γ ⊆ Υ and G(γ) is defined, then G(γ) ∈ Γ. It follows that the map G : Υ → Υ∪{undefined} defined in Theorem 3.3.6 restricts to give F : Γ → Γ ∪ {undefined}. Consequently the map R∗ : (cid:96)1(Υ) → (cid:96)1(Υ) (also defined in 3.3.6) can be restricted to the subspace (cid:96)1(Γ) ⊆ (cid:96)1(Υ) giving S∗ : (cid:96)1(Γ) → (cid:96)1(Γ). S∗ is a bounded linear map on (cid:96)1(Γ) of norm 1 which satisfies (cid:40) S∗e∗ γ = 0 e∗ F (γ) if F (γ) is undefined otherwise for every γ ∈ Γ. Moreover, the dual operator of S∗ restricts to Xk to give a bounded linear operator S : Xk → Xk of norm at most 1, satisfying Sj (cid:54)= 0 for 1 ≤ j ≤ k − 1, Sk = 0. Proof. It is enough to show that when γ ∈ Γ and G(γ) is defined, then G(γ) ∈ Γ. The claims about the operator S∗ follow immediately from the definition of S∗ as the restriction of R∗ and the definition of R∗. The fact that S : Xk → Xk is well defined follows by the same argument as in 3.3.7; indeed it is seen that for δ ∈ Γ, (cid:88) Sdδ = dγ. γ∈F −1(δ) Moreover, by Corollary 3.3.9, we see that (S∗)k = 0 and therefore that Sk = 0. That Sj (cid:54)= 0 for 1 ≤ j ≤ k − 1 is clear from the above formula and consideration of the elements dm for m ∈ ∆1. We use induction on the rank of γ to prove that if γ ∈ Γ and G(γ) is defined, then G(γ) ∈ Γ. This is certainly true when rank γ = 1. Suppose by induction, that whenever γ ∈ Γ, rank γ ≤ n and G(γ) is defined, G(γ) ∈ Γ and consider a γ ∈ Γ, of rank n+1 such that G(γ) is defined. Let us suppose first that this γ has age 1. We can write γ = (n + 1, p, β, b∗) η. Since G(γ) is defined, we where supp b∗ ⊆ Γn \ Γp, and we write b∗ = (cid:80) η∈Γn\Γp aηe∗ 76 have G(γ) = (n + 1, p, β, S∗b∗). We consider further sub-cases. (i.e. γ is an "even weight" element), we see that the only way G(γ) fails to be in Γ is if supp S∗b∗ (cid:42) Γn \ Γp. But If β = m−1 for some j 2j (cid:88) S∗b∗ = aηe∗ F (η) η∈Γn\Γp ∩ {η : F (η) is defined} and by the inductive hypothesis all the F (η) in this sum are in Γn\Γp. So supp S∗b∗ ⊆ Γn\Γp and G(γ) = F (γ) ∈ Γ as required. In the case where β = m−1 2j−1 (i.e. γ is an "odd-weight" element) we must also check that the odd weight element G(γ) is of a permissible form. (p,∞)S∗b∗ (cid:54)= 0. This of course Since G(γ) is defined, in particular we must have that P ∗ η for some η ∈ Γn \ Γp where implies that S∗b∗ (cid:54)= 0 and b∗ (cid:54)= 0. As γ ∈ Γ, b∗ = e∗ weight η = m−1 2j−1. Since S∗b∗ (cid:54)= 0, we must have S∗e∗ η = e∗ F (η) where in particular, F (η) is defined and lies in Γ by the inductive hypothesis. Since G is weight preserving, we have weight F (η) = weight η = m−1 2j−1. Moreover, since G preserves rank, we can now conclude that F (γ) = G(γ) = (n + 1, p, m−1 F (η)) ∈ Γ as required. 2j−1, S∗b∗) = (n + 1, p, m−1 4i < n−2 4i < n−2 2j−1, e∗ When age γ > 1, we can write γ = (n + 1, ξ, β, b∗) where (in particular) supp b∗ ⊆ Γn and ξ ∈ Γp (p < n). If this γ is of even weight, it follows easily from the inductive hyptohesis and arguments similar to the 'age 1' case that G(γ) = F (γ) ∈ Γ (when G(γ) is defined). So we consider only the case when weight γ = β = m−1 2j−1 (for some j) and G(γ) is defined. If G(ξ) = F (ξ) is undefined, then since G(γ) is defined, we must have G(γ) = (n + 1, rank ξ, m−1 S∗b∗ (cid:54)= 0. Again by the restrictions on elements of odd weight, we must have b∗ = e∗ weight η = m−1 and S∗b∗ = S∗e∗ hypothesis, that F (η) ∈ Γn \ Γrank ξ. Furthermore, weight F (η) = weight η = m−1 We conclude that G(γ) = F (γ) ∈ Γ as required. (rank ξ,∞)S∗b∗ (cid:54)= 0. So in particular, b∗ (cid:54)= 0 and η where 2j−1, some k ∈ Σ(ξ). Since S∗b∗ (cid:54)= 0, we must have G(η) = F (η) defined F (η). Since G preserves rank, we see as a consequence of the inductive 2j−1. 2j−1, S∗b∗) and P ∗ 4k < n−2 η = e∗ 4k < n−2 In the case where F (ξ) is defined, G(γ) = (n + 1, F (ξ), m−1 If S∗b∗ = 0 then by the inductive hypothesis, we certainly have G(γ) = F (γ) ∈ Γ and we are done. 2j−1, some k ∈ Σ(ξ) and Otherwise, we again must have b∗ = e∗ 2j−1 and k ∈ Σ(ξ) ⊆ Σ(G(ξ)) by S∗b∗ = e∗ Lemma 3.3.10. F (η). Now, weight F (η) = weight η = m−1 η where weight η = m−1 2j−1, S∗b∗). 4k < n−2 4k < n−2 Later, we will need the following lemma about the elements of odd weight in Γ. i )(cid:1) Lemma 3.3.17. Let γ ∈ Γ be an element of odd weight and age γ > 1. Let(cid:0)p0, (pi, ξi, b∗ be the analysis of γ, where we know each b∗ are i, j, 1 ≤ i < j ≤ age γ := a with b∗ i = e∗ i is either 0 or e∗ ηi and b∗ j = e∗ ηi for some suitable ηi. If there ηj then weight ηj (cid:12) weight ηi. 77 1 = e∗ Proof. We consider the cases when i = 1 and i > 1 separately. Suppose b∗ weight η1 = m−1 m−1 j where 1 ≤ j ≤ p. So weight η1 = m−1 monotonicity of the sequence mj, we get that m−1 η1 where 4i . By construction, elements of rank p are only allowed to have weights 4i =⇒ 4i ≤ rank η1 < rank ξ1 = p1. By the strict j (cid:54)= 0, b∗ Now for any j > 1, if b∗ 4k for some k ∈ Σ(ξj−1). Since the mapping F preserves rank of elements, and σ(θ) > rank θ ∀θ ∈ Γ (Assumption (2) of the σ mapping), it is immediate from the definition of Σ(ξj−1) that k ∈ Σ(ξj−1) =⇒ k > rank ξj−1 ≥ rank ξ1. Now ηj where weight ηj = m−1 4i = weight η1. j = e∗ < m−1 rank ξ1 weight ηj = m−1 4k < m−1 4rank ξj−1 ≤ m−1 4rank ξ1 ≤ weight η1 i = e∗ thus concluding the proof for the case i = 1. The case when i > 1 is easy. In this case for some l ∈ Σ(ξi−1). Now j > i, so (by Lemma we suppose b∗ 3.3.11) Σ(ξj−1) > Σ(ξi−1) so that l < k and therefore weight ηj = m−1 4l = weight ηi, completing the proof. ηi, weight ηi = m−1 4k < m−1 4l 3.4 Rapidly Increasing Sequences and the operator S : Xk → Xk We recall from [2] that special classes of block sequences, namely the rapidly increasing sequences admit good upper estimates. This class of block sequences will also be useful in our construction. We recall the definition: Definition 3.4.1. Let I be an interval in N and let (xk)k∈I be a block sequence (with respect to the FDD (Mn)). We say that (xk) is a rapidly increasing sequence, or RIS, if there exists a constant C such that the following hold: 1. (cid:107)xk(cid:107) ≤ C for all k ∈ I, and there is an increasing sequence (jk) such that, for all k, 2. jk+1 > max ran xk , 3. xk(γ) ≤ Cm−1 i whenever weight γ = m−1 i and i < jk . If we need to be specific about the constant, we shall refer to a sequence satisfying the above conditions as a C-RIS. Remark 3.4.2. We make the following important observation. If (xi)i∈N is a C-RIS, then so also is the sequence (Sxi). We omit the very easy proof. 78 We also note that the estimates of Lemmas and Propositions 5.2 - 5.6 and 5.8 of [2] all still hold. The same proofs go through, with only minor modifications to take account of the fact that p0 doesn't need to be 0 in the evaluation analysis of an element γ in our Γ (see Proposition 3.3.14). For convenience, we state Proposition 5.6 of [2] as we shall be making repeated use of it throughout this chapter:- Proposition 3.4.3. Let (xk) nj0 k=1 be a C-RIS. Then 1. For every γ ∈ Γ with weight γ = m−1 nj0(cid:88) n−1 j0 xk(γ) ≤ (cid:40) h we have 16Cm−1 4Cn−1 j0 j0 m−1 + 6Cm−1 h h k=1 n−1 j0 In particular, if h > j0 and if h < j0 if h ≥ j0 , xk(γ) ≤ 10Cm−2 j0 nj0(cid:88) nj0(cid:88) k=1 k=1 nj0(cid:88) (cid:107)n−1 j0 xk(cid:107) ≤ 10Cm−1 j0 . (cid:88) k∈J λkxk(γ) ≤ C max k∈J λk, (cid:107)n−1 j0 λkxk(cid:107) ≤ 10Cm−2 j0 . 2. If λk (1 ≤ k ≤ nj0) are scalars with λk ≤ 1 and having the property that for every γ of weight m−1 j0 and every interval J ⊆ {1, 2, . . . , nj0}, then k=1 Another result of particular importance to us will be the following proposition of [2]: Proposition 3.4.4. Let Y be any Banach space and T : Xk → Y be a bounded linear operator. If (cid:107)T (xk)(cid:107) → 0 for every RIS (xk)k∈N in Xk then (cid:107)T (xk)(cid:107) → 0 for every bounded block sequence in Xk. The proof exploits the explicit finite dimensional subspace structure provided by the L∞ spaces of Bourgain-Delbaen type; exactly the same argument used by Argyros and Haydon in [2] works for the spaces Xk and we refer the reader to the cited paper for further details. This proposition will be particularly important when we come to prove the operator representation property of the main theorem. We can also make use of Propositions 3.4.3 and 3.4.4 to prove the (cid:96)1 duality of the space Xk. 79 Proposition 3.4.5. The dual of Xk is (cid:96)1(Γ). More precisely the map defined by ϕ : (cid:96)1(Γ) → X∗ k ϕ(x∗)x := (cid:104)x, x∗(cid:105) (where x ∈ Xk ⊆ (cid:96)∞(Γ) = (cid:96)∗ 1(Γ), x∗ ∈ (cid:96)1(Γ)) is an isomorphism. Proof. The argument here can be found in [2], though we write out the details more ex- plicitly. We note that by Theorem 1.4.6 it is enough to show that the basis (dγ)γ∈Γ of Xk is shrinking. Using Proposition 1.4.5 it is enough to show that every bounded block basic sequence with respect to this basis is weakly null. By Proposition 3.4.4, it is sufficient to see that every RIS (xi)i∈N is weakly null. To this end, suppose (xi)i∈N is a C-RIS which fails to be weakly null; without loss of generality we may assume that there exists δ > 0 and an x∗ ∈ X∗ k such that (cid:107)x∗(cid:107) = 1 and x∗(xi) ≥ δ for all i. It follows that nj0(cid:88) i=1 1 nj0 x∗(xi) ≥ δ for any j0 ∈ N. On the other hand, we can choose j0 ∈ N such that 10Cm−1 from Proposition 3.4.3 that for this choice of j0 j0 < δ. It follows nj0(cid:88) i=1 nj0(cid:88) i=1 1 nj0 x∗(xi) ≤ (cid:107) 1 nj0 xi(cid:107) ≤ 10Cm−1 j0 < δ. This contradiction completes the proof. This canonical identification of X∗ k with (cid:96)1(Γ) allows us to prove some important prop- erties of the operator S : Xk → Xk. Lemma 3.4.6. The dual of the operator S : Xk → Xk is precisely the operator S∗ : (cid:96)1(Γ) → (cid:96)1(Γ) under the canonical identification, ϕ, of X∗ k with (cid:96)1(Γ). Moreover, Sj has closed range for every 1 ≤ j ≤ k − 1, and consequently, Sj : Xk → im Sj is a quotient operator. Proof. We will temporarily denote the dual map of S by S(cid:48) : X∗ k so that we don't confuse it with the S∗ mapping on (cid:96)1(Γ). By continuity and linearity, the maps ϕ−1S(cid:48)ϕ γ for γ ∈ Γ. For x ∈ Xk, and S∗ are completely determined by their action on the vectors e∗ k → X∗ (cid:40) 80 (cid:0)S(cid:48)ϕ(e∗ γ)(cid:1) (x) = ϕ(e∗ γ)Sx = (cid:104)Sx, e∗ γ(cid:105) = (cid:104)x, S∗e∗ γ(cid:105) = x(F (γ) 0 if F (γ) is defined otherwise It follows from this that (cid:40) e∗ F (γ) 0 ϕ−1S(cid:48)ϕ(e∗ γ) = = S∗(e∗ γ) if F (γ) is defined otherwise as required. To see that the image of Sj is closed, we make use of Banach's Closed Range Theorem, Theorem 1.4.14; it follows that it is enough to show that the image of (S∗)j is closed. This is easy, as it is clear that the image of (S∗)j is just (cid:96)1(Γ ∩ im F j) ⊆ (cid:96)1(Γ) (with the obvious embedding), so certainly (S∗)j has closed image. Since im Sj is closed, an easy application of the Inverse Mapping Theorem yields that Sj : Xk → im Sj is a quotient operator. Corollary 3.4.7. For 1 ≤ j ≤ k − 1there are aj, bj ∈ R, aj, bj > 0 such that whenever x ∈ Xk, aj(cid:107)Sjx(cid:107) ≤ dist(x, Ker Sj) ≤ bj(cid:107)Sjx(cid:107) Proof. This is immediate from the fact that Sj : Xk → im Sj is a quotient operator. Corollary 3.4.8. Suppose λi ∈ R (0 ≤ i ≤ k − 1) are such that (cid:80)k−1 i=0 λiSi is compact. Then λi = 0 for every i. Consequently, there does not exist j (1 ≤ j ≤ k − 1) such that Sj is a compact perturbation of some linear combination of the operators Sl, l (cid:54)= j. (Equivalently, {I + K(Xk), Sj + K(Xk) : 1 ≤ j ≤ k − 1} is a linearly independent set of vectors in L(Xk)/K(Xk).) Proof. It follows by a standard result that (cid:80)k−1 operator is compact. But the dual operator is just T :=(cid:80)k−1 i=0 λiSi is compact if, and only if, the dual i=0 λi(S∗)i : (cid:96)1(Γ) → (cid:96)1(Γ) and it is now easily seen that this is not compact unless all the λi are 0. Indeed, suppose T is 0) ∈ Γ. compact and consider first the sequence (e∗ Then, for m (cid:54)= n (observing that F (γ0 n=1 ⊆ B(cid:96)1(Γ) where γ0 )∞ n = (n+1, m−1 2 , 0, e∗ γ0 n (cid:107)T e∗ γ0 n − T e∗ γ0 m n) is undefined for every n) we have (cid:107)1 = λ0(cid:107)e∗ (cid:107)1 = 2λ0 − e∗ γ0 n γ0 m We must therefore have that λ0 = 0 in order that the sequence (T e∗ subsequence. Then, considering in turn the sequences (e∗ γj 1, m−1 n be 0 as claimed. n = (n + j ) ∈ Γ (for 1 ≤ j ≤ k − 1), we see by the same arguments that all the λi must n=1 ⊆ B(cid:96)1(Γ) where γj )∞ ) has a convergent 2 , 0, e∗ γ0 n 81 expressible in the form T =(cid:80)k−1 To complete the proof of the main theorem, it still remains for us to see that Xk is HI, the operator S is strictly singular, and that every bounded linear operator T ∈ L(Xk) is uniquely i=0 λiSi + K for some compact operator K : Xk → Xk. We note that the claimed uniqueness of this operator representation follows immediately from the above corollary. We focus now on proving the existence of the representation and aim to prove: Theorem 3.4.9. Let T : Xk → Xk be a bounded linear operator on Xk and (xi)i∈N a RIS in Xk. Then dist(T xi,(cid:104)xi, Sxi . . . , Sk−1xi(cid:105)R) → 0 as i → ∞. The proof is similar to that given in [2]. We will need slight modifications to the definitions of exact pairs and dependent sequences. We find it convenient to define both the 0 (δ = 0 in the definitions that follow) and 1 (δ = 1 in the definitions that follow) exact pairs and dependent sequences below. However, initially, we will only be concerned with the 0 exact pairs and dependent sequences. The 1 exact pairs and dependent sequences will only be needed to establish that the space Xk is hereditarily indecomposable. We also introduce the new, but related notions, of 'weak exact pairs' and 'weak dependent sequences' which will be useful to us later in establishing strict singularity of S. Definition 3.4.10. Let C > 0, δ ∈ {0, 1}. A pair (x, η) ∈ Xk × Γ is said to be a (C, j, δ)- special exact pair if j j for all ξ ∈ Γ 1. (cid:107)x(cid:107) ≤ C ξ, x(cid:105) ≤ Cm−1 2. (cid:104)d∗ 3. weight η = m−1 4. x(η) = δ and Slx(η) = 0 for every 1 ≤ l ≤ k − 1 5. for every element η(cid:48) of Γ with weight η(cid:48) = m−1 Cm−1 Cm−1 x(η(cid:48)) ≤ (cid:40) i (cid:54)= m−1 j , we have i j if i < j if i > j. Given also an ε > 0 we will say a pair (x, η) ∈ Xk × Γ is a (C, j, 0, ε)-weak exact pair if condition (4) is replaced by the following (weaker) condition: (4(cid:48)) Slx(η) ≤ Cε for 0 ≤ l ≤ k − 1. We will say a pair (x, η) ∈ Xk × Γ is a (C, j, 1, ε)-weak exact pair if condition (4) is replaced by condition: 82 (4(cid:48)(cid:48)) x(η) = 1 and Slx(η) ≤ Cε for 1 ≤ l ≤ k − 1. We note that a (C, j, δ) special exact pair is a (C, j, δ, ε) weak exact pair for any ε > 0. Moreover, the definition of a (C, j, δ)-special exact pair is the same as the definition of a (C, j, δ) exact pair given in [2] but with the additional requirement that Sjx(η) = 0 for all j ≥ 1. The remark made in [2] (following the definition of exact pairs) is therefore still valid. In fact it is easily verified that the same remark in fact holds for weak exact pairs. For convenience, we state the remark again as it will be useful to us later: Remark. A (C, j, δ, ε) weak exact pair also satisfies the estimates (cid:104)e∗ η(cid:48), P(s,∞)x(cid:105) ≤ 6Cm−1 6Cm−1 i j if i < j if i > j (cid:40) n2j(cid:88) k=1 n2j(cid:88) for elements η(cid:48) of Γ with weight η(cid:48) = m−1 i (cid:54)= m−1 j . We will need the following method for constructing special exact pairs. 2j (cid:80)n2j k=1 xk. For each k let b∗ k, xk(cid:105) = (cid:104)b∗ k=1 be a skipped-block C-RIS, and let q0 < q1 < q2 < ··· < qn2j Lemma 3.4.11. Let (xk)n2j be natural numbers such that ran xk ⊆ (qk−1, qk) for all k. Let z denote the weighted sum z = m2jn−1 k, xk(cid:105) = 0 and k, Slxk(cid:105) = 0 for all l.Then there exist ζi ∈ ∆qi (1 ≤ i ≤ n2j) such that (cid:104)(S∗)lb∗ the element η = ζn2j has analysis (qi, b∗ i , ζi)1≤i≤n2j and (z, η) is a (16C, 2j, 0)-special exact pair. Proof. The proof is the same as in [2]. We only need to show that Slz(η) = 0 for 1 ≤ l ≤ k − 1. This is easy. k be an element of Bqk−1,qk−1 with (cid:104)b∗ Slz(η) = (cid:104)Slz, e∗ η(cid:105) = (cid:104)Slz, d∗ ζk + m−1 2j P ∗ k(cid:105) (qk−1,qk)b∗ It is clear from the definition of S that ran xk ⊆ (qk−1, qk) =⇒ ran Slxk ⊆ (qk−1, qk) and since rank ζk = qk for every k, it follows that (cid:104)Slz,(cid:80)n2j (cid:105) = 0. We thus see that k=1 d∗ ζk Slz(η) = n−1 2j (cid:104)Slxk, b∗ k(cid:105) = 0 as required. k=1 Definition 3.4.12. Consider the space Xk. We shall say that a sequence (xi)i≤n2j0−1 is a (C, 2j0 − 1, δ)-special dependent sequence if there exist 0 = p0 < p1 < p2 < ··· < pn2j0−1, together with ηi ∈ Γpi−1 \ Γpi−1 and ξi ∈ ∆pi (1 ≤ i ≤ n2j0−1) such that 1. for each k, ran xk ⊆ (pk−1, pk) 83 2. the element ξ = ξn2j0−1 of ∆pn2j0−1 has weight m−1 2j0−1 and analysis (pi, e∗ ηi, ξi) n2j0−1 i=1 3. (x1, η1) is a (C, 4j1, δ)-special exact pair 4. for each 2 ≤ i ≤ n2j0−1, (xi, ηi) is a (C, 4σ(ξi−1), δ)-special exact pair. If we instead ask that (3(cid:48)) (x1, η1) is a (C, 4j1, δ, n−1 (4(cid:48)) (xi, ηi) is a (C, 4σ(ξi−1), δ, n−1 n2j0−1 i=0 we shall say the sequence (xi) 2j0−1) weak exact pair 2j0−1) weak exact pair for 2 ≤ i ≤ n2j0−1 is a weak (C, 2j0 − 1, δ) dependent sequence. In either case, we notice that, because of the special odd-weight conditions, we necessar- 2j0−1, by Lemma 3.3.17 2j0−1, and weight ηi+1 = m−1 = weight η1 < n−2 4σ(ξi) < n−2 ily have m−1 for 1 ≤ i < n2j0−1. 4j1 We also observe that a (C, 2j0 − 1, δ) special dependent sequence is certainly a weak (C, 2j0 − 1, δ) dependent sequence. Lemma 3.4.13. Let (xi)i≤n2j0−1 be a weak (C, 2j0 − 1, 0)-dependent sequence in Xk and let J be a sub-interval of [1, n2j0−1]. For any γ(cid:48) ∈ Γ of weight m2j0−1 we have (cid:88) ξn2j0−1, an element of weight m2j0−1. Let(cid:0)p(cid:48) i∈J each b(cid:48)∗ i is either 0 or e∗ η(cid:48) The proof is easy if all the b(cid:48)∗ i 0, (p(cid:48) for some suitable η(cid:48) i. r are 0, or if xi(γ(cid:48)) ≤ 7C. Proof. Let ξi, ηi, pi, j1 be as in the definition of a dependent sequence and let γ denote i)1≤i≤a(cid:48)(cid:1) be the analysis of γ(cid:48) where i, b(cid:48)∗ i , ξ(cid:48) {weight η(cid:48) r : 1 ≤ r ≤ a(cid:48), b(cid:48)∗ r = e∗ η(cid:48) r So we may suppose that there is some 1 ≤ r ≤ a(cid:48) s.t. b(cid:48)∗ for some i. We choose l maximal such that there exists i with b(cid:48)∗ weight ηl. Clearly we can estimate as follows } ∩ {weight ηi : 1 ≤ i ≤ n2j0−1} = ∅. with weight η(cid:48) i = e∗ η(cid:48) r i r = weight ηi i = and weight η(cid:48) (cid:88) k∈J xk(γ(cid:48)) ≤ (cid:88) k∈J, k<l xk(γ(cid:48)) + xl(γ(cid:48)) + xk(γ(cid:48)). r = e∗ η(cid:48) (cid:88) k∈J, k>l We now estimate the three terms on the right hand side of the inequality separately. xl(γ(cid:48)) ≤ (cid:107)xl(cid:107) ≤ C. Also 84 (cid:88) k∈J, k>l xk(γ(cid:48)) = (cid:88) (cid:88) i≤a(cid:48) l<k∈J, i≤a(cid:48) (cid:104)d∗ ξ(cid:48) (cid:104)d∗ ξ(cid:48) i ≤ n2 i k∈J, k>l 2j0−1 max where weight η(cid:48) i = 0, then Now each b(cid:48)∗ i is 0 or e∗ η(cid:48) contradict maximality of l). If b(cid:48)∗ i , xk(cid:105) + m−1 2j0−1(cid:104)b(cid:48)∗ , xk(cid:105) + m−1 i , P(p(cid:48) 2j0−1(cid:104)b(cid:48)∗ i−1,∞)xk(cid:105) i , P(p(cid:48) i−1,∞)xk(cid:105). i (cid:54)= weight ηk for any k > l (or else we would (cid:104)d∗ ξ(cid:48) i , xk(cid:105) + m−1 2j0−1(cid:104)b(cid:48)∗ i , P(p(cid:48) i−1,∞)xk(cid:105) = (cid:104)d∗ ξ(cid:48) i , xk(cid:105) ≤ Cweight ηk ≤ Cn−2 2j0−1 where the penultimate inequality follows from the definition of a (weak) exact pair, and the final inequality follows from Lemma 3.3.17. Otherwise b(cid:48)∗ (by restrictions on elements of odd weight) weight η(cid:48) exact pair and the remark following it we have where in particular 2j0−1. By the definition of (weak) i = e∗ η(cid:48) i < n−2 i (cid:104)d∗ ξ(cid:48) i , xk(cid:105) + m−1 2j0−1P(pi−1,∞)xk(η(cid:48) i) ≤ Cweight ηk + 6Cm−1 2j0−1 max{weight η(cid:48) i, weight ηk} Finally we consider (cid:80) ≤ 3Cn−2 2j0−1. is proved. So we can suppose l > 1. By definition of l, there exists some i such that b(cid:48)∗ and weight ηl = weight η(cid:48) k∈J, k<l xk(γ(cid:48)). Obviously if l = 1 this sum is zero, and the lemma i = e∗ η(cid:48) i. Now either i = 1 or i > 1. We consider the 2 cases separately. i Suppose first that i = 1. xk(γ(cid:48)) = (cid:104) (cid:88) (cid:88) k∈J, k<l a(cid:48)(cid:88) k∈J, k<l 2j0−1 max ≤ n2 r (cid:105) r−1,∞)b(cid:48)∗ 2j0−1P ∗ (p(cid:48) (cid:105) + m−1 2j0−1(cid:104) (cid:88) k∈J, k<l (cid:88) r≤a(cid:48)s.t. r =e∗ b(cid:48)∗ η(cid:48) r xk, r−1,∞)e∗ P ∗ η(cid:48) (p(cid:48) r (cid:105) r r=1 xk, d∗ ξ(cid:48) + m−1 J(cid:51)k<l, r≤a(cid:48) (cid:104)xk, d∗ ξ(cid:48) 2j0−1(cid:104) (cid:88) xk, r k∈J, k<l (cid:88) r≤a(cid:48)s.t. r =e∗ b(cid:48)∗ η(cid:48) r ≤ C + m−1 r−1,∞)e∗ P ∗ η(cid:48) (p(cid:48) r (cid:105) where the last inequality follows once again from the definition of exact pair. Suppose that for some k ∈ J, k < l, there is an r in {1, 2, . . . a(cid:48)} with b(cid:48)∗ By Lemma 3.3.17, we get weight η(cid:48) weight η(cid:48) to γ(cid:48). Thus there does not exist r in {1, 2, . . . a(cid:48)} with b(cid:48)∗ for some k ∈ J, k < l. Using an argument similar to the above, we finally deduce that r = weight ηk. r > 1. But since γ(cid:48) also has odd weight, this clearly contradicts Lemma 3.3.17 applied r = weight ηk r = weight ηk > weight ηl = weight η(cid:48) 1, i.e. weight η(cid:48) and weight η(cid:48) and weight η(cid:48) r = e∗ η(cid:48) r = e∗ η(cid:48) r r 2j0−1(cid:104) (cid:88) m−1 xk, k∈J, k<l (cid:88) r≤a(cid:48)s.t. b(cid:48)∗ r =e∗ η(cid:48) r 85 r−1,∞)e∗ P ∗ η(cid:48) (p(cid:48) r (cid:105) ≤ 2C and so we get the required result. Finally it remains to consider what happens when i > 1. Recall we are also assuming l > 1 and weight η(cid:48) m−1 4σ(ξl−1), and by restrictions on elements of odd weights, weight η(cid:48) By strict monotonicity of the sequence mj, we deduce that ω = σ(ξl−1) ∈ Σ(ξ(cid:48) Lemma 3.3.12 there are now two possibilites. Either ξl−1 = ξ(cid:48) 1 ≤ j ≤ k− 1, such that F j(ξl−1) = ξ(cid:48) this implies pl−1 = p(cid:48) γ(cid:48) as i = weight ηl. But by definition of a special exact pair, we have weight ηl = i−1). i−1). By i−1 or, if not, there is some j, i−1. In either of these cases, we note that in particular i−1 since F preserves rank and we can write the evaluation analysis of 4ω with ω ∈ Σ(ξ(cid:48) i = m−1 e∗ γ(cid:48) = (S∗)j(e∗ ξl−1 ) + d∗ ξ(cid:48) r + m−1 2j0−1P ∗ (p(cid:48) r−1,p(cid:48) r)b(cid:48)∗ r a(cid:48)(cid:88) r=i for some 0 ≤ j ≤ k − 1. Now, for k < l, since ran xk ⊆ (pk−1, pk) ⊆ (0, pl−1) = (0, p(cid:48) see that i−1), we (cid:104)xk, e∗ γ(cid:48)(cid:105) = (cid:104)xk, (S∗)je∗ l−1(cid:88) ξl−1 ξr + m−1 d∗ (cid:105) = (cid:104)Sjxk, = m−1 ≤ Cn−1 r=1 2j0−1(cid:104)Sjxk, e∗ 2j0−1 ηk (cid:105) 2j0−1P ∗ (pr−1,pr)e∗ ηr(cid:105) by definition of a weak exact pair. It follows that (cid:88) k∈J, k<l xk(γ(cid:48)) ≤ n2j0−1 max k∈J, k<l xk(γ(cid:48)) ≤ C. This completes the proof. As a consequence of the above lemma and Proposition 3.4.3 we obtain the following upper norm estimate for the averages of weak special dependent sequences. Proposition 3.4.14. Let (xi)i≤n2j0−1 be a weak (C, 2j0 − 1, 0) dependent sequence in Xk. Then (cid:107)n−1 2j0−1 xi(cid:107) ≤ 70Cm−2 2j0−1 n2j0−1(cid:88) i=1 Proof. We apply the second part of Proposition 3.4.3, with λi = 1 and 2j0−1 playing the role of j0. Lemma 3.4.13 shows that the extra hypothesis of the second part of Proposition 3.4.3 is satisfied, provided we replace C by 7C. We deduce that (cid:107)n−1 2j0−1 as claimed. (cid:80)n2j0−1 xi(cid:107) ≤ 70Cm−2 2j0−1 i=1 86 We will need the following lemma from elementary linear algebra in the proof of Theorem 3.4.9. Lemma 3.4.15. Suppose for j = 1, . . . , k, uj ∈ Qn ⊆ Rn (where k ≤ n) and moreover, the uj are linearly independent over R. Let V = {v ∈ Rn : (cid:104)v, uj(cid:105) = 0 for j = 1, . . . , k}. Then V ∩ Qn is dense in V . Proof. Consider the linear map L : Qn → Qk, defined by Qn (cid:51) v (cid:55)→ ((cid:104)v, u1(cid:105),(cid:104)v, u2(cid:105), . . . ,(cid:104)v, uk(cid:105)). Thinking of Qn and Qk as vector spaces over Q, and by the assumed linear independence of the uj, we can apply the Rank-Nullity Theorem to L to conclude that V ∩ Qn = ker L is a Q-vector space of dimension n − k. Now, V ∩ Qn (where the closure is taken in Rn), is easily seen to be a real vector space. Moreover, V ∩ Qn ⊆ V = V , and it is easily seen (e.g. by a similar rank-nullity argument as before) that V is an n− k dimensional real vector space. So V ∩ Qn is a real vector space of dimension at most n − k. On the other hand, if q1, . . . ql (l ≤ n) are vectors in Qn which are linearly independent over Q, they must in fact be linearly independent over R. Indeed, without loss of generality, we may assume l = n; if not, working in the Q-vector space, Qn, we may extend the (qi)i≤l to a basis, (qi)1≤i≤n, of Qn and our claim we will be proved if we show this extended set (which is obviously linearly independent over Q) is really linearly independent over R. Now, since q1, . . . , qn is a basis of Qn, the standard basis of Rn, (ej)n j=1 is in the real (in fact, rational) span of q1, . . . , qn. Consequently, Rn is the real span of the q1, . . . qn, and so by basic linear algebra, the q1, . . . qn must in fact be linearly independent over R which is what we wanted. Since we have seen that V ∩ Qn is an n − k dimensional Q-vector space, there are n − k, Q-linearly independent vectors in V ∩ Qn ⊆ V ∩ Qn. By the preceding paragraph, these n − k vectors are linearly independent over R and so V ∩ Qn is a real vector space of dimension at least n − k. Combined with our earlier observation, we conclude that V ∩ Qn is a real vector space of dimension exactly n− k. Now V ∩ Qn ⊆ V and both are real vector spaces of the same dimension, so they must in fact be equal, as required. We can now prove the following analogous result to Lemma 7.1 of [2]. Lemma 3.4.16. Let m < n be natural numbers and let x ∈ Xk ∩ QΓ, y ∈ Xk be such that ran x, ran y are both contained in the interval (m, n]. Suppose that dist(y,(cid:104)x, Sx . . . , Sk−1x(cid:105)R) > δ. Then there exists b∗ ∈ ball (cid:96)1(Γn \ Γm), with rational coordinates, such that (cid:104)b∗, Sjx(cid:105) = 0 for j = 0, 1, . . . k − 1 and b∗(y) > 1 2 δ. 87 (cid:107)y −(cid:80)k−1 j=0 λjSjx(cid:107) ≤ (cid:107)in(cid:107)(cid:107)v −(cid:80)k−1 Proof. It is easily seen (from the construction of S) that since x ∈ Xk ∩ QΓ, with ran x ⊆ (m, n], we have Sjx ∈ Xk ∩ QΓ and ran Sjx ⊆ (m, n] for every j = 0, 1, . . . k − 1. For each j, let uj ∈ (cid:96)∞(Γn \ Γm) be the restriction of Sjx and let v ∈ (cid:96)∞(Γn \ Γm) be the restriction of y. Then Sjx = inuj for every j, y = inv and so, for any scalars (λj)k−1 j=0 , 2 δ and so, by the Hahn -- Banach Theorem in the finite dimensional space (cid:96)∞(Γn \ Γm), there exists a∗ ∈ ball (cid:96)1(Γn \ Γm) with a∗(uj) = 0 for every j and a∗(v) > 1 2 δ. Since Sjx has rational coordinates for every j, our vectors uj are in QΓn\Γm. It follows from Lemma 3.4.15 that we can approximate a∗ arbitrarily well with b∗ ∈ QΓn\Γm retaining the condition b∗(uj) = 0 for every j. j=0 λjuj(cid:107). Hence dist(v,(cid:104)u0, u1, . . . , uk−1(cid:105)R) > 1 The proof of Theorem 3.4.9 is now easy. We obtain the following minor variation of [2, Lemma 7.2]. We omit the proof since exactly the same argument as in [2] works: Lemma 3.4.17. Let T be a bounded linear operator on Xk, let (xi) be a C-RIS in Xk ∩ QΓ and assume that dist(T xi,(cid:104)xi, Sxi . . . , Sk−1xi(cid:105)R) > δ > 0 for all i. Then, for all j, p ∈ N, there exist z ∈ [xi : i ∈ N], q > p and η ∈ ∆q such that 1. (z, η) is a (16C, 2j, 0)-special exact pair with ran z ⊂ (p, q); 2. (T z)(η) > 7 3. (cid:107)(I − P(p,q))T z(cid:107) < m−1 2j δ; 4. (cid:104)P ∗ η, T z(cid:105) > 3 8 δ. 16 δ; (p,q]e∗ Theorem 3.4.9 is now proved in exactly the same way as that of [2, Proposition 7.3]. 3.5 Operators on the Space Xk In this section, we see that all operators on the space Xk are expressible as(cid:80)k−1 j=0 λjSj +K for suitable scalars λj and some compact operator K on Xk. We choose to give an elementary proof of this fact in this chapter. However, in the following chapter, we will construct a space X∞ which has a similar operator representation; S is no longer nilpotent in this j=0 λjSj, where the scalars (λj)∞ j=0 lie in (cid:96)1(N0). The proof of the operator representation in Chapter 4 will be much more technical, relying on a fixed-point-theorem due to Kakutani. We remark that case, so we obtain an infinite sum in the powers of S of the form (cid:80)∞ the proof of the operator representation in Chapter 4 would also work (with only obvious changes) to prove the operator representation we require in this section. With that said, we now proceed to give a more elementary proof. We will need the following lemmas. 88 (cid:48). Certainly then ran yi = ran xi for every (cid:13)(cid:13) ≤ 5(cid:13)(cid:13)(cid:101)xi (cid:48) − xi (cid:13)(cid:13) → 0. (cid:48) − xi Proof. Let ran xi = (pi, qi) and set yi = P(pi,qi)(cid:101)xi i. Note that(cid:0)I − P(pi,qi) (cid:1) xi = 0 and consequently (cid:13)(cid:13)(cid:0)I − P(pi,qi) (cid:1)(cid:101)xi It follows that (cid:48)(cid:13)(cid:13) =(cid:13)(cid:13)(cid:0)I − P(pi,qi) (cid:48) − xi)(cid:13)(cid:13) ≤(cid:13)(cid:13)I − P(pi,qi) (cid:13)(cid:13)(cid:13)(cid:13)(cid:101)xi (cid:1) ((cid:101)xi (cid:13)(cid:13) (cid:107)yi − xi(cid:107) =(cid:13)(cid:13)(cid:101)xi (cid:48) −(cid:0)(I − P(pi,qi))(cid:101)xi (cid:48)(cid:1) − xi (cid:48)(cid:13)(cid:13) → 0. (cid:13)(cid:13) +(cid:13)(cid:13)(cid:0)I − P(pi,qi) ≤(cid:13)(cid:13)(cid:101)xi (cid:1)(cid:101)xi γ(cid:105) = (cid:104)(cid:101)xi γ(cid:105) = (cid:104)SjP(pi,qi)(cid:101)xi (cid:48), e∗ = (cid:104)(cid:101)xi γ(cid:105) = (cid:104)Sj(cid:101)xi (pi,qi)e∗ (cid:48), (S∗)jP ∗ (cid:48) − xi (cid:48), P ∗ k satisfying (cid:48) − xi(cid:107) → 0 as i → ∞ and Sj(cid:101)xi (cid:48) (cid:48) = 0 for every i. Then there is a subsequence Lemma 3.5.1. Let 1 ≤ j ≤ k − 1 and (xi)i∈N be a C-RIS in Xk. Suppose there are (cid:101)xi such that (cid:107)(cid:101)xi (xik )k∈N of (xi) and vectors x(cid:48) (1) ran xik = ran x(cid:48) (2) Sjx(cid:48) (3) (cid:107)x(cid:48) (4) (x(cid:48) We note that in particular, if (xi)i∈N is a C-RIS with Sjxi → 0, then the above hypothesis are satisfied as a consequence of Corollary 3.4.7. k = 0 for every k k − xik(cid:107) → 0 k)k∈N is a 2C-RIS. k Note also that Sjyi = 0 for every i. Indeed, for γ ∈ Γ, Sjyi(γ) = (cid:104)Sjyi, e∗ γ(cid:105) (pi,qi)(S∗)je∗ (cid:48), P ∗ γ(cid:105) = 0 (pi,qi)e∗ (cid:48) = 0 for every i. (We also made use of the fact that P ∗ (pi,qi)(S∗)j = (S∗)jP(pi,qi). This follows from the construction of the operator S∗ and the fact that F is rank-preserving. A more detailed argument of this fact is presented in the proof of Theorem 3.3.6.) since Sj(cid:101)xi So far we have managed to achieve (1) - (3) of the above. We show we can extract a subsequence of the yi, (yik )k∈N say, such that (yik )k∈N is a 2C-RIS. The proof will then be complete if we set x(cid:48) k = yik and take the subsequence (xik ) of the xi. Since (cid:107)xi(cid:107) ≤ C for every i and (cid:107)yi − xi(cid:107) → 0, we can certainly assume (by ignoring some finite number of terms at the beginning of the sequence) that (cid:107)yi(cid:107) ≤ 2C for every i. Let (jk) be the increasing sequence corresponding to the C-RIS (xi), i.e. (i) jk+1 > max ran xk (ii) xk(γ) ≤ Cm−1 i when weight γ = m−1 i and i < jk 89 Set l1 = j1. We can certainly find an i1 ≥ 1 such that (cid:107)yi1 − xi1(cid:107) ≤ Cm−1 weight γ = m−1 w with w < l1, then certainly w < l1 = j1 ≤ ji1 so yi1(γ) ≤ (yi1 − xi1)(γ) + xi1(γ) ≤ Cm−1 l1 + Cm−1 w ≤ 2Cm−1 w l1 . So if γ ∈ Γ, Now set l2 = ji1+1. So l2 > max ran xi1 = max ran yi1. Inductively, suppose we have defined natural numbers l1 ≤ l2 ≤ ··· ≤ ln and i1 < i2 < ··· < in such that (i') lk+1 > max ran yik for all 1 ≤ k ≤ n − 1, and (ii') for all 1 ≤ k ≤ n, yik (γ) ≤ 2Cm−1 w whenever γ ∈ Γ, weight γ = m−1 w with w < lk It is easily seen from the inductive construction that ln+1 ≥ ln and Set ln+1 = jin+1. moreover (by choice of jk), ln+1 > max ran xin = max ran yin. Now we can certainly find in+1 > in such that (cid:107)yin+1 − xin+1(cid:107) ≤ Cm−1 w with w < ln+1 In particular w < jin+1 ≤ jin+1, so by choice of in+1 and the fact that (xi) is a RIS, we see that . So suppose γ ∈ Γ, weight γ = m−1 ln+1 yin+1(γ) ≤ (yin+1 − xin+1)(γ) + xin+1(γ) ≤ Cm−1 ln+1 + Cm−1 w ≤ 2Cm−1 w . Inductively we obtain a subsequence (yik )k∈N which is evidently a 2C-RIS (with the sequence (lk)k∈N satisfying the RIS definition), as required. We will also need the following technical lemma. Lemma 3.5.2. Suppose (xi)i∈N is a normalised sequence in Xk and λj ∈ R (0 ≤ j ≤ k − 1) are scalars such that(cid:80)k−1 j=0 λjSjxi → 0. If Sk−1xi(cid:88)→ 0 then λj = 0 for every j. Otherwise, there is 1 ≤ m ≤ k − 1 such that Smxi → 0 but Sjxi(cid:88)→ 0 if j < m, in which case, we must have that λj = 0 for all j < m. Proof. We consider first the case where Sk−1xi → 0 and choose m ∈ {1, . . . k − 1} minimal such that Smxi → 0 (noting such an m obviously exists). We must observe that λj = 0 for j=0 λjSjxi → 0. If m = 1, this of course implies that λ0 = 0 since the sequence (xi) is normalised and we are done. all j < m. Since Sjxi → 0 for all j ≥ m we in fact know that(cid:80)m−1 Otherwise, we apply the operator Sm−1 to the previous limit. Once again making use Sm−1xi(cid:88)→ 0, we must again have λ0 = 0, and moreover,(cid:80)m−1 of the fact that Sjxi → 0 when j ≥ m, we find that λ0Sm−1xi → 0. Since, by choice of m, j=1 λjSjxi → 0. If m = 2, then this implies λ1Sxi → 0 which implies that λ1 = 0 (since Sxi (cid:88)→ 0). Otherwise, we apply the operator Sm−2. A similar argument concludes once again that we must have λ1 = 0. Continuing in this way, we get that λj = 0 for all j < m as required. 90 required. Theorem 3.5.3. Let T : Xk → Xk be a bounded linear operator on Xk. Then there are j=0 λjSj + K. Proof. We will show that there exist scalars λj such that whenever (xi)i∈N is a RIS, T xi − j=0 λjSjxi → 0 j=0 λjSj is compact. We note that it is enough to show that there are λj ∈ R such that whenever In the case where Sk−1xi (cid:88)→ 0, we notice that in particular this implies Sjxi (cid:88)→ 0 for (cid:80)k−1 every 1 ≤ j ≤ k − 1. Applying the operators Sk−1, Sk−2, . . . S sequentially to the limit, j=0 λjSjxi → 0, first yields that λ0 = 0, then λ1 = 0 etc. So λj = 0 for every j as λj ∈ R (0 ≤ j ≤ k− 1) and a compact operator K : Xk → Xk such that T =(cid:80)k−1 j=0 λjSjxi → 0 as i → ∞. By Proposition 3.4.4, this implies T xi −(cid:80)k−1 (cid:80)k−1 for every block sequence (xi) which, by Proposition 1.4.15, implies that T −(cid:80)k−1 (xi)i∈N is a RIS, we can find some subsequence (xil) with T xil −(cid:80)k−1 Claim 1. Suppose (xi)i∈N is a normalised RIS and that Sk−1xi(cid:88)→ 0 (noting in particular (cid:88)→ 0 for every 1 ≤ j ≤ k − 1). Then there are λj ∈ R (0 ≤ j ≤ k − 1) and a subsequence (xil) of (xi) such that T xil −(cid:80)k−1 i ∈ R such that (cid:107)T xi −(cid:80)k−1 i Sjxi(cid:107) → 0. We first show that the λ0 that this implies that Sjxi j=0 λjSjxil → 0. j=0 λjSjxil → 0. j=0 λj By passing to a subsequence, we may assume that there is some ε > 0 such that (cid:107)Sjxi(cid:107) ≥ ε for every i ∈ N and every 1 ≤ j ≤ k − 1. By Theorem 3.4.9, there are λj i must converge. The argument is similar to that of [2]. If not, by passing to a subsequence, we may assume that 2i+1 − λ0 λ0 2i ≥ δ > 0 for some δ. Since yi := x2i−1 + x2i is a RIS, we deduce from Theorem 3.4.9 that there are µj (cid:16) ≤ (cid:107) k−1(cid:88) i ∈ R with (cid:107)T yi −(cid:80)k−1 (cid:17) 2i−1Sjx2i−1 − T x2i−1(cid:107) + (cid:107)T yi − k−1(cid:88) (cid:17) 2iSjx2i − T x2i(cid:107) + (cid:107) k−1(cid:88) i Sjyi(cid:107) → 0. Now (cid:107) k−1(cid:88) 2i−1 − µj λj Sjx2i−1(cid:107) 2i − µj λj i Sjyi(cid:107) µj k−1(cid:88) Sjx2i + j=0 µj (cid:16) i λj i λj j=0 j=0 j=0 j=0 j=0 and so we deduce that both sides of the inequality converge to 0. Since the sequence (xi) is a block sequence, there exist lk such that P(0,lk]yk = x2k−1 and P(lk,∞)yk = x2k. Recalling that if x ∈ Xk has ran x = (p, q] then ran Sjx ⊆ (p, q], we consequently have (cid:16) (cid:107) k−1(cid:88) 2i − µj λj i j=0 and similarly (cid:17) (cid:17) (cid:16) Sjx2i(cid:107) ≤ (cid:107)P(li,∞)(cid:107)(cid:107) k−1(cid:88) (cid:16) (cid:107) k−1(cid:88) j=0 i 2i − µj λj (cid:17) 2i−1 − µj λj i Sjx2i−1(cid:107) → 0. k−1(cid:88) (cid:16) j=0 (cid:17) Sjx2i + 2i−1 − µj λj i Sjx2i−1(cid:107) → 0 j=0 91 By continuity of S and the fact that Sk = 0, applying Sk−1 to both limits above (and recalling that (cid:107)Sjxi(cid:107) ≥ ε for every 1 ≤ j ≤ k − 1) we obtain (cid:107)(cid:0)λ0 2i − µ0 i (cid:1) Sk−1x2i(cid:107) ≤ 1 (cid:107)Sk−1(cid:107)(cid:107) k−1(cid:88) (cid:16) ε j=0 λ0 2i − µ0 i ≤ 1 ε (cid:17) 2i − µj λj i Sjx2i(cid:107) → 0 and similarly we find that λ0 our assumption. So the λ0 2i−1 → 0 contrary to i converge to some λ0 as claimed. It follows that (cid:107)T xi − λ0xi − i Sjxi(cid:107) → 0. We observe that, since (xi) is normalised and (cid:107)Sk−1xi(cid:107) ≥ ε, applying i → 0. It follows that λ0 (cid:80)k−1 2i−1 − µ0 2i − λ0 j=1 λj Sk−2 to the previous limit, we see that λ1 i ≤ 1 ε (cid:107)λ1 i Sk−1xi(cid:107) ≤ 1 ε (cid:0)(cid:107)Sk−2(cid:107)(cid:107)T xi − λ0xi − k−1(cid:88) i Sjxi(cid:107) + (cid:107)Sk−2T − λ0Sk−2(cid:107)(cid:1) λj j=1 so that in particular the λ1 quence λ1 il (limit λ1) of the λ1 i are bounded. Consequently there is some convergent subse- i . It follows that the corresponding subsequence (xil) satisfies T xil − λ0xil − λ1Sxil − k−1(cid:88) Sjxil → 0 λj il j=2 Now, if k = 2 we are done (the last sum is empty). Otherwise, we can apply Sk−3 to the )∞ previous limit and use the same argument to conclude that (λ2 l=1 is a bounded sequence il of scalars. Continuing in this way, we eventually find (after passing to further subsequences which we relabel as xil) that there are λj with (T −(cid:80)k−1 subsequence (xil) of (xi) such that (T −(cid:80)m0−1 Claim 2. Suppose (xi)i∈N is a normalised C-RIS and that Smxi → 0 for some 1 ≤ m ≤ k−1. Let m0 ≥ 1 be minimal such that Sm0xi → 0. Then there are λj ∈ R (0 ≤ j < m0) and a j=0 λjSj)xil → 0 as required. j=0 λjSj)xil → 0. By minimality of m0, we can assume (by passing to a subsequence if necessary) that (cid:107)Sjxi(cid:107) ≥ ε for all i ∈ N and all j < m0. By Lemma 3.5.1, (with j = m0) there is a 2C-RIS l ∈ Ker Sm0 ⊆ Ker Sj for j ≥ m0 and (x(cid:48) l)l∈N and some subsequence (xil) of (xi) such that x(cid:48) every l. Moreover, (cid:107)xil − x(cid:48) l(cid:107) → 0 (as l → ∞). By Theorem 3.4.9, there are λj l s.t l − k−1(cid:88) (cid:107)T x(cid:48) l − m0−1(cid:88) l Sjx(cid:48) λj l(cid:107) = (cid:107)T x(cid:48) l Sjx(cid:48) λj l(cid:107) → 0. j=0 j=0 We claim the λ0 l must converge to some λ0. The argument is the same as that used in Claim 1, except now we obtain (cid:107) m0−1(cid:88) (cid:16) (cid:17) 2l−1(cid:107) → 0 and (cid:107) m0−1(cid:88) Sjx(cid:48) (cid:16) 2l − µj λj l (cid:17) Sjx(cid:48) 2l(cid:107) → 0. 2l−1 − µj λj l j=0 j=0 92 2k−1 + x(cid:48) 2j−1 → 0 and consequently it follows that λ0 2k also lies in Ker Sm0). We apply Sm0−1, noting that Sjx(cid:48) (We note there are no terms of the form 'Sjy' when j ≥ m0 since the RIS (yk) defined by yk := x(cid:48) l = 0 for every l and j ≥ m0. We reach the same contradiction as in the previous argument, i.e. that λ0 2j − λ0 l must converge to some λ0 as claimed. l → 0. We use the same argument as above k − λ0x(cid:48) It easily follows that T x(cid:48) l=1 for j = 1, 2 . . . , m0 − 1 all have convergent subsequences l )∞ to show that the sequences (λj → 0. and consequently that we can find some subsequence x(cid:48) It follows that l −(cid:80)m0−1 j=0 λjSj)x(cid:48) j=1 λj l Sjx(cid:48) lr − m0−1(cid:88) j=0 (cid:107)T xilr λjSjxilr (cid:107) ≤ (cid:107)(cid:0)T − m0−1(cid:88) ≤ (cid:107)T − m0−1(cid:88) j=0 λjSj(cid:1)(xilr − x(cid:48) λjSj(cid:107)(cid:107)xilr − x(cid:48) lr with (T −(cid:80)m0−1 lr )(cid:107) + (cid:107)(cid:0)T − m0−1(cid:88) lr(cid:107) + (cid:107)(cid:0)T − m0−1(cid:88) λjSj(cid:1)(x(cid:48) λjSj(cid:1)(x(cid:48) j=0 lr )(cid:107) lr )(cid:107) → 0. j=0 j=0 A priori, the λj found in Claims 1 and 2 may depend on the RIS. We see now that this is not the case. Claim 3. There are λj ∈ R (0 ≤ j ≤ k − 1) such that whenever (xi)i∈N is a RIS, there is a subsequence (xil) of (xi) such that T xil −(cid:80)k−1 j=0 λjSjxil → 0. Note that if (xi)i∈N is a RIS with some some subsequence converging to 0 then any λj can be chosen satisfying the conclusion of the claim. So it is sufficient to only consider normalised RIS. We begin by choosing a normalised RIS, (xi)i∈N, with Sk−1xi(cid:88)→ 0. Note that such a RIS must exist. Indeed, if not, then Sk−1xi → 0 whenever (xi)i∈N is a RIS and by the argument in the first paragraph of the proof, this would imply that Sk−1 is compact, contradicting Corollary 3.4.8. It follows from Claim 1, after passing to a subsequence and relabelling if necessary, that there are λj ∈ R with λjSjxi → 0. (1) T xi − k−1(cid:88) j=0 j ∈ R with i − k−1(cid:88) j=0 T x(cid:48) Now suppose (x(cid:48) )l∈N, of (x(cid:48) subsequence, (x(cid:48) proving the claim. relabelling if necessary, that there are λ(cid:48) il i)i∈N is any normalised RIS in Xk. We will show that there is some , thus i) such that Equation (1) holds when xi is replaced by x(cid:48) il It follows from Claims 1 and 2, after passing to a subsequence and λ(cid:48) jSjx(cid:48) i → 0. (2) We pick natural numbers i1 < i2 < . . . and j1 < j2 . . . such that max ran xik < min ran x(cid:48) max ran x(cid:48) < min ran xik+1 for every k and such that the sequence (xik + x(cid:48) ≤ )k∈N is again jk jk jk 93 a RIS. For notational convenience, we (once again) relabel the subsequences (xik ), (x(cid:48) (xi) and (x(cid:48) numbers li such that P(0,li](xi + x(cid:48) i) = xi and P(li,∞)(xi + x(cid:48) Claims 1 and 2 that there are µj and a subsequence (xim + x(cid:48) ) by i)i∈N is a RIS and there are natural i. It follows again from i). So (by choice of the subsequences) (xi + x(cid:48) i) = x(cid:48) ) such that jk im (3) j=0 im) − k−1(cid:88) i − k−1(cid:88) (cid:0)T x(cid:48) i − k−1(cid:88) (cid:0)T x(cid:48) j=0 j=0 im) → 0. µjSj(xim + x(cid:48) k−1(cid:88) (cid:1) + (cid:1) → 0 λ(cid:48) jSjx(cid:48) λ(cid:48) jSjx(cid:48) j=0 i i T (xim + x(cid:48) = P(0,li] P(0,li]T x(cid:48) i = P(0,li] λ(cid:48) jP(0,li]Sjx(cid:48) i We note also that and similarly, P(li,∞)T xi → 0. Passing to the appropriate subsequences of Equations (1) and (2) and substracting them from Equation (3) we see that T x(cid:48) im − k−1(cid:88) T xim − k−1(cid:88) j=0 (µj − λj)Sjxim − k−1(cid:88) im − k−1(cid:88) j)Sjx(cid:48) (µj − λ(cid:48) j=0 j=0 j=0 µjSjx(cid:48) im → 0 µjSjxim → 0. (4) (5) and Finally we apply the projections P(0,lim ] and P(lim ,∞) to Equations (4) and (5) respectively to obtain (using the above observations) that k−1(cid:88) k−1(cid:88) (µj − λj)Sjxim → 0 and (µj − λ(cid:48) j)Sjx(cid:48) im → 0. j=0 j=0 j = µj = λj for every j. So, we can replace λ(cid:48) limits implies that we must have λj = µj for all j. We now consider two cases: (i) Sk−1x(cid:48) Since (xi)i∈N was chosen such that Sk−1xi (cid:88)→ 0, Lemma 3.5.2 and the first of the above im (cid:88)→ 0. By Lemma 3.5.2 and the second of the above limits, we see that we im (cid:88)→ 0 for any j < r. i −(cid:80)k−1 j = µj = λj for all j < r. So, replacing λ(cid:48) i → 0. jSjx(cid:48) → 0, so that must have λ(cid:48) we see that Equation (1) does indeed hold with (xi) replaced by (x(cid:48) i). Again, by Lemma 3.5.2, we must have λ(cid:48) by λj for j < r in Equation (2), we find that T x(cid:48) Since Slx(cid:48) Equation (1) holds with xi replaced by x(cid:48) (ii) There is some 1 ≤ r ≤ k − 1 such that Srx(cid:48) → 0 for all l ≥ r, it is now clear that T x(cid:48) i −(cid:80)r−1 j by λj in Equation (2) and → 0, but Sjx(cid:48) j=0 λjSjx(cid:48) −(cid:80)k−1 j=0 λjSjx(cid:48) j=r λ(cid:48) im im im im . j im This completes the proof of the claim and thus (as noted earlier), the proof. 94 3.6 Strict Singularity of S : Xk → Xk In this section we will prove that S is strictly singular. By Proposition 1.4.19, it is enough to see that S is not an isomorphism when restricted to any infinite dimensional block subspace Z of Xk. We begin by stating a result taken from the paper of Argyros and Haydon, [2, Corollary 8.5]. The reader can check that the same proofs as given in [2] will also work for the spaces Xk constructed here. Lemma 3.6.1. Let Z be a block subspace of Xk, and let C > 2 be a real number. Then Z contains a normalized C-RIS. We will need a variation of Lemma 3.4.11 to be able to construct weak dependent sequences. We first observe that the lower norm estimate for skipped block sequences given in Proposition 4.8 of [2] also holds in the space Xk and exactly the same proof works. We state it for here for convenience: Lemma 3.6.2. Let (xr)a that a ≤ n2j and 2j < min ran x2, then there exists an element γ of weight m−1 r=1 be a skipped block sequence in Xk. If j is a positive integer such 2j satisfying a(cid:88) xr(γ) ≥ 1 2 m−1 2j a(cid:88) (cid:107)xr(cid:107). (cid:107)xr(cid:107). Hence r=1 (cid:107) a(cid:88) r=1 xr(cid:107) ≥ 1 2 m−1 2j r=1 a(cid:88) r=1 Lemma 3.6.3. Let ε > 0, j be a positive integer and let (xi)n2j such that min ran x2 > 2j. Suppose further that one of the following hypotheses holds: (i) (cid:107)Sk−1xi(cid:107) ≥ δ for all i (some δ > 0) (ii) There is some 2 ≤ m ≤ k − 1 (where we are, of course, assuming here that k > 2) i=1 be a skipped-block C-RIS, such that (cid:107)Sm−1xi(cid:107) ≥ δ for all i (some δ > 0) and (cid:107)Smxi(cid:107) ≤ Cm−1 2j ε. Then there exists η ∈ Γ such that x(η) ≥ δ 2 where x is the weighted sum n2j(cid:88) x = m2jn−1 2j xi. Moreover, if hypothesis (i) above holds, the pair (Sx, η) is a (16C, 2j, 0)-special exact pair. Otherwise, hypothesis (ii) holds and (Sx, η) is a (16C, 2j, 0, ε) weak exact pair. i=1 95 Proof. Let us consider first the case where hypothesis (i) holds. Since (Sk−1xi)n2j i=1 is a skipped block sequence, it follows from Lemma 3.6.2 that there is an element γ ∈ Γ of weight m−1 2j satisfying n2j(cid:88) n2j(cid:88) m2jn−1 2j Sk−1xi(γ) ≥ 1 2 n−1 2j i=1 i=1 (cid:107)Sk−1xi(cid:107) ≥ δ 2 Consequently, we must have F k−1(γ) being defined, and x(F k−1(γ)) ≥ δ 2 . We set η = η(cid:105) = 0 for any j ≥ 0 (since, by Lemma F k−1(γ) ∈ Γ. Certainly SjSx(η) = (cid:104)x, (S∗)j+1e∗ 3.3.8, we must have F (η) = F k(γ) being undefined). So conditions 3 and 4 are satisfied for (Sx, η) to be a (16C, 2j, 0)-special exact pair. The other conditions are satisfied since a careful examination of the corresponding argument in [2], i.e. the proof of [2, Lemma 6.2], reveals that the remaining conditions will hold for the weighted sum x, regardless of the specific element η of weight m−1 θ ∈ Γ 2j chosen. We then simply make use of the fact that for any and similarly (cid:104)Sx, e∗ θ(cid:105) = (cid:104)Sx, d∗ θ(cid:105) = (cid:40)(cid:104)x, e∗ (cid:40)(cid:104)x, d∗ 0 F (θ)(cid:105) F (θ)(cid:105) if F (θ) is defined otherwise if F (θ) is defined 0 otherwise In the case where hypothesis (ii) holds, we find by the same argument as above that there is a γ ∈ Γ of weight m−1 2j with F m−1(γ) being defined and x(F m−1(γ)) ≥ δ 2 . We now set η = F m−1(γ). Now, for any 0 ≤ j ≤ k − 1, either SjSx(η) = 0 (if (cid:80)n2j F j+1(η) = F j+m(γ) is undefined) or SjSx(η) = x(F j+m(γ)) = Sj+mx(γ) ≤ (cid:107)Sj+mx(cid:107) ≤ m2jn−1 i=1 (cid:107)Sj+mxi(cid:107) ≤ Cε. The final inequality here is obtained using the fact that (cid:107)Sj+mxi(cid:107) ≤ Cm−1 2j ε for all i and j ≥ 0; indeed, when j = 0, this is the hypothesis in the lemma, and when j > 0, we simply use the fact that S has norm at most 1. We have shown conditions (3) and (4(cid:48)) hold for (Sx, η) to be a (16C, 2j, 0, ε) weak exact pair. The remaining conditions hold once again because they hold for x; the argument is the same as 2j before. Theorem 3.6.4. The operator S : Xk → Xk is strictly singular. Proof. We suppose by contradiction that S is not strictly singular. It follows that there is some infinite dimensional block subspace Y of Xk on which S is an isomorphism, i.e. there is some 0 < δ ≤ 1 such that whenever y ∈ Y,(cid:107)Sy(cid:107) ≥ δ(cid:107)y(cid:107). By Lemma 3.6.1, Y contains a normalised skipped block 3-RIS, (xi)i∈N ⊆ Y . We note that certainly Sxi(cid:88)→ 0 and consider two possibilities. Either Sk−1xi → 0 or it does not. In the latter of these possibilities, passing to a subsequence, we can assume without loss of generality that (cid:107)Sk−1xi(cid:107) ≥ ν > 0 96 for every i (and some ν). Thus, we see by Lemma 3.6.3 that we can construct (48, 2j, 0) special exact pairs (Sx, η) for any j ∈ N, with min ran Sx arbitrarily large and x(η) ≥ ν 2 . Otherwise, we must have k > 2 and there is an m ∈ {2, . . . , (k − 1)} with Sm−1xi(cid:88)→ 0 but Smxi → 0. By passing to a subsequence, we can assume that (cid:107)Sm−1xi(cid:107) ≥ ν for all i. Moreover, for a fixed j0 ∈ N, since Smxi → 0, given any j ∈ N, we can find an Nj ∈ N such that (cid:107)Smxi(cid:107) ≤ Cm−1 2j0−1 for every i ≥ Nj. So by Lemma 3.6.3, we can construct weak (48, 2j, 0, n−1 2j0−1) exact pairs (Sx, η) for any j ∈ N, with min ran Sx arbitrarily large and x(η) ≥ ν 2 . 2j n−1 Now, we choose j0, j1 with m2j0−1 > 6720δ−1ν−1 and m4j1 > n2 and the argument above, there is a y1 ∈ Y, η1 ∈ Γ such that (Sy1, η1) is a (48, 4j1, 0, n−1 weak exact pair and y1(η1) ≥ ν be (p1, 0, m2j0−1, e∗ 2j0−1. By Lemma 3.6.3, 2j0−1)- 2 . We let p1 > rank η1 ∨ max ran y1 and define ξ1 ∈ ∆p1 to Now set j2 = σ(ξ1). Again by Lemma 3.6.3 and the argument above, there is y2 ∈ 2j0−1)-weak exact η2), Y, η2 ∈ Γ with min ran y2 > p1, y2(η2) ≥ ν pair. We pick p2 > rank η2 ∨ max ran y2 and take ξ2 to be the element (p2, ξ1, m2j0−1, e∗ noting that this tuple is indeed in ∆p2. 2 and (Sy2, η2) a (48, 4j2, 0, n−1 η1). Continuing in this way, we obtain a (48, 2j0 − 1, 0, n−1 2j0−1)-weak dependent sequence (Syi). By Proposition 3.4.14 we see that with the final inequality following by the choice of j0. On the other hand, n2j0−1(cid:88) i=1 2j0−1 (cid:107)m2j0−1n−1 n2j0−1(cid:88) n2j0−1(cid:88) Syi(cid:107) ≤ 70 × 48m−1 2j0−1 < δν 2 yi(ξn2j0−1) = m−1 2j0−1 i=1 i=1 yi(ηi) ≥ m−1 2j0−1n2j0−1 ν 2 So, (cid:107)m2j0−1n−1 2j0−1 n2j0−1(cid:88) i=1 n2j0−1(cid:88) n2j0−1(cid:88) i=1 Syi(cid:107) ≥ δ(cid:107)m2j0−1n−1 2j0−1 yi(cid:107) ≥ δm2j0−1n−1 2j0−1 yi(ξn2j0−1) ≥ δν 2 This contradiction completes the proof. i=1 The previous theorem combined with Theorem 3.5.3 and Corollary 1.4.27 immediately yield the following corollary. Corollary 3.6.5. The operators Sj : Xk → Xk (j ≥ 1) are strictly singular. The spaces Xk have few but not very few operators. 97 3.7 The HI Property It only remains to see that the spaces Xk are hereditarily indecomposable. The proof is sufficiently close to the corresponding proof of [2] that we will omit most of the details. We first observe we have the following generalisations of Lemmas 8.8 and 8.9 of [2]. Lemma 3.7.1. Let (xi)i≤n2j0−1 be a (C, 2j0 − 1, 1)−weak dependent sequence in Xk and let J be a sub-interval of [1, n2j0−1]. For any γ(cid:48) ∈ Γ of weight m2j0−1 we have (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:88) i∈J (−1)ixi(γ(cid:48)) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ≤ 7C. n2j0−1(cid:88) i=1 It follows that (cid:107)n−1 2j0−1 n2j0−1(cid:88) i=1 xi(cid:107) ≥ m−1 2j0−1 but (cid:107)n−1 2j0−1 (−1)ixi(cid:107) ≤ 70Cm−2 2j0−1. Proof. The proof of the first claim is sufficiently close to the proof of Lemma 3.4.13 that we omit any more details. The upper bound n2j0−1(cid:88) (cid:107)n−1 2j0−1 (−1)ixi(cid:107) ≤ 70Cm−2 2j0−1 then follows using exactly the same argument as in Proposition 3.4.14. The lower bound is proved in the same way as Lemma 8.9 of [2]; one simply observes that using the notation i=1 d∗ ξi + m−1 2j0−1 n2j0−1(cid:88) i=1 ηi(cid:105) (pi−1,pi)e∗ P ∗ of Definition 3.4.12, n2j0−1(cid:88) xi(ξn2j0−1) = (cid:104) i=1 i=1 = m2j0−1 n2j0−1(cid:88) n2j0−1(cid:88) n2j0−1(cid:88) i=1 xi, i=1 We immediately obtain (cid:107)n−1 2j0−1 n2j0−1(cid:88) i=1 2j0−1. xi(ηi) = n2j0−1m−1 n2j0−1(cid:88) xi(cid:107) ≥ n−1 2j0−1 xi(ξn2j0−1) = m−1 2j0−1. i=1 To see the spaces Xk are HI, we claim it will be enough to see that we have the following lemma: Lemma 3.7.2. Let Y be a block subspace of Xk. There exists δ > 0 such that whenever j, p ∈ N, ε > 0, there exists q ∈ N, x ∈ Y, η ∈ Γ with ran x ⊆ (p, q) and (x, η) a (96δ−1, 2j, 1, ε) weak exact pair. 98 We omit the proof of Lemma 3.7.2. It is essentially the same as Lemma 3.6.3 combined with the proof of [2, Lemma 8.6]. Proposition 3.7.3. Xk is hereditarily indecomposable. Proof. The argument is the same as in [2]; by Proposition 1.4.10 it is enough to show that if Y and Z are block subspaces of Xk, then for each ε > 0 there exists a y ∈ Y and z ∈ Z with (cid:107)y − z(cid:107) < ε(cid:107)y + z(cid:107). By Lemma 3.7.2, given two block subspaces Y and Z of Xk there exists some δ > 0 such that for all j0 ∈ N, we can construct (96δ−1, n2j0−1, 1)−weak dependent sequences, (xi)i≤n2j0−1 with xi ∈ Y when i is odd and xi ∈ Z when i is even. We choose j0 ∈ N such that m2j0−1 > 6720δ−1ε−1 and obtain a (96δ−1, n2j0−1, 1)−weak dependent sequence as in the preceding sentence. We define (cid:88) (cid:88) xi y = xi and z = i odd i even n2j0−1(cid:88) i=1 xi(cid:107) ≥ n2j0−1m−1 2j0−1 (−1)ixi(cid:107) ≤ 70 · 96δ−1n2j0−1m−2 2j0−1 and observe that by Lemma 3.7.1 (cid:107)y + z(cid:107) = (cid:107) while n2j0−1(cid:88) (cid:107)y − z(cid:107) = (cid:107) so that (cid:107)y − z(cid:107) < ε(cid:107)y + z(cid:107) as required. i=1 99 Chapter 4 A Banach space with (cid:96)1 Calkin algebra 4.1 The Main Theorem In this chapter, we push the results of the preceding chapter a little further. We construct another (cid:96)1 predual, X∞, which shares some of the properties of the previously constructed spaces Xk; our new space is still an L∞ space of Bourgain-Delbaen type and we will once again obtain a representation formula for all bounded linear operators on X∞. However, the space X∞ also has some radically different properties from the spaces in the previous chapter - the Calkin algebra is isomorphic as a Banach algebra to (cid:96)1(N0), the space fails to have few operators and moreover, the space is indecomposable, but is not hereditarily indecomposable. Our main result is the following: Theorem 4.1.1. There exists a separable L∞ space with a Schauder basis, X∞, that pos- sesses the following properties: 1. X∗∞ = (cid:96)1. 2. There exists a non-trivial, non-compact operator 'S' on X∞. (By 'non-trivial', we simply mean that the operator S is not a scalar multiple of the identity.) The sequences (Sj)∞ equivalent to the canonical basis of (cid:96)1(N0). j=0 ⊂ L(X∞)/K(X∞) are basic sequences isometrically 3. If T ∈ L(X∞) then there are unique scalars (λi)∞ i=0 and a compact operator K ∈ j=1 ⊂ L(X∞) and (cid:0)[Sj](cid:1)∞ L(X∞) with(cid:80)∞ i=0 λi < ∞ and ∞(cid:88) i=0 T = λiSi + K. The operator S appearing in the above sum is the same operator on X∞ described in Property 2 above. 100 4. The Calkin algebra L(X∞)/K(X∞) is isometric as a Banach algebra to the convolution algebra (cid:96)1(N0). For clarity, we remark that when we talk of (cid:96)1(N0) as a Banach algebra, we are thinking of the Banach space (cid:96)1(N0) equipped with the usual multiplication coming from convolution. That is, if (an)∞ n=0 ∈ (cid:96)1(N0) then the product (an)∞ n=0 ∗ (bn)∞ n=0 is defined by n=0, (bn)∞ (cid:0)(an) ∗ (bn)(cid:1)(m) := m(cid:88) ajbm−j. j=0 It turns out that possessing (cid:96)1 Calkin algebra is a rather remarkable property. Before proving the main theorem, we note some corollaries that demonstrate this. Proposition 4.1.2. If X is a Banach space for which the Calkin algebra is isomorphic (as a Banach algebra) to (cid:96)1(N0) then the strictly singular and compact operators on X coincide. A particular consequence of the proposition is, of course, that the strictly singular and compact operators on X∞ coincide. Consequently X∞ fails to have few operators; the operator S cannot be written as a compact perturbation of the identity by property (2) of the theorem, and since the compact and strictly singular operators coincide, S cannot be written as a strictly singular perturbation of the identity. Before giving the proof of Proposition 4.1.2, we document the following easy lemma: j=0 and b = (bj)∞ Lemma 4.1.3. Let a = (aj)∞ j=0 be non-zero vectors in (cid:96)1(N0) and choose k, l ≥ 0 minimal such that ak (cid:54)= 0 and bl (cid:54)= 0. Then a∗ b(j) is equal to 0 whenever j < k + l. When j = k + l, a ∗ b(j) = akbl. In particular, (cid:107)a ∗ b(cid:107)(cid:96)1(N0) ≥ akbl and consequently (cid:107)an(cid:107)(cid:96)1(N0) ≥ akn. Proof. The proof is an easy computation using the definition of the convolution product. Proof of 4.1.2. If T is strictly singular then by Theorem 1.4.28 n→∞(cid:107)[T ]n(cid:107) 1 lim n = lim n→∞(cid:107)T n + K(X)(cid:107) 1 n = 0 n=0 ∈ (cid:96)1(N0) and the above observation yields that limn→∞ (cid:107)an(cid:107) 1 Since the Calkin algebra is isomorphic as a Banach algebra to (cid:96)1(N0), we can identify [T ] with an element a = (an)∞ n = 0. By the previous lemma, if a (cid:54)= 0, then letting k be minimal such that ak (cid:54)= 0, (cid:107)an(cid:107)(cid:96)1 ≥ akn, so that limn→∞ (cid:107)an(cid:107) 1 n ≥ ak > 0. So we must have that a = 0, and therefore that [T ] = 0. This is the same as saying that T is compact. We have therefore shown that every strictly singular operator is compact. The converse is always true. 101 We can also easily deduce from the previous proposition that the space X∞ fails to be hereditarily indecomposable. Proposition 4.1.4. The space X∞ is not hereditarily indecomposable. Proof. It was shown in [12] that for a real HI space X, the quotient algebra L(X)/SS(X) has dimension at most 4 (where, as usual, SS(X) denotes the ideal of strictly singular operators on X). By the previous proposition, SS(X∞) = K(X∞), and so we see that L(X∞)/SS(X∞) is isomorphic to (cid:96)1(N0) by the final property of the main theorem. Com- bining these observations, it is clear that X∞ is not HI. On the other hand, the fact that the Calkin algebra is (cid:96)1(N0) forces the space to be indecomposable. Proposition 4.1.5. If X is a Banach space for which the Calkin algebra is isomorphic (as a Banach algebra) to (cid:96)1(N0) then X is indecomposable. i=0 ∈ (cid:96)1(N0) It follows that (ai)∞ Proof. If P is a projection on X, then [P ] is an idempotent in the Calkin algebra. We identify [P ] with (ai)∞ i=0 is an idempotent in the Banach algebra (cid:96)1(N0). An easy computation yields that the only idempotents in (cid:96)1(N0) are 0 and (1, 0, 0, . . . ). Thus either [P ] = 0 or [I − P ] = 0, i.e. either P or I − P is compact. In either case, P is certainly a trivial projection. 4.2 The Construction The construction of X∞ is almost identical to that used in the previous chapter to obtain the spaces Xk, though the proof of the operator representation for the new space is somewhat more involved. We continue to work with the fast increasing sequences of natural numbers (mj)∞ 3.3.1 for the precise assumptions on these sequences. j=1 from the previous chapter and refer the reader back to Assumption j=1 and (nj)∞ We begin by working with a Bourgain-Delbaen space generated by a countable set Υ = ∪nΛn which is essentially the same as in the construction of the previous chapter except for the addition of some new elements which will have degenerate corresponding BD functionals. Continuing with the terminology from the previous chapter, we will define these additional elements as having age and weight 0 and refer to them as 'trivial elements' since their corresponding BD functionals will just be 0. Let us be more precise. 102 Definition 4.2.1. We define Υ by the recursion Λ1 = {0}, Λn+1 = (n + 1, p, m−1 , b∗) : 0 ≤ p < n, b∗ ∈ Bp,n j n+1(cid:91) ∪ n−1(cid:91) j=1 (cid:110) p(cid:91) (cid:110) (cid:111) (cid:111) (n + 1, ξ, m−1 j , b∗) : ξ ∈ Λp, weight ξ = m−1 j , age ξ < nj, b∗ ∈ Bp,n p=1 j=1 ∪ {(n + 1, j) : 0 ≤ j ≤ n} The notation is exactly as before, whereby elements of Υ code the corresponding BD functional. For a trivial element, γ = (n + 1, j) ∈ Λn+1, we define the corresponding c∗ γ to be 0 and so d∗ γ. It is readily checked that Theorem 2.2.2 still holds with the addition of this degenerate case (indeed, within the notation of the theorem, we could just as well γ = e∗ consider τ (γ) to be the Type 0 tuple, τ (γ) = (1, 0, 0), where the first coordinate sets ε = 1, the second coordinate denotes the 0 scalar, and the third coordinate denotes the element 0 ∈ Λ1). Consequently we obtain a Bourgain-Delbaen space X(Υ). As in the previous chapter, we bring to the attention of the reader the same remark that was made in the previous chapter (see below). We once again ask the reader to consult Chapter 3 for an explanation of the notation being used. Remark 4.2.2. Later on, we will want to take a suitable subset of Υ. To avoid any ambiguity in notation, all linear combinations b∗ =(cid:80) in the above definition, and throughout the rest of the chapter, Bp,n will denote the set of η aη ≤ 1 and each aη is a aηe∗ η∈Υn\Υp η, where, as before,(cid:80) rational number with denominator dividing Nn!. We have the following theorem, which is the analogue of Theorem 3.3.6 and Lemma 3.3.7. Theorem 4.2.3. There is a map G : Υ → Υ ∪ {undefined} (we say G(γ) is undefined if G(γ) = undefined, otherwise we say G(γ) is defined) and a norm 1, linear mapping R∗ : (cid:96)1(Υ) → (cid:96)1(Υ) satisfying: 1. G(0) is undefined (where we recall Λ1 = {0}). 2. G ((n + 1, j)) = (n + 1, j − 1) whenever 1 ≤ j ≤ n and G ((n + 1, 0)) is undefined. 3. For non-trivial elements γ ∈ Υ \ Λ1, if G(γ) is defined, rank γ = rank G(γ) and weight γ = weight G(γ) (i.e. G preserves weight and rank). Moreover, age G(γ) ≤ age γ (G doesn't increase age). 103 4. 5. R∗(e∗ γ) = R∗(d∗ γ) = (cid:40) (cid:40) e∗ G(γ) 0 d∗ 0 G(γ) if G(γ) is defined otherwise. if G(γ) is defined otherwise. Moreover, the dual operator of R∗ restricts to give a bounded linear operator R : X(Υ) → X(Υ) of norm at most 1. We omit the proof, since it is essentially the same as the proof given for the corresponding results in the previous chapter. The following observation about the operator R∗ will be essential later in the proof. Lemma 4.2.4. For every γ ∈ Υ, there is a unique l ∈ N, 1 ≤ l ≤ rank γ such that Gj(γ) is defined whenever 1 ≤ j < l but Gl(γ) undefined. Consequently, (R∗)je∗ γ = 0 whenever j ∈ N, j ≥ rank γ. γ = (R∗)jd∗ Proof. The proof is an easy modification of the proof of Lemma 3.3.8. As with the constructions in [2] and the previous chapter, for our main construction, we will place some restrictions on the elements of odd-weight via the use of a coding function σ : Υ → N. We make the same assumptions on this coding function as in the previous chapter; we refer the reader back to Section 3.3 for the details. As before, the coding function can be constructed recursively as Υ is constructed. For each γ ∈ Υ, we define a finite set Σ(γ) by Σ(γ) := {σ(γ)} ∪ {σ(δ)} ∞(cid:91) (cid:91) j=1 δ∈G−j (γ) where G−j(γ) := {θ ∈ Υ : Gj(θ) = γ}. (In particular, if θ ∈ G−j(γ) then Gl(θ) ∈ Υ for every l ≤ j.) It is easily checked that Lemmas 3.3.10 and 3.3.11 still hold with (essentially) the same proofs and that we have the following modification of Lemma 3.3.12. Lemma 4.2.5. Suppose γ, δ ∈ Υ. 1 ≤ j < ∞ such that Gj(γ) = δ. If σ(γ) ∈ Σ(δ) then either γ = δ or there is some We are now in a position to describe the main construction. The ideas are exactly the same as in Chapter 3. We take a subset Γ ⊂ Υ by placing restrictions on the elements of odd weight we permit. As a consequence of imposing these additional odd weight restrictions, we are also forced to (roughly speaking) remove those elements (n+1, p, β, b∗) and (n+1, ξ, β, b∗) 104 of Υ for which the support of b∗ is not contained in Γ, in order that we can apply the Bourgain-Delbaen construction to obtain a space X(Γ). Exactly as was the case in Chapter 3, the subset Γ is also constructed inductively, so it will be well defined and consistent with the Bourgain-Delbaen method. We will denote by ∆n the set of all elements in Γ having rank n, and denote by Γn the union Γn = ∪j≤n∆j. The permissible elements of odd weight will be as follows. For an age 1 element of odd weight, γ = (n + 1, p, m−1 2j−1, b∗) we insist that either b∗ = 0 or b∗ = e∗ η where η ∈ Γn \ Γp and weight η = m−1 2j−1. For an odd weight element of age > 1, γ = (n + 1, m−1 η where η ∈ Γn \ Γrank ξ and weight η = m−1 Definition 4.2.6. We define recursively sets ∆n ⊆ Λn. Then Γn := ∪j≤n∆j and Γ := ∪n∈N∆n ⊆ Υ. To begin the recursion, we set ∆1 = Λ1. Then 2j−1, ξ, b∗) we insist that either b∗ = 0 or b∗ = e∗ 4k < n−2 2j−1, k ∈ Σ(ξ). Let us be more precise: 4i < n−2 ∆n+1 = (n + 1, p, m−1 2j , b∗) : 0 ≤ p < n, b∗ ∈ Bp,n ∩ (cid:96)1(Γn) (n + 1, ξ, m−1 2j , b∗) : ξ ∈ ∆p, w(ξ) = m−1 (cid:111) 2j , age ξ < n2j, b∗ ∈ Bp,n ∩ (cid:96)1(Γn) (cid:111) j=1 (cid:110) (cid:98)(n+1)/2(cid:99)(cid:91) (cid:110) (cid:98)p/2(cid:99)(cid:91) ∪ n−1(cid:91) (cid:110) (cid:98)(n+2)/2(cid:99)(cid:91) p=1 j=1 ∪ j=1 ∪ n−1(cid:91) (cid:98)(p+1)/2(cid:99)(cid:91) p=1 j=1 (cid:110) (n + 1, p, m−1 2j−1, b∗) : 0 ≤ p < n, b∗ = 0 or b∗ = e∗ η, η ∈ Γn \ Γp (cid:111) 4i < n−2 2j−1 and w(η) = m−1 (n + 1, ξ, m−1 2j−1, b∗) : ξ ∈ ∆p, w(ξ) = m−1 b∗ = 0 or b∗ = e∗ η with η ∈ Γn \ Γp, w(η) = m−1 2j−1, k ∈ Σ(ξ) 2j−1, age ξ < n2j−1 , 4k < n−2 (cid:111) ∪ {(n + 1, j) : 0 ≤ j ≤ n}. Here the Bp,n are defined as in Remark 4.2.2, and, for brevity, we temporarily write w(ξ) for weight ξ. We define X∞ to be the Bourgain-Delbaen space X(Γ) where Γ is the subset of Υ just defined. Once again, to compute norms we need to work with the evaluation functionals e∗ γ. We note that this is no harder than in the previous constructions. Proposition 4.2.7. For a γ ∈ Γ, either it is a trivial, age 0 element (γ = (n + 1, j)) in γ, or age γ ≥ 1 in which case we once again have the which case we simply have e∗ evaluation analysis given by Proposition 3.3.14. γ = d∗ 105 (cid:40) S∗e∗ γ = 0 e∗ F (γ) if F (γ) is undefined otherwise Proof. If γ is a trivial, age 0 element, there is nothing to prove; the statement follows from our earlier definitions. In the case where γ ∈ Γ and age γ ≥ 1, the same proof as before goes through. We still need to show that there exists an operator 'S' on X∞ having all the properties described in the main theorem. We first define the operator. It is constructed in exactly the same way as in Proposition 3.3.16. Proposition 4.2.8. Γ is invariant under G. More precisely, if γ ∈ Γ ⊆ Υ and G(γ) is defined, then G(γ) ∈ Γ. It follows that the map G : Υ → Υ∪{undefined} defined in Theorem 4.2.3 restricts to give F : Γ → Γ ∪ {undefined}. Consequently the map R∗ : (cid:96)1(Υ) → (cid:96)1(Υ) (also defined in 4.2.3) can be restricted to the subspace (cid:96)1(Γ) ⊆ (cid:96)1(Υ) giving S∗ : (cid:96)1(Γ) → (cid:96)1(Γ). S∗ is a bounded linear map on (cid:96)1(Γ) of norm 1 which satisfies for every γ ∈ Γ. Moreover, the dual operator of S∗ restricts to X∞ to give a bounded linear operator S : X∞ → X∞ of norm at most 1, satisfying Sj (cid:54)= 0 for all j ∈ N. Proof. We once again omit the details, as the proof is only a very minor modification of the proof of the original result. Note that Sjd(n+1,0) = d(n+1,j) whenever 0 ≤ j ≤ n so that Sj (cid:54)= 0 for any j ∈ N. The reader can easily verify that Lemma 3.3.17 of the previous chapter still holds for the odd-weight elements of our new construction. Moreover, we can define a C-RIS in the space X∞ using exactly the same definition as in Chapter 3; we refer the reader to Definition 3.4.1. It is then an easy exercise to check that Lemmas and Propositions 3.4.3, 3.4.4 and 3.4.5 all still hold, with only minor modifications required to the original proofs to take into account the new, weight and age 0 elements. In particular, we note that the basis (dγ)γ∈Γ of the space X∞ is shrinking and X∗∞ is naturally isomorphic to (cid:96)1. Moreover, having established that X∗∞ is naturally isomorphic to (cid:96)1, the same proof as in Lemma 3.4.6 works to show that the dual operator of S : X∞ → X∞ is precisely the operator S∗ : (cid:96)1(Γ) → (cid:96)1(Γ) of Proposition 4.2.8. The strategy used to prove our main result will be to follow the arguments of Gowers and Maurey in [15]. We will thus be interested in the seminorm (cid:107) · (cid:107) on L(X∞), defined by (cid:107)T(cid:107) := sup (xn)n∈N∈L where L denotes the set of all 1-RIS, (xn)n∈N. (cid:107)T xn(cid:107) lim sup n We first document some observations about this seminorm. 106 Lemma 4.2.9. Given T ∈ L(X∞), T is compact ⇐⇒ (cid:107)T(cid:107) = 0. Proof. Suppose first that T is compact. Since the basis (dγ)γ∈Γ is shrinking, it follows by Proposition 1.4.15 that if (xn)n∈N ∈ L, (cid:107)T xn(cid:107) → 0. It follows immediately that (cid:107)T(cid:107) = 0. Conversely, suppose (cid:107)T(cid:107) = 0. If (xn)n∈N is a C-RIS, then ( xn C )n∈N ∈ L so (cid:107)T xn(cid:107) = C(cid:107) T xn C (cid:107) → 0. It follows from Proposition 3.4.4 (and our earlier observation that this result still holds for the space X∞), that whenever (yn)n∈N is a block sequence in X∞, (cid:107)T yn(cid:107) → 0. This implies that T is compact, as we saw in Proposition 1.4.15. j=0 λjSj. Then (cid:107)T(cid:107) = (cid:80)N Lemma 4.2.10. Let T = (cid:80)N (cid:80)N j=0 λj. j=0 λj. Moreover, it is an elementary fact that (cid:107)T(cid:107) = (cid:107)T ∗(cid:107) = (cid:107)(cid:80)N (cid:80)N Proof. We already know (cid:107)S(cid:107) ≤ 1; it follows from the triangle inequality that (cid:107)T(cid:107) ≤ j=0 λj(S∗)j(cid:107). Since j=0 λj and moreover (cid:107)T(cid:107) ≥ 1 6 (N + 1, N ) ∈ Γ, we thus have, (cid:107)T(cid:107) = (cid:107) N(cid:88) j=0 λj(S∗)je∗ (N +1,N )(cid:107)(cid:96)1 λje∗ (N +1,N−j)(cid:107)(cid:96)1 = N(cid:88) j=0 λj j=0 λj(S∗)j(cid:107) ≥ (cid:107) N(cid:88) = (cid:107) N(cid:88) (cid:16)(cid:80) j=0 and we obtain the first equality stated in the lemma. εγ ∈ {1,−1} (for γ ∈ A), (cid:80) Lemma 2.2.3. Consequently, (cid:107)(cid:80) To obtain the final inequality stated in the lemma, note that whenever A ⊆ ∆q, and ; this follows immediately from γ∈A εγdγ(cid:107) ≤ 2. Note that (N + n, N − j) ∈ Γ whenever γ∈A εγdγ = iq γ∈A εγeγ (cid:17) n ≥ 1 and 0 ≤ j ≤ N . We consider the sequence (xn)∞ n=1 defined by N(cid:88) j=0 xn = 1 6 sign λj d(N +n,N−j) 3 ≤ 1 for all n. Moreover, we see We claim (xn) ∈ L. By our first observation, (cid:107)xn(cid:107) ≤ 1 that the local support (see [2, Definition 5.7]) of any xn contains no element of weight > 0. Thus by Lemma 5.8 of [2], whenever γ ∈ Γ and weight γ = m−1 for some h, xn(γ) ≤ m−1 h . It follows that (xn)n∈N is a 1-RIS and is therefore in L. (For the sequence (jk) appearing in the definition of a RIS, we take j1 = 1 and jk+1 = max ran xk + 1 for k ≥ 1.) h 107 Finally, we note that (cid:107)T xn(cid:107) ≥ (cid:104)e∗ (N +n,N ), N(cid:88) j=0 λjSjxn(cid:105) = (cid:104)d∗ (N +n,N ), N(cid:88) j=0 λjSjxn(cid:105) = (cid:104) N(cid:88) j=0 λj(S∗)jd∗ (N +n,N ), xn(cid:105) λjsign λk (cid:104)d∗ (N +n,N−j), d(N +n,N−k)(cid:105) N(cid:88) k=0 λj N(cid:88) N(cid:88) j=0 j=0 = = 1 6 1 6 from which it follows that (cid:107)T(cid:107) ≥ lim supn (cid:107)T xn(cid:107) ≥ 1 6 We aim to prove the following theorem: (cid:80)N j=0 λj as required. Theorem 4.2.11. Every operator in L(X∞) is in the (cid:107) · (cid:107)-closure of the set of operators in the algebra A, generated by the Identity and S. As the proof is long and technical, we first see that our main result, Theorem 4.1.1, almost completely follows from Theorem 4.2.11 and the above lemmas. Proof of Theorem 4.1.1. We have already observed that X∗∞ is isomorphic to (cid:96)1; indeed, the proof is almost identical to the proof for the spaces Xk. We note that the first part of Lemma 4.2.10 is equivalent to saying that (Sj)∞ j=0 is a basic sequence in L(X∞), 1-equivalent to the canonical basis of (cid:96)1 = (cid:96)1(N0). Moreover, it follows from the second part of Lemma 4.2.10 and Lemma 4.2.9 that K(X∞) ⊕ [Sj : j ∈ N0] is an algebraic direct sum in L(X∞). We first show that this direct sum is the entire operator algebra. The proof of this is identical to the arguments used in [15] but for completeness we include it here. We write G for the Banach space obtained by taking the (cid:107) · (cid:107)-completion of A, noting that by the second part of Lemma 4.2.10, (cid:107) · (cid:107) actually defines a norm on A. (Here A is defined as in Theorem 4.2.11.) Given T ∈ L(X∞), by Theorem 4.2.11 we can find a (cid:107)·(cid:107)-Cauchy sequence (Tn)∞ n=1 of operators in A such that (cid:107)T − Tn(cid:107) → 0. Let φ(T ) be the n=1 in G. It is easily checked that, in this way, we get a well-defined, bounded limit of (Tn)∞ linear map φ : L(X∞) → G (of norm at most 1). Note that the restriction of φ to A is the identity (or more accurately, the embdedding of A into G). It is also easily seen that the kernel of φ is the set of T such that (cid:107)T(cid:107) = 0, which by Lemma 4.2.9 is precisely the compact operators, K(X∞). Now, we note that we can define a linear isomorphic embedding I : G → L(X∞), with I(G) = [Sj : j ∈ N0]. Precisely, we define I by defining it on the dense subspace of vectors j=0 λjSj] is the vector j=0 λjSj], where it is to be understood that [(cid:80)N in G of the form [(cid:80)N 108 j=0 λjSj] (cid:55)→(cid:80)N in G obtained under the natural embedding A (cid:44)→ G of the vector(cid:80)N i.e. [(cid:80)N j=0 λjSj (recalling that G is simply the (cid:107) · (cid:107)-completion of A). We can therefore well-define I in the obvious way, j=0 λjSj. By the second part of Lemma 4.2.10, this is a continuous linear map, so we may take the unique continuous linear extension to G. Since (cid:107)·(cid:107) ≤ (cid:107)·(cid:107), it is easily seen that this map is an isomorphic embedding onto the closed linear span [Sj : j ∈ N0], as required. We observe that I ◦ φ(T ) − T is in the kernel of φ and is therefore compact. Indeed, it is clear from the definitions of φ and I that φ ◦ I is the identity on the dense subspace of vectors in G of the form [(cid:80)N j=0 λjSj]. Using this observation and continuity, we see that φ (I ◦ φ(T ) − T ) = φ ◦ I (φ(T )) − φ(T ) = φ(T ) − φ(T ) = 0 as required. Consequently, if T ∈ L(X∞) there exists a compact operator K such that T = I ◦ φ(T ) + K and we have L(X∞) = [Sj : j ∈ N0] ⊕ K(X∞) as claimed. Property (3) of Theorem 4.1.1 now follows immediately from the above and the fact j=0 is a basic sequence in L(X∞). Moreover, if we let e : (cid:96)1(N0) → L(X∞) denote that [Sj]∞ j=0 λjSj and let π : L(X∞) → L(X∞)/K(X∞) denote the isometric embedding (λj)∞ the canonical projection, then the map I : (cid:96)1(N0) → L(X∞)/K(X∞) defined by I = π ◦ e is now easily seen to be a continuous bijection; it is therefore an isomorphism by the Inverse j=0 (cid:55)→(cid:80)∞ Mapping Theorem. It is readily checked that this map is a Banach algebra isomorphism. j=0 ⊂ L(X∞)/K(X∞) is a basic sequence also follows from The fact that the sequence ([Sj])∞ the fact that I is an isomorphism. We have very nearly completed the proof. However, it is claimed in property (4) that L(X∞)/K(X∞) is actually isometric to (cid:96)1(N0). We will show in the next section that the sequence [Sj]∞ j=0 is really 1-equivalent to the canonical basis of (cid:96)1, from which it follows that I is really an isometry. It remains for us to prove Theorem 4.2.11, i.e. that every operator in L(X∞) is in the (cid:107) · (cid:107)−closure of the set of operators in the algebra A generated by the identity operator In order to do so, we will need to make further modifications to the definition and S. of a (C, j, 0, ε) weak exact pair (see Definition 3.4.10) and also to the definition of weak dependent sequences (see Definition 3.4.12). Since these definitions are quite complicated, we will include all the details in our new definitions, although most of the conditions we 109 demand are as before. Moreover, unlike in Chapter 3, we won't need 1-exact pairs and dependent sequences in what follows. Consequently we make a very small change in notation to the previous chapter. Definition 4.2.12. Let C, ε > 0 and ω ∈ N ∪ {∞}. A pair (x, η) ∈ X∞ × Γ is said to be a (C, j, ε, ω)-weak exact pair of order ω if 1. (cid:107)x(cid:107) ≤ C ξ, x(cid:105) ≤ Cm−1 2. (cid:104)d∗ 3. weight η = m−1 4. Slx(η) ≤ Cε for every 0 ≤ l ≤ ω if ω < ∞ and Slx(η) ≤ Cε for every 0 ≤ l < ω if for all ξ ∈ Γ j j ω = ∞ 5. for every element η(cid:48) of Γ with weight η(cid:48) = m−1 Cm−1 Cm−1 x(η(cid:48)) ≤ (cid:40) i i j (cid:54)= m−1 j , we have if i < j if i > j. As before, we observe that we have the following estimates for weak exact pairs (for exactly the same reasons as in [2] and as in the previous chapter). Remark. A (C, j, ε, ω) weak exact pair also satisfies the estimates (cid:40) (cid:104)e∗ η(cid:48), P(s,∞)x(cid:105) ≤ 6Cm−1 6Cm−1 i j if i < j if i > j for elements η(cid:48) of Γ with weight η(cid:48) = m−1 i (cid:54)= m−1 j . We will need the following method for constructing 0 weak exact pairs of order N . Lemma 4.2.13. Let ω ∈ N ∪ {∞}, (xk)n2j k=1 be a skipped-block C-RIS, and let q0 < q1 < q2 < ··· < qn2j be natural numbers such that q1 ≥ 2j and ran xk ⊆ (qk−1, qk) for all k. Let z denote the weighted sum z = m2jn−1 k be an element of Bqk−1,qk−1 with (cid:104)(S∗)lb∗ k, Slxk(cid:105) ≤ Cε for all 0 ≤ l ≤ ω (or 0 ≤ l < ∞ if ω = ∞).Then there exist ζi ∈ ∆qi (1 ≤ i ≤ n2j) such that the element η = ζn2j has analysis (qi, b∗ (cid:80)n2j k=1 xk. For each k let b∗ i , ζi)1≤i≤n2j and (z, η) is a (16C, 2j, ε, ω)-special exact pair. k, xk(cid:105) = (cid:104)b∗ 2j The proof of the lemma is sufficiently close to the proof of Lemma 3.4.11 that we omit it. 110 Definition 4.2.14. We shall say that a sequence (xi)i≤n2j0−1 in X∞ is a (C, 2j0− 1, 0)-weak dependent sequence (of finite order) if there exist sequences of natural numbers 0 = p0 < p1 < p2 < ··· < pn2j0−1 and 1 = N0 < N1 < N2 < ··· < Nn2j0−1 < ∞ together with ηi ∈ Γpi−1 \ Γpi−1 and ξi ∈ ∆pi (1 ≤ i ≤ n2j0−1) such that 1. for each k, ran xk ⊆ (pk−1, pk) 2. for each k, Sjxk(ηk) = 0 whenever j ≥ Nk has weight m−1 2j0−1 and analysis (pi, e∗ ηi, ξi) n2j0−1 i=1 3. the element ξ = ξn2j0−1 of ∆pn2j0−1 4. (x1, η1) is a (C, 4j1, n−1 5. for each 2 ≤ i ≤ n2j0−1, (xi, ηi) is a (C, 4σ(ξi−1), n−1 2j0−1, N0)-weak exact pair 2j0−1, Ni−1)-weak exact pair. In certain applications, we can remove the more complicated conditions involving the se- quence (Nj) dependent sequence (of infinite order) if we remove condition (2) above, and ask that all Specifically, we will say the sequence (xi)i≤n2j0−1 is a (C, 2j0 − 1, 0)-weak n2j0−1 j=0 the weak exact pairs occurring in (4) and (5) have infinite order. In either case, we notice that, because of the special odd-weight conditions, we necessar- 2j0−1, by Lemma 3.3.17 2j0−1, and weight ηi+1 = m−1 = weight η1 < n−2 4σ(ξi) < n−2 ily have m−1 for 1 ≤ i < n2j0−1. 4j1 The following two lemmas are fundamental to proving our main result. Lemma 4.2.15. Let (xi)i≤n2j0−1 be a weak (C, 2j0 − 1, 0)-dependent sequence (of finite or infinite order) in X∞ and let J be a sub-interval of [1, n2j0−1]. For any γ(cid:48) ∈ Γ of weight m2j0−1 we have (cid:88) i∈J xi(γ(cid:48)) ≤ 7C. Proof. The proof is almost the same as the proof of Lemma 3.4.13. We recall the notation used; we let ξi, ηi, pi, Ni (if we are considering a dependent sequence of finite order), j1 be as in the definition of a dependent sequence and let γ denote ξn2j0−1, an element of weight is either 0 or i)1≤i≤a(cid:48)(cid:1) the analysis of γ(cid:48) where each b(cid:48)∗ m2j0−1. We denote by(cid:0)p(cid:48) 0, (p(cid:48) i, b(cid:48)∗ i , ξ(cid:48) i e∗ η(cid:48) i for some suitable η(cid:48) i. We recall that we can estimate as follows (cid:88) k∈J xk(γ(cid:48)) ≤ (cid:88) k∈J, k<l xk(γ(cid:48)) + xl(γ(cid:48)) + (cid:88) k∈J, k>l xk(γ(cid:48)), where we are assuming we have chosen l maximal such that there exists i with b(cid:48)∗ and weight η(cid:48) i = e∗ η(cid:48) i = weight ηl. (We refer the reader back to the proof of Lemma 3.4.13 for the i 111 details on reducing to this case.) The terms xl(γ(cid:48)) and(cid:80) The argument for estimating the remaining term, (cid:80) in exactly the same way as in the proof of Lemma 3.4.13. k∈J, k<l xk(γ(cid:48)), is a little more involved. Obviously if l = 1, the sum we are trying to estimate is just zero, and the lemma is proved. So we can suppose l > 1. By definition of l, there exists some i such that b(cid:48)∗ i = e∗ η(cid:48) and weight ηl = weight η(cid:48) i. Now either i = 1 or i > 1. When i = 1, we can use exactly the same argument as used for the corresponding case in the proof of Lemma 3.4.13 to conclude i k∈J, k>l xk(γ(cid:48)) can be estimated that (cid:80) k∈J, k<l xk(γ(cid:48)) ≤ 3C and we are done. i = weight ηl. But by definition of an exact pair, we have weight ηl = m−1 It only remains to consider what happens when i > 1. Recall we are also assuming l > 1 4σ(ξl−1), i−1). By i−1). By Lemma i−1. We note that in particular i−1 since F preserves rank and we can write the evaluation analysis of and weight η(cid:48) and by restrictions on elements of odd weights, weight η(cid:48) strict monotonicity of the sequence mj, we deduce that ω = σ(ξl−1) ∈ Σ(ξ(cid:48) 4.2.5, there is some j, 0 ≤ j < ∞, such that F j(ξl−1) = ξ(cid:48) this implies pl−1 = p(cid:48) γ(cid:48) as 4ω with ω ∈ Σ(ξ(cid:48) i = m−1 2j0−1P ∗ (p(cid:48) r−1,p(cid:48) Now, for k < l, since ran xk ⊆ (pk−1, pk) ⊆ (0, pl−1) = (0, p(cid:48) γ(cid:48) = (S∗)j(e∗ e∗ + m−1 d∗ ξ(cid:48) ) + ξl−1 r=i r r)b(cid:48)∗ r . i−1), we see that a(cid:48)(cid:88) (cid:104)xk, e∗ γ(cid:48)(cid:105) = (cid:104)xk, (S∗)je∗ l−1(cid:88) ξl−1 ξr + m−1 d∗ (cid:105) = (cid:104)Sjxk, = m−1 r=1 2j0−1(cid:104)Sjxk, e∗ ηk 2j0−1P ∗ (cid:105) ≤ (cid:104)Sjxk, e∗ ηr(cid:105) (pr−1,pr)e∗ (cid:105). ηk Now, if (xi)i≤n2j0−1 is a dependent sequence of infinite order, we have Slxk(ηk) ≤ Cn−1 for every l ∈ N0, so certainly (cid:104)Sjxk, e∗ (cid:105) ≤ Cn−1 2j0−1 and consequently, ηk 2j0−1 (cid:88) k∈J, k<l xk(γ(cid:48)) ≤ n2j0−1 max k∈J, k<l xk(γ(cid:48)) ≤ C. Otherwise (xi)i≤n2j0−1 is a dependent sequence of finite order, and we estimate by consider- ing three possibilities. If j < N1 ≤ Nk−1 for all k ≥ 2, then by definition of weak dependent sequence, (cid:104)Sjxk, η∗ 2j0−1 for k ≥ 2. So, k(cid:105) ≤ Cn−1 xk(γ(cid:48)) ≤ x1(γ(cid:48)) + n2j0−1 max xk(γ(cid:48)) ≤ 2C. k∈J, 2≤k<l (cid:88) (cid:105) = 0 and so (cid:80) k∈J, k<l 112 If j > Nn2j0−1 ≥ Nk for all k, then condition (2) of weak dependent sequence implies that (cid:104)Sjxk, e∗ k∈J, k<l xk(γ(cid:48)) = 0. ηk The only remaining possibility is that there is some r ∈ {1, 2, . . . , n2j0−1 − 1} with Nr ≤ j ≤ Nr+1. Now if k ≤ r, it follows that j ≥ Nr ≥ Nk and so Sjxk(ηk) = 0 If k ≥ r + 2, j ≤ Nr+1 ≤ Nk−1 and so by condition (2) of weak dependent sequence. (cid:104)Sjxk, e∗ (cid:105) ≤ Cn−1 2j0−1. It follows that xk(γ(cid:48)) ≤ xr+1(γ(cid:48)) + (cid:88) xk(γ(cid:48)) + (cid:88) xk(γ(cid:48)) ηk (cid:88) k∈J, k<l k∈J, k<l, k≤r ≤ (cid:107)xr+1(cid:107) + 0 + n2j0−1 max k∈J, k<l, k≥r+2 k∈J, k<l, k≥r+2 (cid:104)Sjxk, e∗ (cid:105) ≤ 2C. ηk This completes the proof. Lemma 4.2.16. If (xi)1≤i≤n2j0−1 is a (C, 2j0 − 1, 0) weak dependent sequence (of finite or infinite order), then (cid:107)n−1 2j0−1 xk(cid:107) ≤ 70Cm−2 2j0−1. n2j0−1(cid:88) k=1 and moreover, if γ ∈ Γ, weight (γ) = m−1 28Cn−1 28Cn−1 n2j0−1(cid:88) xk(γ) ≤ n−1 2j0−1 (cid:40) h , then 2j0−1 + 84Cm−2 2j0−1 + 42Cm−1 h k=1 2j0−1m−1 h if h < 2j0 − 1 if h > 2j0 − 1. Proof. The proof follows from Lemma 4.2.15 (readjusting the RIS constant to 7C), and by the proof of [2, Proposition 5.6]. In the spirit of [15], we aim to prove the following proposition. Proposition 4.2.17. Let T : X∞ → X∞ be a bounded linear operator. Then for every δ > 0, there exists an l ∈ N such that whenever x ∈ X∞ is a block vector satisfying (cid:107)x(cid:107) ≤ 1, min ran x > l and x(γ) ≤ m−1 i whenever weight γ = m−1 i with i < l, dist(T x, l conv{λSjx : j ∈ {0, 1, . . . , l},λ = 1}) ≤ δ. In order to prove the above proposition, we will need the following lemmas, which are analogous to Lemma 7.1 of [2] and Lemma 3.4.17. Lemma 4.2.18. Let m, n and l be natural numbers with m < n and let x, y ∈ X∞ be such that ran x, ran y are both contained in the interval (m, n]. Suppose that dist(y, l conv{λSjx : j ∈ {0, 1, . . . , l},λ = 1}) > δ. Then there exists b∗ ∈ ball (cid:96)1(Γn \ Γm), with rational coordinates, such that (cid:104)Sjx, b∗(cid:105) ≤ (cid:107)y(cid:107) for every j ∈ {0, 1, . . . , l} and b∗(y) > 1 2 δ. l 113 Proof. For 0 ≤ j ≤ l, let uj ∈ (cid:96)∞(Γn\Γm) be the restriction of Sjx, and let v ∈ (cid:96)∞(Γn\Γm) be the restriction of y. Then Sjx = inuj for 0 ≤ j ≤ l, y = inv and for any scalars (µj)l j=0, j=0 µj ≤ l, δ < (cid:107)y −(cid:80)l j=0 µjSjx(cid:107) ≤ (cid:107)in(cid:107)(cid:107)v −(cid:80)l j=0 µjuj(cid:107). Hence with(cid:80)l dist(v, l conv{λuj : j ∈ {0, 1, . . . , l},λ = 1}) > 1 2 δ. Let Cl = conv{λuj : j ∈ {0, 1, . . . , l},λ = 1}. By the Hahn -- Banach Theorem (see Theorem 1.4.7) in the finite dimensional space (cid:96)∞(Γn \ Γm), there exists a∗ ∈ sphere (cid:96)1(Γn \ Γm) with (cid:26) (cid:27) sup (cid:104)a∗, z(cid:105) : z ∈ Cl + 1 2 δB((cid:96)∞(Γn \ Γm) < (cid:104)a∗, v(cid:105). It follows that (cid:104)a∗, v(cid:105) > 1 2 δ < (cid:104)a∗, v(cid:105) ≤ (cid:107)v(cid:107) ≤ (cid:107)y(cid:107). Consequently, (cid:104)a∗, uj(cid:105) < for every 0 ≤ j ≤ l. To complete the proof, we note that by an elementary continuity argument, we may approximate a∗ arbitrarily well by b∗ ∈ ball (cid:96)1(Γn \ Γm) ∩ QΓn\Γm retaining the desired conditions. 2 δ and supa∗(Cl) < supa∗(Cl) + 1 (cid:107)y(cid:107) l The proof of Proposition 4.2.17 will be easy once we have proved the following lemma. Lemma 4.2.19. Let T be a bounded linear operator on X∞. Suppose for this T , there exists some δ > 0 for which the conclusion of Proposition 4.2.17 fails. Then, given any ε > 0 and natural numbers, j, p, N , there exists a block vector z ∈ X∞, a natural number q > p, and an η ∈ ∆q such that ran z ⊆ (p, q) and (1) (z, η) is a (16, 2j, ε, N )-exact pair; 16 δ; (2) (T z)(η) > 7 (3) (cid:107)(I − P(p,q))T z(cid:107) < m−1 2j δ; (4) (cid:104)P ∗ η, T z(cid:105) > 3 8 δ. (p,q]e∗ Proof. Note that under the hypothesis of the lemma, we can obtain a skipped block 1-RIS (xk)k∈N with the property that for every k dist(T xk, k conv{λSjxk : j ∈ {0, 1, . . . , k},λ = 1} > δ. We now fix L ∈ N such that L ≥ N and L > 4(cid:107)T(cid:107)/ε. Observe that if m < n and x ∈ X∞, then m conv{λSjx : j ∈ {0, 1, . . . , m},λ = 1} ⊆ n conv{λSjx : j ∈ {0, 1, . . . , n},λ = 1} so, for k ≥ L, dist(T xk, L conv{λSjxk : j ∈ {0, 1, . . . , L},λ = 1}) ≥ dist(T xk, k conv{λSjxk : j ∈ {0, 1, . . . , k},λ = 1} > δ. 114 We pass to the subsequence (xk)k≥L so that the above inequality holds. Note also that the sequence (T xk) is weakly null (since the basis (dγ)γ∈Γ, is shrinking). Consequently, passing to a further subsequence if necessary, we can assume that there exist p < q0 < q1 < q2 < . . . such that, for all k ≥ 1, ran xk ⊆ (qk−1, qk) and (cid:107)(I − P(qk−1,qk))T xk(cid:107) < 1 2j δ ≤ 1280 δ. It follows from this that dist(P(qk−1,qk)T xk, L conv{λSjxk : j ∈ {0, 1, . . . , L},λ = 1}) > 1279 k ∈ ball (cid:96)1(Γqk−1 \ Γqk−1), with rational coordinates, satisfying 1280 δ. We may apply Lemma 4.2.18 to obtain b∗ 2j δ ≤ 1 80 m−1 5 m−2 1 k, Sjxk(cid:105) ≤ (cid:107)P(qk−1,qk)T xk(cid:107) (cid:104)b∗ L ≤ 4(cid:107)T(cid:107) L < ε whenever 0 ≤ j ≤ L, and (cid:104)b∗ k, P(qk−1,qk)T xk(cid:105) > 1279 2560 δ. k have denominators dividing Nqk−1!, so that b∗ Taking a further subsequence if necessary (and redefining the qk), we may assume that k ∈ Bqk−1,qk−1, and we the coordinates of b∗ may also assume that q1 ≥ 2j. We are thus in a position to apply Lemma 4.2.13, getting elements ξk of weight m−1 has evaluation analysis in ∆qk such that the element η = ξn2j of ∆qn2j 2j n2j(cid:88) e∗ η = d∗ ξi + m−1 2j P ∗ (qi−1,qi)b∗ i . i=1 i=1 n2j(cid:88) n2j(cid:88) z = m2jn−1 2j xi. and such that (z, η) is a (16, 2j, ε, L)-weak exact pair, where z denotes the weighted average Since we also chose L ≥ N , (z, η) is certainly a (16, 2j, ε, N )-weak exact pair as required. We set q = qn2j ; the rest of the proof is now exactly the same as in [2, Lemma 7.2]. i=1 Proof of Proposition 4.2.17. We suppose for contradiction that there exists δ > 0, such that the conclusion of the Proposition 4.2.17 fails. The argument is the same as Proposition 7.3 of [2]. We construct a (16, 2j0 − 1, 0)- weak dependent sequence of finite order (for a suitably chosen j0 ∈ N) by making repeated applications of Lemma 4.2.19. We begin by choosing j0 such that m2j0−1 > 4480(cid:107)T(cid:107)δ−1 and j1 such that m4j1 > n2 Taking p = p0 = 0 and j = 2j1, N = N0 = 1, ε = n−1 q1 and a (16, 4j1, n−1 (cid:107)(I − P(0,q1])(T z1)(cid:107) < m−1 of ∆p1 given by ξ1 = (p1, m−1 Slz1(η1) = 0 whenever l ≥ N1. 2j0−1. 2j0−1 in Lemma 4.2.19, we can find 8 δ and 4j1−2δ. Let p1 = q1 + 1 and let ξ1 be the special Type 1 element η1). By Lemma 4.2.4, we can find an N1 > N0 such that 2j0−1, N0)-exact pair (z1, η1) with rank η1 = q1, (T z1)(η1) > 3 2j0−1, e∗ 115 Now, recursively for 2 ≤ i ≤ n2j0−1, define ji = σ(ξi−1), and use Lemma 4.2.19 again to choose qi and a (16, 4ji, n−1 2j0−1, Ni−1)-exact pair (zi, ηi) with rank ηi = qi, ran zi ⊆ (pi−1, qi), (cid:104)P ∗ δ. We now define pi = qi + 1, choose Ni such that Slxi(ηi) = 0 whenever l ≥ Ni (which is again possible by 4.2.4), and let ξi be the Type 2 element (pi, ξi−1, m−1 8 δ and (cid:107)(I − P(pi,qi])(T zi)(cid:107) < m−1 ηi, T zi(cid:105) > 3 (pi−1,qi]e∗ 4ji ηi) of ∆pi. 2j0−1, e∗ It is clear that we have constructed a (16, 2j0−1, 0)-weak dependent sequence (zi)1≤i≤n2j0−1. We set z = n−1 other hand, the same argument as in the proof of Proposition 7.3 of [2] yields that zi and note that (cid:107)z(cid:107) ≤ 70 · 16m−2 2j0−1 by Lemma 4.2.16. On the 2j0−1 (cid:80)n2j0−1 i=1 (cid:107)T z(cid:107) > m−1 2j0−1δ 4 ≥ δm2j0−1(cid:107)z(cid:107) 4 · 70 · 16 which implies that (cid:107)T(cid:107) > δm2j0−1 4480 , contradicting m2j0−1 > 4480(cid:107)T(cid:107)δ−1. Proposition 4.2.20. Let T be a bounded linear operator on X∞, δ > 0. For this δ, choose l ∈ N as given by Proposition 4.2.17 and let Al = l conv{λSj : j ∈ {0, 1, . . . , l},λ = 1}. Then, there exists U ∈ Al with (cid:107)T − U(cid:107) ≤ 28, 674δ. For the proof, we will need the following fixed-point theorem due to Kakutani. We refer the reader to [19] for the proof and more details. Theorem 4.2.21 (Kakutani Fixed Point Theorem). Let S be an r-dimensional closed simplex and denote by R(S) the set of all non-empty, closed, convex subsets of S. If Φ : S → R(S) is upper semi-continuous, then there exists an x0 ∈ S such that x0 ∈ Φ(x0). We remind the reader of the following definition from [19]: Definition 4.2.22. With the notation as in the preceding theorem, a point-to-set mapping Φ : S → R(S) is called upper semi-continuous if whenever (xn)∞ n=1 is a sequence in S converging to x0 ∈ S, yn ∈ Φ(xn) and yn → y0 ∈ S, then y0 ∈ Φ(x0). Proof of 4.2.20. The proof follows closely that of [15, Lemma 9]. We find it convenient to introduce the following piece of notation. Given a sequence of block vectors (xi)i∈N, we write x1 < x2 < . . . if there exist natural numbers q0 < q1 < . . . with ran xi ⊆ (qi−1, qi) for every i (with similar notation if the sequence has only finite length). It is sufficient to show that if T is an operator on X∞, with matrix representation with respect to the basis (dγ)γ∈Γ possessing the property that there are only finitely many non- zero entries in each row and column, then there is U ∈ Al with (cid:107)T − U(cid:107) ≤ 28, 673δ. (We temporarily call this property P so that we can refer back to this statement.) Indeed, since the basis (dγ)γ∈Γ is shrinking, T can be perturbed by a distance at most δ in operator norm 116 to an operator of this form, and the seminorm, (cid:107) · (cid:107), is bounded above by the operator norm. Having performed the described perturbation, it follows that we may assume T sends block vectors (with respect to the FDD) to block vectors. Note we can also assume (without loss of generality) that if N ∈ N is given, whenever x ∈ X∞ is a block vector with min ran x sufficiently large, min ran T x > N . Indeed, if this were not the case, an elementary argument involving the fact that the matrix of T is assumed to have only finitely many non-zero entries in each row and column would imply that T is in fact a finite-rank, hence compact, operator. We could thus take U = 0 and the proof would be complete by Lemma 4.2.9. Now, if property P is false, then for every U ∈ Al, there exists a 1-RIS (xU (i))i∈N with lim supi (cid:107)(T − U )xU (i)(cid:107) > 28, 673δ. Since subsequences of 1-RIS are again 1-RIS, passing to a subsequence if necessary, we see that for every U ∈ Al, there exists a skipped block 1-RIS (xU (i))i∈N such that (cid:107)(T − U )xU (i)(cid:107) > 28, 673δ for all i. Note that Al is compact. L(X∞), defined by (λ0, λ2, . . . , λl) (cid:55)→(cid:80)l Indeed, consider the (continuous) linear map h : (cid:96)l+1 j=0 λjSj. Then the restriction of h to l B((cid:96)l+1 1 → ) is j=1 of Al by open sets of j=1 be a partition of unity on Al with φj supported inside Uj 1 a homeomorphism onto Al. We can thus choose a covering (Uj)k diameter less than δ. Let (φj)k for each j. For every j = 1, . . . , k, let Uj ∈ Uj and let (xj,i)i∈N be a skipped block 1-RIS with the above property with U = Uj. By the condition on the diameter of Uj, we have (cid:107)(T − U )xj,i(cid:107) > 28, 672δ for all i ∈ N and U ∈ Uj. Moreover, by the remarks made earlier (and the fact that ran x ⊂ (p, q) =⇒ ran Sx ⊆ (p, q)), passing to a subsequence of (xj,i)i∈N if necessary, we may also assume that ((T − U )xj,i)i∈N is a skipped block sequence of successive vectors for every U ∈ Al. For the rest of the proof, we work with j0 ∈ N chosen large enough that (1) 2j0 − 1 > l (2) m2j0n−1 (3) kn2j0−1 ≤ n2j0. 2j0−1 ≤ m−1 l Note that it is certainly possible to choose such a j0. Indeed, by Assumptions (2) and (4) on the sequences (mj) and (nj) (see Assumption 3.3.1), m2jn−1 2j−2 → 0 as j → ∞ and nj+1n−1 j+1 → ∞ as j → ∞. Now, suppose we are given ε > 0, p, r ∈ N. We can select a skipped block 1-RIS of length n2r, (xj,i)M +n2r i=M +1 (M suitably chosen) from the 1-RIS (xj,i)i∈N, such that there are natural numbers p < q0 < q1 < ··· < qn2r with q1 ≥ 2r and ran xj,M +i ⊆ (qi−1, qi). Now, setting j ≥ m2 2j−1n−1 2j−1 ≤ n−1 2j−1 = m2 117 i = 0 ∈ Bqi−1,qi for all i, we can apply Lemma 4.2.13 to see that there exists q = qn2r and b∗ a (16, 2r, ε,∞)-weak exact pair (z, η) where z is the weighted sum z = m2rn−1 2r xj,M +i n2r(cid:88) i=1 and η ∈ ∆q. Note also that ran z ⊆ (p, q) and whenever U ∈ Uj, (cid:107)(T − U )z(cid:107) = m2rn−1 (T − U )xj,M +i(cid:107) 2r (cid:107) n2r(cid:88) i=1 ≥ 1 2 m−1 2r m2rn−1 2r n2r(cid:88) i=1 (cid:107)(T − U )xj,M +i(cid:107) > 14, 336δ where the penultimate inequality follows from Lemma 3.6.2 and the assumption we made earlier about ((T − U )xj,i)i∈N being a skipped block sequence of successive vectors for every U ∈ Al. Using this observation, for each j = 1, . . . , k, we may inductively construct a (16, 2j0 − , with min ran zj,1 > l and (cid:107)(T − 1, 0) weak dependent sequence of infinite order, (zj,i) U )zj,i(cid:107) > 14, 336δ whenever U ∈ Uj, 1 ≤ i ≤ n2j0−1. Moreover, once again making use of the assumptions on T discussed at the beginning of the proof, we may arrange that n2j0−1 i=1 (T − U )zj,1 < (T − U )zj,2,··· < (T − U )zj,n2j0−1 for every j and U ∈ Al and moreover that (T − U )zj,n2j0−1 < (T − U )zj+1,1 1792 m2j0n−1 for every 1 ≤ j < k, and U ∈ Al. Now, for j = 1, . . . , k, we set yj = 1 Observe that min ran yj > l. Moreover, by Lemma 4.2.16, we have (cid:107)yj(cid:107) ≤ 1 16m−2 2j0−1 = 1120 1792 and if γ ∈ Γ, weight γ = m−1 m2j0n−1 2j0−1 + m−1 1344 1792 yj(γ) ≤ 28 × 16 m−1 1792 ≤ 448 1792 l + h h with h < 2j0 − 1, m2j0m−2 84 × 16 1792 2j0−1m−1 h (cid:80)n2j0−1 2j0−1 1792 m2j0 × 70 × i=1 zj,i. For U ∈ Al, let y(U ) =(cid:80)k by the choice of j0. Since j0 was chosen so that l < 2j0 − 1, if h < l < 2j0 − 1, m−1 and we have yj(γ) ≤ m−1 h . j=1 φj(U )yj. We have min ran y(U ) > l,(cid:107)y(U )(cid:107) ≤ 1120 1792 ≤ 1 and y(U )(γ) ≤ m−1 h with h < l (since we have just observed these facts are true when y(U ) is replaced by any yj). We show that y(U ) is also a 'bad vector' for U , by showing that (cid:107)(T − U )y(U )(cid:107) > 4δ. h whenever γ ∈ Γ and weight γ = m−1 l ≤ m−1 h 118 14, 336δ whenever j ∈ J and 1 ≤ i ≤ n2j0−1. Observe that (T − U )y(U ) =(cid:80) To see this, let U ∈ Al be fixed, and let J = {j : φj(U ) > 0}, noting that (cid:107)(T −U )zj,i(cid:107) > j∈J φj(U )(T − φj(U )(T −U )zj,i. By the way in which we constructed U )yj = 1 the zj,i, the final sum is a sum of at most kn2j0−1 ≤ n2j0 skipped, successive vectors. It follows that (cid:80)n2j0−1 (cid:80) j∈J i=1 2j0−1 1792 m2j0n−1 n2j0−1(cid:88) (cid:107)(cid:88) j∈J i=1 φj(U )(T − U )zj,i(cid:107) ≥ 1 2 m−1 2j0 (cid:107)φj(U )(T − U )zj,i(cid:107) (cid:88) j∈J φj(U ) (cid:88) n2j0−1(cid:88) j∈J i=1 > m−1 2j0 = m−1 2j0 n2j0−17168δ n2j0−17168δ and consequently, (cid:107)(T − U )y(U )(cid:107) > 4δ as claimed. The proof of the proposition now follows by the Kakutani Fixed Point Theorem, The- Indeed, for each U ∈ Al, let Φ(U ) be the set of V ∈ Al such that (cid:107)(T − orem 4.2.21. V )y(U )(cid:107) ≤ 4δ. Clearly Φ(U ) is a closed, convex subset of Al, Moreover, by Proposition 4.2.17, and our earlier observations about y(U ), Φ(U ) is non-empty for every U ∈ Al. Clearly, the function Al → X∞ defined by U (cid:55)→ y(U ) is continuous, and this ensures that Φ is upper semi-continuous. Indeed, suppose Un ∈ Al, Un → U , Vn ∈ Φ(Un) and Vn → V . Then (cid:107)(T − V )y(U )(cid:107) ≤ (cid:107)(T − Vn)(cid:0)y(U ) − y(Un)(cid:1)(cid:107) + (cid:107)(T − Vn)y(Un)(cid:107) + (cid:107)(Vn − V )y(U )(cid:107). The first term on the right-hand-side of the inequality converges to 0 by continuity of U (cid:55)→ y(U ) and the fact that Un → U . The third term converges to 0 since Vn → V . The second term is at most 4δ since Vn ∈ Φ(Un). It follows that (cid:107)(T − V )y(U )(cid:107) ≤ 4δ so that V ∈ Φ(U ) as required. Consequently, by the fixed point theorem, there exists some U ∈ Al with U ∈ Φ(U ); this contradicts the fact that y(U ) is a 'bad' vector for U , as was shown earlier. We conclude this section by making the obvious observation that the proof of Theorem 4.2.11 is now an immediate corollary of Proposition 4.2.20. 4.3 L(X∞)/K(X∞) is isometric to (cid:96)1(N0) We have already seen that the Calkin algebra, L(X∞)/K(X∞), is isomorphic to (cid:96)1(N0). In the final section of this chapter, we complete the proof of the main theorem and show that the algebras are in fact isometric. As observed in the proof of Theorem 4.1.1, it will be enough for us to show that the sequence ([Sj])∞ (cid:96)1. j=0 is 1-equivalent to the canonical basis of 119 Lemma 4.3.1. For any x ∈ X∞, (cid:107)Snx(cid:107) → 0 as n → ∞. Proof. Observe that if x is a block vector with ran x ⊆ (p, q) then Snx = 0 whenever n ≥ q. Indeed, since the vectors (d∗ γ)γ∈Γ form a basis for the dual of X∞, it is enough to show (cid:104)d∗ γ, Snx(cid:105) = 0 whenever γ ∈ Γ and n ≥ q. Since ran x ⊆ (p, q) =⇒ ran Snx ⊆ (p, q) for γ, Snx(cid:105) = 0 for every n ∈ N and γ ∈ Γ with rank γ ≥ q. For every n, we certainly have (cid:104)d∗ γ ∈ Γ with rank γ < q, we have (S∗)nd∗ γ = 0 whenever n ≥ q by Lemma 4.2.4, and so (cid:104)d∗ γ, x(cid:105) = 0 as required. γ, Snx(cid:105) = (cid:104)(S∗)nd∗ Now for a general x ∈ X∞, fix ε > 0 and choose q ∈ N such that (cid:107)(I − P(0,q))x(cid:107) ≤ ε. Then for n ≥ q, (cid:107)Snx(cid:107) ≤ (cid:107)SnP(0,q)x(cid:107) + (cid:107)Sn(I − P(0,q))x(cid:107) ≤ (cid:107)(I − P(0,q))x(cid:107) ≤ ε so (cid:107)Snx(cid:107) → 0 as claimed. Lemma 4.3.2. Let K ∈ K(X∞). Then (cid:107)SnK(cid:107) → 0 as n → ∞. 2 -net, x1, . . . xm of K(BX∞). By Lemma 4.3.1, there 2 ε whenever n ≥ N, j = 1, . . . , m. Now suppose 2 ε. For n ≥ N we have 2 ε ≤ ε. Thus (cid:107)SnK(cid:107) → 0 as n → ∞ Proof. Fix ε > 0 and choose a finite ε exists N ∈ N such that (cid:107)Snxj(cid:107) ≤ 1 x ∈ BX∞ and let j ∈ {1, . . . , m} be such that (cid:107)Kx − xj(cid:107) ≤ 1 (cid:107)SnKx(cid:107) ≤ (cid:107)Sn(Kx − xj)(cid:107) + (cid:107)Snxj(cid:107) ≤ (cid:107)Kx − xj(cid:107) + 1 as required. Proposition 4.3.3. Let (λj)∞ quently, the sequence ([Sj])∞ j=0 ∈ (cid:96)1. Then (cid:107)(cid:80)∞ j=0 λjSj +K(X∞)(cid:107) ≤ (cid:107)(cid:80)∞ Proof. Certainly (cid:107)(cid:80)∞ j=0 λj by Lemma 4.2.10. To prove the converse inequality, fix ε > 0 and suppose K ∈ K(X∞). By Lemma 4.3.2, we can find m ∈ N with (cid:107)SmK(cid:107) ≤ ε. It follows that j=0 is 1-equivalent to the canonical basis of (cid:96)1. j=0 λj. Conse- j=0 λjSj + K(X∞)(cid:107) =(cid:80)∞ j=0 λjSj(cid:107) =(cid:80)∞ ∞(cid:88) λjSj+m(cid:107) j=0 λjSj+m(cid:107) − (cid:107)SmK(cid:107) ∞(cid:88) j=0 (cid:107)K + λjSj(cid:107) ≥ (cid:107)SmK + ≥ (cid:107) ∞(cid:88) ∞(cid:88) j=0 λjSj + K(X∞)(cid:107) ≥ (cid:80)∞ ≥ j=0 j=0 λj − ε. It follows that (cid:107)(cid:80)∞ (cid:107)(cid:80)∞ j=0 λjSj + K(X∞)(cid:107) ≥(cid:80)∞ j=0 λj. 120 j=0 λj − ε, and since ε > 0 was arbitrary, Chapter 5 Shift invariant (cid:96)1 preduals 5.1 Introduction We have seen in Chapters 3 and 4 that there exist a multitude of non-isomorphic Banach spaces with (cid:96)1 dual. Moreover, these spaces all possess some interesting Banach space structure; the spaces in Chapter 3 are easily seen to be saturated with infinite dimensional reflexive subspaces and have the property that their operator algebras and Calkin algebras can be considered 'small', whereas the interesting property of the space in Chapter 4 is the fact it has (cid:96)1 Calkin algebra. In addition to the results in this thesis, we also note that there are a number of other exotic (cid:96)1 preduals discussed in the existing literature. For example, it was shown in [5] that there exist isometric preduals of (cid:96)1 which are not isomorphic to a complemented subspace of any C(K) space and, more recently, it has been shown ([13]) that any Banach space with separable dual can be embedded isomorphically into an (cid:96)1 predual. In this chapter, we briefly look at a problem considered by Haydon, Daws, Schlumprecht and White in [9], where (cid:96)1 preduals satisfying an additional 'shift-invariant' condition are studied. For completeness, we begin by discussing the motivation behind this property and define precisely what we mean by a shift-invariant predual of (cid:96)1. We then show that one of the spaces defined in [9] (which was shown to be isomorphic to c0) can in fact be considered (in some sense) as being obtained from a specific Bourgain-Delbaen construction. Thus, whilst none of the results in this chapter are new, our approach is somewhat different to that taken by the authors of [9]. Given the success of constructing exotic Banach spaces via the Bourgain-Delbaen method, the author had hoped that this new approach might lead to new shift invariant preduals of (cid:96)1 possessing some interesting Banach space properties. Unfortunately, at the time of writing, this remains an open problem and the author has been unable to make any further progress. 121 The motivation behind [9] comes from the study of dual Banach algebras. If a Banach algebra, A, is also a dual space of some Banach space, E, i.e. there exists a (Banach space) isomorphism T : A → E∗, we can induce a weak* topology on A, simply by demanding that T should be a weak*-weak* homeomorphism. This is, of course, equivalent to saying that a net (xα)α∈I converges weak* to x in A if and only if T xα converges weak* to T x in E∗. We then have the following definition: Definition 5.1.1. A Banach algebra A is a dual Banach algebra if it is a dual space of some Banach space, E, and the product is separately weak*-continuous (with respect to the induced weak* topology just described). We note that the induced weak* topology, and hence the above definition, depends on the choice of isomorphism T : A → E∗. Therefore, one should be more precise and define a predual of A as a Banach space E together with an isomorphism T : A → E∗. We shall see shortly that we can forget about this technicality for the purposes of this chapter as we will be working with so-called concrete preduals of (cid:96)1 (see Definition 5.1.2). As noted in [9], it is well known that a C∗-algebra, M , which is isometric to a dual space is a von Neumann algebra. In this case, the product and involution are automatically weak* continuous, so that M is a dual Banach algebra. Moreover, the (isometric) predual is unique. However, we observe that isomorphic preduals need not be unique, even in the case of C∗ algebras. Indeed, Pelczy´nski proved that the Banach spaces (cid:96)∞ and L∞[0, 1] are isomorphic. It follows that both (cid:96)1 and L1[0, 1] are isomorphic preduals of the C∗ algebra, (cid:96)∞, while of course (cid:96)1 is not isomorphic to L1[0, 1]. The authors of [9] are specifically concerned with the Banach algebra (cid:96)1(Z) equipped with the natural convolution product, (f ∗ g)(n) = f (k)g(n − k), f, g ∈ (cid:96)1(Z), n ∈ Z (cid:88) k∈Z and the isomorphic preduals of (cid:96)1(Z) with respect to which (cid:96)1(Z) is a dual Banach algebra. We should really be a little more precise. Observe that if we were to equip (cid:96)1(Z) with the zero product, any isomorphic predual of (cid:96)1(Z) makes (cid:96)1(Z) equipped with the zero product into a dual Banach algebra. We are only interested in a predual E of (cid:96)1(Z) with respect to which (cid:96)1(Z) equipped with the convolution product (defined above) is a dual Banach algebra. Consequently, in what follows, we are always thinking of (cid:96)1(Z) as the 'usual' Banach algebra equipped with the multiplication being the convolution just defined. We also remark that the canonical isometry of c0(Z)∗ with (cid:96)1(Z) makes (cid:96)1(Z) into a dual Banach algebra, so the sorts of preduals we are interested in certainly exist. Whilst this happens to be an isometric predual, we emphasise that we are not restricting our attention to just isometric preduals. 122 In [9], the authors are looking for preduals E for which E has some interesting Banach space structure or for which the isomorphism between E and (cid:96)1(Z) induces an exotic weak* topology. Before continuing any further we remark that we shall frequently be making the canoni- cal identification of (cid:96)1(Z)∗ with (cid:96)∞(Z) and find it convenient to denote by θ : (cid:96)1(Z)∗ → (cid:96)∞(Z) the usual isometry. Observe that every predual, E, of (cid:96)1(Z) can be canonically embedded (isomorphically) into (cid:96)∞(Z). Indeed, if JE denotes the canonical isometric embedding of E into its second dual and T : (cid:96)1(Z) → E∗ is an isomorphism, the composition θ ◦ T ∗ ◦ JE : E → (cid:96)∞(Z) is an isomorphic embedding of E into (cid:96)∞(Z). It turns out (see Lemmas 5.1.3 and 5.1.4) that the preduals of (cid:96)1(Z) which make the multiplication separately weak* continuous, i.e. the preduals with respect to which (cid:96)1(Z) is a dual Banach algebra, are precisely those which, when thought of as a subspace of (cid:96)∞(Z) under this embedding, are invariant under the bilateral shift on (cid:96)∞(Z). Consequently, we refer to such preduals as shift-invariant preduals of (cid:96)1(Z). If was shown in [9] that, when considering shift-invariant preduals of (cid:96)1(Z), we lose no generality by considering so called concrete preduals. As well as being a considerable simplification, the (cid:96)1 duality arising from concrete preduals is precisely that which arises naturally from the embedding of (cid:96)1(Z) into the dual of a space of Bourgain-Delbaen type. Consequently, concrete preduals are fundamental, both to the work of the authors in [9] and to enable us to connect the work of this thesis with the results in the aforementioned paper. For this reason, we chose to include the details of concrete preduals now; the proofs and results largely follow the work of the authors in [9]. Given a closed subspace F ⊆ (cid:96)∞(Z), it is well known that the dual space F ∗ is canon- ically isometric to (cid:96)∞(Z)∗/F ⊥, where F ⊥ = {Φ ∈ (cid:96)∞(Z)∗ : (cid:104)Φ, x(cid:105) = 0 ∀x ∈ F}. Let I : (cid:96)∞(Z)∗/F ⊥ → F ∗ denote this identification and π : (cid:96)∞(Z)∗ → (cid:96)∞(Z)∗/F ⊥ be the obvi- ous projection. Definition 5.1.2. Given a closed subspace F ⊆ (cid:96)∞(Z), we define the map ιF : (cid:96)1(Z) → F ∗ to be the composition ιF = Iπ(θ−1)∗J(cid:96)1(Z) and say F is a concrete predual for (cid:96)1(Z) if the map ιF is an isomorphism. It is an easy computation to see that if a ∈ (cid:96)1(Z) and x ∈ F ⊆ (cid:96)∞(Z), then (cid:104)ιF (a), x(cid:105) = (cid:104)x, a(cid:105) (where we are of course thinking of x as an element of (cid:96)1(Z)∗ in the right-hand-side of the equality). Consequently, if F is a concrete predual, the identification of (cid:96)1 with F ∗ under ιF is the same type of identification of (cid:96)1 as a subspace of X∗ when X is a Bourgain-Delbaen space (see, for example, Theorem 2.2.2, Proposition 3.4.5). 123 As stated earlier, we lose no generality in working with concrete preduals for (cid:96)1(Z) and this is the contents of the next lemma (which is proved in [9, Lemma 2.1]). We remark that the lemma states both the Banach space structure and induced weak* topologies on (cid:96)1(Z) are preserved when changing to the setup of a concrete predual. This is fundamental given we are concerned with those preduals which turn (cid:96)1(Z) into a dual Banach algebra. Lemma 5.1.3. Let E be a Banach space and T : (cid:96)1(Z) → E∗ be an isomorphism. Then the map θT ∗JE : E → (cid:96)∞(Z) is an isomorphism onto its range, say F ⊆ (cid:96)∞(Z). Furthermore, ιF is an isomorphism so that F is a concrete predual for (cid:96)1(Z) and the weak* topologies induced by T : (cid:96)1(Z) → E∗ and ιF : (cid:96)1(Z) → F ∗ agree. That is, given a net (aα) in (cid:96)1(Z), we have that limα(cid:104)T (aα), x(cid:105) = 0 for all x ∈ E if and only if limα(cid:104)ιF (aα), y(cid:105) = limα(cid:104)y, aα(cid:105) = 0 for all y ∈ F . Proof. The first statement is obvious. We prove that ιF is an isomorphism and the desired properties of the weak* topologies. Let R denote the isomorphism obtained by thinking of the operator θT ∗JE as a map onto its image, that is, R = θT ∗JE : E → F (where F ⊆ (cid:96)∞(Z) is defined as in the lemma). Note that for x ∈ E, a ∈ (cid:96)1(Z), (cid:104)R∗ιF (a), x(cid:105) = (cid:104)ιF (a), Rx(cid:105) = (cid:104)θ−1Rx, a(cid:105) = (cid:104)T ∗JEx, a(cid:105) = (cid:104)T a, x(cid:105). In other words, R∗ιF = T . It follows that ιF = (R∗)−1T and thus ιF is an isomorphism from (cid:96)1(Z) to F ∗ as required. Next, suppose (aα) is a net in (cid:96)1(Z). We have (aα) weak* convergent to 0 with respect to the topology induced by the isomorphism ιF if and only if (cid:104)ιF (aα), y(cid:105) → 0 for all y ∈ F . Since ιF = (R−1)∗T , this happens if and only (cid:104)(R−1)∗T (aα), y(cid:105) = (cid:104)T aα, R−1y(cid:105) → 0 for all y ∈ F . Since R : E → F is an isomorphism, this happens if and only if (cid:104)T aα, x(cid:105) → 0 for all x ∈ E, that is, if and only if (aα) is weak* convergent to 0 with respect to the weak* topology induced by the isomorphism T : (cid:96)1(Z) → E∗ as required. For completeness, we state and prove two more lemmas. The results and proofs are taken from [9] and the author of this thesis makes no claim to originality of the two results that follow. The first lemma proves that the shift invariant condition discussed at the beginning of this chapter is indeed equivalent to requiring the predual make (cid:96)1 with its natural convolution product into a dual Banach algebra. Of course, since we now know that we lose no generality in working with concrete preduals, we choose to work with these in the lemma. The second lemma provides further evidence that concrete preduals do indeed provide a simplification to the study of the problem originally motivating the authors of 124 [9]. Indeed, we see that concrete preduals allow one to detect if two concrete preduals F1, F2 ⊆ (cid:96)∞(Z) induce the same weak* topology on (cid:96)1(Z); this happens if and only if F1 = F2. Lemma 5.1.4. Let F ⊆ (cid:96)∞(Z) be a concrete predual for (cid:96)1(Z). The following are equivalent: (1) The bilateral shift on (cid:96)1(Z) is weak*-continuous with respect to F . (2) The subspace F is invariant under the bilateral shift on (cid:96)∞(Z). (3) (cid:96)1(Z) is a dual Banach algebra with respect to F . Proof. We show that (1) and (3) are equivalent and (2) and (3) are equivalent. We denote by en the usual vector in (cid:96)1(Z) (i.e. en = (. . . , 0, 0, 1, 0, 0 . . . ) ∈ (cid:96)1(Z) where the 1 occurs in the n'th coordinate) and, for y ∈ (cid:96)1(Z), we denote by ∗y : (cid:96)1(Z) → (cid:96)1(Z) the linear operator defined by x (cid:55)→ x ∗ y. If (3) holds then, in particular, the maps ∗e1 and ∗e−1 on (cid:96)1(Z) are weak* continuous with respect to F . Note that if n ∈ Z, x ∗ e1(n) = (cid:88) x(k)e1(n − k) = x(n − 1) k∈Z so the map ∗e1 is just the right shift on (cid:96)1(Z). It is just as easy to see that ∗e−1 is the left shift operator. So we conclude that the bilateral shifts on (cid:96)1(Z) are weak* continuous with respect to F . Conversely, suppose (1) holds, i.e. the right and left shifts on (cid:96)1(Z) are weak* continuous. We must see that for a fixed y ∈ (cid:96)1(Z) the map ∗y on (cid:96)1(Z) is weak* continuous with respect to F . A similar computation to the one above shows that for j ∈ N, the map ∗ej the right shift operator on (cid:96)1(Z) applied j times and the map ∗e−j (cid:96)1(Z) applied j times. Thus if there is some N ∈ N such that y = (cid:80)j≤N ajej then ∗y =(cid:80)j≤N aj∗ej . For an arbitrary y ∈ (cid:96)1(Z), we can write y = limN→∞(cid:80)j≤N ajej. We write PN y =(cid:80)j≤N ajej. Observe that the map ∗y is weak* continuous since it is just the finite sum of weak* continuous maps; is the left shift on is just (cid:107)x ∗ y − x ∗ PN y(cid:107) = (cid:107)x ∗ (y − PN y)(cid:107) ≤ (cid:107)x(cid:107)(cid:107)y − PN y(cid:107) so that ∗y is the the limit (in the operator norm topology) of the (weak* continuous) operators ∗PN y. It follows by the second part of Lemma 1.4.12 that ∗y is weak* continuous, as required. 125 Now suppose F is invariant under the left shift operator on (cid:96)∞(Z). We claim the right x, shift, σ, on (cid:96)1(Z) is weak*-continuous. Indeed, let (xα) be a net in (cid:96)1(Z) with xα (cid:42)w∗ i.e. (cid:104)θ−1f, xα(cid:105) → (cid:104)θ−1f, x(cid:105) for all f ∈ F ⊆ (cid:96)∞(Z). We are required to see σxα (cid:42)w∗ σx. To see this, let f ∈ F ⊆ (cid:96)∞(Z) and note that (cid:104)θ−1f, σxα(cid:105) = (cid:104)σ∗θ−1f, xα(cid:105) = (cid:104)θ−1 ◦ (θσ∗θ−1)f, xα(cid:105). An easy computation shows that the operator θσ∗θ−1 : (cid:96)∞(Z) → (cid:96)∞(Z) is just the left shift on (cid:96)∞(Z). Indeed, if (am)m∈Z is in (cid:96)∞(Z) and em denotes the m'th standard basis unit vector of (cid:96)1(Z) as before, we have (cid:0)θσ∗θ−1 ((am)m∈Z)(cid:1) (n) = (cid:104)σ∗(cid:0)θ−1(am)(cid:1) , en(cid:105) = (cid:104)θ−1(am), σen(cid:105) = an+1 so θσ∗θ−1 is the left shift on (cid:96)∞ as required. Now, since we are assuming that F is invariant under the left shift, f(cid:48) := θσ∗θ−1f ∈ F . Thus, from our previous calculation, and that xα (cid:42)w∗ x we have, (cid:104)θ−1f, σxα(cid:105) = (cid:104)θ−1f(cid:48), xα(cid:105) → (cid:104)θ−1f(cid:48), x(cid:105) = (cid:104)σ∗θ−1f, x(cid:105) = (cid:104)θ−1f, σx(cid:105) as required. A similar argument yields that if F is invariant under the right shift operator on (cid:96)∞(Z), then the left shift on (cid:96)1(Z) is weak* continuous. We have thus proved that (2) =⇒ (3). Conversely, suppose we know the right shift σ on (cid:96)1(Z) is weak* continuous. We will show F is left shift invariant. We continue with the notation used previously, recalling that θσ∗θ−1 : (cid:96)∞(Z) → (cid:96)∞(Z) is the left shift on (cid:96)∞(Z). Now, if our claim is false, there exists an element f(cid:48) ∈ F such that it's image under the left shift, θσ∗θ−1f(cid:48) ∈ (cid:96)∞(Z)\ F . By the Hahn-Banach Theorem, we may find a γ∗ ∈ (cid:96)∞(Z)∗ such that (i) (cid:104)γ∗, θσ∗θ−1f(cid:48)(cid:105) = 1 (ii) γ∗(F ) = {0} or, equivalently, (cid:104)γ∗, f(cid:105) = 0 for all f ∈ F . Note that θ∗γ∗ ∈ (cid:96)1(Z)∗∗. We can rewrite (i) above as (cid:104)(θ−1)∗θ∗γ∗, θσ∗θ−1f(cid:48)(cid:105) = (cid:104)θ∗γ∗, σ∗θ−1f(cid:48)(cid:105) = 1, and similarly, we can rewrite (ii) as (cid:104)θ∗γ∗, θ−1f(cid:105) = 0 for all f ∈ F . By Goldstine's Theorem, there is a net (xα)α∈I in (cid:96)1(Z) with (cid:107)xα(cid:107)(cid:96)1 ≤ (cid:107)θ∗γ∗(cid:107) for every 126 α and J(cid:96)1xα (cid:42)w∗ limα(cid:104)θ−1f, xα(cid:105) = 0 for all f ∈ F . θ∗γ∗. We conclude that limα(cid:104)σ∗θ−1f(cid:48), xα(cid:105) = limα(cid:104)θ−1f(cid:48), σxα(cid:105) = 1 and Now (xα) is a net in some closed ball of (cid:96)1(Z) (cid:39) F ∗. By the weak* compactness of closed balls in F ∗, (xα) has a subnet (xg(β))β which is weak* convergent to some x ∈ (cid:96)1(Z). Since (cid:104)θ−1f(cid:48), σxg(β)(cid:105) is a subnet of (cid:104)θ−1f(cid:48), σxα(cid:105), the latter of which we know converges to 1, we have limβ(cid:104)θ−1f(cid:48), σxg(β)(cid:105) = 1. Moreover, by the assumed weak* continuity of σ, we have limβ(cid:104)θ−1f(cid:48), σxg(β)(cid:105) = (cid:104)θ−1f(cid:48), σx(cid:105) and therefore, (cid:104)θ−1f(cid:48), σx(cid:105) = 1. On the other hand, by the very definition of the weak* topology induced on (cid:96)1(Z), we have (cid:104)θ−1f, x(cid:105) = limβ(cid:104)θ−1f, xg(β)(cid:105) = limα(cid:104)θ−1f, xα(cid:105) = 0 for all f ∈ F . This is equivalent to saying (cid:104)ιF (x), f(cid:105) = 0 for all f ∈ F , which of course tells us that ιF (x) = 0, and hence x = 0 (ιF is an isomorphism). This clearly contradicts (cid:104)θ−1f(cid:48), σx(cid:105) = 1. Again, one can argue similarly to prove that if the left shift on (cid:96)1(Z) is weak* continuous, then F must be right shift invariant. This completes the proof. Lemma 5.1.5. Let E1 and E2 be preduals of (cid:96)1(Z), and use these to induce concrete preduals F1, F2 ⊆ (cid:96)∞(Z) as described in Lemma 5.1.3. Then E1 and E2 induce the same weak* topology on (cid:96)1(Z) if and only if F1 = F2. In particular, two concrete preduals F1 and F2 induce the same weak* topology on (cid:96)1(Z) if and only if F1 = F2. Proof. If F1 = F2 then it follows immediately from Lemma 5.1.3 that E1 and E2 induce the same weak* topology on (cid:96)1(Z). Conversely, for i=1,2, let Ti : (cid:96)1(Z) → E∗ i be an isomor- phism, and suppose that these induce the same weak* topology on (cid:96)1(Z). Suppose for a contradiction that there exists an x ∈ F2 \ F1. By the Hahn-Banach Theorem, there exists Λ ∈ (cid:96)∞(Z)∗ with (cid:104)Λ, x(cid:105) = 1 and (cid:104)Λ, y(cid:105) = 0 for all y ∈ F1. By Goldstine's Theorem, there is θ∗Λ. It follows that, for all y ∈ F1, we have a bounded net (aα) in (cid:96)1(Z) with J(cid:96)1(Z)aα (cid:42)w∗ (cid:104)θ−1y, aα(cid:105) = lim 0 = (cid:104)Λ, y(cid:105) = (cid:104)θ∗Λ, θ−1y(cid:105) = lim (cid:104)ιF1(aα), y(cid:105) α α so that (T1(aα)) is weak* null in E∗ is weak* null in E∗ 2 , but this contradicts that 1 by Lemma 5.1.3. By assumption, it follows that (T2(aα)) 1 = (cid:104)Λ, x(cid:105) = lim α (cid:104)θ−1x, aα(cid:105) = lim (cid:104)ιF2(aα), x(cid:105). α This shows that F2 ⊆ F1, and an identical argument shows F1 ⊆ F2, as required. 5.2 Connection to the BD construction Having described the motivation behind the work in [9] and understood the importance of concrete preduals, we now move on to discuss a specific class of shift invariant preduals that 127 were introduced in [9]. We follow the same notation used in [9]; for n ≥ 0 in Z, we denote by b(n) the number of ones in the binary expansion of n. (So b(0) = 0, b(1) = 1, b(2) = 1, b(3) = 2 and so on.) Definition 5.2.1. Let λ ∈ R,λ > 1. Define the vector xλ when n ≥ 0 and 0 otherwise. Thus xλ 0 is given by 0 ∈ (cid:96)∞(Z) by xλ 0 (n) = λ−b(n) x0 = (. . . 0, 0, 1, λ−1, λ−1, λ−2, λ−1, λ−2, λ−2, λ−3, λ−1, . . . ). Fλ ⊆ (cid:96)∞(Z) is defined to be the closed, shift-invariant subspace of (cid:96)∞(Z) generated by the vector xλ 0 , i.e. Fλ is the closed linear span of the bilateral shifts of xλ 0 . It was shown in [9] that the collection of subspaces (Fλ)λ>1 of (cid:96)∞(Z) are all concrete preduals of (cid:96)1(Z). Clearly, by construction, they are shift invariant and so by Lemma 5.1.4, they are preduals of (cid:96)1(Z) that make (cid:96)1(Z) with the convolution product a dual Banach algebra. The authors of [9] used a somewhat technical argument involving the Szlenk index to show that as Banach spaces, the spaces Fλ are all in fact isomorphic to c0. The interest in the spaces Fλ is therefore due to the weak* topologies that they induce on (cid:96)1(Z). Since the Fλ are pairwise distinct subspaces of (cid:96)∞(Z), and distinct from c0, it is immediate from Lemma 5.1.5 that these spaces induce an uncountable family of distinct weak*-topologies making (cid:96)1(Z) into a dual Banach algebra. The final result of this thesis, is an observation about the space F2. We will show that we can obtain a 'one-sided' version of the space F2 using the Bourgain Delbaen construction. When viewed in this framework, it is very easy to see that the one-sided version of F2 is isomorphic as a Banach space to c0 and that its dual space is naturally isomorphic to (cid:96)1. In what follows N0 denotes the set of natural numbers including 0. We first set up the notation. We define the vector y0 ∈ (cid:96)∞(N0) by y0(n) = 2−b(n) where, as before, b(n) denotes the number of ones in the binary expansion of n. Thus y0 is given by (1, 2−1, 2−1, 2−2, 2−1, . . . ) We define R2 to be the closed, right-shift-invariant subspace of (cid:96)∞(N0) generated by y0, that is, R2 is the closed linear span of the vector y0 and all its right shifts. We also use Theorem 2.2.2 to construct a space of Bourgain-Delbaen type. Comparing with the notation from Theorem 2.2.2, we take Γ = N0 and define finite subsets ∆n ⊂ N0 as follows. We set ∆0 = {0}, and for n ≥ 1, we set ∆n = {2n−1 + j : 0 ≤ j < 2n−1}. For each k ≥ 1, if n ∈ ∆k, we can uniquely write n = 2k−1 + m where m < 2k−1 and m, k ∈ N0. Again comparing with the notation from Theorem 2.2.2 and making using of the previous 128 observation, we define a τ mapping by setting τ (n) to be the Type 0 tuple (1, 1 2 , m). With this setup, it is easily seen that the Bourgain-Delbaen Theorem (Theorem 2.2.2) applies. n ∈ (cid:96)1(N0). For clarity, m ∈ (cid:96)1(N0) and d∗ We get corresponding vectors c∗ 2 e∗ n = 1 the first few d∗ vectors are as follows: n = e∗ n − c∗ d∗ 0 = (1, 0, 0, 0, 0, 0, . . . ) 1 = (−1/2, 1, 0, 0, 0, 0, . . . ) d∗ 2 = (−1/2, 0, 1, 0, 0, 0, . . . ) d∗ 3 = (0,−1/2, 0, 1, 0, 0, . . . ). d∗ We obtain a L∞ space of Bourgain-Delbaen-type, RBD := X(N0, τ ). We recall that RBD is nothing other than the closed linear span of the biorthogonal vectors [dn : n ∈ N0] ⊆ (cid:96)∞(N0). The main observation of this chapter is the following lemma: Lemma 5.2.2. The spaces R2 and RBD defined above are the same subspace of (cid:96)∞(N0). The vectors (d∗ n)∞ n=1 are in fact a basic sequence in (cid:96)1(N0) equivalent to the canonical basis of (cid:96)1(N0). Consequently, RBD is a right-shift-invariant subspace of (cid:96)∞ which is isomorphic to c0 and has dual space naturally isomorphic to (cid:96)1(N0). Proof. We first show that R2 and RBD are the same subspace of (cid:96)∞(N0). The proof involves nothing other than a number of lengthy computations so we proceed by proving a number of smaller claims. It is convenient for us to re-introduce the τ operator that featured in [9]. Precisely, τ : (cid:96)∞(N0) → (cid:96)∞(N0) is the bounded linear map defined by τ (x)(n) = x( n 0 2 ) n even n odd. We denote by σ : (cid:96)∞(N0) → (cid:96)∞(N0) the right shift, i.e. σx(n) = x(n − 1). It was shown in [9, Lemma 3.1] that τ k(y0) ∈ R2 for all k ∈ N0 and moreover, τ σ = σ2τ . We also define a bounded operator β : (cid:96)∞(N0) → (cid:96)∞(N0) by (cid:40) (cid:40) β(x)(n) = We note that στ = β so that β(y0) ∈ R2. Step 1. d0 = y0. x( n−1 0 2 ) n odd n even. We use induction on n ∈ N0 to prove that d0(n) = (cid:104)d0, e∗ 0(cid:105) = (cid:104)d0, d∗ 0(cid:105) = 1 = 2−b(0) as required. Inductively, assume that (cid:104)d0, e∗ n(cid:105) = 2−b(n) = y0(n). Clearly j(cid:105) = 2−b(j) for all m(cid:105) where n + 1 = 2k + m with m < 2k. (cid:104)d0, e∗ j ≤ n. Then (cid:104)d0, e∗ 0, e∗ n+1(cid:105) = (cid:104)d0, d∗ n+1 + c∗ n+1(cid:105) = 1 2(cid:104)d∗ 129 In particular, m ≤ n, so by the inductive hypothesis, (cid:104)d0, e∗ m < 2k, we can write n + 1 = 2k +(cid:80)k−1 2 2−b(m). Note that since j=0 εj2j where εj ∈ {0, 1} for all j. It follows that n+1(cid:105) = 1 b(n + 1) = b(m) + 1 and so (cid:104)d0, e∗ Step 2. For all j ∈ N0, d2j+1 = βdj = στ dj. n+1(cid:105) = 2−b(n+1), completing the induction. We use induction on n ∈ N0 to prove that d2j+1(n) = βdj(n). It is easily seen that for n < 2j + 1, d2j+1(n) = 0. Moreover, for such n, βdj(n) = 0 if n is even. When n is odd and n < 2j + 1, (n − 1)/2 < j so that βdj(n) = dj((n − 1)/2) = 0. Consequently the inductive statement holds when n < 2j + 1. When n = 2j + 1 it is easily verified that d2j+1(n) = βdj(n) = 1. We now suppose that the inductive statement holds for all k ≤ n and must show that d2j+1(n + 1) = βdj(n + 1). By the preceding argument, we may as well assume that n ≥ 2j + 1, so n + 1 > 2j + 1 ≥ 1. We write n + 1 = 2k + m where 0 ≤ m < 2k and since n + 1 ≥ 2 we must have k ≥ 1. Now d2j+1(n + 1) = (cid:104)d2j+1, c∗ n+1(cid:105) = 1 m(cid:105). Certainly m ≤ n, so we may apply the apply inductive hypothesis which yields 2(cid:104)d2j+1, e∗ (cid:40) 1 2 dj( m−1 d2j+1(n + 1) = 1 2 βdj(m) = 2 ) m odd 0 otherwise. Finally, we consider cases where n + 1 is even or odd separately, noting that n + 1 is even if and only if m is even, since we have seen that k ≥ 1. Consequently, if n + 1 is even, then our above calculation yields that d2j+1(n + 1) = 0 = βdj(n + 1). It remains only to consider 2 dj((m− 1)/2). the possibility that n + 1 is odd, in which case, m is odd and d2j+1(n + 1) = 1 We need to see that this is equal to βdj(n + 1) = dj(n/2). Since we assumed n ≥ 2j + 1, we have n/2 ≥ j + 1/2 and since n/2 ∈ Z, in fact, n/2 ≥ j + 1 > j. Consequently, dj(n/2) = (cid:104)dj, c∗ n/2(cid:105). Recalling that n + 1 = 2k + m, where m must now be odd, we see that n/2 = 2k−1 + m−1 Step 3. For all j ≥ 1, d2j = τ dj. and consequently (cid:104)dj, c∗ 2 ) as required. 2 dj( m−1 n/2(cid:105) = 1 2(cid:104)dj, e∗ (cid:105) = 1 m−1 2 2 Again, we use induction on n and prove that d2j(n) = τ dj(n) for all n ∈ N0. The calculations are very similar to the previous step. When n < 2j it's easy to see that d2j(n) = 0 = τ dj(n) and when n = 2j, d2j(n) = 1 = τ dj(n). So we suppose that the inductive statement holds for all k ≤ n and prove the statement must hold for n + 1. By the previous argument, we can assume n ≥ 2j ≥ 2. Then n + 1 > 2j and d2j(n + 1) = (cid:104)d2j, c∗ m(cid:105) where n + 1 = 2k + m and we must have k ≥ 1 since n + 1 ≥ 3. By the inductive hypothesis we have d2j(n + 1) = 1 2 τ dj(m). Now, if n + 1 is odd, m must also be odd, so d2j(n + 1) = 1 2 τ dj(m) = 0 = τ dj(n + 1). So we assume now that n + 1 is even, (cid:105) so m is even and d2j(n + 1) = 1 2 dj(m/2). In this case, τ dj(n + 1) = dj( n+1 2 ) = (cid:104)dj, c∗ n+1(cid:105) = 1 2(cid:104)d2j, e∗ n+1 2 130 since we recall that n + 1 > 2j. Also n + 1 = 2k + m so that (n + 1)/2 = 2k−1 + m/2. It follows that τ dj(n + 1) = (cid:104)dj, c∗ Step 4. d2j +m = σmd2j whenever j ≥ 1, 1 ≤ m < 2j. 2 dj(m/2) = d2j(n + 1) as required. (cid:105) = 1 n+1 2 To see this, we will show by induction on j ∈ N that d2j +m = σmd2j whenever 1 ≤ m < 2j. The base case, j = 1 follows easily from the previous two steps. Indeed, the only permissible value of m is 1 and d21+1 = στ d1 = σd21 as required. So, suppose by induction, that for some j > 1, d2j−1+m = σmd2j−1 whenever 1 ≤ m < 2j−1. We must show that d2j +m = σmd2j whenever 1 ≤ m < 2j. We will do this by an induction on m. The base case, m = 1, is again easy. By Steps 2 and 3, we have d2j +1 = στ d2j−1 = σd2j as needed. So suppose that d2j +m = σmd2j and that m + 1 < 2j. If m = 2k, we have d2j +m+1 = d2(2j−1+k)+1 = στ d2j−1+k = σd2j +2k = σd2j +m = σm+1d2j . Here we have appealed to the previous two steps and the inductive hypothesis on m. It remains to consider the case m = 2k + 1. Again, we will make repeated use of the previous two steps. We have d2j +m+1 = d2j +2k+2 = d2(2j−1+k+1) = τ d2j−1+k+1. Moreover, m + 1 = 2(k + 1) < 2j so that k + 1 < 2j−1. Thus by our inductive hypothesis on j, d2j−1+k+1 = σk+1d2j−1 and so d2j +m+1 = τ σk+1d2j−1 = σ2k+2τ d2j−1 = σ2k+2d2j = σm+1d2j . (Here we made use of the fact that τ σ = σ2τ .) This completes both inductive steps. Step 5. σdn = dn+1 except if n is of the form n = 2j − 1, with j ≥ 0. This is almost immediate from the previous step. Indeed, suppose j ≥ 1 and 1 ≤ m < 2j − 1. Then if n = 2j + m, σdn = σd2j +m = σm+1d2j = d2j +m+1 = dn+1. We made use of the fact that m + 1 < 2j so that we could apply the previous step to obtain the penultimate equality. If n = 2j with j ≥ 1 then dn+1 = d2(2j−1)+1 = στ d2j−1 by Step 2, and στ d2j−1 = σd2j = σdn by Step 3. We therefore have that σdn = dn+1 whenever n can be written in the form 2j + m with j ≥ 1 and 0 ≤ m < 2j − 1. The only values of n which aren't of the form n = 2j + m with j ≥ 1 and 0 ≤ m < 2j − 1 are precisely those n of the form 2k − 1 for some k ∈ N0. Step 6. For j ≥ 0, (cid:40) d2j (n) = d0(l) 0 if n = 2j(2l + 1) (for some l ∈ N0) otherwise. In particular the vectors {d2j : j ∈ N0} have disjoint supports. We use induction on j ∈ N0. When j = 0 we have d1(n) = στ d0(n) by Step 2. So d1(n) = τ d0(n − 1) = d0( n−1 2 ) 0 if n − 1 even otherwise. (cid:40) 131 It follows that d1(n) = 0 unless n = 2l + 1 in which case d1(n) = d0(l) as required. Now assume the claim holds for some j ≥ 0. We have d2j+1(n) = d2·2j (n) = τ d2j (n) by Step 3. So d2j+1(n) = d2j ( n 2 ) 0 if n even otherwise. (cid:40) From this, and the inductive hypothesis, we see that d2j+1(n) = 0 unless n/2 = 2j(2l + 1) for some l ∈ N0, i.e. if n = 2j+1(2l + 1), in which case we have d2j+1(n) = d0(l). The statement about disjoint supports is obvious. Step 7. For all j ≥ 1, d2j−1(n) = d0(k) 0 if there is k ∈ N0 such that n = 2j(k + 1) − 1 otherwise. (cid:40) We proceed by induction on j. When j = 1, d1(n) = στ d0(n) by Step 2. So d1(n) = τ d0(n − 1); this equals d0(k) if n = 2k + 1 = 2(k + 1) − 1 for some k ∈ N0 and 0 otherwise by the definition of the τ operator. Assuming the claim holds for some j ≥ 1, d2j+1−1(n) = d2(2j−1)+1(n) = στ d2j−1(n) by Step 2. So d2j+1−1(n) = τ d2j−1(n − 1) = (cid:40) d2j−1( n−1 2 ) 0 if n − 1 even otherwise. 0 = d∗ If m = 0, then σd0(n) = σd0(2l + 1) = d0(2l) = d0(l). k=0 2−kd2k (n) =(cid:80)∞ From this, and the inductive hypothesis, we see that d2j+1−1(n) (cid:54)= 0 if and only if (n−1)/2 = Step 8. For all j ≥ 0, σd2j−1 =(cid:80)∞ 2j(k + 1) − 1 for some k ∈ N0 in which case d2j+1−1(n) = d0(k) as required. k=0 2−kd2k+j . (cid:80)∞ We first treat the case j = 0. Since e∗ 0 we easily see that σd0(0) = 0 = k=0 2−kd2k (0). Now, for every n ∈ N, we can write n = 2m(2l + 1) for unique m, l ∈ N0. here and the fact that b(2l) = b(l).) By Step 6,(cid:80)∞ (We have made use of Step 1 k=0 2−kd2k (2l + 1) = n = 2m(2l + 1) and m ≥ 1. In this case, (cid:80)∞ d1(2l + 1) = d0(l) = σd0(n) where the penultimate equality follows by Step 7. Otherwise, other hand, we can write n =(cid:80)N k=0 2−kd2k (n) = 2−md0(l) by Step 6. On the and εm = 1. It follows that n− 1 = 2m − 1 +(cid:80)N k=m εk2k for some N ≥ m, where εk ∈ {0, 1} for k > m k=m+1 εk2k. We see from this that b(n − 1) = b(n) + m − 1. Consequently, σd0(n) = d0(n − 1) = 2−b(n−1) 2−b(l)−m = 2−md0(l) =(cid:80)∞ (by Step 1) = 2−b(n)−m+1. Finally we note that b(n) = b(2l + 1) = b(l) + 1 so, σd0(n) = that σd2j−1(0) = 0 = (cid:80)∞ The calculation for when j ≥ 1 is only slightly more involved. It is again easily seen If n ∈ N, we again write n = 2m(2l + 1) and σd2j−1(n) = d2j−1(n − 1) = d2j−1(2m(2l + 1) − 1). By Step 7, this will be non-zero if and k=m+1 εk2k =(cid:80)m−1 k=0 2k +(cid:80)N k=0 2−kd2k (n) as required. k=0 2−kd2k+j (0). 132 a=0 a=m−j+1 εa2a =(cid:80)m−j−1 only if m ≥ j. Moreover, if m = j, i.e. n = 2m(2l + 1), then we see from the previous calculation and Step 7 that σd2j−1(n) = d0(2l). By Step 1, this equals d0(l). If m > j, then that 2m−j(2l + 1) =(cid:80)N n = 2j(2m−j(2l + 1)) and by Step 7, we have σd2j−1(n) = d0(2m−j(2l + 1) − 1). We note εm−j = 1. It follows that 2m−j(2l + 1) − 1 = 2m−j − 1 +(cid:80)N a=m−j εa2a for some N ≥ m− j, where εa ∈ {0, 1} for a > m− j and (cid:80)N 2a + a=m−j+1 εa2a. This implies that b(2m−j(2l + 1) − 1) = b(2m−j(2l + 1)) + m − j − 1. So, σd2j−1(n) = d0(2m−j(2l + 1) − 1) = 2−b(2m−j (2l+1)−1) = 2−m+j+12−b(2m−j (2l+1)) = It remains to compute (cid:80)∞ 2−m+j+12−b(2l+1) = 2−m+j2−b(l) = 2−m+jd0(l). get (cid:80)∞ we have (cid:80)∞ k=0 2−kd2k+j (n) and see that the formulae we obtain agree with those just found for σd2j−1(n). We note that if n = 2m(2l + 1) then d2k+j (n) will only be non-zero if k + j = m by Step 6. In particular, if n is such that m < j then we k=0 2−kd2k+j (n) = 0 = σd2j−1(n). Otherwise, if n = 2m(2l + 1) is such that m ≥ j k=0 2−kd2k+j (n) = 2−kd0(l) where k is such that k + j = m, by Step 6. The latter expression is therefore equal to 2−m+jd0(l) = σd2j−1(n), as found in the previous paragraph. We are finally in a position to show that R2 = RBD. By Step 5 and Step 8, we clearly (cid:1) = σ(cid:0)[dn : n ∈ N0](cid:1) ⊆ RBD. Since, by Step 1, d0 = y0, we conclude that have that σ(cid:0)RBD RBD is a closed, right-shift invariant subspace of (cid:96)∞(N0) containing y0; we must therefore have R2 ⊆ RBD. On the other hand, since τ k(y0) = τ k(d0) ∈ R2 for all k, and σ2τ = τ σ, it is easily seen by Steps 2 and 3 that for each n ∈ N there are l, m ∈ N0 with dn = σmτ ld0 ∈ R2. Consequently RBD ⊆ R2 and so the two subspaces are indeed equal. n(cid:107) ≤ 1 Finally, we observe that in this construction, (cid:107)c∗ 2 for all n ∈ N0. Appealing to Lemma 2.2.4, we see that RBD is isomorphic to c0. Moreover, the sequence of vectors (d∗ n)∞ n) in (cid:96)1(N0) is equivalent to (e∗ n=1 is a boundedly complete basis n). It follows by Theorem 1.4.6 that (dn) is a shrinking basis for RBD. for (cid:96)1 so also is (d∗ Consequently by the Bourgain-Delbaen Theorem (Theorem 2.2.2), the dual space of RBD is naturally isomorphic to (cid:96)1, i.e. the (cid:96)1 duality obtained here is the same as that which defines the ιF mapping discussed earlier. n) ⊆ (cid:96)1(N0). Since (e∗ 5.3 Concluding Remarks We have obtained a right-shift invariant (cid:96)1 predual using the Bourgain-Delbaen construction which has a striking resemblance to the space F2 constructed in [9]. However, the space R2 in the above lemma does not quite do what we want it to. Tracing through the proof of Lemma 5.1.4, we see that for the natural convolution on (cid:96)1(N0) to be weak* continuous 133 with respect to the topology induced by the natural isomorphism of (cid:96)1 with R∗ need R2 to be left shift invariant. Alas, our shift is in the wrong direction! 2, we would At the time of writing, it is unclear to the author whether or not it is possible to modify the construction in some way so as to obtain a one-sided version of F2 that is left-shift invariant as a subspace of (cid:96)∞(N0). Less clear still, is whether or not it is possible to obtain 2-sided (bilateral) versions of these spaces via the Bourgain-Delbaen construction; indeed, the recursive nature of the Bourgain-Delbaen construction makes the one-sided object a much more natural object to obtain. We make one final remark on shift-invariant preduals in relation to the Bourgain-Delbaen construction. We have seen in Chapters 3 and 4 that the Bourgain-Delbaen construction is much more naturally understood with respect to a countable set Γ whereby elements of Γ code the form of the corresponding c∗ vector. Of course, when working with shift invariant preduals of (cid:96)1, one needs to make some enumeration of Γ so that the shift operator can be defined. This provides yet another obstruction to finding exotic (cid:96)1 preduals via the Bourgain-Delbaen construction. Whilst the work contained in this thesis has allowed us to gain an insight into the structure of Calkin algebras, we note that there are a number of open problems that arise from the work here. Indeed, it is unclear exactly what Banach algebras can be realised as a Calkin algebra. If it is too much to hope to solve this problem, a useful partial result might be to obtain some kind of obstruction that would prevent a Banach algebra being obtained as a Calkin algebra. 134 References [1] F. Albiac and N. J. Kalton, Topics in banach space theory, Springer, 2006. [2] S.A Argyros and R.G Haydon, A Hereditarily Indecomposable L∞ space that solves the scalar-plus-compact problem, Acta Math. 206 (2011), no. 1, 1 -- 54. [3] N. Aronszajn and K.T. Smith, Invariant subspaces of completely continuous operators, Ann. of Math. 60 (1954), no. 2, 345 -- 350. [4] K. Beanland and L. Mitchell, Operators on the L∞-spaces of Bourgain and Delbaen, Quaestiones Mathematicae 33 (2010), 443 -- 448. [5] Y. Benyamini and J. Lindenstrauss, A predual of (cid:96)1 which is not isomorphic to a C(K) space, Israel J. Math. 13 (1972), 246 -- 254. [6] J. Bourgain, New classes of Lp-spaces, Lecture Notes in Math. 889, Springer, 1981. [7] J. Bourgain and F. Delbaen, A class of special L∞-spaces, Acta Math. 145 (1980), 155 -- 176. [8] J. W. Calkin, Two-sided ideals and congruences in the ring of bounded operators in hilbert space, Ann. of Math. 42 (1941), no. 2, 839 -- 873. [9] M. Daws, R. Haydon, Th. Schlumprecht, and S. White, Shift invariant preduals of (cid:96)1(Z), preprint, arXiv:1101.5696. [10] N. Dunford and J.T. Schwartz, Linear Operators. Part 1: General Theory, Wiley- Blackwell, 1988. [11] M. Fabian, P. Habala, P. Hajek, V.M. Santalucia, J. Pelant, and V. Zizler, Functional Analysis and Infinite-Dimensional Geometry, Springer, 2001. [12] V. Ferenczi, Hereditarily finitely decomposable Banach spaces, Studia. Math. 123 (1997), no. 2, 135 -- 149. 135 [13] D. Freeman, E. Odell, and Th. Schlumprecht, The universality of (cid:96)1 as a dual space, Mathematische Annalen 351 (2011), no. 1, 149 -- 186. [14] C. Gohberg, A.S. Markus, and I.A. Fel'dman, Normally solvable operators and ideas associated with them, Bul. Akad. Stiince RSS Moldoven. 76 (1960), no. 10, 51 -- 70, English translation: Amer. Math. Soc. Transl. 61 (1967), 63-84. [15] W. Gowers and B. Maurey, Banach Spaces with Small Spaces of Operators, Mathema- tische Annalen 307 (1997), no. 4, 543 -- 568. [16] W.T. Gowers and B. Maurey, The unconditional basic sequence problem, J. Amer. Math. Soc. 6 (1993), 851 -- 874. [17] R. Haydon, Subspaces of the Bourgain Delbaen space, Studia Math. 139 (2000), 275 -- 293. [18] W.B. Johnson, D. Chen, and B. Zheng, Commutators on ((cid:80) (cid:96)q)p, Studia Math., to appear. [19] S. Kakutani, A generalisation of Brouwer's fixed point theorem, Duke Math. J. 7 (1941), 457 -- 459. [20] N. J. Laustsen, R.J. Loy, and C.J. Read, The lattice of closed ideals in the Banach algebra of operators on certain Banach spaces, J. Funct. Anal. 214 (2004), 106 -- 131. [21] N.J. Laustsen, Th. Schlumprecht, and A. Zs´ak, The lattice of closed ideals in the Banach algebra of operators on a certain dual Banach space, J. Operator Theory 56 (2006), 391 -- 402. [22] J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I and II, Springer, 1996. [23] V.I. Lomonosov, Invariant subspaces of the family of operators that commute with a completely continuous operator, Akademija Nauk SSSR. Funkcional' nyi Analiz iego Prilozenija 7 (1973), 55 -- 56. [24] B. Maurey, Operator Theory and Eexotic Banach Spaces, Online Lecture Notes, http://www.math.jussieu.fr/ maurey/articles/csp.pdf, 1994. [25] R. E. Megginson, An Introduction to Banach Space Theory, Springer, 1998. [26] G.J Murphy, C∗-Algebras and Operator Theory, Academic Press, 1990. [27] Th. Schlumprecht, N. Tomczak-Jaegermann, and V.G. Troitsky, On norm closed ideals in L((cid:96)p, (cid:96)q), Studia. Math. 179 (2007), no. 3, 239 -- 262. 136 [28] M. Tarbard, Hereditarily Indecomposable, separable L∞ Banach spaces with (cid:96)1 dual having few but not very few operators, J. London. Math. Soc 85 (2012), no. 3, 737 -- 764. 137
1605.02873
1
1605
2016-05-10T07:10:28
Recent Progress in Shearlet Theory: Systematic Construction of Shearlet Dilation Groups, Characterization of Wavefront Sets, and New Embeddings
[ "math.FA" ]
The class of generalized shearlet dilation groups has recently been developed to allow the unified treatment of various shearlet groups and associated shearlet transforms that had previously been studied on a case-by-case basis. We consider several aspects of these groups: First, their systematic construction from associative algebras, secondly, their suitability for the characterization of wavefront sets, and finally, the question of constructing embeddings into the symplectic group in a way that intertwines the quasi-regular representation with the metaplectic one. For all questions, it is possible to treat the full class of generalized shearlet groups in a comprehensive and unified way, thus generalizing known results to an infinity of new cases. Our presentation emphasizes the interplay between the algebraic structure underlying the construction of the shearlet dilation groups, the geometric properties of the dual action, and the analytic properties of the associated shearlet transforms.
math.FA
math
RECENT PROGRESS IN SHEARLET THEORY: SYSTEMATIC CONSTRUCTION OF SHEARLET DILATION GROUPS, CHARACTERIZATION OF WAVEFRONT SETS, AND NEW EMBEDDINGS GIOVANNI S. ALBERTI, STEPHAN DAHLKE, FILIPPO DE MARI, ERNESTO DE VITO, AND HARTMUT F UHR Abstract. The class of generalized shearlet dilation groups has recently been developed to allow the unified treatment of various shearlet groups and asso- ciated shearlet transforms that had previously been studied on a case-by-case basis. We consider several aspects of these groups: First, their systematic con- struction from associative algebras, secondly, their suitability for the character- ization of wavefront sets, and finally, the question of constructing embeddings into the symplectic group in a way that intertwines the quasi-regular represen- tation with the metaplectic one. For all questions, it is possible to treat the full class of generalized shearlet groups in a comprehensive and unified way, thus generalizing known results to an infinity of new cases. Our presentation emphasizes the interplay between the algebraic structure underlying the con- struction of the shearlet dilation groups, the geometric properties of the dual action, and the analytic properties of the associated shearlet transforms. 1. Introduction This chapter is concerned with several important aspects of modern signal anal- ysis. Usually, signals are modeled as elements of function spaces such as L2 or Sobolev spaces. To analyze such a signal and to extract the information of interest from it, the first step is always to decompose the signal into suitable building blocks. This is performed by transformation, i.e., the signal is mapped into function spaces on an underlying parameter set, and then the signal is processed and analyzed by studying and modifying the resulting coefficients. By now, a whole zoo of suitable transforms have been introduced and analyzed such as the Fourier transform, the Gabor transform, or the wavelet transform, just to name a few. Some of them have already been very succesful, e.g., the Fourier transform works excellently for sig- nals that are well-localized in the frequency domain, whereas wavelets are often the method of choice for the analysis of piecewise smooth signals with well-localized sin- gularities such as edges in an image. Which transform to choose obviously depends on the application, i.e., on the type of information one wants to detect from the signal. However, in recent years, it has turned out that a serious bottleneck still has to be removed. Most of the classical transforms such as the wavelet transform per- form suboptimally when it comes to the detection of directional information. The reason is very simple: most of these transforms are essentially isotropic, whereas directional information is of anisotropic nature. This observation triggered many Date: May 9, 2016. 2010 Mathematics Subject Classification. 42C15, 42C40, 46F12, 22D10. 1 2 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR innovative studies how to design new building blocks that are particularly tuned to this problem, such as curvelets [4], contourlets [16], ridgelets [3] and many others. In this chapter, we are in particular interested in one specific contribution to this problem, i.e., the shearlet approach. Shearlets are new affine representation sys- tems that are based on translations, shearings, and anisotropic dilations. We refer to the monograph [37] for an overview. Among all the new approaches, the shearlet transform stands out for the following reason: the continuous shearlet transform can be derived from a square-integrable representation of a specific group, the full shearlet group [9, 10, 13]. This property is not only of academic interest but has the important consequence that the whole powerful machinery derived in the realm of square-integrable group representations such as reproducing kernels, inversion for- mulas etc. can directly be employed. This feature of the shearlet transform clearly has strengthened the interest in the shearlet theory, and many important results concerning the group-theoretical background have been derived so far. It is the aim of this chapter to push forward, to clarify and to unify this theory with respect to several important aspects. Our main objectives can be described as follows. After the full shearlet group has been discovered, the question arose if other suitable concepts of shearlet groups could be constructed. A first example was the shearlet Toeplitz group [14], where the shearing part of the group has a Toeplitz structure. As we will see later in Subsection 3.3 of this chapter, the full shearlet group and the shearlet Toeplitz group are in a certain sense the "extreme" cases of a general construction principle. In this view, the full shearlet group corresponds to the nilpotency class n = 2, whereas the Toeplitz case corresponds to the nilpotency class n = d, where d denotes the space dimension. Therefore, one would conjecture that there should be a lot of examples "in between". Indeed, in [30], a positive answer has been given, and a first classification of low-dimensional shearlet groups has been derived. In this chapter, we further extend these results and present an approach to the systematic construction of suitable shearlet groups. The starting point is a general class of shearlet groups introduced in [30]. We say that a dilation group H is a shearlet group if every h ∈ H can be written as h = ±ds, d ∈ D, s ∈ S where D is a diagonal scaling subgroup and S denotes a connected, closed abelian matrix group, the shearing subgroup. The key to understanding and constructing shearing subgroups lies in the realization that their associated Lie algebras carry a very useful associative structure. This associative structure also greatly facilitates the task of identifying the diagonal scaling groups compatible with a given shearing subgroup. Through the notion of Jordan-Holder bases the problem of characterizing all suitable scaling group generators translates to a rather simple linear system of equations. It turns out that all examples known so far are special cases of this general construction. In recent studies, it has also been observed that shearlets provide a very powerful tool in microlocal analysis [32], e.g., to determine the local regularity of a function. In the one-dimensional case, pointwise smoothness can very efficiently be detected by examining the decay of the continuous wavelet transform as the scale parameter a tends to zero [34]. In the multivariate setting, pointwise smoothness does not cover all the geometic information one might be interested in. E.g., if the function under consideration exhibits singularities, one usually wants to know in which direction the function is singular. This can be described by the so-called wavefront set of a distribution. It has turned out that the continuous shearlet transform can be RECENT PROGRESS IN SHEARLET THEORY 3 employed to detect this wavefront set, once again by studying its decay as the scaling parameter tends to zero. This property has been first observed in [36], In [22], this concept has been generalized we also refer to [32] for an overview. to much more general classes of dilation groups. It has been shown that under natural assumptions, a wavefront set can again be detected by the decay of the voice transform. Essentially, two fundamental conditions are needed, that are related with the dual action of the dilation group H: the dual action must be microlocally admissible in direction ξ and it must satisfy the V -cone approximation property at ξ, see Section 4.1 for the precise definitions. If these properties hold for one point ξ0 in the open dual orbit, a characterization of wavefront sets is possible. In this chapter, we show that both properties are satisfied for our general construction of shearlet dilation groups, provided that the infinitesimal generator Y of the scaling subgroup satisfies Y = diag(1, λ2, . . . , λd), 0 < λi < 1, 2 ≤ i ≤ d. Consequently, characterizations of wavefront sets are possible for a huge subclass of our general construction. It is worth mentioning that anisotropic dilations are necessary for the detection of wavefront sets, in particular the classical (isotropic) continuous wavelet transform would not do the job. A third important issue we will be concerned with in this chapter is the relations of our general shearlet groups to other classical groups, in particular to the symplec- tic groups Sp(d, R). The symplectic groups are one of the most important classical groups, because they play a prominent role in classical mechanics. We therefore investigate to which extent our shearlet dilation groups can be embedded into sym- plectic groups, in a way that intertwines the quasi-regular representation with the metaplectic representation. For the full shearlet groups and the shearlet Toeplitz groups, these issues have been studied in [6], see also [35]. Their connected com- ponents can indeed be embedded into the symplectic groups, which yields group isomorphisms of the positive parts of shearlet groups with the so -- called TDS(d) subgroups that have already been studied in [15]. In this chapter, we generalize this result to dilation groups of the form G = Rd ⋊ H, where H is a subgroup of T (d, R)+ = {h ∈ GL(d, R) : h1,1 > 0 and hi,j = 0 for every i > j}. We show that for any such group there exists a group embedding φ : G → Sp(d, R), and that its quasi-regular representation is unitarily equivalent to µ ◦ φ, where µ denotes the metaplectic representation of Sp(d, R). Since the positive part of any shearlet group falls into this general category, the desired embeddings for shearlet groups follow from this result. Let us also mention the following very interesting fact: for the full shearlet dilation groups, such embeddings are never possible. Indeed, in [6] it has been shown that already for the two-dimensional full shearlet group there does not exist an injective continuous homomorphism into Sp(2, R) or into any of its coverings. Let us also mention a nice by-product of our construction. In recent studies [8, 10, 11, 13], an interesting relation of the shearlet approach to the coorbit theory derived by Feichtinger and Grochenig [17, 18, 19, 20] has been established. Based on a square integrable group representation, coorbit space theory gives rise to canonical associated smoothness spaces, where smoothness is measured by the decay of the underlying voice transform. In [8, 10, 11, 13], it has been shown that all the conditions needed in the coorbit setting can be satisfied for the full shearlet and the shearlet Toeplitz groups. In [27], the coorbit approach has been extended to much more general classes of dilation groups, and it turns out that the analysis 4 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR from [27] also carries over to the construction presented in this chapter, so that we obtain many new examples of coorbit spaces. In particular, we refer to [30] for explicit criteria for compactly supported functions that can serve as atoms in the coorbit scheme. This chapter is organized as follows. In Sections 2 and 3, we present our construc- tion of generalized shearlet dilation groups. After discussing the basic notations and definitions in the Subsections 2.1 and 2.2, in Subsection 2.3 we start with the sys- tematic investigation of the Lie algebras of shearing subgroups. One of the main results is Lemma 6 which provides a complete description of a shearing subgroup in terms of the canonical basis of its Lie algebra. This fact can be used to derive linear systems whose nonzero solutions determine the anisotropic scaling subgroups that are compatible with S (Lemma 9). These relationships are then used in Sec- tion 3 to derive a systematic construction principle. The canonical basis can be directly computed from the structure constants of a Jordan-Holder basis (Lemma 13). The power of this approach is demonstrated by several examples. In Section 4, we study the suitability of shearlet dilation groups for the characterization of wave- front sets. Here the main result is Theorem 28 which shows that shearlet groups with anisotropic dilations and suitable infinitesimal generators for the scaling sub- groups do the job. The proof is performed by verifying the basic conditions from [22]. The last section is concerned with the embeddings of shearing dilation groups into symplectic groups. The main result of this section is Theorem 33 which shows that the huge class of semidirect products of the form G = Rd ⋊ H, where H is a subgroup of T (d, R)+ can be embedded into Sp(d, R). 2. Generalities on shearlet dilation groups 2.1. Basic notations and definitions. This chapter is concerned with the con- struction and analysis of large classes of generalized shearlet transforms. These transforms are constructed by fixing a suitable matrix group, the so-called shearlet dilation group. By construction, these groups have a naturally associated isometric continuous wavelet transform, which will be the generalized shearlet transform. In this subsection, we summarize the necessary notation related to general continu- ous wavelet transforms in higher dimensions. We let GL(d, R) denote the group of invertible d × d-matrices. We use Id to denote the d × d identity matrix. The Lie algebra of GL(d, R) is denoted by gl(d, R), which is the space of all d × d matrices, endowed with the Lie bracket [X, Y ] = XY − Y X. Given h ∈ gl(d, R) its (operator) norm is denoted by khk = sup x≤1hx. We let exp : gl(d, R) → GL(d, R) denote the exponential map, defined by exp(X) = X k k! ∞Xk=0 known to converge absolutely for every matrix X. Given a closed subgroup H < GL(d, R), the associated Lie algebra of H is denoted by h, and it is de- fined as tangent space of H at Id, or, equivalently, as the set of all matrices X with exp(RX) ⊂ H. It is a Lie subalgebra of gl(d, R), i.e., it is closed under taking Lie brackets. RECENT PROGRESS IN SHEARLET THEORY 5 A matrix group of particular importance for the following is the group T (d, R) of upper triangular matrices with ones on the diagonal. Elements of T (d, R) are called unipotent. Its Lie algebra is the subspace t(d, R) ⊂ gl(d, R) of all strictly It is well-known that exp : t(d, R) → T (d, R) is a upper triangular matrices. homeomorphism [33]. In particular, whenever s ⊂ t(d, R) is a Lie subalgebra, the exponential image exp(s) is a closed, simply connected and connected matrix group with Lie algebra given by s. Conversely, any connected Lie subgroup S of T (d, R) is closed, simply connected and S = exp(s) where s ⊂ t(d, R) is the corresponding Lie algebra, see Theorem 3.6.2 of [39]. For the definition of generalized wavelet transforms, we fix a closed matrix group H < GL(d, R), the so-called dilation group, and let G = Rd ⋊ H. This is the group of affine mappings generated by H and all translations. Elements of G are denoted by pairs (x, h) ∈ Rd × H, and the product of two group elements is given by (x, h)(y, g) = (x + hy, hg). The left Haar measure of G is given by dµG(x, h) = det(h)−1dxdh, where dx and dh are the Lebesgue measure and the (left) Haar measure of Rd and H, respectively. The group G acts unitarily on L2(Rd) by the quasi-regular representation defined by (1) [π(x, h)f ](y) = det(h)−1/2f(cid:0)h−1(y − x)(cid:1) . We assume that H is chosen irreducibly admissible, i.e. such that π is an (irre- ducible) square-integrable representation. Recall that a representation is irreducible if the only invariant closed subspaces of the representation space are the trivial ones. Square-integrability of the representation means that there exists at least one nonzero admissible vector ψ ∈ L2(Rd) such that the matrix coefficient is in L2(G), which is the L2-space associated to the left Haar measure dµG. In this case the associated wavelet transform (x, h) 7→ hψ, π(x, h)ψi (2) Wψ : L2(Rd) ∋ f 7→ ((x, h) 7→ hf, π(x, h)ψi) ∈ L2(G) is a scalar multiple of an isometry, which gives rise to the wavelet inversion formula (3) f = Wψf (x, h)π(x, h)ψ dµG(x, h) , 1 cψZG where the integral is in the weak sense. We note that the definition of Wψf also makes sense for tempered distributions f , as soon as the wavelet ψ is chosen as a Schwartz function and the L2-scalar product is properly extended to a sesquilinear map S′ × S → C. Analogs of the wavelet inversion formula are not readily available in this general setting, but it will be seen below that the transform has its uses, for example in the characterization of wavefront sets. Most relevant properties of the wavelet transform are in some way or another connected to the dual action, i.e., the (right) linear action Rd × H ∋ (ξ, h) 7→ hT ξ. For example, H is irreducibly admissible if and only if the dual action has a single open orbit O = {hT ξ0 : h ∈ H} ⊂ Rd of full measure (for some ξ0 ∈ O), such that in addition the stabilizer group Hξ0 = {h ∈ H : hT ξ0 = ξ0} is compact [28]. This condition does of course not depend on the precise choice of ξ0 ∈ O. The dual action will also be of central importance to this chapter. 6 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR 2.2. Shearlet dilation groups. The original shearlet dilation group was intro- duced in [9, 10], as H =(cid:26)±(cid:18) a 0 a1/2 (cid:19) : a > 0, b ∈ R(cid:27) . b The rationale behind this choice was that the anisotropic scaling, as prescribed by the exponents 1, 1/2 on the diagonal, combines with the shearing (controlled by the parameter b) to provide a system of generalized wavelets that are able to swiftly adapt to edges of all orientations (except one). A mathematically rigourous formulation of this property is the result, due to Kutyniok and Labate, that the continuous shearlet transform characterizes the wavefront set [36]. Approximation- theoretic properties of a different, more global kind were the subject of the chapter [10] describing the so-called coorbit spaces defined in terms of weighted integrability conditions on the wavelet coefficients. The original shearlet dilation group has since been generalized to higher dimen- sions. Here, the initial construction was introduced in [13], and further studied, e.g., in [11, 5]. It is a matrix group in dimension d ≥ 3 defined by S = ±  a s1 aλ2 . . . sd−1 . . . aλd  : a > 0, s1, . . . , sd−1 ∈ R . Here λ2, . . . , λd are positive exponents, often chosen as λ2 = . . . = λd = 1/2. It should, however, be noted that they can be chosen essentially arbitrarily (even negative), without affecting the wavelet inversion formula. Coorbit space theory is applicable to all these groups as well [11, 27]. Furthermore, it was recently shown that the associated shearlet transform also characterizes the wavefront set [22], as long as the exponents λ2, . . . , λd are strictly between zero and one. A second, fundamentally different class of shearlet groups are the Toeplitz shear- let groups introduced in [14] and further studied in [8]. These groups are given by (5) (4)  : a > 0, s1, . . . , sd−1 ∈ R . H = ±   a s1 a s2 s1 . . . . . . s2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . sd−1 sd−2 ... ... s2 s1 a  Coorbit space theory can be applied to these groups as well [8, 27]. By [22, Lemma 4.10], the fact that H contains nontrivial multiples of the identity implies that H does not characterize the wavefront set. However, it will be shown below that by properly adjusting the diagonal entries, it is possible to construct a closely related group H ′ that does lend itself to the characterization of the wavefront set. A closer inspection of the two higher-dimensional families of shearlet group re- veals several common traits: fix one of the above-listed groups H. Then each h ∈ H RECENT PROGRESS IN SHEARLET THEORY 7 factors as h = ±diag(a, aλ2, . . . , aλd ) · u where the first factor denotes the diagonal matrix with the same diagonal entries as h, and the second factor u is unipotent. In fact, this factorization is necessarily unique. Furthermore, denoting by D the set of all diagonal matrices occurring in such factorizations, and by S the set of all unipotent ones that arise, it is easy to see that D (and consequently S) are closed subgroups of H. Finally, one readily verifies that the groups S that occur in the examples are in fact commutative. We will now use these properties to define a general class of shearlet dilation groups, that we will study in this chapter: Definition 1. Let H < GL(d, R) denote an irreducibly admissible dilation group. H is called generalized shearlet dilation group, if there exist two closed subgroups S, D < H with the following properties: matrix; (i) S is a connected abelian Lie subgroup of T (d, R); (ii) D = {exp(rY ) : r ∈ R} is a one-parameter group, where Y is a diagonal (iii) Every h ∈ H can be written uniquely as h = ±ds, with d ∈ D and s ∈ S. S is called the shearing subgroup of H, and D is called the diagonal complement or scaling subgroup of H. Remark 2. As noted in Subsection 2.1, S is closed, simply connected and the exponential map is a diffeomorphism from its Lie algebra s onto S. Remark 3. The class of shearlet dilation groups was initially defined in [30], and for some of the following results and observations, more detailed proofs can be found in that paper. In particular, it was shown there that coorbit space theory applies to all generalized shearlet dilation groups. In fact, it is possible to construct wavelet frames with compactly supported atoms, with frame expansions that, de- pending on the provenance of the signal, converge in a variety of coorbit space norms simultaneously. As will be seen below, shearlet dilation groups can be constructed systematically. The natural order in finding the constituent subgroups S, D is to first pick a can- didate for S, and then determine the infinitesimal generators of the one-parameter group D that are compatible with S. The details of this programme are given in the next subsections. 2.3. Shearlet dilation groups and their Lie algebras. It is the aim of this sub- section to give an overview of the most important structural properties of shearlet dilation groups. The following proposition gives a first characterization of these groups, see [30, Proposition 4.3]. Proposition 4. Let S denote a connected abelian subgroup of T (d, R). Then the following are equivalent: (i) S is the shearing subgroup of a generalized shearlet dilation group; (ii) There is ξ ∈ Rd such that S acts freely on ST ξ via the dual action, and in (iii) The matrix group A = {rs : s ∈ S, r ∈ R×} is an abelian irreducibly addition, dim(S) = d − 1; admissible dilation group. It is also a shearlet dilation group. 8 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR The fundamental observation made in [26, Remark 9] is that if A is abelian and admissible, as in part (iii) of the above proposition, then its Lie algebra a is in fact an associative subalgebra containing the identity element, hence it is closed under matrix multiplication. This associative structure is in many ways decisive. To begin with, one has the relations a = span(A) , A = a× i.e., A consists precisely of the multiplicatively invertible elements of the associative algebra a. We will see in Subsection 3.1 below that this connection to associative algebras can be used for the systematic -- even exhaustive -- construction of shearing subgroups. There is however a second ingredient, that is more directly related to the prop- erties of the dual action. It is described in the following lemma, see [30, Corollary 4.7]. We use e1, . . . , ed for the canonical basis of Rd. Lemma 5. Let S denote a connected abelian subgroup of T (d, R) of dimension d − 1, with Lie algebra s. Then the following are equivalent: (i) S is a shearing subgroup; (ii) There exists a unique basis X2, . . . , Xd of s with X T i e1 = ei, for all i = 2, . . . , d. We call the basis from part (ii) the canonical basis of s. The canonical basis plays a special role for the description of shearing subgroups. As a first indication of its usefulness, we note that all off-diagonal entries of the elements of shearing groups depend linearly on the entries in the first row. Lemma 6. Let S denote a shearing subgroup with Lie algebra s, and canonical basis X2, . . . , Xd of s. Then the following holds: (a) S = {Id + X : X ∈ s}. (b) Let h ∈ S be written as h = Then  1 h1,2 0 1 0 0 0 0 0 0 0 0 . . . h2,3 . . . 0 0 0 . . . . . . . . . . . . 0 0 . . . . . . ... . . . 1 0 h1,d h2,d ... ... hd−1,d 1 h = Id + h1,iXi . dXi=2 .  Proof. For part (a), denote the right-hand side by S1. Since s is an associative subalgebra consisting of nilpotent matrices, S1 consists of invertible matrices, and it is closed under multiplication. Furthermore, the inverse of any element of S1 can be computed by a Neumann series that breaks off after at most d terms: (Id + X)−1 = Id + d−1Xk=2 (−1)k−1X k , RECENT PROGRESS IN SHEARLET THEORY 9 and the result is again in S1. Hence S1 is a matrix group. It is obviously closed and connected, with tangent space of S1 at the identity matrix given by s. It follows that S1 is a Lie subgroup of T (d, R) and, hence, it is simply connected. Thus S and S1 are closed, connected, and simply connected subgroups sharing the same Lie algebra, hence they are equal. Now part (b) directly follows from (a) and the properties of the canonical basis. (cid:3) We now turn to the question of characterizing the scaling subgroups D that are compatible with a given shearing subgroup S. It is convenient to describe D in terms of its Lie algebra as well. Since D is one-dimensional, we have D = exp(RY ), with a diagonal matrix Y = diag(λ1, λ2, . . . , λd). We then have the following criterion [30, Proposition 4.5]: Proposition 7. Let S < GL(d, R) denote a shearing subgroup. Let Y denote a nonzero diagonal matrix, and let D := exp(RY ) the associated one-parameter group with infinitesimal generator Y . Then the following are equivalent: (i) H = DS ∪ (−DS) is a shearlet dilation group; (ii) For all X ∈ s we have [X, Y ] = XY − Y X ∈ s, and in addition the first diagonal entry of Y is nonzero. Remark 8. The above proposition states that H = S ⋊ R×, so that H is solvable group with two connected components, and each of them is simply connected. Since Y and rY , for nonzero r ∈ R, determine the same one-parameter subgroup, part (ii) of the proposition allows to fix λ1 = 1. Note that part (ii) is trivially fulfilled by isotropic scaling, which corresponds to taking 1 = λ1 = λ2 = . . . = λd. In what follows, we will be particularly interested in anisotropic solutions; our interest in these groups is mainly prompted by the crucial role of anisotropic scaling for wavefront set characterization. It turns out that the relation [Y, s] ⊂ s translates to a fairly transparent system of linear equations. Once again, the canonical basis X2, . . . , Xd of s proves to be particularly useful: As the following lemma shows, the adjoint action s ∋ X 7→ [Y, X] maps s into itself if and only if the Xi are eigenvectors of that map. The lemma uses the notation Ei,j for the matrix having entry one at row i and column j, and zeros everywhere else. Lemma 9. Let s denote the Lie algebra of a shearing subgroup, and let X2, . . . , Xd denote the canonical basis of s, given by (6) Xi = E1,i + dXj=2 dXk=j+1 di,j,kEj,k with suitable coefficients di,j,k. Let Y = diag(1, λ2, . . . , λd) be given. Then [Y, s] ⊂ s if and only if for all i = 2, . . . , d : λi = 1 + µi , and the vector (µ2, . . . , µd) is a solution of the system of linear equations given by for all (i, j, k) ∈ {2, . . . , d}3 with di,j,k 6= 0 : µi + µj = µk . (7) In particular, (µ2, . . . , µd) 7→ (1, 1 + µ2, . . . , 1 + µd) sets up a bijection between the nonzero solutions of (7) on the one hand and the anisotropic scaling subgroups D compatible with S on the other. 10 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR Remark 10. Note that (6) shows that di,j,k = (Xi)jk. Proof. We first note that the Ej,k are eigenvectors under the adjoint action of any diagonal matrix: (8) [Y, Ej,k] = (λj − λk)Ej,k . As a consequence, given any matrix X, the support of the matrix [Y, X] (i.e., the set of indices of its nonzero entries) is contained in the support of X. Note that Y normalizes s if and only if [Y, Xi] ∈ s for i = 2, . . . , d. Now the calculation (9) [Y, Xi] = [Y, E1,i] +X(j,k) di,j,k[Y, Ej,k] = (1 − λi)E1,i +X(j,k) di,j,k(λj − λk)Ej,k shows that the only (potentially) nonzero entry in the first row of [Y, Xi] occurs at the ith column, hence [Y, Xi] is in s if and only if it is a scalar multiple of Xi. In view of (9) and the linear independence of the Ej,k, this holds precisely when for all (i, j, k) ∈ {2, . . . , d}3 with di,j,k 6= 0 : 1 − λi = λj − λk . (10) Rewriting this system for µi = λi − 1, for i = 2, . . . , d, yields (7). (cid:3) Finally, let us return to properties of the associated shearlet transforms. In view of the central role of the dual action, it is important to compute the associated open dual orbit. Here we have the following, see [30, Proposition 4.5]: Proposition 11. Let S be a shearing subgroup, and D any diagonal complement of S. Then H = DS ∪ −DS acts freely on the unique open dual orbit given by O = R× × Rd−1. Note that the dual orbit is the same for all shearing groups. Somewhat surpris- ingly, the same can be said of the admissibility condition [30, Theorem 4.12]: Theorem 12. Let H < GL(Rd) denote a generalized shearlet dilation group. Then ψ ∈ L2(Rd) is admissible iff ZRd bψ(ξ)2 ξ1d dξ < ∞ . 3. A construction method for shearlet dilation groups 3.1. Constructing shearing subgroups. In this subsection we want to describe a general method for the systematic construction of shearing subgroups. Recall that given a shearing subgroup S with Lie algebra s, taking the Lie algebra a = RId ⊕ s and its associated closed matrix group A results in an abelian irreducibly admissible matrix group. Following [26], this entails that a is an associative matrix algebra. Furthermore, note that s consists of strictly upper triangular matrices, which entails that any product of d elements of s vanishes. These features of a can be described in general algebraic terms. Given a finite- dimensional, associative commutative algebra A, we call an element a ∈ A nilpotent if there exists n ∈ N such that an = 0. The set of all nilpotent elements in A is called the nilradical of A, denoted by N . We call A nilpotent if every element of A is nilpotent. N is an ideal in A, i.e., given a ∈ N and an arbitrary b ∈ A, one has (ab)n = anbn = 0 for sufficiently large n, i.e. ab is again in the nilradical. We call the algebra A irreducible (over R) if it has a unit element 1A satisfying 1Ab = b for RECENT PROGRESS IN SHEARLET THEORY 11 all b ∈ A, and such that A = R · 1A ⊕ N holds. Note that N determines A in this case, and we will freely switch between A and N in the following. Now the above considerations show that a is an irreducible associative commu- tative algebra. In the remainder of this subsection, we will be concerned with a converse to this statement, i.e., with the construction of shearing subgroups from an abstractly given irreducible associative algebra. Assume that A is an irreducible commutative associative algebra of dimension d, and denote its nilradical by N . We let n(A) = min{k ∈ N : ak = 0, ∀a ∈ N} , which is called the nilpotency class of A. Letting N k = {a1 . . . ak : ai ∈ N}, for k ≥ 1, and N 0 = N , one can prove that n(A) = min{k ∈ N : N k = {0}} ≤ d . By definition of the nilpotency class, we obtain that N n(A)−1 6= {0}, and for all a ∈ N n(A)−1 and b ∈ N , it follows that ab = 0. Hence, choosing a nonzero ad ∈ N n(A)−1, we find that Id := R · ad is an ideal in N ; in fact, we get NId = {0}. Applying the same reasoning to the algebra N /Id (and choosing any representative modulo Id) produces a second element ad−1 with the property that Id−1 = span(ad−1, ad) fulfills NId−1 ⊂ Id. Further repetitions of this argument finally yield a basis a2, . . . , ad of N , that we supplement by a1 = 1A to obtain a basis of A with the property (11) NIk ⊂ Ik+1 for 1 ≤ k < d , and I2 = N . We call a basis a2, . . . , ad of N satisfying condition (11) a Jordan- Holder basis of N . The existence of a Jordan-Holder basis can be also proved by referring to a general result about nilpotent representations of nilpotent algebras, see Theorem 3.5.3 of [39]. Indeed, regard N as nilpotent algebra and A as a vector space. It is easy to check that the (regular) representation ρ of the Lie algebra N acting on A as ρ(a)b = ab is nilpotent, so that there exists a basis {a1, . . . , ad} of A such that for each a ∈ N , the endomorphism ρ(a) is represented by a strictly upper triangular matrix Ψ(a) ∈ gl(d, R) according to the canonical isomorphism ρ(a)aj = dXk=1 Ψ(a)j,k ak j = 1, . . . , d . Since ρ(a)1A = a, it is always possible to choose a1 = 1A and, by construction, for all a ∈ N and for i = 1, . . . , d − 1 ρ(a)span{ai, . . . , ad} ⊂ span{ai+1, . . . , ad} ρ(a)ad = 0 . These bases provide accesss to an explicit construction of an associated shearing subgroup, explained in detail in the next lemma. Recall the notation Ei,j for the matrix possessing entry one in row i, column j, and zeros elsewhere. Note that the map Ψ : A → gl(d, R) in the following lemma coincides with the identically denoted map that we just introduced. 12 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR Lemma 13. Let A denote an irreducible commutative associative algebra of di- mension d with nilradical N possessing the Jordan-Holder basis Let a1 = 1A, and let ψ : Rd → A denote the induced linear isomorphism a2, . . . , ad ∈ A . ψ((x1, . . . , xd)T ) = xiai . dXi=1 Let Ψ : A → gl(d, R) denote the associated linear map satisfying for all a ∈ A and for all x ∈ Rd: ψ−1(a · ψ(x)) = Ψ(a) · x . (a) The set is a shearing subgroup, with associated Lie algebra given by S = {Id + Ψ(a)T : a ∈ N} s = {Ψ(a)T : a ∈ N} . (b) Defining Xi = Ψ(ai)T , for i = 1, . . . , d, we get that X1 is the identity matrix, and X2, . . . , Xd is the canonical basis of s in the sense of Lemma 5. (c) Let (di,j,k)1≤i,j,k≤d denote the structure constants associated to the basis, defined by the equations (12) (13) Then for all 1 ≤ i, j ≤ d : aiaj = di,j,kak . dXk=1 Xi = di,1,1 ... di,d,1 di,1,2 ... di,d,2 . . . di,1,d ... ... . . . di,d,d  . (d) We note the following nontrivial properties of the di,j,k, valid for all 1 ≤ i, j, k ≤ d: di,j,k = dj,i,k , d1,j,k = δj,k , di,j,k = 0 whenever k ≤ max(i, j) . In particular, we get for 2 ≤ j ≤ d (14) Xi = E1,i + di,j,kEj,k . dXj=2 dXk=j+1 Proof. We start with part (c). Since multiplication with a1 = 1A is the identity operator, the statement about X1 is clear. Let 1 ≤ i, j ≤ d. By definition of ψ, we have ψ(ej) = aj, and hence by definition of Ψ Ψ(ai)ej = ψ−1(ai · ψ(ej)) = ψ−1(aiaj) dXk=1 = ψ−1 dXk=1 di,j,kak! = di,j,kek . Hence the jth column of Ψ(ai) is the vector (di,j,1, . . . , di,j,d)T , and its transpose is the jth row of Ψ(ai)T . This shows (13). RECENT PROGRESS IN SHEARLET THEORY 13 Now, with (c) established, the equation a1aj = aj for i = 2, . . . , d, yields that d1,j,k = δj,k, which also takes care of part (b). Fur- thermore, the fact that aiaj = ajai ensures that di,j,k = dj,i,k. Finally, recall that AIi ⊂ Ii+1 by (11), which entails aiaj ∈ Ii+1, and thus di,j,k = 0 whenever k ≤ i. Since di,j,k = dj,i,k, we then obtain more generally that k ≤ max(i, j) entails di,j,k = 0. Now equation (14) is clear, and (d) is shown. In order to prove (a), we first note that Ψ is a homomorphism of associative algebras, hence s = Ψ(N )T is a commutative associative matrix algebra. In par- ticular, it is also an abelian Lie-subalgebra. Furthermore, the relation di,j,k = 0 whenever k ≤ max(i, j) ensures that the basis X2, . . . , Xd consists of strictly upper triangular matrices. In addition, d1,j,k = δj,k entails that X2, . . . , Xd is indeed a canonical basis, and thus Lemma 5 gives that the associated Lie group is a shearing subgroup. Now part (a) of Lemma 6 yields (a) of the current lemma, and (b) is also shown. (cid:3) Remark 14. It is natural to ask whether the construction of shearing subgroups S from irreducible commutative associative algebras A, as described in Lemma 13, is exhaustive. The answer is yes. To see this, consider the Lie algebra s of a shearing subgroup S. Let X2, . . . , Xd be the canonical basis of s. Since Xj is strictly upper triangular, and the first row of Xi equals eT i , it follows that the first i entries of the first row of XiXj vanish. This product is again in the span of the Xk, hence XiXj =Xk>i di,j,kXk , with suitable coefficients di,j,k. But the fact that the sum on the right-hand side starts with k = i + 1 shows that the basis X2, . . . , Xd is a Jordan-Holder basis of the nilpotent associative matrix algebra s. If one now applies the procedure from Lemma 13 (with ai = Xi), direct calculation allows to verify that Ψ(X)T = X for all X ∈ s. Hence every shearing subgroup arises from the construction in Lemma 13. In particular, the observations concerning the structure constants di,j,k made in part (d) of Lemma 13 also apply to the di,j,k in Lemma 9. Remark 15. A further benefit of the above construction of shearing groups via associative algebras is that it settles the question of conjugacy as a byproduct. By Theorem 13 in [26] and the remarks prior to that result, one sees that two shearing subgroups S1 and S2 are conjugate iff their Lie algebras are isomorphic as associative algebras. In particular, following the observation made in [26, Theorem 15], in dimension d ≥ 7 there exist uncountably many nonconjugate shearing subgroups. 3.2. An inductive approach to shearlet dilation groups. For possible use in inductive proof strategies, we note a further consequence of the block structure: Proposition 16. Let H = ±DS < Gl(d, R) denote a shearlet dilation group, with d ≥ 3, and let H1 =(cid:26)h′ ∈ GL(d − 1, R) : ∃ h ∈ H, z ∈ Rd−1, s ∈ R\ {0} with h =(cid:18) h′ 0 z s (cid:19)(cid:27) . Then H1 is a shearlet dilation group as well. 14 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR Conversely, the elements of H can be described in terms of H1 as follows: There exists a map y : H1 → Rd−1 such that we can write each h ∈ H uniquely as h(h1, r) =  h1 0 . . . . . . 0 r y1(h1) y2(h1) ... yd−2(h1) yd−1(h1) ,  with h1 ∈ H1, r ∈ R. 3.3. Examples. As a result of the previous subsections, we obtain the following general procedure for the systematic construction of shearlet dilation groups: (1) Fix a nilpotent associative algebra N . (2) Pick a Jordan-Holder basis a2, . . . , ad of N , and compute the canonical basis X2, . . . , Xd of the Lie algebra s of the associated shearing subgroup. Note that this amounts to determining the structure constants (di,j,k)1≤i,j,k≤d. The shearing subgroup is then determined as S = Id + s. (3) In order to determine the diagonal scaling groups that are compatible with S, set up and solve the linear system (7) induced by the nonvanishing di,j,k. We will now go through this procedure for several examples or classes of exam- ples. Example 17. We start out with the simplest case of a nilpotent algebra N of dimension d−1, namely that of nilpotency class 2. Here one has ab = 0 for any a, b ∈ N , and it is clear that for two such algebras, any linear isomorphism is an algebra isomorphism as well. Picking any basis a2, . . . , an of N , we obtain Xi = E1,i. In particular, the linear system (7) is trivial. Hence any one-parameter diagonal group can be used as scaling subgroup. We thus recover the groups described in (4). Example 18. Another extreme class of nilpotent algebras of dimension d is that of nilpotency class d. Here there exists b ∈ N with bd−1 6= 0. This implies that b, . . . , bd−1 are linearly independent, and then it is easily seen that ai = bi−1, for i = 2, . . . , d, defines a Jordan-Holder basis of N . In this example, the defining relations read (15) aiaj = ai+j−1, 2 ≤ i, j, i + j − 1 ≤ d, and the resulting canonical Lie algebra basis is then determined as X2 = 0 1 0 1 . . . . . . . . .   . . . 0 1 0 , X3 =  0 0 0 1 0 . . . 1 . . . 0  . . . 0 0 1 0 0 , . . . , RECENT PROGRESS IN SHEARLET THEORY 15 Xd = 0 . . . . . . . . . 0 0 1 0 ... ... ... 0   . Thus we see that the resulting shearing subgroup is that of the Toeplitz shearlet group from (5). The linear system (7) becomes It is easy to see that all solutions of this system are given by µi + µj = µi+j−1, for 2 ≤ i, j, i + j − 1 ≤ d. µj = (j − 1)δ, j = 2, . . . , d with δ an arbitrary real parameter. Thus the scaling subgroups compatible with the Toeplitz dilation group are precisely given by exp(Rdiag(1, 1 + δ, . . . , 1 + (d − 1)δ)) , with δ ∈ R arbitrary. Remark 19. For d = 3, the two above listed cases are all possible examples of shearing subgroups, and not even just up to conjugacy. In particular, we find that all shearing subgroups in dimension 3 are compatible with anisotropic dilations. We now turn to the shearing subgroups in dimension 4, with focus on the groups not covered by (4) and (5). Example 20. Since the nilpotency classes n = 2, 4 are already covered by the previous examples, the remaining 4-dimensional cases of irreducible algebras A all have nilpotency class 3. It is shown in [26] that A ∼= R[Y1, Y2]/(Y 3 1 , Y1Y2), with α ∈ {−1, 0, 1}. Here, R[Y1, Y2] denotes the algebra of polynomials with real coefficients and indeterminates Y1, Y2, and J = (Y 3 2 − αY 2 1 , Y1Y2) denotes the ideal generated by the three polynomials. Then the nilradical N is generated by Y1 +J , Y2 + J . We choose the basis a2 = Y1 +J , a3 = Y2 +J , a4 = Y 2 1 +J , and obtain as the only nonzero relations 2 −αY 2 1 , Y 2 1 , Y 2 a2 2 = a4 , a2 3 = αa4 . This allows to conclude that a2, a3, a4 is indeed a Jordan-Holder basis. Follow- ing Lemma 13 (c), we can read off the canonical basis of the associated shearing subgroup as X2 = 0 0 0 0 1 0 0 0 0 0 0 0 0 1 0 0  , X3 = 0 0 0 0 0 0 1 0 0 0 0 0 α 0 0 0  , X4 = 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0  . We next determine the compatible scaling subgroups. In the case α 6= 0, we obtain the system of equations 2µ2 = µ4, 2µ3 = µ4. Thus the infinitesimal generators of scaling subgroups are of the form Y = diag(1, 1+ δ, 1 + δ, 1 + 2δ), with δ ∈ R arbitrary. 16 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR In the case α = 0, we only get one equation, namely 2µ2 = µ4, showing that here the compatible infinitesimal generators are of the form Y = diag(1, 1 + δ1, 1 + δ2, 1 + 2δ1), with δ1, δ2 ∈ R arbitrary. Finally, we give an example of a shearing subgroup which is only compatible with isotropic scaling. It is based on the same algebra as Example 18 (with d = 4), and as a result the associated shearing subgroups are conjugate. Recall that the groups in Example 18 are compatible with anisotropic scaling. This illustrates an important, somewhat subtle point: While the precise choice of Jordan-Holder basis in the procedure described in Lemma 13 is immaterial if one is just interested in guaranteeing the shearing subgroup property, it may have a crucial influence on the availability of compatible anisotropic scaling subgroups. Example 21. Let A = R[X]/(X 4). We use the Jordan-Holder algebra a2 = X + X 2 + (X 4), a3 = X 2 + (X 4), a4 = X 3 + (X 4). This leads to the following nonzero relations a2 2 = a3 + 2a4, a2a3 = a4, which gives rise to the basis X2 = 0 0 0 0 1 0 0 1 0 0 0 0 0 2 1 0  , X3 = 0 0 0 0 0 1 0 0 0 0 0 0 0 1 0 0  , X4 = 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0  . Now the nonzero entries in the matrix X2 imply that the linear system (7) contains the equations 2µ2 = µ3, 2µ2 = µ4, µ2 + µ3 = µ4. The first two equations imply µ3 = 2µ2 = µ4, and then the third equation yields µ2 = 0. Hence this shearing subgroup is only compatible with isotropic scaling. 4. Anisotropic scaling and wavefront set characterizations In this section we investigate the suitability of the various groups for microlocal analysis. The idea is to verify the criteria derived in [22] that allow to establish the suitability of a dilation group for the characterization of the wavefront set via wavelet coefficient decay. As it will be seen, this property only depends on the scaling subgroup. 4.1. Criteria for wavefront set characterization. Throughout this subsection H is an irreducibly admissible matrix group, i.e. its dual action has a single open orbit O ⊂ Rd, with associated compact fixed groups. We use V ⋐ O to denote that the closure of V inside O is compact. Given R > 0 and x ∈ Rd, we let BR(x) and BR (x) denote the open/closed ball with radius R and center x, respectively. We let Sd−1 ⊂ Rd denote the unit sphere. By a neighborhood of ξ ∈ Sd−1, we will always mean a relatively open set W ⊂ Sd−1 with ξ ∈ W . Given R > 0 and an open set W ⊂ Sd−1, we let C(W ) := {rξ′ : ξ′ ∈ W, r > 0} =(cid:26)ξ ∈ Rd \ {0} : ξ ξ ∈ W(cid:27) , C(W, R) := C(W ) \ BR(0). RECENT PROGRESS IN SHEARLET THEORY 17 Both sets are clearly open subsets of Rd \ {0} and thus of Rd. Given a tempered distribution u, we call (x, ξ) ∈ Rd × Sd−1 a regular directed point of u if there exists ϕ ∈ C∞ c (Rd), identically one in a neighborhood of x, as well as a ξ-neighborhood W ⊂ Sd−1 such that for all N ∈ N there exists a constant CN > 0 with (16) for all ξ′ ∈ C (W ) : cϕu(ξ′) ≤ CN (1 + ξ′)−N . We next formally define the sets Ki and Ko which will allow to associate group elements to directions. Definition 22. Let ∅ 6= W ⊂ Sd−1 be open with W ⊂ O (which implies C (W ) ⊂ O). Furthermore, let ∅ 6= V ⋐ O and R > 0. We define as well as Ki(W, V, R) :=(cid:8)h ∈ H : h−T V ⊂ C(W, R)(cid:9) Ko(W, V, R) :=(cid:8)h ∈ H : h−T V ∩ C(W, R) 6= ∅(cid:9) . If the parameters are provided by the context, we will simply write Ki and Ko. Here, the subscripts i/o stand for "inner/outer". We now define what we mean by dilation groups characterizing the wavefront set. We first extend the continuous wavelet transform to the space of tempered dis- tributions. I.e., we use Wψu, for a Schwartz wavelet ψ and a tempered distribution u. Definition 23. The dilation group H characterizes the wavefront set if there exists a nonempty open subset V ⋐ O with the following property: For all 0 6= ψ ∈ S(Rd) following statements are equivalent: with supp(bψ) ⊂ V , for every u ∈ S′(Rd) and all (x, ξ) ∈ Rd × (O ∩ Sd−1), the (a) (x, ξ) is a regular directed point of u. (b) There exists a neighborhood U of x, some R > 0 and a ξ-neighborhood W ⊂ Sd−1 such that for all N ∈ N there exists a constant CN > 0 such that for all y ∈ U , and for all h ∈ Ko(W, V, R) the following estimate holds: Wψu(y, h) ≤ CNkhkN . Note that the definition excludes a set of directions ξ from the analysis of the wavefront set, namely the directions not contained in O ∩ Sd−1. These directions always constitute a set of measure zero. Recall from Proposition 11 that in the case of shearlet dilation groups, this exceptional set is given by ({0} × Rd−1) ∩ Sd−1. We next recall the sufficient conditions for dilation groups that characterize the wavefront set, as established in [22]. The first one is related to the problem that one would like to interpret the norm as a scale parameter. Definition 24. Let ξ ∈ O ∩ Sd−1 and ∅ 6= V ⋐ O. The dual action is called V - microlocally admissible in direction ξ if there exists a ξ-neighborhood W0 ⊂ Sd−1∩O and some R0 > 0 such that the following hold: (1) There exist α1 > 0 and C > 0 such that kh−1k ≤ C · khk−α1 holds for all h ∈ Ko(W0, V, R0). 18 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR (2) There exists α2 > 0 such that ZKo(W0,V,R0) khkα2 dh < ∞. The dual action is called microlocally admissible in direction ξ if it is V -microlocally admissible in direction ξ for some ∅ 6= V ⋐ O. The second important condition is contained in the following definition. It can be understood as formalizing the ability of the associated wavelet systems to be able to make increasingly fine distinctions between different directions, as the scales go to zero. Definition 25. Let ξ ∈ O∩ Sd−1 and ∅ 6= V ⋐ O. The dual action has the V -cone approximation property at ξ if for all ξ-neighborhoods W ⊂ Sd−1 and all R > 0 there are R′ > 0 and a ξ-neighborhood W ′ ⊂ Sd−1 such that Ko(W ′, V, R′) ⊂ Ki(W, V, R). We now have the following [22, Corollary 4.9]: Theorem 26. Assume that the dual action is V -microlocally admissible at some ξ0 ∈ O and has the V -cone approximation property at ξ0, for some nonempty open subset V ⊂ O. Then H characterizes the wavefront set. Remark 27. The property of characterizing the wavefront set is linked to anisotropic scaling, in the following sense: If H characterizes the wavefront set, then H ∩ R+ · Id = {Id} , by [22, Lemma 4.10]. Hence if H is a shearlet dilation group characterizing the wavefront set, its shearing subgroup must admit at least one anisotropic compatible scaling subgroup. This excludes the shearing group constructed in Example 21. Theorem 26 therefore implies that every group failing the anisotropy criterion H ∩ R+ · Id = {Id} must necessarily fail either the microlocal admissibility or the cone approximation property. It is in fact the latter that breaks down, as noted in [22, Lemma 4.4]. These considerations highlight the importance of understanding when a given shearing groups admits anisotropic scaling. 4.2. Characterization of the wavefront set for shearlet dilation groups. We can now state a very general theorem concerning the ability of shearlet groups to characterize the wavefront set. Note that there are no conditions on the shearing subgroups. Theorem 28. Let H be a shearlet dilation group and let Y = diag(1, λ2, . . . , λd) denote the infinitesimal generator of the scaling subgroup. If 0 < λi < 1 holds, for all 2 ≤ i ≤ d, then H characterizes the wavefront set. Remark 29. We can now quickly go through the examples of shearing subgroups in Subsection 3.3 and show that for most cases, there exists a compatible anisotropic scaling subgroup that allows to characterize the wavefront set. Writing λi = 1 + µi as in Lemma 5, the condition from Theorem 28 translates to −1 < µi < 0, for 2 ≤ i ≤ d. Apart from the group in Example 21, which was specifically constructed to not allow any anisotropic scaling, all other shearing groups can be combined with a compatible scaling group in such a way that the resulting shearlet transform RECENT PROGRESS IN SHEARLET THEORY 19 fulfills the conditions of Theorem 28, and therefore characterizes the wavefront set. Note that this was previously known only for the original shearlet group [36, 22]. In particular, we may combine the Toeplitz shearing subgroup with the scaling subgroup with exponents (1, 1− δ, . . . , 1− (d− 1)δ), and choosing δ ∈ (0, 1/(d− 1)) guarantees that the Toeplitz shearlet transform characterizes the wavefront set. The proof of the Theorem amounts to verifying the cone approximation prop- erty and microlocal admissibility of the dual action, and this will be carried out in the following two propositions. For the remainder of this section, we fix a shear- let dilation group H with infinitesimal generator diag(1, λ2, . . . , λd) of the scaling subgroup. We let λmax = maxi≥2 λi, and λmin = mini≥2 λi. Proposition 30. If λmax < 1, there exists an open subset ∅ 6= V ⋐ O such that the dual action of H on the orbit O has the V -cone approximation property at (1, 0, . . . , 0)T ∈ Sd−1 ∩ O. Proof. We will employ the structural properties of shearing subgroups derived in Section 2.2. We let S and D denote the shearing and scaling subgroups of H, respectively. The infinitesimal generator of D is a diagonal matrix with the entries 1, λ2, . . . , λd. We let X2, . . . , Xd denote the canonical basis of s, consisting of strictly upper triangular matrices Xi. By Lemma 6, each h ∈ S is uniquely described by h = h(t, 1) = Id + tiXi , dXi=2 where t = (t2, . . . , td)T denotes the vector of first row entries of h(t, 1). For H, we thus obtain the global chart h(t, a) = Id + tiXi! sgn(a)diag(a,aλ2 , . . . ,aλd ) ∈ GL(d, R), dXi=2 with (t, a) ∈ Rd−1 × R×. For the purpose of the following computations, it is possible and beneficial to slightly modify this construction and replace h(t, 1) by its inverse. Thus, every h ∈ H can be written (uniquely) as h = ±h(t, 1)−1h(0, a) with t ∈ Rd−1 and a ∈ (0, +∞). The dual action is then given by (h−1)T = ±(h(0, a)−1h(t, 1))T = ± Id + i ! h(0, a−1), tiX T (17) dXi=2 where by construction (18) Id + tiX T i =(cid:18)1 t 0T Id−1 + A(t)T(cid:19) , dXi=2 with A(t) being a (d − 1) × (d − 1) strictly lower-triangular matrix satisfying (19) kA(t)k ≤ Ct with a constant C depending only on H. We now parametrise the open orbit O by the global chart provided by affine coordinates Ω : R× × Rd−1 → O Ω(τ, v) = τ (1, vT )T , 20 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR and Sd−1 ∩ O by the corresponding diffeomorphism to its image . (1, vT )T ω : Rd−1 → Sd−1 ∩ O, ω(v) = p1 + v2 Wǫ = {v ∈ Rd−1 : v < ǫ} = Bǫ(0), Given ǫ > 0, we set since {Wǫ : ǫ > 0} is a neighbourhood basis of the origin in Rd−1 and {ω(Wǫ) : ǫ > 0} is a neighbourhood basis of ξ0 = (1, 0, . . . , 0) ∈ Sd−1 ∩ O. Furthermore, for fixed 0 < τ1 < τ2 and ǫ0 > 0 the set V = Ω( (τ1, τ2) × Wǫ0 ) is an open subset with V ⋐ O. v ∈ Wǫ0 , and we get Given h ∈ H, as in (17), and ξ ∈ V , then ξ = Ω(τ, v) with τ1 < τ < τ2 and (h−1)T ξ = ±τ(cid:18)1 t 0T Id−1 + A(t)T(cid:19)(cid:18)a−1 v′ (cid:19) = ±a−1τ(cid:18) t + (Id−1 + A(t)T )v′′(cid:19) 1 where v′, v′′ ∈ Rd−1 have components given by v′ all i = 2, . . . , d. Hence i = a−λivi and v′′ i = a1−λivi for ǫ0 )(cid:1) (h−1)T (V ) = Ω(cid:0) (±a−1τ1,±a−1τ2) × (t + (Id−1 + A(t)T )W a i = a1−λi vi, v < ǫ0}. ǫ0 = {v′′ ∈ Rd−1 : v′′ where W a Fix now R > 0 and a neighborhood W ⊂ Sd−1 ∩ O of ξ0. Without loss of generality we can assume that W = ω(Wǫ) for some ǫ > 0. Furthermore, since (R, +∞) × Wǫ ⊂ Ω−1(C(ω(Wǫ), R)) ⊂ ( , +∞) × Wǫ, where the last inclusion on the right holds if ǫ ≤ 1, then the V -cone approximation property holds true if there exist R′ > 0 and 0 < ǫ′ ≤ 1 such that for all h ∈ H satisfying , +∞) × Wǫ ⊂ ( √1 + ǫ2 R 2 R (20b) Take R′ > 0 and 0 < ǫ′ < √3, which we will fix later on as functions of R and ǫ, and h ∈ H as in (17). If h = −(h(0, a)−1h(t, 1))T then so that (20a) implies that h = +(h(0, a)−1h(t, 1))T and R′ 2 (cid:18) ( (cid:18) ( , +∞) × Wǫ′(cid:19) ∩(cid:0) (−a−1τ2,−a−1τ1) × (t + (Id−1 + A(t)T )W a , +∞) × Wǫ′(cid:19) ∩(cid:0) (a−1τ1, a−1τ2) × (t + (Id−1 + A(t)T )W a ǫ0(cid:1) 6= ∅. Wǫ′ ∩(cid:0)t + (Idd−1 +A(t)T )W a R′ < 2a−1τ2, ǫ0 )(cid:1) = ∅, ǫ0 )(cid:1) 6= ∅. R′ 2 If we choose R′ > 2τ2, the first inequality gives Hence (21) a < 2τ2 R′ < 1, (20a) it holds that R′ 2 (h−1)T (V ) ∩ Ω(cid:18) ( , +∞) × Wǫ′(cid:19) 6= ∅, (h−1)T (V ) ⊂ Ω ( (R, +∞) × Wǫ ) . RECENT PROGRESS IN SHEARLET THEORY 21 and, since a < 1, setting λmax = max{λ2, . . . , λd}, clearly (22) W a ǫ0 ⊂ Wǫ0a1−λmax . By the above inclusion, since Wǫ′ ∩(cid:0)t + (Id−1 + A(t)T )W a ξ ∈ Wǫ0a1−λmax such that t + ξ + A(t)T ξ < ǫ′. Hence, triangle inequality, (19) and (21) give ǫ0(cid:1) 6= ∅, then there exists t < ǫ′ + (1 + kA(t)Tk)ξ ≤ ǫ′ + (1 + Ct)a1−λmax ǫ0 ≤ ǫ′ + ( 2τ2 R′ )1−λmax (1 + Ct)ǫ0 ≤ 2ǫ′ + 1 2t, where the last inequality holds true provided that (23) R′ > 2τ2 max{1, ( ǫ0 ǫ′ ) 1 1−λmax , (2Cǫ0) 1 1−λmax }. Hence, if (20a) holds true with R′ satisfying (23), then (24a) (24b) (24c) ( 2τ2 R′ < 1 a < t < 4ǫ′ 2τ2 R′ )1−λmax ǫ0 < min{ǫ′, 1 2C }. The condition (20b) is equivalent to (a−1τ1, a−1τ2) × (t + Id−1 + A(t)T )W a ǫ0 ) ⊂ (R, +∞) × Wǫ, which is ensured by a−1τ1 > R and, recalling (22), by t+(Id−1+A(t)T )Wǫ0a1−λmax ⊂ Wǫ. R′ . Taking into account (23), By (24a) the first condition is satisfied if τ1/R > 2τ2 it is sufficient to assume that (25) R′ > 2τ2 max{1, ( ǫ0 ǫ′ ) 1 1−λmax , (2Cǫ0) 1 1−λmax , R τ1 }. To ensure that t+(Id−1 +A(t)T )Wǫ0a1−λmax ⊂ Wǫ, note that, for all ξ ∈ Wǫ0a1−λmax , conditions (19), (24a), and (24b) give t + (Id−1 + A(t)T )ξ ≤ t + (1 + Ct)ξ ≤ t + (1 + Ct)a1−λmax ǫ0 < 4ǫ′ + (1 + C4ǫ′)( ≤ 4ǫ′ + ǫ′ + 2ǫ′ = 7ǫ′, 2τ2 R′ )1−λmax ǫ0 where the last inequality follows from (24c). Hence, with the choice ǫ′ = min{1, ǫ/7} and R′ satisfying (25) for all ξ ∈ Wǫ0a1−λmax , t + (Id−1 + A(t)T )ξ < ǫ, so that (20b) holds true for all h ∈ H satisfying (20a). Remark 31. The proof does not make use of the fact that the shearlet group S is abelian. The proof is based only on the following two properties of S (cid:3) (a) a global smooth chart t 7→ s(t) from Rd−1 onto S; 22 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR (b) for all t ∈ Rd the dual action of s(t) is of the form (s(t)−1)T =(cid:18)1 t B(t)(cid:19) 0T where kB(t)k ≤ C1 + C2t for a suitable choice of C1 and C2. With the cone approximation property already established, the remaining con- dition is quite easy to check. Proposition 32. If 0 < λmin ≤ λmax < 1, there exists an open subset ∅ 6= V ⋐ O such that the dual action of H on the orbit O is V -microlocally admissible in direction (1, 0, . . . , 0) ∈ Sd−1 ∩ O. Proof. We retain the notations from the previous proof, as well as the open set V = Ω( (τ1, τ2) × Wǫ0 ) , with τ1 < 1 < τ2. Since we assume λmax < 1, the cone approximation property holds, and then condition (2) of Definition 24 follows from condition (1) by [22, Lemma 4.7]. In addition, the cone approximation property allows to replace Ko in that condition by the smaller set Ki. In short, it remains to prove the existence of α > 0 and C′′ > 0 such that kh−1k ≤ C′′khk−α holds for all h ∈ Ki(ω(Wǫ), V, R), for suitable ǫ, R > 0. In the following com- putations, we let ǫ = 1 and R > 2. Now assume that h = ±h(t, 1)−1h(0, a) ∈ Ki(ω(Wǫ), V, R), which means that h−T V ⊂ C(ω(Wǫ), R). This implies in partic- ular that h−T 1 0 ... 0  = ±h(t, 1)T a−1 0 ... 0  1 0 ... 0 = ±h(t, 1)T h(0, a)−1  = ±a−1(cid:18) 1 t (cid:19) ∈ C(ω(Wǫ), R) . (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) h−T  2a−1 ≥ (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) 1 0 ... 0 > R > 2 , This implies that the sign is in fact positive. Furthermore, we have t ≤ ǫ = 1, and then which implies a < 1. By using the fact kh(t, 1)k ≤ C(1 + t) ≤ 2C ≤ √2(1 + (1 + C)t) ≤ √2(2 + C), where C was the constant from (19), we can now estimate where we used a < 1 and λmax ≤ 1 to estimate the norm of h(0, a)−1. In addition, khk = kh(t, 1)−1h(0, a)k ≤ kh(t, 1)−1kkh(0, a)k ≤ C′aλmin . kh−1k = kh(0, a)−1h(t, 1)k ≤ kh(0, a)−1kkh(t, 1)k ≤ √2(2 + C)a−1, RECENT PROGRESS IN SHEARLET THEORY 23 Here we used that the set {h(t, 1) : t ≤ 1} ⊂ H is compact to uniformly estimate the norm of the inverses by a suitable C′, and a < 1 to estimate the norm of h(0, a). But these estimates combined yield kh−1k ≤ √2(2 + C)a−1 ≤ √2(2 + C)(C′)1/λminkhk−1/λmin . Since we assume that λmin > 0, the proof is finished. (cid:3) 5. Embeddings into the symplectic group From the analytical point of view, we saw that shearlet dilation groups are a useful tool for the characterization of the wavefront set of distributions. On the other hand, from the algebraic and geometrical points of view, these groups and the associated generalized wavelet representation exhibit an interesting link with the symplectic group and the metaplectic representation. More precisely, in this section we show that the positive part DS of any shearlet dilation group DS ∪ (−DS) may be imbedded into the symplectic group. Note that the full group DS ∪ (−DS) cannot be expected to be imbedded into Sp(d, R) [6, Theorem 3.5]. Moreover, we prove that the wavelet representation is unitarily equivalent to the metaplectic representation, provided that they are restricted to a suitable subspace of L2(Rd). In fact, a much more general class of groups is allowed, see Theorem 33. The relevance of the symplectic group and of the metaplectic representation in this context has already been shown in several works [15, 1, 2, 6]. In particular, the argument given here generalizes [6]. Let T (d, R)+ denote the subgroup of GL(d, R) consisting of the upper triangular matrices with positive entry in position (1, 1), namely T (d, R)+ = {h ∈ GL(d, R) : h1,1 > 0 and hi,j = 0 for every i > j}. We consider the following subspace of L2(Rd): H = {f ∈ L2(Rd) : supp f ⊆ ΘL}, where ΘL = {ξ ∈ Rd : ξ1 ≤ 0}. The main result of this section reads as follows. Theorem 33. Take H < T (d, R)+. The group G = Rd ⋊ H may be embedded into the symplectic group, namely there exists a group embedding φ : G → Sp(d, R). Moreover, the restriction to H of the quasi-regular representation π defined in (1) is unitarily equivalent to µ◦φ restricted to H, where µ is the metaplectic representation of Sp(d, R). The rest of this section is devoted to the proof of this theorem. The embedding φ, the subgroup φ(G), as well as the intertwining operator between the quasi-regular representation and the metaplectic representation will be explicitly constructed. First, we construct the subgroup φ(G) < Sp(d, R) and the map φ. The vectorial part of G = Rd ⋊ H will correspond to the subspace of the d-dimensional symmetric matrices given by We shall need the following preliminary result concerning the map (26) ρ : T (d, R)+ → GL(d, R), Σ := {σb := b1 b2/2 ... bd/2 b2/2 ··· 0 : b ∈ Rd}. bd/2  h 7→ph1,1 h−T . 24 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR Lemma 34. The map ρ is a group homomorphism and for all b ∈ Rd and h ∈ T (d, R)+ there holds (27) ρ(h)−T σbρ(h)−1 = σhb. Proof. The first part is trivial, since the matrices in H are upper triangular with h1,1 > 0. The second part can be proven as follows. Fix b ∈ Rd and h ∈ T (d, R)+. The assertion is equivalent to hσb hT = h1,1σhb. Write for i = 2, . . . , d h1 h2 ... hd ,   h1,1 h′ 1 h′ 0 2 ... ... h′ 0 d (cid:18) b1 b′/2 We have h = hσb = hσb hT = = whence h1,1b1 + h′ 2b′/2 h′ ... db′/2 h′ Therefore, since b′T h′T h1,1h′ 2b′/2 ... db′/2 h1,1h′ i = h′ h1 = [h1,1 h′ 1], hi = [0 h′ i], b′ = b2 ... bd  . h1,1b1 + h′ 2b′/2 h′ ... h′ db′/2 b′T /2 0 (cid:19) =  (cid:18)h1,1 h′T 1 ... 0 0 1b′/2 h1,1b′T /2 1b′/2 h1,1b′T /2 ,  0 h′T 2 ··· ··· 0 h′T d (cid:19) 0 0 ... 0  .  (cid:3) h1,1(h1,1b1 + h′ 1b′/2) + h1,1b′T h′T 1 /2 h1,1b′T h′T 2 /2 ··· h1,1b′T h′T d /2 ib′ for every i and h′ ib′ = hib for i ≥ 2, we obtain 2b′/2 ··· h′ db′/2 h′ 1b′ h1b1 h2b/2 h1,1b1 + h′ 2b′/2 h′ ... db′/2 h′ hσb hT = h1,1 = h1,1 σh h−T(cid:19) ∈ Sp(d, R), hdb/2 ... whence hσb hT = h1,1σhb, as desired. We use the notation g(σ, h) =(cid:18) h h2b/2 ··· 0 0 hdb/2  , σ ∈ Sym(d, R), h ∈ GL(d, R). RECENT PROGRESS IN SHEARLET THEORY 25 The product law is (28) g(σ1, h1)g(σ2, h2) = g(σ1 + h−T 1 σ2h−1 1 , h1h2). In the following result we show that G = Rd ⋊ H is isomorphic to the subgroup of Sp(d, R) given by Σ⋊ρ(H) := g(Σ, ρ(H)). This proves the first part of Theorem 33. Proposition 35. Take H < T (d, R)+. Then the map φ : Rd ⋊ H → g(Σ, ρ(H)) < Sp(d, R), is a group isomorphism. (b, h) 7→ g(σb, ρ(h)) It is worth mentioning that Lemma 2.3 in [6] immediately follows from this result. Proof. Recall that the product in Rd ⋊ H is defined by (b1, h1)(b2, h2) = (b1 + h1b2, h1h2), By definition of φ and using (28) there holds bi ∈ Rd, hi ∈ H. φ(b1, h1)φ(b2, h2) = g(σb1 , ρ(h1))g(σb2 , ρ(h2)) = g(σb1 + ρ(h1)−T σb2 ρ(h1)−1, ρ(h1)ρ(h2)). Therefore, Lemma 34 gives φ(b1, h1)φ(b2, h2) = g(σb1 + σh1b2 , ρ(h1h2)) = g(σb1+h1b2 , ρ(h1h2)) = φ((b1, h1)(b2, h2)), as desired. Note that the fact that g(Σ, ρ(H)) is a subgroup follows a posteriori. (cid:3) Intertwining the quasi-regular representation π, given in (1), with the Fourier transform F : H → L2(ΘL) we obtain the representation π(b, h) := F π(b, h)F −1 on L2(ΘL) given by f ∈ L2(ΘL). The metaplectic representation restricted to Σ ⋊ ρ(H) takes the form π(b, h) f (ξ) = det h1/2e−2πihb,ξi f (hT ξ), (29) µ(φ(b, h)) f (ξ) = det ρ(h)−1/2eπihσbξ,ξi f (ρ(h)−1ξ), f ∈ L2(ΘL). We now show that π and µ are unitarily equivalent, which concludes the proof of Theorem 33. The intertwining operator is given by Ψ : L2(ΘL) → L2(ΘL), Ψ f (ξ) = det JQ−1 (ξ)1/2 f (Q−1(ξ)), where Q : ΘL → ΘL is defined by Q(ξ) = − 1 Proposition 36. Let φ be the group isomorphism given by Proposition 35. For every (b, h) ∈ Rd ⋊ H there holds 2 ξ1ξ. Ψµ(φ(b, h))Ψ−1 = F π(b, h)F −1 = π(b, h). Proof. We start by giving a few identities without proof [6]: (30) (31) (32) (33) 2 −1ξ1− d 2 , d det JQ(ξ) = 21−dξ1d, det JQ−1(ξ) = 2 hσbξ, ξi = −2hb, Q(ξ)i, Q−1(ξ) = √2ξ/p−ξ1. 26 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR By (32) and (27) there holds −2hb, Q(hT ξ)i = hσb hT ξ, hT ξi = hhσb hT ξ, ξi = h1,1hσhbξ, ξi. Therefore, using again (32) we obtain whence (34) −2hb, Q(hT ξ)i = −2h1,1hhb, Q(ξ)i = −2hb, h1,1 hT Q(ξ)i, Q(hT ξ) = h1,1 hT Q(ξ). By using the definition of Ψ, (29), (32), (26) and once again the definition of Ψ, we can now compute for f ∈ L2(ΘL), b ∈ Rd and h ∈ H (35) Ψµ(φ(b, h))Ψ−1 f (ξ) = det JQ−1 (ξ)1/2(µ(φ(b, h))Ψ−1 f )(Q−1(ξ)) = det JQ−1 (ξ)1/2 det ρ(h)−1/2eπihσbQ−1(ξ),Q−1(ξ)iΨ−1 f (ρ(h)−1Q−1(ξ)) = det JQ−1 (ξ)1/2 det ρ(h)−1/2e−2πihb,ξiΨ−1 f (ρ(h)−1Q−1(ξ)) − 1 1,1 hT Q−1(ξ)) = det JQ−1 (ξ)1/2h = det JQ−1 (ξ)1/2h 1,1 det h1/2e−2πihb,ξiΨ−1 f (h 1,1 det h1/2e−2πihb,ξi · det JQ(h 1,1 hT Q−1(ξ))1/2 f (Q(h − d − d − 1 4 2 4 2 Now note that by (34) and by the fact that Q is quadratic there holds (36) Q(h 2 − 1 1,1 hT Q−1(ξ)) = h1,1 hT Q(h 2 − 1 1,1 Q−1(ξ)) = hT Q(Q−1(ξ)) = hT ξ. 2 − 1 1,1 hT Q−1(ξ))). Moreover we have det JQ(h − 1 2 1,1 hT Q−1(ξ)) = 21−d(h = 21−dh − 1 2 2 − d 1,1 hT Q−1(ξ))1d 1,1 ( hT Q−1(ξ))1d − d 1,1 hd 1,1Q−1(ξ)1d 2 ξ1 1,12 2 , d 2 d d 2 = 21−dh = 21−dh where the first equality follows from (30), the third one from the fact that hT is lower triangular and the forth one from (33). Therefore by (31) (37) det JQ−1 (ξ)1/2 det JQ(h 1,1 hT Q−1(ξ))1/2 = 2 2ξ1− d 4 2 4 ξ1 4 = h 1,12 1,1. 2 − d 4 − 1 2 h − 1 d 4 d 4 d d d 1 2 Finally, inserting (36) and (37) into (35) we obtain Ψµ(φ(b, h))Ψ−1 f (ξ) = det h1/2e−2πihb,ξi f ( hT ξ) = π(b, h) f (ξ), as desired. (cid:3) Acknowledgements G. S. Alberti was partially supported by the ERC Advanced Grant Project MULTIMOD-267184. S. Dahlke was supported by Deutsche Forschungsgemeinschaft (DFG), Grant DA 360/19 -- 1. He also acknowledges the support of the Hausdorff Research Institute for Mathematics during the Special Trimester "Mathematics of Signal Processing". F. De Mari and E. De Vito were partially supported by Progetto PRIN 2010-2011 "Variet`a reali e complesse: geome- tria, topologia e analisi armonica". They are members of the Gruppo Nazionale per l'Analisi Matematica, la Probabilit`a e le loro Applicazioni (GNAMPA) of the RECENT PROGRESS IN SHEARLET THEORY 27 Istituto Nazionale di Alta Matematica (INdAM). H. Fuhr acknowledges support from DFG through the grant Fu 402/5-1. Part of the work on this paper was carried out during visits of S. Dahlke and H. Fuhr to Genova, and they thank the Universit`a di Genova for its hospitality. References [1] Alberti, G.S., Balletti, L., De Mari, F., De Vito, E.: Reproducing subgroups of Sp(2, R). Part I: Algebraic classification. J. Fourier Anal. Appl. 19(4), 651 -- 682 (2013) [2] Alberti, G.S., De Mari, F., De Vito, E., Mantovani, L.: Reproducing subgroups of Sp(2, R) Part II: admissible vectors. Monatsh. Math. 173(3), 261 -- 307 (2014) [3] Cand`es, E.J., Donoho, D.L.: Ridgelets: A key to higher-dimensional intermittency? Phil. Trans. R. Soc. 357(1760), 2495 -- 2509 (1999) [4] Cand`es, E.J., Donoho, D.L.: New tight frames of curvelets and optimal representations of objects with piecewise C 2 singularities. Comm. Pure Appl. Math. 57(2), 219 -- 266 (2004) [5] Czaja, W., King, E.J.: Isotropic shearlet analogs for L2(Rk) and localization operators. Numer. Funct. Anal. Optim. 33(7-9), 872 -- 905 (2012) [6] Dahlke, S., De Mari, F., De Vito, E., Hauser, S., Steidl, G., Teschke, G.: Different faces of the shearlet group. J. Geom. Anal. , 1 -- 37 (2015), DOI:10.1007/s12220-015-9605-7 [7] Dahlke, S., Hauser, S., Steidl, G., Teschke, G.: Shearlet coorbit spaces: traces and embed- dings in higher dimensions. Monatsh. Math. 169(1), 15 -- 32 (2013) [8] Dahlke, S., Hauser, S., Teschke, G.: Coorbit space theory for the Toeplitz shearlet transform. Int. J. Wavelets Multiresolut. Inf. Process. 10(4), (2012) [9] Dahlke, S., Kutyniok, G., Maass, P., Sagiv, C., Stark, H.-G., Teschke, G.: The uncertainty principle associated with the continuous shearlet transform. Int. J. Wavelets Multiresolut. Inf. Process. 6(2) , 157 -- 181 (2008) [10] Dahlke, S., Kutyniok, G., Steidl, G., Teschke, G.: Shearlet coorbit spaces and associated banach frames. Appl. Comput. Harmon. Anal. 27(2), 195 -- 214 (2009) [11] Dahlke, S., Steidl, G., Teschke, G.: Multivariate shearlet transform, shearlet coorbit spaces and their structural properties. In: G. Kutyniok, D. Labate (eds.) Shearlets, pp. 105 -- 144. Birkhauser/Springer, New York (2012) [12] Dahlke, S., Steidl, G., Teschke, G.: Shearlet coorbit spaces: compactly supported analyzing shearlets, traces and embeddings. J. Fourier Anal. Appl. 17(6), 1232 -- 1255 (2011) [13] Dahlke, S., Steidl, G., Teschke, G.: The continuous shearlet transform in arbitrary space dimensions. J. Fourier Anal. Appl. 16(3), 340 -- 364 (2010) [14] Dahlke, S., Teschke, G.: The continuous shearlet transform in higher dimensions: Variations of a theme. In: Group Theory: Classes, Representations and Connections, and Applications, pp. 165-175. Nova Science Publishers, (2010) [15] De Mari, F., De Vito, E.: Admissible vectors for mock metaplectic representations. Appl. Comput. Harmon. Anal. 34(2), 163 -- 200 (2013) [16] Do, M.N., Vetterli, M.: Contourlets: a directional multiresolution image representation. In: Proceedings of the International Conference on Image Processing, vol. 1, pp.357 -- 360 (2002) [17] Feichtinger, H.G., Grochenig, K.: A unified approach to atomic decompositions via integrable group representations. In: Proc. Conf. Lund 1986 'Function spaces and applications', Lect. Notes in Math. 1302 (1988), pp 52 -- 73 [18] Feichtinger, H.G., Grochenig, K.: Banach spaces related to integrable group representations and their atomic decompositions. Part I. J. Funct. Anal. 86(2), 307 -- 340 (1989) [19] Feichtinger, H.G., Grochenig, K.: Banach spaces related to integrable group representations and their atomic decompositions. Part II. Monatsh. Math. 108(2-3), 129 -- 148 (1989) [20] Feichtinger, H.G., Grochenig, K.: Non-orthogonal wavelet and Gabor expansions and group representations. In: Ruskai, M.B., et al. (eds.) Wavelets and Their Applications, pp.353-376. Jones and Bartlett, Boston (1992) [21] Feichtinger, H.G., Sun, W., Zhou, X.: Two Banach spaces of atoms for stable wavelet frame expansions. J. Approx. Theory 146(1), 28 -- 70 (2007) [22] Fell, J., Fuhr, H., Voigtlaender, F.: Resolution of the wavefront set using general continuous wavelet transforms. J. Fourier Anal. Appl., 1 -- 62 (2015) [23] Folland, G.B.: A course in abstract harmonic analysis. CRC Press, Boca Raton, FL(1995) 28 G. S. ALBERTI, S. DAHLKE, F. DE MARI, E. DE VITO, AND H. F UHR [24] Frazier, M., Jawerth, B.: Decomposition of Besov spaces. Indiana Univ. Math. J. 34(4), 777 -- 799 (1985) [25] Frazier, M., Jawerth, B., Weiss, G.: Littlewood-Paley theory and the study of function spaces. (1991) [26] Fuhr, H.: Continuous wavelet transforms with abelian dilation groups. J. Math. Phys. 39(8), 3974 -- 3986 (1998) [27] Fuhr, H.: Coorbit spaces and wavelet coefficient decay over general dilation groups. Trans. Amer. Math. Soc. 367(10), 7373 -- 7401 (2015) [28] Fuhr, H.: Generalized Calder´on conditions and regular orbit spaces. Colloq. Math. 120(1), 103 -- 126 (2010) [29] Fuhr, H.: Vanishing moment conditions for wavelet atoms in higher dimensions. Adv. Com- put. Math. 42(1), 127-153 (2016) [30] Fuhr, H., Raisi-Tousi, R.: Simplified vanishing moment criteria for wavelets over general dilation groups, with applications to abelian and shearlet dilation groups. To appear in Appl. Comp. Harm. Anal. DOI:10.1016/j.acha.2016.03.003 Preprint available under http://arxiv.org/abs/1407.0824 (2014) [31] Grochenig, K.: Describing functions: atomic decompositions versus frames. Monatsh. Math. 112(1),1 -- 42 (1991) [32] Grohs, P.: Shearlet and microlocal analysis. In: Kutyniok, G., Labate, D. (eds.) Shearlets: Multiscale Analysis for Multivariate Data, pp. 39-67. Birkhauser, Boston (2012) [33] Hilgert, J., Neeb, K.-H.: Structure and geometry of Lie groups. Springer Monographs in Mathematics. Springer, New York (2012) [34] Jaffard, S., Meyer, Y.: Wavelet Methods for Pointwise Regularity and Local Oscillations of Functions, Mem. Amer. Math. Soc. 123(587), (1996) [35] King, E.J.: Wavelet and frame theory: frame bound gaps, generalized shearlets, Grassman- nian fusion frames, and p-adic wavelets. Dissertation, University of Maryland, College Park (2009) [36] Kutyniok, G., Labate, D.: Resolution of the wavefront set using continuous shearlets. Trans. Amer. Math. Soc. 361(5), 2719 -- 2754 (2009) [37] Kutyniok, G., Labate, D. (eds.): Shearlets: Multiscale Analysis for Multivariate Data. Birkhauser/Springer, New York (2012) [38] Triebel, H.: Characterizations of Besov-Hardy-Sobolev spaces: a unified approach. J. Ap- prox. Theory 52(2), 162 -- 203 (1988) [39] Varadarajan, V. S.: Lie groups, Lie algebras, and their representations. Springer-Verlag, New York (1984) Department of Mathematics, ETH Zurich, Ramistrasse 101, 8092 Zurich, Switzerland E-mail address: [email protected] FB12 Mathematik und Informatik, Philipps-Universitat Marburg, Hans-Meerwein- Strasse, Lahnberge, 35032 Marburg, Germany E-mail address: [email protected] Dipartimento di Matematica, Universit`a di Genova, Via Dodecaneso 35, Genova, Italy E-mail address: [email protected] Dipartimento di Matematica, Universit`a di Genova, Via Dodecaneso 35, Genova, Italy E-mail address: [email protected] Lehrstuhl A fur Mathematik, RWTH Aachen University, 52056 Aachen, Germany E-mail address: [email protected]
1811.09183
1
1811
2018-11-22T14:13:13
On the Maurey--Pisier and Dvoretzky--Rogers theorems
[ "math.FA" ]
A famous theorem due to Maurey and Pisier asserts that for an infinite dimensional Banach space $E$, the infumum of the $q$ such that the identity map $id_{E}$ is absolutely $\left( q,1\right) $-summing is precisely $\cot E$. In the same direction, the Dvoretzky--Rogers Theorem asserts $id_{E}$ fails to be absolutely $\left( p,p\right) $-summing, for all $p\geq1$. In this note, among other results, we unify both theorems by charactering the parameters $q$ and $p$ for which the identity map is absolutely $\left( q,p\right)$-summing. We also provide a result that we call \textit{strings of coincidences} that characterize a family of coincidences between classes of summing operators. We illustrate the usefulness of this result by extending classical result of Diestel, Jarchow and Tonge and the coincidence result of Kwapie\'{n}.
math.FA
math
ON THE MAUREY -- PISIER AND DVORETZKY -- ROGERS THEOREMS G. ARA ´UJO AND J. SANTOS Abstract. A famous theorem due to Maurey and Pisier asserts that for an infinite dimensional Banach space E, the infumum of the q such that the identity map idE is absolutely (q, 1)- summing is precisely cot E. In the same direction, the Dvoretzky -- Rogers Theorem asserts idE fails to be absolutely (p, p)-summing, for all p ≥ 1. In this note, among other results, we unify both theorems by charactering the parameters q and p for which the identity map is absolutely (q, p)-summing. We also provide a result that we call strings of coincidences that characterize a family of coincidences between classes of summing operators. We illustrate the usefulness of this result by extending classical result of Diestel, Jarchow and Tonge and the coincidence result of Kwapie´n. 1. Introduction and background Let 2 ≤ q < ∞. A Banach space E has cotype q (see [5, page 218]) if there is a constant C > 0 such that, no matter how we select finitely many vectors x1, . . . , xn ∈ E, (1.1) 1 q kxkkq! n Xk=1 ≤ C Z[0,1](cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n Xk=1 rk(t)xk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 1/2 dt  , where rk denotes the k-th Rademacher function. In this context we also define cot E := inf{q : E has cotype q}. j=1 ∈ ℓq(F ) whenever (xj)∞ j=1 ∈ ℓw Recall that if 1 ≤ p ≤ q ≤ ∞, for Banach spaces E, F , a linear operator u : E → F is absolutely (q, p)-summing if (u(xj))∞ p (E) is the linear space of the sequences (xj)∞ j=1 ∈ ℓp for every continuous linear functional ϕ : E → K. The class of all absolutely (q, p)-summing operators from E to F will be denoted by Π(q,p)(E; F ). We denote Π(p,p)(E; F ) by Πp(E; F ). From now on, for any p > 1 the symbol p∗ denotes the conjugate of p, i.e., p∗ = p/ (p − 1) . Cotype and absolutely summing operators are closely related by the famous Maurey -- Pisier Theorem (see also [4] for further relations between cotype and absolutely summing operators): j=1 in E such that (ϕ (xj))∞ p (E); we recall that ℓw Theorem 1.1 (Maurey -- Pisier). For every infinite dimensional Banach space E, we have cot E = inf{a : idE is absolutely (a, 1)-summing}. In the same direction, the Dvoretzky -- Rogers Theorem [6] tells us that idE is not absolutely (p, p)-summing, regardless of the p ≥ 1. In a more general version, as stated in [5, Theorem 10.5], it reads as follows: Theorem 1.2 (Dvoretzky -- Rogers). If E is an infinite dimensional Banach space, the identity map idE is not absolutely (q, p)-summing whenever 1 p − 1 q < 1 2. Our first main result revisits and unifies both theorems. Our second main result provides a family of coincidences for the classes of absolutely summing operators that encompasses the following classical results: J. Santos was supported by Conselho Nacional de Desenvolvimento Cient´ıfico e Tecnol´ogico -- CNPq. 2010 Mathematics Subject Classification: 46A32, 47H60. Key words: Absolutely summing operators, Maurey -- Pisier theorem, Dvoretzky -- Rogers theorem. 1 2 ARA ´UJO AND SANTOS Theorem 1.3 (Diestel -- Jarchow -- Tonge). Let E and F be Banach spaces (a) If E has cotype 2, then Π2 (E; F ) = Π1 (E; F ). (b) If E has cotype 2 < q < ∞, then Πr (E; F ) = Π1 (E; F ) for all 1 < r < q∗. Theorem 1.4 ([7], Kwapie´n). Let 1 ≤ p ≤ ∞. Then Π(r(p),1) (ℓ1; ℓp) = L (ℓ1; ℓp) where Moreover, if r < r(p), then Π(r,1) (ℓ1; ℓp) 6= L (ℓ1; ℓp). 1 r(p) − 1 p = 1 −(cid:12)(cid:12)(cid:12)(cid:12) . 1 2(cid:12)(cid:12)(cid:12)(cid:12) 2. Main results The next definition was introduced by M.C. Matos in [8] and, independently, by D. P´erez- Garc´ıa in [9]. Definition 2.1. Let 1 ≤ p ≤ q ≤ ∞. A multilinear operator T : E1 × · · · × Em → F is multiple (q, p)-summing if there exist a constant C > 0 such that ∞ Xj1,...,jm=1(cid:13)(cid:13)(cid:13) j )∞ j=1 ∈ ℓw for all (x(k) summing operators by Πm Πm (q,p) (mE; F ). T (x(1) j1 , . . . , x(m) 1 q ≤ C q! jm )(cid:13)(cid:13)(cid:13) m Yk=1(cid:13)(cid:13)(cid:13) (x(k) j )∞ j=1(cid:13)(cid:13)(cid:13)w,p p (Ek), with k = 1, ..., m. We represent the class of all multiple (q, p)- (q,p) (E1, . . . , Em; F ). When E1 = · · · = Em = E, we denote simply by (zj)∞ Following [1], when we write (xj)∞ j=1 ∈ ℓw The following result of Arregui and Blasco [1, Lemma 3 and Proposition 6] is crucial for us: s (E) such that xj = ajzj for all j. s (E), it means that there are (aj)∞ j=1 ∈ ℓr and j=1 ∈ ℓrℓω Proposition 2.2. Let 1 < r < ∞. Then ℓω In particular, 1 (E) = ℓrℓω r∗(E) if and only if L (c0; E) = Πr (c0; E). (a) If E has cotype 2, then ℓω (b) If E has cotype q > 2, then ℓω 1 (E) = ℓ2ℓω 2 (E). 1 (E) = ℓrℓω r∗(E) for any r > q. ℓrℓω Note that from the previous result it is immediate that if E has cotype 2 then ℓω r∗(E) for any r ≥ 2. Let us to present now an inclusion theorem for multiple summing multilinear operators that, when restricted to the linear case, will be important in the proof of our main result. Before that, let us to recall the classical inclusion theorem for absolutely summing operators [5, Theorem 10.4], which will be very useful to our propose. 1 (E) = Theorem 2.3. If 1 ≤ aj ≤ bj for j = 1, 2 and 1 b1 (a1, b1)-summing operator is also absolutely (a2, b2)-summing. − 1 a1 ≤ 1 b2 − 1 a2 , then every absolutely Lemma 2.4. Let E1, ..., Em and F be infinite dimensional Banach spaces. (a) If E1, ..., Em has cotype 2 and r ≤ 2, then Π(a,r) (E1, ..., Em; F ) ⊂ Π(s,1) (E1, ..., Em; F ) with 1 s (b) If Ei has cotype qi > 2, q = max{qi : i = 1, ..., m} and 1 < r < q∗, then 1 s Π(a,r) (E1, ..., Em; F ) ⊂ Π(s,1) (E1, ..., Em; F ) with ≤ 1 − ≤ 1 − 1 a 1 a 1 r 1 r − − . . Proof. (a) If r ≤ 2, then r∗ ≥ 2 and, from Theorem 2.3, Π2(E; F ) ⊆ Πr∗(E; F ) for all Banach spaces E and F . Thus, since Ei has cotype 2, it follows that L(c0; Ei) = Π2(c0; Ei) ⊆ Πr∗(c0; Ei), ON THE MAUREY -- PISIER AND DVORETZKY -- ROGERS THEOREMS 3 that is, L(c0; Ei) = Πr∗(c0; Ei). From Proposition 2.2 we thus have ℓw if (x(i) j )∞ j=1 ∈ ℓw 1 (Ei), 1 (Ei) = ℓr∗ℓw r (Ei). Then, x(i) j = a(i) j z(i) j with (a(i) j )j ∈ ℓr∗ and (z(i) j )j ∈ ℓw r (Ei). Since from Holder's inequality we have 1 s ≤ 1 r∗ + 1 a , ∞ Xj1,...,jm=1(cid:13)(cid:13)(cid:13) T (cid:16)x(1) = ∞ Xj1,...,jm=1(cid:12)(cid:12)(cid:12) ≤ ∞ Xj1,...,jm=1(cid:12)(cid:12)(cid:12) < ∞. j1 ...a(m) a(1) a(1) j1 ...a(m) j1 , ..., z(m) 1 a a! jm (cid:17)(cid:13)(cid:13)(cid:13) j1 , ..., x(m) 1 s 1 s j1 , ..., z(m) s! jm (cid:17)(cid:13)(cid:13)(cid:13) s! s(cid:13)(cid:13)(cid:13) jm (cid:17)(cid:13)(cid:13)(cid:13) jm (cid:12)(cid:12)(cid:12) T (cid:16)z(1) r∗! r∗ ∞ jm (cid:12)(cid:12)(cid:12) Xj1,...,jm=1(cid:13)(cid:13)(cid:13) T (cid:16)z(1) 1 (b) Follows as in (a) using Proposition 2.2(b). (cid:3) Remark 2.5. This theorem is a generalization of [10, Theorem 10]. Theorem 2.6. Let a, b ∈ [1, ∞) and E be an infinite dimensional Banach space. (i) If b ≥ (cot E)∗, then inf {a : idE is absolutely (a, b) -summing} = ∞. (ii) (ii) If b < (cot E)∗, then inf {a : idE is absolutely (a, b) -summing} = b cot E b + cot E − b cot E . Proof. (i) Let us to prove that idE fails to be (a, (cot E)∗)-summing; this result seems to have been overlooked in the literature; it appears, in a more general form, in the preprint [2] and we sketch the proof below. A famous result of Maurey and Pisier asserts that ℓcot E is finitely representable in E, i.e., for all n ≥ 1, there exists En ⊂ E and an isomorphism Sn : ℓn cot E → En such that kSnk · kS −1 n k ≤ 2. There is no loss of generality in assuming kSnk ≤ 1. For 1 ≤ i ≤ n, let yi = Sn(ei). Note that 1 = keikℓn cot E =(cid:13)(cid:13)S −1 n (yi)(cid:13)(cid:13)ℓn cot E ≤(cid:13)(cid:13)S −1 n (cid:13)(cid:13) kyik ≤ 2 kyik , 4 ARA ´UJO AND SANTOS and thus kyik ≥ 1/2. Moreover sup ϕ∈BE ∗ n Xi=1 hϕ, yii(cot E)∗! 1 (cot E)∗ n = sup ϕ∈BE ∗ sup α∈Bℓn cot E αihϕ, yii αiyii n hϕ, Xi=1 Xi=1 αiyi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)E αiSn(ei)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)E = sup α∈Bℓn cot E sup ϕ∈BE ∗ = sup α∈Bℓn = sup α∈Bℓn = sup α∈Bℓn n n cot E(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 cot E(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 cot E kSn (α)kE We thus conclude that for any a ≥ 1 we have ≤ kSnk sup α∈Bℓn cot E kαkℓn cot E ≤ 1. n 2a ≤ Xj=1 n n kidE(yj)ka and sup ϕ∈BE ∗ Xi=1 hϕ, yii(cot E)∗ ≤ 1 and this means that idE is not absolutely (a, (cot E)∗)-summing. The proof of (i) is done. Now let us prove (ii). Consider If r < λ, then So λ = b cot E b + cot E − b cot E . 1 r > 1 b − 1 + 1 cot E . cot E − rb rb + b − r > 0. Thus, there is 0 < ε < cot E − rb rb+b−r such that 1 b − 1 r < 1 − 1 cot E − ε . From Lemma 2.4, with m = 1 and s = cot E − ε, we conclude that Π(r,b) (E; E) ⊂ Π(cot E−ε,1) (E; E) 6= L (E; E) and thus idE fails to be (r, b)-summing. So λ ≤ inf {a : idE is absolutely (a, b) -summing} . Now, if r > λ then 1 r Therefore, there exist ε > 0 such that < 1 b − 1 + 1 cot E . Thus r > cot E + ε and 1 r ≤ 1 b − 1 + 1 cot E + ε . 1 − 1 cot E + ε ≤ 1 b − 1 r . ON THE MAUREY -- PISIER AND DVORETZKY -- ROGERS THEOREMS 5 The inclusion theorem (Theorem 2.3) asserts that every absolutely (cot E + ε, 1)-summing oper- ator is also absolutely (r, b)-summing. Since idE is absolutely (cot E + ε, 1)-summing, it follows that idE is also absolutely (r, b)-summing. Consequently, λ = inf {a : idE is absolutely (a, b) -summing} , and the proof is done. (cid:3) a cot E 2 1 1 1 (cot E)∗ b − 1 a = 1 (cot E)∗ 2 b idE is not (a, b)-summing idE is (a, b)-summing Figure 1. Graphical overview of Theorem 2.6. Remark 2.7. Note that this result completes the information of the Dvoretzky -- Rogers Theorem (Theorem 1.2) and, when b = 1, we recover the classical result of Maurey and Pisier (Theorem 1.1). In the previous theorem we prove that for b < (cot E)∗ we can fully characterize the classes of (a, b)-summing operators that coincide. The Figure 1 illustrates this result and motivates us to name the next result as "strings of coincidences", which generalizes Theorem 1.3. Theorem 2.8 (Strings of coincidences). Let E and F be infinite dimensional Banach spaces. (a) If E has cotype 2 and a ≤ a1 and r ≤ r1 ≤ 2, then 1 r Π(a,r) (E; F ) = Π(a1,r1) (E; F ) with − (b) If a ≤ a1 and 1 ≤ r ≤ r1 < (cot E)∗, then Π(a,r) (E; F ) = Π(a1,r1) (E; F ) with 1 r − 1 a 1 a = = 1 r1 1 r1 − 1 a1 . − 1 a1 . Proof. Note that (2.1) with Π(a,r) (E; F ) = Π(s,1) (E; F ) = Π(a1,r1) (E; F ) , 1 − 1 s = 1 r − 1 a = 1 r1 − 1 a1 . In fact, the two equalities in (2.1) are immediate consequences of the previous theorem and the inclusion theorem for absolutely summing operators (Theorem 2.3). (cid:3) 6 ARA ´UJO AND SANTOS a 2 1 1 r 2 Figure 2. Strings of coincidence for (a, r)-summability when cot E = 2. Remark 2.9. The strings of coincidence can also be obtained as a consequence of [3, Theorem 2.1]. We finish this section by illustrating how the strings of coincidence can be useful to extend the classical result of Kwapie´n (Theorem 1.4). Theorem 2.10. Let 1 ≤ p ≤ ∞ and a, b ≥ 1. If b ≤ 2, then Π(a,b) (ℓ1; ℓp) = L (ℓ1; ℓp) if, and only if, a ≥ br(p) r(p)+b−br(p) where 1 1 p − 1 r(p) = 1 −(cid:12)(cid:12)(cid:12) . 2(cid:12)(cid:12)(cid:12) Proof. For b ≤ 2 the string of coincidence associated to (r(p), 1), where composed by the pairs (a, b) such that 1 r(p) = 1 −(cid:12)(cid:12)(cid:12) 1 p − 1 , is 2(cid:12)(cid:12)(cid:12) Thus 1 b − 1 a = 1 − 1 r(p) . a = br(p) r(p) + b − br(p) . In view of the optimality of of a is sharp. br(p) r(p)+b−br(p) when b = 1, we can conclude that the above estimate (cid:3) References [1] J.L. Arregui and O. Blasco, (p, q)-summing sequences, J. Math. Anal. Appl. 274 (2002), 812 -- 827. [2] F. Bayart, D. Pellegrino and P. Rueda, On coincidence results for summing multilinear operators: interpo- lation, ℓ1-spaces and cotype, arXiv:1805.12500v1 (2018). [3] A.T. Bernardino, On cotype and a Grothendieck-type Theorem for absolutely summing multilinear opera- tors, Quaestiones Math. 34 (2011), 1 -- 7. [4] G. Botelho, D. Pellegrino and P. Rueda, Cotype and absolutely summing linear operators, Math. Z., 267 (2001), 1 -- 7. [5] J. Diestel, H. Jarchow and A. Tonge, Absolutely summing operators, Cambridge University Press, 1995. [6] A. Dvoretzky and C.A. Rogers, Absolute and unconditional convergence in normed spaces, Proc. Nat. Acad. Sci. USA 36 (1950), 192 -- 197. [7] S. Kwapie´n, Some remarks on (p, q)-absolutely summing operators in ℓp-space, Studia Math. 29 (1968), 327 -- 337. [8] M.C. Matos, Fully absolutely summing and Hilbert-Schmidt multilinear mapping, Collect. Math. 54 (2003), 111 -- 136. [9] D. P´erez-Garc´ıa, Operadores multilineales absolutamente sumantes, PhD Thesis, Universidad Complutense de Madrid, 2003. [10] D. Popa, Reverse inclusions for multiple summing operators, J. Math. Anal. Appl. 350 (2009), 360 -- 368. ON THE MAUREY -- PISIER AND DVORETZKY -- ROGERS THEOREMS 7 (G. Ara´ujo) Departamento de Matem´atica, Universidade Estadual da Para´ıba, 58.429-600 - Campina Grande, Brazil E-mail address: [email protected] (J. Santos) Departamento de Matem´atica, Universidade Federal da Para´ıba, 58.051-900 - Joao Pessoa, Brazil E-mail address: [email protected] or [email protected]
1907.05241
1
1907
2019-07-10T13:59:28
Metrization of probabilistic metric spaces. Applications to fixed point theory and Arzela-Ascoli type theorem
[ "math.FA" ]
Schweizer, Sklar and Thorp proved in 1960 that a Menger space $(G,D,T)$ under a continuous $t$-norm $T$, induce a natural topology $\tau$ wich is metrizable. We extend this result to any probabilistic metric space $(G,D,\star)$ provided that the triangle function $\star$ is continuous. We prove in this case, that the topological space $(G,\tau)$ is uniformly homeomorphic to a (deterministic) metric space $(G,\sigma_D)$ for some canonical metric $\sigma_D$ on $G$. As applications, we extend the fixed point theorem of Hicks to probabilistic metric spaces which are not necessarily Menger spaces and we prove a probabilistic Arzela-Ascoli type theorem.
math.FA
math
METRIZATION OF PROBABILISTIC METRIC SPACES. APPLICATIONS TO FIXED POINT THEORY AND ARZELA-ASCOLI TYPE THEOREM MOHAMMED BACHIR, BRUNO NAZARET Abstract. Schweizer, Sklar and Thorp proved in 1960 that a Menger space (G, D, T ) under a continuous t-norm T , induce a natural topology τ wich is metrizable. We extend this result to any probabilistic metric space (G, D, ⋆) provided that the triangle function ⋆ is continuous. We prove in this case, that the topological space (G, τ ) is uniformly homeomorphic to a (deterministic) metric space (G, σD) for some canonical metric σD on G. As applications, we extend the fixed point theorem of Hicks to probabilistic metric spaces which are not necessarily Menger spaces and we prove a probabilistic Arzela-Ascoli type theorem. Keywords: Metrization of probabilistic metric space; Probabilistic 1-Lipschitz map; Probabilistic Arzela-Ascoli type Theorem; Probabilistic fixed point theorem. msc: 54E70, 46S50. 1. Introduction Let (G, D, T ) be a Menger space equipped with a probabilistic metric D and a t-norm T (the definitions and notation reminders will be given in the details in Section 2). Schweizer and Sklsar [13] defined for ε, λ > 0 and each x ∈ G a neighborhood Nx(ε, λ) as follows Nx(ε, λ) = {y ∈ G : D(x, y)(ε) > 1 − λ}. Schweizer, Sklar and Thorp proved in [14] that, given a t-norm T of a Menger space (G, D, T ) satisfying 1 = supx<1 T (x, x) (in particular if T is continuous), the collection {Nx(ε, λ) : x ∈ G} taken as a neighborhood base at x gives rise to a metrizable topology. In [11] Morrel and Nagata proved the following two extensions: (1) The class of topological Menger spaces coincides with that of semi-metrizable topological spaces. (2) No condition on T weaker than 1 = supx<1 T (x, x) can guarantee that a Menger space, under T , is topological. The aim of the present paper is to prove that, in a general probabilistic metric space (G, D, ⋆), not necessarily being a Menger space, the collection {Nx(ε, λ) : x ∈ X} taken as a neighborhood base at x gives rise to a topology which is uniformly homeomorphic to a metric space, provided that the triangle function ⋆ is continuous (necessarily uniformly continuous by Sibley's result in [16] on the compactness of Date: 05/06/2019. 1991 Mathematics Subject Classification. 54E70, 46S50. 1 2 MOHAMMED BACHIR, BRUNO NAZARET (∆+, dL), where dL denotes the modified L´evy distance and ∆+ denotes the set of all nondecreasing and left-continuous distributions that vanish at 0). We get an even more precise result : if w⋆ : [0, +∞] → [0, +∞] is a modulus of uniform continuity for the triangle function ⋆, then the (deterministic) metric σD on G defined canonically from the probabilistic metric D by ∀x, y ∈ G, σD(x, y) := sup z∈G dL(D(x, z), D(z, y)), satisfies the following inequalities ∀x, y ∈ G, dL(D(x, y), H0) ≤ σD(x, y) ≤ w⋆(dL(D(x, y), H0)). (1) Note from [12] that y ∈ Nx(t, t) if and only if dL(D(x, y), H0) < t, for all t > 0. If moreover we assume that ⋆ is k-Lipschitz given some positive real number k (necessarily k ≥ 1), then we can take w⋆(t) = kt for all t ≥ 0 (see Proposition 4 for examples of such functions). As an immediate consequence of (1), the semi-metric α(x, y) := dL(D(x, y), H0) define a topology on (G, D, ⋆) which is uniformly (resp. Lipschitz) homeomorphic to the metric space (G, σD), whenever ⋆ is continuous (resp. Lipschitz continuous). This result is an extension to non necessarily Menger spaces of the works established in Menger spaces by Schweizer, Sklar and Thorp in [14]. In particular, the formula (1) allows us to transfer several known results from metric space theory to the probabilistic metric theory. For instance, using (1) and the Ekeland variational principle we give some extensions of the fixed point theorem of Hicks (see [4]), or using again (1) we give an Arzela-Ascoli type theorem for the the space of probabilistic 1-Lipschtz maps introduced recently in [1]. Notice that other results such as Baire theorem and all its variants/consequences can be transfered, thanks to our result, to the probabilistic metric framework. This paper is organized as follows. In Section 2, we recall some classical notions related to probabilistic metric space. In Section 3, we treat the metrization of prob- abilistic metric space and prove Theorem 1. We also give some new properties. In Section 4, we establish fixed point theorems (Theorem 2 and Theorem 3) extending a result of Hicks (see [4]). In Section 5, we prove Theorem 5, showing that the set of probabilistic 1-Lipschitz maps introduced in [1] is a compact space for the uniform convergence, giving a probabilistic Arzela-Ascoli theorem. 2. Definitions and notation In this section, we recall some known facts about probabilistic metric spaces, the modified L´evy distance and the weak convergence. All these notions can be found in [12], [5] and [6]. We also recall the notion of probabilistic 1-Lipschitz map introduced in [1], which shall play an important role in the sequel. 2.1. Probabilistic metric space and triangle function. By ∆+ we denote the set of all (cumulative) distribution functions F : [−∞, +∞] −→ [0, 1], nonde- creasing and left-continuous with F (−∞) = 0; F (+∞) = 1 and F (0) = 0. For a ∈ [0, +∞[, we denote Ha(t) = 0 if t ≤ a and Ha(t) = 1, if t > a. In the sequel, we shall write F ≤ G for which defines an ordering relation on ∆+. ∀t ∈ R, F (t) ≤ G(t), METRIZATION OF PROBABILISTIC METRIC SPACES 3 Definition 1. ([12, 4, 5, 6]) A binary operation ⋆ on ∆+ is called a triangle function if and only if it is commutative, associative, non-decreasing in each place, and has H0 as neutral element. In other words: (i) F ⋆ L ∈ ∆+ for all F, L ∈ ∆+. (ii) F ⋆ L = L ⋆ F for all F, L ∈ ∆+. (iii) F ⋆ (L ⋆ K) = (F ⋆ L) ⋆ K, for all F, L, K ∈ ∆+. (iv) F ⋆ H0 = F for all F ∈ ∆+. (v) F ≤ L =⇒ F ⋆ K ≤ L ⋆ K for all F, L, K ∈ ∆+. Definition 2. A t-norm is a function T : [0, 1] × [0, 1] → [0, 1], usually called a triangular norm (see [12, 4, 5, 6]), satisfying • T (x, y) = T (y, x) ( commutativity); • T (x, T (y, z)) = T (T (x, y), z) (associativity); • T (x, y) ≤ T (x, z) whenever y ≤ z (monotonicity ); • T (x, 1) = x (boundary condition). Definition 3. A probabilistic metric space (G, D, ⋆) (an PM-space) is a set G together with a triangle function ⋆ and a function D : G × G → ∆+ satisfying: (i) D(x, y) = H0 iff x = y. (ii) D(x, y) = D(y, x) for all x, y ∈ G (iii) D(x, y) ⋆ D(y, z) ≤ D(x, z) for all x, y, z ∈ G Usually, D(x, y) is denoted by Fx,y in the literature. A probabilistic metric space (G, D, ⋆) is called a Menger space and denoted by (G, D, T ), iff the triangle function ⋆ := ⋆T is defined from a t-norm T as follows: for all F, L ∈ ∆+ and for all t ∈ R, (F ⋆T L)(t) := sup u+v=t T (F (u), L(v)) (2) = sup T (F (u), L(v)) u,v≤0:u+v=t 2.2. L´evy distance and weak convergence. Definition 4. Let F and G be in ∆+. For any h > 0 we set Ah F,G = {t ≥ 0 st. G(t) ≤ F (t + h) + h} . The modified L´evy distance is the map dL defined on ∆+ × ∆+ as G,F(cid:9) . dL(F, G) = inf(cid:8)h > 0 st. [0, h−1[⊂ Ah F,G ∩ Ah Notice that, for all F , G ∈ ∆+, (i) if F ≤ G then AG,F = [0, +∞[, hence dL(F, G) = inf (cid:8)h > 0 st. [0, h−1[⊂ Ah F,G = Ah F,G(cid:9) . (ii) if h ≥ 1, Ah (iii) The usual Levy distance between general cumulative distribution functions G,F = [0, +∞[, hence dL(F, G) ≤ 1. can be expressed as inf (cid:8)h > 0 st. Ah F,G = Ah It is invariant under the action of translations which, as we shall see later, is not the case for the modified version since it somehow does not see the behaviour at infinity. G,F = [0, +∞[(cid:9) . 4 MOHAMMED BACHIR, BRUNO NAZARET Definition 5. Let ⋆ be a triangle function on ∆+. (1) A sequence (Fn) of distributions in ∆+ converges weakly to a function F in ∆+ if (Fn(t)) converges to F (t) at each point t of continuity of F . In this case, we write indifferently Fn w−→ F or limn Fn = F . (2) We say that the law ⋆ is continuous at (F, L) ∈ ∆+ × ∆+ if we have Fn ⋆ w−→ F ⋆ L, whenever Fn w−→ F and Ln w−→ L. We recall the following results due to D. Sibley in [16, Theorem 1. and Theorem Ln 2]. Lemma 1. ([16, 12]) The function dL is a metric on ∆+ and (∆+, dL) is compact. Lemma 2. ([16, 12]) Let (Fn) be a sequence of functions in ∆+, and let F be an element of ∆+. Then (Fn) converges weakly to F if and only if dL(Fn, F ) −→ 0, when n −→ +∞. Remark 1. Thanks to Lemma 2, we shall indifferently use the notations Fn or dL(Fn, F ) −→ 0 to say that (Fn) converges weakly to F . w−→ F Definition 6. Let (G, D, ⋆) be a probabilistic metric space. For x ∈ G and t > 0, the strong t-neighborhood of x is the set and the strong neighborhood system for G is {Nx(t); x ∈ G, t > 0}. Nx(t) = {y ∈ G : D(x, y)(t) > 1 − t}, Lemma 3. ([12, Lemme 4.3.3]) Let t > 0 and x, y ∈ G. Then we have y ∈ Nx(t) if and only if dL(D(x, y), H0) < t. 2.3. Probabilistic 1-Lipschitz map. Definition 7. Let (G, D, ⋆) be a probabilistic metric space and let f be a function f : (G, D, ⋆) −→ (∆+, dL). We say that f is a probabilistic 1-Lipschitz map if : ∀x, y ∈ G, D(x, y) ⋆ f (y) ≤ f (x). We can also define probabilistic k-Lipschitz maps for any nonegative real number k ≥ 0 as the maps f satisfying ∀x, y ∈ G, Dk(x, y) ⋆ f (y) ≤ f (x), where, for all x, y ∈ G and all t ∈ R, Dk(x, y)(t) = D(x, y)( t k ) if k > 0 and D0(x, y)(t) = H0(t) if k = 0. For sake of simplicity, when we use the notion in Definition 7, we shall only treat in this paper the case of probabilistic 1-Lipschitz maps, but our main result result could be easily extended to this more general setting. Examples 1. Let (G, d) be a metric space. Assume that ⋆ is a triangle function on ∆+ satisfying Ha ⋆ Hb = Ha+b for all a, b ∈ R+ (for example if ⋆ = ⋆T where T is a lef-continuous triangular norm). Let (G, D, ⋆) be the probabilistic metric space defined with the probabilistic metric Let L : (G, d) −→ R+ be a real-valued map. Then, L is a non-negative 1-Lipschitz map if and only if f : (G, D, ⋆) −→ ∆+ defined for all x ∈ G by D(p, q) = Hd(p,q). f (x) := HL(x) METRIZATION OF PROBABILISTIC METRIC SPACES 5 is a probabilistic 1-Lipschitz map. This example shows that the framework of probabilistic 1-Lipschitz maps encompasses the classical determinist case. By Lip1 ⋆(G, ∆+) we denote the space of all probabilistic 1-Lipschitz maps Lip1 ⋆(G, ∆+) := {f : G −→ ∆+/D(x, y) ⋆ f (y) ≤ f (x); ∀x, y ∈ G}. For all x ∈ G, by δx we denote the map δx : G −→ ∆+ y 7→ D(y, x). It follows from the properties of the probabilistic metric D that δx is a prob- abilistic 1-Lipschitz for every x ∈ G. We set G(G) := {δx, x ∈ G} and by δ, we denote the operator δ : G −→ G(G) ⊂ Lip1 ⋆(G, ∆+) x 7→ δx. 2.4. Modulus of uniform continuity of a triangle function on ∆+. Let ⋆ : ∆+ × ∆+ → ∆+ be a continuous triangle function (with respect to the modified L´evy distance dL). Since (∆+, dL) is a compact metric space (see Lemma 1) and ⋆ is continuous, then ⋆ is uniformly continuous from ∆+ × ∆+ into ∆+. Let ω⋆ : [0, +∞] → [0, +∞] be a modulus of uniform continuity for ⋆ (limt→0 ω⋆(t) = ω⋆(0) = 0), that is for all (F, L), (F ′, L′) ∈ ∆+ × ∆+ dL(F ⋆ L, F ′ ⋆ L′) ≤ ω⋆(dL(F, F ′) + dL(L, L′)). In particular for all (F, L) ∈ ∆+ × ∆+ dL(F ⋆ L, L) ≤ ω⋆(dL(F, H0)). (3) If moreover the operation ⋆ is k-Lipschitz (with respect to dL) for some positive number k then ω⋆(t) = kt for all t ≥ 0 (necessarily k ≥ 1, by using 3 with L = H0) is a modulus of uniform continuity. We give in Proposition 4 examples of k-Lipschitz triangle function using k-Lipschitz t-norms. 3. Metrization of Probabilistic Metric space. We give below the main result of this section, that is a metrization of probabilistic metric space extending the result of Schweizer, Sklar and Thorp in [14]. Let (G, D, ⋆) be a probabilistic metric space. We define canonically the metric σD on G using the probabilistic metric D as follows: for all x, y ∈ G σD(x, y) := sup z∈K dL(D(x, z), D(y, z)) := sup z∈K dL(δx(z), δy(z)) := d∞(δx, δy) It is easy to see that σD is a metric on G and that for all x, y ∈ G dL(D(x, y), H0) ≤ σD(x, y). Theorem 1. Let (G, D, ⋆) be a probabilistic metric space such that ⋆ is continuous (resp. k-lipschitz). Let ω⋆ be a modulus of uniform continuity of ⋆ on ∆+. Then, the metric σD satisfies: for all x, y ∈ G dL(D(x, y), H0) ≤ σD(x, y) ≤ ω⋆(dL(D(x, y), H0)). (resp. dL(D(x, y), H0) ≤ σD(x, y) ≤ kdL(D(x, y), H0)). 6 MOHAMMED BACHIR, BRUNO NAZARET In particular, the identity map i : (G, τ ) → (G, σD) is an uniform homeomorphism, where τ is the topology induced by the strong neighborhood system {Nx(t); x ∈ G, t > 0} (see Definition 6). This theorem is a mere consequence of the following lemma, that we will also use for proving Theorem 5. Lemma 4. Let (G, D, ⋆) be a probabilistic metric space such that ⋆ is continuous. Let ω⋆ be a modulus of uniform continuity of ⋆ on ∆+. Then, the set Lip1 ⋆(G, ∆+) is uniformly equicontinuous. More precisely, we have ∀x, y ∈ G : sup ⋆(G,∆+) f ∈Lip1 dL(f (x), f (y)) ≤ ω⋆(dL(D(x, y), H0)). Proof. From the formula (3) about the modulus of uniform continuity of ⋆, we have that ∀L ∈ ∆+, ∀x, y ∈ G : dL(D(x, y) ⋆ L, L) ≤ ω⋆(dL(D(x, y), H0)). In particular, we have for all f ∈ Lip1 ⋆(G, ∆+) and all x, y ∈ G, max[dL(D(x, y) ⋆ f (x), f (x)), dL(D(x, y) ⋆ f (y), f (y))] ≤ ω⋆(dL(D(x, y), H0)), hence, it is enough to prove that dL(f (x), f (y)) ≤ max[dL(D(x, y) ⋆ f (x), f (x)), dL(D(x, y) ⋆ f (y), f (y))]. Let h1, h2 > 0 such that, [0, h−1 1 [⊂ Ah1 D(x,y)⋆f (x),f (x) ∩ Ah1 f (x),D(x,y)⋆f (x). that is, for all t ∈]0, h−1 [0, h−1 D(x,y)⋆f (y),f (y) ∩ Ah2 2 [⊂ Ah2 1 [ and all t′ ∈]0, h−1 2 [, we have f (y),D(x,y)⋆f (y). (4) (5) 0 ≤ D(x, y) ⋆ f (x)(t) ≤ f (x)(t + h1) + h1 0 ≤ f (x)(t) ≤ D(x, y) ⋆ f (x)(t + h1) + h1 0 ≤ D(x, y) ⋆ f (y)(t′) ≤ f (y)(t′ + h2) + h2 0 ≤ f (y)(t′) ≤ D(x, y) ⋆ f (y)(t′ + h2) + h2. From the second, the fourth inequalities and the fact that f is 1-Lipschitz, we get that for all t ∈]0, h−1 1 [ and all t′ ∈]0, h−1 2 [ 0 ≤ f (x)(t) ≤ f (y)(t + h1) + h1 0 ≤ f (y)(t′) ≤ f (x)(t′ + h2) + h2. It follows that for all s ∈]0, max(h1, h2)−1[ (a subset of ]0, min(h−1 1 , h−1 2 )[) 0 ≤ f (x)(s) ≤ f (y)(s + max(h1, h2)) + max(h1, h2) 0 ≤ f (y)(s) ≤ f (x)(s + max(h1, h2)) + max(h1, h2). Thus, we have that dL(f (x), f (y)) ≤ max(h1, h2) for all h1, h2 > 0 satisfying (4) and (5). This implies that dL(f (x), f (y)) ≤ max(dL(D(x, y) ⋆ f (x), f (x)), dL(D(x, y) ⋆ f (y), f (y))), and the conclusion. Let us now prove Theorem 1. (cid:3) METRIZATION OF PROBABILISTIC METRIC SPACES 7 Proof of Theorem 1. The inequality at the left is a direct consequence of the defi- nition of σD. To prove the inequality at the right, we use Lemma 4 noticing that G(G) := {δx/x ∈ G} ⊂ Lip1 ⋆(G, ∆+). The second part of the theorem follows from Lemma 3 since, y ∈ Nx(t) if and only if dL(D(x, y), H0) < t for each t > 0 and x, y ∈ G. (cid:3) The notion of probabilistic distance naturally leads to associated metric concepts, such as Cauchy sequence, completeness, separability, density and compatness. Definition 8. A complete probabilistic metric space (K, D, ⋆) is called compact if for all t > 0, the open cover {Nx(t) : x ∈ K} has a finite subcover. Definition 9. In a probabilistic metric space (G, D, ⋆), a sequence (zn) ⊂ G is said to be a Cauchy sequence if for all t ∈ R, lim n,p−→+∞ D(zn, zp)(t) = H0(t). (Equivalently, if D(zn, zp) w−→ H0 or dL(D(zn, zp), H0) → 0, when n, p −→ +∞). A probabilistic metric space (G, D, ⋆) is said to be complete if every Cauchy sequence (zn) ⊂ G weakly converges to some z∞ ∈ G, that is limn→+∞ D(zn, z∞)(t) = H0(t) for all t ∈ R, we will briefly note limn D(zn, z∞) = H0. Corollary 1. Let (G, D, ⋆) be a probabilistic metric space such that ⋆ is continuous. Then, the following assertions hold. (1) (G, D, ⋆) is a probabilistic complete metric space iff (G, σD) is a complete metric space. (2) (G, D, ⋆) is compact as probabilistic metric space iff (G, σD) is a compact metric space. (3) (G, D, ⋆) is separable as probabilistic metric space iff (G, σD) is separable metric space. Proof. It is a direct consequence of Theorem 1 using Lemma 3. (cid:3) Notice that several results in the litterature proved for probabilistic metric spaces could be easily deduced from Corollary 1 and Theorem 1. For instance, recall that a Baire space is a topological space such that every intersection of a countable collection of open dense sets is also dense. In [15], H. Sherwood proved that a complete Menger space under a continuous t-norm, equipped with the topology τ induced by the strong neighborhood system {Nx(t); x ∈ G, t > 0} is a Baire space. Now, Theorem 1 expressing the fact that as soon as the triangle function is continuous then the induced topology is metrizable, we immediately obtain the following result. Proposition 1. Let (G, D, ⋆) be a probabilistic complete metric space such that ⋆ is continuous. Let τ be the topology induced by strong neighborhood system {Nx(t); x ∈ G, t > 0} (see Theorem 1). Then, (G, τ ) is a Baire space. In the same spirit, we also easily recover the following proposition already proven by other means in [10, Theorem 2.2, Theorem 2.3]. Proposition 2. Let (K, D, ⋆) be a probabilistic metric space. Suppose that the triangle function ⋆ is continuous. Then, 8 MOHAMMED BACHIR, BRUNO NAZARET (1) (K, D, ⋆) is compact as probabilistic metric space iff every sequence of K has a convergent subsequence. (2) If (K, D, ⋆) is compact as probabilistic metric space, then it is separable. We end the section by showing that the metric σD is canonical in the following sens. We know that every (complete) metric space induce a probabilistic (complete) metric space. Indeed, if d is a (complete) metric on G and ⋆ is a triangle function on ∆+ satisfying Ha ⋆ Hb = Ha+b for all a, b ∈ R+ (see references [12] and [4]), then (G, D, ⋆) is a probabilistic (complete) metric space, where D(p, q) = Hd(p,q), ∀p, q ∈ G. Using Proposition 3 below, we get that dL(D(p, q), H0) ≤ σD(p, q) := sup z∈G dL(D(p, z), D(z, q)) = sup z∈G ≤ sup z∈G dL(Hd(p,z), Hd(z,q)) min(1, d(p, z) − d(z, q)) = min(1, d(p, q)) = dL(Hd(p,q), H0) = dL(D(p, q), H0) Thus, we have the equality dL(D(p, q), H0) = σD(p, q) = min(1, d(p, q)). It follows that σD(p, q) = d(p, q), for all p, q ∈ G such that d(p, q) ≤ 1. In particular, σD and d coincides if (G, d) is of diameter less than 1. Proposition 3. Let a, b ≥ 0. Then, dL(Ha, Hb) = min(cid:18)1, b − a, 1 min(a, b)(cid:19) , and, in particular, dL(Ha, Hb) ≤ min (1, b − a) = dL(Hb−a, H0). (6) Notice that the inequality (6) expresses the more general fact that, for all λ > 0 and for all F , G ∈ ∆+, which is a consequence of the following property, dL(τλF, τλG) ≤ dL(F, G), ∀λ > 0, (cid:8)h > 0, [0, h−1[⊂ Ah F,G(cid:9) ⊂ (cid:8)h > 0, [0, h−1[⊂ Ah τλF,τλG(cid:9) , where τλF (t) = F (t − λ). This contraction property is an equality for the standard Levy metric while Proposition (3) shows that it is not true for the modified version dL. Proof. In this proof, we will assume without loss of generality that a < b and use the shortened notation Ah a,b := Ah Ha,Hb = {t ≥ 0, Ha(t) ≤ Hb(t + h) + h} , since in this case we have Ha ≥ Hb. Notice that the inequality Ha(t) ≤ Hb(t + h) + h METRIZATION OF PROBABILISTIC METRIC SPACES 9 is immediate for t ∈ [0, a], while if t > a and since h < 1, it is equivalent to that is t + h > b. As a consequence, Hb(t + h) ≥ 1 − h > 0, Ah a,b = [0, a] ∪ (]a, +∞[∩]b − h, +∞[) = [0, a]∪] max(a, b − h), +∞[. We then have 2 cases : • If a > 1, then for all h ≥ a−1, [0, h−1[⊂ [0, a] ⊂ Ah In addition, if a,b if and only if b − h ≤ a, that is h ≥ b − a. a,b. h < a−1, then [0, h−1[⊂ Ah This leads to (cid:8)h > 0, [0, h−1[⊂ Ah a,b(cid:9) = [a−1, +∞[∪[b − a, +∞[= [min(a−1, b − a), +∞[, hence in this case, dL(Ha, Hb) = min(a−1, b − a) = min(1, a−1, b − a). • If a ≤ 1, we have h1 > a for all h ∈]0, 1[, hence [0, h−1[⊂ Ah a,b if and only if b − h ≤ a, that is if h ≥ b − a. It follows that (cid:8)h > 0, [0, h−1[⊂ Ah a,b(cid:9) = [1, +∞]∪(]0, 1[∩[min(1, b − a), +∞[) = [min(1, b−a), +∞[, hence, in this case, dL(Ha, Hb) = min(1, b − a) = min(1, a−1, b − a). This concludes the proof. (cid:3) 4. Fixed point and contraction This section is divided on two subsections. In Subsection 4.1, we give two new fixed point theorems and in Subsection 4.2, we give some general examples of k- Lipschitz triangle functions constructed canonically from k-Lipschitz t-norms. 4.1. Fixed point theorem. Let us start from the following probabilistic notion of contraction introduced by Hicks (see, [4]). Definition 10. Let (G, D, ⋆) be a probabilistic metric space. A map f : G → G is said to be a C-contraction if there exists q ∈ (0, 1) such that for every x, y ∈ G and every t > 0 D(x, y)(t) > 1 − t =⇒ D(f (x), f (y))(qt) > 1 − qt. Lemma 5. A map f : G → G is a C-contraction with constant q iff for all x, y ∈ G, dL(D(f (x), f (y)), H0) ≤ qdL(D(x, y), H0). Proof. From Lemma 3, we have that for every x, y ∈ G, D(x, y)(t) > 1 − t if and only if dL(D(x, y), H0) < t. For every ε > 0, set tε = dL(D(x, y), H0) + ε > 0. Then, D(x, y)(tε) > 1 − tε. Suppose that f is a C-contraction, then we have that D(f (x), f (y))(qtε) > 1 − qtε which is equivalent to dL(D(f (x), f (y), H0) ≤ qtε = q(dL(D(x, y), H0) + ε). Sending ε to 0, we get dL(D(f (x), f (y), H0) ≤ qdL(D(x, y), H0). The converse is straightforward. (cid:3) Hicks proved that a C-contraction map in Menger space under the minimum t-norm TM (a, b) = min(a, b) has a unique fixed point. We can find a extension of this result for generalised C-contraction in Menger space in [4]. We introduce the following new definition of contraction. 10 MOHAMMED BACHIR, BRUNO NAZARET Definition 11. Let (G, D, ⋆) be a probabilistic metric space. Suppose that ⋆ is a continuous triangle function (hence uniformly continuous) and let ω⋆ be a modulus of uniform continuity of ⋆. A map f : G → G is said to be a ω⋆-contraction if there exists q ∈ (0, 1) such that for every x, y ∈ G ω⋆[dL(D(f (x), f (y)), H0)] ≤ qω⋆[dL(D(x, y), H0)]. Remark 2. Using Lemma 5, the notion of ω⋆-contraction concides with the C- contraction, when the triangle function ⋆ is k-Lipschitz since in this case ω⋆(t) = kt for all t ≥ 0 is a modulus of uniform continuity. Examples of k-Lipschitz triangle functions are given in Proposition 4. The original result of Hicks is a particular case corresponding to the 1-Lipschitz triangle function ⋆TM . Using Theorem 1 and the Ekeland variational principle, we give below an exten- sion of the result of Hicks in probabilistic metric spaces which are not necesarily Menger spaces, where the triangle function ⋆ is continuous. Notice that this result seems to be new even in the non probabilistic setting. Theorem 2. Let (G, D, ⋆) be a probabilistic complete metric space, where ⋆ is continuous triangle function with modulus of uniform continuity ω⋆. Let f : G → G be a ω⋆-contraction with a constant of contraction q ∈ (0, 1). Then, f has a unique fixed point x∗ ∈ G. Proof. By assumption, we have for all x, y ∈ G ω⋆(dL(D(f (x), f (y)), H0)) ≤ qω⋆(dL(D(x, y), H0)). Let us consider the function φ : (G, σD) → R defined by φ(x) = ω⋆(dL(D(x, f (x)), H0)) and prove that φ is continuous. Indeed, for x, y ∈ G, from the triangle inequality for dL, the definition of σD and Theorem 1 we have dL(D(x, f (x)), H0) − dL(D(y, f (y)), H0) ≤ dL(D(x, f (x)), D(f (x), y)) +dL(D(f (x), y), D(y, f (y))) ≤ σD(x, y) + σD(f (x), f (y)) ≤ σD(x, y) + ω⋆(dL(D(f (x), f (y)), H0)) ≤ σD(x, y) + qω⋆(dL(D(x, y), H0)). From the continuity of ω⋆, we have that dL(D(x, y), H0) ≤ σD(x, y) → 0 =⇒ ω⋆(dL(D(x, y), H0)) → 0. Thus, from the above inequalities, the function x 7→ dL(D(x, f (x)), H0) is con- tinuous from (G, σD) into R, and by composing it with the uniformly continuous function ω⋆, we get that φ is continuous. Now, by the Ekeland variational principle [3] (since (G, σD) is a complete metric space by Corollary 1), let ε > 0 and u ∈ G such that φ(u) ≤ inf G φ + ε. Then, for all λ > 0 there exists v ∈ G : (i) φ(v) ≤ φ(u) ; (ii) σD(u, v) ≤ λ; (iii) for all x ∈ G x 6= v, φ(v) < φ(x) + ε Now, let us choose ε < 1 − q and set λ = 1. Using Theorem 1 and (iii), we have λ σD(x, v). that φ(v) ≤ φ(x) + εω⋆(dL(D(x, v), H0)), for all x ∈ G. (7) METRIZATION OF PROBABILISTIC METRIC SPACES 11 We claim that x∗ := f (v) is the unique fixed point of f . Indeed, we have ω⋆(dL(D(x∗, f (x∗)), H0)) = ω⋆(dL(D(f (v), f (x∗)), H0)) ≤ qω⋆(dL(D(v, x∗)). (8) By (7) with x∗ and v, we have for all x ∈ G ω⋆(dL(D(v, x∗), H0)) = ω⋆(dL(D(v, f (v)), H0)) ≤ ω⋆(dL(D(x∗, f (x∗)), H0)) + εω⋆(dL(D(x∗, v), H0)).(9) Combining (8) and (9), we get ω⋆(dL(D(v, x∗), H0)) ≤ (q + ε)ω⋆(dL(D(v, x∗), H0)). Since q+ε < 1, we obtain that ω⋆(dL(D(v, x∗), H0)) = 0, which implies by Theorem 1 that σD(v, x∗) = 0, that is x∗ = v. Thus, f (x∗) = f (v) =: x∗. The unicity of the fixed point x∗ is immediate from q < 1. (cid:3) Applying the Banach fixed point we give the following extenstion of Hicks's result with an estimation of convergence of sequences xn+1 = f (xn). Note that we recover the Hicks's result with ⋆ = ⋆TM which is k-Lipschitz, with k = 1. Theorem 3. Let (G, D, ⋆) be a probabilistic complete metric space, where ⋆ is k-Lipschitz triangle function (k ≥ 1). Let f : G → G be a C-contraction with a constant of contraction q ∈]0, 1 k [. Then, f has a unique fixed point x∗ ∈ G. Moreover, every sequence (xn) of G such that xn+1 = f (xn), satisfies: for every t > 0 D(x1, x0)(t) > 1 − t =⇒ D(xn, x∗)( k(kq)n 1 − kq t) > 1 − k(kq)n 1 − kq t, or equivalently, dL(D(xn, x∗), H0) ≤ k(kq)n 1 − kq dL(D(x1, x0), H0). In particular, dL(D(xn, x∗), H0) → 0, when n → +∞. Proof. By Lemma 5, we have that dL(D(f (x), f (y), H0) ≤ qdL(D(x, y), H0), for all x, y ∈ G. Using Theorem 1, we get σD(f (x), f (y)) ≤ qkσD(x, y). Since qk < 1, we can apply the Banach fixed point theorem in the complete metric space (G, σD). Thus, we obtain a unique fixed point x∗ such that, for all n ∈ N σD(xn, x∗) ≤ (kq)n 1 − kq σD(x1, x0). Using again Theorem 1 (the k-Lipschitz part) we give dL(D(xn, x∗), H0) ≤ k(kq)n 1 − kq dL(D(x1, x0), H0), which is equivalent by Lemma 5 to: for all t > 0 D(x1, x0)(t) > 1 − t =⇒ D(xn, x∗)( k(kq)n 1 − kq t) > 1 − k(kq)n 1 − kq t. (cid:3) 12 MOHAMMED BACHIR, BRUNO NAZARET 4.2. k-Lipschitz triangle function. One of the standard way to construct trian- gle function goes in the following way. We refer to [4] for more details. Definition 12. We denote by L the set of all binary operators L on [0, +∞[ which satisfy the following conditions: (i) L maps [0, +∞[2 to [0, +∞[ (ii) L is non-deceasing in both coordinate (iii) L is continuous on [0, +∞[2. For a t-norm T , we define the operation ⋆T,L from ∆+ × ∆+ to ∆+ as follows: for every F, G ∈ ∆+ and every t ≥ 0 (F ⋆T,L G)(t) = sup T (F (u), G(v)). L(u,v)=t In the speciale case where L(u, v) = u + v we obtain ⋆T,L = ⋆T . Theorem 4. commutative, associative, has 0 as identity and satisfy the condition ([4, Theorem 2.15]) if T is a left-continuous t-norm and L ∈ L is if u1 < u2 and v1 < v2 then L(u1, v1) < L(u2, v2), then, ⋆T,L is a triangle function. The above theorem works for example with L(u, v) := L+(u, v) = u + v or L(u, v) := LM (u, v) = max(u, v). Another way to construct a triangle function from a t-norm T is the use the t-conorm T ∗(u, v) = 1 − T (1 − u, 1 − v) as follows : for every F, G ∈ ∆+ and for every s > 0 (F ⋆T ∗ G)(s) = inf u+v=s T ∗(F (u), G(v)). Recall that a t-norm T : [0, 1]×[0, 1] → [0, 1] is k-Lipschitz if there exists k ∈ [0, +∞[ such that, for all a, b, c, d ∈ [0, 1], we have T (a, b) − T (c, d) ≤ k(a − c + b − d). Since T (x, 1) = x, we necessarily have that k ≥ 1. Note also that the minimum t-norm TM (a, b) := min(a, b) is 1-Lipschitz. Other examples of k-Lipschitz t-norms are studied in [7, 8, 9]. In order to give examples of k-Lipschitz triangle functions in Proposition 4, we need the following lemma. Lemma 6. Let T : [0, 1] × [0, 1] → [0, 1] be a k-Lipschitz t-norm. Then, for every (a, b), (c, d) ∈ [0, 1]×[0, 1] and every h ∈ [0, +∞[ such that a ≤ c+h1 and b ≤ d+h2 we have that T (a, b) − T (c, d) ≤ k(h1 + h2) T ∗(a, b) − T ∗(c, d) ≤ k(h1 + h2). Proof. Four cases are discussed. case 1. If a ≤ c and b ≤ d. In this case, since T is a t-norm, then T (a, b) − T (c, d) ≤ 0 ≤ k(h1 + h2). case 2. If a ≤ c and b ≥ d. In this case, since T is a t-norm, then T (a, b) ≤ T (c, b) and so since it is k-Lipschitz we have that T (a, b) − T (c, d) ≤ T (c, b) − T (c, d) ≤ kb − d = k(b − d) ≤ kh1 ≤ k(h1 + h2). METRIZATION OF PROBABILISTIC METRIC SPACES 13 case 3. If a ≥ c and b ≤ d. This case is similar to case 2. case 4. If a ≥ c and b ≥ d. In this case, since T is k-Lipschitz we have that T (a, b) − T (c, d) ≤ k(a − c + b − d) = k(a − c + b − d) ≤ k(h1 + h2). The case of T ∗ comes easily from the case of T . (cid:3) In the following proposition, we consider the cases where L(u, v) := L+(u, v) = u + v and L(u, v) := LM (u, v) = max(u, v). Proposition 4. Let T : [0, 1] × [0, 1] → [0, 1] be a k-Lipschitz t-norm (k ≥ 1). Then, the triangle functions ⋆T ; ⋆T,LM ; ⋆T ∗ and T are k-Lipschitz , where for all F, G ∈ ∆+ and all s > 0 (F ⋆T G)(s) = sup u+v=s T (F (u), G(v)), (F ⋆T ∗ G)(s) = inf u+v=s T ∗(F (u), G(v)), (F ⋆T,LM G)(s) = sup T (F (u), G(v)), max(u,v)=s T(F, G)(s) = T (F (s), G(s)). Proof. We give the prove for ⋆T , the technique is similar for the other triangle functions. Let F, F ′, G, G′ ∈ ∆+. Let h1, h2 > 0 be such that [0, h−1 1 [⊂ Ah1 F,F ′ ∩ Ah1 meanning that for all t ∈]0, h−1 F ′,F and [0, h−1 1 [ and all t′ ∈]0, h−1 2 [⊂ Ah2 2 [ we have: G,G′ ∩ Ah2 G′,G, (10) 0 ≤ F (t) ≤ F ′(t + h1) + h1 0 ≤ F ′(t) ≤ F (t + h1) + h1 0 ≤ G(t′) ≤ G′(t′ + h2) + h2 0 ≤ G′(t′) ≤ G(t′ + h2) + h2. Thus, combining the first and the third (resp. the second and the fourth) in- equalities, and using Lemma 6, we have that for every u, v ∈]0, max(h1, h2)−1[(⊂ ]0, min(h−1 1 , h−1 0 ≤ T (F (u), G(v)) ≤ T (F ′(u + h1), G′(v + h2)) + k(h1 + h2) 0 ≤ T (F ′(u), G′(v)) ≤ T (F (u + h1), G(v + h2)) + k(h1 + h2) 2 )[) Let s ∈]0, max(h1, h2)−1[, taking the supremum over 0 ≤ u, v such that u + v = s in the above inequalities with the fact that k ≥ 1, we get 0 ≤ (F ⋆T G)(s) ≤ (F ′ ⋆T G′)(s + (h1 + h2)) + k(h1 + h2) ≤ (F ′ ⋆T G′)(s + k(h1 + h2)) + k(h1 + h2) 0 ≤ (F ′ ⋆T G′)(s) ≤ (F ⋆T G)(s + (h1 + h2)) + k(h1 + h2) ≤ (F ⋆T G)(s + k(h1 + h2)) + k(h1 + h2) This shows that for all h1, h2 > 0 satisfying (10), we have dL(F ⋆T G, F ′ ⋆T G′) ≤ k(h1 + h2). Thus, taking the infinimum over h1 and h2, we get dL(F ⋆T G, F ′ ⋆T G′) ≤ k(dL(F, F ′) + dL(G, G′)). 14 MOHAMMED BACHIR, BRUNO NAZARET (cid:3) 5. Probabilistic Arzela-Ascoli type theorem This section is divided on two subsections. In subsection 5.1 we give some general definitions of probabilistic function spaces and in subsection 5.2, we give the main result of this section, a probabilistic Arzela-Ascoli type theorem. 5.1. The space of continuous and functions. We are going to define continuity of functions defined from a probabilistic metric space (G, D, ⋆) to a (deterministic) metric space (F, dF ). Definition 13. Let (G, D, ⋆) be a probabilistic metric space, (F, dF ) be a met- ric space and let f be a function f : (G, D, ⋆) −→ (F, dF ). We say that f is (probabilistic) continuous at z ∈ G if dF (f (zn) f (z)) → 0 whenever D(zn, z) w−→ H0 (equivalently dL(D(zn, z), H0) → 0). We say that f is continuous if f is continuous at each point z ∈ G. By C⋆(G, F ) we denote the space of all (probabilistic) continuous functions f : (G, D, ⋆) −→ (F, dF ). By C(G, F ) we denote the space of all (deterministic) continuous functions f : (G, σD) −→ (F, dF ). We both equip the spaces C⋆(G, F ) and C(G, F ) with the uniform metric d∞(f, g) := sup x∈G dF (f (x), g(x)) As in the standard case, the completeness of C⋆(G, F ) only relies on the complete- ness of the arrival space. Proposition 5. Let (G, D, ⋆) be a probabilistic metric space (here ⋆ is not as- sumed to be continuous) and (F, dF ) be a complete metric space. Then, the space (C⋆(G, F ), d∞) is a complete metric space. Proof. Let (fn) be a Cauchy sequence in (C⋆(G, F ), d∞). In particular, for each x ∈ G, (fn(x)) is Cauchy in (F, dF ) which is complete. Thus, there exists a function f : G −→ F such that the sequence (fn) pointwise converges to f on G. It is easy to see that in fact (fn) uniformly converges to f , since it is Cauchy sequence in (C⋆(G, F ), d∞). We need to prove that f is a continuous function from (G, D, ⋆) into (F, dF ). Let x ∈ G and (xk) be a sequence such that dL(D(xk, x), H0) −→ 0, when k −→ +∞. For all ε > 0, there exists Nε ∈ N such that n ≥ Nε =⇒ d∞(fn, f ) := sup x∈G dF (fn(x), f (x)) ≤ ε Using the continuity of fNε, we have that there exists η(ε) > 0 such that dL(D(xk, x), H0) ≤ η(ε) =⇒ dF (fNε(xk), fNε(x)) ≤ ε Using (11) and (12), we have that (11) (12) dF (f (xk), f (x)) ≤ dF (f (xk), fNε(xk)) + dF (fNε(xk), fNε(x)) + dF (fNε(x), f (x)) ≤ 3ε This shows that f is continuous on G. Finally, we proved that every Cauchy sequence (fn) uniformly converges to a continuous function f . In other words, the space (C⋆(G, F ), d∞) is complete. (cid:3) METRIZATION OF PROBABILISTIC METRIC SPACES 15 Since, for all x, y ∈ G, dL(D(x, y), H0) ≤ σD(x, y), We have in general that C⋆(G, F ) ⊂ C(G, F ). Assuming the continuity of the triangle function ⋆, we obtain the equality. Proposition 6. Let (G, D, ⋆) be a probabilistic metric space and (F, d) be a metric space. Suppose that ⋆ is continuous. Then, we have C⋆(G, F ) = C(G, F ). In the case where (F, dF ) = (∆+, dL), we also have that Lip1 ⋆(G, ∆+) ⊂ C⋆(G, ∆+). Proof. Thanks to Theorem 1, we have, for all x, y ∈ G, σD(x, y) ≤ ω⋆ (dL(D(x, y)), H0) , providing the equality between C⋆(G, F ) and C(G, F ). For the second part of the statement, we use Lemma 4 to ensure that, for any f ∈ Lip1 ⋆(G, ∆+) and for any x, y ∈ G, dL(f (x), f (y)) ≤ sup ⋆(G,∆+) g∈Lip1 [dL(g(x), g(y))] ≤ ω⋆ (dL(D(x, y)), H0) , which gives the conclusion. (cid:3) Remark 3. We do not know if Lip1 ⋆(G, ∆+) ⊂ C⋆(G, ∆+) when ⋆ is not continuous. ⋆(K, ∆+). The following 5.2. Arzela-Ascoli type theorem for the space Lip1 proposition gives a canonical way to build probabilistic 1-Lipschitz maps from (G, D, ⋆) into ∆+. Definition 14. A triangle function ⋆ is said to be sup-continuous (see for instance [2]) if for all nonempty set I and all familly (Fi)i∈I of distributions in ∆+ and all L ∈ ∆+, we have sup i∈I (Fi ⋆ L) = sup i∈I (Fi) ⋆ L. Proposition 7. Let (G, D, ⋆) be a probabilistic metric space such that ⋆ is sup- continuous. Let f : (G, D, ⋆) −→ ∆+ be any map and A be any no-empty subset of G. Then, the map fA(x) := supy∈A[f (y) ⋆ D(x, y)], for all x ∈ G is a probabilistic 1-Lipschitz map and we have fA(x) ≥ f (x), for all x ∈ A. Proof. The proof is similar to the standard inf-convolution construction. The fact that fA(x) ≥ f (x) for all x ∈ A is immediate from the definition of fA. Let us now prove that it is probabilistic 1-Lipschiptz. Let x, y ∈ G. Then, for all z ∈ A, we have fA(y) = sup z∈A [f (z) ⋆ D(y, z)] ≥ f (z) ⋆ D(y, z) ≥ f (z) ⋆ (D(y, x) ⋆ D(x, z)) = (f (z) ⋆ D(x, z)) ⋆ D(y, x). We get the conclusion by taking the supremum with respect to z ∈ A and using the sup-continuity of ⋆. (cid:3) Let us now recall the following result from [1]. Proposition 8. ([1, Proposition 3.5]) Let (Fn), (Ln), (Kn) ⊂ (∆+, ⋆). Suppose that (a) the triangle function ⋆ is continuous, (b) Fn w−→ L and Kn w−→ F , Ln w−→ K. 16 MOHAMMED BACHIR, BRUNO NAZARET (c) for all n ∈ N, Fn ⋆ Ln ≤ Kn. Then, F ⋆ L ≤ K. Lemma 7. Let (K, D, ⋆) be a probabilistic compact metric space and (fn) be a sequence of probabilistic 1-Lipschitz maps. Suppose that there exists a function f defined from K into ∆+ such that, for all x ∈ K, dL(fn(x), f (x)) −→ 0, as n → +∞. Then, f is (probabilistic) 1-Lipschitz on K and (fn) converges uniformly to f , that is, d∞(fn, f ) −→ 0, as n → +∞. Proof. Since each fn is a 1-Lipschitz map, we have for all x, y ∈ L and for all n ∈ N: Using Proposition 8, we get that for all x, y ∈ L D(x, y) ⋆ fn(x) ≤ fn(y) D(x, y) ⋆ f (x) ≤ f (y) In other words, f is 1-Lipschitz maps on L (Note that up to now, we have not needed to use the compactness of K). Now, let ε > 0 and, using Lemma 4, let η(ε) be the uniform modulus of equicon- ⋆(K, ∆+). Since (K, D, ⋆) is compact, there exists a finite set tinuity for the set Lip1 A such that K = ∪a∈ANa(η(ε)). Since dL(fn(a), f (a)) −→ 0, as n → +∞ for all a ∈ A. Then, for each a ∈ A, there exists Pa ∈ N such that n ≥ Pa =⇒ dL(fn(a), f (a)) ≤ ε Since A is finite, we have that n ≥ max a∈A Pa =⇒ sup a∈A dL(fn(a), f (a)) ≤ ε Thus, for all x ∈ K = ∪a∈ANa(η(ε)), there exists a ∈ A such that x ∈ Na(η(ε)) and so we have that for all n ≥ maxa∈A Pa : dL(fn(x), f (x)) ≤ dL(fn(x), fn(a)) + dL(fn(a), f (a)) + dL(f (a), f (x)) ≤ 3ε. In other words, n ≥ max a∈A Pa =⇒ d∞(fn, f ) := sup x∈K dL(fn(x), f (x)) ≤ 3ε (cid:3) We give now our main result of this section. For the classical Arzela-Ascoli theorem we refer to the book of L. Schwartz, Analyse I, "Thorie des ensembles et Topologie", page 346. Theorem 5. Let (K, D, ⋆) be a probabilistic complete metric space such that ⋆ is continuous. Then, the following assertions are equivalent. (1) (K, D, ⋆) is compact. (2) The metric space (Lip1 is a compact subset of (C⋆(K, ∆+), d∞) = (C(K, ∆+), d∞)). ⋆(K, ∆+), d∞) is compact (or equivalently, Lip1 ⋆(K, ∆+) Proof. • (1) =⇒ (2) Suppose that (K, D, ⋆) is compact, equivalently (K, σD) is compact by Corollary 1. Using Lemma 4 and Theorem 1, the set Lip1 ⋆(G, ∆+) is uniformly equicontinuous with respect to the metric σD. Moreover, (∆+, dL) is METRIZATION OF PROBABILISTIC METRIC SPACES 17 compact, hence Lip1 Ascoli theorem. On the other hand, by Lemma 7, the set Lip1 (C(G, ∆+), d∞). Hence it is compact. ⋆(G, ∆+) is relatively compact in (C(K, ∆+), d∞) by Arzela- ⋆(G, ∆+) is closed in • (2) =⇒ (1) Suppose that (Lip1 ⋆(K, ∆+), d∞) is compact. Let (xn) be a sequence of K. We need to prove that (xn) has a convergent subsequence. Consider the sequence (δxn) of 1-Lipschitz maps, defined by δxn : x 7→ D(xn, x) for each n ∈ N. By assumption, there exists a subsequence (δxϕ(n)) that converges uniformly to some 1-Lipschitz map, in particular it is a Cauchy sequence. In other words, we have lim p,q−→+∞ sup x∈K dL(δxϕ(p) (x), δxϕ(q) (x)) = 0. In particular we have or equivalently, lim p,q−→+∞ dL(δxϕ(p) (xϕ(q)), H0) = 0, lim p,q−→+∞ dL(D(xϕ(p), xϕ(q)), H0) = 0. This shows that the sequence (xϕ(n)) is Cauchy in (K, σD) (see Theorem 1). Thus, the sequence (xϕ(n)) converges to some point x ∈ K for the metric σD, since (K, σD) is complete. Hence, (K, σD) is compact, equivalently, (K, D, ⋆) is compact. This ends the proof. (cid:3) References 1. M. Bachir, The Space of Probabilistic 1-Lipschitz map, Aequationes Math. (2019) 1-29. 2. S. Cobzas, Completeness with respect to the probabilistic Pompeiy-Hausdorff metric Studia Univ. "Babes-Bolyai", Mathematica, Volume LII, Number 3, (2007) 43-65. 3. I. Ekeland On the variational principle, J. Math. Anal. Appl. 47 (1974), 324-353. 4. O. Hadzi´c and E. Pap Fixed Point Theory in Probabilistic Metric Space, vol. 536 of Math- ematics and Its Applications, Kluwer Academic Publishers, Dordrecht, The Netherlands, 2001. 5. E. P. Klement, R. Mesiar, E. Pap, Triangular norms I: Basic analytical and algebraic prop- erties, Fuzzy Sets and Systems 143 (2004), 5-26. 6. E. P. Klement, R. Mesiar, E. Pap, Triangular Norms. Kluwer, Dordrecht (2000). 7. A. Mesiarov´a: Triangular norms and k-Lipschitz property. In: Proc. EUSFLAT-LFA Con- ference, Barcelona 2005, pp. 922-926. 8. A. Mesiarov´a, k-lp-Lipschitz t-norms, International Journal of Approximate Reasoning 46 (2007) 596-604. 9. A. Mesiarov´a, Lipschitz continuity of triangular norms, in: B. Reusch (Ed.), Computational Intelligence, Theory and Applications (Proc. 9th Fuzzy Days in Dortmund), Springer, Berlin, 2006, pp. 309-321. 10. A. Mbarki, A. Ouahab, R. Naciri, On Compactness of Probabilistic Metric Space, Applied Mathematical Sciences Volume(8), (2014) 1703-1710. No. 124, Paris 1954. 11. B. Morrel, J. Nagata: Statistical metric spaces as related to topological spaces, Gen. Top. Appl. , 9 , (1978 233-237 . 12. B. Schweizer and A. Sklar, Probabilistic metric spaces, North-Holland Series in Probability and Applied Mathematics, North-Holland Publishing Co., New York, 1983. 247 (1958), 2092- 2094. 13. B. Schweizer, A. Sklar Statistical metric spaces, Pacific. J. Math. 10, (1960) 313-334. Math. Soc. 38 (1963), 401-406. 14. B. Schweizer, A. Sklar and E. Thorp, The metrization of statistical metric svaces, Pacific J. Math. 10 (1960), 673-675. 15. H. Sherwood, Complete probabilistic metric spaces and random variables generated spaces, Ph.D. Thesis, University of Arizona (1965). Math. Soc. s1-44, (1969) 441-448 Gebiete 20, (1971) 117-128. 18 MOHAMMED BACHIR, BRUNO NAZARET 16. D. A. Sibley A metric for weak convergence of distribution functions, Rocky Mountain J. Math. l (l971) 427-430. Laboratoire SAMM 4543, Universit´e Paris 1 Panth´eon-Sorbonne, Centre P.M.F. 90 rue Tolbiac, 75634 Paris cedex 13, France E-mail address: [email protected] E-mail address: [email protected]
1704.06722
1
1704
2017-04-21T23:09:43
On a theorem of Terzio\u{g}lu
[ "math.FA" ]
The theory of compact linear operators acting on a Banach space has such a classical core and is familiar to many. Perhaps lesser known is the characterization theorem of Terzio\u{g}lu for compact maps. In this paper we consider Terzio\u{g}lu's theorem and its consequences. We also give a similar characterization theorem in case where there is an approximation scheme on the Banach space.
math.FA
math
ON A THEOREM OF TERZIO GLU ASUMAN G UVEN AKSOY Abstract. The theory of compact linear operators acting on a Banach space has such a classical core and is familiar to many. Perhaps lesser known is the characterization theorem of Terzioglu for compact maps. In this paper we consider Terzioglu's theorem and its consequences. We also give a similar characterization theorem in case where there is an approximation scheme on the Banach space. 1. Introduction Let X and Y be Banach spaces and T : X → Y be an operator. We say T is compact if and only if it maps closed unit ball BX of X into a pre-compact subset of Y . In other words, T is compact if and only if for every norm bounded sequence {xn} of X, the sequence {T xn} has a norm convergent subsequence in Y . Equivalently, T is compact if and only if for every ǫ > 0, there exists elements y1, y2, . . . , yn ∈ Y such that T (BX) ⊆ {yk + ǫBY } n[k=1 where by BX and BY we mean the closed unit balls of X and Y respectively. Every compact linear operator is bounded, hence continuous, but clearly not every bounded linear map is compact since one can take the identity operator on an infinite dimensional space X. Compact operators are natural generalizations of finite rank operators and thus dealing with compact operators provides us with the closest analogy to the usual theorems of finite dimensional spaces. Recall that L(X, Y ) denotes the normed vector space of all continuous operators from X to Y and L(X) stands for L(X, X) and K(X, Y ) is the collection of all compact operators from X to Y . It is well known that if Y is a Hilbert space then any compact T : X → Y is a limit of finite rank operators, in other words if F (X, Y ) denotes the class of finite rank maps then, K(X, Y ) = F (X, Y ) where the closure is taken in the operator norm. However, the situation is quite different for Banach spaces, not every operator between Banach spaces is a uni- form limit of finite rank maps. For further information we refer the reader to a well known example due to P. Enflo [7], in which Enflo constructs a Banach space without the approximation property. The following classical results on compact operators will be used for our discussion later. 2010 Mathematics Subject Classification. Primary 47B07; Secondary 47B06. Key words and phrases. Compact Operators, Approximation Schemes. 1 2 A. AKSOY Theorem 1.1. For Banach spaces X, Y and Z, we have the following: (1) K(X, Y ) is a norm closed vector subspace of L(X, Y ) . (2) If X T−−→ Z are continuous operators and either S or T is compact, S−−→ Y then T S is likewise compact. If one consider the continuos operators on a Banach space X, the above theorem asserts the fact that compact operators on X form a two sided ideal in L(X). The following theorem of Schauder simply states that an operator is compact if and only if its adjoint is compact. Theorem 1.2 (Schauder). A norm bounded operator T : X → Y between Banach spaces is compact if and only if its adjoint T ∗ : Y ∗ → X ∗ is compact. The main idea in proving Schauder's theorem lies in the fact that PnT − T → 0 implies T ∗Pn − T ∗ → 0 where Pn : X → span{e1, . . . , en}. A well known proof of Schauder's theorem may be found in Yosida [[15], p.282]. For our discussion below we also need the following characterization of the compact sets in a Banach space; in some sense, it is a comment on the smallness of compact sets. Theorem 1.3 (Grothendieck). A subset of a Banach space is compact if and only if it is included in the closed convex hull of a sequence that converges in norm to zero. In other words, if we have K a compact subset of a Banach space X, then we can find a sequence {xn} in X such that For a proof we refer the reader to[ [6] , p.3] xn → 0 and K ⊆ co{xn}. 2. Terzioglu's Theorem Theorem 2.1 (Terzioglu [13]). An operator T : X → Y between two Banach spaces is compact if and only if there exists a sequence {un} of linear functionals in X ∗ with un → 0 such that the inequality holds for every x ∈ X. T x ≤ sup < un, x > n Proof. Suppose T : X → Y is compact, then by Schauder's theorem T ∗ : Y ∗ → X ∗ is compact; thus by definition, if V denotes the closed unit ball of Y ∗, T ∗(V ) is a norm totally bounded subset of X ∗. Now applying Grothendieck's result, we have a sequence {un} of elements of X ∗ with un → 0 and T ∗(V ) ⊆ co{un}. In other words, each element of T ∗(V ) can be written of the form αnun with ∞Xn=1 αn ≤ 1. ∞Xn=1 TERZIOGLU THM 3 Thus, for each x ∈ X we have T x = sup v≤1 < T ∗v, x > ≤ ∞Xn=1 αn sup < un, x > . n Suppose T satisfies the inequality T x ≤ supn < un, x > for some sequence {un} ∈ X ∗. For ǫ > 0 choose N such that un < ǫ for n > N and set Mǫ = {x ∈ X : < ui, x >= 0 for i = 1, 2, . . . N}, then one can have T ∗(V ) ⊂ ǫU + M ⊥ ǫ where U denotes the unit ball of X and for each linear subspace M of X, the polar of M denoted by M is a linear subspace of X ∗ defined as: M := {a ∈ X ∗ : < x, a > ≤ 1 for x ∈ M} , this shows that T ∗ is compact and hence T is compact. (cid:3) An application of the Theorem 2.1 yields that every compact mapping of a Banach space into a Pλ-space is ∞-nuclear. Definition 2.2. We say X is a Pλ space, (λ ≥ 1) if every bounded linear operator T from a Banach space Y to X and every Z ⊃ Y there is a linear extension T of Z to X with As illustrated in the following diagram: T ≤ λT . Z Y ∼ T X T If T = T in the above definition, we call X extendible. This property is related to the existence of a global Hahn-Banach type an extension. J. Linden- strauss in [9] examines the problem when is the extension T is compact if T itself is compact and the author's results are diverse and numerous and touches upon many related topics. Next, we define infinite nuclear mappings, this concept was first introduced in [11]. Definition 2.3. Let X and Y be Banach spaces and T : X → Y a linear operator. Then T is said to be infinite-nuclear, if there are sequences {un} ⊂ X ∗ and {yn} ⊂ Y such that lim n un = 0, v≤1( ∞Xn=1 sup v(yn) : v ∈ Y ∗) < +∞ and a sequence {yn} ∈ Y with sup v≤1 form T x = < un, x > yn. ∞Xn=1 ∞Xn=1 4 and for x ∈ X. A. AKSOY T x = ∞Xn=1 < un, x > yn As an application to Terzioglu's Theorem, under the condition that T : X → Y where Y is a Pλ-space, Terzioglu also obtains a precise expression for T x, which we state in the following: Theorem 2.4 ([13]). Let T be a compact mapping of a Banach space X into a Pλ space Y .Then for every ǫ > 0 there exists sequence {un} in X ∗ with lim n un = 0 and sup un ≤ T + ǫ n < v, yn > ≤ λ such that T has the The complete details of the proof can be found in [13]. However, it is worth pointing out that idea of the proof provides a factorization of a compact map through the space c0 as follows: Use Theorem 2.1, choose the sequence {un} in X ∗ satisfying lim n un = 0 and sup n Define linear mapping un ≤ T + ǫ and T x ≤ sup < un, x > . S : X → c0 by Sx = {< un, x >}, and observe that S is compact, then define a linear mapping R0 : S(X) → Y by R0(Sx) = T x, the inequality R0(Sx) = T x ≤ sup < un, x > = Sx implies that R0 ≤ 1. c0 S(E) ∼ R R0 F Since Y is a Pλ space, there exists an extension R of R0 from R :c0 → F with R ≤ λR0 = λ and E S−−→ c0 eR−−→ F, evidently T = eRS. TERZIOGLU THM 5 By considering {en} ∈ c0 and setting yn = eR(en) we obtain < v, yn > ≤ λ, and T x = < un, x > yn. sup v≤1 ∞Xn=1 ∞Xn=1 Using all of the above results of Terzioglu one can find the following conclusions in [14]. Corollary 2.5. (1) Every Pλ space has the approximation property. (2) Every compact linear operator of an L∞ space into a Banach space is infinite-nuclear. (3) Let T be a compact linear map of an infinite-dimensional space X into a Banach space Y . Then there exists an infinite dimensional closed subspace M of X such that TM : M → T (M) is infinite nuclear. 3. Compactness with Approximation Scheme Approximation schemes were introduced by Butzer and Scherer for Banach spaces in 1968 [5] and later by Brudnij and Krugljak [4]. These concepts find its best application in a paper by Pietsch [12], where he defined approximation spaces, proved embedding, reiteration and representation results and established connection to interpolation spaces. Let X be a Banach space and {An} be a sequence of subsets of X satisfying: (1) A1 ⊆ A2 ⊆ · · · ⊆ X (2) λAn ⊆ An for all scalars λ and n = 1, 2, . . . . (3) Am + An ⊆ Am+n for m, n = 1, 2, . . . . For example, for 1 ≤ p < ∞ if we consider the space X = Lp[0, 1], then the collection of sets {An} = {Lp+ 1 Pietsch's approximation spaces X ρ considering the n-th approximation number αn(f, X), where n } form an approximation scheme like above. µ (0 < ρ < ∞, 0 < µ ≤ ∞) is defined by and αn(f, X) := inf{f − a : a ∈ An−1} X ρ µ = {f ∈ X : {nρ− 1 µ αn(f, X)} ∈ ℓµ}. In the same paper [12], embeddings, composition and commutation as well as representation interpolation of such spaces are studied and applications to the distribution of Fourier coefficients and eigenvalues of integral operators are given. In the following we consider for each n ∈ N a family of subsets Qn of X satisfying the very same three conditions stated above. For example for Qn could be the set of all at most n-dimensional subspaces of any Banach space X, or if our Banach space X = L(E), namely the set of all bounded linear operators on another Banach space E, then we can take Qn = Nn(E) the set of all n-nuclear maps on E. Compactness relative to an approximation scheme for bounded sets and linear operators can be studied by using Kolmogorov diameters as follows. Let D ⊂ X be a bounded subset and UX denote the closed unit ball of X. Suppose Q = (Qn(X)n∈N) be an approximation scheme on X, then the nth Kolmogorov 6 A. AKSOY diameter of D with respect to this scheme Q is denoted by δn(D, Q) and defined as δn(D, Q) = inf{r > 0 : D ⊂ rUX + A for some A ∈ Qn(X)}. Let Y be another Banach space and T ∈ L(Y, X), then the nth Kolmogorov diameter of T with respect to this scheme Q is denoted by δn(T, Q) and defined as Definition 3.1. We say D is Q-compact set if δn(T, Q) = δn(T (UX ), Q). and similarly T ∈ L(Y, X) is a Q-compact map, if lim n δn(D, Q) = 0 lim n δn(T, Q) = 0. The following example illustrates that not every Q-compact operator is com- pact. Example 3.2. Let {rn(t)} be the space spanned by the Rademacher functions. It can be seen from the Khinchin inequality [10]that ℓ2 ≈ {rn(t)} ⊂ Lp[0, 1] for all 1 ≤ p ≤ ∞. We define an approximation scheme An on Lp[0, 1] as follows: An = Lp+ 1 n . (3.1) (3.2) n n+1 ⊂ Lp+ 1 gives us An ⊂ An+1. for n = 1, 2, . . . , and it is easily seen that Lp+ 1 An + Am ⊂ An+m for n, m = 1, 2, . . . , and that λAn ⊂ An for all λ. Thus {An} is an approximation scheme. Next, we claim that for p ≥ 2 the projection P : Lp[0, 1] → Rp is a Q-compact map, but not compact, where Rp denotes the closure of the span of {rn(t)} in Lp[0, 1]. Lp P ↓ Rp i−→ L2 ↓ P2 ←− R2 j We know that for p ≥ 2, Lp[0, 1] ⊂ L2[0, 1] and R2 is a closed subspace of L2[0, 1] and P = j ◦ P2 ◦ i where i, j are isomorphisms shown in the above figure. P is not a compact operator, because dimRp = ∞, on the other hand it is a Q-compact operator because, if we let URp, ULp denote the closed unit balls of Rp and Lp respectively, it is easily seen that P (ULp) ⊂ kP kURp. But URp ⊂ CUR where C is a constant follows from the Khinchin inequality. Therefore, P + 1 n P (ULp) ⊂ Lp+ 1 n , which gives δn(P, Q) → 0. Next we give a characterization of Q-compact sets as subsets of the closed convex hull of certain uniform null-sequences. TERZIOGLU THM 7 Definition 3.3. Suppose X is a Banach space with an approximation scheme Qn. A sequence {xn,k}k in X is called an order c0-sequence if (1) ∀n = 1, 2, . . . there exists An ∈ Qn and a sequence {xn,k}k ⊂ An (2) xn,k → 0 as n → ∞ uniformly in k. Theorem 3.4. Let X be a Banach space with an approximation scheme with sets An ∈ Qn satisfy the condition λAn ⊂ An for λ ≤ 1. A bounded subset D of X is Q-compact if and only if there is an order c0-sequence {xn,k}k ⊂ An such that D ⊂( ∞Xn=1 λnxn,k(n) : xn,k(n) ∈ (xn,k) λn ≤ 1) . ∞Xn=1 Proof of the above theorem can be obtained from the one given for p-Banach spaces in [2]. Clearly this is an analogue of Grothendieck's theorem given above in Theorem 1.3 for Q-compact sets. 4. Terzioglu's Theorem for Q-compact Maps Terzioglu's characterization of compact maps relies on both Grothendieck's and Schauder's theorems. Above Theorem 3.2 is Grothendieck's theorem for Q- compact sets, therefore we turn our attention to the relationship between T being Q-compact and its transpose T ∗ being Q-compact. The relationship between the approximation numbers of T and T ∗ was studied by several authors, it is shown in [8] that for T ∈ L(X), we have dist(T, F ) ≤ 3 dist(T ∗, F ∗) where F and F ∗ denote the class of all finite rank operators on X and X ∗ respec- tively. Central to the proof of such result is the assumption of local reflexivity possessed by all Banach spaces, (see [10]). It is not hard to show that if we assume that our space X with approximation scheme Qn satisfies slight modification of this property, called extended local reflexivity principle, then we have αn(T, Q) ≤ 3 αn(T ∗, Q∗). Where, by αn(T, Q) we mean the nth approximation number defined as αn(T, Q) = inf {T − B : B ∈ L(X), B(X) ∈ Qn(X)} . However, we do not have a proof of the Schauder's theorem for Q-compact maps. In the following we present a result analogous to Terzioglu's Theorem for Q-compact maps under the assumption that both T and T ∗ are Q-compact. Theorem 4.1. Let E and F be Banach spaces , T ∈ L(E, F ) and assume that both T and T ∗ are Q-compact maps. Then there exists sequence {un,k} ∈ Qn with un,k → 0 for n → ∞ for all k, such that the inequality T x ≤ sup < un,k(n), x > holds for every x ∈ E. Here Qn is a "special" class of subsets of E∗ with the property that un,k(n) ∈ {un,k}. 8 A. AKSOY Proof. Since T ∗ : F ∗ → E∗ is Q-compact, thus by the Theorem 3.4, T ∗(UF ∗) is a Q-compact set, thus there exists a sequence {un,k}k ⊂ An ∈ Qn such that un,k → 0 as n → ∞ uniformly in k and T ∗(UF ∗) ⊂( ∞Xn=1 Then for each x ∈ E, we have: λnun,k(n) : un,k(n) ∈ (un,k) λn ≤ 1) . ∞Xn=1 λnun,k(n), x > ∞Xn=1 T x = sup v∈UF ∗ < v, T x > = sup v∈UF ∗ < T ∗v, x > = sup n < and thus T x ≤ ∞Xn=1 λn sup n < un,k(n), x > ≤ sup n < un,k(n), x > . (cid:3) Remark 4.2. We say a map T ∈ L(E, F ) is a Q-compact map, if limn δn(T, Q) = 0. To obtain "Schauder's Theorem" for Q-compact maps, one seeks a relationship between δn(T ) and δn(T ∗). K. Astala in [3] proved that under the assumption that the Banach space E has the lifting property and the Banach space F has the extension property, for a map T ∈ L(E, F ), one has γ(T ) = γ(T ∗), where γ(T ) denotes the measure of non-compactness. Since lim n→∞ δn(T ) = γ(T ), by imposing extension and lifting properties on E and F respectively and keeping tract of approximation schemes on these spaces one might obtain Schauder's type of a theorem in this special case. References 1. A. G. Aksoy, Q-Compact sets and Q-compact maps, Math. Japon. 36 (1991), no. 1, 1-7. 2. A. G. Aksoy and M. Nakamura, The approximation numbers γn(T ) and Q-compactness, Math. Japon. 31 (1986), no. 6, 827-840. 3. K. Astala, On measures of non-compactness and ideal variations in Banach spaces, Ann. Acad. Sci. Fenn. Ser. AI Math. Dissertations 29, (1980), 1-42. 4. Y. Brudnij and N. Krugljak, On a family of approximation spaces, Investigation in function theory of several real variables, Yaroslavl State Univ., Yaroslavl, (1978), 15-42. 5. P.L. Butzer and K. Scherer, Approximationsprozesse und Interpolations methoden Bili- ographisches Inst. Mannheim, 1968. 6. J. Diestel, Sequences and Series in Banach Spaces, Spriger-Verlag, Berlin, Heidelberg, New York, 1984. 7. P. Enflo, A counter example to the approximation property in Banach spaces. Acta Math 130, (1973), 309-317. 8. C. V. Hutton On approximation numbers and its adjoint. Math. Ann. 210 (1974), 277-280. 9. J. Lindenstrauss, Extension of compact operators, Mem. Amer. Math. Soc. No. 48 (1964), 112 pp. 10. J. Lindenstrauss and L. Tzafriri, Classical Banach Spaces I, Sequence Spaces Springer- Verlag, Berlin, Heidelberg, New York, 1977. TERZIOGLU THM 9 11. A. Persson and P. Pietsch, p-nucleare une p-integrale Abbildungen in Banachraumen, Studia Math. 33, (1969), 19-62. 12. P. Pietsch Approximation spaces, J. Approx. Theory, 32 (1981), no. 2, 115-134. 13. T. Terzioglu, A characterization of compact linear mappings, Arch. Math. 22 (1971), 76-78. 14. T. Terzioglu, Remarks on Compact and infinite-nuclear mappings Mathematica Balkanica, 2, (1972), 251-255. 15. K. Yosida, Functional analysis Spriger-Verlag, Berlin, Heidelberg, New York, 1965. ∗Department of Mathematics, Claremont McKenna College, 850 Columbia Av- enue, Claremont, CA 91711, USA. E-mail address: [email protected]
1201.2117
1
1201
2012-01-10T17:19:04
Traceability of positive integral operators in the absence of a metric
[ "math.FA" ]
We investigate the traceability of positive integral operators on $L^2(X,\mu)$ when $X$ is a Hausdorff locally compact second countable space and $\mu$ is a non-degenerate, $\sigma$-finite and locally finite Borel measure. This setting includes other cases proved in the literature, for instance the one in which $X$ is a compact metric space and $\mu$ is a special finite measure. The results apply to spheres, tori and other relevant subsets of the usual space $\mathbb{R}^m$.
math.FA
math
Traceability of positive integral operators in the absence of a metric Mario H. Castro, Valdir A. Menegatto and Ana P. Peron 12 Abstract We investigate the traceability of positive integral operators on L2(X, µ) when X is a Hausdorff locally compact second countable space and µ is a non-degenerate, σ-finite and locally finite Borel measure. This setting includes other cases proved in the literature, for instance the one in which X is a compact metric space and µ is a special finite measure. The results apply to spheres, tori and other relevant subsets of the usual space Rm. Mathematics Subject Classification (2010). 15A18, 60G46. 42A82, 45P05, 47B34, 47B65, Keywords. integral operators, positive definite kernels, trace-class, averaging, mar- tingales. 1 Introduction Let X be a Hausdorff locally compact and second countable topological space endowed with a non-degenerate, σ-finite and locally finite Borel measure µ. In this paper, we shall investigate the traceability of integral operators K : L2(X, µ) → L2(X, µ) generated by a suitable kernel K : X × X → C from L2(X × X, µ × µ). The title of the paper refers to the fact that the space X carries a topological structure rather than a metric one. The setting just described allows the space L2(X, µ) to have a countable complete orthonormal subset ([7, p.92]) while the operator K, which is given by the formula K(f ) :=ZX K(·, y)f (y) dµ(y), f ∈ L2(X, µ), (1.1) becomes compact. As so, the spectral theorem for compact operators is applicable and K can be represented in the form K(f ) = ∞ Xn=1 λnhf, fnifn, f ∈ L2(X, µ), (1.2) in which {λn} is a sequence of real numbers (possibly finite) converging to 0 and {fn} is a complete orthonormal sequence in L2(X, µ). The symbol h·, ·i will stand for the usual inner product of L2(X, µ). 1First author partially supported by CAPES 2 Second author partially supported by FAPESP # 2010/19734-6 1 The basic requirement on the kernel K will be its positive definiteness. A kernel K from L2(X × X, µ × µ) is L2(X, µ)-positive definite when the corresponding integral operator K, is positive: hK(f ), f i ≥ 0, f ∈ L2(X, µ). (1.3) Fubini's theorem is all that is need in order to show that a L2(X, µ)-positive definite kernel is hermitian µ × µ-a.e.. As so, the integral operator K is automatically self-adjoint with respect to h·, ·i. In particular, the sequence {λn} mentioned in the previous paragraph needs to be entirely composed of nonnegative numbers. In the present paper, we shall assume they are listed in a decreasing order, with repetitions to account for multiplicities. Under the conditions established above, the specific aim of this paper is to establish additional conditions on K in order that K be trace-class, that is, hK∗K(f ), f i1/2 < ∞ Xf ∈B (1.4) for every orthonormal basis B of L2(X, µ). In the formula above, K∗ is the adjoint of K. We refer the reader to [5, 10, 11] for more information on trace-class operators. The main result in this paper can be seen as a generalization of another one originally proved in [11] for the case X = [a, b]. The proof there used in a key manner the so-called Steklov's smoothing operator to construct an averaging process to generate a convenient approximation to K. The upgrade to the case in which X is a subspace of Rn was discussed in [8] and references therein. By assuming that the Lebesgue measure of nonempty inter- sections of X with open balls of Rn was positive and using auxiliary approximation integral operators generated by an averaging process constructed via the Hardy-Littlewood theory, the main result in [8] described necessary and sufficient conditions for the traceability of the integral operator, under the assumption of positive definiteness of the kernel. The process used in [8] and other references as well provides a way to deal with the generating kernel on the diagonal of X × X and it is convenient when the kernel is not continuous. Despite using a similar average process, another achievement in the present paper is the inclusion of a setting in which the measure does not need to be finite. Since our spaces are no longer metric, the Hardy-Littlewood theory in the average arguments need to be replaced or adapted. We will use techniques involving the construction of auxiliary integral operators based on martingales constructed from special partitions of X, following very closely the development of Brislawn in [2]. A similar construction have appeared in [6] in an attempt to generalize Brislawn results to Lp spaces. The main difference between the construction to be delineated here and those in [2] and [6] is that, in the present one, we need to guarantee that the elements in the partitions belong to the topology of X. This is the exact point where the assumption of local compactness will play an important role. For the sake of completeness we mention references [1, 14] where other characterizations for traceability were obtained. An outline of the paper is as follows. Section 2 contains the basic information on martin- gales used in the paper, along with the key construction we will need in order to introduce 2 approximating auxiliary operators in Section 3. There, the main technical results are estab- lished and proved. Section 4 contains the main results of the paper, including a convenient equivalence for traceability. 2 A special martingale This section contains several results involving a special martingale on X. Some of them are just refined versions of results described in Section 2 of [2]. However, the reader is advised that the basic references we used for the concepts and results either quoted or used here are [4, 15]. Let (X, M, σ) denote a σ-finite measure space and F a sub-σ-algebra of M for which (X, F , σ) is a σ-finite measure space too. If f : X → C is M-measurable, Radon-Nikodyn's theorem asserts that we can find a unique F -measurable function g : X → C so that ZA f dσ =ZA g dσ, A ∈ F . (2.1) The function g is called the conditional expectation of f relative to F and is written g = E(f F ). If {Fn} is a family of sub-σ-algebras of M, a sequence {fn} of M-measurable functions on X is a martingale if every fn is Fn-measurable and E(fnFm) = fm, m < n. Next, we remind the reader about the basic setting we are assuming in the paper: X is a Hausdorff, locally compact and second countable topological space endowed with a non-degenerate, locally finite and σ-finite Borel measure µ. In addition to that, we will write BX to denote the Borel σ-algebra of X. Invoking the first countability axiom, we may infer that every point of X possesses an open neighborhood. Since X is Hausdorff and locally compact, these neighborhoods can be assumed to be the interior of a compact set. Thus, due to the local finiteness of (X, µ), we can assume, in addition, that the open neighborhoods of elements of X have finite measure. We intend to construct a special sequence of partitions of X from an open covering {Ax}x∈X of it, composed of neighborhoods of the type just described, and use them to define a particular martingale. If such a covering has been fixed, Lindeloff's theorem ([13, p.191]) implies that we can extract from it a countable sub-collection {An}, still covering X. Such sub-collection can be used in the construction of a first stage partition P0 of X, following these steps: the first two elements in the partition are A0 and its frontier ∂A0. Observing that {An\A0} is an open and countable covering of X\A0, we pick A1\A0 and ∂A1\A0 to include in the partition. The family {An\A0 ∪ A1} is an open and countable covering of X\A0 ∪ A1. We proceed, including its elements A2\A0 ∪ A1 and ∂A2\A0 ∪ A1 in the partition. Proceeding inductively, we complete the construction of P0, which is countable and entirely composed of Borel sets of finite measure. Since Theorem 7.8 in [9] implies that µ is regular, all the sets of the form ∂An\A0 ∪ · · · ∪ An−1 in P0 have measure zero. In the next step, we construct a sequence {Pn} of partitions of X from P0, using as we can, a countable basis {Un} for the topology of X. For n = 0, 1, . . ., we put Pn+1 = {Un ∩ A : A ∈ Pn} ∪ {(X \ Un) ∩ A : A ∈ Pn} ∪ {∂Un ∩ A : A ∈ Pn}. 3 Clearly, Pn+1 refines Pn and the sequence {Fn} of the corresponding σ-algebras generated by those partitions increases to BX . In addition, every (X, Fn, µ) is σ-finite. It is easy to see that for each x ∈ X and each positive n, there exists a unique set On(x) ∈ Pn such that x ∈ On(x). We denote by N the subset of X containing all x ∈ X for which µ(Om(x)) = 0, for some m ≥ 0. Since the sequence {On(x)} is telescoping, the equality µ(Om(x)) = 0 implies µ(On(x)) = 0, n ≥ m. Being each Pn countable, it is easily seen that µ(N) = 0. The very same arguments used in [15, p.89] show that for every x ∈ X \ N and every positive n, the conditional expectation En(f ) of f relative to Fn is given by the formula En(f )(x) = 1 µ(On(x))ZOn(x) f dµ. (2.2) The sequence {En(f )} defines a martingale generated by just one (measurable) function, the martingale associated with f . Examples related to constructions similar to the one above can be found in [15, p.88]. The section will be completed with a list of results involving the previous formula and the maximal function M f of the martingale associated with f . Such martingale is defined by the formula M f (x) := sup{En(f )(x) : n = 1, 2, . . .}, x ∈ X. (2.3) Since the results are quite general and are not attached to the particular setting introduced above, we will include sketches of the proofs for the convenience of the reader. A classical result concerning the maximal function ([15, p.91]) implies that if p ∈ (0, ∞) then kM f kp ≤ cpkf kp, f ∈ Lp(X, µ), (2.4) where cp is a constant depending on p only and k · kp denotes the usual norm of Lp(X, µ). As for the conditional expectation, it transforms convergence in the mean into convergence µ-a.e. Another basic result ([4, p.53] and [2, p.232]), commonly called Doob's martingale convergence theorem, states that En(f ) converges to f µ-a.e., as long as f ∈ Lp(X, µ) and p ∈ [1, ∞]. Moving forward, the inequalities En(f )(x) ≤ M f (x), x ∈ X \ N, n ≥ 1, and f (x) ≤ f (x) − En(f )(x) + M f (x), x ∈ X \ N, n ≥ 1, (2.5) (2.6) are easily deducted. Combining the last one with Doob's martingale convergence theorem, we are led to the inequality f ≤ M f , µ-a.e.. As for the conditional expectation, we have the following result found in [15, p.90]: if p ∈ [1, ∞] and f ∈ Lp(X, µ) then kEn(f )kp ≤ kf kp, n ≥ 1. As a consequence, the following theorem holds. Theorem 2.1 If p ∈ [1, ∞] then the linear map En : Lp(X, µ) → Lp(X, µ) is bounded. If p = 2, then the previous map is a self-adjoint operator. 4 We close the section with a result for convergence in the mean of the conditional expec- tation. Theorem 2.2 If f ∈ L2(X, µ) then Enf converges to f in the mean. Proof. Now, inequality (2.5) leads to If gn := f − En(f )2, n ≥ 1, the previous theorem yields that {gn} ⊂ L1(X, µ). gn(x) ≤ 2(f (x)2 + En(f )(x)2) ≤ 4M f (x)2, x ∈ X \ N, n ≥ 1. (2.7) Clearly, M f ∈ L2(X, µ) while Doob's convergence theorem gives us gn → 0 µ-a.e.. The dominated convergence theorem connects the final arguments. 3 Approximating kernels This section is entirely composed of technical results involving a family of operators con- structed from the martingale defined in Section 2. Under the notation in Section 2, Theorem 7.20 in [9] informs that the product measure µ × µ is a regular Borel measure on X × X and the sequence {Pn × Pn} of partitions of X × X increases to the Borel σ-algebra BX×X of (X × X, µ × µ). In particular, if K ∈ L1 loc(X ×X, µ×µ), the conditional expectation with respect to the σ-algebra generated by the partition Pn × Pn of X × X can be defined by the formula En(K)(u, v) := 1 σ(On(u))σ(On(v))ZOn(u)ZOn(v) K(x, y) dµ(y)dµ(x). (3.1) Lemma 3.1 below provides information about a limit property regarding the open sets On(x) previously defined. We will use the symbol χA to denote the characteristic function of the subset A of X. We remind the reader that given x ∈ X and n ≥ 1, the construction introduced in the previous section shows that there exists a unique On(x) ⊂ Pn so that x ∈ On(x). Lemma 3.1 If x ∈ X and n ≥ 1 then lim u→u0 χOn(u)(x) = χOn(u0)(x), u0 ∈ X \ N. (3.2) Proof. Fix x ∈ X and n ≥ 1. If u ∈ X then x ∈ On(u) if and only if u ∈ On(x). Since χOn(x)(u) = χOn(u)(x), we can write χOn(u)(x) − χOn(u0)(x) = χOn(x)(u) − χOn(x)(u0), u0 ∈ X. (3.3) Next, if u0 ∈ X \ N, the fact that On(u0) is open, leaves us with two cases: if x ∈ On(u0) then u0 ∈ On(x) and, at the limit, we can assume u ∈ On(u0) = On(x) so that lim u→u0 χOn(x)(u) − χOn(x)(u0) = 1 − 1 = 0. (3.4) 5 If x 6∈ On(u0) then u0 6∈ On(x), and assuming u ∈ On(u0) as we can, we conclude that lim u→u0 χOn(x)(u) − χOn(x)(u0) = 0 − 0 = 0. (3.5) The proof is complete. It is now reasonable that the following result holds. Lemma 3.2 If u0 ∈ X \ N then limu→u0 µ(On(u)) = µ(On(u0)), n = 1, 2, . . . . Proof. Since it follows that µ(On(u)) =ZX µ(On(u)) − µ(On(u0)) ≤ZX χOn(u)(x) dµ(x), u ∈ X, (3.6) χOn(u)(x) − χOn(u0)(x) dµ(x), u ∈ X. (3.7) As so, the assertion of the lemma will be proved if we can show that lim u→u0ZX χOn(u)(x) − χOn(u0)(x) dµ(x) = 0, u0 ∈ X \ N. (3.8) Hence, in view of the previous lemma, it suffices to show that the integral and the limit in the previous equation commute. The family {gu} defined by gu(x) = χOn(u)(x) − χOn(u0)(x), u, x ∈ X, (3.9) and the function g = χOn(u0) belong to L1(X, µ). Since gu ≤ g, µ-a.e., when u → u0, the desired commuting property follows from the dominated convergence theorem. We now turn to kernels of the form Dn(u, x) = 1 µ(On(u)) χOn(u)(x), u, x ∈ X, n = 1, 2, . . . . (3.10) and the corresponding integral operators Dn generated by Dn. For use ahead, we mention the immediate formula En(χOn(u) f ) = Dn(f ), u ∈ X \ N, f ∈ L2(X, µ). (3.11) Initially, we will use the above kernels to prove the following result. Theorem 3.3 If K ∈ L2(X × X, µ × µ) and n ≥ 1 then En(K) is continuous µ × µ-a.e.. Proof. u0, v0 ∈ X \ N. It is not hard to see that It suffices to show that En(K) is continuous in the set (X \ N) × (X \ N). Let En(K)(u, v) =ZXZX Dn(u, x)K(x, y)Dn(v, y) dµ(y)dµ(x), u, v ∈ X, (3.12) 6 and that we can use Lemma 3.1 and Lemma 3.2 to deduce that lim (u,v)→(u0,v0) Dn(u, x)K(x, y)Dn(v, y) = Dn(u0, x)K(x, y)Dn(v0, y), (3.13) for x, y ∈ X a.e.. If (u, v) ∈ On(u0) × On(v0), we have Dn(u, x)K(x, y)Dn(v, y) ≤ 1 µ(On(u0))µ(On(v0)) K(x, y), (3.14) for x, y ∈ X a.e.. So, the continuity at (u0, v0) now follows from the dominated convergence theorem. Next, we will state and prove a list of technical results that will lead to the following conclusion: DnKDn coincides with the integral operator generated by En(K). The sequence of partitions {Pn} was constructed in such a way that each one of them has the following feature: every element of {Ai} is a subset of at most finitely many On(x). That been said, if n and i are fixed, we can write (3.15) Ai ⊂  m(n,i) [j=1 On(xj) [ N(n, i), in which µ(N(n, i)) = 0 and 0 < µ(On(xj)) < ∞, j = 1, 2, . . . , m(n, i). The set N(n, i) is nothing but the union of all elements of Pn for which the intersection with Ai has measure zero. In the next results, we will deal with a continuous function f : X → C with compact support Xf . Since Xf can be covered by finitely many Ai, after re-ordering if necessary, we can find an index l so that l Xf ⊂ [k=1  m(n,k) [j=1 On(xj) [ l [k=1 with µ(∪l k=1 N(n, k)) = 0. In that case, we will write Yf = l m(n,k) [k=1 [j=1 On(xj). N(n, k)! , (3.16) (3.17) Lemma 3.4 Let f : X → C be a function with compact support Xf and K an element of L1 loc(X × X, µ × µ). Then ZX×XZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(z) d(µ × µ)(x, y) =ZXZX×X Dn(u, x)K(x, y)Dn(y, z)f (z) d(µ × µ)(x, y) dµ(z). 7 Proof. Pick M > 0 so that f (x) ≤ M, x ∈ X. We have ZX×XZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(z) d(µ × µ)(x, y) ≤ M µ(On(u))ZOn(u)×X K(x, y) µ(On(y))ZXf χOn(y)(z) dµ(z) d(µ × µ)(x, y). If y 6∈ Yf then y 6∈ Xf and, consequently, On(y) ∩ Xf = ∅. Thus χOn(y) = 0 in Xf and we can take the above integral on On(u) × Yf . Now, if u ∈ X \ N, the local integrability of K implies that ZOn(u)×Yf K(x, y) µ(On(y))ZXf χOn(y)(z) dµ(z) d(µ × µ)(x, y) =ZOn(u)×Yf ≤ZOn(u)×Yf K(x, y) µ(On(y)) µ(On(y) ∩ Xf ) d(µ × µ)(x, y) K(x, y) d(µ × µ)(x, y) < ∞. Fubini's theorem ([12, p.386]) completes the proof. Lemma 3.5 Let f : X → C be a continuous function with compact support Xf and K an element of L1 loc(X × X, µ × µ). If u ∈ X \ N then ZOn(u)×Yf K(x, y)Dn(y, z) d(µ × µ)(x, y) =ZOn(u)ZYf =ZYf ZOn(u) K(x, y)Dn(y, z) dµ(y) dµ(x) K(x, y)Dn(y, z) dµ(x) dµ(y), with Yf as defined in (3.17). Proof. If u ∈ X \ N and z ∈ X then ZOn(u)ZYf K(x, y)Dn(y, z) dµ(y) dµ(x) =ZOn(u)ZYf (cid:12)(cid:12)(cid:12)(cid:12) ≤ZOn(u)ZYf K(x, y) µ(On(y)) K(x, y) dµ(y) dµ(x) χOn(y)(z) µ(On(y))(cid:12)(cid:12)(cid:12)(cid:12) dµ(y) dµ(x). Introducing the decomposition (3.17) in the last expression above and recalling the unique- ness property of the On(x), we deduce that ZOn(u)ZYf K(x, y) µ(On(y)) K(x, y) µ(On(y)) dµ(y) dµ(x) m(n,k) l l Xj=1 ZOn(yj ) dµ(y) dµ(x) =ZOn(u) =ZOn(u) Xj=1 1 ≤ j ≤ m(n, k) (cid:26) Xk=1 Xk=1 max m(n,k) 1 ≤ k ≤ l ≤ K(x, y) dµ(y) dµ(x) 1 µ(On(yj))ZOn(yj ) µ(On(yj))(cid:27) kKkL1(On(u)×Yf ) < ∞. 1 8 Once again, Fubini's theorem leads to the concluding statement. Lemma 3.6 Let f : X → C be a continuous function with compact support Xf and K an element of L1 loc(X × X, µ × µ). Then ZXZXZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(z) dµ(y) dµ(x) =ZX×XZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(z) d(µ × µ)(y, x). Proof. If u ∈ X \ N and M > 0 is a bound for f in X then it is easily seen that ZXZX ZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(z) dµ(y) dµ(x) ≤ ≤ M µ(On(u))ZOn(u)ZYf µ(On(u))ZOn(u)ZYf M K(x, y) µ(On(y)) µ(On(y) ∩ Xf ) dµ(x) dµ(y) K(x, y) dµ(x) dµ(y) < ∞. So, the result follows from Fubini's theorem once again. The proof of the next lemma is analogous and will be omitted. Lemma 3.7 Let f : X → C be a continuous function with compact support and K and element in L1 loc(X × X, µ × µ). Then ZXZXZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(z) dµ(y) dµ(x) =ZXZXZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(x) dµ(y) dµ(z). Recalling that if x, y ∈ X then z ∈ On(y) if and only if y ∈ On(z), the following lemma becomes obvious. Lemma 3.8 If n ≥ 1 then Dn(y, z) = Dn(z, y), y, z ∈ X \ N. Below, E n K will denote the integral operator generated by En(K). Theorem 3.9 If K ∈ L2(X × X, µ × µ) then DnKDn(f ) = E n K(f ), f ∈ L2(X, µ). Proof. Clearly L2(X × X, µ × µ) ⊂ L1 f : X → C is continuous with compact support then the previous lemmas imply that loc(X × X, µ × µ). If K ∈ L2(X × X, µ × µ) and DnKDn(f )(u) =ZXZXZX Dn(u, x)K(x, y)Dn(y, z)f (z) dµ(z) dµ(y) dµ(x) Dn(u, x)K(x, y)Dn(y, z)f (z)dµ(z)dµ(y)dµ(x) =ZOn(u)ZYf ZXf =ZXZXZX = E n K(f )(u), u ∈ X \ N. 9 Dn(u, x)K(x, y)Dn(z, y)f (z) dµ(x) dµ(y) dµ(z) Hence, the result in the statement of the theorem follows from the equality µ(X \ N) = 0 and from a basis approximation theorem from measure theory ([12, p.197]). The last result of the section refers to the positive definiteness of En(K). Theorem 3.10 If K is L2(X, µ)-positive definite then so is En(K). If K is L2(X, µ)-positive definite then both, K and En(K), belong to the space Proof. L2(X × X, µ × µ). On the other hand, Theorem 2.1 and (3.11) imply that Dn(L2(X, µ)) ⊂ L2(X, µ). Thus, an application of Theorem 3.9 leads to hE n K(f ), f i2 = hKDn(f ), Dn(f )i2 ≥ 0, f ∈ L2(X, µ). (3.18) The proof is complete. 4 Traceability This section contains the main results of the paper. They can be interpreted as generaliza- tions of results obtained in [8] and other references quoted here. The traceability results de- scribed here will be obtained via several known results on trace-class operators and singular values of operators. We will quote some of them and just mention others. The construction developed in Section 2 reveals that the diagonal of X is, up to a set of measure zero, a subset of (X \ N) × (X \ N). This remark justify why some of the integrals appearing below are not identically zero. Given K ∈ L2(X × X, µ × µ), we will consider E n K acting like an operator on L2(X, µ). All other operators mentioned here are to be understood acting in the same way. The following lemma is an adapted to our purposes version of Theorem 4.1 in [3]. Lemma 4.1 If K is a continuous (µ × µ-a.e.) L2(X, µ)-positive definite kernel and x ∈ X → K(x, x) is integrable then K is trace-class and tr (K) =ZX K(x, x) dµ(x). Lemma 4.2 Let K be L2(X, µ)-positive definite. If lim sup n→∞ ZX En(K)(x, x) dµ(x) < ∞, then lim supn→∞ tr (E n K) < ∞. (4.1) (4.2) Proof. If (4.2) holds then there exists n0 ∈ N such that x ∈ X → En(K)(x, x) is integrable for n ≥ n0. Theorem 3.10 implies that En(K) is L2(X, µ)-positive definite while Theorem 3.3 shows that En(K) is continuous µ × µ-a.e.. Applying Lemma 4.1 we see that tr (E n K) =ZX En(K)(x, x) dµ(x), n ≥ n0. (4.3) 10 The result follows. Next, we recall some facts involving singular values of an operator. If T is a compact operator on a Hilbert space, a singular value of T is an eigenvalue of (T ∗T )1/2. We shall enumerate the nonzero singular values of T in decreasing order, taking multiplicities into account: s1(T ) ≥ s2(T ) ≥ . . .. If the rank ρ of (T ∗T )1/2 is finite, obviously sj(T ) = 0, j ≥ ρ + 1. If the eigenvalues of T are ordered like l1(T ) ≥ l2(T ) ≥ . . ., then a classical result from operator theory states that sj(T ) = lj(T ), j = 1, 2, . . ., as long as T is either hermitian or normal. If S is another compact operator of same type as T , and assuming the same ordering on the singular values of S, the following inequality holds: sn(T ) − sn(S) ≤ kT − Sk, n = 1, 2 . . .. All of these results can be found with proofs in [10, 11]. In Theorem 4.4, a complement of Lemma 4.2, we also use the following nontrivial result on convergence of operators ([8]). Lemma 4.3 Let {Tn} be a countable set of bounded linear operators on a Hilbert space H such that limn→∞ kTn(f ) − f kH = 0, f ∈ H. If every Tn is self-adjoint and T is a bounded compact operator on H then limn→∞ kTnT Tn − T k = 0. Theorem 4.4 Let K be L2(X, µ)-positive definite. If then K is trace-class. lim sup n→∞ ZX En(K)(x, x) dµ(x) < ∞, (4.4) Proof. Since {sj(E n K)} ⊂ (0, ∞), it is quite clear that sj(E n K) ≤ tr (E n K), k = 1, 2, . . . . (4.5) k Xj=1 Theorem 3.9 and the inequality mentioned before Lemma 4.3 imply that sj(E n K) − sj(K) ≤ kE n K − Kk = kDnKDn − Kk, j = 1, 2, . . . . (4.6) Since each Dn is self-adjoint, K is compact and lim n→∞ kDn(f ) − f k2 = 0, f ∈ L2(X, µ), we are authorized to apply Lemma 4.3 to conclude, from (4.6), that lim n→∞ sj(E n K) = sj(K), j = 1, 2, . . . . (4.7) (4.8) It is now clear that k Xj=1 sj(K) = lim sup n→∞ k Xj=1 and that concludes the proof. sj(E n K) ≤ lim sup n→∞ tr (E n K), k = 1, 2, . . . , (4.9) In order to deal with the converse of the previous result, we will need the following result ([10, p.51]): if S1, S2 and T are bounded linear operators on a Hilbert space and T is compact then so is the composition S1T S2 and sj(S1T S2) ≤ kS1ksj(T )kS2k, j = 1, 2, . . . . 11 Lemma 4.5 Let p ∈ [1, ∞) and K ∈ Lp(X × X, µ × µ). If x ∈ X → K(x, x) is integrable and µ(X) < ∞ then there is a positive integer n0 for which x ∈ X → En(K)(x, x) is integrable when n ≥ n0. Proof. Since En(K)(u, u) ≤ En(K)(u, u) − K(u, u) + K(u, u), u ∈ X \ N, (4.10) we can use Doob's convergence theorem to select a positive integer n0 so that En(K)(u, u) ≤ 1 + K(u, u), u ∈ X \ N, n ≥ n0. (4.11) Our assumptions on X and x ∈ X → K(x, x) imply the result. Theorem 4.6 Let K be L2(X, µ)-positive definite. If x ∈ X → K(x, x) is integrable and µ(X) < ∞ then there is n0 ∈ N so that E n K ∈ B1(L2(X)) and tr (E n K) =ZX En(K)(x, x) dµ(x), n ≥ n0. (4.12) Proof. The previous lemma reveals that x ∈ X → En(K)(x, x) is integrable for n large. As so, the result follows from Theorem 3.3 and Lemma 4.1. Theorem 4.7 Let K ∈ L2(X × X, µ × µ). If K is trace-class then so is every E n number tr (K) is an upper bound for the sequence {tr (E n K. The K)}. Proof. Assume K is trace-class. Since each Dn is bounded, Theorem 3.10 and the comments preceding Lemma 4.5 imply that sj(E n K) = sj(DnKDn) ≤ kDnksj(K)kDnk, n = 1, 2, . . . . Hence, ∞ and the two assertions of the lemma follow. sj(E n K) ≤ kDnk2 Xj=1 sj(K), ∞ Xj=1 The following result is very close to a converse of Theorem 4.4. Theorem 4.8 Let K ∈ L2(X × X, µ × µ). If K is trace-class then Proof. A basic inequality for the trace ([10, p.54]) implies that ∞ lim n→∞ tr (E n K) = tr (K). (4.13) (4.14) tr (E n K) − tr (K) ≤ sj(E n K − K), n = 1, 2, . . . , (4.15) Xj=1 as long as K is trace-class. On the other hand, since (see [10, p.89]) lim n→∞ Theorem 3.9 completes the proof. ∞ Xj=1 sj(DnKDn − K) = 0, (4.16) Next, we move to a proof of the converse of Theorem 4.3 in the case when µ(X) < ∞. 12 Theorem 4.9 Let K be L2(X, µ)- positive definite. If K is trace-class and µ(X) < ∞ then lim n→∞ZX En(K)(x, x) dµ(x) < ∞. (4.17) Proof. Assume K is trace-class. Since the function x ∈ X → K(x, x) is integrable already, if µ(X) < ∞, we can use Theorem 4.6 to find a positive integer n0 such that tr(E n K) =ZX En(K)(x, x) dµ(x), n ≥ n0. (4.18) An application of Theorem 4.8 finishes the proof. At this point, it is very important to remind the reader that the results we have obtained includes the case in which X is either a sphere or a torus. Next, we intend to consider cases in which X has no finite measure. In order to handle that, we use the cover {Am} of X constructed before to define a sequence of subsets of X that increases to X. Precisely, defining Xj = ∪j m=1Am, j ≥ 1, we immediately have the following two properties: X = ∪∞ j=1Xj and if x ∈ X then there exists j0 ≥ 0 such that x ∈ Xj, j ≥ j0. Using the sequence just defined, we now take linear operators Pj : L2(X, µ) → L2(X, µ) defined by the formula Pj(f ) = f χXj , f ∈ L2(X, µ). They are self- adjoint and the uniform boundedness principle shows that the sequence {Pj} is bounded in the space of bounded linear operators on L2(X, µ). Also, the dominated convergence theorem implies that {Pj} converges pointwise to the identity operator on L2(X, µ). The following technical lemma contains a critical information on the sequence {Pj}. Lemma 4.10 If T : L2(X, µ) → L2(X, µ) is trace-class then each PjT Pj is so and the limit formula limj→∞ tr (PjT Pj) = tr (T ) holds. Proof. The first assertion is a consequence of the remark preceding Lemma 4.5. As for the other, it follows from Theorem 11.3 in [10] . The converse of Theorem 4.4 reads as follows. Theorem 4.11 Let K be L2(X, µ)-positive definite. If K is trace-class then the limit lim n→∞ZX En(K)(x, x) dµ(x) (4.19) exists and is finite. Proof. The proof requires the double-indexed operator Qn j given by the formula Qn j (f )(x) =ZXj En(K)(x, y)f (y) dµ(y), x ∈ Xj, f ∈ L2(Xj, µ). (4.20) 13 If f ∈ L2(Xj), let us write f to denote a function on X that coincides with f on Xj and is zero in X \ Xj. It is now clear that Qn ZXj j (f )(x)f (x) dµ(x) = ZXZX = ZXZX = ZX En(K)(x, y) f (y) f (x) dµ(y) dµ(x) En(K)(x, y) f (y) dµ(y) f(x) dµ(x) K( f )(x) f (x) dµ(x), E n f ∈ L2(Xj, µ). j is L2(Xj, µ)-positive Since K is L2(X, µ)-positive definite, Theorem 3.10 guarantees that Qn definite. Also, the fact that K is trace-class implies that x ∈ X → K(x, x) is integrable. Hence, due to Lemma 4.5, there exists n0 ≥ 0 such that x ∈ Xj → En(K)(x, x) whenever n ≥ n0. Recalling Theorem 3.3 and applying Lemma 4.1, we deduce that Qn j is trace-class and En(K)(x, x) dµ(x), (4.21) tr (Qn j ) =ZXj as long as n ≥ n0. Let us keep the previous condition on n in force. If Vj is the closed subspace L2(X, µ) encompassing the functions on X which are zero in X \ Xj and Rn : j Vj → Vj is the operator given by Rn j (f )(x) = χXj (x)ZX En(K)(x, y)χXj (y)f (y) dµ(y), (4.22) with x ∈ X, f ∈ Vj, then Rn previous lemma, j and Qn j possess the same eigenvalues. Having in mind the (PjE n KPj)(f )(x) =ZX En(K)(x, y)χXj ×Xj (x, y)f (y) dµ(y), (4.23) for x ∈ X, f ∈ L2(X), and we can conclude now that Rn eigenvalues. Therefore, j and PjE n KPj have the same tr (PjE n KPj) = tr (Rn j ) = tr (Qn j ) =ZXj En(K)(x, x) dµ(x). (4.24) The monotone convergence theorem leads to tr (E n K) =ZX En(K)(x, x) dµ(x). (4.25) Finally, (4.24) and the observation made before Theorem 4.3 lead to the assertion of the theorem. 14 References [1] Brislawn, C., Kernels of trace class operators. Proc. Amer. Math. Soc. 104 (1988), no. 4, 1181 -- 1190. [2] Brislawn, C., Traceable integral kernels on countably generated measure spaces. Pacific J. Math. 150 (1991), no. 2, 229 -- 240. [3] Castro, M. H.; Menegatto, V. A.; Peron, A. P., Integral operators generated by Mercer- like kernels: non-metric results, preprint. [4] Chatterji, S. D., Les martingales et leurs applications analytiques. (French) ´Ecole d'´Et´e de Probabilit´es: Processus Stochastiques (Saint Flour, 1971), pp. 27 -- 164. Lecture Notes in Math., Vol. 307, Springer, Berlin, 1973. [5] Conway, J. B., A course in operator theory. Graduate Studies in Mathematics, 21. American Mathematical Society, Providence, RI, 2000. [6] Delgado, J., The trace of nuclear operators on Lp(µ) for σ-finite Borel measures on second countable spaces. Integral Equations Operator Theory 68 (2010), no. 1, 61-74. [7] Doob, J. L., Measure theory. Graduate Texts in Mathematics, 143. Springer-Verlag, New York, 1994. [8] Ferreira, J. C.; Menegatto, V. A.; Oliveira, C. P., On the nuclearity of integral opera- tors, Positivity, 13 (2009), no. 3, 519-541. [9] Folland, G. B., Real analysis. Modern techniques and their applications. Second edi- tion. Pure and Applied Mathematics (New York). A Wiley-Interscience Publication. John Wiley & Sons, Inc., New York, 1999. [10] Gohberg, I.; Goldberg, S.; Krupnik, N., Traces and determinants of linear operators. Operator Theory: Advances and Applications, 116. Birkhauser Verlag, Basel, 2000. [11] Gohberg, I. C.; Kreın, M. G., Introduction to the theory of linear nonselfadjoint op- erators. Translated from the Russian by A. Feinstein. Translations of Mathematical Monographs, Vol. 18 American Mathematical Society, Providence, R.I. 1969. [12] Hewitt, E.; Stromberg, K., Real and abstract analysis. A modern treatment of the theory of functions of a real variable. Third printing. Graduate Texts in Mathematics, No. 25. Springer-Verlag, New York-Heidelberg, 1975. [13] Munkres, J. R., Topology. 2nd ed. (English) Upper Saddle River, NJ: Prentice Hall. (2000). [14] Sato, H., Nuclearity of a nonnegative definite integral kernel on a separable metric space. J. Theoret. Probab. 5 (1992), no. 2, 349 -- 353. 15 [15] Stein, E. M., Topics in harmonic analysis related to the Littlewood-Paley theory. An- nals of Mathematics Studies, No. 63 Princeton University Press, Princeton, N.J.; Uni- versity of Tokyo Press, Tokyo 1970. Departamento de Matem´atica, ICMC-USP - Sao Carlos, Caixa Postal 668, 13560-970 Sao Carlos SP, Brasil. [email protected] [email protected] FAMAT-UFU Caixa Postal 593 38400-902 Uberlandia-MG, Brasil [email protected] 16
1302.2483
1
1302
2013-02-11T14:16:06
On numerically hypercyclic operators
[ "math.FA", "math.DS" ]
According to Kim, Peris and Song, a continuous linear operator $T$ on a complex Banach space $X$ is called {\it numerically hypercyclic} if the numerical orbit $\{f(T^nx):n\in\N\}$ is dense in $\C$ for some $x\in X$ and $f\in X^*$ satisfying $\|x\|=\|f\|=f(x)=1$. They have characterized numerically hypercyclic weighted shifts and provided an example of a numerically hypercyclic operator on $\C^2$. We answer two questions of Kim, Peris and Song. Namely, we construct a numerically hypercyclic operator, whose square is not numerically hypercyclic as well as an operator which is not numerically hypercyclic but has two numerical orbits whose union is dense in $\C$. We characterize numerically hypercyclic operators on $\C^2$ as well as the operators similar to a numerically hypercyclic one and those operators whose conjugacy class consists entirely of numerically hypercyclic operators. We describe in spectral terms the operator norm closure of the set of numerically hypercyclic operators on a reflexive Banach space. Finally, we provide criteria for numeric hypercyclicity and decide upon the numerical hypercyclicity of operators from various classes.
math.FA
math
On numerically hypercyclic operators Stanislav Shkarin Abstract According to Kim, Peris and Song, a continuous linear operator T on a complex Banach space X is called numerically hypercyclic if the numerical orbit tfpT nxq : n P Nu is dense in C for some x P X and f P X satisfying }x} " }f} " fpxq " 1. They have characterized numerically hypercyclic weighted shifts and provided an example of a numerically hypercyclic operator on C2. We answer two questions of Kim, Peris and Song. Namely, we construct a numerically hypercyclic operator, whose square is not numerically hypercyclic as well as an operator which is not numerically hypercyclic but has two numerical orbits whose union is dense in C. We characterize numerically hypercyclic operators on C2 as well as the operators similar to a numerically hypercyclic one and those operators whose conjugacy class consists entirely of numerically hypercyclic operators. We describe in spectral terms the operator norm closure of the set of numerically hypercyclic operators on a reflexive Banach space. Finally, we provide criteria for numeric hypercyclicity and decide upon the numerical hypercyclicity of operators from various classes. MSC: 47A16, 37A25 Keywords: Numerically hypercyclic operators, numerical orbit, numerical range 1 Introduction . . . zmk 1 1 , . . . , zn Throughout this article X stands for a Banach space over the field C of complex numbers, while SpXq " tx P X : }x} " 1u and BpXq " tx P X : }x} ď 1u. As usual, R is the field of real numbers, R` is the set of non-negative real numbers, D " tz P C : z ă 1u, T " tz P C : z " 1u, Z` is the set of non-negative integers and N is the set of positive integers. If M is a finite or countable set, the symbol `pMq stands for the set of sequences in ℓ1pMq whose entries are all non-negative real numbers. We say ℓ1 that z1, . . . , zk P T are independent if zm1 k ‰ 1 for every non-zero vector m P Zk. Equivalently, zj " eiθj with θj P R are independent if and only if π, θ1, . . . , θk are linearly independent over the field Q of rational numbers. It is well-known and easy to prove that z1, . . . , zk P T are independent if and kq : n P Z`u is dense in Tn. Recall that a sequence tenunPN in X is called a only if the set tpzn Schauder basis in X if every x P X can be uniquely written as x " cnxn, where cn P C and the series is norm-convergent. The functionals en : X Ñ C, enpxq " cn are automatically continuous and are called the coordinate functionals of the Schauder basis tenunPN. Obviously, enpekq " δn,k for every n, k P N. A sequence tenunPN in X is called a Schauder basic sequence if tenunPN is a Schauder basis in the closed linear span of the set ten : n P Nu. The symbol LpXq stands for the space of bounded linear operators T : X Ñ X and X is the space of continuous linear functionals f : X Ñ C. For T P LpXq, the dual operator T P LpXq acts according to the formula T fpxq " fpT xq. For T P LpXq, σpTq stands for the spectrum of T , while σppTq denotes the point spectrum of T . Recall that T P LpXq is called semi-Fredholm if TpXq is closed in X and either X{TpXq or ker T (or both, in which case T is called Fredholm) is finite dimensional. The number ipTq " dim ker T ´ dim X{TpXq P Z Y t´8,`8u is called the index of T . Note that the set of semi-Fredholm operators is norm-open in LpXq and that the index function i is locally constant. Recall that λ P C is called a normal eigenvalue of T P LpXq if λ is an isolated point of σpTq and the spectral projection corresponding to the clopen subset tλu of σpTq has finite rank. This rank is called the multiplicity of the normal eigenvalue λ. Recall that x P X is called a hypercyclic vector for T P LpXq if the orbit 8řn"1 OpT, xq " tT nx : n P Z`u 1 is dense in X. Similarly, x P X is called a weakly hypercyclic vector for T P LpXq if OpT, xq is dense in X with respect to the weak topology. An operator T , which has a hypercyclic vector is called hypercyclic, while an operator T , which has a weakly hypercyclic vector, is called weakly hypercyclic. For more information on hypercyclicity see the books [2, 7] and references therein. Recall also that T P LpXq is called power bounded if supt}T n} : n P Nu ă 8. Unless stated otherwise, we assume that Cn carries the standard Euclidean norm. We shall routinely identify an operator T P LpCnq with its matrix. For instance, a diagonal operator on Cn is an operator whose matrix is diagonal. The same agreement stands for operators on classical sequence spaces. 1.1 The main concept For T P LpXq and px, fq P X X, the numerical orbit of px, fq is the set OpT, x, fq " fpOpT, xqq " tfpT nxq : n P Z`u Ă C. Note that numerical orbits are also known as weak orbits, see [1] and references therein. The following definition, motivated by the concept of the numerical range of an operator, was introduced by Kim, Peris and Song [10], see also [11] for the generalizations to polynomials on Banach spaces. Definition 1.1. Let ΠpXq " tpx, fq P X X : }x} " }f} " fpxq " 1u. We say that px, fq P ΠpXq is a numerically hypercyclic vector for T P LpXq if the numerical orbit OpT, x, fq is dense in C. An operator T is called numerically hypercyclic if it has a numerically hypercyclic vector. We use the symbol NHpXq to denote the set of all numerically hypercyclic operators T P LpXq. The following proposition collects some elementary observations made in [10]. Proposition 1.2. Let T P LpXq. • If T is power bounded, then it is not numerically hypercyclic. • If T is weakly hypercyclic, then it is numerically hypercyclic. • If the restriction of T to a closed invariant subspace is numerically hypercyclic, then so is T . • If T is numerically hypercyclic, then T is numerically hypercyclic. In particular, if X is reflexive, then T is numerically hypercyclic if and only if T is. • If X is a Hilbert space, then T is numerically hypercyclic if and only if there is x P SpXq such that txT nx, xy : n P Z`u is dense in C, where x,y is the inner product of X. The last observation is due to the fact that if H is a Hilbert space and x P SpHq, then the only functional f P SpHq satisfying fpxq " 1 is given by fpyq " xy, xy. Thus in this case the numerical orbits featuring in the definition of numeric hypercyclicity are naturally labelled by the elements of SpHq. If H is a Hilbert space and T P LpHq, we use the following notation: NOpT, xq " txT nx, xy : n P Z`u for x P SpHq. Recall that forward Fw and backward Bw weighted shifts on ℓppZ`q act according to the rule px0, x1, x2, . . . q FwÞÑp0, w1x0, w2x1, . . . q and px0, x1, x2, . . . q BwÞÑpw1x1, w2x2, . . . q, where w " twnunPN is a bounded sequence of non-zero scalars. A bilateral weighted shift Tw on ℓppZq is defined by pTwxqn " wn`1xn`1, where again w " twnunPZ is a bounded sequence of non-zero scalars. The main results of [10] can be summarized as follows. Theorem KPS1. There is a diagonal numerically hypercyclic operator on C2. Consequently, there is a numerically hypercyclic operator on Cn for every n ě 2. Theorem KPS2. Let 1 ă p ă 8 and T be a backward or forward weighted shift on ℓppNq or a bilateral weighted shift on ℓppZq. Then T is numerically hypercyclic if and only if T is not power bounded. 2 Remark 1.3. Theorem KPS2 is not exactly what is stated in [10]. Namely, the corresponding results in [10] provide a condition on the weight sequence equivalent to the numeric hypercyclicity. For bilateral weighted shifts the authors notice that this condition is equivalent to the operator being not power bounded, while saying nothing for unilateral shifts. However upon reflection, one can readily see that this condition is again equivalent to the operator being not power bounded. Furthermore in the case of unilateral backward shifts the same condition is equivalent to hypercyclicity. One has just co compare it to the characterization of Salas [15] of hypercyclic backward shifts. On the other hand, there are no hypercyclic forward shifts and there are non-hypercyclic bilateral shifts, which are not power bounded. Recall that T P LpXq and S P LpY q are called similar if there is an isomorphism (in the category of topological vector spaces) J : X Ñ Y such that J´1SJ " T . Note that hypercyclicity and weak hypercyclicity are similarity invariants, while the numeric hypercyclicity is not (below we provide plenty of examples substantiating the last claim). This prompts us to consider the following concepts. Definition 1.4. We say that T P LpXq is weakly numerically hypercyclic if T is similar to a numerically hypercyclic operator. We say that T P LpXq is strongly numerically hypercyclic if every operator similar to T is numerically hypercyclic. We denote the sets of weakly numerically hypercyclic operators and of strongly numerically hypercyclic operators on X by the symbols WNHpXq and SNHpXq respectively. The following elementary proposition characterizes weak numeric hypercyclicity. Proposition 1.5. Let T P LpXq. Then T is weakly numerically hypercyclic if and only if there exist x P X and f P X such that OpT, x, fq is dense in C. Proof. The 'only if' part is obvious. Let x P X and f P X be such that OpT, x, fq is dense in C. It remains to show that T P WNHpXq. Replacing x by T nx for an appropriate n P N and normalizing, we can without loss of generality assume that }x} " fpxq " 1. Let D " tx P X : }x} ď 1, fpxq ď 1u and }}1 be the Minkowski functional of D. Since D is bounded and contains a neighborhood of 0 in X, }}1 is a norm on X equivalent to the original norm } }. It is easy to see that }x}1 " }f}1 " fpxq " 1, where }f}1 is the norm of f as a functional on the normed space X1 " pX,} }1q. Thus px, fq P ΠpX1q and therefore px, fq is a numerically hypercyclic vector for T acting on X1. Since }} and }}1 are equivalent, T acting on X is similar to T acting on X1. Thus T is weakly numerically hypercyclic. According to Feldman [4], T P LpXq is called 1-weakly hypercyclic if there is x P X such that fpOpT, xqq is dense in C for each non-zero f P X. Of course, every weakly hypercyclic operator is 1- weakly hypercyclic. We collect a number of straightforward consequences of Definition 1.4 in the following proposition. The proof is elementary and is left for the reader as an exercise. Proposition 1.6. Let T P LpXq. • If T is power bounded, then T R WNHpXq. If T is 1-weakly hypercyclic, then T P SNHpXq. • If a restriction of T to an invariant closed linear subspace is weakly numerically hypercyclic pres- pectively, strongly numerically hypercyclicq, then T is weakly numerically hypercyclic prespectively, strongly numerically hypercyclicq. • If T P WNHpXq, then T P WNHpXq. If additionally X is reflexive, then T P WNHpXq ðñ T P WNHpXq and T P SNHpXq ðñ T P SNHpXq. 1.2 Finite dimensional results The following two questions are raised in [10]. Question KPS3. Let T P NHpXq. Is T n numerically hypercyclic for every n P N? Question KPS4. Let T P LpXq and there exist pxj , fjq P ΠpXq for 1 ď j ď n for which is dense in C. Is T numerically hypercyclic? nŤj"1 OpT, xj , fjq We answer both questions negatively. 3 Theorem 1.7. There exists a diagonal T P NHpC2q such that T 2 R NHpC2q. Theorem 1.8. There exist a diagonal T P LpC4q and x1, x2 P SpC4q such that T R NHpC4q, while NOpT, x1q Y NOpT, x2q is dense in C. The following two results are motivated by Theorem KPS1. They provide sufficient conditions of weak/strong numeric hypercyclicity in terms of the restriction to a finite dimensional invariant subspace. Theorem 1.9. Each of the following conditions is sufficient for the weak numeric hypercyclicity of T P LpXq. (1.9.1) There exist λ1, λ2 P σppTq such that λ1 " λ2 ą 1 and λ1 λ1 (1.9.2) There are independent λ1, λ2 P T such that kerpT ´ λj Iq2 ‰ kerpT ´ λj Iq for j P t1, 2u. (1.9.3) There exist λ1, λ2, λ3 P σppTq such that λ1 " λ2 ą λ3 ą 1, λ1 has infinite order in the group T are independent in T. , λ2 λ2 λ2 and λ1 λ1 , λ3 λ3 are independent. (1.9.4) There exist λ1, λ2 P σppTq and λ3 P T such that kerpT ´ λ3Iq2 ‰ kerpT ´ λ3Iq, λ1 " λ2 ą 1, λ1 λ2 has infinite order in T and λ1 λ1 , λ3 are independent. Theorem 1.10. Each of the following conditions is sufficient for the strong numeric hypercyclicity of T P LpXq. (1.10.1) There exist λ1, . . . , λn P σppTq and c1, . . . , cn P R` such that ! nřj"1 (1.10.2) There exist λ1, λ2, λ3 P σppTq such that λ1 " λ2 " λ3 ą 1 and λ1 λ1 cj λk j : k P Z`) is dense in C. , λ2 λ2 , λ3 λ3 are independent. In the Hilbert space situation we can refine the first two parts of Theorem 1.9. Proposition 1.11. Let H be a Hilbert space and T P LpHq. Assume that there exist λ1, λ2 P σppTq such that λ1 " λ2 ą 1, the eigenspaces kerpT ´ λ1Iq and kerpT ´ λ2Iq are non-orthogonal and λ1 , λ2 λ1 λ2 are independent in T. Then T P NHpHq. Proposition 1.12. Let H be a Hilbert space and T P LpHq. Assume that there exist λ1, λ2 P C such that kerpT ´ λjIq2 ‰ kerpT ´ λj Iq for j P t1, 2u, λ1 " λ2 ě 1 and λ1 are independent in T. Then λ1 T P NHpHq. , λ2 λ2 We characterize NHpC2q, SNHpC2q, WNHpC2q and WNHpC3q. , λ2 λ2 1 ` λk Theorem 1.13. Let T P LpC2q. Then T P WNHpC2q if and only if σpTq " tλ1, λ2u, where λ1 " λ2 ą 1 and λ1 are independent in T. Furthermore, T P SNHpC2q if and only if σpTq " tλ1, λ2u λ1 with tλk 2 : k P Z`u being dense in C. Finally, T P NHpC2q if and only if either T P SNHpC2q or T P WNHpC2q and T is not unitarily equivalent to a diagonal operator. Theorem 1.14. Let T P LpC3q. Then T is weakly numerically hypercyclic if and only if either there are λ1, λ2 P σpTq such that λ1 " λ2 ą 1 and λ1 are independent in T or σpTq " tλ1, λ2, λ3u, where λ1 λ1 " λ2 ą λ3 ą 1, λ1 Remark 1.15. The above results provide easy explicit examples of numerically hypercyclic operators on Cn. For instance, applying Theorems 1.13, Theorem 1.10 and Proposition 1.12 respectively, we have has infinite order in the group T and λ1 λ1 are independent in T. , λ2 λ2 , λ3 λ3 λ2 0 2eiπ log 2 and 1 2eiπ log 3 P NHpC2q, eiπ log 2 1 0 0 0 eiπ log 2 0 0 0 0 eiπ log 3 0 0 1 0 eiπ log 3 2eiπ log 2 0 0 0 2eiπ log 3 0 ‹‹‚P NHpC4q. 4 0 0 2eiπ log 5 ‚P SNHpC3q Ă NHpC3q In relation to Theorem 1.13 it makes sense to mention the following fact. Proposition 1.16. For every R ą 1, pz, wq P T2 : tRnpzn ` wnq : n P Z`u is dense in C( is a dense Gδ-subset of T2 of Lebesgue measure 0. Remark 1.17. Let tx, yu be a fixed linear basis in C2 and R ą 1. For each pz, wq P T2, consider T " Tz,w P LpC2q given by T x " Rzx and T y " Rwy. By Theorem 1.13 and Proposition 1.16, for almost all pz, wq P T2 in the Lebesgue measure sense, Tz,w P WNHpC2qzNHpC2q if xx, yy " 0 and Tz,w P NHpC2qzSNHpC2q if xx, yy ‰ 0. In particular, there are numerically hypercyclic operators on C2 which are not strongly numerically hypercyclic and there are weakly numerically hypercyclic operators on C2 that are not numerically hypercyclic. Due to Le´on-Saavedra and Muller [12], if T P LpXq is hypercyclic or weakly hypercyclic, then so is zT for each z P T. Moreover, Theorem KPS2 ensures that if a weighted shift T is numerically hypercyclic, then so is zT for each z P C satisfying z ě 1. In general, the scalar multiples of numerically hypercyclic operators do not exhibit this nice behavior even in the friendly realm of diagonal operators on C2. Proposition 1.18. For every T P LpC2q, there is w P T such that wT R WNHpC2q. Proposition 1.19. For each z, w P T, Mpz, wq " "r ą 1 : rz 0 0 rw P NHpC2q* is a Gδ-subset of R, being either infinite or empty. There exist w, z P T such that Mpz, wq is a dense Gδ-subset of p1,8q and therefore is uncountable. For every w, z P T, Mpz, wq has zero Lebesgue measure. By Proposition 1.19, not a single diagonal T P LpC2q satisfies rT P NHpC2q for each r ą 1. 1.3 Infinite dimensional results The following two theorems provide further sufficient conditions for weak/strong numeric hypercyclicity. Theorem 1.20. Each of the following conditions is sufficient for the weak numeric hypercyclicity of T P LpXq. (1.20.1) There is a sequence tznunPN in σppTq such that 1 ă z1 ă z2 ă . . . (1.20.2) There is r ą 1 such that σppTq X rT is infinite. (1.20.3) There is a cyclic vector x for T satisfying lim inf nÑ8 (1.20.4) There are x P X and an infinite set A Ď N such that }T nx} Ñ 8 as n Ñ 8, n P A and the sequence 1`}T nx} }T n} " 0. is weakly convergent but is not norm convergent. }T nx}(nPA T nx The following examples show how to apply Theorem 1.20. Example 1.21. I ` V P WNHpL2r0, 1sq, where V fpxq "şx Proof. It is easy to verify (see, for instance, [3]) that }T n} Ñ 8 and }T ng} " op}T n}q, where T " I ` V and gpxq " x for x P r0, 1s. Since g is a cyclic vector for T , (1.20.3) is satisfied and T P WNHpL2r0, 1sq by Theorem 1.20. 0 fptq dt is the classical Volterra operator. Example 1.22. Let a P Cr0, 1s be non-constant and }a} ą 1. Then M P WNHpCr0, 1sq, where M is the multiplication by a operator: M f " af for f P Cr0, 1s. Proof. Let K " tt P r0, 1s : aptq " }a}u. If K ‰ r0, 1s, we can pick f P Cr0, 1s such that K " tt P r0, 1s : fptq " 0u and let X be the closed linear span of OpM, fq. It is easy to see that }M nX} " }a}n for every n P Z` and }T nf} " op}a}nq as n Ñ 8. Thus (1.20.3) is satisfied for MX and therefore MX P WNHpXq. By Proposition 1.2, M P WNHpCr0, 1sq. It remains to consider the case K " r0, 1s. Let R " }a}. Since a is non-constant and a " R, Proposition 1.16 ensures that we can pick w, z P T such that Rz, Rw P apr0, 1sq and tRnpzn` wnq : n P Z`u is dense in C. Choose s, t P r0, 1s such that apsq " Rz and aptq " Rw and let 1 be the constant 1 function on r0, 1s. Set ϕ P Cr0, 1s to be ϕpfq " 1 2pfpsq`fptqq. Clearly, p1, ϕq P ΠpCr0, 1sq and ϕpM n1q " 1 2 Rnpzn ` wnq for n P Z`. Thus OpM, 1, ϕq is dense in C and therefore M P NHpCr0, 1sq Ă WNHpCr0, 1sq. 5 Theorem 1.23. Each of the following conditions is sufficient for the strong numeric hypercyclicity of T P LpXq. (1.23.1) X is reflexive and there are a Schauder basic sequence tenunPN in X and c P ℓ1 `pNq such that T en " λnen with λn P C for each n P N and ! 8řj"1 cj λk j : k P Z`) is dense in C. (1.23.2) X is reflexive and there is a Schauder basic sequence tenunPN in X such that T en " λnen for each n P N, where λn P C are such that λ1 ą 1, the sequence tλnunPN is pmaybe non-strictlyq increasing and the numbers λn λn are pairwise distinct elements of T. (1.23.3) There is λ P C such that λ ě 1 and T ´ λI is semi-Fredholm of positive index. (1.23.4) X is reflexive and there is λ P C such that λ ě 1 and T ´ λI is semi-Fredholm of negative index. The reflexivity condition in the above theorem can not be removed entirely as illustrated by the following examples. cnzk ` such that f acts according to the formula fpyq " řnPA Example 1.24. Let tznunPN be a sequence of finite order elements of T, r ą 1 and T P Lpc0pNqq be the diagonal operator with the numbers rzn on the diagonal. Then T R NHpc0pNqq. Proof. Let px, fq P Πpc0pNqq and A " tn P N : xn " 1u. Clearly A is non-empty and finite and there is c P RA cnxnyn for each y P c0pNq. Hence fpT kxq " rk řnPA n for each k P Z`. Since A is finite and each zn has finite order, OpT, x, fq is contained in the union of finitely many lines in C through the origin. Hence OpT, x, fq is nowhere dense in C and T R NHpc0pNqq. Example 1.25. Let r ą 1 and T R NHpc0pNqq be given by pT xq1 " 0 and pT xqn " rxn´1 if n ą 1. Then T ´ I is Fredholm of index ´1 and T R NHpc0pNqq. Proof. Let px, fq P Πpc0pNqq and the finite set A Ă N and c P RA ` be as in the proof of Example 1.24. It is easy to see that fpT kxq " 0 if k ě maxpAq. Thus OpT, x, fq is a finite set. Hence T R NHpc0pNqq. On the other hand, T ´ I is injective and pT ´ Iqpc0pNqq is exactly the kernel of ϕ P c0pNq given by ϕpxq " r´nxn. Thus T ´ I is Fredholm of index ´1. 8řn"1 Obviously, a self-adjoint operator T on a Hilbert space H can not be numerically hypercyclic. Indeed, NOpT, xq Ă R for each x P SpHq. The following theorem nearly (but not quite) characterizes weakly numerically hypercyclic normal operators. It also allows a funny characterisation of weakly numerically hypercyclic self-adjoint operators. Theorem 1.26. Let H be a Hilbert space, T P LpHq and k P N be such that T k is normal. Then the following statements are true. (1.26.1) T P WNHpHq if there is a sequence tλnunPN in σpTq such that 1 ă λ1 ă λ2 ă . . . (1.26.2) T P NHpHq if there is a sequence tλnunPN in σpTq such that 1 ă λ1 ă λ2 ă . . . and the numbers λj λj are pairwise distinct. (1.26.3) If T k is self-adjoint, then T P WNHpHq if and only if the set t´z : z P σpTq, (1.26.4) T P NHpHq if there exist r ą 1 and a T k-invariant subspace K such that 1 well-ordered by the natural ordering of R. operator with infinite spectrum. z ą 1u is not r T kK P LpKq is a unitary Corollary 1.27. A self-adjoint operator T is weakly numerically hypercyclic if and only if the set t´z : z P σpTq, z ą 1u is not well-ordered by the natural ordering of R. Corollary 1.28. Let r ą 1 and T be a unitary operator with infinite spectrum. Then rT is numerically hypercyclic. 6 Herrero [8, 9] described the norm closure of the set of hypercyclic operators on ℓ2pNq in terms of the spectrum. Prajitura [14] demonstrated that the set of weakly hypercyclic operators on ℓ2pNq has the same closure. We provide a spectral description of the closures of SNHpXq, WNHpXq and NHpXq for an arbitrary reflexive Banach space X. Theorem 1.29. Let X be reflexive and T P LpXq. Then the following statements are equivalent: (1.29.1) T does not belong to the operator norm closure of SNHpXq; (1.29.2) T does not belong to the operator norm closure of NHpXq; (1.29.3) T does not belong to the operator norm closure of WNHpXq; (1.29.4) the set tλ P σpTq : λ ě 1u consists of finitely many normal eigenvalues of multiplicity 1 with pairwise distinct absolute values. 1.4 The structure In Section 2 we prove Theorems 1.7 and 1.8 by means of explicit examples. In the same section we show that the existence of a numerical orbit NOpT, xq dense in a non-empty open subset of C does not guarantee the numeric hypercyclicity of T P LpC2q. In Section 3 we prove Proposition 1.16 and build up a toolbox used later on in order to prove the main results. Section 4 is devoted to the proof of Theorems 1.9 and 1.10, which provide sufficient conditions of weak and of strong numeric hypercyclicity in terms of a restriction to a finite dimensional invariant subspace. In Section 5, we prove Propositions 1.11 and 1.12, which deal with the similar issues in the Hilbert space setting. Theorem 1.20, providing genuinely infinite dimensional sufficient conditions of weak numeric hypercyclicity, is proved in Section 6. We prove Theorems 1.13, describing the numerically hypercyclic operators on C2, and Theorem 1.14, describing the weakly numerically hypercyclic operators on C3, in Section 7. Theorem 1.23, which provides infinite dimensional sufficient conditions for the strong numeric hypercyclicity, is proved in Section 8. In Section 9, we prove Theorem 1.29, which describes the operator norm closure of the set of numerically hypercyclic operators on a reflexive Banach space. Theorem 1.26, dealing with numeric hypercyclicity of normal operators, is proved in Section10. In Section 11, we prove Propositions 1.18 and 1.19 on scalar multiples of numerically hypercyclic operators on C2. In Section 12, we make some further remarks, construct few examples and raise several questions on the numeric hypercyclicity. 2 Proof of Theorems 1.7 and 1.8 Our construction is similar to the proof of Theorem KPS1 in [10]. The main difference is that we need and achieve better control of the numbers in the numerical orbit. Note also that there are natural generalizations and some estimates below are far from optimal. 2.1 Funny numbers As usual, for m, n P N, we write m n if m is a divisor of n. Let p be an odd prime number and νk " νkppq for k P N be powers of p defined recurrently by the formula ν1 " p and νk`1 " νkpk`νk for k P N. Let also w " twkukPN be a fixed sequence of complex numbers satisfying 1 k ď wk ď k for each k P N. We define a sequence tmk " mkpp, wqukPN of positive integers according to the following rule: m1 " 1 and mk " mintm P N : m ą wk´1pk´1 π , 2 m, p ffl mu for k ě 2. (2.1) (2.2) (2.3) 7 Since the gaps between consecutive members of tm P N : 2m, p ffl mu do not exceed 4, we have 0 ď mk ´ wk´1pk´1 π ă 4 for every k ě 2. According to (2.2) and (2.4), By (2.1) and (2.5), the series ř mk νk mk " Opkpkq as k Ñ 8. converges, which allows us to define τ " τpp, wq " 8řk"1 mk νk P R and z " zpp, wq " eπτ i P T. Lemma 2.1. In the above notation the following asymptotic relations hold: lim kÑ8`pνk1 ` zνk ´ wk " 0; nRtνk :kPNu 1 ` zn1{n ě p´1{3. lim inf nÑ8 Proof. Note that for each k P N, (2.4) (2.5) (2.6) (2.7) (2.8) νkτ " Nk ` Since eπNki " ´1, we have zνk " ´ exp´πi 1 ` zνk " ´πi mk`1νk 8řs"k`1 νs νs msνk msνk , where Nk " 8řs"k`1 křs"1 νs ¯. Hence νk`1 ` O´ mk`1νk νk`1 ¯2 8řs"k`2 msνk ` is an odd integer. (2.9) msνk νs as k Ñ 8. Multiplying by pνk and using (2.1), we get pνk1 ` zνk " πmk`1 pk ` O m2 k`1νk νk`1pk ` msνkpνk νs as k Ñ 8. 8řs"k`2 By (2.1) and (2.5), the sequence in the O sign converges to 0 as k Ñ 8. Hence pνk1 ` zνk ´ πmk`1 as k Ñ 8. The inequality (2.4) yields πmk`1 pk ´ wk Ñ 0 as k Ñ 8 and (2.7) follows. We prove (2.8) in two steps. Denote Λ " !νkm : k P N, m P 2N ` 1, m ă b νk`1 pk Ñ 0 νk ). First, we shall verify that If n P N is large enough, there is a unique k " kpnq ě 2 such that ?νk´1νk ă n ď ?νkνk`1. By (2.9), lim nÑ8 nRΛ 1 ` zn1{n " 1. (2.10) 8řs"k`1 Using the estimate n ď ?νkνk`1, (2.1) and (2.5), we obtain k¯. νs " O´ 1 nτ " nNk νk ` msn ν 2 νs msn 8řs"k`1 . If n R Λ, n νk is not an odd integer. Since p ffl Nk, nNk νk nNk νk to the nearest odd integer is at least 1 νk . Then by the last two displays, can not be an odd integer. Hence the distance from 1 ` zn " 1 ` eπnτ i ě 1 2νk if n R Λ is sufficiently large. 8 Since ?νk ă n, 2 ě 1 ` zn ě 1 In order to complete the proof of (2.8) and that of the lemma, it now suffices to show that 2n2 for all sufficiently large n R Λ, which immediately implies (2.10). 1 ` zn1{n ě p´1{3. kÑ8`pνk1 ` zνk ´ wk " 0. Combining it with (2.2), we k " Op1 ` zνkq and 1 ` zνk " Opkp´νkq as k Ñ 8. Hence, we can find a sequence rk of real By the already verified equality (2.7), lim inf nÑ8 nPΛ (2.11) lim have p´νk numbers such that zνk " ´eirk for every k P N, k " Oprkq and rk " Opkp´νkq as k Ñ 8. p´νk The last relation yields rk?νk`1νk Ñ 0 as k Ñ 8. Thus nrk ă π for n " νkm P Λ for all sufficiently large k. Using the inequality eit ´ 1 ě 2 πt if t ď π, we see that if nrk ă π and n " νkm P Λ is sufficiently large, then 1 ` zn " 1 ´ eirkm ě 2 πrkm ě c mp´νk for some constant c ą 0 independent on n (c does exist because p´νk k ě cp´νk k " Oprkq). Hence 1 ` zn1{n ě pc{nq1{n`p´νk1{νkm " pc{nq1{np´1{m ě pc{nq1{np´1{3 Ñ p´1{3 n since m ě 3. Thus the lower limit of 1 ` zn1{n is at least p´1{3 as n Ñ 8, n P Λ. This completes the proof of (2.8) and that of the lemma. Keeping the above notation, for each k P N, we choose qk P t0, 1, . . . , pk ´ 1u such that wk1`zνk wkp1`zνkq ´ exp´ 2πqki wkp1`zνkq ´ exp´ 2πqi The inequality 0 ď qk ă pk and (2.1) ensure that the series ř qk pk ¯ " min" wk1`zνk pkνk pk ¯ : q P t0, 1, . . . , pk ´ 1u*. converges. Thus we can define θ " θpp, wq " Lemma 2.2. In the above notation, 8řk"1 qk pkνk P R, and u " upp, wq " e2πθi P T. (2.12) (2.13) (2.14) Proof. By (2.12), wk1`zνk relations and the estimate wk ď k imply that lim kÑ8`uνk pνkp1 ` zνkq ´ wk " 0. pk ¯ ď π wkp1`zνkq ´ exp´ 2πqki exp´ 2πqki pk ¯pνkp1 ` zνkq ´ wk Ñ 0 as k Ñ 8. uνk exp´´ 2πqki νsps ¯ " 1 ` O´ 8řs"k`1 pk ¯ " exp´ 8řs"k`1 2πνkqki By (2.13), pk . By Lemma 2.1, `pνk1 ` zνk ´ wk Ñ 0. These two (2.15) νkqk νsps¯. 8řs"k`1 νkqk νsps " Opνk{νk`1q " Opp´k´νkq and pνk1 ` zνk " Opwkq " Opkq, we have Since puνk ´ expp2πqki{pkqqpνkp1 ` zqνk Ñ 0, which together with (2.15) implies (2.14). 9 2.2 T P LpC2q with a prescribed numerical orbit The following result shows that the somewhere dense orbit theorem fails miserably for numerical orbits. Proposition 2.3. Let a non-empty open subset U of C be the interior of a closed set. Then there exist a diagonal T P LpC2q and x0 P SpC2q such that U is the interior of NOpT, x0q and for every x P SpC2q either NOpT, xq " NOpT, x0q or NOpT, xq is nowhere dense in C. Proof. Let p be an odd prime number. Choose a sequence w " twkukPN in C such that 1 k ď wk ď k for every k P N and twk : k ě k0u is a dense subset of 2U for some k0 P N. By Lemmas 2.1 and 2.2, (2.14) and (2.8) hold for νk " νkppq, z " zpp, wq and u " upp, wq defined in (2.1), (2.6) and (2.13) respectively. Now let T P LpC2q be the diagonal operator with the numbers pu and puz on the diagonal. If x " pa, bq P SpC2q, then xT nx, xy " pnunpa2 ` b2znq for every n P Z`. If a ‰ b, the above display yields xT nx, xy ě pna2 ´ b2 Ñ 8 and NOpT, xq is nowhere dense in C. If a " b, then a " b " 1?2 and the above display reads 2xT nx, xy " pnunp1 ` znq for every n P Z`. By (2.14), xT νk x, xy ´ wk 2 Ñ 0 as k Ñ 8. Since twk{2 : k ě k0u is a dense subset of U and U is the interior of a closed set, U is the interior of the closure of txT νkx, xy : k P Nu. By the above display and (2.8), xT nx, xy Ñ 8 as n Ñ 8, n R tνk : k P Nu. Hence U is the interior of the closure of txT nx, xy : n P Nu " NOpT, xq. It remains to note that if a " b, then NOpT, xq does not depend on x. 2.3 Proof of Theorem 1.7 Let p be an odd prime number. Choose a sequence w " twkukPN in C such that 1 k ď wk ď k for every k P N and twk : k P Nu is dense in C. By Lemmas 2.1 and 2.2, (2.14) and (2.8) hold for νk " νkppq, z " zpp, wq and u " upp, wq defined in (2.1), (2.6) and (2.13) respectively. Now let T P LpC2q be the diagonal operator with the numbers pu and puz on the main diagonal. , 1?2 P SpC2q, then 2xT nx0, x0y " pnunp1 ` znq for every n P Z`. By (2.14), First, if x0 " ` 1?2 xT νkx0, x0y ´ wk 2 Ñ 0 as k Ñ 8. Since twk : k P Nu is dense in C, txT νk x0, x0y : k P Nu is dense in C. Hence NOpT, xq is dense in C and T P NHpC2q. If a ‰ b, as in the proof of Proposition 2.3, we have xT nx, xy ě Let x " pa, bq P SpC2q. pna2´b2 Ñ 8 and NOpT, xq is nowhere dense in C. Hence NOpT 2, xq, being a subset of NOpT, xq, is nowhere dense in C. If a " b, then 2xpT 2qnx, xy " p2nu2np1 ` z2nq for n P Z`. Since each νk is odd, nÑ8 1`z2n1{2n ě p2{3 ą 1. Hence xpT 2qnx, xy Ñ 8 (2.8) guarantees that lim inf and NOpT 2, xq is nowhere dense in C. Thus T 2 R NHpC2q. The proof of Theorem 1.7 is complete. 2.4 Proof of Theorem 1.8 Let p and q be two odd prime numbers such that p ą q3{2, νk " νkppq and ν1k " νkpqq be as defined in (2.1). Let also w " twkukPN and w1 " tw1kukPN be two sequences of complex numbers such that k ď wk,w1k ď k for every k P N, twk : k P Nu is a dense subset of the half-plane tt P C : Re t ą 0u and tw1k : k P Nu is a dense subset of the half-plane tt P C : Re t ă 0u. Let also z " zpp, wq, z1 " zpq, w1q, u " upp, wq and u1 " upq, w1q be as defined in (2.6) and (2.13) and let T be the diagonal operator on C4 with the numbers pu, puz, qu1 and qu1z1 on the diagonal. First, observe that if x, y P SpC4q are given by x "` 1?2 nÑ8 xpT 2qnx, xy1{2n " p lim inf , 0, 0 and y "`0, 0, 1?2 , 1?2, then , 1?2 1 2xT nx, xy " pnunp1 ` znq and 2xT ny, yy " qnpu1qnp1 ` pz1qnq for every n P Z`. By Lemma 2.2, xT νkx, xy ´ wk 2 : k P Nu is dense in C, NOpT, xq Y NOpT, yq is dense in C. Hence the union of 2 numerical orbits of T corresponding 2 Ñ 0. Since t wk 2 Ñ 0 and xT ν1 2 : k P Nu Y t w1 ky, yy ´ w1 k k 10 to elements of ΠpC4q is dense in C. The proof will be complete if we prove that T is not numerically hypercyclic. In order to do that, let x " pa, b, c, dq be an arbitrary element of SpC4q. It remains to show that NOpT, xq is not dense in C. A direct computation yields xT nx, xy " pnunpa2 ` b2znq ` qnpu1qnpc2 ` d2pz1qnq for every n P Z`. (2.16) Case 1: a ‰ b. By (2.16), xT nx, xy ě a2 ´ b2pnp1 ` Opqn{pnqq Ñ 8 as n Ñ 8 since q ă p. Case 2: a " b ‰ 0 and c ` d ‰ 0. Since p ą q3{2, we can pick r P pq, p2{3q. By (2.8), there is Hence NOpT, xq is non-dense in C. c ą 0 such that pn1 ` zn ě crn for every n P Nztνk : k P Nu. According to (2.16), we have xT nx, xy ě a2crn ´ 2qn Ñ 8 as n Ñ 8, n R tνk : k P Nu. By (2.7), pνk1 ` zνk ´ wk Ñ 0. Since wk ď k, pνk1 ` zνk " Opkq as k Ñ 8. Pick s P p1, q2{3q. If c " d ‰ 0, the obvious observation that ν1k ‰ νm for every k, m P N and (2.8) imply that there is c1 ą 0 such that qνkpu1qνkpc2 ` d2pz1qνkq " qνkc21 ` pz1qνkq ě c1sνk for every k P N. If c ‰ d, then qnpu1qnpc2 `d2pz1qnq ě c2 ´d2qn and therefore qνkpu1qνkpc2 `d2pz1qνkq ě c2 ´ d2qνk ě c2 ´ d2sνk . Thus in any case there is c2 ą 0 such that qνkpu1qνkpc2 ` d2pz1qνkq ě c2sνk for every k P N. Combining these estimates with (2.16), we get xT νk x, xy ě c2sνk ` Opkq Ñ 8 as k Ñ 8. Then by the above display xT nx, xy Ñ 8 as n Ñ 8 and NOpT, xq is non-dense in C. Case 3: a " b " 0 and c ‰ d. By (2.16), xT nx, xy ě c2 ´ d2qn Ñ 8 as n Ñ 8. Hence NOpT, xq is non-dense in C. Case 4: a " b " 0 and c " d. By (2.16), 2xT nx, xy " qnpu1qnp1 ` pz1qnq for n P N. Then by (2.8), the lower limit of xT nx, xy1{n as n Ñ 8, n R tν1k : k P Nu is at least q2{3 ą 1. Hence xT nx, xy Ñ 8 as 2 ¯ " 0 . Since n Ñ 8, n R tν1k : k P Nu. On the other hand, according to Lemma 2.2, each w1k is in the left half-plane, NOpT, xq has no accumulation points in the open right half-plane. Thus NOpT, xq is non-dense in C. a " b and c " d " 0. By (2.16), 2xT nx, xy " pnunp1 ` znq for every n P N. Then according to (2.8), the lower limit of xT nx, xy1{n as n Ñ 8, n R tνk : k P Nu is at least p2{3 ą 1. Hence 2 ¯ " 0. xT nx, xy Ñ 8 as n Ñ 8, n R tνk : k P Nu. On the other hand, by Lemma 2.2, lim Since each wk is in the right half-plane, NOpT, xq has no accumulation points in the open left half-plane. Hence NOpT, xq is non-dense in C. The proof of Theorem 1.8 is complete. kÑ8´xT νkx, xy´ wk k x, xy ´ w1 Case 5: kÑ8´xT ν1 lim k 3 Dense sums of powers We need a number of technical results. First, we state the Universality Theorem, which will be repeatedly used in this and in the subsequent sections. A topological space X is called Baire if the intersection of countably many dense open subsets of X is always dense in X. Recall that a Polish space is a separable topological space X, whose topology can be defined by a metric d such that the metric space pX, dq is complete. By the Baire Theorem, every Polish space is Baire. The following result is a particular case of the Universality Theorem [6, p. 348 -- 349], see also [2, 7]. Theorem U. Let X be a Baire topological space, Y be a Polish space and F " tfa : a P Au be a collection of continuous maps fa : X Ñ Y such that the set tpx, fapxqq : x P X, a P Au is dense in X Y . Then x P X : tfapxq : a P Au is dense in Y( is a dense Gδ-subset of X. In particular, there is x P X such that tfapxq : a P Au is dense in Y . 3.1 Finite sums Lemma 3.1. Let z, w P T be independent and tRnunPN be a sequence of positive numbers such that Rn Ñ 8. Then pa, bq P T2 : tRnpazn ` bwnq : n P Z`u is dense in C( is a dense Gδ-subset of T2. 11 Proof. Obviously, fnpa, bq " Rnpazn ` bwnq is a continuous map from T2 to C. By Theorem U, the proof will be complete if we verify that Λ " tpa, b, Rnpazn ` bwnqq : pa, bq P T2, n P Z`u is a dense subset of T T C. and tn " v2Rn nk P T, ak " sz´nk yk and bk " sw´nk y´1 Pick α, β P T and v P Czt0u and let s " v . Clearly, tn Ñ 0 and v " 2Rnstn for each n P N. v Since z and w are independent, tpzn, wnq : n P Z`u is dense in T2 and there is a strictly increasing sequence tnkukPN of positive integers such that tnk ă 1 for each k, znk Ñ isα´1 and wnk Ñ ´isβ´1 as k Ñ 8. Set yk " tnk ` ib1 ´ t2 k . Then Rnkpakznk ` bkwnkq " 2stnk Rnk " v and therefore pak, bk, vq P Λ for each k P N. Since yk Ñ i, znk Ñ isα´1 and wnk Ñ ´isβ´1, we have ak Ñ α and bk Ñ β. Hence the sequence tpak, bk, vqu of elements of Λ converges to pα, β, vq. Since α, β P T and v P Czt0u are arbitrary, Λ is dense in T T C. Lemma 3.2. Let z, w P T be independent, k P N, m P Z, and tRnunPN, trnunPN be two sequences of positive numbers such that Rn Ñ 8, rn Ñ 8 and Rn pa, bq P T2 : tRnzmnpazkn ` a´1z´knq ` brnwn : n P Z`u is dense in C( is a dense Gδ-subset of T2. Proof. Clearly, fnpa, bq " Rnzmnpazkn ` a´1z´knq ` brnwn is a continuous map from T2 to C. By Theorem U, the proof will be complete if we verify that rn Ñ 8. Then Λ " tpa, b, Rnzmnpazkn ` a´1z´knq ` brnwnq : pa, bq P T2, n P Z`u is a dense subset of T T C. Pick α, β P T and v P C. Since rn Ñ 8, for each sufficiently large n, the circle z´mnv ` rnT intersects the real line at 2 points with exactly one sn being in p0,8q. Let un P T be the unique number such that z´mnv ` rnun " sn. Obviously, un Ñ 1 and sn rn Ñ 1 and rn Rn ď 2 for all sufficiently large n. For such n, we can consider qn " sn Rn Ñ 0. Thus sn i P T. Clearly, qn Ñ i as n Ñ 8. Let an " qnz´kn. It is easy to verify that rn Ñ 1. Denote bn " ´zmnw´nun. Since sn Rn Ñ 0, we have sn 2Rn `b1 ´ s2 n 4R2 n Rnpanzkn ` a´1 n z´knq ` rnbnwn " v for each sufficiently large n. Since z and w are independent, tpzkn, wnz´mnq : n P Z`u is dense in T2. Thus there is a strictly increasing sequence tnjujPN of positive integers such that zknj Ñ iα´1 and wnj z´mnj Ñ ´β´1 as j Ñ 8. Since un Ñ 1 and qn Ñ i, we have bnj " ´w´nj z´mnj unj Ñ β and anj " qnj z´knj Ñ α. By the above display, panj , bnj , vq P Λ and therefore pα, β, vq is the limit of a sequence of elements of Λ. Since α, β P T and v P C are arbitrary, Λ is dense in T T C. Lemma 3.3. Let u, w, z P T be independent and tRnunPN be a sequence of positive numbers such that Rn Ñ 8. Then there exist a, b, c P R` such that tRnpaun ` bwn ` cznq : n P Z`u is dense in C. Proof. Let A " tpa, bq P R2 : a ą 0, b ą 0, a ` b ă 1u. Obviously, A is a non-empty open subset of R2. Clearly, fn : A Ñ C, fnpa, bq " Rnpaun ` bwn ` cznq is continuous. By Theorem U, the proof will be complete if we show that Λ " tpa, b, Rnpaun ` bwn ` p1 ´ a ´ bqznqq : n P Z`, pa, bq P Au is dense in A C. Clearly, M " tpα, β, γq P T3 : α ‰ β, α ‰ γ, β ‰ γu is a dense open subset of T3. Consider F : R2 T3 Ñ C, Fpx, ξq " x1ξ1 ` x2ξ2 ` p1 ´ x1 ´ x2qξ3. Differentiating F , we easily confirm that for every px, ξq P A M , the differentials of both Fpx,q at ξ and Fp, ξq at x have rank 2. (3.1) Furthermore, for every x P A, Fpx, Mq is an open subset of C containing 0. 12 Let x " pa, bq P A and y P C. Pick ξ " pα, β, γq P M such that Fpx, ξq " 0. According to (3.1), a standard implicit function argument yields that for every sequence tξj " pαj , βj , γjqujPN in T3 convergent to ξ and for every sequence tzjujPN in C convergent to 0, there is a sequence txj " paj , bjqujPN in R2 convergent to x such that Fpxj , ξjq " zj for each sufficiently large j P N. (3.2) Rnj Since u, w, z are independent, tpun, wn, znq : n P Z`u is dense in T3. Thus there is a strictly increasing sequence tnjujPN of positive integers such that unj Ñ α, wnj Ñ β and znj Ñ γ. By (3.2), we can pick a sequence txj " paj , bjqujPN in R2 convergent to x such that Fpxj , ξjq " y for each sufficiently large j, where ξj " punj , wnj , znjq. Since A is open, xj P A for each sufficiently large j. By the definition of F , Rnjpaj unj ` bj wnj ` p1 ´ aj ´ bjqznjq " y and therefore paj , bj , yq P Λ for each sufficiently large j. Since paj , bj , yq Ñ pa, b, yq and x " pa, bq P A, y P C are arbitrary, Λ is dense in A C. Lemma 3.4. Let z, w P T be independent, tRnunPN and trnunPN be sequences of positive numbers such that Rn Ñ 8 and Rn rn Ñ 8, U be a non-empty open subset of C and ϕ : U Ñ C be continuous. Then a P U : tpRna ` rnϕpaqqzn ` pRna ` rnϕpaqqwn : n P Nu is dense in C( is a dense Gδ-subset of U . Proof. Clearly, it is enough to prove the statement with U replaced by an arbitrary non-empty open set whose closure is contained in U . Thus, without loss of generality, we can assume that ϕ is bounded: c " suptϕpuq : u P Uu ă 8. Since z, w are independent, we can pick independent s, u P T such that z " us and w " us´1. Then for each n P N, where ρn " rn pRna ` rnϕpaqqzn ` pRna ` rnϕpaqqwn " unRn`pa ` ρnϕpaqqsn ` pa ` ρnϕpaqqs´n Λ " `a, unRn`pa ` ρnϕpaqqsn ` pa ` ρnϕpaqqs´n : a P U, n P N( is dense in U C. . By Theorem U, it suffices to verify that Rn Fix a P Uzt0u and y P Czt0u. Since s, u are independent, tpsn, unq : n P Nu is dense in T2. Hence, we can find a strictly increasing sequence tnkukPN of positive integers such that snk Ñ iaa and unk Ñ y y as k Ñ 8. Since ias´nk Ñ a P U and U is open, there is ε ą 0 such that for each sufficiently large k, is´nkaeix P U for every x P r´ε, εs. Thus for each sufficiently large k, we can define the map ξk : r´ε, εs Ñ R, For every δ P p0, εq, we have ξkpxq " ´ yRnk ` 2 Re`iaeix ` ρnk ϕpis´nkaeixqsnk. min xPr´ε,´δs max xPrδ,εs ξkpxq ě 2a sin δ ´ yRnk ´ cρnk Ñ 2a sin δ ą 0 as k Ñ 8, ξkpxq ď ´2a sin δ ´ yRnk ` cρnk Ñ ´2a sin δ ă 0 as k Ñ 8.. (3.3) By (3.3), ξk takes both positive and negative values for each sufficiently large k. Since each ξk is con- tinuous, the Intermediate Value Theorem guarantees the existence of xk P r´ε, εs such that ξkpxkq " 0. Furthermore, the estimates (3.3) imply that xk Ñ 0 as k Ñ 8. Now we set ak " iaeixk s´nk P U and yk " unk Rnk`pak ` ρnk ϕpakqqsnk ` pak ` ρnk ϕpakqqs´nk P C. Obviously pak, ykq P Λ. Since snk Ñ iaa and xk Ñ 0, we have ak Ñ a. Next, by definition of ak, yk " 2Rnk unk Repaksnk ` ρnk ϕpakqsnkq " 2Rnk unk Repiaeixk ` ρnk ϕpiaeixk s´nkqsnkq. Using the above display together with the equality ξkpxkq " 0 and the definition of ξk, we arrive to yk " unky. Since unk Ñ y , we have yk Ñ y. Thus tpak, ykqu is a sequence in Λ convergent to pa, yq. y Since a P Uzt0u and y P Czt0u are arbitrary, Λ is dense in U C. 13 Lemma 3.5. Let 1 ă r ă R and k, m be distinct integers. Then pz, wq P T2 : tRnpzkn ` zmnq ` rnwn : n P Z`u is dense in C( is a dense Gδ subset of T2. Proof. By Theorem U, it suffices to verify that n ` tm Λ " tpz, w, Rnpzkn ` zmnq ` rnwnq : z, w P T, n P Z`u is dense in T T C. Let a, b P T and y P C. Since k ‰ m, for every n P N, there are sn, tn P T such that sk n " 0 n " 2. Consider the function Gn : T Ñ C, Gnpzq " Rnr´npzk ` zmq ´ r´ny. Then and tk rn ´ yrn ą 1 for each sufficiently large Gnpsnq " r´ny ă 1 for each sufficiently large n and Gnptnq ě 2 Rn n. Thus for n large enough the function Gn must attain the value 1. That is, there is cn P T such that Gnpcnq " dn P T. Hence Rnpck n q ` rndn " y for each sufficiently large n. Now pick zn P T and wn P T such that zn n " y, we have pzn, wn, yq P Λ. Since zn Ñ a, wn Ñ b and a, b P T, y P C are arbitrary, Λ is dense in T T C. n " dn, zn ´ a ă π n and wn ´ b ă π n . Since Rnpznk n q ` rnwn n " cn, wn n ` znm n ` sm n ` cm 3.2 Proof of Proposition 1.16 Let UR " pz, wq P T2 : tRnpzn ` wnq : n P Z`u is dense in C(. We have to show that UR is a dense Gδ-subset of T2 of Lebesgue measure 0. By Theorem U, in order to show that UR is a dense Gδ-subset of T2, it suffices to verify that Λ " tpz, w, Rnpzn ` wnqq : z, w P T, n P Z`u is dense in T T C. Let y P C and z, w P T. Since T ` T Ą 2D, for every n satisfying 2Rn ą y, we can find an, bn P T such n " bn, zn ´ z ă π that Rnpan ` bnq " y. For each such n we can pick zn, wn P T for which zn and wn ´ w ă π nq, we have pzn, wn, yq P Λ for every sufficiently large n. Since pzn, wn, yq Ñ pz, w, yq and y P C, z, w P T are arbitrary, Λ is dense in T T C. By Theorem U, UR is a dense Gδ-subset of T2. n . Since y " Rnpan ` bnq " Rnpzn n " an, wn n ` wn n Now let µ be the normalized Lebesgue measure on T2. A straightforward computation yields µtpz, wq P T2 : zn ` wn ď cu " 2 π arcsin` c 2 for every c P r0, 2s and n P N. (3.4) Pick r P p1, Rq. By (3.4), T2, zn ` wn ą rn Rnzn ` wn Ñ 8 is incompatible with the membership in UR, µpURq " 0, as required. Rn( ă 8. Hence for almost all pz, wq P Rn for all sufficiently large n and therefore Rnzn ` wn Ñ 8. Since the condition µ pz, wq P T2 : zn ` wn ď rn 8řn"1 3.3 Infinite sums and integrals Recall that A Ď Z is called syndetic if there is a finite subset B of Z such that A ` B " Z. Equivalently, A is syndetic if A " tnk : k P Zu, where nk`1 ą nk for each k P Z and tnk`1 ´ nk : k P Zu is bounded. Lemma 3.6. Let G be a compact topological group, U be an open subset of G and g P G. Then the set A " tn P Z : gn P Uu is either empty or syndetic. In particular, if U contains the identity element, then A is always syndetic. Proof. Since G is a compact topological group, the subgroup H " tgm : m P Zu is pre-compact with respect to the induced topology. If V " U X H is empty, then A " ∅. Otherwise, V is a non-empty open subset of H. Hence we can pick m P Z such that gmV is a neighborhood of the identity in H. Since H is pre-compact, there is a finite set C " tgn1 , . . . , gnku Ď H such that CgmV " H. It follows that B ` A " Z, where B " tn1 ` m, . . . , nk ` mu. Hence A is syndetic. Lemma 3.7. Let M be a subset of T such that 1 is an accumulation point of M , A be a syndetic subset of Z and k P N. Then tpzn kq : n P A X N, z1, . . . , zk P Mu is dense in Tk. 1 , . . . , zn 14 Proof. Let u " pu1, . . . , ukq be an arbitrary point in Tk. Then we can pick θj P R such that uj " eθj i for 1 ď j ď k. Fix temporarily δ P p0, 1{2q. Since A is syndetic, there is q P N such that A ` t1, . . . , qu " Z. Since 1 is an accumulation point of M , we can pick b1pδq, . . . , bkpδq P R such that 0 ă bkpδq ă δ, 0 ă bj´1pδq ă δbjpδq for 2 ď j ď k and ebjpδqi P M for 1 ď j ď k. First, we choose the smallest positive integer m1 such that distpm1b1pδq ´ θ1, 2πZq ă b1pδq. Then we pick the smallest integer m2 greater than m1 such that distpm2b2pδq´θ2, 2πZq ă b2pδq. We proceed the same way. Namely, having chosen m1, . . . , mj´1, we set mj to be the smallest integer greater than mj´1 such that distpmj bjpδq ´ θj, 2πZq ă bjpδq. By definition of q, we can pick n P A such that mk ď n ď mk ` q ´ 1. Clearly, mj ´ mj´1 ď 2π bjpδq for 2 ď j ď k. Hence k´j´2řm"1 δm ď q ´ 1 ` 0 ď n ´ mj ď q ´ 1 ` křm"j`1 2π bkpδq ď q ´ 1 ` 2π bjpδq 2πδ bjpδqp1´δq for 1 ď j ď k, where the third inequality is derived from bm´1pδq ă δbmpδq. Hence nbjpδq ´ mjbjpδq ď pq ´ 1qbjpδq ` 2πδ 1´δ ď δpq ´ 1 ` 4πq for 1 ď j ď k j ă ε for 1 ď j ď k. j ´ uj ă ε and z´mzm j ´ u2 since bjpδq ă δ. Since distpmj bjpδq´θj , 2πZq ă bjpδq, we have uj´emj bjpδqi ă bjpδq ă δ. According to the above display, enbjpδqi ´ emj bjpδqi ă δpq ´ 1 ` 4πq. Hence uj ´ zn j ă δpq ` 4πq, where zj " ebjpδqi. Since zj P M , n P A X N and δ P p0, 1{2q is arbitrary, the desired result follows. Corollary 3.8. Let M be an infinite subset of T, A be a syndetic subset of Z, k, n0 P N and u " pu1, . . . , ukq P Tk. Then for every ε ą 0, there exist n, m P A and z, z1, . . . , zk P M such that m ą n ą n0, z´nzn Proof. Since A is syndetic, there is q P N such that A ` t1, . . . , qu " Z. Since T is compact and M is infinite, there is an accumulation point v P T of M . Then for each δ P p0, 1q, 1 is an accumulation point of Mδ " tv´1z : z P M, z ´ v ă δu. By Lemma 3.7, there are n P A and w1, . . . , wk P Mδ such that n ą n0 and wn v with zj P M and zj ´ v ă δ. Since v j ´ uj ă δ, wj " zj is an accumulation point of M , we can find z P M such that z ´ v ă δ and z ´ v ă δ j ă 4δ for 1 ď j ď k. Now pick m P A X t2n, 2n ` 1, . . . , 2n ` q ´ 1u. Since z ´ zj ď z ´ v ` v ´ zj ă 2δ, we have z´2nz2n j ´ uj ă 2δ for 1 ď j ď k. It follows that z´2nz2n ´ zm´2n ď pm ´ 2nqzj ´ z ă 2qδ. Hence j ´ u2 Thus for δ satisfying p2q ` 4qδ ă ε, z´nzn j ´ uj ă δ for 1 ď j ď k. Since wj P Mδ, wj " zj 2n , we have z´nzn j " zm´2n j ´ u2 j ă 4δ ` 2qδ " p2q ` 4qδ. j ´ u2 j ă ε for 1 ď j ď k, as required. j ` z´2nz2n j ´ z´mzm j ´ uj ă ε and z´mzm j ´ z´mzm z´mzm 2n . Since wn j j ă z´2nz2n v j ´ u2 The following lemma looks painfully technical. We need it nonetheless. For each ε ą 0, let for 1 ď j ď 4u. 5 uj ´ e2πji{5 ă ε and u´1 10 u5`j ´ e4πji{5 ă ε Pε " tu P T10 : u´1 (3.5) Lemma 3.9. There exist ε, d ą 0 such that for every u P Pε and every z, w P d D, there are positive numbers a1, . . . , a5 satisfying a1 ` . . . ` a5 " 1, a1u1 ` . . . ` a5u5 " z and a1u6 ` . . . ` a5u10 " w. Proof. For every u P C10, consider the R-linear map Su : R4 Ñ C2 given by the formula Sua "´ 4řj"1puj ´ u5qaj , 4řj"1pu5`j ´ u10qaj¯. Let ξ " e2πi{5 and for α, β P T, let upα, βq " pαξ, αξ2, αξ3, αξ4, α, βξ2, βξ4, βξ, βξ3, βq P T10. Then Supα,βq " STα,β, where Tα,β : C2 Ñ C2, Tαpz, wq " pαz, βwq and S : R4 Ñ C2, Sa " pξ ´ 1qa1 ` pξ2 ´ 1qa2 ` pξ3 ´ 1qa3 ` pξ4 ´ 1qa4 pξ2 ´ 1qa1 ` pξ4 ´ 1qa2 ` pξ ´ 1qa3 ` pξ3 ´ 1qa4 . 15 First, observe that Tα is an isometry. Identifying C2 with R4 in a natural way, we see that S is the linear operator on R4 with the matrix S " cosp2π{5q ´ 1 cosp4π{5q ´ 1 cosp6π{5q ´ 1 cosp8π{5q ´ 1 cosp4π{5q ´ 1 cosp8π{5q ´ 1 cosp2π{5q ´ 1 cosp6π{5q ´ 1 sinp2π{5q sinp4π{5q sinp4π{5q sinp8π{5q sinp6π{5q sinp2π{5q sinp8π{5q sinp6π{5q ‹‹‚. One can verify (with some effort by hand and on the spot using an appropriate software) that the matrix in the above display is invertible. Thus Supα,βq is invertible for each α, β P T. Since the set of invertible R-linear operators from R4 to C2 is open, there is ε0 ą 0 such that Su is invertible for every u P Pε0. Thus for each u P Pε0 and every pz, wq P C2, there exists a unique apu, z, wq P R4 such that Suapu, z, wq " pz ´ u5, w ´ u10q. Obviously, the map pu, z, wq ÞÑ apu, z, wq is continuous on Pε0 C2. It 5. is also straightforward to verify that Supα,βq` 1 5 , 1 By continuity, there exist ε P p0, ε0q and d ą 0 such that whenever u P Pε and z, w P d D, the components of apu, z, wq are positive and sum up to a number less than 1. Now for z, w P dD and u P Pε, we set 4řj"1 aj " ajpu, z, wq for 1 ď j ď 4 and a5 " 1 ´ ajpu, z, wq. It is straightforward to see that the equality Suapu, z, wq " pz ´ u5, w ´ u10q is equivalent to a1u1 ` . . . ` a5u5 " z and a1u6 ` . . . ` a5u10 " w, which completes the proof. 5 " p´α,´βq. Hence apupα, βq, 0, 0q "` 1 5 , 1 5 , 1 5 , 1 5 , 1 5 , 1 ÿzPM Lemma 3.10. Let M be an infinite countable subset of T, ω : R` Ñ R` be a strictly increasing continuous concave function vanishing at 0 and tRnunPN be a sequence of positive numbers such that Rn Ñ 8. Then there exists a P ℓ1 `pMq such that Proof. First, note that QpMq " !a P ℓ1 řzPM azzk : k P Z`) is dense in C. ωpazq ă 8 and !Rk ÿzPM ωpazq ă 8) equipped with the metric ρpa, bq " `pMq : řzPM ωpaz ´ bzq is a separable complete metric space. Let Ω ""a P QpMq : lim inf kÑ8 ÿzPM "a P QpMq :ÿzPM 8ďk"m Then Ω is a non-empty (0 P Ω) Gδ-subset of the Polish space QpMq (it is by no means dense!). Indeed, azzk " 0*. n*. azzk ă azzk " 0 for some k P N). Ω " čm,nPN Ω0 "!a P QpMq : ÿzPM First, we observe that Ω0 is a subset of Ω. Indeed, let a P Ω0. Then there is k P N such that řzPM azzk " 0. Consider the compact metrizable topological group G " TM and g P G defined by gpzq " z for z P M . By Lemma 3.6, for every neighborhood U of 1G in G there are infinitely many n P N such that gn P U . Hence there is a strictly increasing sequence tnjujPN of positive integers such that gnj Ñ 1G. Then Let also 1 lim jÑ8 ÿzPM azzk`nj " ÿzPM azzk " 0 and therefore a P Ω. Let d, ε ą 0 be the numbers provided by Lemma 3.9. First, we shall verify the following claim: For each a P Ω, q P N, δ ą 0 and y P δd D, there exist n P N and b P QpMq such that n ą q, ρpb, 0q ă 5ωpδq, a ` b P Ω0 and řzPMpaz ` bzqzn " y. (3.6) 16 By Lemma 3.9, there exist positive numbers β1, . . . , β5 such that azzn ă yδ ` 1 δpδd ´ yq ă d. δpδd ´ yq " d, y azzn ď yδ ` 1 δ řzPM δ řzPM δ ´ 1 ´ 1 azzm " 1 azzm ă 1 δ řzPM δ řzPM δ ÿzPM 5ÿj"1 y δ ´ j " βj zn 1 βj " 1, 5ÿj"1 Let a P Ω, q P N, δ ą 0 and y P δdD. Clearly, θ " δd ´ y ą 0. Denote A ""n P Z : řzPM U " !f P G : řzPM azzn ă θ*. azfpzq ă θ). Since U is an open subset of the compact topological group G, Since a P Ω, the set A is infinite and therefore non-empty. Clearly, A " tn P Z : gn P Uu, where Lemma 3.6 implies that A is syndetic. Let ξ " e2πi{5. By Corollary 3.8, there exist m, n P A and z, z1, . . . , z5 P M such that m ą n ą q and zn j ´ zmξ2j ă ε{2 for 1 ď j ď 5. Then 1 , . . . , zn pzn 5 q belongs to the set Pε defined in (3.5). Since n, m P A, we have j ´ znξj ă ε{2, zm 1 , . . . , zm 5 , zm azzn and 5ÿj"1 βj zm j " ´ 1 δ ÿzPM azzm. (3.7) Define b P QpMq by setting bz " 0 if z P Mztz1, . . . , z5u and bzj " δβj for 1 ď j ď 5. The equation (3.7) now reads řzPMpaz ` bzqzn " y and řzPMpaz ` bzqzm " 0. The latter implies a ` b P Ω0. Since the non-zero components of b sum up to δ and there are only 5 of them ρpb, 0q ă 5ωpδq. This concludes the proof of Claim (3.6). Since Ω is a non-empty Gδ-subset of a Polish space QpMq, Ω is a Polish space in its own right. Since for every k P N, the map fk : Ω Ñ C, fkpaq " Rk řzPM azzk is continuous, Theorem U guarantees that the proof will be complete if we show that Λ ""´a, Rk ÿzPM azzk¯ : a P Ω, k P N* is dense in Ω C. Since Rn Ñ 8, (3.6) yields that for every δ ą 0, y P C and a P Ω, there are n P N and b P ℓ1 that ρpb, 0q ă δ (and therefore ρpa, a ` bq ă δ), a ` b P Ω and řzPMpaz ` bzqzn " y Since a P Ω, y P C and δ ą 0 are arbitrary, Λ is dense in Ω C, which completes the proof. Lemma 3.11. Let pΩ, F , µq be a measure space. Assume that the essential range of g P L8pµq contains a sequence tλnunPN such that λ1 ą 1, tλnunPN is strictly increasing and λn λn there exist a non-negative real valued function a P L1pµq for which `pMq such Rn . Hence pa ` b, yq P Λ. are pairwise distinct. Then "żΩ gka dµ : k P Z`* is dense in C. Applying the above lemma to the situation of a purely atomic measure space, we immediately obtain the following corollary. Corollary 3.12. Let tλnunPN be a bounded sequence in C such that λ1 ą 1, the sequence tλnunPN is strictly increasing and the numbers λn λn `pNq such that are pairwise distinct. Then there exists a P ℓ1 n : k P Z`) is dense in C. anλk ! 8ÿn"1 Proof of Lemma 3.11. Let rn " λn, zn " λn and r " lim rn. Let Ω0 " tω P Ω : gpωq ă ru and Z be λn the set of all real-valued non-negative f P L1pµq vanishing outside Ω0. Clearly, Z is a closed subset of 17 the Banach space L1pµq. Since for every k P N, the map Fk : Z Ñ C, Fkpfq " şΩ gkf dµ is continuous, Theorem U guarantees that the proof will be complete if we show that Λ ""´a,ż gkf dµ¯ : a P Z, k P N* is dense in Z C. Let U be an arbitrary non-empty open subset of Z C. Since rn Ñ r, we can find pa, yq P U and m ě 2 such that a vanishes outside tω P Ω : gpxq ă rm´1u. Clearly M " tzj : j ą mu is an infinite m´1. Since rm´1 ă rm, we can find subset of T. Since g ď rm´1 on the support of a, ş agk dµ ď }a}1rk 3 and therefore rq m´1}a}1 q P N such that m ` y m ă 1 rq rq y m ´ rk m żΩ 1 rk 1 3 D for each k ě q. An elementary geometric argument ensures that there is ε ą 0 (one can even find the biggest ε) such that 3 D is contained in the convex hull of u, v and w. Corollary 3.8 applied with A " Z guarantees that for every integer s ě q there are 3 D is contained s . Since us, vs, ws P M , there are j1,s, j2,s, j3,s P tm ` 1, m ` 2, . . . u s and agk dµ P v ´ e2πi{3 ă ε and w whenever u, v, w P T satisfy u us, vs, ws P M and ks P N such that ks ą s, uks in the convex hull of uks such that us " zj1,s, vs " zj2,s and ws " zj3,s. Since 1 wks s , by the above display, there exist α1, α2, α3 P R` such that j2,s ` α3zks s ´ e2πi{3 ă ε and uks s ´ e4πi{3 ă ε. Thus 1 3 D is contained in the convex hull of uks y m ´ rks m żΩ 1 rks agks dµ. α1 ` α2 ` α3 " 1 and α1zks j1,s ` α2zks v ´ e4πi{3 ă ε, the disk 1 s , vks s and wks j3,s " s , vks s vks s vks Using the definitions of rn and zn and multiplying by rks m , we get α1rks m rks j1,s λks j1,s ` α2rks m rks j2,s λks j2,s ` j3,s `żΩ λks α3rks m rks j3,s agks dµ " y. Note that for every k P N, the set "żΩ bgk : b P Z, }b} " 1* is convex and is a dense subset of the closed convex hull of the set tzk : z is in the essential range of g and z ă ru. Thus there is bs P Z such that and (3.8) (3.9) }bs}1 " bsgks ´ α1rks m rks j1,s λks j1,s ` żΩ α1rks m rks j1,s α2rks m rks j2,s ` α3rks m rks j3,s ` α2rks m rks j2,s λks j2,s ` α3rks m rks j3,s λks j3,s ă 2´s. The above display together with (3.8) gives ys "żΩpas ` bsqgks dµ Ñ y as s Ñ 8. Since rl,s ą rm for 1 ď l ď 3, (3.9) implies that }bs}1 Ñ 0 as s Ñ 8. By definition of ys, pa ` bs, ysq P Λ for each s ě q. Since a ` bs Ñ a, ys Ñ y and U is open, Λ meets U . Since U is an arbitrary open subset of Z C, Λ is dense in Z C, which completes the proof. 18 For the sake of brevity, we use the following fairly standard notation: pµpnq "żT zn dµpzq for n P Z for a Borel measure µ on T. As usual, for g P L1pµq, gµ stands for the measure absolutely continuous with respect to µ with the density (=Radon -- Nykodim derivative) g. For a non-negative Borel measure on T, we denote the set of all real-valued non-negative f P L1pµq by the symbol L1 Lemma 3.13. Let µ be a purely non-atomic Borel probability measure on T and tRnunPN be a sequence of positive numbers such that Rn Ñ 8. Then there is a P L1 `pµq such that tRnxaµpnq : n P Nu is dense in C. `pµq. Proof. We start by proving the following claim. For every purely non-atomic Borel probability measure ν on T (3.10) By the Fubini theorem, and each ε ą 0, the set tn P Z : pνpnq ă εu is syndetic. w´k dνpwq "żTT´ z w¯k zk dνpzqżT dpν νqpz, wq. pνpkq2 "pνpkqpνpkq "żT m`n´1ÿk"m 1 n For each m P Z and n P N, we sum up these equalities for m ď k ď m ` n ´ 1: pνpkq2 "żTT n´ z w¯m n´1ÿk"0 For each δ P p0, 2q, we can split the above integral: where hm,npz, wq " 1 hm,npz, wq dpν νqpz, wq, zmpwn ´ znq zk wk " n wm`n´1pw ´ zq . hm,npz, wq dpν νqpz, wq `żBδ hm,npz, wq dpν νqpz, wq, 1 n m`n´1ÿk"m pνpkq2 "żAδ żBδ . 2 δn 2 δnpν νqpBδq ď where Aδ " tpz, wq P T2 : z ´ w ă δu and Bδ " tpz, wq P T2 : z ´ w ě δu. Note that w ´ z ě δ and wn ´ zn ď 2 for pz, wq P Bδ. Hence hm,npz, wq ď 2 δn for pz, wq P Bδ. Thus hm,npz, wq dpν νqpz, wq ď On the other hand, for every z, w P T and n P N, hm,npz, wq, being the average of several elements of T, satisfies hm,npz, wq ď 1. Hence hm,npz, wq dpν νqpz, wq ď pν νqpAδq. żAδ m`n´1ÿk"m pνpkq2 ď δÑ0pν νqpAδq " 0. Thus we can find δ P p0, 2q such that Since ν is purely non-atomic, we have lim 2 δn ` pν νqpAδq. Combining the last three displays, we get 1 n pν νqpAδq ă ε2 2 . Having δ fixed, we pick n P N such that 2 δn ă ε2 2 . By the above display, 1 n m`n´1ÿk"m pνpkq2 ă ε2 2 ` ε2 2 " ε2 for every m P Z. 19 We also need the following fact. According to the above display, among n consecutive numbers pνpkq there is at least one strictly less than ε. Hence the set tn P Z : pνpnq ă εu is syndetic, which proves Claim (3.10). Let ν be a Borel probability measure on T with infinite support, r P p0, 1q, m P N, A Ď Z be syndetic, and twnunPA, nąm be a sequence in D. Then there exist a P L1 `pνq and n P A satisfying n ą m, }a}1 " 1 and xaνpnq " rwn. (3.11) zn ´ e2πij{q ă ε{2 for 1 ď j ď q. It follows that zn Let K be the support of ν. For each n P Z denote K n " tzn : z P Ku. It is easy to see that the `pνq, }a}1 " 1u is convex, is contained and is dense in the convex span of K n. set Bn " txaνpnq : a P L1 In particular, Bn contains the interior of the convex span of K n. Since r ă 1, we can pick q P N large enough in such a way that rD is contained in the interior of the convex span of the points e2πij{q for 1 ď j ď q. Then there is ε ą 0 such that rD is contained in the convex span of u1, . . . , uq whenever uj uq ´ e2πij{q ă ε for 1 ď j ď q ´ 1. By Corollary 3.8 we can find n P A and z, z1, . . . , zq P K such that q ´ e2πij{q ă ε for 1 ď j ď q ´ 1. Thus n ą m and zn rD is in the convex span of zn a P L1 `pνq, such that }a}1 " 1 and xaνpnq " rwn, which completes the proof of Claim (3.11). Since each of the maps a ÞÑ Rnxaµpnq from L1 order to complete the proof Lemma 3.13 it suffices to verify that Λ " tpa, Rnxaµpnqq : n P N, a P L1 `pµq, z P C and ε ą 0. By (3.10), A " n P Z : xaµpnq ă ε Rnε ´ xaµpnqε q . Hence rD Ă Bn. In particular, rwn P Bn and therefore there is `pµq to C is continuous, Theorem U ensures that in for every n P A satisfying n ą n0, where wn " z wn P 1 `pνq and n P A satisfying n ą n0, }β}1 " 1 and xβµpnq " wn. Using the definition By (3.11), there is β P L1 of wn, we can rewrite the last equality as xbµpnq " z, where b " a ` εβ. Then pb, zq P Λ and }b ´ a}1 " ε. Since ε ą 0 is arbitrary, pa, zq belongs to the closure of Λ. Since a P L1 `pµq and z P C are arbitrary, Λ is dense in L1 Rn Ñ 8, there is n0 P N such that z 4( is syndetic. Since `pµqu is dense in L1 `pµq C. 4 for each n ě n0. Thus Fix arbitrary a P L1 Rnε ă 1 2 D, j 1 , . . . , zn j zn . `pµq C, which completes the proof. 4 Proof of Theorems 1.9 and 1.10 We start with the following lemma, which must be known. We include the proof for the reader's conve- nience and since we were unable to find it in the literature anyway. Lemma 4.1. Let te1, . . . , enu be a basis in an n-dimensional Banach space X and te1 , . . . , enu be the dual basis in X : ej pekq " δj,k. Then for every c P Rn ` such that c1 ` . . . ` cn " 1, there is px, fq P ΠpXq for which ejpxqfpejq " cj for 1 ď j ď n. Proof. It is easy to check that tpe1pxqfpe1q, . . . , enpxqfpenqq : px, fq P ΠpXqu is a closed subset of Cn. Hence it is enough to prove the result in the case cj ą 0 for every j. From now on we shall just assume that cj ą 0 for 1 ď j ď n. Consider the continuous map F : X Ñ r´8,`8q, Fpwq " cj log ej pwq. nřj"1 Since BpXq is compact and the map t ÞÑ Fptxq increases on r0, 1s for each x P SpXq, FBpXq maximal value at some point x P SpXq. Obviously, ej pxq ‰ 0 for 1 ď j ď n. Consider attains its f P X, fpyq " cj e j pyq e j pxq for y P X. nřj"1 20 Clearly, fpxq " 1 and ej pxqfpejq " cj for 1 ď j ď n. In order to clinch the proof it suffices to show that f is bounded by 1 on SpXq. Indeed, then px, fq P ΠpXq and px, fq satisfies all desired conditions. Assume the contrary. Then there is y P SpXq such that fpyq is real and fpyq ą 1. Clearly, x ` tu P BpXq for every t P r0, 1s, where u " y ´ x. Moreover, fpuq " fpyq ´ fpxq " fpyq ´ 1 ą 0. Since FBpXq attains its maximum at x, Fpx ` tuq ď Fpxq for t P r0, 1s and therefore Fpx`tuq´Fpxq t ď 0. lim sup tÑ0` e j pxq e j pxq let C´ " CzzR` and X´ " tw P X : ej pwq P Choose z P T other than 1 or each of C´ for 1 ď j ď nu. Set log to be the standard (real on the positive half of the real line) branch of the log- cj log ejpwq. Since x` tu P X´ for each sufficiently small t, the function Hptq " Gpx` tuq is defined in a neighborhood of 0. Differentiating, we obtain is exactly the real part of the function G : X´ Ñ C, Gpwq " arithm on C´. Then FX´ nřj"1 , Fpx`tuq´Fpxq t " Re H1p0q " Re lim tÑ0 " Re nřj"1" cj uj e j pxq`te dt logpej pxq ` tejpuqqıt"0 d nřj"1"cj j puqıt"0 " Re nřj"1 cj e e j puq j pxq " Re fpuq " fpuq ą 0. The obvious contradiction between the last two displays completes the proof. 1 ` . . . ` cnλk Lemma 4.2. Let T P LpXq, λ1, . . . , λn P σppTq and c1, . . . , cn P R` satisfy c1 ` . . . ` cn " 1. Then there exists px, fq P ΠpXq such that fpT kxq " c1λk Proof. Without loss of generality we can assume that λj are pairwise distinct. Choose ej P SpXq such that T ej " λjej for 1 ď j ď n. Since λj are pairwise distinct, ej form a basis of the n-dimensional space E " spante1, . . . , enu. Let e1 , . . . , en be the dual basis in E: ekpejq " δi,j. By Lemma 4.1, there is px, gq P ΠpEq such that ej pxqgpejq " cj for 1 ď j ď n. By the Hahn -- Banach Theorem, there is f P X such that fE " g and }f} " 1. Since }x} " 1 and fpxq " gpxq " 1, px, fq P ΠpXq. Finally, using the equalities ejpxqgpejq " cj, we get n for every k P Z`. fpT kxq " gpT nxq " gpe1pxqλk 1 e1 ` . . . ` enpxqλk nenq " c1λk 1 ` . . . ` cnλk n for each k P Z`. 4.1 Proof of Theorem 1.9 Let T P LpXq and λ1, λ2 P σppTq be such that λ1 " λ2 " R ą 1 and z " λ1 R are independent in T. Then we can pick linearly independent e1, e2 P X such that T e1 " λ1e1 and T e2 " λ2e2. By Lemma 3.1, there are a, b P T for which R , w " λ2 O " tRnpazn ` bwnq : n P Z`u " taλn 1 ` bλn 2 : n P Z`u is dense in C. By the Hahn -- Banach Theorem, there is f P X such that fpe1q " a and fpe2q " b. Then OpT, e1`e2, fq " 1 ` bλn taλn 2 : n P Z`u is dense in C. By Proposition 1.5, T is weakly numerically hypercyclic. That is, (1.9.1) is sufficient for the weak numeric hypercyclicity of T . Assume that T P LpXq satisfies (1.9.2). That is, there exist independent λ1, λ2 P T such that kerpT ´ λ1Iq2 ‰ kerpT ´ λ1Iq and kerpT ´ λ2Iq2 ‰ kerpT ´ λ2Iq. An elementary linear algebra argument allows us to pick linearly independent e1, e2, e3, e4 P X such that pT ´λ1Iqe1 " pT ´λ2Iqe3 " 0, pT ´ λ1Iqe2 " λ1e1 and pT ´ λ2Iqe4 " λ2e3. Then T ne2 " λn 2pe4 ` ne3q for every n P N. By Lemma 3.1, there are a, b P T such that tnpaλn 2q : n P Nu is dense in C. By the Hahn -- Banach Theorem, there is f P X for which fpe4q " fpe2q " 0, fpe1q " a and fpe3q " b. Then fpT npe2 ` e4qq " npaλn 2q for every n P N. Hence OpT, e2 ` e4, fq is dense in C and T P WNHpXq according to Proposition 1.5. Thus (1.9.2) is sufficient for the weak numeric hypercyclicity of T . 1pe2 ` ne1q and T ne4 " λn 1 ` bλn 1 ` bλn 21 λ2 has infinite order in T and s " λ1 Assume that T P LpXq satisfies (1.9.3). Then there exist λ1, λ2, λ3 P σppTq such that λ1 " λ2 " r are independent. If s and u " λ2 R ą λ3 " r ą 1, λ1 R are independent, then (1.9.1) is satisfied and therefore T is weakly numerically hypercyclic. It remains to consider the case when s and u are not independent. In this case there are z, f, g P T and l, j P Z such that f and g have finite order, s " f zl and u " gzj and l ě j. Since s u has infinite order, we have l ‰ j and therefore l ą j. Since s and y are independent, z and y are independent. Since f and g have finite order, there is an even d P N such that f d " gd " 1. Then sdn " zknzmn, udn " z´knzmn and ydn " wn for every n P Z`, where k " d 2pl ` jq P Z and w " yd P T. Since z and y are independent, z and w are independent as well. By Lemma 3.2, there exist a, b P T such that 2pl ´ jq P N, m " d R , y " λ3 O " tRdnzmnpazkn ` a´1z´knq ` brdnwn : n P Z`u is dense in C. Since λ1, λ2 and λ3 are distinct eigenvalues of T , we can pick linearly independent e1, e2, e3 P X such that T e1 " λ1e1, T e2 " λ2e2 and T e3 " λ3e3. By the Hahn -- Banach Theorem, we can find h P X for which hpe1q " a, hpe2q " a´1 and hpe3q " b. Then for each n P Z`, hpT dnpe1 ` e2 ` e3qq " aλdn 1 ` a´1λdn 2 ` bλdn 3 " Rdnzmnpazkn ` a´1z´knq ` brdnwn. Thus O is a subset of OpT, e1 ` e2 ` e3, hq. Hence OpT, e1 ` e2 ` e3, hq is dense in C and T P WNHpXq according to Proposition 1.5. That is, (1.9.3) is sufficient for the weak numeric hypercyclicity of T . Finally, assume that T P LpXq satisfies (1.9.4). That is, there exist λ1, λ2 P σppTq and λ3 P T such that λ1 " λ2 " R ą 1, λ1 R , λ3 are independent and kerpT ´ λ3Iq2 ‰ kerpT ´ λ3Iq. Exactly as in the proof of the previous part, we can find d, k P N, m P Z and independent z, w P T such that sdn " zknzmn, udn " z´knzmn and y " wn for every n P Z`, where u " λ2 R and y " λ3. By Lemma 3.2, there exist a, b P T for which has infinite order in the group T, s " λ1 λ2 O " tRdnzmnpazkn ` a´1z´knq ` bdnwn : n P Z`u is dense in C. Pick linearly independent e1, e2, e3, e4 P X such that T e1 " λ1e1, T e2 " λ2e2, pT ´ λ3Iqe3 " 0 and pT ´λ3Iqe4 " λ3e3. By the Hahn -- Banach Theorem, we can find h P X for which hpe1q " a, hpe2q " a´1, hpe3q " b and hpe4q " 0. Then for each n P Z`, 1 ` a´1λdn 3 " Rdnzmnpazkn ` a´1z´knq ` bdnwn. hpT dnpe1 ` e2 ` e4qq " aλdn 2 ` bdnλdn Thus O is a subset of OpT, e1 ` e2 ` e4, hq. Hence OpT, e1 ` e2 ` e4, hq is dense in C and T P WNHpXq according to Proposition 1.5. Thus (1.9.4) is sufficient for the weak numeric hypercyclicity of T . The proof of Theorem 1.9 is complete. 1 ` . . . ` cnλk 4.2 Proof of Theorem 1.10 Assume that T P LpXq satisfies (1.10.1). By Lemma 4.2, there is px, fq P ΠpXq such that fpT kxq " c1λk n for each k P Z`. Hence OpT, x, fq is dense in C and T P NHpXq. Since every operator similar to T satisfies the same conditions, T P SNHpXq. Thus (1.10.1) implies the strong numeric hypercyclicity of T . For further references we prove the following slight modification of this result. Lemma 4.3. Let T P LpCnq be the diagonal operator with the numbers λ1, . . . , λn on the diagonal. Then T P NHpCnq if and only if T P SNHpCnq if and only if there exist c1, . . . , cn P R` such that tc1λk Proof. If there exist c1, . . . , cn P R` such that tc1λk n : k P Z`u is dense in C, (1.10.1) is satisfied, which implies that T P SNHpCnq and therefore T P NHpCnq. Now assume that T P NHpCnq. Then there is x P SpCnq for which NOpT, xq is dense in C. A direct calculation shows that NOpT, xq " tx12λk n : k P Z`u is dense in C with cj " xj2. n : k P Z`u. That is, tc1λk n : k P Z`u is dense in C. 1 ` . . . ` xn2λk 1 ` . . . ` cnλk 1 ` . . . ` cnλk 1 ` . . . ` cnλk Assume that T P LpXq satisfies (1.10.2). Then there exist λ1, λ2, λ3 P σppTq such that λ1 " λ2 " R are independent. By Lemma 3.3, there are c1, c2, c3 P R` such that 1 ` c2λk 3 : k P Z`u is dense in C. Thus (1.10.1) is satisfied and T P SNHpXq. That is, (1.10.2) λ3 " R ą 1 and λ1 tc1λk implies the strong numeric hypercyclicity of T . The proof of Theorem 1.10 is complete. 2 ` c3λk R , λ2 R , λ3 22 5 Proof of Propositions 1.11 and 1.12 Throughout this section H is a Hilbert space and T P LpHq. 5.1 Proof of Proposition 1.11 Assume that λ1, λ2 P σppTq are such that λ1 " λ2 " R ą 1, the eigenspaces kerpT ´ λ1Iq and kerpT ´ λ2Iq are non-orthogonal and λ1 R are independent in T. Pick x, y P SpHq such that T x " λ1x, T y " λ2y and xx, yy " c ą 0. Note that there is ε " εpcq ą 0 such that the set 1`cu 1`cu´1 : u P T( contains Jε " teiθ : ´ε ă θ ă εu. By Lemma 3.1, there is a P Jε such that the set O " tλn ` aµn : n P Z`u is dense in C. Since a P Jε, there is u P T such that a " 1`cu }x`uy} P SpHq and denote b " }x ` uy}´2. Then for each n P Z`, 1`cu´1 . Consider the vector q " x`uy R , λ2 xT nq, qy " bxλnx ` uµny, x ` uyy " bpp1 ` cuqλn ` p1 ` cu´1qµnq " bpp1 ` cu´1qqpλn ` aµnq. Thus the density of O implies the density of NOpT, qq and therefore T P NHpC2q. This completes the proof of Proposition 1.11. , λ2 λ2 5.2 Proof of Proposition 1.12 Assume that λ1, λ2 P C are such that kerpT ´ λj Iq2 ‰ kerpT ´ λjIq for j P t1, 2u, λ1 " λ2 ě 1 and λ1 are independent in T. Since kerpT ´ λj Iq2 ‰ kerpT ´ λj Iq, there are two 2-dimensional λ1 T -invariant subspaces E1 and E2 of H such that E1 X E2 " t0u, pT ´ λj Iq2Ej " 0 and pT ´ λj IqEj ‰ 0 for j P t1, 2u. In particular, Lj " Ej X kerpT ´ λj Iq are one-dimensional for j P t1, 2u. If L1 and L2 are non-orthogonal, T P NHpHq according to Proposition 1.11. It remains to consider the case when L1 and L2 are orthogonal. It is straightforward to see that there are e1, . . . , e4 P SpHq such that e1 P L1, e3 P L2, te1, e2u is a linear basis in E1, te3, e4u is a linear basis in E2, xe1, e2y " xe1, e3y " xe3, e4y " 0, while g1,4 " xe1, e4y, g2,3 " xe2, e3y and g2,4 " xe2, e4y are non-negative real numbers. The above properties of the restrictions of T to E1 and E2 yield the existence of non-zero α, β P C such that T e1 " λ1e1, T e2 " λ1pe2 ` αe1q, T e3 " λ2e3 and T e4 " λ2pe4 ` βe3q. Then 1`a2 ` b2 ` g1,4ad ` g2,3bc ` g2,4bd ` αnbpa ` g1,4dq xT nx, xy " λn 2`c2 ` d2 ` g1,4da ` g2,3cb ` g2,4db ` βndpc ` g2,3bq ` λn for a, b, c, d P C and n P N, where x " ae1 ` be2 ` ce3 ` de4. (5.1) Consider the degree 2 R-polynomial map Φ : C4 Ñ R2 C2, Φ a b c d ‹‹‚" Repa2 ` b2 ` g1,4ad ` g2,3bc ` g2,4bdq Repc2 ` d2 ` g1,4da ` g2,3cb ` g2,4dbq αbpa ` g1,4dq βdpc ` g2,3bq ‹‹‚. First, observe that Φp1, 0, 1, 0q " p1, 1, 0, 0q. Looking at Φ as at a function of 8 real variables Re a, Im a, Re b, Im b, Re c, Im c, Re d and Im d taking values in R6 with the components ordered as Φ1, Φ2, Re Φ3, Im Φ3, Re Φ4, Im Φ4, we can calculate the Jacobi matrix d Φ of Φ at the point p1, 0, 1, 0q: 0 0 g2,3 g2,3 2 0 0 0 0 0 2 0 0 0 Re α ´Im α 0 0 0 0 0 0 Im α Re α 0 0 Re β ´Im β 0 0 0 0 Im β Re β 0 0 g1,4 g1,4 0 0 0 0 0 0 0 0 0 0 ‹‹‹‹‹‹‚ . d Φp1, 0, 1, 0q " Since the complex numbers α and β are non-zero, it is an easy exercise to see that the rank of the above 6 by 8 matrix is 6. Indeed, after removing the two zero columns, we are left with an invertible 6 by 6 23 matrix. The Implicit Function Theorem says that there is ε ą 0 and a smooth (even real-analytic) map Ψ : W " p1´ ε, 1` εqp1´ ε, 1` εq εD εD Ñ C4 such that Ψp1, 1, 0, 0q " p1, 0, 1, 0q and Φ Ψ " IdW . Now we define a, b, c, d : εD Ñ C, papzq, bpzq, cpzq, dpzqq " Ψp1, 1, z, zq. The equality Φ Ψ " IdW yields αbpzqpapzq ` g1,4dpzqq " z, Repapzq2 ` bpzq2 ` g1,4apzqdpzq ` g2,3bpzqcpzq ` g2,4bpzqdpzqq " 1, Repcpzq2 ` dpzq2 ` g1,4dpzqapzq ` g2,3cpzqbpzq ` g2,4dpzqbpzqq " 1 βdpzqpcpzq ` g2,3bpzqq " z, (5.2) (5.3) (5.4) for every z P εD. Denote ϕpzq " apzq2 ` bpzq2 ` g1,4apzqdpzq ` g2,3bpzqcpzq ` g2,4bpzqdpzq for z P εD. Using (5.3) and (5.4), we obtain ϕpzq " cpzq2 ` dpzq2 ` g1,4dpzqapzq ` g2,3cpzqbpzq ` g2,4dpzqbpzq for z P εD. By the above two displays, (5.1) and (5.2), xT nxpzq, xpzqy " λn 1pϕpzq ` nzq ` λn where xpzq " apzqe1 ` bpzqe2` cpzqe3` dpzqe4. Applying Lemma 3.4 with z " λ1 λ1 rn " λ1n and U " εD, we find that there is u P εD such that 2pϕpzq ` nzq for every z P εD and n P N, , w " λ2 λ2 , Rn " nλ1n, O " tλn 1pϕpuq ` nuq ` λn 2pϕpuq ` nuq : n P Nu is dense in C. By the above two displays NOpT, yq, being a positive scalar multiple of O, is dense in C, where y " xpuq }xpuq} Thus T P NHpHq. The proof of Proposition 1.12 is complete. . 6 Proof of Theorem 1.20 Lemma 6.1. Let txnunPN be a weakly convergent sequence in a Banach spaces X, which is not norm- convergent. Then for every sequence trnunPN of positive numbers such that rn Ñ 8, there is f P X such that trnfpxnq : n P Nu is dense in C. Proof. Let sn " suptdistpxm, Lnq : m ą nu, where Ln " spantu, x1, . . . , xnu and u is the weak limit of txnu. Clearly, tsnu is a decreasing sequence of non-negative numbers and therefore sn Ñ 2s ě 0. It is an elementary exercise to show that the equality s " 0 implies the norm convergence of txnu to u. Thus s ą 0 and passing to a subsequence, if necessary, we can assume that distpxn`1, Lnq ě s ą 0 for every n P N. This estimate combined with the Hahn -- Banach Theorem provides fn P X such that fnpxnq " 1, fnpuq " 0, fnpxjq " 0 for j ă n and }fn} ď 1 s . Since fnpuq " 0 and xk Ñ u weakly, we have fnpxmq Ñ 0 as m Ñ 8. Thus we can choose a strictly increasing sequence tnkukPN of positive integers such that fnkpxnmq ă 2´m whenever m ą k and rnk ą 2k for each k. Now we shall verify that the map R : ℓ1pNq Ñ ℓ1pNq defined by pRaqm " defined continuous linear operator. Indeed, since tfnku is bounded, for every a P ℓ1pNq, the series 8řk"1 is absolutely convergent in X and therefore defines ga " is a continuous linear functional on ℓ1pNq. Next, akfnk P X. Thus a ÞÑ pRaqm " gapxnmq akfnkpxnmq is a well- akfnk 8řk"1 8řk"1 }Ra ´ a}1 " 8ÿm"1pRaqm ´ am ď ÿk,mPN, k‰m akfnkpxnmq. 24 Since fnkpxnmq ă 2´m for m ą k and fnkpxnmq " 0 for m ă k, the above display yields }Ra ´ a}1 ď 8ÿk"1 ak 8ÿm"k`1 2´m ď 1 2 8ÿk"1 ak " }a}1 2 . rnk 2 ă 1 and therefore R is invertible. Thus R is bounded. Furthermore, }R ´ I} ď 1 Pick a dense sequence twkukPN in C such that wk ď k for every k P N. Then the sequence b defined by bk " wk belongs to ℓ1pNq since rnk ą 2k. Since R is invertible, there is a P ℓ1pNq such that Ra " b. That is, gapxnmq " bm for every m P N. According to the definition of bm, gaprnk xnkq " wk for each k P N. Hence trngapxnq : n P Nu is dense in C. Lemma 6.2. Let X and Y be Banach spaces and tTnunPN be a sequence of continuous linear operators from Y to X. Assume also that Ω ""py, fq P Y X : lim inf nÑ8 1`fpTnyq ››Tn›› " 0* is not nowhere dense in Y X. Then there is py, fq P Y X such that tfpTnyq : n P Nu is dense in C. Proof. Since Ω is not nowhere dense, Ω X U is dense in U for some non-empty open set U Ă Y X. Let V be an arbitrary non-empty open subset of U . We shall prove that there exist n P N and px, fq P V such that n ą m and fpTnxq " z. (6.1) Since Ω X U is dense in U , we can pick ph, uq P Ω X V . By definition of Ω, there is an infinite subset A of N such that nPA }Tn} " 8 and lim nÑ8 lim nÑ8 nPA hpTnuq }Tn} " 0. For every n P A, choose yn P SpY q for which 2}Tnyn} ě }Tn}. For each n P A, the Hahn -- Banach Theorem provides hn P SpXq such that hnpTnynq " }Tnyn}. In particular, 2hnpTnynq ě }Tn} Ñ 8 as n Ñ 8, n P A. Case 1: lim inf nÑ8 nPA hpTnynq hnpTnynq ą 0. In this case there are an infinite subset B of A and c ą 0 such that hpTnynq ą 2chnpTnynq for each n P B. For n P B denote δn " z´hpTnuq . From the above display and the inequality hpTnynq ą hpTnynq 2chnpTnynq ě c}Tn} it follows that δn Ñ 0 as n Ñ 8, n P B. Then we can choose n P B such that n ą m and pu ` δnyn, hq P V . Plugging the definition of δn into the expression hpTnpu ` δnynqq, we after cancellations get hpTnpu ` δnynqq " z. Thus n and px, fq " pu ` δnyn, hq satisfy (6.1). Case 2: lim inf nÑ8 nPA hnpTnuq hnpTnynq ą 0. In this case there are an infinite subset B of A and c ą 0 such that hnpTnuq ą 2chnpTnynq for each n P B. For n P B denote δn " z´hpTnuq . From the above display and the inequality hpTnynq ą hnpTnuq 2chnpTnynq ě c}Tn} it follows that δn Ñ 0 as n Ñ 8, n P B. Then we can choose n P B such that n ą m and pu, h ` δnhnq P V . Plugging the definition of δn into the expression ph ` δnhnqpTnuq, we after cancellations get ph ` δnhnqpTnuq " z. Thus n and px, fq " pu, h ` δnhnq satisfy (6.1). hnpTnynq Ñ 0 and hnpTnuq Case 3: hpTnynq Fix δ ą 0 such that pu ` δBpY qq ph ` δBpXqq Ď V . In our case hnpTnpu ` δynqq " hnpTnuq ` as n Ñ 8, n P A. hnpTnynq Ñ 0 δhnpTnynq ‰ 0 for all sufficiently large n P A. This allows us to define hnpTnpu`δynqq " z´hpTnuq´δhpTnynq δhnpTnynq`hnpTnuq rn " z´hpTnpu`δynqq . hnpTnynq Ñ 0 and hnpTnuq Since hpTnynq as n Ñ 8, n P A, we have rn Ñ 0 as n Ñ 8, n P A. Thus we can pick n P A such that n ą m and rn ă δ. Set x " u ` δyn and f " h ` rnhn. Since hnpTnynq Ñ 0 25 pu ` δBpY qq ph ` δBpXqq Ď V , we have px, fq P V . Finally, plugging the definition of rn into the expression fpTnxq " ph` rnhnqpTnu` δTnynq, we after cancellations get fpTnxq " z. This completes the proof of Claim (6.1). According to (6.1), Λ " tpy, f, fpTnyqq : py, fq P U, n P Z`u is dense in U C. By Theorem U, W " py, fq P U : tfpTnyq : n P Z`u is dense in C( is a dense Gδ subset of U . particular, W ‰ ∅ and there is py, fq P Y X such that tfpTnyq : n P Nu is dense in C. Corollary 6.3. Let T P LpXq, A be an infinite subset of N and txkukPN be a sequence in X such that T nxk ‰ 0 for every pn, kq P Z` N and In }T nx1} Ñ 8 and for each k P N, }T nxk`1} }T nxk} Ñ 8 as n Ñ 8, n P A. (6.2) Then T P WNHpXq. Proof. Let E " spantxk : k P Nu and Y be the closure of E. By (6.2), }T nx} " o`››T nY›› for each x P E as n Ñ 8, n P A. By Lemma 6.2, applied to Tn " T nY : Y Ñ X, there are px, fq P Y X such that tfpT nxq : n P Z`u " OpT, x, fq is dense in C. Hence T P WNHpXq. }T n} " 0) is dense in X. Then Corollary 6.4. Let T P LpXq be such that !x P X : lim inf T P WNHpXq. Proof. Just apply Lemma 6.2 with Y " X and Tn " T n. 1`}T nx} nÑ8 n zk }T nxk} " zk`1 Assume that (1.20.1) is satisfied. Then there is a sequence tznunPN in σppTq such that 1 ă z1 ă z2 ă . . . Pick xk P SpXq such that T xk " zkxk for each k P N. Then }T nx1} " z1n Ñ 8 and }T nxk`1} Ñ 8 as n Ñ 8. By Corollary 6.3, T P WNHpXq. Thus (1.20.1) implies the weak numeric hypercyclicity of T . Assume that (1.20.3) is satisfied. That is, there is a cyclic vector x for T satisfying lim inf nÑ8 1`}T nx} }T n} " 0. 1`}T ny} Let p P Crzs and y " ppTqx. Then }T ny} ď }ppTq}}T nx} and therefore lim inf }T n} " 0. Since the nÑ8 set tppTqx : p P Crzsu is dense in X, T P WNHpXq by Corollary 6.4. Thus (1.20.3) implies the weak numeric hypercyclicity of T . Assume that (1.20.4) is satisfied. That is, there are x P X and an infinite set A Ď N such that }T nx}(nPA is weakly convergent but is not norm }T nx} Ñ 8 as n Ñ 8, n P A and the sequence T nx convergent. Let rn " }T nx} and xn " T nx . Then trnunPA converges to 8, while txnunPA converges }T nx} weakly but not in the norm. By Lemma 6.1, there is f P X such that trnfpxnq : n P Au is dense in C. Since trnfpxnq : n P Au is a subset of OpT, x, fq, OpT, x, fq is dense in C. By Proposition 1.5, T P WNHpXq. Thus (1.20.4) implies the weak numeric hypercyclicity of T . Finally, assume that (1.20.2) is satisfied. That is, there is r ą 1 such that σppTq X rT is infinite. Then we can choose a countable infinite set M Ă T and ez P SpXq such that T ez " rzez for each z P M . Note that ez are linearly independent. Let E " spantez : z P Mu. Then the closure of E is a closed T -invariant subspace. By Proposition 1.6, it suffices to prove the weak numeric hypercyclicity of the restriction of T to the closure of E. Thus, without loss of generality, we can assume that E is dense 1`}T nx} }T n} " 0 for every x P E since }T nx} " Oprnq. In this case in X. If lim sup nÑ8 }T n}rn ă 8. Then there is c ě 1 T P WNHpXq by Corollary 6.4. It remains to consider the case lim sup nÑ8 such that }T n} ď crn for each n P Z`. For every z P M , consider the unique linear functional ez : E Ñ C such that ezpezq " 1 and ezpewq " 0 if w ‰ z. First, we shall prove that each ez is bounded and satisfies }ez} ď c. Assume the contrary. Then there is z0 P M and x " cj zj (zj are assumed pairwise distinct) }T n}rn " 8, then lim inf nÑ8 mřj"0 26 such that }x} " 1 and c0 " ez0pxq " a ą c. Since }T n} ď crn, we have }xn} ď c for each n P Z`, where xn " z´n 0 ezj . Then cj zn 0 r´nT nx " aez0 ` k´1ÿn"0 j z´n mřj"1 ›››› xn›››› "››››kaez0 ` 0 " ck ` mÿj"1cj k´1ÿn"0 zn j zn ka ď kc ` By the triangle inequality, mÿj"1 cj´k´1ÿn"0 zn j zn 0¯ezj›››› ď ck for each k P N. mÿj"1 j z´k cj1 ´ zk 0 0 ď ck ` 1 ´ zj z´1 mÿj"1 2cj zj ´ z0 for every k P N. azzk : k P N) is dense in C. Since řzPM Since the last sum does not depend on k, it follows that a ď c. This contradiction proves that each ez is bounded and has the norm at most c. Hence we can uniquely extend each ez to X by continuity ?az ă 8 and thus treat them as elements of X. By Lemma 3.10, there is a P ℓ1 and O " !rk řzPM ?az ă 8, }ez} " 1 and }ez} ď c for every z P M , the series x " řzPM ?azez converge absolutely in the Banach spaces X and X respectively. Using the equalities T ez " rzez and ezpewq " δz,w for z, w P M , we have fpT kxq " rkf´ ÿwPM ?azez and f " řzPM ?awwkew¯ " rk ÿz,wPM ?azawwkezpewq " rk ÿzPM `pMq such that řzPM azzk for each k P N. Hence OpT, x, fq " O is dense in C. By Proposition 1.6, T P WNHpXq. That is, (1.20.2) implies the weak numeric hypercyclicity of T . This completes the proof of Theorem 1.20. 7 Proof of Theorems 1.13 and 1.14 First, we shall establish some obstacles for an operator to be weakly numerically hypercyclic. Lemma 7.1. Let T P LpXq and X " Y ' Z, where Y and Z are closed T -invariant subspaces of X such have distinct absolute values. Then T R WNHpXq. that TY P LpY q is power bounded, Z is finite dimensional and the distinct eigenvalues of TZ P LpZq Proof. Since the eigenvalues of TZ have distinct absolute values, we can write σpTZq " tz1, . . . , zmu with z1 ă . . . ă zm. Considering the Jordan normal form of TZ , it is easy to see that for every x P X and f P X there exist y P Y , g P Y and polynomials p1, . . . , pm P Crzs such that fpT nxq " gpT nyq ` pjpnqzn j for every n P Z`. mřj"1 If p1 " . . . " pm " 0, tfpT nxqunPZ` " tgpT nyqunPZ` is bounded since TY is power bounded. Thus OpT, x, fq is bounded in C. Otherwise, we can choose q P N such that 1 ď q ď m, pq ‰ 0 and pj " 0 whenever j ą q. Then q´1řj"1 pjpnq` zj zqn¯ for every n P Z`. fpT nxq " gpT nyq ` zn q´pqpnq ` Since pq ‰ 0, tgpT nyqunPZ` is bounded and zj zq ă 1 for each j in the above sum, the last display ensures that tfpT nxqunPZ` is bounded if zq ă 1 and if zq " 1 and pq is constant and that fpT nxq Ñ 8 as n Ñ 8 otherwise. In any case OpT, x, fq is non-dense in C and therefore T R WNHpXq. Lemma 7.2. Let f be a holomorphic function defined in a neighborhood of T and B be a discrete closed subset of R p"a subset of R with no accumulation pointsq. Then BfpTq is nowhere dense in C. 27 Proof. If f is identically 0, the result is trivial. Otherwise f has only finitely many zeros s1, . . . , sk (if any) in T. Looking at the Taylor expansion of f about sj, we see that for each j, fpzq " ajpz ´ sjqkjp1 ` Opz ´ sjqq as z Ñ sj, where aj P Czt0u and kj P N. Let v P Czt0u be an accumulation point of BfpTq, which does not belong to BfpTq. Then it is easy to see that v " lim rnfpwnq, where trnunPN and twnunPN are sequences in B and T respectively such that rn Ñ 8. Convergence of trnfpwnqu yields fpwnq Ñ 0. Passing to a subsequence, if necessary, we can assume that wn Ñ sj as n Ñ 8 for some j. Then rnfpwnq " rnajpwn ´ sjqkjp1 ` Opwn ´ sjqq as n Ñ 8. This equality together with rnfpwnq Ñ v implies that ajv´1piwjqkj P R. That is v belongs to the union Λ of the lines ajpiwjqkj R for 1 ď j ď k. Hence BfpTq Y Λ is closed in C. Furthermore, for every line L through the origin, which is not in Λ, L X pBfpTq Y Λq is countable. It follows that BfpTq Y Λ has empty interior and therefore, being closed, is nowhere dense in C. Thus BfpTq is nowhere dense in C. Lemma 7.3. Let R ą r ą 1, a, b, c P C and k, m, j P Z. Then the set Ω " tRnpazk ` bzmq ` crnzj : n P Z`, z P Tu is nowhere dense in C. Proof. It is clear that the set of accumulation points of Ω that do not belong to Ω is contained in the set q, where tzquqPN is a sequence in T and tnquqPN is a strictly increasing Λ of lim sequence of positive integers. Since the closure of Ω coincides with ΩYΛ, it suffices to show that ΩYΛ has empty interior. Since the set Ω does not change if we replace z by z2, we can, without loss of generality, assume that k and m are even. qÑ8`Rnqpazk q q ` crnq zj q ` bzm If a ‰ b, then every sequence in the last limit runs off to infinity and therefore Λ " ∅. Same happens if a " b " 0 and c ‰ 0. In these cases ΩY Λ " Ω has empty interior as a union of countably many smooth curves. If c " 0, the desired result follows from Lemma 7.2. The case k " m is trivial. Thus without loss q P Λ. of generality, we can assume that k ‰ m, a " b " 1 and c ‰ 0. Let v " lim qÑ8 Since R ą r ą 1, this limit can only exist if azk q ` bzm q Ñ 0. Passing to a subsequence, if necessary, we may assume that zq Ñ w and awk ` bwm " 0. Since k and m are even, we can write k " s ` t and m " s ´ t, where s " k`m P Zzt0u. As with every pair of elements of T, we can find α, β P T such that a " βαt and b " βα´t. Then P Z and t " k´m q q`crnq zj Rnqpazk q `bzm 2 2 q ` bzm Rnqpazk q q ` crnq zj q, where uq " αzq. q ` u´t q qus q q ` drnq ul q " βα´sRnqput q ` u´t q ` crnq α´j uj q Ñ gv, where l " j ´ s, d " cαs´j β´1 and Then uq Ñ x with xt ` x´t " 0 and Rnqput g " αsβ´1. Since uq Ñ x, we can write uq " xeρq i, where tρqu is a convergent to 0 sequence of real numbers. The equation xt`x´t " 0 implies that xt " i or xt " ´i. Passing to a subsequence, if necessary, we can assume that either each ρq is positive or that each ρq is negative. We consider the case when each ρq is positive and xt " ´i (the other 3 cases are similar). Since 2ixtRnq sinptρqq ` drnq xlelρq i Ñ gv, we have 2Rnq sinptρqq ` drnq xlelρqi Ñ gv. Since elρqi Ñ 1, it follows that dxl " ´h, where h ą 0 (otherwise 2Rnq sinptρqq ` drnq xlelρq i Ñ 8). Thus p2Rnq sinptρqq ´ hrnq cosplρqqq ´ irnq sinplρqq Ñ gv. Hence rnq sinplρqq Ñ ´Impgvq and therefore rnq sinptρqq Ñ ´ t l Impgvq. Then either gv is real or trnq sinptρqqu converges to a non-zero number. In the latter case the boundedness of t2Rnq sinptρqq ´ hrnq cosplρqqu (which follows from the convergence) implies that R " r2. Indeed, if R ‰ r2, one of the sequences t2Rnq sinptρqqu and thrnq cosplρqqu runs to infinity faster than the other. In the case R " r2, by looking at the Taylor series expansions of sin and cos, one sees that 2Rnq sinptρqq ´ hrnq cosplρqq " Opr´nqq and therefore 2Rnq sinptρqq ´ hrnq cosplρqq Ñ 0. Hence Repgvq " 0. Thus in any case gv is either real or purely imaginary. Hence v belongs to the union of two lines g´1R and ig´1R. Since there are finitely many possibilities to choose x, Λ is contained in the union of finitely many lines through the origin. Thus Ω Y Λ is a part of the union of countably many smooth closed curves and finitely many lines. Hence Ω Y Λ has empty interior, as required. Lemma 7.4. Let T P LpCnq be a diagonal operator with the diagonal entries d1, . . . , dn such that d1 " . . . " dn ‰ 0 and for every j, k P t1, . . . , nu, dj dj are not independent. Then T R WNHpCnq. and dk dk 28 If R ď 1, then every numerical orbit of T is bounded and therefore T is not Proof. Let R " d1. weakly numerically hypercyclic. It remains to consider the case R ą 1. Since every pair dj R is not independent, there are w P T, u1, . . . , un P T of finite order and k1, . . . , kn P Z such that dj " Rujwkj for 1 ď j ď n. Since uj have finite order, we can pick m P N such that um n " 1. It is easy to see that every numerical orbit OpT, f, xq has the shape 1 " . . . " um R , dk OpT, f, xq " ta1dk 1 ` . . . ` andk n : k P Z`u, where aj P C. Taking into account that dj " Rujwkj and um 1 " . . . " um n " 1, we see that OpT, f, xq " m´1ďq"0 Mq, where Mq " tRkm`qpa1uq 1wqk1wkmk1 ` . . . ` anuq nwqknwkmknq : k P Z`u. nřj"1 aj uq j wqkj zkj . Clearly Mq Ď fqpTqBq, Now for 0 ď q ď m´1, consider the Laurent polynomials fqpzq " where Bq " tRkm`q : k P Z`u. By Lemma 7.2, each Mq is nowhere dense in C. By the above display OpT, f, xq is nowhere dense in C. Hence T R WNHpCnq. Lemma 7.5. Let T P LpC3q be a diagonal operator with non-zero diagonal entries d1, d2, d3 such that for every j, k P t1, 2, 3u for which dj dk ‰ 0, dj dj Proof. If d1,d2,d3 are pairwise distinct, the result follows from Lemma 7.1. If d1 " d2 " d3, the result follows from Lemma 7.4. Thus we can assume that d1 " d2 ‰ d3. It is easy to see that every numerical orbit OpT, f, xq has the shape OpT, f, xq " ta1dk are not independent. Then T R WNHpC3q. 3 : k P Z`u, where aj P C. 1 ` a2dk 2 ` a3dk and dk dk 1 ` a2dk If d1 " d2 ă d3 and a3 ‰ 0, then the sequence ta1dk 3u is bounded if d3 ď 1 and converges to infinity otherwise, ensuring that OpT, f, xq is non-dense in C. If a3 " 0, then we fall under the jurisdiction of Lemma 7.4 with n " 2 and again OpT, f, xq is non-dense in C. In the case d1 " d2 ą d3 and d3 ď 1, one can proceed similar to the proof of Lemma 7.4 (we leave this easy bit to the reader). It remains to consider the case d1 " d2 ą d3 ą 1 and a3 ‰ 0. Let R " d1 " d2 and r " d3. Since every pair dj is not independent, there is w P T, u1, u2, u3 P T of finite order dj and k1, k2, k3 P Z such that d1 " Ru1wk1, d2 " Ru2wk2 and d3 " ru3wk3. Since uj have finite order, we 1 " um can pick m P N such that um 3 " 1. Then 2 ` a3dk and dk dk 2 " um OpT, f, xq " Mq, where m´1ďq"0 Mq " tRkm`qpa1uq 1wqk1 wkmk1 ` a2uq 2wqk2wkmk2q ` rkm`qa3uq 3wqk3 wkmk3 : k P Z`u. Now each Mq is contained in a set of the shape tRmkpαzk1 ` βzk2q ` rmnγzk3 : k P Z`, z P Tu with α, β, γ P C being constants. Then Lemma 7.3 guarantees that each Mq is nowhere dense in C. By the above display OpT, f, xq is nowhere dense in C. Hence T R WNHpC3q. 7.1 Proof of Theorem 1.13 , λ2 Let T P LpC2q and σpTq " tλ1, λ2u with λ1, λ2 P C. By Theorem 1.9, if λ1 " λ2 ą 1 and λ1 λ1 λ2 are independent, then T P WNHpC2q. If λ1 ‰ λ2 or λ1 " λ2, then T R WNHpC2q by Lemma 7.1. If λ1 " λ2 ď 1 and λ1 ‰ λ2, then T is power bounded and therefore T R WNHpC2q. Finally, if λ1 " λ2 ą 1, λ1 ‰ λ2 and λ1 are not independent, then T R WNHpC2q according to Lemma 7.4. λ1 Thus T P WNHpC2q if and only if λ1 " λ2 ą 1 and λ1 λ1 2 : k P Z`u is dense in C, then T P SNHpC2q by Theorem 1.10. Assume now that T P SNHpC2q. Then T P WNHpC2q and therefore λ1 " λ2 ą 1 and λ1 ‰ λ2. Then T is similar to the are independent in T. 1 ` λk If tλk , λ2 λ2 , λ2 λ2 29 1 ` bλn 2 : n P Z`u is dense in C. Since aλn diagonal operator with the numbers λ1 and λ2 on the diagonal. By Lemma 4.3, there exist a, b ě 0 such 2 ě a ´ bRn, it follows that a " b. Hence 1 ` bλn that taλn 2 : n P Z`u is dense in C. Thus T P SNHpC2q if and only if σpTq " tλ1, λ2u with tλk 1 `λn tλn 2 : k P Z`u being dense in C. Now assume that T P WNHpC2q and T is not unitarily equivalent to a diagonal operator. As we have already shown, λ1 " λ2 ą 1 and λ1 are independent in T. Since T is not unitarily λ1 equivalent to a diagonal operator, the eigenspaces kerpT ´ λ1Iq and kerpT ´ λ2Iq are non-orthogonal. By Proposition 1.11, T P NHpC2q. Finally, assume that T P NHpC2q. Then T P WNHpC2q. If T is unitarily equivalent to a diagonal operator, Lemma 4.3 implies that T P SNHpC2q. These observations amount to the fact that T P NHpC2q if and only if either T P SNHpC2q or T P WNHpC2q and T is not unitarily equivalent to a diagonal operator. The proof of Theorem 1.13 is complete. 1`λk , λ2 λ2 , λ2 λ2 having infinite order in the group T and λ1 λ1 7.2 Proof of Theorem 1.14 Let T P LpC3q. If there are λ1, λ2 P σpTq such that λ1 " λ2 ą 1 and λ1 are independent in λ1 T, then (1.9.1) is satisfied and Theorem 1.9 implies that T P WNHpC3q. If σpTq " tλ1, λ2, λ3u with λ1 " λ2 ą λ3 ą 1, λ1 , λ3 being independent in T, λ3 then (1.9.3) is satisfied and Theorem 1.9 implies that T P WNHpC3q. If distinct members of σpTq have distinct absolute values, then T R WNHpC3q according to Lemma 7.1. If λj are pairwise dependent for λj non-zero λj, Lemma 7.5 guarantees that T R WNHpC3q. Consider the case λ1 ą λ2 " λ3 " R and either R ď 1 or λ2 R are not independent. If R ď 1, then each numerical orbit of T is either bounded or escapes to infinity at the rate λ1n (provided λ1 ą 1). If R ą 1 and λ2 R are not independent, then T R WNHpC3q. Indeed, each numerical orbit of T has the shape taλn 3 : n P Z`u if λ2 ‰ λ3 and taλn 2 : n P Z`u if λ2 " λ3. In any case if a ‰ 0, this is a sequence escaping to infinity. If a " 0 we fall under the jurisdiction of Lemma 7.4 and the numerical orbit is non-dense. If λ3 is 0, the problem is easily reduced to the 2-dimensional situation, already covered by Theorem 1.13. It remains to notice that we have considered all the possibilities (up to the ordering of the eigenvalues). R and λ3 1 ` bλn 2 ` cnλn 2 ` cλn 1 ` bλn λ2 R and λ3 8 Proof of Theorem 1.23 `pNq be such that 8řn"1 Lemma 8.1. Let tenunPN be a Schauder basis in a reflexive Banach space X, tenunPN be the corresponding sequence of coordinate functionals and a P ℓ1 an " 1. Then there is px, fq P ΠpXq such that enpxqfpenq " an for each n P N. Proof. Let Xn " spante1, . . . , enu. By Lemma 4.1, for every n P N, we can find pxn, gnq P ΠpXnq such that ekpxnqgnpekq " ak for 1 ď k ă n and enpxnqgnpenq " 1 ´ a1 ´ . . . ´ an´1. By the Hahn -- Banach Theorem, for each n P N, there is fn P SpXq such that fnXn " gn. Then for every n P N, Since X is reflexive, every bounded sequence in X X has a weakly convergent subsequence. Hence there is a strictly increasing sequence tnmumPN of positive integers such that txnmumPN converges weakly to x P X and tfnmumPN converges weakly to f P X. Since BpXq is weakly closed in X and BpXq is weakly closed in X, we have }x} ď 1 and }f} ď 1. According to the above display, for every k P N, ekpxnmqfnmpekq " ak for all sufficiently large m. Hence, pxn, fnq P ΠpXq and ekpxnqfnpekq " ak for 1 ď k ă n. It follows that ekpxqfpekq " lim mÑ8 fpxq " f´ 8řk"1 ekpxqek¯ " ekpxnmqfnmpekq " ak for each k P N. 8řk"1 ekpxqfpekq " 8řk"1 ak " 1. Since fpxq " 1, }x} ď 1 and }f} ď 1, px, fq P ΠpXq. Thus px, fq satisfies all desired conditions. 30 Assume that T P LpXq satisfies (1.23.3). That is, there exists λ P C be such that λ ě 1 and T ´ λI is a semi-Fredholm operator of positive index. Since the index is locally constant, T ´ zI is a semi-Fredholm operator of positive index for every z P λ ` εD for a sufficiently small ε. Since every semi-Fredholm operator of positive index is non-injective, z ` εD Ď σppTq. Since λ ě 1, there is r ą 1 such that J " pλ`εDqXprTq ‰ ∅. Then J is a non-trivial open arc of the circle rT. By Proposition 1.16, we can find z, w P J Ă σppTq such that tzn ` wn : n P Z`u is dense in C. Thus (1.10.1) is satisfied and therefore T P SNHpXq by Theorem 1.10. Hence (1.23.3) implies the strong numeric hypercyclicity of T . Assume now that T P LpXq satisfies (1.23.4). That is, X be reflexive and there exists λ P C such that λ ě 1 and T ´ λI is a semi-Frdholm operator of negative index. It is well-known and easy to see that if X is reflexive and R P LpXq, then R is semi-Fredholm if and only if R is semi-Fredholm and ipRq " ´ipRq. Thus T ´ λI is a semi-Fredholm operator of positive index. By the previous part of the proof, T P SNHpXq. By Proposition 1.6, T P SNHpXq. Thus (1.23.4) implies the strong numeric hypercyclicity of T . Assume that T P LpXq satisfies (1.23.1). That is, X is reflexive and there is a Schauder basic sequence tenunPN in X such that T en " λnen with λn P C for each n P N and for some c P ℓ1 `pNq, ! 8řj"1 cj λk j : k P Z`) is dense in C. (8.1) Without loss of generality, we may assume that 8řn"1 cn " 1. Let Y be the closed linear span of the sequence tenunPN. By Lemma 8.1, there is px, gq P ΠpY q such that enpxqgpenq " cn for every n P Z`, where en P Y are the coordinate functionals for the Schauder basis tenunPN. By the Hahn -- Banach Theorem, there is f P X such that fY " g and px, fq P ΠpXq. Then 8řn"1 nenpxqgpenq " λk fpT kxq " gpT kxq " g´ 8řn"1 nenpxqen¯ " n for each k P Z`. 8řn"1 cnλk λk By (8.1), OpT, x, fq is dense in C and T P NHpXq. Since every operator similar to T satisfies the same conditions, T P SNHpXq. Thus (1.23.1) implies the strong numeric hypercyclicity of T . Finally, assume that T P LpXq satisfies (1.23.2). That is, X is reflexive and there is a Schauder basic sequence tenunPN in X such that T en " λnen where λn P C are such that λ1 ą 1, the sequence tλnunPN is pmaybe non-strictlyq increasing and the numbers λn are pairwise distinct elements of T. λn Passing to a subsequence, if necessary, we can assume that either tλnunPN is strictly increasing or that tλnunPN is constant. If tλnunPN is strictly increasing, Corollary 3.12 provides c P ℓ1 `pNq satisfying (8.1). If λn " r ą 1 for every n P N, by Lemma 3.10, there is c P ℓ1 `pNq for which (8.1) holds. In any case (1.23.1) is satisfied and therefore T P SNHpXq. Thus (1.23.2) implies the strong numeric hypercyclicity of T . This completes the proof of Theorem 1.23. 9 Proof of Theorem 1.29 Lemma 9.1. Let px, fq P ΠpXq. Then for every y P SpXqXker f , there exists g P X such that gpxq " 0, gpyq " 1 and }g} ď 2. Proof. Let E " spantx, yu. Since fpxq " 1 and fpyq " 0, x and y are linearly independent and therefore form a basis in E. Consider ϕ P E defined by ϕpxq " 0 and ϕpyq " 1. Then u " fpuqx ` ϕpuqy for every u P E. Hence ϕpuq " }u ´ fpuqx} ď }u} ` }f}}u} " 2}u}. That is, }ϕ} ď 2. By the Hahn -- Banach Theorem, there is g P X such that }g} ď 2 and gE " ϕ. Obviously, g satisfies all desired conditions. dimensional subspace E of X such that }TE}LpE,Xq ă ε. Proof. Since TpXq is non-closed, inft}T x} : x P SpXqu " 0. Hence we can pick x P SpXq for which 4 . By the Hahn -- Banach theorem there is f P X such that px, fq P ΠpXq. Next, observe that }T x} ă ε Lemma 9.2. Let T P LpXq be such that TpXq is non-closed in X and ε ą 0. Then there is a 2- 31 inft}T x} : x P SpXqX ker fu " 0. Indeed, otherwise Tpker fq is closed in X and is a subspace of TpXq of codimension at most 1. Since a subspace of a Banach space with a finite codimensional closed subspace is closed itself, that would have implied that TpXq is closed. }T y} ă ε E " spantx, yu is a 2-dimensional subspace of X. Moreover, for each u P BpEq, 4 ` 2 ε Thus inft}T x} : x P SpXq X ker fu " 0 and therefore we can choose y P SpXq X ker f such that 4 . By Lemma 9.1, there is g P X such that }g} ď 2, gpxq " 0 and gpyq " 1. Clearly }T u} " }fpuqT x ` gpuqT y} ď }f}}T x} ` }g}}T y} ă ε 4 " 3ε 4 . 4 ă ε. Lemma 9.3. Let T P LpXq, E be a 2-dimensional subspace of X and S : E Ñ X be a linear map. Then Hence }TE}LpE,Xq ď 3ε there is R P LpXq such that RE " S and }T ´ R} ď 3}S ´ TE}LpE,Xq. Proof. Take any x P SpEq. By the Hahn -- Banach Theorem, there exists f P X such that px, fq P ΠpXq. Pick y P SpEq such that fpyq " 0. By Lemma 9.1, there is g P X such that }g} ď 2, gpxq " 0 and gpyq " 1. Now we define R P LpXq be the formula Ru " T u ` fpuqpSx ´ T xq ` gpuqpSy ´ T yq. It is straightforward to see that Rx " Sx and Ry " Sy. Since tx, yu is a basis in E, RE " S. Next, for every u P X, using the relations }x} " }y} " }f} " 1 and }g} ď 2, we have Hence }T ´ R} ď 3}S ´ TE}LpE,Xq. }pT ´ Rqu} " }fpuqpT ´ Sqx ` gpuqpT ´ Sqy} ď 3}u} }S ´ TE}LpE,Xq. Lemma 9.4. Let E be a 2-dimensional Banach space and T P LpEq be such that σpTq " ts, tu and s " t ě 1. Then there is a sequence tTnunPN in SNHpEq such that }Tn ´ T} Ñ 0. Proof. Since s " t ě 1, by Proposition 1.16, we can choose two sequence tsnunPN and ttnunPN such that sn Ñ s, tn Ñ t and tsk n : k P Z`u is dense in C for every n P N. Since σpTq " ts, tu, we can choose a linear basis tx, yu in E such that T x " sx and T y " ty ` ax for some a P C. For each n P N consider Tn P LpXq given by Tnx " snx and Tny " tny ` ax. Since sn Ñ s and tn Ñ t, we have }Tn ´ T} Ñ 0. On the other hand, σpTnq " tsn, tnu for each n P N. By Theorem 1.13, each Tn belongs to SNHpEq. n ` tk Since WNHpXq Ě NHpXq Ě SNHpXq, we have (1.29.3)ùñ(1.29.2)ùñ(1.29.1). Next, it is easy to see that the set ΩpXq of T P LpXq satisfying (1.29.4) is operator norm open. Thus in order to verify that (1.29.4) implies (1.29.3), it suffices to show that ΩpXq X WNHpXq " ∅. Let T P ΩpXq. Then X " Y ' Z, where Y and Z are closed T -invariant subspaces, TY P LpY q has the spectral radius ă 1, Z has dimension n P Z` and σ`TZ " tz1, . . . , znu satisfies 1 ď z1 ă . . . ă zn. By Lemma 7.1, T R WNHpXq, which proves the implication (1.29.4)ùñ(1.29.3). It remains to verify that (1.29.1) implies (1.29.4). In order to do this, it suffices to show that every T P LpXqzΩpXq is in the operator norm closure of SNHpXq. Let T P LpXqzΩpXq. The relation T R ΩpXq can happen for various reasons. Case 1: There is a 2-dimensional T -invariant subspace E of X such for which σ`TE " ts, tu and }Sn´ TE} Ñ 0. By Lemma 9.3, there is a sequence tRnunPN in LpXq such that RnE " Sn for each n P N s " t ě 1. In this case, according to Lemma 9.4, there is a sequence tSnunPN in SNHpEq such that and }Rn ´ T} Ñ 0. Since the restriction of each Rn to the invariant subspace E is strongly numerically hypercyclic, Proposition 1.6 implies that Rn P SNHpXq for each n P N. Thus T is in the norm closure of SNHpXq. Note that the situation of T having a normal eigenvalue λ of multiplicity ě 2 such that λ ě 1 falls into Case 1. The same happens if there are two distinct s, t P σppTq such that s " t ě 1. This means that if Case 1 does not hold, we can pick λ P σpTq such that λ ě 1, λ is not a normal eigenvalue of T and dim kerpT ´ λIq ď 1. We keep the notation λ for such a spectral point for the rest of the proof. 32 Case 2: subspace E of X such that }pT ´ λIqE} ă ε SE " λIE and }T ´ S} ă 3ε such that }S ´ R} ă ε to the norm closure of SNHpXq. pT ´ λIqpXq is non-closed in X. Let ε ą 0. By Lemma 9.2, we can pick a 2-dimensional 4 . According to Lemma 9.3, there is S P LpXq such that 4 . Now S falls under the jurisdiction of Case 1. Thus there is R P SNHpXq 4 " ε and ε ą 0 is arbitrary, T belongs 4 . Since }T ´ R} ď }T ´ S}`}S ´ R} ă 3ε Thus it remains to consider the case when pT ´ λIqpXq is closed in X. Case 3: kerpT ´ λIq is one-dimensional and pT ´ λIqpXq " X. In this case T ´ λI is a Fredholm operator of index 1. By Theorem 1.23, T P SNHpXq. Case 4: dimpX{pT ´ λIqpXqq ą dimpkerpT ´ λIqq. In this case T ´ λI is a semi-Fredholm operator of negative index. Since X is reflexive, Theorem 1.23 implies that T P SNHpXq. If kerpT ´ λIq " t0u and pT ´ λIqpXq " X, we have λ R σpTq, which is a contradiction. Finally, if dimpkerpT ´ λIqq " dimpX{pT ´ λIqpXqq " 1, λ is a normal eigenvalue of T , which is again a contradiction. This completes the proof of the implication (1.29.1)ùñ(1.29.4) and that of Theorem 1.29. 4 ` ε 10 Proof of Theorem 1.26 Lemma 10.1. Let T be a bounded normal operator on a Hilbert space H. Assume also that there exists a sequence tznunPN in σpTq such that 1 ă z1 ă z2 ă . . . Then T P WNHpXq. Proof. Let rk " 1 3 mintzk ´zk´1,zk`1 ´zku for k P N, where we assume z0 " 1. For each k P N pick a continuous function ϕk : C Ñ r0, 1s such that ϕkpzkq " 1 and ϕk vanishes outside the disk zk ` rkD. The functional calculus for normal operators allows us to consider the operators ϕkpTq. Since zk P σpTq and ϕkpzkq ‰ 0, ϕkpTq ‰ 0 for every k P N. Thus we can pick xk P SpHqX ϕkpTqpHq. The latter inclusion and the fact that ϕk vanishes outside the disk zk ` rkD implies that pzk ´ rkqn ď }T nxk} ď pzk ` rkqn for every k P N and every n P Z`. These estimates together with the inequalities z1 ´ r1 ą 1 and zk`1 ´ rk`1 ą zk ` rk for k P N imply that }T nx1} Ñ 8 and }T nxk`1} }T nxk} Ñ 8 as n Ñ 8. By Corollary 6.3, T P WNHpXq. Lemma 10.2. Let T be a normal operator on Hilbert space H such that the set t´z : z P σpTqu is well-ordered and for every r ą 1, the set σpTq X rT has at most one element. Then for each x P H and f P H, either tfpT nxq : n P Z`u is bounded or fpT nxq Ñ 8. In particular, T R WNHpHq. Proof. Since every well-ordered subset of R is countable, M " tz P σpTq : z ą 1u is (finite or) countable. By the Spectral Theorem, T is unitarily equivalent to the orthogonal direct sum of a normal contraction and a diagonal (with respect to an orthonormal basis) operator with only elements of M on the diagonal. Taking this into account, we see that for each f P H and x P H, fpT nxq " bn ` řtPM attn, where a P ℓ1pMq and b P ℓ8pZ`q. (10.1) If a " 0, then tfpT nxq : n P Z`u " tbn : n P Z`u is bounded. Assume now that a ‰ 0. Since t´z : z P σpTqu is well-ordered, there is c " maxtt : t P M, at ‰ 0u. Since σpTq X cT is a singleton, there is the unique t0 P M satisfying t0 " c. Then 0 ` řtPM, tăc 0 " at0cnp1 ` op1qq as n Ñ 8. fpT nxq " at0cn1 ` bn Since c ą 1, fpT nxq Ñ 8. Lemma 10.3. Let T be a normal operator on Hilbert space H such that there is a sequence tλnunPN in σpTq such that 1 ă λ1 ă λ2 ă . . . and the numbers λj λj are pairwise distinct. Then T P NHpHq. at0 tn attn at0 tn 33 Proof. By the Spectral Theorem we can assume that H " L2pµq, where pΩ, F , µq is a measure space and T f " gf for f P H, where g P L8pµq. Then the spectrum of T is exactly the essential range of g. By Lemma 3.11, we can pick a non-negative real valued function a P L1pµq for which the set "żΩ gka dµ : k P Z`* is dense in C. Without loss of generality, we may assume that a is normalized to satisfy }a}1 " 1. Then x " ?a P SpHq and xT nx, xy "şΩ gka dµ for every n P Z`. Thus NOpT, xq is dense in C and therefore T P NHpHq. Now we are ready to prove Theorem 1.26. Let H be a Hilbert space, T P LpHq and k P N be such that T k is normal. First, assume that there is a sequence tλnunPN in σpTq such that 1 ă λ1 ă λ2 ă . . .. 2 ă . . .. Since T k is normal, Lemma 10.1 Then tλk yields the weak numeric hypercyclicity of T k. Then T is weakly numerically hypercyclic, which completes the proof of (1.26.1). nunPN is a sequence in σpT kq satisfying 1 ă λk 1 ă λk numbers λj λj Next, assume that there is a sequence tλnunPN in σpTq such that 1 ă λ1 ă λ2 ă . . . and the 2 ă . . . : n P N( is infinite. Then there is a strictly increasing sequence tnju of positive integers and the set λk are pairwise distinct. Then tλk λk n nunPN is a sequence in σpT kq satisfying 1 ă λk 1 ă λk are pairwise distinct. Since T k is normal, Lemma 10.3 implies that T k is such that the numbers n λk nj λk nj numerically hypercyclic. Hence T is numerically hypercyclic, which completes the proof of (1.26.2). Now assume that T k is self-adjoint. Then T 2k is self-adjoint and non-negative definite. If t´z : z P σpTq, z ą 1u is not well-ordered, we can pick a sequence tλnunPN in σpTq such that 1 ă λ1 ă λ2 ă . . . z ą 1u is well-ordered. Since By (1.26.1), T P WNHpHq. Assume now that t´z σpT 2kq Ă r0,8q, t´z : z P σpT 2kq, z ą 1u is well-ordered and σpT 2kqX rT has at most one element for each r ą 1. By Lemma 10.2, for every x P H and f P H, OpT 2k, x, fq is either bounded or nowhere dense in C. Now let x P H and f P H. Then OpT, x, fq " OpT 2k, T jx, fq. Hence OpT, x, fq is the union of finitely many sets each of which is either bounded or nowhere dense. Thus OpT, x, fq is not dense in C (the set of accumulation points of OpT, x, fq is bounded). Then T R WNHpHq, which completes the proof of (1.26.3). : z P σpTq, 2k´1Ťj"0 Finally, assume that there exist r ą 1 and a T -invariant subspace K such that U " 1 unitary operator with infinite spectrum. If σppUq is infinite, then there are a sequence tznunPN of pairwise distinct elements of T and an orthonormal sequence tenunPN in K Ď H such that T ken " U en " rznen for every n P N. Thus T k satisfies (1.10.4) and therefore T k P SNHpHq Ă NHpHq by Theorem 1.10. Hence T P NHpHq. It remains to consider the case when σppUq is finite. Since σpUq is infinite, the Spectral Theorem tells us that the restriction of U to an appropriate invariant subspace K0 is unitarily equivalent to the multiplication by the argument operator M P LpL2pµqq, M fpzq " zfpzq, where µ is a Borel probability purely non-atomic measure on T. By Proposition 1.2, in order to prove that T is numerically hypercyclic, it suffices to verify that rM is numerically hypercyclic. By Lemma 3.13, there is a P L1 Thus rM is numerically hypercyclic and therefore T is. This completes the proof of (1.26.4) and that of Theorem 1.26. `pµq such that }a}1 " 1 and the set O " trnxaµpnq : n P Nu is dense in C. Now it is easy to see that x " ?a P SpL2pµqq and xM nx, xy " xaµpnq for each n P Z. Hence NOprM, xq " O is dense in C. r T kK P LpKq is a 11 Proof of Propositions 1.18 and 1.19 Let T P LpC2q. Pick w P T such that at least one of the eigenvalues of wT is real. By Theorem 1.13, wT R WNHpC2q. This proves Proposition 1.18. The rest of this section is devoted to the proof of Proposition 1.19, which is more sophisticated. Denote M ""pr, z, wq P p1,8q T T :rz 0 0 rw P NHpC2q*. 34 By Theorem 1.13, we can rewrite the definition of M in the following equivalent way. Then the sets Mpz, wq featuring in Proposition 1.19 can be written as M " pr, z, wq P p1,8q T T : trnpzn ` wnq : n P Z`u is dense in C(. Mpz, wq ""r ą 1 :rz 0 0 rw P NHpC2q* " tr ą 1 : pr, z, wq P Mu " r ą 1 : trnpzn ` wnq : n P Z`u is dense in C( for z, w P T. If tVkukPN is a base of the topology of C, then we can write M " čk,mPN 8ďn"m pr, z, wq P p1,8q T T : rnpzn ` wnq P Vk(. (11.1) This immediately implies that M is a Gδ subset of p1,8q T T and therefore each Mpz, wq is a Gδ-subset of p1,8q. Lemma 11.1. The set M is a dense Gδ-subset of p1,8q T T. Furthermore, W " pz, wq P T2 : Mpz, wq is dense in p1,8q( is a dense Gδ subset of T2. Proof. By Proposition 1.16, for every r ą 1, the set rT X M is dense in rT. Hence M is dense in p1,8q T T. Let trn : n P Nu be a dense subset of p1,8q. By Proposition 1.16, the set Wn " tpz, wq P T2 : rn P Mpz, wqu is a dense Gδ subset of T2 for every n P N. Hence ŞnPN Wn is a dense Gδ subset of T2, Let tVkukPN is a base of the topology of C. It is easy to verify that being also a subset of W . Hence W is dense in T2. It remains to show that W is a Gδ set. Aa,b,m,k " 8ďn"m pz, wq P T2 : pzn ` wnqra, bs X Vk ‰ ∅( is an open subset of T2 whenever m, k P N and 1 ă a ă b ă 8. It remains to notice that W " čm,kPN ča,bPQ, 1ăaăb Aa,b,m,k to conclude that W is a Gδ subset of T2. a rather typical occurrence. By Lemma 1.13, Mpz, wq " ∅ if z and w are not independent. Actually, the equality Mpz, wq " ∅ is Lemma 11.2. The set pz, wq P T2 : Mpz, wq ‰ ∅( has zero Lebesgue measure in T2. n2( ă 8. Proof. Let µ be the normalized Lebesgue measure on T2. By (3.4), Hence for almost all pz, wq P T2, zn` wn ą n´2 for all sufficiently large n and therefore rnzn` wn Ñ 8 for every r ą 1. Hence Mpz, wq " ∅ for almost all z, w P T2. µ pz, wqPT2 : zn`wn ď 1 8řn"1 On the other hand, it turns out that if Mpz, wq is non-empty, then it is infinite. Lemma 11.3. Let z, w P T, r P Mpz, wq and k be an odd positive integer. Then r1{k P Mpz, wq. 35 Proof. In this proof, we shall use the symbol or for any sequence tαnu of complex numbers satisfying lim supαn1{n ď 1 r . Since r P Mpz, wq, the set tyn : n P Z`u is dense in C, where yn " rnpzn ` wnq. Let A " tn P N : n´1 ď yn ď nu. Then tyn : n P Au is dense in C. The inequality zn ` wn ď nr´n for n P A yields (11.2) wn " ´zn ` or and y2 n " ´yn2z2n ` or as n Ñ 8, n P A. Since k is odd and rnpzkn ` wknq " rnpzn ` wnq and (11.2) imply rnpzkn ` wknq " ynzpk´1qn ` or " yk k´1řj"0p´1qj zjnwpk´1´jqn, the equality yn " rnpzn ` wnq ynk´1 ` or as n Ñ 8, n P A. Hence, n rnpzkn ` wknq " Φpynq ` op1q as n Ñ 8, n P A, where Φ : Czt0u Ñ C, Φpxq " xk xk´1 . Since Φ is a homeomorphism of Czt0u onto itself and tyn : n P Au is dense in C, tΦpynq : n P Au is dense in C. Then by the above display, trnpzkn ` wknq : n P Au is dense in C. Hence the bigger set trn{kpzn ` wnq : n P Z`u is dense in C and therefore r1{k P Mpz, wq. Finally, we shall verify that Mpz, wq can not coincide with p1,8q by showing that each Mpz, wq has zero Lebesgue measure. We need some preparation. Consider the function δ : R Ñ r0, 1s, δpxq " distpx, 2Z ´ 1q. That is, δpxq is the distance from x to the nearest odd integer. Lemma 11.4. Let θ P RzQ, 0 ă a ă b ă 1 and A " tn P N : an ď δpnθq ď bnu. Then řnPA Proof. For every q P N, let ppqq be the unique odd integer p such thatθ´ p q ă 1 q for each q P N. Consider the set q and g.c.d.pq, ppqqq " 1). B "!q P N : αpqq ď bq řqPB q . Denote αpqq "θ´ ppqqq . Then 0 ă αpqq ă 1 First, we shall show that n ă 8. q ă 8. (11.3) 1 1 qj rj and γj ď brj . In particular, ppqjqqj ´ pprjqrj " γj ´ βj. It follows that ppqjqrj´pprjqqj Assume that (11.3) fails. Then there are two strictly increasing sequences tqjujPN and trjujPN in B such that qj ă rj for each j P N and qj rj Ñ 1. Indeed, otherwise members of B written in the increasing order grow exponentially and (11.3) follows. By definition of B, we have θ " ppqjqqj ` βj " pprjqrj ` γj, where qj ` brj βj ď bqj . Hence ppqjqrj ´ pprjqqj ď rjbqj ` qjbrj ă 2rj bqj since qj ă rj. Since qj rj Ñ 1 and b ă 1, we have rj bqj Ñ 0 and therefore ppqjqrj ´ pprjqqj Ñ 0. Since the latter is a sequence of integers, ppqjqrj ´ pprjqqj " 0 for all sufficiently large j. In particular, there is j P N for which ppqjqrj " pprjqqj. Since pprjq and rj are relatively prime, rj must divide qj, which is not possible since rj ą qj. This contradiction proves (11.3). Now n P A if and only if there is an odd integer k such that an ď δpnθq " nθ ´ k ď bn. Let r " g.c.d. pn, kq. Then we can write n " rq and k " rp, where q P N, p is an odd integer and g.c.d. pq, pq " 1. Using this notation we can rewrite the last inequality as arq q , which happens if and only if q P B, p " ppqq and arq rq . Thus, we have obtained the following q ď brq ď bqj rq ď bq qj rj rj description of the set A: rq ď αpqq ď brq A "!qr : q P B, r P 2N ´ 1, arq rq ďθ ´ p rq). rq ď αpqq ď brq m for every m P N. Then Fix d ą 0 small enough in such a way that dm ď am A Ď A1 " tqr : q P B, r P N, drq ď αpqq ď brqu " tqr : q P B, r P N, βpqq ď r ď cβpqqu, 36 where βpqq " log αpqq corresponding integral, one easily sees that q log d and c " log d log b . Obviously, βpqq ą 0 and c ą 1. Estimating a monotonic sum by the řmPN m ď 1 ` log c for every β ą 0 and c ą 1. βďmďcβ 1 Using the above two displays and (11.3), we obtain 1 řrPN βpqqďrďcβpqq 1 1 1 q 1 q ă 8. 1 řnPA 1 n ă 8. n " řqPB r ď p1 ` log cq řqPB n ă 8 and therefore řnPA n ď řnPA1 Corollary 11.5. Let z P T be of infinite order and A " tn P N : can ď 1 ` zn ď dbnu, where 0 ă a ă b ă 1 and c, d ą 0. Then řnPA Proof. Since z has infinite order in T, z " eiπθ with θ P RzQ. Pick real numbers s, t such that 0 ă s ă a ă b ă t ă 1 and let A1 " tn P N : sn ď δpnθq ď tnu. It is an elementary exercise to see that AzA1 is finite. By Lemma 11.4, řnPA1 Lemma 11.6. For every z, w P T, Mpz, wq has Lebesgue measure 0. Proof. If w´1z has finite order in T, Theorem 1.13 implies that Mpz, wq " ∅. Thus we can assume that s " w´1z is of infinite order. If r P Mpz, wq, then the number rnzn ` wn " rn1 ` sn belongs to r1, 2s for infinitely many n P N. Let 0 ă a ă b ă 1. The last observation yields Mpz, wq X rb´1, a´1s Ď Na,b " Let A " tn P N : an ď 1 ` sn ď 2 bnu. By Corollary 11.5, řnPA n ă 8. Note that Kn,a,b " ∅ if n R A. On the other hand, if n P A, then Kn,a,b is contained in the interval r1 ` sn´1{n, 21{n1 ` sn´1{ns, whose length is p21{n ´ 1q1 ` sn´1{n ď 21{n´1 an . Thus, the Lebesgue measure µ of the sets Kn,a,b is Kn,a,b, where Kn,a,b " tr P rb´1, a´1s : 1 ď rn1 ` sn ď 2u. a ď 1 8ďn"m 8čm"1 1 n ă 8. 1 estimated as µpKn,a,bq " 0 if n R A and µpKn,a,bq ď 1 an if n R A. According to the above two displays, for each m P N, µpMpz, wq X rb´1, a´1sq ď µpKn,a,bq ď 8ÿn"m 1 a ÿnPA něm 1 n Ñ 0 as m Ñ 8. Hence µpMpz, wq X rb´1, a´1sq " 0 whenever 0 ă a ă b ă 1. Thus µpMpz, wqq " 0. Proposition 1.19 follows from Lemmas 11.1, 11.3 and 11.6. 12 Further comments and open questions The next proposition provides a peculiar collection of diagonal numerically hypercyclic operators on C3. Proposition 12.1. There is a diagonal T P SN HpC3q such that the restriction of T to each invariant 2-dimensional subspace is not weakly numerically hypercyclic. Proof. Let 1 ă r ă R and k, m be distinct integers. By Lemma 3.5, the set W " pz, wq P T2 : tRnpzkn ` zmnq ` rnwn : n P Z`u is dense in C( is a dense Gδ subset of T2. Pick pz, wq in W and consider the diagonal operator T on C3 with the numbers Rzk, Rzm and rw on the diagonal. Clearly, (1.10.1) is satisfied for T . By Theorem 1.10, T P SNHpC3q. On the other hand, the possible spectra of the restrictions of T to 2-dimensional invariant subspaces are tRzk, Rzmu, tRzk, rwu and tRzm, rwu. In any case, such a restriction is not weakly numerically hypercyclic according to Theorem 1.13. 37 As we have already observed, a self-adjoint operator is never numerically hypercyclic. Funnily enough a scalar multiple of a self-adjoint operator can turn out to be numerically hypercyclic. Proposition 12.2. Let T be a bounded self-adjoint operator on a Hilbert space H such that there exists a sequence tλnunPN in σpTq satisfying 1 ă λ1 ă ´λ2 ă λ3 ă ´λ4 ă . . . Then for any w P T of infinite order, wT P NHpHq. Proof. Let sk " 1 3 mintλk ´ λk´1,λk`1 ´ λku for k P N, where we assume λ0 " 1. Using the spectral theorem, we can find an orthonormal sequence txnunPN in H such that xT kxn, xny is always between pλn ´ snqk and pλn ` snqk and xT mxk, T nxjy " 0 provided k ‰ j. Let E be the linear span of txn : n P Nu and K be the closure of E in H. Fix y P E and t P Czt0u. Let k " maxtj P N : xy, xjy ‰ 0u. Since λj ` sj ă λj`1 ´ sj`1 for each j, xT ny, yy Ñ 8 and xT ny, yy " opxT nxk`1, xk`1yq as n Ñ 8. Furthermore, since λj alternate signs, xT ny, yy and xT nxk`1, xk`1y have opposite signs for every are positive for sufficiently sufficiently large odd positive integer n. Hence the numbers large odd n and form a convergent to 0 sequence. That is, for each sufficiently large odd n, we can find εn ą 0 such that ε2 . Furthermore, εn Ñ 0. The last equality can be easily rewritten as xT npy ` εnxk`1q, y ` εnxk`1y " t. Since w has infinite order, we can find a strictly increasing sequence tnjujPN of odd positive integers such that wnj Ñ t . Then yj Ñ y and tj " xpwTqnj yj, yjy Ñ t, where t yj " y ` εnj xk`1. Thus py, tq is in the closure of the set n " t´xT ny,yy xT nxk`1,xk`1y xT nxk`1,xk`1y t´xT ny,yy Λ " tpy,xpwTqny, yyq : y P K, n P Nu. Since E is dense in K and y P E and t P Czt0u are arbitrary, Λ is dense in K C. By Theorem U, there is y P K such that txpwTqny, yy : n P Nu is dense in C. Hence wT P NHpHq. The normal operator T in the following example is strongly numerically hypercyclic, while even its weak numeric hypercyclicity does not follow from Theorem 1.26. The example illustrates the limitations of Theorem 1.26. Example 12.3. Let trnunPN be a strictly decreasing sequence in p1,8q and A " t1, 2, 3, 4, 5u. Then there exists a sequence taj,nupj,nqPAN of finite order elements of T such that the diagonal operator T on ℓ2pA Nq with the numbers rnaj,n on the diagonal is strongly numerically hypercyclic. Proof. Let ε, d ą 0 be the numbers furnished by Lemma 3.9. For each n P N, let Un be the only cyclic subgroup of T of order n: Un " tz P T : zn " 1u. Clearly, Un is a π n -net in T. Hence there is n0 P N such that for each n ě n0, we can find a1, . . . , a5 P Un for which pa1, . . . , a5, a2 5q P Pε, where Pε is the set defined in (3.5). Fix a sequence tznunPN such that tzn : n P Nu is dense in C and zn ă dn for every n P N. Using the induction with respect to n, we shall construct the sequences taj,nupj,nqPAN in T, tcj,nupj,nqPAN of positive numbers, tknunPN of positive integers and tpnunPN of prime numbers such that (A1) aj,n P Upn for 1 ď j ď 5; 5řj"1 (A2) (A3) pq divides kn ´ 2kq for 1 ď q ă n; (A4) pn ą pn´1 and rkn´kn´1 5řj"1 ą n2n if n ě 2. j,q " 0 and cj,n " n j,n " zn 1, . . . , a2 5řj"1 cj,qa2kn cj,nakn rkn n ; rkn n , n can choose b1,1, . . . , b5,1 P Up1 such that pb1,1, . . . , b5,1, b2 such that p1 does not divide k1 and r´k1 The basis of the induction. Pick an arbitrary prime number p1 such that p1 ě n0. Then we 5,1q P Pε. Since r1 ą 1, there is k1 P N 1 ă 1. Since z1 P d D, Lemma 3.9 provides positive numbers 5řj"1 5řj"1 tj,1 " 1. Since bj,1 P Up1, p1 does not divide t1,1, . . . , t5,1 such that j,1 " bj,1 for 1 ď j ď 5. Set cj,1 " tj,1 k1 and p1 is prime, there are a1,1, . . . , a5,1 P Up1 such that ak1 for 1 ď j ď 5. Now it is straightforward to see that (A1) and (A2) with n " 1 are satisfied. Conditions (A3) and (A4) with n " 1 are satisfied in a trivial manner. Thus we have our basis of induction. tj,1bj,1 " z1, j,1 " 0 and 1,1, . . . , b2 5řj"1 tj,1b2 k1 1 r 38 The induction step. Let n ě 2 and assume that aj,q, cj,q, pq and kq with q ă n satisfying (A1 -- A4) are already constructed. First, we choose an arbitrary prime number pn ą pn´1. Since pn ą p1 ě n0, we can pick b1,n, . . . , b5,n P Upn such that pb1,n, . . . , b5,n, b2 5,nq P Pε. Since rn ą 1 and p1, . . . , pn are pairwise distinct primes, the Chinese Remainder Theorem allows us to choose kn P N such that rkn´kn´1 n P d D, Lemma 3.9 n 5řj"1 tj,n " 1. Since provides positive numbers t1,n, . . . , t5,n such that bj,n P Upn, pn does not divide kn and pn is prime, there are a1,n, . . . , a5,n P Upn such that akn j,n " bj,n for 1 ď j ď 5. Set cj,n " ntj,n for 1 ď j ď 5. Now it is straightforward to see that (A1 -- A4) are satisfied. This completes the construction of the sequences taj,nupj,nqPAN, tcj,nupj,nqPAN, tknunPN and tpnunPN satisfying (A1 -- A4). ą n2n, pn does not divide kn and pq divides kn´2kq for 1 ď q ď n´1. Since zn j,n " 0 and tj,nbj,n " zn n , 1,n, . . . , b2 5řj"1 5řj"1 tj,nb2 rkn n For pj, nq P A N, set λj,n " rnaj,n. Since each m P N, we can write 5řj"1 cj,n " n rkn n and rkn n ě n2n, we have c P ℓ1 `pA Nq. For ÿpj,nqPAN cj,nλkm j,n " ÿnPN αn,m, where αn,m " rkm n cj,nakm j,n. 5ÿj"1 First, note that by (A2), 5ÿj"1 Since for n ă m, pn divides km ´ 2kn and apn αm,m " rkm m cj,makm j,m " rkm m zm m " zm. rkm j,n " 1, (A2) yields αn,m " rkm n 5ÿj"1 cj,nakm j,n " rkm n 5ÿj"1 cj,na2kn j,n " 0 if n ă m. Finally, if n ą m, using (A4) and (A2), we obtain (12.1) (12.2) (12.3) (12.4) Combining (12.1 -- 12.4), we get αn,m " rkm cj,nakm n 5ÿj"1 zm ´ ÿpj,nqPAN n j,n ď rkm j,n " cj,nλkm 5ÿj"1 cj,n " nrkm n n ď rkn n rkn´kn´1 n 1 2n if n ą m. ď 8ÿn"m`1 αn,m ď 8ÿn"m`1 1 2n Ñ 0 as m Ñ 8. Since tzm : m P Nu is dense in C, the above display implies that! řpj,nqPAN j,n : k P N) is dense in C. Thus the diagonal operator T P Lpℓ2pA Nqq with the numbers λj,n on the diagonal satisfies (1.23.1). By Theorem 1.23, T is strongly numerically hypercyclic. cj,nλk Let V be the Volterra operator on L2r0, 1s. Exactly as in Example 1.21, we can show that re´V P WNHpL2r0, 1sq for every r ą 1. Unexpectedly enough it turns out that these operators are not numeri- cally hypercyclic. Example 12.4. Let r ą 1 and T " re´V , where V P LpL2r0, 1sq is the Volterra operator. Then T R NHpL2r0, 1sq. Proof. The key element of the proof is the following claim: }I ´ e´cV } ď 1 for every c P R`. (12.5) 39 We shall derive this estimate from the following observation: for every f P L2r0, 1s, the function Ff : R` Ñ R`, Ffpcq " }pI ´ e´cV qf} is increasing. (12.6) First, we shall verify (12.6). Obviously, it is enough to show that Gf " F 2 f is increasing. Differentiating by c the expression Gfpcq " xpI ´ e´cV qf,pI ´ e´cV qfy, we get G1fpcq " xV gc, gcy ` xgc, V gcy, where gc " pI ´ e´cV qf . Hence G1fpcq " xP gc, gcy, where P " V ` V ‹ and V ‹ is the Hilbert space adjoint of V . It is easy to see that P is the orthogonal projection onto the one-dimensional space of constant functions. Then the self-adjoint operator P is non-negative definite and G1fpcq " xP gc, gcy ě 0 for every c P R`. Hence Gf is increasing and (12.6) follows. Now we shall prove (12.5). Assume the contrary. Then there is a ą 0 such that }I ´ e´aV } ą 1. Then there is f P SpL2r0, 1sq such that }pI ´ e´aV qf} " Ffpaq ą 1. By (12.6), Ffpcq ą 1. On the other hand, in [13] it is shown that }pI ´ V qng} Ñ 0 for each g P L2r0, 1s. According to the main result of [5] the operators I ´ V and e´V are similar. Hence }e´nV g} Ñ 0 for each g P L2r0, 1s. In particular, }e´nV f} Ñ 0. Hence Ffpnq Ñ 1, which contradicts the inequality lim Ffpcq ą 1. This contradiction cÑ8 completes the proof of (12.5). lim cÑ8 Now let r ą 1 and T " re´V . Then for each f P SpL2r0, 1sq, using (12.5), we have 2 xpe´nV `e´nV ‹ RexT nf, fy " rn qqf, fyq ě rnp1´}I´e´nV }q ě 0. Hence NOpT, fq is contained in the right half-plane and therefore can not be dense in C. Thus T R NHpL2r0, 1sq. 2 p2´xppI´e´nV q`pI´e´nV ‹ qf, fy " rn At the expense of a number of technical details entering the proof, one can similarly show that rpI ´ cV q is weakly numerically hypercyclic and not numerically hypercyclic whenever r ą 1 and c ą 0. We shall raise a number of natural questions. Question 12.5. Let T P WNHpCnq. Is it true that at least one of the conditions p1.9.1 -- 1.9.4q is satisfied? Question 12.6. Assume that X is reflexive, T P LpXq is not power bounded and σppTq " ∅. Is it true that T is weakly numerically hypercyclic? Question 12.7. Let T P WNHpXq. Is it true that T n P WNHpXq for every n P N? Question 12.8. Let T P WNHpXq. Is it true that rT P WNHpXq for every r ě 1? Question 12.9. Let T be a unitary operator such that σpTq is infinite and σppTq " ∅. Is rT strongly numerically hypercyclic for each r ą 1? Question 12.10. Let T P LpXq and λ1, λ2 P C are such that kerpT ´ λjIq2 ‰ kerpT ´ λjIq for j P t1, 2u, λ1 " λ2 ě 1 and λ1 Is it true that T must be strongly λ1 numerically hypercyclic? are independent in T. , λ2 λ2 Question 12.11. Characterize normal numerically hypercyclic operators in terms of their spectrum. Question 12.12. Characterize numerically hypercyclic operators on L2r0, 1s commuting with the Volterra operator V . In particular, is I ` V numerically hypercyclic? The last question is in the spirit of the Invariant Subspace Problem and probably is tough. Question 12.13. Is there T P Lpℓ2pNqq such that NOpT, xq is dense in C for each x P Spℓ2pNqq? 40 References [1] C. Badea and V. Muller, On weak orbits of operators, Topology and its Applications 156 (2009), 1381-1385 [2] F. Bayart and E. Matheron, Dynamics of linear operators, Cambridge University Press, 2009 [3] S. Bermudo, A. Montes-Rodr´ıguez and S. Shkarin, Orbits of operators commuting with the Volterra operator, J. Math. Pures Appl. 89 (2008), 145 -- 173 [4] N. Feldman, N -weakly hypercyclic and N -weakly supercyclic operators, J. Functional Analysis [to appear] [5] R. Frankfurt and J. Rovnyak, Finite convolution operators, J. Math. Anal. Appl. 49 (1975), 347-374. [6] K. Grosse-Erdmann, Universal families and hypercyclic operators, Bull. Amer. Math. Soc. 36 (1999), 345 -- 381 [7] K. Grosse-Erdmann and A. Peris, Linear Chaos, Springer, 2011 [8] D. Herrero, Limits of hypercyclic and supercyclic operators, J. Funct. Anal. 99 (1991), 179 -- 190 [9] D. Herrero and Z. Wang, Compact perturbations of hypercyclic and supercyclic operators, Indiana Univ. Math. J. 39 (1990), 819 -- 829 [10] S. Kim, A. Peris and H. Song, Numerically hypercyclic operators, Integral Equations and Operator Theory, 72 (2012), 393 -- 402 [11] S. Kim, A. Peris and H. Song, Numerically hypercyclic polynomials, Archiv der Mathematik, 99 (2012), 443 -- 452 [12] F. Le´on-Saavedra and V. Muller, Rotations of hypercyclic and supercyclic operators, Integral Equations and Operator Theory 50 (2004), 385 -- 391 [13] A. Montes-Rodr´ıguez, J. S´anchez- ´Alvarez and J. Zem´anek, Uniform Abel-Kreiss boundedness and the extremal behaviour of the Volterra operator, Proc. London Math. Soc. 91 (2005), 761-788 [14] G. Prajitura, Limits of weakly hypercyclic and supercyclic operators, Glasg. Math. J. 47 (2005), 255-260 [15] H. Salas, Hypercyclic weighted shifts, Trans. Amer. Math. Soc., 347 (1995), 993 -- 1004 Stanislav Shkarin Queens's University Belfast Department of Pure Mathematics University road, Belfast, BT7 1NN, UK E-mail address: [email protected] 41
1302.6295
2
1302
2013-11-26T23:28:14
Spectral analysis of the truncated Hilbert transform with overlap
[ "math.FA", "math.CA", "math.SP" ]
We study a restriction of the Hilbert transform as an operator $H_T$ from $L^2(a_2,a_4)$ to $L^2(a_1,a_3)$ for real numbers $a_1 < a_2 < a_3 < a_4$. The operator $H_T$ arises in tomographic reconstruction from limited data, more precisely in the method of differentiated back-projection (DBP). There, the reconstruction requires recovering a family of one-dimensional functions $f$ supported on compact intervals $[a_2,a_4]$ from its Hilbert transform measured on intervals $[a_1,a_3]$ that might only overlap, but not cover $[a_2,a_4]$. We show that the inversion of $H_T$ is ill-posed, which is why we investigate the spectral properties of $H_T$. We relate the operator $H_T$ to a self-adjoint two-interval Sturm-Liouville problem, for which we prove that the spectrum is discrete. The Sturm-Liouville operator is found to commute with $H_T$, which then implies that the spectrum of $H_T^* H_T$ is discrete. Furthermore, we express the singular value decomposition of $H_T$ in terms of the solutions to the Sturm-Liouville problem. The singular values of $H_T$ accumulate at both $0$ and $1$, implying that $H_T$ is not a compact operator. We conclude by illustrating the properties obtained for $H_T$ numerically.
math.FA
math
SPECTRAL ANALYSIS OF THE TRUNCATED HILBERT TRANSFORM WITH OVERLAP REEMA AL-AIFARI∗ AND ALEXANDER KATSEVICH† Abstract. We study a restriction of the Hilbert transform as an operator HT from L2(a2, a4) to L2(a1, a3) for real numbers a1 < a2 < a3 < a4. The operator HT arises in tomographic reconstruc- tion from limited data, more precisely in the method of differentiated back-projection (DBP). There, the reconstruction requires recovering a family of one-dimensional functions f supported on compact intervals [a2, a4] from its Hilbert transform measured on intervals [a1, a3] that might only overlap, but not cover [a2, a4]. We show that the inversion of HT is ill-posed, which is why we investigate the spectral properties of HT . We relate the operator HT to a self-adjoint two-interval Sturm-Liouville problem, for which we prove that the spectrum is discrete. The Sturm-Liouville operator is found to commute with HT , which then implies that the spectrum of H ∗ T HT is discrete. Furthermore, we express the singular value decomposition of HT in terms of the solutions to the Sturm-Liouville problem. The singular values of HT accumulate at both 0 and 1, implying that HT is not a compact operator. We conclude by illustrating the properties obtained for HT numerically. 1. Introduction. In tomographic imaging, which is widely used for medical applications, a 2D or 3D object is illuminated by a penetrating beam (usually X- rays) from multiple directions, and the projections of the object are recorded by a detector. Then one seeks to reconstruct the full 2D or 3D structure from this collection of projections. When the beams are sufficiently wide to fully embrace the object and when the beams from a sufficiently dense set of directions around the object can be used, this problem and its solution are well understood [16]. When the data are more limited, e.g. when only a reduced range of directions can be used or only a part of the object can be illuminated, the image reconstruction problem becomes much more challenging. This method goes back to a result by Gelfand and Graev [6]. Reconstruction from limited data requires the identification of specific subsets of line integrals that allow for an exact and stable reconstruction. One class of such configurations that have already been identified, relies on the reduction of the 2D and 3D reconstruction problem to a family of 1D problems. The Radon transform can be related to the 1D Hilbert transform along certain lines by differentiation and back-projection of the Radon transform data (differentiated back-projection or DBP). Inversion of the Hilbert transform along a family of lines covering a sub-region of the object (region of interest or ROI) then allows for the reconstruction within the ROI. Its application to tomography was formulated by Finch [4] and was later made explicit for 2D in [17, 23, 28] and for 3D in [18, 24, 27, 29]. To reconstruct from data obtained by the DBP method, it is necessary to solve a family of 1D problems which consist of inverting the Hilbert transform data on a finite segment of the line. If the Hilbert transform Hf of a 1D function f was given on all of R, then the inversion would be trivial, since H −1 = −H. In case f is compactly supported, it can be reconstructed even if Hf is not known on all of R. Due to an explicit reconstruction formula by Tricomi [22], f can be found from measuring Hf only on an interval that covers the support of f . However, a limited field of view might result in configurations in which the Hilbert transform is known only on a segment that does not completely cover the object support. One example of such a configuration is known as the interior ∗Department of Mathematics, Vrije Universiteit Brussel, Brussels B-1050, Belgium †Department of Mathematics, University of Central Florida, FL 32816, USA 1 problem [1, 10, 12, 25]. Given real numbers a1 < a2 < a3 < a4, the interior problem corresponds to the case in which the Hilbert transform of a function supported on [a1, a4] is measured on the smaller interval [a2, a3]. In this paper, we study a different configuration, namely supp f = [a2, a4] and the Hilbert transform is measured on [a1, a3]. We will refer to this configuration as the truncated problem with overlap: the operator HT we consider is given by P[a1,a3]HP[a2,a4], where H is the usual Hilbert transform acting on L2(R), and PΩ stands for the projection operator (PΩf )(x) = f (x) if x ∈ Ω, (PΩf )(x) = 0 otherwise. For finite intervals Ω1, Ω2 on R, the interior problem corresponds to PΩ1 HPΩ2 for Ω1 ⊂ Ω2. The truncated Hilbert transform with a gap occurs when the intervals Ω1 and Ω2 are separated by a gap, as in [8]. Figure 1.1 shows the different setups. Examples of configurations in which the truncated Hilbert transform with overlap and the interior problem occur are given in Figures 1.2 and 1.3. The truncated problem with overlap arises for example in the "missing arm" problem. This is the case where the field of view is large enough to measure the torso but not the arms. (a) Finite Hilbert transform. (b) Interior problem. (c) Truncated Hilbert transform with overlap. (d) Truncated Hilbert transform with a gap. Fig. 1.1: Different setups for PΩ1 HPΩ2. The upper interval shows the support Ω2 of the function f to be reconstructed. The lower interval is the interval Ω1 where measurements of the Hilbert transform Hf are taken. This paper investigates case (c). 2 (a) (b) Fig. 1.2: Two examples of the truncated problem with overlap, Fig. 1.2(a) shows the missing arm problem. In both cases, the field of view (FOV) does not cover the object support. On the line intersecting the object, measurements can only be taken within the FOV, i.e. from a1 to a3. The Hilbert transform is not measured on [a3, a4]. Consequently, a reconstruction can only be aimed at in the grey-shaded intersection of the FOV with the object support, called the region of interest (ROI). Fig. 1.3: The interior problem. Here, the FOV also does not cover the object support. The line intersecting the object is such that the Hilbert transform is only measured in a subinterval [a2, a3] of the intersection [a1, a4] of the line with the object support. The ROI is the grey-shaded intersection of the FOV with the object support. In this case stable reconstruction of the shaded ROI is impossible unless additional information is available. 3 Fix any four real numbers a1 < a2 < a3 < a4. We define the truncated Hilbert transform with overlap as the operator Definition 1.1. f (y) y − x where p.v. stands for the principal value. In short, p.v. Z a4 (HT f )(x) := 1 π a2 dy, x ∈ (a1, a3) (1.1) HT := P[a1,a3]HP[a2,a4], where H is the ordinary Hilbert transform on L2(R). As we will prove in what follows, the inversion of HT is an ill-posed problem in the sense of Hadamard [3]. In order to find suitable regularization methods for its inversion, it is crucial to study the nature of the underlying ill-posedness, and therefore the spectrum σ(H ∗ T HT ). An important question that arises here is whether the spectrum is purely discrete. This question has been answered for similar operators before, but with two very different answers. In [11], it was shown that the finite Hilbert transform defined as HF = P[a,b]HP[a,b] has a continuous spectrum σ(HF ) = [−i, i]. On the other hand, in [9], we find the result that for the interior problem HI = P[a2,a3]HP[a1,a4], the spectrum σ(H ∗ In addition, we obtain that 0 and 1 are accumulation points of the spectrum. Further- more, we find that the singular value decomposition (SVD) of the operator HT can be related to the solutions of a Sturm-Liouville (S-L) problem. For the actual recon- struction, one would aim at finding f in (1.1) only within a region of interest (ROI), i.e. on [a2, a3]. A stability estimate as well as a uniqueness result for this setup were obtained by Defrise et al in [2]. A possible method for ROI reconstruction is the trun- cated SVD. Thus, it is of interest to study the SVD of HT also for the development of reconstruction algorithms. The main result of this paper is that H ∗ T HT has only discrete spectrum. I HI ) is purely discrete. In [8] and [9], singular value decompositions are obtained for the truncated Hilbert transform with a gap P[a3,a4]HP[a1,a2] and for HI . This is done by relating the Hilbert transforms to differential operators that have discrete spectra. We follow this proce- dure, but obtain a differential operator that is different in nature. In [8] and [9] the discreteness of the spectra follows from standard results of singular S-L theory (see e.g. [26]). In the case of truncated Hilbert transform (1.1) we have to investigate the discreteness of the spectrum of the related differential operator explicitly. The idea is to find a differential operator for which the eigenfunctions are the singular functions of HT on (a2, a4). We define the differential operator similarly to the one in [8], [9], but then the question is which boundary conditions to choose in order to relate the differential operator to HT . To answer this question we first develop an intuition about the singular functions of HT . Let {σn; fn, gn} denote the singular system of HT that we want to find. The problem can be formulated as finding a complete orthonormal system {fn}n∈N in L2(a2, a4) and an orthonormal system {gn}n∈N in L2(a1, a3) such that there exist real numbers σn for which HT fn = σngn, H ∗ T gn = σnfn. 4 At the moment, the gn's only have to be complete in Ran(HT ), but as we will see in Section 5, Ran(HT ) is dense in L2(a1, a3). As will be shown in Section 4, the functions fn and gn (a) can only be bounded or of logarithmic singularity at the points ai, 1 , a− (b) do not vanish at the edges of their supports (a+ We will now make use of the following results from [5], Sections 8.2 and 8.5: 4 for fn, and a+ 2 , a− 3 for gn). Lemma 1.2 (Local properties of the Hilbert transform). Let f be a function with support [b, d] ⊂ R. And let c be in the interior of [b, d]. 1. If f is Holder continuous (for some Holder index α) on [b, d], then close to b the Hilbert transform of f is given by (Hf )(x) = − 1 π f (b+) lnx − b + H0(x) (1.2) where H0 is bounded and continuous in a neighborhood of b. 2. If in a neighborhood of c, the function f is of the form f (x) = f (x) ln x − c for Holder continuous f , then close to the point c its Hilbert transform is of the form (Hf )(x) = H0(x), where H0 is bounded with a possible finite jump discontinuity at c. 3. If f is of the form f (x) = f (x) ln x − b on [b, c], where f is Holder contin- uous, then its Hilbert transform at b has a singularity of the order ln2 x − b if f (b) 6= 0. Suppose fn has a logarithmic singularity at a+ 2 . Since HT integrates over [a2, a4], the function HT fn would have a singularity at a2 of order ln2 x − a2. Hence, this would violate the property of gn at a2. Therefore, fn has to be bounded at a+ 2 . If fn does not vanish at a+ 2 , this leads to logarithmic singularities of HT fn and gn at a2. Using the same argument we conclude that gn is bounded at a− 3 and fn has a logarithmic singularity at a3. On the other hand, since gn is bounded at a− 3 , HT fn is also bounded there. This requires that close to a3, fn = fn,1 + fn,2 lnx− a3 for functions fn,i continuous at a3. A similar argument holds for gn at a2. Close to that point, gn = gn,1 + gn,2 lnx− a2 for functions gn,i continuous at a2. Clearly, HT fn is bounded at a+ 4 . Therefore, fn has to be bounded at a− 1 and HT gn is bounded at a− 4 and gn must be bounded at a+ 1 . Thus, if we want to show the commutation of HT with a differential operator that acts on fn(x), x ∈ (a2, a4), we need to impose boundary conditions at a+ 4 that require boundedness and some transmission conditions at a3 that make the bounded term and the term in front of the logarithm in fn continuous at a3. 2 and a− Having found these properties of the singular functions of HT (in case the SVD for HT exists), in Section 2 we introduce a differential operator and find a self-adjoint extension for this operator. We then show in Section 3 that this self-adjoint differential operator LS has a discrete spectrum. In Section 4 we establish that LS commutes with the operator HT . This allows us to find the SVD of HT . In Section 5 we then 5 (a) Sketch of fn's. (b) Sketch of gn's. Fig. 1.4: Intuition about the singular functions of HT . study the accumulation points of the singular values of HT . In particular, we find that HT is not a compact operator. Finally, we conclude by showing numerical examples in Section 6. 2. Introducing a differential operator. In this section, we find two differen- tial operators LS and LS that will turn out to have a commutation property of the form HT LS = LSHT . (2.1) In order to find the SVD of HT , we will be interested in finding LS and LS with simple discrete spectra. Initially, it is not apparent whether differential operators with such properties exist and if so, how to find them. We do not know of a coherent theory that relates certain integral operators to differential operators via a commutation property as the above. However, there have been examples of integral operators for which -- by what seems to be a lucky accident -- such differential operators exist. One instance is the well-known Landau-Pollak-Slepian (LPS) operator that arises in signal processing in the study of time- and bandlimited representations of signals [20, 13, 14]. There, it is of interest to find the largest eigenvalue of the LPS operator P[−T,T ]F −1P[−W,W ]FP[−T,T ]. Here, F is the Fourier transform, and T and W are some positive numbers. This operator happens to commute with a second order differential operator, of which the eigenfunctions and eigenvalues had been studied long before its connection to the LPS operator was known. The eigenfunctions of this differential operator are the so-called prolate spheroidal wave functions and they turn out to be the eigenfunctions of the LPS operator as well. The work of Landau, Pollak and Slepian has been generalized and extended by Grunbaum et al. [7]. More recent examples of integral operators with commuting differential operators are the interior Radon transform [15] and two instances of the truncated Hilbert transform mentioned earlier [8, 9]. To start our search for LS and LS, we follow the procedure in [8, 9] and define a differential operator Definition 2.1. L(x, dx)ψ(x) := (P (x)ψ′(x))′ + 2(x − σ)2ψ(x) 6 (2.2) where P (x) = 4 Yi=1 (x − ai), σ = 1 4 4 Xi=1 ai. (2.3) The four points ai are all regular singular, and in a complex neighborhood of each ai the functions (x− ai)· P ′(x)/P (x) and (x− ai)2· 2(x− σ)2/P (x) are complex analytic. The term regular singular point is standard in the general theory of differential equa- tions and, as such, is also used in the theory of S-L equations, see e.g. [21] for this and other terminology and basic properties of S-L equations. Consequently, by the method of Fuchs-Frobenius it follows that for λ ∈ C any solution of Lψ = λψ is either bounded or of logarithmic singularity close to any of the points ai, see [21]. Away from the singular points ai the analyticity of the solutions follows from the analyticity of the coefficients of the differential operator L. More precisely, in a left and a right neighborhood of each regular singular point ai, there exist two linearly independent solutions of the form ψ1(x) = x − aiα1 ψ2(x) = x − aiα2 ∞ Xn=0 Xn=0 ∞ bn(x − ai)n, dn(x − ai)n + k lnx − aiψ1(x), (2.4) (2.5) where without loss of generality we can assume b0 = d0 = 1. The exponents α1 and α2 are the solutions of the indicial equation where α2 + (p0 − 1)α + q0 = 0, p0 = lim x→ai q0 = lim x→ai (x − ai)P ′(x)/P (x), (x − ai)2[2(x − σ)2 − λ]/P (x). (2.6) (2.7) With our choice of P , this gives α1 = α2 = 0 which implies k 6= 0. For the bounded solution in (2.4), α1 = 0 results in ψ1(ai) 6= 0. The radius of convergence of the series in (2.4) and (2.5) is the distance to the closest singular point different from ai. In a left and in a right neighborhood of ai, the general form of the solutions of (L−λ)ψ = 0 is ψ1(x) = ℓ0 ψ2(x) = ℓ1 ∞ Xn=0 Xn=0 ∞ bn(x − ai)n dn(x − ai)n + ℓ2 lnx − ai ∞ Xn=0 bn(x − ai)n (2.8) (2.9) for some constants ℓj. Hence we have one degree of freedom for the bounded solution, and two -- for the unbounded solution. Clearly, for the bounded solutions (2.8), the coefficients bn are the same on both sides of ai, since we have assumed b0 = 1. However, the bounded part of the unbounded solutions (2.9) may have different coefficients d− n to the left and to the right of ai respectively. n and d+ 7 2.1. The Maximal and Minimal Domains and Self-Adjoint Realiza- tions. Since we are interested in a differential operator that commutes (on some set to be defined) with HT , we want to consider L on the interval (a2, a4). Due to the regular singular point a3 in the interior of the interval, standard techniques for singular S-L problems are not applicable. It is crucial for our application that we identify a commuting self-adjoint operator, for which the spectral theorem can be ap- plied. We therefore wish to study all self-adjoint realizations; we follow the treatment in Chapter 13 in [26] which gives a characterization of all self-adjoint realizations for two-interval problems, of which problems with an interior singular point are a special case. First of all, one needs to define the maximal and minimal domains on Ij = (aj, aj+1) (see Chapter 9 in [26]). Let ACloc(I) be the set of all functions that are absolutely continuous on all compact subintervals of the open interval I. Then, Dj,max := {ψ : Ij → C : ψ, P ψ′ ∈ ACloc(Ij ); ψ, Lψ ∈ L2(Ij )}, Dj,min := {ψ ∈ Dj,max : supp ψ ⊂ (aj , aj+1)}, and the related maximal and minimal operators are defined as follows: Lj,max := L(Dj,max) : Dj,max → L2(Ij ), Lj,min := L(Dj,min) : Dj,min → L2(Ij ). (2.10) (2.11) (2.12) (2.13) We shall follow essentially the procedure in Chapter 13 in [26], to which we refer for more details. On (a2, a4), the maximal and minimal domains and the corresponding operators are defined as the direct sums: Definition 2.2. The maximal and minimal domains Dmax, Dmin ⊂ L2(a2, a4) and the operators Lmax, Lmin are defined as Dmax := D2,max + D3,max, Lmax := L2,max + L3,max, Dmin := D2,min + D3,min Lmin := L2,min + L3,min and therefore Lmax : Dmax → L2(a2, a4), Lmin : Dmin → L2(a2, a4). (2.14) (2.15) The operator Lmin is a closed, symmetric, densely defined operator in L2(a2, a4) and Lmax, Lmin form an adjoint pair, i.e. L∗ min = Lmax. In order to define a self-adjoint extension of Lmin, we need to introduce the notion of the Lagrange sesquilinear form: max = Lmin and L∗ where, at the singular points, [u, v] := uP v′ − vP u′, [u, v](a+ i ) := lim α→a+ i [u, v](a− i ) := lim α→a− i [u, v](α), [u, v](α). 8 (2.16) (2.17) (2.18) These limits exist and are finite for all u, v ∈ Dmax. If we choose u, v ∈ Dmax such that [u, v](ai) = 1 for all the singular points (a+ 4 ), then the extension of Lmin defined by the following conditions 3 , a+ 2 , a− 3 , a− [ψ, u](a+ [ψ, u](a− [ψ, v](a− 2 ) = 0 = [ψ, u](a− 4 ) 3 ) = [ψ, u](a+ 3 ) 3 ) = [ψ, v](a+ 3 ) (2.19) (2.20) (2.21) is self-adjoint. We refer to (2.19) as boundary conditions, and to (2.20) and (2.21) -- as transmission conditions. The latter connect the two subintervals (a2, a3) and (a3, a4). Motivated by the conditions mentioned in Section 1, we define a self-adjoint extension of Lmin: Lemma 2.3. The extension LS : D(LS) → L2(a2, a4) of Lmin to the domain D(LS) := {ψ ∈ Dmax : [ψ, u](a+ [ψ, u](a− 2 ) = [ψ, u](a− 3 ) = [ψ, u](a+ 4 ) = 0, 3 ), [ψ, v](a− 3 ) = [ψ, v](a+ 3 )} (2.22) with the following choice of maximal domain functions u, v ∈ Dmax u(y) := 1, v(y) := 4 Xi=1 Yj6=i j∈{1,...,4} 1 ai − aj lny − ai, (2.23) (2.24) is self-adjoint. This choice of maximal domain functions gives [u, v](ai) = 1 for i = 1, . . . , 4. The boundary conditions simplify to lim y→a+ 2 P (y)ψ′(y) = lim y→a− 4 P (y)ψ′(y) = 0. (2.25) For an eigenfunction ψ of LS this is equivalent to ψ being bounded at a+ 2 and a− 4 (because the only possible singularity is of logarithmic type). Let φ1 and φ2 be the restrictions of ψ to the intervals (a2, a3) and (a3, a4), respectively. Since ψ is an eigenfunction, on the corresponding intervals φ1 and φ2 are of the form φi(y) = φi1(y) + φi2(y) lny − a3. Here, the functions φij are analytic on (a2, a3) for i = 1 and on (a3, a4) -- for i = 2. Having this, the transmission conditions can be simplified as follows: 3 ) = [ψ, u](a− 3 ) [ψ, u](a+ P (y)ψ′(y) = lim y→a− 3 P (y)ψ′(y) lim y→a+ 3 The condition involving v yields lim y→a− 3 φ12(y) = lim y→a+ 3 φ22(y). lim y→a− 3 [ψ, v](a− [ψ(y) − v(y)(P ψ′)(y)] = lim φ11(y) = lim y→a+ 3 lim y→a− 3 3 ) = [ψ, v](a+ 3 ) [ψ(y) − v(y)(P ψ′)(y)] φ21(y) y→a+ 3 9 (2.26) (2.27) (2.28) Note that on each side of (2.27) the logarithmic terms in φi2 cancel because of the choice of the constants in v. The properties (2.25), (2.26) and (2.28) are the same as the ones found for fn in Section 1. Thus, we have constructed an operator LS for which close to the points a2, a3 and a4, the eigenfunctions behave in the same way that is expected for the fn's. Close to a3, an eigenfunction ψ is given by ψ(y) =   ℓ11 ℓ12 ∞ ∞ Pm=0 Pm=0 d− m(y − a3)m + ℓ21 lny − a3 d+ m(y − a3)m + ℓ22 lny − a3 ∞ ∞ Pm=0 Pm=0 bm(y − a3)m, bm(y − a3)m, y < a3 y > a3 (2.29) where similarly to (2.5), we assume d− conditions require that 0 = d+ 0 = 1 and b0 = 1. The transmission (2.30) (2.31) (2.32) ℓ11 = ℓ12, ℓ21 = ℓ22. We can thus express ψ in a sufficiently small neighborhood of a3 as ∞ ∞ ψ(y) = ℓ11 + ℓ21 lny − a3 Xm=0 bm(y − a3)m + Xm=1 m = ℓ11d− m, when y > a3 and for ℓ− ℓ± m(y − a3)m, m, when y < a3. where ℓ± m stands for ℓ+ m = ℓ11d+ 3. The spectrum of LS. In order to prove that the spectrum of the differential self-adjoint operator LS introduced in Lemma 2.3 is discrete, we need to show that for some z in the resolvent set, (LS − zI)−1 is a compact operator. To do so, it is sufficient to prove that the Green's function G of LS − zI, which for z in the resolvent set exists and is unique, is a function in L2((a2, a4)2). This would allow us to conclude that the integral operator TG with G as its integral kernel is a compact operator from L2(a2, a4) to L2(a2, a4), where TG is equivalent to the inversion of LS − zI. Lemma 3.1. The Green's function G(x, ξ) associated with LS−i is in L2((a2, a4)2) and consequently, (LS − i)−1 : L2(a2, a4) → D(LS) ⊂ L2(a2, a4) is a compact opera- tor. Proof. The self-adjointness of LS is equivalent to LS − i being one-to-one and onto (Theorem VIII.3 in [19]). Moreover, the ai's are limit-circle points and thus, the deficiency index d equals 4 (Theorem 13.3.1 in [26]). This means that if we do not impose boundary and transmission conditions, there are two linearly independent solutions p1 and p2 of (L− i)p = 0 on (a2, a3) as well as two linearly independent solu- tions q1 and q2 of (L − i)q = 0 on (a3, a4). Note that none of these four solutions can be bounded at both of its endpoints because i is not an eigenvalue of the self-adjoint operator Lj,S : D(Lj,S) → L2(Ij ) with D(Lj,S) = {ψ ∈ Dj,max : limy→a+ P (y)ψ′(y) = P (y)ψ′(y) = 0}. By taking appropriate combinations, if necessary, we can limy→a− eliminate the logarithmic singularity at a+ 4 -- of another solution. We can thus assume that 2 of one of the solutions, and at a− j+1 j - on (a2, a3): p1 is bounded at a+ 2 and logarithmic at a− 3 , p2 is logarithmic at both endpoints; 10 - on (a3, a4): q1 is logarithmic at a+ 3 and bounded at a− 4 , q2 is logarithmic at both endpoints. We next check the restrictions imposed by the transmission conditions at a3. Close to a3, both functions p1 and q2 are of the form (2.9). Let ℓ11, ℓ21 denote the free parameters in the expression for p1 and ℓ12, ℓ22 the ones in q2. These can be chosen such that they satisfy (2.30) and (2.31). Thus, there exists a solution h1(x) on (a2, a4) given by q2(x) h1(x) = (cid:26) p1(x) 2 and logarithmic at a− for x ∈ (a2, a3) for x ∈ (a3, a4) that is bounded at a+ 4 . In addition, it is of the form (2.32) close to a3, i.e. it is logarithmic at a3 and satisfies the transmission conditions (2.26), (2.28) there. Similarly, with p2 and q1 we can obtain a solution h2 on (a2, a4) that satisfies the transmission conditions at a3 and is of ln-ln-bounded-type. Thus, imposing only the transmission conditions, we obtain two linearly independent solutions of (L − i)h = 0 on (a2, a3) ∪ (a3, a4). One of them, h1, is of a bounded-ln-ln-type, and the other one, h2, is of a ln-ln-bounded-type, at the points a+ 4 , respectively. We are now in a position to consider the Green's function G(x, ξ) of LS − i. Close to a3, we can write the two functions as hj(x) = hj1(x) + lnx − a3hj2(x) with continuous functions hj1 and hj2. By rescaling if necessary, we can assume h12(a3) = h22(a3). We construct G from h1 and h2 as follows: 2 , a3, a− G(x, ξ) = (cid:26) c1(ξ)h1(x) c2(ξ)h2(x) for x < ξ for x > ξ (3.1) where ξ ∈ (a2, a3) ∪ (a3, a4) and the functions c1(ξ) and c2(ξ) are chosen such that G is continuous at x = ξ and ∂G/∂x has a jump discontinuity of 1/P (ξ) at x = ξ: c1(ξ)h1(ξ) − c2(ξ)h2(ξ) = 0, c1(ξ)h′ 2(ξ) = − 1(ξ) − c2(ξ)h′ 1 P (ξ) . (3.2) (3.3) In other words, G is the solution of (L − i)G = δ, where δ is the Dirac delta function. For ξ away from a3, G(x, ξ) is continuous in ξ but with logarithmic singularities at a+ 2 and a− 4 . This can be seen as follows. Consider ξ close to a2. There, we can write h2(ξ) = h21(ξ) + h22(ξ) lnξ − a2 and, since h1 is bounded close to a+ Wh1,h2 denote the Wronskian of h1 and h2, i.e. Wh1,h2 = h1h′ we obtain 2 , it is of the form (2.8), i.e. h1(a+ 2 − h′ 2 ) 6= 0. Let 1h2. For c1 and c2 c1(ξ) = c2(ξ) = h2(ξ) P (ξ)Wh1,h2(ξ) h1(ξ) P (ξ)Wh1,h2(ξ) , . (3.4) (3.5) The denominator in the above expressions is bounded by P (ξ)(h1(ξ)h′ 2(ξ) − h2(ξ)h′ 1(ξ)) = O((ξ − a2) lnξ − a2) + h1(ξ)h22(ξ)p(ξ), 11 where p(ξ) = P (ξ)/(ξ − a2) and h1(a+ a+ 2 , 2 )h22(a2)p(a2) 6= 0. Thus, in a neighborhood of c1(ξ) = O(ln ξ − a2), c2(ξ) = O(1). Similarly, since h2(a− 4 ) 6= 0, close to a− 4 c1(ξ) = O(1), c2(ξ) = O(ln ξ − a4). (3.6) (3.7) (3.8) (3.9) For each fixed ξ ∈ (a2, a3) ∪ (a3, a4), G(x, ξ) as a function in x is continuous on [a2, a3) ∪ (a3, a4] and has a logarithmic singularity at a3, due to the singularities in h1(x) and h2(x). It remains to check what happens as ξ → a3. We need to make sure that the functions c1(ξ) and c2(ξ) behave in such a way that G ∈ L2((a2, a4)2). Therefore, we derive the asymptotics of c1(ξ) and c2(ξ) as ξ → a− 3 . For ξ = a3 − ǫ and small ǫ > 0, equation (3.2) becomes c1(a3 − ǫ)h1(a3 − ǫ) − c2(a3 − ǫ)h2(a3 − ǫ) = 0. Since close to a3, hi = hi1 + hi2 ln(ǫ) and the hij are continuous, the ratio c1/c2 is of the form a + b ln(ǫ) c + d ln(ǫ) , where b and d are non-zero (because the logarithmic singularity is present). Thus, the ratio tends to the finite limit b/d as ǫ → 0. Conditions (3.2) and (3.3) together imply: c2(a3 − ǫ) = = where h1(a3 − ǫ) P (a3 − ǫ)Wh1,h2(a3 − ǫ) h11(a3 − ǫ) + h12(a3 − ǫ) ln(ǫ) r1(ǫ) + ǫ · r2(ǫ) 1 ǫ P (a3 − ǫ)(cid:0)h21h12 − h22h11(cid:1)(a3 − ǫ), r1(ǫ) = − r2(ǫ) = O(1), If r1(0) 6= 0, then c2 is of order O(ln(ǫ)), removing a possible and h12(a3) 6= 0. obstruction to square integrability of G. Suppose r1(0) = 0, i.e. h21(a3)h12(a3) − h22(a3)h11(a3) = 0. This would imply h11(a3) = C · h21(a3), h12(a3) = C · h22(a3) 12 (3.10) (3.11) for some constant C. By assumption, h12(a3) = h22(a3), so that C = 1. Now if both (3.10) and (3.11) hold for C = 1, the function defined by h(x) = (cid:26) h1(x) h2(x) for x ∈ (a2, a3) for x ∈ (a3, a4) would be a non-trivial solution of (LS − i)h = 0 (fulfilling both boundary and trans- mission conditions), i.e. i would be an eigenvalue of LS. But this contradicts the self- adjointness of LS. We can thus conclude that r1(0) 6= 0. This shows that c2(a3 − ǫ) is of order O(ln(ǫ)) and therefore also c2 · c1 Analogously, we can find the same asymptotics of c1(ξ) and c2(ξ) as ξ → a+ 3 . Therefore, the properties of the Green's function G(x, ξ) can be summarized as fol- lows: = c1 = O(ln(ǫ)). c2 - G(·, ξ) has logarithmic singularities at a+ - G(x,·) is of logarithmic singularity at a3, - away from these singularities G(x, ξ) is continuous in x and ξ. 2 , a3 and a− 4 , Thus, G is in L2((a2, a4)2). Hence, TG : L2(a2, a4) → L2(a2, a4) is a compact Fred- holm integral operator. From this we conclude: Proposition 3.2. The operator LS has only a discrete spectrum, and the asso- ciated eigenfunctions are complete in L2(a2, a4). Proof. By Theorem VIII.3 in [19], the self-adjointness of LS implies that for the operator (LS − i) : D(LS) → L2(a2, a4) we have Ker(LS − i) = {0}, Ran(LS − i) = L2(a2, a4). (3.12) (3.13) Consequently, (LS − i)−1 : L2(a2, a4) → D(LS) is one-to-one and onto. Moreover, it is a normal compact operator and thus we get the spectral representation (LS − i)−1f = ∞ Xn=0 λnhf, fnifn, (3.14) where {fn}n∈N is a complete orthonormal system in L2(a2, a4). This can be trans- formed into the spectral representation for LS: LSf = ∞ ( Xn=0 1 λn + i)hf, fnifn. (3.15) Clearly, the eigenfunctions fn of LS can be chosen to be real-valued. The com- pleteness of {fn}n∈N is essential for finding the SVD of HT . Another property that will be needed for the SVD is that the spectrum of LS is simple, i.e. that each eigen- value has multiplicity 1. 13 Proposition 3.3. The spectrum of LS is simple. Proof. From the compactness of (LS − i)−1, we know that each eigenvalue has finite multiplicity. Suppose f1 and f2 are linearly independent eigenfunctions of LS corresponding to the same eigenvalue λ ∈ R. Then, on all of (a2, a3) ∪ (a3, a4) the following holds f1Lf2 − f2Lf1 = 0. (3.16) Consequently, 0 = f1Lf2 − f2Lf1 = f1(P f ′ = [f1, f2]′. 2)′ − f2(P f ′ 1)′ Thus, [f1, f2] is constant on both (a2, a3) and (a3, a4). From the boundary conditions that f1 and f2 satisfy, we find that [f1, f2](a+ 4 ), which implies [f1, f2] = 0 on (a2, a3) ∪ (a3, a4). Since [f1, f2] = P (f ′ 2 ) = 0 = [f1, f2](a− 2), we get that 1f2 − f1f ′ 2 = 0 on (a2, a3) ∪ (a3, a4). f ′ 1f2 − f1f ′ (3.17) The functions f1 and f2 satisfy the transmission conditions at a3. Consequently, they can be written as f1(x) = f11(x) + f12(x) ln x − a3, f2(x) = f21(x) + f22(x) ln x − a3, in a neighborhood of a3, where fij are continuous. Since the one-sided derivatives f ′ ij are bounded at a3, equation (3.17) implies (cid:0)f12f21 − f11f22(cid:1)(x) x − a3 + O(ln2 x − a3) = 0. (3.18) Note that the terms containing lnx − a3/(x − a3) cancel. Taking the limit x → a3 in (3.18), we obtain f12(a3)f21(a3) − f11(a3)f22(a3) = 0. Thus, for some constant C: (cid:18) f11(a3) f12(a3) (cid:19) = C(cid:18) f21(a3) f22(a3) (cid:19) . If we take f1 on (a2, a3), then f11(a3) and f12(a3) define a singular initial value problem on (a3, a4) that is uniquely solvable (Theorem 8.4.1 in [26]). Thus, f1 = C·f2 on (a3, a4). Now, on the other hand, by considering f1 on (a3, a4), the values f11(a3) and f12(a3) define a singular initial value problem on (a2, a3) which has a unique solution. Hence, f1 = C · f2 on (a2, a3) ∪ (a3, a4) in contradiction to our assumption. 4. Singular value decomposition of HT . Having introduced the differential operator LS, we now want to relate it to the truncated Hilbert transform HT . The main result of this section is that the eigenfunctions of LS fully determine the two 14 families of singular functions of HT . We start by stating the following Proposition 4.1. On the set of eigenfunctions {fn}n∈N of LS, the following commutation relation holds: (HT L(y, dy)fn)(x) = L(x, dx)(HT fn)(x) for x ∈ (a1, a2) ∪ (a2, a3). (4.1) Sketch of proof. This proof follows the same general idea as the proof of Proposition 2.1 in [9]. We therefore provide full details only for those steps where additional care needs to be taken because of the singularity at a3. The steps that are completely analogous to those in the proof of Proposition 2.1 in [9] are only sketched here. Let ψ ∈ {fn}n∈N. The boundedness of ψ at a+ 4 implies that P ψ′ → 0 and P ψ → 0 there. Moreover, the transmission conditions at a3 guarantee that P ψ′ is continuous at a3. With these properties, the commutation relation for x ∈ (a1, a2), i.e. where the Hilbert kernel is not singular, can be shown similarly to the proof of Proposition 2.1 in [8]. 2 and a− Next, let x ∈ (a2, a3). The main difference from the proof of Proposition 2.1 in [9] is that now the eigenfunctions are not in C∞([a2, a4]), but are singular at a3. However, the fact that we exclude the point x = a3 allows us to always have a neighborhood of x away from a3 on which ψ is bounded. We further note that ψ ∈ C∞([a2, a3)∪(a3, a4]). Since the Hilbert kernel is singular, we need to use principal value integration and introduce the following notation: Iǫ(x) := [a2, x − ǫ] ∪ [x + ǫ, a4]. Here ǫ > 0 is so small that (x − ǫ, x + 2ǫ) ⊂ (a2, a3), i.e. the ǫ-neighborhood of x is well separated from a3. Then, π(HT L(y, dy)ψ)(x) = lim ǫ→0+ZIǫ(x)h (P (y)ψ′(y))′ y − x + 2(y − σ)2ψ(y) y − x idy. For the first term under the integral, we integrate by parts twice and plug in the boundary conditions. Again, we use that P ψ′ → 0 and P ψ → 0 at a+ 2 and a− 4 : Z Iǫ(x) (P (y)ψ′(y))′ y − x dy = − (P ψ′)(x − ǫ) + (P ψ′)(x + ǫ) ǫ + (P ψ)(x − ǫ) − (P ψ)(x + ǫ) ǫ2 + Z Iǫ(x) ψ(y) 2P (y) − P ′(y)(y − x) (y − x)3 dy. (4.2) The integral on the right-hand side of (4.2) can be related to the derivatives of In [9] similar relations (cf. eq. R ψ(y)/(y − x)dy. (2.7)) were obtained from the Leibniz integral rule, using explicitly that the integrand was continuous. In our case, the function ψ is no longer continuous because of the singularity at a3. We can generalize the argument of [9] by invoking the dominated convergence theorem and 15 rewrite the last term in (4.2) as follows: 2P (y) − P ′(y)(y − x) dy = dy + ψ′(x − ǫ) + ψ′(x + ǫ) ǫ dy + ψ(x − ǫ) + ψ(x + ǫ) ǫ ψ(x − ǫ) − ψ(x + ǫ) ǫ2 i − i (y − x)3 ψ(y) y − x ψ(y) y − x Iǫ(x) Iǫ(x) ψ(y) Z = P (x)h d2 dx2 Z + P ′(x)h d − Z 2ψ(y) Iǫ(x) dx Z Iǫ(x) (y − σ)2 − (x − σ)2 dy y − x Putting all pieces together, we obtain: π(HT L(y, dy)ψ)(x) = = lim ǫ→0+n − ǫ (P ψ′)(x − ǫ) + (P ψ′)(x + ǫ) ψ′(x − ǫ) + ψ′(x + ǫ) ψ(x − ǫ) + ψ(x + ǫ) + P (x)(cid:2) + P ′(x) ǫ ǫ + (P ψ)(x − ǫ) − (P ψ)(x + ǫ) ǫ2 ǫ2 − + L(x, dx) Z ψ(x − ǫ) − ψ(x + ǫ) (cid:3) dyo ψ(y) y − x Iǫ(x) The eigenfunction ψ is in C∞[a2, x + 2ǫ). Following [9], we can thus express the boundary terms in the above equation by Taylor expansions around x and make use of the fact that the boundary terms consist only of odd functions in ǫ. The boundary terms are then of the order O(ǫ). We thus have π(HT L(y, dy)ψ)(x) = lim ǫ→0+ L(x, dx)ZIǫ(x) ψ(y) y − x dy. (4.3) Since for ǫ > 0 sufficiently small, ψ ∈ C∞([x− ǫ, x + ǫ]), one can interchange the limit with L(x, dx) as in [9]. Because the spectrum of LS is purely discrete, we have thus found an orthonormal basis (the eigenfunctions of LS) {fn}n∈N of L2(a2, a4) for which (4.1) holds. Let us define gn := HT fn/kHT fnkL2(a1,a3). Then, in order to obtain the SVD for HT (with singular functions fn and gn), it is sufficient to prove that the gn's form an orthonormal system of L2(a1, a3) (they will then consequently form an orthonormal basis of L2(a1, a3), see Proposition 5.2). The orthogonality of the gn's will follow from the commutation relation. Since fn is an eigenfunction of LS for some eigenvalue λn, we obtain L(x, dx)gn(x) = λngn(x), x ∈ (a1, a2) ∪ (a2, a3). Similarly to LS, we define a new self-adjoint operator that acts on functions supported on [a1, a3]: 16 Definition 4.2. Let Dmax := D1,max + D2,max and Lmin := L1,min + L2,min. The operator LS : D( LS) → L2(a1, a3) is defined as the self-adjoint extension of Lmin, where D( LS) := {ψ ∈ Dmax : [ψ, u](a+ [ψ, u](a− 1 ) = [ψ, u](a− 2 ) = [ψ, u](a+ 3 ) = 0, 2 ), [ψ, v](a− 2 ) = [ψ, v](a+ 2 )} (4.4) with the maximal domain functions u, v ∈ Dmax as in (2.23), (2.24). The intuition then is the following. The function fn is bounded at a+ 2 and log- arithmic at a3, where it satisfies the transmission conditions. Consequently, as will be shown below, gn is bounded at a− 3 , logarithmic at a2 and satisfies the correspond- ing transmission conditions at a2. Clearly, it is also bounded at a+ 1 . Thus, gn is an eigenfunction of the self-adjoint operator LS. As a consequence, the gn's form an orthonormal system. Proposition 4.3. If LSfn = λnfn, then gn := HT fn/kHT fnkL2(a1,a3) is an eigenfunction of LS corresponding to the same eigenvalue LSgn = λngn. (4.5) Proof. First of all, the commutation relation for fn yields L(x, dx)(HT fn)(x) = (HT L(y, dy)fn)(x), L(x, dx)gn(x) = λngn(x), x ∈ (a1, a2) ∪ (a2, a3). ditions. Therefore, we consider p.v.R a4 What remains to be shown is that gn satisfies the boundary and transmission con- fn(y)/(y − x)dy for x close to a1, a2 and a3. In a neighborhood of a1 away from [a2, a4] this function is clearly analytic. Next, let x be confined to a small neighborhood of a2. Since the discontinuity of fn is away from a2, we can split the above integral into two -- one that integrates over a right neighborhood of a2 and another one that is an analytic function. The first item in Lemma 1.2 then implies that a2 p.v. a4 Z a2 fn(y) y − x dy = gn,1(x) − fn(a+ 2 ) lnx − a2 (4.6) where gn,1(x) is continuous in a neighborhood of x = a2. Thus, gn satisfies the transmission conditions (2.26), (2.28). It remains to check the behavior of gn close to a− 3 . We first express fn as fn(y) = fn,1(y) + fn,2(y) lny − a3, where both fn,1 and fn,2 are Lipschitz continuous. Then, in view of Lemma 1.2, both summands on the right-hand side of the equation p.v. a4 Z a2 fn(y) y − x dy =p.v. a4 Z a2 fn,1(y) y − x dy + p.v. 17 a4 Z a2 fn,2(y) lny − a3 y − x dy remain bounded as x tends to a3. Since the spectrum of LS is simple, we can conclude that the gn's form an or- thonormal system and thus the following holds: Theorem 4.4. The eigenfunctions fn of LS, together with gn := HT fn/kHT fnkL2(a1,a3) and σn := kHT fnkL2(a1,a3) form the singular value decomposition for HT : HT fn = σngn, H ∗ T gn = σnfn. (4.7) (4.8) 5. Accumulation points of the singular values of HT . The main result of this section is that 0 and 1 are accumulation points of the singular values of HT . To find this, we first analyze the nullspace and range of HT , which will also prove the ill-posedness of the inversion of HT . First, we need to state the following Lemma 5.1. If the Hilbert transform of a compactly supported f ∈ L2(a, b) vanishes on an open interval (c, d) disjoint from the object support, then f = 0 on all of R. Sketch of proof. A similar statement (and proof) can be found in [1]. The main difference is that here we consider a more general class of functions f . By dominated convergence, f ∈ L1(a, b) implies that for any z ∈ Ω = C\((−∞, a) ∪ (b,∞)), the function g(z) = R b a f (x)/(x − z)dx is differentiable in a neighborhood of z. Thus, g is analytic on Ω. The statement then follows in the same way as Lemma 2.1 in [1]. With this property of the Hilbert transform, we can obtain results on the nullspace and the range of HT : Proposition 5.2. The operator HT : L2(a2, a4) → L2(a1, a3) has a trivial nullspace and dense range that is not all of L2(a1, a3), i.e. Ker(HT ) = {0}, Ran(HT ) 6= L2(a1, a3), Ran(HT ) = L2(a1, a3). (5.1) (5.2) (5.3) Proof of (5.1). Suppose HT f = 0. Then Hχ[a2,a4]f = 0 on (a1, a2), and by Lemma 5.1, f = 0 on all of [a2, a4]. Thus, f ∈ L2(a2, a4) can always be uniquely determined from HT f . Proof of (5.2). Take any g ∈ L2(a1, a3) that vanishes on (a1, a2) and such that kgkL2(a1,a3) 6= 0. Suppose g ∈ Ran(HT ). By Lemma 5.1, if f ∈ L2(a2, a4) and HT f = g, then f is zero on [a2, a4]. This implies that g = 0 on (a1, a3), which contradicts the assumption kgk 6= 0. Proof of (5.3). The operator H ∗ same general properties. By the above argument, Ker(H ∗ Thus, Ran(HT )⊥ = {0}. T is also a truncated Hilbert transform with the T ) = {0}. 18 Equation (5.2) shows the ill-posedness of the problem. It is not true that for every g ∈ L2(a1, a3) there is a solution f to the equation HT f = g. Since Ran(HT ) is dense, the solution need not depend continuously on the data. Thus, our problem violates two properties of Hadamard's well-posedness criteria [3]. These are the existence of solutions for all data and the continuous dependence of the solution on the data. We now turn to the spectrum of H ∗ T HT . In what follows, k·k denotes the norm associated with L2(R), and h·,·i denotes the L2(R) inner product. We begin with proving the following Lemma 5.3. The operator H ∗ T HT has norm equal to 1. Proof. From kHk = 1, we know that kH ∗ T HTk = sup kH ∗ kψk=1kH ∗ T HT ψk, T HTk ≤ 1. Since finding a sequence ψn with kψnk = 1 and kH ∗ T HT ψnk → 1 would prove the assertion. Take a compactly supported function ψ ∈ L2([−1, 1]) with kψk = 1 and two vanishing moments, R 1 −1 x· ψ(x)dx. From this, we define a family of functions, such that the norm is preserved but the supports decrease. More precisely, for a > 2/(a3 − a2), we set −1 ψ(x)dx = 0 = R 1 ψa(x) = √aψ(a(x − a2 + a3 2 )). (5.4) These functions satisfy kψak = 1 and supp ψa = [ a2+a3 For their Hilbert transforms we obtain 2 − 1 a , a2+a3 2 + 1 a ] ⊂ [a2, a3]. (Hψa)(x) = √a(Hψ)(a(x − a2 + a3 2 )). We can write H ∗ T HT ψa = −χ[a2,a4]Hχ[a1,a3]Hχ[a2,a4]ψa = −χ[a2,a4]H(I − (I − χ[a1,a3]))Hχ[a2,a4]ψa = ψa + χ[a2,a4]H(I − χ[a1,a3])Hχ[a2,a4]ψa; (I − H ∗ T HT )ψa = −χ[a2,a4]H(I − χ[a1,a3])Hχ[a2,a4]ψa. Consider the L2-norm of the last expression (5.5) (5.6) k(I − H ∗ T HT )ψak2 = kχ[a2,a4]H(I − χ[a1,a3])Hχ[a2,a4]ψak2 ≤ k(I − χ[a1,a3])Hχ[a2,a4]ψak2 = Z(−∞,a1)∪(a3,∞) (Hψa)(x)2dx = aZ(−∞,a1)∪(a3,∞) (Hψ)(a(x − = Z a·(a1− a2 +a3 Hψ2dy +Z ∞ ) 2 −∞ a2 + a3 2 )2dx a·(a3− a2 +a3 2 ) Hψ2dy. (5.7) Because of the ordering of the ai's, we have that a1 − a2+a3 2 > 0. Since ψ has two vanishing moments, Hψ asymptotically behaves like 1/y3 and hence, 2 < 0 and a3 − a2+a3 19 both integrals in (5.7) are of the order O(a−5). Thus, given any ǫ > 0, one can find a > 2/(a3 − a2) such that k(I − H ∗ T HT )ψak < ǫ. Consequently, Therefore kH ∗ T HT ψak ≥ kψak − k(I − H ∗ T HT )ψak > 1 − ǫ. which implies that kH ∗ kH ∗ T HTk = 1. T HT ψak → 1 as a → ∞, We are now in a position to prove Theorem 5.4. The values 0 and 1 are accumulation points of the singular values of HT . T HT ). Proof. First of all, 0 and 1 are both elements of the spectrum σ(H ∗ T HT ). For T ) 6= L2(a2, a4). Moreover, since T HT is self-adjoint, the spectral radius is equal to 1. Thus, T HT ) ⊂ Ran(H ∗ T HTk = 1 and H ∗ The second step is to show that 0 and 1 are not eigenvalues of H ∗ the value 0, this follows from Ran(H ∗ kH ∗ 1 ∈ σ(H ∗ 0 is not an eigenvalue: T HT fi = 0. Since Ker(HT ) = 0, this implies f = 0. Thus, Ker(H ∗ 1 is not an eigenvalue: Suppose there exists a non-vanishing function f ∈ L2(a2, a4), such that T HT f = 0, then kHT fk2 = hf, H ∗ T HT ) = {0}. If H ∗ T HT . Then, −χ[a2,a4]Hχ[a1,a3]Hχ[a2,a4]f = f. kHχ[a2,a4]fk2 = kfk2 = −hχ[a2,a4]Hχ[a1,a3]Hχ[a2,a4]f, fi = hχ[a1,a3]Hχ[a2,a4]f, Hχ[a2,a4]fi = kχ[a1,a3]Hχ[a2,a4]fk2. This implies that Hχ[a2,a4]f is identically zero outside [a1, a3]. By Lemma 5.1, this implies f = χ[a2,a4]f = 0, contradicting the assumption f 6≡ 0. Therefore, 0 and 1 are accumulation points of the eigenvalues of H ∗ T HT and consequently, of the singular values of HT . Since the singular values of HT also accumulate at a point other than zero, the operator HT is not compact. 6. Numerical illustration. We want to illustrate the properties of the trun- cated Hilbert transform with overlap obtained above for a specific configuration. We choose a1 = 0, a2 = 1.5, a3 = 6 and a4 = 7.5. First, we consider two different discretizations of HT and calculate the corresponding singular values. We choose the first discretization to be a uniform sampling with 601 partition points in each of the two intervals [0, 6] and [1.5, 7.5]. Let vectors X and Y denote the partition points of [0, 6] and [1.5, 7.5] respectively. To overcome the singularity of the Hilbert kernel the vector X is shifted by half of the sample size. The i-th components of the two vectors 20 100 (i + 1 2 2 ) and Yi = 1.5 + 1 X and Y are given by Xi = 1 100 i; HT is then discretized as (HT )i,j = (1/π)(Xi − Yj), i, j = 0, . . . , 600. Figure 6.1(a) shows the singular values for the uniform discretization. We see a very sharp transition from 1 to 0. The second discretization uses orthonormal wavelets with two vanishing moments. Let φ denote the scaling function. For the discretization we define a finest scale J = −7. The scaling functions on [1.5, 7.5] are taken to be φ−7,k for integers k = 192, . . . , 957, i.e. such that supp φ−7,k ⊂ [1.5, 7.5]. On the interval [0, 6] the scaling for integers ℓ = functions are shifted in the sense that we take them to be φ−7,ℓ+ 1 0, . . . , 765, i.e. such that supp φ−7,ℓ+ 1 2 ⊂ [0, 6]. Figure 6.1(b) shows a plot of the singular values of this wavelet discretization of HT . Although the transition is not as sharp as in 6.1(a), the singular values in both cases very clearly accumulate at 0 and 1. 1.0 0.8 0.6 0.4 0.2 1.0 0.8 0.6 0.4 0.2 100 200 300 400 500 600 100 200 300 400 500 600 700 (a) Uniform discretization of size 601 × 601. (b) Wavelet discretization of size 766 × 766. Fig. 6.1: a1 = 0, a2 = 1.5, a3 = 6, a4 = 7.5. Singular values of two discretizations of HT . Next, we consider the singular functions. Figure 6.2 shows the singular functions of the uniform discretization for singular values in the transmission region between 0 and 1. Figure 6.3 illustrates the behavior of singular functions for small singular values. As anticipated, they are bounded at the two endpoints and singular at the point of truncation. Figure 6.4 gives two examples of the close to linear behavior in a log-linear plot of the singular functions. In agreement with the theory in Section 4, these plots confirm that the singularities are of logarithmic kind. Based on the numerical experiments conducted, we make the following observa- tions on the behavior of the singular functions and singular values. First, the singular functions in Figures 6.2 and 6.3 have the property that two functions with consecu- tive indices have the number of zeros differing by 1. Moreover, the zeros are located only inside one subinterval Ij . Furthermore, the plots show that singular functions with zeros within the overlap region correspond to significant singular values, whereas those which have zeros outside the overlap region correspond to small singular values. Finally, we remark that singular functions for small singular values are concentrated outside the ROI I2 = [a2, a3]. Acknowledgments. RA was supported by an FWO Ph.D. fellowship and would like to thank Prof. Ingrid Daubechies and Prof. Michel Defrise for their supervision and contribution. Also, RA would like to thank the Department of Mathematics at Duke University for hosting several research stays. AK was supported in part by NSF grants DMS-0806304 and DMS-1211164. 21 0.15 0.10 0.05 -0.05 -0.10 -0.15 0.05 1 2 3 4 5 6 7 1 2 3 4 5 6 7 -0.05 -0.10 -0.15 (a) Singular functions f448 and f449. (b) Singular functions f450 and f451. 0.15 0.10 0.05 -0.05 -0.10 -0.15 1 2 3 4 5 6 (c) Singular functions g448 and g449. (d) Singular functions g450 and g451. Fig. 6.2: Consecutive singular functions for the uniform discretization with 3, 2, 1 and no zeros within the overlap region. The corresponding singular values are σ448 = 0.999963, σ449 = 0.998782, σ450 = 0.966192, σ451 = 0.542071. 22 0.2 0.1 -0.1 -0.2 0.2 0.1 -0.1 -0.2 1 2 3 4 5 6 7 0.2 0.1 -0.1 -0.2 1 2 3 4 5 6 7 (a) Singular functions f452 and f453. (b) Singular functions f454 and f455. 1 2 3 4 5 6 0.2 0.1 -0.1 -0.2 1 2 3 4 5 6 (c) Singular functions g452 and g453. (d) Singular functions g454 and g455. Fig. 6.3: Consecutive singular functions for the uniform discretization with 1, 2, 3 and 4 zeros outside the overlap region. The corresponding singular values are σ452 = 6.29189 · 10−3, σ453 = 2.83533 · 10−5, σ454 = 1.18274 · 10−7, σ455 = 4.83357 · 10−10. 0.15 0.10 0.05 0.00 -0.05 -0.10 -0.15 -0.20 0.005 0.010 0.050 0.100 0.500 0.005 0.010 0.050 0.100 0.500 (a) (b) Fig. 6.4: Log-linear plot, i.e. with a logarithmic x scale of the singular functions g450 (left) and g453 (right) on the interval [1.5, 2.3]. 23 REFERENCES [1] M Courdurier, F Noo, M Defrise, and H Kudo. Solving the interior problem of computed tomography using a priori knowledge. Inverse problems, 24, 2008. 065001 (27pp). [2] M Defrise, F Noo, R Clackdoyle, and H Kudo. Truncated Hilbert transform and image recon- struction from limited tomographic data. Inverse Problems, 22(3):1037 -- 1053, 2006. [3] H W Engl, M Hanke, and A Neubauer. Regularization of Inverse Problems, volume 375. Springer, 1996. [4] D V Finch, 2002. Mathematisches Forschungsinstitut Oberwolfach (private conversation). [5] F D Gakhov. Boundary Value Problems. Dover Publications, 1990. [6] I M Gelfand and M I Graev. Crofton function and inversion formulas in real integral geometry. Functional Analysis and its Applications, 25:1 -- 5, 1991. [7] FA Grunbaum, L Longhi, and M Perlstadt. Differential operators commuting with finite con- volution integral operators: some nonabelian examples. SIAM Journal on Applied Math- ematics, 42(5):941 -- 955, 1982. [8] A Katsevich. Singular value decomposition for the truncated Hilbert transform. Inverse Prob- lems, 26, 2010. 115011 (12pp). [9] A Katsevich. Singular value decomposition for the truncated Hilbert transform: part II. Inverse Problems, 27, 2011. 075006 (7pp). [10] E Katsevich, A Katsevich, and G Wang. Stability of the interior problem for polynomial region of interest. Inverse Problems, 28, 2012. 065022. [11] W Koppelman and J D Pincus. Spectral representations for finite Hilbert transformations. Mathematische Zeitschrift, 71(1):399 -- 407, 1959. [12] H Kudo, M Courdurier, F Noo, and M Defrise. Tiny a priori knowledge solves the interior problem in computed tomography. Phys. Med. Biol., 53:2207 -- 2231, 2008. [13] H J Landau and H O Pollak. Prolate spheroidal wave functions, Fourier analysis and Uncer- tainty - II. Bell Syst. Tech. J, 40(1):65 -- 84, 1961. [14] H J Landau and H O Pollak. Prolate spheroidal wave functions, Fourier analysis and Uncer- tainty - III. The dimension of the space of essentially time-and band-limited signals. Bell Syst. Tech. J, 41(4):1295 -- 1336, 1962. [15] P Maass. The interior Radon transform. SIAM J. Appl. Math., 52(3):710 -- 724, June 1992. [16] F Natterer. The Mathematics of Computerized Tomography, volume 32. Society for Industrial Mathematics, 2001. [17] F Noo, R Clackdoyle, and J D Pack. A two-step Hilbert transform method for 2D image reconstruction. Physics in Medicine and Biology, 49(17):3903 -- 3923, 2004. [18] J D Pack, F Noo, and R Clackdoyle. Cone-beam reconstruction using the backprojection of locally filtered projections. IEEE Transactions on Medical Imaging, 24:1 -- 16, 2005. [19] M Reed and B Simon. Methods of Modern Mathematical Physics: Vol. 1: Functional Analysis. Academic Press, 1972. [20] D Slepian and H O Pollak. Prolate spheroidal wave functions, Fourier analysis and Uncertainty - I. Bell Syst. Tech. J, 40(1):43 -- 63, 1961. [21] G Teschl. Ordinary Differential Equations and Dynamical Systems, volume 140. American Mathematical Society, 2012. [22] F G Tricomi. Integral Equations, volume 5. Dover publications, 1985. [23] Y Ye, H Yu, Y Wei, and G Wang. A general local reconstruction approach based on a truncated Hilbert transform. International Journal of Biomedical Imaging, 2007. 63634. [24] Y Ye, S Zhao, H Yu, and G Wang. A general exact reconstruction for cone-beam CT via backprojection-filtration. IEEE Transactions on Medical Imaging, 24:1190 -- 1198, 2005. [25] Y B Ye, H Y Yu, and G Wang. Exact interior reconstruction with cone-beam CT. International Journal of Biomedical Imaging, 2007. 10693. [26] A Zettl. Sturm-Liouville Theory, volume 121 of Mathematical Surveys and Monographs. Amer- ican Mathematical Society, Providence, Rhode Island, 2005. [27] T Zhuang, S Leng, B E Nett, and G-H Chen. Fan-beam and cone-beam image reconstruction via filtering the backprojection image of differentiated projection data. Physics in Medicine and Biology, 49(24):5489 -- 5503, 2004. [28] Y Zou, X Pan, and E Y Sidky. Image reconstruction in regions-of-interest from truncated projections in a reduced fan-beam scan. Physics in Medicine and Biology, 50(1):13 -- 28, 2005. [29] Y Zou and X C Pan. Image reconstruction on PI-lines by use of filtered backprojection in helical cone-beam CT. Physics in Medicine and Biology, 49:2717 -- 2731, 2004. 24
1804.01372
1
1804
2018-04-04T12:43:31
Subsymmetric weak$^*$ Schauder bases and factorization of the identity
[ "math.FA" ]
Let $X^*$ denote a Banach space with a subsymmetric weak$^*$ Schauder basis satisfying condition~\eqref{eq:condition-c}. We show that for any operator $T : X^*\to X^*$, either $T(X^*)$ or $(I-T)(X^*)$ contains a subspace that is isomorphic to $X^*$ and complemented in $X^*$. Moreover, we prove that $\ell^p(X^*)$, $1\leq p \leq \infty$ is primary.
math.FA
math
SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY RICHARD LECHNER Abstract. Let X∗ denote a Banach space with a subsymmetric weak∗ Schauder basis satisfying condition (C). We show that for any operator T : X∗ → X∗, either T (X∗) or (IdX∗ −T )(X∗) contains a subspace that is isomorphic to X∗ and complemented in X∗. Moreover, we prove that 'p(X∗), 1 ≤ p ≤ ∞ is primary. 1. Introduction and results In [3], Casazza and Lin showed that for any bounded linear projection Q on a Banach space S with a subsymmetric basis, either Q(S) or (Id−Q)(S) contains a subspace which is isomorphic to S and complemented in S. Our first main result Theorem 1.1 extends their result to Banach spaces having a subsymmetric weak∗ Schauder basis which satisfies the subsequent condition (C). j=1, whose normalized associated coordinate functionals (e∗ Condition (C). Let U denote a Banach space with normalized unconditional basis j)∞ (ej)∞ j=1 form a uncondi- tional weak∗ Schauder basis of U∗. Given A ⊂ N, we define the bounded projection PA : U∗ → U∗ by (cid:16) ∞(cid:88) j=1 (cid:17) = ∞(cid:88) j∈A PA aje∗ j aje∗ j , (1.1) We say that U∗ together with its unconditional weak∗ Schauder basis (e∗ where the above series converge in the weak∗ topology of U∗. j)∞ j=1 satisfies condition (C), if for every infinite set Λ ⊂ N and every θ > 0, we can find a sequence (Aj)∞ j=1 of pairwise disjoint and infinite subsets of Λ, such that for all j=1 ⊂ U∗ with kx∗ j=1 ∈ '1 with j)∞ (x∗ k(aj)∞ j=1k'1 = 1 such that jkU∗ ≤ 1 there exists a sequence of scalars (aj)∞ (cid:13)(cid:13)(cid:13) ∞(cid:88) j=1 (cid:13)(cid:13)(cid:13)U∗ ≤ θ. ajPAj x∗ j (C) The above series converges in the weak∗ topology of U∗. Theorem 1.1. Let S∗ denote a Banach space with a subsymmetric weak∗ Schauder basis satisfying condition (C). Then for any given bounded operator T : S∗ → S∗, there exist operators M, N : S∗ → S∗ such that for H = T or H = IdS∗ −T, the Date: July 31, 2018. 2010 Mathematics Subject Classification. 46B25,46B26. Key words and phrases. Primary, factorization, subsymmetric, weak∗ Schauder basis. Supported by the Austrian Science Foundation (FWF) Pr.Nr. P28352. 1 2 diagram R. LECHNER S∗ IdS∗ / S∗ kMkkNk ≤ 48K7 uK4 s (1.2) N / S∗ M  S∗ H is commutative. Consequently, H(S∗) contains a subspace which is isomorphic to S∗ and complemented in S∗. n Sn We prove Theorem 1.1 in Section 4. A related concept is the notion of a primary Banach space: a Banach space X is primary, if for every bounded projection Q : X → X either Q(X) or (Id−Q)(X) is isomorphic to X (see e.g. [6]). In [2], Casazza, Kottman and Lin showed that for any Banach space S with a symmetric basis (ej)∞ j=1, the following Banach spaces are primary: c0-sum, and S is not isomorphic to '1; c0-sum, and Sn = span{e1, . . . , en}; . (cid:0)(cid:80) S(cid:1), where the direct sum is either the 'p-sum for 1 < p < ∞ or the . (cid:0)(cid:80) (cid:1), where the direct sum is either the 'p-sum for 1 < p < ∞ or the . (cid:0)(cid:80) '∞(cid:1) p, 1 ≤ p < ∞. 1 and(cid:0)(cid:80) S(cid:1) . (cid:0)(cid:80) S(cid:1) (cid:1) . (cid:0)(cid:80) 1 and(cid:0)(cid:80) These results were later complemented by Capon in [1], who showed that the Banach spaces are primary. Theorem 1.2 below adds 'p(S∗) to that list (S∗ may be non-separable). Theorem 1.2. Let S∗ denote a Banach space with a subsymmetric weak∗ Schauder basis satisfying condition (C). Then for all 1 ≤ p ≤ ∞ and any given bounded operator T : 'p(S∗) → 'p(S∗), there exist operators M, N : 'p(S∗) → 'p(S∗) such that for H = T or H = Id'p(S∗) −T, the diagram ∞, n Sn n Sn (cid:1) ∞, 'p(S∗) M 'p(S∗) Id'p(S∗) H 'p(S∗) N / 'p(S∗) kMkkNk ≤ 48K7 uK4 s , (1.3) is commutative. Consequently, 'p(S∗) is primary. The proof of Theorem 1.2 is given in Section 5. The method of proof for both our main theorems is based on the recent result [4]. 2. Notation Let X be a Banach space and let X∗ denote its dual. We say a sequence (x∗ in X∗ converges weak∗ to x∗ ∈ X∗, if n)∞ n=1 n x ∈ X. hx∗ n, xi = hx∗, xi, lim In this case we write w*lim x∗ n = x∗. j=1 ⊂ X∗ is a weak∗ Schauder basis for X∗, if there exists a j)∞ We say that (e∗ j=1 for X such that the associated coordinate functionals of (ej)j = 1∞ basis (ej)∞ j , eji = 1 and he∗ j=1, i.e. he∗ j , eii = 1, i, j ∈ N, i 6= j. Thus, n(cid:88) are (e∗ j)∞ hx∗, ejie∗ j , x∗ ∈ X∗. x∗ = w*lim n→∞ j=1 /  / O O / /   / O O SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY 3 For more details we refer to [8, 6]. From now on, whenever we encounter a series in the Banach space . X, U or S, the mode of convergence is in the norm topology of X, U or S; . X∗, U∗ or S∗, the mode of convergence is in the weak∗ topology of X∗, U∗ We say the weak∗ Schauder basis (e∗ constant C ≥ 1 such that for all sequences of scalars (aj)∞ j=1 of X∗ is unconditional, if there exists a j)∞ j ∈ X∗ j=1 aje∗ j=1 with(cid:80)∞ or S∗. holds that(cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13)X∗ ≤ C sup γj(cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13)X∗ , γjaje∗ j=1 ∈ '∞. We denote the infimum over all such constants C ≥ 1 by Ku. Note that (γj)∞ aje∗ j∈N j=1 j=1 j j (2.1) (ej)∞ j=1 is unconditional if and only if j)∞ j=1 of X∗ is subsymmetric, if the fol- j)∞ (e∗ j=1 is unconditional. We say that a weak∗ Schauder basis (e∗ lowing two conditions are satisfied: (a) (e∗ (b) Let (nj)∞ j)∞ j=1 is unconditional; sure of span{e∗ j=1 ⊂ N be an increasing sequence, and let Y(nj)∞ }∞ j=1. Then the operator S(nj)∞ nj (cid:16) ∞(cid:88) j=1 (cid:17) = j=1 : X∗ → Y(nj)∞ ∞(cid:88) aje∗ , nj aje∗ j j=1 j=1 denote the weak∗ clo- j=1 given by (2.2) S(nj)∞ j=1 is isomorphic. Note that (ej)∞ We define Ks by j=1 is subsymmetric Ks = sup(cid:8)kS(nj)∞ (cid:110)(yn)∞ j=1k,kS−1 (nj)∞ j=1 k : (nj)∞ if and only if (e∗ j)∞ j=1 is subsymmetric. j=1 ⊂ N is increasing(cid:9). < ∞(cid:111) (cid:17)1/p kynkX . n=1k'p(X) =(cid:16) ∞(cid:88) n=1 The constant Ks is finite (see [8, Chapter II, Theorem 21.2]). For all 1 ≤ p ≤ ∞ and for any Banach space X, we define as usual 'p(X) by n=1 : yn ∈ X and k(yn)∞ We conclude this section with two remarks on condition (C). Remark 2.1. Any Banach space U∗ with an unconditional weak∗ Schauder basis (e∗ j) in which '1 does not embed, satisfies condition (C). Remark 2.2. '∞ with the standard unit vector weak∗ Schauder basis (ej)∞ condition (C). j=1 satisfies First, we will show that Remark 2.1 is indeed correct. To this end, assume that condition (C) does not hold. Then there exists an infinite set Λ ⊂ N and a θ > 0, such that for any sequence (Aj)∞ j=1 of pairwise disjoint and infinite subsets of Λ, j=1 ∈ '1 j=1 ⊂ U∗ with kx∗ we can find (x∗ j)∞ with k(aj)∞ jkU∗ ≤ 1 such that for all scalars (aj)∞ j=1k'1 = 1 we have(cid:13)(cid:13)(cid:13) ∞(cid:88) j=1 (cid:13)(cid:13)(cid:13)U∗ > θ. ajPAj x∗ j 4 R. LECHNER If we put y∗ ality j = PAj x∗ j and (cid:101)y∗ j = y∗ (cid:13)(cid:13)(cid:13) ∞(cid:88) j /ky∗ jkU∗, j ∈ N, then we obtain by uncondition- aj(cid:101)y∗ (cid:13)(cid:13)(cid:13)U∗ > θ/Ku, j j=1 j )∞ j=1 is equivalent to the standard unit vector basis of '1. To see that Remark 2.2 is valid, let Λ ⊂ N denote an infinite set, and let θ > 0. Pick any sequence (Aj)∞ j=1 denote any sequence of vectors in '∞ with kxjk'∞ ≤ 1, j ∈ N. Choose N ∈ N such that N ≤ θ, and define aj = 1 1 j=1 of pairwise disjoint subsets of Λ, and let (xj)∞ N if j ≤ N, and aj = 0 whenever j > N. Then clearly thus, ((cid:101)y∗ (cid:13)(cid:13)(cid:13) ∞(cid:88) j=1 (cid:13)(cid:13)(cid:13)'∞ = 1 N ≤ θ. ajPAj xj 3. Two subspace annihilation lemmata We will present two lemmata, which are used in the proof of our first main result Theorem 1.1 to almost diagonalize an operator T : X∗ → X∗. Although we use Lemma 3.1 only for m = 1 here, we give the proof for general m ∈ N here, for later reference. Lemma 3.1. Let X denote a Banach space with a normalized basis (ej)∞ that the associated coordinate functionals (e∗ X∗. Let Λ0 ⊂ N denote an infinite set, m, n ∈ N, x∗ Then we can find a finite set F ⊂ Λ0 with cardinality F = 2m such that j=1, such j)∞ j=1 form a weak∗ Schauder basis of n ∈ X∗ and η > 0. 1, . . . , x∗ where the set of signs E = E(F) is given by max 1≤j≤n (cid:12)(cid:12)(cid:12)(cid:68)(cid:88) (cid:69)(cid:12)(cid:12)(cid:12) ≤ η, E =(cid:110)(εk)k∈F : εk ∈ {±1}, (cid:88) εkek, x∗ k∈F j ε ∈ E, εk = 0(cid:111) . k∈F (cid:111) Proof. Firstly, we will prove that given x∗ ∈ X∗, there exists an infinite set Λ ⊂ Λ0 such that hek, x∗i − hek0, x∗i ≤ η m , k, k0 ∈ Λ. (3.1) M' =(cid:110) To this end, we define the sets j ∈ Λ0 : η 2m Since hej, x∗i ≤ kx∗k < ∞, there are only finitely many ' ∈ Z for which M' is '∈Z Ml = Λ0, at least one of the sets M', ' ∈ Z has to be infinite. Clearly, this infinite set satisfies (3.1). Now we will just repeatedly use (3.1). We begin by selecting Λ1 ⊂ Λ0 so that non-empty. Hence, since(cid:83) (' − 1) ≤ hej, x∗i < ' ∈ Z. η 2m ' , hek, x∗ 1i − hek0, x∗ 1i ≤ η m , k, k0 ∈ Λ1. Let 1 ≤ i ≤ n− 1 and assume we have already chosen infinite sets Λ0 ⊃ Λ1 ⊃ Λ2 ⊃ ··· ⊃ Λi−1 such that max k, k0 ∈ Λi−1. ji − hek0, x∗ hek, x∗ , ji ≤ η m 1≤j≤i−1 By replacing Λ0 with the infinite set Λi−1 in (3.1), we can find an infinite set Λi ⊂ Λi−1 such that max 1≤j≤i hek, x∗ ji − hek0, x∗ ji ≤ η m , k, k0 ∈ Λi. SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY 5 Stopping the induction after n steps, we obtain ji ≤ η m ji − hek0, x∗ (3.2) We choose any F ⊂ Λn with F = 2m and let ε = (εk)k∈F ∈ E = E(F), k∈F εk = 0. Then since F ⊂ Λn, we obtain from (3.2) k, k0 ∈ Λn. hek, x∗ max 1≤j≤n , i.e. εk ∈ {±1} is such that(cid:80) (cid:12)(cid:12)(cid:12)(cid:88) (cid:69)(cid:12)(cid:12)(cid:12) = (cid:12)(cid:12)(cid:12)(cid:68)(cid:88) εkek, x∗ that j k∈F ji(cid:12)(cid:12)(cid:12) ≤ η, hek, x∗ ji − (cid:88) k∈F εk=−1 hek, x∗ k∈F εk=1 1 ≤ j ≤ n. (cid:3) The following Lemma is an abstract version of an argument which Lindenstrauss used in [5] (see also [6]) to show that '∞ is prime (which means that every infinite dimensional complemented subspace of '∞ is isomorphic to '∞). Lemma 3.2. Let U denote a Banach space with a normalized unconditional basis j)∞ (ej)∞ j=1 form an uncondi- tional weak∗ Schauder basis of U∗. Let Λ ⊂ N denote an infinite set, η > 0 and ϕ ∈ U∗∗. If U∗ together with (e∗ j)∞ j=1 satisfies condition (C), then there exists an infinite set A ⊂ Λ such that j=1, such that the associated coordinate functionals (e∗ sup kx∗kU∗≤1 hϕ, PAx∗i ≤ η. Proof. Let η > 0, ϕ ∈ U∗∗ \ {0}, and assume the conclusion of the Lemma is false. Hence, for all infinite sets A ⊂ Λ there exists x∗ with kx∗kU∗ = 1 such that hϕ, PAx∗i > η. (3.3) j=1 with Aj ⊂ Λ, j ∈ N according to jkU∗ = 1, j ∈ U∗ with kx∗ We define θ = condition (C). By our assumption (3.3), we can find x∗ j ∈ N, such that and choose (Aj)∞ 2kϕkU∗∗ Ku η ji > η. By condition (C) we can find a sequence (aj)∞ that hϕ, PAj x∗ j=1 ∈ '1 with k(aj)∞ (3.4) j=1k'1 = 1 such η 2kϕkU∗∗ Ku . (3.5) But on the other hand, we obtain from (3.4) that (cid:13)(cid:13)(cid:13) ∞(cid:88) j=1 ajPAj x∗ j (cid:13)(cid:13)(cid:13)U∗ ≤ θ = (cid:69) n(cid:88) ajPAj x∗ j (cid:68) ϕ, j=1 > η, for a large enough n ∈ N. Combining the latter estimate with (3.5), we arrive at the contradiction (cid:68) n(cid:88) j=1 η < ϕ, ajPAj x∗ j (cid:69) ≤ kϕkU∗∗ (cid:13)(cid:13)(cid:13) n(cid:88) j=1 (cid:13)(cid:13)(cid:13)U∗ ≤ η/2. ajPAj x∗ j (cid:3) 4. Proof of Theorem 1.1 For convenience of the reader, we restate Theorem 1.1 here below. Theorem 4.1 (Main result Theorem 1.1). Let S∗ denote a Banach space with a subsymmetric weak∗ Schauder basis satisfying condition (C). Then for any given 6 bounded operator T : S∗ → S∗, there exist operators M, N : S∗ → S∗ such that for H = T or H = IdS∗ −T, the diagram R. LECHNER S∗ IdS∗ / S∗ M  S∗ N / S∗ H kMkkNk ≤ 48K7 uK4 s (4.1) is commutative. Consequently, H(S∗) contains a subspace which is isomorphic to S∗ and complemented in S∗. Proof. In this proof, we use the following constants: ηi = K−1 u 4−i−1, i ∈ N. (4.2) Step 1: Inductive construction of the block basis. Let (sj)∞ j=1 denote j)∞ the subsymmetric basis of S, and let (s∗ j=1 denote the associated coordinate functionals, which form a weak∗ Schauder basis for S∗, such that condition (C) is satisfied. We put A1 = N, choose B1 = {1, 2}, and we define 1 − s∗ 2. b1 = s1 − s2 b∗ 1 = s∗ (4.3) and By Lemma 3.2, there exists an infinite collection A2 ⊂ A1 \ {1, 2}, such that sup kx∗kS∗≤1 hT ∗b1, PA2 x∗i ≤ η1. (4.4) Now assume we have already chosen infinite sets A1 ⊃ A2 ⊃ ··· ⊃ Ai, pairwise disjoint finite sets Bj ⊂ Aj with Bj = 2, 1 ≤ j ≤ i − 1 and that we have defined (4.5) for all 1 ≤ j ≤ i − 1. By Lemma 3.1, we can find a set Bi ⊂ Ai with Bi = 2 so that if we put k0, k1 ∈ Bj, k0 < k1, bj = sk0 − sk1 k0 − s∗ k1 , and b∗ j = s∗ k0, k1 ∈ Bi, k0 < k1, bi = sk0 − sk1 we have the estimate such that kx∗kS∗≤1 This completes the inductive step. The estimates (4.4), (4.7) and (4.9) imply i−1(cid:88) ∞(cid:88) j=1 bi, T j=i+1 (cid:12)(cid:12)(cid:12)(cid:68) hbi, T b∗ ji ≤ ηi, (cid:69)(cid:12)(cid:12)(cid:12) ≤ ηi (cid:13)(cid:13)(cid:13) ∞(cid:88) j=i+1 ajb∗ j (cid:13)(cid:13)(cid:13)S∗ , ajb∗ j i ∈ N, i ∈ N, (4.10a) (4.10b) By Lemma 3.2, there exists an infinite set Ai+1 with j=1 and b∗ hbi, T b∗ k0 − s∗ k1 , ji ≤ ηi. i = s∗ i−1(cid:88) Ai+1 ⊂ Ai \(cid:110) k ∈ N : k ≤ max(cid:0) i(cid:91) (cid:12)(cid:12)(cid:12)(cid:68) T ∗bi, PAi+1 x∗(cid:69)(cid:12)(cid:12)(cid:12) ≤ ηi. sup j=1 (cid:1)(cid:111) Bi (4.6) (4.7) (4.8) (4.9) /  / O O j=1 ⊂ R is such that(cid:80)∞ SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY 7 whenever (aj)∞ Step 2: Basic operators. We define the operators B, Q : S∗ → S∗ by j=i+1 ajb∗ j ∈ S∗. ∞(cid:88) ∞(cid:88) j=1 j=1 Bx∗ = Qx∗ = hx∗, sjib∗ j , 1 jkS∗ kb∗ jkS∗ kb∗ Bj hbj, x∗is∗ j , x∗ ∈ S∗, x∗ ∈ S∗. (4.11a) (4.11b) We will now show that B and Q are bounded linear operators. To this end, let (cid:13)(cid:13)(cid:13)S∗ , k1(j) jkS∗ ≥ K−1 (cid:17) u , (cid:13)(cid:13)(cid:13)S∗ 1 jkS∗ kb∗ ajs∗ j=1 ajs∗ k1(j) where {k0(j), k1(j)} = Bj, j ∈ N. By unconditionality, we obtain kb∗ and x∗ =(cid:80)∞ (cid:13)(cid:13)(cid:13)B (cid:16) ∞(cid:88) (cid:13)(cid:13)(cid:13)B j=1 j=1 j=1 j=1 j=1 j u k0(j) k0(j) ajs∗ j ajs∗ ajs∗ ajs∗ 1 jkS∗ kb∗ j. Firstly, note that j=1 ajs∗ (cid:16) ∞(cid:88) (cid:17)(cid:13)(cid:13)(cid:13)S∗ ≤(cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:17)(cid:13)(cid:13)(cid:13)S∗ ≤ K2 (cid:13)(cid:13)(cid:13)B (cid:16) ∞(cid:88) kQx∗kS∗ ≤(cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13) ∞(cid:88) kb∗ jkS∗ Bj hsk0(j), x∗is∗ (cid:13)(cid:13)(cid:13)S∗ + (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:16)(cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13)S∗ + (cid:17)(cid:13)(cid:13)(cid:13)S∗ ≤ 2K2 (cid:13)(cid:13)(cid:13)S∗ . (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13)S∗ + (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13)S∗ + Ks j)∞ j=1 is subsymmetric, hence ajs∗ ajs∗ kb∗ jkS∗ Bj hsk0(j), x∗is∗ uKs k0(j) j j=1 j j=1 j=1 j j=1 obtain Since (s∗ nated by Ks Since kb∗ j=1 j kS∗ Bj ≤ 1, unconditionality yields kQx∗kS∗ ≤ 2KsKukx∗kS∗ . One can easily verify that QB = IdS∗, i.e. the diagram S∗ B IdS∗ S∗ S∗ Q kb∗ jkS∗ Bj hsk1(j), x∗is∗ j kb∗ jkS∗ Bj hsk1(j), x∗is∗ k1(j) j=1 (4.12) (cid:13)(cid:13)(cid:13)S∗ . (cid:13)(cid:13)(cid:13)S∗ . (4.13) (4.14) Furthermore, by (4.8) we have that sk0(j) < sk0(j+1) and sk1(j) < sk1(j+1), j ∈ N. The weak∗ Schauder basis (s∗ Secondly, if we write bj = sk0(j) − sk1(j) for k0, k1 ∈ Bj, k0 < k1, j ∈ N, we j)∞ j=1 is subsymmetric, the right hand side of the latter inequality is domi- is commutative. Consequently, B is an isomorphism onto its range and its range is complemented by BQ. Step 3: Conclusion of the proof. Observe that at least one of the two following sets is infinite: {j ∈ N : hbj, T b∗ {j ∈ N : hbj, (IdS∗ −T)b∗ ji ≥ 1}. ji ≥ 1} or / / ! ! = = 8 R. LECHNER If the left set is infinite we denote it by J and put H = T, and if the left set is finite, we denote the right set by J and we define H = IdS∗ −T. In either case, we obtain J is infinite and hbj, Hb∗ ji ≥ 1, j ∈ J. (4.15) Let Y = B(S∗) and note that (4.12), (4.13) and (4.14) yields S∗ IdS∗ B  Y IdS∗ Now, define P : S∗ → Y by P x∗ =(cid:88) j∈J S∗ QY / Y kBkkQY k ≤ 4K3 uK2 s . (4.16) hbj, x∗i hbj, Hb∗ ji b∗ j , x∗ ∈ S∗, (4.17) and observe that by (4.15), the unconditionality of (s∗ ditionality of (b∗ j)∞ j=1) and the definition of B and Q (see (4.11)), we obtain j)∞ j=1 (and thus, the uncon- Since hbj, yi ≤ kbjkSkykS∗, we obtain aj ≤ kykS∗. Hence, by (4.15) and the crucial off diagonal estimate (4.10a), we obtain aj i∈J identity is true: kP x∗kS∗ ≤ 2KukBQx∗kS∗ , x∗ ∈ S∗. Combining the latter estimate with (4.12) and (4.13) yields x∗ ∈ S∗. skx∗kS∗ , uK2 j∈J:j<i (cid:88) kP x∗kS∗ ≤ 8K4 A straightforward calculation shows that for all y =(cid:80) (cid:10)bi, H(cid:80) P Hy − y =(cid:88) (cid:13)(cid:13)(cid:13)(cid:13)(cid:88) (cid:10)bi, H(cid:80) i +(cid:88) ji hbi, Hb∗ i i b∗ hbi, Hb∗ (cid:13)(cid:13)(cid:13)(cid:13)S∗ hbi, Hb∗ ji i i b∗ hbi, Hb∗ (cid:13)(cid:13)(cid:13) (cid:88) ≤ 2(cid:88) ≤ 2KukykS∗(cid:88) j∈J:j>i ajb∗ (cid:13)(cid:13)(cid:13)(cid:13)S∗ hbi, Hb∗ i i (cid:88) tionality(cid:13)(cid:13)(cid:13)(cid:13)(cid:88) ≤ 2kykS∗(cid:88) i∈J j∈J:j>i i∈J j∈J:j<i j b∗ i (cid:11) i∈N i∈J aj i i∈J ηi ηi. i∈N (4.18) j ∈ Y , the following j∈J ajb∗ j∈J:j>i ajb∗ j hbi, Hb∗ i i b∗ i . (4.19) (cid:11) ηi. (4.20) Using (4.15) and the second off diagonal estimate (4.10b), we obtain by uncondi- (cid:13)(cid:13)(cid:13)S∗ ajb∗ j (4.21) u 4−i−1, i ∈ N. Thus, inserting the esti- Recall that in (4.2) we put ηi = K−1 mates (4.20) and (4.21) into (4.19) yields kP Hy − ykS∗ ≤ 1 (4.22) Let J : Y → S∗ denote the operator given by Jy = y, then, by (4.22), P HJ : 2. Thus, if we define V = (P HJ)−1P, the Y → Y is invertible and k(P HJ)−1k ≤ 3 3kykS∗ , y ∈ Y. / /  / O O SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY 9 diagram Y IdY (P HJ)−1 Y P HJ J S∗ Y H V P / S∗ kJkkV k ≤ 12K4 uK2 s (4.23) is commutative. Merging the diagrams (4.16) and (4.23) concludes the proof: S∗ B Y S∗ M J P HJ IdS∗ IdY Y H S∗ QY Y N (P HJ)−1 P V / S∗ kMkkNk ≤ 48K7 uK4 s . (4.24) (cid:3) 5. Direct sums of Banach spaces with a weak∗ subsymmetric basis We begin this section by introducing notation specific to the two parameter case. Then we extend the subspace annihilation lemmata in Section 3 to the two param- eter case and use them to prove our second main result Theorem 1.2. 5.1. Notation. Let X denote a Banach space with normalized basis (xj)∞ j=1, such that their normalized associated coordinate functionals (x∗ basis of X∗. For each i ∈ N, let qi : 'p(X∗) → X∗ denote the canonical norm 1 coordinate projection given by j)∞ j=1 form a weak∗ Schauder (cid:0)(yn)∞ n=1 (cid:1) = yi, yi (yn)∞ n=1 ∈ 'p(X∗). Let I ⊂ N and define QI : 'p(X∗) → 'p(X∗) as the natural norm 1 projection onto the coordinates indexed by I, i.e. n=1 = yi, i ∈ I qiQI(yn)∞ n=1 ∈ 'p(X∗). For i ∈ N, we will sometimes write Qi instead of Q{i}. for all (yn)∞ For each i ∈ N, we define Ii : X∗ → 'p(X∗) as the canonical isometric embedding of X∗ into the i-th coordinate of 'p(X∗), which is given by k ∈ N, k 6= i, n=1 = 0, i /∈ I, qiQI(yn)∞ qiIix∗ = x∗ qkIix∗ = 0, i ∈ N. and and We define the sequence (eij)∞ i,j=1 ⊂ 'p(X∗) by eij = Iix∗ j , i, j ∈ N; hence, the associated coordinate functionals fij : 'p(X∗) → R are given by hfij, yi = hqiy, xji, Given any sequence of scalars (aij)∞ 'p(X∗), we define ∞(cid:88) aijeij =(cid:16) ∞(cid:88) i,j=1 j=1 aijx∗ j y ∈ 'p(X∗), i, j ∈ N. i,j=1 for which (w*limn→∞(cid:80)n (cid:17)∞ (cid:17)∞ =(cid:16)w*lim n(cid:88) aijx∗ . j n→∞ i=1 i=1 j=1 j=1 aijx∗ i=1 ∈ j)∞ (5.1) / /   > > / ` ` O O   / /  / /   ' ' O O 7 7 / g g O O Y Y 10 Thus, we have the identity ∞(cid:88) hfij, yieij =(cid:16) ∞(cid:88) i,j=1 j=1 R. LECHNER (cid:17)∞ i=1 hqiy, xjix∗ j = y, y ∈ 'p(X∗). For K ⊂ N2, we define the projection RK : 'p(X∗) → 'p(X∗) by ∞(cid:88) RKy = hfij, yi1K(i, j)eij, where 1K(i, j) = 1 if (i, j) ∈ K, and 1K(i, j) = 0 if (i, j) /∈ K. If (x∗ tional, then RK is bounded. j) is uncondi- i,j=1 5.2. Two subspace annihilation lemmata for direct sums. The following lemmata are two parameter versions of the subspace annihilation lemmata presented in Section 3. Lemma 5.1. Let X denote a Banach space with a normalized basis (xj)∞ j=1, such j)∞ j=1 form a weak∗ Schauder that their normalized associated coordinate functionals (x∗ basis of X∗. Let 1 ≤ p ≤ ∞ and let Λ0 ⊂ N denote an infinite set. Then, given m ∈ N, y1, . . . , ym ∈ 'p(X∗), n0 ∈ N, η > 0 and M ∈ N, there exists a finite set F ⊂ Λ0 with F = 2M such that where the set of signs E = E(n0, F) is given by max 1≤'≤m (cid:12)(cid:12)(cid:10)(cid:88) (cid:11)(cid:12)(cid:12) ≤ η, E =(cid:110)(εn0k)k∈F : εn0k ∈ {±1}, (cid:88) εn0kfn0k, y' k∈F k∈F (5.2) ε ∈ E, εn0k = 0(cid:111) . Proof. Since the proof of Lemma 5.1 is completely analogous to that of Lemma 3.1, (cid:3) we will omit it. Lemma 5.2. Let U denote a Banach space with a normalized unconditional basis j)∞ (uj)∞ j=1 form an unconditional weak∗ Schauder basis of U∗. Let U∗ satisfy condition (C), 1 ≤ p ≤ ∞, and let Λi ⊂ N, i ∈ N denote infinite sets. Let ϕ ∈ ('p(U∗))∗ and η > 0. Then there exist infinite sets and Ji ⊂ Λi, i ∈ N such that j=1, such that their normalized associated coordinate functionals (u∗ sup kyk'p(U∗)≤1 hϕ, R{i}×Jiyi ≤ η, i ∈ N. Proof. If we assume Lemma 5.2 is false, then there exists an i0 ∈ N such that sup kyk'p(U∗)≤1 hϕ, R{i0}×Ayi > η, (5.3) (5.4) and choose a sequence j=1 of pairwise disjoint, infinite subsets of J according to (C). By (5.4), we can for every infinite set A ⊂ Λi0. We define θ = (Aj)∞ find a sequence (yj)∞ j=1 in 'p(U∗) with kyjk ≤ 1, j ∈ N such that 2kϕk('p(U∗))∗ Ku η hϕ, R{i0}×Aj yji > η, j ∈ N. j=1 ∈ '1 with k(aj)∞ (5.5) j=1k'1 = 1 Next, define u∗ according to (C) such that j = qi0yj, j ∈ N and choose (aj)∞ (cid:13)(cid:13)(cid:13) ∞(cid:88) j=1 (cid:13)(cid:13)(cid:13)U∗ ≤ θ = ajPAj u∗ j η 2kϕk('p(U∗))∗ Ku . (5.6) SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY11 By (5.5), we obtain an integer n ∈ N such that n(cid:88) Observe that(cid:80)n η < hϕ, j=1 leads to the contradiction j=1 ajR{i0}×Aj yj = Ii0 ajR{i0}×Aj yji ≤ kϕk('p(U∗))∗ (cid:80)n (cid:13)(cid:13)(cid:13) n(cid:88) η ≤ kϕk('p(U∗))∗ ajPAj uj j=1 ajR{i0}×Aj yj j=1 (cid:13)(cid:13)(cid:13) n(cid:88) (cid:13)(cid:13)(cid:13)U∗ ≤ η/2. j=1 ajPAj uj, hence, combining (5.6) and (5.7) (cid:13)(cid:13)(cid:13)'p(U∗) . (5.7) (5.8) (cid:3) 5.3. Proof of main result Theorem 1.2. In this section we prove our second main result Theorem 1.2, which we repeat here below (see Theorem 5.3). The proof involves inductively constructing a block basis in each coordinate of 'p(S∗), 1 ≤ p ≤ ∞. The order by which we proceed is determined by the linear order ≺ on N2, which is defined as follows: Let <' denote the lexicographic order on N2 and define (i0, j0) ≺ (i1, j1) if and only if (i0 + j0, i0) <' (i1 + j1, i1), for all (i0, j0), (i1, j1) ∈ N2 (see Figure 1). Let O≺ : N2 → N denote the unique bijective function that preserves the order ≺, i.e. O≺((i0, j0)) < O≺((i1, j1)) if and only if (i0, j0) ≺ (i1, j1), for all (i0, j0), (i1, j1) ∈ N2. Figure 1. The first 10 Elements of N2 with respect to the order ≺. Theorem 5.3 (Main result Theorem 1.2). Let S∗ denote a Banach space with a subsymmetric weak∗ Schauder basis satisfying condition (C). Then for all 1 ≤ p ≤ ∞ and any given bounded operator T : 'p(S∗) → 'p(S∗), there exist operators M, N : 'p(S∗) → 'p(S∗) such that for H = T or H = Id'p(S∗) −T, the diagram 'p(S∗)Id'p(S∗)/ 'p(S∗) M 'p(S∗) N / 'p(S∗) H kMkkNk ≤ 48K7 uK4 s , (5.9) is commutative. Consequently, 'p(S∗) is primary. (1,1)(1,2)(2,1)(1,3)(2,2)(3,1)(1,4)(2,3)(3,2)(4,1)/   / O O 12 R. LECHNER Proof. For the definition of eij and fij we refer to Section 5.1. Within this proof, we identify ηk ↔ ηk0k1 , Bk ↔ Bk0k1, bk ↔ bk0k1, k ↔ b∗ b∗ k0k1, whenever O≺((k0, k1)) = k. We conclude this preliminary step by defining ηk = K−1 u 4−k−1, k ∈ N. (5.10) Step 1: Inductive construction. We will now inductively (with respect to the order ≺ on N2) construct sequences (b )∞ i,j=1. We begin by putting J1,i = N, i ∈ N, B1 = {(1, 1), (1, 2)} and define i,j=1 and (b ij )∞ (ε) ∗(ε) ij 11 = e11 − e12 (ε) b and ∗(ε) 11 = f11 − f12. b By Lemma 5.2, there exist infinite sets J2,i ⊂ N, i ∈ N with J2,i ⊂ J1,i \ {1, 2}, such that sup kyk'p(S∗)≤1 hT ∗b ∗(ε) 1 , R{i}×J2,i yi ≤ η1, i ∈ N. Let us assume that we have already . selected infinite sets Jk,i ⊂ N, i ∈ N, 1 ≤ k ≤ K and finite sets Bk ⊂ N2 with Bk = 2, 1 ≤ k ≤ K − 1 such that Bk = Bk0k1 ⊂ ({k0} × Jk,k0) \ ({k0} × Jk+1,k0), (cid:47)(cid:110) whenever 1 ≤ k ≤ K − 1 and O≺((k0, k1)) = k, and that j ∈ N : j ≤ max(cid:8)j0 ∈ N : ∃i0 ∈ N, (i0, j0) ∈ k(cid:91) Jk+1,i ⊂ Jk,i (5.14a) (5.14b) (cid:9)(cid:111) , B' '=1 for all i ∈ N and 1 ≤ k ≤ K − 1; . defined b ∗(ε) k by (ε) k , b k = ek0j − ek0j0 (ε) b (5.14c) whenever 1 ≤ k ≤ K−1, O≺((k0, k1)) = k and (k0, j), (k0, j0) ∈ Bk = Bk0k1 with j < j0; ∗(ε) k = fk0j − fk0j0, and b . obtained the estimates (5.11) (5.12) (5.13) We now come to the inductive step. To this end, choose (K0, K1) ∈ N2 with O≺((K0, K1)) = K. The set JK,K0 is infinite (see (5.14a)), hence, by Lemma 5.1 there exists a set F ⊂ JK,K0 with F = 2 such that (cid:12)(cid:12)(cid:10)fK0j − fK0j0, T b (cid:11)(cid:12)(cid:12) ≤ ηK, (ε) ' K−1(cid:88) '=1 j, j0 ∈ F. (5.15) We define the collection (5.16) and note that (with the possible exception of BK ∩ ({K0} × JK+1,K0) = ∅) condi- tion (5.14a) holds true for K + 1. We will take care of the possible exception by BK = BK0K1 = {K0} × F, k−1(cid:88) , R{i}×Jk+1,iy(cid:11)(cid:12)(cid:12) ≤ ηk, ' i ≤ ηk, (ε) ∗(ε) k hb , T b '=1 ∗(ε) k (cid:12)(cid:12)(cid:10)T ∗b sup kyk'p(S∗)≤1 for all 1 ≤ k ≤ K − 1. (5.14d) (5.14e) i ∈ N, '=1 (ε) K , b SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY13 appropriately choosing JK+1,i, i ∈ N at a later stage of the proof. Next, we define the functions b j, j0 ∈ F, j < j0, which is in accordance with (5.14c). Combining (5.15) with (5.18) yields ∗(ε) K and the signs εK0,j, j ∈ F by ∗(ε) K = fK0j − fK0j0, K = eK0j − eK0j0, (ε) b K−1(cid:88) b (5.17) (5.18) (cid:11)(cid:12)(cid:12) ≤ ηK, (cid:12)(cid:12)(cid:10)b ∗(ε) (ε) K , T b ' JK+1,i ⊂ JK,i showing that (5.14d) holds for K + 1, as required. Since all the sets Bk, 1 ≤ k ≤ K are finite, we can apply Lemma 5.2 to find infinite sets JK+1,i, i ∈ N with (cid:47)(cid:110) j ∈ N : j ≤ max(cid:8)j0 ∈ N : ∃i0 ∈ N, (i0, j0) ∈ K(cid:91) (cid:12)(cid:12)(cid:10)T ∗b K , R{i}×JK+1,i y(cid:11)(cid:12)(cid:12) ≤ ηK, i ∈ N, ∗(ε) '=1 sup kyk'p(S∗)≤1 (cid:9)(cid:111) , B' (5.19) (5.20) such that the following estimate is satisfied: Note that (5.19) and (5.20) imply (5.14b) and (5.14e) for K + 1. Moreover, (5.19) also implies BK∩({K0}×JK+1,K0) = ∅, which completes the missing part of (5.14a) for K + 1 (see (5.16)). Altogether, we completed the inductive step, and we know that (5.14) is true for K + 1. Step 2: Basic operators. We define the operators B, Q : 'p(S∗) → 'p(S∗) by ∞(cid:88) ∞(cid:88) i,j=1 i,j=1 By = Qy = 1 hfij, yi kb ij k'p(S∗) (ε) kb ij k'p(S∗) (ε) ∗(ε) Bij ij hb (ε) b ij , y ∈ 'p(S∗), , yieij, y ∈ 'p(S∗). (5.21a) (5.21b) (cid:13)(cid:13)(cid:13) ∞(cid:88) i,j=1 (cid:13)(cid:13)(cid:13)'p(S∗) For the mode of convergence of the above series, we refer to (5.1). y ∈ 'p(S∗) and observe that by (5.14c), kByk'p(S∗) is dominated by We will now show that B and Q are bounded linear operators. To this end, let hfij, yi kb 1 ij k'p(S∗) (ε) ekij 'ij hfij, yi kb 1 ij k'p(S∗) (ε) ekij mij where {(kij, 'ij), (kij, mij)} = Bij, i, j ∈ N. By unconditionality we obtain that ij k'p(S∗) ≥ K−1 kb (ε) kByk'p(S∗) ≤ K2 u , i, j ∈ N, and subsequently hfij, yiekij mij hfij, yiekij 'ij + K2 (cid:13)(cid:13)(cid:13) ∞(cid:88) u u i,j=1 Furthermore, by (5.14b) we have that 'ij < 'i(j+1) and mij < mi(j+1), i, j ∈ N. The weak∗ Schauder basis (s∗ j)∞ j=1 of S∗ is subsymmetric, hence hfij, yiekij j (cid:13)(cid:13)(cid:13) ∞(cid:88) = 2K2 uKs hfij, yieij i,j=1 i,j=1 + (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13)'p(S∗) (cid:13)(cid:13)(cid:13)'p(S∗) i,j=1 (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13) ∞(cid:88) , (cid:13)(cid:13)(cid:13)'p(S∗) (cid:13)(cid:13)(cid:13)'p(S∗) (cid:13)(cid:13)(cid:13)'p(S∗) . (5.22) kByk'p(S∗) ≤ 2K2 uKs uKskyk'p(S∗). i,j=1 = 2K2 14 R. LECHNER i,j=1 (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13) ∞(cid:88) Similarly, we obtain that kQyk'p(S∗) is dominated by + (cid:13)(cid:13)(cid:13)'p(S∗) (cid:13)(cid:13)(cid:13)'p(S∗) i,j=1 (cid:13)(cid:13)(cid:13) ∞(cid:88) (cid:13)(cid:13)(cid:13) ∞(cid:88) kb hfkij 'ij , yieij ij k'p(S∗) (ε) Bij kb ij k'p(S∗) (ε) Bij j=1 is subsymmetric, gives us the following upper bound for kQyk'p(S∗): j)∞ which, since (s∗ ij k'p(S∗) (ε) Bij ij k'p(S∗) (ε) Bij i,j=1 ≤ 1, we obtain by unconditionality ij k'p(S∗) (ε) Bij i,j=1 Using kb hfkij mij , yieimij hfkij 'ij , yiei'ij hfkij mij , yieij (cid:13)(cid:13)(cid:13)'p(S∗) +Ks kb kb Ks (cid:13)(cid:13)(cid:13)'p(S∗) , . kQyk'p(S∗) ≤ 2KsKukyk'p(S∗). One can easily verify that QB = Id'p(S∗), i.e. the diagram 'p(S∗) Id'p(S∗) 'p(S∗) B Q 'p(S∗) (5.23) (5.24) is commutative. Consequently, B is an isomorphism onto its range, and its range is complemented by BQ. Step 3: Factorization of the identity. With i ∈ N fixed, observe that at least one of the two following sets is infinite: Ki = {j ∈ N : hb Thus, clearly one of the following two sets is infinite: Li = {j ∈ N : hb , (Id'p(S∗) −T)b ij i ≥ 1}, (ε) ij i ≥ 1}. (ε) ∗(ε) ij ∗(ε) ij , T b {i ∈ N : Ki is infinite} or {i ∈ N : Li is infinite} If the left set is infinite, we denote it by I and we put Ji = Ki, i ∈ I as well as H = T; if the left set is finite, we denote the right set by I, and we define Ji = Li, i ∈ I as well as H = Id'p(S∗) −T. In either of these two cases, we obtain that (5.25) I is infinite, Let Y = B('p(S∗)) and define A : 'p(S∗) → Y , y 7→ By. Thus, by (5.22), (5.23) and (5.24), the following diagram is commutative: Ji is infinite, i ∈ I and hb ij i ≥ 1, i ∈ I, j ∈ Ji. (ε) ∗(ε) ij , Hb 'p(S∗) Id'p(S∗) A Y Id'p(S∗) 'p(S∗) QY / Y kAkkQY k ≤ 4K3 uK2 s . (5.26) Now, put K = {O≺(i, j) : i ∈ I, j ∈ Ji} and define P : 'p(S∗) → Y by P y = (cid:88) k∈K ∗(ε) hb k ∗(ε) hb k , yi k i b (ε) , Hb (ε) k , y ∈ 'p(S∗). (5.27) ∗(ε) Recall that at the beginning of the proof we identified b k with ∗(ε) k0k1, whenever O≺((k0, k1)) = k. Observe that by (5.25), unconditionality and the b definition of B and Q (see (5.21)), we obtain that (ε) k0k1 and b (ε) k with b kP yk'p(S∗) ≤ 2KukBQyk'p(S∗) / / $ $ : : / /   / O O SUBSYMMETRIC WEAK∗ SCHAUDER BASES AND FACTORIZATION OF THE IDENTITY15 Combining the latter estimate with (5.22) and (5.23) yields (5.28) uK2 skyk'p(S∗). kP yk'p(S∗) ≤ 8K4 A straightforward calculation shows that for all y = (cid:80) ,(cid:80) P Hy − y = (cid:88) following identity is true: k +(cid:88) (cid:10)H∗b (cid:88) , Hb (ε) hb hb ∗(ε) k ∗(ε) k ' i (ε) k i b (ε) a' (ε) b k . (5.29) j kSkyk'p(S∗), we obtain aj ≤ kyk'p(S∗). Hence, by (5.25) (ε) '∈K:'>k a'b , Hb ∗(ε) k k i (ε) k∈K , Hb '∈K:'<k k∈K , yi ≤ kb Since hb and the off-diagonal estimate (5.14d), we obtain (ε) j k∈K akb k ∈ Y , the (ε) (cid:11) (ε) ' ∗(ε) k hb ≤ 2kyk'p(S∗) (cid:88) k∈N ηk. (5.30) (cid:13)(cid:13)(cid:13) (cid:88) '∈K:'>k ηk (cid:13)(cid:13)(cid:13)'p(S∗) . a'b (ε) ' (5.31) k∈K (cid:13)(cid:13)(cid:13)(cid:13)(cid:88) (cid:10)H∗b (cid:13)(cid:13)(cid:13)(cid:13)(cid:88) k∈K (cid:88) '∈K:'<k hb hb ∗(ε) k ∗(ε) k a' , Hb , Hb ∗(ε) k ∗(ε) hb k '∈K:'>k a'b , Hb k i (ε) (ε) ' ,(cid:80) (cid:13)(cid:13)(cid:13) (cid:88) ,(cid:80) b (ε) k (ε) k ' i (ε) k i b (ε) (cid:11) (cid:13)(cid:13)(cid:13)(cid:13)'p(S∗) (cid:13)(cid:13)(cid:13)(cid:13)'p(S∗) ≤ 2(cid:88) (cid:13)(cid:13)(cid:13)'p(S∗) (cid:11) (cid:13)(cid:13)(cid:13)(cid:13)'p(S∗) ≤ 2K3 ≤ 4K3 (ε) k b k∈K Using (5.25), (5.14b), (5.14a) and the other off-diagonal estimate (5.14e) yields By (5.21a), estimate (5.22) and by unconditionality, we obtain a'b (ε) ' '∈K:'>k uKskyk'p(S∗). The latter estimate together with (5.31) yields (cid:10)H∗b (cid:13)(cid:13)(cid:13)(cid:13)(cid:88) k∈K ∗(ε) k ∗(ε) hb k (ε) '∈K:'>k a'b ' , Hb k i (ε) (cid:88) k∈K ηkkyk'p(S∗). (5.32) uKs Recall that we put ηk = K−1 u 4−k−1, k ∈ N, (see (5.10)). Combining the esti- mates (5.30) and (5.32) with (5.29) yields kP Hy − yk'p(S∗) ≤ 1 3kyk'p(S∗), y ∈ Y. (5.33) Let J : Y → 'p(S∗) denote the operator given by Jy = y, then P HJ : Y → Y is invertible by (5.33). Thus, if we define V = (P HJ)−1P, the following diagram is commutative: Y J 'p(S∗) IdY (P HJ)−1 Y H Y V P / 'p(S∗) kJkkV k ≤ 12K4 uK2 s . (5.34) , ,   ; ; / b b O O 16 R. LECHNER The operators J and V are both bounded, thus merging the diagrams (5.26) and (5.34) yields the commutative diagram 'p(S∗) Id'p(S∗) 'p(S∗) A Y J M 'p(S∗) QY Y IdY (P HJ)−1 Y H V P / 'p(S∗) N kMkkNk ≤ 48K7 uK4 s . (5.35) Step 4: 'p(S∗) is primary. Let Q : 'p(S∗) → 'p(S∗) denote a bounded projection. Then by (5.9), we know that 'p(S∗) is either isomorphic to a complemented sub- space of Q('p(S∗)), or isomorphic to a complemented subspace of (Id'p(S∗) −Q)('p(S∗)). Moreover, since 'p(S∗) is the 'p-sum of a Banach space, 'p(S∗) is isomorphic to 'p('p(S∗)). Thus, by Pełczyński's decomposition method (see [7]; see also [9, II.B.24]), we obtain that either Q('p(S∗)) is isomorphic to 'p(S∗), or (Id'p(S∗) −Q)('p(S∗)) (cid:3) is isomorphic to 'p(S∗). Acknowledgments. It is my pleasure to thank P.F.X. Müller for many helpful discussions. Supported by the Austrian Science Foundation (FWF) Pr.Nr. P28352. References [1] M. Capon. Primarité de certains espaces de Banach. Proc. London Math. Soc. (3), 45(1):113 -- 130, 1982. [2] P. G. Casazza, C. A. Kottman, and B. L. Lin. On some classes of primary Banach spaces. Canad. J. Math., 29(4):856 -- 873, 1977. [3] P. G. Casazza and B. L. Lin. Projections on Banach spaces with symmetric bases. Studia [4] R. Lechner. Factorization in SL∞. Israel J. Math., to appear. Preprint available on ArXiv Math., 52:189 -- 193, 1974. https://arxiv.org/abs/1611.00622. [5] J. Lindenstrauss. On complemented subspaces of m. Israel J. Math., 5:153 -- 156, 1967. [6] J. Lindenstrauss and L. Tzafriri. Classical Banach spaces. I. Springer-Verlag, Berlin-New York, 1977. Sequence spaces, Ergebnisse der Mathematik und ihrer Grenzgebiete, Vol. 92. [7] A. Pełczyński. Projections in certain Banach spaces. Studia Math., 19:209 -- 228, 1960. [8] I. Singer. Bases in Banach spaces. I. Springer-Verlag, New York-Berlin, 1970. Die Grundlehren der mathematischen Wissenschaften, Band 154. [9] P. Wojtaszczyk. Banach spaces for analysts, volume 25 of Cambridge Studies in Advanced Mathematics. Cambridge University Press, Cambridge, 1991. Richard Lechner, Institute of Analysis, Johannes Kepler University Linz, Altenberger Strasse 69, A-4040 Linz, Austria E-mail address: [email protected]   / /   , ,   O O ; ; / b b O O Y Y
1310.7626
1
1310
2013-10-28T21:12:03
A new resolvent equation for the S-functional calculus
[ "math.FA" ]
The S-functional calculus is a functional calculus for $(n+1)$-tuples of non necessarily commuting operators that can be considered a higher dimensional version of the classical Riesz-Dunford functional calculus for a single operator. In this last calculus, the resolvent equation plays an important role in the proof of several results. Associated with the S-functional calculus there are two resolvent operators: the left $S_L^{-1}(s,T)$ and the right one $S_R^{-1}(s,T)$, where $s=(s_0,s_1,\ldots,s_n)\in \mathbb{R}^{n+1}$ and $T=(T_0,T_1,\ldots,T_n)$ is an $(n+1)$-tuple of non commuting operators. These two S-resolvent operators satisfy the S-resolvent equations $S_L^{-1}(s,T)s-TS_L^{-1}(s,T)=\mathcal{I}$, and $sS_R^{-1}(s,T)-S_R^{-1}(s,T)T=\mathcal{I}$, respectively, where $\mathcal{I}$ denotes the identity operator. These equations allows to prove some properties of the S-functional calculus. In this paper we prove a new resolvent equation for the S-functional calculus which is the analogue of the classical resolvent equation. It is interesting to note that the equation involves both the left and the right S-resolvent operators simultaneously.
math.FA
math
A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS DANIEL ALPAY, FABRIZIO COLOMBO, JONATHAN GANTNER, AND IRENE SABADINI Contents Introduction 1. 2. Preliminary results 3. The case of several non commuting operators 3.1. Some applications 4. The case of commuting operators and the quaternionic case 4.1. The case of several commuting operators 4.2. The quaternionic setting References 1 4 6 12 20 20 22 22 L (s, T ) and the right one S−1 Abstract. The S-functional calculus is a functional calculus for (n + 1)-tuples of non necessarily commuting operators that can be considered a higher dimensional version of the classical Riesz-Dunford functional calculus for a single operator. In this last calculus, the resolvent equation plays an important role in the proof of several results. Associated with the S-functional calculus there are two resolvent operators: the left S−1 R (s, T ), where s = (s0, s1, . . . , sn) ∈ Rn+1 and T = (T0, T1, . . . , Tn) is an (n + 1)-tuple of non commuting operators. These two S-resolvent operators satisfy the S-resolvent equations S−1 L (s, T ) = I, and sS−1 R (s, T )T = I, respectively, where I denotes the identity operator. These equations allows to prove some properties of the S-functional calculus. In this paper we prove a new resolvent equation for the S-functional calculus which is the analogue of the classical resolvent equation. It is interesting to note that the equation involves both the left and the right S-resolvent operators simultaneously. L (s, T )s − T S−1 R (s, T )−S−1 1. Introduction The S-resolvent operators are a key tool in the definition of the higher dimensional version of the Riesz-Dunford functional calculus called S-functional calculus. This calculus works for (n+1)-tuples (T0, T1, . . . , Tn) of non necessarily commuting operators and is based on the so-called S-spectrum, see [14, 17]. In the case of a single operator the S-functional calculus reduces to the Riesz-Dunford functional calculus (see [21, 32]). When the operators (T0, T1, . . . , Tn) commute among themselves, this calculus admits a commutative version called SC-functional calculus. In this case the S-resolvent operator and the S-spectrum have a simpler expression, see [15]. 1991 Mathematics Subject Classification. MSC: 47S10, 30G35. Key words and phrases. n-tuples of non commuting operators, quaternionic operators, S-spectrum, right S-resolvent operator, left S-resolvent operator, resolvent equation, projectors. 1 2 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI The class of functions on which this calculus is based is the so called set of slice hyperholomorphic (or slice monogenic) functions which are defined on subsets of the Euclidean space Rn+1 and have values in the Clifford algebra Rn. For more details on the S-functional calculus and the function theory on which it is based see the the monograph [19]. As it happens for the classical theory of monogenic functions (see [10, 18, 20, 25]), also in the class of slice hyperholomorphic functions there is the notion of left as well as of right hyperholomorphicity. But despite what happens in the monogenic case, for slice hyperholomorphic functions the Cauchy formulas for left and for right slice hyperholomorphic functions have two different kernels; moreover each of these kernels can be written in two different ways. The calculus admits a quaternionic version, which works for quaternionic linear op- erators and is based on slice hyperholomorphic (or slice regular) functions defined on subsets of the real algebra of quaternions H with values in the quaternions, see [11, 13]. To explain our new result and its consequences, let us focus, at the moment, on the quaternionic setting which is simpler to illustrate. Let us denote by V a two sided quaternionic Banach space and let T : V → V be a bounded right (or left) linear operator. We recall that the S-spectrum is defined as σS(T ) = {s ∈ H : T 2 − 2Re(s)T + s2I is not invertible}, where s = s0 + s1i + s2j + s3k is a quaternion, Re(s) = s0, s2 = s2 left and the right S-resolvent operators are defined as 0 + s2 1 + s2 2 + s2 3. The S−1 L (s, T ) := −(T 2 − 2Re(s)T + s2I)−1(T − sI), s ∈ H \ σS(T ) (1.1) and S−1 R (s, T ) := −(T − sI)(T 2 − 2Re(s)T + s2I)−1, s ∈ H \ σS(T ), (1.2) respectively. The left S-resolvent operator satisfies the equation S−1 L (s, T )s − T S−1 L (s, T ) = I, s ∈ H \ σS(T ), and the right S-resolvent operator satisfies sS−1 R (s, T ) − S−1 R (s, T )T = I, s ∈ H \ σS(T ). (1.3) (1.4) Consider the complex plane CI := R + I R, for I ∈ S, where S is the unit sphere of purely imaginary quaternions. Observe that CI can be identified with a complex plane since I 2 = −1 for every I ∈ S. Let U ⊂ H be a suitable domain that contains the S-spectrum of T . We define for left slice hyperholomorphic functions f : U → H (see the precise definition in the sequel) the quaternionic functional calculus as f (T ) = 1 2π Z∂(U ∩CI ) S−1 L (s, T ) dsI f (s), (1.5) where dsI = −dsI, and for right slice hyperholomorphic functions f : U → H, we define f (T ) = 1 2π Z∂(U ∩CI ) f (s) dsI S−1 R (s, T ). (1.6) These definitions are well posed since the integrals do not depend neither on the open set U and nor on the complex plane CI and can be extended to the case of (n+1)-tuples of operators, using slice hyperholomorphic functions with values in a Clifford algebra. A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 3 Using a similar notion of hyperholomorphicity and the S-spectrum in [23] the authors introduce the continuous functional calculus in a quaternionic Hilbert space. The S-resolvent equations (1.3), (1.4) are useful to prove several properties of the S- functional calculus. However it is natural to ask if it is possible to obtain an analog of the classical resolvent equation (λI − G)−1(µI − G)−1 = (λI − G)−1 − (µI − G)−1 µ − λ , λ, µ ∈ C \ σ(G), (1.7) where G is a complex operator on a Banach space, which might be useful to prove other properties of the calculus. The main goal of this paper it to show that (1.7) can be generalized in this non commutative setting, but it involves both the left and the right S-resolvent operators. Precisely, we will show that S−1 R (s, T )S−1 L (p, T )]p L (p, T )]](p2 − 2s0p + s2)−1, L (p, T ) = [[S−1 − s[S−1 R (s, T ) − S−1 R (s, T ) − S−1 for s, p ∈ H \ σS(T ). It is also worthwhile to mention that the S-resolvent operator plays an important role in the definition of the quaternionic version of the counterpart of the operator (I − zA)−1 in the realization s(z) = D + zC(I − zA)−1B for Schur multipliers, see [2]. The reader is referred to [2, 3, 4] for Schur analysis in the slice hyperholomorphic setting and to [1] and [7] for an overview of Schur analysis in the complex setting. It is interesting to note that in literature there are other cases in which the authors consider two resolvent operators. We mention in particular the case of Schur analysis in the setting of upper triangular operators and in the setting of compact Riemann sur- faces. In the first case, the role of complex numbers is played by diagonal operators and there are two "point evaluations" of an operator at a diagonal, one left and one right, each corresponding to an associated resolvent operator; see [8, (2.4)-(2.6), p. 256], but the resolvent equation is related just with one resolvent at a time; see [8, Corollary 2.9, p 266]. In the setting of compact Riemann surfaces (see [35, 29] for the general setting) there is a resolvent operator associated to every meromorphic function on the given Riemann surface X (see [9, (4.1), p. 307], and one needs two such operators, associated to a pair of functions which generate the field of meromorphic functions on X, to study underlying spaces; see [9, §5]. The same resolvent equation is satisfied by all the resolvent operators; see [9, Theorem 4.2, p. 309]. In this setting both S-resolvent operators enter the resolvent equation. The plan of the paper is as follows. In Section 2 we recall some preliminary results on slice hyperholomorphic functions. In Section 3 we state and prove the new resolvent equation and we show that there are two possible versions which are equivalent. We prove our results for the S-functional calculus for (n + 1)-tuples of non commuting operators and we show some applications of the resolvent equation. In Section 4 we consider the commutative version of the S-functional calculus, the so called SC-functional calculus and we reformulate our main results for the quaternionic functional calculus. Since the proofs follow the lines of the ones for the case of (n + 1)- tuples of non commuting operators we will omit them in both cases. 4 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI 2. Preliminary results In this section we recall the notion of slice hyperholomorphic functions and their Cauchy formulas, see [19]. Let Rn be the real Clifford algebra over n imaginary units e1, . . . , en satisfying the relations eiej + ejei = 0, i 6= j, e2 i = −1. An element in the Clifford algebra will i1 < . . . < ir is a multi-index and eA = ei1ei2 . . . eir , e∅ = 1. An element (x0, x1, . . . , xn) ∈ Rn+1 will be j=1 xjej ∈ Rn called paravector and the real part x0 of x will also be denoted by Re(x). The norm of x ∈ Rn+1 is defined as x2 = x2 j=1 xjej. Let be denoted by PA eAxA where A = {i1 . . . ir} ∈ P{1, 2, . . . , n}, identified with the element x = x0 + x = x0 + Pn n. The conjugate of x is defined by ¯x = x0 − x = x0 −Pn 1 + . . . + x2 0 + x2 S = {x = e1x1 + . . . + enxn x2 1 + . . . + x2 n = 1}; for I ∈ S we obviously have I 2 = −1. Given an element x = x0 + x ∈ Rn+1 let us set Ix = x/x if x 6= 0, and given an element x ∈ Rn+1, the set [x] := {y ∈ Rn+1 : y = x0 + Ix, I ∈ S} is an (n − 1)-dimensional sphere in Rn+1. The vector space R + I R passing through 1 and I ∈ S will be denoted by CI and an element belonging to CI will be indicated by u + Iv, for u, v ∈ R. With an abuse of notation we will write x ∈ Rn+1. Thus, if U ⊆ Rn+1 is an open set, a function f : U ⊆ Rn+1 → Rn can be interpreted as a function of the paravector x. Definition 2.1 (Slice hyperholomorphic functions). Let U ⊆ Rn+1 be an open set and let f : U → Rn be a real differentiable function. Let I ∈ S and let fI be the restriction of f to the complex plane CI. The function f is said to be left slice hyperholomorphic (or slice monogenic) if, for every I ∈ S, on U ∩ CI it satisfies 1 2 (cid:18) ∂ ∂u fI(u + Iv) + I ∂ ∂v fI(u + Iv)(cid:19) = 0. We will denote by SM(U) the set of left slice hyperholomorhic functions on the open set U or by SML(U) when confusion may arise. The function f is said to be right slice hyperholomorphic (or right slice monogenic) if, for every I ∈ S, on U ∩ CI, it satisfies 1 2 (cid:18) ∂ ∂u fI(u + Iv) + ∂ ∂v fI(u + Iv)I(cid:19) = 0. We will denote by SMR(U) the set of right slice hyperholomorphic functions on the open set U. Slice hyperholomorphic functions possess good properties when they are defined on suitable domains which are introduced in the following definition. Definition 2.2 (Axially symmetric slice domain). Let U be a domain in Rn+1. We say that U is a slice domain (s-domain for short) if U ∩ R is non empty and if U ∩ CI is a domain in CI for all I ∈ S. We say that U is axially symmetric if, for all x ∈ U, the (n − 1)-sphere [x] is contained in U. A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 5 Definition 2.3 (Cauchy kernel for left slice hyperholomorphic functions). Let x, s ∈ Rn+1 be such that x 6∈ [s]. Let S−1 L (s, x) be the function defined by S−1 L (s, x) := −(x2 − 2xRe[s] + s2)−1(x − s). (2.8) We say that S−1 written in form I. L (s, x) is the Cauchy kernel (for left slice hyperholomorphic functions) Proposition 2.4. Suppose that x and s ∈ Rn+1 are such that x 6∈ [s]. The following identity holds: − (x − ¯s)(x2 − 2Re(s)x + s2)−1 = (s2 − 2Re(x)s + x2)−1(s − ¯x). (2.9) Remark 2.5. By Proposition 2.4 S−1 L (s, x) can also be written as S−1 L (s, x) := (s − ¯x)(s2 − 2Re(x)s + x2)−1. (2.10) In this case, we will say S−1 L (s, x) is written in form II. Proposition 2.6. The function S−1 x and right slice hyperholomorphic in the variable s for x 6∈ [s]. L (s, x) is left slice hyperholomorphic in the variable The case of the Cauchy kernel for right slice hyperholomorphic functions is similar. Definition 2.7 (Cauchy kernel for right slice hyperholomorphic functions). Let x, s ∈ Rn+1 be such that x 6∈ [s]. The Cauchy kernel S−1 R (s, x) for right slice hyperholomorphic functions is defined by S−1 R (s, x) := −(x − ¯s)(x2 − 2Re(s)x + s2)−1. (2.11) We say that S−1 R (s, x) is written in form I. Remark 2.8. An analog of Proposition 2.4 holds in fact: − (x − ¯s)(x2 − 2Re(s)x + s2)−1 = (s2 − 2Re(x)s + x2)−1(s − ¯x), (2.12) for x, s ∈ Rn+1 such that x 6∈ [s]. Thus S−1 R (s, x) can be written as S−1 R (s, x) = (s2 − 2Re(x)s + x2)−1(s − ¯x), and in this case we say that S−1 R (s, x) is written in form II. Theorem 2.9 (The Cauchy formula with slice hyperholomorphic kernel). Let U ⊂ Rn+1 be an axially symmetric s-domain. Suppose that ∂(U ∩ CI) is a finite union of continuously differentiable Jordan curves for every I ∈ S. Set dsI = −dsI for I ∈ S. • If f is a (left) slice hyperholomorphic function on a set that contains U then f (x) = 1 2π Z∂(U ∩CI ) S−1 L (s, x)dsIf (s) (2.13) and the integral does not depend on U and on the imaginary unit I ∈ S. • If f is a right slice hyperholomorphic function on a set that contains U , then f (x) = 1 2π Z∂(U ∩CI ) f (s)dsIS−1 R (s, x) (2.14) and the integral does not depend on U and on the imaginary unit I ∈ S. 6 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI The above Cauchy formulas are the starting point to define the S-functional calculus. A crucial fact of slice hyperholomorphic functions in the representation formula (also called structure formula). This formula will be used in the sequel to give applications of the new resolvent equation. Theorem 2.10 (Representation Formula). Let U be an axially symmetric s-domain U ⊆ H. • Let f be a (left) slice hyperholomorphic function on U. Choose any J ∈ S. Then the following equality holds for all x = u + Iv ∈ U: f (u + Iv) = 1 2hf (u + Jv) + f (u − Jv)i + I 2hJ[f (u − Jv) − f (u + Jv)]i. 1 (2.15) Moreover, for all u, v ∈ R such that u + vS ⊆ U, there exist Rn-valued functions α, β depending on u, v only such that for all K ∈ S 1 2hf (u + Kv) + f (u − Kv)i = α(u, v) and 2hK[f (u − Kv) − f (u + Kv)]i = β(u, v). (2.16) • Let f be a right slice hyperholomorphic function on U. Choose any J ∈ S. Then 1 1 the following equality holds for all x = u + Iv ∈ U: f (u + Iv) = 1 2hf (u + Jv) + f (u − Jv)i + 2h[f (u − Jv) − f (u + Jv)]JiI. (2.17) Moreover, for all u, v ∈ R such that u + vS ⊆ U, there exist Rn-valued functions α, β depending on u, v only such that for all K ∈ S 1 2hf (u + Kv) + f (u − Kv)i = α(u, v) and 1 2h[f (u − Kv) − f (u + Kv)]Ki = β(u, v). (2.18) 3. The case of several non commuting operators In the sequel, we will consider a Banach space V over R with norm k·k. It is possible to endow V with an operation of multiplication by elements of Rn which gives a two-sided module over Rn. A two-sided module V over Rn is called a Banach module over Rn, if there exists a constant C ≥ 1 such that kvak ≤ Ckvka and kavk ≤ Cakvk for all v ∈ V and a ∈ Rn. By Vn we denote V ⊗ Rn over Rn; Vn turns out to be a two-sided Banach module . i1 < . . . < ir is a multi-index). The multiplications of an element v ∈ Vn with a scalar V . We denote by B(V ) the space of bounded R-homomorphisms of the Banach space V to itself endowed with the natural norm denoted by k · kB(V ). Given TA ∈ B(V ), we An element in Vn is of the type PA vA ⊗ eA (where A = i1 . . . ir, iℓ ∈ {1, 2, . . . , n}, a ∈ Rn are defined by va = PA vA ⊗ (eAa) and av = PA vA ⊗ (aeA). For simplicity, we will write PA vAeA instead of PA vA ⊗ eA. Finally, we define kvk2 Vn = PA kvAk2 can introduce the operator T = PA TAeA and its action on v = P vBeB ∈ Vn as T (v) = PA,B TA(vB)eAeB. The operator T is a right-module homomorphism which is In the sequel, we will consider operators of the form T = T0 + Pn j=1 ejTj where Tj ∈ B(V ) for j = 0, 1, . . . , n. The subset of such operators in B(Vn) will be denoted by B0,1(Vn). We define kT kB0,1(Vn) = Pj kTjkB(V ). Note that, in the sequel, we will omit the subscript B0,1(Vn) in the norm of an operator. Note also that kT Sk ≤ kT kkSk. a bounded linear map on Vn. A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 7 Definition 3.1. Let T ∈ B0,1(Vn). We define the left Cauchy kernel operator series or S-resolvent operator series as S−1 L (s, T ) = Xn≥0 T ns−1−n, and the right Cauchy kernel operator series as S−1 R (s, T ) = Xn≥0 s−1−nT n, for kT k < s. (3.19) (3.20) The Cauchy kernel operator series are the power series expansion of the S-resolvent operators. Their sum is computed in the following result: Theorem 3.2. Let T ∈ B0,1(Vn) and let s ∈ H. Then, for kT k < s, we have T ms−1−m = −(T 2 − 2Re(s)T + s2I)−1(T − sI), s−1−mT m = −(T − sI)(T 2 − 2Re(s)T + s2I)−1. (3.21) (3.22) Xm≥0 Xm≥0 We observe that the sum of the above series are independent of the fact that the components of the paravector operator T commute. Moreover the operators on right hand sides of (3.21) and (3.22) are defined on a subset of Rn+1 that is larger then {s ∈ Rn+1 : kT k < s}. This fact suggests the definition of S-spectrum, of S-resolvent set and of S-resolvent operators. Definition 3.3 (The S-spectrum and the S-resolvent set). Let T ∈ B0,1(Vn). We define the S-spectrum σS(T ) of T as: σS(T ) = {s ∈ Rn+1 : T 2 − 2 Re(s)T + s2I is not invertible}. The S-resolvent set ρS(T ) is defined by ρS(T ) = Rn+1 \ σS(T ). Definition 3.4 (The S-resolvent operators). Let T ∈ B0,1(Vn) and s ∈ ρS(T ). We define the left S-resolvent operator as S−1 L (s, T ) := −(T 2 − 2Re(s)T + s2I)−1(T − sI), and the right S-resolvent operator as S−1 R (s, T ) := −(T − sI)(T 2 − 2Re(s)T + s2I)−1. (3.23) (3.24) The operators S−1 L (s, T ) and S−1 R (s, T ) satisfy the equations below, see [19]: Theorem 3.5. Let T ∈ B0,1(Vn) and let s ∈ ρS(T ). Then, the left S-resolvent operator satisfies the equation and the right S-resolvent operator satisfies the equation S−1 L (s, T )s − T S−1 L (s, T ) = I, sS−1 R (s, T ) − S−1 R (s, T )T = I. (3.25) (3.26) 8 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI Our goal is to establish the analogue of the classical resolvent equation. To this end, we need some preliminary results. A crucial fact is the following Theorem 3.6 that will give us the hint to discover what is the structure of the resolvent equation in this non commutative setting at least in the case the S-resolvent operators are expressed in power series. Theorem 3.6. Let A, B ∈ B(Vn) and let s, p ∈ Rn+1. Then, for p < s, we have and Xm≥0 Xm≥0 pmAs−1−m = −(p2 − 2Re(s)p + s2)−1(pA − As), (3.27) s−1−mBpm = −(Bp − sB)(p2 − 2Re(s)p + s2)−1. (3.28) Moreover, (3.28) can be written as s−1−mBpm = (s2 − 2Re(p)s + p2)−1(sB − Bp). (3.29) Xm≥0 Proof. To verify (3.27) define X := (p2 − 2Re(s)p + s2)Xm≥0 pmAs−1−m and observe that X = Xm≥0 (p2 − 2Re(s)p + s2)pmAs−1−m = p2As−1 − 2Re(s)pAs−1 + s2As−1 + p3As−2 − 2Re(s)p2As−2 + s2pAs−2 + p4As−3 − 2Re(s)p3As−3 + s2p2As−3 + . . . = −(pA − As) + Xm≥2 pmA(s2 − 2Re(s)s + s2)s−1−m. (3.30) Since any paravector s satisfies s2 − 2Re(s)s + s2 = 0 we deduce that X = (p2 − 2Re(s)p + s2)Xm≥0 pmAs−1−m = −(pA − As) and the statement follows. The equality in (3.28) can be verified by setting Y := Xm≥0 s−1−mBpm(p2 − 2Re(s)p + s2) and observing that Y = −(Bp − sB) + Xm≥0 s−1−mBpm(p2 − 2Re(p)p + p2) = −(Bp − sB). With similar computations one can verify equality (3.29). (cid:3) A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 9 Corollary 3.7. Let A, B ∈ B(Vn) and let s, p be paravectors. Then, for p < s, the following equations hold m Xj=0 m Xj=0 and pjAs−1−j = −(p2 − 2Re(s)p + s2)−1(pA − As) + pm+1(p2 − 2Re(s)p + s2)−1(pA − As)s−1−m, s−1−jBpj = −(Bp − sB)(p2 − 2Re(s)p + s2)−1 + s−1−m(Bp − sB)(p2 − 2Re(s)p + s2)−1pm+1. Moreover, (3.32) can also written as m Xj=0 s−1−jBpj = (s2 − 2Re(p)s + p2)−1(sB − Bp) − s−1−m(s2 − 2Re(p)s + p2)−1(sB − Bp)pm+1. (3.31) (3.32) (3.33) Proof. Identity (3.31) follows from m Xj=0 pjAs−1−j = ∞ Xj=0 pjAs−1−j − ∞ Xj=m+1 pjAs−1−j, that can be written as m Xj=0 pjAs−1−j = ∞ Xj=0 pjAs−1−j − pm+1(cid:16) ∞ Xj=0 pjAs−1−j(cid:17)s−1−m, but now we use (3.27) to get the result. Identity (3.32) and (3.33) follow with similar computations. (cid:3) We now prove the new S-resolvent equation. In the proof we first consider the case in which the S-resolvent operators admit the power series expansion S−1 L (s, T ) = Xm≥0 T ms−1−m, S−1 R (s, T ) = Xm≥0 s−1−mT m, that is for kT k < s. Then we verify that such equation holds in general. Theorem 3.8. Let T ∈ B0,1(Vn) and let s and p ∈ ρS(T ). Then we have R (s, T )S−1 S−1 L (p, T ))p−s(S−1 R (s, T )−S−1 L (p, T ) = ((S−1 R (s, T )−S−1 L (p, T )))(p2 −2s0p+s2)−1. (3.34) Moreover, the resolvent equation can also be written as R (s, T )−S−1 R (s, T )S−1 S−1 L (p, T ) = (s2−2p0s+p2)−1(s(S−1 L (p, T ))−(S−1 R (s, T )−S−1 L (p, T ))p). (3.35) Proof. We prove the theorem in two steps. STEP I. First we assume that the S-resolvent operators are expressed in power series. If kT k < p < s then the S-resolvent operators have power series expansion and so S−1 R (s, T )S−1 L (p, T ) = (Xj≥0 s−1−jT j)(Xj≥0 T jp−1−j). (3.36) 10 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI By setting (3.36) can be written as Λm(s, p; T ) := m Xj=0 s−1−j(T mp−1−m)pj S−1 R (s, T )S−1 L (p, T ) = Xm≥0 Λm(s, p; T ). Formula (3.32) with B = T mp−1−m and some computations give Λm(s, p; T ) = −((T mp−1−m)p − s(T mp−1−m))(p2 − 2Re(s)p + s2)−1 + s−1−m((T mp−1−m)p − s(T mp−1−m))(p2 − 2Re(s)p + s2)−1pm+1 = −[(T mp−1−m)p − s(T mp−1−m) + (s−1−mT m)p − s(s−1−mT m)](p2 − 2Re(s)p + s2)−1. (3.37) From the chain of equalities S−1 R (s, T )S−1 L (p, T ) = Xm≥0 Λm(s, p; T ) = −[(Xm≥0 + (Xm≥0 (T mp−1−m)p − sXm≥0 s−1−mT m)p − sXm≥0 (T mp−1−m))] s−1−mT m)](p2 − 2Re(s)p + s2)−1 (3.34) follows. To prove that the resolvent equation can be written in the second form (3.35) observe that Λm(s, p; T ) can also be written using (3.29) as Λm(s, p; T ) = (s2 − 2Re(p)s + p2)−1(s(T mp−1−m) − (T mp−1−m)p) − s−1−m(s2 − 2Re(p)s + p2)−1(s(T mp−1−m) − (T mp−1−m)p)pm+1. (3.39) (3.38) L (p, T ) defined in (3.23) and (3.24), respectively. so taking the sum Pm≥0 Λm(s, p; T ) we get the second version of the resolvent equation. STEP II. We prove that, for s and p ∈ ρS(T ), (3.34) and (3.35) hold with S−1 and S−1 Let us verify (3.34). Since s and p ∈ ρS(T ) the left and right S-resolvent operators defined by (3.23) and (3.24) satisfy the left and the right resolvent equations (3.25) and (3.26), respectively. To verify (3.34) we have to show that S−1 L (p, T )(p2 − 2s0p + s2) equals R (s, T )S−1 R (s, T ) (S−1 R (s, T ) − S−1 L (p, T ))p − s(S−1 R (s, T ) − S−1 L (p, T )). To do this we use the left and the right S-resolvent equations (3.25), (3.26). Indeed, using the left S-resolvent equation, written as S−1 L (p, T )p = T S−1 L (p, T ) + I, A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 11 we have L (p, T )(p2 − 2s0p + s2) = S−1 R (s, T )[S−1 L (p, T )p]p S−1 R (s, T )S−1 − 2s0S−1 = S−1 + s2S−1 R (s, T )S−1 R (s, T )[T S−1 R (s, T )S−1 L (p, T )p + s2S−1 L (p, T ) + I]p − 2s0S−1 L (p, T ) R (s, T )S−1 L (p, T ) R (s, T )[T S−1 L (p, T ) + I] (3.40) and using again the left S-resolvent equation L (p, T )(p2 − 2s0p + s2) = S−1 L (p, T ) + I] + s2S−1 S−1 R (s, T )S−1 − 2s0S−1 R (s, T )[T S−1 R (s, T )T [T S−1 R (s, T )S−1 L (p, T ) L (p, T ) + I] + S−1 R (s, T )p we obtain S−1 R (s, T )S−1 R (s, T )T ]T S−1 L (p, T )(p2 − 2s0p + s2) = [S−1 − 2s0[[S−1 + s2S−1 R (s, T )T ]S−1 R (s, T )S−1 L (p, T ). Now we use the right S-resolvent equation L (p, T ) + S−1 R (s, T )T ] + S−1 R (s, T )p L (p, T ) + S−1 R (s, T )] (3.41) (3.42) (3.43) (3.44) (3.45) we obtain S−1 R (s, T )T = sS−1 R (s, T ) − I S−1 R (s, T )S−1 L (p, T )(p2 − 2s0p + s2) = [[sS−1 R (s, T ) − I]T ]S−1 − 2s0[[sS−1 R (s, T ) − I]S−1 + s2S−1 L (p, T ). R (s, T )S−1 L (p, T ) + sS−1 L (p, T ) + S−1 R (s, T ) − I] + S−1 R (s, T )] R (s, T )p Iterating the use of the above right S-resolvent equation we get S−1 R (s, T )S−1 R (s, T ) − I] − T ]S−1 L (p, T )(p2 − 2s0p + s2) = [s[sS−1 − 2s0[[sS−1 + s2S−1 R (s, T )S−1 R (s, T )S−1 L (p, T ), L (p, T ) + sS−1 R (s, T ) − I] + S−1 R (s, T )p L (p, T ) − S−1 L (p, T )] + S−1 R (s, T )] which leads to S−1 R (s, T )S−1 L (p, T )(p2 − 2s0p + s2) = (s2 − 2s0s + s2)S−1 + [S−1 R (s, T ) − S−1 R (s, T )S−1 L (p, T ) L (p, T )]p − s[S−1 R (s, T ) − S−1 L (p, T )], and since s2 − 2s0s + s2 = 0 we obtain (3.34). With similar computations we can show that also (3.35) holds. (cid:3) We now observe that in the commutative case besides the resolvent equation, also the following relation between the resolvent operators (λI − G)−1(µI − G)−1 = (µI − G)−1(λI − G)−1, for λ, µ ∈ ρ(G) 12 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI holds. In the non commutative case we cannot aspect the validity of such a relation, however we will show that an analogous equation holds for the so-called pseudo S- resolvent operators defined below. Definition 3.9. Let T ∈ B0,1(Vn). We define, for s ∈ ρS(T ), the pseudo S-resolvent operator of T is defined as Qs(T ) := (T 2 − 2Re(s)T + s2I)−1. With the above definition the resolvents S−1 L (s, T ) and S−1 R (s, T ) become and S−1 L (s, T ) := −Qs(T )(T − sI), s ∈ ρS(T ), S−1 R (s, T ) := −(T − sI)Qs(T ), s ∈ ρS(T ). We now prove the following: Theorem 3.10. Let T ∈ B0,1(Vn) and let s, p ∈ ρS(T ). Then we have (T − sI)Qs(T )Qp(T )(T − pI) = (T − sI)Qp(T )Qs(T )(T − pI). Proof. It follows from the fact that (T 2 − 2Re(s)T + s2I)(T 2 − 2Re(p)T + p2I) = (T 2 − 2Re(p)T + p2I)(T 2 − 2Re(s)T + s2I). Since s, p ∈ ρS(T ) we can take the inverse and the statement follows. Remark 3.11. Observe that the function FT (s, p) defined by (3.46) (3.47) (3.48) (cid:3) FT (s, p) := S−1 R (s, T )S−1 L (p, T ) is left slice hyperholomorphic in s and it is right slice hyperholomorphic in p with values in B(Vn). The function GT (s, p) := S−1 L (p, T )S−1 R (s, T ) is not slice hyperholomorphic neither in p nor in s. Remark 3.12. Using the star products left and right in the variables s, p, which will be denoted by ⋆s,lef t, ⋆p,right respectively, see [6], the resolvent equation (3.34) can be written as S−1 R (s, T )S−1 L (p, T ) = [S−1 R (s, T ) − S−1 L (p, T )] ⋆s,lef t (p − s)(p2 − 2Re(s)p + s2)−1I, or S−1 R (s, T )S−1 L (p, T ) = (s − p)(s2 − 2Re(p)s + p2)−1I ⋆p,right [S−1 R (s, T ) − S−1 L (p, T )]. 3.1. Some applications. Here we recall the formulations of the S-functional calculus and the we use the resolvent equation to deduce some results. We first recall two important properties of the S-spectrum. Theorem 3.13 (Structure of the S-spectrum). Let T ∈ B0,1(Vn) and suppose that p = p0 + p belongs σS(T ) with p 6= 0. Then all the elements of the (n − 1)-sphere [p] belong to σS(T ). This result implies that if p ∈ σS(T ) then either p is a real point or the whole (n − 1)- sphere [p] belongs to σS(T ). A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 13 Theorem 3.14 (Compactness of S-spectrum). Let T ∈ B0,1(Vn). Then the S-spectrum σS(T ) is a compact nonempty set. Moreover, σS(T ) is contained in {s ∈ Rn+1 : s ≤ kT k }. Definition 3.15. Let Vn be a two sided Banach module, T ∈ B0,1(Vn)and let U ⊂ Rn+1 be an axially symmetric s-domain that contains the S-spectrum σS(T ) such that ∂(U ∩ CI) is the union of a finite number of continuously differentiable Jordan curves for every I ∈ S. In this case we say that U is a T -admissible open set. We can now introduce the class of functions for which we can define the two versions of the S-functional calculus. Definition 3.16. Let Vn be a two sided Banach module, T ∈ B0,1(Vn) and let W be an open set in Rn+1. (i) A function f ∈ SML(W ) is said to be locally left hyperholomorphic on σS(T ) if there exists a T -admissible domain U ⊂ Rn+1 such that U ⊂ W , on which f is left slice hyperholomorphic. We will denote by SML σS (T ) the set of locally left hyperholomorphic functions on σS(T ). (ii) A function f ∈ SMR(W ) is said to be locally right regular on σS(T ) if there exists a T -admissible domain U ⊂ Rn+1 such that U ⊂ W , on which f is right slice hyperholomorphic. We will denote by SMR σS(T ) the set of locally right slice hyperholomorphic functions on σS(T ). Definition 3.17 (The S-functional calculus). Let Vn be a two sided Banach module and T ∈ B0,1(Vn). Let U ⊂ Rn+1 be a T -admissible domain and set dsI = −dsI. We define and f (T ) = 1 2π Z∂(U ∩CI ) f (T ) = 1 2π Z∂(U ∩CI ) S−1 L (s, T ) dsI f (s), for f ∈ SML σS (T ), (3.49) f (s) dsI S−1 R (s, T ), for f ∈ SMR σS (T ). (3.50) We now define the Riesz projectors for the S-functional calculus. We begin with a preliminary lemma. Lemma 3.18. Let B ∈ B(Vn) and let G be an axially symmetric s-domain such that p ∈ G. Then (sB − Bp)(p2 − 2s0p + s2)−1 = (s2 − 2p0s + p2)−1(sB − Bp), p 6∈ [s], (3.51) and 1 2π Z∂(G∩CI ) dsI(sB − Bp)(p2 − 2s0p + s2)−1 = B. (3.52) 14 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI Proof. Formula (3.51) is obtained by direct computation. Let us prove (3.52), so we write 1 2π Z∂(G∩CI ) dsI(sB − Bp)(p2 − 2s0p + s2)−1 = = + 1 1 1 2π Z∂(G∩CI ) 2π Z∂(G∩CI ) 2π Z∂(G∩CI ) dsI(s2 − 2p0s + p2)−1(sB − Bp) dsI(s2 − 2p0s + p2)−1(s − p)B dsI(s2 − 2p0s + p2)−1(pB − Bp) but observe that 1 2π Z∂(G∩CI ) dsI(s2 − 2p0s + p2)−1(s − p)B = 1 2π Z∂(G∩CI ) dsIS−1 R (s, p)B = B and moreover by the residue theorem it is 1 2π Z∂(G∩CI ) dsI(s2 − 2p0s + p2)−1 = 0 so we get the statement. (cid:3) Theorem 3.19. Let T ∈ B0,1(Vn) and let σS(T ) = σ1S(T ) ∪ σ2S(T ), with dist (σ1S(T ), σ2S(T )) > 0. Let U1 and U2 be two axially symmetric s-domains such that σ1S(T ) ⊂ U1 and σ2S(T ) ⊂ U2, with U 1 ∩ U 2 = ∅. Set Pj := Tj := 1 2π Z∂(Uj ∩CI ) 2π Z∂(Uj ∩CI ) 1 S−1 L (s, T ) dsI, j = 1, 2, S−1 L (s, T ) dsI s, j = 1, 2. (3.53) (3.54) Then Pj are projectors and T Pj = PjT for j = 1, 2. Proof. Let σjS(T ) ⊂ G1 and G2 be two T -admissible open sets such that G1∪∂G1 ⊂ G2 and G2 ∪ ∂G2 ⊂ Uj, for j = 1 or 2. Thanks to the structure of the S-spectrum we will assume that G1 and G2 are axially symmetric and s-domains. Take p ∈ ∂(G1 ∩ CI) and s ∈ ∂(G2 ∩ CI) and observe that, for I ∈ S, we have Pj := 1 2π Z∂(G2∩CI ) dsIS−1 R (s, T ) but we can also write Pj as Pj = 1 2π Z∂(G1∩CI ) S−1 L (p, T )dpI. Now consider P 2 j written as P 2 j = 1 (2π)2 Z∂(G2∩CI ) dsI Z∂(G1∩CI ) S−1 R (s, T )S−1 L (p, T )dpI. A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 15 Using the resolvent equation we write: P 2 = − 1 (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI ) 1 dsI Z∂(G1∩CI ) dsI Z∂(G1∩CI ) [S−1 R (s, T ) − S−1 L (p, T )]p(p2 − 2s0p + s2)−1dpI s[S−1 R (s, T ) − S−1 L (p, T )](p2 − 2s0p + s2)−1dpI. Now observe that 1 (2π)2 Z∂(G2∩CI ) dsIS−1 R (s, T )Z∂(G1∩CI ) p(p2 − 2s0p + s2)−1dpI = 0 and − 1 (2π)2 Z∂(G2∩CI ) dsIsS−1 R (s, T )Z∂(G1∩CI ) (p2 − 2s0p + s2)−1dpI = 0 since the functions p 7→ p(p2 − 2s0p + s2)−1, p 7→ (p2 − 2s0p + s2)−1 are slice hyperholomorphic and do not have singularities inside ∂(G1 ∩ CI). So P 2 be written as j can P 2 j = − = 1 1 dsI Z∂(G1∩CI ) dsI Z∂(G1∩CI ) (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI )Z∂(G1∩CI ) 1 −S−1 L (p, T )p(p2 − 2s0p + s2)−1dpI −sS−1 L (p, T )(p2 − 2s0p + s2)−1dpI, dsI(sS−1 L (p, T ) − S−1 L (p, T )p)(p2 − 2s0p + s2)−1dpI. Applying now Lemma 3.18 with B := S−1 have L (p, T ) and observing that p ∈ G2, we finally P 2 j = 1 2π Z∂(G1∩CI ) S−1 L (p, T )dpI = Pj. Let us now prove that T Pj = PjT . Observe that the functions f (s) = sm, for m ∈ N0 are both right and left slice hyperholomorphic. So the operator T can be written as T = 1 2π Z∂(U ∩CI ) S−1 L (s, T ) dsI s = 1 2π Z∂(U ∩CI ) s dsI S−1 R (s, T ); analogously, as already observed, for the projectors Pj we have Pj = 1 2π Z∂(Uj ∩CI ) S−1 L (s, T ) dsI = 1 2π Z∂(Uj ∩CI ) dsI S−1 R (s, T ). From the identity Tj = 1 2π Z∂(Uj ∩CI ) S−1 L (s, T ) dsI s = 1 2π Z∂(Uj ∩CI ) s dsI S−1 R (s, T ) we can compute T Pj as: T Pj = 1 2π Z∂(Uj ∩CI ) T S−1 L (s, T ) dsI 16 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI and using the resolvent equation (3.25) it follows T Pj = = = 1 1 1 2π Z∂(Uj ∩CI ) 2π Z∂(Uj ∩CI ) 2π Z∂(Uj ∩CI ) [S−1 L (s, T ) s − I] dsI S−1 L (s, T ) s dsI S−1 L (s, T ) dsI s Now consider = Tj. PjT = 1 2π Z∂(Uj ∩CI ) dsI S−1 R (s, T )T and using the resolvent equation (3.26) we obtain PjT = = 1 2π Z∂(Uj ∩CI ) 2π Z∂(Uj ∩CI ) 1 dsI [s S−1 R (s, T ) − I] dsI s S−1 R (s, T ) so the equality PjT = T Pj holds. = Tj, (3.55) (3.56) (cid:3) Remark 3.20. The properties that the Riesz projectors commute with the operator T has been proved for the quaternionic version of the S-functional calculus in [4], while the property that P 2 = P given in [19] is obtained heuristically. This fact shows the importance of this new resolvent equation. As it is well known for hyperholomorphic functions the product of two hyperholomor- phic functions is not in general hyperholomorphic. Here we recall a class of functions for which the pointwise multiplication remains slice hyperholomorphic. Definition 3.21. Let f : U → Rn be a slice hyperholomorphic function, where U is an open set in Rn+1. We define N (U) = {f ∈ SM(U) : f (U ∩ CI) ⊆ CI, ∀I ∈ S}. Proposition 3.22. Let U be an open set in Rn+1. Let f ∈ N (U), g ∈ SM(U), then f g ∈ SM(U). First of all let us observe that functions in the subclass N (U) are both left and right slice hyperholomorphic. When we take the power series expansion of this class of functions at a point on the real line the coefficients of the expansion are real numbers. Now observe that for functions in f ∈ N (U) we can define f (T ) using the left but also the right S-functional calculus. It is f (T ) = = 1 2π Z∂(U ∩CI ) 2π Z∂(U ∩CI ) 1 S−1 L (s, T ) dsI f (s) f (s) dsI S−1 R (s, T ). A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 17 Lemma 3.23. Let B ∈ B(Vn). Let G be an axially symmetric s-domain and assume that f ∈ N (G). Then, for p ∈ G, we have 1 2π Z∂(G∩CI ) f (s)dsI(sB − Bp)(p2 − 2s0p + s2)−1 = Bf (p). Proof. Recalling formula (3.51) we write 1 2π Z∂(G∩CI ) f (s)dsI(sB − Bp)(p2 − 2s0p + s2)−1 = = + 1 1 2π Z∂(G∩CI ) 2π Z∂(G∩CI ) 2π Z∂(G∩CI ) 1 f (s)dsI(s2 − 2p0s + p2)−1(sB − Bp) f (s)dsI(s2 − 2p0s + p2)−1(s − p)B f (s)dsI(s2 − 2p0s + p2)−1(pB − Bp) := J1 + J2 but observe that J1 = = 1 2π Z∂(G∩CI ) 2π Z∂(G∩CI ) 1 f (s)dsI(s2 − 2p0s + p2)−1(s − p)B f (s)dsIS−1 R (s, p)B = f (p)B. Consider now the second integral. Taking s = u + Iv then the solutions of the equation s2 − 2p0s + p2 = 0 are s1 = α and s2 = α where α = p0 + Ip, so J2 = = 1 2π Z∂(G∩CI ) 2π Z∂(G∩CI ) 1 f (s)dsI(s2 − 2p0s + p2)−1(pB − Bp) f (s) (s − α)(s − α) dsI(pB − Bp), by the residues theorem we get J2 = 1 2π Z∂(G∩CI ) f (s)dsI(s2 − 2p0s + p2)−1(pB − Bp) = I 2p [f (p0 − Ip) − f (p0 + Ip)](pB − Bp). Now we recall the structure formula that shows that a slice hyperholomorphic function can be written as f (p) = α(p0, p) + Ipβ(p0, p) where α(p0, p) = β(p0, p) = 1 2 I 2 [f (p0 − Ip) + f (p0 + Ip)], [f (p0 − Ip) − f (p0 + Ip)] 18 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI and in the case of functions f ∈ N the functions α and β are real valued. Observe that J1 + J2 = f (p)B + I 2p [f (p0 − Ip) − f (p0 + Ip)](pB − Bp) = α(p0, p)B + Ipβ(p0, p)B + = α(p0, p)B + Ipβ(p0, p)B + = B(α(p0, p) + Ipβ(p0, p)) = Bf (p), β(p0, p) p β(p0, p) p (pB − Bp) ((p0 − Ipp)B − B(p0 − Ipp)) so we get the statement. (cid:3) Remark 3.24. If we assume that f ∈ N (B(0, r)) where B(0, r) is the open ball in Rn+1 centered at 0 and of radius r > 0 and s ∈ B(0, r), then the proof of the above theorem follows in a shorter way. Indeed we have (sB − Bp)(p2 − 2s0p + s2)−1 = Xm≥0 s−1−mBpm, p < s. So but 1 2π Z∂(G∩CI ) f (s)dsI Xm≥0 s−1−mBpm, p < s, 1 2π Z∂(G∩CI ) Xm≥0 f (s)dsIs−1−mBpm = Xm≥0 1 m! f (m)(0)Bpm and for functions in N (B(0, r)) the derivatives f (m)(0) are real numbers and so they commute with B. We get 1 m! Xm≥0 f (m)(0)Bpm = B Xm≥0 1 m! f (m)(0)pm = Bf (p). We now offer a different proof of the theorem that shows that (f g)(T ) = f (T )g(T ), under suitable assumptions of f, g. Originally, see [19], the proof was based on the Cauchy formula and the resolvent equations (3.25), (3.26). Theorem 3.25. Let T ∈ B0,1(Vn) and assume f ∈ NσS (T ) and g ∈ SMσS (T ). Then we have (f g)(T ) = f (T )g(T ). Proof. Let σS(T ) ⊂ G1 and G2 be two T -admissible open sets such that G1 ∪ ∂G1 ⊂ G2 and G2 ∪ ∂G2 ⊂ U. Take p ∈ ∂(G1 ∩ CI) and s ∈ ∂(G2 ∩ CI) and observe that, for I ∈ S, we have A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 19 f (T )g(T ) = = − − + 1 1 1 (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI ) 1 1 f (s) dsI S−1 R (s, T )Z∂(G1∩CI ) S−1 L (p, T ) dpI g(p) f (s) dsI Z∂(G1∩CI ) f (s) dsI Z∂(G1∩CI ) f (s) dsI Z∂(G1∩CI ) f (s) dsI Z∂(G1∩CI ) S−1 R (s, T )p(p2 − 2s0p + s2)−1dpI g(p) S−1 L (p, T )p(p2 − 2s0p + s2)−1dpI g(p) sS−1 R (s, T )(p2 − 2s0p + s2)−1dpI g(p) sS−1 L (p, T )(p2 − 2s0p + s2)−1dpI g(p) where we have used the resolvent equation. But now observe that 1 (2π)2 Z∂(G2∩CI ) f (s) dsI Z∂(G1∩CI ) S−1 R (s, T )p(p2 − 2s0p + s2)−1dpI g(p) = 0 and − 1 (2π)2 Z∂(G2∩CI ) f (s) dsI Z∂(G1∩CI ) sS−1 R (s, T )(p2 − 2s0p + s2)−1dpI g(p) = 0. so it follows that f (T )g(T ) = − S−1 L (p, T )p(p2 − 2s0p + s2)−1dpI g(p) 1 (2π)2 Z∂(G2∩CI ) (2π)2 Z∂(G2∩CI ) f (s) dsI Z∂(G1∩CI ) f (s) dsI Z∂(G1∩CI ) 1 + sS−1 L (p, T )(p2 − 2s0p + s2)−1dpI g(p) which can be written as 1 f (T )g(T ) = (2π)2 Z∂(G2∩CI ) f (s) dsI Z∂(G1∩CI ) [sS−1 L (p, T ) − S−1 L (p, T )p]× × (p2 − 2s0p + s2)−1dpI g(p). Using Lemma 3.23 we get f (T )g(T ) = 1 2π Z∂(G1∩CI ) S−1 L (p, T )dpI f (p) g(p) which gives the statement. (cid:3) In the original proof of the above theorem we have used the fact that for functions f ∈ NσS (T ) the left S-resolvent equation gives f (T )T m = 1 2π Z∂(U ∩CI ) S−1 L (p, T )dpI f (p)pm from which one obtains f (T )T mt−1−m = 1 2π Z∂(U ∩CI ) S−1 L (p, T )dpI f (p)pmt−1−m. 20 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI By taking the sum and considering t ∈ ρS(T ), we have f (T )S−1 L (t, T ) = 1 2π Z∂(U ∩CI ) S−1 L (p, T )dpI f (p)S−1 L (t, p). Using this equality and the Cauchy formula we obtain the statement. 4. The case of commuting operators and the quaternionic case In this last section we state the resolvent equation in the case of commuting operators and for the quaternionic functional calculus. We also take the occasion to make some comments that show how the S-functional calculus turns out to be a natural extension of the Riesz-Dunford functional calculus. 4.1. The case of several commuting operators. We denote by BC0,1(Vn) the subset of B0,1(Vn) consisting of paravector operators with commuting components. Given an operator in paravector form T = T0 + e1T1 + . . . + enTn, its so-called conjugate T is defined by T = T0 − e1T1 − . . . − enTn. When T ∈ BC0,1(Vn) the operator T T is well defined and T T = T T = T 2 Theorem 4.1. Let T ∈ BC0,1(Vn) and s ∈ Rn+1 be such that s < kT k. Then n and T + T = 2T0. 0 + T 2 1 + . . . + T 2 T ms−1−m = (sI − ¯T )(s2I − s(T + T ) + T T )−1, s−1−mT m = (s2I − s(T + T ) + T T )−1(sI − ¯T ). Xm≥0 Xm≥0 (4.57) (4.58) The above theorem follows from the fact that the Cauchy kernels for slice hyperholo- morphic functions can be written in two possible ways, see Section 2 and [15]. In the case of commuting operators the two expressions are equivalent. The advantage of this approach is that one can work with the so called F-spectrum which is easier to compute than the S-spectrum. In fact it can be computed over a complex plane CI, taking s = u + Iv, and then extended to Rn. This is a consequence of the fact that the F-spectrum takes into account the commutativity of the operators Tj, j = 0, 1, ..., n. The F-spectrum is suggested by Theorem 4.1 and it is described below. Definition 4.2 (The F -spectrum and the F -resolvent sets). Let T ∈ BC0,1(Vn). We define the F -spectrum of T as: σF (T ) = {s ∈ Rn+1 : s2I − s(T + T ) + T T is not invertible }. The F -resolvent set of T is defined by ρF (T ) = Rn+1 \ σF (T ). The main properties of the F -spectrum are similar to those of the S-spectrum as it is proved in the next results: Theorem 4.3 (Structure of the F -spectrum). Let T ∈ BC0,1(Vn) and let p = p0 +p1I ∈ [p0 + p1I] ⊂ Rn+1 \ R, such that p ∈ σF (T ). Then all the elements of the (n − 1)-sphere [p0 + p1I] belong to σF (T ). Theorem 4.4 (Compactness of F -spectrum). Let T ∈ BC0,1(Vn). Then the F -spectrum σF (T ) is a compact nonempty set. Moreover σF (T ) is contained in {s ∈ Rn+1 : s ≤ kT k }. A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 21 The relation between the S-spectrum and the F-spectrum is contained in the following result: Proposition 4.5. Let T ∈ BC0,1(Vn) . Then σF (T ) = σS(T ). Definition 4.6. (The SC-resolvent operator) Let T ∈ BC0,1(V ) and s ∈ ρF (T ). We define the SC-resolvent operator as S−1 C,L(s, T ) := (sI − T )(s2I − s(T + T ) + T T )−1. C,R(s, T ) := (s2I − s(T + T ) + T T )−1(sI − T ). S−1 (4.59) (4.60) Theorem 4.7. Let T ∈ BC0,1(Vn) and s, p ∈ ρF (T ). Then S−1 S-resolvent equation C,L(s, T ) satisfies the left S−1 C,L(s, T )s − T S−1 C,L(s, T ) = I, (4.61) and S−1 C,R(s, T ) satisfies the right S-resolvent equation Moreover, for p 6∈ [s], it satisfies the resolvent equation sS−1 C,R(s, T ) − S−1 C,R(s, T )T = I. S−1 C,R(s, T )S−1 C,L(p, T ) = ((S−1 − s(S−1 C,R(s, T ) − S−1 C,R(s, T ) − S−1 C,L(p, T ))p C,L(p, T )))(p2 − 2s0p + s2)−1. which can be written as S−1 C,R(s, T )S−1 C,L(p, T ) = (s2 − 2p0s + p2)−1(s(S−1 C,L(p, T ))p). C,R(s, T ) − S−1 − (S−1 C,R(s, T ) − S−1 C,L(p, T )) (4.62) We conclude this subsection with a couple of considerations on the case of unbounded operators. (I) Suppose that T is a closed operator with domain D(T ). As one can clearly see, the non commutative version of S−1 L (s, T ), that is S−1 L (s, T ) := −(T 2 − 2Re(s)T + s2I)−1(T − sI), is defined on the domain of T , and not on Vn as it is in the classical case. So we have to consider the extension to Vn writing S−1 L (s, T ) as follows S−1 L (s, T ) := −(T (T 2 − 2Re(s)T + s2I)−1 − (T 2 − 2Re(s)T + s2I)−1sI). In this case S−1 L (s, T ) turns out to be defined on Vn. Observe now that if T is a closed operator with domain D(T ) and with commuting components the left S-resolvent op- erator S−1 C,L(s, T ) turns out to be already defined on Vn. For a more detailed discussion see the original papers [13, 15]. In the case of the right S-resolvent we have the opposite situation. With the above consideration the new resolvent equation remains the same also for unbounded operators. (II) The F-spectrum is also useful to defined the so called F-functional calculus, see [5, 16]. This calculus is defined using the Fueter-Sce-Qian mapping theorem in integral form, see [22, 31, 34]. It is a hyperholomorphic functional calculus in the spirit of A. McIntosh, B. Jefferies and their coauthors (see [27, 28, 30, 33], the monograph [26] and the references therein) who first used the theory of hyperholomorphic functions, see [10, 18, 20, 25], to define a hyperholomorphic functional calculus for n-tuples of operators. 22 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI 4.2. The quaternionic setting. What we have previously proved in the paper can be rephrased also for the quaternionic functional calculus. We simply point out that in this case slice hyperholomorphic functions are defined on an open set U ⊆ H and have values in the quaternions H. The resolvent operators are as in the introduction of this paper. Here it is important to consider right linear operators as well as left linear operators T . The possible formulations of the quaternionic functional calculus has been carried out in [13]. The resolvent equations in Theorem 3.8 hold in this setting where instead of the paravector operator T = T0 + T1e1 + . . . + Tnen we replace quaternionic operators. We finally mention for sake of completeness one more analogy with the classical case. As it is well known the Laplace transform of a semigroup etG where for simplicity we take a bounded operator G defined on a Banach space X is the resolvent operator (λI − G)−1. In the quaternionic case we have the analogue result for the two S-resolvent operators: Let T ∈ B(V ) and let s0 > kT k. Then the left S-resolvent operator S−1 L (s, T ) is given by S−1 L (s, T ) = Z +∞ 0 and S−1 R (s, T ) is given by S−1 R (s, T ) = Z +∞ 0 etT e−ts dt, e−tset T dt. We point out that the theory of the quaternionic evolution operators is developed in [12] where it is also studied the case in which the generator is unbounded. Recently the case of sectorial operators has been treated in [24]. References [1] D. Alpay, The Schur algorithm, reproducing kernel spaces and system theory, American Mathe- matical Society, Providence, RI, 2001. Translated from the 1998 French original by Stephen S. Wilson, Panoramas et Synth`eses. [2] D. Alpay, F. Colombo and I. Sabadini, Schur functions and their realizations in the slice hyper- holomorphic setting, Integral Equations and Operator Theory 72 (2012), 253-289. [3] D. Alpay, F. Colombo and I. Sabadini, Pontryagin de Branges-Rovnyak spaces of slice hyperholo- morphic functions, J. d'Analyse Math´ematique, 2013. [4] D. Alpay, F. Colombo and I. Sabadini, Krein-Langer factorization and related topics in the slice hyperholomorphic setting, J. of Geom. Anal., to appear. [5] D. Alpay, F. Colombo and I. Sabadini, On some notions of convergence for n-tuples of operators, Math. Meth. Appl. Sci. (2013/14), to appear. [6] D. Alpay, F. Colombo, I. Lewkowicz, and I. Sabadini. Realizations of slice hyperholomorphic generalized contractive and positive functions. arXiv:1310.1035. [7] D. Alpay, A. Dijksma, J. Rovnyak, and H. de Snoo. Schur functions, operator colligations, and reproducing kernel Pontryagin spaces, volume 96 of Operator theory: Advances and Applications. Birkhauser Verlag, Basel, 1997. [8] D. Alpay, Y. Peretz, Realizations for Schur upper triangular operators centered at an arbitrary point, Integral Equations Operator Theory 37 (2000), 251–323. [9] D. Alpay and V. Vinnikov, Finite dimensional de Branges spaces on Riemann surfaces. J. Funct. Anal., 189(2):283–324, 2002. [10] F. Brackx, R. Delanghe, F. Sommen, Clifford analysis. Research Notes in Mathematics, 76, Pitman, Advanced Publishing Program, Boston, MA, 1982. [11] F. Colombo, I. Sabadini, On some properties of the quaternionic functional calculus, J. Geom. Anal., 19 (2009), 601–627. A NEW RESOLVENT EQUATION FOR THE S-FUNCTIONAL CALCULUS 23 [12] F. Colombo, I. Sabadini, The quaternionic evolution operator, Adv. Math., 227 (2011), 1772– 1805. [13] F. Colombo, I. Sabadini, On the formulations of the quaternionic functional calculus, J. Geom. Phys., 60 (2010), 1490–1508. [14] F. Colombo, I. Sabadini, The Cauchy formula with s-monogenic kernel and a functional calculus for noncommuting operators, J. Math. Anal. Appl., 373 (2011), 655–679. [15] F. Colombo, I. Sabadini, The F-spectrum and the SC-functional calculus, Proc. Roy. Soc. Edin- burgh Sect. A, 142 (2012), 479–500. [16] F. Colombo, I. Sabadini, F. Sommen, The Fueter mapping theorem in integral form and the F -functional calculus, Math. Meth. Appl. Sci., 33 (2010), 2050–2066. [17] F. Colombo, I. Sabadini, D.C. Struppa, A new functional calculus for noncommuting operators, J. Funct. Anal., 254 (2008), 2255–2274. [18] F. Colombo, I. Sabadini, F. Sommen, D.C. Struppa, Analysis of Dirac Systems and Computational Algebra, Progress in Mathematical Physics, Vol. 39, Birkhauser, Boston, 2004. [19] F. Colombo, I. Sabadini, D. C. Struppa, Noncommutative Functional Calculus. Theory and Appli- cations of Slice Hyperholomorphic Functions, Progress in Mathematics, Birkhauser, Basel, 2011. [20] R. Delanghe, F. Sommen, V. Soucek, Clifford algebra and spinor-valued functions. A function theory for the Dirac operator. Mathematics and its Applications, 53. Kluwer Academic Publishers Group, Dordrecht, 1992. xviii+485 pp. [21] N. Dunford, J. Schwartz, Linear Operators, part I: General Theory , J. Wiley and Sons (1988). [22] R. Fueter, Analytische Funktionen einer Quaternionenvariablen, Comm. Math. Helv., 4 (1932), 9–20. [23] R. Ghiloni, V. Moretti, A. Perotti, Continuous slice functional calculus in quaternionic Hilbert spaces, Reviews in Mathematical Physics, Vol. 25, No. 4 (2013), pp.1350006–1–1350006–83. [24] R. Ghiloni, V. Recupero, Semigroups over real alternative *-algebras: generation theorems and spherical sectorial operators, Preprint 2013. [25] K. Gurlebeck, K. Habetha, W. Sprossig, Holomorphic Functions in the Plane and n-dimensional space, Birkhauser, Basel, 2008. [26] B. Jefferies, Spectral properties of noncommuting operators, Lecture Notes in Mathematics, 1843, Springer-Verlag, Berlin, 2004. [27] B. Jefferies, A. McIntosh, The Weyl calculus and Clifford analysis, Bull. Austral. Math. Soc., 57 (1998), 329–341. [28] B. Jefferies, A. McIntosh, J. Picton-Warlow, The monogenic functional calculus, Studia Math., 136 (1999), 99–119. [29] M.S. Livsic, N. Kravitski, A. Markus, and V. Vinnikov. Commuting nonselfadjoint operators and their applications to system theory. Kluwer, 1995. [30] A. McIntosh, A. Pryde, A functional calculus for several commuting operators, Indiana Univ. Math. J., 36 (1987), 421–439. [31] T. Qian, Generalization of Fueter's result to Rn+1, Rend. Mat. Acc. Lincei, 8 (1997), 111–117. [32] M. Reed, B. Simon, Functional Analysis, Methods of Modern Mathematical Physics, Academic Press; REV edition, 1980. [33] C. Li, A. McIntosh, T. Qian, Clifford algebras, Fourier transforms and singular convolution operators on Lipschitz surfaces, Rev. Mat. Iberoamericana 10 (1994), 665–721. [34] M. Sce, Osservazioni sulle serie di potenze nei moduli quadratici, Atti Acc. Lincei Rend. Fisica , 23 (1957), 220–225. [35] V. Vinnikov, Commuting nonselfadjoint operators and algebraic curves, volume 59 of Operator Theory: Advances and Applications, pages 348–371. Birkhauser Verlag, Basel, 1992. 24 D. ALPAY, F. COLOMBO, J. GANTNER, AND I. SABADINI (DA) Department of Mathematics, Ben-Gurion University of the Negev, Beer-Sheva 84105 Israel E-mail address: [email protected] (FC) Politecnico di Milano, Dipartimento di Matematica, Via E. Bonardi, 9, 20133 Milano, Italy E-mail address: [email protected] (JG) Vienna University of Technology, Institute for Analysis and Scientific Comput- ing, Wiedner Hauptstrasse 8 - 10, 1040 Wien, Austria E-mail address: [email protected] (IS) Politecnico di Milano, Dipartimento di Matematica, Via E. Bonardi, 9, 20133 Milano, Italy E-mail address: [email protected]
1908.07786
1
1908
2019-08-21T10:47:53
On the Banach lattice c_0
[ "math.FA" ]
We show that $c_0$ is not a projective Banach lattice, answering a question of B. de Pagter and A. Wickstead. On the other hand, we show that $c_0$ is complemented in the free Banach lattice generated by itself (seen as a Banach space). As a consequence, the free Banach lattice generated by $c_0$ is not projective.
math.FA
math
ON THE BANACH LATTICE c0 ANTONIO AVIL´ES, GONZALO MART´INEZ-CERVANTES, AND JOS´E DAVID RODR´IGUEZ ABELL ´AN Abstract. We show that c0 is not a projective Banach lattice, answering a question of B. de Pagter and A. Wickstead. On the other hand, we show that c0 is complemented in the free Banach lattice generated by itself (seen as a Banach space). As a consequence, the free Banach lattice generated by c0 is not projective. 9 1 0 2 g u A 1 2 ] . A F h t a m [ 1 v 6 8 7 7 0 . 8 0 9 1 : v i X r a 1. Introduction The purpose of this paper is to answer negatively Question 12.11 proposed by B. de Pagter and A. Wickstead in [3] (notice that this also answers negatively [3, Question 12.10]). We prove that c0, seen as a Banach lattice, is not projective. Moreover, we show that it can be (isometrically) embedded as a Banach lattice into the free Banach lattice generated by itself seen as a Banach space, which is denoted by F BL[c0]. This embedding composed with the natural quotient from F BL[c0] onto c0 gives the identity map on c0. Thus, c0 is complemented in F BL[c0]. As a consequence, we will obtain that F BL[c0] is not projective. The concepts of free and projective Banach lattices were introduced in [3]. If A is a set with no extra structure, the free Banach lattice generated by A, denoted by F BL(A), is a Banach lattice together with a bounded map u : A −→ F BL(A) having the following universal property: for every Banach lattice Y and every bounded map v : A −→ Y there is a unique Banach lattice homomorphism S : F BL(A) −→ Y such that S ◦ u = v and kSk = sup {kv(a)k : a ∈ A}. The same idea is applied by A. Avil´es, J. Rodr´ıguez and P. Tradacete to define the concept of the free Banach lattice generated by a Banach space E, F BL[E]. This is a Banach lattice together with a bounded operator u : E −→ F BL[E] such that for every Banach lattice Y and every bounded operator T : E −→ Y there is a unique Banach lattice homomorphism S : F BL[E] −→ Y such that S ◦ u = v and kSk = kT k. 2010 Mathematics Subject Classification. 46B43, 06BXX. Key words and phrases. c0; F BL[c0]; Banach lattice; Free Banach lattice; Projectivity. Authors supported by project MTM2017-86182-P (Government of Spain, AEI/FEDER, EU) and project 20797/PI/18 by Fundaci´on S´eneca, ACyT Regi´on de Murcia. Third author supported by FPI contract of Fundaci´on S´eneca, ACyT Regi´on de Murcia. 1 2 A. AVIL ´ES, G. MART´INEZ-CERVANTES, AND J.D. RODR´IGUEZ ABELL ´AN In [2] and [3], the corresponding authors show that both objects exist and are unique up to Banach lattices isometries. Moreover, A. Avil´es, J. Rodr´ıguez and P. Tradacete give an explicit description of them in [2]. Let A be a non-empty set. For x ∈ A, let δx : [−1, 1]A −→ [−1, 1] be the evaluation function given by δx(x∗) = x∗(x) for every x∗ ∈ [−1, 1]A, and for every f : [−1, 1]A −→ R define kf k = sup( n Xi=1 f (x∗ i ) : n ∈ N, x∗ 1, . . . , x∗ n ∈ [−1, 1]A, sup x∈A x∗ i (x) ≤ 1) . n Xi=1 The Banach lattice F BL(A) is the Banach lattice generated by the evaluation functions δx inside the Banach lattice of all functions f : [−1, 1]A −→ R with finite norm. The natural identification of A inside F BL(A) is given by the map u : A −→ F BL(A) where u(x) = δx. Since every function in F BL(A) is a uniform limit of such functions, they are all continuous (with respect to the product topology) and positively homogeneous, i.e. they commute with multiplication by positive scalars. Now, let E be a Banach space. For a function f : E∗ −→ R consider the norm kf kF BL[E] = sup( n Xi=1 f (x∗ i ) : n ∈ N, x∗ 1, . . . , x∗ n ∈ E∗, sup x∈BE n Xi=1 x∗ i (x) ≤ 1) . The Banach lattice F BL[E] is the closure of the vector lattice in RE ∗ generated by the evaluations δx : x∗ 7→ x∗(x) with x ∈ E. These evaluations form the natural copy of E inside F BL[E]. All the functions in F BL[E] are positively homogeneous and weak∗-continuous when restricted to the closed unit ball BE ∗. The notions of free and projective objects are closely related in the general theory of In the setting of Banach lattices, de Pagter and Wickstead [3] introduced categories. projectivity in the following form: Definition 1.1. A Banach lattice P is projective if whenever X is a Banach lattice, J a closed ideal in X and Q : X −→ X/J the quotient map, then for every Banach lattice homomorphism T : P −→ X/J and ε > 0, there is a Banach lattice homomorphism T : P −→ X such that T = Q ◦ T and k T k ≤ (1 + ε) kT k. Some examples of projective Banach lattices given in [3] include F BL(A), ℓ1, all finite dimensional Banach lattices and Banach lattices of the form C(K), where K is a compact neighborhood retract of Rn. They also prove that ℓ∞ and c are not projective. In this paper we will focus on c0 and F BL[c0]. ON THE BANACH LATTICE c0 3 2. Non-projectivity of c0 as a Banach lattice In this section we are going to prove that c0, seen as a Banach lattice, is not projective. We will use the following fact (see [1, Proposition 2.1]): Proposition 2.1. Let P be a projective Banach lattice, I an ideal of P and π : P −→ P/I the quotient map. The quotient P/I is projective if and only if for every ε > 0 there exists a Banach lattice homomorphism uε : P/I −→ P such that π ◦ uε = idP/I and kuεk ≤ 1 + ε. Let L = P + f in(ω) = Pf in(ω) \ {∅} be the set of the finite parts of ω without the empty set. For A ∈ L let us define the map χA : L −→ [−1, 1] given by χA(B) = 1 if B ⊂ A and χA(B) = 0 if B 6⊂ A. Let Φ : F BL(L) −→ c0 be the map given by Φ(f ) =(cid:0)f(cid:0)(χA({1}))A∈L(cid:1) , f(cid:0)(χA({2}))A∈L(cid:1) , . . .(cid:1) =(cid:0)f(cid:0)(χA({n}))A∈L(cid:1)(cid:1)n∈N for every f : [−1, 1]L −→ R ∈ F BL(L). Lemma 2.2. The map Φ : F BL(L) −→ c0 has the following properties: (1) Φ(δA) =Pi∈A ei ∈ c0 for every A ∈ L. f ∈ F BL(L). (2) Φ is a Banach lattice homomorphism that is well-defined, i.e., Φ(f ) ∈ c0 for every (3) Φ is surjective. Proof. The first assertion follows from the definition of Φ. For every n ∈ N let gn : L −→ [−1, 1] be the function gn = (χA({n}))A∈L . Then the sequence (gn)n∈N is pointwise convergent to zero, so f (gn) converges to zero and Φ(f ) ∈ c0. Since Φ preserves linear combinations, suprema, infima and kΦ(f )k ≤ kf k for every f ∈ F BL(L), we have that Φ is a Banach lattice homomorphism whose image is in c0. Let us prove property (3). Let x = (x1, x2, . . .) ∈ c0 and suppose, without loss of generality, that x ≥ 0. Fix a sequence of natural numbers (ni)i∈N such that xn1 ≥ xn for every n ∈ N and xni+1 ≥ xn for every n ∈ N \ {n1 . . . , ni}. Now, let Ai = {n1, . . . , ni} and λi = xni − xni+1 for every i = 1, 2, . . . For A ∈ L, if we put eA :=Pi∈A ei, we have that x =P∞ Xj=1 Φ(λjδAj ) = Φ(cid:0) λjΦ(δAj ) = Xj=1 ∞ Xj=1 x = ∞ j=1 λjeAj , and then, ∞ λjδAj(cid:1), 4 A. AVIL ´ES, G. MART´INEZ-CERVANTES, AND J.D. RODR´IGUEZ ABELL ´AN where the last element P∞ j=1 λjδAj is well-defined since P∞ norm one. j=1 λj < ∞ and each δAj has (cid:3) Thus, Φ is a quotient map. We are going to prove that there is no bounded Banach lattice homomorphism ϕ : c0 −→ F BL(L) such that Φ ◦ ϕ = idc0. This fact will be a consequence of the following Lemma: Lemma 2.3. Let A be an infinite set, (x∗ in F BL(A) with the following properties: n)n∈N a sequence in [−1, 1]A and (fn)n∈N a sequence (1) fn ≥ 0 for every n ∈ N; (2) fn(x∗ (3) For every finite set F ⊆ A there is a natural number n such that x∗ n) = 1 for every n ∈ N; nF = 0, i.e. the restriction of x∗ n to F is null. Then for every ε > 0 there is a subsequence (fnk)k∈N such that ≥ m − ε for every m ∈ N. m (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=1 fnk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Proof. Fix ε > 0 and fn1 := f1. Since the elements of F BL(A) are continuous with respect 1 such that f1(x∗) ≥ 1− ε to the product topology, there is a neighborhood U1 of x∗ 2 whenever x∗ ∈ U1. In particular, there is a finite set F1 ⊆ A such that f1(x∗) ≥ 1 − ε 2 whenever x∗F1 = x∗ 1F1. We recursively construct the subsequence (fnk)k∈N and the sequence of sets (Fk)k∈N. Sup- niF1∪F2∪...∪Fi−1 = pose that we have fn1, . . . , fnk and F1, . . . , Fk finite subsets of A such that x∗ 0 and fni(x∗) ≥ 1 − ε 2i whenever x∗Fi = x∗ niFi. Property (3) guarantees the existence of a number nk+1 ∈ N such that x∗ nk+1F1∪F2∪...∪Fk = 0. It follows from property (2) that there is a finite set Fk+1 ⊆ A such that fnk+1(x∗) ≥ 1 − ε 2k+1 whenever x∗Fk+1 = x∗ nk+1Fk+1. For each k ∈ N define y∗ k : A −→ [−1, 1] such that y∗ kFk = x∗ k(x) = 0 whenever 2k for every k ∈ N. On the other hand, if k < k′ and k′(x) = 0. It k) and therefore x∗ nk′ (x) = 0, so y∗ nkFk and y∗ x ∈ A \ Fk. Notice that fnk(y∗ y∗ k(x) 6= 0 then x ∈ Fk (by the definition of y∗ follows that y∗ k) ≥ 1 − ε k and y∗ k′ have disjoint supports. In particular, sup x∈A m Xk=1 y∗ k(x) ≤ 1. Thus, ON THE BANACH LATTICE c0 5 fnk(z∗ : n ∈ N, z∗ 1, . . . , z∗ n ∈ [−1, 1]A, sup x∈A n z∗ i (x) ≤ 1) ≥ m Xk=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) fnk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) m Xi=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) = sup( n Xk=1 Xi=1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Xk=1 ≥ m m i )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) i )(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) for every m ∈ N. fnk(y∗ (1) ≥ fnk (y∗ k) ≥ m Xk=1 m Xk=1(cid:16)1 − Xi=1 2k(cid:17) ≥ m − ε ε (cid:3) Theorem 2.4. The Banach lattice c0 is not projective. Proof. We argue by contradiction. Suppose c0 is projective. Since F BL(L) is projective, it follows from Proposition 2.1 and Lemma 2.2 the existence of a bounded Banach lattice homomorphism ϕ : c0 −→ F BL(L) such that Φ ◦ ϕ = idc0. Set fn := ϕ(en). Since ϕ is a Banach lattice homomorphism and each en is positive, we have that fn = ϕ(en) ≥ 0 for every n ∈ N. It follows from the equality Φ(fn) = (Φ ◦ ϕ)(en) = en and the definition of Φ that for every n ∈ N. Set x∗ F ⊆ L we have x∗ fn(cid:0)(χA({n}))A∈L(cid:1) = en(n) = 1 n = (χA({n}))A∈L for every n ∈ N. Notice that for every finite set n(S) = 0 for every S ∈ F whenever n /∈SS∈F S. Thus, Lemma 2.3 asserts that for every ε > 0 there is a subsequence (fnk)k∈N such that On the other hand, since ϕ is bounded, there is a constant C > 0 such that m (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) fnk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=1 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) enk!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ϕ m Xk=1 m Xk=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) fnk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≥ m − ε for every m ∈ N. m ≤ C(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xk=1 enk(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)∞ = C for every m ∈ N. Thus, m − ε ≤ C for every ε > 0 and every m ∈ N, which yields to a contradiction. (cid:3) 3. Complementability of c0 in F BL[c0] This section is devoted to the proof that c0 is lattice-embeddable in F BL[c0] as a Banach lattice, that is to say, there exist a Banach lattice homomorphism u : c0 −→ F BL[c0] and two constants K, M ≥ 0 such that K kxk∞ ≤ ku(x)kF BL[c0] ≤ M kxk∞ 6 A. AVIL ´ES, G. MART´INEZ-CERVANTES, AND J.D. RODR´IGUEZ ABELL ´AN for every x ∈ c0. Moreover, we will prove that c0 is complemented in F BL[c0]. By [4, Theorem 4.50] we know that the Banach lattice c0 is lattice-embeddable in a Banach lattice E if and only if there exists a disjoint sequence (xn)n∈N ∈ E+ (the positive cone of E) such that a) (xn)n∈N does not converge in norm to zero, and b) the sequence of partial sums of (xn)n∈N is norm bounded, i.e., there exists some M > 0 satisfying kPn i=1 xikE ≤ M for every n ∈ N. Thus, what we are going to do is to construct a sequence (fn)n∈N ∈ F BL[c0] with the desired properties. The following lemma will be very useful: Lemma 3.1. Let A be a set and f : [−1, 1]A −→ R a continuous and positively homogeneous function that depends on a finite amount of coordinates, i.e., there exists a finite subset A0 ⊂ A and f : [−1, 1]A0 −→ R such that f (x∗) = f (x∗A0). Then, f is in F BL(A). Proof. The function f : [−1, 1]A0 −→ R is continuous and positively homogeneous. By [3, Proposition 5.3], f is in F BL(A0). Let T : F BL(A0) −→ F BL(A) be the map given by T (g)(x∗) = g(x∗A0) for every g : [−1, 1]A0 −→ R and x∗ ∈ [−1, 1]A. Clearly, f = T ( f ), and then f is in F BL(A). (cid:3) Let (Nn)n∈N be a strictly increasing sequence of natural numbers. For r ∈ R let r+ = max {r, 0} be the positive part of r, and for every n ∈ N let fn : c∗ 0 −→ R be the map given by fn(x∗) = (x∗ n − Nn max {x∗ m : m < n})+ · Πm>ngnm(x∗) 1, x∗ for every x∗ = (x∗ 2, . . .) ∈ c∗ such that gnm(x∗) = 0 if Nmx∗ gnm( x∗ kx∗k) whenever x∗ 6= 0. 0 = ℓ1, where gnm : c∗ n ≤ x∗ m, gnm(x∗) = 1 if x∗ 0 −→ [0, 1] is any continuous function n and gnm(x∗) = m ≤ (Nm − 1)x∗ Let us see that (fn)n∈N is a disjoint sequence of positive elements which satisfies both properties a) and b): Lemma 3.2. fn ≥ 0 for every n ∈ N and fn ∧ fl = 0 for every n 6= l. ON THE BANACH LATTICE c0 7 Proof. The first assertion is clear. For the second one, suppose, for example, that n < l, and let x∗ ∈ c∗ In particular, x∗ n. Now, if fn(x∗) 6= 0, we have that gnm(x∗) 6= 0 for every m > n, and then, that x∗ 0 such that fl(x∗) 6= 0. We have that x∗ n for every m > n. Taking m = l we have a contradiction. l > Nl max {x∗ m < Nmx∗ m : m < l}. l > Nlx∗ Lemma 3.3. fn is in F BL[c0] for every n ∈ N. Proof. Fix n ∈ N. We are going to find a sequence of functions (hk)k∈N ∈ F BL[c0] such that limk→+∞ khk − fnkF BL[c0] = 0. Then, we will have that fn ∈ F BL[c0]. Let hk : c∗ 0 −→ R be the map given by (cid:3) hk(x∗) = (x∗ n − Nn max {x∗ m : m < n})+ · Πn<m≤n+k gnm(x∗) for every x∗ = (x∗ 1, x∗ 2, . . .) ∈ c∗ 0 = ℓ1. Let us see that hk ∈ F BL[c0] for every k ∈ N. 1 = y∗ 1, . . . , x∗ n+k = y∗ n+k. Let P : F BL(Bc0) −→ Rc∗ Notice that hk is continuous, positively homogeneous and satisfies that hk(x∗) = hk(y∗) whenever x∗ 0 be the map given by P (f )(x∗) = f (( x∗(y) kx∗k )y∈Bc0 ) · kx∗k if x∗ 6= 0 and P (f )(0) = f (0), where 0 denotes the identically zero function in the corresponding space. We have that P (F BL(Bc0)) ⊂ F BL[c0] because P maps the evaluation functions in F BL(Bc0) to the evaluations func- tions in F BL[c0], preserves linear combinations, the lattice structure and kP (f )kF BL[c0] ≤ kf kF BL(Bc0 ). Now, let hk : [−1, 1]Bc0 −→ R be the map given by hk(z∗) = hk((z∗(e1), . . . , z∗(en+k), 0, 0, . . .)) for every z∗ ∈ [−1, 1]Bc0 . Since hk is continuous, positively homogeneous and depends only on finitely many coordinates, by Lemma 3.1 we have that hk ∈ F BL(Bc0). It follows that P ( hk) = hk ∈ F BL[c0]. Now, by definition, we have that khk − fnkF BL[c0] = sup( l Xi=1 (hk − fn)(x∗ i ) : l ∈ N, x∗ 1, . . . , x∗ l ∈ Bℓ1, sup x∈Bc0 x∗ i (x) ≤ 1) . l Xi=1 Take x∗ 1, . . . , x∗ l ∈ Bℓ1, with x∗ i = (x∗ i1, x∗ i2, . . .) for every i = 1, . . . , l and such that (hk − fn)(x∗ i ) 6= 0 for every i = 1, . . . , l and supx∈Bc0Pl i=1 x∗ i (x) ≤ 1. Note that hk ≥ fn, so we can remove absolute values in the previous expression. Then, we have that 8 A. AVIL ´ES, G. MART´INEZ-CERVANTES, AND J.D. RODR´IGUEZ ABELL ´AN in − Nn max {x∗ im : m < n})+ · i ) − Πn<m gnm(x∗ i )(cid:3) in − Nn max {x∗ im : m < n})+ · l Xi=1 (hk − fn)(x∗ i ) = · = · = l l (x∗ (x∗ Xi=1 (cid:2)Πn<m≤n+k gnm(x∗ Xi=1 (cid:2)Πn<m≤n+k gnm(x∗ Xi=1 (x∗ l in − Nn max {x∗ im : m < n})+ · · Πn<m≤n+k gnm(x∗ ≤ l Xi=1 x∗ in(cid:2)1 − Πn+k<m gnm(x∗ i )(cid:3). i )(cid:2)1 − Πn+k<m gnm(x∗ i )(cid:3) i ) − Πn<m≤n+k gnm(x∗ i )Πn+k<m gnm(x∗ i )(cid:3) Since (hk − fn)(x∗ i ) 6= 0 for every i = 1, . . . , l, we have that x∗ i ) 6= 0. Thus, for every i there exists mi > n + k such that gnmi(x∗ 1 − Πn+k<m gnm(x∗ in > Nmi − 1. Since Nmi > Nn+k, this implies that x∗ Nn+k−1x∗ imi that is to say, in < x∗ x∗ in 6= 0 and also that i ) 6= 1, imi. 1 Thus, (hk − fn)(x∗ i ) ≤ l Xi=1 x∗ in < 1 Nn+k − 1 l Xi=1 x∗ imi. l Xi=1 Therefore, since limk→+∞ 1 Nn+k−1 = 0, the proof will follow from the following Claim. Claim. For every x∗ 1, x∗ 2, . . . , x∗ l ∈ Bℓ1 and every natural numbers m1, m2, . . . , ml ∈ N we have x∗ imi ≤ 1 l Xi=1 whenever supx∈Bc0Pl i=1 x∗ i (x) ≤ 1. Proof of the Claim. Fix m = max{mi : 1 ≤ i ≤ l}. We show first that l Xi=1 x∗ imi ≤ max( l Xi=1 (cid:12)(cid:12) m Xj=1 ε(j)x∗ ij(cid:12)(cid:12) : ε ∈ {−1, +1}m) . In fact, ON THE BANACH LATTICE c0 9 max( l Xi=1 (cid:12)(cid:12) m Xj=1 ε(j)x∗ ij(cid:12)(cid:12) : ε ∈ {−1, +1}m) ≥ = ≥ = l m 1 1 l m ε(mi)=ε ε(j)x∗ Xj=1 Xi=1 (cid:12)(cid:12) 2m Xε∈{−1,+1}m ij(cid:12)(cid:12) Xj=1 2m Xε∈{−1,+1} Xi=1 Xε∈{−1,+1}m (cid:12)(cid:12) Xi=1 2 Xε∈{−1,+1} 2m−1(cid:12)(cid:12) Xε∈{−1,+1}m Xi=1 imi. ε(mi)=ε x∗ 1 1 l l ε(j)x∗ ij(cid:12)(cid:12) m Xj=1 ε(j)x∗ ij(cid:12)(cid:12) Let ε ∈ {−1, +1}m be the function which gives the maximum above. We have that x∗ imi ≤ l Xi=1 m Xj=1 l Xi=1 (cid:12)(cid:12) ε(j)x∗ ij(cid:12)(cid:12) = x∗ i (x) ≤ 1, l Xi=1 taking, in the equality, x = (ε(1), ε(2), . . . , ε(m), 0, . . .) ∈ Bc0. i=1 fikF BL[c0] ≤ 1 for every n ∈ N. = sup( k Xj=1 n ( Xi=1 fi)(x∗ j ) : k ∈ N, x∗ 1, . . . , x∗ k ∈ Bℓ1, sup x∈Bc0 (cid:3) x∗ j (x) ≤ 1) . k Xj=1 Proof. By definition, we have that Lemma 3.4. kPn fi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 n Fix x∗ 1, x∗ 2, . . . , x∗ disjoint, for each j = 1, 2, . . . , k there is at most one ij ∈ {1, 2, . . . , n} such that fij (x∗ Thus, k ∈ Bℓ1 with supx∈Bc0Pk j=1 x∗ j (x) ≤ 1. Since the functions fi are j ) 6= 0. k n ( Xj=1 Xi=1 fi)(x∗ j ) = fij (x∗ j ). k Xj=1 Without loss of generality, we suppose that fij (x∗ j ) 6= 0 for every j = 1, 2, . . . , k. 10 A. AVIL ´ES, G. MART´INEZ-CERVANTES, AND J.D. RODR´IGUEZ ABELL ´AN Notice that each fij (x∗ j ) ≤ x∗ jij for every j = 1, 2, . . . , k, so k n ( Xj=1 Xi=1 fi)(x∗ j ) = k Xj=1 k fij (x∗ j ) ≤ x∗ jij . Xj=1 k asserts that Pk Now the Claim in Lemma 3.3 applied to x∗ 1, . . . , x∗ j=1 x∗ jij ≤ 1 and therefore i=1 fikF BL[c0] ≤ 1. Thus, kPn k n ( Xj=1 Xi=1 fi)(x∗ j ) ≤ k Xj=1 x∗ jij ≤ 1. (cid:3) Lemma 3.5. The sequence (fn)n∈N does not converge in norm to zero. Furthermore, kfnkF BL[c0] = 1 for every n ∈ N. Proof. We know that kfnkF BL[c0] = sup( k Xj=1 fn(x∗ j ) : k ∈ N, x∗ 1, . . . , x∗ k ∈ Bℓ1, sup x∈Bc0 x∗ j (x) ≤ 1) . k Xj=1 Taking e∗ n ∈ Bℓ1 we have that fn(e∗ n) = 1, so that kfnkF BL[c0] ≥ 1. In general, since fn(x∗) ≤ x∗(en) for every x∗ ∈ Bℓ1, we have that fn(x∗ j ) ≤ k Xj=1 k Xj=1 x∗ j (en) ≤ sup x∈Bc0 k Xj=1 x∗ j (x) ≤ 1, (cid:3) so kfnkF BL[c0] = 1. Thus, the Banach lattice c0 is lattice-embeddable into the free Banach lattice F BL[c0] [4, Theorem 4.50]. For the sake of completeness we include a proof that, indeed, this embedding has norm one: Theorem 3.6. The operator u : c0 −→ F BL[c0] given by u(x) = P+∞ x = (x1, x2, . . .) ∈ c0 is a Banach lattice embedding with kuk = 1. i=1 xifi for every Proof. It is enough to prove that n Xi=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) aifi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] = max 1≤i≤n ai ON THE BANACH LATTICE c0 11 for every n ∈ N and every a1, a2, . . . , an ∈ R. Fix any a1, a2, . . . , an ∈ R. Without loss of generality, suppose that a1 = max1≤i≤n ai. Since each function fi is positive and has norm one, we have that max 1≤i≤n n ai = a1 = ka1f1kF BL[c0] ≤(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 aifi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) = (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) n n (cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] aifi(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) Xi=1 aifi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] Xi=1 . On the other hand, n Xi=1 (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) aifi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] n =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 aifi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] n a1fi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] ≤ (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 fi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)F BL[c0] = a1(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi=1 ≤ a1 = max 1≤i≤n ai. n (cid:3) 1), f (e∗ The Banach lattice homomorphism T : F BL[c0] −→ c0 given by the formula T (f ) = (f (e∗ 0 −→ R ∈ F BL[c0] is surjective (notice that F LB[c0] contains a natural copy of c0 and T c0 is the identity map), so it is a quotient map. Since T ◦ u = idc0, we have proved the following theorem: 2), . . .) for every f : c∗ Theorem 3.7. The Banach lattice c0 is complemented in F BL[c0]. As a consequence of this fact, that kuk = 1 and that c0 is not projective, applying again Proposition 2.1, we conclude that: Theorem 3.8. F BL[c0] is not projective. References [1] A. Avil´es, J. D. Rodr´ıguez Abell´an, Projectivity of the free Banach lattice generated by a lattice, Archiv der Mathematik. To appear. [2] A. Avil´es, J. Rodr´ıguez, P. Tradacete, The free Banach lattice generated by a Banach space, J. Funct. Anal. 274 (2018), 2955 -- 2977. 12 A. AVIL ´ES, G. MART´INEZ-CERVANTES, AND J.D. RODR´IGUEZ ABELL ´AN [3] B. de Pagter, A. W. Wickstead, Free and projective Banach lattices, Proc. Royal Soc. Edinburgh Sect. A, 145 (2015), 105 -- 143. [4] C. D. Aliprantis, O. Burkinshaw, Positive Operators, Handbook of the Geometry of Banach Spaces, Springer 2006. Universidad de Murcia, Departamento de Matem´aticas, Campus de Espinardo 30100 Mur- cia, Spain. E-mail address: [email protected] Universidad de Murcia, Departamento de Matem´aticas, Campus de Espinardo 30100 Mur- cia, Spain. E-mail address: [email protected] Universidad de Murcia, Departamento de Matem´aticas, Campus de Espinardo 30100 Mur- cia, Spain. E-mail address: [email protected]
1008.3326
3
1008
2011-10-28T06:17:43
On typical properties of Hilbert space operators
[ "math.FA", "math.GN" ]
We study the typical behavior of bounded linear operators on infinite dimensional complex separable Hilbert spaces in the norm, strong-star, strong, weak polynomial and weak topologies. In particular, we investigate typical spectral properties, the problem of unitary equivalence of typical operators, and their embeddability into C_0-semigroups. Our results provide information on the applicability of Baire category methods in the theory of Hilbert space operators.
math.FA
math
On typical properties of Hilbert space operators KdV Institute for Mathematics, University of Amsterdam, P.O. Box 94248, 1090 GE Amsterdam, Tanja Eisner1 The Netherlands. Tam´as M´atrai∗ University of Toronto, Mathematics Department BA 6290, 40 St. George St., M5S 2E4 Toronto, Ontario, Canada. Abstract We study the typical behavior of bounded linear operators on infinite dimensional complex separable Hilbert spaces in the norm, strong-star, strong, weak polynomial and weak topologies. In particular, we investigate typical spectral properties, the problem of unitary equivalence of typical operators, and their embeddability into C0- semigroups. Our results provide information on the applicability of Baire category methods in the theory of Hilbert space operators. Keywords: Baire category, typical behavior, Hilbert space, norm topology, strong-star topology, strong topology, weak polynomial topology, weak topology, contraction, unitary operator, unitary equivalence, C0-semigroup 2000 MSC: 47A05, 47A10, 47A65, 47D03, 54A10, 54H05 ∗The author was partially supported by the OTKA grants K 61600, K 49786 and K 72655, by the NSERC grants 129977, A-7354, RGPIN 3185-10 and 402762, and by the NSF Grant DMS 0600940. Email addresses: [email protected] (Tanja Eisner), [email protected] (Tam´as M´atrai) URL: http://staff.science.uva.nl/~eisner/ (Tanja Eisner), http://www.renyi.hu/~matrait (Tam´as M´atrai) 1The author was supported by the European Social Fund and by the Ministry of Science, Re- search and the Arts Baden-Wurttemberg. Preprint submitted to Elsevier October 22, 2018 1. Introduction Given a property Φ on the points of a Baire space X, we say that a typical point of X satisfies Φ, or simply that Φ is typical, if the set {x ∈ X : x satisfies Φ} is co-meager in X, i.e. if {x ∈ X : x does not satisfy Φ} is of first category in X. Many important and classical results in analysis are concerned with typical properties in particular topological spaces. Examples include the Banach -- Mazurkiewicz theorem (see e.g. [2] and [23]) stating that the set of continuous nowhere differentiable func- tions are residual in (C([0, 1]),k · k∞) (see also [4] for a primer on typical properties of continuous functions), or the famous result by P. R. Halmos [16] and V. A. Rohlin [28] in ergodic theory on the existence of weakly mixing but not strongly mixing transformations. In this paper we continue the investigations of the first author and study the typical properties of contractive linear operators on infinite dimensional complex separable Hilbert spaces in the norm, strong-star, strong, weak polynomial and weak topologies (for the definitions, see Definition 2.1). Typical properties of various classes of operators have been studied previously: see e.g. [5] for typical properties of measure preserving transformations; [3] and [21] for typical mixing properties of Markov semigroups; [9], [12] and [13] for typical stability properties; [19], [27] and [30] for typical spectral properties in various very special families of operators. Our research is different from these works in several respects. We study typical proper- ties of contractions as a whole, and we carry out our analysis in several topologies. Surprisingly, in contrast to classical results, we mostly obtain "good" properties as being typical, and it turns out that the typical properties may change drastically if the reference topology is changed. We obtain the following results. In Section 3, we recall some results obtained by the first author (see [8], [10]) about typical properties in the weak topology. In this topology, a typical contraction is unitary, it has maximal spectrum and empty point spectrum, it can be embedded into a C0-semigroup, and typical contractions are not unitarily equivalent. Our results make use of the theory of typical properties of mea- sures developed by M. G. Nadkarni [24, Chapter 8] (see also [5]). The importance of the weak topology in operator theory is an obvious motivation for our investigations. In Section 4 we consider the weak polynomial topology. Our main observations are that the contractions endowed with this topology form a Polish space, where the set of unitary operators is a co-meager subset. Since on the set of unitary operators the weak and weak polynomial topologies coincide, we conclude that the typical properties of contractions in the weak and weak polynomial topologies coincide. This part of our work is motivated by the increasing interest in this unusual topology. Section 5 treats the strong topology. We show that a typical contraction is 2 unitarily equivalent to the infinite dimensional backward unilateral shift operator. An analogous result for strongly continuous semigroups is obtained as well. In particular, the point spectrum of a typical contraction is the open unit disk, typical contractions are unitarily equivalent and can be embedded into a C0-semigroup. Contrast these results with the behavior in the weak topology. Just as for the weak topology, our interest in the strong topology necessitates no clarification. We study the strong-star topology in Section 6. Our main result gives that the theory of typical properties of contractions in the strong-star topology can be reduced to the theories of typical properties of unitary and positive self-adjoint operators in the strong topology. As a corollary, we obtain that a typical contraction has maximal spectrum and empty point spectrum, and two typical contractions are not unitarily equivalent. We included the strong-star topology in our research because it may play the most important role while extending our investigations into general Banach⋆-algebras. Section 7 contains our results on the norm topology, the only non-separable topol- ogy we consider. We obtain that in this topology there are no such non-trivial typical structural properties as for the separable topologies. Intuitively, the reason for this phenomenon is that the norm topology is fine enough to allow for the coexistence of many different properties on non-meager sets. We close the paper with an outlook to typical properties of operators on general Banach spaces, and with a list of open problems. Before turning our attention to the proofs, let us justify our settings. We choose contractions as the underlying set of operators because it becomes a Baire space in all the five topologies we consider. By writing the set of bounded linear operators as a countable union of scaled copies of the set of contractions, suitable extensions of our result can be easily obtained. In our investigations of contractions, several other important classes of operators (e.g. isometries, positive self-adjoint operators, unitary operators) come into play, and we obtain information about typical properties in these subclasses, as well. We work only in infinite dimensional separable Hilbert spaces because removing any of these assumptions invalidates most of our results. Infinite dimensionality guarantees that the families of operators under consideration are sufficiently rich. Separability is essential for our descriptive set theoretic arguments. The Hilbert space structure not only allows us to use a well-developed spectral theory, but also facilitates the construction of operators using orthogonal decomposition. It is of limited importance that our Hilbert spaces are over the complex field; analogous results hold in real Hilbert spaces, as well. We defer the further discussion of possible extensions of our work until Section 8. 3 To conclude this introduction we note that in research works studying contrac- tions, it is customary to point out that contractions in general are hard to study. The theory of contractions as a whole is often contrasted to the theories of normal, self-adjoint or unitary operators where a satisfactory classification can be obtained, e.g. via spectral measures (see e.g. [24] and [33]). Our results provide an explana- tion of this intuitive observation. As we pointed out above, as far as Baire category methods are concerned, in the four separable topologies we consider, the theory of contractions is reduced to the theory of unitary operators (weak and weak poly- nomial topologies), to the theory of one shift operator (strong topology) or to the theories of unitary and positive self-adjoint operators (strong-star topology). Thus if the oversimplified pictures captured by these separable topologies are dissatisfactory for an analyst, then necessarily the very fine norm topology has to be used, in which case non-separability can be made responsible for being complicated. Since only such properties can be studied using Baire category arguments which are non-trivial in a suitable Baire topology, i.e. which hold at least on a non-meager set, our results outline a limitation of Baire category methods in operator theory. Acknowledgment. We thank Mariusz Lema´nczyk and Nicolas Monod for help- ful discussions and for calling our attention to the problem of unitary equivalence of typical operators, as well as the referee for careful reading and helpful suggestions. 2. Preliminaries As general references, see [20] for descriptive set theory, [33] for functional analy- sis, and [15] for semigroup theory. Recall that a topological space X is a Baire space if every non-empty open set in X is non-meager, or equivalently if the intersection of countably many dense open sets in X is dense (see e.g. [20, (8.1) Proposition and (8.2) Definition p. 41]). Polish spaces, i.e. separable complete metric spaces, are well-known examples of Baire spaces. A set in a topological space is Gδ if it can be obtained as an intersection of countably many open sets. We will often use the observation that in every topological space, every dense Gδ set is co-meager. Let X be a Baire space and let Φ be a property on the points of X. We say that Φ is a typical property on X, or that a typical element of X satisfies Φ if {x ∈ X : x satisfies Φ} is a co-meager subset of X. Note that if Φn (n ∈ N) are typical properties on X then a typical element of X satisfies all Φn (n ∈ N) simultaneously. In the sequel (H,k · k) always denotes an infinite dimensional complex sepa- rable Hilbert space. The scalar product on H is denoted by h·,·i. For every U ⊆ H, span{U} denotes the linear subspace of H generated by U, and U ⊥ = 4 {x ∈ H : hx, ui = 0 (u ∈ U)}. For every U, V ⊆ H, we write U⊥V if hu, vi = 0 (u ∈ U, v ∈ V ). If V ≤ H is a subspace, we define BV = {v ∈ V : kvk ≤ 1}, SV = {v ∈ V : kvk = 1}. Let B(H), C(H), U(H) and P (H) denote the sets of bounded, contractive, uni- tary and contractive positive self-adjoint linear H → H operators. The identity operator is denoted by Id. For every V ≤ H, the orthogonal projection onto V is denoted by PrV . For every A ∈ B(H) the adjoint of A is denoted by A⋆. For every A ∈ B(H) and U ⊆ H we set A[U] = {Ax : x ∈ U}, ker A = {x ∈ H : Ax = 0} and Ran A = A[H]. For every A ∈ B(H), let O(A) = {UAU −1 : U ∈ U(H)}. For every A ∈ B(H) the spectrum, the point spectrum, the continuous spectrum, and the residual spectrum of A is denoted by σ(A), Pσ(A), Cσ(A), and Rσ(A) (see e.g. [33, Definition p. 209]). We recall the definitions and some elementary properties of the weak, weak poly- nomial, strong and strong-star topologies (see e.g. [33, Definition 3 p. 112], [25, Definition p. 1142], [33, Definition p. 69], and [34, Definition p. 220]). Definition 2.1. Let A, An ∈ B(H) (n ∈ N) be arbitrary. 1. We say that {An : n ∈ N} converges to A weakly, A = w-limn∈N An in notation, if for every x, y ∈ H, limn∈NhAnx, yi = hAx, yi. The topology corresponding to this notion of convergence is called the weak topology. Topological notions referring to the weak topology are preceded by w-. 2. We say that {An : n ∈ N} converges to A weakly polynomially, A = pw- n = Ak. The topol- limn∈N An in notation, if for every k ∈ N, w-limn∈N Ak ogy corresponding to this notion of convergence is called the weak polynomial topology. Topological notions referring to the weak polynomial topology are preceded by pw-. 3. We say that {An : n ∈ N} converges to A strongly, A = s-limn∈N An in notation, if for every x ∈ H, limn∈N Anx = Ax. The topology corresponding to this notion of convergence is called the strong topology. Topological notions referring to the strong topology are preceded by s-. 4. We say that {An : n ∈ N} converges to A in the strong-star sense, A = s⋆- n = A⋆. The topology limn∈N An in notation, if s-limn∈N An = A and s-limn∈N A⋆ corresponding to this notion of convergence is called the strong-star topology. Topological notions referring to the strong-star topology are preceded by s⋆-. Proposition 2.2. ([32, Section 2 p. 67], [34, Section 7.f p. 121]) Let {ei : i ∈ N} ⊆ H be an orthonormal basis. 5 1. For every A, B ∈ B(H), set dw(A, B) = Xi,j∈N 2−i−jhAei, eji − hBei, eji. Then dw is a complete separable metric on C(H) which generates the weak topology. 2. For every A, B ∈ B(H), set ds(A, B) =Xi∈N 2−ikAei − Beik. Then ds is a complete separable metric on C(H) which generates the strong topology. The strong-star topology on uniformly bounded sets is generated by the metric ds⋆(A, B) = ds(A, B) + ds(A⋆, B⋆). It is easy to see that ds⋆ is a complete met- ric on C(H). We will see in Section 4 that C(H) endowed with the weak polynomial topology is a Polish space as well. Note that the weak, weak polynomial, strong, strong-star, and norm topologies all refine the preceding topologies in this list. We also need the following. Proposition 2.3. ([32, Remark 4.10 p. 84]) On U(H) the weak, weak polynomial, strong, and strong-star topologies coincide. With this topology, U(H) is a Polish space. 3. The weak topology In the weak topology, the theories of typical properties of contractions and of unitary operators coincide. Theorem 3.1. ([8, Theorem 2.2 p. 2]) A w-typical contraction is unitary. Equivalently, Theorem 3.1 says that U(H) is a w-co-meager subset of C(H). By Proposition 2.3, U(H) endowed with the weak topology is also a Polish space. Hence the notions related to Baire category make sense relative to U(H), and a set A ⊆ C(H) is w-co-meager in C(H) if and only if A∩ U(H) is w-co-meager in U(H). Unitary operators are well-understood. E.g. the theory of spectral measures al- lows a detailed description of the spectral properties and conjugacy classes of unitary 6 operators. We refer to [33, Chapter XI.4 p. 306] and [24, Chapter 2 p. 17] for an introduction to spectral measures. The following proposition briefly summarizes how the spectral properties of w-typical contractions can be obtained from the theory of spectral measures. We set S1 = {λ ∈ C : λ = 1}. For a measure µ on S1, supp µ denotes the closed support of µ. For measures µ, ν on S1, we write µ⊥ν if µ and ν are mutually singular. Proposition 3.2. A w-typical contraction U satisfies Cσ(U) = S1. Proof. Let U be a w-typical contraction. By Theorem 3.1, U is unitary. Let µU denote the maximal spectral type of U (see [24, Section 8.22 p. 55]), which is a Borel measure on S1. E.g. by the uniqueness of the measure class of µU , λ ∈ σ(U) if and only if λ ∈ supp µU , and λ ∈ Pσ(U) if and only if λ is an atom of µU . By [24, 8.25 Theorem (a) p. 56], {U ∈ U(H) : Pσ(U) = ∅} = {U ∈ U(H) : µU is atomless} is a w-co-meager set in U(H). By a similar argument, using [24, 7.7 Corollary p. 46], we obtain that {U ∈ U(H) : σ(U) = S1} = {U ∈ U(H) : supp µU = S1} is also a w-co-meager set in U(H). Since σ(U) = Pσ(U) ∪ Cσ(U) ∪ Rσ(U) and Rσ(U) = ∅ for every unitary operator, we conclude that a w-typical contraction U satisfies Cσ(U) = S1, as required.(cid:4) By analogous applications of spectral measures, one can isolate numerous addi- tional w-typical properties of contractions (see e.g. [24, 8.25 Theorem p. 56]). We refer to e.g. [11, Section IV.3] for asymptotic properties of w-typical contractions, and mention that P. Zorin [35] showed recently that a w-typical contraction admits a fixed cyclic vector. Here we restrict ourselves to pointing out that despite the abundance of w-typical properties, typical contractions are not unitarily equivalent. Proposition 3.3. For every U ∈ C(H), O(U) is w-meager in C(H). In particular, for a w-typical pair of contractions (U1, U2) ∈ C(H)× C(H) we have that U1 and U2 are not unitarily equivalent. Proof. Let U ∈ C(H) be arbitrary. By Theorem 3.1, the statement follows if U /∈ U(H). So for the rest of the argument, we can assume U ∈ U(H). By the conjugacy invariance of the measure class of the maximal spectral type, for every 7 V ∈ U(H) we have that µU and µV U V −1 are mutually absolutely continuous. By [24, 8.25 Theorem (b) p. 56], for every measure ν on S1, the set {V ∈ U(H) : µV ⊥ν} is w-co-meager in U(H). Thus O(U) is w-meager in U(H) and so in C(H), as well. Finally consider the set E = {(U1, U2) ∈ C(H) × C(H) : U1, U2 are unitarily equivalent}. The set E is clearly analytic hence has the Baire property (see e.g. [20, (29.14) Corollary p. 229]). As we have seen above, for every U1 ∈ C(H) we have {U2 ∈ C(H) : (U1, U2) ∈ E} is w-meager in C(H). So by the Kuratowski-Ulam theorem (see e.g. [20, (29.14) Corollary p. 229]) we get that E is w-meager in C(H) × C(H). This completes the proof.(cid:4) 4. The weak polynomial topology The main result of this section is the following. Theorem 4.1. The set C(H) endowed with the weak polynomial topology is a Polish space. Moreover, U(H) is a pw-co-meager pw-Gδ subset of C(H). This result immediately implies that the theories of pw-typical and w-typical properties of contractions coincide. Corollary 4.2. A set C ⊆ C(H) is pw-co-meager in C(H) if and only if C ∩ U(H) is w-co-meager in U(H). In particular, a property Φ of contractions is pw-typical if and only if Φ is w-typical. Proof. By Theorem 4.1, U(H) is a pw-co-meager subset of the Polish space C(H). By Proposition 2.3, the weak and weak polynomial topologies coincide on U(H) and in these topologies U(H) is a Polish space. So the notions related to Baire category make sense relative to U(H). By U(H) ⊆ C(H) being pw-co-meager, a set M ⊆ U(H) is pw-meager in U(H) if and only if M ⊆ C(H) is pw-meager in C(H). We obtained that C ⊆ C(H) is pw-co-meager in C(H) if and only if C ∩ U(H) ⊆ C(H) is pw-co-meager in C(H). This is equivalent to C ∩ U(H) ⊆ U(H) being pw- co-meager in U(H), which is the same as C ∩ U(H) ⊆ U(H) being w-co-meager in U(H). Finally by Theorem 3.1, this is equivalent to C ⊆ C(H) being w-co-meager in C(H). This completes the proof.(cid:4) By Corollary 4.2, for the pw-typical properties of contractions one can refer to Section 3. To prove the first part of Theorem 4.1 we need the following lemmas on the weak topology. 8 Lemma 4.3. Let A, An ∈ B(H) (n ∈ N) satisfy A = w-limn∈N An. If for every x ∈ SH we have kAxk ≥ lim supn∈N kAnxk then A = s-limn∈N An. Proof. For x ∈ SH we have kAx − Anxk2 = kAxk2 + kAnxk2 − 2RehAx, Anxi, so 0 ≤ lim supn∈N kAx − Anxk2 = kAxk2 + lim supn∈N kAnxk2− 2 limn∈N RehAx, Anxi ≤ 2kAxk2 − 2RehAx, Axi = 0. Since x ∈ SH was arbitrary, A = s-limn∈N An follows. (cid:4) Lemma 4.4. Let n > 0, {xi : i < n} ⊆ SH and B ∈ C(H) be arbitrary. Then for every ε > 0 there exists a w-open set W ⊆ C(H) such that B ∈ W and for every A ∈ W we have kAxik ≥ kBxik − ε (i < n). Proof. We prove the statement for n = 1 only; for n > 1 the required set W can be obtained by intersecting the sets Wi satisfying the conditions of the lemma for each xi (i < n) separately. If Bx0 = 0 the statement is trivial; so we can assume Bx0 6= 0. Consider the set W = {A ∈ C(H) : hAx0, Bx0i > kBx0k2 − kBx0kε}; then W is w-open and B ∈ W . For every A ∈ W we have ≥ kBx0k − ε kAx0k ≥ hAx0, Bx0i kBx0k as required. (cid:4) Proof of Theorem 4.1. By [8, Theorem 2.2], U(H) is a w-co-meager w-Gδ subset of C(H). Since the weak polynomial topology is finer than the weak topology, U(H) ⊆ C(H) is pw-Gδ, as well. By [25, Theorem 1 p. 1142], U(H) is pw-dense in C(H), thus U(H) is pw-co-meager in C(H). To prove that C(H) is Polish in the weak polynomial topology we use [20, (8.18) Theorem p. 45] stating that a non-empty, second countable topological space is Polish if and only if it is T1, regular and strong Choquet (for the definition, see [20, (8.14) Definition p. 44.]). Since (C(H), pw) is a metric space, it is second countable, T1 and regular. So in order to show it is Polish, we have to prove it is strong Choquet, i.e. that player II has a winning strategy in the strong Choquet game. 9 We define the strategy for player II as follows. As in Proposition 2.2, let dw denote the complete metric on C(H) which induces the weak topology on C(H). Let {xn : n ∈ N} be a dense subset of SH. Suppose the nth move of player I is (An, Un), where An ∈ Un ⊆ C(H) and Un is a pw-open set. Let Wn be a w-open set with the following properties: 1. An ∈ Wn ⊆ clw(Wn) ⊆ {A : dw(A, An) < 1/(n + 1)}; 2. clw(Wn) ⊆ Wn−1; 3. for every A ∈ Wn we have kAxik ≥ kAnxik − 1/(n + 1) (i ≤ n); such a set exists by Lemma 4.4. Then let player II respond by playing a pw-open set Vn ⊆ Un such that 4. An ∈ Vn ⊆ clpw(Vn) ⊆ Un ∩ Wn. We show that this strategy is winning for player II. Let {(An, Un), Vn : n ∈ N} be a run in the game in which player II follows the above strategy. By An+1 ∈ Vn ⊆ Wn (n ∈ N) and conditions 1 and 2, An is weakly convergent, say A = w- limn∈N An. Again by conditions 1 and 2, A ∈ Wn (n ∈ N); thus by condition 3, kAxik ≥ lim supn∈N kAnxik (i ∈ N). Since {xn : n ∈ N} ⊆ SH is dense and A, An (n ∈ N) are contractions, we get kAxk ≥ lim supn∈N kAnxk (x ∈ SH). Thus by in particular, A = pw-limn∈N An. Since Lemma 4.3, we have A = s-limn∈N An; A ∈ clpw(Vn) ⊆ Vn−1 for each n by condition 4, we get A ∈ Tn∈N Vn. This shows that the strategy is winning for player II, and finishes the proof. (cid:4) Note that, by the same argument as in the above proof, C(H) is Polish for any metrizable topology on C(H) which is finer than the weak topology and coarser than the strong one. In addition, for such topology U(H) is a co-meager Gδ subset of C(H) if and only if U(H) is dense in C(H) in this topology. 5. The strong topology Probably the most surprising observation in the present paper is that some typical properties of contractions in the strong and weak topologies are completely different. While typical contractions in the weak topology are not unitarily equivalent, typical contractions in the strong topology are unitarily equivalent to an infinite dimensional backward unilateral shift operator, hence the investigation of s-typical properties of contractions is reduced to the study of one particular operator. We introduce this operator in the following definition. Since infinite dimensional complex separable 10 Hilbert spaces are isometrically isomorphic, we can restrict ourselves to the study of a particular one. Definition 5.1. Set H = ℓ2(N × N) and denote the canonical orthonormal basis of H by {ei(n) : i, n ∈ N}. We define the infinite dimensional backward unilateral shift operator S ∈ C(H) by Se0(n) = 0 and Sei+1(n) = ei(n) (i, n ∈ N). The main result of this section is the following. Recall O(A) = {UAU −1 : U ∈ U(H)}. Theorem 5.2. The set O(S) is an s-co-meager subset of C(H). By Theorem 5.2, a property of contractions is s-typical if and only if S has this property. In the following corollary, we recall only the properties of S we usually concern in this paper. Corollary 5.3. An s-typical contraction A satisfies that 1. Pσ(A) = {λ ∈ C : λ < 1}, and for every λ ∈ Pσ(A) we have dim ker(λ · Id − A) = ∞; 2. Cσ(A) = S1; 3. A can be embedded into a strongly continuous semigroup. Moreover, typical contractions are unitarily equivalent. Proof. Statements 1 and 2 follow from Theorem 5.2 and the results of [29], while 3 is a corollary of [8, Proposition 4.3]. (cid:4) Our strategy to prove Theorem 5.2 is the following. The main observation is that for an s-typical contraction A, its adjoint A⋆ is an isometry. Then by the Wold decomposition theorem (see e.g. [31, Theorem 1.1 p. 3]), A is unitarily equivalent to a direct sum of unitary and backward unilateral shift operators, and the number of shifts in the direct sum depends on the dimension of ker A. Since an s-typical contraction A is strongly stable, the unitary part is trivial. So we complete the proof by showing dim ker A = ∞. 11 5.1. Elementary observations We collect here some elementary results we need later in our analysis. Lemma 5.4. Let x, y ∈ SH satisfy x 6= −y. Set α = (2 + 2 · Rehx, yi)−1/2. Then kα(x + y)k = 1. Proof. We have kα(x+y)k2 = α2(kxk2+kyk2+2·Rehx, yi) = α2(2+2·Rehx, yi) = 1, as required.(cid:4) Lemma 5.5. Let A ∈ C(H) and let (bn)n∈N ⊆ SH , z ∈ SH satisfy limn∈N Abn = z. Then (bn)n∈N is convergent. Proof. It is enough to prove that for every ε > 0 there is an N ∈ N such that for every n, m ≥ N we have kbn − bmk2 ≤ ε. So let ε > 0 be arbitrary. For every n, m sufficiently large we have bn 6= −bm, so αn,m = (2 + 2 · Rehbn, bmi)−1/2 is defined. By Lemma 5.4 we have kαn,m(bn + bm)k = 1. So since A is a contraction, αn,m · (2 − kAbn − zk − kAbm − zk) ≤ αn,m · k2 · z + (Abn − z) + (Abm − z)k = (1) Let N ∈ N be such that for every n ≥ N we have kAbn − zk ≤ ε/8. Then by (1), for every n, m ≥ N we get αn,m ≤ 1/(2− 2ε/8), i.e. 2 + 2· Rehbn, bmi ≥ 4· (1− ε/8)2. Thus αn,m · kAbn + Abmk = kαn,m · A(bn + bm)k ≤ 1. kbn − bmk2 = 2 − 2 · Rehbn, bmi ≤ 4 − 4 · (1 − ε/8)2 ≤ ε, as required.(cid:4) The following lemma will be helpful to show that the kernel of a typical contrac- tion is infinite dimensional. Lemma 5.6. Let n ∈ N \ {0} and let {ei : i < n} ⊆ H be an orthonormal family. Let {fi : i < n} ⊆ H satisfy kfi − eik < 1/n (i < n). Then {fi : i < n} are linearly independent. Proof. Let αi ∈ C (i < n) be arbitrary satisfying Pi<n αi > 0. We have (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi<n αiei −Xi<n ≤ Xi<n αifi(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 2 2 < αikei − fik!2 αi!2 1 nXi<n ≤ 1 nXi<n αi2 = . 1 n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi<n αiei(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) 12 So by (cid:13)(cid:13)Pi<n αiei(cid:13)(cid:13) > 0 we have (cid:13)(cid:13)Pi<n αifi(cid:13)(cid:13) > 0, as required.(cid:4) Next we point out a trivial sufficient condition for the direct sum of contractions to be a contraction. Lemma 5.7. Let V0, V1 ≤ H be subspaces satisfying V0⊥V1. Let Ai : Vi → H (i < 2) be contractive linear operators such that A0[V0]⊥A1[V1]. Then A : span{V0, V1} → H, A(αv0 + βv1) = αA0v0 + βA1v1 (α, β ∈ C) is also contractive. Proof. Let v ∈ span{V0, V1}, kvk = 1 be arbitrary. Then there are α, β ∈ C with α2 + β2 = 1 and vi ∈ SVi (i < 2) such that v = αv0 + βv1. We have kAvk2 = kA(αv0 + βv1)k2 = kαA0v0 + βA1v1k2 = α2kA0v0k2 + β2kA1v1k2 ≤ α2 + β2 = 1, so the proof is complete.(cid:4) Finally we point out that strongly stable contractions form a s-co-meager s-Gδ subset of C(H). Definition 5.8. A contraction A is strongly stable if s-limn∈N An = 0, i.e., for every x ∈ SH and ε > 0 there is an n ∈ N such that kAnxk < ε. The set of strongly stable contractions is denoted by S. Lemma 5.9. The set of strongly stable contractions is an s-co-meager s-Gδ subset of C(H). Proof. By A = limn∈N(1 − 2−n)A (A ∈ C(H)), the set of contractions A satisfying kAk < 1 is a norm dense and hence an s-dense subset of C(H). Since every such operator is strongly stable, it remains to show that S is s-Gδ. To this end, let {xi : i ∈ N} be a dense subset of SH . Note that for each x ∈ H and A ∈ S, the sequence kAnxk (n ∈ N) is monotonically decreasing, so limn∈N kAnxk = 0 is equivalent to inf n∈N kAnxk = 0. Thus which completes the proof.(cid:4) S =Ti,j∈NSn∈N{A ∈ C(H) : kAnxik < 2−j}, (2) We remark that Lemma 5.9 holds in every separable Banach space. It is interest- ing to note the difference to the weak operator topology in which the set of weakly stable contractions is w-meager in C(H) (see [12, Theorem 4.3]). 13 5.2. Mapping properties of s-typical contractions The purpose of this section is to prove the following. Theorem 5.10. Let G denote the set of contractive operators A satisfying the fol- lowing properties: 1. for every y ∈ SH there exists an x ∈ SH such that Ax = y, i.e., SH ⊆ A[SH ]; 2. dim Ker A = ∞. Then G is an s-co-meager subset of C(H). To prove Theorem 5.10, we need a geometric lemma saying that every contraction defined on a finite dimensional subspace of H can be extended to a contraction which is surjective in a very strong sense. Lemma 5.11. Let V ≤ H be a finite dimensional subspace and let A : V → H be a contractive linear operator. Let W = A[V ] and let Y ≤ H be an arbitrary subspace satisfying Y ⊥W . Then for every X ≤ H satisfying X⊥V and dim X = dim W + dim Y there exists a contractive linear operator A : span{V, X} → H such that AV = A and A[Bspan{V,X}] = Bspan{W,Y }. Proof. We handle first the special Y = {0} case. Let dim V = n and dim W = m. The Gram-Schmidt orthogonalization theorem states that there exists an orthonor- mal base {vi : i < n} ⊆ V such that {Avi : i < n} ⊆ W are pairwise orthogonal, and (i) kAv0k = max{kAvk : v ∈ SV }; (ii) for every j < n − 1, kAvj+1k = max{kAvk : v ∈ SV ∩ span{vi : i ≤ j}⊥}. By (i) and (ii) we have kAvik > 0 if and only if i < m. Set wi = Avi/kAvik (i < m). Let X ≤ H satisfy X⊥V and dim X = m. Fix an orthonormal base {xi : i < m} in X and define A : span{V, X} → H such that AV = A and Axi =p1 − kAvik2wi First we show A is a contraction. Let u ∈ span{V, X} satisfy kuk = 1. Then u =Pi<n αivi +Pi<m βixi where αi ∈ C (i < n), βi ∈ C (i < m) satisfyPi<n αi2 + (i < m). 14 x =Xi<m kxk =Xi<m(cid:16)β2 βi(kAvikvi +p1 − kAvik2xi). ip1 − kAvik2 i kAvik2 + β2 2(cid:17) ≤ 1, Then and Ax =Xi<m 2 = αivi +Xi<m Pi<m βi2 = 1. Then k Auk2 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) βixi!(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) A Xi<n (αikAvikwi + βip1 − kAvik2wi)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:13)(cid:13)(cid:13)(cid:13)(cid:13) Xi<m (αikAvik + βip1 − kAvik2)2 By the Cauchy-Schwarz inequality, for every 0 ≤ p, q, r ≤ 1 we have (pr+q√1 − r2)2 ≤ p2 + q2. So k Auk ≤Pi<m αi2 + βi2 ≤ 1, as required. Next we show that for every y ∈ BW there is an x ∈ Bspan{V,X} such that Ax = y. Let y =Pi<m βiwi where βi ∈ C (i < m) satisfy Pi<m βi2 ≤ 1. Let ≤Xi<m 2 βi(kAvikAvi +p1 − kAvik2 Axi) = Xi<m βi(kAvikkAvikwi +p1 − kAvik2p1 − kAvik2wi) = y, as required. This completes the proof of the special Y = {0} case. In the general case write X = X0 ⊕ X1 where X0⊥X1 and dim X0 = dim W , dim X1 = dim Y . By the special case above, there is a contraction A : span{V, X0} → H such that AV = A and A[Bspan{V,X0}] = BW . Extend further A by setting AX1 : X1 → Y be any isometric isomorphism. By Lemma 5.7, A is a contraction which clearly satisfies A[Bspan{V,X}] = Bspan{W,Y }. This completes the proof.(cid:4) From Lemma 5.11, we immediately get the following. Proposition 5.12. The set of contractive operators A such that for every y ∈ SH there exists an x ∈ SH such that Ax = y is an s-co-meager subset of C(H). 15 Proof. Let M = {A ∈ C(H) : ∀ε > 0, y ∈ SH ∃x ∈ SH (ky − Axk < ε)}. First we show that M is an s-dense s-Gδ subset of C(H). Fix y ∈ SH and ε > 0. The set C(y, ε) = {A ∈ C(H) : ∃x ∈ SH (ky − Axk < ε)} is s-open. We show that it is s-dense. Let U ⊆ C(H) be any non-empty s-open set. By passing to a subset, we can assume U = {A ∈ C(H) : kyi − Axik < εi (i ∈ I)} where xi, yi ∈ H, εi > 0 (i ∈ I) and I is finite. Let V = span{xi : i ∈ I} and take an arbitrary A ∈ U. By restricting A to V we can assume AV ⊥ = 0. Set W = A[V ]. Let Y ≤ H be an at most one dimensional subspace such that Y ⊥W and y ∈ span{W, Y }. Let X ≤ H be a dim W + dim Y dimensional subspace such that X⊥V . By Lemma 5.11, there exists a contraction A : span{V, X} → H such that AV = AV and A[Bspan{V,X}] = Bspan{W,Y }. In particular, there is an x ∈ Bspan{V,X} such that Ax = y. Since A is a contraction and kyk = 1, we get x ∈ SH. Extend further A by setting Aspan{V,X}⊥ = 0. Then A ∈ U ∩ C(y, ε); i.e. we concluded that C(y, ε) is s-dense. Let D ⊆ SH be a countable dense set. We have M =T{C(y, 2−n) : y ∈ D, n ∈ N}, so by the Baire category theorem, M is an s-dense s-Gδ subset of C(H). It now suffices to show that every A ∈ M satisfies SH ⊆ A[SH ]. To this end, let A ∈ M and z ∈ SH be arbitrary. By the definition of M, there is a sequence (bn)n∈N ⊆ SH such that limn∈N Abn = z. By Proposition 5.5, (bn)n∈N is convergent, say limn∈N bn = x. Then Ax = z, which completes the proof. (cid:4) To prove that a typical contraction has infinite dimensional kernel, we need a lemma showing that a typical contraction approximates the zero operator on arbi- trarily large finite dimensional subspaces. Lemma 5.13. The set of contractive operators A such that for every n ∈ N and ε > 0 there exists Z ≤ H with dim Z ≥ n and kAZk < ε is an s-dense s-Gδ subset of C(H). Proof. For every n ∈ N and ε > 0, the set C(n, ε) = {A ∈ C(H) : ∃Z ≤ H (dim Z ≥ n, kAZk < ε)} 16 is s-open. We show that C(n, ε) (n ∈ N, ε > 0) are s-dense. Fix arbitrary n ∈ N and ε > 0. Let U ⊆ C(H) be any non-empty s-open set. By passing to a subset, we can assume U = {A ∈ C(H) : kyi − Axik < εi (i ∈ I)} where xi, yi ∈ H, εi > 0 (i ∈ I) and I is finite. Let A ∈ U be arbitrary. Let V = span{xi : i ∈ I}, and define B ∈ C(H) by BV = AV , BV ⊥ = 0. Since dim V ⊥ = ∞, we obtained B ∈ C(n, ε) ∩ U; i.e. we concluded C(n, ε) is s-dense. The set of contractive operators A such that for every n ∈ N and ε > 0 there exists Z ≤ H with dim Z ≥ n and kAZk < ε is So by the Baire category theorem, this is an s-dense s-Gδ subset of C(H), which completes the proof.(cid:4) T{C(n, 2−m) : n, m ∈ N}. We are ready to prove the second part of Theorem 5.10. Proposition 5.14. The set of contractions A which satisfy dim ker A = ∞ is an s-co-meager subset of C(H). Proof. By Lemma 5.13 and Proposition 5.12, the set of contractions A which satisfy that 1. for every n ∈ N and ε > 0 there exists Z ≤ H with dim Z ≥ n and kAZk < ε/n; 2. for every y ∈ SH there exists an x ∈ SH such that Ax = y; is an s-co-meager subset of C(H). We show that every member A of this set satisfies dim ker A = ∞; this will complete the proof. Fix an arbitrary n ∈ N \ {0}. Set ε = 1 and let Z ≤ H satisfy 1 for this n and ε. Let {ei : i < n} ⊆ Z be an orthonormal family. We find fi ∈ H (i < n) such that kfi − eik < 1/n (i < n) and Afi = 0. Then by Lemma 5.6, {fi : i < n} are linearly independent. Then dim ker A ≥ n, and since n was arbitrary, we concluded dim ker A = ∞. Fix i < n; we define fi as follows. If Aei = 0, set fi = ei. Else observe that by 2, there exists xi ∈ SH with Axi = Aei/kAeik. Define now fi = ei − kAeikxi. Then kfi − eik = kAeik < 1/n and we have Afi = Aei − kAeikAxi = 0, so the construction is complete.(cid:4) Proof of Theorem 5.10. The statement follows from Proposition 5.12 and Proposition 5.14.(cid:4) 17 5.3. Unitary equivalence of s-typical contractions to a shift Operators in the set G introduced in Theorem 5.10 have the following properties. Proposition 5.15. Let A ∈ G. Then AA⋆ = Id and A⋆A is the projection onto Ran A⋆, which is an infinite dimensional and infinite co-dimensional subspace of H. In particular, A⋆ is an isometry, hence A is a co-isometry and in addition, A is an isometry on (Ker A)⊥. Proof. Let {ei : i ∈ N} be an orthonormal basis of H. By the definition of G, for every i ∈ N there is an ai ∈ SH such that Aai = ei. Note that for every i ∈ N, 1 = hei, eii = hAai, Aaii = hA⋆Aai, aii. By A, A⋆ being contractions, this is possible only if A⋆ei = A⋆Aai = ai (i ∈ N), thus AA⋆ei = Aai = ei (i ∈ N). This proves that AA⋆ = Id, Ran A⋆ = span{A⋆ei : i ∈ N} = span{ai : i ∈ N}, and that A⋆A is the projection onto Ran A⋆. Again by A, A⋆ ∈ C(H), this implies that A⋆ is an isometry, and A is an isometry on Ran A⋆. By the definition of G, Ran A⋆ = (Ker A)⊥ is infinite dimensional and infinite co-dimensional. This completes the proof. (cid:4) Proof of Theorem 5.2. By Lemma 5.9 and Theorem 5.10, it is enough to show that every A ∈ S ∩ G is unitarily equivalent to the operator S of Definition 5.1. By Proposition 5.15, A is a co-isometry. So by the Wold decomposition theorem (see e.g. [31, Theorem 1.1 p. 3]), we have H = Hu ⊕ Hs such that AHu is unitary and AHs is unitarily equivalent to the backward unilateral shift operator on l2(N, ker A), i.e. the Hilbert space of square summable N → ker A functions. By A ∈ S we have Hu = {0}. By Theorem 5.10.2, dim ker A = ∞ hence l2(N, ker A) is isometrically isomorphic to ℓ2(N × N). This completes the proof.(cid:4) 5.4. The continuous case In this section we show that a typical strongly continuous contraction semigroup on H is unitarily equivalent to an infinite dimensional backward unilateral shift semigroup. Let C c(H) denote the set of contractive C0-semigroups on H. Here we endow C c(H) with the topology induced by the uniform strong convergence on compact time intervals, i.e., by the metric dc s(T (·), S(·)) =Pj,n∈N 2−(j+n) supt∈[0,n] kT (t)ej − S(t)ejk, where {ej : j ∈ N} is an orthonormal basis of H. With respect to this topology, C c(H) is a Polish space. With an abuse of notation, topological notions referring to this topology are also preceded by s-. 18 We again restrict ourselves without loss of generality to a particular infinite di- mensional complex separable Hilbert space and introduce on it the infinite dimen- sional backward unilateral shift semigroup. Definition 5.16. Set H = L2(R2 unilateral shift semigroup S(·) ∈ C c(H) by [S(t)f ](s, w) = f (s + t, w) (t, s, w ∈ R+, f ∈ H). +). We define the infinite dimensional backward We also set O(S(·)) = {(U −1S(t)U)t∈R+ : U ∈ U(H)} ⊆ C c(H). The following shows that an s-typical contraction semigroup is unitarily equiva- lent to S(·). Theorem 5.17. The set O(S(·)) is an s-co-meager subset of C c(H). Note that for every t > 0, S(t) is unitarily equivalent to the backward unilateral shift operator S of Definition 5.1. So by Theorem 5.17, s-typical contractive C0- semigroups T (·) satisfy that for every t > 0, the operator T (t) has the same properties as S. Moreover, s-typical contraction semigroups are unitarily equivalent. To prove Theorem 5.17, we need to introduce the following not so well-known concept from semigroup theory. Definition 5.18. Let T (·) be a C0-semigroup on H. Let the generator A of T (·) satisfy 1 ∈ ρ(A). Then the operator V ∈ B(H) defined by V = (A + Id)(A − Id)−1 = Id + 2(A − Id)−1 is called the cogenerator of T (·). The cogenerator is a bounded operator which determines the semigroup uniquely. Moreover, it shares many properties of T (·) such as being contractive, unitary, self- adjoint, normal, isometric, strongly stable etc. (see [31, Section III.8-9] for details). We will use the following (see e.g. [31, Theorem III.8.1]). Lemma 5.19. With the notation of Definition 5.18, V is the cogenerator of a con- traction semigroup if and only if V is contractive and satisfies 1 /∈ Pσ(V ). Our key observation is the following. 19 Lemma 5.20. Set V = {V ∈ C(H) : 1 /∈ Pσ(V )}, endowed with the strong topology. Define J : C c(H) → V by J(T (·)) = V where V is the cogenerator of T (·) (T (·) ∈ C c(H)). Then J is a homeomorphism. Proof. The statement follows from Lemma 5.19 and the First Trotter -- Kato Theo- rem (see e.g. [15, Theorem III.4.8]).(cid:4) Proof of Theorem 5.17. By Corollary 5.3, V is s-co-meager in C(H). So with the notation of (2) and Theorem 5.10, by Lemma 5.9 and Theorem 5.10 we have that S ∩ G ∩ V is s-co-meager in V. Hence the set J −1(S ∩ G ∩ V) is s-co-meager in C c(H). So it is enough to show that J −1(S ∩ G ∩ V) ⊆ O(S(·)). Let T (·) ∈ J −1(S∩G∩V) be arbitrary. Let V denote the cogenerator of T (·); then V ∈ S ∩ G ∩ V. By Proposition 5.15, V is a strongly stable co-isometry. Since the cogenerator of T ⋆(·) is V ⋆, and since the semigroup and the cogenerator share strong stability and the isometric property (see [31, Theorem III.9.1]), T (·) is a strongly stable co-isometric contraction semigroup. By Wold's decomposition for semigroups (see [31, Theorem III.9.3]), T (·) is unitarily equivalent to the backward unilateral shift semigroup on L2(R+, Y ), where Y = (Ran V ⋆)⊥. Since (Ran V ⋆)⊥ = Ker V is infinite dimensional, the statement follows. (cid:4) 6. The strong-star topology As we have seen in the previous section, the theory of s-typical properties of contractions is reduced to the study of one particular operator. The reason behind this phenomenon is that s-convergence does not control the adjoint, i.e. the function A 7→ A⋆ is not s-continuous. A straightforward remedy to this problem is to refine the strong topology such that taking adjoint becomes a continuous operation. This naturally leads to the investigation of the strong-star topology. As one may expect, the structure of an s⋆-typical contraction is more complicated than the structure of an s-typical contraction. We show that the theory of s⋆-typical properties of contractions can be reduced to the theories of typical properties of unitary and positive self-adjoint operators in the better understood strong topology. Theorem 6.1. There exist an s⋆-co-meager s⋆-Gδ set H ⊆ C(H) and an s-co- meager s-Gδ set P ⊆ P (H) such that the function Ψ : U(H) × P → H, Ψ(U, P ) = U · P is a homeomorphism, where U(H) and P are endowed with the strong topology and H is endowed with the strong-star topology. Moreover, if (ψ0, ψ1) : H → U(H) × P denotes the inverse of Ψ, then for every A ∈ H and U ∈ U(H) we have UAU −1 ∈ H and ψi(UAU −1) = Uψi(A)U −1 (i < 2). 20 Note that Theorem 6.1 immediately implies the following. Recall O(A) = {UAU −1 : U ∈ U(H)}. Corollary 6.2. For every A ∈ C(H), O(A) is an s⋆-meager subset of C(H). In particular, s⋆-typical contractions are not unitarily equivalent. Proof. We follow the notation of Theorem 6.1. Let A ∈ C(H) be arbitrary. If A /∈ H then by the unitary invariance of H we have O(A) ⊆ C(H) \ H, i.e. O(A) is s⋆-meager. So we can assume A ∈ H. We have Ψ−1(O(A)) = {(Uψ0(A)U −1, Uψ1(A)U −1) : U ∈ U(H)} ⊆ {Uψ0(A)U −1 : U ∈ U(H)} × {Uψ1(A)U −1 : U ∈ U(H)} = O(ψ0(A)) × O(ψ1(A)). By Proposition 3.3, O(ψ0(A)) ⊆ U(H) is s-meager. Hence Ψ−1(O(A)) is s×s-meager in U(H) × P. By Ψ being a homeomorphism, we obtain that O(A) is s⋆-meager in H and so in C(H), as well. The further corollary that unitarily equivalent pairs are s⋆ × s⋆-meager in C(H) × C(H) follows as in the proof of Proposition 3.3.(cid:4) Similarly to our approach in the previous section, for the proof of Theorem 6.1 we need to describe the mapping properties of s⋆-typical contractions. These investiga- tions will also help us to determine the s⋆-typical spectral properties of contractions. 6.1. Mapping properties of s⋆-typical contractions We prove the following. Proposition 6.3. The set T = {A ∈ C(H) : ∀λ ∈ C (Ran (A − λ · I) is dense in H)} is s⋆-co-meager and s⋆-Gδ in C(H). We need the following elementary property of the backward unilateral shift op- erator. It follows from the fact that its adjoint, the unilateral shift operator, on ℓ2 has empty point spectrum (see e.g. [26]). Lemma 6.4. Let {ei : i ∈ N} be an orthonormal base in H. Define D ∈ C(H) by De0 = 0, Dei+1 = ei (i ∈ N). Then for every λ ∈ C, Ran (D − λ · Id) is dense in H. 21 Proof of Proposition 6.3. First we show that T is s⋆-dense in C(H). Let U ⊆ C(H) be a non-empty s⋆-open set. Then there exist {xi : i < n} ⊆ SH, ε > 0 and A ∈ C(H) such that for every B ∈ C(H), kBxi − Axik ≤ ε (i < n) and kB⋆xi − A⋆xik ≤ 2ε (i < n) imply B ∈ U. It is enough to find a B ∈ U such that for every λ ∈ C, Ran (B − λ · Id) is dense in H. Set V = span{xi, Axi, A⋆xi : i < n}. Let Q : V → V be defined by Q = (1 − ε) · PrV AV . Let dim V = m and let {xi(0) : i < m} be an orthonormal base in V . Let {xi(j) : i < m, j ∈ N \ {0}} be an orthonormal base in V ⊥. Define T : V ⊥ → H by T xi(j) = xi(j − 1) (i < m, j ∈ N \ {0}). Set B = Q ⊕ ε · T ; we show that B fulfills the requirements. We have kBk ≤ kBV k + kBV ⊥k = kQk + ε · kTk ≤ 1 − ε + ε = 1, so B ∈ C(H). For every λ ∈ C, Ran (B − λ · Id) ⊇ Ran (BV ⊥ − λ · IdV ⊥) ⊇ Ran (cid:18)T − ε · IdV ⊥(cid:19) . Since T is an m-fold sum of the operator D of Lemma 6.4, Ran (cid:0)T − λ λ dense in H, as required. For every i < n, ε · IdV ⊥(cid:1) is kBxi − Axik = kQxi − Axik = kQxi − PrV AV xik ≤ ε, kB⋆xi − A⋆xik = kQ⋆xi + ε · T ⋆xi − A⋆xik ≤ ε · kT ⋆xik + kQ⋆xi − A⋆xik ≤ ε + kQ⋆xi − PrV A⋆V xik ≤ 2ε, i.e. B ∈ U, as required. and L ≥ 0 set It remains to show that T is s⋆-Gδ in C(H). To this end, for every y ∈ SH , δ > 0 R(y, δ, L) = {A ∈ C(H) : ∃λ ∈ C (λ ≤ L, dist(Ran (A − λ · Id), y) ≥ δ}. (3) We show that R(y, δ, 1) (y ∈ SH, δ > 0) are s-closed. Let {An : n ∈ N} ⊆ R(y, δ, 1) and A ∈ C(H) be such that s-limn∈N An = A; we show that A ∈ R(y, δ, 1). By An ∈ R(y, δ, 1) (n ∈ N), we have λn ∈ C, λn ≤ 1 (n ∈ N) such that dist(Ran (An − λn · Id), y) ≥ δ. By passing to a subsequence, we can assume {λn : n ∈ N} is convergent, say limn∈N λn = λ; then λ ≤ 1. It's enough to prove dist(Ran (A − λ · Id), y) ≥ δ. Suppose this is not the case, i.e. there is an x ∈ H such that kAx − λx − yk < 22 δ. Since limn∈N Anx = Ax and limn∈N λn = λ, for an n sufficiently large we have kAnx − λnx − yk < δ, which contradicts dist(Ran (An − λn · Id), y) ≥ δ. Thus R(y, δ, 1) (y ∈ SH, δ > 0) are s-closed, and so s⋆-closed, as well. Let Y ⊆ SH be a dense countable set. Since T = C(H) \S{R(y, 2−n, 1) : y ∈ Y, n ∈ N}, the statement follows. (cid:4) For technical reasons, we state the following corollary of Proposition 6.3. Corollary 6.5. The set E = {A ∈ C(H) : Ran A is dense in H} is s⋆-co-meager and s⋆-Gδ in C(H). Proof. By Proposition 6.3, E is s⋆-co-meager in C(H). With the notation of (3), E = C(H) \S{R(y, 2−n, 0) : y ∈ Y, n ∈ N}, so the statement follows.(cid:4) We also need a similar result for positive self-adjoint operators. Proposition 6.6. The set P = {P ∈ P (H) : Ran P is dense in H} is s-co-meager and s-Gδ in P (H). Proof. First we show that P is s-dense in P (H). To this end, let U ⊆ P (H) be a non-empty s-open set. Then there exist {xi : i < n} ⊆ SH , ε > 0 and A ∈ P (H) such that for every B ∈ P (H), kBxi − Axik ≤ ε (i < n) implies B ∈ U. We find a B ∈ U such that Ran B is dense in H. Set V = span{xi, Axi : i < n}. Let Q : V → V be an invertible contractive pos- itive self-adjoint operator such that kQ − PrV AV k ≤ ε; such a Q exists, e.g. any Q = (1−δ)·PrV AV +δ′·IdV fulfills the requirements for suitable 0 < δ, δ′ ≤ ε/2. Set B = Q ⊕ IdV ⊥. Then B is a positive self-adjoint operator, kBV k ≤ 1, kBV ⊥k = 1 and B is invertible, hence B ∈ P (H) and Ran B is dense in H. For every i < n, kBxi − Axik = kQxi − Axik = kQxi − PrV AV xik ≤ ε, i.e. B ∈ U, as required. It remains to show that P is s-Gδ. Observe that for every y ∈ SH and δ > 0, the set R(y, δ) = {A ∈ P (H) : dist(Ran A, y) ≥ δ} is s-closed. Let Y ⊆ SH be a dense countable set, then which completes the proof.(cid:4) P = P (H) \S{R(y, 2−n) : y ∈ Y, n ∈ N}, 23 6.2. Proof of Theorem 6.1 With the notation of Corollary 6.5, set H = E ∩ {A⋆ : A ∈ E} = {A ∈ C(H) : Ran A and Ran A⋆ are dense in H}. By Corollary 6.5, H is an s⋆-co-meager s⋆-Gδ set in C(H). Define ψ1 : H → P (H) by ψ1(A) = (A⋆A)1/2 (A ∈ H). For every A ∈ C(H), (A⋆A)1/2 is a positive self- adjoint contraction so the definition makes sense. Moreover, for A ∈ H we have Ran (A⋆A)1/2 ⊇ Ran (A⋆A) is dense in H, so with the notation of Proposition 6.6, ψ1 maps into P. Note that the mapping A 7→ A⋆A is s⋆-continuous. We need that ψ1 is s⋆-continuous. Since we couldn't find a reference, we outline a proof of the following. Lemma 6.7. The function ·1/2 : P (H) → P (H), A 7→ A1/2 (A ∈ P (H)) is s- continuous. Proof. Let A ∈ P (H) be arbitrary; note that σ(A) ⊆ [0, 1]. By [33, Theorem XI.6.1 p. 313], [33, Theorem XI.4.1 p. 307], [33, Proposition XI.5.2 p. 310] and [33, Theorem XI.8.1 p. 319], there is a s-left-continuous function F : [0, 1] → P (H) such that F is projection-valued, F (0) = 0, F (1) = Id, F (ϑ)F (ϑ′) = F (min{ϑ, ϑ′}) (ϑ, ϑ′ ∈ [0, 1]), and for every continuous function f : [0, 1] → R, R 1 0 f (ϑ)dF (ϑ) exists as the s-limit of Riemann-Stieltjes sums and defines a bounded linear operator, denoted by f (A), satisfying kf (A)k ≤ kfk∞. Moreover, by [33, Example XI.12.1 p. 338] and [33, Theorem XI.12.3 p. 343], this defines a functional calculus compatible with power series expansion, in particular A1/2A1/2 = A. Let {An : n ∈ N} ⊆ P (H) satisfy s-limn∈N An = A. Let x ∈ SH and ε > 0 be arbitrary. Let p : [0, 1] → R be a polynomial satisfying maxx∈[0,1] p(x) − x1/2 < ε. n = Ak (k ∈ N), we have limn∈N p(An)x = p(A)x. As we observed Since s-limn∈N Ak above, kA1/2 − p(A)k,kA1/2 n − p(An)k < ε (n ∈ N). This implies lim supn∈N kA1/2x − A1/2 n xk < 2ε; since ε was arbitrary, the statement follows.(cid:4) Corollary 6.8. The function ψ1 : H → P is s⋆-continuous. Proof. Since the strong and strong-star topologies coincide on P (H), the statement follows.(cid:4) Let A ∈ H be arbitrary. As we observed above, ψ1(A) is a positive self-adjoint operator with dense range. Hence ψ1(A)−1 is a closed densely defined positive self- adjoint operator. Consider the densely defined operator A · ψ1(A)−1. It has a dense range, and for every x ∈ Ran ψ1(A) we have hA · ψ1(A)−1x, A · ψ1(A)−1xi = hA⋆A · ψ1(A)−1x, ψ1(A)−1xi = hψ1(A)−1 · A⋆A · ψ1(A)−1x, xi = hx, xi. 24 Hence A · ψ1(A)−1 is a densely defined isometry with dense range, i.e. it is closable and its closure is an unitary operator. Let ψ0(A) ∈ U(H) denote the closure of A · ψ1(A)−1. Note that by definition, A = ψ0(A) · ψ1(A). Corollary 6.9. Every A ∈ H can be written as A = ψ0(A) · ψ1(A) where ψ0(A) ∈ U(H) and ψ1(A) ∈ P. This decomposition is unique and unitary invariant, i.e. for every A ∈ H and U ∈ U(H) we have ψi(UAU −1) = Uψi(A)U −1 (i < 2). Proof. To see uniqueness, let U, V ∈ U(H) and P, Q ∈ P be arbitrary. Then U · P = V · Q implies P 2 = P ⋆U ⋆UP = Q⋆V ⋆V Q = Q2, so P = Q by the uniqueness of positive square root. By P, Q having dense range, we get U = V . If A ∈ H and U ∈ U(H) then UAU −1 ∈ H and UAU −1 = Uψ0(A)U −1 · Uψ1(A)U −1. Hence by the uniqueness of this decomposition, we have ψ0(UAU −1) = Uψ0(A)U −1 and ψ1(UAU −1) = Uψ1(A)U −1. (cid:4) Recall Ψ : U(H) × P → H, Ψ(U, P ) = U · P , which is an s⋆-continuous function. By Corollary 6.9, Ψ is injective. It is obvious that (ψ0, ψ1) is the inverse of Ψ, so the following completes the proof of Theorem 6.1. Lemma 6.10. The function ψ0 is s⋆-continuous. Proof. Let {An : n ∈ N} ⊆ H and A ∈ H be such that s⋆-limn∈N An = A. For every y ∈ Ran ψ1(A), say y = ψ1(A)x, we have kψ0(An)y − ψ0(A)yk = kψ0(An)ψ1(A)x − ψ0(A)ψ1(A)xk ≤ kψ0(An)ψ1(A)x − ψ0(An)ψ1(An)xk + kψ0(An)ψ1(An)x − ψ0(A)ψ1(A)xk ≤ kψ0(An)k · kψ1(A)x − ψ1(An)xk + kAnx − Axk. Since ψ1 is s⋆-continuous, the statement follows from Ran ψ1(A) ⊆ H being dense.(cid:4) 6.3. Spectral properties of s⋆-typical contractions Our main result in this section is the following. Proposition 6.11. An s⋆-typical contraction A satisfies Cσ(A) = {λ ∈ C : λ ≤ 1}. We start with a lemma. Lemma 6.12. The set S = {A ∈ C(H) : ∀λ ∈ C,λ ≤ 1 (λ ∈ σ(A))} is s⋆-co-meager and s⋆-Gδ in C(H). 25 Proof. First we show that S is s⋆-dense in C(H). Let U ⊆ C(H) be a non-empty s⋆-open set. Then there exist {xi : i < n} ⊆ SH , ε > 0 and A ∈ C(H) such that for every B ∈ C(H), kBxi − Axik < ε (i < n) and kB⋆xi − A⋆xik < ε (i < n) imply B ∈ U. It is enough to find a B ∈ U such that for every λ ∈ C with λ ≤ 1, λ ∈ σ(B). Set V = span{xi, Axi, A⋆xi : i < n}, and let Q : V → V be defined by Q = PrV AV . Let T ∈ C(V ⊥) be arbitrary satisfying λ ∈ σ(T ) (λ ∈ C,λ ≤ 1), say the backward unilateral shift operator, and set B = Q ⊕ T . We show that B fulfills the requirements. Since σ(T ) ⊆ σ(B), we have λ ∈ σ(B) (λ ∈ C,λ ≤ 1). For every i < n, kBxi − Axik = kQxi − Axik = k PrV AV xi − Axik = 0 < ε, kB⋆xi − A⋆xik = kQ⋆xi − A⋆xik = k PrV A⋆V xi − A⋆xik = 0 < ε, i.e. B ∈ U, as required. It remains to show that S is s⋆-Gδ in C(H). For every δ > 0, set S(δ) = {A ∈ C(H) : ∃λ ∈ C (λ ≤ 1 and k(A−λ·I)xk,k(A⋆−λ·I)xk ≥ δ (x ∈ SH))}. It is clear that S(δ) (δ > 0) are s⋆-closed. So the proof will be complete if we show S = C(H) \Sn∈N S(2−n). Let A ∈ C(H) be arbitrary. If λ /∈ σ(A) for a λ ∈ C, λ ≤ 1 then (A − λ · I)−1 and (A⋆ − λ · I)−1 exist so A ∈ S(δ) for some sufficiently small δ > 0. This proves C(H) \ S ⊆Sn∈N S(2−n). Similarly, if A ∈ S(δ) for some δ > 0, say λ ∈ C, λ ≤ 1 witnesses this, then λ /∈ Pσ(A) ∪ Cσ(A) and λ /∈ Pσ(A⋆). By Ker (A⋆ − λ · I) = Ran (A − λ · I)⊥, λ /∈ Pσ(A⋆) implies λ /∈ Rσ(A). So we obtained λ /∈ σ(A), i.e. A /∈ S. This proves Sn∈N S(2−n) ⊆ C(H) \ S and finishes the proof.(cid:4) Proof of Proposition 6.11. The map A 7→ A⋆ is a s⋆-homeomorphism of C(H). So by Proposition 6.3, a typical contraction A satisfies that for every λ ∈ C, Ran (A − λ · I) and Ran (A⋆ − λ · I) are dense in H. Since Ker (A − λ · I) = Ran (A⋆ − λ · I)⊥, we get Ker (A − λ · I) = ∅ (λ ∈ C). Hence Pσ(A) = Pσ(A⋆) = ∅ and Rσ(A) = Rσ(A⋆) = ∅. So Cσ(A) = {λ ∈ C : λ ≤ 1} is an immediate corollary of Lemma 6.12.(cid:4) 7. The norm topology In the norm topology it is not possible to give a simple description of the spectral properties of typical operators. An intuitive explanation may be that the norm 26 topology is non-separable, hence several different properties can coexist on non- meager sets. In this section, every topological notion refers to the norm topology. We prove the following. Theorem 7.1. Let λ ∈ C be arbitrary. Then the following sets of operators have non-empty interior. 1. {A ∈ B(H) : λ /∈ σ(A)}; 2. {A ∈ B(H) : λ ∈ Rσ(A)}; 3. {A ∈ B(H) : λ ∈ Pσ(A)}. In particular, the following sets have non-empty interior. 4. {A ∈ B(H) : Ran (A − λ · I) = H}; 5. {A ∈ B(H) : Ran (A − λ · I) is not dense in H}. On the other hand, the following sets of operators are nowhere dense. 6. {A ∈ B(H) : λ ∈ Cσ(A)}; 7. {A ∈ B(H) : Ran (A − λ · I) is dense in H but not equal to H}. We need some terminology in advance. Definition 7.2. Let A ∈ B(H) and λ ∈ C be arbitrary. We say λ ∈ σ(A) is stable if there is an ε > 0 such that for every D ∈ B(H) with kDk < ε we have λ ∈ σ(A + D). Similarly, we say λ ∈ Pσ(A) is stable if there is an ε > 0 such that for every D ∈ B(H) with kDk < ε we have λ ∈ Pσ(A + D). With this terminology, [17, Theorem 2 p. 912] can be reformulated as follows. Proposition 7.3. Let A ∈ B(H) and λ ∈ C be arbitrary. Then the following are equivalent. 1. λ ∈ σ(A) is stable; 2. dim Ker (A − λ · I) 6= dim Ran (A − λ · I)⊥ and there is an ε > 0 such that for every x ∈ Ker (A − λ · I)⊥ we have k(A − λ · I)xk ≥ ε · kxk. 27 We prove a similar result for the point spectrum. Proposition 7.4. Let A ∈ B(H) and λ ∈ C be arbitrary. Then the following are equivalent. 1. λ ∈ Pσ(A) is stable; 2. dim Ran (A − λ · I)⊥ < dim Ker (A − λ · I) and there is a ε > 0 such that for every x ∈ Ker (A − λ · I)⊥ we have k(A − λ · I)xk ≥ ε · kxk. Proof. By linearity, we can assume λ = 0. Suppose first 0 ∈ Pσ(A) is stable. Then 0 ∈ σ(A) is stable so by Proposition 7.3, dim Ker (A) 6= dim Ran (A)⊥ and there is a ε > 0 such that for every x ∈ Ker (A)⊥ we have kAxk ≥ ε · kxk. If dim Ker (A) < dim Ran (A)⊥ then for every ε > 0 there is a D ∈ B(H) such that kDk < ε, DKer (A)⊥ = 0 and for every x ∈ Ker (A)\{0}, Dx ∈ Ran (A)⊥ \{0}. Then for every x ∈ H \{0} we have (A + D)x 6= 0, hence 0 /∈ Pσ(A + D). This contradicts the assumption that 0 ∈ Pσ(A) is stable. Thus dim Ran (A)⊥ < dim Ker (A), as required. To see the converse, suppose the conditions of statement 2 hold. Then we have dim Ker (A) > 0 hence 0 ∈ Pσ(A). We also have that A : Ker (A)⊥ → Ran (A) is a bounded invertible operator, in particular Ran (A) is a closed co-finite dimensional subspace of H. Let k = 1 + dim Ran (A)⊥. Let D ∈ B(H) be arbitrary with kDk < ε/2k2. For every x ∈ Ker (A) consider the following inductive definition of a sequence {xn : n ∈ N} ⊆ H. Set x0 = x. Let n ∈ N be arbitrary and suppose that xn ∈ H is defined. Write Dxn = un + vn where un ∈ Ran (A) and vn ∈ Ran (A)⊥. Set xn+1 = A−1un. This completes the inductive step of the definition of {xn : n ∈ N}. We set ξ(x) = Pn∈N(−1)nxn, ρ(x) = Pn∈N(−1)nvn. Note that kxn+1k ≤ kunk/ε ≤ kDxnk/ε < kxnk/2k2, hence kxnk < kxk/(2k2)n (n ∈ N) so the defi- nitions make sense. Moreover, the functions ξ and ρ are linear, ρ(x) ∈ Ran (A)⊥ and kξ(x)− xk < 1/k (x ∈ BKer (A)). So by Lemma 5.6, if {x(i) : i < k} is an orthonormal system in Ker (A) then {ξ(x(i)) : i < k} are linearly independent. Observe that (A + D)ξ(x) = (A + D)Xn∈N Ax0 + Xn∈N\{0} (−1)nxn =Xn∈N (−1)nAxn +Xn∈N (−1)n(Axn + Dxn) = (−1)n(un + vn) =Xn∈N (−1)nvn = ρ(x). 28 By dim Ker (A) ≥ k, there is an orthonormal system {x(i) : i < k} in Ker (A). Since dim Ran (A)⊥ < k, there is an x ∈ span{x(i) : i < k}\{0} such that ρ(x) = 0. Then (A + D)ξ(x) = 0 shows 0 ∈ Pσ(A + D), so the proof is complete.(cid:4) Proof of Theorem 7.1. By linearity, it is enough to consider the λ = 0 case. The set of invertible operators shows statements 1 and 4. By Proposition 7.4, a neighborhood of the backward unilateral shift operator shows statement 3. The set of statement 7 contains the set of statement 6 so it is sufficient to prove statement 7. Let A ∈ B(H) be arbitrary satisfying Ran (A) is dense in H but not equal to H. Then A cannot be invertible. Moreover, note that condition kAxk ≥ ε · kxk (x ∈ Ker (A)⊥) in Proposition 7.3.2 would imply Ran (A) is closed. Since Ran (A) cannot be closed, A cannot satisfy Proposition 7.3.2. So the set of statement 7 is contained in the boundary of the set of invertible operators hence it is nowhere dense. Finally notice that a neighborhood of the unilateral shift operator D⋆ is contained in {A ∈ B(H) : 0 ∈ σ(A) \ Pσ(A)}, since the stability of 0 ∈ σ(D⋆) is implied by Proposition 7.3 while the stability of 0 /∈ Pσ(D⋆) follows from D⋆ being an isometry. So by statement 6, statements 2 and 5 follow.(cid:4) Observe that the backward unilateral shift operator D of Lemma 6.4 satisfies the conditions of statement 2 of Proposition 7.4 for every λ ∈ C with λ < 1, hence every such λ ∈ Pσ(D) is stable. So by taking direct sums of scaled and translated copies of D we can obtain arbitrarily complicated stable spectra. We also obtain the following. Here an operator T is called embeddable, if it can be embedded into a strongly continuous semigroup, i.e. if T = T (1) holds for some strongly continuous semigroup (T (t))t≥0. Corollary 7.5. The set of embeddable operators and the set of non-embeddable op- erators have non-empty interior. Proof. By [10, Theorem 2.1 p. 452], each operator in a small neighborhood of Id is embeddable. To show a non-empty open set of non-embeddable operators, we aim to use [10, Theorem 3.1 p. 454]. Recall the backward unilateral shift operator D of Lemma 6.4. By Proposition 7.4, each operator A in a small neighborhood of D satisfies 0 ∈ Pσ(A). So we get A is non-embeddable if we show dim Ker (A) ≤ 1. Note that if kA − Dk < 1/√2 then for every x, y ∈ Ker (A) ∩ SH we have kDxk,kDyk < 1/√2. This implies kx − D⋆Dxk,ky − D⋆Dyk > 1/√2, and since D⋆D is an orthogonal projection, we have hx, yi = hx − D⋆Dx, y − D⋆Dyi + hD⋆Dx, D⋆Dyi ≥ hx − D⋆Dx, y − D⋆Dyi − hD⋆Dx, D⋆Dyi > 0. 29 Thus dim Ker (A) ≤ 1, which completes the proof. (cid:4) The results of this section provide no information about the size of the spectrum of a typical operator. We state two questions related to this in Problem 8.2. Here we only point out the following. Proposition 7.6. The following sets are dense. 1. {A ∈ B(H) : Pσ(A) 6= ∅}; 2. {A ∈ B(H) : Rσ(A) 6= ∅}. Proof. Since A 7→ A⋆ (A ∈ B(H)) is a homeomorphism of B(H) and Pσ(A⋆) = Rσ(A) (A ∈ B(H)), it suffices to show 1. Let B ∈ B(H) be arbitrary. Since the approximative point spectrum of B is [6, Proposition VII.6.7 p. 210]), there exists a λ ∈ σ(B) and a nonempty (see e.g. sequence {xn : n ∈ N} ⊆ SH such that limn∈N kBxn − λxnk = 0. For every n ∈ N, let Pn denote the orthogonal projection onto span{xn}, and set An = B(Id−Pn)+λ·Pn. We have Anxn = λxn and kAn − Bk = kλ · Pn − BPnk = kλxn − Bxnk (n ∈ N). Hence Pσ(An) 6= ∅ (n ∈ N) and limn∈N kAn − Bk = 0. This completes the proof.(cid:4) 8. An outlook to general Banach spaces The proofs in the previous section could convince the reader that an attempt to extend our investigations to tackle the typical properties of contractions on arbi- trary Banach spaces could encounter considerable technical difficulties. In addition, such an endeavor has to face some problems of more fundamental nature. Recent developments in the theory of Banach spaces resulted in numerous spaces exhibiting surprising functional analytic properties. On the famous Banach space of S. A. Argy- ros and R. G. Haydon, every bounded linear operator is of the form λ· Id + K where λ ∈ C and K is a compact operator (see [1]). Spaces were constructed with only trivial isometries (see e.g. [7] and the references therein), moreover even a renorming of a Banach space may result in an arbitrary isometry group (see [18] and the refer- ences therein). The relevance of isometries comes from the important role played by unitary operators in the theory developed in the previous section. Observe that renorming does not change the topology of the underlying Banach space, so the five topologies we consider on operators remain unchanged, as well. Instead, renorming affects the size of various classes of operators. So in a sense, the study of typical properties of operators in various topologies is more related to the geometry of the underlying Banach space than to the topology it carries. Therefore, 30 reasonable extensions of our investigations should be pursued in Banach spaces where the geometry of the space is of special significance. The most obvious such spaces are the Lp spaces (1 ≤ p ≤ ∞). The functional analysis of these spaces is well-developed, isometric operators are characterized (see e.g. [22]). Nevertheless, we expect that none of our main results extends to arbitrary Lp spaces. Another promising extension of the theory of typical behavior of operators could tackle Banach⋆-algebras. Developing our results to that generality could separate typical properties which are of operator theoretic nature from typical properties exploiting the geometry of the underlying space. Finally let us propose some concrete problems which stem from the results of the previous section. Problem 8.1. We say A, B ∈ B(H) are similar if there is an invertible operator T ∈ B(H) such that A = T BT −1. 1. Is it true that w-typical contractions are similar? 2. Is it true that s⋆-typical contractions are similar? Problem 8.2. Determine the s-typical properties of self-adjoint and positive self- adjoint operators. Problem 8.3. Is a s⋆-typical contraction embeddable into a strongly continuous semigroup? Problem 8.4. 1. Is the set {A ∈ B(H) : Pσ(A) 6= ∅} co-meager in the norm topology? 2. Is the set {A ∈ B(H) : Cσ(A) = ∅} co-meager in the norm topology? It would be instructive to examine whether suitable analogues of our results hold for strongly continuous semigroups instead of single operators. Let C c(H) denote the set of contractive C0-semigroups. Given any metric d on C(H), one can endow C c(H) with the topology generated by uniform d-convergence on compact time intervals. This topology is induced by the metric dc(T (·), S(·)) =Pn∈N 2−n supt∈[0,n] d(T (t), S(t)). Problem 8.5. Under which of these topologies is Cc(H) a Baire space? What are the typical properties of contractive C0-semigroups in these topologies? Recall that the strong topology case was treated in Section 5.4. Some results related to Problem 8.5 for the weak topology can be found in [8, Section 4], [12] and [14]. 31 References [1] S. A. Argyros, R. G. Haydon, A hereditarily indecomposable L∞-space that solves the scalar-plus-compact problem, Acta Math. 206 (2011), 1-54. [2] S. Banach, Uber die Baire'sche Kategorie gewisser Funktionenmengen, Studia Math. 3 (1931), 174 -- 179. [3] W. Bartoszek and B. Kuna, On residualities in the set of Markov operators on C1, Proc. Amer. Math. Soc. 133 (2005), 2119 -- 2129. [4] A. Bruckner, Differentiation of real functions, Second edition, CRM Monograph Series, 5, American Mathematical Society, Providence, RI, 1994. [5] J. R. Choksi, M. G. Nadkarni, Baire category in spaces of measures, unitary operators, and transformations, Proc. Int. Conference on Invariant Subspaces and Allied Topics (1986), Narosa Publishers, New Delhi. [6] J. B. Conway, A Course in Functional Analysis. Second edition, Graduate Texts in Mathematics 96, Springer-Verlag, New York, 1990. [7] G. Diestel, A. Koldobsky, Sobolev spaces with only trivial isometries, Positivity 10 (2006), no. 1, 135 -- 144. [8] T. Eisner, A "typical" contraction is unitary, Enseign. Math. (2) 56 (2010), 403-410. [9] T. Eisner, Rigidity of contractions on Hilbert spaces, arXiv:0909.4695v1. [10] T. Eisner, Embedding operators into strongly continuous semigroups, Arch. Math. (Basel) 92 (2009), 451 -- 460. [11] T. Eisner, Stability of Operators and Operator Semigroups, Operator Theory: Advances and Applications, 209, Birkhauser Verlag, Basel, 2010. [12] T. Eisner, A. Ser´eny, Category theorems for stable operators on Hilbert spaces, Acta Sci. Math. (Szeged) 74 (2008), 259 -- 270. [13] T. Eisner, A. Ser´eny, Category theorems for stable semigroups, Ergodic Theory Dynamical Systems 29 (2009), 487 -- 494. [14] T. Eisner, A. Ser´eny, On the weak analogue of the Trotter -- Kato theorem, Tai- wanese J. Math. 14 (2010), 1411 -- 1416. 32 [15] K.-J. Engel, R. Nagel, One-parameter semigroups for linear evolution equations, Graduate Texts in Mathematics, 194, Springer-Verlag, New York, 2000. [16] P. R. Halmos, In general a measure preserving transformation is mixing, Ann. Math. 45 (1944), 786 -- 792. [17] J. Feldman, R. V. Kadison, The closure of the regular operators in a ring of operators, Proc. Amer. Math. Soc. 5 (1954), 909 -- 916. [18] V. Ferenczi, E. M. Galego, Countable groups of isometries on Banach spaces, Trans. Amer. Math. Soc. 362 (2010), 4385-4431. [19] S. Jitomirskaya, B. Simon, Operators with singular continuous spectrum. III. Almost periodic Schrdinger operators, Comm. Math. Phys. 165 (1994), no. 1, 201 -- 205. [20] A. S. Kechris, Classical Descriptive Set Theory, Graduate Texts in Mathematics, 156, Springer-Verlag, 1994. [21] B. Kuna, On residualities in the set of Markov continuous semigroups on C1, Demonstratio Math. 39 (2006), no. 2, 439 -- 453. [22] J. Lamperti, On the isometries of certain function-spaces, Pacific J. Math. 8 (1958) 459 -- 466. [23] S. Mazurkiewicz, Sur les fonctions non d´erivables, Studia Math. 3 (1931), 92 -- 94. [24] M. G. Nadkarni, Spectral theory of dynamical systems, Birkhauser Advanced Texts: Basler Lehrbucher, Birkhauser Verlag, Basel, 1998. [25] V. V. Peller, Estimates of operator polynomials in the space Lp with respect to the multiplicative norm, J. Math. Sciences 16 (1981), 1139 -- 1149. [26] W. C. Ridge, Approximate point spectrum of a weighted shift, Trans. Amer. Math. Soc. 147 (1970), 349 -- 356. [27] R. Del Rio, N. Makarov, B. Simon, Operators with singular continuous spec- trum. II. Rank one operators, Comm. Math. Phys. 165 (1994), no. 1, 59 -- 67. [28] V. A. Rohlin, A "general" measure-preserving transformation is not mixing, Doklady Akad. Nauk SSSR 60 (1948), 349 -- 351. 33 [29] A. L. Shields, Weighted shift operators and analytic function theory, Topics in Operator Theory, Math. Surveys No. 13, Amer. Math. Soc., Providence, R.I., 1974, 49 -- 128. [30] B. Simon, Operators with singular continuous spectrum: I. General operators, Annals Math. 141 (1995), 131 -- 145. [31] B. Szokefalvi-Nagy, C. Foia¸s, Harmonic Analysis of Operators on Hilbert Space, North-Holland, Akad´emiai Kiad´o, Amsterdam, Budapest, 1970. [32] M. Takesaki, Theory of operator algebras I, Springer-Verlag, New York- Heidelberg, 1979. [33] K. Yosida, Functional analysis, Sixth edition, Grundlehren der Mathematischen Wissenschaften, 123, Springer-Verlag, Berlin-New York, 1980. [34] Y.-C. Wong, Introductory theory of topological vector spaces, Monographs and Textbooks in Pure and Applied Mathematics, 167, Marcel Dekker, Inc., New York, 1992. [35] P. Zorin-Kranich, Typical bounded operators admit common cyclic vectors, sub- mitted, arXiv:1008.2878. 34
1006.5803
1
1006
2010-06-30T08:36:36
On the two-dimensional moment problem
[ "math.FA" ]
In this paper we obtain an algorithm towards solving the two-dimensional moment problem. This algorithm gives the necessary and sufficient conditions for the solvability of the moment problem. It is shown that all solutions of the moment problem can be constructed using this algorithm.
math.FA
math
On the two-dimensional moment problem. S.M. Zagorodnyuk 1 Introduction. In this paper we analyze the two-dimensional moment problem. Recall that this problem consists of finding a non-negative Borel measure µ in R2 such that xm 1 xn 2 dµ = sm,n, m, n ∈ Z+, (1) ZR2 where {sm,n}m,n∈Z+ is a prescribed sequence of complex numbers. The two-dimensional moment problem and the (closely related to this subject) complex moment problem have an extensive literature, see books [1], [2], [3], surveys [4],[5] and [6]. Some conditions of solvability for this moment problem were obtained by Kilpi and by Stochel and Szafraniec, see e.g. [2] and [6]. However, these conditions are hard to check. To the best of our knowledge, the description of all solutions was not obtained. Firstly, we obtain a solvability criterion for an auxiliary extended two- dimensional moment problem. It is shown that this problem is always de- terminate and its solution can be constructed explicitly. An idea of our algorithm for solving the two-dimensional moment problem is to extend the prescribed sequence of moments (1) to be a moment sequence of the extended moment problem. It is shown that all solutions of the moment problem (1) can be constructed on this way. The method of proof uses an abstract operator approach, see [7]. Roughly speaking, the final algorithm reduces to the solving of finite and infinite linear systems of equations with parameters. Notations. As usual, we denote by R, C, N, Z, Z+ the sets of real numbers, complex numbers, positive integers, integers and non-negative in- tegers, respectively. The real plane will be denoted by R2. For a subset S of R2 we denote by B(S) the set of all Borel subsets of S. Everywhere in this paper, all Hilbert spaces are assumed to be separable. By (·,·)H and k · kH we denote the scalar product and the norm in a Hilbert space H, respectively. The indices may be omitted in obvious cases. For a set M in H, by M we mean the closure of M in the norm k · kH . For {xm,n}m,n∈Z+ , xm,n ∈ H, we write Lin{xm,n}m,n∈Z+ for the set of linear combinations of elements {xm,n}m,n∈Z+ and span{xm,n}m,n∈Z+ = Lin{xm,n}m,n∈Z+. The identity operator in H is denoted by EH . For an arbitrary linear operator 1 A in H, the operators A∗,A,A−1 mean its adjoint operator, its closure and its inverse (if they exist). By D(A) and R(A) we mean the domain and the range of the operator A. The norm of a bounded operator A is denoted by kAk. By P H = PH1 we mean the operator of orthogonal projection in H on a subspace H1 in H. By L2 µ we denote the usual space of square-integrable complex functions f (x1, x2), x1, x2 ∈ R2, with respect to the Borel measure µ in R2. H1 2 The solution of an extended two-dimensional moment problem. Consider the following moment problem: to find a non-negative Borel mea- sure µ in R2 such that ZR2 1 (x1 + i)k(x1 − i)lxn xm 2 (x2 + i)r(x2 − i)tdµ = um,k,l;n,r,t, m, n ∈ Z+, k, l, r, t ∈ Z, (2) where {um,k,l;n,r,t}m,n∈Z+,k,l,r,t∈Z is a prescribed sequence of complex num- bers. This problem is said to be the extended two-dimensional moment problem. We set Ω = {(m, k, l; n, r, t) : m, n ∈ Z+, k, l, r, t ∈ Z}, Ω0 = {(m, k, l; n, r, t) : m, n ∈ Z+, k, l, r, t ∈ Z, k = l = r = t = 0}, Ω′ = Ω\Ω0. Let the moment problem (2) have a solution µ. Choose an arbitrary function αm,k,l;n,r,txm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)t, where all but finite number of complex coefficients αm,k,l;n,r,t are zeros. Then P (x1, x2) = X(m,k,l;n,r,t)∈Ω 0 ≤ZR2 P (x1, x2)2dµ = xm+m′ 1 (x1 + i)k+l′ ∗ZR2 = X (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′ X (x1 − i)l+k′ αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′ um+m′,k+l′,l+k′;n+n′,r+t′,t+r′. (x2 − i)t+r′ dµ xn+n′ 2 (x2 + i)r+t′ 2 Therefore X αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′ um+m′,k+l′,l+k′;n+n′,r+t′,t+r′ ≥ 0, (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω (3) for arbitrary complex coefficients αm,k,l;n,r,t, where all but finite number of αm,k,l;n,r,t are zeros. The latter condition on the coefficients αm,k,l;n,r,t in infinite sums will be assumed in similar situations. We shall use the following important fact (e.g. [8, pp.361-363]). Theorem 1 Let a sequence of complex numbers {um,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfy condition (3). Then there exist a separable Hilbert space H with a scalar product (·,·)H and a sequence {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω in H, such that (xm,k,l;n,r,t, xm′,k′,l′;n′,r′,t′)H = um+m′,k+l′,l+k′;n+n′,r+t′,t+r′ , (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω, (4) and span{xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω = H. Proof (We do not claim the originality of the idea of this proof). Choose an arbitrary infinite-dimensional linear vector space V (for instance, one may choose the space of all complex sequences (un)n∈N, un ∈ C). Let X = {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω be an arbitrary infinite sequence of linear in- dependent elements in V which is indexed by elements of Ω. Set LX = Lin{xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω. Introduce the following functional: [x, y] = αm,k,l;n,r,tβm′,k′,l′;n′,r′,t′ (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω X ∗ um+m′,k+l′,l+k′;n+n′,r+t′,t+r′, (5) for x, y ∈ LX, y = x = X(m,k,l;n,r,t)∈Ω X(m′,k′,l′;n′,r′,t′)∈Ω αm,k,l;n,r,txm,k,l;n,r,t, βm′,k′,l′;n′,r′,t′ xm′,k′,l′;n′,r′,t′, where αm,k,l;n,r,t, βm′,k′,l′;n′,r′,t′ ∈ C. Here all but finite number of indices αm,k,l;n,r,t, βm′,k′,l′;n′,r′,t′ are zeros. The set LX with [·,·] will be a quasi-Hilbert space. Factorizing and making the completion we obtain the desired space H (e.g. [3]). ✷ 3 Let the moment problem (2) be given and the condition (3) hold. By Theorem 1 there exist a Hilbert space H and a sequence {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H, such that relation (4) holds. Set L = Lin{xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω. In- troduce the following operators A0 X(m,k,l;n,r,t)∈Ω B0 X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm,k,l;n,r,t = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm,k,l;n,r,t = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm+1,k,l;n,r,t, αm,k,l;n,r,txm,k,l;n+1,r,t, (6) (7) where all but finite number of complex coefficients αm,k,l;n,r,t are zeros. Let us check that these definitions are correct. Indeed, suppose that x = X(m,k,l;n,r,t)∈Ω We may write αm,k,l;n,r,txm,k,l;n,r,t = X(m′,k′,l′;n′,r′,t′)∈Ω βm′,k′,l′;n′,r′,t′ xm′,k′,l′;n′,r′,t′. (8)  X(m,k,l;n,r,t)∈Ω = X(m,k,l;n,r,t)∈Ω = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm+1,k,l;n,r,t, xa,b,c;d,e,fH αm,k,l;n,r,tum+1+a,k+c,l+b;n+d,r+f,t+e αm,k,l;n,r,t(xm,k,l;n,r,t, xa+1,b,c;d,e,f )H = (x, xa+1,b,c;d,e,f )H , (a, b, c; d, e, f ) ∈ Ω. In the same manner we obtain:  X(m′,k′,l′;n′,r′,t′)∈Ω αm′,k′,l′;n′,r′,t′xm′+1,k′,l′;n′,r′,t′ , xa,b,c;d,e,fH = (x, xa+1,b,c;d,e,f )H , (a, b, c; d, e, f ) ∈ Ω. Therefore the definition of A0 is correct. The correctness of the definition of B0 can be checked in a similar manner. Notice that (A0xm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm+1,k,l;n,r,t, xa,b,c;d,e,f )H 4 = um+1+a,k+c,l+b;n+d,r+f,t+e = (xm,k,l;n,r,t, xa+1,b,c;d,e,f )H = (xm,k,l;n,r,t, A0xa,b,c;d,e,f )H , (m, k, l; n, r, t), (a, b, c; d, e, f ) ∈ Ω. Therefore A0 is symmetric. The same argument implies that B0 is symmet- ric, as well. Suppose that the following conditions hold: um+1+a,k+c,l+b;n+d,r+f,t+e + ium+a,k+c,l+b;n+d,r+f,t+e = um+a,k+1+c,l+b;n+d,r+f,t+e, um+1+a,k+c,l+b;n+d,r+f,t+e − ium+a,k+c,l+b;n+d,r+f,t+e = um+a,k+c,l+1+b;n+d,r+f,t+e, um+a,k+c,l+b;n+1+d,r+f,t+e + ium+a,k+c,l+b;n+d,r+f,t+e = um+a,k+c,l+b;n+d,r+1+f,t+e, (9) (10) (11) um+a,k+c,l+b;n+1+d,r+f,t+e − ium+a,k+c,l+b;n+d,r+f,t+e (12) for all (m, k, l; n, r, t), (a, b, c; d, e, f ) ∈ Ω. These conditions are equivalent to conditions = um+a,k+c,l+b;n+d,r+f,t+1+e, (xm+1,k,l;n,r,t + ixm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm,k+1,l;n,r,t, xa,b,c;d,e,f )H , (xm+1,k,l;n,r,t − ixm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm,k,l+1;n,r,t, xa,b,c;d,e,f )H , (xm,k,l;n+1,r,t + ixm,k,l;n,r,t, xa,b,c;d,e,f )H = (xm,k,l;n,r+1,t, xa,b,c;d,e,f )H , (xm,k,l;n+1,r,t − ixm,k,l;n,r,t, xa,b,c;d,e,f )H (13) (14) (15) (16) for all (m, k, l; n, r, t), (a, b, c; d, e, f ) ∈ Ω. The latter conditions are equiva- lent to the following conditions: = (xm,k,l;n,r,t+1, xa,b,c;d,e,f )H , xm+1,k,l;n,r,t + ixm,k,l;n,r,t = xm,k+1,l;n,r,t, xm+1,k,l;n,r,t − ixm,k,l;n,r,t = xm,k,l+1;n,r,t, xm,k,l;n+1,r,t + ixm,k,l;n,r,t = xm,k,l;n,r+1,t, (17) (18) (19) 5 xm,k,l;n+1,r,t − ixm,k,l;n,r,t = xm,k,l;n,r,t+1, for all (m, k, l; n, r, t) ∈ Ω. The last conditions mean that (A0 + iEH )xm,k,l;n,r,t = xm,k+1,l;n,r,t, (A0 − iEH )xm,k,l;n,r,t = xm,k,l+1;n,r,t, (B0 + iEH)xm,k,l;n,r,t = xm,k,l;n,r+1,t, (B0 − iEH)xm,k,l;n,r,t = xm,k,l;n,r,t+1, (20) (21) (22) (23) (24) for all (m, k, l; n, r, t) ∈ Ω. The latter conditions imply that (A0 ± iEH )L = L, (B0 ± iEH )L = L. Therefore operators A0 and B0 are essentially self-adjoint. The condi- tions (21)-(24) also imply that (A0 + iEH )−1xm,k,l;n,r,t = xm,k−1,l;n,r,t, (A0 − iEH )−1xm,k,l;n,r,t = xm,k,l−1;n,r,t, (B0 + iEH)−1xm,k,l;n,r,t = xm,k,l;n,r−1,t, (B0 − iEH )−1xm,k,l;n,r,t = xm,k,l;n,r,t−1, for all (m, k, l; n, r, t) ∈ Ω. Consider the Cayley transformations of A0 and B0: VA0 = (A0 − iEH )(A0 + iEH )−1, (25) (26) (27) (28) (29) VB0 = (B0 − iEH )(B0 + iEH )−1, D(A0) = D(B0) = L. (30) By virtue of relations (22),(24),(25),(27) we obtain: VA0 VB0 xm,k,l;n,r,t = xm,k−1,l+1;n,r−1,t+1 = VB0 VA0 xm,k,l;n,r,t, for all (m, k, l; n, r, t) ∈ Ω. Therefore VA0 VB0 x = VB0 VA0 x, x ∈ L. (31) By continuity we extend the isometric operators VA0 and VB0 to unitary operators UA0 and VB0 in H, respectively. By continuity we conclude that UA0 UB0 x = UB0 UA0 x, x ∈ H. (32) 6 Set A = A0, B = B0. The Cayley transformations of the self-adjoint oper- rators A and B coincide on L with UA0 and UB0, respectively. Thus, the Cayley transformations of A and B are UA0 and UB0, respectively. There- fore, operators A and B commute. Notice that xm,k,l;n,r,t = Am(A + i)k(A − i)lBn(B + i)r(B − i)tx0,0,0;0,0,0, for all (m, k, l; n, r, t) ∈ Ω. In fact, by induction we can check that (33) xm,k,l;n,r,t = (B − iEH )txm,k,l;n,r,0, t ∈ Z, for any fixed m, n ∈ Z+, k, l, r ∈ Z; xm,k,l;n,r,0 = (B + iEH)rxm,k,l;n,0,0, r ∈ Z, for any fixed m, n ∈ Z+, k, l ∈ Z; xm,k,l;n,0,0 = Bnxm,k,l;0,0,0, n ∈ Z+, for any fixed m ∈ Z+, k, l ∈ Z; xm,k,l;0,0,0 = (A − iEH )lxm,k,0;0,0,0, l ∈ Z, for any fixed m ∈ Z+, k ∈ Z; xm,k,0;0,0,0 = (A + iEH )lxm,0,0;0,0,0, k ∈ Z, for any fixed m ∈ Z+; xm,0,0;0,0,0 = Amx0,0,0;0,0,0, m ∈ Z+, and then by substitution of each relation into previous one we obtain rela- tion (33). For the commuting self-adjoint operators A and B there exists an or- thogonal operator spectral measure E(x) on B(R2) such that A =ZR2 x1dE(x), B =ZR2 x2dE(x). (34) Then =(cid:18)ZR2 um,k,l;n,r,t = (xm,k,l;n,r,t, x0,0,0;0,0,0)H xm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)tdE(x)x0,0,0;0,0,0, x0,0,0;0,0,0(cid:19)H 7 =ZR2 Hence, the Borel measure xm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)td(E(x)x0,0,0;0,0,0, x0,0,0;0,0,0)H . µ = (E(x)x0,0,0;0,0,0, x0,0,0;0,0,0)H , (35) is a solution of the moment problem (2). Theorem 2 Let the extended two-dimensional moment problem (2) be given. The moment problem has a solution if and only if conditions (3) and (9)- (12) are satisfied. If these conditions are satisfied then the solution of the moment problem is unique and can be constructed by (35). Proof. The sufficiency of conditions (3) and (9)-(12) for the existence of a solution of the moment problem (2) was shown before the statement of the Theorem. The necessity of condition (3) was proved, as well. Let us check that conditions (9)-(12) are necessary for the solvability of the moment problem (2). Let µ be a solution of the moment problem (2). Consider the space L2 the following subsets in L2 µ: µ and Lµ = Lin{xm We denote 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)t}(m,k,l;n,r,t)∈Ω, Hµ = Lµ. (36) ym,k,l;n,r,t := xm 1 (x1 + i)k(x1− i)lxn 2 (x2 + i)r(x2− i)t, (m, k, l; n, r, t) ∈ Ω. (37) Notice that (ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)L2 µ = um+m′,k+l′,l+k′;n+n′,r+t′,t+r′, (38) for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Consider the operators of multiplication by the independent variable in L2 µ: Aµf (x1, x2) = x1f (x1, x2), Bµf (x1, x2) = x2f (x1, x2), f ∈ L2 µ. Notice that µ (Aµ + iEL2 (Aµ − iEL2 (Bµ + iEL2 µ µ )ym,k,l;n,r,t = ym,k+1,l;n,r,t, )ym,k,l;n,r,t = ym,k,l+1;n,r,t, )ym,k,l;n,r,t = ym,k,l;n,r+1,t, 8 (39) (40) (41) (42) (Bµ − iEL2 µ )ym,k,l;n,r,t = ym,k,l;n,r,t+1, (43) for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Since conditions (3) are satisfied, by Theorem 1 there exist a Hilbert space H and a sequence of elements {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H, such that relation (4) holds. Repeating arguments after the Proof of Theorem 1 we construct operators A0 and B0 in H. Consider the following operator: W0 X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,tym,k,l;n,r,t = X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,txm,k,l;n,r,t, (44) where all but finite number of complex coefficients αm,k,l;n,r,t are zeros. Let us check that this operator is defined correctly. In fact, suppose that αm,k,l;n,r,tym,k,l;n,r,t = X(m′,k′,l′;n′,r′,t′)∈Ω βm′,k′,l′;n′,r′,t′ ym′,k′,l′;n′,r′,t′, (45) where βm′,k′,l′;n′,r′,t′ ∈ C. We may write x = X(m,k,l;n,r,t)∈Ω 0 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) = = ∗(αm′,k′,l′;n′,r′,t′ − βm′,k′,l′;n′,r′,t′)(ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)L2 µ (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω (αm,k,l;n,r,t − βm,k,l;n,r,t) (αm,k,l;n,r,t − βm,k,l;n,r,t)ym,k,l;n,r,t(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) X(m,k,l;n,r,t)∈Ω X X (αm,k,l;n,r,t − βm,k,l;n,r,t)xm,k,l;n,r,t(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H X(m,k,l;n,r,t)∈Ω Thus, the operator W0 is defined correctly. Ifex ∈ H and ex = X(m,k,l;n,r,t)∈Ω (αm,k,l;n,r,t − βm,k,l;n,r,t) =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) γm,k,l;n,r,tym,k,l;n,r,t, (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω 2 L2 µ . ∗(αm′,k′,l′;n′,r′,t′ − βm′,k′,l′;n′,r′,t′)(xm,k,l;n,r,t, xm′,k′,l′;n′,r′,t′)H 9 (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω αm,k,l;n,r,tγm′,k′,l′;n′,r′,t′ X where γm,k,l;n,r,t ∈ C, then (W0x, W0ex)H = = (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω X ∗(xm,k,l;n,r,t, xm′,k′,l′;n′,r′,t′)H αm,k,l;n,r,tγm′,k′,l′;n′,r′,t′(ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)H By continuity we extend W0 to a unitary operator W which maps Hµ onto H. Observe that . µ = (x,ex)L2 W −1A0W ym,k,l;n,r,t = ym+1,k,l;n,r,t = Aµym,k,l;n,r,t, (46) W −1B0W ym,k,l;n,r,t = ym,k,l;n+1,r,t = Bµym,k,l;n,r,t, (47) for all (m, k, l; n, r, t) ∈ Ω. By using the last relations in relations (40)- (43) we obtain relations (21)-(24). The latter relations are equivalent to conditions (9)-(12). Let us check that the solution of the moment problem is unique. Consider the following transformation T : (x1, x2) ∈ R2 7→ (ϕ, ψ) ∈ [0, 2π) × [0, 2π), eiϕ = x1 + i x1 − i , eiψ = x2 + i x2 − i ; ν(∆) = µ(T −1(∆)), ∆ ∈ B([0, 2π) × [0, 2π)). (48) (49) and set Since T is a bijective continuous transformation, then ν is a non-negative Borel measure on [0, 2π) × [0, 2π). Moreover, we have x1 − i(cid:19)k(cid:18) x2 + i x2 − i(cid:19)l u0,k,−k;0,l,−l =ZR2(cid:18) x1 + i for all k, l ∈ Z. Let eµ be another solution of the moment problem (2) andeν dµ =Z[0,2π)×[0,2π) be defined by eikϕeilψdν, (50) (51) eν(∆) =eµ(T −1(∆)), ∆ ∈ B([0, 2π) × [0, 2π)). 10 By relation (50) we obtain that Z[0,2π)×[0,2π) eikϕeilψdν =Z[0,2π)×[0,2π) eikϕeilψdeν, k, l ∈ Z. (52) By the Weierstrass theorem we can approximate ϕm and ψn, for some fixed m, n ∈ Z+, by trigonometric polynomials Pk(ϕ) and Rk(ψ), respectively: 1 k , ψ∈[0,2π)ψm − Rk(ψ) ≤ max 1 k , k ∈ N. (53) Then ϕ∈[0,2π)ϕm − Pk(ϕ) ≤ max (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z[0,2π)×[0,2π) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z[0,2π)×[0,2π) ϕmψndν −Z[0,2π)×[0,2π) (ϕm − Pk(ϕ))ψndν +Z[0,2π)×[0,2π) ψ∈[0,2π)ψm ν([0, 2π)) + max ≤ max 1 k ϕ∈[0,2π)Pk(ϕ) 1 k Pk(ϕ)Rk(ψ)dν(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Pk(ϕ)(ψn − Rk(ψ))dν(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ν([0, 2π)) as k → ∞. Hence, we conclude that (54) as k → ∞. In the same manner we get ≤ max ψ∈[0,2π)ψm 1 k (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Z[0,2π)×[0,2π) Z[0,2π)×[0,2π) ν([0, 2π)) → 0, k k + max ν([0, 2π)) +(cid:18) 1 ϕmψndeν −Z[0,2π)×[0,2π) ϕmψndν =Z[0,2π)×[0,2π) ϕ∈[0,2π)ϕm(cid:19) 1 Pk(ϕ)Rk(ψ)deν(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) → 0, ϕmψndeν, m, n ∈ Z+. Since the two-dimensional moment problem on a rectangular has a unique solution, we get ν =eν and µ =eµ. ✷ 3 An algorithm towards solving the two-dimensional moment problem. As a first application of our results on the extended two-dimensional moment problem we get the following theorem. 11 Theorem 3 Let the two-dimensional moment problem (1) be given. The moment problem has a solution if and only if there exists a sequence of com- plex numbers {um,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, which satisfies conditions (3), (9)- (12) and (55) um,0,0;n,0,0 = sm,n, m, n ∈ Z+. The proof is obvious and left to the reader. Let the two-dimensional moment problem (1) be given. As it is well known (and can be checked in the same manner as for the relation (3)) the necessary condition for its solvability is the following: Xm,n,m′,n′∈Z+ αm,nαm′,n′sm+m′,n+n′ ≥ 0, (56) for arbitrary complex coefficients αm,n, where all but finite number of αm,n are zeros. We assume that the condition (56) holds. Repeating arguments of the proof of Theorem 1 we can state that there exist a Hilbert space H0 and a sequence {hm,n}m,n∈Z+ such that (hm,n, hm′,n′)H0 = sm+m′,n+n′, m, n, m′, n′ ∈ Z+. Consider the following Hilbert space: (57) (58) H = H0 ⊕ ∞Mj=1 Hj , where Hj are arbitrary one-dimensional Hilbert spaces, j ∈ N. We shall call it the model space for the two-dimensional moment problem. Introduce an arbitrary indexation in the set Ω′ by the unique positive integer index j: j ∈ N 7→ w = w(j) = (m, k, l; n, r, t)(j) ∈ Ω′. (59) Suppose that the two-dimensional moment problem (1) has a solution µ. Consider the space L2 µ and the following subsets in L2 µ: Lµ,0 = Lin{xm 1 xn 2}m,n∈Z+ , Hµ,0 = Lµ,0. We denote Notice that ym,n := xm 1 xn 2 , m, n ∈ Z+. (ym,n, ym′,n′)L2 µ = sm+m′,n+n′, 12 (60) (61) (62) for all m, n, m′, n′ ∈ Z+. We shall also use the notations from (36),(37). Define the following numbers um,k,l;n,r,t :=ZR2 xm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)tdµ, (m, k, l; n, r, t) ∈ Ω. (63) For these numbers conditions (3) hold and repeating arguments after the relation (43) we construct a Hilbert space H and a sequence of elements {xm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H, such that relation (4) holds. We introduce the operator W as after (44). The operator W maps Hµ onto H. Set H0 = span{xm,0,0;n,0,0}m,n∈Z+ ⊆ H. (64) Let us construct a sequence of Hilbert spaces Hj, j ∈ N, in the following way. Step 1. We set f1 = xw(1) − P H H0 xw(1), H1 = span{f1}, where w(·) is the indexation in the set Ω′. Step r, with r ≥ 2. We set fr = xw(r) − P H H0⊕(⊕1≤t≤r−1Ht)xw(1), Hr = span{fr}. Then we get a representation Observe that Hj is either a one-dimensional Hilbert space or Hj = {0}. We denote (68) Λµ = {j ∈ N : Hj 6= {0}}, Λ′ µ = N\Λµ. Then We shall construct a unitary operator U which maps H onto the following subspace of the model space H: (65) (66) (67) (69) (70) H = H0 ⊕ ∞Mj=1 Hj . H = H0 ⊕Mj∈Λµ bH = H0 ⊕Mj∈Λµ Hj . Hj ⊂ H. 13 Choose an arbitrary element x = Xm,n∈Z+ αm,nxm,0,0;n,0,0 + Xj∈Λµ βj fj kfjkH , with αm,n, βj ∈ C. Set U x = Xm,n∈Z+ αm,nhm,n + Xj∈Λµ βj ej , (71) (72) where ej ∈ Hj, kejkH = 1, are chosen arbitrarily. Let us check that this definition is correct. Suppose that x has another representation: (73) fj kfjkH , Xm,n∈Z+ = Xm,n,m′,n′∈Z+ with eαm,n,eβj ∈ C. By orthogonality we have βj = eβj, j ∈ N. Then x = Xm,n∈Z+eαm,nxm,0,0;n,0,0 + Xj∈Λµeβj 0 =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (αm,n −eαm,n)xm,0,0;n,0,0(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (αm,n −eαm,n)(αm′,n′ −eαm′,n′)(xm,0,0;n,0,0, xm′,0,0;n′,0,0)H = Xm,n,m′,n′∈Z+ (αm,n −eαm,n)(αm′,n′ −eαm′,n′)(hm,n, hm′,n′)H =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (αm,n −eαm,n)hm,n(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)H Xm,n∈Z+ Thus, the operator U is defined correctly. If bx ∈ H and bx = Xm,n∈Z+bαm,nxm,0,0;n,0,0 + Xj∈Λµbβj fj kfjkH , 2 H . where bαm,n,bβj ∈ C, then (x,bx)H = Xm,n,m′,n′∈Z+ αm,nbαm′,n′(xm,0,0;n,0,0, xm′,0,0;n′,0,0)H + Xj∈Λµ βjbβj 14 = Xm,n,m′,n′∈Z+ αm,nbαm′,n′(hm,n, hm′,n′)H + Xj∈Λµ = (U x, Ubx)H. βjbβj By continuity we extend U to a unitary operator which maps H onto bH. Then the operator U W is a unitary operator which maps Hµ onto bH. We could define this operator directly, but we prefer to underline an abstract structure of the corresponding spaces and this maybe explains where the model space comes from. We set hm,k,l;n,r,t := U W ym,k,l;n,r,t, (m, k, l; n, r, t) ∈ Ω. Observe that hm,0,0;n,0,0 = hm,n, m, n ∈ Z+; U W Hµ,0 = H0. (74) (75) (76) (77) Since then xw(r) ∈ H0 ⊕ Mj∈Λµ: j≤r Hj , Hj , hw(r) ∈ H0 ⊕ Mj∈Λµ: j≤r r ∈ N. Observe that {ym,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfy relations (17)-(20) (with y in- stead of x). Therefore {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfy relations, as well. Notice that (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (ym,k,l;n,r,t, ym′,k′,l′;n′,r′,t′)L2 µ = um+m′,k+l′,l+k′;n+n′,r+t′,t+r′, (78) for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. Theorem 4 Let the two-dimensional moment problem (1) be given. Choose an arbitrary model space H with a sequence {hm,n}m,n∈Z+, satisfying (57) and fix it. The moment problem has a solution if and only if there exists a sequence {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H such that the following conditions hold: 1) hm,0,0;n,0,0 = hm,n, m, n ∈ Z0; 15 2) hw(r) ∈ H0 ⊕(cid:16)Lj∈Λ: j≤r Hj(cid:17) , Here w(r) is an arbitrary indexation in Ω′ by indices from N. r ∈ N, for some subset Λ ⊆ N. 3) The sequence {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω satisfies conditions (17)-(20) (with h instead of x). 4) There exists a complex function ϕ(m, k, l; n, r, t), (m, k, l; n, r, t) ∈ Ω, such that (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = ϕ(m+m′, k+l′, l+k′; n+n′, r+t′, t+r′), for all (m, k, l; n, r, t), (m′, k′, l′; n′, r′, t′) ∈ Ω. The necessity of conditions 1)-4) for the solvability of the two- Proof. dimensional moment problem was established before the statement of the Theorem. Let conditions 1),3),4) be satisfied. Consider the extended two-dimensional moment problem (2) with um,k,l;n,r,t := ϕ(m, k, l; n, r, t), (m, k, l; n, r, t) ∈ Ω, (79) where ϕ is from the condition 4). Then (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω = (m,k,l;n,r,t),(m′,k′,l′;n′,r′,t′)∈Ω αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′ um+m′,k+l′,l+k′;n+n′,r+t′,t+r′ αm,k,l;n,r,tαm′,k′,l′;n′,r′,t′(hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H X X =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) X(m,k,l;n,r,t)∈Ω αm,k,l;n,r,thm,k,l;n,r,t(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ≥ 0, 2 H for arbitrary complex coefficients αm,k,l;n,r,t, where all but finite number of αm,k,l;n,r,t are zeros. By conditions 3) and 4) we conclude that conditions (9)-(12) hold. By Theorem 2 we obtain that there exists a non-negative Borel measure µ in R2 such that (2) holds. In particular, using conditions 4),1) we get ZR2 1 xn xm 2 dµ = um,0,0;n,0,0 = ϕ(m, 0, 0; n, 0, 0) = (hm,0,0;n,0,0, h0,0,0;0,0,0)H = (hm,n, h0,0)H = sm,n, m, n ∈ Z+. 16 ✷ Observe that condition 2) can be removed from the statement of Theo- rem 4. However, it will be used later. Denote a set of sequences {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, in H satisfying con- ditions 1)-4) by X = X(H). As we have seen in the proof of Theorem 4, for an arbitrary {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H), the unique solution of the extended two-dimensional moment problem with moments (79) gives a solu- tion of the two-dimensional moment problem. Observe that all solutions of the two-dimensional moment problem can be constructed in this manner. In- deed, let µ be an arbitrary solution of the two-dimensional moment problem. Repeating arguments from relation (60) till the statement of Theorem 4 we may write ZR2 xm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)tdµ = (ym,k,l;n,r,t, y0,0,0;0,0,0)L2 µ = (U W ym,k,l;n,r,t, U W y0,0,0;0,0,0)H = (hm,k,l;n,r,t, h0,0,0;0,0,0)H = ϕ(m, k, l; n, r, t) = um,k,l;n,r,t, for all (m, k, l; n, r, t) ∈ Ω. Here the operators U ,W , the sequence {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω, the function ϕ(m, k, l; n, r, t) and the moments um,k,l;n,r,t, of course, depend on the choice of µ. Notice that {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H). For such constructed parameters, the measure µ is a solution of the extended two-dimensional moment problem considered in the proof of Theorem 4. Since the solution of this moment problem is unique, µ will be reconstructed in the above described manner. Notice that condition 4) of Theorem 4 is equivalent to the following conditions: (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (h em,k,l;n,r,t, h em′,k′,l′;n′,r′,t′)H, (80) (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,ek,l;n,r,t, hm′,k′,el′;n′,r′,t′)H, if m, m′,em,em′, n, n′ ∈ Z+, k, l, r, t, k′, l′, r′, t′ ∈ Z: m + m′ = em + em′; if m, m′, n, n′ ∈ Z+, k,ek, l, r, t, k′, l′,el′, r′, t′ ∈ Z: k + l′ =ek +el′; if m, m′, n, n′ ∈ Z+, k, l,el, r, t, k′,ek′, l′, r′, t′ ∈ Z: l + k′ =el +ek′; (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,el;n,r,t, hm′,ek′,l′;n′,r′,t′)H, (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,l;en,r,t, hm′,k′,l′;en′,r′,t′)H, (81) (82) (83) 17 (84) (85) (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,l;n,er,t, hm′,k′,l′;n′,r′,et′)H, if m, m′, n, n′,en,en′ ∈ Z+, k, l, r, t, k′, l′, r′, t′ ∈ Z: n + n′ =en +en′; if m, m′, n, n′ ∈ Z+, k, l, r,er, t, k′, l′, r′, t′,et′ ∈ Z: r + t′ =er +et′; if m, m′, n, n′ ∈ Z+, k, l, r, t,et, k′, l′, r′,er′, t′ ∈ Z: t + r′ =et +er′. (hm,k,l;n,r,t, hm′,k′,l′;n′,r′,t′)H = (hm,k,l;n,r,et, hm′,k′,l′;n′,er′,t′)H, Let {gn}∞ As we can see, the solving of the two-dimensional moment problem re- duces to a construction of the set X(H). Let us describe an algorithm for a construction of sequences from X(H). n=1 be an arbitrary orthonormal basis in H0 obtained by the Gram-Schmidt orthogonalization procedure from the sequence {hm,n}m,n∈Z+ indexed by a unique index. Choose an arbitrary j ∈ N. Let w(j) = (m, k, l; n, r, t)(j) ∈ Ω′. If we had constructed {hm,n}m,n∈Z+ ∈ X(H), then the two-dimensional moment problem has a solution µ and H = kxm 1 (x1 + i)k(x1 − i)lxn 2 (x2 + i)r(x2 − i)tk2 L2 µ dj := khw(j)k2 =ZR2 x2m 1 (x2 1 + 1)k(x2 1 + 1)lx2n 2 (x2 2 + 1)r(x2 2 + 1)tdµ. Therefore dj are bounded by some constants Mj = Mj(S) depending on the prescribed moments S := {sm,n}m,n∈Z+ . (Notice that e.g. (x2 1 + 1)l ≤ 1, for l < 0, and for non-negative m, k, l; n, r, t the values of dj are determined uniquely). Step 0. We set hm,0,0;n,0,0 = hm,n, m, n ∈ Z0. (86) We check that conditions (13)-(16) (with h instead of x) and (80)-(85) are satisfied for hm,0,0;n,0,0, m, n ∈ Z0. If they are not satisfied, the two- dimensional moment problem has no solution and we stop the algorithm. Step 1. We seek for hw(1) in the following form: hw(1) = ∞Xn=1 α1;ngn + β1e1, (87) with some complex coefficients α1;n, β1. 18 Conditions (13)-(16) (with h instead of x) and (80)-(85) which include hw(1) and the already constructed hm,k,l;n,r,t are equivalent to a set L1 of linear equations with respect to α1;n, n ∈ N , and d1 = khw(1)k2 H. Notice that they depend on β1 only by d1. Denote the set of solutions of these equations by S1 = {(α1;n, n ∈ N ; d1) : equations from L1 are satisfied} . (88) Set Finally, we set bS1 =((α1;n, n ∈ N ; d1) ∈ S1 : G1 = α1;ngn + eiθ d1 − ∞Xn=1 ∞Xn=1 ∞Xn=1 α1;n2! 1 2 hw(r) = ∞Xn=1 αr;ngn + (89) α1;n2 ≤ d1, d1 ≤ M1) . e1 : (α1;n, n ∈ N ; d1) ∈ bS1, θ ∈ [0, 2π) rXj=1 βj ej , (91) (90) . Step r, with r ≥ 2. We seek for hw(r) in the following form: with some complex coefficients αr;n, βj. Conditions (13)-(16) (with h instead of x) and (80)-(85) which include hw(r) and the already constructed hm,k,l;n,r,t are equivalent to a set Lr of linear equations with respect to αr;n, n ∈ N , βj, 1 ≤ j ≤ r− 1, and dr = khw(r)k2 H, and depending on parameters (hw(1), hw(2), ..., hw(r−1)) ∈ Gr−1 . Notice that these linear equations depend on βr only by dr. Denote the set of solutions of these equations by (hw(1), hw(2), ..., hw(r−1)) ∈ Gr−1, Sr =(cid:8)(αr;n, n ∈ N ; βj , 1 ≤ j ≤ r − 1; dr; hw(1), hw(2), ..., hw(r−1)) : and equations from Lr with parameters (hw(1), hw(2), ..., hw(r−1)), are satisfied(cid:9) . bSr =(cid:8)(αr;n, n ∈ N ; βj , 1 ≤ j ≤ r − 1; dr; hw(1), hw(2), ..., hw(r−1)) ∈ Sr : (92) Set (93) ∞Xn=1 αr;n2 + r−1Xj=1 βr;j2 ≤ dr, dr ≤ Mr . 19 αr;ngn + r−1Xj=1 ∞Xn=1 1 2 Finally, we set Gr =(cid:8)(cid:0)hw(1), hw(2), ..., hw(r−1), βr;jej + eiθdr − r−1Xj=1 αr;n2 − er : ∞Xn=1 (αr;n, n ∈ N ; βj , 1 ≤ j ≤ r − 1; dr; hw(1), hw(2), ..., hw(r−1)) ∈ bSr, θ ∈ [0, 2π)o . βr;j2 Final step. Consider a space H of sequences (94) h = (h1, h2, h3, ...), hr ∈ H, r ∈ N, with the norm given by khkH = sup r∈N 1 √Mr khrkH < ∞. (95) (96) For arbitrary (h1, ..., hr) ∈ Gr, we put into correspondence elements h ∈ H of the following form h = (h1, ..., hr, gr+1, gr+2, ...) : gj ∈ H, kgjkH ≤pMj , j > r. Thus, the set Gr is mapped onto a set Gr ⊂ H. Observe that all elements of Gr has the norm less or equal to 1. Set (97) G = Gr. ∞\r=1 (98) If G 6= ∅, then to each (g1, g2, ...) ∈ G, we put into correspondence a se- quence H = {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω such that (86) holds and hw(r) := gr, r ∈ N. (99) We state that H ∈ X(H). In fact, conditions (13)-(16) (with h instead of x) and (80)-(85) are satisfied for hm,0,0;n,0,0, m, n ∈ Z+, by Step 0. If one of these equations include hw(r) with r ≥ 1, then we choose the maximal ap- pearing index r. Since (hw(1), ..., hw(r)) ∈ Gr, then this equation is satisfied. Condition 2) is satisfied by the construction. Thus, if G 6= ∅, then using H we can construct a solution of the two- dimensional moment problem in the described above manner. 20 Theorem 5 Let the two-dimensional moment problem (1) be given. Choose an arbitrary model space H with a sequence {hm,n}m,n∈Z+, satisfying (57) and fix it. The moment problem has a solution if and only if conditions (13)- (16) (with h instead of x) and (80)-(85) are satisfied for hm,0,0;n,0,0 := hm,n, m, n ∈ Z+, and (100) G 6= ∅, where G is constructed by (98) according to the algorithm. If the latter conditions are satisfied then to each (g1, g2, ...) ∈ G, we put into correspondence a sequence H = {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω such that (86) holds and (101) This sequence belongs to X(H) and the unique solution of the extended two-dimensional moment problem with moments (79) gives a solution µ of the two-dimensional moment problem. Moreover, all solutions of the two- dimensional moment problem can be obtained in this way. hw(r) := gr, r ∈ N. Proof. The sufficiency of the conditions in the statement of the Theorem for the solvability of the two-dimensional moment problem was shown be- fore the statement of the Theorem. Let us show that these conditions are necessary. Let µ be a solution of the two-dimensional moment problem. By Theorem 4 stead of x) and (80)-(85) are satisfied. In particular, they are satisfied for the set X(H) is not empty. Choose an arbitrarybH = {bhm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ X(H). By conditions 3),4) we see that conditions (13)-(16) (with bh in- bhm,0,0;n,0,0 = hm,n, m, n ∈ Z+. Comparing condition 2) with Steps 1 and r for r ≥ 2, we see that Therefore elements (bhw(1), ...,bhw(r)) ∈ Gr, r ∈ N. (102) (103) (104) where gj ∈ H : kgjkH ≤pMj, for j > r are arbitrary. Thus, the element (bhw(1), ...,bhw(r), gr+1, gr+2, ...) ∈ Gr, bh := (bhw(1),bhw(2),bhw(3), ...) ∈ Gr, r ∈ N, r ∈ N. and Gr = G. (105) bh ∈ \r∈N 21 Therefore G 6= ∅. If the conditions of the Theorem are satisfied then to each g = (g1, g2, ...) ∈ G, we put into correspondence a sequence H = H(g) = {hm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω such that (86) and (101) hold. Then H ∈ X(H), as it was shown before the statement of the Theorem. The sequence H generates a solution of the extended two-dimensional moment problem and of the two-dimensional mo- ment problem, see considerations after the proof of Theorem 4. It remains to show that all solutions of the two-dimensional moment problem can be obtained in this way. Since elements of X(H) generate all solutions of the two-dimensional moment problem (see considerations after the proof of Theorem 4), it remains to prove that {H(g) : g ∈ G} = X(H). (106) On the other hand, choose an arbitrary eH = {ehm,k,l;n,r,t}(m,k,l;n,r,t)∈Ω ∈ Denote the set on the left-hand side by X1. It was shown that X1 ⊆ X(H). X(H). Repeating the construction at the beginning of this proof we obtain that (107) Observe that eh := (ehw(1),ehw(2),ehw(3), ...) ∈ G. H(eh) =eH. Therefore X(H) ⊆ X1 and relation (106) holds. ✷ Remark. The truncated two-dimensional moment problem can be consid- ered in a similar manner. Moreover, the set of indices of the known elements hm,n will be finite in this case and therefore equations in the r-th step of the algorithm will form finite systems of linear equations. Thus, the r-th step could be easily performed using computer. (108) References [1] J. A. Shohat, J. D. Tamarkin, The Problem of Moments, Amer. Math. Soc., New York City, 1943. [2] N. I. Akhiezer, Classical Moment Problem, Fizmatlit., Moscow, 1961. (in Russian). [3] Ju. M. Berezanskii, Expansions in Eigenfunctions of Selfadjoint Op- erators, Amer. Math. Soc., Providence, RI, 1968. (Russian edition: Naukova Dumka, Kiev, 1965). 22 [4] B. Fuglede, The multidimensional moment problem, Expo. Math. 1 (1983), 47 -- 65. [5] Yu. M. Berezansky, Spectral theory of the infinite block Jacobi type nor- mal matrices, orthogonal polynomials on the complex domain, and the complex moment problem, Operator Theory: Advances and Applica- tions 191 (2009), 37 -- 50. [6] J. Stochel, F. H. Szafraniec, The complex moment problem and subnor- mality: a polar decomposition approach, J. of Functional Analysis 159 (1998), 432 -- 491. [7] S. M. Zagorodnyuk, Positive definite kernels satisfying difference equa- tions, Methods of Functional Analysis and Topology 16 (2010), 83 -- 100. [8] N. I. Akhiezer, I.M. Glazman, Theory of Linear Operators in a Hilbert Space, Gos. izdat. teh.-teor. lit., Moscow, Leningrad, 1950. (in Russian). On the two-dimensional moment problem. S.M. Zagorodnyuk In this paper we obtain an algorithm towards solving the two-dimensional moment problem. This algorithm gives the necessary and sufficient condi- tions for the solvability of the moment problem. It is shown that all solutions of the moment problem can be constructed using this algorithm. 23
1809.04829
1
1809
2018-09-13T08:25:56
Weighted composition operators on the Fock space
[ "math.FA" ]
In this paper, we study weighted composition operators on the Fock space. We show that a weighted composition operator is cohyponorma if and only if it is normal. Moreover, we give a complete characterization of closed range weighted composition operators. Finally, we find norms of some weighted composition operators.
math.FA
math
Weighted composition operators on the Fock space Mahsa Fatehi September 14, 2018 Abstract In this paper, we study weighted composition operators on the Fock space. We show that a weighted composition operator is cohyponorma if and only if it is normal. Moreover, we give a complete characteriza- tion of closed range weighted composition operators. Finally, we find norms of some weighted composition operators. 1 1 Introduction Let D denote the open unit disk in the complex plane C. For an analytic map ϕ, let ϕ0 be the identity function, ϕ1 = ϕ and ϕn+1 = ϕ◦ ϕn for n = 1, 2, .... We call them the iterates of ϕ. It is well-known that if ϕ, neither the identity nor an elliptic automorphism of D (i.e., ϕ is an automorphism of D with a fixed point in D), is an analytic map on the unit disk into itself, then there exists a point w in D such that ϕn converges to w uniformly on compact subsets of D. The point w is called the Denjoy-Wolff point of ϕ. The Denjoy- Wolff point w is the unique fixed point of ϕ in D so that ϕ′(w) ≤ 1 (see [7]). Recall that the Fock space F 2 is a Hilbert space of all entire functions on C that are square integrable with respect to the Gaussian measures dµ(z) = π−1e−z2 dA(z), where dA is the usual Lebesgue measure on C. The Fock space F 2 is a reproducing kernel Hilbert space with inner product hf, gi =ZC f (z)g(z)dµ(z) 1AMS Subject Classifications. Primary 47B33. Key words and phrases: Fock space, Weighted composition operator, Normaloid, Cohy- ponormal, Norm, Closed range. 1 (see [16]). and reproducing kernel function Kw(z) = ewz for any w ∈ C. Note that for any w ∈ C, kKwk = ew2/2. For each w ∈ C, we define the normalized reproducing kernel as kw(z) = Kw(z) = ewz−w2/2. For each nonnegative kKwk integer n, let en(z) = zn/√n!. The set {en} is an orthonormal basis for F 2 For entire function ϕ on C, the composition operator Cϕ on F 2 is defined as Cϕ(f ) = f ◦ ϕ for any f ∈ F 2; moreover, for ψ ∈ F 2, the weighted composition operator Cψ,ϕ is defined by Cψ,ϕf = ψ · (f ◦ ϕ). There exists some literature on composition operators acting on the Hardy and Bergman spaces. The books [7] and [11] are the important references. Carswell et al. [4] characterized the bounded and compact composition operators on the Fock space over Cn. Specified to the one-dimensional case, they stated that Cϕ is bounded if and only if ϕ(z) = az, where a = 1 or ϕ(z) = az + b with a < 1. In [12], Ueki found the criteria to charac- terize the boundedness and compactness of weighted composition operators on the Fock space. After that in [10], Le obtained much easier criteria for the boundedness and compactness of weighted composition operators on the Fock space. On the Hardy space, normal weighted composition opera- tors were studied; moreover, unitary weighted composition operators were characterized (see [3]). Also in [6], cohyponormal weighted composition op- erators were obtained. After that in [8], hyponormal weighted composition operators were investigated on the Hardy and weighted Bergman spaces. Unitary weighted composition operators Cψ,ϕ on the Fock space were ob- tained in [13]. Also invertible weighted composition operators on the Fock space were characterized in [14]. In the second section, we find normaloid, hyponormal and cohyponormal composition operators. After that we obtain all hyponormal weighted composition operators Cψ,ϕ, where ψ = Kc for each c ∈ C. Moreover, we study a class of normaloid weighted composition oper- ators. Next, we show that for ϕ(z) = az + b, Cψ,ϕ is cohyponormal if and . Closed range composition operators were studied only if ψ = ψ(0)Kb a−1 on the Hardy and weighted Bergman spaces in [1], [9] and [17]. In the third section, we characterize closed range weighted composition operators on the Fock space. In the fourth section, we find norm of Cψ,ϕ on F 2, when ψ = Kc for any c ∈ C. a−1 2 2 Normaloid weighted composition operators Suppose that T is a bounded operator on a Hilbert space. Throughout this paper, the spectrum of T , the essential spectrum of T , and the point spectrum of T are denoted by σ(T ), σe(T ), σp(T ) respectively. Also the spectral radius of T is denoted by r(T ). Le [10] studied the boundedness of weighted composition operator on the Fock space. His result shows that if Cψ,ϕ is bounded on F 2, then ϕ(z) = az + b, where a ≤ 1; furthermore, he proved that if a = 1 and Cψ,az+b is bounded on F 2, then ψ = ψ(0)K−ab. We use these facts frequently in this paper and so throughout this paper, we assume that ϕ(z) = az + b, where a ≤ 1. [4], [13] and [15] were written on another Fock space (see [16, p. 33]), but their results hold for F 2 which is considered in this paper, with identical arguments. In the following proposition, we investigate σp(Cψ,ϕ), when ϕ(z) = az +b and a < 1. Note that in the case that a = 1, as we saw in the preceding paragraph, ψ = ψ(0)K−ab. Then Cψ,ϕ is a constant multiple of a unitary operator from [13, Corollary 1.2]. Moreover, in [13, Corollary 1.4], the spec- trum of unitary weighted composition operators were characterized. Proposition 2.1. Let ψ and ϕ be entire functions on C and Cψ,ϕ be a bounded weighted composition operator on F 2. Suppose that ϕ(z) = az + b, where a < 1 and b ∈ C. If λ ∈ σp(Cψ,ϕ), then λ ≤ ψ( b 1−a ). Moreover, if ψ( b 1−a ) = 0 and ϕ and ψ are nonconstant, then Cψ,ϕ has no eigenvalues. Proof. Suppose that a < 1. First we find a weighted composition operator C eψ, eϕ which is unitary equivalent to Cψ,ϕ such that the fixed point of eϕ lies in D and eϕ is a self-map of D. There exists a positive integer N such that a + 1 N 1 − a < 1. It is not hard to see that there is a complex number u ∈ C such that u + b N . By [13, Corollary 1.2], we know that Cku,z−u is unitary (the operator Cku,z−u is known as the Weyl unitary). By [10, Proposition 3.1], C∗ku,z−u = Ck−u,z+u, so 1−a < 1 Cku,z−uCψ,ϕCk−u,z+u = = = 1 1 kKuk2 Ceuz,z−uCψ,ϕCe−uz,z+u kKuk2 euz · ψ(z − u) · e(−u(az+b))◦(z−u)C(z+u)◦(az+b)◦(z−u) kKuk2 euz · ψ(z − u) · e−u(az−au+b)Caz+u(1−a)+b. (1) 1 3 Let eϕ(z) = az + u(1 − a) + b and 1 eψ(z) = kKuk2 euz · ψ(z − u) · e−u(az−au+b). (2) It is easy to see that the fixed point of eϕ is u + b 1−a which belongs to D. Because a + u(1 − a) + b < 1, eϕ is a self-map of D. Since Cψ,ϕ is unitary equivalent to C eψ, eϕ, σp(Cψ,ϕ) = σp(C eψ, eϕ). Assume that λ is a nonzero eigenvalue for C eψ, eϕ with corresponding eigenvector h. We obtain for each z ∈ C and positive integer n. For any fixed point z ∈ C, we obtain λnh(z) = Πn−1 j=0 eψ(eϕj(z))h(eϕn(z)) h(eϕn(z)) = hh ◦ eϕn, Kzi ≤ kh ◦ eϕnkkKzk = kh ◦ eϕnke = kC eϕn(h)ke ≤ kC eϕnkkhke z2 2 . z2 2 z2 2 (3) (4) 1−a is the Denjoy-Wolff point of eϕ, eϕj(z) → u + b Since h is not the zero function, we can choose z ∈ D such that h(z) 6= 0. Since u + b 1−a and eψ(eϕj(z)) → eψ(u + b 1−a ) as j → ∞. Take n-th roots of the absolute value each side of Equation (3), use Equation (4) and let n → ∞, we get λ ≤ eψ(u + b 1−a )r(Cϕ). From Equation (2) and [15, Theorem 1.1], we see that λ ≤ eψ(u + b 1−a )r(Cϕ) = ψ( b 1−a ). Then b ). (5) λ ≤ ψ( 1−a ) = ψ( b 1 − a Now assume that eψ(u + b 1−a ) = 0 and ϕ and ψ are nonconstant. Thus, by Equation (5), λ = 0 is the only possible eigenvalue for Cψ,ϕ. Since ψ is not the zero function and ϕ is not constant, the Open Mapping Theo- rem implies that 0 cannot be an eigenvalue for Cψ,ϕ (see idea of the proof of [2, Lemma 4.1]). (cid:3) Suppose that T is a bounded operator. The operator T is hyponor- mal (cohyponormal) if T ∗T ≥ T T ∗ (T ∗T ≤ T T ∗). Also T is normaloid if kTk = r(T ). It is well known that hyponormal (cohyponormal) operators 4 are normaloid. In the following proposition, we characterize hyponormal, cohyponormal and normaloid composition operators on F 2. Proposition 2.2. Let ϕ(z) = az + b, where a ≤ 1 and b ∈ C. Then Cϕ is a bounded normaloid (hyponormal or cohyponormal) operator if and only if b = 0. 1 2 b2 Proof. Let Cϕ be normaloid (hyponormal or cohyponormal). Suppose that b 6= 0. Then [4, Theorem 1] states that a < 1. By [4, Theorem 4], 1−a2 (note that the inner product for F 2 in this paper is different kCϕk = e from [4], so the norm of a composition operator is not exactly the same as [4, Theorem 4]; furthermore, in Remark 4.1, kCϕk will be described). Also [15, Theorem 1.1] implies that r(Cϕ) = 1. Since Cϕ is normaloid (hyponormal b2 1−a2 = 1. Hence b = 0 which is a contradiction. or cohyponormal), e Conversely, suppose that ϕ(z) = az, where a ≤ 1. Invoking [4, Lemma 2], C∗ϕ = Caz. Then Caz is normal and the result follows. (cid:3) 1 2 Suppose that Cψ,ϕ is a bounded weighted composition operator and ϕ(z) = az + b. Note that if a = 1, then from [13, Corollary 1.2] and as we saw in the second paragraph of this section, Cψ,ϕ is a constant multiplie of a unitary operator. Thus, Cψ,ϕ is normal, normaloid, hyponormal and cohyponormal. Hence we state the following proposition for a 6= 1. Proposition 2.3. Let ψ = Kc and ϕ(z) = az + b, where a ≤ 1, a 6= 1 and b, c ∈ C. Suppose that Cψ,ϕ is bounded on F 2. Then the following are equivalent. (a) Cψ,ϕ is hyponormal. (b) Cψ,ϕ is cohyponormal. (c) Cψ,ϕ is normaloid. (d) c = b a−1 a−1 . e−u2 We can see that Cψ,ϕ is unitarily equivalent to ψ( b Proof. There is u ∈ C such that c = u(a − 1). By Equation (1) and some calculation, Cψ,ϕ is unitarily equivalent to C eψ, eϕ, where eψ(z) = euz · ψ(z − u) · e−u(az−au+b) = ψ( b 1−a ) and eϕ(z) = az + u(1 − a) + b. (a) ⇒ (d). Suppose that Cψ,ϕ is hyponormal. Then Caz+u(1−a)+b is hyponormal. Proposition 2.2 implies that u(1 − a) + b = 0. Since u = c a−1 , we conclude that c = b a−1 a−1 . 1−a )Caz+u(1−a)+b. 5 (d) ⇒ (a). Assume that c = b a−1 a−1 . Let u = b and some calculation, Cψ,ϕ is unitarily equivalent to e az. We infer from Proposition 2.2 that C eϕ is hyponormal and so Cψ,ϕ is hyponormal. a−1 . By Equation (1) 1−a C eϕ, where eϕ(z) = b2 (b) ⇔ (d). The idea of the proof is similar to (a) ⇔ (d). (c) ⇔ (d). The idea of the proof is similar to (a) ⇔ (d). Suppose that ϕ(z) = az + b, where a = 1 and ψ is an entire function. If Cψ,ϕ is bounded on F 2, then as we saw in the second paragraph of this section, ψ(z) = ψ(0)K−β , where β = ab. Hence Cψ,ϕ is a constant multiple of the unitary operator (see [13, Corollary 1.2]). It shows that in this case Cψ,ϕ is normaloid and so in the next theorem, we assume that a < 1. (cid:3) Theorem 2.4. Suppose that ψ is an entire function and is not identi- cally zero. Assume that ϕ(z) = az + b, where a < 1 and b ∈ C. Let for each λ ∈ σe(Cψ,ϕ), λ ≤ ψ( b 1−a ). Then Cψ,ϕ is normaloid if and only if ψ = ψ(0)Kb a−1 . a−1 Proof. Let p = b 1−a (note that it is obvious that p is the fixed point of ϕ). By [13, Corollary 1.2], Ckp,z−p is unitary. Furthermore, [10, Proposition 3.1] states that C∗kp,z−p = Ck−p,z+p. We obtain H : = C∗kp,z−pCψ,ϕCkp,z−p = Ck−p,z+pCψ,ϕCkp,z−p = Cq, eϕ, where and eϕ(z) = ϕ(z + p) − p = a(z + p) + b − p = az q(z) = k−p(z)kp(ϕ(z + p))ψ(z + p) = epb+p2(a−1)ep(a−1)zψ(z + p) = ep(a−1)zψ(z + p). (6) (7) (8) It is not Since Cψ,ϕ is unitary equivalent to Cq, eϕ, σe(Cψ,ϕ) = σe(Cq, eϕ). hard to see that q(0) = ψ(p). Now suppose that Cψ,ϕ is normaloid. Then Cq, eϕ is normaloid. Since Cψ,ϕ and Cq, eϕ are unitary equivalent, for each 6 λ ∈ σp(Cq, eϕ), λ ≤ q(0). We infer from [5, Proposition 6.7, p. 210] and [5, Proposition 4.4, p. 359] that r(Cq, eϕ) ≤ q(0). Since Cq, eϕ is normaloid, q(0) ≥ kCq, eϕk ≥ kCq, eϕ(1)k = kqk. We know that { zm : m ≥ 0} is an √m! orthonormal basis for F 2. Then kqk ≥ q(0). It shows that q must be con- stant. Thus, Equation (8) shows that ψ(z) · ep(a−1)(z−p) is constant. Then ψ(z) = ψ(0)e−p(a−1)z = ψ(0)eb a−1 Conversely, suppose that ψ = ψ(0)Kb a−1 . By Equation (8), q is constant. Equations (6) and (7) state that Cψ,ϕ is unitarily equivalent to a constant multiple of Caz. Since by Proposition 2.2, Caz is normaloid, Cψ,ϕ is also normaloid. (cid:3) a−1 z = ψ(0)Kb a−1 (z). a−1 a−1 We know that for each c ∈ C, CKc,ϕ is compact, where ϕ(z) = az + b and a < 1 (see [15, Corollary 2.4 ]). Hence σe(CKc,ϕ) = {0}. Thus, CKc,ϕ satisfies the conditions of Theorem 2.4 and so if CKc,ϕ is normaloid, then c must be b a−1 a−1 (see also Proposition 2.3). Corollary 2.5. Suppose that ψ is an entire function and is not identi- cally zero. Assume that ϕ(z) = az + b, where a < 1 and b ∈ C. Then Cψ,ϕ is compact and normaloid if and only if ψ = ψ(0)Kb a−1 . a−1 Suppose that ψ is an entire function and ϕ(z) = az + b, where a ≤ 1. In the next theorem, we show that Cψ,ϕ is cohyponormal if and only if Cψ,ϕ is normal (see [10, Theorem 3.3] and note that if a = 1 and a 6= 1, then a−1 a−1 = −a). Theorem 2.6. Suppose that ψ is an entire function and ϕ(z) = az + b, a−1 a−1 where a ≤ 1. Then Cψ,ϕ is cohyponormal if and only if ψ = ψ(0)Kb for a 6= 1 and ψ = ψ(0)K−b for a = 1. Proof. We break the proof into two cases. First assume that a = 1. If ψ = ψ(0)K−b, then by [13, Corollary 1.2], Cψ,ϕ is a constant multiple of a unitary operator. Thus, Cψ,ϕ is cohyponormal. Now let Cψ,ϕ be cohyponor- mal. As we saw in the second paragraph of this section ψ = ψ(0)K−b. Now assume that a 6= 1. Suppose that Cψ,ϕ is cohyponormal. Let Cq, eϕ be as in Equation (6), where q and eϕ were obtained in Equations (7) and (8). It is obvious that Cq, eϕ is also cohyponormal. Then kC∗q, eϕK0k ≥ kCq, eϕK0k. We obtain q(0) ≥ kqk. As we saw in the proof of Theorem 2.4, q must be constant, so Equation (8) stats that ψ(z) = ψ(0)eb a−1 a−1 z. 7 The other direction follows easily from Proposition 2.3. (cid:3) 3 Closed range weighted composition operator In this section, we prove that Cψ,ϕ has closed range if and only if Cψ,ϕ is a constant multiple of a unitary operator (see [13, Corollary 1.2 ]). Theorem 3.1. Let ϕ and ψ be entire functions on C such that ψ is not identically zero. Suppose that Cψ,ϕ is bounded on F 2. Then Cψ,ϕ has closed range if and only if and ϕ(z) = az + b, with a = 1, b ∈ C and ψ = ψ(0)K−ab. Proof. First suppose that a = 1 and ψ = (0)K−ab. By [13, Corollary 1.2 ], we have Cψ,ϕ is a constant multiple of a unitary operator. Therefore, Cψ,ϕ has closed range. Conversely, let Cψ,ϕ have closed range on F 2. As we stated in the second paragraph of Section 2, ϕ(z) = az+b, with a ≤ 1. Suppose that a < 1. By Equations (6) and (7), Cψ,ϕ is unitarily equivalent to Cq, eϕ, where eϕ(z) = az, so without loss of generality, we assume that ϕ(z) = az (note that Cψ,ϕ has closed range if and only if Cq, eϕ has closed range). Since Cψ,ϕ is bounded on F 2, Cψ,az(ez) = ψ(z)eaz belongs to F 2. Now we define a bounded linear functional Fψ(z)eaz by Fψ(z)eaz (f ) = hf (z), ψ(z)eazi for each f ∈ F 2. We know that Kw kKwk converges to zero weakly as w → ∞. Then It shows that lim w→∞h Kw kKwk , ψ(z)eazi = 0. ψ(w)eaw ew2/2 lim w→∞ = 0. (9) Now if a = aeiθ, we take w = re−iθ, where r is a positive real number. Then eaw = eaw. Equation (9) shows that r→∞ψ(re−iθ)2ear2−r2 = 0. (10) lim From Equation (10), we obtain lim r→∞kC∗ψ,ϕ Kre−iθ kKre−iθkk2 = lim r→∞ψ(re−iθ)2kereiθazk2 er2 8 = lim r→∞ψ(re−iθ)2er2(a2−1) = 0. (11) Since ψ is not identically zero, it is easy to see that there exits a sequence {rn} such that for any n, rn is a positive real number, rn → ∞ as n → ∞ and Karne−iθ ψ(rne−iθ) 6= 0 for each n. We have C∗ψ,ϕ kKrne−iθk 6= 0 Krne−iθ kKrne−iθk 6∈ Ker(C∗ψ,ϕ). Equation (11) and [5, Proposition 6.1, p. and so 363 ] show that C∗ψ,ϕ does not have closed range. Thus, Cψ,ϕ does not have closed range (see [5, Proposition 6.2, p. 364 ]). Hence a = 1 and the result follows from the second paragraph of Section 2. Krne−iθ kKrne−iθk = ψ(rne−iθ) (cid:3) 4 Norm of weighted composition operator Suppose that ϕ and ψ are entire functions on C and Cψ,ϕ is bounded on F 2. It is well-known that for any w ∈ C, C∗ψ,ϕKw = ψ(w)Kϕ(w). We use this formula in the following remark. Furthermore, in this section for each c ∈ C, MKc is multiplication by the kernel function Kc. Remark 4.1. For ϕ(z) = az + b, where a ≤ 1 and b, c ∈ C, kCϕk was found in [4, Theorem 4]. In [4], the inner product of the Fock space is different from ours. An analogue of [4, Theorem 4] holds for F 2 with our definition. One can follow the outline of the proof of [4, Theorem 4] to find kCϕk on F 2. Moreover, we state another proof for finding kCϕk. We break it into two cases. First assume that ϕ(z) = az+b, where a < 1. By [4, Theorem 2], Cϕ is com- pact. Then [4, Lemma 2] implies that C∗ϕCϕ = MKbCazCaz+b = MKbCa2z+b is compact ( note that by the similar proof which was stated in [4, Lemma 2], we can see that C∗az+b = CKb,az). We know that kCϕk2 = kC∗ϕCϕk = r(MKbCa2z+b). Since MKbCa2z+b is compact, r(MKbCa2z+b) = sup{λ : λ ∈ σp(MKbCa2z+b)} (see [5, Theorem 7.1, p. 214 ]). By Proposition b2 1−a2 . We have 2.1, for each λ ∈ σp(MKbCa2z+b), λ ≤ Kb( (MKbCa2z+b)∗K b . Then e 1−a2 ) = e 1−a2 ∈ σp((MKbCa2z+b)∗) b2 1−a2 K b = e 1−a2 1−a2 b b2 9 b2 1−a2 ∈ σp(MKbCa2z+b). Thus, and so by [5, Theorem 7.1, p. 214 ], e 1 2 b2 1−a2 . kCaz+bk = e Now assume that ϕ(z) = az + b with a = 1. By [4, Theorem 1], b = 0. From [4, Lemma 2], one can easily see that kCϕk2 = kC∗ϕCϕk = kCazCazk = kCa2zk = kCzk = 1. Then in this case kCϕk = 1. In the preceding sections, we saw that among weighted composition op- erators, CKc,ϕ is much important, when c ∈ C. Hence in the following theorem, we try to find kCKc,ϕk. Theorem 4.2. Let ψ = Kc and ϕ(z) = az + b, where a ≤ 1 and b, c ∈ C. Suppose that Cψ,ϕ is bounded on F 2. (a) If a < 1, then kCψ,ϕk = ec b (b) If a = 1 and a 6= 1, then kCψ,ϕk = e (c) If a = 1, then kCψ,ϕk = e 1−ae a−1 1−a2 b2 2 . 1−a. c(1−a) b2 1 2 +b2 . 1 2 c(1−a) +b2 a−1 1−a2 Proof. 1−a )e (see also Remark 4.1). 1−a )Caz+u(1−a)+b, where u = c (a) Assume that a < 1. By the proof of Proposition 2.3, a−1 . Since 1−a )kCaz+u(1−a)+bk = Cψ,ϕ is unitarily equivalent to ψ( b a < 1, [4, Theorem 4] implies that kCψ,ϕk = ψ( b ψ( b (b) Assume that a = 1 and a 6= 1. Again by the proof of Proposition 2.3, kCψ,ϕk = ψ( b 1−a )kCaz+ca+bk. Since a = 1 and CKc,ϕ is bounded, from the second paragraph of Section 2, c = −ab. Then ca + b = 0. Therefore, kCψ,ϕk = ψ( b 1−a ) (see [4, Theorem 4] and Remark 4.1). Thus, kCψ,ϕk = e (c) Assume that a = 1. As we saw in the second paragraph of Section 2, c = −ab = −b. Then ψ(z) = e−bz. Now we must find kCe−bz,z+bk. By [13, Corollary 1.2], e 1−a )kCazk = ψ( b 1−a. 1−a = e a−1 +bk = ψ( b 2 Ce−bz,z+b is unitary. Hence kCψ,ϕk = e 1−a )kC az+ c(1−a) b2 2 . −ab2 −b2 b2 (cid:3) References [1] J. R. Akeroyd and S. R. Fulmer, Closed-range composition operators on weighted Bergman spaces, Integr. Equ. Oper. Theory 72 (2012), 103-114. 10 [2] P. S. Bourdon, Spectra of some composition operators and associated weighted com- position operators, J. Oper. Theory 67 (2) (2012), 537-560. [3] P. S. Bourdon and S. K. Narayan, Normal weighted composition operators on the Hardy space H 2(D), J. Math. Anal. Appl. 367 (2010), 278-286. [4] B. J. Carswell, B. D. MacCluer and A. Schuster, Composition operators on the Fock space, Acta Sci. Math. (Szeged) 69 (2003), 871-887. [5] J. B. Conway, A Course in Functional Analysis, Second Edition, Springer-Verlag, New York, 1990. [6] C. C. Cowen, S. Jung, and E. Ko, Normal and cohyponormal weighted composition operators on H 2, Operator Theory: Advances and Applications 240 (2014), 69-85. [7] C. C. Cowen and B. D. MacCluer, Composition Operators on Spaces of Analytic Functions. Studies in Advanced Mathematics. CRC Press, Boca Raton, FL, 1995. [8] M. Fatehi and M. Haji Shaabani, Norms of hyponormal weighted composition op- erators on the Hardy and weighted Bergman spaces, Operators and Matrices 12(4) (2018), 997-1007. [9] P. Ghatage and M. Tjani, Closed range composition operators on Hilbert function spaces, J. Math. Anal. Appl. 431 (2) (2015), 841-866. [10] T. Le, Normal and isometric weighted composition operators on the Fock space, Bull. London Math. Soc. 46 (2014), 847-856. [11] J. H. Shapiro, Composition Operators and Classical Function Theory, Springer- Verlag, New York, 1993. [12] S. Ueki, Weighted composition operator on the Fock space, Proc. Amer. Math. Soc. 135 (2007), 1405-1410. [13] L. Zhao, Unitary weighted composition operators on the Fock space of Cn, Complex Anal. Oper. Theory 8 (2014), 581-590. [14] L. Zhao, Invertible weighted composition operators on the Fock space of Cn, J. Funct Spaces 2015. Art. ID 250358. [15] L. Zhao and C. Pang, A class of weighted composition operators on the Fock space, Journal of Mathematical Research with Applications, 35 (3) (2015), 303-310. [16] K. Zhu, Analysis on Fock Spaces, Graduate Texts in Mathematics 263, Springer, New York, 2012. [17] N. Zorboska, Compositon operators with closed range, Trans. Amer. Math. Soc. 344 (1994), 791-801. M. Fatehi, Department of Mathematics, Shiraz Branch, Islamic Azad University, Shi- raz, Iran. E-mail: [email protected] 11
1111.3990
2
1111
2012-09-25T17:51:32
Uncertainty Principles in Finitely generated Shift-Invariant Spaces with additional invariance
[ "math.FA" ]
We consider finitely generated shift-invariant spaces (SIS) with additional invariance in $L^2(\R^d)$. We prove that if the generators and their translates form a frame, then they must satisfy some stringent restrictions on their behavior at infinity. Part of this work (non-trivially) generalizes recent results obtained in the special case of a principal shift-invariant spaces in $L^2(\R)$ whose generator and its translates form a Riesz basis.
math.FA
math
UNCERTAINTY PRINCIPLES IN FINITELY GENERATED SHIFT-INVARIANT SPACES WITH ADDITIONAL INVARIANCE ROMAIN TESSERA AND HAICHAO WANG Abstract. We consider finitely generated shift-invariant spaces (SIS) with additional invariance in L2(Rd). We prove that if the generators and their translates form a frame, then they must satisfy some stringent restrictions on their behavior at infinity. Part of this work (non-trivially) generalizes recent results obtained in the special case of a principal shift-invariant spaces in L2(R) whose generator and its translates form a Riesz basis. 1. Introduction Finitely generated shift-invariant spaces have been widely used in approximation theory, numerical analysis, sampling theory and wavelet theory ( see e.g., [1, 3, 4, 9, 11, 12, 15, 16] and the references therein). Shift-invariant spaces with additional invariance have been studied in the context of wavelet analysis and sampling theory [6, 10, 13, 17], and have been given a complete algebraic description in [2] for L2(R) and in [7] for L2(Rd). As a tool for showing our main results, we will prove a slightly different but useful characterization. It is well-known that the Paley-Wiener space P W is translation-invariant. More- over, Shannon's sampling theorem easily implies that P W is principal, i.e. gener- ated by the single function sinc. It turns out that the fact that sinc is non-integrable is not a coincidence. Actually, for a principal shift-invariant space in L2(R) which is translation-invariant, any frame generator is non-integrable (see for instance [5]). This observation holds in any dimension. frame generator has to satisfy for a.e. ω ∈ Rd Indeed, the Fourier transform bφ of a C−11A(ω) ≤ bφ(ω) ≤ C1A(ω) 1 for some C ≥ 1 and some finite measure subset A. In particular, bφ is not continuous. Such a condition also prevents bφ from being in the Sobolev space1 H 2 (Rd) (see 2 −ǫ(Rd) for every ǫ > 0 when A is a Euclidean ball [14]), whereas it belongs to H of positive radius. 1 Our first result is a straightforward generalisation of this fact to shift-invariant spaces generated by several functions. Theorem 1.1. Let Λ be a lattice in Rd. If a finitely generated Λ-invariant space of L2(Rd) is translation-invariant, then at least one of its frame generators has a Key words and phrases. Finitely generated shift-invariant spaces; Frame; Additional invari- ance; Uncertainty principle. 1i.e. R φ(x)2(1 + x)dx = ∞. 1 2 ROMAIN TESSERA AND HAICHAO WANG non continuous Fourier transform, and in particular is not in L1. Moreover this generator satisfies R φ(x)2(1 + x)dx = ∞. The slow spatial-decay of the generators of shift-invariant spaces that are also translation-invariant is a disadvantage for the numerical implementation of some analysis and processing algorithms. This is a motivation for considering instead shift-invariant spaces that are only 1 n Z-invariant and hoping the generators will have better time-frequency localization. Indeed, it was shown in [5] that for every n, one can construct a principal shift-invariant space with an orthonormal genera- tor which is in L1(R) although the space is 1 n Z-invariant. However, there are still obstructions if we require more regularity on the Fourier transform of the gener- ators. Asking for fractional differentiability yields the following Balian-Low type obstructions (see [8] and the reference therein). Compare [5, Theorem 1.2]. Theorem 1.2. Let Λ < Γ be two lattices2 in Rd. Suppose φi ∈ L2(Rd) are such that {φi(· + λ)λ ∈ Λ, i = 1, . . . , r} forms a frame for the closed subspace V Λ(Φ) spanned by these functions. Let ρ (≤ r) be the minimal number of generators of V Λ(Φ). Assume that [Γ : Λ] is not a divisor of ρ, and suppose that V Λ(Φ) is Γ-invariant. Then there exists i0 ∈ {1, . . . , r} such that ZRd φi0 (x)2xd+ǫdx = +∞ for all ǫ > 0. I.e. bφi0 is not in H d 2 +ǫ(Rd). One can also ask for a combinaison of regularity, namely continuity and of control of the decay at infinity. In this spirit, we obtain the following result (compare [5, Theorem 1.3]). Theorem 1.3. Keep the same assumptions as in the last theorem, and suppose moreover that bφi is continuous for every i = 1, . . . , r. Then there exits i0 ∈ {1, . . . , r} such that ω d 2 +ǫbφi0 (ω) 6∈ L∞(Rd) for any ǫ > 0, i.e., ess sup ω∈Rbφi0 (ω)ω 2 +ǫ = +∞. d For d = 1, the exponent is sharp up to the ǫ in both theorems but it does not seem to be the case for larger d. We actually conjecture that the right exponent should be the same for all dimensions. Notice that the condition [Γ : Λ] ∤ ρ in the above theorems is essential. Indeed all regularity constraints trivially disappear when ρ = k[Γ : Λ] for any integer k > 0. To see why, start with some Γ-invariant space generated by exactly k orthogonal generators φ1, . . . , φk, with -- say -- smooth and compactly generated Fourier trans- forms. Then note that as a Λ-invariant space, V Γ(φ1, . . . , φk) is generated by the orthogonal generators φi(· − f ) for f ∈ F and i = 1, . . . , k, where F is a section of Γ/Λ in Γ. The previous results state that under additional invariance there always exists at least one (frame) generator whose Fourier transform has poor regularity. One may wonder if at least some generators can be chosen with good properties. We do not know what the optimal proportion of good generators should be. The following 2The notation Λ < Γ is standard to denote subgroups. UNCERTAINTY PRINCIPLES 3 proposition gives a lower bound on the number of good generators. Observe that this bound gets worse with the dimension. Proposition 1.4. For every d ≥ 1, and every k ∈ N there exists an SIS V (Φ) in L2(Rd) generated by an orthonormal basis Φ consisting of r = (2k)d functions, kd of which have smooth and compactly supported Fourier transforms, and such that V (Φ) is translation-invariant. Moreover all the generators can be chosen so that their Fourier transforms are in H 2 −ǫ(Rd) for all ǫ > 0. 1 We also have the following result, Proposition 1.5. For d ≥ 1, let Γ(> Zd) be a lattice of Rd and r ≥ 1. Then there exists an SIS V (Φ) in L2(Rd) generated by r orthonormal generators φi's, all of which are in L1 (hence they have continuous Fourier transforms) and satisfy ω 2bφi(ω) ∈ L∞(Rd). Moreover, V (Φ) is Γ-invariant. 1 To summarize, while Proposition 1.4 states that it is possible to construct translation- invariant SIS with a portion of the generators having smooth and compactly sup- ported Fourier transforms, Proposition 1.5 shows that we can construct Γ-invariant SIS with all its generators having certain pointwise decay in Fourier domain. Organization. In the following section, we state and prove a convenient charac- terization (similar to the one given in [2]) of SIS with additional invariance. In Section 3, we refine this characterization under the assumption that the genera- tors and their translates form a frame. In Section 4, we prove a useful property of the Gramian under the sole assumption that the generators have continuous Fourier transform. Sections 5 and 6 are dedicated to the proof of the results stated in the introduction. 2. Finitely generated shift-invariant spaces with additional invariance Given a closed subgroup Λ of Rd, we define its Fourier transform Λ∗ as the closed subgroup of Rd defined by Λ∗ = {x ∈ Rd, e2πihλ,xi = 1 ∀λ ∈ Λ}. Note that the map Λ → Λ∗ is an involution. The Fourier transform reverses the inclusions, namely if Λ < Γ then Γ∗ < Λ∗. Observe for instance that for d = 1, the Fourier transform of nZ is 1 n Z, or that for d = 2, the Fourier transform of R × {0} is {0} × R. In this paper, we consider the following general setting: Λ < Γ are two closed subgroups of Rn. Let φ1, . . . , φr ∈ L2(Rd), and denote by Φ the column vector whose components are the φi's. We will denote by V Λ(Φ) the smallest closed Λ- invariant subspace containing the φi's. It is common to call V Λ(Φ) shift-invariant when Λ = Zd, and translation-invariant when Λ = Rd. To allege notation, we will omit the subscript Zd, when Λ = Zd. We will be interested in the situation where V Λ(Φ) is in addition Γ-invariant. As we will see later, this prevents frame generators from having nice decay at infinity. We will start by providing a short and self-contained proof of a result essentially due 4 ROMAIN TESSERA AND HAICHAO WANG to [2, 7]. Observe that this problem is only non-trivial when the quotient Rd/Λ is compact since otherwise, the only finitely generated Λ-invariant space is {0}. This is equivalent to the fact that Λ∗ (hence Γ∗) is discrete. In this paper we will mainly focus on the cases when Λ is a lattice and when Γ is either a (larger) lattice or all of Rd. Before stating this result, let us introduce some notation. The Gramian associated to Φ and Λ is a measurable field of r×r matrices whose general coefficient is defined for a.e. ω ∈ R by For short, GΛ i,j(ω) = Xl∈Λ∗ bφi(ω + l)cφj(ω + l). GΛ(ω) = Xl∈Λ∗ bΦ(ω + l)bΦ∗(ω + l). A(ω) = Xg∈Γ∗ bΦ(ω + g)bΦ∗(ω + g). For a.e. ω ∈ R, let A(ω) to be the r × r matrix defined by Now, let F be a subset of Λ∗ consisting of representatives of the quotient Λ∗/Γ∗. For instance, for d = 1, Λ = Z and Γ = 1 n Z, one has Λ∗ = Z and Γ∗ = nZ, and for F , one can take {0, . . . n − 1}. We have A(ω + f ). (2.1) GΛ(ω) = Xf ∈F Now let us state the main result of this section. Although it could easily be deduced from the main results in [2, 7], we chose to write down a short self-contained proof. Theorem 2.1. Let Λ < Γ be closed cocompact subgroups of Rd. The space V Λ(Φ) is Γ-invariant if and only if the following equality holds for a.e. ω ∈ Rd. rank A(ω + f ). (2.2) rank GΛ(ω) = Xf ∈F Proof. Recall that a function ϕ belongs to V Λ(Φ) if and only if there exist bounded measurable Λ∗-periodic functions P1, . . . , Pr such that bϕ = P1cφ1 + . . . + Prcφr. Fiberwise, this is equivalent to saying that for a.e. ω, the vector (bϕ(ω + l))l∈Λ∗ ∈ ℓ2(Λ∗) lies in the subspace bV (ω) spanned by the r vectors (bφi(ω + l))l∈Λ∗ for i = 1, . . . , r. Now the space V Λ(Φ) is Γ-invariant if for every bounded Γ∗-periodic function Q, and for every i = 1, . . . , r, there are bounded measurable Λ∗-periodic functions Pi,1, . . . , Pi,r such that Qbφi = Pi,1cφ1 + . . . + Pi,rcφr. Again, this is equivalent to saying that for a.e. ω, and for every bounded Γ∗-periodic θ : Λ∗ → C the vector (θ(ω + l)bφi(ω + l))l∈Λ∗ ∈ ℓ2(Λ∗) lies in the subspace bV (ω). Let us denote cW (ω) the subspace of ℓ2(Λ∗) spanned by (θ(ω + l)bφi(ω + l))l∈Λ∗ for all bounded Γ∗-periodic θ, and every i = 1, . . . , r. Clearly bV (ω) is a subspace of cW (ω), and the two coincide for a.e. ω exactly when V Λ(Φ) is Γ-invariant. UNCERTAINTY PRINCIPLES 5 The rest of the proof amounts to showing that the left-hand term in (2.2) corre- sponds to the dimension of bV (ω) (which is obvious) and that the right-hand term of the equality corresponds to the dimension of cW (ω). A basis for the space of bounded Γ∗-periodic functions on Λ∗ consists of the functions (θf )f ∈F , where each θf (l) equals 1 if l ∈ f + Γ∗ and 0 elsewhere. Observe that for all f 6= f ′, and all ϕ, ϕ′ ∈ W bΦ(ω), θf ϕ and θf ′ϕ′ are orthogonal. Hence cW (ω) decomposes as a direct sum cW (ω) = Mf ∈F cW f (ω), where cW f (ω) is the subspace of functions of the form θf ϕ, for ϕ ∈ cW (ω). But it comes out that cW f (ω) is precisely the subspace of ℓ2(Λ∗) spanned by (bφi(ω + f + g))g∈Γ∗ , whose dimension equals the rank of A(ω + f ). This finishes the proof of the theorem. (cid:3) Although we will be mostly interested in the case where both Λ and Γ are lattices, we will also use the following special case of Theorem 2.1, where Λ is a lattice and Γ = Rd (so Γ∗ = {0}). Corollary 2.2. Let Λ be a lattice, and let V Λ(Φ) be a Λ-invariant space generated by φ1, . . . , φr. Then it is translation-invariant if and only if the following equality holds for a.e. ω ∈ Rd. (2.3) (2.4) rank GΛ(ω) = Xf ∈Λ∗ rank A(ω + f ) = {f ∈ Λ∗, Φ(ω + f ) 6= 0} . 3. Frame generators and additional invariance The main goal of this section is to reformulate Theorem 2.1 under the additional assumption that the generators form a frame for the space V Λ(Φ). Recall that the family of all Λ-translates of the generators φ1, . . . , φr form a Riesz basis if and only if there exists s ≥ 1 such that s−1I ≤ GΛ(ω) ≤ sI, a.e. ω ∈ R. (3.1) In this case the functions φi's are called Riesz generators for V Λ(Φ). Similarly, {φi(· + λ) λ ∈ Λ, i = 1, . . . , r} is a frame if and only there exists s ≥ 1 such that s−1GΛ(ω) ≤ (GΛ(ω))2 ≤ sGΛ(ω) (3.2) for almost every ω ∈ Rd (see [9]). Here the φi's are called frame generators for V Λ(Φ). Let us start by an easy lemma. Given a non-negative self-adjoint matrix A, denote qA its associated quadratic form, µ−(A) its smallest non-zero eigenvalue and kA the dimension of its kernel. Denote the unit sphere of a Euclidean space V by SV . Lemma 3.1. Let C = A + B be three d× d non-negative self-adjoint matrices such that rank C = rank A + rank B. Then µ−(C) ≤ min{µ−(A), µ−(B)}. 6 ROMAIN TESSERA AND HAICHAO WANG Proof. Note that we can assume without loss of generality that C has full rank. Observe that KerA and KerB are in direct sum. By the min-max theorem, µ−(A) = min dim V =kA+1 max x∈SV qA(x). Now let V0 be a subspace minimizing the above expression. Let x ∈ V0 ∩ KerB of norm 1. We have µ−(A) ≥ qA(x) = qA(x) + qB(x) = qC (x) ≥ µ−(C). We conclude since the roles of A and B are symmetric. (cid:3) The following theorem will play a central role in the sequel. Theorem 3.2. Let Λ < Γ be closed cocompact subgroups of Rd. Let Φ be finite set of generators of V Λ(Φ). Suppose the space V Λ(Φ) is Γ-invariant and Φ are frame generators. Then, there exists s ≥ 1 such that the formula (2.2) and the following inequality hold for a.e. ω ∈ Rd s−1A(ω) ≤ A(ω)2 ≤ sA(ω). (3.3) Proof. By Theorem 2.1, it is enough to prove that (3.2) implies (3.3) (up to chang- ing the constant s). First observe that the previous lemma can be extended (by induction) to a sum of more than two matrices. We deduce that the lower bound in (3.2) implies that of (3.3). The equivalence between the upper bounds follows from the fact that the number of non-zero matrices in the right-hand side of (2.1) is bounded by r. (cid:3) We immediately get the following corollary. Corollary 3.3. Let Λ be a closed cocompact subgroups of Rd. Assume that the space V Λ(Φ) is translation-invariant and Φ are frame generators. Then T r(A(ω)) is not continuous, nor in H 2 . 1 Proof. By Theorem 3.2, the map ω → T r(A(ω)) is larger than s−1 on a set of positive (but finite) measure, and equals zero elsewhere. In particular it is not continuous and not in H (cid:3) 2 (by [14]). 1 Lemma 3.4. Let (M (x)) x∈Rd be a continuous family of r × r non-negative self- adjoints matrices with complex coefficients. Assume that there exists s ≥ 1 such that s−1M (x) ≤ M 2(x) for all x ∈ Rd. Then the rank of M is constant. Proof. Since x → M (x) is continuous, its rank is lower semi-continuous. Let F be the (closed) subset where rank M reaches its minimal value. Assume that F is not open, which means that there exists a sequence xn ∈ F c converging to some x0 ∈ F . It follows that the rank of M (xn) is strictly larger than the rank of M (x0), which therefore implies that the range of M (xn) intersects non-trivially the kernel of M (x0). Let un be a sequence of unit vectors lying in this intersection. On one hand, because un is in the range of M (xn), we must have kM (xn)unk ≥ s−1. On the other hand, the continuity of M together with the fact that the limit lies in the kernel implies M (xn)un → 0, contradiction. So F is open which implies F = Rd hence the lemma. (cid:3) UNCERTAINTY PRINCIPLES 7 Corollary 3.5. Let Λ < Γ be closed cocompact subgroups of Rd. Assume that the space V Λ(Φ) is Γ-invariant and Φ are frame generators. Then if A(ω) is continuous, its rank is constant. Proof. The corollary follows immediately from Lemma 3.4 and Theorem 3.2. (cid:3) 4. Properties of the Gramian when the bφi's are continuous Although this section mainly serves as preparation for Theorem 1.2, we believe that it is of independent interest and could be useful elsewhere. From now on, Λ will always denote a lattice in Rd. The following statement is essentially trivial (and was observed for instance in [9, 12]). Proposition 4.1. Let Λ be a lattice in Rd, and let Φ = (φ1, . . . , φr) be a generating set for the Λ-invariant space V Λ(Φ). Then the minimal number ρ of generators of V Λ(Φ) equals the essential supremum of rank GΛ(ω). Equivalently, if GΛ(ω) is non invertible a.e., then one can find a generating set Φ′ with r − 1 generators. Proof. This is an essentially trivial statement. Let K be a fundamental domain for the action of Λ∗ (for example take K = [0, 1)d if Λ∗ = Zd). For every ω ∈ K, let V Λ(ω) be the subspace of ℓ2(Λ∗) spanned by the r vectors vi(ω) = bφi(ω + l)l∈Λ∗ for i = 1, . . . , r. Saying that GΛ is non-invertible a.e. amounts to the fact that dim V Λ(ω) ≤ r − 1 for a.e. ω ∈ K. The idea is to remove for a.e. ω one vector vi(ω) and to recombine the other ones in order to get a new set of r − 1 generators for V Λ(Φ). Now, in order to do this in a measurable way, we pick the vi(ω) of minimal index with the property that it lies in the vector space spanned by the other ones. Then we relabel the remaining ones respecting their order: for instance if v2 is removed, then v1 becomes v′ 2 and so on. Now since every point of Rd can be written as ω + l, for some unique (ω, l) ∈ K × Λ∗, we can define our new r − 1 generators by the formula φ′ We deduce from this proposition that the essential supremum of rank G(ω) equals ρ. i(ω))l, for all l ∈ Λ∗. 1, v3 becomes v′ i(ω + l) = (v′ (cid:3) Lemma 4.2. Suppose that the functions bφi are continuous for i = 1, . . . , r. Then the maps ω → rank A(ω) and ω → rank GΛ(ω) from Rd to N are lower semi- continuous. Proof. Since GΛ(ω) = PF A(ω + f ) it is enough to show that rank A(ω) is lower semi-continuous. Observe that A(ω) is the sum over g ∈ Γ∗ of the continuous positive semi-definite matrices bΦ(ω + g)bΦ(ω + g)∗. Therefore the rank of A(ω) is the supremum of the ranks of all partial finite sums. Since lower semi-continuity is stable under taking supremums, we deduce that rank A(ω) is lower semi-continuous. (cid:3) Proposition 4.3. Suppose that the functions bφi are continuous, and that V Λ(Φ) is Γ-invariant for some lattice Γ such that [Γ : Λ] does not divide ρ. Then the subset {ω, rank GΛ(ω) < ρ} is a non-empty closed subset of Rd. 8 ROMAIN TESSERA AND HAICHAO WANG Proof. The fact that {ω, rank GΛ(ω) ≤ ρ − 1} is closed results from Lemma 4.2. We therefore only have to prove that it is non-empty. Let us assume on the contrary that GΛ(ω) has rank ρ for all ω. We know by Theorem 2.1 that for a.e. ω ∈ Rd, (4.1) rank A(ω + f ). ρ = Xf ∈F Moreover the equality is an inequality ≤ for all ω. But again Lemma 4.2 implies that the ω's for which the inequality is strict form an open set, which therefore has to be empty. Now observe that lower semi-continuous functions with integer values can only increase locally. In other words for every ω0 the set {ω, rank A(ω) ≥ rank A(ω0)} = {ω, rank A(ω) > rank A(ω0) − 1/2} is a non-empty open set. Together with (4.1), this immediatly implies that rank A(ω) is locally constant, hence constant on Rd (since Rd is connected!). Let us call m the corresponding integer. We therefore have ρ = m[Γ : Λ], contradiction. (cid:3) 5. Proofs of The main results 5.1. Proof of Theorem 1.1. Immediately follows from Corollary 3.3. d 5.2. Proof of Theorem 1.3. Suppose by contradiction that there exists ǫ > 0 such that ω we conclude by Corollary 3.5: namely since rank A(ω) is constant it must divide the rank of G(ω), which by (2.2) would imply that the index of Λ in Γ divides r. 2 +ǫbφi(ω) ∈ L∞(Rd) for all i. This easily implies that A(ω) is continuous, so 5.3. Proof of Theorem 1.2. The proof of Theorem 1.2 relies on the following result of harmonic analysis. Its proof for d = 1 is essentially the proof of [5, Theorem1.2]. Since it extends without change to any d, we do not reproduce it here. Lemma 5.1. Let Λ be a lattice in Rd. Suppose that a function f ∈ L2(Rd) satisfies the following properties (1) there exists ǫ > 0 such that ZRd f (x)2xd+ǫdx < ∞. Xk∈Λ∗ bf (ω + k)2 ≤ C. (2) There exists a constant C such that for a.e. ω ∈ Rd, (5.1) (3) There exists ω0 is such that Then for all η > 0, there exists δ0 > 0 only depending on η, ǫ and C such that for all 0 < δ ≤ δ0 for all k ∈ Λ∗, bf (ω0 + k) = 0 δd ZB(ω0,δ) Xk∈Λ∗ bf (ω + k)2dω < η. 1 UNCERTAINTY PRINCIPLES 9 Proof of Theorem 1.2. We shall use the notation introduced in Section 3 (before stating Lemma 3.1). Let us suppose by contraction that there exits ǫ > 0 such that ZRd φi(x)2xd+ǫdx < ∞, (5.2) for every i. Note that this condition implies that the φi are in L1, so that their Fourier transforms are uniformly continuous. Define E = {ω, rank GΛ(ω) < ρ}, which by Proposition 4.3, is a non-empty closed subset of Rd. Take an open ball B ⊂ Ec whose boundary interests E in a point ω0. Observe that at least one fourth of the volume of a sufficiently small ball centered in ω0 is contained in Ec. We will use this remark at the end of the proof. Since ω0 ∈ E, there exists an (r − ρ + 1)-dimensional subspace V0 of Rr on which GΛ(ω0) vanishes. Since GΛ(ω) = F (ω)F ∗(ω) where F (ω) is the r × Λ∗ matrix whose (i, l)-coefficient is F (ω)il = bφi(ω + l), we have F ∗(ω0)a = 0 for all a ∈ V0. This implies that rXi=1 1 for all l ∈ Λ∗. aibφi(ω0 + l) = 0 In other words, the function fa(x) = Pr i=1 aiφi(x) satisfies bfa(ω0 + l) = 0 for all l ∈ Λ∗ and for all a ∈ V0. Now clearly the functions fa for a ∈ SV0 satisfy the assumptions of the lemma with C independent of a. We therefore get for δ > 0 small, δd ZB(ω0,δ)Xk∈ bΛ bfa(ω + k)2dω < η. (5.3) Suppose a1, . . . , ar−ρ+1 is an orthonormal basis of V0. Then for a.e. ω ∈ B, the frame condition implies s−1 ≤ µ−(GΛ(ω)) ≤ r−ρ+1Xi=1 hGΛ(ω)ai, aii. By the above remark, we see that this contradicts (5.3) provided η is small enough. (cid:3) 6. Proofs of Proposition 1.4 and Proposition 1.5 6.1. Proof of Proposition 1.4. It is actually enough to provide a construction for d = 1. Indeed given such φ1, . . . , φ2k in L2(R), we define generators in dimension d by considering tensor products of these functions: namely φi1,...,id = φi1 ⊗ . . .⊗ φid . Then we get that V Φ is simply the tensor product V ⊗d Φ , whose generators are clearly orthonormal. Translation invariance of V Φ therefore follows from that of VΦ. Hence, let us focus on the one-dimensional case. Let g be an infinitely-differentiable function that satisfies g(x) = 0 when x ≤ 0, g(x) = 1 when x ≥ 1, and g2(x) + g2(1 − x) = 1 when 0 ≤ x ≤ 1. Define bφ1(ω) = g(ω)g(2 − ω) and bφ2(ω) = g(1 − ω)χ[0,1) − g(ω − 1)χ[1,2). Then for i ≤ r, define bφi(ω) = bφ1(ω − i + 1) when i is odd and bφi(ω) = bφ2(ω − i + 2) when i is even (bφ1 and bφ2 are plotted in Figure 1). It is easy to check that Pl∈Z bφi(ω + l)2 = 1 a.e. ω for all i, and that Pl∈Z bφ1(ω + l)bφ2(ω + l) = 0. We 10 ROMAIN TESSERA AND HAICHAO WANG deduce thatPl∈Z bφi(ω + l)bφj(ω + l) = 0 for i 6= j. This shows that G(ω) = I which amounts to saying that the functions φi are orthonormal generators for V (Φ). 1 0.8 0.6 0.4 0.2 0 −0.2 −0.4 −0.6 −0.8 −1 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 Figure 1. The function bφ1 being red and the function bφ2 being blue. Since all the functions are supported on [0, r], we can show that, for a.e. ω ∈ [0, 1), rank A(ω + f ) = 1 when 0 ≤ f ≤ r − 1 and rank A(ω + f ) = 0 elsewhere. Hence Pf ∈Z rank A(ω + f ) = r = rank GΛ(ω). From Corollary 2.2, V (Φ) is translation- invariant. Observe that when i is odd, bφi is compactly supported and infinitely-differentiable. When i is even, bφi can be written as a product of a compactly supported and infinitely-differentiable function and the characteristic function χ[i−2,i). Since χ[i−2,i) belongs to H 2 −ǫ for any ǫ > 0. 1 1 2 −ǫ, we have bφi ∈ H Z×···× 1 6.2. Proof of Proposition 1.5. We assume that Γ = 1 nd n1 first need the following lemma which treats the case when d = 1 and r = 1. Lemma 6.1 ([5]). For integer n1, there exists a function φ1 ∈ L1 ∩ L2 (and hence bφ1 is continuous), such that φ1 is an orthonormal generator for its generating space V (φ1), V (φ1) is 1 n1 Z-invariant and ω Z. We Z× 1 n2 1 2bφ1(ω) ∈ L∞(R). Since our method relies on the construction of φ1 in Lemma 6.1, we will describe φ1 explicitely here. Let g be the function as in the proof of Proposition 1.4. Define g0(x) = g(x + 1)g(−x + 1) and g1(x) = g(x + 1)g(−2x + 1). Then the Fourier transform of φ11 is defined to be cφ1(ω) = h0(ω) + ∞Xj=1 + ∞Xj=1 4j −1Xl=0 2−jhj(ω − n1(γj + l)) 4j −1Xl=0 2−jhj(−ω − n1(γj + l)), (6.1) UNCERTAINTY PRINCIPLES 11 1 4 , 1 k=0 4k, h0(ω) = g0(4ω) and hj(ω) = g1(2j+1ω − 2j + 1). where γj =Pj−1 It is not hard to check that cφ1 is an even function such that bφ1 ≤ 1 and has support on [− 1 4 ] ∪ (∪∞ 2 − 2j+2 + n1(γj + l)] and E′ From one to several generators (in dimension 1). Now we want to construct a 1 Z-invariant SIS with r orthonormal generators in L1∩L2 satisfying our pointwise n1 decay property in Fourier domain. Define φi for 2 ≤ i ≤ r to be such that bφi(ω) = 0 when ω ∈ [0, n1γ2i−2 − 1] and 1j ), where E1j = ∪4j −1 l=0 [ 1 1 2j − 1 2 − n1(γj + l), 2j + n1(γj + l), 1 2 − n1(γj + l)]. j=1E1j ∪ E′ 1j = ∪4j −1 l=0 [ 1 2 − 1 2j+2 − 1 2−(2i−2)h0(ω − n1(γ2i−2 + l))(cid:17) bφi(ω) = ∞Xj=1 + 1 √2(cid:16) 42i−2−1Xl=0 4j+2i−2−1Xl=0 2−(j+2i−2)hj(ω − n1(γj+2i−2 + l)) (6.2) when ω ≥ n1γ2i−2 − 1, and bφi(ω) = bφi(−ω) when ω ≤ 0. It is easy to see that each bφi obeys the pointwise decay property and that it is an orthonormal generator for the principal SIS V (φi). In fact, Xl∈Z bφi(ω + l)2 =Xl∈Z bφ(ω + l)2 = 1 for a.e. ω. One can also check that for 2 ≤ i ≤ r, bφi has support (cid:16) ∪∞ j=1 Eij ∪ E′ l=0 ij(cid:17)[(cid:16) ∪42i−2−1 1 [− 4 − n1(γ2i−2 + l), [(cid:16) ∪42i−2−1 1 + n1(γ2i−2 + l), [− 4 2 − 1 [ 1 2 − n1(γj+2i−2 + l), 2j + n1(γj+2i−2 + l), 1 2j+2 − 1 l=0 1 1 4 − n1(γ2i−2 + l)](cid:17) + n1(γ2i−2 + l)](cid:17), 1 4 (6.3) l=0 where Eij = ∪4j+2i−2−1 l=0 ij = ∪4j+2i−2−1 2j − 1 [ 1 E′ In particular, each V (φi) is 1 n1 Fourier domain. Hence V (Φ1) is also 1 n1 whose components are the φi's. In fact 2j+2 + n1(γj+2i−2 + l)] and 2 − 1 2 − n1(γj+2i−2 + l)]. -invariant and all the φi's have disjoint support in -invariant, where Φ1 is the column vector V (Φ1) =Mi≤r V (φi) (6.4) From Lemma 6.1, for each j ≥ 2, we can construct V (ψj) with orthonormal gener- ator ψj having the desired properties and that V (ψj ) is 1 nj Higher dimension. Like in the previous section, we let φi be the d-fold tensor product φi ⊗ ψ2 . . .⊗ ψd, for i = 1, . . . , r. By construction, the shift-invariant space generated by these r orthonormal function is Γ-invariant. Since all φi's and ψj's are in L1(R), it follows that the φi's are in L1(Rd). Pointwise decay results from the fact that all these functions have bounded Fourier transforms. Z-invariant. 12 ROMAIN TESSERA AND HAICHAO WANG Acknowledgement The authors would like to thank Akram Aldroubi and Qiyu Sun for valuable discussions. The authors would also like to thank Carolina Mos- quera and Victoria Paternostro for their comments on the proof of Theorem 1.2. References [1] A. Aldroubi, C. A. Cabrelli, D. Hardin, and U. M. Molter, Optimal shift invariant spaces and their parseval frame generators, Appl. Comput. Harmon. Anal., 23 (2007), 273 -- 283. [2] A. Aldroubi, C. A. Cabrelli, C. Heil, K. Kornelson and U. M. Molter, Invariance of a shift invariant space, J. Fourier Anal. and Appl., 16 (2010), 60-75. [3] A. Aldroubi and K. Grochenig, Non-uniform sampling in shift-invariant space, SIAM Rev., 43 (2001), no. 4, 585 -- 620. [4] A. Aldroubi, Q. Sun and W.-S. Tang, Convolution, average sampling, and Calderon resolution of the identity of shift-invariant spaces, J. Fourier Anal. Appl., 11 (2005), 215 -- 244. [5] A. Aldroubi, Q. Sun, and H. Wang, Uncertainty principles and Balian-Low type theorems in principal shift-invariant spaces, Appl. Comput. Harmon. Anal., 30(2011), 337 -- 347. [6] A. Aldroubi, I. Krishtal, R. Tessera, and H. Wang, Principal Shift-Invariant Spaces with Extra Invariance Nearest to Observed Data, Collectanea Mathematica, 63(2012), 393-401. [7] M. Anastasio, C. Cabrelli, and V. Paternostro, Invariance of a shift-invariant space in several variables, Complex Anal. Oper. Theor. 2010. [8] J. Benedetto, W. Czaja, P. Gadzinski and A. Powell, The Balian-Low Theorem and regularity of Gabor systems, J. Geom. Anal., 13 (2003), 239 -- 254. [9] M. Bownik, The structure of shift-invariant subspaces of L2(Rn), J. Funct. Anal., 177(2000), 282 -- 309. [10] C. K. Chui and Q. Sun, Tight frame oversampling and its equivalence to shift-invariance of affine frame operators, Proc. Amer. Math. Soc., 131 (2003), 1527 -- 1538. [11] C. K. Chui and Q. Sun, Affine frame decompositions and shift-invariant spaces, Appl. Com- put. Harmon. Anal., 20 (2006), 74 -- 107. [12] C. de Boor, R.A. DeVore, and A. Ron, The structure of finitely generated shift-invariant spaces in L2(Rd), J. Funct. Anal. 119 (1994), no. 1, 37 -- 78. [13] J. Hogan and J. Lakey, Non-translation-invariance and the synchronization problem in wavelet sampling Acta Appl. Math., 107 (2009), 373 -- 398. [14] M. Kolountzakis and T. Wolff, On the Steinhaus tiling problem. Mathematika 46 (1999), 253 -- 280. [15] A. Ron, Z. Shen, Frames and stable bases for shift invariant subspaces of L2(Rd), Canadian Journal of Mathematics, 47(1995), 1051-1094. [16] M. Unser and T. Blu, Fractional splines and wavelets, SIAM Rev., 42 (2000), 43 -- 67. [17] E. Weber, On the translation invariance of wavelet subspaces, J. Fourier Anal. Appl., 6 (2000), 551 -- 558. (Romain Tessera) Department of Mathematics, Ecole Normale Superieure de Lyon, France E-mail address: [email protected] (Haichao Wang) Department of Mathematics, University of California, Davis, CA 95616 E-mail address: [email protected]
1502.05909
1
1502
2015-02-20T15:43:20
Extensions of the Inequalities of Hardy and Hilbert
[ "math.FA", "math.OA" ]
In this note we produce generalized versions of the classical inequalities of Hardy and of Hilbert and we establish their equivalence. Our methods rely on the H^1-BMOA duality. We produce a class of examples to establish that the generalizations are non-trivial.
math.FA
math
EXTENSIONS OF THE INEQUALITIES OF HARDY AND HILBERT VERN I. PAULSEN AND DINESH SINGH Abstract. In this note we produce generalized versions of the classical inequalities of Hardy and of Hilbert and we establish their equivalence. Our methods rely on the H 1 − BM OA duality. We produce a class of examples to establish that the generalizations are non-trivial. 1. Introduction The classical inequality of Hardy states that there exists a constant A so n=0 anzn is a function in the Hardy space H 1(D), then that if f (z) =P∞ (1.1) an n + 1 ≤ Akfk1. ∞ Xn=0 The well known inequality of Hilbert states that there exists a constant B so that if (αn), (βn) are two sequences in the HIlbert space ℓ2 + = ℓ2(Z+), then (1.2) ∞ ∞ Xm=0 Xn=0 αnβn n + m + 1 ≤ Bk(αn)k2k(βn)k2. In this paper we establish generalizations of both of these inequalities by replacing the sequences (cid:8) 1 n(cid:9) and n 1 n+m+1o with a larger class of se- quences. We then prove that the two generalized inequalities are equivalent. In particular, it follows that the two classical inequalities (1.1) and (1.2) are equivalent. The literature frequently mentions the fact that Hilbert’s inequalitiy (1.2) implies Hardy’s inequality (1.1), but we found no mention of the converse. The inequalities of Hardy (1.1) and Hilbert (1.2) are extremely important for a variety of reasons and they are still of current interest. For some classic expositions, we refer the reader to [1], [2], [3] and [4]. For a sampling of the current research interest, we refer the reader to [5], [6], [7], [8], [9], and [10]. 2000 Mathematics Subject Classification. Primary 26D15; 30H10; Secondary 30H35. Key words and phrases. generalized Hardy inequality; generalized Hilbert inequality; equivalence; H 1 − BM OA duality. Research supported in part by NSF . 1 2 V. I. PAULSEN AND D. SINGH 2. Preliminary Results The Lebesque spaces on the unit circle T in the complex plane shall be de- noted by Lp, 1 ≤ p ≤ ∞. The Hardy spaces H p are the subspaces consisting of all f in Lp whose negative Fourier coefficients vanish. In particular, the +, via the map en → zn, n ≥ 0. Hilbert space H 2 will often be identified with ℓ2 The H p spaces can be identified as spaces of holomorphic functions on the open unit disk D that satisfy the condition: (2.1) sup 0<r<1(cid:18) 1 2π Z 2π 0 f (reiθ)pdθ(cid:19)1/p < ∞, The dual of H 1 is the Banach space BM OA defined as all functions and for f ∈ H p this supremum is equal to the norm of f in Lp. f ∈ H 1 such that kfk∗ = sup I I(f − I(f )) < ∞, where I varies over all subarcs of T, I denotes the normalized arclength of I and I(f ) = f (θ)dθ. 1 IZI The norm of a function f ∈ BM OA is given by (2.2) kfk = f (0) + kfk∗. We recall that by the famous duality theorem of Fefferman, BM OA can be identified isometrically and conjugate linearly with the dual of H 1 via the dual pairing given by hf, pi = 1 2π Z 2π 0 p(eiθ)f (eiθ)dθ, Finally, we shall also need Fefferman’s second characterization of BMOA. for f ∈ BM OA and p ∈ H 1 a polynomial. Namely, that a function f ∈ H 1 is in BM OA if and only if (1−r2)f ′(reiθ)2drdθ is a Carleson measure on D. For the details of these facts, we refer the reader to [1], [2], and [3]. 3. Statement of the Main Results We state our three main results below, together with their immediate corollaries, but we defer their proofs until later sections. We let X denote the space of sequences {cn} such that (3.1) k=0(k + 1)2ck2 k{cn}k2 = sup(cid:26)Pn n + 1 : n ∈ N(cid:27) < ∞. It is not hard to see that X is a Banach space in this norm and that ck = 1 k+1 is an element of unit norm in this space. HARDY AND HILBERT 3 Theorem 3.1 (Generalized Hardy Inequality). There is a constant A so that if (cn) is a sequence in X, then for any f (z) =P∞ ∞ k=0 anzn in H 1, ancn ≤ Ak(cn)kkfk1. Xn=0 Corollary 3.2 (Hardy’s Inequality). There is a constant δ such that for any f (z) =P∞ n=0 anzn in H 1, ∞ an n + 1 ≤ δkfk1. Xn=0 Proof. Set cn = 1 n+1 and apply Theorem 3.1. (cid:3) Theorem 3.3 (Generalized Hilbert Inequality). There is a constant B so that if (cn) ∈ X and (an), (bn) in ℓ2 +, then ∞ Xn,m=0 anbmcn+m ≤ Bk(cn)kk(an)k2k(bn)k. Corollary 3.4 (Hilbert’s Inequality). There is a constant B so that for any (an), (bn) in ℓ2 +, ∞ Xm,n=0 anbm n + m + 1 ≤ Bk(an)k2k(bn)k2. Proof. Apply Theorem 3.3 with cn = 1 n+1 . (cid:3) Theorem 3.5. For each (cn) in X, the corresponding generalized Hardy inequality is equivalent to the corresponding generalized Hilbert inequality. Moreover, the least constant A in the generalized Hardy inequality is equal to the least constant B in the generalized Hilbert inequality. Corollary 3.6. The Hardy inequality and Hilbert inequality are equivalent. 4. Proof of Theorem 3.5 We present the proof of Theorem 3.5 first because it is the most direct and demonstrates the equivalence of the classic Hardy and Hilbert inequalities. We first prove that the generalized Hilbert inequality implies the general- + so that by the generalized ized Hardy inequality. Let (an) and (bn) be in ℓ2 Hilbert inequality(Theorem 3.3), ∞ anbncn+m ≤ Bk(cn)kk(an)k2k(bn)k2, Xn,m=0 where k(cn)k is defined by (3.1). Let f (z) =P∞ as f (z) = g(z)h(z) where g(z) =P∞ n=0 αnzn be in H 1. It is well known that f can be factorized n=0 γnzn are both n=0 βnzn and h(z) =P∞ 4 V. I. PAULSEN AND D. SINGH in H 2 and kfk1 = kgk2khk2. Applying the generlized Hilbert inequality to the sequences (βn) and (γn) we have that ∞ Xn,m=0 βnγncn+m ≤ Bk(cn)kk(βn)k2k(γn)k2. This inequality may be rewritten as ∞ m Xm=0(cid:0) Xk=0 cmβkγm−k(cid:1) ≤ Bk(cn)kkgk2khk2 = Bk(cn)kkf1k. By the Cauchy product formula, αm = m Xk=0 ∞ βkγm−k ≤ m Xk=0 βkγm−k, and hence, αmcm ≤ Bk(cm)kkfk1, Xm=0 so that the generalized Hardy inequality follows. This also shows that A ≤ B. We now prove that the generalized Hardy inequality implies the gener- alized Hilbert inequality. Note that to prove the generalized Hilbert in- equality it is sufficient to consider sequences in ℓ2 + of non-negative terms. Choosing any two sequences, (an), (bn) in ℓ2 + with an ≥ 0 and bn ≥ 0. n=0 bnzn we obtain two functions n=0 dnzn is in H 1, where n=0 anzn and h(z) = P∞ Setting g(z) = P∞ in H 2. Hence the function f (z) = g(z)h(z) = P∞ dn =Pn Xn=0 Thus, by the generalized Hardy inequality(Theorem 3.1), ∞ dncn ≤ Ak(cn)kkfk1 ≤ Ak(cn)kkgk2khk2 = Ak(cn)kk(an)k2k(bn)k2. k=0 akbn−k. Substituting for dn we have ∞ ∞ Xn,m=0 cn+manbm = n Xn=0(cid:0) Xk=0 cnakbn−k(cid:1) ≤ Ak(cn)kk(an)k2k(bn)k2, and so the generalized Hilbert inequality follows. Moreover, we see that B ≤ A and so the best constants must be equal. 5. Proofs of Theorems 3.1 and 3.3 We begin with the proof of the generalized Hardy inequality. Note that K = sup 0≤r<1 r 1 − r2[ 1 1−r ]!2 is finite, where [t] denotes the greatest integer less than or equal to t. HARDY AND HILBERT 5 Given I an arc on the unit circle, we let z R(I) = {z ∈ D : z ∈ I, 1 − z ≤ I}, 2π where I = length(I) denotes the normalized arc length. Recall that a pos- itive Borel measure µ on D is called a Carleson measure if there is a constant k so that µ(R(I)) ≤ kI for all subarcs. to BM OA if and only if The theorems of Fefferman and Garsia say that a function f in H 1 belongs (1 − r2)f ′(reiθ)2rdrdθ is a Carleson measure. Moreover, in this case if we let η(f )2 be the least constant k in the Carleson condition, then η(f ) is essentially, the Garsia norm and hence is a norm on BM OA that is equivalent to kfk∗. Proposition 5.1. Let (cn) ∈ X, and set g(z) = P∞ n=0 cnzn, then g ∈ BM OA and Z ZR(I) (1 − r2)g′(reiθ)2rdrdθ ≤ 2Kk(cn)k2I. k=m k2ck2 ≤ An. Note that for 0 < r < 1, Consequently, the linear map L : X → BM OA defined by L((cn)) = g(z) is bounded. Proof. Set A = k(cn)k2 so thatPm+n Z 2π k2ck2r2k = Xp=0 g′(reiθ)2dθ = m2cm2r2m ≤ Xm=p[ 1 1 − r2[ 1 )2 ≤ AK 1 1 − r ]+2A(p + 1)[ 1 1 − r 1 1 − r ∞ Xp=0 ∞ Xk=1 ] ≤ A[ (p+1)[ 1 1−r ] r2p[ 1 1−r 1 ] 1−r 0 ∞ ]+1 1−r ]r2( . Hence, r + r2dr ≤ 2AKI (1 − r2)g′(reiθ)2rdrdθ ≤ AKZ 1 Z ZR(I) and we have shown the Carleson estimate. Hence g ∈ BM OA with η(g) ≤ √2Kk(cn)k. Proof of the generalized Hardy inequality. Let (cn) ∈ X, and let f (z) = P∞ n=0 anzn ∈ H 1. Write an = αnan where αn = 1. Set g(z) =P∞ so that by Proposition 4.1, g is in BM OA with norm independent of the particular choice of α’s. 1−I (cid:3) n=0 αncnzn We have that hf, gi = lim r→1−Z 2π 0 f (reiθ)g(reiθ)dθ = ∞ Xn=0 ancn. 6 V. I. PAULSEN AND D. SINGH Hence, ∞ ancn ≤ kfk1kgk∗ ≤ kLkkfk1k(cn)k Xn=0 and the proof is complete. Proof of the generalized Hilbert Inequality. Given (an), (bn) ∈ ℓ2 (cn) ∈ X, set f (z) =P∞ H 2 and hence, f (z)h(z) =P∞ Defining g(z) =P∞ Xn,m=0 anbmcn+m = hf h, gi ≤ n=0 anzn and h(z) =P∞ n=0 cnzn and arguing as above, we have that n,m=0 anbmzn+m is in H 1. + and n=0 bnzn. Then f, h are in ∞ kLkk(cn)kkf hk1 ≤ kLkk(cn)kk(an)k2k(bn)k2, and the result follows. 6. Slow Decay in X The importance of the classic Hardy inequality, which corresponds to the sequence cn = 1 n+1 , is that it gives us information about the rate of decay of coefficients of H 1 functions. For this reason it is interesting to seek elements (cn) ∈ X which tend to 0 at a slower rate. In particular, we shall show that there exist sequences in X so that sup{nscn : n ∈ N} = ∞ for any s > 1/2. To this end fix an r, 1/2 ≤ r ≤ 1 and fix β > 1. We define cn inductively as follows: Set c1 = 1 so that 12c12 ≤ 1β and 12c12 + (1+1)2 (1+1)2r ≤ (1 + 1)β Assume that c1, ..., cn have been defined so that m (6.1) k2ck2 ≤ mβ for m = 1, 2, ..., n. Xk=1 k=1 k2ck2 + (n+1)2 (n+1)2r ≤ β(n + 1), then we set cn+1 = 1 (n+1)r and we have that (6.1) holds for m = 1, ..., (n + 1). If this inequality is false, then we set cn+1 = 1 k=1 k2ck2 + (n + 1)2cn+12 ≤ βn + 1 ≤ β(n + 1), so again (6.1) holds for m = 1, ..., (n + 1). In this way we define a sequence cn that satisfies (6.1) for all n with n+1 and observe that Pn If Pn cn = 1 nr or cn = 1 n for each n. Now suppose that ck = 1 k for n ≤ k ≤ n + m. Then k2ck2 + (n + m + 1)2 (n + m + 1)2r = n Xk=1 k2ck2 + (m − 1) + (n + m + 1)2−2r ≤ βn + (m − 1) + (n + m + 1)2−2r. n+m Xk=1 Since, βn + (m − 1) + (n + m − 1)2−2r ≤ β(n + m + 1), HARDY AND HILBERT 7 if and only if (n + m− 1)2−2r ≤ (β − 1)m + β + 1 and 2− 2r < 1 we see that eventually, there will exist an m for which this inequality holds and then cn+m+1 = 1 (n+m+1)r . Thus, cn = 1 nr infinitely often. In particular, sup{nscn : n ∈ N} will be infinite for any s > r ≥ 1/2. References [1] P.L. Duren, The Theory of H p spaces, Academic Press, London-New York, 1970. [2] J.B. Garnet, Bounded Analytic Functions, Academic Press, New York, 1970. [3] G.H. Hardy, J.E. Littlewood, and G. Polya, Inequalities, Cambridge University Press, 1952. [4] H. Helson, Harmonic Analysis, Hindustan Book Agency, 1995. [5] H.L. Montgomery and J.D. Vaaler, A further generalization of Hilbert’s inequality, Mathematika 46(1999), No.1, 35-39. [6] K. Oleszkiewicz, An elementary proof of Hilbert’s inequality, Amer. Math. Monthly 100(1993), No. 3, 276-280. [7] M. Sabahheh, On an argument of Korner and Hardy’s inequality, Analysis Mathe- matica 34(2008), 51-57. [8] S. Saeki and H. Yamaguchi, On Hardy’s inequality and Paley’s gap theorem, Hokkaida Math. Journal 19(1990), No. 2, 289-296. [9] Bicheng Yang, On a relation between Hilbert’s inequality and a Hilbert type inequality, Appl. Math. Letters 21(2008), No. 5, 483-488. [10] Kehe Zhu, Translating inequalities between Hardy and Bergman spaces, Amer. Math. Monthly 111(2004), 520-525. Department of Mathematics, University of Houston, Houston, Texas 77204- 3476, U.S.A. E-mail address: [email protected] Department of Mathematics, University of Delhi, Delhi, India
1312.3042
1
1312
2013-12-11T06:21:10
Perturbation of Browder Spectrum of Upper-triangular Operator Matrices
[ "math.FA" ]
Let $M_{C}=\left(\begin{array}{cc}A&C\\0&B\\\end{array} \right)$ is a 2-by-2 upper triangular operator matrix acting on the Banach space $X\oplus Y$ or Hilbert space $H\oplus K$. For the most import spectra such as spectrum, essential spectrum and Weyl spectrum, the characterizations of the perturbations of $M_{C}$ on Banach space have been presented. However, the characterization for the Browder spectrum on the Banach space is still unknown. The goal of this paper is to present some necessary and sufficient conditions for $M_C$ to be Browder for some $C\in B(Y,X)$ by using an alternative approach based on matrix representation of operators and the ghost index theorem. Moreover, in the Hilbert space case the characterizations of the Fredholm and invertible perturbations of Browder spectrum are also given.
math.FA
math
Perturbation of Browder Spectrum of Upper-triangular Operator Matrices∗ Shifang Zhang1†, Huaijie Zhong1, Lin Zhang 2 1School of Mathematics and Computer Science, Fujian Normal University, Fuzhou 350007, P. R. China 2Institute of Mathematics, Hangzhou Dianzi University, Hangzhou 310018, P. R. China 3 1 0 2 c e D 1 1 ] . A F h t a m [ 1 v 2 4 0 3 . 2 1 3 1 : v i X r a Abstract Let MC = (cid:18) A C 0 B (cid:19) is a 2-by-2 upper triangular operator matrix act- ing on the Banach space X ⊕ Y or Hilbert space H ⊕ K. For the most import spectra such as spectrum, essential spectrum and Weyl spectrum, the characterizations of the perturbations of MC on Banach space have been presented. However, the char- acterization for the Browder spectrum on the Banach space is still unknown. The goal of this paper is to present some necessary and sufficient conditions for MC to be Browder for some C ∈ B(Y, X) by using an alternative approach based on matrix representation of operators and the ghost index theorem. Moreover, in the Hilbert space case the characterizations of the Fredholm and invertible perturbations of Browder spectrum are also given. Keywords: operator matrices, Browder spectrum, left semi-Browder spectrum. 1 Introduction It is well-known that if H is a Hilbert space, T is a bounded linear operator on H, and H1 is an invariant closed subspace of T , then T is of the form T = (cid:18) ∗ ∗ 0 ∗ (cid:19) : H1 ⊕ H ⊥ 1 → H1 ⊕ H ⊥ 1 , which motivated the interest in 2 × 2 upper-triangular operator matrices (see [1-9, 11-20]). The problems related to the perturbation of spectra of 2-by-2 upper triangular operator matrices were first studied by H. K. Du and J. Pan in [5], and they given the characterization of MC to be invertible for some C ∈ B(K, H) in the Hilbert spece case. Some years later, in ∗This work is supported by the NSF of China (Grant Nos. 11301077,11301124, 11226113, 11171066 and 10771191 ) , the FED of Fujian Province(Grant No. JA12074), the NSF of Fujian Province (Grant No.2012J05003). †Corresponding author: E-mail: [email protected] 1 [13] J. K. Han et al. generalized the above result to the Banach space case. In 2002, D. S. Djordjevi´c in [5] future studied this field for some other spectra and presented the necessary and sufficient conditions for MC to be a Fredholm and Weyl operator for some C ∈ B(Y, X), respectively. Also in [5], some sufficient conditions for MC to be a Browder operator were given. Recently, Cao [1] gave the necessary and sufficient condition for MC to be a Browder operator in the Hilbert space case. Till now, although for the most import and widely used spectra such as spectrum, essential spectrum and Weyl spectrum, the characterizations of the perturbations of upper triangular operator matrices on Banach space have been given, the characterization for the Browder spectrum is still unknown. The goal of this paper is to present a necessary and sufficient condition for MC to be Browder for some C ∈ B(Y, X) by using an alternative approach based on matrix representation of operators and the ghost index theorem, and our result is an extension of the main result in [1] and the corresponding results from [5, 22]. Moreover, the characterizations of the Fredholm and invertible perturbations of Browder spectrum are also given. Throughout this paper, let H and K be the complex infinite dimensional separable Hilbert spaces, X and Y be the complex infinite dimensional Banach spaces, and B(X, Y ) be the set of all bounded linear operators from X into Y. For simplicity, we write B(X, X) as B(X). For T ∈ B(X), T is called Drazin invertible if there exists an operator T D ∈ B(X) such that T T D = T DT, T DT T D = T D, T k+1T D = T k, for some nonnegative integer k([5]). T ∈ B(X, Y ) is called regular( also called relatively regular in [5]) if there is an operator B ∈ B(Y, X) such that T = T BT ([4]). It is well-known that T is regular if and only if R(T ) and N(T ), respectively, are closed and complemented subspaces of Y and X. For T ∈ B(X, Y ), let R(T ) and N(T ) be the range and kernel of T , respectively, and denote α(T ) = dim N(T ) and β(T ) = dim Y /R(T ). If T ∈ B(X), the ascent, denoted by asc(T ), and the descent, denoted by des(T ), of T are the non-negative integers k defined respectively by asc(T ) = inf {k : N(T k) = N(T k+1)}, des(T ) = inf {k : R(T k) = R(T k+1)}. If such k does not exist, then asc(T ) = ∞, respectively des(T ) = ∞. If the ascent and the descent of T are finite, then they are equal (see [5]). For T ∈ B(X), if R(T ) is closed and α(T ) < ∞, then T is said to be upper semi-Fredholm, if β(T ) < ∞, then T is said to be semi-Fredholm. If T ∈ B(X) is either upper or lower semi-Fredholm , then T is said to be semi-Fredholm. For a semi-Fredholm operator T , its index ind (T ) is defined as ind (T ) = α(T ) − β(T ). The sets of invertible operators, left invertible operators and right invertible operators on X, respectively, are denoted by G(X), Gl(X) and Gr(X). The sets of all Fredholm operators, Weyl operators, left semi-Fredholm operators and right semi-Fredholm operators on X are defined, respectively, by Φ(X) := {T ∈ B(X) : α(T ) < ∞ and β(T ) < ∞}; Φ0(X) := {T ∈ Φ(X) : ind(T ) = 0}; Φl(X) := {T ∈ B(X) : R(T ) is a closed and complemented subspace of X and α(T ) < ∞}; Φr(X) := {T ∈ B(X) : N(T ) is a closed and complemented subspace of X and β(T ) < ∞}. 2 The sets of all Browder operators, left semi-Browder operators and right semi-Browder operators on X are defined, respectively, by Φb(X) := {T ∈ Φ(X) : asc(T ) = des(T ) < ∞}; Φlb(X) := {T ∈ Φl(X) : asc(T ) < ∞}; Φrb(X) := {T ∈ Φr(X) : des(T ) < ∞}. The spectrum, left spectrum, right spectrum, essential spectrum, Weyl spectrum, left semi- Fredholm spectrum, right semi-Fredholm spectrum, Browder spectrum, left semi-Browder spectrum, right semi-Browder spectrum of an operator T ∈ B(H) are respectively denoted by σ(T ), σl(T ), σr(T ), σe(T ), σw(T ),σle(T ), σre(T ), σb(T ), σlb(T ) and σrb(T ). One of the main results is : MC is a Browder operator for some C ∈ B(Y, X) if and only if the following statements are satisfied (a) A ∈ B(X) is a left semi-Browder operator; (b) B ∈ B(Y ) is a right semi-Browder operator; (c) N(A) × N(B) ∼= X/R(A) × Y /R(B). extension of the corresponding results from [1, 5, 22], that is, Using the above result, we characterize completely the set TC∈B(Y, X) σb(MC), which is the σb(MC) = σlb(A)∪σrb(B)∪{λ ∈ C : N(A−λ)×N(B−λ) 6∼= X/R(A−λ)×Y /R(B−λ)}. \C∈B(Y, X) Moreover, for any given A ∈ B(H) and B ∈ B(K), it is proved that \C∈G(K, H) σb(MC) = \C∈Φ(K, H) σb(MC) = σlb(A) ∪ σrb(B) ∪ {λ ∈ C : α(A − λ) + α(B − λ) 6= β(A − λ) + β(B − λ)}. 2 Main Results and Proofs We begin with some lemmas, which are useful for the proofs of the main results. Lemma 2.1. ([23] or [2]) For any given A ∈ B(X) and B ∈ B(Y ), if MC is Drazin invertible for some C ∈ B(Y, X), then (1) des(B) < ∞ and asc(A) < ∞; (2) des(A∗) < ∞ and asc(B∗) < ∞. Lemma 2.2. ([5] Lemma 2.3) Let M and N be finite dimensional spaces. If dim M = dim N and X × M ∼= Y × N, then X ∼= Y. Lemma 2.3. ([13]) If T ∈ B(X, Y ), S ∈ B(Y, Z) and ST ∈ B(X, Z) are regular, then 3 N(T ) × N(S) × Z/R(ST ) ∼= N(ST ) × Y /R(T ) × Z/R(S). Some well-known facts are collected in the following lemma. Lemma 2.4. ([24]) For any given A ∈ B(X), B ∈ B(Y ) and C ∈ B(Y, X), we have: (1) if any two of operators A, B and MC are invertible (resp., Fredholm, Weyl, Brow- der,Drazin invertible ), then so is the third; (2) if A is Browder, then B is left semi-Browder if and only if so is MC ; (3) if B is Browder, then A is right semi-Browder if and only if so is MC. Lemma 2.5([5]). Let A ∈ B(X) and B ∈ B(Y ) be given and consider the statements: (i) MC is a Weyl operator for some C ∈ B(Y, X) (ii) A ∈ B(X) and B ∈ B(Y ) satisfy the following conditions: (a) A is a left semi-Fredholm operator; (b) B is a right semi-Fredholm operator; (c) N(A) × N(B) ∼= X/R(A) × Y /R(B). Then (i) ⇐⇒ (ii) Lemma 2.6 ([13]). Let A ∈ B(X) and B ∈ B(Y ) be given and consider the statements: (i) MC is invertible for some C ∈ B(Y, X) (ii) A ∈ B(X) and B ∈ B(Y ) satisfy the following conditions: (a) A is left invertible; (b) B is right invertible; (c) N(B) ∼= X/R(A). Then (i) ⇔ (ii) Lemma 2.7. For T ∈ B(X), T is left semi-Browder if and only if T can be decomposed into the following form with respect to space decomposition X = X1 ⊕ X2 T = (cid:18) T1 T12 T2 (cid:19) , 0 where dim(X1) < ∞, T1 is nilpotent, and T2 is left invertible. Proof. Necessity. Suppose that T is left semi-Browder. Then we can assume p = asc(T ) < ∞. Let X1 = N(T p). Then dim X1 < ∞ since T is left semi-Fredholm. Therefore, there exists a closed subspace X2 of X such that X1 ⊕ X2 = X. With respect to the space decomposition X = X1 ⊕ X2, we have T = (cid:18) T1 T12 T2 (cid:19) : X1 ⊕ X2 → X1 ⊕ X2. 0 Obviously, T1 is nilpotent and hence T1 ∈ Φb(X1). Moreover, it follows from Lemma 2.4 that T2 ∈ Φlb(X2). Next, we shall show T2 is injective. If not, there exists some 0 6= x ∈ X2 such 4 that T2x = 0. Put z = (cid:18) 0 x (cid:19) ∈ x (cid:19). Then T p+1(cid:18) 0 x (cid:19) = T p(cid:18) T12x 0 (cid:19) = 0, and so 0 6= (cid:18) 0 N(T p+1). This is in contradiction with the assumption that asc(T ) = p < ∞, thus, T2 is left invertible. Sufficiency. It is evident. Lemma 2.8. For T ∈ B(X), T is right semi-Browder if and only if T can be decomposed into the following form with respect to some space decomposition X = X1 ⊕ X2 T = (cid:18) T1 T12 T2 (cid:19) , 0 where dim(X2) < ∞, T1 is right invertible, and T2 is nilpotent. Proof. Necessity. Suppose T is right semi-Browder. Then we can assume p = des(T ) < ∞. Now put X1 = R(T p). Then dim X/X1 < ∞ since T p is a right semi-Browder operator. Therefore, there exists a closed finite dimensional subspace X2 of X such that X1 ⊕ X2 = X. Corresponding the space decomposition X = X1 ⊕ X2, we have T = (cid:18) T1 T12 T2 (cid:19) : X1 ⊕ X2 → X1 ⊕ X2. 0 Obviously, T1 is surjective and T P Φb(X2), it follows from Lemma 2.4 that T1 ∈ Φrb(X2), and so T1 is right invertible. 2 = 0. Moreover, noting that dim X2 < ∞ implies T2 ∈ Sufficiency. It is evident. One of our main results is as follows Theorem 2.9. Let A ∈ B(X) and B ∈ B(Y ) be given and consider the statements: (i) MC is a Browder operator for some C ∈ B(Y, X) (ii) A ∈ B(X) and B ∈ B(Y ) satisfy the following conditions: (a) A is a left semi-Browder operator; (b) B is a right semi-Browder operator; (c) N(A) × N(B) ∼= X/R(A) × Y /R(B). Then (i) ⇔ (ii). Proof. Necessity. Noting the fact that an operator T is a Browder operator if and only if T is a Drazin invertible Weyl operator, the implication follows from Lemmas 2.1 and 2.5 immediately. Sufficiency. Since A is a left semi-Browder operator, from Lemma 2.7, we have A = (cid:18) A1 A12 0 A2 (cid:19) : X1 ⊕ X2 → X1 ⊕ X2, 5 where A1 is nilpotent, A2 is left invertible and dim(X1) < ∞. Note that A = (cid:18) I 0 A2 (cid:19)(cid:18) A1 A12 I (cid:19) . 0 0 Applying Lemma 2.3, we have 0 N((cid:18) I Since (cid:18) I 0 A2 (cid:19))×N((cid:18) A1 A12 0 A2 (cid:19) is injective, 0 0 I (cid:19))×X/R(A) ∼= X/R(cid:18) I 0 A2 (cid:19)×X/R(cid:18) A1 A12 I (cid:19)×N(A). 0 0 0 N((cid:18) A1 A12 I (cid:19)) × X/R(A) ∼= X/R(cid:18) I I (cid:19) × N(A). It follows from dim(X1) < ∞ that A1 ∈ Φb(X1). By Lemma 2.4, we obtain (cid:18) A1 A12 0 A2 (cid:19) × X/R(cid:18) A1 A12 0 0 0 I (cid:19) ∈ Φb(X1 ⊕ X2), which implies α((cid:18) A1 A12 I (cid:19)) = β((cid:18) A1 A12 I (cid:19)) < ∞. 0 0 Using Lemma 2.2, we have X/R(A) ∼= X/R(cid:18) I 0 A2 (cid:19) × N(A), 0 and hence X/R(A) ∼= X2/R(A2) × N(A). (1) Meanwhile, since B ∈ B(Y ) is a right semi-Browder operator, from Lemma 2.8 we have B = (cid:18) B1 B12 0 B2 (cid:19) : Y1 ⊕ Y2 → Y1 ⊕ Y2, where B2 is nilpotent, B1 is right invertible and dim(Y2) < ∞. Similar to the preceding arguments, we have B = (cid:18) I B12 0 B2 (cid:19)(cid:18) B1 0 I (cid:19) , 0 and N((cid:18) I B12 0 B2 (cid:19)) × N((cid:18) B1 0 I (cid:19)) × Y /R(B) ∼= Y /R(cid:18) I B12 0 B2 (cid:19) × Y /R(cid:18) B1 0 I (cid:19) × N(B). 0 0 6 Since (cid:18) B1 0 0 I (cid:19) is surjective, 0 B2 (cid:19)) × N((cid:18) B1 0 N((cid:18) I B12 0 I (cid:19)) × Y /R(B) ∼= Y /R(cid:18) I B12 0 B2 (cid:19) × N(B). It follows from dim(Y2) < ∞ that B2 ∈ Φb(Y2). By Lemma 2.4, we have which implies (cid:18) I B12 0 B2 (cid:19) ∈ Φb(Y ), α((cid:18) I B12 0 B2 (cid:19)) = β((cid:18) I B12 0 B2 (cid:19)) < ∞. Then it follows from Lemma 2.2 that N(B1) × Y /R(B) ∼= N(B). (2) Combining (1) and (2) with the assumption N(A) × N(B) ∼= X/R(A) × Y /R(B), it is easy to show that N(A) × N(B1) × Y /R(B) ∼= X2/R(A2) × N(A) × Y /R(B). Moreover, since α(A) and β(B) are finite, it results from Lemma 2.2 that N(B1) ∼= X2/R(A2). (3) Observe that B1 is right invertible and A2 is left invertible. It follows from Lemma 2.6 that there exists some C0 ∈ B(Y1, X2) such that (cid:18) A2 C0 0 B1 (cid:19) ∈ G(X2 ⊕ Y1). Define C ∈ B(Y, X) as C = (cid:18) 0 C0 0 (cid:19) : Y1 ⊕ Y2 → X1 ⊕ X2. 0 Then MC can be written as the following form: MC = : X1 ⊕ X2 ⊕ Y1 ⊕ Y2 → X1 ⊕ X2 ⊕ Y1 ⊕ Y2. A1 B12 0 0 A2 C0 0 0 0 0 0 B1 B12 0 0 B2     We conclude from Lemma 2.4 that MC = (cid:18) A C 0 B (cid:19) ∈ Φb(X ⊕ Y ). This completes the proof. 7 The next corollary is immediate from Theorem 2.9, which characterizes the setTC∈B(Y, X) σb(MC). Corollary 2.10. Let A ∈ B(X) and B ∈ B(Y ). Then we have \C∈B(Y, X) σb(MC) = σlb(A)∪σrb(B)∪{λ ∈ C : N(A−λ)×N(B−λ) 6∼= X/R(A−λ)×Y /R(B−λ)}. As a particular case of Corollary 2.10, we can rewrite the main result in [1] Corollary 2.11. Let H and K be the complex infinite dimensional separable Hilbert space, A ∈ B(H) and B ∈ B(K). Then \C∈B(K, H) σb(MC) = σlb(A) ∪ σrb(B) ∪ {λ ∈ C : α(A − λ) + α(B − λ) 6= β(A − λ) + β(B − λ)}. can be equal to spectrum, essential spectrum, Weyl spectrum and so on. Similarly, the sets In [25], we have studied the sets TC∈G(K, H) σ∗(MC) and TC∈Φ(K, H) σ∗(MC), where σ∗(·) TC∈G(K, H) σb(MC) and TC∈Φ(K, H) σb(MC) are studied in the next theorem. Firstly, we give a well-known result which is needed in the proof of the next theorem. Lemma 2.12. [18 Corollary 2.2] Let A ∈ B(H) and B ∈ B(K). If R(B) is closed, then MC is left invertible if and only if both A and C1 are left invertible, where C = (cid:18) C1 C2 C3 C4 (cid:19) : N(B) ⊕ N(B)⊥ → R(A)⊥ ⊕ R(A). Duality, we have Corollary 2.13. Let A ∈ B(H) and B ∈ B(K). If R(A) is closed, then MC is right invertible if and only if both B and C1 are right invertible, where C = (cid:18) C1 C2 C3 C4 (cid:19) : N(B) ⊕ N(B)⊥ → R(A)⊥ ⊕ R(A). Combing Lemma 2.12 and Corollary 2.13, we have Corollary 2.14. Let A ∈ B(H) and B ∈ B(K). Then MC is invertible if and only if A is left invertible, B is right invertible and C1 is invertible, where C = (cid:18) C1 C2 C3 C4 (cid:19) : N(B) ⊕ N(B)⊥ → R(A)⊥ ⊕ R(A). We are now ready to give the last main result. Theorem 2.15. For any given A ∈ B(H) and B ∈ B(K) we have 8 \C∈G(K, H) σb(MC) = \C∈Φ(K, H) σb(MC) = SP R, (4) where SP R = σlb(A) ∪ σrb(B) ∪ {λ ∈ C : α(A − λ) + α(B − λ) 6= β(A − λ) + β(B − λ)}. Proof. Noting the fact \C∈G(K, H) σb(MC) ⊇ \C∈Φ(K, H) σb(MC) ⊇ \C∈B(K, H) σb(MC), it follows from Corollary 2.11 that in order to prove (4), we only need to prove \C∈G(K, H) σb(MC) ⊆ σlb(A) ∪ σrb(B) ∪ {λ ∈ C : α(A − λ) + α(B − λ) 6= β(A − λ) + β(B − λ)}. For this, it suffices to show that, if 0 6∈ σlb(A) ∪ σrb(B) ∪ {λ ∈ C : α(A − λ) + α(B − λ) 6= β(A − λ) + β(B − λ)}, then 0 6∈ \C∈G(K, H) σb(MC). Now suppose that A ∈ Φlb(H) and B ∈ Φrb(K) with α(A) + α(B) = β(A) + β(B). We shall prove that there exists some C ∈ G(K, H) such that MC ∈ Φb(H ⊕ K). From Lemmas 2.7 and 2.8, we have A = (cid:18) A1 A12 0 A2 (cid:19) : H1 ⊕ H2 → H1 ⊕ H2, B = (cid:18) B1 B12 0 B2 (cid:19) : K1 ⊕ K2 → K1 ⊕ K2, where dim(H1) < ∞, dim(K2) < ∞, A2 is left invertible, B1 is right invertible, and both A1 and B2 are nilpotent. Similar to the proof of Theorem 2.9, we have Moreover, considering the operator C ′ ∈ B(K1, H2) with the following expression α(B1) = β(A2). (5) C ′ = (cid:18) C ′ C ′ 1 C ′ 2 3 C ′ 4 (cid:19) : N(B1) ⊕ N(B1)⊥ → R(A2)⊥ ⊕ R(A2), it follows from Corollary 2.11 (replacing operators A, B, C by A1, B1, C ′, respectively) that (cid:18) A2 C ′ 0 B1 (cid:19) is invertible if and only if so is C ′ 1, since A2 is left invertible and B1 is right invertible. Furthermore, since dim N(B1)⊥ = dim R(A2) = ∞, we have dim K2 + dim N(B1)⊥ = dim R(A2) + dim H1 = ∞. (6) 9 It follows from (5) and (6) that there exist two invertible operators Q1 ∈ B(N(B1), R(A2)⊥) and Q2 = (cid:18) T1 T2 T3 T4 (cid:19) : N(B1)⊥ ⊕ K2 → R(A2) ⊕ H1. Define an operator C : K → H by C =   Q1 0 0 0 0 T1 T2 T3 T4   : N(B1) ⊕ N(B1)⊥ ⊕ K2 → R(A2)⊥ ⊕ R(A2) ⊕ H1. Obviously C ∈ B(K, H) is invertible. Next we will prove that MC is a Browder operator. Observing that A1 A12 0 0 A′ 2 0 0 0 0 T3 0 0 0 Q1 0 T1 0 B′ 0 T4 0 T2 1 B12 0 B2   : H1 ⊕ H2 ⊕ N(B1) ⊕ N(B1)⊥ ⊕ K2 → H1 ⊕ R(A2)⊥ ⊕   MC = R(A2) ⊕ K1 ⊕ K2. =   A1 A12 C1 T4 0 A2 C3 C2 0 B1 B12 0 0 0 0 B2   : H1 ⊕ H2 ⊕ K1 ⊕ K2 → H1 ⊕ H2 ⊕ K1 ⊕ K2, where A2 = (cid:18) 0 From Corollary 2.14 we have that (cid:18) A2 C3 2 (cid:19) , C1 = (cid:0) 0 T3 (cid:1) , C3 = (cid:18) Q1 T1 (cid:19) , C4 = (cid:18) 0 T2 (cid:19) , B1 = (cid:0) 0 B′ 1 (cid:1) . 0 B1 (cid:19) is invertibility, and then it follows from A′ 0 0 Lemma 2.4 that MC is a Browder operator. This ends the proof. References [1] X. H. Cao. Browder spectra for upper triangular operator matrices, J. Math. Anal. Appl., 342 (2008), 477-484. [2] D. S. Djordjevi´c. Perturbations of spectra of operator matrices, J. Operator Theory, 48 (2002), 467-486. [3] D. S. Djordjevi´c, M. Z. Kolundzija. Generalized invertibility of operator matrices, Ark. Mat., 50 (2012), 259-267. [4] S. V. Djordjevi´c, Y. M. Han. spectral continuity for operator matrices, Glasg. Math. J., 43 (2001), 487-490. [5] S. V. Djordjevi´c, B. P. Duggal, Drazin invertibility of the diagonal of an operator, Linear Multilinear Algebra 60 (2012), 65-71. 10 [6] H. K. Du, J. Pan. Perturbation of spectrum of 2 × 2 operator matrices, Proc. Amer. Math. Soc., 121 (1994), 761-766. [7] G. J. Hai, A. Chen. Perturbations of right and left spectra for operator matrices, J. Operator Theory, 67 (2012), 207-214. [8] G. J. Hai, A. Chen. On the right (left) invertible completions for operator matrices. Integral Equations Operator Theory, 67 (2010), 79-93. [9] J. K. Han, H. Y. Lee, W. Y. Lee. Invertible completions of 2 × 2 upper triangular operator matrices. Proc. Amer. Math. Soc., 128 (1999), 119-123. [10] R. E. Harte. The ghost of an index theorem. Proc. Amer. Math. Soc., 106 (1989), 1031-1034. [11] I. S. Hwang, W. Y. Lee. The boundedness below of 2 × 2 upper triangular operator matrices, Integr. Equ. Oper. Theory, 39 (2001), 267-276. [12] W. Y. Lee. Weyl's theorem for operator matrices, Integr. equ. oper. theory, 32 (1998), 319-331. [13] W. Y. Lee. Weyl spectra of operator matrices, Proc. Amer. Math. Soc., 129 (2000), 131-138. [14] Y. Li, X. H. Sun, H. K. Du. The intersection of left(right) spectra of 2 × 2 upper triangular operator matrices, Linear Algebra Appl., 418(2006), 112-121. [15] E. H. Zerouali, H. Zguitti. Perturbation of spectra of operator matrices and local spectral theory, J. Math. Anal. Appl., 324(2006), 992-1005. [16] S. F. Zhang, H. J. Zhong, Q. F. Jiang. Drazin spectrum of operator matrices on the Banach space, Linear Algebra Appl., 429 (2008), 2067-2075. [17] S. F. Zhang , Z. Y. Wu, H. J. Zhong, Continuous spectrum, point spectrum and residual spectrum of operator matrices, Linear Algebra Appl., 433 (2010), 653-661. [18] S. F. Zhang, H. J. Zhong, J. D. Wu. Spectra of 2 × 2 Upper Triangular Matrices, Acta Math. Sinica (in Chinese), 54(1) (2011) 41-60. [19] S. F. Zhang, J. D. Wu. Samuel Multiplicites and Browder spectrum of operator matrices, Operators and Matrices, 6 (2012), 169-179. [20] S. F. Zhang, Z. Q. Wu. Characterizations of perturbations of spectra of 2 × 2 upper triangular operator matrices, J. Math. Anal. Appl. 392 (2012), 103-110. 11
1606.06845
1
1606
2016-06-22T08:18:29
Means refinements via convexity
[ "math.FA" ]
The main goal of this article is to find the exact difference between a convex function and its secant, as a limit of positive quantities. This idea will be expressed as a convex inequality that leads to refinements and reversals of well established inequalities treating different means. The significance of these inequalities is to write one inequality that brings together and refine almost all known inequalities treating the arithmetic, geometric, harmonic and Heinz means, for numbers and operators.
math.FA
math
MEANS REFINEMENTS VIA CONVEXITY M. SABABHEH Abstract. The main goal of this article is to find the exact difference between a convex function and its secant, as a limit of positive quantities. This idea will be expressed as a convex inequality that leads to refinements and reversals of well established inequalities treating different means. The significance of these inequalities is to write one inequality that brings together and refine almost all known inequalities treating the arithmetic, geometric, harmonic and Heinz means, for numbers and operators. 1. introduction Convex functions and their inequalities have played a major role in the study of various topics in Mathematics; including applied Mathematics, Mathematical Analysis and Mathematical Physics. Means and their comparison is indeed an important application of convexity. Recall that a function f : I → R, defined on a real interval I, is said to be convex if f (αx1+βx2) ≤ αf (x1)+βf (x2), when x1, x2 ∈ I and α, β ≥ 0 satisfying α+β = 1. On the other hand, f : I → R+ is said to be log-convex if g(x) = log f (x) is convex, or equivalently if f (αx1 + βx2) ≤ f α(x1)f β(x2) for the above parameters. Speaking of means, the comparison between the weighted arithmetic, geometric and harmonic means is an immediate consequence of convexity or log-convexity of the functions x∇ty = (1 − t)x + ty, x#ty = x1−tyt and x!ty = ((1 − t)x−1 + ty−1)−1, x, y > 0, defined for 0 ≤ t ≤ 1. Adopting these notations, we drop t when t = 1 2 . Convexity of the function f (t) = x#ty implies the well known Young's inequality x#t ≤ x∇ty. On the other hand, convexity of the function g(t) = x!ty implies the arithmetic-harmonic mean inequality x!ty ≤ x∇ty, while log-convexity of g implies the geometric-harmonic mean inequality x!ty ≤ x#ty. These inequalities, though very simple, have some significant applications. For example, the above Young's inequality implies the celebrated Holder's inequality kf gk1 ≤ kfkpkgkq for f ∈ Lp(X) and g ∈ Lq(X), for the conjugate exponents p, q, where X is some measure space. Among the most interesting applications of the above mean inequalities is the possible comparison between operators acting on a finite dimensional Hilbert In the sequel, Mn will denote the space of operators acting on an space H. n−deimentional Hilbert space H, M+ n will denotes the cone of semi positive operators in Mn while M++ n will denotes the cone of strictly positive opera- tors in Mn. Then the above numerical inequalities have their operator versions 2010 Mathematics Subject Classification. 15A39, 15B48, 26D15, 26B25, 47A30, 47A63. Key words and phrases. convex functions, means inequalities, norm inequalities. 1 2 M. SABABHEH 2 BA− 1 1 A 1 2 (cid:16)A− 1 n , A∇tB = (1 − t)A + tB and 2 . In this context, we say that A ≤ B for two self- such as A#tB ≤ A#tB, where A, B ∈ M++ A#tB = A adjoint operators A and B if B − A ∈ M+ n . Obtaining the operator versions from the corresponding numerical versions can be done in different approaches, among which is the application of the following lemma [2]. 2(cid:17)t Lemma 1.1. Let X ∈ Mn be self-adjoint and let f and g be continuous real valued functions such that f (t) ≥ g(t) for all t ∈ Sp(X), the spectrum of X. Then f (X) ≥ g(X). Recent studies of the topic have investigated possible refinements of the above inequalities, where adding a positive term to the left side becomes possible. This idea has been treated in [3, 5, 6, 7, 9, ?, 10, 11, 12], where not only refinements have been investigated, but reversed versions and much more have been discussed. Keeping our paper concise, we will not go through the exact results done in the above references now, however we will comment later how the results in this paper generalize almost all results in these references, regarding the refinements and the reverses of the above mean inequalities. The main goal of this article is to avoid dealing with the specific means, and to treat a general convexity argument that leads to these refinements. In particular, we prove that for certain positive quantities Aj(ν)∆jf (ν; a, b), we have f ((1 − ν)a + νb) + N Xj=1 Aj(ν)∆jf (ν; a, b) ≤ (1 − ν)f (a) + νf (b), N ∈ N, for the convex function f : [a, b] → R. This provides N refining terms of the inequality f ((1 − ν)a + νb) ≤ (1 − ν)f (a) + νf (b), which follows from convexity of f . Furthermore, we prove a reversed version and we prove that as N → ∞ the above inequality becomes an equality. As a natural consequence, we obtain some refinements and reverses for log-convex functions. As we will see, the above inequality and its consequences happen to be general- izations that imply almost all inequalities in the references [3, 5, 6, 9, ?, 10, 11, 12]. This is our main motivation behind this work; to find a formula that implies and generalizes all other formulae and hence, to enhance our understanding of these inequalities. We remark that the proof of the first main result in this work is inspired by our recent work in [?]. 2. main results For the rest of the paper, the following notations will be adopted. For 0 ≤ ν ≤ 1 and j ∈ N, let ( kj(ν) = [2j−1ν], rj(ν) = [2jν] and Aj(ν) = (−1)rj(ν)2j−1ν + (−1)rj (ν)+1h rj(ν)+1 2 i . (2.1) MEANS REFINEMENTS VIA CONVEXITY 3 Moreover, if f : [a, b] → R is any function, define ∆jf (ν; a, b) = f(cid:18)(cid:18)1 − − 2f(cid:18)(cid:18)1 − kj(ν) 2j−1 (cid:19) a + 2kj(ν) + 1 2j kj(ν) 2j−1 (cid:19) a + b(cid:19) + f(cid:18)(cid:18)1 − 2kj(ν) + 1 kj(ν) + 1 2j−1 (cid:19) a + b(cid:19) , 0 ≤ ν ≤ 1. 2j kj(ν) + 1 2j−1 b(cid:19) (2.2) 2.1. Convex functions. We discuss first the inequalities that govern convex functions, then we apply these inequalities to log-convex functions. Lemma 2.1. If f : [a, b] → R is convex, then ∆jf (ν; a, b) ≥ 0 for j ∈ N and 0 ≤ ν ≤ 1. 2j−1 b, yj(ν) = (cid:16)1 − kj(ν)+1 Proof. Letting xj(ν) = (cid:16)1 − kj (ν) and zj(ν) = (cid:16)1 − 2kj(ν)+1 ∆jf (ν; a, b) = f (xj(ν)) + f (yj(ν)) − 2f (zj(ν)) ≥ 0, by convexity of f . 2j−1 (cid:17) a + kj (ν)+1 b, it is easy that zj(ν) = xj(ν)+yj (ν) 2j−1(cid:17) a + kj(ν) (cid:17) a + 2kj(ν)+1 2j−1 b . The (cid:3) 2j 2j 2 Remark 2.2. When f : [a, b] → R, we adopt the convention that f (x) = 0 for x 6∈ [a, b]. This convention will be needed, for example, in the next lemma, when N = 1 and ν = 1. Lemma 2.3. Let f : [0, 1] → R be a function and let N ∈ N. Then N (1 − ν)f (0) + νf (1) − Aj(ν)∆jf (ν; 0, 1) Xj=1 = (cid:0)[2N ν] + 1 − 2N ν(cid:1) f(cid:18) [2N ν] 2N (cid:19) +(cid:0)2N ν − [2N ν](cid:1) f(cid:18)[2N ν] + 1 2N (cid:19) . (2.3) 2 , r1(ν) = 0 and k1(ν) = 0. Hence A1(ν) = ν and Proof. We proceed by induction on N. When N = 1 and 0 ≤ ν < 1 ∆1f (ν; a, b) = f (a) + f (b)− 2f(cid:0) a+b 2 (cid:1) . Then direct computations show the result. Now if 1 2 ≤ ν < 1, then r1(ν) = 1 and k1(ν) = 0, hence A1(ν) = 1 − ν and ∆1f (ν; a, b) = f (a)+f (b)−2f(cid:0) a+b 2 (cid:1) . Again, direct computations show the result. When ν = 1, the result follows immediately. Now assume that (2.3) is true for some N ∈ N. We assert its truth for N + 1. 4 M. SABABHEH Notice that, using the inductive step, (1 − ν)f (0) + νf (1) − N +1 Xj=1 Aj(ν)∆jf (ν; 0, 1) N Aj(ν)∆jf (ν; 0, 1) − AN +1(ν)∆N +1f (ν; 0, 1) = (1 − ν)f (0) + νf (1) − Xj=1 = (cid:0)[2N ν] + 1 − 2N ν(cid:1) f(cid:18)[2N ν] 2N (cid:19) +(cid:0)2N ν − [2N ν](cid:1) f(cid:18)[2N ν] + 1 − (cid:18)(−1)[2N +1ν]2N ν + (−1)[2N +1ν]+1(cid:20)[2N +1ν] + 1 (cid:21)(cid:19) × × (cid:18)f(cid:18) [2N ν] (cid:19) − 2f(cid:18) 2[2N ν] + 1 2N +1 (cid:19)(cid:19) . 2N (cid:19) + f(cid:18) [2N ν] + 1 2N 2N 2 (cid:19) (2.4) Now we treat two cases. Case I If [2N +1ν] is odd, then we easily see that [2N ν] = [2N +1ν]−1 2 . Therefore, f(cid:18)[2N ν] + 1 2N (cid:19) = f(cid:18)[2N +1ν] + 1 2N +1 (cid:19) and f(cid:18)2[2N ν] + 1 2N +1 (cid:19) = f(cid:18) [2N +1ν] 2N +1 (cid:19) . Substituting these values in (2.4) and simplifying imply N +1 (1 − ν)f (0) + νf (1) − Aj(ν)∆j f (ν; 0, 1) Xj=1 Xj=1 = (cid:0)2N +1ν − [2N +1ν](cid:1) f(cid:18) [2N +1ν] + 1 2N +1 (cid:19) +(cid:0)[2N +1ν] + 1 − 2N +1ν(cid:1) f(cid:18) [2N +1ν] 2N +1 (cid:19) , which completes the proof, when [2N +1ν] is odd. Case II If [2N +1ν] is even, then 2[2N ν] = [2N +1ν] and f(cid:18)[2N ν] 2N (cid:19) = f(cid:18) [2N +1ν] 2N +1 (cid:19) and f(cid:18) 2[2N ν] + 1 2N +1 (cid:19) = f(cid:18)[2N +1ν] + 1 2N +1 (cid:19) . Substituting these values in (2.4) and simplifying imply N +1 (1 − ν)f (0) + νf (1) − Aj(ν)∆j f (ν; 0, 1) = (cid:0)2N +1ν − [2N +1ν](cid:1) f(cid:18) [2N +1ν] + 1 2N +1 This completes the proof. (cid:19) +(cid:0)[2N +1ν] + 1 − 2N +1ν(cid:1) f(cid:18) [2N +1ν] 2N +1 (cid:19) . (cid:3) Corollary 2.4. Let f : [0, 1] → R be convex and let N ∈ N. Then f (ν) + N Xj=1 Aj(ν)∆jf (ν; 0, 1) ≤ (1 − ν)f (0) + νf (1). (2.5) MEANS REFINEMENTS VIA CONVEXITY 5 Proof. From Lemma 2.3 and convexity of f , we have N (1 − ν)f (0) + νf (1) − Aj(ν)∆j f (ν; 0, 1) Xj=1 = (cid:0)[2N ν] + 1 − 2N ν(cid:1) f(cid:18)[2N ν] ≥ f(cid:18)(cid:0)[2N ν] + 1 − 2N ν(cid:1) 2N (cid:19) +(cid:0)2N ν − [2N ν](cid:1) f(cid:18) [2N ν] + 1 (cid:19) 2N +(cid:0)2N ν − [2N ν](cid:1) [2N ν] + 1 [2N ν] 2N 2N = f (ν). This completes the proof. (cid:19) (cid:3) Now our first main result in its general form can be stated as follows. Theorem 2.5. Let f : [a, b] → R be convex. Then for each N ∈ N and 0 ≤ ν ≤ 1, we have f ((1 − ν)a + νb) + N Xj=1 Aj(ν)∆jf (ν; a, b) ≤ (1 − ν)f (a) + νf (b). (2.6) Proof. For the given f , define g : [0, 1] → R by g(x) = f ((1− x)a + xb). Then g is convex on [0, 1]. Applying Corollary 2.4 on the function g implies the result. (cid:3) Remark 2.6. We remark that a negative version of the above theorem has been recently shown in [8]. Namely, it was proved N (1 + ν)f (a) − νf (b) + (cid:19)  (2.7) for the convex function f : R → R. However, the method of proof is considerably easier than the above proofs and the applications are different. (cid:17) f (a) + f(cid:16) (2j−1−1)a+b Xj=1 ≤ f ((1 + ν)a − νb) , ν ≥ 0, a < b, − f(cid:18)(2j − 1)a + b 2jν  2j−1 2j 2 Our next step is to prove a reversed version of (2.6). Theorem 2.7. Let f : [a, b] → R be convex and let N ∈ N. Then for 0 ≤ ν ≤ 1 2 , N we have f ((1 − ν)a + νb) + (1 − A1(ν))∆1f (ν; a, b) Xj=1 ≥ (1 − ν)f (a) + νf (b) + 2 ≤ ν ≤ 1, we have On the other hand, if 1 f ((1 − ν)a + νb) + (1 − A1(ν))∆1f (ν; a, b) Xj=1 ≥ (1 − ν)f (a) + νf (b) + N Aj(1 − 2ν)∆jf(cid:18)1 − 2ν; a + b 2 , b(cid:19) . Aj(2 − 2ν)∆jf(cid:18)2 − 2ν; a, a + b 2 (cid:19) . 6 M. SABABHEH 2 , we have Proof. For 0 ≤ ν ≤ 1 = 2νf(cid:18) a + b ≥ N f ((1 − ν)a + νb) + (1 − A1(ν))∆1f (ν; a, b) − ((1 − ν)f (a) + νf (b)) 2 (cid:19) + (1 − 2ν)f (b) + f ((1 − ν)a + νb) − 2f(cid:18) a + b 2 (cid:19) + (1 − 2ν)b(cid:19) Aj(1 − 2ν)∆jf(cid:18)1 − 2ν; , b(cid:19) + f(cid:18)2ν a + b a + b 2 2 N Aj(1 − 2ν)∆jf(cid:18)1 − 2ν; Xj=1 2 (cid:19) +f ((1 − ν)a + νb) − 2f(cid:18) a + b Xj=1 2 (cid:19) +f ((1 − ν)a + νb) − 2f(cid:18) a + b Xj=1 Aj(1 − 2ν)∆jf(cid:18)1 − 2ν; a + b a + b 2 2 N = ≥ , b(cid:19) + f (νa + (1 − ν)b) , b(cid:19) , where the last line follows from convexity of f , where one has f (νa + (1 − ν)b) + f ((1 − ν)a + νb) ≥ 2f(cid:18) νa + (1 − ν)b + (1 − ν)a + νb 2 ≤ ν ≤ 1. 2 This completes the proof for 0 ≤ ν ≤ 1 inequality for 1 (cid:19) = 2f(cid:18) a + b 2 (cid:19) . 2 . Similar computations imply the desired (cid:3) In fact, the above reversed version turns out to be equivalent to convexity. Proposition 2.8. Let f : I → R be a function defined on the interval I. Assume that for all a < b in I and all 0 ≤ ν ≤ 1, we have f ((1 − ν)a + νb) + (1 − A1(ν))∆1f (ν; a, b) ≥ (1 − ν)f (a) + νf (b), (2.8) then f is convex on I. Proof. Observe that when 0 ≤ ν ≤ 1 f ((1 − ν)a + νb) + (1 − ν)(cid:18)f (a) + f (b) − 2f(cid:18) a + b 2 , (2.8) is equivalent to or On the other hand, if 1 f(cid:18) a + b 1 2 − 2ν 2 (cid:19) ≤ f ((1 − ν)a + νb) + 2 ≤ ν ≤ 1, (2.8) is equivalent to 2 (cid:19) ≤ 2ν − 1 f (a) + 1 2ν 2ν f(cid:18) a + b 2 (cid:19)(cid:19) ≥ (1 − ν)f (a) + νf (b), 1 − 2ν 2 − 2ν f (b). (2.9) f ((1 − ν)a + νb) . (2.10) MEANS REFINEMENTS VIA CONVEXITY 7 Let x1 < x2 ∈ I and let 0 < λ < 1. We assert that f ((1 − λ)x1 + λx2) ≤ (1 − λ)f (x1) + λf (x2). If 0 < λ ≤ 1 2 , let ν = 1 − 2λ 2(1 − λ) Then one can easily check that when 0 < λ ≤ 1 With these choices, we have , a = (2 − 2λ)x1 + (2λ − 1)x2 and b = x2. 2 , we have 0 < ν ≤ 1 2 and a < b. a + b 2 = (1 − λ)x1 + λx2 and (1 − ν)a + νb = x1. Substituting these quantities in (2.9) implies f ((1− λ)x1 + λx2) ≤ (1− λ)f (x1) + λf (x2). This proves the desired inequality for 0 < λ ≤ 1 2 . Now if 1 2 ≤ λ < 1, let ν = 1 2λ , a = x1 and b = (1 − 2λ)x1 + 2λx2. With these choices, we have 1 quantities in (2.10) implies the desired inequality for 1 the proof. 2 < ν ≤ 1 and a < b. Now substituting these 2 ≤ λ < 1. This completes (cid:3) As for the geometric meaning of these refinements, it turns out we are dealing with the interpolation of the function f over the dyadic partition. Proposition 2.9. Let f : [0, 1] → R be any function, and let N ∈ N. Then, if νi = i 2N for some i = 0, 1,· · · , 2N , we have f (νi) + N Xj=1 Aj(νi)∆jf (νi; 0, 1) = (1 − νi)f (0) + νif (1). (2.11) Proof. Observe that when νi = i we have 2N , we have [2N νi] = 2N νi = i. From Lemma 2.3, N (1 − νi)f (0) + νif (1) − Aj(νi)∆jf (νi; 0, 1) Xj=1 = (cid:0)[2N νi] + 1 − 2N νi(cid:1) f(cid:18)[2N νi] = f(cid:18) i 2N(cid:19) = f (νi). 2N (cid:19) +(cid:0)2N νi − [2N νi](cid:1) f(cid:18) [2N νi] + 1 2N (cid:19) This completes the proof. Proposition 2.10. Let f : [0, 1] → R be a given function. If f is continuous, then (cid:3) f (ν) + lim N→∞ uniformly in ν ∈ [0, 1]. N Xj=1 Aj(ν)∆jf (ν; 0, 1) = (1 − ν)f (0) + νf (1), (2.12) 8 M. SABABHEH Proof. Let N ∈ N and define the function gN (ν) = N Xj=1 Aj(ν)∆jf (ν; 0, 1). From Proposition 2.9, we have g(νi) = (1− νi)f (0) + νif (1)− f (νi), when νi = i for some i = 1,· · · , 2N . Noting the definitions of Aj and ∆jf , one can easily see that gN is linear on each dyadic interval Ii := (cid:2) i 2N(cid:3) , i = 0,· · · , 2N − 1. Now since gN is linear on Ii and gN coincides with the continuous function h(ν) := (1− ν)f (0) + νf (1)− f (ν), it follows that gN is the linear interpolation of h at the dyadic partition of [0, 1]. Since f is continuous, it follows that gN → h uniformly, completing the proof. 2N , i+1 2N (cid:3) 2.2. Log-Convex function. The proof of the following result follows from The- orem 2.5 on replacing f by log f. Corollary 2.11. Let f : [a, b] → (0,∞) be log-convex. Then for 0 ≤ ν ≤ 1 and N ∈ N, we have f ((1 − ν)a + νb) N Yj=1(cid:18) f (xj(ν))f (yj(ν)) f 2(zj(ν)) (cid:19)Aj (ν) ≤ f 1−ν(a)f ν(b), (2.13) where xj(ν), yj(ν) and zj(ν) are as in the proof of Lemma 2.1. On the other hand, applying Theorem 2.7 implies the following. Corollary 2.12. Let f : [a, b] → (0,∞) be log-convex. Then for 0 ≤ ν ≤ 1 N ∈ N, we have f ((1 − ν)a + νb) f (a)f (b) 2 ) !1−A1(ν) ≥ f 1−ν(a)f ν(b) Yj=1(cid:18) f (tj(ν))f (uj(ν)) f 2(wj(ν)) (2.14) where tj(ν), uj(ν) and wj(ν) are obtained from the above xj(ν), yj(ν) and zj(ν) f 2( a+b 2 and (cid:19)Aj (1−2ν) N On the other hand, if 1 on replacing (ν, a, b) by (cid:0)1 − 2ν, a+b 2 , b(cid:1) . 2 ≤ ν ≤ 1, we have 2 ) !1−A1(ν) f ((1 − ν)a + νb) f (a)f (b) ≥ f 1−ν(a)f ν(b) f 2( a+b (2.15) where tj(ν), uj(ν) and wj(ν) are obtained from the above xj(ν), yj(ν) and zj(ν) on replacing (ν, a, b) by (cid:0)2 − 2ν, a, a+b 2 (cid:1) . The following is a squared additive version for log-convex functions. This inequality will help prove some squared versions of certain means. N Yj=1(cid:18) f (tj(ν))f (uj(ν)) f 2(wj(ν)) (cid:19)Aj (2−2ν) , , MEANS REFINEMENTS VIA CONVEXITY 9 Theorem 2.13. Let f : [a, b] → [0,∞) be log-convex. Then for 0 ≤ ν ≤ 1 and N ≥ 2, we have N f 2((1 − ν)a + νb) + A2 1(ν)∆1f 2(ν; a, b) + Xj=2 ≤ ((1 − ν)f (a) + νf (b))2 . Aj(ν)∆j f 2(ν; a, b) Proof. We prove the result for 0 ≤ ν ≤ 1 g = f 2 is log-convex too, and hence is convex. Therefore, Theorem 2.5 implies 2 . Since f is log-convex, it follows that N g((1 − ν)a + νb) + Xj=1 which implies, for 0 ≤ ν ≤ 1 2 , Aj(ν)∆jg(ν; a, b) ≤ (1 − ν)g(a) + νg(b), N f 2((1 − ν)a + νb) + ν2∆1f 2(ν; a, b) + Aj(ν)∆jf 2(ν; a, b) Xj=2 ≤ ((1 − ν)f (a) + νf (b))2 + H(ν; a, b), (2.16) where H(ν; a, b) = (1 − ν)f 2(a) + νf 2(b) + ν2∆1f 2(ν; a, b) − ν∆1f 2(ν; a, b) − ((1 − ν)f (a) + νf (b))2 = 2ν(1 − ν)(cid:18)f 2(cid:18) a + b ≤ 0, 2 (cid:19) − f (a)f (b)(cid:19) where the last inequality follows from log-convexity of f . Since H(ν; a, b) ≤ 0, it follows from (2.16) that f 2((1 − ν)a + νb) + ν2∆1f 2(ν; a, b) + Xj=2 ≤ ((1 − ν)f (a) + νf (b))2 . Similar computations imply the result for 1 2 ≤ ν ≤ 1. N Aj(ν)∆jf 2(ν; a, b) (cid:3) Then reversed squared versions maybe obtained in a similar way from Theorem we have 2.7 as follows. Theorem 2.14. Let f : [a, b] → [0,∞) be log-convex and N ∈ N. If 0 ≤ ν ≤ 1 2 , f 2((1 − ν)a + νb) + (1 − ν)2∆1f 2(ν; a, b) + 2ν(1 − ν)(cid:18)f (a)f (b) − f 2(cid:18) a + b 2 (cid:19)(cid:19) ≥ ((1 − ν)f (a) + νf (b))2 N + Xj=1 Aj(1 − 2ν)∆jf 2(cid:18)1 − 2ν; a + b 2 , b(cid:19) . 10 M. SABABHEH If 1 2 ≤ ν ≤ 1, we have f 2((1 − ν)a + νb) + ν2∆1f 2(ν; a, b) + 2ν(1 − ν)(cid:18)f (a)f (b) − f 2(cid:18) a + b 2 (cid:19)(cid:19) ≥ ((1 − ν)f (a) + νf (b))2 N + Xj=1 Aj(2 − 2ν)∆jf 2(cid:18)2 − 2ν; a, a + b 2 (cid:19) . 3. Application 3.1. Refinements of means inequalities. In this section we present some in- teresting applications of the above inequalities. The first result is the following refinement of Young's inequality. Corollary 3.1. Let x, y > 0, N ∈ N and 0 ≤ ν ≤ 1. Then x#ν y + N Xj=1 Aj(ν)(cid:18) 2jqykj(ν)x2j−1−kj (ν) − 2jqykj(ν)+1x2j−1−kj(ν)−1(cid:19)2 ≤ x∇νy. (3.1) Proof. This follows from Theorem 2.5, on letting f (t) = x1−tyt, a = 0, b = 1. Then f is convex. Moreover, direct computations show that ∆jf (ν; 0, 1) =(cid:18) 2jqykj(ν)x2j−1−kj(ν) − 2jqykj(ν)+1x2j−1−kj (ν)−1(cid:19)2 . (cid:3) The above theorem has been recently proved in [?] as a refinement of Young's inequality. This inequality refines the corresponding refinements appearing in [5] and [11], where the inequality was proved only for N = 1, 2. On the other hand, letting f (t) = x!ty, the weighted harmonic mean, we obtain the following refinement of the arithmetic-harmonic mean inequality. Corollary 3.2. Let x, y > 0, N ∈ N and 0 ≤ ν ≤ 1. Then N x!ν y + Aj(ν)(cid:0)x!αj (ν)y + x!βj(ν)y − 2x!γj (ν)y(cid:1) ≤ x∇ν y, Xj=1 2j−1 , βj(ν) = [2j−1ν]+1 2j−1 2 and γj(ν) = αj (ν)+βj (ν) . where αj(ν) = [2j−1ν] (3.2) This inequality is a significant refinement of the corresponding inequality in [12], where the inequality was proved only for N = 1. Now noting log-convexity of the function t 7→ x!ty on [0, 1], and applying Corollary 2.11, we get the following multiplicative refinement of the geometric- harmonic mean inequality. Corollary 3.3. Let x, y > 0, N ∈ N and 0 ≤ ν ≤ 1, we have ≤ x#ν y, x!ν y N Yj=1(cid:18)(x!αj (ν)y)(x!βj(ν)y) (x!γj (ν)y)2 (cid:19)Aj(ν) MEANS REFINEMENTS VIA CONVEXITY 11 where αj(ν) = [2j−1ν] 2j−1 , βj(ν) = [2j−1ν]+1 2j−1 and γj(ν) = αj (ν)+βj (ν) 2 . When N = 1, Corollary 3.3, reduces to and x#y(cid:19)2ν (x!ν y)(cid:18)x∇y x#y(cid:19)2(1−ν) (x!ν y)(cid:18)x∇y ≤ x#ν y, 0 ≤ ν ≤ 1 2 ≤ x#ν y, 1 2 ≤ ν ≤ 1. x#y(cid:17)2 The constant (cid:16) x∇y appearing in these inequalities is called the Kantorovich constant, and has appeared in recent refinements of these mean inequalities. One can see [6] as a recent reference treating some inequalities using this constant. As for the squared version, applying Theorem 2.13 to the log-convex functions t 7→ x#ty and t 7→ x!ty implies the following. The first inequality refines the corresponding results in [3] and [11], while the other inequality is new. Corollary 3.4. Let x, y > 0, 0 ≤ ν ≤ 1, N ≥ 2 and αj(ν) = Then kj(ν) 2j−1 , βj(ν) = and γj(ν) = kj(ν) + 1 2j−1 αj(ν) + βj(ν) . 2 (x#ν y)2 + A2 1(ν)(x − y)2 + ≤ (x∇ν y)2, and N Xj=2 Aj(ν)(cid:0)x1−αj (ν)yαj(ν) − x1−βj (ν)yβj(ν)(cid:1)2 (x!ν y)2 + 2A2 1(ν)(x2∇y2 − (x!y)2) + ≤ (x∇νy)2. N Xj=2 Aj(ν)(cid:0)(x!αj (ν)y)2 + (x!βj (ν)y)2 − 2(x!γj (ν)y)2(cid:1) have 3.2. Reversed Version. Applying Theorem 2.7 to the function f (t) = x#ty implies the following reversed version of Young's inequality. Corollary 3.5. For x, y > 0, let f (t) = x#ty and let N ∈ N. If 0 ≤ ν ≤ 1 x#ν y + (1 − ν)(√x − √y)2 ≥ x∇ν y + , 1(cid:19) . Xj=1 2 ≤ ν ≤ 1, we have Xj=1 x#ν y + ν(√x − √y)2 ≥ x∇ν y + Aj(2 − 2ν)∆jf(cid:18)2 − 2ν; 0, Aj(1 − 2ν)∆jf(cid:18)1 − 2ν; On the other hand, if 1 1 2(cid:19) . 2 , we 1 2 N N These inequalities refine those in [5] and [11]. Then an arithmetic-harmonic reversed version maybe obtained by applying Theorem 2.7 to the function f (t) = x!ty as follows. 12 M. SABABHEH Xj=1 2 ≤ ν ≤ 1, we have Xj=1 N have Corollary 3.6. For x, y > 0, let f (t) = x#ty and let N ∈ N. If 0 ≤ ν ≤ 1 , 1(cid:19) . Aj(1 − 2ν)∆jf(cid:18)1 − 2ν; x!ν y + (1 − ν) (x + y − 2x!y) ≥ x∇ν y + 1 2 N 2 , we On the other hand, if 1 x!νy + ν(x + y − 2x!y) ≥ x∇νy + Aj(2 − 2ν)∆jf(cid:18)2 − 2ν; 0, 1 2(cid:19) . These inequalities refine those in [6]. Similarly, noting log-convexity of the function f (t) = x!ty, we may apply Corol- lary 2.12 to obtain reversed multiplicative version of the harmonic-geometric mean inequality. We leave the application to the reader. Following the same guideline, we may obtain reversed squared versions by applying Theorem 2.14 to the functions t 7→ x#ty and x 7→ x!ty. Observe that when f (t) = x#ty we have f (a)f (b)−f 2(cid:0) a+b 2 (cid:1) = 0. Therefore, applying Theorem 2.14 implies the following inequalities, which refine the corresponding inequalities in [3] and [11]. Corollary 3.7. Let x, y > 0 and N ≥ 1. If 0 ≤ ν ≤ 1, we have (x#ν y)2 + (1 − ν)2(x − y)2 ≥ (x∇ν y)2 + If 1 Aj(1 − 2ν)∆jf 2(cid:18)1 − 2ν; , 1(cid:19) . Xj=1 1 2 N 2 ≤ ν ≤ 1, we have (x#ν y)2 + ν2(x − y)2 ≥ (x∇ν y)2 + N Xj=1 Aj(2 − 2ν)∆jf 2(cid:18)1 − 2ν; 0, 1 2(cid:19) . Now letting g(t) = x!ty we obtain the following new inequalities for the arithmetic- harmonic means. Corollary 3.8. Let x, y > 0 and N ≥ 1. If 0 ≤ ν ≤ 1, we have (x!ν y)2 + 2(1 − ν)2(x2∇y2 − (x!y)2) + 2ν(1 − ν) xy −(cid:18) 2xy x + y(cid:19)2! ≥ (x∇ν y)2 + N Xj=1 Aj(1 − 2ν)∆jg2(cid:18)1 − 2ν; 1 2 , 1(cid:19) . If 1 2 ≤ ν ≤ 1, we have (x!νy)2 + 2ν2(x2∇y2 − (x!y)2) + 2ν(1 − ν) xy −(cid:18) 2xy x + y(cid:19)2! 2(cid:19) . Aj(2 − 2ν)∆jg2(cid:18)1 − 2ν; 0, ≥ (x∇ν y)2 + 1 N Xj=1 MEANS REFINEMENTS VIA CONVEXITY 13 3.3. Some Lp inequalities. Let (X,M, µ) be a measure space, and let 0 < p < q < r. Then Lp ∩ Lr ⊂ Lq and pkfk1−ν , where f ∈ Lp ∩ Lr and ν = kfkq ≤ kfkν . r q−1 − r−1 p−1 − r−1 This inequality can be modified using Corollary 2.11 and a reversed version can be obtained using Corollary 2.12. Proposition 3.9. Let (X,M, µ) be a measure space, 0 < p < q < r and ν be as above. If f ∈ Lp ∩ Lr and N ∈ N, then we have   Proof. It is easy to check that the function h(t) = kfk1/t is log-convex on [r−1, p−1]. Then direct application of Corollary 2.11 implies the result. (ν)kfky−1 pkfk1−ν ≤ kfkν kfkx−1   Yj=1 kfk2 kfkq Aj (ν) z−1 j (ν) (ν) (cid:3) N . r j j In particular, when N = 1, the above proposition implies kfkq ≤( kfk1−2ν kfk2ν−1 2pr p+r kfk2ν kfk2−2ν 2pr p+r p r , , 2 0 ≤ ν ≤ 1 2 ≤ ν ≤ 1 ) ≤ kfkν 1 pkfk1−ν r . 2 can be interpreted as 2pr p+r ≤ q ≤ r, while 1 2 ≤ ν ≤ 1 The condition 0 ≤ ν ≤ 1 means p ≤ q ≤ 2pr p+r . Moreover, a reversed version maybe obtained using Corollary 2.12. Proposition 3.10. Let (X,M, µ) be a measure space, 0 < p < q < r and ν = q−1−r−1 p−1−r−1 . If 2pr p+r ≤ q ≤ r, f ∈ Lp ∩ Lr and N ∈ N, then Aj (1−2ν) where tj, uj and zj are obtained from xj, yj and zj by replacing (ν, a, b) with kfkq ≥ kfk2−2ν 2pr j j j p N (ν) (ν) w−1 kfk2 Yj=1   kfkt−1 (ν)kfku−1 p+r kfk2ν−1   2pr , p−1(cid:17) . On the other hand, if p ≤ q ≤ 2pr   p+rkfk1−2ν (ν)kfku−1 kfkt−1   Yj=1 kfk2 w−1 (ν) (ν) 2pr N r j j j (cid:16)1 − 2ν, p+r kfkq ≥ kfk2ν p+r , then Aj (1−2ν) ≥ kfk2−2ν p+r kfk2ν−1 2pr p , ≥ kfk2ν p+rkfk1−2ν 2pr r , where tj, uj and zj are obtained from xj, yj and zj by replacing (ν, a, b) with 2pr(cid:17) . (cid:16)2 − 2ν, r−1, p+r Propositions 3.9 and 3.10 have been obtained using log-convexity of the func- tion h(t) = kfkt−1. In fact, noting log-convexity of the function h(t) = kfkt t, we obtain the same results! This is due to the equivalence of log-convexity of the functions t 7→ kfkt−1 and that of kfkt t. We refer the reader to [9] where these relations between the different log-convex function criteria have been discussed. The celebrated three lines lemma of Hadamard states the following. 14 M. SABABHEH Lemma 3.11. Let D = {z ∈ C : 0 ≤ ℜz ≤ 1} and let ϕ : D → C be continuous on D and analytic in the interior of D. Then the function f : [0, 1] → R defined by f (x) = supy ϕ(x + iy) is log-convex. This lemma is an extremely useful tool in the theory of complex functions. In particular, this lemma becomes handy in proving different interpolation versions of bounded linear operators between Lp spaces. Log-convexity implied by Lemma 3.11 allows us to apply our refined and re- versed versions for log-convex functions. In the following proposition, we present one term refinement and reverse. 1 and Proposition 3.12. Let D = {z ∈ C : 0 ≤ ℜz ≤ 1} and let ϕ : D → C be continuous on D and analytic in the interior of D. Then the function f : [0, 1] → R defined by f (x) = supy ϕ(x + iy) satisfies the following f (x) ≤(cid:26) f 1−2x(0)f 2x(cid:0) 1 0 ≤ x ≤ 1 2(cid:1) , f 2x−1(1)f 2−2x(cid:0) 1 2(cid:1) , 2 ≤ x ≤ 1 2(cid:1) , 0 ≤ x ≤ 1 f (x) ≥(cid:26) f 2x−1(1)f 2−2x(cid:0) 1 f 1−2x(0)f 2x(cid:0) 1 2(cid:1) , 2 ≤ x ≤ 1 3.4. Operator versions. The following theorem provides a refinement of the well known Heinz inequality and its reverse. The proof follows immediately noting convexity of the Heinz means, see [1]. Theorem 3.13. For A, B ∈ M+ norm k k, let n , X ∈ Mn, 0 ≤ ν ≤ 1 and any unitarily invariant . 2 2 1 Then we have the following refinement of Heinz inequality f (ν) = kAνXB1−ν + A1−ν XBνk. kAνXB1−ν + A1−νXBνk + N Xj=1 Aj(ν)∆j f (ν; 0, 1) ≤ kAX + XBk. 2 , we have Moreover, if 0 ≤ ν ≤ 1 kAνXB1−ν + A1−ν XBνk + 2(1 − ν)(cid:16)kAX + XBk − k N ≥ kAX + XBk + On the other hand, if 1 2 ≤ ν ≤ 1, we have Xj=1 kAνXB1−ν + A1−ν XBνk + 2ν(cid:16)kAX + XBk − k Xj=1 ≥ kAX + XBk + N √AX√Bk(cid:17) Aj(1 − 2ν)∆jf(cid:18)1 − 2ν; 1 2 , 1(cid:19) . √AX√Bk(cid:17) Aj(2 − 2ν)∆jf(cid:18)2 − 2ν; 0, 1 2(cid:19) . MEANS REFINEMENTS VIA CONVEXITY 15 In [9], it is shown that for A, B ∈ M+ n and X ∈ Mn, the functions t → kAtXB1−tk, t → kAtXB1−tk kA1−tX Btk, t → tr(AtXB1−tX∗) are log-convex on [0, 1]. Therefore, we may apply Corollaries 2.11 and 2.12 to obtain refinements and reversed versions for such functions. For the k k2 norm, we can prove log convexity of the Heinz means, which allows us to obtain further refinements of the Heinz inequality by applying Corollaries 2.11 and 2.12. Proposition 3.14. Let A, B ∈ M+ A1−νXBνk2. Then f is log-convex on [0, 1]. Proof. Since A, B ∈ M+ n , there are diagonal matrices D1 := diag(λi), D2 := diag(µi) and unitarily matrices U, V such that λi, µi ≥ 0, A = U D1U∗ and B = V D2V ∗. Letting Y = U∗XV, we have Aν XB1−ν + A1−νXBν = U(λν n and X ∈ Mn, and define f (ν) = kAνXB1−ν + yijµν j + λ1−ν i j )V ∗. i yijµ1−ν Since k k2 is a unitarily invariant norm, we have j + λ1−ν j + λ1−ν i µν i yijµ1−ν i µ1−ν i f 2(ν) = kU(λν = Xi,j (cid:0)λν 2 yijµν j )V ∗k2 j(cid:1)2 yij2. Notice that each summand is log-convex, being the square of a log-convex func- tion. This implies that f 2 is log-convex. Consequently, f is log-convex. (cid:3) Letting f (ν) = kAνXB1−ν + A1−νXBνk2 and applying Theorem 2.13 imply the following squared version of Heinz inequality. Corollary 3.15. Let A, B ∈ M+ kAνXB1−ν + A1−νXBνk2 n , X ∈ Mn, 0 ≤ ν ≤ 1 and N ≥ 2. Then 2 + 2A2 2 XB 2 − 2kA 1 1(ν)(cid:16)kAX + XBk2 1 2(cid:17) 2k2 + N Xj=2 Aj(ν)∆jf 2(ν; 0, 1) ≤ kAX + XBk2 2. We leave the application of Corollaries 2.11 and 2.12 to the reader. Further operator versions maybe obtained using Lemma 1.1. The following operator versions refine the corresponding results in [5] and [11]. Proposition 3.16. Let A, B ∈ M++ n and 0 ≤ ν ≤ 1. Then for αj(ν) = kj(ν) 2j−1 , βj(ν) = kj(ν)+1 2j−1 , γj(ν) = αj(ν)+βj (ν) 2 and N ∈ N, we have A#νB + N Xj=1 Aj(ν)(cid:0)A#αj (ν) + A#βj(ν) − 2A#γj (ν)B(cid:1) ≤ A∇νB. Proof. In Corollary 3.1, let x = 1, expand the summand and apply Lemma 1.1 with y replaced by X = A− 1 2 . Then the result follows upon conjugating both sides with A (cid:3) 2 BA− 1 2 . 1 16 M. SABABHEH In a similar way one may obtain reversed versions by applying Corollary 3.5. This provides refinements of the reversed versions of [11]. The following is an operator arithmetic-harmonic version, refining the corresponding results in [12]. Proposition 3.17. Let A, B ∈ M++ n and 0 ≤ ν ≤ 1. Then for αj(ν) = kj(ν) 2j−1 , βj(ν) = kj(ν)+1 2j−1 , γj(ν) = αj(ν)+βj (ν) 2 and N ∈ N, we have A!νB + N Xj=1 Aj(ν)(cid:0)A!αj (ν) + A!βj(ν) − 2A!γj (ν)B(cid:1) ≤ A∇νB. The proof follows immediately on applying Lemma 1.1 together with Corollary 3.2. On the other hand, applying Corollary 3.6 implies the following refinement of the corresponding inequalities in [6]. Proposition 3.18. Let A, B ∈ M++ and N ∈ N. If 0 ≤ ν ≤ 1, we have n A!νB + (1 − ν)(A + B − 2A!B) N Xj=1 ≥ A∇νB + 2(cid:16)1 − kj(1−2ν) 2j−1 (cid:17) + kj(1−2ν) where αj(ν) = 1 and γj(ν) = αj (ν)+βj (ν) . On the other hand, if 1 2 ≤ ν ≤ 1, we have 2j−1 2 Aj(1 − 2ν)(cid:0)A!αj (ν)B + A!βj(ν)B − 2A!γj (ν)B(cid:1) , , βj(ν) = 1 2(cid:16)1 − kj(1−2ν)+1 2j−1 (cid:17) + kj (1−2ν)+1 2j−1 A!νB + ν(A + B − 2A!B) N ≥ A∇νB + Xj=1 Aj(2 − 2ν)(cid:0)A!αj (ν)B + A!βj(ν)B − 2A!γj (ν)B(cid:1) , where αj(ν) = kj (2−2ν) 2j−1 , βj(ν) = kj(2−2ν)+1 2j−1 and γj(ν) = αj(ν)+βj (ν) 2 . The following is an interesting one-term multiplicative refinement of the oper- ator geometric-harmonic mean inequality. Theorem 3.19. Let A, B ∈ M++ n and 0 ≤ ν ≤ 1. Then (A!νB)(cid:18) A−1B + 2I + B−1A 4 (cid:19)r ≤ A#ν B, where r = min{ν, 1 − ν}. Proof. We prove the desired inequality for 0 ≤ ν ≤ 1 2√y(cid:17)2ν and x = 1, to get (1!νy)(cid:16) 1+y ≤ 1#νy, or 2 . In Corollary 3.3, let N = 1 1 4ν (cid:0)(1 − ν) + νy−1(cid:1)−1(cid:0)y + 2 + y−1(cid:1)ν ≤ yν. (3.3) Let X = A− 1 2 BA− 1 1 1 = = A− 1 1 1 2 B−1A 2 BA− 1 2 (A!νB)A− 1 2(cid:17)−1(cid:16)A− 1 4ν (cid:16)(1 − ν)I + νA 4ν hA− 1 2 (cid:0)A−1B + 2I + B−1A(cid:1)ν (cid:19)ν 2 (A!ν B)(cid:18) A−1B + 2I + B−1A 2ihA A− 1 2 . 4 1 A− 1 2i 2 + 2I + A 1 2 B−1A 1 2(cid:17)ν (3.4) MEANS REFINEMENTS VIA CONVEXITY 17 2 and apply Lemma 1.1. The left hand side of (3.3) becomes On the other hand, the right hand side of (3.3) is simply (cid:16)A− 1 . This together with (3.4) imply the desired inequality, upon conjugating both sides with A (cid:3) 2 . This completes the proof. 2 BA− 1 1 2(cid:17)ν References [1] R. Bhatia, Matrix analysis, Springer-Verlag, New York, 1997. [2] T. Furuta, J. Micic Hot and J. Pecaric, Mond-Pecaric Method in Operator Inequalities, Element, Zagreb, 2005. [3] O. Hirzallah and F. Kittaneh, Matrix Young inequalities for the Hilbert-Schmidt norm, Linear Algebra Appl. 308 (2000), 77 -- 84. [4] F. Kittaneh, On the convexity of the Heinz mean, Integr. Equ. Oper. Theory 68 (2010), 519 -- 527. [5] F. Kittaneh and Y. Manasrah, Improved Young and Heinz inequalities for matrices, J. Math. Anal. Appl. 36 (2010), 262 -- 269. [6] W. Liao and J. Wu, Reverse arithmetic-harmonic mean and mixed mean operator inequali- ties, J. Inequal. Appl., 2015:215. [7] Y. Manasrah and F. Kittaneh, A generalization of two refined Young inequalities, Positivity 19 (2015), 757 -- 768. [8] M. Sababheh, Convex functions and means of matrices, Math. Ineq. Appl., accepted. [9] M. Sababheh, Log and harmonically log-convex functions related to matrix norms, Operators and Matrices, in press. [10] M. Sababheh, Integral inequalities of the Heinz means as convex functions, J. Math. Ineq., 10 (2) (2016), 313 -- 325. [11] J. Zhao and J. Wu, Operator inequalities involving improved Young and its reverse inequal- ities, J. Math. Anal. Appl. 421 (2015) 1779 -- 1789. [12] H. Zuo, G. Shi and M. Fujii, Refined Young inequality with Kantorovich constant, J. Math. Inequal. 5(4)(2011), 551-556 . Department of Basic Sciences, Princess Sumaya University For Technology, Al Jubaiha, Amman 11941, Jordan. E-mail address: [email protected], [email protected]
1708.06885
1
1708
2017-08-23T05:05:24
Completeness of Unbounded Convergences
[ "math.FA" ]
As a generalization of almost everywhere convergence to vector lattices, unbounded order convergence has garnered much attention. The concept of boundedly uo-complete Banach lattices was introduced by N. Gao and F. Xanthos, and has been studied in recent papers by D. Leung, V.G. Troitsky, and the aforementioned authors. We will prove that a Banach lattice is boundedly uo-complete iff it is monotonically complete. Afterwards, we study completeness-type properties of minimal topologies; minimal topologies are exactly the Hausdorff locally solid topologies in which uo-convergence implies topological convergence.
math.FA
math
COMPLETENESS OF UNBOUNDED CONVERGENCES M.A. TAYLOR Abstract. As a generalization of almost everywhere convergence to vector lattices, unbounded order convergence has garnered much attention. The concept of boundedly uo-complete Banach lattices was introduced by N. Gao and F. Xanthos, and has been studied in recent papers by D. Leung, V.G. Troitsky, and the aforemen- tioned authors. We will prove that a Banach lattice is bound- edly uo-complete iff it is monotonically complete. Afterwards, we study completeness-type properties of minimal topologies; mini- mal topologies are exactly the Hausdorff locally solid topologies in which uo-convergence implies topological convergence. 1. Introduction In the first half of the paper, we study when norm bounded uo- Cauchy nets in a Banach lattice are uo-convergent. The section starts with a counterexample to a question posed in [LC], and culminates in a proof that a Banach lattice is (sequentially) boundedly uo-complete iff it is (sequentially) monotonically complete. This gives the final solution to a problem that has been investigated in [Gao14], [GX14], [GTX17], and [GLX]. The latter half of this paper focuses on the "extremal" topologies of a vector lattice X. For motivation, recall that corresponding to a dual pair hE, E∗i is a family of topologies on E "compatible" with duality. The two most important elements of this family are the weak and Mackey topologies, which are defined by their extremal nature. Analogously, given a vector lattice X, it is often possible to equip X Date: October 3, 2018. 2010 Mathematics Subject Classification. 46A40, 46A16, 46B42. Key words and phrases. uo-convergence, unbounded topology, minimal topology, completeness, boundedly uo-complete, monotonically complete, Levi. The author acknowledges support from NSERC and the University of Alberta. 1 2 M.A. TAYLOR with many topologies compatible (in the sense of being locally solid and Hausdorff) with the lattice structure. It is easy to see that whenever X admits some Hausdorff locally solid topology, the collection of all Riesz pseudonorms on X generates a finest Hausdorff locally solid topology on X. This "greatest" topology appears in many applications. Indeed, analogous to the theory of compatible locally convex topologies on a Banach space - where the norm topology is the Mackey topology - the norm topology on a Banach lattice X is the finest topology on X compatible with the lattice structure. This is [AB03, Theorem 5.20]. On the opposite end of the spectrum, a Hausdorff locally solid topol- ogy on a vector lattice X is said to be minimal if there is no coarser Hausdorff locally solid topology on X; it is least if it is coarser than every Hausdorff locally solid topology on X. Least topologies were introduced in [AB80] and studied in [AB03]; minimal topologies were studied in [Lab87], [Con05], [Tay], and [KT]. An important example of a least topology is the unbounded norm topology on an order con- tinuous Banach lattice. The unbounded absolute weak∗-topology on L∞[0, 1] is a noteworthy example of a minimal topology that is not least. In the next subsection, we briefly recall some facts about mini- mal and unbounded topologies; for a detailed exposition the reader is referred to [Tay] and [KT]. 1.1. Notation. Throughout this paper, all vector lattices are assumed o−→ x if Archimedean. For a net (xα) in a vector lattice X, we write xα (xα) converges to x in order ; that is, there is a net (yβ), possibly over a different index set, such that yβ ↓ 0 and for every β there exists uo−→ x α0 such that xα − x ≤ yβ whenever α ≥ α0. We write xα o−→ 0 for and say that (xα) uo-converges to x ∈ X if xα − x ∧ u every u ∈ X+. For facts on uo-convergence, the reader is referred to [GTX17]. In particular, [GTX17, Theorem 3.2] will be used freely. Recall that a Banach lattice X is (sequentially) boundedly uo- complete if norm bounded uo-Cauchy nets (respectively, sequences) in X are uo-convergent in X. Given a locally solid topology τ on a vector lattice X, one can asso- ciate a topology, uτ , in the following way. If {Ui}i∈I is a base at zero UNBOUNDED CONVERGENCES 3 for τ consisting of solid sets, for each i ∈ I and u ∈ X+ define Ui,u := {x ∈ X : x ∧ u ∈ Ui}. As was proven in [Tay, Theorem 2.3], the collection N0 = {Ui,u : i ∈ I, u ∈ X+} is a base of neighbourhoods at zero for a new lo- cally solid topology, denoted by uτ , and referred to as the unbounded τ -topology . Noting that the map τ 7→ uτ from the set of locally solid topologies on X to itself is idempotent, a locally solid topology τ is called unbounded if there is a locally solid topology σ with τ = uσ or, equivalently, if τ = uτ. The following connection between minimal topologies, unbounded topologies, and uo-convergence was proven in [Tay, Theorem 6.4]. Recall that a locally solid topology τ is Lebesgue if order null nets are τ -null. Theorem 1.1. Let τ be a Hausdorff locally solid topology on a vector lattice X. TFAE: (i) uo-null nets are τ -null; (ii) τ is Lebesgue and unbounded; (iii) τ is minimal. In particular, a vector lattice can admit at most one minimal topology. Interestingly, the process of unbounding a topology can convert the greatest topology into the least topology; this happens with the norm topology on an order continuous Banach lattice. All other undefined terminology is consistent with [AB03]. In par- ticular, we say that a locally solid topology τ on a vector lattice X is Levi if τ -bounded increasing nets in X+ have supremum. Levi and monotonically complete are synonymous; the latter terminology is that of [MN91], and is used in [GLX]. 2. Boundedly uo-complete Banach lattices Results equating the class of boundedly uo-complete Banach lattices to the class of monotonically complete Banach lattices have been ac- quired, under technical assumptions, by N. Gao, D. Leung, V.G. Troit- sky, and F. Xanthos. The sharpest result is [GLX, Proposition 3.1]; it states that a Banach lattice whose order continuous dual separates 4 M.A. TAYLOR points is boundedly uo-complete iff it is monotonically complete. In this section, we remove the restriction on the order continuous dual. The following question was posed as Problem 2.4 in [LC]: Question 2.1. Let (xα) be a norm bounded positive increasing net in a Banach lattice X. Is (xα) uo-Cauchy in X? If Question 2.1 is true, it is easily deduced that a Banach lattice is boundedly uo-complete iff it is monotonically complete. However, the next example answer this question in the negative, even for sequences. Example 2.2. Let S be the set of all non-empty finite sequences of natural numbers. For s ∈ S define λ(s) = length(s). If s, t ∈ S, define s ≤ t if λ(s) ≤ λ(t) and s(i) = t(i) for i = 1, . . . , λ(s). For s ∈ S with λ(s) = n and i ∈ N, define s ∗ i = (s(1), . . . , s(n), i). Put (2.1) X = {x ∈ ℓ∞(S) : lim i→∞ x(s ∗ i) = 1 2 x(s) for all s ∈ S}. It can be verified that X is a closed sublattice of (ℓ∞(S), k · k∞) and for t ∈ S the element et : S → R defined by et(s) =   ( 1 2)λ(s)−λ(t) 0 if t ≤ s otherwise is an element of X with norm 1. Define f1 = e(1), f2 = e(1) ∨ e(2) ∨ e(1,1) ∨ e(1,2) ∨ e(2,1) ∨ e(2,2), and, generally, fn = sup{et : λ(t) ≤ n and t(k) ≤ n ∀k ≤ λ(t)}. The sequence (fn) is increasing and norm bounded by 1; it was shown in [BL88, Example 1.8] that (fn) is not order bounded in X u. There- fore, (fn) cannot be uo-Cauchy in X for if it were then it would be uo-Cauchy in X u and hence order convergent in X u by [GTX17, The- orem 3.10]. Since it is increasing, it would have supremum in X u; this is a contradiction as (fn) is not order bounded in X u. Under some mild assumptions, however, Question 2.1 has a positive solution. Recall that a Banach lattice is weakly Fatou if there exists K ≥ 1 such that whenever 0 ≤ xα ↑ x, we have kxk ≤ K sup kxαk. UNBOUNDED CONVERGENCES 5 Proposition 2.3. Let X be a weakly Fatou Banach lattice. Then every positive increasing norm bounded net in X is uo-Cauchy. If 0 ≤ ( 1 Proof. Let K be such that 0 ≤ xα ↑ x implies kxk ≤ K sup kxαk. Now assume that 0 ≤ uα ↑ and kuαk ≤ 1. Let u > 0 and pick n such that kuk > K n . Therefore, there n . exists 0 < w ∈ X such that ( 1 n uα) ∧ u ≤ u − w for all α. But then n uα) ∧ u ↑α u, then kuk ≤ K n uα) ∧ u(cid:3) ≥ nw > 0 for all α, so that (uα) is (nu − uα)+ = n(cid:2)u − ( 1 and hence uα ↑ bu for some bu ∈ X u. This proves that (uα) is uo-Cauchy dominable. By [AB03, Theorem 7.37], (uα) is order bounded in X u, in X u, hence in X. (cid:3) Proposition 2.4. Let X be a weakly σ-Fatou Banach lattice. Then every positive increasing norm bounded sequence in X is uo-Cauchy. Proof. The proof is similar and, therefore, omitted. (cid:3) Even though Question 2.1 is false, the equivalence between bound- edly uo-complete and Levi still stands. We show that now: Theorem 2.5. Let X be a Banach lattice. TFAE: (i) X is σ-Levi; (ii) X is sequentially boundedly uo-complete; (iii) Every increasing norm bounded uo-Cauchy sequence in X+ has a supremum. xn n=1 1 2n parts. Define e = P∞ Proof. (i)⇒(ii): Let (xn) be a norm bounded uo-Cauchy sequence in X. WLOG, (xn) is positive; otherwise consider positive and negative 1+kxnk and consider Be, the band generated by e. Then (xn) is still norm bounded and uo-Cauchy in Be. Also, Be has the σ-Levi property for if 0 ≤ yn ↑ is a norm bounded sequence in Be, then yn ↑ y for some y ∈ X as X is σ-Levi. Since Be is a band, y ∈ Be and yn ↑ y in Be. We next show that there exists u ∈ Be such that xn uo−→ u in Be, and hence in X. For each m, n, n′ ∈ N, since xn ∧ me − xn′ ∧ me ≤ xn − xn′ ∧ me, the sequence (xn ∧me)n is order Cauchy, hence order converges to some um in Be since the σ-Levi property implies σ-order completeness. The 6 M.A. TAYLOR sequence (um) is increasing and kumk ≤ K sup n kxn ∧ mek ≤ K sup n kxnk < ∞ where we use that σ-Levi implies weakly σ-Fatou. Since Be is σ-Levi, (um) increases to an element u ∈ Be. Fix m. For any N, N ′ define xN,N ′ = supn≥N,n′≥N ′ xn − xn′ ∧ e. Since (xn) is uo-Cauchy, xN,N ′ ↓ 0. Now, for each m, xn ∧ me − xn′ ∧ me ∧ e ≤ xn − xn′ ∧ e ≤ xN,N ′ ∀n ≥ N, n′ ≥ N ′. Taking order limit in n′ yields xn ∧ me − um ∧ e ≤ xN,N ′ Taking order limit in m now yields: xn − u ∧ e ≤ xN,N ′, ∀n ≥ N, from which it follows that xn − u ∧ e in Be since e is a weak unit of Be. o−→ 0 in Be. This yields xn uo−→ u The implication (ii)⇒(iii) is clear. For the last implication it suf- fices, by [AW97, Theorem 2.4], to verify that every norm bounded laterally increasing sequence in X+ has a supremum. Let (xn) be a norm bounded laterally increasing sequence in X+. By [AW97, Propo- sition 2.2], (xn) has supremum in X u, hence is uo-Cauchy in X u. It follows that (xn) is uo-Cauchy in X and, therefore, by assumption, uo-converges to some x ∈ X. It is then clear that xn ↑ x in X. (cid:3) Theorem 2.6. Let X be a Banach lattice. TFAE: (i) X is Levi; (ii) X is boundedly uo-complete; (iii) Every increasing norm bounded uo-Cauchy net in X+ has a supremum. Proof. If X is Levi, then X is boundedly uo-complete by [GLX, Propo- sition 3.1]. It is clear that (ii)⇒(iii) and the proof of (iii)⇒(i) is the same as in the last theorem but with [AW97, Theorem 2.4] replaced with [AW97, Theorem 2.3]. (cid:3) UNBOUNDED CONVERGENCES 7 3. Completeness of minimal topologies Throughout this section, X is a vector lattice and τ denotes a locally solid topology on X. We begin with a brief discussion on relations between minimal topologies and the B-property. Proposition 3.3 will be of importance as many properties of locally solid topologies are stated in terms of positive increasing nets. For minimal topologies, these properties permit a uniform and efficient treatment. The B-property was introduced as property (B,iii) by W.A.J. Lux- emburg and A.C. Zaanen in [LZ64]. It is briefly studied in [AB03] and, in particular, it is shown that the Lebesgue property does not imply the B-property. We prove, however, that if τ is unbounded then this implication does indeed hold true: Definition 3.1. A locally solid vector lattice (X, τ ) satisfies the B- property if it follows from 0 ≤ xn ↑ in X and (xn) τ -bounded that (xn) is τ -Cauchy. An equivalent definition is obtained if sequences are replaced with nets. Proposition 3.2. If X is a vector lattice admitting a minimal topology τ , then τ satisfies the B-property. Proof. Suppose τ is minimal and (xn) is a τ -bounded sequence satisfy- ing 0 ≤ xn ↑. By [AB03, Theorem 7.50], (xn) is dominable. By [AB03, uo−→ u for some Theorem 7.37], (xn) is order bounded in X u so that xn u ∈ X u. In particular, (xn) is uo-Cauchy in X u. It follows that (xn) is uo-Cauchy in X. Since τ is Lebesgue, (xn) is uτ -Cauchy in X. Finally, since τ is unbounded, (xn) is τ -Cauchy in X. (cid:3) Proposition 3.3. Let X be a vector lattice admitting a minimal topol- ogy τ , and (xα) an increasing net in X+. TFAE: (i) (xα) is τ -bounded; (ii) (xα) is τ -Cauchy. Proof. It remains to prove (ii)⇒(i): Let (xα) be an increasing τ -Cauchy net in X+. By [Tay, Corollary 5.7], τ extends to a complete Hausdorff τ u −→ x for some x ∈ X u. Lebesgue topology τ u on X u. It follows that xα Since (xα) is increasing, xα ↑ x in X u. In particular, (xα) is order 8 M.A. TAYLOR bounded in X u, hence dominable in X by [AB03, Theorem 7.37]. By [Tay, Theorem 5.2], (xα) is τ -bounded in X. (cid:3) Recall the following definition, taken from [AB03, Definition 2.43]. Definition 3.4. A locally solid vector lattice (X, τ ) is said to satisfy the monotone completeness property (MCP) if every increasing τ -Cauchy net of X+ is τ -convergent in X. The σ-MCP is defined analogously with nets replaced with sequences. Remark 3.5. By Proposition 3.3, a minimal topology has MCP iff it is Levi. Proposition 3.6. Let τ be a Hausdorff locally solid topology on X. If uτ satisfies MCP then so does τ . If uτ satisfies σ-MCP then so does τ . Proof. Suppose 0 ≤ xα ↑ is a τ -Cauchy net. It is then uτ -Cauchy and τ−→ x. hence uτ -converges to some x ∈ X. Therefore, xα ↑ x and xα Replacing nets with sequences yields the σ-analogue. (cid:3) Recall by [AB03, Theorem 2.46 and Exercise 2.11] that a Hausdorff locally solid vector lattice (X, τ ) is (sequentially) complete iff order intervals are (sequentially) complete and τ has (σ)-MCP. Therefore, since τ -convergence agrees with uτ -convergence on order intervals, uτ being (sequentially) complete implies τ is (sequentially) complete. Lemma 3.7. Let (X, τ ) be a Hausdorff locally solid vector lattice. If τ is unbounded then TFAE: (i) τ has MCP and is pre-Lebesgue; (ii) τ is Lebesgue and Levi. Proof. It is sufficient, by [DL98, Theorem 2.5], to prove that (X, τ ) contains no lattice copy of c0. Suppose, towards contradiction, that X does contain a lattice copy of c0, i.e., there is a homeomorphic Riesz isomorphism from c0 onto a sublattice of X. This leads to a contradic- tion as the standard unit vector basis is not null in c0, but the copy in X is by [Tay, Theorem 4.2]. (cid:3) UNBOUNDED CONVERGENCES 9 Lemma 3.7 is another way to prove that a minimal topology has MCP iff it is Levi. We next present the sequential analogue: Lemma 3.8. Let (X, τ ) be a Hausdorff locally solid vector lattice. If τ is unbounded then TFAE: (i) τ has σ-MCP and is pre-Lebesgue; (ii) τ is σ-Lebesgue and σ-Levi. Proof. (i)⇒(ii) is similar to the last lemma; apply instead [DL98, Propo- sition 2.1 and Theorem 2.4]. (ii)⇒(i): It suffices to show that τ is pre-Lebesgue. For this, suppose that 0 ≤ xn ↑≤ u; we must show that (xn) is τ -Cauchy. Since τ is σ- Levi and order bounded sets are τ -bounded, xn ↑ x for some x ∈ X. Since τ is σ-Lebesgue, xn (cid:3) τ−→ x. Putting pieces together from other papers, we next characterize se- quential completeness of uo-convergence. Theorem 3.9. Let X be a vector lattice. TFAE: (i) X is sequentially uo-complete; (ii) Every positive increasing uo-Cauchy sequence in X uo-converges in X; (iii) X is universally σ-complete. In this case, uo-Cauchy sequences are order convergent. Proof. (i)⇒(ii) is clear. (ii)⇒(iii) by careful inspection of [LC, Propo- sition 2.8], (iii)⇒(i) and the moreover clause follow from [GTX17, The- orem 3.10]. (cid:3) Remark 3.10. Recall that by [AB03, Theorem 7.49], every locally solid topology on a universally σ-complete vector lattice satisfies the pre-Lebesgue property. Using uo-convergence, we give a quick proof of this. Suppose τ is a locally solid topology on a universally σ-complete vector lattice X; we claim that uo-null sequences are τ -null. This fol- lows since τ is σ-Lebesgue and uo and o-convergence agree for sequences by [GTX17, Theorem 3.9]. In particular, since disjoint sequences are uo-null, disjoint sequences are τ -null. 10 M.A. TAYLOR We next give the topological analogue of Theorem 3.9: Lemma 3.11. Let X be a vector lattice admitting a minimal topology τ . TFAE: (i) τ is σ-Levi; (ii) τ has σ-MCP; (iii) X is universally σ-complete. (iv) (X, τ ) is sequentially boundedly uo-complete in the sense that τ -bounded uo-Cauchy sequences in X are uo-convergent in X. in X, and define xn = Pn Proof. (i)⇔(ii) follows from Lemma 3.8. We next deduce (iii). Since τ is σ-Levi, X is σ-order complete; we prove X is laterally σ-complete. Let {an} be a countable collection of mutually disjoint positive vectors k=1 ak. Then (xn) is a positive increasing sequence in X, and it is uo-Cauchy, as an argument similar to [LC, Proposition 2.8] easily shows. By Theorem 1.1, (xn) is τ -Cauchy, hence τ−→ x for some x ∈ X since τ has σ-MCP. Since (xn) is increasing xn and τ is Hausdorff, xn ↑ x. Clearly, x = sup{an}. (iii)⇒(iv) follows from Theorem 3.9; (iv)⇒(i) is easy. (cid:3) The following question(s) remain open: Question 3.12. Let X be a vector lattice admitting a minimal topol- ogy τ . Are the following equivalent? (i) (X, τ ) is sequentially complete; (ii) X is universally σ-complete. Question 3.13. Let (X, τ ) be Hausdorff and Lebesgue. Are the fol- lowing equivalent? (i) Order intervals of X are sequentially τ -complete; (ii) X is σ-order complete. Remark 3.14. Question 3.12 and Question 3.13 are equivalent. In- deed, in both cases it is known that (i)⇒(ii). If Question 3.13 is true then Question 3.12 is true since we have already established that min- imal topologies have σ-MCP when X is universally σ-complete. Sup- pose Question 3.12 is true. If X is σ-order complete, then X is an ideal in its universal σ-completion, X s. Indeed, it is easy to establish that UNBOUNDED CONVERGENCES 11 if Y is a σ-order complete vector lattice sitting as a super order dense sublattice of a vector lattice Z, then Y is an ideal of Z; simply modify the arguments in [AB03, Theorem 1.40]. By [AB03, Theorem 4.22] we may assume, WLOG, that τ is minimal. τ then lifts to the universal completion and can be restricted to X s. Question 3.13 is a special case of Aliprantis and Burkinshaw's [AB78, Open Problem 4.2], which we state as well: Question 3.15. Suppose τ is a Hausdorff σ-Fatou topology on a σ- order complete vector lattice X. Are the order intervals of X sequen- tially τ -complete? The case of complete order intervals is much easier than the sequen- tially complete case. The next result is undoubtedly known, but fits in nicely; we provide a simple proof that utilizes minimal topologies. Proposition 3.16. Suppose τ is a Hausdorff Lebesgue topology on X. Order intervals of X are complete iff X is order complete. Proof. If X is order complete then order intervals are complete by [AB03, Theorem 4.28]. By [AB03, Theorem 4.22] we may assume, WLOG, that τ is minimal. If order intervals are complete then X is an ideal of bX = X u by [AB03, Theorem 2.42] and [Tay, Theorem 5.2]. Since X u is order complete, so is X. (cid:3) Remark 3.17. If X is an order complete and laterally σ-complete vector lattice admitting a minimal topology τ , then τ is sequentially complete. Although these conditions are strong, they do not force X to be universally complete. This can be seen by equipping the vector lattice of [AB03, Example 7.41] with the minimal topology given by restriction of pointwise convergence from the universal completion. The key step in the proof of Theorem 2.5 is [AW97, Theorem 2.4] which states that a Banach lattice is σ-Levi if and only if it is laterally σ-Levi. A sequence (xn) in a vector lattice is said to be laterally increasing if it is increasing and (xm − xn) ∧ xn = 0 for all m ≥ n. 12 M.A. TAYLOR We say that a locally solid vector lattice (X, τ ) has the lateral σ- Levi property if sup xn exists whenever (xn) is laterally increasing and τ -bounded. For minimal topologies, the σ-Levi and lateral σ-Levi properties do not agree, as we now show: Proposition 3.18. Let X be a vector lattice admitting a minimal topol- ogy τ . TFAE: (i) X is laterally σ-complete; (ii) τ has the lateral σ-Levi property; (iii) Every disjoint positive sequence, for which the set of all possible finite sums is τ -bounded, must have a supremum. Proof. (i)⇒(iii) is clear, as is (ii)⇔(iii); we prove (ii)⇒(i). Assume (ii) and let (xn) be a disjoint sequence in X+. Since (xn) is disjoint, (xn) has a supremum in X u. Define yn = x1 ∨ · · · ∨ xn. The sequence (yn) is laterally increasing and order bounded in X u. By [AB03, Theorem 7.37], (yn) forms a dominable set in X+. By [Tay, Theorem 5.2(iv)], (yn) is τ -bounded, and hence has supremum in X by assumption. This implies that (xn) has a supremum in X and, therefore, X is laterally σ-complete. (cid:3) In [Lab84] and [Lab85], many completeness-type properties of locally solid topologies were introduced. For entirety, we classify the remaining properties, which he refers to as "BOB" and "POB". Definition 3.19. A Hausdorff locally solid vector lattice (X, τ ) is said to be boundedly order-bounded (BOB) if increasing τ -bounded nets in X+ are order bounded in X. (X, τ ) satisfies the pseudo-order boundedness property (POB) if increasing τ -Cauchy nets in X+ are order bounded in X. Remark 3.20. It is clear that a Hausdorff locally solid vector lattice is Levi iff it is order complete and boundedly order-bounded. It is also clear that BOB and POB coincide for minimal topologies. Proposition 3.21. Let X be a vector lattice admitting a minimal topol- ogy τ . TFAE: (i) (X, τ ) satisfies BOB; UNBOUNDED CONVERGENCES 13 (ii) X is majorizing in X u. Proof. (i)⇒(ii): Let 0 ≤ u ∈ X u. Since X is order dense in X u, there exists a net (xα) in X such that 0 ≤ xα ↑ u. In particular, (xα) is order bounded in X u, hence dominable in X by [AB03, Theorem 7.37]. By [Tay, Theorem 5.2], (xα) is τ -bounded. By assumption, (xα) is order bounded in X, hence, (xα) ⊆ [0, x] for some x ∈ X+. It follows that u ≤ x, so that X majorizes X u. (ii)⇒(i): Suppose (xα) is an increasing τ -bounded net in X+. It follows from [AB03, Theorem 7.50] that (xα) is dominable, hence order bounded in X u. Since X majorizes X u, (xα) is order bounded in X. (cid:3) Remark 3.22. By [AB03, Theorem 7.15], laterally complete vector lattices majorize their universal completions. Remark 3.23. If τ is a Hausdorff Fatou topology on X, it is easy to see that (X, τ ) satisfies BOB iff every increasing τ -bounded net in X+ is order Cauchy in X. Compare with Question 2.1. We next state the σ-analogue of Proposition 3.21. Proposition 3.24. Let X be an almost σ-order complete vector lattice admitting a minimal topology τ . TFAE: (i) (X, τ ) satisfies σ-BOB; (ii) X is majorizing in the universal σ-completion X s of X. Proof. (i)⇒(ii) is similar to Proposition 3.21. (ii)⇒(i): Suppose (xn) is an increasing τ -bounded sequence in X+. It is then dominable in X, hence in X s by [AB03, Lemma 7.11]. It follows by [AB03, Theorem 7.38] that (xn) is order bounded in X s. Since X is majorizing in X s, (xn) is order bounded in X. (cid:3) The next definition is standard in the theory of topological vector spaces: Definition 3.25. Let (E, σ) be a Hausdorff topological vector space. E is quasi-complete if every σ-bounded σ-Cauchy net is σ-convergent. Remark 3.26. Since Cauchy sequences are bounded, there is no se- quential analogue of quasi-completeness. 14 M.A. TAYLOR We finish with the full characterization of completeness of minimal topologies: Theorem 3.27. Let X be a vector lattice admitting a minimal topology τ . TFAE: (i) X is universally complete; (ii) τ is complete; (iii) τ satisfies MCP; (iv) τ is Levi; (v) τ is quasi-complete; (vi) (X, τ ) is boundedly uo-complete in the sense that τ -bounded uo-Cauchy nets in X are uo-convergent in X. Proof. (i)⇔(ii) by [Tay, Corollary 5.3] combined with [Tay, Theorem 6.4]. Clearly, (ii)⇒(iii)⇔(iv). (iii)⇒(ii) since if τ satisfies MCP then τ is topologically complete by [AB03, Corollary 4.39]. We have thus established that (i)⇔(ii)⇔(iii)⇔(iv). It is clear that (ii)⇒(v), and (v)⇒(iii) by Proposition 3.3. (ii)⇒(vi): Let (xα) be a uo-Cauchy net in X; (xα) is then τ -Cauchy and hence τ -convergent. The claim then follows from [Tay, Remark 2.26]. (vi)⇒(iv): Suppose 0 ≤ xα ↑ is τ -bounded. (xα) is then uo-Cauchy, (cid:3) hence uo-convergent to some x ∈ X. Clearly, x = sup xα. Remark 3.28. This is in good agreement with Proposition 3.6. If the minimal topology satisfies MCP then Proposition 3.6 states that every Hausdorff Lebesgue topology satisfies MCP. Universally complete spaces, however, admit at most one Hausdorff Lebesgue topology by [AB03, Theorem 7.53]. References [AW97] [AB80] [AB78] Y.D. Abramovich and A.W. Wickstead, When each continuous operator is regular, II, Indag. Math., 1997. C.D. Aliprantis and O. Burkinshaw, Minimal topologies and L Illinois J. Math. 24 (1980), 164 -- 172. C.D. Aliprantis and O. Burkinshaw, Locally Solid Riesz Spaces, Aca- demic Press, New York, 1978. p-spaces, UNBOUNDED CONVERGENCES 15 [AB03] [BL88] [Con05] [DL98] C.D. Aliprantis and O. Burkinshaw, Locally Solid Riesz Spaces with Applications to Economics, 2nd ed., AMS, Providence, RI, 2003. G. Buskes and I. Labuda, On Levi-like properties and some of their applications in Riesz space theory, Canad. Math. Bull. Vol. 31 (4), 1988. J.J. Conradie, The Coarsest Hausdorff Lebesgue Topology, Quaestiones Math. 28, 287-304, 2005. L. Drewnowski and I. Labuda, Copies of c0 and ℓ∞ in topological Riesz spaces, Trans. Amer. Math. Soc. 350 , p. 3555-3570, 1998. [Frem74] D. Fremlin, Topological Riesz spaces and measure theory, Cambridge [Gao14] [GLX] University Press, 1974. N. Gao, Unbounded order convergence in dual spaces, J. Math. Anal. Appl., 419(1) (2014), 347-354. N. Gao, D. Leung and F. Xanthos, Duality for unbounded order con- vergence and applications, preprint. arXiv:1705.06143 [math.FA]. [KT] [Lab85] [Lab87] [LC] [GX14] [Lab84] [GTX17] N. Gao, V.G. Troitsky and F. Xanthos, Uo-convergence and its applica- tions to Ces`aro Means in Banach Lattices, Israel J. Math., 220 (2017), 649 -- 689. N. Gao and F. Xanthos, Unbounded order convergence and application to martingales without probability, J. Math. Anal. Appl., 415 (2014), 931-947. M. Kandi´c and M.A. Taylor, Metrizability of minimal and unbounded topologies. I. Labuda, Completeness type properties of locally solid Riesz spaces, Studia Math. 77 (1984), 349-372. I. Labuda, On boundedly order-complete locally solid Riesz spaces, Stu- dia Math. 81 (1985), 245-258. I. Labuda, Submeasures and locally solid topologies on Riesz spaces, Math. Z. 195 (1987), 179-196. H. Li and Z. Chen, Some Loose Ends on Unbounded Order Conver- gence, Positivity, to appear. arXiv:1609.09707v2 [math.FA] P. Meyer-Nieberg, Banach lattices, Springer-Verlag, Berlin, 1991. M.A. Taylor, Unbounded topologies and uo-convergence in locally solid vector lattices, preprint. arXiv:1706.01575 [math.FA]. W.A.J. Luxemburg and A.C. Zaanen, Notes on Banach function space, Nederl. Akad. Wetensch, Proc. Ser. A, Note XI 67, 1964. [MN91] [Tay] [LZ64] Department of Mathematical and Statistical Sciences, University of Alberta, Edmonton, AB, T6G 2G1, Canada. E-mail address: [email protected]
1303.6944
1
1303
2013-03-27T19:56:27
Sharp extensions for convoluted solutions of abstract Cauchy problems
[ "math.FA" ]
In this paper we give sharp extension results for convoluted solutions of abstract Cauchy problems in Banach spaces. The main technique is the use of algebraic structure (for usual convolution product $\ast$) of these solutions which are defined by a version of the Duhamel formula. We define algebra homomorphisms from a new class of test-functions and apply our results to concrete operators. Finally, we introduce the notion of $k$-distribution semigroups to extend previous concepts of distribution semigroups.
math.FA
math
(1.1) (1.2) SHARP EXTENSIONS FOR CONVOLUTED SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA Abstract. In this paper we give sharp extension results for convoluted solutions of abstract Cauchy problems in Banach spaces. The main technique is the use of alge- braic structure (for usual convolution product ∗) of these solutions which are defined by a version of the Duhamel formula. We define algebra homomorphisms from a new class of test-functions and apply our results to concrete operators. Finally, we intro- duce the notion of k-distribution semigroups to extend previous concepts of distribution semigroups. Let A be a closed linear operator on a Banach space X and τ > 0. It is well known (see 1. introduction [6, Theorem 3.1], [3] or [2, Theorem 2.1]) that if for every x ∈ X, the Cauchy problem (u′(t) = Au(t) + x, 0 ≤ t < τ u(0) = 0 has a unique solution u ∈ C 1([0, τ ), X) ∩ C([0, τ ), D(A)) (where D(A) is endowed with the graph norm), then A is the generator of a strongly continuous semigroup. This means that the solutions, initially obtained on [0, τ ), admit extensions to [0, ∞) without loss of regularity. Moreover, these solutions are (uniformly) exponentially bounded in the sense that there exist M > 0, ω ∈ R independent of u(.) such that all solutions u(.) satisfy ku(t)k ≤ M eωt, t ≥ 0. Also note that in this case, A is necessarily densely defined. In 1833, J.M.C. Duhamel considered the following evolution problem corresponding to the initial-boundary value problem for the heat equation in a domain Ω (Ω is an open subset of Rn): = ∆u, ∂u ∂t u(0, x) = u0(x), u(t,·)∂Ω = g(t,·), (t, x) ∈ R+ × Ω x ∈ Ω; t > 0.  2010 Mathematics Subject Classification. Primary 47D62, 47D06; Secondary 44A10, 44A35 . Key words and phrases. Abstract Cauchy problem; convoluted semigroups; Laplace transform, distri- bution semigroups. P.J. Miana and L. S´anchez-Lajusticia have been partly supported by supported by Project MTM2010- 16679, DGI-FEDER, of the MCYTS; Project E-64, D.G. Arag´on, and JIUZ-2012-CIE-12, Universidad de Zaragoza, Spain. This paper was written during several visits of P. J. Miana to the Department of Mathematics, University of Puerto Rico, R´ıo Piedras Camppus. He is grateful to the Department for its hospitality. 1 2 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA and proposed the following formula to express the solution of (1.2). u(t, x) =Z t 0 ∂ ∂t u(λ, t − λ, x)dλ, t > 0, where u(λ, t, x) is a solution of (1.2) for a particular function g(·, λ0) fixed λ0 ([11]). This formula allows one to reduce the Cauchy problem for an in-homogeneous partial differential equation to the Cauchy problem for the corresponding homogeneous equation. This formula (known also as Duhamel's principle) is of widespread use in partial differential equations and has been studied in a large number of papers, see for example [12, 34]. In the [34], the author extends the Duhamel principle to fractional order equations. Let k : (0, T ) → C be a locally integrable function, X a Banach space and x ∈ X. The Cauchy problem (1.3) (v′(t) = Av(t) + K(t)x v(0) = 0, 0 < t < T, 0 k(s)xds for 0 < t < T . is called K-convoluted Cauchy problem where K(t) = R t If there exists a solution of the abstract initial value problem u′(t) = Au(t) for 0 < t < T, u(0) = x then, as usual for a nonhomogeneous equation, we have v = u ∗ K (∗ is the usual convolution in R+). Local k-convoluted semigroups are defined using a version of Duhamel's formula and were introduced in [8, 9]. This class of semigroups includes C0- semigroups and integrated semigroups as particular examples, see a complete treatment in [27, Section 1.3.1], [19, Chapter 2] and other details in [18, 20]. The concept of regularized semigroups is covered by taking k(t) ≡ C, 0 ≤ t < τ, where C is a bounded and injective operator on X. Contrary to what happens in the case of equation (4.12), if k : (0, ∞) −→ C is lo- cally integrable and for every x ∈ X there exists a unique solution u ∈ C 1([0, τ ), X) ∩ C([0, τ ), D(A)) for (1.3), it is generally not the case that these solutions can be extended to [0, ∞), nor that exponential boundedness is achieved in case one can extend the solutions. In this case, we say that A is the generator of a local k−convoluted semigroup. However, there is an underlying algebraic structure of k-convoluted semigroups which leads to the following the extension property: the solution of k-convoluted problem on [0, T ) is used to express the solution of the k ∗ k-convoluted problem on [0, 2T ), see [9, Section 2] and [19, Theorem 2.1.1.9]. Stated otherwise, when (1.3) is well posed on [0, τ ), the equation in which we replace k(.) with (k ∗ k)(.) is well posed on [0, 2τ ). Our result (Theorem 4.4) provide a sharpening of this extension property. Γ(α) with α > 0 defines the α-times integrated semigroup. Originally they were the first example of convoluted semigroups. An extension formula for n-times integrated semigroups (for n ∈ N) was given in [2, Section IV, (4.2)] and for α-times integrated semigroup in [29, Formula (5)] with α > 0. Extensions of local α-times integrated C-semigroups were given in [24, Theorem 6.1] and automatic extension of local regularized semigroups appears in [36, Section 2]. The special case of k(t) = tα−1 The main objective of this paper is to illustrate the algebraic structure of local k- convoluted semigroups. In [18, Section 5], authors consider global exponentially bounded convoluted semigroups and algebra homomorphisms defined via these classes of semi- groups; in fact both concepts are equivalent, see [18, Theorem 5.7]. SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 3 In the context of local convoluted semigroups as well as global non-exponentially bounded convoluted semigroups, this point of view is not so evident. This is due to the fact that the Laplace transform is an essential tool in the global exponential case. First we need some technical identities which involve convolution products in Section 2. In Section 3, we introduce a new test function space, Dk∗∞(in Definition 3.4) which will play a fundamental role in this paper, see Theorem 5.1 and Appendix. In the fourth section, we give one of the main results of this paper. We derive a sharp extension theorem for local convoluted semigroups (Theorem 4.4). In Section 5, we use the extension formula to define algebra homomorphism from the test function space Dk∗∞ via local convoluted semigroup (The- orem 5.1). In Section 6, we apply our results to four concrete operators which generate (local and global) convoluted semigroups. In the global case and under the exponential boundedness assumption, the Laplace transform is used as a crucial tool. This is no longer the case when one consider the local case or the global one without the assumption of ex- ponential boundedness. In the case we consider in this paper, algebras concerned are no longer Banach algebras but only locally convex algebras. Historically distribution semigroups were introduced by J.L. Lions in [25] in the early sixties in the exponential case with the Laplace transform of distributions as an important tool. The paper [7] by J. Chazarain presents an extension to the non exponential case and goes further to introduce the ultradistributional framework (see also the monograph [26]). This class of vector-valued distribution (with a suitable algebraic structure) gives a equivalent approach to local integrated semigroups as was proven in [2, Theorem 7.2]. For local convoluted semigroups, we present a similar approach in Appendix where we intro- duce k-distributions semigroups and we present their connections with local convoluted semigroups. The interest in the local case stems from the fact that for the general classes of generalized semigroups that have been introduced following Lions' paper, by using the local approach, one is able to obtain a Banach space valued formulation that captures almost all the situations involved. The monographs [19], [27] and the references cited therein contain more information on distribution as well as ultradistribution semigroups. They also explore the ways in which they relate to local convoluted semigroups. A similar and independent approach may be followed in the abstract Cauchy problem of second order or wave problem. In this case we need to consider local convoluted co- sine functions and distribution cosine function (mainly algebra homomorphism for cosine convolution product) see more details in [30]. 2. Some identities for convolution products Let L1 loc(R+) the space of complex valued locally integrable functions on R+ and we consider the usual convolution product ∗, given by (f ∗ g)(t) =Z t 0 f (t − s)g(s)ds, t ≥ 0, loc(R+). We write k∗2 instead of k ∗ k and then k∗n = k ∗(cid:0)k∗(n−1)(cid:1) for where f, g ∈ L1 n > 2 is the n−fold convolution power of k. The convolution product is associative and commutative. We also follow the notation ◦ to denote the dual convolution product of ∗ given by (f ◦ g)(t) =Z ∞ t f (s − t)g(s)ds, t ≥ 0, 4 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA where f, g ∈ L1(R+). Note that max{t t ∈ supp(f ◦ g)} ≤ max{t t ∈ supp(g)}, f, g ∈ L1(R+). corresponds to the Heaviside function. Moreover, for α > 0, we set jα(t) = We denote by χ the constant function equal to 1, i.e., χ(t) = 1 for t ∈ R+. This , t ≥ 0. It will be convenient to set j0 = δ0, the Dirac measure concentrated at the origin. Observe that the following semigroup property holds: jα ∗ jβ = jα+β, α, β ≥ 0. The following lemma will be used for the proof of the main result in Section 4. Lemma 2.1. Take 0 ≤ τ ≤ t and f, g ∈ L1 Z t−τ g(t−s) (χ ∗ f ) (s)ds = (g ∗ (χ ∗ f )) (t)−(χ ∗ g) (t−τ ) (χ ∗ f ) (τ ). loc(R+). Then tα−1 Γ(α) 0 0 f (t−s) (χ ∗ g) (s)ds+Z τ dsZ t f (t − u)du = Z τ t−τ d 0 s Proof. Observe that have: Z t Z t−τ 0 f (t − s)Z s 0 parts in the following integral to obtain, for 0 ≤ τ ≤ t. Note that Z t−τ 0 g(s)Z t−s 0 f (t − u)du = −f (t − s) and by simple change of variable we f (u)du. We integrate by f (s)ds and Z t s f (t − u)du =Z t−s 0 0 0 0 t−τ f (t − s)dsZ t−τ g(u)duZ t−τ g(u)du +Z t−τ g(u)duds = −Z t f (t − u)du +Z t−τ = −Z t−τ = − (χ ∗ g) (t − τ ) (χ ∗ f ) (τ ) +Z t−τ f (x)dxds −Z t g(s)Z t g(s)Z s g(s)Z t−s g(s)Z t−s f (x)dxds = Z t g(s)Z t−s 0 0 0 0 0 0 s f (x)dxds f (t − u)duds f (t − u)duds f (u)duds, 0 0 = (g ∗ (χ ∗ f )) (t) −Z τ = (g ∗ (χ ∗ f )) (t) −Z τ 0 0 t−τ g(t − u)Z u 0 f (x)dxdu g(t − u) (χ ∗ f ) (u)du, (cid:3) and this concludes the proof. Taking f = jα and g = jβ with α, β > 0 in Lemma 2.1, we get the equality sβ Γ(β + 1) ds+Z τ 0 (t − s)β−1 Γ(β) sα Γ(α + 1) ds = tα+β Γ(α + β + 1)− (t − τ )β Γ(β + 1) τ α Γ(α + 1) (t − s)α−1 Z t−τ 0 Γ(α) for 0 ≤ τ ≤ t. If we set f = g in Lemma 2.1, we obtain: Corollary 2.2. Take 0 ≤ τ ≤ t and f ∈ L1 loc(R+). Then (2.4) (cid:18)Z t−τ 0 0 (cid:19) f (t − s) (χ ∗ f ) (s)ds = (f ∗ (χ ∗ f )) (t) − (χ ∗ f ) (t − τ ) (χ ∗ f ) (τ ). +Z τ SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 5 Further specializing to f = jα for α > 0, yields the identity: (cid:18)Z t−τ 0 0 (cid:19) (t − s)α−1 +Z τ Γ(α) for 0 ≤ τ ≤ t. sα Γ(α + 1) ds = t2α Γ(2α + 1) − (t − τ )α Γ(α + 1) τ α Γ(α + 1) As a consequence of the last corollary, we obtain another proof of the following result given in [19, Lemma 2.1.12] for continuous functions and in [23, Lemma 3.1] for f (t) = tα−1 Γ(α) Corollary 2.3. For f ∈ L1 and α > 0. in particular for α > 0 and s, u ≥ 0, we get that loc(R+) and s, u ≥ 0 we have 0 −Z u −Z s 0 −Z u −Z s 0 (cid:19) f (u + s − r)f (r)dr = 0; 0 (cid:19) (u + s − r)α−1rα−1dr = 0; 0 (cid:18)Z s+u (cid:18)Z s+u 0 Proof. By change of variable, we write the identity (2.4) as (f ∗ (χ ∗ f )) (t) − (χ ∗ f ) (t − τ ) (χ ∗ f ) (τ ) =(cid:18)Z t τ t−τ(cid:19) f (x)(χ ∗ f )(t − x)dx +Z t for 0 ≤ τ ≤ t. Now we observe that Similarly, we have d dt (f ∗ (χ ∗ f )) (t) = (f ∗ f )(t). d dt ((χ ∗ f )(t − τ )(χ ∗ f )(τ )) = f (t − τ )(χ ∗ f )(τ ), d dtZ τ t f (u)(χ ∗ f )(t − u)du = −f (t)(χ ∗ f )(0) −Z t τ f (u)f (t − u)du = Z τ t f (u)f (t − u)du, and d dtZ t t−τ f (u)(χ ∗ f )(t − u)du = f (t)(χ ∗ f )(0) +Z t 0 f (u)f (t − u)du −f (t − τ )(χ ∗ f )(τ ) −Z t−τ = Z t −Z t−τ f (u)f (t − u)du 0 0 0 f (u)f (t − u)du − f (t − τ )(χ ∗ f )(τ ) f (u)f (t − u)du 6 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA Differentiating with respect to the variable t and using the above, we have: (f ∗ f )(t) = Z t = Z t−τ f (x)f (t − x)dx +Z t f (t − s)f (s)ds +Z τ t−τ 0 0 τ f (x)f (t − x)dx f (t − s)f (s)ds for 0 ≤ τ ≤ t. Now take t = s + u, and τ = s and we conclude the proof. (cid:3) Take k ∈ L1 loc([0, τ )), and we define (kt)t∈[0,τ ) ⊂ L1 (2.5) kt(s) := k(t − s)χ[0,t](s), loc([0, τ )) by s ∈ [0, τ ). A similar result was considered in [18, Proposition 2.2] for functions belong to L1 Here we present a direct proof for L1 Theorem 2.4. Take k ∈ L1 loc([0, τ )) and (kt)t∈[0,τ ) defined by (2.5). Then loc([0, τ )). loc(R+). kt ∗ ks(x) =Z t+s t k(t + s − r)kr(x)dr −Z s 0 k(t + s − r)kr(x)dr, 0 ≤ x < τ, for 0 ≤ s, t ≤ t + s < τ . Proof. We consider (without loss of generality) that 0 ≤ s ≤ t. First we consider 0 ≤ x ≤ s. Then k(t − (x − y))k(s − y)dy, 0 ≤ x < τ, and 0 kt ∗ ks(x) =Z x 0 (cid:19) k(t + s − r)kr(x)dr =(cid:18)Z t+s −Z s k(u)k(t + s − u − x)du −Z s−x k(u)k(t + s − u − x)du =Z x 0 0 t t (cid:18)Z t+s =Z s =Z s s−x 0 x (cid:19) k(t + s − r)k(r − x)dr −Z s k(t + s − x − y)k(y)dy k(s − y)k(t − x + y)dy where we have changed variables in several equalities. The other cases s ≤ x ≤ t, t ≤ In particular we remark that x ≤ t + s and t + s ≤ x < τ are made following ideas. kt ∗ ks(x) = 0 for t + s ≤ x < τ . (cid:3) Remark 2.5. By Proposition 4.3 and Theorem 2.4, we may conclude that (kt)t∈[0,τ ) is a local k-convoluted semigroup in L1 loc([0, τ )). In fact, note that kt = δt ∗ k where (δt)t≥0 is the Dirac measure concentrated at t = 0. In this sense, (kt)t∈[0,τ ) is the canonical local k-convoluted semigroup. 3. The Laplace transform and k-Test function spaces Let k ∈ L1 loc(R+). We write bybk the usual Laplace transform of k, given by e−λtk(t)dt, N→∞Z N 0 bk(λ) = lim SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 7 in the case that there exists for λ ∈ C; abs(k) is defined by abs(k) := inf{ℜλ; existbk(λ)}, see [3, Section 1.4]. In the case that k(t) ≤ M eωt for a.e. t ≥ 0 and M, ω > 0, we have that bk(λ) =Z ∞ 0 e−λtk(t)dt, ℜλ > ω. For λ ∈ C, we write eλ(t) = e−λt, t ≥ 0. Lemma 3.1. Take k ∈ L1 Then In particular, ifbk(λ) = 0 for some ℜλ > ω, then k ◦ e−λ = 0. Proof. Take λ ∈ C with ℜλ > ω and ℜλ > ω. loc(R+) such that k(t) ≤ M eωt for a.e. t ≥ 0 and M, ω > 0. k ◦ e−λ =bk(λ)e−λ, k(s − t)e−λsds =Z ∞ k(u)e−λudue−λt =bk(λ)e−λ(t) (cid:3) 0 k ◦ e−λ(t) =Z ∞ t for t > 0. We write by supp(h) the usual support of a function h defined in R; D is the space of C(∞) functions with compact support on R and D0 is the set of C(∞) functions with compact support on [0,∞), D0 ⊂ D. The space D will be equipped with the Schwartz topology which turns it into a complete topological vector space. We denote the topology by τ. In particular, sequential convergence in D is described by: let (φn)n≥1 ⊂ D, φ ∈ D, then φn →τ φ if and only if (i) there exists a compact subset K ⊂ R such that supp(φn), supp(φ) ⊂ K. (ii) for any j ≥ 0, φ(j) n → φ(j) uniformly on compacts sets. Note that D0 is a closed subspace of D and then (D0, τ ) is a complete topological space (we keep the same notation for the topology τ and its restriction to the subspace D0). We denote by D+ the set of functions defined by φ+ : [0,∞) → C, given by φ+(t) := φ(t) for t ≥ 0 and φ ∈ D and define K : D → D+ by K(φ) = φ+ for φ ∈ D. Due to the extension theorem of R. T. Seeley [32], there exists a linear continuous operator Λ : D+ → D, such that KΛ = ID+; in particular if ψ is a C(∞) functions on [0,∞) and compact support then ψ ∈ D+. The space D+ is also a complete topological vector spaces equipped with the τ -topology of uniform convergence on compact subsets. We define the operator T ′k : D → D by f 7→ T ′k(f ) := k ◦ f, that is, T ′k(f )(t) =Z ∞ t k(s − t)f (s)ds, t ≥ 0. We shall also use the same notation for the restriction to D+ that is, T ′k : D+ → D+; however T ′k : D0 6→ D0. Note that T ′k(fu) = (T ′k(f ))u, where fu(t) = f (u + t) for u, t ≥ 0 and f ∈ D+. In the case that 0 ∈ supp(k), we have that T ′k : D+ → D+ is an injective, linear and continuous homomorphism such that T ′k(f ◦ g) = f ◦ T ′k(g), f, g ∈ D+, 8 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA see [18, Theorem 2.5]. Then, we define the space Dk by Dk := T ′k(D+) ⊂ D+ and the right inverse map of T ′k, i.e., Wk : Dk → D+ by f (t) = T ′k(Wk(f ))(t) =Z ∞ k(s − t)Wkf (s)ds, see [18, Definition 2.7]. Note that the operator Wk : D+ → D+ is a closed operator (D(Wk) = Dk), but we cannot apply the open mapping theorem to conclude that it is continuous. It is clear that the subspace Dk is also a topological algebra: take f, g ∈ Dk, then f ∗ g ∈ Dk ([18, Theorem 2.10]) and the map (f, g) → f ∗ g is continuous in Dk. Moreover Wk(k ◦ f ) = f for f ∈ D+ and fu ∈ Dk, with Wk(fu) = (Wk(f ))u, (3.6) f ∈ Dk, t ≥ 0, t f ∈ Dk, u ≥ 0. We have the following property to the effect that Wk does not increase the support. loc(R+) be such that 0 ∈ supp(k) and let a > 0. Then supp(f ) ⊂ Lemma 3.2. Let k ∈ L1 [0, a] if and only if supp(Wkf ) ⊂ [0, a] for f ∈ Dk. Proof. Take f ∈ Dk such that supp(f ) ⊂ [0, a]. Then 0 = f (t) = k ◦ Wkf (t) =Z ∞ t k(s − t)Wkf (s)ds, t ≥ a, Wkf = l ◦ Wk∗lf, f ∈ Dk∗l. see [18, Lemma 2.8]. A consequence of (3.8) is the following lemma. Lemma 3.3. Take k ∈ L1 (i) Dk∗n ֒→ Dk∗m for n ≥ m ≥ 1. (ii) Wk∗mf = kn−m ◦ Wk∗nf = Wk∗n(kn−m ◦ f ) and if supp(Wk∗nf ) ⊂ I with I an interval in R+ then supp(Wk∗mf ) ⊂ I for f ∈ Dk∗n and n ≥ m ≥ 1. loc(R+) such that 0 ∈ supp(k). Then The next definition gives the test function space will be used later to obtain new dis- tribution spaces and corresponding distribution semigroups. We write t = a + r for t ≥ a and r ≥ 0, 0 =Z ∞ t k(s − t)Wkf (s)ds =Z ∞ 0 k(x)Wkf (x + a + r)dx =Z ∞ r k(x − r)(Wkf )a(x)dx, where (Wkf )a(x) := (Wkf )(x+a) for x > 0. We apply the Titchmarsh-Foia¸s [18, Theorem 2.4] to conclude that (Wkf )a(x) = 0 for x > 0, i.e., supp(Wkf ) ⊂ [0, a]. Conversely, suppose that supp(Wkf ) ⊂ [0, a]. It then follows from the representation f (t) =Z ∞ t k(s − t)Wkf (s)ds, t ≥ 0, (cid:3) that supp(f ) ⊂ [0, a]. Note that in the case that f ∈ Dk then f (n) ∈ Dk and Wk(f (n)) = (Wkf )(n), n ≥ 1; loc(R+) such that 0 ∈ supp(k)∩ supp(l). Then 0 ∈ supp(k ∗ l), Dk∗l ⊂ Dk ∩Dl (3.7) take k, l ∈ L1 and (3.8) SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 9 Definition 3.4. Take k ∈ L1 defined by loc(R+) such that 0 ∈ supp(k). We denote by Dk∗∞ the space Dk∗∞ := Dk∗n. ∞\n=1 It is clear that Dk∗∞ is also a topological algebra (equipped with the τ -topology) and Dk∗∞ ֒→ Dk∗n ֒→ D+. In fact, Dk∗∞ is the inverse (or projective) limit of the family (Dk∗n)n≥1. By Lemma 3.3, Wk∗n : Dk∗∞ → Dk∗∞ and k∗n ◦ Wk∗nf = f for f ∈ Dk∗∞ and n ∈ N. Note that if f ∈ Dk∗∞ then fu ∈ Dk∗∞ for u ≥ 0, see formula (3.6). Example 3.5. In the case that Dk = D+, then Dk∗n = D+ for n ∈ N and Dk∗∞ = D+. Take z ∈ C, ez(t) := ezt for t ≥ 0, Then Dkez = {ezf f ∈ Dk} = Dk and Wkez f = ezWk(e−zf ), f ∈ Dk; in the case that Dk = D+, then Dkez = D+, see [18, Proposition 2.9]. (i) Recall that α > 0 and jα(t) := tα−1 of order α, Wα and Djα = Dj ∗∞ of usual derivation, see more details for example in [31]. α = D+; note that for α ∈ N, Wα = (−1)α dα (ii) Given α > 0 and z ∈ C, we have that Dezjα = D(ezjα)∗∞ = D+ and Γ(α) ; the map Wjα is the Weyl fractional derivative dtα , the α-iterate Wezjαf = ezWα(e−zf ), f ∈ D+; for α = 1, 2 see explicit expressions in [18, Section 2]. (iii) It is straightforward to check that T ′χ(0,1) Dχ(0,1)∗n = D+ and (f )(t) =R t+1 t f (s)ds for f ∈ D+, Dχ(0,1) = Wχ(0,1)f (t) = − ∞Xn=0 f′(t + n), f ∈ D+, t ≥ 0. Now let f, g ∈ Dk. Then f ∗ g ∈ Dk and (3.9) Wk(f ∗ g)(s) =Z s −Z ∞ 0 s Wkg(r)Z s Wkg(r)Z ∞ s−r s k(t + r − s)Wkf (t)dtdr k(t + r − s)Wkf (t)dtdr, see [18, Theorem 2.10]. Under some conditions on the function k, some Banach algebras under the convolution product may be considered as the next theorem shows. Theorem 3.6. ([18, Theorems 3.4 and 3.5]) Let k ∈ L1 supp(k) and abs(k) < ∞. Then the formula kfkk,eβ :=Z ∞ 0 Wkf (t)eβtdt, f ∈ Dk, loc(R+) be a function with 0 ∈ for β > max{abs(k), 0} defines an algebra norm on Dk for the convolution product ∗. We denote by Tk(eβ) the Banach space obtained as the completion of Dk in the norm k · kk,eβ , and then we have Tk(eβ) ֒→ L1(R+). 10 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA Note that in the three examples below, the space Dk = D+. However, as the following result shows, there are functions k such that Dk D+ and Dk∗∞ = {0}. Theorem 3.7. Take k ∈ L1 loc(R+) such that 0 ∈ supp(k) and 0 ≤ abs(k) < ∞. bk(λ0) = 0 for some ℜλ0 > abs(k) then Dk D+ and Dk∗∞ = {0}. Proof. We suppose that Dk = D+ and bk(λ0) = 0 for some ℜλ0 > abs(k). Take β ∈ R such that abs(k) < β < ℜλ0, There exists fn ⊂ D+ such that (3.10) If Wkfn(t) − e−λ0(t)eβtdt → 0, n → ∞, As a consequence of Theorem 3.6, we obtain that fn 6→ 0 in L1(R+). On the other hand, by [18, Theorem 2.5 (ii)] and (3.10), we get that fn(t) − k ◦ e−λ0(t)eβtdt → 0, n → ∞. By Lemma 3.1 k ◦ e−λ0 = 0, and then fn → 0 in L1(R+). We conclude that Dk D+. Now take f ∈ Dk∗∞. Then there exists a sequence (gn) ⊂ D+, such that f = k∗n ∗ gn for n ≥ 1. Then bf (λ0) =(cid:16)bk(λ0)(cid:17)n zero on λ0 of order n at least for any n ≥ 1. We conclude that f = 0. Example 3.8. The following example was presented in [4, Section 5] and appeared later in other references in connection to convoluted semigroups (see [20, Example 6.1] and [19, Example 2.8.1]). Let bgn(λ0) = 0 for any n ≥ 1. We conclude that bf has a (cid:3) Z ∞ 0 Z ∞ 0 K(λ) := 1 λ2 n2 − λ n2 + λ , ℜλ > 0. ∞Yn=0 Then there exists a continuous and exponentially bounded function κ in [0,∞) such that We note that for the cases k(t) = jα(t), t > 0 (corresponding to local integrated semi- bκ = 1/K. Moreover, 0 ∈ supp(κ) and we apply Theorem 3.7 to conclude that Dκ D+. groups) andbk(λ) == in [8]) and will be presented in Section 6), we havebk(λ) 6= 0 for all ℜλ > abs(k). (where l > 0, 0 < a < 1 and which are considered ∞Yj=1 1 + 4. Local convoluted semigroups a!−1 lz 1 j The definition of global k-convoluted semigroups was introduced by the first time by I. Cioranescu [8] and subsequently developed in [9] (see also [17] and the monographs [19] and [27]). We will consider the following definition of local k-convoluted semigroup as appears in [20, Definition 2.1]. Definition 4.1. Let 0 < τ ≤ ∞, k ∈ L1 loc([0, τ )) and A be a closed operator. (Sk(t))t∈[0,τ ) ⊂ B(X) a strongly continuous operator family. The family (Sk(t))t∈[0,τ ) is a local k-convoluted semigroup (or local k-semigroup in short) generated by A if Sk(t)A ⊂ ASk(t),R t 0 Sk(s)xds ∈ D(A) for t ∈ [0, τ ) and x ∈ X and (4.11) AZ t 0 Sk(t)xdt = Sk(t)x −Z t 0 k(s)dsx, x ∈ X, SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 11 for t ∈ [0, τ ); in this case the operator A is called the generator of (Sk(t))t∈[0,τ ). We say that (Sk(t))t∈[0,τ ) is non degenerate if S(t)x = 0 for all 0 ≤ t < τ implies x = 0. Alternatively, in relation to Problem (1.3), we note that when the problem is well posed in the sense that for every x ∈ X, there exists a unique solution v ∈ C 1([0, τ ), X) ∩ C([0, τ ), D(A)), we set S(t)x = v′(t), 0 ≤ t < τ, x ∈ X. It follows from the Closed Graph Theorem that S(t) ∈ B(X), 0 ≤ t < τ. Clearly, t 7→ S(t) is strongly continuous from [0, τ ) to B(X). The local convoluted semigroup defined in this manner is necessarily non degenerate, due to the uniqueness assumption. It is easy to prove that if A generates a k-convoluted semigroup (Sk(t))t∈[0,τ ), then Sk(0) = 0 and Sk(t)x ∈ D(A) for t ∈ [0, τ ) and x ∈ X. See more details, for example in [17] and [20]. Remark 4.2. (i) For α > 0 and k(t) = tα−1 Γ(α) for t > 0, we get α-times integrated semigroups which were introduced in [14]. We follow the usual notation (Sα(t))t∈[0,τ ) for α-times local integrated semigroups. (ii) If C ∈ B(X) is an injective operator and we set k(t) ≡ C, 0 ≤ t < τ then we recover the concept of local C−regularized semigroups. Local C−regularized semigroups were first studied in [33]. (iii) One condition in the definition of local convoluted semigroup, equation (4.11) may be interpreted as a Duhamel formula for the abstract Cauchy problem. More precisely, if we are interested in the (non-homogeneous) initial value problem (u′(t) = Au(t) + F (t), 0 ≤ t < τ u(0) = x ∈ X, where F is an X−valued function and τ ∈ (0, ∞], and K(.) takes values in B(X) with the additional assumptions that K(t)K(s) = K(s)K(t), t, s ∈ [0, τ ); AK(t)x = K(t)Ax, t ∈ [0, τ ), x ∈ D(A), we can consider the regularized problem (v′(t) = Av(t) + K(t)x + FK (t), 0 ≤ t < τ v(0) = 0, in which FK(t) = (K ∗ F )(t) =R t 0 K(t − s)F (s)ds, 0 ≤ t < τ More details can be found in the reference [9]. We hall be concerned only with the situation where K(t) = φ(t)I where I is the identity operator on X. Spectral criteria for the generation of local convoluted semigroups involving the resolvent of the generator can be found in the references [8], [9], [17] and [19]. (iv) Other equivalent definitions of local convoluted semigroup, using the composition property (see Proposition 4.3) or the Laplace transform ([17, Theorem 3.2]) show this algebraic aspect in a straightforward way. The next characterization of local k-semigroups has the advantage to offer an algebraic character which is crucial in the development of the theory as we will see in Theorem 5.1. The proof runs parallel as in the global case presented in [20, Proposition 2.2], see also [19, Proposition 2.1.5]. Proposition 4.3. Let 0 < τ ≤ ∞, k ∈ L1 loc([0, τ )), A a closed linear operator and (Sk(t))t∈[0,τ ) a non-degenerate strongly continuous operator family. Then (Sk(t))t∈[0,τ ) is 12 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA a local k-convoluted semigroup generated by A if and only if Sk(0) = 0 and k(t + s − r)Sk(r)xdr −Z s 0 k(t + s − r)Sk(r)xdr, x ∈ X, (4.12) Sk(t)Sk(s)x =Z t+s for 0 ≤ s, t ≤ t + s < τ . Note that if k ∈ L1 t loc([0, τ )), and we define (kt)t∈[0,τ ) ⊂ L1 loc([0, τ )) by kt(s) := k(t − s)χ[0,t](s), s ∈ [0, τ ). By Proposition 4.3 and Corollary 2.3, we may conclude that (kt)t∈[0,τ ) is a local k- convoluted semigroup in L1 loc([0, τ )). In this paper, we only consider local k-convoluted semigroups which are non-degenerate. The next theorem is the main result in this paper and shows how a local k-convoluted semigroup (Sk(t))t∈[0,τ ) is extended to [0, nτ ); in fact we get a local k∗n-convoluted semi- group in [0, nτ ) for n ∈ N. Note that we improve previous results ([9, Section 2] and [19, Theorem 2.1.1.9]): our approach is sharper that n-iterations of these theorems. Theorem 4.4. Let n ∈ N, 0 < τ ≤ ∞, k ∈ L1 loc([0, (n+1)τ )) and (Sk(t))t∈[0,τ ) be a local k- convoluted semigroup generated by A. Then the family of operators (Sk∗(n+1)(t))t∈[0,(n+1)κ] defined by Sk∗(n+1)(t)x =Z t 0 k(t − s)Sk∗n(s)xds, x ∈ X, for t ∈ [0, nκ] and 0 k(t − s)Sk∗n(s)xds +Z t−nκ Sk∗(n+1)(t)x = Sk∗n(nκ)Sk(t− nκ)x +Z nκ k∗n(t − s)Sk(s)xds, for x ∈ X and t ∈ [nκ, (n + 1)κ] is a local k∗(n+1)-semigroup generated by A for any κ < τ . Then we conclude that A generates a local k∗(n+1)-semigroup (Sk∗(n+1)(t))t∈[0,(n+1)τ ). Proof. Note that the family of operators (Sk∗(n+1)(t))t∈[0,(n+1)κ] is strongly continuous. It is known that (Sk∗(n+1)(t))t∈[0,nκ] is a local k∗(n+1)-semigroup generated by A, see for example [19, Proposition 2.1.3] and [18, Proposition 5.2]. Now take t ∈ [nκ, (n + 1)κ] and x ∈ X. 0 Sk∗(n+1)(r)xdr ∈ D(A). Since 0 Sk∗(n+1)(r)xdr, 0 Sk∗(n+1)(r)xdr =Z nκ It is clear that Sk∗(n+1)(t)A ⊂ ASk∗(n+1)(t) and we show that R t Z t we check thatR t nκ(cid:18)Z nκ +Z t Sk∗(n+1)(r)xdr +Z t nκ Sk∗(n+1)(r)xdr ∈ D(A), i.e, Sk∗n(nκ)Sk(r − nκ)xdr Z t nκ nκ 0 0 k(r − s)Sk∗n(s)xds +Z r−nκ Sk∗n(nκ)Z t nκ 0 Sk(r − nκ)xdr ∈ D(A). As (Sk(t))t∈[0,τ ) is a local k-convoluted semigroup generated by A, we get that k∗n(r − s)Sk(s)xds(cid:19) dr ∈ D(A). SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 13 0 k(r − s)Sk∗n(s)xdsdr ∈ D(A). We apply the Fubini theorem, Now we prove thatR t change the variable u = r − s, to get that 0 nκR nκ k(r − s)Sk∗n(s)xdsdr =Z nκ Z t nκZ nκ k(u)duds =Z t Sk∗n(s)xZ t−s Z nκ k∗n(r − s)Sk(s)xdsdr =Z t Z t nκZ r−nκ nκ−s 0 0 0 0 In a similar way, it is shown that nκ Sk∗n(s)xZ t k(u)Z min{t−u,nκ} k∗n(u)Z t−u max{nκ−u,0} nκ 0 k(r − s)drds Sk∗n(s)xdsdu ∈ D(A). Sk(s)xdsdu ∈ D(A). To finish the proof we prove the equality (4.11) for t ∈ [nκ, (n + 1)κ] and x ∈ X. Note that 0 We apply the Fubini theorem to get that Sk∗(n+1)(s)xds = Sk∗(n+1)(nκ)x −Z nκ AZ t Sk∗(n+1)(s)xds = Sk∗n(nκ)Z t−nκ Z t +Z t−nκ k∗(n+1)(s)dsx + AZ t Sk∗n(r)xZ t k∗n(s − r)dsdr. Sk(u)xdu +Z nκ Sk(r)xZ t (4.13) r+nκ nκ nκ 0 0 0 0 nκ We apply the operator A to the first summand to get that Sk∗(n+1)(s)xds. k(s − r)dsdr Sk(u)xdu = Sk∗n(nκ)Sk(t − nκ)x − Sk∗n(nκ)xZ t−nκ 0 k(u)du. We apply the operator A and the Fubini theorem to obtain that k(u)du. k(u)du−R nκ−r 0 Sk∗n(r)xdrdu 0 nκ k(s− r)ds =R t−r k(u)Z min{nκ,t−u} k∗n(y)dyx(cid:19) du k(u)du 0 0 0 0 0 Sk∗n(nκ)AZ t−nκ In the second summand of (4.13) we writeR t k(u)dudr = AZ t k∗n(y)dyx(cid:19)Z t−nκ k(u)(cid:18)Sk∗n(t − u)x −Z t−u k(u)du +Z nκ k∗n(y)dy(cid:19)(cid:18)Z t−nκ k(u)dudr = Z nκ Sk∗n(r)xZ t−r AZ nκ =(cid:18)Sk∗n(nκ)x −Z nκ +Z t = Sk∗n(nκ)xZ t−nκ −(cid:18)Z nκ Sk∗n(r)xZ nκ−r t−nκ 0 0 0 0 0 0 0 0 AZ nκ 0 Using similar ideas we also get that k(t − r)Sk∗n(r)xdr 0 k(u)du(cid:19) x −Z nκ k(t − r)Z r k(u)(cid:18)Sk∗n(nκ − u)x −Z nκ−u 0 0 k∗(n+1)(r)drx. 0 = Sk∗(n+1)(nκ)x −Z nκ 0 k∗n(y)dydrx. k∗n(r)drx(cid:19) du We conclude that To finish the proof we join together all summands to have that 14 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA We apply the operator A and the Fubini theorem to obtain that In the third summand of (4.13) we write R t k∗n(s − r)dsdr = AZ t k∗n(u)du +Z t−nκ Sk(r)xZ t AZ t−nκ = Sk(t − nk)xZ nκ 0 0 0 0 r r+nκ k∗n(s − r)ds = R t r r −R r+nκ k∗n(u)Z min{t−nκ,t−u} k∗n(t − r)Z r 0 0 0 k∗n(t − r)Sk(r)xdr and finally we get k(y)dy. k∗n(s − r)ds. Sk(r)xdrdu −(cid:18)Z nκ 0 k∗n(u)du(cid:19)(cid:18)Z t−nκ 0 k(u)du(cid:19) −Z t−nκ AZ t−nκ 0 r 0 0 0 r+nκ k∗n(s − r)dsdr AZ t−nκ Sk(r)xZ r+nκ = Sk(t − nκ)xZ nκ Sk(r)xZ t =Z t−nκ k∗n(u)du −(cid:18)Z nκ k∗n(s − r)dsdr k∗n(t − r)Sk(r)xdr −Z t−nκ Sk∗(n+1)(s)xds = Sk∗n(nκ)Sk(t − nκ)x +Z nκ +Z t−nκ −Z nκ k∗n(t − r)Sk(r)xdr −(cid:18)Z nκ k∗n(y)dydrx −Z t−nκ k(t − r)Z r = Sk∗(n+1)(t)x −Z t 0 k∗(n+1)(s)xds, AZ t 0 0 0 0 0 0 0 0 0 k∗n(u)du(cid:19)(cid:18)Z t−nκ 0 k(u)du(cid:19) . k∗n(t − r)Z r 0 k(y)dy. k(t − r)Sk∗n(r)xdr k∗n(y)dy(cid:19)(cid:18)Z t−nκ k∗n(t − r)Z r 0 0 k(u)du(cid:19) x k(y)dydr where we have used the Lemma 2.1. This proves the claim. (cid:3) In fact, the expression of the (Sk∗(n+1)(t))t∈[0,(n+1)κ] is not unique as shown next result. Both proof are similar to the proof of Theorem 4.4 and are left to the reader. Theorem 4.5. Let n ≥ 2, 0 < τ ≤ ∞, k ∈ L1 loc([0, nτ )) and (Sk(t))t∈[0,τ ) is a local k- convoluted semigroup generated by A. Then the family of operators (Sk∗n(t))t∈[0,nκ] defined in Theorem 4.4 verify that Sk∗n(t)x =Z t 0 k∗(n−j)(t − s)Sk∗j (s)xds, x ∈ X, for t ∈ [0, jκ] and Sk∗n(t)x = Sk∗j (nκ)Sk∗(n−j)(t − jκ)x +Z jκ 0 k∗(n−j)(t − s)Sk∗j (s)xds +Z t−jκ 0 for x ∈ X, 1 ≤ j ≤ n − 1 and t ∈ [jκ, nκ] and any κ < τ . k∗j(t − s)Sk∗(n−j)(s)xds, SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 15 Corollary 4.6. Let 0 < τ ≤ ∞, k ∈ L1 loc([0, 2τ )) and (Sk(t))t∈[0,τ ) be a local k-convoluted semigroup. Then the family of operators (Sk∗k(t))t∈[0,2κ] defined in Theorem 4.4 verifies that Sk∗k(t + s)x = Sk(t)Sk(s)x +(cid:18)Z t 0 for t, s ∈ [0, κ) and x ∈ X. 0 (cid:19) k(t + s − u)Sk(u)xdu +Z s The next theorem was given in [29, Theorem 2] in the case n = 1. Corollary 4.7. Let n ∈ N, 0 < τ ≤ ∞, (Sα(t))t∈[0,τ ) is a local α-times integrated semi- group generated by A. Then the family of operators (S(n+1)α(t))t∈[0,(n+1)κ] defined by (t − s)α−1 Γ(α) Snα(s)xds, x ∈ X, S(n+1)α(t)x =Z t S(n+1)α(t)x = Snα(nκ)Sα(t−nκ)x+Z nκ for t ∈ [0, nκ] and 0 0 (t − s)α−1 Γ(α) Snα(s)xds+Z t−nκ 0 (t − s)nα−1 Γ(nα) Sα(s)xds, for x ∈ X and t ∈ [nκ, (n + 1)κ] is a local α-times integrated semigroup generated by A for any κ < τ . Then we conclude that A generates a local α-times integrated semigroup generated by A, (S(n+1)α(t))t∈[0,(n+1)τ ). Remark 4.8. In the case α ∈ N, and t ∈ [nκ, (n + 1)κ], note that and 0 Z nκ Z t−nκ 0 (t − s)α−1 Γ(α) Snα(s)xds = (t − s)nα−1 Γ(nα) Sα(s)xds = α−1Xj=0 nα−1Xj=0 (t − nκ)j j! S(n+1)α−j(nκ) (nκ)j j! S(n+1)α−j(t − nκ) for x ∈ X and we recover the extension given in [2, Theorem 4.1] for the case n = 1. To finish this section, we mention that other extension result (for scalar α-times semigroups in this case) may be found in [21, Lemma 4.4], t ∗ I n I n s = I 2n s+t − n−1Xj=0(cid:18) sj j! I 2n−j t tj j! − I 2n−j s (cid:19) , t, s > 0, where I n t (r) := (t−r)n n! χ[0,t](r), r ∈ R+ and n ∈ N ∪ {0}. 5. Algebra homomorphisms and local convoluted semigroups As the following theorem shows, local k-convoluted semigroups induce algebra homo- morphisms from certain spaces of test functions Dk∗∞. Note that the extension theorem (Theorem 4.4) is necessary to define the algebra homomorphisms from functions defined on R+. The space Dk∗∞ is introduced in Definition 3.4. 16 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA Theorem 5.1. Let k ∈ L1 loc(R+) with 0 ∈ supp(k), and (Sk(t))t∈[0,τ ] a non-degenerate local k-convoluted semigroup generated by A. We define the map Gk : Dk∗∞ → B(X) by Gk(f )x :=Z nτ x ∈ X, f ∈ Dk∗∞, Wk∗nf (t)Sk∗n(t)xdt, 0 where supp(f ) ⊂ [0, nτ ] and (Sk∗n(t))t∈[0,nτ ] is defined in Theorem 4.4 for some n ∈ N. Then the following properties hold. (i) The map Gk is well defined, linear and bounded. (ii) The map Gk(f ∗ g) = Gk(f )Gk(g) for f, g ∈ Dk∗∞. (iii) Gk(f )x ∈ D(A) and AGk(f )x = −Gk(f′)x − f (0)x for any f ∈ Dk∗∞ and x ∈ X. Proof. Take f ∈ Dk∗∞ and supp(f ) ⊂ [0, nτ ] for some n ∈ N. First, we prove that Gk is well defined. Let m ≥ n, k∗m = k∗n ∗ k∗(m−n), and k∗(m−n) ◦ Wk∗mf = Wk∗nf by Lemma 3.3 (ii). Now we apply the Lemma 3.2 to conclude supp(Wk∗mf ) ⊂ [0, nτ ]. By Theorem 4.5 and the Fubini theorem, we get that 0 Z mτ =Z nτ Wk∗mf (t)Sk∗m(t)xdt =Z nτ k∗(m−n) ◦ Wk∗mf (t)Sk∗n(t)xdt =Z nτ 0 0 0 Wk∗mf (t)(k∗(m−n) ∗ Sk∗n)(t)xdt Wk∗nf (t)Sk∗n(t)xdt for x ∈ X. It is direct to check that Gk is linear. Now take (fn)n≥1 ⊂ Dk∗∞, and f ∈ Dk∗∞ such that fn → f . Then there exists n ∈ N such that supp(fn), supp(f ) ⊂ [0, nτ ]. Note that the map t 7→ Sk∗n(t)x, [0, nτ ] → X, is continuous and 0 Wk∗nfn(t)−Wk∗nf (t)kSk∗n(t)xkdt ≤ CxZ nτ kGk(fn)x−Gk(f )xk ≤Z nτ for x ∈ X. Now consider the operator T ′k∗n : L1[0, nτ ] → L1[0, nτ ], f 7→ T ′k∗n(f ) = k∗n ◦ f given in Section 3. By the open mapping theorem, T ′k∗n is open; we conclude that Wk∗nfn → Wk∗nf in L1[0, nτ ], the map Gk is bounded and the part (i) is proved. Take f, g ∈ Dk∗∞, i.e. f, g ∈ Dk∗n and then f ∗ g ∈ Dk∗n (see [18, Theorem 2.10]) for n ≥ 1. Then f ∗ g ∈ Dk∗∞. Now we show that Gk(f ∗ g) = Gk(f )Gk(g). Take n ∈ N such that supp(f ), supp(g) ⊂ [0, nτ ] and by Lemma 3.2, supp(Wk∗2nf ), supp(Wk∗2ng) ⊂ [0, nτ ]. Then supp(f ∗ g) ⊂ [0, 2nτ ] and supp(Wk∗2n(f ∗ g)) ⊂ [0, 2nτ ]. By (3.9) we have that Wk∗n(fn−f )(t)dt 0 0 Gk(f ∗ g)x = Z 2nτ Wk∗2ng(r)Z t = Z 2nτ Z t −Z 2nτ Z 2nτ 0 0 0 t t−r Wk∗2ng(r)Z 2nτ t Wk∗2n(f ∗ g)(t)Sk∗2n (t)xdt k∗2n(s + r − t)Wk∗2nf (s)dsdrSk∗2n(t)xdt k∗2n(s + r − t)Wk∗2nf (s)dsdrSk∗2n(t)xdt. SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 17 By Fubini theorem, we obtain the four integral expressions Gk(f ∗ g)x = Z 2nτ 0 Wk∗2nf (s)Z s+r r k∗2n(s + r − t)Sk∗2n(t)xdtdsdr r s 0 0 k∗2n(s + r − t)Sk∗2n(t)xdtdsdr Wk∗2nf (s)Z s+r Wk∗2nf (s)Z s k∗2n(s + r − t)Sk∗2n(t)xdtdsdr Wk∗2nf (s)Z r k∗2n(s + r − t)Sk∗2n(t)xdtdsdr k∗2n(s + r − t)Sk∗2n(t)xdt(cid:19) dsdr Wk∗2nf (s)(cid:18)Z s+r −Z s k∗2n(s + r − t)Sk∗2n(t)xdt(cid:19) dsdr Wk∗2nf (s)(cid:18)Z s+r −Z r 0 0 0 s r r 0 0 0 0 Wk∗2ng(r)Z 2nτ Wk∗2ng(r)Z r Wk∗2ng(r)Z 2nτ Wk∗2ng(r)Z r +Z 2nτ −Z 2nτ −Z 2nτ Wk∗2ng(r)Z r Wk∗2ng(r)Z nτ Wk∗2ng(r)Sk∗2n(r)Z r Wk∗2ng(r)Sk∗2n(r)Z nτ Wk∗2ng(r)Sk∗2n(r)Z 2nτ 0 0 0 r r 0 0 = Z 2nτ +Z 2nτ = Z 2nτ +Z 2nτ = Z 2nτ 0 0 0 Wk∗2nf (s)Sk∗2n(s)xdsdr Wk∗2nf (s)Sk∗2n(s)xdsdr Wk∗2nf (s)Sk∗2n(s)xdsdr = Gk(g)Gk(f )x, where we have applied the formula (4.12) and the part (ii) is shown. Now we consider f ∈ Dk∗∞, supp(f ) ⊂ [0, nτ ] and x ∈ X. We apply the formulae (3.7) and (4.11) to get 0 AGk(f )x = = −AZ nτ = − Z nτ = −Gk∗n(f′)x −Z nτ (Wk∗nf )′(t)Z t Wk∗nf′(t)(cid:18)Sk∗n(t)x −Z t 0 0 0 0 Sk∗n(s)xdsdt k∗n(s)dsx(cid:19) dt Wk∗nf (t)k∗n(t)dtx = −Gk∗n(f′)x − f (0)x, and the part (iii) is proven. (cid:3) The previous theorem allows to show as a consequence one of main results in [18]. Remark 5.2. When the operator A generates a global convoluted semigroup (Sk(t))t≥0, the homomorphism Gk is defined from Dk to B(X) by Gk(f )x =Z ∞ 0 Wkf (t)Sk(t)xdt, x ∈ X, f ∈ Dk, see [18, Theorem 5.5]. Under some conditions of the boundedness of (Sk(t))t≥0, the ho- momorphism Gk may be extended to a bounded Banach algebra homomorphism, see [18, Theorem 5.6]. 18 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA Let k, l ∈ L1 loc([0, τ )) with 0 ∈ supp(k), and (Sk(t))t∈[0,τ ) a non-degenerate local k- convoluted semigroup generated by A. Then (l ∗ Sk(t))t∈[0,τ ] is a non-degenerate local k-convoluted semigroup generated by A, see a similar proof in [18, Proposition 5.2]. Corollary 5.3. Let k, l ∈ L1 non-degenerate local k-convoluted semigroups generated by A. Then loc(R+) with 0 ∈ supp(k)∩supp(l), and (Sk(t))t∈[0,τ ), (Sk∗l(t))t∈[0,τ ) f ∈ D(k∗l)∗∞, Gk∗l(f ) = Gk(f ), where Gk(f ),Gk∗l are defined in Theorem 5.1. Proof. Take f ∈ D(k∗l)∗∞ and supp(f ) ⊂ [0, nτ ]. By (3.8), f ∈ D(k∗l)∗∞ ⊂ Dk∗∞ and Wk∗n∗l∗nf (t)Sk∗n∗l∗n(t)xdt (l∗n ◦ Wk∗n∗l∗nf )(t)Sk∗n(t)xdt W(k∗l)∗nf (t)S(k∗l)∗n(t)xdt =Z nτ Wk∗n∗l∗nf (t)(l∗n ∗ Sk∗n)(t)xdt =Z nτ 0 0 0 0 Gk∗l(f )x =Z nτ =Z nτ =Z nτ 0 Wk∗nf (t)Sk∗n(t)xdt = Gk(f )x where we have applied the formula (3.8). (cid:3) Corollary 5.4. Let (Sα(t))t∈[0,τ ] a non-degenerate local α-times integrated semigroup gen- erated by A. We define the map Gα : D+ → B(X) by Wαnf (t)Sαn(t)xdt, x ∈ X, f ∈ D+, Gα(f )x :=Z nτ 0 where supp(f ) ⊂ [0, nτ ] and (Snα(t))t∈[0,nτ ] is defined in Theorem 4.4 for some n ∈ N. Then the map Gα is well defined, linear, bounded and Gk(f ∗ g) = Gk(f )Gk(g) for f, g ∈ Dk∗∞. Moreover, Gk(f )x ∈ D(A) and AGk(f )x = −Gk(f′)x − f (0)x, f ∈ D+, x ∈ X. 6. Examples and applications In this section we consider different examples of convoluted semigroups which have been presented in the literature. Our results are applied in all these examples to illustrate its importance. 6.1. Differential operators on Lp(RN ). [1, 3, 14] Let E be one of the spaces Lp(Rn) (1 ≤ p ≤ ∞), C0(Rn), BU C(Rn), or Cb(Rn) and AE the associated operator to a differential operator and defined by Fourier multipliers, see details in [14, Section 4], [3, Chapter 8]. Under some conditions, the operator AE generates an α-times integrated semigroup (Sα(t))t≥0 on E and kSα(t)k ≤ CE(1 + tα) for some constant CE and certain α > 0, see In this case the map Gα : D+ → B(Lp(RN )) [1, Theorem 6.3] and [14, Theorem 4.2]. (Corollary 5.4) extends to a Banach algebra homomorphism Gα : Tα(1 + tα) → B(Lp(RN )) where Tα(1 + tα) is the completion of D+ in the norm W αf (t)(1 + tα)dt, f ∈ D+, kfkα,0,α :=Z ∞ 0 SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 19 where Wα is the Weyl derivation of order α, see [28, Proposition 4.7]. Other examples of global α-times integrated semigroups may be found in [10, Theorem 5.3] and [15, Propo- sition 8.1]. Earlier results on the Schorodinger equation using distribution semigroups can be found in [5] where differential operators are treated using Fourier multipliers in Lp spaces. 6.2. Multiplication local integrated semigroup in ℓ2. [27, Example 1.2.6], [29, Ex- ample 1]. Let ℓ2 be the Hilbert space of all complex sequences x = (xm)m=1 such that ∞Xm=1 xm2 < ∞, with the Euclidean norm kxk :=(cid:0)P∞m=1 xm2(cid:1) 1 T(cid:17)2! 1 + i (cid:18) em m(cid:19)2 −(cid:16) m (t − s)α−1eamsxmds(cid:19)m=1 Γ(α)Z t where i2 = −1. For any α > 0 let (Sα(t))t>0 be defined by Sα(t)x =(cid:18) 1 , m ∈ N, am = m T 0 , 2 . Take T > 0 and define 2 for x ∈ D(Sα(t)) where D(Sα(t)) = {x ∈ ℓ2 ; Sα(t)x ∈ ℓ2}. Then (Sα(t))t∈[0,αT ) is a local α-times integrated semigroup on ℓ2, (Sα(t))t∈[0,αT ) ⊂ B(ℓ2), such that (Sα(t))t∈[0,αT ) cannot be extended to t ≥ αT , see [29, Example 1]. We may apply the Corollary 4.7 to define (Snα(t)) for t < nαT and Corollary 5.4 to define the map Gα : D+ → B(X) by Gα(f )x :=Z nτ 0 Wαnf (t)Sαn(t)xdt, x ∈ X, f ∈ D+, with supp(f ) ⊂ [0, nτ ]. Other examples of local integrated semigroups defined by multi- plication may be found in [2, Example 4.4 (c), (d)]. 6.3. The Laplacian on L2[0, π] with Dirichlet boundary conditions. [4, Section 3, 5], [20, Example 6.1]. The operator −∆ on L2[0, π] with Dirichlet or Neumann boundary conditions generates a polynomially bounded κ-convoluted semigroups (Sκ(t))t≥0, where kSκ(t)k ≤ C(1 + t3) where C > 0; note that κ is given in Example 3.8 and κ(t) ≤ Ceβt for t ≥ 0 and some C, β > 0. By Remarks 5.2, there exists an algebra homomorphism Gκ : Dκ → B(L2[0, π]) such that it extends to a Banach algebra homomorphism Gκ : Tκ(eβ) → B(L2[0, π]), see Theorem 3.6 and [18, Theorem 6.5]. Other examples of generators of k-convoluted semigroups (which does not generate integrated semigroups) may be found in [20, Example 6.2], or [19, Section 2.8], where k(t) = e−a2/4t 2√πt3 = 1 2πiZ r+i∞ r−i∞ eλt/a2− √λdλ, t > 0, for some a > 0 (where in the integral, r is any positive real number). In this case, we have: 1 a e−a√λ, ℜλ > 0. bk(λ) = 20 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA 6.4. Ultradistributions in the Gevrey classes. [9, Section 5. Applications ], [8, 4. Example and final comments]. Let Mk, k = 0, 1, 2, ... be a Gevrey type sequences, i.e., a sequence of positive numbers such that M0 = 1, which is logarithmically convex and non-quasianalytic: M 2 k ≤ Mk−1Mk+1, and ∞Xk=1 Mk−1/Mk < ∞, for example (k!s), (kks) and Γ(1 + ks) for s > 1.Let mk = Mk/Mk−1 for k ∈ N, P (z) = ∞Yj=1(cid:18)1 + z mj(cid:19) , ℜz > 0, (i) Let X = L2(R) and A = i d4 dt4 − d2 differential polynomial. The following two operators are generators of local K-convoluted semigroups and appear in [8, 4. Example and final comments]: and the function K defined by bK(z) = 1/P (z). The entire function P (z) is called an ultra- j2(cid:17) for some l > 0. on [0, τ ), where bK1(z) = 1/P1(z) where P1(z) =Q∞j=1(cid:16)1 + lz In the context of ultradistribution semigroups, this example was first proposed by J. Chazarain ([7, Remarque 6.4]). It was observed in [16] that one can take A = iB where B is the generator of a strongly continuous cosine function. Additional results are given in [19], [20] including the case of Beurling ultradistributions. Other developments are discussed in [19], including a generalization of the abstract Weierstrass formula. dt2 . Then A generates an K1-convoluted semigroup (ii) Let ω be a σ-finite measure, the Lebesgue space X = Lp(Ω) (1 ≤ p ≤ ∞) and m : Ω → C a measurable function. We define Af = mf where D(A) = {f ∈ Lp(Ω); mf ∈ Lp(Ω)} and {z ∈ C;ℜz ≥ αza + β} ⊂ ρ(A). for α, β > 0 and 0 < a < 1. For every τ > 0 there is l > 0 such that A generates a local K2-convoluted semigroup on [0, τ ) where P2(z) = ∞Yj=1 1 + a! , lz 1 j where bK2(z) = 1/P2(z). The following estimates are valid for the Gevrey sequences Mj = (j!s), Mj = (jjs) and Mj = Γ(1 + js) where s > 1 is given. e(lz)a ≤ P (z) ≤ e(Lz)a , ℜz ≥ 0, where L is a positive constant, see e.g. [8, (1.1)]. We apply the Theorem 4.4 to conclude that A generates a local K∗n-convoluted semi- group on [0, nτ ). SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 21 7. Appendix: k-Distribution Semigroups Let X be a Banach space. Vector valued algebraic distributions, i.e., linear and continu- ous maps from a test function space to the space of bounded linear operators, B(X), which satisfy an algebraic property (similar to Theorem 5.1 (ii)) have been studied deeply in a large numbers of papers, see [13, Chapter 8 ]. In this sense, distribution semigroups (in the sense of Lions, DS-L) were introduced by J.L. Lions in [25], see also [7], [2, Definition 7.1]. P.C. Kunstmann considered pre-distribution semigroups (or quasidistribution semi- group in the terminology of S. W. Wang) as linear and continuous maps G : D 7→ B(X), G ∈ D′(B(X)), satisfying see [21, Definition 2.1] and [35, Definition 3.3]. In fact, a pre-distribution G can be regarded as a continuous linear map from D+ into B(X), G : D+ 7→ B(X), such that (i) G(φ ∗ ψ) = G(φ)G(ψ) for φ, ψ ∈ D, (ii) ∩{ker(G(θ)) θ ∈ D0} = {0}, (i) G(φ ∗ ψ) = G(φ)G(ψ) for φ, ψ ∈ D+. (ii) ∩{ker(G(θ)) θ ∈ D+} = {0}. see [35, Remark 3.4]. The differences between quasi-distribution and distribution semi- groups in the sense of Lions may be found in [21, Remark 3.13]. Classes of Distribution semigroups on (0,∞) (in short DS on (0,∞)) were considered in [22, Definition 1]. The subspace D′+(B(X)) ⊂ D′(B(X)) is formed of the elements supported in [0,∞). A distribution semigroup on (0,∞), G ∈ D′+(B(X)), is a continuous linear map from D into B(X) G : D 7→ B(X), supported in [0,∞), such that (i) G(φ ∗ ψ) = G(φ)G(ψ) for φ, ψ ∈ D0. (ii) ∩{ker(G(θ)) θ ∈ D0} = {0}. Distribution semigroups on (0,∞) and quasi-distribution semigroups are not the same concept, see [22, Remark 4]. However, a particular class of distribution semigroups on (0,∞) (simply called distribution semigroups, see [22, Definition 2]) may be identified with quasi-distribution semigroups, [22, Theorem 1]. Keeping in mind these definitions and Theorem 5.1, we introduce the concept of k- distribution semigroup. loc(R+) such that 0 ∈ supp(k). We say that a linear and Definition 7.1. Let k ∈ L1 continuous map Gk : Dk∗∞ 7→ B(X) is a k-distribution semigroup, in short k-DS, if it satisfies the following conditions. (i) Gk(φ ∗ ψ) = Gk(φ)Gk(ψ) for φ, ψ ∈ Dk∗∞. (ii) ∩{ker(Gk(θ)) θ ∈ Dk∗∞} = {0}. In the case that Gk is a k-distribution semigroup on (0,∞), then Dk∗∞ 6= {0} by (ii). Remark 7.2. Let G : D → B(X) be a distribution semigroup (or quasi-distribution semi- group in the sense of Wang). Then G◦Λ is a k-distribution semigroup for any k ∈ L1 loc(R+) such that 0 ∈ supp(k) and Dk∗∞ = D+; in particular G ◦ Λ is a jα − DS for any α > 0 (see definition of Λ and jα in Section 3). For a given k-DS Gk, define the operator A′ by (i) D(A′) := ∪{Im(Gk(θ)) θ ∈ Dk∗∞}. (ii) A′Gk(θ)x := −Gk(θ′)x − θ(0)x, for x ∈ X and θ ∈ Dk∗∞. 22 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA Proposition 7.3. The operator A′ is well defined and closable. Proof. Assume that Gk(φ)x = Gk(ψ)y for some x, y ∈ X and φ, ψ ∈ Dk∗∞. Now take θ ∈ Dk∗∞. Since θ ∗ φ′(t) = θ′ ∗ φ(t) + θ(0)φ(t) − φ′(0)θ(t), t ≥ 0, see, for example [35, Proposition 3.1 (iii)], we get that and hence Similarly, Gk(θ)Gk(φ′)x = Gk(θ′)Gk(φ)x + θ(0)Gk(φ)x − φ′(0)Gk(θ)x (cid:0)−Gk(θ′) − θ(0)(cid:1) Gk(φ)x = Gk(θ)(cid:0)−Gk(φ′)x − φ′(0)x(cid:1) . (cid:0)−Gk(θ′) − θ(0)(cid:1)Gk(ψ)y = Gk(θ)(cid:0)−Gk(ψ′)y − ψ′(0)y(cid:1) . −Gk(φ′)x − φ′(0)x = −Gk(ψ′)y − ψ′(0)y By Definition 7.1 (ii), we conclude that and A′ is well defined. To prove that A′ is closable, let (xn)n≥1 ⊂ D(A′) be such that xn → 0, A′xn → y. We write xn = Gk(φn)zn with (φn)n≥1 ⊂ Dk∗∞ and (zn)n≥1 ⊂ X. Take θ ∈ Dk∗∞ and then Gk(θ)y = lim = lim n→∞Gk(θ)A′xn = lim n→∞Gk(θ)A′Gk(φn)zn n→∞Gk(θ)(cid:0)−Gk(φ′n)zn − φn(0)zn(cid:1) n→∞(cid:0)−Gk(θ′) − θ(0)(cid:1)Gk(φn)zn = lim = lim n→∞(cid:0)−Gk(θ′) − θ(0)(cid:1) xn = 0. This implies that y = 0 by Definition 7.1 (ii). Definition 7.4. The closure of A′, denoted by A, is called the generator of a Gk. (cid:3) Ax = y, Other definitions of generator of distribution semigroups are given using approximate units (see [2, Definition 7.1]) or the distribution −δ′0 ([21, Definition 3.3] and [22, Propo- sition 1]). In our case, given a k-DS Gk and its generator (A, D(A)), then D(A) ⊂ {x ∈ X exists y ∈ X such that Gk(θ)y = −Gk(θ′)x − θ(0)x for any θ ∈ Dk∗∞}; x ∈ D(A), loc(R+) with 0 ∈ supp(k), (Sk(t))t∈[0,τ ] a non-degenerate local k- Theorem 7.5. Let k ∈ L1 convoluted semigroup generated by A and Gk : Dk∗∞ → B(X) the map defined in Theorem 5.1. Then Gk is a k-DS generated by A. Proof. The first condition in Definition 7.1 appears in Theorem 5.1 (ii). Take x ∈ X such that Gk(φ)(x) = 0 for any φ ∈ Dk∗∞. Since φ′ ∈ Dk∗∞, we apply Theorem 5.1 (iii) to conclude 0 = φ(0)x for any φ ∈ Dk∗∞ and then 0 = φu(0)x for any u ≥ 0. We conclude that 0 = φ(u)x for any u ≥ 0 and φ ∈ Dk∗∞. Then x = 0 and the condition Definition 7.1 (ii) holds. Note that AGk(θ)x := −Gk(θ′)x− θ(0)x, for x ∈ X and θ ∈ Dk∗∞, see Theorem 5.1 (iii). As A is a closed operator, we conclude that A is the generator of Gk. A straightforward consequence of Theorem 7.5 is the next corollary. (cid:3) SHARP EXTENSIONS FOR SOLUTIONS OF ABSTRACT CAUCHY PROBLEMS 23 Corollary 7.6. Let (Sα(t))t∈[0,τ ] a non-degenerate local α-times integrated semigroup gen- erated by A and Gα : D+ → B(X) the map defined in Corollary 5.4. Then Gα is a jα-DS generated by A with α > 0. When the generator A is densely defined in Corollary 7.6, it is equivalent that A gener- ates a n-times integrated semigroup for some n ∈ N and A is the generator of a distribution semigroup in the sense of Lions, see [2, Theorem 7.2, Corollary 7.3]. To consider the test-function space Dk∗∞ may be given for a wide class of vector- valued distributions such that different distribution semigroups fall into the scope of this approach. References [1] W. Arendt, H. Kellermann, Integrated solutions of Volterra integrodifferential equations and appli- cations, in: Volterra Integrodifferential Equations in Banach spaces and Applications, Trento, 1987, in: Pitman Res. Notes Math. Ser., vol. 190, Longman sci. Tech., Harlow, 1989, pp. 21-51. [2] W. Arendt, O. El-Mennaoui and V. Keyantuo, Local integrated semigroups: Evolution with jumps of regularity. J. Math. Anal. Appl. 186 (1994), 572-595. [3] W. Arendt, C. Batty, M. Hieber, F. Neubrander. "Vector-valued Laplace Transforms and Cauchy Problems". Monographs in Mathematics. vol.96. Birkhauser, Basel, 2001. [4] B. Baumer, Approximate solutions to the abstract Cauchy problems, in: Evolution Equations and their Application in Physical and Life Sciences, Lecture Notes in Pure and Applied Mathematics Vol 215(Marcel Dekker, New York, 2001), 33 -- 41. [5] M. Balabane, H. A. Emamirad, Lp estimates for Schrodinger evolution equations. Trans. Amer. Math. Soc. 292(1985), 357-373. [6] J. A. van Casteren, "Generators of strongly continuous semigroups". Research Notes in Mathematics, 115. Boston-London-Melbourne: Pitman Advanced Publishing Program, 1985. [7] J. Chazarain, Probl`emes de Cauchy abstraits et applications `a quelques probl`emes mixtes, J. Funct. Anal. 7 (1971), 386-446. [8] I. Cioranescu, Local convoluted semigroups, in: Evolution Equations (Baton Rouge, LA, 1992), 107 -- 122, Dekker, New York, 1995. [9] I. Cioranescu, G. Lumer, Probl`emes d'´evolution r´egularis´es par un noyau g´en´eral K(t). Formule de Duhamel, prolongements, th´eor`emes de g´en´eration, C. R. Acad. Sci. Paris S´er. I Math. 319 (1995), 1273 -- 1278. [10] G. Carron, T. Coulhon and E.-L. Ouhabaz, Gaussian estimates and Lp-boundedness of Riesz means, J. Evol. Equ. 2 (2002) 299-317. [11] J. M. C. Duhamel, M´emoire sur la m´ethode g´en´erale relative au mouvement de la chaleur dans les corps solides plong´es dans les milieux dont la temp´erature varie avec le temps. J. Ec. Polyt. Paris 14, Cah. 22, 20 (1833). [12] R.M. Dubois and G. Lumer, Formule de Duhamel abstraite, Arch. Math. 43 (1984) 49-56. [13] H.O. Fattorini, "The Cauchy Problem", Addison-Wesley Publishing Co., Reading, Mass., 1983. [14] M. Hieber: Integrated semigroups and differential operators on Lp(RN ), Math. Ann. 291 (1991), 1-16. [15] C. Kaiser and L. Weis, Perturbation theorems for α-times integrated semigroups, Arch. Math. 81 (2003) 215-228. [16] V. Keyantuo, Integrated semigroups and related partial differential equations, J. Math. Anal. Appl. 212(1997) 135-153. [17] V. Keyantuo, C. Muller and P. Vieten, The Hille-Yosida theorem for local convoluted semigroups. Proc. Edinb. Math. Soc. 46 (2003), 357-372. [18] V. Keyantuo, C. Lizama, P. Miana, Algebra homomorphisms defined via semigroups and cosine functions, J. Funct. Anal. 257 (2009) 3454-3487. [19] M. Kosti´c, "Generalized semigroups and cosine functions". Mathematical Institute, Belgrade, 2011. [20] M. Kosti´c and S. Pilipovi´c, Global convoluted semigroups. Math. Nachr. 15 (2007), 1727-1743. 24 VALENTIN KEYANTUO, PEDRO J. MIANA, AND LUIS S ´ANCHEZ-LAJUSTICIA [21] P. C. Kunstmann, Distribution semigroups and abstract Cauchy Problem. Trans. Amer. Math. Soc 351 (1999), 837 -- 856. [22] P. C. Kunstmann, M. Mijatovic and S. Pilipovic, Classes of distribution semigroups. Studia Math. 187 (2008), 37 -- 58. [23] C.C. Kuo and S.Y. Shaw, On α-times integrated C-semigroups and the abstract Cauchy problem. Studia Math. 142 (2000), 201 -- 217. [24] Y-C. Li and S-Y Shaw, On local α-times integrated C-semigroup, Abst. Appl. Anal. 207 (2007) 1-18, doi 10.1155/2007/34890. [25] J. L. Lions, Les semi-groupes distributions, Portugalia Math. 19 (1960) 141-164. [26] J. L. Lions and E. Magenes, "Non-homogeneous boundary value problems and applications". Vol. III., Die Grundlehren der mathematischen Wissenschaften, Band 183. Springer-Verlag, New York- Heidelberg, 1973. [27] Melnikova, I., and A. Filinkov: "Abstract Cauchy problems: three approaches", Chapman- Hall/CRC, New York, 2001. [28] P.J. Miana, α-Times integrated semigroups and fractional derivation, Forum Math. 14 (2002), 23-46. [29] P.J. Miana, Local and global solutions of well-posed integrated Cauchy problems. Studia Math. 187 (2008), 219-232. [30] P. J. Miana and V. Poblete, Sharp extensions for convoluted solutions of wave equations, Preprint, 2013. [31] S. G. Samko, A. A. Kilbas and O. I. Marichev: "Fractional Integrals and Derivatives. Theory and Applications". Gordon-Breach, New York (1993). [32] R. T. Seeley, Extension of C∞ functions defined in a half space. Proc. Amer. Math. Soc. 15 (1964), 625-626. [33] N. Tanaka, N. Okazawa, Local C-semigroups and local integrated semigroups. Proc. London Math. Soc. 61 (1990), 63-90. [34] S. Umarov, On fractional Duhamels principle and its applications, J. Differential Equations 252 (2012), 5217-5234. [35] S. W. Wang, Quasi-distribution semigroups and integrated semigroups. J. Funct. Anal. 146 (1997), 352-381. [36] S. W. Wang and M.C. Gao, Automatic extensions of local regularized semigroups and local regular- ized cosine funtions. Proc. Amer. Math. Soc. 127(6) (1999), 1651-1663. University of Puerto Rico, Department of Mathematics, Faculty of Natural Sciences, Box 70377, San Juan PR 00936-8377 U.S.A. E-mail address: [email protected] Universidad de Zaragoza, Departamento de Matem´aticas & I.U.M.A., 50.009, Zaragoza. E-mail address: [email protected] Universidad de Zaragoza, Departamento de Matem´aticas & I.U.M.A., 50.009, Zaragoza. E-mail address: [email protected]
1211.4329
1
1211
2012-11-19T08:18:40
Dimension free boundedness of Riesz transforms for the Grushin operator
[ "math.FA" ]
Let $G = - \Delta_{\xi} - |\xi|^2 \frac{\partial^2}{\partial \eta^2}$ be the Grushin operator on $\R^n \times \R.$ We prove that the Riesz transforms associated to this operator are bounded on $L^p (\R^{n+1}), 1 < p < \infty$ and their norms are independent of the dimension $n$.
math.FA
math
Dimension free boundedness of Riesz transforms for the Grushin operator P. K. Sanjay and S. Thangavelu Abstract. Let G = −∆ξ − ξ2 ∂ 2 ∂η2 be the Grushin operator on Rn × R. We prove that the Riesz transforms associated to this operator are bounded on Lp(Rn+1), 1 < p < ∞ and their norms are independent of the dimension n. 1. Introduction We consider the Grushin operator G = −∆ξ − ξ2 ∂2 ∂η2 on Rn × R which can be written formally as 2 1 0 2 v o N 9 1 ] . A F h t a m [ 1 v 9 2 3 4 . 1 1 2 1 : v i X r a where and where is the scaled Hermite operator on Rn and is the inverse Fourier transform of f in the η variable. In the light of this decomposition, we can define the Riesz transforms associated to the operator G as Gf (ξ, η) = 1 2πZR e−iληH(λ)f λ(ξ)dλ H(λ) = −∆ + λ2ξ2 f λ(ξ) =ZR f (ξ, η)eiληdη Rjf (ξ, η) =Z ∞ R∗j f (ξ, η) =Z ∞ −∞ −∞ e−iληRj(λ)f λ(ξ)dλ e−iληR∗j (λ)f λ(ξ)dλ for j = 1, 2, 3, . . . , n. Here, Rj(λ) = Aj(λ)H(λ)− 1 2 , R∗j (λ) = Aj(λ)∗H(λ)− 1 2 are the Riesz transforms associated to the Hermite operator which has the decompo- sition 1 2 H(λ) = Xj=1 Aj(λ) = − n (Aj(λ)Aj(λ)∗ + Aj(λ)∗Aj(λ)) ∂ ∂ξj + λξj, Aj(λ)∗ = ∂ ∂ξj + λξj 1991 Mathematics Subject Classification. 42C, 42C05 , 43A65. Key words and phrases. Riesz Transforms, Grushin operator, Hermite operator, Transference method. 1 2 P. K. SANJAY AND S. THANGAVELU are the creation and annihilation operators on Rn. The boundedness of these Riesz transforms on Lp(Rn+1) is known [JST, BG]. In [BG], the authors have proved the boundedness of Riesz transforms associated to a much larger class of smooth locally subelliptic diffusion operators on smooth connected non-compact manifold. In [JST] these Riesz transforms were treated as operator valued Fourier multipli- ers for Lp(Rn) valued functions on R and the boundedness was proved with the aid of a result of L. Weis [W] on the operator valued Fourier multipliers. The aim of this paper is to prove the following theorem about the dimension free bound- edness of the vector of Riesz transforms Rf . That is, we consider the operator R = (R1, R2,· · · , Rn, R∗1, R∗2,· · · , R∗n) with Rf (ξ, η) = n Xj=1 Rjf (ξ, η)2 + n Xj=1 and prove: R∗j f (ξ, η)2!1/2 Theorem 1.1. For each 1 < p < ∞, there exists a constant Cp independent of the dimension n, such that for all f ∈ Lp(Rn × R), kRfkp ≤ Cpkfkp. E. M. Stein introduced the notion of the dimension free boundedness for the vec- tor of Riesz transforms in [S1]. Stein considered the Euclidean Riesz transforms and proved the result using the technique of g-functions. Later this technique was adapted for a non-commutative situation by Harboure, et al. [HRST] to prove a similar re- sult for the Riesz transforms associated to the Hermite operator. Alternate proofs for the Euclidean Riesz transforms were provided using the method of rotations by J. Duoandikoetxea and J. L. Rubio de Francia [DR] and using a transference argu- ment by Gilles Pisier [P]. Similar techniques have been used to prove the dimension free boundedness of Riesz transforms in other contexts as well. Coulhon, Muller and Zienkiewicz [CMZ] proved the boundedness for the Riesz transforms associated to the sub-Laplacian on the Heisenberg group. In two works [LP2] and [LP1] Fran- coise Lust-Piquard provided alternate proofs for the Riesz transforms associated to the Hermite operator and the Riesz transforms associated to the sub-Laplacian. Our proof follows a method very similar to that of [CMZ],[LP1] and [LP2]. The method of rotations applied to the Euclidean Riesz transforms [DR] involves expressing the Riesz transforms as the average of certain directional Hilbert transforms and using the boundedness (independent of the direction) of these directional Hilbert transforms. An expression similar to this in the Heisenberg group context is obtained in [CMZ]. In this paper we obtain a similar representation for Riesz transforms associated to Grushin operators . We also note that Theorem 1.1 in turn implies that the Riesz transforms associated to the Hermite operator satisfy dimension free bounds. Since the Riesz transform for the Grushin operator is an operator valued multiplier for the Fourier transform on R, the corresponding multiplier operator for the Fourier series is also bounded by the same norm. This can be proved using a generalisation of a transference result of Karel de Leeuw [L]. The proof is similar to the proof in [ST] of the boundedness of the Riesz transforms on the reduced Heisenberg group using transference from the Heisenberg group. We refer to Theorem 2.1 and Theorem 2.2 of [ST] for details. Then the proof of the Hermite Riesz transforms follows by looking at functions of the form F (ξ, η) = f (ξ)eikη on Rn × [0, 2π). RIESZ TRANSFORMS 3 2. Riesz transforms for the Grushin operator We first consider the individual Riesz transforms Rj and R∗j for j = 1, 2,· · · , n and show that they satisfy dimension free bounds on Lp(Rn+1) for 1 < p < ∞. In order to j and R∗ǫ do this we introduce the operators Rǫ j which we will call the truncated Riesz transforms. We only give details of Rǫ j as the other one is similar. Note that the Riesz transform Rj(λ) associated to the Hermite operator H(λ) can be written as Rj(λ) = Aj(λ)H(λ)−1/2 = e−rH(λ)r−1/2dr. Aj(λ) √π Z ∞ 0 Here e−rH(λ) is the Hermite semigroup. For ǫ > 0, we define the truncated Riesz transforms Rǫ j(λ) by Rǫ j(λ) = e−rH(λ)r−1/2dr. Then the truncated Riesz transforms Rǫ j for the Grushin operator are defined as Aj(λ) ǫ2 √π Z 1/ǫ2 2πZ ∞ −∞ 1 Rǫ jf (ξ, η) = We first prove e−iληRǫ j(λ)f λ(ξ)dλ. Proposition 2.1. For every f ∈ L2(Rn+1), Rǫ Proof. It follows from the definition that jf (ξ, η) − Rjf (ξ, η)2dξ dη =ZR(cid:18)ZRn Rǫ ZRZRn Rǫ The proposition follows once we show that for every λ ∈ R∗ jf → Rjf in L2(Rn+1) as ǫ → 0. j(λ)f λ(ξ) − Rj(λ)f λ(ξ)2dξ(cid:19) dλ. and ZRn Rǫ ZRn Rǫ j(λ)f λ(ξ) − Rj(λ)f λ(ξ)2dξ → 0 as ǫ → 0 j(λ)f λ(ξ) − Rj(λ)f λ(ξ)2dξ ≤ 4ZRn f λ(ξ)2dξ. To see these, we expand f λ in terms of scaled Hermite functions Φλ definition) and use the fact that α (see [T] for Aj(λ)Φλ α = (2αj + 2)1/2λ1/2Φλ α+ej . where ej is the canonical unit vector of Rn with 1 in the jth entry and zero elsewhere. Then Rǫ (2αj + 2)1/2λ1/2 (cid:18)ZAǫ j(λ)f λ(ξ)−Rj(λ)f λ(ξ) = Xα∈Nn where Aǫ = (0, ǫ2) ∪ (1/ǫ2,∞). From this, it follows that 2α + n(f λ, Φλ j(λ)f λ − Rj(λ)f λk2 2αj + 2 kRǫ √π 2 = 1 π Xα∈Nn e−(2α+n)λrr−1/2dr(cid:19) (f λ, Φλ α)Φλ α+ej . α)2 ZAǫ,α e−rr−1/2dr!2 . where Aǫ,α = (0, (2α + n)λǫ2) ∪ ((2α + n)λǫ−2,∞). From the above equation it is clear that kRǫ j(λ)f λ − Rj(λ)f λk2 ≤ 2kf λk2 4 P. K. SANJAY AND S. THANGAVELU We also note that when f λ(ξ) = X(2α+n)λ≤N is a finite linear combination of Φλ α, (f λ, Φλ α)Φλ α(ξ) kRǫ j(λ)f λ − Rj(λ)f λk2 α)2(cid:18)Z(0,N ǫ2)∪(N ǫ−2,∞) which goes to 0 as ǫ → 0. As such functions are dense in L2(Rn), we get π X(2α+n)λ≤N (f λ, Φλ 2 ≤ 1 e−rr−1/2dr(cid:19)2 kRǫ j(λ)f λ − Rj(λ)f λk2 → 0, as ǫ → 0 for any f ∈ L2(Rn+1). For the individual Riesz transforms, we have the following result: Theorem 2.2. For j = 1, 2, 3,· · · , n we have kRjfkp + kR∗j fkp ≤ Cpkfkp, 1 < p < ∞ for all f ∈ Lp(Rn+1) where Cp is independent of the dimension. In order to prove this result, we claim that it is enough to prove (cid:3) kRǫ jfkp + kR∗ǫ j fkp ≤ Cpkfkp for the truncated Riesz transforms. A proof of this will be given in section 5 where jf(cid:1) is Cauchy in Lp(Rn+1) and hence there exists an operator we will show that (cid:0)Rǫ Sj, bounded on Lp(Rn+1) such that Rǫ jf → Sjf as ǫ → 0. In view of Proposition 2.1, Rǫ jf → Rjf for f ∈ L2(Rn+1) and hence Sjf = Rjf . This will prove the stated boundedness of Rj ( and R∗j ). Now to prove the boundedness of Rǫ j and R∗ǫ j we express these operators as a superposition of certain truncated Hilbert transforms. 3. A representation for the truncated Riesz transforms The representation we obtain is very similar to the representation obtained in [CMZ] for the Riesz transforms on the Heisenberg group Hn. Before stating this result, we recall some definitions and notation. For further details and proofs we refer to [T]. Recall that as a manifold Hn = Cn × R and hence we write (z, t), z = x + iy ∈ Cn, t ∈ R to denote the elements of Hn. The sub-Laplacian L on the Heisenberg group Hn can be written as the sum of the differential operators Xj =(cid:16) ∂ ∂t(cid:17) and Yj =(cid:16) ∂ j ). The sub-Laplacian is homogeneous of degree 2 with respect to the non-isotropic dilations δr(z, t) = (rz, r2t) of the Heisenberg group. Hence ps(z, t), the heat kernel associated to L satisfies ∂t(cid:17) as L =Pn ∂yj − 1 j=1(X 2 j + Y 2 + 1 2xj 2 yj ∂xj ∂ ∂ pr2s(z, t) = r−(2n+2)ps(z/r, t/r2). This also follows from the following explicit expression for qs(z, λ), the inverse Fourier transform of ps(z, t) in the t variable, qs(z, λ) =ZR ps(z, t)eiλtdt = (4π)−n(cid:18) λ sinh λs(cid:19)n e− 1 4 λ coth(sλ)z2 . RIESZ TRANSFORMS 5 By πλ, we denote the Schrodinger representation of the Heisenberg group Hn acting on L2(Rn) in the following manner πλ(x + iy, t)φ(ξ) = eiλteiλ(x.ξ+ 1 2 x.y)φ(ξ + y). When t = 0, we will denote πλ(z, t) by πλ(z). πλ also defines a representation of the group algebra L1(Hn) as πλ(f ) =ZHn f (z, t)πλ(z, t)dz dt = Wλ(f λ) where Wλ(g) = RCn g(z)πλ(z)dz is the Weyl transform of g. The representation given in [CMZ] for the Riesz transforms associated to the sub-Laplacian L is Xip1(w, s)H(w,s)f (z, t)dw ds 1 (XiL− 1 2 f )(z, t) = − 4(2π)n+1 ZHn where H(w,s) is the Hilbert transform along a curve in the Heisenberg group. For the Grushin operator, we obtain a similar representation involving certain operators T (z,t) and certain differential operators ǫ with Zj = i Xj + Yj; Z∗j = i Xj − Yj xj 2 ∂t(cid:19) ; Yj =(cid:18) ∂ ∂xj − yj 2 ∂yj + ∂ Xj =(cid:18) ∂ ∂ ∂t(cid:19) . Proposition 3.1. Rǫ jf (ξ, η) = R∗ǫ j f (ξ, η) = f (ξ, η) =Zǫ<r<1/ǫ 1 1 f (ξ, η) Zjp1(z, t)dz dt T (z,t) ǫ √π ZHn √π ZHn f(cid:16)ξ + ry, η + rx · ξ + r2(cid:16)t + T (z,t) ǫ f (ξ, η) Z∗j p1(z, t)dz dt x · y 2 (cid:17)(cid:17) dr r and where T (z,t) ǫ We follow the method in [LP2] in proving this proposition. First we prove the following lemma which will be used in the proof of Proposition 3.1. Lemma 3.2. For any φ ∈ S(Rn), (i) r ∂ ∂ξj (πλ(rz)φ(ξ)) =(cid:16) ∂ (ii) irλξjπλ(rz)φ(ξ) =(cid:16) ∂ + iλr2xj ∂xj − iλr2yj ∂yj 2 (cid:17) (πλ(rz)φ(ξ)) 2 (cid:17) πλ(rz)φ(ξ). Proof. We first look at the derivative of πλ(rz) with respect to the variable yj and see that ∂ ∂yj(cid:0)πλ(rz)φ(ξ)(cid:1) = ∂ ∂yj (cid:16)eiλ(rx·ξ+ 1 2 r2x·y)φ(ξ + ry)(cid:17) πλ(rz)φ(ξ) + eiλ(rx·ξ+ 1 = iλr2xj 2 2 r2x·y) ∂ ∂yj φ(ξ + ry) = iλr2xj 2 πλ(rz)φ(ξ) + rπλ(rz) ∂φ ∂ξj (ξ). 6 Writing πλ(rz) and noting that P. K. SANJAY AND S. THANGAVELU ∂φ ∂ξj (ξ) =(cid:18)πλ(rz) ∂ ∂ξj − ∂ ∂ξj πλ(rz)(cid:19) φ(ξ) + ∂ ∂ξj πλ(rz)φ(ξ) we see that ∂ ∂φ ∂ξj (ξ) ∂ ∂ξj(cid:0)πλ(rz)φ(ξ)(cid:1) = iλrxjπλ(rz)φ(ξ) + πλ(rz) ∂yj(cid:0)πλ(rz)φ(ξ)(cid:1) =(cid:18)r ∂xj(cid:0)πλ(rz)φ(ξ)(cid:1) = iλrξjπλ(rz)φ(ξ) + ∂ ∂ξj − iλr2xj iλr2yj ∂ 2 2 (cid:19)(cid:0)πλ(rz)φ(ξ)(cid:1) πλ(rz)φ(ξ) which proves part (i). Part (ii) of the lemma follows from (cid:3) Since Rǫ Now using the previous lemma, Proof of Proposition 3.1. The heat kernels for the sub-Laplacian and the Her- mite operator are related via the group Fourier transform on the Heisenberg group as follows: ps(z, t)πλ(z, t)dz dt = e−sH(λ) We refer to [T] for a proof of this. Using the homogeneity of the heat kernel we get ps(λ) =ZHn e−r2H(λ) = r−(2n+2)ZHn = r−2nZCn r ∂ξj Rǫ R∗ǫ 1 1 t , r ∂ ∂ξj j(λ) = j (λ) = r2(cid:19) dz dt ,−λr2(cid:17) dz =ZCn πλ(rz)q1(cid:0)z,−λr2(cid:1) dz. j(λ) = 1√π Aj(λ)Rǫ<r<1/ǫ e−r2H(λ)dr, we see that + λξj(cid:19) πλ(rz)q1(cid:0)z,−λr2(cid:1) dz dr; + λξj(cid:19) πλ(rz)q1(cid:0)z,−λr2(cid:1) dz dr. πλ(z, t)p1(cid:18)z πλ(z)q1(cid:16)z √πZǫ<r<1/ǫZCn(cid:18)− √π Zǫ<r<1/ǫZCn(cid:18) ∂ λξjπλ(rz)q1(cid:16)z ,−λr2(cid:17) dz 2 (cid:19) πλ(rz)q1(z,−λr2)dz ∂xj − 2 (cid:19) q1(z,−λr2)dz πλ(rz)(cid:18) ∂ 2 (cid:19)(cid:18)ZR πλ(rz)(cid:18) ∂ πλ(rz)(cid:18)ZR(cid:18) ∂ p1(z, t)eiλr2tdt(cid:19) dz ∂t(cid:19) p1(z, t)eiλr2tdt(cid:19) dz irZCn = ZCn(cid:18) ∂ = −ZCn = −ZCn = −ZCn = −ZHn πλ(rz) Xjp1(z, t)eiλr2tdz dt. ∂xj − iλr2yj iλr2yj iλr2yj ∂ yj 2 ∂xj ∂xj + + r RIESZ TRANSFORMS 7 ∂ξj rZCn(cid:18) ∂ = ZCn(cid:18) ∂ = −ZHn ∂yj πλ(rz)(cid:19) q1(cid:16) z r iλr2xj + ,−λr2(cid:17) dz 2 (cid:19) πλ(rz)q1(z,−λr2)dz πλ(rz) Yjp1(z, t)eiλr2tdz dt. πλ(rz) Zjp1(z, t)eiλr2tdz dt πλ(rz) Z∗j p1(z, t)eiλr2tdz dt dr r dr r . Rǫ R∗ǫ j (λ) = 1 1 j(λ) = √π Zǫ<r<1/ǫZHn √πZǫ<r<1/ǫZHn √π ZR(cid:18)Zǫ<r<1/ǫZHn 1 Similarly, Hence and Thus Rǫ jf (ξ, η) = πλ(rz)f λ(ξ) Zjp1(z, t)eiλr2tdz dt dr r (cid:19) eiληdλ. Now the proposition follows from the fact that ZR πλ(rz)f λ(ξ)eiλr2teiληdλ = f(cid:16)ξ + ry, η + rx · ξ + r2(cid:16)t + x · y 2 (cid:17)(cid:17) . (cid:3) 4. A transference result In the previous section, we obtained a representation for the Riesz transforms as the superposition of certain operators T (z,t) . To prove the boundedness of the vector of Riesz transforms, using the method of rotations, we need to prove that these operators are bounded uniformly in (z, t). ǫ Proposition 4.1. For 1 < p < ∞, there exists a constant Cp independent of (z, t) ∈ Hn, ǫ > 0 and the dimension n such that, ǫ (cid:13)(cid:13)T (z,t) f(cid:13)(cid:13)Lp(Rn+1) ≤ Cp kfkLp(Rn+1) . Proof. This will be proved using Calder´on's method of transferring an operator on Lp(X), for a measure space X, to Lp(G) when the group G acts on X by measure preserving transformations. The measure space Rn+1, the group Hn and the group action U(x + iy, t)(ξ, η) = (ξ − y, η − t + x · y 2 − x.ξ) are all the same as used by Ratnakumar and Thangavelu in [RT] and we follow their notation. Accordingly, for f ∈ Lp(Rn+1) and (ξ, η) ∈ Rn+1, we define the transferred function F(ξ,η) on Hn as F(ξ,η)(z, t) = f (U(z, t)(ξ, η)) . For T ∈ B(Lp(Hn)), the transferred operator T0 on B(Lp(Rn+1)) is defined as T0f (ξ, η) = (T F(ξ,η))(0). 8 P. K. SANJAY AND S. THANGAVELU For a curve γ = {γ(t) ∈ Hn, t ∈ R} and a function f on Hn, the Hilbert transform of f along γ is defined as Hγf (w, s) =ZR f(cid:0)(w, s)γ(r)−1(cid:1) dr r . When γ is the curve (rz, r2t), we will denote Hγ by H(z,t). By H ǫ (z,t) we denote the truncated Hilbert transform where the integration is only over {r ∈ R, ǫ < r < 1/ǫ}. (z,t) to an operator on Lp(Rn+1) under the U action When we transfer this operator H ǫ defined above, we get (H ǫ (z,t))0f (ξ, η) = H ǫ (z,t)F(ξ,η)(0) =Zǫ<r<1/ǫ dr F(ξ,η)(−rz,−r2t) r x · y 2 (cid:17)(cid:17) dr f(cid:16)ξ + ry, η + r(x · ξ) + r2(cid:16)t + = Zǫ<r<1/ǫ r . Hence we see that, (H ǫ (z,t))0 = T (z,t) ǫ . Now our aim is to prove kT (z,t) ǫ fkLp(Rn+1) ≤ CpkfkLp(Rn+1) This can be obtained using the transference method from the uniform boundedness of the operator H ǫ (z,t). To be precise, we use the fact that there exist a finite constant Cp, independent of (z, t) ∈ Hn, n and ǫ > 0 such that kH ǫ (z,t)fkp ≤ Cpkfkp. A proof of the above fact is indicated in [CMZ]. We will present it at the end of this section after proving the boundedness of T (z,t) . Then ǫ ZRn+1T (z,t) ǫ f (ξ, η)pdξ dη = ZRn+1 H ǫ = ZRn+1 H ǫ (z,t)F(ξ,η)(0)pdξ dη (z,t)FU (w,s)(ξ,η)(0)pdξ dη because of the invariance of the measure dξ dη under the U action. Hence ZRn+1T (z,t) ǫ f (ξ, η)pdξ dη = = 1 BR(0)ZBR(0)ZRn+1H ǫ BR(0)ZBR(0)ZRn+1H ǫ 1 (z,t)FU (w,s)(ξ,η)(0)pdξ dη dw ds (z,t)F(ξ,η)(w, s)pdξ dη dw ds where BR(0) is the ball centred at origin in the Heisenberg group and radius R under the Koranyi norm k(w, s)k = (w4 + s2)1/4 . When (w, s) ∈ BR(0), and ǫ < r < 1/ǫ, (w, s) (δr(z, t))−1 ∈ BR+ h (0) where h is the Koranyi norm of (z, t). Hence, in the above equality F(ξ,η) can be replaced by F(ξ,η) = F(ξ,η) · χBR+h/ǫ(0). Now by an application of Fubini, we get ǫ kT (z,t) ǫ fkp p ≤ 1 BR(0)ZRn+1ZHnH ǫ (z,t) F(ξ,η)(w, s)pdw ds dξ dη Now, from the uniform boundedness of the truncated Hilbert transforms, we get, RIESZ TRANSFORMS 9 kT (z,t) ǫ fkp p ≤ Cp = Cp 1 1 BR(0)ZRn+1ZHn F(ξ,η)(w, s)pdw ds dξ dη BR(0)ZBR+h/ǫ(0)ZRn+1 f (U(w, s)(ξ, η))pdξdηdwds BR(0) kfkp p = Cp(cid:18)R + h/ǫ R (cid:19)2n+2 kfkp p = CpBR+h/ǫ(0) again by the invariance dξdη under the U action. Letting R → ∞, we see that kT (z,t) ǫ fkLp(Rn) ≤ CpkfkLp(Rn) Coming back to the boundedness of H ǫ to reduce this to the boundedness of H ǫ Recall that (z,t), we can use a technique in Lemma 3.1 of [S2], γ on Lp(R2) for the curve γ = {(t, t2), t ∈ R}. H(z,t)f (w, s) =ZR f (w − rz, s − r2t − r Im (w.¯z)) dr r . For σ ∈ U(n), define ρ(σ)f (w, s) = f (σw, s). Then ρ(σ−1)H(z,t)(ρ(σ)f )(w, s) = H(z,t)(ρ(σ)f )(σ−1w, s) H(z,t)(ρ(σ)f )(w, s) = ZR = ZR = ZR f (σw − rσz, s − r2t − r Im (w.¯z)) dr r . f (w − rσz, s − r2t − r Im (σ−1w.¯z)) dr r f (w − rσz, s − t − r Im (w. ¯σz)) . dr r . Since kρ(σ)fkp = kfkp and since there exists σ ∈ U(n) such that ρ(σ)(x + iy) = x1e1, it is enough to consider the operator f →ZR f (u1 − rx1, w′, s − r2t − rv1x1) dr r which is equivalent to the operator T(x,t,v)f (u, s) =ZR f (u − rx, s − r2t − rvx) dr r acting on functions defined on R2. For λ1, λ2 > 0, let δλ1,λ2f (u, s) = f (λ1u, λ2s). Then kδλ1,λ2fkp = (λ1λ2)−1/pkfkp and δ1/λ1,1/λ2T(x,t,v)δλ1,λ2f (u, s) =ZR f (u − rλ1x, s − r2λ2t − rλ2vx) dr r Since we can choose λ1, λ2 such that λ1x = λ2t = 1 it is enough to get uniform esti- mates for operators of the form T(1,1,a). That would imply that 10 P. K. SANJAY AND S. THANGAVELU kT(x,t,v)fkp = kδλ1,λ2T(1,1,vx/t)δ1/λ1,1/λ2fkp = (λ1λ2)−1/pkT(1,1,vx/t)δ1/λ1,1/λ2fkp ≤ (λ1λ2)−1/pCpkδ1/λ1,1/λ2fkp = (λ1λ2)−1/pCp(λ1λ2)1/pkfkp. Now consider τaf (u, s) = f (u, s + au) so that kτafkp = kfkp. Since a f )(u − r, s − r2 − ar) (τ−1 f (u − r, s − au − r2) a f )(u, s) =ZR =ZR T(1,1,a)(τ−1 dr r , dr r we have τaT(1,1,a)(τ−1 a f )(u, s) = T(1,1,a)(τ−1 a f )(u, s + au) =ZR f (u − r, s − r2) dr r . This is the Hilbert transform along the parabola γ(r) = (r, r2) in R2. Since γ(r) = δr(1, 1) 0  δ−r(−1, 1) if r > 0 if r = 0 if r < 0 and the linear space spanned by {γ(r)}r>0 and the linear space spanned by {γ(r)}r<0 are the same, namely R2, γ(r) is a two-sided homogeneous curve (see Def 3.1 in [SW, p. 1261]). Now we can appeal to Theorem 11 of [SW, p. 1271] to get the (z,t) on Lp(Hn). That is, uniform boundedness of the truncated Hilbert transforms H ǫ there exist a finite constant Cp, independent of (z, t) ∈ Hn, n and ǫ > 0 such that kH ǫ (z,t)fkp ≤ Cpkfkp. (cid:3) 5. Proof of the main theorem We first prove Theorem 2.2 regarding the boundedness of the individual Riesz transforms Rj and R∗j . Proof of Theorem 2.2. We first consider the operators Rǫ j . In the light of Proposition 3.1 and Proposition 4.1, we just need to prove that there exist a finite C, independent of n such that j and R∗ǫ As mentioned in Lemma 3 of [CMZ], this follows from the fact that k Zjp1(z, t)kL1(Hn),k Z∗j p1(z, t)kL1(Hn) ≤ C. where pn r is the heat kernel on Hn. Now for the operators Rj and R∗j , we need only r(z1,·) ∗ p1 pn r (z1, z2,· · · , zn, t) = p1 jf(cid:1) , (cid:0)R∗ǫ fkp and so it is enough to prove that T (z,t) j f(cid:1) are Cauchy in Lp(Rn+1). We have already seen that f is Cauchy in Lp(Rn+1). r(z2,·),∗ · · · ,∗p1 r(zn,·)(t) ǫ to show that (cid:0)Rǫ jfkp ≤ CkT (z,t) kRǫ ǫ RIESZ TRANSFORMS 11 During the course of the proof of Proposition 4.1, we had observed that the operator T (z,t) ǫ is obtained by applying a transference on H ǫ fkLp(Rn+1) ≤ kH ǫ kT (z,t) (z,t) and that (z,t)fkLp(Hn). ǫ Being truncated Hilbert transforms H ǫ and consequently, so is T (z,t) f . ǫ (z,t)f is Cauchy in Lp(Hn) [SW, Theorem 11], (cid:3) Now we complete the proof of our main theorem. As mentioned during the dis- cussion of Theorem 2.2, we only need to prove the boundedness of the operator Rǫ obtained by replacing Rj and R∗j by their truncated versions. We will be closely following [LP1] in proving this and so will skip some details. Using the property 1 r ∂p1 ∂r = 1 xj ∂p1 ∂xj = 1 yj ∂p1 ∂yj + 2√πZHn (xj − yj)T (z,t) ǫ f (ξ, η) ∂p1 ∂t (z, t)dzdt 1 r 1 1 r 1 where r = z, we can rewrite Rǫ jf (ξ, η) = 1 √π ZHn (ixj + yj)T (z,t) ǫ f (ξ, η) ∂p1 ∂r (z, t)dzdt and R∗ǫ j f (ξ, η) = 1 √πZHn (ixj − yj)T (z,t) ǫ f (ξ, η) ∂p1 ∂r (z, t)dzdt − 2√π ZHn (xj + yj)T (z,t) ǫ f (ξ, η) ∂p1 ∂t (z, t)dzdt For a fixed (ξ, η) we can choose λ1, λ2, . . . , λ2n such that P2n n jf (ξ, η) + λj+nR∗ǫ Rǫf (ξ, η) = Xj=1(cid:16)λjRǫ j=1 λj2 = 1 and j f (ξ, η)(cid:17) Now using the triangle inequality, Holder's inequality and also Lemma 2(a) of [LP1], we get Rǫf (ξ, η) ≤ CkT (z,t) f (ξ, η)kLp( 1 ∂p1 ǫ r ∂r dzdt)kx1kLp′ ( 1 + CkT (z,t) ǫ r ∂p1 ∂r dzdt) f (ξ, η)kLp( ∂p1 ∂t dzdt)kx1kLp′ ( ∂p1 ∂t dzdt) where C is a universal constant. Now the theorem follows from Proposition 4.1 and an application of Minkowski inequality along with the fact (proved in [LP1])that ZHn x1p 1 r ∂p1 ∂r dz dt,ZHn x1p ∂p1 ∂t where Ap is independent of the dimension. dz dt ≤ Ap, p ≥ 0, Acknowledgments The authors are thankful to the referee for pointing out some errors and typos. The work of the first author is supported by the All India Council for Technical Education (AICTE). The work of the second author is supported by J. C. Bose Fellowship from the Department of Science and Technology (DST) and also by a grant from UGC via DSA-SAP. 12 P. K. SANJAY AND S. THANGAVELU References [BG] F. Baudoin and N. Garofalo, A note on the boundedness of Riesz transform for some subel- liptic operators, International Mathematics Research Notices (to appear); Arxiv preprint arXiv:1105.0467 (2012). [CMZ] T. Coulhon, D. Muller, and J. Zienkiewicz, About Riesz transforms on the Heisenberg groups, Math. Ann. 305 (1996), no. 2, 369 -- 379. MR1391221 (97f:22015) [L] Karel de Leeuw, On Lp multipliers, Ann. of Math. (2) 81 (1965), 364 -- 379. MR0174937 (30 #5127) [DR] J. Duoandikoetxea and Jos´e L. Rubio de Francia, Estimations ind´ependantes de la dimension pour les transform´ees de Riesz, C. R. Acad. Sci. Paris S´er. I Math. 300 (1985), no. 7, 193 -- 196. MR780616 (86e:42028) [HRST] E. Harboure, L. de Rosa, C. Segovia, and J. L. Torrea, Lp-dimension free boundedness for Riesz transforms associated to Hermite functions, Math. Ann. 328 (2004), no. 4, 653 -- 682. MR2047645 (2006a:42017) [JST] K. Jotsaroop, P K Sanjay, and S. Thangavelu, Riesz transforms and multipliers for the Grushin operator, Journal d'Analyse Mathematique (to appear); Arxiv preprint arXiv:1105.3227 (2011). [LP1] F. Lust-Piquard, Riesz transforms on generalized Heisenberg groups and Riesz transforms associated to the CCR heat flow, Publ. Mat. 48 (2004), no. 2, 309 -- 333. MR2091008 (2005g:43014) [LP2] , Dimension free estimates for Riesz transforms associated to the harmonic oscillator on Rn, Potential Anal. 24 (2006), no. 1, 47 -- 62. MR2218202 (2006k:42012) [P] G. Pisier, Riesz transforms: a simpler analytic proof of P.-A. Meyer's inequality, S´eminaire de Probabilit´es, XXII, 1988, pp. 485 -- 501. MR960544 (89m:60178) [RT] P. K. Ratnakumar and S. Thangavelu, Spherical means, wave equations, and Hermite- Laguerre expansions, J. Funct. Anal. 154 (1998), no. 2, 253 -- 290. MR1612697 (99h:33047) [S1] E. M. Stein, Some results in harmonic analysis in Rn, for n → ∞, Bull. Amer. Math. Soc. [S2] R. S. Strichartz, Lp harmonic analysis and Radon transforms on the Heisenberg group, J. (N.S.) 9 (1983), no. 1, 71 -- 73. MR699317 (84g:42019) Funct. Anal. 96 (1991), no. 2, 350 -- 406. MR1101262 (92d:22015) [ST] P. K Sanjay and S. Thangavelu, Revisiting Riesz transforms on Heisenberg groups, Revista Mathem´atica Iberoamericana 28 (2012), no. 4, 1091 -- 1108, available at arXiv:1110.3236. [SW] E. M. Stein and S. Wainger, Problems in harmonic analysis related to curvature, Bull. Amer. Math. Soc. 84 (1978), no. 6, 1239 -- 1295. MR508453 (80k:42023) [T] S. Thangavelu, An introduction to the uncertainty principle, Progress in Mathematics, vol. 217, Birkhauser Boston Inc., Boston, MA, 2004. Hardy's theorem on Lie groups, With a foreword by Gerald B. Folland. MR2008480 (2004j:43007) [W] L. Weis, Operator-valued Fourier multiplier theorems and maximal Lp-regularity, Math. Ann. 319 (2001), no. 4, 735 -- 758. MR1825406 (2002c:42016) Department of Mathematics, Indian Institute of Science, Bangalore-560 012. E-mail address, Sanjay P. K.: [email protected] E-mail address, S. Thangavelu: [email protected] Permanent address(Sanjay P. K.):Department of Mathematics, National Institute of Technology, Calicut- 673 601.
1111.6812
4
1111
2012-09-25T14:46:48
Atomic representations in function spaces and applications to pointwise multipliers and diffeomorphisms, a new approach
[ "math.FA" ]
In Chapter 4 of [25] Triebel proved two theorems concerning pointwise multipliers and diffeomorphisms in function spaces of Besov and Triebel-Lizorkin type. In each case he presented two approaches, one via atoms and one via local means. While the approach via atoms was very satisfactory concerning the length and simplicity, only the rather technical approach via local means proved the theorems in full generality. In this paper we generalize two extensions of these atomic decompositions, one by Skrzypczak (see [22]) and one by Triebel and Winkelvoss (see [30]) so that we are able to give a short proof using atomic representations getting an even more general result than in the two theorems in [25]. References: [22] L. Skrzypczak. Atomic decompositions on manifolds with bounded geometry. Forum Math., 10(1):19-38, 1998. [25] H. Triebel. Theory of Function Spaces II. Birkh\"auser, Basel, 1992. [30] H. Triebel and H. Winkelvo{\ss}. Intrinsic atomic characterizations of function spaces on domains. Math. Z., 221(1):647-673, 1996.
math.FA
math
ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH BENJAMIN SCHARF ABSTRACT. In Chapter 4 of [28] Triebel proved two theorems concerning pointwise multipliers and diffeomor- phisms in function spaces Bs p,q(Rn). In each case he presented two approaches, one via atoms and one via local means. While the approach via atoms was very satisfactory concerning the length and simplicity, only the rather technical approach via local means proved the theorems in full generality. p,q(Rn) and Fs In this paper we generalize two extensions of these atomic decompositions, one by Skrzypczak (see [25]) and one by Triebel and Winkelvoss (see [33]) so that we are able to give a short proof using atomic representations getting an even more general result than in the two theorems in [28]. INTRODUCTION The aim of this paper is to generalize the atomic decomposition theorem from Triebel [28, 29] for Besov p,q(Rn) and to present two applications to pointwise multipliers p,q(Rn). For a detailed (historical) p,q(Rn) we refer to Triebel [27, 28], for an introduction to atoms we and Triebel-Lizorkin spaces Bs and diffeomorphisms as continuous linear operators in Bs treatment of the spaces Bs refer to Frazier and Jawerth [5, 6]. p,q(Rn) resp. F s p,q(Rn) and Fs p,q(Rn) and Fs According to Triebel [28] Pj : f 7→ j · f maps Bs vector function j p,q(Rn) into Bs : Rn → Rn p,q(Rn) if s > s p and j ∈ Ck(Rn) with k > s. Furthermore, the superposition with a Dj : f 7→ f ◦ j is a k-diffeomorphism and k is large enough in dependence of s and p. p,q(Rn) if j maps Bs There are similar results for F s p,q(Rn) to Bs p,q(Rn). The main idea for an easy proof is the atomic decomposition theorem. Mainly one has to show that a multiplication of an atom an ,m with a function j is still an atom with similar properties. But there was one problem: If s ≤ s p resp. s ≤ s p,q, then atoms need to fulfil moment conditions, i.e. the superposition with j resp. (1) for L ∈ N0 and L > s p − s resp. L > s p,q − s. But these properties are not preserved by multiplication resp. superposition. By Skrzypczak [25] these moment conditions were replaced by the more general assumptions x a(x) dx = 0 if b ≤ L − 1 b ZRn (cid:12)(cid:12)(cid:12)(cid:12) Zd·Qn ,m y (x)a(x) dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ C · 2−n (cid:16)s+L+n(cid:16)1− 1 p(cid:17)(cid:17)ky CL(Rn)k for all y ∈ CL(Rn). Now the situation changes: These conditions remain true after multiplication resp. superposition. This replacement is typical when thinking of atomic, in particular wavelet representations as repre- sentations of functions not mapping from Rn, but from more general manifolds, see the remarks on the cancellation property in [3, Section 3.1]. 2010 Mathematics Subject Classification. 46E35. Key words and phrases. Atomic decompositions, pointwise multipliers, diffeomorphisms, Sobolev spaces, Besov and Lizorkin- Triebel spaces. Benjamin Scharf: Mathematisches Institut, Fakultät für Mathematik und Informatik, Friedrich-Schiller-Universität Jena, D-07737 Jena, Germany, [email protected]. 1 2 BENJAMIN SCHARF In this paper we go a step further. We show that one can replace the usual CK(Rn)-conditions on atoms by Hölder-conditions (C K(Rn)-spaces) in the following way: A function a : Rn → C is called (s, p)K,L-atom located at Qn ,m if supp a ⊂ d · Qn ,m ka(2−n ·)C K(Rn)k ≤ C · 2−n (s− n p ) and for every y ∈ C L(Rn) it holds (cid:12)(cid:12)(cid:12)(cid:12) Zd·Qn ,m y (x)a(x) dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ C · 2−n (cid:16)s+L+n(cid:16)1− 1 p(cid:17)(cid:17)ky C L(Rn)k. This generalizes the known definitions of atoms from Triebel, Skrzypczak and Winkelvoss [29, 25, 33]. Furthermore, there is an existing theory generalizing the conditions ka(2−n p,p(Rn)k with K > s, mainly in connection with spline representations. For instance, see the books by Kahane and Lemarie-Rieusset [11, part II, Section 6.5], Triebel [31, Section 2.2] and the recent paper by Schneider and Vybiral [22]. Of these, only the first book incorporates the usual moment conditions as in (1). ·)C K(Rn)k by ka(·)BK In Section 3, as corollaries of the atomic representation theorem with these more general atoms from Section 2 we are able to extend the key theorems on pointwise multipliers and diffeomorphisms from [28]. It is not the aim of our observations to give best conditions or even exact characterizations for pointwise p,q(Rn). For this we refer to Strichartz [26], Peetre [18] as multipliers in function spaces Bs well as to Maz'ya and Shaposhnikova [15, 16] for the classical Sobolev spaces, while for Bs p,p(Rn) we refer to Franke [4], Frazier and Jawerth [6], Netrusov [17], Koch, Runst and Sickel [19, 23, 24, 12] as well as to Triebel [31, Section 2.3.3] for general function spaces Bs We obtain for pointwise multipliers with respect to Bs Let 0 < p ≤ ¥ and r > max(s,s p − s). Then there exists a positive number c such that p,q(Rn) and Fs p,q(Rn): p,q(Rn) and F s p,q(Rn). for all j ∈ C r (Rn) and all f ∈ Bs f Bs p,q(Rn). kj p,q(Rn)k ≤ ckj C r (Rn)k · k f Bs p,q(Rn)k For further sufficient results on diffeomorphisms including characterizations for classical Sobolev spaces W k p (Rn) we refer to Gol'dshtein, Reshetnyak, Romanov, Ukhlov and Vodop'yanov [8, 9, 10, 35, 34], [7, p,q(Rn) Chapter 4], Markina [13] as well as to Maz'ya and Shaposhnikova [14, 15], while for Besov spaces Bs with 0 < s < 1 we refer to Vodop'yanov, Bordaud and Sickel [35, 1]. A special case of our result (Lipschitz diffeomorphisms) can be found in Triebel [30, Section 4]. We will prove (in case of Bs Let 0 < p ≤ ¥ , r ≥ 1 and r > max(s,1 +s p − s). If j p,q(Rn)): is a r -diffeomorphism, then there exists a constant c such that k f (j (·))Bs p,q(Rn)k ≤ c · k f Bs p,q(Rn)k. for all f ∈ Bs p,q(Rn). Hence Dj maps Bs p,q(Rn) onto Bs p,q(Rn). Furthermore, at the end of section 2 we are able to give a simple proof of a local mean theorem very similar to Triebel's result [32, Theorem 1.15], which paved the way for the wavelet characterization of p,q(Rn) and F s Bs p,q(Rn) - where we are also using the more general Hölder-space conditions. 1. PRELIMINARIES Let Rn be the euclidean n-space, Z be the set of integers,N be the set of natural numbers and N0 = N ∪ {0}. By x we denote the usual euclidean norm of x ∈ Rn, by kxXk the (quasi)-norm of an element x of a (quasi)-Banach space X. By S (Rn) we mean the Schwartz space on Rn, by S ′(Rn) its dual. The Fourier transform of f ∈ f . The convolution of f ∈ S ′(Rn) and j ∈ S (Rn) S ′(Rn) resp. its inverse will be denoted by f resp. will be denoted by f ∗ j . By Lp(Rn) for 0 < p ≤ ¥ we denote the usual quasi-Banach space of p-integrable complex-valued functions with respect to the Lebesgue measure m with quasi-norm f (x)p dx(cid:19) k f Lp(Rn)k :=(cid:18)ZRn 1 p . Let X,Y be quasi-Banach spaces. By the notation X ֒→ Y we mean that X ⊂ Y and that the inclusion map is bounded. ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH3 Throughout the paper all unimportant constants will be called c,c′,C etc. Only if extra clarity is desirable, the dependency of the parameters will be stated explicitly. The concrete value of these constants may vary in different formulas but remains the same within one chain of inequalities. 1.1. Hölder spaces of differentiable functions. Let k ∈ N0. Then by Ck(Rn) we denote the space of all functions f : Rn → C which are k-times continuously differentiable (continuous, if k = 0) such that the norm k f Ck(Rn)k := (cid:229) a ≤k a sup D f (x) is finite, where the sup is taken over x ∈ Rn. (Rn) is defined by Furthermore, the set C¥ C (Rn) := \k∈N0 Ck(Rn). Definition 1.1. Let 0 < s ≤ 1 and f : Rn → C be continuous. We define f (x) − f (y) s k f lip (Rn)k := sup x,y∈Rn,x6=y s x − y . If s ∈ R, then there are uniquely determined ⌊s⌋ ∈ Z and {s} ∈ (0,1] with s = ⌊s⌋ + {s}. Let s > 0. Then the Hölder space with index s is given by C s(Rn) =n f ∈ C⌊s⌋(Rn) : k f C s(Rn)k < ¥ o with k f C s(Rn)k := k f C⌊s⌋(Rn)k + (cid:229) kD a f lip{s}(Rn)k. a =⌊s⌋ If s = 0, then C 0(Rn) := L¥ (Rn), which is sufficient for the later statements, see e.g. Theorem 2.12. 1.2. Besov and Triebel-Lizorkin function spaces on Rn. Let j j for j ∈ N0 be elements of S (Rn) with (2) supp j 0 ⊂ {x ≤ 2}, supp j j ⊂ {2 j−1 ≤ x ≤ 2 j+1} for j ∈ N, j j(x ) = 1 for all x ∈ Rn, j=0 a j D j(x ) ≤ ca 2− ja for all a ∈ Nn 0. Then we call {j j=0 a smooth dyadic resolution of unity. For instance one can choose Y ∈ S (Rn) with j} (x ) = 1 for x ≤ 1 and supp Y ⊂ {x ≤ 2} and set (x /2) − Y , s ∈ R and {j (x ), Definition 1.2. Let 0 < p ≤ ¥ Bs p,q(Rn) is the collection of all f ∈ S ′(Rn) such that the quasi-norm j 1(x ) := Y , 0 < q ≤ ¥ j 0(x ) := Y (x ), j} j j(x ) := j 1(2− j+1x ) for j ∈ N. j=0 be a smooth dyadic resolution of unity. Then k f Bs p,q(Rn)k := ¥ j=0 2 jsqk(j 1 q j f )Lp(Rn)kq! (modified if q = ¥ ) is finite. Definition 1.3. Let 0 < p < ¥ F s p,q(Rn) is the collection of all f ∈ S ′(Rn) such that the quasi-norm , s ∈ R and {j , 0 < q ≤ ¥ j} j=0 be a smooth dyadic resolution of unity. Then (modified if q = ¥ ) is finite. k f F s p,q(Rn)k :=(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ j=0 2 jsq(cid:12)(cid:12)(j j f )(·)(cid:12)(cid:12) 1 q q! (cid:12)(cid:12) Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ ¥ (cid:229) ¥ Y ¥ (cid:229) ¥ (cid:229) 4 BENJAMIN SCHARF One can show that the introduced quasi-norms1 for two different smooth dyadic resolutions of unity are equivalent for fixed p, q and s, i. e. that the so defined spaces are equal. This follows from Fourier multiplier theorems, see [27], Section 2.3.2., p. 46. Furthermore, the so defined spaces are (quasi)-Banach spaces. The next proposition, the so called Fatou property, is a classical observation for function spaces Bs p,q(Rn) and F s p,q(Rn), see [4]. Definition 1.4. Let A be a quasi-Banach space with S (Rn) ֒→ A ֒→ S ′(Rn). Then we say that A has the Fatou property if there exists a constant c such that: If a sequence { fn}n∈N ⊂ A converges to f with respect to the weak topology in S ′(Rn) and if k fnAk ≤ D, then f ∈ A and k f Ak ≤ c · D. Proposition 1.5. Let s ∈ R,0 < q ≤ ¥ and 0 < p ≤ ¥ . Then Bs the Fatou property. Remark 1.6. If r > 0 and r /∈ N, then C see [28, Sections 1.2.2, 2.6.5] or for the original source [36, Lemma 4]. ,¥ (Rn). This is a classical observation, for instance resp. 0 < p < ¥ p,q(Rn) and F s p,q(Rn) have (Rn) = B r r Moreover, we set where a+ = max(a,0). s p = n(cid:18) 1 p − 1(cid:19)+ , s p,q = n(cid:18) 1 min(p,q) − 1(cid:19)+ , 2. ATOMIC DECOMPOSITIONS At first we describe the concept of atoms as one can find it in [29], Definition 13.3, p. 73, now generalized using ideas from [25] and [33]. In particular, this gives the possibility to omit the distinction between n = 0 and n ∈ N and now the usual parameters K and L are nonnegative real numbers instead of natural numbers. 2.1. General atoms. Let Qn ,m := {x ∈ Rn : xi − 2−n mi ≤ 2−n −1} be the cube with sides parallel to the axes, with center at 2−n m and side length 2−n for m ∈ Zn and n ∈ N0. Definition 2.1. Let s ∈ R, 0 < p ≤ ¥ A function a : Rn → C is called (s, p)K,L-atom located at Qn ,m if (3) , K,L ∈ R and K,L ≥ 0. Furthermore let d > 1, C > 0, n ∈ N0,m ∈ Z. supp a ⊂ d · Qn ,m ka(2−n ·)C K(Rn)k ≤ C · 2−n (s− n p ) Remark 2.2. If L = 0, then condition (5) is neglectable since it follows from (3) and (4) with K = 0. p(cid:17). The constant in the exponent will be shortened by κL := s + L + n(cid:16)1 − 1 If K = 0, then by Definition 1.1 we only require a to be suitably bounded. Later on, we will choose one (s, p)K,L-atom for every n ∈ N0 and m ∈ Zn. Then the parameter d > 1 shall be the same for all these atoms - it describes the overlap of these atoms on one fixed level n ∈ N0. Remark 2.3. The usual formulation of (4) as in [29] was a a(x) ≤ 2−n (cid:16)s− n p(cid:17)+a n for all a ≤ K D (6) for K ∈ N0. The modification here was suggested in [33]. It is easy to see that (4) follows from (6) if K is a natural number, since CK(Rn) ֒→ C K(Rn). Remark 2.4. The usual formulation of (5) as in [29] was (7) for n ∈ N, so n 6= 0. The modification here was suggested in Lemma 1 of [25] for natural numbers L + 1 (using CL(Rn) instead of C L(Rn)). Now we extended this definition to general positive L. For natural L − 1 x a(x) dx = 0 if b ≤ L − 1 ZRn b 1In the following we will use the term "norm" even if we only have quasi-norms for p < 1 or q < 1. (4) and for every y ∈ C L(Rn) it holds (cid:12)(cid:12)(cid:12)(cid:12) Zd·Qn ,m (5) y (x)a(x) dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ C · 2−n (cid:16)s+L+n(cid:16)1− 1 p(cid:17)(cid:17)ky C L(Rn)k. ¥ ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH5 one can derive (5) from (7) using a Taylor expansion, see [25, Lemma 1, (12) and (14)] or the upcoming Lemma 2.8. Hence formulation (5) is a generalization. An alternative formulation of (5) is given by (8) (x − 2−n b m) ≤ C · 2−n κL if b ≤ ⌊L⌋. Zd·Qn ,m (cid:12)(cid:12)(cid:12)(cid:12) a(x) dx(cid:12)(cid:12)(cid:12)(cid:12) Obviously, this condition is covered by condition (5). For the other direction see [25, Lemma 1, (12) and (14)] or the upcoming Remark 2.9, in particular (10). It is also possible to assume this condition for all b ∈ Nn since the statements for b ≥ L follow from the support condition (3) and the boundedness condition included in (4). This shows that both conditions (4) and (5) are ordered in K resp. L, i.e. the conditions get stricter for increasing K resp. L. Now the question will be whether these more general atoms allow analogous results regarding atomic decompositions. 2.2. Sequence spaces. We introduce the sequence spaces bp,q and f p,q, whose use will become clear in the following. For this we refer to [29], Definition 13.5, p. 74. Definition 2.5. Let 0 < p ≤ ¥ , 0 < q ≤ ¥ and l = {l n ,m ∈ C : n ∈ N0,m ∈ Zn} . We set and l bp,q := : kl bp,qk =   : kl f p,qk =(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ n =0 ), where c (p) l f p,q :=  n =0 (cid:229) m∈Zn l n ,mp! l n ,mc (p) n ,m(·)q! m∈Zn 1 q q < ¥  p   Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) < ¥   (cid:12)(cid:12) 1 q (modified in the case p = ¥ or q = ¥ cube Qn ,m, i.e n ,m is the Lp(Rn)-normalized characteristic function of the c (p) n ,m = 2 n n p if x ∈ Qn ,m and c (p) n ,m = 0 if x /∈ Qn ,m. 2.3. Local means. Let N ∈ N0 be given. We choose k0,k ∈ S (Rn) with compact support - e.g. supp k0,supp k ⊂ e · Q0,0 for a suitable e > 0 - such that (9) while k0(0) 6= 0. Furthermore, let there be an e > 0 such that k(x) 6= 0 for 0 < x < e . a k(0) = 0 if a < N, D Such a choice is possible, see [29, 11.2]. We set k j(x) := 2 jnk(2 jx) for j ∈ N. Proposition 2.6. Let N ∈ N0 and N > s. . Then (i) Let 0 < p ≤ ¥ and 0 < q ≤ ¥ k f Bs p,q(Rn)kk0,k := kk0 ∗ f Lp(Rn)k + ¥ j=1 2 jsqkk j ∗ f Lp(Rn)kq! 1 q (modified for q = ¥ ) is an equivalent norm for k · Bs p,q(Rn)k. It holds (ii) Let 0 < p < ¥ and 0 < q ≤ ¥ k f F s Bs . Then p,q(Rn) =(cid:8) f ∈ S ′(Rn) : k f Bs p,q(Rn)kk0,k := kk0 ∗ f Lp(Rn)k +(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ p,q(Rn) =(cid:8) f ∈ S ′(Rn) : k f F s ) is an equivalent norm for k · Fs F s p,q(Rn)kk0,k < ¥ (cid:9) . q! 2 jsq(cid:12)(cid:12)(k j ∗ f )(·)(cid:12)(cid:12) p,q(Rn)kk0,k < ¥ (cid:9) . p,q(Rn)k. It holds j=1 1 q Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12) (modified for q = ¥ ¥ (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) 6 BENJAMIN SCHARF Remark 2.7. This proposition is due to [20]. Some minor technicalities of the proof where modified in the fourth step of [21, Theorem 2.1] (for the more general vector-valued case). 2.4. A general atomic representation theorem. We start with a lemma which helps us to understand the relation between conditions like (5) and (7) and which will be heavily used in the proof of the atomic representation theorem. It also shows that local means and atoms are related, see condition (9). Lemma 2.8. Let j ∈ N0. If k0 and k j = 2 jnk(2 j·) for j ∈ N are local means as in Definition 2.3, then p(cid:17)(cid:17) · k j is an (s, p)K,L-atom located at Q j,0 for arbitrary K > 0 and for L ≤ N + 1 with N from 2− j(cid:16)s+n(cid:16)1− 1 (9). Proof. For j = 0 there is nothing to prove since the moment condition (5) follows from (3) and (4). So we can concentrate on j ∈ N: The support condition (3) follows from the compact support of k. Furthermore, kk j(2− j)C K(Rn)k = 2 jnkkC K(Rn)k ≤ C 2 jn since K is arbitrarily often differentiable. Hence, the Hölder-condition (4) is shown. Now we have to show condition (5). There is nothing to prove for j = 0. Hence, we can use the moment conditions (9). Let L > 0, L = ⌊L⌋ + {L} as in Section 1.1 and let y ∈ C L(Rn). We expand the ⌊L⌋-times continuously differentiable function y into its Taylor series of order ⌊L⌋ − 1. Then there exists a q ∈ (0,1) with y (x) = (cid:229) b ≤⌊L⌋−1 b y (0) · x b D 1 b ! + (cid:229) b =⌊L⌋ 1 b ! b y (q x) · x b D . Hence y (x) − (cid:229) b ≤⌊L⌋ b y (0) · x D 1 b ! (cid:12)(cid:12)(cid:12) b (cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12) b =⌊L⌋ 1 b !(cid:0)D ≤ c · ky C L(Rn)k · xL. b y (q x) − D b y (0)(cid:1)x b (cid:12)(cid:12)(cid:12) Using (9) for k j and ⌊L⌋ ≤ N we can insert the polynomial terms into the integral and get (10) y (x)k j(x) dx(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)Zd·Q j,0 p(cid:17)(cid:17)k j fulfils condition (5). The constant C does not depend on j ∈ N0. Hence 2− j(cid:16)s+n(cid:16)1− 1 Remark 2.9. If we take a look at the proof, we see that instead of (9) it suffices to have ≤ c ky C L(Rn)kZd·Q j,0 k j(x) · xL dx ≤ C · 2− jLky C L(Rn)k. (cid:3) (11) ≤ C · 2− jL if b ≤ ⌊L⌋. b x (cid:12)(cid:12)(cid:12)Zd·Q j,0 k j(x) dx(cid:12)(cid:12)(cid:12) · y C L(Rn)k ≤ C if b ≤ ⌊L⌋, where (Rn) is a cutoff function, i.e. with compact support and y (x) = 1 for x ∈ supp k, hence for x ∈ In fact, this condition is equivalent to condition (5) for k j since kxb y ∈ C¥ supp k j, too. Now we will see what happens if an atom is dilated. Lemma 2.10. Let j ∈ N0 and j ≤ n . If an ,m is an (s, p)K,L-atom located at the cube Qn ,m, then 2 j(s− n an ,m(2− j·) is an (s, p)K,L-atom located at Qn − j,m. p ) · Proof. The support condition (3) and the Hölder-condition (4) are easy to verify. Considering the moment condition (5) we have (cid:12)(cid:12)(cid:12)Zd·Qn − j,m y (x)an ,m(2− jx) dx(cid:12)(cid:12)(cid:12) = 2 jn ·(cid:12)(cid:12)(cid:12)Zd·Qn ,m y (2 jx)an ,m(x) dx(cid:12)(cid:12)(cid:12) ≤ C · 2 jn · 2−n κL · ky (2 j·)C L(Rn)k ≤ C · 2 jn · 2−n κL · 2 jL · ky C L(Rn)k = C · 2−(n − j)κL · 2− j(s− n p ) · ky C L(Rn)k. This is what we wanted to prove. (cid:3) (cid:229) ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH7 Now we come to the essential part - showing the atomic representation theorem. We will use an approach as in Theorem 13.8 of [29]. Using the more general form of the atoms we are able to simplify the proof: One has to estimate Z k j(x − y)an ,m(y) dy, where k j are the local means from Section 2.3 and an ,m are atoms located at Qn ,m. One has to distinguish between j ≥ n and j < n as in the original proof - but now both cases can be proven very similarly with our more general approach of atoms. At first we prove the convergence of the atomic series in S ′(Rn). Lemma 2.11. Let 0 < p ≤ ¥ Then resp. 0 < p < ¥ , 0 < q ≤ ¥ and s ∈ R. Let K ≥ 0, L ≥ 0 with L > s p − s. l n ,man ,m n =0 m∈Zn converges unconditionally in S ′(Rn), where an ,m are (s, p)K,L-atoms located at Qn ,m and l ∈ bp,q or l ∈ f p,q. Proof. Let j ∈ S (Rn). Having in mind (3) and (5) we obtain m (cid:12)(cid:12)(cid:12)(cid:12) ZRn l n ,man ,m(x)j (x) dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ C · 2−n κL (cid:229) m l n ,m · kj n · y (2 · −m)C L(Rn)k, (Rn), y (x) = 1 for x ∈ d · Q0,0 and supp y ∈ (d + 1) · Q0,0. where y ∈ C¥ Observing κL = s + L + n(cid:16)1 − 1 (12) Furthermore, since j ∈ S (Rn) we have kj · y (2n 0 < p ≤ 1 p(cid:17) and L > s p − s we get κL >(0, . n(cid:16)1 − 1 p(cid:17) , 1 < p ≤ ¥ · −m)C L(Rn)k ≤ CM ·(cid:0)1 + 2−n where M ∈ N0 is at our disposal and CM does not depend on n and m. Let at first be 0 < p ≤ 1. Then we choose M = 0 and get , m(cid:1)−M l n ,m ≤ C′ · 2−n κL(cid:18)(cid:229) ≤ C′ · kl bp,¥ k. 1 p . l n ,mp(cid:19) m (13) we get m Summing up over n ∈ N0 using κL > 0 we finally arrive at ≤ C′ · 2−n κL (cid:229) l n ,man ,m(x)j (x) dx(cid:12)(cid:12)(cid:12)(cid:12) m (cid:12)(cid:12)(cid:12)(cid:12) ZRn n l n ,man ,m(x)j (x) dx(cid:12)(cid:12)(cid:12)(cid:12) m (cid:12)(cid:12)(cid:12)(cid:12) ZRn l n ,man ,m(x)j (x) dx(cid:12)(cid:12)(cid:12)(cid:12) ZRn m (cid:12)(cid:12)(cid:12)(cid:12) In the case 1 < p ≤ ¥ we choose M ∈ N0 such that Mp′ > n, where 1 = 1 p + 1 p′ . Using Hölder's inequality m ≤ CM · 2−n κL (cid:229) ≤ C′ · 2−n κL(cid:18)(cid:229) ≤ C′′ · 2−n κL · 2 l n ,m ·(cid:0)1 + 2−n m (cid:0)1 + 2−n p′ ·(cid:18)(cid:229) m(cid:1)−M ·(cid:18)(cid:229) m(cid:1)−M p′(cid:19) l n ,mp(cid:19) n n 1 p′ m 1 p . 1 p l n ,mp(cid:19) m By (12) the exponent is smaller than zero. Hence summing over n ∈ N0 gives the same result as in (13). Since we have shown the absolut and hence unconditional convergence in S ′(Rn). bp,q ֒→ bp,¥ resp. f p,q ֒→ f p,¥ (cid:3) ¥ (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) 8 BENJAMIN SCHARF Theorem 2.12. (i) Let 0 < p ≤ ¥ Then f ∈ S ′(Rn) belongs to Bs , 0 < q ≤ ¥ and s ∈ R. Let K,L ∈ R, K,L ≥ 0, K > s and L > s p − s. p,q(Rn) if and only if it can be represented as f = n =0 m∈Zn l n ,man ,m with convergence in S ′(Rn). Here an ,m are (s, p)K,L-atoms located at Qn ,m (with the same constants d > 1 and C > 0 in Definition 2.1 for all n ∈ N0,m ∈ Z) and kl bp,qk < ¥ . Furthermore, we have in the sense of equivalence of norms k f Bs p,q(Rn)k ∼ inf kl bp,qk, where the infimum on the right-hand side is taken over all admissible representations of f . (ii) Let 0 < p < ¥ , 0 < q ≤ ¥ and s ∈ R. Let K,L ∈ R, K,L ≥ 0, K > s and L > s p,q −s. Then f ∈ S ′(Rn) belongs to F s p,q(Rn) if and only if it can be represented as f = n =0 m∈Zn l n ,man ,m with convergence in S ′(Rn). Here an ,m are (s, p)K,L-atoms located at Qn ,m (with the same constants d > 1 and C > 0 in Definition 2.1 for all n ∈ N0,m ∈ Z) and kl f p,qk < ¥ . Furthermore, we have in the sense of equivalence of norms where the infimum on the right-hand side is taken over all admissible representations of f . k f F s p,q(Rn)k ∼ inf kl f p,qk, Proof. We rely on the proof of Theorem 13.8 of [29], now modified keeping in mind the more general conditions (4) and (5) instead of (6) and (7). There are two directions we have to prove. At first, let us assume that f from Bs p,q(Rn) is given. Then we know from Theorem 13.8 of [29] that f can be written as an atomic decomposition, with atoms now fulfilling conditions (6) and (7) for given natural numbers K′ > s and L′ + 1 > s p − s resp. L′ + 1 > s p,q − s. Hence, because of CK′ (Rn) ⊂ C K′ Conditions (5) are generalizations of the classical moment conditions (7) and are ordered in L, see (Rn), condition (4) is fulfilled for all K ≤ K′. p,q(Rn) or F s Remark 2.4. Thus, every classical (s, p)K′,L′-atom is an (s, p)K,L atom in the sense of definition 2.1 for K ≤ K′ and p,q(Rn) for L ≤ L′ + 1 and this immediately shows that we find a decomposition of f from Bs arbitrary K and L in terms of the general atoms we introduced. p,q(Rn) or Fs Now we come to the essential part of the proof. We have to show that, although we weakened the conditions on the atoms, a linear combination of atoms is still an element of Bs p,q(Rn). We modify the proof of Theorem 13.8 of [29] or into [21] where some minor technical details are modified (for the more general vector-valued case). There one uses the equivalent characterization by local means k0,k j := 2 jnk(2 j·) with a suitably large N(see Proposition 2.6) and distinguishes between the cases j ≥ n and j < n . In both cases the crucial part is the estimate of p,q(Rn) resp. Fs Z k j(x − y)an ,m(y) dy, where an ,m is an (s, p)K,L-atom centered at Qn ,m. The idea now is to use that not only an ,m but also k j can been interpreted as atoms and admit estimates as in (4) and (5), see Lemma 2.8. Let at first be j ≥ n . The function k has compact support and fulfils moment conditions (7). At first we transform the integral, having in mind the form of condition (4) of an ,m, 2 jsZ k j(y)an ,m(x − y) dy = 2 jsZ k j−n (y)an ,m(x − 2−n y) dy. Surely, this integral vanishes for x /∈ c · Qn ,m for a suitable c > 0 because of j ≥ n . So we concentrate on x ∈ c · Qn ,m: By Lemmata 2.8 and 2.10 the function 2−( j−n )(cid:16)s+n(cid:16)1− 1 p(cid:17)(cid:17) · k j−n = 2−( j−n )(cid:16)s+n(cid:16)1− 1 p(cid:17)(cid:17) · 2−n n · k j(2−n ·) is an (s, p)M,N-atom located at Q j−n ,0 for M arbitrarily large and N from (9), so that also N may be arbitrarily large, but fixed. Now we will use the moment condition (5) for k j−n and the Hölder-condition (4) for an ,m. ¥ (cid:229) (cid:229) ¥ (cid:229) (cid:229) ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH9 Hence, with y (y) = an ,m(x − 2−n y) and N ≥ K we have 2 js(cid:12)(cid:12)(cid:12)Z k j−n (y)an ,m(x − 2−n y) dy(cid:12)(cid:12)(cid:12) ≤ C · 2 js · 2−( j−n )K · kan ,m(x − 2−n = C · 2 js · 2−( j−n )K · kan ,m(2−n ≤ C · 2 js · 2−( j−n )K · 2−n (s− n p ) = C′ · 2−( j−n )(K−s) · c (p)(c · Qn ,m), ·)C K(Rn)k ·)C K(Rn)k where c (p)(c ·Qn ,m) is the Lp(Rn)-normalized characteristic function of c ·Qn ,m. This inequality is certainly true for x /∈ c · Qn ,m. Hence (13.37) in [29] is shown. Now let j < n . We will interchange the roles of k j and an ,m using condition (4) now for k j and (5) for an ,m. Hence we start with 2 jsZ k j(x − y)an ,m(y) dy = 2 jsZ k(2 jx − y)an ,m(2− jy) dy. Surely, this integral vanishes for x /∈ c · 2n − j · Qn ,m. So we concentrate on x ∈ c · 2n − j · Qn ,m: By Lemma 2.10 we know that 2 j(s− n p ) · an ,m(2− j·) is an (s, p)K,L-atom located at Qn − j,m while k is an (s, p)M,N -atom located at Q0,0. Thus, using (4) for k with M ≥ L, we get 2 js(cid:12)(cid:12)(cid:12)Z k(2 jx − y)an ,m(2− jy) dy(cid:12)(cid:12)(cid:12) p ) · kk(2 jx − ·)C L(Rn)k ≤ c · 2 js · 2−(n − j)κL · 2− j(s− n ≤ c · 2−(n − j)(L+s) · 2 = c · 2−(n − j)(L+s) · 2 n n p · 2−(n − j)n n n p · 2−(n − j)n · c (c · 2n − j · Qn ,m). where c (c · 2n − j · Qn ,m) is the characteristic function of c · 2n − j · Qn ,m. This estimate is the same as (13.41) combined with (13.42) in [29] or (72) and (73) in [21], observing that we use L instead of L + 1 in the atomic representation theorem. Starting with these two estimates we can follow the steps in [29] or [21] and finish the proof, since K > s and L > s p − s resp. L > s p,q − s. Strictly speaking, we arrive (in the Bs p,q(Rn)-case) at n ≤n 0 m≤m0 (cid:13)(cid:13)(cid:13) l n ,man ,m(cid:12)(cid:12) Bs p,q(Rn)(cid:13)(cid:13)(cid:13) ≤ C · kl bp,qk for all n 0,m0 ∈ N0 with a constant C independent of n 0 and m0. Using Lemma 2.11 and the Fatou property of the spaces Bs p,q(Rn) (see Proposition 1.5) we are finally done, i.e. p,q(Rn) resp. F s n ∈N0 m∈Zn (cid:13)(cid:13)(cid:13) l n ,man ,m(cid:12)(cid:12) Bs p,q(Rn)(cid:13)(cid:13)(cid:13) ≤ C · kl bp,qk. (cid:3) Remark 2.13. The conditions (4) and (5) for the atomic representation theorem can be slightly modified: If K > 0, then it is possible to replace k · C K(Rn)k by k · BK¥ ,¥ (Rn)k in condition (4). This is clear for K /∈ N, see Remark 1.6. If K ∈ N, this follows from C K(Rn) ֒→ BK¥ ,¥ (Rn) ֒→ C K−e (Rn) for e > 0. A similar result holds true for L > 0, L /∈ N and condition (5) by trivial means. If L ∈ N, then k ·C L(Rn)k can be replaced by k · CL(Rn)k, where the condition needs to be true for all y ∈ CL(Rn). This follows from the fact, that both conditions imply (8). Hence they are equivalent. It is not clear to the author whether k · C L(Rn)k can be replaced by k · BL¥ ,¥ (Rn)k for L ∈ N. Remark 2.14. In the proof of Theorem 2.12 we assumed that the local means k j are arbitrarily often differentiable and fulfil as many moment conditions as we wanted. But if we take a look into the proof, we see that we did not use the specific structure k j = 2 jnk(2 j·). It is sufficient to know that there are constants c and C such that for all j ∈ N0 it holds supp k j ⊂ c · Q j,0, that (14) kk j(2− j·)C M(Rn)k ≤ C · 2 jn (cid:229) (cid:229) (cid:229) (cid:229) 10 BENJAMIN SCHARF with M ≥ L and that for every y ∈ C N(Rn) it holds (15) ≤ C · 2− jN · ky C N(Rn)k with N ≥ K because the atomic conditions (4) and (5) are ordered in N and M, see Remark 2.4. As before, condition (15) can be strengthened by Zd·Q j,0 (cid:12)(cid:12)(cid:12)(cid:12) y (x)k j(x) dx(cid:12)(cid:12)(cid:12)(cid:12) Z x b k j(x) dx = 0 for all b < N. Through these considerations the idea arises how to prove a counterpart of Theorem 2.12 for the local mean characterization in [32, Theorem 1.15] without further substantial efforts. This is done in the following Corollary, including some technical issues concerning the definition of a dual pairing (see [32, Remark 1.14]). It is obvious that the original version of Theorem 1.15 in [32] is just some kind of modification of this Corollary. , 0 < q ≤ ¥ and s ∈ R. Let M,N ∈ R, M,N ≥ 0, M > s p − s and N > s. Corollary 2.15. (i) Let 0 < p ≤ ¥ Assume that for all j ∈ N0 it holds that k j ∈ C M(Rn), supp k j ⊂ c · Q j,0 and k j fulfils (14) and (15). Then there is a constant c such that k f Bs p,q(Rn)kk := kk0 ∗ f Lp(Rn)k + ¥ j=1 2 jsqkk j ∗ f Lp(Rn)kq! 1 q ≤ c · k f Bs p,q(Rn)k (modified for q = ¥ ) for all f ∈ Bs (ii) Let 0 < p < ¥ p,q(Rn). and s ∈ R. Let M,N ∈ R, M,N ≥ 0, M > s p,q − s and N > s. Assume that for all j ∈ N0 it holds that k j ∈ C M(Rn), supp k j ⊂ c · Q j,0 and k j fulfils (14) and (15). Then there is a constant c such that , 0 < q ≤ ¥ k f F s p,q(Rn)kk : = kk0 ∗ f Lp(Rn)k +(cid:13)(cid:13)(cid:13)(cid:13)(cid:16) p,q(Rn)k ≤ c · k f Fs p,q(Rn). j=1 2 jsq(cid:12)(cid:12)(k j ∗ f )(·)(cid:12)(cid:12) q(cid:17) 1 q(cid:12)(cid:12)(cid:12) Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13) ) for all f ∈ F s (modified for q = ¥ Proof. There is nearly nothing left to prove because the crucial steps were done in the proof before: Let f ∈ Bs p,q(Rn) (analogously for f ∈ F s p,q(Rn) by an "optimal" atomic decomposition p,q(Rn)) be given. By Theorem 2.12 we can represent f ∈ Bs f = n =0 m∈Zn l n ,man ,m, where an ,m is an (s, p)N,M-atom located at Qn ,m and k f Bs of f ). p,q(Rn)k ∼ kl bp,qk (with constants independent But, by the second step of the proof of Theorem 2.12 and the considerations in the succeeding remark we have (16) for all n 0,m0 ∈ N0 with a constant C independent of n 0 and m0. n ≤n 0 m≤m0 (cid:13)(cid:13)(cid:13) l n ,man ,m(cid:12)(cid:12) Bs p,q(Rn)(cid:13)(cid:13)(cid:13)k ≤ C · kl bp,qk ∼ k f Bs p,q(Rn)k Finally, we use a similar duality argument as in [32, Remark 1.14] or [31, Section 5.1.7] to justify the dual pairing of k j and f . Looking into the proof of Lemma 2.11, we see that (17) n m (cid:12)(cid:12)(cid:12)(cid:12) ZRn l n ,man ,m(x)j (x) dx(cid:12)(cid:12)(cid:12)(cid:12) for j ∈ C M(Rn) with compact support, M − e ≥ 0 and M − e > s p − s, where C′ depends on the support of j . This includes the functions k j for j ∈ N0. Because of this absolut convergence the dual pairing of f and j is given by ≤ C′ · kj C M−e (Rn)k · kl bp,¥ k lim m0,n 0→¥ m≤m0ZRn n ≤n 0 l n ,man ,m(x)j (x) dx. (cid:229) ¥ (cid:229) ¥ (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH11 Furthermore, for two different atomic decompositions of f these limits are the same: By definition of a distribution f ∈ S ′(Rn) and Lemma 2.11 this is valid for j ∈ S (Rn). For arbitrary j ∈ C M(Rn) with compact support this follows by (17) and density arguments because C¥ (Rn) is dense in C M(Rn) with respect to the norm of C M−e C M(Rn) ֒→ BM¥ (Rn). For instance, this can be seen using ,q (Rn) ֒→ BM−e p,q(Rn) if q < ¥ ,¥ (Rn) ֒→ BM−e (Rn) is dense in Bs . ,¥ = C M−e (Rn) for M − e /∈ N0 and the fact that C¥ Hence we have l n ,m (an ,m ∗ k j) (x) → ( f ∗ k j)(x) for n 0,m0 → ¥ n ≤n 0 m≤m0 for all x ∈ Rn. Using the standard Fatou lemma and (16) we finally get k f Bs p,q(Rn)kk ≤ C · kl bp,qk ∼ k f Bs p,q(Rn)k. 3. KEY THEOREMS (cid:3) 3.1. Pointwise multipliers. Triebel proved in Section 4.2 of [28] the following assertion. Theorem 3.1. Let s ∈ R and 0 < q ≤ ¥ (i) Let 0 < p ≤ ¥ and r > max(s,s p − s). Then there exists a positive number c such that . p,q(Rn)k ≤ ckj C r (Rn)k · k f Bs p,q(Rn)k for all j ∈ C r (Rn) and all f ∈ Bs (ii) Let 0 < p < ¥ and r > max(s,s p,q − s). Then there exists a positive number c such that kj kj f Bs p,q(Rn). f F s p,q(Rn). for all j ∈ C r (Rn) and all f ∈ F s p,q(Rn)k ≤ ckj C r (Rn)k · k f Fs p,q(Rn)k is still an (s, p)K,L-atom. He excluded the cases r ∈ N. This is not necessary in our considerations. The very first idea to prove this result is to take an atomic decomposition of f , to multiply it by j and to prove that the resulting sum is again a sum of atoms. Hence one has to check whether a product of an (s, p)K,L-atom and a function j But there was a problem: Moment conditions like (7) are (in general) destroyed by multiplication with j . So the atomic approach in [28] only worked when no moment conditions were required, hence if s > s p resp. s > s p,q, and the full generality of Theorem 3.1 had to be obtained by an approach via local means. Looking at condition (5) instead the situation when multiplying by j Furthermore, the atomic approach only worked for j ∈ Ck(Rn) with k ∈ N and k > s having in mind condition (6). Now we are able to give weaker conditions using the new atomic approach with condition (4). is now different. We start with a first standard analytical observation. Lemma 3.2. Let s ≥ 0. There exists a constant c > 0 such that for all f ,g ∈ C s(Rn) the product f · g belongs to C s(Rn) and it holds k f · gC s(Rn)k ≤ c · k f C s(Rn)k · kgC s(Rn)k. Proof. This can be proven using standard arguments, especially Leibniz formula. (cid:3) Now we are ready to prove Theorem 3.1. This is done by the following lemma together with Theorem 2.12 using the mentioned technique of atomic decompositions. For some further technicalities see the upcoming Remark 3.5 or [28, 4.2.2, Remark 1]. This covers also the well-definedness of the product. Lemma 3.3. There exists a constant c with the following property: For all n ∈ N0, m ∈ Z, all (s, p)K,L- atoms an ,m with support in d · Qn ,m and all j ∈ C r c · kj C (Rn) with r ≥ max(K,L) the product (Rn)k−1 · j r · an ,m is an (s, p)K,L-atom with support in d · Qn ,m. ¥ ¥ (cid:229) (cid:229) 12 BENJAMIN SCHARF Proof. Regarding the conditions (4) on the derivatives Lemma 3.2 gives k(j · a)(2−n ·)C K(Rn)k ≤ c · kj (2−n ·)C K(Rn)k · ka(2−n ·)C K(Rn)k ≤ c′ · kj C K(Rn)k · 2−n (s− n p ). Now we come to the preservation of the moment conditions (5). By our assumptions there exists a constant C > 0 such that for every y ∈ C L(Rn) it holds Using this inequality now for y instead of y (cid:12)(cid:12)(cid:12)(cid:12) y (x)a(x) dx(cid:12)(cid:12)(cid:12)(cid:12) Zd·Qn ,m · j =(cid:12)(cid:12)(cid:12)(cid:12) y (x)(cid:0)j (x) · a(x)(cid:1) dx(cid:12)(cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12) Zd·Qn ,m ≤ C · 2−n κLky C L(Rn)k. together with Lemma 3.2 it follows Zd·Qn ,m(cid:0)y (x) · j (x)(cid:1)a(x) dx(cid:12)(cid:12)(cid:12)(cid:12) ≤ C · 2−n κLky ≤ C′ · 2−n κLky C L(Rn)k · kj C L(Rn)k. · j C L(Rn)k Hence our lemma is shown. (cid:3) Remark 3.4. This is the more general version of part 1 of Lemma 1 in [25] using now the wider atomic approach from 2.1 which yields a stronger result than in [25]. Remark 3.5. As at the end of Corollary 2.15 we have to deal with some technicalities. We concentrate on the Bs p,q(Rn)-case is nearly the same. In principle, Lemma 3.3 shows that p,q(Rn)-case, the F s (18) converges unconditionally in S ′(Rn) where f = (cid:229) n l n ,m(an ,m · j ) l n ,man ,m in S ′(Rn) n m m and the limit belongs to Bs p,q(Rn) if f belongs to Bs p,q(Rn). To define the product of j and f as this limit, we have to show that the limit does not depend on the atomic decomposition we chose for f . Hence we are pretty much in the same situation as at the end of Corollary 2.15: Let at first be j ∈ C¥ . is a continuous operator mapping S ′(Rn) to S ′(Rn). So (18) converges to · f for all choices of atomic decompositions of f . Using Lemma 3.3 and the standard Fatou lemma we get Then the multiplication with j j for all f ∈ Bs p,q(Rn). For arbitrary j ∈ C kj · f Bs p,q(Rn)k ≤ c · kj C r (Rn)k · k f Bs p,q(Rn)k r (Rn) we use a density argument similar to that at the end of Corollary 2.15. We know kj ∗ · f Bs p,q(Rn)k ≤ c · kj ∗C r −e (Rn)k · k f Bs p,q(Rn)k (Rn), r as in Lemma 3.3 and e small enough. Now using the density of C¥ (Rn) in C r (Rn) for j ∗ ∈ C¥ with respect to the norm of C r −e · f Bs p,q(Rn)k ≤ c · kj C (Rn) the uniqueness of the product and p,q(Rn)k ≤ c · kj C (Rn)k · k f Bs r −e r (Rn)k · k f Bs p,q(Rn)k kj follows. Remark 3.6. Since C L(Rn) ֒→ BL¥ (Rn)k by kj B r ,¥ (Rn) ֒→ C L−e r ,¥ (Rn)k, even by kj B (Rn) r for L −e ≥ 0, we can replace kj C in Lemma 3.1. ,q(Rn)k for arbitrary 0 < q ≤ ¥ p,q(Rn)-spaces in Theorem 3.1 can be replaced by r > max(s,s p − s). This is a matter of complex interpolation, see the proof of the corollary in Section 4.2.2 of [28]. The condition r > max(s,s p,q − s) for the F s (cid:229) (cid:229) (cid:229) ¥ ¥ ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH13 r ,¥ (Rn) for r > 0 and r /∈ N. So, let f ∈ Bs Remark 3.7. Our Theorem 3.1 is a special case of Theorem 4.7.1 in [19]: By Remark 1.6 it holds C B Then j ∈ B p,q(Rn) as well as j ∈ C p,q(Rn) or f ∈ F s r ′ ,¥ (Rn) for s < r ′ < r . By Theorem 4.7.1 of [19] it holds p,q(Rn) · B r ′ ,¥ (Rn) ֒→ Bs r ′ ,¥ (Rn) ֒→ F s p,q(Rn) Bs p,q(Rn) · B resp. F s p,q(Rn) r r (Rn) = (Rn) with r > s. if r ′ > s and r ′ > s p − s. p,q(Rn) these are the same conditions as in Theorem 3.1 - in case of F s and s + r ′ > s p ⇔ r ′ > s p,q(Rn) these are even In case of Bs better (no dependency on q). It was not the idea of this paper to give such a detailed and comprehensive treatise as in Runst' and Sickel's book [19] but to show an application of the more general atomic decompositions where the proof is easy to follow (see Triebel [28, Section 4.1]). 3.2. Diffeomorphisms. We want to study the behaviour of the mapping Dj where f is an element of the function space Bs map. : f 7→ f (j (·)), p,q(Rn) resp. Fs p,q(Rn) and j : Rn → Rn is a suitably smooth One would like to deal with this problem analogously to the pointwise multiplier problem in Section 3.1. Hence we start with an atomic decomposition of f and composed with j . Then we are confronted with functions of the form an ,m ◦j originating from the atoms an ,m. This was the idea of Section 4.3.1 in Triebel [28]. But in general, moment conditions of type (7) are destroyed by this operator. So s > s p resp. s > s p,q was necessary. As we will see, conditions like (5) behave more friendly under diffeomorphisms. Furthermore, we are confronted with more difficulties than in section 3.1 because the support of an atom changes remarkably. In particular, after composing with j two or more atoms can be associated with the same cube Qn ,m which is not possible in the atomic representation theorem 2.12. This has not been considered in detail in Section 4.3.1 by Triebel [28] while there is some work done in the proof of Lemma 3 by Skrzypczak [25]. The special case of bi-Lipschitzian maps, also called Lipschitz diffeomorphisms, is treated in Section 4.3 by Triebel [30]. The main theorem there is used to obtain results for characteristic functions of Lipschitz domains as pointwise multipliers in Bs Definition 3.8. Let r ≥ 1. p,q(Rn) and F s p,q(Rn). (i) Let r = 1. We say that the map j : Rn → Rn is a r -diffeomorphism if j is a bi-Lipschitzian map, i.e. that there are constants c1,c2 > 0 such that (19) c1 ≤ j (x) − j (y) x − y ≤ c2. for all x,y ∈ Rn with 0 < x − y ≤ 1. (ii) Let r > 1. We say that the one-to-one map j : Rn → Rn is a r -diffeomorphism if the components r −1(Rn) for all j i of j (x) = (j 1(x), . . . ,j n(x)) have classical derivatives up to order ⌊r ⌋ with i, j ∈ {1, . . . ,n} and if detJ(j )(x) ≥ c for some c > 0 and all x ∈ Rn. Here J(j )(x) stands for the Jacobian matrix of j at the point x ∈ Rn. Remark 3.9. It does not matter, whether we assume (19) for all x,y ∈ Rn with x 6= y or for all x,y ∈ Rn with 0 < x − y < c for a constant c > 0. This is obvious for the upper bound. For the lower bound we have to use the upper bound of the bi-Lipschitzian property of the inverse j −1 of j . Its existence independent of the given exact definition of a bi-Lipschitzian map is shown in the following lemma. Lemma 3.10. Let r ≥ 1. ¶j i ¶ x j ∈ C (i) If j (ii) Let r > 1. If j (iii) If j is a 1-diffeomorphism, then j is bijective and j −1 is a 1-diffeomorphism, too. is a r -diffeomorphism, then j is a r -diffeomorphism, then its inverse j −1 is a r -diffeomorphism as well. is a r ′-diffeomorphism for 1 ≤ r ′ ≤ r . Hence j is a bi- Lipschitzian map. Proof. To prove part (i) we use Brouwer's invariance of domain theorem (see [2]): Since j : Rn → Rn is continuous and injective, the image j (U) of U is an open set if U is open. Otherwise, if U is closed, then also j (U) is closed: If j (xn) → y with xn ∈ U, then xn converges to some x ∈ U by (19) and hence ¥ ¥ ¥ ¥ 14 BENJAMIN SCHARF j (xn) → j (x) = y. Thus j maps Rn to Rn. The inverse j −1 is automatically a bi-Lipschitzian map, see (19). The proof of observation (iii) for r ′ > 1 is trivial. Hence, we have to show that every r -diffeomorphism is a bi-Lipschitzian map for r > 1. The estimate j (x) − j (y) x − y ≤ c2 follows from the fact that the derivatives ¶j i ¶ x j are bounded for all i, j ∈ {1, . . . ,n}. The formula (20) J(j −1)(j (x)) = (J(j )(x))−1 and detJ(j )(x) ≥ c together show that the derivatives of the inverse are bounded for all i, j ∈ {1, . . . ,n}, for instance using the adjugate matrix formula. By the mean value theorem there exists a c > 0 such that ¶ x j ¶ (j −1)i j −1(x) − j −1(y) x − y ≤ c and so part (iii) is shown. Finally, for (ii) we have to show that ¶ (j −1)i ¶ x j part follows from (20) and the boundedness of way as in the inverse function theorem, starting with r −1(Rn) and detJ(j −1)(x) ≥ c for r > 1. The latter . For the first we have to argue inductively in the same ∈ C ¶j i ¶ x j It is well known that J(j −1)(x) =(cid:0)J(j )(j −1(x))(cid:1)−1 A → A−1 is a C¥ (Rn×n)-mapping for invertible A. Together with the upcoming Lemma 3.11 this shows: If the r −1(Rn) and j −1 is an l-diffeomorphism, then the components of J(j −1) components of J(j ) belong to C belong to C min(r −1,l)(Rn) and hence j −1 is a min(l + 1,r )-diffeomorphism. This inductive argument and the induction starting point that j −1 is a 1-diffeomorphism (by part (i) and (iii)) prove that j −1 is a r - diffeomorphism. Thus the lemma is shown. (cid:3) We go on with a second standard analytical observation. Lemma 3.11. Let j be a r -diffeomorphism and let max(1,s) ≤ r . Then there exists a constant C depending on r such that for all f ∈ C s(Rn) it holds k f ◦ j C s(Rn)k ≤ Cj · k f C s(Rn)k. Proof. By definition k f ◦ j C s(Rn)k = k f ◦ j C⌊s⌋(Rn)k + (cid:229) a =⌊s⌋ a kD [ f ◦ j ] lip{s}(Rn)k. The lemma follows now by using the chain rule and Leibniz rule for spaces of differentiable functions and for Hölder spaces C s(Rn). (cid:3) Remark 3.12. As one can easily see, the constant in Lemma 3.11 depends on we have a sequence of functions {j m}m∈N and n(cid:229) ¶j m i n(cid:229) C < ¥ , sup m∈N i=1 j=1(cid:13)(cid:13)(cid:13)(cid:13) ¶ x j (cid:12)(cid:12) r −1(Rn)(cid:13)(cid:13)(cid:13)(cid:13) then there is a universal constant C with Cj m ≤ C, i.e. for all m ∈ N it holds k f ◦ j mC s(Rn)k ≤ C · k f C s(Rn)k. ¶j i ¶ x j C n(cid:229) i=1 n(cid:229) j=1(cid:13)(cid:13)(cid:13) . If r −1(Rn)(cid:13)(cid:13)(cid:13) ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH15 Lemma 3.13. Let 0 < p ≤ ¥ . Let j : Rn → Rn be bijective and let there be a constant c > 0 such that (21) c ≤ j (x) − j (y) x − y for x,y ∈ Rn with x 6= y. Then there is a constant C > 0 such that (22) Proof. If p < ¥ , it suffices to prove (22) for k f ◦ j Lp(Rn)k ≤ C · k f kLp(Rn)k. where a j ∈ C, A j are pairwise disjoint rectangles in Rn and c A j is the characteristic function of A j. We have N(cid:229) f = a jc A j , j=1 Z ( f ◦ j )(x)p dx =Z (cid:12)(cid:12)(cid:12) N(cid:229) j=1 a jc j −1(A j)(x)(cid:12)(cid:12)(cid:12) p dx = N(cid:229) j=1 a jpm (j −1(A j)) because the preimages j −1(A j) are also pairwise disjoint. Hence we have to show: There is a constant C > 0 such that for all rectangles A it holds m (j −1(A)) ≤ C · m (A). (23) To prove this let Br(x0) = {x ∈ Rn : x − x0 < r} be the open ball around x0 ∈ Rn with radius r > 0. Then by (21) we have (24) Hence there is a constant C > 0 such that j −1(Br(x0)) ⊂ B r (j −1(x0)). c for all x0 ∈ Rn, r > 0. Now, we cover a given rectangle A with finitely many open balls {B j}M j=1 such that m (j −1(Br(x0))) < C · m (Br(x0)) (25) m M [j=1 B j! ≤ 2m (A). Afterwards we make use of the following Vitali covering lemma: There exists a subcollection B j1, . . . ,B jm of these balls which are pairwise disjoint and satisfy Using this, (25) and (24) for the balls 3 · B jk finally gives M m B j ⊂ 3 · B jk. [j=1 [k=1 m (j −1(A)) ≤ m j −1 M [j=1 B j!! ≤ m j −1 M [k=1 m (j −1(3 · B jk)) ≤ C · m (3 · B jk) ≤ C · 3n · M(cid:229) k=1 M(cid:229) k=1 = ≤ 2C · 3n · m (A). . This proves the result for 0 < p < ¥ For p = ¥ we have to show j −1(3 · B jk)! 3 · B jk!! = m M [k=1 m (B jk ) M(cid:229) k=1 k f ◦ j L¥ (Rn)k ≤ k f L¥ (Rn)k. This follows from: If m ({x ∈ Rn : f (x) > a)}) = 0, then also m ({x ∈ Rn : f (j (x)) > a}) = 0, which is a consequence of (23): Let M be a measurable set with m (M) = 0. Then also m (j −1(M)) = 0. Hence the lemma is shown for p = ¥ , too. Remark 3.14. A proof of a more general observation using the Radon-Nikodym derivative and the Lebesgue point theorem can be found in Corollary 1.3 and Theorem 1.4 of [35] - but here we wanted to give a direct, more instructive proof for our special situation. (cid:3) 16 BENJAMIN SCHARF Remark 3.15. By the previous proof it is obvious that Condition (23) is equivalent to (22) for 0 < p < ¥ . Condition (23) does not depend on p. For Condition (23) it is necessary that the measure m with m(A) := m (j −1(A)) is absolutely continuous with respect to the Lebesgue measure m . In case of p = ¥ this condition is also sufficient for (23) by the previous proof. Now we are ready for the main theorem of this section. Theorem 3.16. Let s ∈ R, 0 < q ≤ ¥ and r ≥ 1. (i) Let 0 < p ≤ ¥ and r > max(s,1 + s p − s). If j such that is a r -diffeomorphism, then there exists a constant c k f (j (·))Bs p,q(Rn)k ≤ c · k f Bs p,q(Rn)k for all f ∈ Bs p,q(Rn) onto Bs (ii) Let 0 < p < ¥ and r > max(s,1 + s p,q − s). If j p,q(Rn). Hence Dj maps Bs p,q(Rn). is a r -diffeomorphism, then there exists a constant c such that for all f ∈ F s p,q(Rn). Hence Dj maps Fs p,q(Rn) onto Fs p,q(Rn). k f (j (·))F s p,q(Rn)k ≤ c · k f Fs p,q(Rn)k Proof. At first, beside the two conditions (4) and (5) we need to take a closer look at the centres and supports of the atoms. Briefly speaking, the decisive local properties of the set of atoms an ,m are maintained by a superposition with the diffeomorphism j . To be more specific: Let Mn = {x ∈ Rn : x = 2−n m,m ∈ Zn}. Having in mind Lemma 3.10 there is a c2 > 0 with (26) By a simple volume argument for Qn ,m and by 2−n m − 2−n m′ ≥ c · 2−n for m 6= m′ there is a constant M ∼ cn x − y ≤ c2j −1(x) − j −1(y). 2 such that for all n ∈ N0,m ∈ Z. Hence we can take our atomic decomposition and split it into M disjunct sums, i.e. j −1(Mn ) ∩ Qn ,m ≤ M with M f = M(cid:229) j=1 n ∈N0 m∈Mn , j l n ,man ,m Mn , j = Zn, Mn , j ∩ Mn , j′ = /0 for j 6= j′ [j=1 so that for all n ∈ N0, m ∈ Zn and j ∈ {1, . . . ,M} (27) Therefore, not more than one function an ′,m′ ◦ j (cid:8)m′ ∈ Zn : m′ ∈ Mn , j and j −1(2−n The support of a function an ,m ◦ j c1 > 0 with is located at the cube Qn ,m for each of the M sums. is contained in j −1(d · Qn ,m) by (3). By Lemma 3.10 there exists a m′) ∈ Qn ,m(cid:9) ≤ 1. Hence we get j −1(x) − j −1(y) ≤ 1 c1 x − y. j −1(d · Qn ,m) ⊂ c · · B2−n (j −1(2−n m)), d c1 where Br(x0) = {x ∈ Rn : x − x0 ≤ r}. Hence, together with (27) it follows: There is a constant d′ depend- ing on c1 such that for every n ∈ N0 and every j ∈ {1, . . . ,M} there is an injective map F n , j : Mn , j → Zn with supp (an ,m ◦ j ) ⊂ d′ · Qn ,F (28) for all m ∈ Mn , j. The constant d′ does not depend on n or m. n , j (m). (cid:229) (cid:229) ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH17 Thus, if we take the derivative conditions (4) and the moment conditions (5) for an ,m ◦j now for granted (which will be shown later), then f j ◦ j = (cid:229) n ∈N0 m∈Mn , j l n ,m(an ,m ◦ j ) is an atomic decomposition of the function f j ◦ j . Finally, we have to look at the sequence space norms, see Definition 2.5. We will concentrate on the F s p,q(Rn)-case is easier because it does not matter if one changes the order of summation over m. By the atomic representation theorem and (28) we will have p,q(Rn)-case since the Bs k f j ◦ j Fs p,q(Rn)k ≤ c(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ l n ,mc (p) n ,F 1 q n , j(m)(·)q! m∈Mn , j n =0 . Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12) To transfer this into the usual sequence space norm we make use of n , j (m) ⊂ j −1(c · Qn ,m) (29) with a constant c depending on c2 from (26), but independent of n and m. This follows from j −1(2−n m) ∈ Qn ,F n , j (m). Hence assuming that an ,m ◦ j fulfil (4) and (5) we obtain Qn ,F (30) k f j ◦ j Fs p,q(Rn)k ≤ c(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ ≤ c′(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ ≤ c′(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) ¥ l n ,mc (p) n =0 m∈Mn , j l n ,mc (p) n =0 m∈Mn , j 1 q n ,m(j (·))q! n ,m(·)q! n ,m(·)q! 1 q 1 q Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12) Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12) Lp(Rn)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13)(cid:13) (cid:12)(cid:12) l n ,mc (p) m∈Zn n =0 p,q(Rn)k. ≤ c′′k f F s In the first step we used (29), in the second step we used Lemma 3.13 and part (iii) of Lemma 3.10 and in the last step we applied the atomic decomposition theorem for f . As done in the first step, one can replace the characteristic function of c · Qn ,m by the characteristic function of Qn ,m in the sequence space norm getting equivalent norms, see [32, section 1.5.3]. This can be proven using the Hardy-Littlewood maximal function. Finally, we have to take a look at the derivative conditions (4) and the moment conditions (5). The latter part is also considered in Lemma 5 of [25] using the atomic approach with condition (6). Let an ,m be an (s, p)K,L-atom and let r ≥ max(K,L + 1). If we can show that j ◦ an ,m is an (s, p)K,L- atom as well, we are done with the proof since we can choose K and L suitably small enough by the atomic decomposition theorem 2.12. Let Tn (x) := 2−n x and Tn (j ) = T −1n ◦ j ◦ Tn . Then k (an ,m ◦ j ) (2−n ·)C K(Rn)k = kan ,m ◦ j ◦ Tn C K(Rn)k = kan ,m ◦ Tn ◦ Tn (j )C K(Rn)k. By a simple dilation argument for the Hölder spaces C r −1(Rn) it holds ¶ (Tn (j ))i ¶ x j (cid:13)(cid:13)(cid:13)(cid:13) C (cid:12)(cid:12) r −1(Rn)(cid:13)(cid:13)(cid:13)(cid:13) ≤(cid:13)(cid:13)(cid:13)(cid:13) C ¶j i ¶ x j(cid:12)(cid:12) r −1(Rn)(cid:13)(cid:13)(cid:13)(cid:13) for all i, j ∈ {1, . . . ,n} and n ∈ N0. Hence by Lemma 3.11 and Remark 3.12 we find a constant C indepen- dent of n and m such that k (an ,m ◦ j )(2−n ·)C K(Rn)k = kan ,m ◦ Tn ◦ Tn (j )C K(Rn)k ≤ C · kan ,m(2−n ·)C K(Rn)k So the derivative condition (4) is shown. Regarding the moment condition (5) of an ,m ◦j we consider two cases: At first, let j be a r -diffeomorphism with r > 1. Then j and j −1 are differentiable. We use the moment condition of an ,m itself and Lemma (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) 18 3.11 to get BENJAMIN SCHARF Z d′·Qn ,F n , j (m) (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) y (x) · a(j (x)) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) We used the transformation formula for integrals and Z j −1(d·Qn ,m) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) y (x) · a(j (x)) dx(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) y (cid:0)j −1(x)(cid:1) · detj −1(x) · a(x) dx(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) Zd·Qn ,m ≤ C · 2−n κL · kdetj −1(x) ·(cid:0)y ◦ j −1(cid:1) C L(Rn)k det J(cid:0)j −1(cid:1) ∈ C L(Rn) ≤ C′ · 2−n κL · ky C L(Rn)k. since j If r = 1, then L = 0 by our choice of r . This means, that no moment conditions are needed. Hence we is a r -diffeomorphism with r ≥ L + 1. Furthermore, the sign of detJ(cid:0)j −1(cid:1) is constant. have nothing to prove. The choice of r = 1 is only allowed if s p < s < 1 resp. s p,q < s < 1. For some further technicalities similar as in Remark 3.5 see Remark 3.20. (cid:3) Remark 3.17. This has been proven (in a sketchy way) in Lemma 3 in [25] for the more special atomic definition there. Remark 3.18. If s p < s < 1 resp. s p,q < s < 1, then the choice of r = 1 is possible for these values of s. This gives the same result as in Proposition 4.1 in [30], where the notation of Lipschitz diffeomorphisms as in Definition 3.8 is used. This results in Theorem 3.19. Let 0 < q ≤ ¥ . (i) Let 0 < p ≤ ¥ and s p < s < 1. If j : Rn → Rn is a bi-Lipschitzian map, then there exists a constant c such that k f (j (·))Bs p,q(Rn)k ≤ c · k f Bs p,q(Rn)k. for all f ∈ Bs p,q(Rn). Hence Dj maps Bs (ii) Let 0 < p < ¥ and s p,q < s < 1. If j p,q(Rn) onto Bs p,q(Rn). : Rn → Rn is a bi-Lipschitzian map, then there exists a constant c such that for all f ∈ F s p,q(Rn). Hence Dj maps Fs p,q(Rn) onto Fs p,q(Rn). k f (j (·))F s p,q(Rn)k ≤ c · k f Fs p,q(Rn)k. Remark 3.20. We have to deal with some technicalities of the proof of Theorem 3.16. We concentrate on the Bs p,q(Rn)-case is nearly the same. p,q(Rn)-case, the F s Let at first be r > 1. In principle, Theorem 3.16 and Lemma 2.11 show that (31) n converges unconditionally in S ′(Rn), where l n ,m(an ,m ◦ j ) m f = (cid:229) l n ,man ,m in S ′(Rn), n m To define the superposition of f and j as this limit, we have to show that the limit does not depend on p,q(Rn). and the limit belongs to Bs p,q(Rn) if f belongs to Bs the atomic decomposition we chose for f . Let y ∈ C¥ ZRn l n ,m (an ,m ◦ j ) (x)y (x) dx(cid:12)(cid:12)(cid:12)(cid:12) m (cid:12)(cid:12)(cid:12)(cid:12) m (cid:12)(cid:12)(cid:12)(cid:12) ZRn = (cid:229) n n (Rn) with compact support be given. Then l n ,man ,m(x)(cid:2)y (cid:0)j −1(x)(cid:1) · detj −1(x)(cid:3) dx(cid:12)(cid:12)(cid:12)(cid:12) makes sense, see (17), because by Lemma 3.11 the function y (cid:0)j −1(x)(cid:1) · detj −1(x) has compact support and belongs to C M(Rn) for a suitable M > 0 with M > s p −s. Now the achievements at the end of Corollary 2.15 show that this integral limit does not depend on the choice of the atomic decomposition for f . Hence we obtain that the limit in (31) (considered as an element in S ′(Rn)) is the same for all choices of atomic decompositions. (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) (cid:229) ATOMIC REPRESENTATIONS IN FUNCTION SPACES AND APPLICATIONS TO POINTWISE MULTIPLIERS AND DIFFEOMORPHISMS, A NEW APPROACH19 If the choice of r = 1 is allowed, then automatically s > s p and Bs p,q(Rn) consists of regular distributions p,q(Rn) ⊂ Lp(Rn) for 1 ≤ p ≤ ¥ by Sobolev's embedding. Hence the superposition of f ∈ Bs f ∈ p,q(Rn) ⊂ L1(Rn) for 0 < p ≤ 1 with a 1-diffeomorphism j Bs is defined as the superposition of a regular distribution with a 1-diffeomorphism and is continuous as an operator from Lp(Rn) resp. L1(Rn) to Lp(Rn) resp. L1(Rn) by Lemma 3.13. , then atomic decompositions of f ∈ Bs p,q(Rn) converge to f with respect to the norm of Lp(Rn) for 1 ≤ p < ¥ resp. with respect to the norm of L1(Rn) for 0 < p < 1, see [31, Section 2.12]. Hence the limit does not depend on the choice of the atomic decomposition and is equal to the usual definition of the superposition of a regular distribution f and the 1-diffeomorphism j . If p < ¥ resp. , we use the local convergence of the atomic decompositions of f in L¥ (Rn), i.e. we restrict f and its atomic decomposition to a compact subset K of Rn. Then this restricted atomic decomposition converges to the restricted f with respect to the norm of L¥ (K). This suffices to prove uniqueness of the limit which is an L¥ (Rn)-function. Remark 3.21. For fixed s, p and q the constant c in Theorem 3.16 depends on the r -diffeomorphism j . Looking into the proof of Theorem 3.16 and Remark 3.12 the following definition is useful: Definition 3.22. Let r ≥ 1. We call {j m}m∈N a bounded sequence of r -diffeomorphisms if every j m is a r -diffeomorphism, if there are universal constants c1,c2 > 0 with If p = ¥ c1 ≤ j m(x) − j m(y) x − y ≤ c2 for m ∈ N, x,y ∈ Rn with 0 < x − y ≤ 1 and if - for r > 1 - there is a universal constant c with n(cid:229) i=1 ¶j m i ¶ x j (cid:12)(cid:12) n(cid:229) j=1(cid:13)(cid:13)(cid:13)(cid:13) C r −1(Rn)(cid:13)(cid:13)(cid:13)(cid:13) < c. for m ∈ N. Remark 3.23. If {j m}m∈N is a bounded sequence of r -diffeomorphisms, then (j m)−1 exists for all m ∈ N and {(j m)−1}m∈N is a bounded sequence of r -diffeomorphisms, too. This follows by the arguments of Lemma 3.10. Now, by going through the proof of Theorem 3.16 and Remark 3.12 it follows Corollary 3.24. Let s ∈ R, 0 < q ≤ ¥ and r ≥ 1. (i) Let 0 < p ≤ ¥ and r > max(s,1 +s p − s). If {j m}m∈N is a bounded sequence of r -diffeomorphisms, then there exists a constant C such that k f (j m(·))Bs p,q(Rn)k ≤ c · k f Bs p,q(Rn)k for all f ∈ Bs p,q(Rn) and m ∈ N. then there exists a constant C such that (ii) Let 0 < p < ¥ and r > max(s,1+s p,q −s). If {j m}m∈N is a bounded sequence of r -diffeomorphisms, k f (j m(·))F s p,q(Rn)k ≤ c · k f Fs p,q(Rn)k for all f ∈ Fs p,q(Rn) and m ∈ N. REFERENCES [1] G. Bourdaud and W. Sickel. Changes of variable in Besov spaces. Math. Nachr., 198:19 -- 39, 1999. [2] L. E. J. Brouwer. Zur Invarianz des n-dimensionalen Gebiets. Mathematische Annalen, 72(1):55 -- 56, 1912. [3] W. Dahmen. Wavelet methods for PDEs - some recent developments. Journal of Computational and Applied Mathematics, 128(1-2):133 -- 185, 2001. [4] J. Franke. On the spaces of Triebel-Lizorkin type: pointwise multipliers and spaces on domains. Mathematische Nachrichten, 125:29 -- 68, 1986. [5] M. Frazier and B. Jawerth. Decomposition of Besov spaces. Indiana Univ. Math. Journ., 34:777 -- 799, 1985. [6] M. Frazier and B. Jawerth. A discrete transform and decompositions of distribution spaces. J. Funct. Analysis, 93:34 -- 170, 1990. [7] V. M. Gol'dshtein and Y. G. Reshetnyak. Quasiconformal mappings and Sobolev Spaces. Kluwer Academic Publishers, Dor- drecht, 1990. [8] V. M. Gol'dshtein and A. S. Romanov. Transformations that preserve Sobolev spaces. Sib. Math. Zh., 25(3):55 -- 61, 1984. [9] V. M. Gol'dshtein and S. K. Vodop'yanov. Lattice isomorphisms of the spaces L1 n and quasiconformal mappings. Sib. Math. Zh., 16(2):224 -- 246, 1975. [10] V. M. Gol'dshtein and S. K. Vodop'yanov. Functional characteristics of quasiisometric mappings. Sib. Math. Zh., 17(4):768 -- 773, 1976. 20 BENJAMIN SCHARF [11] J. Kahane and P. Lemarie-Rieusset. Fourier Series and Wavelets. Gordon and Breach Publ., Amsterdam, 1995. [12] H. Koch and W. Sickel. Pointwise multipliers of Besov spaces of smoothness zero and spaces of continuous functions. Rev. Mat. Iberoamericana, 18(3):587 -- 626, 2002. [13] I. G. Markina. A change of variable that preserves the differential properties of functions. Sib. Math. Zh., 31(3):422 -- 432, 1990. [14] V. G. Maz'ya. Weak solutions of the Dirichlet and Neumann problems. Tr. Mosk. Mat. O-va., 20:137 -- 172, 1969. [15] V. G. Maz'ya and T. O. Shaposhnikova. Theory of Multipliers in Spaces of Differentiable Functions. Pittman, Boston-London, 1985. [16] V. G. Maz'ya and T. O. Shaposhnikova. Theory of Sobolev Multipliers. Springer, Berlin-Heidelberg, 2009. [17] V. Netrusov. Theorems on traces and multipliers for functions in Lizorkin-Triebel spaces. Zapiski Nauchnykh Seminarov POMI, 200:132 -- 138, 1992. [18] J. Peetre. New Thoughts on Besov Spaces. Duke Univ. Math. Ser., Durham, 1976. [19] T. Runst and W. Sickel. Sobolev spaces of fractional order, Nemytskij operators, and nonlinear partial differential equations. de Gruyter, Berlin. [20] V. S. Rychkov. On a theorem of Bui, Paluszy´nski, and Taibleson. Tr. Mat. Inst. Steklova, 227:286 -- 298, 1999. [21] B. Scharf. Local means and atoms in vector-valued function spaces. Jenaer Schriften zur Mathematik und Informatik, Math/Inf/05/10, 2010. On ArXiv.org: http://arxiv.org/abs/1103.6159v1. [22] C. Schneider and J. Vybiral. Non-smooth atomic decompositions, traces on Lipschitz domains, and pointwise multipliers in function spaces. submitted, 2012. p,q(Rn) in case s p,q < s < n/p. Ann. Mat. Pura Appl., 174:209 -- 250, 1999. [23] W. Sickel. On pointwise multipliers for Fs [24] W. Sickel. Pointwise multipliers of Lizorkin-Triebel spaces. Operator Theory, Advances Appl., 110:295 -- 321, 1999. [25] L. Skrzypczak. Atomic decompositions on manifolds with bounded geometry. Forum Math., 10(1):19 -- 38, 1998. [26] R. Strichartz. Multipliers on fractional Sobolev spaces. Journal of Mathematics and Mechanics, 16:1031 -- 1060, 1967. [27] H. Triebel. Theory of Function Spaces. Birkhäuser, Basel, 1983. [28] H. Triebel. Theory of Function Spaces II. Birkhäuser, Basel, 1992. [29] H. Triebel. Fractals and Spectra. Birkhäuser, Basel, 1997. [30] H. Triebel. Function spaces in Lipschitz domains and on Lipschitz manifolds. Characteristic functions as pointwise multipliers. Rev. Mat. Complut., 15(2):475 -- 524, 2002. [31] H. Triebel. Theory of Function Spaces III. Birkhäuser, Basel, 2006. [32] H. Triebel. Function spaces and wavelets on domains. Publishing House European Math. Soc., Zürich, 2008. [33] H. Triebel and H. Winkelvoss. Intrinsic atomic characterizations of function spaces on domains. Math. Z., 221(1):647 -- 673, 1996. [34] A. D. Ukhlov and S. K. Vodop'yanov. Superposition operators in Sobolev spaces. Izv. Vyssh. Uchebn. Zaved. Mat., 10:11 -- 33, 2002. [35] S. K. Vodop'yanov. Mappings of homogeneous groups and imbeddings of functional spaces. Sib. Math. Zh., 30:25 -- 41, 1989. [36] A. Zygmund. Smooth functions. Duke Math. J., 12(1):47 -- 76, 1945. E-mail address: [email protected] MATHEMATISCHES INSTITUT, FAKULTÄT FÜR MATHEMATIK UND INFORMATIK, FRIEDRICH-SCHILLER-UNIVERSITÄT JENA, D-07737 JENA, GERMANY
1201.2376
2
1201
2012-12-11T17:07:07
Surfaces Meeting Porous Sets in Positive Measure
[ "math.FA", "math.CA", "math.MG" ]
Let n>2 and X be a Banach space of dimension strictly greater than n. We show there exists a directionally porous set P in X for which the set of C^1 surfaces of dimension n meeting P in positive measure is not meager. If X is separable this leads to a decomposition of X into a countable union of directionally porous sets and a set which is null on residually many C^1 surfaces of dimension n. This is of interest in the study of certain classes of null sets used to investigate differentiability of Lipschitz functions on Banach spaces.
math.FA
math
SURFACES MEETING POROUS SETS IN POSITIVE MEASURE GARETH SPEIGHT Abstract. Let X be a Banach space and 2 < n < dim X. We show there exists a directionally porous set P in X for which the set of C 1 surfaces of dimension n meeting P in positive measure is not meager. If X is separable this leads to a decomposition of X into the union of a σ-directionally porous set and a set which is null on residually many C 1 surfaces of dimension n. This is of interest in the study of Γn-null and Γ-null sets and their applications to differentiability of Lipschitz functions. 1. Introduction We investigate the extent to which C1 surfaces meet (directionally) porous sets (Definition 1.1) in positive measure. By definition each point in a porous set sees nearby holes in the set of size proportional to their distance away. It is thus intu- itively clear that porous sets are somehow small. It follows easily from the definition that porous sets are nowhere dense and hence σ-porous sets are meager. Further, if P is a porous (directionally porous) set in a Banach space then the Lipschitz map x 7→ dist(x, P ) is Fr´echet (Gateaux) differentiable at no point of P . If the Banach space is finite dimensional it follows by the classical Rademacher theorem that P has Lebesgue measure zero. Porous sets have been widely studied (see [1] and [2] for surveys of the area) and recently have been used in the study of differentiability. It has been of much inter- est to what extent an analogue of Rademacher's theorem, either with Gateaux or Fr´echet differentiability, holds for Lipschitz functions defined on infinite dimensional Banach spaces (see [3] for an introduction and [4], [5] for some recent developments relevant to us). Since Lebesgue measure is unavailable in infinite dimensional Banach spaces some other notion of null set is needed to say the set of points of non differentiability of a Lipschitz function is small. The classes of most interest to us are the Γn-null sets (Definition 1.2) and Γ-null sets introduced in [4] and [5] respectively. These are sets which are null on typical n dimensional C1 surfaces (or suitably defined infinite dimensional surfaces for the Γ-null case). In separable Banach spaces there are several well known classes of null sets (for example, Aronszajn null, Haar null and Gauss null sets) with respect to which real valued Lipschitz functions are Gateaux differentiable almost everywhere (Theorem 6.42 [3], also see [6], [7] and [8]). The Γ-null sets also have this property (Theorem 2.5 [5]). These classes of null sets form σ-ideals (in other words they are closed under taking subsets and countable unions) so give a reasonable notion of null set. Since x 7→ dist(x, P ) is Gateaux differentiable at no point of a directionally porous set P it follows σ-directionally porous sets must be null in all of the above senses. A Borel set E ⊂ X is Haar null if there exists a Borel probability measure µ on X such that µ(x + E) = 0 for all x ∈ X. It follows from this definition that any σ-directionally porous set is null on many lines in X. On the other hand, it can This work was done while the author was a PhD student of David Preiss and supported by EPSRC funding. I also thank Jaroslav Tiser for suggesting improvements to the presentation. 1 2 GARETH SPEIGHT be shown (Theorem 6.39 [3], see also [9]) that any infinite dimensional separable Banach space contains a σ-porous set whose complement is null on all lines. Such a σ-porous set is not Haar null. This illustrates that in infinite dimensional spaces directionally porous sets are much smaller than porous sets. In finite dimensions, as one might expect, it follows from compactness of the unit sphere that porous and directionally porous sets coincide. Proving existence of Fr´echet derivatives is much more difficult. The main known result is if a Banach space X has separable dual then every real valued Lipschitz function on X is Fr´echet differentiable on a dense set [10]. However, this is not an 'almost everywhere' type result. Indeed, it is not even known if three real valued Lipschitz functions on a separable Hilbert space have a common point of Fr´echet differentiability [4]. It has recently been shown if X ∗ is separable then every real valued Lipschitz map on X is Fr´echet differentiable outside a Γ-null set if and only if every σ-porous subset of X is Γ-null (Corollary 3.12 [5]). This implies, for example, that any real valued Lipschitz function on c0 or C(K) (for K countable compact) is Fr´echet differentiable outside a Γ-null set (Theorem 4.6 [5]). It is known (Theorem 5.4.2 [4]) that Gδ sets are Γ-null if they are Γn-null for infinitely many n. Thus it is desirable to understand when porous sets are either Γn-null or Γ-null. It can be shown if n ≥ dim X then Γn-null and Lebesgue null sets coincide (Theorem 5.3.8 [4]). If n < dim X the situation is much more interesting. Every σ-porous subset of a Banach space with separable dual is Γ1-null (Theorem 10.4.1 [4]) and every σ-directionally porous subset of a separable Banach space is Γ1- null and Γ2-null (Theorem 10.4.2 [4]). We show the situation is very different for higher dimensional surfaces - even directionally porous sets need not be Γn null when 2 < n < dim X (Theorem 1.4) and if X is separable the complement of a σ-directionally porous set may be Γn-null (Theorem 1.5). One might ask if there is any differentiability result using only the notion of Γn-null sets rather than Γ-null sets. It has been shown (Theorem 11.3.6 [4]) that if X ∗ is separable and every porous set in X can be decomposed into the union of a σ-directionally porous set and a Γn-null set of class Gδ, then Lipschitz maps on X into Banach spaces of dimension not exceeding n have points of ε-Fr´echet differentiability for every ε > 0. At the time it was not yet known if porous sets in some infinite dimensional spaces were necessarily Γn-null. By showing even directionally porous sets in infinite dimensional Banach spaces need not be Γn-null we answer a question posed in [4] (pages 186 and 203) and show that for the theorem mentioned to be meaningful the more complicated hypothesis is necessary. That is, it could not instead be more simply assumed porous sets are Γn-null. We now give the formal definitions that will be relevant for us. In what follows B(x, r) will denote the open ball in X with centre x ∈ X and radius r > 0. Definition 1.1. A set P ⊂ X is called porous if there exists 0 < ρ < 1 such that for all x ∈ P and δ > 0 there exists y ∈ X with ky − xk < δ such that B(y, ρky − xk) ∩ P = ∅. We refer to the ball B(y, ρky − xk) as a hole in P . A set P ⊂ X is called directionally porous if there exists 0 < ρ < 1 such that for all x ∈ P there exists v ∈ X with kvk = 1 such that for any δ > 0 there exists t ∈ R with t < δ such that We refer to the constant ρ appearing in the above definitions as a porosity constant of P . B(x + tv, ρt) ∩ P = ∅. SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 3 A set is called σ-porous (σ-directionally porous) if it is a union of countably many porous (directionally porous) sets. Recall a subset of a metric space is called typical, or residual, if its complement is meager. Definition 1.2. A C1 surface of dimension n in X is a C1 map f : [0, 1]n → X. We define the C1 norm by kfkC 1 = max(kfk∞,k∂f/∂x1k∞, . . . ,k∂f/∂xnk∞) where k · k∞ denotes the supremum norm. Denote the space of C1 surfaces of dimension n in X with C1 norm by Γn(X). A set N ⊂ X is called Γn-null if Hn(f ([0, 1]n) ∩ N ) = 0 for residually many f ∈ Γn(X). Remark 1.3. In [4] a set N ⊂ X is defined to be Γn-null if Ln{t ∈ [0, 1]n : f (t) ∈ N} = 0 for residually many f ∈ Γn(X). It is easy to see a set which is Γn-null in this sense is also Γn-null in the sense of Definition 1.2. In fact the two definitions are equivalent for n ≤ dim X. This follows from Lemma 5.3.5 [4] which states that for a typical surface f ∈ Γn(X) the derivative df has rank equal to min(n, dim X) at Lebesgue almost every point of [0, 1]n. We can now state precisely the results of this paper. Theorem 1.4. Let 2 < n < dim X. Then there exists a directionally porous set P ⊂ X which is not Γn-null. Theorem 1.5. Let X be separable and 2 < n < dim X. Then there exists a σ-directionally porous set Q ⊂ X and a Γn-null set N ⊂ X such that X = Q ∪ N . Intuitively the reason for the difference between the cases n ≤ 2 and n > 2 is that for n > 2 modifying a surface to go through a nearby hole causes an area change comparable to the size of the hole (this is stated precisely in Proposition 4.2 which is a corollary of a Poincar´e type inequality). By a careful construction we exploit this fact to construct a porous set where the size of holes a surface meets is controlled by the area of the surface. Fix n > 2 throughout the remainder of the paper. We focus on proving Theorem 1.4 in the case X = Rn+1. Since porous sets and directionally porous sets coincide in finite dimensions it suffices to construct a porous set in Rn+1 which is not Γn- null. In the final section we deduce the result for a general Banach space and see Theorem 1.5 follows. 2. Geometric conditions for the construction Let A be a subset of a complete metric space M and Uk be a sequence of open k=1 Uk ⊂ A then U \ A is meager in M . Hence by the Baire Category Theorem A is not meager in M . We now investigate what this means in terms of surfaces and porous sets. sets which are each dense in a fixed open set U ⊂ M . IfT∞ Define the C1 surface p : [0, 1]n → Rn+1 by p(x) = (x, 0) and and let c = (1/2, . . . , 1/2) ∈ Rn. Fix 0 < r < 1/32 to be chosen small later. If g : B(c, t) → R let G(g) = {(x, g(x)) : x ∈ B(c, t)} denote the graph of g. It will usually be simpler to work with surfaces represented as graphs. 4 GARETH SPEIGHT Lemma 2.1. There exists s > 0 (fixed and independent of r) and δ(r) > 0 such that for all f ∈ BC 1 (p, δ(r)) there exists g : B(c, s) → R of class C1 with kgkC 1 ≤ r and f ([0, 1]n) ∩ (B(c, s) × R) = G(g). Proof. Let e1, . . . , en denote the standard basis of Rn. We apply the Inverse Func- depend only on bounds on its derivative (this is clear from the proof of the Inverse Function Theorem given in [11]). tion Theorem to ef = (f1, . . . , fn) which is a C1 mapping from [0, 1]n to Rn and satisfies k∂ef /∂xi− eik∞ < δ(r) for i = 1, . . . , n. Note the size of the neighbourhood on which ef is invertible can be made independent of the particular function f and Provided δ(r) is sufficiently small (independently of f ) ef is invertible on a region whose image contains c and has a C1 inverse (ef )−1 : B(c, s) → Rn with s > 0 independent of r, f and k∂(ef )−1/∂xik∞ ≤ 2 for i = 1, . . . , n. Define the C1 map g : B(c, s) → R by g = fn+1 ◦ (ef )−1. Since kfn+1kC 1 < δ(r) we have kgkC 1 ≤ r for Let F be a (not necessarily disjoint) collection of open balls in Rn+1 and L > 1. Intuitively F will correspond to the holes of a porous set. Define, for k ≥ 1, (2.1) sufficiently small δ(r). B, (cid:3) Pk = [B∈F diamB<1/k LB, H = [B∈F where LB denotes the ball with the same centre as B but radius enlarged by factor L. Notice Pk is open and Pk+1 ⊂ Pk for all k ≥ 1. Further P = ∞\k=1 Pk! \ H is porous with porosity constant 1/L. InformallyT∞ Given A ⊂ Rn+1 denote the set of n dimensional surfaces which meet A in n k=1 Pk consists of points which have a nearby hole, of radius proportional to its distance from the point, of diameter less than 1/k and H is the union of all the holes (of any size). dimensional measure greater than α ≥ 0 by Sn(A, α) = {f ∈ Γn(X) : Hn(A ∩ f ([0, 1]n)) > α}. To prove Theorem 1.4 in the case X = Rn+1 it suffices to construct F and L > 1 for which there is α > 0 such that, for sufficiently small r, Sn((B(c, s) × R) ∩ Pk, α) is dense in BC 1(p, δ(r)) for all k ≥ 1 (2.2) and (2.3) As the sets Pk are decreasing it follows Hn(f ([0, 1]n) ∩ (B(c, s) × R) ∩ H) ≤ α/4 for all f ∈ BC 1(p, δ(r)). (B(c, s) × R) ∩ Pk, α/2! . ∞\k=1 Sn((B(c, s) × R) ∩ Pk, α) ⊂ Sn ∞\k=1 Since the sets Pk are open each of the sets Sn((B(c, s) × R) ∩ Pk, α) is also open. By (2.2) they are also dense in BC 1(p, δ(r)) so it follows, by the discussion at the start of this section, that the complement of Sn ∞\k=1 (B(c, s) × R) ∩ Pk, α/2! SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 5 is meager in BC 1 (p, δ(r)). Then by (2.3), Sn ∞\k=1 (B(c, s) × R) ∩ Pk, α/2! ⊂ Sn(P, α/4) so the complement of Sn(P, α/4) is meager in BC 1(p, δ(r)). By the Baire Category Theorem, as in the discussion at the start of this section, this will prove Theorem 1.4 in the case Rn+1. Thus proving Theorem 1.4 amounts to constructing smaller and smaller open balls (intuitively the holes of the porous set) whose enlargements mostly cover surfaces in a countable dense set and so that the intersection of all balls with any one surface is kept small. 3. Construction of the porous set One of the requirements on F is that the enlargements of balls in F of arbitrarily small radii must cover a fixed proportion of each surface in a countable dense subset of BC 1(p, δ(r)). To do this it is natural to choose our countable dense subset to be as simple as possible. The following lemma follows from Corollary 10.2.2 [4], which states surfaces in which almost every point has a neighbourhood on which the surface is affine are dense in Γn(Rn+1), and allows us to choose a countable dense subset consisting of surfaces that are mostly covered by countably many planes. Let ωn denote the volume of a unit ball in Rn. Lemma 3.1. There is a countable dense set of surfaces {fl}∞ l=1 in BC 1(p, δ(r)) and a countable collection of affine planes Al(x) = (x, al(x)), where x ∈ [0, 1]n, ∇al is constant and ∇al ≤ r, such that Al([0, 1]n)! < ωnsn/4. fl([0, 1]n) \ Hn ∞[l=1 ∞[l=1 Without loss of generality let f1 = A1 = p. We will define F and L > 1 so that, if Pk is defined as in (2.1), Hn(Al([0, 1]n) ∩ (B(c, s) × R) \ Pk) ≤ ωnsn/2l+2 for all k, l ≥ 1. (3.1) This will imply Hn(fl([0, 1]n) ∩ (B(c, s) × R) ∩ Pk) > ωnsn/2 for all k, l ≥ 1 and hence fl ∈ Sn((B(c, s) × R) ∩ Pk, ωnsn/2) for all k, l ≥ 1 which implies (2.2) with α = ωnsn/2. In Section 5 we then show that if r is sufficiently small our construction ensures and 3P∞ Hn(f ([0, 1]n) ∩ (B(c, s) × R) ∩ H) ≤ ωnsn/8 for all f ∈ BC 1(p, δ(r)) which is (2.3) with α = ωnsn/2. Theorem 1.4 hence follows in the case X = Rn+1. Choose a sequence mk of natural numbers with m1 = 1 in which every natural number is repeated infinitely many times. Fix a sequence εi > 0 with εi < 1/2i i=1 εi ≤ 1/64 to be chosen as small as required later. Let r0 = s. For k ≥ 1 we will inductively define Fk and rk > 0 such that Fk consists of finitely many balls in Rn+1, of radius less then ε3 krk−1 and greater than rk, whose enlargements (by a factor independent of k) mostly cover Amk ([0, 1]n). The family Fk will consist of subfamilies of balls on different levels relative to Amk constructed so there are relatively few balls on any one level. Each of these subfamilies of Fk is formed by lifting families of balls Gl and hence define Fk and rk. Let G0 k = rk−1. k in Rn to Rn+1. Fix k ≥ 1 for which rk−1 has been defined. We show how to define Gl k = ∅ and r0 k inductively 6 GARETH SPEIGHT k of finitely many balls in Rn Lemma 3.2. Fix l ≥ 0. Suppose families G0 and r0 k, . . . ,Gl k > 0 have been constructed so that (if l ≥ 1): k consists of balls of radius rq k, . . . , rl • For each 1 ≤ q ≤ l, Gq • For each 1 ≤ q ≤ l, the balls in {(1/ε3 k)B : B ∈ Gq k} are disjoint and contained inside B(c, s). k ≤ ε3 krq−1 k . • Any two balls from {(1/ε3 k)B : B ∈ G1 k ∪ . . . ∪ Gl k} are either disjoint or one is contained inside the other. Then there exists a family Gl+1 ε3n k /2n+1 proportion of k of finitely many balls in Rn which cover at least B(c, s) \ [B∈G1 k∪...∪Gl k B and 0 < rl+1 krl k ≤ ε3 • Balls in Gl+1 • The balls in k k such that: have radius rl+1 k . k)B : B ∈ Gl+1 k } are disjoint and contained inside B(c, s). {(1/ε3 • Any two balls from {(1/ε3 k)B : B ∈ G1 k ∪ . . . ∪ Gl+1 k } are either disjoint or one is contained inside the other. Proof. Find 0 < rl+1 krl k ≤ ε3 k such that at least half of the points in B(c, s) \ [B∈G1 k∪...∪Gl k B are of distance at least rl+1 k /ε3 k away from the closed set (Rn \ B(c, s)) ∪ [B∈G1 k∪...∪Gl k ∂((1/ε3 k)B). k from centres of previously chosen balls, we can find a finite family Gl+1 k , whose centres are of distance at least of k } are disjoint, contained k)B : B ∈ Gl+1 k /ε3 By inductively choosing balls of radius rl+1 2rl+1 balls of radius rl+1 in k k such that balls in {(1/ε3 B(c, s) \ [B∈G1 k∪...∪Gl k ∂((1/ε3 k)B) and balls in {(2/ε3 k)B : B ∈ Gl+1 k } cover at least half of B(c, s) \ [B∈G1 k∪...∪Gl k B. SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 7 Then Ln [B∈Gl+1 k as required. B = XB∈Gl+1 k = (ε3 k)B) Ln((2/ε3 Ln(B) k/2)n XB∈Gl+1 k/2)nLnB(c, s) \ [B∈G1 k ≥ (ε3 B /2 (cid:3) k∪...∪Gl k Once Gl k has been defined if LnB(c, s) \ [B∈G1 k∪...∪Gl k B ≤ ωnsn/2k+3 we stop, set rk = rl k and let Gk = G1 Otherwise we continue and define Gl+1 as in Lemma 3.2. Since at each stage we cover a fixed proportion (independent of l) of the region not yet covered this process stops in a finite number of steps. To define Fk from Gk we replace each ball k ∪ . . . ∪ Gl k. and rl+1 k k by B(x, t) ∈ Gk B((x, amk (x) + 2t), t) ∈ Fk. Inductively define Fk and rk as above for all k ≥ 1, then let F = ∪∞ L = √10. Proposition 3.3. If Pk are defined from F , L as in (2.1) then k=1Fk and Hn(Amk ([0, 1]n) ∩ (B(c, s) × R) \ Pk) ≤ ωnsn/2k+2 for all k ≥ 1. Proof. Notice if B(x, t) ∈ Gk then, since ∇amk ≤ 1, Hence Amk ([0, 1]n) ∩ (B(x, t) × R) ⊂ B((x, amk (x) + 2t), t√10) ⊂ Pk. Hn(Amk ([0, 1]n) ∩ (B(c, s) × R) \ Pk) ≤ Hn Amk ([0, 1]n) ∩ B(c, s) \ [B∈Gk ≤ 2Ln B(c, s) \ [B∈Gk B! ≤ ωnsn/2k+2. B! × R! (cid:3) Since the sequence mk takes the value of every natural number infinitely often and the sets Pk are decreasing, Proposition 3.3 implies (3.1). 8 GARETH SPEIGHT 4. Area estimates for surfaces BC 1(p, δ(r)) is small. It remains to show the intersection of balls in F with any fixed surface in We now establish several results which later allow us to show how passing through many holes forces the area of a surface f to increase and to distinguish between area increments arising from holes of different sizes. The following lemma is an adaptation of Theorem 1(iii) in Section 5.6.1 of [12]. The proof is essentially the same as in [12] but with the numbers 1 and n/(n − 1) replaced by 2 and its corresponding Sobolev conjugate 2n/(n − 2). Notice the assumption n > 2 is essential. In what follows C will be a positive constant (depending only on n) whose value may be different in different expressions. We will sometimes write 2C instead of C if it is important but not obvious an extra constant term has been added. Lemma 4.1. For each 0 < α ≤ 1 there exists a constant C(α) > 0 such that for all balls B ⊂ Rn and g ∈ W n,2 kgkL2n/(n−2)(B) ≤ C(α)k∇gkL2(B) loc (Rn) satisfying Ln(B ∩ {g = 0}) Ln(B) ≥ α. ≥ α. Proof. Suppose B = B(x, t) ⊂ Rn and g ∈ W n,2 Ln(B ∩ {g = 0}) loc (Rn) satisfies By the triangle inequality Ln(B) kgkL2n/(n−2)(B) ≤ kg − (g)BkL2n/(n−2)(B) + (g)B(Ln(B))(n−2)/2n where (g)B = 1 Ln(B)ZB g. We estimate the two terms individually. Since n > 2, Poincar´e's inequality (Theorem 2 of Section 4.5.2 [12]) states, (cid:18)−ZB g − (g)B2n/(n−2)(cid:19)(n−2)/2n ≤ Ct(cid:18)−ZB ∇g2(cid:19)1/2 . As Ln(B) = ωntn this implies Next notice kg − (g)BkL2n/(n−2)(B) ≤ Ck∇gkL2(B). (g)B(Ln(B))(n−2)/2n ≤ (Ln(B))−(n+2)/2nZB gχ{g6=0} where χA denotes the characteristic function of a set A. Applying Holder's inequal- ity with exponents 2n/(n − 2) and 2n/(n + 2) implies (g)B(Ln(B))(n−2)/2n ≤ (Ln(B))−(n+2)/2nkgkL2n/(n−2)(B)(Ln(B∩{g 6= 0}))(n+2)/2n. The assumption on g then yields (g)B(Ln(B))(n−2)/2n ≤ (1 − α)(n+2)/2nkgkL2n/(n−2)(B). Putting the two estimates together implies kgkL2n/(n−2)(B) ≤ Ck∇gkL2(B) + (1 − α)(n+2)/2nkgkL2n/(n−2)(B). Rearranging this expression and relabelling constants leads to the desired inequality. (cid:3) SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 9 Our use of Lemma 4.1 is expressed in the following proposition. The underlying idea will be that if a surface passes through different vertical levels it creates an area change comparable to holes at those heights. Proposition 4.2. Suppose h > 0 and B ⊂ Rn is a ball of radius at least h. Then ωnhn ≤ CZB∩{g≥h/2} ∇g2 for all g : B → R of class C1 with ∇g ≤ 1 such that Ln(B ∩ {g ≤ h/2}) ≥ 1/2 Ln(B) and g(ex) ≥ h for some ex ∈ B. Proof. Let g : B → R be as in the statement of the proposition. Extend g to a function in W n,2 loc (Rn) and let (g − h/2)+ denote the function which equals g − h/2 if g ≥ h/2 and is zero otherwise. Then (g−h/2)+ belongs to W n,2 loc (Rn) and satisfies Ln(B ∩ {(g − h/2)+ = 0}) Ln(B) ≥ 1/2. Hence by Lemma 4.1 k(g − h/2)+kL2n/(n−2)(B) ≤ C(1/2)k∇(g − h/2)+kL2(B). Since g is C1, ∇g ≤ 1 in B, g(ex) ≥ h and B has radius at least h it follows there is C > 0 such that g ≥ 3h/4 inside a region of measure at least Chn and contained in B. Hence ((h/4)2n/(n−2)Chn)(n−2)/2n ≤ C(1/2) ZB∩{g≥h/2} ∇g2!1/2 . Simplifying this expression and relabelling the constants gives the claimed inequal- ity. (cid:3) Our holes were constructed relative to planes with varying directions. The fol- lowing proposition will later allow us to use the previous result in this context. Proposition 4.3. Let B = B(x, t) ⊂ Rn, g : B → R of class C1 with ∇g ≤ 1 and a : B → R of class C1 with ∇a constant and ∇a ≤ 1. Suppose ε > 0 and g − a ≤ εt in B. Then Proof. Write g = (g − a) + a and expand the terms to obtain ∂a ∂xi (cid:12)(cid:12)(cid:12)(cid:12)ZB (∇g2 − ∇a2 − ∇(g − a)2)(cid:12)(cid:12)(cid:12)(cid:12) ≤ CεLn(B). (∇g2 − ∇a2 − ∇(g − a)2)(cid:12)(cid:12)(cid:12)(cid:12) = 2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZB nXi=1 nXi=1(cid:12)(cid:12)(cid:12)(cid:12)ZB ≤ C ≤ CεLn(B). (cid:12)(cid:12)(cid:12)(cid:12)ZB ∂(g − a) ∂xi (cid:12)(cid:12)(cid:12)(cid:12) ∂(g − a) ∂xi (cid:12)(cid:12)(cid:12)(cid:12)(cid:12) using the Divergence Theorem and the assumption g − a ≤ εt. (cid:3) In what follows we will need to smooth our surfaces in order to distinguish between oscillations arising from different sized holes. The following definition recalls some basic notions [12]. 10 GARETH SPEIGHT Definition 4.4. Define the C∞ function η : Rn → R by η(x) = κ exp(1/(x2 − 1)) for x < 1 and η(x) = 0 for x ≥ 1, where κ > 0 is chosen so that ZRn η(x) dx = 1. For ε > 0 define the standard mollifier ηε : Rn → R by ηε(x) = (1/εn)η(x/ε). If g ∈ L1 loc(U ) and ε > 0 define gε(x) =ZB(x,ε) ηε(x − y)g(y) dy for x ∈ Uε = {x ∈ U : dist(x, ∂U ) > ε}. We informally refer to gε as a smoothing of g. The following facts are either well known [12] or easy to prove using the definition. Lemma 4.5. Let 0 < ε < t and g : B(x, t) → R be of class C1 and satisfy ∇g ≤ 1 in B(x, t). Then gε is of class C∞(B(x, t − ε)) and gε − g ≤ ε. The following proposition gives an estimate on the smoothing of a function whose values are inside a small interval. Proposition 4.6. Let B = B(x, t) ⊂ Rn and 0 < ε < 1/2. Suppose g : B → R is of class C1 with ∇g ≤ 1 and g ≤ ε2t in B. Then Proof. For x ∈ B(x, t − εt), using the formula for the derivative of a convolution, ZB(x,t−εt) ∇(gεt)2 ≤ Cε2Ln(B). (x − y)g(y) dy(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) (x)(cid:12)(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ZB(x,εt) ≤ZB(x,εt) ≤ ωnεntn(C/εn+1tn+1)ε2t = Cε. (C/εn+1tn+1)ε2t dy ∂ηεt ∂xi ∂gεt ∂xi (cid:12)(cid:12)(cid:12)(cid:12) Hence ∇(gεt) ≤ Cε in B(x, t − εt) so the desired inequality follows. (cid:3) We will need to smooth locally at different scales in different regions. The follow- ing proposition gives a function which allows us to interpolate smoothly between two functions. Proposition 4.7. Let B(x, t) ⊂ Rn and 0 < ε < 1/2. Then there exists a function w : Rn → R of class C1 with 0 ≤ w ≤ 1 such that: • w = 1 inside B(x, t − 2εt). • w = 0 outside B(x, t − εt). • ∇w ≤ 3/(εt) in B(x, t − εt) \ B(x, t − 2εt). Suppose g1 : B(x, t − εt) → R and g2 : B(x, t) → R are C1 with bounded deriva- Then the function v : B(x, t) → R defined to be wg1 + (1 − w)g2 in B(x, t − εt) tives and g1 − g2 ≤ ε2t in B(x, t − εt) \ B(x, t − 2εt). and g2 in B(x, t) \ B(x, t − εt) is C1 with ∇v ≤ max(∇g1,∇g2) + 3ε in B(x, t − εt). Proof. Define ew : Rn → R such that: • ew = 1 inside B(x, t − 5εt/3). SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 11 • ew = 0 outside B(x, t − 4εt/3). • ew interpolates between 0 and 1 as a linear function of distance to x. It is easy to see the function w = ewεt/3 then has the properties required. ∇(wg1 + (1 − w)g2) = w∇g1 + (1 − w)∇g2 + ∇w(g1 − g2) Clearly in B(x, t − εt) so in B(x, t − εt). Since w is C1 and equal to zero outside B(x, t − εt) it also follows v is C1 on B(x, t). (cid:3) ∇v ≤ max(∇g1,∇g2) + 3ε 5. Holes are small on surfaces We now need to show (2.3) with α = wnsn/2 which intuitively states that the intersection of surfaces in BC 1(p, δ(r)) with (B(c, s)× R)∩H is small. To do this we control the intersection of any surface f ∈ BC 1(p, δ(r)), which is the graph of some function g over B(c, s), with H usingRB(c,s) ∇g2. This integral is approximately the n dimensional area difference between the image of f over B(c, s) and the flat disc B(c, s)×{0}. The intuition is thus that passing through holes forces the surface to oscillate in a way that increases its area. If x ∈ Rn+1 let x′ ∈ Rn be the point consisting of the first n coordinates of x. If B ⊂ Rn+1 is a ball of radius t > 0 let B = ωntn be the area of an n dimensional cross section of B. Proposition 5.1. There is a constant C ≥ 1 such that B ≤ CZB(c,s) ∇g2 + C for all g : B(c, s) → R of class C1 satisfying kgkC 1 ≤ 1/64. Proof of Proposition 5.1. To prove the proposition we will reformulate it in a way that allows us to use induction (Lemma 5.5 and Lemma 5.6). At times in the argument we will interpolate between pieces of the surface which are smoothed on different scales. Thus it is natural to control the intersection of surfaces with slightly enlarged holes and begin with weaker bounds on ∇g which we tighten after each induction step. For k ≥ 1 let Kk = 2 −Pk i=1(1/2)i. If B = B(x, t) ∈ Fk for some k ≥ 1 denote k and B′ = B(x′, t′) ⊂ Rn. Note this definition makes sense because any t′ = t/ε3 ball can be in Fk for at most one k ≥ 1. Given g : B(c, s) → R of class C1 with ∇g ≤ 1/32 if B ∈ F and G(g) ∩ K1B 6= ∅ define XB∈F ∞Xi=1 G(g)∩B6=∅ εi Rk(B) = B′ ∩ {g − amk > t/4} ⊂ B(c, s). Note it will be clear from the context which map g is meant when Rk(B) appears. Claim 5.2. If k ≥ 1 is fixed and B1, B2 ∈ Fk are distinct balls with G(g)∩ K1B1 6= ∅ and G(g) ∩ K1B2 6= ∅ then Rk(B1) ∩ Rk(B2) = ∅. Proof of Claim 5.2. Let B1 = B(x, t1) and B2 = B(y, t2) ∈ Fk. since {(1/ε3 2 = ∅ so If there exists l ≥ 1 such that both B(x′, t1) and B(y′, t2) belong to Gl k} is a disjoint family, B′ k)B : B ∈ Gl k then, 1 ∩ B′ Rk(B1) ∩ Rk(B2) = ∅. Suppose not and t1 ≥ t2. Then, since radii of balls in different families Gl k differ kt1 ≥ t2. Note, from the definition of Fk in terms k, necessarily ε3 by factor at least ε3 12 GARETH SPEIGHT of Gk, that if B(z, t) ∈ Fk then zn+1 − amk (z′) = 2t. Since G(g) ∩ K1B2 6= ∅ there exists ey ∈ B(y′, t2) such that Hence g(ey) − amk (ey) ≤ 4t2. g − amk ≤ 4t2 + 2t′ 2/16 ≤ t1/4 in B′ has diameter 2t′ 2 since ∇g ≤ 1/32 and ∇amk ≤ 1/32 implies ∇(g − amk ) ≤ 1/16 while B′ 2 2 = ∅ which again implies 2. Hence Rk(B1) ∩ B′ as required. Rk(B1) ∩ Rk(B2) = ∅ For each k ≥ 1 we define two subfamilies of Fk. and F u k = {B ∈ Fk : G(g) ∩ K1B 6= ∅,B ≤ εkLn(Rk(B))} F d k = {B ∈ Fk : G(g) ∩ K1B 6= ∅,B > εkLn(Rk(B))}. Notice {B ∈ Fk : G(g) ∩ K1B 6= ∅} = F u Intuitively F u k consists of holes where the surface stays mostly on the same level around the hole. Since there are relatively few holes on any one level we can show the holes in this family are small. k ∪ F d k . Claim 5.3. For each k ≥ 1, XB∈F u k B ≤ εk. Proof of Claim 5.3. By definition B ≤ εkLn(Rk(B)) for all B ∈ F u Claim 5.2, k . Hence, using (cid:3) (cid:3) (cid:3) B ≤ εk XB∈F u k Ln(Rk(B)) ≤ εkLn [B∈F u k Rk(B) ≤ εk. XB∈F u Intuitively F d k k consists of holes where the surface goes to other levels around the hole. Using results from Section 4 we can show this forces an increase in relative area. Claim 5.4. For each k ≥ 1 and B ∈ F d k , B ≤ CZRk(B) ∇(g − amk )2. Proof of Claim 5.4. Notice if B = B(x, t) ∈ F d k then Ln(B′ ∩ {g − amk ≤ t/4}) = Ln(B′ \ Rk(B)) Ln(B′) Ln(B′) ≥ 1/2. Proposition 4.2 with h = t/2, Since G(g) ∩ B 6= ∅ there existsex ∈ B′ such that g(ex)− amk (ex) ≥ t/2. Hence, by B ≤ CZRk(B) ∇(g − amk )2. Lemma 5.5. Fix g : B(c, s) → R of class C1 satisfying ∇g ≤ 1/32. Then XB∈F1 G(g)∩K1B6=∅ B ≤ CZB(c,s) ∇g2 + ε1. SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 13 Proof of Lemma 5.5. Since Am1 = A1 = p and so am1 = 0 Claim 5.4 implies B ≤ CZ∪B∈Fd 1 R1(B) ∇g2 ≤ CZB(c,s) ∇g2 XB∈F d 1 which, together with Claim 5.3, proves the lemma. (cid:3) Lemma 5.6. Suppose we have shown G(g)∩KkB6=∅ XB∈F1∪...∪Fk B ≤ CZB(c,s) ∇g2 + C kXi=1 for all g : B(c, s) → R of class C1 satisfying ∇g ≤ 1/32 − 3Pk k+1Xi=1 B ≤ CZB(c,s) ∇g2 + C XB∈F1∪...∪Fk+1 Then εi εi G(g)∩Kk+1B6=∅ i=1 εi. i=1 εi. Further, the constants can be chosen so as to remain bounded as the lemma is for all g : B(c, s) → R of class C1 satisfying ∇g ≤ 1/32 − 3Pk+1 repeatedly applied for all k ∈ N. Proof of Lemma 5.6. Fix g : B(c, s) → R of class C1 satisfying ∇g ≤ 1/32 − 3Pk+1 i=1 εi. We show the measure of holes in Fk+1 that the graph of g meets is controlled by the area difference between g and a smoothing eg of g. Since the smoothing is only done on small scales the measure of holes in F1 ∪ . . . ∪ Fk that g meets will be controlled by the area ofeg. Claim 5.7. There exists eg of class C1 with g −eg ≤ εk+1rk and ∇eg ≤ 1/32 − 3Pk B ≤ CZB(c,s) i=1 εi such that XB∈Fk+1 G(g)∩Kk+1B6=∅ (∇g2 − ∇eg2) + Cεk+1. Proof of Claim 5.7. Recall from Claim 5.3, XB∈F u k+1 of F d other we can choose a subfamily F s are disjoint and for each ball B1 ∈ F d k+1 there exists B ∈ F s Since balls in {B′ : B ∈ Fk+1} are either disjoint or one is contained inside the k+1} 1 ⊂ B′. k+1 such that balls in {B′ : B ∈ F s k+1 such that B′ B ≤ εk+1. k+1 Notice, by definition of F d k+1, for each ball B ∈ F s k+1 Ln(B′ ∩ {g − amk+1 > ε3 k+1t′/4}) = Ln(B′ ∩ Rk+1(B)) ≤ B/εk+1 ≤ ωn(ε2 k+1t′)n which, since ∇(g − amk+1) ≤ 1/16, implies g − amk+1 ≤ ε2 k+1t′/2 in B′. 14 GARETH SPEIGHT Hence by Claim 5.4 and Proposition 4.3, XB1∈F d k+1 1⊂B′ B′ B1 ≤ C XB1∈F d k+1 1⊂B′ B′ ZRk+1(B1) ∇(g − amk+1 )2 ≤ CZB′ ∇(g − amk+1)2 ≤ CZB′ Defineeg to be equal to g outside B′, [B∈F s k+1 (∇g2 − ∇amk+12) + Cεk+1Ln(B′). k+1t′ ≤ εk+1rk. k+1t′ in B(x′, t′ − εk+1t′). i=1 εi in B(x′, t′− i=1 εi. ) ≤ 1/32−3Pk+1 g − amk+1 ≤ ε2 k+1t′ in B(x′, t′ − εk+1t′), for B = B(x, t) ∈ F s smoothly as in Proposition 4.7 between g and gεk+1t′ 2εk+1t′). k+1 leteg be equal to gεk+1t′ in B(x′, t′−2εk+1t′) and interpolate in B(x′, t′− εk+1t′)\ B(x′, t′− k+1t′/2 in B′ and smoothing leaves an affine plane un- − amk+1 ≤ ε2 k+1t′/2 in B(x′, t′ − εk+1t′). Hence g − i=1 εi implies ∇(gεk+1t′ Since g − amk+1 ≤ ε2 changed it follows gεk+1t′ gεk+1t′ ≤ ε2 Also ∇g ≤ 1/32−3Pk+1 εk+1t′). Proposition 4.7 hence implieseg is of class C1 with ∇eg ≤ 1/32− 3Pk Further, g −eg ≤ ε2 By Proposition 4.3 and Proposition 4.6, since eg − amk+1 ≤ εk+1t′ in B′ and (cid:12)(cid:12)(cid:12)(cid:12)ZB′ ∇eg2 −ZB′ ∇amk+12(cid:12)(cid:12)(cid:12)(cid:12) ≤ZB′ ∇(eg − amk+1)2 + Cεk+1Ln(B′) ≤ZB(x′,t′−2εk+1t′)(cid:12)(cid:12)(cid:12)∇(cid:16)(g − amk+1 )εk+1t′(cid:17)(cid:12)(cid:12)(cid:12) k+1 XB1∈F d k+1(cid:18)ZB′ k+1(cid:18)ZB′ (∇g2 − ∇eg2) + Cεk+1 B + XB∈F u (∇g2 − ∇amk+12) + Cεk+1Ln(B′)(cid:19) + εk+1 (∇g2 − ∇eg2) + 2Cεk+1Ln(B′)(cid:19) + εk+1 Putting these estimates together and using the definition of F s XB∈Fk+1 B ≤ XB∈F s ≤ C XB∈F s ≤ C XB∈F s ≤ CZB(c,s) ≤ Cεk+1Ln(B′). k+1 gives B k+1 as required. Kk+1B∩G(g)6=∅ k+1 1⊂B′ B′ (cid:3) 2 + 2Cεk+1Ln(B′) Recall balls in F1 ∪ . . . ∪ Fk have radius at least rk. Since g −eg ≤ εk+1rk ≤ (Kk − Kk+1)rk SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 15 it follows G(g) ∩ Kk+1B 6= ∅ implies G(eg) ∩ KkB 6= ∅ for B ∈ F1 ∪ . . . ∪ Fk. Applying the induction assumption toeg and Claim 5.7 gives, XB∈F1∪...∪Fk+1 εi + CZB(c,s) B + XB∈Fk+1 G(g)∩Kk+1B6=∅ G(g)∩Kk+1B6=∅ B G(eg)∩KkB6=∅ B ≤ XB∈F1∪...∪Fk ≤ CZB(c,s) ∇eg2 + C ≤ CZB(c,s) ∇g2 + C (∇g2 − ∇eg2) + Cεk+1 kXi=1 k+1Xi=1 εi. It is clear from above that the constants remain bounded as Lemma 5.6 is re- (cid:3) peatedly applied. together prove the proposition. i=1 εi ≤ 1/64 and Kk > 1 for all k ≥ 1 Lemma 5.5 and Lemma 5.6 (cid:3) Since 3P∞ Proposition 5.1 finally proves Theorem 1.4 in the case X = Rn+1. To see this notice the proposition implies that for 0 < r ≤ 1/64 Hn(f ([0, 1]n) ∩ (B(c, s) × R) ∩ H) ≤ XB∈F ≤ Csnr2 + C f ([0,1]n)∩B6=∅ Hn(f ([0, 1]n) ∩ B) ∞Xi=1 εi for all f ∈ BC 1(p, δ(r)). Provided εi and r are chosen sufficiently small this implies Hn(f ([0, 1]n) ∩ (B(c, s) × R) ∩ H) ≤ ωnsn/8 for all f ∈ BC 1 (p, δ(r)) which is (2.3) with α = ωnsn/2. 6. Decomposition of Banach spaces We now establish the general case of Theorem 1.4 and prove Theorem 1.5. We have shown there exists β > 0, 0 < R < 1 and a directionally porous set P ⊂ Rn+1 such that {f ∈ BC 1(p, (n + 1)R) : Hn(f ([0, 1]n) ∩ P ) < (n + 1)nβ} is meager, where p : [0, 1]n → Rn+1 is the plane p(x) = (x, 0). Let X be a (possibly infinite dimensional) Banach space satisfying dim X > n. The main idea used to prove Theorem 1.4 and Theorem 1.5 will be that a C1 surface of dimension n can be locally approximated by n dimensional affine planes. We use linear maps sending n dimensional affine planes to the plane p in Rn+1 to pull back our directionally porous set in Rn+1 to directionally porous sets in X then rescale. Temporarily fix ε > 0, y ∈ [0, 1 − ε]n and w ∈ X. Let v1, . . . , vn+1 be linearly independent vectors in X and L : Span(v1, . . . , vn+1) → Rn+1 be the corresponding bijective linear map sending v1, . . . , vn+1 to the standard basis e1, . . . , en+1 of Rn+1. Using the Hahn-Banach theorem we can extend L to a linear map eL defined on X with keLk ≤ (n + 1)kLk. Lemma 6.1. Let eL be as defined above. Then the preimage eL−1(P ) of a direction- ally porous set P ⊂ Rn+1 is a directionally porous set in X. 16 GARETH SPEIGHT Proof. Suppose P is directionally porous with porosity constant ρ. Let x ∈ eL−1(P ). Then eL(x) ∈ P so there exists v ∈ Rn+1 with kvk = 1 such that for any δ > 0 there exists t ∈ R with t < δ such that The previous line implies B(eL(x) + tv, ρt) ∩ P = ∅. (cid:3) Hence eL−1(P ) is directionally porous. Consider the map T : Γn(X) → Γn(Rn+1) defined by (cid:19) B(x + tL−1(v), ρt/keLk) ∩eL−1(P ) = ∅. T (f )(x) =eL(cid:18) f (y + εx) − w ε for f ∈ Γn(X) and x ∈ [0, 1]n. The map f 7→ T (f ) +eL(w)/ε is a continuous linear surjection and hence, by the open mapping theorem, is open. It follows that T is continuous and open. Hence if a set A ⊂ Γn(Rn+1) is meager then the preimage T −1(A) ⊂ Γn(X) is also meager. The following lemma follows easily from the definition of T and basic facts about Hausdorff measures. Lemma 6.2. Let q : [0, 1]n → X be the affine plane q(x) = w + (x1 − y1)v1 + . . . + (xn − yn)vn. Define M (ε, y, w, v1, . . . , vn+1) to be the set of surfaces f ∈ Γn(X) such that kf (x) − q(x)k < εR/kLk for x ∈ y + [0, ε]n, k∂f /∂xi(x) − vik < R/kLk for x ∈ y + [0, ε]n and i = 1, . . . , n, and Hn(f (y + [0, ε]n) ∩ (w + εeL−1(P ))) < βεn/kLkn. Then M (ε, y, w, v1, . . . , vn+1) ⊂ T −1{f ∈ BC 1(p, (n+1)R) : Hn(f ([0, 1]n)∩P ) < (n+1)nβ} and hence M (ε, y, w, v1, . . . , vn+1) is meager in Γn(X). It follows immediately from the lemma that for any choice of v1, . . . , vn+1 linearly directionally porous this establishes the general case of Theorem 1.4. independent in X the set Sn(eL−1(P ), 0) is not meager in Γn(X). Since eL−1(P ) is Now we suppose X is separable with countable dense subset F ⊂ X. Let D be the set of all bijective linear maps corresponding to all linearly independent (n + 1)-tuples v1, . . . , vn+1 ∈ F in the way defined earlier. Let Q =[(w + εeL−1(P )) where the union is taken over rational ε > 0, w ∈ F and L ∈ D. Each of the sets show X \ Q is Γn-null. w + εeL−1(P ) is directionally porous and hence Q is σ-directionally porous. We It is stated in Lemma 5.3.5 [4] that if M ′ = {f ∈ Γn(X) : rank df < n on a set of positive measure} then M ′ is meager. SURFACES MEETING POROUS SETS IN POSITIVE MEASURE 17 Lemma 6.3. Define M = M ′ ∪[ M (ε, y, w, v1, . . . , vn+1) where the union is taken over rational ε > 0, y ∈ [0, 1 − ε]n ∩ Qn, w ∈ F and all (n + 1)-tuples v1, . . . , vn+1 of linearly independent vectors in F . Suppose the image of a surface f ∈ Γn(X) meets X \ Q in positive measure. Then f belongs to the meager set M . Proof. Since f is Lipschitz it follows Ln([0, 1]n \ f −1(Q)) > 0. If f does not belong to M ′ ⊂ M then there exists z ∈ (0, 1)n such that rank df (z) = n and z is a Lebesgue density point of [0, 1]n \ f −1(Q). If Lw1,...,wn+1 : Span(w1, . . . , wn+1) → Rn+1 denotes the linear map sending linearly independent vectors w1, . . . , wn+1 ∈ X to e1, . . . , en+1 ∈ Rn+1 it is not hard to show the map defined by is continuous on the open set on which it is defined. (w1, . . . , wn+1) 7→ kLw1,...,wn+1k Since rank df (z) = n we can fix vn+1 ∈ F such that ∂f /∂x1(z), . . . , ∂f /∂xn(z), vn+1 are linearly independent. Hence there exists η > 0 such that if kvi − ∂f /∂xi(z)k < η for i = 1, . . . , n then η ≤ kLv1,...,vn+1k ≤ 1/η. Fix rational ε > 0 to be small enough so that kf (x) − f (z) − ∂f /∂x1(z)(x1 − z1) − . . . − ∂f /∂xn(z)(xn − zn)k < εηR/4 and k∂f /∂xi(x) − ∂f /∂xi(z)k < ηR/2 for x ∈ z + [0, 2ε]n ⊂ [0, 1]n and i = 1, . . . , n and, using the fact that z is a density point of [0, 1]n \ f −1(Q), Hn(f (z + [0, 2ε]n) ∩ Q) < βηnεn. Now fix y ∈ [0, 1 − ε]n ∩ Qn sufficiently close to z so that kf (x) − f (y) − ∂f /∂x1(z)(x1 − y1) − . . . − ∂f /∂xn(z)(xn − yn)k < εηR/2 for x ∈ y + [0, ε]n ⊂ z + [0, 2ε]n. There exists w ∈ F and linearly independent v1, . . . , vn ∈ F with kvi − ∂f /∂xi(z)k < η for i = 1, 2, . . . , n such that if q : [0, 1]n → X is the affine plane q(x) = w + (x1 − y1)v1 + . . . + (xn − yn)vn then and kf (x) − q(x)k < εηR k∂f /∂xi(x) − vik < ηR We note η ≤ kLv1,...,vn+1k ≤ 1/η and observe for x ∈ y + [0, ε]n and i = 1, . . . , n. Hn(f (y + [0, ε]n) ∩ (w + εeL−1 v1,...,vn+1(P ))) ≤ Hn(f (z + [0, 2ε]n) ∩ Q) < βηnεn ≤ βεn/kLv1,...,vn+1kn. Hence f ∈ M (ε, y, w, v1, . . . , vn+1) ⊂ M as required. (cid:3) 18 GARETH SPEIGHT Lemma 6.3 shows X \ Q is Γn-null. Since Q is σ-directionally porous this con- cludes the proof of Theorem 1.5. References [1] L. Zaj´ıcek, Porosity and σ-porosity, Real Anal. Exchange, Volume 13 (1987/1988), Number 2, 314-350. [2] L. Zaj´ıcek, On σ-porous sets in abstract spaces, Abstract and Applied Analysis, Volume 2005, Number 5, 509-534. [3] Y. Benyamini, Geometric Nonlinear Functional Analysis, Volume 1, American Mathematical Society, Colloqium Publications, Volume 48, 1999. [4] J. Lindenstrauss, D. Preiss and J. Tiser, Fr´echet Differentiability of Lipschitz Functions and Porous Sets in Banach Spaces, Annals of Mathematics Studies, 179, Princeton University Press, Princeton, NJ, 2012. [5] J. Lindenstrauss and D. Preiss, On Fr´echet Differentiability of Lipschitz maps between Banach Spaces, Annals of Mathematics, Volume 157 (2003), 257-288. [6] N. Aronszajn, Differentiability of Lipschitzian mappings between Banach spaces, Studia Math, Volume 57 (1976), 147-190. [7] J. P. R. Christensen, Measure theoretic zero sets in infinite dimensional spaces and applica- tions to differentiability of Lipschitz mappings II, Publ Dep Math Lyon, Volume 10 (1973), 29-39. [8] P. Mankiewicz, On the differentiability of Lipschitz mappings in Fr´echet spaces, Studia Math, Volume 45 (1973), 15-29. [9] D. Preiss and J. Tiser, Two unexpected examples concerning differentiability of Lipschitz func- tions on Banach spaces, Operator Theory: Advances and Applications, Volume 77 (1995), 219-238. [10] D. Preiss, Fr´echet Differentiability of Lipschitz functions, Journal of Functional Analysis, Volume 91 (1990), 312-345. [11] W. Rudin, Principles of Mathematical Analysis, McGraw-Hill Science/Engineering/Math, Third Edition, 1976. [12] L. Evans and R. Gariepy, Measure Theory and Fine Properties of Functions, Studies in Advanced Mathematics CRC Press, 1991. E-mail address: [email protected]
1806.02994
1
1806
2018-06-08T07:20:13
Existence of Large Independent-Like Sets
[ "math.FA" ]
Let $G$ be a compact abelian group and $\Gamma$ be its discrete dual group. For $N \in \mathbb{N}$, we define a class of independent-like sets, $N$-PR sets, as a set in $\Gamma$ such that every $\mathbb{Z}_N$-valued function defined on the set can be interpolated by a character in $G$. These sets are examples of $\varepsilon$-Kronecker sets and Sidon sets. In this paper we study various properties of $N$-PR sets. We give a characterization of $N$-PR sets, describe their structures and prove the existence of large $N$-PR sets.
math.FA
math
EXISTENCE OF LARGE INDEPENDENT-LIKE SETS ROBERT (XU) YANG Abstract. Let G be a compact abelian group and Γ be its discrete dual group. For N ∈ N, we define a class of independent-like sets, N -PR sets, as a set in Γ such that every ZN -valued function defined on the set can be interpolated by a character in G. These sets are examples of ε-Kronecker sets and Sidon sets. In this paper we study various properties of N -PR sets. We give a characteri- zation of N -PR sets, describe their structures and prove the existence of large N -PR sets. 8 1 0 2 n u J 8 ] . A F h t a m [ 1 v 4 9 9 2 0 . 6 0 8 1 : v i X r a 1. Introduction Independence is a property prevalent throughout mathematics. In har- monic analysis, independence has been used to produce curious examples. For example, a perfect independent set in R was first constructed by Von Neumann [10]. Independent Cantor sets in non-discrete locally compact abelian groups were constructed by Hewitt and Kakutani [8], in part for showing that M(G) is asymmetric [7]. In these examples, by 'independent' we mean algebraically independent: a set of (non-trivial) characters E is i = 1, then independent if whenever γ1, ..., γN ∈ E, mi ∈ Z, and QN γmi i = 1 for all i. i=1 γmi The Rademacher functions in the dual of the infinite direct product of infinitely many copies of Z2 is a set of characters which is both algebraically and probabilistically independent. These functions have proven to be very useful in harmonic analysis. In particular, they have the property that every ±1-valued function defined on the set is evaluation at some x in the group. A similar interpolation property holds for all algebraically independent sets. A weaker notion than independence is an ε-Kronecker set. Definition. Let ε > 0. The set E is said to be (weak) ε-Kronecker if for every ϕ : E → T there exists x ∈ G such that ϕ(γ) − γ(x) < ε for all γ ∈ E (resp. ϕ(γ) − γ(x) ≤ ε for all γ ∈ E). 2010 Mathematics Subject Classification. Primary 43A25; Secondary 43A46. Key words and phrases. Sidon Sets, Independent Sets, Kronecker Sets. Research was partially supported by NSERC grant 2016-03719. 1 2 ROBERT (XU) YANG This notion was inspired in part by the classical approximation theorem of Kronecker, with early work done by by Hewitt and Kakutani [8] and Rudin [13]. The terminology was introduced by Varapolous in [14]. The set of Rademacher functions is clearly an example of a weak √2- Kronecker set. Kronecker-like sets have been studied intensively, and are known to have many interesting properties. For example, Hare and Ramsey [6] proved that every ε-Kronecker set with ε < 2 is a Sidon set. Graham and Hare in [3] introduced the weaker notion of pseudo-Rademacher sets, sets of characters where every ±1-valued function is point-wise evaluation, in order to study the problem of the existence of Kronecker-like sets. Galindo and Hernandez in [1] and Graham and Lau in [5] both consider interpolation sets of characters of finite order. For other references and further background information we refer the reader to [4]. In this paper, we generalize this notion to N-pseudo-Rademacher sets (or N-PR sets for short), sets of characters with the property that every ZN -valued function on the set is point-evaluation. Of course, a pseudo- Rademacher set is a 2-PR set and we will see that N-PR sets are ε-Kronecker sets for suitable ε = ε(N), which tends to 0 as N → ∞. We give an algebraic characterization of N-PR sets, compare them with ε-Kronecker sets, describe their structures and prove existence theorems (Theorem 4.2) of large N-PR sets. Theorem 4.2 gives a new proof that any uncountable subset in Γ contains a large ε-Kronecker set. 2. Characterization of N -PR sets Throughout this paper G is a compact abelian group and Γ is its discrete dual group. Definition. Let E ⊂ Γ be a subset and N ∈ N. We define E to be an "N-pseudo-Rademacher" set (or N-PR set) if for every ϕ : E → ZN ⊂ T, there exists x ∈ G such that ϕ(γ) = γ(x) for all γ ∈ E. In this section we give an algebraic characterization of N-PR sets. We first establish some useful lemmas. Lemma 2.1 (Quotient and Subgroup Lemma). Let E ⊂ Γ, n ∈ N, γ ∈ Γ and Λ ⊂ Γ be a subgroup. (1) Let q : Γ → Γ/Λ be the quotient map. If q is one-to-one on E and q(E) is n-PR, then E is n-PR. (2) Suppose E ⊂ Λ. Then E is n-PR as a subset of Γ if and only if E is n-PR as a subset of Λ. EXISTENCE OF LARGE INDEPENDENT-LIKE SETS 3 Proof. (1) Suppose q : Γ → Γ/Λ is one-to-one on E and q(E) is n-PR. We will show that E is n-PR. Let ϕ : E → Zn be a function. Because q is one-to-one on E, for each γ, β ∈ E, if β 6= γ, then γ − β /∈ Λ. Thus, we can define ϕ′ : q(E) → Zn via ϕ′(γ + Λ) = ϕ(γ) for γ ∈ E. Since q(E) is n-PR, there exists x ∈ Λ⊥ = dΓ/Λ such that ϕ′(γ + Λ) = x(γ + Λ) for all γ ∈ E. As x ∈ Λ⊥, ϕ(γ) = γ(x) for all γ ∈ E. This means E is n-PR. (2) We first suppose E is an n-PR subset of Γ. Let ϕ : E → Zn be a function. There exists x ∈ G such that ϕ(γ) = γ(x) for all γ ∈ E. Let x + Λ⊥ ∈ G/Λ⊥ =bΛ. Since E ⊂ Λ, for all γ ∈ E we have ϕ(γ) = γ(x) = (x + Λ⊥)(γ). This means E is n-PR as a subset of Λ. The proof of the converse part of (2) is similar. (cid:3) The following proposition is a stronger version of Lemma 3.2 in [9]. Proposition 2.2. Let E ⊂ Γ. The following are equivalent: (1) For all ϕ : E → T with ϕ(γ) ∈ Range(γ) there exists x ∈ G such that ϕ(γ) = γ(x) for γ ∈ E. (2) E is independent. Proof. Assume E is independent and that ϕ : E → T satisfies ϕ(γ) ∈ Range(γ) for γ ∈ E. By similar arguments to the proof of Lemma 2.1, if we can find x ∈dhEi such that ϕ(γ) = γ(x) for all γ ∈ E, then there exists x′ ∈ G such that ϕ(γ) = γ(x′). Thus we may assume Γ = hEi =Lγ∈Ehγi. For each γ ∈ E, there exists xγ ∈ chγi such that γ(xγ) = ϕ(γ). If E is finite, we let x =Qγ∈E xγ and x can interpolate ϕ exactly. For the case that E is infinite, since G is compact, we let x ∈ G be a cluster point of the following set (Yγ∈F xγ : F ⊂ E,F < ∞) , and such an x can interpolate ϕ exactly. 1 i ...γmk k = 1 but γmi Conversely, if E is not independent, then there exist γ1, ..., γk ∈ E and m1, ..., mk ∈ Z such that γm1 6= 1 for all 1 ≤ i ≤ k. Consider the function ϕ : E → T such that ϕ(γi) = 1 for all i > 1, ϕ(γ1)m1 6= 1 and ϕ(γ) ∈ Range(γ) for all γ ∈ E. Notice that such ϕ exists, because γm1 1 ) 6= {1} and hence there exists x ∈ G such that γ1(x)m1 6= 1. Let ϕ(γ1) = γ1(x). This function ϕ cannot be interpolated by any x ∈ G. 6= 1 means Range(γm1 (cid:3) 1 4 ROBERT (XU) YANG Corollary 2.3. Let N ∈ N and E ⊂ Γ. If E is independent and ZN ⊂ Range(γ) for all γ ∈ E, then E is N-PR. Proof. This follows directly from Proposition 2.2. (cid:3) Theorem 2.4. The following are equivalent: (1) E ⊂ Γ is N-PR. i=1 γmi i=1 γmi i = 1 for some i=1 γi(x)mi, whileQn cannot be interpolated by any x ∈ G and therefore (1) fails. Conversely, suppose (2) holds. Let EN = hγN : γ ∈ Ei, the subgroup mi ∈ Z, then N divides mi for all i. Proof. Suppose (2) fails. Then there exist distinct γi ∈ E and integers mi ∈ i = 1, while N does not divide m1. Let f : E → ZN be given by f (γ1) = e2πi/N and f (γ) = 1 for all γ 6= γ1. For any x ∈ G, i=1 f (γi)mi = f (γ1)m1 6= 1. Thus, this function f (2) If γi ∈ E for 1 ≤ i ≤ n are distinct and Qn Z with Qn 1 =Qn generated by(cid:8)γN : γ ∈ E(cid:9), and π : Γ → Γ/EN be the quotient map. Ele- distinct π(γi) and mi ∈ Z such thatQn kj ∈ Z, 1 ≤ j ≤ s and s ≥ n such thatQn γj ∈ E. Hence,Qn ments in EN have the form γk1N for γ1, ...γm ∈ E and k1, ..., km ∈ Z. We claim that π(E) ⊂ π(Γ) is independent. Indeed, suppose γi ∈ E with i ∈ ker(π). Then there exists with distinct = 1 ∈ Γ. By (2), N mi − kiN for all 1 ≤ i ≤ n and hence N mi for all 1 ≤ i ≤ n. This means each γmi i ∈ ker(π) and therefore the set π(E) is independent. Moreover, we claim that π is one-to-one on E. Suppose, otherwise, that 2 ∈ ker(π). Then there exists kj ∈ Z, with distinct γj ∈ E. = 1. Again, (2) gives N 1 − k1N and there are γ1 6= γ2 ∈ E and γ1γ−1 1 ≤ j ≤ s and s ≥ 2 such that γ1γ−1 j=3 γ−kjN We have γ1−k1N N − 1 − k2N, which are not possible. Qs γ−1−k2N 2 2 = Qs i =Qs Similar arguments show that elements of π(E) have order N. Thus π(E) is N-PR from Proposition 2.2. Because π is also one-to-one on E, Lemma 2.1 gives that E is N-PR, proving (1). (cid:3) j=n+1 γ−kjN j i=1 γmi−kiN i i=1 γmi i=1 γmi 1 j ...γkmN m 1 Qs j=1 γkj N j j=1 γkjN j Corollary 2.5. Let a, b ∈ N be co-prime. The subset E ⊂ Γ is (ab)-PR if and only if E is both a-PR and b-PR. Proof. Since Za, Zb ⊂ Zab, if E is (ab)-PR, E is both a-PR and b-PR. To see the converse, we assume E is both a-PR and b-PR. Consider γ1, ..., γn ∈ E i = 1. Since E is both a-PR and b-PR, ami and bmi for all 1 ≤ i ≤ n. Since a and b are co-prime, abmi for all 1 ≤ i ≤ n. Hence, by Theorem 2.4, E is (ab)-PR. and m1, ..., mn ∈ Z such thatQn i=1 γmi (cid:3) EXISTENCE OF LARGE INDEPENDENT-LIKE SETS 5 Corollary 2.6. Suppose E ⊂ Γ is N-PR and hEi ∩hγi = {1}. Then γE is also N-PR. Proof. Let γ1, ...γn ∈ E and m1, ..., mi ∈ Z be such thatQn Since hEi ∩ hγi = {1}, we have Qn 1 ≤ i ≤ n. This shows γE is N-PR by Theorem 2.4. i=1(γγi)mi = 1. i = 1 and therefore Nmi for all i=1 γmi (cid:3) Next we compare N-PR sets with ε-Kronecker sets. We first note that N-PR sets are weak ε-Kronecker sets for appropriate ε depending on N. Proposition 2.7. N-PR sets are weak ε-Kronecker for ε = 1 − e Proof. Assume E ⊂ Γ is N-PR. Let ϕ : E → T. Define ϕN : E → ZN by ϕN (γ) = t, where t ∈ ZN satisfies t − ϕ(γ) ≤ 1 − e N . Let x ∈ G be such that γ(x) = ϕN (γ) for γ ∈ E. We have γ(x) − ϕ(γ) ≤ 1 − e N . N . πi πi πi (cid:3) Of course, not every ε-Kronecker set in a torsion group is N-PR. Here is an example. Example. For every ε > 0 and N ∈ N, there exists an infinite ε-Kronecker set E with every γ ∈ E having finite order a multiple of N, but E is not N-PR. Let N ∈ N and ε > 0. Let (pi)∞ coprime to N. Let Γ = ZN ⊕Li≥1 Zpi. Let S := {(1, 0, 0, ...), (1, 1, 0, 0, ...), (1, 1, 1, 0, 0, ...), ...} . i=1 be an increasing sequence of primes Each element in S has order a multiple of N. It is not hard to see that if we exclude finitely many elements of small orders, we have a co-finite ε-Kronecker set E ⊂ S. From Theorem 2.4 no subsets of S other than singletons are N-PR and therefore E is not N-PR. Remark. Recall that the Bohr compactification of Γ, denoted by Γ, is defined as the dual group of Gd, where Gd is the group G equipped with the discrete topology. Let E be the closure of E in Γ. In [4], Ex. 1.4.3, it is shown that there is an ε-Kronecker set E and an integer M ≥ 1 such that (E ∪ E−1)M has Haar measure 1. That is not the case for N-PR sets. Indeed, if E is any N-PR set with N ≥ 3, then the Haar measure of (E ∪ E−1)M is 0 for all M ≥ 1. The proof is very similar to the proof of Theorem 3.1 in [2] and therefore is omitted. 6 ROBERT (XU) YANG 3. Structure of N -PR sets In this section, we investigate the structure of N-PR sets. We rely heavily on the following structure theorem for general abelian groups. Notation: Let p be a prime number. The group C(p∞) is the group of all pn-th roots of unity for n ≥ 1. Theorem 3.1. [11] Every abelian group Γ is isomorphic to a subgroup of Mα Qα ⊕Mβ C(p∞ β ), where Qα are copies of Q and pβ are prime numbers. Notation: (1) We let Γ0 be the torsion subgroup of Γ and π0 : Γ → Γ/Γ0 be the quotient map. In the notation of Theorem 3.1, π0 :Lα Qα ⊕ Lβ C(p∞ β ) →Lα Qα is the quotient map. (2) Let p ∈ N be a prime number and n ∈ N. We let Γpn be the subgroup of Γ0 containing elements whose orders are not a multiple of pn, or equiva- lently, whose orders are not divisible by pn. Notice Γpn is indeed a subgroup because if γ1 and γ2 have orders m1 and m2, neither a multiple of pn, then lcm(m1, m2), the order of γ1γ2, is not a multiple of pn. Let πpn : Γ → Γ/Γpn be the quotient map. (3) For γ ∈Lα∈A Qα ⊕Lβ∈B C(p∞ the i-th coordinate of γ. In particular, if Γ =Li C(p∞ map πpi can be viewed as πpi = Proji. β ) and i ∈ A ∪ B, we let Proji(γ) be i ) with pi distinct, the Proposition 3.2. Let E ⊂ Γ, n ∈ N and p be a prime number. The follow- ing are equivalent: (1) E is pn-PR. (2) πpk(E) is pn+1−k-PR and πpk is one-to-one on E for all 1 ≤ k ≤ n. (3) πpk(E) is pn+1−k-PR and πpk is one-to-one on E for some 1 ≤ k ≤ n. Proof. We first show (1) implies (2). Fix 1 ≤ k ≤ n and we suppose E ⊂ Γ is pn-PR. We first claim that πpk is one-to-one on E. This is because if γ1, γ2 ∈ E have γ1γ−1 has finite order that is not divisible by pk. But from Theorem 2.4, this implies E is not even pk-PR and therefore contradicts that E is pn-PR. Next, we show that πpk(E) is pn+1−k-PR. We let γ1, ..., γs ∈ E and βi := πpk(γi) for 1 ≤ i ≤ s. Suppose i ∈ Γpk and therefore there exists l ∈ N, which is not divisible by pk, such that i = 1. By Theorem 2.4, pnlmi and hence pn+1−kmi. Theorem 2.4 for some m1, ..., ms ∈ Z we haveQs Qs i = 1. This meansQs 2 ∈ Γpk, then γ1γ−1 implies πpk(E) is pn+1−k-PR. i=1 γlmi i=1 βmi i=1 γmi 2 EXISTENCE OF LARGE INDEPENDENT-LIKE SETS 7 i=1 γmi i = 1. Then Qs Since (2) implies (3) is obvious, it remains to show (3) implies (1). We assume (3) holds for some 1 ≤ k ≤ n. Let γ1, ..., γs ∈ E be distinct and i=1 πpk(γi)mi = 1 and the injectivity of πpk implies πpk (γi) are distinct. Since πpk(E) is pn+1−k-PR, we have pn+1−kmi for all 1 ≤ i ≤ s. i=1 γmi/pk−1 m1, ..., ms ∈ Z such that Qs If n + 1− k ≥ k, we note that order of the productQs ∈ Γpk. This means Qk pk−1 and therefore Qk i=1 γmi/pk−1 divides i=1 πpk(γi)mi/pk−1 = 1 ∈ Γ/Γpk. Hence, by Theorem 2.4, we have pn+1−k divides mi/pk−1 and this gives pn divides mi for all 1 ≤ i ≤ s. Theorem 2.4 gives (1). Otherwise, we have n + 1 − k < k. Notice that the order of the product i=1 γmi/pn+1−k ∈ Γpk. As above, we have pn+1−k divides mi/pn+1−k, which means p2(n+1−k)mi. We continue doing this until we reach r(n + 1 − k) ≥ k for some r ∈ N. The previous case gives (1). divides pn+1−k. Since n + 1 − k < k,Qk i=1 γmi/pn+1−k Qs (cid:3) i i i i Corollary 3.3. Suppose Γ = LC(p∞). Then pn-PR subsets in LC(p∞) are in correspondence with p-PR subsets inLC(p∞) in the following man- ner: Given a set E ⊂LC(p∞), we let Epn :=nγpn−1 : γ ∈ Eo and E1/pn := {ξγ : γ ∈ E}, where ξpn−1 E. If E is p-PR then E1/pn is pn-PR and E = E1/pn. Proof. Since onL C(p∞), the map πpn can be identified as πpn(γ) = γpn−1, = γ. If E is pn-PR then Epn is p-PR and Epn = this follows from Proposition 3.2. (cid:3) γ 1 ...pnk Corollary 3.4. Let N = pn1 E is N-PR if and only if πp all 1 ≤ i ≤ k. Proof. This follows from Corollary 2.5 and Proposition 3.2. (E) is pi-PR and πp ni i ni i for distinct primes pi and ni ∈ N. Then is one-to-one on E for (cid:3) k We thus have the following result about the structure of N-PR sets in the torsion subgroup. Proposition 3.5. Let E ⊂ Γ0 be an N-PR set and N = pn1 k where the pi are distinct prime numbers. There exist pni i ) for 1 ≤ i ≤ k, and bijections f2 : E1 → E2, ..., fk : E1 → Ek such that we can represent E as i -PR sets Ei ⊂L C(p∞ 1 ...pnk E = {γ ⊕ f2(γ) ⊕ ... ⊕ fk(γ) ⊕ βγ : γ ∈ E1} for some βγ ∈Lp6=pi ∀1≤i≤k C(p∞). Proof. For 1 ≤ i ≤ k, we let Ei := πpi(E). Since E is N-PR, from Proposition 3.2, each Ei is pni i -PR and the maps πpi : E → Ei are injective. For 2 ≤ i ≤ k, 8 ROBERT (XU) YANG we define fi on E1 as fi(πp1(γ)) := πpi(γ) for γ ∈ E. The injectivity of πp1 on E ensures the maps are well-defined. Moreover, the injectivity of πpi implies fi is injective. Each fi is clearly surjective by its construction and therefore a bijection. Since each γ ∈ E can be represented as γ = ⊕k i=1πpi(γ) ⊕ βγ for some βγ ∈Lp6=pi ∀1≤i≤k C(p∞), Proposition 3.5 follows. Next, we will discuss the relation between p-PR sets and independent (cid:3) sets. We first note that not every p-PR set is an independent set. The following example shows a p-PR set whose only independent subsets are singletons. Example. Consider E = {γn : n ≥ 2} ⊆ Li≥1 C(p∞), where Proj1(γn) = 1/p2, Projn(γn) = 1/p and Projk(γn) = 0 for all k 6= 1, n. Here we use additive group operation. E is p-PR by Theorem 2.4, but E does not contain any independent subsets other than singletons because for all i 6= j, pγi = pγj 6= 1. Similar to independent sets, we have the following result about the max- imum size of a p-PR set inside a product group. Q ⊕Li∈B2 C(p∞) is p-PR. Then E ≤ B1 + B2. Proof. We identify each element in C(p∞) in the form of a/pn for some Q ⊕ Proposition 3.6. Suppose E ⊆ Li∈B1 a, n ∈ N with additive group operation. Hence, we can identifyLi∈B1 Li∈B2 C(p∞) as a subset of the real vector space RB1+B2. We claim that if E ⊆ Li∈B1 Q ⊕Li∈B2 C(p∞) is p-PR, E is linearly independent in RB1+B2. Indeed, suppose that E is not linearly indepen- dent. Then there exist {γi 1 ≤ i ≤ k} ⊆ E and a1, ..., ak ∈ R such that a1γ1 + ... + akγk = 0, while not all ai's are zero and each γi ∈ RB1+B2 is identified as above. Since the entries of each γi are in Q, we may assume ai ∈ Q for all 1 ≤ i ≤ k. Furthermore, we may assume ai ∈ Z. Notice that if pai for all i, we may replace ai by ai/p. Hence, we may find a choice of such ai's, not all divisible by p. This contradicts that E is p-PR by Theorem 2.4. (cid:3) Thus, E ≤ B1 + B2. 4. Existence of N -PR Sets In this section, we show some existence results about N-PR sets and that large N-PR sets are plentiful. We first prove a lemma. EXISTENCE OF LARGE INDEPENDENT-LIKE SETS 9 Lemma 4.1. Assume E ⊆Lβ∈B Γβ is uncountable. (1) Let p be a prime number. If Γβ = C(p∞) for all β ∈ B, then E contains a p-PR subset of the same cardinality. (2) If Γβ = Q for all β ∈ B, then E contains an independent subset of the same cardinality. Proof. Without loss of generality, we further assume for each β ∈ B there exists γ ∈ E such that Projβ(γ) is non-trivial. (1) We first prove (1) in the special case that every γ ∈ E has order p. We consider the collection C of subsets of E defined as A ∈ C if for all finite subsets F ⊆ A there exists an arrangement F = {γ1, ..., γn} such that for each 1 ≤ k ≤ n there is some β ∈ B with Projβ(γk) non-trivial, but Projβ(γj) trivial for all 1 ≤ j ≤ k. We partially order C by inclusion and Zorn's Lemma gives a maximal S ∈ C. We claim S = E. Indeed, if S < E, we let B1 ⊂ B be given by β ∈ B1 if there exists γ ∈ S such that Projβ(γ) is non-trivial. We thus have B1 < B. Let β0 ∈ B\B1 and γ0 ∈ E be such that Projβ0(γ0) is non-trivial. Since β0 ∈ B\B1, γ0 /∈ S and we form the set S1 := S ∪ {γ0}. It is easy to see S1 ∈ C and this contradicts the maximality of S. Moreover, the construction of S, the assumption that every element has order p and Theorem 2.4 imply S is p-PR, which finishes the proof for the special case. For the general case, we let Ek be defined as the subset in E containing elements of order pk. Then E = ∪k≥0Ek and hence there exists a positive integer K ∈ N+ such that EK = E. If K = 1, the special case finishes the proof, and therefore we suppose K > 1. We have two cases. The first case is that there exists 1 ≤ n0 < K such that πpn0 (EK) = EK, while πpn0+1(EK) < EK. We let B1 ⊂ B be given by β ∈ B1 if there exists γ ∈ EK such that Projβ(γ) has order greater or equal to pn0+1. Since πpn0+1(EK) < EK, we have B1 < B = EK. For a subset C ⊂ B, we define the projection ProjC : Lβ∈B C(p∞) → Lβ∈C C(p∞). We have ProjB\B1(πpn0 (EK)) = EK. Furthermore, each α ∈ ProjB\B1(πpn0 (EK)) has the property that for all β ∈ B\B1, Projβ(α) has order p or 1. Thus, the special case gives a p-PR set F ⊂ ProjB\B1(πpn0 (EK)) such that F = ProjB\B1(πpn0 (EK)) = EK = E and the quotient lemma finishes the proof for this case. The other case is that πpK (EK) = EK. Since πpK (EK) satisfies the special case, the quotient lemma again finishes the proof. (2) The proof of (2) is similar to the first part of the argument of (1), (cid:3) but much simpler. 10 ROBERT (XU) YANG Theorem 4.2. Let E ⊂ Γ be uncountable. Then there exists a prime number p such that E contains a p-PR set of the same cardinality. Q ⊕ ⊕∞ j ), where (Bj)∞ Proof. Embed E ⊂Li∈B0 sets and pj are distinct primes. We assume that for each index i ∈S∞ j=0 are index j=0 Bj, there exists γ ∈ E such that Proji(γ) is non-trivial. Since E is uncountable and the groups Q and C(p∞ j ) are countable, there exists K ∈ N such that BK = E. j=1Li∈Bj C(p∞ If K = 0, from Lemma 4.1 (2) we may extract an independent set F ⊂ π0(E) with F = E. If we choose E′ ⊂ E such that π0 is one-to-one on E′ and π0(E′) = F , then E′ is N-PR for all N ≥ 2 by Lemma 2.1 and Proposition 2.2. Similarly, if K ≥ 1, by Lemma 4.1 (1) we extract a pK-PR subset F ⊂ πpK (E) with F = E and therefore obtain a pK-PR subset in E of the same cardinality. (cid:3) Remark. Since any p-PR set is a Sidon set, Theorem 4.2 implies any un- countable subset in Γ contains a large Sidon set. I0 sets are more special interpolation sets. A set E ⊂ Γ is I0 if every bounded function defined on E can be interpolated as a Fourier transform of a discrete measure. If p ≥ 3, any p-PR set is an I0 set, and therefore in that case any uncountable subset in Γ contains a large I0 set. Theorem 4.3. Let E ⊂ Γ be uncountable, p be a prime number and n ∈ N. Then E contains a pn-PR subset of the same cardinality if and only if πpn(E) = E. Proof. If E contains a pn-PR subset E1 of the same cardinality, by Proposition 3.2 πpn(E) ≥ πpn(E1) = E1 = E. To see the converse, we define the projection π1 :Mi∈B1 Q ⊕Mi∈B2 C(p∞) →Mi∈B2 C(p∞). Assume πpn(E) = E. We claim that either π0(πpn(E)) = E or π1(πpn(E)) = E. If π0(πpn(E)) < E and π1(πpn(E)) < E, then because E is uncount- able, πpn(E) ≤ π0(πpn(E))π1(πpn(E)) < E, which is a contradiction. If π0(πpn(E)) = E, we appeal to (2) in Lemma 4.1 to get a pn-PR set in π0(πpn(E)) and Lemma 2.1 finishes the proof. If π1(πpn(E)) = E, then (1) in Lemma 4.1 similarly finishes the proof. (cid:3) EXISTENCE OF LARGE INDEPENDENT-LIKE SETS 11 Example. If E is countable, Theorem 4.2 may fail. Let E = C(p∞). E is countably infinite, but by Proposition 3.6, E only contains p-PR sets that are singletons. Proposition 4.4. Suppose E = Γ ⊂Li∈B1 Q⊕Li∈B2 C(p∞) is a countably infinite group. Then E contains an infinite p-PR set if and only if E contains an infinite independent set. k Proof. Assume that all independent sets in E are finite. Zorn's Lemma gives a maximal independent set S ⊂ E, where the partial order is inclusion. Our assumption gives that S is finite and the maximality implies hSi = E. Since S is independent, we embed S intoLγ∈S Γγ, where Γγ = Q if γ has infinite order and Γγ = C(p∞) if γ has order some power of p. Since S is maximal, we may extend the embedding to hSi = E. From Proposition 3.6, this implies all p-PR sets in E have cardinality at most S. The converse is trivial. (cid:3) Proposition 4.5. Suppose Γ is an uncountable infinite group and N = 1 ...pmk pm1 is an integer with prime numbers pi, 1 ≤ i ≤ k. Then Γ contains (Γ) = Γ for all 1 ≤ i ≤ k. an N-PR set E with E = Γ if and only if πp Proof. If Γ contains an N-PR set E with E = Γ, then by Corollary 3.4 πp (Γ) = Γ for all 1 ≤ i ≤ k. To see the converse, first of all, if π0(Γ) = Γ, since Γ is uncountable, the conclusion follows by Lemma 4.1 and Lemma 2.1. Thus we may assume π0(Γ) < Γ. Hence Γ0 = Γ and we may further assume Γ is a torsion (Γ) = Γ for all 1 ≤ i ≤ k, then by Lemma 4.1 (1) we let group. If πp (Γ) such that Si is pi-PR with Si = Γ, 1 ≤ i ≤ k. Si be a subset in πp is one-to-one on Ji and For 1 ≤ i ≤ k we let Ji ⊂ Γ be such that πp (Ji) = Si and moreover, we may assume the order of each γ ∈ Ji is πp a power of pi. Since Ji = Γ for all 1 ≤ i ≤ k, we let, for 2 ≤ i ≤ k, fi : J1 → Ji be bijections and we form the set mi i mi i mi i mi i mi i mi i E := {γf2(γ)...fk(γ) : γ ∈ J1} . Then E = Γ. Since for all γ ∈ Ji, 1 ≤ i ≤ k, the order of γ is a power of pi, if γ ∈ J1, fi(γ) ∈ Γp m1 1 for all 2 ≤ i ≤ k and hence (γf2(γ)...fk(γ)) = πpm1 (γ). 1 πpm1 1 Similarly, πp mi i (γf2(γ)...fk(γ)) = πp (fi(γ)), mi i for all 2 ≤ i ≤ k. Hence, πp Corollary 3.4, E is N-PR. mi i (E) = πp mi i (Ji) = Si for 1 ≤ i ≤ k. By (cid:3) 12 ROBERT (XU) YANG Remark. In [3] the terminology "N-large" sets is introduced. A set E ⊂ Γ is N-large if QN (E) < E where QN : Γ → Γ/HN is the quotient map and HN ⊂ Γ is the subgroup of elements of orders divisible by N. Theorem 2.2 in [3] states that if E ⊂ Γ and N is the smallest integer for which E is N-large, then for all primes powers pn dividing N there exists a weak 1 − eπi/pn-Kronecker subset F ⊂ E with F = E. First, we note that the assumption in Theorem 4.3 is weaker than the assumption in Theorem 2.2 in [3]; specifically, if E is infinite and N-large for minimal N = pmpm1 and n ≤ m for some integer n, then πpn(E) = E. To see this, we argue by contradiction and assume πpn(E) < E. Let M := pn−1pm1 k < N. If γ ∈ Γpn ∩ HN and k is the order of γ, then k divides N while pn does not divide k. This implies k divides M and hence γ ∈ HM . Thus Γpn ∩ HN = HM . As a result, the map T : QM (E) → QN (E) × πpn(E) T (QM (γ)) = (QN (γ), πpn(γ)). 1 ...pmk k 1 ...pmk is well-defined and injective. Thus, if πpn(E) < E, then QM (E) ≤ QN (E)πpn(E) < EE = E for infinite E. Thus E is M-large and this contradicts the assumption that N is minimal. Moreover, recall that pn-PR sets are special weak 1 − eπi/pn-Kronecker sets. This shows Theorem 4.3 improves Theorem 2.2 in [3] when E is un- countable. References [1] J. Galindo and S. Hernandez, The concept of boundedness and the Bohr compactifi- cation of a MAP Abelian group, Fundamenta Mathematicae, 159 (3):195 -- 217, 2010. [2] C. C. Graham, K. E. Hare and T. W. Korner, ε-Kronecker and I0 sets in abelian groups, II: sparseness of products of ε-Kronecker sets, Math. Proc. Cambridge Philos. Soc., 140(3): 491-508, 2006. [3] C. C. Graham and K. E. Hare, Existence of large ε-Kronecker and F Z I0(U ) sets in discrete abelian groups, Colloquium Math: 1-15, 2012. [4] C. C. Graham and K. E. Hare, Interpolation and Sidon sets for compact groups, Springer, New York, 2013. [5] C. C. Graham and A. Lau, Relative weak compactness of orbits in Banach spaces associated with locally compact groups, Transactions of the American Mathematical Society, 359 (3):1129-1160, 2007. [6] K. E. Hare and L. T. Ramsey, The Relationship Between ε-Kronecker Sets and Sidon Sets, Canad. Math. Bull. 59: 521-527, 2016. [7] Hewitt, Edwin, The asymmetry of certain algebras of Fourier-Stieltjes transforms, Michigan Math. J. 5 , no. 2: 149-158, 1958. [8] E. Hewitt and S. Kakutani, A class of multiplicative linear functionals on the measure algebra of a locally compact Abelian group, Illinois J. Math. 4: 553-574, 1960. EXISTENCE OF LARGE INDEPENDENT-LIKE SETS 13 [9] K. Kunen and W. Rudin, Lacunarity and the Bohr topology, Math. Proc. Camb. Phil. Soc. 126: 117-137, 1999. [10] J. von Neumann, Ein System algebraisch unabhangiger Zahlen, Ann. of Math. 99: 134-141, 1928. [11] J. J. Rotman, An Introduction to the Theory of Groups, 4th ed., Grad. Texts in Math. 148, Springer, New York, 1995. [12] W. Rudin, Fourier-Stieltjes transforms of measures on independent sets, Bull. Amer. Math. Soc. 66: 199-202, 1960. [13] W. Rudin, Fourier Analysis on Groups, Interscience Tracts in Pure and Applied Math., No. 12. Wiley, New York, 1962. [14] N. Th. Varopoulus, Tensor algebras over discrete spaces, J. Functional Anal., 3: 321-335, 1969. Department of Pure Mathematics University of Waterloo Waterloo ON N2L 3G1, CANADA E-mail address: yangxu [email protected]
1803.10944
1
1803
2018-03-29T07:42:40
Functional Version for Furuta Parametric Relative Operator Entropy
[ "math.FA" ]
Functional version for the so-called Furuta parametric relative operator entropy is here investigated. Some related functional inequalities are also discussed. The theoretical results obtained by our functional approach immediately imply those of operator versions in a simple, fast and nice way.
math.FA
math
FUNCTIONAL VERSION FOR FURUTA PARAMETRIC RELATIVE OPERATOR ENTROPY MUSTAPHA RAISSOULI1 , 2 AND SHIGERU FURUICHI3 Abstract. Functional version for the so-called Furuta parametric relative operator entropy is here investigated. Some related functional inequalities are also discussed. The theoretical results obtained by our functional approach immediately imply those of operator versions in a simple, fast and nice way. 1. Introduction Let H be a complex Hilbert space. We denote by B(H) the C∗-algebra of bounded linear operators acting on H and by B+∗(H) the open cone of all (self- adjoint) positive invertible operators in B(H). Let A, B ∈ B+∗(H) and p ∈ [0, 1] be a real number. The following A∇pB = (1 − p)A + pB, A!pB =(cid:16)(1 − p)A−1 + pB−1(cid:17)−1 , A♯pB = A1/2(cid:16)A−1/2BA−1/2(cid:17)p A1/2 are known in the literature as the weighted arithmetic mean, weighted harmonic mean and weighted geometric mean of A and B, respectively. If p = 1/2 they are simply denoted by A∇B, A!B and A♯B, respectively. The previous operator means satisfy the following relationships A∇pB = B∇1−pA, A!pB = B!1−pA, A♯pB = B♯1−pA. (1.1) It is well-known that the following double inequality (1.2) holds for any A, B ∈ B+∗(H) and p ∈ [0, 1]. Here, the notation T ≤ S means that T, S ∈ B(H) are self-adjoint and S − T is positive semi-definite. A!pB ≤ A♯pB ≤ A∇pB Otherwise, the relative operator entropy S(AB) and the Tsallis relative oper- ator entropy Tp(AB) are, respectively, defined by, see [2, 3, 6] S(AB) = A1/2 log(cid:0)A−1/2BA−1/2(cid:1)A1/2, Tp(AB) = The following double inequality is known in the literature: A♯pB − A p , p 6= 0. A − AB−1A ≤ S(AB) ≤ B − A. (1.3) Date: Received: xxxxxx; Revised: yyyyyy; Accepted: zzzzzz. 2010 Mathematics Subject Classification. Primary 46N10; Secondary 46A20, 47A63, 47N10, 39B62, 52A41. Key words and phrases. Operator inequalities, functional inequalities, operator entropies, convex analysis. 1 2 M. RAISSOULI, S. FURUICHI In [5], Furuta introduced a parametric extension of S(AB) as follows (1.4) Sp(AB) = A1/2(cid:0)A−1/2BA−1/2(cid:1)p log(cid:0)A−1/2BA−1/2(cid:1)A1/2. In fact, Sp(AB) was introduced in [5] for any real number p, but here we restrict ourselves to the case p ∈ [0, 1]. As pointed out in [5], it is not hard to see that S0(AB) = S(AB), S1(AB) = −S(BA) and Sp(AB) = −S1−p(BA). The fundamental goal of this paper is to give an extension of Sp(AB) when the operator variables A and B are (convex) functionals. Some functional rela- tionships and inequalities are provided as well. The related operator versions are deduced in a fast and nice way. 2. Functional Extensions The previous operator concepts have been extended from the case that the vari- ables are positive operators to the case that the variables are convex functionals, see [10]. Let RH be the extended space of all functionals defined from H into R∪{+∞}. Let f, g ∈ RH be two given functionals (convex or not) and p ∈ (0, 1). The following Ap(f, g) = (1 − p)f + pg, Hp(f, g) =(cid:16)(1 − p)f ∗ + pg∗(cid:17)∗ , sin(pπ) tp−1 (1 − t)p Ht(f, g)dt Z 1 0 Gp(f, g) = π are called, by analogy, the weighted functional arithmetic mean, the weighted harmonic mean and the weighted geometric mean of f and g, respectively. Here, the notation f ∗ refers to the Fenchel conjugate of f defined through ∀x∗ ∈ H f ∗(x∗) = sup x∈H(cid:8)ℜehx∗, xi − f (x)(cid:9). (2.1) For p = 1/2 we will denote the previous functional means by A(f, g), H(f, g) and G(f, g), respectively. We extend these means on the whole interval [0, 1] by setting: A0(f, g) = H0(f, g) = G0(f, g) = f, A1(f, g) = H1(f, g) = G1(f, g) = g. (2.2) It is worth mentioning that (2.2) is not immediate from the definition of their related functional means, since our involved functionals f and/or g can take the value +∞, with the convention 0.(+∞) = (+∞) − (+∞) = +∞. With this, analogous relationships of (1.1) for the previous functional means are also valid, i.e. Ap(f, g) = A1−p(g, f ), Hp(f, g) = H1−p(g, f ), Gp(f, g) = G1−p(g, f ). In fact, the two first relations are immediate from their definitions and for the third one see a detailed proof in [9]. Also, analog of (1.2), i.e. Hp(f, g) ≤ Gp(f, g) ≤ Ap(f, g) (2.3) FUNCTIONAL VERSION 3 holds for any f, g ∈ RH and p ∈ [0, 1]. Here the notation f ≤ g means that g(x) − f (x) ≥ 0 for all x ∈ H, with the convention +∞ − (+∞) = +∞ as usual in convex analysis. The double inequality (2.3) implies that the three involved functional means are with finite values whenever f and g are as well. In two earlier papers [10] and [11] we extended S(AB) and Tp(AB) from operators to (convex) functionals, respectively, as follows: S(f g) =Z 1 0 Ht(f, g) − f t dt, Tp(f g) = Gp(f, g) − f p , p 6= 0. The previous functional concepts were constructed as extensions of their related operator versions in the following sense: if O(A, B) is one of the previous operator concepts, its functional extension F (f, g) is such that where the notation fT , for any T ∈ B(H), refers to the quadratic function gener- ated by the operator T , i.e. fT (x) = (1/2)hT x, xi for all x ∈ H. F(cid:0)fA, fB(cid:1) = fO(A,B), (2.4) 3. Needed Tools Let f ∈ RH. We denote by dom f := {x ∈ H, f (x) < +∞} the so-called effective domain of f . The notation int(dom f ) refers to the topological interior of dom f in H. The Fenchel conjugate f ∗ of f defined by (2.1) satisfies f ∗(x∗) := sup x∈dom f(cid:8)ℜehx∗, xi − f (x)(cid:9) for any x∗ ∈ H. As supremum of a family of affine (so convex) functions, f ∗ is always convex even if f is not. The conjugate map f 7−→ f ∗ is point-wise increasing and convex. That is, f ≤ g implies g∗ ≤ f ∗, and the inequality holds for any f, g ∈ RH and p ∈ [0, 1]. (cid:0)(1 − p)f + pg(cid:1)∗ ≤ (1 − p)f ∗ + pg∗ The sub-differential of f at x ∈ dom f is the set ∂f (x) defined through ∂f (x) =(cid:8)x∗ ∈ H; ∀z ∈ H f (z) ≥ f (x) + ℜehx∗, z − xi(cid:9). As it is well known, ∂f (x) is a (possibly empty) convex and closed set. If x ∈ int(dom f ) then ∂f (x) 6= ∅. In the case where ∂f (x) 6= ∅ then we have the equivalence: x∗ ∈ ∂f (x) ⇐⇒ f (x) + f ∗(x∗) = ℜehx∗, xi. As usual we denote by Γ0(H) the cone of all functionals f ∈ RH that are convex, lower semi-continuous and proper (i.e. not identically equal to +∞). It is well-known that f ∗∗ := (f ∗)∗ ≤ f for any f ∈ RH and, f ∈ Γ0(H) if and only if f = f ∗∗ := (f ∗)∗. Moreover, x∗ ∈ ∂f (x) always implies x ∈ ∂f ∗(x∗), with reversed implication provided that f ∈ Γ0(H). 4 M. RAISSOULI, S. FURUICHI The function f is called Gateaux-differentiable (in short G-differentiable) at x if the directional derivative ′ f (x, d) = lim t↓0 f (x + td) − f (x) t of f at x exists in all direction d ∈ H and the map d 7−→ f ′(x, d) is linear. In this case we write f ′(x, d) = ∇f (x).d and ∇f (x) is called the G-derivative of f at x. It is well known that if f is convex and G-differentiable at x then ∂f (x) = {∇f (x)}. For the sake of clearness and simplicity for the reader, we state the following example illustrating the previous concepts. Example 3.1. Let A ∈ B(H) and let fA be the quadratic function associated to A, i.e. fA(x) = (1/2)hAx, xi for all x ∈ H. (i) Assume that A ∈ B+∗(H). Then fA is convex and G-differentiable on H and so ∀x ∈ H ∂fA(x) = {∇fA(x)} = {Ax}. The coefficient 1/2 appearing in fA enjoys a symmetry role, in the aim to have (fA)∗(x∗) = (1/2)hA−1x∗, x∗i for all x∗ ∈ H, or in short (cid:0)fA(cid:1)∗ = fA−1. (ii) For any A, B ∈ B(H) it is easy to check that fA ± fB = fA±B and fA(Bx) = fBAB(x) for any x ∈ H. The following result, which will be needed later, has been proved in [12]. Theorem 3.2. Let f ∈ Γ◦(H) be such that int(dom f ) is nonempty. Then (i) The following inequality sup x∗∈∂f (x)(cid:0)f ∗ − g∗(cid:1)(x∗) ≤ S(f /g)(x) ≤ (g − f )(x) holds true for all x ∈ int(dom f ). (ii) If f is moreover G-differentiable at x, then we have f ∗(cid:0)∇f (x)(cid:1) − g∗(cid:0)∇f (x)(cid:1) ≤ S(f /g)(x) ≤ (g − f )(x). As explained in [12], (3.1) as well as (3.2) is a functional extension of (1.3) from positive operators to convex functionals. For the sake of simplicity for the reader, we need to introduce auxiliary nota- tion. For f, g ∈ RH and p ∈ [0, 1] we set (3.1) (3.2) T ∗ p (f g) = (cid:0)Gp(f, g)(cid:1)∗ − f ∗ p , p 6= 0. (3.3) We have the following result summarizing the elementary properties of T ∗ p (f g). Proposition 3.3. The following assertions hold: (i) For any p ∈ [0, 1) one has T ∗ 1−p(gf ) = (cid:0)Gp(f, g)(cid:1)∗ − g∗ 1 − p . FUNCTIONAL VERSION 5 (ii) For all p ∈ (0, 1], the left hand-side of the following inequality (cid:0)Ap(f, g)(cid:1)∗ (x∗) − f ∗(x∗) p ≤ T ∗ p (f g)(x∗) ≤ g∗(x∗) − f ∗(x∗) holds for any x∗ ∈ H while the right hand-side holds for x∗ such that g∗(x∗) = +∞ or x∗ ∈ dom f ∗. Proof. (i) Follows from (3.3) with the relation Gp(f, g) = G1−p(g, f ). (ii) From (2.3) we obtain by taking the conjugate side to side Remarking that we then deduce the desired result. (cid:0)Ap(f, g)(cid:1)∗ ≤(cid:0)Gp(f, g)(cid:1)∗ ≤(cid:0)Hp(f, g)(cid:1)∗ (cid:0)Hp(f, g)(cid:1)∗ ≤ (1 − p)f ∗ + pg∗ . (cid:3) Proposition 3.4. For any A, B ∈ B+∗(H) and p ∈ (0, 1] there holds Proof. By (3.3) with (2.4) we have T ∗ p (fAfB) = fTp(A−1B−1). T ∗ p (fAfB) = (cid:0)Gp(fA, fB)(cid:1)∗ − f ∗ p A = f ∗ A♯pB − f ∗ A p = f(A♯pB)−1 − fA−1 p = fA−1♯pB−1 − fA−1 p . This, with the fact that αfT = fαT and fT − fS = fT −S for any α ∈ R and T, S ∈ B(H), immediately yields the desired result. (cid:3) 4. Functional Version of Sp(AB) As already pointed before, our aim here is to give an analog of Sp(AB) when the operator arguments A and B are (convex) functionals f and g, respectively. Such analog seems to be hard to define from (1.4), since (1.4) involves product of operators whose analogues for functionals is not known yet. For this, we need to state the following result. Theorem 4.1. The following equalities Sp(AB) = − S(cid:0)A♯pBA(cid:1) p = S(cid:0)A♯pBB(cid:1) 1 − p . (4.1) hold for any A, B ∈ B+∗(H) and p ∈ (0, 1). Proof. Indeed, we have the property T ∗S(AB)T = S(T ∗AT T ∗BT ) for any A, B ∈ B∗(H) and any invertible operator T ∈ B(H) by using Kubo-Ando theory [7] and the integral form S(AB) =Z 1 0 A!tB − A t dt. 6 M. RAISSOULI, S. FURUICHI We thus have the first equality as S (A♯pBA) = A1/2S(cid:0)I♯pA−1/2BA−1/2I(cid:1) A1/2 = A1/2S(cid:0)(A−1/2BA−1/2)pI(cid:1) A1/2 = −A1/2(A−1/2BA−1/2)p log(A−1/2BA−1/2)pA1/2 = −pSp(AB), since S(AI) = −A log A for any A ∈ B+∗(H). The second equality can be proved in a similar manner. (cid:3) Now, to give a functional version of Sp(AB) we use (4.1) which is more appro- priate for our aim, since (4.1) involves only operator concepts (relative operator entropy and operator geometric mean) whose functional extensions were already done. Taking into account to have a symmetric character between p and 1 − p in our desired definition, we then put the following. Definition 4.2. Let f, g ∈ RH and p ∈ [0, 1]. We set Sp(cid:0)f g(cid:1) = with S(cid:16)(cid:0)Gp(f, g)(cid:1)g(cid:17) 2(1 − p) − S(cid:16)(cid:0)Gp(f, g)(cid:1)f(cid:17) 2p , S0(f g) = S(f g) and S1(f g) = −S(gf ). As first result we state the following. Proposition 4.3. Let f, g ∈ RH. Then we have Further, if dom f = dom g = H then the equality S1/2(f g) = S(cid:16)(cid:0)G(f, g)(cid:1)g(cid:17) − S(cid:16)(cid:0)G(f, g)(cid:1)f(cid:17). Sp(f g) = −S1−p(gf ) holds for any p ∈ (0, 1). (4.2) (4.3) (4.4) Proof. The equality (4.3) is immediate from (4.2). However, we mention that (4.4) is not immediate from (4.2), since our involved functionals could take the value +∞. Indeed, we pay attention to the fact that, if φ, ψ ∈ RH, the equality φ−ψ = −(ψ−φ) is not always true, unless dom φ∪dom ψ = H. For this reason we have assumed in our statement that dom f = dom g = H in the aim to guarantee that Ht(cid:0)Gp(f, g), g(cid:1) or Ht(cid:0)Gp(f, g), f(cid:1) is with finite values. With this, (4.4) can be deduced from (4.2) when we refer to the relationship Gp(φ, ψ) = G1−p(ψ, φ) valid for any φ, ψ ∈ RH and p ∈ [0, 1]. (cid:3) A connection relationship between the functional parametric entropy Sp(f g) and the operator parametric entropy Sp(AB) is expressed by the following result. Proposition 4.4. Let A, B ∈ B+∗(H) and p ∈ [0, 1]. Then we have: Proof. By (4.2), with (2.4) and (4.1), we have Sp(cid:0)fAfB(cid:1) = fSp(AB). Sp(cid:0)fAfB(cid:1) = S(cid:0)fA♯pBfB(cid:1) 2(1 − p) − S(cid:0)fA♯pBfA(cid:1) 2p = f S(cid:0)A♯pBB(cid:1) 2(1 − p) − (4.5) . f S(cid:0)A♯pBA(cid:1) 2p FUNCTIONAL VERSION 7 This, with similar arguments as in the proof of Proposition 3.4, implies the desired result. (cid:3) Relationship (4.5) justifies that Sp(f g) is a reasonable extension of Sp(AB), from operators to functionals, in the sense of (2.4). For the sake of simplicity, we use in the next theorem and in its proof the following notations: Gp := Gp(f, g), G∗ We now are in a position to state the following main result. p :=(cid:0)Gp(f, g)(cid:1)∗, ∇Gp := ∇(cid:0)Gp(f, g)(cid:1), ∂Gp := ∂(cid:0)Gp(f, g)(cid:1). Theorem 4.5. Let f, g ∈ RH be such that int(cid:0)dom Gp(f, g)(cid:1) 6= ∅. Then the 2 sup following double inequality x∗∈∂Gp(x) T ∗ 1 1−p(gf )(x∗) + Tp(f g)(x)! ≤ Sp(f g)(x) 2 −T1−p(gf )(x) − sup ≤ 1 T ∗ p (f g)(x∗)! (4.6) x∗∈∂Gp(x) sup holds for any x ∈ int(cid:0)dom Gp(f, g)(cid:1) and p ∈ (0, 1). Proof. Since int(cid:0)dom Gp(f, g)(cid:1) 6= ∅ then ∂Gp(x) 6= ∅ for any x ∈ int(cid:0)dom Gp(f, g)(cid:1). Now, according to Theorem 3.2 we have, for x ∈ int(cid:0)dom Gp(f, g)(cid:1) p − f ∗(cid:1)(x∗) ≤ S(cid:0)Gpf(cid:1)(x) ≤(cid:0)f − Gp(cid:1)(x), p − g∗(cid:1)(x∗) ≤ S(cid:0)Gpg(cid:1)(x) ≤(cid:0)g − Gp(cid:1)(x). x∗∈∂Gp(x)(cid:0)G∗ x∗∈∂Gp(x)(cid:0)G∗ Multiplying (4.7) by −1/p and (4.8) by 1/(1 − p) and then summing side to side we obtain the desired inequalities, after simple manipulations with the help of Proposition 3.3. Details are simple and therefore omitted here. (cid:3) (4.7) (4.8) and sup Remark 4.6. It is worth mentioning that the condition int(cid:0)dom Gp(f, g)(cid:1) 6= ∅ is satisfied if int(cid:0)dom f ∩ dom g(cid:1) 6= ∅, since dom f ∩ dom g ⊂ dom Gp(f, g). Corollary 4.7. Let f, g ∈ RH be such that Gp(f, g) is G-differentiable at x ∈ H. Then the following inequalities (in the point-wise order sense) 1 2(cid:16)T ∗ 1−p(gf )(cid:0)∇Gp(f, g)(cid:1) + Tp(f g)(cid:17) ≤ Sp(f g) 1 hold for any p ∈ (0, 1). ≤ 2(cid:16) − T1−p(gf ) − T ∗ p (f g)(cid:0)∇Gp(f, g)(cid:1)(cid:17) Proof. Since Gp(f, g) is G-differentiable at x then ∂Gp(f, g)(x) = {∇Gp(f, g)(x)}. Substituting this in (4.6) and using the definition of the point-wise order we immediately obtain the desired inequalities. (cid:3) The operator version of the above theorem (and corollary) reads as follows. 8 M. RAISSOULI, S. FURUICHI Corollary 4.8. Let A, B ∈ B+∗(H) and p ∈ (0, 1). Then we have 1 2(cid:16)(A♯pB)T1−p(B−1A−1)(A♯pB) + Tp(AB)(cid:17) ≤ Sp(AB) 1 ≤ 2(cid:16) − T1−p(BA) − (A♯pB)Tp(A−1B−1)(A♯pB)(cid:17). Proof. Combining Corollary 4.7, Proposition 3.4 and Example 3.1,(ii) we obtain the desired operator inequalities after simple manipulations. Details are simple and therefore omitted here. (cid:3) Corollary 4.8 gives the relation between Furuta parametric relative operator entropy and Tsallis relative operator entropy, in more general setting than the result in [4, Theorem 2.3]. Acknowledgement The author (S.F.) was partially supported by JSPS KAKENHI Grant Number 16K05257. References [1] I. Ekeland and R. Temam, Convex Analysis and Variational Problems, SIAM, 1999. [2] J.I. Fujii and E. Kamei, Relative Operator Entropy in noncommutative Information Theory, Math. Japonica, 34 (1989), 341-348. [3] S. Furuichi, Inequalities for Tsallis Relative Entropy and Generalized Skew Information, Linear Multilinear Alg., 59/10 (2011), 1143-1158. [4] S. Furuichi and N. Minculete, Inequalities for relative operator entropies and operator means, Acta. Math. Vietnam (2018). https://doi.org/10.1007/s40306-018-0250-7. [5] T. Furuta, Parameteric extensions of Shannon inequality and its reverse one in Hilbert space operators, Linear Algebra Appl., 381 (2004), 219-235. [6] H. Isa, M. Ito, E. Kamei, H. Tohyama and M. Watanabe, Shannon type inequalities of a relative operator entropy including Tsallis and R´enyi ones, Ann. Funct. Anal., 6/4 (2015), 289-300. [7] F. Kubo and T. Ando, Means of positive linear operators, Math. Ann. 246 (1980), 205-224. [8] P.J. Laurent, Approximation et Optimisation, Hermann, 1972. [9] M. Raıssouli and H. Bouziane,Arithmetico-geometrico-harmonic functional mean in convex analysis, Ann. Sc. Math. Qu´ebec, 30/1 (2006), 79-107. [10] M. Raıssouli, Tsallis relative entropy for convex functionals, Int. J. Pure Appl. Math., 51/4 (2009), 555-563. [11] M. Raıssouli, Relative functional entropy in convex analysis, J. Appl. Anal., 17 (2011), 231-239. [12] M. Raıssouli, Functional versions for some operator entropy inequalities, Ann. Funct. Anal., 6/2 (2015), 204-211. [13] E.Zeidler, Nonlinear Functional Analysis and its Applications III, Springer-Verlag, 1984. 1 Department of Mathematics, Science Faculty, Taibah University, Al Madi- nah Al Munawwarah, P.O.Box 30097, Zip Code 41477, Saudi Arabia. 2 Department of Mathematics, Science Faculty, Moulay Ismail University, Meknes, Morocco. FUNCTIONAL VERSION 9 3Department of Information Science, College of Humanities and Sciences, Ni- hon University, 3-25-40, Sakurajyousui, Setagaya-ku, Tokyo, 156-8550, Japan. E-mail address: [email protected] E-mail address: [email protected]